Sharp global well-posedness for KdV and modified KdV on and

By J. Colliander, M. Keel, G. Staffilani, H. Takaoka, and T. Tao

Abstract

The initial value problems for the Korteweg-de Vries (KdV) and modified KdV (mKdV) equations under periodic and decaying boundary conditions are considered. These initial value problems are shown to be globally well-posed in all -based Sobolev spaces where local well-posedness is presently known, apart from the endpoint for mKdV and the endpoint for KdV. The result for KdV relies on a new method for constructing almost conserved quantities using multilinear harmonic analysis and the available local-in-time theory. Miura’s transformation is used to show that global well-posedness of modified KdV is implied by global well-posedness of the standard KdV equation.

1. Introduction

The initial value problem for the Korteweg-de Vries (KdV) equation,

has been shown to be locally well-posed (LWP) for Kenig, Ponce and Vega Reference 32 extended the local-in-time analysis of Bourgain Reference 5, valid for , to the range by constructing the solution of Equation 1.1 on a time interval with depending upon . Earlier results can be found in Reference 4, Reference 28, Reference 23, Reference 31, Reference 12. We prove here that these solutions exist for in an arbitrary time interval thereby establishing global well-posedness (GWP) of Equation 1.1 in the full range The corresponding periodic -valued initial value problem for KdV

is known Reference 32 to be locally well-posed for . These local-in-time solutions are also shown to exist on an arbitrary time interval. Bourgain established Reference 9 global well-posedness of Equation 1.2 for initial data having (small) bounded Fourier transform. The argument in Reference 9 uses the complete integrability of KdV. Analogous globalizations of the best known local-in-time theory for the focussing and defocussing modified KdV (mKdV) equations ( in Equation 1.1, Equation 1.2 replaced by and , respectively) are also obtained in the periodic and real line ( settings.

The local-in-time theory globalized here is sharp (at least up to certain endpoints) in the scale of -based Sobolev spaces . Indeed, recent examples Reference 33 of Kenig, Ponce and Vega (see also Reference 2, Reference 3) reveal that focussing mKdV is ill-posed for and that -valued KdV () is ill-posed for . (The local theory in Reference 32 adapts easily to the -valued situation.) A similar failure of local well-posedness below the endpoint regularities for the defocusing modified KdV and the -valued KdV has been established Reference 13 by Christ, Colliander and Tao. The fundamental bilinear estimate used to prove the local well-posedness result on the line was shown to fail for by Nakanishi, Takaoka and Tsutsumi Reference 45. Nevertheless, a conjugation of the local well-posedness theory for defocusing mKdV using the Miura transform established Reference 13 a local well-posedness result for KdV at the endpoint . Global well-posedness of KdV at the endpoint and for mKdV in remain open problems.

1.1. GWP below the conservation law

-valued solutions of KdV satisfy conservation: . Consequently, a local well-posedness result with the existence lifetime determined by the size of the initial data in may be iterated to prove global well-posedness of KdV for data Reference 5. What happens to solutions of KdV which evolve from initial data which are less regular than ? Bourgain observed, in a context Reference 8 concerning very smooth solutions, that the nonlinear Duhamel term may be smoother than the initial data. This observation was exploited Reference 8, using a decomposition of the evolution of the high and low frequency parts of the initial data, to prove polynomial-in-time bounds for global solutions of certain nonlinear Schrödinger (NLS) and nonlinear wave (NLW) equations. In Reference 10, Bourgain introduced a general high/low frequency decomposition argument to prove that certain NLS and NLW equations were globally well-posed below , the natural regularity associated with the conserved Hamiltonian. Subsequently, Bourgain’s high/low method has been applied to prove global well-posedness below the natural regularity of the conserved quantity in various settings Reference 20, Reference 50, Reference 48, Reference 34, including KdV Reference 18 on the line. A related argument—directly motivated by Bourgain’s work—appeared in Reference 29, Reference 30 where the presence of derivatives in the nonlinearities leaves a Duhamel term which cannot be shown to be smoother than the initial data. Global rough solutions for these equations are constructed with a slightly different use of the original conservation law (see below).

We summarize the adaptation Reference 18 of the high/low method to construct a solution of Equation 1.1 for rough initial data. The task is to construct the global solution of Equation 1.1 evolving from initial data for with . The argument Reference 18 accomplishes this task for initial data in a subset of consisting of functions with relatively small low frequency components. Split the data with , where is a parameter to be determined. The low frequency part of is in (in fact for all ) with a big norm while the high frequency part is the tail of an function and is therefore small (with large ) in for any . The low frequencies are evolved according to KdV: The high frequencies evolve according to a “difference equation” which is selected so that the sum of the resulting high frequency evolution, and the low frequency evolution solves Equation 1.1. The key step is to decompose , where is the solution operator of the Airy equation. For the selected class of rough initial data mentioned above, one can then prove that and has a small (depending upon ) norm. Then an iteration of the local-in-time theory advances the solution to a long (depending on ) time interval. An appropriate choice of completes the construction.

The nonlinear Duhamel term for the “difference equation” mentioned above is

The local well-posedness machinery Reference 5, Reference 32 allows us to prove that if we have the extra smoothing bilinear estimate

with the space defined below (see Equation 1.11). The estimate Equation 1.3 is valid for functions such that are supported outside , in the range Reference 18, Reference 15. The estimate Equation 1.3 fails for and this places an intrinsic limitation on how far the high/low frequency decomposition technique may be used to extend GWP for rough initial data. Also, Equation 1.3 fails without some assumptions on the low frequencies of and , hence the initial data considered in the high/low argument of Reference 18. We showed that the low frequency issue may indeed be circumvented in Reference 15 by proving Equation 1.1 is GWP in . The approach in Reference 15 does not rely on showing the nonlinear Duhamel term has regularity at the level of the conservation law. We review this approach now and motivate the nontrivial improvements of that argument leading to sharp global regularity results for Equation 1.1 and Equation 1.2.

1.2. The operator and almost conserved quantities

Global well-posedness follows from (an iteration of) local well-posedness (results) provided the successive local-in-time existence intervals cover an arbitrary time interval . The length of the local-in-time existence interval is controlled from below by the size of the initial data in an appropriate norm. A natural approach to global well-posedness in is to establish upper bounds on for solutions which are strong enough to prove that may be covered by iterated local existence intervals. We establish appropriate upper bounds to carry out this general strategy by constructing almost conserved quantities and rescaling. The rescaling exploits the subcritical nature of the KdV initial value problem (but introduces technical issues in the treatment of the periodic problem). The almost conserved quantities are motivated by the following discussion of the conservation property of solutions of KdV.

Consider the following Fourier proof⁠Footnote1 that . By Plancherel,

1

This argument was known previously; see a similar argument in Reference 27.

where

is the (spatial) Fourier transform. Fourier transform properties imply

Since we are assuming is -valued, we may replace by . Hence,

We apply , and we use symmetry and the equation to find

The first expression is symmetric under the interchange of and so may be replaced by . Since we are integrating on the set where , the integrand is zero and this term vanishes. Calculating , the remaining term may be rewritten

On the set where , which we symmetrize to replace in Equation 1.4 by and this term vanishes as well. Summarizing, we have found that -valued solutions of KdV satisfy

and both integrands on the right side vanish.

We introduce the (spatial) Fourier multiplier operator defined via

with an arbitrary -valued multiplier . A formal imitation of the Fourier proof of -mass conservation above reveals that for -valued solutions of KdV we have

The term arising from the dispersion cancels since on the set where . The remaining trilinear term can be analyzed under various assumptions on the multiplier giving insight into the time behavior of . Moreover, the flexibility in our choice of may allow us to observe how the conserved mass is moved around in frequency space during the KdV evolution.

Remark 1.1.

Our use of the multiplier to localize the mass in frequency space is analogous to the use of cutoff functions to spatially localize the conserved density in physical space. In that setting, the underlying conservation law is multiplied by a cutoff function. The localized flux term is no longer a perfect derivative and is then estimated, sometimes under an appropriate choice of the cutoff, to obtain bounds on the spatially localized energy.

Consider now the problem of proving well-posedness of Equation 1.1 or Equation 1.2, with , on an arbitrary time interval . We define a spatial Fourier multiplier operator which acts like the identity on low frequencies and like a smoothing operator of order on high frequencies by choosing a smooth monotone multiplier satisfying

The parameter marks the transition from low to high frequencies. When , the operator is essentially the integration (since ) operator . When , acts like the identity operator. Note that is bounded if . We prove a variant local well-posedness result which shows the length of the local existence interval for Equation 1.1 or Equation 1.2 may be bounded from below by , for an appropriate range of the parameter . The basic idea is then to bound the trilinear term in Equation 1.6 to prove, for a particular small , that

If is huge, Equation 1.7 shows there is at most a tiny increment in as evolves from to . An iteration of the local theory under appropriate parameter choices gives global well-posedness in for certain .

The strategy just described is enhanced with two extra ingredients: a multilinear correction technique and rescaling. The correction technique shows that, up to errors of smaller order in , the trilinear term in Equation 1.6 may be replaced by a quintilinear term improving Equation 1.7 to

where is tiny. The rescaling argument reduces matters to initial data of fixed size: . In the periodic setting, the rescaling we use forces us to track the dependence upon the spatial period in the local well-posedness theory Reference 5, Reference 32.

The main results obtained here are:

Theorem 1.

The -valued initial value problem Equation 1.1 is globally well-posed for initial data

Theorem 2.

The -valued periodic initial value problem Equation 1.2 is globally well-posed for initial data .

Theorem 3.

The -valued initial value problem for modified KdV Equation 9.1 (focussing or defocussing) is globally well-posed for initial data .

Theorem 4.

The -valued periodic initial value problem for modified KdV (focussing or defocussing) is globally well-posed for initial data .

The infinite-dimensional symplectic nonsqueezing machinery developed by S. Kuksin Reference 36 identifies as the Hilbert Darboux (symplectic) phase space for KdV. We anticipate that Theorem 3 will be useful in adapting these ideas to the KdV context. The main remaining issue is an approximation of the KdV flow using finite-dimensional Hamiltonian flows analogous to that obtained by Bourgain Reference 6 in the NLS setting. We plan to address this topic in a forthcoming paper.

We conclude this subsection with a discussion culminating in a table which summarizes the well-posedness theory in Sobolev⁠Footnote2 spaces for the polynomial generalized KdV equations. The initial value problem

2

There are results, e.g. Reference 9, in function spaces outside the -based Sobolev scale.

has the associated Hamiltonian

The replacement shows that the choice is irrelevant when is even, but, when is odd there are two distinct cases in Equation 1.9: is called focussing and is called defocussing. The usefulness of the Hamiltonian in controlling the norm can depend upon the choice in Equation 1.10.

We now summarize the well-posedness theory for the generalized KdV equations. The notation D and F in Table 1 refers to the defocussing and focussing cases. We highlight with the notation ?? some issues which are not yet resolved (as far as we are aware).

Table 1.

-Valued Generalized KdV on Well-posedness Summary Table

kScalingIll-posedL.W.P.G.W.P.
2, Reference 13, Reference 32; Reference 13
3, F:Reference 33, D:Reference 13, Reference 31
4, Reference 33, Reference 25
50F: , Reference 33, D: ??, Reference 31D: Reference 21
F : small Reference 21
F : big blows up Reference 38
F: , Reference 33; D: ??, Reference 31D: , F: small
F: big blows up ??

Our results here and elsewhere Reference 16, Reference 14, Reference 17 suggest that local well-posedness implies global well-posedness in subcritical dispersive initial value problems. In particular, we believe our methods will extend to prove GWP of mKdV in and KdV in and also extend the GWP intervals in the cases . However, our results rely on the fact that we are considering the -valued KdV equation and, due to a lack of conservation laws, we do not know if the local results for the -valued KdV equation may be similarly globalized. An adaptation of techniques from Reference 13 may provide ill-posedness results in the higher power defocussing cases. Blow up in the focussing supercritical ( or, more generally, with ) is expected to occur but no rigorous results in this direction have been so far obtained Reference 39.

1.3. Outline

Sections 2 and 3 describe the multilinear correction technique which generates modified energies. Section 4 establishes useful pointwise upper bounds on certain multipliers arising in the multilinear correction procedure. These upper bounds are combined with a quintilinear estimate, in the setting, to prove the bulk of Equation 1.8 in Section 5. Section 6 contains the variant local well-posedness result and the proof of global well-posedness for Equation 1.1 in We next consider the periodic initial value problem Equation 1.2 with period . Section 7 extends the local well-posedness theory for Equation 1.2 to the -periodic setting. Section 8 proves global well-posedness of Equation 1.2 in . The last section exploits Miura’s transform to prove the corresponding global well-posedness results for the focussing and defocussing modified KdV equations.

1.4. Notation

We will use to denote various time independent constants, usually depending only upon . In case a constant depends upon other quantities, we will try to make that explicit. We use to denote an estimate of the form . Similarly, we will write to mean and . To avoid an issue involving a logarithm, we depart from standard practice and write The notation denotes for an arbitrarily small . Similarly, denotes . We will make frequent use of the two-parameter spaces with norm

For any time interval , we define the restricted spaces by the norm

These spaces were first used to systematically study nonlinear dispersive wave problems by Bourgain Reference 5. Klainerman and Machedon Reference 37 used similar ideas in their study of the nonlinear wave equation. The spaces appeared earlier in a different setting in the works Reference 46, Reference 1 of Rauch, Reed, and M. Beals. We will systematically ignore constants involving in the Fourier transform, except in Section 7. Other notation is introduced during the developments that follow.

2. Multilinear forms

In this section, we introduce notation for describing certain multilinear operators; see for example Reference 41, Reference 40. Bilinear versions of these operators will generate a sequence of almost conserved quantities involving higher order multilinear corrections.

Definition 1.

A k-multiplier is a function . A -multiplier is symmetric if for all , the group of all permutations on objects. The symmetrization of a -multiplier is the multiplier

The domain of is ; however, we will only be interested in on the hyperplane .

Definition 2.

A -multiplier generates a k-linear functional or k-form acting on functions ,

We will often apply to copies of the same function in which case the dependence upon may be suppressed in the notation: may simply be written .

If is symmetric, then is a symmetric -linear functional.

As an example, suppose that is an -valued function. We calculate

The time derivative of a symmetric -linear functional can be calculated explicitly if we assume that the function satisfies a particular PDE. The following statement may be directly verified by using the KdV equation.

Proposition 1.

Suppose satisfies the KdV equation Equation 1.1 and that is a symmetric -multiplier. Then

where

Note that the second term in Equation 2.3 may be symmetrized.

3. Modified energies

Let be an arbitrary even -valued 1-multiplier and define the associated operator by

We define the modified energy by

The name “modified energy” is in part justified since in case We will show later that for of a particular form, certain modified energies enjoy an almost conservation property. By Plancherel and the fact that and are -valued,

Using Equation 2.3, we have

The first term vanishes. We symmetrize the remaining term to get

Note that the time derivative of is a 3-linear expression. Let us denote

Observe that if , the symmetrization results in . This reproduces the Fourier proof of -mass conservation from the introduction.

Form the new modified energy

where the symmetric 3-multiplier will be chosen momentarily to achieve a cancellation. Applying Equation 2.3 gives

We choose

to force the two terms in Equation 3.4 to cancel. With this choice, the time derivative of is a 4-linear expression where

Upon defining

with

we obtain

where

This process can clearly be iterated to generate satisfying

These higher degree corrections to the modified energy may be of relevance in studying various qualitative aspects of the KdV evolution. However, for the purpose of showing GWP in down to and in down to , we will see that almost conservation of suffices.

The modified energy construction process is illustrated in the case of the Dirichlet energy

Define , and use Equation 2.3 to see

where is explicitly obtained from . Noting that on the set , we know that

The choice of results in a cancellation of the terms and

so .

Therefore, is an exactly conserved quantity. The modified energy construction applied to the Dirichlet energy led us to the Hamiltonian for KdV. Applying the construction to higher order derivatives in will similarly lead to the higher conservation laws of KdV.

4. Pointwise multiplier bounds

This section presents a detailed analysis of the multipliers which were introduced in the iteration process of the previous section. The analysis identifies cancellations resulting in pointwise upper bounds on these multipliers depending upon the relative sizes of the multiplier’s arguments. These bounds are applied to prove an almost conservation property in the next section. We begin by recording some arithmetic and calculus facts.

4.1. Arithmetic and calculus facts

The following arithmetic facts may be easily verified:

A related observation for the circle was exploited by C. Fefferman Reference 19 and by Carleson and Sjölin Reference 11 for curves with nonzero curvature. These properties were also observed by Rosales Reference 47 and Equation 4.1 was used by Bourgain in Reference 5.

Definition 3.

Let and be smooth functions of the real variable . We say that is controlled by if is nonnegative and satisfies for and

for all nonzero .

With this notion, we can state the following forms of the mean value theorem.

Lemma 4.1.

If is controlled by and , then

Lemma 4.2.

If is controlled by and , then

We will sometimes refer to our use of Equation 4.4 as applying the double mean value theorem.

4.2. bound

The multiplier was defined in Equation 3.3. In this section, we will generally be considering an arbitrary even -valued 1-multiplier . We will specialize to the situation when is of the form Equation 4.7 below. Recalling that and that is even allows us to re-express Equation 3.3 as

Lemma 4.3.

If is even -valued and is controlled by itself, then, on the set (dyadic),

Proof.

Symmetry allows us to assume . In case , the claimed estimate is equivalent to showing

But this easily follows when we rewrite the left side as and use Equation 4.3. In case , Equation 4.6 may be directly verified.

In the particular case when the multiplier is smooth, monotone, and of the form

we have

4.3. bound.

This subsection establishes the following pointwise upper bound on the multiplier .

Lemma 4.4.

Assume is of the form Equation 4.7. In the region where for dyadic,

We begin by deriving two explicit representations of in terms of . These identities are then analyzed in cases to prove Equation 4.9.

Recall that,

where and

and . We shall ignore the irrelevant constant in Equation 4.10. Therefore,

Recall also from Equation 4.2 that

We can now rewrite the first term in Equation 4.12

The second term in Equation 4.12 is rewritten, using , and the fact the is even,

We record two identities for .

Lemma 4.5.

If is even and -valued, the following two identities for are valid:

Proof.

The identity Equation 4.16 was established above. The identity Equation 4.17 follows from Equation 4.16 upon expanding and writing the second term in Equation 4.16 on a common denominator.

Proof of Lemma 4.4.

The proof consists of a case-by-case analysis pivoting on the relative sizes of . Symmetry properties of permit us to assume that . Consequently, we assume Since for , a glance at Equation 4.12 shows that vanishes when We may therefore assume that . Since , we must also have .

From Equation 4.13, we know that we can replace on the right side of Equation 4.9 by . Suppose Then, and so . Thus, at least one of must be at least of size comparable to . The right side of Equation 4.9 may be re-expressed as

Case 1. .

Term in Equation 4.16 is bounded by , and therefore, after cancelling with one of the , satisfies Equation 4.9. Term is treated next. In case , Equation 4.18 is an upper bound of and the triangle inequality gives since is a decreasing function. If and , we rewrite

Applying the mean value theorem and using gives since and , so this subcase is fine. If and , the double mean value theorem Equation 4.4 applied to term gives the bound

Our assumptions on give the bound which is smaller than Equation 4.18.

The remaining subcases have either precisely one element of the set much smaller than or precisely two elements much smaller than . In the case of just one small , we apply the mean value theorem as above. When there are two small , we apply the double mean value theorem as above.

Case 2.

Certainly, in this region. It is not possible for both and in this region. Indeed, we find then that and which with implies but while . We need to show .

Case 2A. .

Since and , we must have . So and our goal is to show The last three terms in Equation 4.17 are all , which is fine. The first term in Equation 4.17 is

Replacing by and by , we identify three differences poised for the mean value theorem. We find this term equals

with for so . This expression is also .

Case 2B. .

Since and , we must have . We have and here so our desired upper bound is . We recall Equation 4.16 and evaluate when we can to find

The last term is dangerous so we isolate a piece of the first term to cancel it out. Expanding , we see that

The first piece cancels with in Equation 4.19 and the second piece is of size , which is fine. It remains to control

by Expand using Equation 4.13 to rewrite this expression as

(The second term in Equation 4.20 cancelled with part of the first.) The second and third terms in Equation 4.21 are and may therefore be ignored. We rewrite the first term in Equation 4.21 using the fact that is even as

Since , we can apply the double mean value theorem to obtain

with . Therefore, this term is bounded by

which is smaller than as claimed.

Case 2C. .

This case follows from a modification of Case 2A.

Case 2D.

This case does not occur because but is very small which forces to also be small, which is a contradiction.

4.4. bound

The multiplier was defined in Equation 3.9, with Our work on above showed that vanishes whenever vanishes so there is no denominator singularity in . Moreover, we have the following upper bound on in the particular case when is of the form Equation 4.7.

Lemma 4.6.

If is of the form Equation 4.7, then

where

Proof.

This follows directly from Lemma 4.4. Note that allows for the simplification in defining .

5. Quintilinear estimate on

The upper bound contained in Lemma 4.6 and the local well-posedness machinery Reference 31, Reference 5, Reference 32 are applied to prove an almost conservation property of the modified energy . The almost conservation of is the key ingredient in our proof of global well-posedness of the initial value problem for KdV with rough initial data.

Recall that denotes the Bourgain space Reference 5 associated to the cubic on the time interval . We begin with a quintilinear estimate.

Lemma 5.1.

Let be functions of space-time. Then

Proof.

The left side of Equation 5.1 is estimated via Hölder’s inequality by

The first three factors are bounded using a maximal inequality from Reference 31,

(Strictly speaking, Reference 31 contains an estimate for which implies Equation 5.2 by summing over cubic levels using ; see Reference 5 or Reference 22, Reference 24. A similar comment applies to Equation 5.5 below.) The terms are controlled using the smoothing estimate

which is an interpolant between the local-in-time energy estimate

and the Kato smoothing estimate Reference 31, valid for supported outside ,

In the remaining low frequency cases (e.g., when is supported inside ) we have and therefore we may easily control by .

Lemma 5.1 is combined with the upper bound of Lemma 4.6 in the next result.

Lemma 5.2.

Recall the definition Equation 3.1 of the operator . If the associated multiplier is of the form Equation 4.7 with , then

with

Proof.

We may assume that the functions are nonnegative. By a Littlewood-Paley decomposition, we restrict each to a frequency band (dyadic) and sum in the at the end of the argument. The definition of the operator and Equation 4.22 shows that it suffices to prove

We cancel and consider the worst case when throughout.

Note that vanishes when for Hence, we are allowed to assume at least one, and hence two, of the . Symmetry allows us to assume and .

The objective here is to show that

The form Equation 4.7 of with implies that Therefore, we need to control

We break the analysis into three main cases: Case 1. ; Case 2. ; Case 3. .

In Case 1, we have that and . The desired prefactor then appears and Equation 5.1 gives the result claimed.

In Case 2, and we must have so we multiply by and it suffices to bound

which may be done using Equation 5.1.

For Case 3, we have We are certain to have and can therefore multiply by

to again encounter Equation 5.7.

A glance back at Equation 3.8 shows that for solutions of KdV, we can now control the increment of the modified energy .

6. Global well-posedness of KdV on

The goal of this section is to construct the solution of the initial value problem Equation 1.1 on an arbitrary fixed time interval . We first state a variant of the local well-posedness result of Reference 32. Next, we perform a rescaling under which the variant local result has an existence interval of size 1 and the initial data is small. This rescaling is possible because the scaling invariant Sobolev index for is which is much less than . Under the rescaling, we show that Equation 3.8 and Equation 5.6 allow us to iterate the local result many times with an existence interval of size 1, thereby extending the local-in-time result to a global one. This will prove Theorem 1.

6.1. A variant local well-posedness result

The expression , where and is of the form Equation 4.7, is closely related to the norm of . Recall that the definition of in Equation 4.7 depends upon . An adaptation of the local well-posedness result in Reference 32, along the lines of Lemma 5.2 in Reference 16 and Section 12 in Reference 17, establishes the following result.

Proposition 2.

If , the initial value problem Equation 1.1 is locally well-posed for data satisfying . Moreover, the solution exists on a time interval with the lifetime

and the solution satisfies the estimate

We briefly describe why this result follows from the arguments in Reference 32. The norm is connected to the norm by the identity where is the Fourier multiplier operator with symbol

Since is essentially nondecreasing and so acts like a differential operator. The crucial bilinear estimate required to prove Proposition 2 is

which is equivalent to showing

Since , the operator may be moved onto the higher frequency factor inside the parenthesis in the left side of Equation 6.3 and the bilinear estimate of Reference 32 then proves Equation 6.3.

6.2. Rescaling

Our goal is to construct the solution of Equation 1.1 on an arbitrary fixed time interval . We rescale the solution by writing . We achieve the goal if we construct on the time interval . A calculation shows that

The choice of the parameter will be made later but we select now by requiring

We drop the subscript on so that

and the task is to construct the solution of Equation 1.1 on the time interval .

Remark 6.1.

The spatial domain for the initial value problem Equation 1.1 is which is invariant under the rescaling In contrast, the spatial domain for the periodic initial value problem for scales with . The adaptation of our proof of global well-posedness in the periodic context presented in Section 8 requires us to identify the dependence of various estimates on the spatial period.

6.3. Almost conservation

Recall the modified energy This subsection shows that the modified energy of our rescaled local-in-time solution is comparable to the modified energy . Next, as forecasted in Section 5, we use Equation 3.8 and the bound Equation 5.6 to show is almost conserved, implying almost conservation of . Since the lifetime of the local result Equation 6.1 is controlled by , this conservation property permits us to iterate the local result with the same sized existence interval.

Lemma 6.1.

Let be defined with the multiplier of the form Equation 4.7 and Then

Remark 6.2.

The estimate Equation 6.6 is an a priori estimate for functions of alone. The variable appears as a parameter.

Proof.

Since it suffices to prove

We may again assume that the are nonnegative. By the definitions of Equation 3.5, and Equation 3.1, and also Equation 4.1 and Equation 4.5, Equation 6.7 follows if we show

We make a Littlewood-Paley decomposition and restrict attention to the contribution arising from (dyadic), and without loss assume . In case , then So, we can assume . We consider separately the cases: .

I. .

Since and controls itself (recall Lemma 4.1), we may apply Equation 4.3 to show . Of course in this case so we need to bound . But this quantity is bounded by (in fact with a decay in ) and we wish to prove

Let be defined via

The left side of Equation 6.10 may be rewritten

We may now apply Hölder in to bound the left side of Equation 6.10 by

Finally, the form of (and hence ) given in Equation 6.11 allows us to conclude using Sobolev that

Remark 6.3.

The argument reducing the left side of Equation 6.10 to Equation 6.13 by passing through the convolution representation Equation 6.12 will appear many times below. We will often compress this discussion by referring to it as an Hölder application”.

II.

By definition of , we have

Suppose . Then this expression is

Therefore, the multiplier in Equation 6.9 is bounded by

and Hölder finishes off Equation 6.9 and establishes Equation 6.7.

We record here that the preceding calculations imply

We turn our attention to proving Equation 6.8. By Equation 4.9 Equation 4.2, and the definition of Equation 3.7, it suffices to control for (dyadic), with that

The definition of shows the multiplier appearing in the left side of Equation 6.15 is

and for ,

With this upper bound on the multiplier, we bound the left side of Equation 6.15 in via Hölder and Sobolev to obtain the estimate Equation 6.15 and therefore Equation 6.8.

Since our rescaled solution satisfies , we are certain that

and, moreover, that

whenever Using the estimate Equation 5.6 in Equation 3.8, the rescaled solution is seen to satisfy

Consequently, using Equation 6.17, we see that the rescaled solution has

6.4. Iteration

We may now consider the initial value problem for KdV with initial data and, in light of the preceding bound, the local result will advance the solution to time . We iterate this process times and, in place of Equation 6.18, we have

As long as , we will have the bound

and the lifetime of the local results remains uniformly of size 1. We take This process extends the local solution to the time interval . We choose so that

which may certainly be done for This completes the proof of global well-posedness for in .

We make two observations regarding the rescalings of our global-in-time KdV solution:

The almost conservation law and local well-posedness iteration argument presented above implies that provided and are selected correctly

The estimate Equation 6.21 forms a bridge between Equation 6.19 and Equation 6.20 which implies

In fact, the selection of is polynomial in the parameter so Equation 6.22 gives a polynomial-in-time upper bound on .

The choice of . The parameter was chosen above so that

Since, from Equation 6.20, , we see that Equation 6.23 holds provided we choose

The choice of . The parameter is chosen so that

where is the exponent appearing in Equation 5.6 (in the -case just presented, ). This unravels to give a sufficient choice of :

In the range , the numerator of the exponent on is positive. The denominator is positive provided . For so we require better than third order decay with in the local-in-time increment Equation 5.6. With , calculating and inserting the resulting expression for in terms of into Equation 6.22 reveals that, for our global-in-time solutions of Equation 1.1, we have

Remark 6.4.

Observe that the polynomial exponent in Equation 6.27 does not explode as we approach the critical regularity value . This is due to the fact that Equation 5.6 gave us much more decay than required for iterating the local result. In principle, the decay rate in Equation 5.6 could be improved by going further along the sequence of modified energies.

7. Local well-posedness of KdV on

This section revisits the local-in-time theory for periodic KdV developed by Kenig, Ponce and Vega Reference 32 and Bourgain Reference 5. Our presentation provides details left unexposed in Reference 32 and Reference 5 and quantifies the dependence of various implied constants on the length of the spatial period. This quantification is necessary for the adaptation of the rescaling argument used in Section 6 to the periodic setting.

7.1. The -periodic initial value problem for KdV

We consider the -periodic initial value problem for KdV:

We first want to build a representation formula for the solution of the linearization of Equation 7.1 about the zero solution. So, we wish to solve the linear homogeneous -periodic initial value problem

Define to be normalized counting measure on :

Define the Fourier transform of a function defined on by

and we have the Fourier inversion formula

The usual properties of the Fourier transform hold:

and so on. If we apply , to Equation 7.5, we obtain

This, together with Equation 7.6, motivates us to define the Sobolev space with the norm

We will often denote this space by for simplicity. Note that there are about low frequencies in the range where the norm consists of the norm.

The Fourier inversion formula Equation 7.5 allows us to write down the solution of Equation 7.2:

For a function which is -periodic with respect to the variable and with the time variable , we define the space-time Fourier transform for and by

This transform is inverted by

The expression Equation 7.10 may be rewritten as a space-time inverse Fourier transform,

where represents a 1-dimensional Dirac mass at . This recasting shows that has its space-time Fourier transform supported precisely on the cubic in .

We next find a representation for the solution of the linear inhomogeneous -periodic initial value problem

with a given time-dependent -periodic (in ) function. By Duhamel’s principle,

We represent using Equation 7.12, apply Equation 7.10 and rearrange integrations to find

Performing the -integration, we find

The -periodic initial value problem for KdV Equation 7.1 is equivalent to the integral equation

Remark 7.1.

The spatial mean is conserved during the evolution Equation 7.1. We may assume that the initial data satisfies a mean-zero assumption since otherwise we can replace the dependent variable by at the expense of a harmless linear first order term. This observation was used by Bourgain in Reference 5. The mean-zero assumption is crucial for some of the analysis that follows.

7.2. Spaces of functions of space-time

The integral equation Equation 7.17 will be solved using the contraction principle in spaces introduced in this subsection. We also introduce some other spaces of functions of space-time which will be useful in our analysis of Equation 7.17.

We define the spaces for -periodic KdV via the norm

(We will suppress reference to the spatial period in the notation for the space-time function spaces and the related spaces below.)

The study of periodic KdV in Reference 32, Reference 5 has been based around iteration in the spaces . This space barely fails to control the norm. To ensure continuity of the time flow of the solution we construct, we introduce the slightly smaller space defined via the norm

If , then . We will construct the solution of Equation 7.17 by proving a contraction estimate in the space . The mapping properties of Equation 7.15 motivate the introduction of the companion spaces defined via the norm

Let be a nice bump function supported on with on . It is easy to see that multiplication by is a bounded operation on the spaces , , and .

7.3. Linear estimates

Lemma 7.1.

The proof follows easily from the fact that

Lemma 7.2.
Proof.

By applying a smooth cutoff, we may assume that is supported on . Let , where is a smooth bump function supported on which equals 1 on . The identity

valid for and , allows us to rewrite as a linear combination of

and

Consider the contribution Equation 7.24. By Equation 7.21, it suffices to show that

Since the Fourier transform of evaluated at is given by and one can easily verify that , the claimed estimate follows using the definition Equation 7.20.

For Equation 7.25, we discard the cutoff and note that the space-time Fourier transform of evaluated at is equal to . The claimed estimate then follows from the definitions Equation 7.20, Equation 7.19 and the decay estimate for used above.

Proposition 3.

Let be a -periodic function whose Fourier transform is supported on . Then

where

Remark 7.2.

In the limit , Equation 7.26 yields the Strichartz estimate on the line (at least when ),

Proposition 3 (and the related Proposition 4 below) will not be used in our proof of local and global well-posedness of KdV on but may be of relevance in studying other properties of the long period limit of the KdV equation.

Proof.

It suffices to show that

for functions satisfying the hypotheses. Properties of the Fourier transform allow us to re-express the left side as

where may be take to be a positive even Schwarz function. We evaluate the -integration by writing

with also rapidly decreasing. Inserting this into the re-expressed left side and applying Cauchy-Schwarz leads to the upper bound

The first integral may be pulled out of the norm and the term in the second integral is used to integrate in to give

Matters are thus reduced to quantifying the norm above. Let denote . We estimate by counting

In case , the cardinality of the set is so for . Assume now that and rename . The task is to estimate

This set is largest when the parabola is the flattest, i.e., when . We find that

which completes the proof.

Proposition 4.

If is a -periodic function of and the spatial Fourier transform of is supported on , then

where is as it appears in Equation 7.27.

This follows easily by stacking up cubic level sets on which Equation 7.26 holds.

Lemma 7.3.

If is a -periodic function of , then

The estimate Equation 7.31 is a rescaling of the case proven by Bourgain Reference 5.

Remark 7.3.

We can interpolate between Equation 7.30 and Equation 7.31 to obtain

7.4. Bilinear estimate

Proposition 5.

If and are -periodic functions of , also depending upon having zero -mean for all , then

Note that Equation 7.33 implies We will relax the notation by dispensing with various constants involving with the recognition that some of the formulas which follow may require adjusting the constants to be strictly correct.

Proof.

The norms involved allow us to assume that and are nonnegative. There are two contributions to the norm we must control. We begin with the contribution. Duality and an integration by parts shows that it suffices to prove

Writing shows that it suffices to prove

for all having zero -mean. The left side may be rewritten

Note that the mean zero conditions allow us to assume .

Case 1. .

The identity

and the Case 1 defining conditions imply

This of course implies

Inserting Equation 7.36 into Equation 7.35 and using symmetry reduces matters to showing that

After some natural substitutions and undoing the Fourier transform, we see that it suffices to show that

Using Hölder, we split the left side into and apply Equation 7.31 to finish this case. (In fact, we control the left side of Equation 7.37 with )

Case 2.

Derivatives are cheap in this frequency setting. We use Hölder to estimate Equation 7.35 in , and then we apply Equation 7.31 to control the norms. Finally, we use the Case 2 defining conditions and Sobolev to move to on two factors and to on the remaining factor. (Again, we have encountered on two factors.)

Since , the only remaining case to consider is when one of the frequencies is small and the other two are big. Symmetry permits us to focus on

Case 3.

The analog of Equation 7.36 in this case is

Since we are in the -periodic setting and our functions have zero -mean, we have We analyze this expression in two cases: when the term dominates the terms and when the or term dominates the term. In case dominates, it suffices to prove

(Strictly speaking, each that appears in Equation 7.39 should be replaced by but we abuse the notation with the understanding that all smooth cutoff functions are essentially the same within this analysis.) The left side of Equation 7.39 is estimated via Hölder by

Since embeds into , the first two factors are fine. For the third factor, we estimate the norm

and then take the (or ) norm.

Consider next the case when the term dominates the term. (The case when dominates is similar.) It suffices to prove

By Hölder, the left side is controlled by

The first two terms are easily bounded. For the third term, observe that

and

By interpolation, the desired inequality Equation 7.40 follows.

The preceding discussion established that

It remains to control the weighted portion of the norm to complete the proof of Equation 7.33. Since , it suffices to prove that

The left side of Equation 7.42 may be rewritten

The desired estimate may be re-expressed as

Recall that the norm and the various -integrations are with respect to the -dependent measure .

Since we may assume our functions have mean zero, we have that and the identity

implies that

In case is that maximum, we reduce matters to Equation 7.41. Indeed, we rewrite Equation 7.44 as

Cauchy-Schwarz in (with the observation that ) reduces this case to proving

Upon rewriting the left side using duality, we see that an Hölder application using Equation 7.31 finishes off this case. The situation when is the maximum is symmetric so we are reduced to considering the case when is the maximum in Equation 7.46.

In the event that

we get a little help from the 1-denominator in Equation 7.44. We cancel leaving in the denominator and in the numerator. After the natural cancellation using Equation 7.46, we collapse to needing to prove

We apply Cauchy-Schwarz in to obtain the upper bound

which is controlled as desired using Equation 7.31. The case when is symmetric.

All that remains is the situation when

Recalling Equation 7.45, we see here that

which we use to restrict . After performing the natural cancellation using Equation 7.46 on Equation 7.44, we wish to show that

where the set

We apply Cauchy-Schwarz in to bound the left side of Equation 7.49 by

The point here is that the characteristic function appearing in the -integrand above sufficiently restricts the region of integration to prove

uniformly in the parameter . Note that familiar arguments complete the proof of Equation 7.49 (and, hence, Equation 7.42) provided we show Equation 7.50.

Remark 7.4.

The condition Equation 7.48 restricts the functions essentially to the dispersive curve . Suppose for the moment that and we restrict our attention to those satisfying . Observe that the projection of the point set onto the -axis is a set of points which are -separated. Therefore, if we “vertically thicken” these points for , the projected set remains rather sparse on the -axis. The intuition underlying the proof of Equation 7.50 is that a vertical thickening of the set also projects onto a thin set on the axis.

Lemma 7.4.

Fix . For we have for all dyadic that

for some .

Proof.

The hypotheses are symmetric in so we may assume We first consider the situation when . The expression

allows us to conclude that since and . Suppose (dyadic) and (dyadic). We have, for some , that . For to satisfy Equation 7.52, . We make the crude observation that there are at most multiples of in the dyadic block . Hence, the set of possible satisfying Equation 7.52 must lie inside a union of intervals of size , each of which contains an integer multiple of . We have then that

since .

In case , we must have so, if (dyadic), we must have for some and we can repeat the argument presented above.

Remark 7.5.

If we change the setting of the lemma to the case where , we have to adjust the conclusion to read

We use the lemma to prove Equation 7.50. A change of variables leads us to consider

We decompose the integration and use Equation 7.53:

Finally, we crush using the extra decay in to obtain

which proves Equation 7.50.

7.5. Contraction

Consider the -periodic initial value problem Equation 7.1 with periodic initial data We show first that, for arbitrary , this problem is well-posed on a time interval of size provided is sufficiently small. Then we show by a rescaling argument that Equation 7.1 is locally well-posed for arbitrary initial data .

As mentioned before in Remark 7.1, we restrict our attention to initial data having zero -mean.

Fix and for define

The bilinear estimate Equation 7.33 shows that implies so the (nonlinear) operator

is defined on . Observe that is equivalent, at least for , to Equation 7.17, which is equivalent to Equation 7.1.

Claim 1.

We estimate

By Equation 7.21 and Equation 7.23, followed by the bilinear estimate Equation 7.33,

and the claim is proven.

Consider the ball

Claim 2.

is a contraction on if is sufficiently small.

We wish to prove that for some ,

for all . Since , it is not difficult to see that

Since ,

Hence, for fixed , if we take so small that

the contraction estimate is verified.

The preceding discussion establishes well-posedness of Equation 7.1 on a -sized time interval for any initial data satisfying Equation 7.54 .

Finally, consider Equation 7.1 with fixed and . This problem is well-posed on a small time interval if and only if the -rescaled problem

is well-posed on . A simple calculation shows that

Observe that

provided is taken to be sufficiently large. This verifies Equation 7.54 for the problem Equation 7.55 proving well-posedness of Equation 7.55 on the time interval, say . Hence, Equation 7.1 is locally well-posed for .

The preceding discussion reproves the local well-posedness result for periodic KdV in Reference 32. We record the following simple variant which will be used in proving the global result for Equation 1.2. See Section 11 of Reference 17 for a general interpolation lemma related to this proposition.

Proposition 6.

If , the initial value problem Equation 1.2 is locally well-posed for data satisfying . Moreover, the solution exists on a time interval with the lifetime

and the solution satisfies the estimate

8. Almost conservation and global well-posedness of KdV on

This section proves that the 1-periodic initial value problem Equation 7.1 for KdV is globally well-posed for initial data provided . In particular, we prove Theorem 2. The proof is an adaptation of the argument presented for the real line to the periodic setting.

8.1. Quintilinear estimate

The following quintilinear space-time estimate controls the increment of the modified energy during the lifetime of the local well-posedness result.

Lemma 8.1.

Let be -periodic function in also depending upon . Let denote the orthogonal projection onto mean zero functions, Assume that for all . Then

Proof (apart from the endpoint).

We group together and apply Equation 7.34 to control the left side by

The quintilinear estimate Equation 8.1 is thus reduced to proving the trilinear estimate

The estimate Equation 8.2 is implied by the more general fact: For any ,

The multilinear estimate Equation 8.3 is proved in the forthcoming paper Reference 17. Here we indicate the proof for the case of Equation 8.3, namely Equation 8.2, when . The proof for in Reference 17 supplements the discussion presented below with some elementary number theory. The reader willing to accept Equation 8.2 may proceed to Lemma 8.2.

The Fourier transform of equals

where denotes an integration over the set where . We make a case-by-case analysis by decomposing the left side of Equation 8.2 into various regions. We may assume that , are nonnegative -valued functions.

Case 1.

In this case, it suffices to show that

We observe using Sobolev that

and then, by Hölder,

Finally, using Sobolev again and the embedding , we conclude that Equation 8.5 holds. Since for is symmetric with the Case 1 defining condition, we may assume that we are in Case 2.

Case 2. for

The convolution constraints in this case imply that

Therefore, it suffices to show that

This estimate may be recast by wiggling the weights and using duality as

Case 2A. for .

Symmetry allows us to assume and we must have . Since , we also have (see Equation 4.2). Therefore, in this case, the left side of Equation 8.7 is bounded by

Then, we may bound the preceding by

and Equation 8.7 is equivalent to

We recall from Reference 5 that for any . Therefore, we validate Equation 8.8 using a Hölder application in , with the required estimate given by Sobolev and the estimate from Equation 7.31.

Case 2B. for

We bound in this region and control the left side of Equation 8.7 by

If , we may ignore and finish things off with an Hölder argument using for any .

This completes the proof of Equation 8.2 for .

Remark 8.1.

Bourgain has conjectured Reference 5 that . If this estimate were known, the previous discussion could be substantially simplified. Our proof of the case in Reference 17 is partly motivated by an effort to prove this embedding estimate.

Lemma 8.2.

If is of the form Equation 4.7 with , then

The proof is a simple modification of the proof of Lemma 5.2 with Equation 8.1 playing the role of Equation 5.1. The projection appearing in the left side of Equation 8.1 causes no trouble in this application, since, as shown in Lemma 4.6, vanishes when so it also vanishes when Note also that and are systematically replaced by and throughout the argument.

8.2. Rescaling

Our task is to construct the solution of the 1-periodic Equation 7.1 on an arbitrary fixed time interval . This is equivalent to showing the -rescaled problem with corresponding solution has a solution which exists on . The lifetime of the variant local result is controlled by and

We choose so that

This choice guarantees that the local-in-time result for Equation 7.1 is valid for a time interval of size 1.

8.3. Almost conservation and iteration

The local result and Lemma 8.2 imply

Recall that so we may ignore by slightly adjusting . Therefore, since by a (modification of) Equation 6.6 we have that

For small and large we see then that is also of size . We may iterate the local result times until, say, first exceeds , that is, until

The solution of the -periodic Equation 7.1 is thus extended to the interval . We now choose such that This completes the proof that Equation 1.2 is globally well-posed in . Comments similar to those presented in Equation 6.19-Equation 6.26 apply to the periodic case showing that for our solution of Equation 1.2 we have

9. Global well-posedness for modified KdV

The results obtained for KdV are combined with some properties of the Miura transform Reference 42 (see also the survey Reference 43, Reference 44) to prove global well-posedness results for modified KdV (mKdV). This section contains the proofs of Theorems 3 and 4. The initial value problem for -valued mKdV on the line is

The choice of sign distinguishes between the focussing () and defocussing () cases. This problem is known Reference 31 to be locally well-posed in for . The regularity requirement is sharp Reference 33, Reference 13. We establish global well-posedness of Equation 9.1 in the range improving the work of Fonseca, Linares and Ponce Reference 20.

9.1. Defocussing case

Consider the defocussing case of Equation 9.1. The Miura transform of a solution is the function defined by

A calculation shows that solves

Remark 9.1.

Note that the uniqueness of the mKdV evolution is known Reference 31 in a subspace of , obtained by intersection in spaces associated with the maximal function and smoothing effect norms, while the KdV uniqueness holds in , a subspace of which is not naturally identified within the image of under the Miura transform. Nevertheless, the Miura image of the mKdV evolution coincides with the KdV evolution from an element of the Miura image. Let denote the nonlinear solution flow map for the defocussing initial value problem Equation 9.1 and let denote the flow map of Equation 9.3. For smooth enough , we have the intertwining relationship

Since the KdV evolution is uniquely determined in for data in , the next lemma provides the regularity to show the Miura image of the mKdV evolution from , data is the unique solution of KdV. A similar remark applies to the focussing case.

Suppose the initial data for Equation 9.1 is in . We show that is in .

Lemma 9.1.

If , then .

Proof.

The lemma verifies that the initial data for the KdV equation Equation 9.3 is in and since . Therefore, the global well-posedness result for KdV just established applies to Equation 9.3 and we know that the solution exists for all time and satisfies

for some constant . We exploit this polynomial-in-time bound for KdV solutions to control using the Miura transform.

Note that

Since our mKdV solution satisfies -mass conservation, , it suffices to control to control . By Equation 9.2,

Using Equation 9.5,

Summarizing, we have

Lemma 9.2.

Assuming that and , there exists an such that

Assuming the lemma for a moment, observe that combining Equation 9.7 and Equation 9.6 implies a polynomial-in-time upper bound on giving global well-posedness of defocussing mKdV. We now turn to the proof of Equation 9.7.

Proof.

We first consider the case when . The Sobolev estimate in one dimension

is applied with , yielding for

that

We continue the estimate by writing and using Sobolev to get

Finally, we interpolate between and to obtain

where . Using Equation 9.9 and Equation 9.10, we can simplify to find and . Since , we observe that Equation 9.7 holds in case

In case , we begin with a crude step by writing

Modifying the steps in the previous case, we have

Then, by Sobolev and interpolation,

where

It is then clear that for , we have for an appropriate as claimed.

9.2. Focussing case

In the focussing case of -valued modified KdV, the Miura transform has a different form:

The function solves the complex KdV initial value problem

Since the solution of focussing modified KdV is -valued and derivatives are more costly than squaring in one dimension, we take the perspective that is “nearly -valued”. The variant local result for Equation 9.12 has an existence interval determined by . However, Equation 9.12 does not conserve but instead (almost) conserves . An iteration argument showing global well-posedness may proceed if we show that

for functions of the form given by the Miura transform Equation 9.11.

Observe that

Lemma 9.3.

Assuming that for all , and , there exists an such that

If we take the lemma for granted, we deduce from Equation 9.13 and Equation 9.14 that

The equivalence Equation 9.16 links the quantity determining the length of the local existence interval to an almost conserved quantity. Consequently, Equation 9.12 is GWP and , is polynomially bounded in . Since for solutions of focussing modified KdV, the equivalence Equation 9.16 implies , is polynomially bounded in . Therefore, focussing mKdV is globally well-posed in provided we prove the lemma above.

Proof of Lemma 9.3.

We use duality and rewrite the expression to be controlled as

where when and when and we have relaxed to the bilinear situation. The function is introduced to calculate the norm using duality so . We may assume that is nonnegative. Symmetry allows us to assume

Case 1.

In this case, acts like the identity operator and the task is to control . By Sobolev and interpolation,

and we observe that, in this case,

Case 2.

Case 2A.

We decompose the factors dyadically by writing

We focus on a particular dyadic interaction term

In the Case 2A region, . We make this substitution and apply Cauchy-Schwarz in to observe

Multiplying through by leads to

since Of course we can sum over the dyadic scales and retain the claim.

It remains to establish the claim for the Case 2A region when . We rewrite the expression Equation 9.17 differently as

Defining , we observe that the preceding expression is controlled by

We apply Hölder to estimate by

and Sobolev implies . Rewriting this gives

Multiplying by yields

and the prefactor vanishes as , proving the claimed estimate.

Case 2B. .

We multiply Equation 9.17 through by in numerator and denominator. We observe that and use the argument passing through in Case 2A to complete the proof.

9.3. Modified KdV on

Lemmas 9.2 and 9.3 naturally extend to the -periodic setting. These results link the polynomial-in-time upper bound Equation 8.10 for solutions of the -periodic initial value problem Equation 7.1 for KdV to a polynomial-in-time upper bound on for solutions of the -periodic initial value problem for mKdV, implying Theorem 4.

Mathematical Fragments

Equation (1.1)
Equation (1.2)
Equation (1.3)
Equation (1.4)
Equation (1.6)
Equation (1.7)
Equation (1.8)
Equation (1.9)
Equation (1.10)
Equation (1.11)
Proposition 1.

Suppose satisfies the KdV equation Equation 1.1 and that is a symmetric -multiplier. Then

where

Equation (3.1)
Equation (3.3)
Equation (3.4)
Equation (3.5)
Equation (3.7)
Equation (3.8)
Equation (3.9)
Equations (4.1), (4.2)
Lemma 4.1.

If is controlled by and , then

Lemma 4.2.

If is controlled by and , then

Equation (4.5)
Lemma 4.3.

If is even -valued and is controlled by itself, then, on the set (dyadic),

Equation (4.7)
Lemma 4.4.

Assume is of the form Equation 4.7. In the region where for dyadic,

Equation (4.10)
Equation (4.12)
Equation (4.13)
Lemma 4.5.

If is even and -valued, the following two identities for are valid:

Equation (4.18)
Equation (4.19)
Equation (4.20)
Equation (4.21)
Lemma 4.6.

If is of the form Equation 4.7, then

where

Lemma 5.1.

Let be functions of space-time. Then

Equation (5.2)
Equation (5.5)
Lemma 5.2.

Recall the definition Equation 3.1 of the operator . If the associated multiplier is of the form Equation 4.7 with , then

with

Equation (5.7)
Proposition 2.

If , the initial value problem Equation 1.1 is locally well-posed for data satisfying . Moreover, the solution exists on a time interval with the lifetime

and the solution satisfies the estimate

Equation (6.3)
Lemma 6.1.

Let be defined with the multiplier of the form Equation 4.7 and Then

Equation (6.7)
Equation (6.8)
Equation (6.9)
Equation (6.10)
Equation (6.11)
Equations (6.12), (6.13)
Equation (6.15)
Equation (6.17)
Equation (6.18)
Equations (6.19), (6.20)
Equation (6.21)
Equation (6.22)
Equation (6.23)
Equation (6.26)
Equation (6.27)
Equation (7.1)
Equation (7.2)
Equation (7.5)
Equations (7.6), (7.7), (7.8)
Equation (7.10)
Equation (7.12)
Equation (7.15)
Equation (7.17)
Remark 7.1.

The spatial mean is conserved during the evolution Equation 7.1. We may assume that the initial data satisfies a mean-zero assumption since otherwise we can replace the dependent variable by at the expense of a harmless linear first order term. This observation was used by Bourgain in Reference 5. The mean-zero assumption is crucial for some of the analysis that follows.

Equation (7.19)
Equation (7.20)
Lemma 7.1.
Lemma 7.2.
Equation (7.24)
Equation (7.25)
Proposition 3.

Let be a -periodic function whose Fourier transform is supported on . Then

where

Proposition 4.

If is a -periodic function of and the spatial Fourier transform of is supported on , then

where is as it appears in Equation 7.27.

Lemma 7.3.

If is a -periodic function of , then

Proposition 5.

If and are -periodic functions of , also depending upon having zero -mean for all , then

Equation (7.34)
Equation (7.35)
Equation (7.36)
Equation (7.37)
Equation (7.39)
Equation (7.40)
Equation (7.41)
Equation (7.42)
Equation (7.44)
Equation (7.45)
Equation (7.46)
Equation (7.48)
Equation (7.49)
Equation (7.50)
Equation (7.52)
Remark 7.5.

If we change the setting of the lemma to the case where , we have to adjust the conclusion to read

Equation (7.54)
Equation (7.55)
Lemma 8.1.

Let be -periodic function in also depending upon . Let denote the orthogonal projection onto mean zero functions, Assume that for all . Then

Equation (8.2)
Equation (8.3)
Equation (8.5)
Equation (8.7)
Equation (8.8)
Lemma 8.2.

If is of the form Equation 4.7 with , then

Equation (8.10)
Equation (9.1)
Equation (9.2)
Equation (9.3)
Equation (9.5)
Equation (9.6)
Lemma 9.2.

Assuming that and , there exists an such that

Equation (9.9)
Equation (9.10)
Equation (9.11)
Equation (9.12)
Equations (9.13), (9.14)
Lemma 9.3.

Assuming that for all , and , there exists an such that

Equation (9.16)
Equation (9.17)

References

Reference [1]
M. Beals. Self-spreading and strength of singularities for solutions to semilinear wave equations. Ann. of Math. (2), 118(1):187–214, 1983. MR 85c:35057
Reference [2]
B. Birnir, C. Kenig, G. Ponce, N. Svanstedt, L. Vega. On the ill-posedness of the Initial Value Problem for the generalizedKorteweg-de Vries and nonlinear Schrödinger equations. J. London Math. Soc. (2). 53(3):551–559, 1996. MR 97d:35233
Reference [3]
B. Birnir, G. Ponce, N. Svanstedt. The local ill-posedness of the modified KdV equation. Ann. Inst. H. Poincaré Anal. Non-Linéaire. 13(4):529–535, 1996. MR 97e:35152
Reference [4]
J. L. Bona and R. Smith. The initial-value problem for the Korteweg-de Vries equation. Philos. Trans. Roy. Soc. London Ser. A, 278(1287):555–601, 1975. MR 52:6219
Reference [5]
J. Bourgain. Fourier transform restriction phenomena for certain lattice subsets and applications to nonlinear evolution equations I, II. Geom. Funct. Anal., 3:107–156, 209–262, 1993. MR 95d:35160a, MR 95d:35160b
Reference [6]
J. Bourgain. Approximation of solutions of the cubic nonlinear Schrödinger equations by finite-dimensional equations and nonsqueezing properties. Internat. Math. Res. Notices, 2:79–88, 1994. MR 95f:35237
[7]
J. Bourgain. Aspects of long time behaviour of solutions of nonlinear Hamiltonian evolution equations Geom. Funct. Anal., 5:105–140, 1995. MR 96f:35151
Reference [8]
J. Bourgain. On the growth in time of higher Sobolev norms of smooth solutions of Hamiltonian PDE. Internat. Math. Res. Notices, 6:277–304, 1996. MR 97k:35016
Reference [9]
J. Bourgain. Periodic Korteweg de Vries equation with measures as initial data. Selecta Math. (N.S.), 3(2):115–159, 1997. MR 2000i:35173
Reference [10]
J. Bourgain. Refinements of Strichartz’ inequality and applications to 2D-NLS with critical nonlinearity. International Mathematical Research Notices, 5:253–283, 1998. MR 99f:35184
Reference [11]
L. Carleson and P. Sjölin. Oscillatory integrals and a multiplier problem for the disc. Studia Math., 44:287–299. (errata insert), 1972. Collection of articles honoring the completion by Antoni Zygmund of 50 years of scientific activity, III. MR 50:14052
Reference [12]
A. Cohen. Existence and regularity for solutions of the Korteweg-de Vries equation. Arch. Rational Mech. Anal., 71(2):143–175, 1979. MR 80g:35109
Reference [13]
M. Christ, J. Colliander, and T. Tao Asymptotics, frequency modulation and low regularity ill-posedness for canonical defocusing equations. To appear Amer. J. Math., 2002.
Reference [14]
J. Colliander, M. Keel, G. Staffilani, H. Takaoka, and T. Tao. A refined global wellposedness result for Schrödinger equations with derivative. To appear SIAM J. Math. Anal., 2002.
Reference [15]
J. Colliander, M. Keel, G. Staffilani, H. Takaoka, and T. Tao. Global well-posedness for KdV in Sobolev spaces of negative index. Electron. J. Diff. Eqns., 2001(26):1–7, 2001. MR 2001m:35269
Reference [16]
J. Colliander, M. Keel, G. Staffilani, H. Takaoka, and T. Tao. Global wellposedness for Schrödinger equations with derivative. SIAM Journal of Mathematical Analysis, 2001. SIAM J. Math. Anal. 33(3):649–669, 2001. MR 2002j:35278
Reference [17]
J. Colliander, M. Keel, G. Staffilani, H. Takaoka, and T. Tao. Multilinear estimates for periodic KdV equations and applications. To appear J. Funct. Anal., 2002.
Reference [18]
J. E. Colliander, G. Staffilani, and H. Takaoka. Global Wellposedness of KdV below . Mathematical Research Letters, 6(5,6):755–778, 1999. MR 2000m:35159
Reference [19]
C. Fefferman. A note on spherical summation multipliers. Israel J. Math., 15:44–52, 1973. MR 47:9160
Reference [20]
G. Fonseca, F. Linares, and G. Ponce. Global well-posedness for the modified Korteweg-de Vries equation. Comm. Partial Differential Equations, 24(3-4):683–705, 1999. MR 2000a:35210
Reference [21]
G. Fonseca, F. Linares, and G. Ponce. Global existence for the critical generalized KdV equation. Preprint, 2002.
Reference [22]
J. Ginibre. An introduction to nonlinear Schrödinger equations. In Nonlinear waves (Sapporo, 1995), pages 85–133. Gakkōtosho, Tokyo, 1997. MR 99a:35235
Reference [23]
J. Ginibre, Y. Tsutsumi, and G. Velo. Existence and uniqueness of solutions for the generalized Korteweg de Vries equation. Math. Z., 203(1):9–36, 1990. MR 90m:35168
Reference [24]
J. Ginibre. Le problème de Cauchy pour des EDP semi-linéaires périodiques en variables d’espace (d’après Bourgain). Astérisque, 237:Exp. No. 796, 4, 163–187, 1996. Séminaire Bourbaki, Vol. 1994/95. MR 98e:35154
Reference [25]
A. Grünrock. A bilinear Airy-estimate with application to gKdV-3. Preprint, 2001.
[26]
H. Hofer and E. Zehnder. Symplectic invariants and Hamiltonian dynamics. Birkhäuser Verlag, Basel, 1994. MR 96g:58001
Reference [27]
J.-L. Joly, G. Métivier, and J. Rauch. A nonlinear instability for systems of conservation laws. Comm. Math. Phys., 162(1):47–59, 1994. MR 95f:35145
Reference [28]
T. Kato. The Cauchy problem for the Korteweg-de Vries equation. Nonlinear partial differential equations and their applications. Collège de France Seminar, Vol. I (Paris, 1978/1979), pages 293–307. Pitman, Boston, Mass., 1981. MR 82m:35129
Reference [29]
M. Keel and T. Tao. Local and Global Well-Posedness of Wave Maps on for Rough Data. International Mathematical Research Notices, 21:1117–1156, 1998. MR 99k:58180
Reference [30]
M. Keel and T. Tao. Global well-posedness for large data for the Maxwell-Klein-Gordon equation below the energy norm. Preprint, 2000.
Reference [31]
C. Kenig, G. Ponce, and L. Vega. Well-Posedness and Scattering Results for the Generalized Korteweg-de Vries Equation via the Contraction Principle. Communications on Pure and Applied Mathematics, XLVI:527–620, 1993. MR 94h:35229
Reference [32]
C. E. Kenig, G. Ponce, and L. Vega. A bilinear estimate with applications to the KdV equation. J. Amer. Math. Soc., 9:573–603, 1996. MR 96k:35159
Reference [33]
C. E. Kenig, G. Ponce, and L. Vega. On the ill-posedness of some canonical dispersive equations. Duke Math. J., 106(3):617–633, 2001. MR 2002c:35265
Reference [34]
C. E. Kenig, G. Ponce, and L. Vega. Global well-posedness for semi-linear wave equations. Comm. Partial Differential Equations, 25(9-10):1741–1752, 2000. MR 2001h:35128
[35]
S. B. Kuksin. On squeezing and flow of energy for nonlinear wave equations. Geom. Funct. Anal., 5:668–711, 1995. MR 96d:35091
Reference [36]
S. B. Kuksin. Infinite-dimensional symplectic capacities and a squeezing theorem for Hamiltonian PDEs. Comm. Math. Phys. 167(3):531–552, 1995. MR 96e:58060
Reference [37]
S. Klainerman and M. Machedon. Smoothing estimates for null forms and applications. Internat. Math. Res. Notices, 9, 1994. MR 95i:58174
Reference [38]
Y. Martel and F. Merle. Blow up in finite time and dynamics of blow up solutions for the critical generalized KdV equation. J. Amer. Math. Soc. 15(3):617–664, 2002.
Reference [39]
F. Merle. Personal communication. 2002.
Reference [40]
Y. Meyer and R. R. Coifman. Ondelettes et opérateurs. III. Hermann, Paris, 1991. Opérateurs multilinéaires. [Multilinear operators]. MR 93m:42004
Reference [41]
Y. Meyer and R. Coifman. Wavelets. Calderón-Zygmund and multilinear operators, translated from the 1990 and 1991 French originals by David Salinger, Cambridge University Press, Cambridge, 1997. MR 98e:42001
Reference [42]
R. M. Miura. Korteweg-de Vries equation and generalizations. I. A remarkable explicit nonlinear transformation. J. Mathematical Phys., 9:1202–1204, 1968. MR 40:6042a
Reference [43]
R. M. Miura. The Korteweg-de Vries Equation: A Survey of Results. SIAM Review, 18(3):412 – 459, 1976. MR 53:8689
Reference [44]
R. M. Miura. Errata: “The Korteweg-deVries equation: a survey of results” (SIAM Rev. 18 (1976), no. 3, 412–459). SIAM Rev., 19(4):vi, 1977. MR 57:6908
Reference [45]
K. Nakanishi, H. Takaoka, and Y. Tsutsumi. Counterexamples to bilinear estimates related to the KdV equation and the nonlinear Schrödinger equation. Methods of Appl. Anal. 8(4):569–578, 2001.
Reference [46]
J. Rauch and M. Reed. Nonlinear microlocal analysis of semilinear hyperbolic systems in one space dimension. Duke Math. J., 49:397–475, 1982. MR 83m:35098
Reference [47]
R. R. Rosales. I. Exact solution of some nonlinear evolution equations, II. The similarity solution for the Korteweg-de Vries equation and the related Painlevé transcendent. PhD thesis, California Institute of Technology, 1977.
Reference [48]
H. Takaoka. Global Well-posedness for the Kadomtsev-Petviashvili II Equation. Discrete Cont. Dynam. Systems. 6(2):483–499, 1999. MR 2000m:35163
[49]
T. Tao. Multilinear weighted convolution of functions and applications to nonlinear dispersive equations. Amer. J. Math., 123(5):839–908, 2001. MR 2002k:35283
Reference [50]
N. Tzvetkov. Global low regularity solutions for Kadomtsev-Petviashvili equation. Differential Integral Equations., 13(10-12):1289–1320, 2001. MR 2001g:35227

Article Information

MSC 2000
Primary: 35Q53 (KdV-like equations), 42B35 (Function spaces arising in harmonic analysis), 37K10 (Completely integrable systems, integrability tests, bi-Hamiltonian structures, hierarchies)
Keywords
  • Korteweg-de Vries equation
  • nonlinear dispersive equations
  • bilinear estimates
  • multilinear harmonic analysis
Author Information
J. Colliander
Department of Mathematics, University of Toronto, Toronto, ON Canada, M5S 3G3
M. Keel
School of Mathematics, University of Minnesota, Minneapolis, Minnesota, 55455
G. Staffilani
Department of Mathematics, Stanford University, Stanford, California 94305-2125
Address at time of publication: Department of Mathematics, Massachusetts Institute of Technology, 77 Massachusetts Avenue, Cambridge, Massachusetts 02138
MathSciNet
H. Takaoka
Department of Mathematics, Hokkaido University, Sapporo 060-0810, Japan
Address at time of publication: Department of Mathematics, Kobe University, Rokko, Kobe 657-8501, Japan
T. Tao
Department of Mathematics, University of California, Los Angeles, Los Angeles, California 90095-1555
ORCID
MathSciNet
Additional Notes

The first author is supported in part by N.S.F. Grant DMS 0100595.

The second author is supported in part by N.S.F. Grant DMS 9801558.

The third author is supported in part by N.S.F. Grant DMS 9800879 and by a grant from the Sloan Foundation.

The fourth author is supported in part by J.S.P.S. Grant No. 13740087.

The last author is a Clay Prize Fellow and is supported in part by grants from the Packard and Sloan Foundations.

Journal Information
Journal of the American Mathematical Society, Volume 16, Issue 3, ISSN 1088-6834, published by the American Mathematical Society, Providence, Rhode Island.
Publication History
This article was received on , revised on , and published on .
Copyright Information
Copyright 2003 American Mathematical Society
Article References
  • Permalink
  • Permalink (PDF)
  • DOI 10.1090/S0894-0347-03-00421-1
  • MathSciNet Review: 1969209
  • Show rawAMSref \bib{1969209}{article}{ author={Colliander, J.}, author={Keel, M.}, author={Staffilani, G.}, author={Takaoka, H.}, author={Tao, T.}, title={Sharp global well-posedness for KdV and modified KdV on $\mathbb R$ and $\mathbb T$}, journal={J. Amer. Math. Soc.}, volume={16}, number={3}, date={2003-07}, pages={705-749}, issn={0894-0347}, review={1969209}, doi={10.1090/S0894-0347-03-00421-1}, }

Settings

Change font size
Resize article panel
Enable equation enrichment

Note. To explore an equation, focus it (e.g., by clicking on it) and use the arrow keys to navigate its structure. Screenreader users should be advised that enabling speech synthesis will lead to duplicate aural rendering.

For more information please visit the AMS MathViewer documentation.