Cohomologies of Affine Jacobi Varieties and Integrable Systems
Cohomologies of Affine Jacobi Varieties and Integrable Systems
Cohomologies of Affine Jacobi Varieties and Integrable Systems
a b 0
A. Nakayashiki and F.A. Smirnov
a
Graduate School of Mathematics, Kyushu University
Ropponmatsu 4-2-1, Fukuoka 810-8560, Japan
b
Laboratoire de Physique Théorique et Hautes Energies 1
Université Pierre et Marie Curie, Tour 16 1er étage, 4 place Jussieu
75252 Paris cedex 05-France
Abstract. We study the affine ring of the affine Jacobi variety of a hyper-elliptic
curve. The matrix construction of the affine hyper-elliptic Jacobi varieties due
to Mumford is used to calculate the character of the affine ring. By decomposing
the character we make several conjectures on the cohomology groups of the affine
hyper-elliptic Jacobi varieties. In the integrable system described by the family
of these affine hyper-elliptic Jacobi varieties, the affine ring is closely related to
the algebra of functions on the phase space, classical observables. We show that
the affine ring is generated by the highest cohomology group over the action of
the invariant vector fields on the Jacobi variety.
0 Membre du CNRS
1 Laboratoire associé au CNRS.
1
1 Introduction.
Our initial motivation is the study of integrable systems. Consider an integrable
system with 2n degrees of freedom, by definition it possesses n integrals in
involution. The levels of these integrals are n-dimensional tori. This is a general
description, but the particular examples of integrable models that we meet in
practice are much more special. Let us explain how they are organized.
The phase space M is embedded algebraically into the space RN . The
integrals are algebraic functions of coordinates in this space. This situation
allows complexification, the complexified phase space MC is an algebraic affine
variety embedded into CN . The levels of integrals in the complexified case allow
the following beautiful description. The systems that we consider are such that
with every one of them one can identify an algebraic curve X of genus n whose
moduli are defined by the integrals of motion. On the Jacobian J(X) of this
curve there is a particular divisor D (in this paper we consider the case when
this divisor coincides with the theta divisor, but more complicated situations
are possible). The level of integrals is isomorphic to the affine variety J(X) − D.
The real space RN ⊂ CN intersects with every level of integrals by a compact
real sub-torus of J(X) − D.
This structure explains why the methods of algebraic geometry are so impor-
tant in application to integrable models. Closest to the present paper account
of these methods is given in the Mumfords’s book [1].
Let us describe briefly the results of the present paper. We study the struc-
ture of the ring A of algebraic functions (observables) on the phase space of
certain integrable model. The curve X in our case is hyper-elliptic. As is clear
from the description given above this ring of algebraic functions is, roughly, a
product of the functions of integrals of motion by the affine ring of hyper-elliptic
Jacobian. The commuting vector fields defined by taking Poisson brackets with
the integrals of motion are acting on A. We shall show that by the action of
these vector-fields the ring A is generated from finite number of functions corre-
sponding to the highest nontrivial cohomology group of the affine Jacobian. We
conjecture the form of the cohomology groups in every degree and demonstrate
the consistence of our conjectures with the structure of the ring A.
Finally we would like to say that this relation to cohomology groups became
clear analyzing the results of papers [2] and [3] which deal with quantum in-
tegrable models. Very briefly the reason for that is as follows. The quantum
observables are in one-to-one correspondence with the classical ones. Consider a
matrix element of some observable between two eigen-functions of Hamiltonians.
An eigen-functions written in “coordinate” representation (for “coordinates” we
take the angles on the torus) must be considered as proportional to square-root
of the volume form on the torus. The matrix element is written as integral with
respect to “coordinates”, the product of two eigen-functions gives a volume form
on the torus, and the operator itself can be considered, at least semi-classically,
as a multiplier in front of this volume form, i.e. as coefficient of some differen-
tial top form on the torus which is the same as the form of one-half of maximal
dimension on the phase space. Further, those operators which correspond to
2
“exact form” have vanishing matrix elements. This is how the relation to the
cohomologies appears.
The paper, after the introduction, consists of five sections and six appendices
which contain technical details and some proofs.
In section 2 we recall the standard construction of the Jacobi variety which
is valid for any Riemann surface.
An algebraic construction of the affine Jacobi variety J(X) − Θ of a hyper-
elliptic curve X is reviewed in section 3 following the book [1]vol.II. This con-
struction is specific to hyper-elliptic curves or more generally spectral curves
[5].
In section 4 we study the affine ring of J(X) − Θ using the description in
section 3. The main ingredient here is the character of the affine ring. To be
precise we consider the ring A0 corresponding to the most degenerate curve
y 2 = z 2g+1 . The ring A and the affine ring Af of J(X) − Θ for a non-singular
X can be studied using A0 . It is important that A0 is a graded ring and the
character ch(A0 ) is defined. We calculate it by determining explicitly a C-
basis of A0 . The relation between A0 and Af for a non-singular X is given in
Appendix E.
A set of commuting vector fields acting on A is introduced in section 5. This
action descends to the quotients A0 and Af . The action of the vector fields
coincides with the action of invariant vector fields on J(X). With the help
of these vector fields we define the de Rham type complexes (C∗ , d), (C∗0 , d),
(C∗f , d) with the coefficients in A, A0 , Af respectively. The complex (C∗f , d)
is nothing but the algebraic de Rham complex of J(X) − Θ whose cohomology
groups are known to be isomorphic to the singular cohomology groups of J(X)−
Θ. What is interested for us is the cohomology groups of (C∗0 , d). We calculate
the q-Euler characteristic of (C∗0 , d) and show that it coincides with the quotient
of ch(A0 ) by the character ch(D) of the space of commuting vector fields. Then,
by the Euler-Poincaré principle, ch(A0 )/ch(D) is found to be expressible as the
alternating sum of the characters of cohomology groups of (C∗0 , d). Decomposing
independently the explicit formula of ch(A0 ) into the alternating sum, we make
conjectures on the cohomology groups of (C∗0 , d) which are formulated in the
next section.
In section six and in Appendix B we study the singular homology and co-
homology groups of J(X) − Θ. The Riemann bilinear relation plays an impor-
tant role here. We formulate conjectures on the cohomology groups of (C∗ , d),
(C∗0 , d), (C∗f , d).
Acknowledgements. This work was begun during the visit of one of the
authors (A.N.) to LPTHE of Université Paris VI and VII in 1998-1999. We
express our sincere gratitude to this institution for generous hospitality. A.N
thanks K. Cho for helpful discussions.
3
2 Hyper-elliptic curves and their Jacobians.
Consider the hyper-elliptic curve X of genus g described by the equation:
y 2 = f (z),
where
4
where ζ ∈ Cg . The theta function satisfies
for m, n ∈ Zg .
Consider the symmetric product of X, the quotient of the product space by
the action of the symmetric group:
X(n) = X n /Sn .
explicitly given by
g Z pk
X
wj = ωj + ∆,
k=1 ∞
The main subject of our study is the ring A of meromorphic functions on J(X)
with singularities only on Θ. The simple way to describe this ring is provided
by theta functions:
∞
[ Θk (w)
A= , (4)
θ(w)k
k=0
where Θk is the space of theta functions of order k i.e. the space of regular
functions on Cg satisfying
5
Now consider the functions
Θ3 (w)
.
θ(w)3
Obviously, with the help of these functions, we can embed the non-compact
g
variety J(X) − Θ into the complex affine space C3 −1 . Denote the coordinates
in this space by x1 , · · · , x3g −1 , the affine ring of J(X) − Θ is defined as the ring
where (gα ) is the ideal generated by the polynomials {gα } such that {gα = 0}
defines the embedding. It is known that the affine ring is the characteristic of
the non-compact variety J(X)−Θ independent of a particular embedding of this
variety into affine space. Obviously the ring A defined above is isomorphic to
the affine ring. We remark that the above argument on the embedding J(X)−Θ
into an affine space is valid if (J(X), Θ) is replaced by any principally polarized
abelian variety.
Consider X(g) which is mapped to J(X) by the Abel map a. The Riemann
theorem says that
pj
g−1
XZ
θ(w) = 0 iff w= ω+∆
j=1 ∞
which allows to describe Θ in terms of the symmetric product. One easily argues
that the preimage of Θ under the Abel map is described as
D := D∞ ∪ D0 ,
where
The Abel map is not one-to-one, and the compact varieties J(X) and X(g)
are not isomorphic. However, the affine varieties J(X) − Θ and X(g) − D are
isomorphic since the Abel map
a
X(g) − D −→ J(X) − Θ
6
where the matrix elements are polynomials of the form:
a(z) = a 23 z g−1 + a 25 z g−2 + · · · + ag+ 12 , (6)
b(z) = z g + b1 z g−1 + · · · + bg ,
c(z) = z g+1 + c1 z g + c2 z g−1 + · · · + cg+1 .
Later we shall set b0 = c0 = 1. Consider the affine space C3g+1 with coordinates
a 32 , · · · , a g+1 , b1 , · · · , bg , c1 , · · · , cg+1 . Fix the determinant of m(z):
2
and set
yj = a(zj ).
Obviously zj , yj satisfy the equation
yj2 = f (zj ),
which defines the curve X. So, we have constructed a point of X(g) for every
m(z) which satisfies the equations (7). Conversely, for a point (p1 , · · · , pg ) of
X(g), construct the matrix m(z) as
Y g Xg Y z − zk
b(z) = (z − zj ), a(z) = yj ,
j=1 j=1
zj − zk
k6=j
2
−a(z) + f (z)
c(z) = ,
b(z)
where zj = z(pj ) is the z-coordinate of pj . Considering the function b(z) as a
function on X(g) one finds that it has singularities when one of zj equals ∞.
The function a(z) is singular at zj = ∞ and also at the points where zi = zj
but yi = −yj . This is exactly the description of the variety D. The functions
a(z) and c(z) do not add new singularities. Thus we have the embedding of the
affine variety X(g) − D into the affine space:
X(g) − D ֒→ C3g+1 .
Therefore we can profit from the wonderful property of the hyper-elliptic Jaco-
bian: it allows an affine embedding into a space of very small dimension equal
to 3g + 1 (compare with 3g − 1 which we have for any Abelian variety). Actu-
ally, the space C3g+1 occurs foliated with generic leaves isomorphic to the affine
Jacobians.
7
4 Properties of affine ring.
Consider the free polynomial ring A:
A = C [a 32 , · · · , ag+ 12 , b1 , · · · , bg , c1 , · · · , cg+1 ].
On the ring A one can naturally introduce a grading. Prescribe the degree j to
any of generators aj , bj , cj and extend this definition to all monomials in A by
deg(xy) = deg(x) + deg(y).
Every monomial of the ring has positive degree (except for 1 whose degree equals
0). Thus, as a linear space, A splits into
M
A= A(p) ,
2p∈Z+
8
Since F× A is a homogeneous ideal of A, A0 is a graded vector space:
M (p)
A0 = A0 .
2p∈Z+
where
ap , p = half-integer
up =
bp , p = integer,
is a basis of A0 as a vector space, where i1 , · · · , ig are non-negative integers and
l1 , · · · , lg are 0 or 1.
9
which means that Ker(m) = 0. We summarize this in the following:
{zi , yj } = δi,j zi .
Di h = {fg+i , h}, i = 1, · · · g.
10
One can think of these commuting vector-fields as Dj = ∂τ∂ j where τj are
”times” corresponding to the integrals of motion fg+j . The ”times” τj are
coordinates on the Jacobi variety, they are related to w as follows
τ = 21 M w,
where M is the matrix defined in (2). We remark that Di here coincides with
−2Di in the Mumford’s book [1]vol.II. Earlier we have introduced a gradation on
the ring A. We can prescribe the degrees to the vector-fields Dj as deg (Dj ) =
j − 12 because it can be shown that:
1
Dj A(p) ⊂ A(p+j− 2 ) .
with fi1 ,··· ,ik ∈ A. These forms span the linear spaces Ck for k = 0, · · · , g. The
differential
X g
d= dτj Dj ,
j=1
acts from Ck to Ck+1 . As usual applying d we first apply the vector fields Dj
to the coefficients of the differential form and then take exterior product with
dτj . We have the complex
d d d d d
0 −→ C0 −→ C1 −→ · · · −→ Cg−1 −→ Cg −→ 0.
11
The differential d respects the grading. In this case the q-Euler characteristic
can be introduced
which possesses all the essential properties of the usual Euler characteristic.
Using the formula (14) one finds
1 2 2g−1
χq (C∗0 ) = (−1)g q − 2 g 2 ! ch (A0 ) (16)
1 2 [2g + 1]! [ 12 ]
= (−1)g q − 2 g .
[g + 12 ] [g]! [g + 1]!
Consider the cohomology groups H k (C∗0 ). The vector spaces H k (C∗0 ) inherit
a grading from Cj0 . Then
(2g)!
2g
2g
lim χq (C∗0 ) = (−1)g = (−1)g
g − g−1 . (17)
q→1 (g)! (g + 1)!
Certainly the fact that the q-Euler characteristic has a finite limit does not mean
that cohomology groups are finite-dimensional, but we believe that this is the
case. So, we put forward
More explicitly the cohomology groups will be discussed in the next section.
In the situation under consideration there is an important connection be-
tween the algebra A and the highest cohomology group H g (C∗0 ).
12
which implies that X
x=h+ Di xi ,
with h such that Ω0 = h dτ1 ∧ · · · ∧ dτg , xi ∈ A0 . Apply the same procedure
to xi and go on along the same lines. The resulting representation (18) will be
achieved in finite number of steps for the reason of grading. QED.
Proof. We shall prove the proposition by the induction on the degree of x. Since
A(p) = {0} for p < 0, the beginning of induction obviously holds. Suppose that
13
the proposition is true for all elements of degree less than deg(x). By Proposition
2, there exist xj such that
2g+1
X
x = x0 + fi xi , xj ∈ A0 ,
i=1
where deg(x) = deg(x0 ) and deg(xi ) = deg(x) − i < deg(x) for i > 0. By
Proposition 3, there exist polynomials Pα (D1 , · · · , Dg ) with the coefficients in
C such that
X 2g+1
X
x0 = Pα (D1 , · · · , Dg )hα + fi yi , yi ∈ A,
i=1
where deg yi = deg x0 − i. Since deg(yi ) < deg(x0 ), x can be written in the form
(20) by the induction hypothesis. QED.
Proposition 4 represents the most important result of this paper. The possi-
bility of presenting every algebraic function on the phase space of the integrable
model in the form (20) starting from finite number of functions {hα }, which are
representatives of the highest cohomology group, is important both in classical
and in quantum case. The description of null-vectors follows from the one given
above because Di commute with fi .
0 2g+1
and f (z) = z In the case when all fi0 = 0, Af 0 = A0 .
+ f10 z 2g + · · ·+ f2g+1
0
.
2 0
If the curve X: y = f (z) is non-singular, Af 0 is isomorphic to the affine ring of
J(X) − Θ. Since d commutes with F, the complex (C∗ , d) induces the complex
(C∗f 0 , d), where
X
Ckf 0 = Ck ⊗F Cf 0 = Af 0 dτi1 ∧ · · · ∧ dτik .
i1 <···<ik
Recall that
g ∗ Cg g
Cgf 0
H (C ) = , H (C∗f 0 ) = .
dCg−1 dCg−1
f0
14
Thus, tensoring Cf 0 to the exact sequence
d
Cg−1 −→ Cg −→ H g (C∗ ) −→ 0,
we have
H g (C∗ ) ⊗F Cf 0 ≃ H g (C∗f 0 ),
for any f 0 . By Proposition 4 we have
X
H g (C∗ ) = FΩα ,
α
F ⊗C H g (C∗0 ) −→ H g (C∗ ).
We conjecture that this map is in fact injective. In general we put forward the
following conjecture.
H k (C∗ ) ≃ F ⊗C H k (C∗0 ).
H k (C∗0 ) ≃ H k (C∗f 0 ),
15
Since Xaff is affine and connected,
then they generate the cohomology ring H ∗ (Xaff (n), C). Obviously
where
g
1
P
ω= 4 (2k − 1) ν̃k ∧ ν̃−k+1 . (23)
k=1
16
We remark that ω does not depend on the choice of the canonical basis {νj }.
By Proposition 5 the map i′ induces an injective map:
ch(W k ) = Rk − Rk−2 ,
1 2g
Rk = q 2 k(k−2g) .
k
Now we put forward the strong conjecture that the map (24) is in fact
surjective and the character of W k defined here coincides with the character of
H k (C∗0 ):
What does it mean? The divisor D consists of D∞ and D0 . The forms from
W k describe the part of cohomology groups with singularities on D∞ only. Our
conjecture is that this part exhausts the whole space of the cohomology groups,
i.e. that adding exact forms one can move singularities of any form from D0
to D∞ . This is a strong statement which is rather difficult to prove. The first
non-trivial case is g = 3 for which we were able to prove Conjecture 3. The
details of it will be published elsewhere.
For k = 1 we can prove Conjecture 3 (1) for any g. The proof is given in
Appendix D.
Let us now present a simple calculation which shows that Conjectures 2, 3
are consistent with the calculation of q-Euler characteristic of C∗0 . Indeed
This calculation was actually the starting point for Conjectures 2, 3. Certainly it
does not prove anything, but it shows remarkable consistence between different
calculations performed in this paper.
In order to understand the magic of the hyper-elliptic case it is instructive
to compare it with the case of an Abelian variety in generic when the divisor
Θ is non-singular (which rarely the case for Jacobians of algebraic curves). As
17
explained in the Appendix C in the latter case the following can be proven:
W k ≃ H k (J − Θ, C), k ≤ g − 1,
W g ֒→ H g (J − Θ, C).
Actually for g ≥ 3 the space H g (J − Θ, , C) is bigger than W g , the difference of
dimensions being
(2g)!
dim H g (J − Θ, C) − dim W g = g! − .
g!(g + 1)!
We would conjecture that the equality H g (J − Θ, C) ≃ W g specifies hyper-
elliptic Jacobians.
B0 = C [u1 , u 23 , · · · , u g+1 ].
2
We use both this bracket notation and the uj notation to denote a monomial.
The ring B0 is considered as a coefficient in the sequel.
Let us first prove
18
1. If deg(P ) < deg(P ′ ), then P < P ′ .
2. If deg(P ) = deg(P ′ ), compare P and P ′ by the lexicographical order from
the left.
Notice that the product of two elements P = [−m1 , · · · , −mg ] and P ′ =
[−m′1 , · · · , −m′g ] is expressed as
From (25)
ck = −uk + · · · , 1 ≤ k ≤ g + 1,
u2k = · · · , g + 2 ≤ k ≤ 2g + 1. (27)
2
The next lemma describes what kind of monomials appear in the right hand
side of (27).
Lemma A2. The right hand side of (27) is a linear combination of elements
of the form ul , ul um , ul um un with the coefficients in B0 .
Proof. It is sufficient to show that the term like xui1 · · · uir , r ≥ 4, x ∈ B0 does
not appear in the expression. Since (25), (26) are homogeneous, if x 6= 0 and
homogeneous, then
deg(xui1 · · · uir ) = k,
Lemma A3. In each of the cases in Lemma A2 we have the following state-
ments, where x is a homogeneous element in B0 .
1. If u2k/2 = xul + · · · , then u2k/2 > ul .
19
Proof. 1. Since deg(u2k/2 ) = k ≥ g + 2 > g + 1/2 ≥ l = deg(ul ),the claim follows.
2. If k > l + m, there is nothing to be proved. Suppose that k = l + m. Then
l < k < m. Thus comparing by the lexicographical order we have u2k/2 > ul um .
The statement of 3 is similarly proved. QED.
Starting from any element P = [−m1 , · · · , −mg ] we shall show that P can
be reduced to the desired form. If some mj ≥ 2, then rewrite it using (27).
By Lemma A3 every term in the resulting expression is less than P . Repeating
this procedure we finally arrive at the linear combinations of [−n1 , · · · , −ng ],
n1 , · · · , ng = 0, 1 with the coefficients in B0 . Thus Proposition A is proved.
QED.
By Proposition A we have
g g 1
1 2 [2g + 2]! ch(A)
Y Y g+1+k
ch(A0 ) ≤ 1+j 1+q 2 = = . (28)
j=1 2 k=1
g + 12 ! [g]! [g + 1]! ch(F)
20
where the one form κ̃ is given by
1 yi −yj dzj
X
κ̃ = κ(ij) , κ(ij) = 4 zi −zj
dzi
yi + yj
i<j
The form under d in RHS of (29) belongs to Cfk−1 , that is, it has singularity on
the divisor D = D0 ∪ D∞ , but after the differential is applied the singularities
on D0 disappear. Indeed, one easily shows that
g
(j) (i) (i) (j)
dκ(ij) = 1
P
4 (2k − 1) νk ν−k+1 − νk ν−k+1 , (30)
k=1
dz
νj = qj (z) , j = −(g − 1), · · · , g − 1, g,
y
and qj is the following polynomial of degree g − j:
!
y1 z1−j √
qj (z) = resp1 =∞ d z1 .
z1 − z
21
This proves that i′ ω ∧ H k−2 (Xaff (g), C) = 0.
Consider the homology groups of Xaff (g). Taking dual to the relation (22)
we obtain a similar relation for the homology groups:
k
^
Hk (Xaff (g), C) ≃ H1 (Xaff (g), C).
δ̃ = δ (1) + · · · + δ (g) ,
⊗(i−1) ⊗(g−i) g
δ (i) = p0 ⊗ · · · ⊗ δ ⊗ · · · ⊗ p0 ∈ H0 ⊗ H 1 ⊗ H0 ֒→ H1 (Xaff , C),
The meaning of this map is simple: every cycle on X(g) − D is at the same time
a cycle on Xaff (g). There are two subtleties:
1. Nontrivial cycle on X(g) − D can be trivial on Xaff (g) i.e. the map (32) can
have kernel.
2. On Xaff (g) there are cycles that intersect with D0 which means that they are
not cycles on X(g) − D, so, the map (32) can have cokernel.
Let us study the image of the map (32). A k-cycle from Hk (Xaff (g), Z) is a
linear combination of elements of the form:
∆ = δ̃1 ∧ · · · ∧ δ̃k ,
∆′ = δ1 × · · · × δk × p0 × · · · × p0
g g
defines an element of Hk (Xaff , Z). Let π be the projection map Xaff −→
g
Xaff (g) and π∗ the induced map on the homology groups, π∗ : H k (X aff , Z) −→
Hk (Xaff (g), Z). Then ∆ = k! kg π∗ (∆′ ) in Hk (Xaff (g), Z).
Thus the cycle ∆ belongs to Im(i) if ∆′ does not intersect π −1 (D). Recall
that D = D0 ∪ D∞ . By construction ∆′ has no intersection with π −1 (D∞ ).
One easily realizes that ∆′ does not intersect with π −1 (D0 ) iff
δi ◦ σ(δj ) = 0 ∀ i, j.
δi ◦ σ(δj ) = −δi ◦ δj .
22
Hence we come to the following
such that
δi ◦ δj = 0 ∀ i, j.
Define
VZ = H1 (Xaff , Z) = ⊕2g
i=1 ZAi , VC = H1 (Xaff , C) = ⊕2g
i=1 CAi .
Ai ◦ Aj = ±δj,i±g .
By definition
Define
ϕk : ∧k VC −→ ∧k−2 VC ,
X
ϕk (γ1 ∧ · · · ∧ γk ) = (−1)i+j−1 (γi ◦ γj )γ{ij} , (33)
i<j
dk := dim Ker(ϕk ) = 2g 2g
k − k−2 .
23
As a consequence of the lemma one has in particular
and
are related by
for γ ∈ Hk (X(g) − D, C), η ∈ ∧k H 1 (Xaff (g), C). The pairing (36) is given by
the integral:
By (31) we have
Since (36) is non-degenerate and dim i(Wk ) = dim W k , the pairing (38) is
non-degenerate. It easily follows from this that the pairing (37) is also non-
degenerate. Thus we have proved Proposition 5. QED.
dim Wk = 2g 2g
k − k−2 .
24
Proposition C. The dimensions of cohomology groups of J − Θ are given by
k 2g 2g
dimH (J − Θ, C) = − , k ≤ g − 1,
k k−2
2g 2g (2g)!
= − + g! − , k = g,
g g−2 g!(g + 1)!
= 0, k > g.
Proof. Consider the inclusions Θ ⊂ J ⊂ (X, J) and the induced homology exact
sequence:
Taking the dual sequence of this and using the Poincare-Lefschetz duality we
get
Since J − Θ is affine
Hk (J − Θ, C) = 0, k > g.
Then we have
Using the representation theory of sl2 as in the proof of the hard Lefschetz
theorem (c.f.[6]), the map (40) is injective for k ≤ g. Thus by (39) the following
exact sequences hold:
[Θ]∧
0 → H k−2 (Θ, C) −→ H k (J, C) → H k (J − Θ, C) → 0, k < g,
[Θ]∧
0 → H g−2 (Θ, C) −→ H g (J, C) → H g (J − Θ, C) →
→ H g−1 (Θ, C) → H g+1 (J, C) → 0.
QED.
25
By Proposition F 3 the fundamental class of Θ coincides with ω in Proposi-
tion 5 in the hyper-elliptic case. Thus, if we define W k in a similar formula to
(24), we have
W k ≃ H k (J − Θ, C), k ≤ g − 1,
W g ֒→ H g (J − Θ, C).
Proposition D. For any principally polarized Abelian variety (J, Θ) such that
Θ is irreducible we have the isomorphism
H 1 (J, C) ≃ H 1 (J − Θ, C).
In particular
Ω0 (nΘ) = O(nΘ), Ω0 (∗Θ) = O(∗Θ).
We first recall the description of H 1 (J, C) in terms of the differentials of the
first and second kinds. Consider the sheaf exact sequence:
d
0 −→ C −→ O(∗Θ) −→ d O(∗Θ) −→ 0.
26
Since
H k (J, O(∗Θ)) = 0, k ≥ 1,
we have
The numerator in the right hand side of (41) is nothing but the space of differ-
ential one forms of the first and the second kinds on J and the denominator is
the space of globally exact meromorphic one forms.
On the other hand, by the algebraic de Rham theorem, the first cohomology
group of the affine variety J − Θ is described as
27
From the cohomology sequence of it we have the map
Kerπn = 0, n = 0, 1, 2, · · · .
we easily have
The second isomorphism follows from the fact that a meromorphic function on
J which has poles only on Θ of order at most one is a constant. Thus the
Proposition D is proved. QED.
11 Appendix E.
0
Recall that, for a set of complex numbers f 0 = (f10 , · · · , f2g+1 ), the ring Af 0 is
defined by
A
Af 0 = A ⊗F Cf 0 = P2g+1 ,
i=1 A(fi − fi0 )
28
Since deg(fi0 ) = 0, grAf 0 becomes a quotient of A0 . In other words there is a
surjective ring homomorphism
A0 −→ grAf 0 . (47)
A0 ≃ grAf 0 .
Proof. Notice that the map (47) respects the grading and it is surjective at each
grade. By Proposition 2, A ≃ F ⊗C A0 as a C-vector space. Thus we have
A0 ≃ Af 0 as a C-vector space for any f 0 . In particular the basis of A0 given in
(n)
Proposition 1 is also a basis of Af 0 . Denote by {xj } the basis of the degree
(n) (n)
n part A0 . Let us prove that {xj } are linearly independent in grAf 0 by
the induction on n. For n = 0 the statement is obvious. We assume that the
(m)
statement is true for all m satisfying m < n. This means that {xj } is a basis
(m)
of the degree m part of grAf 0 for all m < n. In particular {xj |m < n} is a
basis of Af 0 (n − 21 ). Suppose that the relation
(n)
X
αj xj ∈ Af 0 (n − 12 ).
j
(m)
holds, where some αj 6= 0. Then this means that {xj |m ≤ n} are linearly
(k)
dependent in Af 0 . This contradicts the fact that {xj } is a basis of Af 0 . Thus
(n)
{xj } are linearly independent in grAf 0 . QED.
29
and define
∧k V
Wk = . (48)
ω ∧k−2 V
Assign degrees to the basis elements by
by
g
X
d(P ⊗ vI ∧ ξJ ) = Di P ⊗ vi ∧ vI ∧ ξJ ,
i=1
it induces a map
d
D ⊗C W k −→ D ⊗C W k+1 .
It is easy to check that the map d satisfies d2 = 0.
is exact at D ⊗C W k except k = g.
30
The property 1 is the well known property of the Koszul complex. The
property 2 is also well known and easily proved using the representation theory
of sl2 .
Now suppose that x ∈ D ⊗C ∧k V , k < g and dx = ω ∧ y for some y ∈
D ⊗C ∧k−2 V . Then ω ∧ dy = 0 and thus dy = 0 by the property 2. Then y = dz
for some z by the property 1. Thus we have d(x − ω ∧ z) = 0 and x = dw + ω ∧ z
for some w again by the property 1. QED.
Next we shall define a map from D ⊗C W g to A0 dτ1 ∧ · · · ∧ dτg . To this end
we need to identify W k in this section and that of Section 6, which is defined as
the quotient of the cohomology groups of a Jacobi variety. For this purpose we
describe the cohomology groups of a Jacobi variety in terms of theta functions.
Define
∂
ζi (w) = Di log θ(w) = log θ(w).
∂τi
Then, for each i, the differential dζi (w) defines a meromorphic differential
form on J(X) which has double poles on Θ and which is locally exact. This
means that dζi is a second kind differential on J(X). By (41) first and second
kinds differential one forms define elements of H 1 (J(X), C). The pairing with
H1 (J(X), C) is given by integration. The following proposition can be easily
proved by calculating periods.
Proposition F 2.The first and the second kinds differentials dτ1 ,...,dτg ,
dζ1 ,...,dζg give a basis of the cohomology group H 1 (J(X), C).
vi = dτi , ξi = dζi .
The next proposition can be easily proved by calculating integrals over two
cycles in a similar way to [1](vol.I, p188).
∧g V → Af 0 dτ1 ∧ · · · ∧ dτg ,
vI ∧ ξJ 7→ dτI ∧ dζJ ,
31
where for I = (i1 , · · · , ir ), dτI = dτi1 ∧ · · · ∧ dτir etc. and f 0 is any set of
complex numbers such that y 2 = f 0 (z) defines a non-singular curve X.
This map extends to the map of D modules
D ⊗C ∧g V → Af 0 dτ1 ∧ · · · ∧ dτg ,
in
P the KfollowingKmanner. Let P ∈ D and consider P ⊗ (vI ∧ ξJ ). Write dζJ =
K FJ dτK , FJ ∈ Af 0 . Then we define
X
P ⊗ (vI ∧ ξJ ) −→ P (FJK )dτI ∧ dτK .
Lemma F. Suppose that Conjecture 2 and 3 are true. Then the kernel of ev is
given by
Ker(ev) = d D ⊗ W g−1 .
32
the claim of the lemma follows. QED.
We summarize the result as
Theorem F. Suppose that Conjecture 2, 3 are true. Then the following complex
gives a resolution of A0 dτ1 ∧ · · · ∧ dτg as a D-module:
ev d d
0←−A0 dτ1 ∧ · · · ∧ dτg ←− D ⊗C W g ←− · · · ←− D ⊗C W 0 ←− 0.
References
[1] D. Mumford, Tata Lectures on Theta, vol. I and II, Birkhäuser,
Boston (1983)
[2] O. Babelon, D. Bernard, F.A. Smirnov, Comm. Math. Phys. 186
(1997) 601
[3] F.A. Smirnov, J. Phys. A: Math. Gen. 31 (1998) 8953
[4] E.K. Sklyanin, Separation of Variables, Progress in Theoretical
Physics Supplement 118 (1995) 35
[5] A. Beauville, Acta Math. 164 (1990) 211
[6] P. Griffiths and J. Harris, Principles of Algebraic Geometry, A
Wiley-Interscience publication (1978)
[7] I.G. Macdonald, Topology 1 (1962) 319
[8] M. Atiyah and W.V.D. Hodge, Ann. of Math. 62 (1955) 56
[9] W. Fulton and J. Harris, Representation Theory, Springer, New
York (1991)
33