1 Introduction and main results

In this paper, we study the existence of ground state standing waves with prescribed mass for the nonlinear Schrödinger equation with combined power nonlinearities

$$\begin{aligned} i\partial _t\psi +\Delta \psi +|\psi |^{p-2}\psi +\mu |\psi |^{q-2}\psi =0,\ (t,x)\in {\mathbb {R}}\times {\mathbb {R}}^N, \end{aligned}$$
(1.1)

where \(N\ge 1, \mu >0\) and \(2<q<p\left\{ \begin{array}{ll} < 2^*:=\infty , &{} N=1,2,\\ \le 2^*:= 2N/(N-2), &{}N\ge 3. \end{array} \right. \) Starting from the fundamental contribution by T. Tao, M. Visan and X. Zhang [23], the NLS equation with combined nonlinearities attracted much attention, see for example [1, 6, 7, 11, 12, 15, 18, 19, 26].

Standing waves to (1.1) are solutions of the form \(\psi (t, x) =e^{-i\lambda t}u(x)\), where \(\lambda \in {\mathbb {R}}\) and \(u:{\mathbb {R}}^N\rightarrow {\mathbb {C}}\). Then u satisfies the equation

$$\begin{aligned} -\Delta u=\lambda u+|u|^{p-2}u+\mu |u|^{q-2}u,\ x\in {\mathbb {R}}^{N}. \end{aligned}$$
(1.2)

A possible choice is to fix \(\lambda \in {\mathbb {R}}\) and to search for solutions to (1.2) as critical points of the action functional

$$\begin{aligned} J_{p,q}(u):=\int _{{\mathbb {R}}^N}\left( \frac{1}{2}|\nabla u|^2-\frac{\lambda }{2}|u|^2-\frac{1}{p}|u|^p-\frac{\mu }{q}|u|^q\right) dx, \end{aligned}$$

see for example [2, 17] and the references therein.

Alternatively, one can search for solutions to (1.2) having prescribed mass

$$\begin{aligned} \int _{{\mathbb {R}}^N}|u|^2dx=a^2. \end{aligned}$$
(1.3)

In this direction, define on \(H:= H^1({\mathbb {R}}^N,{\mathbb {C}})\) the energy functional

$$\begin{aligned} E_{p,q}(u)=\frac{1}{2}\int _{{\mathbb {R}}^N}|\nabla u|^2dx-\frac{1}{p}\int _{{\mathbb {R}}^N}|u|^{p}dx -\frac{\mu }{q}\int _{{\mathbb {R}}^N}|u|^{q}dx. \end{aligned}$$

It is standard to check that \(E_{p,q}\in C^1\) and a critical point of \(E_{p,q}\) constrained to

$$\begin{aligned} S_a=\{u\in H^1({\mathbb {R}}^N,{\mathbb {C}}):\int _{{\mathbb {R}}^N}|u|^2=a^2\} \end{aligned}$$

gives rise to a solution to (1.2), satisfying (1.3). Such solution is usually called a normalized solution of (1.2). In this method, the parameter \(\lambda \in {\mathbb {R}}\) arises as a Lagrange multiplier, which depends on the solution and is not a priori given. In this paper, we will focus on the existence of normalized ground state of (1.2), defined as follows:

Definition 1.1

We say that u is a normalized ground state to (1.2) on \(S_a\) if

$$\begin{aligned} E_{p,q}(u)=z_{p,q}:=\inf \{E_{p,q}(v):v\in S_a,\ (E_{p,q}|_{S_a})'(v)=0\}. \end{aligned}$$

The set of the normalized ground states will be denoted by \({\mathcal {Z}}_{p,q}\).

In the study of (1.2-1.3) an important role is played by the so-called \(L^2\)-critical exponent

$$\begin{aligned} {\bar{p}}=2+\frac{4}{N}. \end{aligned}$$

A very complete analysis of the various cases that may happen for (1.2-1.3), depending on the values of (pq), has been provided recently in [4, 9, 10, 21, 22]. See [21] for the cases \(N\ge 1\) and \(p<2^*\), [9, 10, 22] for the cases \(N\ge 3\) and \(p=2^*\), and [4] for the cases \(N=1\), \(p=+\infty \) and \(q\le 6\). See [20] for the Schrödinger equation with combined nonlinearities on metric graphs. For a \(L^2\)-critical or \(L^2\)-supercritical perturbation \(q\ge {\bar{p}}\) and the Sobolev subcritical case \(p<2^*\), [21] obtained the following results to (1.2):

Theorem 1.2

Let \(N\ge 1\), \(a>0\), \(\mu >0\) and \({\bar{p}}\le q<p<2^*\). If \(q={\bar{p}}\), we further assume that \(\mu a^{\frac{4}{N}}<({\bar{a}}_N)^{\frac{4}{N}}\), where \({\bar{a}}_N\) is defined in (2.1). Then \(E_{p,q}|_{S_a}\) has a critical point u at positive level \(E_{p,q}(u)>0\), with the following properties: u is a real-valued positive function in \({\mathbb {R}}^N\), is radially symmetric, is radially non-increasing, solves (1.2) for some \(\lambda <0\), and is a normalized ground state of (1.2) on \(S_a\).

Remark 1.3

In fact, [21] did not consider the case \(q>{\bar{p}}\) of Theorem 1.2, while it also holds by repeating the proof for the case \(q={\bar{p}}\). In this paper, we will give Theorem 1.2 another proof, which is useful to the proof of Theorem 1.4, so we write it here in a unified form.

However, for the \(L^2\)-supercritical and Sobolev critical case \({\bar{p}}<q<p=2^*\), a condition \(\mu a^{N+q-Nq/2}<\alpha (N,q)\) is added to get similar results as to Theorem 1.2, where \(\alpha (N,q)\) is finite for \(N\ge 5\), see [22] for more details. Inspired by the results of the unconstrained problem considered in [14] and [17], we guess that the condition maybe can be removed when q is close to \(2^*\). Fortunately, we succeed to do it in the full interval \({\bar{p}}<q<2^*\) and obtain similar results as Theorem 1.2 for the Sobolev critical problem. Our result settles an open question raised by N. Soave [22].

Theorem 1.4

Let \(N\ge 3\), \(a>0\), \(\mu >0\) and \({\bar{p}}\le q<p=2^*\). If \(q={\bar{p}}\), we further assume that \(\mu a^{\frac{4}{N}}<({\bar{a}}_N)^{\frac{4}{N}}\). Then \(E_{p,q}|_{S_a}\) has a critical point u at positive level \(0<E_{p,q}(u)<\frac{1}{N}S^{\frac{N}{2}}\), with the following properties: u is a real-valued positive function in \({\mathbb {R}}^N\), is radially symmetric, is radially non-increasing, solves (1.2) for some \(\lambda <0\), and is a normalized ground state of (1.2) on \(S_a\). Here S is defined in (3.2).

Remark 1.5

In Theorem 1.4, we only improve the result of [22] for the case \(q>{\bar{p}}\), while it is the same as that of [22] in the case \(q={\bar{p}}\). Since the proof will be done in a uniform way, we write it here.

Remark 1.6

When \(q>{\bar{p}}\), similarly to [22], to prove Theorem 1.4, a key step is to show that \(c_{2^*,q}<\frac{1}{N}S^{\frac{N}{2}}\), which will be obtained by choosing appropriate functions. To do this, in Lemma 6.4 of [22], they first constructed \(u_\epsilon \) and \(v_\epsilon :=a\frac{u_\epsilon (x)}{\Vert u_\epsilon \Vert _2}\), and then estimated the maximum of \(\Psi _{v_\epsilon }(\tau ):=E_{2^*,q}((v_\epsilon )^{\tau })\). In view of the expression of \(\Psi _{v_\epsilon }(\tau )\) and the estimates of \(u_\epsilon \), the lower bound of the maximum point \(\tau _{v_\epsilon }\) of \(\Psi _{v_\epsilon }(\tau )\) was needed and thus a condition \(\mu a^{N+q-Nq/2}<\alpha (N,q)\) was added for \(N\ge 5\). To remove this condition, in this paper, we will use a different transformation to define \(v_\epsilon :=(a^{-1}\Vert u_\epsilon \Vert _2)^{\frac{N-2}{2}}u_\epsilon (a^{-1}\Vert u_\epsilon \Vert _2x)\) and subsequently obtain a different expression of \(\Psi _{v_\epsilon }(\tau )\) (see (3.3)). In this case, by using the estimates of \(u_\epsilon \) and the fact that \(c_{2^*,q}>0\), we can easily show that \(\tau _{v_\epsilon }\in [\tau _0,\tau _1]\) with \(\tau _0,\tau _1>0\) and then obtain the upper bound of \(c_{2^*,q}\) without adding additional conditions, see Lemma 3.3.

Remark 1.7

Following the proof of Theorem 1.7 in [21] word by word, we can show that under the assumptions of Theorems 1.2 or 1.4,

$$\begin{aligned} {\mathcal {Z}}_{p,q}=\{e^{i\theta }|u| \ \mathrm {for \ some}\ \theta \in {\mathbb {R}}\ \mathrm {and}\ |u|>0\ \mathrm {in} \ {\mathbb {R}}^N\} \end{aligned}$$

and for any \(u\in {\mathcal {Z}}_{p,q}\), the standing wave \(e^{-i\lambda t}u(x)\) is strongly unstable.

Remark 1.8

By Lemma 2.6, any normalized ground state u of (1.2) satisfies equation (1.2) with some \(\lambda =\lambda (u)<0\). For such fixed \(\lambda \), it is natural to consider the ground state of (1.2), which is a solution \(w\in H^1({\mathbb {R}}^N,{\mathbb {C}})\backslash \{0\}\) of (1.2) satisfying

$$\begin{aligned} J_{p,q}(w)=\inf \{J_{p,q}(v):v\in H^1({\mathbb {R}}^N,{\mathbb {C}})\backslash \{0\},\ J_{p,q}'(v)=0\}. \end{aligned}$$

It is an open question whether a normalized ground state of (1.2) is a ground state of (1.2) with fixed \(\lambda <0\).

In the proofs of Theorems 1.2 and 1.4, the Pohožaev set

$$\begin{aligned} {\mathcal {P}}_{p,q}=\{u\in S_a:P_{p,q}(u)=0\}, \end{aligned}$$

plays an important role, where

$$\begin{aligned} P_{p,q}(u)=\int _{{\mathbb {R}}^N}|\nabla u|^2dx-\gamma _p\int _{{\mathbb {R}}^N}|u|^{p}dx -\mu \gamma _q\int _{{\mathbb {R}}^N}|u|^{q}dx \end{aligned}$$

and

$$\begin{aligned} \gamma _p=\frac{N(p-2)}{2p}=\frac{N}{2}-\frac{N}{p}. \end{aligned}$$

It is well known that any critical point of \(E_{p,q}|_{S_a}\) belongs to \({\mathcal {P}}_{p,q}\), as a consequence of the Pohožaev identity (we refer for instance to Lemma 2.7 in [8]). Moreover, \(P_{p,q}\) is a natural constraint, see Lemma 2.6. So it is natural to consider the minimizing problem

$$\begin{aligned} c_{p,q}=\inf _{u\in {\mathcal {P}}_{p,q}}E_{p,q}(u) \end{aligned}$$

and define

$$\begin{aligned} {\mathcal {C}}_{p,q}=\{u\in {\mathcal {P}}_{p,q}: E_{p,q}(u)=c_{p,q}\}. \end{aligned}$$

For the Sobolev subcritical problem, we can show that \(c_{p,q}\) is attained by using Schwartz symmetrization rearrangements. For the Sobolev critical problem, we can show that \(c_{p,q}\) is attained, by introducing the Sobolev subcritical approximation method, which has already been used to deal with problems without mass constraint (see [13, 14, 17]). To our knowledge, it is the first time this method is used to discuss mass constrained problems. During the proofs, the following various expressions of \(E_{p,q}(u)\) constrained on \({\mathcal {P}}_{p,q}\)

$$\begin{aligned} \begin{aligned} E_{p,q}(u)&=\left( \frac{1}{2}-\frac{1}{p\gamma _{p}}\right) \int _{{\mathbb {R}}^N}|\nabla u|^2dx+\left( \frac{\gamma _q}{p\gamma _{p}}-\frac{1}{q}\right) \mu \int _{{\mathbb {R}}^N}| u|^qdx\\&=\left( \frac{1}{2}-\frac{1}{q\gamma _q}\right) \int _{{\mathbb {R}}^N}|\nabla u|^2dx+\left( \frac{\gamma _p}{q\gamma _q}-\frac{1}{p}\right) \int _{{\mathbb {R}}^N}| u|^pdx\\&=\left( \frac{\gamma _p}{2}-\frac{1}{p}\right) \int _{{\mathbb {R}}^N}|u|^pdx +\left( \frac{\gamma _q}{2}-\frac{1}{q}\right) \mu \int _{{\mathbb {R}}^N}|u|^qdx \end{aligned} \end{aligned}$$

play an important role.

This paper is organized as follows. In Sect. 2, we cite some preliminaries and give the proof of Theorem 1.2. Section 3 is devoted to the proof of Theorem 1.4.

Notation: For \(t\ge 1\), the \(L^t\)-norm of \(u\in L^t({\mathbb {R}}^N,{\mathbb {C}})\) (or of \(L^t({\mathbb {R}}^N,{\mathbb {R}})\)) is denoted by \(\Vert u\Vert _t\). We simply write H for \(H^1({\mathbb {R}}^N,{\mathbb {C}})\), and \(H^1\) for the subspace of real valued functions \(H^1({\mathbb {R}}^N,{\mathbb {R}})\).

2 Preliminaries and proof of Theorem 1.2

The following Gagliardo-Nirenberg inequality can be found in [24].

Lemma 2.1

Let \(N\ge 1\) and \(2<p<2^*\), then the following sharp Gagliardo-Nirenberg inequality

$$\begin{aligned} \Vert u\Vert _{p}\le C_{N,p}\Vert u\Vert _2^{1-\gamma _p}\Vert \nabla u\Vert _2^{\gamma _p} \end{aligned}$$

holds for any \(u\in H\), where the sharp constant \(C_{N,p}\) is

$$\begin{aligned} C_{N,p}^{p}=\frac{2p}{2N+(2-N)p}\left( \frac{2N+(2-N)p}{N(p-2)}\right) ^{\frac{N(p-2)}{4}}\frac{1}{\Vert Q_{p}\Vert _2^{p-2}} \end{aligned}$$

and \(Q_p\) is the unique positive radial solution of equation

$$\begin{aligned} -\Delta Q+Q=|Q|^{p-2}Q. \end{aligned}$$

In the special case \(p={\bar{p}}\), \(C_{N,{\bar{p}}}^{{\bar{p}}}=\frac{{\bar{p}}}{2}\frac{1}{\Vert Q_{{\bar{p}}}\Vert _2^{4/N}}\), or equivalently,

$$\begin{aligned} \Vert Q_{{\bar{p}}}\Vert _2=\left( \frac{{\bar{p}}}{2C_{N,{\bar{p}}}^{{\bar{p}}}}\right) ^{N/4}=:{\bar{a}}_N. \end{aligned}$$
(2.1)

The following lemma is useful in concerning the uniform bound of radial non-increasing functions, see [3] for its proof.

Lemma 2.2

Let \(N\ge 3\) and \( 1\le t<+\infty \). If \(u\in L^t({\mathbb {R}}^N)\) is a radial non-increasing function (i.e. \(0\le u(x)\le u(y)\) if \(|x|\ge |y|\)), then one has

$$\begin{aligned} |u(x)|\le |x|^{-N/t}\left( \frac{N}{|S^{N-1}|}\right) ^{1/t}\Vert u\Vert _t, \ x\ne 0, \end{aligned}$$

where \(|S^{N-1}|\) is the area of the unit sphere in \({\mathbb {R}}^N\).

For any \(u\in S_a\) and \(\tau >0\), we define

$$\begin{aligned} u^{\tau }(x)=\tau ^{N/2}u(\tau x). \end{aligned}$$
(2.2)

Then \(u^\tau \in S_a\) and for any \(\tau >0\),

$$\begin{aligned} E_{p,q}(u^\tau )=\frac{1}{2}\tau ^2\int _{{\mathbb {R}}^N}|\nabla u|^2dx-\frac{1}{p}\tau ^{\frac{N}{2}p-N}\int _{{\mathbb {R}}^N}|u|^{p}dx -\frac{\mu }{q}\tau ^{\frac{N}{2}q-N}\int _{{\mathbb {R}}^N}|u|^{q}dx \end{aligned}$$
(2.3)

and

$$\begin{aligned} P_{p,q}(u^\tau )=\tau ^2\int _{{\mathbb {R}}^N}|\nabla u|^2dx-\gamma _p\tau ^{\frac{N}{2}p-N}\int _{{\mathbb {R}}^N}|u|^{p}dx -\mu \gamma _q\tau ^{\frac{N}{2}q-N}\int _{{\mathbb {R}}^N}|u|^{q}dx. \end{aligned}$$

The following lemma is about the properties of \(E_{p,q}(u^\tau )\) and \(P_{p,q}(u^\tau )\).

Lemma 2.3

Let \(N\ge 1\), \(a>0\), \(\mu >0\) and

$$\begin{aligned} {\bar{p}}\le q<p\left\{ \begin{array}{ll} < \infty , &{} N=1,2,\\ \le 2^*, &{}N\ge 3. \end{array} \right. \end{aligned}$$

If \(q={\bar{p}}\), we further assume that \(\mu a^{\frac{4}{N}}<({\bar{a}}_N)^{\frac{4}{N}}\). Then for any \(u\in S_a\), there exists a unique \(\tau _0\in (0,\infty )\) such that \(P_{p,q}(u^{\tau _0})=0\). Moreover, \(\tau _0\) is the unique critical point of \(E_{p,q}(u^\tau )\) and \(E_{p,q}(u^{\tau _0})=\max _{\tau \in (0,\infty )}E_{p,q}(u^\tau )\). In particular, if \(P_{p,q}(u)\le 0\), then \(\tau _0\in (0,1]\).

Proof

Set \(P_{p,q}(u^{\tau })=\tau ^2g(\tau )\), where

$$\begin{aligned} g(\tau )=\int _{{\mathbb {R}}^N}|\nabla u|^2dx-\gamma _p\tau ^{\frac{N}{2}p-N-2}\int _{{\mathbb {R}}^N}|u|^{p}dx -\mu \gamma _q\tau ^{\frac{N}{2}q-N-2}\int _{{\mathbb {R}}^N}|u|^{q}dx. \end{aligned}$$

When \({\bar{p}}<q<p\), we have \(\frac{N}{2}p-N-2>\frac{N}{2}q-N-2>0\). When \({\bar{p}}=q<p\) and \(\mu a^{\frac{4}{N}}<({\bar{a}}_N)^{\frac{4}{N}}\), we have \(\frac{N}{2}p-N-2>\frac{N}{2}q-N-2=0\) and by the Gagliardo-Nirenberg inequality,

$$\begin{aligned} \mu \gamma _q\int _{{\mathbb {R}}^N}|u|^{q}dx\le \mu \gamma _qC_{N,q}^qa^{q(1-\gamma _q)}\Vert \nabla u\Vert _2^2<\Vert \nabla u\Vert _2^2. \end{aligned}$$

Hence, in both cases, \(g(\tau )>0\) for \(\tau >0\) small enough, \(g(\tau )<0\) for \(\tau \) large enough, and \(g'(\tau )<0\) for \(\tau \in (0,\infty )\). So \(g(\tau )\) has a unique zero \(\tau _0\) as well as \(P_{p,q}(u^{\tau })\).

By direct calculations, we have \(E_{p,q}'(u^\tau )=\tau ^{-1}P_{p,q}(u^\tau )\), \(E_{p,q}(u^\tau )>0\) for \(\tau >0\) small enough and \(\lim _{\tau \rightarrow \infty }E_{p,q}(u^\tau )=-\infty \). Thus, \(\tau _0\) is the unique critical point of \(E_{p,q}(u^\tau )\) and \(E_{p,q}(u^{\tau _0})=\max _{\tau \in (0,\infty )}E_{p,q}(u^\tau )\). \(\square \)

The following lemmas are about the properties of \(c_{p,q}\) and \({\mathcal {C}}_{p,q}\).

Lemma 2.4

Let \(N\ge 1\), \(a>0\), \(\mu >0\) and

$$\begin{aligned} {\bar{p}}\le q<p\left\{ \begin{array}{ll} < \infty , &{} N=1,2,\\ \le 2^*, &{}N\ge 3. \end{array} \right. \end{aligned}$$

If \(q={\bar{p}}\), we further assume that \(\mu a^{\frac{4}{N}}<({\bar{a}}_N)^{\frac{4}{N}}\). Then \(c_{p,q}>0\).

Proof

By Lemma 2.3, \({\mathcal {P}}_{p,q}\ne \emptyset \).

Case 1 (\(p\ne 2^*\)). For any \(u\in {\mathcal {P}}_{p,q}\), by the Gagliardo-Nirenberg inequality (Lemma 2.1), we have

$$\begin{aligned} \begin{aligned} \int _{{\mathbb {R}}^N}|\nabla u|^2dx&=\gamma _p\int _{{\mathbb {R}}^N}|u|^{p}dx +\mu \gamma _q\int _{{\mathbb {R}}^N}|u|^{q}dx\\&\le \gamma _p C_{N,p}^p\Vert u\Vert _2^{p(1-\gamma _p)}\Vert \nabla u\Vert _2^{p\gamma _p}+\mu \gamma _qC_{N,q}^q\Vert u\Vert _2^{q(1-\gamma _q)}\Vert \nabla u\Vert _2^{q\gamma _q}\\&=\mu \gamma _qC_{N,q}^qa^{q(1-\gamma _q)}\Vert \nabla u\Vert _2^{q\gamma _q}+\gamma _p C_{N,p}^pa^{p(1-\gamma _p)}\Vert \nabla u\Vert _2^{p\gamma _p}. \end{aligned} \end{aligned}$$
(2.4)

If \({\bar{p}}<q<p\), then \(p\gamma _p>q\gamma _q>2\). (2.4) implies that there exists a constant \(C>0\) such that \(\Vert \nabla u\Vert _2^2\ge C\). Consequently,

$$\begin{aligned} \gamma _p\int _{{\mathbb {R}}^N}|u|^{p}dx +\mu \gamma _q\int _{{\mathbb {R}}^N}|u|^{q}dx\ge C. \end{aligned}$$

If \({\bar{p}}=q<p\) and \(\mu a^{\frac{4}{N}}<({\bar{a}}_N)^{\frac{4}{N}}\), then \(p\gamma _p>q\gamma _q=2,\ \mu \gamma _qC_{N,q}^qa^{q(1-\gamma _q)}<1\). (2.4) implies that there exists a constant \(C>0\) such that \(\Vert \nabla u\Vert _2^2\ge C\). Thus, it follows from (2.4) that

$$\begin{aligned} \gamma _p\int _{{\mathbb {R}}^N}|u|^{p}dx\ge \left( 1-\mu \gamma _qC_{N,q}^qa^{q(1-\gamma _q)}\right) \Vert \nabla u\Vert _2^2\ge C\left( 1-\mu \gamma _qC_{N,q}^qa^{q(1-\gamma _q)}\right) . \end{aligned}$$

Any way, there always exists \(C_1>0\) such that for any \(u\in {\mathcal {P}}_{p,q}\),

$$\begin{aligned} E_{p,q}(u)=\left( \frac{\gamma _p}{2}-\frac{1}{p}\right) \int _{{\mathbb {R}}^N}|u|^pdx +\left( \frac{\gamma _q}{2}-\frac{1}{q}\right) \mu \int _{{\mathbb {R}}^N}|u|^qdx\ge C_1, \end{aligned}$$
(2.5)

which implies \(c_{p,q}>0\).

Case 2 (\(p= 2^*\)). Similarly to Case 1, just in (2.4), we estimate the term \(\int _{{\mathbb {R}}^N}|u|^{2^*}dx\) by using

$$\begin{aligned} \int _{{\mathbb {R}}^N}|u|^{2^*}dx\le \left( \frac{\int _{{\mathbb {R}}^N}|\nabla u|^2dx}{S}\right) ^{\frac{N}{N-2}}, \end{aligned}$$

see (3.2). \(\square \)

Lemma 2.5

Let \(N\ge 1\), \(a>0\), \(\mu >0\) and \({\bar{p}}\le q<p<2^*\). If \(q={\bar{p}}\), we further assume that \(\mu a^{\frac{4}{N}}<({\bar{a}}_N)^{\frac{4}{N}}\). Then \(c_{p,q}\) is attained by a real-valued positive, radially symmetric and radially non-increasing function.

Proof

Let \(\{u_n\}_{n=1}^{\infty }\subset {\mathcal {P}}_{p,q}\) be a minimizing sequence of \(c_{p,q}\) and \(|u_n|^*\) be the Schwartz symmetrization rearrangement of \(|u_n|\). From Chapter 3 in [16], we have

$$\begin{aligned} \int _{{\mathbb {R}}^N}|\nabla (|u_n|^*)|^2dx\le \int _{{\mathbb {R}}^N}|\nabla |u_n||^2dx\le \int _{{\mathbb {R}}^N}|\nabla u_n|^2dx \end{aligned}$$

and

$$\begin{aligned} \int _{{\mathbb {R}}^N}||u_n|^*|^tdx=\int _{{\mathbb {R}}^N}|u_n|^tdx,\ t\in \left[ 1, \infty \right) . \end{aligned}$$

Hence \(P_{p,q}(|u_n|^*)\le 0\).

Let \((|u_n|^*)^\tau (x)\) be defined as (2.2). By Lemma 2.3, there exists a unique \(\tau _n\in (0,1]\) such that \(P_{p,q}((|u_n|^*)^{\tau _n})=0\). Hence \(\{(|u_n|^*)^{\tau _n}\}_{n=1}^{\infty }\subset {\mathcal {P}}_{p,q}\). By direct calculations, we have

$$\begin{aligned} \begin{aligned} E&_{p,q}((|u_n|^*)^{\tau _n})\\&=\left( \frac{\gamma _p}{2}-\frac{1}{p}\right) \int _{{\mathbb {R}}^N}|(|u_n|^*)^{\tau _n}|^pdx +\left( \frac{\gamma _q}{2}-\frac{1}{q}\right) \mu \int _{{\mathbb {R}}^N}|(|u_n|^*)^{\tau _n}|^qdx\\&=\tau _n^{\frac{N}{2}p-N}\left( \frac{\gamma _p}{2}-\frac{1}{p}\right) \int _{{\mathbb {R}}^N}||u_n|^*|^pdx+ \tau _n^{\frac{N}{2}q-N}\left( \frac{\gamma _q}{2}-\frac{1}{q}\right) \mu \int _{{\mathbb {R}}^N}||u_n|^*|^qdx\\&\le E_{p,q}(u_n). \end{aligned} \end{aligned}$$
(2.6)

That is, \(\{(|u_n|^*)^{\tau _n}\}_{n=1}^{\infty }\) is a minimizing sequence of \(c_{p,q}\). Reversing the proof of Lemma 2.4, we can show that \(\{(|u_n|^*)^{\tau _n}\}_{n=1}^{\infty }\) is bounded in \(H^1({\mathbb {R}}^N)\). Hence, there exists \(u_0\in H^1({\mathbb {R}}^N)\) such that \((|u_n|^*)^{\tau _n}\rightharpoonup u_0\) weakly in \(H^1({\mathbb {R}}^N)\), \((|u_n|^*)^{\tau _n}\rightarrow u_0\) strongly in \(L^t({\mathbb {R}}^N)\) with \(t\in \left( 2,2^*\right) \) and \((|u_n|^*)^{\tau _n}\rightarrow u_0\) a.e. in \({\mathbb {R}}^N\). Consequently,

$$\begin{aligned}&\int _{{\mathbb {R}}^N}|u_0|^2dx\le \liminf _{n\rightarrow \infty }\int _{{\mathbb {R}}^N}|(|u_n|^*)^{\tau _n}|^2dx=a^2, \\&\quad \int _{{\mathbb {R}}^N}|\nabla u_0|^2dx\le \liminf _{n\rightarrow \infty }\int _{{\mathbb {R}}^N}|\nabla (|u_n|^*)^{\tau _n}|^2dx, \\&\quad E_{p,q}((|u_n|^*)^{\tau _n})\rightarrow \left( \frac{\gamma _p}{2}-\frac{1}{p}\right) \int _{{\mathbb {R}}^N}|u_0|^pdx +\left( \frac{\gamma _q}{2}-\frac{1}{q}\right) \mu \int _{{\mathbb {R}}^N}|u_0|^qdx=c_{p,q}, \end{aligned}$$

which imply that \(u_0\not \equiv 0\) and \(P_{p,q}(u_0)\le 0\).

Set \(\int _{{\mathbb {R}}^N}| u_0|^2dx:=c_0^2\le a^2\) and define \({\tilde{u}}(x)=(c_0a^{-1})^{\frac{2}{p-2}}u_0((c_0a^{-1})^{\frac{2p}{N(p-2)}}x)\). Then

$$\begin{aligned}&\int _{{\mathbb {R}}^N}|{\tilde{u}}|^2dx=a^2,\ \int _{{\mathbb {R}}^N}|{\tilde{u}}|^pdx=\int _{{\mathbb {R}}^N}|u_0|^pdx, \\&\quad \int _{{\mathbb {R}}^N}|{\tilde{u}}|^qdx=(c_0a^{-1})^{\frac{2(q-p)}{p-2}}\int _{{\mathbb {R}}^N}|u_0|^qdx\ge \int _{{\mathbb {R}}^N}|u_0|^qdx, \\&\quad \int _{{\mathbb {R}}^N}|\nabla {\tilde{u}}|^2dx=(c_0a^{-1})^{\frac{2[2N+p(2-N)]}{N(p-2)}}\int _{{\mathbb {R}}^N}|\nabla u_0|^2dx\le \int _{{\mathbb {R}}^N}|\nabla u_0|^2dx. \end{aligned}$$

Hence \(P_{p,q}({\tilde{u}})\le 0\). So there exists \(\tau _0\in (0,1]\) such that \({\tilde{u}}^{\tau _0}\in {\mathcal {P}}_{p,q}\) and

$$\begin{aligned} \begin{aligned} E_{p,q}({\tilde{u}}^{\tau _0})&=\left( \frac{1}{2}-\frac{1}{q\gamma _q}\right) \int _{{\mathbb {R}}^N}|\nabla ({\tilde{u}}^{\tau _0})|^2dx+\left( \frac{\gamma _p}{q\gamma _q}-\frac{1}{p}\right) \int _{{\mathbb {R}}^N}| ({\tilde{u}}^{\tau _0})|^pdx\\&=\left( \frac{1}{2}-\frac{1}{q\gamma _q}\right) \tau _0^2\int _{{\mathbb {R}}^N}|\nabla {\tilde{u}}|^2dx+\left( \frac{\gamma _p}{q\gamma _q}-\frac{1}{p}\right) \tau _0^{\frac{N}{2}p-N}\int _{{\mathbb {R}}^N}| {\tilde{u}}|^pdx\\&= \left( \frac{1}{2}-\frac{1}{q\gamma _q}\right) \tau _0^2(c_0a^{-1})^{\frac{2[2N+p(2-N)]}{N(p-2)}} \int _{{\mathbb {R}}^N}|\nabla u_0|^2dx\\&\qquad +\left( \frac{\gamma _p}{q\gamma _q}-\frac{1}{p}\right) \tau _0^{\frac{N}{2}p-N}\int _{{\mathbb {R}}^N}| u_0|^pdx\\&\le \left( \frac{1}{2}-\frac{1}{q\gamma _q}\right) \int _{{\mathbb {R}}^N}|\nabla u_0|^2dx+\left( \frac{\gamma _p}{q\gamma _q}-\frac{1}{p}\right) \int _{{\mathbb {R}}^N}| u_0|^pdx\\&\le \liminf _{n\rightarrow \infty }\left\{ \left( \frac{1}{2}-\frac{1}{q\gamma _q}\right) \int _{{\mathbb {R}}^N}|\nabla (|u_n|^*)^{\tau _n}|^2dx\right. \\&\qquad \qquad \qquad \qquad \left. +\left( \frac{\gamma _p}{q\gamma _q}-\frac{1}{p}\right) \int _{{\mathbb {R}}^N}| (|u_n|^*)^{\tau _n}|^pdx\right\} \\&= c_{p,q}. \end{aligned} \end{aligned}$$
(2.7)

By the definition of \(c_{p,q}\), we obtain that \(E_{p,q}({\tilde{u}}^{\tau _0})=c_{p,q}\), \(\tau _0=1\) and \(c_0=a\). Hence, \(u_0\in {\mathcal {P}}_{p,q}\) is a real-valued nonnegative, radially symmetric and radially non-increasing minimizer of \(c_{p,q}\). By the strong maximum principle, \(u_0>0\) in \({\mathbb {R}}^N\). \(\square \)

Lemma 2.6

Let \(N\ge 1\), \(a>0\), \(\mu >0\) and

$$\begin{aligned} {\bar{p}}\le q<p\left\{ \begin{array}{ll} < \infty , &{} N=1,2,\\ \le 2^*, &{}N\ge 3. \end{array} \right. \end{aligned}$$

If \({\mathcal {C}}_{p,q}\) is not empty, then for any \(u\in {\mathcal {C}}_{p,q}\), there exists \(\lambda <0\) such that u satisfies equation (1.2). Moreover, \({\mathcal {C}}_{p,q}={\mathcal {Z}}_{p,q}\) and \(|u|\in {\mathcal {C}}_{p,q}\).

Proof

For any \(u\in {\mathcal {C}}_{p,q}\), there exist \(\lambda \) and \(\eta \) such that

$$\begin{aligned} -\Delta u-|u|^{p-2}u-\mu |u|^{q-2}u=\lambda u+\eta [-2\Delta u-p\gamma _p|u|^{p-2}u-\mu q\gamma _q|u|^{q-2}u],\nonumber \\ \end{aligned}$$
(2.8)

or equivalently,

$$\begin{aligned} -(1-2\eta )\Delta u=\lambda u+(1-\eta p\gamma _p)|u|^{p-2}u+\mu (1-\eta q\gamma _q) |u|^{q-2}u. \end{aligned}$$

Next we show \(\eta =0\). Similarly to the definition of \(P_{p,q}(u)\), we obtain

$$\begin{aligned} (1-2\eta )\int _{{\mathbb {R}}^N}|\nabla u|^2dx-(1-\eta p\gamma _p)\gamma _p \int _{{\mathbb {R}}^N}|u|^{p}dx-(1-\eta q\gamma _q)\mu \gamma _q \int _{{\mathbb {R}}^N}|u|^{q}dx=0, \end{aligned}$$

which combined with \(P_{p,q}(u)=0\) gives that

$$\begin{aligned} \eta \left( 2\int _{{\mathbb {R}}^N}|\nabla u|^2dx-p\gamma _p^2\int _{{\mathbb {R}}^N}|u|^{p}dx-\mu q\gamma _q^2\int _{{\mathbb {R}}^N}|u|^{q}dx\right) =0. \end{aligned}$$

If \(\eta \ne 0\), then

$$\begin{aligned} 2\int _{{\mathbb {R}}^N}|\nabla u|^2dx-p\gamma _p^2\int _{{\mathbb {R}}^N}|u|^{p}dx-\mu q\gamma _q^2\int _{{\mathbb {R}}^N}|u|^{q}dx=0, \end{aligned}$$

which combined with \(P_{p,q}(u)=0\) gives that

$$\begin{aligned} \int _{{\mathbb {R}}^N}|u|^{p}dx=\frac{2-q\gamma _q}{\gamma _p(p\gamma _p-q\gamma _q)}\int _{{\mathbb {R}}^N}|\nabla u|^2dx\le 0. \end{aligned}$$

That is a contradiction. So \(\eta =0\).

From (2.8), \(P_{p,q}(u)=0\), \(0< \gamma _q<\gamma _p\le 1\) and \(\mu >0\), we obtain

$$\begin{aligned} \begin{aligned} \lambda a^2&=\int _{{\mathbb {R}}^N}|\nabla u|^2dx- \int _{{\mathbb {R}}^N}|u|^{p}dx-\mu \int _{{\mathbb {R}}^N}|u|^{q}dx\\&=(\gamma _p-1)\int _{{\mathbb {R}}^N}|u|^{p}dx+\mu (\gamma _q-1) \int _{{\mathbb {R}}^N}|u|^{q}dx<0. \end{aligned} \end{aligned}$$

Hence \(\lambda <0\).

Any normalized solution v of (1.2) satisfies \(P_{p,q}(v)=0\). Hence \(E_{p,q}(v)\ge c_{p,q}\) and then \(c_{p,q}=z_{p,q}\), \({\mathcal {C}}_{p,q}={\mathcal {Z}}_{p,q}\). Since \(\int _{{\mathbb {R}}^N}|\nabla |u||^2dx\le \int _{{\mathbb {R}}^N}|\nabla u|^2dx\), we have \(P_{p,q}(|u|)\le 0\). So there exists \(\tau _0\in (0,1]\) such that \(|u|^{\tau _0}\in {\mathcal {P}}_{p,q}\). Similarly to the proof of (2.6), we can show that \(\tau _0=1\) and \(|u|\in {\mathcal {C}}_{p,q}\). \(\square \)

Proof of Theorem 1.2: It follows from Lemmas 2.42.6.

3 Proof of Theorem 1.4

In this section, we first study the properties of \(c_{p,q}\) and then give the proof of Theorem 1.4.

Lemma 3.1

Let \(N\ge 3\), \(a>0\), \(\mu >0\) and \({\bar{p}}\le q<p<2^*\). If \(q={\bar{p}}\), we further assume that \(\mu a^{\frac{4}{N}}<({\bar{a}}_N)^{\frac{4}{N}}\). Then \(\limsup _{p\rightarrow 2^*}c_{p,q}\le c_{2^*,q}\).

Proof

By the definition of \(c_{2^*,q}\), for any fixed \(\epsilon \in (0,1)\), there exists \(u\in {\mathcal {P}}_{2^*,q}\) such that \(E_{2^*,q}(u)<c_{2^*,q}+\epsilon \). By (2.3), there exists \(\tau _0>0\) large enough such that \(E_{2^*,q}(u^{\tau _0})\le -2\). By the Young inequality

$$\begin{aligned} |u|^{p}\le \frac{2^*-p}{2^*-q}|u|^{q}+ \frac{p-q}{2^*-q}|u|^{2^*} \end{aligned}$$
(3.1)

and the Lebesgue dominated convergence theorem, we know

$$\begin{aligned} \frac{1}{p}\tau ^{\frac{N}{2}p-N}\int _{{\mathbb {R}}^N}|u|^{p}dx \end{aligned}$$

is continuous on \(p\in [{\bar{p}},2^*]\) uniformly with \(\tau \in [0,\tau _0]\). Hence, there exists \(\delta >0\) such that \( |E_{p,q}(u^{\tau })-E_{2^*,q}(u^{\tau })|<\epsilon \) for \(2^*-\delta \le p\le 2^*\) and \(0\le \tau \le \tau _0\), which implies that \(E_{p,q}(u^{\tau _0})\le -1\) for all \(2^*-\delta \le p\le 2^*\). In view of \(E_{p,q}(u^{\tau })>0\) for \(\tau \) small enough for every \(p\in [q,2^*]\), it follows from Lemma 2.3 that the unique critical (maximum) point \(\tau _{p,q}\) of \(E_{p,q}(u^{\tau })\) belongs to \((0,\tau _0)\) and \(P_{p,q}(u^{\tau _{p,q}})=0\). Since \(u\in {\mathcal {P}}_{2^*,q}\), we deduce that \(E_{2^*,q}(u)=\max _{\tau >0}E_{2^*,q}(u^{\tau })\). Consequently,

$$\begin{aligned} c_{p,q}\le E_{p,q}(u^{\tau _{p,q}})\le E_{2^*,q}(u^{\tau _{p,q}})+\epsilon \le E_{2^*,q}(u)+\epsilon <c_{2^*,q}+2\epsilon \end{aligned}$$

for any \(2^*-\delta \le p\le 2^*\). Thus, \(\limsup _{p\rightarrow 2^*}c_{p,q}\le c_{2^*,q}\). \(\square \)

Lemma 3.2

Let \(N\ge 3\), \(a>0\), \(\mu >0\) and \({\bar{p}}\le q<p<2^*\). If \(q={\bar{p}}\), we further assume that \(\mu a^{\frac{4}{N}}<({\bar{a}}_N)^{\frac{4}{N}}\). Then \(\liminf _{p\rightarrow 2^*}c_{p,q}>0\).

Proof

By Lemma 2.5, there exists a sequence \(\{u_{p,q}\}_p\subset {\mathcal {P}}_{p,q}\) such that \(E_{p,q}(u_{p,q})=c_{p,q}\). By the Young inequality (3.1), we have

$$\begin{aligned} \begin{aligned} \int _{{\mathbb {R}}^N}|\nabla u_{p,q}|^2dx&=\gamma _p\int _{{\mathbb {R}}^N}|u_{p,q}|^{p}dx +\mu \gamma _q\int _{{\mathbb {R}}^N}|u_{p,q}|^{q}dx\\&\le \left( \gamma _p\frac{2^*-p}{2^*-q}+\mu \gamma _q\right) \int _{{\mathbb {R}}^N}|u_{p,q}|^{q}dx+\gamma _p \frac{p-q}{2^*-q}\int _{{\mathbb {R}}^N}|u_{p,q}|^{2^*}dx. \end{aligned} \end{aligned}$$

Letting \(p\rightarrow 2^*\), similarly to the proof of Lemma 2.4, we can show that there exists \(C>0\) independent of p such that \(\Vert \nabla u_{p,q}\Vert _2^2>C\), subsequently, \(\liminf _{p\rightarrow 2^*}c_{p,q}>0\). \(\square \)

Lemma 3.3

Let \(N\ge 3\), \(a>0\), \(\mu >0\) and \({\bar{p}}\le q<2^*\). If \(q={\bar{p}}\), we further assume that \(\mu a^{\frac{4}{N}}<({\bar{a}}_N)^{\frac{4}{N}}\). Then \(c_{2^*,q}<\frac{1}{N}S^{\frac{N}{2}}\), where S is defined by

$$\begin{aligned} S:=\inf _{ u\in D^{1,2}({\mathbb {R}}^N)\setminus \{0\}}\frac{\int _{{\mathbb {R}}^N}|\nabla u|^2dx}{\left( \int _{{\mathbb {R}}^N}|u|^{2^*}dx\right) ^{\frac{N-2}{N}}}. \end{aligned}$$
(3.2)

Proof

For any \(\epsilon >0\), we define

$$\begin{aligned} u_\epsilon (x)=\varphi (x)U_\epsilon (x), \end{aligned}$$

where

$$\begin{aligned} U_\epsilon (x)=\frac{\left( N(N-2)\epsilon ^2\right) ^{\frac{N-2}{4}}}{\left( \epsilon ^2+|x|^2\right) ^{\frac{N-2}{2}}} \end{aligned}$$

is the ground state of equation

$$\begin{aligned} -\Delta u=|u|^{2^*-2}u,\ x\in {\mathbb {R}}^N, \end{aligned}$$

and \(\varphi (x) \in C_c^{\infty }({\mathbb {R}}^N)\) is a cut off function satisfying:

  1. (a)

    \(0\le \varphi (x)\le 1\) for any \(x\in {\mathbb {R}}^N\);

  2. (b)

    \(\varphi (x)\equiv 1\) in \(B_1\), where \(B_s\) denotes the ball in \({\mathbb {R}}^N\) of center at origin and radius s;

  3. (c)

    \(\varphi (x)\equiv 0\) in \({\mathbb {R}}^N\setminus \overline{B_2}\).

By [5] (see also [25]), we have the following estimates.

$$\begin{aligned}&\int _{{\mathbb {R}}^N}|\nabla u_\epsilon |^2dx=S^{\frac{N}{2}}+O(\epsilon ^{N-2}),\ N\ge 3, \\&\quad \int _{{\mathbb {R}}^N}| u_\epsilon |^{2^*}dx=S^{\frac{N}{2}}+O(\epsilon ^N),\ N\ge 3, \end{aligned}$$

and

$$\begin{aligned} \int _{{\mathbb {R}}^N}| u_\epsilon |^2dx=\left\{ \begin{array}{ll} K_2\epsilon ^2+O(\epsilon ^{N-2}),&{} N\ge 5,\\ K_2\epsilon ^2|\ln \epsilon |+O(\epsilon ^2),&{} N=4,\\ K_2\epsilon +O(\epsilon ^2),&{} N=3, \end{array}\right. \end{aligned}$$

where \(K_2>0\). By direct calculations, for \(t\in (2,2^*)\), there exists \(K_1>0\) such that

$$\begin{aligned} \begin{aligned} \int _{{\mathbb {R}}^N}|u_\epsilon |^tdx&\ge (N(N-2))^{\frac{N-2}{4}t}\epsilon ^{N-\frac{N-2}{2}t}\int _{B_{\frac{1}{\epsilon }}(0)}\frac{1}{(1+|x|^2)^{\frac{N-2}{2}t}}dx\\&\ge \left\{ \begin{array}{ll} K_1\epsilon ^{N-\frac{N-2}{2}t},&{} (N-2)t>N,\\ K_1\epsilon ^{N-\frac{N-2}{2}t}|\ln \epsilon |,&{} (N-2)t=N,\\ K_1\epsilon ^{\frac{N-2}{2}t},&{} (N-2)t<N. \end{array}\right. \end{aligned} \end{aligned}$$

Define \(v_\epsilon (x)=(a^{-1}\Vert u_\epsilon \Vert _2)^{\frac{N-2}{2}}u_\epsilon (a^{-1}\Vert u_\epsilon \Vert _2x) \). Then

$$\begin{aligned} \int _{{\mathbb {R}}^N}|v_\epsilon |^2dx=a^2,\ \int _{{\mathbb {R}}^N}|\nabla v_\epsilon |^2dx= \int _{{\mathbb {R}}^N}|\nabla u_\epsilon |^2dx,\ \int _{{\mathbb {R}}^N}|v_\epsilon |^{2^*}dx=\int _{{\mathbb {R}}^N}|u_\epsilon |^{2^*}dx, \end{aligned}$$

and for \(q\in [{\bar{p}},2^*)\),

$$\begin{aligned} \begin{aligned} \int _{{\mathbb {R}}^N}|v_\epsilon |^{q}dx&=(a^{-1}\Vert u_\epsilon \Vert _2)^{\frac{N-2}{2}q-N}\int _{{\mathbb {R}}^N}|u_\epsilon |^{q}dx\\&\ge a^{N-\frac{N-2}{2}q}\Vert u_\epsilon \Vert _2^{\frac{N-2}{2}q-N}K_1\epsilon ^{N-\frac{N-2}{2}q}\\&\ge \frac{1}{2}a^{N-\frac{N-2}{2}q}K_1K_2^{\frac{N-2}{4}q-\frac{N}{2}}\times \left\{ \begin{array}{ll} 1,&{} N\ge 5,\\ |\ln \epsilon |^{\frac{N-2}{4}q-\frac{N}{2}},&{} N=4,\\ \epsilon ^{\frac{N}{2}-\frac{N-2}{4}q},&{} N=3. \end{array}\right. \end{aligned} \end{aligned}$$

Next we use \(v_\epsilon \) to estimate \(c_{2^*,q}\). By Lemma 2.3, there exists a unique \(\tau _\epsilon \) such that \(P_{2^*,q}((v_\epsilon )^{\tau _\epsilon })=0\) and \(E_{2^*,q}((v_\epsilon )^{\tau _\epsilon })=\sup _{\tau \ge 0}E_{2^*,q}((v_\epsilon )^{\tau })\). Thus, \(c_{2^*,q}\le \sup _{\tau \ge 0}E_{2^*,q}((v_\epsilon )^{\tau })\). By direct calculations, one has

$$\begin{aligned} \begin{aligned}&E_{2^*,q}((v_\epsilon )^{\tau })\\&\quad =\frac{1}{2}\tau ^{2}\int _{{\mathbb {R}}^N}|\nabla v_\epsilon |^2dx-\frac{1}{2^*}\tau ^{\frac{N}{2}2^*-N}\int _{{\mathbb {R}}^N}|v_\epsilon |^{2^*}dx -\frac{\mu }{q}\tau ^{\frac{N}{2}q-N}\int _{{\mathbb {R}}^N}|v_\epsilon |^qdx\\&\quad \le \frac{1}{2}\tau ^{2}\left( S^{\frac{N}{2}}+O(\epsilon ^{N-2})\right) -\frac{1}{2^*}\tau ^{2^*}\left( S^{\frac{N}{2}}+O(\epsilon ^N)\right) \\&\quad -\frac{\mu }{q}\tau ^{\frac{N}{2}q-N}\frac{1}{2}a^{N-\frac{N-2}{2}q}K_1K_2^{\frac{N-2}{4}q-\frac{N}{2}}\times \left\{ \begin{array}{ll} 1,&{} N\ge 5,\\ |\ln \epsilon |^{\frac{N-2}{4}q-\frac{N}{2}},&{} N=4,\\ \epsilon ^{\frac{N}{2}-\frac{N-2}{4}q},&{} N=3. \end{array}\right. \end{aligned} \end{aligned}$$
(3.3)

We claim that there exist \(\tau _0, \tau _1>0\) independent of \(\epsilon \) such that \(\tau _\epsilon \in [\tau _0, \tau _1]\) for \(\epsilon >0\) small. Suppose by contradiction that \(\tau _\epsilon \rightarrow 0\) or \(\tau _\epsilon \rightarrow \infty \) as \(\epsilon \rightarrow 0\). (3.3) implies that \(\sup _{\tau \ge 0}E_{2^*,q}((v_\epsilon )^{\tau })\le 0\) as \(\epsilon \rightarrow 0\) and then \(c_{2^*,q}\le 0\), which contradicts \(c_{2^*,q}>0\). Thus, the claim holds.

In (3.3), \(O(\epsilon ^{N-2})\) can be controlled by the last term for \(\epsilon >0\) small enough. Hence,

$$\begin{aligned} \sup _{\tau \ge 0}E_{2^*,q}((v_\epsilon )^{\tau })<\sup _{\tau \ge 0}\left( \frac{1}{2}\tau ^{2}S^{\frac{N}{2}}-\frac{1}{2^*}\tau ^{2^*}S^{\frac{N}{2}}\right) \le \frac{1}{N}S^{\frac{N}{2}}. \end{aligned}$$

The proof is complete. \(\square \)

Lemma 3.4

Let \(N\ge 3\), \(a>0\), \(\mu >0\) and \({\bar{p}}\le q<2^*\). If \(q={\bar{p}}\), we further assume that \(\mu a^{\frac{4}{N}}<({\bar{a}}_N)^{\frac{4}{N}}\). Then \(c_{2^*,q}\) is attained by a real-valued positive, radially symmetric and radially non-increasing function.

Proof

Let \(p_n\rightarrow 2^{*-}\) as \(n\rightarrow \infty \), by Lemmas 2.5 and 3.1, there exists a sequence of positive and radially non-increasing functions \(\{u_n:=u_{p_n,q}\}\subset {\mathcal {P}}_{p_n,q}\) such that \(E_{p_n,q}(u_n)=c_{p_n,q}\le c_{2^*,q}+1\). If \(q>{\bar{p}}\), we have

$$\begin{aligned} c_{2^*,q}+1\ge E_{p_n,q}(u_n)=\left( \frac{1}{2}-\frac{1}{q\gamma _q}\right) \int _{{\mathbb {R}}^N}|\nabla u_n|^2dx+\left( \frac{\gamma _{p_n}}{q\gamma _q}-\frac{1}{p_n}\right) \int _{{\mathbb {R}}^N}| u_n|^{p_n}dx. \end{aligned}$$

So \(\{u_n\}\) is bounded in \(H^1({\mathbb {R}}^N)\). If \(q={\bar{p}}\), we have

$$\begin{aligned} c_{2^*,q}+1\ge E_{p_n,q}(u_n)=\left( \frac{\gamma _{p_n}}{2}-\frac{1}{p_n}\right) \int _{{\mathbb {R}}^N}| u_n|^{p_n}dx, \end{aligned}$$

which implies that \(\{\int _{{\mathbb {R}}^N}| u_n|^{p_n}dx\}\) is bounded. By the Young inequality

$$\begin{aligned} |u_n|^q\le \frac{p_n-q}{p_n-2}|u_n|^2+\frac{q-2}{p_n-2}|u_n|^{p_n}, \end{aligned}$$

we know that \(\{\int _{{\mathbb {R}}^N}| u_n|^qdx\}\) is bounded. So it follows from the expression

$$\begin{aligned} E_{p_n,q}(u_n)=\left( \frac{1}{2}-\frac{1}{p_n\gamma _{p_n}}\right) \int _{{\mathbb {R}}^N}|\nabla u_n|^2dx+\left( \frac{\gamma _q}{p_n\gamma _{p_n}}-\frac{1}{q}\right) \mu \int _{{\mathbb {R}}^N}| u_n|^qdx \end{aligned}$$

that \(\{u_n\}\) is bounded in \(H^1({\mathbb {R}}^N)\). Thus, there exists a nonnegative and radially non-increasing function \(u\in H^1({\mathbb {R}}^N)\) such that up to a subsequence, \(u_n\rightharpoonup u\) weakly in \(H^1({\mathbb {R}}^N)\), \(u_n\rightarrow u\) strongly in \(L^t({\mathbb {R}}^N)\) for \(t\in (2,2^*)\) and \(u_n\rightarrow u\) a.e. in \({\mathbb {R}}^N\).

By Lemma  2.6, there exists \(\lambda _n<0\) such that \(u_n\) satisfies

$$\begin{aligned} -\Delta u_n=\lambda _n u_n+|u_n|^{p_n-2}u_n+\mu |u_n|^{q-2}u_n,\ x\in {\mathbb {R}}^{N}. \end{aligned}$$
(3.4)

It follows from the expression

$$\begin{aligned} \lambda _na^2=(\gamma _{p_n}-1)\int _{{\mathbb {R}}^N}|u_n|^{p_n}dx+\mu (\gamma _q-1)\int _{{\mathbb {R}}^N}|u_n|^{q}dx \end{aligned}$$

that \(\{\lambda _n\}\) is bounded. So there exists \(\lambda \le 0\) such that up to a subsequence, \(\lim _{n\rightarrow \infty }\lambda _n=\lambda \).

It follows from \(N\ge 3\) that \(\frac{N}{\frac{N-2}{2}(2-1)}\) and \(\frac{N}{\frac{N-2}{2}(2^*-1)} \in (1,\infty )\). Since \(p_n\rightarrow 2^*\) and \(\psi \in L^{r}({\mathbb {R}}^N)\) for \(r\in (1,\infty )\), by the Young inequality, the Hölder inequality and Lemma 2.2 with \(t=2^*\), there exists a constant \(C>0\) independent of n such that

$$\begin{aligned} \begin{aligned} \left| |u_n|^{p_n-2}u_n\psi \right|&\le C\left( |u_n|^{2-1}|\psi |+|u_n|^{2^*-1}|\psi |\right) \\&\le C\left( |x|^{\frac{2-N}{2}(2-1)}|\psi |+|x|^{\frac{2-N}{2}(2^*-1)}|\psi |\right) \in L^{1}({\mathbb {R}}^N). \end{aligned} \end{aligned}$$
(3.5)

Passing to the limit in (3.4) and by using the Lebesgue dominated convergence theorem, we have for any \(\psi \in C_c^{\infty }({\mathbb {R}}^N)\),

$$\begin{aligned} \begin{aligned} 0&=\int _{{\mathbb {R}}^N}(\nabla u_n\nabla \psi -\lambda _nu_n\psi )dx-\int _{{\mathbb {R}}^N}|u_n|^{p_n-2}u_n\psi dx-\mu \int _{{\mathbb {R}}^N}|u_n|^{q-2}u_n\psi dx\\&\rightarrow \int _{{\mathbb {R}}^N}(\nabla u\nabla \psi -\lambda u\psi ) dx-\int _{{\mathbb {R}}^N}|u|^{2^*-2}u\psi dx-\mu \int _{{\mathbb {R}}^N}|u|^{q-2}u\psi dx \end{aligned} \end{aligned}$$

as \(n\rightarrow \infty \). That is, u is a solution of

$$\begin{aligned} -\Delta u=\lambda u+|u|^{2^*-2}u+\mu |u|^{q-2}u,\ x\in {\mathbb {R}}^{N}. \end{aligned}$$

Thus \(P_{2^*,q}(u)=0\).

We claim that \(u\not \equiv 0\). Suppose by contradiction that \(u\equiv 0\). By using \(P_{p_n,q}(u_n)=0\), \(\int _{{\mathbb {R}}^N}|u_n|^{q}=o_n(1)\) and the Young inequality

$$\begin{aligned} |u_n|^{p_n}\le \frac{2^*-p_n}{2^*-q}|u_n|^{q}+ \frac{p_n-q}{2^*-q}|u_n|^{2^*}, \end{aligned}$$

we get that

$$\begin{aligned} \begin{aligned} \int _{{\mathbb {R}}^N}|\nabla u_n|^2dx&=\gamma _{p_n}\int _{{\mathbb {R}}^N}|u_n|^{p_n}dx+o_n(1)\\&\le \gamma _{p_n}\frac{p_n-q}{2^*-q}\int _{{\mathbb {R}}^N}|u_n|^{2^*}dx+o_n(1)\\&\le \gamma _{p_n}\frac{p_n-q}{2^*-q}\left( \frac{\int _{{\mathbb {R}}^N}|\nabla u_n|^2dx}{S}\right) ^{\frac{N}{N-2}}+o_n(1). \end{aligned} \end{aligned}$$

Since \(\liminf _{n\rightarrow \infty }\int _{{\mathbb {R}}^N}|\nabla u_n|^2>0\) (see the proof of Lemma 3.2), we obtain

$$\begin{aligned} \limsup _{n\rightarrow \infty }\Vert \nabla u_n\Vert _2^2\ge S^{\frac{N}{2}}. \end{aligned}$$

Consequently,

$$\begin{aligned} \begin{aligned} c_{2^*,q}&\ge \limsup _{n\rightarrow \infty }c_{p_n,q}\\&=\limsup _{n\rightarrow \infty }\left\{ \left( \frac{1}{2}-\frac{1}{p_n\gamma _{p_n}}\right) \int _{{\mathbb {R}}^N}|\nabla u_n|^{2}dx +\left( \frac{\gamma _q}{p_n\gamma _{p_n}}-\frac{1}{q}\right) \mu \int _{{\mathbb {R}}^N}|u_n|^qdx\right\} \\&=\limsup _{n\rightarrow \infty } \left\{ \left( \frac{1}{2}-\frac{1}{p_n\gamma _{p_n}}\right) \int _{{\mathbb {R}}^N}|\nabla u_n|^2dx\right\} \\&\ge \frac{1}{N}S^{\frac{N}{2}}, \end{aligned} \end{aligned}$$

which contradicts Lemma 3.3. Thus \(u\not \equiv 0\).

Set \(\int _{{\mathbb {R}}^N}|u|^2dx=c^2\le a^2\). Similarly to the proof of (2.7), we define \({\tilde{u}}\in S_a\). Then there exists \(\tau _0\in (0,1]\) such that \(P_{2^*,q}({\tilde{u}}^{\tau _0})=0\) and by Fatou’s lemma,

$$\begin{aligned} \begin{aligned} c_{2^*,q}&\le E_{2^*,q}({\tilde{u}}^{\tau _0})\\&=\left( \frac{1}{2}-\frac{1}{q\gamma _q}\right) \int _{{\mathbb {R}}^N}|\nabla ({\tilde{u}}^{\tau _0})|^2dx+\left( \frac{\gamma _{2^*}}{q\gamma _q}-\frac{1}{2^*}\right) \int _{{\mathbb {R}}^N}| ({\tilde{u}}^{\tau _0})|^{2^*}dx\\&\le \left( \frac{1}{2}-\frac{1}{q\gamma _q}\right) \int _{{\mathbb {R}}^N}|\nabla u|^2dx+\left( \frac{\gamma _{2^*}}{q\gamma _q}-\frac{1}{{2^*}}\right) \int _{{\mathbb {R}}^N}| u|^{2^*}dx\\&\le \liminf _{n\rightarrow \infty }\left\{ \left( \frac{1}{2}-\frac{1}{q\gamma _q}\right) \int _{{\mathbb {R}}^N}|\nabla u_n|^2dx+\left( \frac{\gamma _{p_n}}{q\gamma _q}-\frac{1}{{p_n}}\right) \int _{{\mathbb {R}}^N}| u_n|^{p_n}dx \right\} \\&= \liminf _{n\rightarrow \infty } c_{p_n,q}\le \limsup _{n\rightarrow \infty } c_{p_n,q}\le c_{2^*,q}. \end{aligned} \end{aligned}$$

Hence, \(E_{2^*,q}({\tilde{u}}^{\tau _0})=c_{2^*,q}\). That is \({\tilde{u}}^{\tau _0}\) is a real-valued nonnegative, radially symmetric and radially non-increasing minimizer of \(c_{2^*,q}\). By the strong maximum principle, \({\tilde{u}}^{\tau _0}>0\) in \({\mathbb {R}}^N\). \(\square \)

Proof of Theorem 1.4: It follows from Lemmas 2.42.63.3 and  3.4.