Classification of coadjoint orbits for symplectomorphism groups of surfaces
Abstract
We classify generic coadjoint orbits for symplectomorphism groups of compact symplectic surfaces with or without boundary. We also classify simple Morse functions on such surfaces up to a symplectomorphism.
1 Introduction
The classification problem for coadjoint orbits for the action of symplectic (or area-preserving) diffeomorphisms in two dimensions was known to specialists in view of its application in fluid dynamics since the 1960s, and it was explicitly formulated in [2, see Section I.5] in 1998. The same classification problem also arises in Poisson geometry since coadjoint orbits are symplectic leaves of the Lie-Poisson bracket, and also in representation theory in connection with the orbit method of A.Kirillov [12]. In the recent work [17] the orbit method was applied to the symplectomorphism group of the two-sphere. The classification of generic coadjoint orbits was obtained in [8, 9] for the case of closed surfaces. In [8] there is a list of open questions for the case of surfaces with boundary. We answer all those questions in Sections 3 and 4.
The classification problem for coadjoint orbits for symplectomorphisms of a surface is closely related to a certain classification problem for functions, specifically, with each coadjoint orbit one can associate a (vorticity) function up to a symplectomorphism (see details in Section 4.1). Hence we are going to address the following two problems:
-
1.
Classify generic coadjoint orbits of symplectomorphism groups of surfaces.
-
2.
Classify generic smooth functions on symplectic surfaces up to symplectomorphisms.
There is a number of papers devoted to the classification of functions with non-degenerate critical points on surfaces. In [6] simple Morse functions on surfaces with boundary were classified with respect to smooth left-right equivalence. In [9] simple Morse functions on closed compact symplectic surfaces were classified up to a symplectomorphism. We generalize the results of [9] to the case of surfaces with boundary. The classification of functions is given in Theorem 3.13. Roughy speaking, the classification theorem for functions states that there a one-to-one correspondence between functions up to a symplectomorphism and measured Reeb graphs up to an isomorphism. The classification of coadjoint orbits is given in Theorem 4.10. The classification result for coadjoint orbits in also given in terms of measured Reeb graphs supplemented with some additional data (so called augmented circulation Reeb graphs).
It is worth noting that the classification of functions in [9] is based on the classification of so-called simple Morse fibrations obtained in [5]; on the contrary, the proof in the present paper uses a different method so it gives an alternative proof for Theorem 3.11 from [9] (classification of functions in the case of closed surfaces).
This paper is organised as follows: in Section 2 we give a local classification of Morse functions up to symplectomorphisms. In Section 3 we solve the global classification problem of simple Morse functions. In Section 4 we use the results of Section 3 in order to classify generic coadjoint orbits of the symplectomorphism group and illustrate these results with examples. In Section 5 we discuss some open problems in this area, and in Appendix A we present a hydrodynamical motivation for the main classification theorem.
Acknowledgements
The author is grateful to A.Izosimov, B.Khesin, E.Kudryavtseva, V.Matveev, A.Oshemkov, A. Prishlyak, and the anonymous referee for discussions and valuable comments, which significantly improved this paper. This research is supported in part by the Russian Science Foundation (grant No. 17-11-01303) and the Simons Foundation.
2 Local classification of functions up to a symplectomorphism
2.1 Preliminaries
Throughout this section, let be a compact connected symplectic surface with the boundary
Definition 2.1 ([10]).
A smooth function is called a simple Morse function if the following conditions hold:
-
(i)
All critical points of are non-degenerate;
-
(ii)
does not have critical points on the boundary ;
-
(iii)
The restriction of to the boundary is a Morse function;
-
(iv)
All critical values of and of its restriction are distinct.
Proposition 2.2 ([10]).
Simple Morse functions form an open and dense subset in the space of smooth functions in the -topology.
For a regular point of the function there exists a coordinate chart centered at such that and For a regular point of the restriction there exists a coordinate chart in centered at such that and the boundary is given by the equation Next, we present the normal forms for the pair in a neighbourhood of a critical point for the function and of its restriction
2.2 The case of a singular point inside the surface
Theorem 2.3 ([4, 19]).
Let be a symplectic surface, and be a simple Morse function. Let be a critical point for the function Then there exists a coordinate chart centered at such that and where or Here is a smooth function of one variable defined in some neighborhood of the origin and Moreover:
-
(i)
In the case and the function is uniquely determined by the pair
-
(ii)
In the case only the Taylor series of the function is uniquely determined by the pair In other words, if is another chart as above then (for sufficiently small where is a function of one variable flat111Here flat means that all derivatives of vanish at the origin. at the origin. Furthermore, every function of one variable that is flat at the origin can be obtained in this way.
2.3 The case of a singular point on the boundary
Theorem 2.4.
Let be a symplectic surface, and be a simple Morse function. Let be a regular point of and a non-degenerate critical point of its restriction Then there exists a coordinate chart centered at such that and where or Here is a smooth function of one variable defined in some neighborhood of the origin and In this chart is defined by and the boundary is given by the equation see Figure 1. Moreover:
-
(i)
In the case and the function is uniquely determined by the pair
-
(ii)
In the case only the Taylor series of the function is uniquely determined by the pair In other words, if is another chart as above then Furthermore, every function of one variable that is flat at the origin can be obtained in this way.
Before we proceed with the proof of this theorem let us formulate and prove a lemma.
Lemma 2.5.
Let and be two smooth non-negative functions such that and for Then the following statements are equivalent:
-
(i)
The difference is a function flat at the origin, i.e. the Taylor series and are equal to each other.
-
(ii)
The difference is a smooth function
-
(iii)
The difference is a smooth function flat at the origin.
Proof.
The implication is evident so it enough to show that and Let us start with implication It follows from Hadamard’s lemma that there exist smooth functions and such that and for We have the following formula for the difference
(1) |
for small enough It follows from the formula (1) that the difference is smooth and flat at the origin whenever the difference is flat the origin.
It remains to show that implies (i). Denote by the smooth function Assume that the difference is not flat at the origin. Then there exists a number and a smooth non-zero function such that It is useful to rewrite formula (1) in the following form:
(2) |
Formula (2) implies that the function is flat at the origin whenever the function is flat at the origin. The function is a non-zero function so we conclude that the function is not flat at the origin. Therefore there exists a number and a smooth non-zero function such that Now we take the square of both sides of (2) and obtain the following formula:
(3) |
That gives us a contradiction since the Taylor series of the left hand side starts with an odd power of and the Taylor series of the right hand side starts with an even power of We conclude that the function is flat at the origin. ∎
Now we proceed with the proof of Theorem 2.4.
Proof of Theorem 2.4.
Without loss of generality we can assume that The main part of this theorem on the existence of a coordinate chart was proved in [13, 14].
Let us prove statement of this theorem. We need to prove the equality In this case (see Figure 1, (a)) the region is diffeomorphic to a closed half ball provided that is sufficiently small. Therefore, the area of this region
is well-defined. Let be a coordinate chart centered at such that and Then we can write an explicit formula for the function
So we conclude that the function (and thus the function ) is uniquely determined by the pair
Now let us prove statement of this theorem. Consider the second case where in some local chart centered at we have and Consider a smooth curve such that it is transversal to the right half of the parabola and these two curves intersect each other at exactly one point. In coordinates the curve can be described as a graph of some function
We fix a number and consider the region (see Figure 2) bounded by the boundary curve the right half of the parabola the right half of the parabola and the curve
Then the area of this region
is well-defined. Denote by the -coordinate of the intersection by the -coordinate of the intersection and by the -coordinate of the intersection The coordinates and depend on and also the coordinates and depend on It follows from the implicit function theorem that the coordinate is a smooth function of As for it is explicitly given by Note that provided that is sufficiently small. We have the following formula for the function
(4) | ||||
Now consider some other chart such that and the boundary is given by Then it is follows from above that
(5) |
where is a smooth function of one variable. We want to prove that the Taylor series is equal to the Taylor series of It follows from Lemma 2.5 that the Taylor series is equal to the Taylor series From here we conclude that
It remains to prove the last part of statement . Let be a function of one variable flat at the origin. The goal is to find a symplectomorphism defined in some neighbourhood of the origin such that is a fixed point for the symplectomorphism preserves the boundary and
For this part of the proof we are going to use a different coordinate chart (see Figure 3):
First of all, notice that after this \sayparabolic change of coordinates the symplectic form still has the standard form:
Secondly, in the chart the function \saystraightens i.e. and the boundary becomes a parabola
Now let us proceed with the proof. Apply Moser’s path method and consider the family of functions
for each Instead of looking for one symplectomorphism , we will be looking for a family of Hamiltonian symplectomorphisms such that
(6) |
for each and for each Let be the vector field corresponding to the flow
Differentiating (6) with respect to we obtain the following differential equation
which we rewrite as
Since is a diffeomorphism, it is equivalent to
(7) |
Since the flow of the field has to preserve the symplectic structure we will be looking for the field in the Hamiltonian form
(8) |
where and Substitute the right-hand side of (8) into (7) to obtain the following partial differential equation
Rewrite it as
Consider the family of functions
for each We have so the denominator is non-zero in sufficiently small neighbourhood of the origin. Then our equation assumes the form
(9) |
It is clear that the general solution to this equation has the form
where is a smooth function of one variable. Our goal is to find a particular solution to (9) that is constant along the boundary That implies the following condition on the function
Hence, for any non-positive we have
Now define a function in the following way:
It follows from the flatness of that the function defined as above is a smooth function flat at the origin. Now define the function to be the corresponding solution:
The family of symplectomorphisms can be recovered as the flow of the corresponding field The condition implies that the field has a zero restriction on the boundary and we conclude that the corresponding family preserves the boundary, and for each Now applying the theorem on the smooth dependence of the flow on initial data one can conclude that the flow is well-defined for Hence, the diffeomorphism has the desired properties. ∎
In the next section we are going to use these local results to obtain a global classification of simple Morse functions with respect to the action of the group of symplectomorphisms of
3 Global classification of functions on symplectic surfaces
3.1 The Reeb graph of a function
Throughout this section, let be a compact connected oriented surface with boundary and let be a simple Morse function on . In what follows, by a level we mean a connected component of level sets of Non-critical levels are diffeomorphic to a circle or a line segment. The surface can be considered as a union of levels, and we get a foliation with singularities. The base space of this foliation with the quotient topology is homeomorphic to a finite connected graph (see Figure 4) whose vertices correspond to critical values of or We view this graph as a topological object (rather than combinatorial). This graph is called the Reeb graph222This graph is also called the Kronrod graph of a function, see [18, 1]. of the function By we denote the projection We denote the edges of the Reeb graph by solid lines if they correspond to circle components and by dashed lines if they correspond to segment components. We denote the union of solid (respectively, dashed) edges in by and respectively. We denote the preimages and by and Thus and There are 7 possible types of vertices in the graph (see Table 1). The function on descends to a function on the Reeb graph . It is also convenient to assume that is oriented: edges are oriented in the direction of increasing
Let be a vertex of the Reeb graph Let us fix a number such that
is a proper subset of for each edge incident to Consider the preimage The boundary is a piecewise smooth closed oriented curve. This curve is connected in the case where the vertex is incident only to dashed edges, and its image is a closed oriented curve that passes edges incident to the vertex in a certain cyclic order. This construction is nontrivial only in the case when there are at least three dashed edges incident to the vertex (otherwise, there is only one cyclic order at the set of edges incident to ). Thus for an arbitrary II-vertex or IV-vertex (see Table 1) of the graph we have a natural cyclic order for the edges incident to this vertex. The above properties of the graph make it natural to introduce the following definition of an abstract Reeb graph.
Definition 3.1.
An (abstract) Reeb graph is an oriented connected graph with solid or dashed edges, and a continuous function with the following properties and additional data:
Definition 3.2.
Abstract Reeb graphs and are said to be equivalent by means of the isomorphism if the map
-
(i)
maps solid (respectively, dashed) edges to solid (respectively, dashed) edges;
-
(ii)
preserves the cyclic order on the set of edges incident to each - or -vertex, i.e. if follows in the cyclic order, then follows
-
(iii)
takes the function to the function (i.e. ).
3.2 Recovering the topology of a surface from the Reeb graph
In this subsection we follow [6, Section 5]. Let be a compact connected oriented surface with the boundary and let be a simple Morse function. The restriction of the projection to each boundary component of is a closed curve (a map from a circle to the graph) in the graph Informally speaking, the following definition describes those closed curves for an abstract Reeb graph.
Definition 3.3.
Let be an abstract Reeb graph. A non-empty sequence of edges together with a sequence of vertices is called a boundary cycle if the following three conditions hold:
-
(i)
All edges in the sequence are dashed.
-
(ii)
Each edge is incident to the vertices and for every
-
(iii)
If the vertex has three or more adjacent dashed edges, then the pair ( of consecutive edges is also a consecutive pair of edges with respect to the cyclic order on the set of edges incident to the vertex for every
We call two boundary cycles equivalent if they differ by the action of a cyclic group, i.e. the sets of vertices and define the same topological cycle for each In addition, in the case when a boundary cycle consists only of or -valent vertices (i.e. of vertices of type III and IV) we also call two boundary cycles and equivalent. We denote by the number of (equivalence classes of) boundary cycles in
Example 3.4.
Consider a disk with holes and a torus with one hole, and consider the height function on them (as shown in Figure 5). The corresponding Reeb graphs are identical except for the cyclic orders at the vertices and In case of a disk with with holes there are three boundary cycles: and In case of a torus with one hole there is only one boundary cycle:
Proposition 3.5 ([6, page 12]).
Let be a compact connected oriented surface with the boundary and let be a simple Morse function. Then the number of boundary cycles is equal to the number of boundary components of the surface
Theorem 3.6 ([6, Theorem 5.3]).
The genus of a surface is given by the following formula:
(10) | ||||
where is the Euler characteristic and is the number of boundary cycles.
Theorem 3.6 motivates us to give the following definition.
Definition 3.7.
Let be an abstract Reeb graph. Define the genus as the number from the right-hand side of the formula in Theorem 3.6.
Type | Level Sets | Reeb Graph | Asymptotics |
---|---|---|---|
I | where and . | ||
II | where and . | ||
III | where , , , , and . | ||
IV | where , , , , and . | ||
V | where , , , and . | ||
VI | where , , and . | ||
VII | where and |
3.3 Measured Reeb graphs
Now, fix an area form on the surface Then the natural projection map induces a measure on the graph
Definition 3.8.
A measure on an abstract Reeb graph is called quasi-smooth if the following conditions hold.
-
1.
The measure has a -smooth non-zero density in the complement .
-
2.
In a neighbourhood of each vertex the measure can be expressed by the corresponding formula from Table 1.
Proposition 3.9.
Let be a compact connected symplectic surface with a boundary and let be a simple Morse function. Then the measure is quasi-smooth.
Proof.
For vertices of types VI and VII this was proved in [5, Subsection I.1.2]. The proof is based on Theorem 2.3, the essence of the proof is the study of the area between the non-singular level sets of the function and a singular -level. The proof for other types follows the same lines, with the only difference that it uses both Theorems 2.3 and 2.4. Note that or vertices of types I and VII the function is uniquely (and explicitly) determined by the corresponding function (see Theorems 2.3 and 2.4). In other cases is determined by the corresponding function up to a function flat at the origin, and there is no explicit expression for in terms of (see details in Toulet’s thesis [19, Subsection 2.2]. ∎
The above properties of the measure make it natural to introduce the following definition of an abstract measured Reeb graph.
Definition 3.10.
A measured Reeb graph is a Reeb graph equipped with a quasi-smooth measure
Definition 3.11.
Two measured Reeb graphs and are said to be equivalent by means of the isomorphism if the map
-
(i)
is an isomorphism between the Reeb graphs and
-
(ii)
pushes the measure to the measure
Definition 3.12.
A measured Reeb graph is compatible with if the following conditions hold:
-
(i)
The genus of the graph is equal to the genus of the surface
-
(ii)
The number of boundary cycles is equal to the number of boundary components of the surface
-
(iii)
The volume of with respect to the measure is equal to the area of the surface :
3.4 Classification of simple Morse functions up to a symplectomorphism
Theorem 3.13.
Let be a compact connected oriented surface with boundary Then there is a one-to-one correspondence between simple Morse functions on , considered up to symplectomorphism, and (isomorphism classes of) measured Reeb graphs compatible with . In other words, the following statements hold.
-
i)
Let be two simple Morse functions. Then the following conditions are equivalent:
-
a)
There exists a symplectomorphism such that
-
b)
Measured Reeb graphs of and are isomorphic.
Moreover, every isomorphism can be lifted to a symplectomorphism such that
-
a)
-
ii)
For each measured Reeb graph compatible with there exists a simple Morse function such that the corresponding measured Reeb graph is isomorphic to
Remark 3.14.
Note that the formulation of this theorem is identical to the formulation of Theorem from [9]. The difference, of course, is that all notions in the present paper are extended to cover the case of surfaces with boundary.
Proof.
Let us prove the first statement. The implication is evident, so it suffices to prove the implication Let be an isomorphism of measured Reeb graphs. We need to construct a symplectomorphism such that and
Let be a smooth oriented curve which is transversal to the level sets of the function it does not intersect the singular levels of the function and such that the function is strictly increasing along the curve Consider the Hamiltonian flow corresponding to the function We denote by the time necessary to go from the curve to the point under the action of see Figure 6. The pair of functions forms a coordinate system in some neighborhood of such that (it is a standard computation, see proof in [13, Lemma 4]). In particular, this construction works for the boundary curve see Figure 6. The range of the function along the non-critical level of is a segment in the case when the -level is a segment, and it is a half-interval in the case when the -level is a circle. The function is called a period. It follows from Stokes’ theorem that is equal to the derivative
Let be a dashed edge. The formula defines a symplectomorphism from the interior of to the interior of The condition guarantees that the periods of the functions and coincide and hence the symplectomorphism is well-defined. Now let be a solid edge. Let be a smooth oriented curve which is transversal to the level sets of the function We also assume it does not intersect the singular levels of the function ; and the function is strictly increasing along the curve Then, as above, we obtain a symplectomorphism from the interior of to the interior of By applying the same procedure to all edges of the graph we obtain a symplectomorphism
such that and
Now let be a singular point for the function or its restriction Then there is only one way to define the image of :
Let (respectively, ) be a chart centered at the point (respectively, ) as in Theorem 2.3 or 2.4. Then the condition guarantees that the corresponding functions and are the same or they differ by a function flat at the origin. In the latter case it follows from Theorem 2.3 or 2.4 that we can replace the chart with a chart such that So without loss of generality we may assume that Therefore, one can define in some neighbourhood of by the formula
This local symplectomorphism extends uniquely to a semi-local symplectomorphism
Without loss of generality we may assume that is a ‘‘standard’’ neighbourhood of the singular level (see Table 1), i.e. it is a connected component of the set containing the point and the number is sufficiently small so that these ‘‘standard’’ neighbourhoods for distinct are pairwise disjoint. Denote by the union of all these neighbourhoods. By applying the same procedure to all singular points of the function or its restriction we obtain a symplectomorphism
such that and
So the isomorphism is lifted to a symplectomorphism
and to a symplectomorphism
However, these two symplectomorphisms not necessarily define a global symplectomorphism of the surface Let be a dashed edge. Then the intersection is a disjoint union of two rectangles and the ratio is a symplectic automorphism of this union preserving each component and also preserving the function The only symplectic automorphism of a fibered rectangle is the identity. So on i.e. the symplectomorphisms and agree with each other on the preimage of the edge Now let be a solid edge. Then the intersection is a disjoint union of two open cylinders and the ratio is a symplectic automorphism of this union preserving each component and also preserving the function Any symplectic automorphism of a fibered cylinder is a Hamiltonian automorphism. The same holds for their union. The corresponding Hamiltonian extends (using a bump function) to a smooth function on all of in such a way that its support is in the preimage of the edge for all Let us denote this globally defined Hamiltonian automorphism by Now we have i.e. the symplectomorphisms and do agree with each other on the preimage of the edge By applying the same procedure to all solid edges of we obtain a globally defined symplectomorphism such that and This completes the proof of part .
Now let us prove the second statement of the theorem. Given a triple we need to construct a quadruple such that and If this is done then
and it follows from Moser’s theorem [16] that there is a diffeomorphism such that so that one can take It follows from [6] that there exists a surface with a simple Morse function and a projection such that It remains to construct a symplectic form such that Let be a singular point of the function or its restriction It follows from the proofs of Theorem 2.3 and Theorem 2.4 that there exists a symplectic form such that for some neighborhood of the vertex Using an appropriate partition of unity we construct a symplectic form as a combination of forms such that . ∎
4 Classification of generic coadjoint orbits of symplectomorphism groups
4.1 From Morse functions to coadjoint orbits
Throughout this section, let be a compact connected symplectic surface with boundary By we denote the Lie group333 See [11, Chapter I, Section 1.1] for details on Lie groups and Lie algebras in an infinite-dimensional setting. of all symplectomorphisms of Note that all elements of preserve the boundary but do not necessarily preserve the boundary pointwise. The group has the Lie algebra of divergence-free vector fields on tangent to the boundary The regular dual space can be identified with the space of cosets (see Appendix). Moreover, the natural action of the group on the space of cosets by means of pull-backs coincides with the coadjoint action of the group of symplectomorphisms
where is a symplectomorphism and is a 1-form.
Define the exterior derivative operator on the space of cosets by the formula (This operator is well-defined on the cosets since Consider the following mapping:
defined by taking a vorticity function It is easy to see that if the boundary of the surface is not empty then the mapping is a surjection. In the case of a closed surface there is a relation:
and the mapping is surjective onto the space of zero-mean functions.
Suppose that cosets and belong to the same coadjoint orbit of . Then by definition, there is a symplectomorphism such that and the following diagram is commutative:
Definition 4.1.
A coset is called simple Morse if is a simple Morse functions. A coadjoint orbit is called simple Morse if some (and hence every) coset is simple Morse.
With every simple Morse coset one can associate a measured Reeb graph If two simple Morse cosets and belong to the same coadjoint orbit then the corresponding Reeb graphs are isomorphic.
Suppose that cosets and have isomorphic Reeb graphs. Then it follows from Theorem 3.13 that there exists a symplectomorphism such that Therefore, the 1-form is closed. Since this 1-form is not necessarily exact, the cosets and do not necessarily belong to the same coadjoint orbit. Nevertheless, we conclude that the space of coadjoint orbits corresponding to the same measured Reeb graph is finite-dimensional and its dimension is at most Throughout this section, unless otherwise stated, all (co)homology groups will be with coefficients in
4.2 Circulation functions on a Reeb graph
In [9] the notion of a circulation function was introduced for the case of closed surfaces. In the case of surfaces with boundary, we need a modification of that definition. Take a point which is not a vertex. Then is a circle It is naturally oriented as the boundary of the set of smaller values of the function . The integral of a coset over is well-defined. Thus we obtain a function
defined by
Proposition 4.2 ([9]).
The function has the following properties.
-
(i)
Assume that an are two interior points of some edge and that is pointing from towards . Then satisfies the Newton-Leibniz formula
-
(ii)
for all vertices of which do not belong to the function satisfies the Kirchhoff rule at :
(11) where the notation stands for the set of edges pointing at the vertex , and stands for the set of solid edges pointing away from .
Note that the function on the subgraph can be recovered from the circulation function by the formula: . It follows from Proposition 4.2 that the difference is as an element of the relative homology group
The above properties of the circulation function make it natural to introduce the following definition of an abstract circulation function.
Definition 4.3.
Let be a measured Reeb graph. Any function satisfying properties listed in Proposition 4.2 is called a circulation function (an antiderivative).
Proposition 4.4.
Let be a measured Reeb graph.
-
i)
If the subgraph is not empty, then the pair on admits an antiderivative.
-
ii)
If the subgraph is empty, then the pair on admits an antiderivative if and only if .
-
iii)
If the pair admits an antiderivative, then the set of antiderivatives of is an affine space whose underlying vector space is the relative homology group
Proof.
To prove this result one applies Proposition in [7] to the graph with the set of boundary vertices defined as those vertices that belong to ∎
4.3 Auxiliary classification result
In this subsection we follow [7]. Let be a symplectic surface with boundary Denote by the space of Morse functions on constant on the boundary and without critical points on the boundary Elements of are called functions of type.
Definition 4.5.
A coset is said to be of -type if A coadjoint orbit called to be of -type if some (and hence every) coset is of -type.
All definitions from the present paper such as Reeb graph, compatibility conditions, circulation graph, etc. can be modified for the case of functions and cosets of type, see details in [7]. The result we are interested in can be formulated as follows.
Theorem 4.6 ([7]).
Let be a connected symplectic surface with or without boundary. Then coadjoint orbits of of type are in one-to-one correspondence with (isomorphism classes of) circulation graphs compatible with . In other words, the following statements hold:
-
i)
For a symplectic surface and cosets of type the following conditions are equivalent:
-
a)
and lie in the same orbit of the coadjoint action;
-
b)
circulation graphs and corresponding to the cosets and are isomorphic.
-
a)
-
ii)
For each circulation graph which is compatible with , there exists a generic such that
4.4 Augmented circulation graph
In the case of surfaces with boundary circulation functions do not form a complete set of invariants for coadjoint orbits, i.e. the equality does not in general imply that cosets and belong to the same coadjoint orbit.
Example 4.7.
Consider the disk with two holes from Figure 5(a). In this case there are no circulation functions since there are no solid edges in the Reeb graph. On the other hand, in this case there are no nontrivial symplectomorphisms preserving the function hence the dimension of the space of coadjoint orbits is equal to the first Betti number of the surface, i.e. it is equal to two.
It turns out that it is possible to define some additional invariants: integrals of cosets over certain cycles associated with the pair in an invariant way.
There is a unique way to lift each edge to a smooth oriented (and diffeomorphic to a segment) curve such that
-
i)
-
ii)
for each the regular -level is pointed in the direction of the curve
We define the subset to be the union
We also define the subset to be
where is the set of boundary vertices (i.e. vertices of types I, III, or V) of the graph And, finally, define the subset to be the union of and (see Figure 7).
The set is a topological graph embedded into the surface We denote by the inclusion
Lemma 4.8.
The map is a homotopy equivalence.
Proof.
Consider the graph obtained from by contracting each connected component of a singular -level in to a point. Denote by the projection The map is a homotopy equivalence since each singular -level in is connected and simply connected. The map factors (in a unique way) through i.e. there exists a unique map such that The map is a homeomorphism. We conclude that the map is a homotopy equivalence as a composition of the inclusion the homotopy equivalence and the homeomorphism ∎
Let be a coset of a one-form. There is a natural way to define the restriction First, we define the restriction as a one-cochain such that for each edge Now we take The cohomology class is well-defined since each exact one-form restricts to the exact one-cochain It follows from Lemma 4.8 that is an isomorphism. Hence with each coset we can also associate an element defined by the formula Next, we generalize the notion of a circulation graph from [9].
Definition 4.9.
A measured Reeb graph endowed with a circulation function and an element is called a augmented circulation graph
We demonstrated above that with each coset one can associate an augmented circulation graph Two augmented circulation graphs are isomorphic if they are isomorphic as measured Reeb graphs, and the isomorphism between them preserves all additional data. An augmented circulation graph is compatible with a symplectic surface if the corresponding measured Reeb graph is compatible with (see Definition 3.12).
4.5 Coadjoint orbits of symplectomorphism groups
Theorem 4.10.
Let be a connected symplectic surface with or without boundary. Then generic coadjoint orbits of are in one-to-one correspondence with (isomorphism classes of) augmented circulation graphs compatible with . In other words, the following statements hold:
-
i)
For a symplectic surface and generic cosets the following conditions are equivalent:
-
a)
and lie in the same orbit of the coadjoint action;
-
b)
augmented circulation graphs and corresponding to the cosets and are isomorphic.
-
a)
-
ii)
For each augmented circulation graph which is compatible with , there exists a generic such that .
Corollary 4.11.
The space of coadjoint orbits of the group corresponding to the same measured Reeb graph is a finite-dimensional affine space and its dimension is
Remark 4.12.
It follows from the long exact sequence for the pair that
Therefore, the space of coadjoint orbits of the group corresponding to the same measured Reeb graph has dimension in the case when the subgraph is connected.
Example 4.13.
Consider the torus with one boundary component from Figure 4 with the height function on it, and the corresponding Reeb graph . In this case and Therefore, the corresponding space of coadjoint orbits is one-dimensional.
Before we proceed with the proof of Theorem 4.10 let us formulate and prove two lemmas.
Lemma 4.14.
Let be a connected oriented surface with non-empty boundary, and let be a simple Morse function on . Then
Proof.
Let be the smooth surface obtained from the surface by contracting each circle in to a point. It is clear that
Let be the canonical projection The function descends to a simple Morse function such that The Reeb graph consists only of dashed edges, and it is coincides with Then the surface is homotopy equivalent to graph Therefore,
(12) | ||||
∎
Lemma 4.15.
Let be a connected oriented surface possibly with boundary, and let be a simple Morse function on . Assume that is such that the integral of over any -level vanishes, and is a zero element in Then there exists a function (with zero restriction on the surface ) such that the one-form is closed, and its cohomology class is equal to . Moreover, can be chosen in such a way that the ratio is a smooth function.
Proof.
Denote by the inclusion and denote by the restriction of the projection on the surface Note that the homomorphism is a surjection, and It follows from Lemma 4.14 that
Hence the image of the homomorphism coincides with the kernel of the homomorphism From above we conclude that the homomorphism is an injection, and
Denote by the inclusion Since the integral of over any connected component of any closed -level vanishes and is a zero element in the cohomology class is a zero element in Consider the long exact cohomology sequence for the pair
The cohomology class on belongs to the kernel of the homomorphism Hence it belongs to the image of the homomorphism i.e. there exists a one-form such that and
Denote by the restriction of the projection on the surface The homomorphism is a surjection, and its kernel is generated by those homology classes which are homologous to regular -levels. From above we conclude that the homomorphism is an injection, and where
Therefore, there exists a one-cochain of the form such that
Recall that the function is the pushforward of the function to the graph Consider a continuous function such that
-
i)
it is a smooth function of in a neighborhood of each point ;
-
ii)
it vanishes whenever is sufficiently close to a vertex;
-
iii)
-
iv)
for each edge , we have
Obviously, such a function does exist. Now, lifting to , we obtain a smooth function with the desired properties. ∎
Proof of Theorem 4.10.
Let us prove the first statement. The implication (a) (b) is immediate, so it suffices to prove the implication (b) (a). Let be an isomorphism of augmented circulation graphs. By Theorem 3.13, can be lifted to a symplectomorphism that maps the function to the function . Therefore, the -form defined by
is closed.
Assume that is a symplectomorphism which maps the function to itself and is isotopic to the identity. Then the composition maps to , and
We claim that can be chosen in such a way that is exact, i.e. one has the equality of the cohomology classes
Moreover, we show that there exists a time-independent symplectic vector field that preserves and satisfies
(13) |
where is the flow of . Differentiating (13) with respect to , we get in the left-hand side
since the form is closed and does not change its cohomology class. Thus
(14) |
Since preserves the circulation function, the integrals of over all connected components of -levels vanish. In addition, Therefore, by Lemma 4.15, there exists a smooth function such that
Now we set
It is easy to see that the vector field is zero on symplectic, preserves the levels of , and satisfies the equation (14). Therefore, its phase flow map has the required properties.
Now let us prove the second statement. It follows from Theorem 3.13 that there exists a symplectic surface and a simple Morse function such that Consider the surface and the restriction of the function to the surface The restriction is a Morse function, and it is constant on the boundary since it is formed by some of closed -levels. However, it is not necessarily a function of -type since it has hyperbolic critical points on the boundary whenever the graph has vertices of type In order to apply the Theorem 4.6 we need to ‘cut out’ from these hyperbolic critical points. Let be a vertex of type let be the only solid edge incident to and also let be the only other vertex adjacent to The edge , with endpoints can be uniquely subdivided into two edges, say and , connecting to a new vertex such that After that we cut out the edge Denote by the (abstract) measured Reeb graph obtained by applying the above procedure to all vertices of type in the graph (see Figure 8).
Denote by the preimage It is clear from the above that the restriction is a function of -type. Therefore, it follows from Theorem 4.6 that there exists a one-form on such that It is clear that the form can be extended (along the cylinders ) on all of in such a way that On the other hand, there exists a one-form on such that and since
Using an appropriate partition of unity we construct a one-form (as a combination of one-forms and ) such that and Hence the augmented circulation graph coincides with ∎
5 Conclusion
In this paper we have classified simple Morse functions on symplectic surfaces and generic coadjoint orbits of symplectomorphism groups of surfaces. This allowed us to completely resolve the questions posed in [8]. The answer to Problem on reconstruction of a surface with boundary from its Reeb graph was given in the work [6]. As described in Section 3 the idea is to add a cyclic order for the dashed edges incident to II- or IV-vertices to the structure of an abstract Reeb graph in order to reconstruct the corresponding surface. Table 1 gives an answer to Problem on measure asymptotics on the graph. The required by Problem compatibility conditions are described in Definition 3.12. And finally, Theorem 4.10 describes the required by Problem additional invariants for coadjoint orbits in case of surfaces with boundary.
One should mention two other relevant but not overlapping with us classification results for symplectic surfaces:
-
a)
Dufour, Molino, and Toulet classified in [5] simple Morse fibrations on closed symplectic surfaces under the action of symplectic diffeomorphisms.
- b)
It would be interesting to extend those classifications for surfaces with boundary. (Note that in [15] in the contrast with the present work Hamiltonian functions are assumed to be constant on the boundary.) It also would be very interesting to classify Morse functions and Morse orbits for the action of
-
a)
the group of Hamiltonian diffeomorphisms of a surface
-
b)
the connected component of the identity in the group for the case of surfaces with boundary.
This would generalise the corresponding results of [9] and the present work to these important subgroups of the symplectomorphism groups.
Appendix A Euler’s equation and coadjoint orbits
In this Appendix we describe following [9] the hydrodynamical motivation of the above classification problems. Consider a symplectic surface with boundary We denote by the Lie group of all symplectomorphisms of and by the corresponding Lie algebra of divergence-free vector fields on A linear functional on is called regular if there exists a smooth 1-form such that the value of on a vector field is the pairing between and
The space of regular functionals on is a dense subset in the space of all continues linear functionals on It turns out that the space of regular functionals can be identified with the space of cosets since exact -forms give zero functionals on divergence-free vector fields. Moreover, the natural action of the group on the space of cosets by means of pull-backs coincides with the coadjoint action of the group of symplectomorphisms . The proof of this fact can be found in [2] (see Section I.8). More information about infinite-dimensional Lie groups can be found in [11].
Now let us fix a Riemannian metric on the surface such that the corresponding area form coincides with the symplectic form The motion of an inviscid incompressible fluid on is described by the Euler equation
(15) |
describing an evolution of a divergence-free velocity field of a fluid flow in , where and the field is tangent to the boundary The pressure function entering the Euler equation is defined uniquely modulo an additive constant by this equation along with the divergence-free constraint on the velocity .
The metric allows us to identify the (smooth parts of) the Lie algebra and its dual by means of the so-called inertia operator: given a vector field on one defines the 1-form as the pointwise inner product with vectors of the velocity field : for all The Euler equation (15) rewritten on 1-forms is
for the 1-form and an appropriate function on . In terms of the cosets of 1-forms , the Euler equation looks as follows:
(16) |
on the dual space , where is the Lie derivative along the field .
The Euler equation (16) shows that the coset of 1-forms evolves by an area-preserving change of coordinates, i.e. it remains in the same coadjoint orbit in . This is why invariants of coadjoint orbits of cosets describe first integrals, called Casimirs, of the Euler equation, and their complete classification is important in many areas of ideal fluid dynamics.
References
- [1] G. Adelson-Welsky and A. Kronrode. Sur les lignes de niveau des fonctions continues possédant des dérivées partielles. Comptes rendus (Doklady) de l’Académie des sciences de l’URSS, 49(4):235–237, 1945.
- [2] V. Arnold and B. Khesin. Topological Methods in Hydrodynamics. Springer, 1999.
- [3] A. Bolsinov. A smooth trajectory classification of integrable Hamiltonian systems with two degrees of freedom. Sbornik: Mathematics, 186(1):1, 1995.
- [4] Y. Colin De Verdière and J. Vey. Le lemme de Morse isochore. Topology, 18(4):283–293, 1979.
- [5] J.P. Dufour, P. Molino, and A Toulet. Classification des systèmes intègrables en dimension 2 et invariants des modèles de Fomenko. Comptes rendus de l’Académie des sciences. Série 1, Mathématique, 318(10):949–952, 1994.
- [6] B. Hladysh and A. Prishlyak. Simple Morse functions on an oriented surface with boundary. Zh. Mat. Fiz. Anal. Geom., 15(3):354–368, 2019.
- [7] A. Izosimov and B. Khesin. Characterization of steady solutions to the 2D Euler equation. Int. Math. Res. Not. IMRN, 2017(24):7459–7503, 2017.
- [8] A. Izosimov and B. Khesin. Classification of Casimirs in 2D hydrodynamics. Mosc. Math. J, 17(4):699–716, 2017.
- [9] A. Izosimov, B. Khesin, and M. Mousavi. Coadjoint orbits of symplectic diffeomorphisms of surfaces and ideal hydrodynamics. Ann. Inst. Fourier, 66(6):2385–2433, 2016.
- [10] A. Jankowski and E. Rubinsztejn. Functions with non-degenerate critical points on manifolds with boundary. Comment. Math., 16(1), 1972.
- [11] B. Khesin and R. Wendt. The Geometry of Infinite-Dimensional Groups. Springer, 2008.
- [12] A. Kirillov. The orbit method, II: Infinite-dimensional Lie groups and Lie algebras. Contemp. Math., 145:33–33, 1993.
- [13] I. Kirillov. Morse-Darboux lemma for surfaces with boundary. J. Geom. Phys., 129:34–40, 2018.
- [14] K. Kourliouros. Local diffeomorphisms in symplectic space and Hamiltonian systems with constraints. J. Geom. Phys., 138:206–214, 2019.
- [15] B. Kruglikov. Exact smooth classification of Hamiltonian vector fields on two-dimensional manifolds. Math. Notes, 61(2):146–163, 1997.
- [16] J. Moser. On the volume elements on a manifold. Trans. Amer. Math. Soc., 120(2):286–294, 1965.
- [17] R. Penna. and the orbit method. J. Math. Phys., 61:012301, 2020.
- [18] G. Reeb. Sur les points singuliers dúne forme de Pfaff complètement intégrable ou d’une fonction numérique. Comptes Rendus de l’Académie des Sciences. Paris, 222:847–849, 1946.
- [19] A. Toulet. Classification des systèmes intégrables en dimension 2. PhD thesis, Université de Montpellier 2, 1996.