Sobolev norms of solutions to NLS
Abstract.
We apply inverse spectral theory to study Sobolev norms of solutions to the nonlinear Schrödinger equation. For initial datum and , we prove that there exists a conserved quantity which is equivalent to -norm of the solution.
Key words and phrases:
Dirac operators, NLS, scattering, Sobolev norms2010 Mathematics Subject Classification:
35Q551. Introduction
The classical defocusing nonlinear Schrödinger equation (NLS) [13, 23, 27] on the real line has the form
(1.1) |
It is known that for sufficiently regular initial datum the unique classical solution exists globally in time. For example, if lies in the Schwartz class , then for all . The long-time asymptotics of is known [9, 26, 10]. For less regular initial datum , one can define the solution by an approximation argument (see, e.g., [25]):
Theorem 1.1.
Let , and let converge to in . Denote by the solution of (1.1) corresponding to . We have
for some function that does not depend on the choice of the sequence .
The function in Theorem 1.1 is called the –solution of (1.1) corresponding to the initial datum . It is clear that such a solution is unique. The total energy of the solution is its -norm and it is conserved in time:
By Plancherel’s formula, it is equal to where stands for the Fourier transform. In this paper, we work with Sobolev spaces , . The -norm of a function is defined by
(1.2) |
The space is the completion of with respect to this norm. Equivalently, one can define it by
where is the space of tempered distribution.
In contrast to the linear Schrödinger equation for which all Sobolev norms are conserved, the solutions of NLS can exhibit inflation of Sobolev norm for (see, e.g., [8, 17] for details). Specifically, given an arbitrarily small positive and , there exists a solution to (1.1) that satisfies
(1.3) |
see [8] for that construction. This result is related to the “high-to-low frequency cascade”. It occurs when for initial datum , a part of -norm of , when written on the Fourier side, moves from high to low frequencies as time increases. The Sobolev norms with negative index can be used to capture this phenomenon. Indeed, since is time-invariant and the weight in (1.2) vanishes at infinity when , the transfer of -norm from high to low values of frequency makes the -norm grow.
For NLS, the inflation of -norm can not happen for . In [18], Koch and Tataru discovered the set of conserved quantities which agree with -norm up to a quadratic term for a small value of and . As a corollary, they obtained the bounds on that are uniform in time:
(1.4) |
In [16], Killip, Vişan, and Zhang proved a similar estimate using a different method. The estimates on the growth of -norms are related to questions of well-posedness and ill-posedness of NLS in Sobolev classes which have been extensively studied previously, see, e.g., [14, 18, 17, 5, 6, 7, 20].
In our paper, we use some recent results in the inverse spectral theory [2, 3, 1] to show that there are conserved quantities of NLS which agree with -norm provided that and the value of is under control. We apply our analysis to prove the following theorem.
Theorem 1.2.
Let and let be the solution of (1.1) corresponding to . Then,
(1.5) |
where , and and are two positive absolute constants.
This result shows, in particular, that for a given function whose -norm is concentrated on high frequencies, we will never see a significant part of -norm of the solution moving to the low frequencies. That limits the “high-to-low frequency cascade” we discussed above. The close inspection of construction used in [8] shows that the function in (1.3) has -norm smaller than but its -norm is large when is small. Hence, the bounds in Theorem 1.2 do not contradict the estimates in (1.3) when . We do not know whether Theorem 1.2 holds for .
The main idea of the proof of Theorem 1.2 is based on the analysis of the conserved quantity , , which is a coefficient in the transition matrix for the Dirac equation with potential . We take and show that is related to a certain quantity (see the Lemma 3.3 below) that characterizes both size and oscillation of . Using in the context of NLS is the main novelty of our work. We study and show that it is equivalent to norm of with constants that depend on its -norm. That gives the estimate (1.5) for and the intermediate range of is handled by interpolation. Our analysis relies heavily on the recent results [2, 3, 1] that characterize Krein – de Branges canonical systems and the Dirac operators whose spectral measures belong to the Szegő class on the real line. We also establish the framework that allows working with NLS in the context of well-studied Krein systems.
Notation
-
•
The symbol stands for identity matrix and symbol stands for . Constant matrices , , are defined in (2.2).
-
•
For a measurable set , we say that is for every compact .
-
•
The Fourier transform of a function is defined by
-
•
The symbol , unless we specify explicitly, denotes the absolute constant which can change the value from formula to formula. If we write, e.g., , this defines a positive function of parameter .
-
•
For two non-negative functions and , we write if there is an absolute constant such that for all values of the arguments of and . We define similarly and say that if and simultaneously. If , we will write .
-
•
Symbols are reserved for the standard basis in : , .
-
•
For matrix , the symbol denotes its Hilbert-Schmidt norm: .
2. Preliminaries
Our proof of Theorem 1.1 uses complete integrability of equation (1.1). In that framework, (1.1) can be solved by using the method of inverse scattering which we discuss next following [13].
2.1. The inverse scattering approach to NLS
Given a complex-valued function , define the differential operator
(2.1) |
where we borrow notation for constant matrices , from [13]:
(2.2) |
The expression is one of the forms in which the Dirac operator can be written. In Section 3, we will introduce another form and will show how the two are related. Let us also define
as in [13]. In the free case when , the matrix-function solves . Since , it decays at infinity fast and therefore one can find two solutions such that
(2.3) |
for every . These solutions are called the Jost solutions for . Since both and solve the same ODE, they must satisfy
(2.4) |
where the matrix does not depend on . One can show that it has the form
(2.5) |
The matrix is called the reduced transition matrix for , and the ratio is called the reflection coefficient for . One can obtain in a different way: let , , be the fundamental matrix for , that is,
(2.6) |
Then, we have and the pointwise limits
(2.7) |
exist for every . Moreover, we have on .
The coefficients , and were defined for and they satisfy , for these . However, one can show that is the boundary value of the outer function defined in by the formula (see (6.22) in [13])
which, in view of identity on , can be written as
(2.8) |
That shows, in particular, that defines both and , and defines and .
The map is called the direct scattering transform and its inverse is called the inverse scattering transform. These maps are well-studied when . In particular, we have the following result (see [13] for the proof).
Theorem 2.1.
The map is a bijection from onto the set of complex-valued functions .
The scattering transform has some symmetries:
Lemma 2.1.
If and , then
(dilation): | ||||
(conjugation): | ||||
(translation): | ||||
(modulation): | ||||
(rotation): |
Proof. Indeed, the direct substitution into (2.3) shows that if are Jost solutions for , then
-
are the Jost solutions for ,
-
are the Jost solutions for ,
-
are the Jost solutions for ,
-
are the Jost solutions for ,
-
are the Jost solutions for , .
Now, it is left to use the formula (2.4) which defines . A computation using (2.5) shows how and change under symmetries –. For example, the translation does not change and it multiplies by . The modulation , however, gives . Then, the claim follows from the definition of the reflection coefficient . ∎
The next result (see formula in [13]), along with the previous theorem, shows how the inverse scattering transform can be used to solve (1.1).
Theorem 2.2.
Let and let be the reflection coefficient of . Define the family
(2.9) |
For each , let be the potential in the previous theorem generated by . Then, is the unique classical solution of (1.1) with the initial datum . Moreover, for every , the function lies in .
The solutions to NLS equation
(2.10) |
behave in an explicit way under some transformations. Specifically, we have
These properties can be checked by direct calculation (see, e.g., formula (1.19) in [14] for ) and a simple inspection shows that the bound (1.5) is consistent with all these transformations. The statements of Theorem 2.2 and Lemma 2.1 are consistent with these symmetries as well.
Now, we can explain the idea behind the proof of the Theorem 1.2.
The idea of the proof for Theorem 1.2. One can proceed as follows. First, we assume that and notice that conservation of , , guaranteed by (2.9), yields that is conserved, where is defined for by (2.8). Separately, for every Dirac operator with , we show that is equivalent to some explicit quantity that involves . That quantity was introduced and studied in [2, 3, 1]: it resembles the matrix Muckenhoupt condition and it is equivalent to norm of provided that is under control, e.g., with some fixed . Putting things together, we see that Sobolev norm of does not change much in time provided that the bound holds. Since is time-invariant, we arrive to the statement of Theorem 1.2 for and . For , the claim of Theorem 1.2 is trivial. The intermediate range of is handled by interpolation using Galilean invariance of NLS. The general case when follows by a density argument if one uses the stability of -solutions guaranteed by Theorem 1.1.
There are other methods that use conserved quantities that agree with negative Sobolev norms. The paper [16] uses a representation of through a perturbation determinant. Then, the analysis of the perturbation series allows the authors of [16] to obtain estimates similar to (1.4). It is conceivable that this approach can provide results along the same lines as Theorem 1.2.
To focus on the Dirac operator with , we first consider this operator on half-line in connection to Krein systems that were introduced in [19].
2.2. Operator and Krein system
Let be a function on the positive half-line such that
for every . Recall that we denote the set of such functions by . The Krein system (see the formula in [12]) with the coefficient has the form
(2.11) |
where the derivative is taken with respect to and . Let also
(2.12) |
denote the so-called dual Krein system (see Corollary 5.7 in [12]). Set
(2.13) |
The matrix-function , which was defined in (2.6) for , makes sense if we assume that . In the next lemma, we relate to .
Lemma 2.2.
Let , on , and be the corresponding matrix-valued function defined by (2.13). Then, for and .
Proof. The proof is a computation. We have
Notice, that
Using relation , we obtain
Since and , one has . Therefore,
It follows that matrix-valued functions and solve the same Cauchy problem and thus , as required. ∎
Lemma 2.3.
Let , let on , and let be the corresponding matrix-valued function defined by (2.13). Then, for and .
Proof. Recall that matrices , , are defined in (2.2). Using relations and , we see that , where and . Then, previous lemma applies to , and . It gives . Returning to , we get . ∎
Given , we define the continuous analogs of Wall polynomials (see [15] and Section 7 in [12]) by
(2.14) |
where , , , are the solutions of systems (2.11), (2.12) for the coefficient from Lemma 2.2 and the coefficient from Lemma 2.3, correspondingly. Functions , , , are continuous analogs of polynomials orthogonal on the unit circle, they depend on two parameters: and and they satisfy identities (see formula (4.32) in [12]):
(2.15) |
for real .
We will use the following result (see Lemma 2 in [11] which contains a stronger statement).
Theorem 2.3.
That theorem allows us to define
(2.17) |
for every and for almost every . Moreover, Corollary 12.2 in [12] gives
(2.18) |
for a.e. . For every , we define
Proposition 2.1.
The function is outer in .
Proof. We can write
It is known that are outer (see the formulas (12.9) and (12.29) in [12]) and that satisfy in . The function has a positive real part in and so is an outer function. That shows that is a product of three outer functions and hence it is outer itself. ∎
Proposition 2.2.
Proof. If , the fundamental matrix and the continuous Wall polynomials (2.14) are related by the formula
(2.20) |
Indeed, it is enough to use Lemma 2.2, Lemma 2.3, and the fact that . Our next step is to prove that the limit
(2.21) |
exists in Lebesgue measure when . From (2.15), we obtain
for every and . Similarly,
Hence, the limits
(2.22) |
exist in Lebesgue measure on by Theorem 2.3. Moreover,
and the proposition follows. ∎
We end this section with a few remarks on reflection coefficients of potentials in . We have almost everywhere on due to the fact that almost everywhere on . That can also be established directly using (2.18). Proposition 2.2 then allows to define the reflection coefficient for every . The Lemma 2.1 holds for in that case as well. However, not all results about scattering transform can be generalized from the case to . For example, scattering transform is injective on by Theorem 2.1, but it is not longer so when extended to (see Example 6.1 in Appendix).
3. Another form of Dirac operator, , and the entropy function.
Suppose . The alternative to form of writing Dirac operator on the line is given by an expression
(3.1) |
is densely defined self-adjoint operator on the Hilbert space of functions such that is finite. and defined in (2.1) are related by a simple formula:
One way to study is to focus on Dirac operators on half-line first. Given , we define on by
(3.2) |
on the dense subset of absolutely continuous functions such that , . We will call the Dirac operator defined on the positive half-line with boundary conditions or simply the half-line Dirac operator. Set for , and let , be the solutions of Krein system (2.11) generated by . The Krein system with coefficient and Dirac equation (3.2) are related as follows (see the proof of Lemma 6.1 in Appendix): if solves the Cauchy problem , , then
where the continuous Wall polynomials were defined in (2.14). The Weyl function of the operator coincides (see Lemma 6.1 in Appendix) with
(3.3) |
It is known (see Theorem 7.3 in [12]) that the limit above exists for every and defines an analytic function of Herglotz-Nevanlinna class in . The latter means that . In the next theorem, denotes the nontangential boundary value on which exists Lebesgue almost everywhere. It is understood as a nonnegative function on and it satisfies .
Theorem 3.1.
Proof. Lemma 6.1 in Appendix shows that coincides with the Weyl function for the canonical system with Hamiltonian . Then, the bounds in (3.6) follow from the Theorem 1.2 in [3] (see also Corollary 1.4 in [3]).∎
The quantity will be called the entropy of the Dirac operator on . We now turn to (3.1) to define the entropy for the Dirac operator on the whole line. Take and let and , be the coefficients of Krein systems associated to restrictions of to the half-lines and . As in (3.3), the half-line Weyl functions satisfy
(3.7) |
These Weyl functions can be used to construct the spectral representation for the Dirac operator. Let
(3.8) |
Using , one can show that is a positive definite matrix for . In other words, is the matrix-valued Herglotz function. Therefore, there exists a unique matrix-valued measure taking Borel subsets of into nonnegative matrices such that
where , are constant real matrices, . The importance of becomes clear when we recall the spectral decomposition for . Specifically, let be the solution of the Cauchy problem
(3.9) |
Then, the mapping
(3.10) |
initially defined on the set of compactly supported smooth functions , extends (see Appendix) to the unitary operator between the Hilbert spaces and ,
Moreover, is unitary equivalent to the operator of multiplication by the independent variable in and the unitary equivalence is given by the operator . In fact, these properties of will not be used in the paper, we mention them only to motivate the following definition. Let us define the entropy function by
(3.11) |
where denotes the absolutely continuous part of the spectral measure and it satisfies for a.e. . The quantity will play a crucial role in our considerations. We first relate it to the coefficient of the reduced transition matrix which was introduced in Proposition 2.2.
Lemma 3.1.
We have for almost all . In particular, for all .
Proof. From the definition (or see page 59 in [21]), one has
(3.12) |
Substituting expressions for
into (3.12), we obtain
for almost every and the first claim of the lemma follows. Then, the second claim is immediate because is an outer function as we showed in Proposition 2.1.∎
Consider again the half-line entropy functions
We see that coincides with the entropy (3.4) for the restriction of to (that explains why we use the same notation for the two objects), and for the potential
Our plan now is to relate with and then use the fact that the full line entropy is conserved, see Lemma 3.1. That will eventually lead to the proof of Theorem 1.2.
Lemma 3.2.
Let and let , where and . Denote by the matrix-function in (3.1) corresponding to . Then, , as for every .
Proof. Take . We have
for the corresponding Weyl functions . We also have
by the mean value theorem for harmonic functions. From (3.12), it follows that
Notice that does not depend on because the coefficient in Lemma 3.1 for the potential does not depend on . So, we only need to show that
when and . The second relation follows from , , which hold because tends to zero weakly in as and (see Lemma 6.2 in Appendix). Moreover, relation implies that if and only if
(3.13) |
In the rest of the proof, we will show (3.13). Let , be the limits of continuous Wall polynomials corresponding to . Consider . The formula (12.57) in [12] gives
It implies that when and that when and . Now, we can write
So, it remains to show that as . That holds because as and
where the first equality follows from and the formula in [12]. Thus, (3.13) holds and we are done. ∎
Lemma 3.3.
Let . Denote by the solution of the Cauchy problem , , and set . Consider
(3.14) |
Then, we have
(3.15) |
for some positive absolute constants , .
4. Proof of Theorem 1.2
The following result will play a crucial role in what follows. We postpone its proof to the next section.
Theorem 4.1.
Suppose and let satisfy , where Then,
(4.1) |
where and , are two positive absolute constants.
Proof of Theorem 1.2 in the case . First, assume that and let be the solution of (1.1) with the initial datum . We want to prove that
(4.2) |
We have for all , see formula in [13]. Let denote the coefficient in the matrix (2.5) given by . For each , define by (3.2). Let be as in Lemma 3.3 and be defined by (3.11). Formulas (2.8) and (2.9) show that is constant in and Lemma 3.1 says that is constant in as well. The bound (3.15) yields
(4.3) |
Assume first that . Taking in (4.3) and applying (4.1) to , we get since . Hence, in that case (4.3) can be written as By (4.1), , and so .
If , we use dilation. Consider which solves the same equation and notice that
Let making . Then, for the Sobolev norm, we get
(4.4) |
Since
(4.5) |
one has
In particular, at we get
Since , one can apply the previous bounds to obtain
Then,
Recalling that , we obtain
for all . Finally, having proved (4.2) for , it is enough to use Theorem 1.1 to extend (4.2) to .∎
Our next goal is to prove the estimate
(4.6) |
where , and and are positive absolute constants. For , this bound is trivial. To cover , we will need some auxiliary results first. One of the basic properties of NLS which we discussed in the Introduction has to do with modulation: if solves (2.10), then solves (2.10) for every .
Lemma 4.1.
Let , . Then,
Proof. It is clear that for every and , because is a unimodular constant. We have , . Since , it only remains to change the variable of integration in
to get the statement of the lemma. ∎
The next result is a standard property of convolutions.
Lemma 4.2.
Let and set for . We have
Proof. After comparing the sum to an integral, it is enough to show that
The function on the left-hand side is even and continuous in and , so we can assume that . Then,
where
Combining these bounds proves the lemma. ∎
Proof of Theorem 1.2, the case . We can again assume that . Recall the estimate (1.5) for :
(4.7) |
According to Lemma 4.1, we have
(4.8) |
for . Let , , be the coefficients from Lemma 4.2 with . Then, (4.8) and Lemma 4.2 imply
(4.9) |
In particular, taking gives
(4.10) |
We now apply (4.7) to and use (4.9) and (4.10) to get
(4.11) |
If , we have the statement of our theorem. If , we use dilation transformation like in the previous proof for . Consider which solves the same equation and notice that Let making . Then, for the Sobolev norm, we have
From (4.5),
Then, one has
In particular, taking gives us
Now and we can apply the previous bounds to get
Then,
Recalling that , we obtain
for all . ∎
Our approach also provides the bounds for some positive Sobolev norms. The following proposition slightly improves (1.4) when , is large, and is much smaller than .
Proposition 4.1.
Let and let be the solution of (1.1) corresponding to . Then, for each , we get
(4.12) |
if and
(4.13) |
if .
Proof. In the case when , the proof of proposition repeats the arguments given above to get (4.11) except that the constants in the inequalities depend on and can blow up when . Suppose . Then, for the Sobolev norm, we have
Take and write the following estimate for the integral above:
We use and (4.11) to get . The previous estimate for yields Hence, We can write a lower bound
so
Writing the integral as a sum of two:
and estimating each of them, we get a bound which holds for all :
That is the required upper bound (4.13). ∎
5. Oscillation and Sobolev space .
In this part of the paper, our goal is to prove the Theorem 4.1. Let us recall its statement.
Theorem 5.1.
Suppose that and let satisfy , where Then,
(5.1) |
where and , are two positive absolute constants.
Theorem 5.1 is of independent interest in the spectral theory of Dirac operators. For example, Lemma 3.3 shows that and control the size of .
The strategy of the proof is the following. In the next subsection, we show that -norm of any function can be characterized through BMO-like condition for its “antiderivative”. In Subsection 5.2, we consider solution to Cauchy problem on the interval where zero-trace symmetric and study the quantity , which represents a single term in the sum for . The results in Subsection 5.3 show that small value of guarantees that the “local” norm of is also small. This rough estimate is used in the proof of Theorem 4.1 which is contained in Subsection 5.4.
5.1. One property of Sobolev space
Observe that a function belongs to the Sobolev space if and only if
(5.2) |
Moreover, the last integral is equal to Indeed, recall that stands for the Fourier transform of :
Then, from Plancherel’s identity and formula
we obtain
by properties of convolutions. We will need the following proposition.
Proposition 5.1.
Suppose that . Let be an absolutely continuous function on such that almost everywhere on . Then,
(5.3) |
where , , and the positive constants and are universal.
Proof. Take a function , and let be an absolutely continuous function on such that almost everywhere on . Assume first that has a compact support. The integral under the sum does not change if we add a constant to , so we can suppose without loss of generality that
Upper bound. Given , define by
and recall (see (5.2)) that
(5.4) |
Moreover,
(5.5) |
For each interval , we use (5.5) for the corresponding term in the sum (5.3):
after the Cauchy-Schwarz inequality is applied. Summing these estimates in and using (5.4), we get the upper bound in (5.3) for compactly supported .
Lower bound. Integration by parts gives
Therefore,
Using the inequality , we continue the estimate:
Since
we are left with estimating
Applying the Cauchy-Schwarz inequality for the telescoping sum
we get
Then,
We have
The last sum is finite and does not depend on index . Now, the estimate
proves that
Hence, the lower bound in (5.3) holds for compactly supported .
Now, take any . The definition (1.2) of implies that can be written as for some function . Moreover, this map is a bijection between and and . Taking the inverse Fourier transform of identity , one gets a formula where is understood as a derivative in . Since and , we have and, therefore, is absolutely continuous on with the derivative equal to . Now, take and define the corresponding . Here, is even and
Then, in and so in because the mapping is unitary from onto . Also, each is compactly supported and converges to uniformly on every finite interval. Define , and write (5.3) for . The estimate on the right gives
for each . Sending , the bound
appears. Taking , one has the right estimate in (5.3). In particular, it shows that the sum in (5.3) converges. By construction,
where is a sum of integrals over . Since ,
Hence, and, taking in inequality
one gets the left bound in (5.3). Since all antiderivatives are different by a constant and the integral in (5.3) does not change if we add a constant to , the proof is finished. ∎
5.2. Auxiliary perturbative results for a single interval
Notice that for any real symmetric matrix with zero trace, we have that is also real, symmetric and has zero trace. The converse statement is true as well. Hence, the equation in Theorem 5.1, which is equivalent to , can be written as with having the same properties as . Let denote the solution to
and denote the solution to
Lemma 5.1.
Suppose , where is real-valued, , , and . Then, for , we have
(5.6) | |||
(5.7) |
Proof. Notice that and that every matrix satisfies
(5.8) |
Also, for any real matrix , we have
Hence,
For the second integrand, we have
Then, identities (5.8) imply
and, since and ,
Similarly, Hence,
Now, we use the formula to rewrite the last expression as
Finally, (5.7) follows from by direct inspection after one uses the identities and , which holds for every . ∎
Remark. The integrand in (5.6) is symmetric: because and . Notice also, that
where is an eigenvalue of which explains why the left-hand side in (5.7) is nonnegative.
Lemma 5.2.
Suppose real-valued matrix-function is defined on and satisfies . Consider , where : . Then,
(5.9) |
Proof. The integral equation for is
(5.10) |
By Gronwall’s inequality,
(5.11) |
Iteration of (5.10) gives
Then,
Since , the identity which holds for all matrices , gives
∎
Lemma 5.3.
Suppose real-valued symmetric matrix-functions and are defined on and satisfy
(5.12) | |||
(5.13) | |||
(5.14) |
Consider where . Then, we have
(5.15) |
where
(5.16) |
and is an absolute positive constant. An analogous result holds if and are related by .
Proof. We will use the formula (5.7) for our analysis. Fix and take and which solve and . Iterating the corresponding integral equations, one gets
Taking as in (5.7) leaves us with
(5.17) | |||
(5.18) | |||
(5.19) |
Recall that where satisfies (5.13) and (5.14). These assumptions are to be used in the following proposition. On , we define the partial order
by requiring that and .
Proposition 5.2.
Suppose a matrix-function is defined on and denote
(5.20) |
Let an operator be given by: where and a matrix-function , defined on , satisfies and . Then,
(5.21) |
where is an absolute positive constant, the norms and derivatives are computed with respect to .
Proof. Let . Write
(5.22) |
Then,
Applying Cauchy-Schwarz inequality to the integral, integrating in from to and maximizing in gives
(5.23) |
Then,
and the estimate for the first coordinate in (5.21) follows from Cauchy-Schwarz inequality and (5.23). Since we get and the bound for the second coordinate in (5.21) is obtained. ∎
Continuation of the proof of Lemma 5.3. We apply the proposition to three times with the initial choice of : . That gives rise to taking the third power of matrix : , applying it to , and looking at the first coordinate. As the result, one has . Therefore,
(5.24) |
Similarly, we consider and use the previous proposition four times making the first choice of as . Applying the bound (5.11) to and , we get . This time, we compute the fourth power of matrix , apply it to vector , and look at the first coordinate. In the end, one has
(5.25) |
The first term in (5.17) can be written as
and
(5.26) |
For any three vectors and in , we have an estimate
which follows from the triangle inequality. Multiplying with
we get
Applying it to (5.17) gives
Taking norm in variables and of both sides and using (5.24), (5.25), (5.26) and the Cauchy-Schwartz inequality gives
Recalling the definition (5.16), we get
so
Lemma 5.3 is proved. ∎
Remark. All statements in this subsection can be easily adjusted to any interval but the constants in the inequalities will depend on the size of that interval.
5.3. Rough bound when is small
Lemma 5.4.
Suppose an absolutely continuous function is defined on and satisfies
(5.27) |
Then, where
Proof. There is such that and
Thus, . Then, writing
integrating in and maximizing in , we get
∎
Suppose is real-valued, symmetric matrix-function on with zero trace and . Define , where . Notice that for every constant matrix . Therefore, we can apply Lemma 5.2 to each interval by choosing and get an estimate which explains how controls :
The next lemma shows that controls the convolution of with the exponential.
Lemma 5.5.
Suppose is real-valued, symmetric matrix-function on with zero trace and entries in . Define where and assume that . If , then where is a positive absolute constant.
Proof. Let and . We split the proof into several steps.
1. Bound for a single interval . The definitions (3.5) and (3.14) imply that . From Theorem 1.2 and Theorem 3.2 in [3], we know that admits the following factorization on : where and satisfy conditions:
(5.28) | |||
(5.29) |
and
Since , we have , where is the largest eigenvalue of . If one denotes , then
(5.30) |
In particular, that yields
(5.31) |
The given conditions on and (5.11) yield
where the second estimate follows from the first since . The Hamiltonian is absolutely continuous on and
(5.32) |
on . We claim that and that . Indeed, if satisfies and , then . Moreover, given conditions on and and , we have
(5.33) |
uniformly on . Identity yields
Taking an arbitrary with , we get
which implies . We also have since and the claim is proved. Finally, we have
for since .
Next, let us study and . Since , one has on . Recall that and , so
Since is the largest eigenvalue of and , then (5.30) yields and . Introduce The matrix is unitarily equivalent to and that gives
(5.34) |
We need to study , which is equal to . To do so, notice that
(5.35) |
Hence,
where
The previously obtained estimates give us
(5.36) |
Now, we use (5.34), (5.36) to apply the previous lemma to each component of to obtain
(5.37) |
The formula (5.35) also gives an expression for :
where
and
Since , we have
For smooth matrix-functions , , , we have
Then,
Since and , we have
Summing up, we get
(5.38) |
2. Handling all intervals . Take any . Our immediate goal is to show the bound
(5.39) |
analogous to (5.38) but written for interval . To this end, take the Hamiltonian defined on . For the corresponding , we get as follows from its definition. Since the -characteristics of the Hamiltonians and are equal for every constant matrix , we can instead consider where . Using the arguments in step 1 for , we get (5.39).
5.4. Proof of Theorem 4.1
Denote , , and recall that .
1. Lower bound. Define . By Lemma 5.5, we know that . Next, we apply Lemma 5.3 to each interval . The remainder in that lemma allows the estimate
For each and , we can find a positive such that implies that the remainder is smaller than uniformly in all . For example, one can take
(5.40) |
where is a sufficiently large positive number that depends on . Therefore, for such and some positive constant independent of , we have
where the Proposition 5.1 has been applied to the terms in the right-hand side of (5.15), adjusted to the interval , to show that they are comparable to . Taking , we see that
in the case . If , one uses inequality to get
(5.41) |
which holds for some positive absolute constant due to (5.40). That provides the required lower bound.
2. Upper bound. Let For given value of , apply Lemma 5.3 and Proposition 5.1 to each interval . That gives
with an absolute constant . Since and , one has
∎
6. Appendix
Here we collect some auxiliary results used in the main text.
1. We start with an example that shows that the scattering transform is not injective when defined on . This is an analog of Lemma 17 in [24].
Example 6.1.
There exist potentials such that in but we have a.e. on for their reflection coefficients. In other words, the scattering transform is not injective on .
Proof. Let us consider
and
where , . Note that
Theorem 12.11 in [12] says that for every contractive analytic function on whose boundary values on satisfy there exists a unique coefficient such that , for the continuous Wall polynomials generated by . Moreover, we have
(6.1) |
Applying this result, we see that there exist functions such that , are the limits of their continuous Wall polynomials. Now define potentials by relations
From Proposition 2.2, we conclude that the coefficients , for these potentials satisfy
on . Hence, on . On the other hand, we have and by construction. It follows that and . Since and are nonzero (they have a nonzero -norm as follows from (6.1)), that yields in . ∎
2. Next, we outline how to prove that the spectral representation for the Dirac operator , defined by relation (3.1), is given by the Weyl-Titchmarsh transform (3.10). To this end, we will use the corresponding result for canonical Hamiltonian systems proved in [21].
At first, we note that if is the Hamiltonian from Theorem 3.1, then on and the operator is unitary from onto the Hilbert space
Moreover, coincides with the operator of the canonical Hamiltonian system generated by the Hamiltonian . Thus, the operator on is unitary equivalent to the operator on . Let be the solution of Cauchy problem
(6.2) |
where , , and the differentiation is taken with respect to . The Weyl-Titchmarsh transform for is defined by
on a dense subset of of smooth compactly supported functions. This operator is unitary from onto the space defined in the same way as at the beginning of Section 3. Specifically, we let be the half-line Weyl functions of and define as the representing measure for the matrix-valued Herglotz function in (3.8). It was proved in Theorem 3.21 in [21] that coincides with the operator of multiplication by the independent variable in . We also have
Thus, we only need to check that the Weyl functions used in Section 3 coincide with the half-line Weyl functions of the Hamiltonian . For the -Weyl functions this follows from Lemma 6.1 below. Comparing the formulas for , in the beginning of Section 3, we see that the Weyl function for corresponds to the Weyl function for where . Similarly, in the setting of canonical Hamiltonian systems, the Weyl function for coincides with the Weyl function of , . Therefore, the statement for follows from Lemma 6.1 below and from the relation .
Lemma 6.1.
Proof. The formula
for and is well-known and can be derived from the analysis of Weyl circles by using identity and the invariance of Weyl circles under transforms generated by -unitary matrices (in our setting, the -unitary matrix is : we have on ). See, e.g., [4] or Section 8 in [22] for more details on Weyl circles for canonical Hamiltonian systems. Thus, we focus on the second identity in (6.3) and define
Differentiating, one obtains , . It follows that . In particular, we have
Since , as (see Theorem 12.1 in [12]), and analogous relations hold for and , we have
The lemma is proved. ∎
3. Lemma 6.1 and some known results for canonical systems can be used to show that weak convergence of potentials of the Dirac operator implies convergence of the corresponding Weyl functions.
Lemma 6.2.
Suppose is a bounded sequence in which converges to zero weakly. Let be the associated matrix-functions defined as in Lemma 6.1. Then, the sequence of corresponding Weyl functions converges to locally uniformly in when .
Proof. For , denote by the Hamiltonian generated by as in Lemma 6.1. Then, is the Weyl function for the half-line operators and . Since and converge to zero weakly in as , the Hamiltonians tend to the identity matrix uniformly on compact subsets on . Then, their Weyl functions tend to the Weyl function of the Hamiltonian locally uniformly in by Theorem in [21]. ∎
References
- [1] R. Bessonov and S. Denisov. Szegő condition, scattering, and vibration of Krein strings. preprint https://arxiv.org/abs/2203.07132.
- [2] R. Bessonov and S. Denisov. A spectral Szegő theorem on the real line. Adv. Math., 359:106851, 41, 2020.
- [3] R. Bessonov and S. Denisov. De Branges canonical systems with finite logarithmic integral. Anal. PDE, 14(5):1509–1556, 2021.
- [4] R. Bessonov, M. Lukic, and P. Yuditskii. Reflectionless canonical systems, I. Arov gauge and right limits. preprint arXiv:2011.05261, 2014.
- [5] D. G. Bhimani and R. Carles. Norm inflation for nonlinear Schrödinger equations in Fourier-Lebesgue and modulation spaces of negative regularity. J. Fourier Anal. Appl., 26(6):Paper No.78, 34, 2020.
- [6] R. Carles and T. Kappeler. Norm-inflation with infinite loss of regularity for periodic NLS equations in negative Sobolev spaces. Bull. Soc. Math. France, 145(4):623–642, 2017.
- [7] M. Christ, J. Colliander, and T. Tao. Asymptotics, frequency modulation, and low regularity ill-posedness for canonical defocusing equations. Amer. J. Math., 125(6):1235–1293, 2003.
- [8] M. Christ, J. Colliander, and T. Tao. Ill-posedness for nonlinear Schrödinger and wave equations. preprint arXiv:math/0311048, 2003.
- [9] P. Deift and X. Zhou. Long-time asymptotics for solutions of the NLS equation with initial data in a weighted Sobolev space. volume 56, pages 1029–1077. 2003. Dedicated to the memory of Jürgen K. Moser.
- [10] P. A. Deift and X. Zhou. Long-time asymptotics for integrable systems. Higher order theory. Comm. Math. Phys., 165(1):175–191, 1994.
- [11] S. A. Denisov. On the existence of wave operators for some Dirac operators with square summable potential. Geom. Funct. Anal., 14(3):529–534, 2004.
- [12] S. A. Denisov. Continuous analogs of polynomials orthogonal on the unit circle and Kreĭn systems. IMRS Int. Math. Res. Surv., pages 1–148, 2006, Art. ID 54517.
- [13] L. Faddeev and L. Takhtajan. Hamiltonian methods in the theory of solitons. Classics in Mathematics. Springer, Berlin, english edition, 2007. Translated from the 1986 Russian original by Alexey G. Reyman.
- [14] C. Kenig, G. Ponce, and L. Vega. On the ill-posedness of some canonical dispersive equations. Duke Math. J., 106(3):617–633, 2001.
- [15] S. Khrushchev. Schur’s algorithm, orthogonal polynomials, and convergence of Wall’s continued fractions in . Journal of Approximation Theory, 108(2):161–248, 2001.
- [16] R. Killip, M. Vişan, and X. Zhang. Low regularity conservation laws for integrable PDE. Geom. Funct. Anal., 28(4):1062–1090, 2018.
- [17] N. Kishimoto. A remark on norm inflation for nonlinear Schrödinger equations. Commun. Pure Appl. Anal., 18(3):1375–1402, 2019.
- [18] H. Koch and D. Tataru. Conserved energies for the cubic nonlinear Schrödinger equation in one dimension. Duke Math. J., 167(17):3207–3313, 2018.
- [19] M. G. Kreĭn. Continuous analogues of propositions on polynomials orthogonal on the unit circle. Dokl. Akad. Nauk SSSR (N.S.), 105:637–640, 1955.
- [20] T. Oh and C. Sulem. On the one-dimensional cubic nonlinear Schrödinger equation below . Kyoto J. Math., 52(1):99–115, 2012.
- [21] C. Remling. Spectral Theory of Canonical Systems. De Gruyter Studies in Mathematics Series. Walter De Gruyter GmbH, 2018.
- [22] R. Romanov. Canonical systems and de Branges spaces. preprint arXiv:1408.6022, 2014.
- [23] T. Tao. Nonlinear dispersive equations, volume 106 of CBMS Regional Conference Series in Mathematics. Published for the Conference Board of the Mathematical Sciences, Washington, DC; by the American Mathematical Society, Providence, RI, 2006. Local and global analysis.
- [24] T. Tao and Ch. Thiele. Nonlinear Fourier analysis. preprint arXiv:1201.5129, 2012.
- [25] Y. Tsutsumi. -solutions for nonlinear Schrödinger equations and nonlinear groups. Funkcial. Ekvac., 30(1):115–125, 1987.
- [26] V. E. Zakharov and S. V. Manakov. Asymptotic behavior of non-linear wave systems integrated by the inverse scattering method. Z. Èksper. Teoret. Fiz., 71(1):203–215, 1976.
- [27] V. E. Zakharov and A. B. Shabat. Exact theory of two-dimensional self-focusing and one-dimensional self-modulation of waves in nonlinear media. Ž. Èksper. Teoret. Fiz., 61(1):118–134, 1971.