Skip to main content
Log in

Multiplicity and ellipticity of closed characteristics on compact star-shaped hypersurfaces in \(\mathbf{R}^{2n}\)

  • Published:
Calculus of Variations and Partial Differential Equations Aims and scope Submit manuscript

Abstract

In this paper, we firstly generalize some theories developed by Ekeland and Hofer (Commun Math Phys 113:419–467, 1987) for closed characteristics on compact convex hypersurfaces in \(\mathbf{R}^{2n}\) to star-shaped hypersurfaces. As applications we use Ekeland–Hofer theory and index iteration theory to prove that if a compact star-shaped hypersuface in \(\mathbf{R}^4\) satisfying some suitable pinching condition carries exactly two geometrically distinct closed characteristics, then both of them must be elliptic. We also conclude that the theory developed by Long and Zhu (Ann Math 155:317–368, 2002) still holds for dynamically convex star-shaped hypersurfaces, and combining it with the results in Wang et al. (Duke Math J 139:411–462, 2007), Liu et al. (J Funct Anal 266:5598–5638, 2014) and Wang (Adv Math 297:93–148, 2016), we obtain that there exist at least n closed characteristics on every dynamically convex star-shaped hypersurface in \(\mathbf{R}^{2n}\) for \(n=3, 4\).

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Similar content being viewed by others

References

  1. Abreu, M., Macarini, L.: Multiplicity of periodic orbits for dynamically convex contact forms. J. Fixed Point Theory Appl. 19(1), 175–204 (2017)

    Article  MathSciNet  MATH  Google Scholar 

  2. Berestycki, H., Lasry, J.M., Mancini, G., Ruf, B.: Existence of multiple periodic orbits on starshaped Hamiltonian systems. Commun. Pure Appl. Math. 38, 253–289 (1985)

    Article  MATH  Google Scholar 

  3. Chang, K.C.: Infinite Dimensional Morse Theory and Multiple Solution Problems. Birkhäuser, Boston (1993)

    Book  MATH  Google Scholar 

  4. Cristofaro-Gardiner, D., Hutchings, M.: From one Reeb orbit to two. J. Differ. Geom. 102, 25–36 (2016)

    Article  MathSciNet  MATH  Google Scholar 

  5. Duan, H., Liu, H., Long, Y., Wang, W.: Non-hyperbolic closed characteristics on non-degenerate star-shaped hypersurfaces in \({{\bf R}}^{2n}\). Acta Math. Sin. Engl. Ser. (to appear). arXiv:1510.08648

  6. Ekeland, I.: Convexity Methods in Hamiltonian Mechanics. Springer, Berlin (1990)

    Book  MATH  Google Scholar 

  7. Ekeland, I., Hofer, H.: Convex Hamiltonian energy surfaces and their closed trajectories. Commun. Math. Phys. 113, 419–467 (1987)

    Article  MATH  Google Scholar 

  8. Ekeland, I., Lassoued, L.: Multiplicité des trajectoires fermées d’un systéme hamiltonien sur une hypersurface d’energie convexe. Ann. IHP Anal. Non Linéaire 4, 1–29 (1987)

    Article  Google Scholar 

  9. Fadell, E., Rabinowitz, P.: Generalized cohomological index theories for Lie group actions with an application to bifurcation equations for Hamiltonian systems. Invent. Math. 45, 139–174 (1978)

    Article  MathSciNet  MATH  Google Scholar 

  10. Ginzburg, V., Gören, Y.: Iterated index and the mean Euler characteristic. J. Topol. Anal. 7, 453–481 (2015)

    Article  MathSciNet  MATH  Google Scholar 

  11. Ginzburg, V., Gürel, B.: Lusternik–Schnirelmann theory and closed Reeb orbits. arXiv:1601.03092

  12. Ginzburg, V.L., Hein, D., Hryniewicz, U.L., Macarini, L.: Closed Reeb orbits on the sphere and symplectically degenerate maxima. Acta Math. Vietnam. 38, 55–78 (2013)

    Article  MathSciNet  MATH  Google Scholar 

  13. Girardi, M.: Multiple orbits for Hamiltonian systems on starshaped ernergy surfaces with symmetry. Ann. IHP Anal. Non Linéaire 1, 285–294 (1984)

    Article  MATH  Google Scholar 

  14. Gutt, J., Kang, J.: On the minimal number of periodic orbits on some hypersurfaces in \({{\bf R}}^{2n}\). Ann. Inst. Fourier (Grenoble) 66(6), 2485–2505 (2016)

    Article  MathSciNet  MATH  Google Scholar 

  15. Hu, X., Long, Y.: Closed characteristics on non-degenerate star-shaped hypersurfaces in \({{\bf R}}^{2n}\). Sci. China Ser. A 45, 1038–1052 (2002)

    Article  MathSciNet  MATH  Google Scholar 

  16. Hu, X., Ou, Y.: Stability of closed characteristics on compact convex hypersurfaces in \({{\bf R}}^{2n}\). J. Fixed Point Theory Appl. 19(1), 585–600 (2017)

    Article  MathSciNet  MATH  Google Scholar 

  17. Hofer, H., Wysocki, K., Zehnder, E.: The dynamics on three-dimensional strictly convex energy surfaces. Ann. Math. 148, 197–289 (1998)

    Article  MathSciNet  MATH  Google Scholar 

  18. Hofer, H., Wysocki, K., Zehnder, E.: Finite energy foliations of tight three-spheres and Hamiltonian dynamics. Ann. Math. 157, 125–255 (2003)

    Article  MathSciNet  MATH  Google Scholar 

  19. Liu, C., Long, Y.: Hyperbolic characteristics on star-shaped hypersurfaces. Ann. IHP Anal. Non Linéaire 16, 725–746 (1999)

    Article  MathSciNet  MATH  Google Scholar 

  20. Liu, H., Long, Y.: The existence of two closed characteristics on every compact star-shaped hypersurface in \({\bf R}^4\). Acta Math. Sin. Engl. Ser. 32, 40–53 (2016)

    Article  MathSciNet  MATH  Google Scholar 

  21. Liu, H., Long, Y.: Resonance identities and stability of symmetric closed characteristics on symmetric compact star-shaped hypersurfaces. Calc. Var. PDEs 54, 3753–3787 (2015)

    Article  MathSciNet  MATH  Google Scholar 

  22. Liu, H., Long, Y., Wang, W.: Resonance identities for closed charactersitics on compact star-shaped hypersurfaces in \({{\bf R}}^{2n}\). J. Funct. Anal. 266, 5598–5638 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  23. Long, Y.: Precise iteration formulae of the Maslov-type index theory and ellipticity of closed characteristics. Adv. Math. 154, 76–131 (2000)

    Article  MathSciNet  MATH  Google Scholar 

  24. Long, Y.: Index Theory for Symplectic Paths with Applications. Progress in Mathematics, vol. 207. Birkhäuser, Basel (2002)

    Book  Google Scholar 

  25. Long, Y., Zhu, C.: Closed characteristics on compact convex hypersurfaces in \({{\bf R}}^{2n}\). Ann. Math. 155, 317–368 (2002)

    Article  MathSciNet  MATH  Google Scholar 

  26. Rabinowitz, P.: Periodic solutions of Hamiltonian systems. Commun. Pure. Appl. Math. 31, 157–184 (1978)

    Article  MathSciNet  Google Scholar 

  27. Szulkin, A.: Morse theory and existence of periodic solutions of convex Hamiltonian systems. Bull. Soc. Math. Fr. 116, 171–197 (1988)

    Article  MathSciNet  MATH  Google Scholar 

  28. Viterbo, C.: Une théorie de Morse pour les systèmes hamiltoniens étoilés. Comptes Rendus Acad. Sci. Paris Ser. I Math. 301, 487–489 (1985)

    MathSciNet  MATH  Google Scholar 

  29. Viterbo, C.: Equivariant Morse theory for starshaped Hamiltonian systems. Trans. Am. Math. Soc. 311, 621–655 (1989)

    Article  MathSciNet  MATH  Google Scholar 

  30. Wang, W.: Stability of closed characteristics on compact convex hypersurfaces in \({{\bf R}}^6\). J. Eur. Math. Soc. 11, 575–596 (2009)

    Article  MathSciNet  MATH  Google Scholar 

  31. Wang, W.: Existence of closed characteristics on compact convex hypersurfaces in \({{\bf R}}^{2n}\). Calc. Var. PDEs 55, 1–25 (2016)

    Article  MathSciNet  Google Scholar 

  32. Wang, W.: Closed characteristics on compact convex hypersurfaces in \({{\bf R}}^{8}\). Adv. Math. 297, 93–148 (2016)

    Article  MathSciNet  MATH  Google Scholar 

  33. Weinstein, A.: Periodic orbits for convex Hamiltonian systems. Ann. Math. 108, 507–518 (1978)

    Article  MathSciNet  MATH  Google Scholar 

  34. Wang, W., Hu, X., Long, Y.: Resonance identity, stability and multiplicity of closed characteristics on compact convex hypersurfaces. Duke Math. J. 139, 411–462 (2007)

    Article  MathSciNet  MATH  Google Scholar 

Download references

Acknowledgements

We would like to sincerely thank the referee for her/his careful reading of the manuscript, raising several delicate points which we had overlooked, valuable suggestion and comments to improve the exposition.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Hui Liu.

Additional information

Communicated by P. Rabinowitz.

Huagui Duan: Partially supported by NSFC (Nos. 11671215, 11471169, 11131004), LPMC of MOE of China and Nankai University. Hui Liu: Partially supported by NSFC (Nos. 11401555, 11371339), Anhui Provincial Natural Science Foundation (No. 1608085QA01).

Appendix

Appendix

In the section, we briefly review the equivariant Morse theory and the resonance identities for closed characteristics on compact star-shaped hypersurfaces in \(\mathbf{R}^{2n}\) developed in [22]. Now we fix a \({\Sigma }\in \mathcal{H}_{st}(2n)\) and assume the following condition:

(F) There exist only finitely many geometrically distinct prime closed characteristics \(\quad \{(\tau _j, y_j)\}_{1\le j\le k}\) on \(\Sigma \).

Let \(\hat{\sigma }=\inf _{1\le j\le k}{\sigma _j}\) and T be a fixed positive constant. Then by Section 2 of [22], for any \(a>\frac{\hat{\sigma }}{T}\), we can construct a function \(\varphi _a\in C^{\infty }(\mathbf{R}, \mathbf{R}^+)\) which has 0 as its unique critical point in \([0, +\infty )\). Moreover, \(\frac{\varphi ^{\prime }(t)}{t}\) is strictly decreasing for \(t>0\) together with \(\varphi (0)=0=\varphi ^{\prime }(0)\) and \(\varphi ^{\prime \prime }(0)=1=\lim _{t\rightarrow 0^+}\frac{\varphi ^{\prime }(t)}{t}\). More precisely, we define \(\varphi _a\) and the Hamiltonian function \(\widetilde{H}_a(x)=a{\varphi }_a(j(x))\) via Lemma 2.2 and Lemma 2.4 in [22]. The precise dependence of \(\varphi _a\) on a is explained in Remark 2.3 of [22].

For technical reasons we want to further modify the Hamiltonian, we define the new Hamiltonian function \(H_a\) via Proposition 2.5 of [22] and consider the fixed period problem

$$\begin{aligned} \left\{ \begin{array}{lll} \dot{x}(t) &{}=&{} JH_a^\prime (x(t)), \\ x(0) &{}=&{} x(T). \\ \end{array}\right. \end{aligned}$$
(5.1)

Then \(H_a\in C^{3}(\mathbf{R}^{2n} {\setminus }\{0\},\mathbf{R})\cap C^{1}(\mathbf{R}^{2n},\mathbf{R})\). Solutions of (5.1) are \(x\equiv 0\) and \(x=\rho z(\sigma t/T)\) with \(\frac{{\varphi }_a^\prime (\rho )}{\rho }=\frac{\sigma }{aT}\), where \((\sigma , z)\) is a solution of (1.1). In particular, non-zero solutions of (5.1) are in one to one correspondence with solutions of (1.1) with period \(\sigma <aT\).

For any \(a>\frac{\hat{\sigma }}{T}\), we can choose some large constant \(K=K(a)\) such that

$$\begin{aligned} H_{a,K}(x) = H_a(x)+\frac{1}{2}K|x|^2 \end{aligned}$$
(5.2)

is a strictly convex function, that is,

$$\begin{aligned} (\nabla H_{a, K}(x)-\nabla H_{a, K}(y), x-y) \ge \frac{{\epsilon }}{2}|x-y|^2, \end{aligned}$$
(5.3)

for all \(x, y\in \mathbf{R}^{2n}\), and some positive \({\epsilon }\). Let \(H_{a,K}^*\) be the Fenchel dual of \(H_{a,K}\) defined by

$$\begin{aligned} H_{a,K}^*(y) = \sup \left\{ x\cdot y-H_{a,K}(x)\;|\; x\in \mathbf{R}^{2n}\right\} . \end{aligned}$$

The dual action functional on \(X=W^{1, 2}(\mathbf{R}/{T \mathbf{Z}}, \mathbf{R}^{2n})\) is defined by

$$\begin{aligned} F_{a,K}(x) = \int _0^T{\left[ \frac{1}{2}(J\dot{x}-K x,x)+H_{a,K}^*(-J\dot{x}+K x)\right] dt}. \end{aligned}$$
(5.4)

Then \(F_{a,K}\in C^{1,1}(X, \mathbf{R})\) and for \(KT\not \in 2\pi \mathbf{Z}\), \(F_{a,K}\) satisfies the Palais–Smale condition and x is a critical point of \(F_{a, K}\) if and only if it is a solution of (5.1). Moreover, \(F_{a, K}(x_a)<0\) and it is independent of K for every critical point \(x_a\ne 0\) of \(F_{a, K}\).

When \(KT\notin 2\pi \mathbf{Z}\), the map \(x\mapsto -J\dot{x}+Kx\) is a Hilbert space isomorphism between \(X=W^{1, 2}(\mathbf{R}/{T \mathbf{Z}}; \mathbf{R}^{2n})\) and \(E=L^{2}(\mathbf{R}/(T \mathbf{Z}),\mathbf{R}^{2n})\). We denote its inverse by \(M_K\) and the functional

$$\begin{aligned} \Psi _{a,K}(u)=\int _0^T{\left[ -\frac{1}{2}(M_{K}u, u)+H_{a,K}^*(u)\right] dt}, \quad \forall \,u\in E. \end{aligned}$$
(5.5)

Then \(x\in X\) is a critical point of \(F_{a,K}\) if and only if \(u=-J\dot{x}+Kx\) is a critical point of \(\Psi _{a, K}\).

Suppose u is a nonzero critical point of \(\Psi _{a, K}\). Then the formal Hessian of \(\Psi _{a, K}\) at u is defined by

$$\begin{aligned} Q_{a,K}(v)=\int _0^T\left( -M_K v\cdot v+H_{a,K}^{*\prime \prime }(u)v\cdot v\right) dt, \end{aligned}$$
(5.6)

which defines an orthogonal splitting \(E=E_-\oplus E_0\oplus E_+\) of E into negative, zero and positive subspaces. The index and nullity of u are defined by \(i_K(u)=\dim E_-\) and \(\nu _K(u)=\dim E_0\) respectively. Similarly, we define the index and nullity of \(x=M_Ku\) for \(F_{a, K}\), we denote them by \(i_K(x)\) and \(\nu _K(x)\). Then we have

$$\begin{aligned} i_K(u)=i_K(x),\quad \nu _K(u)=\nu _K(x), \end{aligned}$$
(5.7)

which follow from the definitions (5.4) and (5.5). The following important formula was proved in Lemma 6.4 of [29]:

$$\begin{aligned} i_K(x) = 2n([KT/{2\pi }]+1)+i^v(x) \equiv d(K)+i^v(x), \end{aligned}$$
(5.8)

where the index \(i^v(x)\) does not depend on K, but only on \(H_a\).

By the proof of Proposition 2 of [28], we have that \(v\in E\) belongs to the null space of \(Q_{a, K}\) if and only if \(z=M_K v\) is a solution of the linearized system

$$\begin{aligned} \dot{z}(t) = JH_a''(x(t))z(t). \end{aligned}$$
(5.9)

Thus the nullity in (5.7) is independent of K, which we denote by \(\nu ^v(x)\equiv \nu _K(u)= \nu _K(x)\).

By Proposition 2.11 of [22], the index \(i^v(x)\) and nullity \(\nu ^v(x)\) coincide with those defined for the Hamiltonian \(H(x)=j(x)^\alpha \) for all \(x\in \mathbf{R}^{2n}\) and some \({\alpha }\in (1,2)\). Especially \(1\le \nu ^v(x)\le 2n-1\) always holds.

We have a natural \(S^1\)-action on X or E defined by

$$\begin{aligned} \theta \cdot u(t)=u(\theta +t),\quad \forall \, \theta \in S^1, \, t\in \mathbf{R}. \end{aligned}$$
(5.10)

Clearly both of \(F_{a, K}\) and \(\Psi _{a, K}\) are \(S^1\)-invariant. For any \(\kappa \in \mathbf{R}\), we denote by

$$\begin{aligned} \Lambda _{a, K}^\kappa= & {} \left\{ u\in L^{2}\left( \mathbf{R}/{T \mathbf{Z}}; \mathbf{R}^{2n}\right) \;|\;\Psi _{a,K}(u)\le \kappa \right\} , \end{aligned}$$
(5.11)
$$\begin{aligned} X_{a, K}^\kappa= & {} \left\{ x\in W^{1, 2}\left( \mathbf{R}/(T \mathbf{Z}),\mathbf{R}^{2n}\right) \;|\;F_{a, K}(x)\le \kappa \right\} . \end{aligned}$$
(5.12)

For a critical point u of \(\Psi _{a, K}\) and the corresponding \(x=M_K u\) of \(F_{a, K}\), let

$$\begin{aligned} {\Lambda }_{a,K}(u)= & {} {\Lambda }_{a,K}^{\Psi _{a, K}(u)} = \left\{ w\in L^{2}\left( \mathbf{R}/(T\mathbf{Z}), \mathbf{R}^{2n}\right) \;|\; \Psi _{a, K}(w)\le \Psi _{a,K}(u)\right\} , \end{aligned}$$
(5.13)
$$\begin{aligned} X_{a,K}(x)= & {} X_{a,K}^{F_{a,K}(x)} = \left\{ y\in W^{1, 2}\left( \mathbf{R}/(T\mathbf{Z}), \mathbf{R}^{2n}\right) \;|\; F_{a,K}(y)\le F_{a,K}(x)\right\} . \end{aligned}$$
(5.14)

Clearly, both sets are \(S^1\)-invariant. Denote by \(\mathrm{crit}(\Psi _{a, K})\) the set of critical points of \(\Psi _{a, K}\). Because \(\Psi _{a,K}\) is \(S^1\)-invariant, \(S^1\cdot u\) becomes a critical orbit if \(u\in \mathrm{crit}(\Psi _{a, K})\). Note that by the condition (F), the number of critical orbits of \(\Psi _{a, K}\) is finite. Hence as usual we can make the following definition.

Definition 5.1

Suppose u is a nonzero critical point of \(\Psi _{a, K}\), and \(\mathcal{N}\) is an \(S^1\)-invariant open neighborhood of \(S^1\cdot u\) such that \(\mathrm{crit}(\Psi _{a,K})\cap ({\Lambda }_{a,K}(u)\cap \mathcal{N}) = S^1\cdot u\). Then the \(S^1\)-critical modules of \(S^1\cdot u\) are defined by

$$\begin{aligned} C_{S^1,\; q}\left( \Psi _{a, K}, \;S^1\cdot u\right) =H_{q}\left( \left( \Lambda _{a, K}(u)\cap \mathcal{N}\right) _{S^1},\; \left( \left( \Lambda _{a,K}(u){\setminus } S^1\cdot u\right) \cap \mathcal{N}\right) _{S^1}\right) . \end{aligned}$$

Similarly, we define the \(S^1\)-critical modules \(C_{S^1,\; q}(F_{a, K}, \;S^1\cdot x)\) of \(S^1\cdot x\) for \(F_{a, K}\).

We fix a and let \(u_K\ne 0\) be a critical point of \(\Psi _{a, K}\) with multiplicity \(\mathrm{mul}(u_K)=m\), that is, \(u_K\) corresponds to a closed characteristic \((\tau , y)\subset \Sigma \) with \((\tau , y)\) being m-iteration of some prime closed characteristic. Precisely, we have \(u_K=-J\dot{x}+Kx\) with x being a solution of (5.1) and \(x=\rho y(\frac{\tau t}{T})\) with \(\frac{{\varphi }_a^\prime (\rho )}{\rho }=\frac{\tau }{aT}\). Moreover, \((\tau , y)\) is a closed characteristic on \(\Sigma \) with minimal period \(\frac{\tau }{m}\). By Lemma 2.10 of [22], we construct a finite dimensional \(S^1\)-invariant subspace G of \(L^{2}(\mathbf{R}/{T \mathbf{Z}}; \mathbf{R}^{2n})\) and a functional \(\psi _{a,K}\) on G. For any \(p\in \mathbf{N}\) satisfying \(p\tau <aT\), we choose K such that \(pK\notin \frac{2\pi }{T}\mathbf{Z}\), then the pth iteration \(u_{pK}^p\) of \(u_K\) is given by \(-J\dot{x}^p+pKx^p\), where \(x^p\) is the unique solution of (5.1) corresponding to \((p\tau , y)\) and is a critical point of \(F_{a, pK}\), that is, \(u_{pK}^p\) is the critical point of \(\Psi _{a, pK}\) corresponding to \(x^p\). Denote by \(g_{pK}^p\) the critical point of \(\psi _{a,pK}\) corresponding to \(u_{pK}^p\) and let \(\widetilde{\Lambda }_{a,K}(g_K)=\{g\in G \;|\; \psi _{a, K}(g)\le \psi _{a, K}(g_K)\}\).

Now we use the theory of Gromoll and Meyer, denote by \(W(g_{pK}^p)\) the local characteristic manifold of \(g_{pK}^p\). Then we have

Proposition 5.2

(cf. Proposition 4.2 of [22]) For any \(p\in \mathbf{N}\), we choose K such that \(pK\notin \frac{2\pi }{T}\mathbf{Z}\). Let \(u_K\ne 0\) be a critical point of \(\Psi _{a, K}\) with \(\mathrm{mul}(u_K)=1\), \(u_K=-J\dot{x}+Kx\) with x being a critical point of \(F_{a, K}\). Then for all \(q\in \mathbf{Z}\), we have

$$\begin{aligned}&C_{S^1,\; q}(\Psi _{a,pK},\;S^1\cdot u_{pK}^p) \nonumber \\&\quad \cong \left( \frac{}{}H_{q-i_{pK}(u_{pK}^p)}(W(g_{pK}^p)\cap \widetilde{\Lambda }_{a,pK}(g_{pK}^p),(W(g_{pK}^p) {\setminus }\{g_{pK}^p\})\cap \widetilde{\Lambda }_{a,pK}(g_{pK}^p))\right) ^{\beta (x^p)\mathbf{Z}_p}, \nonumber \\ \end{aligned}$$
(5.15)

where \(\beta (x^p)=(-1)^{i_{pK}(u_{pK}^p)-i_K(u_K)}=(-1)^{i^v(x^p)-i^v(x)}\). Thus

$$\begin{aligned} C_{S^1,q}\left( \Psi _{a,pK},\;S^1\cdot u_{pK}^p\right) =0 \quad if q<i_{pK}\left( u_{pK}^p\right) ~ or ~q>i_{pK}\left( u_{pK}^p\right) +\nu _{pK}\left( u_{pK}^p\right) -1.\nonumber \\ \end{aligned}$$
(5.16)

In particular, if \(u_{pK}^p\) is non-degenerate, i.e., \(\nu _{pK}(u_{pK}^p)=1\), then

$$\begin{aligned} C_{S^1,\; q}(\Psi _{a,pK},\;S^1\cdot u_{pK}^p) = \left\{ \begin{array}{ll} \mathbf{Q}, &{} \quad {\mathrm{if}}\;q=i_{pK}\left( u_{pK}^p\right) \;{\mathrm{and}}\;\beta (x^p)=1, \\ 0, &{}\quad {\mathrm{otherwise}}. \\ \end{array}\right. \end{aligned}$$
(5.17)

We make the following definition:

Definition 5.3

For any \(p\in \mathbf{N}\), we choose K such that \(pK\notin \frac{2\pi }{T}\mathbf{Z}\). Let \(u_K\ne 0\) be a critical point of \(\Psi _{a,K}\) with \(\mathrm{mul}(u_K)=1\), \(u_K=-J\dot{x}+Kx\) with x being a critical point of \(F_{a, K}\). Then for all \(l\in \mathbf{Z}\), let

Here \(k_l(u_{pK}^p)\)’s are called critical type numbers of \(u_{pK}^p\).

By Theorem 3.3 of [22], we obtain that \(k_l(u_{pK}^p)\) is independent of the choice of K and denote it by \(k_l(x^p)\), here \(k_l(x^p)\)’s are called critical type numbers of \(x^p\).

We have the following properties for critical type numbers:

Proposition 5.4

(cf. Proposition 4.6 of [22]) Let \(x\ne 0\) be a critical point of \(F_{a,K}\) with \(\mathrm{mul}(x)=1\) corresponding to a critical point \(u_K\) of \(\Psi _{a, K}\). Then there exists a minimal \(K(x)\in \mathbf{N}\) such that

$$\begin{aligned}&\nu ^v\left( x^{p+K(x)}\right) =\nu ^v(x^p),\quad i^v\left( x^{p+K(x)}\right) -i^v(x^p)\in 2\mathbf{Z}, \quad \forall p\in \mathbf{N}, \end{aligned}$$
(5.18)
$$\begin{aligned}&k_l\left( x^{p+K(x)}\right) =k_l(x^p), \quad \forall p\in \mathbf{N},\;l\in \mathbf{Z}. \end{aligned}$$
(5.19)

We call K(x) the minimal period of critical modules of iterations of the functional \(F_{a, K}\) at x.

For every closed characteristic \((\tau , y)\) on \(\Sigma \), let \(aT>\tau \) and choose \({\varphi }_a\) as above. Determine \(\rho \) uniquely by \(\frac{{\varphi }_a'(\rho )}{\rho }=\frac{\tau }{aT}\). Let \(x=\rho y(\frac{\tau t}{T})\). Then we define the index \(i(\tau , y)\) and nullity \(\nu (\tau , y)\) of \((\tau , y)\) by

$$\begin{aligned} i(\tau , y)=i^v(x), \quad \nu (\tau , y)=\nu ^v(x). \end{aligned}$$

Then the mean index of \((\tau , y)\) is defined by

$$\begin{aligned} \hat{i}(\tau , y) = \lim _{m\rightarrow \infty }\frac{i(m\tau , y)}{m}. \end{aligned}$$
(5.20)

Note that by Proposition 2.11 of [22], the index and nullity are well defined and are independent of the choice of a.

For a closed characteristic \((\tau , y)\) on \(\Sigma \), we simply denote by \(y^m\equiv (m\tau , y)\) the mth iteration of y for \(m\in \mathbf{N}\). By Proposition 3.2 of [22], we can define the critical type numbers \(k_l(y^m)\) of \(y^m\) to be \(k_l(x^m)\), where \(x^m\) is the critical point of \(F_{a, K}\) corresponding to \(y^m\). We also define \(K(y)=K(x)\).

Remark 5.5

(cf. Remark 4.10 of [22]) Note that \(k_l(y^m)=0\) for \(l\notin [0, \nu (y^m)-1]\) and it can take only values 0 or 1 when \(l=0\) or \(l=\nu (y^m)-1\). Moreover, the following facts are useful:

  1. (i)

    \(k_0(y^m)=1\) implies \(k_l(y^m)=0\) for \(1\le l\le \nu (y^m)-1\).

  2. (ii)

    \(k_{\nu (y^m)-1}(y^m)=1\) implies \(k_l(y^m)=0\) for \(0\le l\le \nu (y^m)-2\).

  3. (iii)

    \(k_l(y^m)\ge 1\) for some \(1\le l\le \nu (y^m)-2\) implies \(k_0(y^m)=k_{\nu (y^m)-1}(y^m)=0\).

  4. (iv)

    In particular, only one of the \(k_l(y^m)\hbox {s}\) for \(0\le l\le \nu (y^m)-1\) can be non-zero when \(\nu (y^m)\le 3\).

For a closed characteristic \((\tau ,y)\) on \(\Sigma \), the average Euler characteristic \(\hat{\chi }(y)\) of y is defined by

$$\begin{aligned} \hat{\chi }(y)=\frac{1}{K(y)}\sum _{\begin{array}{c} 1\le m\le K(y)\\ 0\le l\le 2n-2 \end{array}} (-1)^{i(y^{m})+l}k_l(y^{m}). \end{aligned}$$
(5.21)

\(\hat{\chi }(y)\) is a rational number. In particular, if all \(y^m\hbox {s}\) are non-degenerate, then by Proposition 5.4 we have

$$\begin{aligned} \hat{\chi }(y) = \left\{ \begin{array}{ll} (-1)^{i(y)}, &{}\quad {\mathrm{if}}\quad i(y^2)-i(y)\in 2\mathbf{Z}, \\ \frac{(-1)^{i(y)}}{2}, &{}\quad {\mathrm{otherwise}}. \\ \end{array}\right. \end{aligned}$$
(5.22)

We have the following mean index identities for closed characteristics.

Theorem 5.6

Suppose that \({\Sigma }\in \mathcal{H}_{st}(2n)\) satisfies \(\,^{\#}\mathcal{T}({\Sigma })<+\infty \). Denote all the geometrically distinct prime closed characteristics by \(\{(\tau _j,\; y_j)\}_{1\le j\le k}\). Then the following identities hold

$$\begin{aligned} \sum _{\begin{array}{c} 1\le j\le k\\ \hat{i}(y_j)>0 \end{array}}\frac{\hat{\chi }(y_j)}{\hat{i}(y_j)}= & {} \frac{1}{2}, \end{aligned}$$
(5.23)
$$\begin{aligned} \sum _{\begin{array}{c} 1\le j\le k\\ \hat{i}(y_j)<0 \end{array}}\frac{\hat{\chi }(y_j)}{\hat{i}(y_j)}= & {} 0. \end{aligned}$$
(5.24)

Let \(F_{a, K}\) be a functional defined by (5.4) for some \(a, K\in \mathbf{R}\) sufficiently large and let \(\epsilon >0\) be small enough such that \([-\epsilon , 0)\) contains no critical values of \(F_{a, K}\). For b large enough, The normalized Morse series of \(F_{a, K}\) in \( X^{-{\epsilon }}{\setminus } X^{-b}\) is defined, as usual, by

$$\begin{aligned} M_a(t)=\sum _{q\ge 0,\;1\le j\le p} \dim C_{S^1,\;q}(F_{a, K}, \;S^1\cdot v_j)t^{q-d(K)}, \end{aligned}$$
(5.25)

where we denote by \(\{S^1\cdot v_1, \ldots , S^1\cdot v_p\}\) the critical orbits of \(F_{a, K}\) with critical values less than \(-\epsilon \). The Poincaré series of \(H_{S^1, *}( X, X^{-{\epsilon }})\) is \(t^{d(K)}Q_a(t)\), according to Theorem 5.1 of [22], if we set \(Q_a(t)=\sum _{k\in \mathbf{Z}}{q_kt^k}\), then

(5.26)

where I is an interval of \(\mathbf{Z}\) such that \(I \cap [i(\tau , y), i(\tau , y)+\nu (\tau , y)-1]=\emptyset \) for all closed characteristics \((\tau , y)\) on \(\Sigma \) with \(\tau \ge aT\). Then by Section 6 of [22], we have

$$\begin{aligned} M_a(t)-\frac{1}{1-t^2}+Q_a(t) = (1+t)U_a(t), \end{aligned}$$
(5.27)

where \(U_a(t)=\sum _{i\in \mathbf{Z}}{u_it^i}\) is a Laurent series with nonnegative coefficients. If there is no closed characteristic with \(\hat{i}=0\), then

$$\begin{aligned} M(t)-\frac{1}{1-t^2}=(1+t)U(t), \end{aligned}$$
(5.28)

where \(M(t)=\sum _{i\in \mathbf{Z}}{m_it^i}\) denotes the limit of \(M_a(t)\) as a tends to infinity, \(U(t)=\sum _{i\in \mathbf{Z}}{u_it^i}\) denotes the limit of \(U_a(t)\) as a tends to infinity and possesses only non-negative coefficients. Specially, suppose that there exists an integer \(p<0\) such that the coefficients of M(t) satisfy \(m_p>0\) and \(m_q=0\) for all integers \(q<p\). Then (5.28) implies

$$\begin{aligned} m_{p+1} \ge m_p. \end{aligned}$$
(5.29)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Duan, H., Liu, H. Multiplicity and ellipticity of closed characteristics on compact star-shaped hypersurfaces in \(\mathbf{R}^{2n}\) . Calc. Var. 56, 65 (2017). https://doi.org/10.1007/s00526-017-1173-1

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s00526-017-1173-1

Mathematics Subject Classification

Navigation