Skip to Main Content

Stability Properties of Moduli Spaces

Rita Jiménez Rolland
Jennifer C. H. Wilson

Communicated by Notices Associate Editor Steven Sam

Article cover

1. Moduli Spaces and Stability

Moduli spaces are spaces that parameterize topological or geometric data. They often appear in families; for example, the configuration spaces of points in a fixed manifold, the Grassmannians of linear subspaces of dimension in , and the moduli spaces of Riemann surfaces of genus . These families are usually indexed by some geometrically defined quantity, such as the number of points in a configuration, the dimension of the linear subspaces, or the genus of a Riemann surface. Understanding the topology of these spaces has been a subject of intense interest for the last 60 years.

For a family of moduli spaces we ask:

Question 1.1.

How does the topology of the moduli spaces change as the parameter changes?

For many natural examples of moduli spaces , some aspects of the topology get more complicated as the parameter gets larger. For instance, the dimension of frequently increases with as well as the number of generators and relations needed to give a presentation of their fundamental groups. But, maybe surprisingly, there are sometimes features of the moduli spaces that ‘stabilize’ as increases. In this survey we will describe some forms of stability and some examples of where they arise.

1.1. Homology and cohomology

Algebraic topology is a branch of mathematics that uses tools from abstract algebra to classify and study topological spaces. By constructing algebraic invariants of topological spaces, we can translate topological problems into (typically easier) algebraic ones. An algebraic invariant of a space is a quantity or algebraic object, such as a group, that is preserved under homeomorphism or homotopy equivalence. One example is the fundamental group of homotopy classes of loops in a topological space based at the point . Homology and cohomology groups are other examples and are the focus of this article. Their definitions are more subtle than those of homotopy groups like , but they are often more computationally tractable and are widely studied.

Given a topological space and , we can associate groups and , the th homology and cohomology groups (with coefficients in ), where is a commutative ring such as or . These algebraic invariants define functors from the category of topological spaces to the category of -modules: for any continuous map of topological spaces there are induced -linear maps

The cohomology groups in fact have the structure of a graded -algebra with respect to the cup product operation.

The group is the free abelian group on the path components of the topological space and is its dual. If is path-connected, is naturally isomorphic to the abelianization of with respect to any basepoint , and its elements are certain equivalence classes of (unbased) loops in .

For a topological group there exists an associated classifying space for principal -bundles. It is constructed as the quotient of a (weakly) contractible space by a proper free action of . The space is unique up to (weak) homotopy equivalence. If is a discrete group, then is precisely an Eilenberg-MacLane space , i.e., a path-connected topological space with and trivial higher homotopy groups. For example, up to homotopy equivalence, is the circle, is the infinite-dimensional real projective space , and the Grassmanian of -dimensional linear subspaces in is .

Some motivation to study the cohomology of : its cohomology classes define characteristic classes of principal -bundles, invariants that measure the ‘twistedness’ of the bundle. For instance the cohomology algebra can be described in terms of Pontryagin and Stiefel–Whitney classes.

With we can define the group homology and group cohomology of a discrete group by

We can refine Question 1.1 to the following:

Question 1.2.

Given family of moduli spaces or discrete groups, how do the homology and cohomology groups of the th space in the sequence change as the parameter increases?

In this article we discuss Question 1.2 with a particular focus on the families of configuration spaces and braid groups. For further reading⁠Footnote1 we recommend R. Cohen’s survey Coh09 on stability of moduli spaces.

1

A version of this note with an extended reference list is available at https://arxiv.org/abs/2201.04096.

1.2. Homological stability

Definition 1.3.

A sequence of spaces or groups with maps

satisfies homological stability if, for each , the induced map in degree- homology

is an isomorphism for all for some stability threshold depending on . The maps are sometimes called stabilization maps and the set is the stable range.

If the maps are inclusions we define to be the stable group or space. Under mild assumptions, if satisfies homological stability, then

We call the groups the stable homology.

2. An Example: Configuration Spaces and the Braid Groups

2.1. A primer on configuration spaces

Definition 2.1.

Let be a topological space, such as a graph or a manifold. The (ordered) configuration space of particles on is the space

topologized as a subspace of . Notably, is a point and .

Configuration spaces have a long history of study in connection to topics as broad-ranging as homotopy groups of spheres and robotic motion planning.

One way to conceptualize the configuration space is as the complement of the union of subspaces of defined by equations of the form .

Figure 1.

The space is obtained by deleting the diagonal from the square .

Graphic for Figure 1.  without alt text

In other words, we can construct by deleting the “fat diagonal” of , consisting of all -tuples in where two or more components coincide. In the simplest case, when and is the interval , we see that consists of two contractible components, as in Figure 1.

Another way we can conceptualize is as the space of embeddings of the discrete set into , appropriately topologized. We may visualize a point in by labelling points in , as in Figure 2.

Figure 2.

A point in the ordered configuration space of an open surface .

Graphic for Figure 2.  without alt text

From this perspective, we may reinterpret the path components of : one component consists of all configurations where particle 1 is to the left of particle 2, and one component has particle 1 on the right. See Figure 3.

Figure 3.

The path components of .

Graphic for Figure 3.  without alt text

Any path through that interchanges the relative positions of the two particles must involve a ‘collision’ of particles, and hence exit the configuration space . We encourage the reader to verify that, in general, the configuration space is the union of contractible path components, indexed by elements of the symmetric group . See Figure 4.

Figure 4.

A point in in the path component indexed by the permutation in .

Graphic for Figure 4.  without alt text    

In contrast, if is a connected manifold of dimension or more, then is path-connected: given any two configurations, we can construct a path through from one configuration to the other without any ‘collisions’ of particles. In this case for all , and this is our first glimpse of stability in these spaces as .

For any space , the symmetric group acts freely on by permuting the coordinates of an -tuple , equivalently, by permuting the labels on a configuration as in Figure 2. The orbit space is the (unordered) configuration space of particles on . This is the space of all -element subsets of , topologized as the quotient of . The reader may verify that the quotient map (illustrated in Figure 5) is a regular -covering space map. In particular, by covering space theory, the quotient map induces an injective map on fundamental groups.

Figure 5.

The quotient map .

\renewcommand{\arraystretch}{1} \setlength{\unitlength}{1.0pt} \begin{tikzpicture} \matrix(m) [matrix of math nodes,row sep=4.5em,column sep=4em,minimum width=2em] { F_n(M) \\ C_n(M) \coloneq F_n(M) / S_n \\}; \path[-stealth] (m-1-1) edge (m-2-1); \end{tikzpicture} Graphic for Figure 5.  without alt text

In the case that is the complex plane , we can identify with the space of monic degree- polynomials over with distinct roots, by mapping a configuration to the polynomial . For this reason the topology of has deep connections to classical problems about finding roots of polynomials.

We will address Question 1.2 for the families and , but we first specialize to the case when . Although the spaces and are path-connected, in contrast to the configuration spaces of , they have rich topological structures: they are classifying spaces for the braid groups and the pure braid groups, respectively, which we now introduce.

2.2. A primer on the braid groups

Since is path-connected, as an abstract group its fundamental group is independent of choice of basepoint. For path-connected spaces, we sometimes drop the basepoint from the notation for .

Definition 2.2.

The fundamental group is called the braid group and is the pure braid group .

We can understand as follows. Choose a basepoint configuration in , and then we may visualize a loop as a ‘movie’ where the particles continuously move around , eventually returning pointwise to their starting positions. If we represent time by a third spacial dimension, as shown in Figure 6, we can view the particles as tracing out a braid. Note that, up to homeomorphism, we may view as the configuration space of the open 2-disk.

Figure 6.

A visualization of a loop in representing an element of .

Graphic for Figure 6.  without alt text

Loops in are similar, with the crucial distinction that the particles are unlabelled and indistinguishable, and so need only return set-wise to their basepoint configuration.

Figure 7.

A braid on 3 strands.

Graphic for Figure 7.  without alt text

It is traditional to represent elements of the group and its subgroup by equivalence classes of braid diagrams, as illustrated in Figure 7. These braid diagrams depict strings (called strands) in Euclidean 3-space, anchored at their tops at distinguished points in a horizontal plane, and anchored at their bottoms at the same points in a parallel plane. The strands may move in space but may not double back or pass through each other. The group operation is concatenation, as in Figure 8.

Figure 8.

The group structure on .

Graphic for Figure 8.  without alt text

The braid groups were defined rigorously by Artin in 1925, but the roots of this notion appeared in the earlier work of Hurwitz, Firckle, and Klein in the 1890s and of Vandermonde in 1771. This topological interpretation of braid groups as the fundamental groups of configuration spaces was formalized in 1962 by Fox and Neuwirth.

Artin established presentations for the braid group and the pure braid group. His presentation for ,

uses generators corresponding to half-twists of adjacent strands, as in Figure 9.

Figure 9.

Artin’s generator for .

Graphic for Figure 9.  without alt text

Artin also gave a finite presentation for . We will not state it in full, but comment that there are generators , (, corresponding to full twists of each pair of strands, as in Figure 10.

Figure 10.

Artin’s generator for .

Graphic for Figure 10.  without alt text

Corresponding to the regular covering space map of Figure 5, there is a short exact sequence of groups

The quotient map , shown in Figure 11, takes a braid, forgets the strands and simply records the permutation induced on their endpoints. The generator maps to the simple transposition . The kernel is those braids that induce the trivial permutation, i.e., the pure braid group.

Figure 11.

The quotient map .

Graphic for Figure 11.  without alt text

2.3. Homological stability for the braid groups

Arnold calculated some homology groups of in low degree (Table 1).

Table 1.

The homology groups . Empty spaces are zero groups. Stable groups are shaded.

0 1 2 3 4 5
0
1
2
3
4
5
6
7
8
9

The column follows from the fact that is path-connected and the column can be obtained by abelianizing Artin’s presentation of . Even the low-degree calculations in Table 1 suggest a pattern: the homology of in a fixed degree becomes independent of as increases.

Arnold proved the following stability result, in terms of the stabilization map defined by adding an unbraided strand as in Figure 12.

Figure 12.

The stabilization map .

Graphic for Figure 12.  without alt text
Theorem 2.3 (Arnold Arn70).

For each , the induced map

is an isomorphism for .

The family therefore satisfies homological stability. Arnold in fact proved the result for cohomology, and Theorem 2.3 follows from the universal coefficients theorem.

May and Segal proved that the stable braid group has the same homology as the path component of the trivial loop in the double loop space . Fuks calculated the cohomology of braid groups with coefficients in . F. Cohen and Vaĭns̆teĭn computed the cohomology ring with coefficients in (for an odd prime), and described in terms of the groups ( for .

2.4. Homological stability for configuration spaces

For a -manifold , it is possible to visualize homology classes in and concretely. Consider Figure 13. This figure shows a -parameter family of configurations in , in fact (because the two loops do not intersect) it shows an embedded torus . Thus, up to sign, this figure represents an element of . In a sense, the loop traced out by particle arises from the homology of the surface , and the loop traced out by particle arises from the homology of . From the homology of and , it is possible to generate lots of examples of homology classes in . The problem of understanding additive relations among these classes, however, is subtle, and the groups are unknown in most cases.

Figure 13.

A class in .

Graphic for Figure 13.  without alt text

When is (punctured) Euclidean space, the (co)homology groups of were computed by Arnold and Cohen. However, even in the case that is a genus- surface, we currently do not know the Betti numbers . Recently Pagaria computed the asymptotic growth rate in of the Betti numbers in the case is a torus. In the case of unordered configuration spaces, in 2016 Drummond-Cole and Knudsen computed the Betti numbers of for a surface of finite type.

Even though the (co)homology groups of configurations spaces remain largely mysterious, the tools of homological stability give us a different approach to understanding their structure.

Theorem Figure 2.3 on stability for braid groups raises the question of whether the unordered configurations spaces satisfy homological stability for a larger class of topological spaces . Let be a connected manifold. To generalize Theorem Figure 2.3 we must define stabilization maps

Unfortunately, in general there is no way to choose a distinct particle continuously in the inputs , and no continuous map of this form exists. To define the stabilization maps, we must assume extra structure on , for example, assume that is the interior of a manifold with nonempty boundary. Then, if we choose a boundary component, it is possible to define the stabilization map by placing the new particle in a sufficiently small collar neighbourhood of the boundary component. This procedure (illustrated in Figure 14) is informally described as ‘adding a particle at infinity.’

Figure 14.

Stabilization map .

Graphic for Figure 14.  without alt text

In the 1970s McDuff proved that the sequence satisfies homological stability and Segal gave explicit stable ranges.

Theorem 2.4 (McDuff McD75; Segal Seg79).

Let be the interior of a compact connected manifold with nonempty boundary. For each the maps

are isomorphisms for .

Concretely, this theorem states that degree- homology classes arise from subconfigurations on at most particles. Heuristically, these homology classes have the form of Figure 15.

Figure 15.

A homology class after stabilizing by the addition of particles.

Graphic for Figure 15.  without alt text

Moreover, McDuff related the homology of the stable space to the homology of , the space of compactly-supported smooth sections of the bundle over obtained by taking the fibrewise one-point compactification of the tangent bundle of .

3. Other Stable Families

We briefly describe some other significant families satisfying (co)homological stability.

Symmetric groups. In Nak60 Nakaoka proved that the symmetric groups satisfy homological stability with respect to the inclusions . The Barratt–Priddy–Quillen theorem states that the infinite symmetric group has the same homology of , the path-component of the identity in the infinite loop space .

General linear groups. Let be a ring. Consider the sequence of general linear groups with the inclusions given by

In the 1970s Quillen studied the homology of these groups when is a finite field of characteristic in his seminal work on the -theory of finite fields. He computes for prime and determines a vanishing range for .

In 1980 Charney proved homological stability when is a Dedekind domain. Van der Kallen, building on work of Maazen, proved the case that is an associative ring satisfying Bass’s “stable rank condition;” this arguably includes any naturally arising ring.

These results are part of a large stability literature on classical groups that warrants its own survey; see the extended version of this article for further references. Homological stability is known to hold for special linear groups, orthogonal groups, unitary groups, and other families of classical groups. There is ongoing work to study (co)homology with twisted coefficients, and sharpen the stable ranges.

Mapping class groups and moduli space of Riemann surfaces. Let be an oriented surface of genus with one boundary component and let the mapping class group

be the group of isotopy classes of diffeomorphisms of fixing a collar neighbourhood of the boundary. There is a map induced by the inclusion by extending a diffeomorphism by the identity on the complement , as in Figure 16.

Figure 16.

The map is induced by the inclusion .

Graphic for Figure 16.  without alt text

There is also a map induced by gluing a disk on the boundary component of . Harer proved Har85 that the sequence satisfies homological stability with respect to the inclusions and that for large the map induces isomorphisms on homology. The proof and the stable ranges have been improved by the work of Ivanov, Boldsen, and others. Madsen and Weiss computed the stable homology by identifying the homology of mapping class groups, in the stable range, with the homology of a certain infinite loop space.

The rational homology of the mapping class group is the same as that of the moduli space of Riemann surfaces of genus . This moduli space parametrizes:

isometry classes of hyperbolic structures on ,

conformal classes of Riemannian metrics on ,

biholomorphism classes of complex structures on the surface ,

isomorphism classes of smooth algebraic curves homeomorphic to .

One consequence of Harer’s stability theorem and the Madsen–Weiss theorem is their proof of Mumford’s conjecture: the rational cohomology of is a polynomial algebra on generators of degree , the so-called Mumford–Morita–Miller classes, in a stable range depending on . See Tillman’s survey Til13.

Homological stability was established for mapping class groups of non-orientable surfaces by Wahl, for mapping class groups of some -manifolds by Hatcher–Wahl and framed, Spin, and Pin mapping class groups by Randal-Williams.

Automorphism groups of free groups. Let denote the free group of rank . Hatcher and Vogtmann proved that the sequence satisfies homological stability with respect to inclusions . Galatius computed the stable homology by proving that . In particular, for ,

Moduli spaces of high-dimensional manifolds. Let be a smooth compact manifold. The moduli space of manifolds of type is the classifying space . In the last few years Galatius and Randal-Williams proved homological stability for for simply connected manifolds of dimension , with respect to the -fold connected sum with . This generalizes Harer’s result to higher-dimensional manifolds. They also obtained a generalized Madsen–Weiss’s theorem for simply connected manifolds of dimension . Homological stability with respect to connected sum with , for , was established by Perlmutter.

4. A Proof Strategy

There is a well-established strategy for proving homological stability that traces back to unpublished work by Quillen in the 1970s. We describe a simplified version of Quillen’s argument for a family of discrete groups with inclusions.

Recall that a -simplex is a -dimensional polytope defined as the convex hull of points in in general position, called its vertices. For example, a -simplex is a point, a -simplex is a closed line segment, and a 2-simplex is triangle. A face of a simplex is the convex hull of a subset of its vertices. A map is simplicial if it maps vertices to vertices, and takes the form

with , …, the vertices of and , .

A triangulation of a topological space is a decomposition of as a union of simplices, such that the intersection of any pair of simplices in is either empty or equal to a single common face of and . A triangulated space is called a simplicial complex. A map of simplicial complexes is simplicial if it maps simplices to simplices and its restriction to each simplex is simplicial.

A simplicial complex is called -connected if it is nonempty, -connected if it is path-connected, and -connected if it is simply connected. More generally, a nonempty simplicial complex is called -connected if its homotopy groups vanish for all . By the Hurewicz theorem, is -connected () if and only if is simply connected and for all .

With this terminology, we can now describe Quillen’s argument. The following formulation of Theorem 4.1 is due to Hatcher–Wahl HW10, Theorem 5.1.

Theorem 4.1 (Quillen’s argument for homological stability).

Let be a sequence of discrete groups. For each let be a simplicial complex with a simplicial action of satisfying the following properties:

(i)

The simplicial complexes are -connected.

(ii)

For each , the group acts transitively on the set of -simplices.

(iii)

For each simplex in , the stabilizer fixes pointwise.

(iv)

The stabilizer of a -simplex is conjugate in to the subgroup . (By convention if .)

(v)

For each edge in , there exists such that and commutes with all elements of that fix pointwise.

Then the sequence is homologically stable. Specifically, the inclusion induces an isomorphism on degree- homology for and a surjection for .

Theorem 4.1 follows from a formal algebraic argument involving a sequence of spectral sequences associated to the complexes . We remark, for the readers familiar with spectral sequences, that for each we obtain a homology spectral sequence by using to build an approximation to from the spaces for . The th spectral sequence has page

and for .

The assumption that the complexes are highly connected implies that the spectral sequence converges to for . The differential

is the map induced by the inclusion . Under the hypotheses of the theorem, we can argue by induction on that this map is an isomorphism (respectively, a surjection) in the desired range, to complete the proof of Theorem 4.1.

In practice, given Theorem 4.1, the most difficult step in a proof of homological stability is usually the proof that the complexes are highly connected.

In recent years, the argument that we just outlined has been axiomatized by Randal-Williams and Wahl RWW17 and Krannich Kra19 to give a very general framework to prove homological stability results, including (co)homology with twisted abelian and polynomial coefficients. Another axiomatization is due to Hepworth.

4.1. An example: the braid group

Let be the closed disk. Fix marked points in its interior and a distinguished point . Associated to the braid group is an -dimensional simplicial complex called the arc complex which we define combinatorially.

vertices: has a vertex for each isotopy class of embedded arcs in joining with one of the marked points.

-simplices: A set of vertices spans a -simplex if the corresponding isotopy classes can be represented by arcs that are pairwise disjoint except at their starting point .

Figure 17.

The action of on a -simplex of the arc complex .

Graphic for Figure 17.  without alt text

Hatcher and Wahl proved that is -connected (though it is in fact contractible).

The braid group is isomorphic to the group of isotopy classes of diffeomorphisms of the closed disk that stabilize the set of marked points and restrict to the identity on . Thus has an action on that is simplicial and satisfies conditions -. See Figure 17. Theorem 4.1 gives a modern proof of homological stability for (Theorem 2.3), a result originally due to Arnold.

5. Representation Stability

5.1. Configuration spaces revisited

Let us address Question 1.2 for the ordered configuration spaces when is the interior of a compact connected manifold with nonempty boundary. As with the unordered configuration spaces, given a choice of boundary component, we can define a stabilization map that continuously introduces a new particle ‘at infinity.’ See Figure 18.

Figure 18.

Stabilization map .

Graphic for Figure 18.  without alt text

This suggests the question: for a fixed manifold , do the spaces satisfy homological stability? The answer is, in contrast to , they do not, as we will verify directly.

Let , so the homology in degree 1 is the abelianization of the pure braid group . Artin’s presentation implies that is free abelian on the images of the generators of Figure 10. Viewed as a homology class in , we can represent by the loop illustrated in Figure 19. Hence, rank grows quadratically in , and homological stability fails.

Figure 19.

The homology class .

Graphic for Figure 19.  without alt text

Church and Farb, however, proposed a new paradigm for stability in spaces like the ordered configuration spaces of a manifold . Because (co)homology is functorial, the -action on induces an action of on the (co)homology groups. Even though the (co)homology does not stabilize as a sequence of abelian groups, they proposed, it does stabilize as a sequence of -representations.

There are several ways to formalize the idea of stability for a sequence of -representations. One way, which was initially the primary focus of Church and Farb, is to consider the multiplicities of irreducible representations in the rational (co)homology groups. Suppose is a finite-dimensional rational -representation. Because is a finite group, is semisimple: it decomposes as a direct sum of irreducible subrepresentations. The multiplicities of the irreducible components are uniquely defined and determine up to isomorphism.

The irreducible rational -representations are classified, and are in canonical bijection with partitions of . A partition of a positive integer is a set of positive integers (called the parts of ) that sum to . It is traditionally encoded by a Young diagram, a collection of boxes arranged into rows of decreasing lengths equal to the parts of . For example, the Young diagram corresponds to the partition of . If is a partition of (equivalently, a Young diagram of size ), we write to denote the irreducible -representation associated to .

Church and Farb observed a pattern in the rational homology of , which we illustrate in Figure 20 in homological degree .

Figure 20.

The decomposition of the homology groups for some small values of .

Graphic for Figure 20.  without alt text

For , we can recover the decomposition of into irreducible components simply by taking the decomposition of and adding a single box to the top row of each Young diagram. They showed that this pattern holds for all , and Church later proved that it holds for the cohomology groups of the ordered configuration space of a connected oriented manifold of finite type.

Church, Farb, and others observed the same patterns in the (co)homology of a number of other families of groups and spaces. These results raise the question,

Question 5.1.

What underlying structure is responsible for these patterns?

Church, Ellenberg, Farb, Nagpal, and Putman answered this question by developing an algebraic framework that brought their work into a broader field, now called the field of representation stability. Other pioneers of the field, who approached it from different perspectives, include Sam, Snowden, Gan, Li, Djament, Pirashvili, and Vespa.

5.2. -modules

The key to answering Question 5.1 is the concept of an -module. The theory of –modules gives a conceptual framework that explains the ubiquity of the patterns observed in so many naturally arising sequences of -representations, and it also provides algebraic machinery to prove stronger results with streamlined arguments.

Definition 5.2.

Let be the category whose objects are finite sets (including ), and whose morphisms are all injective maps. Given a commutative ring (typically or ), an -module over is a functor from to the category of -modules.

To describe an -module , it is enough to consider the “standard” finite sets in ,

For , we write to denote the image of on . The endomorphisms of in are the symmetric group , so is an -representation. The data of an -module is determined by the sequence of -representations , along with -equivariant maps induced by the inclusion . Figure 21 gives a schematic.

Figure 21.

An -module .

Graphic for Figure 21.  without alt text

We refer to (the morphisms of) the category acting on an -module in the same sense that a ring acts on an -module.

We encourage the reader to verify that the following sequences of -representations form -modules.

the trivial –representations, the identity map.

, permutes the standard basis, .

the polynomial algebra with permuting the variables, the inclusion.

Applying any endofunctor of -modules to an -module will produce another -module, so we can construct more examples (say) by taking tensor products or exterior powers of any of the above.

We leave it as an exercise to the reader to verify that the following sequences of -representations do not form an -module. A hint to this exercise: first verify that if fixes the letters , then must act trivially on the image of in under the map induced by the inclusion .

the alternating representation, i.e. , the identity map.

the regular representation, induced by the inclusion .

Importantly for present purposes, the (co)homology groups of ordered configuration spaces form -modules in many cases. If is any space, there is a contravariant action of on its ordered configuration spaces by continuous maps. If we view a point in as an embedding , then an morphism acts by precomposition,

See Figure 22.

Figure 22.

An morphism and its contravariant action on the configuration spaces .

Graphic for Figure 22.  without alt text

Composing this action with the (contravariant) cohomology functor gives a covariant action of on the cohomology groups .

To obtain a covariant action of on , we need additional assumptions on the space . Let be the interior of a compact manifold of dimension at least 2 with nonempty boundary. Consider an morphism and a configuration in . We relabel particles by their image under , and apply the stabilization map of Section 2.4 to introduce any particles not in in a neighbourhood of a distinguished boundary component. See Figure 23.

Figure 23.

An morphism and its covariant action on the configuration spaces .

Graphic for Figure 23.  without alt text

This action of is only functorial up to homotopy, but this suffices to induce a well-defined -module structure on the sequence of homology groups .

Modules over the category behave in many ways like modules over a ring (technically, they are an abelian category). We define a map of -modules to be a natural transformation, that is, a sequence of maps that commute with the morphisms. The kernels and images of these maps themselves form -modules, and we can define operations like tensor products and direct sums in a natural way. This structure allows us to import many of the standard tools from commutative and homological algebra to the study of -modules.

Church, Ellenberg, and Farb showed the answer to Question 5.1 is that the sequences in question are -modules that are finitely generated.

Definition 5.3.

Let be an -module. A subset generates if the images of under the morphisms span for all . Equivalently, the smallest -submodule of containing is itself. The -module is finitely generated in degree if there is a finite subset of elements that generates .

For example, consider the -module over a ring such that is the submodule of homogeneous degree- polynomials in variables, acts by permuting the variables, and is the inclusion map. We encourage the reader to verify that is finitely generated in degree . Figure 24 shows a finite generating set when .

Figure 24.

A finite generating set for the -module .

Graphic for Figure 24.  without alt text

Another example: from our description of the groups in Figure 19, we see that this -module is generated by the single element shown in Figure 25. Arnold’s description of the homology groups of makes it straightforward to verify finite generation of in every degree .

Figure 25.

The homology class generates the -module .

Graphic for Figure 25.  without alt text

Church–Ellenberg–Farb and (independently) Snowden proved that -modules over satisfy a Noetherian property: submodules of finitely generated modules are themselves always finitely generated. Using this result, Church–Ellenberg–Farb proved that, if is a finitely generated -module, then the sequence of -representations stabilizes in several senses.

Theorem 5.4 (Church–Ellenberg–Farb CEF15).

Let be an -module over , finitely generated in degree . The following hold.

Finite generation. For ,

Polynomial growth. There is a polynomial in of degree that agrees with the dimension for all sufficiently large.

Multiplicity stability. For all the decomposition of into irreducible constituents stabilizes (in the sense illustrated in Figure 20).

Character polynomials. The character of is independent of for all .

The characters of are in fact eventually equal to a character polynomial of degree , independent of ; see CEF15, Section 3.3.

The answer of Question 1.2 for the family is then given by the following result.

Theorem 5.5 (Church Chu12; Church–Ellenberg–Farb CEF15; Miller–Wilson MW19).

Let be the interior of a compact connected smooth manifold of dimension at least 2 with nonempty boundary. In each degree the homology and cohomology of ordered configuration spaces of are finitely generated -modules. In particular the degree- (co)homology groups with rational coefficients stabilize in the sense of Theorem 5.4.

Heuristically, Theorem 5.5 states that the homology of is spanned by classes of the form shown in Figure 26.

Figure 26.

A homology class in the image of .

Graphic for Figure 26.  without alt text

From the -covering relationship (Figure 5) it follows that is equal to the multiplicity of the trivial representation in . Hence Theorem 5.5 implies classical cohomological stability with -coefficients for unordered configuration spaces . Church Chu12 used representation stability techniques to prove rational (co)homological stability results for the unordered configuration spaces even in the case that is a closed manifold, so the isomorphisms are not necessarily induced by natural stabilization maps.

5.3. Other instances of representation stability

The definition of a finitely generated -module makes sense for representations over the integers or other coefficients, even in situations where the representations are not semisimple and multiplicity stability is not well-defined. Moreover, this approach readily generalizes to analogous categories that encode actions by families of groups other than the symmetric groups. Some examples that have been studied are the classical Weyl groups, certain wreath products, various linear groups, and products or decorated variants of . The term “representation stability” now refers to algebraic finiteness results (like finite generation or presentation degree) for a module over one of these categories. For further reading on representation stability, see the introductory notes and article Wil18Sno19Sam20.

The (co)homology of several families of groups and moduli spaces exhibit representation stability.

Generalized ordered configuration spaces and pure braid groups. There is a large and growing body of work on representation stability for the homology of configuration spaces: improving stable ranges, studying configuration spaces of broader classes of topological spaces, or studying alternate stabilization maps.

Other families generalizing the pure braid groups also have representation stable cohomology groups, including the pure virtual braid groups, the pure flat braid groups, the pure cactus groups, and the group of pure string motions.

Pure mapping class groups and moduli spaces of surfaces with marked points. Given a set of labelled marked points in a surface , the mapping class group is the group of isotopy classes of (orientation-preserving if is orientable) diffeomorphisms of that fix and stabilize the set of marked points. The pure mapping class group is the subgroup that fixes the marked points pointwise. These groups also generalize the braid groups since and . There is a short exact sequence

that defines an action of on the (co)homology of . Hatcher and Wahl HW10 proved that the sequence satisfies homological stability and Jiménez Rolland JR19 proved that the groups assemble to a finitely generated -module.

For the moduli space of Riemann surfaces of genus with marked points is a rational model of the classifying space , and the symmetric group acts on by permuting the marked points. Hence, the sequence of -representations stabilizes in the sense of Theorem 5.4.

In contrast, for fixed genus the cohomology groups of the Deligne-Mumford compactification of can grow exponentially in . Thus these sequences cannot be finitely generated as -modules. Tosteson Tos21 proved, however, that the sequences are subquotients of finitely generated -modules, where is the opposite category of the category of finite sets and surjective maps. From this he deduced constraints on the growth rate and on the irreducible -representations that occur.

Flag varieties. Let be a semisimple complex Lie group of type , , , or , with Weyl group and a Borel subgroup. The space is called a generalized flag variety. Representation stability of these cohomology groups (as - or -representations) has been studied by Church–Ellenberg–Farb, Wilson, and others.

Complements of arrangements. The cohomology of hyperplane complements associated to certain reflection groups (and their toric and elliptic analogues) stabilizes as a sequence of -representations by the work of Wilson and Bibby. Representation stability holds for the cohomology of more general linear subspace arrangements with a wider class of groups actions by the work of Gadish.

Congruence subgroups. Let be a commutative ring and a proper two-sided ideal. The level congruence subgroups of are defined to be the kernel of the “reduction modulo map . Representation stability of the sequence of homology groups (as or -representations) has been extensively studied; see the extended version of this article for references.

6. Current Research Directions

Work continues on proving (co)homological stability for new families or new coefficients systems, improving stable ranges, and computing the stable and unstable (co)homology for families known to stabilize.

Recently Galatius, Kupers and Randal-Williams GKRW18 identified and proved a new kind of stabilization result, which they describe by the slogan “the failure of homological stability is itself stable”. They defined homological-degree-shifting stabilization maps and use them to prove secondary homological stability for the homology of mapping class groups and general linear groups outside the stable range of (primary) homological stability. Himes studied secondary stability for unordered configuration spaces. Miller–Patzt–Petersen studied stability with polynomial coefficient systems. Miller–Wilson, Bibby–Gadish, Ho, and Wawrykow studied representation-theoretic analogues of secondary stability for ordered configuration spaces.

For a more in-depth introduction to homological stability and these current research directions, we recommend Kupers’ minicourse notes Kup21 and references therein.

Acknowledgments

We thank Omar Antolín Camarena, Jeremy Miller, and Nicholas Wawrykow for useful feedback on a draft of this article. We thank our referees for their extensive comments. Rita Jiménez Rolland is grateful for the financial support by the CONACYT grant Ciencia Frontera CF-2019/217392. Jennifer Wilson is grateful for the support of NSF grant DMS-1906123. The authors are grateful to Benson Farb and Tom Church for introducing them to the field of representation stability, and to many of the ideas in this survey.

References

[Arn70]
V. I. Arnol′d, Certain topological invariants of algebraic functions (Russian), Trudy Moskov. Mat. Obšč. 21 (1970), 27–46. MR0274462Show rawAMSref\bib{ArnoldStability}{article}{ author={Arnol\cprime d, V. I.}, title={Certain topological invariants of algebraic functions}, language={Russian}, journal={Trudy Moskov. Mat. Ob\v {s}\v {c}.}, volume={21}, date={1970}, pages={27--46}, issn={0134-8663}, review={\MR {0274462}}, } Close amsref.
[Chu12]
Thomas Church, Homological stability for configuration spaces of manifolds, Invent. Math. 188 (2012), no. 2, 465–504, DOI 10.1007/s00222-011-0353-4. MR2909770Show rawAMSref\bib{Church}{article}{ author={Church, Thomas}, title={Homological stability for configuration spaces of manifolds}, journal={Invent. Math.}, volume={188}, date={2012}, number={2}, pages={465--504}, issn={0020-9910}, review={\MR {2909770}}, doi={10.1007/s00222-011-0353-4}, } Close amsref.
[CEF15]
Thomas Church, Jordan S. Ellenberg, and Benson Farb, FI-modules and stability for representations of symmetric groups, Duke Math. J. 164 (2015), no. 9, 1833–1910, DOI 10.1215/00127094-3120274. MR3357185Show rawAMSref\bib{CEF}{article}{ author={Church, Thomas}, author={Ellenberg, Jordan S.}, author={Farb, Benson}, title={FI-modules and stability for representations of symmetric groups}, journal={Duke Math. J.}, volume={164}, date={2015}, number={9}, pages={1833--1910}, issn={0012-7094}, review={\MR {3357185}}, doi={10.1215/00127094-3120274}, } Close amsref.
[Coh09]
Ralph L. Cohen, Stability phenomena in the topology of moduli spaces, Geometry of Riemann surfaces and their moduli spaces, Surv. Differ. Geom., vol. 14, Int. Press, Somerville, MA, 2009, pp. 23–56, DOI 10.4310/SDG.2009.v14.n1.a2. MR2655322Show rawAMSref\bib{CohenSurvey}{article}{ author={Cohen, Ralph L.}, title={Stability phenomena in the topology of moduli spaces}, conference={ title={Geometry of Riemann surfaces and their moduli spaces}, }, book={ series={Surv. Differ. Geom.}, volume={14}, publisher={Int. Press, Somerville, MA}, }, date={2009}, pages={23--56}, review={\MR {2655322}}, doi={10.4310/SDG.2009.v14.n1.a2}, } Close amsref.
[GKRW18]
Søren Galatius, Alexander Kupers, and Oscar Randal-Williams, Cellular -algebras, Preprint, arXiv:1805.07184, 2018.Show rawAMSref\bib{GKRW-Ek}{eprint}{ author={Galatius, S{\o {}}ren}, author={Kupers, Alexander}, author={Randal-Williams, Oscar}, title={Cellular ${E}_k$-algebras}, date={2018}, arxiv={1805.07184}, } Close amsref.
[Har85]
John L. Harer, Stability of the homology of the mapping class groups of orientable surfaces, Ann. of Math. (2) 121 (1985), no. 2, 215–249, DOI 10.2307/1971172. MR786348Show rawAMSref\bib{Harer}{article}{ author={Harer, John L.}, title={Stability of the homology of the mapping class groups of orientable surfaces}, journal={Ann. of Math. (2)}, volume={121}, date={1985}, number={2}, pages={215--249}, issn={0003-486X}, review={\MR {786348}}, doi={10.2307/1971172}, } Close amsref.
[HW10]
Allen Hatcher and Nathalie Wahl, Stabilization for mapping class groups of 3-manifolds, Duke Math. J. 155 (2010), no. 2, 205–269, DOI 10.1215/00127094-2010-055. MR2736166Show rawAMSref\bib{HW}{article}{ author={Hatcher, Allen}, author={Wahl, Nathalie}, title={Stabilization for mapping class groups of 3-manifolds}, journal={Duke Math. J.}, volume={155}, date={2010}, number={2}, pages={205--269}, issn={0012-7094}, review={\MR {2736166}}, doi={10.1215/00127094-2010-055}, } Close amsref.
[JR19]
Rita Jiménez Rolland, Linear representation stable bounds for the integral cohomology of pure mapping class groups, Bull. Belg. Math. Soc. Simon Stevin 26 (2019), no. 5, 641–658, DOI 10.36045/bbms/1579402815. MR4053846Show rawAMSref\bib{JR2}{article}{ author={Jim\'{e}nez Rolland, Rita}, title={Linear representation stable bounds for the integral cohomology of pure mapping class groups}, journal={Bull. Belg. Math. Soc. Simon Stevin}, volume={26}, date={2019}, number={5}, pages={641--658}, issn={1370-1444}, review={\MR {4053846}}, doi={10.36045/bbms/1579402815}, } Close amsref.
[Kra19]
Manuel Krannich, Homological stability of topological moduli spaces, Geom. Topol. 23 (2019), no. 5, 2397–2474, DOI 10.2140/gt.2019.23.2397. MR4019896Show rawAMSref\bib{Krannich}{article}{ author={Krannich, Manuel}, title={Homological stability of topological moduli spaces}, journal={Geom. Topol.}, volume={23}, date={2019}, number={5}, pages={2397--2474}, issn={1465-3060}, review={\MR {4019896}}, doi={10.2140/gt.2019.23.2397}, } Close amsref.
[Kup21]
Alexander Kupers, Homological stability minicourse, Lecture notes for eCHT minicourse, https://www.utsc.utoronto.ca/people/kupers/wp-content/uploads/sites/50/homstab.pdf (2021).Show rawAMSref\bib{Kupers}{article}{ author={Kupers, Alexander}, title={Homological stability minicourse}, date={2021}, journal={Lecture notes for eCHT minicourse, \url {https://www.utsc.utoronto.ca/people/kupers/wp-content/uploads/sites/50/homstab.pdf}}, url={https://www.utsc.utoronto.ca/people/kupers/wp-content/uploads/sites/50/homstab.pdf}, } Close amsref.
[McD75]
Dusa McDuff, Configuration spaces of positive and negative particles, Topology 14 (1975), 91–107, DOI 10.1016/0040-9383(75)90038-5. MR358766Show rawAMSref\bib{McDuff}{article}{ author={McDuff, Dusa}, title={Configuration spaces of positive and negative particles}, journal={Topology}, volume={14}, date={1975}, pages={91--107}, issn={0040-9383}, review={\MR {358766}}, doi={10.1016/0040-9383(75)90038-5}, } Close amsref.
[MW19]
Jeremy Miller and Jennifer C. H. Wilson, Higher-order representation stability and ordered configuration spaces of manifolds, Geom. Topol. 23 (2019), no. 5, 2519–2591, DOI 10.2140/gt.2019.23.2519. MR4019898Show rawAMSref\bib{MW}{article}{ author={Miller, Jeremy}, author={Wilson, Jennifer C. H.}, title={Higher-order representation stability and ordered configuration spaces of manifolds}, journal={Geom. Topol.}, volume={23}, date={2019}, number={5}, pages={2519--2591}, issn={1465-3060}, review={\MR {4019898}}, doi={10.2140/gt.2019.23.2519}, } Close amsref.
[Nak60]
Minoru Nakaoka, Decomposition theorem for homology groups of symmetric groups, Ann. of Math. (2) 71 (1960), 16–42, DOI 10.2307/1969878. MR112134Show rawAMSref\bib{Nakaoka}{article}{ author={Nakaoka, Minoru}, title={Decomposition theorem for homology groups of symmetric groups}, journal={Ann. of Math. (2)}, volume={71}, date={1960}, pages={16--42}, issn={0003-486X}, review={\MR {112134}}, doi={10.2307/1969878}, } Close amsref.
[RWW17]
Oscar Randal-Williams and Nathalie Wahl, Homological stability for automorphism groups, Adv. Math. 318 (2017), 534–626, DOI 10.1016/j.aim.2017.07.022. MR3689750Show rawAMSref\bib{RWW}{article}{ author={Randal-Williams, Oscar}, author={Wahl, Nathalie}, title={Homological stability for automorphism groups}, journal={Adv. Math.}, volume={318}, date={2017}, pages={534--626}, issn={0001-8708}, review={\MR {3689750}}, doi={10.1016/j.aim.2017.07.022}, } Close amsref.
[Sam20]
Steven V. Sam, Structures in representation stability, Notices Amer. Math. Soc. 67 (2020), no. 1, 38–43. MR3970038Show rawAMSref\bib{Sam-Notices}{article}{ author={Sam, Steven V.}, title={Structures in representation stability}, journal={Notices Amer. Math. Soc.}, volume={67}, date={2020}, number={1}, pages={38--43}, issn={0002-9920}, review={\MR {3970038}}, } Close amsref.
[Seg79]
Graeme Segal, The topology of spaces of rational functions, Acta Math. 143 (1979), no. 1-2, 39–72, DOI 10.1007/BF02392088. MR533892Show rawAMSref\bib{SegalRational}{article}{ author={Segal, Graeme}, title={The topology of spaces of rational functions}, journal={Acta Math.}, volume={143}, date={1979}, number={1-2}, pages={39--72}, issn={0001-5962}, review={\MR {533892}}, doi={10.1007/BF02392088}, } Close amsref.
[Sno19]
Andrew Snowden, Algebraic structures in representation stability, MSRI graduate school lecture notes, http://www-personal.umich.edu/~asnowden/msri19/course.pdf (2019).Show rawAMSref\bib{Snowden-FINotes}{article}{ author={Snowden, Andrew}, title={Algebraic structures in representation stability}, date={2019}, journal={MSRI graduate school lecture notes, \url {http://www-personal.umich.edu/~asnowden/msri19/course.pdf}}, } Close amsref.
[Til13]
Ulrike Tillmann, Mumford’s conjecture—a topological outlook, Handbook of moduli. Vol. III, Adv. Lect. Math. (ALM), vol. 26, Int. Press, Somerville, MA, 2013, pp. 399–429. MR3135441Show rawAMSref\bib{Tillman-MumfordSurvey}{article}{ author={Tillmann, Ulrike}, title={Mumford's conjecture---a topological outlook}, conference={ title={Handbook of moduli. Vol. III}, }, book={ series={Adv. Lect. Math. (ALM)}, volume={26}, publisher={Int. Press, Somerville, MA}, }, date={2013}, pages={399--429}, review={\MR {3135441}}, } Close amsref.
[Tos21]
Philip Tosteson, Stability in the homology of Deligne–Mumford compactifications, Compos. Math. 157 (2021), no. 12, 2635–2656, DOI 10.1112/s0010437x21007582. MR4354696Show rawAMSref\bib{Tosteson}{article}{ author={Tosteson, Philip}, title={Stability in the homology of Deligne--Mumford compactifications}, journal={Compos. Math.}, volume={157}, date={2021}, number={12}, pages={2635--2656}, issn={0010-437X}, review={\MR {4354696}}, doi={10.1112/s0010437x21007582}, } Close amsref.
[Wil18]
Jennifer C. H. Wilson, An Introduction to -modules and their generalizations, Summer school lecture notes, http://www.math.lsa.umich.edu/~jchw/FILectures.pdf (2018).Show rawAMSref\bib{Wilson-FINotes}{article}{ author={Wilson, Jennifer C.~H.}, title={An {I}ntroduction to $\mathsf {FI}$-modules and their generalizations}, date={2018}, journal={Summer school lecture notes, \url {http://www.math.lsa.umich.edu/~jchw/FILectures.pdf}}, } Close amsref.

Credits

Opening image is courtesy of shulz via Getty.

Figures 1–26 are courtesy of Jennifer C. H. Wilson.

Photo of Rita Jiménez Rolland is courtesy of Adrián Lozano.

Photo of Jennifer C. H. Wilson is courtesy of Alexander Wright.