Symplectic Hodge Theory On Lie Algebroids: Abstract
Symplectic Hodge Theory On Lie Algebroids: Abstract
Symplectic Hodge Theory On Lie Algebroids: Abstract
SHANE RANKIN
Abstract. We explore the natural analogues of the Brylinksi condition, Strong Lefschetz condition, and
dδ-lemma in Symplectic Geometry originally explored by Brylinksi, Mathieu, Yan, and Guillemin to the
Symplectic Lie Algebroid case. The equivalence of the three conditions is re-established as a purely algebraic
statement along with a primitive notion of the dδ-lemma shown establsihed by Tseng, Yau, and Ho. We then
arXiv:2411.06012v1 [math.SG] 8 Nov 2024
show that the natural analogues of these in the Lie Algebroid setting holds as well with examples given
Contents
1. Introduction 1
1.1. Background 1
1.2. Results 2
2. Lefschetz Modules 3
2.1. The Lie Superalgebra 3
2.2. Equivalence of Surjectivity 6
2.3. Equivalence of Triples 7
2.4. Weakly Lefschetz Modules 10
2.5. Primitive Equivalence 11
3. Lefschetz Algebroids 14
3.1. Symplectic Lie Algebroids 14
3.2. Kahler-Weil Identities 17
3.3. Cohomology with Coefficients 19
4. Examples 21
4.1. Kodaria-Thurston Manifold 21
4.2. Symplectic Lie Algebras of Dimension Four 21
4.3. Six-Dimensional Nilmanifolds 22
4.4. E-manifolds 23
4.5. Kähler Lie Algebroids 24
References 25
1. Introduction
1.1. Background. Hodge theory for Symplectic Manifolds was introduced by Ehresmann and Liberman and
rediscovered by Brylinski [2]; Brylinksi conjectured that on a Symplectic Manifold (M, ω) of dimension 2m
every deRham cohomology class admits a symplectic harmonic representative. Mathieu [14] and Yan [25]
independently showed that the Brylinksi conjecture holds true if and only if (M 2m , ω) satisfies the Strong
Lefschetz Property, that is for all 0 ≤ k ≤ m, the map
[L]k : H m−k (M ) → H m+k (M )
[α] 7→ [ω k ∧ α]
is surjective. This result was then improved by Merkulov in [18], and Guillemin in [7] by showing that these
two conditions are equivalent to the symplectic dδ-lemma in the case that M was a compact manifold using
1
Poincaré duality. In a key departure from traditional Riemannian Hodge theory the Symplectic Laplacian
is no longer an elliptic operator as it vanishes identically, and as such the results that heavily makeup the
theory no longer work. Instead we primarily work with the observation that in the presence of a symplectic
form, the space of differential forms on a smooth manifold admits the structure of an sl2 (R)-module. This
“Symplectic Hodge Theory” has been extended to new settings in work by Lin [11] and others, which require
reestablishing many conditions in specific geometric settings. In this paper we work towards streamlining
this by proving analogous statements about modules in a relevant module category, and applying them
to Symplectic Lie Algebroids, a minimal smooth setting for the notions involved to make sense. We also
generalize the operators ∂+ , ∂− of Tseng and Yau [24] to this strictly algebraic setting (which we rename d
and d respectively), though we do not explore the related Laplacians, or Appelli and Bott-Chern cohomologies
corresponding to certain combinations of these operators. We then relate the original trio of equivalences
back to ∂+ ∂− -lemma of [8] by reestablishing the equivalence of this to the original triple in our algebraic
framework rather than the original topological framework considered. Through this generalization we are
also able to lift the compactness condition, though it was free to assume in the classical case of the tangent
bundle by Poincaré Duality and standard Algebraic Topological facts relating to the relevant conditions.
1.2. Results. We define a category of modules Cn over a suitable Lie Superalgebra with properties that
axiomatize the algebra of differential forms on a Symplectic Manifold much in the spirit of [15].
Definition 1. Let g be the 5 dimensional Lie Superalgebra with basis e, f, h, d, δ with e, f, h degrees 2, −2, 0
respectively and d, δ in degrees 1, −1 respectively all subject to the following relations
[e, f ] = h [e, h] = −2e [f, h] = 2f
[e, d] = 0 [f, d] = δ [h, d] = d
[e, δ] = d [f, δ] = 0 [h, δ] = −δ
[d, d] = 0 [d, δ] = 0 [δ, δ] = 0
where all commutators are graded commutators. As a Lie Superalgebra this is isomorphic to the 5-dimensional
Orthosymplectic Lie Superalgebra
It is well known that the space of forms of any symplectic manifold admits the structure of a g-module
where the even part of the algebra corresponds to the standard sl2 structure, d corresponds to the deRham
differential, and δ is defined in [2]. Since the space of forms admits finitely many degree, 0 through 2n, we
want to restrict our attention to modules for which h acts with finitely many eigenvalues. Moreover, since
h acts on the form integrally and via a diagonalizable action, we want to restrict to modules for which this
occurs as well.
Definition 2. For n ≥ 0, let Cn denote the full subcategory of g-mod such that h acts integrally via
diagonalizable action with finitely many eigenvalues between −n and n
Objects of this category are also referred to as “Modules of Finite h-type” or “Modules of Finite h-spectrum”
in the literature. In the case that the underlying geometric structure is Kähler, or more generally has the
Hard Lefschetz Property, the module enjoys similar properties and is called a “Lefschetz Module”.
Theorem. Let V ∈ Ob(Cn ). Then the following are equivalent
(1) The map [ei ] : H −i (V, d) → H i (V, d) is an isomorphism for all i
(2) The inclusion (ker(δ), d) ֒→ (V, d) is a quasi-isomorphism
(3) The dδ-lemma holds for V
(4) The d d-lemma holds for V
After introducing the notion of the dδ-lemma holding weakly in section 2.4, we also weaken this to the
following theorem
Theorem. Let V ∈ Ob(Cn ). Then the following are equivalent
(1) The dδ-lemma holds for V up to degree s
2
(2) The map [ei ] : H −i (V, d) → H i (V, d) is an isomorphism for all |i| ≥ s
(3) The inclusion (ker(δ), d) ֒→ (V, d) is a quasi-isomorphism in degree |i| ≥ s
Applying the first of these two theorems to a Symplectic Lie Algebroid with representation we have the
following theorem:
Theorem. Let (A → M, ρ, ω) be a Symplectic Lie Algebroid of fiber dimension 2m, with representation
(E, ∇). Then the following are equivalent:
m−k m+k
(1) The map [L]k : HA (M, E) → HA (M, E) given by [ω]k ∧ − is an isomorphism for all k
(2) The inclusion of complexes (ker( ∇ ), ∇) ֒→ (Ω•A (M, E), A ∇) is a quasi-isomorphism
A ∗ A
2. Lefschetz Modules
2.1. The Lie Superalgebra.
Proposition 2.1. If V ∈ Ob(Cn ), then V admits a decomposition
Mn
V = Vi
i=−n
where h acts on Vi as scaling by i, i.e. i · idVi
Lemma 2.1. Every cyclic V ∈ Ob(Cn ) is finite dimensional
P
Proof. Let v generate V as a U (sl2 (R))-module, and decompose v into it’s components v = vi for vi ∈ Vi .
It suffices to show that the submodule each vi generate is finite dimensional. By the Poincaré-Birkhoff-Witt
theorem, this module is spanned by the vectors
hvi i = spanm,r,k {er f m hk vi }
Since h acts by scaling on vi , it suffices to check then on
hvi i = spanm,r {er f m vi }
Now, since there are only finitely many eigenvalues, we have that r, m ≤ n. From this we can note that there
are finitely many combinations of m and r such that
−n ≤ −2m + 2r + i ≤ n
Thus there are finitely many generators for hvi i
Now, we’d like to be able to talk about primitive elements as normal in sl2 representation theory.
L
Definition 3. Suppose V ∈ Ob(Cn ) with decomposition V = i Vi . We call a vector vi ∈ Vi primitive if
f vi = 0 or equivalently ei+1 vi = 0
Using this definition, we reprove a common structure theorem for sl2 -modules
3
Theorem 2.2. Let V ∈ Ob(Cn ). Then every v ∈ V can be written as a finite sum
X
v= ei vi
where the vi are primitive
Proof. Start by replacing V with hvi. This is again an object of Cn , and as it’s cyclic must be finite
dimensional. We can then decompose it into a finite sum of irreducibles, and write v as it’s components
in these, so it suffices to show elements in these irreducibles is of the form desired. Suppose v ∈ V−n
and u ∈ V−n+1 , both of which are primitive by degree considerations. Now, since f ek v = λ1 ek−1 v and
f ek u = λ2 ek−1 u for nonzero constants λi , we know that they span a nonzero submodule together. Since
we’re in the irreducible case, this must be the entire module. Moreover, they all correspond to distinct
eigenvalues of h, so they must be linearly independent, i.e. they form a basis of V .
Another useful fact is the following classical result
L
Proposition 2.2. Suppose V ∈ Ob(Cn ) with decomposition V = i Vi . Then the map ei : V−i → Vi is a
bijection
Proof. This follows from classical sl2 (R) representation theory
As well as the following bracket relations
Lemma 2.3. [ei , δ] = iei−1 d
Proof. We proceed by induction, the base case is given as definition. Suppose this holds for the i − 1 case,
then we have that
ei δ − δei = e(δei−1 + (i − 1)ei−2 d) − ei δ
= eδei−1 + (i − 1)ei−1 d − δei
= (δe + d)ei−1 + (i − 1)ei−1 d − δei
= iei−1 d
Lemma 2.4. [h, ei ] = 2iei
Proof. This holds from definition for i = 1, so suppose it holds for i − 1. Then we have that
hei − ei h = (ei−1 h + 2(i − 1)ei−1 )e − ei h
= ei−1 he + 2(i − 1)ei − ei h
= ei−1 (eh + 2e) + 2(i − 1)ei − ei h
= 2iei
We also want to bring the notion of the ⋆ involution present in classical Symplectic Hodge Theory:
Definition 4. On any V ∈ Ob(Cn ), define the “star operator” ⋆ : Vi → V−i to be the involution such that
δ|Vi = (−1)i+1 ⋆ d⋆
Now, in the spirit of classical deRham cohomology we arrive at the following definitions:
Definition 5. Let V ∈ Cn , and v ∈ V . Then we define the following
(1) v is “closed” if v ∈ ker(d)
(2) v is “exact” if v ∈ im(d)
(3) v is “coclosed” if v ∈ ker(δ)
(4) v is “coexact” if v ∈ im(δ)
4
More generally we can adopt the language of harmonicity as follows
Definition 6. An element v ∈ V , where V ∈ Cn , we say that v is “harmonic” if v ∈ ker(d) ∩ ker(δ). The
submodule of harmonic elements of V will be denoted Vb
Proposition 2.3. For V ∈ Ob(Cn ), Vb is a g-submodule of V .
Proof. This is clearly a vector subspace and closed under the action of d and δ. If v ∈ Vb , then dev = edv = 0
and δev = eδv − dv = 0. Likewise df v = f dv + δv = 0 and δf v = f δv = 0. The closedness under h follows
from e and f
One can regard V as a bi-differential Z-graded complex after writing V out as
d d d d
... Vi−1 Vi Vi+1 ...
δ δ δ δ
L
Definition 7. Let V ∈ Ob(Cn ) with decomposition V = i Vi . The cohomology of V regarded as the above
complex in d at the i-th position is denoted H i (V, d)
L
Definition 8. Let V ∈ Ob(Cn ) with decomposition V = i Vi . The homology of V regarded as the above
complex in in δ at the i-th position is denoted Hi (V, δ)
Lemma 2.5. Let V ∈ Ob(Cn ), and v ∈ V . If v is closed, then ⋆v is coclosed. Similarly, if v is coclosed,
then ⋆v is is closed.
Proof. Suppose that dv = 0. Then we have that
δ(⋆v) = ± ⋆ d ⋆ ⋆v = ± ⋆ dv = 0
Now, suppose δv = 0. Then we have that
δv = 0
±⋆d⋆v =0
±d ⋆ v = ⋆0 = 0
=⇒ d(⋆v) = 0
Lemma 2.6. Let V ∈ Ob(Cn ), and v ∈ V . If v is exact, then ⋆v is coexact. If v is coexact, then ⋆v is exact
Proof. Suppose that v is exact, that is v = dw. Then we have that
⋆v = ⋆dw = ⋆d ⋆ ⋆w = ±δ(⋆w)
Similarly, suppose v = δw. Then we have
⋆v = ⋆δw = ± ⋆ ⋆d ⋆ w = d(± ⋆ w)
Theorem 2.7. For V ∈ Ob(Cn ), there is an isomorphism
H i (ker(δ), d) ∼
= H −i (ker(d), δ)
Given by the ⋆ map
Proof. Define a map
ϕ : Vbi /d(Vi−1 ∩ ker(δ)) → Vb−i /δ(Vi+1 ∩ ker(d))
Given by sending [v] 7→ [⋆v]. This is well-defined by the preceding lemmas, and is surjective as for any class
[v], it has preimage [⋆v]. To see injectivity, suppose that ϕ[v] vanishes, then by definition we have
⋆v = δw = ± ⋆ d ⋆ w
5
For some w ∈ ker(d) ∩ Vi+1 , and the sign depending on the grading. Applying ⋆ to the above relation we
have
v = ±d ⋆ w = d(± ⋆ w)
Which means that [v] was the zero class. The only possible point of confusion is why ± ⋆ w is in the kernel
of δ, but this again follows from the preceding lemmas
2.2. Equivalence of Surjectivity. The following theorem was established independently by Yan and Math-
ieu.
Theorem 2.8 (Yan [25], Mathieu [14]). Let (M 2n , ω) be a symplectic manifold. Then the following are
equivalent:
n−k n+k
(1) The strong Lefschetz map [ω k ] : HdR (M ) → HdR (M ) is surjective
(2) Every deRham Cohomology class has has a (symplectic) harmonic representative
We can rephrase this purely as an algebraic equivalence: Let V ∈ Ob(Cn ), then the following are equivalent:
(1) For all 0 ≤ i ≤ n, the map [ei ] : H −i (V, d) → H i (V, d) is surjection
(2) The inclusion (ker(δ), d) ֒→ (V, d) induces a surjection in cohomology H i (ker(δ), d) → H i (V, d)
Lemma 2.9. [ei ] carries harmonic harmonic representatives to harmonic representatives
Proof. Suppose [v−i ] ∈ H −i (V, d) has a harmonic representative [v−i0
]. We claim that [ei v−i
0
] is a harmonic
i 0
representative of [e v−i ]. First, note that they’re cohomologous as if v−i = v−i + dτ−i−1 , then we have that
ei v−i = ei v−i
0
+ ei dτ−i−1 = ei v−i
0
+ dei τ−i−1
Moreover, we have that ei v−i
0
is harmonic since
δei v−i = ei δv−i − iei−1 dv−i = 0
The following is essentially a rephrasing of Yan’s original proof [25]:
Theorem 2.10 (D. Yan). Suppose V ∈ Ob(Cn ), then the following statements are equivalent
(1) The inclusion (ker(δ), d) ֒→ (V, d) induces a surjection in cohomology H i (ker(δ), d) → H i (V, d)
(2) For all 0 ≤ i ≤ n, the map [ei ] : H −i (V, d) → H i (V, d) is a surjection
Proof. Suppose condition 1 holds, that is the maps H i (ker(δ), d) → H i (V, d) are surjections, or equivalently
we can say we have surjections Vbi → H i (V, d). We also know by classical sl2 -representation theory that
ei : V−i → Vi is a bijection, and this restricts to a bijection on Vb−i → Vbi since the harmonic elements make
up a g-submodule of V . Together these give us the diagram
ei
Vb−i Vbi
H −i (V, d) H i (V, d)
[ei ]
Where the top and vertical arrows are surjections, implying the bottom map is surjective. This is the exact
map in condition 2. Now, suppose that [ei ] is surjective for all i. We’d like to show that the inclusion
map (ker(δ), d) ֒→ (V, d) is a surjection in cohomology, i.e. given any cohomology class [v] ∈ H • (V, d), we
can replace it by a class [w] where w ∈ ker(δ). First, we claim that it suffices to check this on cohomolo-
gies in nonpositive degree. To see this, fix [vi ] ∈ H i (V, d) for i > 0. Then as [ei ] as surjective, we have
that [vi ] = [ei ][w−i ] for some [w−i ] ∈ H −i (V, d), and we know by Propostion 2.9 that [ei ] will carry a har-
monic representative of [w−i ] to a harmonic representative of [vi ]. Now, we claim that for i > 0, H −i (V, d)
decomposes
H −i (V, d) = im([e]) + ker([ei+1 ])
6
Fix a class [v−i ] ∈ H −i (V, d) and consider [ei+1 v−i ] ∈ H i+2 (V, d). Since the [ei ] are surjective, we have that
there is class [β−i−2 ] ∈ H −i−2 (V, d) such that [ei+2 β−i−2 ] = [ei+1 v−i ]. Now, we can (trivially) write
[v−i ] = ([v−i − eβ−i−2 ]) + [e]([β−i−2 ])
It’s clear that the second summand is in the image of [e], we need only to check that the first is in the kernel
of [ei+1 ]. This follows from the definition as
[ei+1 ]([v−i − eβ−i−2 ]) = [ei+1 v−i − ei+2 β−i−2 ] = [ei+1 v−i − ei+1 v−i ] = [0]
We now proceed to prove the claim by induction. First, note that every class in H −n (V, d) and H −n+1 (V, d)
is harmonic. If [v−n ] ∈ H −n (V, d), then δv−n ≡ 0 trivially, and so v−n is harmonic. If [v−n+1 ] ∈ H −n+1 (V, d),
then we have that
δv−n+1 = [d, f ]v−n+1 = df v−n+1 − f dv−n+1 = 0
Since f v−n+1 ∈ V−n−1 ≡ 0, and v−n+1 is closed. So we only need to check on −n + 2 ≤ i ≤ 0, and the −n + 1
degree serves as the base case. Now, suppose the statement holds for H −i−1 (V, d) and we want to show that
it holds on H −i (V, d). Fix [w−i ] ∈ H −i (V, d). By the above decomposition we have that
[β−i ] = [α−i ] + [e]([β−i−2 ])
For [β−i−2 ] ∈ H −i−2 (V, d) and [α−i ] ∈ H −i (V, d) ∩ ker([ei+1 ]). By inductive hypothesis [β−i−2 ] possesses a
harmonic representative, and [e] carries harmonic representatives into harmonic representatives, so we can
freely assume [e]([β−i−2 ] is harmonic. We only need to show that [α−i ] admits a harmonic representative. By
definition of [α−i ] we have that dα−i = 0 and ei+1 α−i = dβi+1 for some βi+1 ∈ Vi+1 . Now, by the bijectivity
of ei+1 : V−i−1 → Vi+1 , there exists γ−i−1 ∈ V−i−1 such that βi+1 = ei+1 γ−i−1 . We claim that α−i − dγ−i−1
is a harmonic representative of [α−i ]. It’s clear that [α−i ] = [α−i − dγ−i−1 ], and that α−i − dγ−i−1 ∈ ker(d),
we only need to show that it’s in ker(δ). Since δ = [d, f ], this reduces to showing that α−i − dγ−i−1 ∈ ker(f ),
i.e. is primitive. Primitivity here is equivalent to showing that α−i − dγ−i−1 ∈ ker(ei+1 ). Computing we
have that
ei+1 (α−i − dγ−i−1 ) = ei+1 α−i − d(ei+1 γ−i−1 ) = ei+1 α−i − dβi+1 = ei+1 α−i − ei+1 α−i = 0
2.3. Equivalence of Triples. In the previous subsection we’ve shown that the following two statements are
equivalent for V ∈ Ob(Cn ):
(1) The map [ei ] : H −i (V, d) → H i (V, d) is a surjection
(2) The inclusion (ker(δ), d) ֒→ (V, d) induces a surjection in cohomology
Which in the original paper(s), Mathieu claims is equivalent to a third condition: The dδ-lemma.
Definition 9. The dδ-lemma holds for V ∈ Ob(Cn ) if the following equalities hold
im(d) ∩ ker(δ) = im(δ) ∩ ker(d) = im(dδ)
This is equivalent to the above two if we replace surjection with isomorphism in both claims; in the original
setting of the tangent bundle of a compact manifold these are equivalent by Poincare duality. The first of the
equivalences - that dδ-lemma is equivalent to the inclusion (ker(δ), d) ֒→ (V, d) follows from the following:
Theorem 2.11. Let V ∈ Ob(Cn ). Then the dδ-lemma is equivalent to the inclusion (ker(δ), d) → (V, d)
inducing an isomorphism in cohomology
Proof. See lemma 5.4.1 in [12]
Proposition 2.4. Let V ∈ Ob(Cn ), and suppose that the dδ-lemma is satisfied. Then [ek ] : H −k (V, d) →
H k (V, d) is an isomorphism for all k
7
Proof. Since dδ-lemma implies that there are harmonic representatives, we have that that [ek ] is surjective
by the prior work. This also allows us to check the validity of this isomorphism by restricting to [ek ] :
H −k (ker(δ), d) → H k (ker(δ), d). So, suppose that [α−k ] ∈ ker([ek ]). Then we have that ek α−k = dτk−1 for
some τk−1 ∈ Vk−1 ∩ ker(δ). Then, note that
δ(ek (α−k )) = δdτk−1 = −dδτk−1 = 0
So, the form ek α−k ∈ im(d) ∩ ker(δ), so by the dδ-lemma we have that there is a ρk such that
ek α−k = dδρk
Now, since ek is a bijection for all k, let φk denote it’s inverse. Then we have that
α−k = φk (dδρk ) = −φk δdρk
Now, since [ek , δ] = kek−1 d, we have that
φk δ = δφk − kφk ek−1 dφk
Meaning that
a−k = −(δφk − kφk ek−1 dφk )(dρk )
= −δφk dρk + kφk ek−1 φk d2 ρk
= −δφk dρk
= dδφk ρk
So α−k ∈ im(d) for δφk ρk ∈ ker(δ)
Proposition 2.5. Let V ∈ Ob(Cn ). If [ei ] : H −i (V, d) → H i (V, d) is an isomorphism for all i, then
ker(d) ∩ im(δ) = im(d) ∩ im(δ).
Proof. What we want to show is that for a general module element v ∈ V , we have that dδv = 0 implies
that δv = dφ for some φ. First, we claim that it suffices to check this on primitive elements. Towards this,
suppose that the statement P holds on primitive elements, and fix arbitrary v ∈ V . We can decompose v into
primitive elements, i.e. v = ei vi where the vi are primitive. Now, since dδv = 0, we have that
X X X
0 = dδv = dδei vi = d(ei δ − ie−i1 d)vi = ei dδvi
Since this sum is direct, this forces ei dδvi = 0 for all i. Since ei is acting on the negative weights in this sum,
it acts injectively, so dδvi = 0 for all i. By hypothesis, we have then that there exist φi such that δvi = dφi ,
and thus
X
δv = δei vi
X
= ei δvi − iei−1 dvi
X
= ei dφi − iei−1 dvi
X
=d ei φi − iei−1 vi
Now, to see that this holds on primitive form we proceed as follows. Suppose v−i ∈ V−i is primitive and
dδv−i = 0. Then we have that
ei+1 δv−i = δei+1 v−i − (i + 1)ei dv−i
However, ei+1 v−i = 0 since v−i is primitive, and we’re left with
ei+1 δv−i = d(−(i + 1)ei v−i )
Rephrased this says that [δv−i ] ∈ ker[ei+1 ], which by the Strong Lefschetz Property is an isomorphism,
meaning that [δv−i ] is the zero class in cohomology, so δv−i = dφ for some φ ∈ V−i−2 . The only degree
where this argument does not hold is V−n , but for any v ∈ V−n , δv ≡ 0, so it’s trivially in the image of d, i.e.
δv = 0 = d(0).
8
Proposition 2.6. Let V ∈ Ob(Cn ), and suppose that [ei ] : H −i (V, d) → H i (V, d) is an isomorphism for all
i. Then ker(d) ∩ im(δ) = ker(δ) ∩ im(d)
Proof. From the preceding proposition, we have that ker(d) ∩ im(δ) = im(d) ∩ im(δ) ⊂ ker(δ) ∩ im(d). So we
need to show that im(d) ∩ ker(δ) ⊂ im(δ) ∩ ker(d). So suppose α ∈ im(d) ∩ ker(δ), that is α = dβ and δα = 0.
Then we have
0 = δα = ± ⋆ d ⋆ α
Where sign depends on degree, but is not of importance for this proof. Since ⋆ is an isomorphism, we have
that d ⋆ α = 0,i.e. ⋆α ∈ ker(d). Now using exactness we have
⋆α = ⋆dβ = ± ⋆ d ⋆ ⋆β = ±δ ⋆ β
Thus ⋆α ∈ ker(d) ∩ im(δ). By the preceding proposition yet again we have then that ⋆α ∈ im(d) ∩ im(δ), and
so ⋆α = dρ. Using this we then have that
α = ⋆ ⋆ α = ⋆dρ = ⋆d ⋆ ⋆ρ = δ(± ⋆ ρ)
Giving us α ∈ im(δ)
Proposition 2.7. Let V ∈ Ob(Cn ), and suppose that [e]i : H −i (V, d) → H i (V, d) is surjective for all i. Then
im(d) ∩ im(δ) = im(dδ)
Proof. Since the two operators anticommute, the ⊃ direction is clear, and we need only to prove the forward
inclusion which we’ll prove by induction. First note that the statement trivially holds elements in Vn or
V−n since if v ∈ im(δ) ∩ Vn , then v = δ0 = 0, and likewise for V−n . Our base case is on Vn−1 , so choose
vn−1 ∈ Vn−1 , and let
vn−1 = dγn−2 = δwn
Now, since dwn is trivially zero it represents a cohomology class, and so we can choose a harmonic represen-
tative of said class giving us wn = wn0 − dτn−1 where wn0 is harmonic. We then have that
vn−1 = δwn
= δ(wn0 − dτn−1 )
= −δdτn−1
= dδτn−1
Now, suppose this holds for Vk for some −n < k < n, and fix vk−1 ∈ Vk−1 ∩ im(d) ∩ im(δ). By definition we
have that
vk−1 = dγk−2 = δβk
From this, define ξk+1 = dβk , and observe that
δξk+1 = δdβk = −d(δβk ) = −d2 γk−2 = 0
So ξk+1 ∈ im(d)∩ker(δ). By the remark preceding this proposition we have that im(d)∩ker(δ) = im(d)∩im(δ),
which by inductive hypothesis extends to im(d) ∩ ker(δ) = im(d) ∩ im(δ) = im(dδ), so ξk+1 = dδφk+1 . Then
note that d(βk − δφk+1 ) = 0, so it admits a harmonic representative, i.e. βk − δφk+1 = βk0 − dτk for some
τk ∈ Vk . Then applying the definition of vk−1 we have
vk−1 = δβk
= δ(βk0 − dτk + δφk+1 )
= −δdτk
= dδτk
At this point, we have enough to establish the equivalence of the triple: dδ-lemma, Strong Lefschetz
property, and Brylinksi property. We’ve already proven all the of the relevant pieces, but for clarity state
them here in one place
9
Theorem 2.12. Let V ∈ Ob(Cn ). The the following conditions are equivalent:
(1) The dδ-lemma holds for V
(2) The map [ei ] : H −i (V, d) → H i (V, d) is an isomorphism for all i
(3) The inclusion (ker(δ), d) ֒→ (V, d) is a quasi-isomorphism
Proof. We have that the first and third conditions are equivalent by Theorem 2.11. Then we have that the
first implies the second by Proposition 2.4, and the reverse direction follows from the Proposition 2.5, 2.6,
and 2.7.
In light of this theorem, we make the following definition
Definition 10. If V ∈ Ob(Cn ) satisfies any of the above equivalent conditions in Theorem 2.12 we call V a
“Lefschetz Module”
Corollary 2.12.1. Let V ∈ Ob(Cn ) be a Lefschetz module. Then H ∗ (V, d) is an sl2 (R)-module
Proof. We can see from the g relations that
[e, d] = 0 [f, d] = δ [h, d] = d
∗
So on H (V, d), we see that e and h pass to cohomology, i.e. it’s always a B-module. The only possible
problem lies in well-definedness of f . Suppose [v] = [w] ∈ H ∗ (V, d), that is v = w + dφ for some φ ∈ V . Then
we have that
f [v] = [f v] = [f (w + dφ)] = f [w] + [f dφ] = f [w] + [(df + δ)φ] = f [w] + [δφ]
Now δφ ∈ im(δ) ∩ ker(d), so since V is Lefschetz, there exists some ϕ such that δφ = dδϕ so
f [v] = f [w] + [dδϕ] = f [w]
2.4. Weakly Lefschetz Modules. There are examples on manifolds where the Lefschetz condition is only
partially met, i.e. the maps [L]k are isomorphisms for some k, but not all 0 ≤ k ≤ n
Definition 11. Given V ∈ Ob(Cn ), we say that V is “s-Lefschetz” if ek is a isomorphism for 0 ≤ s ≤ k ≤ n
The same theorems from before hold in this setting as well, but only on the portions on the modules for
which the maps ek are isomorphisms.
Theorem 2.13. Suppose V ∈ Ob(Cn ) , then the following statements are equivalent
(1) The inclusion (ker(δ), d) ֒→ (V, d) induces a surjection in cohomology H i (ker(δ), d) → H i (V, d) in
degrees |i| ≥ s
(2) For all 0 ≤ s ≤ i ≤ n, the map [ei ] : H −i (V, d) → H i (V, d) is a surjection
Proof. The forward direction follows in the same exact way it does in the proof of Theorem 2.10. The reverse
direction followed by induction that crucially used the fact that we could decompose H −i (V, d) for i > 0.
This decomposition was the only part of the proof that relied on the reverse direction, and the rest of the
proof holds if we only assume that the [ei ] are surjective from V−n up to V−s .
The upgraded equivalence also holds under the added assumption that ei is in fact an isomorphism
Definition 12. Given a module V ∈ Ob(Cn ), we say that V has the “dδ-lemma up to degree s” if
im(d) ∩ ker(δ) = im(dδ) = im(δ) ∩ ker(δ) on Vi for − n ≤ i ≤ s
im(d) ∩ ker(δ) = im(dδ) on Vs+1
for −n ≤ s ≤ −1
Note that by ⋆ duality if this equality holds on Vi for −n ≤ i ≤ −s it then also holds in Vi for s ≤ i ≤ n.
From this observation we’ll only prove statements about it in non-positive degree, as the duality will extend
the results to positive degrees.
10
Theorem 2.14. Let V ∈ Ob(Cn ). Then the following are equivalent
(1) The dδ-lemma holds for V up to degree s
(2) The map [ei ] : H −i (V, d) → H i (V, d) is an isomorphism for all |i| ≥ s
(3) The inclusion (ker(δ), d) ֒→ (V, d) is a quasi-isomorphism in degree |i| ≥ s
Proof. The proof will follow similarly to the original. The first and third condition are equivalent by [12],
whose argument works in each degree. The first implies the second using Proposition 2.4 and acknowledging
the proof is symmetric in it’s degree arguments about the zero weight. Finally, to see that the second
condition implies the first, note that proposition 2.5 and proposition 2.6 are degree agnostic, and proposition
2.7 is done by induction, which we can once again stop early when the hypothesis fail in degree −s and
achieve the result
2.5. Primitive Equivalence. Much of this section is rephrasing of work in Chung-I Ho’s thesis [8] and
Tseng’s paper [24]
L
Definition 13. Suppose V ∈ Ob(Cn ) with decomposition k Vk . Then the space of primitive elements of
degree k will be denoted Pk = ker(f ) ∩ Vk
Throughout the rest of this section, fix V ∈ Ob(Cn ), so that we may speak of elements vi ∈ P i freely.
Lemma 2.15. Suppose that wi ∈ Pi . Then the primitive decomposition of dwi only has two terms
Proof. First, note that
f 2 dwi = f (df + δ)wi = f δwi = δf wi = 0
As wi is primitive. Taking a primitive decomposition of dwi we have
X
dwi = ek vi+1−2k
k≥0
2
So applying f we have X
0= f 2 ek vi+1−2k
k≥0
Where vi+1−2k ∈ Pi+1−2k Which means for all k ≥ 2, we had to have had that ek vi+1−2k = 0, but as ei acts
injectively we have that vi+1−2k = 0 for these k, leaving us with
dwi = vi+1 + evi−1
Definition 14. Given vi ∈ Pi with dvi = vi+1 + evi−1 , define
d vi = vi+1
d vi = vi−1
so on P i we have dvi = vi+1 + evi−1 = d vi + e d vi
These are the same operators as ∂+ and ∂− as introduced in [24].
P These extend to the entire
P module by
taking primitive decomposition of any element v ∈ V , i.e. if v = vi for vi ∈ Vi , and vi = k≥0 ek vi−2k
then we have that
XX
dv = ek d vi−2k
i k≥0
XX
dv = ek d vi−2k
i k≥0
Lemma 2.16. We have that [d, e] = [d, e] = [d, d] = [d, d] = 0 where all are graded commutators, as well as
e(d d) = −e(d d)
On Vn−1 , Vn and [d, d] = 0 on Vk for k < n − 1
11
Proof. This follows the same as it did in [24], Lemma 2.5
Proposition 2.8. For vi ∈ P i , we have that f d vi = 0
Proof. Since vi is primitive, we have that [f, d]vi = f d vi . By the Jacobi identity we have that
1
[f, d] =
[h, [d, f ]] + [f, d]
2
Thus we have to have that h and [d, f ] commute. Checking this on vi gives us that
(i − 1)[d, f ]vi = i[d, f ]vi
Forcing [f, d]vi = 0
Proposition 2.9. For vi ∈ P i δvi = (1 − i) d vi
Proof. Using the definition of δ we have
δvi = f dvi − df vi = f dvi = f d vi + f e d vi
We know the first summand vanishes by the preceding lemma, and on the second we have that
f e d vi = (ef − h) d vi = −h d vi = −(i − 1) d vi
Lemma 2.17. On P i , we have that dδ = (1 − i) d d
Proof. Using the preceding lemmas we have
dδvi = (i − 1)d d vi = (i − 1)(d +e d) d vi = (1 − i) d d vi
Definition 15. We say that v ∈ V is “symphonic” if d v = d v = 0
This was previously called “primitive harmonic” in the literature, see [8], however to avoid the phrase
“harmonic primitive harmonic” we introduce this terminology.
Lemma 2.18. v ∈ V is harmonic if and only if each term of it’s primitive decomposition is symphonic
P
Proof. Let v = vi with vi = k≥0 ek wi−2k for wi−2k ∈ P i−2k . Then if each wi−2k , since d = d +e d it’s clear
that v is closed. Using the commutation relation given in Lemma 2.3, we have
XX
δv = δek wi−2k
i k≥0
XX
= (ek δ − kek−1 d)wi−2k
i k≥0
XX
= ek (1 − i) d wi−2k
i k≥0
=0
P
Now, suppose that v = i vi for vi ∈ Vi is harmonic. this occurs if and only if each vi is harmonic, which
occurs if and only if the term in the primitive decomposition of each vi is harmonic, so it suffices to chow
ci . Then
this for a primitive harmonic element. Towards this, suppose v ∈ P
0 = δv = (i − 1) d v
So v ∈ ker(d). Moreover we have
0 = dv = d v + e d v = d v
Thus v ∈ ker(d)
12
Definition 16. Let V ∈ Ob(Cn ). We say that V satisfies the d d-lemma if the following equalities hold
im(d) ∩ ker(d) ∩ P k = im(d d) ∩ P k −n + 1 ≤ k ≤ −1
im(d) ∩ ker(d) ∩ P −n = im(d d) ∩ P −n
im(d) ∩ ker(d) ∩ P 0 = im(d d) ∩ P 0
Lemma 2.19. The dδ-lemma holds for V ∈ Ob(Cn ) if the d d-lemma does
Proof. It suffices to check this on each weight, so fix vi ∈ Vbi and first suppose that vi ∈ im d so that
vi = dwi−1 . Then taking primitive decompositions of both we have
X X
vi = ek vi−2k wi−1 = ek wi−1−2k
k≥0 k≥0
Forcing vi−2k = dwi−1−2k = d wi−1−2k + e d wi−1−2k for all k. Since vi is harmonic, each vi−2k is symphonic,
so we must have that d wi−1−2k ∈ im d ∩ ker d and d wi−1−2k ∈ im d ∩ ker d. By the d d-lemma then we must
have that there are γi−2k , ρi−2−2k ∈ im(d d) (possibly zero) such that
d wi−1−2k = d d γi−2k d wi−1−2k = d d ρi−2−2k
From here, we have that
X
vi = ek (d wi−1−2k + e d wi−1−2k )
k≥0
X
= ek (d d γi−2k + e d d ρi−2(1+k) )
k≥0
X
γi−2k ρi−2(1+k)
= ek dδ + edδ
1 − (i − 2k) 1 − i + 2(1 + k)
k≥0
X
γi−2k ρi−2(1+k)
= dδ ek +e
1 − (i − 2k) 1 − i + 2(1 + k)
k≥0
Where we repeatedly use the g relations, and make note that the constants that appear in the denominator
are non-vanishing. The only possible point of confusion is why the γi−2k and ρi−2(1+k) are primitive. If
they’re not, then taking a primitive decomposition of them, it’s easy to see that we can adjust by constants
to force this to be true. A similar argument holds for the case when v ∈ im(δ).
Lemma 2.20. The d d-lemma holds for V ∈ Ob(Cn ) if the dδ-lemma does
Proof. Suppose that vi ∈ ker(d) ∩ ker(d) ∩ P i and recall that this forces vi to be harmonic and each of it’s
primitive decomposition terms to be symphonic. First, suppose that vi ∈ im(d), i.e. vi = d βi+1 . Then taking
primitive decompositions we have
X X
vi = ek vi−2k βi+1 ek βi+1−2k
k≥0
Since vi is symphonic, we have that d d βi−1−2k = 0 for all k. We then have that
X
dδβi−1 = dδek βi−1−2k
X
= ek dδβi−1−2k
k≥0
X 1
= ek d d βi−1−2k
2(1 − k) − i
k≥0
X −1
= ek d d βi−1−2k
2(1 − k) − i
k≥0
=0
So δdβi−1 = 0, meaning that dβi−1 ∈ ker(δ) ∩ im(d). By the dδ-lemma then we have that there exist γi such
that dβi−1 = dδγi . Taking a primitive decomposition of γi we have that
X
dδγi = ek dδγi−2k
k≥0
X
= (1 − i + 2k)ek d d γi−2k
k≥0
X
= dd (i − 1 − 2k)ek γi−2k
k≥0
That is, vi ∈ im(d) from which the result follows from the previous case.
Theorem 2.21. Let V ∈ Ob(Cn ). Then the following are equivalent
(1) The dδ-lemma holds for V
(2) The map [ei ] : H −i (V, d) → H i (V, d) is an isomorphism for all i
(3) The inclusion (ker(δ), d) ֒→ (V, d) is a quasi-isomorphism
(4) The d d-lemma holds for V
Proof. The first three are equivalent by 2.12. The first and last are equivalent by Lemmas 2.20 and 2.19
3. Lefschetz Algebroids
3.1. Symplectic Lie Algebroids. Lie Algebroids were originally introduced by Pradines [23] as a general-
ization of the tangent bundle of a manifold to the Differentiable Groupoid setting.
Definition 17. A “Lie Algebroid” (A, ρ, M ) over a C ∞ -manifold M is a vector bundle A → M , equipped
with an “anchor” ρ : A → T M , and a bracket [·, ·]A on sections of A that is bilinear, alternating, and satisfies
the Jacobi identity, such that
14
(1) [X, f Y ]A = f [X, Y ]A + (ρ(X) · f )Y for all X, Y ∈ Γ(A) and f ∈ C ∞ (M )
(2) ρ([X, Y ]A ) = [ρ(X), ρ(Y )]T M for all X, Y ∈ Γ(A)
Definition 18. Given a Lie Algebroid (A, ρ, M ), we can define the “deRham-Chevalley-Eilenberg complex”
of A, where our cochain complex is made of
ΩkA (M ) = Γ(Λk A∗ )
And we define M
Ω∗A (M ) = ΩkA (M )
k
k+1
This comes with an A-differential, dA : ΩkA (M ) → ΩA (M ), defined as
k+1
X
dA η(X1 , ..., Xk+1 ) = ci , ..., Xk+1 )
(−1)i+1 ρ(Xi ) · η(X1 , ..., X
i=1
X
+ ci , ..., X
(−1)i+j η([Xi , Xj ], X1 , ..., X cj , ..., Xk+1 )
i<j
Proposition 3.2. Given Symplectic Lie Algebroids (A1 → M1 , ω1 ) and (A2 → M2 , ω2 ), the product (A1 ×
A2 → M1 × M2 , ω) where ω = π1∗ ω1 + π2∗ ω2 is a Symplectic Lie Algebroid
Proof. The product of two Lie Algebroids has a unique Lie Algebroid structure [17], so we need to check that
ω is a symplectic form. It’s clear that it’s nondegenerate as it’s made from the pullback of two nondegenerate
forms. To see that it’s closed note that πi are Lie Algebroid morphisms, and a map f : A → B is a Lie
Algebroid morphism if and only if the induced map f ∗ : Ω∗B → Ω∗A is a chain map. Using this we have that
πi∗ are chain maps and so
dA1 ×A2 ω = dA1 ×A2 π1∗ ω1 + dA1 ×A2 π2∗ ω2 = π1∗ (dA1 ω1 ) + π2∗ (dA2 ω2 ) = 0
Lemma 3.1. Let (A → M, ρ, ω) be a Symplectic Lie Algebroid of fiber dimension 2m. Then for all p ∈ M ,
there exists a smooth symplectic local frame of A, i.e. there exists a frame {ui , vi }m i=1 ⊂ Γ(A) such that
{ui (q), vi (q)}m
i=1 is a symplectic basis of Aq for all q ∈ U , an open neighborhood of p.
20
(3) The (A, E)-dδ-lemma holds for A
(4) The (A, E)-d d-lemma holds for A
and the first three statements hold equivalence weakly, i.e. in degree as in the statement of Theorem 2.14.
The above corollary is the statement of Theorems 2.14 and 2.21. Taking E to be the trivial line bundle
recovers Theorem 3.4, however it would be interesting to look at what this theorem means for the choice
top
^ top
^
E = QA = A∗ ⊗ T ∗M
As defined in [3] as this always admits the structure of a representation of A.
4. Examples
4.1. Kodaria-Thurston Manifold. One of the first examples of a symplectic manifold that does not admit
a Kähler structure is given by the “Kodaira-Thurston” manifold KT 4 which is constructed as follows. Take
R4 with coordinates (x1 , x2 , x3 , x4 ) and choose a, b, c, d ∈ Z. Then consider the quotient of R4 by identifying
(x1 , x2 , x3 , x4 ) ∼ (x1 + a, x2 + b, x3 + c, x4 + d − bx3 )
The resulting quotient is the connected smooth manifold KT 4. We can see that we have a basis of one forms
{ei }4i=1 given by
e1 = dx1 e2 = dx2 e3 = dx3 e4 = dx4 + x2 dx3
We have a symplectic structure on KT 4 given by ω = e1 ∧ e2 + e3 ∧ e4 , and the deRham cohomology of KT 4
has given by generators
0
HdR (KT 4 ) = h[1]i 1
HdR (KT 4 ) = h[e1 ], [e2 ], [e3 ]i
2
HdR (KT 4 ) = hω, [e12 − e34 ], [e13 ], [e24 ]i 3
HdR (KT 4 ) = h[ω ∧ e1 ], [ω ∧ e2 ], [ω ∧ e4 ]i
4 1
HdR (KT 4 ) = h[ ω 2 ]i
2
Now we can consider the Lefschetz maps
1
[L] : HdR (KT 4 ) → HdR
3
(KT 4 ) 0
[L]2 : HdR (KT 4 ) → HdR
4
(KT 4 )
The map [L]2 is clearly an isomorphism, but [L] is not as it has nontrivial kernel, and doesn’t surject either.
This satisfies the criteria of Theorem 3.4 of degree s = 1 where we take A = T M , the identity as the anchor,
and the differential as the classical deRham differential.
4.2. Symplectic Lie Algebras of Dimension Four. If we consider Symplectic Lie Algebras as Symplectic
Lie Algebroids over the one-point space, the Lie Algebroid Cohomology is exactly the Chevalley-Eilenberg
cohomology, and our theorem can be checked by examining the surjectivity of the maps [L]k : H m−k (g) →
H m+k (g). For clarification sake by a Symplectic Lie Algebra we mean a (real) Lie Algebra g equipped with
V2 ∗
a nondegenerate ω ∈ g that is closed under the Chevalley-Elienberg differential. The 2-dimensional case
has only two isomorphism classes; the abelian one which satisfies Theorem 3.5 and the other one which does
not. Things get more interesting in the 4-dimensional case, where the number of symplectomorphism classes
grows significantly. In these case we need to check the following two maps in cohomology in the diagram:
[L]2
[L]1
4.3. Six-Dimensional Nilmanifolds. Nilmanifolds are homogeneous spaces N/Γ where N is a simply
connected nilpotent real Lie Group, and Γ is a lattice in N of maximal rank. Such spaces have been classified
up to isomorphism in dimension 6, and there are 34 such isomorphism classes. It is known [5] that exactly 26
of the 34 isomorphism classes admit symplectic structures. The data of such a space is typically presented
by two pieces of information: Its class, given as a 6-tuple, and its symplectic form. For example, the tuple
(0, 0, 0, 0, 12, 14 + 25) with symplectic form 13 + 26 + 45 conveys the data of the Lie Algebra with dual
generators {ei }6i=1 such that
de1 = de2 = de3 = de4 = 0
de5 = e1 ∧ e2
de6 = e1 ∧ e4 + e2 ∧ e5
with symplectic form ω = e1 ∧ e3 + e2 ∧ e6 + e4 ∧ e5 . Only one of these classes admit a Kähler structure,
and the other classes have provided interesting (counter)-examples in Symplectic Geometry. The fact that
the Lie Algebra data is sufficient to classify these is evident in the following theorem:
•
Theorem (Nomizu
V• ∗ [21]). The deRham complex Ω (M•) of a nilmanifold M = N/Γ is quasi-isomorphic to
the complex n of left-invariant forms on N , thus HdR (M ) ∼ •
= HCE (n).
Examining the maps [L]k : H 3−k (n) → H 3+k (n) where [α] 7→ [ω k ∧ α] for k = 1, 2, 3 allows to check
when Theorem 3.5 holds. The first of these maps [L]3 is assured to be an isomorphism by compactness
of the nilmanifold, and the remaining two are explicitly calculable given the standard data of a symplectic
nilmanifold described above. By Poincareé Duality, the maps [L]k are between vector spaces of the same
finite dimension, and so being surjective is equivalent to being an isomorphism. Only one class will admit all
22
Signature Symplectic Form [L] [L]2 [L]3
(0,0,12,13,14,15) 16+34-25 No No Yes
(0,0,12,13,14,23+15) 16+24+34-26 No No Yes
(0,0,12,13,23,14) 15+24+34-26 Yes No Yes
(0,0,12,13,23,14-25) 15+24-35+16 Yes No Yes
(0,0,12,13,23,14+25) 15+24+35+16 Yes No Yes
(0,0,12,13,14+23,24+15) 16 + 2 × 34 − 25 No No Yes
(0,0,0,12,13,14+23) 16 − 2 × 34 − 25 No No Yes
(0,0,0,12,13,24) 26+14+35 No No Yes
(0,0,0,12,13,14) 16+24+35 No No Yes
(0,0,0,12,13,23) 15+24+36 No No Yes
(0,0,0,12,14,15+23) 13+26-45 No No Yes
(0,0,0,12,14+15+23+24) 13+26-45 No No Yes
(0,0,0,12,14,15+24) 13+26-45 No No Yes
(0,0,0,12,14,15) 13+26-45 No No Yes
(0,0,0,12,14,13+42) 15+26+34 No No Yes
(0,0,0,12,14,23+24) 16-34+25 No No Yes
(0,0,0,12,14,15+34) 16+35+24 No No Yes
(0,0,0,12,14+23,13+42) 15 + 2 × 26 + 34 No No Yes
(0,0,0,0,12,15) 16+25+34 No No Yes
(0,0,0,0,12,14+25) 13+26+45 No No Yes
(0,0,0,0,12,14+23) 3+26+45 No No Yes
(0,0,0,0,12,34) 15+36+24 No No Yes
(0,0,0,0,12,13) 16+25+34 No No Yes
(0,0,0,0,13+42, 14+23) 16+25+34 No No Yes
(0,0,0,0,0,12) 16+23+45 No No Yes
(0,0,0,0,0,0) 12+34+56 Yes Yes Yes
Table 2. Six-Dimensional Nilmanifold Lefschetz Maps
three isomorphisms, as in this setting all three maps being isomorphisms is equivalent to admitting Kähler
structure [1]. Table 2 lists whether or not a given nilmanifold class has maps [L]k that are isomorphisms; Yes
indicating it’s an isomorphism, and No indicating it’s not an isomorphism.
4.4. E-manifolds. E-manifolds are a generalization of a collection of different geometric settings: b-manifolds,
bm -manifolds, c-manifolds, and Regular Foliations. Miranda and Scott have thoroughly explored the coho-
mology and symplectic geometry of E-manifolds in many settings, e.g. [19]. The definition given in the above
paper of Miranda and Scott of an E-manifold is restated here for convenience.
Definition 31. Let E be a locally free submodule of the C ∞ module X(M ) of vector fields on M . By the
Serre-Swan theorem, there is an E-tangent ∗ bundle E T M whose sections (locally) are sections of E, and an
E-cotangent bundle T M := T M . We will call the global sections of Λp E T ∗ M E-forms of degree p,
E ∗ E
and denote the space of all such sections by E Ωp (M ). If E satisfies the involutivity condition [E, E] ⊆ E,
there is a differential d : E Ωp (M ) → E Ωp+1 (M ) given by
E
X
dη V0 , . . . , Vp = (−1)i Vi η V0 , . . . , V̂i , . . . , Vp
i
X
+ (−1)i+j η Vi , Vj , V0 , . . . , V̂i , . . . , V̂j , . . . , Vp .
i<j
23
The cohomology of this complex is known as E-cohomology and is the same as Lie Algebroid Cohomology
if we view this as a Lie Algebroid with anchor given by inclusion. If we have an E-closed nondegenerate
E-2-form ω, we call this an E-symplectic form and the triple (E, M, ω) an E-symplectic manifold. These
are Symplectic Lie Algebroids. The cohomology of such an object in general is rather difficult to compute,
however Miranda and Scott showed the following
Theorem 4.1 (Miranda and Scott [19]). Let M = R2 , and let E be the involutive subbundle of X(R2 )
generated by
v1 = x∂x + y∂y
v2 = −y∂x + x∂y
With dual bundle generated by
xdx + ydy
v1∗ =
x2 + y 2
−ydx + xdy
v2∗ =
x2 + y 2
Then the E-Cohomology or Lie Algebroid Cohomology is given by
i = 0, 2
R
E k 2
H (M ) = R i=1
0 i≥3
Theorem 4.2. The Symplectic Lie Algebroid defined by Miranda and Scott (E → R2 , ρ = id, ω = v1∗ ∧ v2∗ )
satisfies theorem 3.4
Proof. We only need to check that the map [L] : E H 0 (R2 ) → E H 2 (R2 ) is an isomorphism, which amounts
to showing that ω = v1∗ ∧ v2∗ is not an exact form. Suppose towards a contradiction that it is, i.e. there are
f, g ∈ C ∞ (R2 ) such that ω = E d(f v1∗ + gv2∗ ). Unfolding the definition this implies that
ω = (v1 (g) − v2 (f ))ω
2
Which means for all (x, y) ∈ R we have that
v1 (g) − v2 (f ) = 1
Or equivalently
∂g ∂f ∂g ∂f
− x+ + y=1
∂x ∂y ∂x ∂x
However at (0, 0) this fails, so ω is not exact
4.5. Kähler Lie Algebroids. There has been recent attention towards Kähler Lie Algebroids, see [9]. By
the Hard Lefschetz Theorem, Kälher manifolds are always Lefschetz manifolds. This relationship still holds
true for a suitable class of Kähler Lie Algebroids
Definition 32. A “Kähler Lie Algebroid” (A → M, ρ, g) is a Lie Algebroid with a Hermitian metric g whose
associated canonical (1, 1)-A-form ω ∈ Ω1,1
A (M ) is closed
References
[1] Chal Benson and Carolyn S. Gordon, Kähler structures on compact solvmanifolds, Proc. Amer. Math. Soc. 108 (1990),
no. 4, 971–980. MR993739
[2] Jean-Luc Brylinski, A Differential Complex for Poisson Manifolds, Journal of Differential Geometry 28 (1988), no. 1, 93
–114.
[3] Sam Evens, Jiang-Hua Lu, and Alan Weinstein, Transverse measures, the modular class and a cohomology pairing for Lie
algebroids, Quart. J. Math. Oxford Ser. (2) 50 (1999), no. 200, 417–436. MR1726784
[4] Marisa Fernández, Vicente Muñoz, and Luis Ugarte, The dδ–lemma for weakly lefschetz symplectic manifolds, 2005.
[5] Michel Goze and Yusupdjan Khakimdjanov, Nilpotent lie algebras, Springer, London, 1996.
[6] Victor Guillemin, Hodge theory, 1997. Available at https://math.mit.edu/~vwg/shlomo-notes.pdf.
[7] , Symplectic hodge theory and the dδ-lemma, 2001.
[8] Chung-I Ho, Topological Methods in Symplectic Geometry, ProQuest LLC, Ann Arbor, MI, 2011. Thesis (Ph.D.)–University
of Minnesota. MR2912247
[9] Tengzhou Hu, A Generalized Kodaira Vanishing Theorem for Lie Algebroids and a Riemann-Roch Theorem for Singular
Foliation, ProQuest LLC, Ann Arbor, MI, 2023. Thesis (Ph.D.)–Washington University in St. Louis. MR4722602
[10] Yi Lin, Yiannis Loizides, Reyer Sjamaar, and Yanli Song, Symplectic reduction and a Darboux-Moser-Weinstein theorem
for Lie algebroids, Pure Appl. Math. Q. 19 (2023), no. 4, 2067–2131. MR4671391
[11] Yi Lin and Reyer Sjamaar, Equivariant symplectic Hodge theory and the dG δ-lemma, J. Symplectic Geom. 2 (2004), no. 2,
267–278. MR2108377
[12] Yu.Ĩ. Manin, Three constructions of Frobenius manifolds: a comparative study, Asian J. Math. 3 (1999), no. 1, 179–220.
Sir Michael Atiyah: a great mathematician of the twentieth century. MR1701927
[13] Eduardo Martínez, Lagrangian mechanics on Lie algebroids, Acta Appl. Math. 67 (2001), no. 3, 295–320. MR1861135
[14] Olivier Mathieu, Harmonic cohomology classes of symplectic manifolds, Commentarii Mathematici Helvetici 70
(1995/12/01), no. 1, 1–9.
[15] , Homologies associated with Poisson structures, Deformation theory and symplectic geometry (Ascona, 1996), 1997,
pp. 177–199. MR1480723
[16] Anastasia Matveeva and Eva Miranda, Reduction theory for singular symplectic manifolds and singular forms on moduli
spaces, Adv. Math. 428 (2023), Paper No. 109161, 42. MR4603782
[17] Eckhard Meinrenkin, Lie groupoids and lie algebroids lecture notes, fall 2017, University of Toronto, 2017.
[18] S. A. Merkulov, Formality of canonical symplectic complexes and frobenius manifolds, Internat. Math. Res. Notices 14
(1998), 727–733. MR1637093
[19] Eva Miranda and Geoffrey Scott, The geometry of E-manifolds, Rev. Mat. Iberoam. 37 (2021), no. 3, 1207–1224.
MR4236806
[20] Ryszard Nest and Boris Tsygan, Deformations of symplectic lie algebroids, deformations of holomorphic symplectic struc-
tures, and index theorems, Asian J. Math. 5 (2001), no. 4, 599–635. MR1913813
[21] Katsumi Nomizu, On the cohomology of compact homogeneous spaces of nilpotent Lie groups, Ann. of Math. (2) 59 (1954),
531–538. MR64057
[22] Gabriela Ovando, Four dimensional symplectic Lie algebras, Beiträge Algebra Geom. 47 (2006), no. 2, 419–434. MR2307912
[23] Jean Pradines, Théorie de Lie pour les groupoïdes différentiables. Calcul différenetiel dans la catégorie des groupoïdes
infinitésimaux, C. R. Acad. Sci. Paris Sér. A-B 264 (1967), A245–A248. MR216409
[24] Li-Sheng Tseng and Shing-Tung Yau, Cohomology and Hodge theory on symplectic manifolds: II, J. Differential Geom. 91
(2012), no. 3, 417–443. MR2981844
[25] Dong Yan, Hodge structure on symplectic manifolds, Adv. Math. 120 (1996), no. 1, 143–154. MR1392276
25