Higher order expansions for the probabilistic local Cauchy theory of the cubic nonlinear Schrödinger equation on

By Árpád Bényi, Tadahiro Oh, and Oana Pocovnicu

Abstract

We consider the cubic nonlinear Schrödinger equation (NLS) on with randomized initial data. In particular, we study an iterative approach based on a partial power series expansion in terms of the random initial data. By performing a fixed point argument around the second order expansion, we improve the regularity threshold for almost sure local well-posedness from our previous work. We further investigate a limitation of this iterative procedure. Finally, we introduce an alternative iterative approach, based on a modified expansion of arbitrary length, and prove almost sure local well-posedness of the cubic NLS in an almost optimal regularity range with respect to the original iterative approach based on a power series expansion.

1. Introduction

1.1. Nonlinear Schrödinger equation

We consider the Cauchy problem of the cubic nonlinear Schrödinger equation (NLS) on :⁠Footnote1

1

Our discussion in this paper can be easily adapted to the cubic NLS on (and to other nonlinear dispersive PDEs). For the sake of concreteness, however, we only consider the case in the following. See also the comment after Theorem 1.3 for our particular interest in the three-dimensional problem.

The cubic NLS Equation 1.1 has been studied extensively from both the theoretical and applied points of view. In this paper, we continue our study in Reference 1Reference 2 and further investigate the probabilistic well-posedness of Equation 1.1 with random and rough initial data.

Recall that the equation Equation 1.1 is scaling critical in in the sense that the scaling symmetry

preserves the homogeneous -norm (when it is applied to functions only of ). It is known that the Cauchy problem Equation 1.1 is locally well-posed in for Reference 10. See also Reference 13Reference 18Reference 26Reference 29 for partial⁠Footnote2 global well-posedness and scattering results. On the other hand, Equation 1.1 is known to be ill-posed in for Reference 12. In Reference 2, we studied the probabilistic well-posedness property of Equation 1.1 below the scaling critical regularity under a suitable randomization of initial data; see Equation 1.3 below. In particular, we proved that Equation 1.1 is almost surely locally well-posed in , . Our main goal in this paper is to introduce an iterative procedure to improve this regularity threshold for almost sure local well-posedness. Furthermore, we study a critical regularity associated with this iterative procedure. By introducing a modified iterative approach, we then prove almost sure local well-posedness of Equation 1.1 in an almost optimal range with respect to the original iterative procedure (Theorem 1.8). Beyond the concrete results in this paper, we believe that the iterative procedures based on (modified) partial power series expansions are themselves of interest for further development in the well-posedness theory of dispersive equations with random initial data and/or random forcing.

2

Namely, either for smoother initial data or under an extra hypothesis.

Such a probabilistic construction of solutions to dispersive PDEs first appeared in the works by McKean Reference 34 and Bourgain Reference 5 in the study of invariant Gibbs measures for the cubic NLS on , . In particular, they established almost sure local well-posedness with respect to particular random initial data, basically corresponding to the Brownian motion. (These local-in-time solutions were then extended globally in time by invariance of the Gibbs measures. In the following, however, we restrict our attention to local-in-time solutions.) In Reference 8, Burq-Tzvetkov further elaborated this idea and considered a randomization for any rough initial condition via the Fourier series expansion. More precisely, they studied the cubic nonlinear wave equation (NLW) on a three-dimensional compact Riemannian manifold below the scaling critical regularity. By introducing a randomization via the multiplication of the Fourier coefficients by independent random variables, they established almost sure local well-posedness below the critical regularity. Such randomization via the Fourier series expansion is natural on compact domains Reference 15Reference 35 and more generally in situations where the associated elliptic operators have discrete spectra Reference 45Reference 47. See Reference 2Reference 3 for more references therein.

In the following, we go over a randomization suitable for our problem on . Recall the Wiener decomposition of the frequency space Reference 48:

We employ a randomization adapted to this Wiener decomposition. Let satisfy

Then, given a function on , we have

We define the Wiener randomization of by

where is a sequence of independent mean-zero complex-valued random variables on a probability space . The randomization Equation 1.3 based on the Wiener decomposition of the frequency space is natural in view of time-frequency analysis; see Reference 1 for a further discussion. Almost simultaneously with Reference 1, Lührmann-Mendelson Reference 32 also considered a randomization of the form Equation 1.3 (with cubes being substituted by appropriately localized balls) in the study of NLW on . For a similar randomization used in the study of the Navier-Stokes equations, see also the work of Zhang and Fang Reference 50, preceding Reference 1Reference 32. In Reference 50, the decomposition of the frequency space is not explicitly provided, but their randomization certainly includes randomization based on the decomposition of the frequency space by unit cubes or balls.

In the following, we assume that the probability distribution of satisfies the exponential moment bound

for all and . This condition is satisfied by the standard complex-valued Gaussian random variables and the uniform distribution on the circle in the complex plane.

We now recall the almost sure local well-posedness result from Reference 2 which is of interest to us.⁠Footnote3

3

For simplicity, we only consider positive times. By the time reversibility of the equation, the same analysis applies to negative times.

Theorem A.

Let . Given , let be its Wiener randomization defined in Equation 1.3. Then, the Cauchy problem Equation 1.1 is almost surely locally well-posed with respect to the random initial data . More precisely, there exists a set with such that, for any , there exists a unique function in the class

with such that is a solution to Equation 1.1 on .

Here, denotes the linear Schrödinger operator, and the function space is defined in Section 2 below. Note that the uniqueness statement of Theorem A in the class

is to be interpreted as follows: by setting (see Equation 1.7 below), uniqueness for the residual term holds in . See also Remark 1.1 below.

Recall that while the Wiener randomization Equation 1.3 does not improve differentiability, it improves integrability (see Lemma 4 in Reference 1). See Reference 8Reference 27Reference 42 for the corresponding statements in the context of the random Fourier series. The main idea for proving Theorem A is to exploit this gain of integrability. More precisely, let

denote the random linear solution with as initial data and write

Then, we see that satisfies the following perturbed NLS:

where is viewed as a given (random) source term. The main point is that the gain of space-time integrability of the random linear solution (Lemma 2.1) makes this problem subcritical,⁠Footnote4 and hence we can solve it by a standard fixed point argument. Over the last several years, there have been many results on probabilistic well-posedness of nonlinear dispersive PDEs, using this change of viewpoint.⁠Footnote5 See, for example, Reference 1Reference 2Reference 5Reference 7Reference 8Reference 9Reference 15Reference 25Reference 30Reference 32Reference 33Reference 34Reference 35Reference 37Reference 38Reference 43. In Reference 2, we studied the Duhamel formulation for Equation 1.8,

4

The scaling critical Sobolev regularity for Equation 1.1 is defined by the fact that the -norm remains invariant under the scaling symmetry Equation 1.2. Given , we can also define the scaling critical Sobolev regularity in terms of the -based Sobolev space . A direct computation gives as ). In particular, the gain of integrability of the random linear solution stated in Lemma 2.1 implies that in Equation 1.8 gives rise only to a subcritical perturbation. See Reference 3 for a further discussion on this issue.

Note, however, that if we consider a non-zero initial condition for , then the critical nature of the equation comes into play through the initial condition. See Remark 1.4. This plays an important role in studying the global-in-time behavior of solutions. See, for example, Reference 37.

5

In the field of stochastic parabolic PDEs, this change of viewpoint and solving the fixed point problem for the residual term is called the Da Prato-Debussche trick Reference 16Reference 17.

by carrying out case-by-case analysis on terms of the form , , , etc., in . The main tools were (i) the gain of space-time integrability of and (ii) the bilinear refinement of the Strichartz estimate (Lemma 2.7). This yields Theorem A.

By examining the case-by-case analysis in Reference 2, we see that the regularity restriction in Theorem A comes from the cubic interaction of the random linear solution:

See Proposition 1.2 below. In the following, we discuss an iterative approach to lower this regularity threshold by studying further expansions in terms of the random linear solution . We will also discuss the limitation of this iterative procedure.

Remark 1.1.

(i) The proof of Theorem A, presented in Reference 2, is based on a standard contraction argument for the residual term in a ball of radius in . As such, the argument in Reference 2 only yields uniqueness of in the ball . Noting that ,⁠Footnote6 it follows from Lemma A.8 in Reference 2 that the -norm of is continuous in . Then, by possibly shrinking the local existence time , we can easily upgrade the uniqueness of in the ball to uniqueness of in the entire . Namely, uniqueness of holds in the class Equation 1.5.

6

By convention, our definition of the -space already assumes that functions in belong to . See Definition 2.4 below.

(ii) With the exponential moment assumption Equation 1.4, the proof of Theorem A, presented in Reference 2, allows us to conclude that the set of full probability in Theorem A has the following decomposition:

such that

(a)

there exist and such that

for each ,

(b)

for each , , the function constructed in Theorem A is a solution to Equation 1.1 on .

See the statement of Theorem 1.1 in Reference 2. A similar decomposition of the set of full probability applies to Theorems 1.3 and 1.8 below. We, however, do not state it in an explicit manner in the following.

1.2. Improved almost sure local well-posedness

Let us first state the following proposition on the regularity property of the cubic term defined above.

Proposition 1.2.

Given , let be the Wiener randomization of defined in Equation 1.3 and set .

(i) For any , we have

almost surely. More precisely, there exist an almost surely finite constant and such that

for any . In particular, by taking for small , Equation 1.11 shows that the second order⁠Footnote7 term is smoother (by ) than the first order term , provided .

7

Here, we referred to as the second order term since it corresponds to the second order term appearing in the (formal) power series expansion of solutions to Equation 1.1 in terms of the random initial data. For the same reason, we refer to in Equation 1.19 and in Equation 1.22 as the third and fourth order terms in the following. See Subsection 1.3 below.

(ii) When , there is no smoothing in the second order term in general; there exists such that the estimate Equation 1.11 with fails for any .

In Reference 2, we already proved Equation 1.11 when and , giving the regularity restriction in Theorem A. See Section 3 for the proof of Proposition 1.2 for a general value of . In the proof of Theorem A, in order to carry out the case-by-case analysis for Equation 1.9, we need to have (where we have a deterministic local well-posedness for Equation 1.1). In view of Proposition 1.2, this imposes the regularity restriction . Note, however, that even when , is still a well defined space-time function of spatial regularity . This motivates us to consider the second order expansion

and remove the second order interaction . Indeed, the residual term satisfies the following equation:

where . In terms of the Duhamel formulation, we have

Then, by studying the fixed point problem Equation 1.14 for , we have the following improved almost sure local well-posedness of Equation 1.1.

Theorem 1.3.

Given , let . Given , let be its Wiener randomization defined in Equation 1.3. Then, the cubic NLS Equation 1.1 on is almost surely locally well-posed with respect to the random initial data . More precisely, there exists a set with such that, for any , there exists a unique function in the class

with such that is a solution to Equation 1.1 on .

As in Theorem A, the uniqueness of in the class is to be interpreted as uniqueness of the residual term in . See also Remark 1.1.

By taking , Theorem 1.3 states that Equation 1.1 is almost surely locally well-posed in , provided . This in particular improves the almost sure local well-posedness in Theorem A. On the other hand, by taking , Theorem 1.3 allows us to construct a solution , provided that . In particular, we can take random initial data below the scaling critical regularity , while we construct the residual part in . This opens up a possibility of studying the global-in-time behavior of , using the (non-conserved) energy of :

with random initial data below the scaling critical regularity. We remark that by inspecting the argument in Reference 2, a modification of (the proof of) Theorem A yields almost sure local well-posedness of Equation 1.1 in the class

only for . (This restriction on can be easily seen by setting in Proposition 1.2.) In particular, Theorem A does not allow us to take random initial data below the scaling critical regularity in studying the global-in-time behavior of , namely at the level of the energy . As our focus in this paper is the local-in-time analysis, we do not pursue further this issue on almost sure global well-posedness of Equation 1.1 below the scaling critical regularity in this paper. We, however, point out two recent results Reference 30Reference 37 on almost sure global well-posedness below the energy space for the defocusing energy-critical NLS in higher dimensions.

The main strategy for proving Theorem 1.3 is to study the fixed point problem Equation 1.14 by carrying out case-by-case analysis on

in , where the dual norm is defined by

Note that the number of cases has increased from the case-by-case analysis in the proof of Theorem A, where we had or . One of the main ingredients is the smoothing on stated in Proposition 1.2 above. Note, however, that in order to exploit this smoothing, we need to measure in the -norm, which imposes a certain rigidity on the space-time integrability.⁠Footnote8 In order to prove Theorem 1.3, we also need to exploit a gain of integrability on . In Lemma 3.3, we use the dispersive estimate (see Equation 3.10 below) and the gain of integrability on each of the three factors in Equation 1.10 and show that also enjoys a gain of integrability by giving up some differentiability.

8

Namely, we need to measure in for admissible pairs . See Lemma 2.6.

Remark 1.4.

Given , let be its Wiener randomization defined in Equation 1.3. Given , we can also consider Equation 1.1 with the random initial data of the form :

Then, by slightly modifying the proofs, we see that the analogues of Theorems A and 1.3 (with ) also hold for Equation 1.16. Namely, Equation 1.16 is almost surely locally well-posed, provided . This amounts to considering the following Cauchy problems:

when and

when . In this case, the critical nature of the problem appears through the interaction in the case-by-case analysis Equation 1.15 due to the deterministic (non-zero) initial data at the critical regularity. The required modification is straightforward and thus we omit details. See Proposition 6.3 in Reference 2 and Lemma 6.2 in Reference 37. We point out that the discussion in the next subsection also applies to Equation 1.16.

Remark 1.5.

(i) Let denote the trilinear operator defined by

Then, for , there is no finite constant such that

for all , where . See Appendix A. In particular, this shows that when , there is no deterministic smoothing for ; i.e., Equation 1.18 does not hold for .

(ii) The proof of Proposition 1.2 only exploits “the randomization at the linear level”. Given , let denote the class of functions defined by

Under the regularity assumption and , it follows from the proof of Proposition 1.2 that the left-hand side of Equation 1.18 is finite for any .⁠Footnote9 In other words, the proof of Proposition 1.2 only uses the fact that the random linear solution belongs to almost surely; see the probabilistic Strichartz estimate (Lemma 2.1). The multilinear random structure of in terms of the random linear solution yields further cancellation. See, for example, Lemma 3.6 in Reference 15. Such extra cancellation seems to improve only space-time integrability, and we do not know how to use it to improve the regularity threshold (i.e., differentiability) at this point. A similar comment applies to the unbalanced higher order terms (including below) studied in Proposition 1.7 below.

9

Here, we only need finiteness of the -norm defined in Equation 3.3 for each , . See Remark 3.2.

1.3. Partial power series expansion and the associated critical regularity

In this subsection, we discuss possible improvements over Theorem 1.3 by considering further expansions. For this purpose, we fix in the following. By examining the proof of Theorem 1.3, we see that the regularity restriction comes from the following third order term:

Namely, we have up to permutations. In Lemma 4.1, we show that given , we have . In particular, we have , provided , yielding the regularity threshold in Theorem 1.3.

A natural next step is to remove this non-desirable third order interaction

in the case-by-case analysis in Equation 1.15 by considering the following third order expansion:

In this case, the residual term satisfies the following equation:

We expect that Equation 1.21 is almost surely locally well-posed for , which would be an improvement over Theorem 1.3. The proof will be once again based on case-by-case analysis:

in . Note the increasing number of combinations. In the following, however, we do not discuss details of this particular improvement over Theorem 1.3. Instead, we consider further iterative steps and discuss a possible limitation of this procedure.

Remark 1.6.

We point out that the expansions Equation 1.7, Equation 1.12, and Equation 1.20 correspond to partial power series expansions of the first, second, and third orders,⁠Footnote10 respectively, of a solution to Equation 1.1 in terms of the random initial data. Then, by considering the associated equations Equation 1.8, Equation 1.13, and Equation 1.21 for the residual term , we are recasting the original problem Equation 1.1 as a fixed point problem centered at the partial power series expansions of the first, second, and third orders, respectively.

10

A (partial) power series expansion of a solution to Equation 1.1 in terms of the random initial data can be expressed as a summation of certain multilinear operators over ternary trees. See, for example, Reference 11Reference 36. Here, the term “order” in our context corresponds to “generation (of the associated trees) with the convention that the trivial tree consisting only of the root node is of the zeroth generation. For example, the third order term in Equation 1.19 appears as the summation over all the multilinear operators associated to the ternary trees of the third generation. In terms of the graphical representation in Reference 36, we have

where “  ” denotes the random linear solution and “  ” denotes the trilinear Duhamel integral operator defined in Equation 1.17 with its three children as its arguments, , and .

By drawing an analogy to the previous steps, we expect that the worst contribution comes from the following fourth order terms:

There are basically two contributions to Equation 1.22: or up to permutations. In Lemma 4.2, we show that the contribution from is worse, being responsible for the expected regularity restriction . In order to remove this term, we can consider the fourth order expansion

as in the previous steps and try to solve the following equation for the residual term :

We can obviously iterate this argument and consider the following th order expansion:

In this case, we need to consider the following equation for the residual term :

and hope to construct a solution by carrying out the following case-by-case analysis:

in . Then, a natural question to ask is, Does this iterative procedure work indefinitely, allowing us to arbitrarily lower the regularity threshold for almost sure local well-posedness for Equation 1.1? Or is there any limitation to it?

We now consider a “critical” regularity with respect to this iterative procedure for proving almost sure local well-posedness of Equation 1.1. We simply define the critical regularity for this problem to be the infimum of the values of such that given any , the above iterative procedure⁠Footnote11 shows that Equation 1.1 is almost surely locally well-posed with respect to the Wiener randomization of . This is an empirical notion of criticality; unlike the scaling criticality, we cannot a priori compute this critical regularity . Moreover, our discussion will be based on the estimates on the stochastic multilinear terms (Propositions 1.2 and 1.7 below). In the following, we discuss a (possible) lower bound on , presenting a limitation to our iterative procedure based on partial power series expansions.

11

Namely, we solve Equation 1.24 for with some finite (or infinite) number of steps.

Within the framework of the iterative procedure discussed above, a necessary condition for carrying out the case-by-case analysis Equation 1.25 to study Equation 1.24 in is that the st order term defined in a recursive manner,

belongs to . By the nature of this iterative procedure, we may assume that the lower order terms , , belong to for some but not in , since if any of the lower order terms, say for some , were in , then we would have stopped the iterative procedure at the th step. As in the previous steps, we expect that is responsible for a regularity restriction at the th step of this iterative approach.

It could be a cumbersome task to study the regularity property of due to the increasing number of combinations for satisfying , . In the following, we instead study the regularity property of the st order term⁠Footnote12 of a particular form, corresponding to up to permutations in Equation 1.26. Given an integer , define by setting and

12

In fact, we study the th order term in Proposition 1.7.

As mentioned above, there are only three terms in this sum: up to permutations. Hence, consists of the “unbalanced”⁠Footnote13 -linear terms appearing in the definition Equation 1.26 of . We claim that the st order term is responsible for the regularity restriction at the th step of the iterative procedure. See Theorem 1.8 below.

13

By associating the -linear terms appearing in the summation in Equation 1.26 with ternary trees of the th generation as in Reference 36, the summands in Equation 1.27 correspond to the “unbalanced” trees of the th generation, where two of the three children of the root node are terminal.

In anticipating the alternative expansion Equation 1.33 below, let us study the regularity property of the unbalanced th order term :

For , we have (with appropriate restrictions on the range of )

See Proposition 1.2 and Lemmas 4.1 and 4.2. In general, we have the following proposition.

Proposition 1.7.

Define a sequence of positive real numbers by the recursive relation

with . Given , let be the Wiener randomization of defined in Equation 1.3. Then, we have

for , almost surely.

By solving the recursive relation Equation 1.30, we have

and thus we have , , and . In particular, Proposition 1.7 agrees with Equation 1.29. Moreover, since is increasing and , the regularity restriction in Proposition 1.7 does not cause any issue since our main focus is to study the probabilistic local well-posedness of Equation 1.1 in the range of that is not covered by Theorem 1.3. Namely, we may assume in the following.

In view of Propositions 1.2 and 1.7, one obvious lower bound for the critical regularity for this iterative procedure is given by since there is no gain of regularity when (even in moving from to ). On the other hand, in order to prove almost sure local well-posedness of Equation 1.1 by carrying out the case-by-case analysis Equation 1.25 for the equation Equation 1.24, we need to show that the st order term belongs to . This gives rise to a regularity restriction at the th step of the iterative procedure. By taking , we obtain another “lower” bound⁠Footnote14 on this critical regularity .

14

This “lower” bound is based on the upper bounds obtained in Propositions 1.2 and 1.7. In other words, if one can improve the bounds in Propositions 1.2 and 1.7, then one can lower the value of . See Remark 1.10.

As mentioned above, the case-by-case analysis Equation 1.25 for general may be a combinatorially overwhelming task due to (i) the number of the increasing combinations in Equation 1.25 and (ii) the random multilinear terms , , themselves having non-trivial combinatorial structures which makes it difficult to establish nonlinear estimates; see Equation 1.26. In the following, we instead consider an alternative iterative procedure based on the expansion

in place of Equation 1.23. This expansion allows us to prove the following almost sure local well-posedness of Equation 1.1 for .

Theorem 1.8.

Let . Given , let be its Wiener randomization defined in Equation 1.3. Then, the cubic NLS Equation 1.1 on is almost surely locally well-posed with respect to the random initial data . More precisely, there exists a set with such that, for any , there exists a unique function in the class:

with such that is a solution to Equation 1.1 on . Here, is a unique positive integer such that .

As before, the uniqueness of in the class

is to be interpreted as uniqueness of the residual term in . See also Remark 1.1.

In view of the discussion above, Theorem 1.8 proves almost sure local well-posedness of Equation 1.1 in an almost “optimal”⁠Footnote15 regularity range with respect to the original iterative procedure based on the partial power series expansion Equation 1.23. The proof of Theorem 1.8 is analogous to that of Theorem 1.3. Given , fix such that .⁠Footnote16 Write a solution as in Equation 1.33. Note that as gets closer and closer to the critical value , the expansion Equation 1.33 gets arbitrarily long. In view of Equation 1.28, the residual term satisfies the following equation:

15

Once again, this is based on the estimates in Propositions 1.2 and 1.7. In particular, the “optimality” of the regularity threshold in Theorem 1.8 is with respect to Propositions 1.2 and 1.7. If one can improve the bounds in Propositions 1.2 and 1.7, then one can lower the regularity threshold in Theorem 1.8.

16

In view of Proposition 1.7, the lower bound on guarantees that almost surely, while the upper bound on states that we cannot use Proposition 1.7 to conclude almost surely.

Hence, we need to carry out the following case-by-case analysis:

in . While Theorem 1.8 yields almost sure local well-posedness in an almost optimal range with respect to the original iterative procedure based on the partial power series expansion Equation 1.23, the required analysis is much simpler than that required for the original iterative procedure based on the expansion Equation 1.23. First, note that while the case-by-case analysis Equation 1.25 based on the original iterative procedure involves combinatorially non-trivial , the case-by-case analysis Equation 1.35 only involves the unbalanced th order term which has a much simpler structure than . In particular, Proposition 1.7 shows that , , has a better regularity property than the second order term in Equation 1.10. In terms of space-time integrability, we show that also enjoys a gain of integrability by giving up control on derivatives (Lemma 4.3). Finally, by inspecting the proof of Theorem 1.3 (see Lemma 4.2 and Proposition 5.1 below), we see that, except for with up to permutations,⁠Footnote17 we can bound all the terms appearing in the case-by-case analysis Equation 1.15 in . Hence, we can basically apply the result of the case-by-case analysis Equation 1.15 to our problem at hand. More precisely, by rewriting the case-by-case analysis Equation 1.35 as

17

Namely, the terms constituting the third order term in Equation 1.19.

(with the restriction in Equation 1.35) and using the fact that the terms , , …, , behave better than , the proof of Theorem 1.3 (in particular, Proposition 5.1) can be used to control all the terms in Equation 1.35 except for

(up to permutations). Note that the contribution to Equation 1.36 under the Duhamel integral is precisely given by in Equation 1.27. In particular, Proposition 1.7 with yields the regularity restriction , i.e., the lower bound stated in Theorem 1.8. This allows us to construct a solution by the standard fixed point argument. See Section 6 for details.

We previously conjectured that the st order term in Equation 1.26 would be responsible for a regularity restriction at the th step of the original iterative procedure. Combining Proposition 1.7 and Theorem 1.8, we confirmed this claim; the regularity restriction indeed comes only from the unbalanced st order term in Equation 1.27.

We conclude this introduction with several remarks.

Remark 1.9.

Let . Then, by slightly modifying the proof of Theorem 1.8 based on the modified iterative approach Equation 1.33, we can prove almost sure local well-posedness of the cubic NLS Equation 1.16 with the random initial data of the form for the same range of . See Remark 1.4.

Remark 1.10.

In this paper, we exploit randomness only at the linear level in estimating the stochastic terms. See Remarks 1.5 and 3.2. It may be possible to lower the regularity thresholds in Theorems 1.3 and 1.8 by exploiting randomness at the multilinear level. We, however, do not pursue this direction in this paper since (i) our main purpose is to present the iterative procedures in their simplest forms and (ii) estimating the higher order stochastic terms by exploiting randomness at the multilinear level would require a significant amount of additional work, which would blur the main focus of this paper.

Remark 1.11.

The ill-posedness result in Reference 12 shows that the solution map

is not continuous for Equation 1.1 when . In proving Theorem A, we studied the perturbed NLS Equation 1.8 for . In particular, the proof shows that we can factorize the solution map for Equation 1.1 as⁠Footnote18

18

Similarly, we can factorize the solution map for Equation 1.16 as

such that the second map is continuous in .

where the first map can be viewed as a universal lift map and the second map is the solution map to Equation 1.8, which is in fact continuous in .⁠Footnote19

19

Here, denotes the intersection of suitable space-time function spaces of differentiability at most . In the following, we use in a similar manner.

In the case of Theorem 1.3 (with ), we have the following factorization of the solution map for Equation 1.1:

where the second map is the solution map to Equation 1.13, which is continuous from to .

In the case of Theorem 1.8, we create stochastic objects in the first step, where :

Once again, the second map (which is the solution map to Equation 1.34) is continuous from to . We point out an analogy between these factorizations Equation 1.37, Equation 1.38, and Equation 1.39 of the ill-posed solution maps into

(i)

the first step, involving stochastic analysis, and

(ii)

the second step, where purely deterministic analysis is performed in constructing a continuous map ,

and similar factorizations for studying rough differential equations via the rough path theory Reference 19 and singular stochastic parabolic PDEs Reference 21Reference 23.

In the proof of Theorem 1.8, we could consider the following expansion of an infinite order:

This would allow us to present a single argument that works for all . In this case, the residual part satisfies

In particular, we would need to worry about the convergence issue of infinite series, and hence there seems to be no simplification in considering the infinite order expansion Equation 1.40.

Another strategy would be to treat as one stochastic object and write

It follows from Equation 1.28 that satisfies the following equation:

Noting that

we can rewrite Equation 1.41 as

and thus satisfies

The equations Equation 1.42 and Equation 1.43 do not particularly appear to be in such a friendly format. Namely, studying Equation 1.42 and Equation 1.43 does not seem to provide a simplification over the case-by-case analysis Equation 1.35 for the equation Equation 1.34 for each fixed .

In a recent work Reference 40, the second author (with Tzvetkov and Y. Wang) proved invariance of the white noise for the (renormalized) cubic fourth order NLS on the circle. One novelty of this work is that we introduced an infinite sequence of stochastic -linear objects (depending only on the random initial data) and considered the following expansion of an infinite order:

For this problem, it turned out that satisfies a particularly simple equation.⁠Footnote20 Then by treating as one stochastic object, we wrote a solution as , which led to the factorization

20

In fact, the series corresponds to the power series expansion of the resonant cubic fourth order NLS. We point out that does not belong to the span of Wiener homogeneous chaoses of any finite order.

for , where denotes the Gaussian white noise on the circle.

Remark 1.12.

In Reference 44, the third author (with Y. Wang) recently studied probabilistic local well-posedness of NLS on within the framework of the -based Sobolev spaces, using the dispersive estimate. In the context of the cubic NLS Equation 1.1 on , their result yields almost sure local existence of a unique solution for the randomized initial data , provided . In particular, the argument in Reference 44 allows us to consider random initial data of lower regularities than Theorems A, 1.3, and 1.8. Note that, in Reference 44, it was shown that the solution only belongs to , almost surely. We point out that a slight adaptation of the work Reference 39 by the second and third authors (with Y. Wang) shows that the solution indeed lies in .

As compared to Reference 44, the argument presented in this paper provides extra regularity information, namely the decomposition Equation 1.33 of the solution with the terms of increasing regularities: and . In particular, the residual term lies in the (sub)critical regularity,⁠Footnote21 leaving us a possibility of adapting deterministic techniques to study its further properties.

21

By slightly tweaking the argument, we can easily place in .

This paper is organized as follows. In Section 2, we recall probabilistic and deterministic lemmas along with the definitions of the basic function spaces. In Section 3, we study the regularity properties of the second order term in Equation 1.10. In Section 4, we further investigate the regularity properties of the higher order terms and and the unbalanced higher order terms . Note that the analysis on and contains part⁠Footnote22 of the case-by-case analysis Equation 1.15 needed for proving Theorem 1.3. In Section 5, we then carry out the rest of the case-by-case analysis Equation 1.15 and prove Theorem 1.3. In Section 6, we briefly describe the proof of Theorem 1.8 by indicating how the analysis in the previous sections can lead to the proof. In Appendix A, we prove the deterministic non-smoothing of the Duhamel integral operator discussed in Remark 1.5.

22

Note that defined in Equation 1.22 contains the contribution from with up to permutations.

Notation.

We use (and ) to denote (and , respectively) for arbitrarily small , where an implicit constant is allowed to depend on (and it usually diverges as ).

Given a Banach space of temporal functions, we use the following shorthand notation: . For example, .

Let be an even, smooth cutoff function supported on such that on . Given a dyadic number , we set and

for . Then, we define the Littlewood-Paley projection operator as the Fourier multiplier operator with symbol . In the following, we use the convention that capital letters denote dyadic numbers. For example, for some .

Given dyadic numbers , …, , we set . We also use the shorthand notation . For example, we have .

2. Strichartz estimates and function spaces

2.1. Probabilistic Strichartz estimates

First, we recall the usual Strichartz estimates on for the reader’s convenience. We say that a pair is Schrödinger admissible if it satisfies

with . Then, the following Strichartz estimates are known to hold Reference 20Reference 28Reference 46Reference 49:

It follows from Equation 2.1 and Sobolev’s inequality that

for . We will use the following admissible pairs in this paper:

In particular, by Sobolev’s inequality, we have

One of the important key ingredients for probabilistic well-posedness is the probabilistic Strichartz estimates. Such probabilistic estimates were first exploited by McKean Reference 34 and Bourgain Reference 5. In the following, we state the probabilistic Strichartz estimates under the Wiener randomization Equation 1.3. See Reference 1 for the proofs.

Lemma 2.1.

Given on , let be its Wiener randomization defined in Equation 1.3. Then, given finite and , there exist such that

for all and with (i) if and (ii) if .

A similar estimate holds when (with ), but we will not need it in this paper. See Reference 38. We also need the following lemma on the control of the size of -norm of .

Lemma 2.2.

Given , let be its Wiener randomization defined in Equation 1.3. Then, we have

for all .

2.2. Function spaces and their properties

In this subsection, we go over the basic definitions and properties of the function spaces used for the Fourier restriction norm method (i.e., analysis involving the -spaces introduced in Reference 4) adapted to the space of functions of bounded -variation and its pre-dual, introduced and developed by Tataru, Koch, and their collaborators Reference 22Reference 24Reference 31. We refer readers to Reference 22Reference 24 for the proofs of the basic properties. See also Reference 2.

Let be the set of finite partitions of the real line. By convention, we set if . We use to denote the sharp characteristic function of a set .

Definition 2.3.

Let .

(i) We define a -atom to be a step function of the form

where and with . Furthermore, we define the atomic space by

with the norm

where the infimum is taken over all possible representations for .

(ii) We define to be the space of functions of bounded -variation with the standard -variation norm:

By convention, we impose that the limits exist in .

(iii) Let be the closed subspace of of all right-continuous functions with .

(iv) We define (and , respectively) to be the space of all functions such that is in (and in , respectively) with the norms

The closed subspace is defined in an analogous manner.

Recall the following inclusion relation: for ,

The space is the classical space of functions of bounded -variation, and the space appears as the pre-dual of with , . Their duality relation and the atomic structure of the -space turned out to be very effective in studying dispersive PDEs in critical settings.

We are now ready to define the solution spaces.

Definition 2.4.

(i) Let . We define to be the closure of with respect to the -norm defined by

(ii) Let . We define to be the space of all functions such that the map lies in for any and , where the -norm is defined by

Recall the following embeddings:

for .

Given an interval , we define the local-in-time versions and of these spaces as restriction norms. For example, we define the -norm by

We also define the norm for the non-homogeneous term on an interval :

We conclude this section by presenting some basic estimates involving these function spaces. See Reference 2Reference 22Reference 24 for the proofs.

Lemma 2.5.

Let and . Then, the following linear estimates hold:

for any and .

The transference principle Reference 22, Proposition 2.19 and the interpolation lemma Reference 22, Proposition 2.20 applied on the Strichartz estimates Equation 2.1 and Equation 2.2 imply the following estimates.

Lemma 2.6.

Given any admissible pair with and , we have

Similarly, the bilinear refinement of the Strichartz estimate Reference 6Reference 14Reference 41 implies the following bilinear estimate.

Lemma 2.7.

Let with . Then, we have

for any and .

Proof.

From the bilinear refinement of the Strichartz estimate Reference 6Reference 14 and the transference principle, we have

for all . See Reference 2 for the proof of Equation 2.5. On the other hand, it follows from Hölder’s and Sobolev’s inequalities that

Then, the estimate Equation 2.4 follows from interpolating Equation 2.5 and Equation 2.6.

3. On the second order term

In this and the next sections, we study the regularity properties of the various stochastic terms that appear in the iterative procedures. Given , let be the Wiener randomization of defined in Equation 1.3 and set

In this section, we study the regularity properties of the second order term:

We first present the proof of Proposition 1.2. We follow closely the argument in Reference 2.

Proof of Proposition 1.2.

(i) By Lemma 2.5, the estimate Equation 1.11 follows once we prove

for some almost surely finite constant and , where . In the following, we drop the complex conjugate when it does not play any role.

Define by

Then, by applying the dyadic decomposition, it suffices to prove that

for all , …, with . Once we prove Equation 3.4, the desired estimate Equation 3.2 follows from summing Equation 3.4 over dyadic blocks and applying Lemmas 2.1 and 2.2. Recall our shorthand notation: and .

In the following, we assume that

Case 1 ().

By Hölder’s inequality and Lemma 2.6, we have

provided that Equation 3.5 holds.

Case 2 ().
Subcase 2.a ().

By Cauchy-Schwarz’ inequality and Lemma 2.7 followed by Lemma 2.5,⁠Footnote23 we have

23

In the remaining part of this paper, we repeatedly apply this argument when there is a frequency separation. We shall simply refer to it as the “bilinear Strichartz estimate” argument.

provided that Equation 3.5 holds.

Subcase 2.b ().

By Hölder’s inequality and Lemma 2.6, we have

provided that Equation 3.5 holds.

Subcase 2.c ().

By the bilinear Strichartz estimate, we have

provided that Equation 3.5 holds.

Therefore, putting all the cases together, we obtain Equation 3.4.

(ii) Given and small , consider the following deterministic initial condition whose Fourier transform is given by

where , , and . By taking sufficiently small, we have , and thus we can neglect the effect of the randomization in Equation 1.3 since all three terms on the right-hand side will be multiplied by a common random number . Without loss of generality, we assume that in the following.

We estimate from below the contribution to , where

From Equation 3.1, we have

where the phase function is given by

Then, it follows that the only non-trivial contribution to Equation 3.6 appears if

(up to the permutation ). In this case, we have and thus

for all .

Now, recall the following lemma on the convolution.

Lemma 3.1.

There exists such that

for all .

By applying Lemma 3.1 to Equation 3.6 with Equation 3.8 and Equation 3.9, we obtain

Therefore, for any , we have

as , while remains bounded. This in particular implies that when , the estimate Equation 1.11 cannot hold for any . This proves part (ii).

Remark 3.2.

It follows from the proof of Proposition 1.2(i) that

where is as in Equation 1.17. In particular, the left-hand side of Equation 1.18 is finite for . The only probabilistic component in the proof of Proposition 1.2(i) appears in applying Lemmas 2.1 and 2.2 to control the -norm of in terms of the -norm of . In this sense, we exploit the randomization only at the linear level.

On the one hand, Proposition 1.2 shows that controls almost derivatives. On the other hand, we need to measure in the -norm, which controls only the admissible space-time Lebesgue norms (with derivatives) via Lemma 2.6. The following lemma breaks this rigidity by giving up a control on derivatives. In particular, it allows us to control a wider range of space-time Lebesgue norms of . The main idea is to use the dispersive estimate for the linear Schrödinger operator:

This allows us to reduce the analysis to a product of the random linear solution and apply Lemma 2.1.

Lemma 3.3.

Let . Given , let be its Wiener randomization defined in Equation 1.3. Then, for any finite , we have

for any and . Note that the right-hand side of Equation 3.11 is almost surely finite thanks to the probabilistic Strichartz estimate (Lemma 2.1).

Proof.

We first consider the case . From Equation 3.1 and Equation 3.10, we have

When , we proceed as in Equation 3.12, but we apply Sobolev’s inequality before applying Equation 3.10:

This completes the proof of Lemma 3.3.

4. On the higher order terms

In this section, we study the regularity properties of the higher order terms.

4.1. On the third order term

In this subsection, we study the third order term:

The following lemma shows that the third order term enjoys a gain of extra derivative as compared to the second order term (Proposition 1.2).

Lemma 4.1.

Given , let be the Wiener randomization of defined in Equation 1.3. Then, for any , we have

almost surely. In particular, there exist an almost surely finite constant and such that

for any .

Proof.

First, note that the only possible combination for in Equation 4.1 is up to permutations. The complex conjugate does not play any role in the subsequent analysis, and hence we drop the complex conjugate sign and simply study

By Lemma 2.5, it suffices to prove that

for some almost surely finite constant and , where . Define by

Then, by applying the dyadic decomposition, it suffices to prove that

for all , …, . Once we prove Equation 4.4, the desired estimate Equation 4.3 follows from summing over dyadic blocks and applying Lemmas 2.1 and 2.2 and Proposition 1.2. In the following, we fix

Without loss of generality, we assume that .

Case 1 ().
Subcase 1.a ().

By the bilinear Strichartz estimate and Lemma 2.6, we have