Hostname: page-component-7bb8b95d7b-nptnm Total loading time: 0 Render date: 2024-09-20T06:56:45.519Z Has data issue: false hasContentIssue false

$\mu ^*$-ZARISKI PAIRS OF SURFACE SINGULARITIES

Published online by Cambridge University Press:  05 December 2023

CHRISTOPHE EYRAL*
Affiliation:
Institute of Mathematics Polish Academy of Sciences ul. Śniadeckich 8 00-656 Warsaw Poland
MUTSUO OKA
Affiliation:
Professor Emeritus of Tokyo Institute of Technology 3-19-8 Nakaochiai Shinjuku-ku Tokyo 161-0032 Japan okamutsuo@gmail.com
Rights & Permissions [Opens in a new window]

Abstract

Let $f_0$ and $f_1$ be two homogeneous polynomials of degree d in three complex variables $z_1,z_2,z_3$. We show that the Lê–Yomdin surface singularities defined by $g_0:=f_0+z_i^{d+m}$ and $g_1:=f_1+z_i^{d+m}$ have the same abstract topology, the same monodromy zeta-function, the same $\mu ^*$-invariant, but lie in distinct path-connected components of the $\mu ^*$-constant stratum if their projective tangent cones (defined by $f_0$ and $f_1$, respectively) make a Zariski pair of curves in $\mathbb {P}^2$, the singularities of which are Newton non-degenerate. In this case, we say that $V(g_0):=g_0^{-1}(0)$ and $V(g_1):=g_1^{-1}(0)$ make a $\mu ^*$-Zariski pair of surface singularities. Being such a pair is a necessary condition for the germs $V(g_0)$ and $V(g_1)$ to have distinct embedded topologies.

Type
Article
Copyright
© The Author(s), 2023. Published by Cambridge University Press on behalf of Foundation Nagoya Mathematical Journal

1 Introduction and statement of the result

Let $g_0$ and $g_1$ be two polynomials in three complex variables $z_1,z_2,z_3$ . We assume that they vanish at the origin $\mathbf {0}\in \mathbb {C}^3$ and that the corresponding germs of surfaces, $V(g_0):=g_0^{-1}(0)$ and $V(g_1):=g_1^{-1}(0)$ , have an isolated singularity at $\mathbf {0}$ . It is well known that if $V(g_0)$ and $V(g_1)$ have the same embedded topology (i.e., if the pairs $(\mathbb {C}^3,V(g_0))$ and $(\mathbb {C}^3,V(g_1))$ are homeomorphic in a neighborhood of the origin, or equivalently, by [Reference Saeki28], if the pairs $(\mathbb {S}_\varepsilon ^5,K_{g_0})$ and $(\mathbb {S}_\varepsilon ^5,K_{g_1})$ are diffeomorphic for any $\varepsilon $ small enough), then they have the same Milnor number (see [Reference Lê18], [Reference Milnor23], [Reference Teissier33]). Here, $K_{g_l}$ denotes the link of $g_l$ ( $l\in \{0,1\}$ ), that is, $K_{g_l}:=\mathbb {S}^5_\varepsilon \cap V(g_l)$ for $\varepsilon $ small enough, where $\mathbb {S}_\varepsilon ^5$ is the sphere with radius $\varepsilon $ centered at $\mathbf {0}\in \mathbb {C}^3$ . (Note that the diffeomorphism type of the embedded link $(\mathbb {S}_\varepsilon ^5,K_{g_l})$ is independent of $\varepsilon $ , provided that $\varepsilon $ is small enough.) On the other hand, it is quite possible for two isolated surface singularities $V(g_0)$ and $V(g_1)$ to have the same Milnor number and non-diffeomorphic embedded links. In [Reference Artal-Bartolo3], [Reference Artal-Bartolo4], using Luengo’s theory of superisolated singularities [Reference Luengo20], Artal-Bartolo even showed that the embedded topology of the link of a superisolated surface singularity is not determined by the topology of the abstract link and the characteristic polynomial of the monodromy. However, in practice, given $g_0$ and $g_1$ with the same characteristic polynomial (or equivalently, the same monodromy zeta-function), the same abstract topology, and even with the same Teissier $\mu ^*$ -invariant, it is extremely difficult to determine whether $(\mathbb {S}_\varepsilon ^5,K_{g_0})$ and $(\mathbb {S}_\varepsilon ^5,K_{g_1})$ are diffeomorphic or not. The goal of this paper is to investigate a special class of Lê–Yomdin surface singularities which are “likely to systematically produce” pairs of germs sharing all these invariants but having non-diffeomorphic embedded links. Such pairs are called $\mu ^*$ -Zariski pairs of surface singularities and are defined as follows.

Consider a classical Zariski pair of (reduced) projective curves $C_0=\{f_0=0\}$ and $C_1=\{f_1=0\}$ of degree d in the complex projective plane $\mathbb {P}^2$ , that is, there are regular neighborhoods $N_0$ and $N_1$ of $C_0$ and $C_1$ , respectively, such that $(N_0,C_0)$ and $(N_1,C_1)$ are homeomorphic, while $(\mathbb {P}^2,C_0)$ and $(\mathbb {P}^2,C_1)$ are not. The first example of such a pair was found by Zariski [Reference Zariski36] in the early 1930s, and their systematic study was initiated by Artal-Bartolo [Reference Artal-Bartolo5] in the mid-1990s (for a detailed survey on this topic, see [Reference Artal-Bartolo, Cogolludo and Tokunaga6], [Reference Oka25]). By a linear change of the coordinates $z_1,z_2,z_3$ , we may assume that the singularities of the curves $C_0$ and $C_1$ are not located on the coordinate lines $z_i=0$ ( $1\leq i\leq 3$ ) and that their defining polynomials $f_0$ and $f_1$ are convenientFootnote 1 and Newton non-degenerate on any face $\Delta $ of their (common) Newton diagram if $\Delta $ is not top-dimensional. The fact that the singularities of the curves do not sit on the coordinate lines implies that for any integers $m\geq 1$ and $1\leq i\leq 3$ , the polynomials

$$ \begin{align*} g_0:=f_0+z_i^{d+m} \quad\mbox{and}\quad g_1:=f_1+z_i^{d+m} \end{align*} $$

define an isolated surface singularity at $\mathbf {0}$ (see [Reference Luengo and Melle21, Th. 2]). Such singularities are called m-Lê–Yomdin singularities and were first investigated by Yomdin and Lê in [Reference Lê19], [Reference Iomdin13], respectively. The monodromy zeta-function (or the characteristic polynomial) of such a singularity was computed by Siersma [Reference Siersma29], [Reference Siersma30], Stevens [Reference Stevens31], and Gusein-Zade, Luengo, and Melle-Hernández [Reference Gusein-Zade, Luengo and Melle-Hernández11] (see also [Reference Oka and Papadopoulos26]). (The Milnor number was already known from [Reference Luengo and Melle21].) In [Reference Artal-Bartolo, Cogolludo-Agustín and Martín-Morales7], Artal-Bartolo, Cogolludo-Agustín, and Martín-Morales gave a characterization for the abstract link of a Lê–Yomdin singularity to be a rational homology sphere.

In the special case where $m=1$ , a $1$ -Lê–Yomdin singularity is called a superisolated singularity. Superisolated singularities were introduced by Luengo [Reference Luengo20] to answer important questions and conjectures. For example, in [Reference Luengo20], Luengo gave examples of superisolated surface singularities for which the $\mu $ -constant stratum in the miniversal deformation is not smooth.

Now, let us make precise the notion of Zariski pair of surface singularities. Let $g_0=f_0+z_i^{d+m}$ and $g_1=f_1+z_i^{d+m}$ be two Lê–Yomdin surface singularities obtained from a Zariski pair of curves $f_0$ and $f_1$ as above.

  • We say that $(V(g_0),V(g_1))$ is a weak $\zeta $ -Zariski pair of surface singularities if $g_0$ and $g_1$ have the same monodromy zeta-function (in particular, the same Milnor number).

  • A weak $\zeta $ -Zariski pair for which the germs $V(g_0)$ and $V(g_1)$ (or equivalently, the links $K_{g_0}$ and $K_{g_1}$ ) have the same abstract topology is called a $\zeta $ -Zariski pair (without the adjective “weak”).

  • A (weak) $\zeta $ -Zariski pair is said to be a (weak) $\mu ^*$ -Zariski pair if $g_0$ and $g_1$ have the same $\mu ^*$ -invariant while belonging to distinct path-connected components of the $\mu ^*$ -constant stratum.

  • A (weak) $\mu ^*$ -Zariski pair is called a (weak) $\mu $ -Zariski pair if furthermore $g_0$ and $g_1$ lie in different path-connected components of the $\mu $ -constant stratum.

  • Finally, a (weak) $\zeta $ -Zariski pair is called a (weak) Zariski pair if the germs $V(g_0)$ and $V(g_1)$ (or equivalently, $K_{g_0}$ and $K_{g_1}$ ) have distinct embedded topologies.

Note that a (weak) Zariski pair of surface singularities $V(g_0)$ and $V(g_1)$ sharing the same $\mu ^*$ -invariant is always a (weak) $\mu $ -Zariski pair, and hence a (weak) $\mu ^*$ -Zariski pair. That is, being a (weak) $\mu ^*$ -Zariski pair is a necessary condition for being a (weak) Zariski pair. Indeed, by [Reference Eyral and Oka10, Th. 5.3], if $g_0$ and $g_1$ lie in the same path-connected component of the $\mu ^*$ -constant stratum, then they can always be joined by a piecewise complex-analytic path (defined in the relevant natural way), and by a well-known theorem of Teissier [Reference Teissier32, théorème 3.9], this in turn implies that the diffeomorphism type of the pairs $(\mathbb {S}_\varepsilon ^5,K_{g_0})$ and $(\mathbb {S}_\varepsilon ^5,K_{g_1})$ is identical.

In [Reference Luengo20], Luengo proved that for superisolated singularities (i.e., for $m=1$ ), the abstract links $K_{g_0}$ and $K_{g_1}$ are homeomorphic. The second-named author showed a similar property for $m\geq 1$ if the singularities of the corresponding curves $C_0$ and $C_1$ are Newton non-degenerate (see [Reference Oka27, Th. 24 and Rem. 25]). In [Reference Artal-Bartolo3, théorème 4.4] and [Reference Artal-Bartolo4, théorème 1.6, §1.7, and corollaire 5.6.6], Artal-Bartolo proved that if $m=1$ , then $V(g_0)$ and $V(g_1)$ also share the same characteristic polynomial of the monodromy, and if furthermore the Alexander polynomials of the curves $C_0$ and $C_1$ do not coincide, then $V(g_0)$ and $V(g_1)$ do not have the same embedded topology. In particular, combined with Luengo’s result, this shows that, in this latter case, $(V(g_0),V(g_1))$ is a Zariski pair of surface singularities.

In this paper, we prove the following theorem.

Theorem 1.1. If the singularities of the curves $C_0$ and $C_1$ are Newton non-degenerate in some suitable local coordinates,Footnote 2 then the pair made up of the m-Lê–Yomdin singularities $V(g_0)$ and $V(g_1)$ is a $\mu ^*$ -Zariski pair of surface singularities.

Again, we emphasize that being a $\mu ^*$ -Zariski pair is a necessary condition for being a Zariski pair of surface singularities. We also highlight that in the above theorem, the Alexander polynomials of the curves $C_0$ and $C_1$ may coincide.

We expect that with the assumption of the theorem, $(V(g_0),V(g_1))$ is a $\mu $ -Zariski pair, and in fact, a Zariski pair of surface singularities. As mentioned above, in the special case of superisolated singularities (i.e., $m=1$ ), and provided that the curves have distinct Alexander polynomials (but not necessarily Newton non-degenerate singularities), this is already proved by combining Artal-Bartolo’s [Reference Artal-Bartolo3], [Reference Artal-Bartolo4] and Luengo’s [Reference Luengo20] results.

2 Proof of Theorem 1.1

First, we show that $(V(g_0),V(g_1))$ is a $\zeta $ -Zariski pair of surface singularities, and then we prove that it is in fact a $\mu ^*$ -Zariski pair. To simplify, we assume that $i=1$ , that is, $g_l=f_l+z_1^{d+m}$ ( $l\in \{0,1\}$ ).

To compute the monodromy zeta-function $\zeta _{g_l,\mathbf {0}}(t)$ of $g_l$ , we use the classical formula of Siersma (see [Reference Siersma29, Main theorem, p. 183] and [Reference Siersma30, Th. 3.4 and Rem. 3.6]), Stevens (see [Reference Stevens31, p. 140]), and Gusein-Zade, Luengo, and Melle-Hernández (see [Reference Gusein-Zade, Luengo and Melle-Hernández11, p. 250]) (see also [Reference Oka and Papadopoulos26, Lem. 3.2 and Th. 3.7]). More precisely, the ordinary point blowing up at $\mathbf {0}\in \mathbb {C}^3$ , denoted by $\pi \colon X\to \mathbb {C}^3$ , being a biholomorphism over $\mathbb {C}^3\setminus V(g_l)$ , the tubular Milnor fibration of $g_l$ at $\mathbf {0}$ can be lifted to X, so that the pullback $\pi ^* g_l\equiv g_l\circ \pi $ is a locally trivial fibration which is isomorphic to it. Let $U_1:=\mathbb {P}^2\setminus \{z_1=0\}$ be the standard affine chart of $\mathbb {P}^2$ with coordinates $(z_2/z_1,z_3/z_1)$ . In the corresponding chart $X\cap (\mathbb {C}^3\times U_1)$ of X, with coordinates $\mathbf {y}\equiv (y_1,y_2,y_3):=(z_1,z_2/z_1,z_3/z_1)$ , the pullback $\pi ^*g_l$ is written as

$$ \begin{align*} \pi^*g_l (\mathbf{y}) = y_1^d(f_l(1,y_2,y_3)+y_1^m). \end{align*} $$

The first factor, $y_1^d$ , corresponds to the exceptional divisor $E\simeq \mathbb {P}^2$ , while the second one represents the strict transform $\tilde V(g_l)$ of $V(g_l)$ . Outside of the exceptional divisor, $\tilde V(g_l)$ has no singularities. On the exceptional divisor, it has a finite number of isolated singularities, which are given by the singular points $\mathbf {p}\in \Sigma (C_l)$ of the reduced curve $C_l$ . Then the formula for the zeta-function mentioned above is written as

(2.1) $$ \begin{align} \zeta_{g_l,\mathbf{0}}(t)=\zeta_d(t)\times (1-t^{d})^{\mu^{\scriptscriptstyle{\text{tot}}}(C_l)} \times \prod_{\substack{\mathbf{p}\in\Sigma(C_l)}} \zeta_{\pi^*g_l,\mathbf{p}}(t), \end{align} $$

where $\zeta _d(t)$ is the zeta-function of a Newton non-degenerate homogeneous polynomial of degree d (i.e., $\zeta _d(t)=(1-t^{d})^{-d^2+3d-3}$ ), $\Sigma (C_l)$ is the set of singular points of $C_l$ , and $\mu ^{\scriptscriptstyle{\text{tot}}}(C_l)$ is the total Milnor number of $C_l$ (i.e., the sum of the local Milnor numbers at the singular points of $C_l$ ).

By our assumption, there exist local coordinates $\mathbf {x}=(x_1,x_2,x_3)$ and $\mathbf {u}=(u_1,u_2,u_3)$ near $\mathbf {p}_0\in \Sigma (C_0)$ and $\mathbf {p}_1\in \Sigma (C_1)$ , respectively, where $x_1=u_1=y_1$ and $(x_2,x_3)$ and $(u_2,u_3)$ are analytic coordinate changes of $(y_2,y_3)$ ,Footnote 3 such that

$$ \begin{align*} \pi^* g_0 (\mathbf{x}) = x_1^d(h_0(x_2,x_3)+x_1^m) \quad\mbox{and}\quad \pi^* g_1 (\mathbf{u}) = u_1^d(h_1(u_2,u_3)+u_1^m), \end{align*} $$

where $h_0$ and $h_1$ are Newton non-degenerate. Moreover, if the singularities $(C_1,\mathbf {p}_1)$ and $(C_0,\mathbf {p}_0)$ are topologically equivalent, then we may assume that the Newton diagrams, $\Gamma (h_0)$ and $\Gamma (h_1)$ , of $h_0$ and $h_1$ coincide. It follows that $\pi ^* g_0$ and $\pi ^* g_1$ are Newton non-degenerate with the same Newton diagram, and hence, by Varchenko’s formula (see [Reference Varchenko34, Th. (4.1)]), we have

$$ \begin{align*} \zeta_{\pi^*\!g_0,\mathbf{p}_0}(t)=\zeta_{\pi^*\!g_1,\mathbf{p}_1}(t). \end{align*} $$

Since $(C_0,C_1)$ is a Zariski pair of projective curves, the total Milnor numbers $\mu ^{\scriptscriptstyle{\text{tot}}}(C_0)$ and $\mu ^{\scriptscriptstyle{\text{tot}}}(C_1)$ coincide, and the equality $\zeta _{g_0,\mathbf {0}}(t)=\zeta _{g_1,\mathbf {0}}(t)$ follows immediately from (2.1).

To conclude that $(V(g_0),V(g_1))$ is a $\zeta $ -Zariski pair, it remains to observe that the links $K_{g_0}$ and $K_{g_1}$ have the same abstract topology; this is proved in [Reference Oka27, Th. 24 and Rem. 25].

Now, let us show that $(V(g_0),V(g_1))$ is a $\mu ^*$ -Zariski pair of surface singularities. For that, we must first show that $g_0$ and $g_1$ have the same $\mu ^*$ -invariant at $\mathbf {0}$ . We recall that the $\mu ^*$ -invariant of $g_l$ at $\mathbf {0}$ , introduced by Teissier in [Reference Teissier32], is the triple

$$ \begin{align*} \mu^*_{\mathbf{0}}(g_l):=(\mu_{\mathbf{0}}(g_l),\mu_{\mathbf{0}}({g_l}\vert_{H}),\mbox{mult}_{\mathbf{0}}(g_l)-1), \end{align*} $$

where $\mu _{\mathbf {0}}(g_l)$ is the Milnor number of $g_l$ at $\mathbf {0}$ , $\mu _{\mathbf {0}}({g_l}\vert _{H})$ is the Milnor number at $\mathbf {0}$ of the restriction of $g_l$ to a generic plane H of $\mathbb {C}^3$ through the origin (this number is usually denoted by $\mu _{\mathbf {0}}^{(2)}({g_l})$ ), and $\mbox {mult}_{\mathbf {0}}(g_l)$ is the multiplicity of $g_l$ at $\mathbf {0}$ .

By [Reference Luengo and Melle21, Th. 2], for any $l\in \{0,1\}$ , the Milnor number $\mu _{\mathbf {0}}(g_l)$ is given by

$$ \begin{align*} \mu_{\mathbf{0}}(g_l)=(d-1)^3+m \mu^{\scriptscriptstyle{\text{tot}}}, \end{align*} $$

where $\mu ^{\scriptscriptstyle{\text{tot}}}$ is the (common) total Milnor number of $C_0$ and $C_1$ .

For a generic plane H of $\mathbb {C}^3$ through the origin, the restriction $f_l\vert _{H}$ is a homogeneous polynomial of degree d with an isolated singularity at $\mathbf {0}$ , so that its Milnor number at $\mathbf {0}$ is $\mu _{\mathbf {0}}(f_l\vert _{H})=(d-1)^2$ . Since $f_l\vert _{H}$ is Newton non-degenerate and the term $z_1^{d+m}$ is above the Newton diagram $\Gamma (g_l\vert _{H})=\Gamma (f_l\vert _{H})$ , the restriction $g_l\vert _{H}$ is Newton non-degenerate too. Thus, its Milnor number at $\mathbf {0}$ is determined by $\Gamma (g_l\vert _{H})$ , and hence we have

$$ \begin{align*} \mu_{\mathbf{0}}^{(2)}({g_l}):=\mu_{\mathbf{0}}(g_l\vert_{H})=\mu_{\mathbf{0}}(f_l\vert_{H})=(d-1)^2. \end{align*} $$

Lastly, since the multiplicities of $g_0$ and $g_1$ at $\mathbf {0}$ are equal to d, it follows that $g_0$ and $g_1$ have the same $\mu ^*$ -invariant at $\mathbf {0}$ , namely, for any $l\in \{0,1\}$ , we have

$$ \begin{align*} \mu_{\mathbf{0}}^*(g_l)=((d-1)^3+m \mu^{\scriptscriptstyle{\text{tot}}},(d-1)^2,d-1). \end{align*} $$

Finally, and this is the heart of the proof, we must now show that $g_0$ and $g_1$ lie in different path-connected components of the $\mu ^*$ -constant stratum. To this end, we argue by contradiction. Suppose that $g_0$ and $g_1$ belong to the same component. Then, by [Reference Eyral and Oka10, Th. 5.3], there exists a $\mu ^*$ -constant piecewise complex-analytic family $\{g_s\}_{0\leq s\leq 1}$ connecting $g_0$ and $g_1$ . In particular, the multiplicity $\mbox {mult}_{\mathbf {0}}(g_s)$ of $g_s$ at $\mathbf {0}$ is independent of $s\in [0,1]$ , and the initial polynomial $\mbox {in}(g_s)$ of $g_s$ (i.e., the sum of the monomials of $g_s$ of lowest degree) has degree d.

Lemma 2.1. For each $s\in [0,1]$ , the homogeneous polynomial ${\mathrm{in}}(g_s)$ is reduced, so that the projective curve $C_{s}\subseteq \mathbb {P}^2$ defined by ${\mathrm{in}}(g_s)$ has only isolated singularities.

Proof. We argue by contradiction. Suppose there exists $s_0\in [0,1]$ such that $\mbox {in}(g_{s_0})$ is not reduced (i.e., $C_{s_0}$ has non-isolated singularities). Then, for a generic linear plane H of $\mathbb {C}^3$ , there are coordinates $(x,y)$ for H and linear forms $\ell _1(x,y),\ldots ,\ell _q(x,y)$ such that

$$ \begin{align*} \mbox{in}(g_{s_0})\vert_H(x,y)=\ell_1(x,y)^{p_1}\cdots\ell_q(x,y)^{p_q} \end{align*} $$

with $p_1\geq \cdots \geq p_q$ and $p_1\geq 2$ . By a linear change of coordinates, we may assume that $\ell _1(x,y)\equiv x$ , so that

$$ \begin{align*} \mbox{in}(g_{s_0})\vert_H(x,y)=x^{p_1}h(x,y), \end{align*} $$

where h is a homogeneous polynomial of degree $d-p_1$ (in particular, $\mbox {in}(g_{s_0})\vert _H$ is not convenient with respect to the coordinates $(x,y)$ ). By adding monomials of the form $x^{\alpha }$ and $y^{\beta }$ for $\alpha ,\, \beta $ large enough, we may also assume that $g_{s_0}\vert _H$ is convenient. Now, since the integral point $(1,d-1)$ is not on the Newton diagram $\Gamma (\mbox {in}(g_{s_0})\vert _H)$ of $\mbox {in}(g_{s_0})\vert _H$ with respect to the coordinates $(x,y)$ , it followsFootnote 4 that

$$ \begin{align*} \nu(\Gamma_{\!-}(g_{s_0}\vert_H))>\nu(\Gamma_{\!-}(g_{0}\vert_H)) \end{align*} $$

(see Figure 1, where $\Gamma _{\!+}(\mbox {in}(g_{s_0})\vert _H)$ is the Newton polyhedron of $\mbox {in}(g_{s_0})\vert _H$ in the coordinates $(x,y)$ ). Here, $\nu (\cdot )$ denotes the Newton number (see [Reference Kouchnirenko14] for the definition) and $\Gamma _{\!-}(g_{s_0}\vert _H)$ stands for the cone over $\Gamma (g_{s_0}\vert _H)$ with the origin as vertex. (Again, $\Gamma (g_{s_0}\vert _H)$ denotes the Newton diagram of $g_{s_0}\vert _H$ with respect to the coordinates $(x,y)$ .) The polyhedron $\Gamma _{\!-}(g_{0}\vert _H)$ is defined similarly. Since

$$ \begin{align*} \mu_{\mathbf{0}}(g_{s_0}\vert_H)\geq \nu(\Gamma_{\!-}(g_{s_0}\vert_H)) \end{align*} $$

(see [Reference Kouchnirenko14, théorème 1.10]), altogether we have

$$ \begin{align*} \mu^{(2)}_{\mathbf{0}}(g_{s_0})=\mu_{\mathbf{0}}(g_{s_0}\vert_H)\geq \nu(\Gamma_{\!-}(g_{s_0}\vert_H))>\nu(\Gamma_{\!-}(g_{0}\vert_H))=(d-1)^2=\mu^{(2)}_{\mathbf{0}}(g_{0}), \end{align*} $$

which is a contradiction to the $\mu ^*$ -constancy.

Lemma 2.2. The zeta-function $\zeta _{g_s,\mathbf {0}}(t)$ is independent of $s\in [0,1]$ .

Figure 1 Newton diagrams.

Proof. It is well known that in a $\mu ^*$ -constant piecewise complex-analytic family $\{g_s\}$ , the diffeomorphism type of the embedded link $(\mathbb {S}_\varepsilon ^5,K_{g_s})$ is independent of s (see [Reference Teissier32, théorème 3.9 and remarque 3.12]). Alternatively, we may use [Reference Oka27, Lem. 12], which asserts that in a $\mu $ -constant (a fortiori in a $\mu ^*$ -constant) piecewise complex-analytic family $\{g_s\}$ , the zeta-function $\zeta _{g_s,\mathbf {0}}(t)$ is independent of s.

Now, by the A’Campo formula (see [Reference A’Campo1, théorème 3]), we know that the zeta-function $\zeta _{g_s,\mathbf {0}}(t)$ is uniquely written as

(2.2) $$ \begin{align} \zeta_{g_s,\mathbf{0}}(t)=\prod_{i=1}^{\ell} (1-t^{d_i})^{\nu_i}, \end{align} $$

where $d_1,\ldots ,d_{\ell }$ are mutually disjoint and $\nu _1,\ldots ,\nu _{\ell }$ are nonzero integers. The smallest integer $d_{i_0}$ among $d_1,\ldots ,d_\ell $ is called the zeta-multiplicity of $g_s$ and is denoted by $m_\zeta (g_s)$ . We define the zeta-multiplicity factor of $\zeta _{g_s,\mathbf {0}}(t)$ as the factor $(1-t^{d_{i_0}})^{\nu _{i_0}}$ of (2.2) corresponding to the zeta-multiplicity $d_{i_0}\equiv m_\zeta (g_s)$ . Note that, by Lemma 2.2, the zeta-multiplicity of $g_s$ and the zeta-multiplicity factor of $\zeta _{g_s,\mathbf {0}}(t)$ are independent of s. Moreover, by [Reference Oka27, Prop. 11], we know that $m_\zeta (g_s)\geq \mbox {mult}_{\mathbf {0}}(g_s)=d$ , and the formula (2.1) shows that for $s=0$ we have $m_{\zeta }(g_0)\leq d$ . So, altogether, $m_{\zeta }(g_s)=d$ for any $s\in [0,1]$ .

Lemma 2.3. For any $s\in [0,1]$ , the zeta-multiplicity factor of $\zeta _{g_s,\mathbf {0}}(t)$ is given by

$$ \begin{align*} (1-t^d)^{-d^2+3d-3+\mu^{\scriptscriptstyle{\mathrm{tot}}}(C_{s})}, \end{align*} $$

and since the latter is independent of s, so is the total Milnor number $\mu ^{\scriptscriptstyle{\mathrm{tot}}}(C_{s})$ .

Proof. Here, to compute $\zeta _{g_s,\mathbf {0}}(t)$ , we apply a method developed by the second-named author in [Reference Oka24]. This method, inspired by an approach of Clemens [Reference Clemens8], was used in [Reference Oka24, Chap. I, Proof of Th. 5.2] to generalize the classical zeta-function formula of A’Campo [Reference A’Campo1]. Roughly, the idea is to decompose the lifted Milnor fibration $\pi ^*g_s$ (which is isomorphic to the original Milnor fibration of $g_s$ at $\mathbf {0}$ ) into its restrictions along “controlled” tubular neighborhoods of the strata in a canonical regular stratification of $\pi ^{-1}(V(g_s))$ . Then, by the multiplicativeness of the zeta-function, it suffices to compute the zeta-functions of the induced restricted fibrations. More precisely, let $\mathbf {p}_1,\ldots ,\mathbf {p}_{k_0}$ be the points of the singular set $\Sigma (C_s)$ of $C_s$ , and for each $\mathbf {p}_k$ , let $B_\varepsilon (\mathbf {p}_k)$ be a small ball centered at $\mathbf {p}_k$ . Put

$$ \begin{align*} B:=\bigcup_{k=1}^{k_0}B_\varepsilon(\mathbf{p}_k), \end{align*} $$

and consider tubular neighborhoods $N(C_s)$ and $N(E)$ of $C_s\setminus B$ and $E\setminus (N(C_s)\cup B)$ , respectively. As in [Reference Oka24, Chap. I, p. 56], we assume that the triple

(2.3) $$ \begin{align} \{B,N(C_s),N(E)\}, \end{align} $$

together with its natural associated projections and distance functions, makes a family of “control data” in the sense of Mather [Reference Mather22, §7]. Consider the restrictions of $\hat g_s:=\pi ^*g_s$ to $N(E)$ , $N(C_s)$ and the balls $B_\varepsilon (\mathbf {p}_k)$ , respectively. The relations (5.2.4) and (5.2.5), together with Lemmas (5.3) and (5.4), of [Reference Oka24, Chap. I] say that

(2.4) $$ \begin{align} \zeta_{g_s,\mathbf{0}}(t)\equiv \zeta_{\hat g_s}(t) = \zeta_{\hat g_s\vert_{N(E)}}(t)\cdot \zeta_{\hat g_s\vert_{N(C_s)}}(t)\cdot \prod_{k=1}^{k_0} \zeta_{\hat g_s\vert_{B_\varepsilon(\mathbf{p}_k)}}(t). \end{align} $$

Thus, it suffices to compute each piece $\zeta _{\hat g_s\vert _{N(E)}}(t)$ , $\zeta _{\hat g_s\vert _{N(C_s)}}(t)$ , and $\zeta _{\hat g_s\vert _{B_\varepsilon (\mathbf {p}_k)}}(t)$ separately.

We start with the calculation of the zeta-function $\zeta _{\hat g_s\vert _{N(E)}}(t)$ of the fibration $\hat g_s\vert _{N(E)}$ . For admissible coordinates $\mathbf {x}=(x_1,x_2,x_3)$ in a neighborhood $U_{\mathbf {p}}$ of a point $\mathbf {p}\in E':=E\setminus (N(C_s)\cup B)$ , we may assume that the projection

$$ \begin{align*} p\colon U_{\mathbf{p}}\cap N(E)\to E' \end{align*} $$

associated with the family of control data (2.3) is given by $\mathbf {x}\mapsto (0,x_2,x_3)$ , so that $E'$ is defined by $x_1=0$ and the restriction of $\hat g_s$ to $p^{-1}(\mathbf {p})$ is given by $x_1^d$ . Then, by the relation (5.2.5) of [Reference Oka24, Chap. I], the normal zeta-function $\zeta _{E'}^\bot (t)$ of $\hat g_s$ along $E'$ (see [Reference Oka24, Chap. I, p. 59] for the definition) is given by

$$ \begin{align*} \zeta_{E'}^\bot(t)=(1-t^d)^{-1}. \end{align*} $$

Thus, by [Reference Oka24, Chap. I, Lems. (5.3) and (5.4)], we get

$$ \begin{align*} \zeta_{\hat g_s\vert_{N(E)}}(t) & =(\zeta_{E'}^\bot(t))^{\chi(E\setminus \tilde V(g_s))}=(\zeta_{E'}^\bot(t))^{\chi(\mathbb{P}^2\setminus C_s)}=(\zeta_{E'}^\bot(t))^{\chi(\mathbb{P}^2)-\chi(C_s)}\\ &=(1-t^d)^{-\chi(\mathbb{P}^2)+\chi(C_s)}=(1-t^d)^{-3+\chi(C_s)} =(1-t^d)^{-3+3d-d^2+\mu^{\scriptscriptstyle{\text{tot}}}(C_s)}. \end{align*} $$

Here, $\chi (\cdot )$ denotes the Euler–Poincaré characteristic, and we recall that for a reduced curve $C_s$ of degree d, we have $\chi (C_s)=3d-d^2+\mu ^{\scriptscriptstyle{\text{tot}}}(C_s)$ (see, e.g., [Reference Wall35, Cor. 7.1.4]).

Next, we look at the zeta-function $\zeta _{\hat g_s\vert _{N(C_s)}}(t)$ . This time, for admissible coordinates $\mathbf {x}=(x_1,x_2,x_3)$ in a neighborhood $U_{\mathbf {p}}$ of a point $\mathbf {p}\in C^{\prime }_s:=C_s\setminus B$ , we may assume that the projection

$$ \begin{align*} p'\colon U_{\mathbf{p}}\cap N(C_s)\to C^{\prime}_s \end{align*} $$

associated with the family of control data (2.3) is given by $\mathbf {x}\mapsto (0,x_2,0)$ , so that $C^{\prime }_s$ is defined by $x_1=x_3=0$ and the restriction of $\hat g_s$ to $p^{\prime -1}(\mathbf {p})$ is given by $x_1^dx_3$ . Then, by the relation (5.2.5) of [Reference Oka24, Chap. I], the normal zeta-function of $\hat g_s$ along $C^{\prime }_s$ is given by $\zeta _{C^{\prime }_s}^\bot (t)=1$ , and hence, by [Reference Oka24, Chap. I, Lems. (5.3) and (5.4)] again, we get

$$ \begin{align*} \zeta_{\hat g_s\vert_{N(C_s)}}(t)=1. \end{align*} $$

As for the zeta-function $\zeta _{\hat g_s\vert _{B_\varepsilon (\mathbf {p}_k)}}(t)$ , since the zeta-multiplicity of $g_s$ is d and the (usual) multiplicity of $\hat g_s$ at $\mathbf {p}_{k}$ is greater than or equal to $d+1$ , it follows from [Reference Oka27, Prop. 11] that $\zeta _{\hat g_s\vert _{B_\varepsilon (\mathbf {p}_k)}}(t)$ does not contribute to the zeta-multiplicity factor of $\zeta _{\hat g_s}(t)$ .

So, altogether, the unique contribution to the zeta-multiplicity factor of $\zeta _{\hat g_s}(t)$ comes from the zeta-function $\zeta _{\hat g_s\vert _{N(E)}}(t)$ and is given by $(1-t^d)^{-3+3d-d^2+\mu ^{\scriptscriptstyle{\text{tot}}}(C_s)}$ .

We can now easily complete the proof of Theorem 1.1 thanks to two theorems of Lê. Indeed, we first observe that if there exists $s_0\in [0,1]$ such that the family $\{\mbox {in}(g_s)\}$ has a bifurcation of the singularities in a small ball B centered at a singular point $\mathbf {p}_0$ of $C_{s_0}$ ,Footnote 5 then, by [Reference Lê17, théorème B] (see also [Reference Haş Bey12], [Reference Lazzeri15]), for $s\not =s_0$ near $s_0$ , we have

$$ \begin{align*} \sum_{\mathbf{p}\in B\cap \Sigma(C_s)}\mu_{\mathbf{p}}(\mbox{in}(g_s)) < \mu_{\mathbf{p}_0}(\mbox{in}(g_{s_0})), \end{align*} $$

and hence $\mu ^{\scriptscriptstyle{\text{tot}}}(C_{s})<\mu ^{\scriptscriptstyle{\text{tot}}}(C_{s_0})$ , which contradicts Lemma 2.3. Therefore, there is no such an $s_0$ . But in this case it follows from [Reference Lê16] and the discussion in [Reference Dimca9, pp. 17–18, 121] that the topological type of the pair $(\mathbb {P}^2,C_s)$ is independent of s, so that $(C_0,C_1)$ is not a Zariski pair—again a contradiction.

Figure 2 Bifurcation of singularities.

Footnotes

1 This means that the Newton diagram $\Gamma (f_l)$ of $f_l$ ( $l\in \{0,1\}$ ) meets each coordinate axis.

2 For instance, this is always the case if the singularities are “simple” in the sense of Arnol’d [Reference Arnol’d2].

3 Hereafter, such coordinates will be called admissible coordinates.

4 Let us briefly show it, for instance, in the special case where the Newton boundaries are as in Figure 1, the general case being completely similar. Clearly, in this case,

$$ \begin{align*} \nu(\Gamma_{\!-}(g_{s_0}\vert_H))=2S'-(d+c)-(d+e)+1, \end{align*} $$

where $S'=S+c\, q/2+e\, p/2$ with $p\geq p_1\geq 2$ and S is the area of the triangle $(0,d,d)$ . Similarly, $\nu (\Gamma _{\!-}(g_{0}\vert _H))=2S-2d+1$ . Since $p\geq 2$ , it follows that

$$ \begin{align*} \nu(\Gamma_{\!-}(g_{s_0}\vert_H))-\nu(\Gamma_{\!-}(g_{0}\vert_H))=c(q-1)+e(p-1)>0 \end{align*} $$

(note that if $q=0$ , then $c=0$ , and the above inequality still holds true).

5 That is, $\mathbf {p}_0$ is the only singular point of $C_{s_0}$ in B and it is either a “newly born” singularity or a singularity obtained as a “merging” of several singularities of $C_s$ for $s\not =s_0$ near $s_0$ . In other words, $s_0$ is a point where the natural projection $\{(\mathbf {p},s)\in \mathbb {P}^2\times [0,1]\, ;\, \mathbf {p}\in \Sigma (C_s)\} \to [0,1]$ fails to be a covering map (see Figure 2).

References

A’Campo, N., La fonction zêta d’une monodromie , Comment. Math. Helv. 50 (1975), 233248.10.1007/BF02565748CrossRefGoogle Scholar
Arnol’d, V. I., Normal forms of functions near degenerate critical points, the Weyl groups ${A}_k$ , ${D}_k$ , ${E}_k$ and Lagrangian singularities , Funkcional. Anal. i Priložen. 6 (1972), no. 4, 325.Google Scholar
Artal-Bartolo, E., Sur la monodromie des singularités superisolées , C. R. Acad. Sci. 312 (1991), no. 8, 601604.Google Scholar
Artal-Bartolo, E., Forme de Jordan de la monodromie des singularités superisolées de surfaces, Memoirs of the American Mathematical Society, Vol. 109 (525), American Mathematical Society, Providence, RI, 1994.Google Scholar
Artal-Bartolo, E., Sur les couples de Zariski , J. Algebraic Geom. 3 (1994), 223247.Google Scholar
Artal-Bartolo, E., Cogolludo, J. I. and Tokunaga, H., “A survey on Zariski pairs” in Algebraic geometry in East Asia–Hanoi 2005, Advanced Studies in Pure Mathematics, Vol. 50, Mathematical Society of Japan, Tokyo, 2008, 1100.Google Scholar
Artal-Bartolo, E., Cogolludo-Agustín, J. I. and Martín-Morales, J., “Cremona transformations of weighted projective planes, Zariski pairs, and rational cuspidal curves” in Singularities and their interaction with geometry and low dimensional topology—In honor of András Némethi, Trends in Mathematics, Birkhäuser/Springer, Cham, 2021, 117157.10.1007/978-3-030-61958-9_7CrossRefGoogle Scholar
Clemens, C. H. Jr., Picard–Lefschetz theorem for families of nonsingular algebraic varieties acquiring ordinary singularities , Trans. Amer. Math. Soc. 136 (1969), 93108.10.1090/S0002-9947-1969-0233814-9CrossRefGoogle Scholar
Dimca, A., Singularities and topology of hypersurfaces, Universitext, Springer, New York, 1992.10.1007/978-1-4612-4404-2CrossRefGoogle Scholar
Eyral, C. and Oka, M., On paths in the $\mu$ -constant and ${\mu}^{\ast }$ -constant strata, to appear in Hiroshima Math. J. 54, no. 2.Google Scholar
Gusein-Zade, S. M., Luengo, I. and Melle-Hernández, A., Partial resolutions and the zeta-function of a singularity , Comment. Math. Helv. 72 (1997), no. 2, 244256.10.1007/s000140050014CrossRefGoogle Scholar
Haş Bey, C., Sur l’irréductibilité de la monodromie locale; application à l’équisingularité , C. R. Acad. Sci. Paris Sér. A–B. 275 (1972), A105A107.Google Scholar
Iomdin, I. N., Complex surfaces with a one-dimensional set of singularities , Sibirsk. Mat. Ž. 15 (1974), 10611082, 1181 (in Russian). English translation: Siberian Math. J. 15 (1974), no. 5, 748–762 (1975).Google Scholar
Kouchnirenko, A. G., Polyèdres de Newton et nombres de Milnor , Invent. Math. 32 (1976), 131.10.1007/BF01389769CrossRefGoogle Scholar
Lazzeri, F., “A theorem on the monodromy of isolated singularities” in Singularités à Cargèse (Rencontre Singularités Géom. Anal., Inst. Études Sci. de Cargèse, 1972), Astérisque, Vols. 7–8, Société Mathématique de France, Paris, 1973, 269275.Google Scholar
, D. T., Sur un critère d’équisingularité, C. R. Acad. Sci. Paris Sér. A–B. 272 (1971), A138A140.Google Scholar
, D. T., Une application d’un théorème d’A’Campo à l’équisingularité, Nederl. Akad. Wetensch. Proc. Ser. A. 76 (= Indag. Math. 35) (1973), 403409.10.1016/1385-7258(73)90064-4CrossRefGoogle Scholar
, D. T., “Topologie des singularités des hypersurfaces complexes” in Singularités à Cargèse (Rencontre Singularités Géom. Anal., Inst. Études Sci., Cargèse, 1972), Astérisque, Vols. 7–8, Société Mathématique de France, Paris, 1973, 171182.Google Scholar
, D. T., “Ensembles analytiques complexes avec lieu singulier de dimension un (d’après I. N. Iomdine)” in Seminar on singularities (Paris, 1976/1977), Publications Mathématiques de l’Université Paris VII, Vol. 7, Université Paris VII, Paris, 1980, 8795.Google Scholar
Luengo, I., The $\mu$ -constant stratum is not smooth , Invent. Math. 90 (1987), no. 1, 139152.10.1007/BF01389034CrossRefGoogle Scholar
Luengo, I. and Melle, A., A formula for the Milnor number , C. R. Acad. Sci. 321 (1995), no. 11, 14731478.Google Scholar
Mather, J., Notes on topological stability , Bull. Amer. Math. Soc. (N.S.). 49 (2012), no. 4, 475506.10.1090/S0273-0979-2012-01383-6CrossRefGoogle Scholar
Milnor, J., Singular points of complex hypersurfaces, Annals of Mathematics Studies, Vol. 61, Princeton University Press, Princeton, NJ; University of Tokyo Press, Tokyo, 1968.Google Scholar
Oka, M., Non-degenerate complete intersection singularity, Hermann, Paris, 1997.Google Scholar
Oka, M., “A survey on Alexander polynomials of plane curves” in Singularités Franco-Japonaises, Séminaire et Congrès, Vol. 10, Société Mathématique de France, Paris, 2005, 209232.Google Scholar
Oka, M., “Almost non-degenerate functions and a Zariski pair of links” in Essays in geometry, dedicated to N. A’Campo, edited by Papadopoulos, A., IRMA Lectures in Mathematics and Theoretical Physics, Vol. 34, European Mathematical Society, Zürich, 2023, 601628.10.4171/irma/34/27CrossRefGoogle Scholar
Oka, M., On $\mu$ -Zariski pairs of links , J. Math. Soc. Japan. 75 (2023), no. 4, 12271259.10.2969/jmsj/89138913CrossRefGoogle Scholar
Saeki, O., Topological types of complex isolated hypersurface singularities , Kodai Math. J. 12 (1989), no. 1, 2329.10.2996/kmj/1138038986CrossRefGoogle Scholar
Siersma, D., The monodromy of a series of hypersurface singularities , Comment. Math. Helv. 65 (1990), no. 2, 181197.10.1007/BF02566602CrossRefGoogle Scholar
Siersma, D., “The vanishing topology of non-isolated singularities” in New developments in singularity theory (Cambridge, 2000), NATO Science Series II: Mathematics, Physics and Chemistry, Vol. 21, Kluwer Academic Publishers, Dordrecht, 2001, 447472.10.1007/978-94-010-0834-1_18CrossRefGoogle Scholar
Stevens, J., On the $\mu$ -constant stratum and the $V$ -filtration: An example , Math. Z. 201 (1989), no. 1, 139144.10.1007/BF01162001CrossRefGoogle Scholar
Teissier, B., “Cycles évanescents, sections planes et conditions de Whitney” in Singularités à Cargèse (Rencontre Singularités Géom. Anal., Inst. Études Sci., Cargèse, 1972), Astérisque, Vols. 7–8, Société Mathématique de France, Paris, 1973, 285362.Google Scholar
Teissier, B., “Déformations à type topologique constant” in Quelques problèmes de modules (Sém. de Géométrie Analytique, École Norm. Sup., Paris, 1971–1972), Astérisque, Vol. 16, Société Mathématique de France, Paris, 1974, 215249.Google Scholar
Varchenko, A. N., Zeta-function of monodromy and Newton’s diagram , Invent. Math. 37 (1976), no. 3, 253262.CrossRefGoogle Scholar
Wall, C. T. C., Singular points of plane curves, London Mathematical Society Student Texts, Vol. 63, Cambridge University Press, Cambridge, 2004.10.1017/CBO9780511617560CrossRefGoogle Scholar
Zariski, O., On the problem of existence of algebraic functions of two variables possessing a given branched curve , Amer. J. Math. 51 (1929), no. 2, 305328.CrossRefGoogle Scholar
Figure 0

Figure 1 Newton diagrams.

Figure 1

Figure 2 Bifurcation of singularities.