Research article Special Issues

Stability of the standing waves of the concentrated NLSE in dimension two

  • In this paper we will continue the analysis of two dimensional Schr?dinger equation with a fixed, pointwise, nonlinearity started in [2, 13]. In this model, the occurrence of a blow-up phenomenon has two peculiar features: the energy threshold under which all solutions blow up is strictly negative and coincides with the infimum of the energy of the standing waves; there is no critical power nonlinearity, i.e., for every power there exist blow-up solutions. Here we study the stability properties of stationary states to verify whether the anomalies mentioned before have any counterpart on the stability features.

    Citation: Riccardo Adami, Raffaele Carlone, Michele Correggi, Lorenzo Tentarelli. Stability of the standing waves of the concentrated NLSE in dimension two[J]. Mathematics in Engineering, 2021, 3(2): 1-15. doi: 10.3934/mine.2021011

    Related Papers:

    [1] Delyan Zhelyazov . Numerical spectral analysis of standing waves in quantum hydrodynamics with viscosity. Mathematics in Engineering, 2024, 6(3): 407-424. doi: 10.3934/mine.2024017
    [2] Evangelos Latos, Takashi Suzuki . Quasilinear reaction diffusion systems with mass dissipation. Mathematics in Engineering, 2022, 4(5): 1-13. doi: 10.3934/mine.2022042
    [3] Emilio N. M. Cirillo, Giuseppe Saccomandi, Giulio Sciarra . Compact structures as true non-linear phenomena. Mathematics in Engineering, 2019, 1(3): 434-446. doi: 10.3934/mine.2019.3.434
    [4] Michael Herrmann, Karsten Matthies . Solitary waves in atomic chains and peridynamical media. Mathematics in Engineering, 2019, 1(2): 281-308. doi: 10.3934/mine.2019.2.281
    [5] Giovanni S. Alberti, Yves Capdeboscq, Yannick Privat . On the randomised stability constant for inverse problems. Mathematics in Engineering, 2020, 2(2): 264-286. doi: 10.3934/mine.2020013
    [6] Plamen Stefanov . Conditionally stable unique continuation and applications to thermoacoustic tomography. Mathematics in Engineering, 2019, 1(4): 789-799. doi: 10.3934/mine.2019.4.789
    [7] George Contopoulos . A Review of the “Third” integral. Mathematics in Engineering, 2020, 2(3): 472-511. doi: 10.3934/mine.2020022
    [8] Michele Dolce, Ricardo Grande . On the convergence rates of discrete solutions to the Wave Kinetic Equation. Mathematics in Engineering, 2024, 6(4): 536-558. doi: 10.3934/mine.2024022
    [9] Dheeraj Varma, Manikandan Mathur, Thierry Dauxois . Instabilities in internal gravity waves. Mathematics in Engineering, 2023, 5(1): 1-34. doi: 10.3934/mine.2023016
    [10] Lorenzo Pistone, Sergio Chibbaro, Miguel D. Bustamante, Yuri V. Lvov, Miguel Onorato . Universal route to thermalization in weakly-nonlinear one-dimensional chains. Mathematics in Engineering, 2019, 1(4): 672-698. doi: 10.3934/mine.2019.4.672
  • In this paper we will continue the analysis of two dimensional Schr?dinger equation with a fixed, pointwise, nonlinearity started in [2, 13]. In this model, the occurrence of a blow-up phenomenon has two peculiar features: the energy threshold under which all solutions blow up is strictly negative and coincides with the infimum of the energy of the standing waves; there is no critical power nonlinearity, i.e., for every power there exist blow-up solutions. Here we study the stability properties of stationary states to verify whether the anomalies mentioned before have any counterpart on the stability features.


    The Nonlinear Schrödinger Equation (NLSE) with concentrated nonlinearity in d=2 is the subject of several recent papers, finalizing a research program developed over the last twenty years (see [3,4,8,14] for the NLSE with concentrated nonlinearity and also [15] and [12] for the fractional case and the Dirac equation, respectively).

    Such a research line was originally motivated by some mesoscopic physical models. For instance, in semiconductor theory the effect of electronic charge accumulation in a resonant tunneling in a double barrier heterostructure [20] is typically studied using a concentrated NLSE. More recently, other applications have been suggested: the spontaneous formation of quantum coherent non-dissipative patterns in semiconductor heterostructures with nonlinear properties [11]; the dynamics of the mixed states of statistical physics [23]; the appearance of quantum turbulence in the probability density [9]; the scattering in nuclear physics models for the disexcitation of isomeric states and also the production of weakly bounded states in heavy nuclei close to the instability; the analysis of resonant tunneling diodes, which exhibits intrinsic instability [31] or the fabrication of semiconductor superlattices, for the estimate of the time decay rates for the solutions to the Schrödinger-Poisson equations in the repulsive case [10,25].

    In [13] and [16] the local well-posedness is established, i.e., the problem of existence and uniqueness of the solution for short times, as well as the mass and energy conservation. Global existence is also proven in the defocusing case irrespective of the power of the nonlinearity. In [2] it is studied the occurrence of a blow-up phenomenon for a focusing nonlinearity, with two peculiar features: first, the energy threshold under which all solutions blow up is strictly negative and coincides with the infimum of the energy of standing waves; second, there is no critical power nonlinearity, i.e., for every power there exist blow up solutions. We remark that such a behavior is anomalous compared to the conventional NLSE, also because such anomalies are not a direct consequence of the dimension, or of the concentrated nonlinearity. In fact, there is a critical power for standard nonlinearities in dimension two [21], and there is also a critical power for concentrated nonlinearities in dimension one and three [3,8]. In the present paper we investigate further whether such peculiarities also show up in the stability of stationary states.

    Let us preliminarily recall the results on the standard NLSE [29]: Consider the Cauchy problem for a focusing NLSE, where the word focusing refers to the attractive character of the nonlinearity, with initial data in the energy space

    ıtψ(t,x)+ψ+|ψ|2σψ=0,ψ(0,x)=ψ0(x)H1(Rd).

    In [17], using a variational characterization, it was established the orbital stability of the ground-states in the subcritical case, i.e., for σ<2d. On the other hand, [19,26,30] presents an alternative proof of orbital stability for the subcritical solitary waves and shows the orbital instability in the critical and supercritical case (σd2).

    It turns out that there is a strict relation between blow-up and orbital stability of standing waves [28]. The NLSE admits blow-up solutions if and only if its solitary waves are orbitally unstable. This behavior has some relevant exceptions as in the case of NLSE in bounded domains or in [27] where the key feature of all this models is always the absence of translational invariance in space.

    The analysis of stationary states stability for concentrated nonlinearities traces back to [5,6,7]. For the concentrated NLSE in dimension 2 the scenario is different and, in some sense, surprising. As it will be illustrated in Section 2 there are, at any fixed value of the mass, two branches of stationary states, distinguished by the value of the frequency ω, with opposite orbital stability behavior. To the best of our knowledge, there is no similar behavior for a standard Schrödinger equation on Rd, but some analogy exists with the 1d NLSE in the presence of a point interaction [18] and with NLSE on compact domains [22,24]. In all these cases, the nonlinearity is supercritical.

    As for the case of concentrated NLSE in the defocusing case, the scenario is really puzzling. In Section 3 the analysis of stationary waves reveals that they are stable and moreover that they are ground states.

    The problem under investigation can be formally written as

    {ıψt=(Δ+β|ψ|2σδ0)ψ,in R+×R2,ψ(0)=ψ0,in R2, (1.1)

    where σ>0, βR and δ0 is a Dirac delta function centered at the origin of R2.

    As extensively explained in [2,13], the Cauchy problem in (1.1) can be rigorously formulated in a weak form. To this aim, one has to first introduce the so-called energy space

    V:={χL2(R2):χ=χλ+qGλ,χλH1(R2),qC}, (1.2)

    with λ>0 and Gλ denoting the Green's function of Δ+λ in R2, i.e.,

    Gλ(x):=K0(λx)2π=12πF1[(|k|2+λ)1](x), (1.3)

    (recall that K0 is the Macdonald function of order zero given, e.g., in [1] and that F is the unitary Fourier transform of R2). Note that the parameter λ does not affect the definition of V. Indeed, it is possible to rewrite the space V without the parameter λ, as

    V={χL2(R2),χ=χ0qlog|x|2π,χ0˙H1(R2),qC} (1.4)

    where ˙H1(R2) is the homogeneous Sobolev space. However, as (1.4) is not easily implemented in the expression of the energy, we shall keep using (1.2) throughout.

    Hence, we can define a weak solution of (1.1) as a function ψ such that

    ψ(t)=ϕλ(t)+q(t)GλV,t0, (1.5)

    and such that, for every χ=χλ+qχGλV,

    {ıddtχ,ψ(t)=χλ,ψ(t)+λ(χλ,ϕλ(t)χ,ψ(t))+θλ(|q(t)|)qχq(t),t>0,ψ(0)=ψ0, (1.6)

    where , is the usual scalar product in L2(R2), and θλ:R+R is defined as

    θλ(s):=log(λ/2)+γ2π+βs2σ,

    with γ the Euler-Mascheroni constant (note that the parameter λ does not affect (1.6) too). According to (1.5) we will usually refer to ϕλ(t) as the regular part of ψ(t), to q(t)Gλ as the singular part and to q(t) as the charge.

    It has been proven in [2,13] that, for σ1/2, (1.6) is locally well-posed in V (with the additional assumption ϕλ(0)H1+η, η>0) and that the mass

    M(t)=M(ψ(t)):=ψ(t)2,

    denoting the usual norm in L2(R2), and the energy

    E(t)=E(ψ(t)):=ϕλ(t)2+λ(ϕλ(t)2ψ(t)2)+(β|q(t)|2σσ+1+log(λ/2)+γ2π)|q(t)|2, (1.7)

    which is independent of λ as well, are preserved along the flow. In addition, when β>0, i.e., in the defocusing case, the solution is global in time, whereas when β<0, i.e., in the focusing case, the solution blows up in a finite time. In order to prove these results, one has to require [2] that ϕλ(0) belongs to the Schwartz space, which is only a technical hypothesis, and, more important, its energy satisfies

    E(ψ0)<Λ=Λ(σ,β):=σ4π(σ+1)(4πσβ)1/σ.

    In the following sections we study the problem of the stability of stationary states separately in the focusing and defocusing case.

    In the focusing case, i.e., for β<0, (1.6) admits (see [2]) a unique family of standing waves of the form

    ψω(t,x):=eıωteıηuω(x),ηR,ω(˜ω,+),˜ω:=4e2γ, (2.1)

    where

    uω(x):=q(ω)Gω(x),q(ω):=(log(ω/2)+γ2πβ)1/2σ. (2.2)

    The behavior of q(ω) is depicted in Figure 1a.

    Figure 1.  Plots of q(ω) and E(ω) for ω(˜ω,+), when σ=1 and β=1. Here ˜ω1.26, ¯ω3.43, ¯q0.2 and Λ0.0016.

    Now, plugging (2.2) into (1.7), one finds that the energy of the standing waves as a function of the frequency ω reads

    E(ω):=E(uω)=(σlog(ω/2)+γσ2π(σ+1)14π)(log(ω/2)+γ2πβ)1/σ,ω(˜ω,+). (2.3)

    The behavior of E(ω) is represented in Figure 1b. In addition,

    minω(˜ω,+)E(ω)=E(¯ω)=Λ,where¯ω:=4e2γ+1/σ. (2.4)

    On the other hand, noting that q(˜ω)=0 and that q() is smooth and strictly increasing on (˜ω,+), one can take the inverse q(ω) of the function

    ω(q):=4e2γ4πβq2σ,q>0, (2.5)

    and plug it into (2.3), to obtain the energy as a function of q, i.e.,

    E(q)=q24πσβq2σ+2σ+1. (2.6)

    The behaviors of ω(q) and E(q) are depicted in Figure 2a and b, respectively. This alternative form can be useful in computation since (2.6) is more manageable than (2.3). Furthermore,

    infq>0E(q)=E(¯q)=Λ<0,where¯q:=q(¯ω)=(4πσβ)1/2σ.
    Figure 2.  Plots of ω(q) and E(q) for qR+, when σ=1 and β=1.

    The natural question arising at this point is about the stability of the standing waves. In view of the application of Grillakis-Shatah-Strauss theory [19], it is first necessary to compute the mass M as a function of ω and q. Exploiting (1.3), one has that

    M(ω):=M(uω)=q2(ω)4πω=14πω(log(ω/2)+γ2πβ)1/σ. (2.7)

    On the other hand, one can easily check that

    M(ω)=q2(ω)4πω2[(log(ω/2)+γ)12σ1]=:h(ω),

    whence

    M(ω)>0(resp. M(ω)<0)h(ω)>0(resp. h(ω)<0)˜ω<ω<¯ω(resp. ω>¯ω). (2.8)

    In addition, as limω˜ωM(ω)=limω+M(ω)=0, there results

    supω(˜ω,+)M(ω)=M(¯ω)=e2γ1/σ16π(4πσβ)1/σ=:¯μ.

    As a consequence, for every value of the mass μ(0,¯μ) (or, alternatively, of the energy E(Λ,0)) there exists two distinct families of standing waves uω1,uω2, such that M(ω1)=M(ω2)=μ, with ω1(˜ω,¯ω) and ω2(¯ω,+).

    Analogous results can be obtained writing the mass of the standing waves in terms of q in place of ω, so that

    M(q)=q2e2γ+4πβq2σ16π (2.9)

    and

    supq>0M(q)=M(¯q)=¯μ.

    The qualitative behavior of M(ω) and M(q) is depicted in Figure 3.

    Figure 3.  Plots of M(ω) and M(q) when σ=1 and β=1.

    For any μ>0 there is no ground state of mass μ, i.e., no global minimizer of the energy constrained on

    Vμ:={ψV:M(ψ)=μ}.

    This can be easily seen if one defines a sequence {un}nN such that

    un(x):=2πμnG1(nx),M(un)=μ.

    Indeed, {un}Vμ and

    E(un)=nμ+(β(4πμn)σσ+1+log(n/2)+γ2π)(πμn)1/2n+,

    since β<0. Hence, the stability analysis requires the use of the techniques developed in [19], as shown in Theorem 2.1.

    First, we recall the definition of orbitally stable standing wave. To this aim, preliminarily, we endow the energy space V with a norm. Due to the several possible decompositions of a function χV for different values of the spectral parameter λ>0 (see (1.2)), in order to obtain a suitable norm, one has to fix a value λ=¯λ>0 and then set

    χ2¯λ:=χ¯λ2H1(R2)+|q|24π¯λ.

    Clearly, any other choice of λ gives rise to an equivalent norm. In this section we will set ¯λ=1 for the sake of simplicity.

    Definition 2.1. The standing wave uω is said to be orbitally stable whenever for every ε>0 there exists δ>0 such that: if ψ0eıηuω1<δ, for some ηR, and ψ(t) is a solution of (1.6) on [0,T) with initial condition ψ0, then ψ(t) can be continued to a solution on [0,+) and

    suptR+infηRψ(t)eıηuω1<ε.

    Otherwise the standing wave is called unstable.

    Theorem 2.1 (Stability in the focusing case). Let σ1/2 and β<0. The standing waves defined in (2.1) and (2.2) are orbitally stable if ω(˜ω,¯ω) and unstable if ω>¯ω (where ¯ω is given in (2.4)).

    Note that the previous theorem entails that, for every mass μ(0,¯μ), there is a pair of standing waves of mass μ, where the one with low frequency is stable, while the one with high frequency is unstable.

    Remark 2.1. The assumption σ1/2 is only related to the local well-posedness of (1.6) proved in [13]. It is likely that it could be dropped by means of a more refined analysis of the local well-posedness and hence is not actually relevant in the stability analysis.

    The proof of Theorem 2.1 is based on [19,Theorem 3]. For the sake of completeness, we recall here the statement suitably adapted to (1.6).

    Theorem 2.2. Assume that:

    (A1) there exists a local solution of (1.6), which preserves mass and energy along the flow;

    (A2) there exist ω2>ω1>0 and a family (uω)ω of standing waves of (1.6) such that the a mapping (ω1,ω2)ωuωV is of class C1;

    (A3) letting

    Sω:VR,Sω(u):=E(u)+ωM(u),

    be the action functional associated with (1.6) and defining the operator

    Hω:VV,Hω:=d2Sω(uω) (2.10)

    (d2Sω denoting the second Fréchet differential), suppose that, for every ω(ω1,ω2),

    (i) Hω has exactly one negative simple eigenvalue,

    (ii) the kernel of Hω coincides with the span of uω,

    (iii) the rest of σ(Hω) is positive and bounded away from zero.

    If the function

    D:(ω1,ω2)R,D(ω):=Sω(uω),

    is strictly convex, then uω is orbitally stable. If, on the contrary, D is strictly concave, then uω is unstable.

    Proof of Theorem 2.1. (A1) of Theorem 2.2 has been proven by [13,Theorem 1.1 & Theorem 1.2], while the fulfillment of (A2) is a direct consequence of the form of the standing waves given by (2.1) and (2.2), with ω1=˜ω and ω2=+. Concerning (A3) we argue as follows.

    As d2M(uω)=2×I, it is sufficient to compute only d2E(uω). Since E is a functional of class C2 we can compute the Gˆateaux second differential in place of the Fréchet second differential, i.e.,

    d2E(uω)[h,k]=2E(uω+νh+τk)ντ|ν=τ=0.

    In addition, for the sake of simplicity, we can set λ=ω in the definition of E. Therefore, standard computations yields

    E(uω+νh+τk)τ=2Re{νhω+τkω,kω+ω(νhω+τkω,kωuω+νh+τk,k)++qk(q(ω)+νqh+τqk)log(ω/2)+γ2π+βqk(q(ω)+νqh+τqk)σ(q(ω)+νqh+τqk)σ+1},

    so that

    2E(uω+νh+τk)ντ=2Re{hω,kω+ω(hω,kωh,k)++qkqhlog(ω/2)+γ2π+σβqkqh(q(ω)+νqh+τqk)σ1(q(ω)+νqh+τqk)σ+1++(σ+1)βqkqh(q(ω)+νqh+τqk)σ(q(ω)+νqh+τqk)σ}

    and hence

    2E(uω+νh+τk)ντ|ν=τ=0=2Re{hω,kω+ω(hω,kωh,k)}++log(ω/2)+γπRe{qkqh}+2βq2σ(ω)Re{σqkqh+(σ+1)qkqh}. (2.11)

    Now, if we split each quantity as real and imaginary part, i.e.,

    h=hr+ıhi,k=kr+ıki,hω=hrω+ıhiω,kω=krω+ıkiω,qh=qrh+ıqih,qk=qrk+ıqik,

    then (2.11) reads

    2E(uω+νh+τk)ντ|ν=τ=0=B1[hr,kr]+B1[hi,ki],

    where B1,B2 are two sesquilinear forms given by

    B1[hr,kr]:=2(hrω,krω+ω(hrω,krωhr,kr))+(log(ω/2)+γπ+2β(2σ+1)q2σ(ω))qrkqrh

    and

    B2[hi,ki]:=2(hiω,kiω+ω(hiω,kiωhi,ki))+(log(ω/2)+γπ+2βq2σ(ω))qikqih.

    Furthermore, one notes that B1,B2 are the sesquilinear form (restricted to real-valued functions) associated with the operators Hα1,Hα2:L2(R2)L2(R2) with domains

    dom(Hαj):={ψL2(R2):ψ=ϕλ+qGλ,ϕλH2(R2),qC,ϕλ(0)=(αi+log(λ/2)+γ2π)q},i=1,2, (2.12)

    with λ>0, and action

    (Hαi+λ)ψ:=(Δ+λ)ϕλ,ψdom(Hαi),i=1,2, (2.13)

    where

    α1=(2σ+1)βq2σ(ω),α2=βq2σ(ω). (2.14)

    Summing up,

    2E(uω+νh+τk)ντ|ν=τ=0=2(hr,hi)(Hα100Hα2)(krki),

    whence

    d2E(uω)=2(Hα100Hα2)

    and, consequently,

    Hω=2(Hα1+ω00Hα2+ω).

    In order to verify (ⅰ), (ⅱ) and (ⅲ) of (A3), it suffices to observe that

    σ(Hα1)={ωe8πβσq2σ}[0,+),σ(Hα2)={ω}[0,+),

    with ωe8πβσq2σ and ω simple eigenvalues, and that uω is the eigenfunction associated with ω. Indeed, this entails

    σ(Hω)={ω(1e8πβσq2σ)}{0}[ω,+), (2.15)

    which proves that Hω possesses one simple negative eigenvalue (since 1e8πβσq2σ<0, as β<0), that the kernel of Hω is the span of uω and that the rest of the spectrum is positive and bounded away from zero.

    Finally, it is sufficient to detect for which values of ω the scalar function D(ω) is strictly convex and for which values of ω it is strictly concave. However, as uω is a standing wave, dSω(uω)=0 (dSω denoting the Fréchet differential) so that D(ω)=M(ω). Therefore, recalling (2.8), one can conclude the proof.

    One can easily check that there exists a family of standing waves in the defocusing case β>0 as well:

    ψω(t,x):=eıωteıηuω(x),uω(x):=q(ω)Gω(x),q(ω):=(log(ω/2)+γ2πβ)1/2σ (3.1)

    with ηR, defined for

    ω(0,˜ω),where˜ω:=4e2γ.

    The behavior of q(ω) is shown in Figure 4a.

    Figure 4.  Plots of q(ω) and E(ω) for ω(0,˜ω), when σ=1 and β=1.

    In addition, simple computations show that the form of E(ω) is still given by (2.3), but in this case the function E(ω) is unbounded from below, due to the fact that β>0. The behavior of E(ω) is depicted in Figure 4b.

    From (3.1) one has that the function q(ω) is invertible. Again we get that ω(q) reads as (2.5) and, plugging (2.5) into (2.3), one obtains (2.6) for E(q). The behavior of ω(q) and E(q) is depicted in Figure 5a and b.

    Figure 5.  Plots of ω(q) and E(q) for qR+, when σ=1 and β=1.

    Remark 3.1. Let us point out a relevant difference between the focusing and the defocusing case: M(ω) and M(q), given by (2.7) and (2.9), respectively, are strictly monotone on their domain with range R+. In particular, this means that, in the defocusing case, for every μR+, there exists a unique (up to a phase factor) standing wave uωμ of mass μ.

    Concerning the stability of these standing waves, one can prove the following

    Theorem 3.1 (Stability in the defocusing case). Let σ1/2 and β>0. The standing waves defined by (3.1) are orbitally stable for every ω(0,˜ω).

    The proof of Theorem 3.1 is analogous to that of Theorem 2.1. The main difference is that the key tool now is [19,Theorem 1], instead of [19,Theorem 3]. For the sake of completeness, we recall also this statement (again, suitably adapted to (1.6)).

    Theorem 3.2. Assume that (A1) and (A2) of Theorem 2.2 are satisfied. If, in addition, the operator Hω, defined by (2.10), satisfies (ii) and (iii) of (A3) in Theorem 2.1, then uω is orbitally stable.

    Proof of Theorem 3.1. Arguing as in the proof of Theorem 2.1 one immediately sees that (A1) and (A2) are fulfilled, with ω1=0 and ω2=˜ω.

    In addition, following again the proof of Theorem 2.1, one obtains that

    Hω=2(Hα1+ω00Hα2+ω)

    with Hα1,Hα2 defined in (2.12) and (2.13) and α1,α2 given by (2.14). Hence, the spectrum of Hω is given again by (2.15), but now, as β>0 and ω(0,˜ω), there is no negative eigenvalue so that (ⅱ) and (ⅲ) are satisfied and the proof is complete.

    Moreover, in the defocusing case it is possible to give a further characterization of the standing waves, given by the following

    Theorem 3.3 (Ground states in the defocusing case). Let β>0 and μ>0. Then, the energy functional E restricted to the manifold Vμ has a unique (up to a phase factor) global minimizer, which is of the form (3.1) with ω=ωμ, where ωμ is the unique solution of

    log(2/ω)γ=2πβ(4πωμ)σ. (3.2)

    Proof. Preliminarily, one can see that (3.2) is equivalent to M(ω)=μ with M(ω) defined by (2.7). Hence, by Remark 3.1, there is a unique solution ωμ for any value of μ>0. It is thus clear that, if a minimizer does exist, then it has to be equal to uωμ up to phase factor.

    First, let us fix λ=ωμ in (1.7) and in the definition of the norm of Vμ, which is the same of V. Consider, therefore, a minimizing sequence {ψn}={ϕωμ,n+qnGωμ}Vμ for E. As ψn2=μ and β>0, E is coercive on Vμ and hence ψnωμC for every n. As a consequence there exists ψ=ϕωμ+qGωμV such that, up to subsequences,

    Furthermore, by the weak lower semicontinuity of E

    E(ψ)lim infn+E(ψn),

    and, by the weak lower semicontinuity of the norms, ψ2μ. Hence, if one can prove that ψ2=μ, the proof is complete.

    To this aim, first note that

    E(ψ)ωμψ2+(β|q|2σσ+1+log(ωμ/2)+γ2π)|q|2ωμμ+(β|q|2σσ+1+log(ωμ/2)+γ2π)|q|2=:f(|q|).

    Assuming that q is real-valued (which is not restrictive), one can check that f is minimized for q=q(ωμ) and that

    f(q(ωμ))=q2(ωμ)4πσβq2σ+2(ωμ)σ+1=E(uωμ).

    Therefore,

    E(ψ)E(uωμ)

    and, since M(uωμ)=μ, this implies that uωμ is the minimizer of E on Vμ up to a phase factor.

    The authors declare no conflict of interest.



    [1] Abramovitz M, Stegun IA (1965) Handbook of Mathematical Functions: with Formulas, Graphs, and Mathematical Tables, New York: Dover.
    [2] Adami R, Carlone R, Correggi M, et al. (2020) Blow-up for the pointwise NLS in dimension two: Absence of critical power. J Differ Equations 269: 1-37.
    [3] Adami R, Dell'Antonio G, Figari R, et al. (2003) The Cauchy problem for the Schr?dinger equation in dimension three with concentrated nonlinearity. Ann I H Poincaré Anal Non Linéaire 20: 477-500.
    [4] Adami R, Dell'Antonio G, Figari R, et al. (2004) Blow-up solutions for the Schr?dinger equation in dimension three with a concentrated nonlinearity. Ann I H Poincaré Anal Non Linéaire 21: 121-137.
    [5] Adami R, Noja D (2013) Stability and symmetry-breaking bifurcation for the ground states of a NLS with a δ interaction. Commun Math Phys 318: 247-289.
    [6] Adami R, Noja D, Ortoleva C (2013) Orbital and asymptotic stability for standing waves of a nonlinear Schr?dinger equation with concentrated nonlinearity in dimension three. J Math Phys 54: 013501.
    [7] Adami R, Noja D, Ortoleva C (2016) Asymptotic stability for standing waves of a NLS equation with subcritical concentrated nonlinearity in dimension three: neutral modes. Discrete Contin Dyn Syst 36: 5837-5879.
    [8] Adami R, Teta A (2001) A class of nonlinear Schr?dinger equations with concentrated nonlinearity. J Funct Anal 180: 148-175.
    [9] Azbel MY (1999) Quantum turbulence and resonant tunneling. Phys Rev B 59: 8049-8053.
    [10] Baro M, Neidhardt H, Rehberg J (2005) Current coupling of drift-diffusion models and Schr?dinger-Poisson systems: Dissipative hybrid models. SIAM J Math Anal 37: 941-981.
    [11] Bulashenko OM, Kochelap VA, Bonilla LL (1996) Coherent patterns and self-induced diffraction of electrons on a thin nonlinear layer. Phys Rev B 54: 1537-1540.
    [12] Cacciapuoti C, Carlone R, Noja D, et al. (2017) The 1-D Dirac equation with concentrated nonlinearity. SIAM J Math Anal 49: 2246-2268.
    [13] Carlone R, Correggi M, Tentarelli L (2019) Well-posedness of the two-dimensional nonlinear Schr?dinger equation with concentrated nonlinearity. Ann I H Poincaré Anal Non Linéaire 36: 257-294.
    [14] Carlone R, Figari R, Negulescu C (2017) The quantum beating and its numerical simulation. J Math Anal Appl 450: 1294-1316.
    [15] Carlone R, Finco D, Tentarelli L (2019) Nonlinear singular perturbations of the fractional Schr?dinger equation in dimension one. Nonlinearity 32: 3112-3143.
    [16] Carlone R, Fiorenza A, Tentarelli L (2017) The action of Volterra integral operators with highly singular kernels on H?lder continuous, Lebesgue and Sobolev functions. J Funct Anal 273: 1258-1294.
    [17] Cazenave T, Lions PL (1982) Orbital stability of standing waves for some nonlinear Schr?dinger equations. Commun Math Phys 85: 549-561.
    [18] Fukuizumi R, Ohta M, Ozawa T (2008) Nonlinear Schr?dinger equation with a point defect. Ann I H Poincaré Anal. Non Linéaire 25: 837-845.
    [19] Grillakis M, Shatah J, Strauss W (1987) Stability theory of solitary waves in the presence of symmetry. I. J Funct Anal 74: 160-197.
    [20] Jona-Lasinio G, Presilla C, Sjöstrand J (1995) On Schr?dinger equations with concentrated nonlinearities. Ann Phys 240: 1-21.
    [21] Merle F, Raphael P (2005) The blow-up dynamic and upper bound on the blow-up rate for critical nonlinear Schr?dinger equation. Ann Math 161: 157-222.
    [22] Noris B, Tavares H, Verzini G (2019) Normalized solutions for nonlinear Schr?dinger systems on bounded domains. Nonlinearity 32: 1044.
    [23] Nier F (1998) The dynamics of some quantum open systems with short-range nonlinearities. Nonlinearity 11: 1127-1172.
    [24] Pierotti D, Verzini G (2017) Normalized bound states for the nonlinear Schr?dinger equation in bounded domains. Calc Var Partial Dif 56: 133.
    [25] Sánchez ó, Soler J (2004) Asymptotic decay estimates for the repulsive Schr?dinger-Poisson system. Math Method Appl Sci 27: 371-380.
    [26] Shatah J, Strauss WA (1985) Instability of nonlinear bound states. Commun Math Phys 100: 173-190.
    [27] Sivan Y, Fibich G, Efremidis NK, et al. (2008) Analytic theory of narrow lattice solitons. Nonlinearity 21: 509-536.
    [28] Stubbe J (1991) Global solutions and stable ground states of nonlinear Schr?dinger equations. Phys D 48: 259-272.
    [29] Sulem C, Sulem PL (1999) The Nonlinear Schr?dinger Equation, Self-Focusing and Wave Collapse, New York: Springer-Verlag.
    [30] Weinstein MI (1986) Lyapunov stability of ground states of nonlinear dispersive evolution equations. Commun Pure Appl Math 39: 51-68.
    [31] Woolard DL, Cui HL, Gelmont BL, et al. (2003) Advanced theory of instability in tunneling nanostructures. Int J High Speed Electron Syst 13: 1149-1253.
  • This article has been cited by:

    1. Filippo Boni, Simone Dovetta, Prescribed mass ground states for a doubly nonlinear Schrödinger equation in dimension one, 2021, 496, 0022247X, 124797, 10.1016/j.jmaa.2020.124797
    2. Filippo Boni, Simone Dovetta, Doubly nonlinear Schrödinger ground states on metric graphs, 2022, 35, 0951-7715, 3283, 10.1088/1361-6544/ac7505
    3. William Borrelli, Raffaele Carlone, Lorenzo Tentarelli, Complete Ionization for a Non-autonomous Point Interaction Model in d = 2, 2022, 395, 0010-3616, 963, 10.1007/s00220-022-04447-1
    4. Riccardo Adami, Filippo Boni, Raffaele Carlone, Lorenzo Tentarelli, Ground states for the planar NLSE with a point defect as minimizers of the constrained energy, 2022, 61, 0944-2669, 10.1007/s00526-022-02310-8
    5. Riccardo Adami, Filippo Boni, Raffaele Carlone, Lorenzo Tentarelli, Existence, structure, and robustness of ground states of a NLSE in 3D with a point defect, 2022, 63, 0022-2488, 071501, 10.1063/5.0091334
    6. Riccardo Adami, Filippo Boni, Simone Dovetta, Competing nonlinearities in NLS equations as source of threshold phenomena on star graphs, 2022, 283, 00221236, 109483, 10.1016/j.jfa.2022.109483
    7. Noriyoshi Fukaya, Vladimir Georgiev, Masahiro Ikeda, On stability and instability of standing waves for 2d-nonlinear Schrödinger equations with point interaction, 2022, 321, 00220396, 258, 10.1016/j.jde.2022.03.008
    8. Lorenzo Tentarelli, A general review on the NLS equation with point-concentrated nonlinearity, 2023, 14, 2038-0909, 62, 10.2478/caim-2023-0004
    9. Domenico Finco, Lorenzo Tentarelli, Alessandro Teta, Well–posedness of the three–dimensional NLS equation with sphere–concentrated nonlinearity, 2024, 37, 0951-7715, 015009, 10.1088/1361-6544/ad0aac
    10. Riccardo Adami, Jinyeop Lee, Microscopic derivation of a Schrödinger equation in dimension one with a nonlinear point interaction, 2025, 288, 00221236, 110866, 10.1016/j.jfa.2025.110866
  • Reader Comments
  • © 2021 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(4464) PDF downloads(646) Cited by(10)

Figures and Tables

Figures(5)

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog