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

Characteristic cycles associated to holonomic 𝒟\mathscr{D}-modules

Lei Wu Lei Wu, Department of Mathematics, KU Leuven, Celestijnenlaan 200B, B-3001 Leuven, Belgium [email protected]
Abstract.

We study relative and logarithmic characteristic cycles associated to holonomic 𝒟\mathscr{D}-modules. As applications, we obtain: (1) an alternative proof of Ginsburg’s log characteristic cycle formula for lattices of regular holonomic 𝒟\mathscr{D}-modules following ideas of Sabbah and Briancon-Maisonobe-Merle, and (2) the constructibility of the log de Rham complexes for lattices of holonomic 𝒟\mathscr{D}-modules, which is a natural generalization of Kashiwara’s constructibility theorem.

2010 Mathematics Subject Classification:
14F10, 32S60, 14C17, 32S30, 14A21

1. Introduction

The characteristic variety of a coherent 𝒟\mathscr{D}-module with a good filtration is the support of the associated graded module on the cotangent bundle (see [Kas03]). Characteristic cycles can be obtained with multiplicities taken into account. They can also be considered relative to smooth morphisms (or holomorphic submersions under the analytic setting) and from a logarithmic perspective. See §2.1 and §3 for definitions.

In this paper, we study characteristic cycles of relative 𝒟\mathscr{D}-modules associated to (regular) holonomic 𝒟\mathscr{D}-modules. We apply the relative characteristic cycles to studying the logarithmic characteristic cycles for lattices of regular holonomic 𝒟\mathscr{D}-modules and the constructibility of logarithmic de Rham complexes.

1.1. Constructibility of log de Rham complexes for lattices

For holonomic 𝒟\mathscr{D}-modules on complex manifolds, Kashiwara’s constructibility theorem [Kas75] says that the de Rham complexes of the holonomic modules are constructible (they are indeed perverse, see also [HTT08, Theorem 4.6.6]). Our first two main results are constructibility and perversity of log de Rham complexes of lattices.

Suppose that (X,D)(X,D) is a pair consisting of a complex manifold together with a reduced normal crossing divisor DD, called an analytic smooth log pair. Let 𝒟X,D\mathscr{D}_{X,D} be the sheaf of rings of holomorphic logarithmic differential operators, that is, the sub-sheaf of 𝒟X\mathscr{D}_{X} consisting of differential operators preserving the defining ideal of DD. Let \mathcal{M} be a coherent 𝒟X\mathscr{D}_{X}-module. We now consider a 𝒟X,D\mathscr{D}_{X,D}-lattice ¯\bar{\mathcal{M}} of \mathcal{M}, a special 𝒟X,D\mathscr{D}_{X,D}-module associated to \mathcal{M} (see §3.3 for definition). Typical examples of lattices include Deligne lattices (cf. [WZ21, §4.4]) and lattices given by the graph embedding construction of Malgrange (see §5 for details).

Theorem 1.1.

Suppose that (X,D)(X,D) is an analytic smooth log pair and that \mathcal{M} is a holonomic 𝒟X\mathscr{D}_{X}-module. If ¯\bar{\mathcal{M}} is a 𝒟X,D\mathscr{D}_{X,D}-lattice of \mathcal{M}, then the log de Rham complex DRX,D(¯)\textup{DR}_{X,D}(\bar{\mathcal{M}}) is constructible.

The above theorem naturally generalizes Kashiwara’s constructibility theorem and answers the question at the beginning of [WZ21]. The constructible complex DRX,D(¯)\textup{DR}_{X,D}(\bar{\mathcal{M}}) is not perverse in general and the stratification of the constructible complex DRX,D(¯)\textup{DR}_{X,D}(\bar{\mathcal{M}}) is determined by the stratification of Ch((D))\textup{Ch}(\mathcal{M}(*D)) (see Remark 5.2).

Theorem 1.2.

In the situation of Theorem 1.1, DRX,D(¯(kD))\textup{DR}_{X,D}(\bar{\mathcal{M}}(kD)) are perverse locally on a relative compact open subset of XX ((or globally when XX is algebraic)) for all |k|0|k|\gg 0. Moreover, if \mathcal{M} is regular holonomic, then locally on a relative compact open subset of XX ((or globally when XX is algebraic)) we have natural quasi-isomorphisms

  1. (1)

    DRX,D(¯(kD))q.i.RjDR(|U)\textup{DR}_{X,D}(\bar{\mathcal{M}}(kD))\stackrel{{\scriptstyle q.i.}}{{\simeq}}Rj_{*}\textup{DR}(\mathcal{M}|_{U}),

  2. (2)

    DRX,D(¯(kD))q.i.j!DR(|U)\textup{DR}_{X,D}(\bar{\mathcal{M}}(-kD))\stackrel{{\scriptstyle q.i.}}{{\simeq}}j_{!}\textup{DR}(\mathcal{M}|_{U})

for all integral k0k\gg 0, where j:U=XDXj\colon U=X\setminus D\hookrightarrow X is the open embedding.

Taking the lattice ¯=𝒪X\bar{\mathcal{M}}=\mathscr{O}_{X} in Theorem 1.2 (1), we recover the Grothendieck comparison [Gro66],

[𝒪XΩX1(logD)ΩXn(logD)]q.i.RjU[n],[\mathscr{O}_{X}\rightarrow\Omega^{1}_{X}(\log D)\rightarrow\cdots\rightarrow\Omega^{n}_{X}(\log D)]\stackrel{{\scriptstyle q.i.}}{{\simeq}}Rj_{*}\mathbb{C}_{U}[n],

where nn is the dimension of XX. See also [WZ21, Theorem 1.2]. Theorem 1.2 for lattices given by the graph embedding construction has been used in studying the cohomology support loci of rank one complex local systems [BVWZ21a, BVWZ21b].

The Kashiwara’s constructibility theorem has been extended to Riemann-Hilbert correspondence, the regular case by Kashiwara and Mebkhout independently (see for instance [HTT08, §7]) and the irregular case by D’Agnolo and Kashiwara [DK16]. Under the logarithmic setting, Kato and Nakayama [KN99] and Ogus [Ogu03] studied Riemann-Hilbert correspondence for log connections on smooth log schemes, and Koppensteiner and Taplo [KT19] further studied the theory of logarithmic 𝒟\mathscr{D}-modules on smooth log schemes. Koppensteiner [Kop20] (based on the work of Ogus) augmented the log de Rham complexes to graded complexes on Kato-Nakayama spaces and proved a finiteness result for logarithmic holonomic 𝒟\mathscr{D}-modules. Since smooth log pairs are smooth log schemes, lattices in this paper are special examples of log 𝒟\mathscr{D}-modules in [KT19]. Therefore, one can naturally ask whether Theorem 1.1 together with Theorem 1.2 can be enhanced to a Riemann-Hilbert correspondence on smooth log pairs in the logarithmic category. Notice that our proof of Theorem 1.1 is logarithmic in nature, since it depends on the natural stratification of the normal crossing divisor DD, which gives evidence of the existence of the log Riemann-Hilbert correspondence.

A similar logarithmic Riemann-Hilbert program for log holonomic modules has also been proposed in [KT19]. However, lattices are not logarithmic holonomic [KT19, Definition 3.22] in general and hence Theorem 1.1 is different from the finiteness result in [Kop20]. It would be interesting to further understand relations between lattices and logarithmic holonomic modules.

Our proof of Theorem 1.1 and Theorem 1.2 depends on “transforming" log 𝒟\mathscr{D}-modules to relative 𝒟\mathscr{D}-modules and on the study of relative characteristic cycles. Typical examples of non-trivial relative 𝒟\mathscr{D}-modules arise from the construction of the generalized Kashiwara-Malgrange filtrations. Then we discuss relative characteristic cycles associated to Kashiwara-Malgrange filtrations.

1.2. VV-filtrations along slopes of smooth complete intersections and their relative characteristic cycles

Suppose that XX is a smooth algebraic variety over \mathbb{C} (or a complex manifold) and that YXY\subseteq X is a smooth complete intersection of codimension rr, that is,

Y=j=1rHjY=\bigcap_{j=1}^{r}H_{j}

where jHj\sum_{j}H_{j} is a normal crossing divisor. Let 𝒟X\mathscr{D}_{X} be the sheaf of rings of algebraic (or holomorphic) differential operators. Define a r\mathbb{Z}^{r}-filtration on 𝒟X\mathscr{D}_{X} along YY by

(1) V𝐬𝒟X=j=1rVsjHj𝒟XV_{\mathbf{s}}\mathscr{D}_{X}=\bigcap_{j=1}^{r}V_{s_{j}}^{H_{j}}\mathscr{D}_{X}

for every 𝐬=(s1,s2,sr)r\mathbf{s}=(s_{1},s_{2}\dots,s_{r})\in\mathbb{Z}^{r}, where VHj𝒟XV_{\bullet}^{H_{j}}\mathscr{D}_{X} is the Kashiwara-Malgrange filtration on 𝒟X\mathscr{D}_{X} along HjH_{j} (see Definition 4.1). Following ideas of Sabbah [Sab87a], for a nondegenerate slope LL in the dual cone (0r)(\mathbb{Z}_{\geq 0}^{r})^{\vee}, we define the (generalized) Kashiwara-Malgrange filtration on 𝒟X\mathscr{D}_{X} along LL by

VL(𝐬)L𝒟X=L(𝐬)L(𝐬)V𝐬𝒟X.{}^{L}V_{L(\mathbf{s})}\mathscr{D}_{X}=\sum_{L(\mathbf{s}^{\prime})\leq L(\mathbf{s})}V_{\mathbf{s}^{\prime}}\mathscr{D}_{X}.

This gives a \mathbb{Z}-filtration VL𝒟X{}^{L}V_{\bullet}\mathscr{D}_{X} via the isomorphism r/L\mathbb{Z}\simeq\mathbb{Z}^{r}/L^{\perp}. For a coherent 𝒟X\mathscr{D}_{X}-module \mathcal{M}, one can then define the Kashiwara-Malgrange filtration (or VV-filtration for short) on \mathcal{M} along LL (see Definition 4.14). For degenerate slopes, one can reduce to the nondegenerate ones by ignoring the unrelated HjH_{j}.

The VV-filtration of a coherent 𝒟X\mathscr{D}_{X}-module \mathcal{M} along LL (if exists) contains the “deformation” information of \mathcal{M}. More precisely, the Rees ring RVL𝒟X{}^{L}R_{V}\mathscr{D}_{X} associated to VL𝒟X{}^{L}V_{\bullet}\mathscr{D}_{X} gives the sheaf of differential operators relative to the (algebraic) normal deformation of YY in XX along LL,

φL:X~L,\varphi^{L}:\widetilde{X}^{L}\to\mathbb{C},

where X~L\widetilde{X}^{L} is the ambient space of the deformation. Hence, the Rees module RVL{}^{L}R_{V}\mathcal{M} associated to VL{}^{L}V_{\bullet}\mathcal{M} is a 𝒟\mathscr{D}-module relative to φL\varphi^{L}. See §4.7 for details.

Theorem 1.3 (Sabbah).

Suppose that \mathcal{M} is a holonomic 𝒟X\mathscr{D}_{X}-module and that YY is a smooth complete intersection of codimension rr. Then \mathcal{M} is specializable along every slope vector LL (i.e. the VV-filtration on \mathcal{M} along LL uniquely exists). Moreover, if \mathcal{M} is regular holonomic and the slope vector LL is nondegenerate, then grVL\textup{gr}^{{}^{L}V}_{\bullet}\mathcal{M} gives a regular holonomic 𝒟\mathscr{D}-module on TYXT_{Y}X and we have the following formulas for characteristic cycles,

CCX~L/(RVL)=qL(CC())¯T(X~L/)\textup{CC}_{\widetilde{X}^{L}/\mathbb{C}}({}^{L}R_{V}\mathcal{M})=\overline{q^{*}_{L}(\textup{CC}(\mathcal{M}))}\subseteq T^{*}(\widetilde{X}^{L}/\mathbb{C})

and

CC(gr~VL)=qL(CC(M))¯|TTYXTTYX,\textup{CC}(\widetilde{\textup{gr}}^{{}^{L}V}_{\bullet}\mathcal{M})=\overline{q^{*}_{L}(\textup{CC}(M))}|_{T^{*}T_{Y}X}\subseteq T^{*}T_{Y}X,

where qL:T(X~L/)TTYXTX×TXq_{L}:T^{*}(\widetilde{X}^{L}/\mathbb{C})\setminus T^{*}T_{Y}X\simeq T^{*}X\times\mathbb{C}^{\star}\to T^{*}X is the natural projection.

One can also consider the Rees ring RV𝒟XR_{V}\mathscr{D}_{X} associated to the r\mathbb{Z}^{r}-filtration V𝒟XV_{\bullet}\mathscr{D}_{X}. Similar to RVL𝒟X{}^{L}R_{V}\mathscr{D}_{X}, RV𝒟XR_{V}\mathscr{D}_{X} can be seen as the sheaf of differential operators relative to the (refined) normal deformation of YY in XX (see §4.6),

φ:X~r,\varphi:\widetilde{X}\to\mathbb{C}^{r},

with the r\mathbb{Z}^{r}-grading on RV𝒟XR_{V}\mathscr{D}_{X} induced from the natural toric structure on the base r\mathbb{C}^{r}. Then φL\varphi^{L}, as well as RVL𝒟X{}^{L}R_{V}\mathscr{D}_{X}, is obtained from φ\varphi and RV𝒟XR_{V}\mathscr{D}_{X} respectively through the base-change,

ιL:r\iota_{L}\colon\mathbb{C}\hookrightarrow\mathbb{C}^{r}

induced by the one parameter subgroup of LL. For a 𝒟X\mathscr{D}_{X}-module \mathcal{M} with a good filtration UU_{\bullet}\mathcal{M} over V𝒟XV_{\bullet}\mathscr{D}_{X}, the associated Rees module RUR_{U}\mathcal{M} is then a coherent relative 𝒟\mathscr{D}-module with respect to φ\varphi.

Theorem 1.4 (Sabbah).

Suppose that \mathcal{M} is a regular holonomic 𝒟X\mathscr{D}_{X}-module and that Y is a smooth complete intersection of codimension rr. Let UU_{\bullet}\mathcal{M} be a good r\mathbb{Z}^{r}-filtration over V𝒟XV_{\bullet}\mathscr{D}_{X}. Then we have

CCX~/r(RU)=q(CC())¯T(X~/r)\textup{CC}_{\widetilde{X}/\mathbb{C}^{r}}(R_{U}\mathcal{M})=\overline{q^{*}(\textup{CC}(\mathcal{M}))}\subseteq T^{*}(\widetilde{X}/\mathbb{C}^{r})

where q:T(X~L/)(j=1ruj=0)TX×()rTXq:T^{*}(\widetilde{X}^{L}/\mathbb{C})\setminus(\prod^{r}_{j=1}u_{j}=0)\simeq T^{*}X\times(\mathbb{C}^{\star})^{r}\to T^{*}X is the natural projection and (u1,u2,,ur)(u_{1},u_{2},\dots,u_{r}) are coordinates of r\mathbb{C}^{r}.

Theorem 1.3 and 1.4 are essentially due to Sabbah (see [Sab87b, §2]). Their proof relies on the study of characteristic cycles for relative 𝒟\mathscr{D}-modules (see §4). When L=(1,1,,1r)L=(\underbrace{1,1,\dots,1}_{r}) (use the standard dual basis of (r)(\mathbb{Z}^{r})^{\vee}), the characteristic cycle formula for grVL\textup{gr}^{{}^{L}V}_{\bullet}\mathcal{M} in Theorem 1.3 is due to Ginsburg ([Gin86, Theorem 5.8]). But the precise characteristic cycle formulas in Theorem 1.3 and 1.4 seem to be missing in the literature. It is worth mentioning that Ginsburg use the characteristic cycle formula in [Gin86, Theorem 5.8] to study index theorems. It would be interesting to know whether the characteristic cycle formulas in Theorem 1.3 are related to index theorems or even Fukaya categories (cf. [NZ09]).

1.3. Relative Riemann-Hilbert correspondence

We give an explicit relative Riemann-Hilbert correspondence for RVL{}^{L}R_{V}\mathcal{M}. We assume \mathcal{M} a regular holonomic 𝒟X\mathscr{D}_{X}-module. The relative 𝒟\mathscr{D}-module RVL{}^{L}R_{V}\mathcal{M} has fibers

𝐋iα(RVL)q.i. for 0α and, 𝐋i0(RVL)q.i.gr~VL,\mathbf{L}i_{\alpha}^{*}({}^{L}R_{V}\mathcal{M})\stackrel{{\scriptstyle q.i.}}{{\simeq}}\mathcal{M}\textup{ for $0\not=\alpha\in\mathbb{C}$ and, }\mathbf{L}i_{0}^{*}({}^{L}R_{V}\mathcal{M})\stackrel{{\scriptstyle q.i.}}{{\simeq}}\widetilde{\textup{gr}}^{{}^{L}V}_{\bullet}\mathcal{M},

where q.i.\stackrel{{\scriptstyle q.i.}}{{\simeq}} denotes quasi-isomorphism and iα:X~αL=(φL)1(α)X~Li_{\alpha}:\widetilde{X}^{L}_{\alpha}=(\varphi^{L})^{-1}(\alpha)\hookrightarrow\widetilde{X}^{L} is the closed embedding. Thus, RVL𝒟X{}^{L}R_{V}\mathscr{D}_{X} deforms \mathcal{M} into grVL\textup{gr}^{{}^{L}V}_{\bullet}\mathcal{M}. By Theorem 1.3, RVL{}^{L}R_{V}\mathcal{M} provides an example of relative regular holonomic 𝒟\mathscr{D}-modules (see Definition 2.4 and §2.2). Moreover, using Lemma 2.9 the relative de Rham complex,

ωX~L/𝒟𝐋RVL,\omega_{\widetilde{X}^{L}/\mathbb{C}}\otimes^{\mathbf{L}}_{\mathscr{D}}{}^{L}R_{V}\mathcal{M},

has fibers DR()\textup{DR}(\mathcal{M}) for α0\alpha\not=0 and the central fiber DRTYX(gr~VL)\textup{DR}_{T_{Y}X}(\widetilde{\textup{gr}}^{{}^{L}V}_{\bullet}\mathcal{M}), where ωX~L/\omega_{\widetilde{X}^{L}/\mathbb{C}} is the relative canonical sheaf of φL\varphi_{L} (see §4.3 and Remark 4.13 for explicit formulas). But the central fiber satisfies (applying Theorem 4.5 and Lemma 4.16)

DRTYX(gr~VL)ψTYX(RjLp1(DR(M)))SpTYXL(DR())\textup{DR}_{T_{Y}X}(\widetilde{\textup{gr}}^{{}^{L}V}_{\bullet}\mathcal{M})\simeq\psi_{T_{Y}X}(Rj^{L}_{*}p^{-1}(\textup{DR}(M)))\eqqcolon\textup{Sp}^{L}_{T_{Y}X}(\textup{DR}(\mathcal{M}))

where SpTYXL\textup{Sp}^{L}_{T_{Y}X} is the Verdier specialization along LL by definition, p:X~LTYXXp:\widetilde{X}^{L}\setminus T_{Y}X\to X is the natural projection and jL:X~LTYXX~Lj^{L}:\widetilde{X}^{L}\setminus T_{Y}X\hookrightarrow\widetilde{X}^{L} is the open embedding. Thus, the relative de Rham complex deforms DR()\textup{DR}(\mathcal{M}) into its Verdier specialization along LL. In particular, the relative de Rham complex gives a relative constructible complex (cf. [MFS16]).

In general, a relative regular Riemann-Hilbert correspondence for relative regular holonomic 𝒟\mathscr{D}-modules over curves is established in [FFS20]. However, the relative holonomicity in loc. cit. is more restricted. Motivated by the above example, one might ask whether the relative Riemann-Hilbert correspondence over curves can be extended to the case by using the relative holonomicity in Definition 2.4.

In contrast to RVL{}^{L}R_{V}\mathcal{M}, RUR_{U}\mathcal{M} is not necessarily relatively holonomic because of the main obstruction that RUR_{U}\mathcal{M} is only torsion-free but not necessarily flat over r\mathbb{C}^{r}. Consequently, one cannot “normalize" UU_{\bullet}\mathcal{M} into a r\mathbb{Z}^{r}-indexed VV-filtration in general. See Remark 4.22 for more discussions.

Proposition 1.5.

In the situation of Theorem 1.4, if RUR_{U}\mathcal{M} is flat over r\mathbb{C}^{r}, then RUR_{U}\mathcal{M} is relative holonomic.

1.4. Logarithmic characteristic cycles of lattices

If D=i=1rHjD=\sum_{i=1}^{r}H_{j} is a normal crossing divisor, then

𝒟X,D=V0𝒟X\mathscr{D}_{X,D}=V_{\vec{0}}\mathscr{D}_{X}

where the latter is defined in Eq.(1). This means that if U((D))U_{\bullet}(\mathcal{M}(*D)) is a good filtration over the r\mathbb{Z}^{r}-filtration V𝒟XV_{\bullet}\mathscr{D}_{X}, then each filtrant U𝐬(D)U_{\mathbf{s}}\mathcal{M}(*D) is a lattice of \mathcal{M}. This is an easy relation between log 𝒟\mathscr{D}-modules and relative 𝒟\mathscr{D}-modules. See §3.4 for a complicated relation between them. Our next main result is an alternative proof of Ginsburg’s log characteristic cycle formula based on the complicated relation.

Theorem 1.6 (Ginsburg).

Suppose that (X,D)(X,D) is an analytic smooth log pair and that \mathcal{M} is a regular holonomic 𝒟X\mathscr{D}_{X}-module. If ¯\bar{\mathcal{M}} is a 𝒟X,D\mathscr{D}_{X,D}-lattice of \mathcal{M}, then

CClog(¯)=CC¯log(|U),\textup{CC}_{\log}(\bar{\mathcal{M}})=\overline{\textup{CC}}^{\log}(\mathcal{M}|_{U}),

where CC¯log(|U)\overline{\textup{CC}}^{\log}(\mathcal{M}|_{U}) denotes the closure of CC(U)TU\textup{CC}(\mathcal{M}_{U})\subseteq T^{*}U inside the logarithmic cotangent bundle T(X,D)T^{*}(X,D).

Ginsburg’s proof of the above theorem under the algebraic setting in [Gin89, Appendix A] uses microlocalization of 𝒟\mathscr{D}-modules and resolution of singularities. Our proof under the analytic setting relies on converting log 𝒟\mathscr{D}-modules to relative 𝒟\mathscr{D}-modules, with some ideas due to Sabbah and Briancon-Maisonobe-Merle.

Theorem 1.6 has a long history. A characteristic variety formula of the holonomic system for l=1N(fl+1O)λl\prod_{l=1}^{N}(f_{l}+\sqrt{-1}{O})^{\lambda_{l}} first appears in [KK79]. For the lattice 𝒟X[𝐬](𝐟𝐬m)\mathscr{D}_{X}[\mathbf{s}]({\bf f}^{\mathbf{s}}\cdot m) given by the graph embedding construction (see §5), the formula for characteristic varieties is proved by Briancon-Maisonobe-Merle [BMM02, Théorèm 2.1] and the characteristic cycle formula in this case is obtained in [BVWZ21b]. Ginsburg [Gin89] made it in its most general form as in Theorem 1.6 under the algebraic setting. See [Kas77, Gin86, Gin89, BVWZ21b, Mai16a, Mai16b] for applications of Theorem 1.6. See also [CM99, CMNM02] for related applications.

In a follow-up paper, Theorem 1.6 in its general form is used to obtain the conclusion that the zero loci of Bernstein-Sato ideals for regular holonomic 𝒟\mathscr{D}-modules in general are always of codimension-one [Wu21, Theorem 3.11]. This conclusion in turn plays an important role in the establishment of the Riemann-Hilbert correspondence for Alexander complexes (see [Wu21, §3]).

The following example shows that regularity in both Theorem 1.6 and Theorem 1.3 is needed.

Example 1.

We consider =[t,1/t]e1/t\mathcal{M}=\mathbb{C}[t,1/t]\cdot e^{1/t}, the algebraic irregular holonomic 𝒟X\mathscr{D}_{X}-module generated by the function e1/te^{1/t} for X=X=\mathbb{C}, where tt is the coordinate of the complex plane \mathbb{C}. Let HH be the divisor {0}X\{0\}\subseteq X. Since

tte1/t=e1/t/t,t\partial_{t}\cdot e^{1/t}=-e^{1/t}/t,

\mathcal{M} is coherent over 𝒟X,H\mathscr{D}_{X,H}. Hence, \mathcal{M} is a 𝒟X,H\mathscr{D}_{X,H}-lattice of itself and its VV-filtration along HH is the trivial filtration with grV=0.\textup{gr}^{V}_{\bullet}\mathcal{M}=0. Moreover, since

(1+t2t)e1/t=0,(1+t^{2}\partial_{t})\cdot e^{1/t}=0,

one can see that Chlog()\textup{Ch}_{\log}(\mathcal{M}) has a component over HH.

1.5. Key ideas in proving main results

For Theorem 1.1, we first use direct images of log 𝒟X\mathscr{D}_{X}-modules (see §3.2) under the graph embedding of the defining functions of the normal crossing divisor locally to reduce to the case for lattices 𝒟X[𝐬](𝐟𝐬0)\mathscr{D}_{X}[\mathbf{s}]({\bf f}^{\mathbf{s}}\cdot\mathcal{M}_{0}). The lattices 𝒟X[𝐬](𝐟𝐬0)\mathscr{D}_{X}[\mathbf{s}]({\bf f}^{\mathbf{s}}\cdot\mathcal{M}_{0}) can be treated as relative 𝒟\mathscr{D}-modules over [𝐬]\mathbb{C}[\mathbf{s}] with independent parameters

𝐬=(s1,s2,,sr).\mathbf{s}=(s_{1},s_{2},\dots,s_{r}).

We then use the fact that 𝒟X[𝐬](𝐟𝐬0)\mathscr{D}_{X}[\mathbf{s}]({\bf f}^{\mathbf{s}}\cdot\mathcal{M}_{0}) is indeed relative holonomic, see Theorem 5.1, which generalizes a relative holonomicity result of Maisonobe [Mai16a]. We then identify the log de Rham complex of 𝒟X[𝐬](𝐟𝐬0)\mathscr{D}_{X}[\mathbf{s}]({\bf f}^{\mathbf{s}}\cdot\mathcal{M}_{0}) as the fiber of the relative de Rham complexes over the origin 𝟎r=Spec [𝐬]{\mathbf{0}}\in\mathbb{C}^{r}=\textup{Spec }\mathbb{C}[\mathbf{s}]. Finally, Theorem 1.1 follows from the relative holonomicity result and Kashiwara’s constructibility theorem for complexes of 𝒟\mathscr{D}-modules with holonomic cohomologies. To prove Theorem 1.2, we reduce the required perversity to the flatness of the twisted lattices

𝒟X[𝐬](𝐟𝐬0(kD))\mathscr{D}_{X}[\mathbf{s}]({\bf f}^{\mathbf{s}}\cdot\mathcal{M}_{0}(kD))

over a small (analytic) neighborhood of 𝟎r{\mathbf{0}}\in\mathbb{C}^{r} for |k|0|k|\gg 0 by applying Sabbah’s generalized bb-functions.

The key point of the proof of Theorem 1.6 is to interpret log 𝒟\mathscr{D}-modules as certain relative 𝒟\mathscr{D}-modules. More precisely, we use what we call the log rescaled families (locally) to convert lattices to relative 𝒟\mathscr{D}-modules over the log factor. See §3.4 for constructions. Finally we apply a relative characteristic variety formula of Sabbah and Briancon-Maisonobe-Merle (see Lemma 3.4).

1.6. Relations to singularities in algebraic geometry

We first clarify the Bernsten-Sato polynomials (or bb-functions) in this paper and in the literature.

The bb-functions in Definition 4.14 are the natural generalization of the bb-functions for the usual VV-filtrations (see §4.1 and also [Bj93, §III.7]). For a holomorphic function ff and for the lattices 𝒟X[s](fs0)\mathscr{D}_{X}[s](f^{s}\cdot\mathcal{M}_{0}) (that is, r=1r=1), the associated bb-function is the monic polynomial b(s)[s]b(s)\in\mathbb{C}[s] of the least degree such that

b(s)𝒟X[s](fs0)𝒟X[s](fs+10)=0,b(s)\cdot\frac{\mathscr{D}_{X}[s](f^{s}\cdot\mathcal{M}_{0})}{\mathscr{D}_{X}[s](f^{s+1}\cdot\mathcal{M}_{0})}=0,

where 0\mathcal{M}_{0} is an 𝒪X\mathscr{O}_{X}-coherent submodule of a holonomic 𝒟X\mathscr{D}_{X}-module. See [Bj93, III.2 and VI.1] for this case. If 0=𝒪X\mathcal{M}_{0}=\mathscr{O}_{X}, the above definition gives particularly the bb-function for ff. For the lattice 𝒟X[𝐬](𝐟𝐬0)\mathscr{D}_{X}[\mathbf{s}]({\bf f}^{\mathbf{s}}\cdot\mathcal{M}_{0}) in general, since sis_{i} is identified with titi-t_{i}\partial_{t_{i}}, Theorem 4.23 gives a polynomial b(𝐬)[𝐬]b(\mathbf{s})\in\mathbb{C}[\mathbf{s}] such that

b(𝐬)𝒟X[𝐬](𝐟𝐬0)𝒟X[𝐬](𝐟𝐬+10)=0,b(\mathbf{s})\cdot\frac{\mathscr{D}_{X}[\mathbf{s}]({\bf f}^{\mathbf{s}}\cdot\mathcal{M}_{0})}{\mathscr{D}_{X}[\mathbf{s}]({\bf f}^{\mathbf{s}+\vec{1}}\cdot\mathcal{M}_{0})}=0,

with 1=(1,1,,1)\vec{1}=(1,1,\dots,1). Such b(𝐬)b(\mathbf{s}) are what we mean by Sabbah’s generalized bb-functions. This can be further generalized to the definition of the Bernstein-Sato ideal B𝐟B_{\bf f} of 𝐟{\bf f} (see for instance [Bud15]).

Now we discuss the relations in between VV-filtrations (and/or bb-functions) and singularities in algebraic geometry. The bb-function of ff provides a useful tool to study singularities of the divisor of ff, see for instance [Kas77, ELS+04, BS05, Kol97, Ste88]. For multiple functions, Budur, Mustaţǎ and Saito [BMS06] defined the Bernstein-Sato polynomials (or bb-functions) for arbitrary schemes in X=nX=\mathbb{C}^{n} by considering the VV-filtration along smooth subvarieties (see Definition 4.2).111The VV-filtration in loc. cit. is the \mathbb{Q}-indexed one. One can refine the \mathbb{Z}-index to the \mathbb{Q}-index by a standard procedure using bb-functions. More precisely, they considered the VV-filtration and bb-functions along the slope L=(1,1,,1)L=(1,1,\dots,1) for the holonomic module ι𝐟+𝒪X\iota_{{\bf f}+}\mathscr{O}_{X}, where 𝐟=(f1,,fr){\bf f}=(f_{1},\dots,f_{r}) generate the ideal of an affine scheme. They can reinterpret the log-canonical threshold as well as other jumping coefficients of the multiplier ideals of the scheme [Laz] as roots of the bb-function of the VV-filtration ([BMS06, Theorem 2]). By Theorem 4.23, one can see that if L=(1,,1)L=(1,\dots,1) is not a slope of an adapted cone of certain good r\mathbb{Z}^{r}-filtration on ι𝐟+𝒪X\iota_{{\bf f}+}\mathscr{O}_{X}, then the bb-function of the scheme is irrelevant to the generalized bb-function of Sabbah. In fact, this can be further refined in terms of Bernstein-Sato ideals with the help of a result of Maisonobe. More precisely, by [Mai16a, Résultat 4], if L=(1,,1)L=(1,\dots,1) is not a slope of the codimension one components of the zero locus of B𝐟B_{\bf f}, then the bb-function of the scheme defined by 𝐟{\bf f} is irrelevant to the Bernstein-Sato ideal of 𝐟{\bf f}.

By Theorem 1.3, one can now consider the VV-filtration and the bb-function of ι𝐟+𝒪X\iota_{{\bf f}+}\mathscr{O}_{X} along an arbitrary slope LL. It would be interesting to know whether there exist algebro-geometric interpretations of the jumping indices of the VV-filtration and the roots of the bb-function of ι𝐟+𝒪X\iota_{{\bf f}+}\mathscr{O}_{X} along LL.

Mustaţǎ and Popa [MP19a, MP19b, MP20] defined Hodge ideals (see also [Sai16]) by considering the Hodge filtration of the 𝒟\mathscr{D}-module 𝒪X(D)\mathscr{O}_{X}(*D) using mixed Hodge modules. It is then natural to ask whether there exist connections between VV-filtrations of ι𝐟+𝒪X(D)\iota_{{\bf f}+}\mathscr{O}_{X}(*D) along LL and Hodge ideals of DD, where ifi=0\prod_{i}f_{i}=0 defines the divisor DD.

1.7. Outline

In §2, we discuss the general theory of relative 𝒟\mathscr{D}-modules. Then we discuss log 𝒟\mathscr{D}-modules and the proof of Theorem 1.6 in §3. §4 is about the generalized VV-filtrations and their relations with relative 𝒟\mathscr{D}-modules. Most of the results in §4 are essentially due to Sabbah. We give a down-to-earth exposition in §4, for the reason that the beautiful theory of multi-indexed filtrations of Sabbah seems to be not widely known. Also, the construction of VV-filtrations along arbitrary slopes seems to be missing in the literature, to the best of our knowledge. For instance, as mentioned above only the VV-filtration along the slope L=(1,,1)L=(1,\dots,1) was studied in [BMS06]. Finally, we recall the graph construction of Malgrange and prove the constructibility of log de Rham complexes in §5.

1.8. Convention

Throughout this paper, we discuss sheaves of modules on either algebraic or analytic spaces (or both). If the underlying space is algebraic (resp. analytic), then the sheaves of modules on it are all assumed to be algebraic (resp. analytic). But, when we discuss constructible complexes (or perverse sheaves) on a complex algebraic variety, we always use the Euclidean topology. If f:XYf:X\to Y is a morphism of (algebraic or analytic) schemes, we use f1f^{-1} and ff_{*} to denote the sheaf-theoretical inverse and direct image functors respectively.

Acknowledgement

The author thanks Peng Zhou and Nero Budur for useful discussions, Claude Sabbah for answering questions and Yajnaseni Dutta and Ruijie Yang for useful comments.

2. Relative 𝒟\mathscr{D}-modules

2.1. Relative characteristic cycles

We recall the theory of 𝒟\mathscr{D}-modules under the relative setting. We mainly follow [Sch12, Chapter III. 1.3]. Suppose that φ:𝒳𝒮\varphi\colon\mathcal{X}\to\mathcal{S} is a smooth morphism (i.e. dφd\varphi is surjective everywhere) of complex smooth algebraic varieties (or complex manifolds). We write by 𝒯𝒳/𝒮\mathscr{T}_{\mathcal{X}/\mathcal{S}} the sheaf of vector fields tangent to the leaves of ϕ\phi. We then have an inclusion

𝒯𝒳/𝒮𝒯𝒳.\mathscr{T}_{\mathcal{X}/\mathcal{S}}\hookrightarrow\mathscr{T}_{\mathcal{X}}.

Then the sheaf of rings of relative differential operators associated to φ\varphi is defined to be the subalgebra

𝒟φ=𝒟𝒳/𝒮𝒟X\mathscr{D}_{\varphi}=\mathscr{D}_{\mathcal{X}/\mathcal{S}}\subseteq\mathscr{D}_{X}

generated by 𝒯X/Y\mathscr{T}_{X/Y} and 𝒪X\mathscr{O}_{X}. Similar to the absolute case, 𝒟𝒳/𝒮\mathscr{D}_{\mathcal{X}/\mathcal{S}} is a coherent and noetherian sheaf of rings. Modules over 𝒟𝒳/𝒮\mathscr{D}_{\mathcal{X}/\mathcal{S}} are called relative 𝒟\mathscr{D}-modules over 𝒮\mathcal{S}. We also write by Ω𝒳/𝒮1\Omega^{1}_{\mathcal{X}/\mathcal{S}} the relative cotangent sheaf which is defined to be the 𝒪\mathscr{O}-dual of 𝒯𝒳/𝒮\mathscr{T}_{\mathcal{X}/\mathcal{S}}. Since 𝒟𝒳/𝒮\mathscr{D}_{\mathcal{X}/\mathcal{S}} is not commutative, we have both right and left 𝒟𝒳,𝒮\mathscr{D}_{\mathcal{X},\mathcal{S}}-modules and the side-change operator is given by tensoring ω𝒳/𝒮mΩ𝒳/𝒮1\omega_{\mathcal{X}/\mathcal{S}}\coloneqq\wedge^{m}\Omega^{1}_{\mathcal{X}/\mathcal{S}} with its quasi-inverse by tensoring ω𝒳/𝒮1\omega^{-1}_{\mathcal{X}/\mathcal{S}}, where m=dim𝒳dim𝒮m=\dim\mathcal{X}-\dim\mathcal{S}.

Since φ\varphi is smooth, we have a short exact sequence of cotangent bundles

0𝒳×𝒮T𝒮TXT(𝒳/S)0.0\to\mathcal{X}\times_{\mathcal{S}}T^{*}\mathcal{S}\to T^{*}X\to T^{*}(\mathcal{X}/S)\to 0.

The filtration F𝒟XF_{\bullet}\mathscr{D}_{X} by the orders of differential operators induces on 𝒟𝒳/𝒮\mathscr{D}_{\mathcal{X}/\mathcal{S}} the order filtration F𝒟𝒳/SF_{\bullet}\mathscr{D}_{\mathcal{X}/S}. Then the associated graded sheaf of rings grF𝒟𝒳/S\textup{gr}_{\bullet}^{F}\mathscr{D}_{\mathcal{X}/S} gives the algebraic structure sheaf of T(𝒳/S)T^{*}(\mathcal{X}/S) by the \sim-functor. In the analytic case, 𝒪T(𝒳/S)\mathscr{O}_{T^{*}(\mathcal{X}/S)} is a faithfully flat ring extension of gr~F𝒟𝒳/S\widetilde{\textup{gr}}_{\bullet}^{F}\mathscr{D}_{\mathcal{X}/S} by GAGA.

For a coherent 𝒟𝒳/𝒮\mathscr{D}_{\mathcal{X}/\mathcal{S}}-module \mathscr{M}, a filtration FF_{\bullet}\mathscr{M} over F𝒟𝒳/𝒮F_{\bullet}\mathscr{D}_{\mathcal{X}/\mathcal{S}} is called good if grF\textup{gr}^{F}_{\bullet}\mathscr{M} is coherent over grF𝒟𝒳/𝒮\textup{gr}^{F}_{\bullet}\mathscr{D}_{\mathcal{X}/\mathcal{S}}. Conversely, if there exists a filtration FF_{\bullet}\mathscr{M} satisfying that grF\textup{gr}^{F}_{\bullet}\mathscr{M} is coherent over grF𝒟𝒳/𝒮\textup{gr}^{F}_{\bullet}\mathscr{D}_{\mathcal{X}/\mathcal{S}}, then \mathscr{M} is coherent over 𝒟𝒳/𝒮\mathscr{D}_{\mathcal{X}/\mathcal{S}}. Good filtrations for coherent 𝒟𝒳/𝒮\mathscr{D}_{\mathcal{X}/\mathcal{S}}-modules exist locally in the analytic category and globally in the algebraic category. We define the relative characteristic variety by

Chrel()=Ch𝒳/𝒮()supp(gr~F)T(𝒳/𝒮),\textup{Ch}_{{\textup{rel}}}(\mathscr{M})=\textup{Ch}_{\mathcal{X}/\mathcal{S}}(\mathscr{M})\coloneqq\textup{supp}(\widetilde{\textup{gr}}^{F}_{\bullet}\mathscr{M})\subseteq T^{*}(\mathcal{X}/\mathcal{S}),

where we use the \sim-functor of the affine morphism

π:T(𝒳/𝒮)X.\pi:T^{*}(\mathcal{X}/\mathcal{S})\to X.

By construction, Chrel()\textup{Ch}_{\textup{rel}}(\mathscr{M}) is a conic subvariety of T(𝒳/𝒮)T^{*}(\mathcal{X}/\mathcal{S}), where “conic" means that it is invariant under the natural \mathbb{C}^{\star}-action induced by the grading on grF𝒟𝒳/𝒮\textup{gr}^{F}_{\bullet}\mathscr{D}_{\mathcal{X}/\mathcal{S}}. Each irreducible components of Chrel()\textup{Ch}_{{\textup{rel}}}(\mathscr{M}) has a multiplicity. Similar to the absolute case, Chrel()\textup{Ch}_{{\textup{rel}}}(\mathscr{M}) and the multiplicities are independent of good filtrations. Then the relative characteristic cycle, denoted by CCrel()\textup{CC}_{{\textup{rel}}}(\mathscr{M}) is the associated locally finite cycles of Chrel()\textup{Ch}_{{\textup{rel}}}(\mathscr{M}) with multiplicities. If φ\varphi is an algebraic smooth morphism between smooth varieties over \mathbb{C}, then CCrel()\textup{CC}_{{\textup{rel}}}(\mathscr{M}) is an algebraic cycle inside the algebraic relative cotangent bundle T(𝒳/𝒮)T^{*}(\mathcal{X}/\mathcal{S}).

Similar to the absolute case, for a relative differential operator P𝒟𝒳/𝒮P\in\mathscr{D}_{\mathcal{X}/\mathcal{S}} of order kk, we can define its principal symbol, which gives a section of homogeneous degree kk in grF𝒟𝒳/𝒮\textup{gr}^{F}_{\bullet}\mathscr{D}_{\mathcal{X}/\mathcal{S}}. By [Bj93, 3.24 Definition], we obtain the relative Poisson bracket on grF𝒟𝒳/𝒮\textup{gr}^{F}_{\bullet}\mathscr{D}_{\mathcal{X}/\mathcal{S}} and hence the relative Poisson bracket on 𝒪T(𝒳/𝒮)\mathscr{O}_{T^{*}(\mathcal{X}/\mathcal{S})} (by faithful flatness). A subvariety of T(𝒳/𝒮)T^{*}(\mathcal{X}/\mathcal{S}) is called (relative) involutive if its radical ideal sheaf is closed under the Poisson bracket. Then by Gabber’s involutive theorem (see [Bj93, Appendix III, 3.25 Theorem]), we obtain:

Theorem 2.1 (Gabber’s Involutivity).

Suppose that \mathscr{M} is a coherent 𝒟𝒳/𝒮\mathscr{D}_{\mathcal{X}/\mathcal{S}}-module ((left or right)). Then Chrel()\textup{Ch}_{\textup{rel}}(\mathscr{M}) is ((relative)) involutive.

Notice that the fibers of a relative involutive subvariety 𝒵T(𝒳/𝒮)\mathcal{Z}\subseteq T^{*}(\mathcal{X}/\mathcal{S}) are not necessarily involutive. One reason is that the intersections

𝒵s=𝒵T(𝒳/𝒮)sT𝒳s\mathcal{Z}_{s}=\mathcal{Z}\cap T^{*}(\mathcal{X}/\mathcal{S})_{s}\subseteq T^{*}\mathcal{X}_{s}

are not always proper intersections for s𝒮s\in\mathcal{S}. If additionally 𝒵\mathcal{Z} is smooth over 𝒮\mathcal{S}, then one can easily check that 𝒵sT𝒳s\mathcal{Z}_{s}\subseteq T^{*}\mathcal{X}_{s} is either empty or involutive. However, we have the following relative Bernstein inequality:

Theorem 2.2 (Relative Bernstein Inequality of Maisonobe).

Suppose that \mathscr{M} is a coherent 𝒟𝒳/𝒮\mathscr{D}_{\mathcal{X}/\mathcal{S}}-module ((left or right)). If Chrel()s\textup{Ch}_{\textup{rel}}(\mathscr{M})_{s} is not empty for s𝒮s\in\mathcal{S}, then all the irreducible components of Chrel()s\textup{Ch}_{\textup{rel}}(\mathscr{M})_{s} are of dimension dim𝒳dim𝒮\geq\dim\mathcal{X}-\dim\mathcal{S}.

Proof.

The proof is essentially the same as that of [Mai16a, Proposition 5], where the author only discussed relative 𝒟\mathscr{D}-modules over 𝒮=r\mathcal{S}=\mathbb{C}^{r}. For completeness, we sketch the proof in general. We take a smooth point of Chrel()\textup{Ch}_{\textup{rel}}(\mathscr{M}) and focus on an open neighborhood WW around it. By generic smoothness (or Morse-Sard Theorem for critical values in the analytic case), Chrel()W\textup{Ch}_{\textup{rel}}(\mathscr{M})\cap W is smooth over an open neighborhood UU of 𝒮\mathcal{S} (shrink WW if necessary). Then the relative involutivity in Theorem 2.1 and the relative smoothness imply that (Chrel()W)T𝒳s(\textup{Ch}_{\textup{rel}}(\mathscr{M})\cap W)\cap T^{*}\mathcal{X}_{s} is involutive in T𝒳sT^{*}\mathcal{X}_{s} for s𝒮s\in\mathcal{S}. In particular, the dimension of

(Chrel()W)T𝒳s(\textup{Ch}_{\textup{rel}}(\mathscr{M})\cap W)\cap T^{*}\mathcal{X}_{s}

is dim𝒳dim𝒮\geq\dim\mathcal{X}-\dim\mathcal{S}. Therefore, the required statement follows from the upper semicontinuity of the dimension of fibers of Chrel()\textup{Ch}_{\textup{rel}}(\mathscr{M}). ∎

The following lemma is the relative analogue of [Kas03, Proposition 2.10]. We leave its proof for interested readers. See also [BVWZ21a, Lemma 3.2.2].

Lemma 2.3.

If

0′′00\to\mathscr{M}^{\prime}\longrightarrow\mathscr{M}\longrightarrow\mathscr{M}^{\prime\prime}\to 0

is a short exact sequence of coherent 𝒟𝒳/𝒮\mathscr{D}_{\mathcal{X}/\mathcal{S}}-modules, then

Chrel()=Chrel()Chrel(′′).\textup{Ch}_{\textup{rel}}(\mathscr{M})=\textup{Ch}_{\textup{rel}}(\mathscr{M}^{\prime})\cup\textup{Ch}_{\textup{rel}}(\mathscr{M}^{\prime\prime}).

Following [Sab87b], we define the relative holonomicity as follows.

Definition 2.4 (Relative holonomicity).

A coherent 𝒟𝒳/𝒮\mathscr{D}_{\mathcal{X}/\mathcal{S}}-module \mathscr{M} is called relative holonomic over 𝒮\mathcal{S} (or 𝒪S\mathscr{O}_{S}) if its characteristic variety Chrel()\textup{Ch}_{\textup{rel}}(\mathscr{M}) is relative Lagrangian, that is, the fiber Chrel()s\textup{Ch}_{\textup{rel}}(\mathscr{M})_{s} is either empty or a (possibly reducible) Lagrangian subvariety in T𝒳sT^{*}\mathcal{X}_{s} for every s𝒮s\in\mathcal{S}.222This relative holonomicity is slightly more general than the ones in [Mai16a, MFS16, BVWZ21a, BVWZ21b], where the latter requires additionally that the relative Lagrangian subvarieties are independent of s𝒮=rs\in\mathcal{S}=\mathbb{C}^{r}.

Following from Lemma 2.3 and Theorem 2.2, we immediately have:

Corollary 2.5.

Relative holonomicity is preserved by subquotients and extensions. In particular, the category of relative holonomic modules is abelian.

2.2. Base change for relative 𝒟\mathscr{D}-modules

We now discuss the base change for relative 𝒟X\mathscr{D}_{X}-modules. Suppose we have the following commutative diagram,

(2) 𝒳𝒯{\mathcal{X}_{\mathcal{T}}}𝒳{\mathcal{X}}𝒯{\mathcal{T}}𝒮{\mathcal{S}}μ\scriptstyle{\mu}ν\scriptstyle{\nu}

so that 𝒳𝒯=𝒳×𝒮𝒯\mathcal{X}_{\mathcal{T}}=\mathcal{X}\times_{\mathcal{S}}\mathcal{T}. Suppose \mathscr{M} is a (left) 𝒟𝒳/𝒮\mathscr{D}_{\mathcal{X}/\mathcal{S}}-module. We consider the 𝒪\mathscr{O}-pullback through μ\mu:

μ()=𝒪𝒳𝒯μ1𝒪𝒳μ1.\mu^{*}(\mathscr{M})=\mathscr{O}_{\mathcal{X}_{\mathcal{T}}}\otimes_{\mu^{-1}\mathscr{O}_{\mathcal{X}}}\mu^{-1}\mathscr{M}.

Since μ𝒟𝒳/𝒮=𝒟𝒳𝒯/𝒯\mu^{*}{\mathscr{D}_{\mathcal{X}/\mathcal{S}}}=\mathscr{D}_{\mathcal{X}_{\mathcal{T}}/\mathcal{T}}, μ()\mu^{*}(\mathscr{M}) is naturally a relative 𝒟\mathscr{D}-module over 𝒯\mathcal{T}. We then have the derived pullback functor 𝐋μ\mathbf{L}\mu^{*} for relative 𝒟\mathscr{D}-modules. When the relative 𝒟\mathscr{D}-module structure is forgotten, it is exactly the derived 𝒪\mathscr{O}-module pullback functor. One can easily see that the derived functor 𝐋μ\mathbf{L}\mu^{*} for relative 𝒟\mathscr{D}-module preserves coherence, thanks to the identification μ𝒟𝒳/𝒮=𝒟𝒳𝒯/𝒯\mu^{*}{\mathscr{D}_{\mathcal{X}/\mathcal{S}}}=\mathscr{D}_{\mathcal{X}_{\mathcal{T}}/\mathcal{T}} again.

We write by is:𝒳s𝒳i_{s}\colon\mathcal{X}_{s}\hookrightarrow\mathcal{X} the closed embedding of the fiber over s𝒮s\in\mathcal{S}. A relative holonomic 𝒟𝒳/𝒮\mathscr{D}_{\mathcal{X}/\mathcal{S}}-module \mathcal{M} is regular if 𝐋is()\mathbf{L}i^{*}_{s}(\mathcal{M}) is a complex of 𝒟𝒳s\mathscr{D}_{\mathcal{X}_{s}}-modules with regular holonomic cohomology sheaves for every s𝒮s\in\mathcal{S}. The author is told by C. Sabbah that it is not known whether the category of regular relative holonomic 𝒟\mathscr{D}-module is closed by taking subquotients.

For a closed subvariety 𝒵𝒮\mathcal{Z}\subseteq\mathcal{S}, we denote by i𝒵:𝒳𝒵𝒳i_{\mathcal{Z}}\colon\mathcal{X}_{\mathcal{Z}}\hookrightarrow\mathcal{X} the closed embedding with 𝒳𝒵=𝒳×𝒮𝒵\mathcal{X}_{\mathcal{Z}}=\mathcal{X}\times_{\mathcal{S}}\mathcal{Z}.

Lemma 2.6.

If \mathscr{M} is relative holonomic over 𝒮\mathcal{S} and 𝒵\mathcal{Z} is a smooth subvariety, then 𝐋i𝒵()\mathbf{L}i_{\mathcal{Z}}^{*}(\mathcal{M}) is a complex of relative holonomic cohomology sheaves over 𝒵\mathcal{Z}. In particular, 𝐋kis\mathbf{L}^{k}i_{s}^{*}\mathcal{M} is a holonomic 𝒟𝒳s\mathscr{D}_{\mathcal{X}_{s}}-module for each kk.

Proof.

It is obvious that 𝐋ki𝒵()\mathbf{L}^{k}i_{\mathcal{Z}}^{*}(\mathscr{M}) is coherent over 𝒟𝒳𝒵/𝒵\mathscr{D}_{\mathcal{X}_{\mathcal{Z}}/\mathcal{Z}} and that

Chrel(𝐋ki𝒵())Chrel()|𝒵Chrel()𝒳𝒵.\textup{Ch}_{\textup{rel}}(\mathbf{L}^{k}i_{\mathcal{Z}}^{*}(\mathscr{M}))\subseteq\textup{Ch}_{\textup{rel}}(\mathscr{M})|_{\mathcal{Z}}\coloneqq\textup{Ch}_{\textup{rel}}(\mathscr{M})\cap\mathcal{X}_{\mathcal{Z}}.

But Chrel()|𝒵\textup{Ch}_{\textup{rel}}(\mathscr{M})|_{\mathcal{Z}} is relative Lagrangian over 𝒵\mathcal{Z}. Therefore, 𝐋ki𝒵()\mathbf{L}^{k}i_{\mathcal{Z}}^{*}(\mathscr{M}) is relative holonomic by Theorem 2.2. ∎

Proposition 2.7.

Suppose that ϕ:𝒳𝒮\phi:\mathcal{X}\to\mathcal{S} is smooth with 𝒮\mathcal{H}\subset\mathcal{S} a smooth divisor, and that \mathscr{M} is a coherent 𝒟𝒳/𝒮\mathscr{D}_{\mathcal{X}/\mathcal{S}}-module. If \mathscr{M} has no torsion subsheaf supported on 𝒳\mathcal{X}_{\mathcal{H}} and if the cycle CCrel()\textup{CC}_{\textup{rel}}(\mathscr{M}) does not have components over 𝒳\mathcal{X}_{\mathcal{H}}, then

CCrel(i)=CCrel()|𝒳,\textup{CC}_{\textup{rel}}(i^{*}_{\mathcal{H}}\mathscr{M})=\textup{CC}_{\textup{rel}}(\mathscr{M})|_{\mathcal{X}_{\mathcal{H}}},

where i:𝒳𝒳i_{\mathcal{H}}:\mathcal{X}_{\mathcal{H}}\hookrightarrow\mathcal{X} is the closed embedding with the fiber product 𝒳=𝒳×𝒮\mathcal{X}_{\mathcal{H}}=\mathcal{X}\times_{\mathcal{S}}\mathcal{H}.

Proof.

Since characteristic cycles are local, it is enough to assume \mathcal{H} defined by a regular (or holomorphic) function hh. The torsion-free assumption implies that we have a short exact sequence

0hi()0.0\to\mathscr{M}\xrightarrow{\cdot h}\mathscr{M}\rightarrow i^{*}_{\mathcal{H}}(\mathscr{M})\to 0.

Now we pick a good filtration FF_{\bullet}\mathscr{M} over 𝒟𝒳/𝒮\mathscr{D}_{\mathcal{X}/\mathcal{S}} and the filtration induces a filtered complex

FhF.F_{\bullet}\mathscr{M}\xrightarrow{\cdot h}F_{\bullet}\mathscr{M}.

We then obtain a convergent spectral sequence with the E0E^{0}-page given by the following length 2 complex

η:grFhgrF.\eta\colon\textup{gr}^{F}_{\bullet}\mathscr{M}\xrightarrow{\cdot h}\textup{gr}^{F}_{\bullet}\mathscr{M}.

Then, i()i^{*}_{\mathcal{H}}(\mathscr{M}) has an induced filtration F(i())F_{\bullet}(i^{*}_{\mathcal{H}}(\mathscr{M})) (good over 𝒟𝒳/\mathscr{D}_{\mathcal{X}_{\mathcal{H}}/\mathcal{H}}). By convergence of the spectral sequence (see for instance [Lau83, Lemme 3.5.13] and also [Sab87b, 3.7. Lemme]), we have

[grF(i())]=[Coker(η)][Ker(η)][\textup{gr}^{F}_{\bullet}(i^{*}_{\mathcal{H}}(\mathscr{M}))]=[\textup{Coker}(\eta)]-[\textup{Ker}(\eta)]

in the Grothendieck group K0K_{0}. Since the characteristic cycle is well defined for [grF(i())][\textup{gr}^{F}_{\bullet}(i^{*}_{\mathcal{H}}(\mathscr{M}))], the required statement for characteristic cycles follows from [Bj93, Appendix IV. 3.13 Proposition]. ∎

Suppose that 𝒩\mathscr{N} is a (left) 𝒟𝒳𝒯/𝒯\mathscr{D}_{\mathcal{X}_{\mathcal{T}}/\mathcal{T}}-module (or more generally a complex). We consider the derived pushforward functor 𝐑μ(𝒩)\mathbf{R}\mu_{*}(\mathscr{N}) with μ\mu as in Diag.(2). Since μ𝒟𝒳/𝒮=𝒟𝒳𝒯/𝒯\mu^{*}{\mathscr{D}_{\mathcal{X}/\mathcal{S}}}=\mathscr{D}_{\mathcal{X}_{\mathcal{T}}/\mathcal{T}}, we have

𝐑μ(𝒩)𝐑μ(μ(𝒟𝒳/𝒮)𝒟𝒳𝒯/𝒯𝒩),\mathbf{R}\mu_{*}(\mathscr{N})\simeq\mathbf{R}\mu_{*}(\mu^{*}(\mathscr{D}_{\mathcal{X}/\mathcal{S}})\otimes_{\mathscr{D}_{\mathcal{X}_{\mathcal{T}}/\mathcal{T}}}\mathscr{N}),

and hence the complex 𝐑μ(𝒩)\mathbf{R}\mu_{*}(\mathscr{N}) is a complex of 𝒟𝒳/𝒮\mathscr{D}_{\mathcal{X}/\mathcal{S}}-modules by adjunction.

Proposition 2.8.

Suppose that ν\nu is a proper morphism, 𝒩\mathscr{N} is coherent over 𝒟𝒳𝒯/𝒯\mathscr{D}_{\mathcal{X}_{\mathcal{T}}/\mathcal{T}} ((or more generally a complex of 𝒟𝒳𝒯/𝒯\mathscr{D}_{\mathcal{X}_{\mathcal{T}}/\mathcal{T}}-modules with coherent cohomology sheaves)) and 𝒩\mathcal{N} ((or each of its cohomology sheaves)) admits a good filtration over F𝒟𝒳/𝒮F_{\bullet}\mathscr{D}_{\mathcal{X}/\mathcal{S}}. Then 𝐑iμ(𝒩)\mathbf{R}^{i}\mu_{*}(\mathscr{N}) is coherent over 𝒟𝒳/𝒮\mathscr{D}_{\mathcal{X}/\mathcal{S}} for each ii\in\mathbb{Z}.

Proof.

The idea of the proof of the required coherence result is similar to that of the case for absolute 𝒟\mathscr{D}-modules (see for instance [Bj93, Theorem 2.8.1]). We sketch here its proof for completeness.

By a standard procedure (see for instance the proof of [Bj93, Theorem 1.5.8]), the required statement can be reduced to the case for

𝒩=𝒟𝒳𝒯/𝒯𝒪\mathscr{N}=\mathscr{D}_{\mathcal{X}_{\mathcal{T}}/\mathcal{T}}\otimes_{\mathscr{O}}\mathscr{L}

where \mathscr{L} is a coherent 𝒪\mathscr{O}-module. But this case follows immediately from the Grauert’s direct image theorem for 𝒪\mathscr{O}-modules and the projection formula (since μ𝒟𝒳/𝒮=𝒟𝒳𝒯/𝒯\mu^{*}{\mathscr{D}_{\mathcal{X}/\mathcal{S}}}=\mathscr{D}_{\mathcal{X}_{\mathcal{T}}/\mathcal{T}}). ∎

It is worth mentioning that 𝐑μ\mathbf{R}\mu_{*} does not preserve relative holonomicity under proper base changes in general (compared to Proposition 2.10). See §4.10 for related examples.

2.3. Relative de Rham complexes

We keep notations as in the previous subsection. Let \mathscr{M} be a (left) 𝒟𝒳/𝒮\mathscr{D}_{\mathcal{X}/\mathcal{S}}-module. The relative de Rham complex of \mathscr{M} is defined as

DR𝒳/𝒮()ω𝒳/𝒮𝒟𝒳/𝒮𝐋.\textup{DR}_{\mathcal{X}/\mathcal{S}}(\mathscr{M})\coloneqq\omega_{\mathcal{X}/\mathcal{S}}\otimes^{\mathbf{L}}_{\mathscr{D}_{\mathcal{X}/\mathcal{S}}}\mathscr{M}.
Lemma 2.9.

We have a natural isomorphism

𝐋isDR𝒳/𝒮DR𝒳s𝐋is\mathbf{L}i^{*}_{s}\circ\textup{DR}_{\mathcal{X}/\mathcal{S}}\simeq\textup{DR}_{\mathcal{X}_{s}}\circ\mathbf{L}i^{*}_{s}

for s𝒮s\in\mathcal{S}.

Proof.

The functor 𝐋is\mathbf{L}i^{*}_{s} can be rewritten as 𝒪𝒮𝐋s\otimes^{\mathbf{L}}_{\mathscr{O}_{\mathcal{S}}}\mathbb{C}_{s}, where s\mathbb{C}_{s} is the residue field of s𝒮s\in\mathcal{S}. Since sections of 𝒪𝒮\mathscr{O}_{\mathcal{S}} contains in the center of 𝒟𝒳/𝒮\mathscr{D}_{\mathcal{X}/\mathcal{S}}, the required statement follows. ∎

2.4. Relative direct images

We discuss direct image functors for 𝒟\mathscr{D}-modules under the relative setting. We fix a morphism ff over 𝒮\mathcal{S}

𝒴{\mathcal{Y}}𝒳{\mathcal{X}}𝒮{\mathcal{S}}f\scriptstyle{f}φ1\scriptstyle{\varphi_{1}}φ\scriptstyle{\varphi}

with φ1\varphi_{1} and φ\varphi smooth. We recall the definition of the relative direct image:

f+(𝒩)𝐑f(𝒪ωf/𝒮𝒟f𝒟𝒳/𝒮)f_{+}(\mathscr{N})\coloneqq\mathbf{R}f_{*}(\mathscr{M}\otimes_{\mathscr{O}}\omega_{f/\mathcal{S}}\otimes_{\mathscr{D}}f^{*}\mathscr{D}_{\mathcal{X}/\mathcal{S}})

for a left 𝒟𝒴/𝒮\mathscr{D}_{\mathcal{Y}/\mathcal{S}}-module 𝒩\mathscr{N} (or more generally a complex of left 𝒟𝒴/𝒮\mathscr{D}_{\mathcal{Y}/\mathcal{S}}-modules), where ωf/𝒮=ω𝒴/𝒮f(ω𝒳/𝒮1)\omega_{f/\mathcal{S}}=\omega_{\mathcal{Y}/\mathcal{S}}\otimes f^{*}(\omega^{-1}_{\mathcal{X}/\mathcal{S}}).

The morphism ff over 𝒮\mathcal{S} induces a relative Lagrangian correspondence

T(𝒴/𝒮){T^{*}(\mathcal{Y}/\mathcal{S})}𝒴×𝒳T(𝒳/𝒮){\mathcal{Y}\times_{\mathcal{X}}T^{*}(\mathcal{X}/\mathcal{S})}T(𝒳/𝒮){T^{*}(\mathcal{X}/\mathcal{S})}𝒮{\mathcal{S}}ϕ1\scriptstyle{\phi_{1}}ϱf\scriptstyle{\varrho_{f}}ϖf\scriptstyle{\varpi_{f}}ϕ\scriptstyle{\phi}

See for instance [HTT08, §2.4] for the absolute Lagrangian correspondence.

The following proposition is a generalization of [MFS16, Theorem 1.17(a)].

Proposition 2.10.

With above notations, let 𝒩\mathscr{N} be a coherent 𝒟𝒴/𝒮\mathscr{D}_{\mathcal{Y}/\mathcal{S}}-module with a good filtration over F𝒟𝒴/𝒮F_{\bullet}\mathscr{D}_{\mathcal{Y}/\mathcal{S}}.

  1. (1)

    If ff is proper over 𝒮\mathcal{S}, then f+(𝒩)f_{+}(\mathscr{N}) is a complex of 𝒟𝒳/𝒮\mathscr{D}_{\mathcal{X}/\mathcal{S}}-modules with coherent cohomology sheaves.

  2. (2)

    If moreover 𝒩\mathscr{N} is relative holonomic, then f+(𝒩)f_{+}(\mathscr{N}) is a complex of 𝒟𝒳/𝒮\mathscr{D}_{\mathcal{X}/\mathcal{S}}-modules with relative holonomic cohomology sheaves.

Proof.

Since (𝒩,𝒴)(\mathscr{N},\mathbb{C}_{\mathcal{Y}}) gives us a good relative elliptic pair (see [SS94, Definition 2.14]), the first statement follows from Theorem 4.2 in loc. cit. If moreover 𝒩\mathscr{N} is relative holonomic, then by Corollary 4.3 in loc. cit. we have

Chrel(if+(𝒩))ϖf(ϱf1(Chrel(𝒩))\textup{Ch}_{\textup{rel}}(\mathcal{H}^{i}f_{+}(\mathscr{N}))\subseteq\varpi_{f}(\varrho_{f}^{-1}(\textup{Ch}_{\textup{rel}}(\mathscr{N}))

for each ii. Thus, for s𝒮s\in\mathcal{S}

Chrel(if+(𝒩))ϕ1(s)ϖf(ϱf1(Chrel(𝒩))ϕ1(s)ϖf(ϱf1(Chrel(𝒩)ϕ11(s))).\textup{Ch}_{\textup{rel}}(\mathcal{H}^{i}f_{+}(\mathscr{N}))\cap\phi^{-1}(s)\subseteq\varpi_{f}(\varrho_{f}^{-1}(\textup{Ch}_{\textup{rel}}(\mathscr{N}))\cap\phi^{-1}(s)\subseteq\varpi_{f}(\varrho_{f}^{-1}(\textup{Ch}_{\textup{rel}}(\mathscr{N})\cap\phi_{1}^{-1}(s))).

By definition Chrel(𝒩)ϕ11(s)\textup{Ch}_{\textup{rel}}(\mathscr{N})\cap\phi_{1}^{-1}(s) is Lagrangian and hence isotropic. By [Kas77, (4.9)], ϖf(ϱf1(Chrel(𝒩)ϕ11(s))\varpi_{f}(\varrho_{f}^{-1}(\textup{Ch}_{\textup{rel}}(\mathscr{N})\cap\phi_{1}^{-1}(s)) and hence Chrel(if+(𝒩))ϕ1(s)\textup{Ch}_{\textup{rel}}(\mathcal{H}^{i}f_{+}(\mathscr{N}))\cap\phi^{-1}(s) are both isotropic. Thus, Chrel(if+(𝒩))ϕ1(s)\textup{Ch}_{\textup{rel}}(\mathcal{H}^{i}f_{+}(\mathscr{N}))\cap\phi^{-1}(s) is Lagrangian by Theorem 2.2 and if+(𝒩)\mathcal{H}^{i}f_{+}(\mathscr{N}) is relative holonomic. ∎

The following lemma is immediate by construction, and we skip its proof.

Lemma 2.11.

We have a natural isomorphism

𝐋isf+(fs)+𝐋is,\mathbf{L}i^{*}_{s}\circ f_{+}\simeq(f_{s})_{+}\circ\mathbf{L}i^{*}_{s},

where fs:𝒴s𝒳sf_{s}:\mathcal{Y}_{s}\to\mathcal{X}_{s} is the induced morphism over s𝒮s\in\mathcal{S}.

Corollary 2.12.

If ff is (relative) proper over 𝒮\mathcal{S}, then f+f_{+} preserves relative regular holonomicity.

Proof.

If ff is proper over 𝒮\mathcal{S}, then fsf_{s} is proper for s𝒮s\in\mathcal{S}. Since (fs)+(f_{s})_{+} preserves regular holonomicity, the required statement follows from Lemma 2.11. ∎

3. Logarithmic 𝒟\mathscr{D}-modules

In this section, we recall 𝒟\mathscr{D}-modules under the logarithmic setting. Let XX be a complex manifold of dimension nn and let DD be a normal crossing divisor. We call such (X,D)(X,D) a (analytic) smooth log pair. We write by 𝒟X,D\mathscr{D}_{X,D} the subalgebra of 𝒟X\mathscr{D}_{X} consisting of differential operators preserving the ideal sheaf of DD. In local coordinates (x1,x2,,xn)(x_{1},x_{2},\dots,x_{n}) on an open neighborhood UU with D|U=(i=1rxi=0)D|_{U}=(\prod_{i=1}^{r}x_{i}=0) for some rnr\leq n, 𝒟X,D\mathscr{D}_{X,D} is the subalgebra generated by 𝒪U\mathscr{O}_{U} and

x1x1,,xkxr,xr+1,,xn.x_{1}\partial_{x_{1}},\dots,x_{k}\partial_{x_{r}},\partial_{x_{r+1}},\dots,\partial_{x_{n}}.

Since DD is normal crossing, 𝒟X,D\mathscr{D}_{X,D} is a coherent and noetherian sheaf of rings. Modules over 𝒟X,D\mathscr{D}_{X,D} are called logarithmic (log) 𝒟\mathscr{D}-modules.

The order filtration F𝒟XF_{\bullet}\mathscr{D}_{X} induces the order filtration F𝒟X,DF_{\bullet}\mathscr{D}_{X,D} such that the analytification of grF𝒟X,D\textup{gr}^{F}_{\bullet}\mathscr{D}_{X,D} gives us the structure sheaf of the log cotangent bundle T(X,D)T^{*}(X,D). For a coherent 𝒟X,D\mathscr{D}_{X,D}-module \mathcal{M}, one can define the log characteristic variety

Chlog()T(X,D),\textup{Ch}_{\log}(\mathcal{M})\subseteq T^{*}(X,D),

and the log characteristic cycle CClog()\textup{CC}_{\log}(\mathcal{M}) similar to the relative case. When D=D=\emptyset, we use Ch()\textup{Ch}(\mathcal{M}) (resp. CC()\textup{CC}(\mathcal{M})) to denote the characteristic variety (resp. cycle) of the 𝒟X\mathscr{D}_{X}-module \mathcal{M}.

3.1. Log de Rham complexes

Suppose that \mathcal{M} is a left 𝒟X,D\mathscr{D}_{X,D}-module on a smooth log pair (X,D)(X,D). Similar to the absolute case, ωX(D)\omega_{X}(D) is a right 𝒟X,D\mathscr{D}_{X,D}-module, where ωX\omega_{X} is the sheaf of the top forms on XX. The log de Rham complex of \mathcal{M} is defined as

DRX,D()ωX(D)𝒟X,D𝐋.\textup{DR}_{X,D}(\mathcal{M})\coloneqq\omega_{X}(D)\otimes^{\mathbf{L}}_{\mathscr{D}_{X,D}}\mathcal{M}.

By [WZ21, Lemma 2.3], we have

DRX,D()[Ω1(logD)𝒪Ωn(logD)𝒪]\textup{DR}_{X,D}(\mathcal{M})\simeq[\mathcal{M}\rightarrow\Omega^{1}(\log D)\otimes_{\mathscr{O}}\mathcal{M}\rightarrow\cdots\rightarrow\Omega^{n}(\log D)\otimes_{\mathscr{O}}\mathcal{M}]

where the complex on the right-hand side starts from the degree n-n-term and Ωi(logD)\Omega^{i}(\log D) denotes the sheaf of the degree ii log forms. In local coordinates

(x1,x2,,xn) with D|U=(i=1rxi=0)(x_{1},x_{2},\dots,x_{n})\textup{ with }D|_{U}=(\prod_{i=1}^{r}x_{i}=0)

on an open neighborhood UU for some rnr\leq n, we further have

(3) DRX,D()Kos(;x1x1,,xrxr,xr+1,,xn),\textup{DR}_{X,D}(\mathcal{M})\simeq\textup{Kos}(\mathcal{M};x_{1}\partial_{x_{1}},\dots,x_{r}\partial_{x_{r}},\partial_{x_{r+1}},\dots,\partial_{x_{n}}),

where Kos denotes the Koszul complex of the actions x1x1,,xrxr,xr+1,,xnx_{1}\partial_{x_{1}},\dots,x_{r}\partial_{x_{r}},\partial_{x_{r+1}},\dots,\partial_{x_{n}} on \mathcal{M}.

3.2. Direct image functor

Suppose that f:(X,D)(Y,E)f\colon(X,D)\rightarrow(Y,E) is a morphism of smooth log pairs, that is, f:XYf\colon X\rightarrow Y is a morphism of complex manifolds such that f1EDf^{-1}E\subseteq D. Then we define the derived direct image functor for left log 𝒟\mathscr{D}-modules \mathcal{M} by

f+log()=Rf(𝒪ωf𝒟X,D𝐋f𝒟Y,E),f^{\log}_{+}(\mathcal{M})=Rf_{*}(\mathcal{M}\otimes_{\mathscr{O}}\omega_{f}\otimes^{\mathbf{L}}_{\mathscr{D}_{X,D}}f^{*}\mathscr{D}_{Y,E}),

where ωflog\omega^{\log}_{f} is the canonical sheaf of ff,

ωflog=ωX(D)𝒪f(ωY1(E)).\omega^{\log}_{f}=\omega_{X}(D)\otimes_{\mathscr{O}}f^{*}(\omega^{-1}_{Y}(E)).

The above definition is compatible with the direct images in the relative case in §2.4 if one take DD and EE empty.

Proposition 3.1.

Let f:(X,D)(Y,E)f\colon(X,D)\rightarrow(Y,E) be a morphism of smooth log pairs and let \mathcal{M} be a ((left)) 𝒟X,D\mathscr{D}_{X,D}-modules. Then

RfDRX,D()DRY,E(f+log).Rf_{*}\textup{DR}_{X,D}(\mathcal{M})\simeq\textup{DR}_{Y,E}(f^{\log}_{+}\mathcal{M}).
Proof.

The proof of this proposition is exactly the same as the non-log case (cf. [HTT08, Theorem 4.2.5]). We leave the detail for interested readers. ∎

3.3. Lattices

We recall the definition of lattices in the analytic setting. Suppose that (X,D)(X,D) is an analytic smooth log pair and \mathcal{M} is a coherent 𝒟X\mathscr{D}_{X}-module. We write the algebraic localization of \mathcal{M} along DD by

(D)=𝒪𝒪X(D),\mathcal{M}(*D)=\mathcal{M}\otimes_{\mathscr{O}}\mathscr{O}_{X}(*D),

where

𝒪X(D)=limk𝒪X(kD).\mathscr{O}_{X}(*D)=\lim_{k\to\infty}\mathscr{O}_{X}(kD).

Notice that in general (D)\mathcal{M}(*D) is not even coherent over 𝒟X\mathscr{D}_{X}. However, if \mathcal{M} is holonomic, then so is (D)\mathcal{M}(*D), since holonomicity is preserved under tenser products over 𝒪\mathscr{O}. A coherent 𝒟X,D\mathscr{D}_{X,D}-submodule

¯(D)\bar{\mathcal{M}}\subseteq\mathcal{M}(*D)

is called a 𝒟X,D\mathscr{D}_{X,D}-lattice of (D)\mathcal{M}(*D) (or \mathcal{M}) if ¯|XD=|XD\bar{\mathcal{M}}|_{X\setminus D}=\mathcal{M}|_{X\setminus D}. By definition, lattices of \mathcal{M} do not have torsion subsheaves supported on DD (although, \mathcal{M} might have). The prototype examples of lattices are Deligne lattices of local systems (see [WZ21, §4.4]).

3.4. From log to relative

In this subsection, we discuss the connection between log 𝒟\mathscr{D}-modules and relative 𝒟\mathscr{D}-modules. We focus locally on WW a polydisc,

W=𝔻xk×𝔻tr with coordinates (x1,,xk,t1,,tr)W=\mathbb{D}^{k}_{x}\times\mathbb{D}_{t}^{r}\textup{ with coordinates }(x_{1},\dots,x_{k},t_{1},\dots,t_{r})

and a divisor DD given by t1t2tr=0t_{1}\cdot t_{2}\cdots t_{r}=0. In particular, (W,D)(W,D) is a smooth log pair. We consider the log rescaled families

W~j=W×𝔻yj\widetilde{W}_{j}=W\times\mathbb{D}_{y}^{j}

with y(j)=(y1,,yj)y(j)=(y_{1},\dots,y_{j}) the coordinates on 𝔻yj\mathbb{D}_{y}^{j} for 0jr0\leq j\leq r (W~0=W\widetilde{W}_{0}=W), and the maps

pj:W~jW,(x,t,y(j))(x,ey(j)t),p_{j}\colon\widetilde{W}_{j}\rightarrow W,(x,t,y(j))\mapsto(x,e^{y(j)}t),

where we abbreviate (x1,,xr)(x_{1},\dots,x_{r}) as xx, (ey1t1,,eyjtj,tj+1,,tr)(e^{y_{1}}t_{1},\dots,e^{y_{j}}t_{j},t_{j+1},\dots,t_{r}) as ey(j)xe^{y(j)}x and etc. Then we have the commutative diagram for each 0j<k0\leq j<k

(4) W~j{\widetilde{W}_{j}}W~j+1{\widetilde{W}_{j+1}}W{W}ij\scriptstyle{i_{j}}pj\scriptstyle{p_{j}}pj+1\scriptstyle{p_{j+1}}

where the inclusion is

ij:W~jW~j+1,(x,t,y(j))(x,t,y(j),0).i_{j}:\widetilde{W}_{j}\hookrightarrow\widetilde{W}_{j+1},(x,t,y(j))\mapsto(x,t,y(j),0).

Let U=𝔻xk×(𝔻)trWU=\mathbb{D}^{k}_{x}\times(\mathbb{D}^{\star})^{r}_{t}\subset W with 𝔻\mathbb{D}^{\star} the punctured disk. Then we define U~j=p1(U)\widetilde{U}_{j}=p^{-1}(U), D~j=pj1(D)\widetilde{D}_{j}=p^{-1}_{j}(D). Given a 𝒟W,D\mathscr{D}_{W,D}-lattice ¯\bar{\mathcal{M}} of a 𝒟X\mathscr{D}_{X}-module \mathcal{M}, we consider the pull-backs jpj¯\mathscr{M}_{j}\coloneqq p^{*}_{j}\bar{\mathcal{M}} and write pk=pp_{k}=p, =k\mathscr{M}=\mathscr{M}_{k}, W~=W~k\widetilde{W}=\widetilde{W}_{k}, U~k=U~\widetilde{U}_{k}=\widetilde{U} and D~=D~k\widetilde{D}=\widetilde{D}_{k}.

Lemma 3.2.

With notations above, let \mathcal{M} be a coherent 𝒟X\mathscr{D}_{X}-module. Then

  1. (1)

    pjp_{j} is smooth ((submersive)) for each jj,

  2. (2)

    j\mathscr{M}_{j} is a coherent 𝒟W~j,D~j\mathscr{D}_{\widetilde{W}_{j},\widetilde{D}_{j}}-module for each jj,

  3. (3)

    \mathscr{M} is a coherent 𝒟W~/𝔻tr\mathscr{D}_{\widetilde{W}/\mathbb{D}^{r}_{t}}-module for the natural projection

    πt:W~𝔻tr.\pi_{t}:\widetilde{W}\to\mathbb{D}^{r}_{t}.
Proof.

Part (1) is obvious. For Part (2), we pick a good filtration F¯F_{\bullet}\bar{\mathcal{M}} over F𝒟X,DF_{\bullet}\mathscr{D}_{X,D}. Since pjp_{j} is smooth, pj(F¯)p_{j}^{*}(F_{\bullet}\bar{\mathcal{M}}) is a filtration of j\mathscr{M}_{j} over F𝒟W~j,DjF_{\bullet}\mathscr{D}_{\widetilde{W}_{j},D_{j}}. By construction, we have

titi(1pj1(m))=1pj1(titim)t_{i}\partial_{t_{i}}\cdot(1\otimes p_{j}^{-1}(m))=1\otimes p_{j}^{-1}(t_{i}\partial_{t_{i}}\cdot m)

for sections mm of ¯\bar{\mathcal{M}}. Therefore, pj(grF)p^{*}_{j}(\textup{gr}^{F}_{\bullet}\mathcal{M}) is coherent over grF𝒟W~j,D~j\textup{gr}^{F}_{\bullet}\mathscr{D}_{\widetilde{W}_{j},\widetilde{D}_{j}} and hence pj(F¯)p_{j}^{*}(F_{\bullet}\bar{\mathcal{M}}) is a good filtration of j\mathscr{M}_{j} over F𝒟W~j,DjF_{\bullet}\mathscr{D}_{\widetilde{W}_{j},D_{j}}. In particular, j\mathscr{M}_{j} is coherent over 𝒟W~j,D~j\mathscr{D}_{\widetilde{W}_{j},\widetilde{D}_{j}}. For Part (3), one observes that titiyit_{i}\partial_{t_{i}}-\partial_{y_{i}} annihilates 1pj1(m)1\otimes p_{j}^{-1}(m). Thus, p(grF)p^{*}(\textup{gr}^{F}_{\bullet}\mathcal{M}) is coherent over grF𝒟W~/𝔻tr\textup{gr}^{F}_{\bullet}\mathscr{D}_{\widetilde{W}/\mathbb{D}^{r}_{t}}. In consequence, \mathscr{M} is coherent over 𝒟W~/𝔻tr\mathscr{D}_{\widetilde{W}/\mathbb{D}^{r}_{t}}. ∎

Theorem 3.3.

Let \mathcal{M} be a regular holonomic 𝒟W\mathscr{D}_{W}-module and let ¯\bar{\mathcal{M}} be a 𝒟W,D\mathscr{D}_{W,D}-lattice of \mathcal{M}. Then we have:

  1. (1)

    If r=1r=1, then \mathscr{M} is a relative regular holonomic 𝒟W~/𝔻xr\mathscr{D}_{\widetilde{W}/\mathbb{D}^{r}_{x}}-module.

  2. (2)

    If \mathscr{M} is flat over 𝔻xr\mathbb{D}^{r}_{x} for some r>1r>1, then \mathscr{M} is relative holonomic over 𝔻xr\mathbb{D}^{r}_{x}.

If r2r\geq 2, then \mathscr{M} is not necessarily relative holonomic. See Example 2 in §5.

3.5. Proof of Theorem 1.6

Our first goal is to prove that, the characteristic variety Chlog(¯)\textup{Ch}_{\log}(\bar{\mathcal{M}}) is the closure of Ch(|U)\textup{Ch}(\mathcal{M}|_{U}) in the log cotangent bundle TlogWT_{\log}^{*}W, which is equivalent to prove that Chlog(¯)\textup{Ch}_{\log}(\bar{\mathcal{M}}) has no irreducible components over DD. Then the statement of characteristic cycles follows immediately, since multiplicities are generically defined.

Our main tool is a technical result of Sabbah [Sab87b] (see also [BMM02]). Consider a smooth submersion φ:𝒳𝒮\varphi:\mathcal{X}\to\mathcal{S}. Let 𝒩\mathcal{N} be a regular hononomic 𝒟𝒳\mathscr{D}_{\mathcal{X}}-module, 𝒩rel\mathcal{N}_{{\textup{rel}}} be a coherent 𝒟𝒳/𝒮\mathscr{D}_{\mathcal{X}/\mathcal{S}}-submodule of 𝒩\mathcal{N}, that generates 𝒩\mathcal{N} as D𝒳D_{\mathcal{X}}-module. Suppose Ch(𝒩)=YLTY𝒳\textup{Ch}(\mathcal{N})=\bigcup_{Y\in L}T_{Y}^{*}\mathcal{X}, with a set LL (locally finite) consisting of irreducible subvarieties of 𝒳\mathcal{X}. Let L1LL_{1}\subset L be a subset consisting of YY such that φ(Y)\varphi(Y) contains a non-empty open subset of 𝒮\mathcal{S}. Denote by YsmY^{\textup{sm}} the smooth locus of YY. By generic smoothness, we can assume YsmY^{\textup{sm}} is smooth over 𝒮\mathcal{S} (shrink YsmY^{\textup{sm}} if necessary). Then we obtain the relative conormal bundle of Ysm𝒳Y^{\textup{sm}}\hookrightarrow\mathcal{X} over certain open subset of 𝒮\mathcal{S}. We then write by Tφ|Y(𝒳/𝒮)T^{*}_{\varphi|_{Y}}(\mathcal{X}/\mathcal{S}) the closure of the (generically defined) relative conormal bundle, calling it the relative conormal space of φ|Y\varphi|_{Y}. Notice that relative conormal spaces are not necessarily relative Lagrangian.

Lemma 3.4.

[Sab87b, 3.2.Théorèm], [BMM02, Lemme 2.2] With notations as above, suppose there is a non-constant holomorphic function F:𝒳F:\mathcal{X}\to\mathbb{C}, such that L\LL\backslash L^{\prime} are contained in F1(0)F^{-1}(0), and 𝒩\mathcal{N} are without FF-torsion. Then, we have

Chrel(𝒩rel)=YL1Tφ|Y(𝒳/𝒮),\textup{Ch}_{{\textup{rel}}}(\mathcal{N}_{{\textup{rel}}})=\bigcup_{Y\in L_{1}}T^{*}_{\varphi|_{Y}}(\mathcal{X}/\mathcal{S}),

where Tφ|Y(𝒳/𝒮)T^{*}_{\varphi|_{Y}}(\mathcal{X}/\mathcal{S}) is the relative conormal space of φ|Y\varphi|_{Y}.

Since characteristic cycles are local, it is enough to assume X=WX=W, a polydisc as in §3.4. To apply Lemma 3.4, we let 𝒩=p((D))\mathcal{N}=p^{*}(\mathcal{M}(*D)), 𝒩rel=\mathcal{N}_{{\textup{rel}}}=\mathscr{M}, 𝒳=W~\mathcal{X}=\widetilde{W}, 𝒮=𝔻tr\mathcal{S}=\mathbb{D}^{r}_{t}, φ=πt\varphi=\pi_{t}, and F=t1trF=t_{1}\cdots t_{r}. Suppose we have decomposition

Ch(|U)=YL1TYW\textup{Ch}(\mathcal{M}|_{U})=\bigcup_{Y\in L_{1}}T^{*}_{Y}W

for a set of closed strata L1L_{1} in UU. Then we may define a set of closed strata L~1\widetilde{L}_{1} in W~\widetilde{W}, by sending YL1Y\in L_{1} to Y~=p1Y¯W~\widetilde{Y}=\overline{p^{-1}Y}\subset\widetilde{W}. Since pp is submersive, we have

Ch(p~(D))|U~=Y~L~1TY~W~|U~.\textup{Ch}(p^{*}\widetilde{\mathcal{M}}(*D))|_{\widetilde{U}}=\bigcup_{\widetilde{Y}\in\widetilde{L}_{1}}T^{*}_{\widetilde{Y}}\widetilde{W}|_{\widetilde{U}}.
Corollary 3.5.
Chrel()=Y~L~1Tπt|Y~(W~/𝔻xk).\textup{Ch}_{{\textup{rel}}}(\mathscr{M})=\bigcup_{\widetilde{Y}\in\widetilde{L}_{1}}T^{*}_{\pi_{t}|_{\widetilde{Y}}}(\widetilde{W}/\mathbb{D}^{k}_{x}).
Proof.

By Lemma 3.4, we only need to look at strata of Ch(p(D))\textup{Ch}(p^{*}\mathcal{M}(*D)) that project under πx\pi_{x} to an open set in 𝔻xk\mathbb{D}_{x}^{k}, hence it suffices to consider the strata that intersect with πx1((𝔻x)k)=U~\pi_{x}^{-1}((\mathbb{D}^{*}_{x})^{k})=\widetilde{U} (by the construction of pp). These are labeled exactly by L~1\widetilde{L}_{1}. ∎

Proof of Theorem 3.3.

By Corollary 3.5, Chrel()\textup{Ch}_{\textup{rel}}(\mathscr{M}) does not have components over t1tr=0t_{1}\cdots t_{r}=0. For every point α𝔻tr\alpha\in\mathbb{D}^{r}_{t}, we pick rr general hyperplanes passing α\alpha. By flatness assumption, the relative holonomicity in Part (2) follows from inductively applying Proposition 2.7. Since 𝒩=p((D))\mathscr{M}\subseteq\mathcal{N}=p^{*}(\mathcal{M}(*D)), \mathscr{M} is torsion-free over 𝔻tr\mathbb{D}^{r}_{t}. When r=1r=1, torsion-freeness implies flatness and hence the relative holonomicity of Part (1) follows. For regularity in Part (1), one observes that

iα()pα(|U)i_{\alpha}^{*}(\mathscr{M})\simeq p_{\alpha}^{*}(\mathcal{M}|_{U})

are regular holonomic for 0α𝔻t\textup{for }0\not=\alpha\in\mathbb{D}_{t}, where the morphism pαp_{\alpha} is

pα:𝔻xk×{α}×𝔻y𝔻xk×𝔻t(x,α,y)(x,αey).p_{\alpha}\colon\mathbb{D}_{x}^{k}\times\{\alpha\}\times\mathbb{D}_{y}\to\mathbb{D}_{x}^{k}\times\mathbb{D}_{t}\quad(x,\alpha,y)\mapsto(x,\alpha e^{y}).

When α=0\alpha=0,

i0()p0(|(t=0)).i_{0}^{*}(\mathscr{M})\simeq p_{0}^{*}(\mathcal{M}|_{(t=0)}).

But

[|(t=0)]=[Φt=0()][\mathcal{M}|_{(t=0)}]=[\Phi_{t=0}(\mathcal{M})]

in the Grothendieck group K0K_{0} (by [Gin86, Proposition 1.1.2]). Since regularity is well-defined for objects in K0K_{0}, |(t=0)\mathcal{M}|_{(t=0)} is regular holonomic by Theorem 4.5(1). The regularity in Part (2) is similar by induction. ∎

We use TlogT^{*}_{\log} and TrelT^{*}_{\textup{rel}} to denote the log/relative cotangent bundle.

Proposition 3.6.

We have:

  1. (1)

    Chlog()=ι(Chrel())\textup{Ch}_{\log}(\mathscr{M})=\iota(\textup{Ch}_{{\textup{rel}}}(\mathscr{M})) where

    TlogW~=T𝔻xk×Tlog𝔻tr×T𝔻yr,TrelW~=T𝔻xk×𝔻tr×T𝔻yrT^{*}_{\log}\widetilde{W}=T^{*}\mathbb{D}_{x}^{k}\times T^{*}_{\log}\mathbb{D}^{r}_{t}\times T^{*}\mathbb{D}^{r}_{y},\quad T^{*}_{{\textup{rel}}}\widetilde{W}=T^{*}\mathbb{D}_{x}^{k}\times\mathbb{D}^{r}_{t}\times T^{*}\mathbb{D}^{r}_{y}
    ι:TrelW~TlogW~,(x,t,y,ξx,ξy)(x,t,y,ξx,ξ~t=ξy,ξy),\iota:T^{*}_{{\textup{rel}}}\widetilde{W}\to T^{*}_{\log}\widetilde{W},\quad(x,t,y,\xi_{x},\xi_{y})\mapsto(x,t,y,\xi_{x},\widetilde{\xi}_{t}=\xi_{y},\xi_{y}),

    and ξ~t\widetilde{\xi}_{t} is the coefficient in front of dt/tdt/t.

  2. (2)

    ι~(Chlog(¯))=Chlog()|{y=0}\tilde{\iota}(\textup{Ch}_{\log}(\bar{\mathcal{M}}))=\textup{Ch}_{\log}(\mathscr{M})|_{\{y=0\}}, where ι~:TlogWTlogW~\tilde{\iota}\colon T^{*}_{\log}W\hookrightarrow T^{*}_{\log}\widetilde{W} is given by

    (x,t,ξx,ξ~t)(x,t,ξx,ξ~t,ξy=ξ~t).(x,t,\xi_{x},\widetilde{\xi}_{t})\mapsto(x,t,\xi_{x},\widetilde{\xi}_{t},\xi_{y}=\widetilde{\xi}_{t}).
Proof.

For Part(1), since for any section ss in p1(¯)p^{-1}(\bar{\mathcal{M}}), titiyit_{i}\partial_{t_{i}}-\partial_{y_{i}} annihilate ss, hence on the level of the associated graded modules, we have ξ~tiξyi\widetilde{\xi}_{t_{i}}-\xi_{y_{i}} annihilate grF()\textup{gr}^{F}_{\bullet}(\mathscr{M}).

Now we prove Part(2). By Lemma 3.2, we have

grF(ir())ir(grF())pr1(grF¯).\textup{gr}^{F}_{\bullet}(i_{r}^{*}(\mathscr{M}))\simeq i_{r}^{*}(\textup{gr}^{F}_{\bullet}(\mathscr{M}))\simeq p_{r-1}^{*}(\textup{gr}^{F}_{\bullet}\bar{\mathcal{M}}).

Meanwhile, since grF()\textup{gr}^{F}_{\bullet}(\mathscr{M}) and \mathscr{M} both have no yry_{r}-torsion, we have

supp~ir(grF𝒟W~,D~)(ir(grF()))=Chlog()|{yr=0}TlogW~|yr=0.\widetilde{\textup{supp}}_{i^{*}_{r}(\textup{gr}^{F}_{\bullet}\mathscr{D}_{\widetilde{W},\widetilde{D}})}(i_{r}^{*}(\textup{gr}^{F}_{\bullet}(\mathscr{M})))=\textup{Ch}_{\log}(\mathscr{M})|_{\{y_{r}=0\}}\subseteq T^{*}_{\log}\widetilde{W}|_{y_{r}=0}.

Since ξ~xrξyr\widetilde{\xi}_{x_{r}}-\xi_{y_{r}} annihilates grF()\textup{gr}^{F}_{\bullet}(\mathscr{M}), we further have

supp~ir(grF𝒟W~,D~)(ir(grF()))=ι~(Chlog(r1)).\widetilde{\textup{supp}}_{i^{*}_{r}(\textup{gr}^{F}_{\bullet}\mathscr{D}_{\widetilde{W},\widetilde{D}})}(i_{r}^{*}(\textup{gr}^{F}_{\bullet}(\mathscr{M})))=\widetilde{\iota}(\textup{Ch}_{\log}(\mathscr{M}_{r-1})).

We then do induction backwards until we obtain Part (2). ∎

Finally, we describe the closure in log cotangent bundle.

Lemma 3.7.

Let YUY\subset U be a closed stratum, and let Y~=p1(Y)¯W~\widetilde{Y}=\overline{p^{-1}(Y)}\subset\widetilde{W}. Then ι~(TYU¯log)=ι(TπtY~(W~/𝔻tr))|{y=0}\tilde{\iota}(\overline{T_{Y}^{*}U}^{\log})=\iota(T^{*}_{\pi_{t}\mid\widetilde{Y}}(\widetilde{W}/\mathbb{D}_{t}^{r}))|_{\{y=0\}}.

Proof.

Let us unpackage the definition of Y~\widetilde{Y}. For each local function f(x,t)𝒪Uf(x,t)\in\mathscr{O}_{U} that vanishes on YY, we consider f(x,eyt)f(x,e^{y}t) on W~\widetilde{W}. The relative conormals with a general tt fixed, TπtY~(W~/𝔻tr)T^{*}_{\pi_{t}\mid\widetilde{Y}}(\widetilde{W}/\mathbb{D}_{t}^{r}), are generated by the relative differentials

d/πt(p1f)=(xf)dx+t(tf)eydy.d_{/\pi_{t}}(p^{-1}f)=(\partial_{x}f)dx+t(\partial_{t}f)e^{y}dy.

After we apply ι\iota, we get

ι(d/πt(p1f))=(xf)dx+t(tf)eydtt+t(tf)eydy.\iota(d_{/\pi_{t}}(p^{-1}f))=(\partial_{x}f)dx+t(\partial_{t}f)e^{y}\frac{dt}{t}+t(\partial_{t}f)e^{y}dy.

Then we restrict to {y=0}\{y=0\}, meaning setting yi=0y_{i}=0, and forget ξyi\xi_{y_{i}}.

ι(d/πt(p1f))|{y=0}=(xf)dx+t(tf)dtt.\iota(d_{/\pi_{t}}(p^{-1}f))|_{\{y=0\}}=(\partial_{x}f)dx+t(\partial_{t}f)\frac{dt}{t}.

Indeed, this gives us back the dfdf in the basis section of TlogWT_{\log}^{*}W, i.e. dxdx and dt/tdt/t.

The above argument works generically. Taking closure, the proof is done by the construction of relative conormal spaces. ∎

Now the proof of Theorem 1.6 is accomplished by combining Proposition 3.6 and Lemma 3.7.

4. Relative 𝒟\mathscr{D}-modules and VV-filtrations

In this section, we discuss the Kashiwara-Malgrange filtrations for 𝒟\mathscr{D}-modules in the general sense of Sabbah by using relative 𝒟\mathscr{D}-modules. For simplicity, we focus on the algebraic category unless stated otherwise, that is, all the underlying spaces and sheaves on them are algebraic in this section. See Remark 4.12 for the analytic case.

4.1. Kashiwara-Malgrange filtrations

Definition 4.1.

Suppose that XX is a smooth complex variety and YY is a smooth subvariety of XX with its ideal sheaf denoted by Y\mathscr{I}_{Y}. Then the Kashiwara-Malgrange filtration on 𝒟X\mathscr{D}_{X} is a \mathbb{Z}-indexed increasing filtration defined by

VkY𝒟X{P𝒟X|PYjYjk for every j}V^{Y}_{k}\mathscr{D}_{X}\coloneqq\{P\in\mathscr{D}_{X}|\quad P\mathscr{I}_{Y}^{j}\subseteq\mathscr{I}_{Y}^{j-k}\textup{ for every }j\in\mathbb{Z}\}

where Yj=𝒪X\mathscr{I}_{Y}^{j}=\mathscr{O}_{X} if j0j\leq 0.333In the literature, some authors define the Kashiwara-Malgrange filtration on 𝒟X\mathscr{D}_{X} as the decreasing filtration, that is, VYk𝒟XVkY𝒟XV_{Y}^{k}\mathscr{D}_{X}\coloneqq V^{Y}_{-k}\mathscr{D}_{X}. In particular, if Y=HY=H is a smooth hypersurface V0Y𝒟X=𝒟X,HV^{Y}_{0}\mathscr{D}_{X}=\mathscr{D}_{X,H}, the sheaf of logarithmic differential operators along HH.

We then define the associated Rees ring by

RVY𝒟XkVkY𝒟Xuk𝒟X[u,1/u],R^{Y}_{V}\mathscr{D}_{X}\coloneqq\bigoplus_{k\in\mathbb{Z}}V^{Y}_{k}\mathscr{D}_{X}\cdot u^{k}\subseteq\mathscr{D}_{X}[u,1/u],

where the independent variable uu is used to help remember the grading.

Definition 4.2.

Suppose that \mathcal{M} is a (left) 𝒟X\mathscr{D}_{X}-module. A \mathbb{Z}-indexed increasing filtration Ω\Omega_{\bullet}\mathcal{M} is compatible with V𝒟XV_{\bullet}\mathscr{D}_{X} if

VkY𝒟XΩjΩk+j for all k,j.V^{Y}_{k}\mathscr{D}_{X}\cdot\Omega_{j}\mathcal{M}\subseteq\Omega_{k+j}\mathcal{M}\textup{ for all }k,j\in\mathbb{Z}.

A compatible filtration Ω\Omega_{\bullet}\mathcal{M} is a good filtration over V𝒟XV_{\bullet}\mathscr{D}_{X} if the associated Rees module

RΩkΩkuk[u,1/u]R_{\Omega}\mathcal{M}\coloneqq\bigoplus_{k\in\mathbb{Z}}\Omega_{k}\mathcal{M}\cdot u^{k}\subseteq\mathcal{M}[u,1/u]

is coherent over RVY𝒟XR^{Y}_{V}\mathscr{D}_{X}. A good filtration VV_{\bullet}\mathcal{M} is called the Kashiwara-Malgrange filtration on \mathcal{M} if there exists a monic polynomial b(s)[s]b(s)\in\mathbb{C}[s] with its roots having real parts in [0,1)[0,1) so that

b(ititi+k) annihilates grkVYVk/Vk1 for each k,b(\sum_{i}t_{i}\partial_{t_{i}}+k)\textup{ annihilates }\textup{gr}^{V^{Y}}_{k}\mathcal{M}\coloneqq V_{k}\mathcal{M}/V_{k-1}\mathcal{M}\textup{ for each }k\in\mathbb{Z},

where (t1=t2==tr=0)(t_{1}=t_{2}=\cdots=t_{r}=0) locally defines YY and ti\partial_{t_{i}} are the local vector field along the smooth divisor (ti=0)(t_{i}=0). The monic polynomial b(s)b(s) of the least degree is called the Bernstein-Sato polynomial or bb-function of \mathcal{M} along YY.

One can check that the Kashiwara-Malgrange filtration on \mathcal{M} is unique if exists. It is obvious that the existence of the Kashiwara-Malgrange filtration on \mathcal{M} guarantees that \mathcal{M} is coherent over 𝒟X\mathscr{D}_{X}. A coherent 𝒟X\mathscr{D}_{X}-module \mathcal{M} is called specializable along YY if the Kashiwara-Malgrange filtration exists along YY. Furthermore, it is called RR-specializable if the bb-function b(s)b(s) has roots in RR, where RR is a subring of \mathbb{C}.

Theorem 4.3 (Kashiwara).

If \mathcal{M} is holonomic over 𝒟X\mathscr{D}_{X}, then it is specializable along every submanifold YXY\subseteq X.

We will give a proof of the above fundamental theorem under more general settings (cf. Theorem 4.15).

We now recall the definition of nearby cycles and vanishing cycles along smooth hypersurfaces.

Definition 4.4.

Suppose that HH is a smooth hypersurface and it is defined by (t=0)(t=0) locally. We assume that \mathcal{M} is specializable along HH with VV_{\bullet}\mathcal{M} its Kashiwara-Malgrange filtration. Then the nearby cycle of \mathcal{M} along HH is defined by

ΨH()gr0V\Psi_{H}(\mathcal{M})\coloneqq\textup{gr}^{V}_{0}\mathcal{M}

and the vanishing cycle is

ΦH()gr1V.\Phi_{H}(\mathcal{M})\coloneqq\textup{gr}^{V}_{1}\mathcal{M}.

Since the morphism

t:grkVgrk1Vt:\textup{gr}^{V}_{k}\mathcal{M}\longrightarrow\textup{gr}^{V}_{k-1}\mathcal{M}

is an isomorphism of 𝒟H\mathscr{D}_{H}-modules for all k0k\not=0, we then have

ΦH()grkV\Phi_{H}(\mathcal{M})\simeq\textup{gr}^{V}_{k}\mathcal{M}

for k>0k>0 and

ΨH()grkV\Psi_{H}(\mathcal{M})\simeq\textup{gr}^{V}_{k}\mathcal{M}

for k1k\leq-1. For a perverse sheaf KK (with complex coefficients) on XX, we use ψt(K)\psi_{t}(K) and ϕt(K)\phi_{t}(K) to denote the nearby cycle and vanishing cycle of KK along H=(t=0)H=(t=0) respectively. Let us refer to [KS13, §8.6] for their definitions.

The following theorem of Kashiwara is the Riemann-Hilbert correspondence of nearby and vanishing cycles.

Theorem 4.5.

[Kas83, Theorem 2] Suppose that \mathcal{M} is regular holonomic and VV_{\bullet}\mathcal{M} is the Kashiwara-Malgrange filtration along a smooth hypersurface H=(t=0)H=(t=0). Then

  1. (1)

    grkV\textup{gr}^{V}_{k}\mathcal{M} is a regular holonomic 𝒟H\mathscr{D}_{H}-module for every kk\in\mathbb{Z},

  2. (2)

    DRH(ΨH())ψt(DRX)\textup{DR}_{H}(\Psi_{H}(\mathcal{M}))\simeq\psi_{t}(\textup{DR}_{X}\mathcal{M}) and DRH(ΦH())ϕt(DRX).\textup{DR}_{H}(\Phi_{H}(\mathcal{M}))\simeq\phi_{t}(\textup{DR}_{X}\mathcal{M}).

4.2. Algebraic normal deformation

We now recall the normal deformation algebraically; see [KS13, §4.1] for the topological construction.

Suppose that YXY\subseteq X is a smooth subvariety with the ideal sheaf Y\mathscr{I}_{Y}. We algebraically define a space by

X~YSpec [kYkuk],\widetilde{X}_{Y}\coloneqq\textup{Spec }[\bigoplus_{k\in\mathbb{Z}}\mathscr{I}^{-k}_{Y}\otimes u^{k}],

where uu is an independent variable giving a \mathbb{C}^{\star}-action on X~Y\widetilde{X}_{Y}. Then the natural inclusion [u]kYkuk\mathbb{C}[u]\hookrightarrow\bigoplus_{k\in\mathbb{Z}}\mathscr{I}^{-k}_{Y}\otimes u^{k} gives rise to a smooth family

φY:X~Y\varphi_{Y}\colon\widetilde{X}_{Y}\to\mathbb{C}

so that

  1. (1)

    φY1(u)X\varphi_{Y}^{-1}(u)\simeq X if 0u0\not=u\in\mathbb{C};

  2. (2)

    φY1(0)=Spec [k0Yk/Yk+1]TYX\varphi_{Y}^{-1}(0)=\textup{Spec }[\bigoplus_{k\in\mathbb{Z}^{\geq 0}}\mathscr{I}^{k}_{Y}/\mathscr{I}_{Y}^{k+1}]\eqqcolon T_{Y}X, the algebraic normal bundle of YXY\subseteq X.

By the construction of TYXT_{Y}X and VY𝒟XV^{Y}_{\bullet}\mathscr{D}_{X}, one easily observes

(5) grVY𝒟Xπ𝒟TYX,\textup{gr}_{\bullet}^{V^{Y}}\mathscr{D}_{X}\simeq\pi_{*}\mathscr{D}_{T_{Y}X},

where π:TYXX\pi:T_{Y}X\to X is the natural affine morphism (see [Bj93, §II.10] for the analytical case).

Under the identification (5), titi\sum t_{i}\partial_{t_{i}} gives a global section of 𝒟TYX\mathscr{D}_{T_{Y}X} corresponding to the radial vector field of the bundle TYXT_{Y}X (with respect to the natural \mathbb{C}^{\star}-action on TYXT_{Y}X), denoted by ti¯tigrVY𝒟X\sum\overline{t_{i}\partial}_{t_{i}}\in\textup{gr}^{V^{Y}}_{\bullet}\mathscr{D}_{X}. In particular, ti¯ti\sum\overline{t_{i}\partial}_{t_{i}} is independent of choices of t1,,trt_{1},\dots,t_{r}.

Now we assume that \mathcal{M} is a coherent 𝒟X\mathscr{D}_{X}-module with a good filtration Ω\Omega_{\bullet}\mathcal{M} over VY𝒟XV^{Y}_{\bullet}\mathscr{D}_{X}. Since π\pi is affine, we get a coherent 𝒟TYX\mathscr{D}_{T_{Y}X}-module gr~Ω\widetilde{\textup{gr}}^{\Omega}_{\bullet}\mathcal{M} after applying the \sim-functor. We then say that grΩ\textup{gr}^{\Omega}_{\bullet}\mathcal{M} is holonomic (resp. regular holonomic) over grVY𝒟X\textup{gr}^{V^{Y}}_{\bullet}\mathscr{D}_{X} if gr~Ω\widetilde{\textup{gr}}^{\Omega}_{\bullet}\mathcal{M} is so over 𝒟TYX\mathscr{D}_{T_{Y}X}.

The deformation φY\varphi_{Y} induces a deformation from TXT^{*}X to TTYXT^{*}T_{Y}X as follows. We consider the relative cotangent bundle of φY\varphi_{Y}:

TφY:T(X~Y/).T^{*}\varphi_{Y}:T^{*}(\widetilde{X}_{Y}/\mathbb{C})\to\mathbb{C}.

The fiber of TφYT^{*}\varphi_{Y} over 0 is TTYXT^{*}T_{Y}X and T(X~Y/)TTYXTX×T^{*}(\widetilde{X}_{Y}/\mathbb{C})\setminus T^{*}T_{Y}X\simeq T^{*}X\times\mathbb{C}^{\star}.

The following lemma is essentially due to Sabbah (see [Sab87a, Lemme 2.0.1]). We rephrase it algebraically.

Lemma 4.6 (Sabbah).

For a smooth subvariety YXY\subseteq X, we have a natural isomorphism

RVY𝒟X𝒟X~Y/.R^{Y}_{V}\mathscr{D}_{X}\simeq\mathscr{D}_{\widetilde{X}_{Y}/\mathbb{C}}.
Proof.

We pick a (étale) local coordinates (x1,,xnr,t1,,tr)(x_{1},\dots,x_{n-r},t_{1},\dots,t_{r}) so that (t1=t2==tr=0)(t_{1}=t_{2}=\cdots=t_{r}=0) defines YY, where n=dimXn=\dim X and rr is the codimension of YXY\subseteq X (cf. [HTT08, §A.5]). Since d(tiku)=ktik1dtud(t^{k}_{i}\cdot u)=kt_{i}^{k-1}dt\cdot u (taking differentials over [u]\mathbb{C}[u]), we have a local decomposition of the relative cotangent sheaf

ΩX~Y/1=kYk(i=1rdtiuk1j=1nrdxjuk)\Omega^{1}_{\widetilde{X}_{Y}/\mathbb{C}}=\bigoplus_{k\in\mathbb{Z}}\mathscr{I}_{Y}^{-k}(\bigoplus_{i=1}^{r}dt_{i}\otimes u^{k-1}\bigoplus_{j=1}^{n-r}dx_{j}\otimes u^{k})

and hence

(6) 𝒯X~Y/=kYk(i=1rtiuk+1j=1nrxjuk).\mathscr{T}_{\widetilde{X}_{Y}/\mathbb{C}}=\bigoplus_{k\in\mathbb{Z}}\mathscr{I}_{Y}^{-k}(\bigoplus_{i=1}^{r}\partial_{t_{i}}\otimes u^{k+1}\bigoplus_{j=1}^{n-r}\partial_{x_{j}}\otimes u^{k}).

Since 𝒯X~Y/\mathscr{T}_{\widetilde{X}_{Y}/\mathbb{C}} and 𝒪X~Y\mathscr{O}_{\widetilde{X}_{Y}} generate 𝒟X~Y/\mathscr{D}_{\widetilde{X}_{Y}/\mathbb{C}}, the required isomorphism then follows. ∎

By Lemma 4.6, we immediately have:

Proposition 4.7.
T(X~Y/)=Spec [grFRVY𝒟X],T^{*}(\widetilde{X}_{Y}/\mathbb{C})=\textup{Spec }[\textup{gr}^{F}_{\bullet}R^{Y}_{V}\mathscr{D}_{X}],

where F(RVY𝒟X)F_{\bullet}(R^{Y}_{V}\mathscr{D}_{X}) is the order filtration for relative differential operators induced from the order filtration on 𝒟X\mathscr{D}_{X}.

4.3. Side-change for Rees modules

We discuss side-changes for RVY𝒟XR^{Y}_{V}\mathscr{D}_{X}-modules. The proof of Lemma 4.6 implies that the relative canonical sheaf of φY\varphi_{Y} is

ωX~Y/=kωX𝒪XYkukr,\omega_{\widetilde{X}_{Y}/\mathbb{C}}=\bigoplus_{k\in\mathbb{Z}}\omega_{X}\otimes_{\mathscr{O}_{X}}\mathscr{I}_{Y}^{-k}\otimes u^{k-r},

where we use ω\omega_{-} to denote the canonical sheaf of the smooth variety -. We then immediately have the side-change operators

ωX~Y/𝒪():Modl(RVY𝒟X)Modr(RVY𝒟X)\omega_{\widetilde{X}_{Y}/\mathbb{C}}\otimes_{\mathscr{O}}(\bullet)\colon\textup{Mod}^{l}(R^{Y}_{V}\mathscr{D}_{X})\longrightarrow\textup{Mod}^{r}(R^{Y}_{V}\mathscr{D}_{X})

and

ωX~Y/1𝒪():Modr(RVY𝒟X)Modl(RVY𝒟X),\omega^{-1}_{\widetilde{X}_{Y}/\mathbb{C}}\otimes_{\mathscr{O}}(\bullet)\colon\textup{Mod}^{r}(R^{Y}_{V}\mathscr{D}_{X})\longrightarrow\textup{Mod}^{l}(R^{Y}_{V}\mathscr{D}_{X}),

where Modl(RVY𝒟X)\textup{Mod}^{l}(R^{Y}_{V}\mathscr{D}_{X}) (resp. Modr(RVY𝒟X)\textup{Mod}^{r}(R^{Y}_{V}\mathscr{D}_{X})) is the abelian category of left (resp. right) RV𝒟XR_{V}\mathscr{D}_{X}-modules. Similar to the absolute case, the side-change operators give an equivalence between Modl(RVY𝒟X)\textup{Mod}^{l}(R^{Y}_{V}\mathscr{D}_{X}) and Modr(RVY𝒟X)\textup{Mod}^{r}(R^{Y}_{V}\mathscr{D}_{X}).

Since ωX~Y/[u]0ωTYX\omega_{\widetilde{X}_{Y}/\mathbb{C}}\otimes_{\mathbb{C}[u]}\mathbb{C}_{0}\simeq\omega_{T_{Y}X} and RVY𝒟X[u]0grVY𝒟XR^{Y}_{V}\mathscr{D}_{X}\otimes_{\mathbb{C}[u]}\mathbb{C}_{0}\simeq\textup{gr}_{\bullet}^{V^{Y}}\mathscr{D}_{X} (0\mathbb{C}_{0} is the residue field of 0Spec [u]0\in\textup{Spec }\mathbb{C}[u]), we immediately have the following commutative diagram

(7) Modl(RVY𝒟X){\textup{Mod}^{l}(R^{Y}_{V}\mathscr{D}_{X})}Modr(RVY𝒟X){\textup{Mod}^{r}(R^{Y}_{V}\mathscr{D}_{X})}Modl(grVY𝒟X){\textup{Mod}^{l}(\textup{gr}^{V^{Y}}_{\bullet}\mathscr{D}_{X})}Modr(grVY𝒟X){\textup{Mod}^{r}(\textup{gr}^{V^{Y}}_{\bullet}\mathscr{D}_{X})}ωX~Y/\scriptstyle{\omega_{\tilde{X}_{Y}/\mathbb{C}}\otimes\bullet}[u]0\scriptstyle{\bullet\otimes_{\mathbb{C}[u]}\mathbb{C}_{0}}ωX~Y/1\scriptstyle{\omega_{\tilde{X}_{Y}/\mathbb{C}}^{-1}\otimes\bullet}[u]0\scriptstyle{\bullet\otimes_{\mathbb{C}[u]}\mathbb{C}_{0}}ωTYX\scriptstyle{\omega_{T_{Y}X}\otimes\bullet}ωTYX1\scriptstyle{\omega_{T_{Y}X}^{-1}\otimes\bullet}

where Modl(grVY𝒟X)\textup{Mod}^{l}(\textup{gr}^{V^{Y}}_{\bullet}\mathscr{D}_{X}) (resp. Modr(grVY𝒟X)\textup{Mod}^{r}(\textup{gr}^{V^{Y}}_{\bullet}\mathscr{D}_{X})) are the abelian category of graded left (resp. right) grV𝒟X\textup{gr}^{V}_{\bullet}\mathscr{D}_{X}-modules.

Furthermore, since RV𝒟X[u]1𝒟XR_{V}\mathscr{D}_{X}\otimes_{\mathbb{C}[u]}\mathbb{C}_{1}\simeq\mathscr{D}_{X} where 1\mathbb{C}_{1} is the residue field of 1Spec [u]1\in\textup{Spec }\mathbb{C}[u], we also have the following commutative diagram

(8) Modl(RV𝒟X){\textup{Mod}^{l}(R_{V}\mathscr{D}_{X})}Modr(RV𝒟X){\textup{Mod}^{r}(R_{V}\mathscr{D}_{X})}Modl(𝒟X){\textup{Mod}^{l}(\mathscr{D}_{X})}Modr(𝒟X).{\textup{Mod}^{r}(\mathscr{D}_{X}).}ωX~Y\scriptstyle{\omega_{\tilde{X}_{Y}}\otimes\bullet}[u]1\scriptstyle{\bullet\otimes_{\mathbb{C}[u]}\mathbb{C}_{1}}ωX~Y1\scriptstyle{\omega_{\tilde{X}_{Y}}^{-1}\otimes\bullet}[u]1\scriptstyle{\bullet\otimes_{\mathbb{C}[u]}\mathbb{C}_{1}}ωX\scriptstyle{\omega_{X}\otimes\bullet}ωX1\scriptstyle{\omega_{X}^{-1}\otimes\bullet}

4.4. Characteristic varieties of nearby cycles.

In this subsection, we calculate the characteristic cycles of nearby cycles.

Suppose that \mathcal{M} is specializable along a smooth hypersurface HXH\subseteq X. Then grkV\textup{gr}_{k}^{V}\mathcal{M} is both a coherent 𝒟X,H\mathscr{D}_{X,H}-module and a coherent 𝒟H\mathscr{D}_{H}-module for every kk\in\mathbb{Z}.

Lemma 4.8.

If \mathcal{M} is holonomic, then the Kashiwara-Malgrange filtrations satisfy

Vk=Vk((H))V_{k}\mathcal{M}=V_{k}(\mathcal{M}(*H))

for every k<0k<0, where (H)\mathcal{M}(*H) is the algebraic localization of \mathcal{M} along HH. In particular, VkV_{k}\mathcal{M} is a 𝒟X,H\mathscr{D}_{X,H}-lattice of (H)\mathcal{M}(*H) for every k<0k<0.

Proof.

We consider the exact sequence

0𝒯(H)𝒬00\to\mathcal{T}\rightarrow\mathcal{M}\rightarrow\mathcal{M}(*H)\rightarrow\mathcal{Q}\to 0

where 𝒯\mathcal{T} is the torsion subsheaf of \mathcal{M} supported on \mathcal{H}, namely 𝒯=H0()\mathcal{T}=\mathcal{H}^{0}_{H}(\mathcal{M}), and 𝒬\mathcal{Q} is the quotient module, or 𝒬=H1()\mathcal{Q}=\mathcal{H}^{1}_{H}(\mathcal{M}). Since 𝒯\mathcal{T} and 𝒬\mathcal{Q} are supported on HH, by Kashiwara’s equivalence (cf. [HTT08, Theorem 1.6.1]) Vk𝒯V_{k}\mathcal{T} and Vk𝒬V_{k}\mathcal{Q} are zero for all k<0k<0. The exact sequence induces another exact sequence

0Vk𝒯VkVk(H)Vk𝒬00\to V_{k}\mathcal{T}\rightarrow V_{k}\mathcal{M}\rightarrow V_{k}\mathcal{M}(*H)\rightarrow V_{k}\mathcal{Q}\to 0

for each kk. We thus have obtained the required statement.

The following theorem is equivalent to [Gin86, Theorem 5.5], where in loc. cit. the nearby cycle is alternatively constructed following the algebraic approach of Beilinson and Bernstein. We give it a proof by applying Theorem 1.6.

Theorem 4.9.

Suppose that \mathcal{M} is a regular holonomic 𝒟X\mathscr{D}_{X}-module and that HXH\subseteq X is a smooth hypersurface. Then

CC(|U)¯|HT(X,H)\overline{\textup{CC}(\mathcal{M}|_{U})}|_{H}\subseteq T^{*}(X,H)

is a Lagrangian cycle in THT(X,H)|HT^{*}H\subseteq T^{*}(X,H)|_{H} with U=XHU=X\setminus H. Furthermore, the nearby cycle ΨH()\Psi_{H}(\mathcal{M}) has the characteristic cycle

CC(ΨH())=CC(|U)¯|HTH.\textup{CC}(\Psi_{H}(\mathcal{M}))=\overline{\textup{CC}(\mathcal{M}|_{U})}|_{H}\subseteq T^{*}H.
Proof.

Since characteristic cycles are local, it is enough to assume H=(t=0)H=(t=0) for some local regular (or holomorphic) function tt. Since ΨH()=grkV\Psi_{H}(\mathcal{M})=\textup{gr}^{V}_{k}\mathcal{M} for k<0k<0, we can focus on the short exact sequence of V0𝒟X=𝒟X,HV_{0}\mathscr{D}_{X}=\mathscr{D}_{X,H}-modules

0V1tV1gr1V0.0\to V_{-1}\mathcal{M}\xrightarrow{\cdot t}V_{-1}\mathcal{M}\rightarrow\textup{gr}^{V}_{-1}\mathcal{M}\to 0.

By Lemma 4.8, V1V_{-1}\mathcal{M} is a 𝒟X,H\mathscr{D}_{X,H}-lattice of \mathcal{M}. Then, by Theorem 1.6, we have

CClog(V1)=CC(|U)¯.\textup{CC}_{\log}(V_{-1}\mathcal{M})=\overline{\textup{CC}(\mathcal{M}|_{U})}.

Similar to the proof of Proposition 2.7, considering the above short exact sequence, we conclude that

CClog(gr1V)=CC(|U)¯|HT(X,H)\textup{CC}_{\log}(\textup{gr}^{V}_{-1}\mathcal{M})=\overline{\textup{CC}(\mathcal{M}|_{U})}|_{H}\subseteq T^{*}(X,H)

and

dim(CC(|U)¯|H)=dimX1.\dim(\overline{\textup{CC}(\mathcal{M}|_{U})}|_{H})=\dim X-1.

Now we pick a good filtration F(gr1V)F_{\bullet}(\textup{gr}^{V}_{-1}\mathcal{M}) over F𝒟HF_{\bullet}\mathscr{D}_{H}. Furthermore, we have a closed embedding

THT(X,H)T^{*}H\hookrightarrow T^{*}(X,H)

defined by ξ~t=0\widetilde{\xi}_{t}=0, where ξ~t\widetilde{\xi}_{t} is the symbol of ttt\partial_{t} in grV𝒟X,H\textup{gr}^{V}_{\bullet}\mathscr{D}_{X,H}. Thus, F(grV1)F_{\bullet}(\textup{gr}^{V}_{-1}\mathcal{M}) is also good over F𝒟X,HF_{\bullet}\mathscr{D}_{X,H}. Since characteristic cycles are independent of good filtrations, we therefore have

CC(ΨH())=CClog(grV1)=CC(|U)¯|HTH.\textup{CC}(\Psi_{H}(\mathcal{M}))=\textup{CC}_{\log}(\textup{gr}^{V}_{-1}\mathcal{M})=\overline{\textup{CC}(\mathcal{M}|_{U})}|_{H}\subseteq T^{*}H.

4.5. Generalized Kashiwara-Malgrange filtrations

We discuss refinements of the Kashiwara-Malgrange filtration by using Sabbah’s multi-filtrations.

Suppose that XX is a smooth complex variety and YXY\subseteq X a smooth subvariety of codimension rr such that

Y=j=1rHj,Y=\bigcap_{j=1}^{r}H_{j},

where H1,H2,,HrH_{1},H_{2},\dots,H_{r} are smooth hypersurfaces intersecting transversally (that is, the divisor D=jHjD=\sum_{j}H_{j} has simple normal crossings). We then call YY a smooth complete intersection of H1,,HrH_{1},\dots,H_{r}.

We use VHj𝒟X{V}^{H_{j}}_{\bullet}\mathscr{D}_{X} to denote the Kashiwara-Malgrange filtration of 𝒟X\mathscr{D}_{X} along HjH_{j} for j=1,,rj=1,\dots,r. For 𝐬=(s1,s2,,sr)r\mathbf{s}=(s_{1},s_{2},\dots,s_{r})\in\mathbb{Z}^{r}, we set

V𝐬𝒟X=j=1rVHjsj𝒟X.V_{\mathbf{s}}\mathscr{D}_{X}=\bigcap_{j=1}^{r}{}{V}^{H_{j}}_{s_{j}}\mathscr{D}_{X}.

As the index 𝐬\mathbf{s} varies in r\mathbb{Z}^{r}, we get an increasing r\mathbb{Z}^{r}-filtration of 𝒟X\mathscr{D}_{X} with respect to the natural partial order on r\mathbb{Z}^{r}, denote by V𝒟XV_{\bullet}\mathscr{D}_{X}. One can easily check

V𝟎𝒟X=𝒟X,D,V_{\mathbf{0}}\mathscr{D}_{X}=\mathscr{D}_{X,D},

where the latter is the sheaf of rings of log differential operators. We write the associated Rees ring by

RV𝒟X:=𝐬rV𝐬𝒟Xj=1rujsi,R_{V}\mathscr{D}_{X}:=\bigoplus_{\mathbf{s}\in\mathbb{Z}^{r}}V_{\mathbf{s}}\mathscr{D}_{X}\cdot\prod_{j=1}^{r}u_{j}^{s_{i}},

where the product j=1rujsi\prod_{j=1}^{r}u_{j}^{s_{i}} is used to help us remember the multi-grading of RV𝒟XR_{V}\mathscr{D}_{X}.

For a coherent 𝒟X\mathscr{D}_{X}-module \mathcal{M}, similar to Definition 4.2, we say that a r\mathbb{Z}^{r}-filtration UU_{\bullet}\mathcal{M} is compatible with V𝒟XV_{\bullet}\mathscr{D}_{X} if

V𝐬𝒟XU𝐤U𝐤+𝐬V_{\mathbf{s}}\mathscr{D}_{X}\cdot U_{\bf k}\mathcal{M}\subseteq U_{{\bf k}+\mathbf{s}}\mathcal{M}

for all 𝐤,𝐬r{\bf k},\mathbf{s}\in\mathbb{Z}^{r}. Such a filtration UU_{\bullet}\mathcal{M} is called good over V𝒟XV_{\bullet}\mathscr{D}_{X} if its associated Rees module RUR_{U}\mathcal{M} is coherent over RV𝒟XR_{V}\mathscr{D}_{X}.

4.6. Refinement of normal deformation

We keep the notations as in the previous subsection. Suppose that YXY\subseteq X is a smooth complete intersection of H1,,HrH_{1},\dots,H_{r}. We denote by Hj\mathscr{I}_{H_{j}} the ideal sheaf of HjH_{j} for j=1,2,,rj=1,2,\dots,r. Define

X~Spec (k1,,krjHjkjujkj).\widetilde{X}\coloneqq\textup{Spec }(\bigoplus_{k_{1},\cdots,k_{r}\in\mathbb{Z}}\bigotimes_{j}{\mathscr{I}_{H_{j}}^{-k_{j}}\otimes u_{j}^{k_{j}}}).

Then the natural inclusion

[u1,,ur]k1,,krjHjkjujkj\mathbb{C}[u_{1},\dots,u_{r}]\hookrightarrow\bigoplus_{k_{1},\cdots,k_{r}\in\mathbb{Z}}\bigotimes_{j}{\mathscr{I}_{H_{j}}^{-k_{j}}\otimes u_{j}^{k_{j}}}

gives rise to a smooth family

φ:X~r\varphi\colon\widetilde{X}\to\mathbb{C}^{r}

so that

  1. (1)

    φ1(u1,,ur)X\varphi^{-1}(u_{1},\dots,u_{r})\simeq X if (u1,,ur)()r(u_{1},\dots,u_{r})\in(\mathbb{C}^{\star})^{r};

  2. (2)

    φ1(𝟎)=TYX\varphi^{-1}(\mathbf{0})=T_{Y}X, the algebraic normal bundle of YXY\subseteq X.

The r\mathbb{Z}^{r}-grading of k1,,krjHjkjujkj\bigoplus_{k_{1},\cdots,k_{r}\in\mathbb{Z}}\prod_{j}{\mathscr{I}_{H_{j}}^{-k_{j}}\otimes u_{j}^{k_{j}}} induces ()k(\mathbb{C}^{\star})^{k}-actions on both X~\widetilde{X} and TYXT_{Y}X. Since YY is a complete intersection, the obvious thing is that

TYX=TH1X×X×XTHrXT_{Y}X=T_{H_{1}}X\times_{X}\cdots\times_{X}T_{H_{r}}X

and hence TYXYT_{Y}X\to Y is a split rank rr vector bundle. Moreover, the induced ()r(\mathbb{C}^{\star})^{r}-action on TYXT_{Y}X is given by rescaling the fibers.

Similar to Lemma 4.6, we obtain:

Lemma 4.10.

We have a natural isomorphism

RV𝒟X𝒟X~/r.R_{V}\mathscr{D}_{X}\simeq\mathscr{D}_{\widetilde{X}/\mathbb{C}^{r}}.

From the above lemma, we immediately conclude that RV𝒟XR_{V}\mathscr{D}_{X} is a coherent and noetherian sheaf of rings.

Similar to Proposition 4.7, we have:

Proposition 4.11.
T(X~/r)=Spec [grFRV𝒟X],T^{*}(\widetilde{X}/\mathbb{C}^{r})=\textup{Spec }[\textup{gr}^{F}_{\bullet}R_{V}\mathscr{D}_{X}],

where F(RYV𝒟X)F_{\bullet}(R^{Y}_{V}\mathscr{D}_{X}) is the order filtration for relative differential operators induced from the order filtration on 𝒟X\mathscr{D}_{X}.

Remark 4.12.

In the case that XX is a complex manifold and YY is an analytic smooth complete intersection, one can construct the complex manifold X~\widetilde{X} similar to the topological construction in [KS13, §4.1] or by using open blowups as in [Sab87a, §2.1]. Then 𝒟X~/r\mathscr{D}_{\widetilde{X}/\mathbb{C}^{r}} is a faithfully flat ring extension of RV𝒟XR_{V}\mathscr{D}_{X} by GAGA, or more precisely

𝒟X~/r=𝒪X~RV𝒪XRV𝒟X,\mathscr{D}_{\widetilde{X}/\mathbb{C}^{r}}=\mathscr{O}_{\widetilde{X}}\otimes_{R_{V}\mathscr{O}_{X}}R_{V}\mathscr{D}_{X},

where

RV𝒪X=k1,,krjHjkjujkj.R_{V}\mathscr{O}_{X}=\bigoplus_{k_{1},\cdots,k_{r}\in\mathbb{Z}}\bigotimes_{j}{\mathscr{I}_{H_{j}}^{-k_{j}}\otimes u_{j}^{k_{j}}}.

As a consequence, all the results in this section can be extended to the analytic case.

4.7. Specializability along arbitrary slopes

Let L=(l1,,lr)L=(l_{1},\dots,l_{r}) be a nonzero primitive covector in (r0)(\mathbb{Z}^{r}_{\geq 0})^{\vee}. We also use LL to denote the ray generated by the primitive vector. We call such LL a slope for the smooth complete intersection YXY\subseteq X and that LL is non-degenerate if each ljl_{j} is not zero and degenerate otherwise. We set

YLlj0Hj.Y_{L}\coloneqq\bigcap_{l_{j}\not=0}H_{j}.

By definition, if LL is non-degenerate, then YL=YY_{L}=Y.

Given a nondegenerate slope LL, we have a toric embedding

ι:k, by u(ul1,,ulk).\iota\colon\mathbb{C}\hookrightarrow\mathbb{C}^{k},\textup{ by }u\mapsto(u^{l_{1}},\cdots,u^{l_{k}}).

We can pull-back X~\widetilde{X}, to get a smooth family φL\varphi_{L} in the following Cartesian diagram:

X~L{\widetilde{X}^{L}}X~{\widetilde{X}}{\mathbb{C}}r{\mathbb{C}^{r}}ιL\scriptstyle{\iota_{L}}φL\scriptstyle{\varphi^{L}}φ\scriptstyle{\varphi}

This can be constructed directly, as

X~L=Spec (k1,,krjHjkj(ulj)kj)\widetilde{X}^{L}=\textup{Spec }(\bigoplus_{k_{1},\cdots,k_{r}\in\mathbb{Z}}\bigotimes_{j}{\mathscr{I}_{H_{j}}^{-k_{j}}\otimes(u^{l_{j}})^{k_{j}}})

and the fiber over u=0u=0 is

(φL)1(0)=X~L|u=0=Spec [(k1,,krjHjij(uaj)ij)[u][u]/(u)]TYX.(\varphi^{L})^{-1}(0)=\widetilde{X}_{L}|_{u=0}=\textup{Spec }[(\bigoplus_{k_{1},\cdots,k_{r}\in\mathbb{Z}}\bigotimes_{j}{\mathscr{I}_{H_{j}}^{-i_{j}}\otimes(u^{a_{j}})^{i_{j}}})\otimes_{\mathbb{C}[u]}\mathbb{C}[u]/(u)]\simeq T_{Y}X.

In other words, X~L\widetilde{X}^{L} gives a normal deformation along the slope direction LL. The isomorphism X~L|u=0TYX\widetilde{X}^{L}|_{u=0}\simeq T_{Y}X induces a \mathbb{C}^{\star}-action on TYXT_{Y}X :

(9) λ(y1,,ynr,ξ1,,ξr)=(y1,,ynr,λl1ξ1,,λlrξr)\lambda\cdot(y_{1},\dots,y_{n-r},\xi_{1},\dots,\xi_{r})=(y_{1},\dots,y_{n-r},\lambda^{l_{1}}\cdot\xi_{1},\dots,\lambda^{l_{r}}\cdot\xi_{r})

for λ={0}\lambda\in\mathbb{C}^{\star}=\mathbb{C}\setminus\{0\} and (y1,,ynr,ξ1,,ξr)TYX(y_{1},\dots,y_{n-r},\xi_{1},\dots,\xi_{r})\in T_{Y}X.

The construction of X~L\widetilde{X}^{L} induces the following Cartesian diagram of relative cotangent bundles:

(10) T(X~L/){T^{*}(\widetilde{X}^{L}/\mathbb{C})}T(X~/r){T^{*}(\widetilde{X}/\mathbb{C}^{r})}{\mathbb{C}}r.{\mathbb{C}^{r}.}TφL\scriptstyle{T^{*}\varphi^{L}}Tφ\scriptstyle{T^{*}\varphi}ι\scriptstyle{\iota}

By Lemma 4.10, we see that RV𝒟XR_{V}\mathscr{D}_{X} is flat over [u1,u2,,ur]\mathbb{C}[u_{1},u_{2},\dots,u_{r}]. We then set:

LRV𝒟X:=ιL(RV𝒟X)=𝒟X~L/.{}^{L}R_{V}\mathscr{D}_{X}:=\iota_{L}^{*}(R_{V}\mathscr{D}_{X})=\mathscr{D}_{\widetilde{X}^{L}/\mathbb{C}}.

In particular, LRV𝒟X{}^{L}R_{V}\mathscr{D}_{X} is coherent and noetherian. If LL is degenerate, then one can replace YY by YLY_{L} to reduce to the non-degenerate case.

Remark 4.13.

Similar to ωX~Y/\omega_{\widetilde{X}_{Y}/\mathbb{C}} in §4.3, one can get the explicit formula for the relative canonical sheaves:

ωX~/r=k1,,krωX𝒪(jHjkjujkj1) and ωX~LY/=ιL(ωX~/r).\omega_{\widetilde{X}/\mathbb{C}^{r}}=\bigoplus_{k_{1},\cdots,k_{r}\in\mathbb{Z}}\omega_{X}\otimes_{\mathscr{O}}(\bigotimes_{j}{\mathscr{I}_{H_{j}}^{-k_{j}}\otimes u_{j}^{k_{j}-1}})\textup{ and }\omega_{\widetilde{X}^{L}_{Y}/\mathbb{C}}=\iota_{L}^{*}(\omega_{\widetilde{X}/\mathbb{C}^{r}}).

By construction, we have the explicit formula for LRV𝒟X{}^{L}R_{V}\mathscr{D}_{X}:

LRV𝒟X=kLVk𝒟Xuk{}^{L}R_{V}\mathscr{D}_{X}=\bigoplus_{k\in\mathbb{Z}}{}^{L}V_{k}\mathscr{D}_{X}\cdot u^{k}

where by definition

LVk𝒟X𝐬r,L𝐬=kV𝐬𝒟X.{}^{L}V_{k}\mathscr{D}_{X}\coloneqq\sum_{\mathbf{s}\in\mathbb{Z}^{r},L\cdot\mathbf{s}=k}V_{\mathbf{s}}\mathscr{D}_{X}.

The graded ring LRV𝒟X{}^{L}R_{V}\mathscr{D}_{X} then induces an increasing \mathbb{Z}-filtration LV𝒟X{}^{L}V_{\bullet}\mathscr{D}_{X} on 𝒟X\mathscr{D}_{X}. We might call LV𝒟X{}^{L}V_{\bullet}\mathscr{D}_{X} the Kashiwara-Malgrange filtration of 𝒟X\mathscr{D}_{X} along the slope LL. Since φL\varphi^{L} is smooth and X~L|u=0TYLX\widetilde{X}^{L}|_{u=0}\simeq T_{Y_{L}}X, we have that

(11) grLV𝒟Xπ𝒟TYL𝒟X\textup{gr}^{{}^{L}V}_{\bullet}\mathscr{D}_{X}\simeq\pi_{*}\mathscr{D}_{T_{Y_{L}}\mathscr{D}_{X}}

where π:TYLXYLX\pi:T_{Y_{L}}X\to Y_{L}\hookrightarrow X is the composition (which is an affine morphism). The \mathbb{Z}-grading of grLV𝒟X\textup{gr}^{{}^{L}V}_{\bullet}\mathscr{D}_{X} corresponds to the \mathbb{C}^{\star}-action in (9).

The \mathbb{C}^{\star}-action in (9) induces a radial vector field on TYLXT_{Y_{L}}X, denoted by vLv_{L}. We assume that locally HjH_{j} are defined by tj=0t_{j}=0 where tjt_{j} are among some local coordinate system (x1,,xnr,t1,,tr)(x_{1},\dots,x_{n-r},t_{1},\dots,t_{r}). Then locally

vL=L(t1t1,t2t2,,trtr).v_{L}=L\cdot(t_{1}\partial_{t_{1}},t_{2}\partial_{t_{2}},\dots,t_{r}\partial_{t_{r}}).
Definition 4.14.

Suppose that YXY\subseteq X is a smooth complete intersection of H1,,HrH_{1},\dots,H_{r}, \mathcal{M} is a ((left)) 𝒟X\mathscr{D}_{X}-module and LL is a slope. A \mathbb{Z}-indexed increasing filtration Ω\Omega_{\bullet}\mathcal{M} is compatible with LV𝒟X{}^{L}V_{\bullet}\mathscr{D}_{X} if

LVk𝒟XΩjΩk+j for all k,j.{}^{L}V_{k}\mathscr{D}_{X}\cdot\Omega_{j}\mathcal{M}\subseteq\Omega_{k+j}\mathcal{M}\textup{ for all }k,j\in\mathbb{Z}.

A compatible filtration Ω\Omega_{\bullet}\mathcal{M} is a good filtration over LV𝒟X{}^{L}V_{\bullet}\mathscr{D}_{X} if the associated Rees module

LRΩkΩkuk[u,1/u]{}^{L}R_{\Omega}\mathcal{M}\coloneqq\bigoplus_{k\in\mathbb{Z}}\Omega_{k}\mathcal{M}\cdot u^{k}\subseteq\mathcal{M}[u,1/u]

is coherent over LRV𝒟X{}^{L}R_{V}\mathscr{D}_{X}. A good filtration LV{}^{L}V_{\bullet}\mathcal{M} is called the Kashiwara-Malgrange filtration on \mathcal{M} along the slope LL if there exists a monic polynomial b(s)[s]b(s)\in\mathbb{C}[s] with its roots having real parts in [0,1)[0,1) so that

b(vL+k) annihilates grLVkLVk/LVk1 for each k.b(v_{L}+k)\textup{ annihilates }\textup{gr}^{{}^{L}V}_{k}\mathcal{M}\coloneqq{}^{L}V_{k}\mathcal{M}/{}^{L}V_{k-1}\mathcal{M}\textup{ for each }k\in\mathbb{Z}.

The monic polynomial b(s)b(s) of the least degree is called the Bernstein-Sato polynomial or bb-function of \mathcal{M} along LL.

Kashiwara-Malgrange filtrations along LL are unique if exist. For an arbitrary slope LL, a coherent 𝒟X\mathscr{D}_{X}-module \mathcal{M} is called LL-specializable if the Kashiwara-Malgrange filtration of \mathcal{M} along LL exists. When L=(1,1,,1)L=(1,1,\dots,1), the Definition 4.14 coincides with Definition 4.2. When L=𝐞jL=\mathbf{e}_{j}, the jj-th unit vector in r0\mathbb{Z}^{r}_{\geq 0}, it coincides with Definition 4.2 for the case when Y=HjY=H_{j}. In particular, specializability along LL is compatible with the specializability defined in §4.1. We can then similarly define RR-specializability along LL for RR a subring of \mathbb{C}.

Suppose that UU_{\bullet}\mathcal{M} is a good filtration of \mathcal{M} over V𝒟XV_{\bullet}\mathscr{D}_{X}. Then for a slope LL we can obtain a compatible filtration LU{}^{L}U_{\bullet}\mathcal{M} over LV𝒟X{}^{L}V_{\bullet}\mathscr{D}_{X} defined by

(12) LUk𝐬r,L𝐬=kU𝐬.^{L}U_{k}\mathcal{M}\coloneqq\sum_{\mathbf{s}\in\mathbb{Z}^{r},L\cdot\mathbf{s}=k}U_{\mathbf{s}}\mathcal{M}.

We denote the associated Rees module by

LRUkLUkuk.{}^{L}R_{U}\mathcal{M}\coloneqq\bigoplus_{k\in\mathbb{Z}}^{L}U_{k}\mathcal{M}\cdot u^{k}.

Since LRUιL(RU)/𝒯u{}^{L}R_{U}\mathcal{M}\simeq\iota_{L}^{*}(R_{U}\mathcal{M})/\mathcal{T}_{u} and the pullback functor for relative 𝒟\mathscr{D}-modules preserves coherence, we have that LU{}^{L}U_{\bullet}\mathcal{M} is good over LV𝒟X{}^{L}V_{\bullet}\mathscr{D}_{X}, where 𝒯u\mathcal{T}_{u} is the uu-torsion subsheaf.

The following result is a natural generalization of Theorem 4.3, which is first observed by Sabbah [Sab87a, §3.1]. We provide an alternative proof with the idea essentially due to Björk.

Theorem 4.15.

Suppose that \mathcal{M} is a holonomic 𝒟X\mathscr{D}_{X}-module and LL is a slope. Then \mathcal{M} is splecializable along LL. Moreover, if Ω\Omega_{\bullet}\mathcal{M} is a good filtration over LV𝒟X{}^{L}V_{\bullet}\mathscr{D}_{X}, then grΩ\textup{gr}^{\Omega}_{\bullet}\mathcal{M} is holonomic over grLV𝒟X\textup{gr}^{{}^{L}V}_{\bullet}\mathscr{D}_{X}.

Proof.

We take a good filtration Ω\Omega_{\bullet}\mathcal{M} over LV𝒟X{}^{L}V_{\bullet}\mathscr{D}_{X} locally. Such a filtration exists at least locally by coherence. Then we apply [Bj93, Appendix IV. Theorem 4.10] and conclude that jgrLV𝒟X(grΩ)=j(),j_{\textup{gr}^{{}^{L}V}_{\bullet}\mathscr{D}_{X}}(\textup{gr}^{\Omega}_{\bullet}\mathcal{M})=j(\mathcal{M}), where the first grade number is the graded number of grΩ\textup{gr}^{\Omega}_{\bullet}\mathcal{M} over grLV𝒟X\textup{gr}^{{}^{L}V}_{\bullet}\mathscr{D}_{X}. Since \mathcal{M} is holonomic, by [Bj93, Appendix IV. Proposition 3.5(2)] j()=nj(\mathcal{M})=n, the dimension of XX. Hence jgrLV𝒟X(grΩ)=nj_{\textup{gr}^{{}^{L}V}_{\bullet}\mathscr{D}_{X}}(\textup{gr}^{\Omega}_{\bullet}\mathcal{M})=n and hence grΩ\textup{gr}^{\Omega}_{\bullet}\mathcal{M} is holonomic over grLV𝒟Xπ𝒟TYLX\textup{gr}^{{}^{L}V}_{\bullet}\mathscr{D}_{X}\simeq\pi_{*}\mathscr{D}_{T_{Y_{L}}X} (since dimTYLX=dimX=n\dim T_{Y_{L}}X=\dim X=n).

Now we consider the operator

θLkvL+k\theta_{L}\coloneqq\bigoplus_{k\in\mathbb{Z}}v_{L}+k

on grΩ\textup{gr}^{\Omega}_{\bullet}\mathcal{M}. By construction, one can easily check

θLEndgrLV𝒟X(grΩ).\theta_{L}\in\textup{End}_{\textup{gr}^{{}^{L}V}_{\bullet}\mathscr{D}_{X}}(\textup{gr}^{\Omega}_{\bullet}\mathcal{M}).

Since grΩ\textup{gr}^{\Omega}_{\bullet}\mathcal{M} is holonomic over grLV𝒟X\textup{gr}^{{}^{L}V}_{\bullet}\mathscr{D}_{X}, we conclude that θL\theta_{L} admits a minimal polynomial b(s)[𝐬]b(s)\in\mathbb{C}[\mathbf{s}]. The real parts of roots of b(s)b(s) might not be contained in [0,1)[0,1). Namely, the good filtration Ω\Omega_{\bullet}\mathcal{M} is not the Kashiwara-Malgrange filtration along LL. We then apply the procedure in the proof of [Kas83, Theorem 1(1)] to adjust the roots of b(s)b(s) and the filtration Ω\Omega_{\bullet}\mathcal{M}. The output of the procedure gives us the Kashiwara-Malgrange filtration LV{}^{L}V_{\bullet}\mathcal{M}. By uniqueness, the local construction glues to the global LV{}^{L}V_{\bullet}\mathcal{M}. Therefore, \mathcal{M} is specializable along LL. ∎

4.8. Micro nearby cycles along arbitrary slopes

We keep notations as in the previous subsection and continue to assume YXY\subseteq X a smooth complete intersection and \mathcal{M} a holonomic 𝒟X\mathscr{D}_{X}-module.

By Theorem 4.15, the Kashiwara-Malgrange filtration LV{}^{L}V_{\bullet}\mathcal{M} of \mathcal{M} along every slope LL. For a nondegenerate slope LL, the module pL()p^{*}_{L}(\mathcal{M}) on X~LTYX\widetilde{X}^{L}\setminus T_{Y}X gives rise to a holonomic 𝒟X~L\mathscr{D}_{\widetilde{X}^{L}}-module, denoted by ~L\widetilde{\mathcal{M}}_{L} (that is, ~L=jL(L)\widetilde{\mathcal{M}}_{L}=j^{L}_{*}(\mathcal{M}_{L})), where pL:X~LTYXX×Xp_{L}\colon\widetilde{X}^{L}\setminus T_{Y}X\simeq X\times\mathbb{C}^{\star}\to X is the natural projection and jL:X~LTYXX~Lj^{L}\colon\widetilde{X}^{L}\setminus T_{Y}X\hookrightarrow\widetilde{X}^{L} is the open embedding.

Lemma 4.16.

Suppose that \mathcal{M} is specializable along a slope LL. Then

gr~LVΨu=0(~L).\widetilde{\textup{gr}}^{{}^{L}V}_{\bullet}\mathcal{M}\simeq\Psi_{u=0}(\widetilde{\mathcal{M}}_{L}).
Proof.

The \mathbb{C}^{\star}-action on X~L\widetilde{X}^{L} is induced by the grading of uu by construction. Hence, the \mathbb{C}^{\star}-action induces the action of the Euler vector field along the smooth divisor TYX=(u=0)X~LT_{Y}X=(u=0)\subseteq\widetilde{X}^{L}. But the \mathbb{C}^{\star}-actions on X~L\widetilde{X}^{L} and hence TYLXT_{Y_{L}}X are both induced by the operator

k(L(t1t1,t2t2,,trtr)+k).\bigoplus_{k\in\mathbb{Z}}(L\cdot(t_{1}\partial_{t_{1}},t_{2}\partial_{t_{2}},\dots,t_{r}\partial_{t_{r}})+k).

The required statement then follows by definition. See also [BMS06, §1.3] for the case L=(1,1,,1)L=(1,1,\dots,1). ∎

By the above lemma and Theorem 4.5, we obtain:

Corollary 4.17.

If \mathcal{M} is regular holonomic, then so is grLV\textup{gr}^{{}^{L}V}_{\bullet}\mathcal{M}.

For a holonomic 𝒟X\mathscr{D}_{X}-module \mathcal{M}, we denote

LRΨTYX()gr~LV.{}^{L}R\Psi_{T_{Y}X}(\mathcal{M})\coloneqq\widetilde{\textup{gr}}^{{}^{L}V}_{\bullet}\mathcal{M}.

Following terminology in [Sch12] (and motivated by Lemma 4.16), we call LRΨTYX(){}^{L}R\Psi_{T_{Y}X}(\mathcal{M}) the micro nearby cycle of \mathcal{M} along LL.

Theorem 4.18.

Assume that \mathcal{M} is a regular holonomic 𝒟X\mathscr{D}_{X}-module and LL is a nondegenerate slope. Then

CCX~H/(LRV)=qL(CC())¯ and CC(LRΨTYX())=qL(CC())¯|TTHX.\textup{CC}_{\tilde{X}_{H}/\mathbb{C}}({}^{L}R_{V}\mathcal{M})=\overline{q^{*}_{L}(\textup{CC}(\mathcal{M}))}\textup{ and }\textup{CC}({}^{L}R\Psi_{T_{Y}X}(\mathcal{M}))=\overline{q^{*}_{L}(\textup{CC}(\mathcal{M}))}|_{T^{*}T_{H}X}.

where qL:T(X~L/)TTYXTX×TXq_{L}:T^{*}(\widetilde{X}^{L}/\mathbb{C})\setminus T^{*}T_{Y}X\simeq T^{*}X\times\mathbb{C}^{\star}\to T^{*}X is the natural projection.

Proof.

We apply Lemma 3.4 and let 𝒩rel=LRV\mathcal{N}_{\textup{rel}}={}^{L}R_{V}\mathcal{M}, 𝒩=~L\mathcal{N}=\widetilde{\mathcal{M}}_{L} and F=uF=u. We thus have

CCX~H/(LRV)=qL(CC())¯.\textup{CC}_{\tilde{X}_{H}/\mathbb{C}}({}^{L}R_{V}\mathcal{M})=\overline{q^{*}_{L}(\textup{CC}(\mathcal{M}))}.

To obtain

CC(LRΨTYX())=qL(CC())¯|TTHX,\textup{CC}({}^{L}R\Psi_{T_{Y}X}(\mathcal{M}))=\overline{q^{*}_{L}(\textup{CC}(\mathcal{M}))}|_{T^{*}T_{H}X},

we apply Proposition 2.7. ∎

Theorem 1.3 follows from combining Theorem 4.15, Corollary 4.17 and Theorem 4.18.

4.9. Sabbah’s toric base-change of RUR_{U}\mathcal{M}

We now recall Sabbah’s toric base-changes of Rees modules.

Let us first introduce some notations. We set

M=r,M+=(0)r,M=M× and M+=(0)r.M=\mathbb{Z}^{r},M^{+}=(\mathbb{Z}_{\geq 0})^{r},M_{\mathbb{Q}}=M\times\mathbb{Q}\textup{ and }M^{+}_{\mathbb{Q}}=(\mathbb{Q}_{\geq 0})^{r}.

Let NN be the dual lattice of MM and define N+N^{+}, NN_{\mathbb{Q}} and N+N^{+}_{\mathbb{Q}} similarly. We then fix a simplicial fan Σ\Sigma in NN_{\mathbb{Q}}. We assume that Σ\Sigma is given by a subdivision of N+N^{+}_{\mathbb{Q}}. Then Σ\Sigma gives a toric variety 𝒮Σ\mathcal{S}_{\Sigma} (by making further subdivision, one can assume 𝒮Σ\mathcal{S}_{\Sigma} is smooth) and a projective birational morphism

νΣ:𝒮ΣrSpec [M+].\nu_{\Sigma}\colon\mathcal{S}_{\Sigma}\rightarrow\mathbb{C}^{r}\simeq\textup{Spec }\mathbb{C}[M^{+}].

For a cone ΓΣ\Gamma\in\Sigma, we set

Γˇ{mMm,n0,nΓ}.\check{\Gamma}\coloneqq\{m\in M\mid\langle m,n\rangle\geq 0,\forall n\in\Gamma\}.

Then

𝒮ΓSpec [Γˇ]𝒮Σ\mathcal{S}_{\Gamma}\coloneqq\textup{Spec }\mathbb{C}[\check{\Gamma}]\hookrightarrow\mathcal{S}_{\Sigma}

gives an open affine patch of 𝒮Σ\mathcal{S}_{\Sigma}. We denote a partial ordering induced by Γ\Gamma on MM by

𝐬Γ𝐬𝐬𝐬Γˇ\mathbf{s}\leq_{\Gamma}\mathbf{s}^{\prime}\Leftrightarrow\mathbf{s}^{\prime}-\mathbf{s}\in\check{\Gamma}

and we say

𝐬<Γ𝐬𝐬Γ𝐬 but not 𝐬Γ𝐬.\mathbf{s}<_{\Gamma}\mathbf{s}^{\prime}\Leftrightarrow\mathbf{s}\leq_{\Gamma}\mathbf{s}^{\prime}\textup{ but not }\mathbf{s}^{\prime}\leq_{\Gamma}\mathbf{s}.

We use (Γ)\mathcal{L}(\Gamma) to denote the finite set of all the primitive generators of Γ\Gamma and write

(Σ)ΓΣ(Γ).\mathcal{L}(\Sigma)\coloneqq\bigcup_{\Gamma\in\Sigma}\mathcal{L}(\Gamma).

We suppose that YXY\subseteq X is a smooth complete intersection of H1,,HrH_{1},\dots,H_{r} and \mathcal{M} a 𝒟X\mathscr{D}_{X}-module with a r\mathbb{Z}^{r}-filtration UU_{\bullet}\mathcal{M} that is good over V𝒟XV_{\bullet}\mathscr{D}_{X}. It is now natural to make UU_{\bullet}\mathcal{M} and V𝒟XV_{\bullet}\mathscr{D}_{X} indexed by MrM\simeq\mathbb{Z}^{r}. We consider the following fiber-product diagram

X~ΣX~×r𝒮Σ{\widetilde{X}_{\Sigma}\coloneqq\widetilde{X}\times_{\mathbb{C}^{r}}\mathcal{S}_{\Sigma}}X~{\widetilde{X}}𝒮Σ{\mathcal{S}_{\Sigma}}r.{\mathbb{C}^{r}.}μΣ\scriptstyle{\mu_{\Sigma}}φΣ\scriptstyle{\varphi_{\Sigma}}φ\scriptstyle{\varphi}νΣ\scriptstyle{\nu_{\Sigma}}

Using the flattening theorem relative over toric varieties, Sabbah and Castro proved the following fundamental theorem:

Theorem 4.19.

[Sab87a, A.1.1] Suppose that UU_{\bullet}\mathcal{M} is a good multi-filtration over V𝒟XV_{\bullet}\mathscr{D}_{X} along the smooth complete intersection YXY\subseteq X. Then there exists a simplicial fan Σ\Sigma subdividing +\mathbb{N}^{+}_{\mathbb{Q}} such that RU~μΣRU/𝒯\widetilde{R_{U}\mathcal{M}}\coloneqq\mu^{*}_{\Sigma}R_{U}\mathcal{M}/\mathcal{T} is flat over 𝒮Σ\mathcal{S}_{\Sigma}, where 𝒯\mathcal{T} is the torsion subsheaf of μΣRU\mu^{*}_{\Sigma}R_{U}\mathcal{M} supported over the exceptional locus of νΣ\nu_{\Sigma}.

Such Σ\Sigma in the above theorem is called a fan adapted to UU_{\bullet}\mathcal{M}. By construction, we have a natural inclusion of r\mathbb{Z}^{r}-graded modules

RU[u1,1/u1,,ur,1/ur]R_{U}\mathcal{M}\hookrightarrow\mathcal{M}[u_{1},1/u_{1},\dots,u_{r},1/u_{r}]

or equivalently

RU|X~N=[u1,1/u1,,ur,1/ur].R_{U}\mathcal{M}|_{\widetilde{X}_{N}}=\mathcal{M}[u_{1},1/u_{1},\dots,u_{r},1/u_{r}].

If we have a cone ΓΣ\Gamma\in\Sigma, then we know

μΣRU|X~Γ=μΓRU=[Γˇ][M+]RU.\mu^{*}_{\Sigma}R_{U}\mathcal{M}|_{\widetilde{X}_{\Gamma}}=\mu^{*}_{\Gamma}R_{U}\mathcal{M}=\mathbb{C}[\check{\Gamma}]\otimes_{\mathbb{C}[M^{+}]}R_{U}\mathcal{M}.

Since μΣ\mu_{\Sigma} and μΓ\mu_{\Gamma} are identical over SNS_{N}, using the pullback functor we have an induced natural morphism

(13) [Γˇ][M+]RU[u1,1/u1,,ur,1/ur].\mathbb{C}[\check{\Gamma}]\otimes_{\mathbb{C}[M^{+}]}R_{U}\mathcal{M}\rightarrow\mathcal{M}[u_{1},1/u_{1},\dots,u_{r},1/u_{r}].

The above morphism is neither necessarily surjective nor injective. What is the its image? To answer this question, Sabbah introduced a refined filtration for each cone ΓΣ\Gamma\in\Sigma by

ΓU𝐬=𝐬Γ𝐬U𝐬,𝐬M.{}^{\Gamma}U_{\mathbf{s}}\mathcal{M}=\sum_{\mathbf{s}^{\prime}\leq_{\Gamma}\mathbf{s}}U_{\mathbf{s}^{\prime}}\mathcal{M},\quad\forall\mathbf{s}\in M.

If Γ\Gamma is a unimodular cone of dimesion <k<k, then one easily sees that ΓU𝐬{}^{\Gamma}U_{\mathbf{s}}\mathcal{M} only depends on the image of 𝐬\mathbf{s} in M/ΓM/\Gamma^{\perp}. Hence, in the special case that LL is a ray in Σ\Sigma, we have

LU𝐬=LUL(𝐬),{}^{L}U_{\mathbf{s}}\mathcal{M}={}^{L}U_{L(\mathbf{s})}\mathcal{M},

that is, LU{}^{L}U_{\bullet}\mathcal{M} is indexed by M/Γ\mathbb{Z}\simeq M/\Gamma^{\perp}. Therefore, LU{}^{L}U_{\bullet}\mathcal{M} coincides with the \mathbb{Z}-indexed filtration defined by (12).

We write the associated Rees module by

ΓRUsMΓU𝐬j=1rujsj.{}^{\Gamma}R_{U}\mathcal{M}\coloneqq\bigoplus_{s\in M}{}^{\Gamma}U_{\mathbf{s}}\mathcal{M}\cdot\prod_{j=1}^{r}u_{j}^{s_{j}}.

One observes that ΓRU{}^{\Gamma}R_{U}\mathcal{M} is the image of the natural morphism (13) and that its kernel is the torsion subsheaf 𝒯|X~Γ\mathcal{T}|_{\widetilde{X}_{\Gamma}}. Therefore, we have proved that

RU~|X~ΓΓRU.\widetilde{R_{U}\mathcal{M}}|_{\widetilde{X}_{\Gamma}}\simeq{}^{\Gamma}R_{U}\mathcal{M}.

Furthermore, Sabbah proved:

Lemma 4.20.

[Sab87a, 2.2.2.Lemme] If ΓRU{}^{\Gamma}R_{U}\mathcal{M} is flat over [Γˇ]\mathbb{C}[\check{\Gamma}], then

ΓU𝐬=L(Γ)LUL(𝐬).{}^{\Gamma}U_{\mathbf{s}}\mathcal{M}=\bigcap_{L\in\mathcal{L}(\Gamma)}{}^{L}U_{L(\mathbf{s})}\mathcal{M}.

It is obvious that [Γˇ][M+]RU\mathbb{C}[\check{\Gamma}]\otimes_{\mathbb{C}[M^{+}]}R_{U}\mathcal{M} is coherent over ΓRV𝒟X{}^{\Gamma}R_{V}\mathscr{D}_{X} and hence ΓRU{}^{\Gamma}R_{U}\mathcal{M} is coherent over ΓRV𝒟X{}^{\Gamma}R_{V}\mathscr{D}_{X}. However, it is not always the case that ΓRU{}^{\Gamma}R_{U}\mathcal{M} is coherent over RV𝒟XR_{V}\mathscr{D}_{X} and hence ΓU{}^{\Gamma}U_{\bullet}\mathcal{M} is not necessarily a good filtration over V𝒟XV_{\bullet}\mathscr{D}_{X}. To fix this, Sabbah defined the saturation of UU_{\bullet}\mathcal{M} by

U¯𝐬primitive vectors LN+LUL(𝐬).\bar{U}_{\mathbf{s}}\mathcal{M}\coloneqq\bigcap_{\textup{primitive vectors }L\in N^{+}}{}^{L}U_{L(\mathbf{s})}\mathcal{M}.
Theorem 4.21 (Sabbah).

Suppose that UU_{\bullet}\mathcal{M} is a good multi-filtration over V𝒟XV_{\bullet}\mathscr{D}_{X} and Σ\Sigma is a simplicial fan adapted to UU_{\bullet}\mathcal{M}. Then U¯\bar{U}_{\bullet}\mathcal{M} is good over V𝒟XV_{\bullet}\mathscr{D}_{X} and

U¯𝐬=L(Σ)LUL(𝐬).\bar{U}_{\mathbf{s}}\mathcal{M}=\bigcap_{L\in\mathcal{L}(\Sigma)}{}^{L}U_{L(\mathbf{s})}\mathcal{M}.
Proof.

We take a cone ΓΣ\Gamma\in\Sigma and consider the natural surjection

μΓRURU~|X~ΓΓRU.\mu^{*}_{\Gamma}R_{U}\mathcal{M}\rightarrow\widetilde{R_{U}\mathcal{M}}|_{\widetilde{X}_{\Gamma}}\simeq{}^{\Gamma}R_{U}\mathcal{M}.

By Theorem 4.19, RU~\widetilde{R_{U}\mathcal{M}} is flat over 𝒮Σ\mathcal{S}_{\Sigma}. By Lemma 4.20, we hence know

RU~|X~Γ=sM(L(Γ)LUL(𝐬))j=1rujsj.\widetilde{R_{U}\mathcal{M}}|_{\widetilde{X}_{\Gamma}}=\bigoplus_{s\in M}(\bigcap_{L\in\mathcal{L}(\Gamma)}{}^{L}U_{L(\mathbf{s})}\mathcal{M})\cdot\prod_{j=1}^{r}u_{j}^{s_{j}}.

Since {X~Γ}ΓΣ\{\widetilde{X}_{\Gamma}\}_{\Gamma\in\Sigma} gives a covering of X~Σ\widetilde{X}_{\Sigma}, we hence obtain that

μΣ(RU~)=sM(L(Σ)LUL(𝐬))j=1rujsj.\mu_{\Sigma*}(\widetilde{R_{U}\mathcal{M}})=\bigoplus_{s\in M}(\bigcap_{L\in\mathcal{L}(\Sigma)}{}^{L}U_{L(\mathbf{s})}\mathcal{M})\cdot\prod_{j=1}^{r}u_{j}^{s_{j}}.

Since νΣ\nu_{\Sigma} is projective, by Proposition 2.8 we conclude that μΣ(RU~)\mu_{\Sigma*}(\widetilde{R_{U}\mathcal{M}}) is coherent over RV𝒟XR_{V}\mathscr{D}_{X}.

By construction, we know that

ΓU𝐬primitive vectors LΓN+LUL(𝐬){}^{\Gamma}U_{\mathbf{s}}\mathcal{M}\subseteq\bigcap_{\textup{primitive vectors }L\in\Gamma\cap N^{+}}{}^{L}U_{L(\mathbf{s})}\mathcal{M}

where on the right hand side the intersection is over all primitive vectors in ΓN+\Gamma\cap N^{+} (not just generators of Γ\Gamma) and hence

ΓU𝐬=L(Γ)LUL(𝐬)=primitive vectors LΓN+LUL(𝐬){}^{\Gamma}U_{\mathbf{s}}\mathcal{M}=\bigcap_{L\in\mathcal{L}(\Gamma)}{}^{L}U_{L(\mathbf{s})}\mathcal{M}=\bigcap_{\textup{primitive vectors }L\in\Gamma\cap N^{+}}{}^{L}U_{L(\mathbf{s})}\mathcal{M}

thanks to Lemma 4.20 again. Therefore,

U¯𝐬=ΓΣΓU𝐬=L(Σ)LUL(𝐬).\bar{U}_{\mathbf{s}}\mathcal{M}=\bigcap_{\Gamma\in\Sigma}{}^{\Gamma}U_{\mathbf{s}}\mathcal{M}=\bigcap_{L\in\mathcal{L}(\Sigma)}{}^{L}U_{L(\mathbf{s})}\mathcal{M}.

and

μΣ(RU~)=RU¯.\mu_{\Sigma*}(\widetilde{R_{U}\mathcal{M}})=R_{\bar{U}}\mathcal{M}.

Since we have proved that μΣ(RU~)\mu_{\Sigma*}(\widetilde{R_{U}\mathcal{M}}) is coherent over RV𝒟XR_{V}\mathscr{D}_{X}, U¯\bar{U}_{\bullet}\mathcal{M} is good over V𝒟XV_{\bullet}\mathscr{D}_{X}. ∎

Remark 4.22.

Let \mathcal{M} be a holonomic 𝒟X\mathscr{D}_{X}-module. By Theorem 4.15, we have LV{}^{L}V_{\bullet}\mathcal{M} the Kashiwara-Malgrange filtration of \mathcal{M} for each slope LL. We then fix an adapted fan Σ\Sigma to a good multi-filtration UU_{\bullet}\mathcal{M}. Now one can naively define

ΣV𝐬=L(Σ)LVL(𝐬)𝐬M,{}^{\Sigma}V_{\mathbf{s}}\mathcal{M}=\bigcap_{L\in\mathcal{L}(\Sigma)}{}^{L}V_{L(\mathbf{s})}\mathcal{M}\quad\forall\mathbf{s}\in M,

which gives a r\mathbb{Z}^{r}-filtration ΣV{}^{\Sigma}V_{\bullet}\mathcal{M} over V𝒟XV_{\bullet}\mathscr{D}_{X}. However, it is not necessarily true in general that ΣV{}^{\Sigma}V_{\bullet}\mathcal{M} is good over V𝒟XV_{\bullet}\mathscr{D}_{X} even if \mathcal{M} is regular holonomic; see [Sab87a, §3.3] for further discussions. This means that one cannot define multi-indexed Kashiwara-Malgrange filtrations in general. On the contrary, Bernstein-Sato polynomials can be generalized successfully to the multi-indexed case (see Theorem 4.23).

Using Theorem 4.21, Sabbah proved the following beautiful result about the existence of multi-variable bb-function. We sketch its proof for completeness.

Theorem 4.23 (Existence of Sabbah’s generalized bb-functions).

Suppose that \mathcal{M} is a holonomic 𝒟X\mathscr{D}_{X}-module with a r\mathbb{Z}^{r}-filtration UU_{\bullet}\mathcal{M} good over V𝒟XV_{\bullet}\mathscr{D}_{X} along a smooth complete intersection YXY\subseteq X of H1,,HrH_{1},\dots,H_{r}. Then there exists a simplicial fan Σ\Sigma subdividing N+N^{+} such that for every nonzero vector 𝐚M+{\bf a}\in M^{+} there exist polynomials b𝐚L(s)[s]b^{\bf a}_{L}(s)\in\mathbb{C}[s] ((depending on 𝐚{\bf a})) for all slopes L(Σ)L\in\mathcal{L}(\Sigma) so that locally

L(Σ)b𝐚L(L(t1t1,,trtr)) annihilates U0U𝐚,\prod_{L\in\mathcal{L}(\Sigma)}b^{\bf a}_{L}(L\cdot(t_{1}\partial_{t_{1}},\dots,t_{r}\partial_{t_{r}}))\textup{ annihilates }\frac{U_{\vec{0}}\mathcal{M}}{U_{-{\bf a}}\mathcal{M}},

where tjt_{j} are local defining functions of HjH_{j}.

Proof.

We take a simplicial fan Σ\Sigma adapted to UU_{\bullet}\mathcal{M}, whose existence is guaranteed by Theorem 4.19. By Theorem 4.21, the saturation of UU_{\bullet}\mathcal{M} is

U¯𝐬=L(Σ)LUL(𝐬)𝐬M.\bar{U}_{\mathbf{s}}\mathcal{M}=\bigcap_{L\in\mathcal{L}(\Sigma)}{}^{L}U_{L(\mathbf{s})}\mathcal{M}\quad\forall\mathbf{s}\in M.

Since U¯\bar{U}_{\bullet}\mathcal{M} is good over V𝒟XV_{\bullet}\mathscr{D}_{X}, there exists a vector 𝐤M+{\bf k}\in M^{+} depending on 𝐚{\bf a} such that

U¯kU𝐚.\bar{U}_{-k}\mathcal{M}\subseteq U_{-{\bf a}}\mathcal{M}.

On the other hand, similar to the proof of Theorem 4.15, we conclude that grLU\textup{gr}^{{}^{L}U}_{\bullet}\mathcal{M} is holonomic over grLV𝒟X\textup{gr}^{{}^{L}V}_{\bullet}\mathscr{D}_{X} for each slope LL. Therefore, there exists bL(s)[s]b_{L}(s)\in\mathbb{C}[s] such that bL(θL)b_{L}(\theta_{L}) kills grLU\textup{gr}^{{}^{L}U}_{\bullet}\mathcal{M}. In particular, there exists b𝐚L(s)[s]b^{\bf a}_{L}(s)\in\mathbb{C}[s] so that

b𝐚L(L(t1t1,,trtr))LU0LUL(𝐤)b^{\bf a}_{L}(L\cdot(t_{1}\partial_{t_{1}},\dots,t_{r}\partial_{t_{r}}))\cdot{}^{L}U_{0}\mathcal{M}\subseteq{}^{L}U_{L(-{\bf k})}\mathcal{M}

for each L(Σ)L\in\mathcal{L}(\Sigma). Since U0U¯0U_{\vec{0}}\mathcal{M}\subseteq\bar{U}_{\vec{0}}\mathcal{M} and

U¯k=L(Σ)LUL(𝐤)U𝐚,\bar{U}_{-k}\mathcal{M}=\bigcap_{L\in\mathcal{L}(\Sigma)}{}^{L}U_{L(-{\bf k})}\mathcal{M}\subseteq U_{-{\bf a}}\mathcal{M},

the required statement follows. ∎

4.10. Relative characteristic cycles for RUR_{U}\mathcal{M}

We now prove Theorem 1.4. Assume that \mathcal{M} be a regular holonomic 𝒟X\mathscr{D}_{X}-module with a good filtration UU_{\bullet}\mathcal{M} over V𝒟XV_{\bullet}\mathscr{D}_{X}. We write

p:φ1(()r)X×()rXp\colon\varphi^{-1}((\mathbb{C}^{*})^{r})\simeq X\times(\mathbb{C}^{*})^{r}\longrightarrow X

the natural projection, and j:φ1(()r)X~j\colon\varphi^{-1}((\mathbb{C}^{*})^{r})\hookrightarrow\widetilde{X} the open embedding. We then set 𝒩=j(p)\mathcal{N}=j_{*}(p^{*}\mathcal{M}), 𝒩rel=RU\mathcal{N}_{\textup{rel}}=R_{U}\mathcal{M} and F=jujF=\prod_{j}u_{j}. Thus, Theorem 1.4 follows from Lemma 3.4.

If additionally RUR_{U}\mathcal{M} is flat over r\mathbb{C}^{r}, then we pick an arbitrary point αr\alpha\in\mathbb{C}^{r} and general hyperplanes 1,,r\mathcal{H}_{1},\dots,\mathcal{H}_{r} such that α\alpha is a smooth complete intersection of these hyperplanes. Applying Lemma 3.4 and Proposition 2.7 inductively, we conclude that CCrel(RU)\textup{CC}_{\textup{rel}}(R_{U}\mathcal{M}) is relative Lagrangian and hence that RUR_{U}\mathcal{M} is relative holonomic. We have thus proved Proposition 1.5. Now we pick a simplicial fan adapted to UU_{\bullet}\mathcal{M} as in Theorem 4.19. Then RU~\widetilde{R_{U}\mathcal{M}} is flat over SΣS_{\Sigma}. By a similar argument, we can more generally prove:

Proposition 4.24.

In the situation of Theorem 4.19, if \mathcal{M} is a regular holonomic 𝒟X\mathscr{D}_{X}-module, then RU~\widetilde{R_{U}\mathcal{M}} is relative holonomic over SΣS_{\Sigma}.

Since the saturation U¯\bar{U}_{\bullet}\mathcal{M} is good over V𝒟XV_{\bullet}\mathscr{D}_{X} (Theorem 4.21), Theorem 1.4 specifically implies

CCrel(RU¯)=q1(CC())¯.\textup{CC}_{\textup{rel}}(R_{\bar{U}}\mathcal{M})=\overline{q^{-1}(\textup{CC}(\mathcal{M}))}.

By construction, q1(CC())¯\overline{q^{-1}(\textup{CC}(\mathcal{M}))} is a relative conormal space but not necessarily a relative Lagrangian in general (unless r=1r=1). Since

RU¯=μΣ(RU~),R_{\bar{U}}\mathcal{M}={\mu_{\Sigma}}_{*}(\widetilde{R_{U}\mathcal{M}}),

from Proposition 4.24, we see that the direct image functors for relative 𝒟\mathscr{D}-modules under proper base changes do not necessarily preserve relative holonomicity (cf. §2.2).

5. Graph embedding construction of Malgrange

Let 𝐟=(f1,f2,,fr){\bf f}=(f_{1},f_{2},\dots,f_{r}) be a rr-tuple of regular (or holomorphic) functions on a smooth complex variety (or a complex manifold) YY. We write

j𝐟:UYY(i=1rfi=0)Yj_{\bf f}\colon U_{Y}\coloneqq Y\setminus(\prod_{i=1}^{r}f_{i}=0)\hookrightarrow Y

the open embedding. We consider the graph embedding

ι𝐟:YXY×rx(x,f1(x),,fr(x)).\iota_{{\bf f}}\colon Y\hookrightarrow X\coloneqq Y\times\mathbb{C}^{r}\quad x\mapsto(x,f_{1}(x),\dots,f_{r}(x)).

Let Y\mathcal{M}_{Y} be a holonomic 𝒟Y\mathscr{D}_{Y}-module. We set ~=Y(DY)\widetilde{\mathcal{M}}=\mathcal{M}_{Y}(*D_{Y}) with the divisor DY=(i=1rfi=0)D_{Y}=(\prod_{i=1}^{r}f_{i}=0), which is a holonomic 𝒟Y\mathscr{D}_{Y}-module. We assume

~=𝒟Y0\widetilde{\mathcal{M}}=\mathscr{D}_{Y}\cdot\mathcal{M}_{0}

for some 𝒪Y\mathscr{O}_{Y}-coherent submodule 0~\mathcal{M}_{0}\subseteq\widetilde{\mathcal{M}}. Following the idea of Malgrange [Mal83], we have a coherent 𝒟Y[𝐬]=𝒟Y[𝐬]\mathscr{D}_{Y}[\mathbf{s}]=\mathscr{D}_{Y}\otimes_{\mathbb{C}}\mathbb{C}[\mathbf{s}]-submodule

𝒟Y[𝐬](𝐟𝐬0)~[𝐬]𝐟𝐬\mathscr{D}_{Y}[\mathbf{s}]({\bf f}^{\mathbf{s}}\cdot\mathcal{M}_{0})\subseteq\widetilde{\mathcal{M}}[\mathbf{s}]\cdot{\bf f}^{\mathbf{s}}

where 𝐬=(s1,s2,,sr)\mathbf{s}=(s_{1},s_{2},\dots,s_{r}),

𝐟𝐬=i=1rfisi,{\bf f}^{\mathbf{s}}=\prod_{i=1}^{r}f_{i}^{s_{i}},

and the 𝒟Y[𝐬]\mathscr{D}_{Y}[\mathbf{s}]-module structure is induced by

θ(𝐟𝐬m0)=𝐟𝐬θ(m0)+i=1rsiθ(fi)fi𝐟𝐬m0\theta\cdot({\bf f}^{\mathbf{s}}\cdot m_{0})={\bf f}^{\mathbf{s}}\cdot\theta(m_{0})+\sum_{i=1}^{r}s_{i}\frac{\theta(f_{i})}{f_{i}}{\bf f}^{\mathbf{s}}\cdot m_{0}

for vector fields θ\theta on YY, where m0m_{0} is a section of 0\mathcal{M}_{0}. Since 𝒟Y[𝐬](𝐟𝐬0)\mathscr{D}_{Y}[\mathbf{s}]({\bf f}^{\mathbf{s}}\cdot\mathcal{M}_{0}) is both a [𝐬]\mathbb{C}[\mathbf{s}]-module and a 𝒟Y\mathscr{D}_{Y}-module, it is a coherent relative 𝒟\mathscr{D}-module over [𝐬]\mathbb{C}[\mathbf{s}]. However, ~[𝐬]𝐟𝐬\widetilde{\mathcal{M}}[\mathbf{s}]\cdot{\bf f}^{\mathbf{s}} is not coherent over 𝒟Y[𝐬]\mathscr{D}_{Y}[\mathbf{s}].

We denote by (t1,,tr)(t_{1},\dots,t_{r}) the coordinates of r\mathbb{C}^{r}. The key point is that after identifying sis_{i} with titi-t_{i}\partial_{t_{i}}, we have a 𝒟X\mathscr{D}_{X}-module isomorphism

ι𝐟+(~)ι𝐟(~[𝐬]𝐟𝐬)\iota_{{\bf f}+}(\widetilde{\mathcal{M}})\simeq\iota_{{\bf f}*}({\widetilde{\mathcal{M}}}[\mathbf{s}]{\bf f}^{\mathbf{s}})

with the tit_{i}-action on ι𝐟(~[𝐬]𝐟𝐬)\iota_{{\bf f}*}({\widetilde{\mathcal{M}}}[\mathbf{s}]{\bf f}^{\mathbf{s}}) given by

ti(b(𝐬)𝐟𝐬m0)=b(s1,,si1,si+1,si+1,,sr)fi𝐟𝐬m0.t_{i}\cdot(b(\mathbf{s}){\bf f}^{\mathbf{s}}\cdot m_{0})=b(s_{1},\dots,s_{i-1},s_{i}+1,s_{i+1},\dots,s_{r})f_{i}{\bf f}^{\mathbf{s}}\cdot m_{0}.

Consequently, ι𝐟(𝒟Y[𝐬](𝐟𝐬0))\iota_{{\bf f}*}(\mathscr{D}_{Y}[\mathbf{s}]({\bf f}^{\mathbf{s}}\cdot\mathcal{M}_{0})) is a 𝒟X,D\mathscr{D}_{X,D}-lattice of ι𝐟+(~)\iota_{{\bf f}+}(\widetilde{\mathcal{M}}), where DD is the divisor defined by (t1tr=0)(t_{1}\cdots t_{r}=0). Since ι𝐟(𝒟Y[𝐬](𝐟𝐬0))\iota_{{\bf f}*}(\mathscr{D}_{Y}[\mathbf{s}]({\bf f}^{\mathbf{s}}\cdot\mathcal{M}_{0})) is supported on the graph of YY, abusing notations, we also say 𝒟Y[𝐬](𝐟𝐬0)\mathscr{D}_{Y}[\mathbf{s}]({\bf f}^{\mathbf{s}}\cdot\mathcal{M}_{0}) is a 𝒟X,D\mathscr{D}_{X,D}-lattice of ι𝐟+(~)\iota_{{\bf f}+}(\widetilde{\mathcal{M}}). Then 0\mathcal{M}_{0} generates a holonomic 𝒟X\mathscr{D}_{X}-module

=𝒟Xι𝐟0ι𝐟+~\mathcal{M}=\mathscr{D}_{X}\cdot\iota_{{\bf f}*}\mathcal{M}_{0}\subseteq\iota_{{\bf f}+}\widetilde{\mathcal{M}}

and 𝒟Y[𝐬](𝐟𝐬0)\mathscr{D}_{Y}[\mathbf{s}]({\bf f}^{\mathbf{s}}\cdot\mathcal{M}_{0}) generates a RV𝒟XR_{V}\mathscr{D}_{X}-module

RV𝒟Xι𝐟(𝒟Y[𝐬](𝐟𝐬0)),R_{V}\mathscr{D}_{X}\cdot\iota_{{\bf f}*}(\mathscr{D}_{Y}[\mathbf{s}]({\bf f}^{\mathbf{s}}\cdot\mathcal{M}_{0})),

where the latter induces a r\mathbb{Z}^{r}-filtration on \mathcal{M} with

U0=ι𝐟(𝒟Y[𝐬](𝐟𝐬0)) and U1=ι𝐟(𝒟Y[𝐬](𝐟𝐬+10)).U_{\vec{0}}\mathcal{M}=\iota_{{\bf f}*}(\mathscr{D}_{Y}[\mathbf{s}]({\bf f}^{\mathbf{s}}\cdot\mathcal{M}_{0}))\textup{ and }U_{-\vec{1}}\mathcal{M}=\iota_{{\bf f}*}(\mathscr{D}_{Y}[\mathbf{s}]({\bf f}^{\mathbf{s}+\vec{1}}\cdot\mathcal{M}_{0})).

We then apply Theorem 4.23 and obtain the Sabbah’s generalized bb-function b(𝐬)[𝐬]b(\mathbf{s})\in\mathbb{C}[\mathbf{s}] for 𝒟Y[𝐬](𝐟𝐬0)\mathscr{D}_{Y}[\mathbf{s}]({\bf f}^{\mathbf{s}}\cdot\mathcal{M}_{0}) such that

b(𝐬)𝒟Y[𝐬](𝐟𝐬0)𝒟Y[𝐬](𝐟𝐬+10)=0b(\mathbf{s})\cdot\dfrac{\mathscr{D}_{Y}[\mathbf{s}]({\bf f}^{\mathbf{s}}\cdot\mathcal{M}_{0})}{\mathscr{D}_{Y}[\mathbf{s}]({\bf f}^{\mathbf{s}+\vec{1}}\cdot\mathcal{M}_{0})}=0

with b(𝐬)b(\mathbf{s}) given by a product of polynomials of degree 1. Sabbah’s generalized bb-functions associated to graph embeddings can be further generalized to notions of Bernstein-Sato ideals (see for instance [Bud15]).

In the graph embedding case, we can construct the log rescaled family globally (cf. §3.4):

p:X~Y×rt×yrY×rt,(x,t,y)(x,eyt).p:\widetilde{X}\coloneqq Y\times\mathbb{C}^{r}_{t}\times\mathbb{C}_{y}^{r}\to Y\times\mathbb{C}^{r}_{t},\quad(x,t,y)\mapsto(x,e^{y}t).

We now give a counterexample of Theorem 3.3 without flatness when r=2r=2:

Example 2.

We take Y=2Y=\mathbb{C}^{2} with coordinates (x1,x2)(x_{1},x_{2}) and 𝐟=(x1x2,x2){\bf f}=(x_{1}x_{2},x_{2}). We consider the 𝒟X,D\mathscr{D}_{X,D}-lattice ¯=𝒟Y[𝐬]𝐟𝐬\bar{\mathcal{M}}=\mathscr{D}_{Y}[\mathbf{s}]{\bf f}^{\mathbf{s}}. Its 𝒟X,D\mathscr{D}_{X,D}-annihilator is

Ann𝒟X,D(¯)=(x1x1+t1t1,x2x2+t2t2,t1x1x2,t2x2).{\textup{Ann}}_{\mathscr{D}_{X,D}}(\bar{\mathcal{M}})=(x_{1}\partial_{x_{1}}+t_{1}\partial_{t_{1}},x_{2}\partial_{x_{2}}+t_{2}\partial_{t_{2}},t_{1}-x_{1}x_{2},t_{2}-x_{2}).

Since ¯\bar{\mathcal{M}} and =p(ι𝐟¯)\mathscr{M}=p^{*}(\iota_{{\bf f}*}\bar{\mathcal{M}}) are both acyclic,

Chrel()=(x1ξx1=ξy1,x2ξx2=ξy2,ey1t1=x1x2,ey2t2=x2).\textup{Ch}_{{\textup{rel}}}(\mathscr{M})=(x_{1}\xi_{x_{1}}=-\xi_{y_{1}},x_{2}\xi_{x_{2}}=-\xi_{y_{2}},e^{y_{1}}t_{1}=x_{1}x_{2},e^{y_{2}}t_{2}=x_{2}).

Thus, the fiber of Chrel()\textup{Ch}_{{\textup{rel}}}(\mathscr{M}) over (t1=0,t2=0)(t_{1}=0,t_{2}=0) satisfies

Chrel()(t1=0,t2=0)=(x1ξx1=ξy1,ξy2=0,x2=0)\textup{Ch}_{{\textup{rel}}}(\mathscr{M})\cap(t_{1}=0,t_{2}=0)=(x_{1}\xi_{x_{1}}=-\xi_{y_{1}},\xi_{y_{2}}=0,x_{2}=0)

and hence the dimension of the fiber is 5>45>4. Therefore, Chrel()\textup{Ch}_{\textup{rel}}(\mathscr{M}) is not relative Lagrangian over 2t\mathbb{C}^{2}_{t}.

The following theorem is a generalization of [Mai16a, Résultat 1]. See [Wu21, Theorem 3.3] for the proof of a more generalized result and also [BVWZ21b, Theorem 4.3.4]when 0=𝒪X\mathcal{M}_{0}=\mathscr{O}_{X}.

Theorem 5.1.

If ~\widetilde{\mathcal{M}} is a holonomic 𝒟Y\mathscr{D}_{Y}-module, then every lattice 𝒟Y[𝐬](𝐟𝐬0)\mathscr{D}_{Y}[\mathbf{s}]({\bf f}^{\mathbf{s}}\cdot\mathcal{M}_{0}) is relative holonomic over [𝐬]\mathbb{C}[\mathbf{s}] and

Chrel(𝒟Y[𝐬](𝐟𝐬0))=Ch(~)×r.\textup{Ch}^{\textup{rel}}(\mathscr{D}_{Y}[\mathbf{s}]({\bf f}^{\mathbf{s}}\cdot\mathcal{M}_{0}))=\textup{Ch}(\widetilde{\mathcal{M}})\times\mathbb{C}^{r}.
Proof of Theorem 1.1.

Since constructibility is local, it is enough to assume that

D=(i=1rxi=0)D=(\prod_{i=1}^{r}x_{i}=0)

with local coordinates (x1,x2,,xn)(x_{1},x_{2},\dots,x_{n}) and ~=(D)\widetilde{\mathcal{M}}=\mathcal{M}(*D) satisfies

(D)=𝒟X0\mathcal{M}(*D)=\mathscr{D}_{X}\cdot\mathcal{M}_{0}

for some coherent 𝒪X\mathscr{O}_{X}-submodule 0\mathcal{M}_{0}. Then we take the graph embedding of smooth log pairs

ι𝐟:(X,D)(Z,DZ)\iota_{\bf f}\colon(X,D)\hookrightarrow(Z,D_{Z})

where 𝐟=(x1,x2,,xr){\bf f}=(x_{1},x_{2},\dots,x_{r}) and Z=X×rZ=X\times\mathbb{C}^{r}, DZ=(t1tr=0)D_{Z}=(t_{1}\cdots t_{r}=0). Similar to the non-log case (cf. [HTT08, Example 1.5.23]), we have

ιlog𝐟+(¯)ι𝐟(¯𝒪ωlogι𝐟𝒟X,Df𝒟Z,DZ)\iota^{\log}_{{\bf f}+}(\bar{\mathcal{M}})\simeq\iota_{{\bf f}*}(\bar{\mathcal{M}}\otimes_{\mathscr{O}}\omega^{\log}_{\iota_{\bf f}}\otimes_{\mathscr{D}_{X,D}}f^{*}\mathscr{D}_{Z,D_{Z}})

and

ιlog𝐟+(¯)ι𝐟+(~).\iota^{\log}_{{\bf f}+}(\bar{\mathcal{M}})\hookrightarrow\iota_{{\bf f}+}(\widetilde{\mathcal{M}}).

Thus, ιlog𝐟+(¯)\iota^{\log}_{{\bf f}+}(\bar{\mathcal{M}}) is a 𝒟Z,DZ\mathscr{D}_{Z,D_{Z}}-lattice of the holonomic 𝒟Z\mathscr{D}_{Z}-module ι𝐟,+(~)\iota_{{\bf f},+}(\widetilde{\mathcal{M}}). But the graph embedding gives us different lattices of ι𝐟,+(~)\iota_{{\bf f},+}(\widetilde{\mathcal{M}}),

𝒟X[𝐬](𝐟0(kD))\mathscr{D}_{X}[\mathbf{s}]({\bf f}\cdot\mathcal{M}_{0}(kD))

for all kk\in\mathbb{Z}. The lattices 𝒟X[𝐬](𝐟0(kD))\mathscr{D}_{X}[\mathbf{s}]({\bf f}\cdot\mathcal{M}_{0}(kD)) are relative holonomic over [𝐬]\mathbb{C}[\mathbf{s}] by Theorem 5.1. Meanwhile, we can compare lattices:

ιlog𝐟+(¯)ι𝐟(𝒟X[𝐬](𝐟0))𝒪Z𝒪Z(kDZ)=ι𝐟(𝒟X[𝐬](𝐟0(kD)))\iota^{\log}_{{\bf f}+}(\bar{\mathcal{M}})\subseteq\iota_{{\bf f}*}(\mathscr{D}_{X}[\mathbf{s}]({\bf f}\cdot\mathcal{M}_{0}))\otimes_{\mathscr{O}_{Z}}\mathscr{O}_{Z}(kD_{Z})=\iota_{{\bf f}*}(\mathscr{D}_{X}[\mathbf{s}]({\bf f}\cdot\mathcal{M}_{0}(kD)))

for k0k\gg 0. The above inclusion and Corollary 2.5 together imply that ι1𝐟(ιlog𝐟+(¯))\iota^{-1}_{\bf f}(\iota^{\log}_{{\bf f}+}(\bar{\mathcal{M}})) is relative holonomic over [𝐬]\mathbb{C}[\mathbf{s}]. By Lemma 2.6,

ι1𝐟(ιlog𝐟+(¯))[𝐬]𝐋𝟎𝐋i𝟎(ι1𝐟(ιlog𝐟+(¯)))\iota^{-1}_{\bf f}(\iota^{\log}_{{\bf f}+}(\bar{\mathcal{M}}))\otimes_{\mathbb{C}[\mathbf{s}]}^{\mathbf{L}}\mathbb{C}_{\mathbf{0}}\simeq\mathbf{L}i_{\mathbf{0}}^{*}(\iota^{-1}_{\bf f}(\iota^{\log}_{{\bf f}+}(\bar{\mathcal{M}})))

is a complex of 𝒟X\mathscr{D}_{X}-modules with holonomic cohomology sheaves, where i𝟎:{𝟎}Spec [𝐬]i_{\mathbf{0}}\colon\{\mathbf{0}\}\hookrightarrow\textup{Spec }\mathbb{C}[\mathbf{s}] is the closed embedding and 𝟎\mathbb{C}_{\mathbf{0}} is the residue field. By the construction of the log de Rham complex and Proposition 3.1, we have

(14) ι𝐟(DRX,D(¯))DRZ,DZ(ιlog𝐟+(¯))ι𝐟(DRX(ι1𝐟(ιlog𝐟+(¯))[𝐬]𝐋𝟎))\iota_{{\bf f}*}(\textup{DR}_{X,D}(\bar{\mathcal{M}}))\simeq\textup{DR}_{Z,D_{Z}}(\iota^{\log}_{{\bf f}+}(\bar{\mathcal{M}}))\simeq\iota_{{\bf f}*}(\textup{DR}_{X}(\iota^{-1}_{\bf f}(\iota^{\log}_{{\bf f}+}(\bar{\mathcal{M}}))\otimes_{\mathbb{C}[\mathbf{s}]}^{\mathbf{L}}\mathbb{C}_{\mathbf{0}}))

where the last quasi-isomorphism follows by identifying sis_{i} with titi-t_{i}\partial_{t_{i}}. Since

ι1𝐟(ιlog𝐟+(¯))[𝐬]𝐋𝟎\iota^{-1}_{\bf f}(\iota^{\log}_{{\bf f}+}(\bar{\mathcal{M}}))\otimes_{\mathbb{C}[\mathbf{s}]}^{\mathbf{L}}\mathbb{C}_{\mathbf{0}}

is a complex of 𝒟X\mathscr{D}_{X}-modules with holonomic cohomology sheaves, DRX,D(¯)\textup{DR}_{X,D}(\bar{\mathcal{M}}) is constructible by Kashiwara’s constructibility theorem (cf. [HTT08, Theorem 4.6.3]). ∎

Remark 5.2.

(1) From the proof of Theorem 1.1, DRX,D(¯)\textup{DR}_{X,D}(\bar{\mathcal{M}}) is not necessarily perverse in general, unless ι1𝐟(ι𝐟+log(¯))\iota^{-1}_{{\bf f}}(\iota_{{\bf f}+}^{\log}(\bar{\mathcal{M}})) is flat over a neighborhood of 𝟎r{\mathbf{0}}\in\mathbb{C}^{r}.
(2) The stratification for the constructible complex DRX,D(¯)\textup{DR}_{X,D}(\bar{\mathcal{M}}) is determined by the stratification of Ch((D))\textup{Ch}(\mathcal{M}(*D)) by (14) and Theorem 5.1.

Proof of Theorem 1.2.

Part (1) is the analytification of [WZ21, Theorem 1.1] with the same proof. We now prove Part (2). We keep the notations as in the proof of Theorem 1.1. By picking some k0,k\gg 0, we have an inclusion of lattices

ιlog𝐟+(¯)ι𝐟(𝒟X[𝐬](𝐟𝐬0(kD))).\iota^{\log}_{{\bf f}+}(\bar{\mathcal{M}})\subseteq\iota_{{\bf f}*}(\mathscr{D}_{X}[\mathbf{s}]({\bf f}^{\mathbf{s}}\cdot\mathcal{M}_{0}(kD))).

We then consider the short exact sequence of 𝒟Z,DZ\mathscr{D}_{Z,D_{Z}}-modules

0ιlog𝐟+(¯)ι𝐟(𝒟X[𝐬](𝐟0(kD)))𝒬0,0\to\iota^{\log}_{{\bf f}+}(\bar{\mathcal{M}})\rightarrow\iota_{{\bf f}*}(\mathscr{D}_{X}[\mathbf{s}]({\bf f}\cdot\mathcal{M}_{0}(kD)))\rightarrow\mathcal{Q}\to 0,

where 𝒬\mathcal{Q} is defined to be the quotient module. Applying Theorem 4.23 to 𝒬\mathcal{Q}, there exists b(𝐬)[𝐬]b(\mathbf{s})\in\mathbb{C}[\mathbf{s}] as a product of linear polynomials in 𝐬\mathbf{s} such that

b(𝐬)𝒬=0.b(\mathbf{s})\cdot\mathcal{Q}=0.

Using substitution, we have

b(𝐬+𝐥)𝒬(lDZ)=0b(\mathbf{s}+\mathbf{l})\cdot\mathcal{Q}(-lD_{Z})=0

for 𝐥=(l,l,,l)r\mathbf{l}=(l,l,\dots,l)\in\mathbb{Z}^{r} and for each ll. Chose l0l\gg 0, so that b(𝐬+𝐥)b(\mathbf{s}+\mathbf{l}) does not vanishing at 𝟎r{\mathbf{0}}\in\mathbb{C}^{r}. Thus, 𝒬(lDZ)𝟎=0.\mathcal{Q}(-lD_{Z})\otimes\mathbb{C}_{\mathbf{0}}=0. Considering the above short exact sequence, since

DRZ,DZ(𝒬(lDZ))ι𝐟(DRX(ι1𝐟(𝒬(lDZ))𝟎))\textup{DR}_{Z,D_{Z}}(\mathcal{Q}(-lD_{Z}))\simeq\iota_{{\bf f}*}(\textup{DR}_{X}(\iota^{-1}_{{\bf f}}(\mathcal{Q}(-lD_{Z}))\otimes\mathbb{C}_{\mathbf{0}}))

we hence obtain a quasi-isomorphism

(15) DRZ,DZ(ιlog𝐟+(¯)(lDZ))q.i.DRZ,DZ(ι𝐟(𝒟X[𝐬](𝐟0((kl)D))))\textup{DR}_{Z,D_{Z}}(\iota^{\log}_{{\bf f}+}(\bar{\mathcal{M}})(-lD_{Z}))\xrightarrow{q.i.}\textup{DR}_{Z,D_{Z}}(\iota_{{\bf f}*}(\mathscr{D}_{X}[\mathbf{s}]({\bf f}\cdot\mathcal{M}_{0}((k-l)D))))

for some lkl\gg k. By construction, we have

DRZ,DZ(ι𝐟(𝒟X[𝐬](𝐟0((kl)D))))ι𝐟(DRX(𝒟X[𝐬](𝐟0((kl)D))𝐋[𝐬]𝟎)).\textup{DR}_{Z,D_{Z}}(\iota_{{\bf f}*}(\mathscr{D}_{X}[\mathbf{s}]({\bf f}\cdot\mathcal{M}_{0}((k-l)D))))\simeq\iota_{{\bf f}*}(\textup{DR}_{X}(\mathscr{D}_{X}[\mathbf{s}]({\bf f}\cdot\mathcal{M}_{0}((k-l)D))\otimes^{\mathbf{L}}_{\mathbb{C}[\mathbf{s}]}\mathbb{C}_{\mathbf{0}})).

We then apply [WZ21, Corollary 5.4], and by the quasi-isomorphism (15) obtain

DRZ,DZ(ιlog𝐟+(¯)(lDZ))ι𝐟(j!(DR(|U))).\textup{DR}_{Z,D_{Z}}(\iota^{\log}_{{\bf f}+}(\bar{\mathcal{M}})(-lD_{Z}))\simeq\iota_{{\bf f}*}(j_{!}(\textup{DR}(\mathcal{M}|_{U}))).

By the projection formula,

ιlog𝐟+(¯)(lDZ)ιlog𝐟+(¯(lD)).\iota^{\log}_{{\bf f}+}(\bar{\mathcal{M}})(-lD_{Z})\simeq\iota^{\log}_{{\bf f}+}(\bar{\mathcal{M}}(-lD)).

Since ι𝐟\iota_{{\bf f}} is a closed embedding, the required quasi-isomorphism then follows from Proposition 3.1.

We now prove the perversity statement without the regularity assumption. Using the argument in proving [WZ21, Theorem 5.2 and 5.3] as well as the discussion in [BVWZ21a, §3.6] in dealing with the local analytic case , one obtains that the [𝐬]\mathbb{C}[\mathbf{s}]-modules 𝒟X[𝐬](𝐟𝐬0(kD))\mathscr{D}_{X}[\mathbf{s}]({\bf f}^{\mathbf{s}}\cdot\mathcal{M}_{0}(kD)) are flat over a Zariski neighborhood of 𝟎Spec [𝐬]{\mathbf{0}}\in\textup{Spec }\mathbb{C}[\mathbf{s}] for all |k|0|k|\gg 0. Using Sabbah’s bb-functions, we then conclude that ι1𝐟(ιlog𝐟+(¯(kD)))\iota^{-1}_{\bf f}(\iota^{\log}_{{\bf f}+}(\bar{\mathcal{M}}(kD))) are flat over a Zariski neighborhood of 𝟎Spec [𝐬]{\mathbf{0}}\in\textup{Spec }\mathbb{C}[\mathbf{s}] for all |k|0|k|\gg 0. Consequently, the required perversity follows. ∎

References

  • [Bj93] Jan-Erik Björk, Analytic 𝒟{\mathcal{D}}-modules and applications, Mathematics and its Applications, vol. 247, Kluwer Academic Publishers Group, Dordrecht, 1993. MR 1232191
  • [BMM02] Joël Briançon, Philippe Maisonobe, and Michel Merle, Équations fonctionnelles associées à des fonctions analytiques, Tr. Mat. Inst. Steklova 238 (2002), 86–96.
  • [BMS06] Nero Budur, Mircea Mustaţǎ, and Morihiko Saito, Bernstein-Sato polynomials of arbitrary varieties, Compos. Math. 142 (2006), no. 3, 779–797. MR 2231202
  • [BS05] Nero Budur and Morihiko Saito, Multiplier ideals, V-filtration, and spectrum, J. Algebraic Geometry 14 (2005), 269–282.
  • [Bud15] Nero Budur, Bernstein-sato ideals and local systems, Annales de l’Institut Fourier 65 (2015), no. 2, 549–603 (en).
  • [BVWZ21a] Nero Budur, Robin van der Veer, Lei Wu, and Peng Zhou, Zero loci of Bernstein–Sato ideals, Invent. Math. (2021), 1–28.
  • [BVWZ21b] by same author, Zero loci of Bernstein-Sato ideals II, Sel. Math. New Ser. 27, 32 (2021).
  • [CM99] Francisco J Calderón-Moreno, Logarithmic differential operators and logarithmic de Rham complexes relative to a free divisor, no. 5, 701–714.
  • [CMNM02] Francisco Calderón-Moreno and Luis Narváez-Macarro, The module 𝒟fs\mathcal{D}f^{s} for locally quasi-homogeneous free divisors, Compositio Mathematica 134 (2002), no. 1, 59–74.
  • [DK16] Andrea D’Agnolo and Masaki Kashiwara, Riemann-Hilbert correspondence for holonomic D-modules, Publications mathématiques de l’IHÉS 123 (2016), no. 1, 69–197.
  • [ELS+04] Lawrence Ein, Robert Lazarsfeld, Karen E. Smith, Dror Varolin, et al., Jumping coefficients of multiplier ideals, Duke Mathematical Journal 123 (2004), no. 3, 469–506.
  • [FFS20] Luisa Fiorot, Teresa Monteiro Fernandes, and Claude Sabbah, Relative regular Riemann-Hilbert correspondence, Proceedings of the London Mathematical Society (2020).
  • [Gin86] Victor Ginsburg, Characteristic varieties and vanishing cycles, Invent. Math. 84 (1986), no. 2, 327–402.
  • [Gin89] by same author, Admissible modules on a symmetric space, Astérisque (1989), no. 173-174, 9–10, 199–255, Orbites unipotentes et représentations, III. MR 1021512
  • [Gro66] Alexander Grothendieck, On the de Rham cohomology of algebraic varieties, Publications Mathématiques de l’Institut des Hautes Études Scientifiques 29 (1966), no. 1, 95–103.
  • [HTT08] Ryoshi Hotta, Kiyoshi Takeuchi, and Toshiyuki Tanisaki, DD-modules, perverse sheaves, and representation theory, Progress in Mathematics, vol. 236, Birkhäuser Boston, Inc., Boston, MA, 2008, Translated from the 1995 Japanese edition by Takeuchi. MR 2357361
  • [Kas75] Masaki Kashiwara, On the maximally overdetermined system of linear differential equations, I, Publications of the Research Institute for Mathematical Sciences 10 (1975), no. 2, 563–579.
  • [Kas83] by same author, Vanishing cycle sheaves and holonomic systems of differential equations, Algebraic geometry (Tokyo/Kyoto, 1982), Lecture Notes in Math., vol. 1016, Springer, Berlin, 1983, pp. 134–142. MR 726425
  • [Kas03] by same author, DD-modules and microlocal calculus, Translations of Mathematical Monographs, vol. 217, American Mathematical Society, Providence, RI, 2003, Translated from the 2000 Japanese original by Mutsumi Saito, Iwanami Series in Modern Mathematics. MR 1943036
  • [Kas77] by same author, BB-functions and holonomic systems. Rationality of roots of bb-functions, Invent. Math. 38 (1976/77), no. 1, 33–53. MR 0430304
  • [KK79] Masaki Kashiwara and Takahiro Kawai, On holonomic systems for l=1n(fl+1O)λl\prod_{l=1}^{n}(f_{l}+\sqrt{-1}{O})^{\lambda_{l}}, Publications of the Research Institute for Mathematical Sciences 15 (1979), no. 2, 551–575.
  • [KN99] Kazuya Kato and Chikara Nakayama, Log betti cohomology, log étale cohomology, and log de Rham cohomology of log schemes over C, Kodai Mathematical Journal 22 (1999), no. 2, 161–186.
  • [Kol97] János Kollár, Singularities of pairs, Proceedings of Symposia in Pure Mathematics, vol. 62, American Mathematical Society, 1997, pp. 221–288.
  • [Kop20] Clemens Koppensteiner, The de Rham functor for logarithmic D-modules, Selecta Mathematica 26 (2020), no. 3, 1–44.
  • [KS13] Masaki Kashiwara and Pierre Schapira, Sheaves on manifolds: With a short history.«les débuts de la théorie des faisceaux». by christian houzel, vol. 292, Springer Science & Business Media, 2013.
  • [KT19] Clemens Koppensteiner and Mattia Talpo, Holonomic and perverse logarithmic D-modules, Advances in Mathematics 346 (2019), 510–545.
  • [Lau83] Gérard Laumon, Sur la catégorie dérivée des 𝒟{\mathcal{D}}-modules filtrés, Algebraic geometry (Tokyo/Kyoto, 1982), Lecture Notes in Math., vol. 1016, Springer, Berlin, 1983, pp. 151–237. MR 726427
  • [Laz] Robert Lazarsfeld, Positivity in algebraic geometry. ii, volume 49,(2004), Springer-Verlag 18, no. 385, 8.
  • [Mai16a] Philippe Maisonobe, Filtration Relative, l’Idéal de Bernstein et ses pentes, preprint arXiv:1610.03354 (2016).
  • [Mai16b] by same author, L’Idéal de Bernstein d’un arrangement libre d’hyperplans linéaires, preprint arXiv:1610.03356 (2016).
  • [Mal83] B. Malgrange, Polynômes de Bernstein-Sato et cohomologie évanescente, Analysis and topology on singular spaces, II, III (Luminy, 1981), Astérisque, vol. 101, Soc. Math. France, Paris, 1983, pp. 243–267. MR 737934
  • [MFS16] Teresa Monteiro Fernandes and Claude Sabbah, Riemann-Hilbert correspondence for mixed twistor D-modules, preprint arXiv: 1609.04192 (2016).
  • [MP19a] Mircea Mustaţă and Mihnea Popa, Hodge ideals, Memoirs of the American Mathematical Society 262 (2019), no. 1268.
  • [MP19b] by same author, Hodge ideals for Q-divisors: birational approach, Journal de l’École polytechnique 6 (2019), 283–328.
  • [MP20] by same author, Hodge ideals for Q-divisors, V-filtration, and minimal exponent, Forum of Mathematics, Sigma, vol. 8, 2020.
  • [NZ09] David Nadler and Eric Zaslow, Constructible sheaves and the Fukaya category, J. Amer. Math. Soc. 22 (2009).
  • [Ogu03] Arthur Ogus, On the logarithmic Riemann-Hilbert correspondence, Doc. Math 655 (2003).
  • [Sab87a] Claude Sabbah, Proximité évanescente. I. La structure polaire d’un 𝒟{\mathcal{D}}-module, Compositio Math. 62 (1987), no. 3, 283–328.
  • [Sab87b] by same author, Proximité évanescente. II. Équations fonctionnelles pour plusieurs fonctions analytiques, Compositio Math. 64 (1987), no. 2, 213–241.
  • [Sai16] Morihiko Saito, Hodge ideals and microlocal V-filtration, preprint arXiv:1612.08667 (2016).
  • [Sch12] Pierre Schapira, Microdifferential systems in the complex domain, vol. 269, Springer Science & Business Media, 2012.
  • [SS94] Pierre Schapira and Jean-Pierre Schneiders, Elliptic pairs I. Relative finiteness and duality, Astérisque (1994), no. 224, 5–60.
  • [Ste88] Jan Stevens, On canonical singularities as total spaces of deformations, Abhandlungen aus dem Mathematischen Seminar der Universität Hamburg, vol. 58, Springer, 1988, p. 275.
  • [Wu21] Lei Wu, Riemann-Hilbert correspondence for Alexander complexes, preprint arXiv:2104.06941 (2021).
  • [WZ21] Lei Wu and Peng Zhou, Log D-modules and index theorems, Forum of Mathematics, Sigma, vol. 9, Cambridge University Press, 2021.