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

A-priori estimates for generalized
Korteweg–de Vries equations in H1()H^{-1}(\mathbb{R})

Mihaela Ifrim Department of Mathematics, University of Wisconsin–Madison, Madison, WI 53706, USA ifrim@math.wisc.edu  and  Thierry Laurens Department of Mathematics, University of Wisconsin–Madison, Madison, WI 53706, USA laurens@math.wisc.edu
Abstract.

We prove local-in-time a-priori estimates in H1()H^{-1}(\mathbb{R}) for a family of generalized Korteweg–de Vries equations. This is the first estimate for any non-integrable perturbation of the KdV equation that matches the regularity of the sharp well-posedness theory for KdV. In particular, we show that our analysis applies to models for long waves in a shallow channel of water with an uneven bottom.

The proof of our main result is based upon a bootstrap argument for the renormalized perturbation determinant coupled with a local smoothing norm.

1. Introduction

The Korteweg–de Vries equation

(KdV) tu=u′′′+6uu\partial_{t}u=-u^{\prime\prime\prime}+6uu^{\prime}

(where u=xuu^{\prime}=\partial_{x}u) was originally derived as a model for the propagation of waves in a shallow channel of water [Korteweg1895]. However, over the next century, (KdV) has proved to be a fundamental model for nonlinear dispersive waves, with applications to a wide variety of physical systems spanning many fields of science (see, for example, [Crighton1995]).

Mathematically, (KdV) has also played a central role in our understanding of dispersive equations. In particular, the (KdV) equation was the first discovered example of a completely integrable PDE, a feature that has been heavily exploited in some works. This discovery sparked a pursuit into the well-posedness and dispersive behavior for this system over the next 50 years, and generated a long list of cutting-edge techniques  [Bona1976, Kato1975, Saut1976, Temam1969, Tsutsumi1971, Kenig1991, Bourgain1993, Christ2003, Kenig1996, Colliander2003, Guo2009, Kishimoto2009, KT, MR4706572], some of which rely on complete integrability, and some of which do not. On the integrable side this effort culminated in global well-posedness for initial data in H1H^{-1} on the line and the circle [Killip2019, Kappeler2006], a result that is sharp in the class of HsH^{s} spaces for both geometries [Molinet2011, Molinet2012].

Although (KdV) lied at the center of this effort, it also served as a testing ground for the introduction of new tools and innovative techniques. After their introduction, many of these methods (both integrable and non-integrable) were later adapted to broader classes of dispersive equations. However, this important step has yet to be achieved for the completely integrable methods used to prove the sharp well-posedness results [Killip2019, Kappeler2006].

One physically important class of examples are the KdV equations with variable bottom. As (KdV) has proved to be an effective model for the propagation of waves in a shallow channel of water, many authors have introduced generalizations to describe an uneven bottom [Dingemans1997, Grimshaw1999, Groesen1993, Johnson1973, Miles1979, Pudjaprasetya1996, Pudjaprasetya1999, Yoon1994, Israwi2010, Lannes2013, Tian2001]. For concreteness, consider the model

(1.1) tu=b5u′′′+6uu4bu6bu,\partial_{t}u=-b^{5}u^{\prime\prime\prime}+6uu^{\prime}-4bu^{\prime}-6b^{\prime}u,

which was derived and rigorously justified in [Israwi2010]. Here, b(x)b(x) is defined in terms of a function c(x)c(x) which describes the bottom of the channel:

b(x):=1c(x).b(x):=\sqrt{1-c(x)}.

When c(x)c(x) is a small Schwartz function, the equation (gKdV) closely resembles (KdV). Nevertheless, our understanding of the well-posedness problem is comparatively lacking.

After a change of variables (see (5.15) below), the equation (1.1) can be put in the form

(gKdV) tu=u′′′+6uu+(a1u)+a2u2+a3u+a4u\partial_{t}u=-u^{\prime\prime\prime}+6uu^{\prime}+(a_{1}u^{\prime})^{\prime}+a_{2}u^{2}+a_{3}u^{\prime}+a_{4}u

for certain coefficients aja_{j}. In this work, we will assume that the coefficients aj(t,x)a_{j}(t,x) are given functions that are sufficiently regular and localized in space, and study the corresponding solutions u:[T,T]×u:[-T,T]\times\mathbb{R}\to\mathbb{R}.

Our main contribution is the following local-in-time a-priori estimate in H1H^{-1}:

Theorem 1.1 (A-priori estimate).

There exists ϵ>0\epsilon>0 so that, if the coefficients are given smooth functions that satisfy the decay bounds uniformly for |t|T|t|\leq T and xx\in\mathbb{R}:

(1.2) |aj(t,x)|+|xaj(t,x)|\displaystyle|a_{j}(t,x)|+|\partial_{x}a_{j}(t,x)| ϵ(1+x2)1for j=1,2,3,\displaystyle\leq\epsilon(1+x^{2})^{-1}\quad\text{for }j=1,2,3,
(1.3) |a4(t,x)|+|xa4(t,x)|\displaystyle|a_{4}(t,x)|+|\partial_{x}a_{4}(t,x)| (1+x2)1,\displaystyle\lesssim(1+x^{2})^{-1},

then for any A>0A>0 there exist constants C,T>0C,T>0 so that for any smooth solution uu to (gKdV) in [T,T][-T,T] whose initial data satisfies

(1.4) u(0)H1A,\left\lVert u(0)\right\rVert_{H^{-1}}\leq A,

satisfies the uniform bound

(1.5) sup|t|Tu(t)H1C.\sup_{|t|\leq T}\left\lVert u(t)\right\rVert_{H^{-1}}\leq C.
Remark 1.2.

The lifespan TT for the bounds in the Theorem is more accurately identified in the proof as

T(1+A)4.T\gtrsim(1+A)^{-4}.

On the other hand one might expect that CC should be comparable to AA; however this is not the case because in the proof of the Theorem we work with a weighted form of the H1H^{-1} norm. So instead we get the weaker bound

CA(1+A)2.C\lesssim A(1+A)^{2}.
Remark 1.3.

The assumption that the coefficients aja_{j} are smooth in the above theorem is purely qualitative, and only serves to ensure that we can talk about smooth solutions to (gKdV). Alternatively, if we want to only assume that the coefficients satisfy (1.2) and (1.3), then the conclusion of the theorem would remain valid for rougher solutions, e.g. in L2L^{2}, once local well-posedness is known.

We do not claim that the conditions (1.2)–(1.3) in this theorem are sharp. In fact, our proof will require slightly weaker hypotheses (see (4.1)–(4.4) below for details). The main thrust of this work is to match the sharp regularity of the well-posedness theory for (KdV), and so any hypotheses on the coefficients that work constitutes significant progress. In particular, (1.5) rules out strong forms of instantaneous norm inflation in H1H^{-1}.

Applying our general result to the equation (1.1), we obtain:

Corollary 1.4.

There exists a constant ϵ>0\epsilon>0 so that, if c:c:\mathbb{R}\to\mathbb{R} is a smooth function that satisfies the small pointwise bounds

|xjc(x)|ϵ(1+x2)1for j=0,1,,4,|\partial_{x}^{j}c(x)|\leq\epsilon(1+x^{2})^{-1}\quad\text{for }j=0,1,\dots,4,

then for any A>0A>0 there exist constants ϵ,T>0\epsilon,T>0 so that for any smooth solution uu to (1.1) in [T,T][-T,T] whose initial data satisfies

u(0)H1A,\left\lVert u(0)\right\rVert_{H^{-1}}\leq A,

satisfies the uniform bound

sup|t|Tu(t)H1C.\sup_{|t|\leq T}\left\lVert u(t)\right\rVert_{H^{-1}}\leq C.

We contend that for each of the numerous applications of (KdV), our a-priori estimate (1.5) could also be employed to describe small physical imperfections using similar methods.

Another main way in which (gKdV) arises naturally is in the study of localized perturbations of a background solution to (KdV): if we take our solution u=q+Vu=q+V to be a given background wave VV plus a perturbation qq, then the equation for qq often takes the form (gKdV). For example, if V(t,x)V(t,x) solves (KdV) then qq solves

(1.6) tq=q′′′+6qq+6Vq+6Vq,\partial_{t}q=-q^{\prime\prime\prime}+6qq^{\prime}+6Vq^{\prime}+6V^{\prime}q,

and if V(x)V(x) is a fixed background profile then a forcing term V′′′+6VV-V^{\prime\prime\prime}+6VV^{\prime} is added to the RHS above. Through this lens, the special case (1.6) of gKdV equations have been of great interest in the literature.

The first phase of results for (1.6) addressed the construction of solutions using the inverse scattering transform. Two particularly common choices for VV are profiles that are periodic [Kuznetsov1974, Ermakova1982, Ermakova1982a, Firsova1988], which describe localized defects to periodic wave trains, and step-like [Buslaev1962, Cohen1984, Cohen1987, Kappeler1986], which arise in the study of bore propagation and rarefaction waves. Later, the authors of [Egorova2009, Egorova2009a] established a general framework that encompasses both of these cases, and even a mixture of the two as x+x\to+\infty and xx\to-\infty. Although the main thrust is to prove existence, a key step in many of these works is establishing persistence of regularity for solutions, much like (1.5).

While many of these results work at high regularity, these methods for existence have been adapted to classes of one-sided step-like initial data [Grudsky2014, Rybkin2011, Rybkin2018] and even to one-sided step-like elements of Hloc1()H^{-1}_{\text{loc}}(\mathbb{R}) [Grudsky2015]. Despite the lack of assumptions as xx\to-\infty (the direction in which radiation propagates), these low-regularity results require rapid decay at x+x\to+\infty and global boundedness from below.

An alternative approach was introduced in [Menikoff1972], which proves global existence and uniqueness for initial data that satisfies u(0,x)=o(|x|)u(0,x)=o(|x|) as x±x\to\pm\infty. Here, the background wave VV is evolved according to the inviscid Burger’s equation tV=6VV\partial_{t}V=-6VV^{\prime} using the method of characteristics, and then solutions to the equation for qq are constructed using a family of discretized approximate equations. Existence and uniqueness for qq is then established in a weighted H3H^{3} space.

Inspired by the theory of tidal bores and rarefaction waves (as well as kink solutions to KdV with higher-power nonlinearities), authors began turning their attention to the well-posedness problem for step-like initial data. The first phase of results employed BBM and parabolic regularizations [Bona1994, Iorio1998, Zhidkov2001], which involve adding a term to the LHS of (1.6) that eases the proof of well-posedness (εtx2q-\varepsilon\partial_{t}\partial^{2}_{x}q and εx2q-\varepsilon\partial^{2}_{x}q for ε>0\varepsilon>0, respectively) and then sending ε0\varepsilon\to 0. The former was introduced by Bona and Smith [Bona1975] in the case V0V\equiv 0 and leads to the Benjamin–Bona–Mahony equation, for which the method is named. These approaches culminated in local well-posedness for initial data qHsq\in H^{s} for s>32s>\frac{3}{2}, and was later advanced to s>1s>1 in [Gallo2005] through the incorporation of Strichartz estimates.

Recently, the work [Palacios2023] extends local well-posedness to the range s>12s>\frac{1}{2} using a synthesis of much more modern tools for well-posedness. In addition to the cases of step-like and periodic background waves VV, this result also applies to more general choices of VV with bounded asymptotics as x±x\to\pm\infty, and even more general nonlinearities in (KdV). Nevertheless, it only applies to (gKdV) in the special case where a10a_{1}\equiv 0 and a2a_{2} is constant.

Around the same time, the second author proved that (KdV) is globally well-posed for initial perturbations qH1q\in H^{-1} in [Laurens2023], provided that the background wave VV is a suitable solution to (KdV). These conditions on VV do include regularity but do not impose any assumptions on spatial asymptotics. In particular, this result applies to the important cases of smooth periodic [Laurens2023] and step-like [Laurens2022] initial data.

All of these works for (1.6) focus on the coefficients a3a_{3} and a4a_{4} in (gKdV). The introduction of the coefficient a1a_{1} already significantly changes the analysis. The local well-posedness of this equation for initial data in HsH^{s} with s>32s>\frac{3}{2} and suitable coefficients was demonstrated in [Israwi2013] using energy methods. Recently, local well-posedness was extended to the range s>12s>\frac{1}{2} in [Molinet2023].

In the most general case, the optimal well-posedness results that cover the a2a_{2} term in (gKdV) are [Craig1992, Akhunov2019, Ben] to the best of our knowledge. These works establish well-posedness for initial data at a considerably high regularity, but they apply to very general families of nonlinear equations with a KdV-type dispersion.

Let us now turn to our methods for proving the priori estimate (1.5). At the center of our analysis lies the renormalized perturbation determinant α(κ,u)\alpha(\kappa,u) (see (2.9) below for details). The perturbation determinant is a conserved quantity for (KdV) originating from scattering theory, and is rooted in the complete integrability of this system. However, the authors of [Killip2018] (and independently [Rybkin2010]) discovered a renormalization α\alpha of this quantity that satisfies

(1.7) κα(κ,u)|u^(ξ)|2ξ2+4κ2dξ=:uHκ12for all κuH12\kappa\alpha(\kappa,u)\approx\int\frac{|\widehat{u}(\xi)|^{2}}{\xi^{2}+4\kappa^{2}}\,d\xi=:\left\lVert u\right\rVert_{H^{-1}_{\kappa}}^{2}\quad\text{for all }\kappa\gg\left\lVert u\right\rVert_{H^{-1}}^{2}

and is still conserved, and used this to prove a global-in-time a-priori estimate for solutions to (KdV) in HsH^{s} spaces for 1s<1-1\leq s<1. At the same time, conserved energies in HsH^{s} spaces were independently constructed in [KT] for a full range of Sobolev exponents s1s\geq-1. As it turned out, these energies are also connected to the perturbation determinant, but with a different choice of renormalization. See also the earlier work [B] where similar bounds were proved at slightly higher regularity s>45s>-\frac{4}{5}.

The same quantity α\alpha was then used as a starting point in [Killip2019] to prove that (KdV) is well-posed in H1H^{-1} using the method of commuting flows.

Of course, introducing the coefficients aja_{j} in (gKdV) breaks the conservation of α\alpha, along with all of the other conservation laws of (KdV). In general, the value of α\alpha can grow in time, and consequently we cannot expect a global-in-time estimate. Instead, in order to establish (1.5), we prove that α\alpha must remain finite for a short period of time using a bootstrap/continuity argument.

However, we immediately encounter an obstacle. As an illustrative example, consider the L2L^{2} norm; if uu is a Schwartz solution of (KdV), then the L2L^{2} norm of u(t)u(t) is constant in time. On the other hand, if uu solves (gKdV), then the L2L^{2} norm evolves according to

(1.8) t12u2𝑑x=a1(xu)2+a2u312(xa3)u2+a4u2dx.\partial_{t}\int\tfrac{1}{2}u^{2}\,dx=\int-a_{1}(\partial_{x}u)^{2}+a_{2}u^{3}-\tfrac{1}{2}(\partial_{x}a_{3})u^{2}+a_{4}u^{2}\,dx.

How can we bound the right-hand side solely in terms of uL2\left\lVert u\right\rVert_{L^{2}} in order to close the argument?

Our strategy in this paper is to use local smoothing: due to the dispersive nature of the equation, we expect a gain in the regularity of solutions locally in space on average in time. For (KdV), this effect was discovered by Kato [Kato1983], who proved that L2L^{2} solutions are in fact in H1H^{1} locally in space at almost every time. In particular, this gives us a way to make sense of the RHS of (1.8) even for L2L^{2} solutions uu.

For the a-priori estimate (1.5), we encounter an analogous loss of derivatives phenomenon in computing the time evolution of α\alpha. The local smoothing effect attendant to this problem is to show that an H1H^{-1} solution of (gKdV) is in L2L^{2} locally in space and time:

Theorem 1.5 (Local smoothing estimate).

There exists ϵ>0\epsilon>0 so that, if the smooth coefficients aja_{j} satisfy the bounds in (1.2)–(1.3) uniformly for |t|T|t|\leq T and xx\in\mathbb{R}, then for any A>0A>0 there exist constants C,T>0C,T>0 so that for any smooth solution uu to (gKdV) in [T,T][-T,T] whose initial data satisfies

(1.9) u(0)H1A,\left\lVert u(0)\right\rVert_{H^{-1}}\leq A,

satisfies the local energy bound

(1.10) supx0TTx01x0+1|u(t,x)|2𝑑x𝑑tC.\sup_{x_{0}\in\mathbb{R}}\int_{-T}^{T}\int_{x_{0}-1}^{x_{0}+1}|u(t,x)|^{2}\,dx\,dt\leq C.

We will prove (1.5) and (1.10) simultaneously, using a bootstrap argument in terms of both α\alpha and a local smoothing norm.

In the case where the coefficients aj0a_{j}\equiv 0 vanish, an analogous local smoothing estimate was first proved in [Buckmaster2015]; see also the earlier work in [B], where similar local smoothing estimates are proved for HsH^{s} solutions with s>45s>-\frac{4}{5}. However, our proof is more closely related to the alternative approach in [Killip2019]*Th. 1.2 using the renormalized perturbation determinant. This argument is made possible by the discovery of a density ρ(x;κ,u)\rho(x;\kappa,u) for α\alpha (see (2.9) below) not known during the earlier work [Killip2018]. Moreover, if uu solves (KdV), then ρ\rho satisfies the density-flux equation (or ‘microscopic conservation law’)

(1.11) tρ+xj=0\partial_{t}\rho+\partial_{x}j=0

for a certain current j(x;κ,u)j(x;\kappa,u) (see (2.12)). Integrating this equation in space yields the conservation of α\alpha. Alternatively, first multiplying by a smooth step function and then integrating in space leads to a local smoothing estimate; this is basis of the short proof presented in [Killip2019].

Turning our attention back to (gKdV), we again encounter obstacles. The presence of the coefficients aja_{j} breaks the conservation law (1.11), and instead we have

(1.12) tρ+xj=dρ|u[(a1u)+a2u2+a3u+a4u].\partial_{t}\rho+\partial_{x}j=d\rho|_{u}\big{[}(a_{1}u^{\prime})^{\prime}+a_{2}u^{2}+a_{3}u^{\prime}+a_{4}u\big{]}.

(Here, dρ|ud\rho|_{u} denotes the functional derivative of ρ\rho at uu; see (2.14) for details.) All of terms on the RHS of (1.12) are new, and must that must be controlled in terms of α\alpha and our local smoothing norm.

Many of our estimates are in the spirit of [Bringmann2021], which proves that the fifth-order KdV equation is globally well-posed in H1H^{-1}. In order to prove their result, the authors had to prove an analogous local smoothing estimate for H1H^{-1} solutions of their system and a family of approximate equations. Together, the estimates in [Killip2019, Bringmann2021] provide upper bounds on the density ρ\rho and current jj in terms of the local smoothing and H1H^{-1} norms, the latter of which can then be controlled by α\alpha using (1.7). For example, when κ1\kappa\geq 1 we may bound

(1.13) uHκ12=|u^(ξ)|2ξ2+4κ2𝑑ξuH12\left\lVert u\right\rVert_{H^{-1}_{\kappa}}^{2}=\int\frac{|\widehat{u}(\xi)|^{2}}{\xi^{2}+4\kappa^{2}}\,d\xi\leq\left\lVert u\right\rVert_{H^{-1}}^{2}

and use this to construct α\alpha for all uH121\left\lVert u\right\rVert^{2}_{H^{-1}}\ll 1. For the (KdV) and fifth-order KdV equations, this small-data assumption does not pose a problem; one can use the scaling symmetry to make the initial data arbitrarily small, and then the exact conservation of α\alpha implies that solutions remain small globally in time.

By comparison, (gKdV) does not posses a scaling symmetry in general, and we expect that the H1H^{-1} norm of solutions can be growing in time without bound. This in turn forces us to choose κuH12\kappa\gg\left\lVert u\right\rVert^{2}_{H^{-1}} large, which introduces a large implicit constant in (1.13). Consequently, the estimates in [Killip2019, Bringmann2021] are insufficient, even in estimating the LHS of (1.12). In order to close our bootstrap argument, we must establish new estimates for the density ρ\rho and current jj using the Hκ1H^{-1}_{\kappa} norm, rather than H1H^{-1}, as this is the quantity whose growth we can efficiently estimate using (1.7). In comparison to the estimate (1.13) used throughout [Killip2019, Bringmann2021], this requires a much more careful estimation of high frequency contributions.

This paper is organized as follows. We begin in Section 2 by reviewing the material developed in [Killip2019, Bringmann2021] that we will need, including the key players α\alpha, ρ\rho, and jj mentioned previously as well as the estimates they satisfy. We then proceed in Section 3 by introducing our local smoothing norm and developing the new estimates required for our bootstrap argument.

We employ these ingredients in Sections 4 and 5 to prove more precise versions of our local smoothing and a-priori estimates, Theorem 4.1 and Proposition 5.1, respectively. We then combine these two estimates with a bootstrap argument, and present the full result in Theorem 5.2. Finally, we conclude Section 5 by describing how the statements of Theorems 1.1 and 1.5 and Corollary 1.4 represent special cases of Theorem 5.2.

1.1. Acknowledgments

While working on this project the first author was supported by the CAREER grant DMS-1845037, by the NSF grant DMS-2348908, by the Miller Foundation and by a Simons Fellowship.

2. Preliminaries

Here we recall some of the key objects from [Killip2019] which are used in order to prove that (KdV) is well-posed in H1H^{-1}, and which will serve us well in obtaining the a-priori bounds for our nonintegrable model (gKdV) displayed in Theorem 1.1 and Theorem 1.5. Our starting point is represented by the self-adjoint Lax operator associated to the KdV flow,

Lu=x2+u.L_{u}=-\partial_{x}^{2}+u.

Formally, the spectrum of LuL_{u} is conserved along the KdV flow. More precisely, Lu(t)L_{u(t)} and Lu(s)L_{u(s)} are unitarily equivalent operators. In general LuL_{u} is not invertible, which is one reason it is convenient to replace it with Lu+κ2L_{u}+\kappa^{2}; given uH1u\in H^{-1} the operator Lu+κ2L_{u}+\kappa^{2} is invertible if κ>0\kappa>0 is large enough. One may construct conserved quantities for KdV by looking at the trace of functions of Lu+κ2L_{u}+\kappa^{2}, and in particular the trace of its renormalized logarithm, which can be also viewed as a function of κ\kappa. As shown in [Killip2019], this trace is closely related to the diagonal of the kernel of Lu+κ2L_{u}+\kappa^{2}.

The first step in the analysis is to construct the diagonal of the integral kernel G(x,y)G(x,y) for the resolvent

(2.1) Ru:=(x2+u+κ2)1R_{u}:=(-\partial^{2}_{x}+u+\kappa^{2})^{-1}

of the Lax operator associated to a state uu. With this operator in mind, it will be convenient to work with the norms

fHκs2:=|f^(ξ)|2(ξ2+4κ2)s𝑑ξ,\left\lVert f\right\rVert_{H^{s}_{\kappa}}^{2}:=\int|\widehat{f}(\xi)|^{2}(\xi^{2}+4\kappa^{2})^{s}\,d\xi,

where our convention for the Fourier transform is

f^(ξ)=12πeiξxf(x)𝑑xso thatf,g=f¯(x)g(x)𝑑x=f^,g^.\widehat{f}(\xi)=\frac{1}{\sqrt{2\pi}}\int e^{-i\xi x}f(x)\,dx\quad\text{so that}\quad\langle f,g\rangle=\int\overline{f}(x)g(x)\,dx=\langle\widehat{f},\widehat{g}\rangle.

In the definition of HκsH^{s}_{\kappa}, the constant 44 is simply included in order to make (2.4) an equality. One should see these spaces as versions of the traditional inhomogeneous Sobolev spaces HsH^{s} but adapted to the frequency scale κ\kappa rather than the unit frequency. Topologically these are the same as the classical Sobolev spaces HsH^{s}, but with the implicit constants in the norm equivalence depending on κ\kappa. In particular, we have the elementary estimates

fgHκ±1fH1gHκ±1andfgHκ±1fW1,gHκ±1\left\lVert fg\right\rVert_{H^{\pm 1}_{\kappa}}\lesssim\left\lVert f\right\rVert_{H^{1}}\left\lVert g\right\rVert_{H^{\pm 1}_{\kappa}}\quad\text{and}\quad\left\lVert fg\right\rVert_{H^{\pm 1}_{\kappa}}\lesssim\left\lVert f\right\rVert_{W^{1,\infty}}\left\lVert g\right\rVert_{H^{\pm 1}_{\kappa}}

uniformly for κ1\kappa\geq 1. Notice that the results for Hκ1H^{-1}_{\kappa} follow from those for Hκ1H^{1}_{\kappa} by duality. We will also frequently use the embedding

fLκ12fHκ1uniformly for κ1,\left\lVert f\right\rVert_{L^{\infty}}\lesssim\kappa^{-\frac{1}{2}}\left\lVert f\right\rVert_{H^{1}_{\kappa}}\quad\text{uniformly for }\kappa\geq 1,

which implies the algebra bound

fgHκ1κ12fHκ1gHκ1uniformly for κ1.\left\lVert fg\right\rVert_{H^{1}_{\kappa}}\lesssim\kappa^{-\frac{1}{2}}\left\lVert f\right\rVert_{H^{1}_{\kappa}}\left\lVert g\right\rVert_{H^{1}_{\kappa}}\quad\text{uniformly for }\kappa\geq 1.

The resolvent R0R_{0} associated to u0u\equiv 0,

(2.2) R0(κ):=(x2+κ2)1,R_{0}(\kappa):=(-\partial^{2}_{x}+\kappa^{2})^{-1},

plays an important role in what follows, and in particular has the mapping property

R0:HκsHκs+2,s.R_{0}:H^{s}_{\kappa}\rightarrow H^{s+2}_{\kappa},\quad s\in\mathbb{R}.

This property is also shared by RuR_{u} for uH1u\in H^{-1} for κ\kappa large enough, but only for restricted range of ss. One can formally use R0R_{0} to obtain an expansion for RuR_{u}:

(2.3) Ru==0(1)(R0u)R0,R_{u}=\sum_{\ell=0}^{\infty}(-1)^{\ell}(R_{0}u)^{\ell}R_{0},

which is justified for uH1u\in H^{-1} for κ\kappa large enough.

Our first bound involving R0R_{0} is as follows:

Lemma 2.1 ([Killip2018]).

For κ>0\kappa>0, we have

(2.4) R0fR02=κ12fHκ1.\big{\lVert}\sqrt{R_{0}}f\sqrt{R_{0}}\big{\rVert}_{\mathfrak{I}_{2}}=\kappa^{-\frac{1}{2}}\left\lVert f\right\rVert_{H^{-1}_{\kappa}}.

Here, 2\mathfrak{I}_{2} denotes the class of Hilbert–Schmidt operators on L2()L^{2}(\mathbb{R}). Such operators are automatically continuous due to the inequality

AA2=tr{AA},\left\lVert A\right\rVert_{\mathfrak{I}_{\infty}}\leq\left\lVert A\right\rVert_{\mathfrak{I}_{2}}=\sqrt{\operatorname{tr}\{A^{*}A\}},

where \mathfrak{I}_{\infty} denotes the operator norm. Moreover, the space of such operators forms a two-sided ideal in the space of bounded operators:

ABC2AB2C.\left\lVert ABC\right\rVert_{\mathfrak{I}_{2}}\leq\left\lVert A\right\rVert_{\mathfrak{I}_{\infty}}\left\lVert B\right\rVert_{\mathfrak{I}_{2}}\left\lVert C\right\rVert_{\mathfrak{I}_{\infty}}.

Lastly, the product of two Hilbert–Schmidt operators is trace class, and we have

|tr(AB)|AB1A2B2.|\operatorname{tr}(AB)|\leq\left\lVert AB\right\rVert_{\mathfrak{I}_{1}}\leq\left\lVert A\right\rVert_{\mathfrak{I}_{2}}\left\lVert B\right\rVert_{\mathfrak{I}_{2}}.

Here, 1\mathfrak{I}_{1} denotes the trace class; this consists of operators whose singular values are 1\ell^{1}-summable. In this way, the second inequality above is simply an application of the Cauchy–Schwarz inequality.

The Green’s function is the integral kernel for the resolvent of the Lax operator RuR_{u}, which may be expressed as

G(x,y)=δx,(x2+u+κ2)1δy.G(x,y)=\langle\delta_{x},(-\partial_{x}^{2}+u+\kappa^{2})^{-1}\delta_{y}\rangle.

As explained earlier, its restriction to the diagonal plays a key role in our analysis.

Proposition 2.2 (Diagonal Green’s function [Killip2019]).

There exists a constant C>0C>0 so that the following statements are true for any R>0R>0:

  1. (1)

    For each uHκ1R\left\lVert u\right\rVert_{H^{-1}_{\kappa}}\leq R, the diagonal Green’s function

    g(x;κ,u):=δx,(x2+u+κ2)1δxg(x;\kappa,u):=\langle\delta_{x},(-\partial^{2}_{x}+u+\kappa^{2})^{-1}\delta_{x}\rangle

    exists for all xx\in\mathbb{R} and κ1+CR2\kappa\geq 1+CR^{2}.

  2. (2)

    The mappings

    (2.5) ug12κandu1g2κu\mapsto g-\tfrac{1}{2\kappa}\quad\textrm{and}\quad u\mapsto\tfrac{1}{g}-2\kappa

    are real-analytic functionals from {u:uH1R}\{u:\left\lVert u\right\rVert_{H^{-1}}\leq R\} into H1H^{1} for all κ1+CR2\kappa\geq 1+CR^{2}.

  3. (3)

    We have the estimates

    (2.6) g(κ,u)12κHκ1\displaystyle\left\lVert g(\kappa,u)-\tfrac{1}{2\kappa}\right\rVert_{H^{1}_{\kappa}} κ1uHκ1,\displaystyle\lesssim\kappa^{-1}\left\lVert u\right\rVert_{H^{-1}_{\kappa}},
    (2.7) 1g(κ,u)2κHκ1\displaystyle\big{\lVert}\tfrac{1}{g(\kappa,u)}-2\kappa\big{\rVert}_{H^{1}_{\kappa}} κuHκ1\displaystyle\lesssim\kappa\left\lVert u\right\rVert_{H^{-1}_{\kappa}}

    uniformly for κ1+CR2\kappa\geq 1+CR^{2}.

As a consequence of the expansion of RuR_{u} in (2.3), for gg we have a similar expansion

(2.8) g=12κ+h1+h2+,h(x)=(1)δx,(R0u)R0δx.g=\tfrac{1}{2\kappa}+h_{1}+h_{2}+\dots,\qquad h_{\ell}(x)=(-1)^{\ell}\langle\delta_{x},(R_{0}u)^{\ell}R_{0}\delta_{x}\rangle.

In particular, using the integral kernel δx,R0(κ)δy=12κeκ|xy|\langle\delta_{x},R_{0}(\kappa)\delta_{y}\rangle=\tfrac{1}{2\kappa}e^{-\kappa|x-y|} for the free resolvent, we find that

h1=κ1R0(2κ)u.h_{1}=-\kappa^{-1}R_{0}(2\kappa)u.

As it turns out, it is the function 1/g1/g that is linearly related to the logarithmic renormalized trace of the resolvent. A consequence of the proposition is that the integral of 1/g1/g diverges. Even after removing the constant, its integral still turns out to diverge for uu in any HsH^{s} space. To rectify this, one also needs to remove the linear term in uu, which formally integrates to a multiple of u𝑑x\int u\,dx; this is a conserved quantity for the KdV flow, and thus harmless. A similar renormalization was also implemented in [KT] for the logarithm of the transmission coefficient. Taking these renormalizations into account, we can define the conserved quantity which controls the H1H^{-1} norm of uu:

(2.9) α(κ,u):=ρ𝑑x,ρ(x;κ,u):=12g(x;κ,u)+κ+2κR0(2κ)u.\alpha(\kappa,u):=\int\rho\,dx,\qquad\rho(x;\kappa,u):=-\frac{1}{2g(x;\kappa,u)}+\kappa+2\kappa R_{0}(2\kappa)u.

The formula for α\alpha is the trace of the integral kernel 1/2G(x,y;κ,u)-1/2G(x,y;\kappa,u) with the first two terms of its Taylor series about u0u\equiv 0 removed.

Proposition 2.3 (Introducing α\alpha [Killip2019]).

There exists a constant C>0C>0 so that the following statements are true for any R>0R>0:

  1. (1)

    The quantities ρ(x;κ,u)\rho(x;\kappa,u) and α(κ,u)\alpha(\kappa,u) defined in (2.9) are finite and nonnegative for all xx\in\mathbb{R}, uHκ1R\left\lVert u\right\rVert_{H^{-1}_{\kappa}}\leq R, and κ1+CR2\kappa\geq 1+CR^{2}.

  2. (2)

    The mapping uαu\mapsto\alpha is a real analytic functional on {u:uHκ1R}\{u:\left\lVert u\right\rVert_{H^{-1}_{\kappa}}\leq R\} for all κ1+CR2\kappa\geq 1+CR^{2}.

  3. (3)

    We have

    (2.10) 14κ1uHκ12α(κ,u)κ1uHκ12\tfrac{1}{4}\kappa^{-1}\left\lVert u\right\rVert_{H^{-1}_{\kappa}}^{2}\leq\alpha(\kappa,u)\leq\kappa^{-1}\left\lVert u\right\rVert_{H^{-1}_{\kappa}}^{2}

    for all uHκ1R\left\lVert u\right\rVert_{H^{-1}_{\kappa}}\leq R and κ1+CR2\kappa\geq 1+CR^{2}.

The authors of [Killip2019] also proved that under the (KdV) flow, we have the following density-flux equation

(2.11) tρ+j=0,\partial_{t}\rho+j^{\prime}=0,

where ρ=ρ(x;κ,u)\rho=\rho(x;\kappa,u) is defined in (2.9) and

(2.12) j(x;κ,u):=1g(4κ3g2κ2+u)+2κR0(2κ)(u′′3u2).j(x;\kappa,u):=\tfrac{1}{g}(4\kappa^{3}g-2\kappa^{2}+u)+2\kappa R_{0}(2\kappa)(u^{\prime\prime}-3u^{2}).

We see in particular by integrating (2.11) in space that α\alpha is conserved under the (KdV) flow.

By comparison, the coefficients aja_{j} in (gKdV) break the density-flux equation (2.11). Instead, we obtain an additional term as follows:

(2.13) tρ+j=dρ|u[(a1u)+a2u2+a3u+a4u],\partial_{t}\rho+j^{\prime}=d\rho|_{u}\big{[}(a_{1}u^{\prime})^{\prime}+a_{2}u^{2}+a_{3}u^{\prime}+a_{4}u\big{]},

where

(2.14) dρ|u(f)=ddsρ(x;κ,u+sf)|s=0=12g2dg|u(f)+2κR0(2κ)fd\rho|_{u}(f)=\frac{d}{ds}\rho(x;\kappa,u+sf)\bigg{|}_{s=0}=\tfrac{1}{2g^{2}}dg|_{u}(f)+2\kappa R_{0}(2\kappa)f

is the functional derivative of ρ\rho at uu, and

(2.15) dg|u(f)=G(x,y)f(y)G(y,x)𝑑y.dg|_{u}(f)=-\int G(x,y)f(y)G(y,x)\,dy.

The identity (2.13) lies at the heart of both of the proofs of our a-priori and local smoothing estimates.

In order to leverage local smoothing, we will frequently need to commute slowly varying weights past R0R_{0}, as in the following estimates.

Lemma 2.4.

If w:(0,)w:\mathbb{R}\to(0,\infty) satisfies

(2.16) |w(x)|+|w′′(x)|w(x)andw(y)w(x)e|xy|/2|w^{\prime}(x)|+|w^{\prime\prime}(x)|\lesssim w(x)\quad\textrm{and}\quad\frac{w(y)}{w(x)}\lesssim e^{|x-y|/2}

uniformly for x,yx,y\in\mathbb{R}, then

(2.17) wR01wLpLpκ2,wR01wLpLpκ1,wR01wHκ1Hκ11\left\lVert wR_{0}\tfrac{1}{w}\right\rVert_{L^{p}\to L^{p}}\lesssim\kappa^{-2},\quad\left\lVert w\partial R_{0}\tfrac{1}{w}\right\rVert_{L^{p}\to L^{p}}\lesssim\kappa^{-1},\quad\left\lVert wR_{0}\tfrac{1}{w}\right\rVert_{H^{-1}_{\kappa}\to H^{1}_{\kappa}}\lesssim 1

uniformly for κ1\kappa\geq 1, for any 1p1\leq p\leq\infty.

The proof is elementary; see, for example, [Bringmann2021]*Lem. 2.7.

We will use this lemma in particular for the functions ψ\psi^{\ell}, =1,2,3\ell=1,2,3, where

ψ(x)=sech(x6).\psi(x)=\operatorname{sech}(\tfrac{x}{6}).

The constant 6 is simply included to ensure that ψ\psi^{\ell} satisfies (2.16) for =1,2,3\ell=1,2,3. (In turn, the constant 2 in (2.16) is based on the integral kernel δx,R0δy=12κeκ|xy|\langle\delta_{x},R_{0}\delta_{y}\rangle=\tfrac{1}{2\kappa}e^{-\kappa|x-y|} and the condition κ1\kappa\geq 1.)

Over the course of our analysis, we will also encounter the ‘double commutator’

R0ψ2R0ψR02ψ=[[R0,ψ],R0ψ].R_{0}\psi^{2}R_{0}-\psi R_{0}^{2}\psi=[[R_{0},\psi],R_{0}\psi].

In order to bound this operator, the naive estimate (2.17) unfortunately does not yield enough decay as κ\kappa\to\infty. To this end, we record the following operator estimates, which capture the double commutator structure:

Lemma 2.5.

We have

(2.18) 1ψ(R0ψ2R0ψR02ψ)1ψHκ2Hκ2κ2,\displaystyle\big{\lVert}\tfrac{1}{\psi}(R_{0}\psi^{2}R_{0}-\psi R_{0}^{2}\psi)\tfrac{1}{\psi}\big{\rVert}_{H^{-2}_{\kappa}\to H^{2}_{\kappa}}\lesssim\kappa^{-2},
(2.19) 1ψ(R0ψ2R0ψR022ψ)1ψHκ2Hκ21\displaystyle\big{\lVert}\tfrac{1}{\psi}(R_{0}\partial\psi^{2}R_{0}\partial-\psi R_{0}^{2}\partial^{2}\psi)\tfrac{1}{\psi}\big{\rVert}_{H^{-2}_{\kappa}\to H^{2}_{\kappa}}\lesssim 1

uniformly for κ1\kappa\geq 1.

In comparison to (2.17), the proofs of (2.18) and (2.19) exhibit extra cancellation by computing the double commutator in a symmetric way. For example, (2.18) follows from [Bringmann2021]*Lem. 2.10, and (2.19) can be verified using a similar argument.

3. The local smoothing norm

Here, we will record some useful estimates for the following local smoothing norm:

(3.1) uLSκ=supx0ψx0uLt2Hκ1([T,T]×),\left\lVert u\right\rVert_{\textup{LS}_{\kappa}}=\sup_{x_{0}\in\mathbb{R}}\big{\lVert}\psi_{x_{0}}u^{\prime}\big{\rVert}_{L^{2}_{t}H^{-1}_{\kappa}([-T,T]\times\mathbb{R})},

where T>0T>0 and

ψ(x)=sech(x6),ψx0(x)=ψ(xx0).\psi(x)=\operatorname{sech}(\tfrac{x}{6}),\qquad\psi_{x_{0}}(x)=\psi(x-x_{0}).

We begin with the following elementary observations:

Lemma 3.1.

Given a Schwartz function ϕ(x)\phi(x), we have

(3.2) ϕuLt2Hκ2\displaystyle\left\lVert\phi u\right\rVert_{L^{2}_{t}H^{-2}_{\kappa}} κ1uLt2Hκ1,\displaystyle\lesssim\kappa^{-1}\left\lVert u\right\rVert_{L^{2}_{t}H^{-1}_{\kappa}},
(3.3) (ϕu)Lt2Hκ2+ϕuLt2Hκ2\displaystyle\left\lVert(\phi u)^{\prime}\right\rVert_{L^{2}_{t}H^{-2}_{\kappa}}+\left\lVert\phi u^{\prime}\right\rVert_{L^{2}_{t}H^{-2}_{\kappa}} κ1(uLSκ+uLt2Hκ1),\displaystyle\lesssim\kappa^{-1}\big{(}\left\lVert u\right\rVert_{\textup{LS}_{\kappa}}+\left\lVert u\right\rVert_{L^{2}_{t}H^{-1}_{\kappa}}\big{)},
(3.4) (ϕu)′′Lt2Hκ2+ϕu′′Lt2Hκ2\displaystyle\left\lVert(\phi u)^{\prime\prime}\right\rVert_{L^{2}_{t}H^{-2}_{\kappa}}+\left\lVert\phi u^{\prime\prime}\right\rVert_{L^{2}_{t}H^{-2}_{\kappa}} uLSκ+κ1uLt2Hκ1,\displaystyle\lesssim\left\lVert u\right\rVert_{\textup{LS}_{\kappa}}+\kappa^{-1}\left\lVert u\right\rVert_{L^{2}_{t}H^{-1}_{\kappa}},

uniformly for κ1\kappa\geq 1 and T>0T>0. (All space-time norms are taken over [T,T]×[-T,T]\times\mathbb{R}.)

Proof.

For the estimate (3.2) we simply write

ϕuLt2Hκ2uLt2Hκ2κ1uLt2Hκ1.\|\phi u\|_{L^{2}_{t}H^{-2}_{\kappa}}\lesssim\|u\|_{L^{2}_{t}H^{-2}_{\kappa}}\lesssim\kappa^{-1}\left\lVert u\right\rVert_{L^{2}_{t}H^{-1}_{\kappa}}.

For the estimate (3.3) we use Leibnitz rule (ϕu)=ϕu+ϕu(\phi u)^{\prime}=\phi^{\prime}u+\phi u^{\prime}. For the first term we use (3.2). For the second we write

ψz2(x)𝑑z=c\int\psi_{z}^{2}(x)\,dz=c

for some constant cc, so that

ϕ(x)=1cϕ(x)ψz2(x)𝑑z.\phi(x)=\tfrac{1}{c}\int\phi(x)\psi_{z}^{2}(x)\,dz.

Then, it follows that

ϕuHκ11cϕψz2uHκ1𝑑zϕψzH4ψzuHκ1𝑑z.\left\lVert\phi u^{\prime}\right\rVert_{H^{-1}_{\kappa}}\leq\tfrac{1}{c}\int\left\lVert\phi\psi_{z}^{2}u^{\prime}\right\rVert_{H^{-1}_{\kappa}}\,\,dz\lesssim\int\left\lVert\phi\psi_{z}\right\rVert_{H^{4}}\|\psi_{z}u^{\prime}\|_{H^{-1}_{\kappa}}\,\,dz.

Integrating in zz and noting that

ϕψzH4𝑑zϕ1,\int\left\lVert\phi\psi_{z}\right\rVert_{H^{4}}\,dz\lesssim_{\phi}1,

we obtain

(3.5) ϕuHκ1uLSκ,\|\phi u^{\prime}\|_{H^{-1}_{\kappa}}\lesssim\|u\|_{\textup{LS}_{\kappa}},

and the estimate (3.3) follows, as

ϕuHκ2κ1ϕuHκ1.\|\phi u^{\prime}\|_{H^{-2}_{\kappa}}\lesssim\kappa^{-1}\|\phi u^{\prime}\|_{H^{-1}_{\kappa}}.

For the estimate (3.4) we simply write

(ϕu)′′=(ϕu)+ϕu+ϕ′′u,(\phi u)^{\prime\prime}=(\phi u^{\prime})^{\prime}+\phi^{\prime}u^{\prime}+\phi^{\prime\prime}u,

and we use twice (3.5) and once (3.2). ∎

The operator estimate (3.7) below will play an important role in our analysis. In comparison to (2.4), we now bound the operator norm (rather than Hilbert–Schmidt) and only require a local norm of uu to do so.

Lemma 3.2.

We have

(3.6) fuLt2Hκ1κ12fHκ1uLSκfor all fH1(),\left\lVert fu^{\prime}\right\rVert_{L^{2}_{t}H^{-1}_{\kappa}}\lesssim\kappa^{-\frac{1}{2}}\left\lVert f\right\rVert_{H^{1}_{\kappa}}\left\lVert u\right\rVert_{\textup{LS}_{\kappa}}\quad\textrm{for all }f\in H^{1}(\mathbb{R}),

uniformly for κ1\kappa\geq 1. As a consequence,

(3.7) R0uR0Lt2κ12uLSκ.\big{\lVert}\sqrt{R_{0}}u^{\prime}\sqrt{R_{0}}\big{\rVert}_{L^{2}_{t}\mathfrak{I}_{\infty}}\lesssim\kappa^{-\frac{1}{2}}\left\lVert u\right\rVert_{\textup{LS}_{\kappa}}.
Proof.

We argue by duality. Let hLt2Hκ1h\in L^{2}_{t}H^{1}_{\kappa}. Fix a smooth partition of unity

jχj21,withχj1 on [j,j+1]andsuppχj[j1,j+2].\sum_{j\in\mathbb{Z}}\chi_{j}^{2}\equiv 1,\quad\textrm{with}\quad\chi_{j}\equiv 1\textrm{ on }[j,j+1]\quad\textrm{and}\quad\operatorname{supp}\chi_{j}\subset[j-1,j+2].

Then

h,fu=jχ~j2uχj2fh¯𝑑x,\langle h,fu^{\prime}\rangle=\sum_{j\in\mathbb{Z}}\int\widetilde{\chi}_{j}^{2}u^{\prime}\cdot\chi_{j}^{2}f\overline{h}\,dx,

where χ~j2=χj12+χj2+χj+12\widetilde{\chi}_{j}^{2}=\chi_{j-1}^{2}+\chi_{j}^{2}+\chi_{j+1}^{2} is a fattened bump function. We estimate

TT|h,fu|𝑑t\displaystyle\int_{-T}^{T}|\langle h,fu^{\prime}\rangle|\,dt jχ~j2uLt2Hκ1χj2fh¯Lt2Hκ1\displaystyle\lesssim\sum_{j}\lVert\widetilde{\chi}_{j}^{2}u^{\prime}\rVert_{L^{2}_{t}H^{-1}_{\kappa}}\left\lVert\chi_{j}^{2}f\overline{h}\right\rVert_{L^{2}_{t}H^{1}_{\kappa}}
κ12juLSκχjfHκ1χjhLt2Hκ1\displaystyle\lesssim\kappa^{-\frac{1}{2}}\sum_{j}\left\lVert u\right\rVert_{\textup{LS}_{\kappa}}\left\lVert\chi_{j}f\right\rVert_{H^{1}_{\kappa}}\left\lVert\chi_{j}h\right\rVert_{L^{2}_{t}H^{1}_{\kappa}}
κ12uLSκ(jχjfHκ12)12(jχjhLt2Hκ12)12\displaystyle\lesssim\kappa^{-\frac{1}{2}}\left\lVert u\right\rVert_{\textup{LS}_{\kappa}}\bigg{(}\sum_{j}\left\lVert\chi_{j}f\right\rVert_{H^{1}_{\kappa}}^{2}\bigg{)}^{\frac{1}{2}}\bigg{(}\sum_{j}\left\lVert\chi_{j}h\right\rVert_{L^{2}_{t}H^{1}_{\kappa}}^{2}\bigg{)}^{\frac{1}{2}}
κ12uLSκfHκ1hLt2Hκ1.\displaystyle\lesssim\kappa^{-\frac{1}{2}}\left\lVert u\right\rVert_{\textup{LS}_{\kappa}}\left\lVert f\right\rVert_{H^{1}_{\kappa}}\left\lVert h\right\rVert_{L^{2}_{t}H^{1}_{\kappa}}.

Taking a supremum over hLt2Hκ11\left\lVert h\right\rVert_{L^{2}_{t}H^{1}_{\kappa}}\leq 1, the estimate (3.6) follows.

The second estimate (3.7) now follows immediately:

R0uR0Lt2κ12R0Hκ1L2uLSκR0L2Hκ1κ12uLSκ.\big{\lVert}\sqrt{R_{0}}u^{\prime}\sqrt{R_{0}}\big{\rVert}_{L^{2}_{t}\mathfrak{I}_{\infty}}\lesssim\kappa^{-\frac{1}{2}}\big{\lVert}\sqrt{R_{0}}\big{\rVert}_{H^{-1}_{\kappa}\to L^{2}}\left\lVert u\right\rVert_{\textup{LS}_{\kappa}}\big{\lVert}\sqrt{R_{0}}\big{\rVert}_{L^{2}\to H^{1}_{\kappa}}\lesssim\kappa^{-\frac{1}{2}}\left\lVert u\right\rVert_{\textup{LS}_{\kappa}}.\qed

Our next lemma provides an estimate for the diagonal Green’s function analogous to (2.6) that captures the gain of regularity we expect from our local smoothing norm.

Lemma 3.3.

There exists a constant C>0C>0 so that for any Schwartz function ϕ(x)\phi(x), we have

(3.8) ϕgLt2Hκ1ϕκ1uLSκ\left\lVert\phi g^{\prime}\right\rVert_{L^{2}_{t}H^{1}_{\kappa}}\lesssim_{\phi}\kappa^{-1}\left\lVert u\right\rVert_{\textup{LS}_{\kappa}}

uniformly for

(3.9) κ1+CuLtHκ12.\kappa\geq 1+C\left\lVert u\right\rVert_{L^{\infty}_{t}H^{-1}_{\kappa}}^{2}.
Proof.

Using a ‘continuous partition of unity’ as in the proof of (3.3), it suffices to prove (3.8) in the special case where ϕ=ψ\phi=\psi. We argue by duality. For fLt2Hκ1f\in L^{2}_{t}H^{-1}_{\kappa}, we can use the expansion we have for gg in (2.8) to write

TT|fψg𝑑x|𝑑t1TT|tr{fψ[,(R0u)R0]}|𝑑t.\int_{-T}^{T}\bigg{|}\int f\psi g^{\prime}\,dx\bigg{|}\,dt\leq\sum_{\ell\geq 1}\int_{-T}^{T}\big{|}\operatorname{tr}\big{\{}f\psi[\partial,(R_{0}u)^{\ell}R_{0}]\big{\}}\big{|}\,dt.

When =1\ell=1, there is only one copy of uu on which the derivative \partial may fall. In this case, we write

fψR0u=f[ψR01ψ]ψuf\psi R_{0}u^{\prime}=f[\psi R_{0}\tfrac{1}{\psi}]\psi u^{\prime}

and note that the operator in square brackets is bounded Hκ1Hκ1H^{-1}_{\kappa}\to H^{1}_{\kappa} by (2.17). Therefore

|tr{fψR0uR0}|R0fR02R0ψuR02κ1fHκ1ψuHκ1,\big{|}\operatorname{tr}\big{\{}f\psi R_{0}u^{\prime}R_{0}\big{\}}\big{|}\lesssim\big{\lVert}\sqrt{R_{0}}f\sqrt{R_{0}}\big{\rVert}_{\mathfrak{I}_{2}}\big{\lVert}\sqrt{R_{0}}\psi u^{\prime}\sqrt{R_{0}}\big{\rVert}_{\mathfrak{I}_{2}}\lesssim\kappa^{-1}\left\lVert f\right\rVert_{H^{-1}_{\kappa}}\left\lVert\psi u^{\prime}\right\rVert_{H^{-1}_{\kappa}},

and thus

TT|tr{fψR0uR0}|𝑑tκ1fLt2Hκ1uLSκ.\int_{-T}^{T}\big{|}\operatorname{tr}\big{\{}f\psi R_{0}u^{\prime}R_{0}\big{\}}\big{|}\,dt\lesssim\kappa^{-1}\left\lVert f\right\rVert_{L^{2}_{t}H^{-1}_{\kappa}}\left\lVert u\right\rVert_{\textup{LS}_{\kappa}}.

Next, we turn to the case where 2\ell\geq 2. We distribute the derivative \partial using the product rule [,AB]=[,A]B+A[,B][\partial,AB]=[\partial,A]B+A[\partial,B]. For the one factor of uu^{\prime}, we estimate its contribution in operator norm using (3.7). The requirement 2\ell\geq 2 then guarantees that there is at least one other factor of uu, which together with ff provides us with two terms we can put in Hilbert–Schmidt norm:

2TT|tr{fψ[,(R0u)R0]}|𝑑t\displaystyle\sum_{\ell\geq 2}\int_{-T}^{T}\big{|}\operatorname{tr}\big{\{}f\psi[\partial,(R_{0}u)^{\ell}R_{0}]\big{\}}\big{|}\,dt
2R0fψR0Lt22R0uR0Lt2R0uR0Lt21\displaystyle\qquad\leq\sum_{\ell\geq 2}\ell\big{\lVert}\sqrt{R_{0}}f\psi\sqrt{R_{0}}\big{\rVert}_{L^{2}_{t}\mathfrak{I}_{2}}\big{\lVert}\sqrt{R_{0}}u^{\prime}\sqrt{R_{0}}\big{\rVert}_{L^{2}_{t}\mathfrak{I}_{\infty}}\big{\lVert}\sqrt{R_{0}}u\sqrt{R_{0}}\big{\rVert}_{L^{\infty}_{t}\mathfrak{I}_{2}}^{\ell-1}
2κ1fLt2Hκ1uLSκ(κ12uLtHκ1)1\displaystyle\qquad\lesssim\sum_{\ell\geq 2}\ell\kappa^{-1}\left\lVert f\right\rVert_{L^{2}_{t}H^{-1}_{\kappa}}\left\lVert u\right\rVert_{\textup{LS}_{\kappa}}\big{(}\kappa^{-\frac{1}{2}}\left\lVert u\right\rVert_{L^{\infty}_{t}H^{-1}_{\kappa}}\big{)}^{\ell-1}
κ1fLt2Hκ1uLSκ.\displaystyle\qquad\lesssim\kappa^{-1}\left\lVert f\right\rVert_{L^{2}_{t}H^{-1}_{\kappa}}\left\lVert u\right\rVert_{\textup{LS}_{\kappa}}.

Altogether, we conclude

TT|fψg𝑑x|𝑑tκ1fLt2Hκ1uLSκ.\int_{-T}^{T}\bigg{|}\int f\psi g^{\prime}\,dx\bigg{|}\,dt\lesssim\kappa^{-1}\left\lVert f\right\rVert_{L^{2}_{t}H^{-1}_{\kappa}}\left\lVert u\right\rVert_{\textup{LS}_{\kappa}}.

Taking a supremum over fLt2Hκ11\left\lVert f\right\rVert_{L^{2}_{t}H^{-1}_{\kappa}}\leq 1, the estimate (3.8) follows. ∎

The next few results provide estimates for the terms hh_{\ell} appearing in the series (2.8) for gg. We will start with h1h_{1}, which follows easily from (3.2)–(3.4).

Lemma 3.4.

We have

(3.10) ψh1Lt,x2\displaystyle\left\lVert\psi h_{1}\right\rVert_{L^{2}_{t,x}} κ2uLt2Hκ1,\displaystyle\lesssim\kappa^{-2}\left\lVert u\right\rVert_{L^{2}_{t}H^{-1}_{\kappa}},
(3.11) ψh1Lt,x2\displaystyle\left\lVert\psi h^{\prime}_{1}\right\rVert_{L^{2}_{t,x}} κ2uLSκ,\displaystyle\lesssim\kappa^{-2}\left\lVert u\right\rVert_{\textup{LS}_{\kappa}},
(3.12) ψh1′′Lt,x2\displaystyle\left\lVert\psi h^{\prime\prime}_{1}\right\rVert_{L^{2}_{t,x}} κ1uLSκ+κ2uLt2Hκ1\displaystyle\lesssim\kappa^{-1}\left\lVert u\right\rVert_{\textup{LS}_{\kappa}}+\kappa^{-2}\left\lVert u\right\rVert_{L^{2}_{t}H^{-1}_{\kappa}}

uniformly for κ1\kappa\geq 1 and T>0T>0.

Proof.

Recall that h1=κ1R0(2κ)uh_{1}=-\kappa^{-1}R_{0}(2\kappa)u. Using (2.17), we see that

ψh1Lt,x2κ1ψh1Lt2Hκ1=κ2ψR0(2κ)1ψψuLt2Hκ1κ2uLSκ.\left\lVert\psi h^{\prime}_{1}\right\rVert_{L^{2}_{t,x}}\lesssim\kappa^{-1}\left\lVert\psi h^{\prime}_{1}\right\rVert_{L^{2}_{t}H^{1}_{\kappa}}=\kappa^{-2}\big{\lVert}\psi R_{0}(2\kappa)\tfrac{1}{\psi}\cdot\psi u^{\prime}\big{\rVert}_{L^{2}_{t}H^{1}_{\kappa}}\lesssim\kappa^{-2}\left\lVert u\right\rVert_{\textup{LS}_{\kappa}}.

This proves (3.11). For (3.10) and (3.12), we also use (3.4) and (3.2):

ψh1Lt,x2κ1ψuLt2Hκ2κ2uLt2Hκ1,\displaystyle\left\lVert\psi h_{1}\right\rVert_{L^{2}_{t,x}}\lesssim\kappa^{-1}\left\lVert\psi u\right\rVert_{L^{2}_{t}H^{-2}_{\kappa}}\lesssim\kappa^{-2}\left\lVert u\right\rVert_{L^{2}_{t}H^{-1}_{\kappa}},
ψh1′′Lt,x2κ1ψu′′Lt2Hκ2κ1(uLSκ+κ1uLt2Hκ1).\displaystyle\left\lVert\psi h^{\prime\prime}_{1}\right\rVert_{L^{2}_{t,x}}\lesssim\kappa^{-1}\left\lVert\psi u^{\prime\prime}\right\rVert_{L^{2}_{t}H^{-2}_{\kappa}}\lesssim\kappa^{-1}\big{(}\left\lVert u\right\rVert_{\textup{LS}_{\kappa}}+\kappa^{-1}\left\lVert u\right\rVert_{L^{2}_{t}H^{-1}_{\kappa}}\big{)}.\qed

The following lemma will be useful in estimating all of the cubic and higher order terms of jj (see (4.16) for details). The proof is rather involved, because we need to be efficient in our estimation. Specifically, we need to get a factor on the RHS of (3.13) that is o(κ5)o(\kappa^{-5}). (Otherwise, we would not be able to start at =3\ell=3 in (4.16).)

Lemma 3.5.

Given a Schwartz function ϕ(x)\phi(x), we have

(3.13) 3ϕhLt,x1(Tκ2)34κ112uLtHκ1(uLtHκ12+uLSκ2)\sum_{\ell\geq 3}\left\lVert\phi h_{\ell}\right\rVert_{L^{1}_{t,x}}\lesssim(T\kappa^{2})^{\frac{3}{4}}\kappa^{-\frac{11}{2}}\left\lVert u\right\rVert_{L^{\infty}_{t}H^{-1}_{\kappa}}\big{(}\left\lVert u\right\rVert_{L^{\infty}_{t}H^{-1}_{\kappa}}^{2}+\left\lVert u\right\rVert_{\textup{LS}_{\kappa}}^{2}\big{)}

uniformly for Tκ2T\leq\kappa^{-2} and κ\kappa satisfying (3.9).

Proof.

Let us begin with (3.13) for the case =3\ell=3. We argue by duality: for fL([T,T]×)f\in L^{\infty}([-T,T]\times\mathbb{R}), we have

(3.14) TTfϕh𝑑x𝑑t=(1)TTtr{R0fϕ(R0u)R0}𝑑t.\int_{-T}^{T}\int_{-\infty}^{\infty}f\phi h_{\ell}\,dx\,dt=(-1)^{\ell}\int_{-T}^{T}\operatorname{tr}\big{\{}\sqrt{R_{0}}f\phi(R_{0}u)^{\ell}\sqrt{R_{0}}\big{\}}\,dt.

Using a continuous partition of unity argument as in the proof of (3.4), it suffices to prove (3.13) in the special case when ϕ=ψ3\phi=\psi^{3}. Write

R0fψ3(R0u)3R0=R0f[ψ3R01ψ3]ψu[ψ2R01ψ2]ψu[ψR01ψ]ψuR0.\sqrt{R_{0}}f\psi^{3}(R_{0}u)^{3}\sqrt{R_{0}}=\sqrt{R_{0}}f[\psi^{3}R_{0}\tfrac{1}{\psi^{3}}]\psi u[\psi^{2}R_{0}\tfrac{1}{\psi^{2}}]\psi u[\psi R_{0}\tfrac{1}{\psi}]\psi u\sqrt{R_{0}}.

By (2.17), each operator in square brackets above is bounded Hκ1Hκ1H^{-1}_{\kappa}\to H^{1}_{\kappa}. This allows us to write

R0fψ3(R0u)3R0=(R0fR0)A3(R0ψuR0)A2(R0ψuR0)A1(R0ψuR0)\sqrt{R_{0}}f\psi^{3}(R_{0}u)^{3}\sqrt{R_{0}}=(\sqrt{R_{0}}f\sqrt{R_{0}})A_{3}(\sqrt{R_{0}}\psi u\sqrt{R_{0}})A_{2}(\sqrt{R_{0}}\psi u\sqrt{R_{0}})A_{1}(\sqrt{R_{0}}\psi u\sqrt{R_{0}})

for operators AjA_{j} with Aj1\left\lVert A_{j}\right\rVert_{\mathfrak{I}_{\infty}}\lesssim 1. We seek to estimate the 1\mathfrak{I}_{1} norm of the product, for which it suffices to bound two of the factors in round brackets in \mathfrak{I}_{\infty} and two in 2\mathfrak{I}_{2}. The factor containing ff is directly estimated in \mathfrak{I}_{\infty} by

(3.15) R0fR0κ2fL.\|\sqrt{R_{0}}f\sqrt{R_{0}}\|_{\mathfrak{I}_{\infty}}\lesssim\kappa^{-2}\|f\|_{L^{\infty}}.

It remains to consider the three remaining factors. To estimate them, we will use two bounds. On one hand for the Hilbert–Schmidt norm we use (2.4) which gives

(3.16) R0wR02=κ12wHκ1.\big{\lVert}\sqrt{R_{0}}w\sqrt{R_{0}}\big{\rVert}_{\mathfrak{I}_{2}}=\kappa^{-\frac{1}{2}}\lVert w\rVert_{H^{-1}_{\kappa}}.

On the other hand for the operator norm we combine (3.15) for the low frequencies with (3.16) for the high frequencies and then use an interpolation inequality,

R0wR0κ12w>κHκ1+κ2w<κLκ12w>κHκ1+κ2w<κL212w<κL212,\big{\lVert}\sqrt{R_{0}}w\sqrt{R_{0}}\big{\rVert}_{\mathfrak{I}_{\infty}}\lesssim\kappa^{-\frac{1}{2}}\lVert w_{>\kappa}\rVert_{H^{-1}_{\kappa}}+\kappa^{-2}\|w_{<\kappa}\|_{L^{\infty}}\lesssim\kappa^{-\frac{1}{2}}\lVert w_{>\kappa}\rVert_{H^{-1}_{\kappa}}+\kappa^{-2}\|w_{<\kappa}\|_{L^{2}}^{\frac{1}{2}}\|w^{\prime}_{<\kappa}\|^{\frac{1}{2}}_{L^{2}},

which we can shorten to

(3.17) R0wR0κ12wHκ112wHκ212.\big{\lVert}\sqrt{R_{0}}w\sqrt{R_{0}}\big{\rVert}_{\mathfrak{I}_{\infty}}\lesssim\kappa^{-\frac{1}{2}}\lVert w\rVert_{H^{-1}_{\kappa}}^{\frac{1}{2}}\lVert w^{\prime}\rVert_{H^{-2}_{\kappa}}^{\frac{1}{2}}.

Applying these bounds to h=ψuh=\psi u, we obtain

R0fψ3(R0u)3R01\displaystyle\big{\lVert}\sqrt{R_{0}}f\psi^{3}(R_{0}u)^{3}\sqrt{R_{0}}\big{\rVert}_{\mathfrak{I}_{1}}\lesssim κ232hHκ152hHκ212fL\displaystyle\ \kappa^{-2-\frac{3}{2}}\lVert h\rVert_{H^{-1}_{\kappa}}^{\frac{5}{2}}\lVert h^{\prime}\rVert_{H^{-2}_{\kappa}}^{\frac{1}{2}}\|f\|_{L^{\infty}}
\displaystyle\lesssim κ232hHκ1(c3hHκ22+c1hHκ12)fL\displaystyle\ \kappa^{-2-\frac{3}{2}}\lVert h\rVert_{H^{-1}_{\kappa}}\big{(}c^{3}\lVert h^{\prime}\rVert_{H^{-2}_{\kappa}}^{2}+c^{-1}\lVert h\rVert_{H^{-1}_{\kappa}}^{2}\big{)}\|f\|_{L^{\infty}}

with an arbitrary positive constant cc. Now we integrate in time,

R0fψ3(R0u)3R0Lt11κ232hLtHκ1(c3hLt2Hκ22+c1hLt2Hκ12)fLt,x.\big{\lVert}\sqrt{R_{0}}f\psi^{3}(R_{0}u)^{3}\sqrt{R_{0}}\big{\rVert}_{L^{1}_{t}\mathfrak{I}_{1}}\lesssim\kappa^{-2-\frac{3}{2}}\lVert h\rVert_{L^{\infty}_{t}H^{-1}_{\kappa}}\big{(}c^{3}\lVert h^{\prime}\rVert_{L^{2}_{t}H^{-2}_{\kappa}}^{2}+c^{-1}\lVert h\rVert_{L^{2}_{t}H^{-1}_{\kappa}}^{2}\big{)}\|f\|_{L^{\infty}_{t,x}}.

Now we use (3.3),

R0fψ3(R0u)3R0Lt11\displaystyle\big{\lVert}\sqrt{R_{0}}f\psi^{3}(R_{0}u)^{3}\sqrt{R_{0}}\big{\rVert}_{L^{1}_{t}\mathfrak{I}_{1}}
κ232uLtHκ1(c3κ2uLSκ2+(c1+c3κ2)uLt2Hκ12)fLt,x\displaystyle\qquad\lesssim\kappa^{-2-\frac{3}{2}}\lVert u\rVert_{L^{\infty}_{t}H^{-1}_{\kappa}}\big{(}c^{3}\kappa^{-2}\lVert u\rVert_{\textup{LS}_{\kappa}}^{2}+(c^{-1}+c^{3}\kappa^{-2})\lVert u\rVert_{L^{2}_{t}H^{-1}_{\kappa}}^{2}\big{)}\|f\|_{L^{\infty}_{t,x}}
κ432uLtHκ1(c3uLSκ2+T(c1κ2+c3)uLtHκ12)fLt,x.\displaystyle\qquad\lesssim\kappa^{-4-\frac{3}{2}}\lVert u\rVert_{L^{\infty}_{t}H^{-1}_{\kappa}}\big{(}c^{3}\lVert u\rVert_{\textup{LS}_{\kappa}}^{2}+T(c^{-1}\kappa^{2}+c^{3})\lVert u\rVert_{L^{\infty}_{t}H^{-1}_{\kappa}}^{2}\big{)}\|f\|_{L^{\infty}_{t,x}}.

Finally we make a good choice for cc in order to balance the constants,

(3.18) c=T14κ12.c=T^{\frac{1}{4}}\kappa^{\frac{1}{2}}.

This gives

(3.19) R0fψ3(R0u)3R0Lt11\displaystyle\big{\lVert}\sqrt{R_{0}}f\psi^{3}(R_{0}u)^{3}\sqrt{R_{0}}\big{\rVert}_{L^{1}_{t}\mathfrak{I}_{1}}\lesssim κ112(Tκ2)34uLtHκ1(uLtHκ12+uLSκ2)fLt,x,\displaystyle\ \kappa^{-\frac{11}{2}}(T\kappa^{2})^{\frac{3}{4}}\lVert u\rVert_{L^{\infty}_{t}H^{-1}_{\kappa}}\big{(}\left\lVert u\right\rVert_{L^{\infty}_{t}H^{-1}_{\kappa}}^{2}+\left\lVert u\right\rVert_{\textup{LS}_{\kappa}}^{2}\big{)}\|f\|_{L^{\infty}_{t,x}},

which suffices for the case =3\ell=3.


Next, we turn to the terms with 4\ell\geq 4. Using (3.19), we estimate

R0fψ3(R0u)R0Lt11\displaystyle\big{\lVert}\sqrt{R_{0}}f\psi^{3}(R_{0}u)^{\ell}\sqrt{R_{0}}\big{\rVert}_{L^{1}_{t}\mathfrak{I}_{1}}
R0fψ3(R0u)3R0Lt11R0uR0Lt23\displaystyle\qquad\leq\big{\lVert}\sqrt{R_{0}}f\psi^{3}(R_{0}u)^{3}\sqrt{R_{0}}\big{\rVert}_{L^{1}_{t}\mathfrak{I}_{1}}\big{\lVert}\sqrt{R_{0}}u\sqrt{R_{0}}\big{\rVert}_{L^{\infty}_{t}\mathfrak{I}_{2}}^{\ell-3}
κ112(Tκ2)34fLt,xuLtHκ1(uLtHκ12+uLSκ2)(κ12uLtHκ1)3.\displaystyle\qquad\lesssim\kappa^{-\frac{11}{2}}(T\kappa^{2})^{\frac{3}{4}}\left\lVert f\right\rVert_{L^{\infty}_{t,x}}\left\lVert u\right\rVert_{L^{\infty}_{t}H^{-1}_{\kappa}}\big{(}\left\lVert u\right\rVert_{L^{\infty}_{t}H^{-1}_{\kappa}}^{2}+\left\lVert u\right\rVert_{\textup{LS}_{\kappa}}^{2}\big{)}\big{(}\kappa^{-\frac{1}{2}}\left\lVert u\right\rVert_{L^{\infty}_{t}H^{-1}_{\kappa}}\big{)}^{\ell-3}.

Choosing CC larger if necessary, the condition (3.9) implies

κ12uLtHκ112.\kappa^{-\frac{1}{2}}\left\lVert u\right\rVert_{L^{\infty}_{t}H^{-1}_{\kappa}}\leq\tfrac{1}{2}.

Summing over 3\ell\geq 3 this relation will make our geometric sum convergent. We now can take the supremum over fLt,x1\left\lVert f\right\rVert_{L^{\infty}_{t,x}}\leq 1 and conclude the proof of (3.13). ∎

Using a similar argument, we also obtain estimates in Lt2Hκ1L^{2}_{t}H^{1}_{\kappa} (instead of Lt,x1L^{1}_{t,x}) using only one copy of the local smoothing norm:

Lemma 3.6.

Given a Schwartz function ϕ(x)\phi(x), we have

(3.20) ϕ(g12κh1)Lt2Hκ1\displaystyle\big{\lVert}\phi(g-\tfrac{1}{2\kappa}-h_{1})\big{\rVert}_{L^{2}_{t}H^{1}_{\kappa}} 2ϕhLt2Hκ1\displaystyle\leq\sum_{\ell\geq 2}\left\lVert\phi h_{\ell}\right\rVert_{L^{2}_{t}H^{1}_{\kappa}}
κ52(Tκ2)14uLtHκ1(uLtHκ1+uLSκ)\displaystyle\lesssim\kappa^{-\frac{5}{2}}(T\kappa^{2})^{\frac{1}{4}}\left\lVert u\right\rVert_{L^{\infty}_{t}H^{-1}_{\kappa}}\big{(}\left\lVert u\right\rVert_{L^{\infty}_{t}H^{-1}_{\kappa}}+\left\lVert u\right\rVert_{\textup{LS}_{\kappa}}\big{)}

uniformly for Tκ2T\leq\kappa^{-2} and κ\kappa satisfying (3.9).

Proof.

Using a continuous partition of unity argument as in the proof of (3.4), it suffices to prove (3.20) in the special case ϕ=ψ2\phi=\psi^{2}. We use the expansion (2.8) for gg, which shows that

g12κh1=2h.g-\tfrac{1}{2\kappa}-h_{1}=\sum_{\ell\geq 2}h_{\ell}.

We argue by duality. For fLt2Hκ1([T,T]×)f\in L^{2}_{t}H^{-1}_{\kappa}([-T,T]\times\mathbb{R}), we have

|ϕ(g12κh1),f|2R0fψ2(R0u)2R0Lt11.|\langle\phi(g-\tfrac{1}{2\kappa}-h_{1}),f\rangle|\lesssim\sum_{\ell\geq 2}\|\sqrt{R_{0}}f\psi^{2}(R_{0}u)^{2}\sqrt{R_{0}}\|_{L^{1}_{t}\mathfrak{I}_{1}}.

It remains to estimate the terms on the right. We begin with the case =2\ell=2, where we write

R0fψ2(R0u)2R0=R0f[ψ2R01ψ2]ψu[ψR01ψ]ψuR0.\sqrt{R_{0}}f\psi^{2}(R_{0}u)^{2}\sqrt{R_{0}}=\sqrt{R_{0}}f[\psi^{2}R_{0}\tfrac{1}{\psi^{2}}]\psi u[\psi R_{0}\tfrac{1}{\psi}]\psi u\sqrt{R_{0}}.

By (2.17), each operator in square brackets above is bounded Hκ1Hκ1H^{-1}_{\kappa}\to H^{1}_{\kappa}. This allows us to write

R0fψ2(R0u)2R0=(R0fR0)A2(R0ψuR0)A1(R0ψuR0)\sqrt{R_{0}}f\psi^{2}(R_{0}u)^{2}\sqrt{R_{0}}=\big{(}\sqrt{R_{0}}f\sqrt{R_{0}}\big{)}A_{2}\big{(}\sqrt{R_{0}}\psi u\sqrt{R_{0}}\big{)}A_{1}\big{(}\sqrt{R_{0}}\psi u\sqrt{R_{0}}\big{)}

for operators AjA_{j} with Aj1\left\lVert A_{j}\right\rVert_{\mathfrak{I}_{\infty}}\lesssim 1.

Now we use (3.16) for the ff contribution and for one of the ψu\psi u factors, and (3.17) for the other ψu\psi u factor. This yields

R0fψ2(R0u)2R01\displaystyle\big{\lVert}\sqrt{R_{0}}f\psi^{2}(R_{0}u)^{2}\sqrt{R_{0}}\big{\rVert}_{\mathfrak{I}_{1}}\lesssim κ32fHκ1ψuHκ132(ψu)Hκ212\displaystyle\ \kappa^{-\frac{3}{2}}\|f\|_{H^{-1}_{\kappa}}\|\psi u\|_{H^{-1}_{\kappa}}^{\frac{3}{2}}\|(\psi u)^{\prime}\|_{H^{-2}_{\kappa}}^{\frac{1}{2}}
\displaystyle\lesssim κ32fHκ1ψuHκ1(c1ψuHκ1+c(ψu)Hκ2).\displaystyle\ \kappa^{-\frac{3}{2}}\|f\|_{H^{-1}_{\kappa}}\|\psi u\|_{H^{-1}_{\kappa}}\big{(}c^{-1}\|\psi u\|_{H^{-1}_{\kappa}}+c\|(\psi u)^{\prime}\|_{H^{-2}_{\kappa}}\big{)}.

Finally, we integrate in time and use (3.3) to arrive at

R0fψ2(R0u)2R0Lt11\displaystyle\big{\lVert}\sqrt{R_{0}}f\psi^{2}(R_{0}u)^{2}\sqrt{R_{0}}\big{\rVert}_{L^{1}_{t}\mathfrak{I}_{1}}
κ32fLt2Hκ1ψuLtHκ1(κ1cuLSκ+(c1+κ1c)uLt2Hκ1).\displaystyle\qquad\lesssim\kappa^{-\frac{3}{2}}\lVert f\rVert_{L^{2}_{t}H^{-1}_{\kappa}}\|\psi u\|_{L^{\infty}_{t}H^{-1}_{\kappa}}\big{(}\kappa^{-1}c\lVert u\rVert_{\textup{LS}_{\kappa}}+(c^{-1}+\kappa^{-1}c)\lVert u\rVert_{L^{2}_{t}H^{-1}_{\kappa}}\big{)}.

Then by applying Hölder’s inequality in time, we obtain

R0fψ2(R0u)2R0Lt11κ52fLt2Hκ1ψuLtHκ1(cuLSκ+T12(κc1+c)uLtHκ1).\displaystyle\big{\lVert}\sqrt{R_{0}}f\psi^{2}(R_{0}u)^{2}\sqrt{R_{0}}\big{\rVert}_{L^{1}_{t}\mathfrak{I}_{1}}\lesssim\kappa^{-\frac{5}{2}}\lVert f\rVert_{L^{2}_{t}H^{-1}_{\kappa}}\|\psi u\|_{L^{\infty}_{t}H^{-1}_{\kappa}}\big{(}c\lVert u\rVert_{\textup{LS}_{\kappa}}+T^{\frac{1}{2}}(\kappa c^{-1}+c)\lVert u\rVert_{L^{\infty}_{t}H^{-1}_{\kappa}}\big{)}.

Optimizing the choice of cc as in (3.18) this yields

R0fψ2(R0u)2R0Lt11κ2T14fLt2Hκ1uLtHκ1(uLtHκ1+uLSκ),\displaystyle\big{\lVert}\sqrt{R_{0}}f\psi^{2}(R_{0}u)^{2}\sqrt{R_{0}}\big{\rVert}_{L^{1}_{t}\mathfrak{I}_{1}}\lesssim\kappa^{-2}T^{\frac{1}{4}}\lVert f\rVert_{L^{2}_{t}H^{-1}_{\kappa}}\|u\|_{L^{\infty}_{t}H^{-1}_{\kappa}}\big{(}\left\lVert u\right\rVert_{L^{\infty}_{t}H^{-1}_{\kappa}}+\left\lVert u\right\rVert_{\textup{LS}_{\kappa}}\big{)},

which suffices for =2\ell=2.

For 3\ell\geq 3, we estimate

R0fψ2(R0u)R0Lt11R0fψ2(R0u)2R0Lt11R0uR0Lt22\displaystyle\big{\lVert}\sqrt{R_{0}}f\psi^{2}(R_{0}u)^{\ell}\sqrt{R_{0}}\big{\rVert}_{L^{1}_{t}\mathfrak{I}_{1}}\leq\big{\lVert}\sqrt{R_{0}}f\psi^{2}(R_{0}u)^{2}\sqrt{R_{0}}\big{\rVert}_{L^{1}_{t}\mathfrak{I}_{1}}\big{\lVert}\sqrt{R_{0}}u\sqrt{R_{0}}\big{\rVert}_{L^{\infty}_{t}\mathfrak{I}_{2}}^{\ell-2}
κ2T14fLt2Hκ1uLtHκ1(uLtHκ1+uLSκ)(κ12uLtHκ1)2.\displaystyle\qquad\lesssim\kappa^{-2}T^{\frac{1}{4}}\left\lVert f\right\rVert_{L^{2}_{t}H^{-1}_{\kappa}}\left\lVert u\right\rVert_{L^{\infty}_{t}H^{-1}_{\kappa}}\big{(}\left\lVert u\right\rVert_{L^{\infty}_{t}H^{-1}_{\kappa}}+\left\lVert u\right\rVert_{\textup{LS}_{\kappa}}\big{)}\big{(}\kappa^{-\frac{1}{2}}\left\lVert u\right\rVert_{L^{\infty}_{t}H^{-1}_{\kappa}}\big{)}^{\ell-2}.

Summing over 2\ell\geq 2 (exactly like in the proof of the previous lemma) and taking a supremum over fLt2Hκ11\left\lVert f\right\rVert_{L^{2}_{t}H^{-1}_{\kappa}}\leq 1 yields the second inequality in (3.20). ∎

Altogether, the previous lemmas provide us with the control we need over the functional ugu\mapsto g. However, the other key functional in our analysis is u1gu\mapsto\tfrac{1}{g}; indeed, this is what appears in the definition (2.9) of ρ\rho. The following lemma provides us with an estimate for the quadratic and higher order terms of 1g\tfrac{1}{g}, using the tools that we have already developed.

Lemma 3.7.

Given a Schwartz function ϕ(x)\phi(x), we have

(3.21) ϕ(1g2κ+4κ2h11+2κh1)Lt,x2κ32(Tκ2)14uLtHκ1(uLtHκ1+uLSκ)\big{\lVert}\phi\big{(}\tfrac{1}{g}-2\kappa+\tfrac{4\kappa^{2}h_{1}}{1+2\kappa h_{1}}\big{)}\big{\rVert}_{L^{2}_{t,x}}\lesssim\kappa^{-\frac{3}{2}}(T\kappa^{2})^{\frac{1}{4}}\left\lVert u\right\rVert_{L^{\infty}_{t}H^{-1}_{\kappa}}\big{(}\left\lVert u\right\rVert_{L^{\infty}_{t}H^{-1}_{\kappa}}+\left\lVert u\right\rVert_{\textup{LS}_{\kappa}}\big{)}

uniformly for Tκ2T\leq\kappa^{-2} and κ\kappa as in (3.9).

Proof.

Recall that h1=κ1R0(2κ)uh_{1}=-\kappa^{-1}R_{0}(2\kappa)u. Choosing CC larger if necessary, the condition (3.9) implies

2κh1Lx=2R0(2κ)uLxκ12uHκ1<1.\left\lVert 2\kappa h_{1}\right\rVert_{L^{\infty}_{x}}=\left\lVert 2R_{0}(2\kappa)u\right\rVert_{L^{\infty}_{x}}\lesssim\kappa^{-\frac{1}{2}}\left\lVert u\right\rVert_{H^{-1}_{\kappa}}<1.

Hence,

(3.22) 11+2κh1Lt,x+2κh11+2κh1Lt,x1\left\|\dfrac{1}{1+2\kappa h_{1}}\right\|_{L^{\infty}_{t,x}}+\left\|\dfrac{2\kappa h_{1}}{1+2\kappa h_{1}}\right\|_{L^{\infty}_{t,x}}\lesssim 1

uniformly for κ\kappa satisfying (3.9).

Now, we use the identity

1g2κ+4κ2h11+2κh1=2κg(12κh11+2κh1)(g12κh1)\frac{1}{g}-2\kappa+\frac{4\kappa^{2}h_{1}}{1+2\kappa h_{1}}=-\frac{2\kappa}{g}\left(1-\frac{2\kappa h_{1}}{1+2\kappa h_{1}}\right)\left(g-\frac{1}{2\kappa}-h_{1}\right)

together with the estimates (2.7) and (3.20) to bound

ϕ(1g2κ+4κ2h11+2κh1)Lt,x2\displaystyle\big{\lVert}\phi\big{(}\tfrac{1}{g}-2\kappa+\tfrac{4\kappa^{2}h_{1}}{1+2\kappa h_{1}}\big{)}\big{\rVert}_{L^{2}_{t,x}} κ1gLt,xϕ(g12κh1)Lt,x2\displaystyle\lesssim\kappa\big{\lVert}\tfrac{1}{g}\big{\rVert}_{L^{\infty}_{t,x}}\big{\lVert}\phi\big{(}g-\tfrac{1}{2\kappa}-h_{1}\big{)}\big{\rVert}_{L^{2}_{t,x}}
κψ(g12κh1)Lt2Hκ1\displaystyle\lesssim\kappa\big{\lVert}\psi\big{(}g-\tfrac{1}{2\kappa}-h_{1}\big{)}\big{\rVert}_{L^{2}_{t}H^{1}_{\kappa}}
κ1T14uLtHκ1(uLtHκ1+uLSκ).\displaystyle\lesssim\kappa^{-1}T^{\frac{1}{4}}\left\lVert u\right\rVert_{L^{\infty}_{t}H^{-1}_{\kappa}}\big{(}\left\lVert u\right\rVert_{L^{\infty}_{t}H^{-1}_{\kappa}}+\left\lVert u\right\rVert_{\textup{LS}_{\kappa}}\big{)}.\qed

Before proceeding, we will record one more estimate for the product h1h2h_{1}h_{2}, similar in spirit to the preceding analysis. (This will be useful in estimating (4.15).)

Lemma 3.8.

Given a Schwartz function ϕ(x)\phi(x), we have

(3.23) ϕh1h2Lt,x1κ132(Tκ2)34uLtHκ1(uLtHκ12+uLSκ2)\left\lVert\phi h_{1}h_{2}\right\rVert_{L^{1}_{t,x}}\lesssim\kappa^{-\frac{13}{2}}(T\kappa^{2})^{\frac{3}{4}}\left\lVert u\right\rVert_{L^{\infty}_{t}H^{-1}_{\kappa}}\big{(}\left\lVert u\right\rVert_{L^{\infty}_{t}H^{-1}_{\kappa}}^{2}+\left\lVert u\right\rVert_{\textup{LS}_{\kappa}}^{2}\big{)}

uniformly for Tκ2T\leq\kappa^{-2} and κ1\kappa\geq 1.

Proof.

Using a continuous partition of unity argument as in the proof of (3.4), it suffices to prove (3.23) in the special cases where ϕ=ψ3\phi=\psi^{3}.

We argue by duality: for fL([T,T]×)f\in L^{\infty}([-T,T]\times\mathbb{R}), we have

TTfψ3h1h2𝑑x𝑑t=TTtr{R0fψ3h1(R0u)2R0}𝑑t.\int_{-T}^{T}\int_{-\infty}^{\infty}f\psi^{3}h_{1}h_{2}\,dx\,dt=\int_{-T}^{T}\operatorname{tr}\big{\{}\sqrt{R_{0}}f\psi^{3}h_{1}(R_{0}u)^{2}\sqrt{R_{0}}\big{\}}\,dt.

By (2.17), we may write

R0fψ3h1(R0u)2R0=R0fψh1R0A2R0ψuR0A1R0ψuR0\sqrt{R_{0}}f\psi^{3}h_{1}(R_{0}u)^{2}\sqrt{R_{0}}=\sqrt{R_{0}}f\psi h_{1}\sqrt{R_{0}}A_{2}\sqrt{R_{0}}\psi u\sqrt{R_{0}}A_{1}\sqrt{R_{0}}\psi u\sqrt{R_{0}}

for operators AjA_{j} with Aj1\left\lVert A_{j}\right\rVert_{\mathfrak{I}_{\infty}}\lesssim 1.

Now we use the bounds (3.16) and (3.17) to estimate

R0fψ3h1(R0u)2R01κ32fψh1Hκ1ψuHκ132(ψu)Hκ212.\big{\lVert}\sqrt{R_{0}}f\psi^{3}h_{1}(R_{0}u)^{2}\sqrt{R_{0}}\big{\rVert}_{\mathfrak{I}_{1}}\lesssim\kappa^{-\frac{3}{2}}\|f\psi h_{1}\|_{H^{-1}_{\kappa}}\|\psi u\|_{H^{-1}_{\kappa}}^{\frac{3}{2}}\|(\psi u)^{\prime}\|_{H^{-2}_{\kappa}}^{\frac{1}{2}}.

As h1=κ1R0(2κ)uh_{1}=-\kappa^{-1}R_{0}(2\kappa)u, the first factor on the right is estimated as

fψh1Hκ1κ1fLψh1L2κ3fLψR0uHκ1κ3fLψuHκ1.\|f\psi h_{1}\|_{H^{-1}_{\kappa}}\lesssim\kappa^{-1}\|f\|_{L^{\infty}}\|\psi h_{1}\|_{L^{2}}\lesssim\kappa^{-3}\|f\|_{L^{\infty}}\|\psi R_{0}u\|_{H^{1}_{\kappa}}\lesssim\kappa^{-3}\|f\|_{L^{\infty}}\|\psi u\|_{H^{-1}_{\kappa}}.

Hence we obtain the fixed time bound

R0fψ3h1(R0u)2R01\displaystyle\big{\lVert}\sqrt{R_{0}}f\psi^{3}h_{1}(R_{0}u)^{2}\sqrt{R_{0}}\big{\rVert}_{\mathfrak{I}_{1}}\lesssim κ92fLψuHκ152(ψu)Hκ212\displaystyle\ \kappa^{-\frac{9}{2}}\|f\|_{L^{\infty}}\|\psi u\|_{H^{-1}_{\kappa}}^{\frac{5}{2}}\|(\psi u)^{\prime}\|_{H^{-2}_{\kappa}}^{\frac{1}{2}}
\displaystyle\lesssim κ92fLψuHκ1(c3(ψu)Hκ22+c1ψuHκ12).\displaystyle\ \kappa^{-\frac{9}{2}}\|f\|_{L^{\infty}}\|\psi u\|_{H^{-1}_{\kappa}}\big{(}c^{3}\|(\psi u)^{\prime}\|_{H^{-2}_{\kappa}}^{2}+c^{-1}\|\psi u\|_{H^{-1}_{\kappa}}^{2}\big{)}.

Integrating in time and using (3.3) this gives

R0fψ3h1(R0u)2R0Lt11\displaystyle\big{\lVert}\sqrt{R_{0}}f\psi^{3}h_{1}(R_{0}u)^{2}\sqrt{R_{0}}\big{\rVert}_{L^{1}_{t}\mathfrak{I}_{1}}
κ132fLt,xψuLtHκ1(c3uLSκ2+T(c1κ2+c3)ψuLtHκ12).\displaystyle\qquad\lesssim\kappa^{-\frac{13}{2}}\|f\|_{L^{\infty}_{t,x}}\|\psi u\|_{L^{\infty}_{t}H^{-1}_{\kappa}}\big{(}c^{3}\|u\|_{\textup{LS}_{\kappa}}^{2}+T(c^{-1}\kappa^{2}+c^{3})\|\psi u\|_{L^{\infty}_{t}H^{-1}_{\kappa}}^{2}\big{)}.

Balancing cc as in (3.18) we arrive at

R0fψ3h1(R0u)2R0Lt11\displaystyle\big{\lVert}\sqrt{R_{0}}f\psi^{3}h_{1}(R_{0}u)^{2}\sqrt{R_{0}}\big{\rVert}_{L^{1}_{t}\mathfrak{I}_{1}}\lesssim κ132(Tκ2)34fLt,xψuLtHκ1(uLtHκ12+uLSκ2)\displaystyle\ \kappa^{-\frac{13}{2}}(T\kappa^{2})^{\frac{3}{4}}\|f\|_{L^{\infty}_{t,x}}\|\psi u\|_{L^{\infty}_{t}H^{-1}_{\kappa}}\big{(}\left\lVert u\right\rVert_{L^{\infty}_{t}H^{-1}_{\kappa}}^{2}+\left\lVert u\right\rVert_{\textup{LS}_{\kappa}}^{2}\big{)}

as needed. ∎

We conclude this section with two estimates for the functional derivative of ρ\rho, which appears on RHS(2.13). First, we have the following estimate that does not make use of local smoothing:

Lemma 3.9.

Given T>0T>0, we have

(3.24) dρ|u(f)Lt,x1κ1uLtHκ1fLt1Hκ1\big{\lVert}d\rho|_{u}(f)\big{\rVert}_{L^{1}_{t,x}}\lesssim\kappa^{-1}\left\lVert u\right\rVert_{L^{\infty}_{t}H^{-1}_{\kappa}}\left\lVert f\right\rVert_{L^{1}_{t}H^{-1}_{\kappa}}

uniformly for κ\kappa satisfying (3.9).

Proof.

We argue by duality: let hL([T,T]×)h\in L^{\infty}([-T,T]\times\mathbb{R}). Expanding the series (2.3) in the definition (2.15) of dgdg, we obtain

TTh1g2𝑑g|u(f)dxdt=,m0TT(1)+m+1tr{hg2(R0u)R0f(R0u)mR0}𝑑t.\int_{-T}^{T}\int h\tfrac{1}{g^{2}}dg|_{u}(f)\,dx\,dt=\sum_{\ell,m\geq 0}\int_{-T}^{T}(-1)^{\ell+m+1}\operatorname{tr}\big{\{}\tfrac{h}{g^{2}}(R_{0}u)^{\ell}R_{0}f(R_{0}u)^{m}R_{0}\big{\}}\,dt.

We will insert this into the formula (2.14) for dρd\rho.

For =m=0\ell=m=0, we have

tr{hg2R0fR0}\displaystyle-\operatorname{tr}\big{\{}\tfrac{h}{g^{2}}R_{0}fR_{0}\big{\}} =hg2κ1R0(2κ)f𝑑x\displaystyle=-\int\tfrac{h}{g^{2}}\,\kappa^{-1}R_{0}(2\kappa)f\,dx
=h 4κR0(2κ)f𝑑xh(1g24κ2)κ1R0(2κ)f𝑑x.\displaystyle=-\int h\,4\kappa R_{0}(2\kappa)f\,dx-\int h\big{(}\tfrac{1}{g^{2}}-4\kappa^{2}\big{)}\kappa^{-1}R_{0}(2\kappa)f\,dx.

The first term on the RHS above is canceled out by the contribution of 2κR0(2κ)f2\kappa R_{0}(2\kappa)f in (2.14). For the second term, we use (2.7) to estimate

TT|h(1g24κ2)κ1R0(2κ)f|𝑑x𝑑t\displaystyle\int_{-T}^{T}\int\big{|}h\big{(}\tfrac{1}{g^{2}}-4\kappa^{2}\big{)}\kappa^{-1}R_{0}(2\kappa)f\big{|}\,dx\,dt
hLt,x1g2κLtLx2(1gLt,x+2κ)κ1R0(2κ)fLt1Lx2\displaystyle\qquad\lesssim\left\lVert h\right\rVert_{L^{\infty}_{t,x}}\lVert\tfrac{1}{g}-2\kappa\rVert_{L^{\infty}_{t}L^{2}_{x}}\big{(}\lVert\tfrac{1}{g}\rVert_{L^{\infty}_{t,x}}+2\kappa\big{)}\kappa^{-1}\left\lVert R_{0}(2\kappa)f\right\rVert_{L^{1}_{t}L^{2}_{x}}
κ1hLt,xuLtHκ1fLt1Hκ1.\displaystyle\qquad\lesssim\kappa^{-1}\left\lVert h\right\rVert_{L^{\infty}_{t,x}}\left\lVert u\right\rVert_{L^{\infty}_{t}H^{-1}_{\kappa}}\left\lVert f\right\rVert_{L^{1}_{t}H^{-1}_{\kappa}}.

So it only remains to deal with the terms for which +m1\ell+m\geq 1. In this case, there are at least two operators we can put in Hilbert–Schmidt norm:

+m1hg2(R0u)R0f(R0u)mR0Lt11\displaystyle\sum_{\ell+m\geq 1}\big{\lVert}\tfrac{h}{g^{2}}(R_{0}u)^{\ell}R_{0}f(R_{0}u)^{m}R_{0}\big{\rVert}_{L^{1}_{t}\mathfrak{I}_{1}}
+m1R0hg2R0LtR0fR0Lt12R0uR0Lt2+m\displaystyle\qquad\leq\sum_{\ell+m\geq 1}\big{\lVert}\sqrt{R_{0}}\tfrac{h}{g^{2}}\sqrt{R_{0}}\big{\rVert}_{L^{\infty}_{t}\mathfrak{I}_{\infty}}\big{\lVert}\sqrt{R_{0}}f\sqrt{R_{0}}\big{\rVert}_{L^{1}_{t}\mathfrak{I}_{2}}\big{\lVert}\sqrt{R_{0}}u\sqrt{R_{0}}\big{\rVert}_{L^{\infty}_{t}\mathfrak{I}_{2}}^{\ell+m}
κ52+m1hg2Lt,xfLt1Hκ1(κ12uLtHκ1)+m\displaystyle\qquad\lesssim\kappa^{-\frac{5}{2}}\sum_{\ell+m\geq 1}\big{\lVert}\tfrac{h}{g^{2}}\big{\rVert}_{L^{\infty}_{t,x}}\left\lVert f\right\rVert_{L^{1}_{t}H^{-1}_{\kappa}}\big{(}\kappa^{-\frac{1}{2}}\left\lVert u\right\rVert_{L^{\infty}_{t}H^{-1}_{\kappa}}\big{)}^{\ell+m}
κ1uLtHκ1hLt,xfLt1Hκ1.\displaystyle\qquad\lesssim\kappa^{-1}\left\lVert u\right\rVert_{L^{\infty}_{t}H^{-1}_{\kappa}}\left\lVert h\right\rVert_{L^{\infty}_{t,x}}\left\lVert f\right\rVert_{L^{1}_{t}H^{-1}_{\kappa}}.

Altogether, we have

TT|hdρ|u(f)|dxdtκ1uLtHκ1hLt,xfLt1Hκ1.\int_{-T}^{T}\int\big{|}h\,d\rho|_{u}(f)\big{|}\,dx\,dt\lesssim\kappa^{-1}\left\lVert u\right\rVert_{L^{\infty}_{t}H^{-1}_{\kappa}}\left\lVert h\right\rVert_{L^{\infty}_{t,x}}\left\lVert f\right\rVert_{L^{1}_{t}H^{-1}_{\kappa}}.

Taking a supremum over hLt,x1\left\lVert h\right\rVert_{L^{\infty}_{t,x}}\leq 1, we obtain (3.24). ∎

We can also handle one derivative inside dρd\rho, provided that the input is localized in space and we may estimate one copy of uu in local smoothing norm:

Lemma 3.10.

We have

(3.25) |TTϕx0dρ|u(ψ2f)dxdt|κ1ψfLt2Hκ1(uLtHκ1+uLSκ)\bigg{|}\int_{-T}^{T}\int\phi_{x_{0}}\,d\rho|_{u}(\psi^{2}f)^{\prime}\,dx\,dt\bigg{|}\lesssim\kappa^{-1}\left\lVert\psi f\right\rVert_{L^{2}_{t}H^{-1}_{\kappa}}\big{(}\left\lVert u\right\rVert_{L^{\infty}_{t}H^{-1}_{\kappa}}+\left\lVert u\right\rVert_{\textup{LS}_{\kappa}}\big{)}

uniformly for x0x_{0}\in\mathbb{R}, Tκ2T\leq\kappa^{-2}, and κ\kappa satisfying (3.9), where ϕ\phi is defined by

(3.26) ϕx0(x)=ϕ(xx0)=6tanh(xx06)so thatϕx0=ψx02.\phi_{x_{0}}(x)=\phi(x-x_{0})=6\operatorname{tanh}(\tfrac{x-x_{0}}{6})\quad\textrm{so that}\quad\phi^{\prime}_{x_{0}}=\psi^{2}_{x_{0}}.
Proof.

In the following, we will set ϕ=ϕx0\phi=\phi_{x_{0}} for simplicity. First, we use (2.14) to write

2ϕ𝑑ρ|u(ψ2f)dx\displaystyle 2\int\phi\,d\rho|_{u}(\psi^{2}f)^{\prime}\,dx =ϕ{1g2dg|u(ψ2f)+4κR0(2κ)(ψ2f)}𝑑x\displaystyle=\int\phi\Big{\{}\tfrac{1}{g^{2}}dg|_{u}(\psi^{2}f)^{\prime}+4\kappa R_{0}(2\kappa)(\psi^{2}f)^{\prime}\Big{\}}\,dx
(3.27) =ϕg2{dg|u(ψ2f)+κ1R0(2κ)(ψ2f)}𝑑x\displaystyle=\int\tfrac{\phi}{g^{2}}\Big{\{}dg|_{u}(\psi^{2}f)^{\prime}+\kappa^{-1}R_{0}(2\kappa)(\psi^{2}f)^{\prime}\Big{\}}\,dx
(3.28) κ1ϕ(1g+2κ)(1g2κ+4κ2h11+2κh1)R0(2κ)(ψ2f)𝑑x\displaystyle\quad-\kappa^{-1}\int\phi\big{(}\tfrac{1}{g}+2\kappa\big{)}\big{(}\tfrac{1}{g}-2\kappa+\tfrac{4\kappa^{2}h_{1}}{1+2\kappa h_{1}}\big{)}R_{0}(2\kappa)(\psi^{2}f)^{\prime}\,dx
(3.29) +κ1ϕ(1g+2κ)4κ2h11+2κh1R0(2κ)(ψ2f)𝑑x.\displaystyle\quad+\kappa^{-1}\int\phi\big{(}\tfrac{1}{g}+2\kappa\big{)}\tfrac{4\kappa^{2}h_{1}}{1+2\kappa h_{1}}R_{0}(2\kappa)(\psi^{2}f)^{\prime}\,dx.

We will estimate the contributions of (3.27)–(3.29) individually.

Let us start with (3.28). By (2.7), (3.21), and (2.17) we have

κ1TT|ϕ(1g+2κ)(1g2κ+4κ2h11+2κh1)R0(2κ)(ψ2f)𝑑x|𝑑t\displaystyle\kappa^{-1}\int_{-T}^{T}\bigg{|}\int\phi\big{(}\tfrac{1}{g}+2\kappa\big{)}\big{(}\tfrac{1}{g}-2\kappa+\tfrac{4\kappa^{2}h_{1}}{1+2\kappa h_{1}}\big{)}R_{0}(2\kappa)(\psi^{2}f)^{\prime}\,dx\bigg{|}\,dt
κ1ϕL1g+2κLt,xψ(1g2κ+4κ2h11+2κh1)Lt,x21ψR0ψHκ1L2ψfLt2Hκ1\displaystyle\qquad\lesssim\kappa^{-1}\left\lVert\phi\right\rVert_{L^{\infty}}\big{\lVert}\tfrac{1}{g}+2\kappa\big{\rVert}_{L^{\infty}_{t,x}}\big{\lVert}\psi\big{(}\tfrac{1}{g}-2\kappa+\tfrac{4\kappa^{2}h_{1}}{1+2\kappa h_{1}}\big{)}\big{\rVert}_{L^{2}_{t,x}}\big{\lVert}\tfrac{1}{\psi}R_{0}\partial\psi\big{\rVert}_{H^{-1}_{\kappa}\to L^{2}}\left\lVert\psi f\right\rVert_{L^{2}_{t}H^{-1}_{\kappa}}
κ32(Tκ2)14uLtHκ1(uLtHκ1+uLSκ)ψfLt2Hκ1.\displaystyle\qquad\lesssim\kappa^{-\frac{3}{2}}(T\kappa^{2})^{\frac{1}{4}}\left\lVert u\right\rVert_{L^{\infty}_{t}H^{-1}_{\kappa}}\big{(}\left\lVert u\right\rVert_{L^{\infty}_{t}H^{-1}_{\kappa}}+\left\lVert u\right\rVert_{\textup{LS}_{\kappa}}\big{)}\left\lVert\psi f\right\rVert_{L^{2}_{t}H^{-1}_{\kappa}}.

Next, we turn to (3.29). Integrating by parts and using (2.17), (2.7), (3.11), (3.22), (3.8), and (3.10), we have

κ1TT|ϕ(1g+2κ)4κ2h11+2κh1R0(2κ)(ψ2f)𝑑x|𝑑t\displaystyle\kappa^{-1}\int_{-T}^{T}\bigg{|}\int\phi\big{(}\tfrac{1}{g}+2\kappa\big{)}\tfrac{4\kappa^{2}h_{1}}{1+2\kappa h_{1}}R_{0}(2\kappa)(\psi^{2}f)^{\prime}\,dx\bigg{|}\,dt
κ1ψfLt2Hκ1ψR01ψL2Hκ1ψ[ϕ(1g+2κ)4κ2h11+2κh1]Lt,x2\displaystyle\qquad\lesssim\kappa^{-1}\left\lVert\psi f\right\rVert_{L^{2}_{t}H^{-1}_{\kappa}}\big{\lVert}\psi R_{0}\tfrac{1}{\psi}\big{\rVert}_{L^{2}\to H^{1}_{\kappa}}\big{\lVert}\psi\big{[}\phi\big{(}\tfrac{1}{g}+2\kappa\big{)}\tfrac{4\kappa^{2}h_{1}}{1+2\kappa h_{1}}\big{]}^{\prime}\big{\rVert}_{L^{2}_{t,x}}
κ2ψfLt2Hκ1{1g+2κLt,xψh1Lt,x24κ21+2κh1Lt,x(1+2κh11+2κh1Lt,x)\displaystyle\qquad\lesssim\kappa^{-2}\left\lVert\psi f\right\rVert_{L^{2}_{t}H^{-1}_{\kappa}}\Big{\{}\big{\lVert}\tfrac{1}{g}+2\kappa\big{\rVert}_{L^{\infty}_{t,x}}\left\lVert\psi h^{\prime}_{1}\right\rVert_{L^{2}_{t,x}}\big{\lVert}\tfrac{4\kappa^{2}}{1+2\kappa h_{1}}\big{\rVert}_{L^{\infty}_{t,x}}\big{(}1+\big{\lVert}\tfrac{2\kappa h_{1}}{1+2\kappa h_{1}}\big{\rVert}_{L^{\infty}_{t,x}}\big{)}
+ψgLt,x21gLt,x24κ2h11+2κh1Lt,x+1g+2κLt,x4κ21+2κh1Lt,xψh1Lt,x2}\displaystyle\qquad\qquad+\left\lVert\psi g^{\prime}\right\rVert_{L^{2}_{t,x}}\big{\lVert}\tfrac{1}{g}\big{\rVert}_{L^{\infty}_{t,x}}^{2}\big{\lVert}\tfrac{4\kappa^{2}h_{1}}{1+2\kappa h_{1}}\big{\rVert}_{L^{\infty}_{t,x}}+\big{\lVert}\tfrac{1}{g}+2\kappa\big{\rVert}_{L^{\infty}_{t,x}}\big{\lVert}\tfrac{4\kappa^{2}}{1+2\kappa h_{1}}\big{\rVert}_{L^{\infty}_{t,x}}\left\lVert\psi h_{1}\right\rVert_{L^{2}_{t,x}}\Big{\}}
κ1ψfLt2Hκ1(uLtHκ1+uLSκ).\displaystyle\qquad\lesssim\kappa^{-1}\left\lVert\psi f\right\rVert_{L^{2}_{t}H^{-1}_{\kappa}}\big{(}\left\lVert u\right\rVert_{L^{\infty}_{t}H^{-1}_{\kappa}}+\left\lVert u\right\rVert_{\textup{LS}_{\kappa}}\big{)}.

It remains to estimate (3.27). We write

ϕg2{dg|u(ψ2f)+κ1R0(2κ)(ψ2f)}𝑑x\displaystyle\int\tfrac{\phi}{g^{2}}\Big{\{}dg|_{u}(\psi^{2}f)^{\prime}+\kappa^{-1}R_{0}(2\kappa)(\psi^{2}f)^{\prime}\Big{\}}\,dx
=+m1(1)+m+1tr{ϕg2(R0u)R0ψ2[,ψ2f](R0u)mR0}\displaystyle\qquad=\sum_{\ell+m\geq 1}(-1)^{\ell+m+1}\operatorname{tr}\big{\{}\tfrac{\phi}{g^{2}}(R_{0}u)^{\ell}R_{0}\psi^{2}[\partial,\psi^{2}f](R_{0}u)^{m}R_{0}\big{\}}
=+m1(1)+mtr{ψ2f[,(R0u)R0ϕg2(R0u)mR0]}.\displaystyle\qquad=\sum_{\ell+m\geq 1}(-1)^{\ell+m}\operatorname{tr}\big{\{}\psi^{2}f\big{[}\partial,(R_{0}u)^{\ell}R_{0}\tfrac{\phi}{g^{2}}(R_{0}u)^{m}R_{0}\big{]}\big{\}}.

In the last equality, we cycled the trace (i.e. tr(AB)=tr(BA)\operatorname{tr}(AB)=\operatorname{tr}(BA)). Now, we will distribute the derivative \partial using the product rule [,AB]=[,A]B+A[,B][\partial,AB]=[\partial,A]B+A[\partial,B].

First, let us consider the terms with +m=1\ell+m=1. When the derivative lands on uu, we use (2.17) and (2.7) to estimate:

TT|tr{ψ2fR0uR0ϕg2R0}|+|tr{ψ2fR0ϕg2R0uR0}|dt\displaystyle\int_{-T}^{T}\big{|}\operatorname{tr}\big{\{}\psi^{2}fR_{0}u^{\prime}R_{0}\tfrac{\phi}{g^{2}}R_{0}\big{\}}\big{|}+\big{|}\operatorname{tr}\big{\{}\psi^{2}fR_{0}\tfrac{\phi}{g^{2}}R_{0}u^{\prime}R_{0}\big{\}}\big{|}\,dt
R0ψfR0Lt22R0ψuR0Lt22R0ϕg2R0Lt\displaystyle\qquad\lesssim\big{\lVert}\sqrt{R_{0}}\psi f\sqrt{R_{0}}\big{\rVert}_{L^{2}_{t}\mathfrak{I}_{2}}\big{\lVert}\sqrt{R_{0}}\psi u^{\prime}\sqrt{R_{0}}\big{\rVert}_{L^{2}_{t}\mathfrak{I}_{2}}\big{\lVert}\sqrt{R_{0}}\tfrac{\phi}{g^{2}}\sqrt{R_{0}}\big{\rVert}_{L^{\infty}_{t}\mathfrak{I}_{\infty}}
κ1ψfLt2Hκ1uLSκ.\displaystyle\qquad\lesssim\kappa^{-1}\left\lVert\psi f\right\rVert_{L^{2}_{t}H^{-1}_{\kappa}}\left\lVert u\right\rVert_{\textup{LS}_{\kappa}}.

When the derivative lands on ϕg2\tfrac{\phi}{g^{2}}, we use (2.4), (2.7), and (3.9) to estimate

(3.30) R0(ϕg2)R0Lt2\displaystyle\big{\lVert}\sqrt{R_{0}}(\tfrac{\phi}{g^{2}})^{\prime}\sqrt{R_{0}}\big{\rVert}_{L^{2}_{t}\mathfrak{I}_{\infty}} κ32T12(ϕg2)LtLx2κ32T12(1gLt,x2+1gLt,x3gLt,x2)\displaystyle\lesssim\kappa^{-\frac{3}{2}}T^{\frac{1}{2}}\big{\lVert}(\tfrac{\phi}{g^{2}})^{\prime}\big{\rVert}_{L^{\infty}_{t}L^{2}_{x}}\lesssim\kappa^{-\frac{3}{2}}T^{\frac{1}{2}}\big{(}\big{\lVert}\tfrac{1}{g}\big{\rVert}_{L^{\infty}_{t,x}}^{2}+\big{\lVert}\tfrac{1}{g}\big{\rVert}_{L^{\infty}_{t,x}}^{3}\left\lVert g^{\prime}\right\rVert_{L^{2}_{t,x}}\big{)}
κ12T12(1+uLtHκ1)1.\displaystyle\lesssim\kappa^{\frac{1}{2}}T^{\frac{1}{2}}\big{(}1+\left\lVert u\right\rVert_{L^{\infty}_{t}H^{-1}_{\kappa}}\big{)}\lesssim 1.

This yields

TT|tr{ψ2fR0uR0(ϕg2)R0}|+|tr{ψ2fR0(ϕg2)R0uR0}|dt\displaystyle\int_{-T}^{T}\big{|}\operatorname{tr}\big{\{}\psi^{2}fR_{0}uR_{0}\big{(}\tfrac{\phi}{g^{2}}\big{)}^{\prime}R_{0}\big{\}}\big{|}+\big{|}\operatorname{tr}\big{\{}\psi^{2}fR_{0}\big{(}\tfrac{\phi}{g^{2}}\big{)}^{\prime}R_{0}uR_{0}\big{\}}\big{|}\,dt
R0ψ2fR0Lt22R0uR0Lt2R0(ϕg2)R0Lt2\displaystyle\qquad\lesssim\big{\lVert}\sqrt{R_{0}}\psi^{2}f\sqrt{R_{0}}\big{\rVert}_{L^{2}_{t}\mathfrak{I}_{2}}\big{\lVert}\sqrt{R_{0}}u\sqrt{R_{0}}\big{\rVert}_{L^{\infty}_{t}\mathfrak{I}_{2}}\big{\lVert}\sqrt{R_{0}}(\tfrac{\phi}{g^{2}})^{\prime}\sqrt{R_{0}}\big{\rVert}_{L^{2}_{t}\mathfrak{I}_{\infty}}
κ1ψfLt2Hκ1uLtHκ1.\displaystyle\qquad\lesssim\kappa^{-1}\left\lVert\psi f\right\rVert_{L^{2}_{t}H^{-1}_{\kappa}}\left\lVert u\right\rVert_{L^{\infty}_{t}H^{-1}_{\kappa}}.

Lastly, we turn to the terms with +m2\ell+m\geq 2. When the derivative lands on uu, we use (3.7). When the derivative lands on ϕg2\tfrac{\phi}{g^{2}}, we use (3.30). As +m2\ell+m\geq 2, there are always at least two terms that we can put in Hilbert–Schmidt norm:

TT|tr{ψ2f[,(R0u)R0ϕg2(R0u)mR0]}|𝑑t\displaystyle\int_{-T}^{T}\big{|}\operatorname{tr}\big{\{}\psi^{2}f\big{[}\partial,(R_{0}u)^{\ell}R_{0}\tfrac{\phi}{g^{2}}(R_{0}u)^{m}R_{0}\big{]}\big{\}}\big{|}\,dt
R0ψ2fR0Lt22{R0(ϕg2)R0Lt2R0uR0Lt2+m\displaystyle\qquad\lesssim\big{\lVert}\sqrt{R_{0}}\psi^{2}f\sqrt{R_{0}}\big{\rVert}_{L^{2}_{t}\mathfrak{I}_{2}}\Big{\{}\big{\lVert}\sqrt{R_{0}}\big{(}\tfrac{\phi}{g^{2}}\big{)}^{\prime}\sqrt{R_{0}}\big{\rVert}_{L^{2}_{t}\mathfrak{I}_{\infty}}\big{\lVert}\sqrt{R_{0}}u\sqrt{R_{0}}\big{\rVert}_{L^{\infty}_{t}\mathfrak{I}_{2}}^{\ell+m}
+(+m)R0uR0Lt2R0ϕg2R0LtR0uR0Lt2+m1}\displaystyle\qquad\qquad+(\ell+m)\big{\lVert}\sqrt{R_{0}}u^{\prime}\sqrt{R_{0}}\big{\rVert}_{L^{2}_{t}\mathfrak{I}_{\infty}}\big{\lVert}\sqrt{R_{0}}\tfrac{\phi}{g^{2}}\sqrt{R_{0}}\big{\rVert}_{L^{\infty}_{t}\mathfrak{I}_{\infty}}\big{\lVert}\sqrt{R_{0}}u\sqrt{R_{0}}\big{\rVert}_{L^{\infty}_{t}\mathfrak{I}_{2}}^{\ell+m-1}\Big{\}}
κ1ψfLt2Hκ1(uLtHκ1+uLSκ)(+m+1)(κ12uLtHκ1)+m1.\displaystyle\qquad\lesssim\kappa^{-1}\left\lVert\psi f\right\rVert_{L^{2}_{t}H^{-1}_{\kappa}}\big{(}\left\lVert u\right\rVert_{L^{\infty}_{t}H^{-1}_{\kappa}}+\left\lVert u\right\rVert_{\textup{LS}_{\kappa}}\big{)}(\ell+m+1)\big{(}\kappa^{-\frac{1}{2}}\left\lVert u\right\rVert_{L^{\infty}_{t}H^{-1}_{\kappa}}\big{)}^{\ell+m-1}.

Summing over +m2\ell+m\geq 2 then yields

+m2TT|tr{ψ2f[,(R0u)R0ϕg2(R0u)mR0]}|𝑑tκ1ψfLt2Hκ1(uLtHκ1+uLSκ).\displaystyle\sum_{\ell+m\geq 2}\int_{-T}^{T}\big{|}\operatorname{tr}\big{\{}\psi^{2}f\big{[}\partial,(R_{0}u)^{\ell}R_{0}\tfrac{\phi}{g^{2}}(R_{0}u)^{m}R_{0}\big{]}\big{\}}\big{|}\,dt\lesssim\kappa^{-1}\left\lVert\psi f\right\rVert_{L^{2}_{t}H^{-1}_{\kappa}}\big{(}\left\lVert u\right\rVert_{L^{\infty}_{t}H^{-1}_{\kappa}}+\left\lVert u\right\rVert_{\textup{LS}_{\kappa}}\big{)}.

This was the final term that we needed to estimate, and thus concludes the proof of the lemma. ∎

4. The local smoothing estimate

In this section, we will prove our main local smoothing estimate for solutions to (gKdV):

Theorem 4.1.

Let κ1\kappa\geq 1, and uu a solution to (gKdV) which satisfies the bound (3.9) uniformly in a time interval [0,T][0,T] with Tκ2T\leq\kappa^{-2}. For some ε>0\varepsilon>0 assume that the coefficients a1,a2,a3,a4a_{1},a_{2},a_{3},a_{4} obey the following bounds in [0,T][0,T], uniformly in zz\in\mathbb{R}:

(4.1) ψza1LtHx1𝑑zεorψza1LtWx1,𝑑zε,\displaystyle\int\left\lVert\psi_{z}a_{1}\right\rVert_{L^{\infty}_{t}H^{1}_{x}}\,dz\leq\varepsilon\quad\textrm{or}\quad\int\left\lVert\psi_{z}a_{1}\right\rVert_{L^{\infty}_{t}W^{1,\infty}_{x}}\,dz\leq\varepsilon,
(4.2) ψza2Lt,x𝑑zε,\displaystyle\int\left\lVert\psi_{z}a_{2}\right\rVert_{L^{\infty}_{t,x}}\,dz\leq\varepsilon,
(4.3) ψza3Lt2Hx1𝑑zεorψza3Lt2Wx1,𝑑zε,\displaystyle\int\left\lVert\psi_{z}a_{3}\right\rVert_{L^{2}_{t}H^{1}_{x}}\,dz\leq\varepsilon\quad\textrm{or}\quad\int\left\lVert\psi_{z}a_{3}\right\rVert_{L^{2}_{t}W^{1,\infty}_{x}}\,dz\leq\varepsilon,
(4.4) a4LtHx1<ora4LtWx1,<.\displaystyle\left\lVert a_{4}\right\rVert_{L^{\infty}_{t}H^{1}_{x}}<\infty\quad\textrm{or}\quad\left\lVert a_{4}\right\rVert_{L^{\infty}_{t}W^{1,\infty}_{x}}<\infty.

Then the solution u(t)u(t) to (gKdV) satisfies the local energy estimate

(4.5) uLSκ22uCtHκ12+C(ε+(Tκ2)14+κ2)(uLtHκ12+uLSκ2)\displaystyle\begin{aligned} \left\lVert u\right\rVert_{\textup{LS}_{\kappa}}^{2}\leq{}&2\left\lVert u\right\rVert_{C_{t}H^{-1}_{\kappa}}^{2}+C(\varepsilon+(T\kappa^{2})^{\frac{1}{4}}+\kappa^{-2})\big{(}\left\lVert u\right\rVert_{L^{\infty}_{t}H^{-1}_{\kappa}}^{2}+\left\lVert u\right\rVert_{\textup{LS}_{\kappa}}^{2}\big{)}\end{aligned}

For the definition of ψx0\psi_{x_{0}} and LSκ\textup{LS}_{\kappa}, see (3.1). The rest of the section is devoted to the proof of the theorem.

Proof of Theorem 4.1.

Recall that solutions u(t)u(t) to (gKdV) obey the approximate conservation law (2.13). In order to prove the local smoothing estimate, we will multiply (2.13) by the function ϕ\phi defined in (3.26) and integrate in space and time:

TTψx02j(t,x)𝑑x𝑑t=\displaystyle\int_{-T}^{T}\int\psi^{2}_{x_{0}}j(t,x)\,dx\,dt= ϕx0[ρ(T,x)ρ(T,x)]𝑑x\displaystyle\ \int\phi_{x_{0}}[\rho(T,x)-\rho(-T,x)]\,dx
TTϕx0𝑑ρ|u[(a1u)+a2u2+a3u+a4u]dxdt.\displaystyle-\int_{-T}^{T}\int\phi_{x_{0}}\,d\rho|_{u}\big{[}(a_{1}u^{\prime})^{\prime}+a_{2}u^{2}+a_{3}u^{\prime}+a_{4}u\big{]}\,dx\,dt.

Here we seek to identify the left hand side, respectively the first term on the right, with the left hand side, respectively the first term on the right, in (4.5) modulo acceptable errors.

We begin with the first term on the right, which is easiest. Recall that ρ(t,x)0\rho(t,x)\geq 0, and so ρL1=α(κ,u)\left\lVert\rho\right\rVert_{L^{1}}=\alpha(\kappa,u). For the first term on the RHS, we use (2.10) to estimate

(4.6) κ|ϕx0[ρ(T,x)ρ(T,x)]𝑑x|2uCtHκ12.\kappa\bigg{|}\int\phi_{x_{0}}[\rho(T,x)-\rho(-T,x)]\,dx\bigg{|}\leq 2\left\lVert u\right\rVert_{C_{t}H^{-1}_{\kappa}}^{2}.

We next consider the left hand side, for which we will show that

(4.7) κTTψx02j(t,x)=32ψx0uLt2Hκ12dxdt+Err0,\kappa\int_{-T}^{T}\int\psi^{2}_{x_{0}}j(t,x)=-\tfrac{3}{2}\big{\lVert}\psi_{x_{0}}u^{\prime}\big{\rVert}_{L^{2}_{t}H^{-1}_{\kappa}}^{2}\,dx\,dt+Err_{0},

where for brevity Err0Err_{0} is an expression satisfying the bound

|Err0|((Tκ2)14+κ2)(uLtHκ12+uLSκ2).|Err_{0}|\lesssim({(T\kappa^{2}})^{\frac{1}{4}}+\kappa^{-2})\big{(}\left\lVert u\right\rVert_{L^{\infty}_{t}H^{-1}_{\kappa}}^{2}+\left\lVert u\right\rVert_{\textup{LS}_{\kappa}}^{2}\big{)}.

The last task is then to estimate the second term on the right, and show that

(4.8) κTTϕx0𝑑ρ|u[(a1u)+a2u2+a3u+a4u]dxdt=Err1,\kappa\int_{-T}^{T}\int\phi_{x_{0}}\,d\rho|_{u}\big{[}(a_{1}u^{\prime})^{\prime}+a_{2}u^{2}+a_{3}u^{\prime}+a_{4}u\big{]}\,dx\,dt=Err_{1},

where

|Err1|(ε+κ2)(uLtHκ12+uLSκ2).|Err_{1}|\lesssim(\varepsilon+\kappa^{-2})\big{(}\left\lVert u\right\rVert_{L^{\infty}_{t}H^{-1}_{\kappa}}^{2}+\left\lVert u\right\rVert_{\textup{LS}_{\kappa}}^{2}\big{)}.

The desired estimate (4.5) is obtained by combining (4.6), (4.7) and (4.8) followed by taking the supremum over x0x_{0}\in\mathbb{R}. It remains to prove the bounds (4.7) and (4.8).


We begin with the proof of (4.7). To leading order, jj is quadratic in uu. Specifically, if we insert the series (2.8) for gg, then the quadratic terms of jj are:

j2=8κ4h2h1(16κ5h1+4κ2u)6κR0(2κ)[u2].j_{2}=8\kappa^{4}h_{2}-h_{1}(16\kappa^{5}h_{1}+4\kappa^{2}u)-6\kappa R_{0}(2\kappa)[u^{2}].

It is then natural to expect the leading contribution to come from j2j_{2}, while the cubic and higher contributions coming from jj2j-j_{2} to be perturbative. We do this in the next two lemmas. In the first lemma we examine the contribution of j2j_{2}, and show that this generates the local smoothing norm of uu:

Lemma 4.2.

Assume that κ1\kappa\geq 1 and uu satisfies (3.9) uniformly in [0,T][0,T]. Then we have

ψx0uLt2Hκ12=\displaystyle\big{\lVert}\psi_{x_{0}}u^{\prime}\big{\rVert}_{L^{2}_{t}H^{-1}_{\kappa}}^{2}= 23κTTψx02j2𝑑x𝑑t+O(κ2(uLtHκ12+uLSκ2))\displaystyle-\tfrac{2}{3}\kappa\int_{-T}^{T}\int_{-\infty}^{\infty}\psi_{x_{0}}^{2}j_{2}\,dx\,dt+O\Big{(}\kappa^{-2}\big{(}\left\lVert u\right\rVert_{L^{\infty}_{t}{H^{-1}_{\kappa}}}^{2}+\left\lVert u\right\rVert_{\textup{LS}_{\kappa}}^{2}\big{)}\Big{)}

uniformly for x0x_{0}\in\mathbb{R}.

Proof.

We need κj2\kappa j_{2} to be O(1)O(1), so there must be some cancellation for the terms that are O(κ)O(\kappa) and higher. In order to exhibit this cancellation, we use the identities (cf. [Bringmann2021]*Lem. 2.5)

16κ5h2=3u23κ2(h1′′)220κ4[(h1)2(h12)′′]+4κ42R0(2κ)[(h1)2+2(h12)′′],\displaystyle 16\kappa^{5}h_{2}=3u^{2}-3\kappa^{2}(h^{\prime\prime}_{1})^{2}-20\kappa^{4}\big{[}(h^{\prime}_{1})^{2}-(h_{1}^{2})^{\prime\prime}\big{]}+4\kappa^{4}\partial^{2}R_{0}(2\kappa)\big{[}(h^{\prime}_{1})^{2}+2(h_{1}^{2})^{\prime\prime}\big{]},
(4.9) 16κ5h1+4κ2u=4κ3h1′′\displaystyle 16\kappa^{5}h_{1}+4\kappa^{2}u=4\kappa^{3}h^{\prime\prime}_{1}

to write

2κj2=3κ2(h1′′)2+12κ4(h1)23[u24κ2R0(2κ)u2]4κ42R0(2κ)[(h1)2+2(h12)′′].\displaystyle-2\kappa j_{2}=3\kappa^{2}(h^{\prime\prime}_{1})^{2}+12\kappa^{4}(h^{\prime}_{1})^{2}-3\big{[}u^{2}-4\kappa^{2}R_{0}(2\kappa)u^{2}\big{]}-4\kappa^{4}\partial^{2}R_{0}(2\kappa)\big{[}(h^{\prime}_{1})^{2}+2(h_{1}^{2})^{\prime\prime}\big{]}.

We multiply this by ψx02\psi^{2}_{x_{0}} and integrate in space and time. Working from left to right, we claim that

(4.10) TTψ2,3κ2(h1′′)2𝑑t=3TT(ψu),R02(ψu)𝑑t+O(κ2uLSκ2),\displaystyle\begin{aligned} \int_{-T}^{T}\langle\psi^{2},3\kappa^{2}(h^{\prime\prime}_{1})^{2}\rangle\,dt={}&3\int_{-T}^{T}\langle(\psi u^{\prime})^{\prime},R_{0}^{2}(\psi u^{\prime})^{\prime}\rangle\,dt+O\big{(}\kappa^{-2}\left\lVert u\right\rVert_{\textup{LS}_{\kappa}}^{2}\big{)},\end{aligned}
(4.11) TTψ2,12κ4(h1)2𝑑t=3TTψu,R0ψu(ψu),R02(ψu)dt+O(κ2uLSκ2),\displaystyle\begin{aligned} \int_{-T}^{T}\langle\psi^{2},12\kappa^{4}(h^{\prime}_{1})^{2}\rangle\,dt={}&3\int_{-T}^{T}\langle\psi u^{\prime},R_{0}\psi u^{\prime}\rangle-\langle(\psi u^{\prime})^{\prime},R_{0}^{2}(\psi u^{\prime})^{\prime}\rangle\,dt\\ &+O\big{(}\kappa^{-2}\left\lVert u\right\rVert_{\textup{LS}_{\kappa}}^{2}\big{)},\end{aligned}
(4.12) TTψ2,(rest of 2κj2)𝑑t=O(κ2(uLtHκ12+uLSκ2))\displaystyle\int_{-T}^{T}\langle\psi^{2},(\textrm{rest of }-2\kappa j_{2})\rangle\,dt=O\Big{(}\kappa^{-2}\big{(}\left\lVert u\right\rVert_{L^{\infty}_{t}H^{-1}_{\kappa}}^{2}+\left\lVert u\right\rVert_{\textup{LS}_{\kappa}}^{2}\big{)}\Big{)}

uniformly for x0x_{0}\in\mathbb{R}, where ψ=ψx0\psi=\psi_{x_{0}} and R0=R0(2κ)R_{0}=R_{0}(2\kappa). Adding these together, this would yield

TT2κψ2,j2dt\displaystyle\int_{-T}^{T}-2\kappa\langle\psi^{2},j_{2}\rangle\,dt =3TTψu,R0ψu𝑑t+O(κ12(uLtHκ12+uLSκ2)).\displaystyle=3\int_{-T}^{T}\langle\psi u^{\prime},R_{0}\psi u^{\prime}\rangle\,dt+O\Big{(}\kappa^{-\frac{1}{2}}\big{(}\left\lVert u\right\rVert_{L^{\infty}_{t}H^{-1}_{\kappa}}^{2}+\left\lVert u\right\rVert_{\textup{LS}_{\kappa}}^{2}\big{)}\Big{)}.

The first term on the RHS is exactly ψuLt2Hκ12\lVert\psi u^{\prime}\rVert_{L^{2}_{t}H^{-1}_{\kappa}}^{2}, and so this would finish the proof.

Let us start with  (4.10). As h1=κ1R0(2κ)uh_{1}=-\kappa^{-1}R_{0}(2\kappa)u, we have

κ2ψ2,(h1′′)2=ψR0u′′,ψR0u′′.\kappa^{2}\langle\psi^{2},(h^{\prime\prime}_{1})^{2}\rangle=\langle\psi R_{0}u^{\prime\prime},\psi R_{0}u^{\prime\prime}\rangle.

It remains to replace ψR0\psi R_{0}\partial by R0ψR_{0}\partial\psi above: by (2.19) we have

TT|ψR0u′′,ψR0u′′R0(ψu),R0(ψu)|𝑑t\displaystyle\int_{-T}^{T}\big{|}\langle\psi R_{0}u^{\prime\prime},\psi R_{0}u^{\prime\prime}\rangle-\langle R_{0}(\psi u^{\prime})^{\prime},R_{0}(\psi u^{\prime})^{\prime}\rangle\big{|}\,dt
=TT|ψu,1ψ(R0ψ2R0ψR022ψ)1ψψu|𝑑t\displaystyle\qquad=\int_{-T}^{T}\big{|}\langle\psi u^{\prime},\tfrac{1}{\psi}\big{(}R_{0}\partial\psi^{2}R_{0}\partial-\psi R_{0}^{2}\partial^{2}\psi\big{)}\tfrac{1}{\psi}\cdot\psi u^{\prime}\rangle\big{|}\,dt
ψuLt2Hκ22\displaystyle\qquad\lesssim\|\psi u^{\prime}\|_{L^{2}_{t}H^{-2}_{\kappa}}^{2}
κ2uLSκ2.\displaystyle\qquad\lesssim\kappa^{-2}\left\lVert u\right\rVert_{\textup{LS}_{\kappa}}^{2}.

This proves (4.10).

The proof of (4.11) proceeds in a similar way. We write

ψ2,12κ4(h1)2=12κ2ψR0u,ψR0u.\langle\psi^{2},12\kappa^{4}(h^{\prime}_{1})^{2}\rangle=12\kappa^{2}\langle\psi R_{0}u^{\prime},\psi R_{0}u^{\prime}\rangle.

First, we use (2.18) to replace ψR0\psi R_{0} by R0ψR_{0}\psi above:

12κ2TT|ψR0u,ψR0uR0ψu,R0ψu|𝑑t\displaystyle 12\kappa^{2}\int_{-T}^{T}\big{|}\langle\psi R_{0}u^{\prime},\psi R_{0}u^{\prime}\rangle-\langle R_{0}\psi u^{\prime},R_{0}\psi u^{\prime}\rangle\big{|}\,dt
=12κ2TT|ψu,1ψ(R0ψ2R0ψR02ψ)1ψψu|𝑑t\displaystyle\qquad=12\kappa^{2}\int_{-T}^{T}\big{|}\langle\psi u^{\prime},\tfrac{1}{\psi}\big{(}R_{0}\psi^{2}R_{0}-\psi R_{0}^{2}\psi\big{)}\tfrac{1}{\psi}\cdot\psi u^{\prime}\rangle\big{|}\,dt
ψuLt2Hκ22\displaystyle\qquad\lesssim\left\lVert\psi u^{\prime}\right\rVert_{L^{2}_{t}H^{-2}_{\kappa}}^{2}
κ2uLSκ2.\displaystyle\qquad\lesssim\kappa^{-2}\left\lVert u\right\rVert_{\textup{LS}_{\kappa}}^{2}.

Next we use the identity 4κ2R0=1+2R04\kappa^{2}R_{0}=1+\partial^{2}R_{0} to write

12κ2R0ψu,R0ψu=3ψu,R0ψu3(ψu),R02(ψu),12\kappa^{2}\langle R_{0}\psi u^{\prime},R_{0}\psi u^{\prime}\rangle=3\langle\psi u^{\prime},R_{0}\psi u^{\prime}\rangle-3\langle(\psi u^{\prime})^{\prime},R_{0}^{2}(\psi u^{\prime})^{\prime}\rangle,

and (4.11) follows.

Lastly, we turn to (4.12). There are two terms remaining in 2κj2-2\kappa j_{2} which we want to show make a negligible contribution. For the first term, we use the identity 14κ2R0=2R01-4\kappa^{2}R_{0}=-\partial^{2}R_{0} to write

ψ2,3[u24κ2R0u2]=3ψ2,R0(u2)′′=31ψ2R0(ψ2)′′,ψ2u2.\langle\psi^{2},3\big{[}u^{2}-4\kappa^{2}R_{0}u^{2}\big{]}\rangle=3\langle\psi^{2},R_{0}(u^{2})^{\prime\prime}\rangle=3\langle\tfrac{1}{\psi^{2}}R_{0}(\psi^{2})^{\prime\prime},\psi^{2}u^{2}\rangle.

Therefore, by (2.17) and (3.2)–(3.4) we have for Tκ2T\leq\kappa^{-2}

TT|ψ2,3[u24κ2R0u2]|𝑑t\displaystyle\int_{-T}^{T}\big{|}\langle\psi^{2},3\big{[}u^{2}-4\kappa^{2}R_{0}u^{2}\big{]}\rangle\big{|}\,dt 31ψ2R0(ψ2)′′LxψuLt,x22\displaystyle\leq 3\big{\lVert}\tfrac{1}{\psi^{2}}R_{0}(\psi^{2})^{\prime\prime}\big{\rVert}_{L^{\infty}_{x}}\left\lVert\psi u\right\rVert_{L^{2}_{t,x}}^{2}
κ2(2+4κ2)(ψu)Lt2Hκ22\displaystyle\lesssim\kappa^{-2}\big{\lVert}(-\partial^{2}+4\kappa^{2})(\psi u)\big{\rVert}^{2}_{L^{2}_{t}H^{-2}_{\kappa}}
κ2(uLtHκ12+uLSκ2).\displaystyle\lesssim\kappa^{-2}\big{(}\left\lVert u\right\rVert_{L^{\infty}_{t}{H^{-1}_{\kappa}}}^{2}+\left\lVert u\right\rVert_{\textup{LS}_{\kappa}}^{2}\big{)}.

For the last term of 2κj2-2\kappa j_{2}, we write

ψ2,4κ42R0[(h1)2+2(h12)′′]=4κ2R0(ψ2)′′,(R0u)2+4κ2R0(ψ2)(4),(R0u)2,\langle\psi^{2},4\kappa^{4}\partial^{2}R_{0}[(h^{\prime}_{1})^{2}+2(h_{1}^{2})^{\prime\prime}\big{]}\rangle=4\kappa^{2}\langle R_{0}(\psi^{2})^{\prime\prime},(R_{0}u^{\prime})^{2}\rangle+4\kappa^{2}\langle R_{0}(\psi^{2})^{(4)},(R_{0}u)^{2}\rangle,

in order to estimate

TT|ψ2,4κ42R0[(h1)2+2(h12)′′]|𝑑t\displaystyle\int_{-T}^{T}\big{|}\langle\psi^{2},4\kappa^{4}\partial^{2}R_{0}[(h^{\prime}_{1})^{2}+2(h_{1}^{2})^{\prime\prime}\big{]}\rangle\big{|}\,dt
κ2{1ψ2R0(ψ2)′′LxψR01ψψuLt,x22+1ψ2R0(ψ2)(4)LxψR01ψψuLt,x22}\displaystyle\qquad\lesssim\kappa^{2}\Big{\{}\big{\lVert}\tfrac{1}{\psi^{2}}R_{0}(\psi^{2})^{\prime\prime}\big{\rVert}_{L^{\infty}_{x}}\big{\lVert}\psi R_{0}\tfrac{1}{\psi}\cdot\psi u^{\prime}\big{\rVert}_{L^{2}_{t,x}}^{2}+\big{\lVert}\tfrac{1}{\psi^{2}}R_{0}(\psi^{2})^{(4)}\big{\rVert}_{L^{\infty}_{x}}\big{\lVert}\psi R_{0}\tfrac{1}{\psi}\cdot\psi u\big{\rVert}_{L^{2}_{t,x}}^{2}\Big{\}}
κ2{κ2ψuLt2Hκ22+κ2ψuLt2Hκ22}κ2(uLtHκ12+uLSκ2).\displaystyle\qquad\lesssim\kappa^{2}\big{\{}\kappa^{-2}\left\lVert\psi u^{\prime}\right\rVert_{L^{2}_{t}H^{-2}_{\kappa}}^{2}+\kappa^{-2}\left\lVert\psi u\right\rVert_{L^{2}_{t}H^{-2}_{\kappa}}^{2}\big{\}}\lesssim\kappa^{-2}\big{(}\left\lVert u\right\rVert_{L^{\infty}_{t}{H^{-1}_{\kappa}}}^{2}+\left\lVert u\right\rVert_{\textup{LS}_{\kappa}}^{2}\big{)}.

Altogether, this proves (4.12). ∎

In the next lemma we show that the cubic and higher order terms of jj make a negligible contribution:

Lemma 4.3.

Assume that κ1\kappa\geq 1 and uu satisfies (3.9) uniformly in [0,T][0,T] with Tκ2T\leq\kappa^{-2}. Then we have

(4.13) |TTψx02(jj2)𝑑x𝑑t|κ1(Tκ2)14(uLtHκ12+uLSκ2)\bigg{|}\int_{-T}^{T}\int_{-\infty}^{\infty}\psi^{2}_{x_{0}}(j-j_{2})\,dx\,dt\bigg{|}\lesssim\kappa^{-1}(T\kappa^{2})^{\frac{1}{4}}\big{(}\left\lVert u\right\rVert_{L^{\infty}_{t}{H^{-1}_{\kappa}}}^{2}+\left\lVert u\right\rVert_{\textup{LS}_{\kappa}}^{2}\!\big{)}

uniformly in x0x_{0}\in\mathbb{R}.

Proof.

In order to exhibit cancellation in jj2j-j_{2}, we expand gg as the series (2.8) in the expression (2.12) for jj:

jj2\displaystyle j-j_{2}
(4.14) =1g(4κ3h1+u)+2κR0(2κ)u′′+h1(16κ5h1+4κ2u)\displaystyle={}\tfrac{1}{g}(4\kappa^{3}h_{1}+u)+2\kappa R_{0}(2\kappa)u^{\prime\prime}+h_{1}(16\kappa^{5}h_{1}+4\kappa^{2}u)
(4.15) +(1g2κ)4κ3h2\displaystyle\qquad+(\tfrac{1}{g}-2\kappa)4\kappa^{3}h_{2}
(4.16) +4κ31g3h.\displaystyle\qquad+4\kappa^{3}\tfrac{1}{g}\sum_{\ell\geq 3}h_{\ell}.

We will estimate the contributions (4.14)–(4.16) one at a time.

Let us start with (4.14). Using the identity (4.9), we write

(4.14)\displaystyle\eqref{j3 1} =κ(1g2κ+4κ2h1)h1′′\displaystyle=\kappa\big{(}\tfrac{1}{g}-2\kappa+4\kappa^{2}h_{1}\big{)}h^{\prime\prime}_{1}
=κ(1g2κ+4κ2h11+2κh1)h1′′+8κ41+2κh1h12h1′′.\displaystyle=\kappa\big{(}\tfrac{1}{g}-2\kappa+\tfrac{4\kappa^{2}h_{1}}{1+2\kappa h_{1}}\big{)}h^{\prime\prime}_{1}+\tfrac{8\kappa^{4}}{1+2\kappa h_{1}}h_{1}^{2}h^{\prime\prime}_{1}.

For the first term on the RHS, we use (3.21) and (3.12) to bound:

|TTψ2κ(1g2κ+4κ2h11+2κh1)h1′′𝑑x𝑑t|\displaystyle\bigg{|}\int_{-T}^{T}\int_{-\infty}^{\infty}\psi^{2}\kappa\big{(}\tfrac{1}{g}-2\kappa+\tfrac{4\kappa^{2}h_{1}}{1+2\kappa h_{1}}\big{)}h^{\prime\prime}_{1}\,dx\,dt\bigg{|}
κψ(1g2κ+4κ2h11+2κh1)Lt,x2ψh1′′Lt,x2\displaystyle\qquad\lesssim\kappa\big{\lVert}\psi\big{(}\tfrac{1}{g}-2\kappa+\tfrac{4\kappa^{2}h_{1}}{1+2\kappa h_{1}}\big{)}\big{\rVert}_{L^{2}_{t,x}}\left\lVert\psi h^{\prime\prime}_{1}\right\rVert_{L^{2}_{t,x}}
κ32(Tκ2)14uLtHκ1(uLtHκ12+uLSκ2).\displaystyle\qquad\lesssim\kappa^{-\frac{3}{2}}(T\kappa^{2})^{\frac{1}{4}}\left\lVert u\right\rVert_{L^{\infty}_{t}{H^{-1}_{\kappa}}}\big{(}\left\lVert u\right\rVert_{L^{\infty}_{t}{H^{-1}_{\kappa}}}^{2}+\left\lVert u\right\rVert_{\textup{LS}_{\kappa}}^{2}\big{)}.

For the second term, we use (3.22), (3.10), and (3.12) to bound:

|TTψ28κ41+2κh1h12h1′′𝑑x𝑑t|\displaystyle\bigg{|}\int_{-T}^{T}\int_{-\infty}^{\infty}\psi^{2}\tfrac{8\kappa^{4}}{1+2\kappa h_{1}}h_{1}^{2}h^{\prime\prime}_{1}\,dx\,dt\bigg{|} κ7211+2κh1Lt,xh1LtHκ1ψh1Lt,x2ψh1′′Lt,x2\displaystyle\lesssim\kappa^{\frac{7}{2}}\big{\lVert}\tfrac{1}{1+2\kappa h_{1}}\big{\rVert}_{L^{\infty}_{t,x}}\left\lVert h_{1}\right\rVert_{L^{\infty}_{t}H^{1}_{\kappa}}\left\lVert\psi h_{1}\right\rVert_{L^{2}_{t,x}}\left\lVert\psi h^{\prime\prime}_{1}\right\rVert_{L^{2}_{t,x}}
κ12T12uLtHκ12(uLtHκ1+uLSκ)\displaystyle\lesssim\kappa^{-\frac{1}{2}}T^{\frac{1}{2}}\left\lVert u\right\rVert_{L^{\infty}_{t}H^{-1}_{\kappa}}^{2}\big{(}\left\lVert u\right\rVert_{L^{\infty}_{t}H^{-1}_{\kappa}}+\left\lVert u\right\rVert_{\textup{LS}_{\kappa}}\big{)}
κ32(Tκ2)12uLtHκ1(uLtHκ12+uLSκ2).\displaystyle\lesssim\kappa^{-\frac{3}{2}}(T\kappa^{2})^{\frac{1}{2}}\left\lVert u\right\rVert_{L^{\infty}_{t}H^{-1}_{\kappa}}\big{(}\left\lVert u\right\rVert_{L^{\infty}_{t}H^{-1}_{\kappa}}^{2}+\left\lVert u\right\rVert_{\textup{LS}_{\kappa}}^{2}\big{)}.

Next, we turn to (4.15). We write

(4.15)=4κ3(1g2κ+4κ2h11+2κh1)h216κ51+2κh1h1h2.\eqref{j3 2}=4\kappa^{3}\big{(}\tfrac{1}{g}-2\kappa+\tfrac{4\kappa^{2}h_{1}}{1+2\kappa h_{1}}\big{)}h_{2}-\tfrac{16\kappa^{5}}{1+2\kappa h_{1}}h_{1}h_{2}.

For the first term on the RHS, we use (3.21) and (3.20) to estimate

|TTψ24κ3(1g2κ+4κ2h11+2κh1)h2𝑑x𝑑t|\displaystyle\bigg{|}\int_{-T}^{T}\int_{-\infty}^{\infty}\psi^{2}4\kappa^{3}\big{(}\tfrac{1}{g}-2\kappa+\tfrac{4\kappa^{2}h_{1}}{1+2\kappa h_{1}}\big{)}h_{2}\,dx\,dt\bigg{|}
κ3ψ(1g2κ+4κ2h11+2κh1)Lt,x2ψh2Lt,x2\displaystyle\qquad\lesssim\kappa^{3}\big{\lVert}\psi\big{(}\tfrac{1}{g}-2\kappa+\tfrac{4\kappa^{2}h_{1}}{1+2\kappa h_{1}}\big{)}\big{\rVert}_{L^{2}_{t,x}}\left\lVert\psi h_{2}\right\rVert_{L^{2}_{t,x}}
κ2(Tκ2)12uLtHκ12(uLtHκ12+uLSκ2)\displaystyle\qquad\lesssim\kappa^{-2}(T\kappa^{2})^{\frac{1}{2}}\left\lVert u\right\rVert_{L^{\infty}_{t}{H^{-1}_{\kappa}}}^{2}\big{(}\left\lVert u\right\rVert_{L^{\infty}_{t}{H^{-1}_{\kappa}}}^{2}+\left\lVert u\right\rVert_{\textup{LS}_{\kappa}}^{2}\big{)}
κ32(Tκ2)12uLtHκ1(uLtHκ12+uLSκ2).\displaystyle\qquad{}\lesssim\kappa^{-\frac{3}{2}}(T\kappa^{2})^{\frac{1}{2}}\left\lVert u\right\rVert_{L^{\infty}_{t}{H^{-1}_{\kappa}}}\big{(}\left\lVert u\right\rVert_{L^{\infty}_{t}{H^{-1}_{\kappa}}}^{2}+\left\lVert u\right\rVert_{\textup{LS}_{\kappa}}^{2}\big{)}.

For the second term, we use (3.23) to bound

|TTψ216κ51+2κh1h1h2𝑑x𝑑t|\displaystyle\bigg{|}\int_{-T}^{T}\int_{-\infty}^{\infty}\psi^{2}\tfrac{16\kappa^{5}}{1+2\kappa h_{1}}h_{1}h_{2}\,dx\,dt\bigg{|} κ511+2κh1Lt,xψ2h1h2Lt,x1\displaystyle\lesssim\kappa^{5}\big{\lVert}\tfrac{1}{1+2\kappa h_{1}}\big{\rVert}_{L^{\infty}_{t,x}}\left\lVert\psi^{2}h_{1}h_{2}\right\rVert_{L^{1}_{t,x}}
κ32(Tκ2)34uLtHκ1(uLtHκ12+uLSκ2).\displaystyle\lesssim\kappa^{-\frac{3}{2}}(T\kappa^{2})^{\frac{3}{4}}\left\lVert u\right\rVert_{L^{\infty}_{t}{H^{-1}_{\kappa}}}\big{(}\left\lVert u\right\rVert_{L^{\infty}_{t}{H^{-1}_{\kappa}}}^{2}+\left\lVert u\right\rVert_{\textup{LS}_{\kappa}}^{2}\big{)}.

Finally, we turn to (4.16). Using (3.13), we estimate

|TTψ2(4.16)𝑑x𝑑t|\displaystyle\bigg{|}\int_{-T}^{T}\int_{-\infty}^{\infty}\psi^{2}\eqref{j3 3}\,dx\,dt\bigg{|} κ31gLt,x3ψ2hLt,x1\displaystyle\lesssim\kappa^{3}\big{\lVert}\tfrac{1}{g}\big{\rVert}_{L^{\infty}_{t,x}}\sum_{\ell\geq 3}\left\lVert\psi^{2}h_{\ell}\right\rVert_{L^{1}_{t,x}}
κ32(Tκ2)34uLtHκ1(uLtHκ12+uLSκ2).\displaystyle\lesssim\kappa^{-\frac{3}{2}}(T\kappa^{2})^{\frac{3}{4}}\left\lVert u\right\rVert_{L^{\infty}_{t}{H^{-1}_{\kappa}}}\big{(}\left\lVert u\right\rVert_{L^{\infty}_{t}{H^{-1}_{\kappa}}}^{2}+\left\lVert u\right\rVert_{\textup{LS}_{\kappa}}^{2}\big{)}.

Altogether, we obtain

|TTψx02(jj2)𝑑x𝑑t|κ32(Tκ2)14uLtHκ1(uLtHκ12+uLSκ2).\bigg{|}\int_{-T}^{T}\int_{-\infty}^{\infty}\psi^{2}_{x_{0}}(j-j_{2})\,dx\,dt\bigg{|}\lesssim\kappa^{-\frac{3}{2}}(T\kappa^{2})^{\frac{1}{4}}\left\lVert u\right\rVert_{L^{\infty}_{t}{H^{-1}_{\kappa}}}\big{(}\left\lVert u\right\rVert_{L^{\infty}_{t}{H^{-1}_{\kappa}}}^{2}+\left\lVert u\right\rVert_{\textup{LS}_{\kappa}}^{2}\!\big{)}.

This implies (4.13) by (3.9), and thus finishes the proof of the lemma. ∎

The last step in the proof of the theorem is to prove the estimate (4.8), which shows that the contributions of the source terms in the density flux relation (2.13) are perturbative:

Lemma 4.4.

Assume that κ1\kappa\geq 1 and uu satisfies (3.9) uniformly in [0,T][0,T], and that the coefficients a1,a2,a3,a4a_{1},a_{2},a_{3},a_{4} in (gKdV) satisfy (4.1)–(4.4). Then

(4.17) |TTϕx0dρ|u[(a1u)+a2u2+a3u+a4u]dxdt|κ1(ε+κ2)(uLtHκ12+uLSκ2)\bigg{|}\int_{-T}^{T}\int\phi_{x_{0}}\,d\rho|_{u}\big{[}(a_{1}u^{\prime})^{\prime}+a_{2}u^{2}+a_{3}u^{\prime}+a_{4}u\big{]}\,dx\,dt\bigg{|}\lesssim\kappa^{-1}(\varepsilon+\kappa^{-2})\big{(}\left\lVert u\right\rVert_{L^{\infty}_{t}H^{-1}_{\kappa}}^{2}+\left\lVert u\right\rVert_{\textup{LS}_{\kappa}}^{2}\big{)}

uniformly in x0x_{0}\in\mathbb{R}.

Proof.

We begin with the contribution of (a1u)(a_{1}u^{\prime})^{\prime}. Write

(4.18) a1(t,x)=ca1(t,x)ψz3(x)𝑑za_{1}(t,x)=c\int a_{1}(t,x)\psi_{z}^{3}(x)\,dz

for a constant cc. By (4.1) we have

ψz2a1uLt2Hκ1𝑑zψza1LtXψzuLt2Hκ1𝑑zεuLSκ,\int\left\lVert\psi_{z}^{2}a_{1}u^{\prime}\right\rVert_{L^{2}_{t}H^{-1}_{\kappa}}\,dz\lesssim\int\left\lVert\psi_{z}a_{1}\right\rVert_{L^{\infty}_{t}X}\left\lVert\psi_{z}u^{\prime}\right\rVert_{L^{2}_{t}H^{-1}_{\kappa}}\,dz\lesssim\varepsilon\left\lVert u\right\rVert_{\textup{LS}_{\kappa}},

where X=H1X=H^{1} or X=W1,X=W^{1,\infty}. Therefore, by (3.25) we have

|TTϕdρ|u(a1u)dxdt|εκ1(uLtHκ1+uLSκ)uLSκ,\displaystyle\bigg{|}\int_{-T}^{T}\int\phi\,d\rho|_{u}(a_{1}u^{\prime})^{\prime}\,dx\,dt\bigg{|}\lesssim\varepsilon\kappa^{-1}\big{(}\left\lVert u\right\rVert_{L^{\infty}_{t}{H^{-1}_{\kappa}}}+\left\lVert u\right\rVert_{\textup{LS}_{\kappa}}\big{)}\left\lVert u\right\rVert_{\textup{LS}_{\kappa}},

and this is an acceptable contribution to (4.17).

For the contributions of a2u2a_{2}u^{2}, a3ua_{3}u^{\prime}, and a4ua_{4}u, we use (3.24) to estimate

(4.19) |TTϕdρ|u(a2u2+a3u+a4u)dxdt|\displaystyle\bigg{|}\int_{-T}^{T}\int\phi\,d\rho|_{u}(a_{2}u^{2}+a_{3}u^{\prime}+a_{4}u)\,dx\,dt\bigg{|}
κ1uLtHκ1(a2u2Lt1Hκ1+a3uLt1Hκ1+a4uLt1Hκ1).\displaystyle\qquad\lesssim\kappa^{-1}\left\lVert u\right\rVert_{L^{\infty}_{t}{H^{-1}_{\kappa}}}\big{(}\lVert a_{2}u^{2}\rVert_{L^{1}_{t}H^{-1}_{\kappa}}+\lVert a_{3}u^{\prime}\rVert_{L^{1}_{t}H^{-1}_{\kappa}}+\lVert a_{4}u\rVert_{L^{1}_{t}H^{-1}_{\kappa}}\big{)}.

For a4ua_{4}u, we use (4.4) and (3.9) to bound

a4uLt1Hκ1Ta4LtXuLtHκ1κ2uLtHκ1,\left\lVert a_{4}u\right\rVert_{L^{1}_{t}H^{-1}_{\kappa}}\lesssim T\left\lVert a_{4}\right\rVert_{L^{\infty}_{t}X}\left\lVert u\right\rVert_{L^{\infty}_{t}{H^{-1}_{\kappa}}}\lesssim\kappa^{-2}\left\lVert u\right\rVert_{L^{\infty}_{t}{H^{-1}_{\kappa}}},

where X=H1X=H^{1} or X=W1,X=W^{1,\infty}.

For a3ua_{3}u^{\prime}, we use (4.18) and (4.3) to bound

a3uLt1Hκ1ψz2a3uLt1Hκ1𝑑zuLSκψza3Lt2X𝑑zεuLSκ,\lVert a_{3}u^{\prime}\rVert_{L^{1}_{t}H^{-1}_{\kappa}}\lesssim\int\lVert\psi^{2}_{z}a_{3}u^{\prime}\rVert_{L^{1}_{t}H^{-1}_{\kappa}}\,dz\lesssim\left\lVert u\right\rVert_{\textup{LS}_{\kappa}}\int\big{\lVert}\psi_{z}a_{3}\big{\rVert}_{L^{2}_{t}X}\,dz\lesssim\varepsilon\left\lVert u\right\rVert_{\textup{LS}_{\kappa}},

where X=H1X=H^{1} or X=W1,X=W^{1,\infty}.

Finally, for a2u2a_{2}u^{2}, we use (3.2)–(3.4) to bound

(4.20) ψzuLt,x2uLSκ+κuLt2Hκ1uLSκ+uLtHκ1,\left\lVert\psi_{z}u\right\rVert_{L^{2}_{t,x}}\lesssim\left\lVert u\right\rVert_{\textup{LS}_{\kappa}}+\kappa\left\lVert u\right\rVert_{{L^{2}_{t}}H^{-1}_{\kappa}}\lesssim\left\lVert u\right\rVert_{\textup{LS}_{\kappa}}+\left\lVert u\right\rVert_{L^{\infty}_{t}H^{-1}_{\kappa}},

provided that Tκ2T\leq\kappa^{-2}. Using (4.18), the embedding L1Hκ1L^{1}\hookrightarrow H^{-1}_{\kappa}, and (4.2), we see that

a2u2Lt1Hκ1\displaystyle\left\lVert a_{2}u^{2}\right\rVert_{L^{1}_{t}H^{-1}_{\kappa}} ψz3a2u2Lt1Hκ1𝑑zκ12ψz3a2u2Lt,x1𝑑z\displaystyle\lesssim\int\left\lVert\psi_{z}^{3}a_{2}u^{2}\right\rVert_{L^{1}_{t}H^{-1}_{\kappa}}\,dz\lesssim\kappa^{-\frac{1}{2}}\int\left\lVert\psi_{z}^{3}a_{2}u^{2}\right\rVert_{L^{1}_{t,x}}\!\,dz
κ12ψzuLt,x22ψza2Lt,x𝑑zεκ12(uLtHκ12+uLSκ2).\displaystyle\lesssim\kappa^{-\frac{1}{2}}\int\left\lVert\psi_{z}u\right\rVert_{L^{2}_{t,x}}^{2}\left\lVert\psi_{z}a_{2}\right\rVert_{L^{\infty}_{t,x}}\,dz\lesssim\varepsilon\kappa^{-\frac{1}{2}}\big{(}\left\lVert u\right\rVert_{L^{\infty}_{t}H^{-1}_{\kappa}}^{2}+\left\lVert u\right\rVert_{\textup{LS}_{\kappa}}^{2}\big{)}.

Returning to (4.19), we conclude

|TTϕdρ|u(a2u2+a3u+a4u)dxdt|\displaystyle\bigg{|}\int_{-T}^{T}\int\phi\,d\rho|_{u}(a_{2}u^{2}+a_{3}u^{\prime}+a_{4}u)\,dx\,dt\bigg{|}
κ1(ε+κ2)(uLtHκ12+uLSκ2)+εκ32uLtHκ1(uLSκ2+uLtHκ12).\displaystyle\qquad\lesssim\kappa^{-1}(\varepsilon+\kappa^{-2})\big{(}\left\lVert u\right\rVert_{L^{\infty}_{t}H^{-1}_{\kappa}}^{2}+\left\lVert u\right\rVert_{\textup{LS}_{\kappa}}^{2}\big{)}+\varepsilon\kappa^{-\frac{3}{2}}\left\lVert u\right\rVert_{L^{\infty}_{t}H^{-1}_{\kappa}}\big{(}\left\lVert u\right\rVert_{\textup{LS}_{\kappa}}^{2}+\left\lVert u\right\rVert_{L^{\infty}_{t}H^{-1}_{\kappa}}^{2}\big{)}.

This implies (4.17) by (3.9), and thus completes the proof.

The last three lemmas complete the proof of Theorem 4.1.

5. The a-priori estimates

In this section, we will prove the energy estimate for solutions to (gKdV), and use it to conclude the proof of our main results in Theorem 1.1 and Theorem 1.5.

Proposition 5.1.

For any ε(0,1]\varepsilon\in(0,1], if the coefficients a1,a2,a3,a4a_{1},a_{2},a_{3},a_{4} of (gKdV) satisfy (4.1)–(4.4), then any solution u(t)u(t) to (gKdV) satisfies

(5.1) uCtHκ12\displaystyle\left\lVert u\right\rVert_{C_{t}H^{-1}_{\kappa}}^{2} u(0)Hκ12+(ε+T)(uLtHκ12+uLSκ2)\displaystyle\lesssim\left\lVert u(0)\right\rVert_{H^{-1}_{\kappa}}^{2}+(\varepsilon+T)\big{(}\left\lVert u\right\rVert_{L^{\infty}_{t}H^{-1}_{\kappa}}^{2}+\left\lVert u\right\rVert_{\textup{LS}_{\kappa}}^{2}\big{)}

uniformly for Tκ2T\leq\kappa^{-2} and κ\kappa satisfying (3.9) in [0,T][0,T].

Proof.

Integrating (2.13) in space and using (2.14), we obtain

tα=[12g2dg|u+2κR0(2κ)][(a1u)+a2u2+a3u+a4u]𝑑y.\partial_{t}\alpha=\int\big{[}\tfrac{1}{2g^{2}}dg|_{u}+2\kappa R_{0}(2\kappa)\big{]}\big{[}(a_{1}u^{\prime})^{\prime}+a_{2}u^{2}+a_{3}u^{\prime}+a_{4}u\big{]}\,dy.

Next, we use (2.15) and the identity

G(x,y)G(y,x)2g(y)2𝑑y=g(x)\int\frac{G(x,y)G(y,x)}{2g(y)^{2}}\,dy=g(x)

(see [Killip2019]*Lem. 2.5 for a proof) to write

tα=(g12κ)[(a1u)+a2u2+a3u+a4u]𝑑x.\partial_{t}\alpha=-\int(g-\tfrac{1}{2\kappa})\big{[}(a_{1}u^{\prime})^{\prime}+a_{2}u^{2}+a_{3}u^{\prime}+a_{4}u\big{]}\,dx.

By (2.10) and the fundamental theorem of calculus, this yields

(5.2) uCtHκ12u(0)Hκ12+κTT|(g12κ)[(a1u)+a2u2+a3u+a4u]𝑑x|𝑑t.\left\lVert u\right\rVert_{C_{t}H^{-1}_{\kappa}}^{2}\lesssim\left\lVert u(0)\right\rVert_{H^{-1}_{\kappa}}^{2}+\kappa\int_{-T}^{T}\bigg{|}\int(g-\tfrac{1}{2\kappa})\big{[}(a_{1}u^{\prime})^{\prime}+a_{2}u^{2}+a_{3}u^{\prime}+a_{4}u\big{]}\,dx\bigg{|}\,dt.

We will successively estimate the contribution of each of the terms inside the integral.

We begin with the contribution of a4a_{4}. Using (2.6) and (4.4), we have

TT|(g12κ)a4u𝑑x|𝑑t\displaystyle\int_{-T}^{T}\bigg{|}\int(g-\tfrac{1}{2\kappa})a_{4}u\,dx\bigg{|}\,dt Tg12κLtHκ1a4uLtHκ1\displaystyle\leq T\left\lVert g-\tfrac{1}{2\kappa}\right\rVert_{L^{\infty}_{t}H^{1}_{\kappa}}\left\lVert a_{4}u\right\rVert_{L^{\infty}_{t}H^{-1}_{\kappa}}
κ1Ta4LtXuLtHκ12\displaystyle\lesssim\kappa^{-1}T\left\lVert a_{4}\right\rVert_{L^{\infty}_{t}X}\left\lVert u\right\rVert_{L^{\infty}_{t}H^{-1}_{\kappa}}^{2}
κ1TuLtHκ12,\displaystyle\lesssim\kappa^{-1}T\left\lVert u\right\rVert_{L^{\infty}_{t}H^{-1}_{\kappa}}^{2},

where X=H1X=H^{1} or X=W1,X=W^{1,\infty}.

Next, we address a3a_{3}. Using the continuous partition of unity (4.18) and (4.3), we have

TT|(g12κ)a3u𝑑x|𝑑t\displaystyle\int_{-T}^{T}\bigg{|}\int(g-\tfrac{1}{2\kappa})a_{3}u^{\prime}\,dx\bigg{|}\,dt g12κLtHκ1a3uLt1Hκ1\displaystyle\leq\left\lVert g-\tfrac{1}{2\kappa}\right\rVert_{L^{\infty}_{t}H^{1}_{\kappa}}\left\lVert a_{3}u^{\prime}\right\rVert_{L^{1}_{t}H^{-1}_{\kappa}}
κ1uLtHκ1ψza3Lt2XψzuLt2Hκ1𝑑z\displaystyle\lesssim\kappa^{-1}\left\lVert u\right\rVert_{L^{\infty}_{t}H^{-1}_{\kappa}}\int\left\lVert\psi_{z}a_{3}\right\rVert_{L^{2}_{t}X}\left\lVert\psi_{z}u^{\prime}\right\rVert_{L^{2}_{t}H^{-1}_{\kappa}}\,dz
εκ1uLtHκ1uLSκ.\displaystyle\lesssim\varepsilon\kappa^{-1}\left\lVert u\right\rVert_{L^{\infty}_{t}H^{-1}_{\kappa}}\left\lVert u\right\rVert_{\textup{LS}_{\kappa}}.

For a2a_{2}, we use the embedding Hκ1LH^{1}_{\kappa}\hookrightarrow L^{\infty}, (4.18), and (4.20) to bound

TT|(g12κ)a2u2𝑑x|𝑑t\displaystyle\int_{-T}^{T}\bigg{|}\int(g-\tfrac{1}{2\kappa})a_{2}u^{2}\,dx\bigg{|}\,dt g12κLt,xbu2Lt,x1\displaystyle\leq\left\lVert g-\tfrac{1}{2\kappa}\right\rVert_{L^{\infty}_{t,x}}\left\lVert bu^{2}\right\rVert_{L^{1}_{t,x}}
κ32uLtHκ1ψza2Lt,xψzuLt,x22𝑑z\displaystyle\lesssim\kappa^{-\frac{3}{2}}\left\lVert u\right\rVert_{L^{\infty}_{t}H^{-1}_{\kappa}}\int\left\lVert\psi_{z}a_{2}\right\rVert_{L^{\infty}_{t,x}}\left\lVert\psi_{z}u\right\rVert_{L^{2}_{t,x}}^{2}\,dz
εκ32uLtHκ1(uLSκ2+uLtHκ12)\displaystyle\lesssim\varepsilon\kappa^{-\frac{3}{2}}\left\lVert u\right\rVert_{L^{\infty}_{t}H^{-1}_{\kappa}}\big{(}\left\lVert u\right\rVert_{\textup{LS}_{\kappa}}^{2}+\left\lVert u\right\rVert_{L^{\infty}_{t}H^{-1}_{\kappa}}^{2}\big{)}
εκ1(uLSκ2+uLtHκ12),\displaystyle\lesssim\varepsilon\kappa^{-1}\big{(}\left\lVert u\right\rVert_{\textup{LS}_{\kappa}}^{2}+\left\lVert u\right\rVert_{L^{\infty}_{t}H^{-1}_{\kappa}}^{2}\big{)},

using (3.9) at the last step.

Lastly, we turn to the contribution of a1a_{1}. We integrate by parts once in space, and then we use (4.18), (3.8), and (4.1) to estimate:

TT|(g12κ)(a1u)𝑑x|𝑑t\displaystyle\int_{-T}^{T}\bigg{|}\int(g-\tfrac{1}{2\kappa})(a_{1}u^{\prime})^{\prime}\,dx\bigg{|}\,dt ψzgLt2Hκ1ψzuLt2Hκ1ψza1LtX𝑑z\displaystyle\leq\int\left\lVert\psi_{z}g^{\prime}\right\rVert_{L^{2}_{t}H^{1}_{\kappa}}\left\lVert\psi_{z}u^{\prime}\right\rVert_{L^{2}_{t}H^{-1}_{\kappa}}\left\lVert\psi_{z}a_{1}\right\rVert_{L^{\infty}_{t}X}\,dz
εκ1uLSκ2.\displaystyle\lesssim\varepsilon\kappa^{-1}\left\lVert u\right\rVert_{\textup{LS}_{\kappa}}^{2}.

Altogether, returning to (5.2), we obtain

uCtHκ12\displaystyle\left\lVert u\right\rVert_{C_{t}H^{-1}_{\kappa}}^{2} u(0)Hκ12+(ε+T)(uLtHκ12+uLSκ2).\displaystyle\lesssim\left\lVert u(0)\right\rVert_{H^{-1}_{\kappa}}^{2}+(\varepsilon+T)\big{(}\left\lVert u\right\rVert_{L^{\infty}_{t}H^{-1}_{\kappa}}^{2}+\left\lVert u\right\rVert_{\textup{LS}_{\kappa}}^{2}\big{)}.\qed

At this point we are prepared to complete the proofs of Theorems 1.1 and 1.5, by combining Propositions 4.1 and 5.1 and a continuity argument. We restate the results together in a stronger form:

Theorem 5.2.

There exist ϵ0,c0>0\epsilon_{0},c_{0}>0 so that given coefficients a1,,a4a_{1},\dots,a_{4} in (gKdV) satisfying (4.1)–(4.4) with ε=ε0\varepsilon=\varepsilon_{0} and R1R\geq 1 the solutions u(t)u(t) to (gKdV) and initial data u0u_{0} satisfying

(5.3) u0H1R\|u_{0}\|_{H^{-1}}\leq R

exist up to time

(5.4) T0=c0R4T_{0}=c_{0}R^{-4}

and satisfy in [0,T0][0,T_{0}]

(5.5) uLtHκ1+uLSκR,κ=R2.\left\lVert u\right\rVert_{L^{\infty}_{t}H^{-1}_{\kappa}}+\left\lVert u\right\rVert_{\textup{LS}_{\kappa}}\lesssim R,\qquad\kappa=R^{2}.
Proof.

For any κR2\kappa\gg R^{2}, the condition (3.9) holds at t=0t=0 and thus on some time interval [0,T][0,T]. Let Tκ2T\leq\kappa^{-2} be any time for which the solution exists in [0,T][0,T] and satisfies (3.9) uniformly in [0,T][0,T]. Then both Propositions 4.1 and 5.1 apply in [0,T][0,T].

Consider the quantity

BT:=uCtHκ1([T,T]×)2+14uLSκ([T,T]×)2.B_{T}:=\left\lVert u\right\rVert_{C_{t}H^{-1}_{\kappa}([-T,T]\times\mathbb{R})}^{2}+\tfrac{1}{4}\left\lVert u\right\rVert_{\textup{LS}_{\kappa}([-T,T]\times\mathbb{R})}^{2}.

Adding (4.5) and (5.1), we see that there exists a universal constant C1C\geq 1 so that

(5.6) BTCR2+C(ε0+(Tκ2)14+κ2)BT.B_{T}\leq CR^{2}+C(\varepsilon_{0}+(T\kappa^{2})^{\frac{1}{4}}+\kappa^{-2})B_{T}.

Assuming

(5.7) ϵ01,Tκ2,κ1,\epsilon_{0}\ll 1,\qquad T\ll\kappa^{-2},\qquad\kappa\gg 1,

we conclude that

(5.8) BT2CR2.B_{T}\leq 2CR^{2}.

Now we fix our parameters in order, beginning with ϵ01\epsilon_{0}\ll 1 so that the first relation (5.7) holds, then

κ=1+4CR2,\kappa=1+4CR^{2},

so that the third relation in (5.7) holds, and finally

T0=cκ2,c1T_{0}=c\kappa^{-2},\qquad c\ll 1

so that the second relation of (5.7) holds.

Finally, we run a standard continuity argument. We let T(0,T0]T\in(0,T_{0}] be maximal so that (3.9) holds in [0,T][0,T]. Then by the above reasoning shows that (5.8) holds in [0,T][0,T]. In particular, given the choice of κ\kappa, this implies that (3.9) holds strictly at time TT. But this contradicts the maximality of TT unless T=T0T=T_{0}. We conclude that T=T0T=T_{0}, and thus (5.8) holds in [0,T0][0,T_{0}]. Hence the conclusion of the theorem follows. ∎

Next, we will show that the assumptions (1.2)–(1.3) on the coefficients a1,,a4a_{1},\dots,a_{4} in Theorems 1.1 and 1.5 provide an example of when our hypotheses (4.1)–(4.4) are satisfied.

Lemma 5.3.

Given ε,T(0,1]\varepsilon,T\in(0,1], there exists a constant δ(0,1]\delta\in(0,1] so that, if the coefficients a1,,a4a_{1},\dots,a_{4} satisfy

(5.9) |aj(t,x)|+|xaj(t,x)|\displaystyle|a_{j}(t,x)|+|\partial_{x}a_{j}(t,x)| δ(1+x2)1for j=1,2,3,\displaystyle\leq\delta(1+x^{2})^{-1}\quad\text{for }j=1,2,3,
(5.10) |a4(t,x)|+|xa4(t,x)|\displaystyle|a_{4}(t,x)|+|\partial_{x}a_{4}(t,x)| (1+x2)1\displaystyle\lesssim(1+x^{2})^{-1}

uniformly for |t|T|t|\leq T and xx\in\mathbb{R}, then (4.1)–(4.4) are satisfied for ε\varepsilon.

Proof.

We partition ψ\psi as follows:

ψ2(x)=sech2xn0sech2(n) 1{n|x|<n+1}(x).\psi^{2}(x)=\operatorname{sech}^{2}x\leq\sum_{n\geq 0}\operatorname{sech}^{2}(n)\,1_{\{n\leq|x|<n+1\}}(x).

This yields

|ψz(x)aj(x)|2𝑑xn0sech2(n)n|xz|<n+1|a1(x)|2𝑑xδ2z4,\int|\psi_{z}(x)a_{j}(x)|^{2}\,dx\leq\sum_{n\geq 0}\operatorname{sech}^{2}(n)\int_{n\leq|x-z|<n+1}|a_{1}(x)|^{2}\,dx\lesssim\delta^{2}\langle z\rangle^{-4},

and so

ψzajLx2𝑑zδz2𝑑zδ.\int\left\lVert\psi_{z}a_{j}\right\rVert_{L^{2}_{x}}\,dz\lesssim\delta\int\langle z\rangle^{-2}\,dz\lesssim\delta.

Using the same estimates for xaj\partial_{x}a_{j}, we obtain

ψzajHx1𝑑zδfor j=1,2,3,4.\int\left\lVert\psi_{z}a_{j}\right\rVert_{H^{1}_{x}}\,dz\lesssim\delta\quad\textrm{for }j=1,2,3,4.

For j=4j=4, this shows that (4.4) is satisfied. For j=1,2,3j=1,2,3, this demonstrates that we may choose δ1\delta\leq 1 sufficiently small so that (4.1)–(4.3) hold. ∎

Lastly, we show that our result applies to the model (1.1) for the propagation of waves over a variable channel bottom.

Proof of Corollary 1.4.

If the function c:c:\mathbb{R}\to\mathbb{R} that describes the channel bottom is smooth and satisfies cL<1\left\lVert c\right\rVert_{L^{\infty}}<1, then the function y:y:\mathbb{R}\to\mathbb{R} given by

y(x)=0x1b53(ξ)𝑑ξ=0x1[1c(ξ)]56𝑑ξy(x)=\int_{0}^{x}\frac{1}{b^{\frac{5}{3}}(\xi)}\,d\xi=\int_{0}^{x}\frac{1}{[1-c(\xi)]^{\frac{5}{6}}}\,d\xi

is well-defined and satisfies

(5.11) |y(x)|1uniformly for x.|y^{\prime}(x)|\approx 1\quad\text{uniformly for }x\in\mathbb{R}.

Consequently,

(5.12) |f(y)|2𝑑y|(fy)(x)|2𝑑xand|f(y)|2𝑑y|(fy)(x)|2𝑑x,\int|f(y)|^{2}\,dy\approx\int|(f\circ y)(x)|^{2}\,dx\quad\text{and}\quad\int|f^{\prime}(y)|^{2}\,dy\approx\int|(f\circ y)^{\prime}(x)|^{2}\,dx,

and so

(5.13) fyHκ1fHκ1uniformly for κ1.\left\lVert f\circ y\right\rVert_{H^{1}_{\kappa}}\approx\left\lVert f\right\rVert_{H^{1}_{\kappa}}\quad\text{uniformly for }\kappa\geq 1.

Moreover, notice that y1:y^{-1}:\mathbb{R}\to\mathbb{R} exists by the inverse function theorem, and also satisfies (5.11)–(5.13). A straightforward duality argument then shows

(5.14) gyHκ1gHκ1uniformly for κ1.\left\lVert g\circ y\right\rVert_{H^{-1}_{\kappa}}\approx\left\lVert g\right\rVert_{H^{-1}_{\kappa}}\quad\text{uniformly for }\kappa\geq 1.

Making the change of variables

(5.15) u(t,x)=b53(x)v(t,y(x)4t),u(t,x)=b^{\frac{5}{3}}(x)v(t,y(x)-4t),

we find that if u(t,x)u(t,x) solves (1.1) then v(t,y)v(t,y) solves (gKdV) with coefficients

a1(t,y)\displaystyle a_{1}(t,y) =0,\displaystyle=0,
a2(t,y)\displaystyle a_{2}(t,y) =10b23by1,\displaystyle=10b^{\frac{2}{3}}b^{\prime}\circ y^{-1},
a3(t,y)\displaystyle a_{3}(t,y) =[59b43(b)2103b73b′′+4(1b23)]y1,\displaystyle=\big{[}\tfrac{5}{9}b^{\frac{4}{3}}(b^{\prime})^{2}-\tfrac{10}{3}b^{\frac{7}{3}}b^{\prime\prime}+4(1-b^{-\frac{2}{3}})\big{]}\circ y^{-1},
a4(t,y)\displaystyle a_{4}(t,y) =[103b2(b)3103b3bb′′53b4b′′′383b]y1.\displaystyle=\big{[}\tfrac{10}{3}b^{2}(b^{\prime})^{3}-\tfrac{10}{3}b^{3}b^{\prime}b^{\prime\prime}-\tfrac{5}{3}b^{4}b^{\prime\prime\prime}-\tfrac{38}{3}b^{\prime}\big{]}\circ y^{-1}.

Clearly, we may choose η(0,1]\eta\in(0,1] sufficiently small so that

|xjc(x)|η(1+x2)1for j=0,1,,4|\partial_{x}^{j}c(x)|\leq\eta(1+x^{2})^{-1}\quad\text{for }j=0,1,\dots,4

imply that the coefficients a1,,a4a_{1},\dots,a_{4} above satisfy (5.9)–(5.10). Note that the 44 in (5.15) contributes the constant term 44 in a3a_{3}, which is needed to ensure that this coefficient vanishes as y±y\to\pm\infty.

This demonstrates that vv satisfies the a-priori estimate (5.5). This in turn implies that uu satisfies (5.5) as well, by (5.14). ∎

References