2. Historical Remarks and Comments
The subject of the first volume of Leonhard Euler’s famous manuscript [
1] is the theory of integrals of the following type:
which was developed by means of elementary (albeit sometimes too cumbersome) computational methods. Indeed, the integral (
1) depends on 10 parameters
. Euler discovered that, for some concrete values of the parameters, this integral gives known special functions involving
,
, and others. Of course, this type of integral—under the name
elliptic—was well known to many mathematicians of the seventeenth century as examples of integrals which cannot be expressed as combinations of the elementary functions and operations, and they give, in general, new
transcendental functions.
In the second volume of his fundamental folio, Euler investigated the development in a series of the integral:
where
,
, and
are parameters. As a result, the following expression (some later called it the hypergeometric series) appeared:
Soon afterwards, a great number of mathematical works were devoted to the study of hypergeometric series and functions as well as their analogues and various generalizations. A number of related problems were also investigated by the whole galaxy of outstanding mathematicians of the latter generation, including A. Legendre, C. Gauss (see [
2]), and N. Abel.
It is not difficult to verify that series (
3) satisfies the differential equation of the second order,
which is called the
hypergeometric equation. The second component of the fundamental system of the solution to (
4) is also expressed in terms of the above series. Using elementary transformations, we obtain one of the most known special cases of (
4),
which is often called the
Gauss–Legendre equation. Omitting the rational number
, we get the
Euler equation,
and so on.
Some time later, C. Jacobi introduced the notion of
period integrals when he studied integrals which are similar to (
1) and whose integrands contain polynomial denominators of the fifth or higher degree. Such integrals are often called ultra-elliptic or hyperbolic. At the end of the nineteenth century, L. Pochhammer, Ju. Mellin, and others investigated the integral representation of hypergeometric functions, and G. Lauricella and P. Appell [
3] discovered some new classes of hypergeometric functions of different types, depending on several complex variables. Almost at the same time, R. Birkeland and K. Mayr found series expansions for the roots and their powers in terms of multivariate hypergeometric functions in the sense of Horn (see [
4,
5,
6]). Nowadays, significant progress in the study of such problems has been achieved due to the works of I.M. Gelfand and his followers (see [
7,
8,
9]), K. Aomoto [
10], and many others.
It should be noted that, in view of these results, it is interesting to mention the famous Hilbert’s 21st problem, which concerns a certain class of systems of linear ordinary differential equations in the complex domain. More precisely, assume that the system,
has singularities
. This means that the entries of the matrix
are holomorphic in the complement
, where
is the Riemann sphere and
u is a vector column of unknown functions.
System (
7) is called
Fuchsian at the point if the entries of
have poles at worst of the first order in
. The system under consideration is
Fuchsian if it is Fuchsian at all points
,
, which are called
Fuchsian singularities. It should also be remarked that, in the scientific literature, Hilbert’s 21st problem is commonly referred to as the Riemann–Hilbert problem, which can be formulated as follows:
Does there exist a Fuchsian system with given singularities and a given monodromy? (see [
11] for further explanations and details).
Of course, similar problems arise in the case of several complex variables, where, as in the one-dimensional case of the Riemann sphere, the main role is played by an appropriate class of systems of differential equations for which such questions are relevant and interesting (cf. [
12]). In this paper, we considered a class of systems that appears in the theory of logarithmic connections associated with deformations of isolated singularities. It turns out that, in the multidimensional case, such systems have
logarithmic singularities and they can be considered as an analog of Fuchsian systems in the classical sense.
3. Gauss–Manin Systems and Connections
A. Grothendieck introduced the notion of Gauss–Manin connection and Gauss–Manin systems associated with differential equations considered in the famous proof of the Mordell conjecture by Yu. I. Manin. In the most generality, the corresponding concept was studied in detail by B. Malgrange, P. Deligne, and F. Pham (see [
13,
14]). Then, E. Brieskorn adopted their approach to the study of isolated hypersurface singularities and showed that the theory of singularities gives a clear interpretation of earlier results from quite a general point of view. Thus, among other things, he proved that the Gauss–Manin connection associated with the one-parameter principal deformations of an isolated hypersurface singularity can be presented as a system of ordinary differential equations with
regular singularities. Moreover, he gave an algebraic description of the connection and established its main properties (see [
15]). Furthermore, as it follows from his results, using matrix transformations with meromorphic entries, such systems or equations can be reduced to systems whose coefficients have at most poles of the first order. Afterwards, Brieskorn’s approach was modified and applied to the case of
complete intersections with isolated singularities by G.M. Greuel [
16], and to the case of
nonisolated singularities by H. Hamm [
17].
From another point of view, the concept of Gauss–Manin connection was also investigated by S. Ishiura and M. Noumi [
18]. They described the Gauss–Manin system for an
-singularity with the use of Hamiltonian representations. Such systems naturally appear also in the works of I.M. Gelfand, K. Aomoto, V.P. Palamodov, A.N. Varchenko, and their followers who studied integrals of the following type:
For instance, the case where
,
, are linear functions in variables
was considered in [
10,
19], some special cases where
is a linear deformation of the Pham singularities were studied in [
20] (see also [
21]), the case where
are quasihomogeneous polynomials was investigated in [
8], and so on. Moreover, in some cases, a series of explicit representations of solutions of the corresponding Gauss–Manin systems have been obtained in terms of generalized hypergeometric functions of some special types.
Nevertheless, the problem of the existence of new types of hypergeometric functions remains open, and the investigation of multivariate Picard–Fuchs systems (at least under some additional assumption) looks like a very interesting problem. In addition, there arose nontrivial problems of the description of the integrability condition in terms of the coefficient matrices of the corresponding system. For the simplest types of classical systems this problem is closely related to properties of the fundamental group of the complement of the singular locus of the system in question (see [
22,
23]), and so on.
In fact, a detailed analysis of examples shows that, as a rule, the solutions of multivariate Picard–Fuchs systems can be expressed in terms of known hypergeometric functions. Hence, these integrals satisfy systems of differential equations of hypergeometric type. On the other hand, K. Saito discovered that, in the case of an
-singularity, there are solutions of the corresponding system of uniformization equations which do not have an integral representation of Euler type (
1) (see [
24]). As a consequence, this gives an example of functions rather close to hypergeometric ones which satisfy Gauss–Manin systems of differential equations. In view of this phenomenon, it is natural to ask
whether there exist Gauss–Manin systems associated with the versal deformations of isolated singularities whose solutions cannot be expressed in terms of hypergeometric functions of the known types or their generalizations. 4. Versal Integrals and Horn Functions
The theory of Gauss–Manin systems for the
versal integrals of types
(closely related to the theory of distributions or generalized functions) has been developed by V.P. Palamodov (see [
25,
26]). Among other things, he showed that such systems naturally appear in the study of integrals of the following type:
where
. Here, the contour of integration
can be considered as a regularization of a vanishing cycle from the first homology group of the hyperelliptic curve given by the equation
(see [
27]). More precisely, the corresponding statement can be formulated as follows (see [
25,
28]).
Proposition 1. The versal integrals satisfy the following overdetermined system of differential equations: As an example, we will now describe the case of
-singularities in detail. Let
be a hypersurface given by the equation
Then,
has only one isolated singularity of type
at the origin
. The minimal versal deformation
is represented as the projection of the total deformation space
X defined by the equation
to the base space
of the minimal versal deformation with coordinates
. It is known that
f is a local trivial smooth bundle over the complement to the discriminant
D of the polynomial
and the differential forms
are holomorphic on
. Moreover, the first
k forms determine a
basis of the cohomological bundle
. Next, let us consider the
period integrals,
which satisfy the famous relations,
discovered by K. Mayr in a slightly different context (see [
5]). In fact, in the above notation, one can set
, where
. Next, we consider two vector-columns,
where
and the square matrix of order
which is the cyclic transform of the resultant matrix of
and its derivative
. By definition,
is equal to the discriminant of the polynomial
. It is clear that
As a consequence, we get the Picard–Fuchs system of Pfaffian type over the complement
, which can be represented as follows:
where the
j-th row of the matrix
consists of the following differential forms:
As a result, we obtain a simple algorithm for computing the Gauss–Manin system associated with -singularities.
For example, in the case of an
-singularity, it is not difficult to show that
where
and
. On the other hand, eliminating
and
from (
13), one can reduce the Gauss–Manin system (
14) to the following equation:
where
,
,
, and
are constants. Further transformations give us the Legendre and Euler equations (cf. Equations (
5) and (
6), respectively).
More generally, in a similar way, eliminating all
and
except for
and
from (
13), one can represent (see [
29], §7) the (minimal) equation for the Gauss–Manin system (
14) of an
-singularity in the following form:
which is called a
generalized Legendre equation for an
-singularity (V.P. Palamodov, Moscow State University seminar, 1980).
It should be noted that there is another explicit representation of the corresponding systems of differential equations of the first order as local Picard–Fuchs systems in the standard coordinates (see [
28]). The corresponding Gauss–Manin systems associated with two-parameter
principal deformations of the simple space curve singularities from the list [
30] have been computed by S. Guzev [
31]. However, in these works, the fundamental system of solutions was not determined explicitly.
Herein, we recall how to obtain the fundamental solution of such systems using basic properties of the elliptic and hyperelliptic integrals (see [
14]). More precisely, let
be the polynomial (
11) in one variable
z of degree
with coefficients
. Then, one can define the integral
where
is an analytic path in a domain of the Riemann surface of
. If the parameter
s changes in such a way that the roots of
F do not intersect the path
, then the integral defines an analytic function in
. If
is a cycle on the Riemann surface of
then it is possible to prolong analytically this integral along any path which does not intersect the hypersurface
, the discriminant of the polynomial
. The obtained analytic function is, in fact, a multivalent function on the complement
. In a similar way, one can define the integrals
so that
.
Now we explain how to determine solutions of the holonomic system (
14) for the integral
in an explicit form. First, we discuss the simplest case of the two-parameter deformation of an
-singularity.
Thus, using elementary transformations (see [
29]), it is possible to represent the system (
14) in the following form:
where
for
. As was already remarked in [
32], the fundamental solution to this system can be expressed in terms of the so-called generalized hypergeometric functions. More precisely, let
be a
fractional power series. The range of values of the indices
and
will be defined below. Substituting the series in (
19) and (
20), we obtain
where
,
, and
is the Pochhammer symbol.
As a result, equating all the coefficients to zero, one obtains the relation
Making use of this relation, one can compute
k linearly independent components of the fundamental solution of the above holonomic system, which are determined by their first terms only,
,
, and represent them as follows:
where
. It is possible to express both products in terms of Pochhammer symbols which occur in the generalized hypergeometric series in the following way:
Hence, for every
, the solution
u is the product of a monomial and a generalized hypergeometric series in
x. More precisely, we have
In order to analyze the second example associated with the versal deformation of an
-singularity, we need the following definition due to [
33]. The power series
is called a
k-tuple hypergeometric series if the quotients
are rational functions of the variables
. This series is often called a
k-dimensional hypergeometric series of Horn’s type. In particular, the series in Formula (
24) can be considered as a 1-dimensional hypergeometric series of Horn’s type. The aim of the calculations below is to deduce a similar result for the
minimal versal deformation of an
-singularity. It is possible to verify that the system (
14) can be expressed in the form
where
, are Pochhammer operators of the first order (cf. [
32]). Let us choose the first integrals
and
of Equation (
25) as new variables and suppose that the solutions can be expressed in terms of a fractional power series as follows:
where all the indices
n and
m are non-negative
rational numbers.
Using the following rules, one can obtain the system of finite
difference equations on the coefficients
as follows. Thus, by definition, the Pochhammer operators
,
, and
act on the monomial
as a multiplication by
, by
and
, respectively. After multiplication by
(resp. by
), the index
n (resp.
) of the coefficients increases by 1. As a result, the system of differential Equations (
26)–(
28) is reduced to the following system of finite difference equations:
Since the pattern of this system is a
parallelogram, one can readily find its solutions. It should be remarked that the last equation simplifies calculations essentially because it connects two adjacent vertices of the parallelogram. First, we change indices in the Equation (
32) in the following way:
and multiply Equation (
30) by
. Then,
(resp.
) is eliminated by virtue of Equation (
32) (resp. of Equation (
33)):
Similarly, we can eliminate
from Equation (
31):
so that
Assume now that
. It is evident that Equation (
35) multiplied by
gives Equation (
34). Hence, Equation (
30) can be omitted. Changing the indices in Equation (
32), one obtains the relations
Analogously, Equation (
35) yields
Let us combine Equation (
36) with (
37) and (
38) subsequently. As a result, one obtains the following expressions for nonzero terms of the fractional power series representing two components of the fundamental solution:
or, equivalently,
where
, and
or, equivalently,
where
and
.
The third component of the fundamental solution depends on the arbitrary constant
. More precisely, Equation (
36) implies the relations
while Equation (
35) yields
Now, taking a combination of (
43) and (
44), one gets the following expressions for the nonzero terms of the fractional power series representing the third component of the fundamental solution
or, equivalently,
where
.
It remains to consider the case where
, that is,
. It is clear that in such cases the indices
n and
m are nonintegers. On the other hand, if
, that is,
. Hence, (
30) implies
.
However, the third components of the fundamental solutions have zero coefficients with such indices as were computed before. In addition, it is also possible to verify that the system (
30)–(
32) has no nontrivial solutions with zero terms
,
,
without any restrictions on indices
n,
.
As a result, one obtains the nonzero terms of the three linearly independent components (
40), (
42) and (
46) of the fundamental solution to the maximally overdetermined system of differential Equations (
30)–(
32) at the points of the parameter domain
, contained in the complement of
, the discriminant of the polynomial
. It is given by the equation
which determines the singular locus of the system in question.
It is not difficult to see that the two components (
40)–(
42) are, in fact, 2-tuple hypergeometric series of Horn’s type while the third (
46) can be considered as a fractional power series depending on two variables
It should be also noted that systems (
25)–(
28) were considered in [
32], where the author constructed three fractional power series which are two-dimensional hypergeometric functions of Horn’s type. One of them is a series centered at points of the domain
, while the other two series are centered at points of the line
, contained in the singular locus of the system. Since the discriminant
D coincides with the singular locus of the system, these two series are multi-valued analytic functions branching at the points of
. It was also stated there that these three series represent all linearly independent components of the fundamental solution of the system outside the discriminant. However, whether they are linearly independent in their common domain of analytic continuation remains unclear. In contrast with the results of [
32], three series, (
40), (
42) and (
46), are linearly independent by construction; they are centered at points
outside of the discriminant set
. Therefore, we have just obtained the fundamental system of solutions to (
25)–(
28).
For completeness, it should be noted that other explicit expressions for the fundamental solutions to Picard–Fuchs systems associated with
-singularities,
, can be found in [
21,
34,
35].
5. Logarithmic Connections
Another original approach to the study of the Gauss–Manin connection was developed by K. Saito [
24]. More precisely, he calculated the corresponding system of differential equations in the case of the minimal versal deformation of an isolated hypersurface
-singularity and obtained a nice representation of this system in terms of meromorphic differential forms with
logarithmic poles along the discriminant of the deformation. Among other things, he also gave a simple proof that the connection is
regular singular.
By definition, a meromorphic differential q-form , , on a complex space S is called logarithmic along a divisor if and the total differential have poles along D of order at most one. Equivalently, if is a local equation of D, then both forms and are holomorphic on S.
The corresponding sheaf of logarithmic differential
q-forms is usually denoted by
. Thus,
is contained in the
-module
, which consists of all differential forms having poles of any order along
D. It should be observed that the logarithmic differential forms have many remarkable analytic and algebraic properties (see [
36,
37]).
We will denote by
the sheaf of
logarithmic vector fields along
D on
S; its stalks consist of germs of holomorphic vector fields
on
S such that
. In particular, the vector field
is tangential to
D at its
smooth points. The inner multiplication of vector fields and differential forms induces a natural pairing of
-modules,
Moreover, for , this -bilinear mapping is a nondegenerate pairing so that and become -dual.
Herein, we will consider the case where
and
is a
free -module of rank
m. Then,
,
, and the divisor
D is usually called
free or the
Saito divisor (see [
38,
39]). The following useful statement is due to K. Saito [
24].
Proposition 2. Suppose that there are vector fields such that their coefficients of , , form the -matrix and , where c is a unit. Then is a basis for the free -module . In particular, is a free -module and vice versa.
For example,
is free when
D is the discriminant of the versal deformation of an isolated hypersurface singularity. Following K. Saito, one can exploit this fact as follows (cf. [
24]).
In general, a connection ∇ on a free
-module
with
logarithmic poles along
is defined as a morphism,
satisfying the following two conditions:
It is easy to analyze the case where
. Let
be a basis. Then,
By definition, the connection ∇ is
integrable if the composition
is equal to zero morphism. Thus, in this case,
or, equivalently,
, where
is the
connection matrix of
.
Among other things, Saito’s considerations imply (see [
24]) that, in the case where
is the minimal versal deformation of an isolated hypersurface singularity (or, more generally, of an isolated complete intersection singularity), and
is the discriminant of the deformation, then
and
are
free -modules of
equal rank. Hence, in general, the Gauss–Manin connection,
can be represented as connection (
48) with
.
Furthermore, let
be a holomorphic form representing a class of the relative de Rham cohomology in
, where
B is a sufficiently small ball in the complement
. Let
be any continuous family of cycles over
. Then, the integral
is a holomorphic function on
. Using Stokes’ theorem, one can verify
where
, and
are holomorphic in
. This means that the function
is constant if and only if
for all
. In particular, it follows that the restriction of connection (
49) to
:
can be presented in such a way:
where
and
is a suitable representative of the form
in the corresponding relative cohomology group. Let
,
, be holomorphic forms on
, whose restrictions on every fibre
of
generate the cohomology group
. The corresponding sections
, of the bundle
determine its basis. By construction,
, where
are holomorphic on
. The corresponding holomorphic sections of cohomological bundle can be written as
, where
are holomorphic on
. As a result, we get
where
is a vector-column, and
is a square matrix with entries
. In conclusion, one can integrate the sections
along any continuous family of cycles
.
As before, we denote the corresponding integrals by
. Then, we obtain the following system of differential equations corresponding to connection (
14):
where
is the
connection matrix of ∇ and the entries of
are holomorphic differential 1-form on
. Such a system is often called a
local Picard–Fuchs system.
It is not difficult to see that the entries of the matrix
of the presentation (
14) correspond to the entries of the connection matrix ∇ from the presentation (
48). Moreover, these considerations give us an algorithm for computing the Gauss–Manin system, at least in the case where
is a
free -module of the
same rank as
. Namely, first of all one should have to find all connections ∇ on
, where
is the discriminant of the minimal versal deformation. After this, it remains to select those which correspond to the required connection on
.
6. Examples of Logarithmic Connections
It is not difficult to apply the above observations to the case of an
-singularity. Assume that the polynomial
determines the minimal versal deformation of an
-singularity and the discriminant
is determined by the equation
, where
and
. In order to describe connection (
48), we first observe that the
free -module
is generated by two logarithmic differential 1-forms:
where
(cf. (
15)) is contained in the
torsion submodule of the module of Kähler differentials
on the hypersurface
D (see [
37,
39]). More precisely,
It is not difficult to verify the following relations:
Then, using the natural grading on the modules
,
and
, which is induced by the weights of
and
(equals to 3 and 2, respectively), we see that the connection matrix ∇ has the following form:
where
are
-valued parameters.
The integrability condition
implies
. Thus, we have the three-parameter family depending on
, and
g, which defines the connection (
48). The characteristic (or initial, or indicial) polynomial of the corresponding system is equal to
On the other hand, one can consider two
versal integrals (
18) associated with an
-singularity:
where the contour of the integration is a homology cycle of the elliptic curve given by the equation
. Using the presentations (
11) and (
12), it is not difficult to verify by straightforward computations that the Gauss–Manin system for versal integrals can be presented in the form (
14) with the connection matrix,
where
q is a
-valued parameter,
. Thus,
is a
logarithmic differential form and
. The characteristic polynomial is equal to
Setting
, we obtain the presentation of K. Saito from [
24]:
As a result, the corresponding Picard–Fuchs system is reduced to the differential equation of the second order for the integral
; it is, in fact, the classical Legendre Equation (
5):
where the parameter
t is equal to
up to an invertible factor.
Indeed, the latter equation is a particular case of the usual hypergeometric differential equation. More precisely, it is the so-called
Fuchsian equation with singular points
on the Riemann sphere
(see [
11]). Changing the parameter
and using the Formula ([
11], (7.3.6)), where
we obtain the following
Fuchsian system:
where
denotes the vector-column
of unknown functions. It is not difficult to compute that the characteristic polynomial of the Equation (
51) has the double root
. However, all connections ∇ from the above three-parameter family with
have initial polynomials of the same kind. Consequently, all corresponding Fuchsian systems or equations are reduced to Equation (
51) (cf. also [
40]).
The next good illustration is an
-singularity. First, we recall some computational results due to K. Saito [
24] when he studied connection (
14)
for the versal deformation of an
-singularity.
Similarly to the above notation, let
be the equation of the minimal versal deformation of an
-singularity. Then, its discriminant
is given by the Equation (
47):
Let us now consider the following three vector fields tangent to the discriminant:
The corresponding coefficient matrix
has the following representation:
It is easy to verify that
. In our case, the
-module of logarithmic differential forms
is free; its
dual basis with respect to the vector fields
is defined by the
-minors of the matrix
. More precisely, we have
and the following relations:
In particular, , where is the contraction along the vector field V.
In order to describe the connection on
, we recall that the weights of variables
induce the natural grading on the modules of logarithmic differential forms and vector fields. As in the former example of an
-singularity, one can see that such a connection can be represented with the following connection matrix:
where
are 14 parameters with values in
. The matrix ∇ satisfies the
integrability condition,
Since there is the relation
where
, then the columns of the matrix
define the relations between the total differentials of variables
and the logarithmic forms
:
Performing some calculations, we get the following system of 10 equations on 13 parameters, which are equivalent to the integrability condition (
52):
All equations do not depend on the parameter
. It is clear that the Equations (
1), (
5) and (
7) imply the condition
Further, the determinants of two triples (
2), (
8), (
9) and (
3), (
4), (
6) of linear homogeneous equations with respect to variables
and
, respectively, are equal. Replacing two Equations (
3) and (
8) by the determinant and excluding
, we get the following standard basis of the corresponding ideal:
As a result, all solutions of this system of three equations depending on seven variables determine the four-dimensional space containing a line corresponding to the free parameter .
Now take a look at system (
14):
Recall that (
14) is, in fact, a maximal overdetermined system of differential equations whose solutions are
regular singular. Consequently, the restriction of this system to an arbitrary
generic curve of the base space is reduced to the differential equation with
regular singularities of the type
Changing
, one can transform this system to a similar one with
Moreover, up to such transformations, the above system is defined completely by the
elementary divisors of the corresponding initial polynomial
of the characteristic matrix
associated with the
residue matrix (see [
40], Ch.XI, §10), so that
It is clear that the polynomial of the third degree satisfies one of the following conditions:
- (1)
has one root of multiplicity three;
- (2)
has two separated roots one of which has multiplicity two;
- (3)
has three separated roots.
It turns out that the second and third cases are realized in our situation. As a result, one obtains
two different types of Picard–Fuchs systems. In other words, we get two irreducible components parameterizing the family of all “uniformization equations” in the sense of K. Saito. In fact, he found the following two sets of parameters for the connection ∇ (see [
24], §3). More precisely, the first one (Type I) is defined as follows:
while the second (Type II) is
The characteristic polynomial of the system of type I is equal to
while the system of type II has the characteristic polynomial
. The latter is a slightly corrected expression from [
24].
In both cases, we see that is a -valued parameter and both characteristic polynomials are equal if .
It remains to be shown that the system of type I corresponds to the Gauss–Manin connection associated with an
-singularity. In fact, the generic fibre
of the minimal versal deformation of an
-singularity
is isomorphic to a hyperelliptic curve of genus 1. Hence, its first homology and cohomology groups with coefficients in
both have rank 2. The period integrals
where the contour of the integration
is a homology cycle of a hyperelliptic curve given by the equation
, satisfy the differential equation of the
second order although the Milnor number of the distinguished fibre
is equal to 3. Therefore, the corresponding system consisting of three differential equations has the characteristic polynomial with a
multiple root (cf. [
24]).
To understand the obtained statement better, we will write out explicitly the system (
55) of differential equations for versal integrals. Using the presentation (cf. [
25]) again, it is possible to compute the coefficient matrix
of the Gauss–Manin system (
14) for the versal integrals:
where
q is a
-valued parameter and
,
,
are logarithmic differential forms defined as above. The characteristic polynomial is equal to
that is, in the above notations, we have
.
It should be noted that K. Saito found solutions of the uniformization equation only in the case when
(or, equivalently, for
in our notation). In this case, the solutions can be expressed in terms of the classical Weierstrass elliptic function. In addition, as follows from his calculations, the system of type II, whose solutions “do not have an Euler integral representation” in the sense of relation (
1), arises in an essentially different context, other than Gauss–Manin systems associated to an
-singularity.
The next (highly nontrivial) step was made by J.Sekiguchi, who developed this method in order to describe the fundamental system of solutions to the uniformization equation associated with a
-singularity in terms of the Weierstrass elliptic function, similar to the approach in [
41].
7. Picard–Fuchs Systems of Pfaffian Type
Now we will apply the ideas described above to some special cases. Let
be a holomorphic function on
, and let
be the hypersurface defined by the equation
. Suppose that
has no multiple factors, that is,
D is
reduced. Then the module
of logarithmic (along a divisor
D) differential forms is defined (see
Section 5). It is known [
37] that there exists the following exact sequence of
-modules:
where
is the module of holomorphic differential 1-forms on
S; it is generated by the differentials
over
. Denote by
the module of regular Kähler differentials on
, and by
the torsion submodule of
. The support of
is contained in the singular locus
of the hypersurface
D and it has a system of generators consisting of at least
elements.
Let us consider a system of linear differential equations on
S with
meromorphic coefficients that are contained in the module of logarithmic differential forms
:
where
is a vector-column of unknown functions,
, the differential 1-forms
, correspond via (
57) to nonzero elements of the torsion submodule
, and
, are coefficient matrices with holomorphic entries satisfying the integrability condition:
In particular, this implies the relation
where, for convenience of notation, the total differential
is denoted by
.
The system of linear differential Equation (
59) is, in fact, the local Picard–Fuchs system (cf.
Section 4). The corresponding
global analog (often called a Fuchsian system [
11]) we can relate with local systems as follows.
Assume that
is a homogeneous or, more generally, quasihomogeneous
polynomial relative to the variables
of weights
. Then,
h determines the hypersurface or divisor
D on a compact complex variety
. In the homogeneous case,
V is the
-dimensional projective complex space
while in the second case
V is the
weighted projective space
In both cases, one can consider the Picard–Fuchs system (
59) of linear differential equations given on
V.
The following assertion characterizes a basic property of such systems (cf. [
42]).
Proposition 3. Let be the product of irreducible polynomials with no multiple factors, and let be the corresponding divisor. Then the system (59) has only regular singularities along the divisor . In other words, the system (59) is of Fuchsian type in the broad sense. Proof. First remark that (
59) is regular singular along the
nonsingular part
of the divisor
D denoted by
. Then one can apply a theorem of P. Deligne ([
13], Th. 4.1) in our case. Namely, this theorem asserts that the property of such a system to be regular singular along
implies that it is regular singular along the
whole D. □
The following example is useful. Let
be the local decomposition of a reduced divisor
D at the distinguished point
, and
, the corresponding primary decomposition of the function
. It is evident that differential forms
, are logarithmic and their images via (
57) are contained in the torsion module. Thus, the following Picard–Fuchs system is well defined:
It should be remarked that, if
D is the union of
hyperplanes, then in the general case the system (
61) can be transformed to a similar one with constant matrices
by means of a holomorphic change of variables (see [
43]).
Assume additionally that
is a set of
homogeneous polynomials of degrees
, respectively, defined on the
m-dimensional projective complex space
with homogeneous coordinates
. Suppose also that all irreducible projective hypersurfaces,
are
reduced and
nonsingular, all entries of the matrices
,
are complex numbers and
In fact, this condition is similar to [
11], (1.2.3). Under the above assumptions, the system of linear differential Equation (
61) on the projective space
has been investigated by R. Gerard and A. Levelt in [
44]. Moreover, they gave a classification of such systems that are often called the Pfaffian systems of Fuchs type or simply Fuchsian systems (see also [
11]). For completeness, it should be noted that some cases in their classification were omitted; the remaining cases are described in [
45]. Another special case, where
,
, are the coordinate hyperplanes, was analyzed by M. Yoshida and K. Takano [
46], A.A. Bolibruch [
47], and others.
More generally, one may consider the system (
61), where the equations
define the
singular hypersurfaces
. Of course, any such system is a very special case of (
59). Thus, there arises the problem how one can describe the integrability condition (
60) in terms of the coefficient matrices
.
First, suppose that all entries of the matrices
, are
complex numbers, and
nonsingular divisors
are in a “general position” in the sense of [
44]. This means that, for every couple
, there is a point
such that
for all
. In such cases, the integrability condition (
60) implies the permutability of the matrices
, so that the following commuting relations hold (see [
44]):
Thus, this condition is fulfilled if the divisors have normal crossings, that is, the total differentials and are linearly independent for all .
In fact, commuting relations between matrices
reflect topological properties of the fundamental group
of the complement of the union
, and vice versa. For instance, there is the following useful observation (see [
22]):
Proposition 4. Let be the union of irreducible nonsingular curves in , and the fundamental group is generated by corresponding to the components , . Then, every commutator naturally corresponds to the commutator of the coefficient matrices of the system (
61)
, where are matrices with constant entries, that is, . In view of this result, it is possible to compute the
complete set of commuting relations which are equivalent to the integrability condition for the Fuchsian systems having the same singular loci as the
classical hypergeometric functions of Appell and Kampé de Fériet in two variables (see [
48]):
As a development of this result, one can obtain analogous properties in the case where
D is an arbitrary set of
hyperplanes in
, (see [
23]). Next, the case where
D is the union of
nonsingular hypersurfaces in
of arbitrary degrees was analyzed in [
49].
In view of Proposition 4, it should be noted that there is very probably a kind of “duality” between the module of logarithmic differential 1-forms
and the local fundamental group
of the complement of the divisor
(as was conjectured by K. Saito in [
24]).
It is not difficult to see that the system (
59) is also well defined in the case where
D is the union of
singular hypersurfaces. In general, the problem of describing commuting relations is quite nontrivial and intrigued. However, it is possible to give a partial affirmative answer in the following particular case (see [
24,
36], (2.9)).
Proposition 5. Suppose that , are hypersurfaces whose singular loci have codimensions of at least two satisfying the following two conditions:
- (∗)
has normal crossing with outside of some analytic subset containing in where the codimension of is at least two,
- (∗∗)
for all different triples .
Then, the -module is generated by the following logarithmic differential 1-forms: Indeed, in the case where the hypersurface
D is the union of normally crossing divisors that satisfy the above conditions, there is a set of generators of the torsion
-module
containing the following
differential 1-forms:
where the form
is omitted for some
i. Therefore, in this situation, the system (
61) can be considered as a special case of (
59).
The following statement combined with Proposition 4 enables us to obtain all commuting relations similar to those (
63) of the system (
61) for any divisor with normal crossings
, where
, are irreducible projective hypersurfaces in
(see [
50], (Section 3.1.2)).
Proposition 6. Let D be a complex projective hypersurface in Suppose that there is a subvariety of codimension two such that the singularities of have normal crossings. Then, the complement of D in has an abelian fundamental group .
Consider the logarithmic Picard–Fuchs system (
59) having constant coefficient matrices
that satisfy the condition (
62). In fact, one may regard such system associated to any divisor
satisfying the assumptions of Proposition 6 as a natural generalization of the Pfaffian system of Fuchsian type in the sense of [
44] to the case of arbitrary reduced divisors. At present, it is still unclear how the method of computing the commuting relations with the aid of properties of the fundamental group can be extended to the case of the arbitrary divisor.
In addition, we also note that the theory of logarithmic forms is an effective tool for computing commuting relations for the systems (
59) and (
61) in more general cases. The following example concerns the case of an
irreducible divisor
D when the system (
61) has trivial commuting relations only. On the other hand, the logarithmic Picard–Fuchs system (
59) is quite nontrivial.
Thus, as in the beginning of
Section 6, let us consider the case of an
-singularity. Then the polynomial
determines the discriminant of its minimal versal deformation. The corresponding Picard–Fuchs system has the following form:
where the differential form
corresponds to the element of the torsion submodule
. Assume additionally that all entries of the matrices
,
, are
complex numbers. Set
We already know that the logarithmic differential forms
and
are
free generators of
, and the following relations hold:
Making substitutions in (
60) and using the freeness of the module
one can obtain by straightforward calculations the following commuting relations, which are direct consequences of the integrability condition (
60):
For completeness, it should be remarked that there is a nice classification of Pfaffian systems of Fuchsian type associated to the discriminant of an
-singularity (see details in [
51]).
In the same manner, using elementary properties of logarithmic differential forms, it is possible to obtain commuting relations for Picard–Fuchs system (
59), associated with divisor
D, for which
is a free
-module. For example, consider the case
and suppose that
D has two irreducible components with
non-normal crossing:
Then, the module
is
free again. More exactly, it is generated by two logarithmic differential forms
and
, where
. It is not difficult to check that
Hence, the system (
61) of the form
, where
is, in fact, equivalent to the following Picard–Fuchs system:
Moreover, we have the following relations:
Assume that all entries of the matrices
and
are constant, that is,
. Again, making substitutions in (
60) and using the freeness of the module
, we then obtain the following commuting relations:
which implies the identity
.
As was remarked in [
24], in this case, the fundamental group of the complement
is
not abelian: it is generated by two elements,
and
, corresponding to the irreducible components
and
satisfying the relation
. Moreover, in this situation it is still possible to obtain the commuting relation between matrices
and
using the method from [
22] because all relations between the generators of the fundamental group can be expressed in terms of the commutator
.