Skip to main content
Log in

Existence of closed characteristics on compact convex hypersurfaces in \(\mathbf{R}^{2n}\)

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

Abstract

In this paper, we prove there exist at least \([\frac{n+1}{2}]+1\) geometrically distinct closed characteristics on every compact convex hypersurface \({\Sigma }\) in \(\mathbf{R}^{2n}\), where \(n\ge 2\). In particular, this gives a new proof in the case \(n=3\) to a long standing conjecture in Hamiltonian analysis. Moreover, there exist at least \([\frac{n}{2}]+1\) geometrically distinct non-hyperbolic closed characteristics on \({\Sigma }\) provided the number of geometrically distinct closed characteristics on \({\Sigma }\) is finite.

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. Dell’Antonio, G., D’Onofrio, B., Ekeland, I.: Les systém hamiltoniens convexes et pairs ne sont pas ergodiques en general. C. R. Acad. Sci. Paris Ser. I(315), 1413–1415 (1992)

  2. Ekeland, I.: Une théorie de Morse pour les systèmes hamiltoniens convexes. Ann. IHP. Anal. Non Linéaire. 1, 19–78 (1984)

    MathSciNet  MATH  Google Scholar 

  3. Ekeland, I.: An index throry for periodic solutions of convex Hamiltonian systems. Proc. Symp. Pure Math. 45, 395–423 (1986)

    Article  MathSciNet  Google Scholar 

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

    Book  MATH  Google Scholar 

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

    Article  MathSciNet  MATH  Google Scholar 

  6. Ekeland, I., Lasry, J.: On the number of periodic trajectories for a Hamiltonian flow on a convex energy surface. Ann. Math. 112, 283–319 (1980)

    Article  MathSciNet  MATH  Google Scholar 

  7. 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)

    MathSciNet  Google Scholar 

  8. Fadell, E., Rabinowitz, P.: Generalized cohomological index theories for Lie group actions with an application to bifurcation questions for Hamiltonian systems. Invent. Math. 45(2), 139–174 (1978)

    Article  MathSciNet  MATH  Google Scholar 

  9. Gromoll, D., Meyer, W.: On differentiable functions with isolated critical points. Topology 8, 361–369 (1969)

    Article  MathSciNet  MATH  Google Scholar 

  10. 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 

  11. Horn, J.: Beiträge zur Theorie der kleinen Schwingungen. Z. Math. Phys. 48, 400–434 (1903)

    MATH  Google Scholar 

  12. Hu, X., Ou, Y.: Stability of closed characteristics on compact convex hypersurfaces in \({ R}^{2n}\). arXiv:1405.4057

  13. Liapunov, A.M., Problème général de la stabilité du mouvement. Ann. Fac. Sci. Toulouse 9 (1907) 203–474. Russian original, Kharkov Math. Soc. 1892. Reedited, Princeton U. Press, 1949. Reedited, Gabay, Paris (1989)

  14. Long, Y.: Hyperbolic closed characteristics on compact convex smooth hypersurfaces in \({ R}^{2n}\). J. Differ. Equ. 150, 227–249 (1998)

    Article  MATH  Google Scholar 

  15. 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 

  16. Long, Y.: Index theory for symplectic paths with applications. In: Progress in Math., vol. 207. Birkhäuser, Basel (2002)

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

    Article  MathSciNet  MATH  Google Scholar 

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

    Article  MathSciNet  Google Scholar 

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

    MathSciNet  MATH  Google Scholar 

  20. Wang, W., Hu, X., Long, Y.: Resonance identity, stability and multiplicity of closed characteristics on compact convex hypersurfaces. Duke Math. J. 139(3), 411–462 (2007)

    Article  MathSciNet  MATH  Google Scholar 

  21. Wang, W.: Stability of closed characteristics on compact convex hypersurfaces in \({ R}^6\). J. Eur. Math. Soc. 11(3), 575–596 (2009)

    Article  MathSciNet  MATH  Google Scholar 

  22. Wang, W.: Irrationally elliptic closed characteristics on compact convex hypersurfaces in \({R}^6\). J. Funct. Anal. 267(3), 799–841 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  23. Weinstein, A.: Normal modules for nonlinear Hamiltonian systems. Invent. Math. 20, 45–57 (1973)

    Article  Google Scholar 

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

    Article  MATH  Google Scholar 

Download references

Acknowledgments

I would like to sincerely thank the referees for their careful readings of this paper and for their valuable comments and suggestions. I would like to sincerely thank my Ph.D. thesis advisor, Professor Yiming Long, for introducing me to Hamiltonian dynamics and for his valuable help and encouragement during my research. I would like to say how enjoyable it is to work with him.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Wei Wang.

Additional information

Communicated by P. Rabinowitz.

Partially supported by National Natural Science Foundation of China Nos. 11222105, 11431001, Foundation for the Author of National Excellent Doctoral Dissertation of People’s Republic of China No. 201017.

Appendix: Index iteration theory for closed characteristics

Appendix: Index iteration theory for closed characteristics

In this section, we review briefly the index theory for symplectic paths developed by Long and his coworkers. All the details can be found in [16].

The symplectic group \(\mathrm{Sp}(2n)\) is defined by

$$\begin{aligned} \mathrm{Sp}(2n) = \{M\in \mathrm{GL}(2n,\mathbf{R})|M^TJM=J\}. \end{aligned}$$

For \(\tau >0\) we consider paths in \(\mathrm{Sp}(2n)\):

$$\begin{aligned} \mathcal{P}_{\tau }(2n) = \{{\gamma }\in C([0,\tau ],\mathrm{Sp}(2n))|{\gamma }(0)=I_{2n}\}. \end{aligned}$$

As in [15] we define

$$\begin{aligned} D_{{\omega }}(M) = (-1)^{n-1}\overline{{\omega }}^n\det (M-{\omega }I_{2n}), \quad \forall {\omega }\in \mathbf{U},\, M\in \mathrm{Sp}(2n). \end{aligned}$$

For \({\omega }\in \mathbf{U}\), we define the following codimension 1 hypersurface in \(\mathrm{Sp}(2n)\) as in [15]:

$$\begin{aligned} \mathrm{Sp}(2n)_{{\omega }}^0 = \{M\in \mathrm{Sp}(2n)| D_{{\omega }}(M)=0\}. \end{aligned}$$

For any \(M\in \mathrm{Sp}(2n)_{{\omega }}^0\), we define the co-orientation of \(\mathrm{Sp}(2n)_{{\omega }}^0\) at M by the positive direction \(\frac{d}{dt}Me^{t J}|_{t=0}\) of the path \(Me^{t J}\) with \(0\le t\le {\epsilon }\) and \({\epsilon }>0\) being small enough. Denote by

$$\begin{aligned} \mathrm{Sp}(2n)_{{\omega }}^{*}= & {} \mathrm{Sp}(2n){\setminus }\mathrm{Sp}(2n)_{{\omega }}^0, \\ \mathcal{P}_{\tau ,{\omega }}^{*}(2n)= & {} \{{\gamma }\in \mathcal{P}_{\tau }(2n)|{\gamma }(\tau )\in \mathrm{Sp}(2n)_{{\omega }}^{*}\}, \\ \mathcal{P}_{\tau ,{\omega }}^0(2n)= & {} \mathcal{P}_{\tau }(2n){\setminus }\mathcal{P}_{\tau ,{\omega }}^{*}(2n). \end{aligned}$$

For any two paths \(\xi \) and \(\eta :[0,\tau ]\rightarrow \mathrm{Sp}(2n)\) with \(\xi (\tau )=\eta (0)\), we define as usual:

$$\begin{aligned} \eta *\xi (t) = \left\{ \begin{array}{ll} \xi (2t), &{} \quad \mathrm{if}\;0\le t\le \tau /2, \\ \eta (2t-\tau ), &{} \quad \mathrm{if}\; \tau /2\le t\le \tau . \\ \end{array}\right. \end{aligned}$$

As in [16], for any two \(2m_k\times 2m_k\) matrices of square block form \(M_k=\left( \begin{array}{ll}A_k&{}B_k\\ C_k&{}D_k\\ \end{array}\right) \) with \(k=1, 2\), we define the \(\;\mathrm{\diamond }\)-product of \(M_1\) and \(M_2\) to be the following \(2(m_1+m_2)\times 2(m_1+m_2)\) matrix \(M_1\mathrm{\diamond }M_2\):

$$\begin{aligned} M_1\mathrm{\diamond }M_2=\left( \begin{array}{llllll}A_1&{} 0&{}B_1&{} 0 \\ 0&{}A_2&{} 0&{}B_2 \\ C_1&{} 0&{}D_1&{} 0 \\ 0&{}C_2&{} 0&{}D_2 \\ \end{array}\right) . \end{aligned}$$

Denote by \(M^{\mathrm{\diamond }k}\) the k-fold \(\mathrm{\diamond }\)-product \(M\mathrm{\diamond }\cdots \mathrm{\diamond }M\). One can easily check that the \(\mathrm{\diamond }\)-product of any two symplectic matrices is symplectic. For any two paths \({\gamma }_j\in \mathcal{P}_{\tau }(2n_j)\) with \(j=0\) and 1, let \({\gamma }_0\mathrm{\diamond }{\gamma }_1(t)= {\gamma }_0(t)\mathrm{\diamond }{\gamma }_1(t)\) for \(t\in [0,\tau ]\).

We define a special path \(\xi _n\):

$$\begin{aligned} \xi _n(t) = \left( \begin{array}{ll}2-\frac{t}{\tau } &{} 0 \\ 0 &{} (2-\frac{t}{\tau })^{-1}\\ \end{array}\right) ^{\mathrm{\diamond }n} \quad \mathrm{for}\;0\le t\le \tau . \end{aligned}$$
(4.1)

Definition 4.1

(cf. [15, 16]) For \({\omega }\in \mathbf{U}\) and \(M\in \mathrm{Sp}(2n)\), define

$$\begin{aligned} \nu _{{\omega }}(M)=\dim _{\mathbf{C}}\ker _{\mathbf{C}}(M - {\omega }I_{2n}). \end{aligned}$$
(4.2)

For \(\tau >0\) and \({\gamma }\in \mathcal{P}_{\tau }(2n)\), define

$$\begin{aligned} \nu _{{\omega }}({\gamma })= \nu _{{\omega }}({\gamma }(\tau )). \end{aligned}$$
(4.3)

If \({\gamma }\in \mathcal{P}_{\tau ,{\omega }}^{*}(2n)\), define

$$\begin{aligned} i_{{\omega }}({\gamma }) = [\mathrm{Sp}(2n)_{{\omega }}^0:{\gamma }*\xi _n], \end{aligned}$$
(4.4)

where the right hand side of (4.4) is the usual homotopy intersection number, and the orientation of \({\gamma }*\xi _n\) is its positive time direction under homptopy with fixed end points.

If \({\gamma }\in \mathcal{P}_{\tau ,{\omega }}^0(2n)\), we let \(\mathcal {F}({\gamma })\) be the set of all open neighborhoods of \({\gamma }\) in \(\mathcal{P}_{\tau }(2n)\), and define

$$\begin{aligned} i_{{\omega }}({\gamma }) = \sup _{U\in \mathcal {F}({\gamma })}\inf \{i_{{\omega }}(\beta )| \beta \in U\cap \mathcal{P}_{\tau ,{\omega }}^{*}(2n)\}. \end{aligned}$$
(4.5)

Then the tuple

$$\begin{aligned} (i_{{\omega }}({\gamma }), \nu _{{\omega }}({\gamma })) \in \mathbf{Z}\times \{0,1,\ldots ,2n\}, \end{aligned}$$

is called the index function of \({\gamma }\) at \({\omega }\).

For a symplectic path \({\gamma }\in \mathcal{P}_{\tau }(2n)\) and \(m\in \mathbf{N}\), we define its mth iteration \({\gamma }^m:[0,m\tau ]\rightarrow \mathrm{Sp}(2n)\) to be

$$\begin{aligned} {\gamma }^m(t) = {\gamma }(t-j\tau ){\gamma }(\tau )^j, \quad \mathrm{for}\; j\tau \le t\le (j+1)\tau ,\;j=0,1,\ldots ,m-1. \end{aligned}$$
(4.6)

We also denote the extended path on \([0,+\infty )\) by \({\gamma }\).

Definition 4.2

(cf. [15, 16]) For \({\gamma }\in \mathcal{P}_{\tau }(2n)\), define

$$\begin{aligned} (i({\gamma },m), \nu ({\gamma },m)) = (i_1({\gamma }^m), \nu _1({\gamma }^m)), \quad \forall m\in \mathbf{N}. \end{aligned}$$
(4.7)

We define the mean index \(\hat{i}({\gamma },m)\) per \(m\tau \) for \(m\in \mathbf{N}\) to be

$$\begin{aligned} \hat{i}({\gamma },m) = \lim _{k\rightarrow +\infty }\frac{i({\gamma },mk)}{k}. \end{aligned}$$
(4.8)

For \(M\in \mathrm{Sp}(2n)\) and \({\omega }\in \mathbf{U}\), we define the splitting numbers \(S_M^{\pm }({\omega })\) of M at \({\omega }\) to be

$$\begin{aligned} S_M^{\pm }({\omega }) = \lim _{{\epsilon }\rightarrow 0^+}i_{{\omega }\exp (\pm \sqrt{-1}{\epsilon })}({\gamma }) - i_{{\omega }}({\gamma }), \end{aligned}$$
(4.9)

where \({\gamma }\in \mathcal{P}_{\tau }(2n)\) is any path satisfying \({\gamma }(\tau )=M\).

Given a path \(\gamma \in \mathcal{P}_{\tau }(2n)\), we want to deform it to a new path \(\eta \) in \(\mathcal{P}_{\tau }(2n)\) such that

$$\begin{aligned} i_1(\gamma ^m)=i_1(\eta ^m),\quad \nu _1(\gamma ^m)=\nu _1(\eta ^m), \quad \forall m\in \mathbf{N}, \end{aligned}$$
(4.10)

and that \((i_1(\eta ^m),\nu _1(\eta ^m))\) is easy enough to compute. This leads to finding homotopies \(\delta :[0,1]\times [0,\tau ]\rightarrow \mathrm{Sp}(2n)\) starting from \(\gamma \) in \(\mathcal{P}_{\tau }(2n)\) and keeping the end points of the homotopy always stay in a certain suitably chosen maximal subset of \(\mathrm{Sp}(2n)\) so that (4.10) holds. Actually this set was first discovered in [15] as the path connected component \(\Omega ^0(M)\) containing \(M=\gamma (\tau )\) of the set

$$\begin{aligned} \Omega (M)= & {} \{N\in \mathrm{Sp}(2n)|\sigma (N)\cap \mathbf{U}=\sigma (M)\cap \mathbf{U}\quad \mathrm{and} \nonumber \\&\;\nu _{\lambda }(N)=\nu _{\lambda }(M),\quad \forall \lambda \in \sigma (M)\cap \mathbf{U}\}. \end{aligned}$$
(4.11)

Here \(\Omega ^0(M)\) is called the homotopy component of M in \(\mathrm{Sp}(2n)\).

Note that we have the following \(2\times 2\) and \(4\times 4\) symplectic matrix as basic normal forms (cf. [15, 16]):

$$\begin{aligned} D(\lambda )= & {} \left( \begin{array}{ll}{\lambda }&{} 0\\ 0 &{} {\lambda }^{-1}\\ \end{array}\right) , \quad {\lambda }=\pm 2,\end{aligned}$$
(4.12)
$$\begin{aligned} N_1({\lambda },b)= & {} \left( \begin{array}{ll}{\lambda }&{} b\\ 0 &{} {\lambda }\\ \end{array}\right) , \quad {\lambda }=\pm 1, b=\pm 1, 0, \end{aligned}$$
(4.13)
$$\begin{aligned} R({\theta })= & {} \left( \begin{array}{ll}\cos {\theta }&{} -\sin {\theta }\\ \sin {\theta }&{} \cos {\theta }\\ \end{array}\right) , \quad {\theta }\in (0,\pi )\cup (\pi ,2\pi ), \end{aligned}$$
(4.14)
$$\begin{aligned} N_2({\omega },b)= & {} \left( \begin{array}{ll}R({\theta }) &{} b\\ 0 &{} R({\theta })\\ \end{array}\right) , \quad {\theta }\in (0,\pi )\cup (\pi ,2\pi ), \end{aligned}$$
(4.15)

where \(b=\left( \begin{array}{llll}b_1 &{} b_2\\ b_3 &{} b_4\\ \end{array}\right) \) with \(b_i\in \mathbf{R}\) and \(b_2\not =b_3\). Moreover, we call \( N_2(\omega , b)\) trivial if \((b_2-b_3)\sin \theta >0\); and non-trivial if \((b_2-b_3)\sin \theta <0\).

We have the following properties for Splitting numbers:

Lemma 4.3

(cf. [15] and Lemma 9.1.5 of [16]) Splitting numbers \(S_M^{\pm }({\omega })\) are independent of the choice of the path \({\gamma }\in \mathcal{P}_\tau (2n)\) satisfying \({\gamma }(\tau )=M\) appeared in (4.9). Moreover, for \({\omega }\in \mathbf{U}\) and \(M\in \mathrm{Sp}(2n)\), splitting numbers \(S_N^{\pm }({\omega })\) are constant for all \(N\in {\Omega }^0(M)\).

Lemma 4.4

(cf. [15], Lemma 9.1.5 and List 9.1.12 of [16]) For \(M\in \mathrm{Sp}(2n)\) and \({\omega }\in \mathbf{U}\), we have

$$\begin{aligned} S_M^{\pm }({\omega })= & {} 0, \quad { if}\;{\omega }\not \in {\sigma }(M). \end{aligned}$$
(4.16)
$$\begin{aligned} S_{N_1(1,a)}^+(1)= & {} \left\{ \begin{array}{ll}1, &{}\quad \mathrm{if}\;a\ge 0, \\ 0, &{}\quad \mathrm{if}\; a< 0. \\ \end{array}\right. \end{aligned}$$
(4.17)

For \(M_i\in \mathrm{Sp}(2n_i)\) with \(i=0\) and 1, there holds

$$\begin{aligned} S^{\pm }_{M_0\mathrm{\diamond }M_1}({\omega }) = S^{\pm }_{M_0}({\omega }) + S^{\pm }_{M_1}({\omega }), \quad \forall {\omega }\in \mathbf{U}. \end{aligned}$$
(4.18)

We have the following symplectic additivity property for index functions:

Theorem 4.5

(cf. Theorem 6.1 of [17] or Theorem 6.2.7 of [16]) For any \(\gamma _j\in \mathcal{P}_\tau (2n_j)\) with \(j=0, 1\), we have

$$\begin{aligned} i_\omega ({\gamma }_0\diamond {\gamma }_1)=i_\omega ({\gamma }_0)+i_\omega ({\gamma }_1). \end{aligned}$$
(4.19)

Now we use the above defined index function to study the Morse indices of the Clarke–Ekeland dual functional \(\Phi \) at a critical point. Let \(\Sigma \in \mathcal{H}(2n)\). Using notations in §1, for any closed characteristic \((\tau ,y)\) and \(m\in \mathbf{N}\), we define its mth iteration \(y^m:\mathbf{R}/(m\tau \mathbf{Z})\rightarrow \mathbf{R}^{2n}\) by

$$\begin{aligned} y^m(t) = y(t-j\tau ), \quad \mathrm{for}\; j\tau \le t\le (j+1)\tau , \quad j=0,1,2,\ldots , m-1. \end{aligned}$$
(4.20)

We still denote by y its extension to \([0,+\infty )\).

We define via Definition 4.2 the following

$$\begin{aligned} S^+(y)= & {} S_{{\gamma }_y(\tau )}^+(1), \end{aligned}$$
(4.21)
$$\begin{aligned} (i(y,m), \nu (y,m))= & {} (i({\gamma }_y,m), \nu ({\gamma }_y,m)), \end{aligned}$$
(4.22)
$$\begin{aligned} \hat{i}(y,m)= & {} \hat{i}({\gamma }_y,m), \end{aligned}$$
(4.23)

for all \(m\in \mathbf{N}\), where \({\gamma }_y\) is the associated symplectic path of \((\tau ,y)\).

The following theorem describe the relation between the above defined indices and the indices defined by Ekeland [24]. Thus we can compute the Morse indices of \(\Phi \) at a critical point by those of its associated symplectic paths.

Theorem 4.6

(cf. Lemma 1.1 of [17], Theorem 15.1.1 of [16]) Suppose \((\tau ,y)\) is a closed characteristic on \({\Sigma }\). Then we have

$$\begin{aligned} i(y^m)\equiv i(m\tau ,y)=i(y, m)-n,\quad \nu (y^m)\equiv \nu (m\tau , y)=\nu (y, m), \quad \forall m\in \mathbf{N}, \end{aligned}$$
(4.24)

where \(i(y^m)\) and \(\nu (y^m)\) are the index and nullity defined by Ekeland [24]. In particular, we have \(\hat{i}(y)=\hat{i}(y,\, 1)\), where \(\hat{i}(y)\) is the mean index of \((\tau ,\,y)\) defined by Ekeland.

The following is the precise index iteration formulae for symplectic paths, which is due to Long (cf. Theorem 8.3.1 and Corollary 8.3.2 of [16]). Using this formula, we can compute the indices of any iteration of a fixed symplectic path whenever we know its initial index and end point in \(\mathrm{Sp}(2n)\).

Theorem 4.7

Let \(\gamma \in \mathcal{P}_\tau (2n)\), Then there exists a path \(f\in C([0,1],\Omega ^0(\gamma (\tau ))\) such that \(f(0)=\gamma (\tau )\) and

$$\begin{aligned} f(1)= & {} N_1(1,1)^{\diamond p_-} \diamond I_{2p_0}\diamond N_1(1,-1)^{\diamond p_+} \diamond N_1(-1,1)^{\diamond q_-} \diamond (-I_{2q_0})\diamond N_1(-1,-1)^{\diamond q_+}\nonumber \\&\diamond R(\theta _1)\diamond \cdots \diamond R(\theta _r) \diamond N_2(\omega _1, u_1)\diamond \cdots \diamond N_2(\omega _{r_*}, u_{r_*})\nonumber \\&\diamond N_2({\lambda }_1, v_1)\diamond \cdots \diamond N_2({\lambda }_{r_0}, v_{r_0}) \diamond M_0 \end{aligned}$$
(4.25)

where \( N_2(\omega _j, u_j) \)s are non-trivial and \( N_2({\lambda }_j, v_j)\)s are trivial basic normal forms; \(\sigma (M_0)\cap U=\emptyset \); \(p_-\), \(p_0\), \(p_+\), \(q_-\), \(q_0\), \(q_+\), r, \(r_*\) and \(r_0\) are non-negative integers; \(\omega _j=e^{\sqrt{-1}\alpha _j}\), \( \lambda _j=e^{\sqrt{-1}\beta _j}\); \(\theta _j\), \(\alpha _j\), \(\beta _j\) \(\in (0, \pi )\cup (\pi , 2\pi )\); these integers and real numbers are uniquely determined by \(\gamma (\tau )\). Then using the functions defined in (1.7).

$$\begin{aligned} i(\gamma , m)= & {} m(i(\gamma , 1)+p_-+p_0-r)+2\sum _{j=1}^r E\left( \frac{m\theta _j}{2\pi }\right) -r -p_--p_0\nonumber \\&-\frac{1+(-1)^m}{2}(q_0+q_+)+2\left( \sum _{j=1}^{r_*}\varphi \left( \frac{m\alpha _j}{2\pi }\right) -r_*\right) . \end{aligned}$$
(4.26)
$$\begin{aligned} \nu (\gamma , m)= & {} \nu (\gamma , 1)+\frac{1+(-1)^m}{2}(q_-+2q_0+q_+)+2(r+r_*+r_0)\nonumber \\&-2\left( \sum _{j=1}^{r}\varphi \left( \frac{m\theta _j}{2\pi }\right) + \sum _{j=1}^{r_*}\varphi \left( \frac{m\alpha _j}{2\pi }\right) +\sum _{j=1}^{r_0}\varphi \left( \frac{m\beta _j}{2\pi }\right) \right) \end{aligned}$$
(4.27)
$$\begin{aligned} \hat{i}(\gamma , 1)=i(\gamma , 1)+p_-+p_0-r+\sum _{j=1}^r \frac{\theta _j}{\pi }. \end{aligned}$$
(4.28)

Moreover, we have \(i(\gamma , 1)\) is odd if \(f(1)=N_1(1, 1)\), \(I_2\), \(N_1(-1, 1)\), \(-I_2\), \(N_1(-1, -1)\) and \(R(\theta )\); \(i(\gamma , 1)\) is even if \(f(1)=N_1(1, -1)\) and \( N_2(\omega , b)\); \(i(\gamma , 1)\) can be any integer if \(\sigma (f(1)) \cap \mathbf{U}=\emptyset \).

We have the following iteration inequalities for the index function:

Theorem 4.8

(cf. Theorem 2.3 of [17]) Let \(\gamma \in \mathcal{P}_\tau (2n)\) and \(M = \gamma (\tau )\). Suppose that there exist \(P\in \mathrm{Sp}(2n)\) and \(Q \in \mathrm{Sp}(2n -2)\) such that \(M=P^{-1}(N_1(1, 1)\diamond Q)P\). Then for any \(m\in \mathbf{N}\), there holds

$$\begin{aligned} \nu (\gamma , m)-\frac{e(M)}{2}\!+\!1\le i(\gamma , m+1)-i(\gamma , m)-i(\gamma , 1) \le \nu (\gamma , 1)-\nu (\gamma , m+1)\!+\!\frac{e(M)}{2}, \end{aligned}$$

where e(M) is the total algebraic multiplicity of all eigenvalues of M on the unit circle in the complex plane \(\mathbf{C}\).

The following is the common index jump theorem of Long and Zhu.

Theorem 4.9

(cf. Theorems 4.1–4.4 of [17]) Let \(\gamma _k\in \mathcal{P}_{\tau _k}(2n)\) for \( k = 1,\ldots ,q\) be a finite collection of symplectic paths. Let \(M_k = \gamma _k(\tau _k)\). Suppose that there exist \(P_k\in \mathrm{Sp}(2n)\) and \(Q _k\in \mathrm{Sp}(2n -2)\) such that \(M_k=P_k^{-1}(N_1(1, 1)\diamond Q_k)P_k\) and \(\hat{i}(\gamma _k, 1) > 0\), for all \(k = 1,\ldots ,q\). Then there exist infinitely many \((T,m_1,\ldots ,m_q)\in \mathbf{N}^{q+1}\) such that

$$\begin{aligned} \nu (\gamma _k, 2m_k -1)= & {} \nu (\gamma _k, 1),\end{aligned}$$
(4.29)
$$\begin{aligned} \nu (\gamma _k, 2m_k +1)= & {} \nu (\gamma _k, 1),\end{aligned}$$
(4.30)
$$\begin{aligned} i(\gamma _k, 2m_k -1)+\nu (\gamma _k, 2m_k -1)= & {} 2T-(i(\gamma _k, 1)+2S^+_{M_k}(1)-\nu (\gamma _k, 1)),\end{aligned}$$
(4.31)
$$\begin{aligned} i(\gamma _k, 2m_k+1)= & {} 2T+i(\gamma _k, 1),\end{aligned}$$
(4.32)
$$\begin{aligned} i(\gamma _k, 2m_k)\ge & {} 2T-\frac{e(M_k)}{2}\ge 2T-n,\end{aligned}$$
(4.33)
$$\begin{aligned} i(\gamma _k, 2m_k)+\nu (\gamma _k, 2m_k)\le & {} 2T+\frac{e(M_k)}{2}-1\le 2T+n-1, \end{aligned}$$
(4.34)

for every \(k=1,\ldots ,q\). Moreover we have

$$\begin{aligned} \min \left\{ \left\{ \frac{m_k\theta }{\pi }\right\} ,\,1-\left\{ \frac{m_k\theta }{\pi }\right\} \right\} <\delta , \end{aligned}$$
(4.35)

whenever \(e^{\sqrt{-1}\theta }\in \sigma (M_k)\) and \(\delta \) can be chosen as small as we want (cf. (4.43) of [17]). More precisely, by (4.10) and (4.40) in [17], we have

$$\begin{aligned} m_k=\left( \left[ \frac{T}{M\hat{i}(\gamma _k, 1)}\right] +\chi _k\right) M,\quad 1\le k\le q, \end{aligned}$$
(4.36)

where \(\chi _k=0\) or 1 for \(1\le k\le q\) and \(\frac{M\theta }{\pi }\in \mathbf{Z}\) whenever \(e^{\sqrt{-1}\theta }\in \sigma (M_k)\) and \(\frac{\theta }{\pi }\in \mathbf{Q}\) for some \(1\le k\le q\). Furthermore, given \(M_0\in \mathbf{N}\), by the proof of Theorem 4.1 of [17], we may further require \(M_0|T\) (since the closure of the set \(\{\{Tv\}:T\in \mathbf{N}, \;M_0|T\}\) is still a closed additive subgroup of \(\mathbf T^h\) for some \(h\in \mathbf{N}\), where we use notations as (4.21) in [17]. Then we can use the proof of Step 2 in Theorem 4.1 of [17] to get T).

In fact, Let \(\mu _i=\sum _{\theta \in (0, 2\pi )}S_{M_i}^-(e^{\sqrt{-1}\theta })\) for \(1\le i\le q\) and \(\alpha _{i, j}=\frac{\theta _j}{\pi }\) where \(e^{\sqrt{-1}\theta _j}\in \sigma (M_i)\) for \(1\le j\le \mu _i\) and \(1\le i\le q\). Let \(h=q+\sum _{1\le i\le q}\mu _i\) and

$$\begin{aligned} v= & {} \left( \frac{1}{M\hat{i}(\gamma _1, 1)},\dots , \frac{1}{M\hat{i}(\gamma _q, 1)}, \frac{\alpha _{1, 1}}{\hat{i}(\gamma _1, 1)}, \frac{\alpha _{1, 2}}{\hat{i}(\gamma _1, 1)}, \dots \frac{\alpha _{1, \mu _1}}{\hat{i}(\gamma _1, 1)}, \frac{\alpha _{2, 1}}{\hat{i}(\gamma _2, 1)},\dots , \frac{\alpha _{q, \mu _q}}{\hat{i}(\gamma _q, 1)}\right) \nonumber \\\in & {} \mathbf{R}^h. \end{aligned}$$
(4.37)

Then the above theorem is equivalent to finding a vertex

$$\begin{aligned} \chi =(\chi _1,\ldots ,\chi _q,\chi _{1, 1},\chi _{1, 2},\ldots ,\chi _{1, \mu _1}, \chi _{2, 1},\ldots ,\chi _{q, \mu _q}) \end{aligned}$$

of the unit cube \([0,\,1]^h\) and infinitely many integers \(T\in \mathbf{N}\) such that

$$\begin{aligned} |\{Tv\}-\chi |<\epsilon \end{aligned}$$
(4.38)

for any given \(\epsilon \) small enough (cf. pp. 346 and 349 of [17]).

The next theorem describe how to choose a vertex satisfying (4.38).

Theorem 4.10

(cf. Theorem theo4.2 of [17]) Let H be the closure of the subset \(\{\{mv\}|m\in \mathbf{N}\}\) in \(\mathbf {T}^h=(\mathbf{R}/\mathbf{Z})^h\) and \(V=T_0\pi ^{-1}H\) be the tangent space of \(\pi ^{-1}H\) at the origin in \(\mathbf{R}^h\), where \(\pi :\mathbf{R}^h\rightarrow \mathbf {T}^h\) is the projection map. Define

$$\begin{aligned} A(v)=V{\setminus }\cup _{v_k\in \mathbf{R}{\setminus }\mathbf{Q}}\{x=(x_1,,\ldots ,x_h)\in V | x_k=0\}. \end{aligned}$$
(4.39)

Define \(\psi (x)=0\) when \(x\ge 0\) and \(\psi (x)=1\) when \(x<0\). Then for any \(a=(a_1,\ldots , a_h)\in A(V)\), the vector

$$\begin{aligned} \chi =(\psi (a_1),\ldots ,\psi (a_h)) \end{aligned}$$
(4.40)

makes (4.38) hold for infinitely many \(T\in \mathbf{N}\).

Moreover, we have the following property for the set A(v):

Theorem 4.11

(cf. Theorem 4.2 of [17])

  1. (i)

    If \(v\in \mathbf{R}^h{\setminus }\mathbf{Q}^h\), then \(\dim V\ge 1\), \(0\notin A(v)\subset V\), \(A(v)=-A(v)\) and A(v) is open in V.

  2. (ii)

    If \(\dim V = 1\), then \(A(v) = V {\setminus }\{0\}\).

  3. (iii)

    If \(\dim V \ge 2\), then A(v) is obtained from V by deleting all the coordinate hyperplanes with dimension strictly smaller than \(\dim V\).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Wang, W. Existence of closed characteristics on compact convex hypersurfaces in \(\mathbf{R}^{2n}\) . Calc. Var. 55, 2 (2016). https://doi.org/10.1007/s00526-015-0945-8

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s00526-015-0945-8

Mathematics Subject Classification

Navigation