Soliton resolution for Calogero–Moser derivative nonlinear Schrödinger equation
Abstract.
We consider soliton resolution for the Calogero–Moser derivative nonlinear Schrödinger equation (CM-DNLS). A rigorous PDE analysis of (CM-DNLS) was recently initiated by Gérard and Lenzmann, who demonstrated its Lax pair structure. Additionally, (CM-DNLS) exhibits several symmetries, such as mass-criticality with pseudo-conformal symmetry and a self-dual Hamiltonian. Despite its integrability, finite-time blow-up solutions have been constructed.
The purpose of this paper is to establish soliton resolution for both finite-time blow-up solutions and global solutions in a fully general setting, without imposing radial symmetry or size constraints. To our knowledge, this is the first non-integrable proof of full soliton resolution for Schrödinger-type equations. A key aspect of our proof is the control of the energy of the outer radiation after extracting a soliton, referred to as the energy bubbling estimate. This benefits from two levels of convervation laws, mass and energy, and self-duality. This approach allows us to directly prove continuous-in-time soliton resolution, bypassing time-sequential soliton resolution. Importantly, our proof does not rely on the integrability of the equation, potentially offering insights applicable to other non-integrable models.
Key words and phrases:
Calogero-Moser derivative nonlinear Schrödinger equation, continuum Calogero-Moser model, soliton resolution, asymptotic behaviors, self-duality.2020 Mathematics Subject Classification:
35B40 (primary), 35Q55, 37K10, 37K401. Introduction
In this paper we study the global behaviors of solutions, the soliton resolution, for the Calogero–Moser derivative nonlinear Schrödinger equation (CM-DNLS):
(CM-DNLS) |
Here, is a nonlocal derivative with the projection to positive frequencies, i.e. is the Cauchy–Szegő projection with the Fourier symbol, . (CM-DNLS) is a recently introduced nonlinear Schrödinger-type equation. (CM-DNLS) draws attention as it enjoys rich and interesting mathematical structures. To name a few, (CM-DNLS) is mass-critical (with pseudo-conformal invariance), completely integrable, and has a self-dual Hamiltonian. The purpose of this work is to show soliton resolution theorem for (CM-DNLS). We show the soliton resolution in fully general setting, without size constraints or radial symmetry. Although (CM-DNLS) is completely integrable, we do not use the integrability features, and our proof may give insights toward other non-integrable models.
(CM-DNLS) was introduced in [1] as a continuum model of the classical Calogero–Moser Hamiltonian system, which is known to be completely integrable [9, 10, 45, 47]. It is also known as continuum Calogero–Moser model (CCM) or Calogero–Moser NLS (CM-NLS). There is also a periodic counterpart, Calogero–Sutherland model [53, 54].
A rigorous mathematical analysis was initiated by Gérard and Lenzmann [26]. They verified that (CM-DNLS) is completely integrable by discovering the Lax pairs on the Hardy–Sobolev space :
The flow of (CM-DNLS) preserves the positive frequency condition, , and this distinctive feature is called chirality. For chiral solutions to (CM-DNLS), the equation admits the Lax pair structure:
(1.1) |
with the -dependent operators and defined by 111(1.2) is slightly different but equivalent to what was presented in [26]. This formulation was presented in the introduction of [40].
(1.2) |
There is a similar Lax pairs in the full intermediate NLS [48]. Another feature of integrability on the Hardy space is an explicit formula [40] given by a holomorphic function on the upper half-plane. This is followed by the works of Gérard and collaborators in relevant models [25, 24, 27].222In fact, (1.1) holds true regardless of chirality. See [41, Proposition 3.5]. However, the chirality is required for the explicit formula. This is a global explicit formula for a weak solution.
We briefly recall symmetries and conservation laws. (CM-DNLS) enjoys time and space translation, and phase rotation symmetries. They are associated to the conservation laws of energy, mass, and momentum:
Moreover, it has Galilean invariance
Of particular importance are -scaling symmetry
and pseudo-conformal symmetry
(1.3) |
Associated identities to scaling and pseudo-conformal symmetries are the virial identities (1.9). We note that those symmetries are shared with the mass-critical NLS. But, we remark that is a complete square of first-order form. Regarding the chirality, the Galilean invariance preserves the chirality only if . The pseudo-conformal symmetry is valid for -solutions but does not preserve the chirality.
It is well-known that stationary or static solutions play a pivotal role in the dynamics. Here, a solution to (CM-DNLS) is static if and only if it has zero energy. In [26], the authors showed that , called the ground state or soliton,
is the unique zero energy solution (and thus static) up to scaling, phase rotation, and translation symmetries. Note that this is a chiral solution. More generally, any nonzero traveling wave solutions (i.e., solutions of the form for some ) are given by up to scaling, phase rotation, and translation symmetries [26]. Applying the pseudo-conformal transform (1.3) to , one obtains an explicit finite-time blow-up solution:
It is important to note that is neither of finite energy (), nor chiral ().
Let us briefly discuss previously known results on (CM-DNLS). De Moura and Pilod [14] proved the local well-posedness in for all . It is also locally well-posed in for [26]. The ground state is a threshold for global regularity, i.e. global existence of strong solutions. To this regard, Lax pairs provide significant information for the subthreshold . If in , then the solution is global and moreover, for all [26]. Subsequently, the local well-posedness on was improved to by Killip, Laurens, and Vişan [40]. At the threshold (), by adopting [43], -solutions are global, as [26]. The dynamics above the threshold ( were also studied. Gérard and Lenzmann [26] employed the Lax pair structure to construct -soliton solutions of the form
where the residues and the pairwise distinct poles for solve a complexified version of the classical Calogero–Moser system. These -soliton solutions blow up in infinite-time with as for any . Hogan and Kowalski [28], using the explicit formula, showed the existence of possibly infinite time blow-up solutions with mass arbitrarily close to the threshold . The first finite-time blow-up solutions were constructed by the authors and K. Kim [41], arising from smooth chiral data. This result addresses to the threshold problem for global regularity. Moreover, it is remarkable that although (CM-DNLS) is completely integrable, it admits finite-time blow-up solutions. Additionally, the zero dispersion limit of (CM-DNLS) was investigated by Badreddine [5]. We also refer to [4, 3] for the periodic model.
Our main theorem concerns the global dynamics beyond the threshold, so called, soliton resolution. We show the soliton resolution in a fully general setting, without radial symmetry and size constraints. We use notation for modulated functions in (2.1).
Theorem 1.1 (Soliton resolution for (CM-DNLS)).
Let be a solution to (CM-DNLS) with initial data , where is its maximal forward interval of existence.
(Finite-time blow-up solutions) If , there exist an integer with , a time , modulation parameters for , and asymptotic profile so that admits the decomposition
(1.4) |
and satisfies the following properties:
-
•
(Asymptotic orthogonality, no bubble tree) For all ,
(1.5) -
•
(Convergence of translation parameters) exist for all .
-
•
(Further information about ) We have . In addition, if , then .
-
•
(Bound on the blow-up speed) We have , and
-
•
(Chiral solution) If , then each component in the decomposition is chiral, i.e. .
(Global solutions) If and , then either scatters forward in time,
or there exist an integer with , a time , modulation parameters for , and so that admits the decomposition
and satisfies the following properties:
-
•
(Asymptotic orthogonality, no bubble tree) For all , denoting the translation parameters of , we have
-
•
(Convergence of velocity) exist for all .
-
•
(Further information about ) We have . We also have a further regularity, .
-
•
(Bound on the scale) We have , and
-
•
(Chiral solution) If , then we can choose each component in the decomposition to be chiral.
Soliton resolution is widely believed to occur in many dispersive equations. It suggests that any generic global-in-time solution asymptotically decouples into a sum of solitons (or similar solutions such as breathers) and a radiation term that goes to zero in some sense. For blow-up solutions, in various models, it is believed that each blow-up profile is a sum of modulated solitons. This type of results were conjectured for KdV equation in [23, 57] from the numerical simulations. The rigorous proof of soliton resolution was first demonstrated in various integrable PDEs via the inverse scattering method. To refer to just a few, see [22, 21] for KdV, [50] for mKdV, and [58, 52, 51, 46, 7] for 1-dimensional cubic NLS. Also refer to [37, 38] for Derivative NLS.
For non-integrable dispersive and wave equations, soliton resolution
has been studied in several models. In wave equations, such as the
energy-critical nonlinear wave equation (in various dimensions) and
energy-critical equivariant wave maps [19, 12, 39, 17, 20, 18, 33, 11, 35].
For damped Klein-Gordon equations, soliton resolution for global solutions
was established in [8, 13, 29].
For Schrödinger-type equations, soliton resolution was established
for the equivariant self-dual Chern–Simon–Schrödinger equation
(CSS) in [42]. In the context of parabolic
equations, soliton resolution has been studied by several authors
for the harmonic map heat flow [34, 36]
(or references therein) and the energy-critical semilinear heat equation
[2]. Among others, the authors in [36]
proved a version of continuous-in-time soliton resolution without
symmetry. As seen from the list, most results in non-integrable models
were achieved under symmetry constraints, excluding moving solitons.
Gauge transform.
(CM-DNLS) has an extra structure, so called gauge transform. This property is shared with DNLS. Define the gauge transform
Then, new variable solves
(-CM) |
(-CM) is a gauge transformed Calogero-Moser derivative NLS. All symmetries are transferred accordingly. The conservation laws of energy, mass, and momentum are given by
(1.8) |
where is the Hilbert transform. The virial identities are
(1.9) |
It is noteworthy that (-CM) admits the Hamiltonian formulation
where is a functional derivative with respect to . In other words, is a symplectic derivative with respect to the standard symplectic form . Moreover, as is a complete square, (-CM) is a self-dual Hamiltonian equation. See more details in [41]. The static solution of (CM-DNLS) is transformed as a static solution to (-CM)
Note that we chose the minus sign in the transform to ensure that is positive. is not real-valued, and and exhibit different decays. However is positive real-valued, which is a technical benefit of working with (-CM). In the main body of analysis, we will prove the soliton resolution for (-CM), and then, using the gauge transform and its inverse, we will obtain Theorem 1.1.
Theorem 1.2 (Soliton resolution for (-CM)).
Let be a solution to (-CM) with initial data , where is its maximal forward interval of existence.
(Finite-time blow-up solutions) If , there exist an integer with , a time , modulation parameters for , and asymptotic profile so that admits the decomposition
and satisfies the following properties:
-
•
(Asymptotic orthogonality, no bubble tree) For all ,
(1.10) -
•
(Convergence of translation parameters) exist for all .
-
•
(Further information about ) We have . We also have a further regularity, . In addition, if , then we have .
-
•
(Bound on the blow-up speed) We have , and
(Global solutions) If and , then either scatters forward in time,
or there exist an integer with , a time , modulation parameters for , and so that admits the decomposition
(1.11) |
and satisfies the following properties:
-
•
(Asymptotic orthogonality, no bubble tree) For all , denoting the translation parameters of , we have
-
•
(Convergence of velocity) exist for all .
-
•
(Further information about ) We have . We also have a further regularity, .
-
•
(Bound on the scale) We have , and
1. Novelty and Method. Our proof does not rely on the complete integrability. To our knowledge, this is the first non-integrable proof of soliton resolution in Schrödinger-type equations without radial symmetry and size constraints. We bypass time-sequential soliton resolution and directly prove continuous-in-time soliton resolution. The nonnegativity of energy is crucial. We believe our argument is applicable to other models with nonnegative energy, such as, wave or Schrödinger map, Chern-Simons-Schrödigner, and so on. A similar idea was used in NLS under threshold condition by Merle [43] or Dodson [16].
2. No bubble tree. (1.5) and (1.10) indicate that there is no bubble tree. i.e. any two bubbles maintain a distance larger than the scales of both. We believe this is natural. If a bubble tree existed, there would be a discontinuity in the soliton configuration in (CM-DNLS) along with the gauge transform. See more detail in Remark 5.2. As a similar result, in 1D harmonic map heat flow, there is no finite time bubble tree [56]. So far, finite-time bubble trees are not yet constructed in any model, while there are several results of infinite time bubble tree construction [55, 15, 30, 31, 32].
3. Global solutions. We prove finite-time blow-up cases first, and then take the pseudo-conformal transform to obtain results for global solutions. This is why we need to assume for global solutions, while suffices for finite-time blow-up solutions. After taking the pseudo-conformal transform, the translation parameter becomes the velocity of the Galilean boost, , and the scaling parameter becomes . This results a multi-soliton configuration with moving solitons at constant velocities.
For -solutions, thanks to the pseudo-conformal transform, the linear scattering of radiation part is easily obtained. In particular, this adresses the subthreshold problem, i.e. when and , then has to scatter. A remaining question is whether -global solutions may exhibit different dynamics other than -solutions. At current status, even for small solution in , the (linear or modified) scattering is not known. In view of results of other cubic equations in 1D, it is unclear whether the linear scattering occurs for -solutions.
A multi-soliton example constructed by Gérard–Lenzmann [26] does not belong to due to the slow decay of solitons. Still, their examples meet our criteria in Theorem 1.1. On the other hand, even if , each soliton component in the multi-soliton configuration (1.11), and thus . One might find this decomposition unsatisfactory. If one wants all components to belong to , then one can simply truncate tails of solitons. More precisely, one can replace with ( such that .
Outline of the proof.
As mentioned above, we prove Theorem 1.2 for the finite-time blow-up solutions and then use the pseudo-conformal transform to obtain result for global solutions case. And then we use the gauge transform to obtain Theorem 1.1. Thus, in main analysis we consider a finite-time blow-up solution to (-CM) .
The first ingredient is the variational characterization of . In fact, is the unique zero energy solution up to symmetries, and thus also a static solution. Furthermore, this also tells us the proximity of a small energy function to , as stated in Lemma 3.2. More specifically, if , then with . This is due to the nonnegativity of the energy and achieved near blow-up time. Also note that this is a stronger variational feature than that of mass-critical NLS. This allows us to extract a soliton from .
The second ingredient, one of our main novelty, is the energy bubbling estimate. After the first decomposition with orthogonality conditions on , we have the following improved estimate (Lemma 3.4):
(1.12) |
where is a large parameter depending on and is a smooth cut off on . (1.12) is motivated from a nonlinear coercivity estimate for (CSS) [42]. In (CSS), due to a special property of its nonlinearity they have a stronger estimate on the outer radiation. Then authors in [42] obtain the soliton resolution consisting of a single soliton. (1.12) benefits from the nonnegativity of energy. More importantly, we will take advantage of the energy conservation, which is located above the critical scaling.
For blow-up solutions, since and , the right-hand side of (1.12) is not just small but goes to zero. In particular, we have a quantitative control of the inner part of radiation . However, in general, the outer part of radiation may not go to zero. Also, can be large. Instead, we have good control of . Indeed, we will have a dichotomy: either
or
In fact, if the former is false, then the latter holds true sequentially in time. However, we will prove the latter holds for all . We refer to this as the no-return property, (H1) in Lemma 4.2. We prove (H1) by observing the difference in exterior mass between sequences depending on whether holds or not. Hence, extracting multi-soliton configuration benefits from two levels of conservation laws, mass and energy. Thanks to no-return property, we do not get through time-sequential soliton resolution, but directly prove continuous-in-time soliton resolution.
In the former case of the dichotomy, we have a good quantitative estimate for and so we can verify that converges to an asymptotic profile. This ends the soliton decomposition. In the latter case, we can reapply the decomposition to extract the second soliton from and arrive at the above dichotomy for the outer radiation part. We can iterate this procedure. Since the mass drops by at each step, it ends at finitely many steps. At last, we arrive at the multi-soliton configuration:
The radiation with a uniform bound converges to an asymptotic profile in . Along the way, we also prove behaviors of the modulation parameters, and . Finally, it is noteworthy that we verify a nonradial version of no bubble-tree condition, Proposition 4.4:
This is a consequence of in (1.12), which provides a quantitative bound on the interaction between soliton and . We are not sure if such no-bubble tree property is a special feature of this model. One can observe this no-bubble tree condition is consistent with the gauge transform between (CM-DNLS) and (-CM) (Remark 5.2).
Organization of the paper.
In Section 2, we introduce notation and preliminaries for our analysis. In Section 3, we review a standard variational argument and prove the energy bubbling estimate, which is a core proposition of our analysis. In Section 4, we prove the multi-soliton configuration via an induction argument. In this step, we also show the no-return property and confirm that the multi-soliton configuration holds true continuously in time. In Section 5, we complete the proofs of main theorems by applying the pseudo-conformal transform and the gauge transform.
Acknowledgments.
The authors are partially supported by the National Research Foundation of Korea, NRF-2019R1A5A1028324 and NRF-2022R1A2C1091499. The authors appreciate Kihyun Kim for helpful comments on the first manuscript.
2. Notation and preliminaries
In this section, we collect notations and frequently used formulas. For quantities and , we write if holds for some implicit constant . For , we write if and . If depends on some parameters, then we write them as a subscript. For , we write if for any there exists such that if , then . We also write simply . When the quantities are function of time, the estimates are uniform in time, unless stated otherwise.
Denote a smooth cut-off function by where with and on . It is possible to choose such that for some . We will also frequently use the outer cut-off
and a truncation on the centers of solitons, . We also use the sharp cut-off on a set by . We write the inhomogeneous weight by .
The Fourier transform (on ) is denoted by
with its inverse . We denote by the Fourier multiplier operator with symbol , that is, , and then where is the Hilbert transform:
We note that . We denote by the Cauchy–Szegő projection from onto the Hardy space :
Then, we have
We use the real inner product given by . For given modulation parameters, , we denote a (inversely) modulated function by
(2.1) |
Then, we have and . We also note that
We have some algebraic identities with respect to ,
Lemma 2.1 (Formulas for Hilbert transform).
We have the following:
-
(1)
For , we have
(2.2) in a pointwise sense.
-
(2)
For , we have
(2.3) -
(3)
If and , then we have
(2.4)
Lemma 2.1 is fairly standard. For the proof, see [41] Appendix. Next, we state a useful lemma coming from the nonnegativity of energy. A similar idea was first used in [43] for the threshold dynamics of NLS.
Lemma 2.2 (Nonnegativity of energy).
Let , be real-valued. Then, for a -solution to (-CM), we have
(2.5) |
Proof.
We use (-CM) to compute the left hand side,
We use the nonnegativity of energy for to estimate
Then, the discriminant inequality gives
∎
2.1. Linearization of (-CM)
As mentioned earlier we will proceed all the analysis for the gauged equation (-CM). Here, we review the linearization of (-CM) around . All the material here was investigated in [41], where more details can be found.
The form of energy in (1.8) represents self-duality. Introducing the operator
the energy (1.8) can be rewritten as
In analogy with [6], we call the nonlinear operator the Bogomol’nyi operator, and the soliton solves the Bogomol’nyi equation . Similarly to , is the unique solution to the Bogomol’nyi equation up to symmetries. We first linearize the Bogomol’nyi operator . We write
where the linearized operator and the nonlinear part are given by
The -adjoint operator of with respect to is given by
and using one can write . Now we linearize (-CM) as
where is the linear part, and is the nonlinear part. If one linearize at , then using one derive the self-dual factorization
We will modulate out the kernel directions of and obtain the coercivity of the orthogonal part. For this purpose, we recall kernel information and the coercivity estimate from [41]:
where is the -scaling generator,
Note that each kernel element is a generator of symmetry, phase rotation, scaling, and space translation. For the coercivity estimate, due to a degeneracy of , we need to use an adapted Sobolev space defined by a norm;
Note that with . In this section, we use notation of truncated kernel elements, , , .
Lemma 2.3 (Coercivity for on , [41]).
We have a coercivity estimate
For later use, we will denote a truncated version of adapted Sobolev space, by a norm
(2.6) |
Note that the second term is not truncated by . It covers global interaction of and .
3. Decomposition and energy bubbling
In this section, we take a preliminary decomposition (bubbling) of a blow-up solution. If a solution blows up in finite time, we can decompose the solution into a modulated soliton and radiation such as . This is due to a variational characterization stating that is the unique nontrivial zero energy solution. A similar argument was used in the context of the self-dual Chern–Simons–Schrödinger equation (CSS) in [42]. The proof of this part is fairly standard, so we state them in Lemma 3.2 and 3.3 and postpone the proof to the Appendix A. This is the first step of bubbling of the solution. However, for the soliton resolution, we should be able to extract a sequence of bubbles from the radiation part. In this step, we will use a highly nontrivial energy bubbling estimate (3.4). The estimate (3.4) controls the interaction of and the radiation part at each step and the energy of the exterior part of the radiation. Hence, (3.4) is crucial in the multi-bubble decomposition in the next section. These results are summarized in the following proposition.
Proposition 3.1 (One bubbling).
Let be fixed. There exist small parameters such that the following hold: For satisfying
we have the decomposition as follows;
-
(1)
(Decomposition) There exists unique continuous map satisfying
(3.1) and smallness
Moreover, part is .
-
(2)
(Estimate for ) We have
(3.2) -
(3)
(Energy bubbling) We have
(3.3)
The proof of Proposition 3.1 is divided into three lemmas. First, tube stability (Lemma 3.2) is a consequence of the variational structure of . Next (Lemma 3.3), we decompose into with appropriate orthogonality conditions on , which is an application of the implicit function theorem. The proofs of Lemma 3.2 and Lemma 3.3 are fairly standard and thus postponed to the Appendix A.
Lemma 3.2 (Tube stability for small energy solutions).
For any and , there exists such that the following holds. For nonzero with and , there exist and such that
where .
Since is close to a modulated soliton, the decomposition is possible in a standard way. We define the soliton tube by
Lemma 3.3 (Decomposition).
For any sufficiently small , there exists such that the following hold: For any , we have a unique decomposition,
-
(1)
(Decomposition) There exists unique continuous map satisfying
and smallness
Moreover, part is .
-
(2)
(Estimate for ) We have the estimate (3.2) for .
The remaining part, energy bubbling, is the heart of our analysis. This provides an improved estimate on the radiation . Specifically, it upgrades to . This leads to the vanishing of the inner radiation and , and subsequently deduces the asymptotic decoupling of the soliton and . Observing that is absent in , one can expect that there could be a nontrivial concentration in .
Lemma 3.4 (Energy bubbling).
In fact, (3.4) is a combination of a coercivity estimate of the inner radiation part and the energy control of the outer radiation part . The former is a result of linear coercivity (Lemma 2.3). The latter is obtained from the nonnegativity of energy. Although we do not have a control over , we can instead maintain the smallness of the energy of . We believe that this approach is robust enough to be applied to other models with nonnegative energy. There is a similar (even stronger) nonlinear coercivity inequality for (CSS) [42, Lemma 4.4],
It is worth noting that it also controls the outer radiation . This is due to a special property of (CSS), so called, the defocusing nature at the exterior of a soliton. From this specific property, the authors in [42] were able to achieve soliton resolution with a single bubble. However, in (-CM), we have no coercivity of . But the energy control (3.4) indicates that after extracting a soliton , there could be another bubble from since is still small.
Having Lemmas 3.2, 3.3, and 3.4, we deduce Proposition 3.1. Indeed, for a given , there are and . Then, there exist and to satisfy the statement in Proposition 3.1. We close this section with the proof of Lemma 3.4.
Proof of Lemma 3.4.
We decompose the domain into inner and outer regions. We will choose parameters such that . First, observe from the definition and assumption that
We claim a preliminary bubbling for suitable and
(3.5) |
We will prove (3.5) in Step 1, 2, and 3. Then take an average over to finish the proof.
Step 1. In this step, we decompose the domain into and and claim that
(3.6) |
In (3.6), the first term is the inner term, the second term is the interaction term, and (3.12) is the intermediate and outer term. We will handle (3.12) in Step 2 and Step 3. We start by linearizing the energy;
where is the nonlinear term,
The first term of the is perturbative and absorbed in by
Thus, it suffices to consider
We decompose into inner and outer parts,
(3.7) | ||||
(3.8) |
We first show (3.8) is perturbative. Using and (2.3), we compute the first term of (3.8) by
So, we deduce
We also have
Hence, we have . Now we estimate the main part (3.7),
(3.9) | ||||
(3.10) | ||||
(3.11) | ||||
(3.12) |
For the linear term in the inner region, we use the coercivity, Lemma 2.3, to estimate (3.9)
(3.13) |
We now estimate the interaction term (3.10).
(3.14) |
For (3.14), by integration by parts, we have
Hence, we have
(3.15) |
Now, we control (3.11). We compute
(3.16) | ||||
(3.17) | ||||
(3.18) |
We have
and this implies
(3.19) |
For the last term of (3.19), using and (2.4), we have
and hence, we estimate
(3.20) |
Now, we estimate . Using and (2.2) with and , we deduce
Thus, we have
(3.21) |
Collecting (3.19), (3.20), and (3.21), we have
(3.22) |
Now we finish to prove (3.6) by summarizing from (3.13), (3.15), and (3.22).
Step 2. We estimate (3.12) by decomposing into intermediate and outer regions, and in this step we estimate the intermediate part and summarize in the main claim, (3.32). We first compute
(3.23) | ||||
(3.24) | ||||
(3.25) |
For (3.24), we have
(3.26) |
Using and (2.3), we compute
So, we estimate the second term of (3.26) such as
Therefore, we have
and
(3.25) is the outer term and will be estimated in Step 3. There is an outer part in (3.23), too. We extract it in the following:
(3.27) | ||||
(3.28) |
For the last term (3.28), we have
We use abbreviated notation for the integrals in the intermediate region and outer region. For a function with , e.g. , we denote
Now, we decompose (3.27) into
(3.29) | ||||
(3.30) |
(3.30) is the outer part. We control the intermediate term (3.29). We observe that
On , we have
where
Thus, we have
As a result, we have arrived at
(3.31) | ||||
(3.32) |
(3.31) will be handled in Step 3. The first term of (3.32) is a part of .
Step 3. We control the outer part (3.31), . A crucial feature of this step is to extract the energy of the outer radiation as a lower bound. We observe that the middle term of (3.30) become
Thus, we obtain
(3.33) | ||||
(3.34) |
We observe that (3.33) is a complete square, and the first term of (3.34) is positive. So, discarding and the first term of (3.34), we reduce the outer part to
(3.35) | ||||
(3.36) |
We have to find out the outer energy from (3.35). To do this, we deal with the cut-off error. We claim that
(3.37) |
We divide the first term of (3.35) into the intermediate and outer terms,
(3.38) |
For the outer term, we compute
(3.39) | ||||
(3.40) |
For the second term of (3.39), we deduce
(3.41) |
We control (3.40) as
(3.42) |
For the last term of (3.42), we have
(3.43) |
Here comes from . Therefore, gathering (3.38), (3.41), (3.42), and (3.43), we derive that
(3.44) |
By the pointwise estimate , we have
This implies that
and we conclude the claim (3.37).
Now, using (3.37), we compute (3.35) as
In addition, we check that since
So, discarding for a small , we obtain
(3.45) |
We control (3.36). We have
(3.46) | ||||
(3.47) |
For (3.46), we have
(3.48) |
Now we estimate (3.47). For the second term of (3.47), by using and (2.2) with , we compute
(3.49) |
To cancel out the second term of (3.49), we use (3.25), . We have
(3.50) | ||||
(3.51) |
and
In addition, the second term of (3.49) is canceled by (3.51). Therefore, we derive
(3.52) |
Collecting (3.45), (3.48), and (3.52), and discarding first two positive terms in (3.52), we deduce
(3.53) |
Finally, combining (3.6), (3.32), and (3.53), we arrive at
(3.54) |
where is the implicit constant in (3.54) depending only on and . Taking small , large and small , we can conclude (3.5).
Step 4. Now, we finalize the proof by taking a logarithmic average over . One goal is to eliminate the term . Indeed, we replace by , and take a logarithmic integral of (3.5) on . Then, we have
and by (3.5) this yields (after taking larger if necessary)
(3.55) |
Moreover, we redo all the argument with , by taking smaller depending on . Then we have
(3.56) |
Hence, combining (3.55) and (3.56), and renaming by , and by , we conclude (3.4). This finishes the proof. ∎
4. Multi-soliton configuration
In this section, we establish a multi-soliton decomposition by iterating one bubbling procedure in Proposition 3.1. In this process, we bypass a time-sequential soliton resolution and directly prove the soliton resolution continuously in time. From this section, for given initial data to (-CM), we denote the mass and energy for by . We assume that the solution to (-CM) blows up at the time . Then as , and there exists a time such that on . We apply the decomposition, Proposition (3.1), and then there exist such that
(4.1) |
(4.2) |
and
(4.3) |
(4.3) indicates that we can control the inner part of radiation and the outer radiation energy . However, we cannot control the outer part of radiation . More precisely, it is possible that , and one does not expect that the radiation term to converge to some asymptotic profile.333It is instructive to compare to the self-dual equivariant Chern-Simons-Schrödinger equation (CSS) in [42]. There, the authors obtain and proceed to show the soliton resolution with a single bubble. In this case, for the outer radiation, we sustain near-zero energy . This bound enables us to reapply Proposition 3.1 to extract another bubble from . We will iterate this procedure until the radiation satisfies and thus converges to some asymptotic profile. At each bubbling out, the mass of radiation drops by . So, the iteration should halt in finite steps. Indeed, for the second bubbling, the radiation satisfies either or . In other words, either
The latter case corresponds to , which results in convergence to an asymptotic profile. Here is a single bubble blow-up solution. For the former case, it is possible to reapply Proposition 3.1 to to extract the second bubble in a sequence of times: there exists a sequence such that as and . Then one deduces the sequential small energy for , Hence, we have the decomposition:
where , and . Moreover, we have the (4.3) bound for the next step:
And then, we can check whether or not to proceed to the next step. This allows us to obtain sequential in time soliton resolution. However, we will bypass the sequential soliton resolution and directly prove continuous in time soliton resolution. For this goal, we improve
(4.4) |
Equipped with (4.4), we can derive continuous-in-time soliton resolution as explained above. In the rest of this section, we establish the multi-soliton decomposition by an induction argument. Among other things, as we work in a nonradial setting, a crucial part of the induction step is to control the translation parameters:
4.1. Induction
In order to implement the previously described strategy, we will prove via an induction argument. Fix an initial data and the blow-up solution as above. Fix , and depending on as in Proposition 3.1. At the zeroth step, we use conventions , , and .
For each , we define the induction hypothesis statement for -soliton configuration, , as follows;
For , there exist a time , modulation parameters , and radiation terms on satisfying the following:
-
(H1)
(No-return property) We have
(4.5) -
(H2)
(Further decomposition) There exists a such that there exist modulation parameters and radiation term on which satisfy the following:
with the smallness and the energy bubbling
(4.6) Moreover, for , we have
(4.7) where
(4.8) and
(4.9) -
(H3)
(Soliton decoupling) We have . If , we have
(4.10) Moreover, we have
(4.11) and if for some , we have
(4.12) -
(H4)
(Convergence of the translation) exists and .
Once we have with -soliton decomposition, we encounter a dichotomy for the next step. Define the statements and by the followings:
Now, we are ready to state the initial case and induction step.
Lemma 4.1 (Initial case).
Let be a blow-up solution to (-CM) with initial data . Then is true.
Lemma 4.2 (Induction).
Assume that is true. If is true, then is true.
We also claim that the induction stops within finite steps with respect to the initial value of mass.
Lemma 4.3 (Halting of induction).
There exists an integer with such that and hold true. Moreover, we have uniformly in time.
At each step, we have a dichotomy, or . If holds true (and ), we can extract a soliton. Otherwise, we have and then . By a similar argument to [44], converges to some asymptotic profile.
In the above configuration, it is possible that . This a nonradial version of bubble tree, i.e., . Such a bubble tree naturally appears in statements of other soliton resolutions for blow-up solutions. However, to our knowledge, there is no finite-time bubble tree construction in relevant models. Moreover, there is a specific clue indicating the absence of a bubble tree in (-CM). If there were a bubble tree in (-CM), then after taking the inverse gauge transform , there would be discontinuities in phase at some points for each soliton, which is somewhat unusual. See Remark 5.2 for more details. Fortunately, we are able to verify that there is no bubble tree in (CM-DNLS) and (-CM).
Proposition 4.4 (No bubble tree solutions).
Suppose is true for . For any , we have
(4.13) |
From the continuity of and , we have either is or .
4.2. Proof of the induction
We begin with defining abridged notations for modulation parameters and transformations:
(4.14) |
(4.15) |
Moreover, we denote by ,
(4.16) |
Lemma 4.5.
Let and satisfy relations in (4.9). Then, for any , we have
(4.17) |
Proof.
Since , it suffices to show, for any ,
(4.18) |
We show (4.18) by induction in descending order. For the initial case, using the relation (4.9), we have
(4.19) |
For a fixed , we assume that (4.18) is true for all . Then, we want to show (4.18) for , and it suffices to prove
(4.20) |
Similarly to (4.19), we can show (4.20) for scaling and phase rotation parts. For the translation part, we compute the translation part of RHS of (4.20) by
Using , we have
and this proves (4.20). Therefore, by the induction, we have (4.18), and finish the proof. ∎
When we have -soliton configuration, we naturally have an asymptotic decoupling.
Lemma 4.6 (Decoupling of localized mass).
We suppose that holds true. Let . Then, we have
(4.21) |
Proof.
From the assumption, we have the decomposition (4.7) with (4.10) and (4.12). We compute by
(4.22) |
Our goal is to show . For the first term, if , we deduce
If , thanks to (4.12),
We compute the second term, by using ,
We decompose
(4.23) | ||||
(4.24) |
We have
For, (4.24), we have
Therefore, we arrive at , and conclude
∎
Now, we prove the initial induction step.
Proof of Lemma 4.1.
(H2) was explained in (4.1), (4.2), and (4.3). Since there is only one soliton for now, it only remains to prove the convergence of translation parameter , (H4). In fact, this can be shown by a similar argument in [49]. Instead, we present a different argument that will be employed in later induction step. First, we claim
(4.25) |
Suppose not. Then, there exists a sequence such that . Without loss of generality, we may assume . We omit the time sequence dependency , if there is no confusion. Denote , and then we have . According to (2.5), we have
(4.26) |
Integrating (4.26) on , and taking , we have
(4.27) |
Using (4.1) with the notation in (4.8), we compute
(4.28) | ||||
(4.29) | ||||
By a direct computation, we have
(4.30) |
Thanks to (4.3), we have , and this implies
(4.31) |
Thus, by (4.30) and (4.31) with and , we have
Since , we deduce
Next, we finish showing (H4), that is, and . We assume that does not exist. Then, there exist so that and , with , and , due to (4.25). By symmetry and taking translation, we may assume . We claim that there exist a uniform constant such that
(4.32) |
To show a contradiction, we observe the exterior mass, with . Integrating from to by (2.5), we have
By a similar argument to (4.31), we have . Thus, we have
and . Moreover, since , we have , and . Thus, we derive
(4.33) |
For , we have
(4.34) |
Therefore, by (4.34), we have
(4.35) |
From (4.33) and (4.35), we deduce
Thus, choose arbitrarily small and taking , we have . This prove (4.32). Once we have (4.32), then we can find another sequence such that for any , since is . This obviously makes a contradiction by reapplying the uniform gap (4.32) of and . Hence, we conclude that exists, and we finish the proof. ∎
Proof of Lemma 4.2.
Step 1. (-th sequential decomposition) We assume that and hold true. We first claim -th time-sequential decomposition. From , we have
(4.36) |
and so there is a time sequence so that . Then, (4.6) in implies, for large ,
So, we apply the decomposition (Proposition 3.1) for , and then there exist such that
(4.37) |
with the smallness ,
(4.38) |
and . Moreover, by (3.3), we have
(4.39) |
We denote by , and we just denote by if there is no confusion. From , we have continuous in time -th configuration
Define via (4.8) and (4.9). Using (4.8) and (4.37), we deduce
Therefore, we have the sequential -th decomposition
(4.40) |
Note that the second inequality of (4.40) comes from (4.39) and (4.6) in . In addition, we obtain as by (4.36), (4.38), and the definition of .
Step 2. (Boundedness of translation) Here we show a partial information of translation parameter , (H4) (of ). Indeed, we claim that
(4.41) |
The proof of (4.41) follows a similar argument to (4.25). So here we only sketch it. From , converges to , and for all . If (4.41) fails, passing to a subsequence (again denoting it by ), we have . Without loss of generality, we may assume . By a similar argument to (4.27), taking , we have
(4.42) |
On the other hand, by (4.21), we have
Therefore, taking , and , we derive
(4.43) | ||||
(4.44) | ||||
(4.45) |
We compute (4.43). For , we have
(4.46) |
Here, we used which comes from and . For , we have
(4.47) |
For (4.44), we remind and . We have
(4.48) |
Here, we note that
By (4.46), (4.47), and (4.48), we get to
For the last term (4.45), since on , we have . Therefore, we deduce
and this contradicts to (4.42). This finishes to prove (4.41).
Step 3. (No return property) In this step, we show (H1). This part is crucial to upgrade the sequential soliton resolution to continuous in time resolution. In this step, we use the convergence of for ((H4) in ). Suppose
Then, we can find a subsequence of which is denoted by , and another sequence such that and
(4.49) |
As explained above, along the sequence of , we can extract another soliton , while there is no other soliton along . However, the nonnegativity of energy implies that the exterior mass of the last soliton is Lipschitz in time. This will make a contradiction.
Indeed, passing a subsequence, we have with by the Step 1. Taking translation, we may assume . In order to bring out a contradiction, we investigate the exterior mass, for a sufficiently small to be chosen later. Thanks to (2.5), we have
Integrating this, we deduce
(4.50) |
Now, we estimate on each sequence and . From (4.21), we have
(4.51) |
for any . Firstly, we use (4.51) for to obtain (4.52). Indeed, we take a further decompose by using Proposition 3.1, and then have
We also estimate by a similar argument to (4.44). Then we derive
Moreover, since , after taking large so that , we have
Therefore, we arrive at
(4.52) |
On the sequence , from (4.49), we know . Thanks to energy bubbling (4.6), we also know , and hence we have
From this and energy bubbling , we estimate
(4.53) |
Now, we are ready to bring out the contradiction. From (4.50) and as , we have
(4.54) |
Using DCT, we have
Since as , we deduce . This is the reason what we need the convergence of as . So, we have
(4.55) |
On the other hand, again from (4.21) with , we have
On , by arguing as like (4.52), we have
Thus, we have
(4.56) |
Thanks to (4.53), we have
Therefore, we can take sufficiently small so that
for given small . We remark that the choice of does not depend on . Thus, we obtain
Now, taking , we have for arbitrary small , which leads a contradiction. Hence, we conclude , and this proves (H1).
Step 4. We finish to prove (H2) and (H3). By Step 3, we have (4.5). This means that by (4.6) there exists a with satisfying
Therefore, we can apply the decomposition Proposition 3.1 for on , and we have (H2). Now, we prove (H3). From Proposition 3.1, we have
and deduce by (H1). From , we have , and deduce that , (4.11), and .
Now, we prove (4.10). Assume not. Then, we have and , so there exists a sequence such that and with as . Further taking subsequence if necessary, exists, and by the step 2, we also have . Let , and applying the variational argument to (Lemma A.1), we have
(4.57) |
for some fixed . We have
(4.58) |
and this is a contradiction. So, we derive (4.10).
Now, we show (4.12) by using induction in descending order for . The initial case already was shown in (4.10). We assume that (4.12) holds true for . To show (4.12) for , we use the contradiction argument. We suppose (4.12) is not true for . Then, we can find a sequence such that with and . We have by the assumption . We consider
Using (4.17), we rewrite
Note that is given by (4.16). On the other hand, from the decomposition , we have
where as given above. By the induction hypothesis with , we have for , and
(4.59) |
Moreover, from (4.6) and the assumption with (4.11), we have
(4.60) |
Gathering (4.57) (4.59), and (4.60), we arrive at
Therefore, as in (4.58), we have
which lead a contradiction. Hence, (4.12) holds true for . This finishes the proof of (4.12).
Recall that we have shown that in Step 2. So, it suffices to show converges as . Suppose not. Then, there exist so that and , with . As in the proof of Lemma 4.1, we can show the uniform gap (4.32) of . Indeed, arguing as above, we have
From , we have . Thus, we have
(4.61) |
Now, we estimate . We have
(4.62) |
We also have
(4.63) |
Therefore, by (4.61), (4.62), and (4.63), we have
and taking small and there is a uniform bound on such as . As in the proof of Lemma 4.1, we can conclude (H4). This finishes the proof. ∎
Next, we show that the induction has to stop in finite steps, since at each step, the mass drops by soliton mass .
Proof of Lemma 4.3.
Let be such that . Assume that for all , is true. Then, by Lemma 4.1 and 4.2, we have for all , and
(4.64) |
Thanks to (4.21) with , we have
There exist small and such that for , we have
By (4.64), there exists a time sequence such that and . Let be
Then, we have
Now, applying Lemma A.1, we have , and this is a contradiction. Thus, there exists a such that is true.
Now, we prove . It suffices to show . Since is true, we have , and combining with (4.6), we conclude . ∎
We now end this section with proving that there is no bubble tree, (4.13).
Proof of Proposition 4.4.
Since for , it suffices to show as for . We first claim that, for ,
(4.65) |
Suppose not. Then, there exists a time sequence and a constant such that
We remark that and . To reach a contradiction, we consider . Thanks to (4.6), we have
(4.66) |
From the decomposition , we obtain
By renormalizing with , we have for large ,
Here, again according to (4.6), we have
This implies that as . Thus, we derive
for large . However, from the definition of and as , we have
and this is a contradiction. Thus, we have (4.65).
Now, we show the general case using an induction. We assume that for all . We will show that for ,
(4.67) |
Again, we take further induction on in descending order, as we already proved it for from (4.65). We assume that (4.67) is true for . Then, we want to show that (4.67) holds for . Suppose not, then we again find a sequence and a constant such that
In a similar manner to (4.65), we consider to derive a contradiction. The estimate such as (4.66) also holds true. From the decomposition for , and (4.17), we have
(4.68) |
Here, we recall given in (4.16), and we note that
Substituting with , we compute
(4.69) |
Here, we used that, for ,
which also can be shown by (4.17). We write
So, on , we again have, from ,
(4.70) |
We have . In addition, we have . Therefore, we reduce the right side of (4.70) to
By the induction hypothesis, for all . Thus we have
Thus, the right side of (4.70) becomes
and this means that
and we obtain . This makes a contradiction. Therefore, we conclude that , and by the induction, we deduce (4.67). This finishes the proof. ∎
5. Proof of Theorem 1.1 and 1.2
In this section we complete the proof of Theorem 1.1 and 1.2 based on the multi-soliton configuration derived in Section 4. We first prove the blow-up case of Theorem 1.2 and then use the pseudo-conformal transform to show the global solution case. For this part, we require that . Theorem 1.1 is proved using the gauge transform and from Theorem 1.2.
Proof of Theorem 1.2.
For a finite-time blow-up solution to (-CM), in Section 4, we have a multi-soliton configuration. That is, there exists an such that and are true. More precisely, there exists a such that for , there exist that satisfy the followings:
(5.1) |
and
(5.2) |
Moreover, we have
(5.3) |
To finish the proof for finite-time blow-up solutions, we remain to show that , in , and satisfies and
Step 1. Convergence of and regularity of .
Our first goal is to show that converges to an asymptotic profile in . This proof is motivated by [44]. We will truncate the outer region of each contracting soliton and show the convergence of in the truncated outer region. Then, using , we can conclude the convergence in . Define a smooth cutoff away from the centers of solitons for , by
Then, we have the outer convergence of the radiation.
Lemma 5.1 (Outer convergence).
There exists such that for any , we have in as .
Proof.
We first claim that for any , there exist and such that
(5.4) |
Denote . Then, we have
Here, . From the Duhamel formula, we have for ,
If we have
(5.5) |
then we have
So, taking sufficiently close to and then choosing to be small, we deduce (5.4). Here, we use the continuity of the flow at . Thus, we reduce (5.4) to (5.5).
Now, we prove (5.5). If necessary, we take larger so that
(5.6) |
Using (5.2), we have
(5.7) |
Moreover, interpolating (5.7) and mass conservation law, we have
(5.8) |
For the commutator term, similar to (5.7), we have
(5.9) |
For the quadratic term in the nonlinear terms, using (5.8), we have
The estimate of the nonlocal part is not as simple as above since the Hilbert transform may interfere with the truncation . In fact, we need to look inside and use the multi-soliton configuration. We will use some special relation between and . Moreover, we will use the commutation relation with . We first simplify by (5.8),
and use the decomposition
Using the pointwise bound from , we have
Now, we estimate interaction terms,
(5.10) |
It suffices to estimate
(5.11) |
for . Changing the variable with , we have
We recall the definition of , (4.16). By (2.4), we have
(5.12) |
We decompose (5.12) into
(5.13) | ||||
(5.14) |
For (5.13), applying (2.3), we have
Using and Hölder inequality, we have
For (5.14), further using (2.3), we have
(5.15) | ||||
(5.16) |
We first control (5.16). By integrating by parts and Hölder inequality, we have
Therefore, we have
For (5.15), from , we have
Therefore, we deduce
Next, the estimate of is performed in a similar manner. Indeed, we have
(5.17) |
Again thanks to (2.4), we have
Finally, is estimated as
Hence, we have
(5.18) |
Now, let us finish the proof. We first claim that is Cauchy in for any . We want to show that for any fixed and , there exists such that
(5.19) |
By (5.1) and (5.4), to prove (5.19), it suffices to show that for all . Given the definition of , (5.3), (5.6), and as for all , we have for all . Therefore, we conclude (5.19). That is, is Cauchy in for any . Thus, there exist for each such that in . By the uniqueness of the limit, we conclude that there exists such that in for any . Since is uniformly bounded in , we also have , and we finish the proof. ∎
Now, we show that in . We first prove that in for some . For any sequence , we know that is bounded. Therefore, there exists a subseqeunce such that for some . Moreover, by the Rellich–Kondrachov theorem, for any , we have in . In view of Lemma 5.1 and the uniqueness of the limit, we have in . Thus, by taking arbitrary large, we have .
We show further information of . From (4.21) with , we deduce . If we further assume , then from the virial identity, we have that is bounded for . Therefore, by Lemma 5.1 and Fatou’s lemma, we estimate
and we conclude .
Step 2. Pseudo-conformal bound .
Let . From (2.5), we deduce and
(5.20) |
We have
We also estimate for ,
Similarly, we have for ,
Using the decomposition of , we estimate
Since in , we have
Using (5.20), and then we obtain
Now, we take and estimate
(5.21) |
On the other hand, by Young’s inequality, we obtain
(5.22) |
Since in , . Hence, by (5.21) and (5.22) we have , or equivalently . From , we have for all . This finishes the proof of Theorem 1.2 for finite time blow-up solutions.
Step 3. Global solution case.
We use the pseudo-conformal transform to extend results to a global solution in . According to the local theory, if , then for its lifespan. Suppose that is a solution to (-CM) on the time interval . If scatters forward in time, then there is nothing to prove. Suppose that does not scatter forward in time. Denote . Then is a solution to (-CM) on . We claim that does not converge to in as . If were to converge, then we would have
Taking the inverse of the pseudo-conformal transform (indeed, , we would derive that
and this is a contradiction. Therefore, does not converge in . Let be the maximal forward time of existence of . We claim that , i.e. blows up at . If , then we have by a standard Cauchy theory with . Thus, we have as in . This is a contradiction and thus blows up at . We can apply Theorem 1.2 for the finite-time blow-up case. Hence, for some , has the decomposition
where the modulation parameters and satisfy the properties stated in Theorem 1.2. We can rewrite by as
Taking the inverse of the pseudo-conformal transform (indeed, , we have
(5.23) |
where and . Here, we denote the velocity of soliton by , in accordance with Galilean transform notation. Moreover, from the properties stated in Theorem 1.2, we have , and
for all . Moreover, we have
Using , and the virial identity (1.9), we have . Thus, we obtain as .
Now, we re-express in (5.23) as a canonical form using Galilean boost. Set , and we compute
Since , and using DCT in the asymptotics as , we can replace the pseudo-conformal factor with . Finally, from and , we have . Note that exists. On the blow-up side, is a position parameter, but on the pseudo-conformal side, is a velocity of each soliton. We rename as in the theorem statement. This finishes the proof of global solution case. ∎
Next, we provide the proof of Theorem 1.1. This is accomplished using the gauge transform and its inverse .
Proof of Theorem 1.1.
The gauge transform is a fixed-time nonlinear transform and a diffeomorphism in . Therefore, it suffices to justify the multi-soliton configuration under . In the procedure, since is a nonlinear transform, we need to handle the sum of solitons with care. Recall the inverse gauge transform,
In view of , and by the local mass decoupling (4.21) with , we decompose the phase
Here, does not depend on . Then, using we obtain
(5.24) | ||||
(5.25) |
We now compute the asymptotics of the phase functions in (5.24) and (5.25). From , , and in , we have
On the other hand, we have
Using ,
where
Now, we show that converges to a constant, independent of as . For , using and , we have
(5.26) |
as . By Proposition 4.4, we have or , the right side of (5.26) is or . If , we use and , we have
and this is also or . Therefore, again by DCT, converges to a constant, which we denote by , and we have
(5.27) |
Next, we consider (5.25). From in , we have in . By a similar argument, we have
in pointwise sense. Hence, we have
Thus, we have (1.4). Since the modulation parameters and remain unchanged, we also have (1.5), exists with , and . We also have , which implies . If , then , which means that . However, due to discontinuities at , we do not have control of . This completes the proof for finite-time blow-up solutions.
For global solutions in , we argue similarly to the proof of Step 3 of Theorem 1.2 to obtain the scattering or the soliton decomposition
(5.28) |
for some modulation parameters and . Using , we have .
Now, we show that for chiral solutions , we can maintain the chirality in the multi-soliton configuration. For the case , since we know in (1.4), and thus , there is nothing to prove. For the case , the solitons contain the Galilean transforms . Therefore, we might need to adjust the soliton configurations. Note that . So, we obtain
(5.29) |
In view of (5.29), if for some , then . If for all , then the multi-soliton configuration becomes asymptotically chiral. Therefore, by choosing the valid time large enough, we are fine. If for some , we might need to make a suitable change of the configuration. We first claim that
(5.30) |
Assume not, i.e. . We have for , from (5.29),
(5.31) |
This means that, applying to (5.28), since , we have the following decoupled norm,
However, we have
(5.32) |
which leads to a contradiction, thus proving (5.30). Moreover, (5.32) also shows that and . Under the condition , using (5.30) and (5.31), we are able to replace with a chiral soliton in the configuration. This finishes the proof. ∎
Remark 5.2.
One can observe that the no bubble tree property, as stated in Proposition 4.4, is natural. Indeed, in the proof of Theorem 1.1, if there were to be a bubble tree, i.e., , we could not derive (5.27) for . Instead, we would have
This implies that converges to a step function with a discontinuity. This means that in the multi-soliton configuration, some solitons have discontinuities in phase. We believe this is quite unnatural. Note that, so far, there is no finite-time bubble tree construction in any models. Therefore, the nonexistence of a bubble tree in (CM-DNLS) provides sharper information on the soliton resolution.
Appendix A Decomposition
We provide the proofs of Lemmas 3.2 and 3.3 . They are consequences of variational structure of the ground state .
Lemma A.1 ([26]).
Suppose that is a sequence such that
Then, there exist a sequence , and such that
Now, we prove the tube stability.
Proof of Lemma 3.2.
By scaling, we may assume . We suppose the tube stability fails. Then, there exist and a sequence in such that
and
(A.1) |
By Lemma A.1, we have weakly in . We write . So, we have weakly in , and we also have in and for . We have
(A.2) |
where
Since and , we have
Using (2.2) with , we deduce
We have . For any , from the isometric property of with , we can know that , and this means that . Therefore, we have as . This leads to . Now, thanks to the subcoercivity of ([41], Lemma A.3), we have in , and we also have in . This contradicts (A.1). ∎
Proof of Lemma 3.3.
(1) We equip with the metric , and equip with the induced metric from . We first decompose by . We define
where and . Then, we can check , and for , and for . Thus, by the implicit function theorem, is for and invertible at . Let . For given , there exist and maps such that for given , is unique solution to in . We also have that and .
Now, we want to prove the uniqueness of g on for some . From the -scaling, phase rotation, and translation invariances, we may assume . We choose and so that and . We assume that with and for . Then, by the scaling, rotation, and translation symmetries and changing the roles of and if necessary, we may assume , , , and . Then, we have and we deduce
So, we have . Since , we have , and thus we derive . By the uniqueness which comes from the implicit function theorem, we conclude .
(2) Let . Then, we have . Since , we have and (3.2). This finishes the proof. ∎
References
- [1] A. G. Abanov, E. Bettelheim, and P. Wiegmann. Integrable hydrodynamics of Calogero-Sutherland model: bidirectional Benjamin-Ono equation. J. Phys. A, 42(13):135201, 24, 2009.
- [2] S. Aryan. Soliton resolution for the energy-critical nonlinear heat equation in the radial case. preprint arXiv:2405.06005, 2024.
- [3] R. Badreddine. Traveling waves & finite gap potentials for the Calogero-Sutherland derivative nonlinear Schrödinger equation. preprint arXiv:2307.01592, to appear in Ann. Inst. H. Poincaré C Anal. Non Linéaire, 2023.
- [4] R. Badreddine. On the global well-posedness of the Calogero-Sutherland derivative nonlinear Schrödinger equation. Pure Appl. Anal., 6(2):379–414, 2024.
- [5] R. Badreddine. Zero dispersion limit of the Calogero-Moser derivative NLS equation. preprint arXiv:2403.00119, to appear in SIAM J. Math. Anal., 2024.
- [6] E. B. Bogomol’nyĭ. The stability of classical solutions. Jadernaja Fiz., 24(4):861–870, 1976.
- [7] M. Borghese, R. Jenkins, and K. D. T.-R. McLaughlin. Long time asymptotic behavior of the focusing nonlinear Schrödinger equation. Ann. Inst. H. Poincaré C Anal. Non Linéaire, 35(4):887–920, 2018.
- [8] N. Burq, G. Raugel, and W. Schlag. Long time dynamics for damped Klein-Gordon equations. Ann. Sci. Éc. Norm. Supér. (4), 50(6):1447–1498, 2017.
- [9] F. Calogero. Solution of the one-dimensional -body problems with quadratic and/or inversely quadratic pair potentials. J. Mathematical Phys., 12:419–436, 1971.
- [10] F. Calogero and C. Marchioro. Exact solution of a one-dimensional three-body scattering problem with two-body and/or three-body inverse-square potentials. J. Mathematical Phys., 15:1425–1430, 1974.
- [11] C. Collot, T. Duyckaerts, C. Kenig, and F. Merle. Soliton Resolution for the Radial Quadratic Wave Equation in Space Dimension 6. Vietnam J. Math., 52(3):735–773, 2024.
- [12] R. Côte. On the soliton resolution for equivariant wave maps to the sphere. Comm. Pure Appl. Math., 68(11):1946–2004, 2015.
- [13] R. Côte, Y. Martel, and X. Yuan. Long-time asymptotics of the one-dimensional damped nonlinear Klein-Gordon equation. Arch. Ration. Mech. Anal., 239(3):1837–1874, 2021.
- [14] R. P. de Moura and D. Pilod. Local well posedness for the nonlocal nonlinear Schrödinger equation below the energy space. Adv. Differential Equations, 15(9-10):925–952, 2010.
- [15] M. del Pino, M. Musso, and J. Wei. Existence and stability of infinite time bubble towers in the energy critical heat equation. Anal. PDE, 14(5):1557–1598, 2021.
- [16] B. Dodson. A determination of the blowup solutions to the focusing, quintic NLS with mass equal to the mass of the soliton. Anal. PDE, 17(5):1693–1760, 2024.
- [17] T. Duyckaerts, H. Jia, C. Kenig, and F. Merle. Soliton resolution along a sequence of times for the focusing energy critical wave equation. Geom. Funct. Anal., 27(4):798–862, 2017.
- [18] T. Duyckaerts, C. Kenig, Y. Martel, and F. Merle. Soliton resolution for critical co-rotational wave maps and radial cubic wave equation. Comm. Math. Phys., 391(2):779–871, 2022.
- [19] T. Duyckaerts, C. Kenig, and F. Merle. Classification of radial solutions of the focusing, energy-critical wave equation. Camb. J. Math., 1(1):75–144, 2013.
- [20] T. Duyckaerts, C. Kenig, and F. Merle. Soliton resolution for the radial critical wave equation in all odd space dimensions. Acta Math., 230(1):1–92, 2023.
- [21] W. Eckhaus. The long-time behaviour for perturbed wave-equations and related problems. In Trends in applications of pure mathematics to mechanics (Bad Honnef, 1985), volume 249 of Lecture Notes in Phys., pages 168–194. Springer, Berlin, 1986.
- [22] W. Eckhaus and P. Schuur. The emergence of solitons of the Korteweg-de Vries equation from arbitrary initial conditions. Math. Methods Appl. Sci., 5(1):97–116, 1983.
- [23] E. Fermi, P. Pasta, and S. Ulam. Los Alamos scientific laboratory report LA-1940. Collected Papers of Enrico Fermi, 2, 1955.
- [24] P. Gérard. An explicit formula for the Benjamin-Ono equation. Tunis. J. Math., 5(3):593–603, 2023.
- [25] P. Gérard and S. Grellier. An explicit formula for the cubic Szegő equation. Trans. Amer. Math. Soc., 367(4):2979–2995, 2015.
- [26] P. Gérard and E. Lenzmann. The Calogero-Moser Derivative Nonlinear Schrödinger equation. preprint arXiv:2208.04105, to appear in Comm. Pure Appl. Math., 2023.
- [27] P. Gérard and A. Pushnitski. The Cubic Szegő Equation on the Real Line: Explicit Formula and Well-Posedness on the Hardy Class. Comm. Math. Phys., 405(7):Paper No. 167, 2024.
- [28] J. Hogan and M. Kowalski. Turbulent Threshold for Continuum Calogero-Moser models. preprint arXiv:2401.16609, 2024.
- [29] K. Ishizuka. Long-time asymptotics of the damped nonlinear Klein-Gordon equation with a delta potential. preprint arXiv:2402.14381, 2024.
- [30] J. Jendrej. Construction of two-bubble solutions for the energy-critical NLS. Anal. PDE, 10(8):1923–1959, 2017.
- [31] J. Jendrej. Construction of two-bubble solutions for energy-critical wave equations. Amer. J. Math., 141(1):55–118, 2019.
- [32] J. Jendrej and A. Lawrie. Two-bubble dynamics for threshold solutions to the wave maps equation. Invent. Math., 213(3):1249–1325, 2018.
- [33] J. Jendrej and A. Lawrie. Soliton resolution for energy-critical wave maps in the equivariant case. preprint arXiv:2106.10738, to appear in J. Amer. Math. Soc., 2022.
- [34] J. Jendrej and A. Lawrie. Bubble decomposition for the harmonic map heat flow in the equivariant case. Calc. Var. Partial Differential Equations, 62(9):Paper No. 264, 36, 2023.
- [35] J. Jendrej and A. Lawrie. Soliton resolution for the energy-critical nonlinear wave equation in the radial case. Ann. PDE, 9(2):Paper No. 18, 117, 2023.
- [36] J. Jendrej, A. Lawrie, and W. Schlag. Continuous in time bubble decomposition for the harmonic map heat flow. preprint arXiv:2304.05927, 2023.
- [37] R. Jenkins, J. Liu, P. Perry, and C. Sulem. Soliton resolution for the derivative nonlinear Schrödinger equation. Comm. Math. Phys., 363(3):1003–1049, 2018.
- [38] R. Jenkins, J. Liu, P. Perry, and C. Sulem. The derivative nonlinear Schrödinger equation: global well-posedness and soliton resolution. Quart. Appl. Math., 78(1):33–73, 2020.
- [39] H. Jia and C. Kenig. Asymptotic decomposition for semilinear wave and equivariant wave map equations. Amer. J. Math., 139(6):1521–1603, 2017.
- [40] R. Killip, T. Laurens, and M. Vişan. Scaling-critical well-posedness for continuum Calogero-Moser models. preprint arXiv:2311.12334, 2023.
- [41] K. Kim, T. Kim, and S. Kwon. Construction of smooth chiral finite-time blow-up solutions to Calogero–Moser derivative nonlinear Schrödinger equation. preprint arXiv:2404.09603, 2024.
- [42] K. Kim, S. Kwon, and S.-J. Oh. Soliton resolution for equivariant self-dual Chern-Simons-Schrödinger equation in weighted Sobolev class. preprint arXiv:2202.07314, to appear in Amer. J. Math., 2022.
- [43] F. Merle. Determination of blow-up solutions with minimal mass for nonlinear Schrödinger equations with critical power. Duke Math. J., 69(2):427–454, 1993.
- [44] F. Merle and P. Raphael. Profiles and quantization of the blow up mass for critical nonlinear Schrödinger equation. Comm. Math. Phys., 253(3):675–704, 2005.
- [45] J. Moser. Three integrable Hamiltonian systems connected with isospectral deformations. Advances in Math., 16:197–220, 1975.
- [46] V. J. Novokšenov. Asymptotic behavior as of the solution of the Cauchy problem for a nonlinear Schrödinger equation. Dokl. Akad. Nauk SSSR, 251(4):799–802, 1980.
- [47] M. A. Olshanetsky and A. M. Perelomov. Completely integrable Hamiltonian systems connected with semisimple Lie algebras. Invent. Math., 37(2):93–108, 1976.
- [48] D. E. Pelinovsky and R. H. J. Grimshaw. A spectral transform for the intermediate nonlinear Schrödinger equation. J. Math. Phys., 36(8):4203–4219, 1995.
- [49] P. Raphael. Stability of the log-log bound for blow up solutions to the critical non linear Schrödinger equation. Math. Ann., 331(3):577–609, 2005.
- [50] P. C. Schuur. Asymptotic analysis of soliton problems, volume 1232 of Lecture Notes in Mathematics. Springer-Verlag, Berlin, 1986. An inverse scattering approach.
- [51] H. Segur. Asymptotic solutions and conservation laws for the nonlinear Schrödinger equation. II. J. Mathematical Phys., 17(5):714–716, 1976.
- [52] H. Segur and M. J. Ablowitz. Asymptotic solutions and conservation laws for the nonlinear Schrödinger equation. I. J. Mathematical Phys., 17(5):710–713, 1976.
- [53] B. Sutherland. Exact results for a quantum many-body problem in one dimension. Phys. Rev. A, 4:2019–2021, Nov 1971.
- [54] B. Sutherland. Exact results for a quantum many-body problem in one dimension. ii. Phys. Rev. A, 5:1372–1376, Mar 1972.
- [55] P. Topping. An example of a nontrivial bubble tree in the harmonic map heat flow. In Harmonic morphisms, harmonic maps, and related topics (Brest, 1997), volume 413 of Chapman & Hall/CRC Res. Notes Math., pages 185–191. Chapman & Hall/CRC, Boca Raton, FL, 2000.
- [56] R. van der Hout. On the nonexistence of finite time bubble trees in symmetric harmonic map heat flows from the disk to the 2-sphere. J. Differential Equations, 192(1):188–201, 2003.
- [57] N. J. Zabusky and M. D. Kruskal. Interaction of ”solitons” in a collisionless plasma and the recurrence of initial states. Phys. Rev. Lett., 15:240–243, Aug 1965.
- [58] V. E. Zakharov and A. B. Shabat. Exact theory of two-dimensional self-focusing and one-dimensional self-modulation of waves in nonlinear media. Soviet Physics JETP, 35(1):62–69, 1972.