Let A(t) be a continuous path of self-adjoint Fredholm operators, we derive a decomposition formula of spectral flow if the path is invariant under a matrix-like cogredient. As applications, we give the generalized Bott-type iteration formula for linear Hamiltonian systems.
Citation: Xijun Hu, Li Wu. Decomposition of spectral flow and Bott-type iteration formula[J]. Electronic Research Archive, 2020, 28(1): 127-148. doi: 10.3934/era.2020008
[1] | Xijun Hu, Li Wu . Decomposition of spectral flow and Bott-type iteration formula. Electronic Research Archive, 2020, 28(1): 127-148. doi: 10.3934/era.2020008 |
[2] | Hui Guo, Tao Wang . A note on sign-changing solutions for the Schrödinger Poisson system. Electronic Research Archive, 2020, 28(1): 195-203. doi: 10.3934/era.2020013 |
[3] | Ling-Xiong Han, Wen-Hui Li, Feng Qi . Approximation by multivariate Baskakov–Kantorovich operators in Orlicz spaces. Electronic Research Archive, 2020, 28(2): 721-738. doi: 10.3934/era.2020037 |
[4] | Jintao Li, Jindou Shen, Gang Xu . The global supersonic flow with vacuum state in a 2D convex duct. Electronic Research Archive, 2021, 29(2): 2077-2099. doi: 10.3934/era.2020106 |
[5] |
Hongyong Cui, Peter E. Kloeden, Wenqiang Zhao .
Strong |
[6] | Liqian Bai, Xueqing Chen, Ming Ding, Fan Xu . A generalized quantum cluster algebra of Kronecker type. Electronic Research Archive, 2024, 32(1): 670-685. doi: 10.3934/era.2024032 |
[7] | Jinguo Jiang . Algebraic Schouten solitons associated to the Bott connection on three-dimensional Lorentzian Lie groups. Electronic Research Archive, 2025, 33(1): 327-352. doi: 10.3934/era.2025017 |
[8] | Qing Zhao, Shuhong Chen . Regularity for very weak solutions to elliptic equations of $ p $-Laplacian type. Electronic Research Archive, 2025, 33(6): 3482-3495. doi: 10.3934/era.2025154 |
[9] | Kai Jiang, Wei Si . High-order energy stable schemes of incommensurate phase-field crystal model. Electronic Research Archive, 2020, 28(2): 1077-1093. doi: 10.3934/era.2020059 |
[10] | Mónica Clapp, Angela Pistoia . Yamabe systems and optimal partitions on manifolds with symmetries. Electronic Research Archive, 2021, 29(6): 4327-4338. doi: 10.3934/era.2021088 |
Let A(t) be a continuous path of self-adjoint Fredholm operators, we derive a decomposition formula of spectral flow if the path is invariant under a matrix-like cogredient. As applications, we give the generalized Bott-type iteration formula for linear Hamiltonian systems.
In this paper, we consider the decomposition of the spectral flow for a path of self-adjoint Fredholm operators. Let
Some basic properties of spectral flow such as homotopy invariance, path additivity, direct sum e.t. are well known, please refer to the Appendix. We give the proof for another basic property which is called cogredient invariance property of spectral flow. For convenience, we first introduce some notations. Let
Lemma 1.1. Let
M∗sAsMs∈C([a,b],FS(H1)), | (1.1) |
and we have
sf(As;s∈[a,b])=sf(M∗sAsMs;[a,b]). | (1.2) |
Remark 1.2. The cogredient invariance property is a nature property. In the case
Our main result is the decomposition formula based on the cogredient invariance property. Let
∑1≤i≤mHi=H1+⋯Hm |
which is the subspace spanned by
H=∑1≤i≤nker(g−λi)m. | (1.3) |
We denote by
Hλ:=ker(g−λ)m, |
and denote
Fλ={Hλ,ifλ∈U;Hλ+Hˉλ−1,ifλ∉U. | (1.4) |
Then we have
H=∑1≤i≤kFλi. |
Moreover, let
Theorem 1.3. Let
g∗Asg=As,fors∈[0,1], | (1.5) |
then we have
sf(As)=sf(As|Fλ1)+⋯+sf(As|Fλj)+12(dimker(A1|ˆF)−dimker(A0|ˆF)). | (1.6) |
In [18], by assuming
sf(As)=sf(As|ker(g−λ1)+⋯+sf(As|ker(g−λj)) | (1.7) |
under the condition
Asg=gAs. | (1.8) |
Obviously, we give a generalization of (1.7). In fact, a significant difference is that we do not assume
The second main result is a generalization for the Bott-type iteration formula which is a powerful tool to study the multiplicity and stability of periodic orbits in Hamiltonian systems. In 1956, Bott got his celebrated iteration formula for the Morse index of closed geodesics [5], and it was generalized by [3,10,9,11]. The precise iteration formula of the general Hamiltonian system was established by Long [22,23]. In fact, the iteration could be regarded as a unitary group action. Motivated by the symmetry orbits in
Based on Theorem 1.3, we prove the Bott-type iteration formula which cover all the previous cases and moreover give some new generalizations. Our generalized formula could be applied to the closed geodesics on Semi-Riemanian manifold and heteroclinic orbits with reversible symmetry.
Now we consider the linear Hamiltonian system
˙x(t)=JB(t)x(t),t∈I, | (1.9) |
where
(x(a),x(b))∈Λ∈Lag(4n). | (1.10) |
In the case
σ(JB(±∞))∩iR=∅. | (1.11) |
Let
A:=−Jddt:E⊂H→H. |
and
g(E)=E,g∗Ag=A,g∗Bg=B, | (1.12) |
we have
It is well known that spectral flow is equal to Maslov index, and this is also true for the unbounded domain, see [6,27,26,7,15] and reference therein. The Maslov index is associated integer to a pair of continuous path
This paper is organized as follows. We prove Lemma 1.1 in Section 2 and Theorem 1.3 in Section 3. In Section 4, we list
Let
Definition 2.1. Let
sf(A(t);[0,1])=∑a<t≤bdimker(A(t)). | (2.1) |
Let
Let
We assume that
(a) For any
(b) For any
Then we have the following lemma.
Lemma 2.2. Let
sf(A(t);t∈[0,1])=∑1≤k≤nsf(fk(A(t));t∈[0,1]). | (2.2) |
Proof. Since the spectral flow satisfies the path additivity property, we only need to prove (2.2) locally. Let
dim ker(A(t0)+δK)=0. |
It follows that
dim ker(A(t)+δK)=0,∀t∈[t0−δ1,t0+δ1]. |
It follows that
{sf(A(t)+δK,t∈[t0−δ1,t0+δ1])=0sf(hk(δ,t),t∈[t0−δ1,t0+δ1])=0. |
By homotopy invariance of spectral flow, we have
sf(A(t);t∈[t0−δ1,t0+δ1])=sf(A(t0−δ1+sK);s∈[0,δ])−sf(A(t0+δ1+sK);s∈[0,δ]) | (2.3) |
and
sf(hk(0,t);t∈[t0−δ1,t0+δ1])=sf(hk(s,t0−δ1);s∈[0,δ])−sf(hk(s,t0+δ1);s∈[0,δ]). | (2.4) |
Note that
sf(A(t0±δ1)+sK;s∈[0,δ])=∑0<s≤δdim ker(A(t0±δ1)+sK)=∑0<s≤δ∑1≤k≤ndim ker(hk(s,t0±δ1))=∑1≤k≤nsf(hk(s,t0±δ1);s∈[0,δ]). |
This completes the proof.
Please note that Lemma 2.2 can be considered as a generalization of direct sum property of spectral flow.
In the next, we will prove that the spectral flow is invariant under the cogredient. The next Lemma is contained in [13], but for reader's convenience, we give details here.
Lemma 2.3. Let
Proof. Since
Since
Recall that the gap topology can be induced by the gap distance
ˆδ(M,N)=max{δ(M,N),δ(N,M)}, | (2.5) |
where
δ{M,N}:={supu∈SMdist(u,N),ifM≠{0}0,ifM={0}. |
The gap distance has the following properties:
Lemma 2.4. Let
Proof. Without loss of generality, we assume that
‖x−Qy‖≤‖x−Py‖+‖Qy−Py‖≤‖P‖‖P−1x−y‖+‖Q−P‖‖y‖≤‖P‖dist(P−1x,N)+‖Q−P‖‖P−1‖≤‖P‖δ(M,N)+‖Q−P‖‖P−1‖. |
It follows that
Lemma 2.5. Suppose
Proof. We only need to show that
Gr(M∗sAsMs)={(M∗sAsx,M−1sx)|x∈Es}. |
Let
ˆδ(Gr(M∗s0As0Ms0,M∗sAsMs)=ˆδ(Qs0Gr(As0),QsGr(As))≤C1ˆδ(As0,As)+C2‖Qs−Qs0‖. |
By the continuity of
Now we give the proof of Lemma 1.1.
Proof of Lemma 1.1. Please note that (1.1) is from Lemma 2.5. We first prove the case
dim ker(M∗AM)=dim (M−1kerA)=dim kerA |
for each
sf(As;s∈[a,b])=sf(M∗AsM;s∈[a,b])forM∈L∗(H1,H2). |
Now we consider the two family
sf(M∗aAsMa)=sf(M∗sAsMs)−sf(M∗a+t(b−a)AbMa+t(b−a)). |
Note that
sf(M∗a+t(b−a)AbMa+t(b−a))=0. |
It follows that
sf(M∗aAsMa)=sf(M∗sAsMs). |
This completes the proof.
As an example, we consider the one parameter family of linear Hamiltonian systems
˙z(t)=JBs(t)z(t),(s,t)∈[0,1]×[0,T], | (2.6) |
where
(xs(0),xs(T))∈Λs∈Lag(4n), |
where we assume
Let
E(Λs)={x∈W1,2([0,T],C2n),(x(0),x(T))∈Λs}. |
We define
˙γs(t)=JBs(t)γs(t), | (2.7) |
then
γs(t)∈Sp(2n):={P∈L∗(R2n),P∗JP=J}, |
which implies
−sf(As−Bs)=μ(Λs,Gr(γs(T))). |
Let
(P∗s)−1(As−Bs)P−1s∈FS(H) |
with domain
PsEs={x∈W1,2([0,T],C2n),(x(0),x(T))∈ˆPs(T)Λs}, |
where
(P∗s)−1(−Jddt|E(Λs)−Bs)P−1s=As−ˆBs, |
where
sf(−Jddt|E(ˆPs(T)Λs)−ˆBs)=sf((P∗s)−1(As−Bs)P−1s)=sf(As−Bs). | (2.8) |
From (6.6), we can express the left of (2.8) as Maslov index. In fact, the fundamental solution is
sf(−Jddt|E(ˆPs(T)Λs)−ˆBs)=μ(ˆPs(T)Λs,ˆPs(T)Gr(γs(T))). |
Formula (2.8) implies that
μ(Λs,Gr(γs(T)))=μ(ˆPs(T)Λs,ˆPs(T)Gr(γs(T))), |
which is just the symplectic invariance property (6.4) of Maslov index.
In this section, we will prove the decomposition formula for spectral flow. Suppose
H=∑1≤i≤nHλi, | (3.1) |
where
x=p1(g)(g−λ1)mx+p2(g)(g−λ2)mx=0. |
Similarly we have
Lemma 3.1.
Proof. We only need to show that
n∑i=1ai(λ)Gi(λ)=1. |
It follows that
H=∑1≤i≤nGi(g)H. |
We also have
We have the following lemmas.
Lemma 3.2. Let
g∗Ag=A,gE=E, |
then
(Ax,y)=0,ifx∈Hλ∩E,y∈Hμ∩E. | (3.2) |
Proof. Let
(Ax,y)=(Agx,gy)=λˉμ(Ax,y) |
implies
(Ax,y)=(Agx,gy)=(A(g−λ)x,gy)+(Aλx,(g−μ)y)+λˉμ(Ax,y)=λˉμ(Ax,y). |
Since
Lemma 3.3. Under the condition of Lemma 3.2, we have
Proof. Note that
g∗Ag(kerA)=AkerA=0. |
It follows that
For
Fλ={Hλ,ifλ∈U;Hλ+Hˉλ−1,ifλ∉U, |
then we have
Let
((x1,x2,⋯,xk),(y1,y2,⋯,yk))=∑1≤i≤k(xi,yi), |
where
M:(x1,x2,⋯,xk)→∑1≤i≤kxi |
is a homeomorphism from
Please note that
ker(A|Fλi)=ker(AM)∩M−1(Fλi)=M−1(kerA∩Fλi). |
Proposition 3.4. Suppose
sf(As)=sf(As|Fλ1)+⋯+sf(As|Fλk). | (3.3) |
Proof. By Lemma 1.1, we have
sf(As)=sf(M∗AsM). |
Note that
sf(As)=sf(M∗AsM)=∑1≤i≤ksf(As|Fλi). |
This complete the proof.
Lemma 3.5. If
sf(As|Fλ)=12(dimker(A1|Fλ)−dimker(A0|Fλ)). | (3.4) |
Proof. Recall that
(QM∗AsMQ(x1+y1),(x2+y2))=−(M∗AsM(x1+y1),(x2+y2)). |
It follows that
2sf(As|Fλ)=sf(As|Fλ)+sf(QAs|FλQ)=sf(As)+sf(−As)=dimker(A1|Fλ)−dimker(A0|Fλ). |
The lemma then follows.
Proof of Theorem 1.3. By Proposition 3.4 and Lemma 3.5, we only need to show that
12(dimker(A1|ˆF)−dimker(A0|ˆF))=∑λ∉U12(dimker(A1|Fλ)−dimker(A0|Fλ)). |
In fact
kerA1∩ˆF=∑λ∉Uker(A1)∩Fλ. |
It follows that
Corollary 3.6. Under the condition of Theorem 1.3, if
sf(As)=12(dimker(A1)−dimker(A0)). | (3.5) |
If the path is closed, then
sf(As)=0. |
Remark 3.7. In the case
I(A,A−B)=−sf(A−sB;s∈[0,1]). | (3.6) |
Especially, when
In this section, we will give the applications for Hamiltonian systems. We list
For
˙z(t)=JB(t),(z(0),z(T))∈Λ, | (4.1) |
where
EΛ={x∈W1,2([0,T],C2n),(x(0),x(T))∈Λ}, |
then
g∗Ag=A,g∗Bg=B,gEΛ=EΛ. | (4.2) |
In order to make
((gx)(0),(gx)(T))∈Λif(x(0),x(T))∈Λ. |
Hence we have
g∗(A−sB)g=A−sB,s∈R. |
and get the decomposition formula (1.7).
It is well known that for
Case 1. For
(gx)(t)=Px(t), | (4.3) |
then it is obvious that
σ(g)=σ(P). |
Let
Case 2. For
z(0)=Sz(T), | (4.4) |
and moreover we assume
S∗B(0)S=B(T). | (4.5) |
We assume (4.1) with
(gx)(t)={Px(t+Tk),t∈[0,k−1kT];S−1Px(t+Tk−T),t∈[k−1nT,T]. | (4.6) |
Easy computation show that
Lemma 4.1. The adjoint operator
(g∗x)(t)={P∗(S∗)−1x(t+T−Tk),t∈[0,Tk];P∗x(t−Tk),t∈[Tk,T]. | (4.7) |
Proof. Let
∫k−1kT0(Px(t+T/k),y(t))dt=∫TT/k(x(t),P∗y(t−T/k))dt, |
and
∫Tk−1kT(S−1Px(t+T/k−T),y(t))dt=∫T/k0(x(t),P∗(S∗)−1y(t+T−T/k)). |
Then we have checked
B(t)={P∗(S∗)−1B(t+T−Tk)S−1P,t∈[0,Tn];P∗B(t−Tk)P,t∈[Tn,T]. | (4.8) |
Please note that (4.8) implies (4.5), and (4.2) is satisfied. Since
(gkx)(t)=S−1Pkx(t), |
which is a multiplicity operator on
σ(gk)=σ(S−1Pk). |
To simplify the notation, for
Ω1k={z∈C,zk∈Ω}. |
By this notation, we have
Case 3. We consider the generalized reversible symmetry. We call a matrix
M∗JM=−J. | (4.9) |
We denote by
M1M2∈Sp(2n),M1M3∈Spa(2n). |
We list some basic property of
Lemma 4.2. If
Proof. Note that
dim ker(M−ˉλ)=dim ker(M∗−λ)=ker(M−1+λ)=dim ker(M+λ−1). |
And we also have
¯det(M−ˉλ)=det(−M−1−λ)=det(−M−1λ)det(M+λ−1). |
It follows that
Similar with the symplectic matrix, we have the following results.
Lemma 4.3. Let
Proof. Let
(Jx,y)=−(JMx,My)=−(J(M−λ)x,My)−(Jλx,(M−μ)y)−λˉμ(Jx,y)=−λˉμ(Jx,y). |
Since
We assume (4.1) admits a
(gx)(t)=Nx(T−t). | (4.10) |
We assume
(Nx(T),Nx(0))∈Λ,if(x(0),x(T))∈Λ, | (4.11) |
then
N∗B(T−t)N=B(t), | (4.12) |
then (4.2) is satisfied.
Please note that for the
x(0)∈V0,x(T)∈V1, |
where
NV0=V1,NV1=V0, |
then
Obviously, we have
(g2x)(t)=N2x(t), |
hence
σ(g)=(σ(N2))12. |
For
From theorem 1.3, we get the decomposition of spectral flow. Since on the finite interval
I(A,A−B)=m∑i=1I(A|Fλi,A|Fλi−B|Fλi)+12(dimker((A−B)|ˆF)−dimker(A|ˆF)). | (4.13) |
All the above discussions can be applied to Sturm-Liouville systems, so we don't give the details in all cases, instead we only consider the following two cases which have clear background.
Case 4. We consider the one parameter family Sturm-Liouville system
−(Gs(t)˙x)′+Rs(t)x(t)=0,x(0)=Sx(T),˙x(0)=S˙x(T),s∈[0,1] | (4.14) |
where
Gs(t)={P∗(S∗)−1Gs(t+T−Tn)S−1P,t∈[0,Tn];P∗Gs(t−Tn)P,t∈[Tn,T]. | (4.15) |
Rs(t)={P∗(S∗)−1Rs(t+T−Tn)S−1P,t∈[0,Tn];P∗Rs(t−Tn)P,t∈[Tn,T]. | (4.16) |
Then
g∗(−(Gs(t)ddt)′+Rs)g=−(Gs(t)ddt)′+Rs, |
and we could give the decomposition of spectral flow from Theorem 1.3.
This case includes the Bott-type formula of Semi-Riemann manifold [17]. Let
g(ei,ej)={0,i≠j;1,1≤i=j≤n−ν;−1,n−ν≤i=j≤n |
and
(e1(0),⋯,en(0))=(e1(T),⋯,en(T))P, |
then
Writing the
−G¨u+R(t)u(t)=0,t∈[0,T], | (4.17) |
where
u(0)=Pu(T). |
For
E2ω,T:={u∈W2,2([0,T],Cn)|u(0)=ωPu(T),˙u(0)=ωP˙u(T)}, |
then
Aωs,T=−Gd2dt2+R(t)+sG |
are self-adjoint Fredholm operators on
It has proved in [17] that there exist
iωspec(c):=sf(Aωs,T;s∈[0,+∞)). |
Let
iωspec(c(m)):=sf(Aωs,mT;s∈[0,+∞)). |
Let
σ(g)={ω}1m. |
Let
Hωj=ker(g−ωj)={u(t)=ωjPu(t+T)}. |
We have
sf(Aωs,mT;s∈[0,+∞))=∑ωmj=ωsf(Aωjs,T;s∈[0,+∞)). |
Hence we get the Bott-type iteration formula [17]
iωspec(c(m))=∑ωmj=ωiωjspec(c). | (4.18) |
Obviously, we can consider the case of reversible symmetry, since it is similar, we omit the detail.
Case 5. Now we consider the case of heteroclinic orbits, for the one parameter family linear Hamiltonian system
˙x=JBs(t)x(t),t∈R,s∈[0,1]. | (4.19) |
Let
σ(JBs(±∞))∩iR=∅,s∈[0,1]. |
Let
We assume
N∗Bs(−t)N=Bs(t), |
then
g∗(A−Bs)g=A−Bs,s∈[0,1]. |
Obviously, we have
(g2x)(t)=N2x(t), |
hence
σ(g)=(σ(N2))12. |
For
In the case
H±=kerg∓I={x∈H,Nx(−t)=±x(t)}, | (4.20) |
we have
sf(A−Bs)=sf(A|H+−Bs|H+)+sf(A|H−−Bs|H−). | (4.21) |
Now we consider the case of homoclinic orbits. For the linear Hamiltonian system
˙x(t)=JB(t)x(t),t∈R, | (4.22) |
assume
I(A−B∗,A−B)=−sf(A−B∗+s(B−B∗)). |
In the case that (4.22) is a linear system of homoclinic orbits
i(z)=I(A−B∗,A−B). |
Assume
I(A−B∗,A−B)=I(A|H+−B∗|H+,A|H+−B|H+)+I(A|H−−B∗|H−,A|H−−B|H−). | (4.23) |
Case 6. We consider the one parameter Sturm-Liouville system on
−(G(t)˙x)′+R(t)x(t)=0,t∈R | (4.24) |
where
δ1<R(t)<δ2,fort≥|T|. | (4.25) |
The Morse index of
m−(A)=sf(A+sG(0);s∈[0,+∞)). |
We assume there exist
N∗R(−t)N=R(t),N∗G(−t)N=G(t). |
Let
gx(t)=Nx(−t), |
then
m−(A)=j∑i=1m−(A|Fλi)−dimkerA|ˆF. |
In the case
m−(A)=m−(A|H+)+m−(A|H−). | (4.26) |
In this section, we will give some Bott-type iteration formulas of Maslov-index. In what follows
gm=ωI |
for some
In cases 2, 3, 5,
(Tf)(x)=f(x),x∈ˆI. | (5.1) |
Lemma 5.1. Suppose for
sf(As|Hi;s∈[a,b])=sf(ˆAis;s∈[a,b]). | (5.2) |
Proof. Note that
Now we consider Case 2. We assume
Hi=ker(g−ωi)={x∈H,ωix(t)=Ps(t+Tm)}. |
We choose
TEi={x∈W1,2([0,Tm],C2n),ωix(0)=Ps(Tm)}. |
From Corollary 6.2, we have
−sf(As)=μ(Gr(ωS),Gr(γ(t));t∈[0,T]),−sf(ˆA(i)s)=μ(Gr(ωi),Gr(γ(t));t∈[0,T/m]). |
Then we have
μ(Gr(S−1),Gr(γ(t));t∈[0,T])=m∑i=1μ(Gr(ωiP−1),Gr(γ(t));t∈[0,T/m]). | (5.3) |
Remark 5.2. In the case
For Case 3. We assume
H±=kerg∓I={x∈H,Nx(T−t)=±x(t)}, |
then
For the
TE±={x∈W1,2([0,T/2],C2n),x(0)∈V±(SN),x(T/2)∈V±(N)}. |
We have
μ(Gr(S),Gr(γ(t));t∈[0,T])=μ(V+(N),γ(t)V+(SN);t∈[0,T/2])+μ(V−(N),γ(t)V−(SN);t∈[0,T/2]). |
Similarly, if the boundary condition is given by
μ(V1,γ(t)V0;t∈[0,T])=μ(V+(N),γ(t)V0;t∈[0,T/2])+μ(V−(N),γ(t)V0;t∈[0,T/2]). | (5.4) |
Remark 5.3. To our knowledge, in the case
Now we consider Case 5. For
˙γλ(τ,t)=JBλ(t)γλ(τ,t),γλ(τ,τ)=I2n. | (5.5) |
Let
Vsλ(τ):={v∈R2n|limτ→∞γλ(τ,t)v=0},Vuλ(τ):={v∈R2n|limτ→−∞γλ(τ,t)v=0}, |
be the stable and unstable paths, then
Recall that in this case,
Aλ:=−Jddt−Bs(t):E=W1,2(R,R2n)⊂H→H. |
Let
TE±={x∈W1,2(R−,R2n),x(0)∈V±(N)}. |
Let
From Prop 3.7 of [15], we have
−sf(Aλ;λ∈[0,1])=μ(Vsλ(0),Vuλ(0);λ∈[0,1]). |
Similarly
−sf(A±λ;λ∈[0,1])=μ(V±(N),Vuλ(0);λ∈[0,1]). |
Then we have
μ(Vsλ(0),Vuλ(0);λ∈[0,1])=μ(V+(N),Vuλ(0);λ∈[0,1])+μ(V−(N),Vuλ(0);λ∈[0,1]). | (5.6) |
In the case of homoclinic orbits, let
−sf(Aλ;λ∈[0,1])=μ(Vs(+∞),Vu(t);t∈R)=−μ(Vs(t),Vu(−t);t∈R+). | (5.7) |
We have
−sf(A±λ;λ∈[0,1])=μ(V±(N),Vu(t);t∈R−)=−μ(V±(N),Vu(−t);t∈R+). | (5.8) |
Compare (5.7) and (5.8), we have
μ(Vs(+∞),Vu(t);t∈R)=μ(V+(N),Vu(t);t∈R−)+μ(V−(N),Vu(t);t∈R−), | (5.9) |
or equivalently
μ(Vs(t),Vu(−t);t∈R+)=μ(V+(N),Vu(−t);t∈R+)+μ(V−(N),Vu(−t);t∈R+). | (5.10) |
Obviously, we can use
−sf(A±λ;λ∈[0,1])=μ(Vsλ(0),V±(N);λ∈[0,1]). | (5.11) |
μ(Vsλ(0),Vuλ(0);λ∈[0,1])=μ(Vsλ(0),V+(N);λ∈[0,1])+μ(Vsλ(0),V−(N);λ∈[0,1]). | (5.12) |
In the case of homoclinic orbits,
μ(Vs(t),Vu(−∞);t∈R)=μ(Vs(t),Vu(−t);t∈R+)=μ(Vs(t),V+(N);t∈R+)+μ(Vs(t),V−(N);t∈R+). |
In study the stability problem of homographic solution in planar
Spectral flow was introduced by Atiyah, Patodi and Singer in their study of index theory on manifold with boundary [2]. Let
(Stratum homotopy relative to the ends) If
sf(As,0;s∈[a,b])=sf(As,1;s∈[a,b]). |
(Path additivity) If
sf(A1t∗A2t;t∈[a,b])=sf(A1t;t∈[a,b])+sf(A2t;t∈[a,b]) |
where
(Direct sum) If
sf(A1t⊕A2t;t∈[a,b])=sf(A1t;t∈[a,b])+sf(A2t;t∈[a,b]). |
(Nullity) If
(Reversal) Denote the same path travelled in the reverse direction in
sf(At;t∈[a,b])=−sf(ˆAt;t∈[−b,−a]). |
The spectral flow is related to Maslov index in Hamiltonian systems. We now briefly reviewing the Maslov index theory [1,6,25]. Let
(Reparametrization invariance) Let
μ(L1(t),L2(t))=μ(L1(ϕ(τ)),L2(ϕ(τ))). | (6.1) |
(Homotopy invariant with end points) For two continuous family of Lagrangian path
μ(L1(0,t),L2(0,t))=μ(L1(1,t),L2(1,t)). | (6.2) |
(Path additivity) If
μ(L1(t),L2(t))=μ(L1(t),L2(t)|[a,c])+μ(L1(t),L2(t)|[c,b]). | (6.3) |
(Symplectic invariance) Let
μ(L1(t),L2(t))=μ(γ(t)L1(t),γ(t)L2(t)). | (6.4) |
(Symplectic additivity) Let
L1,L2∈C([a,b],Lag(W1))andˆL1,ˆL2∈C([a,b],Lag(W2)), |
then
μ(L1(t)⊕ˆL1(t),L2(t)⊕ˆL2(t))=μ(L1(t),L2(t))+μ(ˆL1(t),ˆL2(t)). | (6.5) |
The next Theorem give the relation of spectral flow and Maslov index.
Theorem 6.1.
−sf(As−Bs)=μ(Λs,Gr(γs(T))) | (6.6) |
The above Theorem is well known [15], we like to give a direct proof here.
Proof. We use the idea of Lemma 2.2. We only need to prove the theorem locally. Let
ker(As−Bs+I)=0,∀s∈[0,1]. |
Let
˙z(t)=J(Bs(t)−rI)z(t),(s,t)∈[0,1]×[0,T],r∈[0,1]. | (6.7) |
Recall that
μ(Λs,Gr(γs,1(T)))=0. |
Then use the homotopy invariant property of spectral flow and Maslov index, we have
{sf(As−Bs,s∈[0,1])=sf(A0−B0+rI)−sf(A1−B1+rI)μ(Λs,Gr(γs(T)))=μ(Λ0,Gr(γ0,r(T)))−μ(Λ1,Gr(γ0,r(T))). | (6.8) |
Note that
sf(A0−B0+rI)=∑0<r≤1dim ker(A0−B0+rI). |
Let
Note that
∂∂t(γs(t)−1∂∂sγs(t))=−γs(t)−1∂∂tγs(t)γs(t)−1∂∂sγs(t)+γs(t)−1∂∂s(∂∂tγs(t))=−γs(t)−1∂∂tγs(t)γs(t)−1∂∂sγs(t)+γs(t)−1∂∂s(JBs)γs(t)+γs(t)−1JBs∂∂sγs(t)=γs(t)−1∂∂s(JBs)γs(t). |
Then we have
−Jγ0,r(t)−1∂∂rγ0,r(t)=−TI. |
Thus we have
μ(Λ0,Gr(γ0,r(T)))=−∑0<r≤1dim (Λ0∩Gr(γ0,r(T)))=−∑0<r≤1dim (ker(A0−B0+rI))=−sf(A0−B0+rI). |
Similarly
μ(Λ1,Gr(γ1,r(T)))=−sf(A1−B1+rI). |
Then by (6.8), we get (6.6).
From the homotopy invariance of Maslov index, we have
Corollary 6.2.
−sf(A−sB)=μ(Λ,Gr(γ(t)),t∈[0,T]). | (6.9) |
We would like to thank Professor Alessandro Portaluri some helpful discussion with us for the spectral flow.
[1] |
V. I. Arnol'd, On a characteristic class entering into conditions of quantization, Funkcional. Anal. i Priložen, 1 (1967), 1–14. doi: 10.1007/BF01075861
![]() |
[2] |
Spectral asymmetry and Riemannian geometry. Ⅲ. Math. Proc. Cambridge Philos. Soc. (1976) 79: 71-99. ![]() |
[3] |
W. Ballmann, G. Thorbergsson and W. Ziller, Closed geodesics on positively curved manifolds, Ann. of Math. (2), 116 (1982), 213-247. doi: 10.2307/2007062
![]() |
[4] |
Unbounded Fredholm operators and spectral flow. Canad. J. Math. (2005) 57: 225-250. ![]() |
[5] |
On the iteration of closed geodesics and the Sturm intersection theory. Comm. Pure Appl. Math. (1956) 9: 171-206. ![]() |
[6] |
On the Maslov index. Comm. Pure Appl. Math. (1994) 47: 121-186. ![]() |
[7] |
Maslov index for homoclinic orbits of Hamiltonian systems. Ann. Inst. H. Poincaré Anal. Non Linéaire (2007) 24: 589-603. ![]() |
[8] |
A. Chenciner and R. Montgomery, A remarkable periodic solution of the three-body problem in the case of equal masses. Ann. of Math. (2), 152 (2000), 881-901. doi: 10.2307/2661357
![]() |
[9] |
Morse-type index theory for flows and periodic solutions for Hamiltonian equations. Comm. Pure Appl. Math. (1984) 37: 207-253. ![]() |
[10] |
The behavior of the index of a periodic linear Hamiltonian system under iteration. Advances in Math. (1977) 23: 1-21. ![]() |
[11] |
I. Ekeland, Convexity Methods in Hamiltonian Mechanics, Results in Mathematics and Related Areas, 19, Springer-Verlag, Berlin, 1990. doi: 10.1007/978-3-642-74331-3
![]() |
[12] |
On the existence of collisionless equivariant minimizers for the classical n-body problem. Invent. Math. (2004) 155: 305-362. ![]() |
[13] | P. Fitzpatrick, J. Pejsachowicz and C. Stuart, Spectral flow for paths of unbounded operators and bifurcation of critical points, work in progress, (2006). |
[14] |
Collision index and stability of elliptic relative equilibria in planar n-body problem. Comm. Math. Phys. (2016) 348: 803-845. ![]() |
[15] |
X. Hu and A. Portaluri, Index theory for heteroclinic orbits of Hamiltonian systems, Calc. Var. Partial Differential Equations, 56 (2017), 24pp. doi: 10.1007/s00526-017-1259-9
![]() |
[16] |
X. Hu, A. Portaluri and R. Yang, A dihedral Bott-type iteration formula and stability of symmetric periodic orbits, Calc. Var. Partial Differential Equations, 59 (2020). doi: 10.1007/s00526-020-1709-7
![]() |
[17] |
Instability of semi-Riemannian closed geodesics. Nonlinearity (2019) 32: 4281-4316. ![]() |
[18] |
Index and stability of symmetric periodic orbits in Hamiltonian systems with application to figure-eight orbit. Comm. Math. Phys. (2009) 290: 737-777. ![]() |
[19] |
Maslov (P, ω)-index theory for symplectic paths. Adv. Nonlinear Stud. (2015) 15: 963-990. ![]() |
[20] |
Iteration theory of L-index and multiplicity of brake orbits. J. Differential Equations (2014) 257: 1194-1245. ![]() |
[21] |
Seifert conjecture in the even convex case. Comm. Pure Appl. Math. (2014) 67: 1563-1604. ![]() |
[22] |
Bott formula of the Maslov-type index theory. Pacific J. Math. (1999) 187: 113-149. ![]() |
[23] |
Y. Long, Index Theory for Symplectic Paths with Applications, Progress in Mathematics, 207, Birkhäuser Verlag, Basel, 2002. doi: 10.1007/978-3-0348-8175-3
![]() |
[24] |
Multiple brake orbits in bounded convex symmetric domains. Adv. Math. (2006) 203: 568-635. ![]() |
[25] |
The Maslov index for paths. Topology (1993) 32: 827-844. ![]() |
[26] |
The spectral flow and the Maslov index. Bull. London Math. Soc. (1995) 27: 1-33. ![]() |
[27] |
Maslov-type index theory for symplectic paths and spectral flow. (Ⅰ). Chinese Ann. Math. Ser. B (1999) 20: 413-424. ![]() |
1. | Duanzhi Zhang, Zhihao Zhao, Maslov-type (L,P)-index and subharmonic P-symmetric brake solutions for Hamiltonian systems, 2025, 415, 00220396, 383, 10.1016/j.jde.2024.09.037 |