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

The number of traveling wave families in a running water with Coriolis force

Zhiwu Lin School of Mathematics, Georgia Institute of Technology, 30332, Atlanta, GA, USA [email protected] Dongyi Wei School of Mathematical Science, Peking University, 100871, Beijing, P. R. China [email protected] Zhifei Zhang School of Mathematical Science, Peking University, 100871, Beijing, P. R. China [email protected]  and  Hao Zhu Department of Mathematics, Nanjing University, 210093, Nanjing, Jiangsu, P. R. China [email protected]
Abstract.

In this paper, we study the number of traveling wave families near a shear flow under the influence of Coriolis force, where the traveling speeds lie outside the range of the flow uu. Under the β\beta-plane approximation, if the flow uu has a critical point at which uu attains its minimal (resp. maximal) value, then a unique transitional β\beta value exists in the positive (resp. negative) half-line such that the number of traveling wave families near the shear flow changes suddenly from finite to infinite when β\beta passes through it. On the other hand, if uu has no such critical points, then the number is always finite for positive (resp. negative) β\beta values. This is true for general shear flows under mildly technical assumptions, and for a large class of shear flows including a cosine jet u(y)=1+cos(πy)2u(y)={1+\cos(\pi y)\over 2} (i.e. the sinus profile) and analytic monotone flows unconditionally. The sudden change of the number of traveling wave families indicates that long time dynamics around the shear flow is much richer than the non-rotating case, where no such traveling wave families exist.

1. Introduction

The earth’s rotation influences dynamics of large-scale flows significantly. Under the β\beta-plane approximation, the motion for such a flow could be described by 2-D incompressible Euler equation with rotation

(1.1) tv+(v)v=PβyJv,v=0,\displaystyle\partial_{t}\vec{v}+(\vec{v}\cdot\nabla)\vec{v}=-\nabla P-\beta yJ\vec{v},\quad\nabla\cdot\vec{v}=0,

where v=(v1,v2)\vec{v}=(v_{1},v_{2}) is the fluid velocity, PP is the pressure, J=(0110)J=\begin{pmatrix}0&-1\\ 1&0\end{pmatrix} is the rotation matrix, and β\beta is the Rossby number. Here we study the fluid in a periodic finite channel Ω=DT=𝕋T×[y1,y2]\Omega=D_{T}=\mathbb{T}_{T}\times{[y_{1},y_{2}]}, 𝕋T=𝐑/(T𝐙)\mathbb{T}_{T}=\mathbf{R}/(T\mathbf{Z}) with non-permeable boundary condition on Ω\partial\Omega:

(1.2) v2=0ony=y1,y2.\displaystyle v_{2}=0\quad\text{on}\quad y=y_{1},y_{2}.

The β\beta-plane approximation is commonly used for large-scale motions in geophysical fluid dynamics [35, 36]. The vorticity form of (1.1) takes

(1.3) tω+(v)ω+βv2=0,\partial_{t}\omega+(\vec{v}\cdot\nabla)\omega+\beta v_{2}=0,

where ω=xv2yv1\omega=\partial_{x}v_{2}-\partial_{y}v_{1}. By the incompressible condition, we introduce the stream function ψ\psi such that v=ψ=(yψ,xψ)\vec{v}=\nabla^{\perp}\psi=(\partial_{y}\psi,-\partial_{x}\psi). Consider the shear flow (u(y),0)(u(y),0), which is a steady solution of (1.3). The linearized equation of (1.3) around (u(y),0)(u(y),0) takes

(1.4) tω+uxω(βu′′)xψ=0,\displaystyle\partial_{t}\omega+u\partial_{x}\omega-(\beta-u^{\prime\prime})\partial_{x}\psi=0,

which was derived in [42].

In the study of long time dynamics near a shear flow, the most rigid case is the nonlinear inviscid damping (to a shear flow), a kind of asymptotic stability. This means that if the initial velocity is taken close enough to the given shear flow in some function space, then the velocity tends asymptotically to a nearby shear flow in this space. The existence of nearby non-shear steady states or traveling waves means that nonlinear inviscid damping (to a shear flow) is not true, and long time dynamics near the shear flow may be richer and fruitful. To understand the richer long time dynamics near the shear flow in this situation, an important step is to clarify whether the number of curves of nearby traveling waves with traveling speeds converging to different points is infinite. Indeed, if the number is finite, then the velocity might tend asymptotically to some nonlinear superpositions of finite such non-shear states when the initial data is taken close to the flow, and quasi-periodic nearby solutions are expected, which indicates new but not so complicated dynamics. If the number is infinite, then the evolutionary velocity might tend asymptotically to superpositions of infinite such non-shear states, and almost periodic nearby solutions potentially exist, which predicts complicated even chaotic long time behavior near the flow. Similar phenomena were observed numerically in the study of Vlasov-Poisson system, a model describing collisionless plasmas [9, 4, 5, 25]. This model shares many similarities with the 2D incompressible Euler equation. By numerical simulations, it was found that for some initial perturbation near homogeneous states, the asymptotic state toward which the system evolves can be described by a superposition of BGK modes [9]. This offers a hint for further numerical study in the 2D Euler case. It is very challenging to study long time dynamics near a shear flow in a fully analytic way when such non-shear steady states or traveling waves exist. The first step towards this direction is to construct nonlinear superpositions of traveling waves as in the Vlasov-Poisson case.

When there is no Coriolis force, long time dynamics near monotone flows is relatively rigid in strong topology, while it is still highly non-trivial to give a mathematical confirmation. A first step is to understand the linearized equation. Orr [38] observed the linear damping for Couette flow, and Case predicted the decay of velocity for monotone shear flows. Recently, their predictions are confirmed in [29, 48, 49, 44, 14, 19, 20] and are extended to non-monotone flows in [45, 46]. Meanwhile, great progress has been made in the study of nonlinear dynamics near shear flows. Bedrossian and Masmoudi [3] proved nonlinear inviscid damping near Couette flow for the initial perturbation in some Gevrey space on 𝕋×𝐑\mathbb{T}\times\mathbf{R}. Ionescu and Jia [17] extended the above asymptotic stability to a periodic finite channel 𝕋×[1,1]\mathbb{T}\times[-1,1] under the assumption that the initial vorticity perturbation is compacted supported in the interior of the channel. Later, nonlinear inviscid damping was proved near a class of Gevrey smooth monotone shear flows in a periodic finite channel if the perturbation is taken in a suitable Gevrey space, where u′′(y)u^{\prime\prime}(y) is compactly supported [18, 34]. It is still challenging to study the long time behavior near general, rough, monotone or non-monotone shear flows. On the other hand, inviscid damping (to a shear flow) depends on the regularity of the perturbation, and the existence of non-shear stationary structures is shown near some specific flows. Lin and Zeng [29] found cats’ eyes flows near Couette for H<32H^{<{3\over 2}} vorticity perturbation in a periodic finite channel, while no non-shear traveling waves near Couette exist if the regularity is H>32H^{>{3\over 2}}, in contrast to the linear level, where damping is always true for any initial vorticity in L2L^{2}. For Kolmogorov flows, which is non-monotone, Coti Zelati, Elgindi and Widmayer [8] constructed non-shear stationary states near Kolmogorov at analytic regularity on the square torus, while there are no nearby non-shear steady states at regularity H3H^{3} for velocity on a rectangular torus. They also proved that any traveling wave near Poiseuille must be shear for H>5H^{>5} vorticity perturbation in a periodic finite channel.

As indicated in [36], the study of the dynamics of large-scale oceanic or atmospheric motions must include the Coriolis force to be geophysically relevant, and once the Coriolis force is included a host of subtle and fascinating dynamical phenomena are possible. By numerical computation, Kuo [24] found the boundary of barotropic instability for the sinus profile, which is far from linear instability in no Coriolis case. Later, based on Hamiltonian index theory and spectral analysis, Lin, Yang and Zhu theoretically confirmed large parts of the boundary and corrected the rest. New traveling waves, which are purely due to the Coriolis effects, are found near the sinus profile [28]. Barotropic instability of other geophysical shear flows has also attracted much attention. For instance, by looking for the neutral solutions, most of the stability boundary, which is again different from no Coriolis case, of bounded and unbounded Bickley jet is found numerically and analytically in [30, 16, 32, 2, 12]. More fruitful geophysical fluid dynamics, such as Rossby wave and baroclinic instability, could be found in [24, 36, 10, 31, 35]. On the other hand, similar to no Coriolis case, linear inviscid damping is still true for a large class of flows and moreover, the same decay estimates of the velocity can be obtained for a class of monotone flows [47]. Elgindi and Widmayer [11] viewed Coriolis effect as one mechanism helping to stabilize the motion of an ideal fluid, and proved the almost global stability of the zero solution for the β\beta-plane equation. Global stability of the zero solution is further to be confirmed in [37].

When Coriolis force is involved, long time dynamics near a shear flow becomes fruitful. One of the main reasons is that, comparing with no Coriolis case, there are new traveling waves with fluid trajectories moving in one direction. This paper is devoted to studying the number of such traveling wave families near a general shear flow uu under the influence of Coriolis force. Here, a traveling wave family roughly includes the sets of nearby traveling waves with traveling speeds converging to a same number outside the range of the flow, see Definition 2.8 for details. Precisely, we prove that if the flow uu has a critical point at which uu attains its minimal (resp. maximal) value, then a unique transitional β\beta value β+\beta_{+} (resp. β\beta_{-}) exists in the positive (resp. negative) half-line, through which the number of traveling wave families changes suddenly from finite to infinite. The transitional β\beta values are defined in (1.11)–(1.12). If the flow uu has no such critical points, then the number of traveling wave families is always finite for positive (resp. negative) β\beta values. This is true for general shear flows under mildly technical assumptions. Based on Hamiltonian structure and index theory, we unconditionally prove the above results for a flow in class 𝒦+\mathcal{K}^{+}, which is defined as follows.

Definition 1.1.

A flow uu in class 𝒦+\mathcal{K}^{+} means that uH4(y1,y2)u\in H^{4}(y_{1},y_{2}) is not a constant function, and for any βRan(u′′)\beta\in\text{Ran}(u^{\prime\prime}), there exists uβRan(u)u_{\beta}\in\text{Ran}(u) such that Kβ=(βu′′)/(uuβ)K_{\beta}=(\beta-u^{\prime\prime})/(u-u_{\beta}) is positive and bounded on [y1,y2][y_{1},y_{2}].

A typical example of such a flow is a cosine jet u(y)=1+cos(πy)2,u(y)={\frac{1+\cos(\pi y)}{2}}, y[1,1]y\in[-1,1] (i.e. the sinus profile), which was studied in geophysical literature [23, 24, 36]. For β=0\beta=0 and a general shear flow uC2([y1,y2])u\in C^{2}([y_{1},y_{2}]), Rayleigh [39] gave a necessary condition for spectral instability that u′′(y0)=0u^{\prime\prime}(y_{0})=0 for some y0(y1,y2)y_{0}\in(y_{1},y_{2}), and even under this condition, Fjørtoft [13] provided a sufficient condition for spectral stability that (uu(y0))u′′0(u-u(y_{0}))u^{\prime\prime}\geq 0 on (y1,y2)(y_{1},y_{2}). For β0\beta\neq 0 and uC2([y1,y2])u\in C^{2}([y_{1},y_{2}]), the above two conditions can be extended as βu′′(yβ)=0\beta-u^{\prime\prime}(y_{\beta})=0 for some yβ(y1,y2)y_{\beta}\in(y_{1},y_{2}) and (βu′′)(uu(yβ))0(\beta-u^{\prime\prime})(u-u(y_{\beta}))\leq 0 on (y1,y2)(y_{1},y_{2}), respectively. See, for example, (6.3)–(6.4) in [24]. For a flow in class 𝒦+\mathcal{K}^{+}, the extended Rayleigh’s condition implies that βRan(u′′)\beta\in\text{Ran}(u^{\prime\prime}) is necessary for spectral instability, but the flow does not satisfy the extended Fjørtoft’s sufficient condition for spectral stability. The sharp condition for spectral stability indeed depends on β\beta and the wave number α\alpha, which was obtained in [27] for β=0\beta=0 and in [28] for β0\beta\neq 0.

Consider a class of general shear flows satisfying

(H1)uH4(y1,y2),u′′0onu’scritical level{u=0}.\displaystyle(\textbf{H1})\quad u\in H^{4}(y_{1},y_{2}),\,\,u^{\prime\prime}\neq 0\,\,\text{on}\,\,u\text{'s}\,\,\text{critical level}\;\{u^{\prime}=0\}.

A flow uu in class 𝒦+\mathcal{K}^{+} satisfies the assumption (H1). In fact, it is trivial for 0Ran(u′′)0\notin\text{Ran}(u^{\prime\prime}); if 0Ran(u′′)0\in\text{Ran}(u^{\prime\prime}) and y0[y1,y2]y_{0}\in[y_{1},y_{2}] satisfies u(y0)=0u^{\prime}(y_{0})=0 and u′′(y0)=0u^{\prime\prime}(y_{0})=0, then u(y0)u0=1K0(y0)u′′(y0)=0u(y_{0})-u_{0}=-{1\over K_{0}(y_{0})}u^{\prime\prime}(y_{0})=0. Thus, φuu0\varphi\equiv u-u_{0} solves φ′′+K0φ=0\varphi^{\prime\prime}+K_{0}\varphi=0, φ(y0)=φ(y0)=0\varphi(y_{0})=\varphi^{\prime}(y_{0})=0. Then uu0u\equiv u_{0}, which is a contradiction.

To state our main results with few restriction, we first consider flows in class 𝒦+\mathcal{K}^{+}, and left the extension to general shear flows satisfying (H1) in Section 2.

Theorem 1.2.

Let β0\beta\neq 0 and the flow uu be in class 𝒦+\mathcal{K}^{+}.

  • (1)

    If {u=0}{u=umin}\{u^{\prime}=0\}\cap\{u=u_{\min}\}\neq\emptyset, then there exists β+(0,)\beta_{+}\in(0,\infty) such that there exist at most finitely many traveling wave families near (u,0)(u,0) for β(0,β+]\beta\in(0,\beta_{+}], and infinitely many traveling wave families near (u,0)(u,0) for β(β+,)\beta\in(\beta_{+},\infty). Moreover, β+\beta_{+} is specified in (1.11).

  • (2)

    If {u=0}{u=umax}\{u^{\prime}=0\}\cap\{u=u_{\max}\}\neq\emptyset, then there exists β(,0)\beta_{-}\in(-\infty,0) such that there exist at most finitely many traveling wave families near (u,0)(u,0) for β[β,0)\beta\in[\beta_{-},0), and infinitely many traveling wave families near (u,0)(u,0) for β(,β)\beta\in(-\infty,\beta_{-}). Moreover, β\beta_{-} is specified in (1.12).

  • (3)

    If {u=0}{u=umin}=\{u^{\prime}=0\}\cap\{u=u_{\min}\}=\emptyset, then there exist at most finitely many traveling wave families near (u,0)(u,0) for β(0,)\beta\in(0,\infty).

  • (4)

    If {u=0}{u=umax}=\{u^{\prime}=0\}\cap\{u=u_{\max}\}=\emptyset, then there exist at most finitely many traveling wave families near (u,0)(u,0) for β(,0)\beta\in(-\infty,0).

Here, the precise description of a traveling wave family near (u,0)(u,0) is given in Definition 2.8.

Unless otherwise specified, “near (u,0)(u,0)” always means “in a (velocity) H3H^{3} neighborhood of (u,0)(u,0)” in Theorem 1.2 and the rest of this paper, as indicated in Definition 2.4. These traveling wave families do not exist if there is no Coriolis force. By Theorem 1.2, Coriolis force and its magnitude indeed bring fascinating dynamics near the shear flow. On the one hand, for flows having no critical point which is meanwhile a minimal point, the number of traveling wave families is always finite no matter how much magnitude of Coriolis force, which is a mild Coriolis effect. On the other hand, for flows having such a critical point, there is a surprisingly sharp difference, namely, when the Coriolis parameter passes through the transitional point β+\beta_{+}, the number of traveling wave families changes suddenly from finite to infinite. In particular, quasi-periodic solutions to (1.1)–(1.2) can be expected near the shear flow for β(0,β+]\beta\in(0,\beta_{+}], while almost periodic solutions potentially exist for β(β+,)\beta\in(\beta_{+},\infty). This could be regarded as a strong Coriolis effect and predicts chaotic long time dynamics near these flows.

The same dynamical phenomena are true for general shear flows under some mildly technical assumptions. The explicit result is stated in Theorem 2.2. For β>0\beta>0, the technical assumption for flows having a critical and meanwhile minimal point is that uminu_{\min} is not an embedding eigenvalue of the linearized Euler operator for small wave numbers. The assumption for flows having no such critical points is some regularized condition near the endpoints of uu. Note that the first spectral assumption has only restriction for one point uminu_{\min}, no matter whether the interior of Ran(u)\text{Ran}\,(u) has embedding eigenvalues. The second assumption is more generic and quite easy to verify. Both the two technical assumptions are used for ruling out eigenvalues’ oscillation for Rayleigh-Kuo boundary value problem (BVP) as the parameter cc tends to uminu_{\min}, see Subsection 2.2 for details.

Let us give some remarks on properties of such traveling waves near the flow uu.

  • The traveling waves have fluid trajectories moving in one direction, see (5.6) in the proof of Lemma 2.5. Thus unlike the constructed steady flow near Couette flow in [29], the streamlines here have no cat’s eyes structure.

  • The traveling waves can be constructed near a smooth shear flow for H3H^{\geq{3}} (including H>6H^{>6}) velocity perturbation when the Coriolis parameter is large, see Corollary 2.6. In contrast, in the case of no Coriolis force, no traveling waves could be found near Couette flow for H>52H^{>{5\over 2}} velocity perturbation [29] and near Poiseuille flow for H>6H^{>6} velocity perturbation [8].

  • Let {u=0}{u=umin}\{u^{\prime}=0\}\cap\{u=u_{\min}\}\neq\emptyset and β>98κ+\beta>{9\over 8}\kappa_{+}. The directions of vertical velocities of the nearby traveling waves might change frequently with small amplitude as the traveling speeds converge to uminu_{\min}^{-}, see Remark 5.2.

We apply the main results to analytic monotone flows (including Couette flow) and the sinus profile. For an analytic monotone flow, there exist at most finitely many nearby traveling wave families for β0\beta\neq 0, see Corollary 2.3. For the sinus profile, as mentioned above, it is in class 𝒦+\mathcal{K}^{+}, and so applying Theorem 1.2 (1)-(2) we get that {u=0}{u=umin}={±1}\{u^{\prime}=0\}\cap\{u=u_{\min}\}=\{\pm 1\}, {u=0}{u=umax}={0}\{u^{\prime}=0\}\cap\{u=u_{\max}\}=\{0\}, β+=916π2\beta_{+}={9\over 16}\pi^{2}, β=12π2\beta_{-}=-{1\over 2}\pi^{2}, there exist at most finitely many traveling wave families near the sinus profile for β[12π2,916π2]\beta\in[-{1\over 2}\pi^{2},{9\over 16}\pi^{2}], and infinitely many nearby traveling wave families for β[12π2,916π2]\beta\notin[-{1\over 2}\pi^{2},{9\over 16}\pi^{2}]. Moreover, we will give a systematical study on the number of isolated real eigenvalues of the linearized Euler operator and traveling wave families near the sinus profile on the whole (α,β)(\alpha,\beta)’s region in Section 7 (here α\alpha is the wave number in the xx-direction), which plays an important role in further study on its long time dynamics. We make a comparison with the previous work in [28]. By Theorem 2.1, the number of isolated real eigenvalues of the linearized Euler operator (i.e. non-resonant modes) determines that of traveling wave families. The explicit number of isolated real eigenvalues in the region (α,β)(0,)×[π22,π22](\alpha,\beta)\in(0,\infty)\times[-{\pi^{2}\over 2},{\pi^{2}\over 2}] can be obtained in [28], but no information can be concluded outside this region, see the discussion below Figure 4 in [28]. Our new contribution for the sinus profile in this paper is that we calculate the explicit number of isolated real eigenvalues in the remaining area (α,β)(0,)×(,π22)(π22,)(\alpha,\beta)\in(0,\infty)\times(-\infty,-{\pi^{2}\over 2})\cup({\pi^{2}\over 2},\infty), and thus completely get the number of traveling wave families near the sinus profile on the whole (α,β)(\alpha,\beta)’s region. For the sinus profile, the novelty is that we give the asymptotic behavior of the nn-th eigenvalue λn(c)\lambda_{n}(c) of the Rayleigh-Kuo BVP (2.6) as c0c\to 0^{-} for β(π22,)\beta\in({\pi^{2}\over 2},\infty) and as c1+c\to 1^{+} for β(,π22)\beta\in(-\infty,-{\pi^{2}\over 2}), from which we find the transitional β\beta values such that the number of traveling wave families changes suddenly from finite to infinite. For general shear flows satisfying (𝐇𝟏){\bf(H1)}, the key is to study whether λn(c)\lambda_{n}(c) is unbounded from below as cc is close to uminu_{\min} (or umaxu_{\max}) in Theorem 2.9 and to rule out the oscillation of λn(c)\lambda_{n}(c) in Theorems 2.11-2.13. In this paper, we focus on the description of the eigenvalues of the Rayleigh-Kuo BVP (2.6), which in turn, by Theorem 2.1, yields information on traveling wave families.

The rest of this paper is organized as follows. In Section 22, we extend Theorem 1.2 to general shear flows and give the outline of the proof. In Sections 3-4, we study the asymptotic behavior of the nn-th eigenvalue of Rayleigh-Kuo BVP, where we determine the transitional values for the nn-th eigenvalue of Rayleigh-Kuo BVP in Section 3, and rule out oscillation of the nn-th eigenvalue in Section 4. In Section 55, we establish the correspondence between a traveling wave family and an isolated real eigenvalue of the linearized Euler operator. In Section 66, we prove the main Theorems 2.2 and 1.2. As a concrete application, we thoroughly study the number of traveling wave families near the sinus profile in the last section.

Notation

We provide the notations that we use in this paper. Let umin=min(u)u_{\min}=\min(u) and umax=max(u)u_{\max}=\max(u) for uC([y1,y2])u\in C([y_{1},y_{2}]). For a shear flow uu satisfying (H1)(\textbf{H1}), we use the following characteristic quantities of the flow. If {u=0}{u=umin},\{u^{\prime}=0\}\cap\{u=u_{\min}\}\neq\emptyset, we define

(1.5) κ+\displaystyle\kappa_{+} :=min{u′′(y)|y[y1,y2] such that u(y)=0 and u(y)=umin}.\displaystyle:=\min\{u^{\prime\prime}(y)|y\in[y_{1},y_{2}]\text{ such that }u^{\prime}(y)=0\text{ and }u(y)=u_{\min}\}.

If {u=0}{u=umax},\{u^{\prime}=0\}\cap\{u=u_{\max}\}\neq\emptyset, we define

(1.6) κ\displaystyle\kappa_{-} :=max{u′′(y)|y[y1,y2] such that u(y)=0 and u(y)=umax}.\displaystyle:=\max\{u^{\prime\prime}(y)|y\in[y_{1},y_{2}]\text{ such that }u^{\prime}(y)=0\text{ and }u(y)=u_{\max}\}.

Note that κ+(0,)\kappa_{+}\in(0,\infty) and κ(,0)\kappa_{-}\in(-\infty,0) in (1.5)–(1.6). In fact, (𝐇𝟏){\bf(H1)} implies u′′(y0)>0u^{\prime\prime}(y_{0})>0 for y0A:={y[y1,y2]|u(y)=0 and u(y)=umin}y_{0}\in A:=\{y\in[y_{1},y_{2}]|u^{\prime}(y)=0\text{ and }u(y)=u_{\min}\}. Then y0y_{0} is an isolated point of AA. Thus, AA is a finite set and κ+(0,)\kappa_{+}\in(0,\infty) in (1.5). Similarly, κ(,0)\kappa_{-}\in(-\infty,0) in (1.6). Besides (1.5)–(1.6), we define

(1.7) κ+:=, if\displaystyle\kappa_{+}:=\infty,\quad\text{ if } {u=0}{u=umin}=,\displaystyle\{u^{\prime}=0\}\cap\{u=u_{\min}\}=\emptyset,
(1.8) κ:=, if\displaystyle\kappa_{-}:=-\infty,\;\text{ if } {u=0}{u=umax}=.\displaystyle\{u^{\prime}=0\}\cap\{u=u_{\max}\}=\emptyset.

If {u=umin}(y1,y2),\{u=u_{\min}\}\cap(y_{1},y_{2})\neq\emptyset, we define

(1.9) μ+\displaystyle\mu_{+} :=min{u′′(y)|y(y1,y2) such that u(y)=umin}.\displaystyle:=\min\{u^{\prime\prime}(y)|y\in(y_{1},y_{2})\text{ such that }u(y)=u_{\min}\}.

If {u=umax}(y1,y2),\{u=u_{\max}\}\cap(y_{1},y_{2})\neq\emptyset, we define

(1.10) μ\displaystyle\mu_{-} :=max{u′′(y)|y(y1,y2) such that u(y)=umax}.\displaystyle:=\max\{u^{\prime\prime}(y)|y\in(y_{1},y_{2})\text{ such that }u(y)=u_{\max}\}.

Note that μ+[κ+,)\mu_{+}\in[\kappa_{+},\infty) and μ(,κ]\mu_{-}\in(-\infty,\kappa_{-}] in (1.9)–(1.10). Then we define

(1.11) β+:={min{98κ+,μ+},if {u=umin}(y1,y2),98κ+,if {u=umin}(y1,y2)=,\displaystyle\beta_{+}:=\begin{cases}\min\{{9\over 8}\kappa_{+},\mu_{+}\},&\text{if }\{u=u_{\min}\}\cap(y_{1},y_{2})\neq\emptyset,\\ {9\over 8}\kappa_{+},&\text{if }\{u=u_{\min}\}\cap(y_{1},y_{2})=\emptyset,\end{cases}

and

(1.12) β:={max{98κ,μ},if {u=umax}(y1,y2),98κ,if {u=umax}(y1,y2)=.\displaystyle\beta_{-}:=\begin{cases}\max\{{9\over 8}\kappa_{-},\mu_{-}\},&\text{if }\{u=u_{\max}\}\cap(y_{1},y_{2})\neq\emptyset,\\ {9\over 8}\kappa_{-},&\text{if }\{u=u_{\max}\}\cap(y_{1},y_{2})=\emptyset.\end{cases}

We denote

(1.13) (E+)umin is not an embedding eigenvalue of α,β,\displaystyle(\textbf{E${}_{+}$})\quad\quad\quad u_{\min}\text{ is not an embedding eigenvalue of }\mathcal{R}_{\alpha,\beta},
(1.14) (E-)umax is not an embedding eigenvalue of α,β,\displaystyle(\textbf{E${}_{-}$})\quad\quad\quad u_{\max}\text{ is not an embedding eigenvalue of }\mathcal{R}_{\alpha,\beta},

where α,β\mathcal{R}_{\alpha,\beta} is defined in (2.5). Moreover, we define

(1.17) mβ\displaystyle m_{\beta} :={{a(y1,y2)|u(a)=umin,u′′(a)β<0},if 0<β98κ+,{a(y1,y2)|u(a)=umax,u′′(a)β>0},if98κβ<0,\displaystyle:=\left\{\begin{array}[]{ll}\sharp\{a\in(y_{1},y_{2})|u(a)=u_{\min},u^{\prime\prime}(a)-\beta<0\},&\text{if}\;0<\beta\leq{9\over 8}\kappa_{+},\\ \sharp\{a\in(y_{1},y_{2})|u(a)=u_{\max},u^{\prime\prime}(a)-\beta>0\},&\text{if}\;{9\over 8}\kappa_{-}\leq\beta<0,\end{array}\right.

and

(1.20) Mβ\displaystyle M_{\beta} :={infc(,umin)λmβ+1(c),if 0<β98κ+,infc(umax,)λmβ+1(c),if98κβ<0,\displaystyle:=\left\{\begin{array}[]{ll}-\inf\limits_{c\in(-\infty,u_{\min})}\lambda_{m_{\beta}+1}(c),&\text{if}\;0<\beta\leq{9\over 8}\kappa_{+},\\ -\inf\limits_{c\in(u_{\max},\infty)}\lambda_{m_{\beta}+1}(c),&\text{if}\;{9\over 8}\kappa_{-}\leq\beta<0,\end{array}\right.

where λmβ+1(c)\lambda_{m_{\beta}+1}(c) is the (mβ+1)(m_{\beta}+1)-th eigenvalue of the Rayleigh-Kuo BVP (2.6).

𝐑\mathbf{R}, 𝐙\mathbf{Z} and 𝐙+\mathbf{Z}^{+} denote the set of all the real numbers, integers and positive integers, respectively. (K)\sharp(K) or K\sharp\,K is the cardinality of the set KK. Let LL be a linear operator from a Banach space XX to XX. XX^{*} is the dual space of XX. σ(L)\sigma(L), σe(L)\sigma_{e}(L) and σd(L)\sigma_{d}(L) are the spectrum, essential spectrum and discrete spectrum of the operator LL, respectively. For ψL2(DT)\psi\in L^{2}(D_{T}), the Fourier transform of ψ\psi in xx is denoted by ψ^\widehat{\psi}.

2. Extension to general shear flows and outline of the proof

In this section, we first extend the main Theorem 1.2 to general shear flows under mild assumptions, and then discuss our approach in its proof.

2.1. Main results for general shear flows

For a shear flow in H4(y1,y2)H^{4}(y_{1},y_{2}), we give the exact number of traveling wave families near the flow.

Theorem 2.1.

Let α=2π/T\alpha=2\pi/T, β0\beta\neq 0 and uH4(y1,y2)u\in H^{4}(y_{1},y_{2}). Then (k1(σd(kα,β)𝐑))\sharp(\bigcup_{k\geq 1}(\sigma_{d}(\mathcal{R}_{k\alpha,\beta})\cap\mathbf{R})) is exactly the number of traveling wave families near (u,0)(u,0), where kα,β\mathcal{R}_{k\alpha,\beta} is defined in (2.5) and the precise description of a traveling wave family near (u,0)(u,0) is given in Definition 2.8.

Then we state our main theorem for a shear flow satisfying (H1)(\textbf{H1}).

Theorem 2.2.

Let β0\beta\neq 0 and uu satisfy (𝐇𝟏){\bf(H1)}.

  • (1)

    If {u=0}{u=umin}\{u^{\prime}=0\}\cap\{u=u_{\min}\}\neq\emptyset and (𝐄+)\bf{(E_{+})} holds for every α(0,Mβ]{2kπT|k𝐙+}\alpha\in(0,\sqrt{M_{\beta}}]\cap\{{2k\pi\over T}|k\in\mathbf{Z}^{+}\} and β(0,98κ+)\beta\in(0,{9\over 8}\kappa_{+}), then there exists β+(0,)\beta_{+}\in(0,\infty) such that there exist at most finitely many traveling wave families near (u,0)(u,0) for β(0,β+)\beta\in(0,\beta_{+}), and infinitely many traveling wave families near (u,0)(u,0) for β(β+,)\beta\in(\beta_{+},\infty), where κ+\kappa_{+}, (𝐄+)\bf{(E_{+})} and MβM_{\beta} are defined in (1.5), (1.13) and (1.20), respectively. Moreover, β+\beta_{+} is specified in (1.11).

  • (2)

    If {u=0}{u=umax}\{u^{\prime}=0\}\cap\{u=u_{\max}\}\neq\emptyset and (𝐄)\bf{(E_{-})} holds for every α(0,Mβ]{2kπT|k𝐙+}\alpha\in(0,\sqrt{M_{\beta}}]\cap\{{2k\pi\over T}|k\in\mathbf{Z}^{+}\} and β(98κ,0)\beta\in({9\over 8}\kappa_{-},0), then there exists β(,0)\beta_{-}\in(-\infty,0) such that there exist at most finitely many traveling wave families near (u,0)(u,0) for β(β,0)\beta\in(\beta_{-},0), and infinitely many traveling wave families near (u,0)(u,0) for β(,β)\beta\in(-\infty,\beta_{-}), where κ\kappa_{-} and (𝐄)\bf{(E_{-})} are defined in (1.6) and (1.14), respectively. Moreover, β\beta_{-} is specified in (1.12).

Assume that u(y1)u(y2)u(y_{1})\neq u(y_{2}) and for i=1,2i=1,2, there exist δ>0,C>0\delta>0,\ C>0 and mi>0m_{i}>0 such that (i)(\rm{i}) u′′(y)=βiu^{\prime\prime}(y)=\beta_{i} for y(yiδ,yi+δ)[y1,y2]y\in(y_{i}-\delta,y_{i}+\delta)\cap[y_{1},y_{2}] or (ii)(\rm{ii}) C1|yyi|mi|u′′(y)βi|C|yyi|miC^{-1}|y-y_{i}|^{m_{i}}\leq|u^{\prime\prime}(y)-\beta_{i}|\leq C|y-y_{i}|^{m_{i}} for y(yiδ,yi+δ)[y1,y2]y\in(y_{i}-\delta,y_{i}+\delta)\cap[y_{1},y_{2}] or (iii)(\rm{iii}) βiu(yi)(1)i0\beta_{i}u^{\prime}(y_{i})(-1)^{i}\geq 0, where βi=u′′(yi)\beta_{i}=u^{\prime\prime}(y_{i}).

  • (3)

    If {u=0}{u=umin}=\{u^{\prime}=0\}\cap\{u=u_{\min}\}=\emptyset, then there exist at most finitely many traveling wave families near (u,0)(u,0) for β(0,)\beta\in(0,\infty).

  • (4)

    If {u=0}{u=umax}=\{u^{\prime}=0\}\cap\{u=u_{\max}\}=\emptyset, then there exist at most finitely many traveling wave families near (u,0)(u,0) for β(,0)\beta\in(-\infty,0).

Here, the precise description of a traveling wave family near (u,0)(u,0) is given in Definition 2.8.

As mentioned in Introduction, (𝐄+)\bf{(E_{+})} or (𝐄)\bf{(E_{-})} is “one spectral point” assumption for small wave numbers. Note that if 2πT>Mβ{2\pi\over T}>\sqrt{M_{\beta}}, then (0,Mβ]{2kπT|k𝐙+}=(0,\sqrt{M_{\beta}}]\cap\{{2k\pi\over T}|k\in\mathbf{Z}^{+}\}=\emptyset, and (𝐄±)\bf{(E_{\pm})} is not needed in Theorem 2.2 (1)–(2). One of the conditions (i)–(iii) is the “good” endpoints assumption and rather generic. For example, if uCm([y1,y2])u\in C^{m}([y_{1},y_{2}]), m3m\geq 3 and u(ki)(yi)0u^{(k_{i})}(y_{i})\neq 0 for some 3kim3\leq k_{i}\leq m, then (ii)(\rm{ii}) is true for mi=ki2m_{i}=k_{i}-2. Thus, for analytic flows, (ii) holds if u(ki)(yi)0u^{(k_{i})}(y_{i})\neq 0 for some ki3k_{i}\geq 3 and (i) holds otherwise. Applying Theorem 2.2 (3)-(4) to analytic monotone flows, we have the following result.

Corollary 2.3.

Let uu be an analytic monotone flow: u(y)0u^{\prime}(y)\neq 0 for y[y1,y2]y\in[y_{1},y_{2}]. Then there exist at most finitely many traveling wave families near (u,0)(u,0) for β0\beta\neq 0.

2.2. Outline and our approach in the proof

Non-parallel steady flows or traveling waves may be bifurcated from a shear flow if the linearized Euler operator has an embedding or isolated real eigenvalues [1, 29, 28]. Based on the existence of an embedding eigenvalue for a class of monotone shear flows near Couette flow, cat’s eyes steady states are bifurcated from these flows [29]. When the Coriolis force is involved, non-parallel traveling waves are bifurcated from the sinus profile on account of the existence of an isolated real eigenvalue [28]. The traveling speeds lie outside the range of the sinus profile and are contiguous to the isolated real eigenvalue. Now, we consider such bifurcation theorem for general shear flows, namely, using an isolated real eigenvalue of the linearized Euler operator, we prove that such traveling waves can be bifurcated from general shear flows. We use the following concept.

Definition 2.4.

{uε(xcεt,y)=(uε(xcεt,y),vε(xcεt,y))|ε(0,ε0) for some ε0>0}\{\vec{u}_{\varepsilon}\left(x-c_{\varepsilon}t,y\right)=(u_{\varepsilon}\left(x-c_{\varepsilon}t,y\right),v_{\varepsilon}\left(x-c_{\varepsilon}t,y\right))|\varepsilon\in(0,\varepsilon_{0})\text{ for some }\varepsilon_{0}>0\} is called a set of traveling wave solutions near (u,0)(u,0) with traveling speeds converging to c0c_{0}, if for each ε(0,ε0)\varepsilon\in(0,\varepsilon_{0}), uε(xcεt,y)=(uε(xcεt,y),vε(xcεt,y))\vec{u}_{\varepsilon}\left(x-c_{\varepsilon}t,y\right)=(u_{\varepsilon}\left(x-c_{\varepsilon}t,y\right),v_{\varepsilon}\left(x-c_{\varepsilon}t,y\right)) is a traveling wave solution to (1.1)–(1.2) which has period TT in xx such that

(2.1) (uε,vε)(u,0)H3(DT)ε,\displaystyle\|(u_{\varepsilon},v_{\varepsilon})-(u,0)\|_{H^{3}(D_{T})}\leq\varepsilon,

vεL2(DT)0\|v_{\varepsilon}\|_{L^{2}\left(D_{T}\right)}\neq 0, cεRan(u)c_{\varepsilon}\notin\text{Ran}(u) and cεc0c_{\varepsilon}\rightarrow c_{0}.

Then we give the bifurcation result for general shear flows.

Lemma 2.5.

Let α=2π/T\alpha=2\pi/T, β0\beta\neq 0 and uH4(y1,y2)u\in H^{4}(y_{1},y_{2}). Assume that c0k1(σd(kα,β)𝐑)c_{0}\in\bigcup_{k\geq 1}(\sigma_{d}(\mathcal{R}_{k\alpha,\beta})\cap\mathbf{R}), where kα,β\mathcal{R}_{k\alpha,\beta} is defined in (2.5). Then there exists a set of traveling wave solutions near (u,0)(u,0) with traveling speeds converging to c0c_{0}. Moreover, we have uε(x,y)cε0u_{\varepsilon}\left(x,y\right)-c_{\varepsilon}\neq 0.

Here, we mention some differences from the construction of traveling waves in the literature. First, the horizontal period of constructed traveling waves in Proposition 7 of [28] is not the given period TT, and for the sinus profile, the period of traveling waves is modified to TT by adjusting the traveling speed in Theorem 7 of [28]. But the price is an additional condition, namely, the isolated eigenvalue c0c_{0} can not be an extreme point of λ1\lambda_{1} (i.e. α0Λβ\alpha_{0}\neq\sqrt{\Lambda_{\beta}} in Theorem 7 (ii) of [28]), where λn\lambda_{n} is the nn-th eigenvalue of (2.6). In Lemma 2.5, we can construct traveling waves for general flows no matter whether c0c_{0} is an extreme point of λn\lambda_{n}, and thus improve the result in Theorem 7 of [28] even for the sinus profile. Second, it is possible that c0σd(0,β)𝐑c_{0}\in\sigma_{d}(\mathcal{R}_{0,\beta})\cap\mathbf{R}, which makes it subtle to guarantee that the bifurcated solutions near the flow uu is not a shear flow. Thus, the extension of the bifurcation result for the sinus profile in [28] to general shear flows in Lemma 2.5 is still non-trivial, since we have to treat the unsolved case that c0c_{0} is an extreme point of λn0\lambda_{n_{0}} for some n0𝐙+n_{0}\in\mathbf{Z}^{+} or c0σd(0,β)𝐑c_{0}\in\sigma_{d}(\mathcal{R}_{0,\beta})\cap\mathbf{R}. To overcome the difficulty, we carefully modify the flow uu to a suitable shear flow, which satisfies that λn0\lambda_{n_{0}} is locally monotone near c0c_{0} and c0σd(0,β)𝐑c_{0}\notin\sigma_{d}(\mathcal{R}_{0,\beta})\cap\mathbf{R}, and then study the bifurcation at the suitable shear flow. Finally, the minimal horizontal periods of constructed traveling waves are possibly less than 2π/α{2\pi/\alpha} if c0σd(α,β)𝐑.c_{0}\in\sigma_{d}(\mathcal{R}_{\alpha,\beta})\cap\mathbf{R}. In fact, the Sturm-Liouville operator c0\mathcal{L}_{c_{0}} could indeed have more than one negative eigenvalues (e.g., if κ+<\kappa_{+}<\infty, β>9κ+/8\beta>{9\kappa_{+}/8} and c0c_{0} is close to uminu_{\min}), where c0\mathcal{L}_{c_{0}} is defined in (2.6). In this case, we give sufficient condition to guarantee that the minimal period is 2π/α{2\pi/\alpha} in Lemma 5.3. In contrast, the minimal period must be 2π/α{2\pi/\alpha} in Theorem 5.1 of [26] and Theorem 1 of [29], since the corresponding Sturm-Liouville operator has only one negative eigenvalue.

Since the isolated real eigenvalue c0c_{0} lies outside the range of the flow uu, by a similar proof of Lemma 2.5 we can improve the regularity of traveling waves as follows.

Corollary 2.6.

Let α=2π/T\alpha=2\pi/T, β0\beta\neq 0, uC([y1,y2])u\in C^{\infty}([y_{1},y_{2}]) and s3s\geq 3. Assume that c0k1(σd(kα,β)𝐑)c_{0}\in\bigcup_{k\geq 1}(\sigma_{d}(\mathcal{R}_{k\alpha,\beta})\cap\mathbf{R}), where kα,β\mathcal{R}_{k\alpha,\beta} is defined in (2.5). Then the conclusion in Lemma 2.5 holds true with (2.1) replaced by (uε,vε)(u,0)Hs(DT)ε.\|(u_{\varepsilon},v_{\varepsilon})-(u,0)\|_{H^{s}(D_{T})}\leq\varepsilon.

One naturally asks whether the assumption c0k1(σd(kα,β)𝐑)c_{0}\in\bigcup_{k\geq 1}(\sigma_{d}(\mathcal{R}_{k\alpha,\beta})\cap\mathbf{R}) in Lemma 2.5 is necessary. By studying the asymptotic behavior of traveling speeds and L2L^{2} normalized vertical velocities for nearby traveling waves, we confirm that it is true.

Lemma 2.7.

Let α=2π/T\alpha=2\pi/T, β0\beta\neq 0 and uH4(y1,y2)u\in H^{4}(y_{1},y_{2}). Assume that {uε(xcεt,y)=(uε(xcεt,y),vε(xcεt,y))|ε(0,ε0)}\{\vec{u}_{\varepsilon}\left(x-c_{\varepsilon}t,y\right)=(u_{\varepsilon}\left(x-c_{\varepsilon}t,y\right),v_{\varepsilon}\left(x-c_{\varepsilon}t,y\right))|\varepsilon\in(0,\varepsilon_{0})\} is a set of traveling wave solutions near (u,0)(u,0) with traveling speeds converging to c0c_{0}. Then c0k1(σd(kα,β)𝐑){umin,umax},c_{0}\in\bigcup_{k\geq 1}(\sigma_{d}(\mathcal{R}_{k\alpha,\beta})\cap\mathbf{R})\cup\{u_{\min},u_{\max}\}, where kα,β\mathcal{R}_{k\alpha,\beta} is defined in (2.5). Moreover, if c0k1(σd(kα,β)𝐑)c_{0}\in\bigcup_{k\geq 1}(\sigma_{d}(\mathcal{R}_{k\alpha,\beta})\cap\mathbf{R}), then there exists φc0ker(𝒢c0)\varphi_{c_{0}}\in\ker(\mathcal{G}_{c_{0}}) such that

(2.2) v~εφc0inH2(DT),\displaystyle{\tilde{v}_{\varepsilon}}\longrightarrow\varphi_{c_{0}}\ \ \text{in}\ \ {H^{2}\left(D_{T}\right)},

where the operator 𝒢c0\mathcal{G}_{c_{0}} is defined by

(2.3) 𝒢c0=Δβu′′(y)u(y)c0:H2(DT)L2(DT)\displaystyle\mathcal{G}_{c_{0}}=-\Delta-\frac{\beta-u^{\prime\prime}\left(y\right)}{u\left(y\right)-c_{0}}\;:\;H^{2}(D_{T})\to L^{2}(D_{T})

with periodic boundary condition in xx and Dirichlet boundary condition in yy, and v~ε=vε/vεL2(DT)\tilde{v}_{\varepsilon}={{v}_{\varepsilon}/\|{v}_{\varepsilon}\|_{L^{2}\left(D_{T}\right)}}.

The limit function φc0\varphi_{c_{0}} in Lemma 2.7 is a superposition of finite normal modes, see Remark 5.1. If c0k1(σd(kα,β)𝐑)c_{0}\in\bigcup_{k\geq 1}(\sigma_{d}(\mathcal{R}_{k\alpha,\beta})\cap\mathbf{R}) in Lemma 2.7, the vertical velocities of the nearby traveling waves have simple asymptotic behavior as seen in (2.2). However, if c0{umin,umax}c_{0}\in\{u_{\min},u_{\max}\}, then the asymptotic behavior of vertical velocities might be complicated, see Remark 5.2. The proofs of Lemmas 2.5-2.7 are given in Section 5.

By Lemma 2.7, for any set of traveling waves near (u,0)(u,0) with traveling speeds converging to c0c_{0}, c0c_{0} must be an isolated real eigenvalue of the linearized Euler operator (besides uminu_{\min} and umaxu_{\max}). By Lemma 2.5, every isolated real eigenvalue is contiguous to the speeds of nearby traveling waves. As the minimal periods of traveling waves in xx can be less than 2π/α{2\pi/\alpha}, there might be two or more sets of traveling wave solutions near (u,0)(u,0) with traveling speeds converging to a same isolated real eigenvalue. For example, if ((i+1)α)2=λni(c0)((i+1)\alpha)^{2}=-\lambda_{n_{i}}(c_{0}), λni\lambda_{n_{i}} is monotone near c0c_{0} for i=1,2i=1,2, and (kα)2λn(c0)(k\alpha)^{2}\neq\lambda_{n}(c_{0}) for k{2,3}k\notin\{2,3\} and n{n1,n2}n\notin\{n_{1},n_{2}\}, then an application to Lemma 2.5 (see Case 1 in its proof) gives two sets of traveling wave solutions, which has minimal periods π/α\pi/\alpha and 2π/(3α)2\pi/(3\alpha) respectively, near (u,0)(u,0) with traveling speeds converging to c0c_{0}. Moreover, traveling wave solutions could be bifurcated from nearby shear flows, which might induce more sets of traveling wave solutions near (u,0)(u,0) with traveling speeds converging to c0c_{0}. This suggests us to define a traveling wave family near (u,0)(u,0) by an equivalence class as follows.

Definition 2.8.

A traveling wave family near (u,0)(u,0) is defined by an equivalence class under \sim, where if {ui,ε(xci,εt,y)=(ui,ε(xci,εt,y),vi,ε(xci,εt,y))|ε(0,εi)}\{\vec{u}_{i,\varepsilon}\left(x-c_{i,\varepsilon}t,y\right)=(u_{i,\varepsilon}\left(x-c_{i,\varepsilon}t,y\right),v_{i,\varepsilon}\left(x-c_{i,\varepsilon}t,y\right))|\varepsilon\in(0,\varepsilon_{i})\}, i=1,2i=1,2, are two sets of traveling wave solutions near (u,0)(u,0) with traveling speeds converging to ciRan(u)c_{i}\notin\text{Ran}(u), then {u1,ε(xc1,εt,y)|ε(0,ε1)}\{\vec{u}_{1,\varepsilon}\left(x-c_{1,\varepsilon}t,y\right)|\varepsilon\in(0,\varepsilon_{1})\} and {u2,ε(xc2,εt,y)|ε(0,ε2)}\{\vec{u}_{2,\varepsilon}\left(x-c_{2,\varepsilon}t,y\right)|\varepsilon\in(0,\varepsilon_{2})\} are equivalent, {u1,ε(xc1,εt,y)|ε(0,ε1)}{u2,ε(xc2,εt,y)|ε(0,ε2)}\{\vec{u}_{1,\varepsilon}\left(x-c_{1,\varepsilon}t,y\right)|\varepsilon\in(0,\varepsilon_{1})\}\sim\{\vec{u}_{2,\varepsilon}\left(x-c_{2,\varepsilon}t,y\right)|\varepsilon\in(0,\varepsilon_{2})\}, if c1=c2.c_{1}=c_{2}.

By Lemma 2.7, there exists φiker(𝒢ci)\varphi_{i}\in\ker(\mathcal{G}_{c_{i}}) such that v~i,εφi{\tilde{v}_{i,\varepsilon}}\longrightarrow\varphi_{i} in H2(DT){H^{2}\left(D_{T}\right)}, where v~i,ε=vi,ε/vi,εL2(DT)\tilde{v}_{i,\varepsilon}={{v}_{i,\varepsilon}/\|{v}_{i,\varepsilon}\|_{L^{2}\left(D_{T}\right)}} and vi,ε,ε(0,εi),{v}_{i,\varepsilon},\varepsilon\in(0,\varepsilon_{i}), are given in Definition 2.8. By Lemmas 2.5 and 2.7, we obtain the exact number of traveling wave families near a flow uH4(y1,y2)u\in H^{4}(y_{1},y_{2}) in Theorem 2.1.

Thus, the number of isolated real eigenvalues of the linearized Euler operator plays an important role in counting the traveling wave families near the shear flow. In terms of the stream function ψ\psi, (1.4) can be written as tΔψ+uxΔψ+(βu′′)xψ=0.\partial_{t}\Delta\psi+u\partial_{x}\Delta\psi+(\beta-u^{\prime\prime})\partial_{x}\psi=0. By taking Fourier transform in xx, we have

(2.4) (y2α2)tψ^=iα((u′′β)u(y2α2))ψ^.\displaystyle(\partial^{2}_{y}-\alpha^{2})\partial_{t}\widehat{\psi}=i\alpha((u^{\prime\prime}-\beta)-u(\partial^{2}_{y}-\alpha^{2}))\widehat{\psi}.

For α>0\alpha>0 and β𝐑\beta\in\mathbf{R}, the linearized Euler operator is given by

(2.5) α,β:=(y2α2)1((u′′β)u(y2α2)).\displaystyle\mathcal{R}_{\alpha,\beta}:=-(\partial^{2}_{y}-\alpha^{2})^{-1}((u^{\prime\prime}-\beta)-u(\partial^{2}_{y}-\alpha^{2})).

Then (2.4) becomes 1iαtψ^=α,βψ^.-\frac{1}{i\alpha}\partial_{t}\widehat{\psi}=\mathcal{R}_{\alpha,\beta}\widehat{\psi}. Recall that σe(α,β)=Ran(u)\sigma_{e}(\mathcal{R}_{\alpha,\beta})=\text{Ran}(u). Then the set of isolated real eigenvalues σd(α,β)𝐑(,umin)(umax,)\sigma_{d}(\mathcal{R}_{\alpha,\beta})\cap\mathbf{R}\subset(-\infty,u_{\min})\cup(u_{\max},\infty). Moreover, it is well-known that σd(α,β)𝐑=\sigma_{d}(\mathcal{R}_{\alpha,\beta})\cap\mathbf{R}=\emptyset if β=0\beta=0; σd(α,β)(umax,)=\sigma_{d}(\mathcal{R}_{\alpha,\beta})\cap(u_{\max},\infty)=\emptyset if β>0\beta>0; and σd(α,β)(,umin)=\sigma_{d}(\mathcal{R}_{\alpha,\beta})\cap(-\infty,u_{\min})=\emptyset if β<0\beta<0, see [23, 43, 36, 28]. Therefore, we only need to study (σd(α,β)(,umin))\sharp(\sigma_{d}(\mathcal{R}_{\alpha,\beta})\cap(-\infty,u_{\min})) for β>0\beta>0 and (σd(α,β)(umax,))\sharp(\sigma_{d}(\mathcal{R}_{\alpha,\beta})\cap(u_{\max},\infty)) for β<0\beta<0. We mainly study (σd(α,β)(,umin))\sharp(\sigma_{d}(\mathcal{R}_{\alpha,\beta})\cap(-\infty,u_{\min})) for β>0\beta>0, since the other is similar. cσd(α,β)c\in\sigma_{d}(\mathcal{R}_{\alpha,\beta}) if and only if its corresponding eigenfunction ψc\psi_{c} satisfies the Rayleigh-Kuo BVP:

(2.6) cϕ:=ϕ′′+u′′βucϕ=λϕ,ϕ(y1)=ϕ(y2)=0,\displaystyle\mathcal{L}_{c}\phi:=-\phi^{\prime\prime}+{u^{\prime\prime}-\beta\over u-c}\phi=\lambda\phi,\;\;\;\;\phi(y_{1})=\phi(y_{2})=0,

where ϕH01H2(y1,y2)\phi\in H_{0}^{1}\cap H^{2}(y_{1},y_{2}) and λ=α2\lambda=-\alpha^{2}. This equation is formulated by Kuo [23]. For c<uminc<u_{\min}, it follows from [40] that the nn-th eigenvalue of (2.6) is

(2.7) λn(c)=\displaystyle\lambda_{n}(c)= infdimVn=nsupϕH01,ϕVn,ϕL2=1cϕ,ϕ\displaystyle\inf_{\dim V_{n}=n}\;\;\sup_{\phi\in H_{0}^{1},\phi\in V_{n},\|\phi\|_{L^{2}}=1}\langle\mathcal{L}_{c}\phi,\phi\rangle
=\displaystyle= infdimVn=nsupϕH01,ϕVn,ϕL2=1y1y2(|ϕ|2+u′′βuc|ϕ|2)𝑑y.\displaystyle\inf_{\dim V_{n}=n}\;\;\sup_{\phi\in H_{0}^{1},\phi\in V_{n},\|\phi\|_{L^{2}}=1}\int_{y_{1}}^{y_{2}}\left(|\phi^{\prime}|^{2}+{u^{\prime\prime}-\beta\over u-c}|\phi|^{2}\right)dy.

In this way, we have

σd(α,β)(,umin)=n1{c<umin:λn(c)=α2}.\displaystyle\sigma_{d}(\mathcal{R}_{\alpha,\beta})\cap(-\infty,u_{\min})=\bigcup_{n\geq 1}\{c<u_{\min}:\lambda_{n}(c)=-\alpha^{2}\}.

To determine whether (σd(α,β)(,umin))\sharp(\sigma_{d}(\mathcal{R}_{\alpha,\beta})\cap(-\infty,u_{\min})) is finite, we need to study the number of solutions c<uminc<u_{\min} such that λn(c)=α2\lambda_{n}(c)=-\alpha^{2} for n1n\geq 1. Since limcλn(c)=n24π2>0\lim_{c\to-\infty}\lambda_{n}(c)={{n^{2}}\over 4}\pi^{2}>0 by Proposition 4.2 in [28] and λn(c)\lambda_{n}(c) is real-analytic on (,umin)(-\infty,u_{\min}), the only possibility such that (σd(α,β)(,umin))=\sharp(\sigma_{d}(\mathcal{R}_{\alpha,\beta})\cap(-\infty,u_{\min}))=\infty is that there exists a sequence {cj(α,β)}j=1n1{c<umin:λn(c)=α2}\{c_{j}(\alpha,\beta)\}_{j=1}^{\infty}\subset\bigcup_{n\geq 1}\{c<u_{\min}:\lambda_{n}(c)=-\alpha^{2}\} such that cj(α,β)uminc_{j}(\alpha,\beta)\to u_{\min}^{-}. Thus, the key is to study the asymptotic behavior of λn(c)\lambda_{n}(c) as cuminc\to u_{\min}^{-}. We divide it into two steps.
Step 1. We study how many nn’s such that λn(c)\lambda_{n}(c)\to-\infty as cuminc\to u_{\min}^{-}. We determine a transitional β\beta value such that the number {n1:λn(c) as cumin}\sharp\{n\geq 1:\lambda_{n}(c)\to-\infty\text{ as }c\to u_{\min}^{-}\} changes suddenly from finite to infinite when β\beta passes through it.

Theorem 2.9.

Let uu satisfy (𝐇𝟏){\bf(H1)}. (1)(1) Let 0<β98κ+0<\beta\leq{9\over 8}\kappa_{+}. Then

  • (i)

    limcuminλn(c)=,  1nmβ;\lim\limits_{c\to u_{\min}^{-}}\lambda_{n}(c)=-\infty,\;\;1\leq n\leq m_{\beta};

  • (ii)

    Mβ<;M_{\beta}<\infty;

  • (iii)

    there exists an integer Nβ>mβN_{\beta}>m_{\beta} such that infc(,umin)λNβ(c)0.\inf\limits_{c\in(-\infty,u_{\min})}\lambda_{N_{\beta}}(c)\geq 0.

(2)(2) Let 98κβ<0{9\over 8}\kappa_{-}\leq\beta<0. Then

  • (i)

    limcumax+λn(c)=,  1nmβ;\lim\limits_{c\to u_{\max}^{+}}\lambda_{n}(c)=-\infty,\;\;1\leq n\leq m_{\beta};

  • (ii)

    Mβ<;M_{\beta}<\infty;

  • (iii)

    there exists an integer Nβ>mβN_{\beta}>m_{\beta} such that infc(umax,)λNβ(c)0.\inf\limits_{c\in(u_{\max},\infty)}\lambda_{N_{\beta}}(c)\geq 0.

(3)(3) Let β>98κ+\beta>{9\over 8}\kappa_{+}. Then limcuminλn(c)=\lim_{c\to u_{\min}^{-}}\lambda_{n}(c)=-\infty for n1n\geq 1.
(4)(4) Let β<98κ\beta<{9\over 8}\kappa_{-}. Then limcumax+λn(c)=\lim_{c\to u_{\max}^{+}}\lambda_{n}(c)=-\infty for n1n\geq 1.
Here, κ±\kappa_{\pm}, mβm_{\beta} and MβM_{\beta} are defined in (1.5)–(1.8), (1.17) and (1.20), respectively.

The transitional value β=98κ+\beta={9\over 8}\kappa_{+} is illustrated in Figure 1. We give a simple example to explain why such a transitional value exists. Consider the flow u=12y2u={1\over 2}y^{2} on [0,1][0,1] and β>0\beta>0. If c<0c<0 is very close to 0, then the energy quadratic form in (2.7) roughly looks like

cϕ,ϕ01|ϕ|2+22βy2|ϕ|2dy.\displaystyle\langle\mathcal{L}_{c}\phi,\phi\rangle\sim\int_{0}^{1}|\phi^{\prime}|^{2}+{2-2\beta\over y^{2}}|\phi|^{2}dy.

Thus, if 22β>14β<982-2\beta>-{1\over 4}\Leftrightarrow\beta<{9\over 8}, by Hardy type inequality (Lemma 3.1) we have cϕ,ϕ\langle\mathcal{L}_{c}\phi,\phi\rangle is bounded from below for any test functions ϕ\phi with ϕ(0)=0\phi(0)=0. From this formal observation, we may expect λ1(c)\lambda_{1}(c) is bounded from below. If 22β<14β>982-2\beta<-{1\over 4}\Leftrightarrow\beta>{9\over 8}, cϕ,ϕ\langle\mathcal{L}_{c}\phi,\phi\rangle is unbounded from below by looking at the test functions y12+εy^{{1\over 2}+\varepsilon} with ε0+\varepsilon\to 0^{+}. We will construct test functions motivated by the function y12y^{1\over 2} to show that all the eigenvalues are unbounded from below.

[Uncaptioned image]  [Uncaptioned image]
    0<β98κ+0<\beta\leq{9\over 8}\kappa_{+}                       β>98κ+\beta>{9\over 8}\kappa_{+}

Figure 1.

For general flows, the main idea in the proof of Theorem 2.9 (1)-(2) is to control u′′βuc|ϕ|2𝑑y\int{u^{\prime\prime}-\beta\over u-c}|\phi|^{2}dy using the L2L^{2} norm of ϕ\phi^{\prime} near a singular point (see Lemma 3.2), which involves very delicate and careful localized analysis. The transitional β\beta values 98κ±{9\over 8}\kappa_{\pm} are essentially due to the optimal Hardy type inequality (3.1). The idea in the proof of Theorem 2.9 (3)-(4) is to construct suitable test functions such that the functional in (2.7) converges to -\infty as cuminc\to u_{\min}^{-} or umax+u_{\max}^{+}, see (3.22). This is inspired by the “eigenfunction” y12y^{1\over 2} for the optimal Hardy type equality and a support-separated technique. The proof of Theorem 2.9 is given in Section 3.

Then we give sharp criteria for λ1(c)\lambda_{1}(c)\to-\infty as cuminc\to u_{\min}^{-} if β[98κ,98κ+]\beta\in[{9\over 8}\kappa_{-},{9\over 8}\kappa_{+}]. By Theorem 2.1, the number of traveling wave families is to count the union of (σd(kα,β)𝐑)\sharp(\sigma_{d}(\mathcal{R}_{k\alpha,\beta})\cap\mathbf{R}) for all k1k\geq 1. Thus, the number of traveling wave families is infinity provided that λ1(c)\lambda_{1}(c)\to-\infty as cuminc\to u_{\min}^{-}. By Theorem 2.9 (1)-(2), we get the sharp criteria for λ1(c)\lambda_{1}(c)\to-\infty as cuminc\to u_{\min}^{-}.

Corollary 2.10.

Let uu satisfy (𝐇𝟏){\bf(H1)}.

  • (1)

    If {u=umin}(y1,y2)\{u=u_{\min}\}\cap(y_{1},y_{2})\neq\emptyset, then a transitional β\beta value min{98κ+,μ+}\min\{{9\over 8}\kappa_{+},\mu_{+}\} exists in (0,98κ+](0,{9\over 8}\kappa_{+}] such that infc(,umin)λ1(c)>\inf_{c\in(-\infty,u_{\min})}\lambda_{1}(c)>-\infty for β(0,min{98κ+,μ+}]\beta\in(0,\min\{{9\over 8}\kappa_{+},\mu_{+}\}] and limcumin\lim_{c\to u_{\min}^{-}} λ1(c)=\lambda_{1}(c)=-\infty for β(min{98κ+,μ+},98κ+]\beta\in(\min\{{9\over 8}\kappa_{+},\mu_{+}\},{9\over 8}\kappa_{+}].

  • (2)

    If {u=umax}(y1,y2)\{u=u_{\max}\}\cap(y_{1},y_{2})\neq\emptyset, then a transitional β\beta value max{98κ,μ}\max\{{9\over 8}\kappa_{-},\mu_{-}\} exists in [98κ,0)[{9\over 8}\kappa_{-},0) such that infc(umax,)λ1(c)>\inf_{c\in(u_{\max},\infty)}\lambda_{1}(c)>-\infty for β[max{98κ,μ},0)\beta\in[\max\{{9\over 8}\kappa_{-},\mu_{-}\},0) and limcumax+λ1(c)=\lim_{c\to u_{\max}^{+}}\\ \lambda_{1}(c)=-\infty for β[98κ,max{98κ,μ})\beta\in[{9\over 8}\kappa_{-},\max\{{9\over 8}\kappa_{-},\mu_{-}\}).

  • (3)

    If {u=umin}(y1,y2)=\{u=u_{\min}\}\cap(y_{1},y_{2})=\emptyset, then infc(,umin)λ1(c)>\inf_{c\in(-\infty,u_{\min})}\lambda_{1}(c)>-\infty for β(0,98κ+]\beta\in(0,{9\over 8}\kappa_{+}].

  • (4)

    If {u=umax}(y1,y2)=\{u=u_{\max}\}\cap(y_{1},y_{2})=\emptyset, then infc(umax,)λ1(c)>\inf_{c\in(u_{\max},\infty)}\lambda_{1}(c)>-\infty for β[98κ,0)\beta\in[{9\over 8}\kappa_{-},0).

Here, κ±\kappa_{\pm} and μ±\mu_{\pm} are defined in (1.5)–(1.10).

Here, a key point for Corollary 2.10 (1) and (3) is that infc(,umin)λ1(c)>\inf_{c\in(-\infty,u_{\min})}\lambda_{1}(c)>-\infty if and only if mβ=0m_{\beta}=0 and β98κ+.\beta\leq{9\over 8}\kappa_{+}.

Step 2. We rule out the oscillation of λn(c)\lambda_{n}(c) as cuminc\to u_{\min}^{-} (or cumax+c\to u_{\max}^{+}). By Theorem 2.9 (1), we get that for 1nmβ1\leq n\leq m_{\beta}, λn(c)=α2\lambda_{n}(c)=-\alpha^{2} has only finite number of solutions cc on (,umin)(-\infty,u_{\min}). Moreover, if nNβn\geq N_{\beta}, no solutions exist for λn(c)=α2\lambda_{n}(c)=-\alpha^{2} on c(,umin)c\in(-\infty,u_{\min}). Now, we consider whether ({λn(c)=α2,c(,umin)})<\sharp(\{\lambda_{n}(c)=-\alpha^{2},c\in(-\infty,u_{\min})\})<\infty for mβ<n<Nβm_{\beta}<n<N_{\beta}. Indeed, we rule out the oscillation of λn(c)\lambda_{n}(c) under the spectral assumption (𝐄±)\bf{(E_{\pm})}, or under the “good” endpoints assumption (i.e. one of the conditions (i)–(iii) in Theorem 2.2), or for flows in class 𝒦+\mathcal{K}^{+}. The oscillation of λn(c)\lambda_{n}(c) is illustrated in Figure 2.

[Uncaptioned image]

Figure 2.

Case 1. Under the spectral assumption, the main argument to rule out oscillation is to prove uniform H1H^{1} bound for corresponding eigenfunctions, and the proof is in Subsection 4.1. In this case, 98κ±{9\over 8}\kappa_{\pm} are also transitional β\beta values for the number of isolated real eigenvalues of the linearized Euler operator if |κ±|<|\kappa_{\pm}|<\infty.

Theorem 2.11.

Assume that uu satisfies (𝐇𝟏)\bf{(H1)} and α>0\alpha>0.

  • (1)

    If 0<β<98κ+0<\beta<{9\over 8}\kappa_{+}, 0<α2Mβ0<\alpha^{2}\leq M_{\beta} and (𝐄+)\bf{(E_{+})} holds for this α\alpha, then

    (2.8) mβ(σd(α,β)(,umin))<.\displaystyle m_{\beta}\leq\sharp(\sigma_{d}(\mathcal{R}_{\alpha,\beta})\cap(-\infty,u_{\min}))<\infty.

    If 0<β98κ+0<\beta\leq{9\over 8}\kappa_{+}, then (2.8) holds for α2>Mβ\alpha^{2}>M_{\beta}.

  • (2)

    If 98κ<β<0{9\over 8}\kappa_{-}<\beta<0, 0<α2Mβ0<\alpha^{2}\leq M_{\beta} and (𝐄)\bf{(E_{-})} holds for this α\alpha, then

    (2.9) mβ(σd(α,β)(umax,))<.\displaystyle m_{\beta}\leq\sharp(\sigma_{d}(\mathcal{R}_{\alpha,\beta})\cap(u_{\max},\infty))<\infty.

    If 98κβ<0{9\over 8}\kappa_{-}\leq\beta<0, then (2.9) holds for α2>Mβ\alpha^{2}>M_{\beta}.

  • (3)

    If β>98κ+\beta>{9\over 8}\kappa_{+}, then (σd(α,β)(,umin))=\sharp(\sigma_{d}(\mathcal{R}_{\alpha,\beta})\cap(-\infty,u_{\min}))=\infty.

  • (4)

    If β<98κ\beta<{9\over 8}\kappa_{-}, then (σd(α,β)(umax,))=\sharp(\sigma_{d}(\mathcal{R}_{\alpha,\beta})\cap(u_{\max},\infty))=\infty.

Here, κ±\kappa_{\pm}, (𝐄±)\bf{(E_{\pm})}, mβm_{\beta} and MβM_{\beta} are defined in (1.5)–(1.8), (1.13)–(1.14), (1.17) and (1.20), respectively.

In fact, by Theorem 2.9 we have Mβ<M_{\beta}<\infty for 0<β98κ+0<\beta\leq{9\over 8}\kappa_{+} or 98κβ<0{9\over 8}\kappa_{-}\leq\beta<0. Here, we focus on sufficient conditions of (2.8) and (2.9), it is unclear whether (2.8) is true for the case β=98κ+\beta={9\over 8}\kappa_{+} with 0<α2Mβ0<\alpha^{2}\leq M_{\beta}, or the case 0<β<98κ+0<\beta<{9\over 8}\kappa_{+} with 0<α2Mβ0<\alpha^{2}\leq M_{\beta} but no assumption (𝐄+)\bf{(E_{+})}.

Note that Theorem 2.11 (3)-(4) is a direct consequence of Theorem 2.9 (3)-(4).
Case 2. Under the “good” endpoints assumption (i.e. one of the conditions (i)(\rm{i})(iii)(\rm{iii}) in Theorem 2.2), a delicate analysis near the endpoints is involved to rule out oscillation, and the proof is in Subsection 4.2. In this case, we get that no transitional β\beta values exist if |κ±|=|\kappa_{\pm}|=\infty.

Theorem 2.12.

Let α>0\alpha>0 and uu satisfy (𝐇𝟏){\bf(H1)}. Assume that u(y1)u(y2)u(y_{1})\neq u(y_{2}), and one of the conditions (i)(\rm{i})(iii)(\rm{iii}) in Theorem 2.2 holds. Then

  • (1)

    (σd(α,β)𝐑)<\sharp(\sigma_{d}(\mathcal{R}_{\alpha,\beta})\cap\mathbf{R})<\infty for all β(0,)\beta\in(0,\infty) if and only if {u=0}{u=umin}=\{u^{\prime}=0\}\cap\{u=u_{\min}\}=\emptyset;

  • (2)

    (σd(α,β)𝐑)<\sharp(\sigma_{d}(\mathcal{R}_{\alpha,\beta})\cap\mathbf{R})<\infty for all β(,0)\beta\in(-\infty,0) if and only if {u=0}{u=umax}=\{u^{\prime}=0\}\cap\{u=u_{\max}\}=\emptyset.

Consequently, (σd(α,β)𝐑)<\sharp(\sigma_{d}(\mathcal{R}_{\alpha,\beta})\cap\mathbf{R})<\infty for all β𝐑\beta\in\mathbf{R} if and only if {u=0}{u=umin}={u=0}{u=umax}=\{u^{\prime}=0\}\cap\{u=u_{\min}\}=\{u^{\prime}=0\}\cap\{u=u_{\max}\}=\emptyset.

Note that if α>Mβ\alpha>\sqrt{M_{\beta}}, then (2.8) and (2.9) are true, and the “good” endpoints assumption (i.e. one of the conditions (i)(\rm{i})(iii)(\rm{iii}) in Theorem 2.2) is not needed in Theorem 2.12. Consequently, if 2πT>Mβ{2\pi\over T}>\sqrt{M_{\beta}}, then Theorem 2.2 (3)–(4) hold true without this assumption (see their proof).

Let uu be an analytic monotone flow and α>0\alpha>0. Then (σd(α,β)𝐑)<\sharp(\sigma_{d}(\mathcal{R}_{\alpha,\beta})\cap\mathbf{R})<\infty for β0\beta\neq 0. This is a corollary of Theorem 2.12, and can also be deduced by the method used in Lemma 3.2 and Theorem 4.1 of [41].

Case 3. For flows in class 𝒦+\mathcal{K}^{+}, the main tools to rule out oscillation are Hamiltonian structure and index formula, and the proof is in Subsection 4.3. This is also the main reason why the spectral and “good” endpoints assumptions can be removed in Theorem 1.2.

Theorem 2.13.

Let uu be in class 𝒦+\mathcal{K}^{+} and α>0\alpha>0. Then mβ(σd(α,β)(,umin))<m_{\beta}\leq\sharp(\sigma_{d}(\mathcal{R}_{\alpha,\beta})\cap(-\infty,u_{\min}))<\infty for 0<β98κ+0<\beta\leq{9\over 8}\kappa_{+}; mβ(σd(α,β)(umax,))<m_{\beta}\leq\sharp(\sigma_{d}(\mathcal{R}_{\alpha,\beta})\cap(u_{\max},\infty))<\infty for 98κβ<0{9\over 8}\kappa_{-}\leq\beta<0; and (σd(α,β)𝐑)=\sharp(\sigma_{d}(\mathcal{R}_{\alpha,\beta})\cap\mathbf{R})=\infty for β[98κ,98κ+]\beta\notin[{9\over 8}\kappa_{-},{9\over 8}\kappa_{+}]. Here, κ±\kappa_{\pm} and mβm_{\beta} are defined in (1.5)–(1.8) and (1.17), respectively.

The idea of the proof is as follows. The linearized equation has Hamiltonian structure and the energy quadratic form has finite negative directions. The key observation is that oscillation of λn(c)\lambda_{n}(c) brings infinite times of sign-changes of λn(c)\lambda_{n}^{\prime}(c). This contributes infinite negative directions of quadratic form for non-resonant neutral modes, which is a contradiction to the index formula. Thus, the oscillation of λn(c)\lambda_{n}(c) can be ruled out unconditionally for flows in class 𝒦+\mathcal{K}^{+}.

3. Transitional values β=98κ±\beta={9\over 8}\kappa_{\pm} for the nn-th eigenvalue of Rayleigh-Kuo BVP

We begin to study the asymptotic behavior of the nn-th eigenvalue λn(c)\lambda_{n}(c) of Rayleigh-Kuo BVP. In this section, we focus on the number {n1:λn(c) as cumin(or cumax+)}\sharp\{n\geq 1:\lambda_{n}(c)\to-\infty\text{ as }c\to u_{\min}^{-}\;(\text{or }c\to u_{\max}^{+})\}. We prove that the number is finite for β[98κ,98κ+]\beta\in[{9\over 8}\kappa_{-},{9\over 8}\kappa_{+}] and it is infinite for β[98κ,98κ+]\beta\notin[{9\over 8}\kappa_{-},{9\over 8}\kappa_{+}], which is stated precisely in Theorem 2.9.

3.1. Finite number for β[98κ,98κ+]\beta\in[{9\over 8}\kappa_{-},{9\over 8}\kappa_{+}]

The optimal constant in the following Hardy type inequality plays an important role in discovering the transitional values β=98κ±\beta={9\over 8}\kappa_{\pm}.

Lemma 3.1.

Let ϕH1(a,b)\phi\in H^{1}(a,b) and ϕ(y0)=0\phi(y_{0})=0 for some y0[a,b]y_{0}\in[a,b]. Then

(3.1) ϕyy0L2(a,b)2+ab|yy0|1ϕ(y)2𝑑ymax(by0,y0a)4ϕL2(a,b)2.\displaystyle\left\|\phi\over y-y_{0}\right\|_{L^{2}(a,b)}^{2}+\frac{\int_{a}^{b}|y-y_{0}|^{-1}{\phi(y)^{2}}{}dy}{\max(b-y_{0},y_{0}-a)}\leq 4\|\phi^{\prime}\|_{L^{2}(a,b)}^{2}.

Here the constant 44 is optimal.

Proof.

Suppose that ϕ\phi is real-valued without loss of generality. Let ε=1/max(by0,y0a)\varepsilon=1/\max(b-y_{0},y_{0}-a). First, we consider the integration on [y0,b][y_{0},b] (if y0<by_{0}<b). Since

(3.2) |ϕ(y)2yy0|ϕL2(y0,y)2(yy0)yy0=ϕL2(y0,y)20,asyy0+,\displaystyle\left|\phi(y)^{2}\over y-y_{0}\right|\leq{\|\phi^{\prime}\|^{2}_{L^{2}(y_{0},y)}(y-y_{0})\over y-y_{0}}=\|\phi^{\prime}\|^{2}_{L^{2}(y_{0},y)}\to 0,\;\;as\;\;y\to y_{0}^{+},

we have

ϕyy0L2(y0,b)2=y0bϕ(y)2d(1yy0)=ϕ(y)2yy0|y0b+y0b2ϕ(y)ϕ(y)yy0𝑑y,\displaystyle\left\|\phi\over y-y_{0}\right\|_{L^{2}(y_{0},b)}^{2}=-\int_{y_{0}}^{b}\phi(y)^{2}d\left(1\over y-y_{0}\right)=-{\phi(y)^{2}\over y-y_{0}}\big{|}_{y_{0}}^{b}+\int_{y_{0}}^{b}{2\phi(y)\phi^{\prime}(y)\over y-y_{0}}dy,
y0b2ϕ(y)ϕ(y)𝑑y=ϕ(y)2|y0b=ϕ(b)2,\displaystyle\int_{y_{0}}^{b}{2\phi(y)\phi^{\prime}(y)}dy={\phi(y)^{2}}\big{|}_{y_{0}}^{b}=\phi(b)^{2},
ϕyy0L2(y0,b)2+ϕyy0εϕ2ϕL2(y0,b)2\displaystyle\left\|\phi\over y-y_{0}\right\|_{L^{2}(y_{0},b)}^{2}+\left\|{\phi\over y-y_{0}}-\varepsilon\phi-2\phi^{\prime}\right\|_{L^{2}(y_{0},b)}^{2}
=\displaystyle= 2ϕ(b)2by0+2εϕ(b)22εy0bϕ(y)2yy0𝑑y+ε2ϕL2(y0,b)2+4ϕL2(y0,b)2\displaystyle-2{\phi(b)^{2}\over b-y_{0}}+2\varepsilon\phi(b)^{2}-2\varepsilon\int_{y_{0}}^{b}\frac{\phi(y)^{2}}{y-y_{0}}dy+\varepsilon^{2}\|\phi\|_{L^{2}(y_{0},b)}^{2}+4\|\phi^{\prime}\|_{L^{2}(y_{0},b)}^{2}
\displaystyle\leq εy0bϕ(y)2yy0𝑑y+4ϕL2(y0,b)2.\displaystyle-\varepsilon\int_{y_{0}}^{b}\frac{\phi(y)^{2}}{y-y_{0}}dy+4\|\phi^{\prime}\|_{L^{2}(y_{0},b)}^{2}.

Here we used ε1by0\varepsilon\leq{1\over b-y_{0}} and ϕ(y)2yy0|y0b=ϕ(b)2by0.{\phi(y)^{2}\over y-y_{0}}\big{|}_{y_{0}}^{b}={\phi(b)^{2}\over b-y_{0}}. Thus, ϕyy0L2(y0,b)2+εy0bϕ(y)2yy0𝑑y4ϕL2(y0,b)2.\big{\|}{\phi\over y-y_{0}}\big{\|}_{L^{2}(y_{0},b)}^{2}+\varepsilon\int_{y_{0}}^{b}\frac{\phi(y)^{2}}{y-y_{0}}dy\leq 4\|\phi^{\prime}\|_{L^{2}(y_{0},b)}^{2}. Similarly, ϕyy0L2(a,y0)2+εay0ϕ(y)2y0y𝑑y4ϕL2(a,y0)2\big{\|}{\phi\over y-y_{0}}\big{\|}_{L^{2}(a,y_{0})}^{2}+\varepsilon\int_{a}^{y_{0}}\frac{\phi(y)^{2}}{y_{0}-y}dy\leq 4\|\phi^{\prime}\|_{L^{2}(a,y_{0})}^{2}. This gives (3.1). Letting y0=ay_{0}=a, ϕ(y)=(ya)12+ε1\phi(y)=(y-a)^{{1\over 2}+\varepsilon_{1}} and sending ε10+\varepsilon_{1}\to 0^{+}, we see that the constant 44 is optimal. ∎

For other versions of Hardy type inequality, the readers are referred to [15, 33]. To study the lower bound of the nn-th eigenvalue λn(c)\lambda_{n}(c) of Rayleigh-Kuo BVP for cc close to uminu_{\min}^{-}, it is important to estimate the energy expression (2.7) near singular points. To this end, we need the following lemma.

Lemma 3.2.

Assume that a[y1,y2],u(a)=umin,ϕH01(y1,y2),c<umina\in[y_{1},y_{2}],\ u(a)=u_{\min},\ \phi\in H_{0}^{1}(y_{1},y_{2}),\ c<u_{\min} and β>0\beta>0. Then there exists a constant δ0>0\delta_{0}>0 (depending only on uu and aa) such that for 0<δδ00<\delta\leq\delta_{0},

  • (1)

    if (i)(\rm{i}) u(a)0u^{\prime}(a)\neq 0 or (ii)(\rm{ii}) u(a)=0,β9u′′(a)/8,ϕ(a)=0u^{\prime}(a)=0,\ \beta\leq 9u^{\prime\prime}(a)/8,\ \phi(a)=0, then

    (3.3) [aδ,a+δ][y1,y2](|ϕ|2+u′′βuc|ϕ|2)𝑑y0;\displaystyle\int_{[a-\delta,a+\delta]\cap[y_{1},y_{2}]}\left(|\phi^{\prime}|^{2}+{u^{\prime\prime}-\beta\over u-c}|\phi|^{2}\right)dy\geq 0;
  • (2)

    if (iii)(\rm{iii}) a(y1,y2),u(a)=0,βu′′(a)a\in(y_{1},y_{2}),\ u^{\prime}(a)=0,\ \beta\leq u^{\prime\prime}(a), then

    (3.4) [aδ,a+δ][y1,y2](|ϕ|2+u′′βuc|ϕ|2)𝑑yCδϕL2([aδ,a+δ][y1,y2])2.\displaystyle\int_{[a-\delta,a+\delta]\cap[y_{1},y_{2}]}\left(|\phi^{\prime}|^{2}+{u^{\prime\prime}-\beta\over u-c}|\phi|^{2}\right)dy\geq-C_{\delta}\|\phi\|_{L^{2}([a-\delta,a+\delta]\cap[y_{1},y_{2}])}^{2}.

    Here CδC_{\delta} is a positive constant depending only on u,au,a and δ\delta.

Proof.

First, we assume (i). Then we have a{y1,y2}a\in\{y_{1},y_{2}\} and thus ϕ(a)=0\phi(a)=0. Without loss of generality, we assume that a=y1a=y_{1}. In this case, u(y1)>0u^{\prime}(y_{1})>0. Choose δ1(0,y2y1)\delta_{1}\in(0,y_{2}-y_{1}) small enough such that u(y)>u(y1)2u^{\prime}(y)>{u^{\prime}(y_{1})\over 2} for y(y1,y1+δ1)y\in(y_{1},y_{1}+\delta_{1}), and thus, there exists ξy(y1,y)\xi_{y}\in(y_{1},y) such that u(y)c>u(y)umin=u(ξy)(yy1)>u(y1)2(yy1)>0u(y)-c>u(y)-u_{\min}=u^{\prime}(\xi_{y})(y-y_{1})>{u^{\prime}(y_{1})\over 2}(y-y_{1})>0 for c<uminc<u_{\min}. Note that for y(y1,y1+δ)y\in(y_{1},y_{1}+\delta),

|ϕ(y)|2=|y1yϕ(s)𝑑s|2ϕL2(y1,y)2(yy1).\displaystyle|\phi(y)|^{2}=\left|\int_{y_{1}}^{y}\phi^{\prime}(s)ds\right|^{2}\leq\|\phi^{\prime}\|_{L^{2}(y_{1},y)}^{2}(y-y_{1}).

Now we take δ0(0,δ1)\delta_{0}\in(0,\delta_{1}) to be small enough such that y1y1+δ02|u′′β|u(y1)𝑑y1.\int_{y_{1}}^{y_{1}+\delta_{0}}{2|u^{\prime\prime}-\beta|\over u^{\prime}(y_{1})}dy\leq 1. Then for 0<δδ00<\delta\leq\delta_{0},

|y1y1+δu′′βuc|ϕ|2𝑑y|ϕL2(y1,y1+δ)2y1y1+δ2|u′′β|u(y1)𝑑yϕL2(y1,y1+δ)2,\displaystyle\left|\int_{y_{1}}^{y_{1}+\delta}{u^{\prime\prime}-\beta\over u-c}|\phi|^{2}dy\right|\leq\|\phi^{\prime}\|_{L^{2}(y_{1},y_{1}+\delta)}^{2}\int_{y_{1}}^{y_{1}+\delta}{2|u^{\prime\prime}-\beta|\over u^{\prime}(y_{1})}dy\leq\|\phi^{\prime}\|_{L^{2}(y_{1},y_{1}+\delta)}^{2},

which implies (3.3) since [aδ,a+δ][y1,y2]=[y1,y1+δ][a-\delta,a+\delta]\cap[y_{1},y_{2}]=[y_{1},y_{1}+\delta].

Now we assume (ii), then u′′(a)>0u^{\prime\prime}(a)>0. Let δ1(0,max(y2a,ay1))\delta_{1}\in(0,\max(y_{2}-a,a-y_{1})) be small enough such that u′′(y)>u′′(a)2>0u^{\prime\prime}(y)>{u^{\prime\prime}(a)\over 2}>0 for y[aδ1,a+δ1][y1,y2]y\in[a-\delta_{1},a+\delta_{1}]\cap[y_{1},y_{2}]. Since uH4(y1,y2)C3([y1,y2])u\in H^{4}(y_{1},y_{2})\subset C^{3}([y_{1},y_{2}]), we have |u′′(y)u′′(a)|C|ya||u^{\prime\prime}(y)-u^{\prime\prime}(a)|\leq C|y-a| and |1/u′′(y)1/u′′(a)|C|ya||1/u^{\prime\prime}(y)-1/u^{\prime\prime}(a)|\leq C|y-a| for y[aδ1,a+δ1][y1,y2]y\in[a-\delta_{1},a+\delta_{1}]\cap[y_{1},y_{2}]. Then there exists ξy{z:|za|<|ya|}\xi_{y}\in\{z:|z-a|<|y-a|\} such that u(y)c>u(y)umin=u′′(ξy)2(ya)2u(y)-c>u(y)-u_{\min}={u^{\prime\prime}(\xi_{y})\over 2}(y-a)^{2} >2β9(ya)2>0>{2\beta\over 9}(y-a)^{2}>0 for c<uminc<u_{\min} and y[aδ1,a+δ1][y1,y2]y\in[a-\delta_{1},a+\delta_{1}]\cap[y_{1},y_{2}], and thus

0\displaystyle 0 <1u(y)c<2u′′(ξy)(ya)22+C|ξya|u′′(a)(ya)22+C|ya|u′′(a)(ya)2,\displaystyle<{1\over u(y)-c}<{2\over u^{\prime\prime}(\xi_{y})(y-a)^{2}}\leq{2+C|\xi_{y}-a|\over u^{\prime\prime}(a)(y-a)^{2}}\leq{2+C|y-a|\over u^{\prime\prime}(a)(y-a)^{2}},
u′′(y)β\displaystyle u^{\prime\prime}(y)-\beta u′′(a)C|ya|9u′′(a)/8=u′′(a)/8C|ya|,\displaystyle\geq u^{\prime\prime}(a)-C|y-a|-9u^{\prime\prime}(a)/8=-u^{\prime\prime}(a)/8-C|y-a|,
u′′(y)βu(y)c\displaystyle{u^{\prime\prime}(y)-\beta\over u(y)-c} u′′(a)/8C|ya|u(y)cu′′(a)82+C|ya|u′′(a)(ya)2C|ya|(2β/9)(ya)2\displaystyle\geq{-u^{\prime\prime}(a)/8-C|y-a|\over u(y)-c}\geq-{u^{\prime\prime}(a)\over 8}{2+C|y-a|\over u^{\prime\prime}(a)(y-a)^{2}}-{C|y-a|\over(2\beta/9)(y-a)^{2}}
1+C0|ya|4(ya)2.\displaystyle\geq-{1+C_{0}|y-a|\over 4(y-a)^{2}}.

Now we take δ0=min(δ1,C01)>0\delta_{0}=\min(\delta_{1},C_{0}^{-1})>0. For 0<δδ00<\delta\leq\delta_{0}, we have

(3.5) [aδ,a+δ][y1,y2]u′′βuc|ϕ|2𝑑y[aδ,a+δ][y1,y2]1+C0|ya|4(ya)2|ϕ|2𝑑y\displaystyle-\int_{[a-\delta,a+\delta]\cap[y_{1},y_{2}]}{u^{\prime\prime}-\beta\over u-c}|\phi|^{2}dy\leq\int_{[a-\delta,a+\delta]\cap[y_{1},y_{2}]}{1+C_{0}|y-a|\over 4(y-a)^{2}}|\phi|^{2}dy
\displaystyle\leq [aδ,a+δ][y1,y2]1+δ1|ya|4(ya)2|ϕ|2𝑑yϕL2([aδ,a+δ][y1,y2])2,\displaystyle\int_{[a-\delta,a+\delta]\cap[y_{1},y_{2}]}{1+\delta^{-1}|y-a|\over 4(y-a)^{2}}|\phi|^{2}dy\leq\left\|\phi^{\prime}\right\|_{L^{2}([a-\delta,a+\delta]\cap[y_{1},y_{2}])}^{2},

which implies (3.3). Here we used Lemma 3.1 in the last step.

Next, we assume (iii). In this case, u(a)=0u^{\prime}(a)=0. Let β~=u′′(a)>0\widetilde{\beta}=u^{\prime\prime}(a)>0. Then β~β\widetilde{\beta}\geq\beta. Let δ1(0,min(y2a,ay1))\delta_{1}\in(0,\min(y_{2}-a,a-y_{1})) be small enough such that u′′(y)>u′′(a)2>0u^{\prime\prime}(y)>{u^{\prime\prime}(a)\over 2}>0 and |u(y)umin|1/2|u(y)-u_{\min}|\leq 1/2 for y[aδ1,a+δ1]y\in[a-\delta_{1},a+\delta_{1}]. Then 0<u(y)c<10<u(y)-c<1 for y[aδ1,a+δ1]y\in[a-\delta_{1},a+\delta_{1}] and c(umin1/2,umin)c\in(u_{\min}-1/2,u_{\min}). Now we assume 0<δδ10<\delta\leq\delta_{1}. Direct computation implies

aδa+δu′′β~uc|ϕ|2𝑑y=aδa+δu′′β~u|ϕ|2d(ln(uc))\displaystyle\int_{a-\delta}^{a+\delta}{u^{\prime\prime}-\widetilde{\beta}\over u-c}|\phi|^{2}dy=\int_{a-\delta}^{a+\delta}{u^{\prime\prime}-\widetilde{\beta}\over u^{\prime}}|\phi|^{2}d(\ln(u-c))
=\displaystyle= (u′′β~u|ϕ|2(ln(uc)))|aδa+δaδa+δln(uc)(u′′β~u|ϕ|2)𝑑y=Ic,δ(ϕ)+IIc,δ(ϕ).\displaystyle\left({u^{\prime\prime}-\widetilde{\beta}\over u^{\prime}}|\phi|^{2}(\ln(u-c))\right)\big{|}_{a-\delta}^{a+\delta}-\int_{a-\delta}^{a+\delta}\ln(u-c)\left({u^{\prime\prime}-\widetilde{\beta}\over u^{\prime}}|\phi|^{2}\right)^{\prime}dy=I_{c,\delta}(\phi)+II_{c,\delta}(\phi).

Since β~u′′(a)=0\widetilde{\beta}-u^{\prime\prime}(a)=0, it follows from the proof of Lemma 3.7 in [47] that u′′β~uH1(aδ1,a+δ1){u^{\prime\prime}-\widetilde{\beta}\over u^{\prime}}\in H^{1}(a-\delta_{1},a+\delta_{1}). By interpolation, we have ϕL(aδ,a+δ)CδϕL2(aδ,a+δ)+ϕL2(aδ,a+δ)\|\phi\|_{L^{\infty}(a-\delta,a+\delta)}\leq C_{\delta}\|\phi\|_{L^{2}(a-\delta,a+\delta)}+\|\phi^{\prime}\|_{L^{2}(a-\delta,a+\delta)}, and thus

|(u′′β~u|ϕ|2)(a+δ)(u′′β~u|ϕ|2)(aδ)|C(u′′β~u|ϕ|2)L2(aδ,a+δ)δ12\displaystyle\left|\left({u^{\prime\prime}-\widetilde{\beta}\over u^{\prime}}|\phi|^{2}\right)({a+\delta})-\left({u^{\prime\prime}-\widetilde{\beta}\over u^{\prime}}|\phi|^{2}\right)({a-\delta})\right|\leq C\left\|\left({u^{\prime\prime}-\widetilde{\beta}\over u^{\prime}}|\phi|^{2}\right)^{\prime}\right\|_{L^{2}(a-\delta,a+\delta)}\delta^{1\over 2}
\displaystyle\leq C(u′′β~u)|ϕ|2+(u′′β~u)2ϕϕL2(aδ,a+δ)δ12\displaystyle C\left\|\left({u^{\prime\prime}-\widetilde{\beta}\over u^{\prime}}\right)^{\prime}|\phi|^{2}+\left({u^{\prime\prime}-\widetilde{\beta}\over u^{\prime}}\right)2\phi\phi^{\prime}\right\|_{L^{2}(a-\delta,a+\delta)}\delta^{1\over 2}
\displaystyle\leq C(ϕL(aδ,a+δ)2+ϕL(aδ,a+δ)ϕL2(aδ,a+δ))δ12\displaystyle C\left(\|\phi\|^{2}_{L^{\infty}(a-\delta,a+\delta)}+\|\phi\|_{L^{\infty}(a-\delta,a+\delta)}\|\phi^{\prime}\|_{L^{2}(a-\delta,a+\delta)}\right)\delta^{1\over 2}
\displaystyle\leq (CδϕL2(aδ,a+δ)2+CϕL2(aδ,a+δ)2)δ12\displaystyle(C_{\delta}\|\phi\|^{2}_{L^{2}(a-\delta,a+\delta)}+C\|\phi^{\prime}\|^{2}_{L^{2}(a-\delta,a+\delta)})\delta^{1\over 2}

and |(u′′β~u|ϕ|2)(aδ)|CϕL(aδ,a+δ)2CδϕL2(aδ,a+δ)2+CϕL2(aδ,a+δ)2\left|\left({u^{\prime\prime}-\widetilde{\beta}\over u^{\prime}}|\phi|^{2}\right)(a-\delta)\right|\leq C\|\phi\|^{2}_{L^{\infty}(a-\delta,a+\delta)}\leq C_{\delta}\|\phi\|^{2}_{L^{2}(a-\delta,a+\delta)}+C\|\phi^{\prime}\|^{2}_{L^{2}(a-\delta,a+\delta)}. Then

Ic,δ(ϕ)=\displaystyle I_{c,\delta}(\phi)= (u′′β~u|ϕ|2)|aδa+δln(u(a+δ)c)+(u′′β~u|ϕ|2)|aδln(uc)|aδa+δ\displaystyle\left({u^{\prime\prime}-\widetilde{\beta}\over u^{\prime}}|\phi|^{2}\right)\big{|}_{a-\delta}^{a+\delta}\ln(u(a+\delta)-c)+\left({u^{\prime\prime}-\widetilde{\beta}\over u^{\prime}}|\phi|^{2}\right)|_{a-\delta}\ln(u-c)\big{|}_{a-\delta}^{a+\delta}
\displaystyle\leq (CδϕL2(aδ,a+δ)2+CϕL2(aδ,a+δ)2)(δ12|ln(u(a+δ)c)|+|ln(uc)|aδa+δ|).\displaystyle(C_{\delta}\|\phi\|^{2}_{L^{2}(a-\delta,a+\delta)}+C\|\phi^{\prime}\|^{2}_{L^{2}(a-\delta,a+\delta)})(\delta^{1\over 2}|\ln(u(a+\delta)-c)|+|\ln(u-c)\big{|}_{a-\delta}^{a+\delta}|).

Note that C(ya)2|u(y)u(a)|C′′(ya)2C^{\prime}(y-a)^{2}\leq|u(y)-u(a)|\leq C^{\prime\prime}(y-a)^{2} for y[aδ,a+δ]y\in[a-\delta,a+\delta]. Then |ln(u(y)c)|=ln(u(y)c)ln(u(y)u(a))ln(C(ya)2)|\ln(u(y)-c)|=-\ln(u(y)-c)\leq-\ln(u(y)-u(a))\leq-\ln(C^{\prime}(y-a)^{2}) and |ln(u(a+δ)c)|ln(Cδ2)C(|lnδ|+1)|\ln(u(a+\delta)-c)|\leq-\ln(C^{\prime}\delta^{2})\leq C(|\ln\delta|+1) for |cumin|<1/2|c-u_{\min}|<1/2 and y[aδ,a+δ]y\in[a-\delta,a+\delta]. Let u+=max(u(a+δ),u(aδ))u_{+}=\max(u(a+\delta),u(a-\delta)) and u=min(u(a+δ),u(aδ))u_{-}=\min(u(a+\delta),u(a-\delta)). Then u+u>u(a)u_{+}\geq u_{-}>u(a), and

|ln(uc)|aδa+δ|=lnu+cuclnu+u(a)uu(a)=|lnu(a+δ)u(a)u(aδ)u(a)|\displaystyle|\ln(u-c)\big{|}_{a-\delta}^{a+\delta}|=\ln{u_{+}-c\over u_{-}-c}\leq\ln{u_{+}-u(a)\over u_{-}-u(a)}=\left|\ln{u(a+\delta)-u(a)\over u(a-\delta)-u(a)}\right|

for c<umin=u(a)c<u_{\min}=u(a). Thus,

(3.6) |Ic,δ(ϕ)|(CδϕL2(aδ,a+δ)2+CϕL2(aδ,a+δ)2)Ψ(δ)\displaystyle|I_{c,\delta}(\phi)|\leq(C_{\delta}\|\phi\|^{2}_{L^{2}(a-\delta,a+\delta)}+C\|\phi^{\prime}\|^{2}_{L^{2}(a-\delta,a+\delta)})\Psi(\delta)

for c(umin1/2,umin)c\in(u_{\min}-1/2,u_{\min}), where

Ψ(δ)=δ12(|lnδ|+1)+|lnu(a+δ)u(a)u(aδ)u(a)|.\displaystyle\Psi(\delta)=\delta^{1\over 2}(|\ln\delta|+1)+\left|\ln{u(a+\delta)-u(a)\over u(a-\delta)-u(a)}\right|.

Note that

limδ0+|lnu(a+δ)u(a)u(aδ)u(a)|=limδ0+|lnu′′(ξa+δ)u′′(ξaδ)|=0,\displaystyle\lim_{\delta\to 0^{+}}\left|\ln{u(a+\delta)-u(a)\over u(a-\delta)-u(a)}\right|=\lim_{\delta\to 0^{+}}\left|\ln{u^{\prime\prime}(\xi_{a+\delta})\over u^{\prime\prime}(\xi_{a-\delta})}\right|=0,

where ξa+δ(a,a+δ)\xi_{a+\delta}\in(a,a+\delta) and ξaδ(aδ,a)\xi_{a-\delta}\in(a-\delta,a). Therefore limδ0+Ψ(δ)=0\lim_{\delta\to 0^{+}}\Psi(\delta)=0.

Next, we claim that ln(uc)\ln(u-c), c(umin1/2,umin)c\in(u_{\min}-1/2,u_{\min}), is uniformly bounded in Lp(aδ,a+δ)L^{p}(a-\delta,a+\delta) for 1<p<1<p<\infty. The proof is similar as that in Lemma 3.7 of [47]. Note that |ln(u(y)c)|ln(C(ya)2)|\ln(u(y)-c)|\leq-\ln(C^{\prime}(y-a)^{2}) for |cumin|<1/2|c-u_{\min}|<1/2 and y[aδ,a+δ]y\in[a-\delta,a+\delta]. Therefore,

aδa+δ|ln(uc)|pdyCaδa+δ(|ln(ya)2|p+1)dyCδδ(|ln|z|2|p+1)dzC.\displaystyle\int_{a-\delta}^{a+\delta}|\ln(u-c)|^{p}dy\leq C\int_{a-\delta}^{a+\delta}(|\ln(y-a)^{2}|^{p}+1)dy\leq C\int_{-\delta}^{\delta}(|\ln|z|^{2}|^{p}+1)dz\leq C.

Now, we consider IIc,δ(ϕ)II_{c,\delta}(\phi).

(3.7) |IIc,δ(ϕ)|\displaystyle|II_{c,\delta}(\phi)|\leq (2δ)14ln(uc)L4(aδ,a+δ)(u′′βu|ϕ|2)L2(aδ,a+δ)\displaystyle(2\delta)^{1\over 4}\|\ln(u-c)\|_{L^{4}(a-\delta,a+\delta)}\left\|\left({u^{\prime\prime}-\beta\over u^{\prime}}|\phi|^{2}\right)^{\prime}\right\|_{L^{2}(a-\delta,a+\delta)}
\displaystyle\leq (CδϕL2(aδ,a+δ)2+CϕL2(aδ,a+δ)2)δ14.\displaystyle(C_{\delta}\|\phi\|^{2}_{L^{2}(a-\delta,a+\delta)}+C\|\phi^{\prime}\|^{2}_{L^{2}(a-\delta,a+\delta)})\delta^{1\over 4}.

Combining (3.6) and (3.7), we get for c(umin1/2,umin)c\in(u_{\min}-1/2,u_{\min}),

|aδa+δu′′β~uc|ϕ|2𝑑y|\displaystyle\left|\int_{a-\delta}^{a+\delta}{u^{\prime\prime}-\widetilde{\beta}\over u-c}|\phi|^{2}dy\right|\leq (CδϕL2(aδ,a+δ)2+C1ϕL2(aδ,a+δ)2)(δ14+Ψ(δ)).\displaystyle(C_{\delta}\|\phi\|^{2}_{L^{2}(a-\delta,a+\delta)}+C_{1}\|\phi^{\prime}\|^{2}_{L^{2}(a-\delta,a+\delta)})(\delta^{1\over 4}+\Psi(\delta)).

Since limδ0+Ψ(δ)=0\lim_{\delta\to 0^{+}}\Psi(\delta)=0, we have limδ0+(δ14+Ψ(δ))=0\lim_{\delta\to 0^{+}}(\delta^{1\over 4}+\Psi(\delta))=0, and there exists δ0(0,δ1)\delta_{0}\in(0,\delta_{1}) such that C1(δ14+Ψ(δ))<1C_{1}(\delta^{1\over 4}+\Psi(\delta))<1. Then for 0<δδ00<\delta\leq\delta_{0} and c(umin1/2,umin)c\in(u_{\min}-1/2,u_{\min}), we have

[aδ,a+δ][y1,y2]u′′βuc|ϕ|2𝑑y[aδ,a+δ][y1,y2]u′′β~uc|ϕ|2𝑑y\displaystyle\int_{[a-\delta,a+\delta]\cap[y_{1},y_{2}]}{u^{\prime\prime}-\beta\over u-c}|\phi|^{2}dy\geq\int_{[a-\delta,a+\delta]\cap[y_{1},y_{2}]}{u^{\prime\prime}-\widetilde{\beta}\over u-c}|\phi|^{2}dy
\displaystyle\geq (CδϕL2(aδ,a+δ)2+C1ϕL2(aδ,a+δ)2)(δ14+Ψ(δ))\displaystyle-(C_{\delta}\|\phi\|^{2}_{L^{2}(a-\delta,a+\delta)}+C_{1}\|\phi^{\prime}\|^{2}_{L^{2}(a-\delta,a+\delta)})(\delta^{1\over 4}+\Psi(\delta))
\displaystyle\geq CδϕL2(aδ,a+δ)2(δ14+Ψ(δ))ϕL2(aδ,a+δ)2,\displaystyle-C_{\delta}\|\phi\|^{2}_{L^{2}(a-\delta,a+\delta)}(\delta^{1\over 4}+\Psi(\delta))-\|\phi^{\prime}\|^{2}_{L^{2}(a-\delta,a+\delta)},

which implies (3.4) since [aδ,a+δ][y1,y2]=[aδ,a+δ][a-\delta,a+\delta]\cap[y_{1},y_{2}]=[a-\delta,a+\delta]. On the other hand, for 0<δδ00<\delta\leq\delta_{0} and cumin1/2c\leq u_{\min}-1/2, we have

[aδ,a+δ][y1,y2]u′′βuc|ϕ|2𝑑y[aδ,a+δ][y1,y2]u′′β~uc|ϕ|2𝑑y\displaystyle\int_{[a-\delta,a+\delta]\cap[y_{1},y_{2}]}{u^{\prime\prime}-\beta\over u-c}|\phi|^{2}dy\geq\int_{[a-\delta,a+\delta]\cap[y_{1},y_{2}]}{u^{\prime\prime}-\widetilde{\beta}\over u-c}|\phi|^{2}dy
\displaystyle\geq 2[aδ,a+δ][y1,y2]|u′′β~||ϕ|2𝑑yCϕL2(aδ,a+δ)2,\displaystyle-2\int_{[a-\delta,a+\delta]\cap[y_{1},y_{2}]}|u^{\prime\prime}-\widetilde{\beta}||\phi|^{2}dy\geq-C\|\phi\|^{2}_{L^{2}(a-\delta,a+\delta)},

which implies (3.4). This completes the proof.∎

Now, we are ready to prove Theorem 2.9 (1)-(2).

Proof of Theorem 2.9 (1)-(2).

We first give the proof of (1), and (2) can be proved similarly. Consider 1nmβ1\leq n\leq m_{\beta}. It suffices to show that limcuminλmβ(c)=\lim_{c\to u_{\min}^{-}}\lambda_{m_{\beta}}(c)=-\infty. Let

(3.8) {a(y1,y2):u=umin,u′′(a)β<0}={a1,,amβ},\displaystyle\{a\in(y_{1},y_{2}):u=u_{\min},u^{\prime\prime}(a)-\beta<0\}=\{a_{1},\cdots,a_{m_{\beta}}\},

and

(3.11) η(x)={μexp(11x2),x(1,1),0,x(1,1),\displaystyle\eta(x)=\left\{\begin{array}[]{ll}\mu\exp\left({-1\over 1-x^{2}}\right),\;\;\;\;\;\;\;\;\;\;x\in(-1,1),\\ 0,\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;x\notin(-1,1),\end{array}\right.

where μ>0\mu>0 is a constant such that 11η(x)2𝑑x=1\int_{-1}^{1}\eta(x)^{2}dx=1. Then ηC0(𝐑)\eta\in C_{0}^{\infty}(\mathbf{R}). Define

φi(y)=δ012η(yaiδ0),y[y1,y2],\displaystyle\varphi_{i}(y)=\delta_{0}^{-{1\over 2}}\eta\left({y-a_{i}\over\delta_{0}}\right),\;\;y\in[y_{1},y_{2}],

where 1imβ1\leq i\leq m_{\beta}, and δ0>0\delta_{0}>0 is small enough such that (aiδ0,ai+δ0)(ajδ0,aj+δ0)=(a_{i}-\delta_{0},a_{i}+\delta_{0})\cap(a_{j}-\delta_{0},a_{j}+\delta_{0})=\emptyset for iji\neq j and u′′(y)β<0u^{\prime\prime}(y)-\beta<0 for all y1imβ(aiδ0,ai+δ0)(y1,y2)y\in\cup_{1\leq i\leq m_{\beta}}(a_{i}-\delta_{0},a_{i}+\delta_{0})\subset(y_{1},y_{2}). Then φiL2(y1,y2)=1\|\varphi_{i}\|_{L^{2}(y_{1},y_{2})}=1 and supp(φi)=(aiδ0,ai+δ0)\text{supp}\,(\varphi_{i})=(a_{i}-\delta_{0},a_{i}+\delta_{0}). Thus, φiφj\varphi_{i}\bot\varphi_{j} in the L2L^{2} sense for iji\neq j. Let Vmβ=span{φ1,,φmβ}.V_{m_{\beta}}=\text{span}\{\varphi_{1},\cdots,\varphi_{m_{\beta}}\}. Then VmβH01(y1,y2)V_{m_{\beta}}\subset H_{0}^{1}(y_{1},y_{2}). By (2.7), there exist bi,c𝐑b_{i,c}\in\mathbf{R}, i=1,,mβi=1,\cdots,m_{\beta}, with i=1mβ|bi,c|2=1\sum_{i=1}^{m_{\beta}}|b_{i,c}|^{2}=1 such that φc=i=1mβbi,cφiVmβ\varphi_{c}=\sum_{i=1}^{m_{\beta}}b_{i,c}\varphi_{i}\in V_{m_{\beta}} with φcL22=1\|\varphi_{c}\|^{2}_{L^{2}}=1, and

λmβ(c)\displaystyle\lambda_{m_{\beta}}(c)\leq supϕL2=1,ϕVmβy1y2(|ϕ|2+u′′βuc|ϕ|2)𝑑y=y1y2(|φc|2+u′′βuc|φc|2)𝑑y\displaystyle\sup_{\|\phi\|_{L^{2}}=1,\phi\in V_{m_{\beta}}}\int_{y_{1}}^{y_{2}}\left(|\phi^{\prime}|^{2}+{u^{\prime\prime}-\beta\over u-c}|\phi|^{2}\right)dy=\int_{y_{1}}^{y_{2}}\left(|\varphi_{c}^{\prime}|^{2}+{u^{\prime\prime}-\beta\over u-c}|\varphi_{c}|^{2}\right)dy
=\displaystyle= i=1mβ|bi,c|2aiδ0ai+δ0(|φi|2+u′′βuc|φi|2)𝑑y\displaystyle\sum_{i=1}^{m_{\beta}}|b_{i,c}|^{2}\int_{a_{i}-\delta_{0}}^{a_{i}+\delta_{0}}\left(|\varphi_{i}^{\prime}|^{2}+{u^{\prime\prime}-\beta\over u-c}|\varphi_{i}|^{2}\right)dy
(3.12) \displaystyle\leq max1imβaiδ0ai+δ0(|φi|2+u′′βuc|φi|2)𝑑yascumin.\displaystyle\max_{1\leq i\leq m_{\beta}}\int_{a_{i}-\delta_{0}}^{a_{i}+\delta_{0}}\left(|\varphi_{i}^{\prime}|^{2}+{u^{\prime\prime}-\beta\over u-c}|\varphi_{i}|^{2}\right)dy\to-\infty\;\;\text{as}\;\;c\to u_{\min}^{-}.

Next, we prove (ii). Let δ1>0\delta_{1}>0 be a sufficiently small constant such that (aδ1,a+δ1)[y1,y2](a-\delta_{1},a+\delta_{1})\subset[y_{1},y_{2}] for a{u=umin}{y1,y2}a\in\{u=u_{\min}\}\setminus\{y_{1},y_{2}\}, and |ab|>2δ1|a-b|>2\delta_{1} for a,b{u=umin}a,b\in\{u=u_{\min}\} and aba\neq b. There are four cases for zeros of a{u=umin}a\in\{u=u_{\min}\} as follows:

Case 1. a{y1,y2}a\in\{y_{1},y_{2}\} and u(a)0u^{\prime}(a)\neq 0;

Case 2. a{y1,y2}a\in\{y_{1},y_{2}\}, u(a)=0u^{\prime}(a)=0 (thus β98κ+9u′′(a)/8\beta\leq{9\over 8}\kappa_{+}\leq 9u^{\prime\prime}(a)/8);

Case 3. a(y1,y2)a\in(y_{1},y_{2}) and βu′′(a)\beta\leq u^{\prime\prime}(a);

Case 4. a(y1,y2)a\in(y_{1},y_{2}) and u′′(a)<β9u′′(a)/8u^{\prime\prime}(a)<\beta\leq 9u^{\prime\prime}(a)/8.
Then we divide our proof into four cases as above. In fact, for Cases 1–2, by Lemma 3.2 (1) there exists δ(a)>0\delta(a)>0 such that for 0<δδ(a)0<\delta\leq\delta(a), c<uminc<u_{\min} and ϕH01\phi\in H_{0}^{1},

(3.13) [aδ,a+δ][y1,y2](|ϕ|2+u′′βuc|ϕ|2)𝑑y0.\displaystyle\int_{[a-\delta,a+\delta]\cap[y_{1},y_{2}]}\left(|\phi^{\prime}|^{2}+{u^{\prime\prime}-\beta\over u-c}|\phi|^{2}\right)dy\geq 0.

For Case 3, by Lemma 3.2 (2) there exists δ(a)>0\delta(a)>0 such that for 0<δδ(a)0<\delta\leq\delta(a), c<uminc<u_{\min} and ϕH01\phi\in H_{0}^{1},

(3.14) [aδ,a+δ][y1,y2](|ϕ|2+u′′βuc|ϕ|2)𝑑yC(δ,a)[aδ,a+δ][y1,y2]|ϕ|2𝑑y.\displaystyle\int_{[a-\delta,a+\delta]\cap[y_{1},y_{2}]}\left(|\phi^{\prime}|^{2}+{u^{\prime\prime}-\beta\over u-c}|\phi|^{2}\right)dy\geq-C(\delta,a)\int_{[a-\delta,a+\delta]\cap[y_{1},y_{2}]}|\phi|^{2}dy.

Here C(δ,a)C(\delta,a) depends only on u,a,δu,a,\delta. Moreover, if ϕ(a)=0\phi(a)=0, then by Lemma 3.2 (1),

(3.15) [aδ,a+δ][y1,y2](|ϕ|2+u′′βuc|ϕ|2)𝑑y0.\displaystyle\int_{[a-\delta,a+\delta]\cap[y_{1},y_{2}]}\left(|\phi^{\prime}|^{2}+{u^{\prime\prime}-\beta\over u-c}|\phi|^{2}\right)dy\geq 0.

For Case 4, we have a{a1,,amβ}.a\in\{a_{1},\cdots,a_{m_{\beta}}\}. By Lemma 3.2 (1), there exists δ(a)>0\delta(a)>0 such that for 0<δδ(a)0<\delta\leq\delta(a), c<uminc<u_{\min}, ϕH01\phi\in H_{0}^{1} and ϕ(a)=0\phi(a)=0,

(3.16) [aδ,a+δ][y1,y2](|ϕ|2+u′′βuc|ϕ|2)𝑑y0.\displaystyle\int_{[a-\delta,a+\delta]\cap[y_{1},y_{2}]}\left(|\phi^{\prime}|^{2}+{u^{\prime\prime}-\beta\over u-c}|\phi|^{2}\right)dy\geq 0.

Now let δ0=min({δ(a):a{u=umin}}{δ1})\delta_{0}=\min(\{\delta(a):a\in\{u=u_{\min}\}\}\cup\{\delta_{1}\}). Define

(q1(y),q10(y))\displaystyle(q_{1}(y),q_{1}^{0}(y)) ={(u′′(y)βu(y)c,1)y[y1,y2]a{u=umin}(aδ0,a+δ0),(0,0)ya{u=umin}((aδ0,a+δ0)[y1,y2]),\displaystyle=\begin{cases}({u^{\prime\prime}(y)-\beta\over u(y)-c},1)&\;y\in[y_{1},y_{2}]\setminus\cup_{a\in\{u=u_{\min}\}}(a-\delta_{0},a+\delta_{0}),\\ (0,0)&y\in\cup_{a\in\{u=u_{\min}\}}\left((a-\delta_{0},a+\delta_{0})\cap[y_{1},y_{2}]\right),\end{cases}
(q2(y),q20(y))\displaystyle(q_{2}(y),q_{2}^{0}(y)) =(u′′(y)βu(y)cq1(y),1q10(y))y[y1,y2].\displaystyle=({u^{\prime\prime}(y)-\beta\over u(y)-c}-q_{1}(y),1-q_{1}^{0}(y))\;\;y\in[y_{1},y_{2}].

Then there exists C0>0C_{0}>0 such that for c<uminc<u_{\min},

|q1(y)|C0fory[y1,y2].\displaystyle|q_{1}(y)|\leq C_{0}\;\;\text{for}\;\;y\in[y_{1},y_{2}].

For ϕH01\phi\in H_{0}^{1} and ϕL2=1\|\phi\|_{L^{2}}=1,

y1y2(|ϕ|2+u′′βuc|ϕ|2)𝑑y\displaystyle\int_{y_{1}}^{y_{2}}\left(|\phi^{\prime}|^{2}+{u^{\prime\prime}-\beta\over u-c}|\phi|^{2}\right)dy
=\displaystyle= y1y2(q10|ϕ|2+q1|ϕ|2)𝑑y+y1y2(q20|ϕ|2+q2|ϕ|2)𝑑y=Ic(ϕ)+IIc(ϕ).\displaystyle\int_{y_{1}}^{y_{2}}\left(q_{1}^{0}|\phi^{\prime}|^{2}+q_{1}|\phi|^{2}\right)dy+\int_{y_{1}}^{y_{2}}\left(q_{2}^{0}|\phi^{\prime}|^{2}+q_{2}|\phi|^{2}\right)dy=I_{c}(\phi)+II_{c}(\phi).

Let us first consider Ic(ϕ)I_{c}(\phi). For ϕH01\phi\in H_{0}^{1} and ϕL2=1\|\phi\|_{L^{2}}=1, we have

(3.17) Ic(ϕ)y1y2(C0|ϕ|2)𝑑yC0\displaystyle I_{c}(\phi)\geq\int_{y_{1}}^{y_{2}}\left(-C_{0}|\phi|^{2}\right)dy\geq-C_{0}

for c<uminc<u_{\min}. We proceed to consider IIc(ϕ)II_{c}(\phi).

(3.18) IIc(ϕ)=\displaystyle II_{c}(\phi)= a{u=umin}[aδ0,a+δ0][y1,y2](|ϕ|2+u′′βuc|ϕ|2)𝑑y\displaystyle\sum_{a\in\{u=u_{\min}\}}\int_{[a-\delta_{0},a+\delta_{0}]\cap[y_{1},y_{2}]}\left(|\phi^{\prime}|^{2}+{u^{\prime\prime}-\beta\over u-c}|\phi|^{2}\right)dy
=\displaystyle= (Case 1++Case 4)[aδ0,a+δ0][y1,y2](|ϕ|2+u′′βuc|ϕ|2)𝑑y.\displaystyle\left(\sum_{\text{Case 1}}+\cdots+\sum_{\text{Case 4}}\right)\int_{[a-\delta_{0},a+\delta_{0}]\cap[y_{1},y_{2}]}\left(|\phi^{\prime}|^{2}+{u^{\prime\prime}-\beta\over u-c}|\phi|^{2}\right)dy.

Recall that a1,,amβa_{1},\cdots,a_{m_{\beta}} are defined in (3.8). For any (mβ+1)(m_{\beta}+1)-dimensional subspace V=span{ψ1,,ψmβ+1}V=\text{span}\{\psi_{1},\cdots,\psi_{m_{\beta}+1}\} in H01(y1,y2)H_{0}^{1}(y_{1},y_{2}), there exists 0(ξ1,,ξmβ+1)𝐑mβ+10\neq(\xi_{1},\cdots,\xi_{m_{\beta}+1})\in\mathbf{R}^{m_{\beta}+1} such that ξ1ψ1(ai)++ξmβ+1ψmβ+1(ai)=0,i=1,,mβ.\xi_{1}\psi_{1}(a_{i})+\cdots+\xi_{m_{\beta}+1}\psi_{m_{\beta}+1}(a_{i})=0,i=1,\cdots,m_{\beta}. Define ψ~=ξ1ψ1++ξmβ+1ψmβ+1.\tilde{\psi}=\xi_{1}\psi_{1}+\cdots+\xi_{m_{\beta}+1}\psi_{m_{\beta}+1}. Then ψ~(ai)=0\tilde{\psi}(a_{i})=0, i=1,,mβi=1,\cdots,m_{\beta}. We normalize ψ~\tilde{\psi} such that ψ~L2(y1,y2)=1\|\tilde{\psi}\|_{L^{2}(y_{1},y_{2})}=1. Then by (3.13), (3.14), (3.16) and (3.18), we have

IIc(ψ~)\displaystyle II_{c}(\tilde{\psi})\geq Case 3[aδ0,a+δ0][y1,y2](|ψ~|2+u′′βuc|ψ~|2)𝑑y\displaystyle\sum_{\text{Case 3}}\int_{[a-\delta_{0},a+\delta_{0}]\cap[y_{1},y_{2}]}\left(|\tilde{\psi}^{\prime}|^{2}+{u^{\prime\prime}-\beta\over u-c}|\tilde{\psi}|^{2}\right)dy
\displaystyle\geq Case 3C(δ0,a)[aδ0,a+δ0][y1,y2]|ψ~|2𝑑ymaxCase 3C(δ0,a).\displaystyle-\sum_{\text{Case 3}}C(\delta_{0},a)\int_{[a-\delta_{0},a+\delta_{0}]\cap[y_{1},y_{2}]}|\tilde{\psi}|^{2}dy\geq-\max_{\text{Case 3}}C(\delta_{0},a).

This, along with (2.7) and (3.17), yields that infc(,umin)λmβ+1(c)maxCase 3C(δ0,a)C0.\inf_{c\in(-\infty,u_{\min})}\lambda_{m_{\beta}+1}(c)\geq-\max\limits_{\text{Case 3}}C(\delta_{0},a)-C_{0}. This proves (ii).

Finally, we prove (iii). Let q1,q2,Ic(ϕ),IIc(ϕ)q_{1},q_{2},I_{c}(\phi),II_{c}(\phi) and C0C_{0} be defined as in (ii). Let μ1([a,b])\mu_{1}([a,b]) be the principal eigenvalue of

ϕ′′=λϕ,ϕ(a)=ϕ(b)=0.\displaystyle-\phi^{\prime\prime}=\lambda\phi,\;\;\phi(a)=\phi(b)=0.

Then we have μ1([a,b])=|π/(ba)|2,\mu_{1}([a,b])=|\pi/(b-a)|^{2}, and

(3.19) ab|ϕ|2𝑑yμ1([a,b])ab|ϕ|2𝑑yforϕH01(a,b).\displaystyle\int_{a}^{b}|\phi^{\prime}|^{2}dy\geq\mu_{1}([a,b])\int_{a}^{b}|\phi|^{2}dy\ \ \text{for}\ \phi\in H_{0}^{1}(a,b).

Let δ2=π/C012\delta_{2}=\pi/C_{0}^{\frac{1}{2}}. Then we have μ1([a,b])C0\mu_{1}([a,b])\geq C_{0} for 0<baδ20<b-a\leq\delta_{2}. Let

M=({nδ2:n𝐙}{a+b:a{u=umin},b{δ0,0,δ0}}{y1,y2})[y1,y2].\displaystyle M=(\{n\delta_{2}:n\in\mathbf{Z}\}\cup\{a+b:a\in\{u=u_{\min}\},b\in\{-\delta_{0},0,\delta_{0}\}\}\cup\{y_{1},y_{2}\})\cap[y_{1},y_{2}].

Then MM is a finite set, and we can write its elements in the increasing order

M={a0,,aNβ},y1=a0<<aNβ=y2.\displaystyle M=\{a_{0}^{\prime},\cdots,a_{N_{\beta}}^{\prime}\},\ y_{1}=a_{0}^{\prime}<\cdots<a_{N_{\beta}}^{\prime}=y_{2}.

Then 0<ak+1akδ20<a^{\prime}_{k+1}-a_{k}^{\prime}\leq\delta_{2} and μ1([ak,ak+1])C0\mu_{1}([a_{k}^{\prime},a^{\prime}_{k+1}])\geq C_{0} for 0k<Nβ0\leq k<N_{\beta}. Let

M1={k𝐙:0k<Nβ,[ak,ak+1](aδ0,a+δ0)=,a{u=umin}}.\displaystyle M_{1}=\{k\in\mathbf{Z}:0\leq k<N_{\beta},\ [a_{k}^{\prime},a^{\prime}_{k+1}]\cap(a-\delta_{0},a+\delta_{0})=\emptyset,\ \forall\ a\in\{u=u_{\min}\}\}.

Then we have [y1,y2](a{u=umin}(aδ0,a+δ0))=kM1[ak,ak+1].[y_{1},y_{2}]\setminus(\cup_{a\in\{u=u_{\min}\}}(a-\delta_{0},a+\delta_{0}))=\cup_{k\in M_{1}}[a_{k}^{\prime},a^{\prime}_{k+1}].

For any NβN_{\beta}-dimensional subspace V=span{ψ1,,ψNβ}V=\text{span}\{\psi_{1},\cdots,\psi_{N_{\beta}}\} in H01(y1,y2)H_{0}^{1}(y_{1},y_{2}), there exists 0(ξ1,,ξNβ)𝐑Nβ0\neq(\xi_{1},\cdots,\xi_{N_{\beta}})\in\mathbf{R}^{N_{\beta}} such that ξ1ψ1(ai)++ξNβψNβ(ai)=0,i=1,,Nβ1.\xi_{1}\psi_{1}(a_{i}^{\prime})+\cdots+\xi_{N_{\beta}}\psi_{N_{\beta}}(a_{i}^{\prime})=0,i=1,\cdots,N_{\beta}-1. Define ψ~=ξ1ψ1++ξNβψNβ.\tilde{\psi}=\xi_{1}\psi_{1}+\cdots+\xi_{N_{\beta}}\psi_{N_{\beta}}. Then ψ~(ai)=0\tilde{\psi}(a_{i}^{\prime})=0, i=1,,Nβ1i=1,\cdots,N_{\beta}-1. We normalize ψ~\tilde{\psi} such that ψ~L2(y1,y2)=1\|\tilde{\psi}\|_{L^{2}(y_{1},y_{2})}=1. Since ψ~H01(y1,y2)\tilde{\psi}\in H_{0}^{1}(y_{1},y_{2}), we also have ψ~(a0)=ψ~(y1)=0,ψ~(aNβ)=ψ~(y2)=0\tilde{\psi}(a_{0}^{\prime})=\tilde{\psi}(y_{1})=0,\ \tilde{\psi}(a_{N_{\beta}}^{\prime})=\tilde{\psi}(y_{2})=0, and thus ψ~(ai)=0\tilde{\psi}(a_{i}^{\prime})=0 for i=0,,Nβ,i=0,\cdots,N_{\beta}, i.e. ψ~|M=0.\tilde{\psi}|_{M}=0. By (3.19), we have

(3.20) akak+1|ψ~|2𝑑yμ1([ak,ak+1])akak+1|ψ~|2𝑑yC0akak+1|ψ~|2𝑑y,k=0,,Nβ1.\displaystyle\int_{a_{k}^{\prime}}^{a_{k+1}^{\prime}}|\tilde{\psi}^{\prime}|^{2}dy\geq\mu_{1}([a_{k}^{\prime},a^{\prime}_{k+1}])\int_{a_{k}^{\prime}}^{a_{k+1}^{\prime}}|\tilde{\psi}|^{2}dy\geq C_{0}\int_{a_{k}^{\prime}}^{a_{k+1}^{\prime}}|\tilde{\psi}|^{2}dy,\ k=0,\cdots,N_{\beta}-1.

First, we consider Ic(ψ~)I_{c}(\tilde{\psi}). By (3.20), we have

(3.21) Ic(ψ~)\displaystyle I_{c}(\tilde{\psi}) [y1,y2](a{u=umin}(aδ0,a+δ0))(|ψ~|2C0|ψ~|2)𝑑y\displaystyle\geq\int_{[y_{1},y_{2}]\setminus(\cup_{a\in\{u=u_{\min}\}}(a-\delta_{0},a+\delta_{0}))}\left(|\tilde{\psi}^{\prime}|^{2}-C_{0}|\tilde{\psi}|^{2}\right)dy
=kM1akak+1(|ψ~|2C0|ψ~|2)𝑑y0.\displaystyle=\sum_{k\in M_{1}}\int_{a_{k}^{\prime}}^{a_{k+1}^{\prime}}\left(|\tilde{\psi}^{\prime}|^{2}-C_{0}|\tilde{\psi}|^{2}\right)dy\geq 0.

Next, we consider IIc(ψ~)II_{c}(\tilde{\psi}). For a{u=umin},a\in\{u=u_{\min}\}, we have aMa\in M and ψ~(a)=0.\tilde{\psi}(a)=0. Then by (3.13), (3.15), (3.16) and (3.18), we have IIc(ψ~)0.II_{c}(\tilde{\psi})\geq 0. This, along with (2.7) and (3.21), yields that infc(,umin)λNβ(c)0.\inf_{c\in(-\infty,u_{\min})}\lambda_{N_{\beta}}(c)\geq 0. This proves (iii).∎

3.2. Infinite number for β[98κ,98κ+]\beta\notin[{9\over 8}\kappa_{-},{9\over 8}\kappa_{+}]

In this subsection, we prove Theorem 2.9 (3)-(4). The proof is based on construction of suitable test functions such that the energy in (2.7) converges to -\infty as cuminc\to u_{\min}^{-} or cumax+c\to u_{\max}^{+}.

Proof of Theorem 2.9 (3)-(4).

We only prove Theorem 2.9 (3), since (4) can be proved similarly. Let β>98κ+\beta>{9\over 8}\kappa_{+}. Then there exists a[y1,y2]a\in[y_{1},y_{2}] such that β/u′′(a)>9/8\beta/u^{\prime\prime}(a)>9/8, u(a)=0u^{\prime}(a)=0 and u(a)=uminu(a)=u_{\min}. If a[y1,y2)a\in[y_{1},y_{2}), our analysis is completely on [a,a+δ][y1,y2)[a,a+\delta]\subset[y_{1},y_{2}) for δ>0\delta>0 small enough. If a=y2a=y_{2}, the analysis is only on [aδ,a][a-\delta,a] and the proof is similar as a[y1,y2)a\in[y_{1},y_{2}). Now we assume that a=0[y1,y2)a=0\in[y_{1},y_{2}). Then u′′(0)>0u^{\prime\prime}(0)>0 and there exists ε0>0\varepsilon_{0}>0 such that u′′(z)>0u^{\prime\prime}(z)>0 and

2(u′′(y)β)u′′(z)<14ε0\displaystyle{2(u^{\prime\prime}(y)-\beta)\over u^{\prime\prime}(z)}<-{1\over 4-\varepsilon_{0}}

for y,z[0,δ][y1,y2)y,z\in[0,\delta]\subset[y_{1},y_{2}) and δ>0\delta>0 small enough. Let ν0=minz[0,δ]{u′′(z)}>0\nu_{0}=\min_{z\in[0,\delta]}\{u^{\prime\prime}(z)\}>0 and J(x)=η(2x1),x𝐑,J(x)=\eta(2x-1),x\in\mathbf{R}, where η\eta is defined in (3.11). Define

φi,R(y)={y12J(lnyR+i+1),y(0,y2],0,y[y1,0],\varphi_{i,R}(y)=\begin{cases}y^{1\over 2}J\left({\ln y\over R}+i+1\right),&y\in(0,y_{2}],\\ 0,&y\in[y_{1},0],\end{cases}

where i=1,,ni=1,\cdots,n, and RR is large enough such that eR<δe^{-R}<\delta. Then φi,RH01(y1,y2)\varphi_{i,R}\in H_{0}^{1}(y_{1},y_{2}) and suppφi,R=[e(i+1)R,eiR]\text{supp}\,\varphi_{i,R}=[e^{-(i+1)R},e^{-iR}], i=1,,ni=1,\cdots,n. Thus, φi,Rφj,R\varphi_{i,R}\perp\varphi_{j,R} in the L2L^{2} sense for iji\neq j. Note that u′′(y)β<0u^{\prime\prime}(y)-\beta<0 for y[0,δ]y\in[0,\delta]. For 1in1\leq i\leq n, we define

φ~i,R=1φi,RL2φi,R,andV~n,R=span{φ~1,R,,φ~n,R}.\tilde{\varphi}_{i,R}={1\over\|\varphi_{i,R}\|_{L^{2}}}\varphi_{i,R},\;\;\text{and}\;\;\tilde{V}_{n,R}=\text{span}\{\tilde{\varphi}_{1,R},\cdots,\tilde{\varphi}_{n,R}\}.

Choose R>0R>0 such that 2(uminc)=ε0ν0e2(n+1)R82(u_{\min}-c)={\varepsilon_{0}\nu_{0}e^{-2(n+1)R}\over 8}. Then cuminRc\to u_{\min}^{-}\Leftrightarrow R\to\infty. We shall show that for 1in,1\leq i\leq n,

(3.22) limcuminy1y2(|φ~i,R|2+u′′βuc|φ~i,R|2)𝑑y=.\displaystyle\lim_{c\to u_{\min}^{-}}\int_{y_{1}}^{y_{2}}\left(|\tilde{\varphi}_{i,R}^{\prime}|^{2}+{u^{\prime\prime}-\beta\over u-c}|\tilde{\varphi}_{i,R}|^{2}\right)dy=-\infty.

Assume that (3.22) is true. Similar to (3.12), there exist di,c𝐑,i=1,,n,d_{i,c}\in\mathbf{R},i=1,\cdots,n, with i=1n|di,c|2=1\sum_{i=1}^{n}|d_{i,c}|^{2}=1 such that

(3.23) λn(c)i=1n|di,c|2y1y2(|φ~i,R|2+u′′βuc|φ~i,R|2)𝑑yascumin.\displaystyle\lambda_{n}(c)\leq\sum_{i=1}^{n}|d_{i,c}|^{2}\int_{y_{1}}^{y_{2}}\left(|\tilde{\varphi}_{i,R}^{\prime}|^{2}+{u^{\prime\prime}-\beta\over u-c}|\tilde{\varphi}_{i,R}|^{2}\right)dy\to-\infty\;\;\text{as}\;\;c\to u_{\min}^{-}.

Now we prove (3.22). Direct computation gives

u′′(y)βu(y)c=\displaystyle{u^{\prime\prime}(y)-\beta\over u(y)-c}= u′′(y)βu(y)umin+uminc=2(u′′(y)β)u′′(ξy)y2+2(uminc)<2(u′′(y)β)(u′′(ξy)+ε0u′′(ξy)8)y2\displaystyle{u^{\prime\prime}(y)-\beta\over u(y)-u_{\min}+u_{\min}-c}={2(u^{\prime\prime}(y)-\beta)\over u^{\prime\prime}(\xi_{y})y^{2}+2(u_{\min}-c)}<{2(u^{\prime\prime}(y)-\beta)\over\left(u^{\prime\prime}(\xi_{y})+{\varepsilon_{0}u^{\prime\prime}(\xi_{y})\over 8}\right)y^{2}}
<\displaystyle< 14ε0u′′(ξy)(u′′(ξy)+ε0u′′(ξy)8)y2=1(4ε0)(1+ε08)y2=1(4ε1)y2\displaystyle{-{1\over 4-\varepsilon_{0}}u^{\prime\prime}(\xi_{y})\over\left(u^{\prime\prime}(\xi_{y})+{\varepsilon_{0}u^{\prime\prime}(\xi_{y})\over 8}\right)y^{2}}=-{1\over(4-\varepsilon_{0})\left(1+{\varepsilon_{0}\over 8}\right)y^{2}}=-{1\over(4-\varepsilon_{1})y^{2}}

for y[e(i+1)R,eiR]y\in[e^{-(i+1)R},e^{-iR}] and 2(uminc)=ε0ν0e2(n+1)R82(u_{\min}-c)={\varepsilon_{0}\nu_{0}e^{-2(n+1)R}\over 8}, where ξy(0,y)\xi_{y}\in(0,y) and ε1=ε02+ε028\varepsilon_{1}={\varepsilon_{0}\over 2}+{\varepsilon^{2}_{0}\over 8}. Then

(3.24) y1y2(|φi,R|2+u′′βuc|φi,R|2)𝑑y=\displaystyle\int_{y_{1}}^{y_{2}}\left(|\varphi_{i,R}^{\prime}|^{2}+{u^{\prime\prime}-\beta\over u-c}|\varphi_{i,R}|^{2}\right)dy= e(i+1)ReiR(|φi,R|2+u′′βuc|φi,R|2)𝑑y\displaystyle\int_{e^{-(i+1)R}}^{e^{-iR}}\left(|\varphi_{i,R}^{\prime}|^{2}+{u^{\prime\prime}-\beta\over u-c}|\varphi_{i,R}|^{2}\right)dy
<\displaystyle< e(i+1)ReiR(|φi,R|21(4ε1)y2|φi,R|2)𝑑y.\displaystyle\int_{e^{-(i+1)R}}^{e^{-iR}}\left(|\varphi_{i,R}^{\prime}|^{2}-{1\over(4-\varepsilon_{1})y^{2}}|\varphi_{i,R}|^{2}\right)dy.

Note that for y[e(i+1)R,eiR]y\in[e^{-(i+1)R},e^{-iR}],

|φi,R(y)|2=\displaystyle|\varphi_{i,R}^{\prime}(y)|^{2}= |J(x)(2x1)4x2(x1)21Ryy12+12y12J(x)|2\displaystyle\left|J(x){-(2x-1)\over 4x^{2}(x-1)^{2}}{1\over Ry}y^{1\over 2}+{1\over 2}y^{-{1\over 2}}J(x)\right|^{2}
=\displaystyle= J(x)2(2x1)216x4(x1)41R2yJ(x)22x14x2(x1)21Ry+14yJ(x)2,\displaystyle J(x)^{2}{(2x-1)^{2}\over 16x^{4}(x-1)^{4}}{1\over R^{2}y}-J(x)^{2}{2x-1\over 4x^{2}(x-1)^{2}}{1\over Ry}+{1\over 4y}J(x)^{2},

where x=lnyR+i+1x={\ln y\over R}+i+1. Since |J(x)2(2x1)216x4(x1)4|C\left|J(x)^{2}{(2x-1)^{2}\over 16x^{4}(x-1)^{4}}\right|\leq C and |J(x)22x14x2(x1)2|C\left|J(x)^{2}{2x-1\over 4x^{2}(x-1)^{2}}\right|\leq C for x[0,1]x\in[0,1], we get

|e(i+1)ReiR(J(x)2(2x1)216x4(x1)41R2yJ(x)22x14x2(x1)21Ry)𝑑y|\displaystyle\left|\int_{e^{-(i+1)R}}^{e^{-iR}}\left(J(x)^{2}{(2x-1)^{2}\over 16x^{4}(x-1)^{4}}{1\over R^{2}y}-J(x)^{2}{2x-1\over 4x^{2}(x-1)^{2}}{1\over Ry}\right)dy\right|
\displaystyle\leq CR2e(i+1)ReiR1y𝑑y+CRe(i+1)ReiR1y𝑑y=CR+CC.\displaystyle{C\over R^{2}}\int_{e^{-(i+1)R}}^{e^{-iR}}{1\over y}dy+{C\over R}\int_{e^{-(i+1)R}}^{e^{-iR}}{1\over y}dy={C\over R}+C\leq C.

Then we infer from (3.24) that

(3.25) y1y2(|φi,R|2+u′′βuc|φi,R|2)𝑑yC+e(i+1)ReiR(14yJ(x)21(4ε1)y2|φi,R|2)𝑑y\displaystyle\int_{y_{1}}^{y_{2}}\left(|\varphi_{i,R}^{\prime}|^{2}+{u^{\prime\prime}-\beta\over u-c}|\varphi_{i,R}|^{2}\right)dy\leq C+\int_{e^{-(i+1)R}}^{e^{-iR}}\left({1\over 4y}J(x)^{2}-{1\over(4-\varepsilon_{1})y^{2}}|\varphi_{i,R}|^{2}\right)dy
=\displaystyle= C+e(i+1)ReiRε14(4ε1)J(x)21y𝑑y=Cε1R8(4ε1)<0\displaystyle C+\int_{e^{-(i+1)R}}^{e^{-iR}}{-\varepsilon_{1}\over 4(4-\varepsilon_{1})}J(x)^{2}{1\over y}dy=C-{\varepsilon_{1}R\over 8(4-\varepsilon_{1})}<0

when RR is large enough. Direct computation gives

(3.26) φi,RL22=\displaystyle\|\varphi_{i,R}\|_{L^{2}}^{2}= e(i+1)ReiRyJ(lnyR+i+1)2𝑑y=01Re2R(xi1)J(x)2𝑑x\displaystyle\int_{e^{-(i+1)R}}^{e^{-iR}}yJ\left({\ln y\over R}+i+1\right)^{2}dy=\int_{0}^{1}Re^{2R(x-i-1)}J(x)^{2}dx
\displaystyle\leq CR01e2R(xi1)𝑑x=C(e2iRe2(i+1)R)Ce2iR.\displaystyle CR\int_{0}^{1}e^{2R(x-i-1)}dx=C(e^{-2iR}-e^{-2(i+1)R})\leq Ce^{-2iR}.

Combining (3.25) and (3.26), we have

y1y2(|φ~i,R|2+u′′βuc|φ~i,R|2)𝑑y=1φi,RL22y1y2(|φi,R|2+u′′βuc|φi,R|2)𝑑y\displaystyle\int_{y_{1}}^{y_{2}}\left(|\tilde{\varphi}_{i,R}^{\prime}|^{2}+{u^{\prime\prime}-\beta\over u-c}|\tilde{\varphi}_{i,R}|^{2}\right)dy={1\over\|\varphi_{i,R}\|_{L^{2}}^{2}}\int_{y_{1}}^{y_{2}}\left(|\varphi_{i,R}^{\prime}|^{2}+{u^{\prime\prime}-\beta\over u-c}|\varphi_{i,R}|^{2}\right)dy
\displaystyle\leq (Cε1R8(4ε1))1φi,RL22(Cε1R8(4ε1))e2iRC\displaystyle\left(C-{\varepsilon_{1}R\over 8(4-\varepsilon_{1})}\right){1\over\|\varphi_{i,R}\|_{L^{2}}^{2}}\leq\left(C-{\varepsilon_{1}R\over 8(4-\varepsilon_{1})}\right){e^{2iR}\over C}\to-\infty

as RR\to\infty. This proves (3.22). ∎

4. Rule out oscillation of the nn-th eigenvalue of Rayleigh-Kuo BVP

Let β[98κ,98κ+]\beta\in[{9\over 8}\kappa_{-},{9\over 8}\kappa_{+}]. By Theorem 2.9 (1)-(2), λn(c)=α2\lambda_{n}(c)=-\alpha^{2} has only finite number of solutions cc outside the range of uu for 1nmβ1\leq n\leq m_{\beta}, and no solutions exist for nNβn\geq N_{\beta}. It is non-trivial to study whether the number of solutions is finite for mβ<n<Nβm_{\beta}<n<N_{\beta}. Recall that NβN_{\beta} is obtained in Theorem 2.9 such that infc(,umin)λNβ(c)0\inf\limits_{c\in(-\infty,u_{\min})}\lambda_{N_{\beta}}(c)\geq 0 for 0<β98κ+0<\beta\leq{9\over 8}\kappa_{+}, infc(umax,)λNβ(c)0\inf\limits_{c\in(u_{\max},\infty)}\lambda_{N_{\beta}}(c)\geq 0 for 98κβ<0{9\over 8}\kappa_{-}\leq\beta<0. The main difficulty is that λn(c)\lambda_{n}(c) might oscillate when cc is close to uminu_{\min} or umaxu_{\max}. In this section, we rule out the oscillation in the following three cases.

4.1. Rule out oscillation under the spectral assumption

We rule out the oscillation of λn(c)\lambda_{n}(c) under the spectral assumption (𝐄±)\bf{(E_{\pm})}, which is stated in Theorem 2.11 (1)-(2). To this end, we first consider the compactness near a class of singular points.

Lemma 4.1.

Let cRan(u)c\in\text{Ran}(u), y0u1{c}(y1,y2)y_{0}\in u^{-1}\{c\}\cap(y_{1},y_{2}), u(y0)=0u^{\prime}(y_{0})=0 and δ>0\delta>0 so that (u′′(y0)β)(u′′(y)β)>0(u^{\prime\prime}(y_{0})-\beta)(u^{\prime\prime}(y)-\beta)>0 on [y0δ,y0+δ][y1,y2][y_{0}-\delta,y_{0}+\delta]\subset[y_{1},y_{2}] and [y0δ,y0+δ]u1{c}={y0}[y_{0}-\delta,y_{0}+\delta]\cap u^{-1}\{c\}=\{y_{0}\}. Assume that β/u′′(y0)<9/8.\beta/u^{\prime\prime}(y_{0})<9/8. Let ϕn,ωnH1(y0δ,y0+δ)\phi_{n},\omega_{n}\in H^{1}(y_{0}-\delta,y_{0}+\delta) and cn𝐂c_{n}\in\mathbf{C} so that cni>0c_{n}^{i}>0, cncc_{n}\to c, ϕn0,ωn0\phi_{n}\rightharpoonup 0,\omega_{n}\to 0 in H1(y0δ,y0+δ)H^{1}(y_{0}-\delta,y_{0}+\delta) and

(ucn)(ϕn′′α2ϕn)(u′′β)ϕn=ωn\displaystyle(u-c_{n})(\phi_{n}^{\prime\prime}-\alpha^{2}\phi_{n})-(u^{\prime\prime}-\beta)\phi_{n}=\omega_{n}

holds on [y0δ,y0+δ][y_{0}-\delta,y_{0}+\delta]. Then ϕn0\phi_{n}\to 0 in H1(y0δ,y0+δ)H^{1}(y_{0}-\delta,y_{0}+\delta).

Here cni=Im(cn)c_{n}^{i}=\text{Im}(c_{n}). The proof of Lemma 4.1 is the same as that of Lemma 3.4 in [47], where we only used the condition β/u′′(y0)<9/8\beta/u^{\prime\prime}(y_{0})<9/8 rather than the stronger condition:

(H)uH4(y1,y2),u′′(yc)0,β/u′′(yc)<9/8at critical points u(yc)=0.\displaystyle(\textbf{H})\quad u\in H^{4}(y_{1},y_{2}),\,\,u^{\prime\prime}(y_{c})\neq 0,\,\,\beta/u^{\prime\prime}(y_{c})<9/8\,\,\text{at critical points }u^{\prime}(y_{c})=0.

Since otherwise, we can construct u~\widetilde{u} such that u~H4(y1,y2),u~|[y0δ,y0+δ]=u|[y0δ,y0+δ]\widetilde{u}\in H^{4}(y_{1},y_{2}),\ \widetilde{u}|_{[y_{0}-\delta,y_{0}+\delta]}=u|_{[y_{0}-\delta,y_{0}+\delta]} and (u~)1{0}={y0}(\widetilde{u}^{\prime})^{-1}\{0\}=\{y_{0}\}. Recall that all the conditions and conclusions depend only on u|[y0δ,y0+δ]u|_{[y_{0}-\delta,y_{0}+\delta]}. Then we prove the uniform H1H^{1} bound for the eigenfunctions. More precisely, we have the following result.

Proposition 4.2.

Let 0<β<98κ+0<\beta<{9\over 8}\kappa_{+}. Assume that mβ<n<Nβm_{\beta}<n<N_{\beta}, {ck}(,umin)\{c_{k}\}\subset(-\infty,u_{\min}), ckuminc_{k}\to u_{\min}^{-} and λn(ck)=α2>0-\lambda_{n}(c_{k})=\alpha^{2}>0, where mβm_{\beta} and NβN_{\beta} are given in (1.17) and Theorem 2.9, and λn(ck)\lambda_{n}(c_{k}) is the nn-th eigenvalue of

(4.1) ψk′′+u′′βuckψk=λn(ck)ψk,ψk(y1)=ψk(y2)=0\displaystyle-\psi_{k}^{\prime\prime}+{u^{\prime\prime}-\beta\over u-c_{k}}\psi_{k}=\lambda_{n}(c_{k})\psi_{k},\;\;\;\;\psi_{k}(y_{1})=\psi_{k}(y_{2})=0

with the L2L^{2} normalized eigenfunction ψk\psi_{k}. Then

(4.2) ψkH1(y1,y2)C,k1.\displaystyle\|\psi_{k}\|_{H^{1}(y_{1},y_{2})}\leq C,\;\;k\geq 1.
Proof.

Suppose that (4.2) is not true. Up to a subsequence, we can assume that ψkH1(y1,y2)k.\|\psi_{k}\|_{H^{1}(y_{1},y_{2})}\\ \geq k. Let ψ^k=ψkψkH1(y1,y2)\hat{\psi}_{k}={\psi_{k}\over\|\psi_{k}\|_{H^{1}(y_{1},y_{2})}} on [y1,y2][y_{1},y_{2}]. Then ψ^k′′+u′′βuckψ^k=α2ψ^k-\hat{\psi}_{k}^{\prime\prime}+{u^{\prime\prime}-\beta\over u-c_{k}}\hat{\psi}_{k}=-\alpha^{2}\hat{\psi}_{k} on [y1,y2][y_{1},y_{2}], ψ^kH1(y1,y2)=1\|\hat{\psi}_{k}\|_{H^{1}(y_{1},y_{2})}\\ =1 and ψ^kL2(y1,y2)=1/ψkH1(y1,y2)1/k0\|\hat{\psi}_{k}\|_{L^{2}(y_{1},y_{2})}=1/\|\psi_{k}\|_{H^{1}(y_{1},y_{2})}\leq 1/k\to 0. Thus, ψ^k0\hat{\psi}_{k}\rightharpoonup 0 in H1(y1,y2)H^{1}(y_{1},y_{2}).

Similar to Lemma 3.1 in [47], we have ψ^k0\hat{\psi}_{k}\to 0 in H1((aδ,a+δ)[y1,y2])H^{1}((a-\delta,a+\delta)\cap[y_{1},y_{2}]) for a{u=umin}{u0}.a\in\{u=u_{\min}\}\cap\{u^{\prime}\neq 0\}. Similar to Lemma 3.5 and Remark 3.6 in [47], we have ψ^k0\hat{\psi}_{k}\to 0 in H1((aδ,a+δ)[y1,y2])H^{1}((a-\delta,a+\delta)\cap[y_{1},y_{2}]) for a{u=umin}{y1,y2}{u=0}{u′′β}.a\in\{u=u_{\min}\}\cap\{y_{1},y_{2}\}\cap\{u^{\prime}=0\}\cap\{u^{\prime\prime}\neq\beta\}. Similar to Lemma 3.7 and Remark 3.8 in [47], we have ψ^k0\hat{\psi}_{k}\to 0 in H1((aδ,a+δ)[y1,y2])H^{1}((a-\delta,a+\delta)\cap[y_{1},y_{2}]) for a{u=umin}{u=0}{u′′=β}.a\in\{u=u_{\min}\}\cap\{u^{\prime}=0\}\cap\{u^{\prime\prime}=\beta\}. The main difference is that ck𝐑c_{k}\in\mathbf{R} rather than Im(ck)>0\text{Im}({c}_{k})>0, and we can overcome this difficulty by perturbation of ck{c}_{k} as in the next case.

If a{u=umin}(y1,y2){u′′β}a\in\{u=u_{\min}\}\cap(y_{1},y_{2})\cap\{u^{\prime\prime}\neq\beta\}, then 0<β<98κ+9u′′(a)/80<\beta<{9\over 8}\kappa_{+}\leq 9u^{\prime\prime}(a)/8. Take δ(0,min(y2a,ay1))\delta\in(0,\min(y_{2}-a,a-y_{1})) small enough so that (u′′(a)β)(u′′(y)β)>0(u^{\prime\prime}(a)-\beta)(u^{\prime\prime}(y)-\beta)>0 on [aδ,a+δ][y1,y2][a-\delta,a+\delta]\subset[y_{1},y_{2}] and [aδ,a+δ]{u=umin}={a}[a-\delta,a+\delta]\cap\{u=u_{\min}\}=\{a\}. Noting that ψ^k,u′′βuckH1(aδ,a+δ)\hat{\psi}_{k},{u^{\prime\prime}-\beta\over u-c_{k}}\in H^{1}(a-\delta,a+\delta), we have ψ^k′′α2ψ^k=u′′βuckψ^kH1(aδ,a+δ),\hat{\psi}_{k}^{\prime\prime}-\alpha^{2}\hat{\psi}_{k}={u^{\prime\prime}-\beta\over u-c_{k}}\hat{\psi}_{k}\in H^{1}(a-\delta,a+\delta), and there exists ϵk>0\epsilon_{k}>0 such that ϵk(1+ψ^k′′α2ψ^kH1(aδ,a+δ))0.\epsilon_{k}(1+\|\hat{\psi}_{k}^{\prime\prime}-\alpha^{2}\hat{\psi}_{k}\|_{H^{1}(a-\delta,a+\delta)})\to 0. Let c~k=ck+iϵk\widetilde{c}_{k}=c_{k}+i\epsilon_{k} and ωk=iϵk(ψ^k′′α2ψ^k)\omega_{k}=-i\epsilon_{k}(\hat{\psi}_{k}^{\prime\prime}-\alpha^{2}\hat{\psi}_{k}). Then we have (uc~k)(ψ^k′′α2ψ^k)(u′′β)ψ^k=ωk,ωkH1(aδ,a+δ)0,c~kumin,Im(c~k)>0(u-\widetilde{c}_{k})(\hat{\psi}_{k}^{\prime\prime}-\alpha^{2}\hat{\psi}_{k})-(u^{\prime\prime}-\beta)\hat{\psi}_{k}=\omega_{k},\ \|\omega_{k}\|_{H^{1}(a-\delta,a+\delta)}\to 0,\ \widetilde{c}_{k}\to u_{\min},\ \text{Im}(\widetilde{c}_{k})>0 and ψ^k0\hat{\psi}_{k}\rightharpoonup 0 in H1(aδ,a+δ)H^{1}(a-\delta,a+\delta). By Lemma 4.1, we have ψ^k0\hat{\psi}_{k}\to 0 in H1((aδ,a+δ)[y1,y2])H^{1}((a-\delta,a+\delta)\cap[y_{1},y_{2}]). Note that ψ^k0\hat{\psi}_{k}\to 0 in Cloc2([y1,y2]{u=umin})C^{2}_{loc}([y_{1},y_{2}]\setminus\{u=u_{\min}\}). Therefore, ψ^k0\hat{\psi}_{k}\to 0 in H1(y1,y2)H^{1}(y_{1},y_{2}), which contradicts ψ^kH1(y1,y2)=1\|\hat{\psi}_{k}\|_{H^{1}(y_{1},y_{2})}=1. Thus, (4.2) is true.∎

Following Definition 3.10 in [47], we call uminu_{\min} (or umaxu_{\max}) to be an embedding eigenvalue of α,β\mathcal{R}_{\alpha,\beta} if there exists a nontrivial ψH01(y1,y2)\psi\in H_{0}^{1}(y_{1},y_{2}) such that for any φH01(y1,y2)\varphi\in H_{0}^{1}(y_{1},y_{2}) and suppφ(y1,y2){y(y1,y2):u(y)=umin,u′′(y)β}\text{supp}\ \varphi\subset(y_{1},y_{2})\setminus\{y\in(y_{1},y_{2}):u(y)=u_{\min},u^{\prime\prime}(y)\neq\beta\},

y1y2(ψφ+α2ψφ)𝑑y+p.v.y1y2(u′′β)ψφuumin𝑑y=0.\int_{y_{1}}^{y_{2}}(\psi^{\prime}\varphi^{\prime}+\alpha^{2}\psi\varphi)dy+p.v.\int_{y_{1}}^{y_{2}}{(u^{\prime\prime}-\beta)\psi\varphi\over u-u_{\min}}dy=0.

Equivalently, uminu_{\min} (or umaxu_{\max}) is an embedding eigenvalue of the linearized operator of (1.1) (in velocity form) defined on L2×L2L^{2}\times L^{2}. In fact, v=(ψ,iαψ)0\vec{v}=(\psi^{\prime},-i\alpha\psi)\neq 0 is the corresponding eigenfunction.

We are now in a position to prove Theorem 2.11 (1)-(2).

Proof of Theorem 2.11 (1)-(2).

First, we prove Theorem 2.11 (1). Suppose (σd(α,β)(,umin))=\sharp(\sigma_{d}(\mathcal{R}_{\alpha,\beta})\cap(-\infty,u_{\min}))=\infty. Then by Theorem 2.9, there exist mβ<n<Nβm_{\beta}<n<N_{\beta} and {ck}(,umin)\{c_{k}\}\subset(-\infty,u_{\min}) with ckuminc_{k}\to u_{\min}^{-} such that λn(ck)=α2\lambda_{n}(c_{k})=-\alpha^{2} is the nn-th eigenvalue of (4.1) with the L2L^{2} normalized eigenfunction ψk\psi_{k}. By the definition of MβM_{\beta} we have α2=λn(ck)Mβ\alpha^{2}=-\lambda_{n}(c_{k})\leq M_{\beta}, which implies the second statement of Theorem 2.11 (1). To prove the first statement, we now assume that 0<α2Mβ0<\alpha^{2}\leq M_{\beta} and 0<β<98κ+0<\beta<{9\over 8}\kappa_{+}. By Proposition 4.2, up to a subsequence, there exists ψ0H01(y1,y2)\psi_{0}\in H_{0}^{1}(y_{1},y_{2}) such that ψkψ0\psi_{k}\rightharpoonup\psi_{0} in H01(y1,y2)H_{0}^{1}(y_{1},y_{2}). Similar to (3.28) in [47],

limkaδa+δ(u′′β)ψkφuck𝑑y=p.v.aδa+δ(u′′β)ψ0φuumin𝑑y\lim_{k\to\infty}\int_{a-\delta}^{a+\delta}{(u^{\prime\prime}-\beta)\psi_{k}\varphi\over u-c_{k}}dy=p.v.\int_{a-\delta}^{a+\delta}{(u^{\prime\prime}-\beta)\psi_{0}\varphi\over u-u_{\min}}dy

for any a{u=umin}(y1,y2){u′′=β}a\in\{u=u_{\min}\}\cap(y_{1},y_{2})\cap\{u^{\prime\prime}=\beta\} and φH01(aδ,a+δ)\varphi\in H_{0}^{1}(a-\delta,a+\delta). Since (u′′β)ψkφuck(u′′β)ψ0φuumin{(u^{\prime\prime}-\beta)\psi_{k}\varphi\over u-{c_{k}}}\to{(u^{\prime\prime}-\beta)\psi_{0}\varphi\over u-u_{\min}} in Cloc0((y1,y2){u=umin})C^{0}_{\text{loc}}((y_{1},y_{2})\setminus\{u=u_{\min}\}), taking limits in

y1y2(ψkφ+α2ψkφ)+(u′′β)ψkφuckdy=0\int_{y_{1}}^{y_{2}}(\psi_{k}^{\prime}\varphi^{\prime}+\alpha^{2}\psi_{k}\varphi)+{(u^{\prime\prime}-\beta)\psi_{k}\varphi\over u-c_{k}}dy=0

for any φH01(y1,y2)\varphi\in H_{0}^{1}(y_{1},y_{2}) and suppφ(y1,y2){y(y1,y2):u(y)=umin,u′′(y)β}\text{supp}\ \varphi\subset(y_{1},y_{2})\setminus\{y\in(y_{1},y_{2}):u(y)=u_{\min},u^{\prime\prime}(y)\neq\beta\}, we get

y1y2(ψ0φ+α2ψ0φ)𝑑y+p.v.y1y2(u′′β)ψ0φuumin𝑑y=0.\int_{y_{1}}^{y_{2}}(\psi_{0}^{\prime}\varphi^{\prime}+\alpha^{2}\psi_{0}\varphi)dy+p.v.\int_{y_{1}}^{y_{2}}{(u^{\prime\prime}-\beta)\psi_{0}\varphi\over u-u_{\min}}dy=0.

If ψ0\psi_{0} is nontrivial, uminu_{\min} is an embedding eigenvalue of α,β\mathcal{R}_{\alpha,\beta}, which is a contradiction. Therefore, ψkψ00\psi_{k}\rightharpoonup\psi_{0}\equiv 0 in H1(y1,y2)H^{1}(y_{1},y_{2}), which contradicts that ψkL2(y1,y2)=1\|\psi_{k}\|_{L^{2}(y_{1},y_{2})}=1, k1k\geq 1. Thus, (σd(α,β)(,umin))<\sharp(\sigma_{d}(\mathcal{R}_{\alpha,\beta})\cap(-\infty,u_{\min}))<\infty. Theorem 2.11 (2) can be proved similarly. ∎

4.2. Rule out oscillation under “good” endpoints assumption

We rule out the oscillation of λn(c)\lambda_{n}(c) under the “good” endpoints assumption (i.e. one of the conditions (i)–(iii) in Theorem 2.2). The statement is given in Theorem 2.12. To this end, we need the following two lemmas.

Lemma 4.3.

Let uC2([y1,y2])u\in C^{2}([y_{1},y_{2}]), u(y1)=uminu(y_{1})=u_{\min} and u(y1)0u^{\prime}(y_{1})\neq 0. For fixed γ(0,1/2],\gamma\in(0,1/2], there exist constants C>0C>0 and δ(0,y2y1)\delta\in(0,y_{2}-y_{1}) such that if δ1(0,δ],z=y1+δ1, 0<uminc<1,\delta_{1}\in(0,\delta],\ z=y_{1}+\delta_{1},\ 0<u_{\min}-c<1, ϕC2([y1,z])\phi\in C^{2}([y_{1},z]) and ϕ′′=F\phi^{\prime\prime}=F, then

(4.3) |ϕ|L(z)C(δ1γ|(uc)2γF|L(z)+|ϕ(z)|+δ1γ|ϕ(z)|),\displaystyle|\phi|_{L^{\infty}(z)}\leq C(\delta_{1}^{\gamma}|(u-c)^{2-\gamma}F|_{L^{\infty}(z)}+|\phi(z)|+\delta_{1}^{\gamma}|\phi^{\prime}(z)|),
(4.4) |(uc)1γϕ|L(z)+δ1γ|(uc)22γϕ|L(z)\displaystyle|(u-c)^{1-\gamma}\phi|_{L^{\infty}(z)}+\delta_{1}^{\gamma}|(u-c)^{2-2\gamma}\phi^{\prime}|_{L^{\infty}(z)}
\displaystyle\leq C(δ1γ|(uc)32γF|L(z)+|ϕ(z)|+δ1γ|ϕ(z)|),\displaystyle C(\delta_{1}^{\gamma}|(u-c)^{3-2\gamma}F|_{L^{\infty}(z)}+|\phi(z)|+\delta_{1}^{\gamma}|\phi^{\prime}(z)|),
(4.5) |(uc)γ1ϕ|L(z)\displaystyle|(u-c)^{\gamma-1}\phi|_{L^{\infty}(z)}
\displaystyle\leq C(|(uc)γ+1F|L(z)+|(uminc)γ1ϕ(y1)|+|ϕ(z)|+|ϕ(z)|),\displaystyle C(|(u-c)^{\gamma+1}F|_{L^{\infty}(z)}+|(u_{\min}-c)^{\gamma-1}\phi(y_{1})|+|\phi(z)|+|\phi^{\prime}(z)|),
(4.6) |(uc)1ϕ|L(z)C(δ1γ|(uc)1γF|L(z)+|(uminc)1ϕ(y1)|+|ϕ(z)|),\displaystyle|(u-c)^{-1}\phi|_{L^{\infty}(z)}\leq C(\delta_{1}^{\gamma}|(u-c)^{1-\gamma}F|_{L^{\infty}(z)}+|(u_{\min}-c)^{-1}\phi(y_{1})|+|\phi^{\prime}(z)|),

where |f|L(z):=supy[y1,z]|f(y)||f|_{L^{\infty}(z)}:=\sup_{y\in[y_{1},z]}|f(y)|.

Proof.

Since 0<uminc<10<u_{\min}-c<1 and Cuminu(y)C,-C\leq u_{\min}\leq u(y)\leq C, we have 0<u(y)cC0<u(y)-c\leq C for y[y1,z].y\in[y_{1},z]. Let Aμ=|(uc)μF|L(z)A_{\mu}=|(u-c)^{\mu}F|_{L^{\infty}(z)} and Bμ=|(uc)μϕ|L(z)B_{\mu}=|(u-c)^{\mu}\phi^{\prime}|_{L^{\infty}(z)} for μ𝐑\mu\in\mathbf{R}. Let δ>0\delta>0 be small enough such that u(y)>u(y1)2>0u^{\prime}(y)>{u^{\prime}(y_{1})\over 2}>0 for y[y1,z][y1,y1+δ][y1,y2].y\in[y_{1},z]\subset[y_{1},y_{1}+\delta]\subset[y_{1},y_{2}]. Then

yz(u(s)c)μ𝑑s2u(y1)yzu(s)(u(s)c)μ𝑑s=2(u(s)c)1μ|s=zs=yu(y1)(μ1)2(u(y)c)1μu(y1)(μ1)\displaystyle\int_{y}^{z}(u(s)-c)^{-\mu}ds\leq\frac{2}{u^{\prime}(y_{1})}\int_{y}^{z}u^{\prime}(s)(u(s)-c)^{-\mu}ds=\frac{2(u(s)-c)^{1-\mu}|_{s=z}^{s=y}}{u^{\prime}(y_{1})(\mu-1)}\leq\frac{2(u(y)-c)^{1-\mu}}{u^{\prime}(y_{1})(\mu-1)}

for fixed μ>1\mu>1 and y[y1,z]y\in[y_{1},z], and thus

(4.7) yz(u(s)c)μ𝑑s2(u(y)c)1μu(y1)(μ1)C(u(y)c)1μ.\displaystyle\int_{y}^{z}(u(s)-c)^{-\mu}ds\leq\frac{2(u(y)-c)^{1-\mu}}{u^{\prime}(y_{1})(\mu-1)}\leq C(u(y)-c)^{1-\mu}.

Similarly, for fixed μ<1\mu<1 and y[y1,z]y\in[y_{1},z], we have

(4.8) y1y(u(s)c)μ𝑑s2(u(s)c)1μ|s=y1s=yu(y1)(1μ)C(u(y)c)1μ.\displaystyle\int_{y_{1}}^{y}(u(s)-c)^{-\mu}ds\leq\frac{2(u(s)-c)^{1-\mu}|_{s=y_{1}}^{s=y}}{u^{\prime}(y_{1})(1-\mu)}\leq C(u(y)-c)^{1-\mu}.

Since u(s)cu(s)uminu(y1)(sy1)/2u(s)-c\geq u(s)-u_{\min}\geq u^{\prime}(y_{1})(s-y_{1})/2 and u(s)cu(y)cu(s)-c\geq u(y)-c for y1ysz,y_{1}\leq y\leq s\leq z, we have for fixed μ1γ,\mu\geq 1-\gamma,

(4.9) yz(u(s)c)μ𝑑syz(u(y1)(sy1)/2)γ1(u(y)c)1γμ𝑑s\displaystyle\int_{y}^{z}(u(s)-c)^{-\mu}ds\leq\int_{y}^{z}(u^{\prime}(y_{1})(s-y_{1})/2)^{\gamma-1}(u(y)-c)^{1-\gamma-\mu}ds
\displaystyle\leq C(u(y)c)1γμyz(sy1)γ1𝑑sC(u(y)c)1γμ(zy1)γ\displaystyle C(u(y)-c)^{1-\gamma-\mu}\int_{y}^{z}(s-y_{1})^{\gamma-1}ds\leq C(u(y)-c)^{1-\gamma-\mu}(z-y_{1})^{\gamma}
=\displaystyle= Cδ1γ(u(y)c)1γμ.\displaystyle C\delta_{1}^{\gamma}(u(y)-c)^{1-\gamma-\mu}.

For fixed μ>1,\mu>1, using (4.7) and the definition of AμA_{\mu}, we have for y[y1,z]y\in[y_{1},z],

|ϕ(y)ϕ(z)|yz|ϕ(s)|ds=yz|F(s)|dsyz(u(s)c)μAμdsC(u(y)c)1μAμ,\displaystyle|\phi^{\prime}(y)-\phi^{\prime}(z)|\leq\int_{y}^{z}|\phi^{\prime\prime}(s)|ds=\int_{y}^{z}|F(s)|ds\leq\int_{y}^{z}(u(s)-c)^{-\mu}A_{\mu}ds\leq C(u(y)-c)^{1-\mu}A_{\mu},

and

|ϕ(y)||ϕ(z)|+C(u(y)c)1μAμ,\displaystyle|\phi^{\prime}(y)|\leq|\phi^{\prime}(z)|+C(u(y)-c)^{1-\mu}A_{\mu},
|(u(y)c)μ1ϕ(y)|(u(y)c)μ1|ϕ(z)|+CAμC|ϕ(z)|+CAμ.\displaystyle|(u(y)-c)^{\mu-1}\phi^{\prime}(y)|\leq(u(y)-c)^{\mu-1}|\phi^{\prime}(z)|+CA_{\mu}\leq C|\phi^{\prime}(z)|+CA_{\mu}.

Then by the definition of BμB_{\mu}, we have

(4.10) Bμ1=supy[y1,z]|(u(y)c)μ1ϕ(y)|CAμ+C|ϕ(z)|for fixedμ>1.\displaystyle B_{\mu-1}=\sup_{y\in[y_{1},z]}|(u(y)-c)^{\mu-1}\phi^{\prime}(y)|\leq CA_{\mu}+C|\phi^{\prime}(z)|\ \ \text{for fixed}\ \mu>1.

Similarly, for fixed μ1γ,\mu\geq 1-\gamma, using (4.9) and the definition of AμA_{\mu}, we have for y[y1,z]y\in[y_{1},z],

|ϕ(y)ϕ(z)|yz|ϕ(s)|dsyz(u(s)c)μBμdsCδ1γ(u(y)c)1γμBμ.\displaystyle|\phi(y)-\phi(z)|\leq\int_{y}^{z}|\phi^{\prime}(s)|ds\leq\int_{y}^{z}(u(s)-c)^{-\mu}B_{\mu}ds\leq C\delta_{1}^{\gamma}(u(y)-c)^{1-\gamma-\mu}B_{\mu}.

This implies

(4.11) |(uc)μ+γ1ϕ|L(z)Cδ1γBμ+C|ϕ(z)|for fixedμ1γ.\displaystyle|(u-c)^{\mu+\gamma-1}\phi|_{L^{\infty}(z)}\leq C\delta_{1}^{\gamma}B_{\mu}+C|\phi(z)|\ \ \text{for fixed}\ \mu\geq 1-\gamma.

Using (4.11) for μ=1γ\mu=1-\gamma and (4.10) for μ=2γ\mu=2-\gamma, we have

|ϕ|L(z)Cδ1γB1γ+C|ϕ(z)|Cδ1γ(A2γ+|ϕ(z)|)+C|ϕ(z)|,\displaystyle|\phi|_{L^{\infty}(z)}\leq C\delta_{1}^{\gamma}B_{1-\gamma}+C|\phi(z)|\leq C\delta_{1}^{\gamma}(A_{2-\gamma}+|\phi^{\prime}(z)|)+C|\phi(z)|,

which implies (4.3) by recalling the definition of AμA_{\mu}. Using (4.11) for μ=22γ\mu=2-2\gamma and (4.10) for μ=32γ\mu=3-2\gamma, we have

|(uc)1γϕ|L(z)+δ1γB22γCδ1γB22γ+C|ϕ(z)|Cδ1γ(A32γ+|ϕ(z)|)+C|ϕ(z)|,\displaystyle|(u-c)^{1-\gamma}\phi|_{L^{\infty}(z)}+\delta_{1}^{\gamma}B_{2-2\gamma}\leq C\delta_{1}^{\gamma}B_{2-2\gamma}+C|\phi(z)|\leq C\delta_{1}^{\gamma}(A_{3-2\gamma}+|\phi^{\prime}(z)|)+C|\phi(z)|,

which implies (4.4) by recalling the definition of AμA_{\mu} and BμB_{\mu}.

For fixed μ<1,\mu<1, using (4.8) and the definition of AμA_{\mu}, we have for y[y1,z]y\in[y_{1},z],

|ϕ(y)ϕ(y1)|y1y|ϕ(s)|dsy1y(u(s)c)μBμdsC(u(y)c)1μBμ,\displaystyle|\phi(y)-\phi(y_{1})|\leq\int_{y_{1}}^{y}|\phi^{\prime}(s)|ds\leq\int_{y_{1}}^{y}(u(s)-c)^{-\mu}B_{\mu}ds\leq C(u(y)-c)^{1-\mu}B_{\mu},

and

|ϕ(y)||ϕ(y1)|+C(u(y)c)1μBμ,\displaystyle|\phi(y)|\leq|\phi(y_{1})|+C(u(y)-c)^{1-\mu}B_{\mu},
|(u(y)c)μ1ϕ(y)|(u(y)c)μ1|ϕ(y1)|+CBμ(uminc)μ1|ϕ(y1)|+CBμ,\displaystyle|(u(y)-c)^{\mu-1}\phi(y)|\leq(u(y)-c)^{\mu-1}|\phi(y_{1})|+CB_{\mu}\leq(u_{\min}-c)^{\mu-1}|\phi(y_{1})|+CB_{\mu},

where we used u(y)cuminc>0u(y)-c\geq u_{\min}-c>0 and μ1<0.\mu-1<0. Thus,

(4.12) |(uc)μ1ϕ|L(z)(uminc)μ1|ϕ(y1)|+CBμfor fixedμ<1.\displaystyle|(u-c)^{\mu-1}\phi|_{L^{\infty}(z)}\leq(u_{\min}-c)^{\mu-1}|\phi(y_{1})|+CB_{\mu}\ \ \text{for fixed}\ \mu<1.

Using (4.12) for μ=γ\mu=\gamma and (4.10) for μ=1+γ\mu=1+\gamma, we have

|(uc)γ1ϕ|L(z)\displaystyle|(u-c)^{\gamma-1}\phi|_{L^{\infty}(z)} (uminc)γ1|ϕ(y1)|+CBγ\displaystyle\leq(u_{\min}-c)^{\gamma-1}|\phi(y_{1})|+CB_{\gamma}
(uminc)γ1|ϕ(y1)|+C(A1+γ+|ϕ(z)|),\displaystyle\leq(u_{\min}-c)^{\gamma-1}|\phi(y_{1})|+C(A_{1+\gamma}+|\phi^{\prime}(z)|),

which implies (4.5) by recalling the definition of AμA_{\mu}. Using (4.9) for μ=1γ\mu=1-\gamma and the definition of AμA_{\mu}, we have for y[y1,z]y\in[y_{1},z],

|ϕ(y)ϕ(z)|yz|ϕ(s)|ds=yz|F(s)|dsyz(u(s)c)γ1A1γdsCδ1γA1γ.\displaystyle|\phi^{\prime}(y)-\phi^{\prime}(z)|\leq\int_{y}^{z}|\phi^{\prime\prime}(s)|ds=\int_{y}^{z}|F(s)|ds\leq\int_{y}^{z}(u(s)-c)^{\gamma-1}A_{1-\gamma}ds\leq C\delta_{1}^{\gamma}A_{1-\gamma}.

Then by the definition of BμB_{\mu}, we have

(4.13) B0=supy[y1,z]|ϕ(y)|Cδ1γA1γ+C|ϕ(z)|.\displaystyle B_{0}=\sup_{y\in[y_{1},z]}|\phi^{\prime}(y)|\leq C\delta_{1}^{\gamma}A_{1-\gamma}+C|\phi^{\prime}(z)|.

Using (4.12) for μ=0\mu=0 and (4.13), we have

|(uc)1ϕ|L(z)\displaystyle|(u-c)^{-1}\phi|_{L^{\infty}(z)} (uminc)1|ϕ(y1)|+CB0(uminc)1|ϕ(y1)|+C(δ1γA1γ+|ϕ(z)|),\displaystyle\leq(u_{\min}-c)^{-1}|\phi(y_{1})|+CB_{0}\leq(u_{\min}-c)^{-1}|\phi(y_{1})|+C(\delta_{1}^{\gamma}A_{1-\gamma}+|\phi^{\prime}(z)|),

which implies (4.6) by recalling the definition of AμA_{\mu}. ∎

Lemma 4.4.

Let uC2([y1,y2])u\in C^{2}([y_{1},y_{2}]), u(y1)=uminu(y_{1})=u_{\min} and u(y1)0u^{\prime}(y_{1})\neq 0. For fixed γ(0,1/2],\gamma\in(0,1/2], there exist constants C>0C>0 and δ1>0\delta_{1}>0 such that if z=y1+δ1, 0<uminc<1,z=y_{1}+\delta_{1},\ 0<u_{\min}-c<1, ϕC2([y1,z])\phi\in C^{2}([y_{1},z]) and

(4.14) ϕ+uβucϕ=α2ϕFon[y1,z],\displaystyle-\phi^{\prime\prime}+{u^{\prime\prime}-\beta\over u-c}\phi=-\alpha^{2}\phi-F\ \ \text{on}\ [y_{1},z],

then the inequalities (4.3)–(4.6) are still true.

Proof.

Let F~=uβucϕ+α2ϕ+F\widetilde{F}={u^{\prime\prime}-\beta\over u-c}\phi+\alpha^{2}\phi+F and δ1(0,δ]\delta_{1}\in(0,\delta] be given in Lemma 4.3. Then ϕ=F~\phi^{\prime\prime}=\widetilde{F} on [y1,z][y_{1},z]. By Lemma 4.3, (4.3)–(4.6) are still true with FF replaced by F~\widetilde{F}. As |uβ|C|u^{\prime\prime}-\beta|\leq C and |uc|C|u-c|\leq C, we have |FF~|C|ϕ/(uc)||F-\widetilde{F}|\leq C|\phi/(u-c)| for y[y1,z]y\in[y_{1},z]. Thus, for μ𝐑\mu\in\mathbf{R}, we have

(4.15) |(uc)μF~|L(z)\displaystyle|(u-c)^{\mu}\widetilde{F}|_{L^{\infty}(z)} |(uc)μF|L(z)+|(uc)μ(FF~)|L(z)\displaystyle\leq|(u-c)^{\mu}{F}|_{L^{\infty}(z)}+|(u-c)^{\mu}(F-\widetilde{F})|_{L^{\infty}(z)}
|(uc)μF|L(z)+C|(uc)μ1ϕ|L(z).\displaystyle\leq|(u-c)^{\mu}{F}|_{L^{\infty}(z)}+C|(u-c)^{\mu-1}\phi|_{L^{\infty}(z)}.

Using (4.3) with FF replaced by F~\widetilde{F} and (4.15) for μ=2γ\mu=2-\gamma, we have

(4.16) |ϕ|L(z)C(δ1γ|(uc)2γF~|L(z)+|ϕ(z)|+δ1γ|ϕ(z)|)\displaystyle|\phi|_{L^{\infty}(z)}\leq C(\delta_{1}^{\gamma}|(u-c)^{2-\gamma}\widetilde{F}|_{L^{\infty}(z)}+|\phi(z)|+\delta_{1}^{\gamma}|\phi^{\prime}(z)|)
\displaystyle\leq Cδ1γ|(uc)1γϕ|L(z)+C(δ1γ|(uc)2γF|L(z)+|ϕ(z)|+δ1γ|ϕ(z)|)\displaystyle C\delta_{1}^{\gamma}|(u-c)^{1-\gamma}\phi|_{L^{\infty}(z)}+C(\delta_{1}^{\gamma}|(u-c)^{2-\gamma}{F}|_{L^{\infty}(z)}+|\phi(z)|+\delta_{1}^{\gamma}|\phi^{\prime}(z)|)
\displaystyle\leq C1δ1γ|ϕ|L(z)+C(δ1γ|(uc)2γF|L(z)+|ϕ(z)|+δ1γ|ϕ(z)|).\displaystyle C_{1}\delta_{1}^{\gamma}|\phi|_{L^{\infty}(z)}+C(\delta_{1}^{\gamma}|(u-c)^{2-\gamma}{F}|_{L^{\infty}(z)}+|\phi(z)|+\delta_{1}^{\gamma}|\phi^{\prime}(z)|).

Using (4.4) with FF replaced by F~\widetilde{F} and (4.15) for μ=32γ\mu=3-2\gamma, we have

(4.17) |(uc)1γϕ|L(z)+δ1γ|(uc)22γϕ|L(z)\displaystyle|(u-c)^{1-\gamma}\phi|_{L^{\infty}(z)}+\delta_{1}^{\gamma}|(u-c)^{2-2\gamma}\phi^{\prime}|_{L^{\infty}(z)}
\displaystyle\leq C(δ1γ|(uc)32γF~|L(z)+|ϕ(z)|+δ1γ|ϕ(z)|)\displaystyle C(\delta_{1}^{\gamma}|(u-c)^{3-2\gamma}\widetilde{F}|_{L^{\infty}(z)}+|\phi(z)|+\delta_{1}^{\gamma}|\phi^{\prime}(z)|)
\displaystyle\leq C(δ1γ|(uc)32γF|L(z)+|ϕ(z)|+δ1γ|ϕ(z)|)+Cδ1γ|(uc)22γϕ|L(z)\displaystyle C(\delta_{1}^{\gamma}|(u-c)^{3-2\gamma}F|_{L^{\infty}(z)}+|\phi(z)|+\delta_{1}^{\gamma}|\phi^{\prime}(z)|)+C\delta_{1}^{\gamma}|(u-c)^{2-2\gamma}\phi|_{L^{\infty}(z)}
\displaystyle\leq C(δ1γ|(uc)32γF|L(z)+|ϕ(z)|+δ1γ|ϕ(z)|)+C1δ1γ|(uc)1γϕ|L(z).\displaystyle C(\delta_{1}^{\gamma}|(u-c)^{3-2\gamma}F|_{L^{\infty}(z)}+|\phi(z)|+\delta_{1}^{\gamma}|\phi^{\prime}(z)|)+C_{1}\delta_{1}^{\gamma}|(u-c)^{1-\gamma}\phi|_{L^{\infty}(z)}.

Using (4.6) with FF replaced by F~\widetilde{F} and (4.15) for μ=1γ\mu=1-\gamma, we have

(4.18) |(uc)1ϕ|L(z)C(δ1γ|(uc)1γF~|L(z)+|(uminc)1ϕ(y1)|+|ϕ(z)|)\displaystyle|(u-c)^{-1}\phi|_{L^{\infty}(z)}\leq C(\delta_{1}^{\gamma}|(u-c)^{1-\gamma}\widetilde{F}|_{L^{\infty}(z)}+|(u_{\min}-c)^{-1}\phi(y_{1})|+|\phi^{\prime}(z)|)
\displaystyle\leq C(δ1γ|(uc)1γF|L(z)+|(uminc)1ϕ(y1)|+|ϕ(z)|)+Cδ1γ|(uc)γϕ|L(z)\displaystyle C(\delta_{1}^{\gamma}|(u-c)^{1-\gamma}F|_{L^{\infty}(z)}+|(u_{\min}-c)^{-1}\phi(y_{1})|+|\phi^{\prime}(z)|)+C\delta_{1}^{\gamma}|(u-c)^{-\gamma}\phi|_{L^{\infty}(z)}
\displaystyle\leq C(δ1γ|(uc)1γF|L(z)+|(uminc)1ϕ(y1)|+|ϕ(z)|)+C1δ1γ|(uc)1ϕ|L(z).\displaystyle C(\delta_{1}^{\gamma}|(u-c)^{1-\gamma}F|_{L^{\infty}(z)}+|(u_{\min}-c)^{-1}\phi(y_{1})|+|\phi^{\prime}(z)|)+C_{1}\delta_{1}^{\gamma}|(u-c)^{-1}\phi|_{L^{\infty}(z)}.

Here, C1>0C_{1}>0 is a constant depending only on γ,α,β,u,δ\gamma,\ \alpha,\ \beta,\ u,\ \delta (and independent of δ1\delta_{1}). Taking δ1(0,δ]\delta_{1}\in(0,\delta] small enough such that C1δ1γ1/2C_{1}\delta_{1}^{\gamma}\leq 1/2 in (4.16)–(4.18), we obtain (4.3), (4.4) and (4.6).

Note that γ>0\gamma>0 and 2γγ+1.2-\gamma\geq\gamma+1. Using (4.15) for μ=γ+1\mu=\gamma+1 and (4.3), we have

(4.19) |(uc)γ+1F~|L(z)|(uc)γ+1F|L(z)+C|(uc)γϕ|L(z)\displaystyle|(u-c)^{\gamma+1}\widetilde{F}|_{L^{\infty}(z)}\leq|(u-c)^{\gamma+1}F|_{L^{\infty}(z)}+C|(u-c)^{\gamma}\phi|_{L^{\infty}(z)}
\displaystyle\leq |(uc)γ+1F|L(z)+C|ϕ|L(z)\displaystyle|(u-c)^{\gamma+1}F|_{L^{\infty}(z)}+C|\phi|_{L^{\infty}(z)}
\displaystyle\leq |(uc)γ+1F|L(z)+C(δ1γ|(uc)2γF|L(z)+|ϕ(z)|+δ1γ|ϕ(z)|)\displaystyle|(u-c)^{\gamma+1}F|_{L^{\infty}(z)}+C(\delta_{1}^{\gamma}|(u-c)^{2-\gamma}F|_{L^{\infty}(z)}+|\phi(z)|+\delta_{1}^{\gamma}|\phi^{\prime}(z)|)
\displaystyle\leq C(|(uc)γ+1F|L(z)+|ϕ(z)|+|ϕ(z)|).\displaystyle C(|(u-c)^{\gamma+1}F|_{L^{\infty}(z)}+|\phi(z)|+|\phi^{\prime}(z)|).

Using (4.5) with FF replaced by F~\widetilde{F} and (4.19), we have

|(uc)γ1ϕ|L(z)\displaystyle|(u-c)^{\gamma-1}\phi|_{L^{\infty}(z)}
\displaystyle\leq C(|(uc)γ+1F~|L(z)+|(uminc)γ1ϕ(y1)|+|ϕ(z)|+|ϕ(z)|)\displaystyle C(|(u-c)^{\gamma+1}\widetilde{F}|_{L^{\infty}(z)}+|(u_{\min}-c)^{\gamma-1}\phi(y_{1})|+|\phi(z)|+|\phi^{\prime}(z)|)
\displaystyle\leq C(|(uc)γ+1F|L(z)+|(uminc)γ1ϕ(y1)|+|ϕ(z)|+|ϕ(z)|).\displaystyle C(|(u-c)^{\gamma+1}F|_{L^{\infty}(z)}+|(u_{\min}-c)^{\gamma-1}\phi(y_{1})|+|\phi(z)|+|\phi^{\prime}(z)|).

Thus, (4.5) is also true. ∎

We are now in a position to prove Theorem 2.12.

Proof of Theorem 2.12.

We only prove (1), and the proof of (2) is similar. If {u=0}{u=umin}\{u^{\prime}=0\}\cap\{u=u_{\min}\}\neq\emptyset, then 0<κ+<0<\kappa_{+}<\infty. By Theorem 2.11 (3), (σd(α,β)𝐑)=\sharp(\sigma_{d}(\mathcal{R}_{\alpha,\beta})\cap\mathbf{R})=\infty for β>98κ+\beta>{9\over 8}\kappa_{+}. If {u=0}{u=umin}=\{u^{\prime}=0\}\cap\{u=u_{\min}\}=\emptyset, then {u=umin}{y1,y2}\{u=u_{\min}\}\subset\{y_{1},y_{2}\} (i.e. u=uminu=u_{\min} can be achieved only at the endpoints). We assume that u(y1)=uminu(y_{1})=u_{\min}. Then u(y2)>u(y1)u(y_{2})>u(y_{1}) and u(y1)>0u^{\prime}(y_{1})>0. By taking δ(0,y2y1)\delta\in(0,y_{2}-y_{1}) smaller, we can assume that u>u(y1)2u^{\prime}>{u^{\prime}(y_{1})\over 2} on y[y1,y1+δ]y\in[y_{1},y_{1}+\delta]. Let ψc,c𝐂Ran(u)\psi_{c},c\in\mathbf{C}\setminus\text{Ran}(u), be the solution of

(4.20) y2ψc+uβucψc=α2ψcon[y1,y2],ψc(y2)=0,yψc(y2)=1.\displaystyle-\partial_{y}^{2}\psi_{c}+{u^{\prime\prime}-\beta\over u-c}\psi_{c}=-\alpha^{2}\psi_{c}\ \ \text{on}\ \ [y_{1},y_{2}],\ \psi_{c}(y_{2})=0,\ \partial_{y}\psi_{c}(y_{2})=1.

Note that for c(,umin)c\in(-\infty,u_{\min}), cσd(α,β)c\in\sigma_{d}(\mathcal{R}_{\alpha,\beta}) if and only if ψc(y1)=0\psi_{c}(y_{1})=0. Suppose that σd(α,β)(,umin)={ck}k=1\sigma_{d}(\mathcal{R}_{\alpha,\beta})\cap(-\infty,u_{\min})=\{c_{k}\}_{k=1}^{\infty}. Then ckuminc_{k}\to u_{\min}^{-}. Note that if (iii) is true for i=1i=1, then β10\beta_{1}\leq 0 and thus ββ1\beta\neq\beta_{1} for all β>0\beta>0. So we divide the discussion into two cases.

Case 1. β=β1\beta=\beta_{1} and (i) holds for i=1i=1.

Case 2. β=β1\beta=\beta_{1} and (ii) holds for i=1i=1; or ββ1\beta\neq\beta_{1}.

If Case 1 is true, then uβ=0u^{\prime\prime}-\beta=0 on [y1,y1+δ][y_{1},y_{1}+\delta]. By (4.20), ψc\psi_{c} can be extended to an analytic function in 𝐂u([y1+δ,y2]).\mathbf{C}\setminus u([y_{1}+\delta,y_{2}]). Since uminu([y1+δ,y2]),u_{\min}\not\in u([y_{1}+\delta,y_{2}]), ψc(y1)\psi_{c}(y_{1}) has a finite number of zeros in a neighborhood of c=uminc=u_{\min}, which contradicts that (σd(α,β)(,umin))=\sharp(\sigma_{d}(\mathcal{R}_{\alpha,\beta})\cap(-\infty,u_{\min}))={\infty}.

Now, we assume Case 2 is true. If ββ1\beta\neq\beta_{1}, define m=0;m=0; if β=β1\beta=\beta_{1} and (ii) is true, define m=m1.m=m_{1}. Then m0,m\geq 0, by taking δ>0\delta>0 smaller, we can assume that

(4.21) C1|yy1|m|u(y)β|C|yy1|mfory[y1,y1+δ].\displaystyle C^{-1}|y-y_{1}|^{m}\leq|u^{\prime\prime}(y)-\beta|\leq C|y-y_{1}|^{m}\quad\text{for}\quad y\in[y_{1},y_{1}+\delta].

As |u(y)β|/|yy1|mC([y1+δ,y2]),|u^{\prime\prime}(y)-\beta|/|y-y_{1}|^{m}\in C([y_{1}+\delta,y_{2}]), |u(y)β|C|yy1|m|u^{\prime\prime}(y)-\beta|\leq C|y-y_{1}|^{m} is also true for y[y1+δ,y2]y\in[y_{1}+\delta,y_{2}]. Since uC1([y1,y2])u\in C^{1}([y_{1},y_{2}]), we have u(y)umin=u(y)u(y1)=y1yu(z)dz=(yy1)v(y)u(y)-u_{\min}=u(y)-u(y_{1})=\int_{y_{1}}^{y}u^{\prime}(z)dz=(y-y_{1})v(y), here v(y)=01u(y1+s(yy1))dsv(y)=\int_{0}^{1}u^{\prime}(y_{1}+s(y-y_{1}))ds and vC([y1,y2])v\in C([y_{1},y_{2}]). If y(y1,y2]y\in(y_{1},y_{2}], we have u(y)>uminu(y)>u_{\min} and v(y)>0v(y)>0. If y=y1y=y_{1}, we have v(y)=u(y1)>0v(y)=u^{\prime}(y_{1})>0. Thus, v(y)>0v(y)>0 in [y1,y2][y_{1},y_{2}] and there exists a constant C>1C>1 such that C1v(y)C,C^{-1}\leq v(y)\leq C, which implies

(4.22) C1|yy1|u(y)uminC|yy1|,\displaystyle C^{-1}|y-y_{1}|\leq u(y)-u_{\min}\leq C|y-y_{1}|,
(4.23) C1|yy1|u(y)cforcumin.\displaystyle C^{-1}|y-y_{1}|\leq u(y)-c\quad\text{for}\quad c\leq u_{\min}.

Let n𝐍n\in\mathbf{N} and ψc,n=cnψc.\psi_{c,n}=\partial_{c}^{n}\psi_{c}. By Rolle’s Theorem, there exists {ck,n}k=1(,umin)\{c_{k,n}\}_{k=1}^{\infty}\subset(-\infty,u_{\min}) such that ψck,n,n(y1)=0\psi_{c_{k,n},n}(y_{1})=0 and ck,numinc_{k,n}\to u_{\min}^{-} as kk\to\infty. For fixed c<umin,c<u_{\min}, let k>0k>0 be large enough such that ck,n(c,umin).c_{k,n}\in(c,u_{\min}). Then ψc,n(y1)=ck,ncsψs,n(y1)ds=ck,ncψs,n+1(y1)ds,\psi_{c,n}(y_{1})=\int_{c_{k,n}}^{c}\partial_{s}\psi_{s,n}(y_{1})ds=\int_{c_{k,n}}^{c}\psi_{s,n+1}(y_{1})ds, and

(4.24) |ψc,n(y1)|cck,n|ψs,n+1(y1)|dscumin|ψs,n+1(y1)|ds.\displaystyle|\psi_{c,n}(y_{1})|\leq\int_{c}^{c_{k,n}}|\psi_{s,n+1}(y_{1})|ds\leq\int_{c}^{u_{\min}}|\psi_{s,n+1}(y_{1})|ds.

Moreover, ψc,n\psi_{c,n} satisfies

(4.25) y2ψc,n+uβucψc,n=α2ψc,nFc,n,Fc,n=k=1nn!(nk)!uβ(uc)k+1ψc,nk,\displaystyle-\partial_{y}^{2}\psi_{c,n}+{u^{\prime\prime}-\beta\over u-c}\psi_{c,n}=-\alpha^{2}\psi_{c,n}-F_{c,n},\quad F_{c,n}=\sum_{k=1}^{n}{n!\over(n-k)!}{u^{\prime\prime}-\beta\over(u-c)^{k+1}}\psi_{c,n-k},

where ψc,0=ψc\psi_{c,0}=\psi_{c} and Fc,0=0F_{c,0}=0. Note that ψc(y)\psi_{c}(y) is continuous on (𝐂Ran(u))×[y1,y2](\mathbf{C}\setminus\text{Ran}(u))\times[y_{1},y_{2}], and analytic in cc. Let u+(y)=infu([y,y2])u_{+}(y)=\inf u([y,y_{2}]). Then ψc(y)\psi_{c}(y) can be extended to a continuous function on D1:={(c,y):c<u+(y),y[y1,y2]},D_{1}:=\{(c,y):c<u_{+}(y),\ y\in[y_{1},y_{2}]\}, still satisfying (4.20) in D1.D_{1}. Moreover, u+u_{+} is increasing and continuous on [y1,y2],[y_{1},y_{2}], u+(y1)=uminu_{+}(y_{1})=u_{\min} and u+(y)>uminu_{+}(y)>u_{\min} for y(y1,y2].y\in(y_{1},y_{2}]. By standard theory of ODE, cnψc=ψc,nC2(D1)\partial_{c}^{n}\psi_{c}=\psi_{c,n}\in C^{2}(D_{1}) and ψc,n|D1\psi_{c,n}|_{D_{1}} is real-valued for n𝐍n\in\mathbf{N}. Using this extension, ψc|c=umin\psi_{c}|_{c=u_{\min}} is well-defined and satisfies (4.20) for y(y1,y2].y\in(y_{1},y_{2}]. For fixed δ1(0,δ]\delta_{1}\in(0,\delta] and n𝐍n\in\mathbf{N}, we have

(4.26) |ψc,n(y)|+|yψc,n(y)|C(n,δ1)fory[y1+δ1,y2] and 0uminc1,\displaystyle|\psi_{c,n}(y)|+|\partial_{y}\psi_{c,n}(y)|\leq C(n,\delta_{1})\ \text{for}\ y\in[y_{1}+\delta_{1},y_{2}]\text{ and }0\leq u_{\min}-c\leq 1,

since a continuous function is bounded in a compact set. Let m0𝐙m_{0}\in\mathbf{Z} be such that m01<mm0m_{0}-1<m\leq m_{0} and γ=(m+1m0)/2\gamma=(m+1-m_{0})/2. Then m00m_{0}\geq 0 and γ(0,1/2].\gamma\in(0,1/2]. We claim that the following uniform bounds

(4.27) |ψc,n|C,|ψc,0|C|uc|,|ψc,1|C|uc|1γ,|yψc,m0+2|C|uc|2γ2\displaystyle|\psi_{c,n}|\leq C,\ |\psi_{c,0}|\leq C|u-c|,\ |\psi_{c,1}|\leq C|u-c|^{1-\gamma},\ |\partial_{y}\psi_{c,m_{0}+2}|\leq C|u-c|^{2\gamma-2}

hold for 0<uminc<1,y[y1,y2]0<u_{\min}-c<1,\ y\in[y_{1},y_{2}] and n𝐙[0,m0+1].n\in\mathbf{Z}\cap[0,m_{0}+1]. Assume that the uniform bounds (4.27) are true, which will be verified later. Let Wc,n=yψc,nψcyψcψc,n.W_{c,n}=\partial_{y}\psi_{c,n}\psi_{c}-\partial_{y}\psi_{c}\psi_{c,n}. Then we get by (4.25) that yWc,n=Fc,nψc.\partial_{y}W_{c,n}=F_{c,n}\psi_{c}. By (4.27) and using u(y1)=uminu(y_{1})=u_{\min}, we have for 0<uminc<10<u_{\min}-c<1,

|ψc(y1)|cumin|ψs,1(y1)|dsCcumin|umins|1γdsC|uminc|2γ,\displaystyle|\psi_{c}(y_{1})|\leq\int_{c}^{u_{\min}}|\psi_{s,1}(y_{1})|ds\leq C\int_{c}^{u_{\min}}|u_{\min}-s|^{1-\gamma}ds\leq C|u_{\min}-c|^{2-\gamma},

and thus

|yψc,m0+2ψc|(y1)=|yψc,m0+2(y1)||ψc(y1)|C|uminc|2γ2|uminc|2γ=C|uminc|γ,\displaystyle|\partial_{y}\psi_{c,m_{0}+2}\psi_{c}|(y_{1})=|\partial_{y}\psi_{c,m_{0}+2}(y_{1})||\psi_{c}(y_{1})|\leq C|u_{\min}-c|^{2\gamma-2}|u_{\min}-c|^{2-\gamma}=C|u_{\min}-c|^{\gamma},

which implies limcuminyψc,m0+2ψc(y1)=0.\lim\limits_{c\to u_{\min}^{-}}\partial_{y}\psi_{c,m_{0}+2}\psi_{c}(y_{1})=0. Since ψck,m0+2,m0+2(y1)=0\psi_{c_{k,m_{0}+2},m_{0}+2}(y_{1})=0 for k1k\geq 1 and ck,m0+2uminc_{k,m_{0}+2}\to u_{\min}^{-}, we have lim infcumin|yψcψc,m0+2|(y1)=0,\liminf\limits_{c\to u_{\min}^{-}}|\partial_{y}\psi_{c}\psi_{c,m_{0}+2}|(y_{1})=0, and thus lim infcumin|Wc,m0+2|(y1)=0.\liminf\limits_{c\to u_{\min}^{-}}|W_{c,m_{0}+2}|(y_{1})=0.

Since ψc(y2)=0,yψc(y2)=1\psi_{c}(y_{2})=0,\ \partial_{y}\psi_{c}(y_{2})=1 and recall that ψc,n=cnψc,\psi_{c,n}=\partial_{c}^{n}\psi_{c}, we have ψc,n(y2)=0,yψc,n\psi_{c,n}(y_{2})=0,\ \partial_{y}\psi_{c,n} (y2)=0(y_{2})=0 and Wc,n(y2)=0W_{c,n}(y_{2})=0 for n>0.n>0. Thus, Wc,n(y1)=y1y2yWc,n(y)dy=y1y2Fc,nψc(y)dy.-W_{c,n}(y_{1})=\int_{y_{1}}^{y_{2}}\partial_{y}W_{c,n}(y)dy=\int_{y_{1}}^{y_{2}}F_{c,n}\psi_{c}(y)dy. Note that

Fc,nn!uβ(uc)n+1ψc=k=1n1n!(nk)!uβ(uc)k+1ψc,nk.\displaystyle F_{c,n}-n!{u^{\prime\prime}-\beta\over(u-c)^{n+1}}\psi_{c}=\sum_{k=1}^{n-1}{n!\over(n-k)!}{u^{\prime\prime}-\beta\over(u-c)^{k+1}}\psi_{c,n-k}.

If n𝐙[2,m0+2]n\in\mathbf{Z}\cap[2,m_{0}+2], by (4.27) and (4.21), we have for 0<uminc<10<u_{\min}-c<1 and y[y1,y2]y\in[y_{1},y_{2}],

|Fc,nn!(uβ)ψc(uc)n+1|k=1n1n!(nk)!|uβ|(uc)k+1|ψc,nk|\displaystyle\left|F_{c,n}-{n!(u^{\prime\prime}-\beta)\psi_{c}\over(u-c)^{n+1}}\right|\leq\sum_{k=1}^{n-1}{n!\over(n-k)!}{|u^{\prime\prime}-\beta|\over(u-c)^{k+1}}|\psi_{c,n-k}|
\displaystyle\leq Ck=1n2|yy1|m(uc)k+1+C|yy1|m|uc|1γ(uc)nC|yy1|m|uc|n1+γ,\displaystyle C\sum_{k=1}^{n-2}{|y-y_{1}|^{m}\over(u-c)^{k+1}}+C{|y-y_{1}|^{m}|u-c|^{1-\gamma}\over(u-c)^{n}}\leq C{|y-y_{1}|^{m}\over|u-c|^{n-1+\gamma}},
|Fc,nψcn!(uβ)ψc2(uc)n+1|C|yy1|m|ψc||uc|n1+γC|yy1|m|uc||uc|n1+γ=C|yy1|m|uc|n2+γ,\displaystyle\left|F_{c,n}\psi_{c}-{n!(u^{\prime\prime}-\beta)\psi_{c}^{2}\over(u-c)^{n+1}}\right|\leq C{|y-y_{1}|^{m}|\psi_{c}|\over|u-c|^{n-1+\gamma}}\leq C{|y-y_{1}|^{m}|u-c|\over|u-c|^{n-1+\gamma}}=C{|y-y_{1}|^{m}\over|u-c|^{n-2+\gamma}},

and thus using mm0=2γ1m-m_{0}=2\gamma-1 and (4.23), we have

|Fc,m0+2ψc(m0+2)!(uβ)ψc2(uc)m0+3|C|yy1|m|uc|m0+γC|yy1|m|yy1|m0+γ=C|yy1|γ1.\displaystyle\left|F_{c,m_{0}+2}\psi_{c}-{(m_{0}+2)!(u^{\prime\prime}-\beta)\psi_{c}^{2}\over(u-c)^{m_{0}+3}}\right|\leq C{|y-y_{1}|^{m}\over|u-c|^{m_{0}+\gamma}}\leq C{|y-y_{1}|^{m}\over|y-y_{1}|^{m_{0}+\gamma}}=C|y-y_{1}|^{\gamma-1}.

Integrating it on [y1,y2][y_{1},y_{2}] and using Wc,n(y1)=y1y2Fc,nψc(y)dy-W_{c,n}(y_{1})=\int_{y_{1}}^{y_{2}}F_{c,n}\psi_{c}(y)dy, we have for 0<uminc<10<u_{\min}-c<1,

|y1y2(m0+2)!(uβ)ψc2(uc)m0+3dy|\displaystyle\left|\int_{y_{1}}^{y_{2}}{(m_{0}+2)!(u^{\prime\prime}-\beta)\psi_{c}^{2}\over(u-c)^{m_{0}+3}}dy\right| |y1y2Fc,m0+2ψcdy|+Cy1y2|yy1|γ1dy\displaystyle\leq\left|\int_{y_{1}}^{y_{2}}F_{c,m_{0}+2}\psi_{c}dy\right|+C\int_{y_{1}}^{y_{2}}|y-y_{1}|^{\gamma-1}dy
|Wc,m0+2|(y1)+C,\displaystyle\leq|W_{c,m_{0}+2}|(y_{1})+C,

and as lim infcumin|Wc,m0+2|(y1)=0\liminf\limits_{c\to u_{\min}^{-}}|W_{c,m_{0}+2}|(y_{1})=0, we have

lim infcumin|y1y2(m0+2)!(uβ)ψc2(uc)m0+3dy|C,\displaystyle\liminf\limits_{c\to u_{\min}^{-}}\left|\int_{y_{1}}^{y_{2}}{(m_{0}+2)!(u^{\prime\prime}-\beta)\psi_{c}^{2}\over(u-c)^{m_{0}+3}}dy\right|\leq C, lim infcumin|y1y2(uβ)ψc2(uc)m0+3dy|C.\displaystyle\liminf\limits_{c\to u_{\min}^{-}}\left|\int_{y_{1}}^{y_{2}}{(u^{\prime\prime}-\beta)\psi_{c}^{2}\over(u-c)^{m_{0}+3}}dy\right|\leq C.

Since uβu^{\prime\prime}-\beta is continuous and real-valued, and C1|yy1|m|u(y)β|C^{-1}|y-y_{1}|^{m}\leq|u^{\prime\prime}(y)-\beta|, it does not change sign on y[y1,y1+δ]y\in[y_{1},y_{1}+\delta]. Then for 0<uminc<10<u_{\min}-c<1,

|y1y1+δ(uβ)ψc2(uc)m0+3dy|=y1y1+δ|uβ|ψc2(uc)m0+3dyC1y1y1+δ|yy1|mψc2(uc)m0+3dy.\displaystyle\left|\int_{y_{1}}^{y_{1}+\delta}{(u^{\prime\prime}-\beta)\psi_{c}^{2}\over(u-c)^{m_{0}+3}}dy\right|=\int_{y_{1}}^{y_{1}+\delta}{|u^{\prime\prime}-\beta|\psi_{c}^{2}\over(u-c)^{m_{0}+3}}dy\geq C^{-1}\int_{y_{1}}^{y_{1}+\delta}{|y-y_{1}|^{m}\psi_{c}^{2}\over(u-c)^{m_{0}+3}}dy.

As |ψc|C,|uβ|C|\psi_{c}|\leq C,\ |u^{\prime\prime}-\beta|\leq C, ucuuminC1u-c\geq u-u_{\min}\geq C^{-1} for y[y1+δ,y2],y\in[y_{1}+\delta,y_{2}], and 0<uminc<10<u_{\min}-c<1, we have

|y1+δy2(uβ)ψc2(uc)m0+3dy|y1+δy2CdyC.\displaystyle\left|\int_{y_{1}+\delta}^{y_{2}}{(u^{\prime\prime}-\beta)\psi_{c}^{2}\over(u-c)^{m_{0}+3}}dy\right|\leq\int_{y_{1}+\delta}^{y_{2}}Cdy\leq C.

Thus,

lim infcuminy1y1+δ|yy1|mψc2(uc)m0+3dyClim infcumin|y1y1+δ(uβ)ψc2(uc)m0+3dy|\displaystyle\liminf\limits_{c\to u_{\min}^{-}}\int_{y_{1}}^{y_{1}+\delta}{|y-y_{1}|^{m}\psi_{c}^{2}\over(u-c)^{m_{0}+3}}dy\leq C\liminf\limits_{c\to u_{\min}^{-}}\left|\int_{y_{1}}^{y_{1}+\delta}{(u^{\prime\prime}-\beta)\psi_{c}^{2}\over(u-c)^{m_{0}+3}}dy\right|
\displaystyle\leq Clim infcumin|y1y2(uβ)ψc2(uc)m0+3dy|+Clim supcumin|y1+δy2(uβ)ψc2(uc)m0+3dy|C.\displaystyle C\liminf\limits_{c\to u_{\min}^{-}}\left|\int_{y_{1}}^{y_{2}}{(u^{\prime\prime}-\beta)\psi_{c}^{2}\over(u-c)^{m_{0}+3}}dy\right|+C\limsup\limits_{c\to u_{\min}^{-}}\left|\int_{y_{1}+\delta}^{y_{2}}{(u^{\prime\prime}-\beta)\psi_{c}^{2}\over(u-c)^{m_{0}+3}}dy\right|\leq C.

Since ψcC(D1)\psi_{c}\in C(D_{1}), we have for fixed y(y1,y1+δ]y\in(y_{1},y_{1}+\delta],

limcuminψc(y)=ψumin(y),\displaystyle\lim\limits_{c\to u_{\min}^{-}}\psi_{c}(y)=\psi_{u_{\min}}(y), limcumin|yy1|mψc2(uc)m0+3=|yy1|mψumin2(uumin)m0+3.\displaystyle\lim\limits_{c\to u_{\min}^{-}}{|y-y_{1}|^{m}\psi_{c}^{2}\over(u-c)^{m_{0}+3}}={|y-y_{1}|^{m}\psi_{u_{\min}}^{2}\over(u-u_{\min})^{m_{0}+3}}.

Thus, by Fatou’s Lemma, we have

y1y1+δ|yy1|mψumin2(uumin)m0+3dy=y1y1+δlimcumin|yy1|mψc2(uc)m0+3dylim infcuminy1y1+δ|yy1|mψc2(uc)m0+3dyC.\displaystyle\int_{y_{1}}^{y_{1}+\delta}{|y-y_{1}|^{m}\psi_{u_{\min}}^{2}\over(u-u_{\min})^{m_{0}+3}}dy=\int_{y_{1}}^{y_{1}+\delta}\!\!\lim\limits_{c\to u_{\min}^{-}}{|y-y_{1}|^{m}\psi_{c}^{2}\over(u-c)^{m_{0}+3}}dy\leq\liminf\limits_{c\to u_{\min}^{-}}\int_{y_{1}}^{y_{1}+\delta}{|y-y_{1}|^{m}\psi_{c}^{2}\over(u-c)^{m_{0}+3}}dy\leq C.

By (4.22), we have

|yy1|mψumin2(uumin)m0+3|yy1|mψumin2(C(yy1))m0+3C1ψumin2|yy1|m0m+3C1ψumin2|yy1|3\displaystyle{|y-y_{1}|^{m}\psi_{u_{\min}}^{2}\over(u-u_{\min})^{m_{0}+3}}\geq{|y-y_{1}|^{m}\psi_{u_{\min}}^{2}\over(C(y-y_{1}))^{m_{0}+3}}\geq{C^{-1}\psi_{u_{\min}}^{2}\over|y-y_{1}|^{m_{0}-m+3}}\geq{C^{-1}\psi_{u_{\min}}^{2}\over|y-y_{1}|^{3}}

for y(y1,y1+δ]y\in(y_{1},y_{1}+\delta], where we used mm0m\leq m_{0}. Thus,

y1y1+δψumin2|yy1|3dyCy1y1+δ|yy1|mψumin2(uumin)m0+3dyC.\displaystyle\int_{y_{1}}^{y_{1}+\delta}{\psi_{u_{\min}}^{2}\over|y-y_{1}|^{3}}dy\leq C\int_{y_{1}}^{y_{1}+\delta}{|y-y_{1}|^{m}\psi_{u_{\min}}^{2}\over(u-u_{\min})^{m_{0}+3}}dy\leq C.

Now we take φ=ψumin\varphi=\psi_{u_{\min}}. Then φ\varphi is real-valued and for y(y1,y2]y\in(y_{1},y_{2}], it satisfies

(4.28) φ+uβuuminφ=α2φ,φ(y2)=0,φ(y2)=1,y1y1+δφ2|yy1|3dyC.\displaystyle-\varphi^{\prime\prime}+{u^{\prime\prime}-\beta\over u-u_{\min}}\varphi=-\alpha^{2}\varphi,\ \varphi(y_{2})=0,\ \varphi^{\prime}(y_{2})=1,\ \int_{y_{1}}^{y_{1}+\delta}{\varphi^{2}\over|y-y_{1}|^{3}}dy\leq C.

Thus, φ,φ/|yy1|L2(y1,y1+δ).\varphi,\varphi/|y-y_{1}|\in L^{2}(y_{1},y_{1}+\delta). By (4.22), we have for y(y1,y1+δ]y\in(y_{1},y_{1}+\delta],

|uβuuminφ|C|φ|uuminC|φ||yy1|,uβuuminφL2(y1,y1+δ).\displaystyle\left|{u^{\prime\prime}-\beta\over u-u_{\min}}\varphi\right|\leq{C|\varphi|\over u-u_{\min}}\leq{C|\varphi|\over|y-y_{1}|},\quad{u^{\prime\prime}-\beta\over u-u_{\min}}\varphi\in L^{2}(y_{1},y_{1}+\delta).

Thus, φL2(y1,y1+δ),φH2(y1,y1+δ)\varphi^{\prime\prime}\in L^{2}(y_{1},y_{1}+\delta),\ \varphi\in H^{2}(y_{1},y_{1}+\delta) and φC1([y1,y1+δ])\varphi\in C^{1}([y_{1},y_{1}+\delta]) by defining φ(y1)=limyy1+φ(y)\varphi(y_{1})=\lim_{y\to y_{1}^{+}}\varphi(y). If φ(y1)0,\varphi(y_{1})\neq 0, then there exists δ1(0,δ]\delta_{1}\in(0,\delta] such that |φ(y)||φ(y1)|/2C1|yy1||\varphi(y)|\geq|\varphi(y_{1})|/2\geq C^{-1}|y-y_{1}| for y(y1,y1+δ1]y\in(y_{1},y_{1}+\delta_{1}]. If φ(y1)=0\varphi(y_{1})=0 and φ(y1)0,\varphi^{\prime}(y_{1})\neq 0, then there exists δ1(0,δ]\delta_{1}\in(0,\delta] such that |φ(y)||φ(y1)|/2|\varphi^{\prime}(y)|\geq|\varphi^{\prime}(y_{1})|/2 for y(y1,y1+δ1]y\in(y_{1},y_{1}+\delta_{1}], and |φ(y)|=|φ(y)φ(y1)|=|(yy1)φ(ξy)||yy1||φ(y1)|/2|\varphi(y)|=|\varphi(y)-\varphi(y_{1})|=|(y-y_{1})\varphi^{\prime}(\xi_{y})|\geq|y-y_{1}||\varphi^{\prime}(y_{1})|/2 for ξy(y1,y)\xi_{y}\in(y_{1},y) and y(y1,y1+δ1]y\in(y_{1},y_{1}+\delta_{1}]. Therefore, if φ(y1)0\varphi(y_{1})\neq 0 or φ(y1)0,\varphi^{\prime}(y_{1})\neq 0, then there exists δ1(0,δ]\delta_{1}\in(0,\delta] and C>0C>0 such that |φ(y)|C1|yy1||\varphi(y)|\geq C^{-1}|y-y_{1}| for y(y1,y1+δ1]y\in(y_{1},y_{1}+\delta_{1}], and

y1y1+δφ2|yy1|3dyC2y1y1+δ1|yy1|2|yy1|3dy=+,\displaystyle\int_{y_{1}}^{y_{1}+\delta}{\varphi^{2}\over|y-y_{1}|^{3}}dy\geq C^{-2}\int_{y_{1}}^{y_{1}+\delta_{1}}{|y-y_{1}|^{2}\over|y-y_{1}|^{3}}dy=+\infty,

which contradicts (4.28). Thus, we must have φ(y1)=φ(y1)=0.\varphi(y_{1})=\varphi^{\prime}(y_{1})=0. Then by the proof of Lemma 3 in [28], we have φ0\varphi\equiv 0 on [y1,y2][y_{1},y_{2}], which contradicts φ(y2)=1\varphi^{\prime}(y_{2})=1. This proves (1) for Case 2.

It remains to prove (4.27). Let δ1(0,δ]\delta_{1}\in(0,\delta] be fixed such that Lemma 4.4 is true. Recall that z=y1+δ1z=y_{1}+\delta_{1} and |f|L(z)=supy[y1,z]|f(y)|.|f|_{L^{\infty}(z)}=\sup_{y\in[y_{1},z]}|f(y)|. By (4.26) we know that (4.27) is true for y[z,y2]y\in[z,y_{2}]. Now we assume that y[y1,z],0<uminc<1,y\in[y_{1},z],0<u_{\min}-c<1, and that (4.3)–(4.6) are used for FF satisfying (4.14) (i.e. the condition in Lemma 4.4). The proof of (4.27) for y[y1,z]y\in[y_{1},z] is divided into 7 steps as follows.
Step 1. |ψc,n|C|\psi_{c,n}|\leq C for y[y1,z]y\in[y_{1},z] and n𝐙[0,m0].n\in\mathbf{Z}\cap[0,m_{0}].

For n=0,n=0, by (4.25), (4.3), (4.26) and Fc,0=0F_{c,0}=0, we have |ψc,0|L(z)C(|ψc,0(z)|+δ1γ|ψc,0(z)|)C|\psi_{c,0}|_{L^{\infty}(z)}\leq C(|\psi_{c,0}(z)|+\delta_{1}^{\gamma}|\psi_{c,0}^{\prime}(z)|)\leq C, and thus |ψc,0|C|\psi_{c,0}|\leq C for y[y1,z].y\in[y_{1},z]. Now, we prove the result by induction. Assume that n𝐙(0,m0]n\in\mathbf{Z}\cap(0,m_{0}] and |ψc,k|C|\psi_{c,k}|\leq C for k𝐙[0,n)k\in\mathbf{Z}\cap[0,n) and y[y1,z].y\in[y_{1},z]. Then by (4.25), (4.21), (4.23) and mm0=2γ1m-m_{0}=2\gamma-1, we have for y[y1,z]y\in[y_{1},z],

(4.29) |Fc,n|Ck=1n|yy1|m|ψc,nk|(uc)k+1Ck=1n|yy1|m(uc)k+1C(uc)m(uc)n+1,\displaystyle|F_{c,n}|\leq C\sum_{k=1}^{n}{|y-y_{1}|^{m}|\psi_{c,n-k}|\over(u-c)^{k+1}}\leq C\sum_{k=1}^{n}{|y-y_{1}|^{m}\over(u-c)^{k+1}}\leq C{(u-c)^{m}\over(u-c)^{n+1}},
(4.30) |(uc)2γFc,n|C(uc)2γ+mn1C(uc)mm0+1γ=C(uc)γC.\displaystyle|(u-c)^{2-\gamma}F_{c,n}|\leq C(u-c)^{2-\gamma+m-n-1}\leq C(u-c)^{m-m_{0}+1-\gamma}=C(u-c)^{\gamma}\leq C.

By (4.3), (4.26) and (4.30), we have

(4.31) |ψc,n|L(z)C(δ1γ|(uc)2γFc,n|L(z)+|ψc,n(z)|+δ1γ|ψc,n(z)|)C,\displaystyle|\psi_{c,n}|_{L^{\infty}(z)}\leq C(\delta_{1}^{\gamma}|(u-c)^{2-\gamma}F_{c,n}|_{L^{\infty}(z)}+|\psi_{c,n}(z)|+\delta_{1}^{\gamma}|\psi_{c,n}^{\prime}(z)|)\leq C,

which means |ψc,n|C|\psi_{c,n}|\leq C for y[y1,z].y\in[y_{1},z]. Thus, the result in Step 1 is true.
Step 2. |ψc,n|C|uc|γ1|\psi_{c,n}|\leq C|u-c|^{\gamma-1} for y[y1,z]y\in[y_{1},z] and n=m0+1.n=m_{0}+1.

Let n=m0+1.n=m_{0}+1. By Step 1, we know that |ψc,k|C|\psi_{c,k}|\leq C for y[y1,z]y\in[y_{1},z] and k𝐙[0,n)k\in\mathbf{Z}\cap[0,n). Thus, (4.29) is still true and for y[y1,z]y\in[y_{1},z],

|(uc)32γFc,n|C(uc)32γ+mn1=C(uc)mm0+12γ=C,\displaystyle|(u-c)^{3-2\gamma}F_{c,n}|\leq C(u-c)^{3-2\gamma+m-n-1}=C(u-c)^{m-m_{0}+1-2\gamma}=C,

which, along with (4.4) and (4.26), implies that

|(uc)1γψc,n|L(z)C(δ1γ|(uc)32γFc,n|L(z)+|ψc,n(z)|+δ1γ|ψc,n(z)|)C.\displaystyle|(u-c)^{1-\gamma}\psi_{c,n}|_{L^{\infty}(z)}\leq C(\delta_{1}^{\gamma}|(u-c)^{3-2\gamma}F_{c,n}|_{L^{\infty}(z)}+|\psi_{c,n}(z)|+\delta_{1}^{\gamma}|\psi_{c,n}^{\prime}(z)|)\leq C.

Then |(uc)1γψc,n|C|(u-c)^{1-\gamma}\psi_{c,n}|\leq C, and thus |ψc,n|C|uc|γ1|\psi_{c,n}|\leq C|u-c|^{\gamma-1} for y[y1,z].y\in[y_{1},z].
Step 3. |ψc,0|C|uc|1γ|\psi_{c,0}|\leq C|u-c|^{1-\gamma} for y[y1,z]y\in[y_{1},z].

If m0=0,m_{0}=0, then m=0m=0 and γ=1/2.\gamma=1/2. By Step 2, we have |ψc,1|C|uc|γ1=C|uc|γ|\psi_{c,1}|\leq C|u-c|^{\gamma-1}=C|u-c|^{-\gamma} for y[y1,z]y\in[y_{1},z]. If m0>0,m_{0}>0, then by Step 1, we have |ψc,1|CC|uc|γ|\psi_{c,1}|\leq C\leq C|u-c|^{-\gamma} for y[y1,z]y\in[y_{1},z]. Thus, |ψc,1|C|uc|γ|\psi_{c,1}|\leq C|u-c|^{-\gamma} is always true for y[y1,z]y\in[y_{1},z]. Then by (4.24) and u(y1)=uminu(y_{1})=u_{\min}, we have |ψc,0(y1)|cumin|ψs,1(y1)|dsCcumin|umins|γdsC|uminc|1γ|\psi_{c,0}(y_{1})|\leq\int_{c}^{u_{\min}}|\psi_{s,1}(y_{1})|ds\leq C\int_{c}^{u_{\min}}|u_{\min}-s|^{-\gamma}ds\leq C|u_{\min}-c|^{1-\gamma}. By (4.5), (4.26) and Fc,0=0F_{c,0}=0, we have |(uc)γ1ψc,0|L(z)C(|(uminc)γ1ψc,0(y1)|+|ψc,0(z)|+|ψc,0(z)|)C|(u-c)^{\gamma-1}\psi_{c,0}|_{L^{\infty}(z)}\leq C(|(u_{\min}-c)^{\gamma-1}\psi_{c,0}(y_{1})|+|\psi_{c,0}(z)|+|\psi_{c,0}^{\prime}(z)|)\leq C. Then |(uc)γ1ψc,0|C|(u-c)^{\gamma-1}\psi_{c,0}|\leq C, and thus |ψc,0|C|uc|1γ|\psi_{c,0}|\leq C|u-c|^{1-\gamma} for y[y1,z].y\in[y_{1},z].
Step 4. |ψc,n|C|\psi_{c,n}|\leq C for y[y1,z]y\in[y_{1},z] and n=m0+1.n=m_{0}+1.

Let n=m0+1.n=m_{0}+1. By (4.25), (4.21), (4.22), Step 1 and Step 3, we have for y[y1,z]y\in[y_{1},z],

|Fc,n|Ck=1n|yy1|m|ψc,nk|(uc)k+1Ck=1n1|yy1|m(uc)k+1+C|yy1|m|ψc,0|(uc)n+1\displaystyle|F_{c,n}|\leq C\sum_{k=1}^{n}{|y-y_{1}|^{m}|\psi_{c,n-k}|\over(u-c)^{k+1}}\leq C\sum_{k=1}^{n-1}{|y-y_{1}|^{m}\over(u-c)^{k+1}}+C{|y-y_{1}|^{m}|\psi_{c,0}|\over(u-c)^{n+1}}
\displaystyle\leq C(uc)m(uc)n+C|yy1|m|uc|1γ(uc)n+1C(uc)mγn,\displaystyle C{(u-c)^{m}\over(u-c)^{n}}+C{|y-y_{1}|^{m}|u-c|^{1-\gamma}\over(u-c)^{n+1}}\leq C(u-c)^{m-\gamma-n},
|(uc)2γFc,n|C(uc)22γ+mn=C(uc)mm0+12γ=C.\displaystyle|(u-c)^{2-\gamma}F_{c,n}|\leq C(u-c)^{2-2\gamma+m-n}=C(u-c)^{m-m_{0}+1-2\gamma}=C.

Here, we used n=m0+1n=m_{0}+1 and mm0=2γ1m-m_{0}=2\gamma-1. Thus, (4.31) is still true for n=m0+1,n=m_{0}+1, i.e. |ψc,n|C|\psi_{c,n}|\leq C for y[y1,z]y\in[y_{1},z].
Step 5. |ψc,0|C|uc||\psi_{c,0}|\leq C|u-c| for y[y1,z]y\in[y_{1},z].

By Step 1 and Step 4, we have |ψc,1|C|\psi_{c,1}|\leq C for y[y1,z]y\in[y_{1},z]. Then by (4.24), we have |ψc,0(y1)|cumin|ψs,1(y1)|dsCcumindsC|uminc||\psi_{c,0}(y_{1})|\leq\int_{c}^{u_{\min}}|\psi_{s,1}(y_{1})|ds\leq C\int_{c}^{u_{\min}}ds\leq C|u_{\min}-c|. By (4.6), (4.26) and Fc,0=0F_{c,0}=0, we have |(uc)1ψc,0|L(z)C(|(uminc)1ψc,0(y1)|+|ψc,0(z)|)C|(u-c)^{-1}\psi_{c,0}|_{L^{\infty}(z)}\leq C(|(u_{\min}-c)^{-1}\psi_{c,0}(y_{1})|+|\psi_{c,0}^{\prime}(z)|)\leq C, which gives |(uc)1ψc,0|C|(u-c)^{-1}\psi_{c,0}|\leq C and |ψc,0|C|uc||\psi_{c,0}|\leq C|u-c| for y[y1,z].y\in[y_{1},z].
Step 6. |yψc,n|C|uc|2γ2|\partial_{y}\psi_{c,n}|\leq C|u-c|^{2\gamma-2} and |ψc,n|C|uc|γ1|\psi_{c,n}|\leq C|u-c|^{\gamma-1} for y[y1,z]y\in[y_{1},z] and n=m0+2.n=m_{0}+2.

Since n=m0+2n=m_{0}+2 and mm0=2γ1m-m_{0}=2\gamma-1, we have by (4.25), (4.21), (4.23), Step 1 and Steps 4-5 that for y[y1,z]y\in[y_{1},z],

|Fc,n|Ck=1n|yy1|m|ψc,nk|(uc)k+1Ck=1n1|yy1|m(uc)k+1+C|yy1|m|ψc,0|(uc)n+1\displaystyle|F_{c,n}|\leq C\sum_{k=1}^{n}{|y-y_{1}|^{m}|\psi_{c,n-k}|\over(u-c)^{k+1}}\leq C\sum_{k=1}^{n-1}{|y-y_{1}|^{m}\over(u-c)^{k+1}}+C{|y-y_{1}|^{m}|\psi_{c,0}|\over(u-c)^{n+1}}
\displaystyle\leq C(uc)m(uc)n+C|yy1|m|uc|(uc)n+1C(uc)mn=C(uc)mm02,\displaystyle C{(u-c)^{m}\over(u-c)^{n}}+C{|y-y_{1}|^{m}|u-c|\over(u-c)^{n+1}}\leq C(u-c)^{m-n}=C(u-c)^{m-m_{0}-2},
|(uc)32γFc,n|C(uc)12γ+mm0=C.\displaystyle|(u-c)^{3-2\gamma}F_{c,n}|\leq C(u-c)^{1-2\gamma+m-m_{0}}=C.

Then by (4.4) and (4.26), we have

|(uc)1γψc,n|L(z)+δ1γ|(uc)22γyψc,n|L(z)\displaystyle|(u-c)^{1-\gamma}\psi_{c,n}|_{L^{\infty}(z)}+\delta_{1}^{\gamma}|(u-c)^{2-2\gamma}\partial_{y}\psi_{c,n}|_{L^{\infty}(z)}
\displaystyle\leq C(|(uc)32γFc,n|L(z)+|ψc,n(z)|+|yψc,n(z)|)C.\displaystyle C(|(u-c)^{3-2\gamma}F_{c,n}|_{L^{\infty}(z)}+|\psi_{c,n}(z)|+|\partial_{y}\psi_{c,n}(z)|)\leq C.

Therefore, |yψc,n|Cδ1γ|uc|2γ2C|uc|2γ2|\partial_{y}\psi_{c,n}|\leq C\delta_{1}^{-\gamma}|u-c|^{2\gamma-2}\leq C|u-c|^{2\gamma-2} and |ψc,n|C|uc|γ1|\psi_{c,n}|\leq C|u-c|^{\gamma-1} for y[y1,z].y\in[y_{1},z].
Step 7. |ψc,1|C|uc|1γ|\psi_{c,1}|\leq C|u-c|^{1-\gamma} for y[y1,z]y\in[y_{1},z].

If m0=0,m_{0}=0, then m=0m=0 and γ=1/2.\gamma=1/2. By Step 6, we have |ψc,2|C|uc|γ1=C|uc|γ|\psi_{c,2}|\leq C|u-c|^{\gamma-1}=C|u-c|^{-\gamma} for y[y1,z]y\in[y_{1},z]. If m0>0,m_{0}>0, then by Step 1 and Step 4, we have |ψc,2|CC|uc|γ|\psi_{c,2}|\leq C\leq C|u-c|^{-\gamma} for y[y1,z]y\in[y_{1},z]. Thus, |ψc,2|C|uc|γ|\psi_{c,2}|\leq C|u-c|^{-\gamma} is always true for y[y1,z]y\in[y_{1},z]. Then by (4.24) and u(y1)=uminu(y_{1})=u_{\min}, we have |ψc,1(y1)|cumin|ψs,2(y1)|dsCcumin|umins|γdsC|uminc|1γ|\psi_{c,1}(y_{1})|\leq\int_{c}^{u_{\min}}|\psi_{s,2}(y_{1})|ds\leq C\int_{c}^{u_{\min}}|u_{\min}-s|^{-\gamma}ds\leq C|u_{\min}-c|^{1-\gamma}. By (4.25), (4.21) and Step 5, we have for y[y1,z]y\in[y_{1},z],

|Fc,1|C|yy1|m|ψc,0|(uc)2C|yy1|m|uc|(uc)2=C|yy1|muc,\displaystyle|F_{c,1}|\leq C{|y-y_{1}|^{m}|\psi_{c,0}|\over(u-c)^{2}}\leq C{|y-y_{1}|^{m}|u-c|\over(u-c)^{2}}=C{|y-y_{1}|^{m}\over u-c},
|(uc)γ+1Fc,1|C(uc)γ|yy1|mC.\displaystyle|(u-c)^{\gamma+1}F_{c,1}|\leq C(u-c)^{\gamma}|y-y_{1}|^{m}\leq C.

Then by (4.5) and (4.26), we have

|(uc)γ1ψc,1|L(z)\displaystyle|(u-c)^{\gamma-1}\psi_{c,1}|_{L^{\infty}(z)}
\displaystyle\leq C(|(uc)γ+1Fc,1|L(z)+|(uminc)γ1ψc,1(y1)|+|ψc,1(z)|+|yψc,1(z)|)C,\displaystyle C(|(u-c)^{\gamma+1}F_{c,1}|_{L^{\infty}(z)}+|(u_{\min}-c)^{\gamma-1}\psi_{c,1}(y_{1})|+|\psi_{c,1}(z)|+|\partial_{y}\psi_{c,1}(z)|)\leq C,

which gives |(uc)γ1ψc,1|C|(u-c)^{\gamma-1}\psi_{c,1}|\leq C and |ψc,1|C|uc|1γ|\psi_{c,1}|\leq C|u-c|^{1-\gamma} for y[y1,z].y\in[y_{1},z].

By Step 1 and Steps 4-7 we know that (4.27) is true for y[y1,z]=[y1,y1+δ1]y\in[y_{1},z]=[y_{1},y_{1}+\delta_{1}]. This completes the proof of (4.27) and thus Case 2. ∎

4.3. Rule out oscillation for flows in class 𝒦+\mathcal{K}^{+}

We rule out the oscillation of λn(c)\lambda_{n}(c) for flows in class 𝒦+\mathcal{K}^{+}, which is stated in Theorem 2.13. The proof is based on Hamiltonian structure and index theory.

Proof of Theorem 2.13.

The assumption (H1) is satisfied for a flow uu in class 𝒦+\mathcal{K}^{+}. By Theorem 2.11, it suffices to prove (σd(α,β)(,umin))<\sharp(\sigma_{d}(\mathcal{R}_{\alpha,\beta})\cap(-\infty,u_{\min}))<\infty for 0<α2Mβ0<\alpha^{2}\leq M_{\beta} and 0<β98κ+0<\beta\leq{9\over 8}\kappa_{+}. Similar proof is valid for 98κβ<0{9\over 8}\kappa_{-}\leq\beta<0. First, we consider βRan(u)(0,98κ+]\beta\in\text{Ran}(u)\cap(0,{9\over 8}\kappa_{+}]. Define the non-shear space

X:={ωL2(DT):0Tω(x,y)dx=0,T-periodic in x}.X:=\{\omega\in L^{2}(D_{T}):\int_{0}^{T}\omega(x,y)dx=0,\text{T-periodic in }x\}.

Note that as ω=xv2yv1\omega=\partial_{x}v_{2}-\partial_{y}v_{1}, 0Tω(x,y)dx=0\int_{0}^{T}\omega(x,y)dx=0 is equivalent to 0Tv1(x,y)dx=\int_{0}^{T}v_{1}(x,y)dx=constant. Thus, 0Tv1(x,y)dx=0\int_{0}^{T}v_{1}(x,y)dx=0 implies 0Tω(x,y)dx=0\int_{0}^{T}\omega(x,y)dx=0.

The linearized equation (1.4) has a Hamiltonian structure in the traveling frame (xuβt,y,t)(x-u_{\beta}t,y,t):

ωt=(βu)x(ω/Kβψ)=JLω,\displaystyle\omega_{t}=-(\beta-u^{\prime\prime})\partial_{x}\left({\omega}/{K_{\beta}}-\psi\right)=JL\omega,

where J=(βu)x:XX,L=1/Kβ(Δ)1:XX.J=-(\beta-u^{\prime\prime})\partial_{x}:X^{*}\rightarrow X,L={1}/{K_{\beta}}-\left(-\Delta\right)^{-1}:X\rightarrow X^{*}. Let Jα=iα(βu)J_{\alpha}=-i\alpha(\beta-u^{\prime\prime}) and Lα=1Kβ(d2dy2+α2)1L_{\alpha}=\frac{1}{K_{\beta}}-(-\frac{d^{2}}{dy^{2}}+\alpha^{2})^{-1} on L1Kβ2L_{\frac{1}{K_{\beta}}}^{2}. It follows from Theorem 3 in [28] that

kc+kr+ki0=n(Lα),k_{c}+k_{r}+k_{i}^{\leq 0}=n^{-}(L_{\alpha}),

where n(Lα)n^{-}(L_{\alpha}) is the Morse index of LαL_{\alpha}, kr{k}_{r} is the sum of algebraic multiplicities of positive eigenvalues of JαLα{J}_{\alpha}{L}_{\alpha}, kc{k}_{c} is the sum of algebraic multiplicities of eigenvalues of JαLα{J}_{\alpha}{L}_{\alpha} in the first and the fourth quadrants and ki0{k}_{i}^{\leq 0} is the total number of non-positive dimensions of Lα,\langle{L}_{\alpha}\cdot,\cdot\rangle restricted to the generalized eigenspaces of nonzero purely imaginary eigenvalues of JαLα{J}_{\alpha}{L}_{\alpha}.

Suppose that (σd(α,β)(,umin))=\sharp(\sigma_{d}(\mathcal{R}_{\alpha,\beta})\cap(-\infty,u_{\min}))=\infty. Then it follows from Theorem 2.9 that there exists mβ<n<Nβm_{\beta}<n<N_{\beta} such that ({λn(c)=α2,c<umin})=\sharp(\{\lambda_{n}(c)=-\alpha^{2},c<u_{\min}\})=\infty. Let c<uminc^{*}<u_{\min} be a solution of λn(c)=α2\lambda_{n}(c)=-\alpha^{2} with eigenfunction ϕ\phi^{*}. cc^{*} can be chosen sufficiently close to uminu_{\min}. Then iα(cuβ)-i\alpha(c^{*}-u_{\beta}) is a purely imaginary eigenvalue of JαLαJ_{\alpha}L_{\alpha} with eigenfunction ω=ϕ+α2ϕ\omega^{*}=-{\phi^{*}}^{\prime\prime}+\alpha^{2}\phi^{*}. By Theorem 4 in [28],

Lαω,ω=(cuβ)λn(c).\langle L_{\alpha}\omega^{*},\omega^{*}\rangle=-(c^{*}-u_{\beta})\lambda^{\prime}_{n}(c^{*}).

Note that cuβc-u_{\beta} does not change sign when c<uminc<u_{\min} is sufficiently close to uminu_{\min}. Then

({(cuβ)λn(c)0,c<umin}{λn(c)=α2,c<umin})=.\displaystyle\sharp(\{-(c-u_{\beta})\lambda^{\prime}_{n}(c)\leq 0,c<u_{\min}\}\cap\{\lambda_{n}(c)=-\alpha^{2},c<u_{\min}\})=\infty.

Hence, ki0=k_{i}^{\leq 0}=\infty. This contradicts that

ki0n(Lα)=n(L~0+α2)n(L~0)<,\displaystyle k_{i}^{\leq 0}\leq n^{-}(L_{\alpha})=n^{-}(\tilde{L}_{0}+\alpha^{2})\leq n^{-}(\tilde{L}_{0})<\infty,

where L~0=d2dy2Kβ:H2H01L2\tilde{L}_{0}=-{d^{2}\over dy^{2}}-K_{\beta}:H^{2}\cap H_{0}^{1}\to L^{2}. Therefore, (σd(α,β)(,umin))<\sharp(\sigma_{d}(\mathcal{R}_{\alpha,\beta})\cap(-\infty,u_{\min}))<\infty.

Then, we consider β(0,98κ+]Ran(u)\beta\in(0,{9\over 8}\kappa_{+}]\setminus\text{Ran}(u^{\prime\prime}). By Corollary 1 in [28], λn(c)\lambda_{n}(c) is decreasing on c(,umin)c\in(-\infty,u_{\min}) for any fixed n1n\geq 1. By Theorem 2.9, (σd(α,β)(,umin))<Nβ\sharp(\sigma_{d}(\mathcal{R}_{\alpha,\beta})\cap(-\infty,u_{\min}))<N_{\beta}. ∎

5. Relations between a traveling wave family and an isolated real eigenvalue

In this section, we establish the correspondence between a traveling wave family near a shear flow and an isolated real eigenvalue of kα,β\mathcal{R}_{k\alpha,\beta}. For a given isolated real eigenvalue c0c_{0}, we prove that there exists a set of traveling wave solutions near (u,0)(u,0) with traveling speeds converging to c0c_{0}, which is stated precisely in Lemma 2.5. We assume k,k0𝐙k,k_{0}\in\mathbf{Z}.

Proof of Lemma 2.5.

We assume that β>0\beta>0, and the case for β<0\beta<0 is similar. Since c0σd(k0α,β)𝐑c_{0}\in\sigma_{d}(\mathcal{R}_{k_{0}\alpha,\beta})\cap\mathbf{R} for some k01k_{0}\geq 1, we have c0<uminc_{0}<u_{\min} and we choose δ0>0\delta_{0}>0 such that c0+δ0<uminc_{0}+\delta_{0}<u_{\min}. By (1.3), u(xct,y)\vec{u}\left(x-ct,y\right) is a solution of (1.1)–(1.2) if and only if (ψ,c)(\psi,c) solves

(5.1) (ω+βy,ψcy)(x,y)=0\displaystyle\frac{\partial\left(\omega+\beta y,\psi-cy\right)}{\partial\left(x,y\right)}=0

and ψ\psi takes constant values on {y=yi}\left\{y=y_{i}\right\}, where i=1,2i=1,2, ω=curlu\omega=\operatorname{curl}\vec{u} and u=(yψ,xψ)\vec{u}=(\partial_{y}\psi,-\partial_{x}\psi). Let ψ0\psi_{0} be a stream function associated with the shear flow (u,0)\left(u,0\right), i.e., ψ0=u\psi_{0}^{\prime}=u. Since uc>0u-c>0 for c[c0δ0,c0+δ0]c\in[c_{0}-\delta_{0},c_{0}+\delta_{0}], ψ0cy\psi_{0}-cy is increasing on [y1,y2][y_{1},y_{2}]. Let Ic={ψ0(y)cy:y[y1,y2]}I_{c}=\{\psi_{0}\left(y\right)-cy:y\in[y_{1},y_{2}]\} for c[c0δ0,c0+δ0]c\in[c_{0}-\delta_{0},c_{0}+\delta_{0}], and then we can define a function f~cC2(Ic)\tilde{f}_{c}\in C^{2}(I_{c}) such that

(5.2) f~c(ψ0(y)cy)=ω0(y)+βy=ψ0(y)+βy.\tilde{f}_{c}\left(\psi_{0}\left(y\right)-cy\right)=\omega_{0}\left(y\right)+\beta y=-\psi_{0}^{\prime\prime}\left(y\right)+\beta y.

Moreover,

f~c(ψ0(y)cy)=βu(y)u(y)c=:𝒦c(y)\tilde{f}^{\prime}_{c}\left(\psi_{0}\left(y\right)-cy\right)=\frac{\beta-u^{\prime\prime}\left(y\right)}{u\left(y\right)-c}=:\mathcal{K}_{c}\left(y\right)

for c[c0δ0,c0+δ0]c\in[c_{0}-\delta_{0},c_{0}+\delta_{0}]. We extend f~c\tilde{f}_{c} to fcC02(𝐑)f_{c}\in C_{0}^{2}\left(\mathbf{R}\right) such that fc=f~cf_{c}=\tilde{f}_{c} on IcI_{c} and z2cfc(z)\partial_{z}^{2}\partial_{c}f_{c}(z) is continuous for c[c0δ0,c0+δ0]c\in[c_{0}-\delta_{0},c_{0}+\delta_{0}] and z𝐑z\in\mathbf{R}. Taking cc as the bifurcation parameter, we now construct steady solutions u=(yψ,xψ)\vec{u}=\left(\partial_{y}\psi,-\partial_{x}\psi\right) near (u,0)\left(u,0\right) by solving the elliptic equations

(5.3) Δψ+βy=fc(ψcy)\displaystyle-\Delta\psi+\beta y=f_{c}\left(\psi-cy\right)

with the boundary conditions that ψ{\psi} takes constant values on {y=yi}\left\{y=y_{i}\right\}, i=1,2i=1,2. Define the perturbation of the stream function by

ϕ(x,y)=ψ(x,y)ψ0(y).\phi\left(x,y\right)={\psi}\left(x,y\right)-\psi_{0}\left(y\right).

Then by (5.2)–(5.3), we have

Δϕ(fc(ϕ+ψ0cy)fc(ψ0cy))=0.-\Delta\phi-\left(f_{c}(\phi+\psi_{0}-cy)-f_{c}\left(\psi_{0}-cy\right)\right)=0.

Define the spaces

B={φH4(DT): φ(x,yi)=0,i=1,2,φ is even and T-periodic in x}\displaystyle B=\{\varphi\in H^{4}(D_{T}):\text{ }\varphi(x,y_{i})=0,\,i=1,2,\;\varphi\text{ is even and }T\text{-periodic in }x\}

and

C={φH2(DT): T-periodic in x},C=\left\{\varphi\in H^{2}(D_{T}):\text{ }T\text{-periodic in }x\right\},

where T=2π/αT={2\pi/\alpha}. Consider the mapping

F:B×[c0δ0,c0+δ0]C,\displaystyle F:B\times[c_{0}-\delta_{0},c_{0}+\delta_{0}]\longrightarrow C,
(ϕ,c)Δϕ(fc(ϕ+ψ0cy)fc(ψ0cy)).\displaystyle\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;(\phi,c)\longmapsto-\Delta\phi-\left(f_{c}(\phi+\psi_{0}-cy)-f_{c}\left(\psi_{0}-cy\right)\right).

Then F(0,c)=0F(0,c)=0 for c[c0δ0,c0+δ0]c\in[c_{0}-\delta_{0},c_{0}+\delta_{0}]. We study the bifurcation near the trivial solution (0,c0)(0,c_{0}) of the equation F(ϕ,c)=0F(\phi,c)=0 in BB, whose solutions give steady flows of (5.1).

For fixed c[c0δ0,c0+δ0]c\in[c_{0}-\delta_{0},c_{0}+\delta_{0}], by linearizing FF around ϕ=0\phi=0, we have

ϕF(0,c)=Δfc(ψ0cy)=Δ𝒦c=𝒢c|B,\partial_{\phi}F(0,c)=-\Delta-f_{c}^{\prime}(\psi_{0}-cy)=-\Delta-\mathcal{K}_{c}=\mathcal{G}_{c}|_{B},

where 𝒢c|B\mathcal{G}_{c}|_{B} is the restriction of 𝒢c\mathcal{G}_{c} in BB and 𝒢c\mathcal{G}_{c} is defined in (2.3). Then we divide the discussion of bifurcation near (0,c0)(0,c_{0}) of the equation F(ϕ,c)=0F(\phi,c)=0 into three cases. Since c0σd(k0α,β)𝐑c_{0}\in\sigma_{d}(\mathcal{R}_{k_{0}\alpha,\beta})\cap\mathbf{R}, there exists n01n_{0}\geq 1 such that (k0α)2=λn0(c0)(k_{0}\alpha)^{2}=-\lambda_{n_{0}}(c_{0}), where λn0(c0)\lambda_{n_{0}}(c_{0}) is the n0n_{0}-th eigenvalue of c0\mathcal{L}_{c_{0}} and c0\mathcal{L}_{c_{0}} is defined in (2.6). Let

(5.4) k=\displaystyle k_{*}= maxk1{k:there exists nk1 such that (kα)2=λnk(c0)}.\displaystyle\max\limits_{k\geq 1}\{k:\text{there exists }n_{k}\geq 1\text{ such that }-(k\alpha)^{2}=\lambda_{n_{k}}(c_{0})\}.

Then kk_{*} exists by our assumption and 1k0k<1\leq k_{0}\leq k_{*}<\infty. Now we denote n=nk.n_{*}=n_{k_{*}}.
Case 1. λn(c0)0\lambda_{n_{*}}^{\prime}(c_{0})\neq 0 (the transversal crossing condition) and c0σd(0,β)𝐑c_{0}\notin\sigma_{d}(\mathcal{R}_{0,\beta})\cap\mathbf{R}.

In this case, we have 0σ(c0)0\notin\sigma(\mathcal{L}_{c_{0}}). Let B={φB:2πkα-periodic in x}B_{*}=\{\varphi\in B:{2\pi\over k_{*}\alpha}\text{-periodic in }x\} and C={φC:2πkα-periodic in x}C_{*}=\{\varphi\in C:{2\pi\over k_{*}\alpha}\text{-periodic in }x\}. Consider the restriction F|BF|_{B_{*}} and 𝒢c|B\mathcal{G}_{c}|_{B_{*}}. Then by the definition of kk_{*}, we have

(5.5) ker(𝒢c0|B)=span{ϕc0,n(y)cos(kαx)}anddim(ker(𝒢c0|B))=1,\displaystyle\ker({\mathcal{G}_{c_{0}}|_{B_{*}}})={\rm{span}}\{\phi_{c_{0},n_{*}}(y)\cos(k_{*}\alpha x)\}\quad\text{and}\quad\dim(\ker(\mathcal{G}_{c_{0}}|_{B_{*}}))=1,

where ϕc0,n\phi_{c_{0},n_{*}} is a real-valued eigenfunction of λn(c0)σ(c0)\lambda_{n_{*}}(c_{0})\in\sigma(\mathcal{L}_{c_{0}}). Note that

cϕF(0,c0)(ϕc0,n(y)cos(kαx))=βu(uc0)2ϕc0,n(y)cos(kαx).\displaystyle\partial_{c}\partial_{\phi}F(0,c_{0})\left(\phi_{c_{0},n_{*}}(y)\cos(k_{*}\alpha x)\right)=-{\beta-u^{\prime\prime}\over(u-c_{0})^{2}}\phi_{c_{0},n_{*}}(y)\cos(k_{*}\alpha x).

Then by Lemma 1111 in [28], we have

0Ty1y2ϕc0,n(y)cos(kαx)[cϕF(0,c0)(ϕc0,n(y)cos(kαx))]dydx\displaystyle\int_{0}^{T}\int_{y_{1}}^{y_{2}}\phi_{c_{0},n_{*}}(y)\cos(k_{*}\alpha x)\left[\partial_{c}\partial_{\phi}F(0,c_{0})\left(\phi_{c_{0},n_{*}}(y)\cos(k_{*}\alpha x)\right)\right]dydx
=\displaystyle= 0Ty1y2βu(uc0)2|ϕc0,n(y)|2cos2(kαx)dydx=παλn(c0)0,\displaystyle-\int_{0}^{T}\int_{y_{1}}^{y_{2}}{\beta-u^{\prime\prime}\over(u-c_{0})^{2}}|\phi_{c_{0},n_{*}}(y)|^{2}\cos^{2}(k_{*}\alpha x)dydx={\pi\over\alpha}\lambda_{n_{*}}^{\prime}(c_{0})\neq 0,

where we used that ϕc0,n\phi_{c_{0},n_{*}} is real-valued. By (5.5), we have ϕc0,n(y)cos(kαx)ker(𝒢c0|B)\phi_{c_{0},n_{*}}(y)\cos(k_{*}\alpha x)\in\ker({\mathcal{G}_{c_{0}}|_{B_{*}}}) and thus, cϕF(0,c0)(ϕc0,n(y)cos(kαx))\partial_{c}\partial_{\phi}F(0,c_{0})\left(\phi_{c_{0},n_{*}}(y)\cos(k_{*}\alpha x)\right)\notin Ran(𝒢c0|B){\rm Ran}\,({\mathcal{G}_{c_{0}}|_{B_{*}}}). Then by Theorem 1.71.7 in [7], there exist δ>0\delta>0 and a nontrivial C1C^{1} bifurcating curve {(ϕγ,c(γ)),γ(δ,δ)}\{\left(\phi_{\gamma},c(\gamma)\right),\gamma\in(-\delta,\delta)\} of F(ϕ,c)=0F(\phi,c)=0, which intersects the trivial curve (0,c)\left(0,c\right) at c=c0c=c_{0}, such that

ϕγ(x,y)=γϕc0,n(y)cos(kαx)+o(|γ|).\displaystyle\phi_{\gamma}(x,y)=\gamma\phi_{c_{0},n_{*}}(y)\cos(k_{*}\alpha x)+o(|\gamma|).

So the stream functions take the form

ψγ(x,y)=ψ0(y)+ϕγ(x,y)=ψ0(y)+γϕc0,n(y)cos(kαx)+o(|γ|).\displaystyle\psi_{\gamma}(x,y)=\psi_{0}(y)+\phi_{\gamma}(x,y)=\psi_{0}(y)+\gamma\phi_{c_{0},n_{*}}(y)\cos(k_{*}\alpha x)+o(|\gamma|).

Let the velocity uγ=(u(γ),v(γ))=(yψγ,xψγ).\vec{u}_{\gamma}=(u_{(\gamma)},v_{(\gamma)})=\left(\partial_{y}\psi_{\gamma},-\partial_{x}\psi_{\gamma}\right). Since c0<uminc_{0}<u_{\min}, we have

(5.6) u(γ)(x,y)c(γ)=yψγ(x,y)c(γ)\displaystyle u_{(\gamma)}(x,y)-c{(\gamma)}=\partial_{y}\psi_{\gamma}(x,y)-c{(\gamma)}
=\displaystyle= u(y)c(γ)+γϕc0,n(y)cos(kαx)+o(|γ|)>0,\displaystyle u(y)-c{(\gamma)}+\gamma\phi^{\prime}_{c_{0},n_{*}}(y)\cos(k_{*}\alpha x)+o(|\gamma|)>0,

and

(5.7) v(γ)(x,y)=xψγ(x,y)=kαγϕc0,n(y)sin(kαx)+o(|γ|)0\displaystyle v_{(\gamma)}(x,y)=-\partial_{x}\psi_{\gamma}(x,y)=k_{*}\alpha\gamma\phi_{c_{0},n_{*}}(y)\sin(k_{*}\alpha x)+o(|\gamma|)\neq 0

when γ\gamma is small. Moreover, (u(γ),v(γ))(u,0)H3(DT)+|c(γ)c0|C0γ\|(u_{(\gamma)},v_{(\gamma)})-(u,0)\|_{H^{3}(D_{T})}+|c({\gamma})-c_{0}|\leq C_{0}\gamma for some constant C0>0C_{0}>0 large enough. Thus, we can take δ>0\delta>0 smaller and ε0=C0δ\varepsilon_{0}=C_{0}\delta such that for ε(0,ε0)\varepsilon\in(0,\varepsilon_{0}), (uε,vε,cε):=(u(γ),v(γ),c(γ))|γ=ε/C0(u_{\varepsilon},v_{\varepsilon},c_{\varepsilon}):=(u_{(\gamma)},v_{(\gamma)},c({\gamma}))|_{\gamma=\varepsilon/C_{0}} satisfies that (uε,vε)(u,0)H3(DT)ε\|(u_{\varepsilon},v_{\varepsilon})-(u,0)\|_{H^{3}(D_{T})}\leq\varepsilon, cεc0c_{\varepsilon}\to c_{0}, uε(x,y)cε>0u_{\varepsilon}(x,y)-c_{\varepsilon}>0 and vεL2(DT)0\|v_{\varepsilon}\|_{L^{2}(D_{T})}\neq 0. By (5.7), v~εα/πϕc0,n(y)sin(kαx){\tilde{v}_{\varepsilon}}\longrightarrow\sqrt{\alpha/\pi}\phi_{c_{0},n_{*}}(y)\sin(k_{*}\alpha x) in H2(DT){H^{2}\left(D_{T}\right)}, where v~ε=vε/vεL2(DT)\tilde{v}_{\varepsilon}=v_{\varepsilon}/\|v_{\varepsilon}\|_{L^{2}(D_{T})}.
Case 2. λn(c0)=0\lambda_{n_{*}}^{\prime}(c_{0})=0 and c0σd(0,β)𝐑c_{0}\notin\sigma_{d}(\mathcal{R}_{0,\beta})\cap\mathbf{R}.

In this case, there exist δ1(0,δ0]\delta_{1}\in(0,\delta_{0}] and a{±1}a\in\{\pm 1\} such that aλna\lambda_{n_{*}} is increasing in [c0,c0+δ1][c_{0},c_{0}+\delta_{1}], and thus,

(5.8) aλn(c)>aλn(c0)=a(kα)2,c(c0,c0+δ1].\displaystyle a\lambda_{n_{*}}(c)>a\lambda_{n_{*}}(c_{0})=-a(k_{*}\alpha)^{2},\quad\forall\ c\in(c_{0},c_{0}+\delta_{1}].

Let ζ1C([y1,y2])\zeta_{1}\in C^{\infty}([y_{1},y_{2}]) be a positive function, u1u_{1} be a solution of the regular ODE

(5.9) u1(uc0)(uβ)u1=ζ1on[y1,y2],\displaystyle u_{1}^{\prime\prime}(u-c_{0})-(u^{\prime\prime}-\beta)u_{1}=\zeta_{1}\quad\text{on}\quad[y_{1},y_{2}],

and τ0>0\tau_{0}>0 be such that [c0,c0+δ1]Ran(u+τu1)=[c_{0},c_{0}+\delta_{1}]\cap\text{Ran}(u+\tau u_{1})=\emptyset for τ[τ0,τ0]\tau\in[-\tau_{0},\tau_{0}]. Since uH4(y1,y2)u\in H^{4}(y_{1},y_{2}) and ζ1C([y1,y2])\zeta_{1}\in C^{\infty}([y_{1},y_{2}]), we have u1H4(y1,y2)u_{1}\in H^{4}(y_{1},y_{2}). Let λn(c,τ)\lambda_{n}(c,\tau) denote the nn-th eigenvalue of c,τ:H2H01(y1,y2)L2(y1,y2)\mathcal{L}_{c,\tau}:H^{2}\cap H_{0}^{1}(y_{1},y_{2})\longrightarrow L^{2}(y_{1},y_{2}) defined by

c,τϕ=ϕ+u+τu1βu+τu1cϕ\displaystyle\mathcal{L}_{c,\tau}\phi=-\phi^{\prime\prime}+{u^{\prime\prime}+\tau u_{1}^{\prime\prime}-\beta\over u+\tau u_{1}-c}\phi

for c[c0,c0+δ1]c\in[c_{0},c_{0}+\delta_{1}] and τ[τ0,τ0]\tau\in[-\tau_{0},\tau_{0}]. Then by (5.9) and the fact that ζ1\zeta_{1} is a positive function, we have

τλn(c0,0)=\displaystyle\partial_{\tau}\lambda_{n_{*}}(c_{0},0)= y1y2τ(u+τu1βu+τu1c0)|τ=0ϕn,c02dy\displaystyle\int_{y_{1}}^{y_{2}}\partial_{\tau}\left({u^{\prime\prime}+\tau u_{1}^{\prime\prime}-\beta\over u+\tau u_{1}-c_{0}}\right)\big{|}_{\tau=0}\phi_{n_{*},c_{0}}^{2}dy
=\displaystyle= y1y2u1(uc0)(uβ)u1(uc0)2ϕn,c02dy=y1y2ζ1(uc0)2ϕn,c02dy>0,\displaystyle\int_{y_{1}}^{y_{2}}{u_{1}^{\prime\prime}(u-c_{0})-(u^{\prime\prime}-\beta)u_{1}\over(u-c_{0})^{2}}\phi_{n_{*},c_{0}}^{2}dy=\int_{y_{1}}^{y_{2}}{\zeta_{1}\over(u-c_{0})^{2}}\phi_{n_{*},c_{0}}^{2}dy>0,

where ϕn,c0\phi_{n_{*},c_{0}} is a L2L^{2} normalized eigenfunction of λn(c0)σ(c0)\lambda_{n_{*}}(c_{0})\in\sigma(\mathcal{L}_{c_{0}}). By the definition of kk_{*}, (kα)2σ(c0,0)-(k\alpha)^{2}\notin\sigma(\mathcal{L}_{c_{0},0}) for k>kk>k_{*}. Since c0σd(0,β)𝐑c_{0}\notin\sigma_{d}(\mathcal{R}_{0,\beta})\cap\mathbf{R}, we have 0σ(c0,0)0\notin\sigma(\mathcal{L}_{c_{0},0}). By the continuity of τλn\partial_{\tau}\lambda_{n_{*}} and the small perturbation of σ(c,τ)\sigma(\mathcal{L}_{c,\tau}), we can take τ0>0\tau_{0}>0 and δ1>0\delta_{1}>0 smaller such that τλn(c,τ)>0\partial_{\tau}\lambda_{n_{*}}(c,\tau)>0 and

(5.10) 0σ(c,τ)and(kα)2σ(c,τ),k>k\displaystyle 0\notin\sigma(\mathcal{L}_{c,\tau})\quad\text{and}\quad-(k\alpha)^{2}\notin\sigma(\mathcal{L}_{c,\tau}),\quad\forall\ k>k_{*}

for (c,τ)[c0,c0+δ1]×[τ0,τ0](c,\tau)\in[c_{0},c_{0}+\delta_{1}]\times[-\tau_{0},\tau_{0}]. By taking δ1>0\delta_{1}>0 smaller and the Implicit Function Theorem, there exists γ~C1([c0,c0+δ1])\widetilde{\gamma}\in C^{1}([c_{0},c_{0}+\delta_{1}]) such that λn(c,γ~(c))=λn(c0,0)=(kα)2,γ~(c0)=0\lambda_{n_{*}}(c,\widetilde{\gamma}(c))=\lambda_{n_{*}}(c_{0},0)=-(k_{*}\alpha)^{2},\ \widetilde{\gamma}(c_{0})=0 and |γ~(c)|τ0|\widetilde{\gamma}(c)|\leq\tau_{0} for c[c0,c0+δ1]c\in[c_{0},c_{0}+\delta_{1}]. By (5.8), we have λn(c,γ~(c))=λn(c0,0)λn(c,0)\lambda_{n_{*}}(c,\widetilde{\gamma}(c))=\lambda_{n_{*}}(c_{0},0)\neq\lambda_{n_{*}}(c,0) and γ~(c)0\widetilde{\gamma}(c)\neq 0 for c(c0,c0+δ1]c\in(c_{0},c_{0}+\delta_{1}]. Then for fixed τ(0,τ0]\tau\in(0,\tau_{0}], there exists cτ[c0,c0+δ1]c_{\tau}\in[c_{0},c_{0}+\delta_{1}] such that γ~(cτ)0\widetilde{\gamma}^{\prime}(c_{\tau})\neq 0 and |γ~(cτ)|τ|\widetilde{\gamma}(c_{\tau})|\leq\tau. Note that 0=ddc[λn(c,γ~(c))]=cλn(c,γ~(c))+γ~(c)τλn(c,γ~(c))0=\frac{d}{dc}[\lambda_{n_{*}}(c,\widetilde{\gamma}(c))]=\partial_{c}\lambda_{n_{*}}(c,\widetilde{\gamma}(c))+\widetilde{\gamma}^{\prime}(c)\partial_{\tau}\lambda_{n_{*}}(c,\widetilde{\gamma}(c)). Let τ1=γ~(cτ)\tau_{1}=\widetilde{\gamma}(c_{\tau}). Then we have cλn(cτ,τ1)=γ~(cτ)τλn(cτ,τ1)0.\partial_{c}\lambda_{n_{*}}(c_{\tau},\tau_{1})=-\widetilde{\gamma}^{\prime}(c_{\tau})\partial_{\tau}\lambda_{n_{*}}(c_{\tau},\tau_{1})\neq 0.

Fix any ε(0,1)\varepsilon\in(0,1). Then we can choose τ(0,τ0]\tau\in(0,\tau_{0}] and δ1>0\delta_{1}>0 smaller such that for τ1=γ~(cτ)\tau_{1}=\widetilde{\gamma}(c_{\tau}),

(5.11) (u+τ1u1,0)(u,0)H3(y1,y2)τ1u1H3(y1,y2)<ε2and|cτc0|<δ1<ε2.\displaystyle\|(u+\tau_{1}u_{1},0)-(u,0)\|_{H^{3}(y_{1},y_{2})}\leq\tau_{1}\|u_{1}\|_{H^{3}(y_{1},y_{2})}<{\varepsilon\over 2}\quad\text{and}\quad|c_{\tau}-c_{0}|<\delta_{1}<{\varepsilon\over 2}.

By (5.10), λn(cτ,τ1)=(kα)2\lambda_{n_{*}}(c_{\tau},\tau_{1})=-(k_{*}\alpha)^{2} and cλn(cτ,τ1)0\partial_{c}\lambda_{n_{*}}(c_{\tau},\tau_{1})\neq 0, we can apply Case 1 to the shear flow (u+τ1u1,0)(u+\tau_{1}u_{1},0): there exists a traveling wave solution (uε(xcεt,y),vε(xcεt,y))(u_{\varepsilon}(x-c_{\varepsilon}t,y),v_{\varepsilon}(x-c_{\varepsilon}t,y)) to (1.1)–(1.2) which has period T=2π/αT={2\pi}/{\alpha} in xx,

(5.12) (uε,vε)(u+τ1u1,0)H3(DT)ε2and|cεcτ|ε2,\displaystyle\|(u_{\varepsilon},v_{\varepsilon})-(u+\tau_{1}u_{1},0)\|_{H^{3}(D_{T})}\leq{\varepsilon\over 2}\quad\text{and}\quad|c_{\varepsilon}-c_{\tau}|\leq{\varepsilon\over 2},

uε(x,y)cε0u_{\varepsilon}\left(x,y\right)-c_{\varepsilon}\neq 0 and vεL2(DT)0\|v_{\varepsilon}\|_{L^{2}\left(D_{T}\right)}\neq 0. Then by (5.11)–(5.12), we have (uε,vε)(u,0)H3(DT)\|(u_{\varepsilon},v_{\varepsilon})-(u,0)\|_{H^{3}(D_{T})} <ε<{\varepsilon} and |cεc0|<ε.|c_{\varepsilon}-c_{0}|<{\varepsilon}.
Case 3. c0σd(0,β)𝐑c_{0}\in\sigma_{d}(\mathcal{R}_{0,\beta})\cap\mathbf{R}.

In this case, 0σ(c0)0\in\sigma(\mathcal{L}_{c_{0}}) and there exists j0>n0nj_{0}>n_{0}\geq n_{*} such that λj0(c0)=0\lambda_{j_{0}}(c_{0})=0. There exist δ1(0,δ0]\delta_{1}\in(0,\delta_{0}] and a,b{±1}a,b\in\{\pm 1\} such that both aλna\lambda_{n_{*}} and bλj0b\lambda_{j_{0}} are decreasing in [c0,c0+δ1][c_{0},c_{0}+\delta_{1}].

Since ϕn,c02\phi_{n_{*},c_{0}}^{2} is linearly independent of ϕj0,c02\phi_{j_{0},c_{0}}^{2}, there exists ξ1C([y1,y2])\xi_{1}\in C^{\infty}([y_{1},y_{2}]) such that

(5.13) y1y2ξ1ϕn,c02(uc0)2dy=aandy1y2ξ1ϕj0,c02(uc0)2dy=b.\displaystyle\int_{y_{1}}^{y_{2}}\xi_{1}{\phi_{n_{*},c_{0}}^{2}\over(u-c_{0})^{2}}dy=a\quad\text{and}\quad\int_{y_{1}}^{y_{2}}\xi_{1}{\phi_{j_{0},c_{0}}^{2}\over(u-c_{0})^{2}}dy=-b.

Let u1u_{1} be a solution of (5.9) with ζ1=ξ1\zeta_{1}=\xi_{1}, and τ0>0\tau_{0}>0 be such that [c0,c0+δ1]Ran(u+τu1)=[c_{0},c_{0}+\delta_{1}]\cap\text{Ran}(u+\tau u_{1})=\emptyset for τ[τ0,τ0]\tau\in[-\tau_{0},\tau_{0}]. Then by (5.13), we have aτλn(c0,0)=a2>0a\partial_{\tau}\lambda_{n_{*}}(c_{0},0)=a^{2}>0 and bτλj0(c0,0)=b2<0.b\partial_{\tau}\lambda_{j_{0}}(c_{0},0)=-b^{2}<0. As in Case 2, we can take τ0>0\tau_{0}>0 and δ1>0\delta_{1}>0 smaller such that

aτλn(c,τ)>0and(kα)2σ(c,τ),k>k\displaystyle a\partial_{\tau}\lambda_{n_{*}}(c,\tau)>0\quad\text{and}\quad-(k\alpha)^{2}\notin\sigma(\mathcal{L}_{c,\tau}),\quad\forall\ k>k_{*}

for (c,τ)[c0,c0+δ1]×[τ0,τ0](c,\tau)\in[c_{0},c_{0}+\delta_{1}]\times[-\tau_{0},\tau_{0}]. Note that λj01(c0,0)<λj0(c0,0)=0<λj0+1(c0,0)\lambda_{j_{0}-1}(c_{0},0)<\lambda_{j_{0}}(c_{0},0)=0<\lambda_{j_{0}+1}(c_{0},0). By the continuity of τλj0\partial_{\tau}\lambda_{j_{0}}, λj01\lambda_{j_{0}-1} and λj0+1\lambda_{j_{0}+1}, we can choose τ0>0\tau_{0}>0 and δ1>0\delta_{1}>0 smaller such that

(5.14) bτλj0(c,τ)<0andλj01(c,τ)<0<λj0+1(c,τ)\displaystyle b\partial_{\tau}\lambda_{j_{0}}(c,\tau)<0\quad\text{and}\quad\lambda_{j_{0}-1}(c,\tau)<0<\lambda_{j_{0}+1}(c,\tau)

for (c,τ)[c0,c0+δ1]×[τ0,τ0](c,\tau)\in[c_{0},c_{0}+\delta_{1}]\times[-\tau_{0},\tau_{0}].

As aτλn(c0,0)>0a\partial_{\tau}\lambda_{n_{*}}(c_{0},0)>0 and aλn(,0)a\lambda_{n_{*}}(\cdot,0) is decreasing in [c0,c0+δ1][c_{0},c_{0}+\delta_{1}], we can choose τ1(0,τ0]\tau_{1}\in(0,\tau_{0}] such that aλn(c0+δ1,τ1)<aλn(c0,0)=a(kα)2<aλn(c0,τ1)a\lambda_{n_{*}}(c_{0}+\delta_{1},\tau_{1})<a\lambda_{n_{*}}(c_{0},0)=-a(k_{*}\alpha)^{2}<a\lambda_{n_{*}}(c_{0},\tau_{1}). Then there exists cτ1(c0,c0+δ1)c_{\tau_{1}}\in(c_{0},c_{0}+\delta_{1}) such that λn(cτ1,τ1)=(kα)2.\lambda_{n_{*}}(c_{\tau_{1}},\tau_{1})=-(k_{*}\alpha)^{2}.

Since bτλj0(c,τ)<0b\partial_{\tau}\lambda_{j_{0}}(c,\tau)<0 and bλj0(,0)b\lambda_{j_{0}}(\cdot,0) is decreasing in [c0,c0+δ1][c_{0},c_{0}+\delta_{1}], we have bλj0(c,τ)<bλj0(c,0)<bλj0(c0,0)=0,b\lambda_{j_{0}}(c,\tau)<b\lambda_{j_{0}}(c,0)<b\lambda_{j_{0}}(c_{0},0)=0, which, along with (5.14), gives

bλj0+b(c,τ)>0>bλj0(c,τ),(c,τ)(c0,c0+δ1]×(0,τ0].\displaystyle b\lambda_{j_{0}+b}(c,\tau)>0>b\lambda_{j_{0}}(c,\tau),\quad(c,\tau)\in(c_{0},c_{0}+\delta_{1}]\times(0,\tau_{0}].

Since (cτ1,τ1)(c0,c0+δ1]×(0,τ0](c_{\tau_{1}},\tau_{1})\in(c_{0},c_{0}+\delta_{1}]\times(0,\tau_{0}], we have 0σ(cτ1,τ1)0\notin\sigma(\mathcal{L}_{c_{\tau_{1}},\tau_{1}}).

Now, we can construct a desired traveling wave solution (uε(xcεt,y),vε(xcεt,y))(u_{\varepsilon}(x-c_{\varepsilon}t,y),v_{\varepsilon}(x-c_{\varepsilon}t,y)) by first perturbing the shear flow (u,0)(u,0) to (u+τ1u1,0)(u+\tau_{1}u_{1},0) and then applying Case 1 or Case 2 to (u+τ1u1,0)(u+\tau_{1}u_{1},0) as in (5.11)–(5.12). ∎

To prove Corollary 2.6, we only need to modify the spaces BB and CC from H4H^{4} and H2H^{2} to Hs+1H^{s+1} and Hs1H^{s-1} in the proof of Lemma 2.5. We also use the fact that f~cC(Ic)\tilde{f}_{c}\in C^{\infty}(I_{c}), fcC0(𝐑)f_{c}\in C_{0}^{\infty}(\mathbf{R}) and u1C([y1,y2])u_{1}\in C^{\infty}([y_{1},y_{2}]) due to the assumption that uC([y1,y2])u\in C^{\infty}([y_{1},y_{2}]).

Conversely, for a set of traveling wave solutions near (u,0)(u,0) with traveling speeds converging to c0c_{0}, we show that c0c_{0} is an isolated real eigenvalue besides uminu_{\min} and umaxu_{\max}, which is given in Lemma 2.7.

Proof of Lemma 2.7.

It suffices to show that if c0{umax,umin}c_{0}\notin\{u_{\max},u_{\min}\}, then c0k1(σd(kα,β)𝐑)c_{0}\in\bigcup_{k\geq 1}(\sigma_{d}(\mathcal{R}_{k\alpha,\beta})\cap\mathbf{R}) and (2.2) holds. Note that (uε,vε)(u_{\varepsilon},v_{\varepsilon}) solves

(5.15) (uεcε)xωε+vεyωε+βvε=0.(u_{\varepsilon}-c_{\varepsilon})\partial_{x}\omega_{\varepsilon}+v_{\varepsilon}\partial_{y}\omega_{\varepsilon}+\beta v_{\varepsilon}=0.

Moreover,

ωεω0H2(DT)C(uε,vε)(u,0)H3(DT)Cε.\displaystyle\|\omega_{\varepsilon}-\omega_{0}\|_{H^{2}(D_{T})}\leq C\|(u_{\varepsilon},v_{\varepsilon})-(u,0)\|_{H^{3}(D_{T})}\leq C\varepsilon.

By taking ε0>0\varepsilon_{0}>0 smaller,

(5.16) |uεcε||ucε||uuε|C1\displaystyle|u_{\varepsilon}-c_{\varepsilon}|\geq|u-c_{\varepsilon}|-|u-u_{\varepsilon}|\geq C^{-1}

for ε(0,ε0)\varepsilon\in(0,\varepsilon_{0}) and y[y1,y2]y\in[y_{1},y_{2}]. Note that πy2y1vεL2(DT)vεL2(DT){\pi\over y_{2}-y_{1}}\|v_{\varepsilon}\|_{L^{2}(D_{T})}\leq\|\nabla v_{\varepsilon}\|_{L^{2}(D_{T})}. By Sobolev embedding, we have

(5.17) vεL4(DT)CvεH1(DT)CvεL2(DT),\displaystyle\|v_{\varepsilon}\|_{L^{4}(D_{T})}\leq C\|v_{\varepsilon}\|_{H^{1}(D_{T})}\leq C\|\nabla v_{\varepsilon}\|_{L^{2}(D_{T})},
(5.18) y(ωεω0)L4(DT)Cy(ωεω0)H1(DT)Cωεω0H2(DT)Cε.\displaystyle\|\partial_{y}(\omega_{\varepsilon}-\omega_{0})\|_{L^{4}(D_{T})}\leq C\|\partial_{y}(\omega_{\varepsilon}-\omega_{0})\|_{H^{1}(D_{T})}\leq C\|\omega_{\varepsilon}-\omega_{0}\|_{H^{2}(D_{T})}\leq C\varepsilon.

Since xωε=x(xvεyuε)=Δvε\partial_{x}\omega_{\varepsilon}=\partial_{x}(\partial_{x}v_{\varepsilon}-\partial_{y}u_{\varepsilon})=\Delta v_{\varepsilon}, we get by (5.15) that

(5.19) Δv~ε+y(ωεω0)uεcεv~ε+βuuεcεv~ε=0,\displaystyle\Delta\tilde{v}_{\varepsilon}+{\partial_{y}(\omega_{\varepsilon}-\omega_{0})\over u_{\varepsilon}-c_{\varepsilon}}\tilde{v}_{\varepsilon}+{\beta-u^{\prime\prime}\over u_{\varepsilon}-c_{\varepsilon}}\tilde{v}_{\varepsilon}=0,

where v~ε=vε/vεL2(DT)\tilde{v}_{\varepsilon}={v_{\varepsilon}/\|v_{\varepsilon}\|_{L^{2}(D_{T})}}. By (5.16), we have |βuuεcε|C\left|{\beta-u^{\prime\prime}\over u_{\varepsilon}-c_{\varepsilon}}\right|\leq C for y[y1,y2]y\in[y_{1},y_{2}] and

Δv~εL2(DT)Cy(ωεω0)L4(DT)v~εL4(DT)+Cv~εL2(DT)\displaystyle\|\Delta\tilde{v}_{\varepsilon}\|_{L^{2}(D_{T})}\leq C\|\partial_{y}(\omega_{\varepsilon}-\omega_{0})\|_{L^{4}(D_{T})}\|\tilde{v}_{\varepsilon}\|_{L^{4}(D_{T})}+C\|\tilde{v}_{\varepsilon}\|_{L^{2}(D_{T})}
\displaystyle\leq Cεv~εH1(DT)+Cv~εL2(DT)Cεv~εH2(DT)1/2+C,\displaystyle C\varepsilon\|\tilde{v}_{\varepsilon}\|_{H^{1}(D_{T})}+C\|\tilde{v}_{\varepsilon}\|_{L^{2}(D_{T})}\leq C\varepsilon\|\tilde{v}_{\varepsilon}\|_{H^{2}(D_{T})}^{1/2}+C,

where we used (5.17)–(5.18) and v~εH1(DT)Cv~εH2(DT)1/2v~εL2(DT)1/2\|\tilde{v}_{\varepsilon}\|_{H^{1}(D_{T})}\leq C\|\tilde{v}_{\varepsilon}\|_{H^{2}(D_{T})}^{1/2}\|\tilde{v}_{\varepsilon}\|_{L^{2}(D_{T})}^{1/2}. Since vε(x,yi)=0v_{\varepsilon}(x,y_{i})=0 for i=1,2i=1,2, we have v~εH2(DT)CΔv~εL2(DT).\|\tilde{v}_{\varepsilon}\|_{H^{2}(D_{T})}\leq C\|\Delta\tilde{v}_{\varepsilon}\|_{L^{2}(D_{T})}. Thus, v~εH2(DT)C\|\tilde{v}_{\varepsilon}\|_{H^{2}(D_{T})}\leq C and

(5.20) v~εC0([0,T]×[y1,y2])Cv~εH2(DT)C,\displaystyle\|\tilde{v}_{\varepsilon}\|_{C^{0}([0,T]\times[y_{1},y_{2}])}\leq C\|\tilde{v}_{\varepsilon}\|_{H^{2}(D_{T})}\leq C,
(5.21) xv~εL4(DT)+yv~εL4(DT)Cv~εH2(DT)C.\displaystyle\|\partial_{x}\tilde{v}_{\varepsilon}\|_{L^{4}(D_{T})}+\|\partial_{y}\tilde{v}_{\varepsilon}\|_{L^{4}(D_{T})}\leq C\|\tilde{v}_{\varepsilon}\|_{H^{2}(D_{T})}\leq C.

Up to a subsequence, there exists v~0H2(DT)\tilde{v}_{0}\in{H^{2}(D_{T})} such that v~εv~0\tilde{v}_{\varepsilon}\rightharpoonup\tilde{v}_{0} in H2(DT)H^{2}(D_{T}), v~εv~0\tilde{v}_{\varepsilon}\rightarrow\tilde{v}_{0} in H1(DT){H^{1}(D_{T})} and v~0L2(DT)=1\|\tilde{v}_{0}\|_{L^{2}(D_{T})}=1. Taking derivative in (5.19) with respect to xx and yy, we get by (5.18) and (5.20)–(5.21) that

xΔv~εL2(DT)\displaystyle\|\partial_{x}\Delta\tilde{v}_{\varepsilon}\|_{L^{2}(D_{T})}
\displaystyle\leq x(y(ωεω0)uεcε)v~ε+y(ωεω0)uεcεxv~ε+x(βuuεcε)v~ε+βuuεcεxv~εL2(DT)\displaystyle\left\|\partial_{x}\left({\partial_{y}(\omega_{\varepsilon}-\omega_{0})\over u_{\varepsilon}-c_{\varepsilon}}\right)\tilde{v}_{\varepsilon}+{\partial_{y}(\omega_{\varepsilon}-\omega_{0})\over u_{\varepsilon}-c_{\varepsilon}}\partial_{x}\tilde{v}_{\varepsilon}+\partial_{x}\left({\beta-u^{\prime\prime}\over u_{\varepsilon}-c_{\varepsilon}}\right)\tilde{v}_{\varepsilon}+{\beta-u^{\prime\prime}\over u_{\varepsilon}-c_{\varepsilon}}\partial_{x}\tilde{v}_{\varepsilon}\right\|_{L^{2}(D_{T})}
\displaystyle\leq C(xy(ωεω0)L2(DT)+y(ωεω0)L2(DT))v~εC0([0,T]×[y1,y2])+\displaystyle C\left(\|\partial_{xy}(\omega_{\varepsilon}-\omega_{0})\|_{L^{2}(D_{T})}+\|\partial_{y}(\omega_{\varepsilon}-\omega_{0})\|_{L^{2}(D_{T})}\right)\|\tilde{v}_{\varepsilon}\|_{C^{0}([0,T]\times[y_{1},y_{2}])}+
Cy(ωεω0)L4(DT)xv~εL4(DT)+Cv~εL2(DT)+Cxv~εL2(DT)C,\displaystyle C\|\partial_{y}(\omega_{\varepsilon}-\omega_{0})\|_{L^{4}(D_{T})}\|\partial_{x}\tilde{v}_{\varepsilon}\|_{L^{4}(D_{T})}+C\|\tilde{v}_{\varepsilon}\|_{L^{2}(D_{T})}+C\|\partial_{x}\tilde{v}_{\varepsilon}\|_{L^{2}(D_{T})}\leq C,

and

yΔv~εL2(DT)\displaystyle\|\partial_{y}\Delta\tilde{v}_{\varepsilon}\|_{L^{2}(D_{T})}
\displaystyle\leq y(y(ωεω0)uεcε)v~ε+y(ωεω0)uεcεyv~ε+y(βuuεcε)v~ε+βuuεcεyv~εL2(DT)\displaystyle\left\|\partial_{y}\left({\partial_{y}(\omega_{\varepsilon}-\omega_{0})\over u_{\varepsilon}-c_{\varepsilon}}\right)\tilde{v}_{\varepsilon}+{\partial_{y}(\omega_{\varepsilon}-\omega_{0})\over u_{\varepsilon}-c_{\varepsilon}}\partial_{y}\tilde{v}_{\varepsilon}+\partial_{y}\left({\beta-u^{\prime\prime}\over u_{\varepsilon}-c_{\varepsilon}}\right)\tilde{v}_{\varepsilon}+{\beta-u^{\prime\prime}\over u_{\varepsilon}-c_{\varepsilon}}\partial_{y}\tilde{v}_{\varepsilon}\right\|_{L^{2}(D_{T})}
\displaystyle\leq C(y2(ωεω0)L2(DT)+y(ωεω0)L2(DT))v~εC0([0,T]×[y1,y2])+\displaystyle C\left(\|\partial_{y}^{2}(\omega_{\varepsilon}-\omega_{0})\|_{L^{2}(D_{T})}+\|\partial_{y}(\omega_{\varepsilon}-\omega_{0})\|_{L^{2}(D_{T})}\right)\|\tilde{v}_{\varepsilon}\|_{C^{0}([0,T]\times[y_{1},y_{2}])}+
Cy(ωεω0)L4(DT)yv~εL4(DT)+Cv~εL2(DT)+Cyv~εL2(DT)C,\displaystyle C\|\partial_{y}(\omega_{\varepsilon}-\omega_{0})\|_{L^{4}(D_{T})}\|\partial_{y}\tilde{v}_{\varepsilon}\|_{L^{4}(D_{T})}+C\|\tilde{v}_{\varepsilon}\|_{L^{2}(D_{T})}+C\|\partial_{y}\tilde{v}_{\varepsilon}\|_{L^{2}(D_{T})}\leq C,

which implies that v~εH3(DT)C\|\tilde{v}_{\varepsilon}\|_{H^{3}(D_{T})}\leq C and thus, v~εv~0\tilde{v}_{\varepsilon}\longrightarrow\tilde{v}_{0} in H2(DT)H^{2}(D_{T}). For any ϕH1(DT)\phi\in H^{1}(D_{T}) with periodic boundary condition in xx and Dirichlet boundary condition in yy, we have

(5.22) 0Ty1y2(v~εϕ+y(ωεω0)uεcεv~εϕ+βuuεcεv~εϕ)dydx=0.\displaystyle\int_{0}^{T}\int_{y_{1}}^{y_{2}}\left(-\nabla\tilde{v}_{\varepsilon}\cdot\nabla\phi+{\partial_{y}(\omega_{\varepsilon}-\omega_{0})\over u_{\varepsilon}-c_{\varepsilon}}\tilde{v}_{\varepsilon}\phi+{\beta-u^{\prime\prime}\over u_{\varepsilon}-c_{\varepsilon}}\tilde{v}_{\varepsilon}\phi\right)dydx=0.

Since v~εL4(DT)Cv~εH1(DT)C\|\tilde{v}_{\varepsilon}\|_{L^{4}(D_{T})}\leq C\|\tilde{v}_{\varepsilon}\|_{H^{1}(D_{T})}\leq C, we have by (5.16) and (5.18) that

|0Ty1y2y(ωεω0)uεcεv~εϕdydx|C0Ty1y2|y(ωεω0)||v~ε||ϕ|dydx\displaystyle\left|\int_{0}^{T}\int_{y_{1}}^{y_{2}}{\partial_{y}(\omega_{\varepsilon}-\omega_{0})\over u_{\varepsilon}-c_{\varepsilon}}\tilde{v}_{\varepsilon}\phi dydx\right|\leq C\int_{0}^{T}\int_{y_{1}}^{y_{2}}|\partial_{y}(\omega_{\varepsilon}-\omega_{0})||\tilde{v}_{\varepsilon}||\phi|dydx
\displaystyle\leq Cy(ωεω0)L4(DT)v~εL4(DT)ϕL2(DT)CεϕL2(DT)0 as ε0+.\displaystyle C\|\partial_{y}(\omega_{\varepsilon}-\omega_{0})\|_{L^{4}(D_{T})}\|\tilde{v}_{\varepsilon}\|_{L^{4}(D_{T})}\|\phi\|_{L^{2}(D_{T})}\leq C\varepsilon\|\phi\|_{L^{2}(D_{T})}\longrightarrow 0\ \ \text{ as }\varepsilon\to 0^{+}.

Noting that v~εv~0\tilde{v}_{\varepsilon}\longrightarrow\tilde{v}_{0} in H2(DT){H^{2}(D_{T})} and sending ε0+\varepsilon\to 0^{+} in (5.22), we have

0Ty1y2(v~0ϕ+βuuc0v~0ϕ)dydx=0.\displaystyle\int_{0}^{T}\int_{y_{1}}^{y_{2}}\left(-\nabla\tilde{v}_{0}\cdot\nabla\phi+{\beta-u^{\prime\prime}\over u-c_{0}}\tilde{v}_{0}\phi\right)dydx=0.

Thus, v~0H2(DT)\tilde{v}_{0}\in H^{2}(D_{T}) is a weak solution of

(5.23) 𝒢c0v~0=Δv~0βuuc0v~0=0.\displaystyle\mathcal{G}_{c_{0}}\tilde{v}_{0}=-\Delta\tilde{v}_{0}-{\beta-u^{\prime\prime}\over u-c_{0}}\tilde{v}_{0}=0.

Since c0Ran(u)c_{0}\notin\text{Ran}(u), we have |βuuc0|C\left|{\beta-u^{\prime\prime}\over u-c_{0}}\right|\leq C for y[y1,y2]y\in[y_{1},y_{2}]. Then by elliptic regularity theory, we have v~0\tilde{v}_{0} is a classical solution of (5.23). Thus, φc0:=v~0ker(𝒢c0)\varphi_{c_{0}}:=\tilde{v}_{0}\in\ker(\mathcal{G}_{c_{0}}). Since Δϕ=0-\Delta\phi=0 has no nontrivial solutions satisfying the boundary conditions, we have |c0|<|c_{0}|<\infty. Since φc0=k𝐙φ^c0,k(y)eikαx0\varphi_{c_{0}}=\sum_{k\in\mathbf{Z}}\widehat{\varphi}_{{c_{0}},k}(y)e^{ik\alpha x}\neq 0 solves (5.23), there exists k0𝐙k_{0}\in\mathbf{Z} such that φ^c0,k00\widehat{\varphi}_{{c_{0}},k_{0}}\neq 0 solves

φ^c0,k0+(k0α)2φ^c0,k0βuuc0φ^c0,k0=0\displaystyle-\widehat{\varphi}_{{c_{0}},k_{0}}^{\prime\prime}+(k_{0}\alpha)^{2}\widehat{\varphi}_{{c_{0}},k_{0}}-{\beta-u^{\prime\prime}\over u-c_{0}}\widehat{\varphi}_{{c_{0}},k_{0}}=0

with φ^c0,k0(y1)=φ^c0,k0(y2)=0\widehat{\varphi}_{{c_{0}},k_{0}}(y_{1})=\widehat{\varphi}_{{c_{0}},k_{0}}(y_{2})=0. Now we show that k00k_{0}\neq 0. Let P0f(x,y)=1T0Tf(x,y)dxP_{0}f(x,y)={1\over T}\int_{0}^{T}f(x,y)dx for fL2(DT)f\in L^{2}(D_{T}). Then P0P_{0} is a bounded linear operator on L2(DT)L^{2}(D_{T}). Since v~ε=vε/vεL2(DT)=xψε/vεL2(DT)\tilde{v}_{\varepsilon}=v_{\varepsilon}/\|v_{\varepsilon}\|_{L^{2}(D_{T})}=-\partial_{x}\psi_{\varepsilon}/\|v_{\varepsilon}\|_{L^{2}(D_{T})}, we have P0v~ε=P0vε=0P_{0}\tilde{v}_{\varepsilon}=P_{0}v_{\varepsilon}=0. Taking limit as ε0+\varepsilon\to 0^{+}, we have P0φc0=P0v~0=0P_{0}\varphi_{c_{0}}=P_{0}\tilde{v}_{0}=0 and thus, φ^c0,0(y)=P0φc0(x,y)0\widehat{\varphi}_{{c_{0}},0}(y)=P_{0}\varphi_{c_{0}}(x,y)\equiv 0, which implies that k00k_{0}\neq 0. Thus, c0k0(σd(kα,β)𝐑)=k1(σd(kα,β)𝐑)c_{0}\in\bigcup_{k\neq 0}(\sigma_{d}(\mathcal{R}_{k\alpha,\beta})\cap\mathbf{R})=\bigcup_{k\geq 1}(\sigma_{d}(\mathcal{R}_{k\alpha,\beta})\cap\mathbf{R}), where we used the fact that σd(kα,β)=σd(kα,β).\sigma_{d}(\mathcal{R}_{k\alpha,\beta})=\sigma_{d}(\mathcal{R}_{-k\alpha,\beta}).

We give two remarks to Lemma 2.7: the first is to study the Fourier expansion of the limit function φc0\varphi_{c_{0}}, and the second is to show that the asymptotic behavior of L2L^{2} normalized vertical velocities v~ε\tilde{v}_{\varepsilon} might be complicated if cεc0{umin,umax}c_{\varepsilon}\rightarrow c_{0}\in\{u_{\min},u_{\max}\}.

Remark 5.1.

The function φc0\varphi_{c_{0}} in Lemma 2.7 is a superposition of finite normal modes. In fact, since 0φc0(x,y)=k𝐙φ^c0,k(y)eikαxker(𝒢c0)0\neq\varphi_{c_{0}}(x,y)=\sum_{k\in\mathbf{Z}}\widehat{\varphi}_{{c_{0}},k}(y)e^{ik\alpha x}\in\ker{(\mathcal{G}_{c_{0}})} and infσ(c0)>\inf\sigma(\mathcal{L}_{c_{0}})>-\infty, we have n:=({k𝐙:(kα)2σ(c0)})[1,)n_{*}:=\sharp(\{k\in\mathbf{Z}:-(k\alpha)^{2}\in\sigma(\mathcal{L}_{c_{0}})\})\in[1,\infty). Let {k𝐙:(kα)2σ(c0)}={kn:1nn}\{k\in\mathbf{Z}:-(k\alpha)^{2}\in\sigma(\mathcal{L}_{c_{0}})\}=\{k_{n}:1\leq n\leq n_{*}\}. Then φc0(x,y)=n=1nφ^c0,kn(y)eiknαx\varphi_{c_{0}}(x,y)=\sum_{n=1}^{n_{*}}\widehat{\varphi}_{{c_{0}},k_{n}}(y)e^{ik_{n}\alpha x} and φ^c0,kn\widehat{\varphi}_{{c_{0}},k_{n}} is an eigenfunction of (knα)2σ(c0)-(k_{n}\alpha)^{2}\in\sigma(\mathcal{L}_{c_{0}}).

Remark 5.2.

Consider a flow uu satisfying (𝐇𝟏)(\bf{H1}), {u=0}{u=umin}\{u^{\prime}=0\}\cap\{u=u_{\min}\}\neq\emptyset, α>0\alpha>0 and β>98κ+\beta>{9\over 8}\kappa_{+}. By Theorem 2.9 (3)(3), there exists {cn}n=1=σd(α,β)𝐑\{c_{n}\}_{n=1}^{\infty}=\sigma_{d}(\mathcal{R}_{\alpha,\beta})\cap\mathbf{R} such that cnuminc_{n}\to u_{\min}^{-}, cn+1>cnc_{n+1}>c_{n} and α2=λn(cn)\alpha^{2}=-\lambda_{n}(c_{n}) for n1n\geq 1. By Lemma 2.5, we can choose εn0+\varepsilon_{n}\to 0^{+} and nearby traveling wave solutions uεn(xcεnt,y)=(uεn(xcεnt,y),\vec{u}_{\varepsilon_{n}}\left(x-c_{\varepsilon_{n}}t,y\right)=(u_{\varepsilon_{n}}\left(x-c_{\varepsilon_{n}}t,y\right), vεn(xcεnt,y))v_{\varepsilon_{n}}\left(x-c_{\varepsilon_{n}}t,y\right)) with period 2π/α{2\pi/\alpha} in xx such that (uεn,vεn)(u,0)H3(DT)εn,\|(u_{\varepsilon_{n}},v_{\varepsilon_{n}})-(u,0)\|_{H^{3}(D_{T})}\leq\varepsilon_{n}, |cεncn|0+|c_{\varepsilon_{n}}-c_{n}|\to 0^{+}, vεnL2(DT)0\|v_{\varepsilon_{n}}\|_{L^{2}\left(D_{T}\right)}\neq 0 and there exists φcnker(𝒢cn)\varphi_{c_{n}}\in\ker(\mathcal{G}_{c_{n}}) such that v~εnφcnC0\left\|\tilde{v}_{\varepsilon_{n}}-\varphi_{c_{n}}\right\|_{C^{0}} is small enough for large nn, where v~εn=vεn/vεnL2(DT)\tilde{v}_{\varepsilon_{n}}={{v}_{\varepsilon_{n}}/\|{v}_{\varepsilon_{n}}\|_{L^{2}\left(D_{T}\right)}}. Then cεnuminc_{\varepsilon_{n}}\to u_{\min}^{-}. A possible case is that there exists a subsequence {nj}j=1\{n_{j}\}_{j=1}^{\infty} such that cnjk>1(σd(kα,β)𝐑)c_{n_{j}}\notin\bigcup_{k>1}(\sigma_{d}(\mathcal{R}_{k\alpha,\beta})\cap\mathbf{R}) for j1j\geq 1. In this case, φcnj(x,y)=α/πϕcnj(y)sin(αx)\varphi_{c_{n_{j}}}(x,y)=\sqrt{\alpha/\pi}\phi_{c_{n_{j}}}(y)\sin(\alpha x) since it is odd in xx (see the construction in Lemma 2.5), where ϕcnj\phi_{c_{n_{j}}} is a L2L^{2} normalized eigenfunction of λnj(cnj)=α2σ(cnj)\lambda_{n_{j}}(c_{n_{j}})=-\alpha^{2}\in\sigma(\mathcal{L}_{c_{n_{j}}}). Since ϕcnj\phi_{c_{n_{j}}} has nj1n_{j}-1 sign-changed zeros in (y1,y2)(y_{1},y_{2}), v~εnj\tilde{v}_{\varepsilon_{n_{j}}} oscillates frequently in the yy-direction for large jj.

The minimal period of any nearby traveling wave solution in xx can be determined under the following condition.

Lemma 5.3.

Under the assumption of Lemma 2.7, if

(5.24) c0σd(α,β)𝐑andc0k2(σd(kα,β)𝐑),\displaystyle c_{0}\in\sigma_{d}(\mathcal{R}_{\alpha,\beta})\cap\mathbf{R}\quad\text{and}\quad c_{0}\notin\bigcup_{k\geq 2}(\sigma_{d}(\mathcal{R}_{k\alpha,\beta})\cap\mathbf{R}),

then uε(xcεt,y)\vec{u}_{\varepsilon}\left(x-c_{\varepsilon}t,y\right) has minimal period 2π/α{2\pi/\alpha} in xx for ε>0\varepsilon>0 small enough.

Proof.

Let the minimal horizontal period of the traveling wave solution uε(xcεt,y)\vec{u}_{\varepsilon}\left(x-c_{\varepsilon}t,y\right) be T/nεT/n_{\varepsilon} for ε(0,ε0)\varepsilon\in(0,\varepsilon_{0}), where nε𝐙+n_{\varepsilon}\in\mathbf{Z}^{+} and T=2π/αT={{2\pi}/\alpha}. Fix ε(0,ε0)\varepsilon\in(0,\varepsilon_{0}) and let v~ε=vε/vεL2(DT)\tilde{v}_{\varepsilon}=v_{\varepsilon}/\|v_{\varepsilon}\|_{L^{2}(D_{T})}. Since v~ε(x,y)=v~ε(x+T/nε,y)\tilde{v}_{\varepsilon}(x,y)=\tilde{v}_{\varepsilon}(x+T/n_{\varepsilon},y) for x𝐑x\in\mathbf{R} and y[y1,y2]y\in[y_{1},y_{2}], we have 0Teiαxv~ε(x,y)dx=eiαT/nε0Teiαxv~ε(x,y)dx\int_{0}^{T}e^{i\alpha x}\tilde{v}_{\varepsilon}(x,y)dx=e^{i\alpha T/n_{\varepsilon}}\int_{0}^{T}e^{i\alpha x}\tilde{v}_{\varepsilon}(x,y)dx. Thus, if nε>1n_{\varepsilon}>1, then eiαT/nε=e2πi/nε1e^{i\alpha T/n_{\varepsilon}}=e^{2\pi i/n_{\varepsilon}}\neq 1 and 0Teiαxv~ε(x,y)dx=0\int_{0}^{T}e^{i\alpha x}\tilde{v}_{\varepsilon}(x,y)dx=0 for y[y1,y2]y\in[y_{1},y_{2}].

Suppose that there exists a sequence {εk:k1}(0,ε0)\{\varepsilon_{k}:k\geq 1\}\subset(0,\varepsilon_{0}) such that εk0+\varepsilon_{k}\to 0^{+} and uεk(xcεkt,y)\vec{u}_{\varepsilon_{k}}\left(x-c_{\varepsilon_{k}}t,y\right) has minimal period T/nεk<T{T/n_{\varepsilon_{k}}}<T in xx for k1k\geq 1. Then 0Teiαxv~εk(x,y)dx=0\int_{0}^{T}e^{i\alpha x}\tilde{v}_{\varepsilon_{k}}(x,y)dx=0 for k1k\geq 1 and y[y1,y2]y\in[y_{1},y_{2}]. By Lemma 2.7, v~εkφc0{\tilde{v}_{\varepsilon_{k}}}\longrightarrow\varphi_{c_{0}} in H2(DT){H^{2}\left(D_{T}\right)}, where φc0ker(𝒢c0)\varphi_{c_{0}}\in\ker(\mathcal{G}_{c_{0}}) and 𝒢c0\mathcal{G}_{c_{0}} is defined in (2.3). Then

(5.25) 0Teiαxφc0(x,y)dx=limk0Teiαxv~εk(x,y)dx=0\displaystyle\int_{0}^{T}e^{i\alpha x}\varphi_{c_{0}}(x,y)dx=\lim_{k\to\infty}\int_{0}^{T}e^{i\alpha x}\tilde{v}_{\varepsilon_{k}}(x,y)dx=0

for y[y1,y2]y\in[y_{1},y_{2}]. By (5.24) and φ^c0,0=0\widehat{\varphi}_{c_{0},0}=0, we have φc0(x,y)=φ^c0,1(y)eiαx+φ^c0,1(y)¯eiαx\varphi_{c_{0}}(x,y)=\widehat{\varphi}_{{c_{0}},1}(y)e^{i\alpha x}+\overline{\widehat{\varphi}_{{c_{0}},1}(y)}e^{-i\alpha x}, where φ^c0,10\widehat{\varphi}_{{c_{0}},1}\neq 0 is an eigenfunction of α2σ(c0)-\alpha^{2}\in\sigma(\mathcal{L}_{c_{0}}). On the other hand, we have

0Teiαxφc0(x,y)dx=Tφ^c0,1(y)¯0,\displaystyle\int_{0}^{T}e^{i\alpha x}\varphi_{c_{0}}(x,y)dx=T\overline{\widehat{\varphi}_{{c_{0}},1}(y)}\not\equiv 0,

which contradicts (5.25). ∎

Remark 5.4.

(1)(1) Let c0k1(σd(kα,β)𝐑)c_{0}\in\bigcup_{k\geq 1}(\sigma_{d}(\mathcal{R}_{k\alpha,\beta})\cap\mathbf{R}) and kk_{*} be defined in (5.4). It follows from Lemma 5.3 that under the assumption of Lemma 2.7, if uε(xcεt,y)\vec{u}_{\varepsilon}\left(x-c_{\varepsilon}t,y\right) has period 2π/(kα){2\pi}/({k_{*}\alpha}) in xx for ε(0,ε0)\varepsilon\in(0,\varepsilon_{0}), then the period 2π/(kα){2\pi}/({k_{*}\alpha}) is minimal for ε>0\varepsilon>0 small enough.

(2)(2) In Proposition 77 of [28], it should be corrected that the minimal period of constructed traveling wave solutions in xx might be less than 2π/α02\pi/\alpha_{0}, since it is possible that (c0,kα0,β,ϕk)(c_{0},k\alpha_{0},\beta,\phi_{k}) is a non-resonant neutral mode for some k2k\geq 2 and ϕkH01H2(y1,y2)\phi_{k}\in H_{0}^{1}\cap H^{2}(y_{1},y_{2}). Consequently, the minimal period of constructed traveling wave solutions near the sinus profile in Theorem 77 (i)(\rm{i}) of [28] might be less than 2π/α02\pi/\alpha_{0}, see Example 7.1 for systematic study of traveling wave families near the sinus profile. If (5.24) holds true for α=α0\alpha=\alpha_{0}, then the minimal period of these traveling wave solutions in Proposition 77 and Theorem 77 (i)(\rm{i}) of [28] is 2π/α02\pi/\alpha_{0}.

6. The number of traveling wave families near a shear flow

In this section, we prove the main theorems-Theorems 2.2 and 1.2. The proof is based on the study on the number of isolated real eigenvalues of the linearized Euler operator in Sections 3-4, and correspondence between a traveling wave family near the shear flow and an isolated real eigenvalue in Section 5. We only prove Theorem 2.2, since the other is similar.

Proof of Theorem 2.2..

Let the number of traveling wave families near (u,0)(u,0) be denoted by θ\theta. By Theorem 2.1, θ=(k1(σd(kα,β)𝐑))\theta=\sharp(\bigcup_{k\geq 1}(\sigma_{d}(\mathcal{R}_{k\alpha,\beta})\cap\mathbf{R})). Here α=2π/T\alpha=2\pi/T.
Proof of (1): Since {u=0}{u=umin}\{u^{\prime}=0\}\cap\{u=u_{\min}\}\neq\emptyset, we have 0<κ+<0<\kappa_{+}<\infty. First, let {u=umin}(y1,y2)\{u=u_{\min}\}\cap(y_{1},y_{2})\neq\emptyset and we divide the discussion into two cases.
Case 1a. β(0,min{98κ+,μ+})\beta\in(0,\min\{{9\over 8}\kappa_{+},\mu_{+}\}).

By Corollary 2.10 (1),

(6.1) infc(,umin)λ1(c)>.\displaystyle\inf_{c\in(-\infty,u_{\min})}\lambda_{1}(c)>-\infty.

Thus, there exists 1k0<1\leq k_{0}<\infty such that

(6.2) k>k0(σd(kα,β)𝐑)=.\displaystyle\bigcup_{k>k_{0}}(\sigma_{d}(\mathcal{R}_{k\alpha,\beta})\cap\mathbf{R})=\emptyset.

By Theorem 2.11 (1), we have

(6.3) θ=(k1(σd(kα,β)𝐑))k1(σd(kα,β)𝐑)=k=1k0(σd(kα,β)𝐑)<.\displaystyle\theta=\sharp\big{(}\bigcup_{k\geq 1}(\sigma_{d}(\mathcal{R}_{k\alpha,\beta})\cap\mathbf{R})\big{)}\leq\sum_{k\geq 1}\sharp(\sigma_{d}(\mathcal{R}_{k\alpha,\beta})\cap\mathbf{R})=\sum_{k=1}^{k_{0}}\sharp(\sigma_{d}(\mathcal{R}_{k\alpha,\beta})\cap\mathbf{R})<\infty.

Case 1b. β(min{98κ+,μ+},)\beta\in(\min\{{9\over 8}\kappa_{+},\mu_{+}\},\infty).

By Corollary 2.10 (1) for β(min{98κ+,μ+},98κ+]\beta\in(\min\{{9\over 8}\kappa_{+},\mu_{+}\},{9\over 8}\kappa_{+}] and Theorem 2.9 (3) for β(98κ+,)\beta\in({9\over 8}\kappa_{+},\infty), we have

(6.4) limcuminλ1(c)=.\displaystyle\lim_{c\to u_{\min}^{-}}\lambda_{1}(c)=-\infty.

Thus, there exists ckσd(kα,β)𝐑c_{k}\in\sigma_{d}(\mathcal{R}_{k\alpha,\beta})\cap\mathbf{R} such that ck<ck+1<uminc_{k}<c_{k+1}<u_{\min} for every k1k\geq 1, and ckuminc_{k}\to u_{\min}^{-}. Then θ=(k1(σd(kα,β)𝐑))=\theta=\sharp(\bigcup_{k\geq 1}(\sigma_{d}(\mathcal{R}_{k\alpha,\beta})\cap\mathbf{R}))=\infty.

Next, let {u=umin}(y1,y2)=\{u=u_{\min}\}\cap(y_{1},y_{2})=\emptyset and we separate the proof into two cases.
Case 2a. β(0,98κ+)\beta\in(0,{9\over 8}\kappa_{+}).

By Corollary 2.10 (3), we obtain (6.1). Thus, there exists 1k0<1\leq k_{0}<\infty such that (6.2) holds. By Theorem 2.11 (1), we obtain (6.3).
Case 2b. β(98κ+,)\beta\in({9\over 8}\kappa_{+},\infty).

By Theorem 2.9 (3), we have

(6.5) limcuminλn(c)=,n1.\displaystyle\lim_{c\to u_{\min}^{-}}\lambda_{n}(c)=-\infty,\;\;n\geq 1.

Using (6.5) for n=1n=1 and the fact that θ=(k1(σd(kα,β)𝐑))\theta=\sharp(\bigcup_{k\geq 1}(\sigma_{d}(\mathcal{R}_{k\alpha,\beta})\cap\mathbf{R})), it can be proved that θ=\theta=\infty by a similar way as in Case 1b. This completes the proof of (1). The proof of (2) is similar.
Proof of (3): Since {u=0}{u=umin}=\{u^{\prime}=0\}\cap\{u=u_{\min}\}=\emptyset, we have {u=umin}(y1,y2)=\{u=u_{\min}\}\cap(y_{1},y_{2})=\emptyset and κ+=\kappa_{+}=\infty. Fix β(0,)\beta\in(0,\infty). By Corollary 2.10 (3), we obtain (6.1), and thus, there exists 1k0<1\leq k_{0}<\infty such that (6.2) holds. By Theorem 2.12 (1), we obtain (6.3). This completes the proof of (3). The proof of (4) is similar. ∎

Remark 6.1.

In Cases 1b1b and 2b2b of the above proof, the infinitely many traveling wave families are produced by the asymptotic behavior of the first eigenvalue λ1(c)\lambda_{1}(c) of c\mathcal{L}_{c}, see (6.4). There could be many other traveling wave families in general, which are produced by the asymptotic behavior of λn(c)\lambda_{n}(c) for n2n\geq 2, see (6.5). In fact, if β(98κ+,)\beta\in({9\over 8}\kappa_{+},\infty), for fixed n2n\geq 2, there exists ck,nσd(kα,β)𝐑c_{k,n}\in\sigma_{d}(\mathcal{R}_{k\alpha,\beta})\cap\mathbf{R} such that λn(ck,n)=(kα)2\lambda_{n}(c_{k,n})=-(k\alpha)^{2} for every k1k\geq 1. If {ck,n:k1,n2}{ck:k1}\{c_{k,n}:k\geq 1,n\geq 2\}\setminus\{c_{k}:k\geq 1\}\neq\emptyset, then a simple application of Theorem 2.1 yields other traveling wave families.

7. Application to the sinus profile

In this section, we apply our main results to the sinus profile. Moreover, we calculate the explicit number of isolated real eigenvalues of α,β\mathcal{R}_{\alpha,\beta} and traveling wave families near sinus profile.

Example 7.1.

The sinus profile is u(y)=1+cos(πy)2,u(y)={\frac{1+\cos(\pi y)}{2}}, y[1,1]y\in[-1,1]. We determine (σd(α,β)\sharp\big{(}\sigma_{d}(\mathcal{R}_{\alpha,\beta})\cap 𝐑)\mathbf{R}\big{)} and the number of traveling wave families for the sinus profile on the (α,β)(\alpha,\beta)’s region. For the sinus profile, we have umin=0u_{\min}=0, umax=1u_{\max}=1, {u=0}{u=umin}={±1}\{u^{\prime}=0\}\cap\{u=u_{\min}\}=\{\pm 1\}, {u=0}{u=umax}={0}\{u^{\prime}=0\}\cap\{u=u_{\max}\}=\{0\}, κ+=u(±1)=12π2\kappa_{+}=u^{\prime\prime}(\pm 1)={1\over 2}\pi^{2} and κ=u(0)=12π2\kappa_{-}=u^{\prime\prime}(0)=-{1\over 2}\pi^{2}. We divide the plane into nine parts as follows.

β\betaα\alpha916π2-{9\over 16}\pi^{2}12π2-{1\over 2}\pi^{2}βl\beta_{l}012π2{1\over 2}\pi^{2}916π2{9\over 16}\pi^{2}314π2{\sqrt{3}-1\over 4}\pi^{2}IIVIIIVIIIIIIIIIIIIIIVIVIXIXVVVIIVIIVIVI

Figure 3.

In Figure 33,

I\displaystyle I ={(α,β)|α>0,β<916π2},\displaystyle=\{(\alpha,\beta)|\alpha>0,\beta<-{9\over 16}\pi^{2}\},
II\displaystyle II ={(α,β)|0<α<πr2r+34,916π2β<12π2,r[14,12)},\displaystyle=\{(\alpha,\beta)|0<\alpha<\pi\sqrt{-r^{2}-r+{3\over 4}},-{9\over 16}\pi^{2}\leq\beta<-{1\over 2}\pi^{2},r\in[{1\over 4},{1\over 2})\},
III\displaystyle III ={(α,β)|απr2r+34,916π2β<12π2,r[14,12)}\displaystyle=\{(\alpha,\beta)|\alpha\geq\pi\sqrt{-r^{2}-r+{3\over 4}},-{9\over 16}\pi^{2}\leq\beta<-{1\over 2}\pi^{2},r\in[{1\over 4},{1\over 2})\}\cup
{(α,β)|0<α<32π,β=12π2},\displaystyle\{(\alpha,\beta)|0<\alpha<{\sqrt{3}\over 2}\pi,\beta=-{1\over 2}\pi^{2}\},
IV\displaystyle IV ={(α,β)|0<α<Λβ,12π2<β<βl},\displaystyle=\{(\alpha,\beta)|0<\alpha<\sqrt{\Lambda_{\beta}},-{1\over 2}\pi^{2}<\beta<\beta_{l}\},
V\displaystyle V ={(α,β)|π1r2<α<Λβ,314π2<β<12π2},\displaystyle=\{(\alpha,\beta)|\pi\sqrt{1-r^{2}}<\alpha<\sqrt{\Lambda_{\beta}},{\sqrt{3}-1\over 4}\pi^{2}<\beta<{1\over 2}\pi^{2}\},
VI\displaystyle VI ={(α,β)|0<α<πr2r+34,12π2<β916π2,r[14,12)},\displaystyle=\{(\alpha,\beta)|0<\alpha<\pi\sqrt{-r^{2}-r+{3\over 4}},{1\over 2}\pi^{2}<\beta\leq{9\over 16}\pi^{2},r\in[{1\over 4},{1\over 2})\},
VII\displaystyle VII ={(α,β)|0<α<Λβ,0<β314π2}\displaystyle=\{(\alpha,\beta)|0<\alpha<\sqrt{\Lambda_{\beta}},0<\beta\leq{\sqrt{3}-1\over 4}\pi^{2}\}\cup
{(α,β)|0<απ1r2,314π2<β<12π2,r(12,32)}\displaystyle\{(\alpha,\beta)|0<\alpha\leq\pi\sqrt{1-r^{2}},{\sqrt{3}-1\over 4}\pi^{2}<\beta<{1\over 2}\pi^{2},r\in({1\over 2},{\sqrt{3}\over 2})\}\cup
{(α,β)|πr2r+34α<π1r2,12π2β916π2,r[14,12]},\displaystyle\{(\alpha,\beta)|\pi\sqrt{-r^{2}-r+{3\over 4}}\leq\alpha<\pi\sqrt{1-r^{2}},{1\over 2}\pi^{2}\leq\beta\leq{9\over 16}\pi^{2},r\in[{1\over 4},{1\over 2}]\},
VIII\displaystyle VIII ={(α,β)|α>0,β>916π2},\displaystyle=\{(\alpha,\beta)|\alpha>0,\beta>{9\over 16}\pi^{2}\},
IX\displaystyle IX ={(α,β)|α32π,β=12π2}{(α,β)|α>Λβ,12π2<β<12π2}{(α,β)|\displaystyle=\{(\alpha,\beta)|\alpha\geq{\sqrt{3}\over 2}\pi,\beta=-{1\over 2}\pi^{2}\}\cup\{(\alpha,\beta)|\alpha>\sqrt{\Lambda_{\beta}},-{1\over 2}\pi^{2}<\beta<{1\over 2}\pi^{2}\}\cup\{(\alpha,\beta)|
α=Λβ,0<β314π2}{(α,β)|απ1r2,12π2β916π2,r[14,12]},\displaystyle\alpha=\sqrt{\Lambda_{\beta}},0<\beta\leq{\sqrt{3}-1\over 4}\pi^{2}\}\cup\{(\alpha,\beta)|\alpha\geq\pi\sqrt{1-r^{2}},{1\over 2}\pi^{2}\leq\beta\leq{9\over 16}\pi^{2},r\in[{1\over 4},{1\over 2}]\},

where Λβ=supc(0,1)max{λ1(c),0},\Lambda_{\beta}=\sup_{c\notin(0,1)}\max\{-\lambda_{1}(c),0\}, r=14+916+βπ2r={1\over 4}+\sqrt{{9\over 16}+{\beta\over\pi^{2}}} for 916π2β<0-{9\over 16}\pi^{2}\leq\beta<0, r=14+916βπ2r={1\over 4}+\sqrt{{9\over 16}-{\beta\over\pi^{2}}} for 0<β916π20<\beta\leq{9\over 16}\pi^{2}, and βl\beta_{l} is given by Theorem 66 of [28]. Moreover, mβ=0m_{\beta}=0 and max{Mβ,0}=Λβ\max\{M_{\beta},0\}=\Lambda_{\beta} for β[12π2,0)(0,916π2]\beta\in[-{1\over 2}\pi^{2},0)\cup(0,{9\over 16}\pi^{2}], and mβ=1m_{\beta}=1 and Mβ=(r2r+34)π2M_{\beta}=(-r^{2}-r+{3\over 4})\pi^{2} for β[916π2,12π2)\beta\in[-{9\over 16}\pi^{2},-{1\over 2}\pi^{2}).

The explicit number (σd(α,β)𝐑)\sharp\left(\sigma_{d}(\mathcal{R}_{\alpha,\beta})\cap\mathbf{R}\right) is given as follows:

(α,β)IX\displaystyle(\alpha,\beta)\in IX (σd(α,β)𝐑)=0;\displaystyle\Longrightarrow\sharp\left(\sigma_{d}(\mathcal{R}_{\alpha,\beta})\cap\mathbf{R}\right)=0;
(α,β)IIIVII\displaystyle(\alpha,\beta)\in III\cup VII (σd(α,β)𝐑)=1;\displaystyle\Longrightarrow\sharp\left(\sigma_{d}(\mathcal{R}_{\alpha,\beta})\cap\mathbf{R}\right)=1;
(7.1) (α,β)IVVVI\displaystyle(\alpha,\beta)\in IV\cup V\cup VI (σd(α,β)𝐑)=2;\displaystyle\Longrightarrow\sharp\left(\sigma_{d}(\mathcal{R}_{\alpha,\beta})\cap\mathbf{R}\right)=2;
(α,β)II\displaystyle(\alpha,\beta)\in II (σd(α,β)𝐑)=3;\displaystyle\Longrightarrow\sharp\left(\sigma_{d}(\mathcal{R}_{\alpha,\beta})\cap\mathbf{R}\right)=3;
(α,β)IVIII\displaystyle(\alpha,\beta)\in I\cup VIII (σd(α,β)𝐑)=.\displaystyle\Longrightarrow\sharp\left(\sigma_{d}(\mathcal{R}_{\alpha,\beta})\cap\mathbf{R}\right)=\infty.

In addition, (σd(α,β)𝐑)=1\sharp\left(\sigma_{d}(\mathcal{R}_{\alpha,\beta})\cap\mathbf{R}\right)=1 if (α,β)Γ:={(α,β)|α=Λβ,12π2<β<βl or (\alpha,\beta)\in\Gamma:=\{(\alpha,\beta)|\alpha=\sqrt{\Lambda_{\beta}},-{1\over 2}\pi^{2}<\beta<\beta_{l}\text{ or } 314π2<β<π22}{\sqrt{3}-1\over 4}\pi^{2}<\beta<{\pi^{2}\over 2}\}. Now we fix α=2π/T\alpha=2\pi/T.

The number (denoted by θ\theta) of traveling wave families near the sinus profile is given by

(α,β)IX\displaystyle(\alpha,\beta)\in IX θ=0;\displaystyle\Longrightarrow\theta=0;
(7.2) (α,β)VII\displaystyle(\alpha,\beta)\in VII 1θ<;\displaystyle\Longrightarrow 1\leq\theta<\infty;
(α,β)IVVVI\displaystyle(\alpha,\beta)\in IV\cup V\cup VI 2θ<;\displaystyle\Longrightarrow 2\leq\theta<\infty;
(α,β)IIIIIIVIII,βπ22\displaystyle(\alpha,\beta)\in I\cup II\cup III\cup VIII,\ \beta\neq-\frac{\pi^{2}}{2} θ=.\displaystyle\Longrightarrow\theta=\infty.

In addition, θ=(σd(α,β)𝐑)=1\theta=\sharp(\sigma_{d}(\mathcal{R}_{\alpha,\beta})\cap\mathbf{R})=1 if (α,β)Γ(\alpha,\beta)\in\Gamma. Moreover,

(7.3) (α,β)IIIIVVVIIθ=k1(σd(kα,β)𝐑).\displaystyle(\alpha,\beta)\in III\cup IV\cup V\cup VII\Longrightarrow\theta=\sum_{k\geq 1}\sharp(\sigma_{d}(\mathcal{R}_{k\alpha,\beta})\cap\mathbf{R}).

If (α,β)IIIIVVVIIΓ(\alpha,\beta)\in III\cup IV\cup V\cup VII\cup\Gamma, then

(7.4) c0σd(α,β)𝐑uε(xcεt,y) has minimal period 2π/α in x\displaystyle c_{0}\in\sigma_{d}(\mathcal{R}_{\alpha,\beta})\cap\mathbf{R}\Longrightarrow\vec{u}_{\varepsilon}\left(x-c_{\varepsilon}t,y\right)\text{ has minimal period }{2\pi/\alpha}\text{ in }x

for ε>0\varepsilon>0 small enough, where uε(xcεt,y)\vec{u}_{\varepsilon}\left(x-c_{\varepsilon}t,y\right) satisfies the assumption of Lemma 2.7.

To prove (7.1)–(7.4), we need the following asymptotic behavior, signatures and monotonicity of λn\lambda_{n}.

(1)(\rm{1}) limc±λn(c)=n24π2>0\lim_{c\to\pm\infty}\lambda_{n}(c)={n^{2}\over 4}\pi^{2}>0 for β𝐑\beta\in\mathbf{R};

(2)(\rm{2}) limc0λn(c)=\lim_{c\to 0^{-}}\lambda_{n}(c)=-\infty and λn\lambda_{n} is decreasing on (,0)(-\infty,0) for β(916π2,)\beta\in({9\over 16}\pi^{2},\infty);

(3)(\rm{3}) limc0λn(c)=λn(0)\lim_{c\to 0^{-}}\lambda_{n}(c)=\lambda_{n}(0), λ1(0)<λ2(0)0\lambda_{1}(0)<\lambda_{2}(0)\leq 0, λ3(0)>0\lambda_{3}(0)>0 and λn\lambda_{n} is decreasing on (,0)(-\infty,0) for β[12π2,916π2]\beta\in[{1\over 2}\pi^{2},{9\over 16}\pi^{2}];

(4)(\rm{4}) limc0λn(c)=λn(0)\lim_{c\to 0^{-}}\lambda_{n}(c)=\lambda_{n}(0), λ1(0)<0\lambda_{1}(0)<0, λ2>0\lambda_{2}>0 on (,0](-\infty,0], there exist c1<c2(,0)c_{1}<c_{2}\in(-\infty,0) such that λ1(c)λ1(c1)=0\lambda_{1}(c)\geq\lambda_{1}(c_{1})=0 for c(,c1)c\in(-\infty,c_{1}), λ1\lambda_{1} is decreasing on (c1,c2)(c_{1},c_{2}), λ1(c2)=infc(,0)λ1(c)\lambda_{1}(c_{2})=\inf_{c\in(-\infty,0)}\lambda_{1}(c) and λ1\lambda_{1} is increasing on (c2,0)(c_{2},0) for β(314π2,12π2)\beta\in({\sqrt{3}-1\over 4}\pi^{2},{1\over 2}\pi^{2});

(5)(\rm{5}) limc0λn(c)=λn(0)\lim_{c\to 0^{-}}\lambda_{n}(c)=\lambda_{n}(0), λ1(0)<0\lambda_{1}(0)<0, λ2>0\lambda_{2}>0 on (,0](-\infty,0], there exists c1(,0)c_{1}\in(-\infty,0) such that λ1(c)λ1(c1)=0\lambda_{1}(c)\geq\lambda_{1}(c_{1})=0 for c(,c1)c\in(-\infty,c_{1}) and λ1\lambda_{1} is decreasing on (c1,0)(c_{1},0) for β(0,314π2]\beta\in(0,{\sqrt{3}-1\over 4}\pi^{2}];

(6)(\rm{6}) λ10\lambda_{1}\geq 0 on (1,)(1,\infty) for β[βl,0)\beta\in[\beta_{l},0);

(7)(\rm{7}) limc1+λn(c)=λn(1)\lim_{c\to 1^{+}}\lambda_{n}(c)=\lambda_{n}(1), λ1(1)>0\lambda_{1}(1)>0, λ2>0\lambda_{2}>0 on (1,)(1,\infty), there exist c1<c2<c3(1,)c_{1}<c_{2}<c_{3}\in(1,\infty) such that λ1(c)λ1(c1)=λ1(c3)=0\lambda_{1}(c)\geq\lambda_{1}(c_{1})=\lambda_{1}(c_{3})=0 for c(1,c1)(c3,)c\in(1,c_{1})\cup(c_{3},\infty), λ1\lambda_{1} is decreasing on (c1,c2)(c_{1},c_{2}), λ1(c2)=infc(1,)λ1(c)\lambda_{1}(c_{2})=\inf_{c\in(1,\infty)}\lambda_{1}(c) and λ1\lambda_{1} is increasing on (c2,c3)(c_{2},c_{3}) for β(12π2,βl)\beta\in(-{1\over 2}\pi^{2},\beta_{l});

(8)(\rm{8}) limc1+λn(c)=λn(1)\lim_{c\to 1^{+}}\lambda_{n}(c)=\lambda_{n}(1), λ1(1)<0\lambda_{1}(1)<0, λ2(1)=0\lambda_{2}(1)=0 and λn\lambda_{n} is increasing on (1,)(1,\infty) for β=12π2\beta=-{1\over 2}\pi^{2};

(9)(\rm{9}) limc1+λ1(c)=\lim_{c\to 1^{+}}\lambda_{1}(c)=-\infty, limc1+λn+1(c)=λn(1)\lim_{c\to 1^{+}}\lambda_{n+1}(c)=\lambda_{n}(1), λ1(1)=λ2(1)<0\lambda_{1}(1)=\lambda_{2}(1)<0, λ3(1)>0\lambda_{3}(1)>0 and λn\lambda_{n} is increasing on (1,)(1,\infty) for β[916π2,12π2)\beta\in[-{9\over 16}\pi^{2},-{1\over 2}\pi^{2});

(10)(\rm{10}) limc1+λn(c)=\lim_{c\to 1^{+}}\lambda_{n}(c)=-\infty and λn\lambda_{n} is increasing on (1,)(1,\infty) for β(,916π2)\beta\in(-\infty,-{9\over 16}\pi^{2}),
where n1n\geq 1, λn(0)=((r+n12)21)π2\lambda_{n}(0)=\left(\left(r+{n-1\over 2}\right)^{2}-1\right)\pi^{2} for β(0,916π2]\beta\in(0,{9\over 16}\pi^{2}], λn(1)=((r12+n2)21)\lambda_{n}(1)=\left(\left(r-{1\over 2}+\lceil{n\over 2}\rceil\right)^{2}-1\right) π2\pi^{2} for β[916π2,12π2)(12π2,0)\beta\in[-{9\over 16}\pi^{2},-{1\over 2}\pi^{2})\cup(-{1\over 2}\pi^{2},0), and λn(1)=(n241)π2\lambda_{n}(1)=({n^{2}\over 4}-1)\pi^{2} for β=12π2\beta=-{1\over 2}\pi^{2} by Proposition 11 in [28].

Assertions (1)(1)(10)(10) provide pictures of the negative eigenvalues of c\mathcal{L}_{c} for fixed β\beta. Assume that (1)(1)(10)(10) are true. Note that σd(α,β)(1,)=\sigma_{d}(\mathcal{R}_{\alpha,\beta})\cap(1,\infty)=\emptyset if α𝐑\alpha\in\mathbf{R} and β>0\beta>0, and σd(α,β)(,0)=\sigma_{d}(\mathcal{R}_{\alpha,\beta})\cap(-\infty,0)=\emptyset if α𝐑\alpha\in\mathbf{R} and β<0\beta<0. Then we claim that

(7.5) (σd(kα,β)𝐑)(σd(jα,β)𝐑)=,\displaystyle(\sigma_{d}(\mathcal{R}_{k\alpha,\beta})\cap\mathbf{R})\cap(\sigma_{d}(\mathcal{R}_{j\alpha,\beta})\cap\mathbf{R})=\emptyset,
(7.6) σd(kα,β)𝐑={c𝐑[0,1]|λ1(c)=(kα)2}\displaystyle\sigma_{d}(\mathcal{R}_{k\alpha,\beta})\cap\mathbf{R}=\{c\in\mathbf{R}\setminus[0,1]|\lambda_{1}(c)=-(k\alpha)^{2}\}

for k,j𝐙+k,j\in\mathbf{Z}^{+}, kjk\neq j and (α,β)IIIIVVVIIΓ(\alpha,\beta)\in III\cup IV\cup V\cup VII\cup\Gamma. In fact, (7.6) implies (7.5). If (α,β)III(\alpha,\beta)\in III, then by (8)(8)(9)(9) we have λ2>λ2(1)α2\lambda_{2}>\lambda_{2}(1)\geq-\alpha^{2} on (1,)(1,\infty), which gives (7.6). If (α,β)IVΓ(\alpha,\beta)\in IV\cup\Gamma with β<0\beta<0, then by (7)(7) we have λ2>0\lambda_{2}>0 on (1,)(1,\infty) and thus, c\mathcal{L}_{c} has at most one negative eigenvalue for c(1,)c\in(1,\infty), which gives (7.6). Similarly, we can prove (7.6) for (α,β)VVIIΓ(\alpha,\beta)\in V\cup VII\cup\Gamma with 0<β<π220<\beta<{\pi^{2}\over 2} by (4)(4)(5)(5). If (α,β)VII(\alpha,\beta)\in VII with βπ22\beta\geq{\pi^{2}\over 2}, then by (3)(3) we have λ2>λ2(0)α2\lambda_{2}>\lambda_{2}(0)\geq-\alpha^{2} on (,0)(-\infty,0), which gives (7.6).

By applying Theorem 2.1, we get (7.1)–(7.3). (7.4) is a direct consequence of Lemma 5.3. Here, (7.5) is used in the proof of (7.3)–(7.4).

Using (7.3) we can evaluate θ\theta for (α,β)VI(\alpha,\beta)\not\in VI as follows.

β(,12π2)(916π2,+)\displaystyle\beta\in(-\infty,-{1\over 2}\pi^{2})\cup({9\over 16}\pi^{2},+\infty) θ=;\displaystyle\Longrightarrow\theta=\infty;
(7.7) β=12π2\displaystyle\beta=-{1\over 2}\pi^{2} θ=3π2α1;\displaystyle\Longrightarrow\theta=\lceil\frac{\sqrt{3}\pi}{2\alpha}\rceil-1;
12π2<β<βl\displaystyle-{1\over 2}\pi^{2}<\beta<\beta_{l} θ=Λβα+Λβα1;\displaystyle\Longrightarrow\theta=\lfloor\frac{\sqrt{\Lambda_{\beta}}}{\alpha}\rfloor+\lceil\frac{\sqrt{\Lambda_{\beta}}}{\alpha}\rceil-1;
βlβ0\displaystyle\beta_{l}\leq\beta\leq 0 θ=0;\displaystyle\Longrightarrow\theta=0;
0<β314π2\displaystyle 0<\beta\leq{\sqrt{3}-1\over 4}\pi^{2} θ=Λβα1;\displaystyle\Longrightarrow\theta=\lceil\frac{\sqrt{\Lambda_{\beta}}}{\alpha}\rceil-1;
314π2<β<12π2\displaystyle{\sqrt{3}-1\over 4}\pi^{2}<\beta<{1\over 2}\pi^{2} θ=Λβα+Λβαπ1r2α1;\displaystyle\Longrightarrow\theta=\lfloor\frac{\sqrt{\Lambda_{\beta}}}{\alpha}\rfloor+\lceil\frac{\sqrt{\Lambda_{\beta}}}{\alpha}\rceil-\lfloor\frac{\pi\sqrt{1-r^{2}}}{\alpha}\rfloor-1;
πr2r+34α,12π2β916π2\displaystyle\pi\sqrt{-r^{2}-r+{3\over 4}}\leq\alpha,\ {1\over 2}\pi^{2}\leq\beta\leq{9\over 16}\pi^{2} θ=π1r2α1.\displaystyle\Longrightarrow\theta=\lceil\frac{{\pi\sqrt{1-r^{2}}}}{\alpha}\rceil-1.

The case (α,β)VI(\alpha,\beta)\in VI is more complicated. By (3)(3), we have θ=(A1A2)\theta=\sharp(A_{1}\cup A_{2}) with Ai={c<0|λi(c)=(kα)2,k𝐙+}A_{i}=\{c<0|\lambda_{i}(c)=-(k\alpha)^{2},\ k\in\mathbf{Z}^{+}\}. Then by (3)(3) and the expression of λi(0)\lambda_{i}(0), i=1,2i=1,2, we have (A1)=π1r2α1\sharp(A_{1})=\lceil\frac{{\pi\sqrt{1-r^{2}}}}{\alpha}\rceil-1, (A2)=πr2r+34α1\sharp(A_{2})=\lceil\frac{\pi\sqrt{-r^{2}-r+{3\over 4}}}{\alpha}\rceil-1, and (A1)θ(A1)+(A2)\sharp(A_{1})\leq\theta\leq\sharp(A_{1})+\sharp(A_{2}), i.e. π1r2α1θπ1r2α+πr2r+34α2\lceil\frac{{\pi\sqrt{1-r^{2}}}}{\alpha}\rceil-1\leq\theta\leq\lceil\frac{{\pi\sqrt{1-r^{2}}}}{\alpha}\rceil+\lceil\frac{\pi\sqrt{-r^{2}-r+{3\over 4}}}{\alpha}\rceil-2. In fact, θ=(A1)+(A2)(A1A2)\theta=\sharp(A_{1})+\sharp(A_{2})-\sharp(A_{1}\cap A_{2}), but it seems difficult to give an explicit formula if A1A2.A_{1}\cap A_{2}\neq\emptyset.

Now, we prove (1)(1)(10)(10). (1)(1) and (4)(4)(7)(7) are a summary of spectral results in Section 44 of [28]. Monotonicity of λn\lambda_{n} for β(,12π2][12π2,)\beta\in(-\infty,-{1\over 2}\pi^{2}]\cup[{1\over 2}\pi^{2},\infty) is due to Corollary 11 in [28]. Asymptotic behavior of λn\lambda_{n} in (2)(2) and (10)(10) is obtained by Corollary 2.10. Signatures of λn\lambda_{n} in (3)(3) and (8)(8)(9)(9) are due to Proposition 11 in [28] and simple computation.

The rest is to prove the asymptotic behavior of λn\lambda_{n} in (3)(3) and (8)(8)(9)(9). First, we consider β=±12π2\beta=\pm{1\over 2}\pi^{2}. We only prove that limc0λn(c)=λn(0)\lim_{c\to 0^{-}}\lambda_{n}(c)=\lambda_{n}(0) for β=12π2\beta={1\over 2}\pi^{2} and n1n\geq 1. Note that βuu=π2{\beta-u^{\prime\prime}\over u}=\pi^{2} and βuucβuuL1(1,1)=π2cucL1(1,1)\left\|{\beta-u^{\prime\prime}\over u-c}-{\beta-u^{\prime\prime}\over u}\right\|_{L^{1}(-1,1)}=\pi^{2}\left\|{c\over u-c}\right\|_{L^{1}(-1,1)} for c<0c<0. Let 0<δ<10<\delta<1. Then

cucL1(1δ,1)=2π0cos(π2(1δ))cz2c11z2dzCδc0\displaystyle\left\|{c\over u-c}\right\|_{L^{1}(1-\delta,1)}={2\over\pi}\int_{0}^{\cos({\pi\over 2}(1-\delta))}{-c\over z^{2}-c}{1\over\sqrt{1-z^{2}}}dz\leq C_{\delta}\sqrt{-c}\to 0

as c0c\to 0^{-}. Similarly, cucL1(1,1+δ)0\left\|{c\over u-c}\right\|_{L^{1}(-1,-1+\delta)}\to 0. Clearly, cucL1(1+δ,1δ)0\left\|{c\over u-c}\right\|_{L^{1}(-1+\delta,1-\delta)}\to 0. Then cucL1(1,1)0\left\|{c\over u-c}\right\|_{L^{1}(-1,1)}\to 0. It follows from Theorem 2.12.1 in [22] that limc0λn(c)=λn(0)\lim_{c\to 0^{-}}\lambda_{n}(c)=\lambda_{n}(0) for n1n\geq 1.

We then consider β(12π2,916π2)\beta\in({1\over 2}\pi^{2},{9\over 16}\pi^{2}). We use the eigenfunctions of λn(β,0)\lambda_{n}(\beta,0) in Proposition 11 (iii)(\rm{iii}) of [28]. Here, we rewrite λn(β,c)=λn(c)\lambda_{n}(\beta,c)=\lambda_{n}(c) to indicate its dependence on β\beta if necessary. There exist ϕn(y)=ϕn(β,0)(y)=cos2r(π2y)Pn1(sin(π2y))\phi_{n}(y)=\phi_{n}^{(\beta,0)}(y)=\cos^{2r}(\frac{\pi}{2}y)P_{n-1}(\sin(\frac{\pi}{2}y)), n1n\geq 1, satisfying

ϕnβuuϕn=λn(0)ϕnon(1,1),ϕn(±1)=0.\displaystyle-\phi_{n}^{\prime\prime}-\frac{\beta-u^{\prime\prime}}{u}\phi_{n}=\lambda_{n}(0)\phi_{n}\quad\text{on}\quad(-1,1),\quad\phi_{n}(\pm 1)=0.

Here, λn(β,0)=((r+n12)21)π2\lambda_{n}(\beta,0)=\left(\left(r+{n-1\over 2}\right)^{2}-1\right)\pi^{2}, r=14+916βπ2r={1\over 4}+\sqrt{{9\over 16}-{\beta\over\pi^{2}}} and Pn1()P_{n-1}(\cdot) is a polynomial with order n1n-1. Moreover, ϕnH01(1,1)\phi_{n}\in H_{0}^{1}(-1,1) is real-valued, and we normalize it such that ϕnL2(1,1)=1.\|\phi_{n}\|_{L^{2}(-1,1)}=1. Then we have for m,n1m,n\geq 1,

11ϕnϕmdy=δmn={1ifn=m,0ifnm,11(ϕnϕm+uβuϕnϕm)dy=λn(0)δmn.\displaystyle\int_{-1}^{1}\phi_{n}\phi_{m}dy=\delta_{mn}=\left\{\begin{array}[]{l@{\ \text{if}\ }l}1&n=m,\\ 0&n\neq m,\end{array}\right.\quad\int_{-1}^{1}\left(\phi_{n}^{\prime}\phi_{m}^{\prime}+{u^{\prime\prime}-\beta\over u}\phi_{n}\phi_{m}\right)dy=\lambda_{n}(0)\delta_{mn}.

Note that ϕn\phi_{n} has n1n-1 zeros in (1,1)(-1,1), and we denote Zn:={y(1,1):ϕn(y)=0}={an,1,,an,n1}Z_{n}:=\{y\in(-1,1):\phi_{n}(y)=0\}=\{a_{n,1},\cdots,a_{n,n-1}\}. For any nn-dimensional subspace V=span{ψ1,,ψn}V=\text{span}\{\psi_{1},\cdots,\psi_{n}\} in H01(1,1)H_{0}^{1}(-1,1), there exists 0(ξ1,,ξn)𝐑n0\neq(\xi_{1},\cdots,\xi_{n})\in\mathbf{R}^{n} such that ξ1ψ1(an,i)++ξnψn(an,i)=0,i=1,,n1.\xi_{1}\psi_{1}(a_{n,i})+\cdots+\xi_{n}\psi_{n}(a_{n,i})=0,i=1,\cdots,n-1. Define ψ~=ξ1ψ1++ξnψn.\tilde{\psi}=\xi_{1}\psi_{1}+\cdots+\xi_{n}\psi_{n}. Then ψ~(an,i)=0\tilde{\psi}(a_{n,i})=0, i=1,,n1i=1,\cdots,n-1, i.e. ψ~|Zn=0.\tilde{\psi}|_{Z_{n}}=0. We normalize ψ~\tilde{\psi} such that ψ~L2(1,1)=1\|\tilde{\psi}\|_{L^{2}(-1,1)}=1. Since ψ~H01(1,1)\tilde{\psi}\in H_{0}^{1}(-1,1), we have ψ~(±1)=0\tilde{\psi}(\pm 1)=0. Similar to (3.2), we have |ψ~(y)|2ϕn(y)/ϕn(y)0|\tilde{\psi}(y)|^{2}\phi_{n}^{\prime}(y)/\phi_{n}(y)\to 0 as yan,iy\to a_{n,i} or y1+y\to-1^{+} or y1y\to 1^{-}, where 1in11\leq i\leq n-1. Integration by parts gives

ψ~ψ~ϕnϕnL2(1,1)2=11(|ψ~|2+ϕnϕn|ψ~|2)dy=11(|ψ~|2βuu|ψ~|2λn(0)|ψ~|2)dy.\displaystyle\left\|\tilde{\psi}^{\prime}-\tilde{\psi}{\phi_{n}^{\prime}\over\phi_{n}}\right\|_{L^{2}(-1,1)}^{2}=\int_{-1}^{1}\left(|\tilde{\psi}^{\prime}|^{2}+{\phi_{n}^{\prime\prime}\over\phi_{n}}|\tilde{\psi}|^{2}\right)dy=\int_{-1}^{1}\left(|\tilde{\psi}^{\prime}|^{2}-\frac{\beta-u^{\prime\prime}}{u}|\tilde{\psi}|^{2}-\lambda_{n}(0)|\tilde{\psi}|^{2}\right)dy.

If c<0c<0, then using βu>0\beta-u^{\prime\prime}>0 and uc>0u-c>0 for y[1,1]y\in[-1,1], we have

11(|ψ~|2βuuc|ψ~|2)dy11(|ψ~|2βuu|ψ~|2)dy11λn(0)|ψ~|2dy=λn(0).\displaystyle\int_{-1}^{1}\left(|\tilde{\psi}^{\prime}|^{2}-\frac{\beta-u^{\prime\prime}}{u-c}|\tilde{\psi}|^{2}\right)dy\geq\int_{-1}^{1}\left(|\tilde{\psi}^{\prime}|^{2}-\frac{\beta-u^{\prime\prime}}{u}|\tilde{\psi}|^{2}\right)dy\geq\int_{-1}^{1}\lambda_{n}(0)|\tilde{\psi}|^{2}dy=\lambda_{n}(0).

This, along with (2.7), yields that infc(,0)λn(c)λn(0).\inf_{c\in(-\infty,0)}\lambda_{n}(c)\geq\lambda_{n}(0). Now, we consider the upper bound. Let Vn=span{ϕ1,,ϕn}.V_{n}=\text{span}\{\phi_{1},\cdots,\phi_{n}\}. Then VnH01(1,1)V_{n}\subset H_{0}^{1}(-1,1). By (2.7), there exist bi,c𝐑b_{i,c}\in\mathbf{R}, i=1,,ni=1,\cdots,n, with i=1n|bi,c|2=1\sum_{i=1}^{n}|b_{i,c}|^{2}=1 such that φc=i=1nbi,cϕiVn\varphi_{c}=\sum_{i=1}^{n}b_{i,c}\phi_{i}\in V_{n} with φc2L2=1\|\varphi_{c}\|^{2}_{L^{2}}=1, and

λn(c)\displaystyle\lambda_{n}(c)\leq supϕL2=1,ϕVn11(|ϕ|2+uβuc|ϕ|2)dy=11(|φc|2+uβuc|φc|2)dy\displaystyle\sup_{\|\phi\|_{L^{2}}=1,\phi\in V_{n}}\int_{-1}^{1}\left(|\phi^{\prime}|^{2}+{u^{\prime\prime}-\beta\over u-c}|\phi|^{2}\right)dy=\int_{-1}^{1}\left(|\varphi_{c}^{\prime}|^{2}+{u^{\prime\prime}-\beta\over u-c}|\varphi_{c}|^{2}\right)dy
=\displaystyle= i=1n|bi,c|211(|ϕi|2+uβuc|ϕi|2)dy+1i<jn2bi,cbj,c11(ϕiϕj+uβucϕiϕj)dy\displaystyle\sum_{i=1}^{n}|b_{i,c}|^{2}\int_{-1}^{1}\left(|\phi_{i}^{\prime}|^{2}+{u^{\prime\prime}-\beta\over u-c}|\phi_{i}|^{2}\right)dy+\!\!\!\!\sum_{1\leq i<j\leq n}\!\!\!2b_{i,c}b_{j,c}\int_{-1}^{1}\left(\phi_{i}^{\prime}\phi_{j}^{\prime}+{u^{\prime\prime}-\beta\over u-c}\phi_{i}\phi_{j}\right)dy
\displaystyle\leq max1in11(|ϕi|2+uβuc|ϕi|2)dy+1i<jn|11(ϕiϕj+uβucϕiϕj)dy|\displaystyle\max_{1\leq i\leq n}\int_{-1}^{1}\left(|\phi_{i}^{\prime}|^{2}+{u^{\prime\prime}-\beta\over u-c}|\phi_{i}|^{2}\right)dy+\!\!\!\!\sum_{1\leq i<j\leq n}\!\left|\int_{-1}^{1}\left(\phi_{i}^{\prime}\phi_{j}^{\prime}+{u^{\prime\prime}-\beta\over u-c}\phi_{i}\phi_{j}\right)dy\right|
\displaystyle\to max1in11(|ϕi|2+uβu|ϕi|2)dy+1i<jn|11(ϕiϕj+uβuϕiϕj)dy|\displaystyle\max_{1\leq i\leq n}\int_{-1}^{1}\left(|\phi_{i}^{\prime}|^{2}+{u^{\prime\prime}-\beta\over u}|\phi_{i}|^{2}\right)dy+\!\!\!\!\sum_{1\leq i<j\leq n}\!\left|\int_{-1}^{1}\left(\phi_{i}^{\prime}\phi_{j}^{\prime}+{u^{\prime\prime}-\beta\over u}\phi_{i}\phi_{j}\right)dy\right|
=\displaystyle= max1inλi(0)+1i<jn0=λn(0),asc0.\displaystyle\max_{1\leq i\leq n}\lambda_{i}(0)+\!\!\sum_{1\leq i<j\leq n}0=\lambda_{n}(0),\;\text{as}\;\;c\to 0^{-}.

Combining the upper and lower bounds, we have limc0λn(c)=λn(0)\lim_{c\to 0^{-}}\lambda_{n}(c)=\lambda_{n}(0) for β(12π2,916π2).\beta\in({1\over 2}\pi^{2},{9\over 16}\pi^{2}).

Now, we consider β=916π2\beta={9\over 16}\pi^{2}. By Corollary 11 (i)(\rm{i}) in [28], we have for fixed c<0c<0, λn(β,c)λn(β,c)\lambda_{n}(\beta,c)\leq\lambda_{n}(\beta^{\prime},c) if β<β\beta^{\prime}<\beta. As λn(β,c)λn(β,c)\lambda_{n}(\beta^{\prime},c)\geq\lambda_{n}(\beta^{\prime},c^{\prime}) if c<c<0c<c^{\prime}<0 (see Corollary 11 (iv)(\rm{iv}) in [28]) and limc0λn(β,c)=λn(β,0)\lim_{c\to 0^{-}}\lambda_{n}(\beta^{\prime},c)=\lambda_{n}(\beta^{\prime},0), we have for fixed β(12π2,916π2)\beta^{\prime}\in({1\over 2}\pi^{2},{9\over 16}\pi^{2}), λn(β,c)λn(β,0)\lambda_{n}(\beta^{\prime},c)\geq\lambda_{n}(\beta^{\prime},0) if c<0c<0. Then

limc0λn(β,c)lim infββlimc0λn(β,c)=lim infββλn(β,0)=λn(β,0),\displaystyle\lim_{c\to 0^{-}}\lambda_{n}(\beta,c)\leq\liminf_{\beta^{\prime}\to\beta^{-}}\lim_{c\to 0^{-}}\lambda_{n}(\beta^{\prime},c)=\liminf_{\beta^{\prime}\to\beta^{-}}\lambda_{n}(\beta^{\prime},0)=\lambda_{n}(\beta,0),
limc0λn(β,c)=limc0limββλn(β,c)limββλn(β,0)=λn(β,0).\displaystyle\lim_{c\to 0^{-}}\lambda_{n}(\beta,c)=\lim_{c\to 0^{-}}\lim_{\beta^{\prime}\to\beta^{-}}\lambda_{n}(\beta^{\prime},c)\geq\lim_{\beta^{\prime}\to\beta^{-}}\lambda_{n}(\beta^{\prime},0)=\lambda_{n}(\beta,0).

Here, we used the left continuity of λn(,0)\lambda_{n}(\cdot,0) at β=916π2.\beta={9\over 16}\pi^{2}. Thus, limc0λn(β,c)=λn(β,0).\lim_{c\to 0^{-}}\lambda_{n}(\beta,c)=\lambda_{n}(\beta,0).

Next, we consider β(916π2,12π2)\beta\in(-\frac{9}{16}\pi^{2},-\frac{1}{2}\pi^{2}). By Proposition 11 (iv)(\rm{iv}) in [28], there exists ϕn(y)=ϕn(β,1)(y)=|sin(π2y)|2rPn(cos(π2y))\phi_{n}(y)=\phi_{n}^{(\beta,1)}(y)=|\sin(\frac{\pi}{2}y)|^{2r}P_{n}(\cos(\frac{\pi}{2}y)) if nn is odd; there exists ϕn(y)=ϕn(β,1)(y)=sign(y)\phi_{n}(y)=\phi_{n}^{(\beta,1)}(y)=\text{sign}(y) |sin(π2y)|2rPn1(cos(π2y))|\sin(\frac{\pi}{2}y)|^{2r}P_{n-1}(\cos(\frac{\pi}{2}y)) if nn is even; and for n1n\geq 1,

ϕnβuu1ϕn=λn(1)ϕnon(1,1){0},ϕn(±1)=0.\displaystyle-\phi_{n}^{\prime\prime}-\frac{\beta-u^{\prime\prime}}{u-1}\phi_{n}=\lambda_{n}(1)\phi_{n}\quad\text{on}\quad(-1,1)\setminus\{0\},\quad\phi_{n}(\pm 1)=0.

Here, λn(β,1)=((r12+n2)21)π2\lambda_{n}(\beta,1)=\left(\left(r-{1\over 2}+\lceil{n\over 2}\rceil\right)^{2}-1\right)\pi^{2} and r=14+916+βπ2r={1\over 4}+\sqrt{{9\over 16}+{\beta\over\pi^{2}}}. Moreover, ϕnH01(1,1)\phi_{n}\in H_{0}^{1}(-1,1) is real-valued and we normalize it such that ϕnL2(1,1)=1.\|\phi_{n}\|_{L^{2}(-1,1)}=1. Then for m,n1m,n\geq 1,

11ϕnϕmdy=δmn,11(ϕnϕm+uβu1ϕnϕm)dy=λn(1)δmn.\displaystyle\int_{-1}^{1}\phi_{n}\phi_{m}dy=\delta_{mn},\quad\int_{-1}^{1}\left(\phi_{n}^{\prime}\phi_{m}^{\prime}+{u^{\prime\prime}-\beta\over u-1}\phi_{n}\phi_{m}\right)dy=\lambda_{n}(1)\delta_{mn}.

If n1n\geq 1 is odd, then ϕn\phi_{n} has nn zeros in (1,1)(-1,1), and 0Zn={y(1,1):ϕn(y)=0}={an,1,,an,n}0\in Z_{n}=\{y\in(-1,1):\phi_{n}(y)=0\}=\{a_{n,1},\cdots,a_{n,n}\}. For any (n+1)(n+1)-dimensional subspace V=span{ψ1,,ψn+1}V=\text{span}\{\psi_{1},\cdots,\psi_{n+1}\} in H01(1,1)H_{0}^{1}(-1,1), there exists 0(ξ1,,ξn+1)𝐑n+10\neq(\xi_{1},\cdots,\xi_{n+1})\in\mathbf{R}^{n+1} such that ξ1ψ1(an,i)++ξn+1ψn+1(an,i)=0,i=1,,n.\xi_{1}\psi_{1}(a_{n,i})+\cdots+\xi_{n+1}\psi_{n+1}(a_{n,i})=0,i=1,\cdots,n. Define ψ~=ξ1ψ1++ξn+1ψn+1.\tilde{\psi}=\xi_{1}\psi_{1}+\cdots+\xi_{n+1}\psi_{n+1}. Then ψ~(an,i)=0\tilde{\psi}(a_{n,i})=0, i=1,,ni=1,\cdots,n, i.e. ψ~|Zn=0.\tilde{\psi}|_{Z_{n}}=0. We normalize ψ~\tilde{\psi} such that ψ~L2(1,1)=1\|\tilde{\psi}\|_{L^{2}(-1,1)}=1. Since ψ~H01(1,1)\tilde{\psi}\in H_{0}^{1}(-1,1), we have ψ~(±1)=0\tilde{\psi}(\pm 1)=0. Integration by parts gives

ψ~ψ~ϕnϕnL2(1,1)2=11(|ψ~|2+ϕnϕn|ψ~|2)dy=11(|ψ~|2βuu1|ψ~|2λn(1)|ψ~|2)dy.\displaystyle\left\|\tilde{\psi}^{\prime}-\tilde{\psi}{\phi_{n}^{\prime}\over\phi_{n}}\right\|_{L^{2}(-1,1)}^{2}=\int_{-1}^{1}\left(|\tilde{\psi}^{\prime}|^{2}+{\phi_{n}^{\prime\prime}\over\phi_{n}}|\tilde{\psi}|^{2}\right)dy=\int_{-1}^{1}\left(|\tilde{\psi}^{\prime}|^{2}-\frac{\beta-u^{\prime\prime}}{u-1}|\tilde{\psi}|^{2}-\lambda_{n}(1)|\tilde{\psi}|^{2}\right)dy.

If c>1c>1, then using βu<0\beta-u^{\prime\prime}<0 and uc<0u-c<0 for y[1,1]y\in[-1,1], we have

11(|ψ~|2βuuc|ψ~|2)dy11(|ψ~|2βuu1|ψ~|2)dy11λn(1)|ψ~|2dy=λn(1).\displaystyle\int_{-1}^{1}\left(|\tilde{\psi}^{\prime}|^{2}-\frac{\beta-u^{\prime\prime}}{u-c}|\tilde{\psi}|^{2}\right)dy\geq\int_{-1}^{1}\left(|\tilde{\psi}^{\prime}|^{2}-\frac{\beta-u^{\prime\prime}}{u-1}|\tilde{\psi}|^{2}\right)dy\geq\int_{-1}^{1}\lambda_{n}(1)|\tilde{\psi}|^{2}dy=\lambda_{n}(1).

This, along with (2.7), yields that infc(1,+)λn+1(c)λn(1).\inf_{c\in(1,+\infty)}\lambda_{n+1}(c)\geq\lambda_{n}(1). If n1n\geq 1 is even, then λn+1(c)λn(c)λn1(1)=λn(1)\lambda_{n+1}(c)\geq\lambda_{n}(c)\geq\lambda_{n-1}(1)=\lambda_{n}(1). Thus, infc(1,+)λn+1(c)λn(1)\inf_{c\in(1,+\infty)}\lambda_{n+1}(c)\geq\lambda_{n}(1) is always true. Now, we consider the upper bound. Let Vn+1=span{ϕ0,ϕ1,,ϕn},V_{n+1}=\text{span}\{\phi_{0},\phi_{1},\cdots,\phi_{n}\}, here ϕ0=η|[1,1]\phi_{0}=\eta|_{[-1,1]} is defined in (3.11), and ϕ1,,ϕn\phi_{1},\cdots,\phi_{n} are L2L^{2} normalized eigenfunctions. Then Vn+1H01(1,1)V_{n+1}\subset H_{0}^{1}(-1,1). By (2.7), for c>1,c>1, there exists φcVn+1\varphi_{c}\in V_{n+1} with φc2L2=1\|\varphi_{c}\|^{2}_{L^{2}}=1, and

λn+1(c)\displaystyle\lambda_{n+1}(c)\leq supϕL2=1,ϕVn+111(|ϕ|2+uβuc|ϕ|2)dy=11(|φc|2+uβuc|φc|2)dy.\displaystyle\sup_{\|\phi\|_{L^{2}}=1,\phi\in V_{n+1}}\int_{-1}^{1}\left(|\phi^{\prime}|^{2}+{u^{\prime\prime}-\beta\over u-c}|\phi|^{2}\right)dy=\int_{-1}^{1}\left(|\varphi_{c}^{\prime}|^{2}+{u^{\prime\prime}-\beta\over u-c}|\varphi_{c}|^{2}\right)dy.

Since Vn+1H01(1,1)V_{n+1}\subset H_{0}^{1}(-1,1) is finite dimensional, there exist φ1Vn+1\varphi_{1}\in V_{n+1} and cm1+c_{m}\to 1^{+} such that φcmφ1\varphi_{c_{m}}\to\varphi_{1} in H01(1,1).H_{0}^{1}(-1,1). Then φ12L2=1,\|\varphi_{1}\|^{2}_{L^{2}}=1, 11|φcm|2dy11|φ1|2dy\int_{-1}^{1}|\varphi_{c_{m}}^{\prime}|^{2}dy\to\int_{-1}^{1}|\varphi_{1}^{\prime}|^{2}dy and

λn+1(cm)\displaystyle\lambda_{n+1}(c_{m})\leq 11(|φcm|2+uβucm|φcm|2)dy.\displaystyle\int_{-1}^{1}\left(|\varphi_{c_{m}}^{\prime}|^{2}+{u^{\prime\prime}-\beta\over u-c_{m}}|\varphi_{c_{m}}|^{2}\right)dy.

Since uβ>0u^{\prime\prime}-\beta>0 and uc<0u-c<0 for y[1,1]y\in[-1,1] and c>1c>1, by Fatou’s Lemma, we have

lim supm11uβucm|φcm|2dy11lim supmuβucm|φcm|2dy=11uβu1|φ1|2dy.\displaystyle\limsup_{m\to\infty}\int_{-1}^{1}{u^{\prime\prime}-\beta\over u-c_{m}}|\varphi_{c_{m}}|^{2}dy\leq\int_{-1}^{1}\limsup_{m\to\infty}{u^{\prime\prime}-\beta\over u-c_{m}}|\varphi_{c_{m}}|^{2}dy=\int_{-1}^{1}{u^{\prime\prime}-\beta\over u-1}|\varphi_{1}|^{2}dy.

In particular, if φ1(0)0\varphi_{1}(0)\neq 0, then

11uβu1|φ1|2dy=,lim supm11uβucm|φcm|2dy=,lim supmλn+1(cm)=.\displaystyle\int_{-1}^{1}{u^{\prime\prime}-\beta\over u-1}|\varphi_{1}|^{2}dy=-\infty,\quad\limsup_{m\to\infty}\int_{-1}^{1}{u^{\prime\prime}-\beta\over u-c_{m}}|\varphi_{c_{m}}|^{2}dy=-\infty,\quad\limsup_{m\to\infty}\lambda_{n+1}(c_{m})=-\infty.

If φ1(0)=0\varphi_{1}(0)=0, then φ1span{ϕ1,,ϕn}\varphi_{1}\in\text{span}\{\phi_{1},\cdots,\phi_{n}\} and

lim supmλn+1(cm)11(|φ1|2+uβu1|φ1|2)dy.\displaystyle\limsup_{m\to\infty}\lambda_{n+1}(c_{m})\leq\int_{-1}^{1}\left(|\varphi_{1}^{\prime}|^{2}+{u^{\prime\prime}-\beta\over u-1}|\varphi_{1}|^{2}\right)dy.

As φ12L2=1,\|\varphi_{1}\|^{2}_{L^{2}}=1, there exist bi𝐑b_{i}\in\mathbf{R}, i=1,,ni=1,\cdots,n, with i=1n|bi|2=1\sum_{i=1}^{n}|b_{i}|^{2}=1 such that φ1=i=1nbiϕiVn+1\varphi_{1}=\sum_{i=1}^{n}b_{i}\phi_{i}\in V_{n+1} and

11(|φ1|2+uβu1|φ1|2)dy\displaystyle\int_{-1}^{1}\left(|\varphi_{1}^{\prime}|^{2}+{u^{\prime\prime}-\beta\over u-1}|\varphi_{1}|^{2}\right)dy =i=1n|bi|211(|ϕi|2+uβu1|ϕi|2)dy=i=1n|bi|2λi(1)\displaystyle=\sum_{i=1}^{n}|b_{i}|^{2}\int_{-1}^{1}\left(|\phi_{i}^{\prime}|^{2}+{u^{\prime\prime}-\beta\over u-1}|\phi_{i}|^{2}\right)dy=\sum_{i=1}^{n}|b_{i}|^{2}\lambda_{i}(1)
max1inλi(1)=λn(1).\displaystyle\leq\max_{1\leq i\leq n}\lambda_{i}(1)=\lambda_{n}(1).

Therefore, if φ1(0)=0\varphi_{1}(0)=0, then lim supmλn+1(cm)λn(1);\limsup\limits_{m\to\infty}\lambda_{n+1}(c_{m})\leq\lambda_{n}(1); if φ1(0)0\varphi_{1}(0)\neq 0, this is clearly true since the limit is -\infty in this case. By monotonicity of λn\lambda_{n}, we have limc1+λn+1(c)λn(1).\lim\limits_{c\to 1^{+}}\lambda_{n+1}(c)\leq\lambda_{n}(1). Combining the upper and lower bounds, we have limc1+λn+1(c)=λn(1)\lim\limits_{c\to 1^{+}}\lambda_{n+1}(c)=\lambda_{n}(1) for β(916π2,12π2)\beta\in(-\frac{9}{16}\pi^{2},-\frac{1}{2}\pi^{2}).

For β=916π2\beta=-\frac{9}{16}\pi^{2}, the limits limc1+λn+1(c)=λn(1)\lim\limits_{c\to 1^{+}}\lambda_{n+1}(c)=\lambda_{n}(1), n1n\geq 1, can be proved similarly as in the case β=916π2.\beta=\frac{9}{16}\pi^{2}. Finally, the limit limc1+λ1(c)=\lim_{c\to 1^{+}}\lambda_{1}(c)=-\infty for β[916π2,12π2)\beta\in[-\frac{9}{16}\pi^{2},-\frac{1}{2}\pi^{2}) follows from Theorem 2.9 (2)(2).

Acknowledgement

Z. Lin is partially supported by the NSF grants DMS-1715201 and DMS-2007457. Z. Zhang is partially supported by NSF of China under Grant 11425103.

References

  • [1] A. Barcilon, P. G. Drazin, Nonlinear waves of vorticity, Stud. Appl. Math., 106 (2001), 437-479.
  • [2] N. J. Balmforth, C. Piccolo, The onset of meandering in a barotropic jet, J. Fluid Mech., 449 (2001), 85-114.
  • [3] J. Bedrossian, N. Masmoudi, Inviscid damping and the asymptotic stability of planar shear flows in the 2D Euler equations, Publ. Math. Inst. Hautes Études Sci., 122 (2015), 195-300.
  • [4] M. Buchanan, J. J. Dorning, Superposition of nonlinear plasma waves, Phys. Rev. Lett., 70 (1993), 3732-3735.
  • [5] M. Buchanan, J. J. Dorning, Near equilibrium multiple-wave plasma states, Phys. Rev. E, 50 (1994), 1465-1478.
  • [6] K.M. Case, Stability of inviscid plane Couette flow, Phys. Fluids 3, (1960), 143-148.
  • [7] M. Crandall, P. Rabinowitz, Bifurcation from simple eigenvalues, J. Funct. Anal., 8 (1971), 321-340.
  • [8] M. Coti Zelati, T. M. Elgindi, K. Widmayer, Stationary Structures near the Kolmogorov and Poiseuille Flows in the 2d Euler Equations, Preprint, arXiv:2007.11547.
  • [9] L. Demeio, P. F. Zweifel, Numerical simulations of perturbed Vlasov equilibria, Phys. Fluids B, 2 (1990), 1252-1255.
  • [10] P. G. Drazin, Introduction to hydrodynamic stability, Cambridge Texts in Applied Mathematics, Cambridge University Press, Cambridge, 2002.
  • [11] T. M. Elgindi, K. Widmayer, Long time stability for solutions of a β\beta-plane equation, Comm. Pure Appl. Math., 70 (2017), 1425-1471.
  • [12] L. Engevik, A note on the barotropic instability of the Bickley jet, J. Fluid Mech., 499 (2004), 315-326.
  • [13] R. Fjørtoft, Application of integral theorems in deriving criteria of stability of laminar flow and for baroclinic circular vortex, Geofys. Publ. Norske Vid.-Akad. Oslo, 17 (1950), 1-52.
  • [14] E. Grenier, T.T. Nguyen, F. Rousset and A. Soffer, Linear inviscid damping and enhanced viscous dissipation of shear flows by using the conjugate operator method, J. Funct. Anal., 278 (2020), 108339.
  • [15] G. H. Hardy, J. E. Littlewood, G. Pólya, Inequalities, Cambridge Mathematical Library, Cambridge University Press,Cambridge (1988); reprint of the 1952 edition.
  • [16] L. N. Howard, P. G. Drazin, On instability of a parallel flow of inviscid fluid in a rotating system with variable Coriolis parameter, J. Math. Phys., 43 (1964), 83-99.
  • [17] A. Ionescu, H. Jia, Inviscid damping near the Couette flow in a channel, Comm. Math. Phys., 374 (2020), 2015-2096.
  • [18] A. Ionescu, H. Jia, Nonlinear inviscid damping near monotonic shear flows, Preprint, arXiv:2001.03087.
  • [19] H. Jia, Linear inviscid damping in Gevrey spaces, Arch. Ration. Mech. Anal., 235 (2020), 1327-1355.
  • [20] H. Jia, Linear inviscid damping near monotone shear flows, SIAM J. Math. Anal., 52 (2020), 623-652.
  • [21] T. Kato, Perturbation theory for linear operators, second edition, Springer-Verlag, Heidelberg, 1980.
  • [22] Q. Kong, H. Wu, A. Zettl, Dependence of the nn-th Sturm-Liouville eigenvalue on the problem, J. Differential Equations, 156 (1999), 328-354.
  • [23] H. L. Kuo, Dynamic instability of two-dimensional non-divergent flow in a barotropic atmosphere, J. Meteor., 6 (1949), 105-122.
  • [24] H. L. Kuo, Dynamics of quasi-geostrophic flows and instability theory, Adv. Appl. Mech., 13 (1974), 247-330.
  • [25] C. Lancellotti, J. J. Dorning, Nonlinear Landau damping, Transport Theor. Stat., 38 (2009), 1-146.
  • [26] C. Y. Li, Z. Lin, A resolution of the Sommerfeld Paradox, SIAM J. Math. Anal., 43 (2011), 1923-1954.
  • [27] Z. Lin, Instability of some ideal plane flows, SIAM J. Math. Anal., 35 (2003), 318-356.
  • [28] Z. Lin, J. Yang, H. Zhu, Barotropic instability of shear flows, Stud. Appl. Math., 144 (2020), 289-326.
  • [29] Z. Lin, C. Zeng, Inviscid dynamic structures near Couette flow, Arch. Rat. Mech. Anal., 200 (2011), 1075-1097.
  • [30] F. B. Lipps, The barotropic stability of the mean winds in the atmosphere, J. Fluid Mech., 12 (1962), 397-407.
  • [31] A. Majda, Introduction to PDEs and waves for the atmosphere and ocean, Courant Lecture Notes in Mathematics, vol. 9, New York University, Courant Institute of Mathematical Sciences, New York; American Mathematical Society, Providence, RI, 2003.
  • [32] S. A. Maslowe, Barotropic instability of the Bickley jet, J. Fluid Mech., 229 (1991), 417-426.
  • [33] N. Masmoudi, About the Hardy inequality. In: Schleicher D., Lackmann M. (eds) An Invitation to Mathematics. Springer, Berlin, Heidelberg.
  • [34] N. Masmoudi, W. Zhao, Nonlinear inviscid damping for a class of monotone shear flows in finite channel, Preprint, arXiv:2001.08564.
  • [35] J. C. McWilliams, Fundamentals of geophysical fluid dynamics, Cambridge University Press, 2006.
  • [36] J. Pedlosky, Geophysical fluid dynamics, 2nd edn. Springer, New York (1987).
  • [37] F. Pusateri, K. Widmayer, On the global stability of a β\beta-plane equation, Anal. PDE, 11 (2018), 1587-1624.
  • [38] W. Orr, Stability and instability of steady motions of a perfect liquid, Proc. Ir. Acad. Sect. A: Math Astron. Phys. Sci., 27(1907), 9-66.
  • [39] L. Rayleigh, On the stability or instability of certain fluid motions, Proc. London Math. Soc., 9 (1880), 57-70.
  • [40] M. Reed, B. Simon, Methods of modern mathematical physics. IV. Analysis of operators. Academic Press [Harcourt Brace Jovanovich, Publishers], New York-London, 1978. xv+396 pp.
  • [41] S. I. Rosencrans, D. H. Sattinger, On the spectrum of an operator occuring in the theory of hydrodynamic stability, J. Math. Phys., 45 (1966), 289-300.
  • [42] C. G. Rossby, Relation between variations in the intensity of the zonal circulation of the atmosphere and the displacements of the semi-permanent centers of action, J. Mar. Res., 2 (1939), 38-55.
  • [43] K. K. Tung, Barotropic instability of zonal flows, J. Atmos. Sci., 38 (1981), 308-321.
  • [44] D. Wei, Z. Zhang and W. Zhao, Linear inviscid damping for a class of monotone shear flow in Sobolev spaces, Commun. Pure Appl. Math., 71 (2018), 617-687.
  • [45] D. Wei, Z. Zhang and W. Zhao, Linear inviscid damping and vorticity depletion for shear flows, Ann. PDE, 5 (2019), 3, 101 pp.
  • [46] D. Wei, Z. Zhang and W. Zhao, Linear inviscid damping and enhanced dissipation for the Kolmogorov flow, Adv. Math., 362 (2020), 106963, 103 pp.
  • [47] D. Wei, Z. Zhang, H. Zhu, Linear inviscid damping for the β\beta-plane equation, Comm. Math. Phys., 375 (2020), 127-174.
  • [48] C. Zillinger, Linear inviscid damping for monotone shear flows, Trans. Am. Math. Soc., 369 (2017), 8799-8855.
  • [49] C. Zillinger, Linear inviscid damping for monotone shear flows in a finite periodic channel, boundary effects, blow-up and critical Sobolev regularity, Arch. Ration. Mech. Anal., 221 (2016), 1449-1509.