Abstract:
For the case of algebraic curves (compact Riemann surfaces), it is shown that
de Rham cohomology group $H^1_{\mathrm{dR}}(X,\mathbb{C})$ of a genus $g$
of the Riemann surface $X$ has a natural structure of a symplectic vector space.
Every choice of a non-special effective divisor $D$ of degree $g$ on $X$
defines a symplectic basis of $H^1_{\mathrm{dR}}(X,\mathbb{C})$ consisting
of holomorphic differentials and differentials of the second kind with poles
on $D$. This result, which is the algebraic de Rham theorem, is used to describe
the tangent space to Picard and Jacobian varieties of $X$
in terms of differentials of the second kind, and to define a natural
vector fields on the Jacobian of the curve $X$ that move points of the divisor $D$.
In terms of the Lax formalism on algebraic curves, these vector fields
correspond to the Dubrovin equations in the theory of integrable systems,
and the Baker–Akhierzer function is naturally obtained by the integration along
the integral curves.
Keywords:Riemann surface, divisor, line bundle, Riemann–Roch theorem,
differentials of the second kind, algebraic de Rham theorem, Picard
and Jacobian varieties, vector field on the Jacobian variety, Lax
representation, Dubrovin equation, Baker–Akhiezer function.
Let $X$ be a smooth algebraic variety over $\mathbb{C}$ with the classical topology of a complex manifold. According to Atiyah and Hodge [1], a closed meromorphic $p$-form $\varphi$ on $X$ is called differential of a second kind if it has zero residues on open subsets $U=X\setminus D$ for sufficiently large divisors $D$. A far-reaching generalization of Atiyah and Hodge results was given by Grothendieck [2]. The quotient groups
$$
\begin{equation*}
\frac{\{p\text{-forms of the second kind}\}}{\{\text{exact forms}\}}
\end{equation*}
\notag
$$
have a natural interpretation in terms of a spectral sequence of certain complex of sheaves of meromorphic forms on $X$ (see [3], Ch. 3, § 5). In particular,
$$
\begin{equation*}
H_{\mathrm{dR}}^1(X,\mathbb{C}) \simeq \frac{\{1\text{-forms of the second kind}\}}{\{\text{exact forms}\}}.
\end{equation*}
\notag
$$
When $X$ is a smooth algebraic curve of genus $g$, this isomorphism follows from the Riemann–Roch theorem. It turns out that in this case the space of differentials of the second kind carries a natural skew-symmetric bilinear form, which is non-degenerate when taking a quotient by the subspace of exact forms. Using this bilinear form, in Theorem 1 we give a more explicit formulation of the algebraic de Rham theorem. Specifically, we show that each non-special effective divisor $D$ of degree $g$ on the Riemann surface $X$ defines a symplectic basis of $H_{\mathrm{dR}}^1(X,\mathbb{C})$, which makes it possible to explicitly describe a complement to the Lagrangian subspace of holomorphic $1$-forms on $X$ as the subspace of differentials of the second kind with the poles in $D$.
In § 4, it will be shown that every non-special effective divisor $D$ of degree $g$ on $X$ defines an explicit isomorphism between the vector space $H^{0,1}(X,\mathbb{C})$ and the Lagrangian subspace of differentials of the second kind with poles in $D$. This allows us to explicitly describe the tangent space to the Picard variety (and its “incarnations”, Albanese and Jacobian varieties) in pure algebro-geometric terms.
It is quite remarkable that this formalism is connected with the theory of integrable systems. In the standard approach (see, for example, [4]), an integrable system is described by the zero curvature equation
where $L(x,t,\lambda)$ and $M(x,t,\lambda)$ are certain $r\times r$ matrix-valued rational functions of the spectral parameter $\lambda\in \mathbb{C}\mathbb{P}^1$ which also depend on extra arguments $x$ and $t$ (physical space and time variables). In [5], [6] by the first author, the zero curvature formalism was extended to the case when spectral parameter varies on an algebraic curve.
Namely, it was shown in [5] that a natural framework for this generalization is provided by the Hitchin system expressed in terms of Tyurin parameters for stable holomorphic vector bundles of rank $r$ and degree $rg$ on an algebraic curve. Correspondingly, the rational functions $L$ on $\mathbb{C}\mathbb{P}^1$ become special $r\times r$ meromorphic matrix $1$-forms $L(z)\,dz$ on an algebraic curve, and the set of such matrices is parameterized by the moduli space (more precisely, by its Zariski open subset) of stable holomorphic vector bundles of rank $r$ and degree $rg$ (see [5] for details). A similar explicit description is given for $r\times r$ meromorphic matrix-valued functions $M(z)$.
It turns out that, in the simplest case $r=1$, this formalism is still non-trivial and is naturally connected with Theorem 1 (see the related discussion in § 3). Namely, as shown in § 5, the meromorphic $1$-forms $L(z)\,dz$ become differentials of the first kind on an algebraic curve $X$, while analogs of $M(z)$ are meromorphic functions $f$ defined using two non-special effective divisors $D$ and $D_0$ of degree $g$ on $X$. The varying divisors $D$ parameterize the Jacobian of $X$ with the base point $D_0$, and the vector fields describing the motion of points of $D$ are naturally expressed in terms of the meromorphic functions $f$.
Remarkably, in the case when $X$ is a hyperelliptic curve, equations (5.3) for the integral curves of these vector fields coincide with the Dubrovin equations arising in the theory of finite-gap integration of the Korteweg–de Vries equation [7]. Moreover, integrating meromorphic functions $f$ along these integral curves and using the Dubrovin equations, we naturally obtain the Baker–Akhiezer function, a fundamental object in the algebro-geometric approach to integrable systems, which was introduced by the first author in [8].
The second author is grateful to the referee for constructive remarks and suggestions.
§ 2. Differentials of the second kind
Let $X$ be a connected compact Riemann surface of genus $g$ with classical topology. We denote by $\mathcal{O}_X$ the sheaf of germs of holomorphic functions on $X$. Let $\mathcal{M}_X$ be the sheaf of germs of meromorphic functions on $X$, and let $\mathcal{M}$ be the vector space of meromorphic functions on $X$. For every divisor $D$ on $X$, let $L=\mathcal{O}(D)$ be the holomorphic line bundle associated with $D$, and let $H^0(X,L)$ be the vector space of holomorphic sections of $L$ over $X$. The isomorphism
From the Riemann–Roch theorem together with the Kodaira–Serre duality it follows that
$$
\begin{equation*}
h^0(L)-h^0(K_X-L)=\deg L +1-g,
\end{equation*}
\notag
$$
where $h^0(L)=\dim_{\mathbb{C}} H^0(X,L)$, $\deg L$ is the degree of $L$, and $K_X$ is the canonical class of $X$ (the holomorphic cotangent bundle to $X$).
Let $\mathrm{d}$ be the exterior derivative on $X$. The sheaf $\mathrm{d} \mathcal{M}_X$ is a sheaf of germs of differentials of the second kind on $X$ and $\Omega^{(\mathrm{2nd})}= H^0(X,\mathrm{d} \mathcal{M}_X)$ is the infinite-dimensional vector space of the differentials of the second kind (the meromorphic $1$-forms on $X$ with zero residues).
The infinite-dimensional vector space $\Omega^{(\mathrm{2nd})}$ has the natural skew-symmetric bilinear form1[x]1Analogues of the skew-symmetric bilinear form $\omega_X$ and of the algebraic de Rham theorem for meromorphic quadratic differentials will be considered in a forthcoming paper by the second author.
where $\mathrm{d}^{-1}\theta_1$ denotes any locally defined function $f$ such that $\mathrm{d} f = \theta_1$ (the local antiderivative). The ambiguity in the choice of $f$ does not matter.
Indeed, it is clear that the bilinear form $\omega_X$ is defined by a finite sum and the choice of an additive constant in the definition of a local antiderivative is irrelevant. The skew-symmetry of $\omega_X$ follows from the basic property
Using the bilinear form $\omega_X$, we can make isomorphism (3.1) more concrete. Namely, we have the following result (see [10], Ch. 6, § 8, [11], Ch. III, § 5.3, § 5.4, and [12], Theorem 4).
Theorem 1. The following assertions hold.
(i) The restriction of the bilinear form $\omega_X$ to $\Omega^{(\mathrm{2nd})}/\mathrm{d}\mathcal{M}$ is non-degenerate and
(iii) Let $D=P_1 + \dots +P_g$ be a non-special divisor of degree $g$ with distinct points. For every choice of local coordinates in the neighbourhoods of $P_i$, the vector space $\Omega^{(\mathrm{2nd})}\cap H^0(X, K_X+2D)$ has the basis $\{\vartheta_i,\tau_i\}_{i=1}^g$ symplectic with respect to the bilinear from $\omega_X$,
This basis consists of differentials of the first kind $\vartheta_i$ and differentials of the second kind $\tau_i $ uniquely characterized by the conditions
where $z_j=z(P_j)$ for a local coordinate $z$ at $P_j$, and $i,j=1,\dots,g$.
Proof. Let $(\theta)_{\infty}=\sum_{i=1}^{l}n_iQ_i$ be the polar divisor of $\theta\in\Omega^{(\mathrm{2nd})}$, $n_i\geqslant 2$. Since $D$ is non-special, $h^0(K_X-D)=0$, and by Riemann–Roch formula we have $h^0(D+nQ_i)=n+1$ for $n\geqslant 0$. Thus if $Q_i$ is not a point of $D$, there exists a meromorphic function $f_i\in\mathcal{L}_{D+(n_i-1)Q_i}$ such that
(In this case, $h^0(D)=1$, and hence one can not adjust the principle part of $df_i$ at $Q_i$ to cancel possible second order pole of $\theta$.) So, for $f=\sum_{i=1}^{l}f_i$, we have
The divisor $D$ is non-special, and hence the map $L$ is injective, and therefore, is an isomorphism, and we define $\vartheta_i$ and $\tau_i$ to have only non-zero components of $L$ to be, respectively, $\alpha_i=1$ and $\beta_i=1$. This proves Theorem 1.
Remark 1. The choice of a non-special effective divisor $D$ on $X$ with $g$ distinct points $P_i$ and local coordinates is as an algebraic analogue of the choice of $a$-cycles on a Riemann surface. Correspondingly, the differentials $\tau_i$ are analogues of differentials of the second kind with second-order poles, zero $a$-periods and normalized $b$-periods. The symplectic property of the basis $\{\vartheta_i,\tau_i\}_{i=1}^g$ is an analogue of the reciprocity laws for differentials of the first kind and the second kind (see [13], Ch. 5, § 1, and [14], Ch. VI, § 3).
Remark 2. Let $\Omega^{(\mathrm{2nd})}(2D)$ be the subspace in $\Omega^{(\mathrm{2nd})}$ spanned by $\tau_i$,
Then $\Omega^{(\mathrm{2nd})}(2D)$ and $H^0(X,K_X)$ are Lagrangian subspaces in $\Omega^{(\mathrm{2nd})}/\mathrm{d}\mathcal{M}$ dual with respect to the pairing given by the symplectic form $\omega_X$.
where $c\in H_1(X,\mathbb{Z}$), defines the canonical inclusion of the lattice $H_1(X,\mathbb{Z})$ into the vector space $H^{1,0}(X,\mathbb{C})^{\vee}$, which is the dual space of $H^{1,0}(X,\mathbb{C})$, and defines the Albanese variety
Using the Dolbeault isomorphism and the exponential exact sequence of sheaves on $X$, we have, for the Picard variety of line bundles of degree $0$ on $X$,
Thus, the holomorphic tangent space to $\mathrm{Pic}^0(X)$ can be identified with the vector space $H^{0,1}(X,\mathbb{C})$.
However, Theorem 1 allows us to describe the tangent space to the Picard variety in purely algebro-geometric terms. Namely, the following simple result holds.
Proposition 1. Each non-special effective divisor $D$ of degree $g$ on a Riemann surface $X$ defines an isomorphism
for any $\vartheta\in H^0(X,K_X)$, and is an isomorphism. This proves the proposition.
By identifying $H^{1,0}(X,\mathbb{C})^{\vee}$ with $\Omega^{(\mathrm{2nd})}(2D)$, we get an inclusion of the lattice $H_1(X,\mathbb{Z})$ into the vector space $\Omega^{(\mathrm{2nd})}(2D)$, which is defined as follows. Let $\theta_c$ be the $(0,1)$-component of the Poincaré dual of a cycle $c\in H_1(X,\mathbb{Z})$, so that
with vector space $\Omega^{(\mathrm{2nd})}(2D)$ of the differentials of the second kind with poles in $D$. Correspondingly, the holomorphic cotangent space is naturally identified with the vector space of $H^{1,0}(X, \mathbb{C})$ of differentials of the first kind, and the pairing with $\Omega^{(\mathrm{2nd})}(2D)$ is given by the symplectic form $\omega_X$.
§ 5. The Baker–Akhiezer function
Given a fixed non-special effective divisor $D_0=Q_1 \,{+}\,{\cdots}\,{+}\, Q_g$ of degree $g$ on $X$, let $\{\vartheta_i\}_{i=1}^g$ be the basis of $H^0(X, K_X)$ from Theorem 1 specialized to the divisor $D_0$. Consider the Abel–Jacobi map
We choose local coordinates at the points $P_i$ and put $z_i=z(P_i)$. It follows from (5.1) that $1$-forms $dz_i$ on $\mathrm{Jac}(X)$ at the base point $\mu^g(D_0)$ correspond to the differentials $\vartheta_i$, and the vector fields $\partial/\partial z_i$ correspond to the differentials of the second kind $\tau_i$ from Theorem 1. If divisor $D$ is also non-special, it follows from the group law on the Jacobian and Theorem 1 that $\mathrm{d} z_i$ and $\partial/\partial z_i$ at a point $\mu^{(g)}(D)\in \mathrm{Jac}(X)$ are given by the symplectic basis of $\Omega^{(\mathrm{2nd})}\cap H^0(X, K_X+2D)$ from Theorem 1.
Equivalently, these vector fields on $\mathrm{Jac}(X)$ can be described using the formalism of Lax equations on algebraic curves, developed by the first author in [5] and [6]. The main ingredients in [5] and [6] are stable vector bundles of rank $r$ and degree $rg$, Lax operators, certain meromorphic $r\times r$ matrix-valued $1$-forms $L(z)\,dz$ on a Riemann surface $X$, and $r\times r$ meromorphic matrix-valued functions $M(z)$.
Specialization to the Jacobian corresponds to the case $r=1$ and substantially simplifies construction in [5] and [6]. Namely, the meromorphic $1$-forms $L(z)\, \mathrm{d} z$ become differentials of the first kind $\vartheta\in H^0(X,K_X)$, while analogs of the meromorphic functions $M(z)$ are defined as follows.
It follows from the Riemann–Roch theorem that $\dim_{\mathbb{C}}\mathcal{L}_{D+D_0}=g+1$. Thus for any fixed choice of principal parts of $f$ at the points of $D_0$, not all of them are equal to zero, there is a unique, up to an inessential additive constant, function $f\in\mathcal{L}_{D+D_0}$ satisfying
at all points of the divisor $D=P_1+\dots+P_g$. Functions $f$, parametrized by their fixed principal parts at $D_0$, play the role of meromorphic functions $M(z)$ in case $r=1$; coefficients $\alpha_i$ depend on the principal parts at $D_0$.
between the rational functions $f\in\mathcal{L}_{D+D_0}$ and the vector fields on $\mathrm{Jac}(X)$. Along the integral curve $D(t)=P_1(t)+\dots +P_g(t)$ of $\mathscr{L}_{f}$, where $D(0)=D$, we have
where the dot stands for the $t$ derivative. If $X$ is a hyperelliptic curve, equations (5.3) are the classical Dubrovin equations arising in the theory of finite-gap integration for the Korteweg–de Vries equation [7], written in terms of the Abel transform. Using the Dubrovin equations, we see that, along the integral curve, equations (5.2) take the form
we see from (5.4) that $\Psi$ is a meromorphic function on $X\setminus D_0$ with simple poles only at $D$, simple zeros only at $D(T)$, and essential singularities at $D_0$.
The function $\Psi$ is nothing but the celebrated Baker–Akhiezer function, which was introduced by the first author in [8].
We leave it to the interested reader to describe by explicit formulas this connection between algebraic de Rham theorem and the integrable systems.
Bibliography
1.
W. V. D. Hodge and M. F. Atiyah, “Integrals of the second kind on an algebraic variety”, Ann. of Math. (2), 62 (1955), 56–91
2.
A. Grothendieck, “On the de Rham cohomology of algebraic varieties”, Inst. Hautes Études Sci. Publ. Math., 29 (1966), 95–103
3.
Ph. Griffiths and J. Harris, Principles of algebraic geometry, Pure Appl. Math., Wiley-Interscience [John Wiley & Sons], New York, 1978
4.
L. D. Faddeev and L. A. Takhtajan, Hamiltonian methods in the theory of solitons, Classics Math., Reprint of the 1987 original, Springer, Berlin, 2007
5.
I. Krichever, “Vector bundles and Lax equations on algebraic curves”, Comm. Math. Phys., 229:2 (2002), 229–269
6.
I. M. Krichever, “Isomonodromy equations on algebraic curves, canonical transformations and Whitham equations”, Mosc. Math. J., 2:4 (2002), 717–752
7.
B. A. Dubrovin, “Periodic problems for the Korteweg–de Vries equation in the class of finite band potentials”, Funct. Anal. Appl., 9:3 (1975), 215–223
8.
I. M. Krichever, “Integration of nonlinear equations by the methods of algebraic geometry”, Funct. Anal. Appl., 11:1 (1977), 12–26
9.
O. Forster, Riemannsche Flächen, Heidelberger Taschenbucher, 184, Springer-Verlag, Berlin–New York, 1977
10.
C. Chevalley, Introduction to the theory of algebraic functions of one variable, Math. Surveys, VI, Amer. Math. Soc., New York, 1951
11.
M. Eichler, Introduction to the theory of algebraic numbers and functions, Transl. from the German, Pure Appl. Math., 23, Academic Press, New York–London, 1966
12.
L. A. Takhtajan, “Quantum field theories on algebraic curves. I. Additive bosons”, Izv. Math., 77:2 (2013), 378–406
13.
K. Iwasawa, Algebraic functions, Transl. from the Japan., Transl. Math. Monogr., 118, Amer. Math. Soc., Providence, RI, 1993
14.
I. Kra, Automorphic forms and Kleinian groups, Math. Lecture Note Ser., W. A. Benjamin, Inc., Reading, MA, 1972
Citation:
I. M. Krichever, L. A. Takhtadzhyan, “Algebraic de Rham theorem and Baker–Akhiezer function”, Izv. Math., 88:3 (2024), 506–514