DEFORMATION RETRACTION of the GROUP of STRICT CONTACTOMORPHISMS of the THREE-SPHERE to the UNITARY GROUP
Abstract.
We prove that the group of strict contactomorphisms of the standard tight contact structure on the three-sphere deformation retracts to its unitary subgroup .
The group of strict contactomorphisms of the standard tight contact structure on the three-sphere is known to be the total space of a fiber bundle over the group of orientation-preserving, area-preserving diffeomorphisms of the two-sphere , where the projection is the map given by descending under the Hopf map , and the fiber is the circle subgroup of diffeomorphisms of which rotate all Hopf circles within themselves by the same amount.
Theorem A.
In the category of Fréchet Lie groups and maps, the fiber bundle
deformation retracts to its finite-dimensional subbundle
where the fibers move rigidly during the deformation.
It was already known that this bundle inclusion is a homotopy equivalence, and we improve on that by showing how to lift Mu-Tao Wang’s deformation retraction of the group onto its subgroup to one of onto its subgroup .
Here are the basic definitions.
The Hopf fibration of the three-sphere is a fiber bundle whose fibers are the oriented unit circles on the complex lines through the origin in . The Hopf vector field on is the unit vector field tangent to these oriented great circles. The group of automorphisms of is the subgroup of consisting of diffeomorphisms which permute the oriented great circle fibers of , not necessarily rigidly. They are all orientation-preserving. The subgroup of strict automorphisms of is the subgroup of permuting Hopf fibers rigidly,
The standard tight contact structure on is the field of tangent two-planes which are everywhere orthogonal to the great circle fibers of the Hopf fibration. The standard contact one-form is the inner product with the Hopf vector field, so that , and therefore . The group is the subgroup of consisting of diffeomorphisms whose differential permutes the tangent -planes of , meaning that maps the tangent 2-plane of at to the tangent 2-plane of at , for all . We write and call a a contactomorphism. We have that for some smooth (always meaning here) real-valued nowhere zero function on . If on the nose, then we call a strict contactomorphism or quantomorphism, and denote the group of these by .
We will show in Proposition 3.1 that the group consists precisely of those diffeomorphisms of which simultaneously preserve the Hopf fibration and the standard tight contact structure , that is,
This, in turn, will help us in the proof of the main theorem.

1. Introduction
We give some historical context. The study of the homotopy types of the groups of diffeomorphisms of smooth manifolds and their subgroups has a rich history. By a result of Smale, the diffeomorphism group deformation retracts to the orthogonal group , and by the celebrated Smale Conjecture proved by Hatcher in [Hat83], the diffeomorphism group deformation retracts to the orthogonal group . It is natural to consider diffeomorphisms of which preserve extra structure there and the interplay and homotopy types of resulting moduli spaces. In this paper we are focused on the subgroup of which consists of strict contactomorphisms.
The following exact sequences and bundles have been studied in the literature, and we have also presented detailed self-contained proofs in Part 2.
The exact sequence of Fréchet Lie algebras, equivalently, tangent spaces at the identity
The exact sequence of Fréchet Lie groups
The Fréchet fiber bundle
The finite-dimensional subsequence of (2) and finite-dimensional subbundle of (3)
Leslie [Les67] introduced a differential structure on the group of diffeomorphisms of a differentiable manifold, converting it into a Fréchet Lie group. Banyaga [Ban78b], [Ban78a] presented the above exact sequence (2) of Fréchet Lie groups, attributing this to Souriau [Sou70], and noted its finite-dimensional exact subsequence (4).
The fiber bundle result (3) was proved by Ratiu and Schmid in the Sobolev category [RS81], building on work of Kostant [Kos70], Souriau [Sou70], Ebin and Marsden [EM70], Omori [Omo74], and Banyaga [Ban78b], [Ban78a]. They attribute the exact sequences (1) and (2) of Fréchet Lie algebras and Lie groups to Kostant [Kos70], and used these to derive the bundle result.
Vizman [Viz97] worked in the category, and obtained the exact sequences (1) and (2) above of Fréchet Lie algebras and Lie groups, as well as the Fréchet fiber bundle (3). Casals and Spacil [CS16] also worked in the category, attributed the Fréchet fiber bundle (3) above to Vizman, and showed that the inclusion of the finite-dimensional subbundle (4) into this bundle is a homotopy equivalence. Their further conclusions depended on a result of Eliashberg [Eli92] which was only stated though not proved by him, but later proved by Eliashberg and Mishachev [EM21].
Mu-Tao Wang [Wan01, Wan13] showed how to deformation retract the group of orientation-preserving, area-preserving diffeomorphisms of the two-sphere to its subgroup of orthogonal transformations by applying mean curvature flow in simultaneously to the graphs of all orientation-preserving, area-preserving diffeomorphisms of to itself. Our Theorem A will be proved by lifting this to a deformation retraction of to .
Organization of the paper and plan of the proof of Theorem A
Part I begins with Section 2 where we regard as the group of unit quaternions, and quickly review left-invariant vector fields and differential forms on . We also give a very brief overview of Fréchet spaces, manifolds and Lie groups, which provide the setting for this paper. In Section 3 we will examine the behavior of diffeomorphisms which lie in the group of strict contactomorphisms, and show that this group is the intersection of the groups and . After that, here is the plan for proving the main theorem in Section 4.
We must show that the fiber bundle deformation retracts to its finite-dimensional subbundle . We will start with Wang’s deformation retraction [Wan01] of the base space to , and show how to lift this to the desired deformation retraction of the total space to , in a way that moves fibers to fibers rigidly at all times, while keeping the fibers of the subbundle pointwise fixed.
To begin, we will put the standard Riemannian metric on and show that every smooth path in can be lifted to a smooth horizontal path in , meaning one that is everywhere orthogonal to the fiber direction, and is unique once we specify its starting point.
It is natural to aim to lift Mu-Tao Wang’s deformation retraction of the base space to a deformation retraction of the total space by simply lifting the path followed by each point in to the horizontal path followed by each point in the fiber above it. The problem is that although we can see that the various lifted paths in are smooth in the time direction, we don’t yet know that they are smooth in the transverse direction.
To address this, we will start with Mu-Tao Wang’s deformation in and, using the local product structure from the fiber bundle, define smooth “local lifts” of it to , ignoring the fact that they do not fit together coherently to a global lift. Instead, we will show how to smoothly “adjust” these smooth local lifts to the desired global “horizontal-in-time” lifts there, and so conclude that these horizontal-in-time lifts are themselves smooth.
In Part II, we begin by describing nearest neighbor maps, horizontal lifts and quantitative holonomy in Section 5. In Section 6 we compute the tangent spaces at the identity of our various Lie groups. In Section 7, Section 8 and Section 9, we give independent, self-contained proofs of the exactness of sequences (1) and (2) of Lie algebras and Lie groups described above, and the bundle structure of sequence (3). Finally, we give some background on Fréchet manifolds in Appendix A.
Acknowledgements
It is a pleasure to acknowledge the contributions to this project arising from conversations with Alexander Kupers and Jim Stasheff. We thank Ziqi Fang for perceptive comments on a draft of this paper. Merling acknowledges partial support from NSF DMS grants CAREER 1943925 and FRG 2052988. Wang acknowledges partial support from NSF GRFP 1650114.
Part I Deformation retraction of the strict contactomorphism group
2. Preliminaries
2.1. Fréchet spaces and manifolds
Fréchet spaces, manifolds and Lie groups provide the setting for extending the theory of finite-dimensional differentiable manifolds and maps between them to the infinite-dimensional case. We give a very brief overview here and for more details, we refer the reader to the two papers of Eells [Eel58, Eel66] and that of Leslie [Les67] for early developments and to Hamilton’s paper [Ham82] and the book [KM97] of Kriegl and Michor for a good overview with details. We highlight in Appendix A some results which we use in the present paper.
A Fréchet space is a complete metrizable vector space whose topology is induced by a countable family of semi-norms, where a semi-norm behaves like a norm except that does not imply that . A simple example is the space of maps from the interval to the real numbers. Another example is the space of vector fields on a compact manifold , and yet another example is the space of divergence-free vector fields on with respect to a Riemannian metric on . The semi-norms are the usual norms for on the versions of these spaces.
As in the finite-dimensional case, maps between open subsets of Fréchet spaces are defined in terms of the convergence of various difference quotients. A Fréchet manifold modeled on a Fréchet space is a Hausdorff topological space with an atlas of charts which are homeomorphisms from open sets in into such that the change of coordinate maps are maps. Basic examples are the space of diffeomorphisms of a compact finite-dimensional manifold , equipped with the topology, which is modeled on the Fréchet space , and the space of orientation-preserving, volume-preserving diffeomorphisms of a compact Riemannian manifold, modeled on the Fréchet space . Both are Fréchet Lie groups, meaning that multiplication via composition as well as inversion are smooth maps.
In this paper, we focus on the Fréchet Lie group of strict contactomorphisms (or quantomorphisms) of the standard tight contact structure on the 3-sphere . Since is the total space of an -bundle over , it is modeled, just like , on the Fréchet space . This in turn is isomorphic to the Fréchet space of real-valued functions on the two-sphere . The Fréchet Lie groups of automorphisms of the Hopf fibration of , and of contactomorphisms of the standard tight contact structure on , appear briefly in this paper in the proposition that their intersection is precisely the group , which in turn helps us to better understand . We will study in a forthcoming paper.
2.2. Left-invariant vector fields and differential forms on
We view the 3-sphere as the space of unit quaternions and make the following definitions. Let
(2.1) |
be the standard left invariant vector fields given by right multiplication by . Any smooth vector field on can be written in the basis from Equation 2.1 as
(2.2) |
where and are smooth real-valued functions on .
The Lie brackets of these vector fields satisfy
(2.3) |
The dual left-invariant one-forms to , and on with respect to the standard metric will be denoted by and , so that
and likewise for and . Their exterior derivatives are given by
(2.4) |
We choose the great circle orbits of the vector field as the fibers of our Hopf fibration , so that . It then follows that is the Reeb vector field of , i.e., and .
Viewing , and as directional derivative operators, the differential operators and , acting on a vector field as above, are
(2.5) |
(2.6) |
The gradient of a smooth function is
(2.7) |
The formula for is derived from the identities , , and , which can be verified directly, together with the Leibniz rule
3. The group of strict contactomorphisms
In this section we characterize the strict contactomorphisms of the standard tight contact structure .
Proposition 3.1.
The group of strict contactomorphisms is the intersection of the contactomorphism group with the automorphism group of the Hopf fibration,
[Proof]We begin by showing that . Suppose , so by definition is a diffeomorphism of that satisfies . Recall that the Reeb vector field associated with the 1-form is uniquely characterized by
We consider the pushforward of the vector field by the diffeomorphism , and note that
and
By uniqueness of Reeb vector fields, we have , so . Therefore is in the intersection .
Next, suppose that , so that
where and are smooth real-valued, positive functions on . This gives us that
while at the same time
so it follows that . We now show that , so that , which will imply .
Recall that is the left-invariant orthonormal frame field on . Note that
while at the same time
Thus . Similarly, we can show . Lastly,
and hence the function must be constant. But since Hopf fibers are taken to Hopf fibers with for constant , then must be identically 1. Thus , and so .
We collect a few more useful properties of strict contactomorphisms. Note that the following proposition falls out of the proof of Proposition 3.1, where we showed that for diffeomorphisms , we must have , namely they permute Hopf fibers rigidly.
Proposition 3.2.
The strict contactomorphism group is a subgroup of the strict automorphism group of the Hopf fibration.
Lastly, we record how elements in the simultaneous automorphism group of the Hopf fibration and the standard tight contact structure behave with respect to volume on and area on .
Proposition 3.3.
The diffeomorphisms of in are volume-preserving on and project to area-preserving diffeomorphisms of under the Hopf projection map .
[Proof]Let , so that . Hence takes the contact tangent 2-plane distribution to itself. We show that takes these tangent 2-planes to one another in an area-preserving way, as follows.
Recall the formulas that the dual forms to satisfy from Equation 2.4 and note that the area form on the tangent 2-planes in the distribution is . We compute
Thus indeed takes the 2-planes in the distribution to one another in an area preserving way.
We finish as follows. By Proposition 3.1, , telling us that permutes Hopf fibers rigidly. And as we just saw above, is area-preserving on the tangent 2-planes orthogonal to the Hopf fibers. So it follows that is volume-preserving on . Finally, since the Hopf projection is up to scale a Riemannian submersion (it doubles lengths in orthogonal to the Hopf fibers), it follows that the diffeomorphism of projects to an area-preserving diffeomorphism of , as claimed.
4. Bundle deformation retraction
Since we have shown that , we know that each diffeomorphism of which lies in also lies in , which means that it permutes the fibers of the Hopf fibration and hence induces a diffeomorphism of . The projection map is then defined by . We know from [Viz97] that this is a bundle with fiber . We now begin the proof of our main theorem, namely that the bundle
(4.1) |
deformation retracts to its finite-dimensional subbundle
(4.2) |
We will prove this by starting with Wang’s deformation retraction [Wan01] of the bigger base space to the smaller base space and lifting it to the desired deformation retraction of the bigger total space to the smaller one, , in a way that moves fibers rigidly throughout the deformation retraction while keeping fibers of the subbundle fixed pointwise.
4.1. The standard Riemannian metric on
To facilitate the lifting, we equip with the Riemannian metric. Let and be vector fields on . We define their inner product as
(4.3) |
where the point ranges over and where is the Euclidean volume element on . The scale factor lets unit vector fields on have length equal to 1, since the volume of is , and this will simplify expressions later on.
The left-invariant vector field on , given by , lies in , and its length is 1. By contrast, the left-invariant vector fields and on do not lie in .
At other points , an element of the tangent space is a vector field in along the diffeomorphism , meaning that it assigns to each point a tangent vector to at the point . We will denote such an element of by the symbol , where is a smooth vector field on and where assigns to each point the tangent vector .
Then our Riemannian metric at the point is given by
and is well-defined because all the diffeomorphisms of which lie in are volume-preserving by Proposition 3.3.
This is a smooth, weak Riemannian metric on in the sense that the topology induced on by the norm has fewer open sets than the topology.
The Riemannian metric on is right-invariant, but not left-invariant. The subgroup of which rotates all Hopf fibers by the same amount consists of isometries in this metric, and is the center of the group . The Fréchet group is a Fréchet Lie subgroup of the Fréchet Lie group of volume-preserving and orientation-preserving diffeomorphisms of .
4.2. Lifting a single curve in S to
A path in will be said to be horizontal if it is everywhere orthogonal to the fiber direction with respect to the Riemannian metric. Lifting paths in to horizontal paths in will play a key role in the proof of our main theorem, so we establish the following lemma first.
Lemma 4.1 (Lifting Lemma).
Let be a smooth path in with , and let be an element of such that . Then there exists a unique horizontal path such that and .

[Proof]We start with the quantomorphism bundle
(4.4) |
and then use the map to construct the pullback bundle
(4.5) |
over the interval . The familiar pullback construction extends to the category of Fréchet manifolds and smooth maps [KM97]. The points of the total space are, as usual, the pairs , where and . Since the base space is an interval, the total space is trivial, that is, an annulus diffeomorphic to the product . The bundle map is defined by .
The smooth Riemannian metric on the Fréchet manifold pulls back to a smooth Riemannian metric on the annulus . The horizontal tangent hyperplane distribution on is by definition the orthogonal complement to the one-dimensional vertical fiber direction there. It pulls back to a smooth tangent line field on the annulus which is transverse to the vertical fiber direction there. Though it may not look horizontal to Euclidean eyes, we will say that this line field is “horizontal” on .
Since is finite-dimensional, by the usual existence and uniqueness theorems for ordinary differential equations we get a horizontal path on which begins at the point . In particular, it is a cross-section of the pullback bundle .
Pushing this horizontal path in forward by the bundle map , we get the desired lift of to a horizontal path in which begins at the given point in the fiber .
This completes the proof of the lifting lemma for single curves.
Remark 4.2.
If we let vary over all the points in the -fiber , we get a circle’s worth of disjoint lifts of which are carried to one another by the action of the subgroup of .
4.3. Lifting families of curves in S to
Let
be any smooth deformation of within itself, meaning that , without any other requirements. Then for each we have a smooth path in , and these paths vary smoothly with the choice of initial point . By the Lifting Lemma 4.1, we can lift each of these paths uniquely to a horizontal path in once we specify its initial point
We know that each lifted path is smooth in the time parameter , but we do not yet know that the collection of lifts is smooth in the “transverse direction”, meaning smoothly dependent on the initial points . In this section we prove smooth dependence on initial points.
As mentioned earlier, our plan is to define smooth “local lifts” of these paths, ignoring the fact that they do not fit together coherently to a global lift, and then show how to smoothly “adjust” these to the desired “horizontal-in-time” lifts, which are defined globally, and so conclude that they are indeed smoothly dependent on their initial points.
We start with the following lemma.
Lemma 4.3.
For each point in , there is an open neighborhood of and a partition of the interval such that each image lies in an open set in over which the bundle is trivial.

[Proof]It follows from the continuity of our deformation that for each point in its domain, there is an open neighborhood of in and a real number such that the image lies in an open set in over which our -bundle is trivial.
By compactness finitely many of these open intervals cover , and we can simply let be the intersection of the finitely many corresponding open sets , and choose a partition of subordinate to this covering of . This proves the lemma.
Defining the smooth local lifts
We choose any point and focus on one of the pieces of our tubular neighborhood of the curve . By Lemma 4.3, there is an open set which contains this piece and over which the bundle
is trivial. Let be a smooth trivialization of this bundle over . Then, picking and fixing any point , we have a smooth local lift
of the piece of to the total space of our bundle, as desired.
Adjusting the local lifts to prove that the global horizontal lift is smooth
To simplify the notation, let , and define
We want to adjust each such diffeomorphism along the fiber through it in by an angle so that the corrected family of diffeomorphisms
is horizontal with respect to in the Riemannian metric on for each . In this notation, denotes the diffeomorphism of which rotates all fibers through the angle . For simplicity of notation we will write instead of , and tacitly understand dependence of this angle on the initial diffeomorphism of .
Regard the right side of the equation
as a product in the group , and apply the Leibniz Rule when differentiating it with respect to time to get
where is the time-dependent vector field on generated by the one-parameter family of diffeomorphisms of , so that , and where is a vector field along .
In the last equation, the first term is a vector field on , and so is the second term , even though it may not look so at first glance. The vector field when evaluated at a point , lies in , is tangent to the Hopf fiber through that point, and is scaled to have length . That is the same as the vector field at the point , scaled to length . So we can write
or dropping the point from the notation, we have
Inserting this into the last term of our above computation of the derivative , we get
and will continue on from here.
We keep in mind that our goal is to find the family of rotations of which will make the “adjusted” curves of quantomorphisms
horizontal in time with respect to the Riemannian metric on . To this end, we consider the tangent space to at any point , and let denote its projection to the one-dimensional “vertical” subspace tangent to the -fiber direction,
Then we write
thanks to the invariance of our metric under the action of , and to the fact that is a unit vector tangent to at .
Now we apply this vertical projection to our earlier equation, and set the result equal to zero to require it to be horizontal,
We drop the vertical vector from above and save only its coefficient, recall that , and are left with the scalar equation
or equivalently
We recall that , so Inserting this above, we get
Thus
This makes sense because we want to eliminate the vertical component of in order to move to . Integrating, we get
The last equation tells us that the adjusting angle depends smoothly on the initial angle and on , which itself depends smoothly on the diffeomorphism and the time .
On each subinterval , the initial angle depends smoothly on by the above construction for the preceding time interval , and we start with .
Thus the adjusted family of quantomorphisms
with , and , is horizontal in time and depends smoothly on , which is exactly what we were aiming for.
4.4. Completing the proof of the main theorem
We adjust notation as follows. Let be a smooth path in , with , beginning at the point , and let be a point in . Then we will write
to designate the horizontal path in which covers and which begins at the point . This is the path of quantomorphisms that we called above.
Now we have all the ingredients we need to complete the proof of our main theorem, which we restate.
Theorem 4.4.
In the category of Fréchet Lie groups and maps, the fiber bundle
deformation retracts to its finite-dimensional subbundle
where the fibers move rigidly during the deformation.

[Proof]Let be the deformation retraction of to the orthogonal group given by Mu-Tao Wang’s theorem from [Wan01]. We lift to a deformation retraction
(4.6) |
of to its subgroup by defining
(4.7) |
where is the path in which starts at the point and follows Wang’s deformation retraction, , and where is the horizontal lift of defined above.
The deformation retraction of moves along horizontal curves which cover the corresponding paths of the deformation retraction of . The subgroup of which rotates all Hopf fibers by the same amount consists of isometries in this metric, and so carries horizontal paths to horizontal paths. Thus the fibers of move rigidly among themselves during the deformation retraction
At the end of the deformation retraction, has compressed to the orthogonal group , and has compressed to the unitary group .
A point in which starts out in the subgroup does not move during this process, and likewise a point in which starts out in the subgroup does not move. This completes the proof of our main theorem.
Part II The Fréchet bundle structure of the space of strict contactomorphisms
We begin this part of our paper by describing nearest neighbor maps, horizontal lifts and quantitative holonomy, and then compute the tangent spaces at the identity of our various Lie groups. After that, we give independent, self-contained proofs of the exactness of our sequence of Lie algebras and the exactness and bundle structure of our sequence of Lie groups, proved earlier by the many mathematicians cited in Section 1.
5. Nearest neighbor maps, horizontal lifts and quantitative holonomy
5.1. Nearest neighbor maps
Let and be two Hopf fibers on which are not orthogonal to one another, or equivalently, whose projections to are not antipodal. These two Hopf fibers are a constant distance, say apart on .
Thus, each point on has a unique nearest neighbor on , which is the point that minimizes the distance between and . Similarly, on has on as its nearest neighbor there. Furthermore, the correspondence between on and on is an isometry between these two circles.
The nearest neighbor map between the Hopf fibers and takes the point
to the point on , as depicted in Figure 5.

The composition of nearest neighbor maps is not necessarily the nearest neighbor map , and if we move along a succession of nearest neighbor maps out from and eventually back again to , the composition will be some rotation of . In related settings, a similar phenomenon is called holonomy, so we will use that term here as well.
5.2. Horizontal lifts
Consider the Hopf projection and let be a smooth curve. Given a point on the Hopf fiber , there exists a smooth curve which is unique and runs always orthogonal to Hopf fibers, covers in the sense that and satisfies . We refer to as a horizontal lift of because we think of Hopf fibers as being “vertical” and the orthogonal tangent 2-planes as being “horizontal”. In fact, viewing as a principal -bundle over , the horizontal lift is parallel transport with respect to the connection defined by the 1-form . If is a geodesic in between the non-antipodal points and , then the horizontal lifts of give us the nearest neighbor map between the Hopf fibers and .
5.3. Quantitative holonomy
In the Hopf fibration , we choose radius for the base 2-sphere, so that the projection map is a Riemannian submersion, meaning that its differential takes tangent 2-planes orthogonal to the Hopf fibers isometrically to their images in .

In Figure 6 we consider a loop in based at the point , and the region of that it bounds. We pick a point , and consider the horizontal lift of beginning at .111We warn the reader about the very similar notation for paths, which are denoted by and the dual form to , which is denoted by , since they both appear in this subsection.
The holonomy here is illustrated by the fact that when the lift returns to the fiber , it does so at a point of that fiber, displaced by an angle from the starting point . So followed by the arc on from to is a loop in . This loop bounds a region in , which projects down via to the region on
We claim that the holonomy angle is given by
and confirm this as follows:
using the fact that the Hopf projection is a Riemannian submersion, and so is area-preserving on the 2-form , down to the usual area form on From Equation 2.4 we have that , and hence .
Using Stokes’ theorem, we get
Now consists of two pieces, the arc followed by the arc on from to . Since the arc is horizontal, the one-form is identically zero along it, so we get no contribution to the last integral above. And since the angle along the Hopf great circle measured from to is , the integral of along this arc in the opposite direction is .
Putting all this together, we have
Hence the holonomy of horizontal transport in induced by the loop on is given by the
as claimed above.
Example 5.1.
The equator on bounds a hemisphere of area . The inverse image of is a Clifford torus in , filled with Hopf fibers. The orthogonal trajectories are Hopf fibers of the opposite handedness and are horizontal with respect to the original Hopf fibration. Starting at any location along any original Hopf fiber on this Clifford torus and then following a horizontal circle will bring us back to the antipodal point on the starting fiber. So the holonomy angle in this case is , which is twice the area of .
6. Computation of the Lie algebras
In this section, we give an explicit description of the Lie algebras, or equivalently, the tangent spaces at the identity, of the various Fréchet Lie groups we consider.
Proposition 6.1.
The tangent spaces at the identity to our various subgroups of are as follows.
The tangent space consists of vector fields such that
The tangent space consists of vector fields such that
The tangent space consists of vector fields such that
and these vector fields are divergence-free.
Remark 6.2.
In view of Proposition 3.1, it is natural to ask whether the pair of Lie algebras and are the same. The left side is certainly contained in the right side, and we leave it to the reader to establish the reverse inclusion by manipulating the conditions in parts (a) and (b) of Proposition 6.1.
We start with an intermediary proposition that gives conditions on the vector fields which are in the tangent spaces of interest.
Proposition 6.3.
The tangent spaces at the identity to our various subgroups of admit the following descriptions.
-
,
-
,
-
.
We will see from the proof of Proposition 6.1 that the functions appearing in parts (a) and (b) above have the property that they integrate to zero over each Hopf fiber. Furthermore, for any such function , there exists a vector field on for which , and similarly there exists a vector field on for which .
We prove part (a) here. Parts (b) and (c) can be found in [Gei08, Lemma 1.5.8]. Before we delve into the proof, we make some remarks about the definition of Lie derivatives. Let and be smooth vector fields on the smooth manifold , let and let be the local one-parameter group generated by , meaning that
(6.1) |
Then, the Lie derivative is traditionally defined as
In this definition, the one-parameter group of diffeomorphisms provides the service of pulling the tangent vector in the tangent space to at back to a vector in the tangent space to at , so that one can subtract from it the tangent vector living there. But it is easy to check that any smooth curve satisfying Equation 6.1 can be used to define the Lie derivative as above, and that requiring to be a one-parameter subgroup is just a convention, but not essential. Of course, when is not a one-parameter group, the pullback of can only be defined to be .
[Proof of Proposition 6.3(a)]Let be a smooth vector field which lies in By definition, this means that there is a smooth curve in with and such that for all . Then, as discussed above, the Lie derivative is defined as
But note that here is a path in with , so we can write
since each takes Hopf fibers to Hopf fibers. Therefore in the definition of , for any given , both terms in the numerator are multiples of , so we can factor out of the limit, and we get that for a smooth .
Conversely, suppose is a smooth vector field on with for some smooth function . Let be the one-parameter group of diffeomorphisms of generated by the vector field , i.e., and for each we have Using the group property of this flow, which says that , we compute
Thus holds for all not just .
We need to show that the one-parameter group lies entirely in . Let us use local coordinates in a tubular neighborhood of a Hopf fiber, with and and with as the unit vector field along the Hopf fibers.
We write the vector field in local coordinates as
Then we can compute the Lie derivative
From this we see that and , so the functions and only depend on and and not on . We incorporate this by writing
We also note from above that , which integrates to zero around Hopf circles, and hence
Thus locally the flow covers a flow on the -plane and takes vertical circles to vertical circles, which tells us that each diffeomorphism takes Hopf circles to Hopf circles, and hence , as desired.
Now we turn to the proof of Proposition 6.1, and prove each of its parts separately.
[Proof of Proposition 6.1(a)]Let be a smooth vector field on , written in terms of the orthonormal basis of left-invariant vector fields and on , following the conventions introduced in Section 2. By Proposition 6.3(a), lies in if and only if for some smooth real-valued function on . We compute to see what constraints this conditions imposes on the coefficients and .
Notationally, we switch from Lie derivatives to Lie brackets and compute
using the bracket relations from Equation 2.3.
Therefore, if and only if for some smooth real-valued function , and and . This completes the proof of Proposition 6.1(a).
Note in this proof that since is the negative of the directional derivative of the coefficient around a Hopf circle, we see why must integrate to zero around the Hopf fibers.
[Proof of Proposition 6.1(b)] Again, let be a smooth vector field on . By Proposition 6.3(b), lies in if and only if for some smooth .
Suppose lies in so that for some . Rewrite as and then differentiate to get
Thus
Using the computation for from part (a), we get
thus
Analogously to the computation of in part (a), we can compute
Proceeding as before with rewriting the equations and as and , and differentiating, we get
(6.2) |
Combining with the computations of and above, we get that
as desired.
Conversely, assuming the coefficients of satisfy the conditions in Proposition 6.1(b), using the computations of , and , and working backwards from the computations of the differentiation of the brackets we get
so
Note again that is the directional derivative of the coefficient around Hopf circles, so we reaffirm the observation made after part (a) that must integrate to zero around Hopf fibers.
[Proof of Proposition 6.1(c)]
Let be a smooth vector field on . By Proposition 6.3(c), lies in if and only if .
Suppose lies in so that . Just as in Proposition 6.1(b), rewriting as and then differentiating, we get
But again, by the computation for from part (a), we have thus Just as in part (b), combining the computations for and from part (b) with Equation 6.2, we get
as desired.
Conversely, if we assume that the conditions in Proposition 6.1(c) hold, as we saw in the proof of (b), we get that . Thus if , we immediately get , so by Proposition 6.3(c), lies in .
Lastly, we check that any is divergence free. We have
Remark 6.4.
The conditions on the coefficients of in Proposition 6.1 may seem mysterious at first glance, and it is a rewarding exercise to try to decode their geometric meaning. We give some hints. In part (a), you can take the conditions on the coefficients and and differentiate again in the -direction to show that as the flow of moves a Hopf fiber off itself, it assumes a coiling shape so as to approximate a nearby Hopf fiber. In part (b), another approach to describing is to observe that a vector field is in this space if and only if and both lie in the 2-plane spanned by and , and then compute with Lie brackets.
Having given in Proposition 6.1 a description of the tangent space at the identity to our various subgroups of , we end Section 6 now with a similar description of the tangent spaces and .
We can doubly appreciate our ability to write vector fields on in terms of left-invariant vector fields and when we turn to and seek a similar description there. But we can use the Hopf projection to uniquely lift smooth vector fields on to smooth horizontal fields on , that is, vector fields which are orthogonal to the Hopf fibers and, by virtue of lifting from , twist around each Hopf fiber so they lie in . This allows us to think of and as subspaces of , and therefore rely on expressions in terms of and to describe the vector fields therein. With this identification in mind, we prove the following proposition.
Proposition 6.5.
The tangent spaces and have the following descriptions.
The tangent space consists of vector fields such that
The tangent space consists of vector fields such that
[Proof]For part (a), note that we know from Proposition 6.1(a) that the tangent space consists of vector fields such that is any smooth function on , , and If is horizontal, then . Thus the conditions in part (a) are certainly necessary for to be the horizontal lift of a vector field in .
Conversely, suppose that a vector field on satisfies the conditions in part (a), and since , write .
We claim that the horizontal vector field is the lift of a vector field on if and only if . To see this, note that the left-invariant vector field on is the infinitesimal generator of the one parameter subgroup of consisting of diffeomorphisms of , for , which uniformly rotate all Hopf fibers by the same amount. Then the horizontal vector field is the lift of a vector field on if and only if , which is equivalent to .
From our computation of in the proof of Proposition 6.1(a), and setting , we have
which is equal to by our conditions in part (a). Thus is the lift of a vector field on . So the stated conditions are both necessary and sufficient for to lie in .
For part (b) of our current proposition, it is easy to check that consists of all divergence-free vector fields on . We claim that a vector field is divergence-free if and only if its horizontal lift to is divergence-free, which in turn is equivalent to the condition that =0. This will show that the extra condition in part (b) is both necessary and sufficient for a vector field from part (a) to actually lie in the subspace .
To prove the claim, let be the one-parameter group of diffeomorphisms of generated by the vector field , and let be their lifts to a one-parameter group of diffeomorphisms of generated by the lifted vector field .
If we assume that the lifted field is divergence-free, then the diffeomorphisms are volume-preserving on . Moreover, since is orthogonal to the Hopf fibers, the diffeomorphisms take Hopf fibers rigidly to Hopf fibers. It then follows that the diffeomorphisms must be area-preserving on and their generating vector field must be divergence-free on .
Conversely, if we assume that the vector field on is divergence-free, it follows that the diffeomorphisms are area-preserving there. Then, since the horizontally lifted vector field on is the infinitesimal generator of a one-parameter subgroup of diffeomorphisms of which take Hopf fibers rigidly to one another, and which cover the area-preserving diffeomorphisms of , the diffeomorphisms must be volume-preserving on and hence the vector field must be divergence-free there.
This proves the claim, and completes the proof of Proposition 6.5.
7. The exact sequence of Lie algebras
In this section, we establish the exactness of the following sequence on the level of Lie algebras. We note that this result also appears in [RS81], where Ratiu and Schmid attribute it to [Kos70], but give their own proof. We give our own version of a proof here, building on our explicit computation from the previous section.
Proposition 7.1.
The sequence of tangent spaces
is an exact sequence of Lie algebras.
Before turning to the proof, we give explicit descriptions of the tangent spaces in the sequence, which are computed in detail in Section 6. Writing a smooth vector field on as as in Section 2, the conditions on the coefficients and , which describe membership in the tangent spaces in question are as follows:
-
(1)
if and only if
-
(2)
if and only if
We view as horizontal vector fields on , which push forward consistently along Hopf fibers to divergence-free vector fields on , where by “consistently” we mean that for all in the same Hopf fiber. With this interpretation, we get the following description.
-
(3)
if and only if
It is easy to see (1), whereas (2) is proved as part (c) of Proposition 6.1 and (3) is Proposition 6.5.
[Proof of Proposition 7.1] We start by showing that the maps and do restrict to maps between tangent spaces. First, in order for , must be constant, so we have .
For , we have . To show that this lives in , we need to verify that if and satisfy the conditions in (2), then satisfy the conditions in (3). Using the description of and from (2), note that the condition is equivalent to This equality can be seen to be true using the bracket formula and the fact that is also assumed to be 0. In a similar fashion, we can show that . Lastly, again using the description of and from (2), we get that .
Now we turn to exactness of the sequence. The map is injective, so we have exactness at . To see exactness at , first note that by definition it follows immediately that . To see the reverse inclusion, suppose and suppose . Then and . But then Thus is constant on , and .
Lastly, to verify exactness at we need to check that is surjective. Suppose that , so the coefficients satisfy the conditions in (3). We need to find a smooth function such that the vector field lies in , i.e., so that and satisfy the equations in (2). Combining the conditions on and from (2) and (3), we have , and .
Plugging this into the gradient formula from Equation 2.7, we are seeking so that
On we can solve for if and only if . From Equation 2.6, after simplifying, we get
Differentiating the equations and with respect to , we get and , thus our equation reduces to
Furthermore, using the equations , and , we conclude that , and thus for some smooth function , as desired. This completes the proof of the proposition, namely that our sequence of tangent spaces is exact.
8. The exact sequence of Fréchet Lie groups
In this section we establish the exactness of the sequence on the level of Lie groups. More precisely, we give an independent proof of the following theorem, originally due to Banyaga [Ban78b, Ban78a], Souriau [Sou70] and Kostant [Kos70].
Theorem 8.1.
The sequence of Fréchet Lie groups
(8.1) |
is exact.
The subgroup in the above exact sequence is the set of diffeomorphisms which rotate the Hopf fibers within themselves by the same angle. The projection starts with a diffeomorphism in and then records the resulting permutation of the Hopf fibers. We can write
(8.2) |
where is the Hopf map.
The proof of Theorem 8.1 is broken down into two lemmas, corresponding to the two main challenges: proving that the kernel of is no larger than the subgroup , and proving that the map is onto . The map from into is just the inclusion, so exactness there is automatic.
Lemma 8.2.
The sequence from Equation 8.1 is exact at .
[Proof]The map takes the subgroup of to the identity of , because the elements of this subgroup just rotate the fibers within themselves, and so induce the identity map of to itself. Thus, to confirm exactness at , the challenge is to show that the kernel of is no larger than this subgroup.

We start with an element which takes each Hopf fiber rigidly to itself, and show that it rotates each fiber within itself by the same amount.
We consider two Hopf fibers and , and connect the points and of by a geodesic arc there. We can assume these points and are not antipodal, since we only need to show that the amount each Hopf fiber is rotated by is locally constant. With this choice, the geodesic arc connecting and is unique, and we have a well-defined nearest neighbor map between and .
Then we choose two points and on the fiber , and consider the two horizontal lifts and of which begin at and . These horizontal lifts are geodesics in , and they end on the fiber at the points and which are the nearest neighbors there to the points and , respectively, on .
Since the nearest neighbor map from to is an isometry between Hopf fibers, the angle between and on the first fiber is the same as the angle between and on the second fiber.
Now given , we choose to be . Since is a contactomorphism, it permutes the contact tangent -planes among themselves, and so in particular takes horizontal curves to horizontal curves in .
It follows that , and in particular . This means that the angle between the points and on the Hopf fiber is the same as the angle between the points and on the Hopf fiber . Thus, rotates all fibers by the same amount, which means that , which is what we wanted to prove. This confirms exactness of our sequence of Fréchet Lie groups at .
We turn now to exactness at , following the approach introduced by Ratiu and Schmid in [RS81]. Given the Hopf projection and a path in , we denote by
(8.3) |
the horizontal transport along , in which each point of the first fiber moves along the horizontal lift of to a point on the second fiber, as introduced in Section 5.2. This rigid motion between great circle fibers is the continuous analog of our nearest neighbor maps. Recall from Section 5.2 that if the path in is a geodesic arc, then the map in Equation 8.3 is precisely the nearest neighbor map between these two Hopf fibers.
Recall that the subgroup of strict automorphisms of is the subgroup of permuting Hopf fibers rigidly,
The following lemma characterizes the strict automorphisms of the Hopf fibration which commute with horizontal transport.
Lemma 8.3.
Let induce through . Then if and only if
(8.4) |
for all smooth curves in .

[Proof]If , then takes horizontal curves in to horizontal curves. In particular, in Figure 8, takes the horizontal curve labeled , which runs from to , to the horizontal curve labeled , which runs from to . Thus, .
Conversely, suppose that for all smooth curves in . Then given any point , choose two horizontal curves through whose tangent vectors at span the tangent -plane . Since takes horizontal curves in to horizontal curves, its differential must take to , which means . Since we started out with , we have .
Lemma 8.4.
[Proof]We start out with a diffeomorphism , which we want to lift to an automorphism .

We fix a point to serve as our base point throughout the proof and then begin the definition of the diffeomorphism of by requiring that it take the Hopf fiber rigidly to the Hopf fiber in an orientation-preserving but otherwise arbitrary way. We let
(8.5) |
be this map, which is determined up to a rigid rotation.
Next, consider an arbitrary point and its projection in . We connect and with an arbitrary smooth path in , so that and , and let denote its unique horizontal lift to a path in which ends at , meaning , as in Figure 9.
Let be the beginning point of this lifted path, so that lies somewhere on the Hopf fiber . In the notation of horizontal transport, we can write . The diffeomorphism has already been defined on this “base” Hopf fiber, so we know the point .
Now consider the smooth path in , which runs from to . The unique horizontal lift of this path which begins at is shown in the figure. Horizontal transport in along this horizontal lift takes the point to the point that we will define to be , that is,
(8.6) |
We will show that the definition of does not depend on the choice of the path from to in , and this will follow from the fact that the diffeomorphism of is area-preserving. To that end, let be another smooth path in from to , shown in Figure 9.
We must show that
(8.7) |
Consider the loop in based at that runs through and then backwards. The image under of this loop is the loop based at . Then a little transposing of terms in Equation 8.7 gives us
(8.8) |
Since is area-preserving, the areas enclosed by the loops and are the same. Hence, by the results of Section 5.3, the holonomy experienced by the horizontal lifts of these loops are equal, and preserved by the rigid motion between the fibers. This confirms Equation 8.8, and hence that does not depend on the choice of the path in running from to . A different choice of basepoint in in this construction would result in a new map which differs from by a uniform rotation on all Hopf fibers.
We note that by construction covers , i.e., . Equation 8.6, which defines , together with Lemma 8.3 show that is in . Since takes Hopf fibers rigidly to Hopf fibers and covers the diffeomorphism , its differential at each point cannot have a nontrivial kernel. Hence is a submersion from to itself, thus a covering map, and since is simply connected, is a diffeomorphism. We leave the proof of smoothness of for Appendix A.
This concludes the proof of exactness of the sequence of Fréchet Lie groups stated in Theorem 8.1.
9. The fiber bundle structure
The goal of this section is to give an independent proof of the following theorem, originally due to Vizman [Viz97].
Theorem 9.1.
The sequence
(9.1) |
is a fiber bundle in the Fréchet category.
[Proof]This amounts to constructing slices over small open sets in , and then using the action of the subgroup to promote these slices to the product neighborhood needed to confirm the bundle structure.
First, we note that , which was defined by the formula
(9.2) |
depends smoothly on . This follows from the fact that the composition map
(9.3) |
is smooth in the Fréchet category, together with the fact that is smooth as a function of , and (see Proposition A.1 and Proposition A.5).
Second, we restrict attention to a small neighborhood of the identity , for example the set
(9.4) |
where we regard as the sphere of radius so that the Hopf projection is a Riemannian submersion. Restricting to this open set will let us uniquely define the nearest neighbor map from to to serve as the map .
To construct our slice, define by
(9.5) |
Note that the nearest neighbor map between Hopf fibers depends smoothly on [Eel66], and is chosen as the (unique) shortest geodesic connecting and , which is possible since .
Hence is a smooth map of Fréchet manifolds, with
(9.6) |
This is the slice over for the proposed bundle (9.1). We now promote this slice to a product neighborhood in over by using the action of the circle group as follows. Let
(9.7) |
where the right hand side takes the element of and either follows or precedes it (same result) by uniformly rotating all Hopf fibers through the angle . Since multiplication in the Fréchet Lie group is smooth, it follows that (9.7) is a smooth map of Fréchet manifolds. To check that it gives the local product structure required to confirm that (9.1) is a Fréchet fiber bundle, we write down its inverse explicitly and check that it is also smooth.

To define , let be any diffeomorphism of lying in the tube and let . Then define , and since , the diffeomorphisms and lie in the same circular fiber , separated by some angle . We identify this angle by . Define
(9.8) |
where . Since depends smoothly on and depends smoothly on , and since inversion and multiplication in the Fréchet Lie group are smooth maps, we see that also depends smoothly on . The equations
(9.9) |
confirm that and are indeed inverses of each other, and this proves that is a diffeomorphism, so that we have a bundle structure over the open neighborhood of the identity in .
Finally, the fact that the map is a smooth homomorphism of Fréchet Lie groups provides the homogeneity needed to transfer the above argument to small open sets throughout . This completes our proof that is a fiber bundle in the world of Fréchet manifolds and smooth maps between them.
Appendix A Fréchet spaces and manifolds
For convenience, we give a brief introduction to Fréchet spaces and manifolds in this appendix. After that, we prove some technical results which are used in the proof of the main theorem. For more on this subject, we refer the reader to [Ham82] and [Omo74].
A.1. Fréchet spaces
Let be a vector space. A seminorm on is a function satisfying the following properties:
-
(1)
;
-
(2)
.
If implies , then is called a norm.
An arbitrary collection of seminorms on induces a unique topology on by declaring that a sequence in converges to if and only if for all . From this, we declare that a subset is closed if it contains its limit points. This topology makes into a topological vector space, in the sense that the operations of addition and multiplication by scalars are continuous.
Fix a collection of seminorms on and let be the topology generated by them. We say that two collections of seminorms are equivalent if they generate the same topology. Then is metrizable if and only if it admits an equivalent countable family of seminorms, . In this case, we can define an explicit metric by
(A.1) |
In this paper, we are interested in the metrizable case, so we work under this assumption from now on. The topology is Hausdorff if and only if for all implies , and it is complete if every Cauchy sequence converges. A sequence in is Cauchy if, for each fixed , we have as .
A vector space equipped with a countable family of seminorms is a Fréchet space provided that the topology induced by , as described above, is Hausdorff and complete.
Let and be Fréchet spaces and be an open set. We say that a continuous map is differentiable at in the direction provided that the limit
(A.2) |
exists. If this limit exists for all and all , we can form the map
(A.3) |
If is continuous, as a map from with the product topology into , then we say is or continuously differentiable. We avoid thinking of as a map into , since this is usually not a Fréchet space in a natural way. This definition is weaker than the one usually given for maps between Banach spaces.
Proceeding inductively, we define the second derivative of as
(A.4) |
and say that is provided that the map
(A.5) |
exists and is continuous, and likewise for .
We say that is smooth provided it is for all . This notion of smoothness agrees with the standard one in the case where and are finite dimensional.
A standard example of a Fréchet space is , the set of all smooth functions from to , equipped with the family of seminorms given by
(A.6) |
for , with the convention that . One can readily check the Hausdorff and completeness conditions.
A.2. Fréchet manifolds
A Fréchet manifold modeled on is a Hausdorff topological space with an atlas of homeomorphisms between open sets of and of such that the transition maps
are smooth maps between Fréchet spaces.
Let be a Fréchet manifold and a closed subset of . We say that is a Fréchet submanifold of if for every , there exists a coordinate chart of with and a subspace of such that
(A.7) |
We say that is a coordinate chart adapted to .
At any point , the tangent space can be defined as follows. First, consider the set of all triples , where is a local chart at and . We say that two triples , , are equivalent if
Then is the set of all such equivalence classes. Although this is a rather cumbersome description of the tangent space, in many situations a much more concrete one is available, as we shall see below. In what follows, we describe in detail a number of Fréchet manifolds that are used throughout the paper.
A.3. Examples
Let be a smooth, closed, finite-dimensional manifold. Then the group of all diffeomorphisms from to itself, equipped with the topology, is a Fréchet manifold. Following [Eel66], we describe an atlas for , modeled on Fréchet spaces of vector fields.
Let be the space of all smooth vector fields on . Choose a Riemannian metric on and let denote its Levi-Civita connection. For each , let
(A.8) |
where
(A.9) |
The vector space equipped with the collection of seminorms is a Fréchet space (cf [Ham82]). More generally, given , we let
(A.10) |
The set of vector fields along is again a Fréchet space, and the map is a linear isomorphism between and .
Let be the exponential map associated with the Riemannian metric on . Given a diffeomorphism , there exists an open neighborhood containing the zero section, and an open neighborhood containing such that
(A.11) |
is a homeomorphism ([Les67], [Omo74], [KM97]). We see from the definition that the transition maps are smooth. The collection of maps cover , and the maximal atlas compatible with this collection defines the manifold structure on . Furthermore, this manifold structure makes a Fréchet Lie group, in the sense that the natural operations of multiplication
(A.12) |
and inversion
(A.13) |
are smooth. We remark that it is possible to model as a Banach manifold, if we choose to work with the topology, or a Hilbert manifold, using Sobolev topologies. In this case, we could construct coordinate charts in the same way as (A.11). However, the resulting Banach or Hilbert manifold would not be a Lie group: both the composition and the inversion maps above would be continuous but not differentiable.
On the other hand, a disadvantage of working in the Fréchet category, as opposed to the Banach or Hilbert category, is that the classical Inverse Function Theorem is no longer true. Instead, it must be replaced by the celebrated Nash-Moser Inverse Function Theorem; see [Ham82] for a detailed account of this. We will not need this theorem here.
The propositions to follow, Proposition A.1 through Proposition A.5, are there to help us prove that the diffeomorphism from Lemma 8.4 and Theorem 9.1 depends smoothly on the point , the path and the diffeomorphism , the ingredients which went into its construction.
Proposition A.1.
The space of maps from the interval into is a Fréchet manifold.
[Proof]Fix a curve . Then we can parametrize nearby curves in by the Fréchet space
of vector fields on along , where is the projection from the tangent bundle of to . The correspondence between these vector fields and curves near is given by the Riemannian exponential map
where is the subset of vector fields along with magnitude less than .
The inverse of this map is given as follows. If is a curve close to , meaning that the spherical distance for all , then there exists a unique geodesic from to with initial velocity . By construction,
This proves that is a Fréchet manifold (cf [Ham82, Example 4.2.3]).

Proposition A.2.
The set
is a Fréchet manifold, and a smooth submanifold of .
[Proof]Fix a point . We will show that points in near can be parametrized by vectors in the Fréchet space
Choose a local trivialization of the Hopf fibration
containing . Using this trivialization, for each we write
(A.14) |

Now, given with and sufficiently small, we first let
as before, so is a curve in near the original . Then write according to the decomposition Equation A.14, and set
where is the exponential map in . The geodesic is horizontal to the Hopf fibers, since it starts that way and is a Riemannian submersion. The map
(A.15) |
is our coordinate chart for . It is clear that any pair sufficiently close to can be obtained in this way as the image of some under .
Proposition A.3.
The map , which takes a pair to the unique horizontal lift of starting at , is smooth.
[Proof]Let . By definition, is the unique solution of the system
(A.16) |
which depends smoothly on the initial condition and the parameter . We will compute this dependence explicitly when lifting curves from to .

Borrowing notation from the proof of Proposition A.3, we let be the map that sends to the endpoint of its lift. Then this is also a smooth map.
Proposition A.4.
The map is smooth.
[Proof]Note that , where
The map is smooth: its first derivative at any is
which is a bounded map between Fréchet spaces. The same remark applies for higher derivatives. Since is a composition of smooth maps, it is also smooth by the Chain rule.
We now turn to our main goal in this appendix, which is to prove explicitly that the map defined in Equation 8.6 is smooth. Recall that to define this map, we first fix a point and a rigid motion
between the Hopf fibers and , where is a given area-preserving diffeomorphism. Then, is given by the composition
(A.17) |
where is any path in between and and the maps are the horizontal transport maps defined in Equation 8.3.
Proposition A.5.
The map is smooth as a function of the point , the path and the diffeomorphism .
[Proof]It suffices to check that each of the factors in Equation A.17 is smooth. To do that, we first focus on the points with close to the base point . Given such an , choose to be the unique shortest geodesic between and . Then depends smoothly on and
(A.18) |
in turn depends smoothly on , by Proposition A.3 and Proposition A.4. Similarly,
(A.19) |
and since is a fixed rigid motion, it follows that is smooth, at least on a neighborhood of the fiber . To treat the case where is far away from this fiber, it suffices to note that we can choose a different base point whose fiber is close to , since this new choice of base point will yield the same map up to a uniform rotation of all fibers. Thus, is everywhere smooth.
References
- [Ban78a] Augustin Banyaga, The group of diffeomorphisms preserving a regular contact form, Topology and algebra (Proc. Colloq., Eidgenöss. Tech. Hochsch., Zurich, 1977), Monograph. Enseign. Math., vol. 26, Univ. Genève, Geneva, 1978, pp. 47–53. MR 511781
- [Ban78b] by same author, Sur la structure du groupe des difféomorphismes qui préservent une forme symplectique, Comment. Math. Helv. 53 (1978), no. 2, 174–227. MR 490874
- [CS16] Roger Casals and Oldřich Spáčil, Chern-Weil theory and the group of strict contactomorphisms, J. Topol. Anal. 8 (2016), no. 1, 59–87. MR 3463246
- [Eel58] James Eells, Jr., On the geometry of function spaces, Symposium internacional de topología algebraica International symposium on algebraic topology, Universidad Nacional Autónoma de México and UNESCO, Mexico City, 1958, pp. 303–308. MR 0098419
- [Eel66] by same author, A setting for global analysis, Bull. Amer. Math. Soc. 72 (1966), 751–807. MR 203742
- [Eli92] Yakov Eliashberg, Contact -manifolds twenty years since J. Martinet’s work, Ann. Inst. Fourier (Grenoble) 42 (1992), no. 1-2, 165–192. MR 1162559
- [EM70] David G. Ebin and Jerrold Marsden, Groups of diffeomorphisms and the motion of an incompressible fluid, Ann. of Math. (2) 92 (1970), 102–163. MR 271984
- [EM21] Yakov Eliashberg and Nikolai Mishachev, The space of tight contact structures on is contractible, arXiv:2108.09452 (2021).
- [Gei08] Hansjörg Geiges, An introduction to contact topology, Cambridge Studies in Advanced Mathematics, vol. 109, Cambridge University Press, Cambridge, 2008. MR 2397738
- [Ham82] Richard S. Hamilton, The inverse function theorem of Nash and Moser, Bull. Amer. Math. Soc. (N.S.) 7 (1982), no. 1, 65–222. MR 656198
- [Hat83] Allen E. Hatcher, A proof of the Smale conjecture, , Ann. of Math. (2) 117 (1983), no. 3, 553–607. MR 701256
- [KM97] Andreas Kriegl and Peter W. Michor, The convenient setting of global analysis, Mathematical Surveys and Monographs, vol. 53, American Mathematical Society, Providence, RI, 1997. MR 1471480
- [Kos70] Bertram Kostant, Quantization and unitary representations. I. Prequantization, Lectures in modern analysis and applications, III, 1970, pp. 87–208. Lecture Notes in Math., Vol. 170. MR 0294568
- [Les67] J. A. Leslie, On a differential structure for the group of diffeomorphisms, Topology 6 (1967), 263–271. MR 210147
- [Omo74] Hideki Omori, Infinite dimensional Lie transformation groups, Lecture Notes in Mathematics, Vol. 427, Springer-Verlag, Berlin-New York, 1974. MR 0431262
- [RS81] Tudor S. Raţiu and Rudolf Schmid, The differentiable structure of three remarkable diffeomorphism groups, Math. Z. 177 (1981), no. 1, 81–100. MR 611471
- [Sou70] J.-M. Souriau, Structure des systèmes dynamiques, Dunod, Paris, 1970, Maîtrises de mathématiques. MR 0260238
- [Viz97] Cornelia Vizman, Some remarks on the quantomorphism group, Proceedings of the Third International Workshop on Differential Geometry and its Applications and the First German-Romanian Seminar on Geometry (Sibiu, 1997), vol. 5, 1997, pp. 393–399. MR 1723630
- [Wan01] Mu-Tao Wang, Deforming area preserving diffeomorphism of surfaces by mean curvature flow, Math. Res. Lett. 8 (2001), no. 5-6, 651–661. MR 1879809
- [Wan13] by same author, Mean curvature flows and isotopy problems, Surveys in differential geometry. Geometry and topology, Surv. Differ. Geom., vol. 18, Int. Press, Somerville, MA, 2013, pp. 227–235. MR 3087921