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

Index theory for traveling waves in reaction diffusion systems with skew gradient structure

Qin Xing School of Mathematics, Shandong University, Jinan, Shandong, The People’s Republic of China [email protected]
Abstract.

A unified geometric approach for the stability analysis of traveling pulse solutions for reaction diffusion equations with skew-gradient structure has been established in a previous paper [9], but essentially no results have been found in the case of traveling front solutions. In this work, we will bridge this gap. For such cases, a Maslov index of the traveling wave is well-defined, and we will show how it can be used to provide the spectral information of the waves. As an application, we use the same index providing the exact number of unstable eigenvalues of the traveling front solutions of FitzHugh-Nagumo equations.

Keywords: heteroclinic orbits; stability of traveling wave; Maslov index; spectral flow; FitzHugh–Nagumo equations

Partially supported by NSFC No.11425105

1. Introduction and main results

The setting of this paper is the systems of reaction-diffusion equations of the form

wt=wxx+QF(w),w_{t}=w_{xx}+Q\nabla F(w), (1.1)

where x,tx,t\in\mathbb{R} are space and time, respectively, wnw\in\mathbb{R}^{n} and F\nabla F are the gradients of the function F:nF:\mathbb{R}^{n}\rightarrow\mathbb{R}, and QMat(n,)Q\in\mathrm{Mat}\,(n,\mathbb{R}) has the form Q=[Ir00Inr]Q=\begin{bmatrix}I_{r}&0\\ 0&-I_{n-r}\end{bmatrix}, where Mat(n,)\mathrm{Mat}\,(n,\mathbb{R}) is the set of all n×nn\times n matrices. Such a system is of the activator-inhibitor type and will be referred to as a skew-gradient system [18]. A traveling wave solution ww^{*} of one variable ξ=xct\xi=x-ct is a solution to (1.1) with asymptotic behavior

w± as ξ±.w_{\pm}\text{ as }\xi\rightarrow\pm\infty.

Here, w±w_{\pm} are the constant equilibria of (1.1): that is, F(w±)=0\nabla F(w_{\pm})=0. If w+=ww_{+}=w_{-}, then ww^{*} is called a pulse (and a front otherwise). In this paper, we are focused primarily on fronts. We will always consider the case of c>0c>0: otherwise, the direction of motion can be reversed.

Traveling waves being solutions of (1.1) is an important subjection in dynamical systems, and many aspects of traveling waves have been studied in extensive works[11, 6]. One of the key issues of concern is whether a steady state is stable with respect to disturbances under initial conditions since this directly determines whether it can be observed in nature. Motivated by [9], in this paper, we establish a unified geometric approach for the stability analysis of traveling front solutions for (1.1). Written in a moving frame, a traveling front solution ww^{*} of (1.1) can be regarded as a heteroclinic solution ww^{*} of the following equation:

{wξξ+cwξ+QF(w)=0,limξ±w(ξ)=w±.\begin{cases}w_{\xi\xi}+cw_{\xi}+Q\nabla F(w)&=0,\\ \lim_{\xi\rightarrow\pm\infty}w(\xi)=w_{\pm}.\end{cases} (1.2)

The stability analysis is directly related to the spectral information of the operator L:=d2dξ2+cddξ+QB(ξ)L:=\frac{\operatorname{d}^{2}}{\operatorname{d}\xi^{2}}+c\frac{\operatorname{d}}{\operatorname{d}\xi}+QB(\xi) by linearizing (1.2) along ww^{*}, where B(ξ)=2F(w)B(\xi)=\nabla^{2}F(w^{*}); note that the matrices B±limξ±B(ξ)B_{\pm}\coloneqq\lim\limits_{\xi\to\pm\infty}B(\xi) are well-defined. Moreover, the previous description guarantees that there exists C>C>, such that

QB(ξ)v,vC|v|2 for all (ξ,v)×n.\displaystyle\left\langle QB(\xi)v,v\right\rangle\leqslant C|v|^{2}\text{ for all }(\xi,v)\in\mathbb{R}\times\mathbb{R}^{n}. (1.3)

For zz\in\mathbb{C}, denote by z\Re z and z\mathrm{\mathfrak{I}}z the real and imaginary parts of zz, respectively. +(0,+)\mathbb{R}^{+}\coloneqq(0,+\infty), (,0)\mathbb{R}^{-}\coloneqq(-\infty,0), +{z|z>0}\mathbb{C}^{+}\coloneqq\{z\in\mathbb{C}|\Re z>0\} and {z|z<0}\mathbb{C}^{-}\coloneqq\{z\in\mathbb{C}|\Re z<0\}. The pair (n,,)(\mathbb{R}^{n},\langle,\rangle) denotes the nn-dimensional Euclidean space. Moreover, #¯\overline{\#} denotes the closure of the set #\#.

Since a wave solution of (1.1) possesses a translation invariance property, it is said to be nondegenerate if zero is a simple eigenvalue of LL.

Definition 1.1.

A nondegenerate wave solution of (1.1) is spectrally stable if all the nonzero eigenvalues of LL are in \mathbb{C}^{-}.

In this paper, we study the following special case.

(H1).

Suppose that σ(QB±)\sigma\left(QB_{\pm}\right)\subset\mathbb{C}^{-}.

Thus, w+w_{+} and ww_{-} are both stable equilibria of (1.1).

From now on, we focus on studying the eigenvalue problem

Lϕ=λϕ.L\phi=\lambda\phi. (1.4)

Denoted by σp(L)\sigma_{p}(L) is the set of isolated eigenvalues with finite multiplicity, and σess(L)=σ(L)\σp(L)\sigma_{ess}(L)=\sigma(L)\backslash\sigma_{p}(L) is the essential spectrum of LL. It is known (Cf. Lemma 2.3) that σess(L)\sigma_{ess}(L)\subset\mathbb{C}^{-}. As was commonly performed in [9], we set y=[ϕ˙ϕ]y=\begin{bmatrix}\dot{\phi}\\ \phi\end{bmatrix} to convert (1.4) to

y˙=Aλ(ξ)y,\dot{y}=A_{\lambda}(\xi)y, (1.5)

where Aλ(ξ)=[cλIQB(ξ)I0]A_{\lambda}(\xi)=\begin{bmatrix}-c&\lambda I-QB(\xi)\\ I&0\end{bmatrix}, and there are well-defined matrices Aλ(±)=limξ±Aλ(ξ)A_{\lambda}(\pm\infty)=\lim\limits_{\xi\to\pm\infty}A_{\lambda}(\xi). Throughout this paper, the dots denote differentiation with respect to ξ\xi.

For any MMat(,n)M\in\mathrm{Mat}\,(\mathbb{R},\mathbb{R}^{n}), denote by V+(M)V^{+}(M) (V(M)V^{-}(M)) the positive (negative) spectral space corresponding to the eigenvalues of MM having positive (negative) real parts. We provide the next two conditions:

(H2).

(QB±v,v)<0(QB_{\pm}v,v)<0, for all vV(Q)\{0}v\in V^{-}(Q)\backslash\{0\},

(H2’).

(QB±v,v)<0(QB_{\pm}v,v)<0, for all vn\{0}v\in\mathbb{R}^{n}\backslash\{0\}.

Letting BMat(n,)B\in\mathrm{Mat}\,(n,\mathbb{R}), in the following remark, we will explain the relationship between (H(H1)) and (H(H2’)).

Remark 1.2.

(1) We claim that (H(H2’)) implies that (H(H1)). Let λ\lambda be the eigenvalue of QBQB with eigenvector vv. If λ\lambda\in\mathbb{R}, then we have that λ|v|2=λv,v=QBv,v<0\lambda|v|^{2}=\lambda\left\langle v,v\right\rangle=\left\langle QBv,v\right\rangle<0, which implies that λ<0\lambda<0. If λ\\lambda\in\mathbb{C}\backslash\mathbb{R}, then λ¯\bar{\lambda} is also the eigenvalue of QBQB with eigenvector v¯\bar{v}, where λ¯\bar{\lambda} and v¯\bar{v} denote the complex conjugate of λ\lambda and vv, respectively. Now, let λ=a+bi\lambda=a+bi and v=u+iwv=u+iw: then, we have that

λ(|u|2+|w|2)=λv,v=QBv,v=QBu,uiQBu,w+iQBw,u+QBw,w.\displaystyle\lambda(|u|^{2}+|w|^{2})=\lambda\left\langle v,v\right\rangle=\left\langle QBv,v\right\rangle=\left\langle QBu,u\right\rangle-i\left\langle QBu,w\right\rangle+i\left\langle QBw,u\right\rangle+\left\langle QBw,w\right\rangle. (1.6)

Similarly, we have that

λ¯(|u|2+|w|2)=QBu,u+iQBu,wiQBw,u+QBw,w.\bar{\lambda}(|u|^{2}+|w|^{2})=\left\langle QBu,u\right\rangle+i\left\langle QBu,w\right\rangle-i\left\langle QBw,u\right\rangle+\left\langle QBw,w\right\rangle.

We thus determine that a=12(λ+λ¯)=QBu,u+QBw,w|u|2+|w|2<0.a=\frac{1}{2}(\lambda+\bar{\lambda})=\frac{\left\langle QBu,u\right\rangle+\left\langle QBw,w\right\rangle}{|u|^{2}+|w|^{2}}<0. This implies that λ\lambda has a negative real part.

(2) We remark that (H(H2’)) cannot be derived from (H(H1)). We consider BMat(2,)B\in\mathrm{Mat}\,(2,\mathbb{R}) with the form B=[13543542]B=\begin{bmatrix}1&\frac{\sqrt{35}}{4}\\ \frac{\sqrt{35}}{4}&2\end{bmatrix} and Q=[1001]Q=\begin{bmatrix}1&0\\ 0&-1\end{bmatrix}. It is easy to determine that σ(QB)\sigma(QB)\subset\mathbb{C}^{-}, but QB[10],[10]=1>0\left\langle QB\begin{bmatrix}1\\ 0\end{bmatrix},\begin{bmatrix}1\\ 0\end{bmatrix}\right\rangle=1>0.

(3) Supposing that σ(QB)\sigma(QB)\subset\mathbb{C}^{-}, we claim that there exists a nonsingular matrix TMat(n,)T\in\mathrm{Mat}\,(n,\mathbb{R}), such that

T1QBTv,v<0,vn\{0}.\langle T^{-1}QBTv,v\rangle<0,\ \forall v\in\mathbb{R}^{n}\backslash\{0\}.

We only consider a special case, where the general case is analogous. Letting QBMat(6,)QB\in\mathrm{Mat}\,(6,\mathbb{R}), suppose that λ1\lambda_{1}\in\mathbb{R} and λ2=a+ib\lambda_{2}=a+ib are generalized eigenvalues of QBQB and have the same algebraic multiplicity 2: then, for ϵ<max{λ1,a}\epsilon<-\max\{\lambda_{1},a\}, there exists TϵT_{\epsilon}, such that

Tϵ1QBTϵ=[λ1ϵ00000λ1000000abϵ000ba0ϵ0000ab0000ba].T_{\epsilon}^{-1}QBT_{\epsilon}=\begin{bmatrix}\lambda_{1}&\epsilon&0&0&0&0\\ 0&\lambda_{1}&0&0&0&0\\ 0&0&a&-b&\epsilon&0\\ 0&0&b&a&0&\epsilon\\ 0&0&0&0&a&-b\\ 0&0&0&0&b&a\end{bmatrix}.

By a simple calculation, we have that

Tϵ1QBTϵv,v\displaystyle\langle T_{\epsilon}^{-1}QBT_{\epsilon}v,v\rangle =λ1v12+λ1v22+ϵv1v2+av32+av42+av52+av62+ϵv3v5+ϵv4v6\displaystyle=\lambda_{1}v_{1}^{2}+\lambda_{1}v^{2}_{2}+\epsilon v_{1}v_{2}+av^{2}_{3}+av^{2}_{4}+av^{2}_{5}+av^{2}_{6}+\epsilon v_{3}v_{5}+\epsilon v_{4}v_{6} (1.7)
i=12(λ1+12ϵ)vi2+j=36(a+12ϵ)vj2<0.\displaystyle\leqslant\sum_{i=1}^{2}(\lambda_{1}+\frac{1}{2}\epsilon)v_{i}^{2}+\sum_{j=3}^{6}(a+\frac{1}{2}\epsilon)v^{2}_{j}<0. (1.8)

Following Definition 1.1, we focus our attention on λ+¯\lambda\in\overline{\mathbb{R}^{+}}: suppose that Φτ,λ(ξ)\Phi_{\tau,\lambda}(\xi) is the matrix solution of (1.5), such that Φτ,λ(τ)=I\Phi_{\tau,\lambda}(\tau)=I. We recall that the stable and unstable subspaces of (1.5) are

Eλs(τ):={v2n|limτ+Φτ,λ(ξ)v=0}and Eλu(τ):={v2n|limτΦτ,λ(ξ)v=0}.E_{\lambda}^{s}(\tau):=\left\{v\in\mathbb{R}^{2n}|\lim_{\tau\rightarrow+\infty}\Phi_{\tau,\lambda}(\xi)v=0\right\}\ \text{and \ }E_{\lambda}^{u}(\tau):=\left\{v\in\mathbb{R}^{2n}|\lim_{\tau\rightarrow-\infty}\Phi_{\tau,\lambda}(\xi)v=0\right\}. (1.9)

Throughout this paper, we abbreviate E0s(τ)E_{0}^{s}(\tau) and E0u(τ)E^{u}_{0}(\tau) as Es(τ)E^{s}(\tau) and Eu(τ)E^{u}(\tau), respectively.

To realize the symplectic structure, we introduce the matrix J[0QQ0]J\coloneqq\begin{bmatrix}0&-Q\\ Q&0\end{bmatrix}. Since Q2=IQ^{2}=I and QT=QQ^{T}=Q, it follows that J2=IJ^{2}=-I and JT=JJ^{T}=-J. Therefore, JJ is a complex structure on 2n\mathbb{R}^{2n}, and ω(,)J,\omega(\cdot,\cdot)\coloneqq\left\langle J\cdot,\cdot\right\rangle defines a syplectic form on 2n\mathbb{R}^{2n}. Lag(n)\mathrm{Lag}(n) denotes the set of all Lagrangian subspaces of (2n,ω)(\mathbb{R}^{2n},\omega).

By invoking [13, Lemma 3.1], Remarks 2.4 and 2.6 imply that the subspaces Eλs(τ)E^{s}_{\lambda}(\tau) and Eλu(τ)E^{u}_{\lambda}(\tau) are both Lagrangian for (τ,λ)×+¯(\tau,\lambda)\in\mathbb{R}\times\overline{\mathbb{R}^{+}}. Moreover, those results can be obtained from the same discussion in [9, Theorem 2.3].

We denote by 𝒫([0,1];2n)\mathscr{P}([0,1];\mathbb{R}^{2n}) the space of all ordered pairs of continuous maps of Lagrangian subspaces L:[0,1]tL(t)(L1(t),L2(t))Lag(n)×Lag(n)L:[0,1]\ni t\longmapsto L(t)\coloneqq\big{(}L_{1}(t),L_{2}(t)\big{)}\in\mathrm{Lag}(n)\times\mathrm{Lag}(n) equipped with the compact-open topology. Following the authors in [4], we are in the position to briefly recall the definition of the Maslov index for pairs of Lagrangian subspaces, which will be denoted throughout the paper by the symbol ιCLM\operatorname{\iota^{\scriptscriptstyle{\mathrm{CLM}}}}. Loosely speaking, given the pair L=(L1,L2)𝒫([0,1];2n)L=(L_{1},L_{2})\in\mathscr{P}([0,1];\mathbb{R}^{2n}), this index counts with signs and multiplicities the number of instants t[0,1]t\in[0,1] that L1(t)L2(t){0}L_{1}(t)\cap L_{2}(t)\neq\{0\}.

Definition 1.3.

The CLM-index is the unique integer valued function

ιCLM:𝒫([0,1];2n)LιCLM(L;[0,1])\operatorname{\iota^{\scriptscriptstyle{\mathrm{CLM}}}}:\mathscr{P}([0,1];\mathbb{R}^{2n})\ni L\longmapsto\operatorname{\iota^{\scriptscriptstyle{\mathrm{CLM}}}}(L;[0,1])\in\mathbb{Z}

satisfying Properties I-VI given in [4, Section 1].

For the sake of the reader, we list a couple of properties of the ιCLM\operatorname{\iota^{\scriptscriptstyle{\mathrm{CLM}}}}-index that we shall use throughout the paper.

  • (Reversal) Let L(L1,L2)𝒫([a,b];2n)L\coloneqq(L_{1},L_{2})\in\mathscr{P}([a,b];\mathbb{R}^{2n}). Denoting by L^𝒫([b,a];2n)\widehat{L}\in\mathscr{P}([-b,-a];\mathbb{R}^{2n}) the path traveled in the opposite direction, and by setting L^(L1(s),L2(s))\widehat{L}\coloneqq(L_{1}(-s),L_{2}(-s)), we obtain

    ιCLM(L^;[b,a])=ιCLM(L;[a,b]).\operatorname{\iota^{\scriptscriptstyle{\mathrm{CLM}}}}(\widehat{L};[-b,-a])=-\operatorname{\iota^{\scriptscriptstyle{\mathrm{CLM}}}}(L;[a,b]).
  • (Stratum homotopy relative to the ends) Given a continuous map

    L:[a,b]sL(s)𝒫([a,b];2n) where L(s)(t)(L1(s,t),L2(s,t))L:[a,b]\ni s\rightarrow L(s)\in\mathscr{P}([a,b];\mathbb{R}^{2n})\text{ where }L(s)(t)\coloneqq(L_{1}(s,t),L_{2}(s,t))

    such that dimL1(s,a)L2(s,a)\dim L_{1}(s,a)\cap L_{2}(s,a) and dimL1(s,b)L2(s,b)\dim L_{1}(s,b)\cap L_{2}(s,b) are both constant, and then,

    ιCLM(L(0);[a,b])=ιCLM(L(1);[a,b]).\operatorname{\iota^{\scriptscriptstyle{\mathrm{CLM}}}}(L(0);[a,b])=\operatorname{\iota^{\scriptscriptstyle{\mathrm{CLM}}}}(L(1);[a,b]).

Moreover, one efficient way to study the Maslov index is via the crossing form introduced by [17] as follows.

Let L(t):[0,1]Lag(n)L(t):[0,1]\to\mathrm{Lag}(n) be a smooth curve with L(0)=L0L(0)=L_{0}. Let WW be a fixed Lagrangian complement of L(t)L(t). For vL0v\in L_{0} and small tt, define w(t)Ww(t)\in W by v+w(t)L(t)v+w(t)\in L(t). The form Q(v)=ddt|t=0ω(v,w(t))Q(v)=\left.\frac{\operatorname{d}}{\operatorname{d}t}\right|_{t=0}\omega(v,w(t)) is independent of the choice of WW[17]. A crossing for L(t)L(t) is some tt for which L(t)L(t) intersects VLag(n)V\in\mathrm{Lag}(n) nontrivially. At each crossing, the crossing form is defined as

Γ(L(t),V;t)=Q|L(t)V.\displaystyle\Gamma(L(t),V;t)=\left.Q\right|_{L(t)\cap V}. (1.10)

A crossing is called regular if the crossing form is nondegenerate. For a quadratic form QQ, we use the notation sign(Q)\text{sign}(Q) for its signature. We also write m+(Q)m^{+}(Q) and m(Q)m^{-}(Q) for the positive and negative indices of inertia of QQ, so that

sign(Q)=m+(Q)m(Q).\text{sign}(Q)=m^{+}(Q)-m^{-}(Q).

From [20], if the path L(t)L(t) has only regular crossing with respect to VV, then we have that

ιCLM(V,L(t),t[a,b])=m+(Γ(L(a),V;a))+a<t<bsignΓ(L(t),V;t)m(Γ(L(b),V;b)).\displaystyle\operatorname{\iota^{\scriptscriptstyle{\mathrm{CLM}}}}(V,L(t),t\in[a,b])=m^{+}(\Gamma(L(a),V;a))+\sum\limits_{a<t<b}\text{sign}\Gamma(L(t),V;t)-m^{-}(\Gamma(L(b),V;b)). (1.11)

We recall the definition of Maslov index for a traveling pulse wave ww^{*} of (1.1) defined in [9]. The only requirement is that τ0\tau_{0} is large enough so that

Es(τ0)Eu(τ)={0} for all ττ0.\displaystyle E^{s}(\tau_{0})\cap E^{u}(\tau)=\{0\}\text{ for all }\tau\geqslant\tau_{0}. (1.12)

Moreover, since the transformation used in the conversion of (1.4) to (1.5) is different from that in [9], then the crossing form defined in the previous paper [9] differs from ours by a negative sign, and so, we provide [9, Definition 3.6] in the following form:

Definition 1.4.

[9, Definition 3.6]Let ww^{*} be a traveling pulse solution of (1.1), where the Maslov index of ww^{*} is given by

𝐌𝐚𝐬𝐥𝐨𝐯(w)τ<τ0signΓ(Eu(τ),Es(τ0);τ)m(Γ(Eu(τ),Es(τ0);τ0)).\mathbf{Maslov}(w^{*})\coloneqq\sum\limits_{\tau<\tau_{0}}\mathrm{sign}\Gamma\left(E^{u}(\tau),E^{s}(\tau_{0});\tau\right)-m^{-}(\Gamma\left(E^{u}(\tau),E^{s}(\tau_{0});\tau_{0}\right)).

By using the Maslov index, this paper [9] provides a unified geometric treatment for the stability analysis of the traveling pulses solution and has been successfully employed in [11, 10] to obtain some elegant stability results. Motivated by [9], it is natural to consider the case of the traveling front solutions. We remark that Definition 1.4 is dependent on the condition (1.12). It is easy to check that this condition may be inefficient for the traveling front solutions, and we sidestep this difficulty by following [13, Definition 1.2] and define the Maslov index for the traveling waves as following.

Definition 1.5.

Let ww^{*} be a traveling wave of (1.1), and define the Maslov index of ww^{*} as

ι(w)=ιCLM(Es(τ),Eu(τ);τ+¯).\operatorname{\iota}(w^{*})=-\operatorname{\iota^{\scriptscriptstyle{\mathrm{CLM}}}}\left(E^{s}(\tau),E^{u}(-\tau);\tau\in\overline{\mathbb{R}^{+}}\right).
Remark 1.6.

If Definition 1.4 is well-defined, then it should be independent of the choice of τ0\tau_{0}. It was shown in [7] that this definition is independent of τ0\tau_{0}, as long as (1.12) is satisfied. However, for the traveling front solution, there may be an intersection point between Es(τ)E^{s}(\tau) and Eu(τ)E^{u}(-\tau) at ++\infty. Thus, (1.12) may be inefficient. We sidestep this difficulty by following [13, Definition 1.2], and the following proposition shows that Definition 1.5 generalizes Definition 1.4 to the front case.

Proposition 1.7.

Letting ww^{*} be a traveling pulse solution of (1.1) and supposing that (H1) holds, then there exists τ0>0\tau_{0}>0, such that Eu()Es(+)={0}E^{u}(-\infty)\cap E^{s}(+\infty)=\{0\} for all ττ0\tau\geqslant\tau_{0} and

ι(w)=Maslov(w).\operatorname{\iota}(w^{*})=\mathrm{Maslov}(w^{*}). (1.13)

Lemmas 2.10 and 2.5 show that the set of nonnegative, real eigenvalues of LL is bounded above, and then, the spectrum of LL in +¯\overline{\mathbb{C}^{+}} consists of isolated eigenvalues of finite multiplicity (Cf. page 172 of [2]), and so, it follows that the quantities

  • N+(L)the number of real, positive eigenvalues of L counting algebraic multiplicityN_{+}(L)\coloneqq\text{the number of real, positive eigenvalues of $L$ counting algebraic multiplicity}

  • N¯+(L)the number of real, nonnegative eigenvalues of L counting algebraic multiplicity\overline{N}_{+}(L)\coloneqq\text{the number of real, nonnegative eigenvalues of $L$ counting algebraic multiplicity}

are both well-defined.

Now, we introduce a symplectic invariant called triple index (Cf. Definition 4.3) which can be used to compute the Maslov index, and we denote by ι(L1,L2;L3)\operatorname{\iota}(L_{1},L_{2};L_{3}) the triple index for any L1,L2,L3Lag(n)L_{1},L_{2},L_{3}\in\mathrm{Lag}(n).

Theorem 1.8.

If (H(H1)) and (H(H2)) hold, then

|ι(w)+ι(Eu(),Es(+);LR)|N¯0(L), where LR={[pq]|pV+(Q)andqV(Q)}.|\operatorname{\iota}(w^{*})+\operatorname{\iota}\left(E^{u}(-\infty),E^{s}(+\infty);L_{R}\right)|\leqslant\overline{N}_{0}(L),\text{ where }L_{R}=\left\{\left.\begin{bmatrix}p\\ q\end{bmatrix}\right|p\in V^{+}(Q)\ \mathrm{and}\ q\in V^{-}(Q)\right\}. (1.14)
Remark 1.9.

Under the conditions (H1) and (H2), if (1.12) holds, then Definition 1.4 is well-defined for the traveling front solution. Using the notation of Maslov box in [9, Figure 4.1], by the same discussion in the proof of Proposition 1.7, it is easy to prove that the Maslov index contribution along the “bottom shelf” α4\alpha_{4} is equal to ι(Eu(),Es(+);LR)\operatorname{\iota}\left(E^{u}(-\infty),E^{s}(+\infty);L_{R}\right). Based on this discussion and Theorem 1.8, the distinction between the pulse and front is nontrivial, and we generalize [9, Theorem 4.1] to the front case.

Now, we present the central result of this paper.

Theorem 1.10.

If (H(H2’)) holds, then we have that

|ι(w)|N¯0(L).|\operatorname{\iota}(w^{*})|\leqslant\overline{N}_{0}(L). (1.15)
Remark 1.11.

If (H(H2’)) holds, by Lemmas 3.4 and 3.5, we have that Es(+)Eu()={0}E^{s}(+\infty)\cap E^{u}(-\infty)=\{0\}, and this implies that if (1.12) holds, then Definition 1.4 is well-defined. By invoking (3.11) and Lemma 3.6, (H(H2’)) provides a sufficient condition for the Maslov index contribution along the “bottom shelf” when α4\alpha_{4} is equal to 0. This can be easily verified for λ=0\lambda=0, and it remains true for λ>0\lambda>0. This is important for practitioners who want to use the Maslov index to prove stability. As shown in [11], an employed strategy can be used to compute Maslov(w)\text{Maslov}(w^{*}) for the doubly-diffusive FitzHugh-Nagumo system. Based on those, Definition 1.4 is more practical than Definition 1.5 if one wants to compute the Maslov index for this case. Moreover, from the same discussion in the proof of Proposition 1.7, under (H(H2’)), Definitions 1.4 and 1.5 are equivalent. The strategy employed in [11] may be valid for computing ι(w)\operatorname{\iota}(w^{*}) for the doubly diffusive FitzHugh-Nagumo system, such as to consider a traveling front solution established in [11, Theorem 2.2].

For the following FitzHugh-Nagumo equations

{ut=uxx+1d(f(u)v),vt=vxx+uγv,\displaystyle\begin{cases}u_{t}&=u_{xx}+\frac{1}{d}(f(u)-v),\\ v_{t}&=v_{xx}+u-\gamma v,\end{cases} (1.16)

where d,γ>0d,\ \gamma>0 and f(u)=u(1u)(ua)f(u)=u(1-u)(u-a) with 0<a<120<a<\frac{1}{2}. If γ\gamma is large enough, then the equation u=γf(u)u=\gamma f(u) has three solutions 0=u1<u2<u30=u_{1}<u_{2}<u_{3}, and we remark that it is easy to check that f(0)<0f^{\prime}(0)<0 and f(u3)<0f^{\prime}(u_{3})<0. Now, we consider a traveling front solution ww^{*} of the FitzHugh-Nagumo equation (1.16): its existence has been established in [6], and we remark that such waves are obtained as local minimizers of an energy functional, and based on the variational characterization, authors [5] utilize the spectral flow to define and calculate a stability index. As a supplement, we provide a geometric insight into the stability of the traveling front solution for the FitzHugh-Nagumo equation (1.16).

Making a comparison with (1.1), then, (1.16) can be expressed as

wt=wxx+QDF(w),\displaystyle w_{t}=w_{xx}+QD\nabla F(w), (1.17)

where D=[1d001]D=\begin{bmatrix}\frac{1}{d}&0\\ 0&1\end{bmatrix}, Q=[1001]Q=\begin{bmatrix}1&0\\ 0&-1\end{bmatrix} and F(w)=[f(u)vγvu].\nabla F(w)=\begin{bmatrix}f(u)-v\\ \gamma v-u\end{bmatrix}. The eigenvalue problem

ϕ¨+cϕ˙+QD2F(w)ϕ=λϕ\ddot{\phi}+c\dot{\phi}+QD\nabla^{2}F(w^{*})\phi=\lambda\phi

or its equivalent eigenvalue problem

Lϕ=λϕ\displaystyle L\phi=\lambda\phi (1.18)

is studied to determine the stability of ww^{*}, where L=d2dξ2+cddξ+QB(ξ)L=\frac{\operatorname{d}^{2}}{\operatorname{d}\xi^{2}}+c\frac{\operatorname{d}}{\operatorname{d}\xi}+QB(\xi) and B(ξ)=D122F(w)D12B(\xi)=D^{\frac{1}{2}}\nabla^{2}F(w^{*})D^{\frac{1}{2}}.

Theorem 1.12.

Letting ww^{*} be a traveling front solution of the FitzHugh-Nagumo equation (1.16) satisfying the following asymptotic condition:

limξw=(u3,u3γ),limξ+w=(0,0)\lim\limits_{\xi\to-\infty}w^{*}=(u_{3},\frac{u_{3}}{\gamma}),\ \lim\limits_{\xi\to+\infty}w^{*}=(0,0)

and d>γ2d>\gamma^{-2}, then we have that

N+(L)=ι(u).N_{+}(L)=\operatorname{\iota}(u^{*}). (1.19)

2. Preliminary: an index formula

As is commonly done for the Sturm-Liouville operators [7, 14], in this section, by applying [13, Theorem 1], we derive an index formula (Cf. Proposition 2.12) for the Hamiltonian system (2.4), which plays a crucial role in obtaining the spectral information of LL.

The notation of spectral flow was introduced by Atiyah, Patodi and Singer in their study of index theory on manifolds with a boundary [3]. For the reader’s convenience, we provide a brief description of the basic properties of spectral flow. Suppose that EE is a real separable Hilbert space, and denote by 𝒞sa(E)\mathscr{CF}^{sa}(E) the space of all closed self-adjoint and Fredholm operators equipped with the gap topology. Let A:[0,1]𝒞sa(E)A:[0,1]\rightarrow\mathscr{CF}^{sa}\left(E\right) be a continuous curve. The Sf(At;t[0,1)]\mathrm{Sf\,}(A_{t};t\in[0,1)] counts the algebraic multiplicities of the spectral flow of AtA_{t} across the line t=ϵt=-\epsilon with some small positive number ϵ\epsilon.

For each AtA_{t}, there is an orthogonal splitting

E=E(At)E0(At)E+(At),E=E_{-}(A_{t})\oplus E_{0}(A_{t})\oplus E_{+}(A_{t}),

where E0E_{0} is the null space of AtA_{t}, Atv,v0\langle A_{t}v,v\rangle\geqslant 0 if vE+v\in E_{+} and Atv,v0\langle A_{t}v,v\rangle\leqslant 0 if vEv\in E_{-}. There is an efficient way to compute the spectral flow through what are called crossing forms. Let PtP_{t} be the orthogonal projector from EE to E0(At)E_{0}\big{(}A_{t}\big{)}. When E0(At0){0}E_{0}\big{(}A_{t_{0}}\big{)}\neq\{0\}, we call the instant t0t_{0} a crossing instant. In this case, we defined the crossing from Cr[At0]\mathrm{Cr}[A_{t_{0}}] as

Cr[At0]Pt0tPt0:E0(At0)E0(At0).\mathrm{Cr}[A_{t_{0}}]\coloneqq P_{t_{0}}\dfrac{\partial}{\partial t}P_{t_{0}}:E_{0}\big{(}A_{t_{0}}\big{)}\to E_{0}\big{(}A_{t_{0}}\big{)}.

We call the crossing instant t0t_{0} regular if the crossing form Cr(At0)\mathrm{Cr}(A_{t_{0}}) is nondegenerate. In this case we define the signature simply as

sgn(Cr(At0))dimE+(Cr(At0))dimE(Cr(At0)).\operatorname{sgn}\big{(}\mathrm{Cr}(A_{t_{0}})\big{)}\coloneqq\dim E_{+}\big{(}\mathrm{Cr}(A_{t_{0}})\big{)}-\dim E_{-}\big{(}\mathrm{Cr}(A_{t_{0}})\big{)}.

We assume that all of the crossings are regular. Then, the crossing instants are isolated (and, hence, those on a compact interval are finite in number), and the spectral flow is given by the following formula

Sf(At;t[0,1])=t0𝒮sgn(Cr(At0))dimE(Cr(A0))+dimE+(Cr(A1))\mathrm{Sf\,}(A_{t};t\in[0,1])=\sum_{t_{0}\in\mathcal{S}_{*}}\operatorname{sgn}\big{(}\mathrm{Cr}(A_{t_{0}})\big{)}-\dim E_{-}\big{(}\mathrm{Cr}(A_{0})\big{)}+\dim E_{+}\big{(}\mathrm{Cr}(A_{1})\big{)} (2.1)

where 𝒮𝒮(a,b)\mathcal{S}_{*}\coloneqq\mathcal{S}\cap(a,b), and 𝒮\mathcal{S} denotes the set of all crossings.

We list some basic properties of spectral flow:

Proposition 2.1.
  1. (1)

    If t0[0,1]t_{0}\in[0,1], then

    Sf(At;t[0,1])=Sf(At;t[0,t0])+Sf(At;t[t0,1]).\mathrm{Sf\,}(A_{t};t\in[0,1])=\mathrm{Sf\,}(A_{t};t\in[0,t_{0}])+\mathrm{Sf\,}(A_{t};t\in[t_{0},1]).
  2. (2)

    If 𝒮𝒮(0,1)\mathcal{S}_{*}\coloneqq\mathcal{S}\cap(0,1) and 𝒮\mathcal{S} denotes the set of all crossings, then

    Sf(At;t[0,1])t𝒮dimE0(At).\mathrm{Sf\,}(A_{t};t\in[0,1])\leqslant\sum\limits_{t\in\mathcal{S}_{*}}\dim E_{0}(A_{t}).

Now, we return to the problem of stability analysis. If there is not a Hamiltonian structure for LL by the presence of the ddξ\frac{\operatorname{d}}{\operatorname{d}\xi} term, we can circumvent this by considering instead the operator

𝕃ecξ2Lecξ2=d2dξ2c24I+QB,\mathbb{L}\coloneqq e^{\frac{c\xi}{2}}Le^{-\frac{c\xi}{2}}=\frac{\operatorname{d}^{2}}{\operatorname{d}\xi^{2}}-\frac{c^{2}}{4}I+QB, (2.2)

as is done in [12, 9]. We now proceed to the study of the following eigenvalue problem

𝕃φ=λφ,\mathbb{L}\varphi=\lambda\varphi, (2.3)

Setting ψ=φ˙12cφ\psi=\dot{\varphi}-\frac{1}{2}c\varphi and z=[ψφ]z=\begin{bmatrix}\psi\\ \varphi\end{bmatrix} converts Equation (2.3) to the following Hamiltonian system

z˙=JHλ(ξ)z,\dot{z}=JH_{\lambda}(\xi)z, (2.4)

where Hλ=JAλ12cJ=[Q12cQ12cQBλQ].H_{\lambda}=-JA_{\lambda}-\frac{1}{2}cJ=\begin{bmatrix}Q&\frac{1}{2}cQ\\ \frac{1}{2}cQ&B-\lambda Q\end{bmatrix}. We define the self-adjoint operators

FλJddξHλ:domFλL2(,2n)L2(,2n),\displaystyle{F}_{\lambda}\coloneqq-J\frac{\operatorname{d}}{\operatorname{d}\xi}-H_{\lambda}:\operatorname{dom}{F}_{\lambda}\subset L^{2}(\mathbb{R},\mathbb{R}^{2n})\subset L^{2}(\mathbb{R},\mathbb{R}^{2n}), (2.5)

where domFλ=W1,2(,2n)\operatorname{dom}{F}_{\lambda}=W^{1,2}(\mathbb{R},\mathbb{R}^{2n}) and λ[0,C]\lambda\in[0,C]. Here, CC is given in (1.3). The following Proposition 2.2 shows that the Hamiltonian system (2.4) has close contact with the system (1.5).

Proposition 2.2.

Letting λ+¯\lambda\in\overline{\mathbb{C}^{+}} and zH2(,2n)z\in H^{2}(\mathbb{R},\mathbb{C}^{2n}), then zker(ddξAλ)z\in\ker\left(\frac{\operatorname{d}}{\operatorname{d}\xi}-A_{\lambda}\right) if and only if e12cξzker(JddξHλ)e^{\frac{1}{2}c\xi}z\in\ker\left(-J\frac{\operatorname{d}}{\operatorname{d}\xi}-H_{\lambda}\right).

It is well-known [16, Lemma 3.1.10] that the essential spectrum is given by

σess(L)={λ|Aλ(+)orAλ()hasapureimaginaryeigenvalue}.\sigma_{ess}(L)=\{\lambda\in\mathbb{C}|A_{\lambda}(+\infty)\mathrm{\ or\ }A_{\lambda}(-\infty)\mathrm{\ has\ a\ pure\ imaginary\ eigenvalue}\}.

Following the same method as that in paper [11, Section 2], we here give the proof of the following Lemma for convenience of the reader.

Lemma 2.3.

Under the condition (H(H1)), we have that

  • (1)

    σess(L)\sigma_{ess}(L)\subset\mathbb{C}^{-},

  • (2)

    if λ+¯\lambda\in\overline{\mathbb{C}^{+}}, then Aλ(±)A_{\lambda}(\pm\infty) are hyperbolic.

Proof.

A simple calculation shows the eigenvalues of Aλ(+)A_{\lambda}(+\infty) and Aλ()A_{\lambda}(-\infty), such that

μ(λ)=12{c±c2+4(λα)} and ν(λ)=12{c±c2+4(λβ),} respectively,\mu(\lambda)=\frac{1}{2}\{-c\pm\sqrt{c^{2}+4(\lambda-\alpha)}\}\textrm{ and }\nu(\lambda)=\frac{1}{2}\{-c\pm\sqrt{c^{2}+4(\lambda-\beta),}\}\textrm{ respectively}, (2.6)

where α\alpha is the eigenvalue of QB+QB_{+}, and β\beta is the eigenvalue of QBQB_{-}. To prove that (1) holds, we only need to show that Aλ(+)A_{\lambda}(+\infty) and Aλ()A_{\lambda}(-\infty) have no purely imaginary eigenvalues if λ+¯\lambda\in\overline{\mathbb{C}^{+}}, which is equivalent to showing that c2+4(λα)c\mathrm{\mathfrak{R}}\sqrt{c^{2}+4(\lambda-\alpha)}\neq-c and c2+4(λβ)c\mathrm{\mathfrak{R}}\sqrt{c^{2}+4(\lambda-\beta)}\neq-c. From the formula a+ib={12a2+b2+a}12,\mathrm{\mathfrak{R}}\sqrt{a+ib}=\{\frac{1}{2}\sqrt{a^{2}+b^{2}}+a\}^{\frac{1}{2}}, we have that

c2+4(λα)={12[(c2+4(λα))2+((λα))2]12+c2+4(λα)}12,\mathrm{\mathfrak{R}}\sqrt{c^{2}+4(\lambda-\alpha)}=\{\frac{1}{2}[{\left(c^{2}+4\mathrm{\mathfrak{R}}\left(\lambda-\alpha\right)\right)^{2}+\left(\mathrm{\mathfrak{I}}(\lambda-\alpha)\right)^{2}}]^{\frac{1}{2}}+c^{2}+4\mathrm{\mathfrak{R}}\left(\lambda-\alpha\right)\}^{\frac{1}{2}}, (2.7)

we note that (λα)>0\mathrm{\mathfrak{R}}\left(\lambda-\alpha\right)>0, so Equation (2.7) implies that

c2+4(λα)>c.\mathrm{\mathfrak{R}}\sqrt{c^{2}+4(\lambda-\alpha)}>c. (2.8)

This calculation actually proves that Aλ(+)A_{\lambda}(+\infty) has exactly nn eigenvalues of the positive real part and nn eigenvalues of the negative real part for λ0\mathrm{\mathfrak{R}}\lambda\geqslant 0. Additionally, Aλ()A_{\lambda}(-\infty). We label the eigenvalues of Aλ(±)A_{\lambda}(\pm\infty) in order of increasing real part and observe that

μ1(λ)μn(λ)<c<0<μn+1(λ)μ2n(λ),\mathrm{\mathfrak{R}}\mu_{1}(\lambda)\leqslant\cdots\leqslant\mathrm{\mathfrak{R}}\mu_{n}(\lambda)<-c<0<\mathrm{\mathfrak{R}}\mu_{n+1}(\lambda)\leqslant\cdots\leqslant\mathrm{\mathfrak{R}}\mu_{2n}(\lambda), (2.9)
ν1(λ)νn(λ)<c<0<νn+1(λ)ν2n(λ).\mathrm{\mathfrak{R}}\nu_{1}(\lambda)\leqslant\cdots\leqslant\mathrm{\mathfrak{R}}\nu_{n}(\lambda)<-c<0<\mathrm{\mathfrak{R}}\nu_{n+1}(\lambda)\leqslant\cdots\leqslant\mathrm{\mathfrak{R}}\nu_{2n}(\lambda). (2.10)

From those, we complete the proof. ∎

Remark 2.4.

We note that JHλ(±)=Aλ(±)+12cJH_{\lambda}(\pm\infty)=A_{\lambda}(\pm\infty)+\frac{1}{2}c: from Lemma 2.3, if (H(H1)) holds, then JHλ(±)JH_{\lambda}(\pm\infty) are both hyperbolic for all λ+¯\lambda\in\overline{\mathbb{C}^{+}}, and this, together with [14, Theorem 1], tells us that for each λ[0,C]\lambda\in[0,C], the operator Fλ{F}_{\lambda} is a self-adjoint Fredholm operator, and in particular, it is possible to associate it to path

λFλ\displaystyle\lambda\to F_{\lambda} (2.11)

the topological invariant: spectral flow.

The proof of Proposition 2.2.

If zker(ddξAλ)z\in\ker\left(\frac{\operatorname{d}}{\operatorname{d}\xi}-A_{\lambda}\right). By a simple calculation,

ddξ(e12cξz)=12ce12cξz+e12cξz˙=J(12cJJAλ(ξ))e12cξz=JHλ(ξ)e12cξz.\displaystyle\frac{\operatorname{d}}{\operatorname{d}\xi}\left(e^{\frac{1}{2}c\xi}z\right)=\frac{1}{2}ce^{\frac{1}{2}c\xi}z+e^{\frac{1}{2}c\xi}\dot{z}=J\left(-\frac{1}{2}cJ-JA_{\lambda}(\xi)\right)e^{\frac{1}{2}c\xi}z=JH_{\lambda}(\xi)e^{\frac{1}{2}c\xi}z. (2.12)

From Lemma 2.3, zz must decay as fast as eνn(λ)ξe^{\nu_{n}(\lambda)\xi} as ξ+\xi\to+\infty, and we note that if νn(λ)<c\nu_{n}(\lambda)<-c, then e12cξze^{\frac{1}{2}c\xi}z and e12cξz˙e^{\frac{1}{2}c\xi}\dot{z} both exponentially decay to 0 as ξ±\xi\to\pm\infty, and hence, e12cξzL2(,2n)e^{\frac{1}{2}c\xi}z\in L^{2}(\mathbb{R},\mathbb{C}^{2n}) and e12cξz˙L2(,2n)e^{\frac{1}{2}c\xi}\dot{z}\in L^{2}(\mathbb{R},\mathbb{C}^{2n}). Then, e12cξzH2(,2n)e^{\frac{1}{2}c\xi}z\in H^{2}(\mathbb{R},\mathbb{C}^{2n}): this together with Equation (2.12) implies that e12cξzker(JddξHλ)e^{\frac{1}{2}c\xi}z\in\ker\left(-J\frac{\operatorname{d}}{\operatorname{d}\xi}-H_{\lambda}\right).

Conversely, if e12cξzker(JddξHλ)e^{\frac{1}{2}c\xi}z\in\ker\left(-J\frac{\operatorname{d}}{\operatorname{d}\xi}-H_{\lambda}\right), then by a simple calculation,

z˙=12cz+z˙12cz=e12cξ(12ce12cξz+e12cξz˙)12cz=e12cξddξ(e12cξz)12cz=Aλz.\displaystyle\dot{z}=\frac{1}{2}cz+\dot{z}-\frac{1}{2}cz=e^{-\frac{1}{2}c\xi}\left(\frac{1}{2}ce^{\frac{1}{2}c\xi}z+e^{\frac{1}{2}c\xi}\dot{z}\right)-\frac{1}{2}cz=e^{-\frac{1}{2}c\xi}\frac{\operatorname{d}}{\operatorname{d}\xi}\left(e^{\frac{1}{2}c\xi}z\right)-\frac{1}{2}cz=A_{\lambda}z. (2.13)

We note that when JHλ(±)=Aλ(±)+12cJH_{\lambda}(\pm\infty)=A_{\lambda}(\pm\infty)+\frac{1}{2}c, for each μ^σ(JHλ(+))+\hat{\mu}\in\sigma(JH_{\lambda}(+\infty))\cap\mathbb{C}^{+}, invoking Lemma 2.5, it is easy to check that μ^>12c\Re\hat{\mu}>\frac{1}{2}c: then, e12cξze^{\frac{1}{2}c\xi}z must decay at least as fast as e12cξe^{\frac{1}{2}c\xi} as ξ\xi\to-\infty. We thus determine that zz and z˙\dot{z} both exponentially decay as ξ±\xi\to\pm\infty, and then, z,z˙L2(,2n)z,\ \dot{z}\in L^{2}(\mathbb{R},\mathbb{C}^{2n}), and this together with Equation (2.13), implies that zker(ddξAλ)z\in\ker\left(\frac{\operatorname{d}}{\operatorname{d}\xi}-A_{\lambda}\right).

As a direct consequence, the following result holds.

Corollary 2.5.

Letting λ+¯\lambda\in\overline{\mathbb{C}^{+}} and ϕH2(,n)\phi\in H^{2}(\mathbb{R},\mathbb{C}^{n}), then ϕker(LλI)\phi\in\ker(L-\lambda I) if and only if ecξ2ϕker(𝕃λI)e^{\frac{c\xi}{2}}\phi\in\ker(\mathbb{L}-\lambda I).

Remark 2.6.

Under the condition (H(H1)), if yy satisfies (1.5) with limξ+y=0\lim\limits_{\xi\to+\infty}y=0, then a similar discussion in the proof of Proposition 2.2 guarantees that e12ξye^{\frac{1}{2}\xi}y satisfies (2.4) with limξ+e12ξy=0\lim\limits_{\xi\rightarrow+\infty}e^{\frac{1}{2}\xi}y=0. Then, {e12cτy(τ)|y solve (1.5) and y0 as τ+}\{e^{\frac{1}{2}c\tau}y(\tau)|y\text{ solve }(1.5)\text{ and }y\to 0\text{ as }\tau\to+\infty\} of the Hamiltonian system (2.4), so we can say that Eλs(τ)E^{s}_{\lambda}(\tau) is also the stable space of the Hamiltonian system (2.4). By the same reasoning, Eλu(τ)E^{u}_{\lambda}(\tau) is also the unstable space of the Hamiltonian system (2.4). Moreover, letting Eλs(+){v2n|limξ+exp(ξAλ(+))v=0} and Eλu(){v2n|limξexp(ξAλ())v=0}E_{\lambda}^{s}(+\infty)\coloneqq\left\{v\in\mathbb{R}^{2n}\left|\lim\limits_{\xi\to+\infty}\exp\left(\xi A_{\lambda}(+\infty)\right)v=0\right.\right\}\text{ and }E_{\lambda}^{u}(-\infty)\coloneqq\left\{v\in\mathbb{R}^{2n}\left|\lim\limits_{\xi\to-\infty}\exp\left(\xi A_{\lambda}(-\infty)\right)v=0\right.\right\}, a similar discussion in Proposition 2.2 shows that the following facts
Eλs(+)={v2n|limξ+exp(ξJHλ(+))v=0}E^{s}_{\lambda}(+\infty)=\left\{v\in\mathbb{R}^{2n}\left|\lim\limits_{\xi\to+\infty}\exp\left(\xi JH_{\lambda}(+\infty)\right)v=0\right.\right\} and Eλu()={v2n|limξexp(ξJHλ())v=0}E^{u}_{\lambda}(-\infty)=\left\{v\in\mathbb{R}^{2n}\left|\lim\limits_{\xi\to-\infty}\exp\left(\xi JH_{\lambda}(-\infty)\right)v=0\right.\right\} hold.

Under the condition (H(H1)), we determine that

Eλs(+)=limτ+Eλs(τ) and Eλu()=limτEλu(τ),E^{s}_{\lambda}(+\infty)=\lim\limits_{\tau\to+\infty}E^{s}_{\lambda}(\tau)\text{ and }E^{u}_{\lambda}(-\infty)=\lim\limits_{\tau\to-\infty}E^{u}_{\lambda}(\tau), (2.14)

where the convergence is meant in the gap (norm) topology of the Lagrangian Grassmannian (Cf. [1] for further details).

Given τ0\tau\geqslant 0, let B1(ξ)=B(τ+ξ)B_{1}(\xi)=B(\tau+\xi) with ξ+¯\xi\in\overline{\mathbb{R}^{+}} and B2(ξ)=B(ξτ)B_{2}(\xi)=B(\xi-\tau) with ξ¯\xi\in\overline{\mathbb{R}^{-}}. The aim of the next part is to provide some sufficient condition for the coefficient of (2.11) to obtain the nondegeneracy.

Lemma 2.7.

Let

𝕃λ,M+d2dξ2{c24+λ}I+QB1:W2,2(+¯,n)L2(+¯,n)\mathbb{L}^{+}_{\lambda,M}\coloneqq\frac{\operatorname{d}^{2}}{\operatorname{d}\xi^{2}}-\left\{\frac{c^{2}}{4}+\lambda\right\}I+QB_{1}:W^{2,2}\left(\overline{\mathbb{R}^{+}},\mathbb{R}^{n}\right)\to L^{2}\left(\overline{\mathbb{R}^{+}},\mathbb{R}^{n}\right) (2.15)

and

𝕃λ,Md2dξ2{c24+λ}I+QB2:W2,2(¯,n)L2(¯,n).\mathbb{L}^{-}_{\lambda,M}\coloneqq\frac{\operatorname{d}^{2}}{\operatorname{d}\xi^{2}}-\left\{\frac{c^{2}}{4}+\lambda\right\}I+QB_{2}:W^{2,2}\left(\overline{\mathbb{R}^{-}},\mathbb{R}^{n}\right)\to L^{2}\left(\overline{\mathbb{R}^{-}},\mathbb{R}^{n}\right). (2.16)

With CC given in (1.3), assuming that (H(H1)) holds, and λC\lambda\geqslant C, then the system

{𝕃λ,M+φ1=0=𝕃λ,Mφ2,φ1(0)=φ2(0),φ˙1(0)=φ˙2(0)\displaystyle\begin{cases}\mathbb{L}^{+}_{\lambda,M}\varphi_{1}=0=\mathbb{L}^{-}_{\lambda,M}\varphi_{2},\\ \varphi_{1}(0)=\varphi_{2}(0),\ \dot{\varphi}_{1}(0)=\dot{\varphi}_{2}(0)\end{cases} (2.17)

has only the zero solution.

Proof.

Assuming that the system has a solution (φ1,φ2)(\varphi_{1},\varphi_{2}), then we have that

𝕃λ,M+φ1,φ1L2+𝕃λ,Mφ2,φ2L2=0\displaystyle\langle\mathbb{L}^{+}_{\lambda,M}\varphi_{1},\varphi_{1}\rangle_{L^{2}}+\langle\mathbb{L}^{-}_{\lambda,M}\varphi_{2},\varphi_{2}\rangle_{L^{2}}=0 (2.18)

Integrating by part, we obtain

𝕃λ,M+φ1,φ1L2=φ˙1L22{c24+λ}φ1L22+0+QB1φ1,φ1dξφ1(0),φ˙1(0)\displaystyle\langle\mathbb{L}^{+}_{\lambda,M}\varphi_{1},\varphi_{1}\rangle_{L^{2}}=-\left\|\dot{\varphi}_{1}\right\|^{2}_{L^{2}}-\left\{\frac{c^{2}}{4}+\lambda\right\}\left\|\varphi_{1}\right\|^{2}_{L^{2}}+\int_{0}^{+\infty}\langle QB_{1}\varphi_{1},\varphi_{1}\rangle\operatorname{d}\xi-\langle\varphi_{1}(0),\dot{\varphi}_{1}(0)\rangle (2.19)
𝕃λ,Mφ2,φ2L2=φ˙2L22{c24+λ}φ2L22+0QB2φ2,φ2dξ+φ2(0),φ˙2(0)\displaystyle\langle\mathbb{L}^{-}_{\lambda,M}\varphi_{2},\varphi_{2}\rangle_{L^{2}}=-\left\|\dot{\varphi}_{2}\right\|^{2}_{L^{2}}-\left\{\frac{c^{2}}{4}+\lambda\right\}\left\|\varphi_{2}\right\|^{2}_{L^{2}}+\int_{-\infty}^{0}\langle QB_{2}\varphi_{2},\varphi_{2}\rangle\operatorname{d}\xi+\langle\varphi_{2}(0),\dot{\varphi}_{2}(0)\rangle (2.20)

Let Iλ,1φ˙1L22{c24+λ}φ1L22+0+QB1φ1,φ1dξI_{\lambda,1}\coloneqq-\left\|\dot{\varphi}_{1}\right\|^{2}_{L^{2}}-\left\{\frac{c^{2}}{4}+\lambda\right\}\left\|\varphi_{1}\right\|^{2}_{L^{2}}+\int_{0}^{+\infty}\langle QB_{1}\varphi_{1},\varphi_{1}\rangle\operatorname{d}\xi and Iλ,2φ˙2L22{c24+λ}φ2L22+0QB2φ2,φ2dξI_{\lambda,2}\coloneqq-\left\|\dot{\varphi}_{2}\right\|^{2}_{L^{2}}-\left\{\frac{c^{2}}{4}+\lambda\right\}\left\|\varphi_{2}\right\|^{2}_{L^{2}}+\int_{-\infty}^{0}\langle QB_{2}\varphi_{2},\varphi_{2}\rangle\operatorname{d}\xi. It is easy to see that Iiφ˙iL22{c24+λC}φiL22I_{i}\leqslant-\left\|\dot{\varphi}_{i}\right\|^{2}_{L^{2}}-\left\{\frac{c^{2}}{4}+\lambda-C\right\}\left\|\varphi_{i}\right\|^{2}_{L^{2}}. Then, by using the second condition in the above boundary value problem, we obtain

0=I1+I2i=1,2(φ˙iL22{c24+λC}φiL22).0=I_{1}+I_{2}\leqslant\sum\limits_{i=1,2}\left(-\left\|\dot{\varphi}_{i}\right\|^{2}_{L^{2}}-\left\{\frac{c^{2}}{4}+\lambda-C\right\}\left\|\varphi_{i}\right\|^{2}_{L^{2}}\right). (2.21)

If λC\lambda\geqslant C, then we infer that

I1+I2=0I_{1}+I_{2}=0

if and only if φi=φ˙i=0\varphi_{i}=\dot{\varphi}_{i}=0 for i=1,2i=1,2. This concludes the proof.

Let us now consider the associated first order differential operators Fλ,M+F_{\lambda,M}^{+} and Fλ,MF_{\lambda,M}^{-} of 𝕃λ,M+\mathbb{L}^{+}_{\lambda,M} and 𝕃λ,M\mathbb{L}^{-}_{\lambda,M}. A similar result holds.

Lemma 2.8.

With CC given in (1.3), assuming that (H(H1)) holds, and λC\lambda\geqslant C, then the system

{Fλ,M+z1=Fλ,Mz2=0z1(0)=z2(0)\displaystyle\begin{cases}F_{\lambda,M}^{+}z_{1}=F_{\lambda,M}^{-}z_{2}=0\\ z_{1}(0)=z_{2}(0)\end{cases} (2.22)

has only the zero solution.

Lemma 2.9.

With CC given in (1.3), assuming that (H(H1)) holds, we have that

Eλs(τ)Eλu(τ)={0} for all (λ,τ)[C,+)×+¯.E^{s}_{\lambda}(\tau)\cap E^{u}_{\lambda}(-\tau)=\{0\}\text{ for all }(\lambda,\tau)\in[C,+\infty)\times\overline{\mathbb{R}^{+}}. (2.23)
Proof.

Let B1(xi)=B(ξ+τ)B_{1}(xi)=B(\xi+\tau) with ξ+¯\xi\in\overline{\mathbb{R}^{+}} and B2(ξ)=B(ξτ)B_{2}(\xi)=B(\xi-\tau) with ξ¯\xi\in\overline{\mathbb{R}^{-}}. Then, the stable subspace of the equation Fλ,M+=0F_{\lambda,M}^{+}=0 at 0 is EλsE_{\lambda}^{s}, the unstable subspace of the equation Fλ,M=0F_{\lambda,M}^{-}=0 at 0 is EλuE_{\lambda}^{u}, and there exists a linear bijection from the set of solutions of the system

{Fλ,M+z1=Fλ,Mz2=0z1(0)=z2(0)\begin{cases}F_{\lambda,M}^{+}z_{1}=F_{\lambda,M}^{-}z_{2}=0\\ z_{1}(0)=z_{2}(0)\end{cases} (2.24)

with the subspace Eλs(τ)Eλu(τ)E^{s}_{\lambda}(\tau)\cap E^{u}_{\lambda}(-\tau). By invoking once again Lemma 2.8 and Lemma 2.7, we conclude that the initial value problem only admits the trivial solution for every λC\lambda\geqslant C. This concludes the proof. ∎

By setting B1(ξ)=B(ξ){B}_{1}(\xi)={B}(\xi) for every ξ0\xi\geqslant 0 and B2(ξ)=B(ξ){B}_{2}(\xi)={B}(\xi) for every ξ0\xi\leqslant 0, the following result holds.

Lemma 2.10.

With CC given in (1.3), assuming that (H(H1)) holds, if λC\lambda\geqslant C, then ker(𝕃λ)={0}\ker\left(\mathbb{L}-\lambda\right)=\{0\} and kerFλ={0}\ker F_{\lambda}=\{0\}.

From Lemma 2.10, we determine that the following result holds.

Corollary 2.11.

With CC given in (1.3), assuming that (H(H1)) holds, and if λC\lambda\geqslant C, then we have that kerSλ={0}\ker S_{\lambda}=\{0\}, where SλQd2dξ2+c24QB+λQS_{\lambda}\coloneqq-Q\frac{\operatorname{d}^{2}}{\operatorname{d}\xi^{2}}+\frac{c^{2}}{4}Q-{B}+\lambda Q.

It is well-known that for each λ[0,C]\lambda\in[0,C], the operator SλS_{\lambda} is closed and self-adjoint with dense domain in L2(,n)L^{2}(\mathbb{R},\mathbb{R}^{n}). As a byproduct of condition (H(H1)) and [14, Theorem 1], SλS_{\lambda} is also a Fredholm operator.

Finally, from [13, Theorem 1], we obtain that

Sf(Fλ,λ[0,C])=\displaystyle\mathrm{Sf\,}(F_{\lambda},\lambda\in[0,C])= ιCLM(ECs(τ),ECu(τ);τ+¯)ιCLM(Es(τ),Eu(τ);τ+¯)\displaystyle\operatorname{\iota^{\scriptscriptstyle{\mathrm{CLM}}}}(E^{s}_{C}(\tau),E^{u}_{C}(-\tau);\tau\in\overline{\mathbb{R}^{+}})-\operatorname{\iota^{\scriptscriptstyle{\mathrm{CLM}}}}(E^{s}(\tau),E^{u}(-\tau);\tau\in\overline{\mathbb{R}^{+}}) (2.25)
ιCLM(Eλs(+),Eλu();λ[0,C]).\displaystyle-\operatorname{\iota^{\scriptscriptstyle{\mathrm{CLM}}}}(E^{s}_{\lambda}(+\infty),E^{u}_{\lambda}(-\infty);\lambda\in[0,C]). (2.26)

As a direct consequence of Lemma 2.9 and Equation (2.25), we obtain the following result.

Proposition 2.12.

Under the previous notations and assuming that (H(H1)) holds, the following equation holds:

Sf(Fλ,λ[0,C])=ιCLM(Es(τ),Eu(τ);τ+¯)+ιCLM(Eλs(+),Eλu();λ[0,C]).\displaystyle-\mathrm{Sf\,}(F_{\lambda},\lambda\in[0,C])=\operatorname{\iota^{\scriptscriptstyle{\mathrm{CLM}}}}(E^{s}(\tau),E^{u}(-\tau);\tau\in\overline{\mathbb{R}^{+}})+\operatorname{\iota^{\scriptscriptstyle{\mathrm{CLM}}}}(E^{s}_{\lambda}(+\infty),E^{u}_{\lambda}(-\infty);\lambda\in[0,C]). (2.27)

3. The proof of the main result

The goal of this section is to provide a detailed proof of Theorem 1.8, Theorem 1.10 and Theorem 1.12. Before showing the proof, we start by analyzing the distribution of the eigenvalues of 𝕃\mathbb{L}.

As in [8], the same discussion can be used here to obtain the distribution of eigenvalues of 𝕃\mathbb{L}, which serves a crucial role of counting the number of nonnegative eigenvalues of LL via spectral flow. Let Q+Q^{+} and QQ^{-} be the orthogonal projections from EE to E+(Q)E_{+}(Q) and E(Q)E_{-}(Q), respectively. Define 𝒮1=Q+SQ+\mathscr{S}_{1}=Q^{+}SQ^{+}, 𝒮2=QSQ\mathscr{S}_{2}=Q^{-}SQ^{-} and 𝒮3=Q+SQ\mathscr{S}_{3}=Q^{+}SQ^{-}: in other words, SS can be decomposed as

[𝒮1𝒮3𝒮3𝒮2],\begin{bmatrix}\mathscr{S}_{1}&\mathscr{S}_{3}\\ \mathscr{S}_{3}^{*}&\mathscr{S}_{2}\end{bmatrix}, (3.1)

where 𝒮3=𝒮¯3T\mathscr{S}_{3}^{*}=\bar{\mathscr{S}}_{3}^{T} and 𝒮¯3\bar{\mathscr{S}}_{3} denotes the complex conjugate of 𝒮3\mathscr{S}_{3}. For a linear self-adjoint operator AA defined on a Hilbert space EE, denoted by A>0A>0, if Av,v>\langle Av,v\rangle> for all vE\{0}v\in E\backslash\{0\}, two linear operators AA and A^\hat{A} are denoted by A>A^A>\hat{A} if AA^>0A-\hat{A}>0.

Lemma 3.1.

[8] Supposing that 𝒮1>0\mathscr{S}_{1}>0 and I>𝒮3𝒮12𝒮3I>\mathscr{S}_{3}^{*}\mathscr{S}_{1}^{-2}\mathscr{S}_{3}, then σ(𝕃)C+¯\sigma(\mathbb{L})\cap\overline{{C}^{+}}\subset\mathbb{R}. The same assertion holds if 𝒮2>0-\mathscr{S}_{2}>0 and I>𝒮3(𝒮2)2𝒮3I>\mathscr{S}_{3}^{*}(-\mathscr{S}_{2})^{-2}\mathscr{S}_{3}.

Proposition 3.2.

Under the condition (H(H1)), we have that
(1) if 𝒮2>0-\mathscr{S}_{2}>0 and I>𝒮3(𝒮2)2𝒮3I>\mathscr{S}_{3}^{*}(-\mathscr{S}_{2})^{-2}\mathscr{S}_{3}, then Sf(Sλ;λ[0,C])=N+(L)\mathrm{Sf\,}(S_{\lambda};\lambda\in[0,C])=N_{+}(L),
(2) if 𝒮1>0\mathscr{S}_{1}>0 and I>𝒮3𝒮12𝒮3I>\mathscr{S}_{3}^{*}\mathscr{S}_{1}^{-2}\mathscr{S}_{3}, then Sf(Sλ;λ[0,C])=N¯+(L)\mathrm{Sf\,}(S_{\lambda};\lambda\in[0,C])=-\overline{N}_{+}(L).

Proof.

We only prove (1), while the other is analogous. From Remark 2.6, we know that N+(L)=N+(𝕃)N_{+}(L)=N_{+}(\mathbb{L}). Next, we prove that Sf(Sλ;λ[0,C])=N+(𝕃)\mathrm{Sf\,}(S_{\lambda};\lambda\in[0,C])=N_{+}(\mathbb{L}). Suppose that along the spectral flow, there is a crossing at SλS_{\lambda} for some λ[0,C]\lambda\in[0,C] and ϕkerSλ\phi\in\ker S_{\lambda}: that is,

Qϕ¨+(14c2QB+λQ)ϕ=0.\displaystyle-Q\ddot{\phi}+\left(\frac{1}{4}c^{2}Q-B+\lambda Q\right)\phi=0. (3.2)

Letting ϕ+=Q+ϕ\phi_{+}=Q^{+}\phi and ϕ=Qϕ\phi_{-}=Q^{-}\phi, we can rewrite Equation (3.2) as

𝒮1ϕ++𝒮3ϕ=λϕ+\displaystyle\mathscr{S}_{1}\phi_{+}+\mathscr{S}_{3}\phi_{-}=-\lambda\phi_{+} (3.3)
𝒮3ϕ++𝒮2ϕ=λϕ.\displaystyle\mathscr{S}^{*}_{3}\phi_{+}+\mathscr{S}_{2}\phi_{-}=\lambda\phi_{-}. (3.4)

Solving Equation (3.4), we obtain that ϕ=(λ𝒮2)1𝒮3ϕ+\phi_{-}=(\lambda-\mathscr{S}_{2})^{-1}\mathscr{S}_{3}^{*}\phi_{+}, and this, together with Equation (3.3), obtains

ddλSλϕ,ϕ=Qϕ,ϕ=ϕ+,ϕ+ϕ,ϕ=ϕ+,ϕ+𝒮3(λ𝒮2)2𝒮3ϕ+,ϕ+\displaystyle\frac{\operatorname{d}}{\operatorname{d}\lambda}\langle S_{\lambda}\phi,\phi\rangle=\langle Q\phi,\phi\rangle=\langle\phi_{+},\phi_{+}\rangle-\langle\phi_{-},\phi_{-}\rangle=\langle\phi_{+},\phi_{+}\rangle-\langle\mathscr{S}_{3}^{*}(\lambda-\mathscr{S}_{2})^{-2}\mathscr{S}_{3}\phi_{+},\phi_{+}\rangle (3.5)

We note that I>𝒮3(𝒮2)2𝒮3>𝒮3(λ𝒮2)2𝒮3I>\mathscr{S}_{3}^{*}(-\mathscr{S}_{2})^{-2}\mathscr{S}_{3}>\mathscr{S}_{3}^{*}(\lambda-\mathscr{S}_{2})^{-2}\mathscr{S}_{3} for all λ0\lambda\geqslant 0. This indicates that the sign of the crossing form has to be positive whenever a crossing occurs at λ[0,C]\lambda\in[0,C]. In view of Equation (2.1), we conclude from Lemma that

Sf(Sλ;λ[0,C])=λ(0,C]dimkerSλ.\mathrm{Sf\,}(S_{\lambda};\lambda\in[0,C])=\sum\limits_{\lambda\in(0,C]}\dim\ker S_{\lambda}. (3.6)

For (2), a slightly modified argument shows that the sign of the crossing operator must be negative if a crossing occurs at λ[0,C]\lambda\in[0,C], and then

Sf(Sλ;λ[0,C])=λ[0,C]dimkerSλ.\mathrm{Sf\,}(S_{\lambda};\lambda\in[0,C])=-\sum\limits_{\lambda\in[0,C]}\dim\ker S_{\lambda}. (3.7)

This completes the proof. ∎

The aim of the next part is to prove some transversal properties about some invariant subspaces that are useful in our proof.

Lemma 3.3.

Under the conditions (H(H1)) and (H(H2)), we have that

Eλs(+)LR and Eλu()LR,E^{s}_{\lambda}(+\infty)\pitchfork L_{R}\text{ and }E^{u}_{\lambda}(-\infty)\pitchfork L_{R}, (3.8)

.

Proof.

We provide the proof of Eλs(+)LRE^{s}_{\lambda}(+\infty)\pitchfork L_{R} in completely similar fashion. Let [pq]Eλs(+)LR\begin{bmatrix}p\\ q\end{bmatrix}\in E^{s}_{\lambda}(+\infty)\cap L_{R}, and noting that Eλs(+)E^{s}_{\lambda}(+\infty) is invariant under JHλ(+)JH_{\lambda}(+\infty), then JHλ(+)[pq]Eλs(+)JH_{\lambda}(+\infty)\begin{bmatrix}p\\ q\end{bmatrix}\in E^{s}_{\lambda}(+\infty). From (H(H2)), a direct computation yields that

0\displaystyle 0 =ω(JHλ(+)[pq],[pq])=[Q12cQ12cQB+λQ1][pq],[pq]\displaystyle=\omega\left(JH_{\lambda}(+\infty)\begin{bmatrix}p\\ q\end{bmatrix},\begin{bmatrix}p\\ q\end{bmatrix}\right)=-\left\langle\begin{bmatrix}Q&\frac{1}{2}cQ\\ \frac{1}{2}cQ&B_{+}-\lambda Q^{-1}\end{bmatrix}\begin{bmatrix}p\\ q\end{bmatrix},\begin{bmatrix}p\\ q\end{bmatrix}\right\rangle (3.9)
=[p12cq12cp+λq+B+q],[pq]=p,pλq,q+QB+q,q0,\displaystyle=-\left\langle\begin{bmatrix}p-\frac{1}{2}cq\\ \frac{1}{2}cp+\lambda q+B_{+}q\end{bmatrix},\begin{bmatrix}p\\ q\end{bmatrix}\right\rangle=-\left\langle p,p\right\rangle-\lambda\left\langle q,q\right\rangle+\left\langle QB_{+}q,q\right\rangle\leqslant 0, (3.10)

we determine that Eλs(+)LRE^{s}_{\lambda}(+\infty)\pitchfork L_{R}. This completes the proof. ∎

Now, following Lemma 4.4 and Equation (4.3), we have that

ιCLM(Eλs(+),Eλu();λ[0,C])=\displaystyle\operatorname{\iota^{\scriptscriptstyle{\mathrm{CLM}}}}\left(E^{s}_{\lambda}(+\infty),E^{u}_{\lambda}(-\infty);\lambda\in[0,C]\right)= ι(ECu(),ECs(+);LR)ι(Eu(),Es(+);LR)\displaystyle\operatorname{\iota}\left(E^{u}_{C}(-\infty),E^{s}_{C}(+\infty);L_{R}\right)-\operatorname{\iota}\left(E^{u}(-\infty),E^{s}(+\infty);L_{R}\right) (3.11)
=\displaystyle= m+(𝔔(ECu(),ECs(+);LR))m+(𝔔(Eu(),Es(+);LR)).\displaystyle m^{+}(\mathfrak{Q}\left(E_{C}^{u}(-\infty),E_{C}^{s}(+\infty);L_{R}\right))-m^{+}(\mathfrak{Q}\left(E^{u}(-\infty),E^{s}(+\infty);L_{R}\right)). (3.12)

We recall that a Lagrangian frame for a Lagrangian subspace LL is an injective linear map T:n2nT:\mathbb{R}^{n}\to\mathbb{R}^{2n} whose image is LL (Cf. page 828 of [17]). Such a frame has the form

T=[XY],T=\begin{bmatrix}X\\ Y\end{bmatrix},

where XX and YY are both n×nn\times n matrices and XTQY=YTQXX^{T}QY=Y^{T}QX.

We introduce the following notations Tλ+=[Xλ,+Yλ,+0I2I10Yλ,+TZλ,+]T_{\lambda}^{+}=\begin{bmatrix}X_{\lambda,+}&Y_{\lambda,+}\\ 0&I_{2}\\ I_{1}&0\\ Y^{T}_{\lambda,+}&Z_{\lambda,+}\end{bmatrix}, Tλ=[Xλ,Yλ,0I2I10Yλ,TZλ,]T_{\lambda}^{-}=\begin{bmatrix}X_{\lambda,-}&Y_{\lambda,-}\\ 0&I_{2}\\ I_{1}&0\\ Y^{T}_{\lambda,-}&Z_{\lambda,-}\end{bmatrix}, Mλ,+=[Xλ,+Yλ,+TYλ,+Zλ,+]M_{\lambda,+}=\begin{bmatrix}X_{\lambda,+}&Y^{T}_{\lambda,+}\\ Y_{\lambda,+}&Z_{\lambda,+}\end{bmatrix} and Mλ,=[Xλ,Yλ,TYλ,Zλ,]M_{\lambda,-}=\begin{bmatrix}X_{\lambda,-}&Y_{\lambda,-}^{T}\\ Y_{\lambda,-}&Z_{\lambda,-}\end{bmatrix}, where Xλ,±X_{\lambda,\pm} are r×rr\times r matrices, Yλ,±Y_{\lambda,\pm} are (nr)×r(n-r)\times r matrices, Zλ,±Z_{\lambda,\pm} are (nr)×(nr)(n-r)\times(n-r) matrices, I1I_{1} is the r×rr\times r identity matrix, and I2I_{2} is the (nr)×(nr)(n-r)\times(n-r) identity matrix. Here, r=dimV+(Q)r=\dim V^{+}(Q). For each λ+¯\lambda\in\overline{\mathbb{R}^{+}}, from Lemma 3.3, we can use notations Tλ,+T_{\lambda,+} and Tλ,T_{\lambda,-} for the Lagrangian frames of Eλs(+)E^{s}_{\lambda}(+\infty) and Eλu()E^{u}_{\lambda}(-\infty), respectively.

We now consider the following operator:

Fλ+JddξHλ(+)F^{+}_{\lambda}\coloneqq-J\frac{\operatorname{d}}{\operatorname{d}\xi}-H_{\lambda}(+\infty) (3.13)

and the associated second order operator 𝕃λ+\mathbb{L}_{\lambda}^{+} and let φ\varphi be a solution of 𝕃λ+φ=0\mathbb{L}_{\lambda}^{+}\varphi=0, where 𝕃λ,M+\mathbb{L}_{\lambda,M}^{+} denotes the operator 𝕃λ+\mathbb{L}_{\lambda}^{+} defined on the maximal domain W2,2(+¯,n)W^{2,2}(\overline{\mathbb{R}^{+}},\mathbb{R}^{n}). Then, the map ϕ(ϕ˙T(0)12cϕT(0),ϕT(0))T\phi\mapsto(\dot{\phi}^{T}(0)-\frac{1}{2}c\phi^{T}(0),\phi^{T}(0))^{T} provides a linear bijection from ker𝕃λ,M+\ker\mathbb{L}_{\lambda,M}^{+} to Eλs(+)=V(JHλ(+))E^{s}_{\lambda}(+\infty)=V^{-}(JH_{\lambda}(+\infty)).

We note that for each zEλs(+)z\in E^{s}_{\lambda}(+\infty), there exists u=[pq]V+(Q)V(Q)nu=\begin{bmatrix}p\\ q\end{bmatrix}\in V^{+}(Q)\oplus V^{-}(Q)\cong\mathbb{R}^{n}, such that z=Tλ,+(u)z=T_{\lambda,+}(u).

Let φ(ξ)ker𝕃λ+\varphi(\xi)\in\ker\mathbb{L}_{\lambda}^{+} with (ϕ˙T(0)12cϕT(0),ϕT(0))T=Tλ+uEλs(+)(\dot{\phi}^{T}(0)-\frac{1}{2}c\phi^{T}(0),\phi^{T}(0))^{T}=T^{+}_{\lambda}u\in E^{s}_{\lambda}(+\infty), where u=[pq]V+(Q)V(Q)nu=\begin{bmatrix}p\\ q\end{bmatrix}\in V^{+}(Q)\oplus V^{-}(Q)\cong\mathbb{R}^{n}. A simple calculation shows that

0\displaystyle 0 =𝕃λ+φ(ξ),φ(ξ)L2=φ˙,φ˙L2cφ,φ˙L2+c24φL22+λφL22QB+φ,φL2+φ˙(0)c2φ(0),φ(0)\displaystyle=\langle\mathbb{L}_{\lambda}^{+}\varphi(\xi),\varphi(\xi)\rangle_{L^{2}}=\langle\dot{\varphi},\dot{\varphi}\rangle_{L^{2}}-c\langle\varphi,\dot{\varphi}\rangle_{L^{2}}+\frac{c^{2}}{4}\left\|\varphi\right\|^{2}_{L^{2}}+\lambda\left\|\varphi\right\|^{2}_{L^{2}}-\langle QB_{+}\varphi,\varphi\rangle_{L^{2}}+\langle\dot{\varphi}(0)-\frac{c}{2}\varphi(0),\varphi(0)\rangle (3.14)
=φ˙c2φL220+QB+φ,φdξ+λφL22+Mλ,+[pq],[pq]\displaystyle=\left\|\dot{\varphi}-\frac{c}{2}\varphi\right\|^{2}_{L^{2}}-\int_{0}^{+\infty}\left\langle QB_{+}\varphi,\varphi\right\rangle\operatorname{d}\xi+\lambda\left\|\varphi\right\|^{2}_{L^{2}}+\left\langle M_{\lambda,+}\begin{bmatrix}p\\ q\end{bmatrix},\begin{bmatrix}p\\ q\end{bmatrix}\right\rangle (3.15)
φ˙c2φL22+(λC)φL22+Mλ,+u,u,\displaystyle\geqslant\left\|\dot{\varphi}-\frac{c}{2}\varphi\right\|^{2}_{L^{2}}+(\lambda-C)\left\|\varphi\right\|^{2}_{L^{2}}+\left\langle M_{\lambda,+}u,u\right\rangle, (3.16)

this equation, together with (H(H1)), (H(H2)) and (H(H2’)), shows the following Lemma:

Lemma 3.4.

With CC given in (1.3), the following results hold:
(1) if (H(H1)) and (H(H2)) hold, we have that Mλ,+(s)M_{\lambda,+}(s) is negative definite for all λC\lambda\geqslant C.
(2) if (H(H2’)) holds, we have that Mλ,+(s)M_{\lambda,+}(s) is negative definite for all λ0\lambda\geqslant 0.

Similarly, we have that

Lemma 3.5.

With CC given in (1.3), the following results hold:
(1) if (H(H1)) and (H(H2)) hold, we have that Mλ,M_{\lambda,-} is positive definite for all λC\lambda\geqslant C.
(2) if (H(H2’)) holds, we have that Mλ,M_{\lambda,-} is positive definite for all λ0\lambda\geqslant 0.

Letting zEλu()z\in E^{u}_{\lambda}(-\infty), then there is u=[pq]V+(Q)V(Q)nu=\begin{bmatrix}p\\ q\end{bmatrix}\in V^{+}(Q)\oplus V^{-}(Q)\cong\mathbb{R}^{n}, such that z=Tλ,+uz=T_{\lambda,+}u, and zz can be rewritten as z=Tλ,+u+(Tλ,Tλ,+)uEλs+LRz=T_{\lambda,+}u+(T_{\lambda,-}-T_{\lambda,+})u\in E^{s}_{\lambda}+L_{R}. From a simple calculation, we have that

𝔔(Eλu(),Eλs(+);LR)(u,u)\displaystyle\mathfrak{Q}\left(E^{u}_{\lambda}(-\infty),E^{s}_{\lambda}(+\infty);L_{R}\right)(u,u) (3.17)
=\displaystyle= ω([Xλ,+p+Yλ,+TqqpYλ,+p+Zλ,+q],[(Xλ,Xλ,+)p+(Yλ,TYλ,+T)q00(Yλ,Yλ,+)p+(Zλ,Zλ,+)q])\displaystyle\omega\left(\begin{bmatrix}X_{\lambda,+}p+Y^{T}_{\lambda,+}q\\ q\\ p\\ Y_{\lambda,+}p+Z_{\lambda,+}q\end{bmatrix},\begin{bmatrix}\left(X_{\lambda,-}-X_{\lambda,+}\right)p+\left(Y^{T}_{\lambda,-}-Y^{T}_{\lambda,+}\right)q\\ 0\\ 0\\ \left(Y_{\lambda,-}-Y_{\lambda,+}\right)p+\left(Z_{\lambda,-}-Z_{\lambda,+}\right)q\end{bmatrix}\right) (3.18)
=\displaystyle= (Xλ,+Xλ,)p,p+2(Yλ,+TYλ,T)p,q+(Zλ,+Zλ,)q,q\displaystyle\left\langle(X_{\lambda,+}-X_{\lambda,-})p,p\right\rangle+2\left\langle(Y^{T}_{\lambda,+}-Y^{T}_{\lambda,-})p,q\right\rangle+\left\langle(Z_{\lambda,+}-Z_{\lambda,-})q,q\right\rangle (3.19)
=\displaystyle= [Xλ,+Xλ,Yλ,+TYλ,TYλ,+Yλ,Zλ,+Zλ,][pq],[pq]=(Mλ,+Mλ,)u,u.\displaystyle\left\langle\begin{bmatrix}X_{\lambda,+}-X_{\lambda,-}&Y^{T}_{\lambda,+}-Y^{T}_{\lambda,-}\\ Y_{\lambda,+}-Y_{\lambda,-}&Z_{\lambda,+}-Z_{\lambda,-}\end{bmatrix}\begin{bmatrix}p\\ q\end{bmatrix},\begin{bmatrix}p\\ q\end{bmatrix}\right\rangle=\left\langle(M_{\lambda,+}-M_{\lambda,-})u,u\right\rangle. (3.20)

From Lemma 3.4, Lemma 3.5, Equation (3.17) and Equation (4.3), the following result holds.

Lemma 3.6.

With CC given in (1.3), the following results hold:
(1) if (H(H1)) and (H(H2)) hold, then we have that

ι(Eλu(),Eλs(+),LR)=0for λC.\operatorname{\iota}\left(E^{u}_{\lambda}(-\infty),E^{s}_{\lambda}(+\infty),L_{R}\right)=0\ \text{for }\lambda\geqslant C. (3.21)

(2) if (H(H2’)) holds, then we have that

ι(Eλu(),Eλs(+),LR)=0for λ0.\operatorname{\iota}\left(E^{u}_{\lambda}(-\infty),E^{s}_{\lambda}(+\infty),L_{R}\right)=0\ \text{for }\lambda\geqslant 0. (3.22)

Before finishing the preparation of our proof of our main results, we recall the definition of positive curve.

Definition 3.7.

[15] Let A:[0,1]𝒞sa(E)A:[0,1]\rightarrow\mathscr{CF}^{sa}\left(E\right) be a continuous curve. The curve AA is named a positive curve if {t|kerAλ0}\Set{t}{\ker A_{\lambda}\neq 0} is finite and

Sf(At;t[0,1])=0<t1dimkerAt.\mathrm{Sf\,}(A_{t};t\in[0,1])=\sum_{0<t\leqslant 1}\dim\ker A_{t}.
The Proof of Proposition 1.7.

For some aa\in\mathbb{R}, we construct the following homotopy Lagrangian path

(Es(τ+sa),Eu(τ+sa)),(τ,s)+×[0,1].\displaystyle\left(E^{s}(\tau+sa),E^{u}(-\tau+sa)\right),(\tau,s)\in\mathbb{R}^{+}\times[0,1]. (3.23)

We point out that dim(Es(sa)Eu(sa))\dim\left(E^{s}(sa)\cap E^{u}(sa)\right) is constant for all s[0,1]s\in[0,1] and Es(+)Eu()E^{s}(+\infty)\pitchfork E^{u}(-\infty).

By the stratum homotopy invariance property of the Maslov index, we have that

ιCLM(Es(τ),Eu(τ);τ+)=ιCLM(Es(τ+a),Eu(τ+a);τ+)\displaystyle\operatorname{\iota^{\scriptscriptstyle{\mathrm{CLM}}}}\left(E^{s}(\tau),E^{u}(-\tau);\tau\in\mathbb{R}^{+}\right)=\operatorname{\iota^{\scriptscriptstyle{\mathrm{CLM}}}}\left(E^{s}(\tau+{a}),E^{u}(-\tau+{a});\tau\in\mathbb{R}^{+}\right) (3.24)
=\displaystyle= ιCLM(Es(τ+2a),Eu(τ);τ[a,+)).\displaystyle\operatorname{\iota^{\scriptscriptstyle{\mathrm{CLM}}}}\left(E^{s}(\tau+2{a}),E^{u}(-\tau);\tau\in[-{a},+\infty)\right). (3.25)

We know that Es(+)Eu()E^{s}(+\infty)\pitchfork E^{u}(-\infty) and Es(τ)Es(+)E^{s}(\tau)\to E^{s}(+\infty) as τ+\tau\to+\infty under the gap topology of the Lagrangian Grassmannian, so we can choose τ0\tau_{0}, such that Es(τ)Eu()E^{s}(\tau)\pitchfork E^{u}(-\infty) for all ττ0\tau\geqslant\tau_{0}, and the path Eu(τ):(,τ0]Lag(n)E^{u}(\tau):(-\infty,\tau_{0}]\to\mathrm{Lag}(n) has only regular crossing with respect to Es(τ0)E^{s}(\tau_{0}). Letting a=τ0{a}=\tau_{0}, we construct the following homotopy Lagrangian path:

(Es(τ0+s(τ0+τ)),Eu(τ)),(τ,s)[τ0,+)×[0,1].\left(E^{s}(\tau_{0}+s(\tau_{0}+\tau)),E^{u}(-\tau)\right),\ (\tau,s)\in[-\tau_{0},+\infty)\times[0,1].

By the stratum homotopy invariance, reversal property of Maslov index and Equation (1.11), we determine that

ιCLM(Es(τ+2τ0),Eu(τ);τ[τ0,+))=ιCLM(Es(τ0),Eu(τ);τ[τ0,+))\displaystyle\operatorname{\iota^{\scriptscriptstyle{\mathrm{CLM}}}}\left(E^{s}(\tau+2\tau_{0}),E^{u}(-\tau);\tau\in[-\tau_{0},+\infty)\right)=\operatorname{\iota^{\scriptscriptstyle{\mathrm{CLM}}}}\left(E^{s}(\tau_{0}),E^{u}(-\tau);\tau\in[-\tau_{0},+\infty)\right) (3.26)
=ιCLM(Es(τ0),Eu(τ);τ(,τ0])=Maslov(w),\displaystyle=-\operatorname{\iota^{\scriptscriptstyle{\mathrm{CLM}}}}\left(E^{s}(\tau_{0}),E^{u}(\tau);\tau\in(-\infty,\tau_{0}]\right)=-\mathrm{Maslov}(w^{*}), (3.27)

and this, together with Equation (3.24), completes the proof.

The Proof of Theorem 1.8.

We first prove that Sf(Sλ;λ[0,C])=Sf(Fλ;λ[0,C])\mathrm{Sf\,}(S_{\lambda};\lambda\in[0,C])=\mathrm{Sf\,}(F_{\lambda};\lambda\in[0,C]). We start by introducing the continuous map

f:𝒞sa(L2(,n))𝒞sa(L2(,2n)) defined by f(Sλ)Fλ.f:\mathscr{CF}^{sa}\left(L^{2}\left(\mathbb{R},\mathbb{R}^{n}\right)\right)\rightarrow\mathscr{CF}^{sa}\left(L^{2}\left(\mathbb{R},\mathbb{R}^{2n}\right)\right)\textrm{ defined by }f(S_{\lambda})\coloneqq F_{\lambda}.

Let h(λ,s)=f(Sλ+sI)h(\lambda,s)=f(S_{\lambda}+sI) for (λ,s)[0,C]×[0,ϵ](\lambda,s)\in[0,C]\times[0,\epsilon]. Then, for every λ[0,C]\lambda\in[0,C], h(λ,s)h(\lambda,s) is a positive curve. Let λ0[0,C]\lambda_{0}\in[0,C] be a crossing instant for the path λSλ\lambda\mapsto S_{\lambda}, meaning that kerSλ0{0}\ker S_{\lambda_{0}}\neq\{0\}, and let us consider the positive path sSλ0+sIs\mapsto S_{\lambda_{0}}+sI. Thus, there exists δ>0\delta>0, such that ker(Sλ0+δI)={0}\ker\left(S_{\lambda_{0}}+\delta I\right)=\{0\}, which is equivalent to kerh(λ0,δ)={0}\ker h(\lambda_{0},\delta)=\{0\}. Since Sλ0+δIS_{\lambda_{0}}+\delta I is a Fredholm operator, then there exists δ1>0\delta_{1}>0, such that ker(Sλ+δI)={0}\ker(S_{\lambda}+\delta I)=\{0\} for every λ[λ0δ1,λ0+δ1]\lambda\in[\lambda_{0}-\delta_{1},\lambda_{0}+\delta_{1}]. By this argument, we determine that kerh(λ,δ)={0}\ker h(\lambda,\delta)=\{0\} for every λ[λ0δ1,λ0+δ1]\lambda\in[\lambda_{0}-\delta_{1},\lambda_{0}+\delta_{1}]. Then, we determine that

{Sf(Sλ+δI,λ[λ0δ1,λ0+δ1])=0,Sf(h(λ,δ),λ[λ0δ1,λ0+δ1])=0.\begin{cases}\mathrm{Sf\,}(S_{\lambda}+\delta I,\lambda\in[\lambda_{0}-\delta_{1},\lambda_{0}+\delta_{1}])=0,\\ \mathrm{Sf\,}(h(\lambda,\delta),\lambda\in[\lambda_{0}-\delta_{1},\lambda_{0}+\delta_{1}])=0.\end{cases} (3.28)

By the homotopy invariance of the spectral flow, we infer that

Sf(Sλ,λ[λ0δ1,λ0+δ1])=Sf(Sλ0δ1+sI,s[0,δ])Sf(Sλ0+δ1+sI,s[0,δ])\mathrm{Sf\,}(S_{\lambda},\lambda\in[\lambda_{0}-\delta_{1},\lambda_{0}+\delta_{1}])=\mathrm{Sf\,}(S_{\lambda_{0}-\delta_{1}}+sI,s\in[0,\delta])-\mathrm{Sf\,}(S_{\lambda_{0}+\delta_{1}}+sI,s\in[0,\delta]) (3.29)

and

Sf(h(λ,0),λ[λ0δ1,λ0+δ1])=Sf(h(λ0δ1,s),s[0,δ])Sf(h(λ0δ1,s),s[0,δ])\mathrm{Sf\,}(h(\lambda,0),\lambda\in[\lambda_{0}-\delta_{1},\lambda_{0}+\delta_{1}])=\mathrm{Sf\,}(h(\lambda_{0}-\delta_{1},s),s\in[0,\delta])-\mathrm{Sf\,}(h(\lambda_{0}-\delta_{1},s),s\in[0,\delta]) (3.30)

We observe that sSλ0±δ1+sIs\mapsto S_{\lambda_{0}\pm\delta_{1}}+sI and sh(λ0±δ1,s)s\mapsto h(\lambda_{0}\pm\delta_{1},s) are both positive curves. It follows that

Sf(Sλ0±δ1+sI,s[0,δ])\displaystyle\mathrm{Sf\,}(S_{\lambda_{0}\pm\delta_{1}}+sI,s\in[0,\delta]) =0<sδdimker(Sλ0±δ1+sI)=0<sδdimkerh(λ0±δ1,s)\displaystyle=\sum_{0<s\leqslant\delta}\dim\ker\left(S_{\lambda_{0}\pm\delta_{1}}+sI\right)=\sum_{0<s\leqslant\delta}\dim\ker h(\lambda_{0}\pm\delta_{1},s) (3.31)
=Sf(h(λ0±δ1,s),s[0,δ])\displaystyle=\mathrm{Sf\,}(h(\lambda_{0}\pm\delta_{1},s),s\in[0,\delta]) (3.32)

from Equations (3.29), (3.30) and (3.31), we have that

Sf(Sλ,λ[λ0δ1,λ0+δ1])=Sf(h(λ,0),λ[λ0δ1,λ0+δ1])\displaystyle\mathrm{Sf\,}(S_{\lambda},\lambda\in[\lambda_{0}-\delta_{1},\lambda_{0}+\delta_{1}])=\mathrm{Sf\,}(h(\lambda,0),\lambda\in[\lambda_{0}-\delta_{1},\lambda_{0}+\delta_{1}]) (3.33)
=\displaystyle= Sf(Fλ,λ[λ0δ1,λ0+δ1]).\displaystyle\mathrm{Sf\,}({F}_{\lambda},\lambda\in[\lambda_{0}-\delta_{1},\lambda_{0}+\delta_{1}]). (3.34)

the crossing instants are isolated, and those on a compact interval are finite in number. From Equation (3.33) and the path additivity of spectral flow, we determine that Sf(Sλ;λ[0,C])=Sf(Fλ;λ[0,C])\mathrm{Sf\,}(S_{\lambda};\lambda\in[0,C])=\mathrm{Sf\,}(F_{\lambda};\lambda\in[0,C]).

From Proposition 2.12 and Equation (3.11), we determine that

Sf(Sλ;λ[0,C])=ιCLM(Es(τ),Eu(τ);τ+¯)ι(Eu(),Es(+);LR).\displaystyle-\mathrm{Sf\,}(S_{\lambda};\lambda\in[0,C])=\operatorname{\iota^{\scriptscriptstyle{\mathrm{CLM}}}}\left(E^{s}(\tau),E^{u}(-\tau);\tau\in\overline{\mathbb{R}^{+}}\right)-\operatorname{\iota}\left(E^{u}(-\infty),E^{s}(+\infty);L_{R}\right). (3.35)

Since (H(H1)) and (H(H2)) hold, from Lemma 3.6, Equation (3.35) and Proposition 2.1, then

|ι(w)+ι(Eu(),Es(+);LR)|N¯+(L),|\operatorname{\iota}(w^{*})+\operatorname{\iota}\left(E^{u}(-\infty),E^{s}(+\infty);L_{R}\right)|\leqslant\overline{N}_{+}(L), (3.36)

The proof of Theorem 1.10.

From Remark 1.2, (H(H2’)) implies that (H(H1)) holds, and then, from Theorem 1.8 and Lemma 3.6, we have that

|ι(w)|N¯+(L).|\operatorname{\iota}(w^{*})|\leqslant\overline{N}_{+}(L). (3.37)

The proof of Theorem 1.12.

We note that S=[d2dt2+c24f(u)d1d1dd2dt2c24γ]S=\begin{bmatrix}-\frac{\operatorname{d}^{2}}{\operatorname{d}t^{2}}+\frac{c^{2}}{4}-\frac{f^{\prime}(u^{*})}{d}&\frac{1}{\sqrt{d}}\\ \frac{1}{\sqrt{d}}&\frac{\operatorname{d}^{2}}{\operatorname{d}t^{2}}-\frac{c^{2}}{4}-\gamma\end{bmatrix}, and then, 𝒮3=1d\mathscr{S}_{3}=\frac{1}{\sqrt{d}}, 𝒮2=d2dt2c24γ\mathscr{S}_{2}=\frac{\operatorname{d}^{2}}{\operatorname{d}t^{2}}-\frac{c^{2}}{4}-\gamma. Since γ>0\gamma>0, it is easy to see that 𝒮2>0\mathscr{S}_{2}>0. Moreover, if d>γ2d>\gamma^{-2}, then we have that I>1d(d2dt2+c24+γ)2=𝒮3(𝒮2)2𝒮3I>\frac{1}{d}\left(-\frac{\operatorname{d}^{2}}{\operatorname{d}t^{2}}+\frac{c^{2}}{4}+\gamma\right)^{-2}=\mathscr{S}^{*}_{3}(-\mathscr{S}_{2})^{-2}\mathscr{S}_{3}, so Proposition 3.2 (1) holds, and then, we have that Sf(Sλ;λ[0,C])=N+(L)\mathrm{Sf\,}(S_{\lambda};\lambda\in[0,C])=N_{+}(L). Moreover, by a simple calculation and the facts f(0)<0f^{\prime}(0)<0 and f(u3)<0f^{\prime}(u_{3})<0, it is easy to check that the condition (H(H2’)) holds. Then, from Equation (3.35) and Lemma 3.6, we complete the proof. ∎

4. The triple and Hörmander index

Recently, Zhu et al., in the interesting paper [19], deeply investigated the Hörmander index, particularly its relation with respect to the so-called triple index in a slightly generalized (in fact, isotropic) setting. Given three isotropic subspaces α,β\alpha,\beta and δ\delta in (2n,ω)(\mathbb{R}^{2n},\omega), we define the quadratic form 𝔔\mathfrak{Q} as follows:

𝔔𝔔(α,β;δ):α(β+δ) given by 𝔔(x1,x2)=ω(y1,z2),\mathfrak{Q}\coloneqq\mathfrak{Q}(\alpha,\beta;\delta):\alpha\cap(\beta+\delta)\to\mathbb{R}\quad\textrm{ given by }\quad\mathfrak{Q}(x_{1},x_{2})=\omega(y_{1},z_{2}), (4.1)

where for j=1,2j=1,2, xj=yj+zjα(β+δ)x_{j}=y_{j}+z_{j}\in\alpha\cap(\beta+\delta) and yjβy_{j}\in\beta, zjδz_{j}\in\delta. By invoking [19, Lemma 3.3], in the particular case in which α,β,δ\alpha,\beta,\delta are Lagrangian subspaces, we obtain

ker𝔔(α,β;δ)=αβ+αδ.\ker\mathfrak{Q}(\alpha,\beta;\delta)=\alpha\cap\beta+\alpha\cap\delta. (4.2)

By [19, Lemma 3.13], we are in position to define the triple index in terms of the quadratic form 𝔔\mathfrak{Q} defined above.

Definition 4.1.

Let α,β\alpha,\beta and κ\kappa be three Lagrangian subspaces of symplectic vector space (2n,ω)(\mathbb{R}^{2n},\omega). Then, the triple index of the triple (α,β,κ)(\alpha,\beta,\kappa) is defined by

ι(α,β,κ)=m+(𝔔(α,β;κ))+dim(ακ)dim(αβκ)).\iota(\alpha,\beta,\kappa)=m^{+}\big{(}\mathfrak{Q}(\alpha,\beta;\kappa)\big{)}+\dim\big{(}\alpha\cap\kappa)-\dim(\alpha\cap\beta\cap\kappa)\big{)}. (4.3)

where m+m^{+} is the Morse positive index of a quadratic form Q.

Another closely related symplectic invariant is the so-called Hörmander index, which is particularly important for measuring the difference in the (relative) Maslov index computed with respect to two different Lagrangian subspaces (we refer the interested reader to the celebrated and beautiful paper [17] and the references therein).

Let V0,V1,L0,L1V_{0},V_{1},L_{0},L_{1} be four Lagrangian subspaces and L𝒞0([0,1],Lag(n))L\in\mathscr{C}^{0}\big{(}[0,1],\mathrm{Lag}(n)\big{)} be such that L(0)=L0L(0)=L_{0} and L(1)=L1L(1)=L_{1}.

Definition 4.2.

Letting L,V𝒞0([0,1],Lag(n))L,\ V\in\mathscr{C}^{0}([0,1],\mathrm{Lag}(n)) be such that L(0)=L0,L(1)=L1,V(0)=V0L(0)=L_{0},\ L(1)=L_{1},\ V(0)=V_{0} and V(1)=V1V(1)=V_{1}, the Hörmander index is the integer defined by

s(L0,L1;V0,V1)\displaystyle s(L_{0},L_{1};V_{0},V_{1}) =ιCLM(V1,L(t);t[0,1])ιCLM(V0,L(t));t[0,1]\displaystyle=\operatorname{\iota^{\scriptscriptstyle{\mathrm{CLM}}}}(V_{1},L(t);t\in[0,1])-\operatorname{\iota^{\scriptscriptstyle{\mathrm{CLM}}}}(V_{0},L(t));t\in[0,1] (4.4)
=ιCLM(V(t),L1;t[0,1])ιCLM(V(t),L0;t[0,1])\displaystyle=\operatorname{\iota^{\scriptscriptstyle{\mathrm{CLM}}}}(V(t),L_{1};t\in[0,1])-\operatorname{\iota^{\scriptscriptstyle{\mathrm{CLM}}}}(V(t),L_{0};t\in[0,1]) (4.5)
Remark 4.3.

As a direct consequence of the fixed endpoints homotopy invariance of the ιCLM\operatorname{\iota^{\scriptscriptstyle{\mathrm{CLM}}}}-index, it is actually possible to prove that Definition 4.2 is well-posed, meaning that it is independent of the path LL joining the two Lagrangian subspaces L0,L1L_{0},L_{1}. (Cf. [17] for further details).

Let us now be given four Lagrangian subspaces, namely λ1,λ2,κ1,κ2\lambda_{1},\lambda_{2},\kappa_{1},\kappa_{2} of symplectic vector space (2n,ω)(\mathbb{R}^{2n},\omega). By [19, Theorem 1.1], the Hörmander index s(λ1,λ2;κ1,κ2)s(\lambda_{1},\lambda_{2};\kappa_{1},\kappa_{2}) can be expressed in terms of the triple index as follows

s(λ1,λ2;κ1,κ2)=ι(λ1,λ2,κ2)ι(λ1,λ2,κ1)=ι(λ1,κ1,κ2)ι(λ2,κ1,κ2).s(\lambda_{1},\lambda_{2};\kappa_{1},\kappa_{2})=\iota(\lambda_{1},\lambda_{2},\kappa_{2})-\iota(\lambda_{1},\lambda_{2},\kappa_{1})=\iota(\lambda_{1},\kappa_{1},\kappa_{2})-\iota(\lambda_{2},\kappa_{1},\kappa_{2}). (4.6)
Lemma 4.4.

[14] Let L1(t)L_{1}(t) and L2(t)L_{2}(t) be two paths in Lag(n)\mathrm{Lag}(n) with t[0,1]t\in[0,1], and assume that L1(t)L_{1}(t) and L2(t)L_{2}(t) are both transversal to the (fixed) Lagrangian subspace LL. We then obtain

ιCLM(L1(t),L2(t);t[0,1])=ι(L2(1),L1(1);L)ι(L2(0),L1(0);L).\operatorname{\iota^{\scriptscriptstyle{\mathrm{CLM}}}}\big{(}L_{1}(t),L_{2}(t);t\in[0,1]\big{)}=\iota\big{(}L_{2}(1),L_{1}(1);L)-\iota(L_{2}(0),L_{1}(0);L).

Acknowledgments

The author is grateful to Professors Xijun Hu and Li Wu for their helpful discussion during the preparation of this article. Finally, the author thanks the anonymous referee for reading the paper carefully and providing thoughtful comments.

References

  • [1] Alberto Abbondandolo and Pietro Majer “Ordinary Differential Operators in Hilbert Spaces and Fredholm Pairs” In Math. Z. 243.3, 2003, pp. 525–562 DOI: 10.1007/s00209-002-0473-z
  • [2] J. Alexander, R. Gardner and C. Jones “A Topological Invariant Arising in the Stability Analysis of Travelling Waves” In J. Reine Angew. Math. 410, 1990, pp. 167–212
  • [3] M.. Atiyah, V.. Patodi and I.. Singer “Spectral Asymmetry and Riemannian Geometry. III” In Math. Proc. Cambridge Philos. Soc. 79.1, 1976, pp. 71–99
  • [4] Sylvain E. Cappell, Ronnie Lee and Edward Y. Miller “On the Maslov Index” In Comm. Pure Appl. Math. 47.2, 1994, pp. 121–186 DOI: 10.1002/cpa.3160470202
  • [5] Chao-Nien Chen, Y S Choi and Xijun Hu “An Index Method for Stability Analysis of Traveling and Standing Waves” In preparation, pp. 21
  • [6] Chao-Nien Chen and Y.. Choi “Front Propagation in Both Directions and Coexistence of Traveling Fronts and Pulses”, 2018 arXiv:1807.01832 [math]
  • [7] Chao-Nien Chen and Xijun Hu “Maslov Index for Homoclinic Orbits of Hamiltonian Systems” In Ann. Inst. H. Poincaré Anal. Non Linéaire 24.4, 2007, pp. 589–603 DOI: 10.1016/j.anihpc.2006.06.002
  • [8] Chao-Nien Chen and Xijun Hu “Stability Analysis for Standing Pulse Solutions to FitzHugh-Nagumo Equations” In Calc. Var. Partial Differential Equations 49.1-2, 2014, pp. 827–845 DOI: 10.1007/s00526-013-0601-0
  • [9] Paul Cornwell “Opening the Maslov Box for Traveling Waves in Skew-Gradient Systems: Counting Eigenvalues and Proving (in)Stability” In Indiana Univ. Math. J. 68.6, 2019, pp. 1801–1832 DOI: 10.1512/iumj.2019.68.7831
  • [10] Paul Cornwell and Christopher K… Jones “A Stability Index for Travelling Waves in Activator-Inhibitor Systems” In Proc. Roy. Soc. Edinburgh Sect. A 150.1, 2020, pp. 517–548 DOI: 10.1017/prm.2018.92
  • [11] Paul Cornwell and Christopher K… Jones “On the Existence and Stability of Fast Traveling Waves in a Doubly Diffusive FitzHugh-Nagumo System” In SIAM J. Appl. Dyn. Syst. 17.1, 2018, pp. 754–787 DOI: 10.1137/17M1149432
  • [12] P. Howard and A. Sukhtayev “The Maslov and Morse Indices for Schrödinger Operators on [0,1][0,1] In J. Differential Equations 260.5, 2016, pp. 4499–4549 DOI: 10.1016/j.jde.2015.11.020
  • [13] Xijun Hu and Alessandro Portaluri “Index Theory for Heteroclinic Orbits of Hamiltonian Systems” In Calc. Var. Partial Differential Equations 56.6, 2017, pp. Paper No. 167\bibrangessep24 DOI: 10.1007/s00526-017-1259-9
  • [14] Xijun Hu, Alessandro Portaluri, Li Wu and Qin Xing “Morse Index Theorem for Heteroclinic Orbits of Lagrangian Systems”, 2020 arXiv:2004.08643 [math]
  • [15] Xijun Hu and Li Wu “Decomposition of Spectral Flow and Bott-Type Iteration Formula” In Electronic Research Archive 28.1 American Institute of Mathematical Sciences, 2020, pp. 127 DOI: 10.3934/era.2020008
  • [16] Todd Kapitula and Keith Promislow “Spectral and Dynamical Stability of Nonlinear Waves” 185, Applied Mathematical Sciences New York, NY: Springer New York, 2013 DOI: 10.1007/978-1-4614-6995-7
  • [17] Joel Robbin and Dietmar Salamon “The Maslov Index for Paths” In Topology 32.4, 1993, pp. 827–844 DOI: 10.1016/0040-9383(93)90052-W
  • [18] Eiji Yanagida “Mini-Maximizers for Reaction-Diffusion Systems with Skew-Gradient Structure” In J. Differential Equations 179.1, 2002, pp. 311–335 DOI: 10.1006/jdeq.2001.4028
  • [19] Yuting Zhou, Li Wu and Chaofeng Zhu “Hörmander Index in Finite-Dimensional Case” In Front. Math. China 13.3, 2018, pp. 725–761 DOI: 10.1007/s11464-018-0702-3
  • [20] Chaofeng Zhu and Yiming Long “Maslov-Type Index Theory for Symplectic Paths and Spectral Flow. I” In Chinese Ann. Math. Ser. B 20.4, 1999, pp. 413–424 DOI: 10.1142/S0252959999000485