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

Formal deformations of algebraic spaces and generalizations of the motivic Igusa-zeta function.

Andrew R. Stout Borough of Manhattan Community College, CUNY 199 Chambers Street
New York, NY 10007 [email protected]
Abstract.

We generalize the notion of the auto-Igusa zeta function to formal deformations of algebraic spaces. By incorporating data from all algebraic transformations of local coordinates, this function can be viewed as a generalization of the traditional motivic Igusa zeta function. Furthermore, we introduce a new series, which we term the canonical auto-Igusa zeta function, whose coefficients are given by the quotient stacks formed from the coefficients of the auto-Igusa zeta function modulo change of coordinates. We indicate the current state of the literature on these generalized Igusa-zeta functions and offer directions for future research.

2010 Mathematics Subject Classification:
Primary 14H20, Secondary 14H50, 14E18.
Support for this project was provided by a PSC-CUNY Award (PSC-Grant Traditional B, # 60784-00 48), jointly funded by the Professional Staff Congress and The City University of New York.

1. Introduction

Progress has been made recently in understanding certain motivic generating series associated to the germ of a variety. Originally introduced by Schoutens in [Scha] and [Schb], the auto-Igusa zeta series is a motivic generating series whose coefficients are determined by the algebra automorphisms of the local ring corresponding to a point on a variety. First calculations of this series were completed by Schoutens and secondary calculations were carried out by the author of this paper with the use of computer algebra software in [Sto17]. This led to several conjectures concerning the structure of the auto-Igusa zeta function. The author later confirmed that these conjectures are indeed true in [Sto18] for plane curve singularities, yet the veracity of these conjectures for broader classes of singularities remains unknown.

In this paper, the author generalizes the auto-Igusa zeta function to stacks. However, we choose to frame almost all of the work in the language of algebraic spaces, as there is a bottleneck for Artin stacks related to the occasional lack of representability of the hom functor for general Artin stacks. The representability results are discussed in §2.

One notable development in this paper concerns the introduction of a new type of moduli stack, which takes into account the canonical group action on the space of endormophism by conjugation of all automorphisms. We call this the moduli stack of Jordan-Normal forms. This space is in many ways more natural than the space of all endomrophism (from which the auto-Igusa zeta series is formed), yet it is unclear to the author under what conditions this moduli stack is itself an algebraic space or even when it is an Artin stack. Regardless, one may form a generating series from this stack, which we then term the canonical auto-Igusa zeta function.

Another important development in this paper is that it is shown in §6 that both the canonical zeta function and the auto-Igusa zeta function may both be regarded as generalizations of the motivic Igusa zeta function, which has been studied by Denef, Loeser, Cluckers, Nicaise, Mustata, and many others. It is hoped that the motivic Igusa zeta series would offer a key to proving the monodromy conjecture first introduced by Igusa. However, as of yet, the general conjecture remains unproven.

In §7, the author rephrases his work in [Sto18] in this new context and offers possible directions for future research. As mentioned at the beginning of this section, the possible rationality of the auto-Igusa zeta function for varieties (and other conjectures) other than plane curves remains mostly elusive to direct proof. It is the hope of the author that the work carried out in this paper will offer progress toward such a proof as well as possible keys to further understanding of local zeta functions in general.

2. Representability Results Concerning Hom

Let 𝐂\mathbf{C} be a site. It follows from the Yoneda lemma that the Grothendieck topos τ=Sh(𝐂)\tau=\mbox{Sh}(\mathbf{C}) is a locally cartesian closed monoidal category. In particular, given two objects 𝒳\mathcal{X} and 𝒴\mathcal{Y} of τ\tau, we have a presheaf which sends an object UU of 𝐂\mathbf{C} to the set Homτ(𝒳U,𝒴U)\mbox{Hom}_{\tau}(\mathcal{X}\otimes U,\mathcal{Y}\otimes U), and, moreover, this presheaf is in fact a sheaf – i.e., it is an object of τ\tau, which we denote by Hom¯τ(𝒳,𝒴)\underline{\mbox{Hom}}_{\tau}(\mathcal{X},\mathcal{Y}). In other words, there is a natural bijection

(2.1) Homτ(𝒳U,𝒴U)Homτ(U,Hom¯τ(𝒳,𝒴)).\mbox{Hom}_{\tau}(\mathcal{X}\otimes U,\mathcal{Y}\otimes U)\xrightarrow{\sim}\mbox{Hom}_{\tau}(U,\underline{\mbox{Hom}}_{\tau}(\mathcal{X},\mathcal{Y})).

In fact, a similar result holds in higher category theory–i.e., if τ\tau is an (,1)(\infty,1)-topos, then τ\tau is a locally cartesian closed monoidal (,1)(\infty,1)-category. This follows from the (,1)(\infty,1)-Yoneda lemma (cf. Proposition 5.1.3.1 of [Lur09]). In particular, if 𝒳\mathcal{X} and 𝒴\mathcal{Y} are two stacks over an algebraic space SS, then one defines Hom¯S(𝒳,𝒴)\underline{\mbox{Hom}}_{S}(\mathcal{X},\mathcal{Y}) to be the fibered category of groupoids over the site of SS-schemes (𝐒𝐜𝐡/S)J(\mathbf{Sch}/S)_{J} with a given Grothendieck topology JJ, which to any obect USU\to S of (𝐒𝐜𝐡/S)J(\mathbf{Sch}/S)_{J} associates the groupoid of functors from 𝒳×SU\mathcal{X}\times_{S}U to 𝒴×SU\mathcal{Y}\times_{S}U over UU, and, moreover, this fibered category Hom¯S(𝒳,𝒴)\underline{\mbox{Hom}}_{S}(\mathcal{X},\mathcal{Y}) is in fact a stack. It is possible to show that Hom¯S(𝒳,𝒴)\underline{\mbox{Hom}}_{S}(\mathcal{X},\mathcal{Y}) is an Artin stack (or, Deligne-Mumford stack, or algebraic space) depending on certain conditions on 𝒳\mathcal{X} and 𝒴\mathcal{Y} (cf. Theorem 1.1 of [Ols06]). For example, we have the following result due to Artin.

Theorem 2.1.

If SS is a locally Noetherian algebraic space and if XX and YY are algebraic spaces over SS with XSX\to S proper and flat and YSY\to S is separated and of finite type, then Hom¯S(X,Y)\underline{\mbox{Hom}}_{S}(X,Y) is a separated algebraic space locally of finite type over SS.

This result can be found in [Art69], and generalizes111Essentially, Artin relaxed the condition that XSX\to S is projective to merely proper, but this came at the cost of moving from the category of schemes to the larger category of algebraic spaces. the following result of Grothendieck. If SS is a locally Noetherian scheme and if XX and YY are in 𝐒𝐜𝐡/S\mathbf{Sch}/S with XX projective and flat over SS and YY quasi-projective over SS, then the functor from 𝐒𝐜𝐡/S\mathbf{Sch}/S to 𝐒𝐞𝐭𝐬\mathbf{Sets} defined by sending UU to HomS(X×SU,Y×SU)\mbox{Hom}_{S}(X\times_{S}U,Y\times_{S}U) is represented by a scheme Hom¯S(X,Y)\underline{\mbox{Hom}}_{S}(X,Y) which is a separated and locally of finite type over SS. This result can be found at the end of page 221-19 of [Gro62].

In particular, if XX is finite and flat over a locally Noetherian scheme SS, then for any quasi-projective object YY in 𝐒𝐜𝐡/S\mathbf{Sch}/S, Hom¯S(X,Y)\underline{\mbox{Hom}}_{S}(X,Y) is a scheme over SS. This is found in Proposition 5.7 on pages 221-27 of [Gro62], and it usually goes by the name Weil restriction or restriction of scalars. If YSY\to S is separated (respectively, affine, finite type), then Hom¯S(X,Y)S\underline{\mbox{Hom}}_{S}(X,Y)\to S is separated (respectively, affine, finite type) over SS (cf. Section 7 on pages 221-27 and 221-28 of [Gro62]).

Remark 2.2.

In fact, regarding either Artin’s result or Grothendieck’s result, the assumption that SS is locally Noetherian can be relaxed if we assume that XSX\to S and YSY\to S are locally of finite presentation. Note that a finite type morphism with a locally Noetherian target is locally of finite presentation.

3. Projective Systems of Hilbert Spaces

Let XSX\to S be a proper and flat morphism of algebraic spaces with SS locally Noetherian. There is a functor HilbX/S:𝐒𝐜𝐡/S𝐒𝐞𝐭𝐬\mbox{Hilb}_{X/S}:\mathbf{Sch}/S\to\mathbf{Sets} defined by

(3.1) HilbX/S(T)={[j:ZX×ST]ZT proper and flat, and j a closed immersion},\mbox{Hilb}_{X/S}(T)=\{[j:Z\hookrightarrow X\times_{S}T]\mid Z\to T\mbox{ proper and flat, and }j\mbox{ a closed immersion}\},

where [j:ZX×ST[j:Z\hookrightarrow X\times_{S}T denotes the isomorphism class of jj. This is in fact represented by a separated algebraic space locally of finite type over SS (cf. Section 6 of [Art69]) which we denote by Hilb¯X/S\underline{\mbox{Hilb}}_{X/S}. The natural transformation ι:HomS(X,Y)Hilb(X×SY)/S\iota:\mbox{Hom}_{S}(X,Y)\hookrightarrow\mbox{Hilb}_{(X\times_{S}Y)/S} defined by

(3.2) ι(T):HomS(X,Y)(T)Hilb(X×SY)/S(T)[f:X×STY×ST][ΓfX×S×Y×ST]\begin{split}\iota(T):\mbox{Hom}_{S}(X,Y)(T)&\hookrightarrow\mbox{Hilb}_{(X\times_{S}Y)/S}(T)\\ [f:X\times_{S}T\to Y\times_{S}T]&\mapsto[\Gamma_{f}\hookrightarrow X\times_{S}\times Y\times_{S}T]\end{split}

where Γf\Gamma_{f} denotes the graph of ff induces an open immersion of algebraic spaces Hom¯S(X,Y)Hilb¯(X×SY)/S\underline{\mbox{Hom}}_{S}(X,Y)\hookrightarrow\underline{\mbox{Hilb}}_{(X\times_{S}Y)/S} provided XSX\to S is proper and flat and YSY\to S is separated and of finite type.

Now, if f:S′′Sf:S^{\prime\prime}\rightarrow S^{\prime} is an SS-morphism of algebraic spaces, then there is a natural pullback ff^{*} given by the natural transformation f:HilbS/SHilbS′′/Sf^{*}:\mbox{Hilb}_{S^{\prime}/S}\to\mbox{Hilb}_{S^{\prime\prime}/S} defined by

(3.3) f(T):HilbS/S(T)HilbS′′/S(T)ZZ×S×ST(S′′×ST).\begin{split}f^{*}(T):\mbox{Hilb}_{S^{\prime}/S}(T)&\to\mbox{Hilb}_{S^{\prime\prime}/S}(T)\\ Z&\mapsto Z\times_{S^{\prime}\times_{S}T}(S^{\prime\prime}\times_{S}T).\end{split}

Given two SS-morphisms f:XXf:X^{\prime}\to X and g:YYg:Y^{\prime}\to Y of algebraic spaces with XSX^{\prime}\to S and XSX\to S proper and flat and with YSY^{\prime}\to S and YSY\to S separated and finite type, we have the canonical morphism (f,g):X×SYX×SY(f,g):X^{\prime}\times_{S}Y^{\prime}\to X\times_{S}Y and the corresponding pullback morphism of Hilbert spaces

(3.4) (f,g):Hilb¯X×SY/SHilb¯X×SY/S.(f,g)^{*}:\underline{\mbox{Hilb}}_{X\times_{S}Y/S}\to\underline{\mbox{Hilb}}_{X^{\prime}\times_{S}Y^{\prime}/S}.

If, in addition, g:YYg:Y^{\prime}\hookrightarrow Y is a closed immersion, then the restriction of (f,g)(f,g)^{*} to Hom¯(X,Y)\underline{\mbox{Hom}}(X,Y) induces a well-defined morphism of algebraic spaces

(3.5) (f,g):Hom¯S(X,Y)Hom¯S(X,Y).(f,g)^{*}:\underline{\mbox{Hom}}_{S}(X,Y)\to\underline{\mbox{Hom}}_{S}(X^{\prime},Y^{\prime}).

Note that Hom¯S(X,Y)\underline{\mbox{Hom}}_{S}(X,Y) is contravariant in the first argument and covariant in the second; however, the morphism (f,g)(f,g)^{*} sends a morphism h:XYh:X\to Y to hf:XYh\circ f:X^{\prime}\to Y and then restricts the range of hfh\circ f to the closed subset YY^{\prime} of YY.

An injective system {Pi}iI\{P_{i}\}_{i\in I} of algebraic spaces over SS with affine transition maps and with PiSP_{i}\to S proper and flat gives rise to a projective system {Hilb¯Pi/S}iI\{\underline{\mbox{Hilb}}_{P_{i}/S}\}_{i\in I} of algebraic spaces with affine transition maps, and it is therefore the case that limHilb¯Pi/S\varprojlim\underline{\mbox{Hilb}}_{P_{i}/S} exists in the category of algebraic spaces over SS. In particular, if {Yi}iI\{Y_{i}\}_{i\in I} is an injective system whose transition maps are closed immersions with YiSY_{i}\to S separated and finite type and if {Xi}iI\{X_{i}\}_{i\in I} is any other injective system with XiSX_{i}\to S proper and flat, then limHom¯S(Xi,Yi)limHilb¯Xi×SYi/S\varprojlim\underline{\mbox{Hom}}_{S}(X_{i},Y_{i})\hookrightarrow\varprojlim\underline{\mbox{Hilb}}_{X_{i}\times_{S}Y_{i}/S} is an open immersion of algebraic spaces over SS.

Furthermore, if {Xi}iI\{X_{i}\}_{i\in I} is an injective system of algebraic spaces over SS whose transitions maps are closed immersions and where each XiSX_{i}\to S is proper and flat then there is a projective system of open immersions

(3.6) Aut¯S(Xi)End¯S(Xi)\underline{\mbox{Aut}}_{S}(X_{i})\hookrightarrow\underline{\mbox{End}}_{S}(X_{i})

where Aut¯(Xi)\underline{\mbox{Aut}}(X_{i}) is the algebraic space which represents the subfunctor of HomS(Xi,Xi)\mbox{Hom}_{S}(X_{i},X_{i}) which associates to TST\to S the set of all automorphisms of Xi×STX_{i}\times_{S}T. Here, we write End¯S(Xi)\underline{\mbox{End}}_{S}(X_{i}) in place of Hom¯S(Xi,Xi).\underline{\mbox{Hom}}_{S}(X_{i},X_{i}). Note that End¯S(Xi)\underline{\mbox{End}}_{S}(X_{i}) is a monoid object and Aut¯S(Xi)\underline{\mbox{Aut}}_{S}(X_{i}) is a group object in the category of algebraic spaces over SS. Since the transition maps are closed immersions, the inclusion

(3.7) limAut¯S(Xi)limEnd¯S(Xi)\varprojlim\underline{\mbox{Aut}}_{S}(X_{i})\hookrightarrow\varprojlim\underline{\mbox{End}}_{S}(X_{i})

is an open immersion of algebraic spaces over SS.

Remark 3.1.

As in Remark 2.2, we may relax the condition that SS is locally Noetherian in this section by assuming that all morphisms are locally of finite presentation.

4. Finite Free Algebras

Lemma 4.1.

Let XSX\to S be a finite and flat morphism of algebraic spaces with SS locally Noetherian. Then, XSX\to S is finite locally free. In particular, XSX\to S is faithfully flat.

Proof.

Note that flat is local on the source and target in the étale topology (cf. Remark 5.4.13 of [Ols16]). Therefore, it is enough to check when SS is affine by Lemma 45.3 of [Sta17]. Since XSX\to S is finite and hence also an affine morphism, XX is affine. Hence, we reduce to the case where (A,𝔪)(A,\mathfrak{m}) is a Noetherian local ring with residue field k=A/𝔪k=A/\mathfrak{m} and a flat morphism ABA\to B, which induces a surjective morphism AnBA^{n}\to B of AA-modules (by finiteness). Therefore, we have a short exact sequence

0IAnB00\to I\to A^{n}\to B\to 0

where II is finitely generated (since AA is Noetherian). Tensoring this short exact sequence by kk gives the short exact sequence

0IAkknB/𝔪B00\to I\otimes_{A}k\to k^{n}\to B/\mathfrak{m}B\to 0

since BB is flat over AA. Since B/𝔪BB/\mathfrak{m}B is a free kk-module, we may choose nn above (i.e., generators b1,,bnb_{1},\ldots,b_{n} of BB over AA) so that B/𝔪BB/\mathfrak{m}B is isomorphic to knk^{n}, which implies that I/𝔪I=IAk=0I/\mathfrak{m}I=I\otimes_{A}k=0. Therefore, by Nakayama’s Lemma (cf, Part (2) of Lemma 10.19.1 of [Sta17]), I=0I=0, and BB is a free AA-module. ∎

Remark 4.2.

The Noetherian condition on SS in Lemma 4.1 may be relaxed. In fact, the morphism XSX\to S of algebraic spaces is a finite, flat, and locally of finite presentation if and only if it is a finite locally free morphism. This follows from 1.4.7 of [GD67], Lemma 45.3 [Sta17], and the fact that flatness and locally of finite presentation are both local on the source and target in the étale topology (cf. Remark 5.4.13 of [Ols16]).

Assume that AA is a Noetherian local ring with BB finite and flat over AA, and whence free. Let AnBA^{n}\xrightarrow{\sim}B be a presentation of BB as a free AA-algebra obtained by sending the standard basis elements eie_{i} of AnA^{n} to the generating elements bib_{i} of BB for i=1,,ni=1,\ldots,n. Consider the polynomial ring R=A[x1,,xn]R=A[x_{1},\ldots,x_{n}] and the surjective homomorphism ϕ:RB\phi:R\to B defined by ϕ(xi)=bi\phi(x_{i})=b_{i} for i=1,,ni=1,\ldots,n. Then, by Hilbert’s Nullstellensatz, RR is Noetherian and therefore the kernel Ker(ϕ)\mbox{Ker}(\phi) is finitely generated – i.e., Ker(ϕ)=(f1,,fs)\mbox{Ker}(\phi)=(f_{1},\ldots,f_{s}). Moreover, there is a canonical isomorphism R/Ker(ϕ)BR/\mbox{Ker}(\phi)\xrightarrow{\sim}B in the category of AA-algebras. Therefore, an AA-algebra endomorphism γ\gamma of BB is given by a choice γ(xi)=Pi\gamma(x_{i})=P_{i} with PiP_{i} any element of B=A[x1,,xn]/(f1,,fs)B=A[x_{1},\ldots,x_{n}]/(f_{1},\ldots,f_{s}) for each i=1,,ni=1,\ldots,n such that fj(P1,,Pn)=0f_{j}(P_{1},\ldots,P_{n})=0 in BB for j=1,,sj=1,\ldots,s.

Lemma 4.3.

Let S=Spec(A)S=\mbox{Spec}(A) with AA is a reduced Noetherian local ring and let X=Spec(B)X=\mbox{Spec}(B) with B=A[x1,,xd]/(x1,,xd)nB=A[x_{1},\ldots,x_{d}]/(x_{1},\ldots,x_{d})^{n}. Set r=d(1)r=d\cdot(\ell-1), where \ell is the (generic) rank of BB over AA. Then, XSX\to S is finite and flat, and we have the following isomorphisms

(4.1) (End¯S(X))red𝔸Srand(Aut¯S(X))redGLd,S×S𝔸Srd2.(\underline{\mbox{End}}_{S}(X))^{\mathrm{red}}\cong\mathbb{A}_{S}^{r}\quad\mbox{and}\quad(\underline{\mbox{Aut}}_{S}(X))^{\mathrm{red}}\cong\mbox{GL}_{d,S}\times_{S}\mathbb{A}_{S}^{r-d^{2}}.
Proof.

Consider the AA-algebra endomorphism of A[x1,,xd]/(x1,,xd)nA[x_{1},\ldots,x_{d}]/(x_{1},\ldots,x_{d})^{n} defined by xiPix_{i}\mapsto P_{i}, where we define Pi=|j|<nai,jxjP_{i}=\sum_{|j|<n}a_{i,j}x^{j} where jj is a multi-index (i.e., j=(j1,,jd)j=(j_{1},\ldots,j_{d}), xj=s=1dxsjsx^{j}=\prod_{s=1}^{d}x_{s}^{j_{s}}, and |j|=s=1djs<n|j|=\sum_{s=1}^{d}j_{s}<n) and where ai,jAa_{i,j}\in A for i=1,,di=1,\ldots,d. Then, the equations defining the space of SS-endomorphisms of XX are given by

0=Pin=(ai,0+Qi)n,i=1,,d0=P_{i}^{n}=(a_{i,0}+Q_{i})^{n},\quad\forall i=1,\ldots,d

where 0 is treated as the multi-index (0,,0)(0,\ldots,0) and Qi(x1,,xd)/(x1,,xd)nQ_{i}\in(x_{1},\ldots,x_{d})/(x_{1},\ldots,x_{d})^{n}. This implies that 0=(ai,0)n0=(a_{i,0})^{n} for all i=1,,di=1,\ldots,d. Thus, in the reduction, 0=ai,00=a_{i,0} for all i=1,,di=1,\ldots,d since AA is reduced. Clearly, Qn=0Q^{n}=0 for all Q(x1,,xd)/(x1,,xd)nQ\in(x_{1},\ldots,x_{d})/(x_{1},\ldots,x_{d})^{n} from which the first isomorphism follows.

Note that those endomorphisms above which also lie in (Aut¯S(X))red(\underline{\mbox{Aut}}_{S}(X))^{\mathrm{red}} are given by xij=1dai,jxj+Tix_{i}\mapsto\sum_{j=1}^{d}a_{i,j}x_{j}+T_{i} with Ti(x1,,xd)2/(x1,,xd)nT_{i}\in(x_{1},\ldots,x_{d})^{2}/(x_{1},\ldots,x_{d})^{n} and where the d×dd\times d-matrix M=(ai,j)M=(a_{i,j}) is invertible over AA – i.e., det(M)\mbox{det}(M) is a unit of AA. The second isomorphism then follows from this fact. ∎

Remark 4.4.

In the case where d=1d=1, we obtain

(4.2) (End¯S(X))red𝔸Sn1and(Aut¯S(X))red𝔾m,S×S𝔸Sn2.(\underline{\mbox{End}}_{S}(X))^{\mathrm{red}}\cong\mathbb{A}_{S}^{n-1}\quad\mbox{and}\quad(\underline{\mbox{Aut}}_{S}(X))^{\mathrm{red}}\cong\mathbb{G}_{m,S}\times_{S}\mathbb{A}_{S}^{n-2}.

One simple case where Lemma 4.1 may be applied is when S=Spec(A)S=\mbox{Spec}(A) with (A,𝔪)(A,\mathfrak{m}) an Artinian local ring. This is because if BB is a finitely generated AA-module with AA an Artinian local ring, then BB will have finite length over the residue field k=A/𝔪k=A/\mathfrak{m} and hence BB will also be an Artinian local ring. In particular, BB is automatically free over AA. Thus, we may form the full subcategory 𝐅𝐚𝐭/A\mathbf{Fat}/A of 𝐒𝐜𝐡/A\mathbf{Sch}/A defined by all finite maps XSpec(A)X\to\mbox{Spec}(A). We call222In general, if SS is an algebraic space, we call the full subcategory 𝐂\mathbf{C} of 𝐒𝐜𝐡/S\mathbf{Sch}/S whose objects are finite, flat, and locally finitely presented over SS the category of fat points over SS, and we denote this category by 𝐅𝐚𝐭/S\mathbf{Fat}/S. this category the category of fat points over AA. Usually, we assume A=kA=k is a field (and, in this case, Lemma 4.3 will apply), and often, we will further assume that it has characteristic zero and that it is algebraically closed.

In general, if XX is proper and flat over an algebraic space SS, then the group action

(4.3) End¯S(X)red×SAut¯S(X)redEnd¯S(X)red\underline{\mbox{End}}_{S}(X)^{\mathrm{red}}\times_{S}\underline{\mbox{Aut}}_{S}(X)^{\mathrm{red}}\to\underline{\mbox{End}}_{S}(X)^{\mathrm{red}}

given by conjugation is well-defined in the category of algebraic spaces over SS. The resulting quotient stack

(4.4) JN(X):=[End¯S(X)red/Aut¯S(X)red]\mathcal{M}_{\mathrm{JN}}(X):=[\underline{\mbox{End}}_{S}(X)^{\mathrm{red}}/\underline{\mbox{Aut}}_{S}(X)^{\mathrm{red}}]

is termed the moduli stack of Jordan norm forms of XX over SS, and it is of general interest. In the case of Lemma 4.3, assuming further that AA is an algebraically closed field kk, the points of JN(X)\mathcal{M}_{\mathrm{JN}}(X) correspond to d×dd\times d Jordan normal forms over kk.

5. Auto-Arc Spaces of Formal Deformations

Definition 5.1.

Let SS be an algebraic space and let II be a directed countable set. Let {Xi}iI\{X_{i}\}_{i\in I} be an injective system of objects of 𝐅𝐚𝐭/S\mathbf{Fat}/S such that all transition maps XiXjX_{i}\hookrightarrow X_{j} are closed immersions. We say such a system is an admissible system of fat points over SS if

1) the structure morphism j:X0Sj:X_{0}\to S is surjective étale and

2) the underlying topological spaces of X0X_{0} and XiX_{i} are homeomorphic (iI)(\forall i\in I).

Definition 5.2.

We say that a system fi:YiYi+1f_{i}:Y_{i}\hookrightarrow Y_{i+1} of closed SS-immersions is a formal deformation over an algebraic space SS if there is an admissible system of fat points {Xi}iI\{X_{i}\}_{i\in I} over SS such that for all ii in II, there are flat morphisms φi:YiXi\varphi_{i}:Y_{i}\to X_{i} which make the following diagram commutative:

Y0\textstyle{Y_{0}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}φ0\scriptstyle{\varphi_{0}}f0\scriptstyle{f_{0}}Y1\textstyle{Y_{1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}φ1\scriptstyle{\varphi_{1}}f1\scriptstyle{f_{1}}Y2\textstyle{Y_{2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}φ2\scriptstyle{\varphi_{2}}f2\scriptstyle{f_{2}}\textstyle{\cdots}X0\textstyle{X_{0}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}j\scriptstyle{j}X1\textstyle{X_{1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}X2\textstyle{X_{2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\textstyle{\cdots}S\textstyle{S}

and where φi\varphi_{i} induces an isomorphism

(5.1) Yi1Yi×XiXi1.Y_{i-1}\cong Y_{i}\times_{X_{i}}X_{i-1}.

Given an admissible system {Xi}iI\{X_{i}\}_{i\in I} over SS and a morphism YSY\to S of algebraic spaces, we have the trivial formal deformation given by Yi:=Y×SXiY_{i}:=Y\times_{S}X_{i}. More generally, given an ind-object 𝒴\mathcal{Y} in the category of algebraic spaces over SS – i.e., the filtered colimit 𝒴=limiIYi\mathcal{Y}=\varinjlim_{i\in I}Y_{i} where the transition maps YiYi+1Y_{i}\hookrightarrow Y_{i+1} are closed immersions of algebraic spaces – and, an admissible system {Xi}iI\{X_{i}\}_{i\in I} with 𝒳=limiIXi\mathcal{X}=\varinjlim_{i\in I}X_{i}, we let 𝙳𝚎𝚏𝒳\mathtt{Def}_{\mathcal{X}} denote the fibered category of formal deformations over SS with respect to the admissible system {Xi}\{X_{i}\}. Thus, in particular, if YSY\to S is a morphism of algebraic space then the category 𝙳𝚎𝚏𝒳(Y)\mathtt{Def}_{\mathcal{X}}(Y) has at least one object.

An object 𝒴=limiIYi\mathcal{Y}=\varprojlim_{i\in I}Y_{i} of 𝙳𝚎𝚏𝒳\mathtt{Def}_{\mathcal{X}} gives rise to a projective system {𝒜i(𝒴)}iI\{\mathcal{A}_{i}(\mathcal{Y})\}_{i\in I} of algebraic spaces over SS defined by

(5.2) 𝒜i(𝒴):=Hom¯S(Xi,Yi)red.\mathcal{A}_{i}(\mathcal{Y}):=\underline{\mbox{Hom}}_{S}(X_{i},Y_{i})^{\mathrm{red}}.

We call 𝒜i(𝒴)\mathcal{A}_{i}(\mathcal{Y}) the truncated auto-arc space of 𝒴\mathcal{Y} at level ii. Moreover, we may form the projective limit

(5.3) 𝒜(𝒴)=limiI𝒜i(𝒴),\mathcal{A}(\mathcal{Y})=\varprojlim_{i\in I}\mathcal{A}_{i}(\mathcal{Y}),

which is termed the infinite auto-arc space of 𝒴\mathcal{Y}. If 𝒴\mathcal{Y} is the trivial deformation of YSY\to S, then

(5.4) 𝒜i(𝒴)=(Hom¯S(Xi,Y)red×SredEnd¯S(Xi)red)red\mathcal{A}_{i}(\mathcal{Y})=(\underline{\mbox{Hom}}_{S}(X_{i},Y)^{\mathrm{red}}\times_{S^{\mathrm{red}}}\underline{\mbox{End}}_{S}(X_{i})^{\mathrm{red}})^{\mathrm{red}}

In general, the group scheme Aut¯S(Xi)red\underline{\mbox{Aut}}_{S}(X_{i})^{\mathrm{red}} acts on 𝒜i(𝒴)\mathcal{A}_{i}(\mathcal{Y}) by conjugation for any formal deformation 𝒴\mathcal{Y}, and this give rise to the quotient stack

(5.5) i(𝒴):=[𝒜i(𝒴)/Aut¯S(Xi)red].\mathcal{M}_{i}(\mathcal{Y}):=[\mathcal{A}_{i}(\mathcal{Y})/\underline{\mbox{Aut}}_{S}(X_{i})^{\mathrm{red}}].

In particular, when 𝒴\mathcal{Y} is the trivial deformation, then

(5.6) i(𝒴)𝒜i(Y)×SredJN(Xi),\mathcal{M}_{i}(\mathcal{Y})\cong\mathcal{A}_{i}(Y)\times_{S^{\mathrm{red}}}\mathcal{M}_{JN}(X_{i}),

where JN(Xi)\mathcal{M}_{JN}(X_{i}) is the moduli stack of Jordan normal forms introduced at the end of section 4. The projective limit of quotient stacks

(5.7) (𝒴)=limiIi(𝒴)\mathcal{M}(\mathcal{Y})=\varprojlim_{i\in I}\mathcal{M}_{i}(\mathcal{Y})

is of general interest.

6. A generalization of the motivic Igusa zeta function

Definition 6.1.

Let SS be an algebraic space. Let 𝐀𝐥𝐠/S\mathbf{Alg}/S denote the category of finitely presented algebraic spaces over SS. The Grothendieck ring of algebraic spaces over SS, denoted by K0(𝐀𝐥𝐠/S)\mbox{K}_{0}(\mathbf{Alg}/S), is the ring formed by introducing relations on the free abelian group of isomorphism classes X\langle X\rangle of objects XX of 𝐀𝐥𝐠/S\mathbf{Alg}/S:

1. X=XY+Y whenever YX is a locally closed S-immersion.\langle X\rangle=\langle X\setminus Y\rangle+\langle Y\rangle\mbox{ whenever }Y\hookrightarrow X\mbox{ is a locally closed }S\mbox{-immersion.}

2. Z=X×S𝔸Sn whenever ZX is a vector bundle of constant rank n\langle Z\rangle=\langle X\times_{S}\mathbb{A}_{S}^{n}\rangle\mbox{ whenever }Z\to X\mbox{ is a vector bundle of constant rank }n. We denote the class of an algebraic space XX in K0(𝐀𝐥𝐠/S)K_{0}(\mathbf{Alg}/S) by [X][X]. The multiplicative structure is defined by [X][Y]:=[X×SY][X]\cdot[Y]:=[X\times_{S}Y].

Remark 6.2.

Note that when S=Spec(k)S=\mbox{Spec}(k) with kk a field, then K0(𝐀𝐥𝐠/S)\mbox{K}_{0}(\mathbf{Alg}/S) is isomorphic to the Grothendieck ring of algebraic varieties K0(𝐕𝐚𝐫/S)\mbox{K}_{0}(\mathbf{Var}/S). This isomorphism is induced by the natural inclusion of sets 𝐕𝐚𝐫/S𝐀𝐥𝐠/S\mathbf{Var}/S\hookrightarrow\mathbf{Alg}/S. Moreover, relations of the form given in 2 above are superfluous in this case (cf., Section 1 of [Eke09]).

We define 𝕃:=[𝔸S1]\mathbb{L}:=[\mathbb{A}_{S}^{1}] and call it the Leftshetz motive over SS. We may invert this element to obtain the localized Grothendieck ring 𝒢S:=K0(𝐀𝐥𝐠/S)[𝕃1]\mathcal{G}_{S}:=\mbox{K}_{0}(\mathbf{Alg}/S)[\mathbb{L}^{-1}] of algebraic spaces over SS. We say that an element of the power series ring 𝒢S[[t]]\mathcal{G}_{S}[[t]] is a motivic generating series of algebraic spaces over SS in one variable, or, more briefly, we will call it a motivic generating series.

Definition 6.3.

Let 𝒳=limiIXi\mathcal{X}=\varinjlim_{i\in I}X_{i} be given by an admissible system of fat points {X}iI\{X\}_{i\in I} over SS. Let 𝒴=limiIYi\mathcal{Y}=\varinjlim_{i\in I}Y_{i} be an object of 𝙳𝚎𝚏X\mathtt{Def}_{X} with YiY_{i} an object of 𝐀𝐥𝐠/S\mathbf{Alg}/S for all iIi\in I. We define the auto-Igusa zeta function of 𝒴\mathcal{Y} to be the motivic generating series given by

(6.1) ζ𝒴(t)=iI[𝒜i(𝒴)]𝕃niti\zeta_{\mathcal{Y}}(t)=\sum_{i\in I}[\mathcal{A}_{i}(\mathcal{Y})]\mathbb{L}^{-n_{i}}t^{i}

where ni=dimS(Yi)(rankS(Xi)1)+ei(rankS(Xi)1)n_{i}=\mbox{dim}_{S}(Y_{i})(\mbox{rank}_{S}(X_{i})-1)+e_{i}(\mbox{rank}_{S}(X_{i})-1) where eie_{i} is the whole number which makes the coefficient [𝒜i(𝒴)]𝕃ni[\mathcal{A}_{i}(\mathcal{Y})]\mathbb{L}^{-n_{i}} dimensionless333It is anticipated that ei=rankS(ΩX0/S)e_{i}=\mbox{rank}_{S}(\Omega_{X_{0}/S}) for all iIi\in I..

Let 𝒴\mathcal{Y} be as in Definition 6.3 and assume further that i(𝒴)\mathcal{M}_{i}(\mathcal{Y}) is an algebraic space for all iIi\in I, we may form the motivic generating function

(6.2) η𝒴(t)=iI[i(𝒴)]𝕃miti\eta_{\mathcal{Y}}(t)=\sum_{i\in I}[\mathcal{M}_{i}(\mathcal{Y})]\mathbb{L}^{-m_{i}}t^{i}

where mi=dimS(Yi)(rankS(Xi)1)+eim_{i}=\mbox{dim}_{S}(Y_{i})(\mbox{rank}_{S}(X_{i})-1)+e_{i} where eie_{i} is chosen to be the whole number which makes the coefficient [i(𝒴)]𝕃mi[\mathcal{M}_{i}(\mathcal{Y})]\mathbb{L}^{-m_{i}} dimensionless. We call η𝒴(t)\eta_{\mathcal{Y}}(t) the canonical auto-Igusa zeta function of 𝒴\mathcal{Y}.

Example 6.4.

Let 𝒴\mathcal{Y} be the trivial formal deformation of a morphism YSY\to S where S=Spec(A)S=\mbox{Spec}(A) with AA reduced with respect to the admissible system {Xi}iI\{X_{i}\}_{i\in I} with Xi=Spec(A[t]/ti+1)X_{i}=\mbox{Spec}(A[t]/t^{i+1}). Then,

(6.3) 𝒜i(𝒴)i(Y)×S𝔸Si\mathcal{A}_{i}(\mathcal{Y})\cong\mathcal{L}_{i}(Y)\times_{S}\mathbb{A}_{S}^{i}

where i(Y)\mathcal{L}_{i}(Y) denotes the classical jet space of YY over SS. Thus,

(6.4) ζ𝒴(t)=iI[i(Y)]𝕃diti\zeta_{\mathcal{Y}}(t)=\sum_{i\in I}[\mathcal{L}_{i}(Y)]\mathbb{L}^{-d\cdot i}t^{i}

where d=dimS(Y)d=\mbox{dim}_{S}(Y). This is the classical motivic Igusa zeta function of YY over SS. Moreover, in this case, the canonical auto-Igusa zeta function is also the classical motivic Igusa-zeta function–i.e., in this case, we have

(6.5) η𝒴(t)=ζ𝒴(t).\eta_{\mathcal{Y}}(t)=\zeta_{\mathcal{Y}}(t).
Example 6.5.

Let S=Spec(k)S=\mbox{Spec}(k) with kk an algebraically closed field of characteristic zero. Let 𝒴\mathcal{Y} be the trivial deformation of the point Spec(k)X0=S\mbox{Spec}(k)\to X_{0}=S, and let Xi=Spec(𝒪C,O/𝔪Oi+1)X_{i}=\mbox{Spec}(\mathcal{O}_{C,O}/\mathfrak{m}_{O}^{i+1}) where CC is the cuspidal cubic defined by y2=x3y^{2}=x^{3} over kk and OO is the origin given by the singular point (0,0)(0,0). Then,

(6.6) 𝒜i(𝒴)2(i3)(C)×S𝔸S7\mathcal{A}_{i}(\mathcal{Y})\cong\mathcal{L}_{2(i-3)}(C)\times_{S}\mathbb{A}_{S}^{7}

for all i>3i>3 (cf. Theorem 6.1 of [Sto17]). Moreover, in Section 7 of loc. cit., we use formula 6.6 to explicitly calculate the auto-Igusa zeta function to obtain

(6.7) ζ𝒴(t)=𝕃1+𝕃t+𝕃2t2+(𝕃7𝕃6)t3+𝕃7t4+𝕃7t7(1𝕃t3)(1t).\zeta_{\mathcal{Y}}(t)=\mathbb{L}^{-1}+\mathbb{L}t+\mathbb{L}^{2}t^{2}+\frac{(\mathbb{L}^{7}-\mathbb{L}^{6})t^{3}+\mathbb{L}^{7}t^{4}+\mathbb{L}^{7}t^{7}}{(1-\mathbb{L}t^{3})(1-t)}.

7. Motivic Rationality for Plane Curve Singularities.

We will need to refer to the following Set-up below.

Set-up 7.1.

Let S=Spec(k)S=\mbox{Spec}(k) where kk is an algebraically closed field of characteristic zero. Let CC be an algebraic curve on a smooth surface and let pp be a point on CC. We form the admissible injective system of fat points {Xi}i\{X_{i}\}_{i\in\mathbb{N}} over SS by defining XiX_{i} to be the formal neighborhood Spec(𝒪C,p/𝔪pi+1)\mbox{Spec}{(\mathcal{O}_{C,p}/\mathfrak{m}_{p}^{i+1})} of the point pp on CC. Let 𝒴=limiYi\mathcal{Y}=\varinjlim_{i\in\mathbb{N}}Y_{i} be any object of 𝙳𝚎𝚏𝒳\mathtt{Def}_{\mathcal{X}} such that φi:YiXi\varphi_{i}:Y_{i}\to X_{i} is smooth with YiY_{i} of pure dimension did_{i} for all ii\in\mathbb{N}.

The main rationality result the author has obtained is the following.

Theorem 7.2.

Let 𝒴\mathcal{Y} be a formal deformation as in Set-up 7.1. Then,

(7.1) ζ𝒴(t)=p(t)i=1n(1𝕃aitbi)1,\zeta_{\mathcal{Y}}(t)=p(t)\cdot\prod_{i=1}^{n}(1-\mathbb{L}^{a_{i}}t^{b_{i}})^{-1},

where p(t)𝒢k[t]p(t)\in\mathcal{G}_{k}[t] and where aia_{i}\in\mathbb{Z} and bib_{i}\in\mathbb{N}.

Proof.

By assumption, φi:YiXi\varphi_{i}:Y_{i}\to X_{i} is smooth, and, by shrinking YiY_{i} if necessary, φi\varphi_{i} factors through an étale morphism444The existence of an étale morphism ψ\psi follows from the proof of Theorem 8.3.3, page 182 of [Ols16] whose argument is taken from the proof Theorem 8.1 of [LMB00]. In fact, one may generalize by letting XiX_{i} can be merely a Deligne-Mumford Stack. ψ:Yi𝔸Xidi\psi:Y_{i}\to\mathbb{A}_{X_{i}}^{d_{i}}. Therefore,

(7.2) Hom¯S(Xi,Yi)Hom¯S(Xi,𝔸Xidi)×𝔸XidiYiEnd¯S(Xi)×S𝔸Sdi(rankS(Xi)1)×SYi\underline{\mbox{Hom}}_{S}(X_{i},Y_{i})\cong\underline{\mbox{Hom}}_{S}(X_{i},\mathbb{A}_{X_{i}}^{d_{i}})\times_{\mathbb{A}_{X_{i}}^{d_{i}}}Y_{i}\cong\underline{\mbox{End}}_{S}(X_{i})\times_{S}\mathbb{A}_{S}^{d_{i}(\mathrm{rank}_{S}(X_{i})-1)}\times_{S}Y_{i}

Since (Yi)red=(Y0)red(Y_{i})^{\mathrm{red}}=(Y_{0})^{\mathrm{red}}, we have

(7.3) [𝒜i(𝒴)]𝕃di(rankS(Xi)1)ei(rankS(Xi)1)=[Y0][End¯S(Xi)]𝕃ei(rankS(Xi)1)[\mathcal{A}_{i}(\mathcal{Y})]\mathbb{L}^{-d_{i}(\mathrm{rank}_{S}(X_{i})-1)-e_{i}(\mathrm{rank}_{S}(X_{i})-1)}=[Y_{0}][\underline{\mbox{End}}_{S}(X_{i})]\mathbb{L}^{-e_{i}(\mathrm{rank}_{S}(X_{i})-1)}

Multiplying tit^{i} on both sides and summing over ii in 𝒢k[[t]]\mathcal{G}_{k}[[t]], we have

(7.4) ζ𝒴(t)=[Y0]i=0[End¯S(Xi)]𝕃ei(rankS(Xi)1)ti.\zeta_{\mathcal{Y}}(t)=[Y_{0}]\sum_{i=0}^{\infty}[\underline{\mbox{End}}_{S}(X_{i})]\mathbb{L}^{-e_{i}(\mathrm{rank}_{S}(X_{i})-1)}t^{i}.

Note that then that ζ𝒴(t)=[Y0]ζC,p(t)\zeta_{\mathcal{Y}}(t)=[Y_{0}]\zeta_{C,p}(t), where ζC,p(t)\zeta_{C,p}(t) is the auto-Igusa zeta of the algebraic germ (C,p)(C,p) defined in [Sto18]. In loc. cit., we show that this series is of the form displayed in formula 7.1. ∎

Part of the argument in [Sto18] relies on rationality results of [DF98] and [DL99] and their later generalizations in [CL08] and [CL15]. In fact, the first part of this argument holds whenever the morphisms φi:YiXi\varphi_{i}:Y_{i}\to X_{i} are smooth with {Xi}iI\{X_{i}\}_{i\in I} any admissible injective system of fat points over any algebraic space SS – i.e., regardless of whether or not XiX_{i} is subscheme of a plane curve CC, formula 7.4 holds provided φi:YiXi\varphi_{i}:Y_{i}\to X_{i} is smooth for all iIi\in I. However, as of yet, rationality is only proven for the case when 𝒴\mathcal{Y} is as in Set-up 7.1. Thus, there are several directions for future results:

Question 7.3.

Let 𝒴\mathcal{Y} be a formal deformation such that φi:YiXi\varphi_{i}:Y_{i}\to X_{i} are smooth for all iIi\in I with {Xi}iI\{X_{i}\}_{i\in I} any admissible system of fat points over an algebraic space SS. Under what conditions on the admissible system {Xi}iI\{X_{i}\}_{i\in I} and on SS will ζ𝒴(t)\zeta_{\mathcal{Y}}(t) be rational – i.e., when will ζ𝒴(t)\zeta_{\mathcal{Y}}(t) have a similar form as in formula 7.1?

Question 7.4.

Assuming that the appropriate conditions are placed on {Xi}iI\{X_{i}\}_{i\in I} and SS through investigations of the type discussed in Question 7.3, which formal deformations 𝒴\mathcal{Y} over {Xi}iI\{X_{i}\}_{i\in I} will have a zeta function ζ𝒴(t)\zeta_{\mathcal{Y}}(t) of the form given in formula 7.1?

Question 7.5.

Note that formula 6.5 gives a relationship between η𝒴(t)\eta_{\mathcal{Y}}(t) and ζ𝒴(t)\zeta_{\mathcal{Y}}(t) in the most basic case (i.e., they are both equal to the classical motivic Igusa-zeta function). In general, what is the relationship between η𝒴(t)\eta_{\mathcal{Y}}(t) and ζ𝒴(t)\zeta_{\mathcal{Y}}(t)? In particular, is there a way to transfer results vis-a-vis Question 7.3 and Question 7.4 to corresponding statments about η𝒴(t)\eta_{\mathcal{Y}}(t)? In general, if ζ𝒴(t)\zeta_{\mathcal{Y}}(t) is rational, must η𝒴(t)\eta_{\mathcal{Y}}(t) also be rational?

Question 7.6.

In the case of Set-up 7.1, there are cases where explicit computations of the auto Igusa-zeta function are possible. Are similar computations possible for the canonical Igusa-zeta function η𝒴(t)\eta_{\mathcal{Y}}(t)?

References

  • [Art69] Michael Artin, Algebraization of formal moduli: I, Global Analysis (Papers in Honor of K. Kodaira), Univ. Tokyo Press, Tokyo, 1969, pp. 21–71.
  • [CL08] R. Cluckers and F. Loeser, Constructible motivic functions and motivic integration, Invent. Math. 173 (2008), no. 1, 23–121.
  • [CL15] by same author, Motivic integration in all residue field characteristics for Henselian discretely valued fields of characteristic zero, J. Reine Angew. Math. 701 (2015), 1–31.
  • [DF98] J. Denef and Loeser F., Motivic Igusa Zeta Functions, J. Algebraic Geometry 7 (1998), no. 3, 505–537.
  • [DL99] J. Denef and F. Loeser, Germs of arcs of singular varieties and motivic integration, Inven. Math. 135 (1999), 201–232.
  • [Eke09] Torsten Ekedahl, The grothendieck group of algebraic stacks, arXiv:0903.3143v2, 2009.
  • [GD67] Alexander Grothendieck and Jean Dieudonné, Éléments de géométrie algébrique IV, Publications Mathématiques, vol. 20, 24, 28, 32, Institute des Hautes Études Scientifiques., 1964-1967.
  • [Gro62] Alexander Grothendieck, Fondements de la géométrie algébrique, Secrétariat mathématique, 1962.
  • [LMB00] G. Laumon and L. Moret-Bailly, Champs algébriques, Ergebnisse der Mathematik, 2000.
  • [Lur09] Jacob Lurie, Higher topos theory (am-170), Princeton University Press, 2009.
  • [Ols06] Martin C. Olsson, Hom¯\underline{Hom}-stacks and restriction of scalars, Duke Math. J. 134 (2006), no. 1, 139–164.
  • [Ols16] by same author, Algebraic spaces and stacks, American Mathematical Society, 2016.
  • [Scha] H. Schoutens, Schemic Grothendieck Rings I, websupport1.citytech.cuny.edu/faculty/hschoutens/PDF/SchemicGrothendieckRingPartII.pdf, accessed on 4/01/2017.
  • [Schb] by same author, Schemic Grothendieck Rings II, websupport1.citytech.cuny.edu/faculty/hschoutens/PDF/SchemicGrothendieckRingPartII.pdf, accessed on 4/01/2017.
  • [Sta17] The Stacks Project Authors, Stacks Project, http://stacks.math.columbia.edu, 2017.
  • [Sto17] Andrew R. Stout, On the auto Igusa-zeta function of an algebraic curve, Journal of Symbolic Computation 79 (2017), no. Part 1, 156 – 185, SI: MEGA 2015.
  • [Sto18] Andrew R. Stout, The auto Igusa-zeta function of a plane curve singularity is rational, Proc. Amer. Math. Soc. In Press. (2018).