This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

On the inhomogeneous NLS with inverse-square potential

LUCCAS CAMPOS Luccas Campos Department of Mathematics, UFMG, Brazil. luccasccampos@gmail.com  and  CARLOS M. GUZMÁN CARLOS M. GUZMÁN Department of Mathematics, Fluminense Federal University, BRAZIL carlos.guz.j@gmail.com
Abstract.

We consider the inhomogeneous nonlinear Schrödinger equation with inverse-square potential in N\mathbb{R}^{N}

iut+au+λ|x|b|u|αu=0,a=Δa|x|2,iu_{t}+\mathcal{L}_{a}u+\lambda|x|^{-b}|u|^{\alpha}u=0,\;\;\mathcal{L}_{a}=\Delta-\frac{a}{|x|^{2}},

where λ=±1\lambda=\pm 1, α,b>0\alpha,b>0 and a>(N2)24a>-\frac{(N-2)^{2}}{4}. We first establish sufficient conditions for global existence and blow-up in Ha1(N)H^{1}_{a}(\mathbb{R}^{N}) for λ=1\lambda=1, using a Gagliardo-Nirenberg-type estimate. In the sequel, we study local and global well-posedness in Ha1(N)H^{1}_{a}(\mathbb{R}^{N}) in the H1H^{1}-subcritical case, applying the standard Strichartz estimates combined with the fixed point argument. The key to do that is to establish good estimates on the nonlinearity. Making use of these estimates, we also show a scattering criterion and construct a wave operator in Ha1(N)H^{1}_{a}(\mathbb{R}^{N}), for the mass-supercritical and energy-subcritical case.

Key words and phrases:
Inhomogeneous nonlinear Schrödinger equation; Local well-posedness; Global well-posedness; Blow-up; Inverse-square potential

1. Introduction

In this paper, we study the initial value problem (IVP) associated to the inhomogeneous nonlinear Schrödinger equation (INLSa for short)

{ituau+λ|x|b|u|αu=0,t,xN,u(0,x)=u0(x),\begin{cases}i\partial_{t}u-\mathcal{L}_{a}u+\lambda|x|^{-b}|u|^{\alpha}u=0,\;\;\;t\in\mathbb{R},\;x\in\mathbb{R}^{N},\\ u(0,x)=u_{0}(x),\end{cases} (1.1)

where N3N\geq 3, a=Δ+a|x|2\mathcal{L}_{a}=-\Delta+\frac{a}{|x|^{2}} with a>(N2)24a>-\frac{(N-2)^{2}}{4} and α,b>0\alpha,b>0. The equation is called “focusing INLSa” when λ=+1\lambda=+1 and “defocusing INLSa” when λ=1\lambda=-1. The restriction on aa comes from the sharp Hardy inequality, namely

(N2)24|x|2|u(x)|2𝑑x|u(x)|𝑑x,uH1,\frac{(N-2)^{2}}{4}\int|x|^{-2}|u(x)|^{2}dx\leq\int|\nabla u(x)|dx,\;\;\forall\;u\in H^{1}, (1.2)

which guarantees that a\mathcal{L}_{a} is a positive operator. It is a model from various physical contexts, for example, in nonlinear optical systems with spatially dependent interactions (see e.g. [1] and the references therein). In particular, when a=0a=0, it can be thought of as modeling inhomogeneities in the medium in which the wave propagates (see for instance [18]). When b=0b=0, the equation (1.1) appears in several physical settings, such a quantum field equations or black hole solutions of the Einstein’s equations (see e.g. [17]).

The INLSa model can be seen as an extension to the Schrödinger equation. For instance, when a=b=0a=b=0 we have the classical nonlinear Schrödinger (denoted by NLS) equation, extensively studied over the three decades. (see [2], [6], [20], [26] and the references therein). The case b=0b=0 and a0a\neq 0 is known as the NLS equation with inverse square potential, denoted by NLSa equation, that is,

ituau+|u|αu=0.i\partial_{t}u-\mathcal{L}_{a}u+|u|^{\alpha}u=0. (1.3)

This equation has also been intensively studied in recent years (see for instance, [3], [21], [22], [23], [27]). Moreover, when a=0a=0 and b0b\neq 0 we have the inhomogeneous NLS equation, denoted by (INLS), i.e.,

itu+Δu+|x|b|u|αu=0,i\partial_{t}u+\Delta u+|x|^{-b}|u|^{\alpha}u=0, (1.4)

which also has received substantial attention recently (see e.g. [16], [11], [12], [4], [8], [21]).

Similarly to the INLS, the INLSa equation is invariant under the scaling, uμ(t,x)=μ2bαu(μ2t,μx)u_{\mu}(t,x)=\mu^{\frac{2-b}{\alpha}}u(\mu^{2}t,\mu x) for μ>0\mu>0. A straightforward computation yields

u0,μH˙s=μsN2+2bαu0H˙s,\|u_{0,\mu}\|_{\dot{H}^{s}}=\mu^{s-\frac{N}{2}+\frac{2-b}{\alpha}}\|u_{0}\|_{\dot{H}^{s}},

implying that the scale-invariant Sobolev space is H˙sc(N)\dot{H}^{s_{c}}(\mathbb{R}^{N}), with sc=N22bαs_{c}=\frac{N}{2}-\frac{2-b}{\alpha}, the so called critical Sobolev index. If sc=0s_{c}=0 (or α=42bN\alpha=\frac{4-2b}{N}) the (IVP) is known as mass-critical or L2L^{2}-critical; if sc<0s_{c}<0 (or 0<α<42bN0<\alpha<\frac{4-2b}{N}) it is called mass-subcritical or L2L^{2}-subcritical; if 0<sc<20<s_{c}<2, (1.1) is known as mass-supercritical and energy-subcritical (or intercritical). In additional, solutions to (1.1) conserve their mass and energy, defined respectively by

M[u(t)]=N|u(t,x)|2𝑑x=M[u0],M[u(t)]=\int_{\mathbb{R}^{N}}|u(t,x)|^{2}dx=M[u_{0}], (1.5)
Ea[u(t)]=12N|u(t,x)|2𝑑x+a2N|x|2|u(t,x)|2𝑑x+λα+2N|u|α+2𝑑x=Ea[u0].E_{a}[u(t)]=\frac{1}{2}\int_{\mathbb{R}^{N}}|\nabla u(t,x)|^{2}dx+\frac{a}{2}\int_{\mathbb{R}^{N}}|x|^{-2}|u(t,x)|^{2}dx+\frac{\lambda}{\alpha+2}\int_{\mathbb{R}^{N}}|u|^{\alpha+2}dx=E_{a}[u_{0}]. (1.6)

When a=0a=0, we denote EaE_{a} by E0E_{0}.

We review now some recent developments in H1(N)H^{1}(\mathbb{R}^{N}), for the particular cases, i.e., NLSa and INLS models, starting with the NLSa equation. Okazawa-Suzuki-Yokota [23], by the energy method, showed local well-posedness in the energy-subcritical case, for N3N\geq 3 and a>(N2)24a>-\tfrac{(N-2)^{2}}{4}. They also proved that the solutions are global if λ=1\lambda=-1 or 0<α<4N0<\alpha<\frac{4}{N} and λ=1\lambda=1. Zhang-Zheng [27] studied the defocusing case, establishing well posedness and scattering for 4N<α<4N2\tfrac{4}{N}<\alpha<\tfrac{4}{N-2}, assuming a0a\geq 0 and N=3N=3, or a(N2)24+4(α+2)2a\geq-\tfrac{(N-2)^{2}}{4}+\tfrac{4}{(\alpha+2)^{2}} and N4N\geq 4. Recently, Killip-Murphy-Visan-Zheng in [21] considered the focusing 3D3D cubic NLSa. They established well-posedness and scattering in H1(N)H^{1}(\mathbb{R}^{N}) if a>14a>-\tfrac{1}{4}. Later, Lu-Miao-Murphy [22] extended the result of [21] to all L2L^{2}-supercritical, energy subcritical nonlinearities, in dimensions, 3N63\leq N\leq 6, with

{a>14ifN=3,43<α2a>(N2)24+(N221α)2if   3N6,max{2N2,4N}<α<4N2.\begin{cases}a>-\tfrac{1}{4}\;\;\;\;\;\qquad\qquad\qquad\qquad\textnormal{if}\;\;\;N=3,\;\;\tfrac{4}{3}<\alpha\leq 2\\ a>-\frac{(N-2)^{2}}{4}+(\frac{N-2}{2}-\frac{1}{\alpha})^{2}\;\;\textnormal{if}\;\;\;3\leq N\leq 6,\;\;\max\{\frac{2}{N-2},\frac{4}{N}\}<\alpha<\frac{4}{N-2}.\end{cases} (1.7)

On the other hand, the INLS equation was first studied by Genoud-Stuart [15] via the energy method. For N1N\geq 1 and 0<b<min{N,2}0<b<\min\{N,2\}, they showed local well posedness in H1(N)H^{1}(\mathbb{R}^{N}) for the H1H^{1}-subcritical case and global well possedness in the mass-subcritical case. In the mass-critical case, Genoud in [14] established global well-posedness in H1(N)H^{1}(\mathbb{R}^{N}), provided that the mass of the initial data is below that of the associated ground state. This result was extended in the intercritical case by Farah [10] and he also showed that the solution blows-up in finite time (see also [9]). The second author in [16], using Kato’s method, proved local well-posedness in H1(N)H^{1}(\mathbb{R}^{N}), for the energy subcritical case in dimensions N4N\geq 4 and 0<b<20<b<2. Cho-Lee [7] treated the case N=3N=3 for 0<b<320<b<\frac{3}{2} and Dinh [8] the case N=2N=2 for 0<b<10<b<1. Furthermore, in the intercritical case, the second author in [16] also established a small data global theory in H1(N)H^{1}(\mathbb{R}^{N}) for N4N\geq 4, the first author [4] treated the case N=3N=3 and Cardoso-Farah-Guzmán [5] the case N=2N=2. In all these works the range of bb is the same one where local well-posedness was obtained.

Motivated by the aforementioned papers, our main interest in this paper is to prove similar results for the INLSa equation in Ha1(N)H^{1}_{a}(\mathbb{R}^{N}). The well-posedness theory was already studied by Suzuki [25]. Using the energy method, the author showed that, if111It is worth mentioning that in [25] the author considered (1.1) with a=(N2)24a=-\frac{(N-2)^{2}}{4}, the critical coefficient. The proof for the case a>(N2)24a>\frac{(N-2)^{2}}{4} is an immediate consequence of the previous one.

N3,0<α<42bN2a>(N2)24,and0<b<2,N\geq 3,\qquad 0<\alpha<\tfrac{4-2b}{N-2}\qquad a>-\tfrac{(N-2)^{2}}{4},\qquad\textnormal{and}\qquad 0<b<2, (1.8)

then (1.1) is locally well-posed in Ha1(N)H^{1}_{a}(\mathbb{R}^{N}) satisfying uC([0,T);H1(N))C1([0,T);H1(N))u\in C\left([0,T);H^{1}(\mathbb{R}^{N})\right)\cap C^{1}\left([0,T);H^{-1}(\mathbb{R}^{N})\right) for some T>0T>0. It was also proved that any local solution of the IVP (1.1) with u0Ha1(N)u_{0}\in H^{1}_{a}(\mathbb{R}^{N}) extends globally in time if either λ=1\lambda=-1 (defocusing case) or 0<α<42bN0<\alpha<\tfrac{4-2b}{N} for λ=1\lambda=1 (focusing, L2L^{2}-subcritical case).

Our first goal here is to study global existence and blow-up in H1(N)H^{1}(\mathbb{R}^{N}) for λ=1\lambda=1 for both the L2L^{2}-critical and the intercritical (L2L^{2}-supercriticial and H1H^{1}-subcritical) case. For that, we apply a sharp Gagliardo-Nirenberg type estimate. In order to do so, we prove the existence of a ground state.

Proposition 1.1.

Let N3N\geq 3 and α\alpha, aa, bb as in (1.8) and λ=1\lambda=1. There exists a positive solution to the elliptic equation

aQ+Q|x|b|Q|αQ=0\mathcal{L}_{a}Q+Q-|x|^{-b}|Q|^{\alpha}Q=0

in Ha1(N)H^{1}_{a}(\mathbb{R}^{N}). Moreover, all possible solutions have the same mass, the same H˙a1\dot{H}_{a}^{1} norm and the same energy.

Using the variational analysis associated to Proposition 1.1, we have thresholds for global existence and blow-up.

Theorem 1.2 (L2L^{2}-critical case).

Let NN, aa, bb as in (1.8), λ=1\lambda=1 and α=42bN\alpha=\frac{4-2b}{N}. Suppose that uu is the solution to (1.1) with initial data u0Ha1(N)u_{0}\in H^{1}_{a}(\mathbb{R}^{N}). If

  • a)

    (Global existence) u0L2<QL2\|u_{0}\|_{L^{2}}<\|Q\|_{L^{2}}, then the solution is uniformly bounded in Ha1H^{1}_{a}, and therefore extends globally in time;

  • b)

    (Blow-up) E[u0]<0E[u_{0}]<0 and either |x|u0L2(N)|x|u_{0}\in L^{2}(\mathbb{R}^{N}) or u0u_{0} is radial, then uu blows-up in finite positive and negative times.

Theorem 1.3 (Intercritical case).

Let NN, aa, bb as in (1.8), λ=1\lambda=1 and 42bN<α<42bN2\frac{4-2b}{N}<\alpha<\frac{4-2b}{N-2}. Suppose that uu is the solution of (1.1) with initial data u0Ha1(N)u_{0}\in H^{1}_{a}(\mathbb{R}^{N}). Assume

M(u0)1scEa(u0)sc<M(Q)1scEa(Q)sc.M(u_{0})^{1-s_{c}}E_{a}(u_{0})^{s_{c}}<M(Q)^{1-s_{c}}E_{a}(Q)^{s_{c}}. (1.9)

If

  • a)

    (Global existence)

    u0L21scau0L2sc<QL21scaQL2sc,\|u_{0}\|^{1-s_{c}}_{L^{2}}\|\sqrt{\mathcal{L}_{a}}u_{0}\|^{s_{c}}_{L^{2}}<\|Q\|^{1-s_{c}}_{L^{2}}\|\sqrt{\mathcal{L}_{a}}Q\|^{s_{c}}_{L^{2}}, (1.10)

    then the solution is uniformly bounded in Ha1H^{1}_{a}, and therefore extends globally in time;

  • b)

    (Blow-up)

    u0L21scscau(t)L2>QL21scaQL2sc\|u_{0}\|^{\frac{1-s_{c}}{s_{c}}}_{L^{2}}\|\sqrt{\mathcal{L}_{a}}u(t)\|_{L^{2}}>\|Q\|^{1-s_{c}}_{L^{2}}\|\sqrt{\mathcal{L}_{a}}Q\|^{s_{c}}_{L^{2}} (1.11)

    and, in addition, if |x|u0L2(N)|x|u_{0}\in L^{2}(\mathbb{R}^{N}) or u0u_{0} is radial, then uu blows-up in finite positive and negative times.

The main tools for proving Theorems 1.2 and 1.3 are the coercivity given by the variational analysis, and the virial identities. The main difficulty is to appropriately control the error terms appearing when one truncates the virial identity in the radial case. The L2L^{2}-critical case is especially delicate, since the criticality gives us less room for error terms.

Note that the local well-posedness showed in [25] as well as the global results, Theorems 1.2a) and 1.3a) ensure the existence of solutions to (1.1). However we do not know whether or not the solutions satisfy uLq(I;Ha1,r)u\in L^{q}(I;H^{1,r}_{a}) for any L2L^{2}-admissible pair (q,rq,r), which is a key property to study other problems such as scattering for example. To obtain this extra information and working towards the proof of scattering in a next work, we establish the local and global well-posedness for (1.1) via Kato’s method, which is based on the contraction mapping principle and the Strichartz estimates. We start with the local theory for the energy-subcritical case.

Theorem 1.4.

Assume that N3N\geq 3 and 0b<min{N2,2}0\leq b<\min\{\frac{N}{2},2\}. If u0Ha1(N)u_{0}\in H^{1}_{a}(\mathbb{R}^{N}) and

{a>(N2)24if22bN<α22bN2and  0b<1,a>(N2)24+(α(N2)(22b)2(α+1))2ifmax{0,22bN2}<α<42bN2.\begin{cases}a>-\tfrac{(N-2)^{2}}{4}\;\;\;\;\qquad\qquad\qquad\qquad\ \textnormal{if}\;\;\;\;\tfrac{2-2b}{N}<\alpha\leq\;\frac{2-2b}{N-2}\quad\textnormal{and}\;\;0\leq b<1,\\ a>-\tfrac{(N-2)^{2}}{4}+\left(\tfrac{\alpha(N-2)-(2-2b)}{2(\alpha+1)}\right)^{2}\;\;\textnormal{if}\;\;\;\max\{0,\tfrac{2-2b}{N-2}\}<\alpha<\tfrac{4-2b}{N-2}.\end{cases} (1.12)

Then there exists T=T(u0Ha1,N,α,b)T=T(\|u_{0}\|_{H^{1}_{a}},N,\alpha,b) and a unique solution of (1.1) satisfying

uC([0,T];Ha1(N))Lq([0,T];Ha1,r(N)),u\in C\left([0,T];H^{1}_{a}(\mathbb{R}^{N})\right)\cap L^{q}\left([0,T];H^{1,r}_{a}(\mathbb{R}^{N})\right),

where (q,rq,r) is any L2L^{2}-admissible pair.

New challenges and technical obstructions appear with the presence of the functions |x|2|x|^{-2} and |x|b|x|^{-b} in (INLSa), related especially to the problem of equivalence of Sobolev spaces. This leads us to impose some technical restrictions on the parameters α\alpha, bb and aa, given in (1.12).

Remark 1.5.

Observe that Theorem 1.1 also holds for b=0b=0, thus we have that (NLSa) is locally well-posed in H1(N)H^{1}(\mathbb{R}^{N}). In this particular case, we have a local result a little diferent from Luo-Miao-Murphy in [22], they showed local well posedness for 3N63\leq N\leq 6, assuming (1.7). Our result holds for any dimensions N3N\geq 3, however the condition on aa is weaker than Luo-Miao-Murphy’result. Note also that we improve the range of the parameter α\alpha to222Note that, in Theorem 1.4 we have the condition α>22bN\alpha>\frac{2-2b}{N}, however when b=0b=0 we can even consider α=2N\alpha=\frac{2}{N} (see Lemma 4.3). αN2\alpha\geq\frac{N}{2}. On the other hand, if b<1b<1 then we have a lower bound for the parameter α\alpha in Theorem 1.4 and if b1b\geq 1 we then need α>0\alpha>0.

As an immediately consequence we obtain the following result.

Corollary 1.6.

Assume one of the following conditions:

  • (i)

    N3N\geq 3, 1b<min{N2,4}1\leq b<\min\left\{\frac{N}{2},4\right\}, 0<α<42bN20<\alpha<\frac{4-2b}{N-2} and a>(N2)24+(α(N2)(22b)2(α+1))2a>-\frac{(N-2)^{2}}{4}+\left(\frac{\alpha(N-2)-(2-2b)}{2(\alpha+1)}\right)^{2};

  • (ii)

    N=3N=3, α=2\alpha=2, 0<b<10<b<1 and a>14+b29a>-\frac{1}{4}+\frac{b^{2}}{9}.

If u0Ha1(N)u_{0}\in H^{1}_{a}(\mathbb{R}^{N}), then the same result of Theorem 1.4 holds.

It is worth mentioning that Corollary 1.6 (ii) can be seen as an extension of a local result by Killip-Murphy-Visan-Zheng [21] to the INLSa model.

Remark 1.7.

The range of the parameters bb and aa in Theorem 1.4 are more restricted than in [25]. That is, applying the energy method we obtain a better result than using the Kato method, however in Theorem 1.4 we obtain an extra information on the solution.

In the sequel we establish small data global results in Ha1(N)H^{1}_{a}(\mathbb{R}^{N}), for N3N\geq 3 and 42bN<α<42bN2\tfrac{4-2b}{N}<\alpha<\frac{4-2b}{N-2}. Since we use the Strichartz estimates (see Section 2) we show global well-posedness for radial and non-radial initial data. Here, eitau0e^{-it\mathcal{L}_{a}}u_{0} denotes the solution to the linear problem associated to (1.1) and the Strichartz norm S(H˙sc)\|\cdot\|_{S(\dot{H}^{s_{c}})} is defined in Section 2.1.

Theorem 1.8 (Radial small data theory).

Let a>0a>0, 0<b<min{0,N2}0<b<\min\{0,\frac{N}{2}\} and 42bN<α<42bN2\frac{4-2b}{N}<\alpha<\frac{4-2b}{N-2} (α<32b(\alpha<3-2b if N=3)N=3). If u0Ha1(N)u_{0}\in H^{1}_{a}(\mathbb{R}^{N}) radial with u0H1M\|u_{0}\|_{H^{1}}\leq M then there exists a unique global solution uu of (1.1) such that if eitau0S(H˙sc)<δsd\|e^{-it\mathcal{L}_{a}}u_{0}\|_{S(\dot{H}^{s_{c}})}<\delta_{sd}, there exists a unique global solution uu of (1.1) such that

uS(H˙sc)2eitau0S(H˙sc)anduS(L2)+auS(L2)2cu0Ha1,\|u\|_{S(\dot{H}^{s_{c}})}\leq 2\|e^{-it\mathcal{L}_{a}}u_{0}\|_{S(\dot{H}^{s_{c}})}\quad\textnormal{and}\quad\|u\|_{S\left(L^{2}\right)}+\|\sqrt{\mathcal{L}_{a}}u\|_{S\left(L^{2}\right)}\leq 2c\|u_{0}\|_{H^{1}_{a}},

for some universal constant c>0c>0.

In Theorem 1.8, when N=3N=3, we have an extra restriction on α\alpha, namely α<32b\alpha<3-2b. To reach α<42b\alpha<4-2b, we need to restrict the parameter bb. This restriction comes from the need of α>1\alpha>1 in the fixed point argument. To this end, we use the norm uLta¯Lxr¯\|u\|_{L^{\bar{a}}_{t}L^{\bar{r}}_{x}} denoted by

uS~(H˙sc)=uLta¯Lxr¯,\|u\|_{\widetilde{S}(\dot{H}^{s_{c}})}=\|u\|_{L^{\bar{a}}_{t}L^{\bar{r}}_{x}}, (1.13)

where (a¯,r¯)\bar{a},\bar{r}) only satisfies r¯2\bar{r}\geq 2 and the relation of H˙sc\dot{H}^{s_{c}}-admissible pair, i.e., 2a¯=N2Nr¯sc\frac{2}{\bar{a}}=\frac{N}{2}-\frac{N}{\bar{r}}-s_{c} and not the remaining conditions (see Section 2).

Theorem 1.9.

Let N=3N=3, a>0a>0 and 0<b<320<b<\tfrac{3}{2}. Assume u0Ha1(N)u_{0}\in H^{1}_{a}(\mathbb{R}^{N}) and one of the following conditions:

  • (i)

    max{1,42bN}<α<42b\max\{1,\tfrac{4-2b}{N}\}<\alpha<4-2b;

  • (ii)

    42bN<α<42b\tfrac{4-2b}{N}<\alpha<4-2b  and  0<b<120<b<\frac{1}{2};

  • (iii)

    32bα<42b3-2b\leq\alpha<4-2b  and  0<b<10<b<1,

then the same result as in Theorem 1.8 holds, replacing S(H˙sc)\|\cdot\|_{S(\dot{H}^{s_{c}})} by S~(H˙sc)\|\cdot\|_{\widetilde{S}(\dot{H}^{s_{c}})}.

We remark that Theorem 1.9 holds for general initial data. On the other hand, Theorem 1.9-(ii) shows global well posedness in the full intercritical regime, however with b<12b<\frac{1}{2}. The gap 12b<32\frac{1}{2}\leq b<\frac{3}{2} is still an open problem. Moreover, in the particular case333The case 42b3<α<32b\frac{4-2b}{3}<\alpha<3-2b was obtained in Theorem 1.8. 32bα<42b3-2b\leq\alpha<4-2b, we have a better range for bb than in (ii).

The next result holds for non-radial data and a<0a<0, however only for dimensions N=3,4,5N=3,4,5. Here, we also use the norm S~(H˙sc)\|\cdot\|_{\widetilde{S}(\dot{H}^{s_{c}})}.

Theorem 1.10 (Small data theory).

Let 0<b<6N20<b<\tfrac{6-N}{2} and u0Ha1u_{0}\in H^{1}_{a} with u0Ha1M\|u_{0}\|_{H^{1}_{a}}\leq M, for some M>0M>0. Assume that (N,a,α)(N,a,\alpha) satisfy

{a>14ifN=3,42b3<α 22band  0b<12,a>(N2)24+(α(N2)(22b)2(α+1))2if  3N5,max{42bN,22bN2,1}<α<42bN2.\begin{cases}a>-\tfrac{1}{4}\;\quad\qquad\qquad\qquad\qquad\qquad\ \textnormal{if}\;\;N=3,\;\;\;\tfrac{4-2b}{3}<\alpha\leq\;2-2b\quad\textnormal{and}\;\;0\leq b<\tfrac{1}{2},\\ a>-\tfrac{(N-2)^{2}}{4}+\left(\tfrac{\alpha(N-2)-(2-2b)}{2(\alpha+1)}\right)^{2}\;\;\textnormal{if}\;\;3\leq N\leq 5,\;\;\;\max\{\tfrac{4-2b}{N},\frac{2-2b}{N-2},1\}<\alpha<\tfrac{4-2b}{N-2}.\end{cases} (1.14)

Then there exists δsd=δsd(M)>0\delta_{sd}=\delta_{sd}(M)>0 such that if eitau0S~(H˙sc)<δsd\|e^{-it\mathcal{L}_{a}}u_{0}\|_{\widetilde{S}(\dot{H}^{s_{c}})}<\delta_{sd}, then there exists a unique global solution uu of (1.1) such that

uS~(H˙sc)2eitau0S~(H˙sc)anduS(L2)+auS(L2)2cu0Ha1,\|u\|_{\widetilde{S}(\dot{H}^{s_{c}})}\leq 2\|e^{-it\mathcal{L}_{a}}u_{0}\|_{\widetilde{S}(\dot{H}^{s_{c}})}\quad\textnormal{and}\quad\|u\|_{S\left(L^{2}\right)}+\|\sqrt{\mathcal{L}_{a}}u\|_{S\left(L^{2}\right)}\leq 2c\|u_{0}\|_{H^{1}_{a}}, (1.15)

for some universal constant c>0c>0.

Remark 1.11.

The results above still hold, with the same proof, if one restricts the time interval to [t0,+)[t_{0},+\infty) or (,t0](-\infty,t_{0}], instead of \mathbb{R}, where u(t0)=u0u(t_{0})=u_{0}. By time-translation invariance, we assume t0=0t_{0}=0 in Theorem 1.10.

As mentioned above the main tool to show the local and global well-posedness is the Fixed Point Theorem, which is based on the Strichartz estimates. Similarly as in the local theory the main difficulty here is to look for admissible pairs to establish the equivalence of Sobolev spaces, mainly when a<0a<0.

Once global results are proved, the natural route is to study the asymptotic behavior of such global solutions as t±t\rightarrow\pm\infty. We show that our solutions scatter to a solution of the linear problem in Ha1(N)H^{1}_{a}(\mathbb{R}^{N}). In addition, we construct the wave operator associated to eq. (1.1). This is the reciprocal problem of the scattering theory, which consists in constructing a solution with a prescribed scattering state.

Theorem 1.12.

Assume the assumptions in Theorems 1.8, 1.9 and 1.10.

  • (i)

    Let u(t)u(t) be a global solution of (1.1) with initial data u0Ha1(N)u_{0}\in H^{1}_{a}(\mathbb{R}^{N}). If suptu(t)Ha1M\sup\limits_{t\in\mathbb{R}}\|u(t)\|_{H^{1}_{a}}\leq M and uS(H˙sc)<+(oruS~(H˙sc)<+)\|u\|_{S(\dot{H}^{s_{c}})}<+\infty\;\;(\textnormal{or}\;\|u\|_{\widetilde{S}(\dot{H}^{s_{c}})}<+\infty), then u(t)u(t) scatters in Ha1H^{1}_{a}, that is, there exists ϕ±Ha1\phi^{\pm}\in H^{1}_{a} such that

    limt±u(t)eitaϕ±Ha1=0.\lim_{t\rightarrow\pm\infty}\|u(t)-e^{-it\mathcal{L}_{a}}\phi^{\pm}\|_{H^{1}_{a}}=0.
  • (ii)

    For any ϕHa1(N)\phi\in H^{1}_{a}\left(\mathbb{R}^{N}\right), there exist T0>0T_{0}>0 and uC([T0,):Ha1(N))u\in C([T_{0},\infty):H^{1}_{a}(\mathbb{R}^{N})) solution of (1.1) satisfying

    limtu(t)eitaϕHa1=0.\lim_{t\rightarrow\;\infty}\|u(t)-e^{-it\mathcal{L}_{a}}\phi\|_{H^{1}_{a}}=0.

    The analogous statement holds backward in time.

The rest of the paper is organized as follows. In Section 2, we introduce some notations and give a review of the Strichartz estimates. In Section 3 we discuss the existence of a ground state and establish global existence as well as blow up in Ha1(N)H^{1}_{a}(\mathbb{R}^{N}), for L2L^{2}-critical and intercritical cases. In Section 4 we study the local and global well-posedness applying the contraction mapping principle. Finally, in Section 5 we prove Theorem 1.12.

2. Notation and Preliminaries

In this section, we introduce the notation used throughout the paper and list some useful results. We use CC to denote various constants that may vary line by line. If aa and bb be positive real numbers, the notation aba\lesssim b means that there exists a positive constant C>0C>0 such that444The constant CC may depend on parameters, such as the dimension NN, as well on a priori estimates on the solution, but never on the solution itself or on time aCba\leq Cb. The notation aba\sim b means aba\lesssim b and bab\lesssim a. Given a real number rr, we use r±r^{\pm} to denote r±εr\pm\varepsilon for some ε>0\varepsilon>0 sufficiently small. For a subset ANA\subset\mathbb{R}^{N}, its complement is denoted by AC=N\AA^{C}=\mathbb{R}^{N}\backslash A and the characteristic function χA\chi_{A} denotes the function that has value 11 at points of AA and 0 at points of ACA^{C}. Given x,yNx,y\in\mathbb{R}^{N}, xyx\cdot y denotes the usual inner product of xx and yy in N\mathbb{R}^{N}.

The norm in the Sobolev spaces H˙s,r=H˙s,r(N)\dot{H}^{s,r}=\dot{H}^{s,r}(\mathbb{R}^{N}) and Hs,r=Hs,r(N)H^{s,r}=H^{s,r}(\mathbb{R}^{N}), are defined by

fH˙s,r:=DsfLrandfHs,r:=DsfLr,\|f\|_{\dot{H}^{s,r}}:=\|D^{s}f\|_{L^{r}}\quad\textnormal{and}\quad\|f\|_{H^{s,r}}:=\|\langle D\rangle^{s}f\|_{L^{r}}, (2.1)

where Dsf:=Δsf=(|ξ|sf^)D^{s}f:=\sqrt{-\Delta}^{s}f=(|\xi|^{s}\widehat{f})^{\vee} and =(1+||2)12\langle\cdot\rangle=(1+|\cdot|^{2})^{\frac{1}{2}}. If r=2r=2 we denote Hs,2H^{s,2} and H˙s,2\dot{H}^{s,2} simply by HsH^{s} and H˙s\dot{H}^{s}, respectively. Similarly, we define Sobolev spaces H˙as,r\dot{H}_{a}^{s,r} and Has,rH_{a}^{s,r} associated to a\mathcal{L}_{a} by the closure of 0(N\{0})\mathbb{C}^{\infty}_{0}(\mathbb{R}^{N}\backslash\{0\}) under the norms

uH˙as,r:=asuLranduHas,r:=asuLr.\|u\|_{\dot{H}_{a}^{s,r}}:=\|\sqrt{\mathcal{L}_{a}}^{s}u\|_{L^{r}}\quad\textnormal{and}\quad\|u\|_{H_{a}^{s,r}}:=\|\langle\sqrt{\mathcal{L}_{a}}\rangle^{s}u\|_{L^{r}}. (2.2)

We abbreviate H˙as(N)=H˙as,2(N))\dot{H}_{a}^{s}(\mathbb{R}^{N})=\dot{H}_{a}^{s,2}(\mathbb{R}^{N})) and Has(N)=Has,2(N)H_{a}^{s}(\mathbb{R}^{N})=H_{a}^{s,2}(\mathbb{R}^{N}). Note that, by the sharp Hardy inequality, one has

uH˙a1uH˙1fora>(N22)2.\|u\|_{\dot{H}_{a}^{1}}\sim\|u\|_{\dot{H}^{1}}\quad\textnormal{for}\quad a>-\left(\frac{N-2}{2}\right)^{2}. (2.3)

We also define, for 1p<1\leq p<\infty, the weighted Sobolev space Lbp=Lbp(N)={f:fp,b<+}L^{p}_{b}=L^{p}_{b}(\mathbb{R}^{N})=\{f:\;\|f\|_{p,b}<+\infty\}, where

fp,b=[|x|b|f(x)|p𝑑x]1p.\|f\|_{p,b}=\left[\int|x|^{-b}|f(x)|^{p}\,dx\right]^{\frac{1}{p}}. (2.4)

Let q,r>0q,r>0, ss\in\mathbb{R}, and II\subset\mathbb{R} an interval; the mixed norms in the spaces LIqLxrL^{q}_{I}L^{r}_{x} and LIqHxsL^{q}_{I}H^{s}_{x} of a function f=f(t,x)f=f(t,x) are defined as

fLIqLxr=(If(t,)Lxrq𝑑t)1qandfLIqHxs=(If(t,)Hxsq𝑑t)1q,\|f\|_{L^{q}_{I}L^{r}_{x}}=\left(\int_{I}\|f(t,\cdot)\|^{q}_{L^{r}_{x}}dt\right)^{\frac{1}{q}}\qquad\mbox{and}\qquad\|f\|_{L^{q}_{I}H^{s}_{x}}=\left(\int_{I}\|f(t,\cdot)\|^{q}_{H^{s}_{x}}dt\right)^{\frac{1}{q}},

with the usual modifications if either q=q=\infty or r=r=\infty. When the space 99+-6integration is restricted to a subset ANA\subset\mathbb{R}^{N} then the mixed norm will be denoted by fLIqLxr(A)\|f\|_{L_{I}^{q}L^{r}_{x}(A)}. Moreover, if I=I=\mathbb{R} we shall use the notations fLtqLxr\|f\|_{L_{t}^{q}L^{r}_{x}} and fLtqHxs\|f\|_{L_{t}^{q}H^{s}_{x}}.

Next, we recall some important inequalities. To state the estimates below, it is useful to introduce the parameter

ρ=(N2)(N2)2+4a2.\rho=\frac{(N-2)-\sqrt{(N-2)^{2}+4a}}{2}. (2.5)
Lemma 2.1 (Equivalence of Sobolev spaces).

Fix N3N\geq 3, a(N22)2a\geq-(\frac{N-2}{2})^{2}, and 0<s<20<s<2. If 1<p<1<p<\infty satisfies s+ρN<1p<min{1,NρN}\frac{s+\rho}{N}<\frac{1}{p}<\min\{1,\frac{N-\rho}{N}\}, then

DsfLpas2fLpfor allf0(N\{0}).\|D^{s}f\|_{L^{p}}\lesssim\|\mathcal{L}_{a}^{\frac{s}{2}}f\|_{L^{p}}\;\textnormal{for all}\;f\in\mathbb{C}^{\infty}_{0}(\mathbb{R}^{N}\backslash\{0\}). (2.6)

If max{sN,ρN}<1p<min{1,NρN}\max\{\frac{s}{N},\frac{\rho}{N}\}<\frac{1}{p}<\min\{1,\frac{N-\rho}{N}\}, then

as2fLpDsfLpfor allf0(N\{0}).\|\mathcal{L}_{a}^{\frac{s}{2}}f\|_{L^{p}}\lesssim\|D^{s}f\|_{L^{p}}\;\textnormal{for all}\;f\in\mathbb{C}^{\infty}_{0}(\mathbb{R}^{N}\backslash\{0\}). (2.7)
Proof.

See [19]

Remark 2.2.

Let 0<s<20<s<2. It is easy to see that, if a>0a>0 then fHas,rfHs,r\|f\|_{H_{a}^{s,r}}\sim\|f\|_{H^{s,r}}, provided that 1<r<Ns1<r<\frac{N}{s}. When (N2)24a<0-\frac{(N-2)^{2}}{4}\leq a<0 we have 0<ρ<(N2)20<\rho<\frac{(N-2)}{2} and so fHas,rfHs,r\|f\|_{H_{a}^{s,r}}\sim\|f\|_{H^{s,r}} if NNρ<r<Ns+ρ\frac{N}{N-\rho}<r<\frac{N}{s+\rho}.

Lemma 2.3.

If fH1(N)f\in H^{1}(\mathbb{R}^{N}) is radial, N2N\geq 2, then, for any R>0R>0,

fL{|x|R}RN12fL212fL212.\|f\|_{L^{\infty}_{\left\{|x|\geq R\right\}}}\lesssim R^{-\frac{N-1}{2}}\|f\|_{L^{2}}^{\frac{1}{2}}\|\nabla f\|_{L^{2}}^{\frac{1}{2}}. (2.8)
Proof.

See Strauss [24]. ∎

The next lemma implies that the Strichartz estimates (Lemma 2.6) hold.

Lemma 2.4 (Dispersive estimate).

Let f be a radial function.

  • (i)

    If a0a\geq 0, then we have

    eitafL|t|N2fL1.\|e^{it\mathcal{L}_{a}}f\|_{L^{\infty}}\lesssim|t|^{-\frac{N}{2}}\|f\|_{L^{1}}. (2.9)
  • (ii)

    If (N2)24<a<0-\frac{(N-2)^{2}}{4}<a<0, then

    (1+|x|ρ)1eitafL1+|t|ρ|t|N2(1+|x|ρ)fL1,\|(1+|x|^{-\rho})^{-1}e^{it\mathcal{L}_{a}}f\|_{L^{\infty}}\lesssim\frac{1+|t|^{\rho}}{|t|^{\frac{N}{2}}}\|(1+|x|^{-\rho})f\|_{L^{1}}, (2.10)

    with ρ\rho being as in (2.5).

Proof.

See Zheng [29]. ∎

2.1. Strichartz-Type Estimates

Before stating the Strichartz estimates, we need the following definitions.

We say the pair (q,r)(q,r) is Schrödinger admissible (S-admissible or L2L^{2}-admissible for short) if it satisfies

2q=N2Nr\frac{2}{q}=\frac{N}{2}-\frac{N}{r} (2.11)

where

{2r2NN2ifN3,2r<+ifN=1,2.\begin{cases}2\leq r\leq\frac{2N}{N-2}\;\;\textnormal{if}\;\;\;N\geq 3,\\ 2\leq r<+\infty\;\hskip 5.69046pt\textnormal{if}\;\;\;N=1,2.\end{cases} (2.12)

Also, given a real number s>0s>0, the pair (q,r)(q,r) is called H˙s\dot{H}^{s}-admissible if555It is worth mentioning that the pair (,2NN2sc)\left(\infty,\frac{2N}{N-2s_{c}}\right) also satisfies the relation (2.13), however, in our work we will not make use of this pair when we estimate the nonlinearity. See Section 5.

2q=N2Nrs\frac{2}{q}=\frac{N}{2}-\frac{N}{r}-s (2.13)

with

{2NN2sr<2NN2ifN3,2NN2sr<+ifN=1,2.\begin{cases}\frac{2N}{N-2s}\leq r<\frac{2N}{N-2}\;\textnormal{if}\;\;\;N\geq 3,\\ \frac{2N}{N-2s}\leq r<+\infty\;\hskip 2.84544pt\textnormal{if}\;\;\;N=1,2.\end{cases} (2.14)

We set666The restriction for (q,r)(q,r) H˙0\dot{H}^{0}-admissible is given by (2.12). 𝒜s:={(q,r);(q,r)isH˙s-admissible}\mathcal{A}_{s}:=\{(q,r);\;(q,r)\;\textnormal{is}\;\dot{H}^{s}\textnormal{-admissible}\}. Also, given (q,r)𝒜s(q,r)\in\mathcal{A}_{s}, by (q,r)(q^{\prime},r^{\prime}) we denote its dual pair, that is, 1q+1q=1\frac{1}{q}+\frac{1}{q^{\prime}}=1 and 1r+1r=1\frac{1}{r}+\frac{1}{r^{\prime}}=1. We define the Strichartz norm by

uS(H˙s)=sup(q,r)𝒜suLtqLxr\|u\|_{S(\dot{H}^{s})}=\sup_{(q,r)\in\mathcal{A}_{s}}\|u\|_{L^{q}_{t}L^{r}_{x}}

and the dual Strichartz norm by

uS(H˙s)=inf(q,r)𝒜suLtqLxr.\|u\|_{S^{\prime}(\dot{H}^{-s})}=\inf_{(q,r)\in\mathcal{A}_{-s}}\|u\|_{L^{q^{\prime}}_{t}L^{r^{\prime}}_{x}}.

If s=0s=0 then 𝒜0\mathcal{A}_{0} is the set of all SS-admissible pairs. We denote S(H˙0)S(\dot{H}^{0}) by S(L2)S(L^{2}). We write S(H˙s)S(\dot{H}^{s}) or S(H˙s)S^{\prime}(\dot{H}^{-s}) if the mixed norm is evaluated over ×N\mathbb{R}\times\mathbb{R}^{N}. To indicate the restriction to a time interval I(,)I\subset(-\infty,\infty) or a subset ANA\subset\mathbb{R}^{N}, we will use the notations S(H˙s(A);I)S(\dot{H}^{s}(A);I) and S(H˙s(A);I)S^{\prime}(\dot{H}^{-s}(A);I).

Finally, we define the norm777It was mentioned in the introduction, see (1.13). uS~(H˙sc)=uLta¯Lxr¯\|u\|_{\widetilde{S}(\dot{H}^{s_{c}})}=\|u\|_{L^{\bar{a}}_{t}L^{\bar{r}}_{x}}, where (a¯,r¯)\bar{a},\bar{r}) only satisfies r¯2\bar{r}\geq 2 and the relation (2.13) but not necessarily (2.14). Specifically, the number r¯\bar{r} does not need to satisfy the condition r¯<2NN2\bar{r}<\frac{2N}{N-2}.

We end this section by recalling the Strichartz estimates for the linear flow eitae^{-it\mathcal{L}_{a}}. They were first proved by Burq-Planchon-Stalker-Tahvildar-Zadeh888They showed Strichartz estimates for eitae^{-it\mathcal{L}_{a}} except the endpoint (q,r)=(2,2NN2)(q,r)=(2,\frac{2N}{N-2}). in [3]. Zhang-Zheng [28] confirmed the double endpoint case.

Lemma 2.5.

If N3N\geq 3 and a>(N2)24a>-\frac{(N-2)^{2}}{4}. Then,

eitafS(L2;I)\displaystyle\|e^{-it\mathcal{L}_{a}}f\|_{S(L^{2};I)} \displaystyle\lesssim fL2,\displaystyle\|f\|_{L^{2}}, (2.15)
eitafS(H˙s;I)\displaystyle\|e^{-it\mathcal{L}_{a}}f\|_{S(\dot{H}^{s};I)} \displaystyle\lesssim fH˙as\displaystyle\|f\|_{\dot{H}^{s}_{a}} (2.16)
t0tei(tt)ag(,t)𝑑tS(L2;I)\displaystyle\left\|\int_{t_{0}}^{t}e^{-i(t-t^{\prime})\mathcal{L}_{a}}g(\cdot,t^{\prime})dt^{\prime}\right\|_{S(L^{2};I)} \displaystyle\lesssim gS(L2;I)\displaystyle\|g\|_{S^{\prime}(L^{2};I)} (2.17)
Proof.

See [3] and [27]. ∎

Lemma 2.6.

If N3N\geq 3, a0a\geq 0 and gg a radial function. Then,

t0tei(tt)ag(,t)𝑑tS(H˙s;I)gS(H˙s;I),\left\|\int_{t_{0}}^{t}e^{-i(t-t^{\prime})\mathcal{L}_{a}}g(\cdot,t^{\prime})dt^{\prime}\right\|_{S(\dot{H}^{s};I)}\lesssim\|g\|_{S^{\prime}(\dot{H}^{-s};I)}, (2.18)

where II\subset\mathbb{R} be an interval and t0It_{0}\in I.

Proof.

The estimate is readily obtained from the main result in Foschi [13], given the L1LL^{1}\rightarrow L^{\infty} estimate in Lemma 2.4, and the invariance of the L2L^{2} norm by eitae^{-it\mathcal{L}_{a}}. ∎

Remark 2.7.

As usual, if I=(T,+)I=(T,+\infty) then in Lemma 2.6 one may replace the integral t0t\int_{t_{0}}^{t} by t+\int_{t}^{+\infty}. A similar statement holds if I=(,T)I=(-\infty,T).

Remark 2.8.

In the case, when s=0s=0, we have the norms uS(L2)=sup(q,r)𝒜0uLtqLxr\|u\|_{S(L^{2})}=\sup_{(q,r)\in\mathcal{A}_{0}}\|u\|_{L^{q}_{t}L^{r}_{x}} and uS(L2)=inf(q,r)𝒜0uLtqLxr\|u\|_{S^{\prime}(L^{2})}=\inf_{(q,r)\in\mathcal{A}_{0}}\|u\|_{L^{q^{\prime}}_{t}L^{r^{\prime}}_{x}}.

3. Global well-posedness and blow-up in Ha1H^{1}_{a}

In this section, we prove results about the ground state QQ, together with a dichotomy between global existence and finite-time blow-up below a mass-energy threshold. We first show the existence of a ground state (Proposition 1.1).

3.1. Existence of a ground state

We start by proving a compact embedding result.

Lemma 3.1 (Compactness of an immersion).

If N3N\geq 3, a>(N2)24a>-\tfrac{(N-2)^{2}}{4}, 0<b<20<b<2 and 0<α<42bN20<\alpha<\tfrac{4-2b}{N-2}, then Ha1(N)H^{1}_{a}(\mathbb{R}^{N}) is compactly embedded in Lbα+2(N)L^{\alpha+2}_{b}(\mathbb{R}^{N}).

Proof.

Let {fn}n\{f_{n}\}_{n} be a bounded sequence in Ha1H^{1}_{a}. Since a>(N2)24a>-\frac{(N-2)^{2}}{4}, Hardy’s inequality yields that {fn}n\{f_{n}\}_{n} is also bounded in the standard space H1H^{1}, so we may assume that fnfH1f_{n}\rightharpoonup f\in H^{1} weakly in H1H^{1}. Now, for R,ϵ>0R,\epsilon>0,

|x|b|fnf|α+2𝑑x(|x|R|x|bNb+ϵ𝑑x)b+ϵNfnfLN(α+2)Nbϵα+2+1Rb(fnLα+2+fLα+2)α+2.\int|x|^{-b}|f_{n}-f|^{\alpha+2}dx\lesssim\left(\int_{|x|\leq R}|x|^{-\frac{bN}{b+\epsilon}}dx\right)^{\frac{b+\epsilon}{N}}\|f_{n}-f\|_{L^{\frac{N(\alpha+2)}{N-b-\epsilon}}}^{\alpha+2}+\frac{1}{R^{b}}(\|f_{n}\|_{L^{\alpha+2}}+\|f\|_{L^{\alpha+2}})^{\alpha+2}. (3.1)

Thus, by Sobolev and Rellich-Kondrachov, since

2<α+2<N(α+2)Nbϵ<2NN2,2<\alpha+2<\frac{N(\alpha+2)}{N-b-\epsilon}<\frac{2N}{N-2}, (3.2)

if 0<α<42bN20<\alpha<\tfrac{4-2b}{N-2} and ϵ\epsilon is small, the result follows.∎

Lemma 3.2 (Adapted Gagliardo-Nirenberg inequality).

Let N3N\geq 3. If fHa1f\in H^{1}_{a}, 0<b<20<b<2 and 0<α<42bN20<\alpha<\frac{4-2b}{N-2}, then

|x|b|f|α+2𝑑xCaafL2Nα+2b2fL242bα(N2)2.\int|x|^{-b}|f|^{\alpha+2}\,dx\leq C_{a}\|\sqrt{\mathcal{L}_{a}}f\|_{L^{2}}^{\tfrac{N\alpha+2b}{2}}\|f\|_{L^{2}}^{\tfrac{4-2b-\alpha(N-2)}{2}}. (3.3)

Equality in the bound above is attained by a function QHa1Q\in H^{1}_{a}, which is a positive solution to the elliptic equation

aQQ+|x|b|Q|αQ=0.\mathcal{L}_{a}Q-Q+|x|^{-b}|Q|^{\alpha}Q=0. (3.4)
Proof.

We mimic the classic proof for a=b=0a=b=0, and exploit the compactness given by the immersion Ha1Lbα+2H^{1}_{a}\hookrightarrow L^{\alpha+2}_{b}. Let {fn}nHa1\{f_{n}\}_{n}\subset H^{1}_{a} be a minimizing sequence for the Weinstein functional

Ja(f)=afL2Nα+2b2fL242bα(N2)2|x|b|f|α+2𝑑x.J_{a}(f)=\frac{\|\displaystyle\sqrt{\mathcal{L}_{a}}f\|_{L^{2}}^{\tfrac{N\alpha+2b}{2}}\|f\|_{L^{2}}^{\tfrac{4-2b-\alpha(N-2)}{2}}}{\displaystyle\int|x|^{-b}|f|^{\alpha+2}\,dx}. (3.5)

By choosing λn>0\lambda_{n}>0 and μn>0\mu_{n}>0 such that gn(x)=λnfn(μnx)g_{n}(x)=\lambda_{n}f_{n}(\mu_{n}x) satisfy agnL2=gnL2=1\|\sqrt{\mathcal{L}_{a}}g_{n}\|_{L^{2}}=\|g_{n}\|_{L^{2}}=1, and noting that Ja(fn)=Ja(gn)J_{a}(f_{n})=J_{a}(g_{n}) for all nn, we may assume, by compactness, that {gn}n\{g_{n}\}_{n} converges strongly to gg in Lbα+2L^{\alpha+2}_{b} (see Lemma 3.1) and by reflexiveness, weakly in Ha1H^{1}_{a}. Moreover, since ||f|||f||\nabla|f||\leq|\nabla f| for all fHa1f\in H^{1}_{a}, we can assume that g0g\geq 0. Furthermore, by Hölder, Hardy and Sobolev, we see that Ja=inffHa1\{0}J(f)>0J^{*}_{a}=\displaystyle\inf_{f\in H^{1}_{a}\backslash\{0\}}J(f)>0, and since {gn}n\{g_{n}\}_{n} is a minimizing sequence,

|x|b|g(x)|α+2𝑑x=1Ja(0,+).\int|x|^{-b}|g(x)|^{\alpha+2}\,dx=\frac{1}{J^{*}_{a}}\in(0,+\infty). (3.6)

This shows that g0g\neq 0. Now note that agL2=gL2=1\|\sqrt{\mathcal{L}_{a}}g\|_{L}^{2}=\|g\|_{L}^{2}=1, since otherwise it would contradict the minimality of Ja(g)=JaJ_{a}(g)=J^{*}_{a}. Defining Ca:=(Ja)1C_{a}:=(J_{a}^{*})^{-1}, one sees that (3.3) holds.

The Euler-Lagrange equation for gg gives

JaNα+2b2ag+Ja[42bα(N2)2]g(α+2)gα+1=0.J_{a}^{*}\tfrac{N\alpha+2b}{2}\mathcal{L}_{a}g+J_{a}^{*}[\tfrac{4-2b-\alpha(N-2)}{2}]g-(\alpha+2)g^{\alpha+1}=0. (3.7)

Finally, defining g(x)=λQ(μx)g(x)=\lambda Q(\mu x), with

λ={[42bα(N2)]Ja2(α+2)}1α,\lambda=\left\{\tfrac{[4-2b-\alpha(N-2)]J^{*}_{a}}{2(\alpha+2)}\right\}^{\frac{1}{\alpha}}, (3.8)

and

μ=42bα(N2)Nα+2b,\mu=\tfrac{4-2b-\alpha(N-2)}{N\alpha+2b}, (3.9)

one sees that QQ solves (3.4). ∎

In the following lemma we obtain Pohozaev-type identities which are satisfied by any solution of (3.4). The proof follows multiplying (3.4) by Q and xQx\cdot\nabla Q and using integration by parts. We omit the details.

Lemma 3.3 (Pohozaev identities).

If QHa1Q\in H^{1}_{a} is a solution to (3.4), then the following identities hold

|aQ|2𝑑x=(Nα+2b42bα(N2))QL22,\displaystyle\int\left|\sqrt{\mathcal{L}_{a}}Q\right|^{2}\,dx=\left(\tfrac{N\alpha+2b}{4-2b-\alpha(N-2)}\right)\|Q\|_{L^{2}}^{2}, (3.10)
|x|b|Q|α+2𝑑x=(2(α+2)42bα(N2))QL22.\displaystyle\int|x|^{-b}|Q|^{\alpha+2}\,dx=\left(\tfrac{2(\alpha+2)}{4-2b-\alpha(N-2)}\right)\|Q\|_{L^{2}}^{2}. (3.11)

In particular, one can write the mass (and therefore the quantities in (3.10) and (3.11)) in terms of the sharp constant CaC_{a}. Namely,

M[Q]={2(α+2)Nα+2b[42bα(N2)]Nα(42b)4Ca}1α+2.M[Q]=\left\{\tfrac{2(\alpha+2)}{N\alpha+2b}\tfrac{[4-2b-\alpha(N-2)]^{\tfrac{N\alpha-(4-2b)}{4}}}{C_{a}}\right\}^{\frac{1}{\alpha+2}}. (3.12)

With the previous results we show Proposition 1.1.

Proof of Proposition 1.1.

The existence part of Proposition 1.1 follows from Lemma 3.2, and the uniqueness of the mass, H˙a1\dot{H}^{1}_{a} and energy follows from Lemma 3.3. ∎

3.2. Global behavior in the mass-critical case

We now study the global existence and blow-up in Ha1H^{1}_{a} of (1.1), when α=42bN\alpha=\tfrac{4-2b}{N} and λ=1\lambda=1. We start by proving the global well-posedness.

3.2.1. Global well-posedness

Proof of Theorem 1.2a).

We make use here of the sharp Gagliardo-Nirenberg-type inequality (Lemma 3.2) in the case α=42bN\alpha=\tfrac{4-2b}{N}:

|x|b|f|42bN+2𝑑x22b+NNafL22(fL2QL2)42bN.\int|x|^{-b}|f|^{\frac{4-2b}{N}+2}\,dx\leq\tfrac{2-2b+N}{N}\|\sqrt{\mathcal{L}_{a}}f\|_{L^{2}}^{2}\left(\frac{\|f\|_{L^{2}}}{\|Q\|_{L^{2}}}\right)^{\tfrac{4-2b}{N}}. (3.13)

By energy conservation:

E[u0]\displaystyle E[u_{0}] =12au(t)L22N42b+2N|x|b|u(t)|42bN+2𝑑x\displaystyle=\frac{1}{2}\|\sqrt{\mathcal{L}_{a}}u(t)\|_{L^{2}}^{2}-\frac{N}{4-2b+2N}\int|x|^{-b}|u(t)|^{\frac{4-2b}{N}+2}\,dx (3.14)
12au(t)L22[1(u0L2QL2)42bN].\displaystyle\geq\frac{1}{2}\|\sqrt{\mathcal{L}_{a}}u(t)\|_{L^{2}}^{2}\left[1-\left(\frac{\|u_{0}\|_{L^{2}}}{\|Q\|_{L^{2}}}\right)^{\tfrac{4-2b}{N}}\right]. (3.15)

Therefore, if u0L2<QL2\|u_{0}\|_{L^{2}}<\|Q\|_{L^{2}}, the corresponding solution uu extends globally in time. ∎

3.2.2. Blow-up

The main tool to prove the blow-up results is the Virial identity.

Lemma 3.4 (Virial identity).

Let φ:N\varphi:\mathbb{R}^{N}\rightarrow\mathbb{R} be a real weight. Define

V(t)=φ|u(t)|2𝑑x.V(t)=\int\varphi|u(t)|^{2}dx. (3.16)

Then, if uu is a solution to (1.4), we have the following identities

V(t)=2Imu¯uφdx,V^{\prime}(t)=2\operatorname{Im}\int\bar{u}\nabla u\cdot\nabla\varphi\,dx,
V′′(t)\displaystyle V^{\prime\prime}(t) =(4α+22)|x|b|u|α+2Δφ𝑑x4bα+2|x|b2|u|α+2xφdx\displaystyle=\left(\frac{4}{\alpha+2}-2\right)\int|x|^{-b}|u|^{\alpha+2}\Delta\varphi dx-\frac{4b}{\alpha+2}\int|x|^{-b-2}|u|^{\alpha+2}x\cdot\nabla\varphi dx (3.17)
+4Rei,jφiju¯iuj𝑑x+4axφ|x|4|u|2𝑑x\displaystyle\quad+4\operatorname{Re}\sum_{i,j}\int\varphi_{ij}\bar{u}_{i}u_{j}dx+4a\int\frac{x\cdot\nabla\varphi}{|x|^{4}}|u|^{2}dx (3.18)
|u|2ΔΔφ𝑑x.\displaystyle\quad-\int|u|^{2}\Delta\Delta\varphi dx. (3.19)
Proof.

The proof is considered standard, see for instance [12, 21]. ∎

Proof of Theorem 1.2b).

The blow-up for fast-decaying, negative-energy solutions is proved using Glassey’s argument, with the Virial identity by taking ϕ=|x|2\phi=|x|^{2}. It means,

tt2|x|2|u(t)|2𝑑x=16E[u0]<0.\partial^{2}_{tt}\int|x|^{2}|u(t)|^{2}dx=16E[u_{0}]<0. (3.20)

We now show for radial solutions. Define ψ:(0,+)\psi:(0,+\infty)\to\mathbb{R} as

ψ(r)={r2,r1,r2(r1)42,1<r272,r>2.\psi(r)=\begin{cases}r^{2},&r\leq 1,\\ r^{2}-\frac{(r-1)^{4}}{2},&1<r\leq 2\\ \frac{7}{2},&r>2.\end{cases} (3.21)

Note that ψ\psi and all of its (weak) derivatives are essentially bounded. Define also, for R>0R>0, ϕR(x)=R2ψ(x/R)\phi_{R}(x)=R^{2}\psi(x/R) and

VR(t)=ϕR|u(t)|2𝑑x.V_{R}(t)=\int\phi_{R}|u(t)|^{2}dx. (3.22)

By Lemma 3.4 and the radiality of uu, in the case α=42bN\alpha=\tfrac{4-2b}{N}, we have

VR′′(t)\displaystyle V_{R}^{\prime\prime}(t) =16E[u(t)]+42b2b+N(2NΔϕR)|x|b|u|42bN+2𝑑x\displaystyle=16E[u(t)]+\frac{4-2b}{2-b+N}\int(2N-\Delta\phi_{R})|x|^{-b}|u|^{\tfrac{4-2b}{N}+2}dx (3.23)
+2Nb2b+N(2xϕR|x|2)|x|b|u|42bN+2𝑑x\displaystyle\quad+\frac{2Nb}{2-b+N}\int(2-\frac{x\cdot\nabla\phi_{R}}{|x|^{2}})|x|^{-b}|u|^{\tfrac{4-2b}{N}+2}dx (3.24)
4(2ψ′′(|x|/R))|u|2𝑑x4a(2xϕR|x|2)1|x|2|u|2𝑑xΔΔϕR|u|2𝑑x\displaystyle\quad-4\int(2-\psi^{\prime\prime}(|x|/R))|\nabla u|^{2}dx-4a\int(2-\frac{x\cdot\nabla\phi_{R}}{|x|^{2}})\frac{1}{|x|^{2}}|u|^{2}dx-\int\Delta\Delta\phi_{R}|u|^{2}dx (3.25)
16E[u0]4(2ψ′′(|x|/R))|u|2𝑑x+O(1Rb|x|R[ηR(x)|u|]42bN|u|2𝑑x)+O(1R2|u|2𝑑x),\displaystyle\leq 16E[u_{0}]-4\int(2-\psi^{\prime\prime}(|x|/R))|\nabla u|^{2}dx+O(\frac{1}{R^{b}}\int_{|x|\geq R}[\eta_{R}(x)|u|]^{\tfrac{4-2b}{N}}|u|^{2}dx)+O(\frac{1}{R^{2}}\int|u|^{2}dx), (3.26)

where

ηR(x)={12b+N[(42b)(2NΔϕR(x))+2Nb(2xϕR(x)|x|2)]}N42b.\eta_{R}(x)=\left\{\frac{1}{2-b+N}\left[(4-2b)(2N-\Delta\phi_{R}(x))+2Nb\left(2-\frac{x\cdot\nabla\phi_{R}(x)}{|x|^{2}}\right)\right]\right\}^{\frac{N}{4-2b}}. (3.27)

By mass conservation, we are left to control the third term in the last inequality, if R>0R>0 is large enough. We use Strauss and Young inequalities:

1Rb|x|R[ηR(x)|u|]42bN|u|2𝑑x\displaystyle\frac{1}{R^{b}}\int_{|x|\geq R}[\eta_{R}(x)|u|]^{\tfrac{4-2b}{N}}|u|^{2}dx 1RbηRuL{|x|R}42bN|u|2𝑑x\displaystyle\lesssim\frac{1}{R^{b}}\|\eta_{R}u\|_{L^{\infty}_{\{|x|\geq R\}}}^{\tfrac{4-2b}{N}}\int|u|^{2}dx (3.28)
1R(2b)(N1)+NbN((ηRu)L22bNuL22bN)|u|2𝑑x\displaystyle\lesssim\frac{1}{R^{\frac{(2-b)(N-1)+Nb}{N}}}(\|\nabla(\eta_{R}u)\|_{L^{2}}^{\frac{2-b}{N}}\|u\|^{\frac{2-b}{N}}_{L^{2}})\int|u|^{2}dx (3.29)
ϵηRuL22+C(ϵ)R2uL22.\displaystyle\leq\epsilon\|\eta_{R}\nabla u\|_{L^{2}}^{2}+\frac{C(\epsilon)}{R^{2}}\|u\|_{L^{2}}^{2}. (3.30)

We therefore obtain

VR′′(t)16E[u0][4(2ψ′′(|x|/R))ϵηR2(x))]|u|2+O(C(ϵ)R2).V^{\prime\prime}_{R}(t)\leq 16E[u_{0}]-\int\left[4(2-\psi^{\prime\prime}(|x|/R))-\epsilon\eta_{R}^{2}(x))\right]|\nabla u|^{2}+O(\frac{C(\epsilon)}{R^{2}}). (3.31)

By the definition of ψ\psi and ηR\eta_{R}, one can choose ϵ>0\epsilon>0 such that, almost everywhere and independently on RR,

4(2ψ′′(/R))ϵηR2)0.4(2-\psi^{\prime\prime}(\cdot/R))-\epsilon\eta_{R}^{2})\geq 0. (3.32)

Choosing, afterwards, R>0R>0 large enough, we have,

VR′′(t)15E[u0]<0,V^{\prime\prime}_{R}(t)\leq 15E[u_{0}]<0, (3.33)

which implies finite-time blow-up.

3.3. Global behavior for the intercritical case

We now state some coercivity-type (also known as energy-trapping) results for the INLSa, which are necessary for Theorem 1.3.

Lemma 3.5.

Let N3N\geq 3, aa, α\alpha and bb as in Theorem 1.3, and fHa1(N)f\in H^{1}_{a}(\mathbb{R}^{N}). Assume that, for some δ0>0\delta_{0}>0,

M(f)1scscEa(f)(1δ0)M(Q)1scscEa(Q),M(f)^{\tfrac{1-s_{c}}{s_{c}}}E_{a}(f)\leq(1-\delta_{0})M(Q)^{\tfrac{1-s_{c}}{s_{c}}}E_{a}(Q), (3.34)

then there exists δ=δ(δ0,N,α,Q,a,b)\delta=\delta(\delta_{0},N,\alpha,Q,a,b) such that

|fL21scscafL2QL21scscaQL2|δQL21scscaQL2.\left|\|f\|^{\frac{1-s_{c}}{s_{c}}}_{L^{2}}\|\sqrt{\mathcal{L}_{a}}f\|_{L^{2}}-\|Q\|^{\frac{1-s_{c}}{s_{c}}}_{L^{2}}\|\sqrt{\mathcal{L}_{a}}Q\|_{L^{2}}\right|\geq\delta\|Q\|^{\frac{1-s_{c}}{s_{c}}}_{L^{2}}\|\sqrt{\mathcal{L}_{a}}Q\|_{L^{2}}. (3.35)

In particular, if

fL21scscafL2<QL21scscaQL2,\|f\|^{\frac{1-s_{c}}{s_{c}}}_{L^{2}}\|\sqrt{\mathcal{L}_{a}}f\|_{L^{2}}<\|Q\|^{\frac{1-s_{c}}{s_{c}}}_{L^{2}}\|\sqrt{\mathcal{L}_{a}}Q\|_{L^{2}}, (3.36)

then

fL21scscafL2(1δ)QL21scscaQL2.\|f\|^{\frac{1-s_{c}}{s_{c}}}_{L^{2}}\|\sqrt{\mathcal{L}_{a}}f\|_{L^{2}}\leq(1-\delta)\|Q\|^{\frac{1-s_{c}}{s_{c}}}_{L^{2}}\|\sqrt{\mathcal{L}_{a}}Q\|_{L^{2}}. (3.37)

Similarly, if

fL21scscafL2>QL21scscaQL2,\|f\|^{\frac{1-s_{c}}{s_{c}}}_{L^{2}}\|\sqrt{\mathcal{L}_{a}}f\|_{L^{2}}>\|Q\|^{\frac{1-s_{c}}{s_{c}}}_{L^{2}}\|\sqrt{\mathcal{L}_{a}}Q\|_{L^{2}}, (3.38)

then

fL21scscafL2(1+δ)QL21scscaQL2.\|f\|^{\frac{1-s_{c}}{s_{c}}}_{L^{2}}\|\sqrt{\mathcal{L}_{a}}f\|_{L^{2}}\geq(1+\delta)\|Q\|^{\frac{1-s_{c}}{s_{c}}}_{L^{2}}\|\sqrt{\mathcal{L}_{a}}Q\|_{L^{2}}. (3.39)
Proof.

Using Lemma 3.2, we write

(1δ0)M[Q]1scscEa[Q]\displaystyle(1-\delta_{0})M[Q]^{\frac{1-s_{c}}{s_{c}}}E_{a}[Q] M[f]1scscEa[f]\displaystyle\geq M[f]^{\frac{1-s_{c}}{s_{c}}}E_{a}[f] (3.40)
=M[f]1scsc2|af|2𝑑xM[f]1scscα+2|x|b|f|α+2𝑑x\displaystyle=\frac{M[f]^{\frac{1-s_{c}}{s_{c}}}}{2}\int\left|\sqrt{\mathcal{L}_{a}}f\right|^{2}dx-\frac{M[f]^{\frac{1-s_{c}}{s_{c}}}}{\alpha+2}\int|x|^{-b}|f|^{\alpha+2}dx (3.41)
12(fL21scscafL2)2Caα+2(fL21scscafL2)Nα+2b2.\displaystyle\geq\frac{1}{2}\left(\|f\|_{L^{2}}^{\frac{1-s_{c}}{s_{c}}}\|\sqrt{\mathcal{L}_{a}}f\|_{L^{2}}\right)^{2}-\frac{C_{a}}{\alpha+2}\left(\|f\|_{L^{2}}^{\frac{1-s_{c}}{s_{c}}}\|\sqrt{\mathcal{L}_{a}}f\|_{L^{2}}\right)^{\frac{N\alpha+2b}{2}}. (3.42)

Let P(y)=12y2Caα+2yNα+2b2P(y)=\frac{1}{2}y^{2}-\frac{C_{a}}{\alpha+2}y^{\frac{N\alpha+2b}{2}} and y(f)=fL21scscafL2y(f)=\|f\|_{L^{2}}^{\frac{1-s_{c}}{s_{c}}}\|\sqrt{\mathcal{L}_{a}}f\|_{L^{2}}. By Lemmas 3.2 and 3.3, we see that, if f=Qf=Q, then y:=y(Q)y^{*}:=y(Q) satisfies P(y)=M[Q]1scscEa[Q]P(y^{*})=M[Q]^{\frac{1-s_{c}}{s_{c}}}E_{a}[Q]. Since PP is continuous, increasing if y<yy<y^{*}, and decreasing if y>yy>y^{*}, we conclude that there exists δ0>0\delta_{0}>0 such that.

P(y)(1δ0)P(y)|yy|δy.P(y)\leq(1-\delta_{0})P(y^{*})\implies|y-y^{*}|\geq\delta y^{*}. (3.44)

By continuity, we conclude (3.35). ∎

Lemma 3.6.

Under the conditions (3.35) and (3.38) of the previous lemma, one also has, for some η=η(δ0,Q,M(f))>0\eta=\eta(\delta_{0},Q,M(f))>0 and ϵ=ϵ(δ0,Q)>0\epsilon=\epsilon(\delta_{0},Q)>0,

(1+ϵ)|af|2𝑑x+(Nbα+2N2)|x|b|f|α+2𝑑xη<0(1+\epsilon)\int\left|\sqrt{\mathcal{L}_{a}}f\right|^{2}dx+\left(\frac{N-b}{\alpha+2}-\frac{N}{2}\right)\int|x|^{-b}|f|^{\alpha+2}dx\leq-\eta<0 (3.45)
Proof.

Under the notation of the proof of the previous lemma, if

M(f)1scscEa(f)(1δ0)M(Q)1scscEa(Q)M(f)^{\tfrac{1-s_{c}}{s_{c}}}E_{a}(f)\leq(1-\delta_{0})M(Q)^{\tfrac{1-s_{c}}{s_{c}}}E_{a}(Q)

and

fL21scscafL2QL21scscaQL2,\|f\|^{\frac{1-s_{c}}{s_{c}}}_{L^{2}}\|\sqrt{\mathcal{L}_{a}}f\|_{L^{2}}\geq\|Q\|^{\frac{1-s_{c}}{s_{c}}}_{L^{2}}\|\sqrt{\mathcal{L}_{a}}Q\|_{L^{2}}, (3.46)

we write

(1+ϵ)|af|2+\displaystyle(1+\epsilon)\int\left|\sqrt{\mathcal{L}_{a}}f\right|^{2}+ (Nbα+2N2)|x|b|f|α+2\displaystyle\left(\frac{N-b}{\alpha+2}-\frac{N}{2}\right)\int|x|^{-b}|f|^{\alpha+2} (3.47)
=Nα+2b2Ea[f][N2(α42bN)ϵ]|af|2\displaystyle=\frac{N\alpha+2b}{2}E_{a}[f]-\left[\frac{N}{2}\left(\alpha-\frac{4-2b}{N}\right)-\epsilon\right]\int\left|\sqrt{\mathcal{L}_{a}}f\right|^{2} (3.48)
(1δ0)Nα+2b2M[f]1scscM[Q]1scscEa[Q]\displaystyle\leq(1-\delta_{0})\frac{N\alpha+2b}{2}M[f]^{-\frac{1-s_{c}}{s_{c}}}M[Q]^{\frac{1-s_{c}}{s_{c}}}E_{a}[Q] (3.49)
[N2(α42bN)ϵ]M[f]1scsc(y)2\displaystyle\quad\quad\quad\quad\quad-\left[\frac{N}{2}\left(\alpha-\frac{4-2b}{N}\right)-\epsilon\right]M[f]^{-\frac{1-s_{c}}{s_{c}}}(y^{*})^{2} (3.50)
=M[f]1scscM[Q]1scsc(|aQ|2+(Nbα+2N2)|x|b|Q|α+2)=0\displaystyle=M[f]^{-\frac{1-s_{c}}{s_{c}}}M[Q]^{\frac{1-s_{c}}{s_{c}}}\underbrace{\left(\int\left|\sqrt{\mathcal{L}_{a}}Q\right|^{2}+\left(\frac{N-b}{\alpha+2}-\frac{N}{2}\right)\int|x|^{-b}|Q|^{\alpha+2}\right)}_{=0} (3.51)
M[f]1scsc(Nα+2b2δ0M[Q]1scscEa[Q]ϵ(y)2).\displaystyle\quad\quad\quad\quad\quad-M[f]^{-\frac{1-s_{c}}{s_{c}}}\left(\frac{N\alpha+2b}{2}\delta_{0}M[Q]^{\frac{1-s_{c}}{s_{c}}}E_{a}[Q]-\epsilon(y^{*})^{2}\right). (3.52)

Hence, by taking, say,

η=M[f]1scscδ0M[Q]1scscEa[Q]Nα+2b4>0,\eta=M[f]^{-\frac{1-s_{c}}{s_{c}}}\delta_{0}M[Q]^{\frac{1-s_{c}}{s_{c}}}E_{a}[Q]\frac{N\alpha+2b}{4}>0,

the lemma is proved. ∎

An immediate consequence of Lemma 3.5 and mass and energy conservation we obtain the following.

Lemma 3.7.

Let N3N\geq 3, aa, α\alpha and bb as in Theorem 1.3, and u0H1(N)u_{0}\in H^{1}(\mathbb{R}^{N}). Denote by II the maximal time of existence of the corresponding solution uu to (1.1). If u0u_{0} satisfies (3.35) and (3.36), then

suptIu0L21scscau(t)L2(1δ)QL21scscaQL2.\sup_{t\in I}\|u_{0}\|^{\frac{1-s_{c}}{s_{c}}}_{L^{2}}\|\sqrt{\mathcal{L}_{a}}u(t)\|_{L^{2}}\leq(1-\delta)\|Q\|^{\frac{1-s_{c}}{s_{c}}}_{L^{2}}\|\sqrt{\mathcal{L}_{a}}Q\|_{L^{2}}. (3.53)

In particular, suptIu(t)Ha1<+\sup_{t\in I}\|u(t)\|_{H^{1}_{a}}<+\infty, which implies I=(,+)I=(-\infty,+\infty). Alternatively, if u0u_{0} satisfies (3.35) and (3.38), then for all tIt\in I,

u0L21scscau(t)L2(1+δ)QL21scscaQL2.\|u_{0}\|^{\frac{1-s_{c}}{s_{c}}}_{L^{2}}\|\sqrt{\mathcal{L}_{a}}u(t)\|_{L^{2}}\geq(1+\delta)\|Q\|^{\frac{1-s_{c}}{s_{c}}}_{L^{2}}\|\sqrt{\mathcal{L}_{a}}Q\|_{L^{2}}. (3.54)
Proof of Theorem 1.3.

Theorem 1.3a) follows directly from Lemma 3.7.

Now we prove Theorem 1.3b). Given Lemmas 3.6 and 3.7, the result for |x|u0L2|x|u_{0}\in L^{2} follows from Lemma 3.4, which gives, in this context:

tt2|x|2|u(t)|2𝑑x=8[|af|2𝑑x+(Nbα+2N2)|x|b|f|α+2𝑑x].\partial^{2}_{tt}\int|x|^{2}|u(t)|^{2}dx=8\left[\int\left|\sqrt{\mathcal{L}_{a}}f\right|^{2}dx+\left(\frac{N-b}{\alpha+2}-\frac{N}{2}\right)\int|x|^{-b}|f|^{\alpha+2}dx\right]. (3.55)

By (3.45), we then have, for all tt in the interior of the maximal interval of existence of uu,

tt2|x|2|u(t)|2𝑑x8η<0,\partial^{2}_{tt}\int|x|^{2}|u(t)|^{2}dx\leq-8\eta<0, (3.56)

which shows that uu blows up in both finite positive and negative times.

To prove blow-up in the radial case, we employ Lemma 3.4 again. Define ψ:(0,+)\psi:(0,+\infty)\to\mathbb{R} as a smooth function satisfying

ψ(r)={r2,r1,4,r2\psi(r)=\begin{cases}r^{2},&r\leq 1,\\ 4,&r\geq 2\end{cases} (3.57)

and we also impose that ψ′′(r)2\psi^{\prime\prime}(r)\leq 2 for all r>0r>0. Define also, for R>0R>0, ϕR(x)=R2ψ(x/R)\phi_{R}(x)=R^{2}\psi(x/R) and

VR(t)=ϕR|u(t)|2𝑑x.V_{R}(t)=\int\phi_{R}|u(t)|^{2}dx. (3.58)

By Lemma 3.4, the non-negativity of ψ′′\psi^{\prime\prime} and the radiality of uu,

VR′′(t)\displaystyle V_{R}^{\prime\prime}(t) =8|af|2𝑑x+8(Nbα+2N2)|x|b|f|α+2𝑑x\displaystyle=8\int\left|\sqrt{\mathcal{L}_{a}}f\right|^{2}dx+8\left(\frac{N-b}{\alpha+2}-\frac{N}{2}\right)\int|x|^{-b}|f|^{\alpha+2}dx (3.59)
(4α+22)R|x|2RΔϕR|x|b|u|α+2𝑑x+4bα+2R|x|2RxϕR|x|b2|u|α+2dx\displaystyle-\left(\frac{4}{\alpha+2}-2\right)\int_{R\leq|x|\leq 2R}\Delta\phi_{R}|x|^{-b}|u|^{\alpha+2}dx+\frac{4b}{\alpha+2}\int_{R\leq|x|\leq 2R}x\cdot\nabla\phi_{R}|x|^{-b-2}|u|^{\alpha+2}dx (3.60)
4R|x|2Rψ′′(|x|/R)|u|2𝑑x4aR|x|2RxϕR|x|4|u|2ΔΔϕR|u|2𝑑x\displaystyle-4\int_{R\leq|x|\leq 2R}\psi^{\prime\prime}(|x|/R)|\nabla u|^{2}dx-4a\int_{R\leq|x|\leq 2R}\frac{x\cdot\nabla\phi_{R}}{|x|^{4}}|u|^{2}-\int\Delta\Delta\phi_{R}|u|^{2}dx (3.61)
8[|af|2𝑑x+(Nbα+2N2)|x|b|f|α+2𝑑x]+O(1Rb|x|R|u|α+2)+O(1R2|u|2).\displaystyle\leq 8\left[\int\left|\sqrt{\mathcal{L}_{a}}f\right|^{2}dx+\left(\frac{N-b}{\alpha+2}-\frac{N}{2}\right)\int|x|^{-b}|f|^{\alpha+2}dx\right]+O(\frac{1}{R^{b}}\int_{|x|\geq R}|u|^{\alpha+2})+O(\frac{1}{R^{2}}\int|u|^{2}). (3.62)

By the coercivity lemmas and mass conservation, we are left to control the middle term in the last inequality, if R>0R>0 is large enough. We use Strauss and Young inequalities:

1Rb|x|R|u|α+2𝑑x\displaystyle\frac{1}{R^{b}}\int_{|x|\geq R}|u|^{\alpha+2}dx 1RbuL{|x|R}α|u|2𝑑x1Rα(N1)+2b2(uL2α2uL2α2)|u|2𝑑x\displaystyle\lesssim\frac{1}{R^{b}}\|u\|_{L^{\infty}_{\{|x|\geq R\}}}^{\alpha}\int|u|^{2}dx\lesssim\frac{1}{R^{\frac{\alpha(N-1)+2b}{2}}}(\|\nabla u\|_{L^{2}}^{\frac{\alpha}{2}}\|u\|^{\frac{\alpha}{2}}_{L^{2}})\int|u|^{2}dx (3.63)
ϵuL22+C(ϵ)(1Rα(N1)+2b2uL22+α2dx)44α.\displaystyle\lesssim\epsilon\|\nabla u\|_{L^{2}}^{2}+C(\epsilon)\left(\frac{1}{R^{\frac{\alpha(N-1)+2b}{2}}}\|u\|_{L^{2}}^{2+\frac{\alpha}{2}}dx\right)^{\frac{4}{4-\alpha}}. (3.64)

Thus, by Lemma 3.6, by choosing R>0R>0 depending only on the mass of uu, we have, for all times

VR′′(t)7η<0,V_{R}^{\prime\prime}(t)\leq-7\eta<0, (3.65)

which implies blowup in finite positive and negative times. ∎

4. Well-posedness theory via Kato’s method

In this section we prove the well-posedness results using the Kato method. The proofs follow from a contraction mapping argument based on the Strichartz estimates. In view of the singular factor |x|b|x|^{-b} in the nonlinearity, we will divide our analysis in two regions. Indeed, consider a unit ball, B={xN;|x|1}B=\{x\in\mathbb{R}^{N};|x|\leq 1\}. A simple computation reveals that

|x|bLγ(B)<+ifNγb>0and|x|bLγ(BC)<+ifNγb<0.\||x|^{-b}\|_{L^{\gamma}(B)}<+\infty\quad\textnormal{if}\quad\frac{N}{\gamma}-b>0\quad\textnormal{and}\quad\||x|^{-b}\|_{L^{\gamma}(B^{C})}<+\infty\quad\textnormal{if}\quad\frac{N}{\gamma}-b<0. (4.1)

Moreover, if F(x,z)=|x|b|z|αzF(x,z)=|x|^{-b}|z|^{\alpha}z, then

|F(x,z)F(x,w)||x|b(|z|α+|w|α)|zw|.|F(x,z)-F(x,w)|\lesssim|x|^{-b}(|z|^{\alpha}+|w|^{\alpha})|z-w|. (4.2)

Before stating the lemmas, we define the norm

auS(L2;J)=uS(L2;J)+auS(L2;J),\|\langle\sqrt{\mathcal{L}_{a}}\rangle u\|_{S(L^{2};J)}=\|u\|_{S(L^{2};J)}+\|\sqrt{\mathcal{L}_{a}}u\|_{S(L^{2};J)}, (4.3)

where JJ\subseteq\mathbb{R}. For this norms, it is worth mentioning the following:

Remark 4.1.

Since we use the equivalence of Sobolev spaces in auS(L2)\|\sqrt{\mathcal{L}_{a}}u\|_{S(L^{2})}, for these terms we only employ SS-admissible pairs which satisfy the conditions of Remark 2.2, that is, 1<r<Ns1<r<\frac{N}{s} if a>0a>0 and NNρ<r<Ns+ρ\frac{N}{N-\rho}<r<\frac{N}{s+\rho} if a<0a<0. We do the same with auS(L2)\|\sqrt{\mathcal{L}_{a}}u\|_{S^{\prime}(L^{2})}.

4.1. Local Theory

First, we establish good estimates for the nonlinearity in the Strichartz spaces.

Lemma 4.2.

Let N3N\geq 3, 0<b<min{N2,2}0<b<\min\{\frac{N}{2},2\} and a>(N2)24+(α(N2)(22b)2(α+1))2a>-\frac{(N-2)^{2}}{4}+\left(\frac{\alpha(N-2)-(2-2b)}{2(\alpha+1)}\right)^{2}. If max{0,22bN2}<α<42bN2\max\left\{0,\frac{2-2b}{N-2}\right\}<\alpha<\frac{4-2b}{N-2}, then the following statement holds

  • (i)

    |x|b|u|αvLI2Lx2NN+2c(Tθ1+Tθ2)auS(L2;I)αvS(L2;I)\left\||x|^{-b}|u|^{\alpha}v\right\|_{L^{2}_{I}L_{x}^{\frac{2N}{N+2}}}\;\leq\;c\;(T^{\theta_{1}}+T^{\theta_{2}})\|\sqrt{\mathcal{L}_{a}}u\|^{\alpha}_{S(L^{2};I)}\|v\|_{S(L^{2};I)}

  • (ii)

    a(|x|b|u|αu)LI2Lx2NN+2c(Tθ1+Tθ2)auS(L2;I)α+1,\left\|\sqrt{\mathcal{L}_{a}}(|x|^{-b}|u|^{\alpha}u)\right\|_{L^{2}_{I}L_{x}^{\frac{2N}{N+2}}}\;\leq\;c\;(T^{\theta_{1}}+T^{\theta_{2}})\|\sqrt{\mathcal{L}_{a}}u\|^{\alpha+1}_{S(L^{2};I)},

where c,θ1,θ2>0c,\theta_{1},\theta_{2}>0 and I=[0,T]I=[0,T].

Proof.

We start with (ii). By using the equivalence of Sobolev spaces (Remark 2.2) and dividing the estimate in BB and BCB^{C}, we have

a(|x|b|u|αu)LI2Lx2NN+2(|x|b|u|αu)LI2Lx2NN+2(|x|b|u|αu)LI2Lx2NN+2(BC)+(|x|b|u|αu)LI2Lx2NN+2(B).\displaystyle\begin{split}\left\|\sqrt{\mathcal{L}_{a}}(|x|^{-b}|u|^{\alpha}u)\right\|_{L^{2}_{I}L_{x}^{\frac{2N}{N+2}}}&\lesssim\left\|\nabla(|x|^{-b}|u|^{\alpha}u)\right\|_{L^{2}_{I}L_{x}^{\frac{2N}{N+2}}}\\ &\leq\left\|\nabla(|x|^{-b}|u|^{\alpha}u)\right\|_{L^{2}_{I}L_{x}^{\frac{2N}{N+2}}(B^{C})}+\left\|\nabla(|x|^{-b}|u|^{\alpha}u)\right\|_{L^{2}_{I}L_{x}^{\frac{2N}{N+2}}(B)}.\end{split} (4.4)

Let ANA\subset\mathbb{R}^{N} denotes either BB or BCB^{C}. Applying Hölder’s inequality first in space and then in time, we deduce

(|x|b|u|αu)LI2Lx2NN+2(A)|x|bLγ(A)(|u|αu)Lxβ+(|x|b)Ld(A)uLx(α+1)eα+1LI2|x|bLγ(A)uLxαr1αuLxr+|x|b1Ld(A)uLxrα+1LI2T1q(|x|bLγ(A)uLIqLxrα+1+|x|b1Ld(A)uLIqLxrα+1),\begin{split}\left\|\nabla(|x|^{-b}|u|^{\alpha}u)\right\|_{L^{2}_{I}L_{x}^{\frac{2N}{N+2}}(A)}&\leq\left\|\||x|^{-b}\|_{L^{\gamma}(A)}\|\nabla(|u|^{\alpha}u)\|_{L^{\beta}_{x}}+\|\nabla(|x|^{-b})\|_{L^{d}(A)}\|u\|^{\alpha+1}_{L^{(\alpha+1)e}_{x}}\right\|_{L^{2}_{I}}\\ &\lesssim\left\|\||x|^{-b}\|_{L^{\gamma}(A)}\|u\|^{\alpha}_{L_{x}^{\alpha r_{1}}}\|\nabla u\|_{L_{x}^{r}}+\||x|^{-b-1}\|_{L^{d}(A)}\|\nabla u\|^{\alpha+1}_{L^{r}_{x}}\right\|_{L^{2}_{I}}\\ &\lesssim T^{\frac{1}{q^{*}}}\left(\||x|^{-b}\|_{L^{\gamma}(A)}\|\nabla u\|^{\alpha+1}_{L^{q}_{I}L_{x}^{r}}+\||x|^{-b-1}\|_{L^{d}(A)}\|\nabla u\|^{\alpha+1}_{L^{q}_{I}L^{r}_{x}}\right),\end{split} (4.5)

where we also have used the Sobolev inequality. Here, we must have the relations

{N+22N=1γ+1β=1d+1e,1β=1r1+1r1=NrNαr1=NrN(α+1)e,r<N12=1q+α+1q,\begin{cases}\frac{N+2}{2N}=\frac{1}{\gamma}+\frac{1}{\beta}=\frac{1}{d}+\frac{1}{e},\qquad\frac{1}{\beta}=\frac{1}{r_{1}}+\frac{1}{r}\\ 1=\frac{N}{r}-\frac{N}{\alpha r_{1}}=\frac{N}{r}-\frac{N}{(\alpha+1)e}\;,\qquad r<N\\ \frac{1}{2}=\frac{1}{q^{*}}+\frac{\alpha+1}{q},\end{cases}

which in turn are equivalent to

{Nγ=N+22N(α+1)r+αNd=N+22N(α+1)r+α+11q=12α+1q.\begin{cases}\frac{N}{\gamma}=\frac{N+2}{2}-\frac{N(\alpha+1)}{r}+\alpha\\ \frac{N}{d}=\frac{N+2}{2}-\frac{N(\alpha+1)}{r}+\alpha+1\\ \frac{1}{q^{*}}=\frac{1}{2}-\frac{\alpha+1}{q}.\end{cases} (4.6)

Our goal is to find a pair (q,r)(q,r) L2L^{2}-admissible such that |x|bLγ(B)\||x|^{-b}\|_{L^{\gamma}(B)} and |x|b1Ld(B)\||x|^{-b-1}\|_{L^{d}(B)} are finite (see (4.1)), r<Nr<N and 1q>0\frac{1}{q^{*}}>0. Let (q±,r±)(q_{\pm},r_{\pm}) defined by999It is easy to see that q2q\geq 2. Moreover, note that the denominator of qq is positive for α>0\alpha>0, if b1b\geq 1 and α>22bN2\alpha>\frac{2-2b}{N-2}, if b<1b<1.

r±=2N(α+1)N+2+2α2b±εandq±=4(α+1)α(N2)(22b)ε,r_{\pm}=\frac{2N(\alpha+1)}{N+2+2\alpha-2b\pm\varepsilon}\quad\textnormal{and}\quad q_{\pm}=\frac{4(\alpha+1)}{\alpha(N-2)-(2-2b)\mp\varepsilon},

for ε>0\varepsilon>0 sufficiently small. Indeed, we can easily see that (q±,r±)(q_{\pm},r\pm) is SS-admissible, r±<Nr_{\pm}<N (here we need to use that b<N2b<\frac{N}{2}) and

1q±=12α+1q±=42bα(N2)±ε4.\frac{1}{q^{*}_{\pm}}=\frac{1}{2}-\frac{\alpha+1}{q_{\pm}}=\frac{4-2b-\alpha(N-2)\pm\varepsilon}{4}.

Now, if A=BCA=B^{C} we choose (q,r)=(q+,r+)(q,r)=(q_{+},r_{+}) and θ1=1q+\theta_{1}=\frac{1}{q^{*}_{+}}. Then, Nγb>0\frac{N}{\gamma}-b>0 and Ndb1>0\frac{N}{d}-b-1>0, and consequently, |x|bLγ(BC),|x|b1Ld(BC)<\||x|^{-b}\|_{L^{\gamma}(B^{C})},\||x|^{-b-1}\|_{L^{d}(B^{C})}<\infty, by (4.1). On the other hand, if A=BA=B we choose the pair (q,r)=(q,r)(q,r)=(q_{-},r_{-}) and θ2=1q\theta_{2}=\frac{1}{q^{*}_{-}}, so101010Since, α<42bN2\alpha<\frac{4-2b}{N-2} we have θ1,θ2>0\theta_{1},\theta_{2}>0. we also get |x|bLγ(B),|x|b1Ld(B)<\||x|^{-b}\|_{L^{\gamma}(B)},\||x|^{-b-1}\|_{L^{d}(B)}<\infty. Hence, the relations (4.4) and (4.5) implies that

a(|x|b|u|αu)LI2Lx2NN+2(Tθ1+Tθ2)uLIqLxrα+1.\left\|\sqrt{\mathcal{L}_{a}}(|x|^{-b}|u|^{\alpha}u)\right\|_{L^{2}_{I}L_{x}^{\frac{2N}{N+2}}}\lesssim(T^{\theta_{1}}+T^{\theta_{2}})\|\nabla u\|^{\alpha+1}_{L^{q}_{I}L^{r}_{x}}.

Finally, if NNρ<r<N1+ρ\frac{N}{N-\rho}<r<\frac{N}{1+\rho} and applying again Remark 2.2 we complete the proof of part (ii). Indeed, using the value of ρ\rho given in (2.5), it is easy to see that r>NNρr>\frac{N}{N-\rho}. In addition, r<N1+ρr<\frac{N}{1+\rho} is equivalent to 2ρ(α+1)<N2b2\rho(\alpha+1)<N-2b with b<N2b<\frac{N}{2}, which is true assuming our hypothesis on aa.

The proof of (i) is essentially the same as in (ii). It means, we have

|x|b|u|αvLI2Lx2NN+2(A)\displaystyle\left\||x|^{-b}|u|^{\alpha}v\right\|_{L^{2}_{I}L_{x}^{\frac{2N}{N+2}}(A)} \displaystyle\lesssim |x|bLγ(A)uLxαr1αvLxrLI2\displaystyle\left\|\||x|^{-b}\|_{L^{\gamma}(A)}\|u\|^{\alpha}_{L_{x}^{\alpha r_{1}}}\|v\|_{L_{x}^{r}}\right\|_{L^{2}_{I}}
\displaystyle\lesssim T1q|x|bLγ(A)uLIqLxrαvLIqLxr,\displaystyle T^{\frac{1}{q^{*}}}\||x|^{-b}\|_{L^{\gamma}(A)}\|\nabla u\|^{\alpha}_{L^{q}_{I}L_{x}^{r}}\|v\|_{L^{q}_{I}L_{x}^{r}},

where Nγ=N+22N(α+1)r+α\frac{N}{\gamma}=\frac{N+2}{2}-\frac{N(\alpha+1)}{r}+\alpha and 12=1q+α+1q\frac{1}{2}=\frac{1}{q^{*}}+\frac{\alpha+1}{q}. Choosing (q,r)(q,r) and arguing exactly as in (ii) we obtain (i). ∎

In the next lemma we consider the case, α22bN2\alpha\leq\frac{2-2b}{N-2} for 0<b<10<b<1.

Lemma 4.3.

Let N3N\geq 3, 0<b<10<b<1 and a>(N2)24a>-\frac{(N-2)^{2}}{4}. If 22bN<α22bN2\frac{2-2b}{N}<\alpha\leq\frac{2-2b}{N-2}, then the following statement hold

  • (i)

    |x|b|u|αvLI2Lx2NN+2c(Tθ1+Tθ2)auS(L2;I)αvS(L2;I)\left\||x|^{-b}|u|^{\alpha}v\right\|_{L^{2}_{I}L_{x}^{\frac{2N}{N+2}}}\;\leq\;c\left(T^{\theta_{1}}+T^{\theta_{2}}\right)\|\langle\sqrt{\mathcal{L}_{a}}\rangle u\|^{\alpha}_{S(L^{2};I)}\|v\|_{S(L^{2};I)}

  • (ii)

    a(|x|b|u|αu)LI2Lx2NN+2c(Tθ1+Tθ2)auS(L2;I)α+1,\left\|\sqrt{\mathcal{L}_{a}}(|x|^{-b}|u|^{\alpha}u)\right\|_{L^{2}_{I}L_{x}^{\frac{2N}{N+2}}}\;\leq\;c\left(T^{\theta_{1}}+T^{\theta_{2}}\right)\|\langle\sqrt{\mathcal{L}_{a}}\rangle u\|^{\alpha+1}_{S(L^{2};I)},

where c,θ1,θ2>0c,\theta_{1},\theta_{2}>0 and I=[0,T]I=[0,T].

Proof.

We first estimate (ii). We divide in two regions, BB and BCB^{C}. Similarly as the previous lemma, it follows that

(|x|b|u|αu)LI2Lx2NN+2(B)|x|bLγ(B)uLxαr1αuLx2+|x|b1Ld(B)uLxαr1αuLx2NN2LI2T1q(uLIqLxrα+uLIqLxrα)uLx2,\begin{split}\left\|\nabla(|x|^{-b}|u|^{\alpha}u)\right\|_{L^{2}_{I}L_{x}^{\frac{2N}{N+2}}(B)}&\lesssim\left\|\||x|^{-b}\|_{L^{\gamma}(B)}\|u\|^{\alpha}_{L_{x}^{\alpha r_{1}}}\|\nabla u\|_{L_{x}^{2}}+\||x|^{-b-1}\|_{L^{d}(B)}\|u\|^{\alpha}_{L^{\alpha r_{1}}_{x}}\|u\|_{L^{\frac{2N}{N-2}}_{x}}\right\|_{L^{2}_{I}}\\ &\lesssim T^{\frac{1}{q^{*}}}\left(\|\nabla u\|^{\alpha}_{L^{q}_{I}L_{x}^{r}}+\|\nabla u\|^{\alpha}_{L^{q}_{I}L^{r}_{x}}\right)\|\nabla u\|_{L_{x}^{2}},\end{split} (4.7)

where

{N+22N=1γ+1r1+1r=1d+1r1+N22N1=NrNαr112=1q+αq,\begin{cases}\frac{N+2}{2N}=\frac{1}{\gamma}+\frac{1}{r_{1}}+\frac{1}{r}=\frac{1}{d}+\frac{1}{r_{1}}+\frac{N-2}{2N}\\ 1=\frac{N}{r}-\frac{N}{\alpha r_{1}}\\ \frac{1}{2}=\frac{1}{q^{*}}+\frac{\alpha}{q},\end{cases}

which implies that

{Nγ=Nd1=1Nαr+α1q=12α+1q.\begin{cases}\frac{N}{\gamma}=\frac{N}{d}-1=1-\frac{N\alpha}{r}+\alpha\\ \frac{1}{q^{*}}=\frac{1}{2}-\frac{\alpha+1}{q}.\end{cases} (4.8)

For ε>0\varepsilon>0 small, by choosing the SS-admissible pair (q,r)(q,r) defined by111111Note that, the denominator of rr is positive since b<1b<1.

q=4(22b+ε)ε(N2)andr=N(22b+ε)(1b)N+ε,q=\frac{4(2-2b+\varepsilon)}{\varepsilon(N-2)}\quad\textnormal{and}\quad r=\frac{N(2-2b+\varepsilon)}{(1-b)N+\varepsilon},

we deduce that Nγb>0\frac{N}{\gamma}-b>0 and Nd1b>0\frac{N}{d}-1-b>0 (assuming, α22bN2\alpha\leq\frac{2-2b}{N-2}). It leads to |x|bLγ(B)\||x|^{-b}\|_{L^{\gamma}(B)} and |x|b1Ld(B)\||x|^{-b-1}\|_{L^{d}(B)} are finite. Moreover, it is easy to see that 1q>0\frac{1}{q*}>0. On the other hand, since NNρ<r<N1+ρ\frac{N}{N-\rho}<r<\frac{N}{1+\rho} and121212Using the value of ρ\rho, it is easy to check r>NNρr>\frac{N}{N-\rho}. Moreover, chossing ε<(1b)(N2)2+4aρ\varepsilon<\frac{(1-b)\sqrt{(N-2)^{2}+4a}}{\rho} we have r<N1+ρr<\frac{N}{1+\rho}. by the equivalence of Sobolev spaces one has

a(|x|b|u|αu)LI2Lx2NN+2(B)Tθ1auLIqLxrαuLIHa1,\left\|\sqrt{\mathcal{L}_{a}}(|x|^{-b}|u|^{\alpha}u)\right\|_{L^{2}_{I}L_{x}^{\frac{2N}{N+2}}(B)}\lesssim T^{\theta_{1}}\|\sqrt{\mathcal{L}_{a}}u\|^{\alpha}_{L^{q}_{I}L^{r}_{x}}\|u\|_{L^{\infty}_{I}H^{1}_{a}},

where θ1=1q\theta_{1}=\frac{1}{q^{*}}. We now consider the estimate on BCB^{C}. The Hölder inequality, equivalence of Sobolev spaces and Sobolev embedding imply that

(|x|b|u|αu)LI2Lx2NN+2(BC)\displaystyle\left\|\nabla(|x|^{-b}|u|^{\alpha}u)\right\|_{L^{2}_{I}L_{x}^{\frac{2N}{N+2}}(B^{C})} |x|bLγ(BC)uLxαr1αuLx2+|x|b1Ld(BC)uLxαr1αuLx2NN2LI2\displaystyle\lesssim\left\|\||x|^{-b}\|_{L^{\gamma}(B^{C})}\|u\|^{\alpha}_{L_{x}^{\alpha r_{1}}}\|\nabla u\|_{L_{x}^{2}}+\||x|^{-b-1}\|_{L^{d}(B^{C})}\|u\|^{\alpha}_{L^{\alpha r_{1}}_{x}}\|u\|_{L^{\frac{2N}{N-2}}_{x}}\right\|_{L^{2}_{I}}
T12uLIHx1αuLILx2Tθ2uLIHa1α+1,\displaystyle\lesssim T^{\frac{1}{2}}\|u\|^{\alpha}_{L^{\infty}_{I}H^{1}_{x}}\|\nabla u\|_{L^{\infty}_{I}L_{x}^{2}}\lesssim T^{\theta_{2}}\|u\|^{\alpha+1}_{L^{\infty}_{I}H^{1}_{a}},

where θ2=12\theta_{2}=\frac{1}{2} and

Nγ=Nd1=1Nr1.\frac{N}{\gamma}=\frac{N}{d}-1=1-\frac{N}{r_{1}}.

We need to show that H1Lαr1H^{1}\hookrightarrow L^{\alpha r_{1}} and Nγb=Ndb1<0\frac{N}{\gamma}-b=\frac{N}{d}-b-1<0. To this end, we choose r1=N1b+εr_{1}=\frac{N}{1-b+\varepsilon} for ε>0\varepsilon>0 small, we have 2<αr1<2NN22<\alpha r_{1}<\frac{2N}{N-2} by hypothesis131313Note that in the particular case, b=0b=0, if α2N\alpha\geq\frac{2}{N} then αr12\alpha r_{1}\geq 2, so H1Lαr1H^{1}\hookrightarrow L^{\alpha r_{1}}. That is, in this case we can consider α=2N\alpha=\frac{2}{N}. 2(1b)N<α2(1b)N2\frac{2(1-b)}{N}<\alpha\leq\frac{2(1-b)}{N-2}.

To show (i) is only replace |x|b|u|αu|x|^{-b}|u|^{\alpha}\nabla u by |x|b|u|αv|x|^{-b}|u|^{\alpha}v in the proof of (ii). This completes the proof of Lemma 4.3. ∎

Now, with the previous lemmas in hand we are in a position to prove Theorem 1.4.

Proof of Theorem 1.4.

We use the contraction mapping principle. To do so, we define

X=C([0,T];Ha1(N))Lq([0,T];Ha1,r(N)),X=C\left([0,T];H_{a}^{1}(\mathbb{R}^{N})\right)\bigcap L^{q}\left([0,T];H_{a}^{1,r}(\mathbb{R}^{N})\right),

where (q,r)(q,r) is any SS-admissible pair and T>0T>0 will be determined properly later. We shall show that

G(u)(t)=eitau0+iλ0tei(tt)a|x|b|u(t)|αu(t)𝑑tG(u)(t)=e^{-it\mathcal{L}_{a}}u_{0}+i\lambda\int_{0}^{t}e^{-i(t-t^{\prime})\mathcal{L}_{a}}|x|^{-b}|u(t^{\prime})|^{\alpha}u(t^{\prime})dt^{\prime} (4.9)

is a contraction on the complete metric space S(m,T)={uX:auS2(L2;I)m}S(m,T)=\{u\in X:\|\langle\sqrt{\mathcal{L}_{a}}\rangle u\|_{S^{2}(L^{2};I)}\leq m\} with the metric

dT(u,v)=uvS(L2;I),d_{T}(u,v)=\|u-v\|_{S\left(L^{2};I\right)},

where I=[0.T]I=[0.T] and auS2(L2;I)\|\langle\sqrt{\mathcal{L}_{a}}\rangle u\|_{S^{2}(L^{2};I)} is defined in (4.3).

Let us first show that GG is well defined from S(m,T)S(m,T) to S(m,T)S(m,T). Indeed, it follows from the Strichartz inequalities in Lemma 2.5 together with Lemmas 4.2 and 4.3 that

aG(u)S(L2;I)cau0L2+a(|x|b|u|αu)LI2Lx2NN+2cu0Ha1+c(Tθ1+Tθ2)auS(L2;I)α+1,\begin{split}\|\langle\sqrt{\mathcal{L}_{a}}\rangle G(u)\|_{S\left(L^{2};I\right)}&\leq c\|\langle\sqrt{\mathcal{L}_{a}}\rangle u_{0}\|_{L^{2}}+\left\|\langle\sqrt{\mathcal{L}_{a}}\rangle(|x|^{-b}|u|^{\alpha}u)\right\|_{L^{2}_{I}L_{x}^{\frac{2N}{N+2}}}\\ &\leq c\|u_{0}\|_{H^{1}_{a}}+c\left(T^{\theta_{1}}+T^{\theta_{2}}\right)\|\langle\sqrt{\mathcal{L}_{a}}\rangle u\|^{\alpha+1}_{S\left(L^{2};I\right)},\end{split} (4.10)

Consequently, by choosing m=2cu0Ha1m=2c\|u_{0}\|_{H^{1}_{a}} and T>0T>0 such that

cmα(Tθ1+Tθ2)<14,cm^{\alpha}(T^{\theta_{1}}+T^{\theta_{2}})<\frac{1}{4}, (4.11)

we obtain G(u)S(m,T)G(u)\in S(m,T). Hence, GG is well defined on S(m,T)S(m,T).

To prove that GG is a contraction on S(m,T)S(m,T) with respect to the metric dTd_{T}. As before and by (4.2) we deduce

dT(G(u),G(v))c|x|b(|u|αu|v|αv)LI2Lx2NN+2cTθ(auS(L2;I)α+avS(L2;I)α)uvS(L2;I).\begin{split}d_{T}(G(u),G(v))&\leq c\left\||x|^{-b}(|u|^{\alpha}u-|v|^{\alpha}v\right)\|_{L^{2}_{I}L_{x}^{\frac{2N}{N+2}}}\\ &\leq cT^{\theta}\left(\|\sqrt{\mathcal{L}_{a}}u\|^{\alpha}_{S\left(L^{2};I\right)}+\|\sqrt{\mathcal{L}_{a}}v\|^{\alpha}_{S\left(L^{2};I\right)}\right)\|u-v\|_{S\left(L^{2};I\right)}.\end{split}

So, if u,vS(m,T)u,v\in S(m,T), then dT(G(u),G(v))c(Tθ1+Tθ2)aαdT(u,v).d_{T}(G(u),G(v))\leq c(T^{\theta_{1}}+T^{\theta_{2}})a^{\alpha}d_{T}(u,v). Therefore, from (4.11), GG is a contraction on S(m,T)S(m,T) and by the contraction mapping principle we have a unique fixed point uS(m,T)u\in S(m,T) of GG. This completes the proof of the theorem. ∎

4.2. Small Global Theory

In this subsection, we turn our attention to prove the small data global results. Similarly as in the local theory, we establish suitable estimates on the nonlinearity141414When u=vu=v, we denote F(x,u,v)F(x,u,v) by F(x,u)F(x,u). F(x,u,v)=|x|b|u|αvF(x,u,v)=|x|^{-b}|u|^{\alpha}v. It is worth mentioning that, since (2.18) holds for radial data, we obtain global results for radial initial data and also nonradial data. To this end, we use the norms uS(H˙sc)\|u\|_{S(\dot{H}^{s_{c}})} and uS~(H˙sc)\|u\|_{\tilde{S}(\dot{H}^{s_{c}})}, respectively.

We first obatin estimates for a>0a>0 and in the sequel for a<0a<0. For a>0a>0, we use the results obtained by the second author [16]. Recalling the numbers used in [16].

q^=4α(α+2θ)α(Nα+2b)θ(Nα4+2b),r^=Nα(α+2θ)α(Nb)θ(2b),\widehat{q}=\frac{4\alpha(\alpha+2-\theta)}{\alpha(N\alpha+2b)-\theta(N\alpha-4+2b)},\;\;\;\widehat{r}\;=\;\frac{N\alpha(\alpha+2-\theta)}{\alpha(N-b)-\theta(2-b)}, (4.12)

and

a~=2α(α+2θ)α[N(α+1θ)2+2b](42b)(1θ),a^=2α(α+2θ)42b(N2)α.\widetilde{a}\;=\;\frac{2\alpha(\alpha+2-\theta)}{\alpha[N(\alpha+1-\theta)-2+2b]-(4-2b)(1-\theta)},\;\;\;\widehat{a}=\frac{2\alpha(\alpha+2-\theta)}{4-2b-(N-2)\alpha}. (4.13)

It is easy to see that (q^,r^)(\widehat{q},\widehat{r}) is L2L^{2}-admissible, (a^,r^)(\widehat{a},\widehat{r}) is H˙sc\dot{H}^{s_{c}}-admissible and (a~,r^)(\widetilde{a},\widehat{r}) is H˙sc\dot{H}^{-s_{c}}-admissible.

The first lemma will be used to prove the global well posedness in the radial case.

Lemma 4.4.

Let N3N\geq 3, 42bN<α<42bN2\frac{4-2b}{N}<\alpha<\frac{4-2b}{N-2} and 0<b<min{N2,2}0<b<\min\{\frac{N}{2},2\}. If a>0a>0, then there exists θ(0,α)\theta\in(0,\alpha) sufficiently small such that

  • (i)

    F(x,u,v)S(H˙sc)uLtHa1θuS(H˙sc)αθvS(H˙sc)\left\|F(x,u,v)\right\|_{S^{\prime}(\dot{H}^{-s_{c}})}\;\lesssim\;\|u\|^{\theta}_{L^{\infty}_{t}H^{1}_{a}}\|u\|^{\alpha-\theta}_{S(\dot{H}^{s_{c}})}\|v\|_{S(\dot{H}^{s_{c}})}

  • (ii)

    F(x,u,v)S(L2)uLtHa1θuS(H˙sc)αθvS(L2)\left\|F(x,u,v)\right\|_{S^{\prime}(L^{2})}\;\lesssim\;\|u\|^{\theta}_{L^{\infty}_{t}H^{1}_{a}}\|u\|^{\alpha-\theta}_{S(\dot{H}^{s_{c}})}\|v\|_{S(L^{2})}

  • (iii)

    aF(x,u)Ltq^Lxr^uLtHa1θuS(H˙sc)αθauS(L2)\left\|\sqrt{\mathcal{L}_{a}}F(x,u)\right\|_{L^{\widehat{q}^{\prime}}_{t}L^{\widehat{r}^{\prime}}_{x}}\;\lesssim\;\|u\|^{\theta}_{L^{\infty}_{t}H^{1}_{a}}\|u\|^{\alpha-\theta}_{S(\dot{H}^{s_{c}})}\|\sqrt{\mathcal{L}_{a}}u\|_{S(L^{2})}  if  N4N\geq 4

  • (iv)

    aF(x,u)Lt2Lx65uLtHa1θuS(H˙sc)αθauS(L2)\left\|\sqrt{\mathcal{L}_{a}}F(x,u)\right\|_{L^{2}_{t}L_{x}^{\frac{6}{5}}}\;\lesssim\;\|u\|^{\theta}_{L^{\infty}_{t}H^{1}_{a}}\|u\|^{\alpha-\theta}_{S(\dot{H}^{s_{c}})}\|\sqrt{\mathcal{L}_{a}}u\|_{S(L^{2})}   if  N=3N=3  and  α<32b\alpha<3-2b.

Proof.

See [16, Lemmas 4.14.1 and 4.24.2, with s=1s=1] to show (i) and (ii), respectively151515To show (i), the pair used was (a~,r^)\widetilde{a},\widehat{r}) H˙sc\dot{H}^{-s_{c}}-admissible.. To show (iii) we used the estimate used in [16, Lemma 4.3] with s=1s=1, i.e.,

F(x,u)Ltq^Lxr^uLtHa1θuLta^Lxr^αθuLtq^Lxr^.\left\|\nabla F(x,u)\right\|_{L^{\widehat{q}^{\prime}}_{t}L^{\widehat{r}^{\prime}}_{x}}\lesssim\|u\|^{\theta}_{L^{\infty}_{t}H^{1}_{a}}\|u\|^{\alpha-\theta}_{L^{\widehat{a}}_{t}L^{\widehat{r}}_{x}}\|\nabla u\|_{L^{\widehat{q}}_{t}L^{\widehat{r}}_{x}}.

We notice that, if N4N\geq 4 then r^<N\widehat{r}<N and r^<N\widehat{r}^{\prime}<N, so the equivalence of Sobolev spaces (Remark 2.2 with s=1s=1) implies (iii). Finally, we consider (iv). To this end, we use the following numbers161616We use other admissible pairs since r^<N\widehat{r}<N is not true for N=3N=3..

qε=412ε,rε=31+ε\displaystyle q_{\varepsilon}\;=\;\frac{4}{1-2\varepsilon}\;,\;\;\;\;r_{\varepsilon}=\frac{3}{1+\varepsilon} (4.14)

and

a=8(αθ)1+2ε,r=6α(αθ)α(32b2ε)2θ(2b),\displaystyle a\;=\;\frac{8(\alpha-\theta)}{1+2\varepsilon}\;,\;\;\;\;\;\;\;r=\frac{6\alpha(\alpha-\theta)}{\alpha(3-2b-2\varepsilon)-2\theta(2-b)}, (4.15)

where ε\varepsilon is small. Observe that rε<3r_{\varepsilon}<3 and the denominator of rr is positive if b<32b<\frac{3}{2}. Moreover, an easy computation shows that (a,r)(a,r) is H˙sc\dot{H}^{s_{c}}-admissible if171717The condition α<32b\alpha<3-2b implies that r<6r<6, condition of S(H˙sc)S(\dot{H}^{s_{c}})-admissible pair, see (2.14). α<32b\alpha<3-2b and (qε,rε)(q_{\varepsilon},r_{\varepsilon}) is SS-admissible.

Let E(t)=(|x|b|u|αu)Lx65(A)E(t)=\left\|\nabla(|x|^{-b}|u|^{\alpha}u)\right\|_{L_{x}^{\frac{6}{5}}(A)}, where AA denotes either BB or BCB^{C}. The Hölder inequality and the Sobolev embedding lead to

E(t)|x|bLγ(A)uLxθr1θuLxrαθuLxrε+|x|b1Ld(A)uLxθr1θuLxrαθuLxeuLxθr1θuLxrαθuLxrε+uLxθr1θuLxrαθuLxrε,\begin{split}E(t)&\leq\||x|^{-b}\|_{L^{\gamma}(A)}\|u\|^{\theta}_{L_{x}^{\theta r_{1}}}\|u\|^{\alpha-\theta}_{L_{x}^{r}}\|\nabla u\|_{L_{x}^{r_{\varepsilon}}}+\||x|^{-b-1}\|_{L^{d}(A)}\|u\|^{\theta}_{L_{x}^{\theta r_{1}}}\|u\|^{\alpha-\theta}_{L^{r}_{x}}\|u\|_{L_{x}^{e}}\\ &\lesssim\|u\|^{\theta}_{L_{x}^{\theta r_{1}}}\|u\|^{\alpha-\theta}_{L_{x}^{r}}\|\nabla u\|_{L_{x}^{r_{\varepsilon}}}+\|u\|^{\theta}_{L_{x}^{\theta r_{1}}}\|u\|^{\alpha-\theta}_{L^{r}_{x}}\|\nabla u\|_{L_{x}^{r_{\varepsilon}}},\end{split}

where (using 1=3rε3e1=\frac{3}{r_{\varepsilon}}-\frac{3}{e})

{3γ=32+13r13(αθ)r3rε3d=32+13r13(αθ)r3e=32+13r13(αθ)r(3rε1),\begin{cases}\frac{3}{\gamma}=\frac{3}{2}+1-\frac{3}{r_{1}}-\frac{3(\alpha-\theta)}{r}-\frac{3}{r_{\varepsilon}}\\ \frac{3}{d}=\frac{3}{2}+1-\frac{3}{r_{1}}-\frac{3(\alpha-\theta)}{r}-\frac{3}{e}=\frac{3}{2}+1-\frac{3}{r_{1}}-\frac{3(\alpha-\theta)}{r}-(\frac{3}{r_{\varepsilon}}-1),\end{cases} (4.16)

which implies using the definition of the numbers rr and rεr_{\varepsilon} (see (4.14)-(4.15)) that

3γb=3db1=θ(2b)α3r1.\frac{3}{\gamma}-b=\frac{3}{d}-b-1=\frac{\theta(2-b)}{\alpha}-\frac{3}{r_{1}}. (4.17)

If A=BA=B we choose θr1=2NN2\theta r_{1}=\frac{2N}{N-2}, so that Nγb=θ(1sc)>0\frac{N}{\gamma}-b=\theta(1-s_{c})>0 (recall that sc<1s_{c}<1). On the other hand, if A=BCA=B^{C} we choose θr1=2\theta r_{1}=2, so that Nγb=θsc<0\frac{N}{\gamma}-b=-\theta s_{c}<0. Thus, the quantities |x|bLγ(A)\||x|^{-b}\|_{L^{\gamma}(A)}, |x|b1Ld(A)<\||x|^{-b-1}\|_{L^{d}(A)}<\infty and Hx1Lθr1H^{1}_{x}\hookrightarrow L^{\theta r_{1}}. Therefore, since 12=αθa+1qε\frac{1}{2}=\frac{\alpha-\theta}{a}+\frac{1}{q_{\varepsilon}} one has

F(x,u)Lt2Lx65uLtHx1θuLtaLxrαθuLtqεLxrε.\left\|\nabla F(x,u)\right\|_{L^{2}_{t}L^{\frac{6}{5}}_{x}}\lesssim\|u\|^{\theta}_{L^{\infty}_{t}H^{1}_{x}}\|u\|^{\alpha-\theta}_{L^{a}_{t}L^{r}_{x}}\|\nabla u\|_{L^{q_{\varepsilon}}_{t}L^{r_{\varepsilon}}_{x}}.

Hence, applying the equivalence of Sobolev spaces (since rϵ<3r_{\epsilon}<3) we conclude with the proof of (iv). ∎

Our goal now is to show the radial small data result (Theorem 1.8).

Proof of Theorem 1.8.

As usual, our proof is based on the contraction mapping principle. Indeed, define

S={u:uS(H˙sc)2eitau0S(H˙sc)andauS(L2)2cu0Ha1}.S=\{u:\;\|u\|_{S(\dot{H}^{s_{c}})}\leq 2\|e^{-it\mathcal{L}_{a}}u_{0}\|_{S(\dot{H}^{s_{c}})}\quad\textnormal{and}\quad\|\langle\sqrt{\mathcal{L}_{a}}\rangle u\|_{S(L^{2})}\leq 2c\|u_{0}\|_{H^{1}_{a}}\}.

We shall show that G=Gu0G=G_{u_{0}} defined in (4.9) is a contraction on SS equipped with the metric

d(u,v)=uvS(L2)+uvS(H˙sc).d(u,v)=\|u-v\|_{S(L^{2})}+\|u-v\|_{S(\dot{H}^{s_{c}})}.

Indeed, we deduce by the Strichartz inequalities (2.15), (2.16), (2.17) and (2.18)

G(u)S(H˙sc)eitau0S(H˙sc)+cF(x,u)S(H˙sc)\|G(u)\|_{S(\dot{H}^{s_{c}})}\leq\|e^{-it\mathcal{L}_{a}}u_{0}\|_{S(\dot{H}^{s_{c}})}+c\|F(x,u)\|_{S^{\prime}(\dot{H}^{-s_{c}})} (4.18)
G(u)S(L2)cu0L2+cF(x,u)S(L2)\|G(u)\|_{S(L^{2})}\leq c\|u_{0}\|_{L^{2}}+c\|F(x,u)\|_{S^{\prime}(L^{2})} (4.19)

and

aG(u)S(L2)cau0L2+caF(x,u)S(L2),\|\sqrt{\mathcal{L}_{a}}G(u)\|_{S(L^{2})}\leq c\|\sqrt{\mathcal{L}_{a}}u_{0}\|_{L^{2}}+c\|\sqrt{\mathcal{L}_{a}}F(x,u)\|_{S^{\prime}(L^{2})}, (4.20)

where F(x,u)=|x|b|u|αuF(x,u)=|x|^{-b}|u|^{\alpha}u. On the other hand, it follows from Lemma 4.4 and the three last inequalities that, for uSu\in S

G(u)S(H˙sc)\displaystyle\|G(u)\|_{S(\dot{H}^{s_{c}})}\leq eitau0S(H˙sc)+cuLtHa1θuS(H˙sc)αθuS(H˙sc)\displaystyle\|e^{-it\mathcal{L}_{a}}u_{0}\|_{S(\dot{H}^{s_{c}})}+c\|u\|^{\theta}_{L^{\infty}_{t}H^{1}_{a}}\|u\|^{\alpha-\theta}_{S(\dot{H}^{s_{c}})}\|u\|_{S(\dot{H}^{s_{c}})}
\displaystyle\leq eitau0S(H˙sc)+2α+1cθ+1Mθeitau0S(H˙sc)αθ+1\displaystyle\|e^{-it\mathcal{L}_{a}}u_{0}\|_{S(\dot{H}^{s_{c}})}+2^{\alpha+1}c^{\theta+1}M^{\theta}\|e^{-it\mathcal{L}_{a}}u_{0}\|^{\alpha-\theta+1}_{S(\dot{H}^{s_{c}})}

and

aG(u)S(L2)cu0Ha1+cuLtHa1θuS(H˙sc)αθauS(L2)cu0Ha1+cθ+22α+1Mθeitau0S(H˙sc)αθu0Ha1.\begin{split}\|\langle\sqrt{\mathcal{L}_{a}}\rangle G(u)\|_{S(L^{2})}&\leq c\|u_{0}\|_{H^{1}_{a}}+c\|u\|^{\theta}_{L^{\infty}_{t}H^{1}_{a}}\|u\|^{\alpha-\theta}_{S(\dot{H}^{s_{c}})}\|\langle\sqrt{\mathcal{L}_{a}}\rangle u\|_{S(L^{2})}\\ &\leq c\|u_{0}\|_{H^{1}_{a}}+c^{\theta+2}2^{\alpha+1}M^{\theta}\|e^{-it\mathcal{L}_{a}}u_{0}\|^{\alpha-\theta}_{S(\dot{H}^{s_{c}})}\|u_{0}\|_{H^{1}_{a}}.\end{split} (4.21)

Now if U(t)u0S(H˙sc)<δ\|U(t)u_{0}\|_{S(\dot{H}^{s_{c}})}<\delta with

δmin12cθ+12α+1Mθαθ,\delta\leq\min\sqrt[\alpha-\theta]{\frac{1}{2c^{\theta+1}2^{\alpha+1}M^{\theta}}}, (4.22)

where A>0A>0 is a number such that u0HsA\|u_{0}\|_{H^{s}}\leq A, then

G(u)S~(H˙sc)2eitau0S~(H˙sc)andaG(u)S(L2)2cu0Ha1,\|G(u)\|_{\widetilde{S}(\dot{H}^{s_{c}})}\leq 2\|e^{-it\mathcal{L}_{a}}u_{0}\|_{\widetilde{S}(\dot{H}^{s_{c}})}\quad\mbox{and}\quad\|\langle\sqrt{\mathcal{L}_{a}}\rangle G(u)\|_{S(L^{2})}\leq 2c\|u_{0}\|_{H^{1}_{a}},

that is G(u)SG(u)\in S.

To prove that GG is a contraction on SS, we repeat the above computations. Indeed, taking u,vSu,v\in S

G(u)G(v)S(H˙sc)2α+1cθ+1Mθeitau0B(H˙sc)αθuvS(H˙sc)\|G(u)-G(v)\|_{S(\dot{H}^{s_{c}})}\leq 2^{\alpha+1}c^{\theta+1}M^{\theta}\|e^{-it\mathcal{L}_{a}}u_{0}\|^{\alpha-\theta}_{B(\dot{H}^{s_{c}})}\|u-v\|_{S(\dot{H}^{s_{c}})} (4.23)

and

G(u)G(v)S(L2)2α+1cθ+1Mθeitau0B(H˙sc)αθuvS(L2).\|G(u)-G(v)\|_{S(L^{2})}\leq 2^{\alpha+1}c^{\theta+1}M^{\theta}\|e^{-it\mathcal{L}_{a}}u_{0}\|^{\alpha-\theta}_{B(\dot{H}^{s_{c}})}\|u-v\|_{S(L^{2})}. (4.24)

By using the last inequalities and (4.22) we obtain

d(G(u),G(v))2α+1cθ+1Mθeitau0B(H˙sc)αθd(u,v)12d(u,v),d(G(u),G(v))\leq 2^{\alpha+1}c^{\theta+1}M^{\theta}\|e^{-it\mathcal{L}_{a}}u_{0}\|^{\alpha-\theta}_{B(\dot{H}^{s_{c}})}\;d(u,v)\leq\frac{1}{2}d(u,v),

i.e., GG is a contraction.

Therefore, by the Banach Fixed Point Theorem, GG has a unique fixed point uSu\in S, which is a global solution of (1.1). ∎

Remark 4.5.

Observe that Theorem 1.4 (when N=3N=3) shows global well posedness with an extra condition on α\alpha, i.e., α<32b\alpha<3-2b. We now establish suitable estimates on F(x,u,v)F(x,u,v) for the full intercritical regime, that is, 42b3<α<42b\frac{4-2b}{3}<\alpha<4-2b.

Before stating the lemma, we first define the following numbers

a¯=4(αθ)1+2εr¯=6α(αθ)α(32b2ε)θ(42b)\displaystyle\bar{a}=\frac{4(\alpha-\theta)}{1+2\varepsilon}\qquad\bar{r}=\frac{6\alpha(\alpha-\theta)}{\alpha(3-2b-2\varepsilon)-\theta(4-2b)} (4.25)

and

q=412εr=31+ε.\displaystyle q=\frac{4}{1-2\varepsilon}\qquad r=\frac{3}{1+\varepsilon}. (4.26)

Observe that (q,r(q,r) is SS-admissible and r<Nr<N. Moreover, 12=αθa¯+1q\frac{1}{2}=\frac{\alpha-\theta}{\bar{a}}+\frac{1}{q}.

Lemma 4.6.

Let N=3N=3 and a>0a>0. If 0<b<320<b<\frac{3}{2} and max{42b3,1}<α<42b\max\{\frac{4-2b}{3},1\}<\alpha<4-2b, then the following statements hold

  • (i)

    F(x,u,v)Lt2Lx65uLtHa1θuLta¯Lxr¯αθvS(L2)\left\|F(x,u,v)\right\|_{L^{2}_{t}L_{x}^{\frac{6}{5}}}\;\lesssim\;\|u\|^{\theta}_{L^{\infty}_{t}H^{1}_{a}}\|u\|^{\alpha-\theta}_{L^{\bar{a}}_{t}L^{\bar{r}}_{x}}\|v\|_{S(L^{2})}

  • (ii)

    aF(x,u)Lt2Lx65uLtHa1θuLta¯Lxr¯αθauS(L2)\left\|\sqrt{\mathcal{L}_{a}}F(x,u)\right\|_{L^{2}_{t}L_{x}^{\frac{6}{5}}}\;\lesssim\;\|u\|^{\theta}_{L^{\infty}_{t}H^{1}_{a}}\|u\|^{\alpha-\theta}_{L^{\bar{a}}_{t}L^{\bar{r}}_{x}}\|\sqrt{\mathcal{L}_{a}}u\|_{S(L^{2})}

  • (iii)

    ascF(x,u)Lt2Lx65uLtHa1θuLta¯Lxr¯αθauS(L2)\left\|\sqrt{\mathcal{L}_{a}}^{s_{c}}F(x,u)\right\|_{L^{2}_{t}L_{x}^{\frac{6}{5}}}\;\lesssim\;\|u\|^{\theta}_{L^{\infty}_{t}H^{1}_{a}}\|u\|^{\alpha-\theta}_{L^{\bar{a}}_{t}L^{\bar{r}}_{x}}\|\langle\sqrt{\mathcal{L}_{a}}\rangle u\|_{S(L^{2})},

where θ\theta is sufficiently small.

Proof.

Let us prove (i). Similarly as in Lemma 4.4 (iv), it follows that (since 12=αθa¯+1q\frac{1}{2}=\frac{\alpha-\theta}{\bar{a}}+\frac{1}{q})

F(x,u,v)Lt2Lx65(A)|x|bLγ(A)uLtLxθr1θuLta¯Lxr¯αθvLtqLxr,\begin{split}\left\|F(x,u,v)\right\|_{L^{2}_{t}L_{x}^{\frac{6}{5}}(A)}&\leq\||x|^{-b}\|_{L^{\gamma}(A)}\|u\|^{\theta}_{L^{\infty}_{t}L_{x}^{\theta r_{1}}}\|u\|^{\alpha-\theta}_{L_{t}^{\bar{a}}L_{x}^{\bar{r}}}\|v\|_{L^{q}_{t}L_{x}^{r}},\end{split}

where AA denotes either BB or BCB^{C}. Moreover 3γ=523r13(αθ)r¯3r\frac{3}{\gamma}=\frac{5}{2}-\frac{3}{r_{1}}-\frac{3(\alpha-\theta)}{\bar{r}}-\frac{3}{r}. From (4.25), (4.26) we have

3γb=θ(2b)α3r1,\frac{3}{\gamma}-b=\frac{\theta(2-b)}{\alpha}-\frac{3}{r_{1}},

which is the same relation as in (4.17), so choosing r1r_{1} as in Lemma 4.4 (iv) we deduce that |x|bLγ(A)|x|^{-b}\in L^{\gamma}(A) and Hx1Lθr1H^{1}_{x}\hookrightarrow L^{\theta r_{1}}. Hence, in view of (q,rq,r) is SS-admissible we obatin (i). We now consider (ii). The equivalence of Sobolev spaces implies that

a(|x|b|u|αu)Lt2Lx65(|x|b|u|αu)Lt2Lx65.\left\|\sqrt{\mathcal{L}_{a}}(|x|^{-b}|u|^{\alpha}u)\right\|_{L_{t}^{2}L^{\frac{6}{5}}_{x}}\leq\left\|\nabla(|x|^{-b}|u|^{\alpha}u)\right\|_{L_{t}^{2}L^{\frac{6}{5}}_{x}}. (4.27)

The derivate of FF leads to two terms, one of the form |x|b|u|αu|x|^{-b}|u|^{\alpha}\nabla u and one of the form |x|b1|u|αu|x|^{-b-1}|u|^{\alpha}u, so for estimating the first one we replace vv by u\nabla u in (i). Thus,

|x|b|u|αuLt2Lx65\displaystyle\left\||x|^{-b}|u|^{\alpha}\nabla u\right\|_{L^{2}_{t}L^{\frac{6}{5}}_{x}} \displaystyle\lesssim uLtHa1θuLta¯Lxr¯αθuLtqLxr.\displaystyle\|u\|^{\theta}_{L^{\infty}_{t}H^{1}_{a}}\|u\|^{\alpha-\theta}_{L^{\bar{a}}_{t}L^{\bar{r}}_{x}}\|\nabla u\|_{L^{q}_{t}L^{r}_{x}}.

Furthermore, as in Lemma 4.4 (iv) we get

|x|b1|u|αuLt2Lx65(A)|x|b1Ld(A)uLxθr1θuLx(αθ)r2αθuLxr3Lt2uLtHx1θuLta¯Lxr¯αθuLtqLxr,\begin{split}\left\||x|^{-b-1}|u|^{\alpha}u\right\|_{L_{t}^{2}L^{\frac{6}{5}}_{x}(A)}&\leq\left\|\||x|^{-b-1}\|_{L^{d}(A)}\|u\|_{L_{x}^{\theta r_{1}}}^{\theta}\|u\|^{\alpha-\theta}_{L_{x}^{(\alpha-\theta)r_{2}}}\|u\|_{L^{r_{3}}_{x}}\right\|_{L^{2}_{t}}\\ &\lesssim\|u\|_{L^{\infty}_{t}H_{x}^{1}}^{\theta}\|u\|^{\alpha-\theta}_{L^{\bar{a}}_{t}L^{\bar{r}}_{x}}\|\nabla u\|_{L^{q}_{t}L^{r}_{x}},\end{split}

where 3db1=θ(2b)α3r1\frac{3}{d}-b-1=\frac{\theta(2-b)}{\alpha}-\frac{3}{r_{1}}. Again the last relation is the same relation as in (4.17). Therefore, applying the equivalence of Sobolev spaces (since r<3r<3) together with (4.27), one has

a(|x|b|u|αu)Lt2Lx65uLtHa1θuLta¯Lxr¯αθauLtqLxr.\left\|\sqrt{\mathcal{L}_{a}}(|x|^{-b}|u|^{\alpha}u)\right\|_{L_{t}^{2}L^{\frac{6}{5}}_{x}}\lesssim\|u\|_{L^{\infty}_{t}H_{a}^{1}}^{\theta}\|u\|^{\alpha-\theta}_{L^{\bar{a}}_{t}L^{\bar{r}}_{x}}\|\sqrt{\mathcal{L}_{a}}u\|_{L^{q}_{t}L^{r}_{x}}.

Finally, we show (iii). Indeed, the interpolation inequality and (i)-(ii) imply that

ascF(x,u)Lt2Lx65F(x,u)Lt2Lx651scaF(x,u)Lt2Lx65scuLtHa1θuLta¯Lxr¯αθuS(L2)1scauS(L2)sc.\begin{split}\left\|\sqrt{\mathcal{L}_{a}}^{s_{c}}F(x,u)\right\|_{L^{2}_{t}L^{\frac{6}{5}}_{x}}&\lesssim\left\|F(x,u)\right\|^{1-s_{c}}_{L^{2}_{t}L^{\frac{6}{5}}_{x}}\left\|\sqrt{\mathcal{L}_{a}}F(x,u)\right\|^{s_{c}}_{L^{2}_{t}L^{\frac{6}{5}}_{x}}\\ &\lesssim\|u\|^{\theta}_{L^{\infty}_{t}H^{1}_{a}}\|u\|^{\alpha-\theta}_{L^{\bar{a}}_{t}L^{\bar{r}}_{x}}\|u\|^{1-s_{c}}_{S(L^{2})}\|\sqrt{\mathcal{L}_{a}}u\|^{s_{c}}_{S(L^{2})}.\end{split}

This completes the proof of the lemma. ∎

Remark 4.7.

Note that, in Lemma 4.6 the pair (a¯,r¯)\bar{a},\bar{r}) is not S(H˙sc)S(\dot{H}^{s_{c}})-admissible due to r¯<6\bar{r}<6 not being true for α32b\alpha\geq 3-2b. However, we obtain a small data global result assuming eitau0Lta¯Lxr¯\|e^{-it\mathcal{L}_{a}}u_{0}\|_{L^{\bar{a}}_{t}L^{\bar{r}}_{x}} sufficiently small. For the proof of this result (Theorem 1.9) we do not use the Strichartz estimate (2.18) in view of (a¯,r¯)(\bar{a},\bar{r}) not being H˙sc\dot{H}^{s_{c}}-admissible. Instead, use the Sobolev embedding and apply the Strichartz estimate (2.17). This is possible since the SS-admissible pair used satisfies the conditions (2.12). It is also worth mentioning that, since we do not use (2.18) then the result holds for nonradial data.

Proof of Theorem 1.9.

We only show (i) since (ii) and (ii) are immediate consequences. As before, define

S={u:uS~(H˙sc)2eitau0S~(H˙sc)andauS(L2)2cu0Ha1}.S=\{u:\;\|u\|_{\tilde{S}(\dot{H}^{s_{c}})}\leq 2\|e^{-it\mathcal{L}_{a}}u_{0}\|_{\tilde{S}(\dot{H}^{s_{c}})}\quad\textnormal{and}\quad\|\langle\sqrt{\mathcal{L}_{a}}\rangle u\|_{S(L^{2})}\leq 2c\|u_{0}\|_{H^{1}_{a}}\}.

We prove that G=Gu0G=G_{u_{0}} defined in (4.9) is a contraction on SS equipped with the metric

d(u,v)=uvS(L2).d(u,v)=\|u-v\|_{S(L^{2})}.

Combining the Sobolev embedding, equivalence of Sobolev spaces and Strichartz estimate (2.17), it follows that

G(u)S~(H˙sc)\displaystyle\|G(u)\|_{\tilde{S}(\dot{H}^{s_{c}})} \displaystyle\leq eitau0S~(H˙sc)+cDsc0tei(ts)aF(x,u)𝑑sLta¯Lxp¯\displaystyle\|e^{-it\mathcal{L}_{a}}u_{0}\|_{\tilde{S}(\dot{H}^{s_{c}})}+c\left\|D^{s_{c}}\int_{0}^{t}e^{-i(t-s)\mathcal{L}_{a}}F(x,u)ds\right\|_{L_{t}^{\bar{a}}L_{x}^{\bar{p}}}
\displaystyle\leq eitau0S~(H˙sc)+cascF(x,u)Lt2Lx2NN+2,\displaystyle\|e^{-it\mathcal{L}_{a}}u_{0}\|_{\tilde{S}(\dot{H}^{s_{c}})}+c\left\|\sqrt{\mathcal{L}_{a}}^{s_{c}}F(x,u)\right\|_{L^{2}_{t}L_{x}^{\frac{2N}{N+2}}},

where181818It is easy to see that 2<p¯<3sc2<\bar{p}<\frac{3}{s_{c}} and since 42b3<α<42b\frac{4-2b}{3}<\alpha<4-2b we have that (a¯,p¯)(\bar{a},\bar{p}) is SS-admissible.

p¯=6α(αθ)α(32b2ε)+2αsc(αθ)θ(42b).\bar{p}=\frac{6\alpha(\alpha-\theta)}{\alpha(3-2b-2\varepsilon)+2\alpha s_{c}(\alpha-\theta)-\theta(4-2b)}. (4.28)

In addition,

G(u)S(L2)cu0L2+cF(x,u)Lt2Lx2NN+2\|G(u)\|_{S(L^{2})}\leq c\|u_{0}\|_{L^{2}}+c\|F(x,u)\|_{L^{2}_{t}L_{x}^{\frac{2N}{N+2}}} (4.29)

and

aG(u)S(L2)cau0L2+caF(x,u)Lt2Lx2NN+2.\|\sqrt{\mathcal{L}_{a}}G(u)\|_{S(L^{2})}\leq c\|\sqrt{\mathcal{L}_{a}}u_{0}\|_{L^{2}}+c\|\sqrt{\mathcal{L}_{a}}F(x,u)\|_{L^{2}_{t}L_{x}^{\frac{2N}{N+2}}}. (4.30)

An application of Lemma 4.6 together with the last inequalities yield, for any uSu\in S,

G(u)S~(H˙sc)\displaystyle\|G(u)\|_{\tilde{S}(\dot{H}^{s_{c}})} \displaystyle\leq eitau0S~(H˙sc)+cuLtHa1θuS~(H˙sc)αθauS(L2)\displaystyle\|e^{-it\mathcal{L}_{a}}u_{0}\|_{\tilde{S}(\dot{H}^{s_{c}})}+c\|u\|^{\theta}_{L^{\infty}_{t}H^{1}_{a}}\|u\|^{\alpha-\theta}_{\widetilde{S}(\dot{H}^{s_{c}})}\|\langle\sqrt{\mathcal{L}_{a}}\rangle u\|_{S(L^{2})}
\displaystyle\leq eitau0S~(H˙sc)+cθ+22α+1Mθ+1eitau0S~(H˙sc)αθ\displaystyle\|e^{-it\sqrt{\mathcal{L}_{a}}}u_{0}\|_{\tilde{S}(\dot{H}^{s_{c}})}+c^{\theta+2}2^{\alpha+1}M^{\theta+1}\|e^{-it\mathcal{L}_{a}}u_{0}\|^{\alpha-\theta}_{\widetilde{S}(\dot{H}^{s_{c}})}

and

aG(u)S(L2)cu0Ha1+cuLtHa1θuS~(H˙sc)αθauS(L2)cu0Ha1+cθ+22α+1Mθeitau0S~(H˙sc)αθu0Ha1.\begin{split}\|\langle\sqrt{\mathcal{L}_{a}}\rangle G(u)\|_{S(L^{2})}&\leq c\|u_{0}\|_{H^{1}_{a}}+c\|u\|^{\theta}_{L^{\infty}_{t}H^{1}_{a}}\|u\|^{\alpha-\theta}_{\widetilde{S}(\dot{H}^{s_{c}})}\|\langle\sqrt{\mathcal{L}_{a}}\rangle u\|_{S(L^{2})}\\ &\leq c\|u_{0}\|_{H^{1}_{a}}+c^{\theta+2}2^{\alpha+1}M^{\theta}\|e^{-it\mathcal{L}_{a}}u_{0}\|^{\alpha-\theta}_{\widetilde{S}(\dot{H}^{s_{c}})}\|u_{0}\|_{H^{1}_{a}}.\end{split} (4.31)

If eitau0S~(H˙sc)<δ\|e^{it\mathcal{L}_{a}}u_{0}\|_{\widetilde{S}(\dot{H}^{s_{c}})}<\delta with191919Note that, to define δ\delta, we need the condition α>1\alpha>1. (since α>1\alpha>1 and θ\theta small)

δmin{12cθ+22α+1Mθ+1α1θ,12cθ+12α+1Mθαθ},\delta\leq\min\left\{\sqrt[\alpha-1-\theta]{\frac{1}{2c^{\theta+2}2^{\alpha+1}M^{\theta+1}}},\sqrt[\alpha-\theta]{\frac{1}{2c^{\theta+1}2^{\alpha+1}M^{\theta}}}\right\}, (4.32)

implies that

G(u)S~(H˙sc)2eitau0S~(H˙sc)andaG(u)S(L2)2cu0Ha1.\|G(u)\|_{\widetilde{S}(\dot{H}^{s_{c}})}\leq 2\|e^{-it\mathcal{L}_{a}}u_{0}\|_{\widetilde{S}(\dot{H}^{s_{c}})}\quad\mbox{and}\quad\|\langle\sqrt{\mathcal{L}_{a}}\rangle G(u)\|_{S(L^{2})}\leq 2c\|u_{0}\|_{H^{1}_{a}}.

therefore, G(u)SG(u)\in S.

To show that GG is a contraction on SS, we repeat the above computations. Indeed, let u,vSu,v\in S one has

G(u)G(v)S(L2)2α+1cθ+1Mθeitau0B(H˙sc)αθuvS(L2).\|G(u)-G(v)\|_{S(L^{2})}\leq 2^{\alpha+1}c^{\theta+1}M^{\theta}\|e^{-it\mathcal{L}_{a}}u_{0}\|^{\alpha-\theta}_{B(\dot{H}^{s_{c}})}\|u-v\|_{S(L^{2})}. (4.33)

By using the last inequalities and (4.32) we obtain

d(G(u),G(v))2α+1cθ+1Mθeitau0B(H˙sc)αθd(u,v)12d(u,v),d(G(u),G(v))\leq 2^{\alpha+1}c^{\theta+1}M^{\theta}\|e^{-it\mathcal{L}_{a}}u_{0}\|^{\alpha-\theta}_{B(\dot{H}^{s_{c}})}\;d(u,v)\leq\frac{1}{2}d(u,v),

which implies that GG is also a contraction. Therefore, by the contraction mapping principle, GG has a unique fixed point uSu\in S. ∎

We end this section with the proof of Theorem 1.10. To this end, we establish good estimates on the nonlinearity, for negative values of aa.

Lemma 4.8.

Let 0<b<6N20<b<\frac{6-N}{2}. Assume that (N,a,α)(N,a,\alpha) satisfy

{a>(N2)24ifN=3,42b3<α 22band  0b<12,a>(N2)24+(α(N2)(22b)2(α+1))2if  3N5,max{42bN,22bN2,1}<α<42bN2.\begin{cases}a>-\frac{(N-2)^{2}}{4}\;\;\;\;\qquad\qquad\qquad\qquad\ \textnormal{if}\;\;N=3,\;\;\;\frac{4-2b}{3}<\alpha\leq\;2-2b\quad\textnormal{and}\;\;0\leq b<\frac{1}{2},\\ a>-\frac{(N-2)^{2}}{4}+\left(\frac{\alpha(N-2)-(2-2b)}{2(\alpha+1)}\right)^{2}\;\;\textnormal{if}\;\;3\leq N\leq 5,\;\;\;\max\{\frac{4-2b}{N},\frac{2-2b}{N-2},1\}<\alpha<\frac{4-2b}{N-2}.\end{cases} (4.34)

Then the following statements hold

  • (i)

    F(x,u,v)Lt2Lx2NN+2uLtHa1θuLta¯Lxr¯αθvS(L2)\left\|F(x,u,v)\right\|_{L^{2}_{t}L_{x}^{\frac{2N}{N+2}}}\;\lesssim\;\|u\|^{\theta}_{L^{\infty}_{t}H^{1}_{a}}\|u\|^{\alpha-\theta}_{L^{\bar{a}}_{t}L^{\bar{r}}_{x}}\|v\|_{S(L^{2})}

  • (ii)

    aF(x,u)Lt2Lx2NN+2uLtHa1θuLta¯Lxr¯αθauS(L2)\left\|\sqrt{\mathcal{L}_{a}}F(x,u)\right\|_{L^{2}_{t}L_{x}^{\frac{2N}{N+2}}}\;\lesssim\;\|u\|^{\theta}_{L^{\infty}_{t}H^{1}_{a}}\|u\|^{\alpha-\theta}_{L^{\bar{a}}_{t}L^{\bar{r}}_{x}}\|\sqrt{\mathcal{L}_{a}}u\|_{S(L^{2})}

  • (iii)

    ascF(x,u)Lt2Lx2NN+2uLtHa1θuLta¯Lxr¯αθauS(L2)\left\|\sqrt{\mathcal{L}_{a}}^{s_{c}}F(x,u)\right\|_{L^{2}_{t}L_{x}^{\frac{2N}{N+2}}}\;\lesssim\;\|u\|^{\theta}_{L^{\infty}_{t}H^{1}_{a}}\|u\|^{\alpha-\theta}_{L^{\bar{a}}_{t}L^{\bar{r}}_{x}}\|\langle\sqrt{\mathcal{L}_{a}}\rangle u\|_{S(L^{2})},

where θ\theta is sufficiently small.

Proof.

In view of (4.34) we have two cases:
Case 11. For N=3N=3, 42bN<α22b\frac{4-2b}{N}<\alpha\leq 2-2b and 0b<120\leq b<\frac{1}{2}, we define

a¯=2(αθ)1θ,r¯=3α(αθ)α(1b)θ(2bα),q=2θandr=632θ.\bar{a}=\frac{2(\alpha-\theta)}{1-\theta}\;,\quad\;\bar{r}=\frac{3\alpha(\alpha-\theta)}{\alpha(1-b)-\theta(2-b-\alpha)}\;,\quad\;q=\frac{2}{\theta}\;\quad\textnormal{and}\;\quad r=\frac{6}{3-2\theta}. (4.35)

Case 22. For 3N53\leq N\leq 5 and max{22bN2,42bN,1}<α<42bN2\max\{\frac{2-2b}{N-2},\frac{4-2b}{N},1\}<\alpha<\frac{4-2b}{N-2}, define

a¯=4(α+1)(αθ)42bα(N4)r¯=2Nα(α+1)(αθ)α2(N2b)θ(42b)(α+1)\bar{a}=\frac{4(\alpha+1)(\alpha-\theta)}{4-2b-\alpha(N-4)}\qquad\quad\bar{r}=\frac{2N\alpha(\alpha+1)(\alpha-\theta)}{\alpha^{2}(N-2b)-\theta(4-2b)(\alpha+1)} (4.36)

and

q=4(α+1)α(N2)2+2br=2N(α+1)2(α+1)+N2b.q=\frac{4(\alpha+1)}{\alpha(N-2)-2+2b}\quad\qquad r=\frac{2N(\alpha+1)}{2(\alpha+1)+N-2b}. (4.37)

Note that, the hypothesis (4.34) yields202020Recalling, the equivalence of Sobolev spaces (Remark 2.2) holds if NNρ<r<N1+ρ\frac{N}{N-\rho}<r<\frac{N}{1+\rho}. NNρ<r<N1+ρ\frac{N}{N-\rho}<r<\frac{N}{1+\rho} and (q,r)(q,r) is S-admissible. Moreover, since θ\theta is small, b<N2b<\frac{N}{2} and α>22bN2\alpha>\frac{2-2b}{N-2}, we get 2<r<N2<r<N.

The proof is similar as in Lemma 4.6. We first consider (i). Let A={B,BC}A=\{B,B^{C}\}. It follows that (using 12=αθa¯+1q\frac{1}{2}=\frac{\alpha-\theta}{\bar{a}}+\frac{1}{q})

F(x,u,v)Lt2Lx2NN+2(A)|x|bLγ(A)uLtLxθr1θuLta¯Lxr¯αθvLtqLxr,\begin{split}\left\|F(x,u,v)\right\|_{L^{2}_{t}L_{x}^{\frac{2N}{N+2}}(A)}&\leq\||x|^{-b}\|_{L^{\gamma}(A)}\|u\|^{\theta}_{L^{\infty}_{t}L_{x}^{\theta r_{1}}}\|u\|^{\alpha-\theta}_{L_{t}^{\bar{a}}L_{x}^{\bar{r}}}\|v\|_{L^{q}_{t}L_{x}^{r}},\end{split}

where Nγ=N+22Nr1N(αθ)r¯Nr\frac{N}{\gamma}=\frac{N+2}{2}-\frac{N}{r_{1}}-\frac{N(\alpha-\theta)}{\bar{r}}-\frac{N}{r}. Hence, the values of r¯\bar{r} and rr imply

Nγb=θ(2b)αNr1,\frac{N}{\gamma}-b=\frac{\theta(2-b)}{\alpha}-\frac{N}{r_{1}},

which is the same relation as in (4.17), so arguing as in Lemma 4.6 and since (q,rq,r) is SS-admissible we deduce (i). Let us prove (ii). From the equivalence of Sobolev spaces we know that

a(|x|b|u|αu)Lt2Lx2NN+2(|x|b|u|αu)Lt2Lx2NN+2.\left\|\sqrt{\mathcal{L}_{a}}(|x|^{-b}|u|^{\alpha}u)\right\|_{L_{t}^{2}L^{\frac{2N}{N+2}}_{x}}\leq\left\|\nabla(|x|^{-b}|u|^{\alpha}u)\right\|_{L_{t}^{2}L^{\frac{2N}{N+2}}_{x}}. (4.38)

We only estimate |x|b1|u|αu|x|^{b-1}|u|^{\alpha}u. We deduce that (using 1=NrNr31=\frac{N}{r}-\frac{N}{r_{3}}, with r<Nr<N)

|x|b1|u|αuLt2Lx2NN+2(A)|x|b1Ld(A)uLxθr1θuLx(αθ)r2αθuLxr3Lt2uLtHx1θuLta¯Lxr¯αθuLtqLxr,\begin{split}\left\||x|^{-b-1}|u|^{\alpha}u\right\|_{L_{t}^{2}L^{\frac{2N}{N+2}}_{x}(A)}&\leq\left\|\||x|^{-b-1}\|_{L^{d}(A)}\|u\|_{L_{x}^{\theta r_{1}}}^{\theta}\|u\|^{\alpha-\theta}_{L_{x}^{(\alpha-\theta)r_{2}}}\|u\|_{L^{r_{3}}_{x}}\right\|_{L^{2}_{t}}\\ &\lesssim\|u\|_{L^{\infty}_{t}H_{x}^{1}}^{\theta}\|u\|^{\alpha-\theta}_{L^{\bar{a}}_{t}L^{\bar{r}}_{x}}\|\nabla u\|_{L^{q}_{t}L^{r}_{x}},\end{split}

where we get the same relation as in (4.17), i.e., Ndb1=θ(2b)αNr1\frac{N}{d}-b-1=\frac{\theta(2-b)}{\alpha}-\frac{N}{r_{1}}. Therefore, again the equivalence of Sobolev spaces together with (4.38) lead to

a(|x|b|u|αu)Lt2Lx2NN+2uLtHa1θuLta¯Lxr¯αθauLtqLxr.\left\|\sqrt{\mathcal{L}_{a}}(|x|^{-b}|u|^{\alpha}u)\right\|_{L_{t}^{2}L^{\frac{2N}{N+2}}_{x}}\lesssim\|u\|_{L^{\infty}_{t}H_{a}^{1}}^{\theta}\|u\|^{\alpha-\theta}_{L^{\bar{a}}_{t}L^{\bar{r}}_{x}}\|\sqrt{\mathcal{L}_{a}}u\|_{L^{q}_{t}L^{r}_{x}}.

In addition, applying the interpolation inequality and (i) - (ii) we have (iii), that is

ascF(x,u)Lt2Lx2NN+2uLtHa1θuLta¯Lxr¯αθuS(L2)1scauS(L2)sc.\begin{split}\left\|\sqrt{\mathcal{L}_{a}}^{s_{c}}F(x,u)\right\|_{L^{2}_{t}L^{\frac{2N}{N+2}}_{x}}&\lesssim\|u\|^{\theta}_{L^{\infty}_{t}H^{1}_{a}}\|u\|^{\alpha-\theta}_{L^{\bar{a}}_{t}L^{\bar{r}}_{x}}\|u\|^{1-s_{c}}_{S(L^{2})}\|\sqrt{\mathcal{L}_{a}}u\|^{s_{c}}_{S(L^{2})}.\end{split}

Proof of Theorem 1.10.

First, we define

p¯=Nα(αθ)αsc(αθ)+α(1b)θ(2bα),\bar{p}=\frac{N\alpha(\alpha-\theta)}{\alpha s_{c}(\alpha-\theta)+\alpha(1-b)-\theta(2-b-\alpha)}, (4.39)

if Case 11 of Lemma 4.8 holds and

p¯=2Nα(α+1)(αθ)2αsc(α+1)(αθ)+α2(N2b)θ(42b)(α+1),\bar{p}=\frac{2N\alpha(\alpha+1)(\alpha-\theta)}{2\alpha s_{c}(\alpha+1)(\alpha-\theta)+\alpha^{2}(N-2b)-\theta(4-2b)(\alpha+1)}, (4.40)

if Case 22 of Lemma 4.8 holds. Note that, One has that (a¯,p¯)(\bar{a},\bar{p}) is SS-admissible and sc=Np¯Nr¯s_{c}=\frac{N}{\bar{p}}-\frac{N}{\bar{r}}. Moreover, p¯<Nsc\bar{p}<\frac{N}{s_{c}}, for θ\theta small enough and 2<p¯<2NN22<\bar{p}<\frac{2N}{N-2}. Hence, applying Sobolev embedding we get

0tei(ts)aF(x,u)𝑑sLta¯Lxr¯\displaystyle\|\int_{0}^{t}e^{-i(t-s)\mathcal{L}_{a}}F(x,u)ds\|_{L_{t}^{\bar{a}}L_{x}^{\bar{r}}} \displaystyle\leq Dsc0tei(ts)aF(x,u)𝑑sLta¯Lxp¯.\displaystyle\left\|D^{s_{c}}\int_{0}^{t}e^{-i(t-s)\mathcal{L}_{a}}F(x,u)ds\right\|_{L_{t}^{\bar{a}}L_{x}^{\bar{p}}}.

where212121Note that a¯2\bar{a}\geq 2 since α>1\alpha>1, which is important to conclude that (a¯,p¯)(\bar{a},\bar{p}) is SS-admissible. a¯\bar{a} and r¯\bar{r} are defined in Lemma 4.8. Thus, with the previous lemma in hand the rest the proof is exactly the same as in Theorem 1.9. ∎

5. Scattering and existence of wave operator

Our goal here is to show Theorem 1.12. It gives us a criterion to establish scattering and the existence of wave operator. To this end, we use the estimates on the nonlinearity obtained in Section 4. Before proving the theorem itself, we must point out that our estimates in Lemmas 4.4, 4.25, and 4.34 also hold if we replace the norms (in time) on the whole \mathbb{R} by a interval II\subset\mathbb{R}. To see this it is sufficient to observe that in all results the only estimates in time we used was the Hölder inequality. We will only prove the result by assuming the assumptions of Theorem 1.8 holds. The other cases are dealt with similarly.

Proof of Theorem 1.12.

Let’s first prove show (i). We claim that if uS(H˙sc)<+\|u\|_{S(\dot{H}^{s_{c}})}<+\infty, then

auS(L2)<+.\|\langle\sqrt{\mathcal{L}_{a}}\rangle u\|_{S(L^{2})}<+\infty. (5.1)

Indeed, we only show auS(L2;[0,))<+\|\langle\sqrt{\mathcal{L}_{a}}\rangle u\|_{S(L^{2};[0,\infty))}<+\infty. A similar analysis may be performed to see that auS(L2;(,0])<+\|\langle\sqrt{\mathcal{L}_{a}}\rangle u\|_{S(L^{2};(\infty,0])}<+\infty. Given δ>0\delta>0 we decompose the interval [0,)[0,\infty) into nn intervals Ij=[tj,tj+1)I_{j}=[t_{j},t_{j+1}) such that uS(H˙sc;Ij)<δ\|u\|_{S(\dot{H}^{s_{c}};I_{j})}<\delta, for all j=1,,nj=1,\ldots,n. On the time interval IjI_{j} we consider the integral equation

u(t)=ei(ttj)au(tj)+iλtjtei(ts)a(|x|b|u|αu)(s)𝑑s.u(t)=e^{-i(t-t_{j})\mathcal{L}_{a}}u(t_{j})+i\lambda\int_{t_{j}}^{t}e^{-i(t-s)\mathcal{L}_{a}}(|x|^{-b}|u|^{\alpha}u)(s)ds. (5.2)

Combining the Strichartz estimates (2.15) and (2.17) together with Lemma 4.4, it follows that

auS(L2;Ij)cu(tj)Ha1+ca(|x|b|u|αu)LIj2Lx2NN+2cu(tj)Ha1+cuLIjHx2θuS(H˙sc;Ij)αθuS(L2;Ij)cu(tj)Ha1+cMθδαθauS(L2;Ij),\begin{split}\|\langle\sqrt{\mathcal{L}_{a}}\rangle u\|_{S(L^{2};I_{j})}&\leq c\|u(t_{j})\|_{H^{1}_{a}}+c\left\|\langle\sqrt{\mathcal{L}_{a}}\rangle(|x|^{-b}|u|^{\alpha}u)\right\|_{L^{2}_{I_{j}}L_{x}^{\frac{2N}{N+2}}}\\ &\leq c\|u(t_{j})\|_{H^{1}_{a}}+c\|u\|^{\theta}_{L^{\infty}_{I_{j}}H^{2}_{x}}\|u\|^{\alpha-\theta}_{S(\dot{H}^{s_{c}};I_{j})}\|u\|_{S(L^{2};I_{j})}\\ &\leq c\|u(t_{j})\|_{H^{1}_{a}}+cM^{\theta}\delta^{\alpha-\theta}\|\langle\sqrt{\mathcal{L}_{a}}\rangle u\|_{S(L^{2};I_{j})},\end{split} (5.3)

where we have used the assumption suptIju(t)Ha1M\sup_{t\in I_{j}}\|u(t)\|_{H^{1}_{a}}\leq M. Taking δ>0\delta>0 such that ηθδαθ<12c\eta^{\theta}\delta^{\alpha-\theta}<\frac{1}{2c} we obtain that auS(L2;Ij)2cM,\|\langle\sqrt{\mathcal{L}_{a}}\rangle u\|_{S(L^{2};I_{j})}\leq 2cM, and by summing over the n intervals, we conclude the proof of (5.1).

Returning to the proof of the theorem, define

ϕ+=u0+iλ0+ei(s)a(|x|b|u|αu)(s)𝑑s.\phi^{+}=u_{0}+i\lambda\int\limits_{0}^{+\infty}e^{i(s)\mathcal{L}_{a}}\left(|x|^{-b}|u|^{\alpha}u\right)(s)ds.

Note that ϕ+Ha1(N)\phi^{+}\in H^{1}_{a}(\mathbb{R}^{N}). Following the above steps, one has

aϕ+L2cu0Ha1+cuLtHa1θuS(H˙sc)αθauS(L2).\|\langle\sqrt{\mathcal{L}_{a}}\rangle\phi^{+}\|_{L^{2}}\leq c\|u_{0}\|_{H^{1}_{a}}+c\|u\|^{\theta}_{L^{\infty}_{t}H^{1}_{a}}\|u\|^{\alpha-\theta}_{S(\dot{H}^{s_{c}})}\|\langle\sqrt{\mathcal{L}_{a}}\rangle u\|_{S(L^{2})}. (5.4)

Hence, applying (5.1) we get the desired result.

Since uu is a solution of (1.1), a simple inspection gives

u(t)eitaϕ+=iλt+ei(ts)a|x|b(|u|αu)(s)𝑑s,u(t)-e^{-it\mathcal{L}_{a}}\phi^{+}=-i\lambda\int\limits_{t}^{+\infty}e^{-i(t-s)\mathcal{L}_{a}}|x|^{-b}(|u|^{\alpha}u)(s)ds,

thus as before

u(t)eitaϕ+Ha1cuLtHa1θuS(H˙sc;[t,))αθauS(L2).\|u(t)-e^{-it\mathcal{L}_{a}}\phi^{+}\|_{H^{1}_{a}}\leq c\|u\|^{\theta}_{L^{\infty}_{t}H^{1}_{a}}\|u\|^{\alpha-\theta}_{S(\dot{H}^{s_{c}};[t,\infty))}\|\langle\sqrt{\mathcal{L}_{a}}\rangle u\|_{S(L^{2})}.

By using (5.1) we get uS(H˙sc;[t,))0\|u\|_{S(\dot{H}^{s_{c}};[t,\infty))}\rightarrow 0 as t+t\rightarrow+\infty, this implies that

u(t)eitaϕ+Ha10,ast+.\|u(t)-e^{-it\mathcal{L}_{a}}\phi^{+}\|_{H^{1}_{a}}\rightarrow 0,\,\,\textnormal{as}\,\,t\rightarrow+\infty. (5.5)

In the same way, let ϕ=u0+iλ0ei(s)a(|x|b|u|αu)(s)𝑑s\phi^{-}=u_{0}+i\lambda\int\limits_{0}^{-\infty}e^{i(s)\mathcal{L}_{a}}\left(|x|^{-b}|u|^{\alpha}u\right)(s)ds, so that we obtain ϕHa1\phi^{-}\in H^{1}_{a} and u(t)eitaϕ+Ha10,ast\|u(t)-e^{-it\mathcal{L}_{a}}\phi^{+}\|_{H^{1}_{a}}\rightarrow 0,\,\,\textnormal{as}\,\,t\rightarrow-\infty. This completes the proof of (i).

We now consider (ii). We will divide the proof in two parts. We first look for a fixed point for the operator

G(w)(t)=iλt+ei(ts)La(|x|b|w+eitLaϕ|α(w+eitLaϕ)(s)ds,tIT,G(w)(t)=-i\lambda\int_{t}^{+\infty}e^{-i(t-s)L_{a}}(|x|^{-b}|w+e^{-itL_{a}}\phi|^{\alpha}(w+e^{-itL_{a}}\phi)(s)ds,\;\;t\in I_{T}, (5.6)

where IT=[T,+)I_{T}=[T,+\infty) for T1T\gg 1. In the sequel, we show that uu defined by

u(t)=eitaϕ+w(t)u(t)=e^{-it\mathcal{L}_{a}}\phi+w(t)

is a solution of (1.1). To this end, let us start by proving that GG has a fixed point in S(T,ρ)S(T,\rho) given by

S(T,ρ)={wC(IT;Ha1(N)):wT:=wS(H˙sc;IT)+awS(L2;IT)ρ}.S(T,\rho)=\{w\in C\left(I_{T};H^{1}_{a}(\mathbb{R}^{N})\right):\;\|w\|_{T}:=\|w\|_{S(\dot{H}^{s_{c}};I_{T})}+\|\langle\sqrt{\mathcal{L}_{a}}\rangle w\|_{S(L^{2};I_{T})}\leq\rho\}.

The Strichartz estimates (2.17), (2.18) and Lemmas 4.4 yield

G(w)S(H˙sc;IT)\displaystyle\|G(w)\|_{S(\dot{H}^{s_{c}};I_{T})}\;\lesssim\; w+eitaϕLTHa1θw+eitaϕS(H˙sc;IT)αθw+eitaϕS(H˙sc;IT)\displaystyle\|w+e^{-it\mathcal{L}_{a}}\phi\|^{\theta}_{L^{\infty}_{T}H^{1}_{a}}\|w+e^{-it\mathcal{L}_{a}}\phi\|^{\alpha-\theta}_{S(\dot{H}^{s_{c}};I_{T})}\|w+e^{-it\mathcal{L}_{a}}\phi\|_{S(\dot{H}^{s_{c}};I_{T})} (5.7)

and

aG(w)S(L2;IT)\displaystyle\|\langle\sqrt{\mathcal{L}_{a}}\rangle G(w)\|_{S(L^{2};I_{T})}\;\lesssim\; w+eitaϕLTHa1θw+eitaϕS(H˙sc;IT)αθa(w+eitaϕ)S(L2;IT).\displaystyle\|w+e^{-it\mathcal{L}_{a}}\phi\|^{\theta}_{L^{\infty}_{T}H^{1}_{a}}\|w+e^{-it\mathcal{L}_{a}}\phi\|^{\alpha-\theta}_{S(\dot{H}^{s_{c}};I_{T})}\|\langle\sqrt{\mathcal{L}_{a}}\rangle(w+e^{-it\mathcal{L}_{a}}\phi)\|_{S(L^{2};I_{T})}. (5.8)

Hence,

G(w)T\displaystyle\|G(w)\|_{T} \displaystyle\lesssim w+eitaϕLTHx1θw+eitaϕS(H˙sc;IT)αθw+eitaϕT.\displaystyle\|w+e^{-it\mathcal{L}_{a}}\phi\|^{\theta}_{L^{\infty}_{T}H^{1}_{x}}\|w+e^{-it\mathcal{L}_{a}}\phi\|^{\alpha-\theta}_{S(\dot{H}^{s_{c}};I_{T})}\|w+e^{-it\mathcal{L}_{a}}\phi\|_{T}.

In view of222222Note that (5.9) might not be true in the norm LITLx2NN2scL^{\infty}_{I_{T}}L^{\frac{2N}{N-2s_{c}}}_{x} and for this reason we remove (,2NN2sc)\left(\infty,\frac{2N}{N-2s_{c}}\right) in the definition of H˙sc\dot{H}^{s_{c}}-admissible pair. More precisely, as we use Lemma 4.4 to the proof and we did not use this pair to prove it.

eitaϕS(H˙sc;IT)0\|e^{-it\mathcal{L}_{a}}\phi\|_{S(\dot{H}^{s_{c}};I_{T})}\rightarrow 0 (5.9)

as T+T\rightarrow+\infty, we can find T0>0T_{0}>0 large enough and ρ>0\rho>0 small enough such that GG is well defined on S(T0,ρ)S(T_{0},\rho). In the same way, we show that GG is a contraction on B(T0,ρ)B(T_{0},\rho). Therefore, there exists a unique wS(T0,ρ)w\in S(T_{0},\rho) such that G(w)=wG(w)=w.

On the other hand, by using the fact that w+eitaϕLTHa1wH1+ϕH1<+\|w+e^{-it\mathcal{L}_{a}}\phi\|_{L^{\infty}_{T}H^{1}_{a}}\leq\|w\|_{H^{1}}+\|\phi\|_{H^{1}}<+\infty and (5.7) we deduce

wS(H˙sc;IT)\displaystyle\|w\|_{S(\dot{H}^{s_{c}};I_{T})} \displaystyle\lesssim w+eitaϕS(H˙sc;IT)αθw+eitaϕS(H˙sc;IT)\displaystyle\|w+e^{-it\mathcal{L}_{a}}\phi\|^{\alpha-\theta}_{S(\dot{H}^{s_{c}};I_{T})}\|w+e^{-it\mathcal{L}_{a}}\phi\|_{S(\dot{H}^{s_{c}};I_{T})}
\displaystyle\lesssim AwS(H˙sc;IT)+AeitaϕS(H˙sc;IT),\displaystyle A\|w\|_{S(\dot{H}^{s_{c}};I_{T})}+A\|e^{-it\mathcal{L}_{a}}\phi\|_{S(\dot{H}^{s_{c}};I_{T})},

where A=w+eitaϕS(H˙sc;IT)αθA=\|w+e^{-it\mathcal{L}_{a}}\phi\|^{\alpha-\theta}_{S(\dot{H}^{s_{c}};I_{T})}. Observe that, if ρ\rho has been chosen small enough and since eitaϕS(H˙sc;IT)\|e^{-it\mathcal{L}_{a}}\phi\|_{S(\dot{H}^{s_{c}};I_{T})} is also sufficiently small for TT large, it follows that

AcwS(H˙sc;IT)αθ+ceitaϕS(H˙sc;IT)αθ<12.A\leq c\|w\|^{\alpha-\theta}_{S(\dot{H}^{s_{c}};I_{T})}+c\|e^{-it\mathcal{L}_{a}}\phi\|^{\alpha-\theta}_{S(\dot{H}^{s_{c}};I_{T})}<\frac{1}{2}.

Thus, using the last two inequalities we obtain 12wS(H˙sc;IT)AeitaϕS(H˙sc;IT),\frac{1}{2}\|w\|_{S(\dot{H}^{s_{c}};I_{T})}\lesssim A\|e^{-it\mathcal{L}_{a}}\phi\|_{S(\dot{H}^{s_{c}};I_{T})}, consequently

wS(H˙sc;IT)0asT+.\|w\|_{S(\dot{H}^{s_{c}};I_{T})}\rightarrow 0\;\;\;\;\textnormal{as}\;\;\;\;T\rightarrow+\infty. (5.10)

This also implies that 232323Here, we use the relations (5.8) and (5.10)\eqref{EWO5}. awS(L2;IT)0asT+,\|\langle\sqrt{\mathcal{L}_{a}}\rangle w\|_{S(L^{2};I_{T})}\rightarrow 0\;\;\;\;\textnormal{as}\;\;\;\;T\rightarrow+\infty, and so

wT0asT+.\|w\|_{T}\rightarrow 0\;\;\textnormal{as}\;\;T\rightarrow+\infty. (5.11)

We now prove that u(t)=eitaϕ+w(t)u(t)=e^{-it\mathcal{L}_{a}}\phi+w(t) satisfies (1.1) in the time interval [T0,)[T_{0},\infty). Indeed, we need to show that

u(t)=ei(tT0)au(T0)+iλT0tei(ts)a(|x|b|u|αu)s𝑑s,u(t)=e^{-i(t-T_{0})\mathcal{L}_{a}}u(T_{0})+i\lambda\int_{T_{0}}^{t}e^{-i(t-s)\mathcal{L}_{a}}(|x|^{-b}|u|^{\alpha}u)sds, (5.12)

for all t[T0,)t\in[T_{0},\infty). To do that, in view of

w(t)=iλtei(ts)a|x|b|w+eitaϕ|α(w+eitaϕ)(s)𝑑s,w(t)=-i\lambda\int_{t}^{\infty}e^{-i(t-s)\mathcal{L}_{a}}|x|^{-b}|w+e^{-it\mathcal{L}_{a}}\phi|^{\alpha}(w+e^{-it\mathcal{L}_{a}}\phi)(s)ds,

we have

ei(T0t)aw(t)\displaystyle e^{-i(T_{0}-t)\mathcal{L}_{a}}w(t) =\displaystyle= iλtei(T0s)a|x|b|w+eitaϕ|α(w+eitaϕ)(s)𝑑s\displaystyle-i\lambda\int_{t}^{\infty}e^{-i(T_{0}-s)\mathcal{L}_{a}}|x|^{-b}|w+e^{-it\mathcal{L}_{a}}\phi|^{\alpha}(w+e^{-it\mathcal{L}_{a}}\phi)(s)ds
=\displaystyle= iλT0tei(T0s)a|x|b|w+eitaϕ|α(w+eitaϕ)(s)𝑑s+w(T0),\displaystyle i\lambda\int_{T_{0}}^{t}e^{-i(T_{0}-s)\mathcal{L}_{a}}|x|^{-b}|w+e^{-it\mathcal{L}_{a}}\phi|^{\alpha}(w+e^{-it\mathcal{L}_{a}}\phi)(s)ds+w(T_{0}),

and so applying ei(tT0)ae^{-i(t-T_{0})\mathcal{L}_{a}} and adding eitaϕe^{-it\mathcal{L}_{a}}\phi on both sides, we obtain (5.12). Finally, adding eitaϕe^{-it\mathcal{L}_{a}}\phi to both sides of the last equation, we deduce (5.12).

Finally, since u(t)=eitaϕ+wu(t)=e^{-it\mathcal{L}_{a}}\phi+w and applying (5.11), we conclude that

u(t)eitaϕLTHx1=wLTHx1cwS(L2;IT)+cwS(L2;IT)cwT0asT,\displaystyle\|u(t)-e^{-it\mathcal{L}_{a}}\phi\|_{L^{\infty}_{T}H^{1}_{x}}=\|w\|_{L^{\infty}_{T}H^{1}_{x}}\leq c\|w\|_{S(L^{2};I_{T})}+c\|\nabla w\|_{S(L^{2};I_{T})}\leq c\|w\|_{T}\rightarrow 0\;\;\textnormal{as}\;\;T\rightarrow\infty,

which completes the proof.

References

  • [1] J. Belmonte-Beitia, V. M. Pérez-García, V. Vekslerchik, and P. J. Torres. Lie symmetries and solitons in nonlinear systems with spatially inhomogeneous nonlinearities. Physical review letters, 98(6):064102, 2007.
  • [2] J. Bourgain. Global solutions of nonlinear Schrödinger equations, volume 46 of American Mathematical Society Colloquium Publications. American Mathematical Society, Providence, RI, 1999.
  • [3] N. Burq, F. Planchon, J. G. Stalker, and A. S. Tahvildar-Zadeh. Strichartz estimates for the wave and Schrödinger equations with the inverse-square potential. J. Funct. Anal., 203(2):519–549, 2003.
  • [4] L. Campos. Scattering of radial solutions to the inhomogeneous nonlinear Schrödinger equation. Nonlinear Anal., 202:1–17, 2021.
  • [5] M. Cardoso, L. G. Farah, and C. M. Guzmán. On well-posedness and concentration of blow-up solutions for the intercritical inhomogeneous nls equation. arXiv preprint arXiv:2004.06706, 2020.
  • [6] T. Cazenave. Semilinear Schrödinger equations, volume 10 of Courant Lecture Notes in Mathematics. New York University, Courant Institute of Mathematical Sciences, New York; American Mathematical Society, Providence, RI, 2003.
  • [7] Y. Cho and M. Lee. On the orbital stability of inhomogeneous nonlinear Schrödinger equations with singular potential. Bull. Korean Math. Soc, 2019.
  • [8] V. D. Dinh. Scattering theory in a weighted l2l^{2} space for a class of the defocusing inhomogeneous nonlinear schrödinger equation. arXiv preprint arXiv:1710.01392, 2017.
  • [9] V. D. Dinh. Blowup of H1H^{1} solutions for a class of the focusing inhomogeneous nonlinear Schrödinger equation. Nonlinear Anal., 174:169–188, 2018.
  • [10] L. G. Farah. Global well-posedness and blow-up on the energy space for the inhomogeneous nonlinear Schrödinger equation. J. Evol. Equ., 16(1):193–208, 2016.
  • [11] L. G. Farah and C. M. Guzmán. Scattering for the radial 3D cubic focusing inhomogeneous nonlinear Schrödinger equation. J. Differential Equations, 262(8):4175–4231, 2017.
  • [12] L. G. Farah and C. M. Guzmán. Scattering for the radial focusing inhomogeneous NLS equation in higher dimensions. Bull Braz Math Soc, New Series (in press), 2019.
  • [13] D. Foschi. Inhomogeneous Strichartz estimates. J. Hyperbolic Differ. Equ., 2(1):1–24, 2005.
  • [14] F. Genoud. An inhomogeneous, L2L^{2}-critical, nonlinear Schrödinger equation. Z. Anal. Anwend., 31(3):283–290, 2012.
  • [15] F. Genoud and C. A. Stuart. Schrödinger equations with a spatially decaying nonlinearity: existence and stability of standing waves. Discrete Contin. Dyn. Syst., 21(1):137–186, 2008.
  • [16] C. M. Guzmán. On well posedness for the inhomogeneous nonlinear Schrödinger equation. Nonlinear Anal. Real World Appl., 37:249–286, 2017.
  • [17] H. Kalf, U.-W. Schmincke, J. Walter, and R. Wüst. On the spectral theory of Schrödinger and Dirac operators with strongly singular potentials. In Spectral theory and differential equations, pages 182–226. Lecture Notes in Math., Vol. 448, 1975.
  • [18] Y. V. Kartashov, B. A. Malomed, V. A. Vysloukh, M. R. Belic, and L. Torner. Rotating vortex clusters in media with inhomogeneous defocusing nonlinearity. Optic Letters., 42(3):446–449, 2017.
  • [19] R. Killip, C. Miao, M. Visan, J. Zhang, and J. Zheng. Sobolev spaces adapted to the Schrödinger operator with inverse-square potential. Math. Z., 288(3-4):1273–1298, 2018.
  • [20] F. Linares and G. Ponce. Introduction to nonlinear dispersive equations. Universitext. Springer, New York, second edition, 2015.
  • [21] J. Lu, C. Miao, and J. Murphy. Scattering in H1H^{1} for the intercritical NLS with an inverse-square potential. J. Differential Equations, 264(5):3174–3211, 2018.
  • [22] J. Lu, C. Miao, and J. Murphy. Scattering in H1H^{1} for the intercritical NLS with an inverse-square potential. J. Differential Equations, 264(5):3174–3211, 2018.
  • [23] N. Okazawa, T. Suzuki, and T. Yokota. Energy methods for abstract nonlinear Schrödinger equations. Evol. Equ. Control Theory, 1(2):337–354, 2012.
  • [24] W. A. Strauss. Existence of solitary waves in higher dimensions. Comm. Math. Phys., 55(2):149–162, 1977.
  • [25] T. Suzuki. Solvability of nonlinear Schrödinger equations with some critical singular potential via generalized Hardy-Rellich inequalities. Funkcial. Ekvac., 59(1):1–34, 2016.
  • [26] T. Tao. Nonlinear dispersive equations, volume 106 of CBMS Regional Conference Series in Mathematics. Published for the Conference Board of the Mathematical Sciences, Washington, DC; by the American Mathematical Society, Providence, RI, 2006. Local and global analysis.
  • [27] J. Zhang and J. Zheng. Scattering theory for nonlinear Schrödinger equations with inverse-square potential. J. Funct. Anal., 267(8):2907–2932, 2014.
  • [28] J. Zhang and J. Zheng. Global-in-time strichartz estimates and cubic schrodinger equation on metric cone. arXiv preprint arXiv:1702.05813, 2017.
  • [29] J. Zheng. Focusing NLS with inverse square potential. J. Math. Phys., 59(11):111502, 14, 2018.