This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

groupoids derived from the simple elliptic singularities

Chuangqiang Hu Stephen S.-T. Yau  and  Huaiqing Zuo
Abstract.

K. Saito’s classification of simple elliptic singularities includes three families of weighted homogeneous singularities: E~6,E~7\tilde{E}_{6},\tilde{E}_{7}, and E~8\tilde{E}_{8}. For each family, the isomorphism classes can be distinguished by K. Saito’s jj-functions. By applying the Mather-Yau theorem, which states that the isomorphism class of an isolated hypersurface singularity is completely determined by its kk-th moduli algebra, M. Eastwood demonstrated explicitly that one can directly recover K. Saito’s jj-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 jj-functions from the kk-th moduli algebras but also elucidate the automorphism structure within the kk-th moduli algebras. We derive the automorphisms using the methodology of the kk-th Yau algebra and establish a Torelli-type theorem for the E~7\tilde{E}_{7}-family when k=1k=1. In contrast, we find that the Torelli-type theorem is inapplicable for the first Yau algebra in the E~6\tilde{E}_{6}-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

Beijing Institute of Mathematical Sciences and Applications, Beijing 101408, P. R. China
E-mail: [email protected]
Corresponding author: Stephen S.-T. Yau
Beijing Institute of Mathematical Sciences and Applications, Beijing 101408, P. R. China
E-mail: [email protected]
Department of Mathematical Sciences, Tsinghua University, Beijing 100084, P. R. China
E-mail: [email protected]
Zuo was supported by NSFC Grant 12271280. Yau was supported by Tsinghua University start-up fund and Tsinghua University Education Foundation fund (042202008).
Data availability: Data sharing not applicable to this article as no datasets were generated or analysed during the current study.

1. Introduction

We present necessary definitions and auxiliary results concerning the moduli algebras of singularities. Let 𝒪n:=[[x1,,xn]]\mathcal{O}_{n}:=\mathbb{C}[[x_{1},\ldots,x_{n}]] be a formal power series ring over \mathbb{C} with maximal ideal mm. Let 𝒱f\mathcal{V}_{f} be a germ of an isolated hypersurface singularity at the origin in n\mathbb{C}^{n} represented as the zero locus of polynomial (or analytic function) f=f(x1,,xn)f=f(x_{1},\ldots,x_{n}).

Definition 1.

The kk-th moduli algebra of 𝒱f\mathcal{V}_{f} is defined by

𝒜k(𝒱f)=𝒪n/f,mkJ(f),\mathcal{A}^{k}(\mathcal{V}_{f})=\mathcal{O}_{n}/\langle f,m^{k}J(f)\rangle,

where J(f)=1(f),,n(f)J(f)=\langle\partial_{1}(f),\ldots,\partial_{n}(f)\rangle denotes the Jacobi ideal of ff.

It is convenient to denote the local function algebra (or coordinate ring) of 𝒱f\mathcal{V}_{f} by

𝒜(𝒱f)=𝒪n/f.\mathcal{A}^{\infty}(\mathcal{V}_{f})=\mathcal{O}_{n}/\langle f\rangle.

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 kk-th Tjurina algebra of weighted homogeneous singularities. For further related studies on on Tjurina algebra, we refer to [3, 4, 5, 6].

The kk-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 fif_{i} for i=0,1i=0,1 denote analytic functions associated with isolated singularities 𝒱fi\mathcal{V}_{f_{i}} respectively. Then the following conditions are equivalent:

  1. (1)

    The function f1f_{1} is contact equivalent to f2f_{2}, i.e., there exists some automorphism ϕAut(𝒪n)\phi\in\operatorname{Aut}(\mathcal{O}_{n}) and a unit u𝒪nu\in\mathcal{O}_{n}^{*} such that

    ϕ(f1)=uf2.\phi(f_{1})=u\cdot f_{2}.
  2. (2)

    For all k0k\geqslant 0, there exists some isomorphism 𝒜k(𝒱f1)𝒜k(𝒱f2)\mathcal{A}^{k}(\mathcal{V}_{f_{1}})\cong\mathcal{A}^{k}(\mathcal{V}_{f_{2}}) of \mathbb{C}-algebras.

  3. (3)

    There exists some k0k\geqslant 0 such that 𝒜k(𝒱f1)𝒜k(𝒱f2)\mathcal{A}^{k}(\mathcal{V}_{f_{1}})\cong\mathcal{A}^{k}(\mathcal{V}_{f_{2}}) as \mathbb{C}-algebras.

Among all isolated hypersurface singularities, weighted homogeneous singularities have been of particular interest. Recall that a polynomial f(x1,,xn)f(x_{1},\cdots,x_{n}) is weighted homogeneous of type (w1,w2,,wn)(w_{1},w_{2},\cdots,w_{n}), where w1,w2,,wnw_{1},w_{2},\cdots,w_{n} are fixed positive rational numbers, if it can be expressed as a linear combination of monomials x1i1x2i2xninx_{1}^{i_{1}}x_{2}^{i_{2}}\cdots x_{n}^{i_{n}} for which

i1w1+i2w2++inwn=Wi_{1}w_{1}+i_{2}w_{2}+\cdots+i_{n}w_{n}=W

for some constant WW.

We view 𝒪n\mathcal{O}_{n} as a graded algebra by assigning the weight of xix_{i} to be wiw_{i}. Let I1,I2I_{1},I_{2} be two homogeneous ideals. For the local algebra 𝒪n/Ii\mathcal{O}_{n}/I_{i} with i=0,1i=0,1, we say that homomorphism ϕ:𝒪n/I1𝒪n/I2\phi:\mathcal{O}_{n}/I_{1}\to\mathcal{O}_{n}/I_{2} is homogeneous if ϕ\phi preserves the weights of 𝒪n\mathcal{O}_{n}. For the case where both f1f_{1} and f2f_{2} are weighted homogeneous, a contact equivalence (ϕ,u)(\phi,u) 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:

  1. (4)

    (Suppose that f1f_{1} and f2f_{2} are weighted homogeneous.) There exists some homogeneous automorphism ϕAut(𝒪n)\phi\in\operatorname{Aut}(\mathcal{O}_{n}) and a constant λ\lambda\in\mathbb{C}^{*} such that

    ϕ(f1)=λf2.\phi(f_{1})=\lambda\cdot f_{2}.

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) E~6:\displaystyle\tilde{E}_{6}: x3+y3+z3+txyz=0,for t(E~6);\displaystyle x^{3}+y^{3}+z^{3}+txyz=0,\quad\text{for $t\in\mathbb{C}(\tilde{E}_{6})$};
E~7:\displaystyle\tilde{E}_{7}: x4+y4+z2+tx2y2=0,for t(E~7);\displaystyle x^{4}+y^{4}+z^{2}+tx^{2}y^{2}=0,\quad\text{for $t\in\mathbb{C}(\tilde{E}_{7})$};
E~8:\displaystyle\tilde{E}_{8}: x6+y3+z2+tx4y=0,for t(E~8);\displaystyle x^{6}+y^{3}+z^{2}+tx^{4}y=0,\quad\text{for $t\in\mathbb{C}(\tilde{E}_{8})$};

where the restrictions on tt ensure the singularities are isolated:

E~6:t(E~6)\displaystyle\tilde{E}_{6}:t\in\mathbb{C}(\tilde{E}_{6}) ={t|t327};\displaystyle=\{t\in\mathbb{C}|t^{3}\not=27\};
E~7:t(E~7)\displaystyle\tilde{E}_{7}:t\in\mathbb{C}(\tilde{E}_{7}) ={t|t24};\displaystyle=\{t\in\mathbb{C}|t^{2}\not=4\};
E~8:t(E~8)\displaystyle\tilde{E}_{8}:t\in\mathbb{C}(\tilde{E}_{8}) ={t|4t327}.\displaystyle=\{t\in\mathbb{C}|4t^{3}\not=-27\}.

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 GG-functions for these families. Saito determined the isomorphism classes of these families by computing the jj-functions and concluded the criteria that two singular with coefficients tt and ss are contact equivalent if and only if j(E~i;t)=j(E~i;s)j(\tilde{E}_{i};t)=j(\tilde{E}_{i};s), where

j(E~6;t)\displaystyle j(\tilde{E}_{6};t) =t3(t3216)1728(t3+27)3;\displaystyle=-\frac{t^{3}(t^{3}-216)}{1728(t^{3}+27)^{3}};
j(E~7;t)\displaystyle j(\tilde{E}_{7};t) =(12+t2)3108(t24)2;\displaystyle=\frac{(12+t^{2})^{3}}{108(t^{2}-4)^{2}};
j(E~8;t)\displaystyle j(\tilde{E}_{8};t) =4t34t3+27.\displaystyle=\frac{4t^{3}}{4t^{3}+27}.

For i=6,7,8i=6,7,8, we denote by 𝒜k(E~i;t)\mathcal{A}^{k}(\tilde{E}_{i};t) with t(E~i)t\in\mathbb{C}(\tilde{E}_{i}) the corresponding kk-th moduli algebra (resp. local function algebra) of E~i\tilde{E}_{i}-families listed above. With the help of the Mather-Yau theorem, Eastwood [12] demonstrated explicitly that one can recover directly Saito’s jj-functions from the zeroth moduli algebra 𝒜0(E~i;t)\mathcal{A}^{0}(\tilde{E}_{i};t).

In this paper, we aim to recover Saito’s jj-functions from the kk-th moduli algebra 𝒜k(E~i;t)\mathcal{A}^{k}(\tilde{E}_{i};t) with k1k\geqslant 1 by applying the kk-th version of the Mather-Yau theorem. Furthermore, we pose a finer question: how can we completely describe the homogeneous isomorphisms from 𝒜k(E~i;t)\mathcal{A}^{k}(\tilde{E}_{i};t) to 𝒜k(E~i;s)\mathcal{A}^{k}(\tilde{E}_{i};s)? To understand the algebraic structure in this context, we introduce the groupoid Grpk(Ei)\operatorname{Grp}^{k}(E_{i}) of simple elliptic singularity. Our complete description of the groupoid Grpk(Ei)\operatorname{Grp}^{k}(E_{i}) 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 GG acts on the set XX, then we can form the action groupoid Grpac(X|g1,,gn)\operatorname{Grp}^{\operatorname{ac}}(X|g_{1},\ldots,g_{n}) (or transformation groupoid) as follows.

  1. (1)

    The objects are the elements of XX.

  2. (2)

    For any two elements x,yXx,y\in X, the morphisms from xx to yy correspond to the elements gGg\in G such that gx=ygx=y.

  3. (3)

    Composition of morphisms interprets the binary operation of GG.

Definition 4.

For i=6,7,8i=6,7,8 and k0{}k\in\mathbb{Z}_{\geqslant 0}\cup\{\infty\}, we define the groupoid Grpk(E~i)\operatorname{Grp}^{k}(\tilde{E}_{i}) of simple elliptic singularities as follows.

  1. (1)

    The objects of Grpk(E~i)\operatorname{Grp}^{k}(\tilde{E}_{i}) are precisely the coefficients t(E~i)t\in\mathbb{C}(\tilde{E}_{i}).

  2. (2)

    For t,s(E~i)t,s\in\mathbb{C}(\tilde{E}_{i}), the morphisms of tt to ss are the homogeneous isomorphisms from 𝒜k(E~i;t)\mathcal{A}^{k}(\tilde{E}_{i};t) to 𝒜k(E~i;s)\mathcal{A}^{k}(\tilde{E}_{i};s) modulo the scalars in \mathbb{C}^{*}.

This paper devotes to the explicit computation of the groupoids Grpk(E~i)\operatorname{Grp}^{k}(\tilde{E}_{i}). From the lifting property, any morphism of Grpk(E~i)\operatorname{Grp}^{k}(\tilde{E}_{i}) lifts to an automorphism of weighted projective spaces. Up to the \mathbb{C}^{*}-action, the morphism is essentially contained in the weighted projective general linear group PGL3𝐰()\operatorname{PGL}_{3}^{\mathbf{w}}(\mathbb{C}), which is the quotient of the general linear group GL3()\operatorname{GL}_{3}(\mathbb{C}) by its subgroup of diagonal matrices

(λw1000λw2000λw3)\begin{pmatrix}\lambda^{w_{1}}&0&0\\ 0&\lambda^{w_{2}}&0\\ 0&0&\lambda^{w_{3}}\end{pmatrix}

for some weights (w1,w2,w3)=:𝐰(w_{1},w_{2},w_{3})=:\mathbf{w}. When kk is large, we obtain

Grpk(E~i)=Grp(E~i),\operatorname{Grp}^{k}(\tilde{E}_{i})=\operatorname{Grp}^{\infty}(\tilde{E}_{i}),

since the morphism in Grpk(E~i)\operatorname{Grp}^{k}(\tilde{E}_{i}) preserves the defining functions (see Lemma 7).

For k=0k=0 or 11, Grpk(E~i)\operatorname{Grp}^{k}(\tilde{E}_{i}) contains more morphisms due to the fact that fmJ(f)f\in mJ(f) when ff is weighted homogeneous. Finally, we also note that there exist some coefficients t(E~i)t\in\mathbb{C}(\tilde{E}_{i}) that behave weirdly, referred to as jump points according to references[15, 16]. The jump points are listed as follows.

  1. (1)

    For the E~6\tilde{E}_{6}-family, t=0,6,6ρ,6ρ2t=0,6,6\rho,6\rho^{2};

  2. (2)

    For the E~7\tilde{E}_{7}-family, t=0,±6t=0,\pm 6;

  3. (3)

    For the E~8\tilde{E}_{8}-family, t=0t=0.

Here and afterwards, ρ\rho denotes a primitive cubic root of unity. In the following subsections, we will formulate the main results in detail.

1.1. Groupoid of the E~6\tilde{E}_{6}-family

In [17], Chen, Seeley and Yau provided a detailed characterization of homogeneous isomorphisms of the moduli algebras arising from the E~6\tilde{E}_{6}-family, i.e., the morphisms in Grp0(E~6)\operatorname{Grp}^{0}(\tilde{E}_{6}). Our newly elaborated technique for studying the isomorphisms of Grp1(E~6)\operatorname{Grp}^{1}(\tilde{E}_{6}) 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 {Lt}\{L_{t}\} associated to the zeroth moduli algebras of E~7\tilde{E}_{7} and E~8\tilde{E}_{8}-families. They proved the Torelli-type theorem saying that the Lie algebras LsL_{s} and LtL_{t} 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 E~6\tilde{E}_{6}-family, we consider the derivation Lie algebra, written as L1(E~6;t)L^{1}(\tilde{E}_{6};t), associated with the first moduli algebra 𝒜1(E~6;t)\mathcal{A}^{1}(\tilde{E}_{6};t). By equipping L1(E~6;t)L^{1}(\tilde{E}_{6};t) with a natural module structure over the moduli algebra, we discover that the invariants arising from the Lie algebra L1(E~6;t)L^{1}(\tilde{E}_{6};t), together with 𝒜(E~6;t)\mathcal{A}(\tilde{E}_{6};t), are applicable in determining the isomorphisms of the E~6\tilde{E}_{6}-family. It turns out that the groupoids Grp0(E~6)\operatorname{Grp}^{0}(\tilde{E}_{6}) and Grp1(E~6)\operatorname{Grp}^{1}(\tilde{E}_{6}) coincide indicating a simpler proof for Chen-Seeley-Yau’s result in [17]. However, we note that the Lie algebra L1(E~6;t)L^{1}(\tilde{E}_{6};t) itself (without the module structure) is insufficient to determine the isomorphism classes of the E~6\tilde{E}_{6}-family. In other words, there is no Torelli-type theorem for L1(E~6;t)L^{1}(\tilde{E}_{6};t) 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 GG of PGL3()\operatorname{PGL}_{3}(\mathbb{C}) generated by the matrices

A1:=(ρ00010001),A2:=(ρρ21ρ2ρ1111),A3:=(010100001),A4:=(100001010).\displaystyle A_{1}:=\begin{pmatrix}\rho&0&0\\ 0&1&0\\ 0&0&1\end{pmatrix},A_{2}:=\begin{pmatrix}\rho&\rho^{2}&1\\ \rho^{2}&\rho&1\\ 1&1&1\end{pmatrix},A_{3}:=\begin{pmatrix}0&1&0\\ 1&0&0\\ 0&0&1\end{pmatrix},A_{4}:=\begin{pmatrix}1&0&0\\ 0&0&1\\ 0&1&0\end{pmatrix}.

Note that A3A_{3} and A4A_{4} generates the symmetric subgroup on three letters. One can show by direct computation that GG contains totally 216216 matrices and all the entries belong to {0,ρ,ρ2,1}\{0,\rho,\rho^{2},1\}. Let HH be the group of fractional linear transformations generated by

P:tρtP:t\mapsto\rho t

and

R:t183t3+t.R:t\mapsto\frac{18-3t}{3+t}.

It can be checked that HH is a group of order 1212, isomorphic to the alternative group on four letters. In addition, there exists a group homomorphism π:GH\pi:G\to H which sends A1,A2,A3,A4A_{1},A_{2},A_{3},A_{4} to P,R,Id,IdP,R,\operatorname{Id},\operatorname{Id} respectively.

We restate and generalize the main result in [17] in the following form.

Theorem A.

(1) For the case k=0k=0 or 11, the groupoid Grpk(E~6)\operatorname{Grp}^{k}(\tilde{E}_{6}) restricted to the objects (E~6){0,6,6ρ,6ρ2}\mathbb{C}(\tilde{E}_{6})\setminus\{0,6,6\rho,6\rho^{2}\} is represented by the action groupoid

Grpac((E~6){0,6,6ρ,6ρ2}|G),\operatorname{Grp}^{\operatorname{ac}}(\mathbb{C}(\tilde{E}_{6})\setminus\{0,6,6\rho,6\rho^{2}\}\,|\,G),

where AiA_{i} acts on (E~6)\mathbb{C}(\tilde{E}_{6}) through the morphism π\pi defined above.

(2)For k2k\geqslant 2 or k=k=\infty, we find that Grpk(E~6)\operatorname{Grp}^{k}(\tilde{E}_{6}) equals the action groupoid Grpac((E~6)|G)\operatorname{Grp}^{\operatorname{ac}}(\mathbb{C}(\tilde{E}_{6})\,|\,G).

(3)(The Failure of a Torelli-Type Theorem) The first Yau algebra of the E~6\tilde{E}_{6}-family is independent on the parameter tt and gives rise to a continuous family of 1111-dimensional representations of a solvable Lie algebra.

We remark that for k=0k=0 or 11, Grpac((E~6)|G)\operatorname{Grp}^{\operatorname{ac}}(\mathbb{C}(\tilde{E}_{6})\,|\,G) is a subcategory of Grpk(E~6)\operatorname{Grp}^{k}(\tilde{E}_{6}), but it is not fully faithful. Indeed, it can be concluded that Grpk(E~6)(0,0)\operatorname{Grp}^{k}(\tilde{E}_{6})(0,0) is the group generated by A2,A3,A4A_{2},A_{3},A_{4} and the diagonal matrices. As a consequence of Theorem A, K. Saito’s jj-function of the E~6\tilde{E}_{6}-family can be obtained by observing that the integral functions on (E~6)\mathbb{C}(\tilde{E}_{6}) fixed by GG (or HH equivalently) are generated by j(E~6;t)j(\tilde{E}_{6};t).

1.2. Groupoid of the E~7\tilde{E}_{7}-family

Consider the group HH^{\prime} consisting of the following six fractional linear transformations

t±t,t±(122t2+t),t±(12+2t2t).t\mapsto\pm t,t\mapsto\pm\left(\frac{12-2t}{2+t}\right),t\mapsto\pm\left(\frac{12+2t}{2-t}\right).

Note that HH^{\prime} is isomorphic to the symmetric group S3S^{3}. According to [15], instead of jj-function j(E~7;t)j(\tilde{E}_{7};t), we have the criteria that the coefficient tt is equivalent to ss in Grpk(E~7)\operatorname{Grp}^{k}(\tilde{E}_{7}) if and only if s=h(t)s=h(t) for a unique hHh\in H^{\prime}. We define subgroup GG^{\prime} of PGL2(1,1)()\operatorname{PGL}_{2}^{(1,1)}(\mathbb{C}) generated by

(11ii),(100i),(0110).\begin{pmatrix}1&1\\ i&-i\end{pmatrix},\begin{pmatrix}1&0\\ 0&i\end{pmatrix},\begin{pmatrix}0&1\\ 1&0\end{pmatrix}.

This group is to shown to preserve the polynomial x4+y4x^{4}+y^{4} and is isomorphic to the symmetry group S4S^{4}. Thus, there are totally 2424 elements of GG^{\prime} which are represented by the following matrices

Bα,0:=(100α),B0,β:=(0β10),Bα,β:=(1βααβ),B_{\alpha,0}:=\begin{pmatrix}1&0\\ 0&\alpha\end{pmatrix},B_{0,\beta}:=\begin{pmatrix}0&\beta\\ 1&0\end{pmatrix},B_{\alpha,\beta}:=\begin{pmatrix}1&\beta\\ \alpha&-\alpha\beta\end{pmatrix},

where α\alpha and β\beta range over {±1,±i}\{\pm 1,\pm i\}.

Notice that S4S^{4} has a Klein four-group as a proper normal subgroup, namely the even transpositions Id,(12)(34),(13)(24),(14)(23){\operatorname{Id},(12)(34),(13)(24),(14)(23)}, with quotient S3S^{3}. In according to the mapping S4S3S^{4}\to S^{3}, there exists a homomorphism

π:GH,\pi^{\prime}:G^{\prime}\to H^{\prime},

defined as

(2) Bα,β\displaystyle B_{\alpha,\beta} [tβ2(122α2t)2+tα2],\displaystyle\mapsto\left[t\mapsto\frac{\beta^{2}(12-2\alpha^{2}t)}{2+t\alpha^{2}}\right],
Bα,0,B0,α\displaystyle B_{\alpha,0},B_{0,\alpha} [tα2t].\displaystyle\mapsto[t\mapsto\alpha^{2}t].

In particular, the kernel of π\pi^{\prime} gives a subgroup

G0={B1,0=Id,B1,0,B0,1,B0,1}G^{\prime}_{0}=\{B_{1,0}=\operatorname{Id},B_{-1,0},B_{0,1},B_{0,-1}\}

which is isomorphic to the Klein four-group. Using the notation

B~α,β,γ:=(Bα,β00γ)\tilde{B}_{\alpha,\beta,\gamma}:=\begin{pmatrix}B_{\alpha,\beta}&0\\ 0&\gamma\end{pmatrix}

for γ\gamma\in\mathbb{C}^{*}, we obtain the following result which parallels to Theorem A:

Theorem B.

(1) For k=0,1k=0,1, the groupoid Grpk(E~7)\operatorname{Grp}^{k}(\tilde{E}_{7}) restricted to E~7{0,±6}\tilde{E}_{7}\setminus\{0,\pm 6\} is identical to the action groupoid

Grpac((E~7){0,±6}|{B~α,β,γ,γ}).\operatorname{Grp}^{\operatorname{ac}}(\mathbb{C}(\tilde{E}_{7})\setminus\{0,\pm 6\}\,|\,\{\tilde{B}_{\alpha,\beta,\gamma},\gamma\in\mathbb{C}^{*}\}).

Here α\alpha and β\beta take values from {0,±1,±i}\{0,\pm 1,\pm i\}, and they are not simultaneously zero; and the matrices B~α,β,γ\tilde{B}_{\alpha,\beta,\gamma} act on the points of (E~7)\mathbb{C}(\tilde{E}_{7}) through π\pi^{\prime}, restricting to the (x,y)(x,y) coordinates.

(2)If k2k\geqslant 2 or k=k=\infty, then

Grpk(E~7)(t,s)={{Id,B~1,0,1,B~0,1,1,B~0,1,1}when s=t;{B~α,0,1,B~0,α,1}when s=α2t;{B~α,β,±2+α2t}when s=β2(122α2t)2+α2t.\operatorname{Grp}^{k}(\tilde{E}_{7})(t,s)=\begin{cases}\{\operatorname{Id},\tilde{B}_{-1,0,1},\tilde{B}_{0,1,1},\tilde{B}_{0,-1,1}\}&\text{when $s=t$};\\ \{\tilde{B}_{\alpha,0,1},\tilde{B}_{0,\alpha,1}\}&\text{when $s=\alpha^{2}t$};\\ \{\tilde{B}_{\alpha,\beta,\pm\sqrt{2+\alpha^{2}t}}\}&\text{when $s=\frac{\beta^{2}(12-2\alpha^{2}t)}{2+\alpha^{2}t}$}.\\ \end{cases}

(3) (Torelli-Type Theorem)For t(E~7){0,±6}t\in\mathbb{C}(\tilde{E}_{7})\setminus\{0,\pm 6\}, the first Yau algebra determines the isomorphism class of the E~7\tilde{E}_{7}-family.

Similar to the E~6\tilde{E}_{6}-family, the matrices of the form B~α,β,γ\tilde{B}_{\alpha,\beta,\gamma} generate a proper sub-groupoid of Grpk(E~7)\operatorname{Grp}^{k}(\tilde{E}_{7}) when k=0k=0 or 11. The integral functions on (E~7)\mathbb{C}(\tilde{E}_{7}) fixed by B~α,β,γ\tilde{B}_{\alpha,\beta,\gamma} (or HH^{\prime} equivalently) are generated by j(E~7;t)j(\tilde{E}_{7};t).

1.3. Groupoid of the E~8\tilde{E}_{8}-family

The criteria for the isomorphism class of the E~8\tilde{E}_{8}-family is much simpler. It follows from the jj-function j(E~8;t)j(\tilde{E}_{8};t) that tt is isomorphic to ss in Grpk(E~8)\operatorname{Grp}^{k}(\tilde{E}_{8}) if and only if s=ats=at with a3=1a^{3}=1. Let Cρi,γC_{\rho^{i},\gamma} be the matrix of the form

(1000ρi000γ)\begin{pmatrix}1&0&0\\ 0&\rho^{i}&0\\ 0&0&\gamma\end{pmatrix}

with γ\gamma\in\mathbb{C}^{*}. The collection of matrices {Cρi,γ}\{C_{\rho^{i},\gamma}\} with i=0,1,2i=0,1,2, cc\in\mathbb{C}^{*} forms a subgroup of PGL3𝐰()\operatorname{PGL}_{3}^{\mathbf{w}}(\mathbb{C}) with 𝐰=(1,2,3)\mathbf{w}=(1,2,3). For k=0k=0 or 11, the matrix Cρi,γC_{\rho^{i},\gamma} induces an isomorphism from 𝒜k(E~8,t)\mathcal{A}^{k}(\tilde{E}_{8},t) to 𝒜k(E~8,ρit)\mathcal{A}^{k}(\tilde{E}_{8},\rho^{i}t), represented by the transformation

x=x,y=ρiy,z=γz.x=x^{\prime},y=\rho^{i}y^{\prime},z=\gamma z^{\prime}.
Theorem C.

(1) For k=0k=0 or 11,the groupoid Grpk(E~8)\operatorname{Grp}^{k}(\tilde{E}_{8}) is given by the union of the action groupoid

Grpac((E~8){0}|{Cρi,γ|γ})\operatorname{Grp}^{\operatorname{ac}}(\mathbb{C}(\tilde{E}_{8})\setminus\{0\}\,|\,\{C_{\rho^{i},\gamma}|\gamma\in\mathbb{C}^{*}\})

and the group of diagonal matrices (as a groupoid with a single object 0(E~8)0\in\mathbb{C}(\tilde{E}_{8})).

(2) While for k2k\geqslant 2 or k=k=\infty, the groupoid Grpk(E~8)\operatorname{Grp}^{k}(\tilde{E}_{8}) is given by the action groupoid Grpac((E~8)|{Cρi,1})\operatorname{Grp}^{\operatorname{ac}}(\mathbb{C}(\tilde{E}_{8})\,|\,\{C_{\rho^{i},1}\}).

2. Preliminary and notations

2.1. Homogeneous isomorphism

In this paper, we denote by 𝒪n\mathcal{O}_{n} the formal local ring [[x1,,xn]]\mathbb{C}[[x_{1},\ldots,x_{n}]], and assume that the weight of xix_{i} equals wiw_{i}. For simplicity, we assume the weights wiw_{i} are positive integers. Given ItI_{t} and IsI_{s}, the two homogeneous ideals of 𝒪n\mathcal{O}_{n}, i.e., the generators in ItI_{t} and IsI_{s} are weighted homogeneous, we consider the graded analytic algebras 𝒪n/It\mathcal{O}_{n}/I_{t} and 𝒪n/Is\mathcal{O}_{n}/I_{s}. By the well-known lifting lemma, the isomorphism from 𝒪n/It\mathcal{O}_{n}/I_{t} to 𝒪n/Is\mathcal{O}_{n}/I_{s} is given by ϕAut(𝒪n)\phi\in\operatorname{Aut}(\mathcal{O}_{n}) such that

ϕ(It)Is,and ϕ1(Is)It.\phi(I_{t})\subseteq I_{s},\text{and }\phi^{-1}(I_{s})\subseteq I_{t}.

One can express ϕ:𝒪n𝒪n\phi:\mathcal{O}_{n}\to\mathcal{O}_{n} as

xjϕj(0)(x1,,xn)+ϕj(+)(x1,,xn) for j=1,,n,x_{j}\mapsto\phi^{(0)}_{j}(x_{1},\ldots,x_{n})+\phi^{(+)}_{j}(x_{1},\ldots,x_{n})\text{ for $j=1,\ldots,n$},

such that the monomials in ϕj(0)(x1,,xn)\phi^{(0)}_{j}(x_{1},\ldots,x_{n}) have weight wjw_{j}, while those in ϕj(0)(x1,,xn)\phi^{(0)}_{j}(x_{1},\ldots,x_{n}) have weight >wj>w_{j}. The homogeneous component ϕ(0)\phi^{(0)} of ϕ\phi, given by ϕ(0)=(ϕj(0))j\phi^{(0)}=(\phi_{j}^{(0)})_{j}, forms an isomorphism if and only if ϕ\phi is an isomorphism. Consequently, each isomorphism from 𝒪n/It\mathcal{O}_{n}/I_{t} to 𝒪n/Is\mathcal{O}_{n}/I_{s} induces a homogeneous isomorphism.

A \mathbb{C}^{*}-action is present on 𝒪n\mathcal{O}_{n} defined as:

λ:xjλwixj,\lambda\in\mathbb{C}^{*}:x_{j}\to\lambda^{w_{i}}x_{j},

resulting in a trivial automorphism of 𝒪/It\mathcal{O}/I_{t}. Moreover, if ϕ(x1,,xn)\phi(x_{1},\ldots,x_{n}) represents a homogeneous isomorphism, then ϕ(λw1x1,,λwnxn)\phi(\lambda^{w_{1}}x_{1},\ldots,\lambda^{w_{n}}x_{n}) also signifies a homogeneous isomorphism. This establishes the \mathbb{C}^{*}-action within the set of homogeneous isomorphisms.

2.2. Yau algebra with graded module structure

A graded Lie algebra is an ordinary Lie algebra gg together with a gradation of vector spaces

𝔤=i𝔤i\mathfrak{g}=\oplus_{i\in\mathbb{Z}}\mathfrak{g}_{i}

such that the Lie bracket respects this gradation:

[𝔤i,𝔤j]𝔤i+j.[\mathfrak{g}_{i},\mathfrak{g}_{j}]\subseteq\mathfrak{g}_{i+j}.
Definition 5.

For an isolated hypersurface singularity 𝒱f\mathcal{V}_{f} determined by a polynomial f=f(x1,,xn)f=f(x_{1},\ldots,x_{n}), we define the kk-th Yau algebra of 𝒱f\mathcal{V}_{f} by

Lk(𝒱f):=Der(𝒜k(𝒱f),𝒜k(𝒱f));L^{k}(\mathcal{V}_{f}):=\operatorname{Der}(\mathcal{A}^{k}(\mathcal{V}_{f}),\mathcal{A}^{k}(\mathcal{V}_{f}));

namely the derivation Lie algebra of the kk-th moduli algebra.

We need to introduce a graded 𝒜k(𝒱f)\mathcal{A}^{k}(\mathcal{V}_{f})-module structure of Lk(𝒱f)L^{k}(\mathcal{V}_{f}). Denote by i\partial_{i} the partial derivation with respect to xix_{i}. By definition, a derivation δLk(𝒱f)\delta\in L^{k}(\mathcal{V}_{f}) is of the form

δ=i=1naii,\delta=\sum_{i=1}^{n}a_{i}\partial_{i},

where the coefficients ai𝒜k(𝒱f)a_{i}\in\mathcal{A}^{k}(\mathcal{V}_{f}) satisfy the condition

δf,mkJ(f)f,mkJ(f).\delta\cdot\langle f,m^{k}J(f)\rangle\subseteq\langle f,m^{k}J(f)\rangle.

In this way, we obtain the embedding map of 𝒜k(𝒱f)\mathcal{A}^{k}(\mathcal{V}_{f})-modules:

(3) Lk(𝒱f)𝒜k(𝒱f)1,,nL^{k}(\mathcal{V}_{f})\to\mathcal{A}^{k}(\mathcal{V}_{f})\langle\partial_{1},\ldots,\partial_{n}\rangle

where 𝒜k(𝒱f)1,,n\mathcal{A}^{k}(\mathcal{V}_{f})\langle\partial_{1},\ldots,\partial_{n}\rangle denotes the free 𝒜k(𝒱f)\mathcal{A}^{k}(\mathcal{V}_{f})-module with generator 1,,n\partial_{1},\ldots,\partial_{n}.

With the assumption that ff is a weighted homogeneous polynomial, the algebra 𝒜k(𝒱f)\mathcal{A}^{k}(\mathcal{V}_{f}) admits a graded algebra structure since the ideal f,mkJ(f)\langle f,m^{k}J(f)\rangle is homogeneous. For a monomial \mathcal{M} in 𝒜k(𝒱f)\mathcal{A}^{k}(\mathcal{V}_{f}), we define the notion of degree as follows:

deg(xi)=weight()weight(xi).\deg(\mathcal{M}\partial_{x_{i}})=\operatorname{weight}(\mathcal{M})-\operatorname{weight}(x_{i}).

Based on this grading, the free 𝒜k(𝒱f)\mathcal{A}^{k}(\mathcal{V}_{f})-module 𝒜k(𝒱f)1,,n\mathcal{A}^{k}(\mathcal{V}_{f})\langle\partial_{1},\ldots,\partial_{n}\rangle is regarded as a graded Lie algebra. Similarly, the kk-th Yau algebra Lk(𝒱f)L^{k}(\mathcal{V}_{f}) inherits a graded Lie algebra structure.

We utilize the crucial observation that a homogeneous isomorphism of kk-th moduli algebras leads to an isomorphism of the corresponding kk-th Yau algebras. For weighted homogeneous polynomials f1f_{1} and f2f_{2}, a homogeneous isomorphism

ϕ:𝒜k(𝒱f1)𝒜k(𝒱f2)\phi:\mathcal{A}^{k}(\mathcal{V}_{f_{1}})\to\mathcal{A}^{k}(\mathcal{V}_{f_{2}})

induces a Lie-algebraic homomorphism

ϕ:Lk(𝒱f1)Lk(𝒱f2).\phi_{*}:L^{k}(\mathcal{V}_{f_{1}})\to L^{k}(\mathcal{V}_{f_{2}}).

Such homomorphism ϕ\phi_{*} is ϕ\phi-equivalent meaning that

ϕ(ge1)=ϕ(g)ϕ(e1)\phi_{*}(ge_{1})=\phi(g)\phi_{*}(e_{1})

and

ϕ[e1,e2]=[ϕe1,ϕe2]\phi_{*}[e_{1},e_{2}]=[\phi_{*}e_{1},\phi_{*}e_{2}]

hold for g𝒜k(𝒱f1)g\in\mathcal{A}^{k}(\mathcal{V}_{f_{1}}) and e1,e2Lk(𝒱f1)e_{1},e_{2}\in L^{k}(\mathcal{V}_{f_{1}}).

2.3. Notation of homomorphisms for Lie algebra

When focusing on the case where 𝒱f\mathcal{V}_{f} from E~i\tilde{E}_{i}-families, we consistently employ the subsequent notations to denote homogeneous isomorphisms of the kk-th moduli algebras of simple elliptic singularities throughout this paper.

Notation 6.

Suppose that (x,y,z)(x^{\prime},y^{\prime},z^{\prime}) are the coordinates of 𝒜k(E~i;s)\mathcal{A}^{k}(\tilde{E}_{i};s) and (x,y,z)(x,y,z) are the coordinates of 𝒜k(E~i;t)\mathcal{A}^{k}(\tilde{E}_{i};t). We denote a homogeneous isomorphism ϕ\phi from 𝒜k(E~i,t)\mathcal{A}^{k}(\tilde{E}_{i},t) to 𝒜k(E~i,s)\mathcal{A}^{k}(\tilde{E}_{i},s) through coordinate transformations as:

(4) {x=ϕ1(x,y,z);y=ϕ2(x,y,z);z=ϕ3(x,y,z),\begin{cases}x=\phi_{1}(x^{\prime},y^{\prime},z^{\prime});\\ y=\phi_{2}(x^{\prime},y^{\prime},z^{\prime});\\ z=\phi_{3}(x^{\prime},y^{\prime},z^{\prime}),\end{cases}

where ϕi\phi_{i}’s are homogeneous functions.

Specifically, when all wi=1w_{i}=1, ϕ\phi uniquely corresponds to a nonsingular linear transformation of 3\mathbb{C}^{3}. One can represent ϕ\phi as a linear transform

(5) (xyz)=(a1a2a3b1b2b3c1c2c3)(xyz).\begin{pmatrix}x\\ y\\ z\end{pmatrix}=\begin{pmatrix}a_{1}&a_{2}&a_{3}\\ b_{1}&b_{2}&b_{3}\\ c_{1}&c_{2}&c_{3}\end{pmatrix}\begin{pmatrix}x^{\prime}\\ y^{\prime}\\ z^{\prime}\end{pmatrix}.

Its inverse can be expressed as

(6) (xyz)=(a1a2a3b1b2b3c1c2c3)(xyz).\begin{pmatrix}x^{\prime}\\ y^{\prime}\\ z^{\prime}\end{pmatrix}=\begin{pmatrix}a_{1}^{\prime}&a_{2}^{\prime}&a_{3}^{\prime}\\ b_{1}^{\prime}&b_{2}^{\prime}&b_{3}^{\prime}\\ c_{1}^{\prime}&c_{2}^{\prime}&c_{3}^{\prime}\end{pmatrix}\begin{pmatrix}x\\ y\\ z\end{pmatrix}.

We define 𝒜k(E~i;t)x,y,z\mathcal{A}^{k}(\tilde{E}_{i};t)\langle\partial_{x},\partial_{y},\partial_{z}\rangle to be the free 𝒜k(E~i;t)\mathcal{A}^{k}(\tilde{E}_{i};t)-module generated by x,y,z\partial_{x},\partial_{y},\partial_{z}.

As in (3), we possess the embedding map

Lk(E~i;t)𝒜k(E~i;t)x,y,z.L^{k}(\tilde{E}_{i};t)\to\mathcal{A}^{k}(\tilde{E}_{i};t)\langle\partial_{x},\partial_{y},\partial_{z}\rangle.

According to Leibniz’s Rule, we derive the formulas for the differential of ϕ\phi:

(7) ϕ(x)\displaystyle\phi_{*}(\partial_{x}) =x(x)x+x(y)y+x(z)z=a1x+b1y+c1z;\displaystyle=\partial_{x}(x^{\prime})\partial_{x^{\prime}}+\partial_{x}(y^{\prime})\partial_{y^{\prime}}+\partial_{x}(z^{\prime})\partial_{z^{\prime}}=a^{\prime}_{1}\partial_{x^{\prime}}+b^{\prime}_{1}\partial_{y^{\prime}}+c^{\prime}_{1}\partial_{z^{\prime}};
ϕ(y)\displaystyle\phi_{*}(\partial_{y}) =y(x)x+y(y)y+y(z)z=a2x+b2y+c2z;\displaystyle=\partial_{y}(x^{\prime})\partial_{x^{\prime}}+\partial_{y}(y^{\prime})\partial_{y^{\prime}}+\partial_{y}(z^{\prime})\partial_{z^{\prime}}=a^{\prime}_{2}\partial_{x^{\prime}}+b^{\prime}_{2}\partial_{y^{\prime}}+c^{\prime}_{2}\partial_{z^{\prime}};
ϕ(z)\displaystyle\phi_{*}(\partial_{z}) =z(x)x+z(y)y+z(z)z=a3x+b3y+c3z.\displaystyle=\partial_{z}(x^{\prime})\partial_{x^{\prime}}+\partial_{z}(y^{\prime})\partial_{y^{\prime}}+\partial_{z}(z^{\prime})\partial_{z^{\prime}}=a^{\prime}_{3}\partial_{x^{\prime}}+b^{\prime}_{3}\partial_{y^{\prime}}+c^{\prime}_{3}\partial_{z^{\prime}}.

In this way, we obtain a ϕ\phi-equivalent map

ϕ:\displaystyle\phi_{*}: Lk(E~i;t)Lk(E~i;s)\displaystyle L^{k}(\tilde{E}_{i};t)\to L^{k}(\tilde{E}_{i};s)
ϕ(xaybzci)=\displaystyle\phi_{*}(x^{a}y^{b}z^{c}\partial_{i})= ϕ1(x,y,z)aϕ2(x,y,z)bϕ3(x,y,z)cϕ(i),\displaystyle\phi_{1}(x^{\prime},y^{\prime},z^{\prime})^{a}\phi_{2}(x^{\prime},y^{\prime},z^{\prime})^{b}\phi_{3}(x^{\prime},y^{\prime},z^{\prime})^{c}\phi_{*}(\partial_{i}),

which preserves the Lie structures.

2.4. Basis property of groupoid

We examine some fundamental properties of the simple elliptic groupoid Grpk(E~i)\operatorname{Grp}^{k}(\tilde{E}_{i}) associated to 𝒜k(E~i;t)\mathcal{A}^{k}(\tilde{E}_{i};t). Let ftf_{t} be the defining function of the E~i\tilde{E}_{i}-family as described in Equation (1). Recall that the kk-th moduli algebra is defined by the formal local ring 𝒪\mathcal{O} modulo the ideal

Ik(t)=ft,mkJ(ft).I_{k}(t)=\langle f_{t},m^{k}J(f_{t})\rangle.

In this context, a morphism ϕ\phi of Grpk(E~i)(t,s)\operatorname{Grp}^{k}(\tilde{E}_{i})(t,s) is expressed as an automorphism in PGL3𝐰()\operatorname{PGL}_{3}^{\mathbf{w}}(\mathbb{C}) such that ϕ(Ik(t))=Ik(s)\phi(I_{k}(t))=I_{k}(s).

Lemma 7.
  1. (1)

    The groupoid Grp(E~i)\operatorname{Grp}^{\infty}(\tilde{E}_{i}) is a sub-groupoid of Grpk(E~i)\operatorname{Grp}^{k}(\tilde{E}_{i}) for each k0k\geqslant 0.

  2. (2)

    When ll is sufficiently large, we have

    Grpk(E~i)=Grp(E~i)\operatorname{Grp}^{k}(\tilde{E}_{i})=\operatorname{Grp}^{\infty}(\tilde{E}_{i})

    for klk\geqslant l.

  3. (3)

    The groupoid Grp0(E~i)\operatorname{Grp}^{0}(\tilde{E}_{i}) is a sub-groupoid of Grp1(E~i)\operatorname{Grp}^{1}(\tilde{E}_{i}).

Proof.

(1) The first assertion is deduced from the fact that the isomorphism of the coordinate ring induces weighted homogeneous isomorphisms of the kk-th moduli algebras.

(2) For sufficiently large kk, we can assume that the weights of the generators in mkJ(ft)m^{k}J(f_{t}) exceed the weight of ftf_{t}. Consequently, the homogeneous polynomial ftf_{t}, as a generator of the defining ideal Ik(t)I_{k}(t) of the algebra 𝒜(E~i;t)\mathcal{A}(\tilde{E}_{i};t), has the minimal weight among the generators. Therefore, each morphism in Grpk(E~i)\operatorname{Grp}^{k}(\tilde{E}_{i}) yields a transformation ϕ\phi of 𝒪n\mathcal{O}_{n} which satisfies the equation

ϕft(x,y,z)=λfs(x,y,z)\phi f_{t}(x,y,z)=\lambda\cdot f_{s}(x^{\prime},y^{\prime},z^{\prime})

for some constant λ\lambda. This relationship yields that ϕ\phi is essentially a weighted homogeneous isomorphism between kk-moduli algebras.

(3) When k=0,1k=0,1, we have ftmJ(ft)f_{t}\in mJ(f_{t}) for any weighted homogeneous polynomial ftf_{t}. Then

I1(t)=ft,mJ(ft)=mJ(ft)=mI0(t).I_{1}(t)=\langle f_{t},mJ(f_{t})\rangle=\langle mJ(f_{t})\rangle=mI_{0}(t).

Note that a homogeneous isomorphism ϕ\phi preserves the maximal ideal mm. Therefore, a homogeneous isomorphism ϕGrp0(E~i)\phi\in\operatorname{Grp}^{0}(\tilde{E}_{i}) is also an isomorphism in the context of the first moduli algebras. ∎

3. Groupoid of the E~8\tilde{E}_{8}-family

In this section, we focus on the isomorphisms of the E~8\tilde{E}_{8}-family. Recall that the E~8\tilde{E}_{8}-family is defined by the polynomial

ft=x6+y3+z2+tx4y=0,where 4t327,f_{t}=x^{6}+y^{3}+z^{2}+tx^{4}y=0,\quad\text{where }4t^{3}\neq-27,

which is weighted homogeneous of type (1,2,3;6)(1,2,3;6). By definition, the kk-th moduli algebra takes the form

𝒜k(E~8;t)=[[x,y,z]]/Ik(t),\mathcal{A}^{k}(\tilde{E}_{8};t)=\mathbb{C}[[x,y,z]]/I_{k}(t),

where

Ik(t)=x6+y3+z2+tx4y,mk(3x5+2tx3y),mk(3y2+tx4),mkz.I_{k}(t)=\left\langle x^{6}+y^{3}+z^{2}+tx^{4}y,\,m^{k}(3x^{5}+2tx^{3}y),\,m^{k}(3y^{2}+tx^{4}),\,m^{k}z\right\rangle.

Our goal is to determine the isomorphism classes of 𝒜k(E~8;t)\mathcal{A}^{k}(\tilde{E}_{8};t) for various values of tt. Since the monomials y,x2y,x^{2} are both have weight 22 and xx is the unique monomial of weight 11, any homogeneous automorphism ϕ\phi of E~8\tilde{E}_{8}-families can be expressed as

{x=λx,y=λ2(ay+bx2),z=λ3cz+dx3+exyλ,a,c,b\begin{cases}x=\lambda x^{\prime},\\ y=\lambda^{2}(ay^{\prime}+bx^{\prime 2}),\\ z=\lambda^{3}cz^{\prime}+dx^{\prime 3}+ex^{\prime}y^{\prime}\end{cases}\lambda,a,c\in\mathbb{C}^{*},b\in\mathbb{C}

in accordance with the notation in (4). Up to scaling, we can always assume that λ=1\lambda=1. The possible generators with weight less than k+3k+3 are xkzx^{k}z and fsf_{s} (when 3<k3<k). From the fact

xkz=0modIk(t),x^{k}z=0\mod I_{k}(t),

we have

xk(z+dx3+exy)=0mod(xkz,fs).x^{\prime k}(z^{\prime}+dx^{\prime 3}+ex^{\prime}y^{\prime})=0\mod(x^{\prime k}z^{\prime},f_{s}).

This implies that d=e=0d=e=0.

Similarly, we have

xk(3y2+tx4)=0modIk(t),x^{k}(3y^{2}+tx^{4})=0\mod I_{k}(t),

and the possible generators in Ik(s)I_{k}(s) with weight k+4\leqslant k+4 are xk(3y2+tx4),zx^{k}(3y^{2}+tx^{4}),z and fsf_{s}.

xk(3(ay+bx2)2+tx4)=0mod(fs,z,xk(3y2+sx4))x^{\prime k}(3(ay^{\prime}+bx^{\prime 2})^{2}+tx^{\prime 4})=0\mod(f_{s},z,x^{\prime k}(3y^{\prime 2}+sx^{\prime 4}))

This implies that b=0b=0. In conclusion, we find that the automorphism ϕ\phi is represented by a diagonal matrix.

3.1. Case k2k\geqslant 2

The generator in Ik(t)I_{k}(t) with weight less than or equal to 66 are ftf_{t}, x3zx^{3}z (for k=3k=3), and xyzxyz, x2(3y2+tx4)x^{2}(3y^{2}+tx^{4}) (for k=2k=2). From this, we obtain the relation

x6+y3+z2+tx4y=c2(x6+y3+z2+sx4y)mod(fs,x3z,xyz,x2(3y2+sx4)).x^{6}+y^{3}+z^{2}+tx^{4}y=c^{2}(x^{\prime 6}+y^{\prime 3}+z^{\prime 2}+sx^{\prime 4}y^{\prime})\mod(f_{s},x^{\prime 3}z^{\prime},x^{\prime}y^{\prime}z^{\prime},x^{\prime 2}(3y^{2}+sx^{\prime 4})).

This simplifies to:

x6+y3+z2+tx4y=c2(x6+y3+z2+sx4y).x^{6}+y^{3}+z^{2}+tx^{4}y=c^{2}(x^{\prime 6}+y^{\prime 3}+z^{\prime 2}+sx^{\prime 4}y^{\prime}).

Since x=ϕ1(x)=xx=\phi_{1}(x^{\prime})=x^{\prime}, we find that c=1c=1. Hence, we have

a=ρi,b=0,c=1,s=tρi.a=\rho^{i},b=0,c=1,s=t\rho^{i}.

This completes the proof of the case k=2k=2 in Theorem C.

3.2. Case k=0k=0

From relations

{3x5+2tx3y=3y2+tx4=0mod(I0(t))3x5+2sx3y=3y2+sx4=0mod(I0(s)),\begin{cases}3x^{5}+2tx^{3}y=3y^{2}+tx^{4}=0\mod(I_{0}(t))\\ 3x^{\prime 5}+2sx^{\prime 3}y=3y^{\prime 2}+sx^{\prime 4}=0\mod(I_{0}(s)),\end{cases}

we obtain

3x5+2tx3y=\displaystyle 3x^{5}+2tx^{3}y= 3x5+2sx3y\displaystyle 3x^{\prime 5}+2sx^{\prime 3}y
3y2+tx4=\displaystyle 3y^{2}+tx^{4}= 3y2+sx4.\displaystyle 3y^{\prime 2}+sx^{\prime 4}.

Therefore, we have

t=sa, and t=sa2.t=sa,\text{ and }t=\frac{s}{a^{2}}.

For t0t\not=0, we obtain

a3=1,s=ta.a^{3}=1,s=ta.

Note that in this case, there is no essential restriction on ϕ(z)=cz\phi(z)=cz^{\prime}, allowing cc to be any arbitrary element in \mathbb{C}^{*}. For t=0t=0, then we obtain s=0s=0 and there are no essential restrictions on aa and cc.

3.3. Case k=1k=1

This case is essentially the same with k=0k=0 by replacing J(ft)J(f_{t}) with mJ(ft)mJ(f_{t}). This completes the proof of Theorem C.

4. Groupoid of the E~6\tilde{E}_{6}-family

Now we turn our attention to the study of simple elliptic singularity E~6\tilde{E}_{6}. For the parameter t(E~6)t\in\mathbb{C}(\tilde{E}_{6}) associated with the E~6\tilde{E}_{6}-family, the isolated hypersurface singularity is defined by the polynomial

ft:=x3+y3+z3+txyz.f_{t}:=x^{3}+y^{3}+z^{3}+txyz.

Note that the Jacobi ideal of ftf_{t} is given by

J(ft)=3x2+tyz,3y2+txz,3z2+txy.J(f_{t})=\langle 3x^{2}+tyz,3y^{2}+txz,3z^{2}+txy\rangle.

We have

𝒜0(E~6;t)=[[x,y,z]]/3x2+tyz,3y2+txz,3z2+txy,\mathcal{A}^{0}(\tilde{E}_{6};t)=\mathbb{C}[[x,y,z]]/\langle 3x^{2}+tyz,3y^{2}+txz,3z^{2}+txy\rangle,

and

𝒜1(E~6;t)=[[x,y,z]]/3x3+txyz,3y3+txyz,3z3+txyz,x2y,x2z,y2x,y2z,z2x,z2y.\mathcal{A}^{1}(\tilde{E}_{6};t)=\mathbb{C}[[x,y,z]]/\langle 3x^{3}+txyz,3y^{3}+txyz,3z^{3}+txyz,x^{2}y,x^{2}z,y^{2}x,y^{2}z,z^{2}x,z^{2}y\rangle.

For k2k\geqslant 2, it is easy to see that

mk+2mkJ(ft)mk+2.m^{k+2}\subseteq m^{k}J(f_{t})\subseteq m^{k+2}.

Hence, mkJ(ft)=mk+2m^{k}J(f_{t})=m^{k+2}. This implies that for k2k\geqslant 2,

𝒜k(E~6;t)=[[x,y,z]]/ft,mkJ(ft)=[[x,y,z]]/ft,mk+2.\mathcal{A}^{k}(\tilde{E}_{6};t)=\mathbb{C}[[x,y,z]]/\langle f_{t},m^{k}J(f_{t})\rangle=\mathbb{C}[[x,y,z]]/\langle f_{t},m^{k+2}\rangle.

Applying (2) in Lemma 7 it implies that for k2k\geqslant 2,

(8) Grpk(E~6)=Grp(E~6).\operatorname{Grp}^{k}(\tilde{E}_{6})=\operatorname{Grp}^{\infty}(\tilde{E}_{6}).

In order to prove Theorem A, we need to study the first Yau Algebra of the E~6\tilde{E}_{6}-family, defined as

L1(E~6;t):=Der(𝒜1(E~6;t),𝒜1(E~6;t)).L^{1}(\tilde{E}_{6};t):=\operatorname{Der}(\mathcal{A}^{1}(\tilde{E}_{6};t),\mathcal{A}^{1}(\tilde{E}_{6};t)).

By direct calculation, we find that for the case t(E~6){0,6,6ρ,6ρ2}t\in\mathbb{C}(\tilde{E}_{6})\setminus\{0,6,6\rho,6\rho^{2}\}, the algebra L1(E~6;t)L^{1}(\tilde{E}_{6};t) is 2222-dimensional, with a \mathbb{C}-linear basis represented as

{xx+yy+zz}m2𝒜1(𝒱t)x,y,z.\{x\partial_{x}+y\partial_{y}+z\partial_{z}\}\cup m^{2}\mathcal{A}^{1}(\mathcal{V}_{t})\langle\partial_{x},\partial_{y},\partial_{z}\rangle.

One may choose the basis of 𝒜1(E~6;t)\mathcal{A}^{1}(\tilde{E}_{6};t):

{1,x,y,z,x2,y2,z2,xy,xz,yz,xyz}\{1,x,y,z,x^{2},y^{2},z^{2},xy,xz,yz,xyz\}

with multiplication rules:

x3\displaystyle x^{3} =y3=z3=t3xyz,\displaystyle=y^{3}=z^{3}=-\frac{t}{3}xyz,
x2y\displaystyle x^{2}y =x2z=y2x=y2z=z2x=z2y=0.\displaystyle=x^{2}z=y^{2}x=y^{2}z=z^{2}x=z^{2}y=0.

Then the set

{xx+yy+zz}{x2i,y2i,z2i,xyi,xzi,yzi,xyzi},\displaystyle\{x\partial_{x}+y\partial_{y}+z\partial_{z}\}\cup\{x^{2}\partial_{i},y^{2}\partial_{i},z^{2}\partial_{i},xy\partial_{i},xz\partial_{i},yz\partial_{i},xyz\partial_{i}\},

where i=x,y,z\partial_{i}=\partial_{x},\partial_{y},\partial_{z}, forms a basis of L1(E~6;t)L^{1}(\tilde{E}_{6};t) (see Section 4.3 for another complete list of bases). In the case of jump points, i.e., t{0,6,6ρ,6ρ2}t\in\{0,6,6\rho,6\rho^{2}\}, the dimension of L1(E~6;t)L^{1}(\tilde{E}_{6};t) equals 2424.

Through direct computation, we find that each matrix AiA_{i} in Theorem A is a morphism in Grpk(E~6)\operatorname{Grp}^{k}(\tilde{E}_{6}) for k0k\geqslant 0 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. (1)

    The action groupoid Grpac({0,6,6ρ,6ρ2}|G)\operatorname{Grp}^{\operatorname{ac}}(\{0,6,6\rho,6\rho^{2}\}\,|\,G) is a sub-groupoid of Grpk(E~6)\operatorname{Grp}^{k}(\tilde{E}_{6}). In particular, Grpac({0,6,6ρ,6ρ2}|G)\operatorname{Grp}^{\operatorname{ac}}(\{0,6,6\rho,6\rho^{2}\}\,|\,G) becomes a fully faithful sub-groupoid of Grpk(E~6)\operatorname{Grp}^{k}(\tilde{E}_{6}) when k2k\geqslant 2.

  2. (2)

    When t,s(E~6){0,6,6ρ,6ρ2}t,s\in\mathbb{C}(\tilde{E}_{6})\setminus\{0,6,6\rho,6\rho^{2}\}, the equality

    Grpk(E~6)(t,s)=Grp(E~6)(t,s)\operatorname{Grp}^{k}(\tilde{E}_{6})(t,s)=\operatorname{Grp}^{\infty}(\tilde{E}_{6})(t,s)

    holds for k0k\geqslant 0.

  3. (3)

    When t,s(E~6){0,6,6ρ,6ρ2}t,s\in\mathbb{C}(\tilde{E}_{6})\setminus\{0,6,6\rho,6\rho^{2}\}, each morphism in Grp1(E~6)(t,s)\operatorname{Grp}^{1}(\tilde{E}_{6})(t,s) is generated by A1,,A4A_{1},\ldots,A_{4}.

Proof of the first assertion.

It is straightforward to verify that the group GG induces morphisms of Grpk(E~6)\operatorname{Grp}^{k}(\tilde{E}_{6}), and that the objects 0,6,6ρ,6ρ20,6,6\rho,6\rho^{2} are stable under GG.

For k2k\geqslant 2, the morphisms in Grpk(E~6)(0,0)\operatorname{Grp}^{k}(\tilde{E}_{6})(0,0) preserve the cubic polynomial

f0=x3+y3+z3.f_{0}=x^{3}+y^{3}+z^{3}.

Therefore, Grpk(E~6)(0,0)\operatorname{Grp}^{k}(\tilde{E}_{6})(0,0) is represented by scaling matrices (equivalent to Id\operatorname{Id} in PGL3()\operatorname{PGL}_{3}(\mathbb{C})) and permutations of three letters. By composing with A1iA2A_{1}^{i}A_{2}, it becomes apparent that Grpk(E~6)(t,s)\operatorname{Grp}^{k}(\tilde{E}_{6})(t,s) for t,s{0,6,6ρ,6ρ2}t,s\in\{0,6,6\rho,6\rho^{2}\} is generated by A1,,A4A_{1},\ldots,A_{4}. ∎

4.1. The second assertion

Notation 9.
  1. (1)

    For an element 𝐰\mathbf{w} in 𝒜1(E~6;t)\mathcal{A}^{1}(\tilde{E}_{6};t) of weight one, we introduce the symbol

    𝐟t(x,y,z)(𝐰)=ft(α,β,γ)\mathbf{f}_{t}^{(x,y,z)}(\mathbf{w})=f_{t}(\alpha,\beta,\gamma)

    where α,β,γ\alpha,\beta,\gamma are coefficients satisfying 𝐰=αx+βy+γz\mathbf{w}=\alpha x+\beta y+\gamma z. Additionally, we will utilize the notation 𝐟s(x,y,z)(𝐰)\mathbf{f}_{s}^{(x^{\prime},y^{\prime},z^{\prime})}(\mathbf{w}) in reference to alternative coordinates x,y,zx^{\prime},y^{\prime},z^{\prime}.

  2. (2)

    For linear subspace U,VL1(E~6;t)U,V\subseteq L^{1}(\tilde{E}_{6};t), we denote by [U,V]t[U,V]_{t} the vector space spanned by all elements [u,v]t[u,v]_{t} where uUu\in U, and vVv\in V. For a vector vv, we also define [v,U]t[v,U]_{t} in the obvious manner.

To prove the second assertion of Lemma 8, we need the following technique lemma characterizing the relation between Grp1(E~6)\operatorname{Grp}^{1}(\tilde{E}_{6}) and Grp(E~6)\operatorname{Grp}^{\infty}(\tilde{E}_{6}).

Lemma 10.

Assume that ϕ\phi is contained in Grp1(E~6)(t,s)\operatorname{Grp}^{1}(\tilde{E}_{6})(t,s) with t,s(E~6){0,6,6ρ,6ρ2}t,s\in\mathbb{C}(\tilde{E}_{6})\setminus\{0,6,6\rho,6\rho^{2}\} and 𝐰\mathbf{w} is an element of weight one in 𝒜(E~6;t)\mathcal{A}(\tilde{E}_{6};t).

  1. (1)

    The quantity

    𝐟ϕ:=𝐟18s(x,y,z)(ϕ(𝐰))𝐟18t(x,y,z)(𝐰)\mathbf{f}_{\phi}:=\frac{\mathbf{f}_{\frac{-18}{s}}^{(x^{\prime},y^{\prime},z^{\prime})}(\phi(\mathbf{w}))}{\mathbf{f}_{\frac{-18}{t}}^{(x,y,z)}(\mathbf{w})}

    is independent on the choice of 𝐰\mathbf{w} (but depends on t,st,s).

  2. (2)

    The transpose of the matrix ϕ\phi is contained in Grp(E~6)(18s,18t)\operatorname{Grp}^{\infty}(\tilde{E}_{6})(\frac{-18}{s},\frac{-18}{t}).

Proof.

(1) Denote by [,]t[-,-]_{t} the Lie bracket of L1(E~6;t)L^{1}(\tilde{E}_{6};t). Let DtD_{t} be the \mathbb{C}-linear space with linear basis x,y,z\partial_{x},\partial_{y},\partial_{z}.

For an element 𝐰=αx+βy+γz\mathbf{w}=\alpha x+\beta y+\gamma z in 𝒜1(E~6;t)\mathcal{A}^{1}(\tilde{E}_{6};t) of weight one, we define 𝐰:=𝐰(xx+yy+zz)\mathcal{H}_{\mathbf{w}}:=\mathbf{w}(x\partial_{x}+y\partial_{y}+z\partial_{z}). Then

[𝐰,L1(E~6;t)]t=xyz(βx+αy),xyz(γx+αz).[\mathcal{H}_{\mathbf{w}},L^{1}(\tilde{E}_{6};t)]_{t}=\langle xyz(-\beta\partial_{x}+\alpha\partial_{y}),xyz(-\gamma\partial_{x}+\alpha\partial_{z})\rangle.

We obtain a linear subspace of DtD_{t} parameterized by 𝐰\mathbf{w}:

Dt′′(𝐰):=βx+αy,γx+αz.D_{t}^{\prime\prime}(\mathbf{w}):=\langle-\beta\partial_{x}+\alpha\partial_{y},-\gamma\partial_{x}+\alpha\partial_{z}\rangle.

For arbitrary =ξ1(βx+αy)+ξ2(γx+αz)Dt′′(𝐰)\partial=\xi_{1}(-\beta\partial_{x}+\alpha\partial_{y})+\xi_{2}(-\gamma\partial_{x}+\alpha\partial_{z})\in D_{t}^{\prime\prime}(\mathbf{w}) with ξ1,ξ2\xi_{1},\xi_{2}\in\mathbb{C}, we have

[𝐰2,𝐰]t\displaystyle[\mathbf{w}^{2}\partial,\mathcal{H}_{\mathbf{w}}]_{t}
=[𝐰2ξ1(βx+αy)+𝐰2ξ2(γx+αz),𝐰(xx+yy+zz)]t\displaystyle=[\mathbf{w}^{2}\xi_{1}(-\beta\partial_{x}+\alpha\partial_{y})+\mathbf{w}^{2}\xi_{2}(-\gamma\partial_{x}+\alpha\partial_{z}),\mathbf{w}(x\partial_{x}+y\partial_{y}+z\partial_{z})]_{t}
(9) =t3𝐟18t(𝐰)xyz.\displaystyle=\frac{t}{3}\mathbf{f}_{\frac{-18}{t}}(\mathbf{w})\cdot xyz\partial.

Assume that ϕ\phi is represented by the matrix (5) with inverse (6). Let us investigate the images of quantities 𝐰\mathbf{w}, 𝐰\mathcal{H}_{\mathbf{w}}, Dt′′(𝐰)D_{t}^{\prime\prime}(\mathbf{w}) and [𝐰2,𝐰]t[\mathbf{w}^{2}\partial,\mathcal{H}_{\mathbf{w}}]_{t} under the isomorphism ϕ\phi.

  • For the image of 𝐰\mathbf{w}, we obtain

    (10) ϕ(𝐰)=(αa1+βb1+γc1)x+(αa2+βb2+γc2)y+(αa3+βb3+γc3)z.\phi(\mathbf{w})=(\alpha a_{1}+\beta b_{1}+\gamma c_{1})x^{\prime}+(\alpha a_{2}+\beta b_{2}+\gamma c_{2})y^{\prime}+(\alpha a_{3}+\beta b_{3}+\gamma c_{3})z^{\prime}.
  • For 𝐰\mathcal{H}_{\mathbf{w}}, we have

    ϕ(𝐰)=ϕ(𝐰)(xx+yy+zz).\phi_{*}(\mathcal{H}_{\mathbf{w}})=\phi(\mathbf{w})(x^{\prime}\partial_{x^{\prime}}+y^{\prime}\partial_{y^{\prime}}+z^{\prime}\partial_{z^{\prime}}).
  • For Dt′′(𝐰)D_{t}^{\prime\prime}(\mathbf{w}), we compute the Lie bracket of ϕ(𝐰)\phi_{*}(\mathcal{H}_{\mathbf{w}}) and Ls1L_{s}^{1} to obtain

    (11) ϕ(Dt′′(𝐰))=\displaystyle\phi_{*}(D_{t}^{\prime\prime}(\mathbf{w}))=\langle (αa2+βb2+γc2)x+(αa1+βb1+γc1)y,\displaystyle-(\alpha a_{2}+\beta b_{2}+\gamma c_{2})\partial_{x^{\prime}}+(\alpha a_{1}+\beta b_{1}+\gamma c_{1})\partial_{y^{\prime}},
    (αa3+βb3+γc3)x+(αa1+βb1+γc1)z.\displaystyle-(\alpha a_{3}+\beta b_{3}+\gamma c_{3})\partial_{x^{\prime}}+(\alpha a_{1}+\beta b_{1}+\gamma c_{1})\partial_{z^{\prime}}\rangle.
  • If =ϕ()ϕ(Dt′′(𝐰))\partial_{*}=\phi_{*}(\partial)\in\phi_{*}(D_{t}^{\prime\prime}(\mathbf{w})), then

    (12) [ϕ(𝐰2),ϕ(𝐰)]s=[ϕ(𝐰2),ϕ(𝐰)]s=s3𝐟s18(x,y,z)(ϕ(𝐰))xyz.[\phi_{*}(\mathbf{w}^{2}\partial),\phi_{*}(\mathcal{H}_{\mathbf{w}})]_{s}=[\phi(\mathbf{w}^{2})\partial_{*},\phi_{*}(\mathcal{H}_{\mathbf{w}})]_{s}=\frac{s}{3}\mathbf{f}_{\frac{-s}{18}}^{(x^{\prime},y^{\prime},z^{\prime})}(\phi(\mathbf{w}))x^{\prime}y^{\prime}z^{\prime}\partial_{*}.

Note that ϕ\phi_{*} preserves the Lie brackets. Combining Equations (9) and (12), we find

t3𝐟18t(x,y,z)(𝐰)ϕ(xyz)\displaystyle\frac{t}{3}\mathbf{f}_{\frac{-18}{t}}^{(x,y,z)}(\mathbf{w})\cdot\phi_{*}(xyz)\partial_{*} =t3𝐟18t(x,y,z)(𝐰)ϕ(xyz)\displaystyle=\frac{t}{3}\mathbf{f}_{\frac{-18}{t}}^{(x,y,z)}(\mathbf{w})\cdot\phi_{*}(xyz\partial)
=ϕ[𝐰2,w]t\displaystyle=\phi_{*}[\mathbf{w}^{2}\partial,\mathcal{H}_{w}]_{t}
=[ϕ(𝐰2),ϕ(𝐰)]s\displaystyle=[\phi_{*}(\mathbf{w}^{2}\partial),\phi_{*}(\mathcal{H}_{\mathbf{w}})]_{s}
=s3𝐟s18(x,y,z)(ϕ(𝐰))xyz.\displaystyle=\frac{s}{3}\mathbf{f}_{\frac{-s}{18}}^{(x^{\prime},y^{\prime},z^{\prime})}(\phi(\mathbf{w}))x^{\prime}y^{\prime}z^{\prime}\partial_{*}.

Thus, for any element 𝐰\mathbf{w} in 𝒜1(E~6;t)\mathcal{A}^{1}(\tilde{E}_{6};t) of weight one, we obtain the equality

(13) ϕ(xyz)=s𝐟18s(x,y,z)(ϕ(𝐰))t𝐟18t(x,y,z)(𝐰)xyz.\phi(xyz)=\frac{s\mathbf{f}_{\frac{-18}{s}}^{(x^{\prime},y^{\prime},z^{\prime})}(\phi(\mathbf{w}))}{t\mathbf{f}_{\frac{-18}{t}}^{(x,y,z)}(\mathbf{w})}\cdot x^{\prime}y^{\prime}z^{\prime}.

So the factor

s𝐟18s(x,y,z)(ϕ(𝐰))t𝐟18t(x,y,z)(𝐰)\frac{s\mathbf{f}_{\frac{-18}{s}}^{(x^{\prime},y^{\prime},z^{\prime})}(\phi(\mathbf{w}))}{t\mathbf{f}_{\frac{-18}{t}}^{(x,y,z)}(\mathbf{w})}

is invariant regardless of the selection of 𝐰\mathbf{w}. Consequently, the first assertion is validated.

(2) From the first assertion, any element 𝐰\mathbf{w} in 𝒜1(E~6;t)\mathcal{A}^{1}(\tilde{E}_{6};t) of weight one satisfies

(14) 𝐟18s(x,y,z)(ϕ(𝐰))=𝐟ϕ𝐟18t(x,y,z)(𝐰)\mathbf{f}_{\frac{-18}{s}}^{(x^{\prime},y^{\prime},z^{\prime})}(\phi(\mathbf{w}))=\mathbf{f}_{\phi}\cdot\mathbf{f}_{\frac{-18}{t}}^{(x,y,z)}(\mathbf{w})

for some constant 𝐟ϕ\mathbf{f}_{\phi}. From the definition 𝐟18t(x,y,z)(𝐰)=f18t(α,β,γ)\mathbf{f}_{\frac{-18}{t}}^{(x,y,z)}(\mathbf{w})=f_{\frac{-18}{t}}(\alpha,\beta,\gamma) and the explicit expansion of the term 𝐟18s(x,y,z)(ϕ(𝐰))\mathbf{f}_{\frac{-18}{s}}^{(x^{\prime},y^{\prime},z^{\prime})}(\phi(\mathbf{w})), we see that the transformation 𝐰ϕ(𝐰)\mathbf{w}\mapsto\phi(\mathbf{w}) represents a morphism in Grp(E~6)(18s,18t)\operatorname{Grp}^{\infty}(\tilde{E}_{6})(\frac{-18}{s},\frac{-18}{t}). The equality (10) can be rewritten as

ϕ(𝐰)=αx+βy+γz\phi(\mathbf{w})=\alpha^{\prime}x^{\prime}+\beta^{\prime}y^{\prime}+\gamma^{\prime}z^{\prime}

where

(αβγ)=(a1b1c1a2b2c2a3b3c3)(αβγ).\begin{pmatrix}\alpha^{\prime}\\ \beta^{\prime}\\ \gamma^{\prime}\end{pmatrix}=\begin{pmatrix}a_{1}&b_{1}&c_{1}\\ a_{2}&b_{2}&c_{2}\\ a_{3}&b_{3}&c_{3}\end{pmatrix}\begin{pmatrix}\alpha\\ \beta\\ \gamma\end{pmatrix}.

It means that 𝐰ϕ(𝐰)\mathbf{w}\mapsto\phi(\mathbf{w}) is represented as the transpose matrix of ϕ\phi, and thus the lemma holds. ∎

Proof of the second assertion of Lemma 8.

We have known from (8) that the second assertion holds for k2k\geqslant 2. It suffices to consider the cases k=0k=0 and k=1k=1.

We have already shown in (1) of Lemma 7 that

Grp(E~6)(t,s)Grp0(E~6)(t,s).\operatorname{Grp}^{\infty}(\tilde{E}_{6})(t,s)\subseteq\operatorname{Grp}^{0}(\tilde{E}_{6})(t,s).

It follows from (3) of Lemma 7 that

Grp0(E~6)(t,s)Grp1(E~6)(t,s).\operatorname{Grp}^{0}(\tilde{E}_{6})(t,s)\subseteq\operatorname{Grp}^{1}(\tilde{E}_{6})(t,s).

For a subset SPGL3()S\subseteq\operatorname{PGL}_{3}(\mathbb{C}), we denote all the transposes of matrices in SS by STS^{T}. We obtain from Lemma 10 that

Grp1(E~6)(t,s)TGrp(E~6)(18s,18t).\operatorname{Grp}^{1}(\tilde{E}_{6})(t,s)^{T}\subseteq\operatorname{Grp}^{\infty}(\tilde{E}_{6})(-\frac{18}{s},-\frac{18}{t}).

Note that this implies

Grp1(E~6)(t,s)\displaystyle\operatorname{Grp}^{1}(\tilde{E}_{6})(t,s) Grp(E~6)(18s,18t)T\displaystyle\subseteq\operatorname{Grp}^{\infty}(\tilde{E}_{6})(-\frac{18}{s},-\frac{18}{t})^{T}
Grp1(E~6)(18s,18t)T\displaystyle\subseteq\operatorname{Grp}^{1}(\tilde{E}_{6})(-\frac{18}{s},-\frac{18}{t})^{T}
Grp(E~6)(t,s).\displaystyle\subseteq\operatorname{Grp}^{\infty}(\tilde{E}_{6})(t,s).

Finally, we conclude that

Grp0(E~6)(t,s)=Grp1(E~6)(t,s)=Grp(E~6)(t,s).\operatorname{Grp}^{0}(\tilde{E}_{6})(t,s)=\operatorname{Grp}^{1}(\tilde{E}_{6})(t,s)=\operatorname{Grp}^{\infty}(\tilde{E}_{6})(t,s).

4.2. The third assertion

Lemma 11.

For t,s(E~6){0,6,6ρ,6ρ2}t,s\in\mathbb{C}(\tilde{E}_{6})\setminus\{0,6,6\rho,6\rho^{2}\}, we assume that the morphism ϕ\phi in Grp1(E~6)(t,s)\operatorname{Grp}^{1}(\tilde{E}_{6})(t,s) is presented as the matrix (5) with inverse (6). Then

(15a) a1=(a126sa2a3)/f18s(a1,a2,a3),a_{1}^{\prime}=(a_{1}^{2}-\frac{6}{s}a_{2}a_{3})/f_{\frac{-18}{s}}(a_{1},a_{2},a_{3}),
(15b) b1=(a226sa1a3)/f18s(a1,a2,a3),b_{1}^{\prime}=(a_{2}^{2}-\frac{6}{s}a_{1}a_{3})/f_{\frac{-18}{s}}(a_{1},a_{2},a_{3}),
(15c) c1=(a326sa1a2)/f18s(a1,a2,a3).c_{1}^{\prime}=(a_{3}^{2}-\frac{6}{s}a_{1}a_{2})/f_{\frac{-18}{s}}(a_{1},a_{2},a_{3}).
Proof.

We maintain the notations of Lemma 10. If we select 𝐰=x𝒜1(E~6;t)\mathbf{w}=x\in\mathcal{A}^{1}(\tilde{E}_{6};t), then 𝐰:=𝐰(xx+yy+zz)\mathcal{H}_{\mathbf{w}}:=\mathbf{w}(x\partial_{x}+y\partial_{y}+z\partial_{z}). Define N(𝐰):=𝐰2x,𝐰2y,𝐰2zN(\mathbf{w}):=\langle\mathbf{w}^{2}\partial_{x},\mathbf{w}^{2}\partial_{y},\mathbf{w}^{2}\partial_{z}\rangle and Z=x,y,zZ=\langle\mathcal{H}_{x},\mathcal{H}_{y},\mathcal{H}_{z}\rangle. Then,

V(𝐰):={uN(𝐰):[u,Z]t=0}=𝐰2x.V(\mathbf{w}):=\{u\in N(\mathbf{w}):[u,Z]_{t}=0\}=\langle\mathbf{w}^{2}\partial_{x}\rangle.

This implies that we identify a 11-dimensional subspace Dt(𝐰):=xD_{t}^{\prime}(\mathbf{w}):=\langle\partial_{x}\rangle parameterized by 𝐰\mathbf{w}. By computing the Lie bracket of 𝐰\mathcal{H}_{\mathbf{w}} and L1(E~6;t)L^{1}(\tilde{E}_{6};t), we obtain the subspace

Dt′′(𝐰)=y,z.D_{t}^{\prime\prime}(\mathbf{w})=\langle\partial_{y},\partial_{z}\rangle.

Therefore, 𝐰\mathbf{w} leads to a decomposition of DtD_{t} as follows:

Dt=Dt(𝐰)Dt′′(𝐰)=xy,z.D_{t}=D_{t}^{\prime}(\mathbf{w})\oplus D_{t}^{\prime\prime}(\mathbf{w})=\langle\partial_{x}\rangle\oplus\langle\partial_{y},\partial_{z}\rangle.

In other words, we generate the vector space x\langle\partial_{x}\rangle through the influence of 𝐰\mathbf{w}. Thus, we achieve a decomposition of xx+yy+zzx\partial_{x}+y\partial_{y}+z\partial_{z} relative to 𝐰\mathbf{w}, given by

(16) xx+yy+zz=𝐰x+(yy+zz).x\partial_{x}+y\partial_{y}+z\partial_{z}=\mathbf{w}\partial_{x}+(y\partial_{y}+z\partial_{z}).

It yields that x\partial_{x} (not just the linear space x\langle\partial_{x}\rangle) is determined by 𝐰\mathbf{w}.

We turn to study the corresponding decomposition of DtD_{t} and image of x\partial_{x} under ϕ\phi_{*}. By definition, we have

ϕ(𝐰)=a1x+a2y+a3z.\phi(\mathbf{w})=a_{1}x^{\prime}+a_{2}y^{\prime}+a_{3}z^{\prime}.

Then the image of 𝐰\mathcal{H}_{\mathbf{w}} equals

ϕ(𝐰)=(a1x+a2y+a3z)(xx+yy+zz).\phi_{*}(\mathcal{H}_{\mathbf{w}})=(a_{1}x^{\prime}+a_{2}y^{\prime}+a_{3}z^{\prime})({x^{\prime}}\partial_{x^{\prime}}+{y^{\prime}}\partial_{y^{\prime}}+z^{\prime}\partial_{z^{\prime}}).

Evidently,

ϕN(𝐰)=ϕ(𝐰)2x,ϕ(𝐰)2y,ϕ(𝐰)2z.\phi_{*}N(\mathbf{w})=\langle\phi(\mathbf{w})^{2}\partial_{x^{\prime}},\phi(\mathbf{w})^{2}\partial_{y^{\prime}},\phi(\mathbf{w})^{2}\partial_{z^{\prime}}\rangle.

Since ϕZ=Z\phi_{*}Z=Z, it follows that

ϕV(𝐰)\displaystyle\phi_{*}V(\mathbf{w}) ={uϕN(𝐰):[u,Z]s=0}\displaystyle=\{u\in\phi_{*}N(\mathbf{w}):[u,Z]_{s}=0\}
=ϕ(𝐰)2((sa126a2a3)x+(sa226a1a3)y+(sa326a1a2)z).\displaystyle=\langle\phi(\mathbf{w})^{2}\left((sa_{1}^{2}-6a_{2}a_{3})\partial_{x^{\prime}}+(sa_{2}^{2}-6a_{1}a_{3})\partial_{y^{\prime}}+(sa_{3}^{2}-6a_{1}a_{2})\partial_{z^{\prime}}\right)\rangle.

Hence, we have

ϕDt(𝐰)=(sa126a2a3)x+(sa226a1a3)y+(sa326a1a2)z.\phi_{*}D_{t}^{\prime}(\mathbf{w})=\langle(sa_{1}^{2}-6a_{2}a_{3})\partial_{x^{\prime}}+(sa_{2}^{2}-6a_{1}a_{3})\partial_{y^{\prime}}+(sa_{3}^{2}-6a_{1}a_{2})\partial_{z^{\prime}}\rangle.

Recalling the computation of ϕDt′′(𝐰)\phi_{*}D_{t}^{\prime\prime}(\mathbf{w}) as obtained in Equation (11):

ϕDt′′(𝐰)=a2x+a1y,a3x+a1z,\phi_{*}D_{t}^{\prime\prime}(\mathbf{w})=\langle-a_{2}\partial_{x^{\prime}}+a_{1}\partial_{y^{\prime}},-a_{3}\partial_{x^{\prime}}+a_{1}\partial_{z^{\prime}}\rangle,

we arrive at the decomposition

Ds=ϕDt(𝐰)ϕDt′′(𝐰).D_{s}=\phi_{*}D_{t}^{\prime}(\mathbf{w})\oplus\phi_{*}D_{t}^{\prime\prime}(\mathbf{w}).

Thus, the decomposition of xx+yy+zzx^{\prime}\partial_{x^{\prime}}+y^{\prime}\partial_{y^{\prime}}+z^{\prime}\partial_{z^{\prime}} is given by

ϕ(xx+yy+zz)\displaystyle\phi_{*}(x\partial_{x}+y\partial_{y}+z\partial_{z})
=xx+yy+zz\displaystyle=x^{\prime}\partial_{x^{\prime}}+y^{\prime}\partial_{y^{\prime}}+z^{\prime}\partial_{z^{\prime}}
=ϕ(𝐰)f18s(a1,a2,a3)((a126sa2a3)x+(a226sa1a3)y+(a326sa1a2)z)+Δ𝐰,\displaystyle=\frac{\phi(\mathbf{w})}{f_{\frac{-18}{s}}(a_{1},a_{2},a_{3})}\left((a_{1}^{2}-\frac{6}{s}a_{2}a_{3})\partial_{x^{\prime}}+(a_{2}^{2}-\frac{6}{s}a_{1}a_{3})\partial_{y^{\prime}}+(a_{3}^{2}-\frac{6}{s}a_{1}a_{2})\partial_{z^{\prime}}\right)+\Delta_{\mathbf{w}},

for some Δ𝐰ϕDt′′(𝐰)\Delta_{\mathbf{w}}\in\phi_{*}D_{t}^{\prime\prime}(\mathbf{w}). Comparing this with (16), we see the image of x\partial_{x} is given by

ϕ(x)=1f18s(a1,a2,a3)((a126sa2a3)x+(a226sa1a3)y+(a326sa1a2)z).\phi_{*}(\partial_{x})=\frac{1}{f_{\frac{-18}{s}}(a_{1},a_{2},a_{3})}\left((a_{1}^{2}-\frac{6}{s}a_{2}a_{3})\partial_{x^{\prime}}+(a_{2}^{2}-\frac{6}{s}a_{1}a_{3})\partial_{y^{\prime}}+(a_{3}^{2}-\frac{6}{s}a_{1}a_{2})\partial_{z^{\prime}}\right).

Notice that

ϕ(x)\displaystyle\phi_{*}(\partial_{x}) =a1x+b1y+c1z.\displaystyle=a_{1}^{\prime}\partial_{x^{\prime}}+b_{1}^{\prime}\partial_{y^{\prime}}+c_{1}^{\prime}\partial_{z^{\prime}}.

So we obtain the equalities (15a)(15b)(15c). ∎

Lemma 12.

Assume that t,s(E~6){0,6,6ρ,6ρ2}t,s\in\mathbb{C}(\tilde{E}_{6})\setminus\{0,6,6\rho,6\rho^{2}\}. Let ϕ\phi be a homogeneous isomorphism contained in Grp1(E~6)(t,s)\operatorname{Grp}^{1}(\tilde{E}_{6})(t,s). Denote by a1,a2,a3a_{1},a_{2},a_{3} the entries in the first row of the representation matrix (5) of ϕ\phi. Then only one the following two cases may occur:

  1. (1)

    If all elements a1,a2,a3a_{1},a_{2},a_{3} are nonzero, then a13=a23=a33a_{1}^{3}=a_{2}^{3}=a_{3}^{3};

  2. (2)

    If at least two elements of {a1,a2,a3}\{a_{1},a_{2},a_{3}\} are equal to zero.

Proof.

(1) Firstly, we assume that a1,a2,a3a_{1},a_{2},a_{3} are nonzero. Let ϕ\phi be a morphism contained in Grp1(E~6)(t,s)\operatorname{Grp}^{1}(\tilde{E}_{6})(t,s). Lemma 10 implies that the transpose matrix of ϕ1\phi^{-1} represents a morphism in Grp1(E~6)(18t,18s)\operatorname{Grp}^{1}(\tilde{E}_{6})(\frac{-18}{t},\frac{-18}{s}). Analogous to the equalities (15a)-(15c), we deduce the equations

(17a) a1=(a12+s3b1c1)/fs(a1,b1,c1),a_{1}=(a_{1}^{\prime 2}+\frac{s}{3}b_{1}^{\prime}c_{1}^{\prime})/f_{s}(a_{1}^{\prime},b_{1}^{\prime},c_{1}^{\prime}),
(17b) a2=(b12+s3a1c1)/fs(a1,b1,c1),a_{2}=(b_{1}^{\prime 2}+\frac{s}{3}a_{1}^{\prime}c_{1}^{\prime})/f_{s}(a_{1}^{\prime},b_{1}^{\prime},c_{1}^{\prime}),
(17c) a3=(c12+s3a1b1)/fs(a1,b1,c1).a_{3}=(c_{1}^{\prime 2}+\frac{s}{3}a_{1}^{\prime}b_{1}^{\prime})/f_{s}(a_{1}^{\prime},b_{1}^{\prime},c_{1}^{\prime}).

Combining these equations shows that the values aia_{i} with i=1,2,3i=1,2,3 verify the equality

ai3(s2+2)+(s33+36)1ai3a12a22a32=Λ\displaystyle a_{i}^{3}(s^{2}+2)+(\frac{s^{3}}{3}+36)\frac{1}{a_{i}^{3}}a_{1}^{2}a_{2}^{2}a_{3}^{2}=\Lambda

where Λ\Lambda denotes a constant of the form

Λ:=2(a13+a23+a33)+s2(f18s(a1,a2,a3))2fs(a1,b1,c1).\Lambda:=2(a_{1}^{3}+a_{2}^{3}+a_{3}^{3})+s^{2}(f_{\frac{-18}{s}}(a_{1},a_{2},a_{3}))^{2}f_{s}(a_{1}^{\prime},b_{1}^{\prime},c_{1}^{\prime}).

Equivalently, we find that a13,a23,a33a_{1}^{3},a_{2}^{3},a_{3}^{3} are roots of G(X)G(X) defined as

(18) G(X):=(s2+2)X+(s33+36)a12a22a321XΛ.G(X):=(s^{2}+2)\cdot X+(\frac{s^{3}}{3}+36)a_{1}^{2}a_{2}^{2}a_{3}^{2}\cdot\frac{1}{X}-\Lambda.

Since the function G(X)G(X) can have at most two distinct roots, at least two of the values ai3a_{i}^{3} must coincide. If our lemma does not hold, then we can assume without loss of generality that a13=a23a33a_{1}^{3}=a_{2}^{3}\not=a_{3}^{3}. From the expression in (18), we observe that both (s33+36)(\frac{s^{3}}{3}+36) and (s2+2)(s^{2}+2) are nonzero, which implies that

a13a33=a23a33=s3+1083s2+6a12a22a32.a_{1}^{3}a_{3}^{3}=a_{2}^{3}a_{3}^{3}=\frac{s^{3}+108}{3s^{2}+6}\cdot a_{1}^{2}a_{2}^{2}a_{3}^{2}.

This can be simplified to the relation

(19) a3=s3+1083s2+6a22a1.a_{3}=\frac{s^{3}+108}{3s^{2}+6}\cdot\frac{a_{2}^{2}}{a_{1}}.

We now consider another morphism ϕ\phi^{\prime} in Grp1(E~6)(t,s)\operatorname{Grp}^{1}(\tilde{E}_{6})(t,s) represented by the matrix

(a1a2ρa3b1b2ρb3c1c2ρc3)=(a1a2a3b1b2b3c1c2c3)(1000ρ0001)\begin{pmatrix}a_{1}&a_{2}\rho&a_{3}\\ b_{1}&b_{2}\rho&b_{3}\\ c_{1}&c_{2}\rho&c_{3}\end{pmatrix}=\begin{pmatrix}a_{1}&a_{2}&a_{3}\\ b_{1}&b_{2}&b_{3}\\ c_{1}&c_{2}&c_{3}\end{pmatrix}\cdot\begin{pmatrix}1&0&0\\ 0&\rho&0\\ 0&0&1\end{pmatrix}

It is clear that a13=(a2ρ)3a33a_{1}^{3}=(a_{2}\rho)^{3}\not=a_{3}^{3}. The arguments above applied to ϕ\phi^{\prime} lead us to the expression

(20) a3=6(sρ)3+64818(sρ)2+36(a2ρ)2a1a_{3}=\frac{6(s\rho)^{3}+648}{18(s\rho)^{2}+36}\cdot\frac{(a_{2}\rho)^{2}}{a_{1}}

which parallels (19). Combining Equations (19) with (20) yields that

(sρ)3+1083(sρ)2+6ρ2=s3+1083s2+6,\frac{(s\rho)^{3}+108}{3(s\rho)^{2}+6}\cdot\rho^{2}=\frac{s^{3}+108}{3s^{2}+6},

leading to

ρ2=1,\rho^{2}=1,

which contradicts the assumption ρ3=1\rho^{3}=1. So we obtain a13=a23=a33a_{1}^{3}=a_{2}^{3}=a_{3}^{3}.

(2) In the second case, if at least one of the values aia_{i} equals zero, we can set a3=0a_{3}=0 without loss of generality. The equations (15a)-(15c) imply:

a1\displaystyle a_{1}^{\prime} =a12/f18s(a1,a2,a3),\displaystyle=a_{1}^{2}/f_{\frac{-18}{s}}(a_{1},a_{2},a_{3}),
b1\displaystyle b_{1}^{\prime} =a22/f18s(a1,a2,a3),\displaystyle=a_{2}^{2}/f_{\frac{-18}{s}}(a_{1},a_{2},a_{3}),
c1\displaystyle c_{1}^{\prime} =6sa1a2/f18s(a1,a2,a3).\displaystyle=-\frac{6}{s}a_{1}a_{2}/f_{\frac{-18}{s}}(a_{1},a_{2},a_{3}).

Assuming both a1a_{1} and a2a_{2} are nonzero, we obtain a1,b1,c1a_{1}^{\prime},b_{1}^{\prime},c_{1}^{\prime} are also nonzero. Applying the result from the first case to the transpose matrix of ϕ1\phi^{-1}, we get

(21) a13=b13=c13.a_{1}^{\prime 3}=b_{1}^{\prime 3}=c_{1}^{\prime 3}.

Now Equation (17c)\eqref{eq:a3} implies

0=a3=(a12+s3b1c1)/fs(a1,b1,c1).0=a_{3}=(a_{1}^{\prime 2}+\frac{s}{3}b_{1}^{\prime}c_{1}^{\prime})/f_{s}(a_{1}^{\prime},b_{1}^{\prime},c_{1}^{\prime}).

It follows that a12=s3b1c1a_{1}^{\prime 2}=-\frac{s}{3}b_{1}^{\prime}c_{1}^{\prime}. Combining this with Equations (21), we have s3=27s^{3}=-27. This contradicts our initial assumption about ss. Therefore, the assumption about a1a_{1} and a2a_{2} must be false, ensuring that either a1a_{1} or a2a_{2} is zero. In conclusion, the second case of the lemma is validated. ∎

We are in the position to give a complete description for Grp1(E~6)(t,s)\operatorname{Grp}^{1}(\tilde{E}_{6})(t,s).

Proof of the third assertion of Lemma 8.

We are employing the notations of Lemma 12.

Case (1): There exists some index i=1,2,3i=1,2,3 such that ai=0a_{i}=0.

For instance, if we set a3=0a_{3}=0, applying Lemma 12 reveals that either a1a_{1} or a2a_{2} is zero. For the case where a10,a2=a3=0a_{1}\not=0,a_{2}=a_{3}=0, Equations (15a)(15b)(15c) can be rewritten as

a1=sa12/θ,b1=0,c1=0.a_{1}^{\prime}=sa_{1}^{2}/\theta,\quad b_{1}^{\prime}=0,\quad c_{1}^{\prime}=0.

Combining these equalities with the original constraints

b1a1+b2b1+b3c1=0b_{1}a_{1}^{\prime}+b_{2}b_{1}^{\prime}+b_{3}c_{1}^{\prime}=0

and

c1a1+c2b1+c3c1=0c_{1}a_{1}^{\prime}+c_{2}b_{1}^{\prime}+c_{3}c_{1}^{\prime}=0

derided from the definition, we have b1=c1=0b_{1}=c_{1}=0. Upon extending the above arguments to {b1,b2,b3}\{b_{1},b_{2},b_{3}\} and {c1,c2,c3}\{c_{1},c_{2},c_{3}\} respectively, we conclude that ϕ\phi is a matrix of types \@slowromancapi@ or \@slowromancapii@, where

\@slowromancapi@=(λ1000λ2000λ3) and \@slowromancapii@=(λ10000λ20λ30).\text{\@slowromancap i@}=\begin{pmatrix}\lambda_{1}&0&0\\ 0&\lambda_{2}&0\\ 0&0&\lambda_{3}\end{pmatrix}\text{ and \@slowromancap ii@}=\begin{pmatrix}\lambda_{1}&0&0\\ 0&0&\lambda_{2}\\ 0&\lambda_{3}&0\end{pmatrix}.

If we also consider the cases a30,a1=a2=0a_{3}\not=0,a_{1}=a_{2}=0 and a20,a1=a3=0a_{2}\not=0,a_{1}=a_{3}=0, then we obtain four additional types:

\@slowromancapiii@=(0λ10λ20000λ3),\@slowromancapiv@=(0λ1000λ2λ300),\@slowromancapv@=(00λ1λ2000λ30),\@slowromancapvi@=(00λ10λ20λ300).\text{\@slowromancap iii@}=\begin{pmatrix}0&\lambda_{1}&0\\ \lambda_{2}&0&0\\ 0&0&\lambda_{3}\end{pmatrix},\text{\@slowromancap iv@}=\begin{pmatrix}0&\lambda_{1}&0\\ 0&0&\lambda_{2}\\ \lambda_{3}&0&0\end{pmatrix},\text{\@slowromancap v@}=\begin{pmatrix}0&0&\lambda_{1}\\ \lambda_{2}&0&0\\ 0&\lambda_{3}&0\end{pmatrix},\text{\@slowromancap vi@}=\begin{pmatrix}0&0&\lambda_{1}\\ 0&\lambda_{2}&0\\ \lambda_{3}&0&0\end{pmatrix}.

The first assertion of Lemma 10 states that

𝐟ϕ:=𝐟18s(x,y,z)(ϕ(𝐰))𝐟18t(x,y,z)(𝐰)\mathbf{f}_{\phi}:=\frac{\mathbf{f}_{\frac{-18}{s}}^{(x^{\prime},y^{\prime},z^{\prime})}(\phi(\mathbf{w}))}{\mathbf{f}_{\frac{-18}{t}}^{(x,y,z)}(\mathbf{w})}

is invariant for any 𝐰\mathbf{w} of weight one. By choosing 𝐰=x,y,z\mathbf{w}=x,y,z respectively, we get the equality

(22) f18s(a1,a2,a3)=f18s(b1,b2,b3)=f18s(c1,c2,c3).f_{\frac{-18}{s}}(a_{1},a_{2},a_{3})=f_{\frac{-18}{s}}(b_{1},b_{2},b_{3})=f_{\frac{-18}{s}}(c_{1},c_{2},c_{3}).

Substituting the values of ai,bi,cia_{i},b_{i},c_{i} from the six types \@slowromancapi@-\@slowromancapvi@, we conclude that

λ13=λ23=λ33\lambda_{1}^{3}=\lambda_{2}^{3}=\lambda_{3}^{3}

in each type respectively. There are exactly 5454 matrices contained in these six types up to scalar multiplication and each matrix is generated by A1,A3,A4A_{1},A_{3},A_{4}.

Case (2): Suppose that a1,a2,a3a_{1},a_{2},a_{3} are nonzero.

From Case (1), we know that all entries of ϕ\phi are nonzero. Applying Lemma 12 to the sets {a1,a2,a3}\{a_{1},a_{2},a_{3}\}, {a1,b1,c1}\{a_{1},b_{1},c_{1}\}, {a2,b2,c2}\{a_{2},b_{2},c_{2}\} and {a3,b3,c3}\{a_{3},b_{3},c_{3}\} respectively, we conclude that all entries of ϕ\phi are contained in the set {a1,a1ρ,a1ρ2}\{a_{1},a_{1}\rho,a_{1}\rho^{2}\}.

From Equation (22), we get

(23) a1a2a3=b1b2b3=c1c2c3.a_{1}a_{2}a_{3}=b_{1}b_{2}b_{3}=c_{1}c_{2}c_{3}.

By considering the transpose of ϕ\phi, we deduced from Equation (23) that

(24) a1b1c1=a2b2c2=a3b3c3.a_{1}b_{1}c_{1}=a_{2}b_{2}c_{2}=a_{3}b_{3}c_{3}.

Fix a1a_{1}\in\mathbb{C}^{*}. Computational checks reveal that only 162×3162\times 3 non-singular matrices verify Equations (23) and (24). Up to scalar multiplication, there are exactly 162162 matrices. One may check directly that each matrix is generated by A1,A2,A3,A4A_{1},A_{2},A_{3},A_{4}. ∎

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 tt. For the entirety of this section, we assume that t30,216,27t^{3}\not=0,216,-27.

Recalling that the basis of 𝒜1(E~6;t)\mathcal{A}^{1}(\tilde{E}_{6};t) is given by

1,x,y,z,x2,y2,z2,xy,xz,yz,xyz.1,x,y,z,x^{2},y^{2},z^{2},xy,xz,yz,xyz.

We choose an alternative basis for L1(E~6;t)L^{1}(\tilde{E}_{6};t) as follows.

  1. (1)

    Degree 0:
    e1:=xx+yy+zze_{1}:=x\partial_{x}+y\partial_{y}+z\partial_{z};

  2. (2)

    Degree 1:
    e2:=3x2+tyzxe_{2}:=3x^{2}+tyz\partial_{x}, e3:=3y2+txzxe_{3}:=3y^{2}+txz\partial_{x}, e4:=3z2+txyxe_{4}:=3z^{2}+txy\partial_{x}, e5:=3x2+tyzye_{5}:=3x^{2}+tyz\partial_{y}, e6:=3y2+txzye_{6}:=3y^{2}+txz\partial_{y}, e7:=3z2+txyye_{7}:=3z^{2}+txy\partial_{y}, e8:=3x2+tyzze_{8}:=3x^{2}+tyz\partial_{z}, e9:=3y2+txzze_{9}:=3y^{2}+txz\partial_{z}, e10:=3z2+txyze_{10}:=3z^{2}+txy\partial_{z}, e11:=x2x+2xyye_{11}:=x^{2}\partial_{x}+2xy\partial_{y}, e12:=x2x+2xzxe_{12}:=x^{2}\partial_{x}+2xz\partial_{x},
    e13:=2xyx+y2ye_{13}:=2xy\partial_{x}+y^{2}\partial_{y}, e14:=y2y+2yzze_{14}:=y^{2}\partial_{y}+2yz\partial_{z}, e15:=2yzy+z2ze_{15}:=2yz\partial_{y}+z^{2}\partial_{z},
    e16:=2xzx+z2ze_{16}:=2xz\partial_{x}+z^{2}\partial_{z}, e17:=(1/t)x2xe_{17}:=(1/t)x^{2}\partial_{x}, e18:=(1/t)y2ye_{18}:=(1/t)y^{2}\partial_{y}, e19:=(1/t)z2ze_{19}:=(1/t)z^{2}\partial_{z};

  3. (3)

    Degree 2:
    e20:=xyzxe_{20}:=xyz\partial_{x}, e21:=xyzye_{21}:=xyz\partial_{y}, e22:=xyzze_{22}:=xyz\partial_{z}.

Observe that L1(E~6;t)L^{1}(\tilde{E}_{6};t) is a graded Lie algebra with the decomposition:

Lt1=D0D1D2,L^{1}_{t}=D_{0}\oplus D_{1}\oplus D_{2},

where DiD_{i}’s are generated by homogeneous derivations of degree ii respectively. Since there are no derivations of degree three, we have [D1,D2]=0[D_{1},D_{2}]=0. The subspace D0D_{0} is generated by the Euler derivation e1e_{1}. We obtain [e1,u]=u[e_{1},u]=u for uD1u\in D_{1} and [e1,u]=2u[e_{1},u]=2u for uD2u\in D_{2}. Next, we define the subspaces U:=e2,,e10U:=\langle e_{2},\cdots,e_{10}\rangle, V:=e11,,e16V:=\langle e_{11},\cdots,e_{16}\rangle, and W:=e17,e18,e19W:=\langle e_{17},e_{18},e_{19}\rangle. This leads to the decomposition:

D1=UVW.D_{1}=U\oplus V\oplus W.

One can check that

[U,U]=0,[U,V]=0,[W,W]=0,[V,W]=0.[U,U]=0,[U,V]=0,[W,W]=0,[V,W]=0.

Furthermore, the matrix representation for Lie bracket [U,W][U,W] is given by

(2e20002e21002e220002e20002e21002e220002e20002e21002e22),\left(\begin{matrix}2e_{20}&0&0&2e_{21}&0&0&2e_{22}&0&0\\ 0&2e_{20}&0&0&2e_{21}&0&0&2e_{22}&0\\ 0&0&2e_{20}&0&0&2e_{21}&0&0&2e_{22}\end{matrix}\right),

and the representation for [V,V][V,V] is given by

(0004e2204e21004e2204e21004e22004e2004e2200004e2004e214e200004e21004e2000).\left(\begin{matrix}0&0&0&4e_{22}&0&-4e_{21}\\ 0&0&-4e_{22}&0&4e_{21}&0\\ 0&4e_{22}&0&0&-4e_{20}&0\\ -4e_{22}&0&0&0&0&4e_{20}\\ 0&-4e_{21}&4e_{20}&0&0&0\\ 4e_{21}&0&0&-4e_{20}&0&0\end{matrix}\right).

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 tt. We have finished the proof of Theorem A.

5. Groupoid of the E~7\tilde{E}_{7}-family

5.1. Outline of the proof of Theorem B

Recall that the E~7\tilde{E}_{7}-family is defined as the zero locus of polynomial

ft=x4+y4+z2+tx2y2,where t24.f_{t}=x^{4}+y^{4}+z^{2}+tx^{2}y^{2},\text{where $t^{2}\not=4$}.

Firstly, we can directly verify that each matrix B~α,β,γ\tilde{B}_{\alpha,\beta,\gamma} in Theorem B induces an isomorphism from 𝒜k(E~7;t)\mathcal{A}^{k}(\tilde{E}_{7};t) to 𝒜k(E~7;s))\mathcal{A}^{k}(\tilde{E}_{7};s)), where s=π(Bα,β)(t)s=\pi^{\prime}(B_{\alpha,\beta})(t) and k0{}k\in\mathbb{Z}_{\geqslant 0}\cup\{\infty\}. For example, let ϕ\phi be the morphism represented as B~α,β,γ\tilde{B}_{\alpha,\beta,\gamma} with γ=2+tα2\gamma=\sqrt{2+t\alpha^{2}}. In terms of transformations, we can express this as:

(xyz)=(1β0ααβ000γ)(xyz).\begin{pmatrix}x\\ y\\ z\\ \end{pmatrix}=\begin{pmatrix}1&\beta&0\\ \alpha&-\alpha\beta&0\\ 0&0&\gamma\end{pmatrix}\begin{pmatrix}x^{\prime}\\ y^{\prime}\\ z^{\prime}\end{pmatrix}.

Restricting to the coordinates (x,y)(x,y), we get

h=π(Bα,β)=[tβ2(122α2t)2+tα2]h=\pi^{\prime}(B_{\alpha,\beta})=\left[t\mapsto\frac{\beta^{2}(12-2\alpha^{2}t)}{2+t\alpha^{2}}\right]

by the definition of π\pi^{\prime} in (2). A direct computation shows

ft(x,y,z)\displaystyle f_{t}(x,y,z) =(x+βy)4+(xβy)4+γ2z2+α2t(x2β2y2)2\displaystyle=(x^{\prime}+\beta y^{\prime})^{4}+(x^{\prime}-\beta y^{\prime})^{4}+\gamma^{2}z^{\prime 2}+\alpha^{2}t(x^{\prime 2}-\beta^{2}y^{\prime 2})^{2}
=2x4+2y4+12β2x2y2+γ2z2+α2t(x42x2β2y2+y4)\displaystyle=2x^{\prime 4}+2y^{\prime 4}+12\beta^{2}x^{\prime 2}y^{\prime 2}+\gamma^{2}z^{\prime 2}+\alpha^{2}t(x^{\prime 4}-2x^{\prime 2}\beta^{2}y^{\prime 2}+y^{\prime 4})
=(2+α2t)fs(x,y,z)\displaystyle=(2+\alpha^{2}t)f_{s}(x^{\prime},y^{\prime},z^{\prime})

where

s=β2(122α2t)2+tα2=h(t).s=\frac{\beta^{2}(12-2\alpha^{2}t)}{2+t\alpha^{2}}=h(t).

It follows that B~α,β,γ\tilde{B}_{\alpha,\beta,\gamma} is a morphism in groupoid Grp(E~7)\operatorname{Grp}^{\infty}(\tilde{E}_{7}).

From [15], we know that for t{0,±6}t\in\{0,\pm 6\}, we encounter the jump point case, in which the zeroth Yau algebra L0(E~7;t)L^{0}(\tilde{E}_{7};t) is 1212-dimensional. In contrast, for the cases where t0,±6t\not=0,\pm 6, L0(E~7;t)L^{0}(\tilde{E}_{7};t) is 1111-dimensional.

The proof for t=0,±6t=0,\pm 6 of Theorem B follows the same argument as in the assertion (1) of Lemma 8. Therefore, we omit the details.

When k2k\geqslant 2, we conclude from Lemma 7 that

Grpk(E~7)=Grp(E~7)Grpi(E~7)\operatorname{Grp}^{k}(\tilde{E}_{7})=\operatorname{Grp}^{\infty}(\tilde{E}_{7})\subseteq\operatorname{Grp}^{i}(\tilde{E}_{7})

for i=0,1i=0,1. So the assertion (2) in Theorem B is concluded by the assertion (1).

It remains to prove the case k=0,1k=0,1 and t0,±6t\not=0,\pm 6. We will compute the groupoid Grp0(E~7)\operatorname{Grp}^{0}(\tilde{E}_{7}) by restricting to the coordinates (x,y)(x,y). In fact, the complete collection of isomorphism with respect to (x,y)(x,y) are exactly the group GG^{\prime}, which is also described in [12] without proof. We utilize the Torelli-type theorem from [15] to show that the isomorphisms of the E~7\tilde{E}_{7}-family preserve the four lines in the zeroth Yau algebra. In this sense, the group GG^{\prime} is indeed the symmetric group acting on the four lines.

For the groupoid Grp1(E~7)\operatorname{Grp}^{1}(\tilde{E}_{7}), 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 Grp1(E~7)\operatorname{Grp}^{1}(\tilde{E}_{7}) are derived from Grp(E~7)\operatorname{Grp}^{\infty}(\tilde{E}_{7}). As a consequence of the Torelli-type theorem, we can reduce the groupoid Grp1(E~7)\operatorname{Grp}^{1}(\tilde{E}_{7}) to Grp(E~7)\operatorname{Grp}^{\infty}(\tilde{E}_{7}), which completes the proof of Theorem B.

5.2. The zeroth Yau algebra

We assume that t20,4,36t^{2}\not=0,4,36. The moduli algebra is written as

𝒜0(E~7;t)=[[x,y]]/I0(t)\displaystyle\mathcal{A}^{0}(\tilde{E}_{7};t)=\mathbb{C}[[x,y]]/I_{0}(t)

where, as usual,

I0(t)=4x3+2txy2,4y3+2tx2y,z.I_{0}(t)=\left\langle 4x^{3}+2txy^{2},4y^{3}+2tx^{2}y,z\right\rangle.

Notice that for t(E~7)t\in\mathbb{C}(\tilde{E}_{7}), we have

(4t2)x3y=y(4x3+2txy2)t2x(4y3+2tx2y)=0modI0(t)(4-t^{2})x^{3}y=y(4x^{3}+2txy^{2})-\frac{t}{2}x(4y^{3}+2tx^{2}y)=0\mod I_{0}(t)

and similarly

(4t2)y3x=x(4y3+2tyx2)t2y(4x3+2ty2x)=0modI0(t).(4-t^{2})y^{3}x=x(4y^{3}+2tyx^{2})-\frac{t}{2}y(4x^{3}+2ty^{2}x)=0\mod I_{0}(t).

The monomials

1,x,y,x2,xy,y2,x2y,y2x,x2y21,x,y,x^{2},xy,y^{2},x^{2}y,y^{2}x,x^{2}y^{2}

form a basis of 𝒜0(E~7;t)\mathcal{A}^{0}(\tilde{E}_{7};t) with multiplication rule

x3=t2xy2,y3=t2x2y,x3y=y3x=0.x^{3}=-\frac{t}{2}xy^{2},\quad y^{3}=-\frac{t}{2}x^{2}y,\quad x^{3}y=y^{3}x=0.

From [15], we know that the jump points of the E~7\tilde{E}_{7}-family are t=0,±6t=0,\pm 6, in which case the zeroth Yau algebra L0(E~7;t)L^{0}(\tilde{E}_{7};t) is 1212-dimensional. In contrast, for the case t0,±6t\not=0,\pm 6, the Lie algebra L0(E~7;t)L^{0}(\tilde{E}_{7};t) is 11-dimensional and there exists a basis for L0(E~7;t)L^{0}(\tilde{E}_{7};t):

  • Degree 0:
    e0=xx+yye_{0}=x\partial_{x}+y\partial_{y};

  • Degree 1:
    e1=x2x+xyye_{1}=x^{2}\partial_{x}+xy\partial_{y}, e2=yxx+y2ye_{2}=yx\partial_{x}+y^{2}\partial_{y}, e3=(t212)xyx+4tx2ye_{3}=(t^{2}-12)xy\partial_{x}+4tx^{2}\partial_{y},
    e4=4ty2x+(t212)xyye_{4}=4ty^{2}\partial_{x}+(t^{2}-12)xy\partial_{y};

  • Degree 2:
    e5=x2yxe_{5}=x^{2}y\partial_{x}, e6=xy2ye_{6}=xy^{2}\partial_{y}, e7=xy2xe_{7}=xy^{2}\partial_{x}, e8=x2yye_{8}=x^{2}y\partial_{y};

  • Degree 3:
    e9=x2y2xe_{9}=x^{2}y^{2}\partial_{x}, e10=x2y2ye_{10}=x^{2}y^{2}\partial_{y}.

Choose a solution CtC_{t} of the equation

(25) Ct2+1Ct2=12t.C_{t}^{2}+\frac{1}{C_{t}^{2}}=\frac{12}{t}.

We define l1(t),,l4(t)l_{1}(t),\ldots,l_{4}(t) (without orderings) to be the four lines on the plane e9,e10\langle e_{9},e_{10}\rangle whose slopes are exactly the solutions of equation (25), specifically:

l1(t)=\displaystyle l_{1}(t)= Cte9+e10,\displaystyle\left\langle C_{t}e_{9}+e_{10}\right\rangle,
l2(t)=\displaystyle l_{2}(t)= e9+Cte10,\displaystyle\left\langle e_{9}+C_{t}e_{10}\right\rangle,
l3(t)=\displaystyle l_{3}(t)= Cte9+e10,\displaystyle\left\langle-C_{t}e_{9}+e_{10}\right\rangle,
l4(t)=\displaystyle l_{4}(t)= e9Cte10.\displaystyle\left\langle e_{9}-C_{t}e_{10}\right\rangle.

The set L4L^{4} 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 S4S^{4} acting on the four lines, the 2424 orderings provide six different values of the cross-ratio:

6+t2t,2t6+t,6+t6t,6t6+t,2tt6,t62t\frac{6+t}{2t},\frac{2t}{6+t},\frac{6+t}{6-t},\frac{6-t}{6+t},\frac{2t}{t-6},\frac{t-6}{2t}

which constitute the six continuous invariants of L0(E~7;t)L^{0}(\tilde{E}_{7};t). The set of the six continuous invariants is equivalent to the jj-function j(E~7;t)j(\tilde{E}_{7};t). In addition, the transformations preserve L4L^{4} and leaves the six continuous invariants unchanged generate the Klein four-group. We adopt the approach that the morphisms in Grp0(E~7)\operatorname{Grp}^{0}(\tilde{E}_{7}) induce linear maps on the plane e9,e10\langle e_{9},e_{10}\rangle. In terms of matrices, the morphisms in Grp0(E~7)\operatorname{Grp}^{0}(\tilde{E}_{7}) essentially derive the action of GG^{\prime} on the four lines when restricted to the coordinates xx and yy.

Lemma 13.

The group GG^{\prime} acts on the set L4={l1(t),,l4(t)}L^{4}=\{l_{1}(t),\ldots,l_{4}(t)\} in one-to-one correspondence with the action of symmetric group S4S^{4}. Specifically, we have:

ϕ(li(t))=lσ(i)(s)\phi_{*}(l_{i}(t))=l_{\sigma(i)}(s)

for σS4\sigma\in S^{4}, ϕGrp0(t,s)G\phi\in\operatorname{Grp}^{0}(t,s)\cap G^{\prime} and t,st,s contained in some domain of (E~7)\mathbb{C}(\tilde{E}_{7}). In particular, the subgroup G0GG^{\prime}_{0}\subseteq G^{\prime} acts on {l1(t),,l4(t)}\{l_{1}(t),\ldots,l_{4}(t)\} as the Klein four-group.

Proof.

We only check for the isomorphism ϕ\phi expressed as the matrix Bα,βB_{\alpha,\beta} acting on l1l_{1}. Notice that

Bα,β1=(1212α12β12αβ).B_{\alpha,\beta}^{-1}=\begin{pmatrix}\frac{1}{2}&\frac{1}{2\alpha}\\ \frac{1}{2\beta}&-\frac{1}{2\alpha\beta}\end{pmatrix}.

Following the notation in (7) we have

ϕ(e9)=ϕ(x2y2)ϕ(x)=(x+βy)2(αxαβy)2(12x+12βy)\phi_{*}(e_{9})=\phi(x^{2}y^{2})\phi_{*}(\partial_{x})=(x^{\prime}+\beta y^{\prime})^{2}(\alpha x^{\prime}-\alpha\beta y^{\prime})^{2}(\frac{1}{2}\partial_{x^{\prime}}+\frac{1}{2\beta}\partial_{y^{\prime}})

and

ϕ(e10)=ϕ(x2y2)ϕ(y)=(x+βy)2(αxαβy)2(12αx12αβy).\phi_{*}(e_{10})=\phi(x^{2}y^{2})\phi_{*}(\partial_{y})=(x^{\prime}+\beta y^{\prime})^{2}(\alpha x^{\prime}-\alpha\beta y^{\prime})^{2}(\frac{1}{2\alpha}\partial_{x^{\prime}}-\frac{1}{2\alpha\beta}\partial_{y^{\prime}}).

Then the line ϕ(l1(t))\phi_{*}(l_{1}(t)) has a direction vector

ϕ(Cte9+e10)=λ(αβCt+β)e9+λ(αCt1)e10,\phi_{*}(C_{t}e_{9}+e_{10})=\lambda(\alpha\beta C_{t}+\beta)e_{9}+\lambda(\alpha C_{t}-1)e_{10},

for some constant λ\lambda. So the slope of ϕ(l1(t))\phi_{*}(l_{1}(t)) is given by

η=αβCt+βαCt1.\eta=\frac{\alpha\beta C_{t}+\beta}{\alpha C_{t}-1}.

Therefore, we obtain

η2+1η2\displaystyle\eta^{2}+\frac{1}{\eta^{2}} =(αβCt+β)2(αCt1)2+(αCt1)2(αβCt+β)2\displaystyle=\frac{(\alpha\beta C_{t}+\beta)^{2}}{(\alpha C_{t}-1)^{2}}+\frac{(\alpha C_{t}-1)^{2}}{(\alpha\beta C_{t}+\beta)^{2}}
=(αβCt+β)4+(αCt1)4(αβCt+β)2(αCt1)2\displaystyle=\frac{(\alpha\beta C_{t}+\beta)^{4}+(\alpha C_{t}-1)^{4}}{(\alpha\beta C_{t}+\beta)^{2}(\alpha C_{t}-1)^{2}}
=2Ct4+12α2Ct2+2β2(Ct42α2Ct2+1)\displaystyle=\frac{2C_{t}^{4}+12\alpha^{2}C_{t}^{2}+2}{\beta^{2}(C^{4}_{t}-2\alpha^{2}C_{t}^{2}+1)}
=212t+12α2β212t2α2β2=12+6tα26β2α2tβ2.\displaystyle=\frac{2\frac{12}{t}+12\alpha^{2}}{\beta^{2}\frac{12}{t}-2\alpha^{2}\beta^{2}}=\frac{12+6t\alpha^{2}}{6\beta^{2}-\alpha^{2}t\beta^{2}}.

Since s=β2(122α2t)2+tα2s=\frac{\beta^{2}(12-2\alpha^{2}t)}{2+t\alpha^{2}}, we can express

12s=122+tα2β2(122α2t)=η2+1η2,\frac{12}{s}=12\cdot\frac{2+t\alpha^{2}}{\beta^{2}(12-2\alpha^{2}t)}=\eta^{2}+\frac{1}{\eta^{2}},

which coincides with Equation (25). It yields that η\eta is one of the conjugate elements of CsC_{s}, and therefore the image of l1(t)l_{1}(t) is just one of the lines l1(s),,l4(s)l_{1}(s),\ldots,l_{4}(s). ∎

Now we are in the position to prove the case k=0k=0 in the assertion (1) of Theorem B.

Partial proof of Theorem B.

It suffices to show that the restriction of the morphism of Grp0(E~7)\operatorname{Grp}^{0}(\tilde{E}_{7}) to (x,y)(x,y) are contained in the group GG^{\prime}. Given a morphism ϕ\phi of Grp0(E~7)\operatorname{Grp}^{0}(\tilde{E}_{7}), we know that ϕ\phi induces a morphism ϕ\phi_{*} on the zero Yau algebra, which preserves the set L4L^{4} of four lines (not necessary in order). By composing with some matrix in GG^{\prime} if necessary, we can assume the ϕ\phi_{*} preserves the set L4L^{4} and the jj-function. We have argued that only the transformations arising from the Klein four-group leaves the jj-function invariants unchanged. According to Lemma 13, the Klein four-group are presented by G0G_{0}^{\prime} with respect to coordinates (x,y)(x,y). Thus, we conclude that the transformations on (x,y)(x,y) preserving the set of the four lines generate the group GG^{\prime}, which establishes the case k=0k=0 of Theorem B. ∎

5.3. The first Yau algebra of the E~7\tilde{E}_{7}-family

The first moduli algebra of the E~7\tilde{E}_{7}-family is given by

𝒜1(E~7;t)=[[x,y,z]]/I1(t)\mathcal{A}^{1}(\tilde{E}_{7};t)=\mathbb{C}[[x,y,z]]/I_{1}(t)

where

I1(t)=2x4+tx2y2,2y4+tx2y2,xy3,yx3,xz,yz,z2.I_{1}(t)=\left\langle 2x^{4}+tx^{2}y^{2},2y^{4}+tx^{2}y^{2},xy^{3},yx^{3},xz,yz,z^{2}\right\rangle.

For t{0,±6}t\not\in\{0,\pm 6\}, we construct the \mathbb{C}-linear basis of L1(t)L_{1}(t) given by e1,,e23e_{1},\ldots,e_{23}:

  • degree 0:
    e1=zze_{1}=z\partial_{z} and e2=xx+yye_{2}=x\partial_{x}+y\partial_{y};

  • degree 1:
    e3=x2xe_{3}=x^{2}\partial_{x}, e4=y2xe_{4}=y^{2}\partial_{x}, e5=xyxe_{5}=xy\partial_{x}, e6=x2ye_{6}=x^{2}\partial_{y}, e7=y2ye_{7}=y^{2}\partial_{y}, e8=xyye_{8}=xy\partial_{y}, e9=zxe_{9}=z\partial_{x},e10=zye_{10}=z\partial_{y}, e11=(2x3+txy2)ze_{11}=(2x^{3}+txy^{2})\partial_{z}, e12=(2y3+tx2y)ze_{12}=(2y^{3}+tx^{2}y)\partial_{z};

  • degree 2:
    e13=x3xe_{13}=x^{3}\partial_{x}, e14=x2yxe_{14}=x^{2}y\partial_{x}, e15=xy2xe_{15}=xy^{2}\partial_{x}, e16=y3xe_{16}=y^{3}\partial_{x},
    e17=x3ye_{17}=x^{3}\partial_{y}, e18=x2yye_{18}=x^{2}y\partial_{y}, e19=xy2ye_{19}=xy^{2}\partial_{y}, e20=y3ye_{20}=y^{3}\partial_{y}, e21=x2y2ze_{21}=x^{2}y^{2}\partial_{z};

  • degree 3:
    e22=x2y2xe_{22}=x^{2}y^{2}\partial_{x} and e23=x2y2ye_{23}=x^{2}y^{2}\partial_{y}.

We denote by [,][-,-] the Lie bracket of L1(E~7;t)L^{1}(\tilde{E}_{7};t) and use the notations [U,V][U,V] and [v,U][v,U] as in Notation 9. We aim to construct four distinct lines in a canonical order from the Lie structure of L1(E~7;t)L^{1}(\tilde{E}_{7};t).

Lemma 14.

There exists a sequence (H1,H2,H3,H4)(H_{1},H_{2},H_{3},H_{4}) of four lines in L1(E~7;t)L^{1}(\tilde{E}_{7};t) which is invariant up to homogeneous isomorphisms.

Proof.

Denote by DiD^{i} the weighted homogeneous part of L1(E~7;t)L^{1}(\tilde{E}_{7};t) of degree ii. The existence of the four lines can be verified through the following construction procedure:

  1. (1)

    The subspace e1\langle e_{1}\rangle of D0D^{0}.

  2. (2)

    The decomposition of D1D^{1}:

    (26) D1=D01D11D11D^{1}=D^{1}_{0}\oplus D^{1}_{1}\oplus D^{1}_{-1}

    where D01=e3,,e8D_{0}^{1}=\langle e_{3},\ldots,e_{8}\rangle, and D11=e9,e10D_{1}^{1}=\langle e_{9},e_{10}\rangle and D11=e11,e12D_{-1}^{1}=\langle e_{11},e_{12}\rangle.

  3. (3)

    The 44-dimension subspace VV of D01D_{0}^{1} generated by

    (27) z1=xe2=e3+e8,z2=ye2=e5+e7,z_{1}=xe_{2}=e_{3}+e_{8},\quad z_{2}=ye_{2}=e_{5}+e_{7},

    and

    (28) v1=4te4+8e3+(t24)e8,v2=4te6+8e7+(t24)e5.v_{1}=4te_{4}+8e_{3}+(t^{2}-4)e_{8},\quad v_{2}=4te_{6}+8e_{7}+(t^{2}-4)e_{5}.
  4. (4)

    The 22-dimensional subspace ZZ of VV generated by z1z_{1} and z2z_{2} with property [Z,Z]=0[Z,Z]=0.

  5. (5)

    The line H1H_{1} of [V,V]D2[V,V]\subseteq D^{2} with director vector

    h1\displaystyle h_{1} =(tx2+2y2)yx(ty2+2x2)xy\displaystyle=(tx^{2}+2y^{2})y\partial_{x}-(ty^{2}+2x^{2})x\partial_{y}
    =te14+2e162e17te19.\displaystyle=te_{14}+2e_{16}-2e_{17}-te_{19}.
  6. (6)

    The 44-dimensional subspace UU containing ZZ with basis z1,z2z_{1},z_{2} and

    (29) u1=t3e3+2e42t3e8,u2=t3e7+2e62t3e5,u_{1}=\frac{t}{3}e_{3}+2e_{4}-\frac{2t}{3}e_{8},\quad u_{2}=\frac{t}{3}e_{7}+2e_{6}-\frac{2t}{3}e_{5},

    which verifies the property Z=UVZ=U\cap V.

  7. (7)

    The line H2H_{2}, being the intersection of [U,U][U,U] and [V,Z][V,Z].

  8. (8)

    The complementary space VV^{*} of ZZ in VV, and H3=[V,V]H_{3}=[V^{*},V^{*}].

  9. (9)

    The complementary space UU^{*} of ZZ in UU, and H4:=[V,U]H1,H2H_{4}:=[V,U^{*}]\cap\langle H_{1},H_{2}\rangle.

The detailed constructs are given as follows:

(1) By restricting the Lie bracket to

[,]:D0×D3D3,[-,-]:D^{0}\times D^{3}\to D^{3},

we find that the vectors vD0v\in D^{0} subject to the property

[v,D3]=0[v,D^{3}]=0

are contained in e1\langle e_{1}\rangle. Therefore, e1\langle e_{1}\rangle is an invariant linear subspace.

(2) The vector e1e_{1} induces the linear map:

[e1,]:D1D1[e_{1},-]:D^{1}\to D^{1}

with eigenvalues 0,1,10,1,-1. Hence, there exists the Jordan decomposition (26), where the eigenspace D01,D11,D11D_{0}^{1},D_{1}^{1},D_{-1}^{1} correspond to the eigenvalues 0,1,10,1,-1.

(3) The subspace VD01V\subseteq D^{1}_{0} is defined as the collection of vectors vD01v\in D^{1}_{0} such that

[v,D11]=0.[v,D^{1}_{-1}]=0.

We can verify that VV is spanned by v1,v2,z1,z2v_{1},v_{2},z_{1},z_{2} in Equations (28) and (27).

(4) We define ZZ to be the collection of xVx\in V such that there exists some yVy\in V, linearly independent on xx, subject to

[x,y]=0.[x,y]=0.

Equivalently, for such xx, the kernel of [x,]:VV[x,-]:V\to V is at least 22-dimensional. It is clearly that ZZ is invariant up to isomorphism. Notice that [z1,z2]=0[z_{1},z_{2}]=0, so ZZ contains the linear subspace generated by z1,z2z_{1},z_{2}. The equality Z=z1,z2Z=\langle z_{1},z_{2}\rangle can be checked by observing that the two bilinear maps

[,]:V/Z×ZD2[-,-]:V/Z\times Z\to D^{2}

and

[,]:V/Z×V/ZD2/[V,Z][-,-]:V/Z\times V/Z\to D^{2}/[V,Z]

are non-degenerate.

(5) We define H1H_{1} as the collection of x[V,V]x\in[V,V] such that [x,V]=0[x,V]=0. A direct computation shows that H1H_{1} is 11-dimensional with basis h1h_{1}.

(6) It can be checked that

h1=[u2,z1][u1,z2]=(tx2y+2y3)x(ty2x+2x3)yh_{1}=[u_{2},z_{1}]-[u_{1},z_{2}]=(tx^{2}y+2y^{3})\partial_{x}-(ty^{2}x+2x^{3})\partial_{y}

where u1,u2D1u_{1},u_{2}\in D^{1} are defined in (29). Therefore, the space UU spanned by u1,u2,z1,z2u_{1},u_{2},z_{1},z_{2} is the minimal 44-dimensional space such that

H1[U,Z] and ZU.H_{1}\subseteq[U,Z]\text{ and }Z\subseteq U.

By construction, UU is also invariant.

(7) It is straightforward to show that

H2:=[U,U][V,Z]H_{2}:=[U,U]\cap[V,Z]

is 11-dimensional with basis

h2\displaystyle h_{2} =[v1,z2][v2,z1]=32[u1,u2]\displaystyle=[v_{1},z_{2}]-[v_{2},z_{1}]=\frac{3}{2}[u_{1},u_{2}]
=((t212)x2y4ty3)x((t212)y2x4tx3)y\displaystyle=((t^{2}-12)x^{2}y-4ty^{3})\partial_{x}-((t^{2}-12)y^{2}x-4tx^{3})\partial_{y}
=(t212)e154te16+4te17(t212)e18.\displaystyle=(t^{2}-12)e_{15}-4te_{16}+4te_{17}-(t^{2}-12)e_{18}.

(8) The space V={vV|[v,H2]=0}V^{*}=\{v\in V|[v,H_{2}]=0\} is an invariant subspace of VV with basis v1,v2v_{1},v_{2} as defined in (28). Then the linear subspace H3:=[V,V]H_{3}:=[V^{*},V^{*}] of D2D^{2} is spanned by

h3:=[v1,v2]=4t(t236)3h1t2+123h2.h_{3}:=[v_{1},v_{2}]=\frac{4t(t^{2}-36)}{3}h_{1}-\frac{t^{2}+12}{3}h_{2}.

(9) Similarly, the space U={uU|[u,H3]=0}U^{*}=\{u\in U|[u,H_{3}]=0\} is an invariant subspace of VV. The intersection of [V,U][V,U^{*}] and H1,H2\langle H_{1},H_{2}\rangle is spanned by the vector

h4=5(t2+12)2h1+4t(t236)h2.h_{4}=5(t^{2}+12)^{2}h_{1}+4t(t^{2}-36)h_{2}.

The line H4H_{4} spanned by h4h_{4} is obviously invariant. ∎

With the aid of these four invariant lines, we can establish the Torelli-type theorem for L1(E~7;t)L^{1}(\tilde{E}_{7};t).

Proof of part (3) in Theorem B.

From Lemma 14, the sequence of lines (H1,H2,H3,H4)(H_{1},H_{2},H_{3},H_{4}) is invariant under homogeneous isomorphisms. Moreover, the cross ratio of (H1,H2,H3,H4)(H_{1},H_{2},H_{3},H_{4}) is given by

516(t2+12)3t2(t236)2=516(t2+12)3(t2+12)3108(t24)2=516j(E~7)1j(E~7),-\frac{5}{16}\frac{(t^{2}+12)^{3}}{t^{2}(t^{2}-36)^{2}}=-\frac{5}{16}\frac{(t^{2}+12)^{3}}{(t^{2}+12)^{3}-108(t^{2}-4)^{2}}=\frac{5}{16}\frac{j(\tilde{E}_{7})}{1-j(\tilde{E}_{7})},

which aligns with the jj-function j(E~7)j(\tilde{E}_{7}). Therefore, the first Yau algebra determines the isomorphism classes of the E~7\tilde{E}_{7}-family. ∎

Finally, we would like to utilize the invariance of H1H_{1} to compute the groupoid Grp1(E~7)\operatorname{Grp}^{1}(\tilde{E}_{7}) for the case k=1k=1 of the assertion (1) in Theorem B.

Complete the proof of Theorem B.

It suffices to demonstrate that, when restricting to the coordinates (x,y)(x,y), the groupoid Grp(E~7)(t,s)\operatorname{Grp}^{\infty}(\tilde{E}_{7})(t,s) coincides with Grp1(E~7)(t,s)\operatorname{Grp}^{1}(\tilde{E}_{7})(t,s). Specifically, the restriction of a morphism of Grp1(E~7)(t,s)\operatorname{Grp}^{1}(\tilde{E}_{7})(t,s) to (x,y)(x,y) induces a homogeneous homomorphism of x4+y4+tx2y2x^{4}+y^{4}+tx^{2}y^{2}. Notice that this, together with Lemma 7, implies that Grpk(E~7)(t,s)\operatorname{Grp}^{k}(\tilde{E}_{7})(t,s) are the same up to transformations along the zz-axis.

Let ϕ\phi be a morphism in Grp1(E~7)(t,s)\operatorname{Grp}^{1}(\tilde{E}_{7})(t,s), represented as the matrix

(xy)=(abcd)(xy)\begin{pmatrix}x\\ y\end{pmatrix}=\begin{pmatrix}a&b\\ c&d\end{pmatrix}\begin{pmatrix}x^{\prime}\\ y^{\prime}\end{pmatrix}

in the coordinates (x,y)(x,y). Let Δ=abbc\Delta=ab-bc be the determinant. Then the inverse of ϕ\phi is given by

(xy)=1Δ(dbca)(xy).\begin{pmatrix}x^{\prime}\\ y^{\prime}\end{pmatrix}=\frac{1}{\Delta}\begin{pmatrix}d&-b\\ -c&a\end{pmatrix}\begin{pmatrix}x\\ y\end{pmatrix}.

Then, the partial derivatives can be expressed as

x=dΔxcΔy\partial_{x}=\frac{d}{\Delta}\partial_{x^{\prime}}-\frac{c}{\Delta}\partial_{y^{\prime}}

and

y=bΔx+aΔy.\partial_{y}=-\frac{b}{\Delta}\partial_{x^{\prime}}+\frac{a}{\Delta}\partial_{y^{\prime}}.

The image of the director vector h1h_{1} of H1H_{1} under ϕ\phi_{*} is given by

ϕ(h1)\displaystyle\phi_{*}(h_{1}) =(tx2y+2y3)(dΔxcΔy)(ty2x+2x3)(bΔx+aΔy)\displaystyle=(tx^{2}y+2y^{3})(\frac{d}{\Delta}\partial_{x^{\prime}}-\frac{c}{\Delta}\partial_{y^{\prime}})-(ty^{2}x+2x^{3})(-\frac{b}{\Delta}\partial_{x^{\prime}}+\frac{a}{\Delta}\partial_{y^{\prime}})
=(dΔ(tx2y+2y3)+bΔ(ty2x+2x3))x(cΔ(tx2y+2y3)+aΔ(ty2x+2x3))y.\displaystyle=(\frac{d}{\Delta}(tx^{2}y+2y^{3})+\frac{b}{\Delta}(ty^{2}x+2x^{3}))\partial_{x^{\prime}}-(\frac{c}{\Delta}(tx^{2}y+2y^{3})+\frac{a}{\Delta}(ty^{2}x+2x^{3}))\partial_{y^{\prime}}.

Since the line H1H_{1} is invariant under ϕ\phi_{*}, we may assume the equality

ϕ(h1(t))=λh1(s)\phi_{*}(h_{1}(t))=\lambda\cdot h_{1}(s)

holds for some constant λ\lambda. This is equivalent to

(dΔ(tx2y+2y3)+bΔ(ty2x+2x3))=λ(sx2y+2y3),(\frac{d}{\Delta}(tx^{2}y+2y^{3})+\frac{b}{\Delta}(ty^{2}x+2x^{3}))=\lambda(sx^{\prime 2}y^{\prime}+2y^{\prime 3}),

and

(cΔ(tx2y+2y3)+aΔ(ty2x+2x3))=λ(sy2x+2x3).(\frac{c}{\Delta}(tx^{2}y+2y^{3})+\frac{a}{\Delta}(ty^{2}x+2x^{3}))=\lambda(sy^{\prime 2}x^{\prime}+2x^{\prime 3}).

Therefore, we obtain:

2λ(x4+y4+sx2y2)=\displaystyle 2\lambda(x^{\prime 4}+y^{\prime 4}+sx^{\prime 2}y^{\prime 2})= λ(sx2y+2y3)y+λ(sy2x+2x3)x\displaystyle\lambda(sx^{\prime 2}y^{\prime}+2y^{\prime 3})y^{\prime}+\lambda(sy^{\prime 2}x^{\prime}+2x^{\prime 3})x^{\prime}
=\displaystyle= (tx2y+2y3)y+(ty2x+2x3)x\displaystyle(tx^{2}y+2y^{3})y+(ty^{2}x+2x^{3})x
=\displaystyle= 2(x4+y4+tx2y2).\displaystyle 2(x^{4}+y^{4}+tx^{2}y^{2}).

It follows that ϕ\phi comes from Grp(E~7)(t,s)\operatorname{Grp}^{\infty}(\tilde{E}_{7})(t,s). Therefore, the case k=1k=1 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 kk-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 kk-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 E~6\tilde{E}_{6} 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.