We consider the following standard Sobolev space and spatially weighted Lebesgue space: \(\mathbf{W}^{q}_{p}:=\{u:\|u\|_{p}^{q}:=\sum_{|\beta|\leq p}\|D^{\beta}u\|_{\mathbf{L}^{q}}^{q}<\infty\}\), \(\mathbf{L}_{m}^{n}:=\{u:\| u\|_{m}^{n}:=\int_{\mathbf{R}^{3}}\rho^{m}(x)u^{n}(x)dx<\infty\}\) with the weighted function \(\rho(x)=\sqrt{1+|x|^{2}}\). We denote \(\mathbf {W}^{2}_{p}\) by \(\mathbf{H}^{p}\) with the norm \(\|u\|_{\mathbf{H}^{p}}=\sum_{|\beta|\leq p}\|D^{\beta}u\|_{\mathbf{L}^{2}}^{2}\). When \(q=2\), the Fourier transform is an isomorphism between \(\mathbf{H}^{p}\) and \(\mathbf {L}_{p}^{2}\) with \(\|u\|_{\mathbf{L}_{p}^{2}}=\|\rho^{p}u\|_{\mathbf{L}^{2}}\). We also introduce the space \(\mathbf{X}:=\{u=(u_{n})_{n\in\mathbf{Z}}:\|u\| _{\mathbf{X}}<\infty\}\) and the weighted spaces \(\mathcal{L}_{s}^{p}=\mathbf{L}_{s}^{p}\times\mathbf {L}_{s}^{p}\), \(\mathcal{H}^{m}=\mathbf{H}^{m}\times\mathbf{H}^{m}\), \(\mathcal {X}=\mathbf{X}\times\mathbf{X}\), which are equipped with the norms \(\|u\|_{\mathbf{X}}=\sum_{n\in\mathbf{Z}}\|u_{n}\|_{\mathbf{H}^{m}}\), \(\| \varphi\|_{\mathcal{X}}:=\|u\|_{\mathbf{X}}+\|v\|_{\mathbf{X}}\), \(\| \varphi\|_{\mathcal{L}_{m}^{p}}:=\|u\|_{\mathbf{L}_{m}^{p}}+\|v\|_{\mathbf {L}_{m}^{p}}\), \(\|\varphi\|_{\mathcal{H}^{m}}:=\|u\|_{\mathbf{H}^{m}}+\|v\| _{\mathbf{H}^{m}}\) for \(\forall\varphi=(u,v)^{T}\in\mathcal{L}_{s}^{p}\) or \(\mathcal{X}\).
Next, we look for 2π time-periodic solutions of
$$\begin{aligned} \Xi\frac{d\varphi}{dt}+\mathcal{N}\varphi+G(\varphi)=F(\varphi), \end{aligned}$$
(2.1)
where
$$\begin{aligned} \Xi= \begin{pmatrix} \xi_{1}& 0 \\ 0 &\xi_{2} \end{pmatrix},\qquad \mathcal{N}= \begin{pmatrix} \nu_{0}\triangle+c_{1}\partial_{x_{1}}+\nu_{0}'c_{1}\partial_{x_{1}x_{1}}& -\partial _{x_{3}} \\ 0 &\kappa_{0}\triangle+c_{1}\partial_{x_{1}}+\kappa_{0}'c_{1}\partial_{x_{1}x_{1}} \end{pmatrix}, \end{aligned}$$
and
$$\begin{aligned} G(\varphi)= \begin{pmatrix} g^{1}\\ g^{2} \end{pmatrix} ,\qquad F(\varphi)= \begin{pmatrix} g^{3}\\ g^{4} \end{pmatrix} \end{aligned}$$
with
$$\begin{aligned}& \begin{aligned}[b] g^{1}={}&\nabla\cdot \bigl(\omega_{\alpha}u^{T}-u_{\alpha} \omega^{T}-\omega u_{\alpha }^{T}+u\omega_{\alpha}^{T}- \nu_{0}'(\nabla\times T_{\alpha})\nabla u^{T}+\nu _{0}'T_{\alpha}\nabla \omega^{T} \\ &{}-\nu_{0}'(\nabla\times v)\nabla u_{\alpha}^{T}+ \nu_{0}'v\nabla\omega_{\alpha }^{T} \bigr), \end{aligned} \\& g^{2}=T_{\alpha}\cdot\nabla v+u\cdot\nabla T_{\alpha}- \kappa_{0}'\nabla \cdot(v\nabla T_{\alpha}-T_{\alpha} \nabla v), \\& g^{3}=-\nabla\cdot \bigl(\omega u^{T}-u \omega^{T}-\nu_{0}'(\nabla\times v)\nabla u^{T}+v\nabla\omega^{T} \bigr), \end{aligned}$$
(2.2)
$$\begin{aligned}& g^{4}=-u\cdot\nabla v-\kappa_{0}' \nabla \cdot(v\nabla v). \end{aligned}$$
(2.3)
According to the classical result in [8], we know that the essential spectrum of the operator \(\mathcal{N}+G\) is a relatively compact perturbation of \(\mathcal{N}\). It has the essential spectrum
$$\begin{aligned}[b] &\operatorname{essspec}(\widehat{\mathcal{N}})\\ &\quad= \bigl\{ \lambda\in \mathcal{C}^{2}: \lambda= \bigl(-\nu _{0}|y|^{2}- \nu_{0}'c_{1}|y_{1}|^{2}+icy_{1},- \kappa_{0}|y|^{2}-\kappa _{0}'c_{1}|y_{1}|^{2}+icy_{1} \bigr), y\in\mathbf{R}^{3} \bigr\} . \end{aligned} $$
Moreover, the spectra of \(\mathcal{N}+G\) and \(\mathcal{N}\) only differ by isolated eigenvalues of finite multiplicity. The above spectrum properties are critical to prove our main result.
For convenience, we rewrite (2.1) as
$$\begin{aligned}& \xi_{1}\omega_{t}=M_{1} \omega+g^{3}( \omega,u,v), \end{aligned}$$
(2.4)
$$\begin{aligned}& \xi_{2}v_{t}=M_{2}v+g^{4}( \omega,u,v), \end{aligned}$$
(2.5)
where \(g_{3}\) and \(g_{4}\) defined in (2.2)-(2.3),
$$\begin{aligned}& M_{1}\omega=\overline{M_{1}} \omega+g^{1}= \nu_{0}\triangle\omega+\nu _{0}'c_{1} \partial_{x_{1}x_{1}}\omega+c_{1}\partial_{x_{1}}\omega- \partial _{x_{3}}v+g^{1}, \end{aligned}$$
(2.6)
$$\begin{aligned}& M_{2}v=\overline{M_{2}}v+g^{2}= \kappa_{0}\triangle v+\kappa_{0}'c_{1} \partial _{x_{1}x_{1}}v+c_{1}\partial_{x_{1}}v+g^{2}. \end{aligned}$$
(2.7)
We make the ansatz \(\omega(x,t)=\sum_{n\in\mathbf{Z}}\omega_{n}(x)e^{int}\) and \(v(x,t)=\sum_{n\in\mathbf{Z}}v_{n}(x)e^{int}\) to (2.4)-(2.5). Then we obtain
$$\begin{aligned}& (in \xi_{1}-M_{1})\omega_{n}=g^{3}_{n}( \omega,u,v), \end{aligned}$$
(2.8)
$$\begin{aligned}& (in \xi_{2}-M_{2})v_{n}=g^{4}_{n}( \omega,u,v), \end{aligned}$$
(2.9)
where \(g^{3}(\omega,u,v)(x,t)=\sum_{n\in\mathbf{Z}} g^{3}_{n}(\omega,u,v)e^{int}\), \(g^{4}(\omega,u,v)(x,t)=\sum_{n\in\mathbf{Z}}g^{4}_{n}(\omega,u,v)e^{int}\).
Note that we are interested in a real valued solution only. So we will always suppose that \((\omega_{n},v_{n})=(\omega_{-n},v_{-n})\) for \(n\in \mathbf{Z}\). These series are uniformly convergent on \(\mathbf{R}^{3}\times[0,2\pi]\) in the spaces which we have chosen. More precisely, we have the following three results, which are taken from [7, 9].
Lemma 2.1
A linear operator
\(\mathbf{J}:\mathcal{X}\longrightarrow\mathbf {C}_{b}^{0}(\mathbf{R}^{3}\times[0,\pi],\mathbf{C}^{2})\)
is defined by
$$\begin{aligned} (\mathbf{J}u) (x,t)=\tilde{u}(x,t):=\sum_{n\in\mathbf {Z}}u_{n}(x)e^{int}, \quad u=(u_{n})_{n\in\mathbf{Z}}\in\mathcal{X}. \end{aligned}$$
Then
J
is bounded. Here
\(\mathbf{C}^{2}\)
denotes the twice continuous differentiable function space.
Lemma 2.2
For
\(u=(u_{n})_{n\in\mathbf{Z}}\), \(v=(v_{n})_{n\in\mathbf{Z}}\in\mathbf{X}\), the convolution
\(u*v\in\mathbf{X}\)
is defined by
$$\begin{aligned} (u*v)_{n}=\sum_{k\in\mathbf{Z}}u_{n-k}v_{k}, \quad n\in\mathbf{Z}. \end{aligned}$$
Then there exists
\(C>0\)
such that
\(\|u*v\|_{\mathcal{X}}\leq C\|u\|_{\mathcal{X}}\|v\|_{\mathcal{X}}\).
Lemma 2.3
Let a linear operator
\(M:\mathbf{X}\longrightarrow\mathbf{X}\)
be defined component-wise as
\((Mu)_{n}=M_{n}u_{n}\)
for
\(u=(u_{n})_{n\in\mathbf {Z}}\). Then
\(\|Mu\|_{\mathbf{X}}=(\|M_{0}\|_{\mathbf{H}^{m}\longrightarrow \mathbf{H}^{m}}+\sup_{n\in\mathbf{Z}\backslash\{0\}} \|M\|_{\mathbf{H}^{m}\longrightarrow\mathbf{H}^{m}})\|u\|_{\mathbf{X}}\).
By (H2) and (H3), we know that the operators \(M_{1}\) and \(M_{2}\) have two eigenvalues \(\lambda_{0}^{\pm}(\alpha)\) and \(\mu_{0}^{\pm }(\alpha)\), respectively, and all other eigenvalues of \(M_{1}\) and \(M_{2}\) are strictly bounded away from the imaginary axis in the left half plane. Thus we construct the \(M_{j}\)-invariant (\(j=1,2\)) projections \(\mathbf{P}_{\pm1,c}\) by
$$\begin{aligned}& P_{1,c}\omega= \bigl(\psi^{+,*},\omega \bigr)_{\mathbf{L}^{2}} \psi ^{+},\qquad P_{-1,c}\omega= \bigl(\psi^{-,*}, \omega \bigr)_{\mathbf{L}^{2}}\psi^{-}, \end{aligned}$$
(2.10)
$$\begin{aligned}& P_{1,c}v= \bigl(\psi^{+,*},v \bigr)_{\mathbf{L}^{2}}\psi^{+}, \qquad P_{-1,c}v= \bigl(\psi ^{-,*},v \bigr)_{\mathbf{L}^{2}} \psi^{-}, \end{aligned}$$
(2.11)
where \(\psi^{\pm}\) denotes the associated normalized eigenfunctions, \(\psi^{\pm1,*}\) denotes the associated normalized eigenfunctions of the adjoint operator \(M_{j}^{*}\). The bounded ‘stable’ part of the projection is \(\mathbf{P}_{\pm1,s}=I-\mathbf{P}_{\pm1,c}\), and we also know that \(\mathbf{P}_{\pm,c}M_{j}=M_{j}\mathbf{P}_{\pm,c}\) and \(\mathbf{P}_{\pm ,s}M_{j}=M_{j}\mathbf{P}_{\pm,s}\). Thus we can split \(\omega_{\pm1}\) and \(v_{\pm1}\) as \(\omega_{1}=\omega_{1,c}+\omega_{1,s}\), \(\omega _{-1}=\omega_{-1,c}+\omega_{-1,s}\), \(v_{1}=v_{1,c}+v_{1,s}\), and \(v_{-1}=v_{-1,c}+v_{-1,s}\), where \(\omega_{\pm1,c}=\mathbf{P}_{\pm1,c}\omega_{1}\), \(\omega_{\pm1,s}=\mathbf {P}_{\pm1,s}\omega_{1}\), \(v_{\pm1,c}=\mathbf{P}_{\pm1,c}v_{1}\), and \(v_{\pm1,s}=\mathbf{P}_{\pm1,s}v_{1}\). Applying the above decompositions to (2.8)-(2.9), we have
$$\begin{aligned}& (in \xi_{1}-M_{1})\omega_{n}=g^{3}_{n}( \omega,u,v),\quad n=\pm2,\pm3,\ldots, \end{aligned}$$
(2.12)
$$\begin{aligned}& (in \xi_{2}-M_{2})v_{n}=g^{4}_{n}( \omega,u,v),\quad n=\pm2,\pm3,\ldots, \end{aligned}$$
(2.13)
$$\begin{aligned}& M_{1}\omega_{0}=g^{3}_{0}( \omega,u,v), \quad n=0, \end{aligned}$$
(2.14)
$$\begin{aligned}& M_{2}v_{0}=g^{4}_{0}( \omega,u,v),\quad n=0, \end{aligned}$$
(2.15)
$$\begin{aligned}& (\pm i\xi_{1}-M_{1})\omega_{\pm1,s}= \mathbf{P}_{\pm1,s}g^{3}_{\pm1}(\omega ,u,v), \end{aligned}$$
(2.16)
$$\begin{aligned}& (\pm i\xi_{2}-M_{2})v_{\pm1,s}= \mathbf{P}_{\pm1,s}g^{4}_{\pm1}( \omega ,u,v), \end{aligned}$$
(2.17)
$$\begin{aligned}& (\pm i\xi_{1}-M_{1})\omega_{\pm1,c}= \mathbf{P}_{\pm1,c}g^{3}_{\pm1}(\omega ,u,v), \end{aligned}$$
(2.18)
$$\begin{aligned}& (\pm i\xi_{2}-M_{2})v_{\pm1,c}= \mathbf{P}_{\pm1,c}g^{4}_{\pm1}( \omega,u,v). \end{aligned}$$
(2.19)
The organization of the proof of Theorem 1.1 is as follows: we first solve (2.14)-(2.15), then we use the fixed point theorem to solve (2.12)-(2.13) and (2.16)-(2.17). This is a nontrivial process due to the nonlinear terms \(g^{3}_{n}(\omega,u,v)\) and \(g^{4}_{n}(\omega,u,v)\). Finally, we employ the implicit function theorem to solve (2.18)-(2.19). The process of solving (2.18)-(2.19) is inspired by the classical Hopf-Bifurcation result [10].
Equations (2.12)-(2.19) can be rewritten as
$$\begin{aligned}& (in \Xi+\mathcal{N}+G)\varphi_{n}=F_{n}( \varphi,u), \quad n=\pm2,\pm3,\ldots, \end{aligned}$$
(2.20)
$$\begin{aligned}& (\mathcal{N}+G)\varphi_{0}=F_{0}( \varphi,u),\quad n=0, \end{aligned}$$
(2.21)
$$\begin{aligned}& (\pm i\Xi+\mathcal{N}+G)\varphi_{\pm1,s}= \mathbf{P}_{\pm1,s}F_{\pm 1}( \varphi,u), \end{aligned}$$
(2.22)
$$\begin{aligned}& (\pm i\Xi+\mathcal{N}+G)\varphi_{\pm1,c}= \mathbf{P}_{\pm1,c}F_{\pm 1}( \varphi,u). \end{aligned}$$
(2.23)
Next we solve (2.21). The linear operator \(\mathcal{N}\) has an essential spectrum up to the imaginary axis, and it is an invertible operator in the following sense. One may refer to [9] for the details of the proof.
Lemma 2.4
For
\(j=1,2\)
and
\(f=(f^{1},f^{2})^{T}\in\mathcal{H}^{m-1}\cap\mathcal{L}^{1}\), the equation
\(\mathcal{N}\varphi=\partial_{j}f\)
has a unique solution
\(\varphi=\mathcal{N}^{-1}\partial_{j}f\in\mathcal {H}^{m}\). Moreover, \(\|\varphi\|_{\mathcal{H}^{m}}\leq C\|f\|_{\mathcal{H}^{m-1}\cap\mathcal{L}^{1}}\).
This lemma tells us that \(\widehat{\mathcal{N}}(iy_{i},iy_{i})^{T}\) is a bounded compact operator from \(\mathcal{L}^{2}_{m}\) to itself. Furthermore, the spectra of \(\widehat{\mathcal{N}}+\widehat{G}\) and \(\widehat{\mathcal{N}}\) only differ by isolated eigenvalues of finite multiplicity (see the book of Henry [8, p.136]). The following three lemmas give the solvability of (2.21); they are similar to Lemma 6-8 in [9]. Here we omit their proofs.
Lemma 2.5
Assume that (H1)-(H3) hold. Then (2.21) has a unique solution
\(\varphi_{0}=(\mathcal{N}+G)^{-1}F_{0}(\varphi,u)\). Moreover, \(\|\varphi_{0}\|_{\mathcal{H}^{m}}\leq C\|y_{j}^{-1}I_{2\times2}\widehat{F_{0}} (\hat{\varphi},\hat{u})\|_{\mathcal{L}_{m}^{2}}\), where
\(I_{2\times 2}\)
and
\(\widehat{F_{0}}\)
denote the
\(2\times2\)
unit matrix and the application of Fourier transform to
\(F_{0}\), respectively.
Lemma 2.6
There exists a constant
\(C>0\)
such that
$$\begin{aligned} \|u\|_{\mathbf{H}^{m}}\leq C\|\omega\|_{\mathbf{H}^{m}}, \qquad\|\partial _{x_{i}}u \|_{\mathbf{H}^{m}}\leq C\|\omega\|_{\mathbf{H}^{m}}. \end{aligned}$$
Lemma 2.7
For
\(m>\frac{3}{2}\), there exists a positive constant
C
such that
$$\begin{aligned} \|\hat{\omega}*\hat{u}\|_{\mathbf{L}_{m}^{2}}\leq C\|\hat{\omega}\|_{\mathbf {L}_{m}^{2}}\| \hat{u}\|_{\mathbf{L}_{m}^{2}}. \end{aligned}$$
Applying the Fourier transform to \(g^{3}\) and \(g^{4}\) in (2.2)-(2.3), we get
$$\begin{aligned}& \widehat{g^{3}}=-iy \bigl(\hat{\omega}* \hat{u}^{T}-\hat{u}* \hat{\omega }^{T}+\nu_{0}'|y|^{2} \hat{v}*\hat{u}^{T}+iy\hat{v}*\hat{\omega} \bigr), \end{aligned}$$
(2.24)
$$\begin{aligned}& \widehat{g^{4}}=-iy\hat{u}*\hat{v}-2 \kappa_{0}'|y|^{2} \hat{v}*\hat{v}. \end{aligned}$$
(2.25)
From the form of the nonlinear terms \(g^{3}\) and \(g^{4}\), it is critical to estimate the term as uv and \(u^{2}\). For convenience, we derive some estimates as regards the nonlinear terms \(N^{1}(\varphi)=\varphi^{2}\) and \(N^{2}(\varphi,\psi)=\varphi\psi\). Since the proof of the next lemma is similar to Lemma 4 in [7], we omit it.
Lemma 2.8
Define
\(N^{1}:\mathcal{X}\longrightarrow\mathcal{X}\)
by
\(N^{1}(\varphi )_{n}=N^{1}_{n}(\mathbf{J}\varphi)\)
and
\(N^{2}:\mathcal{X}\times\mathcal {X}\longrightarrow\mathcal{X}\)
by
\(N^{2}(\varphi)_{n}=N^{2}_{n}(\mathbf {J}\varphi,\mathbf{J}\psi)\)
for
\(\varphi,\psi\in\mathcal{X}\). Then we have
$$\begin{aligned} \bigl\| N^{1}(\varphi)\bigr\| _{\mathcal{X}}\leq C\|\varphi\|^{2}_{\mathcal{X}}, \qquad \bigl\| N^{2}(\varphi,\psi)\bigr\| _{\mathcal{X}}\leq C\|\psi \|_{\mathcal{X}}\| \varphi\| _{\mathcal{X}} \end{aligned}$$
for
\(\varphi,\psi\in\mathcal{X}\)
with
\(\|\varphi\|_{\mathcal{X}}\leq1\)
and
\(\|\psi\|_{\mathcal{X}}\leq1\). Moreover, we have
$$\begin{aligned}& \bigl\| N^{1} \bigl(\varphi^{1} \bigr)-N^{1} \bigl( \varphi^{2} \bigr)\bigr\| _{\mathcal{X}}\leq C \bigl(\bigl\| \varphi^{1} \bigr\| _{\mathcal{X}}+\bigl\| \varphi^{2}\bigr\| _{\mathcal{X}} \bigr)\bigl\| \varphi^{1}-\varphi^{2}\bigr\| _{\mathcal{X}}, \\& \begin{aligned}[b] \bigl\| N^{2} \bigl(\varphi^{1}, \psi^{1} \bigr)-N^{2} \bigl(\varphi^{2}, \psi^{2} \bigr) \bigr\| _{\mathcal{X}}\leq{}& C \bigl(\bigl\| \varphi^{1} \bigr\| _{\mathcal{X}}+\bigl\| \varphi^{2}\bigr\| _{\mathcal{X}}+\bigl\| \psi ^{1} \bigr\| _{\mathcal{X}}+\bigl\| \psi^{2}\bigr\| _{\mathcal{X}} \bigr) \\ &{}\times \bigl(\bigl\| \varphi^{1}-\varphi^{2}\bigr\| _{\mathcal{X}}+ \bigl\| \psi^{1}-\psi^{2}\bigr\| _{\mathcal{X}} \bigr), \end{aligned} \end{aligned}$$
for
\(\varphi^{1},\varphi^{2},\psi^{1},\psi^{2}\in\mathcal{X}\)
with
\(\|\varphi ^{1}\|_{\mathcal{X}},\|\varphi^{2}\|_{\mathcal{X}},\|\psi^{1}\|_{\mathcal {X}},\|\psi^{2}\|_{\mathcal{X}}\leq1\).
By a small modification of the proof of Lemma 7 in [7], we have the following result.
Lemma 2.9
Assume that
ξ
is close enough to
\(\xi_{0}\). Then there exists a constant
\(C>0\)
such that
$$\begin{aligned}& \bigl\| (in \xi_{i}-\overline{M_{i}})^{-1} \bigr\| _{\mathbf{H}^{m}\longrightarrow\mathbf {H}^{m}}\leq C, \qquad\bigl\| (in \xi_{i}-\overline{M_{i}})^{-1} \nabla^{j}\cdot\bigr\| _{\mathbf {H}^{m}\longrightarrow\mathbf{H}^{m}}\leq C, \\& \bigl\| (in \xi_{i}-M_{i})^{-1} \bigr\| _{\mathbf{H}^{m}\longrightarrow\mathbf{H}^{m}}\leq C,\qquad \bigl\| (in \xi_{i}-M_{i})^{-1} \nabla^{j} \cdot\bigr\| _{\mathbf{H}^{m}\longrightarrow \mathbf{H}^{m}}\leq C, \\& \bigl\| (in \xi_{i}-M_{i})^{-1} \mathbf{P}_{\pm1,s} \bigr\| _{\mathbf{H}^{m}\longrightarrow \mathbf{H}^{m}}\leq C,\qquad \bigl\| (in \xi_{i}-M_{i})^{-1} \nabla^{j}\cdot\mathbf{P}_{\pm 1,s}\bigr\| _{\mathbf{H}^{m}\longrightarrow\mathbf{H}^{m}}\leq C, \end{aligned}$$
for
\(n\neq0\)
and
\(j=1,2\). Here
\(\overline{M_{i}}\)
and
\(M_{i}\)
are defined in (2.6)-(2.7), respectively.
Thus by Lemma 2.9, we can rewrite (2.20) and (2.22) as
$$\begin{aligned}& \varphi_{n}=(in \Xi+\mathcal{N}+G)^{-1}F_{n}( \varphi,u),\quad n=\pm2,\pm 3,\ldots, \\& \varphi_{\pm1,s}=(\pm i\Xi+\mathcal{N}+G)^{-1} \mathbf{P}_{\pm 1,s}F_{\pm1}(\varphi,u), \end{aligned}$$
i.e.
$$\begin{aligned}& \omega_{n}=(in \xi_{1}-M_{1})^{-1}g^{3}_{n}( \omega,u,v), \quad n=\pm2,\pm3,\ldots, \end{aligned}$$
(2.26)
$$\begin{aligned}& v_{n}=(in \xi_{2}-M_{2})^{-1}g^{4}_{n}( \omega,u,v),\quad n=\pm2,\pm3,\ldots, \end{aligned}$$
(2.27)
$$\begin{aligned}& \omega_{\pm1,s}=(\pm i\xi_{1}-M_{1})^{-1} \mathbf{P}_{\pm1,s}g^{3}_{\pm 1}(\omega,u,v), \end{aligned}$$
(2.28)
$$\begin{aligned}& v_{\pm1,s}=(\pm i\xi_{2}-M_{2})^{-1} \mathbf{P}_{\pm1,s}g^{4}_{\pm1}(\omega,u,v). \end{aligned}$$
(2.29)
Using Lemmas 2.8 and 2.9, we obtain the solvability of (2.26)-(2.29). Since the proof is similar to [9], we omit it.
Lemma 2.10
Assume that there exist
\(\sigma_{1}, \sigma_{2}>0\)
such that for all
\(\xi _{1},\xi_{2}>0\)
with
\(|\xi_{1}-\xi_{0}|,|\xi_{2}-\xi_{0}|\leq\sigma_{1}\)
and all
\(\omega_{\pm1,c},v_{\pm1,c}\in\mathbf{H}^{m}\)
with
\(\|\omega_{\pm1,c}\| _{\mathbf{H}^{m}},\|v_{\pm1,c}\|_{\mathbf{H}^{m}}\leq\sigma_{2}\). Then (2.26)-(2.29) has a unique solution
\((\tilde{\omega},\tilde {v})=\Phi(\omega_{c},v_{c})\in\mathcal{X}\), where
\(\omega_{c}=(\omega_{-1,c},\omega_{1,c})\), \(v_{c}=(v_{-1,c},v_{1,c})\), \(\tilde{\omega}=(\ldots,\omega_{-2},\omega_{-1,c}+\omega_{-1,s},\omega _{0},\omega_{1,c}+\omega_{1,s},\omega_{2},\ldots)\), \(\tilde{v}=(\ldots ,v_{-2},v_{-1,c}+v_{-1,s},v_{0},v_{1,c}+v_{1,s},v_{2},\ldots)\). Moreover, there exists a positive
C
such that
$$\begin{aligned}& \Phi(0,0)=(0,0), \qquad\|\tilde{\omega}-\omega_{c} \|_{\mathbf{X}} \leq C \bigl(\| \omega_{-1,c}\|^{2}_{\mathbf{H}^{m}}+ \| \omega_{1,c}\|^{2}_{\mathbf{H}^{m}} \bigr), \end{aligned}$$
(2.30)
$$\begin{aligned}& \|\tilde{v}-v_{c}\|_{\mathbf{X}}\leq C \bigl( \|v_{-1,c} \|^{2}_{\mathbf{H}^{m}}+\| v_{1,c} \|^{2}_{\mathbf{H}^{m}} \bigr), \end{aligned}$$
(2.31)
with
\(\tilde{\omega}-\omega_{c}:=(\ldots,0,\omega_{-1,c},0,\omega _{1,c},0,\ldots)\)
and
\(\tilde{v}-v_{c}:=(\ldots,0,v_{-1,c},0,v_{1,c},0,\ldots)\).
Proof
For fixed \(\xi_{1},\xi_{2}>0\) so close to \(\xi_{0}\) and given \(\omega_{\pm 1,c},v_{\pm1,c}\in\mathbf{H}^{m}\) with \(\|\omega_{\pm1,c}\|_{\mathbf {H}^{m}}, \|v_{\pm1,c}\|_{\mathbf{H}^{m}}\leq\sigma_{2}\). Define the operator
$$\begin{aligned} \Gamma: \bigl(\tilde{\omega}^{*},\tilde{v}^{*} \bigr) \longmapsto&(\tilde{\omega}, \tilde {v}) \\ =& \bigl(\tilde{\omega}^{*}+(\ldots,0,\omega_{-1,c},0, \omega_{1,c},0,\ldots ),\tilde{v}^{*}+(\ldots,0,v_{-1,c},0,v_{1,c},0, \ldots) \bigr) \\ \longmapsto&(\omega,v)\longmapsto \bigl(\tilde{\omega}^{**},\tilde {v}^{**} \bigr)=\mbox{the right hand side of (2.26)-(2.29)}, \end{aligned}$$
where \((\omega,v)=(\mathbf{J}\tilde{\omega},\mathbf{J}\tilde{v})\) are defined in Lemma 2.1 and
$$\begin{aligned}& \bigl(\tilde{\omega}^{*},\tilde{v}^{*} \bigr)= \bigl((\ldots,\omega_{-2}, \omega _{-1,s},\omega_{0},\omega_{1,s}, \omega_{2},\ldots),(\ldots ,v_{-2},v_{-1,s},v_{0},v_{1,s},v_{2}, \ldots) \bigr), \\& \begin{aligned}[b] (\tilde{\omega},\tilde{v})&= \bigl(\tilde{\omega}^{*}+ \omega_{c},\tilde {v}^{*}+v_{c} \bigr) \\ &= \bigl(\tilde{\omega}^{*}+(\ldots,0,\omega_{-1,c},0, \omega_{1,c},0,\ldots ),\tilde{v}^{*}+(\ldots,0,v_{-1,c},0,v_{1,c},0, \ldots) \bigr). \end{aligned} \end{aligned}$$
Now we prove the operator Γ is a self-map of a sufficiently small ball in \(\mathcal{X}\). By Lemmas 2.6-2.9 and the form of the nonlinear terms \(\widehat{g^{3}}\), \(\widehat{g^{4}}\) in (2.24)-(2.25), we derive
$$\begin{aligned} \bigl\| \tilde{\omega}^{**}\bigr\| _{\mathbf{X}}\leq{}&C\sup \bigl\{ \bigl\| (in \xi_{1}-M_{1})^{-1}\bigr\| _{\mathbf{H}^{m}\longrightarrow\mathbf{H}^{m}}, \bigl\| (\pm i \xi_{1}-M_{1})^{-1} \mathbf{P}_{\pm1,s} \bigr\| _{\mathbf {H}^{m}\longrightarrow\mathbf{H}^{m}}, \\ &{}\bigl\| (in \xi_{1}-M_{1})^{-1} \nabla^{j} \bigr\| _{\mathbf{H}^{m}\longrightarrow\mathbf{H}^{m}}, \bigl\| (\pm i\xi_{1}-M_{1})^{-1} \nabla^{j}\mathbf{P}_{\pm1,s}\bigr\| _{\mathbf {H}^{m}\longrightarrow\mathbf{H}^{m}} \\ &{}:n\in\mathbf{Z}\setminus\{-1,1\} \bigr\} \times\bigl\| \bigl(\tilde{g}^{3}_{n}( \omega ,u,v) \bigr)_{n\in\mathbf{Z}}\bigr\| _{\mathbf{X}} \\ \leq{}&C\bigl\| \tilde{g}^{3}(\tilde{\omega},u,\tilde{v})\bigr\| _{\mathbf{X}} \\ \leq{}&C\bigl(\|\tilde{\omega}\|_{\mathbf{X}}\|u\|_{\mathbf{X}}+\|\tilde {\omega} \|_{\mathbf{X}}\|\tilde{v}\|_{\mathbf{X}}+\|u\|_{\mathbf{X}}\| \tilde{v} \|_{\mathbf{X}}\bigr) \\ \leq{}&C \bigl(\|\tilde{\omega}\|^{2}_{\mathbf{X}}+\|\tilde{\omega} \|_{\mathbf {X}}\|\tilde{v}\|_{\mathbf{X}} \bigr) \\ \leq{}&C \bigl(\bigl\| \tilde{\omega}^{*}\bigr\| ^{2}_{\mathbf{X}}+\| \omega_{-1,c}\|^{2}_{\mathbf {H}^{m}}+\|\omega_{1,c} \|^{2}_{\mathbf{H}^{m}} +\bigl\| \tilde{v}^{*}\bigr\| ^{2}_{\mathbf{X}}+ \|v_{-1,c}\|^{2}_{\mathbf{H}^{m}}+\| v_{1,c} \|^{2}_{\mathbf{H}^{m}} \bigr) \\ \leq{}&C \bigl(\bigl\| \tilde{\omega}^{*}\bigr\| ^{2}_{\mathbf{X}}+\bigl\| \tilde{v}^{*} \bigr\| ^{2}_{\mathbf {X}}+\sigma_{2}^{2} \bigr), \end{aligned}$$
(2.32)
$$\begin{aligned} \bigl\| \tilde{v}^{**}\bigr\| _{\mathbf{X}}\leq{}&C\sup \bigl\{ \bigl\| ( in \xi_{2}-M_{2})^{-1}\bigr\| _{\mathbf{H}^{m}\longrightarrow\mathbf{H}^{m}}, \bigl\| (\pm i \xi_{2}-M_{2})^{-1}\mathbf{P}_{\pm1,s} \bigr\| _{\mathbf {H}^{m}\longrightarrow\mathbf{H}^{m}}, \\ &{}\bigl\| (in \xi_{2}-M_{2})^{-1} \nabla^{j} \bigr\| _{\mathbf{H}^{m}\longrightarrow\mathbf{H}^{m}}, \bigl\| (\pm i\xi_{2}-M_{2})^{-1} \nabla^{j}\mathbf{P}_{\pm1,s}\bigr\| _{\mathbf {H}^{m}\longrightarrow\mathbf{H}^{m}} \\ &{}:n\in\mathbf{Z}\setminus\{-1,1\} \bigr\} \times\bigl\| \bigl(\tilde{g}^{4}_{n}( \omega ,u,v) \bigr)_{n\in\mathbf{Z}}\bigr\| _{\mathbf{X}} \\ \leq{}&C\bigl\| \tilde{g}^{4}(\tilde{\omega},u,\tilde{v})\bigr\| _{\mathbf{X}} \\ \leq{}&C \bigl(\|\tilde{v}\|^{2}_{\mathbf{X}}+\|\tilde{\omega} \|_{\mathbf{X}}\| \tilde{v}\|_{\mathbf{X}} \bigr) \\ \leq{}&C \bigl(\bigl\| \tilde{\omega}^{*}\bigr\| ^{2}_{\mathbf{X}}+\| \omega_{-1,c}\|^{2}_{\mathbf {H}^{m}}+\|\omega_{1,c} \|^{2}_{\mathbf{H}^{m}} +\bigl\| \tilde{v}^{*}\bigr\| ^{2}_{\mathbf{X}}+ \|v_{-1,c}\|^{2}_{\mathbf{H}^{m}}+\| v_{1,c} \|^{2}_{\mathbf{H}^{m}} \bigr) \\ \leq{}&C \bigl(\bigl\| \tilde{\omega}^{*}\bigr\| ^{2}_{\mathbf{X}}+\bigl\| \tilde{v}^{*} \bigr\| ^{2}_{\mathbf {X}}+\sigma_{2}^{2} \bigr), \end{aligned}$$
(2.33)
where
$$\begin{aligned}& \tilde{g}^{3}=2\omega\cdot u^{T}+\nu_{0}'v \cdot u^{T}+v\cdot\omega^{T}, \\& \tilde{g}^{4}=u\cdot v^{T}+2\kappa_{0}'v \cdot v^{T}. \end{aligned}$$
Thus, for \(\sigma_{2}\leq\frac{1}{\sqrt{2C}}\) and \((\tilde{\omega }^{*},\tilde{v}^{*})\in\mathcal{X}\) with \(\|(\tilde{\omega}^{*},\tilde{v}^{*})\| _{\mathcal{X}}\leq\frac{1}{\sqrt{2C}}\), we have
$$\begin{aligned} \bigl\| \Gamma \bigl(\tilde{\omega}^{*},\tilde{v}^{*} \bigr)\bigr\| _{\mathcal{X}} =&\bigl\| \tilde{ \omega }^{**}\bigr\| _{\mathcal{X}}+\bigl\| \tilde{v}^{**} \bigr\| _{\mathbf{X}} \\ \leq&C \bigl( \bigl(\bigl\| \tilde{\omega}^{*}\bigr\| _{\mathbf{X}}+\bigl\| \tilde{v}^{*} \bigr\| _{\mathbf {X}} \bigr)^{2}+\sigma_{2}^{2} \bigr) \leq1, \end{aligned}$$
which implies that for sufficient small \(\sigma_{2}>0\). Hence, by the Banach fixed point theorem, we obtain a unique fixed point \((\tilde{\theta}^{*},\tilde{v}^{*})\in \mathcal{X}\) of Γ, which means that (2.26)-(2.29) have a solution of \((\tilde{\omega},\tilde{v})=(\tilde{\omega}^{*}+\omega _{c},\tilde{v}^{*}+v_{c})\). Moreover, if \((\omega_{\pm1,c},v_{\pm 1,c})=(0,0)\), then \(\Phi(0,0)=(0,0)\). Next we prove the second inequality in (2.30) and (2.31). Note that \((\tilde{\omega}^{*},\tilde{v}^{*})=\Gamma(\tilde{\omega}^{*},\tilde {v}^{*})=(\tilde{\theta}^{**},\tilde{v}^{**})\). This, combined with (2.32)-(2.33), allows us to derive
$$\begin{aligned}& \begin{aligned}[b] \|\tilde{\omega}-\omega_{c}\|_{\mathbf{X}}&=\bigl\| \tilde{ \omega}^{*}\bigr\| _{\mathbf{X}}=\bigl\| \tilde{\omega}^{**} \bigr\| _{\mathbf{X}} \\ &\leq C \bigl(\bigl\| \tilde{\omega}^{*}\bigr\| ^{2}_{\mathbf{X}}+\| \omega_{-1,c}\| ^{2}_{\mathcal{H}^{m}}+\|\omega_{1,c} \|^{2}_{\mathcal{H}^{m}} \bigr), \end{aligned} \\& \begin{aligned}[b] \|\tilde{v}-v_{c}\|_{\mathbf{X}}&=\bigl\| \tilde{v}^{*} \bigr\| _{\mathbf{X}}=\bigl\| \tilde {v}^{**} \bigr\| _{\mathbf{X}} \\ &\leq C \bigl(\bigl\| \tilde{v}^{*}\bigr\| ^{2}_{\mathbf{X}}+\|v_{-1,c} \|^{2}_{\mathcal{H}^{m}}+\| v_{-1,c}\|^{2}_{\mathcal{H}^{m}} \bigr). \end{aligned} \end{aligned}$$
Thus we deduce that for a sufficient small ball \(\mathbf{B}_{r}(0)\subset \mathbf{B}_{1}(0)\),
$$\begin{aligned}& \|\tilde{\omega}-\omega_{c}\|_{\mathbf{X}}\leq C \bigl(\| \omega_{-1,c}\| ^{2}_{\mathbf{H}^{m}}+\|\omega_{1,c} \|^{2}_{\mathbf{H}^{m}} \bigr), \\& \|\tilde{v}-v_{c}\|_{\mathbf{X}}\leq C \bigl(\|v_{-1,c} \|^{2}_{\mathbf{H}^{m}}+\| v_{1,c}\|^{2}_{\mathbf{H}^{m}} \bigr), \end{aligned}$$
where
$$\begin{aligned}& \tilde{\omega}-\omega_{c}:=(\ldots,0,\omega_{-1,c},0, \omega _{1,c},0,\ldots), \\& \tilde{v}-v_{c}:=(\ldots,0,v_{- 1,c},0,v_{1,c},0, \ldots). \end{aligned}$$
This completes the proof. □
It remains to analyze (2.18)-(2.19). We restate the equations:
$$\begin{aligned}& (\pm i\xi_{1}-M_{1})\omega_{\pm1,c}= \mathbf{P}_{\pm1,c}g^{3}_{\pm1}(\omega ,u,v), \\& (\pm i\xi_{2}-M_{2})v_{\pm1,c}=\mathbf{P}_{\pm1,c}g^{4}_{\pm1}( \omega,u,v). \end{aligned}$$
It follows from \((\omega_{-1},v_{-1})=(\overline{\omega_{1}},\overline {v_{1}})\) and \((g^{3}_{-1},g^{4}_{-1})=(\overline{g^{3}_{1}},\overline {g^{4}_{1}})\) that the ‘−’ equation is the complex conjugate of the ‘+’ equation. By Lemma 2.1, we can denote \((\omega,v)=(\mathbf {J}\tilde{\omega},\mathbf{J}\tilde{v})\) by means of \((\tilde{\omega},\tilde{v})=\Phi(\omega_{c},v_{c})=\Phi((\overline{\omega _{1,c}},\omega_{1,c}),(\overline{v_{1,c}},v_{1,c}))\). Our target is to find \((\xi_{1},\alpha)\) and \((\xi_{2},\alpha)\) close to \((\xi_{0},\alpha_{c})\) and a nontrivial solution \((\omega _{1,c},v_{1,c})=(\omega_{1,c},v_{1,c})(x)\) of
$$\begin{aligned}& -i\xi_{1}\omega_{1,c}+M_{1} \omega_{1,c}+ \mathbf{P}_{1,c}g^{3}_{1} \bigl(\mathbf {J}\Phi( \overline{\omega_{1,c}},\omega_{1,c}, \overline {v_{1,c}},v_{1,c}) \bigr)=0, \end{aligned}$$
(2.34)
$$\begin{aligned}& -i\xi_{2}v_{1,c}+M_{2}v_{1,c}+ \mathbf{P}_{1,c}g^{4}_{1} \bigl(\mathbf{J}\Phi ( \overline{\omega_{1,c}},\omega_{1,c},\overline{v_{1,c}},v_{1,c}) \bigr)=0. \end{aligned}$$
(2.35)
Since \(\omega_{1,c},v_{1,c}\in\mathrm{C}\psi^{+}\) and \((M_{1}\psi^{+},M_{2}\psi ^{+})=(\lambda_{0}^{+}(\alpha)\psi^{+},\mu_{0}^{+}(\alpha)\psi^{+})\), we can write \(\omega_{1,c}=\eta\psi^{+}\) and \(v_{1,c}=\delta\psi^{+}\). Then by (2.34)-(2.35), for some \(\eta,\delta\in\mathrm {C}\setminus\{0\}\), we obtain
$$\begin{aligned}& -i\xi_{1}\eta\psi^{+}+\lambda_{0}^{+}(\alpha) \eta\psi^{+}+\mathbf {P}_{1,c}g^{3}_{1} \bigl( \mathbf{J}\Phi \bigl( \overline{\eta\psi^{+}},\eta\psi ^{+},\overline{\delta\psi^{+}}, \delta\psi^{+} \bigr) \bigr)=0, \end{aligned}$$
(2.36)
$$\begin{aligned}& -i\xi_{2}\delta\psi^{+}+\mu_{0}^{+}(\alpha) \delta\psi^{+}+\mathbf {P}_{1,c}g^{4}_{1} \bigl( \mathbf{J}\Phi \bigl( \overline{\eta\psi^{+}},\eta\psi ^{+},\overline{\delta\psi^{+}}, \delta\psi^{+} \bigr) \bigr)=0. \end{aligned}$$
(2.37)
Now we introduce \((p_{1,c},\theta_{1,c})\) by \((\mathbf{P}_{1,c}\omega,\mathbf{P}_{1,c}v)=(p_{1,c}(\omega)\psi^{+}, \theta_{1,c}(v)\psi^{+})\). Then (2.36)-(2.37) can be simplified to
$$\begin{aligned} -i\xi_{1}\eta+\lambda_{0}^{+}(\alpha) \eta+g^{3}( \alpha,\eta,\delta )=0,\quad \mbox{for some }\eta\in\mathrm{C}, \end{aligned}$$
(2.38)
$$\begin{aligned} -i\xi_{2}\delta+\mu_{0}^{+}(\alpha) \delta+g^{4}( \alpha,\eta,\delta )=0,\quad \mbox{for some }\delta\in\mathrm{C}, \end{aligned}$$
(2.39)
where the cubic coefficient \(a_{1}\neq0\) and \(a_{2}\neq0\) in
$$\begin{aligned}& g^{3}(\alpha,\eta,\delta):=p_{1,c} \bigl(g^{3}_{1} \bigl(\mathbf{J}\Phi \bigl(\overline {\eta\psi^{+}}, \eta\psi^{+},\overline{\delta \psi^{+}},\delta\psi^{+} \bigr) \bigr)\bigr), \end{aligned}$$
(2.40)
$$\begin{aligned}& g^{4}(\alpha,\eta,\delta):=\theta_{1,c} \bigl(g^{4}_{1} \bigl(\mathbf{J}\Phi \bigl(\overline{\eta\psi^{+}}, \eta\psi^{+},\overline{\delta \psi^{+}},\delta\psi ^{+} \bigr) \bigr)\bigr). \end{aligned}$$
(2.41)
Note that \(|p_{1,c}(\omega)|\leq C\|\mathbf{P}_{1,c}\omega\|_{\mathbf {H}^{m}}\leq C\|\omega\|_{\mathbf{H}^{m}}\) and \(|\theta_{1,c}(v)|\leq C\|\mathbf{P}_{1,c}v\|_{\mathbf{H}^{m}}\leq C\|v\| _{\mathbf{H}^{m}}\). Thus by (2.30)-(2.31) and (2.40)-(2.41), we derive
$$\begin{aligned}& \bigl|p_{1,c} \bigl(g^{3}_{1} \bigl(\mathbf{J}\Phi \bigl( \overline{\eta\psi^{+}},\eta\psi ^{+},\overline{\delta\psi^{+}},\delta\psi^{+} \bigr) \bigr)\bigr)\bigr|\leq C \bigl(|\eta|^{2}+|\delta |^{2} \bigr), \\& \bigl|\theta_{1,c} \bigl(g^{4}_{1} \bigl(\mathbf{J}\Phi \bigl(\overline{\eta\psi^{+}},\eta\psi ^{+},\overline{\delta\psi^{+}},\delta\psi^{+} \bigr) \bigr)\bigr)\bigr| \leq C \bigl(|\eta|^{2}+|\delta|^{2} \bigr), \end{aligned}$$
where we use the following relation: \((\tilde{\omega},\tilde{v})=\Phi(\omega_{c},v_{c})=\Phi(\overline{\eta\psi ^{+}},\eta\psi^{+},\overline{\delta\psi^{+}},\delta\psi^{+})\).
Analogously to the case of the classical Hopf-bifurcation result [10], the Banach fixed point theorem can be applied to (2.38)-(2.39), as soon as the zero solution is divided out. It is sufficient to find the real value solutions (i.e.
\((\gamma_{1},\gamma_{2})=(\eta ,\delta)\in\mathbf{R}^{2}\)) of (2.38)-(2.39). Hence we define the complex-valued smooth function
$$\begin{aligned}& \Upsilon^{1}(\gamma_{1},\gamma_{2};\varrho, \epsilon):=\left \{ \begin{array}{@{}l} -i(\xi_{0}+\varrho)+\lambda_{0}^{+}(\alpha_{c}+\epsilon)+\gamma _{1}^{-1}g^{3}(\alpha_{c}+\epsilon,\gamma_{1},\gamma_{2}), \qquad \gamma_{1}\neq0,\\ -i(\xi_{0}+\varrho)+\lambda_{0}^{+}(\alpha_{c}+\epsilon),\qquad \gamma_{1}=0, \end{array} \right . \\& \Upsilon^{2}(\gamma_{1},\gamma_{2};\varrho, \epsilon):=\left \{ \begin{array}{@{}l} -i(\xi_{0}+\varrho)+\mu_{0}^{+}(\alpha_{c}+\epsilon)+\gamma_{2}^{-1}g^{4}(\alpha _{c}+\epsilon,\gamma_{1},\gamma_{2}),\qquad \gamma_{2}\neq0,\\ -i(\xi_{0}+\varrho)+\mu_{0}^{+}(\alpha_{c}+\epsilon), \qquad \gamma_{2}=0. \end{array} \right . \end{aligned}$$
It follows from \((\lambda^{+}_{0}(\alpha_{c}),\mu^{+}_{0}(\alpha_{c}))=(i\xi_{0},i\xi _{0})\) that \((\Upsilon^{1}(0,0,0,0),\Upsilon^{2}(0,0,0,0))=(0,0)\). Moreover, by assumption (H2) for the Jacobi matrix
$$\begin{aligned}& \mathbf{D}_{\rho,\epsilon}\Upsilon^{1}(\gamma_{1}, \gamma_{2};\varrho,\epsilon )|_{\gamma_{1}=\gamma_{2}=\varrho=\epsilon=0}= \begin{pmatrix} 0& \frac{d}{d\beta}\operatorname{Re}\lambda_{0}^{+}(\alpha)|_{\alpha=\alpha_{c}} \\ -1 & \frac{d}{d\beta}\operatorname{Im}\lambda_{0}^{+}(\alpha)|_{\alpha=\alpha_{c}} \end{pmatrix}, \\& \mathbf{D}_{\rho,\epsilon}\Upsilon^{2}(\gamma_{1}, \gamma_{2};\varrho,\epsilon )|_{\gamma_{1}=\gamma_{2}=\varrho=\epsilon=0}= \begin{pmatrix} 0& \frac{d}{d\beta}\operatorname{Re}\mu_{0}^{+}(\alpha)|_{\alpha=\alpha_{c}} \\ -1 & \frac{d}{d\beta}\operatorname{Im}\mu_{0}^{+}(\alpha)|_{\alpha=\alpha_{c}} \end{pmatrix} \end{aligned}$$
with respect to ρ, ϵ one has \(\operatorname{det}\mathbf{D}_{\rho,\epsilon}\Upsilon^{1}(\gamma_{1},\gamma_{2};\varrho ,\epsilon)|_{\gamma_{1}=\gamma_{2}=\varrho=\epsilon=0}=\frac{d}{d\beta }\operatorname{Re}\lambda_{0}^{+}(\alpha)|_{\alpha=\alpha_{c}}>0\) and \(\operatorname{det}\mathbf{D}_{\rho,\epsilon}\Upsilon^{2}(\gamma,\gamma_{2};\varrho,\epsilon )|_{\gamma_{1}=\gamma_{2}=\varrho=\epsilon=0}=\frac{d}{d\beta}\operatorname{Re}\mu _{0}^{+}(\beta)|_{\alpha=\alpha_{c}}>0\). Thus for sufficient small \(\gamma_{1},\gamma_{2}>0\), we can find a function \(\gamma_{1}\mapsto(\varrho(\gamma_{1}),\epsilon(\gamma_{1}))\) and \(\gamma _{2}\mapsto(\varrho(\gamma_{2}),\epsilon(\gamma_{2}))\) with \(\varrho (0)=\epsilon(0)=0\) such that
$$\begin{aligned}& -i \bigl(\xi_{0}+\varrho(\gamma_{1}) \bigr)+ \lambda_{0}^{+} \bigl(\alpha_{c}+\epsilon(\gamma _{1}) \bigr)-\gamma_{1}^{-1}g^{3} \bigl( \alpha_{c}+\epsilon(\gamma_{1}),\gamma_{1},\alpha _{c}+\epsilon(\gamma_{2}),\gamma_{2} \bigr)=0, \\& -i \bigl(\xi_{0}+\varrho(\gamma_{2}) \bigr)+ \mu_{0}^{+} \bigl(\alpha_{c}+\epsilon(\gamma _{2}) \bigr)-\gamma_{2}^{-1}g^{4} \bigl( \alpha_{c}+\epsilon(\gamma_{1}),\gamma_{1},\alpha _{c}+\epsilon(\gamma_{2}),\gamma_{2} \bigr)=0. \end{aligned}$$
Note the degree of nonlinearity. Then it follows from differentiating this equation that \(\epsilon^{(i)}\neq0\) for some first i. Hence the function \(\gamma_{1}\mapsto\epsilon(\gamma_{1})\) and \(\gamma_{1}\mapsto \epsilon(\gamma_{2})\) can locally be inverted to yield a function \(\epsilon\mapsto\gamma_{1}(\epsilon)\) and \(\epsilon\mapsto\gamma_{2}(\epsilon)\). This implies that the following equation holds:
$$\begin{aligned}& -i \bigl(\xi_{0}+\varrho \bigl(\gamma_{1}(\epsilon) \bigr) \bigr)\gamma_{1}(\epsilon)+\lambda _{0}^{+}( \alpha_{c}+\epsilon)\gamma_{1}(\epsilon)-g^{3} \bigl(\alpha_{c}+\epsilon,\gamma _{1}(\epsilon), \gamma_{2}(\epsilon) \bigr)=0, \\& -i \bigl(\xi_{0}+\varrho \bigl(\gamma_{2}(\epsilon) \bigr) \bigr)\gamma_{2}(\epsilon)+\mu_{0}^{+}(\alpha _{c}+ \epsilon)\gamma_{2}(\epsilon)-g^{4} \bigl( \alpha_{c}+\epsilon,\gamma_{1}(\epsilon ), \gamma_{2}(\epsilon) \bigr)=0, \end{aligned}$$
for sufficient small \(\epsilon>0\).
Therefore we obtain the desired solutions of (2.34)-(2.35) by setting \((\xi_{1},\xi_{2})=(\xi_{0}+\varrho(\gamma_{1}(\epsilon )),\xi_{0}+\varrho(\gamma_{2}(\epsilon)))\), \(\alpha=\alpha_{c}+\epsilon\), and \((\omega_{1,c},v_{1,c})=(\gamma_{1}(\epsilon)\psi^{+}_{\alpha_{c}+\epsilon },\gamma_{2}(\epsilon)\psi^{+}_{\alpha_{c}+\epsilon})(x)\). This result combined with Lemmas 3.2, 3.8, and (1.10) gives the proof of Theorem 1.1.