You are on page 1of 23

An approach to Hamiltonian Floer theory for

maps from surfaces


arXiv:2404.12249v1 [math.SG] 18 Apr 2024

Ronen Brilleslijper and Oliver Fabert

April 19, 2024

Abstract

In n-dimensional classical field theory one studies maps from n-


dimensional manifolds in such a way that classical mechanics is recovered
for n = 1. In previous papers we have shown that the standard polysym-
plectic framework in which field theory is described, is not suitable for
variational techniques. In this paper, we introduce for n = 2 a Lagrange-
Hamilton formalism that allows us to define a generalization of Hamil-
tonian Floer theory. As an application, we prove a cuplength estimate
for our Hamiltonian equations that yields a lower bound on the number
of solutions to Laplace equations with nonlinearity. We also discuss the
relation with holomorphic Floer theory.

Contents

1 Introduction 2

1.1 Two-dimensional Bridges regularization . . . . . . . . . . . . . . 3

1.2 Outline of the article . . . . . . . . . . . . . . . . . . . . . . . . . 5

2 Complex structures and the Lagrangian formalism 5

2.1 Relation to holomorphic Hamiltonian systems . . . . . . . . . . . 6

2.2 Lagrangian approach . . . . . . . . . . . . . . . . . . . . . . . . . 7

3 Cuplength proof 9

3.1 C 0 -bounds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9

1
3.2 Floer curves . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10

3.3 Proof of theorem 3.1 . . . . . . . . . . . . . . . . . . . . . . . . . 12

4 Fredholm theory 15

4.1 Linearization of the Floer map . . . . . . . . . . . . . . . . . . . 15

4.2 Transversality . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17

5 Compactness 18

A Proof of lemma 5.3 20

1 Introduction

This articles continues in the lines of [BF23b, BF23a], developing a Hamiltonian


formalism for (respectively Minkowski and Euclidean) field theories that is fit
for variational calculus. In the current article, the focus lies on the Laplace
equation with non-linearity on the two-dimensional torus. The aim of this arti-
cle is twofold. First of all, we will provide a connection between the Hamilton
equations and their Lagrangian counterpart. In this process a nice connection
will emerge with complex structures and we will highlight a relation to holomor-
phic Hamiltonian systems. Secondly, we will prove a cuplength result similar to
[BF23a] in the two-dimensional case using Floer theory. The novel contributions
of this paper are the Fredholm theory and compactness results for the moduli
space of Floer curves (see sections 4 and 5). Working out these aspects puts us
a step closer to the development of a full Floer theory. Furthermore the analysis
should be useful also for defining a holomorphic Floer theory.

The Hamiltonian formalism used in this article arises from what we call Bridges
regularization. This formalism can be found in [Bri06] and provides a Hamilto-
nian framework for, amongst others, the Laplace equation that pertains its nice
analytical properties. To motivate its merits, let us start by looking at Newton’s
equation

d2
− q(t) = Vt′ (q(t)), (1)
dt2
for a particle q : R → Q subject to some potential Vt : Q → R on a manifold
Q. To formulate this as a first-order equation, we can introduce the momentum
d
p(t) = dt q(t) and the Hamiltonian Ht (u) = 12 |p|2 + Vt (q), for u = (q, p) and

2
write equation (1) as

d
J u(t) = ∇Ht (u(t)). (2)
dt
 
0 −Id
Here J = is the standard complex structure on the cotangent bundle
Id 0
T ∗ Q. Note that this so-called Hamiltonian equation is elliptic and that (J dt
d 2
) =
d2
− dt2 . In Floer theory, one uses this Hamiltonian formalism to rewrite solutions
to equation (1) as critical points of some action functional AH . One then studies
gradient-lines of AH to conclude existence results for solutions to Newton’s
equation. A well-known result in Floer theory is that the number of critical
points of AH is bounded below by the topology of Q. See [AH16] for more
explanation and [Cie94] for the specific case of cotangent bundles. Crucial in
the proofs of these results is the ellipticity of equation (2).

Note that equation (2) comes from a Lagrangian formalism. Periodic solutions
to equation (1) arise as critical points of the Lagrangian action functional
Z
S[q] = L(q(t), q ′ (t), t) dt,
S1

where L(q, q ′ , t) = 21 (q ′ )2 − Vt (q). When we define p = ∂q


∂L ′
′ and H = pq − L, the

Euler-Lagrange equations for the Lagrangian action coincide with equation (2).
Here, in order to view Ht as a function of q and p, we need the Legendre
condition
∂2L
det ′2 ̸= 0.
∂q
We will see a similar condition when connecting the Lagrangian and Hamiltonian
approaches for the Laplace equation (see section 2.2).

1.1 Two-dimensional Bridges regularization

In this article, the starting point is the Laplace equation on the 2-dimensional
torus. Let q : T2 → Td satisfy the Laplace equation with non-linearity
dV
−(∂12 + ∂22 )q(t) = (q(t)), (3)
dq

where t = (t1 , t2 ) ∈ T2 , ∂i := ∂ti and V : Td → R is a non-linearity. We will


prove a cuplength result for the number of solutions to equation (3) and actually
for more general equations. In comparison with Newton’s equation we will first
try to formulate 3 as a first-order equation. The standard way of doing this,
is by introducing ’momentum-like’ variables for the different derivatives. The

3
following set of equations follows this so-called de Donder-Weyl approach:i
∂1 q = p1
∂2 q = p2 (4)
dV
−∂1 p1 −∂2 p2 = .
dq
Even though equation (4) can be written using a Hamiltonian H as K∂ u(t) =
∇H(u(t)) with u = (q, p1 , p2 ) for a first-order differential operator K∂ , this
equation on its own has some major disadvantages. Most notably, the equation
is not elliptic anymore, whereas equation (3) is. For Floer theoretic purposes
this is highly undesirable. Moreover, K∂ has an infinite-dimensional kernel on
the space of maps on T2 and it does not square to the Laplacian in contrast to
d 2 d2
the mechanical case, where (J dt ) = − dt 2 . As proven in [BF23b] for the case

of the wave equation, the Floer curves corresponding to equation (4) do not
necessarily converge to solutions. Also see [BF23a] for a similar problem for the
Laplace equation on a 3-dimensional torus.

Fortunately, there is a different way of writing equation (3) as a first-order


system that solves these problems. First proposed in [BD99] and later worked
out in [Bri06], this new set of equations follows from adding some extra terms
to equation (4), depicted in blue.
∂1 q1 +∂2 q2 = p1
−∂1 q2 +∂2 q1 = p2
dV1 (5)
−∂1 p1 − ∂2 p2 =
dq1
dV2
∂1 p2 − ∂2 p1 =
dq2
Note that we changed the variable q from equation (4) into q1 here, to stress that
this first-order system of equations actually describes two Laplace equations
dV1
−(∂12 + ∂22 )q1 (t) = (q1 (t))
dq1
dV2
−(∂12 + ∂22 )q2 (t) = (q2 (t)).
dq2
The underlying geometric structure behind equation (5), which we call Bridges’
equations is explained in [Bri06]. A shorter version in the context of maps
between tori can also be found in [BF23b, BF23a], so we do not go into details on
this in the present article. However, we do want to stress again that equation (5)
is an elliptic first-order system. We can write it as
J∂ Z(t) = ∇H(Z(t)),
i Note that the de Donder-Weyl equations are typically used as a first-order approach to

the wave equation, so the signs are different. See a.o. [Kan93, Gün87].

4
where Z = (q1 , q2 , p1 , p2 ), H(Z) = 12 p21 + 12 p22 + V1 (q1 ) + V2 (q2 ) and J∂ is a first-
order operator satisfying J∂2 = −(∂12 + ∂22 ). Note that only constant functions
lie in the kernel of J∂ .
Remark 1.1. Note that there is a hyperkähler structure on the target space
hidden in these equations. When we write out J∂ = J1 ∂1 + J2 ∂2 ,then J1 and J2
are both complex structures and their product is too. In remark 3.5 we will note
that this yields a relation between our Floer curves and the Fueter equation. The
complex structure J1 pairs qi with pi whereas J2 pairs qi with pj for i ̸= j. Their
product J3 = J1 J2 therefore pairs q1 with q2 and p1 with p2 . In section 2 we will
see that indeed equation (5) can be written using some complex multiplication
on T2d .

Equation (5) can also be described in terms of the polysymplectic form


Ω = ω1 ⊗ ∂1 + ω2 ⊗ ∂2 , (6)
where ω1 = dp1 ∧ dq1 − dp2 ∧ dq2 and ω2 = dp2 ∧ dq1 + dp1 ∧ dq2 . This description
will be of use later, when we compare the equations to holomorphic Hamiltonian
systems. Note that both ω1 and ω2 are symplectic forms. Heuristically, ωi de-
termines the derivative in the direction of ti . In other words, Z = (q1 , q2 , p1 , p2 )
is a solution to equation (5) if and only if
ω1 (·, ∂1 Z) + ω2 (·, ∂2 Z) = dH.
See [Gün87] for an exposition of polysymplectic geometry.

1.2 Outline of the article

This paper is organized as follows. In section 2 we start by rewriting the Bridges


equations using language from complex analysis. After relating our equations
to holomorphic Hamiltonian systems, we discuss the underlying Lagrangian for-
malism. In section 3 we state the main theorem 3.1 together with its implication
for Laplace equatinos with nonlinearity. The C 0 -bounds are given for both the
orbits and the Floer curves. The section concludes with a proof of the theo-
rem, modulo the necessary Fredholm theory and compactness results. These
are treated in respectively sections 4 and 5. The proof of one technical lemma
needed for compactness is deferred to the appendix.

2 Complex structures and the Lagrangian for-


malism

Before motivating the Hamiltonian approach to Laplace’s equation mentioned


above from the corresponding Lagrangian system, we first want to note that we

5
can rewrite equation (5) in terms of complex structures. Note that for H ≡ 0
(i.e. a vanishing right-hand side in equation (5)), Bridges’ equations reduce
to a set of two Cauchy Riemann equations for the pairs (q1 , q2 ) and (p1 , p2 ).
This leads us to define t = t1 + it2 and correspondingly t̄ = t1 − it2 . Also let
q = q1 + iq2 , p = p1 + ip2 and correspondingly q̄ = q1 − iq2 , p̄ = p1 − ip2 . Then
equation (5) is equivalent to the set of equations
1 ∂H
∂t q = p̄ =
2 ∂p
(7)
∂V ∂H
−∂t p = = ,
∂q ∂q

where ∂t = 21 (∂1 − i∂2 ), ∂q = 21 ( ∂q∂ 1 − i ∂q∂ 2 ) and V (q, q̄) = V1 (q1 ) + V2 (q2 ) ∈ R.
Also H = H(q, q̄, p, p̄) = 12 pp̄ + V (q, q̄) is still the same real-valued function as
before, but now seen as a function of these new coordinates.

While the Hamiltonian formalism for the Laplace equation might come across as
purely notation at this point, we will show in the section 2.2 that these equations
actually relate to a Lagrangian version of a minimization problem by means of
a Legendre transform. This will also provide us with a much bigger class of
examples then just Laplace equations. Before that, the next section explains
the relation to holomorphic Hamiltonian systems.

2.1 Relation to holomorphic Hamiltonian systems

Equation (7) looks very similar to the holomorphic Hamiltonian systems (HHS)
studied in [Wag23]. The main difference is that their Hamiltonians are holo-
morphic functions with values in C, whereas our Hamiltonian is real-valued.
We crucially need the derivative ∂H
∂p to actually depend on p̄ in order to get the
Laplace equation
∂V dV1 dV2
−∆(q1 + iq2 ) = −4∂t̄ ∂t q = −2∂t̄ p̄ = 2 = +i . (8)
∂ q̄ dq1 dq2
Notice also that our Hamiltonians are not necessarily the real or imaginary parts
of a holomorphic Hamiltonian, since we do not require them to be harmonic.

Interestingly, both the HHS’s from Wagner and the Bridges equations used in
this article can be described as special cases of a broader framework. Recall the
polysymplectic form Ω from equation (6). In the complex notation it can be
written as Ω = ωC ⊗ ∂t + ω̄C ⊗ ∂t̄ , where ωC = dp ∧ dq and ω̄C = dp̄ ∧ dq̄. Just as
in [Wag23], the form ωC (respectively ω̄C ) induces an isomorphisms between the
holomorphic (respectively anti-holomorphic) tangent and cotangent bundles

ωC♭ : T (1,0) X → T ∗(1,0) X

ω̄C♭ : T (0,1) X → T ∗(0,1) X,

6
where X = T ∗ T2d . Thus, the polysymplectic form Ω induces an isomorphism

ωC♭ ⊕ ω̄C♭ : TC X → TC∗ X, where the complexified tangent and cotangent bun-
dles are viewed as the direct sums of their holomorphic and anti-holomorphic
subbundles. In principle this means that for any function H : X → C we get
a complex vector field by taking the inverse of dH under this map. In the
study of HHS’s H is taken to be a holomorphic function, which means that
dH ∈ T ∗(1,0) X. Thus the Hamiltonian equations only take into account the
holomorphic vector field (ωC♭ )−1 (dH) and the anti-holomorphic vector field van-
ishes. In our case the Hamiltonians are real-valued meaning that its holomor-
phic and anti-holomorphic derivatives are equal. Thus dH lies in the diagonal of
TC∗ X = T ∗(1,0) X ⊕T ∗(0,1) X and we actually get two sets of conjugate equations.
The analysis in this article is aimed at the case of real-valued Hamiltonians, but
we expect similar techniques to work in the holomorphic setting.
Remark 2.1. Note that equation (7) is written in terms of local coordinates
on the torus. However, just as for HHS’s, the equations can be formulated on
more general manifolds. For equation (7) to make sense the domain should be
a Riemann surface and the target a complex manifold. From a dynamical point
of view this could be of particular interest for string theory, where the Riemann
surface is viewed as the closed world-sheet of a 1-dimensional string and the
target manifold is Calabi-Yau. The flatness of the Calabi-Yau manifold will be
crucial for the compactness result to hold, just like in [HNS09].

2.2 Lagrangian approach

We observe that solutions to the Laplace equations arise as critical points of the
functional
 dt ∧ dt̄
Z
S[q] = − 2∂t q∂t̄ q̄ − V (q, q̄) ,
T2 2i
where we use the notation of equation (8) (for a proof see lemma 2.2 below). In
this section, we will develop a connection between the Lagrangian picture and
the Hamiltonian formalism sketched above.

Let L = L(q, q̄, ∂t q, ∂t̄ q̄, t, t̄) be a Lagrangian with values in R and
dt ∧ dt̄
Z
S[q] = − L(q, q̄, ∂t q, ∂t̄ q̄, t, t̄) (9)
T 2 2i
the corresponding action. Note that S is real-valued as well. Our main example
is the Lagrangian L = 2∂t q∂t̄ q̄ − V (q, q̄) with action given above.
Lemma 2.2. The Euler-Lagrange equations for the action (9) are given by
∂L ∂L ∂L ∂L
= ∂t = ∂t̄ . (10)
∂q ∂(∂t q) ∂ q̄ ∂(∂t̄ q̄)
Note that these two equations are each others conjugates as L is real.

7
Proof. Let us calculate the critical points of S. To this extent we take a family
of maps qs where s ∈ R such that q0 = q and write dq ds |s=0 = q̇. Then
s

Z  
d ∂L ∂L ∂L ∂L dt ∧ dt̄
S[qs ] = − q̇ + ˙
q̄ + ∂t q̇ + ˙
∂t̄ q̄
ds s=0 T 2 ∂q ∂ q̄ ∂(∂ t q) ∂(∂ t̄ q̄) 2i
   !
dt ∧ dt̄
Z
∂L ∂L ∂L ∂L
=− q̇ − ∂t + q̄˙ − ∂t̄ .
T 2 ∂q ∂(∂t q) ∂ q̄ ∂(∂ t̄ q̄) 2i

We see that this derivative vanishes precisely when equation (10) is satisfied.

In order to get back to a Hamiltonian formalism we define a momentum-like


∂L ∂L
variable p = ∂(∂ t q)
and correspondingly p̄ = ∂(∂ t̄ q̄)
. This yields

∂L ∂L ∂L ∂L
dL = d(p∂t q + p̄∂t̄ q̄) − ∂t qdp − ∂t̄ q̄dp̄ + dq + dq̄ + dt + dt̄, (11)
∂q ∂ q̄ ∂t ∂ t̄
which motivates the definition
H(q, q̄, p, p̄, t, t̄) = p∂t q + p̄∂t̄ q̄ − L(q, q̄, ∂t q, ∂t̄ q̄, t, t̄),
where (∂t q, ∂t̄ q̄) is seen as a function of (q, q̄, p, p̄, t, t̄) by the inverse function
theorem. This requires the Legendre-like condition
 !2 
2 2 2
∂ L ∂ L ∂ L 
det  − ̸= 0.
∂(∂t q)∂(∂t̄ q̄) ∂(∂t q)2 ∂(∂t̄ q̄)2

Note again that H is real-valued.


Lemma 2.3. The Euler-Lagrange equations (10) are equivalent to the Hamil-
tonian equations
∂H
∂t q =
∂p
(12)
∂H
−∂t p = .
∂q

Proof. By the relation (11) above we get that


dL = d(p∂t q + p̄∂t̄ q̄) − dH.
So
∂H ∂L ∂H
=− = ∂t q.
∂q ∂q ∂p
Plugging in the Euler-Lagrange equations and the definition of p, the result
follows immediately.
Example 2.4. In the example L = 2∂t q∂t̄ q̄ − V (q, q̄) of the Laplace equations,
we get H = 21 pp̄ + V (q, q̄) as before and we recover equation (7).

8
3 Cuplength proof

In this section we will prove a cuplength result for solutions to the Hamiltonian
equations (12). Recall that we view t = t1 + it2 and its conjugate as the
coordinates on T2 and q = q1 + iq2 and its conjugate as the coordinates on
T2d . Consequently, Z := (q, q̄, p, p̄) denote the coordinates on T ∗ T2d , where
p = p1 + ip2 .
Theorem 3.1. Let H : T2 × T ∗ T2d → R be given by Ht,t̄ (Z) = 12 pp̄ + Ft,t̄ (Z),
where F is smooth, real-valued and has finite C 3 -norm. Then there exist at least
(2d + 1) solutions to equation (12).
Corollary 3.2. The set of Laplace equations (8) has at least (2d + 1) solutions,
when V1 and V2 have finite C 3 -norm.

Theorem 3.1 will be proven using a Floer theoretic argument along the lines
of [Sch98, AH16]. Note that similar statements have been proven in [HNS09,
GH12, GH13, AH16] for closed target manifolds and the proof for maps on a 3-
dimensional torus was given in [BF23a]. In the latter article, the proof was given
as a corollary of results from [GH12] after establishing the necessary C 0 -bounds.
This proof does not use Floer theoretic arguments. In the present article we aim
to establish a connection to the Floer theory of these systems and will prove the
necessary Fredholm and compactness results (see sections 4 and 5 respectively).
As pointed out in section 2.1, the techniques used here should also be of use for
studying holomorphic Hamiltonian systems in the sense of [Wag23] using Floer
theory.

3.1 C 0 -bounds

First, we will prove the necessary C 0 -bounds. The proofs are analogous to the
ones given in [BF23a], but as the notation with complex structures is different
in the present article and we treat the 2-dimensional torus here, the proofs will
be briefly be repeated below. First of all, define the action by
Z

AH (Z) = p∂t q + p̄∂t̄ q̄ − Ht,t̄ (Z) dV, (13)
T2

where dV = − dt∧d
2i

is the standard volume form on T2 . In the case that H
actually comes from a Lagrangian as in section 2.2, this action coincides with
equation (9). We denote
P(H) = {Z : T2 → T ∗ T2d | Z is nullhomotopic and satisfies equation (12)}
Pa (H) = {Z ∈ P(H) | a ≥ AH (Z)}.
Lemma 3.3. Let Ht,t̄ (Z) = 12 pp̄ + Ft,t̄ (Z), where F is smooth, real-valued and
has finite C 2 -norm. Then for any a ∈ R the set Pa (H) is bounded in L∞ .

9
Proof. We start by proving that Pa (H) is bounded in L2 . Note that for solutions
of equation (12) we get
Z  
∂H ∂H
AH (Z) = p + p̄ − Ht,t̄ (Z) dV
T2 ∂p ∂ p̄
Z  
1 ∂F ∂F
= pp̄ + p + p̄ − Ft,t̄ dV
T2 2 ∂p ∂ p̄
≥ h0 ||p||2L2 − h1 ,

for some h0 > 0 and h1 ≥ 0. The last inequality comes from the boundedness
of F and its derivatives. The assumption a ≥ AH (Z) now yields a uniform
L2 -bound on p. As q maps into T2d it is L2 -bounded by construction. So indeed
Pa (H) is bounded in L2 .

We now proceed by a bootstrapping argument. For Z ∈ Pa (H) we haveii


Z  
||dZ||2L2 = 4 |∂t q|2 + |∂t p|2 dV
T2
!
Z 2 2
∂H ∂H
=4 + dV
T2 ∂p ∂q
!
Z 2 2
2 ∂F ∂F
≤ ||p||L2 + 4 + dV.
T2 ∂q ∂p

As the first derivatives of F are bounded and p is uniformly bounded in L2 this


induces a uniform L2 -bound on dZ. Therefore Pa (H) is bounded in H 1 . A
similar argument shows that the H 1 -norm of dZ is bounded by the H 1 -norm
of Z and the C 2 -norm of F . Thus, we get that Pa (H) is bounded in H 2 . As
2 · 2 > 2 we get boundedness in L∞ by Sobolev’s embedding theorem.

From lemma 3.3 it follows that we can actually use a cut-off version of F .
Let χρ : [0, ∞) → R a smooth cut-off function such that χρ ≡ 1 on [0, ρ − 1]
and χρ ≡ 0 on [ρ, ∞). We can define the cut-off Hamiltonian as H̃t,ρ t̄ (Z) :=
1 ρ
2 pp̄ + F̃t,t̄ (Z)
:= 21 pp̄ + χρ (|p|2 )Ft,t̄ (Z).

Corollary 3.4. For any a ∈ R there exists some ρ > 0 such that Pa (H) =
Pa (H̃ ρ ).

3.2 Floer curves

Now that we have the necessary bounds on the orbits, we turn to Floer curves.
A Floer curve is defined as a negative L2 -gradient flow line of the action (13).
ii In the first line we use partial integration, using that Z is nullhomotopic.

10
Note that in the complex notation the L2 -inner product is given by
Z
1 
⟨X, Y ⟩L2 = Xq Yq̄ + Xq̄ Yq + Xp Yp̄ + Xp̄ Yp dV.
2 T2
It follows that the Floer curves are given by
1 ∂ H̃ ρ 1 ∂ H̃ ρ
∂s q − ∂t̄ p̄ = ∂s q̄ − ∂t p =
2 ∂ q̄ 2 ∂q
1 ∂ H̃ ρ 1 ∂ H̃ ρ
∂s p + ∂t̄ q̄ = ∂s p̄ + ∂t q= , (14)
2 ∂ p̄ 2 ∂p
where Z = (q, q̄, p, p̄) : R × T2 → T ∗ T2d and we have taken the Hamiltonian to
be H̃ ρ in view of corollary 3.4. Note that s ∈ R is a real coordinate, so that ∂s
denotes the ordinary derivative of functions on R, whereas ∂t and ∂t̄ still denote
the complex derivative and its conjugate on T2 .
Remark 3.5. Note that in the notation of remark 1.1 we can write the Floer
equation with vanishing Hamiltonian as ∂s Z + J1 ∂1 Z + J2 ∂2 Z = 0. This is
equivalent to the Fueter equation
J3 ∂s Z + J2 ∂1 Z − J1 ∂2 Z = 0,
where J3 = J1 J2 is the third complex structure from the hyperkähler structure
on the target space.

The following lemma shows that the image of a Floer curve actually sits within
a compact subset of T ∗ T2d (compare with Lemma 3.4 from [BF23a]).
Lemma 3.6. Let Z : R × T2 → T ∗ T2d be a solution to equation (14) which
converges to critical points of AH̃ ρ for |s| → ∞. Then |p(s, t)|2 ≤ ρ for any
(s, t) ∈ R × T2 .

Proof. Suppose |p(s0 , t0 )|2 > ρ for some (s0 , t0 ) ∈ R × T2 . Locally around
(s0 , t0 ) the Hamiltonian reduces to H̃ ρ (Z(s, t)) = 12 p(s, t)p̄(s, t). Therefore, on
this neighbourhood
∂s2 p = ∂s p − 2∂s ∂t̄ q̄
= ∂s p − 4∂t ∂t̄ p.
In the second equality, we use that ∂s q̄ = 2∂t p. It follows that
(∂s2 + ∆)p = ∂s2 p + 4∂t ∂t̄ p = ∂s p.

Define ϕ = 21 pp̄. Then


1 1
(∂s2 + ∆)ϕ = |∂s p|2 + 2|∂t p̄|2 + 2|∂t p|2 + p∂s p̄ + p̄∂s p
2 2
≥ ∂s ϕ.

11
The maximum principle now tells us that ϕ cannot attain a maximum in the
neighbourhood of (s0 , t0 ), contradicting the assumption that the Floer curve
converges as |s| → ∞.

3.3 Proof of theorem 3.1

Having established the C 0 -bounds we can prove theorem 3.1. The necessary
Fredholm and compactness results will be deferred to sections 4 and 5. For the
proof we will follow the lines of the paper [AH16].

To start the proof, assume there are finitely many solutions to equation (12),
since otherwise we would be done. Then their action is necessarily bounded.
Corollary 3.4 implies that we may replace H by H̃ ρ for some ρ and by lemma 3.6
also the Floer curves will take value in some compact set T2d × B ⊆ T ∗ T2d . Let
M ⊆ C ∞ (T2 , T2d × B) be the subset of nullhomotopic maps. Just like in [AH16,
§3.1] we define a smooth family of functions βr : R → [0, 1] for r ≥ 0 such that

• βr (s) = 0 for s ≤ −1 and s ≥ (2d + 1)r + 1 for all r ≥ 0,


• βr |[0,(2d+1)r] ≡ 1 for r ≥ 1,

• 0 ≤ βr′ (s) ≤ 2 for s ∈ (−1, 0) and


0 ≥ βr′ (s) ≥ −2 for s ∈ ((2d + 1)r, (2d + 1)r + 1),
• limr→0+ βr = 0 in the strong topology on C ∞ .

See [AH16] for the precise conditions on βr and figure 1 for an idea of what they
look like. We will use the functions βr to interpolate between the quadratic
Hamiltonian with and without non-linearity. To this extent define

ρ,r 1
H̃s,t,t̄ (Z) = pp̄ + βr (s)F̃t,ρt̄ (Z).
2

One can see the Hamiltonian H̃ ρ,r as the quadratic Hamiltonian H 0 (Z) = 21 pp̄
with the non-linearity F̃ ρ switched on just on an interval whose length is deter-
mined by r. Correspondingly, we have the action Gr,s := AH̃ ρ,r , interpolating
between the actions AH 0 and AH̃ ρ .

With this setup we can introduce the moduli space

M = {(r, Z) | r ≥ 0, Z : R → M, (⋆1) − (⋆3)}, (15)

where (⋆1) − (⋆3) are given below. Note that we can view Z as a contractible
map R × T2 → T2d × B. The condition (⋆1) is that Z is a negative gradient line

12
Figure 1: The functions βr . Source: [AH16]. In our case we take k = cl(T2d ) =
2d.

of Gr,s , meaning a Floer curve for H̃ ρ,r . In formula this means


 
∂q̄
1 ∂q  ρ,r
∂s Z + ∂ Z =  
 H̃ (Z), (⋆1)
2 ∂p̄ 
∂p
where  
0 0 0 −∂t̄
0 0 −∂t 0 
∂ = 
0

∂t̄ 0 0 
∂t 0 0 0
and as before Z = (q, q̄, p, p̄). Condition (⋆2) is finiteness of energy
Z ∞Z
E(Z) = |∂s Z(s, t)|2 dV ds < +∞ (⋆2)
−∞ T2

and the final condition is convergence of the p-coordinates


p(s, t) → 0 as |s| → ∞. (⋆3)
We also introduce the subspaces
M[0,R] (T2d ) := {(r, Z) ∈ M | 0 ≤ r ≤ R and lim Z(s) ∈ T2d × {0}}
s→±∞

MR (T2d ) := {Z | (R, Z) ∈ M[0,R] (T2d )},


where T2d × {0} ⊆ M denotes the set of constant maps. This coincides with the
set of critical points of AH 0 . Note that for R = 0 we get that M0 (T2d ) is equal
to this set of critical points.

It follows from theorem 4.1 and theorem 5.5 that M is a compact 1-dimensional
manifold, possibly after a small perturbation of F . Also the spaces M[0,R] (T2d )
and MR (T2d ) will then be compact manifolds (again after possibly perturbing
F , see [AH16]). We have an evaluation map
evR : MR (Z) → (T2d × B)2d
Z 7→ (Z(R, t0 , t̄0 ), Z(2R, t0 , t̄0 ), . . . , Z(2dR, t0 , t̄0 )),

13
where (t0 , t̄0 ) ∈ T2 is some basepoint. For R = 0 this gives the diagonal embed-
ding T2d ,→ (T2d × B)2d .

The rest of the proof goes analogously to [AH16]. Take Morse functions
f1 , . . . , f2d , f∗ on T2d and extend to Morse functions f¯α for α = 1, . . . , 2d on
T2d × B using a positive definite quadratic form on a tubular neighbourhood
of T2d in the direction normal to T2d . We also choose Riemannian metrics
g1 , . . . , g2d , g∗ on T2d × B. For critical points xα ∈ T2d of fα and x± ∗ ∈ T
2d
of
f∗ we define
n o
M(R, x1 , . . . , x2d , x− + 2d
∗ , x∗ ) := Z ∈ MR (T ) (⋆4) − (⋆6) ,

where
Y
evR ∈ W s (xα , f¯α , gα ) (⋆4)
α
lim Z(s) ∈ W u (x−
∗ , f∗ , g∗ ) (⋆5)
s→−∞

lim Z(s) ∈ W s (x+


∗ , f∗ , g∗ ). (⋆6)
s→+∞

Here, W u and W s denote the unstable and stable manifolds of the criti-
cal points for the gradient flows of the indicated Morse functions. Again,
by choice of perturbations, we may assume these are all smooth manifolds.
Note that M(0, x1 , . . . , x2d , x− + s
(xα , fα , gα ) ∩ W u (x−
T
∗ , x∗ ) = α W ∗ , f∗ , g∗ ) ∩
W s (x+
∗ , f ,
∗ ∗g ).

Define the following operation on Morse co-chain groups


O
θR : CM ∗ (fα ) ⊗ CM∗ (f∗ ) → CM∗ (f∗ )
α
X
x1 ⊗ · · · ⊗ x2d ⊗ x−
∗ 7→ #2 M(R, x1 , . . . , x2d , x− + +
∗ , x∗ ) · x∗ .
x+
∗ ∈Crit(f∗ )

Here, #2 denotes the parity of the set if it is discrete and 0 if it not. For
generic choices of Morse functions and Riemannian metrics, the map θ0 realizes
the usual cup-product in cohomology (see [AM10]). This map does not vanish
as the cuplength of T2d is 2d. Moreover, θR is chain homotopic to θ0 and can
therefore not vanish either. Thus there have to be critical points xα of fα and x± ∗
of f∗ of index at least 1 such that the moduli spaces M(n, x1 , . . . , x2d , x− +
∗ , x∗ )
are non-empty for all n ∈ N.

Let Zn ∈ M(n, x1 , . . . , x2d , x− +


∗ , x∗ ) and

Z (α) = lim Zn (· + nα),


n→∞

for α = 1, . . . , 2d. By the compactness discussed in section 5 these limits exist



in Cloc (R, M ). The Z (α) satisfy the Floer equation (14) for the Hamiltonian H̃ ρ

14
and they converge to critical points yα± of AH̃ ρ for s → ±∞ such that

AH̃ ρ (y1− ) ≥ AH̃ ρ (y1+ ) ≥ AH̃ ρ (y2− ) ≥ AH̃ ρ (y2+ ) ≥ . . . ≥ AH̃ ρ (y2d
− +
) ≥ AH̃ ρ (y2d ).

The only thing left to do is show that at least 2d+1 of these points are different,
by showing that the inequalities AH̃ ρ (yα− ) > AH̃ ρ (yα+ ) are strict.

Note that since Crit(AH̃ ρ ) is finite by assumption, we may assume critical points
of AH̃ ρ do not intersect the stable manifolds W s (xα , f¯α , gα ). If AH̃ ρ (yα− ) =
AH̃ ρ (yα+ ), then Z (α) is constant and therefore a critical point of AH̃ ρ . Also note
that by definition Zn (nα) ∈ W s (xα , f¯α , gα ) for any n. Therefore Z (α) (0) =
limn→∞ Zn (nα) lies in the closure of a stable manifold, which itself is a union
of stable manifolds. This is in contradiction with Z (α) being a critical point of
AH̃ ρ . Thus indeed the inequalities AH̃ ρ (yα− ) > AH̃ ρ (yα+ ) are strict and we get
that y1− , y2− , . . . , y2d
− +
and y2d form 2d + 1 different critical points of AH̃ ρ . These
correspond to 2d + 1 distinct solutions of equation (12)

4 Fredholm theory

In this section we will prove the following theorem.


Theorem 4.1. The moduli space M from equation (15) is a 1-dimensional
manifold for generic choice of F .

As is usual in Floer theory, this will be proven by describing M as the zero set
of a Fredholm map. The results are standard, so we will not go into full detail.
However, we would like to give the proof of the Fredholmness, to stress that in
our setting elementary Fourier arguments can be used.

We let F denote the Floer map


 
∂q̄
1 ∂q 
F(r, Z) = F r (Z) = ∂s Z + ∂ Z −   ρ,r

2 ∂p̄  H̃ (Z),
∂p

where as before Z = (q, q̄, p, p̄). This means that M = F −1 (0). We start by
proving F is a nonlinear Fredholm map when acting on the appropriate spaces.

4.1 Linearization of the Floer map

Let r be fixed. Then the linearization of F r at Z is given by

(dF r )Z = D − βr (s)SZ , (16)

15
where D = 12 ∂s + ∂ − 1
2P,
   
0 0 0 0 0 1 0 0
0 0 0 0 1 0 0 0
P =  and SZ =   HessZ F̃ ρ .
0 0 1 0 0 0 0 1
0 0 0 1 0 0 1 0
3
By H k we denote the Sobolev space W k,2 . Note that for k > 2 we have an
embedding H k (R × T2 , R4d ) ,→ C 0 (R × T2 , R4d ).
Lemma 4.2. The operator D : H k (R × T2 , R4d ) → H k−1 (R × T2 , R4d ) is a
bijection for k > 23 .

Proof. For Y ∈ H k := H k (R × T2 , R4d ) consider the Fourier expansion


Z X
Y (s, t, t̄) = Ŷ (ξ, m1 , m2 )eiξs eim1 t1 eim2 t2 dξ
m1 ,m2 ∈Z
Z
1 1
X
= Ŷ (ξ, m1 , m2 )eiξs e 2 t(im1 +m2 ) e 2 t̄(im1 −m2 ) dξ.
m1 ,m2 ∈Z

Then on the Fourier coefficients the operator D acts as


 
0 0 0 −im1 + m2
1 1 0 0 −im 1 − m2 0  1
 − P.
D̂(ξ, m1 , m2 ) = iξ + 
2 2 0 im1 − m2 0 0  2
im1 + m2 0 0 0
1
We compute det(D̂(ξ, m1 , m2 )) = 16 (m21 + m22 + ξ 2 + iξ)2 . This means only
constant functions can be in the kernel of D, however the only constant function
that is square-integrable on R × T2 is the zero function. So ker D = 0.

Now, for surjectivity let W ∈ H k−1 and define


(
D̂(ξ, m1 , m2 )−1 Ŵ (ξ, m1 , m2 ) (ξ, m1 , m2 ) ̸= (0, 0, 0)
Ŷ (ξ, m1 , m2 ) =
0 (ξ, m1 , m2 ) = (0, 0, 0).

Note that by the calculation of the determinant above D̂(ξ, m1 , m2 )−1 exists.
Now by definition we have that DY = W provided Y ∈ H k . To see this, note
that the eigenvalues of D̂(ξ, m1 , m2 ) are
 
1
q
λ± (ξ, m1 , m2 ) = i i + 2ξ ± i 1 + 4m21 + 4m22 .
4
For m21 + m22 large enough the norms of these eigenvalues can be bounded from
below. In other words, there exists some N such that for m21 + m22 > N , we
have
1 1 1
|λ± (ξ, m1 , m2 )|2 ≥ ξ 2 + m21 + m22 .
4 8 8

16
Let pr denote the projection onto the subspace where m21 + m22 > N . Then the
H k -norm of pr Y is
Z X
2
||pr Y ||H k = (1 + ξ 2 + m21 + m22 )k |Ŷ (ξ, m1 , m2 )|2 dξ
m21 +m22 >N

(1 + ξ 2 + m21 + m22 )k
Z X
2
≤ 1 2 1 2 1 2 |Ŵ (ξ, m1 , m2 )| dξ
m21 +m22 >N 4 ξ + 8 m1 + 8 m2

1 + ξ 2 + m21 + m22
≤ |||W ||2H k−1 sup 1 2 1 2 1 2 < +∞.
m21 +m22 >N 4 ξ + 8 m1 + 8 m2

Here, the third line follows from Hölders inequality. Thus Y ∈ H k , proving that
D is surjective.
Corollary 4.3. The Floer map F is a Fredholm map of index 1.

Proof. Note that the operator βr (s)SZ is compact as a map H k → H k−1 . Thus,
by equation (16) and lemma 4.2 the differential of F r is Fredholm of index 0.
Varying r as well, we see that (dF)(r,Z) must be Fredholm of index 1.

4.2 Transversality

In this section we will show that for generic choice of non-linearity the differ-
ential of the Floer map is surjective. Combined with corollary 4.3 this yields
theorem 4.1.

Note that the moduli space M depends on the chosen non-linearity βr (s)F̃tρ .
By varying these non-linearities for each l, we get the universal moduli space
[
M
f := M(Fs,t ),
F ∈S

where S = Ccl (R × T2 × T2d × B, R) is the space of compactly supported smooth


functions Fs,t on T2d × B ⊆ T ∗ T2d . We denote by M(Fs,t ) the moduli space
M of equation (15), where the non-linearity βr (s)F̃tρ is replaced by Fs,t . Note
that in this case Fs,t is not necessarily the product of a function in s, with a
time-dependent non-linearity on T2d × B.

This universal moduli space can be seen as the zero set of the map
 
∂q̄
r 1 ∂q  F
F(r,
e Z, F ) = Fe (Z, F ) = ∂s Z + ∂ Z −   Hs,t (Z),
∂p̄ 
2
∂p

17
where Hs,tF
(Z) = 21 pp̄ + Fs,t (Z) and F vanishes for s ∈
/ [−1, (2d + 1)r + 1]. The
result would follow from Sard-Smale’s theorem, if we know that the differential
of Fer is surjective at every (Z, F ) ∈ M.f Let (Z, F ) ∈ M f and take a variation
(Y, G) ∈ H k (R × T2 , R4d ) × Ccl (R × T2 × T2d × B, R). Then the linearization of
Fer at (Z, F ) can be decomposed as

(dFer )(Z,F ) · (Y, G) = DF1 G + DZ


2
Y.

Assume we have some non-zero W ∈ H k−1 (R × T2 , R4d ) in the orthogonal com-


plement of the image of (dFer )(Z,F ) . Then ⟨W, DZ 2
Y ⟩ = 0 for all Y , so that
2 ∗

W ∈ ker (DZ ) . Note that for s ∈ / [−1, (2d + 1)r + 1] the non-linearity Fs,t
2 ∗
vanishes so that (DZ ) = 12 ∂s − ∂ − 21 P . This operator satisfies a maximum prin-
ciple, similar to the one used in lemma 3.6. Therefore, by standard arguments
W (s, t) ̸= 0 for some (s, t) ∈ [−1, (2d+1)r+1]×T2 . Thus, we can find some vari-
ation G with support for s ∈ [−1, (2d + 1)r + 1], such that ⟨W, DF1 G⟩ = ̸ 0. This
contradicts the assumption that ⟨W, (dFer )(Z,F ) · (Y, G)⟩ = 0 for any (Y, G).
We must conclude (dFer )(Z,F ) is surjective. This finishes the proof of theo-
rem 4.1.

5 Compactness

In this section we will prove compactness of the moduli space (15) that is needed
in the proof of theorem 3.1. We start by proving a uniform energy bound, using
a standard argument.
Lemma 5.1. For all (r, Z) ∈ M we have
Z !
E(Z) ≤ 2||F̃ ρ ||Hofer := 2 sup F̃tρ (Z ′ ) − inf F̃tρ (Z ′ ) dV.
T2 Z ′ ∈T ∗ T2d Z ′ ∈T ∗ T2d

Proof. By definition
Z ∞Z Z ∞
2
E(Z) = |∂s Z(s, t)| dV ds = − ⟨grad AH̃ ρ,r (Z), ∂s Z⟩ ds,
−∞ T2 −∞

since Z is a negative gradient line of AH̃ ρ,r . Writing this out further we get
Z ∞

E(Z) = − dAH̃ ρ,r Zs (∂s Z) ds
−∞
Z ∞ Z ∞
d ∂AH̃ ρ,r
=− (AH̃ ρ,r (Zs )) ds + (Zs ) ds
−∞ ds −∞ ∂s
Z ∞Z
= − lim AH̃ ρ,r (Zs ) + lim AH̃ ρ,r (Zs ) − βr′ (s)F̃tρ (Zs ) dV ds.
s→∞ s→−∞ −∞ T2

18
Note that the first two terms on the last line vanish. Also, βr′ vanishes every-
where outside [−1, 0] ∪ [(2d + 1)r, (2d + 1)r + 1] and is positive and negative
respectively on those two intervals. Therefore,
Z 0Z Z (2d+1)r+1 Z
ρ
E(Z) = − ′
βr (s)F̃t (Zs ) dV ds − βr′ (s)F̃tρ (Zs ) dV ds
−1 T2 (2d+1)r T2
Z !
≤2 sup F̃tρ (Z ′ ) − inf F̃tρ (Z ′ ) dV
T2 Z ′ ∈T ∗ T2d Z ′ ∈T ∗ T2d

=: 2||F̃ ρ ||Hofer .

Now we want to relate the energy to the L2 -norm of the derivatives of Z. To


this end we define the energy density
1 1
eZ (s, t) = |dZ(s, t)|2 = |∂s Z(s, t)|2 + |∂t Z(s, t)|2 + |∂t̄ Z(s, t)|2 .
2 2
Lemma 5.2. Let K = [−M, M ] × T2 ⊆ R × T2 . Then there exists a constant
C = C(K) such that for any (r, Z) ∈ M
Z
eZ (s, t) dV ds ≤ E(Z) + C(K).
K

Proof. We calculateiii
Z Z Z 
1 
eZ (s, t) dV ds = |∂s Z|2 dV ds + 2 |∂t q|2 + |∂t p|2 dV ds
K 2 K K
!
Z Z 2 2
1 ∂H 1 1 ∂H
= |∂s Z|2 dV ds + 2 − ∂s p̄ + ∂s q̄ − dV ds
2 K K ∂p 2 2 ∂q
!
Z 2 2
∂H ∂H
≤ E(Z) + 2 + dV ds
K ∂q ∂p
 
2 2
Z Z ρ ρ
1  βr (s) ∂ F̃ ∂ F̃
≤ E(Z) + |p|2 dV ds + 2 + βr (s)  dV ds.
2 K K ∂q ∂p

Now, as in lemma 3.6 we have that |p(s, t)|2 < ρ for all (s, t) ∈ K. Combined
with the C 1 -boundedness of F , we can estimate the right-hand side purely in
terms of E(Z) and the volume of K. This proves the lemma.

In order to get point-wise bounds on the derivatives, we want to use the Heinz
trick (theorem 5.4). We first need to show that the energy density satisfies an
appropriate inequality.
iii The first equality follows from partial integration, using that Z is nullhomotopic.

19
Lemma 5.3. The energy density e := eZ satisfies the inequality
3
(∂s2 + ∆)e ≥ −c(1 + e 2 ), (17)
for some c > 0.

Proof. See appendix A.


Theorem 5.4 (Heinz trick, see [HNS09]). Let M be an (m + 1)-dimensional
Riemannian manifold for 23 < m+3
m+1 and e M → R≥0 a smooth function
: satis-
R
fying equation (17). Then for K ⊆ M compact supK e is bounded by K e.

Now we are ready to prove the desired compactness theorem.



Theorem 5.5. The moduli space M is relatively compact in the Cloc -topology.

Proof. Let (rn , Zn ) ∈ M for all n ∈ N and K = [−M, M ] × T2 ⊆ R × T2 . By


lemma 3.6 the sequence Zn |K is a bounded family of smooth functions
R on K.
By combining lemmas 5.1 and 5.2 we find a uniform bound on K eZn . Then
the Heinz trick (theorem 5.4) combined with the density estimate in lemma 5.3
yields a uniform pointwise bound on eZn (s, t) for all (s, t) ∈ K. By the mean-
value theorem it follows that the sequence (Zn ) is uniformly equicontinuous.
0
The theorem of Arzelà-Ascoli then yields a converging subsequence in the Cloc -
iv
topology. Finally, the ellipticity of the Floer operator yields convergence of

the subsequence in Cloc .

A Proof of lemma 5.3

The proof of lemma 5.3 is easier to write out in the ’real’ notation using T =
(t1 , t2 ) ∈ T2 and Z = (q1 , q2 , p1 , p2 ). Note that it is just a matter of notation so
that the proof still holds in the complex notation as well. Recall that J∂ is an
operator satisfying J∂2 = −(∂12 + ∂22 ). We can write J∂ = J1 ∂1 + J2 ∂2 for some
matrices J1 , J2 satisfying J12 = J22 = −Id and J1 J2 + J2 J1 = 0 (see [Bri06]). For
any (r, Z) in M, it holds that Z satisfies the Floer equation
∂s Z + J∂ Z = ∇H̃ ρ,r (Z), (18)
ρ,r
where H̃s,T (Z) = 21 (p21 + p22 ) + βr (s)F̃Tρ (Z). We define t0 := s and ξi := ∂i Z for
i = 0, 1, 2. From the identity
2
X
(∂s − J∂ )(∂s + J∂ ) = L := ∂i2
i=0
iv Notethat ( 21 ∂s − ∂)( 12 ∂s
 +
∂) = 1 2

4 s
+ ∂t ∂t̄ , so the Floer operator is a factor of the
Laplacian on R × T2 .

20
we see that the ξi satisfy the equation
Lξi = (∂s − J∂ )(σZ ξi ), (19)

where σZ is the Hessian of H̃ ρ,r at Z.

Note that in this notation the energy density is given by


2
1X
e(s, T ) = |ξi (s, T )|2 .
2 i=0

We calculate
2
X 2
X
Le = |∂j ξi |2 + ⟨ξi , Lξi ⟩
i,j=0 i=0
2
X 2
X
= |∂j ξi |2 + ⟨ξi , (∂s − J∂ )(σZ ξi )⟩,
i,j=0 i=0

where we have used equation (19) in the second equality. Now note that the
chain rule gives that
∂j (σZ ξi ) = T σZ · ξj · ξi + σZ ∂j ξi .
Denote J0 = −Id. Then
2 
X 
Le = |∂j ξi |2 − ⟨(Jj σZ )T ξi , ∂j ξi ⟩ − ⟨ξi , Jj T σZ · ξj · ξi ⟩
i,j=0
2
!
2 2
X 1 1
= ∂j ξi − (Jj σZ )T ξi − (Jj σZ )T ξi − ⟨ξi , Jj T σZ · ξj · ξi ⟩ .
i,j=0
2 4

Note that since F is C 3 -bounded, we have that both σZ and T σZ are bounded.
Thus
2
(Jj σZ )T ξi ≤ c1 |ξi |2 ≤ c2 e
and
3
⟨ξi , Jj T σZ · ξj · ξi ⟩ ≤ c3 |ξi |2 |ξj | ≤ c4 e 2 .
Combined we get that  
3
Le ≥ −c5 e + e 2 .
3
Finally, Young’s inequality with p = 3, q = 2 gives that
1 2 3
e≤ + e2 ,
3 3
so that
3
Le ≥ −c(1 + e 2 ).

21
References
[AH16] Peter Albers and Doris Hein. Cuplength estimates in Morse cohomol-
ogy. Journal of Topology and Analysis, 8(02):243–272, 2016.
[AM10] Peter Albers and Al Momin. Cup-length estimates for leaf-wise inter-
sections. In Mathematical Proceedings of the Cambridge Philosophical
Society, volume 149, pages 539–551. Cambridge University Press, 2010.
[BD99] Thomas J Bridges and Gianne Derks. Unstable eigenvalues and the
linearization about solitary waves and fronts with symmetry. Proceed-
ings of the Royal Society of London. Series A: Mathematical, Physical
and Engineering Sciences, 455(1987):2427–2469, 1999.
[BF23a] Ronen Brilleslijper and Oliver Fabert. From Euclidean field theory
to hyperkähler Floer theory via regularized polysymplectic geometry.
arXiv preprint 2311.18485, 2023.
[BF23b] Ronen Brilleslijper and Oliver Fabert. Regularized polysymplectic ge-
ometry and first steps towards Floer theory for covariant field theories.
Journal of Geometry and Physics, 183:104703, 2023.
[Bri06] Thomas J Bridges. Canonical multi-symplectic structure on the total
exterior algebra bundle. Proceedings of the Royal Society A: Math-
ematical, Physical and Engineering Sciences, 462(2069):1531–1551,
2006.
[Cie94] Kai Cieliebak. Pseudo-holomorphic curves and periodic orbits on
cotangent bundles. Journal de Mathématiques Pures et Appliquées,
73:251–278, 1994.
[GH12] Viktor L Ginzburg and Doris Hein. Hyperkähler Arnold conjec-
ture and its generalizations. International Journal of Mathematics,
23(08):1250077, 2012.
[GH13] Viktor L Ginzburg and Doris Hein. The Arnold conjecture for Clifford
symplectic pencils. Israel Journal of Mathematics, 196:95–112, 2013.
[Gün87] Christian Günther. The polysymplectic Hamiltonian formalism in field
theory and calculus of variations. I. The local case. Journal of differ-
ential geometry, 25(1):23–53, 1987.
[HNS09] Sonja Hohloch, Gregor Noetzel, and Dietmar A Salamon. Hypercon-
tact structures and Floer homology. Geometry & Topology, 13(5):2543–
2617, 2009.
[Kan93] Igor V Kanatchikov. On the canonical structure of the De Donder-
Weyl covariant Hamiltonian formulation of field theory I. Graded Pois-
son brackets and equations of motion. arXiv preprint hep-th/9312162,
1993.

22
[Sch98] Matthias Schwarz. A quantum cup-length estimate for symplectic fixed
points. Inventiones mathematicae, 133(2):353–397, 1998.
[Wag23] Luiz Frederic Wagner. Pseudo-Holomorphic Hamiltonian Systems.
arXiv preprint 2303.09480, 2023.

23

You might also like