\documentclass[reqno]{amsart} \usepackage{psfig} \begin{document} {\noindent\small {\em Electronic Journal of Differential Equations}, Vol.~2000(2000), No.~58, pp.~1--32.\newline ISSN: 1072-6691. URL: http://ejde.math.swt.edu or http://ejde.math.unt.edu \newline ftp ejde.math.swt.edu \quad ejde.math.unt.edu (login: ftp)} \thanks{\copyright 2000 Southwest Texas State University and University of North Texas.} \vspace{1cm} \title[\hfilneg EJDE--2000/58\hfil Three-dimensional Kolmogorov problem] {Steady-state bifurcations of the three-dimensional Kolmogorov problem} \author[Z. M. Chen \& S. Wang\hfil EJDE--2000/58\hfilneg] {Zhi-Min Chen \& Shouhong Wang} \address{Zhi-Min Chen \hfill\break Department of Ship Science, Southampton University, Southampton SO17 1BJ, UK \hfill\break and: Department of Mathematics, Tianjin University, China} \email{zhimin@ship.soton.ac.uk} \address{Shouhong Wang \hfill\break Department of Mathematics, Indiana University, Bloomington, IN 47405, USA} \email{showang@indiana.edu} \date{} \thanks{Submitted February 12, 1999. Revised May 26, 2000. Published August 30, 2000.} \subjclass{35Q30, 76D05, 58J55, 35B32} \keywords{3D Navier-Stokes equations, Kolmogorov flow, multiple steady states, \hfill\break\indent supercritical pitchfork bifurcation, continuous fractions} \begin{abstract} This paper studies the spatially periodic incompressible fluid motion in $\mathbb R^3$ excited by the external force $k^2(\sin kz, 0,0)$ with $k\geq 2$ an integer. This driving force gives rise to the existence of the unidirectional basic steady flow $u_0=(\sin kz,0, 0)$ for any Reynolds number. It is shown in Theorem 1.1 that there exist a number of critical Reynolds numbers such that $u_0$ bifurcates into either 4 or 8 or 16 different steady states, when the Reynolds number increases across each of such numbers. Thanks to the Rabinowitz global bifurcation theorem, all of the bifurcation solutions are extended to global branches for $\lambda \in (0, \infty)$. Moreover we prove that when $\lambda$ passes each critical value, a) all the corresponding global branches do not intersect with the trivial branch $(u_0,\lambda)$, and b) some of them never intersect each other; see Theorem 1.2. \end{abstract} \maketitle \renewcommand{\theequation}{\arabic{section}.\arabic{equation}} \renewcommand\thefigure{\arabic{section}.\arabic{figure}} \newtheorem{Theorem}{Theorem}[section] \newtheorem{Lemma}{Lemma}[section] \newtheorem{Remark}{Remark}[section] \def\dbH{\dot{\mathbb H}} \def\tilde{\widetilde} \section{Introduction} \setcounter{equation}{0} The three-dimensional Kolmogorov problem, which was first formulated by Kolmogorov (see \cite{AM}), refers to the Navier-Stokes equations defining the spatially periodic fluid motion in the following form: \begin{gather} \frac{\partial u}{\partial t} -\Delta u+\lambda (u\!\cdot\!\nabla) u + \lambda\nabla p =k^2(\sin kz,0,0), \label{ns1}\\ \text{div }u = 0, \label{ns2}\\ u(t,x,y,z) =u(t,x+2\pi,y,z)=u(t,x,y+2\pi,z)=u(t,x,y,z+2\pi), \label{ns3}\\ \int_{{\mathbb T}^3} u\,dx\,dy\,dz= 0\,. \label{ns4} \end{gather} Here $u=(u_1,u_2,u_3)$ is the velocity field, $p$ the pressure, $\lambda >0$ the Reynolds number in this dimensionless formulation, $k$ a positive integer, and ${\mathbb T}^n={{\mathbb R}}^n/(2 \pi {\mathbb Z})^n$ ($n=2, 3$) the n-dimensional flat torus. In particular $u_0=(\sin kz,0,0)$ solves this problem for all $\lambda$. There have been extensive mathematical and physical studies for the Kolmogorov problem as well as for general fluid equations; see for instance \cite{CP1,FSV, Green,Iu,Kir, L1, Mar, MS,Obu,Oka,Sirov,Rab, Rab68, Rab73,serrin, velte}, and the references therein. From the bifurcation point of view, most of the literatures are devoted to the existence of secondary steady-states of the Navier-Stokes equations, and in particular the detailed rigorous mathematical analysis on the bifurcation of the solutions of the Kolmogorov problem as well as the general Navier-Stokes equations is still in its early stage, due partially to the difficulty in estimating the eigenvalues of the associated linearized operators. The main objective of this paper is to establish the existence of multi-branches of steady-state solutions bifurcated respectively from $u_0$ at some critical Reynolds numbers. The multi-branches are obtained in some flow invariant subspaces, which are defined using special Fourier modes in terms of some $n$--tuple integer vectors. Let $\mathbb N$ be the set of all positive integers, and $\mathbb Z$ be the integer set. The following condition for an integer vector $(l,j,i,k)$ will be used often to define some invariant function spaces: \begin{equation} \label{4-tuple} \gathered (l^2+j^2+i^2-k^2)(l^2+j^2+(k-i)^2-k^2)<0, \\ 0\leq l 0 \mbox{ and }j\{n_0i\}=0,\\ 16,& \mbox{ when } j\{n_0i\}\neq 0. \end{array}\right. \end{equation} Here $\{n\}$ is defined by \begin{equation} \label{mod-k} \{n\} \equiv n \mbox{\rm \quad (mod $k$)}. \end{equation} Let $H^m({\mathbb T}^3)$ be the usual Sobolev space of scalar and periodic functions endowed with the usual $H^m$ norm $|| \cdot ||_{H^m}$, and let ${\mathbb H}^m({\mathbb T}^3)=H^m({\mathbb T}^3)^3$ be the vectorial Sobolev space. For $m\ge 1$, we use the following function spaces of divergence-free vector fields: $${\mathbb H}^m_{\sigma}= \left\{u\in {\mathbb H}^{m}({\mathbb T}^3) \, |\,\text{ div } u=0 \right\}.$$ Let $\dot{H}^m({\mathbb T}^3)=H^m({\mathbb T}^3)/{\mathbb R}$, $\dbH^m({\mathbb T}^3) = {\mathbb H}^m({\mathbb T}^3)/ {\mathbb R}$, and $\dbH^2_{\sigma}= {\mathbb H}^2_{\sigma}/{\mathbb R}$. When $m=2$, we can use $\|u\|_{H^2_\sigma}=\|\Delta u\|_{L^2}$ as the norm of $\dbH^2_{\sigma}$. \begin{figure} $$\psfig{figure=fig1.eps,width=10cm}$$ \caption{The global branch $u_{l,j,i,k,n,\lambda}$ in $\mathcal C_1$. Here the dotted horizontal lines represent $ \| u\|_{H^1} = c_2 k^2 $, and dotted vertical line is $\displaystyle \lambda = \frac{c_1}{k^2}$.} \end{figure} \begin{figure} $$\psfig{figure=fig2.eps,width=10cm}$$ \caption{The global branch $u_{l,j,i,k,n,\lambda}$ in $\mathcal C_2$. Here the dotted vertical line is $\displaystyle \lambda = \frac{c_1}{k^2}$, while the dotted curves represent $\| u\|_{H^2} = c_3 k^2 (\lambda^2 k^4 + 1)$.} \end{figure} As we mentioned before, the main objective of this article is to find and possibly classify steady-state bifurcations of the Kolmogorov problem in some subspaces of $\dbH^2_{\sigma}$ invariant to the Navier-Stokes equations. Now we state two main theorems of this article without specify invariant subspaces, which will be explicitly given in Sections 4--6. Thanks to the Rabinowitz global bifurcation theorem, we obtain in these two theorems $i_0$ different global branches of steady-state solutions bifurcating from $u_0$ at a single critical value. Each of those branches undergoes local supercritical bifurcation around the bifurcation point, whereas half ($i_0/2$) of those branches never touch each other away from the bifurcation point. \begin{Theorem} Let $(l,j,i,k)$ be an integer vector satisfying (\ref{4-tuple}). Then there exists a critical Reynolds number $\lambda_{l,j,i,k}>0$ such that (\ref{ns1}-\ref{ns4}) with $0<\lambda <\infty $ admit $i_0$ steady-state solutions $u_{l,j,i,k,n,\lambda}$ $(n=1,...,i_0)$ branching off $(\lambda_{l,j,i,k},u_0)$ continuously such that: \footnote{See (\ref{5.1}--\ref{5.5}) for specific spaces where these bifurcation solutions are located.} \begin{enumerate} \item[a).] if $\lambda_{l,j,i,k}<\lambda$, then $$ u_{l,j,i,k,n,\lambda} =u_0,$$ \item[b).] if $\lambda_{l,j,i,k}<\lambda $ and $\|\Delta(u_{l,j,i,k,n,\lambda}-u_0)\|_{L^2} + |\lambda-\lambda_{l,j,i,k}| <\epsilon $ for some small $\epsilon> 0 $, then \begin{equation}\label{1.3} u_0\neq u_{l,j,i,k,n,\lambda} \neq u_{l,j,i,k,n',\lambda}\neq u_0. \end{equation} \end{enumerate} \end{Theorem} We remark that for one value of $\lambda \in (0, \infty)$, there might be more than one steady-state solutions on the global bifurcation branch as shown in Figures 1.1 and 1.2. It is easy to see that all steady-state solutions of the Kolmogorov problem lie in both the following subsets of $\dbH^1_\sigma\times (0, \infty)$ and $\dbH^2_\sigma\times (0, \infty)$: \begin{align} \mathcal C_1= & \left\{ (u, \lambda) \in \dbH^1_\sigma \times (0, \infty) \quad | \quad \| u\|_{H^1} \le c_2 k^2, \quad \lambda \ge \frac{c_1}{k^2} \right\} \\ & \cup \left\{ (0, \lambda) \quad | \quad 0 < \lambda \le \frac{c_1}{k^2} \right\}, \nonumber \\ \mathcal C_2= & \left\{ (u, \lambda) \in \dbH^2_\sigma \times (0, \infty) \quad | \quad \| u\|_{H^2} \le c_3 k^2 (\lambda^2 k^4 + 1), \quad \lambda \ge \frac{c_1}{k^2} \right\} \\ & \cup \left\{ (0, \lambda)\quad | \quad 0 < \lambda \le \frac{c_1}{k^2} \right\}. \nonumber \end{align} Here the absolute constants $c_1$, $c_2$ and $c_3$ are given in Lemma 3.2 in Section 3. The set $\mathcal C_1$ is as shown in Figure 1.1. Since stronger $H^2$ norm is used in $\mathcal C_2$, the schematic picture of $\mathcal C_2$ shown in Figure 1.2 is similar to $\mathcal C_1$ depicted in Figure 1.1 but with the dotted horizontal lines replaced by quadratic line of $\lambda$ given by $ \| u\|_{H^2} = c_3 k^2 (\lambda^2 k^4 + 1).$ \begin{Theorem} Let $(l,j,i,k)$ be an integer vector satisfying (\ref{4-tuple}). Then all bifurcation solutions $u_{l,j,i,k,n,\lambda}$ given by Theorem 1.1 satisfying the following properties: \begin{enumerate} \item each branch $u_{l,j,i,k,n,\lambda}$ extends to $\lambda=\infty$ in $\mathcal C_1 \cap \mathcal C_2$ as shown in Figures 1.1 and 1.2; \item each branch $u_{l,j,i,k,n,\lambda}$ intersects with the $\lambda$--axis only at the line segment $pq$. In particular, $u_{l,j,i,k,n,\lambda}$ never touches the $\lambda$--axis for $\lambda > \lambda_{l,j,i,k}$; \item for $ j \neq 0$ and for $i_1=0$ or $i_0/2$, \begin{equation} u_{l,j,i,k,n,\lambda}\neq u_{l,j,i,k,n',\lambda},\label{1.4} \end{equation} if $$\lambda_{l,j,i,k}<\lambda<\infty,\, i_1+1\leq n\leq i_1+\frac{i_0}{4}0$ $$\gathered \frac{d\rho_{l,j,i,k}(\lambda)}{d\lambda} >0, \\ \lim_{\lambda\rightarrow \infty}\rho_{l,j,i,k}(\lambda)=\infty,\\ \lim_{\lambda\rightarrow 0^+}\rho_{l,j,i,k}(\lambda)=-l^2-j^2-\min\{i^2, (k-i)^2\}. \endgathered $$ Furthermore for any $\lambda >0$ and $\rho > -(l^2+j^2+\min\{i^2,(k-i)^2\})$, \begin{equation}\nonumber \begin{split} & \mbox{\rm dim}\bigcup_{m\in {\mathbb N}} \{u\in {\mathbb E}_{l,j,i,k}|\,\,(\Delta -\lambda A -\rho)^m u=0\}\leq 1 ,\\ & \mbox{\rm dim} \bigcup_{m\in {\mathbb N}} \{u\in \tilde{{\mathbb E}}_{l,j,i,k}|\,\, (\Delta -\lambda A -\rho)^m u=0\} \leq 1, \\ & \mbox{\rm dim}\bigcup_{m\in {\mathbb N}} \{u\in {\mathbb E}_{l,j,k-i,k}|\,\, (\Delta -\lambda A-\rho )^mu=0\}\leq 1 \quad \mbox{\rm if }i\neq 0, \\ & \mbox{\rm dim}\bigcup_{m\in {\mathbb N}} \{u\in \tilde{{\mathbb E}}_{l,j,k-i,k}|\,\, (\Delta -\lambda A -\rho)^m u=0\} \leq 1 \quad \mbox{ if }i\neq 0, \end{split} \end{equation} where the equalities hold if and only if $\rho=\rho_{l,j,i,k}(\lambda)$. \end{Theorem} The proof of Theorem 2.1 will be accomplished by converting the eigenvalue problem $ \Delta u -\lambda A u = \rho u$ to coupled continuous fraction equations. This approach was first used by Meshalkin and Sinai \cite{MS} in a stability problem of fluid flows. To proceed, we first recall a theorem on continuous fraction \cite{MS}. Consider three term recurrence equations: \begin{equation} d_n e_n +e_{n-1} -e_{n+1}=0, \quad n \in {\mathbb Z}, \label{2.5} \end{equation} where $d_n$ and $e_n$ are complex numbers. Then we have \begin{Theorem} Assume \begin{equation}\label{2.6} \gathered \text{Re} \, d_n > 0, for \,n\not= 0, \\ \lim_{|n| \to \infty} \text{Re} \, d_n = \infty. \endgathered \end{equation} Then the following two assertions are equivalent: \begin{enumerate} \item[(i)] There exists a nontrivial solution $\{ e_n\}_{n\in {\mathbb Z}}$ of (\ref{2.5}) such that \begin{equation} \sum_{n\in {\mathbb Z}}|e_n|^2 < \infty. \label{2.7} \end{equation} \item[(ii)] The following equation holds true: \begin{equation} \label{2.8} d_0 + \displaystyle\frac{1} { \displaystyle d_{-1} +\displaystyle\frac{1}{\displaystyle d_{-2}+ \displaystyle\frac{1}{\ddots}}} = \displaystyle\frac{- 1} { \displaystyle d_{1} +\displaystyle\frac{1}{\displaystyle d_{2}+ \displaystyle\frac{1}{\ddots}}}. \end{equation} \end{enumerate} Furthermore, the solution obtained in (i) or (ii) is unique up to a constant factor and satisfies that \begin{equation} \text{ either $e_n = 0$ for all $n\in {\mathbb Z}$, or $e_n \not= 0$ for all $n \in {\mathbb Z}$.} \label{2.9} \end{equation} \end{Theorem} \begin{proof} 1. Let $\{e_n\}_{n\in {\mathbb Z}}$ be a nontrivial solution of (\ref{2.5}). If $e_{n_0}=0$ for some $n_0\geq 0$, multiplying the $n$-th equation of (\ref{2.5}) by $\bar e_n$ and summing up the resultant equations for $n\geq n_0+1$ yields $$\sum_{n\geq n_0+1}\mbox{Re} d_n|e_n|^2=0.$$ This together with (\ref{2.6}) gives $e_{n}=0$ for $n\geq n_0+1$, and thus (\ref{2.5}) shows $e_n=0$ for all $n\in {\mathbb Z}$. In the same way, the assumption $e_{n_0}=0$ for some $n_0\leq -1$ implies $e_n=0$ for $n\leq n_0-1$ and then $e_n=0$ for all $n\in {\mathbb Z}$. Namely (\ref{2.9}) holds true. 2. Dividing the $n$th and $-n$th equation of (\ref{2.5}) by $e_n$ and $e_{-n}$ respectively yields that for any $n\ge 0$, \begin{align} & \frac{e_{n}}{e_{n-1}}=\frac{-1}{d_n-\displaystyle\frac{e_{n+1}}{e_n} }, \nonumber \\ & \frac{e_{-n}}{e_{-n+1}}=\frac{1} {d_{-n}+\displaystyle\frac{e_{-n-1}}{e_{-n}}}. \nonumber \end{align} Namely, for any $n\ge 0$, $$ \frac{e_{\pm n}}{e_{\pm (n-1)}}=\frac{\mp 1} {d_{\pm n}+\displaystyle\frac{1}{d_{\pm(n+1)} +\displaystyle\frac{1}{\ddots +\displaystyle\frac{1}{d_{\pm (n+m)} \mp\displaystyle\frac{e_{\pm( n+m+1)}}{e_{\pm (n+m)}} } }} }.$$ This together with (\ref{2.6}-\ref{2.7}) implies that for any $n\ge 0$, \begin{equation}\label{2.10} \frac{e_{\pm n}}{e_{\pm (n-1)}} =\gamma_{\pm n}\overset{ \text{def} }{=} \frac{\mp 1} {d_{\pm n}+\displaystyle\frac{1}{d_{\pm(n+1)} + \displaystyle\frac{1}{d_{\pm (n+2)} +\displaystyle\frac{1}{\ddots}}}}. \end{equation} In particular, (\ref{2.8}) follows from $(e_0/e_{-1})^{-1}=e_{-1}/e_0$. 3. Furthermore we infer from (\ref{2.10}) that for any $n\geq 1$, \begin{equation}\nonumber \begin{split} &e_n=e_0\gamma_1\cdots\gamma_{n}, \\ &e_{-n}=e_0\gamma_{-1}\cdots \gamma_{-n}, \end{split} \end{equation} which are uniquely determined up to constants, i.e. an arbitrary choice of $e_0$. 4. Moreover, if (\ref{2.8}) is valid, we can also define $\{\gamma_n\}_{n\in {\mathbb Z}}$ and a nontrivial solution $\{e_n\}_{n\in {\mathbb Z}}$ as above. \end{proof} \begin{proof}[Proof of Theorem 2.1] We divide the proof into several steps. {\sl Step 1.} We observe that the condition (\ref{4-tuple}) is equivalent to either \begin{equation}\label{4-tuple1} \gathered l^2+j^2+i^2 < k^2 , \\ l^2+j^2+(k-i)^2 > k^2 , \\ 0\leq i < k/2, \\ j \in {\mathbb Z}, \\ 0\leq l < k, \endgathered \end{equation} or \begin{equation}\label{4-tuple2} \gathered l^2+j^2+i^2 > k^2 , \\ l^2+j^2+(k-i)^2 < k^2 , \\ k/2 < i < k, \\ j \in {\mathbb Z}, \\ 0\leq l < k. \endgathered \end{equation} We first consider the case where (\ref{4-tuple1}) holds true. {\sl Step 2. An equivalent form of the non homogeneous problem $(-\Delta +\rho) u +\lambda A u=(-\Delta +\rho)f $.} For either $u,\,f \in {\mathbb E}_{l,j,i,k}$ or $u,\,f\in \tilde{{\mathbb E}}_{l,j,i,k}$, we write $$ u=\sum_{n\in {\mathbb Z}} (\xi_n,\eta_n,\zeta_n)\phi_n,\,\,\, f=\sum_{n\in {\mathbb Z}} (a_n,b_n,c_n)\phi_n, $$ where $$\phi_n =\sin(lx+jy+iz+nkz)\,\,\mbox{ if } \,\, u,\,f \in {\mathbb E}_{l,j,i,k},$$ $$\phi_n =\cos(lx+jy+iz+nkz) \,\,\mbox{ if } \,\, u,\,f\in \tilde{{\mathbb E}}_{l,j,i,k}.$$ Here $$ (\xi_n,\eta_n,\zeta_n)=(\xi_n(l,j,i,k,\lambda), \eta_n(l,j,i,k,\lambda),\zeta_n (l,j,i,k,\lambda)),$$ $$(a_n,b_n,c_n)=(a_n(l,j,i,k,\lambda), b_n(l,j,i,k,\lambda),c_n (l,j,i,k,\lambda)).$$ Let \begin{equation} \label{beta-n} \beta_n=\beta_n(l,j,i,k)=l^2+j^2+(nk+i)^2. \end{equation} Then direct computation yields that $(-\Delta +\rho) u +\lambda A u=(-\Delta +\rho)f $ can be specified as follows: \begin{gather} \sum_{n\in {\mathbb Z}}\! \left((\beta_n+\rho)(\xi_n-a_n) +\!\frac{\lambda l}{2}(\xi_{n-1}-\xi_{n+1})+\!\frac{\lambda k}{2} (\zeta_{n-1}+\zeta_{n+1})\right)\phi_n = -\partial_{x}p, \label{2.11}\\ \sum_{n\in {\mathbb Z}}\! \left((\beta_n+\rho)(\eta_n -b_n) +\!\frac{\lambda l}{2}(\eta_{n-1}-\eta_{n+1})\right)\phi_n = -\partial_{y}p, \label{2.12}\\ \sum_{n\in {\mathbb Z}}\! \left( (\beta_n+\rho)(\zeta_n-c_n) +\!\frac{\lambda l}{2}(\zeta_{n-1}-\zeta_{n+1})\right)\phi_n = -\partial_{z}p, \label{2.13}\\ \sum_{n\in {\mathbb Z}}(l\xi_n+j\eta_n+(i+nk)\zeta_n)\phi_n = 0, \label{2.14} \\ \sum_{n\in {\mathbb Z}}(la_n+jb_n+(i+nk)c_n)\phi_n = 0. \label{n2.14} \end{gather} It follows from (\ref{2.11}-\ref{2.12}) that \begin{eqnarray*} \lefteqn{ \sum_{n\in {\mathbb Z}}\! \left((\beta_n+\rho)(l\xi_n\!+\!j\eta_n-la_n-jb_n) \!+\!\frac{\lambda l}{2}(l\xi_{n-1}\!+\!j\eta_{n-1} -l\xi_{n+1}-j\eta_{n+1}) \right)\phi_n }\\ \lefteqn{+ \sum_{n\in {\mathbb Z}}\frac{\lambda lk}{2} (\zeta_{n-1}\!+\!\zeta_{n+1})\phi_n }\\ &=& -l\partial_{x}p-j\partial_{y}p. \hspace{8cm} \end{eqnarray*} Applying the operator $\partial_{z}$ to this equation and the operator $-l\partial_{x}-j\partial_{y} $ to (\ref{2.13}) respectively, and summing the resultant equations, we have \begin{eqnarray*} \lefteqn{\left((\beta_n\!+\!\rho)(l\xi_n\!+\!j\eta_n-la_n-jb_n) \!+\!\frac{\lambda l}{2}(l\xi_{n-1}\!+\!j\eta_{n-1}\!-\!l\xi_{n+1}\!-\!j\eta_{n+1}) \right)(nk+i) }\\ \lefteqn{ +\frac{\lambda lk}{2} (\zeta_{n-1}\!+\!\zeta_{n+1})(nk+i) }\\ &=& \left((\beta_n+\rho)(\zeta_n -c_n) +\!\frac{\lambda l}{2}(\zeta_{n-1}\!-\!\zeta_{n+1})\right) (l^2+j^2),\; n\in {\mathbb Z}. \hspace{2.5cm} \end{eqnarray*} Moreover, eliminating the pressure function $p$ from (\ref{2.11}-\ref{2.12}) yields \begin{align*} & \left((\beta_n+\rho)(\xi_n -a_n) +\!\frac{\lambda l}{2}(\xi_{n-1}\!-\!\xi_{n+1})+\!\frac{\lambda k}{2} (\zeta_{n-1}+\zeta_{n+1})\right)j\\ & =\left((\beta_n+\rho)(\eta_n -b_n) +\!\frac{\lambda l}{2}(\eta_{n-1}\!-\!\eta_{n+1})\right)l,\,\,n\in {\mathbb Z}. \end{align*} Thus by (\ref{2.14}-\ref{n2.14}), we see that the non homogeneous problem $$(-\Delta+\rho) u +\lambda A u =(-\Delta +\rho)f,\,\,\,\, u,\,f \in {\mathbb E}_{l,j,i,k} \mbox{ or } u,\,f \in \tilde{{\mathbb E}}_{l,j,i,k} $$ is equivalent to the following set of coupled algebraic equations \begin{align} & \frac{2\beta_n(\beta_n+\rho)}{\lambda l}\zeta_n +(\beta_{n-1}-k^2)\zeta_{n-1}-(\beta_{n+1}-k^2)\zeta_{n+1} \label{2.15} \\ & \qquad = \frac{2\beta_n(\beta_n+\rho)}{\lambda l}c_n, \nonumber \\ & \frac{2(\beta_n+\rho)}{\lambda }(j\xi_n-l\eta_n) +l(j\xi_{n-1}-l\eta_{n-1})-l(j\xi_{n+1}-l\eta_{n+1}) \label{2.16}\\ & \qquad =-kj (\zeta_{n-1}+\zeta_{n+1}) + \frac{2(\beta_n+\rho)}{\lambda }(ja_n-lb_n), \nonumber \\ & l\xi_n+j\eta_n=-(i+nk)\zeta_n, \label{2.17} \\ & l a_n+jb_n=-(i+nk)c_n, \label{n2.17} \end{align} for any $n\in {\mathbb Z}$. {\sl Step 3. Claim: $\{\xi_n\}_{n\in {\mathbb Z}}$ and $\{\eta_n\}_{n\in {\mathbb Z}}$ are uniquely determined by $\{ja_n -lb_n\}_{n\in {\mathbb Z}}$ and $\{\zeta_n\}_{n\in {\mathbb Z}}$ when $\rho>-(l^2+j^2+i^2)$. } Let $\mathcal S$ be \begin{equation} \label{ss} \mathcal S=\left\{\{\tau_n\}_{n\in {\mathbb Z}} \, |\, \sum_{n\in {\mathbb Z}}|\tau_n|^2<\infty\right\}, \end{equation} and $L:\mathcal S \to \mathcal S$ be a linear operator defined by $$L\{\tau_n\}_{n\in {\mathbb Z}} =\left\{\frac{\tau_{n-1}-\tau_{n+1}} {\beta_n+\rho}\right\}_{n\in {\mathbb Z}}.$$ Then (\ref{2.16}) can be rewritten as $$ \left( \frac{2}{\lambda l }+L\right) \{j\xi_n-l\eta_n\}_{n\in {\mathbb Z}} =-\left\{\frac{kj (\zeta_{n-1}+\zeta_{n+1})}{(\beta_n +\rho)l} \right\}_{n\in {\mathbb Z}}+ \left\{ \frac{2}{\lambda l} (ja_n-lb_n)\right\}_{n\in {\mathbb Z}}. $$ It is easy to see that $L:\mathcal S \to \mathcal S$ is a compact operator for $\rho>-\beta_0$. We now show that $-2/\lambda l$ is not an eigenvalue of $L$ or the coupled algebraic equations \begin{equation} \label{2.18} \frac{2(\beta_n+\rho)}{\lambda l}\tau_n +\tau_{n-1}-\tau_{n+1}=0,\,\,n\in {\mathbb Z} \end{equation} has no nontrivial solution $\{\tau_n\}_{n\in {\mathbb Z}} \in \mathcal S$. Otherwise, let $\{ \tau_n\}_{n\in {\mathbb Z}}$ be a nontrivial solution, then by Theorem 2.2, \begin{equation}\label{2.19} \nonumber -\displaystyle\frac{2(\beta_0+\rho)}{\lambda l} = \displaystyle\frac{1} {\displaystyle\frac{2(\beta_1+\rho)}{\lambda l} +\displaystyle\frac{1}{\displaystyle\frac{2(\beta_{2}+\rho)}{\lambda l}+ \displaystyle\frac{1}{\ddots}}} + \displaystyle\frac{1} {\displaystyle\frac{2(\beta_{-1}+\rho)}{\lambda l} +\displaystyle\frac{1}{\displaystyle\frac{2(\beta_{-2}+\rho)}{\lambda l}+ \displaystyle\frac{1}{\ddots}}}. \end{equation} Since $-(\beta_0+\rho)/(\lambda l)<0$ while the right-hand side of this equation is positive, this leads to a contradiction. Hence (\ref{2.18}) has no nontrivial solution. By the Riesz--Schauder theory, it is easy to see that $(2/\lambda l+L)^{-1}:\mathcal S \to \mathcal S$ is a bounded operator, and $$\{j\xi_n-l\eta_n\}_{n\in {\mathbb Z}}=- \big(\frac{2}{\lambda l}+ L\big)^{-1} \big(\big\{\frac{kj(\zeta_{n-1} +\zeta_{n+1})}{(\beta_n +\rho)l}\big\}_{n\in {\mathbb Z}}+ \big\{\frac{2}{\lambda l}(ja_n-lb_n)\big\}_{n\in bZ}\big). $$ This together with (\ref{2.17}-\ref{n2.17}) implies the desired assertion. {\sl Step 4. Existence and uniqueness of $\rho=\rho_{l,j,i,k}(\lambda)$ for the eigenvalue problem \begin{equation} \label{eig} \Delta u -\lambda A u =\rho u,\,\,\, u\in {\mathbb E}_{l,j,i,k}\cup \tilde{{\mathbb E}}_{l,j,i,k}. \end{equation} } It follows from (\ref{4-tuple1}) and (\ref{beta-n}) that $\beta_0-k^2<0$ and $\beta_n-k^2>0$ for $n\neq 0$. By Steps 2 and 3, (\ref{eig}) is uniquely determined by (\ref{2.15}) with $c_n=0$ or set of equations \begin{equation} \frac{2\beta_n(\beta_n+\rho)}{\lambda l}\zeta_n +(\beta_{n-1}-k^2)\zeta_{n-1}-(\beta_{n+1}-k^2)\zeta_{n+1} =0,\label{3-term} \end{equation} which is in the form of (\ref{2.5}) with $\tau_n =(\beta_n-k^2)\zeta_n$. By Theorem 2.2, it becomes to show the existence and uniqueness of $\rho=\rho_{l,j,i,k}$ in the following algebraic equation \begin{eqnarray}\label{2.20} -\displaystyle\frac{2\beta_0(\beta_0+\rho)}{\lambda l(\beta_0-k^2)} &=& \displaystyle\frac{1} {\displaystyle\frac{2\beta_1(\beta_1+\rho)}{\lambda l(\beta_1-k^2)} +\displaystyle\frac{1}{\displaystyle\frac{2\beta_2(\beta_{2}+\rho)} {\lambda l(\beta_2-k^2)}+ \displaystyle\frac{1}{\ddots}}} \\ &&+ \displaystyle\frac{1} {\displaystyle\frac{2\beta_{-1}(\beta_{-1}+\rho)}{\lambda l(\beta_{-1}-k^2)} +\displaystyle\frac{1}{\displaystyle\frac{2\beta_{-2} (\beta_{-2}+\rho)}{\lambda l (\beta_{-2}-k^2)}+ \displaystyle\frac{1}{\ddots}}}. \nonumber \end{eqnarray} Multiplying this equation by $-l(\beta_0-k^2)(\beta_0(\beta_0+\rho))^{-1}$ gives \begin{equation}\label{2.22} \frac{2}{\lambda }= \displaystyle\frac{1}{\displaystyle\frac{g_1(\rho)}{\lambda} + \displaystyle\frac{1}{\displaystyle\frac{g_2(\rho)}{\lambda} + \displaystyle\frac{1}{\ddots}}} +\displaystyle\frac{1}{\displaystyle\frac{g_{-1}(\rho)}{\lambda} + \displaystyle\frac{1}{\displaystyle\frac{g_{-2}(\rho)}{\lambda} + \displaystyle\frac{1}{\ddots}}}, \end{equation} where \begin{equation} \nonumber g_{\pm n}(\rho)= \left\{ \begin{split} & \frac{2\beta_{\pm n}\beta_0(\beta_{\pm n}+\rho)(\beta_0+\rho)} {l^2(k^2-\beta_0) (\beta_{\pm n}-k^2)},\,\,\mbox{ if $n$ is odd},\\ & \frac{2\beta_{\pm n}(\beta_{\pm n}+\rho)(k^2-\beta_0)} {\beta_0(\beta_{\pm n}-k^2)(\beta_{0}+\rho)},\,\,\mbox{ if $n$ is even}. \end{split} \right. \end{equation} Let $G(\lambda,\rho)$ be the right-hand side of (\ref{2.22}), and for $n\geq 1$ we set $$G_{\pm n}(\lambda,\rho)= \displaystyle\frac{1}{\displaystyle\frac{g_{\pm 1}(\rho)}{\lambda} + \displaystyle\frac{1}{\displaystyle\frac{g_{\pm 2}(\rho)}{\lambda} + \displaystyle\frac{1}{\ddots+ \displaystyle\frac{1}{\displaystyle\frac{g_{\pm n}(\rho)}{\lambda}}}}}.$$ We have $$G(\lambda, \rho)=\lim_{n\to\infty}G_{n}(\lambda,\rho)+ \lim_{n\to\infty}G_{-n}(\lambda,\rho) . $$ For any $\lambda >0$, $\epsilon >0$ and $M > 0$ arbitrarily large, we see that $\{G_{n}(\lambda, \cdot)\}_{n\geq 1}$ and $\{G_{-n}(\lambda, \cdot)\}_{n\geq 1}$ are Cauchy sequences of the Banach space $C^1([-\beta_0+\epsilon, M])$, and thus $G(\lambda,\cdot) \in C^1([-\beta+\epsilon, M])$, \begin{eqnarray*} \frac{\partial G(\lambda,\rho)}{\partial \rho} &=&\lim_{n\to \infty}\frac{\partial G_{n}(\lambda,\rho)}{\partial \rho} + \lim_{n\to \infty}\frac{\partial G_{-n}(\lambda,\rho)} {\partial \rho} \\ &=& \sum_{n=1}^\infty \frac{(-1)^n}{\lambda} \frac{\mbox{d}g_{n}(\rho)}{\mbox{d}\rho} \hat{G}_1^2(\lambda,\rho) \cdots \hat{G}_n^2(\lambda,\rho) \\ & & + \sum_{n=1}^\infty \frac{(-1)^n}{\lambda} \frac{\mbox{d}g_{-n}(\rho)}{\mbox{d}\rho} \hat{G}_{-1}^2(\lambda,\rho) \cdots \hat{G}_{-n}^2(\lambda,\rho), \end{eqnarray*} with $$\hat{G}_{\pm n}(\lambda,\rho)= \displaystyle\frac{1}{\displaystyle\frac{g_{\pm n}(\rho)}{\lambda} + \displaystyle\frac{1}{\displaystyle\frac{g_{\pm (n+1)}(\rho)}{\lambda} + \displaystyle\frac{1}{\ddots}}}. $$ By direct calculation, we infer from (\ref{4-tuple1}) that for any $n \ge 1$, ${\mbox{d}g_{\pm n}(\rho)}/{\mbox{d}\rho} >0$ when $n$ is odd, and ${\mbox{d}g_{\pm n} (\rho)}/ {\mbox{d}\rho}<0$ when $n$ is even. Therefore when $\rho >- \beta_0,$ we obtain \begin{equation}\label{2.23} \frac{\partial G(\lambda,\rho)}{\partial\rho} <0. \end{equation} Hence the observation $$\lim_{\rho\searrow -\beta_0}G(\lambda,\rho)= \infty\mbox{ and } \lim_{\rho\rightarrow \infty}G(\lambda,\rho)= 0$$ implies the existence and uniqueness of $\rho=\rho_{l,j,i,k}(\lambda)>- \beta_0$ satisfying \begin{equation} \frac{2}{\lambda }=G(\lambda, \rho_{l,j,i,k}(\lambda)).\label{2.24} \end{equation} {\sl Step 5.} Notice that $$ \lambda G(\lambda,\rho)= \displaystyle\frac{1}{ \displaystyle\frac{g_1(\rho)}{\lambda^2} + \displaystyle\frac{1}{g_2(\rho) +\displaystyle\frac{1}{\displaystyle\frac{g_3(\rho)}{\lambda^2} + \displaystyle\frac{1}{\ddots}}} } +\displaystyle\frac{1}{ \displaystyle\frac{g_{-1}(\rho)}{\lambda^2} + \displaystyle\frac{1}{g_{-2}(\rho) +\displaystyle\frac{1}{\displaystyle\frac{g_{-3}(\rho)}{\lambda^2} + \displaystyle\frac{1}{\ddots}}} }. $$ Arguing as in the derivation of (\ref{2.23}), we have $$\frac{\partial (\lambda G(\lambda,\rho))}{\partial \lambda } >0\,\ \mbox{ for } \lambda >0 \,\mbox{ and } \rho>-\beta_0.$$ (\ref{2.24}) gives for $\rho=\rho_{l,j,i,k}(\lambda)$, \begin{eqnarray*} 0&=&\frac{\mbox{d}(\lambda G(\lambda,\rho))} {\mbox{d}\lambda}\\ &=&\frac{\partial (\lambda G(\lambda,\rho))}{\partial\lambda}+ \lambda\frac{\partial G(\lambda,\rho)}{\partial\rho} \frac{\mbox{d}\rho}{\mbox{d}\lambda} \\ &>&\lambda\frac{\partial G(\lambda,\rho)} {\partial\rho}\frac{\mbox{d}\rho}{\mbox{d}\lambda}. \end{eqnarray*} This shows, by (\ref{2.23}), that $\mbox{d}\rho_{l,j,i,k}/\mbox{d}\lambda>0.$ Furthermore, multiplying (\ref{2.20}) by $\lambda$ gives \begin{eqnarray} \lefteqn{ -\frac{2\beta_0(\beta_0+\rho_{l,j,i,k}(\lambda))} { l(\beta_0-k^2)} }\\ &=& \displaystyle\frac{1} {\displaystyle\frac{2\beta_1(\beta_1+\rho_{l,j,i,k}(\lambda))} {\lambda^2l(\beta_1-k^2)} +\displaystyle\frac{1}{\displaystyle\frac{2\beta_2(\beta_{2}+ \rho_{l,j,i,k}(\lambda))} {l(\beta_2-k^2)}+ \displaystyle\frac{1}{\ddots}}}\nonumber \\ &&+ \displaystyle\frac{1} {\displaystyle\frac{2\beta_{-1}(\beta_{-1}+\rho_{l,j,i,k}(\lambda))} {\lambda^2 l(\beta_{-1}-k^2)} +\displaystyle\frac{1}{\displaystyle\frac{2\beta_{-2}(\beta_{-2} +\rho_{l,j,i,k}(\lambda))}{l (\beta_{-2}-k^2)}+ \displaystyle\frac{1}{\ddots}}}. \nonumber \end{eqnarray} Passing to the limits $\lambda\rightarrow 0^+$ and $\lambda\rightarrow \infty$ respectively, we obtain immediately \begin{equation} \nonumber \begin{split} & \lim_{\lambda\rightarrow 0^+}\rho_{l,j,i,k}(\lambda)=-l^2-j^2-i^2, \\ & \lim_{\lambda\rightarrow \infty}\rho_{l,j,i,k}(\lambda)=\infty. \end{split} \end{equation} In addition, it follows from (\ref{2.15}) with $\rho=\rho_{l,j,i,k}(\lambda)$ and Theorem 2.2 that $$\gamma_{\pm n} =\frac{(\beta_{\pm n}-k^2)\zeta_{\pm n}} {(\beta_{\pm (n-1)}-k^2)\zeta_{\pm(n-1)}} \,\, \mbox{ for } n\in N$$ with $\zeta_m$ subject to the following condition $$ \begin{array}{lcll} \zeta_0&=&c,&\vspace{2mm}\\ \zeta_{\pm n}&=&c\displaystyle\frac{(\beta_0-k^2) \gamma_{\pm 1}\cdots\gamma_{\pm n}}{\beta_{\pm n}-k^2}, &n\in N. \end{array} $$ Here $c$ is an arbitrary constant. Notice that $\gamma_{\pm n}$ are uniquely determined by $\rho_{l, i, j, k}(\lambda)$ and $\beta_0, \beta_{\pm 1}, \cdots$. Thus all the eigenfunctions of the spectral problem $\Delta u -\lambda A u=\rho_{l,j,i,k}u$ form a one-dimensional subspace in ${\mathbb E}_{l,j,i,k}$ and $\tilde{{\mathbb E}}_{l,j,i,k}$ respectively. {\sl Step 6. Claim: ker$(\Delta-\lambda A-\rho)^m =\mbox{ker}(\Delta-\lambda A-\rho)$ $(m>1,\,\, \rho >-\beta_0).$} It suffices to consider the operator with $\rho=\rho_{l,j,i,k}$ in the space ${\mathbb E}_{l,j,i,k}$. We start with the case $m=2$. Let $$u=\sum_{n\in {\mathbb Z}}(\xi_n,\eta_n,\zeta_n) \sin(lx+jy+iz+nkz)\in {\mathbb E}_{l,j,i,k}$$ such that $$(\Delta-\lambda A -\rho)^2u=0,$$ or equivalently, $$(1 -\lambda (\Delta-\rho)^{-1}A )^2u=0.$$ This implies the existence of $$u'=\sum_{n\in {\mathbb Z}}(\xi'_n,\eta'_n,\zeta'_n) \sin(lx+jy+iz+nkz)\in {\mathbb E}_{l,j,i,k}$$ such that \begin{equation}\label{ge} \gathered (-\Delta +\rho)u +\lambda A u = (-\Delta +\rho) u', \\ (-\Delta +\rho)u'+\lambda Au' = 0. \endgathered \end{equation} Thanks to (\ref{2.15}-\ref{n2.17}), the above system (\ref{ge}) is equivalent to the following set of coupled algebraic equations: \begin{align} & \frac{2\beta_n(\beta_n+\rho)}{\lambda l}\zeta_n +(\beta_{n-1}-k^2)\zeta_{n-1}-(\beta_{n+1}-k^2)\zeta_{n+1} \label{n2.15} \\ & \qquad = \frac{2\beta_n(\beta_n+\rho)}{\lambda l}\zeta'_n,\nonumber \\ & \frac{2(\beta_n+\rho)}{\lambda }(j\xi_n-l\eta_n) +l(j\xi_{n-1}-l\eta_{n-1})-l(j\xi_{n+1}-l\eta_{n+1}) \\ & \qquad =-kj(\zeta_{n-1}+\zeta_{n+1}) +\frac{2(\beta_n+\rho)}{\lambda }(j\xi'_n-l\eta'_n) , \nonumber \\ & l\xi_n+j\eta_n=-(i+nk)\zeta_n, \\ & \frac{2\beta_n(\beta_n+\rho)}{\lambda l}\zeta'_n +(\beta_{n-1}-k^2)\zeta'_{n-1}-(\beta_{n+1}-k^2)\zeta'_{n+1} = 0,\label{nn2.15} \\ & \label{nn2.16} \frac{2(\beta_n+\rho)}{\lambda }(j\xi'_n-l\eta'_n) +l(j\xi'_{n-1}-l\eta'_{n-1})-l(j\xi'_{n+1}-l\eta'_{n+1}) \\ & \qquad =-kj(\zeta'_{n-1}+\zeta'_{n+1}), \nonumber \\ & \label{nn2.17} l\xi_n'+j\eta_n'=-(i+nk)\zeta_n', \end{align} for any $n\in {\mathbb Z}$. Thus it amounts to showing that $\zeta'_n=\eta'_n=\xi'_n =0$ for any $n\in {\mathbb Z}$. Indeed, (\ref{n2.15}) gives the following equation in $\mathcal S$ defined by (\ref{ss}) \begin{equation}\label{ttt} (1+T)\{(\beta_n-k^2)\zeta_n\}_{n\in {\mathbb Z}} =\{(\beta_n-k^2)\zeta'_n\}_{n\in {\mathbb Z}}, \end{equation} with the operator $T$ in the form $$ T\{\tau _n\}_{n\in {\mathbb Z}} =\left\{ \frac{\lambda l(\beta_n-k^2) }{2\beta_n(\beta_n+\rho)}(\tau_{n-1}-\tau_{n+1})\right\}_ {n\in {\mathbb Z}}.$$ Since $T: \mathcal S\mapsto \mathcal S$ is compact, by the Riesz-Schauder theory, (\ref{ttt}) is solvable if and only if $$ \sum_{n\in {\mathbb Z}}(\beta_n-k^2)\zeta'_n \tau_n=0\mbox{ for all } \{\tau_n\}_{n\in{\mathbb Z}}\in \mathcal S \mbox{ with } (1+T^*)\{\tau_n\}_{n\in {\mathbb Z}}=0, $$ where $T^*$ denotes the dual operator of $T$. To specify the kernel of $1+T^*$, we note that $$T^*\{\tau _n\}_{n\in {\mathbb Z}} =\left\{ -\frac{\lambda l(\beta_{n-1}-k^2)} {2\beta_{n-1}(\beta_{n-1}+\rho)}\tau_{n-1} +\frac{\lambda l(\beta_{n+1}-k^2)} {2\beta_{n+1}(\beta_{n+1}+\rho)}\tau_{n+1} \right\}_ {n\in {\mathbb Z}}.$$ Thus $(1+T^*)\{\tau_n\}_{n\in {\mathbb Z}}=0$ becomes, for $n\in {\mathbb Z}$, $$\frac{2\beta_n(\beta_n+\rho)}{\lambda l(\beta_n-k^2)} \sigma_n +\sigma_{n-1}-\sigma_{n+1}=0,\,\,\,\, \sigma_m= (-1)^m\frac{\lambda l(\beta_{m}-k^2)} {2\beta_{m}(\beta_{m}+\rho)}\tau_{m}. $$ or $\{\sigma_n\}_{n\in{\mathbb Z}}\in \mbox{ker}(1+T)$. If $\{\zeta'_n\}_{n\in{\mathbb Z}}\neq 0$, applying Theorem 2.2 to (\ref{nn2.15}) gives $\zeta'_n\neq 0$ for all $n\in {\mathbb Z}$ and $$\{\sigma_n\}_{n\in {\mathbb Z}}= \left\{ (-1)^n\frac{\lambda l(\beta_{n}-k^2)} {2\beta_{n}(\beta_{n}+\rho)}\tau_{n}\right\}_{n\in {\mathbb Z}} =c\{(\beta_n-k^2)\zeta'_n\}_{n\in{\mathbb Z}} $$ for some constant $c\neq 0$, whenever $\{\tau_n\}_{n\in {\mathbb Z}}\neq 0$. This implies \begin{equation}\label{ab} \sum_{n\in{\mathbb Z}}(\beta_n-k^2)\zeta'_n\tau_n =c\sum_{n\in{\mathbb Z}}(-1)^n \frac{2\beta_{n}(\beta_{n}+\rho)(\beta_n-k^2)}{\lambda l}{\zeta'}^2_n. \end{equation} On the other hand, multiplying by $(\beta_n-k^2)\zeta'_n$ the $n$th equation of (\ref{nn2.15}) and summing up the resultant equations yield $$\sum_{n\in{\mathbb Z}} \frac{2\beta_{n}(\beta_{n}+\rho)(\beta_n-k^2)}{\lambda l}{\zeta'}^2_n=0. $$ This together with (\ref{ab}) implies $$\sum_{n\in{\mathbb Z}}(\beta_n-k^2)\zeta'_n\tau_n =-c\sum_{n\in{\mathbb Z}} \frac{2\beta_{2n+1}(\beta_{2n+1}+\rho)(\beta_{2n+1}-k^2) }{\lambda l}{\zeta'}^2_{2n+1} \neq 0. $$ This leads to a contradiction, and thus $\{\zeta'_n\}_{n\in{\mathbb Z}}=0$. From (\ref{nn2.16}-\ref{nn2.17}) and Step 3, we have $\{\eta'_n\}_{n\in{\mathbb Z}}=0$ and $\{\xi'_n\}_{n\in{\mathbb Z}}=0$. Hence we have $u'=0$ or $$\mbox{ker}(\Delta-\lambda A -\rho)^2 =\mbox{ker}(\Delta-\lambda A-\rho).$$ To reach the case where $m>2$, suppose $(1+\lambda(-\Delta+\rho)^{-1} A)^mu=0$ for some $u\in {\mathbb E}_{l,j,i,k}$. Then there exits a $u'\in {\mathbb E}_{l,j,i,k}$ so that \begin{eqnarray*} (-\Delta +\rho)v +\lambda A v&=&(-\Delta +\rho)u',\\ (\Delta -\lambda A -\rho)u'&=&0, \\ (1+\lambda (-\Delta+\rho)^{-1}A )^{m-2}u&=&v. \end{eqnarray*} It follows from the argument for the case $m=2$ that $u'=0$. Hence, we obtain the case for $m>2$ by induction. {\sl Step 7.} We now consider the case where (\ref{4-tuple2}) holds true. Obviously, it follows from the Steps 2 and 3 that spectral problem $\Delta u -\lambda A u=\rho u$ with $u\in {\mathbb E}_{l,j,i,k}\cup\tilde{{\mathbb E}}_{l,j,i,k}$ is uniquely determined by the system $$ \frac{2\beta_{-n-1}(\beta_{-n-1}+\rho)}{\lambda l}\zeta_{-n-1} +(\beta_{-n-2}-k^2)\zeta_{-n-2}-(\beta_{-n}-k^2)\zeta_{-n} =0, \,\,\,\mbox{ for } n\in {\mathbb Z}$$ with $\beta_n=\beta_n(l,j,i,k)$. Setting $\beta'_n=\beta_{n}(l,j,k-i,k)$ and observing that \\ $\beta_{-n-1}(l,j,i,k)=\beta_{n}(l,j,k-i,k)$, we see that $$ \frac{2\beta'_{n}(\beta'_{n}+\rho)}{\lambda l}\zeta_{-n-1} +(\beta'_{n+1}-k^2)\zeta_{-n-2}-(\beta'_{n-1}-k^2)\zeta_{-n} =0, \,\,\,\mbox{ for } n\in {\mathbb Z}.$$ By setting $\zeta'_n =(-1)^n\zeta_{-n-1}$, we have $$\frac{2\beta'_{n}(\beta'_{n}+\rho)}{\lambda l}\zeta'_{n} +(\beta'_{n-1}-k^2)\zeta'_{n-1}-(\beta'_{n+1}-k^2)\zeta'_{n+1} =0, \,\,\,\mbox{ for } n\in {\mathbb Z}.$$ Thus the above argument implies the unique existence of the eigenvalue $\rho_{l,j,i,k}(\lambda)$ so that $\Delta u -\lambda A u=\rho_{l,j,i,k}u$ with $k/2 k^2, \quad \forall n\geq 2,$$ and $$l^2+j^2+(k-i)^2\geq k^2 \quad \mbox{\rm implies }\quad n^2(l^2+j^2)+(k-\{ni\})^2> k^2, \quad \forall n\geq 2.$$ \end{Lemma} \begin{proof} Notice that the desired conclusion holds true if $n^2(l^2+j^2)>k^2$. It suffices to consider the integers $n$ subject to the condition $ n^2(l^2+j^2)\leq k^2$. In the case where $l^2+j^2+i^2\geq k^2$, we have $$i^2\geq k^2-l^2-j^2\geq \frac{(n^2-1)k^2}{n^2}> \frac{(n-1)^2k^2}{n^2},$$ which yields $$ nk> ni =(n-1)k + \{ni\} > (n-1)k.$$ Namely $\{ni\}=ni-(n-1)k>0$. Hence \begin{eqnarray*} n^2(l^2+j^2)+\{ni\}^2&=&n^2(l^2+j^2)+(ni-(n-1)k)^2\\ &=& n^2(l^2+j^2+i^2)+n^2k^2-2nk^2-2n(n-1)ki+k^2\\ &\geq & 2n(n-1)k(k-i)+k^2>k^2. \end{eqnarray*} For the second case where $l^2+j^2+(k-i)^2\geq k^2$, we have $$(k-i)^2\geq k^2-l^2-j^2\geq \frac{(n^2-1)k^2}{n^2}>\frac{(n-1)^2k^2}{n^2},$$ which shows $\{ni\}=ni 0$, $n\in {\mathbb N}$ with $n>n_0$ and $\rho >-(l^2+j^2+\min\{i^2,(k-i)^2\})$, \begin{equation} \nonumber \begin{split} & \mbox{\rm dim}\{u\in {\mathbb E}_{nl,nj,\{ni\},k} \cup \tilde{{\mathbb E}}_{nl,nj,\{ni\},k} |\,\,\Delta u -\lambda A u-\rho u=0\}=0,\\ & \mbox{\rm dim}\{u\in {\mathbb E}_{nl,nj,k- \{ni\},k} \cup\tilde{{\mathbb E}}_{nl,nj,k- \{ni\},k} |\,\,\Delta u -\lambda A u-\rho u=0\}=0, \text{ if } \{ni\}\neq 0. \end{split} \end{equation} (ii). There exist $n_0 $ eigenvalues $\rho_{nl,nj,\{ni\},k}:(0,\infty)\rightarrow {\mathbb R}$ for $n=1,...,n_0$ such that \begin{equation}\nonumber \begin{split} & \frac{d\rho_{nl,nj,\{ni\},k}(\lambda)}{d\lambda} >0, \\ & \lim_{\lambda\rightarrow \infty}\rho_{nl,nj,\{ni\},k} (\lambda)=\infty, \\ & \lim_{\lambda\rightarrow 0^+}\rho_{nl,nj,\{ni\},k}(\lambda)= -n^2(l^2+j^2)-\min\{\{ni\}^2,(k-\{ni\})^2\}, \end{split} \end{equation} and \begin{equation}\nonumber \begin{split} & \mbox{\rm dim}\bigcup_{m\in {\mathbb N}} \{u\in {\mathbb E}_{nl,nj,\{ni\},k} |\,\,(\Delta -\lambda A -\rho)^m u=0\}\leq 1, \\ & \mbox{\rm dim}\bigcup_{m\in{\mathbb N}} \{u\in \tilde{{\mathbb E}}_{nl,nj,\{ni\},k}|\,\,(\Delta -\lambda A - \rho )^mu=0\}\leq 1, \\ & \mbox{\rm dim}\bigcup_{m\in{\mathbb N}} \{u\in {\mathbb E}_{nl,nj,k-\{ni\},k} |\,\,(\Delta -\lambda A -\rho)^m u=0\}\leq 1 \,\mbox{\rm if } \,\{ni\} \neq 0, \\ & \mbox{\rm dim}\bigcup_{m\in{\mathbb N}} \{u\in \tilde{{\mathbb E}}_{nl,nj,k-\{ni\},k}|\,\, (\Delta -\lambda A - \rho )^mu=0\}\leq 1 \,\,\mbox{\rm if } \,\{ni\}\neq 0, \end{split} \end{equation} where the equalities hold if and only if $\rho=\rho_{nl,nj,\{ni\},k}(\lambda)$ for $\lambda >0$. \end{Theorem} \begin{proof} (i). Let $n\geq n_0+1$. Without loss of generality, we may assume that $\{ni\} n^2(l^2+ j^2)+\{ni\}^2-k^2 \ge 0. $$ As in the proof of Theorem 2.1, for $u\in {\mathbb E}_{nl,nj,\{ni\},k}\cup\tilde{{\mathbb E}}_{nl,nj,\{ni\},k},$ the spectral problem $\Delta u -\lambda A u-\rho u=0 $, is uniquely determined by the following analog of (\ref{2.15}): \begin{equation} \label{2.26} \frac{2\beta_m(\beta_m+\rho)}{\lambda n l}\zeta_m +(\beta_{m-1}-k^2)\zeta_{m-1}-(\beta_{m+1}-k^2)\zeta_{m+1} =0,\,\,m\in {\mathbb Z}, \end{equation} where $\beta_m$ now equals $\beta_m(nl,nj,\{ni\},k)=n^2(l^2+j^2)+(mk+\{ni\})^2$. Multiplying the $m$--th equation of (\ref{2.26}) by $(\beta_m-k^2)\zeta_m$ and summing the resultant equations yield $$ \sum_{m\in {\mathbb Z}} \frac{2\beta_m(\beta_m+\rho)(\beta_m-k^2)}{\lambda n l}|\zeta_m|^2 =0 . $$ Notice that $\beta_m-k^2>0$ for $m\neq 0$. Thus $\zeta_m\equiv 0$ for $m\neq 0$, and $\zeta_0=0$ as well in view of (\ref{2.26}). In the case where $\{ni\}\neq 0$ and $u\in{\mathbb E}_{nl,nj,k-\{ni\},k}\cup\tilde{{\mathbb E}}_{nl,nj,k-\{ni\},k},$ the spectral problem $\Delta u -\lambda A u-\rho u=0 $ is uniquely determined by \begin{equation} \nonumber \frac{2\beta_m(\beta_m+\rho)}{\lambda n l}\zeta_m +(\beta_{m-1}-k^2)\zeta_{m-1}-(\beta_{m+1}-k^2)\zeta_{m+1} =0,\,\,m\in {\mathbb Z}, \end{equation} with $\beta_m(nl,nj,k-\{ni\},k)=n^2(l^2+j^2)+(mk+k-\{ni\})^2$. Then as before, it is easy to see that $\zeta_m\equiv 0$. (ii). For $ 1\leq n\leq n_0$, it follows from Lemma 2.1 and the definition of the integer $n_0$ that $(nl,nj,\{ni\},k)$ satisfies (\ref{4-tuple}). Thus Assertion (ii) is a direct consequence of Theorem 2.1. \end{proof} \setcounter{equation}{0} \section{Invariant Subspaces and Properties of the Steady-State Solutions} In the previous section, we found that the eigenspaces of the spectral problem $\Delta u -\lambda Au=\rho u$ in $\dbH^2_\sigma$ are even-dimensional. Thus it will be difficult to obtain steady-state bifurcations in the whole space $\dbH^2_\sigma$. Fortunately, (\ref{ns1}-\ref{ns4}) admit many flow invariant subspaces, in which it will be convenient to examine the bifurcation phenomena. First, by the divergence free condition $\mbox{dvi} u =0$, we observe that the bilinear term $B(u,u)=P[u\cdot \nabla u]$ can be alternatively given by \begin{eqnarray*} B(u,u)&=&u\cdot\nabla u- \nabla \Delta^{-1}( \partial_x(u\cdot\nabla u_1) +\partial_y(u\cdot\nabla u_2) +\partial_z(u\cdot\nabla u_1) ) \\ &=& \nabla\cdot (u\otimes u)-\nabla \Delta^{-1}\nabla^2\cdot (u\otimes u) \end{eqnarray*} with $$\nabla\cdot (u\otimes u) =\left( \sum_{n=1}^3\partial_n(u_n u_1), \sum_{n=1}^3\partial_n(u_n u_2), \sum_{n=1}^3\partial_n(u_n u_3)\right)$$ and $$\nabla^2\cdot (u\otimes u) = \sum_{n=1}^3\sum_{m=1}^3\partial_n\partial_m(u_n u_m)$$ for $(\partial_1,\partial_2,\partial_3)= (\partial_x,\partial_y,\partial_z)$. Let the integers $l\geq 0$, $j\in {\mathbb Z}$ and $0\leq i0$ such that problem (\ref{2.1}) admits two different global branches of steady-state solutions $ u^1_{l,j,i,k,\lambda},\,\, u^2_{l,j,i,k,\lambda}\in \dbH^2_{n_0l,n_0j,\{n_0i\},k}\subset \dbH^2_{l,j,i,k}\subset \dbH^2_\sigma $ branching off the bifurcation point $(\lambda_{l,j,i,k}, u_0)$ continuously such that for $m=1,\,2$, \begin{enumerate} \item each branch $u^m_{l,j,i,k,\lambda}$ extends to $\lambda=\infty$ in $\mathcal C_1 \cap \mathcal C_2$ as shown in Figures 1.1 and 1.2; \item each branch $u^m_{l,j,i,k,\lambda}$ starting from the bifurcation point $(\lambda_{l,j,i,k},u_0)$ never touches the $\lambda$--axis for $\lambda \neq \lambda_{l,j,i,k}$. \end{enumerate} \label{Th4.1} \end{Theorem} \begin{proof} For the integer $n_0$ given in (\ref{n-zero}), we see that $\dbH^2_{l,j,i,k}\supset \dbH^2_{n_0l,n_0j,\{n_0i\},k}$, which is in the following form \begin{eqnarray*} &&\left\{u\in \dbH^2_\sigma\,|\,u= \sum_{n\in {\mathbb N}}(\xi_{n,1},\xi_{n,2},\xi_{n,3})\sin nkz \right.\\ &&+\sum_{m\geq 2}\sum_{n\in {\mathbb Z}} (\eta_{m,n,1},\eta_{m,n,2},\eta_{m,n,3}) \sin (mn_0lx+mn_0jy+\{mn_0i\}z+knz)\\ &&+\left.\sum_{n\in {\mathbb Z}} (\zeta_{n,1},\zeta_{n,2},\zeta_{n,3}) \sin (n_0lx+n_0jy+\{n_0i\}z+knz)\right\}. \end{eqnarray*} It is obvious that the spectral problem \begin{equation} \label{4.2} \Delta u -\lambda A u=\rho u\,\quad \mbox{ for }\, \rho >-\beta_0 \mbox{ and } u\in \dbH^2_{n_0l,n_0j,\{n_0i\},k} \end{equation} has no eigenfunction in the form $$\sum_{n\in {\mathbb N}}(\xi_{n,1},\xi_{n,2},\xi_{n,3})\sin nkz. $$ Thus it follows from Theorem 2.3 that (\ref{4.2}) has a unique eigenvalue $\rho_{n_0l,n_0j,\{n_0i\},k}$ transversal across the imaginary axis at the origin, and the choice of the integer $n_0$ ensures the simplicity of this eigenvalue. Let us denote by $\lambda_{l,j,i,k}$ the critical Reynolds number such that \begin{equation} \label{4.3} \rho_{n_0l,n_0j,\{n_0i\},k}(\lambda_{l,j,i,k})=0. \end{equation} Taking $ u'+ u_0$ in place of $ u$ in the steady-state equations with respect to (\ref{2.1}) reduced in $\dbH^2_{n_0l,n_0j,\{n_0i\},k}$ and omitting the prime, we have \begin{equation}\label{4.4} u =\lambda\Delta^{-1} A u +\lambda\Delta^{-1} B( u),\quad u\in \dbH^2_{n_0l,n_0j,\{n_0i\},k}, \end{equation} of which the linear part \begin{equation} \label{4.5} u -\lambda\Delta^{-1} A u=0,\quad u\in \dbH^2_{n_0l,n_0j,\{n_0i\},k} \end{equation} has a simple eigenvalue $1/\lambda_{l,j,i,k}$ or a simple characteristic value $\lambda_{l,j,i,k}$ in the notation of Rabinowitz \cite{Rab}. It follows from \cite[Theorem 1.40]{Rab} that (\ref{4.4}) admit two continuous branches of solutions $(\lambda,u_{l,j,i,k,\lambda}^1-u_0)$ and $(\lambda, u_{l,j,i,k,\lambda}^2-u_0)$ other than the branch $(\lambda, \,0)$ in the space $\mathbb R \times \dbH^2_{n_0l,n_0j,\{n_0i\},k}$ under the norm $\|(\lambda,u)\|= (\lambda^2+\|\Delta u\|_{L^2})^{1/2}$. Each of those two branches meets $(\lambda_{l,j,i,k},\,0)$ and either \begin{enumerate} \item[(i)] meets $\infty $ in $\mathbb R\times \dbH^2_{n_0l,n_0j,\{n_0i\},k}$, or \item[(ii)] meets $(\tilde\lambda,\,0)$, where $\tilde\lambda\neq \lambda_{l,j,i,k}$ is a characteristic value of (\ref{4.5}). \end{enumerate} Observe that $(\lambda,0)$ is the unique solution of (\ref{4.4}) when $0 < \lambda < {c}/{k^2}$. Thus these two branches of solutions are contained in the space $(0,\infty)\times \dbH^2_{n_0l,n_0j,\{n_0i\},k}$. Meanwhile, the {\it a priori} bounds of the the steady-state solutions of (\ref{2.1}) for every $\lambda >0$ show that both branches of the bifurcation solutions extends to $\lambda=\infty $ in $\mathcal C_1 \cap \mathcal C_2$; see Figures 1.1 and 1.2. Furthermore, the above second alternative is excluded due to the fact that $\lambda_{l,j,i,k}$ is the only characteristic value of (\ref{4.5}) for $\lambda \in (0,\infty)$. Hence, each branch $u^m_{l,j,i,k,\lambda}$ intersects with the $\lambda$--axis only at the line segment $pq$ shown in Figures 1.1 and 1.2. In particular, each branch $u^m_{l,j,i,k,\lambda}$ starting from the bifurcation point $(\lambda_{l,j,i,k},u_0)$ never touches the $\lambda$--axis for $\lambda \neq \lambda_{l,j,i,k}$. \end{proof} \setcounter{equation}{0} \section{Supercritical Pitchfork Bifurcation} In order to reach our main results, it is necessary to give the local behavior of the global branches of the steady-state solutions $u^1_{l,j,i,k,\lambda}$ and $u^2_{l,j,i,k,\lambda}$ close to the bifurcation point $(\lambda_{l,j,i,k},u_0)$. More precisely, we have the following supercritical pitchfork bifurcation theorem. \begin{Theorem} \label{Th5.1} Let the steady-state solutions $u^1_{l,j,i,k,\lambda}$, $u^2_{l,j,i,k,\lambda}\in \dbH^2_{n_0l,n_0j,\{n_0i\},k}$ be given in Theorem \ref{Th4.1}. Then for $m=1,$ $2$, \begin{equation} \label{t5.1} \begin{split} & u^m_{l,j,i,k,\lambda}= u_0 \qquad \qquad \hspace{25mm} \mbox{ for } \lambda <\lambda_{l,j,i,k}, \\ & u_0\neq u^1_{l,j,i,k,\lambda}\neq u^2_{l,j,i,k,\lambda} \neq u_0\qquad \mbox{ for }\lambda_{l,j,i,k}<\lambda, \end{split} \end{equation} provided that \begin{equation}\label{t5.2} \|\Delta(u^m_{l,j,i,k\lambda}-u_0)\|_{L^2}+|\lambda-\lambda_{l,j,i,k}| <\epsilon \end{equation} for some small constant $\epsilon >0$. \end{Theorem} For simplicity, we set \begin{equation}\label{t5.3} \lambda_0=\lambda_{l,j,i,k},\,\,\,\, u^+_\lambda=u^1_{l,j,i,k,\lambda},\,\,\,\, u^-_\lambda =u^2_{l,j,i,k,\lambda}. \end{equation} The proof of Theorem 5.1 will be achieved by analyzing the local asymptotic expansion of $u^\pm_\lambda$ in terms of $\lambda$. First, notice that he bifurcation phenomenon of the Navier-Stokes system (\ref{2.1}) reduced to the subspace $\dbH^2_{n_0l,n_0j,\{n_0i\},k}$ is excited by the external force $k^2u_0$ and the unstable subspace of the bifurcation point $(\lambda_0,u_0)$ is contained in $ {\mathbb E}_{n_0l,n_0j,\{n_0i\},k} $. The unstable subspace is given by the nontrivial solution $u^* \in {\mathbb E}_{n_0l,n_0j,\{n_0i\},k}$ of the spectral problem $\Delta u^*-\lambda_0Au^*=0$: \begin{equation}\label{u*} \gathered u^*=\sum_{n\in {\mathbb Z}} (\xi_n,\eta_n,\zeta_n)\sin(n_0lx+n_0jy+\{n_0i\}z+nkz) \in {\mathbb E}_{n_0l,n_0j,\{n_0i\},k}, \\ \zeta_0 =1, \\ \Delta u^*-\lambda_0Au^*=0, \endgathered \end{equation} By Theorem 2.3, for any $n$, $\zeta_n \not=0$. Hence there exists a unique nontrivial solution $u^*$ of (\ref{u*}) with $\zeta_0 =1.$ Then we examine the nonlinear interaction of the unstable direction given by $u^*$. To this end, we notice that $$B(u^*,u^*) \in {\mathbb E}_{0,0,0,k}\oplus {\mathbb E}_{2n_0l,2n_0j,\{2n_0i\},k}.$$ Hence, we can decompose $B(u^*,u^*)$ as \begin{equation}\label{interaction} \begin{split} & B(u^*,u^*)= B_0(u^*,u^*)+ B_2(u,u)+B_3(u,u), \\ & B_0(u^*,u^*)=\frac{u_0}{4\pi^3} \int_{{\mathbb T}^3} B(u^*,u^*)\cdot u_0\,dx\,dy\,dz , \\ & B_2(u^*,u^*) \in {\mathbb E}_{0,0,0,k}\,\,\mbox{ with }\,\, \int_{{\mathbb T}^3} B_2(u^*,u^*)\cdot u_0\,dx\,dy\,dz =0, \\ & B_3(u^*,u^*) \in {\mathbb E}_{2n_0l,2n_0j,\{2n_0i\},k}. \end{split} \end{equation} To prove Theorem 5.1, it suffices for us to prove the following stronger result, providing a detailed local asymptotic expansion of the bifurcation solutions in terms of $\lambda$. \begin{Theorem} If (\ref{t5.2}) holds true for some small number $\epsilon>0$, then the two branches of steady-state solutions $(\lambda, u^\pm_\lambda)$ close to the bifurcation point $(\lambda_0,u_0)$ undergo the supercritical pitchfork bifurcation in the following sense: \begin{equation} \label{expansion} u^\pm_\lambda = \cases u_0 & \lambda<\lambda_0, \\[2pt] u_0 \pm \alpha \frac{\sqrt{\lambda-\lambda_0}}{\lambda} u^* +\frac{\lambda-\lambda_0}{\lambda} [ -u_0 + \alpha^2 \Delta^{-1} B_2(u^*,u^*) \\ +\alpha^2 (-\Delta +\lambda_0 A)^{-1}B_3(u^*,u^*) ]+o(\lambda-\lambda_0) & \lambda>\lambda_0\,.\endcases \end{equation} Here the eigenfunction $u^*$ is given by (\ref{u*}) and the constant $\alpha$ is defined \begin{equation}\label{alpha} \alpha= \frac{2k\pi^{3/2}}{\displaystyle \left( \int_{{\mathbb T}^3} B(u^*,u^*)\cdot u_0 \,dx\,dy\,dz \right)^{1/2}}. \end{equation} \end{Theorem} The definition of the constant $\alpha$ in (\ref{alpha}) is justified in the following lemma. \begin{Lemma} $$\int_{{\mathbb T}^3}B(u^*,u^*)\cdot u_0 \,dx\,dy\,dz >0.$$ \end{Lemma} \begin{proof}[Proof of Lemma 5.1] By the divergence free condition and integration by parts, \begin{eqnarray} \label{n5.6} \lefteqn{ \int_{{\mathbb T}^3}B(u^*,u^*)\cdot u_0 \,dx\,dy\,dz }\\ &=&-\int_{{\mathbb T}^3}B(u^*,u_0)\cdot u^*\,dx\,dy\,dz \nonumber\\ &=& -k\sum_{n,\,m\in {\mathbb Z}}\xi_n\zeta_m\int_{{\mathbb T}^3} \sin(n_0lx+n_0jy+\{n_0i\}z+nkz)\nonumber\\ &&\times \sin(n_0lx+n_0jy+\{n_0i\}z+mkz) \cos kz \,dx\,dy\,dz\nonumber \\ &=& -2k\pi^3 \sum_{n\in {\mathbb Z}} \xi_n(\zeta_{n-1}+\zeta_{n+1}). \nonumber \end{eqnarray} On the other hand, recall from (\ref{2.15})-(\ref{2.17}) that $u^*$ satisfies the following set of coupled algebraic equations \begin{align} & \frac{2\beta_n^2}{\lambda_{l,j,i,k} n_0l}\zeta_n +(\beta_{n-1}-k^2)\zeta_{n-1}-(\beta_{n+1}-k^2)\zeta_{n+1} =0,\label{n5.7} \\ & \label{n5.8} \frac{2\beta_n}{\lambda_{l,j,i,k} }(j\xi_n-l\eta_n) +n_0l(j\xi_{n-1}-l\eta_{n-1})-n_0l(j\xi_{n+1}-l\eta_{n+1}) \\ & \qquad =-kj (\zeta_{n-1}+\zeta_{n+1}), \nonumber \\ & -n_0j\eta_n=n_0l\xi_n+(\{n_0i\}+nk)\zeta_n, \label{n5.9} \end{align} for any $n\in {\mathbb Z}$, where $\beta_n = n_0^2l^2 +n_0^2j^2 +(\{n_0i\} +nk)^2$. Multiplying (\ref{n5.7}) by $\zeta_n$ and summing up the resultant equations yield \begin{eqnarray} \label{n5.10} \sum_{n\in{\mathbb Z}} \frac{2\beta_n^2}{\lambda_{l,j,i,k} n_0l}\zeta_n^2 &=&-\sum_{n\in {\mathbb Z}}(\beta_n-\beta_{n+1})\zeta_n\zeta_{n+1}\\ &=&\sum_{n\in{\mathbb Z}}k(2\{n_0i\}+2nk+k)\zeta_n\zeta_{n+1}.\nonumber \end{eqnarray} Moreover, multiplying (\ref{n5.8}) by $j\xi_n-l\eta_n$ and summing up the resultant equations yield \begin{equation} \label{nn5.10} \sum_{n\in {\mathbb Z}}\frac{2\beta_n}{\lambda_{l,j,i,k} }(j\xi_n-l\eta_n)^2 =-k\sum_{n\in{\mathbb Z}}(j^2\xi_n-lj\eta_n)(\zeta_{n-1}+\zeta_{n+1}). \end{equation} We infer then from (\ref{n5.6}), (\ref{n5.9}), (\ref{n5.10}) and (\ref{nn5.10}) that \begin{eqnarray*} \lefteqn{ \sum_{n\in {\mathbb Z}}\frac{2\beta_n}{\lambda_{l,j,i,k} }(j\xi_n-l\eta_n)^2 }\\ &=& -k\sum_{n\in{\mathbb Z}}(j^2\xi_n+l^2\xi_n +\frac{l}{n_0} (\{n_0i\}+nk)\zeta_n)(\zeta_{n-1}+\zeta_{n+1})\\ &=& -k(l^2+j^2)\sum_{n\in{\mathbb Z}}\xi_n(\zeta_{n-1}+\zeta_{n+1}) -\frac{kl}{n_0}\sum_{n\in{\mathbb Z}}(\{n_0i\}+nk)\zeta_n (\zeta_{n-1}+\zeta_{n+1})\\ &=&\frac{l^2+j^2}{2\pi^3}\int_{{\mathbb T}^3}B(u^*,u^*)\cdot u_0\,dx\,dy\,dz -\frac{kl}{n_0}\sum_{n\in{\mathbb Z}}(2\{n_0i\}+2nk+k)\zeta_n\zeta_{n+1}\\ &=&\frac{l^2+j^2}{2\pi^3}\int_{{\mathbb T}^3}B(u^*,u^*)\cdot u_0\,dx\,dy\,dz - \sum_{n\in{\mathbb Z}}\frac{2\beta_n^2}{\lambda_{l,j,i,k} n_0^2 }\zeta_n^2. \end{eqnarray*} We thus have $$ \int_{{\mathbb T}^3}B(u^*,u^*)\cdot u_0\,dx\,dy\,dz =\frac{4\pi^3}{(l^2+j^2)\lambda_{l,j,i,k}n_0^2} \sum_{n\in {\mathbb Z}}(\beta_nn_0^2(j\xi_n-l\eta_n)^2 +\beta_n^2\zeta_n^2) >0. $$ \end{proof} Now in order to facilitate the understanding of the problem, we consider the infinite mode truncation model by projecting (\ref{2.1}) onto the unstable space Span$\{u_0\}\oplus {\mathbb E}_{n_0l,n_0j,\{n_0i\},k}$: \begin{equation}\label{proj} \gathered \mbox{ Find } u\in W = \mbox{ Span}\{u_0\}\oplus {\mathbb E}_{n_0l,n_0j,\{n_0i\},k} \mbox{ such that } \\ \int_{{\mathbb T}^3}(-\Delta(u-u_0) +\lambda B(u,u) ) \cdot \phi \,dx\,dy\,dz =0, \qquad \forall \,\phi \in W. \endgathered \end{equation} \begin{Lemma} The parameter $\lambda=\lambda_0$ is a supercritical bifurcation point of (\ref{proj}) with the two bifurcation branches given by \begin{equation}\label{asmpt} u= \cases u_0 & \mbox{ when } \lambda\leq \lambda_0,\\ u_0 \pm \alpha \frac{\sqrt{\lambda-\lambda_0}} {\lambda} u^* - \frac{\lambda-\lambda_0}{\lambda}u_0 & \mbox{ when } \lambda> \lambda_0, \endcases \end{equation} for $ \|\Delta(u -u_0)\|_{L^2} +|\lambda-\lambda_0|<\epsilon$ for some constant $\epsilon >0$. \end{Lemma} \begin{proof}[Proof of Lemma 5.2] Decompose $u$ as $$u=\mu u_0 + v \,\,\mbox{ with } \mu\in {\mathbb R} ,\,\, v\in {\mathbb E}_{n_0l,n_0j,\{n_0i\},k}.$$ Then (\ref{proj}) becomes \begin{gather} \label{trun1} -\Delta v +\lambda \mu A v = 0,\\ k^2(\mu -1)+\frac{\lambda }{4\pi^3} \int_{{\mathbb T}^3} B(v,v)\cdot u_0 \,dx\,dy\,dz =0, \label{trun2} \end{gather} It follows from Theorem 2.3 that $ u=\mu u_0+v$ solves (\ref{trun1})-(\ref{trun2}) if and only if $$\mu\lambda =\lambda_0 \,\,\mbox{ and } \,\, v=cu^* \,\,\mbox{ for some }\,\,c\in {\mathbb R},$$ where the eigenfunction $u^*$ is defined by (\ref{u*}). Thus (\ref{trun2}) becomes $$ \frac{4k^2\pi^3(\lambda -\lambda_0)}{\lambda^2 }= c^2 \int_{{\mathbb T}^3} B(u^*,u^*)\cdot u_0 \,dx\,dy\,dz.$$ This together with Lemma 5.1 implies $$c=\cases 0 & \mbox{ if } \lambda <\lambda_0, \\[2pt] \pm \alpha \frac{\sqrt{\lambda-\lambda_0}}{\lambda} & \mbox{ if } \lambda >\lambda_0\,. \endcases $$ Thus (\ref{asmpt}) follows, and the lemma is proved. \end{proof} \begin{proof}[Proof of Theorem 5.2] Lemma 5.2 implies that the bifurcation solutions $u^\pm_\lambda $ close to the bifurcation point $(\lambda_0,u_0)$ are in the following form: \begin{equation} \label{comp} u^\pm_\lambda = u_0+ \frac{\sqrt{|\lambda-\lambda_0|}}{\lambda}w_1 + \frac{\lambda-\lambda_0}{\lambda}w +o(|\lambda-\lambda_0|), \end{equation} with $w_1,\, w\in \dbH^2_{n_0l,n_0j,\{n_0i\},k}$ independent of $\lambda$. Here $ w$ can be decomposed as $$\gathered w=-u_0 + w_2+w_3 \\ w_2\in {\mathbb E}_{0,0,0,k}, \quad \int_{{\mathbb T}^3}w_2\cdot u_0 \,dx\,dy\,dz =0, \\ \int_{{\mathbb T}^3}w_3\cdot \phi\,dx\,dy\,dz =0 \qquad \forall \phi\in {\mathbb E}_{0,0,0,k} \,. \endgathered$$ Thus \begin{eqnarray*} \lefteqn{\lambda B(u^\pm_\lambda,u^\pm_\lambda)}\\ &=& \sqrt{|\lambda-\lambda_0|}\left(B(u_0 +\frac{\lambda-\lambda_0}{\lambda} (w_2-u_0),w_1)+B (w_1,u_0+\frac{\lambda-\lambda_0}{\lambda}(w_2-u_0)) \right) \\ && +(\lambda-\lambda_0)\left(B(u_0 +\frac{\lambda-\lambda_0}{\lambda}(w_2-u_0) ,w_3)+ B(w_3,u_0+\frac{\lambda-\lambda_0}{\lambda}(w_2-u_0))\right) \\ &&+ \frac{|\lambda-\lambda_0|}{\lambda} B(w_1,w_1)+ o(|\lambda-\lambda_0|)\\ &=& \frac{\sqrt{|\lambda-\lambda_0|}}{\lambda}\lambda_0(B(u_0,w_1)+B(w_1,u_0))\\ &&+\frac{\sqrt{|\lambda-\lambda_0|}}{\lambda}\lambda_0(B(u_0,w_3)+B(w_3,u_0)) + \frac{|\lambda-\lambda_0|}{\lambda} B(w_1,w_1)+ o(|\lambda-\lambda_0|). \end{eqnarray*} Hence the stationary Navier-Stokes system \begin{equation} \label{st} -\Delta(u^\pm_\lambda-u_0) +\lambda B(u^\pm_\lambda,u^\pm_\lambda)=0 \,\,\,\, \,\mbox{ in } \dbH^2_{n_0l,n_0j,\{n_0i\},k} \end{equation} becomes \begin{eqnarray*} 0&=& -\frac{\sqrt{|\lambda-\lambda_0|}}{\lambda}\Delta w_1 -\frac{\lambda-\lambda_0}{\lambda}\Delta (w_2-u_0) -\frac{\lambda-\lambda_0}{\lambda}\Delta w_3 \\ && +\frac{\sqrt{|\lambda-\lambda_0|}}{\lambda}\lambda_0 A w_1 +\frac{(\lambda-\lambda_0)}{\lambda}\lambda_0 Aw_3 + \frac{|\lambda-\lambda_0|}{\lambda} B(w_1,w_1)+ o(|\lambda-\lambda_0|). \end{eqnarray*} This yields \begin{eqnarray*} -\Delta w_1 +\lambda_0 Aw_1&=&0,\\ (\lambda-\lambda_0)(\Delta(w_2-u_0+w_3) -\lambda_0Aw_3) &=& |\lambda-\lambda_0|B(w_1,w_1). \end{eqnarray*} By Theorem 2.3, we see $w_1=cu^*\in {\mathbb E}_{n_0l,n_0j,\{n_0i\},k}$ for some constant $c$ with $u^*$ defined by (\ref{u*}). Hence $$B(w_1,w_1)=c^2B(u^*,u^*) \in {\mathbb E}_{0,0,0,k}\oplus {\mathbb E}_{2n_0l,2n_0j,\{2n_0i\},k}, $$ with $B(u^*,u^*)$ decomposed by (\ref{interaction}). Then (\ref{st}) yields \begin{eqnarray} \nonumber 0&=&-\Delta u^* +\lambda_0 Au^*,\\ \label{n5.16} (\lambda-\lambda_0) k^2 u_0 &=& c^2 |\lambda-\lambda_0| B_0(u^*,u^*)\\ \nonumber (\lambda-\lambda_0)\Delta w_2 &=&c^2 |\lambda-\lambda_0|B_2(u^*,u^*),\\ (\lambda-\lambda_0)(-\Delta w_3 +\lambda Aw_3) &=& -c^2|\lambda-\lambda_0|B_3(u^*,u^*). \nonumber \end{eqnarray} By Lemma 5.1 and (\ref{n5.16}), we obtain that for $\lambda<\lambda_0$, $c=0$. For $\lambda>\lambda_0$, we have \begin{eqnarray} \nonumber 0&=& -\Delta u^* +\lambda_0 Au^*,\\ \label{CCC} c &=&\pm \alpha,\\ \nonumber w_2 &=&c^2 \Delta^{-1} B_2(u^*,u^*)\,\,\in {\mathbb E}_{0,0,0,k},\\ \nonumber w_3&=& c^2(\Delta -\lambda_0 A)^{-1}B_3(u^*,u^*) \,\, \in {\mathbb E}_{2n_0l,2n_0j,\{2n_0i\},k}. \end{eqnarray} Here we have used Theorem 2.3, the definition of $n_0$ and the Riesz-Schauder theory to obtain the boundedness of the operator $(-\Delta +\lambda_0 A)^{-1} $ in ${\mathbb E}_{2n_0l,2n_0j,\{2n_0i\},k}$. Thus the two branches of steady-state solutions $(\lambda, u^\pm_\lambda)$ close to the bifurcation point $(\lambda_0,u_0)$ undergo the supercritical pitchfork bifurcation in the sense given by (\ref{expansion}). The theorem is proved. \end{proof} To copy the results of Theorems 4.1 and 5.1 to another space $\tilde{\dbH}^2_{n_0l,n_0j,\{n_0i\},k}$. we notice that $\dbH^2_\sigma\supset \tilde{\dbH}^2_{n_0l,n_0j,\{n_0i\},k},$ which is in the form $$\aligned \bigg\{&u\in \dbH^2_\sigma\,|\, u=\sum_{n\in {\mathbb N}} (\xi_{n,1},\xi_{n,2},\xi_{n,3}) \sin nkz \\ &+\sum_{m\geq 2}\sum_{n\in {\mathbb Z}} (\eta_{m,n,1},\eta_{m,n,2},\eta_{m,n,3}) \cos ((2m-1)(n_0lx+n_0jy) \\ &\hspace{18mm} +\{2mn_0i-n_0i\}z+knz)\\ &+\sum_{n\in {\mathbb Z}} (\zeta_{n,1},\zeta_{n,2},\zeta_{n,3}) \sin (2m(n_0lx+n_0jy)+\{2mn_0i\}z+knz) \\ &+\sum_{n\in {\mathbb Z}} (\eta_{1,n,1},\eta_{1,n,2},\eta_{1,n,3}) \cos (n_0lx+n_0jy+\{n_0i\}z+knz)\bigg\}, \endaligned$$ and $$\tilde{\dbH}^2_{l,j,i,k}\supset \tilde\dbH^2_{n_0l,n_0j,\{n_0i\},k}\,\mbox{ for } n_0 \mbox{ odd }.$$ %(THE LINES FROM NEXT ONE TO THE ONE ABOVE THEOREM 4.2 ARE NEW ) We see that for each $(l,j,i,k)$ satisfying (\ref{4-tuple}), $\tilde{\mathbb E}_{n_0l,n_0j,\{n_0i\},k}\subset\tilde \dbH^2_{n_0l,n_0j,\{n_0i\},k}$. It follows from Theorem 2.3 that $\rho_{n_0l,n_0j,\{n_0i\},k}$ is simple and is unique eigenvalue of the spectral problem $$ \Delta u -\lambda A u=\rho u\,\quad \mbox{ for }\, \rho >-\beta_0 \mbox{ and } u\in\tilde \dbH^2_{n_0l,n_0j,\{n_0i\},k}$$ which is transversal across the imaginary axis at the origin. Then the argument as that given in the above proof of Theorems \ref{Th4.1} and \ref{Th5.1} shows that there exist two steady-state solutions $$\tilde u_{l,j,i,k,\lambda}^1,\,\,\tilde u_{l,j,i,k,\lambda}^2 \in \tilde \dbH^2_{n_0l,n_0j,\{n_0i\},k}$$ globally branching off the bifurcation point $(\lambda_{l,j,i,k},u_0)$, where $\lambda_{l,j,i,k}$ is defined by (\ref{4.3}). More precisely, we have the following Theorem. \begin{Theorem} For the critical Reynolds number $\lambda_{l,j,i,k}>0$ obtained by (\ref{4.3}), the problem (\ref{2.1}) with $0<\lambda <\infty$ admits also two steady-state solutions $$\tilde{u}^1_{l,j,i,k,\lambda}, \,\,\tilde{ u}^2_{l,j,i,k,\lambda}\in \tilde{\dbH}^2_{n_0l,n_0j,\{n_0i\},k}\subset \dbH^2_\sigma $$ branching off the point $(\lambda_{l,j,i,k}, u_0)$ continuously such that for $m=1,\,2$, $$ \begin{array}{ll} \displaystyle\tilde{ u}^m_{l,j,i,k,\lambda}= u_0 & \mbox{ for } \lambda\leq \lambda_{l,j,i,k}, \\[1pt] \displaystyle u_0\neq \tilde{ u}^1_{l,j,i,k,\lambda}\neq \tilde{ u}^2_{l,j,i,k,\lambda}\neq u_0 & \mbox{ for }\lambda>\lambda_{l,j,i,k}, \end{array} $$ provided that, for some small constant $\epsilon>0$, $$\|\Delta(\tilde {u}^m_{l,j,i,k,\lambda}-u_0)\|_{L^2} + |\lambda-\lambda_0|<\epsilon.$$ Furthermore, both of the global branches of bifurcation solutions $ \tilde u^m_{l,j,i,k,\lambda}$ satisfying the following properties: \begin{enumerate} \item each branch $ \tilde u^m_{l,j,i,k,\lambda}$ extends to $\lambda=\infty$ in $\mathcal C_1 \cap \mathcal C_2$ as shown in Figures 1.1 and 1.2; \item each branch $ \tilde u^m_{l,j,i,k,\lambda}$ intersects with the $\lambda$--axis only at the line segment $pq$. In particular, $ \tilde u^m_{l,j,i,k,\lambda}$ never touches the $\lambda$--axis for $\lambda > \lambda_{l,j,i,k}$. \end{enumerate} \end{Theorem} \setcounter{equation}{0} \section{Proof of Theorems 1.1 and 1.2} For an given integer vector $(l,j,i,k)$ satisfying (\ref{4-tuple}), let $u^m_{l,\pm j,i,k,\lambda}$, $\tilde u^m_{l,\pm j,i,k,\lambda}$, $u^m_{l,\pm j,k-i,k,\lambda}$ and $\tilde{u}^m_{l,\pm j,k-i,k,\lambda}$ be the bifurcation solutions obtained in Theorems 4.1, 5.1 and 5.3, and let \begin{equation} \label{5.1} u_{l,j,i,k,n,\lambda}= u^n_{l,j,i,k,\lambda} \in \dbH^2_{n_0l,n_0j,\{n_0i\},k},\mbox{ for } n=1,2. \end{equation} For $\{n(k-i)\}=k-\{n_0i\}$, we define other bifurcation solutions $u_{l,j,i,k,n,\lambda}$ as follows: \begin{enumerate} \item[a).] {\sl Case $\{n_0i\}=j=0$}: \begin{equation}\label{5.2} u_{l,0,i,k,n,\lambda}= \tilde{u}^{n-2}_{l,0,i,k,\lambda} \in \tilde \dbH^2_{n_0l,0,0,k} \mbox{ for } n=3,4; \end{equation} \item[b).] {\sl Case $\{n_0i\}=0, j\neq 0$}: \begin{equation}\label{5.3} u_{l,j,i,k,n,\lambda}=\left\{\begin{array}{ll} u^{n-2}_{l,-j,i,k,\lambda} \in \dbH^2_{n_0l,-n_0 j,0,k} &\mbox{ for } n=3,4,\\[2pt] \tilde{u}^{n-4}_{l,j,i,k,\lambda} \in \tilde \dbH^2_{n_0l,n_0 j,0,k} &\mbox{ for } n=5,6,\\[2pt] \tilde{u}^{n-6}_{l,-j,i,k,\lambda} \in \tilde \dbH^2_{n_0l,-n_0 j,0,k} &\mbox{ for } n=7,8. \end{array}\right. \end{equation} \item[c).] {\sl Case $\{n_0i\}\not=0, j=0$}: \begin{equation}\label{5.4} u_{l,0,i,k,n,\lambda}=\left\{\begin{array}{ll} u^{n-2}_{l,0,k-i,k,\lambda} \in \dbH^2_{n_0l,0, k-\{n_0 i\},k}, &\mbox{ for } n=3,4,\\[2pt] \tilde{u}^{n-4}_{l,0,i,k,\lambda} \in \tilde \dbH^2_{n_0l,0, \{n_0 i\},k} &\mbox{ for } n=5,6, \\[2pt] \tilde{u}^{n-6}_{l,0,k-i,k,\lambda} \in \tilde \dbH^2_{n_0l,0, k-\{n_0 i\},k} &\mbox{ for } n=7,8. \vspace{3mm} \end{array}\right. \end{equation} \item[d).] {\sl Case $\{n_0i\}\not=0, \,j\neq 0$}: \begin{equation} \label{5.5} u_{l,j,i,k,n,\lambda}=\left\{\begin{array}{ll} u^{n-2}_{l,j,k-i,k,\lambda} \in \dbH^2_{n_0l,n_0 j,k-\{n_0i\},k}, &\mbox{ for } n=3, 4, \\[2pt] u^{n-4}_{l,-j,i,k,\lambda} \in \dbH^2_{n_0l,-n_0 j,\{n_0i\},k}, &\mbox{ for } n=5, 6, \\[2pt] u^{n-6}_{l,-j,k-i,k,\lambda} \in \dbH^2_{n_0l,-n_0 j,k-\{n_0i\},k}, &\mbox{ for } n=7, 8, \\[2pt] \tilde{u}^{n-8}_{l,j,i,k,\lambda} \in \tilde \dbH^2_{n_0l,n_0 j,\{n_0i\},k} \\[2pt] \tilde{u}^{n-10}_{l,j,k-i,k,\lambda} \in \tilde \dbH^2_{n_0l,n_0 j,k-\{n_0i\},k} &\mbox{ for } n=11, 12, \\[2pt] \tilde{u}^{n-12}_{l,-j,i,k,\lambda} \in \tilde \dbH^2_{n_0l,-n_0 j,\{n_0i\},k} &\mbox{ for } n=13, 14, \\[2pt] \tilde{u}^{n-14}_{l,-j,k-i,k,\lambda} \in \tilde \dbH^2_{n_0l,-n_0 j,k-\{n_0i\},k} &\mbox{ for } n=15, 16. \end{array}\right. \end{equation} \end{enumerate} \begin{Lemma} The above defined solutions $u_{l,j,i,k,n,\lambda}$ bifurcate from the same critical Reynolds number $\lambda_{l,j,i,k}$. \end{Lemma} \begin{proof} By Theorem 2.3, the fact $\{n_0(k-i)\}=k-\{n_0i\}$ in case of $\{n_0i\}\neq 0$ and the proof of Theorem 2.1, it is easy to see that $\rho_{n_0l,n_0j,0,k}=\rho_{n_0l,-n_0j,0,k}$ when $\{n_0i\}=0$, and $\rho_{n_0l,n_0j,\{n_0i\},k}=\rho_{n_0l,n_0j,k-\{n_0i\},k} =\rho_{n_0l,-n_0j,\{n_0i\},k}=\rho_{n_0l,-n_0j,k-\{n_0i\},k}$ when $\{n_0i\}\neq 0.$ Therefore we infer from (\ref{4.3}) that $\lambda_{l,j,i,k}=\lambda_{l,-j,i,k}$ when $\{n_0i\}=0$, and $\lambda_{l,j,i,k}=\lambda_{l,j,k-i,k} = \lambda_{l,-j,i,k}=\lambda_{l,-j,k-i,k}$ when $\{n_0i\}\neq 0.$ \end{proof} \begin{proof}[Proof of Theorem 1.1] It suffices to prove (\ref{1.3}). To this end, set $$(l',j',i')=(n_0l,n_0j,\{n_0i\}).$$ Then we consider the following function spaces: {\sl Case $i'=j=0$}: \begin{equation*} {\mathbb E}_{l',j',i',k}\subset \dbH^2_{l',j',i',k},\qquad \tilde{{\mathbb E}}_{l',j',i',k}\subset \tilde \dbH^2_{l',j',i',k}; \end{equation*} {\sl Case $j\neq 0,\,i'=0$}: \begin{equation*} \begin{split} & {\mathbb E}_{l',j',i',k}\subset \dbH^2_{l',j',i',k}, \qquad \tilde{{\mathbb E}}_{l',j',i',k}\subset \tilde \dbH^2_{l',j',i',k},\\ & {\mathbb E}_{l',-j',i',k}\subset \dbH^2_{l',-j',i',k}, \qquad \tilde{{\mathbb E}}_{l',-j',i',k}\subset \tilde \dbH^2_{l',-j',i',k}; \end{split} \end{equation*} {\sl Case $j=0,\,i'>0$}: \begin{equation*} \begin{split} & {\mathbb E}_{l',j',i',k}\subset \dbH^2_{l',j',i',k}, \qquad \tilde{{\mathbb E}}_{l',j',i',k}\subset \tilde \dbH^2_{l',j',i',k}, \\ & {\mathbb E}_{l',j',k-i',k}\subset \dbH^2_{l',j',k-i',k}, \qquad \tilde{{\mathbb E}}_{l',j',k-i',k}\subset \tilde \dbH^2_{l',j',k-i',k}; \end{split} \end{equation*} {\sl Case $ji'\neq 0$}: \begin{equation*} \begin{split} & {\mathbb E}_{l',j',i',k}\subset \dbH^2_{l',j',i',k}, \qquad \tilde{{\mathbb E}}_{l',j',i',k}\subset \tilde \dbH^2_{l',j',i',k}, \\ & {\mathbb E}_{l',-j',i',k}\subset \dbH^2_{l',-j',i',k}, \qquad \tilde{{\mathbb E}}_{l',-j',i',k}\subset \tilde \dbH^2_{l',-j',i',k}\\ & {\mathbb E}_{l',j',k-i',k}\subset \dbH^2_{l',j',k-i',k}, \qquad \tilde{{\mathbb E}}_{l',j',k-i',k}\subset \tilde \dbH^2_{l',j',k-i',k}, \\ & {\mathbb E}_{l',-j',k-i',k}\subset \dbH^2_{l',-j',k-i',k}, \qquad \tilde{{\mathbb E}}_{l',-j',k-i',k}\subset \tilde \dbH^2_{l',-j',k-i',k}. \end{split} \end{equation*} It is easy to see that for any integer vector $(l,j,i,k)$ satisfying (\ref{4-tuple}) and for each case, \begin{enumerate} \item each of these subspaces contains an unstable subspace of (\ref{2.1}), generating the bifurcation solutions; and \item these subspaces are orthogonal to each other in $\dbH^2_\sigma$. \end{enumerate} Hence (\ref{1.3}), consequently Theorem 1.1, follows. \end{proof} \begin{proof}[Proof of Theorem 1.2] 1. By Theorems 4.1, 5.1 and 5.2 and (\ref{5.1}-\ref{5.5}), it remains to show (\ref{1.4}). For the integer vector $(l,j,i,k)$ given by (\ref{4-tuple}), we set $${\mathbb Y}=\bigg\{ u\in \dbH^2_{l,j,i,k}|\, u= \sum^{\infty}_{n=1}(\xi_n,\eta_n,\zeta_n)\sin knz\bigg\}.$$ It is easy to see that $u_0=(\sin kz,0,0)$ is the unique steady-state solution of (\ref{2.1}) in ${\mathbb Y}$. With this observation in mind, the cases where $\{n_0i\}=0$ and $\{n_0i\}\neq 0$ for (\ref{1.4}) are shown respectively as follows: 2. In the case where $\{n_0i\}=0$ and $j \neq 0$, by (\ref{i-zero}), $i_0=4$. Then we need to check, for $i_1=0,\,4$, \begin{equation}\label{5.6} u_{l,j,i,k,n,\lambda}\neq u_{l,j,i,k,n',\lambda}, \mbox{ if } \lambda_{l,j,i,k}<\lambda<\infty,\, i_1+1\leq n\leq i_1+2\lambda_{l,j,i,k}.$$ This gives (\ref{5.6}) with $i_1=0$. For $i_1=4$ we see that $$ {\mathbb Y} = \tilde \dbH^2_{n_0l,n_0 j,0,k} \cap \tilde \dbH^2_{n_0l,-n_0 j,0,k}, $$ and from (\ref{5.3}), $$ u_{l,j,i,k,n,\lambda}\in \tilde \dbH^2_{n_0l,n_0 j,0,k},$$ $$u_{l,j,i,k,n',\lambda}\in \tilde \dbH^2_{n_0l,-n_0 j, 0,k} . $$ By Assertion 2 of Theorem 1.2, $$u_{l,j,i,k,n,\lambda}\neq u_0\neq u_{l,j,i,k,n',\lambda} \mbox{ for } \lambda>\lambda_{l,j,i,k}.$$ This gives (\ref{5.6}) with $i_1=4$. 3. In the case where $j\{n_0i\}\neq 0$, we see that $i_0=16$. When $i_1=0$, we notice that \begin{eqnarray*} {\mathbb Y} &=&\dbH^2_{n_0l,n_0j,\{n_0i\},k}\cap \left( \dbH^2_{n_0l,-n_0j,\{n_0i\},k}\cup \dbH^2_{n_0l,-n_0j,k-\{n_0i\},k} \right)\\ &=&\dbH^2_{n_0l,n_0j,k-\{n_0i\},k}\cap \left( \dbH^2_{n_0l,-n_0j,\{n_0i\},k}\cup \dbH^2_{n_0l,-n_0j,k-\{n_0i\},k} \right). \end{eqnarray*} From (\ref{5.4}) it follows that $$ u_{l,j,i,k,n,\lambda}\in \dbH^2_{n_0l,n_0j,\{n_0i\},k}\cup \dbH^2_{n_0l,n_0 j,k-\{n_0i\},k} \mbox{ for } 1\leq n\leq 4,$$ and $$ u_{l,j,i,k,n',\lambda}\in \dbH^2_{n_0l,-n_0j,\{n_0i\},k}\cup \dbH^2_{n_0l,-n_0 j,k-\{n_0i\},k} \mbox{ for } 5\leq n\leq 8.$$ By Assertion 2 of Theorem 1.2, we thus have $$ u_{l,j,i,k,n,\lambda}\neq u_{l,j,i,k,n',\lambda} \mbox{ for } 1\leq n\leq 4\lambda_{l,j,i,k} \mbox{ and } 9\leq n\leq 12