Flat Surfaces and Stability Structures
Flat Surfaces and Stability Structures
CONTENTS
1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 247
1.1. Spaces of stability structures and flat surfaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 247
1.2. Fukaya categories of surfaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 248
1.3. Finite vs. infinite area, finite length property . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 249
1.4. Meromorphic connections on T∗ C and Nevanlinna’s Kernpolygon . . . . . . . . . . . . . . . . . . . . . . . 249
1.5. Comparison with the work of Bridgeland–Smith . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 250
2. Moduli spaces of flat surfaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 251
2.1. Grading of surfaces and curves . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 251
2.2. Flat surfaces with conical singularities . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 253
2.3. Saddle connections . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 256
2.4. Horizontal strip decompositions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 259
2.5. Complex-analytic point of view . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 260
2.6. Voronoi and Delaunay partitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 267
3. The topological Fukaya category of a surface . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 268
3.1. Preliminaries on A∞ -categories . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 268
3.2. Surfaces with marked boundary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 271
3.3. Minimal A∞ -category of an arc system . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 272
3.4. Formal generators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 276
3.5. Localization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 278
3.6. Cosheaf of categories . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 280
4. Tameness and geometricity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 283
4.1. Twisted complexes from curves . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 283
4.2. Representations of nets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 286
4.3. Minimal twisted complexes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 293
4.4. Classification of objects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 294
5. Comparison of moduli spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 297
5.1. Stability structures on categories . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 297
5.2. Charge lattice . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 298
5.3. Statement of main theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 301
5.4. Support property . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 302
5.5. Harder–Narasimhan filtrations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 303
5.6. The map M(X) → Stab(F (X)) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 306
6. Cluster-like structures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 308
6.1. S-graphs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 308
6.2. Example: Dynkin type An . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 311
6.3. Example: affine type A n . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 313
7. Open problems and further directions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 315
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 315
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 316
1. Introduction
DOI 10.1007/s10240-017-0095-y
248 F. HAIDEN, L. KATZARKOV, M. KONTSEVICH
map Z : K0 (C ) → C, the central charge. These have to satisfy a number of axioms, see
Section 5.1. Although the definition seems somewhat strange at first from a mathemat-
ical point of view, it leads to the remarkable fact, proven by Bridgeland, that Stab(C )
naturally has the structure of a complex manifold, possibly infinite-dimensional. Further-
more, Stab(C ) comes with an action of Aut(C ) and the universal cover of GL+ (2, R).
Thanks to the efforts of a number of people, the structure of Stab(C ) is understood in
many particular cases, see e.g. [7, 12, 13, 15, 29, 34].
Moduli spaces of very similar nature appear in the theory of complex curves
with a quadratic differential, also known as half-translation surfaces or simply flat surfaces,
as they carry a flat Riemannian metric with conical singularities. Each of these moduli
spaces, M(S), like Stab(C ), comes with a wall-and-chamber structure and an action of
GL+ (2, R). An explanation for this similarity was proposed by the third named author
and Soibelman in [31]: The spaces M(S) should be realized as moduli spaces of stability
structures for a suitably defined Fukaya category of the surface S. The main achievement
of the present paper is to make precise and verify this claim, leading to the following
result (Theorem 5.3 in the main text).
Theorem. — Let S be a marked surface of finite type, M(S) the space of marked flat structures
on S, and F (S) the Fukaya category of S. The there is a natural map
(1.1) M(S) → Stab F (S)
which is bianalytic onto its image, which is a union of connected components.
For now it remains an open question whether the image is in fact all of Stab(F (S)).
The method of proof is new and relies on a detailed understanding of the category F (S).
Theorem. — Isomorphism classes of objects in F (S) are classified by isotopy classes of certain
immersed curves on S together with local system of vector spaces on them. In particular, these categories
are of tame representation type.
FLAT SURFACES AND STABILITY STRUCTURES 249
Another way to say this is that all objects are “geometric”. This is not just a direct
consequence of the definition, since F (S) is required to be closed under taking cones.
We reduce the problem to the classification of certain pairs of filtrations on vector spaces
with subquotients identified.
An important tool, e.g. for proving the equivalence with other definitions and com-
puting K-theory, are formal generators of F (S) with endomorphism algebras given by
graded quivers with quadratic monomial relations. These exist only for some S, but the
other cases are obtained by localization of F (S). They are graded versions of the gentle
algebras introduced by Assem–Skowroński [3].
where f and g are meromorphic and f has a pole of order one greater than D. We
call this an exponential singularity of a quadratic differential. In terms of flat geometry, this
singularity gives rise to n infinite-angle singularities of the metric, where n is the order
of the pole of f . Such a singularity is modeled on the metric completion of the universal
cover of the punctured Euclidean plane, and can be regarded as a limit of the conical
singularity when the cone angle goes to infinity. We show the following:
Theorem. — Let S be a half-translation surface of finite type with conical singularities, possi-
bly infinite-angle ones, then S comes from a compact Riemann surface with quadratic differential with
exponential singularities.
See Theorem 2.3 in the main text. Thus allowing exponential singularities is quite
natural from a geometric point of view, and turns out to be necessary to get stability
structures on partially wrapped Fukaya categories of surfaces.
The previous theorem is similar in nature and has some overlap with an old result
of Nevanlinna [38] on the class of functions which have Schwarzian derivative a polyno-
mial. Indeed, ideas of his are found in our proof. The analysis of the topological structure
of the branched covering defined by these functions leads him to a certain combinatorial
structure, the Kernpolygon, which translates into the t-structure of the stability structure for
a quadratic differential exp(P(z))dz2 .
order pole. Some quadratic differentials with simple poles do appear in the limit. It is
expected that their results extend in some form to holomorphic quadratic differentials
with simple zeros, but the stability structures are not described by quivers with potential
and would require different methods to construct. In this paper we do indeed construct
stability structures from these quadratic differentials, though not on CY3 categories.
It is somewhat puzzling that the same geometry should give rise to stability struc-
tures on, at least superficially, very different categories. The relation between the two
categories (the wrapped Fukaya category of the surface and the CY3 category) is still to
be clarified. It seems to be simpler in the case when all poles have order ≥ 3. One can
show that there is a non-commutative divisor in the sense of Seidel [42] with total space
the wrapped category and central fiber the CY3 category. We suspect that in the case of
double poles an additional deformation is necessary.
More recently, an extension of the results of Bridgeland–Smith to certain poly-
nomial quadratic differentials on C was carried out in [14, 27]. There, CYn categories
appear for n ≥ 2.
In this section we define the moduli spaces M(X) of marked flat structures. Sec-
tion 2.1 discusses some topological preliminaries which are mostly relevant later when
we define Fukaya categories of surfaces, but also play a small role in the definition of
M(X). Next, in Section 2.2 we define flat surfaces with conical singularities and their
moduli spaces and state a theorem about their local structure. This theorem is proven
in the case of finite area in Section 2.3, using triangulations by saddle connections. Sec-
tion 2.4 deals with the case of infinite area, where the main tool is the horizontal strip
decomposition. In Section 2.5 we show that our class of flat surfaces may be described in
terms of quadratic differentials on compact Riemann surfaces. Finally, in Section 2.6 we
review the dual Voronoi and Delaunay partitions.
where α · β denotes concatenation of paths. Every graded surface X has a shift automor-
phism given by the pair (idX , σ ) where σ restricts, for every x ∈ X, to the generator of
π1 (P(Tx X)) given by the orientation of X, i.e. the path which rotates a line counterclock-
wise by an angle of π . Anticipating the connection with triangulated categories, we write
this automorphism as [1] and its integer powers as [n].
For an oriented surface X observe that the set
(2.2) G := π0 X, P(TX)
The above isomorphism simply takes a foliation η and counts how many times it rotates
along the loop {0} × S1 ⊂ X. Call this number deg(η), then the map (x, θ ) → (−x, θ )
from the cylinder to itself acts on the foliation by switching the sign of deg(η). From this
one can see that |deg(η)| ∈ Z≥0 is a complete invariant of the graded surface.
We have the following homotopy classes of paths in P(Tp X): (1) c̃1 (t1 ) from η(p) to ċ1 (t1 ),
(2) c̃2 (t2 ) from η(p) to ċ2 (t2 ), (3) κ from ċ1 (t1 ) to ċ2 (t2 ) given by counterclockwise rotation
in Tp X by an angle < π . Define the intersection index of c1 , c2 at p
(2.5) ip (c1 , c2 ) = c̃1 (t1 ) · κ · c̃2 (t2 )−1 ∈ π1 P(Tp X) ∼= Z.
FLAT SURFACES AND STABILITY STRUCTURES 253
More precisely, this notation is correct only if c1 , c2 pass through p exactly once, in par-
ticular if ci are in general position, otherwise the index may depend on the ti as well. We
note that
and
(2.7) ip c1 [m], c2 [n] = ip (c1 , c2 ) + m − n.
Another observation which will be used repeatedly, is that if graded curves c1 and c2
intersect in p with ip (c1 , c2 ) = 1, then we may perform a kind of smoothing near p which
produces again a graded curve (see Figure 1). (If the intersection index is not 1 then the
grading does not extend continuously along the modified curves.)
A graded surface (X, η) has a canonical double cover τ of X with fiber over p ∈ X
the set of orientations of η(p). We will consider τ as a locally constant sheaf and write
(2.8) Zτ := Z ⊗Z/2 τ
and similarly for Rτ and Cτ . Singular (co)homology with coefficients in any of these local
systems is defined, as is deRham cohomology with coefficients in Rτ or Cτ . Integration
provides a pairing
where H∗dR (M, ∂M; E) is the cohomology of forms with coefficients in E which vanish
when pulled back to ∂M.
Definition. — Let Y be a smooth surface with boundary. The structure of a real blow-up
of a flat surface on Y is given by the structure of a smooth flat surface on Y \ ∂Y such that each
component of ∂Y has a neighborhood diffeomorphic to a neighborhood of the boundary of some C n or C
∞
by a diffeomorphism preserving the flat structure on the interior, and satisfying the following completeness
condition: The space X obtained by identifying points lying in the same component of ∂Y is the metric
completion of Y \ ∂Y.
Definition. — A flat surface is a complete metric space together with a partition X = Xsm ∪ Xsg
into smooth and singular points and the structure of a smooth flat surface on Xsm (i.e. direction of
horizontal foliation) such that near each x ∈ Xsg the space is locally isometric to one of Cn or C∞ .
Note that in the above definition x ∈ Xsg is not necessarily a singular point of the
metric if near x the model is C2 . One can consider it as an arbitrary marked point in the
metrically smooth part. As the terminology suggests, the real blow-up of a flat surface has
the structure of a real blow-up of a flat surface. Conversely, one obtains a flat surface from the
real blow-up by taking the metric completion of the interior. (Topologically, this contracts
each boundary component to a point.)
Definition. — Fixing a graded surface S we consider the moduli space of marked flat sur-
faces M(S). An element of M(S) is represented by a real blow-up of a flat surface Y with underlying
oriented surface S together with a morphism of graded surfaces f : S → Y with underlying map the
identity on S. Two such pairs (Y, f ), (Z, g) are equivalent if there is a diffeomorphism h : Y → Z
preserving the flat structure so that h ◦ f and g are isotopic as maps of graded surfaces. We give flat
structures (before identification) the topology of pointwise convergence on the real blow-up, and M(S) the
quotient topology.
Note that the grading on S fixes the number and types of singular points and
marked points of any flat surfaces in M(S). Explicit descriptions of M(S) when S is a
disk or annulus are found in Section 6.
If Y is the real blow-up of a flat surface then the square-roots of the quadratic
differential on Y \ ∂Y may be viewed as a closed 1-form α with values in Cτ , and as such
it extends to the boundary. In particular we get a class [α] ∈ H1dR (Y, ∂Y; Cτ ) so using the
pairing (2.9) we get a homomorphism H1 (Y, ∂Y; Zτ ) → C which represents the periods
256 F. HAIDEN, L. KATZARKOV, M. KONTSEVICH
of the flat surface. We say that a graded surface Y is of finite type if Y and ∂Y have a finite
number of connected components and H1 (Y, ∂Y; Zτ ) is finitely generated.
The strategy of the proof will depend on whether the area of the flat surface is
finite or infinite and will be completed in the following two subsections.
Proposition 2.1. — Let X be a flat surface of finite type, then the size of any geodesic arc system
is bounded above by a constant depending only on the topology of the real blow-up Y = X.
Proof. — Let A be a finite geodesic arc system on X and E = |A| its cardinality.
Each saddle connection in A together with an arbitrary choice of orientation gives a non-
zero class in H1 (X, Xsg , Z). A complete set of relations between these classes is given by
the polygons cut out by A. Let F be the number of these polygons. Any bi-gon cut out
by A must contain a conical point with cone angle π in its interior, so their number, D,
is bounded by the number of compact boundary components of Y. Assigning boundary
edges to polygons we find that
2
(2.17) F − D ≤ E.
3
By assumption, B = rkH1 (X, Xsg , Z) is finite, and since E − F ≤ B we get
(2.18) E ≤ 3(D + B)
Proposition 2.2. — Suppose A is a maximal geodesic arc system on a flat surfaces X. Let K be
the union of Xsg , the saddle connections in A, and the triangles cut out by A. Then K is the convex hull
of Xsg , thus independent of A.
Proof. — From the local geometry above we see that any geodesic which leaves K
stays in X \ K for all subsequent times, so K is convex. It contains Xsg by definition. If K
is convex and contains Xsg , then it must also contain all saddle connections, in particular
those in A, and thus all triangles cut out by A, so K ⊆ K .
We will refer to the convex hull of Xsg as the core of the flat surface X, denoted
Core(X). The previous proposition shows that, for X of finite type, it is covered by a finite
number of singular points, saddle connections, and geodesic triangles, thus compact.
258 F. HAIDEN, L. KATZARKOV, M. KONTSEVICH
Proposition 2.3. — Let X be a flat surface with the property that each component contains at
least one singular point, and let K be the core of X. Then K is a deformation retract of X.
Proof. — The assumption ensures that each p ∈ X has finite distance to Core(X).
A deformation retraction can be defined by moving p ∈ X along a geodesic path to the
unique point in Core(X) which is closest to p.
Let X be a connected flat surface with finite area and assume that X is not smooth,
which excludes only the case of the torus. A maximal geodesic arc system on X gives a tri-
angulation with some triangles possibly degenerate. By definition, a triangle is degenerate
if two of its edges coincide, so that they meet in a conical singularity with cone angle π .
Using triangulations by saddle connections, we can now give a proof of Theorem 2.1 in
the case when the flat surface has finite area. This result is classical, see e.g. Veech [50],
but we include a proof here for completeness. The proof in the case of infinite area (but
still finite type) will be given in the next section.
Proof of Theorem 2.1, finite area case. — Fix a graded surface X of finite type admitting
flat structures with finite total area. This means that X is compact as a surface with
boundary. We also assume that there is at least one singularity, excluding the case of
the torus which is easily handled directly. Let = H1 (X, ∂X; Zτ ) and : M(X) →
Hom(, C) be the period map. We want to show that is a local homeomorphism.
Let a0 ∈ M(X) and equip X with this flat structure. Pick a maximal collection
A of geodesic arcs on X. These arcs give a triangulation of X, since the flat metric has
finite area by assumption. Fix a grading on each arc in A, then each α ∈ A gives a class
[α] ∈ . A complete set of relations between these classes comes from the list of triangles.
The number (a0 )([α]) ∈ C records the length and slope of an arc α ∈ A.
Since the number of triangles is finite, there is a contractible neighborhood U of
(a0 ) ∈ Hom(, C) with the following property: For any b ∈ U and any three classes
α1 , α2 , α3 in corresponding to the edges of a triangle of A, the vectors (b)([αi ]) form
the edges of a non-degenerate triangle in C. This ensures that for each b ∈ U we get
an actual flat structure on X essentially by direct construction. Namely, we cut X into its
geodesic triangles, deform each of the triangles according to the values (b)([α]), α ∈ A,
then glue the triangles back together in the same combinatorial manner. In this way, we
have defined a continuous section of over U.
To complete the proof, it remains to be shown that (U) is an open neighborhood
of a0 . Note that there is a neighborhood W of a0 ∈ M(X) such that all saddle connections
in A persist throughout W, i.e. do not break along a singularity. Making W sufficiently
small, we can assume that W ⊂ −1 (U). Then the same collection of arcs A, up to
homotopy, again form a maximal collection of geodesic arcs for any flat surface in W.
This shows that W ⊂ (U), so (U) is a neighborhood of a0 , in fact open by the same
arguments.
FLAT SURFACES AND STABILITY STRUCTURES 259
Proposition 2.4. — Let X be flat surface of finite type with infinite area. Assume that X is
connected and has at least one conical point. Then, after possibly rotating the horizontal direction, the
leaves converging towards a conical point cut the surface into horizontal strips as above.
Proof. — Assume that no horizontal leaf is a saddle connection. This can always
be achieved by rotating the horizontal direction, as there are at most countably many
slopes of saddle connections. In fact, the set of slopes of saddle connections is closed
under the present assumptions, but this is not needed. Note that no leaf can be a closed
geodesic either, as a cylinder foliated by closed geodesics is bounded by one or more
saddle connections. Arguments of Strebel [47] show that there are then only two types of
leaves. Generic leaves which are closed and intersect the core of X in a compact interval, and
Critical leaves which converge towards a conical point in one direction and eventually leave
the core in the other. Indeed, the boundary of the closure of a leaf remaining entirely in
the core would be a union of leaves which are saddle connections, which is impossible. In
particular, the closure of a critical leaf adds only a single conical point. The components
of the union of all generic leaves are open parts of horizontal strips. The exponential maps
at the various conical points give identifications of the closures of these open components
with the standard horizontal strips of finite or infinite height.
Let us say a bit more about the horizontal strip decompositions appearing in the
previous proposition. A conical point with finite cone angle lies on finitely many hori-
zontal strips, while a conical point with infinite cone angle lies on infinitely many, all but
finitely many of which will have infinite height. Each horizontal strip of finite height con-
tains a unique saddle connection: the straight line segment connecting 0 and z. These
are exactly the saddle connections α1 , . . . , αm , which do not intersect any of the critical
leaves. We will refer to them as simple saddle connections. Following the leaves and collapsing
the horizontal strips of infinite height we construct a deformation retraction of the sur-
face to the union of the αi and the conical points. This shows that the αi give a basis of
H1 (X, ∂X; Zτ ), after choosing gradings.
The graph formed by the αi is identified with part of the Hausdorff version of the
leaf space, L, of the horizontal foliation. We may realize L inside X by adding the positive
260 F. HAIDEN, L. KATZARKOV, M. KONTSEVICH
Proof of Theorem 2.1, infinite area case. — Let a0 ∈ M(X) be a flat structure and assume
first that there are no leaves which are saddle connections. By the considerations above,
the critical leaves cut X into horizontal strips. Let α1 , . . . , αm be the saddle connections
corresponding to the horizontal strips of finite height. Also choose the grading on each
αi so that Im (a0 )(αi ) > 0. This is possible since αi is not horizontal and shifting the
grading flips the sign of (a0 )(αi ). We get a basis [α1 ], . . . , [αm ] of = H1 (X, ∂X; Zτ ).
Consider the open subset U ⊂ Hom(, C) of maps sending each [αi ] to the upper half
plane. For each b ∈ U we construct a flat surface (b) with the same combinatorial type
of horizontal strip decomposition as a0 , but different slopes and lengths of the αi given
by b. The map defines a section of M(X) → Hom(, C) over U. To see that (U) is
open, note that any flat surface sufficiently close to one in (U) has the same horizontal
strip decomposition.
If a0 has horizontal saddle connections, we argue as follows. By assumption = 0,
so R ⊂ GL + (2, R) acts on M(X) by rotation of the horizontal direction and each orbit
contains a flat structure with only critical and generic leaves considered before. As
is equivariant with respect to the action it must be a local homeomorphism near a0 as
well.
coordinate z with
√ a
(2.20) ϕ = z−n/2 + dz
z
√
when n ≥ 4 and ϕ = za dz when n = 2. In any case, we note that the flat metric near a
higher order pole has infinite area and is complete, i.e. higher order poles do not lead to
any additional singularities of the metric.
Theorem 2.2. — Let C be a compact Riemann surface with non-zero meromorphic quadratic
differential ϕ with set of zeros and poles D. Then |ϕ| gives C \ D the structure of the smooth part of a
flat surface of finite type with conical points which have cone angle an integer multiple of π . Conversely,
any flat surface of finite type without infinite-angle conical points is obtained in such a way.
Remark. — Let Tg,n be the Teichmüller space of compact Riemann surfaces of genus
g with n marked points. The cotangent space to Tg,n at C is naturally identified with the
space of quadratic differentials on C with at most simple poles at the marked points.
Thus every point in the cotangent bundle of Tg,n corresponds to some flat surface of
genus g with at most n conical points with cone angle π (and other conical points with
bigger cone angle). The moduli spaces M(S) defined above correspond to strata in the
cotangent bundle. The ribbon graph attached to a Strebel differential is a special case of
the core of a flat surface introduced here.
Proof. — It remains to show the second part, so let X be a flat surface of finite
type without infinite-angle conical points. First of all, we have a complex structure and
non-vanishing holomorphic quadratic differential on the smooth part Xsm . We know that
near a conical point the flat structure comes from a zero or first order pole of a quadratic
differential, so the complex structure extends to all of X and ϕ extends meromorphically.
If X has finite area, then it is compact and so C = X, D = Xsg , and we are done.
For the case of infinite area we need to find a compact K ⊂ X such that the com-
ponents of the complement X \ K are isometric to some punctured neighborhoods of
higher order poles. As a first step, take K to be the core of X. The boundary of K in X is
a sequence of saddle connections meeting in exterior angles φp ≥ π , by convexity of K.
The condition on the global monodromy of the metric (existence of horizontal foliation)
implies that
(2.21) (φp − π ) = nπ
p
where p runs over the corners of a fixed component of the boundary of K, and n ∈ Z≥0
depends on that component.
Points in the complement of K are parametrized by pairs (l, d) where l is a straight
line starting at some point in the boundary of K, meeting K only in the one endpoint,
262 F. HAIDEN, L. KATZARKOV, M. KONTSEVICH
Exponential-type singularities
An exponential-type singularity is a transcendental singularity of a quadratic differential
of the form
then the completion of D with respect to |ϕ| has n additional points, all of which are ∞-angle singular-
ities.
FLAT SURFACES AND STABILITY STRUCTURES 263
Proof. — 1. Define η(z) = (max{1, |Im(z)|})λ e−Re(z) , then for sufficiently small
ε > 0, the boundary of {η ≥ ε} in A is given by the curve
(2.27) Re(z) = λ log max 1, |Im(z)| − log(ε).
Moreover, for distinct values of ε, these curves have positive distance with respect to the
Euclidean metric.
2. We claim that e−Re(z) h(z) is bounded below on {η ≥ ε} for any ε > 0. By assump-
tion,
Case λ < 0:
λ
(2.32) C1 |z|λ e−Re(z) ≥ C1 |Re(z)|e−Re(z)/λ + |Im(z)|e−Re(z)/λ
The first term is bounded above, as Re(z) is bounded above by − log(ε), the second term
is bounded above by ε1/λ .
3. We claim that for every ε > 0, the set {η ≥ ε} is complete with respect to the
metric g = (e−Re(z) h(z))2 |dz|2 , thus any Cauchy sequence zj ∈ A without limit must sat-
isfy η(zj ) → 0. Namely, we have g ≥ Cgeucl on {η ≥ ε/2} be the previous step, and the
boundary curves {η = ε}, {η = ε/2} have some positive distance δ, so
for z1 , z2 ∈ {η ≥ ε}. Hence, any sequence which is Cauchy with respect to g is Cauchy for
the standard metric.
264 F. HAIDEN, L. KATZARKOV, M. KONTSEVICH
Case λ > 0:
∞ λ
(2.36) I≤ |Re(z) + t|e−(Re(z)+t)/λ + |Im(z)|e−(Re(z)+t)/λ dt
0
As η(z) ≤ ε, Re(z) ≥ − log(ε) > 0, the first summand is bounded above by (− log(ε) +
t)ε1/λ e−t/λ , and the second by ε1/λ e−t/λ , so
∞
λ
(2.37) I≤ε 1 − log(ε) + t e−t dt
0
(2.38) ≤ Cεe(1−log(ε))/2
√
(2.39) = C eε
(2.41) = η(z) ≤ ε
To prove the claim, it remains to show that for different values of z1 , z2 , the corresponding
horizontal curves starting at z1 , z2 have vanishing distance. Let β(t) = x + it, t ∈ [y1 , y2 ],
x > ρ, then
y2
(2.42) l(β) = e−x h(x + it)dt
y1
y2
−x x→∞
(2.43) ≤ C2 e |x + it|λ dt −−→ 0
y1
Proof of Proposition 2.5. — After a change of coordinates w = eπi/n /z, ϕ is of the form
ϕ = e−w h(w)dw2
n
(2.44)
FLAT SURFACES AND STABILITY STRUCTURES 265
n−1
(2.46) A=B\ Vn .
k=0
|ϕ| = e−r
n cos(nφ)
(2.47) |hdz2 | ≥ C|dz|2
for some C > 0. Since the sets Vk have positive distance with respect to the standard
metric, this shows that every Cauchy sequence in B without limit is eventually contained
in one of the Vk .
On Vk we perform a change of coordinates u = wn , so
(2.50) B̄ =: B ∪ {a1 , . . . , an }
is a complete metric space. Let a = ak , the we must show that a is an ∞-angle singularity
of B̄.
As Vk is simply connected, we may choose a holomorphic f : Vk → C with
(df )2 = ϕ. Choose r > 0 such that D∗2r (0) ⊂ Vk . Then for any path α in D∗2r with endpoint
z1 , z2 ∈ D∗r (0) we compute
z2 √
(2.51) f (z2 ) − f (z1 ) = ϕ ≤ |ϕ|
z1 α
hence
Next we prove a kind of converse to the previous proposition, which is the main
result of this subsection.
Theorem 2.3. — Let X be a flat surface of finite type. Then there is a compact Riemann surface
C with quadratic differential ϕ of exponential type giving C \ D(ϕ) the structure of the smooth part of
a flat surface isomorphic to Xsm .
Proof. — Let K := Core(X), which we may assume to contain more than one point.
Define a boundary walk to be a piecewise geodesic path which follows the boundary of K
so that S \ K lies to the right (and possibly also to the left), takes the rightmost possible
direction at every singularity, and is maximal with these properties. By rightmost direction
we mean the following. From each singularity there is a finite set E of directions which lie
in ∂K, and they are either cyclically or totally ordered. So the direction from which we
approach the singularity has some successor, by definition the rightmost direction, or is a
maximal element, and the boundary walk ends. Hence a boundary walk is either closed
or starts and ends at infinite-angle singularities, and there is a unique boundary walk
starting/ending at each infinite-angle singularity. If the boundary walk starts and ends at
an infinite-angle singularity then t cuts X into two pieces, one of which is topologically a
disk lying completely outside the core K. This follows from the description of X in terms
of the core above.
Let si ∈ Xsg , i ∈ Z/k be a cyclic sequence of infinite-angle singularities so that there
is a boundary path from si to si+1 .
Choose a closed embedded curve α on Xsm which is a smoothing of the cyclic
sequence of boundary paths and cuts X into two pieces X , X so that X contains the si
and no other singularities. The smooth part of X is topologically an annulus. It consists,
up to modifications coming from the smoothing, of several pieces of the complement of
K which are topological disks. The smoothing cases these disks to be glued together at
infinite-angle singularities to form X . To complete the proof, it suffices to show that X
corresponds to some punctured neighborhood of an exponential singularity.
Fix some direction ri from the singularity si . For each integer n > 0 consider the
pair of geodesics Li , Ri which start at si in the directions ri + nπ/2 and ri − nπ/2 respec-
FLAT SURFACES AND STABILITY STRUCTURES 267
By the previous results about flat surfaces without infinite-angle singularities, there
is a biholomorphic fn : Xn → D∗ and meromorphic quadratic differential ϕn on D with a
pole at 0 and k zeros, such that fn becomes an isometry of flat surfaces. Since D ⊂ C is
bounded, the fn form a normal family and, after passing to a subsequence, converge to
a holomorphic f : Xsm → C. For degree reasons, f cannot be constant and consequently
has image contained in D∗ . By a standard application of Hurwitz’s theorem, f is injective
as the limit of injective functions. We wish to show that the image of f is all of D∗ , and
that the ϕn converge to a quadratic differential with exponential singularity at the origin.
Let gn (z)dz2 = ϕn (z) and consider hn = (∂gn /∂z)/gn , then hn is meromorphic on the
unit disk with k + 1 simple poles, one of which is at the origin. Convergence of fn implies
convergence of hn , possibly after passing to a subsequence, to a meromorphic function
h with k + 1 poles, counted with multiplicity. As X has k infinite-angle singularities we
must have a single pole of order k + 1 at the origin, so that ϕn converge to ϕ of the
form ep(z) q(z)dz with p, q meromorphic and p having a pole of order k at the origin.
Completeness of X requires the image of f to be D∗ .
Lemma 2.2. — Let X be a flat surface of finite type, and C ≥ 0 the maximal length of any
saddle connection on the boundary of Core(X). Then the length of any Delaunay edge is bounded above
by max(C, 2ρ(X)).
(3.2) μ1 (1X ) = 0
(3.3) μ2 (a, 1X ) = (−1)|a| μ2 (1Y , a) = a, a ∈ Hom(X, Y)
(3.4) μk (. . . , 1X , . . .) = 0 for k ≥ 3
Indeed, in this case the first three A∞ -relations correspond to d 2 = 0, the Leibniz rule,
and associativity of the product.
An A∞ -functor F : A → B is given by a map ObA → ObB and multilinear maps
We will mostly consider strict A∞ -functors, i.e. those with Fd = 0 for d > 1.
Twisted complexes
We briefly recall the construction of the category of twisted complexes, TwA, over
an A∞ -category A. Twisted complexes can be thought of as formal vector bundles with
flat connection.
270 F. HAIDEN, L. KATZARKOV, M. KONTSEVICH
The first step is to form addZA whose objects are formal sums of tensor products
of the form
(3.9) V= VX ⊗ X
X∈ObA
with VX finite-dimensional graded vector spaces, zero for all but finitely many X. Mor-
phism spaces are defined by
Note that addZA has a natural shift functor and formal finite direct sums. Strictly speak-
ing, Ob(addZA) is not a set, but since the category of finite-dimensional vector spaces is
essentially small, this is a non-issue.
An object in TwA is given by a pair (V, δ) with V ∈ addZA and δ ∈ Hom 1
(V, V).
The first condition is that there is a direct sum decomposition of V (i.e. of VX as an
ObA × Z-graded vector space) so that δ is strictly upper triangular. This ensures that
μk (δ, . . . , δ) = 0 for k big, so that the second condition, which is
(3.12) μk (δ, . . . , δ) = 0
k≥1
where ai ∈ Hom((Vi−1 , δi−1 ), (Vi , δi )). Again, this sum is actually finite by our require-
ment on the δi .
The main fact we need about TwA is that its homotopy category, H0 (TwA), is tri-
angulated. When we write K0 (TwA), we always mean the K0 -group of this triangulated
category.
FLAT SURFACES AND STABILITY STRUCTURES 271
FIG. 3. — Full system of arcs in a marked surface and dual ribbon graph (dashed)
is independent of the choice of grading on c. The numbers d(h) have the property that if
v is a vertex of , then
where the sum is over all half-edges pointing to v and val(v) is the valency of v.
Conversely, given a ribbon graph together with integers d(h) for every half-edge
h satisfying (3.16) one constructs a graded marked surface with arc system by gluing
polygons with suitable grading foliation which is tangent to the arcs. This is a convenient
way to specify any compact graded marked surface by a finite amount of data.
(3.19) μn (ban , . . . , a1 ) = b
for paths b with ban = 0, and μn vanishes on all sequences of paths not of the
above forms. The lemma below ensures that this is really well-defined.
Lemma 3.1. — For a sequence of composable basis morphisms an , . . . , a1 there is at most one
factorization a1 = a1 b with a1 , a2 , . . . , an a disk sequence. If such a factorization exists with b not
an identity, then there is no factorization an = can with a1 , . . . , an−1 , an a disk sequence and c not an
identity. The dual statement also holds.
Proof. — To see the first statement note that if we have such a factorization and αi is
the arc on which ai ends, then the concatenation a1 · α1 · · · an · αn is a null-homotopic loop.
Hence, b must be homotopic (relative endpoints) to the concatenation a1 · α1 · · · an · αn and
is thus uniquely determined as a morphism.
For the second statement, assume such a factorization exists. Then c = b by the
same argument as before, but b and an end at different endpoints of αn , contradicting
an = can .
Proposition 3.1. — With the structure defined above, FA (S) is a strictly unital A∞ -category.
the boundary arcs, with some arbitrary grading. The A∞ -category FA (S)
has objects Ek ,
k ∈ Z/n, and morphisms ak : Ek → Ek+1 of some degrees |ak | ∈ Z with |ak | = n − 2
which, together with the identity morphisms, form a basis of all morphisms. The only
non-zero A∞ -terms come from strict unitality and
(3.21) μn (ak+n−1 , . . . , ak ) = 1Ek .
See also [36] for a discussion of these categories.
We make a simple observation which will be essential in what follows. Namely, that
the twisted complex
a1 an−2
E1 −→ E2 [a1 ] −→ · · · −
−→ En−1 [a1 + · · · + an−2 ]
is isomorphic to En [−|an |], the inverse isomorphisms being given by an−1 and an . In other
words, n − 1 of the boundary arcs already generate all of Tw(FA (S)). This also shows that
Tw(FA (S)) is the bounded derived category of an An−1 quiver with all arrows oriented in
the same direction. To set up this equivalence send the objects E1 , . . . , En−1 to the simple
objects corresponding to vertices of the quiver, up to some shift.
Morita invariance
Fix a graded marked surface S. If A ⊂ B are full arc systems, i.e. all arcs in A
also belong to B, then it is evident from the definition that FA (S) is a full subcategory of
FB (S).
Lemma 3.2. — The inclusion functor FA (S) → FB (S) for full arc systems A ⊂ B is a
Morita equivalence.
Proof. — We can assume that B is obtained from A by adding a single graded arc Y.
Let D be one of the disks cut out by B and with Y on its boundary. Considering D as a
graded marked surface with minimal full arc system, we have a morphism f : D → S
inducing a functor of A∞ -categories. From the discussion above, we know that in F (D)
the object corresponding to Y is isomorphic to a twisted complex over the other boundary
arcs, Y1 , . . . Yn , of D. As A is a full arc system, there is exactly one boundary of D which
gets mapped to Y, hence Y1 , . . . , Yn get mapped to arcs in A. Applying the functor f∗ we
thus find that Y is isomorphic to a twisted complex over objects in FA (S). In particular,
the inclusion functor FA (S) → FB (S) is a Morita equivalence.
Proposition 3.3. — The Morita equivalence class of FA (S) is independent of A. The equiva-
lences are canonical in the higher categorical sense (explained in the proof).
Proof. — Suppose A, B are full systems of graded arc. As established in the previous
lemma we get a canonical Morita equivalence when A ⊂ B. The same is also clearly true
when A and B just differ by grading shift.
276 F. HAIDEN, L. KATZARKOV, M. KONTSEVICH
Consider the set A of full arc systems up to isotopy on (S, M), partially ordered
by inclusion. If we view A as a category in the usual way, then the arguments above
show that we get a functor A → FA (S) from A to the category of strictly unital A∞
categories and Morita equivalences. From contractability of the classifying space |A| of
full arc systems or ribbon graphs (see [25, 26]), it now follows that the various categories
FA (S) are canonically Morita equivalent.
Thus, TwFA (S) is essentially independent of A and we may drop it from the nota-
tion and just write F (S), the (topological) Fukaya category of S.
Lemma 3.3. — Let S be a compact and connected graded marked surface. If S has at least one
boundary arc, then it has a full formal system of arcs.
Proof. — This follows from the fact that we can find a system of arcs on S which
cuts it into a single disk, the boundary of which must include all boundary arcs of S. This
in turn is seen from the dual ribbon graph. If it has more than one vertex we can contract
some edges to reduce the number of vertices to one eventually.
We say a graded quiver with relations is of type F1 if it arises as above from a
full formal system of arcs. It is not difficult to see that a graded quiver with quadratic
monomial relations is of type F1 if and only if
1. There are no cycles a1 , . . . , an , n ≥ 1, with ai ai+1 = 0 for all i ∈ Z/n.
2. Each vertex has at most two incoming and outgoing arrows.
3. Let a, b = c be arrows. If ab, ac are defined, then ab = 0 or ac = 0, but not both.
If ba, ca are defined, then ba = 0 or ca = 0, but not both.
is exact.
278 F. HAIDEN, L. KATZARKOV, M. KONTSEVICH
Proposition 3.5. — Let Q be a graded quiver with quadratic monomial relations, A = KQ/R
its path algebra.
1. Suppose there are no cyclic paths α1 · · · αn in Q with (αi , αi+1 ) ∈ R, i ∈ Z/n, then A is
homologically smooth.
2. Suppose there are no cyclic paths α1 · · · αn in Q with (αi , αi+1 ) ∈
/ R, i ∈ Z/n, then A is
proper.
Proof. — 1. Under the stated condition on cyclic paths, we see that Mn = 0 for
n 0, hence A is perfect as an Aop ⊗ A-module.
2. The condition implies that only finitely many paths are non-zero in A, so A has
finite rank over K.
We apply these results to graded linear models of F (S). In this case, the first condi-
tion of Proposition 3.5 is always satisfied, while the second is satisfied if S has no boundary
components without corners.
Corollary 3.1. — Let S be a compact graded marked surface without boundary components
diffeomorphic to S1 , then F (S) is homologically smooth and proper.
Question. — If S is a compact graded marked surface which has a full arc system
that gives an acyclic quiver, then the category F (S) has a full strong exceptional collection
formed by these arcs. Are there simple conditions on S which allows one to find such an
arc system?
3.5. Localization
We begin by recalling a version of Drinfeld’s construction for strictly unital
A∞ -categories, studied in detail in [33]. Let A be a strictly unital A∞ -category over a
field K and E ∈ Ob(A). It will be convenient to use the notation A(X, Y) for Hom(X, Y)
in this subsection. We will define the quotient category A/E = B of A by E, which will
again be a strictly unital A∞ -category. Informally, B is obtained by freely adjoining a
morphism ε ∈ B −1 (E, E) with μ1 (ε) = 1E . Set
(3.27) B (X, Y) = A(X, Y)⊕
⊗n
⊕ A(E, Y) ⊗ K[1] ⊗ A(E, E) ⊗ K[1] ⊗ A(X, E)
n≥0
as
an εan−1 ε · · · a1
Proposition 3.6. — Let A be an arc system on a graded marked surface S and E ∈ A a boundary
arc. By the modification as above we get S with arc system A containing a null arc. Then there is a
natural equivalence of A∞ -categories
(3.31) FA (S)/E ∼ = FA S
under FA (S).
compatible with the functors from FA (S). It is obtained by factoring a boundary path
in M into its pieces which are alternately contained in M and E. The functor G is not
280 F. HAIDEN, L. KATZARKOV, M. KONTSEVICH
an isomorphism of A∞ -categories, but becomes one once we drop E from both the tar-
get and the source category. By general properties, E ∈ H∗ (FA (S)/E) is a zero-object,
and it is easily seen that the same is true in H∗ (FA (S )), implying that G is a quasi-
equivalence.
Theorem 3.1. — Let S be a graded marked surface with ribbon graph G dual to a full graded
system of arcs. Then F (S) represents global sections of the constructible cosheaf E on G defined above.
each p ∈ P, and keeping the same arrows and relations. Passing to the associated path-
categories P , Q, R, where P is thought of as a quiver without arrows, we get a coequalizer
diagram
f1
(3.33) P Q R
f2
Proposition 3.7. — The diagram (3.33) is a homotopy coequalizer in the category of dg-
categories.
of Q making the
Proof. — To show this, we will make a cofibrant replacement Q
diagram
f1
(3.34) P
Q
f2
cofibrant by the condition on f1 f2 , and verify and the coequalizer of (3.34) is quasi-
isomorphic to R.
The graded linear path-category A of Qn has a cofibrant replacement given by the
bar-cobar resolution B = BA. The dg-category B is described explicitly by a dg-quiver
with n vertices, arrows βij from the i-th to the j-th vertex, 1 ≤ i < j ≤ n, with degrees
and differential
extended to paths by the graded Leibniz rule. The functor P : B → A sends βi,i+1 to
αi,i+1 and all other arrows to zero. The right inverse I sending αi,i+1 to βi,i+1 is not mul-
tiplicative, but an inverse of P on the level of cohomology, as the homotopy H defined
by
(−1)|βik−1 ,ik | βik−2 ,ik · · · βi1 ,i2 if ik = ik−1 + 1, k ≥ 3
(3.37) H(βik−1 ,ik · · · βi1 ,i2 ) =
0 else
and
Proof of Theorem 3.1. — Let S be a graded marked surface, which we may as well
assume to be connected. Suppose first that S has at least one boundary arc. Then S
has a full formal system of arcs A with dual ribbon graph G. Recall that (G, E ) is the
homotopy colimit of a diagram with the various categories Ce , Cv and functors between
them. Removing the categories Ce for e an edge ending in a boundary arc from the
diagram does not change the colimit, since such Ce are being included as subcategories
into some Cv only once. In the reduced diagram there is now, by formality of the arc
system, at least one object in each Cv which is not in the image of any functor Ce → Cv .
We remove such an object from each Cv , producing a Morita-equivalent diagram. Note
that the resulting diagram is now a diagram of graded linear categories. It is essentially
a diagram of the form (3.33) where P is the coproduct of the Ce (with P identified with
the set of arcs in A) and Q is the coproduct of the Cv . The coequalizer R is just the
graded linear model for F (S), and we know it is a homotopy coequalizer by the previous
proposition.
FLAT SURFACES AND STABILITY STRUCTURES 283
The previous arguments also show that (G, E ) is independent of G, and hence
that the theorem holds in fact for any full system of arcs on a surface with boundary arc.
Namely, it suffices to check that global sections do not change when an edge of G is
contracted. This can be checked in a neighborhood of that edge, so we can assume that
G has only one internal edge and is dual to a full arcs system A on the disk D with single
internal arc. Let G be the graph obtained by contracting the internal edge in G, and E
its cosheaf of categories. As it is already established that both (G, E ) and (G , E ) are
represented by F (D), they must be equivalent.
What remains is the case when S has no boundary arcs. As discussed in Section 3.5,
the category F (S) is a localization of a category F (S ) where S is obtained from S by
inserting a boundary arc on some boundary component. So to complete the proof of the
theorem it suffices to check that the corresponding categories of global sections have the
same relation.
Choose some ribbon graph G on S dual to a full system of graded arcs. To get a
ribbon graph G for S we just need to remove the edge ending on the unique boundary
arc. This means that in the diagram computing global sections we change some Cv from
type An to type An−1 , which is indeed just localization by the boundary arc. Finally, taking
the quotient by some object is a special case of a homotopy push out, thus commutes with
colimits, so that (G, E ) is the quotient of (G , E ) by the boundary arc.
In this section we deal with the problem of classifying objects in F (S). First, in
Section 4.1, we assign objects in F (S) to certain immersed curves in S with a local system
of vector spaces. The purpose of the rest of the section is to prove that we get all objects
in this way. In Section 4.2 we introduce nets and study their representations. Section 4.3
discusses a minimality condition on twisted complexes which ensures uniqueness up to
isomorphism. The proof of the classification is completed in Section 4.4.
Suppose that S is compact so that we have a category F (S), well defined up to canonical
equivalence. The purpose of this subsection is to show that an admissible curve together
with a local system of finite dimensional K-vector spaces (on its domain) gives an equiva-
lence class of objects in F (S).
First, in the case when S is topologically a disk any admissible curve c is a graded
arc. Thus we can find a full arc system A which includes c, so that c is an object of
FA (S) by definition. The isomorphism class of that object is clearly well-defined in F (S)
independently of the arc system.
Returning to the case of general S, we will first deal with admissible curves c which
have domain [0, 1]. For any such c we can find a graded marked surface S which is of
disk-type and with a map f to S so that c is the image of an admissible curve c̃ under f . To
see this, consider the universal cover S of S and lift a full arc system A on S to
A on
S. Lift
c to c̃ on S and take as S a closed disk which is cut out by arcs in A and which contains
c̃. Now, c̃ gives an equivalence class of objects in F (S ), and the image under the functor
F (S ) → F (S) is independent of the choice of (S ). This follows from the fact that if we
have S , S as above, then the maps S → S, S → S both factor through a third S → S
as can be seen by looking at the universal cover again.
Suppose now instead that c is an admissible curve with domain S1 and local sys-
tem V of finite-dimensional vector spaces on it. We will follow the same strategy as be-
fore and assume first that (S, M) is of annular type, i.e. topologically a compact annulus
with corners on each boundary component. Choose a cyclic sequence of disjoint non-
isotopic arcs Xi , i ∈ Z/n so that at least one connects the two components of ∂S and such
that every component of M contains either exactly two endpoints of the arcs, belonging
Xi , Xi+1 for some i, or none of the endpoints (see Figure 5). Thus we get a sequence ai ,
i ∈ Z/n of distinct boundary paths so that ai connects endpoints of Xi , Xi+1 . If we fol-
low X0 , a0 , X1 , a1 , . . . we get a path which after suitable smoothing near the intersection
points becomes a simple closed loop isotopic to c. It is possible to choose grading and a
local systems on Xi , ai , so that the smoothed path is isotopic to c as a graded curve with
local system. As a result, each ai will be morphism of degree 1 either from Xi to Xi+1 or
in the other direction. Further we get a vector space Vi of sections of over Xi and parallel
transport Ti : Vi → Vi+1 along ai . Consider the twisted complex
(4.1) Vi ⊗ Xi , T±1
i ⊗ ai
FLAT SURFACES AND STABILITY STRUCTURES 285
where the signs are determined by the direction of the morphisms ai . This is the object
of F (S, M, ) we assign to c.
Lemma 4.1. — The equivalence class of the twisted complex constructed depends only on the
isotopy class of the graded curve c with local system.
Proof. — We claim first that we can replace X1 , . . . , Xn−1 by a single arc Y with
the pair X0 , Y giving an isomorphic twisted complex. Indeed, cutting S along X0 we
get a surface of disk type S in which the sequence of arcs X1 , . . . , Xn−1 concatenates
to a single graded arc Y. The corresponding isomorphism of twisted complexes formed
from X1 , . . . , Xn−1 and Y respectively was established in Section 3.3, and the claimed
isomorphism of twisted complexes formed from X0 , . . . , Xn−1 and X0 , Y follows.
We have reduced the problem to the case of two arcs. Any two pairs of arcs satisfy-
ing our requirements are related by Dehn twists (automorphisms) of S. It is clear that the
isotopy class of c is invariant under Dehn twists. To finish the proof we need to show that
the twisted complex associated with c is invariant under the induced autoequivalence, up
to isomorphism. This can be checked by direct computation, or by using the equivalence
(4.2) H0 Tw FX0 ,Y (S) = Db P1
We have excluded above the case when one or both boundary components are en-
tirely contained in M. These cases can be handled fairly easily directly, or alternatively by
the localization construction of the previous section. The case of general S is handled by
finding maps from surfaces of annular type. Here the argument uses the annular covering
associated with c instead of the universal covering.
286 F. HAIDEN, L. KATZARKOV, M. KONTSEVICH
(4.3) f1 ◦ α = α ◦ f1
(4.4) f2 (Domβ) ⊂ Domβ
(4.5) f2 (B \ Domβ) ⊂ B \ Domβ
(4.6) f2 |Domβ ◦ β = β ◦ f2 |Domβ
(4.7) f2 (Bi ) ⊂ B f1 (i) and f2 |Bi is increasing
α β α β α
(4.9) • • • • ··· •
and
α β α β α
(4.10) • • • • ··· •
β
gri f
gri V gri W
φi ψi
grβ(i) f
grβ(i) V grβ(i) W
commutes for all i ∈ Dom(β). We note that the category of representations of a net as
in (4.9) is the category of finite dimensional vector spaces, while for (4.10) we get the
category of finite dimensional representations of Z (up to equivalence).
Given a morphism of nets f = (f1 , f2 ) : X = (A, α, B, β) → X = (A , α , B , β )
with f2 injective on each Bi , there is an induced pushforward functor f∗ on the correspond-
ing categories of representations. If V = (V, Fi , φi ) is a representation of X , then we
produce a representation f∗ V = W = (W, Gi , ψi ) of X with W = V as vector spaces and
(4.11) Wj = Vi , j ∈ A /α
i∈A/α,f1 (i)=j
(4.12) Gl = Fk , where k = max r ∈ Bi | f2 (r) ≤ l
i∈A,l∈Bf (i)
1
If f2 fails to be injective on some Bi ’s, then one can still define f∗ V given additional choices
on V (i.e. in a non-functorial way). Namely, for each i ∈ A and j ∈ Bf1 (i) ∩ Dom(β ) with
A
X Y
B
of parallel surjective linear maps between finite dimensional vector spaces. First, we have
two increasing filtrations on X:
k−1
(4.17) A−1 0 ⊂ A−1 BA−1 0 ⊂ A−1 BA−1 BA−1 0 ⊂ · · · ⊂ A−1 BA−1 0 ⊂ ···
=:Fk
k−1
(4.18) B−1 0 ⊂ B−1 AB−1 0 ⊂ B−1 AB−1 AB−1 0 ⊂ · · · ⊂ B−1 AB−1 0 ⊂ ···
=:Gk
B−1 A
Mi,j Mi−1,j+1
A−1 B
FLAT SURFACES AND STABILITY STRUCTURES 289
for i ≥ 0, j ≥ 1, where
Fi ∩ Gj
(4.20) Mi,j =
(Fi−1 ∩ Gj ) + (Fi ∩ Gj−1 )
for i, j ≥ 1. In particular,
• Ker(B) has an exhaustive filtration F ∩ Ker(B) with associated graded pieces
Mi,1 ,
• Ker(A) has an exhaustive filtration G ∩ Ker(A) with associated graded pieces
M1,j ,
• there are isomorphisms Mi,1 −→ M1,i .
We proceed with the proof of the theorem.
Proof. — Let (V, {Fi }, φi ) be a representation of a net X = (A, α, B, β). The proof
is by induction over
(4.21) dim Vi
|Bi |≥2
for all nets simultaneously. If h(X ) = 1 we are done, so let us assume that h(X ) ≥ 2. We
can also assume, by passing to subsets of A and the Bi , that all associated graded gri V are
non-zero and all Bi are non-empty.
Let r ∈ A with |Br | = h(X ). The goal is to separate a “strand” of the representation
going through r, thus decreasing the quantity (4.21). Let n = max Br and find the unique
k ∈ Bα(r) such that
Set X1 = F<k + F<n , X2 = Fk ∩ (F<k + F<n ) and consider the refinements of the two
filtrations of Vr :
such that (4.30) holds. We also have a direct sum X1 ⊕ U = Vr , but (4.29) may fail.
However, we can ensure (4.29) after shearing U in the directions of F<k , which does not
affect (4.30).
The choice of U gives an identification
p2
grk V Fk /X2
By the discussion preceding the proof we get first of all a pair of non-exhaustive filtrations
(4.33) 0 = G0 ⊂ G1 ⊂ · · · ⊂ Gm , 0 = H0 ⊂ H1 ⊂ · · · ⊂ Hm = Gm
of grn V, where G1 = Ker(p1 ) and H1 = φn−1 (Ker(p2 )). For the associated double graded
Mij , 1 ≤ i, j, i + j ≤ m + 1, we have isomorphisms Mij → Mi+1,j−1 induced by φn . Further,
we have restricted filtrations on Ker(p1 ), Ker(p2 ) lifting to refinements of the filtrations
on Vr :
with
where in Br (resp. Bα(r) ) the new elements replace n (resp. k) in the total order. There is a
morphism of nets f : X → X with 1, (1, i, j) → n, and 2, (2, i, j) → k.
Let G be the preimage of Gm = Hm under the projection Fn → grn V. V determines
a representation V = (V , Fi , ψi ) of X with
the restrictions of the filtrations (4.34) to X1 , and ψ1 , ψ1,i,j induced by φn . We claim that
f∗ V is isomorphic to V. Choose a complement Y to G ⊂ Fn with Y ⊂ Fk . Further, let
Y1,1 , . . . , Y1,m ⊂ grn V with
Also we can use the X1,j and Xi,1 in the definition of f∗ V . Choose a complement Z to
X1 ⊂ G with Z ⊂ Fk , allowing us to lift Yi,j to Zi,j ⊂ Fk . Combining the various splittings
we obtain an isomorphism
(4.46) X1 ⊕ (Fn /G) ⊕ Mij → Vr
i,j≥2
i+j<m+1
Call a net as in (4.10) a cycle. Note that for any cycle and any n ≥ 2 there is a
morphism from another cycle which is n : 1, i.e. an n-fold “covering”. These, and iso-
morphisms, are the only morphisms between connected nets of height 1. We use this to
formulate a strengthening of the previous theorem.
Lemma 4.2. — The functor T : addZ(A) → addZ(Ae ) induced by the augmentation reflects
isomorphisms.
Proposition 4.1.
1. Every twisted complex A ∈ TwA is isomorphic to a direct sum A = Am ⊕ Ac with Am
minimal and Ac contractible.
2. Any homotopy equivalence between minimal twisted complexes is an isomorphism.
Proof. — 1. Let (M, δ) be a twisted complex over A, (M, δ) its image in Tw(Ae ). By
semisimplicity of Ae , (M, δ) is isomorphic to a direct sum Bc ⊕ Bm where Bc is contractible
and Bm has trivial differential. Therefore, (M, δ) is isomorphic to a twisted complex of
the form
⎛ ⎛ ⎞⎞
δ11 δ12 δ13
⎝K ⊕ K[−1] ⊕ L, ⎝ 1K + δ21 δ22 δ23 ⎠⎠
δ31 δ32 δ33
(4.50) 1M − gf = δ M G + Gδ M , 1N − fg = δ N H + Hδ N .
Theorem 4.3. — Let S be a compact graded marked surface, then the construction of Section 4.1
sets up a bijection between isomorphism classes of indecomposable objects in H0 (F (S)) and isotopy
classes of admissible curves with indecomposable local system.
Remark. — In the special case when H0 (F (S)) is the bounded derived category
of a gentle algebra the classification of indecomposable objects was found by Drozd–
Burban [16] and Bekkert–Merklen [8]. To classify objects in F (S) in general we need to
consider Z-graded algebras and the language of twisted complexes instead of projective
resolutions. Consider for example the Dynkin quiver with two parallel arrows which do
not have the same degree.
with total order so that (i, k + d(αi ) − 1) < (i + 1, k) on {1, . . . , n} × Z and (1, n) <
(2, n) < (3, n) on {1, 2, 3} × Z. In fact, we get an embedding of the category of minimal
twisted complexes into the category of representations of the above net as a full subcate-
gory. We do not get an equivalence of categories, since we only consider δ which decrease
the filtration (4.54).
Turn now to the case of general Q. Let D be the set of maximal non-zero paths in
KQ. Then for each element of D there is a sub-quiver of Q which is of the simple form
above, and Q is obtained from their disjoint union by identifying some pairs of vertices.
The corresponding net is now given by
(4.57) A = D × {1, 2} × Z
(4.58) α = id × (12) × id
(4.59) B = (v, d) ∈ Q0 × D | v on d D × {1, 2, 3} × Z =: Bg Ba
with partial order defined so that for an arrow α in d ∈ D we have (∂1 (α), n + d(α) − 1) <
(∂0 (α), n) and (d, 1, n) < (d, 2, n) < (d, 3, n). Further,
(4.60) β = τ id × (13) × id
This section contains our main result, identifying M(S) with an open and closed
subset of Stab(F (S)). We begin by reviewing Bridgeland’s axioms in Section 5.1. In
Section 5.2 we compute K0 (F (S)) in terms of singular homology. The main theorems
are stated in Section 5.3 and proven in the remaining subsections. Sections 5.4 and 5.5
deal with the support property and Harder–Narasimhan filtrations, respectively. In the
final Section 5.6 we prove that the map on moduli spaces is complex bianalytic onto an
open and closed subset.
0 = E0 E1 ··· En−1 En = E
A1 An
with 0 = Ai ∈ C φi and φ1 > φ2 > · · · > φn . The Ai are called the semistable com-
ponents of E.
4. If 0 = E ∈ C φ then Z(E) := Z(cl([E])) ∈ R>0 eπiφ .
5. The support property: For some norm . on ⊗R and C > 0 we have an estimate
(5.1) cl(E) ≤ C|Z(E)|
for E ∈ C φ .
The set Stab(C , ) of all stability structures has a natural topology which is induced by
the metric
mσ2 (E)
(5.2) d(σ1 , σ2 ) = sup |φσ2 (E) − φσ1 (E)|, |φσ2 (E) − φσ1 (E)|, log
− − + +
0=E∈C m (E)
σ1
298 F. HAIDEN, L. KATZARKOV, M. KONTSEVICH
cf. [11]. Here, for an object E ∈ C with semistable components A1 , . . . , An with phases
φ1 > · · · > φn one defines m(E) = |Z(Ai )|, φ − (E) = φn , and φ + (E) = φ1 . The main
result of [11] (see also [31]) is that the projection
Stab(C , ) → Hom(, C)
Theorem 5.1. — Let S = (S, M, η) be a compact graded marked surface, then there is a
natural isomorphism of abelian groups
K0 F (S) ∼ = H1 (S, M; Zτ ).
Proof. — Choose
√ a full system A of graded arcs on S. Each arc defines a class in
H = H1 (S, M; Z ). From cellular homology it follows that H is the group generated
by these arcs and with relations of the form ±X1 ± · · · ± Xn = 0 where X1 , . . . , Xn are
the arcs bounding a disk cut out by A. Correspondingly, it is also clear that the arcs in A
generate
(5.3) K := K0 F (S) = K0 Tw FA (S)
and satisfy the same relations, which follows from the observation in Section 3.3. By
construction we get a surjective homomorphism H → K which is independent of the
choice of arcs and natural with respect to maps of graded marked surfaces. It remains to
show that no additional relations are needed to present K.
Suppose first that M has no components diffeomorphic to S1 . Choosing a formal
collection A of arcs, we can present F (S) as the category Tw(KQ) of twisted complexes
over the path category of a graded quiver Q with quadratic monomial relations. Note
that H is freely generated by A, so we need to verify that K is freely generated by the
vertices Q0 of Q. One way to see this is by using the explicit resolution of the diagonal
of KQ, showing that, as a bimodule, KQ is a repeated cone of Yoneda bimodules. As a
consequence, Tw(KQ) ∼ = Perf(KQ), and an inverse of the map
(5.4) ZQ0 → K0 Perf(KQ)
FLAT SURFACES AND STABILITY STRUCTURES 299
sending a vertex to the corresponding simple module, is induced by the dimension vector
of a module.
For a general graded marked surface S there is a S obtained by adding two corners
to each component of ∂S diffeomorphic to S1 . Let N denote set of boundary arcs of S cre-
ated in this process, and N the corresponding full triangulated subcategory of H0 (F (S ))
generated by N. Each arc in N has endomorphism algebra of the form K[x]/x2 with
|x| ∈ Z, so N is a product of categories of the form H0 (Tw(K[x]/x2 )). We get a commu-
tative diagram
√ √
(5.5) ZN H1 S , M , Z H1 (S, M, Z ) 0
K0 ( N ) K 0 F S K0 F (S) 0
where we use Proposition 3.6 to get the bottom row. We claim that the rows are exact.
For the top row this is just the exact sequence of a triple. For the bottom row this uses the
fact that N is idempotent complete, thus a thick subcategory so that Proposition 3.1 in
[24] can be applied. Since the left and middle horizontal maps are isomorphisms, so is
the right one.
When a flat surface S comes from a quadratic differential with higher order poles,
the category F (S) includes some objects which have infinite length since
1
(5.6) x−n/2 dx = ∞
0
for n ≥ 2. We need to pass to the subcategory of F (S) which does not include these
objects. Formally, we want to allow S = (S, M) to have boundary circles which are
unmarked, i.e. not belonging to M. Thus, define F (S, M) as the full subcategory of
F (S, M ) of objects corresponding to immersed curves avoiding M \ M, where M is
the union of M and all boundary circles.
Lemma 5.1. — The full subcategory F (S, M) ⊂ F (S, M ) is triangulated, i.e. closed under
shifts and cones.
300 F. HAIDEN, L. KATZARKOV, M. KONTSEVICH
Finally, if both endpoints of A lie on B, then IB (A) is a sum of two graded vector spaces
as above. The map on morphisms is defined in the obvious way so that IB (a) = 0 if a is a
boundary path which is not in B. One checks that this gives a well defined A∞ -functor.
Proposition 5.1. — Let S be a compact graded marked surface which is allowed to have un-
marked boundary circles. Then there is a natural map
(5.8) K0 F (S) −→ H1 (S, M; Zτ ).
The group K0 (F (S)) is often infinitely generated, and so the above map will fail
to be injective. It may also fail to be surjective, e.g. for the cylinder with non-standard
grading, where F (S) is trivial if both boundary circles are unmarked.
Proof. — Let S be the graded marked surface obtained from S by replacing each
unmarked boundary circle with one marked and one unmarked boundary arc. Then
F (S) is also a subcategory of F (S ), as follows from localization. Note that
(5.9) ι : H1 (S, M; Zτ ) −→ H1 S , M ; Zτ ∼ = K 0 F S
is an inclusion and the image of ι contains the image of K0 (F (S)) −→ K0 (F (S )). Hence
we have the desired natural map.
FLAT SURFACES AND STABILITY STRUCTURES 301
Theorem 5.2. — Let X be a flat surface of finite type. Define subcategories C φ ⊂ H0 (F (X))
so that isomorphism classes of indecomposable objects correspond to unbroken graded geodesics of phase φ
with indecomposable local system, and let
(5.10) Z : K0 F (S) −→ H1 (X, ∂X; Zτ ) =: → C
be the period map. Then this data satisfies the axioms of a stability structure. In particular, for a graded
surface of finite type X there is a continuous map
(5.11) M(X) −→ Stab F (X)
Theorem 5.3 (Main theorem). — Let X be a graded surface of finite type, then the map
M(X) → Stab(F (X)) is injective and its image open and closed, thus a union of components of
Stab(F (X)).
Remark. — When C is compact, then necessarily g(C) = 1. In this case one can
use homological mirror symmetry for elliptic curves [41] and Atiyah’s classification [4]
to derive a version of Theorem 5.2. The computation of Stab(Db (E)) in [11] shows that
M(X) = Stab(X).
In the case where C is a punctured torus, the mirror is a nodal elliptic curve E.
The space of stability conditions on Db (E) was computed by Burban–Kreußler [17].
Our theorem gives a different description of the same space.
The proofs of Theorem 5.2 and Theorem 5.3 will occupy the rest of this section.
302 F. HAIDEN, L. KATZARKOV, M. KONTSEVICH
Proposition 5.2. — There is a norm . on R , C > 0 such that for γ ∈ the class of a
finite geodesic on X there is an estimate
(5.13) γ ≤ C|Z(γ )|
where Z : → C is the period map.
Proof. — The locally constant sheaf Rτ is given by flat sections of a flat line bundle
with metric for which we use the same notation. We have
(5.14) R∗ = H1 (Y, ∂Y; Rτ )∗
(5.15) = H1 (Y, ∂Y; Rτ )
(5.16) = H1dR (Y, ∂Y; Rτ )
be the closure of
Choose forms ω1 , . . . , ωn representing a basis of H1dR (Y, ∂Y; Rτ ). Let K
the preimage of K = Core(X) under the inclusion Y \ ∂Y → X. Define a norm on R by
and set
(5.18) C= ωi |K ∞
i
where the norm is taken with respect to the flat metric. Let α be a finite geodesic on X
with class γ ∈ , then
(5.19) l(α) = |Z(γ )|
where l(α) is the length. Thus
where δi has non-zero coefficients corresponding to the singular points where the geodesic
representative of α passes through without changing phase, which is equal to φi . Also,
(5.27) E= Aj , δ
j
where δ has non-zero coefficients corresponding to all the singular points the geodesic
representative passes through (i.e. where αi meets αi+1 ). Note that this really gives mor-
phisms of degree 1 since the concatenation of the saddle connections smooths to a graded
curve. To show that the twisted complex representation of E in terms of the Bi is a HN-
tower, we need to check that components of δ increase phase. For this we use that fact
that the concatenation of the αi is a geodesic, and so αi , αi+1 necessarily meet at a singular
point in an angle φ ≥ π , thus φ > π if the phase jumps. The corresponding morphism,
without loss of generality from Ai to Ai+1 , has degree 1, and so
φ
(5.28) 1 = φ(Ai ) − φ(Ai+1 ) +
π
which implies φ(Ai ) < φ(Ai+1 ).
Suppose now that X is a general flat surface of finite type. The first claim is that
Hom(A, B) = 0 for a pair of graded saddle connections A, B with φ(A) > φ(B). To see
this, let Y be the flat surface so that its real blow-up Note
Y is the universal cover of X.
that all singularities of Y have infinite cone angle. Lift A, B to graded curves A,
B on Y
with the same phases. We have
(5.29) Hom(A, B) = Hom( A, g
B)
)
g∈π1 (X
and hence it suffices to prove the claim for Y. But any two saddle connections in Y are
objects in some subcategory of F (Y) which is the Fukaya category of some surface of
disk-type as above, and so our previous arguments can be applied.
Second, we claim that the HN-property holds for admissible curves c with domain
[0, 1]. To see this, lift c to a graded curve c̃ in the universal cover Y. We work in the
category generated by all the saddle connections in a geodesic representative of c̃. It
corresponds to some surface of disk-type in the universal cover. We get a HN-tower for c̃
in this category and push it forward to F (X) to get a tower for c in that category.
We turn to the case when X is of annular type, i.e. Xsm is diffeomorphic to an an-
nulus with grading such that a simple closed loop around it is gradable, and all singular-
ities have infinite cone angles. Let S ⊂ X be the corresponding compact graded marked
surface. The category F (X) is the bounded derived category of finite-dimensional repre-
sentations of a quiver which is a cyclic chain of arrows, all in degree zero. The number of
arrows oriented in one or the other way is equal to the number of marked boundary arcs
FLAT SURFACES AND STABILITY STRUCTURES 305
FIG. 6. — The cylinder Z, cut along a dotted line, closed geodesics α, β, saddle connections Si , arcs Ei
that φ(Si ) < φ(B). Namely, there is an isotopy moving β to the concatenation of the Si
and Ei , and it should move any point of intersection with ∂Z so that Z lies to the right.
From what we have shown in the annular case, it follows that Hom(A, Si ) = 0 for all i.
Further, A is orthogonal to any Ei , so Hom(A, B) = 0 follows. This completes the proof
of Theorem 5.2.
which is a union of horizontal rays starting at points of the form log(Z(E)), E semistable.
The support property ensures that the central charges Z(E) ∈ C, E semistable, form a
discrete subset, hence CF,σ ⊂ R2 is closed.
for all 0 = E ∈ C . For a subset A of a metric space let B (A) be the set of points with dis-
tance ≤ to A. To show that CF,σ2 ⊆ B (CF,σ1 ) it suffices to find, for every σ2 -semistable
object E with Hk (F(E)) = 0, a σ1 -semistable object E with Hk (F(E )) = 0 and
(5.34) |φσ2 (E) − φσ1 E | ≤ , log |Zσ1 E | ≤ log |Zσ2 (E)| + .
this will imply the claim. Consider the functor I : F (X) → D(Mod(K)) associated with
s as before. We have closed subsets Ci = CI,σi ⊂ R2 converging to C = CI,σ in the Haus-
dorff topology. Let D denote the set of directions (of geodesics) from s. The choice of
308 F. HAIDEN, L. KATZARKOV, M. KONTSEVICH
grading of marked boundary as above gives an identification D ∼ = R/nπ when s has cone
∼
angle nπ and D = R when s has infinite cone angle. By the support of Ci , Supp(Ci ),
we mean the closure of its projection to D . It follows from compactness of Core(Fi )
that each Ci is compactly supported. By convergence, there is compact J ⊂ D contain-
ing all Supp(Ci ) and Supp(C). Let ri (φ) ∈ R≥0 ∪ {+∞} be the length of the maximal
geodesic starting at s in direction φ ∈ D and which is entirely contained in Vs,i . Note
that the functions ri are determined by Ci , thus σi alone, are continuous, and converge
pointwise to r : D → R≥0 ∪ {+∞} determined by σ . Define qi (φ) ∈ R≥0 similarly as the
length of the maximal geodesic starting at s in direction φ ∈ D and which is entirely con-
tained in Vs,i ∩ Core(Fi ). We need to shows that qi are uniformly bounded. By definition,
qi (φ) ≤ ri (φ), and
(5.36) (φ − π, φ + π ) ∩ J = ∅ =⇒ qi (φ) = 0.
Suppose, for contradiction, that there is a sequence φi ∈ D with qi (φi ) → ∞. By (5.36)
the φi are contained in a compact subset of D and we may assume φi → φ after passing
to a subsequence. Now since qi (φi ) → ∞ we must have r(φ) = ∞, which means that the
interval (φ − π, φ + π ) is disjoint from Supp(C). Thus also
% &
π π
(5.37) φ − ,φ + ∩ Supp(Ci ) = ∅
2 2
for sufficiently large i. The sector bounded by directions φ − π/2 and φ + π/2 in
Dom(exps ) cannot map entirely to Core(Fi ), as this contradicts ρ(Fi ) < ∞. Hence
the part of the sector which maps to Core(Fi ) is a triangle or just the point s. The
length of the side of the triangle opposite the vertex s is bounded above by β. Thus also
qi |(φ − π/2, φ + π/2) has an upper bound in terms of β, which is independent of i, a
contradiction.
We have established that ρ(Fi ) and β(Fi ) have upper bounds independent of i. By
Lemma 2.2 the length the Delaunay edges on all the Fi is bounded above. Thus, there
are only a finite number of polygonal subdivisions up to isotopy arising as Delaunay
partitions in the sequence Fi . Passing to a subsequence, we may hence assume that the
Delaunay subdivisions for the Fi are all isotopic, i.e. combinatorially equivalent. Since
σi → σ , lengths of edges converge to positive numbers as i → ∞. It may still happen
that some polygons become degenerate in the limit, i.e. their area goes to zero. If we glue
back together the set of non-degenerate limiting polygons, we get a flat structure F which
is the limit of the Fi and corresponds to σ .
6. Cluster-like structures
6.1. S-graphs
In this section we specialize to the case of flat surfaces with infinite area. It turns
out that for generic choice of horizontal direction the heart of the t-structure of the cor-
FLAT SURFACES AND STABILITY STRUCTURES 309
responding stability structure is Artinian with a finite number of simple objects and de-
scribed by a certain type of finite graph. Tilting of the t-structure leads to mutation of the
graph.
Suppose X is a flat surface of finite type with infinite area. Recall from Section 2.4
that, for generic choice of horizontal direction, the leaves converging to the conical singu-
larities cut the surface into horizontal strips of finite or infinite height. The combinatorial
structure of this horizontal strip decomposition can be encoded in terms of a finite graph,
G, with set of vertices Xsg and set of edges the horizontal strips of finite height. An edge
is attached to the singularities which lie on the boundary of the corresponding horizontal
strip. We also need to record two additional pieces of data:
1. For each v ∈ Xsg with finite (resp. infinite) cone angle a cyclic (resp. total) order
on the set Hv of half-edges meeting that vertex given by the counter-clockwise
order of horizontal strips around s.
2. For each pair of successive half-edges a < b ∈ Hv a positive integer, d(a, b), so
that d(a, b) − 1 is the number of horizontal strips of infinite height between a
and b as we go around v.
We call a graph as above, i.e. a finite graph with orders on Hv and integers d(a, b),
an S-graph. In order to record the entire flat structure we attach a number Z(e) in the
upper half-plane, H, to each edge e, given by the vector between the singularities on the
boundary of the corresponding horizontal strip. We summarize the construction in the
following proposition.
The data of an S-graph G may be encoded in terms of a graded quiver with rela-
tions Q. Vertices of Q are edges of G. Arrows of Q are given by pairs (h1 , h2 ) of successive
half edges h1 < h2 ∈ Hv in degree d(h1 , h2 ). Relations are quadratic: (h1 , h2 ) · (h3 , h4 ) = 0
when h2 = h3 belong to the same edge. Note that in the graded linear path category
A = KQ we have
K if A = B
(6.1) Hom (A, B) = 0,
<0
Hom (A, B) =
0
0 else
These properties ensure that the extension closure of the objects of A in TwA = PerfA
is the heart of an Artinian bounded t-structure, and objects of A are precisely the simple
ones.
310 F. HAIDEN, L. KATZARKOV, M. KONTSEVICH
Proposition 6.2. — Let X be a flat surface with infinite area such that no horizontal leaves have
finite length (which is true for generic choice of horizontal direction), and A be the heart of the t-structure
for the stability structure on F (X) determined by the flat metric, then A is Artinian with a finite number of
simple objects. In terms of the S-graph, A is described by the graded quiver with relations considered above.
Proof. — Choose a family Xt , t ∈ [0, 1] of flat structures so that X0 = X, none of
the Xt have finite horizontal leaves, and the simple saddle connections of X1 are vertical.
All the Xt have horizontal strip decompositions described by the same S-graph, only the
parameters Z(e) change, and are in iR>0 for X1 . The corresponding stability structures
σt all have the same heart of the t-structure, as none of them have semistable objects with
real central charge (by the condition on the horizontal leaves), i.e. there is no wall-crossing
of the second kind.
The flat surface X1 has a purely one-dimensional core, which is just the union K of
the simple saddle connections. This is clear, since K contains all singularities, and exterior
angles are ≥ π . We conclude that X1 has no closed geodesics and all saddle connections
are simple. Thus, stable objects of σ1 in the heart of the t-structure are just the simple
geodesics. Also, the central charge of σ1 takes values in iR, so simple objects in the heart
of the t-structure are the same as stable ones.
The quiver determined by the S-graph is a special case of the graded quiver with
relations describing End(G) of a formal generator of F (X) (Section 3.4).
Mutation
We have assumed above that there are no horizontal saddle connections. Points in
the moduli space where this does happen, i.e. some stable object has real central charge,
are by definition on walls of the second kind. Let us consider a generic path γ : (−, ) →
M(X) which crosses one of these walls at time t = 0. There are two possibilities, either
the foliation of γ (0) has a single horizontal saddle connection, e, or some of the leaves are
closed loops foliating a cylinder. The latter is a more complicated kind of wall-crossing,
and we will restrict attention to the former case.
Making smaller, we may assume that the horizontal foliations of γ (t), t = 0, have
no finite-length leaves. Let G− , G+ be the S-graphs for γ (t) for t < 0, t > 0 respectively.
The saddle connection e becomes simple for small non-zero values of t, and so can be
identified with an edge of both G− and G+ . The graphs G− , G+ are related by mutation,
which we want to describe explicitly. There are two cases: Left mutation if the phase of
e is moving clockwise, right mutation if it is moving counterclockwise. Let us assume the
former. Looking at the horizontal strip decompositions one sees that S+ is obtained from
S− as follows. For a half-edge a denote by S(a) and S−1 (a) its successor and predecessor
respectively.
1. Let a, b be the half-edges of e. Decrement d(a, S(a)) and d(b, S(b)) by one.
2. If d(a, S(a)) = 0, then slide the end of S(a) along e so that it becomes the new
predecessor, c, of b with d(c, b) = 1. Otherwise, increment d(S−1 (b), b) by one.
3. Do the previous step with a and b switched.
FLAT SURFACES AND STABILITY STRUCTURES 311
The process is illustrated in Figure 7. Right mutation is then simply the inverse opera-
tion.
Remark. — In the special case when all zeros are simple we get an ideal triangulation
(for generic horizontal direction) by picking a single horizontal leaf from the interior of
each horizontal strip as in [23]. A mutation of the S-graph corresponds then to a flip of
the triangulation. In general however the mutations described here are not mutations in
the sense of cluster theory.
Proof. — The proof consists of two parts, first to show that M(S) ∼ = Cn as complex
manifolds, and then to show that the map M(S) → Stab(F (S)) is surjective.
1. We may assume that the interior of S is identified with C, and the grading given
by the standard horizontal foliation on C. Define an analytic map u : Cn → M(S) by
(6.3) (a0 , . . . , an−1 ) → exp zn+1 + an−1 zn−1 + · · · + a0 dz2 , a0 .
This requires some explanation. By the results of Section 2.5, the quadratic differentials
above define a flat structure on C, the completion of which is, as a topological space,
independent of the parameters ai . We assume that S is the real blow-up of this comple-
tion, so that we get a flat structure on S. To determine a point in M(S) we also need a
312 F. HAIDEN, L. KATZARKOV, M. KONTSEVICH
homotopy class of paths from the horizontal foliation of the quadratic differential to the
standard horizontal foliation on z, which is what a0 is used for. Since S is contractible,
such a homotopy class of paths is determined by its restriction to a single point, which we
take to be the origin, where a0 defines a path exp((1 − t)a0 )dz2 in (T∗0 C)⊗2 , t ∈ [0, 1].
An analytic automorphism of C acting on the set of polynomials of the form
z + an−1 zn−1 + · · · must be of the form z → ζ z, ζ n+1 = 1. All these extend to auto-
n+1
morphisms of S, but only for ζ = 1 is it isotopic to the identity. This implies injectivity
of u. For surjectivity, let S be a contractible flat surface with n + 1 conical points, all with
infinite cone angle. By the results of Section 2.5, S comes from a quadratic differential ϕ
with exponential singularities on a compact Riemann surface C. Since S is contractible,
C must be CP1 and ϕ must have a single exponential singularity. After applying a suit-
able Möbius transformation, the exponential singularity is at ∞ and ϕ is of the form
exp(zn+1 + an−1 zn−1 + · · · + a0 )dz2 .
2. Let σ ∈ Stab(F (S)) be a stability structure, A the heart of its t-structure. As
F (S) has only a finite number of indecomposable objects up to shift, A must be Artinian
with n = rkK0 (F (S)) simple objects E1 , . . . , En , up to isomorphism. Each Ei corresponds
to an isotopy class of graded arcs in S, which we denote by the same symbol. Since
Ext≤1 (Ei , Ej ) = 0 for i = j, we can choose representatives of the Ei which intersect only
at the endpoints. Because the Ei give a basis of H1 (S, ∂S; Zτ ), they necessarily form a
full formal system of graded arcs. Its structure is given by an S-graph which is a tree with
n edges. Rotating σ slightly by an angle , we may assume that no central charge lies
in R. Using Proposition 6.1 we get a flat surface X corresponding to σ , after rotating the
horizontal direction back by −. We already showed in the first part that X belongs to
M(S).
Proposition 6.3. — Let S be the graded marked surface of An -type, n > 1, then the canonical
map MCG(S) → Aut(F (S)) from the mapping class group of graded automorphisms to the group
of autoequivalences up to natural equivalence, is an isomorphism. The group is abelian, isomorphic to
Z2 /Z(2, n + 1).
Proof. — The assumption n > 1 guarantees that distinct graded boundary arcs of
S correspond to distinct isomorphism classes of objects in F (S). Since any element of
MCG(S), which are all represented by rotations, is determined by its action on graded
boundary arcs, this implies injectivity.
Note that (isomorphism classes of) objects E ∈ F (S) corresponding to graded
boundary arcs are intrinsically characterized, e.g. by being indecomposables with F (S)/E
equivalent to the bounded derived category of an An−1 quiver. Thus an automorphism
of F (S) induces a map on the set B of graded boundary arcs. B has a natural total
FLAT SURFACES AND STABILITY STRUCTURES 313
FIG. 8. — Possible topologies of K when n = 3 for a generic stability structure. Note that the number of stable objects up
to shift is 6, 5, 4, 3 respectively
Remark. — Fix a grading on each boundary arc of S and let B1 , . . . , Bn+1 be the
Yoneda-duals of the corresponding objects in F (S). Their sum is a functor B : F (S) →
C∼= Perf(Kn+1 ). In order to make the statement of the above proposition work for n = 1
one needs to replace Aut(F (S)) by the group of pairs of autoequivalences, one of F (S),
one of C , which are compatible with B, up to natural equivalence. Something like this
also seems to be needed in order to generalize the proposition to other marked surfaces.
In the context of Fukaya–Seidel categories, B is the restriction to the distinguished fiber.
Corollary 6.1. — Let S be the graded marked surface of An -type, n > 1, then
(6.4) Stab F (S) /Aut F (S) = M(S)/MCG(S)
as complex orbifolds.
n
6.3. Example: affine type A
Fix positive integers p, q and let n + 1 = p + q. Let S be a graded surface with
interior an annulus and p (resp. q) marked boundary components, each diffeomorphic
314 F. HAIDEN, L. KATZARKOV, M. KONTSEVICH
to an open interval, on the left (resp. right) end, and grading so that a simple closed
loop around the annulus has vanishing Maslov index, i.e. is gradable. Such S is uniquely
determined up to isomorphism. For concreteness we take S to be the real blow-up of
the completion of C∗ with respect to the metric |ϕ|, where ϕ = exp(zp + z−q )dz2 , and
grading given by the horizontal foliation of ϕ. The Fukaya category F (S) is equivalent to
the bounded derived category of any quiver with underlying graph the extended Dynkin
diagram An and p (resp. q) arrows directed counterclockwise (resp. clockwise).
n -type,
Theorem 6.2. — For S as above of A
(6.7) M(S) = Stab F (S) ∼= Cn+1 .
Proof. — As in the Dynkin case the proof consists of two parts, first to show
that M(S) = Cn+1 as complex manifolds, and then to show that the map M(S) →
Stab(F (S)) is surjective.
1. To a point (ap−1 , . . . , a−q ) ∈ Cn+1 we assign the flat surface given by the quadratic
differential
(6.8) exp zp + ap−1 zp−1 + · · · + a−q+1 z−q+1 + exp(a−q )z−q dz2
heart of the t-structure is Artinian. From here we can proceed as in the proof of Theo-
rem 6.1.
Moduli of objects
For a partially wrapped Fukaya category F of a surface S our classification de-
scribes Ob(F ) as a set in terms of curves on S. However, Ob(F ) has a much richer
structure of a derived stack.
Also, the case of Fukaya categories of closed surfaces remains. This would be an
A-side generalization of Atiyah’s classification for the elliptic curve. However, for g > 1
the category is only Z/2-graded and there are no t-structures or stability structures.
Problem. — Classify objects of F (S) when S is a closed surface with g(S) > 1, assuming this
is a tame problem.
Higher dimensions
There is at present no complete conjectural picture extending the results of this
paper to higher dimensional symplectic manifolds. Investigations into this direction were
started by Thomas [48] and Thomas–Yau [49] with the aim of finding an algebraic no-
tion of stability describing special Lagrangian submanifolds and their mean curvature
flow. More recently, their proposal has been updated by Joyce [28], taking into account
recent developments on Fukaya categories, mean curvature flow, and Bridgeland’s ax-
iomatics. The conjecture is that given a compact Calabi–Yau M one has a stability struc-
ture on its Fukaya category, F (M), such that singular special Lagrangians in M, with ad-
ditional data to make them objects in F (M), are (all?) stable objects. The central charge
is given by integration of the holomorphic volume form. Mean curvature flow applied to
an object in F (S), at least with sufficiently small phase variation, should converge to the
components of the Harder–Narasimhan filtration. It is however still an open problem to
describe (a component of) Stab(F (M)) in general, even conjecturally.
Acknowledgements
We also thank Mohammed Abouzaid, Denis Auroux, Tom Bridgeland, Igor Burban,
Tobias Dyckerhoff, Alexander Efimov, Alexander Goncharov, Dominic Joyce, Mikhail
Kapranov, Gabriel Kerr, Andrew Neitzke, Tony Pantev, Yan Soibelman, Pawel Sosna,
and Zachary Sylvan for discussions relating to this paper. Our thanks also go to the
anonymous referees for carefully reading the manuscript and providing helpful sugges-
tions. The authors were partially supported by the FRG grant DMS-0854977, the re-
search grants DMS-0854977, DMS-0901330, DMS-1201321, DMS-1201475, DMS-
1265230, OISE-1242272 PASI from the National Science Foundation, the FWF grant
P24572-N25, by an ERC GEMIS grant, and by an RF Government grand, ag. No.
14.641.31.0001. We are grateful to the IHES for providing perfect working conditions.
The authors were supported by two Simons Foundation grants.
REFERENCES
1. M. ABOUZAID, On the Fukaya categories of higher genus surfaces, Adv. Math., 217 (2008), 1192–1235.
2. M. ABOUZAID and P. SEIDEL, An open string analogue of Viterbo functoriality, Geom. Topol., 14 (2010), 627–718.
3. I. ASSEM and A. SKOWROŃSKI, Iterated tilted algebras of type Ãn , Math. Z., 195 (1987), 269–290.
4. M. F. ATIYAH, Vector bundles over an elliptic curve, Proc. Lond. Math. Soc., 3 (1957), 414–452.
5. D. AUROUX, Fukaya categories and bordered Heegaard–Floer homology, in Proceedings of the International Congress of
Mathematicians, vol. II, pp. 917–941, Hindustan Book Agency, Gurugram, 2010.
6. M. BARDZELL, The alternating syzygy behavior of monomial algebras, J. Algebra, 188 (1997), 69–89.
7. A. BAYER, E. MACRÌ and Y. TODA, Bridgeland stability conditions on threefolds I: Bogomolov-Gieseker type inequali-
ties, J. Algebraic Geom., 23 (2014), 117–163.
8. V. BEKKERT and H. A. MERKLEN, Indecomposables in derived categories of gentle algebras, Algebr. Represent. Theory, 6
(2003), 285–302.
9. R. BOCKLANDT, Noncommutative mirror symmetry for punctured surfaces, Trans. Am. Math. Soc., 368 (2016), 429–469.
10. J. P. BOWMAN and F. VALDEZ, Wild singularities of flat surfaces, Isr. J. Math., 197 (2013), 69–97.
11. T. BRIDGELAND, Stability conditions on triangulated categories, Ann. Math., 166 (2007), 317–345.
12. T. BRIDGELAND, Stability conditions on K3 surfaces, Duke Math. J., 141 (2008), 241–291.
13. T. BRIDGELAND, Stability conditions and Kleinian singularities, Int. Math. Res. Not., 21 (2009), 4142–4157.
14. T. BRIDGELAND, Y. QIU and T. SUTHERLAND, Stability conditions on the A2 quiver, arXiv:1406.2566.
15. T. BRIDGELAND and I. SMITH, Quadratic differentials as stability conditions, Publ. Math. IHÉS, 121 (2015), 155–278.
16. I. BURBAN and Y. DROZD, On derived categories of certain associative algebras, in Representations of Algebras and Related
Topics, Fields Inst. Commun., vol. 45, pp. 109–128, Am. Math. Soc., Providence, 2005.
17. I. BURBAN and B. KREUSSLER, Derived categories of irreducible projective curves of arithmetic genus one, Compos.
Math., 142 (2006), 1231–1262.
18. G. DIMITROV, F. HAIDEN, L. KATZARKOV and M. KONTSEVICH, Dynamical systems and categories, in The Influence of
Solomon Lefschetz in Geometry and Topology: 50 Years of Mathematics at CINVESTAV, Con. Math., vol. 621, pp. 133–170, Am.
Math. Soc., Providence, 2014.
19. G. DIMITROV and L. KATZARKOV, Stability conditions on the acyclic triangular quiver, arXiv:1410.0904.
20. T. DYCKERHOFF, A1-homotopy invariants of topological Fukaya categories of surfaces, arXiv:1505.06941.
21. T. DYCKERHOFF and M. KAPRANOV, Triangulated surfaces in triangulated categories, arXiv:1306.2545.
22. K. FUKAYA, Y.-G. OH, H. OHTA and K. ONO, Lagrangian Intersection Floer Theory, Anomaly and Obstruction, Parts I, II,
AMS/IP Studies in Adv. Math., vol. 46.1, 2009.
23. D. GAIOTTO, G. W. MOORE and A. NEITZKE, Wall-crossing, Hitchin systems, and the WKB approximation, Adv. Math.,
234 (2013), 239–403.
24. A. GROTHENDIECK, Groupes de classes des categories abeliennes et triangulees. Complexes parfaits (Redige par I.
Bucur), in Semin. Geom. Algebr. Bois-Marie 1965–1966, SGA 5, Lect. Notes Math., vol. 589, Expose No. VIII, pp. 351–
371, 1977.
FLAT SURFACES AND STABILITY STRUCTURES 317
25. J. L. HARER, Stability of the homology of the mapping class groups of orientable surfaces, Ann. Math. (2), 121 (1985),
215–249.
26. J. L. HARER, The virtual cohomological dimension of the mapping class group of an orientable surface, Invent. Math.,
84 (1986), 157–176.
27. A. IKEDA, Stability conditions on CYn categories associated to An -quivers and period maps, arXiv:1405.5492.
28. D. JOYCE, Conjectures on Bridgeland stability for Fukaya categories of Calabi–Yau manifolds, special Lagrangians,
and Lagrangian mean curvature flow, arXiv:1401.4949.
29. H. KAJIURA, K. SAITO and A. TAKAHASHI, Matrix factorization and representations of quivers. II. Type ADE case, Adv.
Math., 211 (2007), 327–362.
30. M. KONTSEVICH, Homological algebra of mirror symmetry, in Proceedings of the International Congress of Mathematicians,
vols. 1, 2, Zürich, 1994, pp. 120–139, Birkhäuser, Basel, 1995.
31. M. KONTSEVICH and Y. SOIBELMAN, Stability structures, motivic Donaldson-Thomas invariants and cluster transfor-
mations, arXiv:0811.2435.
32. M. KONTSEVICH and Y. SOIBELMAN, Notes on A∞ -algebras, A∞ -categories and non-commutative geometry, in Homo-
logical Mirror Symmetry, Lecture Notes in Phys., vol. 757, pp. 153–219, Springer, Berlin, 2009.
33. V. LYUBASHENKO and S. OVSIENKO, A construction of quotient A∞ -categories, Homol. Homotopy Appl., 8 (2006), 157–
203.
34. E. MACRÌ, Stability conditions on curves, Math. Res. Lett., 14 (2007), 657–672.
35. H. MASUR and J. SMILLIE, Hausdorff dimension of sets of nonergodic measured foliations, Ann. Math. (2), 134 (1991),
455–543.
36. D. NADLER, Cyclic symmetries of An -quiver representations, arXiv:1306.0070.
37. L. A. NAZAROVA and A. V. ROĬTER, A certain problem of I.M. Gel’fand, Funkc. Anal. Prilozh., 7 (1973), 54–69.
38. R. NEVANLINNA, Über Riemannsche Flächen mit endlich vielen Windungspunkten, Acta Math., 58 (1932), 295–373.
39. S. OKADA, Stability manifold of P1 , J. Algebraic Geom., 15 (2006), 487–505.
40. J. PASCALEFF and N. SIBILLA, Topological fukaya category and mirror symmetry for punctured surfaces, arXiv:1604.
06448.
41. A. POLISHCHUK and E. ZASLOW, Categorical mirror symmetry: the elliptic curve, Adv. Theor. Math. Phys., 2 (1998),
443–470.
42. P. SEIDEL, Fukaya A∞ structures associated to Lefschetz fibrations. II, arXiv:1404.1352.
43. P. SEIDEL, Graded Lagrangian submanifolds, Bull. Soc. Math. Fr., 128 (2000), 103–149.
44. P. SEIDEL, Fukaya Categories and Picard-Lefschetz Theory, Zurich Lectures in Advanced Mathematics, European Mathe-
matical Society (EMS), Zürich, 2008.
45. N. SIBILLA, D. TREUMANN and E. ZASLOW, Ribbon graphs and mirror symmetry, Sel. Math. New Ser., 20 (2014), 979–
1002.
46. I. SMITH, Quiver algebras as Fukaya categories, arXiv:1309.0452.
47. K. STREBEL, Quadratic Differentials, Ergebnisse der Mathematik und ihrer Grenzgebiete (3) [Results in Mathematics and
Related Areas (3)], vol. 5, Springer, Berlin, 1984
48. R. P. THOMAS, Moment maps, monodromy and mirror manifolds, in Symplectic Geometry and Mirror Symmetry, Seoul, 2000,
pp. 467–498, World Scientific, River Edge, 2001.
49. R. P. THOMAS and S.-T. YAU, Special Lagrangians, stable bundles and mean curvature flow, Commun. Anal. Geom., 10
(2002), 1075–1113.
50. W. A. VEECH, Flat surfaces, Am. J. Math., 115 (1993), 589–689.
F. H.
Department of Mathematics, Science Center,
Harvard University,
One Oxford Street,
Cambridge, MA 02138, USA
[email protected]
318 F. HAIDEN, L. KATZARKOV, M. KONTSEVICH
L. K.
Fakultät für Mathematik,
Universität Wien,
Oskar-Morgenstern-Platz 1,
1090 Wien, Austria
[email protected]
and
HSE Moscow,
Moscow, Russia
M. K.
Institut des Hautes Études Scientifiques,
35 route de Chartres,
91440 Bures-sur-Yvette, France
[email protected]