In differential geometry, a branch of mathematics, the Moser's trick (or Moser's argument) is a method to relate two differential forms and on a smooth manifold by a diffeomorphism such that , provided that one can find a family of vector fields satisfying a certain ODE.

More generally, the argument holds for a family and produce an entire isotopy such that .

It was originally given by Jürgen Moser in 1965 to check when two volume forms are equivalent,[1] but its main applications are in symplectic geometry. It is the standard argument for the modern proof of Darboux's theorem, as well as for the proof of Darboux-Weinstein theorem[2] and other normal form results.[2][3][4]

General statement

edit

Let   be a family of differential forms on a compact manifold  . If the ODE   admits a solution  , then there exists a family   of diffeomorphisms of   such that   and  . In particular, there is a diffeomorphism   such that  .

Proof

edit

The trick consists in viewing   as the flows of a time-dependent vector field, i.e. of a smooth family   of vector fields on  . Using the definition of flow, i.e.   for every  , one obtains from the chain rule that   By hypothesis, one can always find   such that  , hence their flows   satisfies  . In particular, as   is compact, this flows exists at  .

Application to volume forms

edit

Let   be two volume forms on a compact  -dimensional manifold  . Then there exists a diffeomorphism   of   such that   if and only if  .[1]

Proof

edit

One implication holds by the invariance of the integral by diffeomorphisms:  .


For the converse, we apply Moser's trick to the family of volume forms  . Since  , the de Rham cohomology class   vanishes, as a consequence of Poincaré duality and the de Rham theorem. Then   for some  , hence  . By Moser's trick, it is enough to solve the following ODE, where we used the Cartan's magic formula, and the fact that   is a top-degree form: However, since   is a volume form, i.e.  , given   one can always find   such that  .

Application to symplectic structures

edit

In the context of symplectic geometry, the Moser's trick is often presented in the following form.[3][4]

Let   be a family of symplectic forms on   such that  , for  . Then there exists a family   of diffeomorphisms of   such that   and  .

Proof

edit

In order to apply Moser's trick, we need to solve the following ODE

 where we used the hypothesis, the Cartan's magic formula, and the fact that   is closed. However, since   is non-degenerate, i.e.  , given   one can always find   such that  .

Corollary

edit

Given two symplectic structures   and   on   such that   for some point  , there are two neighbourhoods   and   of   and a diffeomorphism   such that   and  .[3][4]

This follows by noticing that, by Poincaré lemma, the difference   is locally   for some  ; then, shrinking further the neighbourhoods, the result above applied to the family   of symplectic structures yields the diffeomorphism  .

Darboux theorem for symplectic structures

edit

The Darboux's theorem for symplectic structures states that any point   in a given symplectic manifold   admits a local coordinate chart   such that While the original proof by Darboux required a more general statement for 1-forms,[5] Moser's trick provides a straightforward proof. Indeed, choosing any symplectic basis of the symplectic vector space  , one can always find local coordinates   such that  . Then it is enough to apply the corollary of Moser's trick discussed above to   and  , and consider the new coordinates  .[3][4]

Application: Moser stability theorem

edit

Moser himself provided an application of his argument for the stability of symplectic structures,[1] which is known now as Moser stability theorem.[3][4]

Let   a family of symplectic form on   which are cohomologous, i.e. the deRham cohomology class   does not depend on  . Then there exists a family   of diffeomorphisms of   such that   and  .

Proof

edit

It is enough to check that  ; then the proof follows from the previous application of Moser's trick to symplectic structures. By the cohomologous hypothesis,   is an exact form, so that also its derivative   is exact for every  . The actual proof that this can be done in a smooth way, i.e. that   for a smooth family of functions  , requires some algebraic topology. One option is to prove it by induction, using Mayer-Vietoris sequences;[3] another is to choose a Riemannian metric and employ Hodge theory.[1]

References

edit
  1. ^ a b c d Moser, Jürgen (1965). "On the volume elements on a manifold". Transactions of the American Mathematical Society. 120 (2): 286–294. doi:10.1090/S0002-9947-1965-0182927-5. ISSN 0002-9947.
  2. ^ a b Weinstein, Alan (1971-06-01). "Symplectic manifolds and their lagrangian submanifolds". Advances in Mathematics. 6 (3): 329–346. doi:10.1016/0001-8708(71)90020-X. ISSN 0001-8708.
  3. ^ a b c d e f McDuff, Dusa; Salamon, Dietmar (2017-06-22). Introduction to Symplectic Topology. Vol. 1. Oxford University Press. doi:10.1093/oso/9780198794899.001.0001. ISBN 978-0-19-879489-9.
  4. ^ a b c d e Cannas Silva, Ana (2008). Lectures on Symplectic Geometry. Springer. doi:10.1007/978-3-540-45330-7. ISBN 978-3-540-42195-5.
  5. ^ Sternberg, Shlomo (1964). Lectures on Differential Geometry. Prentice Hall. pp. 140–141. ISBN 9780828403160.