groupoids derived from the simple elliptic singularities
Abstract.
K. Saito’s classification of simple elliptic singularities includes three families of weighted homogeneous singularities: , and . For each family, the isomorphism classes can be distinguished by K. Saito’s -functions. By applying the Mather-Yau theorem, which states that the isomorphism class of an isolated hypersurface singularity is completely determined by its -th moduli algebra, M. Eastwood demonstrated explicitly that one can directly recover K. Saito’s -functions from the zeroth moduli algebras. This research aims to generalize M. Eastwood’s result through meticulous computation of the groupoids associated with simple elliptic singularities. We not only directly retrieve K. Saito’s -functions from the -th moduli algebras but also elucidate the automorphism structure within the -th moduli algebras. We derive the automorphisms using the methodology of the -th Yau algebra and establish a Torelli-type theorem for the -family when . In contrast, we find that the Torelli-type theorem is inapplicable for the first Yau algebra in the -family. By considering the first Yau algebra as a module rather than solely as a Lie algebra, we can impose constraints on the coefficients of the transformation matrices, which facilitates a straightforward identification of all isomorphisms. Our new approach also provides a simple verification of the result by Chen, Seeley, and Yau concerning the zeroth moduli algebras.
MSC(2020): Primary14B05; Secondary 32S05
Keywords: simple elliptic singularity, Tjurina algebra, Yau algebra, weighted homogeneous singularity, Groupoid
E-mail: [email protected]
Beijing Institute of Mathematical Sciences and Applications, Beijing 101408, P. R. China
E-mail: [email protected]
E-mail: [email protected]
1. Introduction
We present necessary definitions and auxiliary results concerning the moduli algebras of singularities. Let be a formal power series ring over with maximal ideal . Let be a germ of an isolated hypersurface singularity at the origin in represented as the zero locus of polynomial (or analytic function) .
Definition 1.
The -th moduli algebra of is defined by
where denotes the Jacobi ideal of .
It is convenient to denote the local function algebra (or coordinate ring) of by
The zeroth moduli algebra is also called Tjurina algebra, and its dimension is referred to as the Tjurina number. It is well known that the moduli algebra can serve as a base space of versal deformation of singularities [1]. In our recent work [2], we computed the dimension of the -th Tjurina algebra of weighted homogeneous singularities. For further related studies on on Tjurina algebra, we refer to [3, 4, 5, 6].
The -th moduli algebras are important due to the Mather-Yau theorem [7]. The extension of Mather-Yau theorem to positive characteristic was studied in [8]. We restate the theorem in a slightly different version [3].
Theorem 2 (Mather-Yau theorem).
Let for denote analytic functions associated with isolated singularities respectively. Then the following conditions are equivalent:
-
(1)
The function is contact equivalent to , i.e., there exists some automorphism and a unit such that
-
(2)
For all , there exists some isomorphism of -algebras.
-
(3)
There exists some such that as -algebras.
Among all isolated hypersurface singularities, weighted homogeneous singularities have been of particular interest. Recall that a polynomial is weighted homogeneous of type , where are fixed positive rational numbers, if it can be expressed as a linear combination of monomials for which
for some constant .
We view as a graded algebra by assigning the weight of to be . Let be two homogeneous ideals. For the local algebra with , we say that homomorphism is homogeneous if preserves the weights of . For the case where both and are weighted homogeneous, a contact equivalence as described in (1) of the Mather-Yau theorem induces a homogeneous homomorphism by truncating the higher-order terms. Therefore, we obtain the equivalent condition of the Mather-Yau theorem:
-
(4)
(Suppose that and are weighted homogeneous.) There exists some homogeneous automorphism and a constant such that
In [9], Saito defined a simple elliptic singularity to be a normal surface singularity such that the exceptional set of the minimal resolution is a smooth elliptic curve, and classified those singularities that are hypersurface singularities into the following three weighted homogeneous cases:
(1) | ||||
where the restrictions on ensure the singularities are isolated:
Using a result of Noumi and Yamada [10] on the flat structure on the space of versal deformations of these singularities, Strachan [11] explicitly constructed the -functions for these families. Saito determined the isomorphism classes of these families by computing the -functions and concluded the criteria that two singular with coefficients and are contact equivalent if and only if , where
For , we denote by with the corresponding -th moduli algebra (resp. local function algebra) of -families listed above. With the help of the Mather-Yau theorem, Eastwood [12] demonstrated explicitly that one can recover directly Saito’s -functions from the zeroth moduli algebra .
In this paper, we aim to recover Saito’s -functions from the -th moduli algebra with by applying the -th version of the Mather-Yau theorem. Furthermore, we pose a finer question: how can we completely describe the homogeneous isomorphisms from to ? To understand the algebraic structure in this context, we introduce the groupoid of simple elliptic singularity. Our complete description of the groupoid provides more information than the results presented in [12].
A (small) groupoid is a small category in which every morphism is invertible. The notion of groupoid generalizes the concept of a group by replacing the binary operation with a partial function. In this paper, we are interesting in the groupoid [13, 14] presented in the following manner.
Definition 3 (Action Groupoid).
If the group acts on the set , then we can form the action groupoid (or transformation groupoid) as follows.
-
(1)
The objects are the elements of .
-
(2)
For any two elements , the morphisms from to correspond to the elements such that .
-
(3)
Composition of morphisms interprets the binary operation of .
Definition 4.
For and , we define the groupoid of simple elliptic singularities as follows.
-
(1)
The objects of are precisely the coefficients .
-
(2)
For , the morphisms of to are the homogeneous isomorphisms from to modulo the scalars in .
This paper devotes to the explicit computation of the groupoids . From the lifting property, any morphism of lifts to an automorphism of weighted projective spaces. Up to the -action, the morphism is essentially contained in the weighted projective general linear group , which is the quotient of the general linear group by its subgroup of diagonal matrices
for some weights . When is large, we obtain
since the morphism in preserves the defining functions (see Lemma 7).
For or , contains more morphisms due to the fact that when is weighted homogeneous. Finally, we also note that there exist some coefficients that behave weirdly, referred to as jump points according to references[15, 16]. The jump points are listed as follows.
-
(1)
For the -family, ;
-
(2)
For the -family, ;
-
(3)
For the -family, .
Here and afterwards, denotes a primitive cubic root of unity. In the following subsections, we will formulate the main results in detail.
1.1. Groupoid of the -family
In [17], Chen, Seeley and Yau provided a detailed characterization of homogeneous isomorphisms of the moduli algebras arising from the -family, i.e., the morphisms in . Our newly elaborated technique for studying the isomorphisms of differs from the previous approach used in [17]. In fact, our idea is motivated by the work of Seeley and Yau [15], in which the authors studied the derivation Lie algebras associated to the zeroth moduli algebras of and -families. They proved the Torelli-type theorem saying that the Lie algebras and are isomorphic if and only if the corresponding singularities are contact equivalent to each other. In [18], the Lie algebras of simple elliptic singularity were computed along with several elaborate applications to deformation theory.
For the -family, we consider the derivation Lie algebra, written as , associated with the first moduli algebra . By equipping with a natural module structure over the moduli algebra, we discover that the invariants arising from the Lie algebra , together with , are applicable in determining the isomorphisms of the -family. It turns out that the groupoids and coincide indicating a simpler proof for Chen-Seeley-Yau’s result in [17]. However, we note that the Lie algebra itself (without the module structure) is insufficient to determine the isomorphism classes of the -family. In other words, there is no Torelli-type theorem for analogous to that in [15]. The failure of a Torelli-type theorem for the zeroth Yau algebra can also be found in [19].
We define the subgroup of generated by the matrices
Note that and generates the symmetric subgroup on three letters. One can show by direct computation that contains totally matrices and all the entries belong to . Let be the group of fractional linear transformations generated by
and
It can be checked that is a group of order , isomorphic to the alternative group on four letters. In addition, there exists a group homomorphism which sends to respectively.
We restate and generalize the main result in [17] in the following form.
Theorem A.
(1) For the case or , the groupoid restricted to the objects is represented by the action groupoid
where acts on through the morphism defined above.
(2)For or , we find that equals the action groupoid .
(3)(The Failure of a Torelli-Type Theorem) The first Yau algebra of the -family is independent on the parameter and gives rise to a continuous family of -dimensional representations of a solvable Lie algebra.
We remark that for or , is a subcategory of , but it is not fully faithful. Indeed, it can be concluded that is the group generated by and the diagonal matrices. As a consequence of Theorem A, K. Saito’s -function of the -family can be obtained by observing that the integral functions on fixed by (or equivalently) are generated by .
1.2. Groupoid of the -family
Consider the group consisting of the following six fractional linear transformations
Note that is isomorphic to the symmetric group . According to [15], instead of -function , we have the criteria that the coefficient is equivalent to in if and only if for a unique . We define subgroup of generated by
This group is to shown to preserve the polynomial and is isomorphic to the symmetry group . Thus, there are totally elements of which are represented by the following matrices
where and range over .
Notice that has a Klein four-group as a proper normal subgroup, namely the even transpositions , with quotient . In according to the mapping , there exists a homomorphism
defined as
(2) | ||||
In particular, the kernel of gives a subgroup
which is isomorphic to the Klein four-group. Using the notation
for , we obtain the following result which parallels to Theorem A:
Theorem B.
(1) For , the groupoid restricted to is identical to the action groupoid
Here and take values from , and they are not simultaneously zero; and the matrices act on the points of through , restricting to the coordinates.
(2)If or , then
(3) (Torelli-Type Theorem)For , the first Yau algebra determines the isomorphism class of the -family.
Similar to the -family, the matrices of the form generate a proper sub-groupoid of when or . The integral functions on fixed by (or equivalently) are generated by .
1.3. Groupoid of the -family
The criteria for the isomorphism class of the -family is much simpler. It follows from the -function that is isomorphic to in if and only if with . Let be the matrix of the form
with . The collection of matrices with , forms a subgroup of with . For or , the matrix induces an isomorphism from to , represented by the transformation
Theorem C.
(1) For or ,the groupoid is given by the union of the action groupoid
and the group of diagonal matrices (as a groupoid with a single object ).
(2) While for or , the groupoid is given by the action groupoid .
2. Preliminary and notations
2.1. Homogeneous isomorphism
In this paper, we denote by the formal local ring , and assume that the weight of equals . For simplicity, we assume the weights are positive integers. Given and , the two homogeneous ideals of , i.e., the generators in and are weighted homogeneous, we consider the graded analytic algebras and . By the well-known lifting lemma, the isomorphism from to is given by such that
One can express as
such that the monomials in have weight , while those in have weight . The homogeneous component of , given by , forms an isomorphism if and only if is an isomorphism. Consequently, each isomorphism from to induces a homogeneous isomorphism.
A -action is present on defined as:
resulting in a trivial automorphism of . Moreover, if represents a homogeneous isomorphism, then also signifies a homogeneous isomorphism. This establishes the -action within the set of homogeneous isomorphisms.
2.2. Yau algebra with graded module structure
A graded Lie algebra is an ordinary Lie algebra together with a gradation of vector spaces
such that the Lie bracket respects this gradation:
Definition 5.
For an isolated hypersurface singularity determined by a polynomial , we define the -th Yau algebra of by
namely the derivation Lie algebra of the -th moduli algebra.
We need to introduce a graded -module structure of . Denote by the partial derivation with respect to . By definition, a derivation is of the form
where the coefficients satisfy the condition
In this way, we obtain the embedding map of -modules:
(3) |
where denotes the free -module with generator .
With the assumption that is a weighted homogeneous polynomial, the algebra admits a graded algebra structure since the ideal is homogeneous. For a monomial in , we define the notion of degree as follows:
Based on this grading, the free -module is regarded as a graded Lie algebra. Similarly, the -th Yau algebra inherits a graded Lie algebra structure.
We utilize the crucial observation that a homogeneous isomorphism of -th moduli algebras leads to an isomorphism of the corresponding -th Yau algebras. For weighted homogeneous polynomials and , a homogeneous isomorphism
induces a Lie-algebraic homomorphism
Such homomorphism is -equivalent meaning that
and
hold for and .
2.3. Notation of homomorphisms for Lie algebra
When focusing on the case where from -families, we consistently employ the subsequent notations to denote homogeneous isomorphisms of the -th moduli algebras of simple elliptic singularities throughout this paper.
Notation 6.
Suppose that are the coordinates of and are the coordinates of . We denote a homogeneous isomorphism from to through coordinate transformations as:
(4) |
where ’s are homogeneous functions.
Specifically, when all , uniquely corresponds to a nonsingular linear transformation of . One can represent as a linear transform
(5) |
Its inverse can be expressed as
(6) |
We define to be the free -module generated by .
As in (3), we possess the embedding map
According to Leibniz’s Rule, we derive the formulas for the differential of :
(7) | ||||
In this way, we obtain a -equivalent map
which preserves the Lie structures.
2.4. Basis property of groupoid
We examine some fundamental properties of the simple elliptic groupoid associated to . Let be the defining function of the -family as described in Equation (1). Recall that the -th moduli algebra is defined by the formal local ring modulo the ideal
In this context, a morphism of is expressed as an automorphism in such that .
Lemma 7.
-
(1)
The groupoid is a sub-groupoid of for each .
-
(2)
When is sufficiently large, we have
for .
-
(3)
The groupoid is a sub-groupoid of .
Proof.
(1) The first assertion is deduced from the fact that the isomorphism of the coordinate ring induces weighted homogeneous isomorphisms of the -th moduli algebras.
(2) For sufficiently large , we can assume that the weights of the generators in exceed the weight of . Consequently, the homogeneous polynomial , as a generator of the defining ideal of the algebra , has the minimal weight among the generators. Therefore, each morphism in yields a transformation of which satisfies the equation
for some constant . This relationship yields that is essentially a weighted homogeneous isomorphism between -moduli algebras.
(3) When , we have for any weighted homogeneous polynomial . Then
Note that a homogeneous isomorphism preserves the maximal ideal . Therefore, a homogeneous isomorphism is also an isomorphism in the context of the first moduli algebras. ∎
3. Groupoid of the -family
In this section, we focus on the isomorphisms of the -family. Recall that the -family is defined by the polynomial
which is weighted homogeneous of type . By definition, the -th moduli algebra takes the form
where
Our goal is to determine the isomorphism classes of for various values of . Since the monomials are both have weight and is the unique monomial of weight , any homogeneous automorphism of -families can be expressed as
in accordance with the notation in (4). Up to scaling, we can always assume that . The possible generators with weight less than are and (when ). From the fact
we have
This implies that .
Similarly, we have
and the possible generators in with weight are and .
This implies that . In conclusion, we find that the automorphism is represented by a diagonal matrix.
3.1. Case
The generator in with weight less than or equal to are , (for ), and , (for ). From this, we obtain the relation
This simplifies to:
Since , we find that . Hence, we have
This completes the proof of the case in Theorem C.
3.2. Case
From relations
we obtain
Therefore, we have
For , we obtain
Note that in this case, there is no essential restriction on , allowing to be any arbitrary element in . For , then we obtain and there are no essential restrictions on and .
3.3. Case
This case is essentially the same with by replacing with . This completes the proof of Theorem C.
4. Groupoid of the -family
Now we turn our attention to the study of simple elliptic singularity . For the parameter associated with the -family, the isolated hypersurface singularity is defined by the polynomial
Note that the Jacobi ideal of is given by
We have
and
For , it is easy to see that
Hence, . This implies that for ,
Applying (2) in Lemma 7 it implies that for ,
(8) |
In order to prove Theorem A, we need to study the first Yau Algebra of the -family, defined as
By direct calculation, we find that for the case , the algebra is -dimensional, with a -linear basis represented as
One may choose the basis of :
with multiplication rules:
Then the set
where , forms a basis of (see Section 4.3 for another complete list of bases). In the case of jump points, i.e., , the dimension of equals .
Through direct computation, we find that each matrix in Theorem A is a morphism in for and induces an isomorphism of the first Yau algebra. Based on this observation, the assertions (1) and (2) of Theorem A can be divided into the three assertions of the following Lemma.
Lemma 8.
-
(1)
The action groupoid is a sub-groupoid of . In particular, becomes a fully faithful sub-groupoid of when .
-
(2)
When , the equality
holds for .
-
(3)
When , each morphism in is generated by .
Proof of the first assertion.
It is straightforward to verify that the group induces morphisms of , and that the objects are stable under .
For , the morphisms in preserve the cubic polynomial
Therefore, is represented by scaling matrices (equivalent to in ) and permutations of three letters. By composing with , it becomes apparent that for is generated by . ∎
4.1. The second assertion
Notation 9.
-
(1)
For an element in of weight one, we introduce the symbol
where are coefficients satisfying . Additionally, we will utilize the notation in reference to alternative coordinates .
-
(2)
For linear subspace , we denote by the vector space spanned by all elements where , and . For a vector , we also define in the obvious manner.
To prove the second assertion of Lemma 8, we need the following technique lemma characterizing the relation between and .
Lemma 10.
Assume that is contained in with and is an element of weight one in .
-
(1)
The quantity
is independent on the choice of (but depends on ).
-
(2)
The transpose of the matrix is contained in .
Proof.
(1) Denote by the Lie bracket of . Let be the -linear space with linear basis .
For an element in of weight one, we define . Then
We obtain a linear subspace of parameterized by :
For arbitrary with , we have
(9) |
Assume that is represented by the matrix (5) with inverse (6). Let us investigate the images of quantities , , and under the isomorphism .
-
–
For the image of , we obtain
(10) -
–
For , we have
-
–
For , we compute the Lie bracket of and to obtain
(11) -
–
If , then
(12)
Note that preserves the Lie brackets. Combining Equations (9) and (12), we find
Thus, for any element in of weight one, we obtain the equality
(13) |
So the factor
is invariant regardless of the selection of . Consequently, the first assertion is validated.
(2) From the first assertion, any element in of weight one satisfies
(14) |
for some constant . From the definition and the explicit expansion of the term , we see that the transformation represents a morphism in . The equality (10) can be rewritten as
where
It means that is represented as the transpose matrix of , and thus the lemma holds. ∎
4.2. The third assertion
Lemma 11.
Proof.
We maintain the notations of Lemma 10. If we select , then . Define and . Then,
This implies that we identify a -dimensional subspace parameterized by . By computing the Lie bracket of and , we obtain the subspace
Therefore, leads to a decomposition of as follows:
In other words, we generate the vector space through the influence of . Thus, we achieve a decomposition of relative to , given by
(16) |
It yields that (not just the linear space ) is determined by .
We turn to study the corresponding decomposition of and image of under . By definition, we have
Then the image of equals
Evidently,
Since , it follows that
Hence, we have
Recalling the computation of as obtained in Equation (11):
we arrive at the decomposition
Thus, the decomposition of is given by
for some . Comparing this with (16), we see the image of is given by
Notice that
Lemma 12.
Assume that . Let be a homogeneous isomorphism contained in . Denote by the entries in the first row of the representation matrix (5) of . Then only one the following two cases may occur:
-
(1)
If all elements are nonzero, then ;
-
(2)
If at least two elements of are equal to zero.
Proof.
(1) Firstly, we assume that are nonzero. Let be a morphism contained in . Lemma 10 implies that the transpose matrix of represents a morphism in . Analogous to the equalities (15a)-(15c), we deduce the equations
(17a) | |||
(17b) | |||
(17c) |
Combining these equations shows that the values with verify the equality
where denotes a constant of the form
Equivalently, we find that are roots of defined as
(18) |
Since the function can have at most two distinct roots, at least two of the values must coincide. If our lemma does not hold, then we can assume without loss of generality that . From the expression in (18), we observe that both and are nonzero, which implies that
This can be simplified to the relation
(19) |
We now consider another morphism in represented by the matrix
It is clear that . The arguments above applied to lead us to the expression
(20) |
which parallels (19). Combining Equations (19) with (20) yields that
leading to
which contradicts the assumption . So we obtain .
(2) In the second case, if at least one of the values equals zero, we can set without loss of generality. The equations (15a)-(15c) imply:
Assuming both and are nonzero, we obtain are also nonzero. Applying the result from the first case to the transpose matrix of , we get
(21) |
Now Equation implies
It follows that . Combining this with Equations (21), we have . This contradicts our initial assumption about . Therefore, the assumption about and must be false, ensuring that either or is zero. In conclusion, the second case of the lemma is validated. ∎
We are in the position to give a complete description for .
Proof of the third assertion of Lemma 8.
We are employing the notations of Lemma 12.
Case (1): There exists some index such that .
For instance, if we set , applying Lemma 12 reveals that either or is zero. For the case where , Equations (15a)(15b)(15c) can be rewritten as
Combining these equalities with the original constraints
and
derided from the definition, we have . Upon extending the above arguments to and respectively, we conclude that is a matrix of types \@slowromancapi@ or \@slowromancapii@, where
If we also consider the cases and , then we obtain four additional types:
The first assertion of Lemma 10 states that
is invariant for any of weight one. By choosing respectively, we get the equality
(22) |
Substituting the values of from the six types \@slowromancapi@-\@slowromancapvi@, we conclude that
in each type respectively. There are exactly matrices contained in these six types up to scalar multiplication and each matrix is generated by .
Case (2): Suppose that are nonzero.
From Case (1), we know that all entries of are nonzero. Applying Lemma 12 to the sets , , and respectively, we conclude that all entries of are contained in the set .
From Equation (22), we get
(23) |
By considering the transpose of , we deduced from Equation (23) that
(24) |
Fix . Computational checks reveal that only non-singular matrices verify Equations (23) and (24). Up to scalar multiplication, there are exactly matrices. One may check directly that each matrix is generated by . ∎
4.3. Continue family of representations of the first Yau algebra
In this section, we prove assertion (3) of Theorem A. Following the approach in [20], we construct a basis in which the coefficients of the Lie bracket are independent of . For the entirety of this section, we assume that .
Recalling that the basis of is given by
We choose an alternative basis for as follows.
-
(1)
Degree 0:
; -
(2)
Degree 1:
, , , , , , , , , , ,
, , ,
, , , ; -
(3)
Degree 2:
, , .
Observe that is a graded Lie algebra with the decomposition:
where ’s are generated by homogeneous derivations of degree respectively. Since there are no derivations of degree three, we have . The subspace is generated by the Euler derivation . We obtain for and for . Next, we define the subspaces , , and . This leads to the decomposition:
One can check that
Furthermore, the matrix representation for Lie bracket is given by
and the representation for is given by
This implies that all coefficients of Lie bracket under our basis are constant, indicating that the first Yau algebra first Yau algebra does not depend on the parameter . We have finished the proof of Theorem A.
5. Groupoid of the -family
5.1. Outline of the proof of Theorem B
Recall that the -family is defined as the zero locus of polynomial
Firstly, we can directly verify that each matrix in Theorem B induces an isomorphism from to , where and . For example, let be the morphism represented as with . In terms of transformations, we can express this as:
Restricting to the coordinates , we get
by the definition of in (2). A direct computation shows
where
It follows that is a morphism in groupoid .
From [15], we know that for , we encounter the jump point case, in which the zeroth Yau algebra is -dimensional. In contrast, for the cases where , is -dimensional.
The proof for of Theorem B follows the same argument as in the assertion (1) of Lemma 8. Therefore, we omit the details.
When , we conclude from Lemma 7 that
for . So the assertion (2) in Theorem B is concluded by the assertion (1).
It remains to prove the case and . We will compute the groupoid by restricting to the coordinates . In fact, the complete collection of isomorphism with respect to are exactly the group , which is also described in [12] without proof. We utilize the Torelli-type theorem from [15] to show that the isomorphisms of the -family preserve the four lines in the zeroth Yau algebra. In this sense, the group is indeed the symmetric group acting on the four lines.
For the groupoid , we need to prove the assertion (3) in Theorem B, namely the Torelli-type theorem of the first Yau algebra, which implies that the morphism of groupoid are derived from . As a consequence of the Torelli-type theorem, we can reduce the groupoid to , which completes the proof of Theorem B.
5.2. The zeroth Yau algebra
We assume that . The moduli algebra is written as
where, as usual,
Notice that for , we have
and similarly
The monomials
form a basis of with multiplication rule
From [15], we know that the jump points of the -family are , in which case the zeroth Yau algebra is -dimensional. In contrast, for the case , the Lie algebra is 11-dimensional and there exists a basis for :
-
–
Degree 0:
; -
–
Degree 1:
, , ,
; -
–
Degree 2:
, , , ; -
–
Degree 3:
, .
Choose a solution of the equation
(25) |
We define (without orderings) to be the four lines on the plane whose slopes are exactly the solutions of equation (25), specifically:
The set of these four lines is shown to be invariant under homogeneous isomorphisms. These lines have a cross-ratio, relative to a choice of orderings. Considering the symmetric group acting on the four lines, the orderings provide six different values of the cross-ratio:
which constitute the six continuous invariants of . The set of the six continuous invariants is equivalent to the -function . In addition, the transformations preserve and leaves the six continuous invariants unchanged generate the Klein four-group. We adopt the approach that the morphisms in induce linear maps on the plane . In terms of matrices, the morphisms in essentially derive the action of on the four lines when restricted to the coordinates and .
Lemma 13.
The group acts on the set in one-to-one correspondence with the action of symmetric group . Specifically, we have:
for , and contained in some domain of . In particular, the subgroup acts on as the Klein four-group.
Proof.
We only check for the isomorphism expressed as the matrix acting on . Notice that
Following the notation in (7) we have
and
Then the line has a direction vector
for some constant . So the slope of is given by
Therefore, we obtain
Since , we can express
which coincides with Equation (25). It yields that is one of the conjugate elements of , and therefore the image of is just one of the lines . ∎
Now we are in the position to prove the case in the assertion (1) of Theorem B.
Partial proof of Theorem B.
It suffices to show that the restriction of the morphism of to are contained in the group . Given a morphism of , we know that induces a morphism on the zero Yau algebra, which preserves the set of four lines (not necessary in order). By composing with some matrix in if necessary, we can assume the preserves the set and the -function. We have argued that only the transformations arising from the Klein four-group leaves the -function invariants unchanged. According to Lemma 13, the Klein four-group are presented by with respect to coordinates . Thus, we conclude that the transformations on preserving the set of the four lines generate the group , which establishes the case of Theorem B. ∎
5.3. The first Yau algebra of the -family
The first moduli algebra of the -family is given by
where
For , we construct the -linear basis of given by :
-
–
degree 0:
and ; -
–
degree 1:
, , , , , , ,, , ; -
–
degree 2:
, , , ,
, , , , ; -
–
degree 3:
and .
We denote by the Lie bracket of and use the notations and as in Notation 9. We aim to construct four distinct lines in a canonical order from the Lie structure of .
Lemma 14.
There exists a sequence of four lines in which is invariant up to homogeneous isomorphisms.
Proof.
Denote by the weighted homogeneous part of of degree . The existence of the four lines can be verified through the following construction procedure:
-
(1)
The subspace of .
-
(2)
The decomposition of :
(26) where , and and .
-
(3)
The -dimension subspace of generated by
(27) and
(28) -
(4)
The -dimensional subspace of generated by and with property .
-
(5)
The line of with director vector
-
(6)
The -dimensional subspace containing with basis and
(29) which verifies the property .
-
(7)
The line , being the intersection of and .
-
(8)
The complementary space of in , and .
-
(9)
The complementary space of in , and .
The detailed constructs are given as follows:
(1) By restricting the Lie bracket to
we find that the vectors subject to the property
are contained in . Therefore, is an invariant linear subspace.
(2) The vector induces the linear map:
with eigenvalues . Hence, there exists the Jordan decomposition (26), where the eigenspace correspond to the eigenvalues .
(3) The subspace is defined as the collection of vectors such that
We can verify that is spanned by in Equations (28) and (27).
(4) We define to be the collection of such that there exists some , linearly independent on , subject to
Equivalently, for such , the kernel of is at least -dimensional. It is clearly that is invariant up to isomorphism. Notice that , so contains the linear subspace generated by . The equality can be checked by observing that the two bilinear maps
and
are non-degenerate.
(5) We define as the collection of such that . A direct computation shows that is -dimensional with basis .
(6) It can be checked that
where are defined in (29). Therefore, the space spanned by is the minimal -dimensional space such that
By construction, is also invariant.
(7) It is straightforward to show that
is -dimensional with basis
(8) The space is an invariant subspace of with basis as defined in (28). Then the linear subspace of is spanned by
(9) Similarly, the space is an invariant subspace of . The intersection of and is spanned by the vector
The line spanned by is obviously invariant. ∎
With the aid of these four invariant lines, we can establish the Torelli-type theorem for .
Proof of part (3) in Theorem B.
From Lemma 14, the sequence of lines is invariant under homogeneous isomorphisms. Moreover, the cross ratio of is given by
which aligns with the -function . Therefore, the first Yau algebra determines the isomorphism classes of the -family. ∎
Finally, we would like to utilize the invariance of to compute the groupoid for the case of the assertion (1) in Theorem B.
Complete the proof of Theorem B.
It suffices to demonstrate that, when restricting to the coordinates , the groupoid coincides with . Specifically, the restriction of a morphism of to induces a homogeneous homomorphism of . Notice that this, together with Lemma 7, implies that are the same up to transformations along the -axis.
Let be a morphism in , represented as the matrix
in the coordinates . Let be the determinant. Then the inverse of is given by
Then, the partial derivatives can be expressed as
and
The image of the director vector of under is given by
Since the line is invariant under , we may assume the equality
holds for some constant . This is equivalent to
and
Therefore, we obtain:
It follows that comes from . Therefore, the case of the assertion (1) in Theorem B is proved. ∎
References
- [1] V. I. Arnold, A. N. Varchenko, and S. M. Guseĭn-Zade, “Singularities of differentiable maps,” “Nauka”, Moscow, p. 336, 1984.
- [2] C. Hu, S. S.-T. Yau, and H. Zuo, “On the -th tjurina number of weighted homogeneous singularities,” arXiv preprint arXiv:2409.09384, 2024.
- [3] G.-M. Greuel, C. Lossen, and E. Shustin, “Introduction to singularities and deformations,” Springer, 2006.
- [4] L. D. Tráng and C. P. Ramanujam, “The invariance of milnor’s number implies the invariance of the topological type,” American Journal of Mathematics, vol. 98, no. 1, pp. 67–78, 1976.
- [5] J. Milnor and P. Orlik, “Isolated singularities defined by weighted homogeneous polynomials,” Topology, vol. 9, no. 4, pp. 385–393, 1970.
- [6] N. Hussain, Z. Liu, S. S.-T. Yau, and H. Zuo, “k-th milnor numbers and k-th tjurina numbers of weighted homogeneous singularities,” Geometriae Dedicata, vol. 217, no. 2, p. 34, 2023.
- [7] J. N. Mather and S. S.-T. Yau, “Classification of isolated hypersurface singularities by their moduli algebras,” Inventiones Mathematicae, vol. 69, no. 2, pp. 243–251, 1982.
- [8] G.-M. Greuel and T. H. Pham, “Mather-yau theorem in positive characteristic,” Journal of Algebraic Geometry, vol. 26, pp. 347–355, 2017.
- [9] K. Saito, “Einfach-elliptische singularitäten,” Inventiones Mathematicae, vol. 23, no. 3-4, pp. 289–325, 1974.
- [10] M. Noumi and Y. Yamada, “Notes on the flat structures associated with simple and simply elliptic singularities,” Integrable systems and algebraic geometry (Kobe/Kyoto, 1997), pp. 373–383, 1997.
- [11] I. A. B. Strachan, “Simple elliptic singularities: a note on their g-function,” Asian Journal of Mathematics, vol. 16, no. 3, pp. 409–426, 2010.
- [12] M. G. Eastwood, “Moduli of isolated hypersurface singularities,” Asian Journal of Mathematics, vol. 8, no. 2, pp. 305–314, 2004.
- [13] P. J. Higgins, Categories and groupoids. Citeseer, 1971.
- [14] R. Brown, “Topology and groupoids,” 2006.
- [15] C. Seeley and S. S.-T. Yau, “Variation of complex structures and variation of Lie algebras,” Inventiones Mathematicae, vol. 99, no. 1, pp. 545–565, 1990.
- [16] N. Hussain, S. S.-T. Yau, and H. Zuo, “New -th yau algebras of isolated hypersurface singularities and weak torelli-type theorem,” Mathematical Research Letters, vol. 29, no. 2, pp. 455–486, 2022.
- [17] H. Chen, C. Seeley, and S. S.-T. Yau, “Algebraic determination of isomorphism classes of the moduli algebras of singularities,” Mathematische Annalen, vol. 318, no. 4, pp. 637–666, 2000.
- [18] M. Benson and S. S.-T. Yau, “Lie algebra and their representations arising from isolated singularities: Computer method in calculating the lie algebras and their cohomology,” Advance studies in Pure Mathematics, vol. 8, pp. 3–58, 1986.
- [19] A. Elashvili and G. Khimshiashvili, “Lie algebras of simple hypersurface singularities,” Journal of Lie Theory, vol. 4, no. 4, pp. 621–650, 2006.
- [20] S. S.-T. Yau, “Continuous family of finite-dimensional representations of a solvable Lie algebra arising from singularities,” Proceedings of the National Academy of Sciences of the United States of America, vol. 80, no. 24, pp. 7694–7696, 1983.