Standing wave solutions of a quasilinear Schrödinger equation in the small frequency limit
Abstract.
This article is concerned with the quasilinear Schrödinger equation
where , and or and . After proving uniqueness and non-degeneracy of the positive solution for all , our main results establish the asymptotic behavior of in the limit . Three different regimes arise, termed ‘subcritical’, ‘critical’ and ‘supercritical’, corresponding respectively (when ) to , and . In each case a limit equation is exhibited which governs, in a suitable scaling, the behavior of in the limit . The critical case is the most challenging, technically speaking. In this case, the limit equation is the famous Lane-Emden-Fowler equation. A substantial part of our efforts is dedicated to the study of the function . We find that, for small , is increasing if and decreasing if . In the supercritical case, the monotonicity of depends on the dimension, except in the regime , where is always decreasing close to . The crucial role played by for the orbital stability of the standing wave , and for the uniqueness of normalized ground states, is discussed in the introduction.
1. Introduction
This work is concerned with the quasilinear Schrödinger equation
(1.1) |
and the associated stationary equation satisfied by standing waves ,
() |
We consider the problem in with , and .
This equation belongs to a family of quasilinear Schrödinger equations of the form
(1.2) |
which appear in several physical situations (see [9] and references therein). Here and are given smooth functions. When and , we obtain () which is relevant in various problems in plasma physics and nonlinear optics (see [43, 30, 20]). The local and global well-posedness of the Cauchy problem associated to (1.2) have been studied by Poppenberg in [42] for smooth initial data (belonging to the space ). In [12, 13], the authors improved the local well-posedness result for initial data in for with the integer part of . More precisely, in [12], equation (1.2) is solved locally for smooth nonlinearities and such that there exists a positive constant with for any , while [13] deals with the case and .
The parameter is a coupling constant relevant to describe the strength of the quasilinear interaction in the associated physical models. In several works about (), it is simply set equal to one. In the series of papers [1, 5, 4, 2], the asymptotic behavior of solutions of () is studied in the limit . It is worth noting here that a simple scaling transforms () with into
Some of the results obtained in [1, 5, 4, 2] can thus be recovered from the present approach, which we believe to be much more straightforward. Furthermore, as explained below, our main focus in this work is the asymptotic behavior of the function defined in (1.11), which is not addressed in detail in [1, 5, 4, 2].
Concerning the existence of solutions to (), the following result has been proved by Colin, Jeanjean and Squassina [13].
Theorem 1.1.
Remark 1.
The main idea used to prove Theorem 1.1 is to remark that finding a solution of the equation () is equivalent to finding a solution of the semilinear equation
(1.3) |
with a suitably chosen function (see Section 2 for more details). Variational arguments are then used to deduce an existence result for (1.3). Finally, a solution to () is obtained by setting .
Note that that for , is a necessary condition for the existence of nontrivial solutions . This is a consequence of the following integral identities, which will play an important role in our analysis. They can be proved as sketched in [13]. Let
(1.4) |
and
(1.5) |
Proposition 1.1.
As already noted above, it follows from Proposition 1.1 that, in dimensions , is a necessary condition for the existence of nontrivial solutions .
The main goal of this paper is to study the qualitative properties of a branch of solutions of (), parametrized by . In particular, we want to investigate the uniqueness of positive solutions and their non-degeneracy, i.e. the fact that the kernel of the linearized operators is trivial, modulo phase and space translations.
The study of uniqueness of positive solutions to semilinear Schrödinger equations of the form
with a well-behaved nonlinearity, has a very long history, see e.g. [11, 29, 10, 24, 37, 38, 41, 44, 48, 32]. Quasilinear equations have attracted less attention, and fewer results on the uniqueness or non-degeneracy of positive solutions of equations of this type are available, see e.g. [47, 19, 3, 31].
A first important result that will be proved here is the uniqueness and non-degeneracy of the positive solution to (), for any .
Theorem 1.2 (Uniqueness and non-degeneracy).
As a consequence of the non-degeneracy of , the following proposition will also be established.
The non-degeneracy property is also crucial for the study of the mass
(1.11) |
of the unique positive solution . An important motivation for studying this quantity is the central role played by the function in the Grillakis-Shatah-Strauss theory of stability [52, 21, 22, 14, 15], which can be applied to the standing wave solutions of the time-dependent equation (1.1). In particular, the standing wave is expected to be orbitally stable when and unstable when . Therefore, it is important to be able to identify the regions where is increasing, which correspond to stable solutions, and those where is decreasing, corresponding to unstable solutions. For the classical nonlinear Schrödinger equation (NLS) with a single-power nonlinearity, which corresponds to () with , and , or and , the mass is an explicit function of . Indeed, for , the solution for can be obtained from the unique positive solution to by the simple scaling . This leads to
However, when , the presence of the quasilinear term prevents one from using a scaling argument and any hope of obtaining a simple expression for vanishes. A similar situation occurs in the case of the double-power nonlinearity considered in [32]. Those two problems seem rather different at first sight, but they share a common feature, in the form of an extra term living at another spatial scale.
Hence, in the same spirit as the works of Lewin, Rota Nodari [32] and Moroz, Muratov [39], our main theorem regarding () establishes the behavior of and as .
Theorem 1.3.
(i) Suppose and , or and . Then, as , the rescaled function
(1.12) |
converges in to the ground state of the stationary nonlinear Schrödinger equation
(1.13) |
Furthermore, as ,
(1.14) |
In a neighborhood of , is increasing if and decreasing if .
(ii) Suppose and . Then there exists a function such that, as , and the rescaled function
(1.15) |
converges in to the function
(1.16) |
which is the (up to dilations) unique positive radial-decreasing solution of the Lane-Emden-Fowler equation
(1.17) |
The scaling function can be chosen so that, as ,
(1.18) |
Furthermore,
(1.19) |
In a neighborhood of , is decreasing.
(iii) Suppose and . Then, as , converges in to the unique positive radial-decreasing solution of the equation
(1.20) |
Furthermore, as .
If , we have and in
as .
Finally,
(1.21) |
and
(1.22) |
In a neighborhood of , is decreasing in dimension . In dimensions , is decreasing for .
Remark 2.
In the subcritical case (i), the classical NLS scaling (1.12) kills the quasilinear term in the limit , yielding the single-power NLS (1.13) as limit equation. As in the case of the NLS with a double-power nonlinearity [32], it seems natural to use an implicit function argument to recover the branch of solutions starting from the ground state of (1.13). In the case , the presence of the quasilinear term in () requires a clever choice of functional framework.
In the supercritical case (iii), the limit equation is formally obtained by letting in (), which yields (1.20). However, (1.20) has no nontrivial solutions if , as can be seen by combining (1.7) and (1.6) (see (4.41)). In the critical case (ii), revealing the asymptotic behavior of as is more delicate. Our approach follows that laid down in [39], based on the concentration-compactness principle. It shows that, in the rescaled variable (1.15), only two terms survive as , thus yielding the limit equation (1.17). As in [39], both in the critical and the supercritical cases, we will take advantage of the variational characterization of solutions of the auxiliary semilinear problem (1.3) to deduce the desired convergence of . The main difficulty is thus to deal with the function whose explicit expression is not known.
Remark 3.
In the supercritical case, as for the double-power nonlinearity [32], although we are able to prove that admits a finite limit when in dimension , we cannot determine its sign in the full range of parameters. However, if
then for small enough, see Proposition 4.6 below. This condition is probably not optimal but it allows us to conclude that is a decreasing function in a neighborhood of whenever , for any .
As a consequence of Theorem 1.3, for any and ( if ), the positive solution , for small enough, lies on an unstable branch of solutions, which is in agreement with the result of instability by blow-up obtained by Colin, Jeanjean, Squassina [13, Theorem 1.5]. 111For simplicity, we speak of the stability of when actually referring to the orbital stability of the associated standing wave of (1.1).
In [13], it has been conjectured that whenever , the positive solution to () is orbitally stable. On the one hand, our analysis for close to confirms this conjecture for since is an increasing function in a neighborhood of in this case. On the other hand, when , the function is decreasing in a neighborhood of , at least in dimension . This implies that , the positive solution of (), should be unstable at least for small enough in the latter case.
The original conjecture from Colin, Jeanjean and Squassina was supported by the results obtained for the so-called normalized solutions, i.e. solutions of () with a prescribed -norm. In particular, for a given , one can look for solutions of
(1.23) |
where is the energy functional defined by
(1.24) |
with . To each minimizer of (1.23) corresponds a Lagrange multiplier such that (after an appropriate space translation), where is the unique positive solution to ().
The minimization problem (1.23) has been studied extensively over the last decade (see [13, 25, 26, 54]). The known results can be summarized as follows:
-
(i)
For all , if and if .
-
(ii)
If , then for all , and admits a minimizer.
-
(iii)
If , there exists such that for all and is negative and strictly decreasing on . Moreover, if then admits a minimizer if and only if . If , then admits a minimizer if and only if .
-
(iv)
If , then either or . As a consequence, has no minimizers for all .
-
(v)
The set of minimizers of (1.23), when it is not empty, is orbitally stable.
Unfortunately, the orbital stability of the set of minimizers does not allow one to deduce the orbital stability of a single solution of () for a fixed . On the one hand, for fixed , one should first determine whether a solution of () is also a solution of the minimization problem (1.23) with . Thanks to the non-degeneracy of and the spectral properties of , we know from [52, App. E] and [27, Theorem 5.3.2] (see also [35, 36]) that, is a strict local minimum of at fixed mass , if , whereas the solution is not a local minimum when . As a consequence, a solution lying on a decreasing branch of the function cannot be a solution of (1.23), so its stability cannot be deduced from the orbital stability of the set of minimizers. On the other hand, the set of minimizers may contain more than one solution. More precisely, the uniqueness of positive solutions to () at fixed does not imply the uniqueness of energy minimizers. As already mentioned, any minimizer, when it exists, is positive and solves () for some Lagrange multiplier . The difficulty here is that the Lagrange multiplier is a priori not uniquely determined : for a given mass , there can be several minimizers that share the same energy but give rise to different Lagrange multipliers . In other words, there may not be a one-to-one mapping . Nevertheless, any candidate to be a Lagrange multiplier in this situation must be a solution to the equation , hence the importance of studying the behavior of the function and its variations. In particular, if one is able to find a region of ’s where the function is one-to-one, then the uniqueness of energy minimizers follows for such ’s.
To conclude, we mention the following result, that is a corollary of Theorem 1.3.
Corollary 1.1.
Thanks to this result, we can conclude that if , then the solution , for close to , cannot be a global minimizer of since its energy is strictly positive. Thus, we cannot deduce its stability from the orbital stability of the set of minimizers. In fact, our conjecture is that , for small, is unstable for any (at least in dimensions ).
All the previous remarks emphasize the importance of studying the function and, in particular, the number of sign changes of . Nevertheless, getting a complete picture on the behavior of seems out of reach for the moment. As in [32], a first step will be to understand the exact behavior of at the two endpoints of its interval of definition : the current paper was devoted to the limit , while the limit will be discussed in a follow-up paper currently in preparation.
Organization of the paper
Section 2 is devoted to basic properties of the change of variables and of the auxiliary semilinear equation (1.3). In Section 3, the uniqueness and non-degeneracy stated in Theorem 1.2 are proved, as well as Proposition 1.2. Finally, the proof of Theorem 1.3 is given in Section 4, where it is split into three subsections, dealing with cases (i), (ii) and (iii), respectively.
2. Reformulation of the problem: from a quasilinear to a semilinear equation
To prove Theorem 1.2, and Theorem 1.3 (ii) and (iii), we will use a change of variables borrowed from [34], given by
Since for all , it follows that has an inverse function , which is on and satisfies the first order Cauchy problem
(2.1) |
We now extend to the whole real line by letting when .
Lemma 2.1.
The odd function has the following properties:
and is strictly increasing on , with for all ;
for all and is a Lipschitz function on ;
for all ,
Proof.
(i) and (ii) follow immediately from (2.1). To prove (iii), integrating (2.1) yields
(whence the formula for if you rather decided to define by (2.1)). Since, by construction, as and as , the limits in (iii) then easily follow from the above relation. To prove (iv), we consider the function defined by for . Clearly, and
As a consequence, we get
which implies (iv). ∎
Remark 4.
For , we now define ,
(2.2) |
and we consider the nonlinear elliptic equation on
() |
Note that the nonlinearity is of the form with
(2.3) |
As a consequence, for any ,
(2.4) |
Furthermore, it follows from Lemma 2.1 that there exists a constant such that
(2.5) |
For all , we attack problem () via a standard variational approach. Let222The normalization factor in (2.6) is introduced for coherence with the classical minimization problem for the critical Sobolev inequality which will be used in Section 4.3.
(2.6) |
where, using (2.1),
(2.7) |
It follows by classical results of Berestycki-Lions [8, Theorem 2] and Berestycki-Gallouët-Kavian [7] that, for all , the minimization problem (2.6) has a solution which is spherically symmetric and radially nonincreasing. Furthermore, there is a corresponding Lagrange multiplier such that
(2.8) |
Then, by elliptic regularity, , with exponentially as . Hence, for all . Furthermore, the following classical integral identities (Nehari and Pohozaev) provide a relation between and :
(2.9) |
Indeed, if satisfies and , it follows that
Once a solution of (2.8) is obtained, one gets a solution of () by a simple dilation,
3. Uniqueness and non-degeneracy
Equipped with the change of variables in Lemma 2.2, we can now prove uniqueness and non-degeneracy of . Our proof is a consequence of results from McLeod [37] and Lewin, Rota Nodari [32] (see also Adachi et al. [3]).
Proof of Theorem 1.2.
It is straightforward to check that satisfies the hypotheses of [17, Theorem 4.1]. As a consequence, if is a positive solution to (), then it is radial decreasing with respect to some point in . Next a direct computation shows that satisfies the hypotheses of [37, Theorem 2] when for or for , and of [32, Theorem 1] when (see also [3]). Thus, we deduce that the solution is the unique positive radial solution of (), modulo translations. Furthermore, it is non-degenerate:
(3.1) |
Since satisfies the hypotheses of [18, Theorem 2], all positive classical solutions of () that tend to zero at infinity are radial decreasing about some point in . As a consequence, the solution is the unique positive solution of (), modulo translations.
Thanks to the monotonicity of , we can conclude that is the unique positive solution of (), modulo translations.
To prove the non-degeneracy of , we have to compute the linearized operator of () around . A straightforward computation gives with and defined by (1.9) and (1.10) respectively. Note the first eigenvalue of the operators and , when it exists, is necessarily simple with a positive eigenvalue. This can be proved for instance by using that and Harnack’s inequality.
Since is a solution of (), it is clear that . Moreover, . As a consequence, is the first eigenfunction of and .
Since , the multiplier is bounded away from and we deduce that if and only if . Hence, if and only if . As a consequence, the non-degeneracy of implies that if and only if . As a conclusion, using that for all , we deduce that
This concludes the proof of Theorem 1.2. ∎
The next proposition will be useful to prove parts (ii) and (iii) of Theorem 1.3.
Proposition 3.1.
The linearized operator has exactly one negative eigenvalue.
Proof.
We shall use again the relation
with . In this proof we will also denote by ′ differentiation with respect to . We shall use the same notation for and its radial counterpart, i.e. , with , and similarly for other spherically symmetric functions. On the one hand, by taking , we see that
As a consequence, has at least one negative eigenvalue. Let be the first eigenvalue of .
Note that, for any eigenvalue of , there exists such that while the corresponding solves
that is . Since the operator commutes with space rotations, it may be written as a direct sum with
For any and , in the sense of quadratic forms. As a consequence, for all and . Hence, to the first eigenvalue of , there corresponds , , such that
Moreover, using that , we can easily show that . A classical bootstrap argument shows that for any . This, thanks to Sobolev inequalities, implies . Finally, using the ODE satisfied by and the fact that , we can then show that .
Next, suppose, by contradiction, that has a second eigenvalue . As above, there exists such that
Furthermore, from the proof of [32, Lemma 7.3], we know that the unique solution to
vanishes exactly once and is such that .
Since , using Sturm’s comparison theorem, we deduce that vanishes at least once in . Let such that . Since , we can apply twice Sturm’s comparison theorem and deduce that has at least one zero in and at least one zero in . This contradicts the fact that vanishes exactly once and concludes the proof. ∎
We conclude this section with the proof of Proposition 1.2, which is a consequence of the implicit function theorem.
Proof of Proposition 1.2.
Let such that and define as
Using the fact that , we deduce that for any and any . Then we can easily show that is well-defined and continuously Fréchet differentiable.
For any , let be the unique positive solution to (). Then and, since is non-degenerate, is one-to-one. Hence, it remains to prove that is an isomorphism from onto . Since is in the Schwartz space , the operator can be seen as a compact perturbation of which is itself an isomorphism from onto .
Applying the implicit function theorem, we conclude that there exists a map defined in a neighborhood of of solutions to ().
To conclude that , it suffices to prove that is positive, which can be done via a spectral theory argument. For each , consider the operator
where
Note that and has a zero eigenvalue with eigenfunction . Moreover, when , and is an isolated simple eigenvalue at the bottom of the spectrum. Hence, for close to , the zero eigenvalue of must also be at the bottom of the spectrum. Hence, by applying [46, Theorem XIII.46], we deduce that must be positive for any close to . ∎
4. Proof of Theorem 1.3
We shall decompose the proof of Theorem 1.3 into three parts: first (i), then (iii), and finally (ii), which is more involved.
4.1. Subcritical case
In the subcritical case, the behavior of as can be determined by a straightforward application of the implicit function theorem.
Proof of Theorem 1.3 (i).
When or and , the rescaled function defined by (1.12) solves
(4.1) |
and it converges to the unique positive solution to the NLS equation (1.13). More precisely, the implicit function theorem gives
(4.2) |
where and . Using (4.2), we obtain
(4.3) |
To compute the derivative of , we note that and, thanks to the non-degeneracy of , we can write
Using the change of variables defined in (1.12), we obtain
with
Since converges to in as , we deduce that converges to in the norm resolvent sense (see [45, Theorem VIII.25]). Since , we obtain the convergence
(4.4) |
in norm. More precisely, using (4.2), we have
where is understood as a multiplication operator. By iterating the resolvent identity, we obtain
Note that decays at the same rate as so that tends to at infinity even if . As a consequence,
(4.5) |
With a scaling argument, we can easily compute
for the NLS case, so that
This leads to
and gives the first-order term in whenever .
To compute the next-order term in and we proceed as follows. Let . We know that
with
On the one hand,
On the other hand,
so that
(4.6) |
Formulas (4.1) and (4.6) prove (1.3). (Note that, alternatively, (4.6) can be obtained by using directly the explicit expression of .)
Similarly, we consider and we compute
Hence,
with . As a consequence, it is enough to evaluate at to conclude. In particular, this gives
For , we obtain
The monotonicity properties of the map follow from these formulas. This concludes the proof of part (i) of Theorem 1.3. ∎
4.2. Supercritical case
To prove part (iii) of Theorem 1.3, we shall take advantage of another variational characterization of solutions of (). We define the functional by
where
It follows from [8] that . We next observe that is invariant under dilations: if and
Proposition 4.1.
Proof.
Let be a minimizer for (2.6):
It follows that and . Hence, . Suppose by contradiction that . One can then find such that . Letting , the dilation satisfies , so that
This contradiction shows that, indeed, .
Finally, if is such that , again the dilation factor yields , and so is a minimizer for (2.6). ∎
We shall now consider the so-called ‘zero-mass case’, . We let
(4.7) |
where and are still defined by (2.2) and (2.7), respectively. The existence of a spherically symmetric and radially nonincreasing minimizer of (4.7) follows from [8]. It is also known (see [6, 16]) that decays like as . Furthermore, as in the case , there exists a Lagrange multiplier such that
(4.8) |
The integral identities (2.9) still hold for , and it follows that
Again, the dilation
produces a solution of
() |
and the change of variables yields a solution of (1.20). Note that, as in the case , the above changes of variables give a one-to-one correspondence between and . Furthermore, and have the same regularity and decay at infinity. The main difference here compared to the case is that, in dimensions , , due to its slow decay as . Only in dimensions does .
Proposition 4.2.
We also have uniqueness, up to translations, and non-degeneracy of the minimizer , as a consequence of the following theorem.
Theorem 4.1.
Suppose . Then the following properties hold true.
(4.8) has a unique positive radial solution in .
non-degenerate:
where
(4.9) |
The proof of the uniqueness is a direct application of [51, Theorem 2] while the non-degeneracy in can be proved as in [32, Lemma A.1].
We can now state and prove the first main result of this section.
Proposition 4.3.
Suppose . Consider a sequence , such that . Then
(4.10) |
and
(4.11) |
We will prove Proposition 4.3 using several lemmas. The first one is technical.
Lemma 4.1.
Let . Consider a function , a number and a cut-off function such that on and on . Suppose there exists a constant such that
Then we have the following asymptotics as :
(4.12) |
Proof.
Let be a continuous function such that for all . For , we have
If , choosing yields
If , we choose and we get
This gives the desired lower bounds. The upper bounds are straightforward. ∎
Lemma 4.2.
As , there holds
Proof.
Firstly, for and ,
Hence, using the minimizer as a test function for , we have
Now, if , we have that . Furthermore, since , it follows by continuity that, for small enough, . Therefore, we can use as a test function for , which yields
By Remark 4, and we have
which concludes the proof in case .
For , we let and we introduce a cut-off function such that for , for , for , and for all . We shall simply write for the function . We have and we now use it as a test function for . As above, we have
First note that as by dominated convergence. Next,
Now, for any , there exists such that
Since decays like as , so does . Hence,
This leads to
(4.13) |
where we observe that since . As consequence, we have
Furthermore, by Lemma 4.1,
(4.14) |
For , by (4.13) and (4.14) we have
We now conclude the proof in the following way.
If , we let and we have
If , we let and we have
which completes the proof. ∎
Lemma 4.3.
Consider a sequence , such that . For all , let be a minimizer for (2.6) with . Then
(4.15) |
Proof.
The next two lemmas provide classical results that are crucial in our analysis, and which will be proved in the Appendix for completeness.
Lemma 4.4 (Radial Lemma).
Let and be a radial nonincreasing function. Then,
(4.17) |
where .
Let be a sequence of radial nonincreasing functions, such that
Then there exists such that, up to a subsequence:
(4.18) |
Lemma 4.5.
Let , . Let be the minimizer obtained above, which solves the Euler-Lagrange equation
(4.19) |
There exists and such that
(4.20) |
There exists such that for all there exists a constant such that
(4.21) |
Note that the result in (ii) is trivial for since . For , the proof follows a Moser-type iteration argument, which will be given in the Appendix.
We are now in a position to prove Proposition 4.3.
Proof of Proposition 4.3.
Suppose and let be defined as in Lemma 4.3. By (4.15), is bounded in . Hence, there exists and a subsequence of (still denoted by ) such that
Furthermore, by Lemma 4.2 and Lemma 4.5 (ii), for all there exists a constant such that
(4.22) |
Let . We will first show that, up to a further subsequence,
(4.23) |
Let . By the Radial Lemma, we already know that, up to a subsequence,
(4.24) |
We now fix . Since is bounded in , there exists such that, up to a subsequence, weakly in and a.e. in . But then a.e. and we conclude that . As a consequence, by (4.17) and (4.22), there is a constant such that
Hence, for . Since and a.e. in , it follows by dominated convergence that
(4.25) |
Next, using (4.23) with , we have
so that
Hence, using the constraint for all , it follows by Lemma 4.3 that
(4.26) |
We deduce that and satisfies the constraint in (4.7).
Furthermore, by weak lower semi-continuity of on ,
Hence, is a minimizer for (4.7). It follows that is a positive solution of (4.8) with , and so by Theorem 4.1.
We now prove that in . We have
On the one hand,
On the other, using (4.8) we find
By Hölder’s inequality, we have
Therefore,
We conclude that
Next, by Lemma 4.5 (i), the sequence is bounded in . It then follows by standard elliptic theory arguments (see e.g. the proof of Proposition 4.14 below) that in .
To complete the proof, we now turn to . Observing that
we deduce from the conclusions obtained for the sequence that, up to a subsequence,
Finally, since the only possible limit point is , a proof by contradiction shows that, in fact, the whole sequence converges to . ∎
Going back to the variable , we obtain the following.
Proposition 4.4.
Suppose . Consider a sequence , such that . Then
(4.27) |
and
(4.28) |
Proof.
First, (4.28) is a direct reformulation of (4.11). Next, in (4.27), the -convergence of follows directly from the -convergence of in (4.10) and the fact that is Lipschitz. The -convergence follows from the -convergence of by dominated convergence. Finally, -convergence follows from the -convergence of , using the -bound on (Lemma 4.5 (i)) and the fact that , and are Lipschitz on compact subsets of . ∎
4.2.1. Asymptotic behavior of as
Proposition 4.5.
Let and . As , we have
More precisely,
(4.29) |
for all .
Moreover, letting for any , there holds
(4.30) |
in for all .
Remark 5.
The strong convergence of , and holds in for all . This includes only in dimensions .
Proof.
Let us start with the case . First, we show that . Indeed, suppose by contradiction there is a sequence such that is bounded, where . Then using Proposition 4.3 is bounded in and (using Rellich-Kondrachov and Radial Lemma) up to a subsequence, in , hence , a contradiction. As a consequence, as .
Next, by Lemma 4.5 (i), there exist such that for all . Then, by Lemma 2.1, for all ,
and
This implies, as .
Since a.e. and is bounded in , dominated convergence yields as , for any and . Furthermore, a maximum principle argument as in [32] allows one to improve the classical bound
to
with independent of . This provides an upper bound in , for , if is chosen small enough, and it follows by dominated convergence that as .
Next, by Lemma 2.1,
Finally, let and ,
The first term of the right-hand side of this inequality is simply and goes to zero as . Since a.e. and , the second term also goes to zero by dominated convergence. We conclude that as . ∎
4.2.2. Asymptotic behavior of as .
Proposition 4.6.
Let and . Then, for small enough, we have
(4.31) |
where
Moreover,
-
(i)
if , then
-
(ii)
if , then
for all small enough;
-
(iii)
if and
(4.32) then
for all small enough.
Remark 6.
Proof.
For any , let be defined by (1.9). We know from Proposition 3.1 that, for any , has exactly one negative eigenvalue. Then we define the symmetric matrix given by the restriction of to the finite dimensional space spanned by .
Since
straightforward computations give
Next, for any , let . As a consequence of Proposition 1.1, we can write
with , and deduce that
(4.33) |
As a consequence,
and
Moreover,
(4.34) |
In particular, using , this leads to
(4.35) |
since .
Now, a tedious but straightforward computation gives
and, using again , we obtain from (4.35) that
Since has a unique negative eigenvalue, we deduce that the determinant of is negative:
This gives the estimate (4.6).
Using again (4.35), we have
with
The function is a second order polynomial and its maximum on is reached at . In particular,
for . This, together with Proposition 4.5, concludes the proof for .
For , and vanishes twice in at
Note that and , so that for or . As a consequence, for such ’s, whenever is small enough. ∎
Proposition 4.6 gives a complete description of the asymptotic behavior of for . We conclude this section with the study of for . We know from above that in , being the unique positive radial-decreasing solution to (1.20). More precisely, for any , and in with and the unique positive radial-decreasing solution to
To discuss the asymptotic behavior of as , we proceed as follows. On the one hand,
with . Now, since is a solution to (), we have
This implies
with .
On the other hand, from Theorem 4.1, we know that the limiting linearized operator
has a trivial kernel so that can be defined using functional calculus. Moreover, since converges to in as , we deduce that converges to in the norm resolvent sense (see [45, Theorem VIII.25]). More precisely, we have the following lemma.
Lemma 4.6.
Let and . For any , define . Then, as ,
(4.36) |
for all if , for all if .
Proof.
First of all, using and the properties of , we have
so that
and
Next, for any , can be written as
A direct computation gives, for any ,
while
As a consequence, for any ,
(4.37) |
where we used that is uniformly bounded for small enough. This, together with the -convergence of , implies
Finally,
Next, using Proposition 4.5 and (4.37), we deduce that
for any if or if . Hence, to prove that in , it is enough to show that
is uniformly bounded. This is the case for any . Indeed, using again that is uniformly bounded for small enough and that there exist independent of such that
for all (see proof of Proposition 4.5), we have
for any . Since, in any case, , this completes the proof. ∎
Since is the limit, in the norm resolvent sense, of and has exactly one negative eigenvalue (see Proposition 3.1), we deduce that has one negative eigenvalue and is otherwise positive and unbounded from above. As a consequence, we expect that, in the limit , behaves like .
As in [32, Lemma 4.2], the first step to give a proper interpretation of the quadratic form
is to prove that the quadratic form domain of is the same as for the free Laplacian if . More precisely, we have the following lemma.
Lemma 4.7.
Let and . There exists a constant such that
The proof of this lemma is exactly the same as in [32, Lemma 4.2] and we do not reproduce it here. The key ingredients are the regularity of the potential and the fact has exactly one negative eigenvalue. In particular, since and , we have that behaves like at infinity. As a consequence, for any which is the regularity used in the proof of the lemma. Using that has exactly one negative eigenvalue, we then deduce that the operator is invertible (within the sector of radial functions) and that can be written as
Next, we give the upper bound on .
Proposition 4.7.
Let and . Then
(4.38) |
in the sense of quadratic forms and where . In dimensions the right side equals whereas it is finite for .
As above, the proof of this lemma is exactly the same as in [32, Lemma 4.3]. Once again we use the convergence of in the norm resolvent sense and the strong convergence of in for . The fact that the right side of (4.38) is infinite in dimension and finite for depends on the behavior of . In particular, it depends on the behavior at infinity of . On the one hand, we have
On the other hand, using Lemma 2.1, we have
As a consequence, using the decay of at infinity, we deduce that, for any ,
as in [32, Lemma 4.3].
Proposition 4.8.
Let and . Then
(4.39) |
The proof of this lemma is similar to that of [32, Lemma 4.4]. The main difference is linked to the particular form of which contains a term of the form .
Here the quadratic form should be interpreted as with .
Proof.
We start by noticing that can be written as
(4.40) |
where
Thanks to Lemma 4.6, in . This, together with the Hardy-Littlewood-Sobolev inequality, implies that
in operator norm. Since the operator is compact and converges strongly to the identity, we deduce that in operator norm. This implies that the spectrum of converges to the spectrum of and, since is invertible, we deduce that is bounded and converges to in operator norm. As a consequence, we can invert (4.40) and write
and
To conclude, it remains to prove that converges to in . From the HLS inequality, we have
which is finite for . More precisely, if , Proposition 4.5 gives the strong convergence of towards in as goes to . As a consequence,
which tends to zeros. Hence, we can pass to the limit and obtain
as expected. ∎
To sum up, we finally gather the main results for the supercritical case.
4.3. Critical case
We now consider the case , i.e. . Combining the two identities from Proposition 1.1 with , we find that
(4.41) |
for any solution . Hence, for , the limit equation (1.20) has no nontrivial solution in . In fact, we will see that, in the critical case, a rescaled version of the solution converges, as , to the unique (up to scaling) radial positive solution of the critical Lane-Emden-Fowler equation
On the formal level, considering a scaling parameter , if is a solution of (), it follows that
satisfies
Thus, choosing such that yields the limit equation (1.17).
In order to establish rigorously the convergence of in the limit we shall rather work with , as in Subsection 4.2. We will compare, in the limit , the minimization problem and its minimizers with the following one:
where
(4.42) |
is invariant under the scaling
(4.43) |
It is well-known [50] that is attained on the family of positive radial functions
where is the Aubin-Talenti function defined in (1.16). The following properties of the family of minimizers are noteworthy:
(4.44) |
(4.45) |
Proposition 4.9.
Suppose . There exists a scaling function with the following properties. For any sequence such that , letting
and
(4.46) |
we have
(4.47) |
The proof of Proposition 4.9 will use several lemmas. The following bounds are direct consequences of Lemma 2.1 and (4.45).
Lemma 4.8.
We next establish the lemma corresponding to Lemma 4.2 in the supercritical case.
Lemma 4.9.
As , there holds
(4.50) |
Proof.
First, since satisfies
and
we obtain
(4.51) |
Now, if , we have and we can use the family as test functions for . By Lemma 4.8, we see that, if and ,
i.e. . Choosing , it follows from Lemma 4.8 that
We deduce that
which concludes the proof in case .
For , we take and use the same cut-off function as in the proof of Lemma 4.2. Then but can be used as a test function for . We have
First, by dominated convergence, as .
Next, similarly to (4.48), we obtain
Furthermore, by (4.45),
Using Lemma 4.1 and proceeding as in (4.49), we have
Now
We now conclude the proof in the following way. Let with to be chosen later.
If , we let , so that
and
provided that . This implies, by taking ,
If , we let , so that
and
This implies, by taking ,
The proof is complete.
∎
Remark 7.
We can now determine the behavior of as .
Lemma 4.10.
Consider a sequence , such that . For all , let be a minimizer for (2.6) with . Then, as ,
(4.52) |
and
(4.53) |
Proof.
We can now give the
Proof of Proposition 4.9.
4.3.1. Further convergence properties of the solutions
We first prove the upper bounds in (1.18).
Proposition 4.10.
The scaling function can be chosen so that, as ,
(4.54) |
Proof.
Consider . By the proof of Lemma 4.9, we have
(4.55) |
The upper bounds (4.54) follow directly from (4.55) and the following estimates:
(4.56) |
(4.57) |
To prove (4.56), we first observe that
(4.58) |
By Lemma 4.5 (i), there exists and such that
Hence,
and it follows by Lemma 2.1 that
Thus,
and (4.58) yields
The proof of (4.56) will be complete if we show that
(4.59) |
To this end, we use as a test function for :
Since
we deduce that
It follows by a first order Taylor expansion that
We next turn to the lower bounds in (1.18), which are more involved.
Proposition 4.11.
The scaling function can be chosen so that, as ,
(4.60) |
Remark 8.
For , the upper bounds (4.54) and the lower bounds (4.60) come with the same power of . For , they do not match. This is due to the rough estimate (4.57) used to derive the upper bounds from (4.56) and also, in case , to the non-optimal decay of that was already pointed out in Remark 7. For , estimate (4.57) will be improved in Lemma 4.16 and the optimal lower bounds on will be given in Proposition 4.12 and Proposition 4.13, respectively.
The proof of Proposition 4.11 will use the following lemmas.
Lemma 4.11.
Proof.
In this proof we use the Nehari and Pohozaev integral identities derived from the equation
(4.61) |
which respectively read
and
We deduce from these identities that
(4.62) |
Suppose by contradiction there exist and a sequence such that and , as . For the remainder of the proof, we shall again abbreviate the notation as , , and .
First suppose that . From Lemma 4.5 (i), we know there exists such that for all . It follows that
whence pointwise as , which contradicts .
Lemma 4.12.
As , there holds
Proof.
Lemma 4.13.
As , there holds
(4.66) |
Proof.
Lemma 4.14.
Consider , and let . Then there exists a constant such that
(4.68) |
Proof.
Apply the classical Gagliardo-Nirenberg inequality (see [40]) to . ∎
We are now in a position to prove Proposition 4.11.
Proof of Proposition 4.11.
Extending (4.68) to by density and applying it to with exponents and , we obtain
(4.69) |
Furthermore, the Hölder inequality yields a constant such that
Since in , it follows that
Therefore, (4.69) shows that is bounded away from zero. Hence, by (4.56) and (4.66),
(4.70) |
The conclusion now follows from (4.55). ∎
As was pointed out in Remark 8, for all the upper and lower bounds on have the same blow-up rate. We shall now tighten the upper bounds on for .
Lemma 4.15.
For all , there exist and such that
(4.71) |
Proof.
As a consequence of (4.71), using the same arguments as in [39, Lemma 4.9, Lemma 4.11], we obtain the following.
Lemma 4.16.
As , there holds
(4.72) |
This is enough to conclude in the case .
Proposition 4.12.
Let . As , there holds
(4.73) |
Proof.
The case is slightly more involved. Indeed, to obtain a matching lower and upper bound for we need to improve the upper bound for .
Proposition 4.13.
Let . As , there holds
(4.74) |
Remark 9.
Proof.
From the proof of Lemma 4.9, we know that, for ,
provided that , and . Hence, we let and with to be chosen later. We have
and
This implies, by taking ,
so that
Arguing as in the proof of Proposition 4.11, we get the lower bound
On the other hand, estimates (4.59), (4.56) and (4.72) imply that
which leads to
as claimed. ∎
Lemma 4.17.
For all ,
(4.75) |
Proof.
As a consequence of Lemma 4.17, the following -bound can be proved similarly to Lemma 4.5 (i), using (4.61) instead of (5.7).
Lemma 4.18.
There exists and such that
(4.76) |
Thanks to this estimate, we can now establish -convergence of .
Proposition 4.14.
Consider such that as . Then
Proof.
Fix an arbitrary . We will first prove that
(4.77) |
This convergence follows by interpolation between
(4.78) |
and
(4.79) |
To prove (4.78), we split into
On the one hand, thanks to Lemma 4.18, by dominated convergence. On the other, by Lemma 4.4 (ii).
To prove (4.79), let and write
Using the elliptic equations (4.61) and (4.44) satisfied by and , respectively, we have
Since as by (4.54), it is easy to show that
(4.80) |
It follows by dominated convergence and Lemma 4.18 that , for any fixed .
As for , we now show that it can be made as small as desired by choosing large enough. On the one hand, since , we deduce that
(4.81) |
On the other hand, since
we have that
Furthermore, since is bounded in , it follows by Lemma 4.4 (i) that there exists a constant such that
Thus, using again as , we find that
(4.82) |
Since , it follows from (4.81) and (4.3.1) that, given any , we can choose so large that
Since , we deduce that in . We now bootstrap this to in by using the ODE’s satisfied by the radial functions and :
(4.83) |
(4.84) |
where ′ denotes differentiation with respect to . We use the same notation for , and their radial counterparts, i.e. , , with .
We first note that, using , the pointwise convergence in (4.80) can be improved to
(4.85) |
Subtracting (4.83) from (4.84), multiplying by and integrating yields
With the change of variables , this identity becomes
and it follows that
Hence, by (4.85),
(4.86) |
It then follows from (4.83), (4.84) and (4.86) that
Since we already know that in , this completes the proof. ∎
4.3.2. Asymptotic behavior of and as .
As a consequence of Lemma 4.12 and the exact asymptotic behavior of , we deduce the following proposition which gives the behavior of as goes to .
Proposition 4.15.
Let and . As , we have
More precisely,
(4.87) |
Proof.
To deduce the asymptotic behavior of , we proceed as in Proposition 4.6.
Proposition 4.16.
Let and . Then, for small enough, we have
(4.88) |
where
Moreover,
Proof.
As in the proof of Proposition 4.6, for any , let be defined by (1.9) and be the symmetric matrix given by the restriction of to the finite dimensional space spanned by .
When , the same arguments detailed above give
with .
In the critical case, the integral identities of Proposition 1.1 give
(4.89) |
As a consequence,
and
for all .
Since has a unique negative eigenvalue, we deduce that the determinant of is negative:
This gives the estimate (4.88).
In the limit , goes which implies . As a consequence,
and, for small enough,
for some positive constant . This implies,
∎
5. Appendix
Proof of Lemma 4.4.
For the whole proof, we denote by etc., the functions on defined by , etc., for . Any radial has a representative such that is continuous on , and is related to by
Part (i) follows from the estimate
To prove (ii), we fix an arbitrary . We first address the convergence in . Since is bounded, there exists such that, up to a subsequence, converges to , weakly in and a.e. on , as . Applying (4.17) with and the Sobolev embedding theorem, there is a constant (independent of ) such that
Hence, for large,
Dominated convergence then implies in .
To prove convergence in , we let and we show that
We first apply (4.17) again with , using the fact that is bounded. Given , there exist and a constant such that,
Hence, we only need to show that as , uniformly for . By an Arzela-Ascoli type argument which will be made explicit below, this is a consequence of the equicontinuity of the sequence on , which we prove now.
Let . Since is bounded in , there exists such that,
(5.1) | ||||
(5.2) | ||||
(5.3) |
Hence, is equicontinuous on .
We now claim that as for all . We will prove below that the convergence is uniform. To complete the proof, the claim can be proved by similar arguments, using convergence almost everywhere and equicontinuity of on .
Suppose by contradiction that does not converge uniformly to on : there exists such that, for all there exists such that
There exists a subsequence and a point such that as . Suppose without loss of generality that . By equicontinuity, there exists such that
Now choose such that for all . It follows that,
This contradicts the pointwise convergence and finishes the proof. ∎
Proof of Lemma 4.5.
We start by proving part (ii), which will be used in the proof of part (i). To prove estimate (4.21), we follow the scheme of proof laid down in [1, Lemma 5.5]. Let
Since , there holds
Multiplying both sides of (4.22) by and integrating by parts, one has
(5.4) |
Since , it follows by Lemma 4.2/4.9 and the Sobolev embedding theorem that there exists and such that
Hence, using (2.5) and Lemma 4.2/4.9, there is a constant such that, for all ,
(5.5) |
On the other hand,
(5.6) |
Let
A direct calculation shows that
By the Sobolev embedding, there is a constant such that
Therefore,
By (5.6), there exists a constant such that
The proof now follows by iteration. Let
For each , multiplying both sides of (4.22) by yields
instead of . Next, the above argument shows that
and we conclude that there exists a constant such that
Finally, it is an easy exercise to show that as . The conclusion thus follows by Hölder interpolation.
We next prove part (i). First recall that, for any , is positive, radial decreasing, and satisfies the elliptic equation
(5.7) |
where
By Lemma 4.9, there exist and such that
Furthermore, by part (ii) proved above, for any there exists a constant such that
(5.8) |
Hence, by the Radial Lemma, taking larger if needed,
Thus,
uniformly for . Choosing , it is possible to find , so that and hence is bounded, for . Then, by Theorem 5.1 and Remark 5.1 in [49], there is a constant , independent of , such that
(5.9) |
Since converges in as , it follows that is bounded, for . Finally, it follows by part (ii) of Lemma 4.4 that is also bounded, which completes the proof. ∎
References
- [1] S. Adachi, M. Shibata, and T. Watanabe, Blow-up phenomena and asymptotic profiles of ground states of quasilinear elliptic equations with -supercritical nonlinearities, J. Differential Equations, 256 (2014), pp. 1492–1514.
- [2] , Blow-up phenomena and asymptotic profiles of ground states of quasilinear elliptic equations with -supercritical nonlinearities, J. Differential Equations, 256 (2014), pp. 1492–1514.
- [3] S. Adachi, M. Shibata, and T. Watanabe, A note on the uniqueness and the non-degeneracy of positive radial solutions for semilinear elliptic problems and its application, Acta Math. Sci. Ser. B (Engl. Ed.), 38 (2018), pp. 1121–1142.
- [4] S. Adachi, M. Shibata, and T. Watanabe, Asymptotic property of ground states for a class of quasilinear Schrödinger equation with -critical growth, Calc. Var. Partial Differential Equations, 58 (2019), pp. Paper No. 88, 29.
- [5] S. Adachi and T. Watanabe, Asymptotic properties of ground states of quasilinear Schrödinger equations with -subcritical exponent, Adv. Nonlinear Stud., 12 (2012), pp. 255–279.
- [6] , Asymptotic uniqueness of ground states for a class of quasilinear Schrödinger equations with h1-supercritical exponent, J. Differential Equations, 260 (2016), pp. 3086–3118.
- [7] H. Berestycki, T. Gallouet, and O. Kavian, Équations de champs scalaires euclidiens non linéaires dans le plan, C. R. Acad. Sci. Paris Ser I Math., 297 (1983), pp. 307–310.
- [8] H. Berestycki and P.-L. Lions, Nonlinear scalar field equations. I. Existence of a ground state, Arch. Rational Mech. Anal., 82 (1983), pp. 313–345.
- [9] L. Brüll, H. Lange, and E. de Jager, Stationary, oscillatory and solitary wave type solution of singular nonlinear Schrödinger equations, Mathematical Methods in the Applied Sciences, 8 (1986), pp. 559–575.
- [10] C. C. Chen and C. S. Lin, Uniqueness of the ground state solutions of in , Comm. Partial Differential Equations, 16 (1991), pp. 1549–1572.
- [11] C. V. Coffman, Uniqueness of the ground state solution for and a variational characterization of other solutions, Arch. Rational Mech. Anal., 46 (1972), pp. 81–95.
- [12] M. Colin, On the local well posedness of quasilinear Schrödinger equations in arbitrary space dimension, Comm. Partial Differential Equations, 27 (2002), pp. 325–354.
- [13] M. Colin, L. Jeanjean, and M. Squassina, Stability and instability results for standing waves of quasi-linear Schrödinger equations, Nonlinearity, 23 (2010), p. 1353.
- [14] S. De Bièvre, F. Genoud, and S. Rota Nodari, Orbital stability: analysis meets geometry, in Nonlinear optical and atomic systems. At the interface of physics and mathematics, Cham: Springer; Lille: Centre européen pour les mathématiques, la physique et leurs interactions (CEMPI), 2015, pp. 147–273.
- [15] S. De Bièvre and S. Rota Nodari, Orbital stability via the energy-momentum method: the case of higher dimensional symmetry groups, Arch. Rational Mech. Anal., 231 (2019), pp. 233–284.
- [16] M. Flucher and S. Müller, Radial symmetry and decay rate of variational ground states in the zero mass case, SIAM J. Math. Anal., 29 (1998), pp. 712–719.
- [17] R. L. Frank, Ground states of semi-linear PDE. Lecture notes from the “Summerschool on Current Topics in Mathematical Physics”, CIRM Marseille, Sept. 2013.
- [18] B. Gidas, W. M. Ni, and L. Nirenberg, Symmetry of positive solutions of nonlinear elliptic equations in , Adv. in Math. Suppl. Stud., 7 (1981), pp. 369–402.
- [19] F. Gladiali and M. Squassina, Uniqueness of ground states for a class of quasi-linear elliptic equations, Adv. Nonlinear Anal., 1 (2012), pp. 159–179.
- [20] M. V. Goldman, Strong turbulence of plasma waves, Rev. Mod. Phys., 56 (1984), pp. 709–735.
- [21] M. Grillakis, J. Shatah, and W. Strauss, Stability theory of solitary waves in the presence of symmetry. I, J. Funct. Anal., 74 (1987), pp. 160–197.
- [22] , Stability theory of solitary waves in the presence of symmetry. II, J. Funct. Anal., 94 (1990), pp. 308–348.
- [23] R. D. Hajer Bahouri, Jean-Yves Chemin, Fourier Analysis and Nonlinear Partial Differential Equations, Grundlehren der mathematischen Wissenschaften, Springer Berlin, Heidelberg, 2011.
- [24] J. Jang, Uniqueness of positive radial solutions of in , , Nonlinear Anal., 73 (2010), pp. 2189–2198.
- [25] L. Jeanjean and T. Luo, Sharp nonexistence results of prescribed -norm solutions for some class of Schrödinger–Poisson and quasi-linear equations, Z. Angew. Math., 64 (2013), pp. 937–954.
- [26] L. Jeanjean, T. Luo, and Z.-Q. Wang, Multiple normalized solutions for quasi-linear Schrödinger equations, J. Differential Equations, 259 (2015), pp. 3894–3928.
- [27] T. Kapitula and K. Promislow, Orbital Stability of Waves in Hamiltonian Systems, Springer New York, New York, NY, 2013, pp. 117–157.
- [28] R. Killip, T. Oh, O. Pocovnicu, and M. Vişan, Solitons and scattering for the cubic–quintic nonlinear Schrödinger equation on , Arch. Rational Mech. Anal., 225 (2017), pp. 469–548.
- [29] M. K. Kwong, Uniqueness of positive solutions of in , Arch. Rational Mech. Anal., 105 (1989), pp. 243–266.
- [30] E. W. Laedke, K. H. Spatschek, and L. Stenflo, Evolution theorem for a class of perturbed envelope soliton solutions, J. Math. Phys., 24 (1983), pp. 2764–2769.
- [31] M. Lewin and S. Rota Nodari, Uniqueness and non-degeneracy for a nuclear nonlinear Schrödinger equation, NoDEA Nonlinear Differential Equations Appl., 22 (2015), pp. 673–698.
- [32] M. Lewin and S. Rota Nodari, The double-power nonlinear Schrödinger equation and its generalizations: uniqueness, non-degeneracy and applications, Calc. Var. Partial Differential Equations, 59 (2020), pp. Paper No. 197, 49.
- [33] P.-L. Lions, The concentration-compactness principle in the calculus of variations. The limit case. I, Rev. Mat. Iberoamericana, 1 (1985), pp. 145–201.
- [34] J.-Q. Liu, Y.-Q. Wang, and Z.-Q. Wang, Soliton solutions for quasilinear Schrödinger equations, ii, J. Differential Equations, 187 (2003), pp. 473–493.
- [35] J. H. Maddocks, Restricted quadratic forms and their application to bifurcation and stability in constrained variational principles, SIAM J. Math. Anal., 16 (1985), pp. 47–68.
- [36] , Errata: Restricted quadratic forms and their application to bifurcation and stability in constrained variational principles, SIAM J. Math. Anal., 19 (1988), pp. 1256–1257.
- [37] K. McLeod, Uniqueness of positive radial solutions of in , II, Trans. Amer. Math. Soc., 339 (1993), pp. 495–505.
- [38] K. McLeod and J. Serrin, Uniqueness of positive radial solutions of in , Arch. Rational Mech. Anal., 99 (1987), pp. 115–145.
- [39] V. Moroz and C. B. Muratov, Asymptotic properties of ground states of scalar field equations with a vanishing parameter, J. Eur. Math. Soc. (JEMS), 16 (2014).
- [40] L. Nirenberg, On elliptic partial differential equations, Ann. Scuola Norm. Sup. Pisa Cl. Sci., Ser. 3, 13 (1959), pp. 115–162.
- [41] L. A. Peletier and J. Serrin, Uniqueness of positive solutions of semilinear equations in , Arch. Rational Mech. Anal., 81 (1983), pp. 181–197.
- [42] M. Poppenberg, On the local well posedness of quasilinear Schrödinger equations in arbitrary space dimension, J. Differential Equations, 172 (2001), pp. 83–115.
- [43] M. Porkolab and M. V. Goldman, Upper‐hybrid solitons and oscillating‐two‐stream instabilities, The Physics of Fluids, 19 (1976), pp. 872–881.
- [44] P. Pucci and J. Serrin, Uniqueness of ground states for quasilinear elliptic operators, Indiana Univ. Math. J., 47 (1998), pp. 501–528.
- [45] M. Reed and B. Simon, Methods of Modern Mathematical Physics. I. Functional analysis, Academic Press, 1972.
- [46] , Methods of Modern Mathematical Physics. IV. Functional analysis, Academic Press, 1972.
- [47] A. Selvitella, Uniqueness and nondegeneracy of the ground state for a quasilinear Schrödinger equation with a small parameter, Nonlinear Anal., 74 (2011), pp. 1731–1737.
- [48] J. Serrin and M. Tang, Uniqueness of ground states for quasilinear elliptic equations, Indiana Univ. Math. J., 49 (2000), pp. 897–923.
- [49] G. Stampacchia, Le problème de Dirichlet pour les équations elliptiques du second ordre à coefficients discontinus, Ann. Inst. Fourier (Grenoble), 15 (1965), pp. 189–258.
- [50] G. Talenti, Best constant in Sobolev inequality, Ann. Mat. Pura Appl. (4), 110 (1976), pp. 353–372.
- [51] M. Tang, Uniqueness and global structure of positive radial solutions for quasilinear elliptic equations, Comm. Partial Differential Equations, 26 (2001), pp. 909–938.
- [52] M. I. Weinstein, Modulational stability of ground states of nonlinear Schrödinger equations, SIAM J. Math. Anal., 16 (1985), pp. 472–491.
- [53] M. Willem, Minimax theorems, vol. 24 of Progress in Nonlinear Differential Equations and their Applications, Birkhäuser Boston, Inc., Boston, MA, 1996.
- [54] H. Ye and Y. Yu, The existence of normalized solutions for -critical quasilinear Schrödinger equations, J. Math. Anal. Appl., 497 (2021), p. 124839.