Surface mapping class group actions on 3-manifolds
Abstract.
For each circle bundle over a surface with genus , there is a natural surjection . When is the unit tangent bundle , it is well-known that splits. On the other hand does not split when the Euler number is not divisible by the Euler characteristic by [CT23]. In this paper we show that this homomorphism does not split in many cases where divides .
1. Introduction
Let be a closed oriented surface of genus , and let denote the oriented -bundle over with Euler number . Let be the group of orientation-preserving homeomorphisms of that act trivially on the center of , and let denote the mapping class group.
The (generalized) Nielsen realization problem for asks whether the surjective homomorphism
splits over subgroups of . In this paper we study a closely related problem. For each there is a surjection . Consider the composition
Problem 1.1.
Does spilt?
If , then is the unit (co)tangent bundle, and does split; see [Sou10, §1]. On the other hand, if is not divisible by , then the surjection does not split by work of the second two authors [CT23], so also does not split in these cases. Given this, it remains to study the case when is divisible by and . In these cases does split [CT23], but we prove does not split in many cases.
Theorem A.
Fix a surface of genus and . Assume that where and is not a power of , and assume that is divisible by where is an odd prime dividing . Then the natural surjective homomorphism does not split.
For example, if , we find that does not split when .
Theorem A solves the Nielsen realization problem for subgroups of in the cases of the theorem. Specifically, if is divisible by , then [CT23], and every subgroup of is the image of a splitting of . By Theorem A, does not split over any of these subgroups.
Theorem A has the following topological consequence. When divides , there is a “tautological” -bundle whose monodromy
splits the surjection (c.f. [CT23, §1]). One can ask whether or not the bundle is flat. Recall that an -bundle is flat if there is a homomorphism and an -bundle isomorphism . Such bundles are characterized by the existence of a horizontal foliation on , or, equivalently, by the property that their monodromy lifts to . When , the bundle is flat because of the splitting of in this case. When does not split, we deduce that is not flat.
Corollary 1.2.
Fix as in the statement of Theorem A. Then the tautological -bundle is not flat.
Short proof sketch of Theorem A. The proof strategy is similar to an argument of Chen–Salter [CS22] that shows that does not split when . Theorem A is proved by contradiction: assuming the existence of a splitting , first we obtain, by lifting, an action of the based mapping class group on the cover corresponding to the center of . The conditions on and in Theorem A guarantee the existence of a subgroup of for which we can show the action on has a fixed circle. Denoting a generator of by , we show that is generated by the centralizers of and . This shows that the entire group acts on with a fixed circle, which contradicts the fact that the point-pushing subgroup acts freely (by deck transformations) on .
Other questions. Related to the action on the unit tangent bundle , we pose the following question.
Question 1.3.
Do either of the following surjections split?
If one includes orientation-reversing diffeomorphisms and mapping classes, then if , then does not split by Souto [Sou10, Thm. 1].
Acknowledgement. The authors LC and BT are supported by NSF grants DMS2203178, DMS-2104346, and DMS-2005409.
2. Proof of Theorem A
Fix and as in the theorem statement, and set and . Suppose for a contradiction that there is a homomorphism
whose composition with is the identity.
2.1. Step 1: lifting argument
Consider the covering space of , where is the universal cover. This is the covering corresponding to the center of . Given the action of on , we consider the set of all lifts of homeomorphisms in this action to . This is an action of the pointed mapping class group on . To explain this, we start with the following general proposition.
Proposition 2.1.
Let be a closed manifold. Let be the center of the fundamental group, and denote . Let be the covering space with . Fix a basepoint . Assume that the evaluation map
induces a surjection . Then there is a commutative diagram
(1) |
whose rows are exact, where the bottom row is the (generalized) Birman exact sequence. Furthermore, this diagram is a pullback diagram.
We prove Proposition 2.1 after explaining how it gives the desired lifting. In our situation, the center of is the kernel of since has trivial center. Thus Proposition 2.1 gives us the following diagram.
The splitting defines a subgroup and a splitting of over this subgroup. Since the top row is a pullback of the middle row, it follows that splits over (this uses only general facts about pullbacks). Denote this splitting by
Under this splitting the point-pushing subgroup acts by deck transformations.
Remark 2.2.
If and has a fixed point , then after choosing a lift of , one can lift canonically elements of to by choosing the unique lift that fixes . This implies that can be lifted to so that has a fixed point.
Proof of Proposition 2.1.
First we recall the construction of the bottom row of diagram (1). Evaluation at defines a fibration
The long exact sequence of homotopy groups gives an exact sequence
In general the image of is contained in the center of ; see e.g. [Hat02, §1.1, Exer. 20]. By assumption, surjects onto the center, so we obtain the short exact sequence in the bottom row of (1). The homomorphism is the so-called “point-pushing” homomorphism. It sends (basepoint= ) to the time-1 map of an isotopy that pushes around in reverse (this follows directly from the definition of the connecting homomorphism in the long exact sequence; note that it makes sense for the reverse of to appear in defining this homomorphism since concatenation of paths is left-to-right, while composition of functions is right-to-left).
Next we define . Fix a point that covers the basepoint . Given . Choose a path from to and let denote the composition . By isotopy extension, there exists an isotopy where and for each . Define
The map is well-defined. The choice of is unique only up to an element of . This implies that the isotopy class is only well-defined up to composition by a point-pushing mapping class by an element of , but such a point-push is trivial by assumption.
It is a straightforward exercise to check that is a homomorphism. The right square in diagram (1) commutes because and are isotopic by construction. It is easy to see that the left square in the diagram commutes by applying the definition of to deck transformations.
Finally, regarding the claim that the diagram is a pullback, we show that the map to the fibered product
is an isomorphism. The codomain consists of pairs such that is isotopic to a representative of the isotopy class .
We define an inverse to . Given in the fibered product, choose an isotopy from to a homeomorphism representing . Lift to an isotopy such that fixes , and define . The reader can check that the maps and are inverses. ∎
2.2. Step 2: finite group action rigidity
Recall that and is not a power of 2; let be an odd prime dividing . From Step 1, we have homomorphism that descends to a splitting . In this section we describe the action of a particular finite subgroup of on .
Proposition 2.3.
There exists an element of order such that the fixed sets of , , and coincide and are equal to an embedded circle .
It is worth noting that the fixed set of a finite-order, orientation-preserving homeomorphism of a 3-manifold can be wildly embedded [MZ54].
In order to prove Proposition 2.3 we first construct the specific element . Then we prove (Proposition 2.5) a weaker version of Proposition 2.3 with the additional assumption that the action is smooth. Finally, we combine this with a result of Pardon [Par21] and Smith theory to prove Proposition 2.3.
Construction of . We obtain as an element in a dihedral subgroup of , where denotes the dihedral group of order . The dihedral action we use has quotient homeomorphic to and the quotient has a single branch point; it is determined by the homomorphism
By Riemann–Hurwitz, the genus of is . The orbifold has fundamental group
and there is a short exact sequence
(2) |
This sequence induces a homomorphism . We take , where , as defined above, is an odd prime dividing , which exists by assumption. Then is an element of order in the subgroup of .
Remark 2.4.
The argument that follows works equally well when is a genus- surface and has a single branched point. This provides more values of for which the conclusion of Theorem A holds.
Smooth case. Here we prove the following proposition.
Proposition 2.5.
Fix as above. Suppose that and is a splitting of over . Then fixes a unique circle on . Consequently, the fixed set of is nonempty.
The last part of the statement of Proposition 2.5 follows from the preceding statement because the image of under is .
Proof of Proposition 2.5.
First we reduce to a more geometric setting. By Meeks–Scott [MS86, Thm. 2.1], the smooth(!) action preserves some geometric metric on . There are two possibilities for the geometry: if , then has -geometry, and if , then has -geometry. We treat these cases in parallel.
The universal cover (with the induced geometric structure) is either or . In either case, has an isometric foliation by lines whose leaf space is isometric to , and this foliation is preserved by , so there is a homomorphism . Let be the group whose action on the leaves and on the leaf space are both orientation preserving. There is an exact sequence
(3) |
See also [Sco83, §4].
Next consider the group of all lifts of elements of to . This yields an exact sequence
The action of on induces an action of on , where is the center of . This action extends to an action of , and there is a homomorphism
The last map is an isomorphism by the general formula for discrete subgroups .
To prove the proposition, we first identify with (Claim 2.6). Then it is a formal consequence of our setup that , and after showing (Claim 2.7), we show that fixes a unique circle in (Claim 2.8).
Claim 2.6.
Proof of Claim 2.6.
Recall the map from (3). First we identify with . For this, it suffices to show that fits into a short exact sequence
(4) |
where the “monodromy” has image the given subgroup . This implies that because is an extension of the same form (see (2)), and extensions of are determined by their monodromy [Bro82, §IV.3].
To construct the extension (4), first note that the restriction of (3) to is the short exact sequence
The group is normal in because is normal in . Furthermore, the surjection induces a surjection .
The quotient map , which is equivariant with respect to descends to a map that’s equivariant with respect to .
Since is a realization, the induced action of on is a realization of the , and in particular the action on is faithful. Therefore, , and the monodromy of the associated extension
is the given inclusion . This concludes the proof that is isomorphic to .
To finish the proof of Claim 2.6, it remains to show that the intersection of with is . We do this by showing (i) is the center of , and (ii) the center of is contained in . Together with the obvious containment , (i) and (ii) imply .
(i): First note that is central because is central in . On the other hand, the center of is contained in because the center of has trivial center.
(ii): To show the center of is contained in , we show that the center of projects trivially to . This is true because factors through , which has trivial center.∎
We summarize the relation between the relevant groups in Diagram (5).
(5) |
By Claim 2.6, , so takes the form
By construction, this homomorphism is the restriction of to . Since , to show the fixed set of is a circle, it suffices to show the same statement for . To prove this, we first compute .
Claim 2.7.
The group is isomorphic to .
Proof of Claim 2.7.
First note that there is an extension
induced from (3). This sequence is obviously split when since is a product.
Assume now that , and write where is a nonzero integer. Let denote the kernel of the universal cover homomorphism .
We claim that . To see this, note that the extension
pulled back under a Fuchsian representation induces the extension of the unit tangent bundle group , which has Euler number , and there is an -fold fiberwise cover , so the center of is generated by the -the power of the generator of the center of , i.e. .
The inclusion of in (given by left-multiplication) descends to a homomorphism
that defines a splitting of the sequence
Claim 2.8.
Let be an odd prime dividing . If is divisible by , then the fixed set of is a circle.
Before proving the claim, we explain how the factors of act on .
Remark 2.9.
Consider the isomorphism from Claim 2.7. In each case ( or ) the action of on covers the identity of and acts freely by rotation on the circle fibers of . For the action, when , then is a metric product, and the action of is trivial on the factor and is the natural action on . If is nonzero, then
and with respect to this isomorphism, the action of on is induced from left multiplication of on . Identifying with the unit tangent , we can also view as the quotient of by the action that covers the identity of and rotates each fiber.
Proof of Claim 2.8.
Write for some integer .
First note that since has finite order, the induced isometry of has a unique fixed point, so preserves a unique circle of the fibering . The same is true for , and we will show that acts trivially on .
Since is a commutator in and is abelian, we find that the projection
is trivial. Therefore, the action of on factors through acting on . This action is described in Remark 2.9. If , since acts trivially on the factor of , we conclude that acts trivially on . If , then acts as a a rotation by on , so acts as a rotation by , which is trivial. ∎
This completes the proof of Proposition 2.5. ∎
Homeomorphism case. Here prove Proposition 2.3.
Proof of Proposition 2.3.
By Pardon [Par21, Thm. 1.1], there is a sequence of smooth actions converging in to the given action of on . Sufficiently close approximates also give a splitting of over because is locally path connected [EK71].
For each of the smooth approximations of , the fixed set of is nonempty by Proposition 2.5. This implies that has a fixed point (a sequence of fixed points, one for each smooth action, sub-converges to a fixed point of the action). By Remark 2.2, there exists a lift of to a finite order element so that has a fixed point. Since has a unique conjugacy class of finite subgroup of order , the subgroups and are conjugate, so the fixed set of is nonempty.
It remains to show the fixed set of is a circle, and that this circle is the same as the fixed sets of and .
First we show (using Smith theory) that both and have fixed set a single circle (we are not yet claiming/arguing that the fixed sets of and are the same). To see this, we focus on for concreteness. Consider the group of all lifts of powers of to the universal cover . This group is an extension
which is central and split; hence . It is central because acts orientation-preservingly on fibers of (otherwise, the action of on would reverse orientation, contrary to the construction); it splits because has a fixed point. The (lifted) action of on has fixed set a line (i.e. embedded copy of ) by Smith theory and local Smith theory [Bre12, Theorem 20.1], and this line is preserved and acted properly by ; thus acts on with a circle in its fixed set. Furthermore, each component of the fixed set of acting on corresponds to a distinct conjugacy class of order- subgroup of . Since there is only one subgroup of , the fixed set of is connected, i.e. a single circle. The same argument111Smith theory applies to prime-order finite cyclic group actions, so we cannot apply this argument directly to . works for .
Now we determine the fixed set of . First observe that preserves the fixed set of and has a fixed point there (the fixed set of is nonempty and contained in the fixed set of ). The only action on the circle with a fixed point is the trivial action, so in fact the fixed sets of and are the same. The same argument applies to and . This proves Proposition 2.3. ∎
2.3. Step 3: centralizer argument
Recall that we have defined as an element of order in . In this step we prove that is generated by the centralizers of and .
Proposition 2.10 (centralizer property).
Let be the element of order constructed above. Then
where denotes the centralizer in .
Strategy for proving Proposition 2.10. Set . Our method for showing , which is similar to the proof of [CS22, Thm. 1.1], is to inductively build subsurfaces
(6) |
such that for each and fills (i.e. each boundary component of is inessential in ). The fact that fills implies that , so then by the last step in the inductive argument.
In order to ensure that , the subsurface is obtained from by an operation known as subsurface stabilization. If is a subsurface and is a simple closed curve that intersects in a single arc, then the stabilization of along is the subsurface , where is a regular neighborhood of . It is easy to show that is generated by and the Dehn twist [CS22, Lem. 4.2], so if and , then . Therefore, for the proof, it suffices to find a sequence of subsurface stabilizations along curves whose Dehn twist belongs to .
Model for the action. Our proof of Proposition 2.10 makes use of an explicit model for with its action, which is pictured below in the case and (recall that and is an odd prime dividing ).
2pt \pinlabel at 87 154 \pinlabel at 87 208 \pinlabel at 87 262 \pinlabel at 87 320 \pinlabel at 87 373 \pinlabel at 87 429
at 173 154 \pinlabel at 173 208 \pinlabel at 173 262 \pinlabel at 173 320 \pinlabel at 173 373 \pinlabel at 173 429
at 326 154 \pinlabel at 326 208 \pinlabel at 326 262 \pinlabel at 326 320 \pinlabel at 326 373 \pinlabel at 326 429
at 373 154 \pinlabel at 373 208 \pinlabel at 373 262 \pinlabel at 373 320 \pinlabel at 373 373 \pinlabel at 373 429
at 433 154 \pinlabel at 433 208 \pinlabel at 433 262 \pinlabel at 433 320 \pinlabel at 433 373 \pinlabel at 433 429
at 480 154 \pinlabel at 480 208 \pinlabel at 480 262 \pinlabel at 480 320 \pinlabel at 480 373 \pinlabel at 480 429
at 637 154 \pinlabel at 637 208 \pinlabel at 637 262 \pinlabel at 637 320 \pinlabel at 637 373 \pinlabel at 637 429
at 722 154 \pinlabel at 722 208 \pinlabel at 722 262 \pinlabel at 722 320 \pinlabel at 722 373 \pinlabel at 722 429
at 820 218
\pinlabel at 400 485
\pinlabel at 400 460
\pinlabel at 400 123
\pinlabel at 287 291
\pinlabel at 520 291
\pinlabel at 130 485
\pinlabel at 130 460
\pinlabel at 130 123
\pinlabel at 680 485
\pinlabel at 680 460
\pinlabel at 680 123
\endlabellist
The surface is built out of two copies of the standard action of on and one copy of a free action of on . We glue each copy of to along free orbits by an equivariant connected sum. In the figure, acts by vertical translation. Note that the fixed points of on are not pictured in the figure – they are at along the -axis.
To derive this model, recall that has abelianization with kernel . Then there is a sequence of regular covers
The cover is unbranched and is the -homology cover of (in particular, is also a torus). The cover is branched over four points; the local monodromy around the branched points is at two of the branched points and at the other two. Choosing branched cuts joining branched points in pairs gives a model for , and one can check that this model is equivalent to the one described above. (The spheres in Figure 1 arise from pre-images under of neighborhoods of the branch cuts.)
By Remark 2.2, since has a fixed point, the subgroup , which lifts , has a fixed point. The different lifts of to a finite subgroup of are in one-to-one correspondence to fixed points of . Since these fixed points are permuted transitively by the action of , the different lifts of are conjugate. Consequently, for the purpose of our argument, we can choose to be any one of the four fixed points of and prove Proposition 2.10 for this choice, without loss of generality. (It will also be evident from the argument that a similar argument applies if is changed to another fixed point.)
Remark 2.11.
We do not known how generally the relation holds. For example, it may hold for every subgroup of . We do not know a general (abstract) approach to this problem.
Proof of Proposition 2.10.
Symmetry breaking. In preparation for constructing a sequence of subsurface stabilizations, in this paragraph we find a suitable collection of Dehn twists that belong to . The obvious way for to belong to is if is preserved by either or . More generally, we use a process that we call symmetry breaking to show for certain . We formulate this in the following lemma, which is similar to [CS22, Lem. 3.2].
Lemma 2.12 (Symmetry breaking).
Assume that are simple closed curves that intersect once and . Suppose that either (i) is disjoint from and or (ii) the curves are disjoint from and the curves are pairwise disjoint. Then .
Proof of Lemma 2.12.
We prove case (i) of the statement; case (ii) is similar. Since Dehn twists about disjoint curves commute,
The left hand side of the equation is in because by assumption and because the curves are permuted by and are disjoint (so their twists commute), and thus . Since and intersect once, the braid relation implies that also belongs to . Since this implies that , as desired. ∎
Remark 2.13.
When applying Lemma 2.12(i) or (ii) we refer to it as the - or -symmetry breaking, respectively.
Lemma 2.14.
Dehn twists about the curves in Figure 2 are in .
2pt \pinlabel at 87 154 \pinlabel at 87 208 \pinlabel at 87 262 \pinlabel at 87 320 \pinlabel at 87 373 \pinlabel at 87 429
at 173 154 \pinlabel at 173 208 \pinlabel at 173 262 \pinlabel at 173 320 \pinlabel at 173 373 \pinlabel at 173 429
at 326 154 \pinlabel at 326 208 \pinlabel at 326 262 \pinlabel at 326 320 \pinlabel at 326 373 \pinlabel at 326 429
at 373 154 \pinlabel at 373 208 \pinlabel at 373 262 \pinlabel at 373 320 \pinlabel at 373 373 \pinlabel at 373 429
at 433 154 \pinlabel at 433 208 \pinlabel at 433 262 \pinlabel at 433 320 \pinlabel at 433 373 \pinlabel at 433 429
at 480 154 \pinlabel at 480 208 \pinlabel at 480 262 \pinlabel at 480 320 \pinlabel at 480 373 \pinlabel at 480 429
at 637 154 \pinlabel at 637 208 \pinlabel at 637 262 \pinlabel at 637 320 \pinlabel at 637 373 \pinlabel at 637 429
at 722 154 \pinlabel at 722 208 \pinlabel at 722 262 \pinlabel at 722 320 \pinlabel at 722 373 \pinlabel at 722 429
at 820 218
at 400 459 \pinlabel at 400 123 \pinlabel at 287 291 \pinlabel at 520 291
at 130 459 \pinlabel at 130 123
at 680 459 \pinlabel at 680 123
at 410 308 \pinlabel at 354 133 \pinlabel at 385 186 \pinlabel at 130 140 \pinlabel at 507 189 \pinlabel at 409 405 \pinlabel at 130 210

In Figure 2, we illustrate the case . The corresponding curves in the general case belong to by the exact same argument.
Proof of Lemma 2.14.
First observe that and are in because each is invariant under . We deduce that using -symmetry breaking with . Each of and are in by -symmetry breaking with . Finally, both and are in by -symmetry breaking with . ∎
Surface stabilization sequence. We stabilize with the sequence of curves represented in Figure 3. To get the initial subsurface we can take the subsurface spanned by the chain of curves . This subsurface has genus 2 and one boundary component, and these curves are Humphries generators for [FM12, Fig. 4.10]. Next we extend this chain with the curves ; at this point the left genus-0 subsurface has been filled. Next we stabilize with and the curves () that fill the right genus-0 subsurface; there is some choice in the order of curves we stabilize, but this is not important. Finally we stabilize with the curves and that generate .
2pt \pinlabel at 820 218
at 400 460 \pinlabel at 400 123 \pinlabel at 287 291 \pinlabel at 520 291
at 130 460 \pinlabel at 130 123
at 680 460 \pinlabel at 680 123
2pt
at 45 430 \pinlabel at 88 430 \pinlabel at 100 403 \pinlabel at 88 375 \pinlabel at 100 340 \pinlabel at 88 320 \pinlabel at 100 293 \pinlabel at 88 265 \pinlabel at 100 235 \pinlabel at 88 205 \pinlabel at 102 180 \pinlabel at 88 155 \pinlabel at 130 141 \pinlabel at 173 155 \pinlabel at 188 180 \pinlabel at 173 205 \pinlabel at 188 235 \pinlabel at 173 265 \pinlabel at 188 293 \pinlabel at 173 320 \pinlabel at 188 347 \pinlabel at 173 375 \pinlabel at 188 400 \pinlabel at 173 430
at 230 310
at 637 320 \pinlabel at 622 349 \pinlabel at 637 375 \pinlabel at 622 402 \pinlabel at 637 430
at 678 439
at 723 430 \pinlabel at 740 400 \pinlabel at 723 375 \pinlabel at 740 347 \pinlabel at 723 320
at 637 155 \pinlabel at 622 180 \pinlabel at 637 208 \pinlabel at 622 237 \pinlabel at 637 265
at 678 141
at 723 155 \pinlabel at 740 180 \pinlabel at 723 208 \pinlabel at 740 237 \pinlabel at 723 265
at 302 190 \pinlabel at 418 400

2.4. Step 4: conclusion
References
- [Bre12] Glen E Bredon. Sheaf theory, volume 170. Springer Science & Business Media, 2012.
- [Bro82] K. S. Brown. Cohomology of groups, volume 87 of Graduate Texts in Mathematics. Springer-Verlag, New York-Berlin, 1982.
- [CS22] L. Chen and N. Salter. Global fixed points of mapping class group actions and a theorem of Markovic. J. Topol., 15(3):1311–1324, 2022.
- [CT23] Lei Chen and Bena Tshishiku. Mapping class groups of circle bundles over a surface, 2023.
- [EK71] R. Edwards and R. Kirby. Deformations of spaces of imbeddings. Ann. of Math. (2), 93:63–88, 1971.
- [FM12] B. Farb and D. Margalit. A primer on mapping class groups, volume 49 of Princeton Mathematical Series. Princeton University Press, Princeton, NJ, 2012.
- [Hat02] A. Hatcher. Algebraic topology. Cambridge University Press, Cambridge, 2002.
- [MS86] W. H. Meeks, III and P. Scott. Finite group actions on -manifolds. Invent. Math., 86(2):287–346, 1986.
- [MZ54] D. Montgomery and L. Zippin. Examples of transformation groups. Proc. Amer. Math. Soc., 5:460–465, 1954.
- [Par21] J. Pardon. Smoothing finite group actions on three-manifolds. Duke Math. J., 170(6):1043–1084, 2021.
- [Sco83] P. Scott. The geometries of -manifolds. Bull. London Math. Soc., 15(5):401–487, 1983.
- [Sou10] J. Souto. A remark on the action of the mapping class group on the unit tangent bundle. Ann. Fac. Sci. Toulouse Math. (6), 19(3-4):589–601, 2010.