Weinstein Manifolds Revisited: Yakov Eliashberg
Weinstein Manifolds Revisited: Yakov Eliashberg
Yakov Eliashberg∗
arXiv:1707.03442v2 [math.SG] 29 Aug 2017
Abstract
This is a very biased and incomplete survey of some basic notions, old and new results, as
well as open problems concerning Weinstein symplectic manifolds.
1
1 Weinstein manifolds, domains, cobordisms 2
Xb → R such that X 0 ⊂ X b →X
b is the image of X under the time 1 map ψ : X b of
the flow of the vector field hZ on the completion X.
b The completions of the radially
0
equivalent Liouville domains L and L are canonically isomorphic.
The space of Liouville structures for (X, ω) is convex, and hence any two Liouville
structures are canonically homotopic. Given a homotopy of completed Liouville
structures (X, b → X
b ωt , λt ) there exists an isotopy φt : X b such that φ∗ ωt = ω0 .
t
Moreover, one can always arrange that φ∗t λt = λ0 + dHt , see [11], Sections 11.1 and
11.2. In particular on completed Liouville manifolds it is always sufficient to con-
sider homotopies fixing the symplectic form, and, moreover, changing the Liouville
form by adding an exact form. Homotopic non-completed Liouville domains are
symplectomorphic up to radial deformation.
Given a Liouville domain L = (X, ω, λ) consider a compact set
\
Core(L) = Z−t (X),
t>0
the attractor of the negative flow of the Liouville vector field Z. We will call Core(L)
the core, or the skeleton of the Liouville structure L. While Core(L) has obviously its
2n-dimensional Lebesgue measure equal to 0, it still can be pretty large if no extra
conditions are imposed on the Liouville structure. For instance, McDuff constructed
in [31] a Liouville structure on T ∗ Sg \ Sg for a closed surface Sg of genus g > 1, whose
core has codimension 1.
However, the situation changes if one requires existence of a Lyapunov function for
the Liouville vector field Z. A Weinstein structure on a domain X is a Liouville
structure L together with a function φ : X → R which is Lyapunov for the Liouville
vector field Z, i.e.
(L1) dφ(Z) > c||Z||2 for a positive constant c and some Riemann metric on X.
Note that condition (L1) implies that Core(X, λ) is the union of Z-stable manifolds
of critical points of φ (i.e. points converging to the critical locus in forward time).
In [11] it was required in addition that φ is either Morse or generalized Morse (i.e.
may have death-birth critical points). Under these assumptions it was shown in [11],
see also [20, 15], that
(L2) the core is stratified by isotropic for λ, and hence for ω submanifolds.
F. Laudenbach proved, see [30], that if the flow of Z is Morse-Smale (i.e. stable and
unstable manifolds of critical points intersect transversely) and near critical points
1 Weinstein manifolds, domains, cobordisms 3
Core(Σ)
the vector field Z is gradient with respect to an Euclidean metric, then the skeleton
can be further Whitney substratified. It is likely that the Whitney condition also
holds if near its zeroes the vector field Z is gradient with respect to any Riemannian
metric. However, as far as know, this was never verified in the literature. The
Whitney condition need not hold if eigenvalues of the linearization of Z at critical
points have non-vanishing imaginary parts, as a spiraling phenomenon of trajectories
may occur.1
Condition (L2) holds for a much more general class of taming functions (e.g. when
φ is Morse-Bott), and hence for the the purposes of this paper we will take the
following working definition of a Weinstein structure, extending the class considered
in [11]: W = (X, λ, Z, φ) is Weinstein if it satisfies conditions (L1) and (L2) with
the Whitney condition and also condition
function with respect to some (not necessarily integrable) almost complex structure
J, see [11], Chapter 1, for the details.
Remark 1.2. Note that not every closed subset C of a symplectic manifold which is
stratified by isotropic strata may serve as the skeleton for an appropriately chosen
Weinstein structure on a neighborhood of C (compatible with the given ambient
symplectic form). Examples of this kind exist already in R2 . For instance, let
C := {x = 0, y ≥ 0} ∪ {x = y 2 , y ≥ 0} ∪ {y = 0, x ≥ 0} ∪ {y = x2 , x ≥ 0}
be the union of 4 arcs emanating from the origin. Then there is no Liouville structure
on a neighborhood U 3 0 which has C ∩ U as a part of its skeleton. Indeed it is
straightforward to check that for any 1-form λ vanishing on C ∩ U one has (dλ)0 = 0.
Problem 1.3. Find a necessary and sufficient condition on a compact subset C of
a symplectic manifold to serve as the skeleton of some
a) Liouville, or
b) Weinstein
+W
2 2
W W
−W
Fig. 1.2: Sutured Weinstein cobordism W with corners.
(ii) Pages of open books. According to Giroux’s theorem [25], any contact manifold
admits an open book decomposition whose pages are Weinstein hypersurfaces.
Problem 2.3. Is there an analog of the Legendrian algebra LHA(Λ) for a general
Weinstein hypersurface?
Let us return to the case of the Legendrian homology algebra of a Legendrian sub-
manifold Λ and pick a contact form λ such that its Reeb vector field is tangent to
the contact submanifold ∆ := ∂Σ(Λ). We also choose an almost complex structure
J on ξ such that ξ ∩ T (∆) are J-invariant. This allows us to define a deformation
(A[t], D) of the Legendrian differentialP
algebra (A, ∂) := LHA(Λ) as follows. For a
generating chord c ∈ A define D(c) = (∂k c)tk , where ∂0 = ∂ and ∂k c counts holo-
k≥0
morphic curves with the intersection index k with the symplectization of ∆. This
symplectization is a complex hypersurface in the symplectization of Y , and hence
k ≥ 0. The sum defining differential D is finite due to the Gromov compactness.
Problem 2.4. Explore whether the above construction yields a genuinely new in-
variant of a Legendrian submanifold.
3
I thank Sheel Ganatra and Tobias Ekholm for the discussion of this problem.
2 Weinstein hypersurfaces and Weinstein pairs 8
Let us first recall a few basic facts about convex hypersurfaces in contact manifolds.
If a germ ξ of a contact structure along a closed hypersurface V in a (2n − 1)-
dimensional manifold admits a transverse contact vector field v then we canonically
can construct a contact structure ξb on V × R which is invariant with respect to
translations along the second factor and whose germ along any slice V × t, t ∈ R, is
isomorphic to ξ. We will call ξb the invariant extension of the convex germ ξ.
Lemma 2.6. Let V be a closed (2n − 2)-dimensional manifold and ξ a contact
structure on Y = V ×[0, ∞) which admits a contact vector field v inward transverse to
V ×0 and such that its trajectories intersecting V ×0 fill the whole manifold Y (we do
not require v to be complete). Then (Y, ξ) is contactomorphic to (V ×[0, ∞), ξ),
b where
ξb is the invariant extension of the germ of ξ along V × 0. Moreover, for any compact
set C ⊂ Y , Int C ⊃ V ×0, there exists a contactomorphism h : (Y, ξ) → (V ×[0, ∞), ξ)
b
which is equal to the identity on V × 0 and which sends the contact vector field v|C
∂
to the vector field ∂t .
vector field obtained by cutting off ve1 outside C1 but inside C2 and denote by h2
the time T2 > T1 + 1 flow of v2 , where T2 is chosen such that h2 (V × 0) ⊂ C2 \ C1 .
Denote by C e1 the domain bounded by V × 0 and h2 (V × 0) and by ve2 the contact
e1 and to the push-forward vector field (h1 )∗ ve1 on Y \ C
vector field equal to v2 on C e1 .
Continuing this process we construct a sequence of contact vector fields ve1 , ve2 , . . . ,
which stabilize on compact sets C1 , C2 , . . . and converge to the contact vector field
ve on Y with the required properties.
∂
Proof of Proposition 2.5. The contact vector field v = −ZΣ − u ∂u is transverse to
∂U (σ) and retracts U (Σ) to Core(Σ, λΣ ), and hence the contact structure on U (Σ) \
Core(Σ, λΣ ) is canonically isomorphic to ∂U (Σ) × [0, ∞) endowed with the invariant
extension ξb of the germ of contact structure
Sξ along ∂U (Σ). On the other hand, v is
transverse to ∂Uδ (σ) for each δ ≤ ε and ∂Uδ (Σ) = U (Σ) \ Σ. Hence, applying
δ∈(0,ε]
Lemma 2.6 we conclude that (U (Σ) \ Σ, ξ) is contactomorphic to (∂U (Σ) × [0, ∞), ξ),
b
and the claim follows.
Remark 2.7. One of the corollaries of Lemma 2.6 is that any open domain in the
2n+1
P
standard contact (R , dz + Pxi dyi − yi dxi ) which is star-shaped with respect to
∂
the contact vector field 2z ∂z + xi ∂x∂ i + yi ∂y∂ i is contactomorphic to R2n+1 . On the
other hand, in the standard contact R3 any open domain diffeomorphic to R3 is
contactomorphic to R3 , see [17].
Problem 2.8. Is there a domain in the standard contact R2n+1 , n > 2, which is
diffeomorphic to the closed ball, has convex in contact sense boundary, but whose
interior is not contactomorphic to the standard R2n+1 ? Or even are there any open
domains in the standard contact R2n+1 , n > 2, which are diffeomorphic but not
contactomorphic to R2n+1 ?
Weinstein pairs
A Weinstein pair (W, Σ) consists of a Weinstein domain W = (X, λ, φ) together with
a Weinstein hypersurface (Σ, λ|Σ ) in its boundary ∂X. Equivalently, a Weinstein pair
can be viewed as a Weinstein manifold with cylindrical end, together with a Weinstein
hypersurface in its ideal contact boundary.
2 Weinstein hypersurfaces and Weinstein pairs 10
be its saturation by the trajectories of the Liouville vector field Z. The union
Core(X, Σ) := Core(X) ∪ Λ
b
Proposition 2.9. Given a Weinstein pair (W, Σ), W = (X, λ, φ), there exist a
Liouville form λ0 for ω and a function φ0 : X → R such that
To construct the adjusted Liouville field Z0 let us write the form λ near ∂X as
s(du + λΣ ) near U (Σ). Note that the Hamiltonian vector field Y for a function
∂ ∂
su near U (Σ) coincides with −s ∂s + u ∂u + ZΣ , and hence by appropriately cutting
off the function su outside a neighborhood of U (Σ) and subtracting the differential
dg of the resulting function g to the Liouville form λ we get the Liouville form λ0
with the required properties. Note that the form λ0 |U is no more contact. Instead,
λ0 |U = π ∗ (λ|Σ ).
Suppose that λ0 , φ0 are adjusted to the Weinstein pair (W, Σ). Recall that φ0 |∂U (Σ) =
ε2 . Denote X0 = {φ0 ≤ ε2 }. We note that φ0 has no critical points in X \ Int X0 ,
and hence X0 is a manifold with boundary with a corner along ∂U (Σ) which is
homeomorphic to X. We will sometime refer to (X0 , λ0 , φ0 ) as the cornered version
of the Weinstein pair (W, Σ).4 For instance, the cornered version of the standard
Weinstein ball B 2n is the cotangent ball bundle of Dn . Thus, it is always possible
to go back and forth between the original and adjusted (cornered) versions of a
Weinstein pair, and we will be using the term “Weinstein pair” for both versions.
Ζ Σ Ζ0 Σ Ζ− Σ
Remark 2.10. There are several other useful adjustments of a Weinstein pair struc-
ture. Ekholm and Lekili in [14], Section B.3, are doing a similar to the cornered
version construction by deforming the boundary ∂X near U (Σ) without changing Z,
as on Fig. 2.2. Without defining here Sylvan’s stop structure we just say that for a
given Weinstein pair there is a contractible space of choices of stop structures on the
completion.
One can also transform a Weinstein pair into a Weinstein cobordism whose negative
boundary is U (Σ), see Fig. 2.1:
4
The completion of the cornered version of a Weinstein pair is a special case of a Liouville sector
in the sense of [26].
3 Operations on Weinstein pairs 12
- there exists a hypersurface (S, ∂S) ⊂ (P, ∂P ) which is Weinstein for the restricted
Liouville form λ|S , tangent to the vector field Z and intersects all leaves of
the characteristic foliation F of the hypersurface P ; we will refer to S as the
Weinstein soul of the splitting hypersurface P and denote it by Soul(P ).
3 Operations on Weinstein pairs 13
Note that the latter condition together with Lemma 2.6 imply that P is contacto-
morphic to the contact surrounding of its Weinstein soul.
It follows that (S, λ|S , φ|S ; ∂S) is a codimension two Weinstein subdomain of X and
Core(S, λ|S , φ|S ) = Core(W) ∩ P. Moreover, (W± ; S), where W± := (X± , λ|X± , φ|X± )
are cornered Weinstein pairs and Core(W± ; S) = Core(W) ∩ X± .
The gluing construction reverses the splitting. This operation was considered by
Avdek in [3] in the context of Liouville hypersurfaces. Let (W, Σ) and (W0 , Σ0 )
be two Weinstein pairs and (X0 , λ0 , φ0 ), (X00 , λ00 , φ00 ) their cornered forms. Let F :
(Σ, λ|Σ , φ|Σ ) → (Σ0 , λ0 |Σ0 , φ0 |Σ0 ) be a Weinstein isomorphism. We extend F to a
contactomorphism U (Σ) → U (Σ0 ), still denoted by F , and use it to define a domain
Then the Liouville forms λ0 and λ00 , as well as Lyapunov functions φ0 : X0 → R and
φ00 : X00 → R, can be glued together to define a Weinstein structure (W, Σ) ∪(W0 , Σ0 ) :=
F
(XF , λF , φF ), see Fig. 3.1.
F
X Σ Σ X U( Σ ) =F(U( Σ))
XU X
F
Note that
Core(XF , λF , φF ) = Core(X, Σ) ∪ Core(X0 , Σ0 ).
F|Core(Σ)
Note that the constructed Weinstein domain XF contains U (Σ) as its splitting hy-
persurface. Applying the above described splitting construction we get back the
Weinstein pairs (W, Σ) and (W0 , Σ0 ).
3 Operations on Weinstein pairs 14
In the case when (X 0 , Σ0 ) is the Weinstein pair (T ∗ Dk , T ∗ S k−1 ) the product operation
is called the stabilization (or k-stabilization). It was first proposed in a slightly
different form by M. Kontsevich, [28]. The core of the k-stabilized pair (W, Σ) is
equal to Core(W, Σ) × Dk .
It is important to stress the point that the result of the stabilization is always a
Weinstein pair with a non-empty hypersurface in the boundary, even if we begin
with the absolute case of a Weinstein domain.
Proof. Note that the corresponding to λ e Liouville vector field is given by the formula
∂
Ze = Zt + (u + ḣt ) ∂u , where Zt is the Liouville vector field corresponding to λt . Define
the function φe by the formula φe = φt + k2 (u + ḣt )2 , where a positive constant k will
be chosen later. Then we have
Not that |dφt (Zt ) ≥ a||Zt ||2 and |dḣt (Zt )| ≤ b||Zt || for some constants a, b > 0.
Denoting X := ||Zt ||, Y := u + ḣt we can write
The stable manifold of a critical point (x0 , t0 , u0 ) projects to the stable manifold of the
critical point x0 of φt0 . Its u-coordinate can be found by solving the inhomogeneous
linear ODE
du(γ(s))
= u(γ(s)) + ḣt0 (γ(s))
ds
with the asymptotic boundary condition lim u(γ(s)) = u0 , where γ(s) is a trajectory
s→∞
of Xt0 converging to the critical point x0 .
At first glance Theorem 4.1 implies that symplectic topology of flexible Weinstein
4 Looseness and Flexibility 18
manifolds is quite boring. This is also confirmed by the fact that symplectic homology
in all its flavors of a flexible Weinstein manifold is trivial. However, as we will see
below in Section 7 the contact boundaries of flexible Weinstein domains have a rich
contact topology.
The looseness property of a Legendrian submanifold can be naturally extended to
Weinstein hypersurfaces of contact manifolds. A Weinstein hypersurface Σ of a
contact manifold Y of dimension 2n + 1 ≥ 5 is called loose if for each n-dimensional
strata S of the skeleton Core(Σ) there is a ball BS ⊂ Y \ (Core(Σ) \ S) such that
BS ∩ S is loose in BS relative ∂(BS ∩ S). A canonical Weinstein thickening of a loose
Legendrian knot is loose. However, it is unclear whether looseness is preserved under
Weinstein isotopy.
(W, Σ) × (W0 , Σ0 ) := (X × X 0 , λ ⊕ λ0 , Σ
e := (Σ × X 0 ; Σ × Σ0 ) t(X × Σ0 , Σ × Σ0 ))
Id
e is loose in ∂(X × X 0 ).
be their product. Suppose that Σ is loose in ∂X. Then Σ
Indeed, this is straightforward from the following fact: given any contact manifold
(Y, {α = 0}), a Liouville manifold (U, µ), a loose Legendrian Λ ⊂ Y and a Lagrangian
L ⊂ U with µ|L = 0, then the Legendrian Λ × L ⊂ (Y × U, {α ⊕ µ = 0}) is loose as
well.
Let us stress the point that while flexibility of a Weinstein manifold is its intrinsic
property, the looseness of a Weinstein hypersurface depends on its embedding in the
contact manifold. However, the above fact about the looseness of a product shows
that flexibility always implies looseness (I thank the referee for this argiment).
Problem 4.5. Suppose that the stabilization of a Weinstein pair is flexible. Does this
imply that the Weinstein pair itself is flexible? More generally, does existence of a
homotopy between stabilizations of two Weinstein (pair) structures implies existence
of a homotopy between the structures themselves?
- Σ− is loose in ∂X and
Then the glued pair (W, Σ) ∪ 0 (W0 , Σ0 ) is flexible. In particular, the result of gluing
F,T,T
of two flexible Weinstein domains along Weinstein hypersurfaces one of which is
loose is flexible.
5 Lagrangian submanifolds of Weinstein domains 20
This follows from the fact that the gluing operations of two Weinstein pairs can be
decomposed into a sequence of handle attachments, and the looseness assumption
for the Weinstein hypersurface in one of the glued parts implies that all the critical
handles are attached along loose knots.
As a corollary Proposition 4.6 implies the following generalization of the following
result of E. Murphy and K. Siegel, [37]:
Proposition 4.7. The product of two Weinstein pairs, one of which is flexible, is
flexible.
Indeed, the product of two Weinstein pairs can always be built by a sequence of
gluing of various stabilizations of the first pair.
While flexible Weinstein structures enjoy a full parametric h-principle, there is plenty
of symplectic rigidity and fine symplectic invariants of non-flexible ones. I will not
discuss in this survey any such invariants and just mention that until recently most
examples of formally homotopic but not symplectomorphic Weinstein manifolds were
distinguished by their (possibly appropriately deformed) symplectic cohomology. For
instance, there are infinitely many non-symplectomorphic Weinstein structures on
R2n for any n > 2 ([41, 33]) and by taking connected sums of these examples with
flexible Weinstein manifolds one gets infinitely many non-symplectomorphic Wein-
stein structures on any given “almost Weinstein” (i.e. an almost complex manifold
of homotopy type of a half-dimensional CW-complex) manifolds, see [2].
Note that Theorem 5.2 can also be used for constructing exotic Weinstein structures.
In particular,
Theorem 6.1 ([18]). Let L be a closed 3-manifold. Then there exists a unique
up so symplectomorphism Weinstein structure W(L) = (ωL , ZL , φL ) on T ∗ S 3 which
contains L as its flexible Lagrangian submanifold in the homology class of the 0-
section (with Z/2-coefficients in the non-orientable case). Moreover, infinitely many
of these W(L) are pairwise non-symplectomorphic.
Note that there exists only 1 homotopy class of almost complex structures on T ∗ S 3 .
While the symplectic structure of W(L) carries a lot of information about the topol-
ogy of L, the following problem is open:
Problem 6.2. Suppose W(L) is symplectomorphic to W(L0 )? Does it imply that L
is diffeomorphic to L0 ?
Problem 6.3. Can the uniquenes results from [1] be extended to a more general
class of Weinstein structures on T ∗ S n ?
The proof of Theorem 5.2 yields also the following slightly stronger result.
(X, ω, λ, φ) → (T ∗ S 3 , ωX , λX , φX )
Theorem 7.1 ([31]). Any Weinstein filling of the standard contact sphere (S 2n−1 , ξstd )
is diffeomorphic to the ball B 2n .
7 Topology of Weinstein fillings 23
In fact, both Theorems 7.1 and 7.2 hold in a stronger form for a more general class
of symplectic, and not necessarily Weinstein fillings. We also note that while it fol-
lows from Theorem 4.1 that all completed subcritical Weinstein fillings of a given
contact manifold are symplectomorphic (we note that the (n − 1)-connectedness of
the inclusion map ∂X ,→ X implies that the homotopy class of an almost complex
structure on a subcritical manifold is determined by the homotopy class of its re-
striction to the boundary), it is unknown for n > 2 whether all completed fillings of
a contact manifold admitting a subcritical filling (e.g. the standard contact sphere)
are symplectomorphic.
The following theorem of Oleg Lazarev constrains topology of flexible Weinstein
manifolds.
Theorem 7.3 ([29]). All flexible fillings of of a contact manifold (Y, ξ) with c1 (Y, ξ) =
0 have canonically isomorphic integral homology.
There are further constraints on the topology of flexible Weinstein fillings. In par-
ticular,
Theorem 7.5 ([19]). Let (S 4n−1 , ξ) be a flexibly fillable contact structure. Then the
signature of its flexible filling is uniquely determined by the contact structure ξ.
8 Nadler’s program of arborealization 24
We note that the diffeomorphism type of X together with the homotopy class [ω]
determine a flexible Weinstein structure up to Weinstein homotopy, and hence the
positive answer to a) and b) would imply that the contact structure (Y, ξ) remember
the symplectomorphism type of the completion of its flexible filling.
+ −
Fig. 8.1: Arboreal singularities labeled by rooted decorated trees. The picture repre-
sents Lagrangian skeleta themselves, and not their front projections. Free
boundaries of vertical strata form Legendrian trees, while their traces at
the horizontal plane are fronts of these trees.
P
λ|Πj = dqNj +1 + pi dqi is a contact form. Cyclically ordering coordinates
i∈{1,...,n},i6=Nj +1
qNj +2 , . . . , qn , q1 , . . . , qNj and taking the coordinate qNj +1 as z we identify Πj with
T ∗ Rn−1 × R. Consider Asign(σj ) (Tj , εj , n − 1) ⊂ Πj .
Denote
k
[
∗ n bsign(σj ) (Tj , εj , n − 1))}.
B(T, ε, n) := {(tp, q) ∈ T R ; t ∈ [0, ∞), (p, q) ∈ A
j=1
Finally, consider the right models on Fig. 8.1. The models are contained in the
standard symplectic R4 with canonical coordinates (p1 , q1 , p2 , q2 ), and we have Π =
Π1 = {p1 = 1} The tree T1 in this case consists of two vertices, and identifying Π
with the standard symplectic R2 , we find that
b± (T1 , 1) = {q1 = p2 = 0} ∪ {q1 = ∓p22 , p2 = ±q2 , p2 ≥ 0}.
A
8 Nadler’s program of arborealization 28
Note that the second stratum in the union can also be written as {q1 = ∓q22 , p2 =
±q2 , p2 ≥ 0} Thus we have
q22
B(T, +1, 2) = {p2 = 0, q1 = 0, p1 ≥ 0} ∪ {q1 = − , p2 = p1 q2 , p1 , p2 ≥ 0},
2
q22
B(T, −1, 2) = {p2 = 0, q1 = 0, p1 ≥ 0} ∪ {q1 = , p2 = −p1 q2 , p1 , p2 ≥ 0}
2
q2
Note that B(T, ±1, 2) ∩ {p = 0} = {q2 = 0} ∪ {q1 = ∓ 22 } is the front of the
Legendrian tree Ab± (T1 , 1), while B(T, ±1, 2) is the positive conormal of this front
co-oriented by the vector field ∂q∂1 .
A general arboreal singularity is associated to a double decorated rooted tree with
an additional decoration β which assigns 0 or 1 to all terminal vertices of the tree T .
We extend β to all vertices by setting β(v) = 0 for all non-terminal vertices. Primary
arboreal singularities correspond to the case when the decoration β is identically 0.
P
We denote |β| := β(v), where the sum is taken over all terminal vertices v of
the tree T . With each double decorated tree (T, ε, β) we associate a unique up to
symplectomorphism model A(T, ε, β, m) ⊂ T ∗ Rm for each m ≥ |T | + |β| − 1. In
dimension m ≥ n := |T | + |β| − 1 we have
A(T, ε, β, m) = A(T, ε, β, n) × Rm−n ⊂ T ∗ Rn × T ∗ Rm−n = T ∗ Rm .
Note that the model A(T, ε, β, m) inherits a smooth structure (i.e. the algebra of
smooth functions) from the ambient space R2n . By an n-dimensional arboreal com-
plex we mean a set covered by charts diffeomorphic to one of the models A(T, ε, β, n).
Hence, every arboreal complex can be canonically stratified by strata ST,ε,β of dimen-
sion n − |T | − |β| + 1. A diffeomorphism f : C → C 0 between two arboreal complexes
induces a diffeomorphism between the corresponding strata, but not every contin-
uous map f : C → C 0 which is a diffeomorphism on the corresponding strata is a
diffeomorphism of arboreal complexes C and C 0 .
8 Nadler’s program of arborealization 29
Under some topological constraints on the manifold X one can further restrict the
list of necessary singularities.
References
[5] J. Bowden, D. Crowley and A. Stipsicz, The topology of Stein fillable manifolds
in high dimensions I, arXiv:arXiv:1306.2746.
[10] K. Cieliebak, Handle attaching in symplectic homology and the Chord Conjec-
ture, J. Eur. Math. Soc.. 4(2002), 115–142.
[11] K. Cieliebak and Y. Eliashberg, From Stein to Weinstein and Back – Symplectic
Geometry of Affine Complex Manifolds, Colloquium Publications Vol. 59, Amer.
Math. Soc. (2012).
[13] T. Ekholm, Rational SFT, linearized Legendrian contact homology, and La-
grangian Floer cohomology, Perspectives in analysis, geometry, and topology,
109–145, Progr. Math., 296, Birkhäuser/Springer, New York, 2012.
[14] T. Ekholm and Y. Lekili, Duality between Lagrangian and Legendrian invariants,
arXiv:1701.01284.
[22] Y. Eliashberg and E. Murphy, Lagrangian caps, Geom. and Funct. Anal.,
23(2013), Volume 23, 1483–1514.
[31] D. McDuff, Symplectic manifolds with contact type boundaries, Invent. Math.
103(1991), 651–671.
8 Nadler’s program of arborealization 32
[32] M. Maydanskiy and P. Seidel, Lefschetz fibrations and exotic symplectic struc-
tures on cotangent bundles of spheres, J. Topol. 3 (2010), 157–180.
[34] M. Kervaire, and J.W. Milnor, Groups of homotopy spheres: I. Ann. of Math.,
77(3), 504–537.
[41] P. Seidel and I. Smith, The symplectic topology of Ramanujams surface, Com-
ment. Math. Helv. 80(2005), 859–881.
[43] S. Smale, On the structure of manifolds, Amer. J. Math., 84(1962) pp. 387–399.