Moser's trick

In differential geometry, a branch of mathematics, the Moser's trick (or Moser's argument) is a method to relate two differential forms α 0 {\displaystyle \alpha _{0}} and α 1 {\displaystyle \alpha _{1}} on a smooth manifold by a diffeomorphism ψ D i f f ( M ) {\displaystyle \psi \in \mathrm {Diff} (M)} such that ψ α 1 = α 0 {\displaystyle \psi ^{*}\alpha _{1}=\alpha _{0}} , provided that one can find a family of vector fields satisfying a certain ODE.

More generally, the argument holds for a family { α t } t [ 0 , 1 ] {\displaystyle \{\alpha _{t}\}_{t\in [0,1]}} and produce an entire isotopy ψ t {\displaystyle \psi _{t}} such that ψ t α t = α 0 {\displaystyle \psi _{t}^{*}\alpha _{t}=\alpha _{0}} .

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

Let { ω t } t [ 0 , 1 ] Ω k ( M ) {\displaystyle \{\omega _{t}\}_{t\in [0,1]}\subset \Omega ^{k}(M)} be a family of differential forms on a compact manifold M {\displaystyle M} . If the ODE d d t ω t + L X t ω t = 0 {\displaystyle {\frac {d}{dt}}\omega _{t}+{\mathcal {L}}_{X_{t}}\omega _{t}=0} admits a solution { X t } t [ 0 , 1 ] X ( M ) {\displaystyle \{X_{t}\}_{t\in [0,1]}\subset {\mathfrak {X}}(M)} , then there exists a family { ψ t } t [ 0 , 1 ] {\displaystyle \{\psi _{t}\}_{t\in [0,1]}} of diffeomorphisms of M {\displaystyle M} such that ψ t ω t = ω 0 {\displaystyle \psi _{t}^{*}\omega _{t}=\omega _{0}} and ψ 0 = i d M {\displaystyle \psi _{0}=\mathrm {id} _{M}} . In particular, there is a diffeomorphism ψ := ψ 1 {\displaystyle \psi :=\psi _{1}} such that ψ ω 1 = ω 0 {\displaystyle \psi ^{*}\omega _{1}=\omega _{0}} .

Proof

The trick consists in viewing { ψ t } t [ 0 , 1 ] {\displaystyle \{\psi _{t}\}_{t\in [0,1]}} as the flows of a time-dependent vector field, i.e. of a smooth family { X t } t [ 0 , 1 ] {\displaystyle \{X_{t}\}_{t\in [0,1]}} of vector fields on M {\displaystyle M} . Using the definition of flow, i.e. d d t ψ t = X t ψ t {\displaystyle {\frac {d}{dt}}\psi _{t}=X_{t}\circ \psi _{t}} for every t [ 0 , 1 ] {\displaystyle t\in [0,1]} , one obtains from the chain rule that d d t ( ψ t ω t ) = ψ t ( d d t ω t + L X t ω t ) . {\displaystyle {\frac {d}{dt}}(\psi _{t}^{*}\omega _{t})=\psi _{t}^{*}{\Big (}{\frac {d}{dt}}\omega _{t}+{\mathcal {L}}_{X_{t}}\omega _{t}{\Big )}.} By hypothesis, one can always find X t {\displaystyle X_{t}} such that d d t ω t + L X t ω t = 0 {\displaystyle {\frac {d}{dt}}\omega _{t}+{\mathcal {L}}_{X_{t}}\omega _{t}=0} , hence their flows ψ t {\displaystyle \psi _{t}} satisfies ψ t ω t = c o n s t = ψ 0 ω 0 = ω 0 {\displaystyle \psi _{t}^{*}\omega _{t}=\mathrm {const} =\psi _{0}^{*}\omega _{0}=\omega _{0}} . In particular, as M {\displaystyle M} is compact, this flows exists at t = 1 {\displaystyle t=1} .

Application to volume forms

Let α 0 , α 1 {\displaystyle \alpha _{0},\alpha _{1}} be two volume forms on a compact n {\displaystyle n} -dimensional manifold M {\displaystyle M} . Then there exists a diffeomorphism ψ {\displaystyle \psi } of M {\displaystyle M} such that ψ α 1 = α 0 {\displaystyle \psi ^{*}\alpha _{1}=\alpha _{0}} if and only if M α 0 = M α 1 {\displaystyle \int _{M}\alpha _{0}=\int _{M}\alpha _{1}} .[1]

Proof

One implication holds by the invariance of the integral by diffeomorphisms: M α 0 = M ψ α 1 = ψ ( M ) α 1 = M α 1 {\displaystyle \int _{M}\alpha _{0}=\int _{M}\psi ^{*}\alpha _{1}=\int _{\psi (M)}\alpha _{1}=\int _{M}\alpha _{1}} .


For the converse, we apply Moser's trick to the family of volume forms α t := ( 1 t ) α 0 + t α 1 {\displaystyle \alpha _{t}:=(1-t)\alpha _{0}+t\alpha _{1}} . Since M ( α 1 α 0 ) = 0 {\displaystyle \int _{M}(\alpha _{1}-\alpha _{0})=0} , the de Rham cohomology class [ α 0 α 1 ] H d R n ( M ) {\displaystyle [\alpha _{0}-\alpha _{1}]\in H_{dR}^{n}(M)} vanishes, as a consequence of Poincaré duality and the de Rham theorem. Then α 1 α 0 = d β {\displaystyle \alpha _{1}-\alpha _{0}=d\beta } for some β Ω n 1 ( M ) {\displaystyle \beta \in \Omega ^{n-1}(M)} , hence α t = α 0 + t d β {\displaystyle \alpha _{t}=\alpha _{0}+td\beta } . By Moser's trick, it is enough to solve the following ODE, where we used the Cartan's magic formula, and the fact that α t {\displaystyle \alpha _{t}} is a top-degree form:

0 = d d t α t + L X t α t = d β + d ( ι X t α t ) + ι X t ( d α t ) = d ( β + ι X t α t ) . {\displaystyle 0={\frac {d}{dt}}\alpha _{t}+{\mathcal {L}}_{X_{t}}\alpha _{t}=d\beta +d(\iota _{X_{t}}\alpha _{t})+\iota _{X_{t}}({\cancel {d\alpha _{t}}})=d(\beta +\iota _{X_{t}}\alpha _{t}).}
However, since α t {\displaystyle \alpha _{t}} is a volume form, i.e. T M n 1 T M , X t ι X t α t {\textstyle TM\xrightarrow {\cong } \wedge ^{n-1}T^{*}M,\quad X_{t}\mapsto \iota _{X_{t}}\alpha _{t}} , given β {\displaystyle \beta } one can always find X t {\displaystyle X_{t}} such that β + ι X t α t = 0 {\displaystyle \beta +\iota _{X_{t}}\alpha _{t}=0} .

Application to symplectic structures

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

Let { ω t } t [ 0 , 1 ] Ω 2 ( M ) {\displaystyle \{\omega _{t}\}_{t\in [0,1]}\subset \Omega ^{2}(M)} be a family of symplectic forms on M {\displaystyle M} such that d d t ω t = d σ t {\displaystyle {\frac {d}{dt}}\omega _{t}=d\sigma _{t}} , for { σ t } t [ 0 , 1 ] Ω 1 ( M ) {\displaystyle \{\sigma _{t}\}_{t\in [0,1]}\subset \Omega ^{1}(M)} . Then there exists a family { ψ t } t [ 0 , 1 ] {\displaystyle \{\psi _{t}\}_{t\in [0,1]}} of diffeomorphisms of M {\displaystyle M} such that ψ t ω t = ω 0 {\displaystyle \psi _{t}^{*}\omega _{t}=\omega _{0}} and ψ 0 = i d M {\displaystyle \psi _{0}=\mathrm {id} _{M}} .

Proof

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

0 = d d t ω t + L X t ω t = d σ t + ι X t ( d ω t ) + d ( ι X t ω t ) = d ( σ t + ι X t ω t ) , {\displaystyle 0={\frac {d}{dt}}\omega _{t}+{\mathcal {L}}_{X_{t}}\omega _{t}=d\sigma _{t}+\iota _{X_{t}}({\cancel {d\omega _{t}}})+d(\iota _{X_{t}}\omega _{t})=d(\sigma _{t}+\iota _{X_{t}}\omega _{t}),}
where we used the hypothesis, the Cartan's magic formula, and the fact that ω t {\displaystyle \omega _{t}} is closed. However, since ω t {\displaystyle \omega _{t}} is non-degenerate, i.e. T M T M , X t ι X t ω t {\textstyle TM\xrightarrow {\cong } T^{*}M,\quad X_{t}\mapsto \iota _{X_{t}}\omega _{t}} , given σ t {\displaystyle \sigma _{t}} one can always find X t {\displaystyle X_{t}} such that σ t + ι X t ω t = 0 {\displaystyle \sigma _{t}+\iota _{X_{t}}\omega _{t}=0} .

Corollary

Given two symplectic structures ω 0 {\displaystyle \omega _{0}} and ω 1 {\displaystyle \omega _{1}} on M {\displaystyle M} such that ( ω 0 ) x = ( ω 1 ) x {\displaystyle (\omega _{0})_{x}=(\omega _{1})_{x}} for some point x M {\displaystyle x\in M} , there are two neighbourhoods U 0 {\displaystyle U_{0}} and U 1 {\displaystyle U_{1}} of x {\displaystyle x} and a diffeomorphism ϕ : U 0 U 1 {\displaystyle \phi :U_{0}\to U_{1}} such that ϕ ( x ) = x {\displaystyle \phi (x)=x} and ϕ ω 1 = ω 0 {\displaystyle \phi ^{*}\omega _{1}=\omega _{0}} .[3][4]

This follows by noticing that, by Poincaré lemma, the difference ω 1 ω 0 {\displaystyle \omega _{1}-\omega _{0}} is locally d σ {\displaystyle d\sigma } for some σ Ω 1 ( M ) {\displaystyle \sigma \in \Omega ^{1}(M)} ; then, shrinking further the neighbourhoods, the result above applied to the family ω t := ( 1 t ) ω 0 + t ω 1 {\displaystyle \omega _{t}:=(1-t)\omega _{0}+t\omega _{1}} of symplectic structures yields the diffeomorphism ϕ := ψ 1 {\displaystyle \phi :=\psi _{1}} .

Darboux theorem for symplectic structures

The Darboux's theorem for symplectic structures states that any point x {\displaystyle x} in a given symplectic manifold ( M , ω ) {\displaystyle (M,\omega )} admits a local coordinate chart ( U , x 1 , , x n , y 1 , , y n ) {\displaystyle (U,x^{1},\ldots ,x^{n},y^{1},\ldots ,y^{n})} such that

ω | U = i = 1 n d x i d y i . {\displaystyle \omega |_{U}=\sum _{i=1}^{n}dx^{i}\wedge dy^{i}.}
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 ( T x M , ω x ) {\displaystyle (T_{x}M,\omega _{x})} , one can always find local coordinates ( U ~ , x ~ 1 , , x ~ n , y ~ 1 , , y ~ n ) {\displaystyle ({\tilde {U}},{\tilde {x}}^{1},\ldots ,{\tilde {x}}^{n},{\tilde {y}}^{1},\ldots ,{\tilde {y}}^{n})} such that ω x = i = i n ( d x ~ i d y ~ i ) | x {\displaystyle \omega _{x}=\sum _{i=i}^{n}(d{\tilde {x}}^{i}\wedge d{\tilde {y}}^{i})|_{x}} . Then it is enough to apply the corollary of Moser's trick discussed above to ω 0 = ω | U ~ {\displaystyle \omega _{0}=\omega |_{\tilde {U}}} and ω 1 = i = i n d x ~ i d y ~ i {\displaystyle \omega _{1}=\sum _{i=i}^{n}d{\tilde {x}}^{i}\wedge d{\tilde {y}}^{i}} , and consider the new coordinates x i = x ~ i ϕ , y i = y ~ i ϕ {\displaystyle x^{i}={\tilde {x}}^{i}\circ \phi ,y^{i}={\tilde {y}}^{i}\circ \phi } .[3][4]

Application: Moser stability theorem

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 { ω t } t [ 0 , 1 ] Ω 2 ( M ) {\displaystyle \{\omega _{t}\}_{t\in [0,1]}\subset \Omega ^{2}(M)} a family of symplectic form on M {\displaystyle M} which are cohomologous, i.e. the deRham cohomology class [ ω t ] H d R 2 ( M ) {\displaystyle [\omega _{t}]\in H_{dR}^{2}(M)} does not depend on t {\displaystyle t} . Then there exists a family ψ t {\displaystyle \psi _{t}} of diffeomorphisms of M {\displaystyle M} such that ψ ω t = ω 0 {\displaystyle \psi ^{*}\omega _{t}=\omega _{0}} and ψ 0 = i d M {\displaystyle \psi _{0}=\mathrm {id} _{M}} .

Proof

It is enough to check that d d t ω t = d σ t {\textstyle {\frac {d}{dt}}\omega _{t}=d\sigma _{t}} ; then the proof follows from the previous application of Moser's trick to symplectic structures. By the cohomologous hypothesis, ω t ω 0 {\displaystyle \omega _{t}-\omega _{0}} is an exact form, so that also its derivative d d t ( ω t ω 0 ) = d d t ω t {\textstyle {\frac {d}{dt}}(\omega _{t}-\omega _{0})={\frac {d}{dt}}\omega _{t}} is exact for every t {\displaystyle t} . The actual proof that this can be done in a smooth way, i.e. that d d t ω t = d σ t {\textstyle {\frac {d}{dt}}\omega _{t}=d\sigma _{t}} for a smooth family of functions σ t {\displaystyle \sigma _{t}} , 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

  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.