Zariski-Van Kampen Method and Monodromy in Complexified Integrable Systems

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 14

Zariski-van Kampen Method and Monodromy in

Complexified Integrable Systems


∗a
Zhiyuan Liu
a
School of Mathematical Science, Capital Normal University

Abstract
We computed the fundamental groups of non-singular sets of some complexified Hamiltonian integrable
arXiv:2312.12761v2 [math.DS] 21 Dec 2023

systems by Zariski-van Kampen method. By our computation, we determined all possible monodromy in
complexified planar Kepler problem and spherical pendulum, that is their monodromy groups. We also gave
an answer to the conjecture proposed by You and Sun.
Key Words: Complexification; Kepler Problem; Algebraic Curve; Braid Monodromy; Zariski-Van Kam-
pen Method.
AMS Classification Code: 37J35; 37N05; 53D22.

1 Introduction
In a Liouville integrable system (M, ω, H, µ), the famous Liouville-Arnold theorem asserts that the generic
fibers of the first integrals µ are compact Lagrangian tori, these tori are also called angle coordinates, they play
an important role in Hamiltonian dynamics and symplectic topology. If the angle coordinates exist globally,
that implies geometrically, the symplectic manifold (M, ω) can be endowed with a global Hamiltonian torus
action (the dimension of the torus is half of M ), and the associated moment map µ coincides with the first
integrals µ. However, in general, such coordinates cannot be constructed globally, because the Lagrangian toric
fibrations are not always trivial.
To be specific, suppose µ = (F1 , ..., Fn ) are n Poisson commuting almost independent first integrals of the
Hamiltonian dynamical system (M, ω, H) (where H = F1 ) , let B ⊂ Rn be the set of regular values such that
for each b ∈ B the fiber µ−1 (b) is generic (the complement of regular values will be called singular set in this
paper), then the restriction of µ gives a toric fibration (see figure 1):

µ : M ′ −→ B

where M ′ is the primage of B. The fiber µ−1 (b) ∼ = Tn is a Lagrangian torus (n = 1


2 dim M ). The action-angle
coordinate exists globally if and only if the fibration is trivial.

Figure 1: Toric fibration determined by Liouville-Arnold theorem

In 1980, Duistmaat developed a notion in [1] called monodromy of a Liouville integrable system, which can
perfectly describe the obstructions of the existence of global angle coordinates.
∗ 2200502078@cnu.edu.cn

1
To define the Hamiltonian monodromy, one can consider the period lattice bundle Λ over B, where each
fiber at b ∈ B is replaced by the lattice (called period lattice, see [2, §21.2]) Λb of the original torus µ−1 (b) ∼
= Tn .
Notice that when we restrict the period lattice bundle Λ on some loop γ(t) ∈ π1 (B, b) based at b ∈ B, we
gain a Z⊕n -bundle over γ ∼ = S1 , namely

Λb −→ γ ∼
a
Λ|γ := = S1
b∈γ

The bundle Λ|γ is determined by a matrix g(b) ∈ Aut(Λb ) = GLn (Z), hence it defines a group homomor-
phism, called the monodromy map:

Mon : π1 (B, b) −→ GLn (Z), γ 7→ g(γ(0))

The image of Mon is called the monodromy group of the integrable system (M, ω, H, µ).
Monodromy is very important in studying Liouville integrable systems. Indeed, Duistmaat showed in [1]
that the action-angle coordinate of a system (M, ω, H, Fi ) exits globally precisely if its monodromy map is trivial
(see also [3, p. 393]). After that, various examples of non-trivial monodromy were found, the first such example
was the spherical pendulum, found by Cushman in [3] and computed by Duistmaat in [1]. A typical example
of a system with trivial monodromy is the Kepler problem, which describes the motion of two bodies under the
grivation with one body fixed at the origin. The triviality can be proved by using Pauli and Delauny variables,
see [4, 5] for reference. Besides, as was indicated by [6, 7], the monodromy can explain some quantum effects as
well.
Some studies in recent decades illustrate that the complexification of some mechanical systems will let
us have more benefits. For example, the authors in [8] used the complexification of planar N −body problem
proved the finiteness of configurations in planar 5-body problem. The authors in [9] computed the monodromy
of complexified spherical pendulum, it has one more rank than the real case. In a recent work [10] of You and
Sun, they discovered the non-trivial monodromy phenomenon in complexified planar Kepler problem.
In this paper, we will keep studying the monodromy behavior in complexified planar Kepler problem and
spherical pendulum. In particular, we will determine all possible monodromy, that is the monodromy groups
of them, by the tools from algebraic geometry and topology. In particular, our computation for complexified
planar Kepler problem will give an answer to the conjecture proposed by You and Sun at the end of [10].
In practice, it is difficult to determine all possible monodromy, the main reason is due to the difficulties
in computing the fundamental group π1 (B, b), especially in the complex settings. In a complexified integrable
system, the set B will be the complement of some hypersurfaces in the affine space Cn , and it’s hard to imagine its
figure intuitively. Fortunately, in our cases, the regular sets B are both the complement of some affine algebraic
curves in C2 , we can compute its fundamental group by tools from algebraic geometry, namely Zariski-van
Kampen method, it is a powerful method in studying the topology of algebraic varieties, for example [11, 12].
Our main result is:
Theorem 1.1 (Theorem 3.1 and Theorem 4.1). (1). The fundamental group of non-singular set of complexified
spherical pendulum is Z, and its monodromy group is Z.
(2). The fundamental group of non-singular set of complexified planar Kepler problem is

π 1 C2 \ S = Z ⊕ Z


Hence consequently, the monodromy group is Z2 ⊕ Z2 .


The arrangement of the rest of this paper is as follows:
In section 2, we will introduce the preliminary knowledge in algebraic geometry, including the notion of a
holomorphic integrable system. The main tool, that is Zariski-van Kampen method, will be introduced in detail
in §2.2. The monodromy of complexified spherical pendulum will be studied in section 3. This is the easier case
because all functions are polynomials, hence the complexification is directly. Then in section 4, we will study
the complexified planar Kepler problem, the complexification in this case will need a bit modification.

2 Preliminaries
2.1 Complex Integrable System
We first recall that, an Abelian variety T is a compact complex torus T = Cg /Λ which can be embedded
to some projection space, Λ ∼ = Z2g is called the lattice of T .
It is well-known that not all compact complex tori admit a projective embedding, it depends on the
properties of lattice Λ. The conditions on the lattice such that the torus is projective are called Riemann
conditions.

2
Theorem 2.1 (Riemann Conditions, [13]). A compact complex torus T = Cg /Λ is an Abelian variety if and
only if Λ admits a Riemann form, that is a positive-definite Hermitian form H on Cg such that the image part
Im H(Λ, Λ) ⊂ Q.
Now, we can formulate what is a complex integrable system. Let (M, ω) be a real symplectic manifold of
dimension 2n, H a Hamiltonian defined on M .
Definition 2.1 (Complex integrable system, [14]). A real Hamiltonian system (M, ω, H) is called a complex
integrable system, if there exists a complex 2n−dimensional holomorphic symplectic manifold (X, τ ), that is X
is a complex manifold together with a symplectic (2, 0)−form ω, and a holomorphic map

µ = (f1 , ..., fn ) : X −→ B ⊂ Cn

onto a Zariski dense open subset B, such that


(1). (M, ω) is the real part of (X, τ ), and ω is the restriction of τ along ω.
(2). µ is submersive and proper.
(3). f1 , ..., fn are Poisson commuting functions on (X, τ ).

Remark 2.1. (1). In algebraic geometry, an algebraic integrable system is just referred to an algebraic symplec-
tic variety, together with n Poisson commuting functions, such that the generic fibers are Lagrangian Abelian
varieties. Here, we take the Mumford’s definition because all complex integrable systems that will be considered
in this paper are the complexification from the real case. However, the Mumford’s definition implies that the
generic fibers are affine parts (or can be extended) to Abelian varieties. For a discussion of some other possible
definitions of complex integrable systems, we refer to [15, §5].
(2).Different from the Kähler manifolds (where ω is a (1,1)-form), holomorphic symplectic manifolds have
more special geometry, see [16, 17] for more detail. ♣
There are various examples of complex integrable systems. The simplest examples can be constructed from
the complexification of some real Liouville integrable systems, and the complexified spherical pendulum and
planar Kepler problem are such examples. More interesting example is the Hitchin systems [18], which are
defined on the cotangent bundle of the moduli space of Higgs bundles.

2.2 Braid Monodromy and Zariski-van Kampen Method


This section will introduce the necessary tools from algebraic geometry and algebraic topology, namely
braid monodromy and Zariski-van Kampen method, which will be used in proving our main results. A good
reference in this subject can be found in [19].

2.2.1 Braid Groups


We first introduce what are geometric braids, for more detailed contents on braid groups can be found
in [20].

Figure 2: geometric braids

Let Π1 , Π2 be two parallel planes in R3 , in particular, we assume they both parallel to the xOy-plane, and
Π2 is above Π1 in the sense that Π2 has larger z−component. There are n marked ordered positions on each
plane Πi , namely 1, 2, ..., n, we assume the lines joining the corresponding positions are all vertical to the both
planes Πi , see figure 2.

3
Definition 2.2 ( [20]). A geometric braid on n strings is a family of simple arcs (β1 , β2 , ..., βn ) in R3 , where
each arc β is connecting the i−th position in π2 and the σ(i)−th position in Π1 for some σ ∈ Sn , and each two
arcs do not intersect.
If all strings in a braid are connecting correspondingly by the order, we call such a braid the identity,
denoted by 1. Two n−string geometric braid (β1 , ..., βn ) and (γ1 , ..., γn ) are called equivalent, if they can be
continuously deformed to each other.
The product or composition of two braids is defined to be the juxtaposition, see figure 3. Clearly, the braids
product satisfy the associative law, and any braid product with the identity is itself. Hence the collection of all
braids with n strings forms a group, called braid groups, denoted by Bn .

Figure 3: composition of two braids

Although braids can be complicated, they can be write as the products of a sequel of simple braids, denoted
σi . σi is the braid with just the i−th and the (i + 1)−th position interchange and only once, see figure 4. These
σi forms the generators in the braid group Bn , and the generation is given by [20]
* ( +
σi σj = σj σi |i − j| ≥ 2
Bn = σ1 , ..., σn−1
σi σi+1 σi = σi+1 σi σi+1 1 ≤ i ≤ n

Figure 4: simple braid σi

Remark 2.2 (Pure Braids). If we request every arc βi in the n-string braid (β1 , ..., βn ) to having same starting-
end position, then such a braid will be called the pure braid, the group formed by all pure braids are called
pure braid group, denoted by Pn , it is clearly that we have the group exact sequence
1 −→ Pn −→ Bn −→ Sn −→ 1
. ♣
There are some classical models of braid groups Bn which will be needed in this paper.
Example 2.1. Suppose there are n orderless points moving on the complex plane C without collisions, the
motion of these n points forms a configuration space X, that is
X = {(z1 , ..., zn ) ∈ Cn |zi ̸= zj , i ̸= j} /Sn
= (Cn \ {zi = zj , i ̸= j}) /Sn := X ′ /Sn
The fundamental group of this configuration space is actually the braid group
π1 (X, x0 ) ∼
= Bn

4
In fact, note that the quotient X ′ −→ X defines an Sn -bundle over X, if we choose a section s such that
s(x0 ) = x̃0 ∈ X ′ , then by the homotopy lifting property, each loop in π1 (X, x0 ) can be lifted to a path in X ′
starting at x̃0 . Notice that every path in X ′ is the motion from one complex plane with ordered positions to
the other, and it is in fact an n-string braid illustrated in figure 2, hence π1 (X, x0 ) is actually the braid group
Bn . ♣

Example 2.2 (Artin representation). Another interesting model is the mapping class group Mn0,1 of the
compact Riemann surface with 1 genus, 1 boundary and n marked points, that is the n−punctured disk
D \ {a1 , ..., an }, it was proved in [21] that
Mn0,1 ∼
= Bn
Mapping class group has its natural action on the fundamental group

π1 (D \ {a1 , ..., an }, a0 ) = Z
| ∗ ·{z
· · ∗ Z} := Fn
n−times

If we order the generators in Fn in an appropriate way, say g1 , ..., gn , then the action coincides with the Artin
representation [21]:
α : Mn0,1 ∼
= Bn −→ Aut (Fn )

gi+1
 j=i
(1)
α(σi )(gj ) = gj gi gj−1 j = i + 1

gj otherwise

This model will be used in constructing the braid monodromy.♣

Example 2.3. Let C[y]n be the set of monic polynomials with degree n, let ∆ ⊂ C[y]n be the subset consisting
of polynomials with multiple roots, i.e those monic polynomials f with vanishing discriminants ∆f = 0, then
the fundamental group π1 (C[y]n \ ∆) is exactly Bn .
Indeed, by taking coefficients, the set C[y]n \ ∆ can be identified with the set of n−tuples in Cn with
non-zero discriminants, and this is equivalent to the set of n distinct complex roots, it then becomes the model
stated in example 2.1, hence its fundamental group is obviously Bn . ♣

2.2.2 Fundamental Group of the Complement of an Algebraic Curve


Let
n
X
f (x, y) = y n + ai (x)y n−1 ∈ C[y]n
i=1

be a monic degree n polynomial, let C be the algebraic curve

(x, y) ∈ C2 |f (x, y) = 0 ⊂ C2


We will introduce Zariski-van Kampen method to compute the fundamental group π1 (C2 \ C).
We first define the projection onto the first component

p : C2 \ C −→ C
(x, y) 7→ x

The fiber of the projection is C \ {n points} except for those x such that f (x, y) = fx (y) has multiple roots.
These points are precisely the branch points of the projection or the singularities of the curve C. Let S =
{x1 , ..., xs } ⊂ C be the set of all such points. Denoted by Lk = p−1 (xk ), and let
s
[
L= Lk
k=1

be the union of all such lines, they will be called the singular fibers, see figure 5.
Remark 2.3. Without loss of generality, we shall assume that all singular fibers Lk are in generic position.
That is every Lk intersects transversally with the curve C, and cannot contain other singularities or branch
points. If some Lk is not generic, we can change of the variables on C so that the projection onto the first
component in the new coordinates has the generic singular fibers.

5
Figure 5: projection onto the first component

Proposition 2.1 ( [11]). The restriction of the projection p:


p : C2 \ (C ∪ L) −→ C \ S
is a fibration with fiber F = C \ {n points}. The structure group of this fiber bundle is precisely the braid group
Bn .
Choose a base point x0 ∈ C \ S and y0 ∈ F , the fundamental groups of the base manifold and the fiber are
simply:
π1 (C \ S, x0 ) ∼
= Fs , π1 (F, y0 ) ∼
= Fn
The fundamental group of the base manifold has an action on the fundamental group of the fiber, this
action is called braid monodromy, denoted by ρ:
ρ : Fs −→ Aut (Fn )
The construction of the braid monodromy ρ is as follows.
We take the algebraic curve f (x, y) = 0 as a map:
f : C −→ C[y]n ∼ = Cn
x 7→ fx (y) = f (x, y)
Let ∆ be the subset containing the polynomials with vanishing discriminants (see example 2.3), observe that
f −1 (∆) is precisely S, hence the restriction of f on C \ S induces a homomorphism of fundamental groups:

f∗ : π1 (C \ S, x0 ) π1 (Cn \ ∆)

Fs Bn

The braid monodromy is given by the composition of Artin representation α (see example 2.2) and f∗
ρ = α ◦ f∗ : Fs −→ Aut (Fn )
Hence now, we can apply homotopy exact sequence of fibration:
p∗
1 π1 (C \ S, x0 ) π1 (C2 \ C) π1 (F, y0 ) 1

Fs Fn

If we denoted by γ1 , ..., γs , g1 , ..., gs the generators of the fundamental groups π1 (C \ S, x0 ) and π1 (F, y0 )
respectively, see figure 6, then we can state Zariski-van Kampen method as follows:
Theorem 2.2 (Zariski-van Kampen, [19]). The fundamental group of total space is the semi-product along braid
monodromy:
π1 C2 \ (C ∪ L) ∼

= Fs ⋊ ρ Fn
In particular, it has the presentation
π1 C2 \ (C ∪ L) = γ1 , ..., γs , g1 , ..., gn |γk−1 gi γk = ρ(γk )(gi )


Moreover, the fundamental group of C2 \ C is the quotient by the normal closure generated by γ1 , ..., γs , and it
has the presentation
π1 (C2 \ C) = ⟨g1 , ..., gn |gi = ρ(γk )(gi )⟩

6
Figure 6: generators


Example 2.4 (The Riemann surface of z). As a simple example, let’s consider the fundamental group of the
complement of the curve X√z : y 2 = x, which is the Riemann surface of the square root function.
Observe that the curve X√z has only one branched point (0, 0) in projecting onto the first component,

hence π1 (C \ S) ∼= Z, and braid monodromy will be given√
in B 2 . Now choose γ(t) = e 2π −1
t be the generator in
π −1t
π1 (C \ S), observe that the change of y is simply ±e , hence it yields the braid as σ1 , by (1), the generators
g1 , g2 on the fiber should satisfy the relation g1 = g2 , hence

π1 C2 \ X√z ∼

=Z


More generally, we have
Theorem 2.3 ( [19]). For any irreducible smooth affine algebraic curve C in C2 , the fundamental group

π1 C 2 \ C ∼

=Z

proof. We assume the curve C has degree d. Since it is irreducible and smooth, the projection onto the first
component will only have branch points with branching order d. Therefore, near each branch point, the curve
has the local expression y d = x, that gives a braid as σ1 σ2 · · ·σd−1 ∈ Bd . Therefore, by Zariski-van Kampen
method, we can compute the generators g1 , ..., gd of π1 C2 \ C by

gi = σ1 · · · σd−1 (gi ) ⇒ g1 = g2 = · · · = gd

hence the fundamental group is Z, as was to be shown. ♣


More interesting examples can be found in [19].

3 Complexified Spherical Pendulum


3.1 Complexification
The configuration space of spherical pendulum is the unit sphere
( 3
)
X
2 3 2
S = (x1 , x2 , x3 ) ∈ R xi = 1
i=1

The configuration space is the cotangent bundle T ∗ C2 . If we denote ⟨·, ·⟩ the Euclidean inner product on R3 ,
the cotangent bundle can be write as

T ∗ S2 = {(x, v) ∈ S2 × R3 |⟨x, v⟩ = 0}

The Hamiltonian is
1
H(x, v) =∥v∥2 + x3 : T ∗ S2 −→ R
2
Lie group SO(2) has a Hamiltonian action on T ∗ S2

SO(2) × T ∗ S2 −→ T ∗ S2
(A, (x, v)) 7→ (Ax, Av)

7
with the moment map
J(x, v) = x1 v2 − x2 v1 : T ∗ S2 −→ R
This moment J is also known as the angular moment. Together with the Hamiltonian H, µ = (H, J)
defines the first integrals of the spherical pendulum, we call µ the energy-momentum map.
Since everything here is polynomial, we can complexify every directly to a complex integrable system.
Let ⟨·, ·⟩ be the standard symmetric bilinear form on C3 , the configuration space is now a complexified
sphere:
CS2 = x ∈ C3 |⟨x, x⟩ = 1


and the phase space is simply

T ∗ CS2 = {(x, v) ∈ S2 × C3 |⟨x, v⟩ = 0} ⊂ C3 × C3

It is now a holomorphic symplectic manifold, in particular, it is an affine variety in C3 × C3 .


The complex Lie group
   
a −b
SO(2, C) = A = ∈ GL2 (C) a2 + b2 = 1 ∼ = C∗
b a

has a natural Hamiltonian action on T ∗ CS2 given by


    
A 0 A 0
A · (x, v) := x, v
0 1 0 1

the moment map, called the complexified angular moment, is given by

J(x, v) = x1 v2 − x2 v1 : T ∗ CS2 −→ C

The complexified energy-momentum map will still be denoted by µ = (H, J). Hence now, the spherical pendulum
becomes a complex integrable system in sense of definition 2.1.

3.2 Singular Set and Monodromy


Following [9], we define new coordinates (w1 , w2 , w, z1 , z2 , z) := (w, z) on C3 × C3 :
√ √
w1 = x1 + −1x2 , w2 = x1 − −1x2
√ √
z1 = v1 + −1v2 , z2 = v1 − −1v2

hence the defining equation for T ∗ CS2 is


(
∗ 2 w1 w2 + w 2 = 0
T CS :
w1 z2 + w2 z1 + 2wz = 0

The energy-momentum map can be re-write under the new coordinates as:

z1 z2 + z 2
 
w2 z1 − w1 z2
µ(w, z) = (x, y) := + w, √
2 −1
Proposition 3.1 ( [9]). Under the notation above, we have
(1). For each c = (x, y) ∈ C2 , the fiber µ−1 (c) of the energy-momentum map is a C∗ −bundle over the
punctured elliptic curve

Rc := (w, z) ∈ C2 |z 2 = 2 w2 − 1 (w − x) − y 2 \ {(±1, 0)}


 
(2)

(2). The singular set such that the energy-moment map µ has non-generic fibers is the algebraic curve
 
2 27 4 2 2 2 2
S = (x, y) ∈ C y + 2xy (x − 9) − 4(x − 1) = 0 (3)
4

proof. (1). Recall that the map



 
a −b
7→ a + b −1 := λ
b a
gives an isomorphism between SO(2, C) and C∗ , and the Hamiltonian SO(2, C) action can be reduced to the C∗
action by scalar product.

8
From [9, Proposition 2.1], we know the image of the quotient map
π : T ∗ CS2 −→ T ∗ CS2 /SO(2, C)
is the affine variety
(w, z, x, y) ∈ C4 |z 2 = 2 w2 − 1 (w − x) − y 2 \ {(±1, 0, ±1, 0)}
 

By Noether’s theorem, energy-momentum map µ is constant on the orbits, hence it can be descent to the
quotient space T ∗ CS2 /SO(2, C), denoted by µ̃. Therefore, µ−1 (c) is an C∗ −bundle over µ̃−1 (c). Notice that,
for each fixed c ∈ C2 , the fiber µ̃−1 (c) is exactly the Rc defined in (2), that proves the claim.
(2). Let R be the elliptic curve defined in (2), we notice that R is nonsingular, if and only if the discriminant
of
fc (w) = 2 w2 − 1 (w − x) − y 2


vanishes. Its vanishing locus is precisely formulated by (3). Let M ′ be the primage of C2 \ S under µ, to obtain
the desired conclusion, we also need to show
µ : M ′ −→ C2 \ S
has maximal rank everywhere, it follows from [9, Proposition 2.2]. ♣
Theorem 3.1. The fundamental group of the non-singular set C2 \S is Z, and the monodromy group of complex
spherical pendulum is Z
proof . We notice that the defining equation (3) for S is smooth and irreducible, hence by theorem 2.3 we
have
π1 C 2 \ S ∼

=Z
Denoted by Γ a generator of the fundamental group.
To compute the monodromy group, we define the period lattice in the following way.
Fix a c = (x, y) ∈ 2 \ S, first give Mc = µ−1 (c) a partial compactification M c by compactifying the elliptic
curve R defined in (2), let Rc be the compactification of Rc . Then we will need following results:
Lemma 3.1 ( [9]). (1). The forms

dw √ wz − −1y dw1
ω1 = , ω2 = −1 dw −
z z(w2 − 1) w1
can be analytically continued to M c .
(2). The following set  Z Z  

Λc := ω1 , ω2 γ ∈ H1 M c; Z (4)
γ γ

is a lattice in C2 , its Z−rank is 3, the generating vectors are


(λ1 , µ1 ), (λ2 , µ2 ), (0, 2π)
where √
√ wz − −1y
Z Z
dw
λi = , µi = −1 2
dw
γi z γi z(w − 1)
∼ Z ⊕ Z.
here γi ’s are the generators of H1 (Rc ; C) =
(3). The Hamiltonian monodromy matrix of the lattice Λc along Γ is
 
1 −1 0
Mon(Γ) = 0 1 0
0 0 1
Then our conclusion follows directly from lemma 3.1. For a detailed proof of lemma 3.1 can be found
in [9, §4 and §6]. ♣
For more interesting computations of Hamiltonian monodromy of spherical pendulum, we refer to [22].
Remark 3.1. The lattice Λc defined in (4) comes from the generalised Abel-Jacobi map AJ [23], which is
defined as
AJ : M c −→ C2
Z p Z p 
p 7→ ω1 , ω2
p0 p0

where p0 is a fixed point in M c . It was shown in [9, Theorem 5.1] that the Abel-Jacobi map AJ has a globally
defined inverse. ♣

9
4 Complexified Planar Kepler Problem
Let’s first recall the data in the usual planar Kepler problem, more detailed contents can be found in [4, 5].
The phase space is the symplectic manifold T ∗ (R2 \ {0}) ∼ = (R2 \ {0}) × R2 endowed with the standard
symplectic form:
ω = dp1 ∧ dq1 + dp2 ∧ dq2
where (q1 , q2 , p1 , p2 ) := (q, p) is the coordinate in T ∗ (R2 \ {0}).
The Hamiltonian H is defined by
1 2 1
p + p22 − p 2

H(q, p) = (5)
2 1 q1 + q22

Same as the spherical pendulum, planar Kepler problem endowed with a natural Hamiltonian action of
SO(2), the moment map is the angular moment

J : T ∗ R2 \ {0} −→ R


J(q, p) = q1 p2 − q2 p1

hence it has two Poisson commuting linearly independent first integrals, namely the Hamiltonian H and the
angular moment J, it is a Liouville integrable system.

4.1 Complexification
We cannot simply replace the configuration space R2 \ {0} by C2 \ {0}, since the square root term in (5)
will cause ambiguities. We should define the complexified configuration space to be the algebraic surface

Q = (w, z1 , z2 ) ∈ C3 w2 = z12 + z22 \ {0}




Proposition 4.1. The configuration space Q is a 2 dimensional holomorphic complex symplectic manifold. Its
cotangent bundle is trivial, that is
T ∗ Q = Q × C2
proof. Note that the algebraic surface w2 = z12 + z22 has the only singularity at the origin, and our Q just
removes it, that makes Q a holomorphic complex 2 dimensional manifold.
There are 2 holomorphic charts (Q1 , φ1 ) and (Q2 , φ2 ) on Q, namely

Q1 = {(w, z1 , z2 ) ∈ Q|w ∈ C \ R>0 }



=
φ1 (w, z1 , z2 ) = (z1 , z2 ) : Q1 −→ C2
(6)
Q2 = {(w, z1 , z2 ) ∈ Q|w ∈ C \ R<0 }

=
φ2 (w, z1 , z2 ) = (z1 , z2 ) : Q2 −→ C2

−1θ
if we denoted by z12 +z22 = re , where θ ∈ [−π, π), then the inverses on each coordinate chart are respectively
given by √ √ 
−1
φ−1
1 (z1 , z2 ) = re 2 (θ+2π)
, z1 , z2
√ √−1 
φ−1
2 (z1 , z2 ) = re 2 θ , z1 , z2

Clearly, the transition φ2 ◦ φ−1 1 : C2 −→ C2 on the overlap is identity, thus Q has trivial tangent and
cotangent bundle. Consequently, our complexified phase space is simply T ∗ Q = Q × C2 , the local coordinate
will be denoted by (z1 , z2 , w1 , w2 ) := (z, w). Under our notation, the standard symplectic form on T ∗ Q can be
write as
ω = dz1 ∧ dw1 + dz2 ∧ dw2
In particular, ω is a holomorphic (2, 0)−form on T ∗ Q, hence a holomorphic symplectic manifold. ♣
Let (z1 , z2 ) be the coordinate on Qi , now the complexified Hamiltonian

1 2 (−1)i
w1 + w22 − p 2

H(z, w) = , i = 1, 2
2 z1 + z22

becomes a holomorphic single-valued function in O(T ∗ Q).


Same as before, the complex Lie group SO(2, C) has a natural Hamiltonian action on T ∗ Q by

10
      
a −b a −b a −b
· (z1 , z2 , w1 , w2 ) := (z1 , z2 ) , (w1 , w2 )
b a b a b a
and the moment map associated to this action is the complexified angular moment:

J : T ∗ Q −→ C
J(z, w) = z1 w2 − z2 w1

Similar to the real case, the system (T ∗ Q, ω, H, J) becomes a complex integrable system in sense of definition
2.1. The energy-momentum map is
µ : T ∗ Q −→ C2
(z, w) 7→ (H(z, w), J(z, w))

4.2 Singular Set and Monodromy


We shall first determine all non-generic values of the energy-momentum map µ : T ∗ Q −→ C2 .
On each coordinate chart (Qi , φi ) of Q (defined in (6)), we define the following variables:
q √
w := (−1)i z12 + z22 , z := z1 w1 + z2 w2 + −1(z2 w1 − z1 w2 )
(7)
1 2  1
x := z1 w2 − z2 w1 , y := w + w22 −
2 1 w
Note that the new variables x, y are just angular moment and total energy respectively.
The fiber of the energy-momentum map µ has the following description:
Proposition 4.2 ( [10]). Under the notation above
(1). For each value c = (x, y) ∈ C2 , the fiber Mc = µ−1 (c) is a C∗ -bundle over a punctured algebraic curve
Rc which is defined by
√ √
(w, z) ∈ C2 2yw2 + 2w = z 2 + 2 −1xz \ (0, 0), 0, −2 −1x
  
(8)

(2). The collection of c ∈ C2 such that the fiber Mc is not generic forms an algebraic curve

C = (x, y) ∈ C2 |y(1 + 2xy 2 ) = 0




proof . The proof of (1) is similar to the proof in proposition 3.1. One can also find it in [10, Proposition
3.1].
To show the second part, let c = (x, y) be fixed, let R be the curve defined by the polynomial

p(z, w) = 2yw2 + 2w − z 2 − 2 −1xz ∈ C[z, w] (9)
Discuss into cases:
(1). If y = 0, then R is
√ √ 2
2w − z 2 − 2 −1xz = 2w − x2 − z + −1x = 0

the curve is isomorphic to C.


(2). If y ̸= 0, but ∆ = 2(1 + 2x2 y) = 0, then R becomes to
2


1 2
2y w + − z + −1x = 0
2y
the curve is isomorphic to a singular cone, and this case correspond to the singular fiber.
(3). If y ̸= 0 and ∆ = 2(1 + 2x2 y) ̸= 0, then R becomes to
√ 2
2y(w − r)(w − s) − z + −1x =0

for some r ̸= s, then the curve R is isomorphic to C∗ . This is the case such that the fiber Mc is a generic fiber.
To sum up, the singular set of µ is indeed the curve y(1 + 2x2 y) = 0. ♣
Theorem 4.1. The fundamental group of non-singular set C2 \ C is

π 1 C2 \ C ∼

=Z⊕Z

and the monodromy group of the complexified Kepler problem (T ∗ Q, ω, H, J) is Z2 ⊕ Z2 .

11
proof . Let C1 be the curve 1 + 2x2 y = 0 and ℓ the asymptotic line y = 0.
Consider C2 \ (C1 ∪ ℓ) in the projective plane P2 by

(x, y) 7→ [x : y : 1] ∈ P2

hence z = 0 becomes line at infinity of C2 , denoted by ℓ∞ . We shall use C¯1 and ℓ̄ represent for the projectivization
of C1 and ℓ in P2 respectively. In particular, C¯1 and ℓ̄ intersect at [1 : 0 : 0] ∈ ℓ∞ .
Recall that the projective plane P2 is just C2 together with its line at infinity, hence we have

π1 C2 \ (C1 ∪ ℓ) = π1 P2 \ C¯1 ∪ ℓ̄ ∪ ℓ∞
 

Now, we change the coordinates in P2 by


 
1 0 0
A = 0 0 1 ∈ PGL3 (C)
0 1 0

Observe that the line ℓ̄ becomes to the line at infinity ℓ∞ after changing coordinates by A. Let C¯1′ , ℓ′∞ be the
new curve after changing the coordinates, they are

C¯1′ : y 3 + 2x2 z = 0, ℓ′∞ : y = 0

Since projective transformation doesn’t impact on the topology, we have

π1 P2 \ C¯1 ∪ ℓ̄ ∪ ℓ∞ ∼ = π1 P2 \ C¯1′ ∪ ℓ′∞ ∪ ℓ∞


 

= π1 P2 \ ℓ∞ \ C¯1′ ∪ ℓ′∞
 

Notice that P2 \ ℓ∞ is the affine part of P2 , we can use the coordinate chart:

ψ : P2 \ ℓ∞ −→ C2
x y
[x : y : z] 7→ ,
z z
hence P2 \ ℓ∞ \ C¯1′ ∪ ℓ′∞ can be write under ψ by
 

P2 \ ℓ∞ \ C¯1′ ∪ ℓ′∞ ∼= C2 \ y(2x2 + y 3 ) = 0


  

Next, we will use Zariski-van Kampen method to compute the fundamental group of the complement of
the curve y(2x2 + y 3 ) = 0.
The curve y(2x2 + y 3 ) = 0 has the only singularity at the origin (0, 0), where the image of real part is
illustrated in figure 7.

Figure 7: the curve y(2x2 + y 3 ) = 0

Hence the singular set S is just one point, and the fundamental group π1 (C \ S, x0 ) is just Z. For each
x ∈ C \ S, since the curve has degree 4, the fundamental group of the fiber is just

π1 (C \ {4 points}) ∼
= F4

Their generators will be√ denoted by g1 , g2 , g3 , and g4 .


Choose γ(t) = e2π −1t ∈ π1 (C \ S, x0 ), the generator (see figure 7), the change on the fiber will give the
braid as σ1 σ3 σ2 σ1 ∈ B4 , see figure 8

12
Figure 8: the braids

Now, We can compute by (1):


g1 = σ1 σ3 σ2 σ1 (g1 ) = g4
g2 = σ1 σ3 σ2 σ1 (g2 ) = g1 g2 g1−1 ⇒ g1 g2 = g2 g1
g3 = σ1 σ3 σ2 σ1 (g3 ) = g4
g4 = σ1 σ3 σ2 σ1 (g4 ) = g3 g4 g3−1
By theorem 2.2, we finally know that
π1 C2 \ y(1 + 2x2 y) = 0 ∼= π1 C2 \ y 4 + 2x2 y = 0 = Z ⊕ Z
   

The first part is proved.  


The generators in the π1 C2 \ y(1 + 2x2 y) = 0 can be found in the following way: The curve x = 1
intersects the curve y(1 + 2x2 y) = 0 at q1 = (1, −1/2) and q2 = (1, 0), let γ1 , γ2 be the meridians in the complex
plane {(1, y)|y ∈ C} based at (1, −1/4) and go around q1 , q2 by a small circle respectively. These γ1 , γ2 will be
the generators.
To compute the Hamiltonian monodromy group, we will use the following fact
Lemma 4.1 ( [10]). (1). The 1-forms ω1 , ω2 :
√  √
−1 x −1
ω1 = − √ dw − dξ
w w(z + −1x) ξ
(10)
w
ω2 = √ dw
z + −1x

are holomorphic 1-forms defined on an open set of Mc where z+ −1x ̸= 0, and they can be extended analytically
to the whole Mc .
(2). The following set  Z Z  
Λc := ω1 , ω2 Γ ∈ H1 (Mc ; Z)
Γ Γ
is the period lattice of Mc . The generating vectors are (2π, 0) and (u, v), where
Z √ 
−1
Z
x w
u= − √ dw, v = √ dw
γ w w(z + −1x) γ z+ −1x
here γ is the generator of H1 (R; Z) ∼
= Z (R is defined in (9)).
(3). The Hamiltonian monodromy matrices of the lattice Λc along γ1 , γ2 ∈ π1 (C2 \ C) are the conjugations
 
−1 0
Mon(γi ) = Pi Pi−1 , i = 1, 2
−2 1
Then the monodromy group follows directly from lemma 4.1, and we refer to [10, §5] for a detailed proof.

Acknowledgement
The author would like to thank to Mr. Yunpeng Meng and Mr. Hongjie Zhou for the meaningful discussions,
thank to Prof. Shanzhong Sun for his generous guidance, and thank to Prof. José I. Cogolludo-Augustı́n for
his kind and patient explanation in using Zariski-van Kampen method.

13
Reference
[1] Johannes J Duistermaat. On global action-angle coordinates. Communications on pure and applied mathematics,
33(6):687–706, 1980.
[2] Otto Forster. Lectures on Riemann surfaces, volume 81. Springer Science & Business Media, 2012.
[3] Richard H Cushman and Larry M Bates. Global aspects of classical integrable systems, volume 94. Springer, 1997.
[4] Alain Albouy. Lectures on the two-body problem, 2002.
[5] Bruno Cordani. The Kepler Problem: Group Theotretical Aspects, Regularization and Quantization, with Application
to the Study of Perturbations, volume 29. Springer Science & Business Media, 2003.
[6] MS Child. Quantum monodromy and molecular spectroscopy. Contemporary Physics, 55(3):212–221, 2014.
[7] RH Cushman, HR Dullin, A Giacobbe, DD Holm, M Joyeux, P Lynch, DA Sadovskiı́, and BI Zhilinskiı́. CO2
molecule as a quantum realization of the 1:1:2 resonant swing-spring with monodromy. Physical review letters,
93(2):024302, 2004.
[8] Alain Albouy and Vadim Kaloshin. Finiteness of central configurations of five bodies in the plane. Annals of
mathematics, pages 535–588, 2012.
[9] Frits Beukers and Richard Cushman. The complex geometry of the spherical pendulum. Contemporary mathematics,
292:47–70, 2002.
[10] Shanzhong Sun and Peng You. Monodromy of complexified planar kepler problem. arXiv preprint arXiv:2209.00146,
2022.
[11] Daniel C Cohen and Alexander I Suciu. The braid monodromy of plane algebraic curves and hyperplane arrange-
ments. Commentarii Mathematici Helvetici, 72:285–315, 1997.
[12] Enrique Artal Bartolo, Jorge Carmona Ruber, and José Ignacio Cogolludo-Agustı́n. Braid monodromy and topology
of plane curves. Duke Mathematical Journal, 118(2):261–278, 2003.
[13] Phillip Griffiths and Joseph Harris. Principles of algebraic geometry. John Wiley & Sons, 2014.
[14] David Mumford. Tata Lectures on Theta II: Jacobian theta functions and differential equations. Springer, 2007.
[15] P Vanhaecke. Integrable Systems in the Realm of Algebraic Geometry. Springer, 2001.
[16] Arnaud Beauville. Holomorphic symplectic geometry: a problem list. In Complex and Differential Geometry:
Conference held at Leibniz Universität Hannover, September 14–18, 2009, pages 49–63. Springer, 2011.
[17] Misha Verbitsky. Hyperkaehler and holomorphic symplectic geometry. arXiv preprint alg-geom/9307009, 1993.
[18] Nigel J Hitchin. The self-duality equations on a riemann surface. Proceedings of the London Mathematical Society,
3(1):59–126, 1987.
[19] José Ignacio Cogolludo-Agustı́n. Braid monodromy of algebraic curves. In Annales mathématiques Blaise Pascal,
volume 18, pages 141–209, 2011.
[20] V Hansen. Braids and Coverings: Selected Topics. Cambridge University Press, 1989.
[21] Joan S Birman. Mapping class groups and their relationship to braid groups. Communications on Pure and Applied
Mathematics, 22(2):213–238, 1969.
[22] Michele Audin. Hamiltonian monodromy via picard-lefschetz theory. Communications in mathematical physics,
229:459–489, 2002.
[23] Igor Reider. Toward abel-jacobi theory for higher dimensional varieties. In Algebraic Geometry: Proceedings of the
International Conference held in L’Aquila, Italy, May 30–June 4, 1988, pages 287–300. Springer, 2006.

14

You might also like