1 Introduction

In the present paper we study well-posedness, asymptotic behavior of solutions and stability of traveling waves to the Cauchy problem of the system of the Schrödinger equations with derivative nonlinearities on the Euclidean space \(\mathbb {R}^d\):

$$\begin{aligned} {\left\{ \begin{array}{ll} i\partial _tu_1+\alpha \Delta u_1=-(\nabla \cdot u_3)u_2,&{} t\in I_{\max },\ x\in \mathbb {R}^d,\\ i\partial _tu_2+\beta \Delta u_2=-(\nabla \cdot \overline{u_3})u_1,&{} t\in I_{\max },\ x\in \mathbb {R}^d,\\ i\partial _tu_3+\gamma \Delta u_3=\nabla (u_1\cdot \overline{u_2}),&{} t\in I_{\max },\ x\in \mathbb {R}^d,\\ (u_1,u_2,u_3)|_{t=0}=(u_{1,0},u_{2,0},u_{3,0}),&{} x\in \mathbb {R}^d, \end{array}\right. } \end{aligned}$$
(1.1)

where \(d\in \{1,2,3\}\) is the spatial dimension, \(\partial _t:=\partial /\partial _t\) is the time derivative, \(\partial _{x_j}:=\partial /\partial _{x_j}\) is the spatial derivative with respect to the spatial variable \(x_j\) with \(j\in \{1,\cdots ,d\}\), \(\nabla :=(\partial _{x_1},\cdots ,\partial _{x_d})\) is the nabla, and \(\Delta :=\sum _{j=1}^d\partial _{x_j}^2\) is the Laplace operator on \(\mathbb {R}^d\). Here \(\alpha \), \(\beta \), \(\gamma \in \mathbb {R}\backslash \{0\}\) are real constants, \(I_{\max }:=(-T_{\min },T_{\max })\) is the maximal existence time interval, \(T_{\max }\in (0,\infty ]\) (resp. \(-T_{\min }\)) is the forward (resp.backward) existence time of the function \((u_1,u_2,u_3)\), \(u_1\), \(u_2\), \(u_3:I_{\max }\times \mathbb {R}^d\rightarrow \mathbb {C}^d\) are unknown d-dimensional complex vector-valued functions and \(u_{1,0}\), \(u_{2,0}\), \(u_{3,0}:\mathbb {R}^d\rightarrow \mathbb {C}^d\) are prescribed d-dimensional complex vector-valued functions. Let \(u_j^{(k)}:I_{\max }\times \mathbb {R}^d\rightarrow \mathbb {C}\) denote the k-th component of \(u_j\) for \(j=1,2,3\) and \(k=1,\cdots ,d\). Namely, \(u_j=(u_j^{(1)},\cdots , u_j^{(d)})\).

The system (1.1) was first derived from the bi-fluid Euler-Maxwell system by Colin and Colin [4, P301] as a model of laser-plasma interaction, in which \(u_1\) and \(u_3\) denote the scattered light and the electronic-plasma wave, respectively and \(u_2\) is the sum of the incident laser field and the gauged Brillouin component.

The system (1.1) is invariant under the following scale transformation

$$\begin{aligned} U_{\lambda }(t,x):=\lambda ^{-1}U(\lambda ^{-2}t,\lambda ^{-1}x),\ \ \ (U=(u_1,u_2,u_3),\ \lambda >0). \end{aligned}$$

More presicely if U is a solution to (1.1), then so is \(U_{\lambda }\) with the rescaled initial data \(\lambda ^{-1}U_0(\lambda ^{-1}x)\). Moreover, we calculate

$$\begin{aligned} \Vert U_{\lambda }(0)\Vert _{\dot{H}^s}=\lambda ^{\frac{d}{2}-1-s}\Vert U_0\Vert _{\dot{H}^s}, \end{aligned}$$

where \(\Vert \cdot \Vert _{\dot{H}^s}\) is the s-th order and \(L^2\)-based homogeneous Sobolev norm. If s satisfies \(s=d/2-1\), then \(\Vert U_{\lambda }(0)\Vert _{\dot{H}^s}=\Vert U_0\Vert _{\dot{H}^s}\) for any \(\lambda >0\). Therefore \(d/2-1\) is called the scaling critical Sobolev index and the current cases where \(d=1,2,3\) belong to the \(H^1\)(energy)-subcritical case.

For \(s\in \mathbb {R}\), we denote the inhomogeneous \(L^2\)-based s-th order Sobolev space by \(H^{s}=H^{s}(\mathbb {R}^d)\) endowed with the norm

$$\begin{aligned} \Vert f\Vert _{H^{s}}:=\left\| \langle \nabla \rangle ^sf\right\| _{L^2}=\left\| \langle \xi \rangle ^{s}\hat{f}\right\| _{L^{2}}, \end{aligned}$$

where \(\langle \cdot \rangle :=1+\vert \cdot \vert \). We introduce the function space \(\mathcal {H}^s:=(H^s(\mathbb {R}^d))^{3d}\) and give the initial data \((u_{1,0},u_{2,0},u_{3,0})\) belonging to \(\mathcal {H}^s\). In [4, 15], the well-posedness of (1.1) in \(\mathcal {H}^s\) was considered by using the energy method, and they obtained the local well-posedness for \(s>d/2+3\) under the condition \(\beta \gamma > 0\). The global well-posedness for sufficiently small initial data was also obtained in [15] if the condition \(\beta +\gamma \ne 0\) holds. On the other hand, the low regularity well-posedness of (1.1) in \(\mathcal {H}^s\) was studied in [11,12,13,14] by using the iteration argument in the space of the Fourier restriction norm. In particular, the well-posedness in the energy space \(\mathcal {H}^1\) was obtained as follows.

Theorem 1.1

(Local well-posedness in \(\mathcal {H}^1\), [11, 14]) Let \(d=1,2\), or 3. Assume that \(\alpha ,\beta ,\gamma \in \mathbb {R}\backslash \{0\}\) satisfy \((\alpha -\gamma )(\beta +\gamma )\ne 0\). Then, (1.1) is locally well-posed in \(\mathcal {H}^1\). The forward (resp. backward) maximal existence time \(T_{\max }\) (resp. \(T_{\min }\)) depends only on \(\Vert U_0\Vert _{\mathcal {H}^1}\). Furthermore, if \(\alpha \), \(\beta \), and \(\gamma \) have same sign, then the local solution for sufficiently small initial data in \(\mathcal {H}^1\) can be extended globally in time.

Remark 1.1

The global well-posedness in Theorem 1.1 was obtained by using the conservation law of the charge and the energy, which will be defined below.

Remark 1.2

For the case \(d=1\), which is \(L^2\)-subcritical case, we can obtain the global well-posedness in \(\mathcal {H}^1\) without smallness condition on initial data by using the Gagliardo-Nirenberg-Sobolev inequality. However for the case \(d=2\) or 3, the global well-posedness without smallness condition for initial data was not known.

Remark 1.3

For the case \(d=4\), which is \(H^1\)-critical case, the global well-posedness and the scattering of solution for small initial data in \(\mathcal {H}^1\) are obtained in [11]

As mentioned above remarks, there are only a few results for global solutions. In [4], the authors proposed open problems, which are existence and stability of solitary waves for the system (1.1). We address these problems in the case of \(d=1,2,3\) in this paper.

For vector valued functions \(f=(f^{(1)},\cdots ,f^{(d)})\) and \(g=(g^{(1)},\cdots ,g^{(d)})\in (H^1(\mathbb {R}^d))^d\), we define the derivatives, gradient, \(L^2\)-norm, \(H^1\)-norm, and \(L^2\)-inner product as

$$\begin{aligned} \begin{aligned}&\partial _lf:=(\partial _lf^{(1)},\cdots ,\partial _lf^{(d)})\in (L^2(\mathbb {R}^d))^d,\ \ \nabla f:=(\partial _1f,\cdots ,\partial _df)\in (L^2(\mathbb {R}^d))^{d\times d},\\&\Vert f\Vert _{L^2(\mathbb {R}^d)}^2:=\sum _{k=1}^d\Vert f^{(k)}\Vert _{L^2(\mathbb {R}^d)}^2,\ \ \Vert \nabla f\Vert _{L^2 (\mathbb {R}^d)}^2:=\sum _{l=1}^d\Vert \partial _l f\Vert _{L^2(\mathbb {R}^d)}^2 =\sum _{l=1}^d\sum _{k=1}^d\Vert \partial _lf^{(k)}\Vert _{L^2(\mathbb {R}^d)}^2,\\&\Vert f\Vert _{H^1(\mathbb {R}^d)}^2:=\Vert f\Vert _{L^2(\mathbb {R}^d)}^2+\Vert \nabla f\Vert _{L^2(\mathbb {R}^d)}^2,\ \ (f,g)_{L^2(\mathbb {R}^d)}:=\sum _{k=1}^d\int _{\mathbb {R}^d}f^{(k)}(x)\overline{g^{(k)}(x)}dx. \end{aligned} \end{aligned}$$

We introduce a new unknown function \(U:I_{\max }\times \mathbb {R}^d\rightarrow \mathbb {C}^d\times \mathbb {C}^d\times \mathbb {C}^d\) and a new initial data \(U_0:\mathbb {R}^d\rightarrow \mathbb {C}^d\times \mathbb {C}^d\times \mathbb {C}^d\) defined respectively by

$$\begin{aligned} U:=(u_1,u_2,u_3),\ \ \ \text {and}\ \ U_0:=(u_{1,0},u_{2,0},u_{3,0}). \end{aligned}$$

We define the function space \(\mathcal {H}^s:=(H^s(\mathbb {R}^d))^{3d}\) with the norm \(\Vert \cdot \Vert _{\mathcal {H}^s}\) as

$$\begin{aligned} \Vert U\Vert _{\mathcal {H}^s}^2:=\Vert u_1\Vert _{H^s(\mathbb {R}^d)}^2+\Vert u_2\Vert _{H^s(\mathbb {R}^d)}^2+\Vert u_3\Vert _{H^s(\mathbb {R}^d)}^2. \end{aligned}$$

We define the charge \(Q:(L^2(\mathbb {R}^d))^{3d}\rightarrow \mathbb {R}_{\ge 0}\) and the energy \(E:\mathcal {H}^1\rightarrow \mathbb {R}\) for the system (1.1) as

$$\begin{aligned} \begin{aligned} Q(U)&:=\Vert u_1\Vert _{L^2(\mathbb {R}^d)}^2+\frac{1}{2}\Vert u_2\Vert _{L^2(\mathbb {R}^d)}^2+\frac{1}{2}\Vert u_3\Vert _{L^2(\mathbb {R}^d)}^2,\\ E(U)&:=L(U)+N(U), \end{aligned} \end{aligned}$$

where L and N are the kinetic energy and the potential energy respectively defined by

$$\begin{aligned} \begin{aligned} L(U)&:= \frac{\alpha }{2}\Vert \nabla u_1\Vert _{L^2(\mathbb {R}^d)}^2+\frac{\beta }{2}\Vert \nabla u_2\Vert _{L^2(\mathbb {R}^d)}^2 +\frac{\gamma }{2}\Vert \nabla u_3\Vert _{L^2(\mathbb {R}^d)}^2,\\ N(U)&:= \textrm{Re}\left( u_3,\nabla (u_1\cdot \overline{u_2})\right) _{L^2(\mathbb {R}^d)}. \end{aligned} \end{aligned}$$

Furthermore, we define the momentum \(\textbf{P}:\mathcal {H}^1\rightarrow \mathbb {R}^d\) as

$$\begin{aligned} \textbf{P}(U):=\textbf{p}(u_1)+ \textbf{p}(u_2)+ \textbf{p}(u_3), \end{aligned}$$

where

$$\begin{aligned} \textbf{p}(f):=(P_1(f),\cdots P_d(f)),\ \ P_k(f):=-\frac{1}{2}\textrm{Re}(if,\partial _kf)_{L^2(\mathbb {R}^d)} \ \ (k=1,\cdots d) \end{aligned}$$

for \(f=(f_1,\cdots f_d)\in (H^1(\mathbb {R}^d))^d\). We note that Q, E, L, N are real valued and well-defined functionals on \(\mathcal {H}^1\) and \(\textbf{P}\) is vector valued and well-defined functional on \(\mathcal {H}^1\). In particular, Q(U), E(U), and \(\textbf{P}(U)\) are conserved quantities under the flow of (1.1). Namely, if U is a smooth and rapidly decaying solution to (1.1), then it holds that

$$\begin{aligned} Q(U(t))=Q(U_0),\ \ E(U(t))=E(U_0),\ \ \textbf{P}(U(t))=\textbf{P}(U_0) \end{aligned}$$

for all \(t\in I_{\max }\). For the conservation law of the charge and the energy, see Section 7 in [11]. The conservation law of the momentum is implied by the following calculation.

$$\begin{aligned} \begin{aligned} \frac{d}{dt}P_k(U)&=-\sum _{j=1}^3\textrm{Re}(i\partial _tu_j,\partial _ku_j)_{L^2(\mathbb {R}^d)}\\&=\textrm{Re}(\nabla \cdot u_3,\partial _ku_1\cdot \overline{u_2})_{L^2(\mathbb {R}^d)} +\textrm{Re}(\nabla \cdot \overline{u_3},\overline{u_1}\cdot \partial _ku_2)_{L^2(\mathbb {R}^d)} -\textrm{Re}(\nabla (u_1\cdot \overline{u_2}),\partial _ku_3)_{L^2(\mathbb {R}^d)}\\&=0 \end{aligned} \end{aligned}$$

To guarantee the definiteness of the kinetic energy L(U), we assume \(\alpha ,\beta ,\gamma >0\) or \(\alpha ,\beta ,\gamma <0\). Note that we can assume that \(\alpha >0\) without loss of generality. Indeed, if \((u_1,u_2,u_3)\) is a solution to (1.1), then \((\widetilde{u}_1,\widetilde{u}_2,\widetilde{u}_3)\) defined by \(\widetilde{u}_j(t,x):=u_j(-t,-x)\) \((j=1,2,3)\) is a solution to

$$\begin{aligned} {\left\{ \begin{array}{ll} i\partial _t\widetilde{u}_1-\alpha \Delta \widetilde{u}_1=-(\nabla \cdot \widetilde{u}_3)\widetilde{u}_2,&{}-t\in I_{\max },\ x\in \mathbb {R}^d,\\ i\partial _t\widetilde{u}_2-\beta \Delta \widetilde{u}_2=-(\nabla \cdot \overline{\widetilde{u}_3})\widetilde{u}_1,&{}-t\in I_{\max },\ x\in \mathbb {R}^d,\\ i\partial _t\widetilde{u}_3-\gamma \Delta \widetilde{u}_3=\nabla (\widetilde{u}_1\cdot \overline{\widetilde{u}_2}),&{}-t\in I_{\max },\ x\in \mathbb {R}^d,\\ (\widetilde{u}_1(0,x),\widetilde{u}_2(0,x),\widetilde{u}_3(0,x))=(u_{1,0}(-x),u_{2,0}(-x),u_{3,0}(-x)), &{}x\in \mathbb {R}^d. \end{array}\right. } \end{aligned}$$

If \(\alpha <0\), then it suffices to consider this Cauchy problem instead of (1.1). Therefore, we always consider the case \(\alpha ,\beta ,\gamma >0\) from here.

For \(\omega >0\) and \(\textbf{c}\in \mathbb {R}^d\), we define the functionals \(S_{\omega ,\textbf{c}}\), \(K_{\omega ,\textbf{c}}\), and \(L_{\omega ,\textbf{c}}\) on \(\mathcal {H}^1\) as

$$\begin{aligned} S_{\omega ,\textbf{c}}(U)&:=E(U)+\omega Q(U)+\textbf{c}\cdot \textbf{P}(U), \end{aligned}$$
(1.2)
$$\begin{aligned} K_{\omega ,\textbf{c}}(U)&:=\partial _{\lambda }S_{\omega ,\textbf{c}}(\lambda U)|_{\lambda =1}=2L(U)+3N(U)+2\omega Q(U)+2 \textbf{c}\cdot \textbf{P}(U), \end{aligned}$$
(1.3)
$$\begin{aligned} L_{\omega ,\textbf{c}}(U)&:= K_{\omega , \textbf{c}}(U)-3N(U)=2L(U)+2\omega Q(U)+2\textbf{c}\cdot \textbf{P}(U). \end{aligned}$$
(1.4)

We note that

$$\begin{aligned} S_{\omega ,\textbf{c}}(U)=\frac{1}{2}L_{\omega ,\textbf{c}}(U)+N(U)=\frac{1}{3}K_{\omega , \textbf{c}}(U)+\frac{1}{6}L_{\omega ,\textbf{c}}(U) \end{aligned}$$
(1.5)

and

$$\begin{aligned} N(U)=-2S_{\omega ,\textbf{c}}(U)+K_{\omega ,\textbf{c}}(U) \end{aligned}$$
(1.6)

hold. When \(d=1\), we write \(\textbf{P}(U)=P(U)\), \(\textbf{c}=c\), and \(\textbf{c}\cdot \textbf{P}(U)=cP(U)\).

We define the derivatives \(D_j\) of the functional \(S_{\omega ,\textbf{c}}\) as

$$\begin{aligned} \langle D_{j}S_{\omega ,\textbf{c}}(\Phi ),\psi \rangle :=\lim _{\epsilon \rightarrow 0}\frac{S_{\omega ,\textbf{c}}(\Phi +\epsilon E_j(\psi ))-S_{\omega ,\textbf{c}}(\Phi )}{\epsilon },\ \ \Phi \in \mathcal {H}^1,\ \psi \in (H^1(\mathbb {R}^d))^{d},\ j=1,2,3, \end{aligned}$$

where

$$\begin{aligned} E_1(\psi )=(\psi ,\textbf{0},\textbf{0}),\ E_2(\psi )=(\textbf{0},\psi ,\textbf{0}),\ E_3(\psi )=(\textbf{0},\textbf{0},\psi ) \in \mathcal {H}^1,\ \ \textbf{0}=(0,\cdots ,0)\in (H^1(\mathbb {R}^d))^d. \end{aligned}$$

By the simple calculation, we can see that

$$\begin{aligned} \begin{aligned} \langle D_{1}S_{\omega ,\textbf{c}}(\Phi ),\psi \rangle&=\textrm{Re}( -\alpha \Delta \varphi _1+2\omega \varphi _1+i(\textbf{c}\cdot \nabla )\varphi _1-(\nabla \cdot \varphi _3)\varphi _2,\psi )_{L^2(\mathbb {R}^d)},\\ \langle D_{2}S_{\omega ,\textbf{c}}(\Phi ),\psi \rangle&=\textrm{Re}( -\beta \Delta \varphi _2+\omega \varphi _2+i(\textbf{c}\cdot \nabla )\varphi _2-(\nabla \cdot \overline{\varphi _3})\varphi _1,\psi )_{L^2(\mathbb {R}^d)},\\ \langle D_{3}S_{\omega ,\textbf{c}}(\Phi ),\psi \rangle&=\textrm{Re}( -\gamma \Delta \varphi _3+\omega \varphi _3+i(\textbf{c}\cdot \nabla )\varphi _3+\nabla (\varphi _1\cdot \overline{\varphi _2}),\psi )_{L^2(\mathbb {R})^d} \end{aligned} \end{aligned}$$

for smooth \(\Phi =(\varphi _1,\varphi _2,\varphi _3)\) and smooth \(\psi \). Therefore, solution \(\Phi =(\varphi _1,\varphi _2,\varphi _3)\) to the elliptic system

$$\begin{aligned} {\left\{ \begin{array}{ll} -\alpha \Delta \varphi _1+2\omega \varphi _1+i(\textbf{c}\cdot \nabla )\varphi _1=(\nabla \cdot \varphi _3)\varphi _2,\\ -\beta \Delta \varphi _2+\omega \varphi _2+i(\textbf{c}\cdot \nabla )\varphi _2=(\nabla \cdot \overline{\varphi _3})\varphi _1,\\ -\gamma \Delta \varphi _3+\omega \varphi _3+i(\textbf{c}\cdot \nabla )\varphi _3=-\nabla (\varphi _1\cdot \overline{\varphi _2}) \end{array}\right. } \end{aligned}$$
(1.7)

satisfies

$$\begin{aligned} D_{1}S_{\omega ,\textbf{c}}(\Phi )= D_{2}S_{\omega ,\textbf{c}}(\Phi )= D_{3}S_{\omega ,\textbf{c}}(\Phi )=0. \end{aligned}$$
(1.8)

Definition 1.1

(Weak solution to the stationary problem) We say that a triple of functions \(\Phi =(\varphi _1,\varphi _2,\varphi _3)\in ~\mathcal {H}^1\) is a weak solution to the stationary problem (1.7) if it satisfies

$$\begin{aligned} {\left\{ \begin{array}{ll} \alpha ( \nabla \varphi _1,\nabla \psi )_{L^2(\mathbb {R}^d)} +2\omega (\varphi _1,\psi )_{L^2(\mathbb {R}^d)} +(i(\textbf{c}\cdot \nabla )\varphi _1,\psi )_{L^2(\mathbb {R}^d)} =((\nabla \cdot \varphi _3)\varphi _2,\psi )_{L^2(\mathbb {R}^d)},\\ \beta ( \nabla \varphi _2,\nabla \psi )_{L^2(\mathbb {R}^d)} +\omega (\varphi _2,\psi )_{L^2(\mathbb {R}^d)} +(i(\textbf{c}\cdot \nabla )\varphi _2,\psi )_{L^2(\mathbb {R}^d)} =((\nabla \cdot \overline{\varphi _3})\varphi _1,\psi )_{L^2(\mathbb {R}^d)},\\ \gamma ( \nabla \varphi _3,\nabla \psi )_{L^2(\mathbb {R}^d)} +\omega (\varphi _3,\psi )_{L^2(\mathbb {R}^d)} +(i(\textbf{c}\cdot \nabla )\varphi _3,\psi )_{L^2(\mathbb {R}^d)}= -(\nabla (\varphi _1 \cdot \overline{\varphi _2}),\psi )_{L^2(\mathbb {R}^d)} \end{array}\right. } \end{aligned}$$

for any \(\psi \in (H^1(\mathbb {R}^d))^d\), and then write \(S_{\omega ,\textbf{c}}'(\Phi )=0\).

We note that \(\Phi \in \mathcal {H}^1\) becomes a weak solution to (1.7) if and only if \(\Phi \) satisfies (1.8). In the next section, we will show that a weak solution \(\Phi \in \mathcal {H}^1\) becomes smooth and satisfies the elliptic system (1.7) under the suitable condition for \(\omega \) and \(\textbf{c}\) (\(\omega >\frac{\sigma |\textbf{c}|^2}{4}\)). (See Corollary 2.2 and Remark 2.1 below.) We define the sets of weak solutions to (1.7) as

$$\begin{aligned} \begin{aligned} \mathcal {E}_{\omega ,\textbf{c}}&:=\{\Psi \in \mathcal {H}^1\backslash \{(\textbf{0},\textbf{0},\textbf{0})\}\ |\ \ S_{\omega ,\textbf{c}}'(\Psi )=0\},\\ \mathcal {G}_{\omega ,\textbf{c}}&:=\{\Psi \in \mathcal {E}_{\omega ,\textbf{c}}\ |\ \ S_{\omega ,\textbf{c}}(\Psi )\le S_{\omega ,\textbf{c}}(\Theta )\ \text {for any}\ \Theta \in \mathcal {E}_{\omega ,\textbf{c}}\}. \end{aligned} \end{aligned}$$

We call an element \(\Psi \in \mathcal {G}_{\omega ,\textbf{c}}\) a ground state to the elliptic system (1.7).

Remark 1.4

If \(\Psi \in \mathcal {E}_{\omega ,\textbf{c}}\), then it holds \(K_{\omega ,\textbf{c}}(\Psi )=0\). Indeed, if \(\Psi =(\psi _1,\psi _2,\psi _3)\in \mathcal {E}_{\omega ,\textbf{c}}\), then we have

$$\begin{aligned} K_{\omega ,\textbf{c}}(\Psi )=\partial _{\lambda }S_{\omega ,\textbf{c}}(\lambda \Psi )\big |_{\lambda =1} =\sum _{j=1}^3\langle D_jS_{\omega ,\textbf{c}}(\Psi ),\psi _j\rangle =0. \end{aligned}$$

Remark 1.5

For \(\Psi \in \mathcal {H}^1\), we set \(\Psi _{\omega }\) for \(\omega >0\) as

$$\begin{aligned} \Psi _{\omega }(x):=\omega ^{\frac{1}{2}}\Psi (\omega ^{\frac{1}{2}}x). \end{aligned}$$

By a simple calculation, we can see that \(\Psi \in \mathcal {E}_{1,\frac{\textbf{c}}{\sqrt{\omega }}}\) is equivalent to \(\Psi _{\omega }\in \mathcal {E}_{\omega ,\textbf{c}}\) and it holds that

$$\begin{aligned} Q(\Psi _{\omega })=\omega ^{1-\frac{d}{2}}Q(\Psi ),\ \ E(\Psi _{\omega })=\omega ^{2-\frac{d}{2}}E(\Psi ),\ \ \textbf{P}(\Psi _{\omega })=\omega ^{\frac{3-d}{2}}\textbf{P}(\Psi ). \end{aligned}$$

These identities imply the equality

$$\begin{aligned} S_{\omega ,\textbf{c}}(\Psi _{\omega })=\omega ^{2-\frac{d}{2}}S_{1,\frac{\textbf{c}}{\sqrt{\omega }}}(\Psi ). \end{aligned}$$

Remark 1.6

For \(\theta \in \mathbb {R}\), we define an unitary operator \(\Lambda (\theta )\) as

$$\begin{aligned} \Lambda (\theta )\Phi :=(e^{2i\theta }\varphi _1, e^{i\theta }\varphi _2, e^{i\theta }\varphi _3). \end{aligned}$$

If we set \(U(t,x)=U_{\omega ,\textbf{c}}(t,x):=\Lambda (\omega t)\Phi (x-\textbf{c}t)\) for \(\Phi =(\phi _1,\phi _2,\phi _3)\in \mathcal {E}_{\omega ,\textbf{c}}\), then U is a solution to the system (1.1) with \(U_0(x)=\Phi (x)\). We call this solution U a two-parameter solitary wave solution.

We study the minimization problem

$$\begin{aligned} \mu _{\omega ,\textbf{c}}:= \inf \{S_{\omega ,\textbf{c}}(\Psi )\ |\ \Psi \in \mathcal {H}^1\backslash \{(\textbf{0},\textbf{0},\textbf{0})\},\ K_{\omega ,\textbf{c}}(\Psi )=0\} \end{aligned}$$

and define the set of minimizers

$$\begin{aligned} \begin{aligned} \mathcal {M}_{\omega ,\textbf{c}}:= \{\Psi \in \mathcal {H}^1\backslash \{(\textbf{0},\textbf{0},\textbf{0})\}\ |\ S_{\omega ,\textbf{c}}(\Psi )=\mu _{\omega ,\textbf{c}},\ K_{\omega ,\textbf{c}}(\Psi )=0\}. \end{aligned} \end{aligned}$$

We note that if \(S_{\omega ,\textbf{c}}(\Psi )<\mu _{\omega ,\textbf{c}}\) and \(K_{\omega ,\textbf{c}}(\Psi )=0\) hold, then \(\Psi =(\textbf{0},\textbf{0},\textbf{0})\) by the definition of \(\mu _{\omega ,\textbf{c}}\). Furthermore, for \(\eta >0\), we define a subset of \(\mathcal {M}_{\omega ,\textbf{c}}\) as

$$\begin{aligned} \begin{aligned} \mathcal {M}_{\omega ,\textbf{c}}^*(\eta ):= \{\Psi \in \mathcal {M}_{\omega ,\textbf{c}}\ |\ G(\Psi )\ge \eta \}, \end{aligned} \end{aligned}$$

where \(G:\mathcal {H}^1\rightarrow \mathbb {R}\) is defined by

$$\begin{aligned} G(\Psi ):=(4-2d)\omega Q(\Psi )+(3-d)\textbf{c}\cdot \textbf{P}(\Psi ). \end{aligned}$$
(1.9)

Then

$$\begin{aligned} \mathcal {M}_{\omega ,c}^*(\eta ) = {\left\{ \begin{array}{ll} \{\Psi \in \mathcal {M}_{\omega ,c}|\ 2\omega Q(\Psi )+2cP(\Psi )\ge \eta \},&{}\text {if}\ d=1,\\ \{\Psi \in \mathcal {M}_{\omega ,\textbf{c}}\ |\ \textbf{c}\cdot \textbf{P}(\Psi )\ge \eta \},&{}\text {if}\ d=2,\\ \emptyset ,&{}\text {if}\ d=3. \end{array}\right. } \end{aligned}$$

For \(\alpha ,\beta ,\gamma >0\), we introduce a positive number \(\sigma \) given by

$$\begin{aligned} \sigma :=\max \left\{ \frac{1}{2\alpha },\frac{1}{\beta },\frac{1}{\gamma }\right\} =\left( \min \{2\alpha ,\beta ,\gamma \}\right) ^{-1}>0. \end{aligned}$$

The main results in the present paper are the followings.

Theorem 1.2

Let \(d\in \{1,2,3\}\) and \(\alpha ,\beta ,\gamma >0\). We assume that \((\omega ,\textbf{c})\in \mathbb {R}\times \mathbb {R}^d\) satisfies \(\omega >\frac{\sigma |\textbf{c}|^2}{4}\). Then we have \(\mathcal {G}_{\omega ,\textbf{c}}=\mathcal {M}_{\omega ,\textbf{c}}\ne \emptyset \). Namely, there exists at least one of ground state.

According to Theorem 1.1, the local well-posedness of (1.1) in \(\mathcal {H}^1\) is obtained under the condition \((\alpha -\gamma )(\beta +\gamma )\ne 0\). We note that \(\beta +\gamma =0\) does not occur for \(\alpha ,\beta ,\gamma >0\).

Theorem 1.3

(Global well-posedness below the ground state level and uniform boundedness) Let \(d\in \{1,2,3\}\) and \(\alpha ,\beta ,\gamma >0\). We assume that \(\alpha -\gamma \ne 0\) and \((\omega ,\textbf{c})\in \mathbb {R}\times \mathbb {R}^d\) satisfies \(\omega >\frac{\sigma |\textbf{c}|^2}{4}\). If \(U_0\in \mathcal {H}^1\) satisfies

$$\begin{aligned} S_{\omega ,\textbf{c}}(U_0)<\mu _{\omega ,\textbf{c}},\ \ K_{\omega ,\textbf{c}}(U_0)>0, \end{aligned}$$

then \(I_{\max }=\mathbb {R}\), that is, the local solution \(U(t)\in \mathcal {H}^1\) to (1.1) with \(U(0)=U_0\) can be extended globally in time. Moreover for any \(t\in \mathbb {R}\), the estimate \(S_{\omega ,\textbf{c}}(U(t))<\mu _{\omega ,\textbf{c}}\) holds.

When \(d=2\), which is \(L^2\)-critical case, the above global well-posedness result depending on the parameters \(\omega , \textbf{c}\) can be written as the following form without the parameters.

Corollary 1.4

(Global well-posedness below the ground state without parameters in 2d) Let \(d=2\) and \(\alpha ,\beta ,\gamma >0\). We assume that \(\alpha -\gamma \ne 0\). If \(U_0\in \mathcal {H}^1\) satisfies

$$\begin{aligned} Q(U_0)<Q(\Phi )-E(\Phi ) \end{aligned}$$
(1.10)

for some \(\Phi \in \mathcal {M}_{1,\textbf{c}_0}\) and some \(\textbf{c}_0\in \mathbb {R}^2\) with \(|\textbf{c}_0|<\frac{2}{\sqrt{\sigma }}\), namely

$$\begin{aligned} Q(U_0)<\sup _{|\textbf{c}_0|<\frac{2}{\sqrt{\sigma }}} \sup _{\Phi \in \mathcal {M}_{1,\textbf{c}_0}}\left( Q(\Phi )-E(\Phi )\right) , \end{aligned}$$

then the local solution \(U(t)\in \mathcal {H}^1\) to (1.1) with \(U(0)=U_0\) can be extended globally in time.

Remark 1.7

When \(d=2\), the estimate \(Q(\Phi )-E(\Phi )>0\) holds for \(\Phi \in \mathcal {M}_{1,\textbf{c}_0}\). This can be seen as follows. For \(\Phi \in \mathcal {M}_{1,\textbf{c}_0}\), the estimate

$$\begin{aligned} \textbf{c}_0\cdot \textbf{P}(\Phi )=-2E(\Phi ) \end{aligned}$$

holds by Proposition 2.4 below. Therefore, we have

$$\begin{aligned} 0<\mu _{1,\textbf{c}_0} =E(\Phi )+Q(\Phi )+\textbf{c}_0\cdot \textbf{P}(\Phi ) =Q(\Phi )-E(\Phi ) \end{aligned}$$

because \(\mu _{1,\textbf{c}_0}>0\) by Proposition 3.2 below.

Remark 1.8

If \(\Phi \in \mathcal {M}_{1,\textbf{0}}\), then the condition (1.10) is equivalent to \(Q(U_0)<Q(\Phi )\) because \(\Phi \in \mathcal {M}_{1,\textbf{0}}\) with \(d=2\) satisfies \(E(\Phi )=0\) by Proposition 2.4 below.

Remark 1.9

Because the smallness condition for initial data is not necessary except (1.10) in Corollary 1.4, this result is the improvement of global result in Theorem 1.1 for \(d=2\). (See, also Remark 1.2.)

Next we state our result of orbital stability of the set \(\mathcal {M}_{\omega ,\textbf{c}}^*(\eta )\) for arbitrary \(\eta >0\), which solves one of the open problems proposed in [4, P301].

Theorem 1.5

(Stability of the set \(\mathcal {M}_{\omega ,\textbf{c}}^*(\eta )\)) Let \(d\in \{1,2\}\) and \(\alpha ,\beta ,\gamma >0\). We assume that \(\alpha -\gamma \ne 0\) and \((\omega ,\textbf{c})\in \mathbb {R}\times \mathbb {R}^d\) satisfies \(\omega >\frac{\sigma |\textbf{c}|^2}{4}\). For \(\eta >0\), if \(\mathcal {M}_{\omega ,\textbf{c}}^*(\eta )\ne \emptyset \) holds, then the set \(\mathcal {M}_{\omega ,\textbf{c}}^*(\eta )\) is orbitally stable. More precisely, for any \(\epsilon >0\), there exists \(\delta >0\) such that if \(U_0\in \mathcal {H}^1\) satisfies

$$\begin{aligned} \inf _{\Phi \in \mathcal {M}_{\omega ,\textbf{c}}^*(\eta )}\Vert U_0-\Phi \Vert _{\mathcal {H}^1}<\delta , \end{aligned}$$

then the solution U(t) to (1.1) with \(U(0)=U_0\) exists globally in time and satisfies

$$\begin{aligned} \sup _{t\ge 0}\inf _{\Phi \in \mathcal {M}_{\omega ,\textbf{c}}^*(\eta )}\Vert U(t)-\Phi \Vert _{\mathcal {H}^1}<\epsilon . \end{aligned}$$

Next we give orbital stability of the set \(\mathcal {M}_{\omega ,c}\) for 1-dimensional setting as a Corollary of Theorem 1.5.

Corollary 1.6

Let \(d=1\). Then the ground-state set \(\mathcal {M}_{\omega ,c}\) is orbitally stable if |c| is small enough. More precisely, for any \(\omega >0\), there exists \(c_0=c_0(\omega )\in (0,2\sqrt{\frac{\omega }{\sigma }})\) such that if \(|c|\le c_0\), then \(\mathcal {M}_{\omega ,c}\) is orbitally stable.

To prove the stability of the ground-states set, we will use the variational argument for the function \(\mu (\omega ,\textbf{c}):=\mu _{\omega ,\textbf{c}}\) with respect to \((\omega , \textbf{c})\), whose argument is used in [6]. In this argument, the second derivatives of \(\mu \) play an important role. However in our problem, there is a difficulty that the ground state \(\Phi \in \mathcal {M}_{\omega ,\textbf{c}}\) cannot be written explicitly. Therefore, we cannot calculate the second derivatives of \(\mu \) specifically. To avoid this difficulty, we use the argument by using a so-called scaling curve, which was used by Hayashi in [10]. Here the scaling curve is given as a map \(\tau \mapsto \left( (\sqrt{\omega }-\tau )^2,\frac{\textbf{c}}{\sqrt{\omega }} (\sqrt{\omega } -\tau )\right) \). By considering the restriction of \(\mu \) on the scaling curve, we can treat \(\mu \) such as a polynomial function and calculate the second derivative. Thanks to these properties and the Pohozaev’s identity (Proposition 2.4) below, we can give the sufficient condition for ground states to establish the stability. Furthermore, by checking the sufficient condition when \(d=1\) (see, Proposition 4.7), we can obtain the stability of the ground-states set \(\mathcal {M}_{\omega ,\textbf{c}}\) for small \(\textbf{c}\). In particular, our results contain the stability of the set of traveling waves.

We introduce similar results such as existence of ground state and stability of solitary wave for other Schrödinger type equations. Guo-Wu [9] and Colin-Ohta [6] studied orbital stability of a two-parameter family of solitary waves \(\{u_{\omega ,c}\}_{(\omega ,c)\in \mathbb {R}\times \mathbb {R}}\) to the single derivative nonlinear Schrödinger equation in one spatial dimension

$$\begin{aligned} i\partial _tu +\partial _x^2u+i\partial _x(|u|^2u)=0,\ \ \ (t,x)\in \mathbb {R}\times \mathbb {R}, \end{aligned}$$
(1.11)

where \(u_{\omega ,c}\) for \((\omega ,c)\in \mathbb {R}\times \mathbb {R}\) is the form of

$$\begin{aligned} u_{\omega ,c}(t,x):=\phi _{\omega ,c}(x-ct)\exp \left\{ i\omega t+i\frac{c}{2}(x-ct)-\frac{3i}{4}\int _{-\infty }^{x-ct}|\phi _{\omega ,c}(\eta )|^2d\eta \right\} . \end{aligned}$$

Here \(\phi _{\omega ,c}\) is the unique ground state of the elliptic equation

$$\begin{aligned} -\partial _{x}^2\phi +\left( \omega -\frac{c^2}{4}\right) \phi +\frac{c}{2}\phi ^3-\frac{3}{16}\phi ^5=0. \end{aligned}$$

If \(c^2<4\omega \), then \(\phi _{\omega ,c}\) shows an exponential decay and is explicitly written as

$$\begin{aligned} \phi _{\omega ,c}(x):=\left[ \frac{\sqrt{\omega }}{4\omega -c^2}\left\{ \cosh (\sqrt{(4\omega -c^2)x})-\frac{c}{\sqrt{4\omega }}\right\} \right] ^{-1/2} \end{aligned}$$

and the charge of \(\phi _{\omega ,c}\) is given by

$$\begin{aligned} \Vert \phi _{\omega ,c}\Vert _{L^2(\mathbb {R})}^2=8\tan ^{-1}\sqrt{\frac{\sqrt{4\omega }+c}{\sqrt{4\omega }-c}}<4\pi . \end{aligned}$$

The orbital stability of those solitons was proved in [9] for \(c<0\) and \(c^2<4\omega \) and in [6] for any \(c^2<4\omega \). See Kwon-Wu [16] for the endpoint case \(c^2=4\omega \). Fukaya-Hayashi-Inui [7] gave a sufficient condition for global existence of solutions to the generalized version of (1.11) by a variational argument (see also [2, 19]). For the system of nonlinear Schrödinger equations without derivative nonlinearities, see [1, 5, 8, 17] and their references. In particular, in [8], similar results as Theorems 1.2 and  1.3 are obtained for the following system of Schrödinger equations:

$$\begin{aligned} {\left\{ \begin{array}{ll} i\partial _tu+\frac{1}{2m}\Delta u=\overline{u}v,\\ i\partial _tv+\frac{1}{2M}\Delta v=u^2, \end{array}\right. } (t,x)\in \mathbb {R}\times \mathbb {R}^d, \end{aligned}$$
(1.12)

where \(m,M\in \mathbb {R}\backslash \{0\}\) are constants. In [8], the authors also proved the global well-posedness of (1.12) with \(d=4\) (which is \(L^2\)-critical case) for arbitrary large initial data with modification of oscillations.

We use the shorthand \(A\lesssim B\) to denote the estimate \(A\le CB\) with some constant \(C>0\). The notation \(A\sim B\) stands for \(A\lesssim B\) and \(B\lesssim A\). We give a notation table (Table 1 below) for reader’s convenience.

Table 1 Notation table

2 Properties of weak solutions to the stationary problem

In this section we study properties of weak solutions (not necessarily ground state) to stationary problem.

Proposition 2.1

Let \(d\in \{1,2,3\}\) and \(\alpha ,\beta ,\gamma >0\). Assume that \((\omega ,\textbf{c})\in \mathbb {R}\times \mathbb {R}^d\) satisfies \(\omega >\frac{\sigma |\textbf{c}|^2}{4}\). If \(\Phi =(\varphi _1,\varphi _2,\varphi _3)\in \mathcal {H}^1\) be a weak solution to (1.7), then \(\Phi \in (W^{2,p}(\mathbb {R}^d))^{3d}\) holds for any \(p\in [1,2]\), where \(W^{2,p}(\mathbb {R}^d)\) denotes the second order \(L^p\)-based Sobolev space, that is, \(W^{2,p}(\mathbb {R}^d):=\{f\in L^p(\mathbb {R}^d)\ |\ \sum _{|\alpha |\le 2}\Vert \partial _x^{\alpha }f\Vert _{L^p}<\infty \}\). In particular, \(\varphi _j \in (H^2(\mathbb {R}^d))^d\) \((j=1,2,3)\) holds.

Proof

We set \(p_0:=\max \{1,\frac{d}{2}\}\) and \(q_0:=2p_0/(2-p_0)\). Then, we have

$$\begin{aligned} \frac{1}{p_0}=\frac{1}{2}+\frac{1}{q_0}. \end{aligned}$$

We note that \(1\le p_0<2\), \(q_0\ge 2\) and

$$\begin{aligned} 1\ge \frac{d}{2} -\frac{d}{q_0} \end{aligned}$$

hold because \(d\le 3\). By the Hölder inequality and the Sobolev inequality, we have

$$\begin{aligned} \Vert (\nabla \cdot \varphi _3)\varphi _2\Vert _{L^{p_0}} \le \Vert \nabla \cdot \varphi _3\Vert _{L^2}\Vert \varphi _2\Vert _{L^{q_0}} \lesssim \Vert \varphi _3\Vert _{H^1}\Vert \varphi _2\Vert _{H^1}<\infty . \end{aligned}$$

By the first equation of (1.7), we obtain

$$\begin{aligned} \Vert \varphi _1\Vert _{W^{2,p_0}}= \Vert (-\alpha \Delta +2\omega +i\textbf{c}\cdot \nabla )^{-1} \{(\nabla \cdot \varphi _3)\varphi _2\}\Vert _{W^{2,p_0}} \lesssim \Vert (\nabla \cdot \varphi _3)\varphi _2\Vert _{L^{p_0}}<\infty \end{aligned}$$

because

$$\begin{aligned} \alpha |\xi |^2+2\omega -\textbf{c}\cdot \xi =\alpha \left| \xi -\frac{\textbf{c}}{2\alpha }\right| ^2+2\left( \omega -\frac{|\textbf{c}|^2}{8\alpha }\right) \sim 1+|\xi |^2 \end{aligned}$$

holds when \(\omega >\frac{\sigma |\textbf{c}|^2}{4}\). Therefore, we have \(\varphi _1\in (W^{2,p_0}(\mathbb {R}^d))^d\). By the same argument, we also obtain \(\varphi _2, \varphi _3\in (W^{2,p_0}(\mathbb {R}^d))^d\). Since \(p_0\ge \frac{d}{2}\), the relation \(W^{2,p_0}(\mathbb {R}^d)\subset L^q(\mathbb {R}^d)\) holds for any \(q\in [2,\infty ]\) by the Sobolev embedding theorem. This implies \(\Phi \in (L^q(\mathbb {R}^d))^{3d}\) for any \(q\in [2,\infty ]\).

For any \(p\in [1,2]\), we put

$$\begin{aligned} \frac{1}{q}:=\frac{1}{p}-\frac{1}{2}. \end{aligned}$$

We note that \(q\ge 2\). Therefore, we have \(\varphi _2 \in L^q(\mathbb {R}^d)\) and \( \Vert \varphi _1\Vert _{W^{2,p}} \lesssim \Vert (\nabla \cdot \varphi _3)\varphi _2\Vert _{L^{p}} <\infty \) by the same argument as above. Similarly, we also have \( \Vert \varphi _2\Vert _{W^{2,p}}<\infty ,\ \ \Vert \varphi _3\Vert _{W^{2,p}}<\infty . \)

Corollary 2.2

Let \(d\in \{1,2,3\}\) and \(\alpha ,\beta ,\gamma >0\). Assume that \((\omega ,\textbf{c})\in \mathbb {R}\times \mathbb {R}^d\) satisfies \(\omega >\frac{\sigma |\textbf{c}|^2}{4}\). If \(\Phi =(\varphi _1,\varphi _2,\varphi _3)\in \mathcal {H}^1\) be a weak solution to (1.7), then it holds

$$\begin{aligned} \varphi _1,\ \varphi _2,\ \varphi _3\in \left( \bigcap _{m=1}^{\infty }H^m(\mathbb {R}^d)\right) ^d. \end{aligned}$$

Proof

We first prove that if \(\varphi _j\in (H^m(\mathbb {R}^d))^d\) \((j=1,2,3)\) for some \(m\in \mathbb {N}\) with \(m\ge 2\), then \(\varphi _j \in (H^{m+1}(\mathbb {R}^d))^d\) \((j=1,2,3)\) holds. We note that \(H^m(\mathbb {R}^d)\) is Banach algebra since \(d<4\) and \(m\ge 2\). Therefore, by the third equation of (1.7) and the Hölder inequality, we have

$$\begin{aligned} \begin{aligned} \Vert \partial _k \varphi _3\Vert _{H^m}&=\Vert \partial _k (-\gamma \Delta +\omega +i\textbf{c} \cdot \nabla )^{-1}\nabla (\varphi _1\cdot \overline{\varphi _2})\Vert _{H^m} \lesssim \Vert \varphi _1 \cdot \varphi _2\Vert _{H^m} \lesssim \Vert \varphi _1\Vert _{H^m}\Vert \varphi _2\Vert _{H^m} <\infty \end{aligned} \end{aligned}$$

for \(k=1,\cdots , d\) because \(\omega >\frac{\sigma |\textbf{c}|^2}{4}\). This means that \(\varphi _3\in (H^{m+1}(\mathbb {R}^d))^d\). Next, the first equation of (1.7) and the Hölder inequality, we have

$$\begin{aligned} \begin{aligned} \Vert \partial _k \varphi _1\Vert _{H^m}&=\Vert \partial _k (-\alpha \Delta +2\omega +i\textbf{c} \cdot \nabla )^{-1}\{(\nabla \cdot \varphi _3)\varphi _2\}\Vert _{H^m} \lesssim \Vert (\nabla \cdot \varphi _3)\varphi _2\Vert _{H^m}\\&\quad \lesssim \Vert \varphi _3\Vert _{H^{m+1}}\Vert \varphi _2\Vert _{H^m} <\infty \end{aligned} \end{aligned}$$

for \(k=1,\cdots ,d\) because \(\omega >\frac{\sigma |\textbf{c}|^2}{4}\). This means that \(\varphi _1\in (H^{m+1}(\mathbb {R}^d))^d\). Similarly, we also have \(\varphi _2\in (H^{m+1}(\mathbb {R}^d))^d\).

By Proposition 2.1, we have \(\varphi _1\), \(\varphi _2, \varphi _3\in (H^2(\mathbb {R}^d))^d\). Therefore, we obtain \(\varphi _1\), \(\varphi _2, \varphi _3\in (H^m(\mathbb {R}^d))^d\) for any \(m\in \mathbb {N}\) by the induction. \(\square \)

Remark 2.1

By Corollary 2.2, the weak solution \(\Phi =(\varphi _1,\varphi _2,\varphi _3)\in \mathcal {H}^1\) is also a classical solution and derivatives of \(\Phi \) vanish at infinity because the embedding

$$\begin{aligned} H^{m}(\mathbb {R}^d) \subset \left\{ f\in C^r(\mathbb {R}^d)\left| \ \displaystyle \lim _{|x|\rightarrow \infty }\partial ^{\alpha }f(x)=0,\ \alpha \in \mathbb {Z}^d:\text {multi index with}\ |\alpha |\le r\right. \right\} \end{aligned}$$

holds for \(m>\frac{d}{2}+r\). In particular, \(\Phi \in (L^{\infty }(\mathbb {R}^d))^{3d}\) holds and \(\varphi _3\) satisfies

$$\begin{aligned} \lim _{|x|\rightarrow \infty }\nabla \cdot \varphi _3(x)=0. \end{aligned}$$
(2.1)

We can also obtain the exponential decay of the solution to (1.7).

Proposition 2.3

Let \(d\in \{1,2,3\}\) and \(\alpha ,\beta ,\gamma >0\). Assume that \((\omega ,\textbf{c})\in \mathbb {R}\times \mathbb {R}^d\) satisfies \(\omega >\frac{\sigma |\textbf{c}|^2}{4}\). If \((\varphi _1,\varphi _2,\varphi _3)\in \mathcal {H}^1\) be a weak solution to (1.7), then we have

$$\begin{aligned} \sum _{j=1}^3\int _{\mathbb {R}^d}e^{p|x|} \left( |\varphi _j(x)|^2+|\nabla \varphi _j (x)|^2\right) dx<\infty \end{aligned}$$
(2.2)

for

$$\begin{aligned} 0<p<\sqrt{4\omega \sigma _0}\left( 1-\sqrt{\frac{\sigma }{4\omega }}|c|\right) , \end{aligned}$$
(2.3)

where

$$\begin{aligned} \sigma _0:=\min \left\{ \frac{2}{\alpha },\frac{1}{\beta },\frac{1}{\gamma }\right\} . \end{aligned}$$

Proof

For \(\epsilon >0\), we define \(\theta _{\epsilon }:\mathbb {R}^d\rightarrow \mathbb {R}_{> 0}\) as

$$\begin{aligned} \theta _{\epsilon }(x):=\exp \left( \frac{p\sqrt{1+|x|^2}}{1+\epsilon \sqrt{1+|x|^2}}\right) . \end{aligned}$$

We note that \(\theta _{\epsilon }\) and \(\nabla \theta _{\epsilon }\) are bounded on \(\mathbb {R}^d\). Indeed, by a simple calculation, we can see that \(0\le \theta _{\epsilon }(x)\le e^{\frac{p}{\epsilon }}\) and

$$\begin{aligned} |\nabla \theta _{\epsilon }(x)|\le p\theta _{\epsilon }(x) \end{aligned}$$
(2.4)

for any \(x\in \mathbb {R}^d\). We define the real-valued inner product \((\cdot ,\cdot )\) on \((L^2(\mathbb {R}^d))^d\) as \((f,g):=\textrm{Re}(f,g)_{L^2(\mathbb {R}^d)}\) for \(f,g\in (L^2(\mathbb {R}^d))^d\). Because \((\varphi _1,\varphi _2,\varphi _3)\) be a weak solution to (1.7), we have \(\varphi _1\), \(\varphi _2\), \(\varphi _3\in \left( \bigcap _{m=1}^{\infty }H^m(\mathbb {R}^d)\right) ^d\) and

$$\begin{aligned} \begin{aligned} 0&=(-\alpha \Delta \varphi _1+2\omega \varphi _1+i(\textbf{c}\cdot \nabla ) \varphi _1 -(\nabla \cdot \varphi _3)\varphi _2,\theta _{\epsilon }\varphi _1),\\ 0&=(-\beta \Delta \varphi _2+\omega \varphi _2+i(\textbf{c}\cdot \nabla ) \varphi _2 -(\nabla \cdot \overline{\varphi _3})\varphi _1,\theta _{\epsilon }\varphi _2),\\ 0&=(-\gamma \Delta \varphi _3+\omega \varphi _3+i(\textbf{c}\cdot \nabla ) \varphi _3 +\nabla (\varphi _1\cdot \overline{\varphi _2}),\theta _{\epsilon }\varphi _3) \end{aligned} \end{aligned}$$

by Corollary 2.2. By using the integration by parts for the first terms of right hand side, these equations can be written as

$$\begin{aligned} \begin{aligned} 0&=\alpha (\nabla \varphi _1,\theta _{\epsilon }\nabla \varphi _1) +\alpha I_{1,1} +2\omega (\varphi _1,\theta _{\epsilon }\varphi _1) +I_{2,1} -J_1,\\ 0&=\beta (\nabla \varphi _2,\theta _{\epsilon }\nabla \varphi _2) +\beta I_{1,2} +\omega (\varphi _2,\theta _{\epsilon }\varphi _2) +I_{2,2} -J_2,\\ 0&=\gamma (\nabla \varphi _3,\theta _{\epsilon }\nabla \varphi _3) +\gamma I_{1,3} +\omega (\varphi _3,\theta _{\epsilon }\varphi _3) +I_{2,3} -J_3, \end{aligned} \end{aligned}$$

where

$$\begin{aligned} I_{1,j}= \sum _{k=1}^d(\nabla \varphi _{j}^{(k)}\cdot \nabla \theta _{\epsilon },\varphi _{j}^{(k)}),\ \ \ I_{2,j}=(i(\textbf{c}\cdot \nabla )\varphi _j,\theta _{\epsilon }\varphi _j),\ \ (\varphi _j=(\varphi _j^{(1)},\cdots ,\varphi _j^{(d)})) \end{aligned}$$

for \(j=1,2,3\) and

$$\begin{aligned} J_1:=((\nabla \cdot \varphi _3)\varphi _2,\theta _{\epsilon }\varphi _1),\ \ J_2:=((\nabla \cdot \overline{\varphi _3})\varphi _1,\theta _{\epsilon }\varphi _2),\ \ J_3:=-(\nabla (\varphi _1\cdot \overline{\varphi _2}),\theta _{\epsilon }\varphi _3). \end{aligned}$$

Then, we obtain

$$\begin{aligned} \begin{aligned} J_1&=\alpha \Vert \theta _{\epsilon }^{\frac{1}{2}}|\nabla \varphi _1|\Vert _{L^2}^2 +2\omega \Vert \theta _{\epsilon }^{\frac{1}{2}}|\varphi _1|\Vert _{L^2}^2 +\alpha I_{1,1}+I_{2,1},\\ J_2&=\beta \Vert \theta _{\epsilon }^{\frac{1}{2}}|\nabla \varphi _2|\Vert _{L^2}^2 +\omega \Vert \theta _{\epsilon }^{\frac{1}{2}}|\varphi _2|\Vert _{L^2}^2 +\beta I_{1,2}+I_{2,2},\\ J_3&=\gamma \Vert \theta _{\epsilon }^{\frac{1}{2}}|\nabla \varphi _3|\Vert _{L^2}^2 +\omega \Vert \theta _{\epsilon }^{\frac{1}{2}}|\varphi _3|\Vert _{L^2}^2 +\gamma I_{1,3}+I_{2,3}. \end{aligned} \end{aligned}$$

By (2.4) and the Young inequality, we have

$$\begin{aligned} |I_{1,j}|\le \sum _{k=1}^d\int _{\mathbb {R}^d}|\nabla \varphi _j^{(k)}(x)|p\theta _{\epsilon }(x)|\varphi _j^{(k)}(x)|dx \le p\left( \frac{\delta _j}{2}\Vert \theta _{\epsilon }^{\frac{1}{2}}|\nabla \varphi _j|\Vert _{L^2}^2 +\frac{1}{2\delta _j}\Vert \theta _{\epsilon }^{\frac{1}{2}}|\varphi _j|\Vert _{L^2}^2\right) \end{aligned}$$

and

$$\begin{aligned} |I_{2,j}|\le \int _{\mathbb {R}^d}|\textbf{c}||\nabla \varphi _j(x)|\theta _{\epsilon }(x)|\varphi _j(x)|dx \le |\textbf{c}|\left( \frac{\eta _j}{2}\Vert \theta _{\epsilon }^{\frac{1}{2}}|\nabla \varphi _j|\Vert _{L^2}^2 +\frac{1}{2\eta _j}\Vert \theta _{\epsilon }^{\frac{1}{2}}|\varphi _j|\Vert _{L^2}^2\right) , \end{aligned}$$

where \(\delta _j>0\), \(\eta _j>0\) \((j=1,2,3)\) will be chosen later. Therefore, we get

$$\begin{aligned} \begin{aligned} |J_1|&\ge \left( \alpha -\frac{\alpha p\delta _1}{2}-\frac{|\textbf{c}|\eta _1}{2}\right) \Vert \theta _{\epsilon }^{\frac{1}{2}}|\nabla \varphi _1|\Vert _{L^2}^2 +\left( 2\omega -\frac{\alpha p}{2\delta _1}-\frac{|\textbf{c}|}{2\eta _1}\right) \Vert \theta _{\epsilon }^{\frac{1}{2}}|\varphi _1|\Vert _{L^2}^2\\&\quad =:A_1\Vert \theta _{\epsilon }^{\frac{1}{2}}|\nabla \varphi _1|\Vert _{L^2}^2+B_1\Vert \theta _{\epsilon }^{\frac{1}{2}}|\varphi _1|\Vert _{L^2}^2,\\ |J_2|&\ge \left( \beta -\frac{\beta p\delta _2}{2}-\frac{|\textbf{c}|\eta _2}{2}\right) \Vert \theta _{\epsilon }^{\frac{1}{2}}|\nabla \varphi _2|\Vert _{L^2}^2 +\left( \omega -\frac{\beta p}{2\delta _2}-\frac{|\textbf{c}|}{2\eta _2}\right) \Vert \theta _{\epsilon }^{\frac{1}{2}}|\varphi _2|\Vert _{L^2}^2\\&=:A_2\Vert \theta _{\epsilon }^{\frac{1}{2}}|\nabla \varphi _2|\Vert _{L^2}^2+B_2\Vert \theta _{\epsilon }^{\frac{1}{2}}|\varphi _2|\Vert _{L^2}^2,\\ |J_3|&\ge \left( \gamma -\frac{\gamma p\delta _3}{2}-\frac{|\textbf{c}|\eta _3}{2}\right) \Vert \theta _{\epsilon }^{\frac{1}{2}}|\nabla \varphi _2|\Vert _{L^2}^2 +\left( \omega -\frac{\gamma p}{2\delta _3}-\frac{|\textbf{c}|}{2\eta _3}\right) \Vert \theta _{\epsilon }^{\frac{1}{2}}|\varphi _2|\Vert _{L^2}^2\\&=:A_3\Vert \theta _{\epsilon }^{\frac{1}{2}}|\nabla \varphi _3|\Vert _{L^2}^2+B_3\Vert \theta _{\epsilon }^{\frac{1}{2}}|\varphi _3|\Vert _{L^2}^2. \end{aligned} \end{aligned}$$

We choose \(\delta _1,\delta _2,\delta _3>0\) as

$$\begin{aligned} \delta _1=\sqrt{\frac{\alpha }{2\omega }},\ \ \delta _2=\sqrt{\frac{\beta }{\omega }},\ \ \delta _3=\sqrt{\frac{\gamma }{\omega }}. \end{aligned}$$

Then, we can see that

$$\begin{aligned} \frac{\alpha -\frac{\alpha p\delta _1}{2}}{\frac{|\textbf{c}|}{2}}>\frac{\frac{|\textbf{c}|}{2}}{2\omega -\frac{\alpha p}{2\delta _1}},\ \ \frac{\beta -\frac{\beta p\delta _2}{2}}{\frac{|\textbf{c}|}{2}}>\frac{\frac{|\textbf{c}|}{2}}{\omega -\frac{\beta p}{2\delta _2}},\ \ \frac{\gamma -\frac{\gamma p\delta _3}{2}}{\frac{|\textbf{c}|}{2}}>\frac{\frac{|\textbf{c}|}{2}}{\omega -\frac{\gamma p}{2\delta _3}}. \end{aligned}$$

by (2.3). Therefore, we can choose \(\eta _1,\eta _2,\eta _3>0\) such that \(A_1,A_2,A_3>0\) and \(B_1,B_2,B_3>0\).

On the other hand, by (2.1), there exists \(R>0\) such that for any \(x\in \mathbb {R}^d\) with \(|x|>R\),

$$\begin{aligned} |\nabla \cdot \varphi _3(x)|\le \delta M \end{aligned}$$
(2.5)

holds, where \(M:=\min _{1\le j\le 3}\{A_j,B_j\}>0\) and \(\delta >0\) is small constant. For \(j=1,2\), we divide the integral in the definition of \(J_j\) as

$$\begin{aligned} \begin{aligned} |J_j|&\le \int _{\mathbb {R}^d}|\nabla \cdot \varphi _3|\theta _{\epsilon }|\varphi _1||\varphi _2|dx =\int _{|x|\ge R}|\nabla \cdot \varphi _3|\theta _{\epsilon }|\varphi _1||\varphi _2|dx\\&\quad +\int _{|x|<R}|\nabla \cdot \varphi _3|\theta _{\epsilon }|\varphi _1||\varphi _2|dx =:J_{\infty }+J_{0}. \end{aligned} \end{aligned}$$

For the first term, by (2.5) and the Young inequality, we obtain

$$\begin{aligned} J_{\infty } \le \delta M\int _{|x|\ge R}\theta _{\epsilon }|\varphi _1||\varphi _2|dx \le \frac{\delta M}{2}\left( \Vert \theta _{\epsilon }^{\frac{1}{2}}|\varphi _1|\Vert _{L^2}^2 +\Vert \theta _{\epsilon }^{\frac{1}{2}}|\varphi _2|\Vert _{L^2}^2\right) . \end{aligned}$$

For the second term, by the Hölder inequality and the Sobolev inequality, there exists \(C>0\) such that

$$\begin{aligned} J_{0}\le \Vert \theta _{\epsilon }\Vert _{L^{\infty }(|x|<R)} \Vert \nabla \cdot \varphi _3\Vert _{L^2}\Vert \varphi _1\Vert _{L^4}\Vert \varphi _2\Vert _{L^4} \le C\theta _{\epsilon }(R)\prod _{j=1}^3\Vert \varphi _j\Vert _{H^1}. \end{aligned}$$

For \(j=3\), we note that by the integration by parts,

$$\begin{aligned} \begin{aligned} J_3&=(\varphi _1\cdot \overline{\varphi _2},\nabla \cdot (\theta _{\epsilon }\varphi _3)) = (\varphi _1\cdot \overline{\varphi _2},\theta _{\epsilon }\nabla \cdot \varphi _3) +(\varphi _1\cdot \overline{\varphi _2},(\nabla \theta _{\epsilon })\cdot \varphi _3)\\&=(\theta _{\epsilon }\varphi _1,(\nabla \cdot \varphi _3)\varphi _2) +(\varphi _1\cdot \overline{\varphi _2},(\nabla \theta _{\epsilon })\cdot \varphi _3) \\&=J_1+(\varphi _1\cdot \overline{\varphi _2},(\nabla \theta _{\epsilon })\cdot \varphi _3). \end{aligned} \end{aligned}$$

For the second term, we have

$$\begin{aligned} \begin{aligned} |(\varphi _1\cdot \overline{\varphi _2},(\nabla \theta _{\epsilon })\cdot \varphi _3)|&\le p\int _{\mathbb {R}^d}\theta _{\epsilon }|\varphi _1||\varphi _2||\varphi _3|dx \\&\quad =p\left( \int _{|x|\ge R}\theta _{\epsilon }|\varphi _1||\varphi _2||\varphi _3|dx +\int _{|x|<R}\theta _{\epsilon }|\varphi _1||\varphi _2||\varphi _3|dx\right) \\&\le \frac{p\delta M}{2}\left( \Vert \theta _{\epsilon }^{\frac{1}{2}}|\varphi _1|\Vert _{L^2}^2 +\Vert \theta _{\epsilon }^{\frac{1}{2}}|\varphi _2|\Vert _{L^2}^2\right) +pC\theta _{\epsilon }(R)\prod _{j=1}^3\Vert \varphi _j\Vert _{H^1} \end{aligned} \end{aligned}$$

by (2.4) and the same argument as for \(j=1,2\) because \(\varphi _3\) satisfies

$$\begin{aligned} \lim _{|x|\rightarrow \infty }\varphi _3(x)=0. \end{aligned}$$

As a result, we obtain

$$\begin{aligned} \begin{aligned}&M\sum _{j=1}^3\left( \Vert \theta _{\epsilon }^{\frac{1}{2}}|\nabla \varphi _j|\Vert _{L^2}^2+\Vert \theta _{\epsilon }^{\frac{1}{2}}|\varphi _j|\Vert _{L^2}^2\right) \le |J_1|+|J_2|+|J_3|\\&\le \frac{3+p}{2}\delta M \left( \Vert \theta _{\epsilon }^{\frac{1}{2}}|\varphi _1|\Vert _{L^2}^2 +\Vert \theta _{\epsilon }^{\frac{1}{2}}|\varphi _2|\Vert _{L^2}^2\right) +(3+p)C\theta _{\epsilon }(R)\prod _{j=1}^3\Vert \varphi _j\Vert _{H^1}. \end{aligned} \end{aligned}$$

By choosing \(\delta =\frac{1}{3+p}\), we have

$$\begin{aligned} \sum _{j=1}^3\left( \Vert \theta _{\epsilon }^{\frac{1}{2}}|\nabla \varphi _j|\Vert _{L^2}^2+\Vert \theta _{\epsilon }^{\frac{1}{2}}|\varphi _j|\Vert _{L^2}^2\right) \le 2(3+p)C\theta _{\epsilon }(R)\prod _{j=1}^3\Vert \varphi _j\Vert _{H^1}. \end{aligned}$$

Let \(\epsilon \rightarrow +0\) and by using the monotone convergence theorem, the desired estimate (2.2) is obtained. \(\square \)

Remark 2.2

Because the solution \(\Phi =(\varphi _1,\varphi _2,\varphi _3)\) to (1.7) satisfies

$$\begin{aligned} \sum _{j=1}^3|\Delta \varphi _j|^2 \lesssim \sum _{j=1}^3\left( |\varphi _j|^2+|\nabla \varphi _j|^2 +\Vert \Phi \Vert _{L^{\infty }}^2|\nabla \varphi _j|^2\right) \end{aligned}$$

We also obtain

$$\begin{aligned} \sum _{j=1}^3\int _{\mathbb {R}^d}e^{p|x|} |\Delta \varphi _j (x)|^2dx<\infty \end{aligned}$$

by using (2.2).

The following proposition (Pohozaev’s identity) plays an important role to prove the stability in Sect. 4.

Proposition 2.4

(Pohozaev’s identity) Let \(d\in \{1,2,3\}\) and \(\alpha ,\beta ,\gamma >0\). Assume that \((\omega ,\textbf{c})\in \mathbb {R}\times \mathbb {R}^d\) satisfies \(\omega >\frac{\sigma |\textbf{c}|^2}{4}\). If \(\Phi =(\varphi _1,\varphi _2,\varphi _3)\in \mathcal {H}^1\) is a weak solution to (1.7), then \(\Phi \) satisfies

$$\begin{aligned} 2L(\Phi )+\left( \frac{d}{2}+1\right) N(\Phi ) +\textbf{c}\cdot \textbf{P}(\Phi )=0. \end{aligned}$$

Proof

For \(\lambda >0\), we set

$$\begin{aligned} \Phi ^{\lambda }(x):=\lambda ^{\frac{d}{2}}\Phi (\lambda x). \end{aligned}$$
(2.6)

Then, we have

$$\begin{aligned} \frac{d}{d\lambda }S_{\omega ,\textbf{c}}(\Phi ^{\lambda })\big |_{\lambda =1} =\sum _{j=1}^3\left. \left\langle (D_jS_{\omega ,\textbf{c}})(\Phi ^{\lambda }), \frac{d}{d\lambda }\Phi ^{\lambda }\right\rangle \right| _{\lambda =1} =\sum _{j=1}^3\left\langle D_jS_{\omega ,\textbf{c}}(\Phi ), \frac{d}{2}\Phi +(x\cdot \nabla ) \Phi \right\rangle . \end{aligned}$$

Because \(\Phi \) be a weak solution to (1.7), \(\nabla \varphi _j\) and \(\Delta \varphi _j\) \((j=1,2,3)\) are decaying exponentially as \(|x|\rightarrow \infty \) by Proposition 2.3 (and Remark 2.2). Namely, \(\frac{d}{2}\Phi +(x\cdot \nabla ) \Phi \in \mathcal {H}^1\) holds. Therefore, we obtain

$$\begin{aligned} \frac{d}{d\lambda }S_{\omega ,\textbf{c}}(\Phi ^{\lambda })\big |_{\lambda =1} =0. \end{aligned}$$

On the other hand, we can easily see that

$$\begin{aligned} L(\Phi ^{\lambda })=\lambda ^2L(\Phi ),\ \ N(\Phi ^{\lambda })=\lambda ^{\frac{d}{2}+1}N(\Phi ),\ \ Q(\Phi ^{\lambda })=Q(\Phi ),\ \ \textbf{P}(\Phi ^{\lambda })=\lambda \textbf{P}(\Phi ). \end{aligned}$$

It implies that

$$\begin{aligned} \begin{aligned} \frac{d}{d\lambda }S_{\omega ,\textbf{c}}(\Phi ^{\lambda })|_{\lambda =1}&=\frac{d}{d\lambda }\left. \left( L(\Phi ^{\lambda })+N(\Phi ^{\lambda })+\omega Q(\Phi ^{\lambda }) +\textbf{c}\cdot \textbf{P}(\Phi ^{\lambda })\right) \right| _{\lambda =1}\\&=2L(\Phi )+\left( \frac{d}{2}+1\right) N(\Phi ) +\textbf{c}\cdot \textbf{P}(\Phi ) \end{aligned} \end{aligned}$$

and we get the desired identity. \(\square \)

3 Existence of a ground state to the stationary problem

In this section we give a proof of existence of a ground state to (1.7), which addresses a open problem proposed by Colin-Colin [4].

We prove coercivity of the functional \(L_{\omega , \textbf{c}}\) given by (1.4), which plays an important role to prove Proposition 3.2 and Lemma 4.6 below.

Proposition 3.1

(Coecivity of the functional \(L_{\omega ,\textbf{c}}\)) Let \(d\in \{1,2,3\}\) and \(\alpha ,\beta ,\gamma >0\). Assume that \((\omega ,\textbf{c})\in \mathbb {R}\times \mathbb {R}^d\) satisfies \(\omega >\frac{\sigma |\textbf{c}|^2}{4}\). Then, there exists a constant \(C=C_{\omega }>0\) independent of \(\textbf{c}\) such that the estimate

$$\begin{aligned} L_{\omega , \textbf{c}}(U)\ge C(4\omega -\sigma |\textbf{c}|^2)\Vert U\Vert _{\mathcal {H}^1}^2 \end{aligned}$$

holds for any \(U\in \mathcal {H}^1\), where \(L_{\omega , \textbf{c}}\) is defined by (1.4).

Proof

Let \(U=(u_1,u_2,u_3)\in \mathcal {H}^1\) and \(\textbf{c}=(c_1,\cdots ,c_d)\in \mathbb {R}^d\). For parameters \(A_1,A_2,A_3>0\), we set

$$\begin{aligned} I_{j,k}:=\frac{1}{A_j} \left\| A_j\partial _ku_j-\frac{c_k}{4}iu_j\right\| _{L^2(\mathbb {R}^d)}^2\ \ (j=1,2,3,\ k=1,\cdots , d). \end{aligned}$$

By a direct computation, for \(j\in \{1,2,3\}\), the following identities hold:

$$\begin{aligned} \textbf{c}\cdot \textbf{p}(u_j) =-\frac{1}{2}\sum _{k=1}^dc_k\textrm{Re}(iu_j,\partial _ku_j)_{L^2(\mathbb {R}^d)} =\sum _{k=1}^dI_{j,k}-A_j\Vert \nabla u_j\Vert _{L^2(\mathbb {R}^d)}^2 -\frac{|\textbf{c}|^2}{16A_j}\Vert u_j\Vert _{L^2(\mathbb {R}^d)}^2. \end{aligned}$$

This implies that by (1.4), the following holds:

$$\begin{aligned} \begin{aligned} L_{\omega , \textbf{c}}(U)&=(\alpha -2A_1)\Vert \nabla u_1\Vert _{L^2(\mathbb {R}^d)}^2 +(\beta -2A_2)\Vert \nabla u_2\Vert _{L^2(\mathbb {R}^d)}^2 +(\gamma -2A_3)\Vert \nabla u_3\Vert _{L^2(\mathbb {R}^d)}^2\\&\ \ \ \ +\left( 2\omega -\frac{|\textbf{c}|^2}{8A_1}\right) \Vert u_1\Vert _{L^2(\mathbb {R}^d)}^2 +\left( \omega -\frac{|\textbf{c}|^2}{8A_2}\right) \Vert u_2\Vert _{L^2(\mathbb {R}^d)}^2 +\left( \omega -\frac{|\textbf{c}|^2}{8A_3}\right) \Vert u_3\Vert _{L^2(\mathbb {R}^d)}^2\\&\ \ \ \ +\sum _{k=1}^d2(I_{1,k}+I_{2,k}+I_{3,k}). \end{aligned} \end{aligned}$$

Because \(\omega >\frac{\sigma |\textbf{c}|^2}{4}\), there exist \(A_1,A_2,A_3>0\) such that

$$\begin{aligned} \begin{aligned}&\alpha -2A_1>0,\ \ \beta -2A_2>0,\ \ \gamma -2A_3>0,\\&2\omega -\frac{|\textbf{c}|^2}{8A_1}>0,\ \ \omega -\frac{|\textbf{c}|^2}{8A_2}>0,\ \ \omega -\frac{|\textbf{c}|^2}{8A_3}>0. \end{aligned} \end{aligned}$$
(3.1)

Indeed, if we choose \(A_1,A_2,A_3>0\) as

$$\begin{aligned} A_1=\frac{1}{4}\left( \alpha +\frac{|\textbf{c}|^2}{8\omega }\right) ,\ \ A_2=\frac{1}{4}\left( \beta +\frac{|\textbf{c}|^2}{4\omega }\right) ,\ \ A_3=\frac{1}{4}\left( \gamma +\frac{|\textbf{c}|^2}{4\omega }\right) , \end{aligned}$$

then (3.1) holds and we can get

$$\begin{aligned} L_{\omega ,\textbf{c}}(U) \gtrsim (4\omega -\sigma |\textbf{c}|^2)\Vert U\Vert _{\mathcal {H}^1}^2, \end{aligned}$$

where the implicit constant does not depend on \(\textbf{c}\).

Next we prove positivity of \(\mu _{\omega ,\textbf{c}}\) and a property of the Nehari functional \(K_{\omega ,\textbf{c}}\) under \(\omega >\frac{\sigma |\textbf{c}|^2}{4}\).

Proposition 3.2

Let \(d\in \{1,2,3\}\) and \(\alpha ,\beta ,\gamma >0\). Assume that \((\omega ,\textbf{c})\in \mathbb {R}\times \mathbb {R}^d\) satisfies \(\omega >\frac{\sigma |\textbf{c}|^2}{4}\). Then, the following properties hold:

  1. (i)

    \(\mu _{\omega ,\textbf{c}}>0\).

  2. (ii)

    If \(U\in \mathcal {H}^1\backslash \{(\textbf{0},\textbf{0},\textbf{0})\}\) satisfies \(K_{\omega ,\textbf{c}}(U)<0\) (resp \(\le 0)\), then \(L_{\omega ,\textbf{c}}(U)>6\mu _{\omega ,\textbf{c}}\) (resp \(\ge 6\mu _{\omega ,\textbf{c}})\).

Proof

We first prove (i). By (1.5) and the definition of \(\mu _{\omega ,\textbf{c}}\), it holds

$$\begin{aligned} \mu _{\omega ,\textbf{c}}=\frac{1}{6}\inf \{L_{\omega ,\textbf{c}}(\Psi )|\ \Psi \in \mathcal {H}^1\backslash \{(\textbf{0},\textbf{0},\textbf{0})\},\ K_{\omega ,\textbf{c}}(\Psi )=0\}. \end{aligned}$$
(3.2)

Therefore, it suffices to show that there exists \(C_1>0\) such that

$$\begin{aligned} L_{\omega ,\textbf{c}}(\Psi )\ge C_1 \end{aligned}$$

holds for any \(\Psi =(\psi _1,\psi _2,\psi _3)\in \mathcal {H}^1\backslash \{(\textbf{0},\textbf{0},\textbf{0})\}\) with \(K_{\omega ,\textbf{c}}(\Psi )=0\). If \(K_{\omega ,\textbf{c}}(\Psi )=0\), then by the definition of \(L_{\omega ,\textbf{c}}\), the Hölder inequality, and the Sobolev inequality, we have

$$\begin{aligned} L_{\omega ,\textbf{c}}(\Psi )=-3N(\Psi ) \lesssim \Vert \nabla \cdot \psi _3\Vert _{L^2}\Vert \psi _1\Vert _{L^4}\Vert \psi _2\Vert _{L^4} \lesssim \Vert \Psi \Vert _{\mathcal {H}^1}^3. \end{aligned}$$

Furthermore by Proposition 3.1, we obtain

$$\begin{aligned} \Vert \Psi \Vert _{\mathcal {H}^1}^2\lesssim L_{\omega ,\textbf{c}}(\Psi )\lesssim \Vert \Psi \Vert _{\mathcal {H}^1}^3. \end{aligned}$$

This implies that there exists \(\widetilde{C_1}>0\) such that \(\Vert \Psi \Vert _{\mathcal {H}^1}\ge \widetilde{C_1}\) for any \(\Psi \in \mathcal {H}^1\backslash \{(\textbf{0},\textbf{0},\textbf{0})\}\) with \(K_{\omega ,\textbf{c}}(\Psi )=0\). By setting \(C_1:=\widetilde{C_1}^2C(4\omega -\sigma |\textbf{c}|^2)\), where \(C=C_{\omega }>0\) appears in Proposition 3.1, we get \(L_{\omega ,\textbf{c}}(\Psi )\ge C_1\).

Next, we prove (ii). We assume \(U\in \mathcal {H}^1\backslash \{(\textbf{0},\textbf{0},\textbf{0})\}\) and \(K_{\omega ,\textbf{c}}(U)<0\). For \(\lambda \in \mathbb {R}\), by a simple calculation, we have

$$\begin{aligned} K_{\omega ,\textbf{c}}(\lambda U) =L_{\omega ,\textbf{c}}(\lambda U)+3N(\lambda U) =\lambda ^2L_{\omega ,\textbf{c}}(U)+3\lambda ^3N(U). \end{aligned}$$

Therefore, if we set \(\lambda \) as

$$\begin{aligned} \lambda :=-\frac{L_{\omega ,\textbf{c}}(U)}{3N(U)}, \end{aligned}$$

then \(K_{\omega ,\textbf{c}}(\lambda U)=0\) holds. We can see that \(\lambda \in (0,1)\) from the fact that \(L_{\omega ,\textbf{c}}(U)+3N(U)=K_{\omega ,\textbf{c}}(U)<0\) and \(L_{\omega ,\textbf{c}}(U)>0\). As a result, by using (3.2), we get

$$\begin{aligned} \mu _{\omega ,\textbf{c}}\le \frac{1}{6}L_{\omega ,\textbf{c}}(\lambda U) =\frac{\lambda ^2}{6}L_{\omega ,\textbf{c}}(U)<\frac{1}{6}L_{\omega ,\textbf{c}}(U). \end{aligned}$$

If we assume \(K_{\omega ,\textbf{c}}(U)\le 0\) instead of \(K_{\omega ,\textbf{c}}(U)<0\), then there is a possibility that \(\lambda =1\) holds. Therefore, we can only get \(\mu _{\omega ,\textbf{c}}\le \frac{1}{6}L_{\omega ,\textbf{c}}(U)\).

Proposition 3.3

(Existence of a minimizer of \(\mathcal {M}_{\omega ,\textbf{c}}\)) Let \(d\in \{1,2,3\}\) and \(\alpha ,\beta ,\gamma >0\). Assume that \((\omega ,\textbf{c})\in \mathbb {R}\times \mathbb {R}^d\) satisfies \(\omega >\frac{\sigma |\textbf{c}|^2}{4}\). Let \(\{U_n\}\subset \mathcal {H}^1\) be a sequence satisfying

$$\begin{aligned} \lim _{n\rightarrow \infty }S_{\omega ,\textbf{c}}(U_n)=\mu _{\omega ,\textbf{c}},\ \ \lim _{n\rightarrow \infty }K_{\omega ,\textbf{c}}(U_n)=0. \end{aligned}$$

Then, there exists \(\{y_n\}\subset \mathbb {R}^d\) such that the sequence \(\{U_n(\cdot -y_n)\}\) has a subsequence which converges to some \(V\in \mathcal {H}^1\backslash \{(\textbf{0},\textbf{0},\textbf{0})\}\) strongly in \(\mathcal {H}^1\). Furthermore it holds that \(V\in \mathcal {M}_{\omega ,\textbf{c}}\). Namely \(\mathcal {M}_{\omega ,\textbf{c}}\) is not empty.

In addition, if \(\{U_n\}\) satisfies

$$\begin{aligned} \limsup _{n\rightarrow \infty }G(U_n)\ge \eta \end{aligned}$$
(3.3)

for some \(\eta >0\), where G is defined by (1.9), then it holds that \(V\in \mathcal {M}_{\omega ,\textbf{c}}^*(\eta )\).

We remark that the latter statement in this proposition will be applied to prove Theorem 1.5 in Sect. 4.2.

To prove Proposition 3.3, we use the following Lieb’s compactness theorem.

Lemma 3.4

(Lieb’s compactness theorem [18]) Let \(\{F_n\}\) be a bounded sequence in \(\mathcal {H}^1\). Assume that there exists \(p\in (0,2^*)\) such that \(\limsup _{n\rightarrow \infty }\Vert F_n\Vert _{L^p}>0\). Then there exist \(\{y_n\}\subset \mathbb {R}^d\) and \(F\in \mathcal {H}^1\backslash \{(\textbf{0},\textbf{0},\textbf{0})\}\) such that \(\{F_n(\cdot -y_n)\}\) has a subsequence that converges to F weakly in \(\mathcal {H}^1\), where \(2^*=\infty \) if \(d=1\) or 2, \(2^*=\frac{2d}{d-2}\) if \(d\ge 3\).

Lemma 3.5

(Special case of Brezis–Lieb’s lemma [3]) Assume that \(\{f_n\}\) be a bounded sequence in \(L^2(\mathbb {R}^d)\) and \(f_n\rightharpoonup f\) weakly in \(L^2(\mathbb {R}^d)\). Then it holds that

$$\begin{aligned} \Vert f_n\Vert _{L^2}^2-\Vert f_n-f\Vert _{L^2}^2-\Vert f\Vert _{L^2}^2\ \rightarrow \ 0. \end{aligned}$$

Remark 3.1

Lemma 3.5 also holds if we replace \(L^2(\mathbb {R}^d)\) by \((L^2(\mathbb {R}^d))^d\).

Proof of Proposition 3.3

We divide the proof into 6 steps as follows.

Step 1. We prove the boundedness of \(\{U_n\}\) in \(\mathcal {H}^1\). By (1.5), we have

$$\begin{aligned} L_{\omega ,\textbf{c}}(U_n)=6S_{\omega ,\textbf{c}}(U_n)-2K_{\omega ,\textbf{c}}(U_n) \ \rightarrow \ 6\mu _{\omega ,\textbf{c}}. \end{aligned}$$
(3.4)

Thus \(\{L_{\omega ,\textbf{c}}(U_n)\}\) is a bounded sequence. This and Proposition 3.1 imply that \(\{U_n\}\) is bounded in \(\mathcal {H}^1\).

Step 2. We prove \(\limsup _{n\rightarrow \infty }\Vert U_n\Vert _{L^4}>0\). We write \(U_n=(u_{1n},u_{2n},u_{3n})\) (\(u_{jn}\in (H^1(\mathbb {R}^d))^d\)). By Step 1, there exists \(C>0\) such that \(\Vert U_n\Vert _{\mathcal {H}^1}\le C\) holds for any \(n\in \mathbb {N}\). Therefore, by the Hölder inequality, we have

$$\begin{aligned} N(U_n)=\left| \textrm{Re}(u_{3n},\nabla (u_{1n}\cdot \overline{u_{2n}})_{L^2}\right| \le \Vert \nabla \cdot u_{3n}\Vert _{L^2}\Vert u_{1n}\Vert _{L^4}\Vert u_{2n}\Vert _{L^4} \le C\Vert U_n\Vert _{L^4}^2. \end{aligned}$$

Assume \(\lim _{n\rightarrow \infty }\Vert U_n\Vert _{L^4}=0\). Then we obtain \(\lim _{n\rightarrow \infty }N(U_n)=0\). This implies that

$$\begin{aligned} L_{\omega ,\textbf{c}}(U_n)=K_{\omega ,\textbf{c}}(U_n)-3N(U_n)\ \rightarrow \ 0. \end{aligned}$$

This contradicts to (3.4) because \(\mu _{\omega ,\textbf{c}}>0\) by Proposition 3.2 (i). Therefore we get \(\lim _{n\rightarrow \infty }\Vert U_n\Vert _{L^4}\ne 0\). Namely,

$$\begin{aligned} \limsup _{n\rightarrow \infty }\Vert U_n\Vert _{L^4}>0. \end{aligned}$$

Step 3. In this step, we apply Lemma 3.4. By Step 1 and Step 2, \(\{U_n\}\) satisfies the assumptions in Lemma 3.4 for \(p=4\). Therefore, there exist \(\{y_n\}\subset \mathbb {R}^d\) and \(V\in \mathcal {H}^1\backslash \{(\textbf{0},\textbf{0},\textbf{0})\}\) such that \(\{U_n(\cdot -y_n)\}\) has a subsequence \(\{V_{n}\}\) that converges to V weakly in \(\mathcal {H}^1\). We set

$$\begin{aligned} \begin{aligned}&V_n=(v_{1n},v_{2n},v_{3n}):=U_n(\cdot -y_n)\ \ (v_{jn}\in (H^1(\mathbb {R}^d))^d),\\&V=(v_{1},v_{2},v_{3})\ \ (v_{j}\in (H^1(\mathbb {R}^d))^d). \end{aligned} \end{aligned}$$

We can assume \(V_n\ne V\) for any \(n\in \mathbb {N}\) without loss of generality.

Step 4. We prove

$$\begin{aligned} L_{\omega ,\textbf{c}}(V_n)-L_{\omega ,\textbf{c}}(V_n-V)-L_{\omega ,\textbf{c}}(V)\ \rightarrow \ 0 \end{aligned}$$
(3.5)

and

$$\begin{aligned} K_{\omega ,\textbf{c}}(V_n)-K_{\omega ,\textbf{c}}(V_n-V)-K_{\omega ,\textbf{c}}(V)\ \rightarrow \ 0. \end{aligned}$$
(3.6)

First, we show (3.5). As in the proof of Proposition 3.1, we can write

$$\begin{aligned} \begin{aligned} L_{\omega ,\textbf{c}}(V)&=\sum _{j=1}^3\left( C_{1,j}\Vert \nabla v_j\Vert _{L^2}^2+C_{2,j}\Vert v_j\Vert _{L^2}^2+ \sum _{k=1}^dC_{3,j,k}\Vert \partial _kv_j+c_{j,k}v_j\Vert _{L^2}^2\right) \\&=\sum _{j=1}^3\left( C_{1,j}\sum _{k=1}^d\Vert \partial _k v_j\Vert _{L^2}^2+C_{2,j}\Vert v_j\Vert _{L^2}^2+ \sum _{k=1}^dC_{3,j,k}\Vert \partial _kv_j+c_{j,k}v_j\Vert _{L^2}^2\right) \end{aligned} \end{aligned}$$
(3.7)

for some positive constants \(C_{1,j}\), \(C_{2,j}\), \(C_{3,j,k}\) and constants \(c_{j,k}\in \mathbb {C}\). Because \(V_n\rightharpoonup V\) weakly in \(\mathcal {H}^1\), the sequences \(\{v_{jn}\}\) and \(\{\partial _kv_{jn}\}\) are bounded in \((L^2(\mathbb {R}^d))^d\), and it holds that

$$\begin{aligned} \partial _k v_{jn}\rightharpoonup \partial _k v_j,\ \ \ \ v_{jn}\rightarrow v_j\ \ \ \ \textrm{weakly}\ \textrm{in}\ L^2(\mathbb {R}^d)^d \end{aligned}$$
(3.8)

by taking subsequences. Therefore, by applying Lemma 3.5, we get (3.5).

Next, we show (3.6). We can write

$$\begin{aligned} 2N(V)=2\textrm{Re}(v_3,\nabla (v_1\cdot v_2))_{L^2} =\Vert \nabla \cdot v_3-v_1\cdot \overline{v_2}\Vert _{L^2}^2 -\Vert \nabla \cdot v_3\Vert _{L^2}^2-\Vert v_1\cdot \overline{v_2}\Vert _{L^2}^2. \end{aligned}$$
(3.9)

Because \(V_n\rightarrow V\) weakly in \(\mathcal {H}^1\), the sequences \(\{v_{1n}\}\), \(\{v_{2n}\}\), and \(\{v_{3n}\}\) are bounded in \((H^1(\mathbb {R}^d))^d\). This implies that the sequences \(\{\nabla \cdot v_{3n}\}\) and \(\{v_{1n}\cdot \overline{v_{2n}}\}\) are bounded in \(L^2(\mathbb {R}^d)\) because

$$\begin{aligned} \Vert \nabla \cdot v_{3n}\Vert _{L^2}\le \Vert v_{3n}\Vert _{H^1} \end{aligned}$$

and

$$\begin{aligned} \Vert v_{1n}\cdot \overline{v_{2n}}\Vert _{L^2} \le \Vert v_{1n}\Vert _{L^4}\Vert v_{2n}\Vert _{L^4} \le \Vert v_{1n}\Vert _{H^1}\Vert v_{2n}\Vert _{H^1} \end{aligned}$$

hold by the Hölder inequality and the Sobolev inequality. Therefore, by taking subsequences, we obtain

$$\begin{aligned} \nabla \cdot v_{3n}\rightharpoonup \nabla \cdot v_{3},\ \ \ \ v_{1n}\cdot \overline{v_{2n}}\rightharpoonup v_1\cdot \overline{v_2}\ \ \ \ \textrm{weakly}\ \textrm{in}\ L^2(\mathbb {R}^d). \end{aligned}$$

Furthermore, by (3.9) and applying Lemma 3.5, we have

$$\begin{aligned} N(V_n)-N(V_n-V)-N(V)\ \rightarrow \ 0. \end{aligned}$$
(3.10)

Because \(K_{\omega ,\textbf{c}}(U)=L_{\omega ,\textbf{c}}(U)+3N(U)\), we get (3.6) by (3.5) and (3.10).

Step 5. In this step, we prove \(K_{\omega ,\textbf{c}}(V)\le 0\) by contradiction. We assume \(K_{\omega ,\textbf{c}}(V)>0\). (Then, \(V\ne 0\) by the definition of \(K_{\omega ,\textbf{c}}\).) Because

$$\begin{aligned} K_{\omega ,\textbf{c}}(V_n)=K_{\omega ,\textbf{c}}(U_n)\rightarrow 0, \end{aligned}$$

we have

$$\begin{aligned} \begin{aligned} K_{\omega ,\textbf{c}}(V_n-V)&=-K_{\omega ,\textbf{c}}(V)+K_{\omega ,\textbf{c}}(V_n) -(K_{\omega ,\textbf{c}}(V_n)-K_{\omega ,\textbf{c}}(V_n-V)-K_{\omega ,\textbf{c}}(V))\\&\rightarrow -K_{\omega ,\textbf{c}}(V)<0 \end{aligned} \end{aligned}$$

by (3.6). Therefore, if \(n\in \mathbb {N}\) is large enough, then it holds \(K_{\omega ,\textbf{c}}(V_n-V)<0\). Since \(V_n-V\ne 0\), this and Proposition 3.2 (ii) imply that \(L_{\omega ,\textbf{c}}(V_n-V)>6\mu _{\omega ,\textbf{c}}\) for sufficiently large \(n\in \mathbb {N}\). Because by (3.4),

$$\begin{aligned} L_{\omega ,\textbf{c}}(V_n)=L_{\omega ,\textbf{c}}(U_n)\rightarrow 6\mu _{\omega ,\textbf{c}} \end{aligned}$$
(3.11)

holds and we obtain

$$\begin{aligned} \begin{aligned} L_{\omega ,\textbf{c}}(V)&=L_{\omega ,\textbf{c}}(V_n)-L_{\omega ,\textbf{c}}(V_n-V) -(L_{\omega ,\textbf{c}}(V_n)-L_{\omega ,\textbf{c}}(V_n-V)-L_{\omega ,\textbf{c}}(V))\\&<L_{\omega ,\textbf{c}}(V_n)-6\mu _{\omega ,\textbf{c}} -(L_{\omega ,\textbf{c}}(V_n)-L_{\omega ,\textbf{c}}(V_n-V)-L_{\omega ,\textbf{c}}(V))\\&\rightarrow 0 \end{aligned} \end{aligned}$$

by (3.5). But this leads to a contradiction because by Proposition 3.1 and \(V\ne 0\),

$$\begin{aligned} L_{\omega ,\textbf{c}}(V)\gtrsim \Vert V\Vert _{\mathcal {H}^1}^2>0 \end{aligned}$$

holds.

Step 6. Finally. we prove \(V_n\rightarrow V\) strongly in \(\mathcal {H}^1\) and \(V\in \mathcal {M}_{\omega ,\textbf{c}}\). By Step 5 and Proposition 3.2 (ii), we have \(L_{\omega , \textbf{c}}(V)\ge 6\mu _{\omega ,\textbf{c}}\). On the other hand, by (3.7) and (3.8), we obtain

$$\begin{aligned} \begin{aligned} L_{\omega ,\textbf{c}}(V)&\le \sum _{j=1}^3\left( C_{1,j}\sum _{k=1}^d \liminf _{n\rightarrow \infty }\Vert \partial _k v_{jn}\Vert _{L^2}^2+C_{2,j} \liminf _{n\rightarrow \infty }\Vert v_{jn}\Vert _{L^2}^2+ \sum _{k=1}^dC_{3,j,k} \liminf _{n\rightarrow \infty }\Vert \partial _kv_{jn}+c_{j,k}v_{jn}\Vert _{L^2}^2\right) \\&\le \lim _{n\rightarrow \infty } \sum _{j=1}^3\left( C_{1,j}\sum _{k=1}^d \Vert \partial _k v_{jn}\Vert _{L^2}^2+C_{2,j} \Vert v_{jn}\Vert _{L^2}^2+ \sum _{k=1}^dC_{3,j,k} \Vert \partial _kv_{jn}+c_{j,k}v_{jn}\Vert _{L^2}^2\right) \\&= \lim _{n\rightarrow \infty }L_{\omega ,\textbf{c}}(V_n). \end{aligned} \end{aligned}$$

Therefore, we get \(L_{\omega ,\textbf{c}}(V)\le 6\mu _{\omega ,\textbf{c}}\) by (3.11). As a result, we have \(L_{\omega ,\textbf{c}}(V)= 6\mu _{\omega ,\textbf{c}}\). Furthermore, by Proposition 3.1, (3.5), and (3.11), we obtain

$$\begin{aligned} \begin{aligned} \Vert V_n-V\Vert _{\mathcal {H}^1}^2&\lesssim L_{\omega ,\textbf{c}}(V_n-V)\\&=L_{\omega ,\textbf{c}}(V_n)-L_{\omega ,\textbf{c}}(V) -(L_{\omega ,\textbf{c}}(V_n)-L_{\omega ,\textbf{c}}(V_n-V)-L_{\omega ,\textbf{c}}(V))\\&\rightarrow 0. \end{aligned} \end{aligned}$$

Namely \(V_n\rightarrow V\) strongly in \(\mathcal {H}^1\). Because \(L_{\omega ,\textbf{c}}(V)= 6\mu _{\omega ,\textbf{c}}\) (In particular \(L_{\omega ,\textbf{c}}(V)\le 6\mu _{\omega ,\textbf{c}}\)), we have \(K_{\omega ,\textbf{c}}(V)\ge 0\) by Proposition 3.2 (ii). This and Step 5 imply \(K_{\omega ,\textbf{c}}(V)=0\). Thus, we obtain

$$\begin{aligned} S_{\omega ,\textbf{c}}(V) =\frac{1}{3}K_{\omega ,\textbf{c}}(V)+\frac{1}{6}L_{\omega ,\textbf{c}}(V)=\mu _{\omega ,\textbf{c}} \end{aligned}$$

by (1.5) and it implies \(V\in \mathcal {M}_{\omega ,\textbf{c}}\).

Now, we assume that the additional condition (3.3) holds. Namely, we assume

$$\begin{aligned} \limsup _{n\rightarrow \infty }G(U_n)= \limsup _{n\rightarrow \infty }\left\{ (4-2d)\omega Q(U_n)+(3-d)\textbf{c}\cdot \textbf{P}(U_n)\right\} \ge \eta . \end{aligned}$$

By the definitions of Q and \(\textbf{P}\),

$$\begin{aligned} \begin{aligned} |G(V_n)-G(V)|&\lesssim \Vert V_n-V\Vert _{L^2}^2 +\Vert V_n-V\Vert _{L^2}\Vert \nabla (V_n-V)\Vert _{L^2} +\Vert V\Vert _{L^2}\Vert \nabla (V_n-V)\Vert _{L^2}\\&\quad +\Vert V_n-V\Vert _{L^2}\Vert \nabla V\Vert _{L^2}\\&\lesssim \Vert V_n-V\Vert _{\mathcal {H}^1}^2+\Vert V_n-V\Vert _{\mathcal {H}^1} \ \rightarrow \ 0 \end{aligned} \end{aligned}$$

holds. We also have

$$\begin{aligned} G(V)=G(V_n)-(G(V_n)-G(V))=G(U_n)-(G(V_n)-G(V)). \end{aligned}$$

These imply \(G(V)=\limsup _{n\rightarrow \infty }G(U_n)\ge \eta \) and we get \(V\in \mathcal {M}_{\omega ,\textbf{c}}^*(\eta )\).

Proposition 3.6

Let \(d\in \{1,2,3\}\) and \(\alpha ,\beta ,\gamma >0\). Assume that \((\omega ,\textbf{c})\in \mathbb {R}\times \mathbb {R}^d\) satisfies \(\omega >\frac{\sigma |\textbf{c}|^2}{4}\). Then, we have \(\mathcal {M}_{\omega ,\textbf{c}}=\mathcal {G}_{\omega ,\textbf{c}}\).

Proof

We first prove \(\mathcal {M}_{\omega ,\textbf{c}}\subset \mathcal {G}_{\omega ,\textbf{c}}\). Let \(\Psi =(\psi _1,\psi _2,\psi _3)\in \mathcal {M}_{\omega ,\textbf{c}}\). Since \(\Psi \) is a minimizer of \(S_{\omega ,\textbf{c}}(\Phi )\) under the condition \(K_{\omega ,\textbf{c}}(\Phi )=0\), there exists a Lagrange multiplier \(\eta \in \mathbb {R}\) such that

$$\begin{aligned} D_jS_{\omega ,\textbf{c}}(\Psi )=\eta D_jK_{\omega ,\textbf{c}}(\Psi ),\ \ j=1,2,3. \end{aligned}$$

Therefore, we have

$$\begin{aligned}{} & {} 0=K_{\omega ,\textbf{c}}(\Psi ) =\partial _{\lambda }S_{\omega ,\textbf{c}}(\lambda \Psi )\big |_{\lambda =1} =\sum _{j=1}^3\langle D_jS_{\omega ,\textbf{c}}(\Psi ),\psi _j\rangle \\{} & {} \quad =\eta \sum _{j=1}^3\langle D_jK_{\omega ,\textbf{c}}(\Psi ),\psi _j\rangle =\eta \partial _{\lambda }K_{\omega ,\textbf{c}}(\lambda \Psi )\big |_{\lambda =1}. \end{aligned}$$

We also note that the following identities hold:

$$\begin{aligned} \begin{aligned} \partial _{\lambda }K_{\omega ,\textbf{c}}(\lambda \Psi )\big |_{\lambda =1}&=\partial _{\lambda }(2L(\lambda \Psi )+3N(\lambda \Psi )+2\omega Q(\lambda \Psi )+2\textbf{c}\cdot \textbf{P}(\lambda \Psi ))\big |_{\lambda =1}\\&=4L(\Psi )+9N(\Psi )+4\omega Q(\Psi )+4\textbf{c}\cdot \textbf{P}(\Psi )\\&=3K_{\omega ,\textbf{c}}(\Psi )-L_{\omega ,\textbf{c}}(\Psi )=-L_{\omega ,\textbf{c}}(\Psi ). \end{aligned} \end{aligned}$$

Because \(\Psi \ne 0\), we obtain \(\partial _{\lambda }K_{\omega ,\textbf{c}}(\lambda \Psi )\big |_{\lambda =1}<0\) by Proposition 3.1. This implies \(\eta =0\) and we get

$$\begin{aligned} D_jS_{\omega ,\textbf{c}}(\Psi )=0,\ \ j=1,2,3. \end{aligned}$$

This means \(\Psi \in \mathcal {E}_{\omega ,\textbf{c}}\). Furthermore, for any \(\Theta \in \mathcal {E}_{\omega ,\textbf{c}}\), we have \(K_{\omega ,\textbf{c}}(\Theta )=0\) by Remark 1.4. Thus, by the definition of \(\mu _{\omega ,\textbf{c}}\), we obtain

$$\begin{aligned} S_{\omega ,\textbf{c}}(\Psi )=\mu _{\omega ,\textbf{c}}\le S_{\omega ,\textbf{c}}(\Theta ). \end{aligned}$$

Therefore, we get \(\Psi \in \mathcal {G}_{\omega ,\textbf{c}}\).

Next, we prove \(\mathcal {G}_{\omega ,\textbf{c}}\subset \mathcal {M}_{\omega ,\textbf{c}}\). Let \(\Psi \in \mathcal {G}_{\omega ,\textbf{c}}\). By Proposotion 3.3, there exists \(\Phi \in \mathcal {M}_{\omega ,\textbf{c}}\). Because \(\mathcal {M}_{\omega ,\textbf{c}}\subset \mathcal {G}_{\omega ,\textbf{c}}\subset \mathcal {E}_{\omega ,\textbf{c}}\), we have \(\Phi \in \mathcal {E}_{\omega ,\textbf{c}}\). This implies that

$$\begin{aligned} S_{\omega ,\textbf{c}}(\Psi )\le S_{\omega ,\textbf{c}}(\Phi )=\mu _{\omega ,\textbf{c}}. \end{aligned}$$

On the other hand, by Remark 1.4, we have \(K_{\omega ,\textbf{c}}(\Psi )=0\) since \(\Psi \in \mathcal {G}_{\omega ,\textbf{c}}\subset \mathcal {E}_{\omega ,\textbf{c}}\). Therefore, by the definition of \(\mu _{\omega ,\textbf{c}}\), we get

$$\begin{aligned} \mu _{\omega ,\textbf{c}}\le S_{\omega ,\textbf{c}}(\Psi ). \end{aligned}$$

As a result, we obtain \(S_{\omega ,\textbf{c}}(\Psi )=\mu _{\omega ,\textbf{c}}\) and this implies \(\Psi \in \mathcal {M}_{\omega ,\textbf{c}}\).

Proof of Theorem 1.2

Theorem 1.2 can be proved by combining Proposition 3.3 and Proposition 3.6.

Next we prepare fundamental identities about the ground-state energy level \(\mu _{\omega ,\textbf{c}}\), which will be applied to prove the stability results in Sect. 4.

Proposition 3.7

Let \(d\in \{1,2,3\}\) and \(\alpha ,\beta ,\gamma >0\). Assume that \((\omega ,\textbf{c})\in \mathbb {R}\times \mathbb {R}^d\) satisfies \(\omega >\frac{\sigma |\textbf{c}|^2}{4}\). Then, we have

$$\begin{aligned} \mu _{\omega ,\textbf{c}}=\omega ^{2-\frac{d}{2}}\mu _{1,\frac{\textbf{c}}{\sqrt{\omega }}}. \end{aligned}$$

Proof

By Propositions 3.3 and 3.6, there exists \(\Psi \in \mathcal {M}_{1,\frac{\textbf{c}}{\sqrt{\omega }}}=\mathcal {G}_{1,\frac{\textbf{c}}{\sqrt{\omega }}}\), which implies that \(\Psi \in \mathcal {E}_{1,\frac{\textbf{c}}{\sqrt{\omega }}}\). Furthermore, we set

$$\begin{aligned} \Psi _{\omega }(x):=\omega ^{\frac{1}{2}}\Psi (\omega ^{\frac{1}{2}}x). \end{aligned}$$

Then, we have \(\Psi _{\omega }\in \mathcal {E}_{\omega ,\textbf{c}}\) and by Remark 1.5,

$$\begin{aligned} S_{\omega ,\textbf{c}}(\Psi _{\omega })=\omega ^{2-\frac{d}{2}}S_{1,\frac{\textbf{c}}{\sqrt{\omega }}}(\Psi ). \end{aligned}$$
(3.12)

Let \(\Theta \in \mathcal {E}_{\omega ,\textbf{c}}\). We put

$$\begin{aligned} \Theta ^{\omega }(x):=\omega ^{-\frac{1}{2}}\Theta (\omega ^{-\frac{1}{2}}x). \end{aligned}$$

Then, we can see that \(\Theta ^{\omega }\in \mathcal {E}_{1,\frac{\textbf{c}}{\sqrt{\omega }}}\) and

$$\begin{aligned} S_{1,\frac{\textbf{c}}{\sqrt{\omega }}}(\Theta ^{\omega })=\omega ^{\frac{d}{2}-2}S_{\omega ,\textbf{c}}(\Theta ). \end{aligned}$$
(3.13)

On the other hand, because \(\Psi \) is a minimizer of \(\mathcal {M}_{1,\frac{\textbf{c}}{\sqrt{\omega }}}\), we obtain

$$\begin{aligned} S_{1,\frac{\textbf{c}}{\sqrt{\omega }}}(\Psi )\le S_{1,\frac{\textbf{c}}{\sqrt{\omega }}}(\Theta ^{\omega }) \end{aligned}$$

This implies that by (3.12) and (3.13), \(S_{\omega ,\textbf{c}}(\Psi _{\omega })\le S_{\omega ,\textbf{c}}(\Theta )\). This implies \(\Psi _{\omega }\in \mathcal {G}_{\omega ,\textbf{c}}=\mathcal {M}_{\omega ,\textbf{c}}\). As a result, we get

$$\begin{aligned} \mu _{\omega ,\textbf{c}}=S_{\omega ,\textbf{c}}(\Psi _{\omega }) =\omega ^{2-\frac{d}{2}}S_{1,\frac{\textbf{c}}{\sqrt{\omega }}}(\Psi ) =\omega ^{2-\frac{d}{2}}\mu _{1,\frac{\textbf{c}}{\sqrt{\omega }}}. \end{aligned}$$

Proposition 3.8

Let \(d\in \{1,2,3\}\) and \(\alpha ,\beta ,\gamma >0\). Assume that \((\omega ,\textbf{c})\in \mathbb {R}\times \mathbb {R}^d\) satisfies \(\omega >\frac{\sigma |\textbf{c}|^2}{4}\). If \(\Phi \in \mathcal {M}_{\omega ,\textbf{c}}\), then it holds that

$$\begin{aligned} 2\omega Q(\Phi ) +\textbf{c}\cdot \textbf{P}(\Phi ) =(4-d)\omega ^{2-\frac{d}{2}} \mu _{1,\frac{\textbf{c}}{\sqrt{\omega }}}. \end{aligned}$$
(3.14)

Proof

Because \(\Phi \) is a minimizer of \(\mu _{\omega ,\textbf{c}}\), it holds

$$\begin{aligned} K_{\omega , \textbf{c}}(\Phi ) =2L(\Phi )+3N(\Phi )+2\omega Q(\Phi )+2\textbf{c}\cdot \textbf{P}(\Phi )=0. \end{aligned}$$
(3.15)

We also note that \(\Phi \) is a weak solution to (1.7) by Proposition 3.6. Therefore, by Proposition 2.4, the identity

$$\begin{aligned} 2L(\Phi )+\left( \frac{d}{2}+1\right) N(\Phi )+\textbf{c}\cdot \textbf{P}(\Phi )=0 \end{aligned}$$
(3.16)

holds. The identities (3.15) and (3.16) imply

$$\begin{aligned} L(\Phi )=\frac{d+2}{4-d}\omega Q(\Phi )+\frac{d-1}{4-d}\textbf{c}\cdot \textbf{P}(\Phi ),\ \ \ \text {and}\ \ \ N(\Phi )=-\frac{2}{4-d}\left( 2\omega Q(\Phi ) +\textbf{c}\cdot \textbf{P}(\Phi )\right) . \end{aligned}$$

Hence we obtain

$$\begin{aligned} \begin{aligned} \mu _{\omega ,\textbf{c}} =S_{\omega ,\textbf{c}}(\Phi )&=L(\Phi )+N(\Phi )+\omega Q(\Phi )+\textbf{c}\cdot \textbf{P}(\Phi ) =\frac{1}{4-d}\left( 2\omega Q(\Phi )+\textbf{c}\cdot \textbf{P}(\Phi )\right) . \end{aligned} \end{aligned}$$

On the other hand, by Proposition 3.7, it holds that \(\mu _{\omega ,\textbf{c}}=\omega ^{2-\frac{d}{2}}\mu _{1,\frac{\textbf{c}}{\sqrt{\omega }}}\). Therefore, we get (3.14).

4 Proof of global well-posednessand stability results

In this section, we give the proofs of Theorems 1.3,  1.5 and Corollaries 1.4,  1.6.

4.1 Proof of global well-posedness

We define the subsets in the energy space \(\mathcal {H}^1\)

$$\begin{aligned} \begin{aligned} \mathcal {A}^{+}_{\omega ,\textbf{c}}&:=\left\{ \Psi \in \mathcal {H}^1\backslash \{(\textbf{0},\textbf{0},\textbf{0})\}\ |\ S_{\omega ,\textbf{c}}(\Psi )< \mu _{\omega ,\textbf{c}},\ K_{\omega ,\textbf{c}}(\Psi )> 0\right\} ,\\ \mathcal {A}^{-}_{\omega ,\textbf{c}}&:=\left\{ \Psi \in \mathcal {H}^1\backslash \{(\textbf{0},\textbf{0},\textbf{0})\}\ |\ S_{\omega ,\textbf{c}}(\Psi )< \mu _{\omega ,\textbf{c}},\ K_{\omega ,\textbf{c}}(\Psi )< 0\right\} . \end{aligned} \end{aligned}$$
(4.1)

We prove that \(\mathcal {A}^{\pm }_{\omega ,\textbf{c}}\) is invariant under the flow of (1.1).

Proposition 4.1

Let \(d\in \{1,2,3\}\) and \(\alpha ,\beta ,\gamma >0\). Assume that \((\omega ,\textbf{c})\in \mathbb {R}\times \mathbb {R}^d\) satisfies \(\omega >\frac{\sigma |\textbf{c}|^2}{4}\). Then the sets \(\mathcal {A}^{\pm }_{\omega ,\textbf{c}}\) are invariant under the flow of (1.1). More precisely, if \(U_0\) belongs to \(\mathcal {A}^{+}_{\omega ,\textbf{c}}\) (resp \(\mathcal {A}^{-}_{\omega ,\textbf{c}})\), then the solution U(t) of (1.1) with \(U(0)=U_0\) also belongs to \(\mathcal {A}^{+}_{\omega ,\textbf{c}}\) (resp \(\mathcal {A}^{-}_{\omega ,\textbf{c}})\) for all \(t\in I_{\max }\).

Proof

Let \(U_0\in \mathcal {A}^{+}_{\omega ,\textbf{c}}\) and U be the solution to (1.1) with \(U(0)=U_0\) on \(I_{\max }\). Because \(S_{\omega ,\textbf{c}}(U_0)<\mu _{\omega ,\textbf{c}}\) holds and \(S_{\omega ,\textbf{c}}\) is a conserved quantity, we have \(S_{\omega ,\textbf{c}}(U(t))<\mu _{\omega ,\textbf{c}}\) for all \(t\in I_{\max }\). Furthermore, because \(U_0\ne 0\) and Q is a conserved quantity, we obtain \(Q(U(t))=Q(U_0)\ne 0\). It implies that \(U(t)\ne 0\) for all \(t\in I_{\max }\). Therefore, it suffices to show that \(K_{\omega ,\textbf{c}}(U(t))>0\) for all \(t\in I_{\max }\). Assume \(K_{\omega ,\textbf{c}}(U(t_*))<0\) for some \(t_*\in I_{\max }\). Then, by the continuity of \(t\mapsto K_{\omega ,\textbf{c}}(U(t))\), there exists \(t_0\in (0,t_*)\) such that \(K_{\omega ,\textbf{c}}(U(t_0))=0\). This implies \(U(t_0)=0\) because \(S_{\omega ,\textbf{c}}(U(t_0))<\mu _{\omega ,\textbf{c}}\). It is a contradiction. The invariance of \(\mathcal {A}^{-}_{\omega ,\textbf{c}}\) can be proved by the same manner.

Now we prove our global well-posedness results (Theorem 1.3 and Corollary 1.4).

Proof of Theorem 1.3

Let U be a solution on the maximal existence time interval \(I_{\max }\) to (1.1) with \(U(0)=U_0\in \mathcal {H}^1\backslash \{0\}\) satisfying

$$\begin{aligned} S_{\omega ,\textbf{c}}(U_0)<\mu _{\omega ,\textbf{c}},\ \ K_{\omega ,\textbf{c}}(U_0)>0. \end{aligned}$$

That is \(U_0\in \mathcal {A}^{+}_{\omega ,\textbf{c}}\). By Proposition 4.1, \(U(t)\in \mathcal {A}^{+}_{\omega ,\textbf{c}}\) holds for any \(t\in I_{\max }\). This implies \(K_{\omega ,\textbf{c}}(U(t))>0\) and by (1.5) we have

$$\begin{aligned} L_{\omega ,\textbf{c}}(U(t))\le 6S_{\omega ,\textbf{c}}(U(t)). \end{aligned}$$

Thus by Proposition 3.1, we obtain

$$\begin{aligned} \Vert U(t)\Vert _{\mathcal {H}^1}^2 \le \frac{6}{C}S_{\omega ,\textbf{c}}(U(t)) =\frac{6}{C}S_{\omega ,\textbf{c}}(U_0) \end{aligned}$$

for any \(t\in I_{\max }\) and some \(C>0\) independent of t, since \(S_{\omega ,\textbf{c}}\) is a conserved quantity, which implies that \(|I_{\max }|=\infty \) by the blow-up alternative.

Proof of Corollary 1.4

We assume \(\Phi \in \mathcal {M}_{1,\textbf{c}_0}\) with \(|\textbf{c}_0|<\frac{2}{\sqrt{\sigma }}\) and set \(\textbf{c}:=\sqrt{\omega }\ \textbf{c}_0\) for \(\omega >0\). Furthermore, we can assume \(Q(U_0)>0\) because \(U_0=0\) when \(Q(U_0)=0\). Thanks to Theorem 1.3, it is enough to show that

$$\begin{aligned} S_{\omega ,\textbf{c}}(U_0)<\mu _{\omega ,\textbf{c}},\ \ K_{\omega ,\textbf{c}}(U_0)>0 \end{aligned}$$
(4.2)

hold for some \(\omega >0\). Because

$$\begin{aligned} \begin{aligned} K_{\omega ,\textbf{c}}(U_0)&=2L(U_0)+3N(U_0)+2\omega Q(U_0)+\textbf{c}\cdot \textbf{P}(U_0)\\&=2\omega Q(U_0)+\sqrt{\omega }\ \textbf{c}_0\cdot \textbf{P}(U_0) +2L(U_0)+3N(U_0) \end{aligned} \end{aligned}$$

and \(Q(U_0)>0\), the second inequality in (4.2) holds for sufficiently large \(\omega >0\).

Next, to prove the first inequality in (4.2), we define the function \(f:(0,\infty )\rightarrow \mathbb {R}\) as

$$\begin{aligned} f(\omega ):=\mu _{\omega ,\textbf{c}}-S_{\omega ,\textbf{c}}(U_0) =\mu _{\omega ,\sqrt{\omega }\ \textbf{c}_0}-S_{\omega ,\sqrt{\omega }\ \textbf{c}_0}(U_0). \end{aligned}$$

Because \(\Phi \in \mathcal {M}_{1,\textbf{c}_0}\), we have

$$\begin{aligned} \textbf{c}_0\cdot \textbf{P}(\Phi )=-2(L(\Phi )+N(\Phi ))=-2E(\Phi ) \end{aligned}$$

by Proposition 2.4 with \(d=2\). This identity and Proposition 3.7 with \(d=2\) imply that

$$\begin{aligned} \mu _{\omega ,\sqrt{\omega }\ \textbf{c}_0} =\omega \mu _{1,\textbf{c}_0} =\omega S_{1,\textbf{c}_0}(\Phi ) =\omega (Q(\Phi )-E(\Phi )). \end{aligned}$$

Therefore, we obtain

$$\begin{aligned} \begin{aligned} f(\omega )&=\omega (Q(\Phi )-E(\Phi )) -(E(U_0)+\omega Q(U_0)+\sqrt{\omega }\ \textbf{c}_0\cdot \textbf{P}(U_0))\\&=\omega (Q(\Phi )-E(\Phi )-Q(U_0))-\sqrt{\omega }\ \textbf{c}_0\cdot \textbf{P}(U_0)-E(U_0). \end{aligned} \end{aligned}$$

This says that if \(Q(U_0)<Q(\Phi )-E(\Phi )\), then \(f(\omega )>0\) holds for sufficiently large \(\omega >0\).

We define the subsets in the energy space \(\mathcal {H}^1\)

$$\begin{aligned} \begin{aligned} \mathcal {B}^{+}_{\omega ,\textbf{c}}&:=\left\{ \Psi \in \mathcal {H}^1 \backslash \{(\textbf{0},\textbf{0},\textbf{0})\}\ |\ S_{\omega ,\textbf{c}}(\Psi )< \mu _{\omega ,\textbf{c}},\ N(\Psi )> -2\mu _{\omega ,\textbf{c}}\right\} ,\\ \mathcal {B}^{-}_{\omega ,\textbf{c}}&:=\left\{ \Psi \in \mathcal {H}^1\backslash \{(\textbf{0},\textbf{0},\textbf{0})\}\ |\ S_{\omega ,\textbf{c}}(\Psi )< \mu _{\omega ,\textbf{c}},\ N(\Psi )< -2\mu _{\omega ,\textbf{c}}\right\} . \end{aligned} \end{aligned}$$
(4.3)

The following Proposition will be used to prove the stability of ground states set in next subsection.

Proposition 4.2

Let \(d\in \{1,2,3\}\) and \(\alpha ,\beta ,\gamma >0\). Assume that \((\omega ,\textbf{c})\in \mathbb {R}\times \mathbb {R}^d\) satisfies \(\omega >\frac{\sigma |\textbf{c}|^2}{4}\). Then, \(\mathcal {A}^{+}_{\omega ,\textbf{c}}=\mathcal {B}^{+}_{\omega ,\textbf{c}}\) and \(\mathcal {A}^{-}_{\omega ,\textbf{c}}=\mathcal {B}^{-}_{\omega ,\textbf{c}}\) hold, where \(\mathcal {A}^{+}_{\omega ,\textbf{c}}\) and \(\mathcal {A}^{-}_{\omega ,\textbf{c}}\) are defined in (4.1).

Proof

We first prove \(\mathcal {A}^{+}_{\omega ,\textbf{c}}= \mathcal {B}^{+}_{\omega ,\textbf{c}}\). Let \(U\in \mathcal {A}^{+}_{\omega ,\textbf{c}}\). Because \(S_{\omega ,\textbf{c}}(U)<\mu _{\omega ,\textbf{c}}\) and \(K_{\omega ,\textbf{c}}(U)>0\) hold we obtain \( N(U)>-2\mu _{\omega ,\textbf{c}}\) by (1.6). This says \(U\in \mathcal {B}^{+}_{\omega ,\textbf{c}}\). Conversely, let \(V\in \mathcal {B}^{+}_{\omega ,\textbf{c}}\). We assume \(K_{\omega ,\textbf{c}}(V)\le 0\). Then, by Proposition 3.2 (ii), we have \(L_{\omega ,\textbf{c}}(V)\ge 6\mu _{\omega ,\textbf{c}}\). Since \(N(V)>-2\mu _{\omega ,\textbf{c}}\), this and (1.5) imply

$$\begin{aligned} S_{\omega ,\textbf{c}}(V)=\frac{1}{2}L_{\omega ,\textbf{c}}(V)+N(V)>\mu _{\omega ,\textbf{c}}, \end{aligned}$$

which contradicts to \(V\in \mathcal {B}^{+}_{\omega ,\textbf{c}}\). Therefore, we get \(K_{\omega ,\textbf{c}}(V)>0\), which implies \(V\in \mathcal {A}^{+}_{\omega ,\textbf{c}}\).

Next, we prove \(\mathcal {A}^{-}_{\omega ,\textbf{c}}= \mathcal {B}^{-}_{\omega ,\textbf{c}}\). Let \(U\in \mathcal {A}^{-}_{\omega ,\textbf{c}}\). Then, \(K_{\omega ,\textbf{c}}(U)<0\) holds and we have \(L_{\omega ,\textbf{c}}(U)>6\mu _{\omega ,\textbf{c}}\) by Proposition 3.2 (ii). Therefore, we obtain

$$\begin{aligned} N(U)=\frac{1}{3}(K_{\omega ,\textbf{c}}(U)-L_{\omega ,\textbf{c}}(U))<-2\mu _{\omega ,\textbf{c}}. \end{aligned}$$

This says \(U\in \mathcal {B}^{-}_{\omega ,\textbf{c}}\). Conversely, let \(V\in \mathcal {B}^{-}_{\omega ,\textbf{c}}\). Then, by (1.6) and \(S_{\omega ,\textbf{c}}(V)<\mu _{\omega ,\textbf{c}}\), we have

$$\begin{aligned} K_{\omega ,\textbf{c}}(V)=2S_{\omega ,\textbf{c}}(V)+N(V)<0. \end{aligned}$$

This says \(V\in \mathcal {A}^{-}_{\omega ,\textbf{c}}\).

Remark 4.1

Let \(\alpha \), \(\beta \), \(\gamma >0\), \(\omega >\frac{\sigma |c|^2}{4}\), and \(U_0\in \mathcal {B}^{+}_{\omega ,\textbf{c}}\). Then, we can obtain the global solution to (1.1) with \(U|_{t=0}=U_0\) by Theorem 1.3 and Proposition 4.2. We note that \(U_0\in \mathcal {B}^{+}_{\omega ,\textbf{c}}\) satisfies either

$$\begin{aligned} \mathrm{(i)}\ -2\mu _{\omega ,\textbf{c}}<N(U_0)<0\ \ \ \ \textrm{or}\ \ \ \ \mathrm{(ii)}\ N(U_0)\ge 0\ \ (then\ E(U_0)>0). \end{aligned}$$

While the case (i) corresponds to the focusing case with small nonlinear effect, the case (ii) corresponds to the defocusing case.

4.2 Proof of stability of ground states set

In this subsection, we write \(\mu _{\omega ,\textbf{c}}=\mu (\omega ,\textbf{c})\). For fixed \((\omega , \textbf{c})\in \mathbb {R}\times \mathbb {R}^d\) with \(\omega >\frac{\sigma |\textbf{c}|^2}{4}\), we define the function \(h=h_{\omega ,\textbf{c}}\) on \(\bigl (-\infty ,\sqrt{\omega }\bigl )\) as

$$\begin{aligned} h(\tau ):= \mu \bigg ((\sqrt{\omega }-\tau )^2,\frac{\textbf{c}}{\sqrt{\omega }}(\sqrt{\omega }-\tau )\bigg ) =(\sqrt{\omega }-\tau )^{4-d} \mu \left( 1,\frac{\textbf{c}}{\sqrt{\omega }}\right) , \end{aligned}$$

where to obtain the second equality, we have used Proposition 3.7. Namely, the function h is a restriction of \(\mu \) on a so-called scaling curve

$$\begin{aligned} \mathcal {C}_{\omega ,\textbf{c}}:=\left\{ \left. \left( (\sqrt{\omega }-\tau )^2,\frac{\textbf{c}}{\sqrt{\omega }}(\sqrt{\omega }-\tau )\right) \ \right| \ \tau < \sqrt{\omega }\ \right\} . \end{aligned}$$

We note that

$$\begin{aligned} \begin{aligned} h'(\tau )&=-(4-d)(\sqrt{\omega }-\tau )^{3-d}\mu \left( 1,\frac{\textbf{c}}{\sqrt{\omega }}\right) ,\\ h''(\tau )&=(3-d)(4-d)(\sqrt{\omega }-\tau )^{2-d}\mu \left( 1,\frac{\textbf{c}}{\sqrt{\omega }}\right) , \end{aligned} \end{aligned}$$

and h is strictly decreasing on \(\bigl (-\infty ,\sqrt{\omega }\bigl )\). By using these identities and (3.14), the following lemma is obtained.

Lemma 4.3

Let \(d\in \{1,2\}\) and \(\alpha ,\beta ,\gamma >0\). Assume that \((\omega ,\textbf{c})\in \mathbb {R}\times \mathbb {R}^d\) satisfies \(\omega >\frac{\sigma |\textbf{c}|^2}{4}\). If \(\Phi \in \mathcal {M}_{\omega ,\textbf{c}}\), then it holds that

$$\begin{aligned} \begin{aligned} h(0)&=\mu (\omega ,\textbf{c})=S_{\omega ,\textbf{c}}(\Phi ),\\ h'(0)&=-\frac{1}{\sqrt{\omega }}(2\omega Q(\Phi )+\textbf{c}\cdot \textbf{P}(\Phi )),\\ h''(0)&=\frac{3-d}{\omega }(2\omega Q(\Phi )+\textbf{c}\cdot \textbf{P}(\Phi )). \end{aligned} \end{aligned}$$

For \(\eta >0\), we define the function \(F=F_{\omega ,\textbf{c},\eta }\) on \(\bigl (-\infty ,\sqrt{\omega }\bigl )\) as

$$\begin{aligned} F(\tau ):=\inf _{\Phi \in \mathcal {M}_{\omega ,\textbf{c}}^*(\eta )} \left( \frac{h''(\tau )}{2}-Q(\Phi )\right) . \end{aligned}$$

Lemma 4.4

Let \(d\in \{1,2\}\) and \(\alpha ,\beta ,\gamma >0\). Assume that \((\omega ,\textbf{c})\in \mathbb {R}\times \mathbb {R}^d\) satisfies \(\omega >\frac{\sigma |\textbf{c}|^2}{4}\). There exists \(\tau _*=\tau _*(\omega ,\textbf{c},\eta )>0\) such that \(F(\tau )\ge \frac{\eta }{4\omega }\) holds for any \(\tau \in (-\tau _*,\tau _*)\).

Proof

Because the definition of the function h is independent of \(\Phi \in \mathcal {M}_{\omega ,\textbf{c}}^*(\eta )\), we have

$$\begin{aligned} F(\tau )=\frac{h''(\tau )}{2}-\sup _{\Phi \in \mathcal {M}_{\omega ,\textbf{c}}^*(\eta )}Q(\Phi ). \end{aligned}$$

In particular, F is continuous because \(h''(\tau )\) is polynomial. On the other hand, we have

$$\begin{aligned} \frac{h''(0)}{2}-Q(\Phi )=\frac{1}{2\omega }\left\{ (4-2d)\omega Q(\Phi )+(3-d)\textbf{c}\cdot \textbf{P}(\Phi ))\right\} \ge \frac{\eta }{2\omega }\ (>0) \end{aligned}$$

for any \(\Phi \in \mathcal {M}_{\omega ,\textbf{c}}^*(\eta )\) by Lemma 4.3. This implies \(F(0)\ge \frac{\eta }{2\omega }\). Therefore, we obtain the conclusion by the continuity of F on \(\tau =0\).

For \(\tau _0\in \Bigl (0,\sqrt{\omega }\Bigl )\), we set

$$\begin{aligned} \omega _{\pm }=\omega _{\pm }(\tau _0):=(\sqrt{\omega }\pm \tau _0)^2,\ \ \textbf{c}_{\pm }=\textbf{c}_{\pm }(\tau _0):=\frac{\textbf{c}}{\sqrt{\omega }}(\sqrt{\omega } \pm \tau _0), \end{aligned}$$

where the double-signs correspond. The following proposition plays an important role to prove the stability result.

Proposition 4.5

Let \(d\in \{1,2\}\), \(\alpha ,\beta ,\gamma >0\), and \(\eta >0\). Assume that \((\omega ,\textbf{c})\in \mathbb {R}\times \mathbb {R}^d\) satisfies \(\omega >\frac{\sigma |\textbf{c}|^2}{4}\) and \(\mathcal {M}_{\omega ,\textbf{c}}^*(\eta )\ne \emptyset \). Then for any \(\tau _0\in (0,\tau _*)\), there exists \(\delta =\delta (\tau _0,\tau _*,\omega ,\textbf{c},\eta )>0\) such that if \(U_0\in \mathcal {H}^1\) satisfies

$$\begin{aligned} \inf _{\Phi \in \mathcal {M}_{\omega ,\textbf{c}}^*(\eta )}\Vert U_0-\Phi \Vert _{\mathcal {H}^1}<\delta , \end{aligned}$$
(4.4)

then it holds \(U_0\in \mathcal {B}^{+}_{\omega _+,\textbf{c}_+}\cap \mathcal {B}^{-}_{\omega _-,\textbf{c}_-}\), where \(\tau _*>0\) is given in Lemma 4.4 and \(\mathcal {B}^{\pm }_{\omega _{\pm },\textbf{c}_{\pm }}\) are defined in (4.3). Furthermore, such \(\delta \) tends to 0 as \(\tau _0\rightarrow 0\).

Proof

We assume that \(U_0\in \mathcal {H}^1\) satisfies (4.4) with sufficiently small \(\delta \in (0,1)\), which will be chosen later. By the definition of the infimum, there exists \(\Phi _{\omega ,\textbf{c}}\in \mathcal {M}_{\omega ,\textbf{c}}^*(\eta )\) such that

$$\begin{aligned} \Vert U_0-\Phi _{\omega ,\textbf{c}}\Vert _{\mathcal {H}^1}<2\delta . \end{aligned}$$

Therefore, we have

$$\begin{aligned} \begin{aligned} |Q(U_0)-Q(\Phi _{\omega ,\textbf{c}})|&\lesssim (\Vert U_0\Vert _{\mathcal {H}^1}+\Vert \Phi _{\omega ,\textbf{c}}\Vert _{\mathcal {H}^1}) \Vert U_0-\Phi _{\omega ,\textbf{c}}\Vert _{\mathcal {H}^1}\lesssim \delta ,\\ |P(U_0)-P(\Phi _{\omega ,\textbf{c}})|&\lesssim (\Vert U_0\Vert _{\mathcal {H}^1}+\Vert \Phi _{\omega ,c}\Vert _{\mathcal {H}^1}) \Vert U_0-\Phi _{\omega ,\textbf{c}}\Vert _{\mathcal {H}^1}\lesssim \delta ,\\ |N(U_0)-N(\Phi _{\omega ,\textbf{c}})|&\lesssim (\Vert U_0\Vert _{\mathcal {H}^1}^2 +\Vert \Phi _{\omega ,\textbf{c}}\Vert _{\mathcal {H}^1}^2) \Vert U_0-\Phi _{\omega ,\textbf{c}}\Vert _{\mathcal {H}^1} \lesssim \delta \end{aligned} \end{aligned}$$
(4.5)

and

$$\begin{aligned} |S_{\omega ,c}(U_0)-S_{\omega ,\textbf{c}}(\Phi _{\omega ,\textbf{c}})|\lesssim (\Vert U_0\Vert _{\mathcal {H}^1}+\Vert \Phi _{\omega ,\textbf{c}}\Vert _{\mathcal {H}^1} +\Vert U_0\Vert _{\mathcal {H}^1}^2+\Vert \Phi _{\omega ,\textbf{c}}\Vert _{\mathcal {H}^1}^2) \Vert U_0-\Phi _{\omega ,\textbf{c}}\Vert _{\mathcal {H}^1}\lesssim \delta . \end{aligned}$$
(4.6)

Here we note that the above implicit constants do not depend on \(U_0\) and \(\Phi _{\omega ,\textbf{c}}\). Indeed, by Proposition 3.1 and (1.5), there exists \(C=C(\omega ,\textbf{c})\) such that

$$\begin{aligned} \Vert \Phi _{\omega ,\textbf{c}}\Vert _{\mathcal {H}^1}^2 \le CL_{\omega ,\textbf{c}}(\Phi _{\omega ,\textbf{c}})=6C\mu _{\omega ,\textbf{c}}. \end{aligned}$$

This implies that

$$\begin{aligned} \Vert U_0\Vert _{\mathcal {H}^1}+\Vert \Phi _{\omega ,\textbf{c}}\Vert _{\mathcal {H}^1} \le \Vert U_0-\Phi _{\omega ,\textbf{c}}\Vert _{\mathcal {H}^1} +2\Vert \Phi _{\omega ,\textbf{c}}\Vert _{\mathcal {H}^1} \le \delta +2\sqrt{6C\mu _{\omega ,\textbf{c}}}. \end{aligned}$$

We divide the proof into the following two steps:

Step 1. We first prove that for any \(\tau _0 \in (0,\tau _*)\),

$$\begin{aligned} S_{\omega _{\pm },\textbf{c}_{\pm }}(U_0)<\mu (\omega _{\pm },\textbf{c}_{\pm }). \end{aligned}$$
(4.7)

By (4.5), (4.6), and Lemma 4.3, we obtain

$$\begin{aligned} \begin{aligned} S_{\omega _{\pm },\textbf{c}_{\pm }}(U_0)&=E(U_0)+\omega _{\pm }Q(U_0)+\textbf{c}_{\pm }\cdot \textbf{P}(U_0)\\&=S_{\omega ,\textbf{c}}(U_0) \pm \frac{\tau _0}{\sqrt{\omega }}\left( 2\omega Q(U_0)+\textbf{c}\cdot \textbf{P}(U_0)\right) +\tau _0^2Q(U_0)\\&=S_{\omega ,\textbf{c}}(\Phi _{\omega ,\textbf{c}}) \pm \frac{\tau _0}{\sqrt{\omega }} \left( 2\omega Q(\Phi _{\omega ,\textbf{c}})+\textbf{c}\cdot \textbf{P}(\Phi _{\omega ,\textbf{c}})\right) +\tau _0^2Q(\Phi _{\omega ,\textbf{c}})+O(\delta )\\&=h(0)\mp \tau _0 h'(0)+\tau _0^2Q(\Phi _{\omega ,\textbf{c}})+O(\delta ). \end{aligned} \end{aligned}$$

We can assume that the coefficient which appears in \(O(\delta )\) does not depend on \(\tau _0\) (but depends on \(\tau _*\)) by using \(0<\tau _o<\tau _*\).

On the other hand, by the Taylor expansion, there exists \(\theta \in (-\tau _0, \tau _0)\) such that

$$\begin{aligned} h(\mp \tau _0)=h(0)\mp \tau _0 h'(0)+\frac{\tau _0^2}{2}h''(\theta ). \end{aligned}$$

Therefore, we obtain

$$\begin{aligned} \begin{aligned} S_{\omega _{\pm },\textbf{c}_{\pm }}(U_0) =h(\mp \tau _0)-\tau _0^2\left( \frac{h''(\theta )}{2}-Q(\Phi _{\omega ,\textbf{c}})\right) +O(\delta ) \le h(\mp \tau _0)-\tau _0^2F(\theta )+O(\delta ). \end{aligned} \end{aligned}$$

Since \(|\theta |<\tau _0<\tau _*\), by Lemma 4.4, we have \(F(\theta )\ge \frac{\eta }{4\omega }\). Here we choose \(\delta =\delta (\tau _0,\tau _*,\omega ,\textbf{c},\eta )>0\) as

$$\begin{aligned} -\frac{\tau _0^2\eta }{4\omega }+O(\delta )<0. \end{aligned}$$

Then we have

$$\begin{aligned} S_{\omega _{\pm },\textbf{c}_{\pm }}(U_0) <h(\mp \tau _0). \end{aligned}$$

Therefore we get (4.7) by the definition of h. We also note that \(\delta \) tends to 0 as \(\tau _0\rightarrow 0\) because the coefficient which appears in \(O(\delta )\) does not depend on \(\tau _0\).

Step 2. Next, we prove that for any \(\tau _0\in (0,\tau _*)\),

$$\begin{aligned} -2\mu (\omega _+,\textbf{c}_+)<N(U_0)<-2\mu (\omega _-,\textbf{c}_-). \end{aligned}$$
(4.8)

Because h is a decreasing function on \(\Bigl (-\infty ,\sqrt{\omega }\Bigl )\), it holds that

$$\begin{aligned} h(\tau )<h(0)<h(-\tau ) \end{aligned}$$

for any \(0<\tau <\tau _0\). Since \(K_{\omega ,\textbf{c}}(\Phi _{\omega ,\textbf{c}})=0\), by using (1.6), we have

$$\begin{aligned} -2h(-\tau )<N(\Phi _{\omega ,\textbf{c}})<-2h(\tau ). \end{aligned}$$
(4.9)

Therefore, by the third inequality of (4.5), there exists \(C=C(\omega ,\textbf{c})>0\) such that

$$\begin{aligned} -C\delta -2h(-\tau )<N(U_0)<C\delta -2h(\tau ). \end{aligned}$$

By choosing \(\delta =\delta (\tau _0,\tau _*,\omega ,\textbf{c})>0\) further smaller as

$$\begin{aligned} C\delta <2\min \{h(\tau )-h(\tau _0),h(-\tau _0)-h(-\tau )\}(>0), \end{aligned}$$

we have

$$\begin{aligned} -2h(-\tau _0)<N(U_0)<-2h(\tau _0), \end{aligned}$$

which is equivalent to (4.8). This completes the proof of the proposition.

For \(n\in \mathbb {N}\), we set

$$\begin{aligned} \omega _{n\pm }:=\omega _{\pm }(1/n)=\left( \sqrt{\omega }\pm \frac{1}{n}\right) ^2,\ \ \textbf{c}_{n\pm }:=\textbf{c}_{\pm }(1/n)=\frac{\textbf{c}}{\sqrt{\omega }}\left( \sqrt{\omega } \pm \frac{1}{n}\right) . \end{aligned}$$

Let \(\tau _*\) be given in Lemma 4.4. Then by Proposition 4.5, for \(n>\frac{1}{\tau _*}\), there exists \(\delta _n=\delta (1/n,\tau _{*},\eta ,\omega ,\textbf{c})>0\) satisfying \(\delta _n\rightarrow 0\) as \(n\rightarrow \infty \), such that if \(U_0\in \mathcal {H}^1\) satisfies

$$\begin{aligned} \inf _{\Phi \in \mathcal {M}_{\omega ,\textbf{c}}^*(\eta )}\Vert U_0-\Phi \Vert _{\mathcal {H}^1}<\delta _n, \end{aligned}$$

then \(U_0\in \mathcal {B}^{+}_{\omega _{n+},\textbf{c}_{n+}}\cap \mathcal {B}^{-}_{\omega _{n-},\textbf{c}_{n-}}\).

Now we give a proof of the stability of the ground-states set \(\mathcal {M}_{\omega ,\textbf{c}}^*(\eta )\).

Proof of Theorem 1.5

We prove the stability by contradiction. We assume that the ground-states set \(\mathcal {M}_{\omega ,\textbf{c}}^*(\eta )\) is unstable. Then, there exists \(\epsilon >0\) such that the following holds:

For each \(n\in \mathbb {N}\) sufficiently large with \(n>\frac{1}{\tau _*}\), there exists \(U_{n,0}\in \mathcal {H}^1\) such that

$$\begin{aligned} \inf _{\Phi \in \mathcal {M}_{\omega ,\textbf{c}}^*(\eta )}\Vert U_{n,0}-\Phi \Vert _{\mathcal {H}^1}<\delta _n, \end{aligned}$$
(4.10)

and one of the following (i) or (ii) are satisfied, where \(\delta _n>0\) is given as above.

  1. (i)

    The time global solution to (1.1) with initial data \(U_{n,0}\) does not exists.

  2. (ii)

    There exist a time global solution \(U_n\) to (1.1) with initial data \(U_{n,0}\) and \(t_n>0\) such that

    $$\begin{aligned} \inf _{\Phi \in \mathcal {M}_{\omega ,\textbf{c}}^*(\eta )}\Vert U_n(t_n)-\Phi \Vert _{\mathcal {H}^1}\ge \epsilon . \end{aligned}$$
    (4.11)

Because \(U_{n,0}\in \mathcal {B}^{+}_{\omega _{n+},\textbf{c}_{n+}}\cap \mathcal {B}^{-}_{\omega _{n-},\textbf{c}_{n-}}\) (in particular \(U_{n,0}\in \mathcal {B}^{+}_{\omega _{n+},\textbf{c}_{n+}}\)) by (4.10) and Proposition 4.5, there exists a solution to (1.1) with initial data \(U_{n,0}\) globally in time by Theorem 1.3 and Proposition 4.2. Therefore, the case (i) does not occur.

Now we assume the case (ii). By Propositions 4.1 and  4.2, we have \(U_n(t_n)\in \mathcal {B}^{+}_{\omega _{n+},\textbf{c}_{n+}}\cap \mathcal {B}^{-}_{\omega _{n-},\textbf{c}_{n-}}\) because \(U_{n,0}\in \mathcal {B}^{+}_{\omega _{n+},\textbf{c}_{n+}}\cap \mathcal {B}^{-}_{\omega _{n-},\textbf{c}_{n-}}\). This implies

$$\begin{aligned} -2h\left( -\frac{1}{n}\right)<N(U_n(t_n))<-2h\left( \frac{1}{n}\right) . \end{aligned}$$

Thus as \(n\rightarrow \infty \) by the continuity of h, we obtain

$$\begin{aligned} N(U_n(t_n))\ \rightarrow \ -2h(0)=-2\mu _{\omega ,\textbf{c}}. \end{aligned}$$
(4.12)

Furthermore, by (4.10) and the definition of the infimum, there exists \(\Phi _{\omega ,\textbf{c},n}\in \mathcal {M}_{\omega ,\textbf{c}}^*(\eta )\) such that

$$\begin{aligned} \Vert U_{n,0}-\Phi _{\omega ,\textbf{c},n}\Vert _{\mathcal {H}^1}<2\delta _n. \end{aligned}$$
(4.13)

Since \(S_{\omega ,\textbf{c}}\) is a conserved quantity and \(\Phi _{\omega ,\textbf{c},n}\) is a minimizer of \(\mu _{\omega ,\textbf{c}}\), by (4.13), we have

$$\begin{aligned} |S_{\omega ,\textbf{c}}(U_n(t_n))-\mu _{\omega ,\textbf{c}}| =|S_{\omega ,\textbf{c}}(U_{n,0})-S_{\omega ,\textbf{c}} (\Phi _{\omega ,\textbf{c},n})| \ \rightarrow \ 0 \end{aligned}$$
(4.14)

as \(n\rightarrow \infty \). By using (1.6), (4.12), and (4.14), we get

$$\begin{aligned} K_{\omega ,c}(U_n(t_n))=N(U_n(t_n))+2S_{\omega ,\textbf{c}}(U_n(t_n))\ \rightarrow \ 0. \end{aligned}$$
(4.15)

as \(n\rightarrow \infty \). In addition, because Q and \(\textbf{P}\) are conserved quantities, by (4.13), we obtain

$$\begin{aligned} \begin{aligned} G(U_n(t_n))&=(4-2d)\omega Q(U_n(t_n))+(3-d)\textbf{c}\cdot \textbf{P}(U_n(t_n))\\&\ =(4-2d)\omega Q(U_n(0))+(3-d)\textbf{c}\cdot \textbf{P}(U_n(0))\\&\ =(4-2d)\omega Q(\Phi _{\omega ,\textbf{c},n}) +(3-d)\textbf{c}\cdot \textbf{P}(\Phi _{\omega ,\textbf{c},n}) +O(\delta _n). \end{aligned} \end{aligned}$$

These identities and the relation \(\Phi _{\omega ,\textbf{c},n}\in \mathcal {M}_{\omega ,\textbf{c}}^*(\eta )\) imply

$$\begin{aligned} \displaystyle \limsup _{n\rightarrow \infty }G(U_n(t_n)) =\displaystyle \limsup _{n\rightarrow \infty }G(\Phi _{\omega ,\textbf{c},n})\ge \eta . \end{aligned}$$
(4.16)

Therefore, by (4.14), (4.15), (4.16), and Proposition 3.3, there exist \(\{y_n\}\subset \mathbb {R}^d\) and \(V\in \mathcal {M}_{\omega ,\textbf{c}}^*(\eta )\) such that \(\{U_n(t_n,\cdot -y_n)\}\) has a subsequence which converges strongly to V in \(\mathcal {H}^1\). Therefore, by setting \(\Phi _n:=V(\cdot +y_n)\) and taking subsequence, we obtain

$$\begin{aligned} \Vert U_n(t_n)-\Phi _n\Vert _{\mathcal {H}^1}\ \rightarrow \ 0. \end{aligned}$$

This contradicts to (4.11) since \(\Phi _n\in \mathcal {M}_{\omega ,\textbf{c}}^*(\eta )\).

4.3 Stability of ground state with small speed for \(d=1\)

To prove Corollary 1.6, we introduce \(\widetilde{\mathcal {M}}_{\omega ,c_*}\) for \(c_*>0\) defined by

$$\begin{aligned} \widetilde{\mathcal {M}}_{\omega ,c_*}:=\bigcup _{|c|\le c_*}\mathcal {M}_{\omega , c}, \end{aligned}$$

and prepare the following lemma.

Lemma 4.6

Let \(d=1\) and \(\alpha ,\beta ,\gamma >0\). Assume that \((\omega ,c)\in \mathbb {R}\times \mathbb {R}\) satisfies \(\omega >\frac{\sigma c^2}{4}\). Then the following holds:

  1. (i)

    There exists \(c_1=c_1(\omega )>0\) such that \(\widetilde{\mathcal {M}}_{\omega ,c_1(\omega )}\) is bounded in \(\mathcal {H}^1\). Namely, it holds that

    $$\begin{aligned} A_{\omega }:=\sup _{\Phi \in \widetilde{\mathcal {M}}_{\omega ,c_1(\omega )}} \Vert \Phi \Vert _{\mathcal {H}^1}\in (0,\infty ). \end{aligned}$$
  2. (ii)

    There exists \(c_2=c_2(\omega )>0\) such that it holds that

    $$\begin{aligned}B_{\omega }:=\inf _{\Phi \in \widetilde{\mathcal {M}}_{\omega ,c_2(\omega )}}\omega Q(\Phi )>0. \end{aligned}$$

Proof

We first prove (i). By Theorem 1.2, there exists an element \(\Phi _{\omega }\in \mathcal {M}_{\omega ,0}\). Thus we see that \(A_{\omega }>0\). Because \(K_{\omega , 0}(\Phi _{\omega })=0\) and \(N(\Phi _{\omega })=-2\mu _{\omega ,0}\) by (1.6), for \(\lambda \in \mathbb {R}\), we have

$$\begin{aligned} \begin{aligned} K_{\omega ,c}(\lambda \Phi _{\omega })&=2\lambda ^2L(\Phi _{\omega })+3\lambda ^3N(\Phi _{\omega }) +2\lambda ^2\omega Q(\Phi _{\omega }) +2\lambda ^2cP(\Phi _{\omega })\\&=\lambda ^2K_{\omega ,0}(\Phi _{\omega }) +\lambda ^2\{3(\lambda -1)N(\Phi _{\omega })+2cP(\Phi _{\omega })\}\\&=2\lambda ^2\{-3(\lambda -1)\mu _{\omega ,0}+cP(\Phi _{\omega })\}. \end{aligned} \end{aligned}$$

Since \(\mu _{\omega ,0}> 0\) by Proposition 3.2, we can choose \(\lambda \) as \(\lambda =\lambda _0\), where \(\lambda _0:=1+\frac{1}{3\mu _{\omega ,0}}cP(\Phi _{\omega })\). Then, it holds that \(K_{\omega ,c}(\lambda _0 \Phi _{\omega })=0\). Because

$$\begin{aligned} \frac{1}{3\mu _{\omega ,0}}|cP(\Phi _{\omega })| \le \frac{1}{3\mu _{\omega ,0}}|c|\cdot 3\Vert \Phi _{\omega }\Vert _{\mathcal {H}^1}^2 =\frac{|c|}{\mu _{\omega ,0}}\Vert \Phi _{\omega }\Vert _{\mathcal {H}^1}^2, \end{aligned}$$

if \(|c|< \mu _{\omega ,0}\Vert \Phi _{\omega }\Vert _{\mathcal {H}^1}^{-2}\) holds, then we have \(0<\lambda _0 <2\). Therefore, by the definitions of \(\mu _{\omega ,c}\) and \(\lambda _0\), we get

$$\begin{aligned} \begin{aligned} \mu _{\omega ,c}&\le S_{\omega ,c}(\lambda _0 \Phi _{\omega })\\&=\lambda _0^2L(\Phi _{\omega })+\lambda _0^3N(\Phi _{\omega }) +\lambda _0^2\omega Q(\Phi _{\omega })+\lambda _0^2cP(\Phi _{\omega })\\&=\lambda _0^2S_{\omega ,0}(\Phi _{\omega }) +\lambda _0^2\{(\lambda _0-1)N(\Phi _{\omega })+cP(\Phi _{\omega })\}\\&=\lambda _0^3\mu _{\omega ,0}<8\mu _{\omega ,0}. \end{aligned} \end{aligned}$$

On the other hand, by (1.5) and Proposition 3.1, it holds that

$$\begin{aligned} \mu _{\omega ,c}=\frac{1}{6}L_{\omega ,c}(\Phi ) \ge C_{\omega }(4\omega -\sigma c^2)\Vert \Phi \Vert _{\mathcal {H}^1}^2 \end{aligned}$$
(4.17)

for any \(\Phi \in \mathcal {M}_{\omega ,c}\), where \(C_{\omega }>0\) is a constant which depends on \(\omega \) but does not depend on c.

By the above argument, we put

$$\begin{aligned} c_{1}=c_1(\omega ):=\min \left\{ \frac{\mu _{\omega ,0}}{2\Vert \Phi _{\omega }\Vert _{\mathcal {H}^1}^2}, \sqrt{\frac{2\omega }{\sigma }}\right\} . \end{aligned}$$

Then, for any \(\Phi \in \mathcal {M}_{\omega ,c}\) with \(|c|\le c_1\), we have

$$\begin{aligned} \Vert \Phi \Vert _{\mathcal {H}^1}^2 \le \frac{\mu _{\omega ,c}}{C_{\omega }(4\omega -\sigma c^2)} < \frac{4}{C_{\omega }\omega }\mu _{\omega ,0}. \end{aligned}$$

Next, we prove (ii). Let \(\Phi \in \mathcal {M}_{\omega ,c}\). By (1.6), the Hölder inequality, and the Sobolev inequality, we obtain

$$\begin{aligned} \mu _{\omega ,c} =-\frac{1}{2}N(\Phi ) \le C\Vert \Phi \Vert _{\mathcal {H}^1}^3, \end{aligned}$$

where the constant \(C>0\) does not depend on \(\omega \) and c. Therefore, by (4.17), we have

$$\begin{aligned} \frac{C_{\omega }(4\omega -\sigma c^2)}{C}\le \Vert \Phi \Vert _{\mathcal {H}^1} \end{aligned}$$

because \(\Phi \ne 0\) in \(\mathcal {H}^1\). In particular, if \(|c|\le \sqrt{\frac{2\omega }{\sigma }}\), then it holds that \(\frac{2C_{\omega }\omega }{C}\le \Vert \Phi \Vert _{\mathcal {H}^1}\). This implies that

$$\begin{aligned} L(\Phi )+\omega Q(\Phi ) \ge C'(1+\omega )\Vert \Phi \Vert _{\mathcal {H}^1}^2 \ge \frac{4C'C_{\omega }^2\omega ^2(1+\omega )}{C^2} =:\widetilde{B}_{\omega }, \end{aligned}$$

where \(C'=\min \{1,\frac{\alpha }{2},\frac{\beta }{2},\frac{\gamma }{2}\}\). Hence, at least one of

$$\begin{aligned} 2\omega Q(\Phi )\ge \widetilde{B}_{\omega }\ \ \ \ \textrm{or}\ \ \ \ 2L(\Phi )\ge \widetilde{B}_{\omega } \end{aligned}$$

holds. If the former estimate holds, then by setting \(c_2=c_2(\omega ):=\sqrt{\frac{2\omega }{\sigma }}>0\), the proof of (ii) is completed as \(B_{\omega }=\frac{\widetilde{B}_{\omega }}{2}\). On the other hand, if the latter estimate holds, then by Proposition 2.4 with \(d=1\) and \(K_{\omega ,c}(\Phi )=0\), the estimates \(\omega Q(\Phi )=L(\Phi )\ge \frac{\widetilde{B}_{\omega }}{2}\) hold.

Corollary 1.6 follows from Theorem 1.5 and the following proposition.

Proposition 4.7

(Stability of the set \(\mathcal {M}_{\omega ,c}\)) Let \(d=1\) and \(\alpha ,\beta ,\gamma >0\). Assume that \((\omega ,c)\in \mathbb {R}\times \mathbb {R}\) satisfies \(\omega >\frac{\sigma c^2}{4}\). Then, there exist \(\eta _0=\eta _0(\omega )>0\) and \(c_0=c_0(\omega )>0\) such that if \(|c|\le c_0\), then \(\mathcal {M}_{\omega ,c}^*(\eta _0)=\mathcal {M}_{\omega ,c}(\ne \emptyset )\) holds. Namely, the estimate

$$\begin{aligned} 2\omega Q(\Phi )+2cP(\Phi )\ge \eta _0 \end{aligned}$$

holds for any \(\Phi \in \mathcal {M}_{\omega ,c}\) with \(|c|\le c_0\).

Proof

Let \(c_1=c_1(\omega )>0\) and \(c_2=c_2(\omega )>\) are the constants given in Lemma 4.6. If \(|c|\le \min \{c_1,c_2\}\), then by Lemma 4.6, the estimates \(\omega Q(\Phi )\ge B_{\omega }\) and \(|P(\Phi )|\le 3\Vert \Phi \Vert _{\mathcal {H}^1}^2\le 3A_{\omega }^2 \) hold for any \(\Phi \in \mathcal {M}_{\omega ,c}\), which implies that \(2\omega Q(\Phi )+2cP(\Phi ) \ge 2B_{\omega }-6|c|A_{\omega }^2\). Here we set

$$\begin{aligned} \eta _{0}=\eta _0(\omega ):=B_{\omega },\ \ c_0=c_0(\omega ):=\min \left\{ c_1(\omega ),c_2(\omega ), \frac{B_{\omega }}{6A_{\omega }^2} \right\} >0. \end{aligned}$$

Then the estimate \(\omega Q(\Phi )+cP(\Phi ) \ge \eta _0\) holds for any \(\Phi \in \mathcal {M}_{\omega ,c}\) with \(|c|\le c_0\).