Moser's trick

From HandWiki

In differential geometry, a branch of mathematics, the Moser's trick (or Moser's argument) is a method to relate two differential forms [math]\displaystyle{ \alpha_0 }[/math] and [math]\displaystyle{ \alpha_1 }[/math] on a smooth manifold by a diffeomorphism [math]\displaystyle{ \psi \in \mathrm{Diff}(M) }[/math] such that [math]\displaystyle{ \psi^* \alpha_1 = \alpha_0 }[/math], provided that one can find a family of vector fields satisfying a certain ODE. More generally, the argument holds for a family [math]\displaystyle{ \{ \alpha_t \}_{t \in [0,1]} }[/math] and produce an entire isotopy [math]\displaystyle{ \psi_t }[/math] such that [math]\displaystyle{ \psi_t^* \alpha_t = \alpha_0 }[/math].

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 [math]\displaystyle{ \{ \omega_t \}_{t \in [0,1]} \subset \Omega^k (M) }[/math] be a family of differential forms on [math]\displaystyle{ M }[/math]. If the ODE [math]\displaystyle{ \frac{d}{dt} \omega_t + \mathcal{L}_{X_t} \omega_t = 0 }[/math] admits a solution [math]\displaystyle{ \{ X_t \}_{t \in [0,1]} \subset \mathfrak{X}(M) }[/math], then there exists a family [math]\displaystyle{ \{ \psi_t \}_{t \in [0,1]} }[/math] of diffeomorphisms of [math]\displaystyle{ M }[/math] such that [math]\displaystyle{ \psi_t^*\omega_t = \omega_0 }[/math] and [math]\displaystyle{ \psi_0 = \mathrm{id}_M }[/math]. In particular, there is a diffeomorphism [math]\displaystyle{ \psi := \psi_1 }[/math] such that [math]\displaystyle{ \psi^*\omega_1 = \omega_0 }[/math].

Proof

The trick consists in viewing [math]\displaystyle{ \{ \psi_t \}_{t \in [0,1]} }[/math] as the flows of a time-dependent vector field, i.e. of a smooth family [math]\displaystyle{ \{ X_t \}_{t \in [0,1]} }[/math] of vector fields on [math]\displaystyle{ M }[/math]. Using the definition of flow, i.e. [math]\displaystyle{ \frac{d}{dt} \psi_t = X_t \circ \psi_t }[/math] for every [math]\displaystyle{ t \in [0,1] }[/math], one obtains from the chain rule that [math]\displaystyle{ \frac{d}{dt} (\psi_t^* \omega_t) = \psi_t^* \Big( \frac{d}{dt} \omega_t + \mathcal{L}_{X_t}\omega_t \Big). }[/math] By hypothesis, one can always find [math]\displaystyle{ X_t }[/math] such that [math]\displaystyle{ \frac{d}{dt} \omega_t + \mathcal{L}_{X_t} \omega_t = 0 }[/math], hence their flows [math]\displaystyle{ \psi_t }[/math] satisfies [math]\displaystyle{ \psi_t^* \omega_t = \mathrm{const} = \psi_0^* \omega_0 = \omega_0 }[/math].

Application to volume forms

Let [math]\displaystyle{ \alpha_0, \alpha_1 }[/math] be two volume forms on a compact [math]\displaystyle{ n }[/math]-dimensional manifold [math]\displaystyle{ M }[/math]. Then there exists a diffeomorphism [math]\displaystyle{ \psi }[/math] of [math]\displaystyle{ M }[/math] such that [math]\displaystyle{ \psi^*\alpha_1 = \alpha_0 }[/math] if and only if [math]\displaystyle{ \int_M \alpha_0 = \int_M \alpha_1 }[/math].[1]

Proof

One implication holds by the invariance of the integral by diffeomorphisms: [math]\displaystyle{ \int_M \alpha_0 = \int_M \psi^*\alpha_1 = \int_{\psi(M)} \alpha_1 = \int_M \alpha_1 }[/math].

For the converse, we apply Moser's trick to the family of volume forms [math]\displaystyle{ \alpha_t := (1-t) \alpha_0 + t \alpha_1 }[/math]. Since [math]\displaystyle{ \int_M (\alpha_1 - \alpha_0) = 0 }[/math], the de Rham cohomology class [math]\displaystyle{ [\alpha_0 - \alpha_1] \in H^n_{dR}(M) }[/math] vanishes, as a consequence of Poincaré duality and the de Rham theorem. Then [math]\displaystyle{ \alpha_1 - \alpha_0 = d\beta }[/math] for some [math]\displaystyle{ \beta \in \Omega^{n-1} (M) }[/math], hence [math]\displaystyle{ \alpha_t = \alpha_0 + t d\beta }[/math]. By Moser's trick, it is enough to solve the following ODE, where we used the Cartan's magic formula, and the fact that [math]\displaystyle{ \alpha_t }[/math] is a top-degree form:[math]\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 + i_{X_t} \alpha_t). }[/math]However, since [math]\displaystyle{ \alpha_t }[/math] is a volume form, i.e. [math]\displaystyle{ TM \xrightarrow{\cong} \wedge^{n-1} T^*M, \quad X_t \mapsto \iota_{X_t} \alpha_t }[/math], given [math]\displaystyle{ \beta }[/math] one can always find [math]\displaystyle{ X_t }[/math] such that [math]\displaystyle{ \beta + \iota_{X_t} \alpha_t = 0 }[/math].

Application to symplectic structures

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

Let [math]\displaystyle{ \{ \omega_t \}_{t \in [0,1]} \subset \Omega^2 (M) }[/math] be a family of symplectic forms on [math]\displaystyle{ M }[/math] such that [math]\displaystyle{ \frac{d}{dt} \omega_t = d \sigma_t }[/math], for [math]\displaystyle{ \{ \sigma_t \}_{t \in [0,1]} \subset \Omega^1 (M) }[/math]. Then there exists a family [math]\displaystyle{ \{ \psi_t \}_{t \in [0,1]} }[/math] of diffeomorphisms of [math]\displaystyle{ M }[/math] such that [math]\displaystyle{ \psi_t^*\omega_t = \omega_0 }[/math] and [math]\displaystyle{ \psi_0 = \mathrm{id}_M }[/math].

Proof

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

[math]\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), }[/math]where we used the hypothesis, the Cartan's magic formula, and the fact that [math]\displaystyle{ \omega_t }[/math] is closed. However, since [math]\displaystyle{ \omega_t }[/math] is non-degenerate, i.e. [math]\displaystyle{ TM \xrightarrow{\cong} T^*M, \quad X_t \mapsto \iota_{X_t} \omega_t }[/math], given [math]\displaystyle{ \sigma_t }[/math] one can always find [math]\displaystyle{ X_t }[/math] such that [math]\displaystyle{ \sigma_t + \iota_{X_t} \omega_t = 0 }[/math].

Corollary

Given two symplectic structures [math]\displaystyle{ \omega_0 }[/math] and [math]\displaystyle{ \omega_1 }[/math] on [math]\displaystyle{ M }[/math] such that [math]\displaystyle{ (\omega_0)_x = (\omega_1)_x }[/math] for some point [math]\displaystyle{ x \in M }[/math], there are two neighbourhoods [math]\displaystyle{ U_0 }[/math] and [math]\displaystyle{ U_1 }[/math] of [math]\displaystyle{ x }[/math] and a diffeomorphism [math]\displaystyle{ \phi: U_0 \to U_1 }[/math] such that [math]\displaystyle{ \phi(x) = x }[/math] and [math]\displaystyle{ \phi^*\omega_1 = \omega_0 }[/math].[3][4]

This follows by noticing that, by Poincaré lemma, the difference [math]\displaystyle{ \omega_1 - \omega_0 }[/math] is locally [math]\displaystyle{ d\sigma }[/math] for some [math]\displaystyle{ \sigma \in \Omega^1 (M) }[/math]; then, shrinking further the neighbourhoods, the result above applied to the family [math]\displaystyle{ \omega_t := (1-t) \omega_0 + t \omega_1 }[/math] of symplectic structures yields the diffeomorphism [math]\displaystyle{ \phi := \psi_1 }[/math].

Darboux theorem for symplectic structures

The Darboux's theorem for symplectic structures states that any point [math]\displaystyle{ x }[/math] in a given symplectic manifold [math]\displaystyle{ (M,\omega) }[/math] admits a local coordinate chart [math]\displaystyle{ (U, x^1,\ldots,x^n,y^1,\ldots,y^n) }[/math] such that[math]\displaystyle{ \omega|_U = \sum_{i=1}^n dx^i \wedge dy^i. }[/math]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 [math]\displaystyle{ (T_x M,\omega_x) }[/math], one can always find local coordinates [math]\displaystyle{ (\tilde{U}, \tilde{x}^1,\ldots,\tilde{x}^n,\tilde{y}^1,\ldots,\tilde{y}^n) }[/math] such that [math]\displaystyle{ \omega_x = \sum_{i=i}^n (d\tilde{x}^i \wedge d\tilde{y}^i) |_x }[/math]. Then it is enough to apply the corollary of Moser's trick discussed above to [math]\displaystyle{ \omega_0 = \omega |_{\tilde{U}} }[/math] and [math]\displaystyle{ \omega_1 = \sum_{i=i}^n d\tilde{x}^i \wedge d\tilde{y}^i }[/math], and consider the new coordinates [math]\displaystyle{ x^i = \tilde{x}^i \circ \phi, y^i = \tilde{y}^i \circ \phi }[/math].[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 [math]\displaystyle{ \{ \omega_t \}_{t \in [0,1]} \subset \Omega^2 (M) }[/math] a family of symplectic form on [math]\displaystyle{ M }[/math] which are cohomologous, i.e. the deRham cohomology class [math]\displaystyle{ [\omega_t] \in H^2_{dR}(M) }[/math] does not depend on [math]\displaystyle{ t }[/math]. Then there exists a family [math]\displaystyle{ \psi_t }[/math] of diffeomorphisms of [math]\displaystyle{ M }[/math] such that [math]\displaystyle{ \psi^*\omega_t = \omega_0 }[/math] and [math]\displaystyle{ \psi_0 = \mathrm{id}_M }[/math].

Proof

It is enough to check that [math]\displaystyle{ \frac{d}{dt} \omega_t = d \sigma_t }[/math]; then the proof follows from the previous application of Moser's trick to symplectic structures. By the cohomologous hypothesis, [math]\displaystyle{ \omega_t - \omega_0 }[/math] is an exact form, so that also its derivative [math]\displaystyle{ \frac{d}{dt} (\omega_t - \omega_0) = \frac{d}{dt} \omega_t }[/math] is exact for every [math]\displaystyle{ t }[/math]. The actual proof that this can be done in a smooth way, i.e. that [math]\displaystyle{ \frac{d}{dt} \omega_t = d \sigma_t }[/math] for a smooth family of functions [math]\displaystyle{ \sigma_t }[/math], 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