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

Logarithmic intersections of double ramification cycles

David Holmes and Rosa Schwarz
Abstract

We describe a theory of logarithmic Chow rings and tautological subrings for logarithmically smooth algebraic stacks, via a generalisation of the notion of piecewise-polynomial functions. Using this machinery we prove that the double-double ramification cycle lies in the tautological subring of the (classical) Chow ring of the moduli space of curves, and that the logarithmic double ramification cycle is divisorial (as conjectured by Molcho, Pandharipande, and Schmitt).

1 Introduction

If C/SC/S is a family of smooth algebraic curves and {\mathcal{L}} on CC a line bundle, the double ramification cycle measures the locus of points sSs\in S where the line bundle {\mathcal{L}} becomes trivial upon restriction to the fibre CsC_{s}. More formally, 𝖣𝖱()\mathsf{DR}({\mathcal{L}}) is a virtual fundamental class of this locus, living in the Chow group of codimension gg cycles on SS. Extending this class in a natural way to families of (pre)stable curves, and giving a tautological formula, has been the subject of much recent research, including [FP05, Hai13, GZ14, Dud15, FP16, Sch16, JPPZ17, MW20, JPPZ18, HKP18, Hol19a, HS19, HPS19]. In particular, [BHP+20] gives a definition of a double ramification cycle 𝖣𝖱()\mathsf{DR}({\mathcal{L}}) for any line bundle {\mathcal{L}} on any family C/SC/S of prestable curves, and proves a tautological formula for this cycle.

1.1 Double-double ramification cycles are tautological

Suppose now we have two line bundles 1{\mathcal{L}}_{1}, 2{\mathcal{L}}_{2} on a smooth curve C/SC/S. Then the double-double ramification cycle 𝖣𝖱(1,2)\mathsf{DR}({\mathcal{L}}_{1},{\mathcal{L}}_{2}) measures the locus of sSs\in S such that both 1{\mathcal{L}}_{1} and 2{\mathcal{L}}_{2} become trivial on the fibre CsC_{s} – of course, this is just the intersection of the corresponding cycles 𝖣𝖱(1)\mathsf{DR}({\mathcal{L}}_{1}) and 𝖣𝖱(2)\mathsf{DR}({\mathcal{L}}_{2}). The key insight of [HPS19] was that this naive intersection is the ‘wrong’ way to extend this class to a family of (pre)stable curves. Instead, one should construct a new virtual class for the product, and in general it will not equal the product of the virtual classes of the two factors:

𝖣𝖱(1,2)𝖣𝖱(1)𝖣𝖱(2).\mathsf{DR}({\mathcal{L}}_{1},{\mathcal{L}}_{2})\neq\mathsf{DR}({\mathcal{L}}_{1})\cdot\mathsf{DR}({\mathcal{L}}_{2}). (1.1.1)

Why is this new construction better than simply taking the intersection of the classes? One way to see this is to consider what happens when one tensors the line bundles 1{\mathcal{L}}_{1} and 2{\mathcal{L}}_{2} together. For a family of smooth curves one sees easily the formula

𝖣𝖱(1)𝖣𝖱(2)=𝖣𝖱(1)𝖣𝖱(12);\mathsf{DR}({\mathcal{L}}_{1})\mathsf{DR}({\mathcal{L}}_{2})=\mathsf{DR}({\mathcal{L}}_{1})\mathsf{DR}({\mathcal{L}}_{1}\otimes{\mathcal{L}}_{2}); (1.1.2)

this also holds in compact type, which plays a key role in the construction of quadratic double ramification integrals and the noncommutative KdV hierarchy in [BR19]. However, eq. 1.1.2 fails for general families of (pre)stable curves, obstructing the extension of quadratic double ramification integrals beyond the compact-type case (see [HPS19, §8] for an explicit example of this failure). On the other hand, the formula

𝖣𝖱(1,2)=𝖣𝖱(1,12)\mathsf{DR}({\mathcal{L}}_{1},{\mathcal{L}}_{2})=\mathsf{DR}({\mathcal{L}}_{1},{\mathcal{L}}_{1}\otimes{\mathcal{L}}_{2}) (1.1.3)

does hold for arbitrary families, giving hope of extending the results of [BR19] beyond compact-type. This is a particular instance of a GL2()\operatorname{GL}_{2}({\mathbb{Z}})-invariance property for the double-double ramification cycles, which we generalise in theorem 5.6 to GLr()\operatorname{GL}_{r}({\mathbb{Z}})-invariance for rr-fold products.

While the cycle 𝖣𝖱(1,2)\mathsf{DR}({\mathcal{L}}_{1},{\mathcal{L}}_{2}) is in some ways better behaved than the product 𝖣𝖱(1)𝖣𝖱(2)\mathsf{DR}({\mathcal{L}}_{1})\mathsf{DR}({\mathcal{L}}_{2}), until now the question of whether it is a tautological cycle has remained open, and is important to address if we hope to study quadratic double ramification integrals. Our first main theorem resolves this question:

Theorem 1.1.

Let gg, nn be non-negative integers, rr a positive integer, and 1,,r{\mathcal{L}}_{1},\dots,{\mathcal{L}}_{r} be line bundles on the universal curve over ¯g,n\overline{{\mathcal{M}}}_{g,n}. Then the rr-fold double ramification cycle

𝖣𝖱(1,,r)\mathsf{DR}({\mathcal{L}}_{1},\dots,{\mathcal{L}}_{r}) (1.1.4)

lies in the tautological subring of the Chow ring of ¯g,n\overline{{\mathcal{M}}}_{g,n}.

This theorem opens up the possibility of giving an explicit formula for the class 𝖣𝖱(1,,r)\mathsf{DR}({\mathcal{L}}_{1},\dots,{\mathcal{L}}_{r}) in terms of the standard generators of the tautological ring, as was done in [JPPZ17] for the case r=1r=1 (that 𝖣𝖱()\mathsf{DR}({\mathcal{L}}) lies in the tautological ring was proven earlier by Faber and Pandharipande [FP05], but no formula was given at that time).

Remark 1.2.

Ranganathan and Molcho have an independent approach (in a paper to appear soon) to theorem 1.1, by studying the virtual strict transforms of the DR cycle.

1.2 Logarithmic Chow rings

The fundamental reason for the failure of the product formula eq. 1.1.2 for stable curves is that 𝖣𝖱()\mathsf{DR}({\mathcal{L}}) should not really be viewed as a cycle on ¯g,n\overline{{\mathcal{M}}}_{g,n}, but rather it lives naturally on a log blowup of ¯g,n\overline{{\mathcal{M}}}_{g,n} — essentially an iterated blowup in reduced boundary strata. To avoid having to make a choice of blowup, we work on LogCH(¯g,n)\operatorname{LogCH}(\overline{{\mathcal{M}}}_{g,n}), which is defined to be the colimit of all log blowups of ¯g,n\overline{{\mathcal{M}}}_{g,n} and comes with a proper pushforward ν:LogCH(¯g,n)CH(¯g,n)\nu_{*}\colon\operatorname{LogCH}(\overline{{\mathcal{M}}}_{g,n})\to\operatorname{CH}(\overline{{\mathcal{M}}}_{g,n}), which is a group homomorphism but not a ring homomorphism. The construction of 𝖣𝖱()\mathsf{DR}({\mathcal{L}}) can be upgraded (see definition 4.4) to give a cycle 𝖫𝗈𝗀𝖣𝖱()LogCH(¯g,n)\mathsf{LogDR}({\mathcal{L}})\in\operatorname{LogCH}(\overline{{\mathcal{M}}}_{g,n}), whose pushforward to CH(¯g,n)\operatorname{CH}(\overline{{\mathcal{M}}}_{g,n}) is 𝖣𝖱()\mathsf{DR}({\mathcal{L}}). The formula

𝖫𝗈𝗀𝖣𝖱(1)𝖫𝗈𝗀𝖣𝖱(2)=𝖫𝗈𝗀𝖣𝖱(1)𝖫𝗈𝗀𝖣𝖱(12)\mathsf{LogDR}({\mathcal{L}}_{1})\mathsf{LogDR}({\mathcal{L}}_{2})=\mathsf{LogDR}({\mathcal{L}}_{1})\mathsf{LogDR}({\mathcal{L}}_{1}\otimes{\mathcal{L}}_{2}) (1.2.1)

is not hard to prove in LogCH(¯g,n)\operatorname{LogCH}(\overline{{\mathcal{M}}}_{g,n}) (see theorem 5.6). We then define

𝖣𝖱(1,,r)=ν(𝖫𝗈𝗀𝖣𝖱(1)𝖫𝗈𝗀𝖣𝖱(r)),\mathsf{DR}({\mathcal{L}}_{1},\dots,{\mathcal{L}}_{r})=\nu_{*}\left(\mathsf{LogDR}({\mathcal{L}}_{1})\cdots\mathsf{LogDR}({\mathcal{L}}_{r})\right), (1.2.2)

from which the product formula eq. 1.1.3 is immediate. The fact that eq. 1.1.2 fails is then just a symptom of the fact that proper pushforward ν:LogCH(¯g,n)CH(¯g,n)\nu_{*}\colon\operatorname{LogCH}(\overline{{\mathcal{M}}}_{g,n})\to\operatorname{CH}(\overline{{\mathcal{M}}}_{g,n}) is not a ring homomorphism.

1.3 Logarithmic tautological rings

Our proof of theorem 1.1 (that double-double ramification cycles are tautological) will run via showing that 𝖫𝗈𝗀𝖣𝖱()\mathsf{LogDR}({\mathcal{L}}) is tautological; but first we have to decide what it means for a cycle in LogCH(¯g,n)\operatorname{LogCH}(\overline{{\mathcal{M}}}_{g,n}) to be tautological.

In fact, we need to do something slightly more general. Our proof that 𝖫𝗈𝗀𝖣𝖱()\mathsf{LogDR}({\mathcal{L}}) is tautological for a line bundle {\mathcal{L}} on the universal curve over ¯g,n\overline{{\mathcal{M}}}_{g,n} proceeds by reduction to the fact that the usual double ramification cycle is tautological. However, for this reduction step it will be necessary to modify the universal curve (so that it is no longer stable, only prestable), and also to modify the line bundle {\mathcal{L}}. This leads us to study double ramification cycles on the total-degree-zero111In [BHP+20] we do not assume total degree zero, but the DR cycle is supported in total degree zero, so this is only a superficial change. Picard stack 𝔍𝔞𝔠\mathfrak{Jac} of the universal curve over the stack 𝔐\mathfrak{M} of all prestable marked curves, exactly the setting considered in [BHP+20].

It is then necessary to define a tautological subring of LogCH(𝔍𝔞𝔠)\operatorname{LogCH}(\mathfrak{Jac}), which is slightly delicate as this smooth algebraic stack is neither Deligne-Mumford nor quasi-compact. For this we develop a theory of piecewise-polynomial functions on any log algebraic stack, and for log smooth stacks over a field or dedekind scheme we construct a map from the space of piecewise-polynomial functions to the log Chow ring. We then define the tautological subring as the ring generated by image of this map together with pullbacks of classes in the usual tautological ring on 𝔍𝔞𝔠\mathfrak{Jac} (as described in [BHP+20, definition 4]). This leads to our main technical result, from which theorem 1.1 follows easily:

Theorem 1.3.

𝖫𝗈𝗀𝖣𝖱\mathsf{LogDR} lies in the tautological subring of LogCH(𝔍𝔞𝔠)\operatorname{LogCH}(\mathfrak{Jac}).

In fact we prove a stronger result (corollary 4.19); if {\mathcal{L}} is the universal line bundle on the universal curve π:C𝔍𝔞𝔠\pi\colon C\to\mathfrak{Jac}, we define the class

η=π(c1()2)CH(𝔍𝔞𝔠),\eta=\pi_{*}(c_{1}({\mathcal{L}})^{2})\in\operatorname{CH}(\mathfrak{Jac}), (1.3.1)

and prove that 𝖫𝗈𝗀𝖣𝖱\mathsf{LogDR} lies in the subring of LogCH(𝔍𝔞𝔠)\operatorname{LogCH}(\mathfrak{Jac}) generated by boundary divisors and the class η\eta.

1.4 𝖫𝗈𝗀𝖣𝖱\mathsf{LogDR} is divisorial

Double ramification cycles in logarithmic Chow rings are also studied in the recent paper [MPS21], with a particular emphasis on the case of the trivial bundle (corresponding to the top chern class of the Hodge bundle on the moduli space of curves). The objective there is to understand the complexity of 𝖣𝖱(𝒪C)\mathsf{DR}({\mathcal{O}}_{C}) in the Chow ring, in particular to understand when it can be written as a polynomial in divisor classes. It is shown that 𝖣𝖱(𝒪C)\mathsf{DR}({\mathcal{O}}_{C}) cannot be written as polynomial in divisor classes, and conjectured that 𝖫𝗈𝗀𝖣𝖱()\mathsf{LogDR}({\mathcal{L}}) can be written as a polynomial in divisors for all {\mathcal{L}}. As a byproduct of the proof of theorem 1.3 we obtain something a little more general. The ring LogCH(S)\operatorname{LogCH}(S) is graded by codimension, and we write divLogCH(S)\operatorname{divLogCH}(S) for the subring generated in degree 1. Since 𝖫𝗈𝗀𝖣𝖱\mathsf{LogDR} lies in the ring generated by η\eta and boundary divisors, we immediately obtain

Theorem 1.4.
𝖫𝗈𝗀𝖣𝖱divLogCH(𝔍𝔞𝔠).\mathsf{LogDR}\in\operatorname{divLogCH}(\mathfrak{Jac}).

By pulling back to ¯g,n\overline{{\mathcal{M}}}_{g,n} this proves [MPS21, Conjecture C].

1.5 Strategy of proof

As with many things in life, our strategy is best illustrated by carrying it out over ¯1,2\overline{{\mathcal{M}}}_{1,2}. We write CC for the universal curve with markings p1p_{1}, p2p_{2}, and we let =𝒪(2p12p2){\mathcal{L}}={\mathcal{O}}(2p_{1}-2p_{2}). Then 𝖣𝖱()\mathsf{DR}({\mathcal{L}}) is invariant under various changes to {\mathcal{L}}; these are listed quite exhaustively in [BHP+20, §0.6]. In particular, let DD be the prime divisor on CC given by the rational tails (via the isomorphism C=¯1,3C=\overline{{\mathcal{M}}}_{1,3} this is the closure of the locus of curves with a single rational tail and all markings on the tail). Then Invariance V of [BHP+20, §0.6] states that

𝖣𝖱()=𝖣𝖱((D)).\mathsf{DR}({\mathcal{L}})=\mathsf{DR}({\mathcal{L}}(D)). (1.5.1)

Our toehold on 𝖫𝗈𝗀𝖣𝖱\mathsf{LogDR} is obtained by realising that it should satisfy a stronger invariance property, corresponding to twisting by vertical divisors which only exist after blowing up ¯1,2\overline{{\mathcal{M}}}_{1,2}. Let x¯1,2x\in\overline{{\mathcal{M}}}_{1,2} be the 2-marked 2-gon (fig. 1), and let ~1,2\widetilde{{\mathcal{M}}}_{1,2} be the blowup of ¯1,2\overline{{\mathcal{M}}}_{1,2} in xx (fig. 2), with C~\tilde{C} the pullback of CC. The curve CxC_{x} has two irreducible components Y1Y_{1}, Y2Y_{2} (say Y1Y_{1} carries p1p_{1}), and the pullbacks of these to C~\tilde{C} are prime divisors supported over the exceptional locus of the blowup, which we denote Y~i\tilde{Y}_{i}. We would like to say that 𝖫𝗈𝗀𝖣𝖱\mathsf{LogDR} satisfies the invariance

𝖫𝗈𝗀𝖣𝖱()=𝖫𝗈𝗀𝖣𝖱((Y~1)),\mathsf{LogDR}({\mathcal{L}})=\mathsf{LogDR}({\mathcal{L}}(\tilde{Y}_{1})), (1.5.2)

but this makes no sense because Y~1\tilde{Y}_{1} is only a Weil divisor, not a Cartier divisor over the ‘danger’ points marked in fig. 2. To rectify this we blow up C~\tilde{C} quite carefully so that the result C~~\tilde{\tilde{C}} is still a prestable curve, but now Y~1\tilde{Y}_{1} is a Cartier divisor on C~~\tilde{\tilde{C}}. Then the invariance

𝖫𝗈𝗀𝖣𝖱()=𝖫𝗈𝗀𝖣𝖱((Y~1))\mathsf{LogDR}({\mathcal{L}})=\mathsf{LogDR}({\mathcal{L}}(\tilde{Y}_{1})) (1.5.3)

makes sense on C~~\tilde{\tilde{C}}, and is moreover true.

It is at this point perhaps not clear what we have gained; we have replaced the rather simple bundle {\mathcal{L}} on CC by the rather complicated (Y~1){\mathcal{L}}(\tilde{Y}_{1}) on C~~\tilde{\tilde{C}}. The magic is that (Y~1){\mathcal{L}}(\tilde{Y}_{1}) has multidegree 0¯{\underline{0}} — that is, it has degree zero on every irreducible component of every fibre of C~~\tilde{\tilde{C}} (with the exception of the danger points, which we will sweep under the carpet for now). Now, for a line bundle of multidegree 0¯{\underline{0}} the cycle 𝖫𝗈𝗀𝖣𝖱\mathsf{LogDR} is just the pullback of the corresponding 𝖣𝖱\mathsf{DR} (lemma 4.7), and we know the latter to be tautological by Pixton’s formula [BHP+20].

Figure 1: Curve over ¯1,2\overline{{\mathcal{M}}}_{1,2}
danger¯1,2\overline{{\mathcal{M}}}_{1,2}~1,2\widetilde{{\mathcal{M}}}_{1,2}= exceptional locus
Figure 2: ~1,2¯1,2\widetilde{{\mathcal{M}}}_{1,2}\to\overline{{\mathcal{M}}}_{1,2}

1.6 Interpretation as a new invariance of 𝖫𝗈𝗀𝖣𝖱\mathsf{LogDR}

Dimitri Zvonkine asked us whether the six invariance properties listed in [BHP+20, §0.6], together with knowledge of 𝖣𝖱\mathsf{DR} for families C/SC/S, {\mathcal{L}} of multidegree zero, would be enough to determine 𝖣𝖱\mathsf{DR} completely. The answer is no, essentially because the invariances in [BHP+20] do not allow us to twist by vertical divisors on CC coming from non-separating edges. We saw above how to rectify this in the case of ¯1,2\overline{{\mathcal{M}}}_{1,2}; here we give a more general statement of the new invariance satisfied by 𝖫𝗈𝗀𝖣𝖱\mathsf{LogDR}.

Let C/SC/S be a log curve with SS log regular, and {\mathcal{L}} on CC a line bundle. We say C/SC/S is twistable222We thank Rahul Pandharipande for suggesting this terminology if there exists a Cartier divisor DD on CC supported over the boundary of SS and such that (D){\mathcal{L}}(D) has multidegree 0¯{\underline{0}}. We write 𝖫𝗈𝗀𝖣𝖱()\mathsf{LogDR}({\mathcal{L}}) for the pullback of 𝖫𝗈𝗀𝖣𝖱\mathsf{LogDR} from 𝔍𝔞𝔠\mathfrak{Jac} along the map S𝔍𝔞𝔠S\to\mathfrak{Jac} induced by {\mathcal{L}}, and we write 𝖣𝖱((D))\mathsf{DR}({\mathcal{L}}(D)) for the pullback of 𝖣𝖱\mathsf{DR} from 𝔍𝔞𝔠\mathfrak{Jac} along the map S𝔍𝔞𝔠S\to\mathfrak{Jac} induced by (D){\mathcal{L}}(D). Viewing 𝖣𝖱((D))\mathsf{DR}({\mathcal{L}}(D)) as an element of LogCH(S)\operatorname{LogCH}(S) by pullback, our new invariance states

𝖫𝗈𝗀𝖣𝖱()=𝖣𝖱((D)).\mathsf{LogDR}({\mathcal{L}})=\mathsf{DR}({\mathcal{L}}(D)). (1.6.1)

That this invariance holds is quite straightforward once the definitions are set up correctly, see lemma 4.7. However, there are not enough twistable families that 𝖫𝗈𝗀𝖣𝖱\mathsf{LogDR} is determined by 𝖣𝖱\mathsf{DR} and the invariance eq. 1.6.1; requiring multidegree 0¯{\underline{0}} over every point in SS is too restrictive a condition (e.g. it fails over the ‘danger’ points in ~1,2\widetilde{{\mathcal{M}}}_{1,2} mentioned above). Because of this we introduce in definition 4.10 a notion of almost twistable families. In lemma 4.12 we show the analogue of eq. 1.6.1 for almost twistable families, and in lemma 4.16 we show that there are enough almost-twistable families to completely determine 𝖫𝗈𝗀𝖣𝖱\mathsf{LogDR} from 𝖣𝖱\mathsf{DR}.

1.7 Notation and conventions

We work with algebraic stacks in the sense of [Sta13], and with log structures in the sense of Fontaine-Illusie-Kato, for which we find [Ogu18] and [Kat89b] particularly useful general references. The sheaf of monoids on a log scheme (or stack) XX is denoted 𝖬X\mathsf{M}_{X}, and the characteristic (or ghost) sheaf is denoted 𝖬¯X\bar{\mathsf{M}}_{X}, with groupifications 𝖬X𝗀𝗉\mathsf{M}_{X}^{\mathsf{gp}} and 𝖬¯X𝗀𝗉\bar{\mathsf{M}}_{X}^{\mathsf{gp}}. Occasionally we write X¯{\underline{X}} for the underlying scheme (or algebraic stack) of XX.

We work over a field or Dedekind scheme kk equipped with trivial log structure. We work in the category of fine saturated (fs) log schemes (and stacks) over kk. In theorems 5.5, 4.7 and 4.6 we assume that kk has characteristic zero, so that we can apply the results of [BHP+20]; it is plausible that the results would become false were this assumption omitted.

A log algebraic stack is an algebraic stack equipped with an (fs) log structure.

We work almost exclusively with operational Chow groups with rational coefficients, as defined in [BHP+20, §2], denoted CH\operatorname{CH}.

1.8 Acknowledgements

We are very grateful to Younghan Bae, Lawrence Barrott, Samouil Molcho, Giulio Orecchia, Rahul Pandharipande, Dhruv Ranganathan, Johannes Schmitt, Pim Spelier, and Jonathan Wise for numerous discussions of double ramification cycles on Picard stacks and logarithmic Chow groups. The idea of extending the multiplication formulae in [HPS19] to a GLr()\operatorname{GL}_{r}({\mathbb{Z}})-invariance property came up in a discussion with Adrien Sauvaget, and was further developed at the AIM workshop on Double ramification cycles and integrable systems.

The first-named author is very grateful to Alessandro Chiodo for many extensive discussions on computing the double ramification cycle on blowups of ¯g,n\overline{{\mathcal{M}}}_{g,n}, which provided key motivation and examples.

Both authors are supported by NWO grant 613.009.103.

2 Logarithmic Chow rings

2.1 Logarithmic Chow rings of algebraic stacks

In this section we make a slight generalisation of some of the ideas from [HPS19], see also [MPS21]. We work extensively with log schemes (and stacks) which are both regular and log regular; equivalently, with log structures that are induced by normal crossings divisors (see [Niz06]). We make quite some effort in this and other sections to avoid unnecessary separatedness or quasi-compactness assumptions, and to work with algebraic stacks in place of (for example) schemes. This is not (primarily?) due to a particular personal preference, but rather because the objects we consider (such as the stack of log curves, or its universal Picard space) make this essential.

Definition 2.1 ([ALT18, Example 4.3]).

A morphism f:XYf\colon X\to Y of log algebraic stacks is a monoidal alteration if it is proper, log étale, and is an isomorphism over the locus in YY where the log structure is trivial.

Examples of monoidal alterations are log blowups and root stacks. We expect that every monoidal alteration can be dominated by a composition of log blowups and root stacks, but have not written down a proof.

Definition 2.2.

Let XX be an algebraic stack locally of finite type over kk. We define CH(X)\operatorname{CH}(X) to be the operational Chow group of XX with rational coefficients, using finite-type algebraic spaces as test objects, see [BHP+20, §2] for details.

Remark 2.3.

If in addition XX is smooth and Deligne-Mumford then the intersection pairing induces an isomorphism from the usual Chow ring of XX (as defined by Vistoli [Vis89]) to the operational Chow ring CH(X)\operatorname{CH}(X).

Definition 2.4.

Let XX be a log smooth stack of finite type over kk. We define the (operational) log Chow ring of XX to be

LogCH(X)=colimX~CH(X~),\operatorname{LogCH}(X)=\operatorname{colim}_{\tilde{X}}\operatorname{CH}(\tilde{X}), (2.1.1)

where the colimit runs over monoidal alterations X~X\tilde{X}\to X with X~\tilde{X} smooth over kk.

Definition 2.5.

Let zLogCH(X)z\in\operatorname{LogCH}(X) and let UXU\hookrightarrow X be a quasi-compact open. We say the restriction of zz to UU is determined on a monoidal alteration U~U\tilde{U}\to U if there exists a cycle zCH(U~)z^{\prime}\in\operatorname{CH}(\tilde{U}) in the equivalence class of zz as defined in the above remark; we then call zz^{\prime} the determination of zz on U~\tilde{U}.

Remark 2.6.

Because we work with rational coefficients, taking the colimit over log blowups would yield the same operational Chow ring; in particular, our Log Chow ring is the same as that in [HPS19, §9]. Throughout the paper we use the possibility of determining a cycle on a (smooth) log blowup without further comment.

Remark 2.7.

The idea of allowing monoidal alterations rather than just log blowups was suggested to the authors by Leo Herr. It will play little role in most of the paper, but is absolutely essential in section 4.6, where it allows us to apply ideas of [ALT18] on canonical resolution of singularities.

Remark 2.8.

The ring colimX~CH(X~)\operatorname{colim}_{\tilde{X}}\operatorname{CH}(\tilde{X}) can be realised concretely as the disjoint union of the rings CH(X~)\operatorname{CH}(\tilde{X}), modulo the equivalence relation where we set cycle z1z_{1} on X~1\tilde{X}_{1} and z2z_{2} on X~2\tilde{X}_{2} to be equivalent if there exists a monoidal alteration X~\tilde{X} dominating both X~1\tilde{X}_{1} and X~2\tilde{X}_{2} and on which the pullbacks of z1z_{1} and z2z_{2} coincide.

2.2 Operations on the logarithmic Chow ring

Throughout this subsection XX and YY are log smooth stacks of finite type over kk.

Definition 2.9 (LogCH\operatorname{LogCH} is a CH\operatorname{CH}-algebra).

If X~X\tilde{X}\to X is a log blowup then pullback gives a ring homomorphism CH(X)CH(X~)\operatorname{CH}(X)\to\operatorname{CH}(\tilde{X}). These assemble into a ring homomorphism CH(X)LogCH(X)\operatorname{CH}(X)\to\operatorname{LogCH}(X).

Definition 2.10 (Pullback for LogCH\operatorname{LogCH}).

Let f:XYf\colon X\to Y be a morphism and let zLogCH(Y)z\in\operatorname{LogCH}(Y). Let Y~Y\tilde{Y}\to Y be a log blowup on which zz is determined, with Y~\tilde{Y} smooth, and let X~X×YY~\tilde{X}\to X\times_{Y}\tilde{Y} be a log blowup which is smooth. Then the composite X~X\tilde{X}\to X is a log blowup, and f~:X~Y~\tilde{f}\colon\tilde{X}\to\tilde{Y} is lci, and we have a pullback f~!zCH(X~)\tilde{f}^{!}z\in\operatorname{CH}(\tilde{X}). This class f~!z\tilde{f}^{!}z is independent of all choices, and the construction yields a ring homomorphism

f!:LogCH(Y)LogCH(X). f^{!}\colon\operatorname{LogCH}(Y)\to\operatorname{LogCH}(X).  (2.2.1)
Lemma 2.11.

Let f:XYf\colon X\to Y be a morphism, then the following diagram commutes:

CH(Y){\operatorname{CH}(Y)}LogCH(Y){\operatorname{LogCH}(Y)}CH(X){\operatorname{CH}(X)}LogCH(X){\operatorname{LogCH}(X)}f!\scriptstyle{f^{!}}f!\scriptstyle{f^{!}} (2.2.2)
Definition 2.12 (Pushforward from LogCH\operatorname{LogCH} to CH\operatorname{CH}).

Suppose XX is smooth, and let zLogCH(X)z\in\operatorname{LogCH}(X). Let X~X\tilde{X}\to X be a log blowup on which zz is determined, with X~\tilde{X} smooth. Then π:X~X\pi\colon\tilde{X}\to X is proper and lci, so we have a proper pushforward on operational Chow π:CH(X~)CH(X)\pi\colon\operatorname{CH}(\tilde{X})\to\operatorname{CH}(X). These assemble into a pushforward

LogCH(X)CH(X).\operatorname{LogCH}(X)\to\operatorname{CH}(X). (2.2.3)

2.3 Extension to the non-quasi-compact case

Definition 2.13.

Let XX be a log smooth log algebraic stack over kk (we no longer assume XX to be quasi-compact). Let qOp(X)\operatorname{qOp}(X) denote the category of open substacks of XX which are quasi-compact. We define the (operational) log Chow ring of XX to be

LogCH(X)=limUqOp(X)LogCH(U).\operatorname{LogCH}(X)=\operatorname{lim}_{U\in\operatorname{qOp}(X)}\operatorname{LogCH}(U). (2.3.1)
Remark 2.14.

Morally, we can think of an element of LogCH(X)\operatorname{LogCH}(X) as an (operational) cycle on the valuativisation333See for example [Kat89a]. of XX, which can be everywhere-locally represented on some finite log blowup of XX. In the absence of a good theory of Chow groups of valuativisations of algebraic stacks, we make the above definition.

Remark 2.15.

All of the constructions of section 2.2 carry through to this setting by restricting to suitable quasi-compact opens. We will make use of these extensions without further comment.

3 Tautological subrings of logarithmic Chow rings

In this section we develop a fairly general theory of piecewise-polynomial functions on log algebraic stacks, generalising the theory for toric varieties (for which see [Pay06] and the references therein). We use these piecewise-polynomial functions to build tautological subrings of the log Chow ring. Once again we need only log blowups in this section, root stacks are unnecessary.

In the toric case one can hope to realise every element of the Chow ring in terms of piecewise-polynomial functions, which is far from the case in the our context; for example, all piecewise-polynomials functions are zero on a scheme equipped with the trivial log structure, but the Chow ring can be large and interesting. However, in the presence of a non-trivial log structure the piecewise-polynomial functions can still generate many interesting Chow elements.

The theory in this section was largely developed before we became aware of the related work of Molcho, Pandharipande and Schmitt [MPS21], where ‘normally decorated strata classes’ approximately correspond to classes coming from our piecewise-polynomial functions. Their approach is probably better for writing formulae for (log) tautological classes, and ours has the advantage that piecewise-polynomials on opens can be glued (which is very useful when working on large algebraic stacks as we do in this paper; as far as we are aware the theory in [MPS21] has so far only been developed for finite-type Deligne-Mumford stacks).

3.1 Piecewise polynomial functions

Let (X,𝒪X)(X,{\mathcal{O}}_{X}) be a ringed site and {\mathcal{M}} a sheaf of 𝒪X{\mathcal{O}}_{X}-modules. We write Sym\operatorname{Sym}{\mathcal{M}} for the sheafification of the presheaf USym(M(U))U\mapsto\operatorname{Sym}(M(U)); it is a sheaf of 𝒪X{\mathcal{O}}_{X}-algebras. If XX is any site and 𝒜{\mathcal{A}} a sheaf of abelian groups, then we view 𝒜{\mathcal{A}} as a sheaf of modules for the constant sheaf of rings {\mathbb{Z}}, yielding a sheaf Sym𝒜\operatorname{Sym}{\mathcal{A}} of graded {\mathbb{Z}}-algebras.

Example 3.1.

If XX is a scheme and 𝒜{\mathcal{A}} is the constant sheaf n{\mathbb{Z}}^{n} of abelian groups, then Sym𝒜\operatorname{Sym}{\mathcal{A}} is the constant sheaf [x1,,xn]{\mathbb{Z}}[x_{1},\dots,x_{n}].

Definition 3.2.

We define the sheaf of piecewise-polynomial functions on a log algebraic stack SS as

PPSSym𝖬¯S𝗀𝗉.\operatorname{PP}_{S}\coloneqq\operatorname{Sym}\bar{\mathsf{M}}_{S}^{\mathsf{gp}}. (3.1.1)

we write

PPSn=Symn𝖬¯S𝗀𝗉,\operatorname{PP}^{n}_{S}=\operatorname{Sym}^{n}\bar{\mathsf{M}}_{S}^{\mathsf{gp}}, (3.1.2)

for the graded pieces, and piecewise-linear functions are

PPS1=Sym1𝖬¯S𝗀𝗉=𝖬¯S𝗀𝗉.\operatorname{PP}^{1}_{S}=\operatorname{Sym}^{1}\bar{\mathsf{M}}_{S}^{\mathsf{gp}}=\bar{\mathsf{M}}_{S}^{\mathsf{gp}}. (3.1.3)
Remark 3.3.
  1. 1.

    The sheaf 𝖬¯S𝗀𝗉\bar{\mathsf{M}}_{S}^{\mathsf{gp}} makes sense on the big strict étale site of SS, so the same holds for the sheaf PPS\operatorname{PP}_{S}.

  2. 2.

    There is natural map Symn(𝖬¯S𝗀𝗉(S))PPSn(S)\operatorname{Sym}^{n}(\bar{\mathsf{M}}_{S}^{\mathsf{gp}}(S))\to\operatorname{PP}_{S}^{n}(S), but is in general not surjective unless n=1n=1, see example 3.4; this will play a prominent role in what follows.

  3. 3.

    Given a map of log algebraic stacks f:SSf\colon S^{\prime}\to S there is a natural map f𝖬¯S𝖬¯Sf^{*}\bar{\mathsf{M}}_{S}\to\bar{\mathsf{M}}_{S^{\prime}}, inducing a natural map of sheaves of {\mathbb{Z}}-algebras fPPSPPSf^{*}\operatorname{PP}_{S}\to\operatorname{PP}_{S^{\prime}}.

Example 3.4.

Let S¯=k2{\underline{S}}={\mathbb{P}}_{k}^{2}, and let EE be an irreducible nodal cubic in S¯{\underline{S}}, with complement i:US¯i\colon U\hookrightarrow{\underline{S}}. We define 𝖬S=i𝒪U\mathsf{M}_{S}=i_{*}{\mathcal{O}}_{U}, so that 𝖬¯S(S)=\bar{\mathsf{M}}_{S}(S)={\mathbb{N}}, and Sym(𝖬¯S𝗀𝗉(S))=[e]\operatorname{Sym}(\bar{\mathsf{M}}_{S}^{\mathsf{gp}}(S))={\mathbb{Z}}[e], where ee corresponds to the divisor EE. There is an étale chart for SS at the singular point of EE given by k[a,b]k[\left<a,b\right>] where aa, bb correspond to the two branches of EE through the singular point. The image of Sym2(𝖬¯S𝗀𝗉(S))\operatorname{Sym}^{2}(\bar{\mathsf{M}}_{S}^{\mathsf{gp}}(S)) is the free module (a+b)2{\mathbb{Z}}\left<(a+b)^{2}\right>. However, there is another global section of PPS2\operatorname{PP}^{2}_{S} given by abab, and in fact PPS2(S)=(a+b)2,ab=a2+b2,ab\operatorname{PP}^{2}_{S}(S)={\mathbb{Z}}\left<(a+b)^{2},ab\right>={\mathbb{Z}}\left<a^{2}+b^{2},ab\right>.

3.2 Simple log algebraic stacks

3.2.1 Barycentric subdivision

If SS is a regular log regular log algebraic stack then by [Niz06, 5.2] there exists a unique normal crossings divisor ZZ on SS (the boundary divisor of SS) with complement i:USi\colon U\to S and 𝖬S=i𝒪U\mathsf{M}_{S}=i_{*}{\mathcal{O}}_{U}. We write the irreducible components of ZZ as (Di)iI(D_{i})_{i\in I}.

If SS is a regular log regular atomic444In the sense of [AW18]: SS has a unique stratum that is closed and connected, and the restriction of the characteristic monoid to this stratum is a constant sheaf. log scheme then we define the barycentric ideal sheaf to be the product

JI(jJDj),\prod_{J\subseteq I}{\mathcal{I}}(\bigcap_{j\in J}D_{j}),

and the barycentric subdivision of SS to be the blowup of SS along the barycentric ideal sheaf. This blowup is stable under strict smooth pullback, defining a barycentric subdivision of any log regular log algebraic stack. A more explicit description can be found in [MPS21, §5.3]

3.2.2 Simple log algebraic stacks

Definition 3.5.

If SS is a regular log regular log algebraic stack with boundary divisor Z=iIDiZ=\bigcup_{i\in I}D_{i}, we say SS is simple if for every JIJ\subseteq I the fibre product

DJ×jJ,SDjD_{J}\coloneqq\bigtimes_{j\in J,S}D_{j} (3.2.1)

is regular and the natural map on sets of connected components π0(DJ)π0(S)\pi_{0}(D_{J})\to\pi_{0}(S) is injective. The closed connected substacks DJD_{J} are the closed strata of SS.

This condition is more restrictive than requiring the boundary divisor to be a strict normal crossings divisor; consider the union of a line and a smooth conic in 2{\mathbb{P}}^{2} meeting at two points, then the intersection is not connected.

3.2.3 Simplifying blowups

Lemma 3.6.

Let SS be a log regular log algebraic stack. Then there exists a log blowup S~S\tilde{S}\to S such that S~\tilde{S} is simple.

The proof consists of three observations:

  1. 1.

    By [IT14] there exists a functorial resolution algorithm for log regular log schemes, hence there exists a log blowup of SS which is regular and log regular;

  2. 2.

    If SS is regular log regular then the barycentric subdivision has strict normal crossings boundary (i.e. the substacks DJD_{J} of eq. 3.2.1 are regular);

  3. 3.

    If S is regular log regular with strict normal crossings boundary then the barycentric subdivision is simple.

3.2.4 Global generation on simple log schemes

In this section and the next we prove the key technical result on piecewise-polynomial functions on log schemes and stacks. The version for stacks implies that for schemes, but the proof is a little fiddly, so for expository reasons we treat the case of schemes first (the proofs in the two cases are similar).

If SS is a simple log stack with irreducible boundary divisors DbD_{b}, we write DSD_{S} for the sheafification of the presheaf on the small Zariski site SZarS_{Zar} associating to an open USU\hookrightarrow S the free abelian group on those DbD_{b} such that DbUD_{b}\cap U\neq\emptyset. On connected opens the values of the sheaf and the presheaf coincide.

Theorem 3.7.

Let SS be a quasi-compact simple log scheme, and let n0n\in{\mathbb{Z}}_{\geq 0}. Then the natural map of {\mathbb{Z}}-modules

Symn(𝖬¯S𝗀𝗉(S))(Symn𝖬¯S𝗀𝗉)(S)\operatorname{Sym}^{n}(\bar{\mathsf{M}}_{S}^{\mathsf{gp}}(S))\to(\operatorname{Sym}^{n}\bar{\mathsf{M}}_{S}^{\mathsf{gp}})(S) (3.2.2)

is surjective.

In other words, PPS(S)\operatorname{PP}_{S}(S) is just the symmetric algebra on 𝖬¯S(S)\bar{\mathsf{M}}_{S}(S); every global piecewise-polynomial function can be written globally as a polynomial in piecewise-linear functions.

Proof.

To simplify notation we assume SS connected (hence irreducible). For n=0n=0 and n=1n=1 the result is obvious. We write 𝕊{\mathbb{S}} for the constant sheaf on the {\mathbb{Z}}-algebra Symn(𝖬¯S𝗀𝗉(S))\operatorname{Sym}^{n}(\bar{\mathsf{M}}_{S}^{\mathsf{gp}}(S)). Restriction gives a natural surjection

φ:𝕊Sym(𝖬¯S𝗀𝗉)\varphi\colon{\mathbb{S}}\to\operatorname{Sym}(\bar{\mathsf{M}}_{S}^{\mathsf{gp}}) (3.2.3)

whose kernel we denote KK, a sheaf of sub-{\mathbb{Z}}-modules of 𝕊{\mathbb{S}}. Fixing an ordering on the index set BB of the DbD_{b}, we identify 𝕊{\mathbb{S}} with the constant sheaf on the free abelian group on monomials in the DbD_{b}. If USU\subseteq S is a connected (equivalently, non-empty) open then K(U)K(U) is the free abelian group on monomials aADa\prod_{a\in A}D_{a} such that aADaU=\cap_{a\in A}D_{a}\cap U=\emptyset.

To prove the theorem it suffices to show that H𝖾´𝗍1(S,K)=0H_{\mathsf{\acute{e}t}}^{1}(S,K)=0. For this we choose a finite Zariski cover 𝒰=(Ui)I{\mathcal{U}}=(U_{i})_{I} of SS by atomic log schemes (this exists because we assume SS simple). These UiU_{i} are connected and SS is irreducible, so all intersections UijU_{ij}, UijkU_{ijk}, … among UiU_{i} are also connected. The UiU_{i} are evidently acyclic for KK, so it is enough to prove vanishing of the Cech cohomology group H𝒰1(S,K)H^{1}_{\mathcal{U}}(S,K). We consider the relevant piece of the ordered Cech complex Cˇord\check{C}_{ord}^{\bullet}

iIK(Ui)d1i<jIK(Uij)d2i<j<kIK(Uijk).\prod_{i\in I}K(U_{i})\stackrel{{\scriptstyle d_{1}}}{{\longrightarrow}}\prod_{i<j\in I}K(U_{ij})\stackrel{{\scriptstyle d_{2}}}{{\longrightarrow}}\prod_{i<j<k\in I}K(U_{ijk}).

Since KK is a subsheaf of a constant sheaf of free abelian groups, the kernel of d2d_{2} is generated by elements ss where, for some triple of indices i0<j0<k0Ii_{0}<j_{0}<k_{0}\in I, we have

  • sij=0s_{ij}=0 unless (i,j)=(i0,j0)(i,j)=(i_{0},j_{0}) or (i,j)=(j0,k0)(i,j)=(j_{0},k_{0});

  • si0j0s_{i_{0}j_{0}} and sj0k0s_{j_{0}k_{0}} map to the same monomial in K(Ui0j0k0)K(U_{i_{0}j_{0}k_{0}}).

Choose such an element ss, where to simplify the notation we assume i0=1,i_{0}=1, j0=2,j_{0}=2, k0=3k_{0}=3. Again remembering that KK is a subsheaf of the constant sheaf 𝕊{\mathbb{S}} and that U123U_{123} is connected, we can assume that s12s_{12} and s2,3s_{2,3} are both given by the monomial aADa\prod_{a\in A}D_{a}. Then necessarily

aADaU12==aADaU23\bigcap_{a\in A}D_{a}\cap U_{12}=\emptyset=\bigcap_{a\in A}D_{a}\cap U_{23}

and

aADaUij unless (i,j)=(1,2) or (i,j)=(2,3).\bigcap_{a\in A}D_{a}\cap U_{ij}\neq\emptyset\text{ unless }(i,j)=(1,2)\text{ or }(i,j)=(2,3). (3.2.4)

Further, since aADa\bigcap_{a\in A}D_{a} is irreducible555If it is empty there is nothing to check. we see that at least one of

aADaU1 and aADaU2\bigcap_{a\in A}D_{a}\cap U_{1}\text{ and }\bigcap_{a\in A}D_{a}\cap U_{2} (3.2.5)

is empty, and at least one of

aADaU2 and aADaU3\bigcap_{a\in A}D_{a}\cap U_{2}\text{ and }\bigcap_{a\in A}D_{a}\cap U_{3} (3.2.6)

is empty. Hence if aADaU2\bigcap_{a\in A}D_{a}\cap U_{2}\neq\emptyset then both

aADaU1 and aADaU3\bigcap_{a\in A}D_{a}\cap U_{1}\text{ and }\bigcap_{a\in A}D_{a}\cap U_{3} (3.2.7)

are empty, hence aADaU13=\bigcap_{a\in A}D_{a}\cap U_{13}=\emptyset, a contradiction. We see that the element siIK(Ui)s^{\prime}\in\prod_{i\in I}K(U_{i}) given by s2=aADas^{\prime}_{2}=\prod_{a\in A}D_{a} and si=0s^{\prime}_{i}=0 for i2i\neq 2 has d1(s)=sd_{1}(s^{\prime})=s as required. ∎

3.2.5 Global generation on simple log algebraic stacks

We now prove the analogous result for log algebraic stacks. We strongly encourage the reader to skip the proof; it is almost identical to that for schemes, except that we have to work with smooth covers.

Theorem 3.8.

Let SS be a quasi-compact simple log algebraic stack, and let n0n\in{\mathbb{Z}}_{\geq 0}. Then the natural map of {\mathbb{Z}}-modules

Symn(𝖬¯S𝗀𝗉(S))(Symn𝖬¯S𝗀𝗉)(S)\operatorname{Sym}^{n}(\bar{\mathsf{M}}_{S}^{\mathsf{gp}}(S))\to(\operatorname{Sym}^{n}\bar{\mathsf{M}}_{S}^{\mathsf{gp}})(S) (3.2.8)

is surjective.

Proof.

As in the proof of theorem 3.7 we assume SS connected, we write 𝕊{\mathbb{S}} for the constant sheaf on the {\mathbb{Z}}-algebra Symn(𝖬¯S𝗀𝗉(S))\operatorname{Sym}^{n}(\bar{\mathsf{M}}_{S}^{\mathsf{gp}}(S)), and

φ:𝕊Sym(𝖬¯S𝗀𝗉)\varphi\colon{\mathbb{S}}\to\operatorname{Sym}(\bar{\mathsf{M}}_{S}^{\mathsf{gp}}) (3.2.9)

for the natural surjection whose kernel we denote KK, a sheaf of sub-{\mathbb{Z}}-modules of 𝕊{\mathbb{S}}. Fixing an ordering on the index set BB of the DbD_{b}, we identify 𝕊{\mathbb{S}} with the constant sheaf on the free abelian group on monomials in the DbD_{b}. If f:USf\colon U\to S is a smooth map from a connected scheme then K(U)K(U) is the free abelian group on monomials ADa\prod_{A}D_{a} such that ADaf(U)=\cap_{A}D_{a}\cap f(U)=\emptyset.

To prove the theorem it suffices to show that Hsm1(S,K)=0H_{sm}^{1}(S,K)=0. For this we first choose a strict smooth map f:SSf\colon S^{\prime}\to S from a log scheme. Shrinking SS^{\prime}, we may assume that the fibre of SS^{\prime} over the generic point of any stratum of SS is connected. This condition has two crucial consequences:

  1. 1.

    SS^{\prime} is simple;

  2. 2.

    If (Ui)iI(U_{i})_{i\in I} is any collection of opens of SS^{\prime}, and ABA\subseteq B any subset, then

    aADaiIf(Ui)\bigcap_{a\in A}D_{a}\cap\bigcap_{i\in I}f(U_{i})\neq\emptyset (3.2.10)

    if and only if

    aADaf(iIUi).\bigcap_{a\in A}D_{a}\cap f\left(\bigcap_{i\in I}U_{i}\right)\neq\emptyset. (3.2.11)

We choose a Zariski cover of SS^{\prime} by atomic schemes, inducing a smooth cover 𝒰=(Ui)I{\mathcal{U}}=(U_{i})_{I} of SS by atomic log schemes. This cover has the property that the fibre product of any number of the UiU_{i} over SS is connected.

The proof now proceeds exactly as in the case when SS was a scheme (theorem 3.7), though we write the details for completeness. The UiU_{i} are acyclic for KK, so it is enough to prove vanishing of the Cech cohomology group H𝒰1(S,K)H^{1}_{\mathcal{U}}(S,K). The relevant piece of the ordered Cech complex Cˇord\check{C}_{ord}^{\bullet} is

iIK(Ui)d1i<jIK(Uij)d2i<j<kIK(Uijk).\prod_{i\in I}K(U_{i})\stackrel{{\scriptstyle d_{1}}}{{\longrightarrow}}\prod_{i<j\in I}K(U_{ij})\stackrel{{\scriptstyle d_{2}}}{{\longrightarrow}}\prod_{i<j<k\in I}K(U_{ijk}).

Since KK is a subsheaf of a constant sheaf of free abelian groups, the kernel of d2d_{2} is generated by elements ss where, for some triple of indices i0<j0<k0Ii_{0}<j_{0}<k_{0}\in I, we have

  • sij=0s_{ij}=0 unless (i,j)=(i0,j0)(i,j)=(i_{0},j_{0}) or (i,j)=(j0,k0)(i,j)=(j_{0},k_{0});

  • si0j0s_{i_{0}j_{0}} and sj0k0s_{j_{0}k_{0}} map to the same monomial in K(Ui0j0k0)K(U_{i_{0}j_{0}k_{0}}).

Choose such an element ss, where to simplify the notation we assume i0=1,i_{0}=1, j0=2,j_{0}=2, k0=3k_{0}=3. Suppose that s12s_{12} and s2,3s_{2,3} are given by the monomial aADa\prod_{a\in A}D_{a}, so necessarily

aADaf(U12)==aADaf(U23)\bigcap_{a\in A}D_{a}\cap f(U_{12})=\emptyset=\bigcap_{a\in A}D_{a}\cap f(U_{23})

and

aADAf(Uij) unless (i,j)=(1,2) or (i,j)=(2,3)\bigcap_{a\in A}D_{A}\cap f(U_{ij})\neq\emptyset\text{ unless }(i,j)=(1,2)\text{ or }(i,j)=(2,3) (3.2.12)

by injectivity of the restriction maps. Then we apply property (2) above to see that at least one of

aADaf(U1) and aADaf(U2)\bigcap_{a\in A}D_{a}\cap f(U_{1})\text{ and }\bigcap_{a\in A}D_{a}\cap f(U_{2}) (3.2.13)

is empty, and at least one of

aADaf(U2) and aADaf(U3)\bigcap_{a\in A}D_{a}\cap f(U_{2})\text{ and }\bigcap_{a\in A}D_{a}\cap f(U_{3}) (3.2.14)

is empty. Hence if aADaf(U2)\bigcap_{a\in A}D_{a}\cap f(U_{2})\neq\emptyset then both

aADaf(U1) and aADaf(U3)\bigcap_{a\in A}D_{a}\cap f(U_{1})\text{ and }\bigcap_{a\in A}D_{a}\cap f(U_{3}) (3.2.15)

are empty, hence aADaf(U13)=\bigcap_{a\in A}D_{a}\cap f(U_{13})=\emptyset, a contradiction. We see that the element siIK(Ui)s^{\prime}\in\prod_{i\in I}K(U_{i}) given by s2=aADas^{\prime}_{2}=\prod_{a\in A}D_{a} and si=0s^{\prime}_{i}=0 for i2i\neq 2 has d1(s)=sd_{1}(s^{\prime})=s as required. ∎

3.3 Map to the Chow group

3.3.1 Map on divisors

For an algebraic stack XX, we write Div(X)\operatorname{Div}(X) for the monoid of isomorphism classes of pairs (,)({\mathcal{L}},\ell) where {\mathcal{L}} is a line bundle on XX and (X)\ell\in{\mathcal{L}}(X) a section, with monoid operation given by tensor product.

Let SS be a log algebraic stack and m𝖬¯S(S)m\in\bar{\mathsf{M}}_{S}(S). The preimage 𝒪S(m)×{\mathcal{O}}_{S}(-m)^{\times} of mm in 𝖬S\mathsf{M}_{S} is an 𝒪S×{\mathcal{O}}_{S}^{\times}-torsor and the log structure equips it with a map to 𝒪S(m)×𝒪S{\mathcal{O}}_{S}(-m)^{\times}\to{\mathcal{O}}_{S}. This map admits a unique 𝒪S×{\mathcal{O}}_{S}^{\times}-equivariant extension to a map of line bundles 𝒪S(m)𝒪S{\mathcal{O}}_{S}(-m)\to{\mathcal{O}}_{S}, where we built 𝒪S(m){\mathcal{O}}_{S}(-m) from 𝒪S(m)×{\mathcal{O}}_{S}(-m)^{\times} by filling in the zero section. Dualising gives a map 𝒪S𝒪S(m)𝒪S(m){\mathcal{O}}_{S}\to{\mathcal{O}}_{S}(m)\coloneqq{\mathcal{O}}_{S}(-m)^{\vee}, and the image of the section 11 of 𝒪S{\mathcal{O}}_{S} gives a section of 𝒪S(m){\mathcal{O}}_{S}(m). This defines a map

𝒪S():𝖬¯S(S)Div(S).{\mathcal{O}}_{S}(-)\colon\bar{\mathsf{M}}_{S}(S)\to\operatorname{Div}(S).

This can be upgraded to a monoidal functor of fibred symmetric monoidal categories, see [BV12, §3.1]. Taking the (operational) first chern class yields a group homomorphism

Φ1:𝖬¯S𝗀𝗉(S)CH1(S),\Phi^{1}\colon\bar{\mathsf{M}}_{S}^{\mathsf{gp}}(S)\to\operatorname{CH}^{1}(S), (3.3.1)

with image contained in the subgroup generated by Cartier divisors.

3.3.2 The case of simple finite-type stacks

Let SS be a simple log algebraic stack, smooth666If kk is a field of characteristic zero then being smooth is here equivalent to being locally of finite type (since simple implies regular). over kk. Since SS is regular, the intersection pairing equips the Chow group CH(S)\operatorname{CH}(S) with a commutative ring structure. As such, the map

Φ1:𝖬¯S(S)CH1(S)\Phi^{1}\colon\bar{\mathsf{M}}_{S}(S)\to\operatorname{CH}^{1}(S) (3.3.2)

of eq. 3.3.1 extends uniquely to a ring homomorphism

Φ:Sym(𝖬¯S(S))CH(S).\Phi^{\prime}\colon\operatorname{Sym}(\bar{\mathsf{M}}_{S}(S))\to\operatorname{CH}(S). (3.3.3)

Since

Symn(𝖬¯S𝗀𝗉(S))(Symn𝖬¯S𝗀𝗉)(S)\operatorname{Sym}^{n}(\bar{\mathsf{M}}_{S}^{\mathsf{gp}}(S))\to(\operatorname{Sym}^{n}\bar{\mathsf{M}}_{S}^{\mathsf{gp}})(S) (3.3.4)

is surjective, and any element of the kernel evidently maps to 0 in CH(S)\operatorname{CH}(S), this map Φ\Phi^{\prime} descends to a unique ring homomorphism

Φ:(Sym𝖬¯S)(S)=PPS(S)CH(S),\Phi\colon(\operatorname{Sym}\bar{\mathsf{M}}_{S})(S)=\operatorname{PP}_{S}(S)\to\operatorname{CH}(S), (3.3.5)

whose degree 1 part is Φ1\Phi^{1}.

3.3.3 The case of log smooth finite-type stacks

Let SS be a quasi-compact log smooth log algebraic stack over kk. By lemma 3.6 there exists a log blowup π:S~S\pi\colon\tilde{S}\to S with S~\tilde{S} simple. We define

ΦS:(Sym𝖬¯S)(S)=PPS(S)CH(S)\Phi_{S}\colon(\operatorname{Sym}\bar{\mathsf{M}}_{S})(S)=\operatorname{PP}_{S}(S)\to\operatorname{CH}(S) (3.3.6)

as the composite

(Sym𝖬¯S)(S)Sym𝖬¯S~(S~)ΦS~CH(S~)πCH(S).(\operatorname{Sym}\bar{\mathsf{M}}_{S})(S)\to\operatorname{Sym}\bar{\mathsf{M}}_{\tilde{S}}(\tilde{S})\stackrel{{\scriptstyle\Phi_{\tilde{S}}}}{{\longrightarrow}}\operatorname{CH}(\tilde{S})\stackrel{{\scriptstyle\pi_{*}}}{{\longrightarrow}}\operatorname{CH}(S). (3.3.7)

Lemma 3.6 actually yields a canonical choice of log blowup π\pi, but we should still check that the map ΦS\Phi_{S} is independent of the choice of π\pi (for example, if SS was already simple, we don’t want to have changed the map by blowing up).

Lemma 3.9.

Let π:S~S\pi\colon\tilde{S}\to S be a log blowup with SS and S~\tilde{S} simple. The diagram

PPS~(S~){\operatorname{PP}_{\tilde{S}}(\tilde{S})}CH(S~){\operatorname{CH}(\tilde{S})}PPS(S){\operatorname{PP}_{S}(S)}CH(S){\operatorname{CH}(S)}ΦS~\scriptstyle{\Phi_{\tilde{S}}}π\scriptstyle{\pi_{*}}π\scriptstyle{\pi^{*}}ΦS\scriptstyle{\Phi_{S}} (3.3.8)

commutes.

Proof.

Since SS is simple it is enough to check this for a monomial in elements of 𝖬¯S(S)\bar{\mathsf{M}}_{S}(S) corresponding to prime boundary divisors on SS, say aADa\prod_{a\in A}D_{a}. Applying π\pi^{*} corresponds to taking the total transforms of these divisors up to S~\tilde{S}. We then need to show that

π(aπDa)=aDa,\pi_{*}(\prod_{a}\pi^{*}D_{a})=\prod_{a}D_{a}, (3.3.9)

which follows from the projection formula and the fact that ππ\pi_{*}\pi^{*} is the identity. ∎

Lemma 3.10.

For any log regular SS, the map ΦS\Phi_{S} is independent of the choice of log blowup π:S~S\pi\colon\tilde{S}\to S.

Proof.

Reduce to one blowup dominating another, then apply lemma 3.9. ∎

Example 3.11.

We resume example 3.4, and recall that PP2(S)=(a+b)2,ab\operatorname{PP}^{2}(S)={\mathbb{Z}}\left<(a+b)^{2},ab\right>. Then (a+b)2(a+b)^{2} maps to E2CH2(S)E^{2}\in\operatorname{CH}^{2}(S), and abab maps to the class of the singular point of EE in CH2(S)\operatorname{CH}^{2}(S).

3.3.4 The case of log regular stacks locally of finite type

Let S¯{\underline{S}} be an algebraic stack locally of finite type over kk, and write qOp(S¯)\operatorname{qOp}({\underline{S}}) for the category of open substacks of S¯{\underline{S}} which are quasi-compact over kk, with maps over S¯{\underline{S}}. Then one sees easily that

CH(S¯)=limUqOp(S¯)CH(U).\operatorname{CH}({\underline{S}})=\operatorname{lim}_{U\in\operatorname{qOp}({\underline{S}})}\operatorname{CH}(U). (3.3.10)
Lemma 3.12.

Let i:S1S2i\colon S_{1}\hookrightarrow S_{2} be a strict open immersion of quasi-compact log smooth log algebraic stacks over kk. The diagram

PPS2(S2){\operatorname{PP}_{S_{2}}(S_{2})}CH(S2){\operatorname{CH}(S_{2})}PPS1(S1){\operatorname{PP}_{S_{1}}(S_{1})}CH(S1){\operatorname{CH}(S_{1})}Φ2\scriptstyle{\Phi_{2}}i\scriptstyle{i^{*}}i\scriptstyle{i^{*}}Φ1\scriptstyle{\Phi_{1}} (3.3.11)

commutes.

Proof.

A simplifying blowup for S2S_{2} pulls back to one for S1S_{1}, so we may assume both SiS_{i} simple. Then it is enough to check the result for divisors (since both maps ii^{*} are ring homomorphisms), but this is easy. ∎

Now let SS be a log smooth log algebraic stack over kk. Given pPPS(S)p\in\operatorname{PP}_{S}(S) and any USU\hookrightarrow S quasi-compact, we restrict pp to pUPPU(U)p_{U}\in\operatorname{PP}_{U}(U), yielding an element ΦU(pU)CH(U)\Phi_{U}(p_{U})\in\operatorname{CH}(U). By lemma 3.12 these glue, yielding a map

ΦS:PPS(S)CH(S).\Phi_{S}\colon\operatorname{PP}_{S}(S)\to\operatorname{CH}(S). (3.3.12)

3.4 Subdivided piecewise-polynomials and the log-tautological ring

For a log algebraic stack SS we define the group of subdivided piecewise-polynomial functions as

sPP(S)=colimS~SPP(S~),\operatorname{sPP}^{\prime}(S)=\operatorname{colim}_{\tilde{S}\to S}\operatorname{PP}(\tilde{S}), (3.4.1)

where S~S\tilde{S}\to S runs over all log blowups of SS.

Lemma 3.13.

The pullback PP(S)PP(S~)\operatorname{PP}(S)\to\operatorname{PP}(\tilde{S}) is injective for S~S\tilde{S}\to S any log blowup, so the natural maps to the colimit are injective.

Proof.

It suffices to show this locally, so we reduce to the atomic case. It is then enough to check that the natural map 𝖬¯S(S)𝖬¯S~(S~)\bar{\mathsf{M}}_{S}(S)\to\bar{\mathsf{M}}_{\tilde{S}}(\tilde{S}) is injective. This is clear from the construction of the blowup in the toric case, but any log blowup is locally a strict base-change of a toric blowup. ∎

We define the sheaf of subdivided piecewise-polynomials sPP\operatorname{sPP} on the small strict étale site of SS as the sheafification of the presheaf of rings sPP:UsPP(U)\operatorname{sPP}^{\prime}\colon U\mapsto\operatorname{sPP}^{\prime}(U). This sheaf property then immediately yields

sPP(S)=limUqOp(S)sPP(U).\operatorname{sPP}(S)=\operatorname{lim}_{U\in\operatorname{qOp}(S)}\operatorname{sPP}(U). (3.4.2)

The natural maps

Φi𝗅𝗈𝗀:sPP(Ui)colimU~iCH(U~i)\Phi_{i}^{\mathsf{log}}\colon\operatorname{sPP}(U_{i})\to\operatorname{colim}_{\tilde{U}_{i}}\operatorname{CH}(\tilde{U}_{i}) (3.4.3)

then assemble into a ring homomorphism

Φ𝗅𝗈𝗀:sPP(S)LogCH(S).\Phi^{\mathsf{log}}\colon\operatorname{sPP}(S)\to\operatorname{LogCH}(S). (3.4.4)
Remark 3.14.

The presheaf sPP\operatorname{sPP}^{\prime} is always separated, and is a sheaf if SS is quasi-compact and quasi-separated; we make the above construction to avoid having to worry about finding common refinements of blowups of very large stacks.

3.4.1 The log-tautological ring

Definition 3.15.

Let SS be a smooth log smooth log algebraic stack over kk and let TCH(S)T\subseteq\operatorname{CH}(S) be a subring. We define T𝗅𝗈𝗀LogCH(S)T^{\mathsf{log}}\subseteq\operatorname{LogCH}(S) to be the sub-TT-algebra of LogCH(S)\operatorname{LogCH}(S) generated by the image of

Φ𝗅𝗈𝗀:sPP(S)LogCH(S).\Phi^{\mathsf{log}}\colon\operatorname{sPP}(S)\to\operatorname{LogCH}(S). (3.4.5)

A natural application is to take S=¯g,nS=\overline{{\mathcal{M}}}_{g,n} and TT to be the usual tautological subring of the Chow ring. We want to ensure that after carrying out our logarithmic constructions and pushing back down to ¯g,n\overline{{\mathcal{M}}}_{g,n} we still have tautological classes.

Definition 3.16.

We say a subring TCH(S)T\subseteq\operatorname{CH}(S) is tectonic777It contains many strata, which are formed by overlaps of other strata, perhaps after some things blow up… if the pushforward of T𝗅𝗈𝗀T^{\mathsf{log}} from LogCH(S)\operatorname{LogCH}(S) to CH(S)\operatorname{CH}(S) is equal to TT.

Giving criteria for when a subring TCH(S)T\subseteq\operatorname{CH}(S) is tectonic is somewhat subtle. Certainly if a subring is tectonic then it must contain all boundary strata, and the converse holds if SS is simple, but not in general. Fortunately for us a precise criterion has been worked out in [MPS21] for the case where SS is Deligne-Mumford and quasi-compact, which will be enough for our applications888It seems likely that their results (perhaps with slight modification) will also hold in the setting of smooth log smooth algebraic stacks, but we have not verified the details. . Their criterion goes by way of defining certain normally decorated strata class in CH(S)\operatorname{CH}(S); the definition is somewhat lengthy, and the details will not be so important for us. We need only the following lemma, and the fact that the usual tautological ring of ¯g,n\overline{{\mathcal{M}}}_{g,n} contains these normally decorated strata classes.

Lemma 3.17.

Suppose SS is Deligne-Mumford and quasi-compact. Then a subring TCH(S)T\subseteq\operatorname{CH}(S) is tectonic if and only if it contains all normally decorated strata class.

Proof.

Let ν:S~S\nu\colon\tilde{S}\to S be a log blowup. We denote by R(S)R^{\star}(S) the ring of normally decorated strata classes on SS, and similarly for S~\tilde{S}. By definition R(S~)R^{\star}(\tilde{S}) contains999If we take S~\tilde{S} simple then R(S~)R^{\star}(\tilde{S}) is in fact generated by strata, but this is not needed for our argument. all boundary strata of S~\tilde{S}, and ν(R(S~))R(S)\nu_{*}(R^{\star}(\tilde{S}))\subseteq R^{\star}(S) by [MPS21, Theorem 13]. ∎

3.5 Constructing a class

Suppose we are given the following data:

  1. 1.

    a quasi-separated log smooth log algebraic stack S/kS/k which is stratified by global quotients101010In practise this last condition means we must be exclude 𝔐1,0\mathfrak{M}_{1,0} from our results; but this is fairly harmless since we can just consider the corresponding cycle on 𝔐1,1\mathfrak{M}_{1,1} with zero weighting on the new marking. ;

  2. 2.

    f:XSf\colon X\to S a birational representable log étale morphism;

  3. 3.

    𝖩/k\mathsf{J}/k an algebraic stack and i:e𝖩i\colon e\rightarrowtail\mathsf{J} a regularly embedded closed substack (or more generally an lci morphism);

  4. 4.

    σ:X𝖩\sigma\colon X\to\mathsf{J} a morphism over kk.

Suppose also that the base-change X×𝖩eX\times_{\mathsf{J}}e is proper over SS. Then we construct a class [σe]f,𝗅𝗈𝗀LogCH(S)[\sigma^{*}e]_{f,\mathsf{log}}\in\operatorname{LogCH}(S) — we often omit the ff from the notation when it is clear from context.

The simplest case of our construction is when XX is smooth and f:XSf\colon X\to S is a log blowup (then X×𝖩eSX\times_{\mathsf{J}}e\to S is automatically proper). Then σ!e\sigma^{!}e is a well-defined class on XX, and automatically gives an element of LogCH(S)\operatorname{LogCH}(S), which we denote [σe]f,𝗅𝗈𝗀[\sigma^{*}e]_{f,\mathsf{log}}.

In the general case a little more care is needed. Because LogCH(S)\operatorname{LogCH}(S) is defined as a limit over quasi-compact opens, we may assume that SS is quasi-compact. Then by lemma 3.21 there exist log blowups S~S\tilde{S}\to S and X~X\tilde{X}\to X and a strict open immersion f~:X~S~\tilde{f}\colon\tilde{X}\hookrightarrow\tilde{S} over ff; after further log blowup we may also assume S~\tilde{S} (and hence X~\tilde{X}) to be smooth.

Definition 3.18.

We call such an S~S\tilde{S}\to S a sufficiently fine log blowup (for f:XSf\colon X\to S), and X~S~\tilde{X}\hookrightarrow\tilde{S} the lift of XX.

Set Z=X~×σ,𝖩,ieZ=\tilde{X}\times_{\sigma,\mathsf{J},i}e, and consider the diagram

Z{Z}e{e}X~{\tilde{X}}𝖩{\mathsf{J}}S~{\tilde{S}}j\scriptstyle{j}i\scriptstyle{i}σ\scriptstyle{\sigma}f~\scriptstyle{\tilde{f}} (3.5.1)

We then define

[σe]f,𝗅𝗈𝗀=ji![X~].[\sigma^{*}e]_{f,\mathsf{log}}=j_{*}i^{!}[\tilde{X}]. (3.5.2)

To unravel this formula, recall that ii is lci so we have a gysin morphism i!:A(X~)AZi^{!}\colon A^{*}(\tilde{X})\to A^{*}Z. The composite j:ZS~j\colon Z\to\tilde{S} is a closed immersion, in particular projective, so we have a pushforward j:AZAS~j_{*}\colon A^{*}Z\to A^{*}\tilde{S}. Finally, S~\tilde{S} is smooth, so the intersection product furnishes a map A(S)CH(S~)A^{*}(S)\to\operatorname{CH}(\tilde{S}), and we have a natural inclusion CH(S~)LogCH(S)\operatorname{CH}(\tilde{S})\hookrightarrow\operatorname{LogCH}(S).

To see that the above construction is independent of the choice of S~\tilde{S} we use that gysin pullbacks along lci maps commute with each other and with projective pushforward [Kre99, Theorem 2.1.12 (xi)]

Remark 3.19.

Let S~S\tilde{S}\to S be a sufficiently fine log blowup. Then [σe]f,𝗅𝗈𝗀[\sigma^{*}e]_{f,\mathsf{log}} is determined on S~\tilde{S} (in the sense of definition 2.5).

Lemma 3.20.

Let φ:XX\varphi\colon X^{\prime}\to X be another birational representable log étale morphism, and write f:XSf^{\prime}\colon X^{\prime}\to S, σ:X𝖩\sigma^{\prime}\colon X^{\prime}\to\mathsf{J} for the obvious composites. Then

[σe]f,𝗅𝗈𝗀=[σe]f,𝗅𝗈𝗀LogCH(S).[\sigma^{*}e]_{f,\mathsf{log}}=[\sigma^{\prime*}e]_{f^{\prime},\mathsf{log}}\in\operatorname{LogCH}(S). (3.5.3)
Proof.

Let TST\to S be a blowup that is sufficiently fine for XSX^{\prime}\to S, and which dominates a sufficiently fine blowup for XSX\to S. Unravelling the definitions one sees that the classes agree already in CH(T)\operatorname{CH}(T). ∎

The key to the above construction is the following lemma, whose proof is essentially the same as that of Lemma 6.1 of [Hol19a].

Lemma 3.21.

Let SS be a regular log regular qcqs stack and f:XSf\colon X\to S birational separated log étale representable. Then there exist log blowups S~S\tilde{S}\to S and X~X\tilde{X}\to X and a strict open immersion X~S~\tilde{X}\to\tilde{S} over SS.

Probably this lemma is false if one drops either the quasi-compactness or quasi-separatedness assumptions, but we have not managed to write down an example, and would be interested to see one.

Proof.

Consider first the case where SS is an affine toric variety, given by some cone crc\stackrel{{\scriptstyle\sim}}{{\longrightarrow}}{\mathbb{N}}^{r}. Then XX is given by a fan FF consisting of a collection of cones contained in cc. Let F¯\bar{F} be a complete fan in cc such that every cone in FF is a union of cones in cc; after further refinement of F¯\bar{F} we can assume that it corresponds to a log blowup S~S\tilde{S}\to S. The restriction of F¯\bar{F} to FF gives a log blowup X~X\tilde{X}\to X, and a strict open immersion X~S~\tilde{X}\to\tilde{S}.

In the case where SS is an atomic log scheme we can follow essentially the same procedure, where the cone crc\stackrel{{\scriptstyle\sim}}{{\longrightarrow}}{\mathbb{N}}^{r} is replaced by the stalk of the ghost sheaf over the closed stratum of SS.

In the general case we can find a smooth cover SS by finitely many atomic patches (by quasi-compactness), and each intersection can be covered by finitely many atomic patches (by quasi-separatedness). A strict map of atomic patches just corresponds to some inclusion of a cone as a face of another cone: rs{\mathbb{N}}^{r}\hookrightarrow{\mathbb{N}}^{s}.

Given a face inclusion rs{\mathbb{N}}^{r}\hookrightarrow{\mathbb{N}}^{s} and a subdivision of s{\mathbb{N}}^{s}, we can pull the subdivision back to a (unique) subdivision of r{\mathbb{N}}^{r}. But also, given a subdivision of r{\mathbb{N}}^{r} we can turn it into a subdivision of s{\mathbb{N}}^{s} in a canonical way, by taking the product. Hence if rs{\mathbb{N}}^{r}\hookrightarrow{\mathbb{N}}^{s} is a face map (where we allow r=sr=s) and we have subdivisions of r{\mathbb{N}}^{r} and s{\mathbb{N}}^{s}, then we can find ‘common refinements’ to subdivisions of r{\mathbb{N}}^{r} and s{\mathbb{N}}^{s} which agree along the face map. Moreover if both starting subdivisions were log blowups then so are these common refinements.

To conclude the proof, we just need to extend this ‘common refinement’ procedure from a single face map to any diagram DD of face maps with finitely many objects. Such a diagram necessarily also has finitely many morphisms (since there are only finitely many face maps between any two cones), hence the same is true for the category DD^{\prime} obtained by formally inverting all the maps in DD. By the discussion in the previous paragraph we can pull back a log blowup along any map in DD^{\prime}.

We are given a log blowup of each cone in DD. For a fixed cone cc there are only finitely many pairs d,fd,f where dd is another cone and f:dcf\colon d\to c is a morphism in DD^{\prime}. We then give cc the log blowup which is the superposition over all these pairs (f,d)(f,d) of the pullback along ff of the given log blowup of dd. In this way we equip every object of DD with a log blowup, in such a way that these are compatible along all the face maps in DD. These then glue to a global log blowup of SS, which pulls back to a global log blowup of XX. ∎

4 Logarithmic double ramification cycles

4.1 Notation

Here we introduce notation needed for applying the machinery developed in the previous two sections to moduli of curves and to double ramification cycles.

  1. 1.

    𝔐g,n\mathfrak{M}_{g,n} denotes the (smooth, algebraic) stack of prestable curves of genus gg with nn ordered disjoint smooth markings. This has a normal crossings boundary, inducing a log smooth log structure. Equivalently, this is the stack of log curves of genus gg and nn markings, with a choice of total ordering on the markings (see [Kat00], [GS13, Appendix A]; the underlying algebraic stack is then given by the machinery of minimal log structures [Gil12]).

  2. 2.

    ¯g,n\overline{{\mathcal{M}}}_{g,n} is the open substack of 𝔐g,n\mathfrak{M}_{g,n} consisting of Deligne-Mumford-Knudsen stable curves.

  3. 3.

    𝔐=g,n𝔐g,n\mathfrak{M}=\bigsqcup_{g,n}\mathfrak{M}_{g,n} denotes the stack of all log curves with a choice of total ordering on their markings. Often the genus and markings will not be so important to us, so we can use this more compact notation.

  4. 4.

    CC is the universal curve over 𝔐\mathfrak{M}. We will abusively use the same notation for the tautological curve over any stack over 𝔐\mathfrak{M} (so for example, for the universal curve over ¯g,n\overline{{\mathcal{M}}}_{g,n}).

  5. 5.

    𝔓𝔦𝔠\mathfrak{Pic} is the relative Picard stack of CC over 𝔐\mathfrak{M}; objects are pairs of a curve and a line bundle on the curve. This is smooth over 𝔐\mathfrak{M} with relative inertia 𝔾m{\mathbb{G}}_{m}; we equip it with the strict (pullback) log structure over 𝔐\mathfrak{M}.

  6. 6.

    𝔍𝔞𝔠\mathfrak{Jac} denotes the connected component of 𝔓𝔦𝔠\mathfrak{Pic} corresponding to line bundles of (total) degree 0 on every fibre.

  7. 7.

    𝖩\mathsf{J} is the relative coarse moduli space over 𝔐\mathfrak{M} of the fibrewise connected component of identity in 𝔍𝔞𝔠\mathfrak{Jac} (or equivalently in 𝔓𝔦𝔠)\mathfrak{Pic}). Over the locus of smooth curves in 𝔐\mathfrak{M} this is an abelian variety, the classical jacobian. In general it is a semiabelian variety over 𝔐\mathfrak{M} which parametrises isomorphism classes of line bundles on /𝔐\mathfrak{C}/\mathfrak{M} which have degree 0 on every irreducible component of every geometric fibre (sometimes we refer to this condition as having multidegree 0¯{\underline{0}}). The morphism 𝖩𝔐\mathsf{J}\to\mathfrak{M} is separated, quasi-compact, and relatively representable by algebraic spaces (none of which hold for 𝔓𝔦𝔠\mathfrak{Pic} or 𝔍𝔞𝔠\mathfrak{Jac}).

  8. 8.

    𝖩¯\bar{\mathsf{J}} is the relative coarse moduli space of 𝔍𝔞𝔠\mathfrak{Jac} over 𝔐\mathfrak{M}; it can be defined analogously to 𝖩\mathsf{J} except that we require total degree 0 instead of multidegree 0¯{\underline{0}}. In particular we have an open immersion 𝖩𝖩¯\mathsf{J}\hookrightarrow\bar{\mathsf{J}}, which is an isomorphism over the locus of irreducible curves.

Lemma 4.1.

𝔍𝔞𝔠{\mathfrak{Jac}} is quasi-separated.

Proof.

First we check that 𝔐\mathfrak{M} is quasi-separated; equivalently, that the diagonal is qcqs. In other words, if C/SC/S is a prestable curve, then IsomS(C)\operatorname{Isom}_{S}(C) is qcqs over SS - but this is well-known.

Now we show that 𝔍𝔞𝔠{\mathfrak{Jac}} is quasi-separated over 𝔐\mathfrak{M}. In other words, we fix a prestable curve C/SC/S and a line bundle {\mathcal{L}} on CC, and look at the automorphisms of {\mathcal{L}} over CC; but this is just 𝔾m{\mathbb{G}}_{m}. ∎

4.1.1 Piecewise linear functions

If C/SC/S is a log curve and α𝖬¯C𝗀𝗉(C)\alpha\in\bar{\mathsf{M}}_{C}^{\mathsf{gp}}(C), the outgoing slope at a marked section cc of C/SC/S is the image of α\alpha in the stalk of the relative characteristic monoid 𝖬¯C/S,s=\bar{\mathsf{M}}_{C/S,s}={\mathbb{N}}.

Definition 4.2.

A piecewise-linear (PL) function on a log curve C/SC/S is an element α𝖬¯C𝗀𝗉(C)\alpha\in\bar{\mathsf{M}}_{C}^{\mathsf{gp}}(C) (cf. eq. 3.1.3) with all outgoing slopes vanishing111111It would be cleaner to work with vertical (‘unmarked’) log curves, but we will make use of smooth sections of C/SC/S in other places in our arguments, so we do not wish to impose verticality. .

The preimage of α\alpha in the exact sequence

1𝒪C×𝖬C𝗀𝗉𝖬¯C𝗀𝗉11\to{\mathcal{O}}_{C}^{\times}\to\mathsf{M}_{C}^{\mathsf{gp}}\to\bar{\mathsf{M}}_{C}^{\mathsf{gp}}\to 1 (4.1.1)

defines an associated 𝔾m{\mathbb{G}}_{m}-torsor 𝒪×(α){\mathcal{O}}^{\times}(\alpha), which we compactify to a line bundle 𝒪(α){\mathcal{O}}(\alpha) by glueing in the \infty section (this is just a choice of sign; it corresponds to 𝒪(p){\mathcal{O}}(-p) being an ideal sheaf, rather than its dual).

The bundle 𝒪(α){\mathcal{O}}(\alpha) always has total degree zero, but rarely multidegree 0¯{\underline{0}}; more precisely, it has multidegree 0¯{\underline{0}} if and only if 𝒪(α){\mathcal{O}}(\alpha) is a pullback from SS, if and only if α\alpha is constant on geometric fibres.

4.2 Defining 𝖫𝗈𝗀𝖣𝖱\mathsf{LogDR}

Before defining the logarithmic double ramification cycle it seems useful to summarise the construction of the ‘usual’ double ramification cycle from [BHP+20]. Various constructions of double ramification cycles are given in various places in the literature in various levels of generality (e.g. [Hol19a, MW20, HKP18, BHP+20]). They are mostly121212With the exception of classes constructed by means of admissible covers, which only apply in cases when working with the sheaf of differentials. equivalent, but descriptions of the relations between the constructions are scattered across various sketches in various papers at various levels of generality131313This is in fairly large part the responsibility of the first-named author., making it troublesome to assemble a complete picture. Here we attempt to rectify this by giving a precise and general statement of the relation between the two most widely-used definitions, that of the first-named author by resolving rational maps, and that of Marcus and Wise via tropical divisors (in the form used in [BHP+20]).

4.2.1 Tropical divisors

If C/SC/S is a log curve and α\alpha a PL function on CC, then the line bundle 𝒪C(α){\mathcal{O}}_{C}(\alpha) determines a map S𝔍𝔞𝔠S\to\mathfrak{Jac}. In this way we have an Abel-Jacobi map from the stack of pairs (C/S,α)(C/S,\alpha) to 𝔍𝔞𝔠\mathfrak{Jac}. We can see this Abel-Jacobi map as a first approximation of the double ramification cycle, but the map has relative dimension 1 (a section α\alpha admits no non-trivial automorphisms, whereas the line bundle 𝒪C(α){\mathcal{O}}_{C}(\alpha) has a 𝔾m{\mathbb{G}}_{m} worth of automorphisms), and hence will not induce a good Chow class on 𝔍𝔞𝔠\mathfrak{Jac}. To fix this we need a little more setup. Given a log scheme S=(S,MS)S=(S,M_{S}), we write

𝔾m𝗍𝗋𝗈𝗉(S)=Γ(S,𝖬¯Sgp),\mathbb{G}_{m}^{\mathsf{trop}}(S)=\Gamma(S,\bar{\mathsf{M}}_{S}^{gp}),

which we call the tropical multiplicative group; it can naturally be extended to a presheaf on the category 𝐋𝐒𝐜𝐡S{\mathbf{LSch}}_{S} of log schemes over SS. A tropical line on SS is a 𝔾m𝗍𝗋𝗈𝗉{\mathbb{G}}_{m}^{\mathsf{trop}} torsor on SS for the strict étale topology. Then a point of 𝐃𝐢𝐯{\mathbf{Div}} is a quadruple

(C/S,P,α,)(C/S,P,\alpha,{\mathcal{M}}) (4.2.1)

where C/SC/S is a log curve, PP a tropical line on SS, α:CP\alpha\colon C\to P a morphism over SS with zero outgoing slopes, and {\mathcal{M}} is a line bundle on SS. An isomorphism

(π:CS,P,α,)(π:CS,P,α,)(\pi\colon C\to S,P,\alpha,{\mathcal{M}})\to(\pi\colon C\to S,P^{\prime},\alpha^{\prime},{\mathcal{M}}^{\prime}) (4.2.2)

in 𝐃𝐢𝐯{\mathbf{Div}} is an isomorphism π(α)π(α)\pi^{*}{\mathcal{M}}(\alpha)\to\pi^{*}{\mathcal{M}}^{\prime}(\alpha), and the Abel-Jacobi map 𝖺𝗃:𝐃𝐢𝐯𝔍𝔞𝔠\mathsf{aj}\colon{\mathbf{Div}}\to\mathfrak{Jac} sends (π:CS,P,α,)(\pi\colon C\to S,P,\alpha,{\mathcal{M}}) to π(α)\pi^{*}{\mathcal{M}}(\alpha).

In [BHP+20] we defined 𝖣𝖱CH(𝔍𝔞𝔠)\mathsf{DR}\in\operatorname{CH}(\mathfrak{Jac}) to be the fundamental class of the proper log monomorphism 𝐃𝐢𝐯𝔍𝔞𝔠{\mathbf{Div}}\to\mathfrak{Jac} (we describe this in more detail in definition 4.3. )

4.2.2 Universal σ\sigma-extending morphisms

Over the locus of irreducible curves in 𝔍𝔞𝔠\mathfrak{Jac} the notions of total degree and multidegree coincide, so that 𝖩\mathsf{J} comes with a tautological map from 𝔍𝔞𝔠\mathfrak{Jac}. We can think of this as a rational map

σ:𝔍𝔞𝔠𝖩\sigma\colon\mathfrak{Jac}\dashrightarrow\mathsf{J} (4.2.3)

(rational because it is only defined on the open locus of irreducible curves).

We call a map t:T𝔍𝔞𝔠t\colon T\to\mathfrak{Jac} of algebraic stacks σ\sigma-extending if141414The analogous definition in [Hol19a] had the additional assumption that TT be normal. At the time this was needed in order to be able to apply [Hol19b, Theorem 4.1] at a certain critical step in the arguments, but since then Marcus and Wise have proven the analogue of [Hol19b, Theorem 4.1] with no regularity assumptions, see [MW20, Corollary 3.6.3]. This can then be use to modify they theory of [Hol19a] without a normality assumption. Alternatively one can reinstate the condition that TT be normal, and all of the subsequent discussion will go through unchanged except that we will have to insert normalisations in various places. By Costello’s Theorem [HW21] this will have no effect on the resulting cycles, but will make things much less readable, which is why we prefer to omit the condition.

  1. 1.

    The pullback along tt of the locus of line bundles on smooth curves is schematically dense in TT;

  2. 2.

    The rational map T𝖩T\dashrightarrow\mathsf{J} induced by σ\sigma extends to a morphism (necessarily unique if exists, by separatedness of 𝖩\mathsf{J} over 𝔐\mathfrak{M}).

One can then show just as in [Hol19a] that the category of σ\sigma-extending stacks over 𝔍𝔞𝔠\mathfrak{Jac} has a terminal object, which we denote 𝔍𝔞𝔠{\mathfrak{Jac}}^{\lozenge}. The natural map

f:𝔍𝔞𝔠𝔍𝔞𝔠f\colon{\mathfrak{Jac}}^{\lozenge}\to\mathfrak{Jac} (4.2.4)

is separated, relatively representable by algebraic spaces, of finite presentation, and an isomorphism over the locus of smooth (even treelike) curves, but it is not in general proper. The construction equips it with a map

σ:𝔍𝔞𝔠𝖩.\sigma\colon{\mathfrak{Jac}}^{\lozenge}\to\mathsf{J}. (4.2.5)

4.2.3 The functor of points of 𝔍𝔞𝔠\mathfrak{Jac}^{\lozenge}

One can describe the functor of points (on the category of log schemes) of 𝔍𝔞𝔠\mathfrak{Jac}^{\lozenge} in a manner very similar to the definition of 𝐃𝐢𝐯{\mathbf{Div}}. Namely, a point of 𝔍𝔞𝔠\mathfrak{Jac}^{\lozenge} is a quadruple

(C/S,P,α,)(C/S,P,\alpha,{\mathcal{L}}) (4.2.6)

where C/SC/S is a log curve, PP a tropical line on SS, α:CP\alpha\colon C\to P a morphism over SS with zero outgoing slopes, and {\mathcal{L}} a line bundle on CC such that the line bundle (α){\mathcal{L}}(\alpha) has multidegree 0¯{\underline{0}} on every fibre of C/SC/S.

Given such (C/S,P,α,)(C/S,P,\alpha,{\mathcal{L}}), the map S𝖩S\to\mathsf{J} given by (α){\mathcal{L}}(\alpha) is an extension of σ\sigma, so by the universal property of 𝔍𝔞𝔠\mathfrak{Jac}^{\lozenge} we obtain a map from the stack of such quadruples to 𝔍𝔞𝔠\mathfrak{Jac}^{\lozenge}. To show this is an isomorphism, we may work locally (so assume C/SC/S to be nuclear in the sense of [HMOP20], and smooth over a dense open of SS), then PP can be taken trivial, and it is enough to show that the extension of σ\sigma is given by a PL function; but this follows from [MW20, Corollary 3.6.3].

4.2.4 Comparison

The key actor in section 3.5 is the fibre product of the diagram

𝔐{\mathfrak{M}}𝔍𝔞𝔠{\mathfrak{Jac}^{\lozenge}}𝖩.{\mathsf{J}.}e\scriptstyle{e}σ\scriptstyle{\sigma} (4.2.7)

From the functor-of-points description of 𝔍𝔞𝔠\mathfrak{Jac} this fibre product is exactly given by tuples (C/S,P,α,)𝔍𝔞𝔠(C/S,P,\alpha,{\mathcal{L}})\in\mathfrak{Jac}^{\lozenge} such that (α){\mathcal{L}}(-\alpha) is the pullback of some line bundle on SS, say (α)=π{\mathcal{L}}(-\alpha)=\pi^{*}{\mathcal{M}}. Giving the data of {\mathcal{L}} or of {\mathcal{M}} is exactly equivalent, and the tuple (C/S,P,α,)(C/S,P,\alpha,{\mathcal{M}}) is exactly a point of 𝐃𝐢𝐯{\mathbf{Div}}; in other words we have a pullback square

𝐃𝐢𝐯{{\mathbf{Div}}}𝔐{\mathfrak{M}}𝔍𝔞𝔠{\mathfrak{Jac}^{\lozenge}}𝖩.{\mathsf{J}.}e\scriptstyle{e}σ\scriptstyle{\sigma} (4.2.8)

Marcus and Wise show that the composite 𝐃𝐢𝐯𝔍𝔞𝔠{\mathbf{Div}}\to\mathfrak{Jac} is proper, and in [BHP+20] we define 𝖣𝖱\mathsf{DR} to be the associated cycle on 𝔍𝔞𝔠\mathfrak{Jac}. The full construction of the operational class is a little subtle (see [BHP+20, §2] for details), but is easy to describe for a smooth stack SS mapping to 𝔍𝔞𝔠\mathfrak{Jac}.

Definition 4.3.

Let SS be a smooth stack and φ:S𝔍𝔞𝔠\varphi\colon S\to\mathfrak{Jac} a morphism. Then φ\varphi is lci, so we have a gysin pullback φ!:A(𝐃𝐢𝐯)A(𝐃𝐢𝐯×𝔍𝔞𝔠S)\varphi^{!}\colon A^{*}({\mathbf{Div}})\to A^{*}({\mathbf{Div}}\times_{\mathfrak{Jac}}S), and following [Sko12] a proper pushforward i:A(𝐃𝐢𝐯×𝔍𝔞𝔠S)A(S)i_{*}\colon A^{*}({\mathbf{Div}}\times_{\mathfrak{Jac}}S)\to A^{*}(S). Since SS is smooth the intersection pairing furnishes a map :A(S)CH(S)\cap\colon A^{*}(S)\to\operatorname{CH}(S), and we define

φ𝖣𝖱=(iφ![𝐃𝐢𝐯])cCH(S),\varphi^{*}\mathsf{DR}=\cap(i_{*}\varphi^{!}[{\mathbf{Div}}])c\in\operatorname{CH}(S), (4.2.9)

where [𝐃𝐢𝐯][{\mathbf{Div}}] denotes the fundamental class of 𝐃𝐢𝐯{\mathbf{Div}} as a cycle on itself.

4.2.5 Defining 𝖫𝗈𝗀𝖣𝖱\mathsf{LogDR}

We construct the cycle 𝖫𝗈𝗀𝖣𝖱\mathsf{LogDR} in LogCH(𝔍𝔞𝔠)\operatorname{LogCH}(\mathfrak{Jac}) as hinted at in [BHP+20, §3.8]. We apply the construction of section 3.5, taking S=𝔍𝔞𝔠S=\mathfrak{Jac}, X=𝔍𝔞𝔠X=\mathfrak{Jac}^{\lozenge}, 𝖩=𝖩\mathsf{J}=\mathsf{J}, and σ=σ\sigma=\sigma. We need the natural map

𝔍𝔞𝔠×𝖩𝔐𝔍𝔞𝔠\mathfrak{Jac}^{\lozenge}\times_{\mathsf{J}}\mathfrak{M}\to\mathfrak{Jac} (4.2.10)

to be proper; this can be proven in the same way as in [Hol19a, §5], or follows by the comparison to the construction of Marcus-Wise in section 4.2.4.

Definition 4.4.

The construction in section 3.5 yields a class 𝖫𝗈𝗀𝖣𝖱[σe]f,𝗅𝗈𝗀LogCH(𝔍𝔞𝔠)\mathsf{LogDR}\coloneqq[\sigma^{*}e]_{f,\mathsf{log}}\in\operatorname{LogCH}(\mathfrak{Jac}), the log double ramification cycle.

Comparing the constructions yields

Lemma 4.5.

Applying the pushforward ν:LogCH(𝔍𝔞𝔠)CH(𝔍𝔞𝔠)\nu_{*}\colon\operatorname{LogCH}(\mathfrak{Jac})\to\operatorname{CH}(\mathfrak{Jac}) of definition 2.12 to 𝖫𝗈𝗀𝖣𝖱LogCH(𝔍𝔞𝔠)\mathsf{LogDR}\in\operatorname{LogCH}(\mathfrak{Jac}) recovers the double ramification cycle 𝖣𝖱CH(𝔍𝔞𝔠)\mathsf{DR}\in\operatorname{CH}(\mathfrak{Jac}) of [BHP+20].

4.3 Invariance of 𝖫𝗈𝗀𝖣𝖱\mathsf{LogDR} in twistable families

Throughout this subsection, C/SC/S is a log curve over a smooth log smooth base, and {\mathcal{L}} is a line bundle on CC.

Definition 4.6.

We say the pair (C/S,)(C/S,{\mathcal{L}}) is twistable if there exists a PL function α\alpha on CC such that (α){\mathcal{L}}(\alpha) has multidegree 0¯{\underline{0}}; we call such an α\alpha a twisting function.

Being twistable is equivalent to the existence of a Cartier divisor on DD on CC supported over the boundary of SS and such that (D){\mathcal{L}}(D) has multidegree 0¯{\underline{0}}; see [Hol19b, Theorem 4.1] or [MW20, Corollary 3.6.3].

Lemma 4.7.

Let (C/S,)(C/S,{\mathcal{L}}) be twistable with α\alpha a twisting function. Write φ:S𝔍𝔞𝔠\varphi_{\mathcal{L}}\colon S\to\mathfrak{Jac} for the map induced by {\mathcal{L}}, and φ(α):S𝔍𝔞𝔠\varphi_{{\mathcal{L}}(\alpha)}\colon S\to\mathfrak{Jac} for the map induced by (α){\mathcal{L}}(\alpha). Then

φ𝖫𝗈𝗀𝖣𝖱=φ(α)𝖣𝖱\varphi_{{\mathcal{L}}}^{*}\mathsf{LogDR}=\varphi_{{\mathcal{L}}(\alpha)}^{*}\mathsf{DR} (4.3.1)

in LogCH(S)\operatorname{LogCH}(S) (where we view φ(α)𝖣𝖱\varphi_{{\mathcal{L}}(\alpha)}^{*}\mathsf{DR} in LogCH(S)\operatorname{LogCH}(S) by pullback, cf. section 2.2).

Proof.

Write σ:S𝖩\sigma\colon S\dashrightarrow\mathsf{J} for the rational map induced by {\mathcal{L}}. Then the identity on SS is the universal σ\sigma-extending morphism! More precisely, the extension is given by (α):S𝖩{\mathcal{L}}(\alpha)\colon S\to\mathsf{J}, and it is easily seen to be universal among extensions. We have a pullback diagram

S×Je{S\times_{J}e}e{e}S{S}𝖩,{\mathsf{J},}j\scriptstyle{j}i\scriptstyle{i}σ\scriptstyle{\sigma} (4.3.2)

and the definitions of φ𝖫𝗈𝗀𝖣𝖱\varphi_{{\mathcal{L}}}^{*}\mathsf{LogDR} and φ(α)𝖣𝖱\varphi_{{\mathcal{L}}(\alpha)}^{*}\mathsf{DR} simplify to

φ𝖫𝗈𝗀𝖣𝖱=ji![S]andφ(α)𝖣𝖱=jσ![e],\varphi_{{\mathcal{L}}}^{*}\mathsf{LogDR}=j_{*}i^{!}[S]\;\;\;\text{and}\;\;\;\varphi_{{\mathcal{L}}(\alpha)}^{*}\mathsf{DR}=j_{*}\sigma^{!}[e], (4.3.3)

which are equal since σ![e]=i![s]\sigma^{!}[e]=i^{!}[s] (commutativity of the intersection pairing). ∎

Remark 4.8.

If (C/S,)(C/S,{\mathcal{L}}) is twistable then α\alpha is not unique, but the line bundle (α){\mathcal{L}}(\alpha) is uniquely determined up to pullback from SS. Hence φ(α)𝖣𝖱\varphi_{{\mathcal{L}}(\alpha)}^{*}\mathsf{DR} does not depend on the choice of α\alpha.

Remark 4.9.

Unfortunately the notion of twistable families seems a little too restrictive; not enough of them seem to exist to determine 𝖫𝗈𝗀𝖣𝖱\mathsf{LogDR} from 𝖣𝖱\mathsf{DR} (though we have not written down a proof). Because of this we now introduce a weaker notion.

Definition 4.10.

We say (C/S,)(C/S,{\mathcal{L}}) is almost twistable if there exist a PL function α\alpha on CC and a dense open USU\subseteq S such that:

  1. 1.

    the restriction of α\alpha to UU is a twisting function;

  2. 2.

    for any trait TT with generic point η\eta and any map TST\to S sending η\eta to a point in UU, if the map η𝖩\eta\to\mathsf{J} induced by (α){\mathcal{L}}(\alpha) can be extended to a map T𝖩T\to\mathsf{J} then the map TST\to S factors via UU.

Remark 4.11.

Condition (2) implies that UU is the largest open of SS such that (CU/U,|CU(C_{U}/U,{\mathcal{L}}|_{C_{U}}) is twistable. However, it is not equivalent to this; the definition also captures the possibility that the family might become twistable after some blowup.

Condition (2) is equivalent to the following: suppose uu lies in the closure of UU in 𝔍𝔞𝔠\mathfrak{Jac}, and that there exists a point in the closure of the unit section of 𝔍𝔞𝔠\mathfrak{Jac} with the same multidegree as uu; then uUu\in U. In fact for the invariance below it would be enough to replace this with the weaker condition that the intersection of the closure of UU with the closure of the unit section is contained in UU.

Lemma 4.12.

Let (C/S,)(C/S,{\mathcal{L}}) be almost twistable with α\alpha a twisting function. Write φ:S𝔍𝔞𝔠\varphi_{\mathcal{L}}\colon S\to\mathfrak{Jac} for the map induced by {\mathcal{L}}, and φ(α):S𝔍𝔞𝔠\varphi_{{\mathcal{L}}(\alpha)}\colon S\to\mathfrak{Jac} for the map induced by (α){\mathcal{L}}(\alpha). Then

φ𝖫𝗈𝗀𝖣𝖱=φ(α)𝖣𝖱\varphi_{{\mathcal{L}}}^{*}\mathsf{LogDR}=\varphi_{{\mathcal{L}}(\alpha)}^{*}\mathsf{DR} (4.3.4)

in LogCH(S)\operatorname{LogCH}(S).

Proof.

Write σ:S𝖩\sigma\colon S\dashrightarrow\mathsf{J} for the rational map induced by {\mathcal{L}}. Then the inclusion USU\hookrightarrow S is the universal σ\sigma-extending morphism. More precisely, the extension is given by (α):U𝖩{\mathcal{L}}(\alpha)\colon U\to\mathsf{J}, and the second property of definition 4.10 shows it to be universal among extensions. Since (α){\mathcal{L}}(\alpha) is of total degree 0 over the whole of SS, it defines a map σ¯:S𝖩¯\bar{\sigma}\colon S\to\bar{\mathsf{J}} over the whole of SS. Consider the diagram

U×𝖩e{U\times_{\mathsf{J}}e}e{e}U{U}𝖩{\mathsf{J}}S{S}J¯{\bar{J}}j\scriptstyle{j}σ\scriptstyle{\sigma}σ¯\scriptstyle{\bar{\sigma}} (4.3.5)

where both squares are pullbacks (the top by construction, the bottom by the defining property of UU), so that U×𝖩e=S×𝖩¯eU\times_{\mathsf{J}}e=S\times_{\bar{\mathsf{J}}}e. In the notation of section 3.5 we take S=S~S=\tilde{S} and X=X~=UX=\tilde{X}=U. Then the definitions of φ𝖫𝗈𝗀𝖣𝖱\varphi_{{\mathcal{L}}}^{*}\mathsf{LogDR} and φ(α)𝖣𝖱\varphi_{{\mathcal{L}}(\alpha)}^{*}\mathsf{DR} simplify to

φ𝖫𝗈𝗀𝖣𝖱=ji![U]andφ(α)𝖣𝖱=jσ¯![e]=jσ![e],\varphi_{{\mathcal{L}}}^{*}\mathsf{LogDR}=j_{*}i^{!}[U]\;\;\;\text{and}\;\;\;\varphi_{{\mathcal{L}}(\alpha)}^{*}\mathsf{DR}=j_{*}\bar{\sigma}^{!}[e]=j_{*}\sigma^{!}[e], (4.3.6)

which are equal by the commutativity of the intersection pairing. ∎

The hard work remaining in this paper is to show that there are ‘enough’ almost-twistable families for lemma 4.12 to determine 𝖫𝗈𝗀𝖣𝖱\mathsf{LogDR} from 𝖣𝖱\mathsf{DR}.

4.4 Extending piecewise-linear functions

Let C/SC/S be a log curve. The key to showing the existence of enough almost-twistable families will be to extend PL functions over open subsets of SS to PL functions over the whole of SS, perhaps after some monoidal alteration.

Lemma 4.13.

Let C/SC/S be a log curve with SS a smooth log smooth log algebraic stack. Then there exist

  1. 1.

    a monoidal alteration S~S\tilde{S}\to S;

  2. 2.

    a subdivision C~C×SS~\tilde{C}\to C\times_{S}\tilde{S}

with C~/S~\tilde{C}/\tilde{S} a log curve and C~\tilde{C} regular.

Proof.

This follows from [ALT18]. More precisely, their Theorem 4.4 gives a canonical monoidal resolution over schemes, which therefore applies to stacks. The argument in the proof of their Theorem 4.5 then shows that this monoidal resolution has C~\tilde{C} regular. ∎

After applying the above lemma we will show that PL functions always extend. We start by considering the case where the base SS is very small (nuclear in the sense of [HMOP20]), after which we will glue to a global solution.

Lemma 4.14.

Let C/SC/S be a regular log curve over a (regular) log regular base, with C/SC/S nuclear. Let USU\hookrightarrow S be strict dense open and let α\alpha be a PL function on CU/UC_{U}/U. Then we construct an extension α¯\bar{\alpha} to a PL function on C/SC/S, and this construction is compatible with strict open base-change.

Proof.

Let rr be the rank of 𝖬¯S,s\bar{\mathsf{M}}_{S,s}, and let D1,,DrD_{1},\dots,D_{r} be the divisorial strata of the boundary of SS. Let Γ\Gamma be the graph of C/SC/S over the closed stratum, and let Γi\Gamma_{i} be the graph over the generic point of DiD_{i} (obtained by contracting those edges of Γ\Gamma whose lengths differ from DiD_{i}).

On each DiD_{i} with non-empty intersection with UU we equip Γi\Gamma_{i} with the PL function from α\alpha, and for the other DiD_{i} we put the zero PL function.

Now let zz be a stratum of SS, with graph Γz\Gamma_{z}, and let Nz{1,,r}N_{z}\subseteq\{1,\dots,r\} be such that {z}¯=iNzDi\overline{\{z\}}=\bigcap_{i\in N_{z}}D_{i}. Then 𝖬¯S,z=iNzDi\bar{\mathsf{M}}_{S,z}=\bigoplus_{i\in N_{z}}{\mathbb{N}}\cdot D_{i}; write fi:Di𝖬¯S,zf_{i}\colon{\mathbb{N}}\cdot D_{i}\to\bar{\mathsf{M}}_{S,z} for the natural inclusion. Let vv be a vertex of Γz\Gamma_{z}, and for each iNzi\in N_{z} let viv_{i} be its image in Γi\Gamma_{i} under specialisation. Then we define

α¯(v)=iNzfi(α(vi)).\bar{\alpha}(v)=\sum_{i\in N_{z}}f_{i}(\alpha(v_{i})). (4.4.1)

To check that α¯\bar{\alpha} is a PL function on Γz\Gamma_{z}, suppose that ee is an edge of Γz\Gamma_{z} between vertices uu and vv. By regularity of CC we know that the length of ee is DiD_{i} for some iNzi\in N_{z}; suppose it is D1D_{1}. Then fi(u)=fi(v)f_{i}(u)=f_{i}(v) for every i1i\neq 1, and D1α(f1(u))α(f1(v))D_{1}\mid\alpha(f_{1}(u))-\alpha(f_{1}(v)). It is easy to see that α¯\bar{\alpha} restricts to α\alpha over UU.

Suppose that SSS^{\prime}\to S is a strict open map such that CS/SC_{S^{\prime}}/S^{\prime} is also nuclear and CSC_{S^{\prime}} is regular. Let ss^{\prime} be the closed stratum of SS^{\prime}; it is enough to check the result for the restriction of α¯\bar{\alpha} to Γs\Gamma_{s^{\prime}}. Let N{,,r}N^{\prime}\subseteq\{,\dots,r\} be the set of those DiD_{i} meeting the image of ss^{\prime}. Then each of those DiD_{i} meet the image of SS^{\prime}, and their pullbacks are exactly the divisorial strata on SS^{\prime} (so in particular the rank of 𝖬¯S,s\bar{\mathsf{M}}_{S^{\prime},s^{\prime}} is #N\#N^{\prime}). Then α¯\bar{\alpha} on Γs\Gamma_{s^{\prime}} is constructed by interpolating the values of α\alpha on the DiD_{i} for iNi\in N^{\prime}, regardless of whether we compute this on SS or on SS^{\prime}; in particular, these give the same result. ∎

Lemma 4.15.

Let C/SC/S be a log curve with CC (and hence SS) regular log regular, SS a log algebraic stack. Let USU\hookrightarrow S be a strict dense open immersion and α\alpha a PL function on CUC_{U}. Then there exists a PL function α¯\bar{\alpha} on SS restricting to α\alpha.

Proof.

In lemma 4.14 we give a canonical choice of extension in the case where C/SC/S is nuclear, and these are compatible with smooth base-change, so descend to algebraic stacks. ∎

4.5 𝖫𝗈𝗀𝖣𝖱\mathsf{LogDR} from 𝖣𝖱\mathsf{DR}

We wish to compute 𝖫𝗈𝗀𝖣𝖱\mathsf{LogDR} in LogCH(𝔍𝔞𝔠)\operatorname{LogCH}(\mathfrak{Jac}). Let i:S𝔍𝔞𝔠i\colon S\hookrightarrow\mathfrak{Jac} be a strict open immersion with SS quasi-compact, and write C/SC/S for the universal curve and {\mathcal{L}} on CC for the universal line bundle.

Lemma 4.16.

There exist

  1. 1.

    a monoidal alteration ψ:S~S\psi\colon\tilde{S}\to S;

  2. 2.

    a subdivision C~\tilde{C} of C×SS~C\times_{S}\tilde{S};

such that the pair (C~/S~,ψ)(\tilde{C}/\tilde{S},\psi^{*}{\mathcal{L}}) is almost twistable.

Proof.

Write σ:S𝖩\sigma\colon S\dashrightarrow\mathsf{J} for the rational map induced by {\mathcal{L}}, and let ψ1:SS\psi_{1}\colon S^{\lozenge}\to S be the universal σ\sigma-extending morphism.

Claim: there exists a representable monoidal alteration SS^{\lozenge\lozenge} of SS^{\lozenge} over which the map σ:S𝖩\sigma\colon S^{\lozenge}\to\mathsf{J} can be represented as ψ1(α)\psi_{1}^{*}{\mathcal{L}}(\alpha) for some PL function α\alpha over SS^{\lozenge\lozenge}.

The claim is clear from section 4.2.3 locally on SS^{\lozenge}, but these PL functions are only unique up to addition of a PL function from the base, and so need not glue. We define SS^{\lozenge\lozenge} to be the subfunctor of SS^{\lozenge} where the maps α:CP\alpha\colon C\to P (in the notation of section 4.2.3) can be chosen such that their set of values is totally ordered (in the ordering on PP induced by the monoid structure on 𝖬¯𝗀𝗉\bar{\mathsf{M}}^{\mathsf{gp}}). That this subfunctor is a representable monoidal alteration of SS^{\lozenge} is proven exactly as in [MW20, Theorem 5.3.4] for the map 𝐑𝐮𝐛𝐃𝐢𝐯{\mathbf{Rub}}\to{\mathbf{Div}}, to which it is closely analogous.

Now we can construct these α\alpha locally on SS^{\lozenge\lozenge} just as before, but with P=𝔾M𝗍𝗋𝗈𝗉P={\mathbb{G}}_{M}^{\mathsf{trop}} and the additional requirement that the smallest value taken by α\alpha on any vertex is zero. This makes the α\alpha unique, hence they glue to a global PL function, proving the claim.

Now let SSS^{\blacklozenge}\to S be a sufficiently fine log blowup for SSS^{\lozenge\lozenge}\to S, and let USU\hookrightarrow S^{\blacklozenge} be the lift of SS^{\lozenge\lozenge}.

Writing C/SC^{\blacklozenge}/S^{\blacklozenge} for the pullback of C/SC/S, we apply lemma 4.13 to the C/SC^{\blacklozenge}/S^{\blacklozenge} to construct a monoidal alteration S~S\tilde{S}\to S^{\blacklozenge} and a subdivision C~\tilde{C} of C×SS~=C×SS~C^{\blacklozenge}\times_{S^{\blacklozenge}}\tilde{S}=C\times_{S}\tilde{S} with C~\tilde{C} (and hence S~\tilde{S}) regular. Writing ψ:S~S\psi\colon\tilde{S}\to S for the composite, we claim that the pair (C~/S~,ψ)(\tilde{C}/\tilde{S},\psi^{*}{\mathcal{L}}) is almost twistable.

Let U~\tilde{U} be the pullback of UU to S~\tilde{S} (a twistable open), and let α\alpha be a twisting function over U~\tilde{U}. Then lemma 4.15 implies that this α\alpha can be extended to a PL function over the whole of S~\tilde{S}.

Now let TST\to S be a trait with generic point η\eta landing in U~\tilde{U}. Suppose that the map η𝖩\eta\to\mathsf{J} given by (ψ1)(α)(\psi_{1}^{*}{\mathcal{L}})(\alpha), then it is proven in [Hol19a, Lemma 4.3] that this cannot be extended to a map T𝖩T\to\mathsf{J} unless it can already be extended to a map TUT\to U. This shows that (C~/S~,ψ)(\tilde{C}/\tilde{S},\psi^{*}{\mathcal{L}}) is almost twistable, with U~S~\tilde{U}\hookrightarrow\tilde{S} the largest twistable open. ∎

Let (C~/S~(\tilde{C}/\tilde{S}, ψ)\psi^{*}{\mathcal{L}}) be as in the statement of lemma 4.16, with twisting function α\alpha over S~\tilde{S}. Then we have maps

φ:S~𝔍𝔞𝔠 and φ(α):S~𝔍𝔞𝔠\varphi_{{\mathcal{L}}}\colon\tilde{S}\to\mathfrak{Jac}\;\;\;\text{ and }\;\;\;\varphi_{{\mathcal{L}}(\alpha)}\colon\tilde{S}\to\mathfrak{Jac} (4.5.1)

induced by ψ\psi^{*}{\mathcal{L}} and ψ(α)\psi^{*}{\mathcal{L}}(\alpha) respectively. Then

Theorem 4.17.

We have an equality of cycles

φ𝖫𝗈𝗀𝖣𝖱=φ(α)𝖣𝖱\varphi_{{\mathcal{L}}}^{*}\mathsf{LogDR}=\varphi_{{\mathcal{L}}(\alpha)}^{*}\mathsf{DR} (4.5.2)

in LogCH(S)\operatorname{LogCH}(S).

Proof.

Immediate from lemmas 4.12 and 4.16. ∎

4.6 LogDR is tautological

If {\mathcal{L}} is the universal line bundle on the universal curve π:C𝔍𝔞𝔠\pi\colon C\to\mathfrak{Jac}, we define the class

η=π(c1()2)CH(𝔍𝔞𝔠).\eta=\pi_{*}(c_{1}({\mathcal{L}})^{2})\in\operatorname{CH}(\mathfrak{Jac}). (4.6.1)

In [BHP+20, Definition 4] we defined a tautological subring of CH(𝔍𝔞𝔠)\operatorname{CH}(\mathfrak{Jac}); it is the {\mathbb{Q}}-span of certain decorated prestable graphs of degree 0, as described in [BHP+20, Section 0.3.3]; in particular, it includes the class η\eta from eq. 4.6.1. Here we prove that 𝖫𝗈𝗀𝖣𝖱\mathsf{LogDR} is contained in the corresponding tautological subring of LogCH(𝔍𝔞𝔠)\operatorname{LogCH}(\mathfrak{Jac}). In fact, we can prove something stronger151515This improvement was suggested to us by Johannes Schmitt, to whom we are very grateful for permission to include it. . We write [η]CH(𝔍𝔞𝔠){\mathbb{Q}}[\eta]\subseteq\operatorname{CH}(\mathfrak{Jac}) for the sub-{\mathbb{Q}}-algebra of the Chow ring generated by the class η\eta of eq. 4.6.1, and recall from definition 3.15 that [η]𝗅𝗈𝗀{\mathbb{Q}}[\eta]^{\mathsf{log}} denotes the corresponding subring of LogCH(𝔍𝔞𝔠)\operatorname{LogCH}(\mathfrak{Jac}); we show that 𝖫𝗈𝗀𝖣𝖱\mathsf{LogDR} lies in [η]𝗅𝗈𝗀{\mathbb{Q}}[\eta]^{\mathsf{log}}.

Continuing in the notation of the previous subsection, we can pull back [η]𝗅𝗈𝗀{\mathbb{Q}}[\eta]^{\mathsf{log}} along φ:S𝔍𝔞𝔠\varphi_{\mathcal{L}}\colon S\to\mathfrak{Jac} to give a subring of LogCH(S)\operatorname{LogCH}(S). Since φ\varphi_{\mathcal{L}} is strict this is equivalent to pulling back [η]CH(𝔍𝔞𝔠){\mathbb{Q}}[\eta]\subseteq\operatorname{CH}(\mathfrak{Jac}) to CH(S)\operatorname{CH}(S), then taking the corresponding subring of LogCH(S)\operatorname{LogCH}(S). We denote the resulting subring [η]S𝗅𝗈𝗀LogCH(S){\mathbb{Q}}[\eta]_{S}^{\mathsf{log}}\subseteq\operatorname{LogCH}(S).

Lemma 4.18.

Suppose kk has characteristic zero. The cycle φ(α)𝖣𝖱\varphi_{{\mathcal{L}}(\alpha)}^{*}\mathsf{DR} lies in [η]S𝗅𝗈𝗀LogCH(S){\mathbb{Q}}[\eta]_{S}^{\mathsf{log}}\subseteq\operatorname{LogCH}(S).

We are grateful to Johannes Schmitt for pointing out an omission in an earlier version of the proof (as well as the strengthening mentioned above).

Proof.

This is an easy consequence of Pixton’s formula for 𝖣𝖱\mathsf{DR} on 𝔍𝔞𝔠{\mathfrak{Jac}}^{\lozenge}, as stated in equation (56) of [BHP+20, §0.7]. The formula expresses 𝖣𝖱\mathsf{DR} as a polynomial in the following classes:

  1. 1.

    The class π(c1(ψ(α~))2)\pi_{*}(c_{1}(\psi^{*}{\mathcal{L}}(\tilde{\alpha}))^{2});

  2. 2.

    Classes ψh+ψh\psi_{h}+\psi_{h^{\prime}} where hh, hh^{\prime} are the two half-edges forming an edge of a graph of C/S~C/\tilde{S}.

It hence suffices to show that the above classes lie in [η]S𝗅𝗈𝗀{\mathbb{Q}}[\eta]_{S}^{\mathsf{log}}; we treat them in order:

  1. 1.

    The class π(c1(ψ(α~))2)\pi_{*}(c_{1}(\psi^{*}{\mathcal{L}}(\tilde{\alpha}))^{2}) can be expanded as a sum

    π(c1(ψ)2)+π(c1(ψ)c1(𝒪C(α~))+π(c1(𝒪C(α~))2).\pi_{*}(c_{1}(\psi^{*}{\mathcal{L}})^{2})+\pi_{*}(c_{1}(\psi^{*}{\mathcal{L}})c_{1}({\mathcal{O}}_{C}(\tilde{\alpha}))+\pi_{*}(c_{1}({\mathcal{O}}_{C}(\tilde{\alpha}))^{2}).

    The first summand is the pullback of the tautological class η=π(c1()2)\eta=\pi_{*}(c_{1}({\mathcal{L}})^{2}) from CH(𝔍𝔞𝔠)\operatorname{CH}(\mathfrak{Jac}), hence is in [η]S𝗅𝗈𝗀{\mathbb{Q}}[\eta]_{S}^{\mathsf{log}}. For the second summand, we can reduce to computing π(c1(ψ)D)\pi_{*}(c_{1}(\psi^{*}{\mathcal{L}})D) where DD is some vertical prime divisor on C/S~C/\tilde{S}, say with image a prime divisor ZZ on S~\tilde{S}. Then for dimension reasons we see that π(c1(ψ)D)\pi_{*}(c_{1}(\psi^{*}{\mathcal{L}})D) is an integer multiple of the class of the boundary divisor ZZ, in particular is in [η]S𝗅𝗈𝗀{\mathbb{Q}}[\eta]_{S}^{\mathsf{log}}.

    Finally, the class c1(𝒪C(α~))c_{1}({\mathcal{O}}_{C}(\tilde{\alpha})) can be written as a sum of vertical boundary divisors on C/S~C/\tilde{S}. If DD and EE are distinct vertical prime divisors then DD and EE meet properly, and their locus of intersection is a union of vertical codimension 2 loci in CC (which push down to zero on S~\tilde{S} for dimension reasons) and horizontal boundary strata which push forward to boundary classes on S~\tilde{S}.

    It remains to show that π(D2)\pi_{*}(D^{2}) is tautological. For this, let ZZ be the prime divisor in S~\tilde{S} which is the image of DD, and let EE be the vertical divisor lying over ZZ such that πZ=D+E\pi^{*}Z=D+E. Then π(D2)=π(D(πZE))=π(DE)\pi_{*}(D^{2})=\pi_{*}(D\cdot(\pi^{*}Z-E))=\pi_{*}(D\cdot E), which reduces us to the previous case.

  2. 2.

    These are exactly the first chern classes of conormal bundles to boundary divisors on S~\tilde{S}. As such they can be realised as self-intersections of these boundary divisors (as we assume S~\tilde{S} simple), hence are in [η]S𝗅𝗈𝗀{\mathbb{Q}}[\eta]_{S}^{\mathsf{log}}. ∎

Putting together theorem 4.17 and lemma 4.18 we obtain

Corollary 4.19.

Suppose kk has characteristic zero. Write TCH(𝔍𝔞𝔠)T\subseteq\operatorname{CH}(\mathfrak{Jac}) for the tautological ring as in [BHP+20, Definition 4], and [η]T{\mathbb{Q}}[\eta]\subseteq T for the subring generated by the class η\eta from eq. 4.6.1. Then the class 𝖫𝗈𝗀𝖣𝖱LogCH(𝔍𝔞𝔠)\mathsf{LogDR}\in\operatorname{LogCH}(\mathfrak{Jac}) lies in [η]𝗅𝗈𝗀T𝗅𝗈𝗀{\mathbb{Q}}[\eta]^{\mathsf{log}}\subseteq T^{\mathsf{log}}.

4.7 Conjecture C

In this section we prove Conjecture C of [MPS21]; we again thank Johannes Schmitt for corrections and improvements to this argument. We first set up some general notation: if XX is any log smooth log algebraic stack over kk we write divCH(X)\operatorname{divCH}(X) for the subring of CH(X)\operatorname{CH}(X) generated by classes of degree 1 (in other words, by divisor classes; we call it the divisorial subring). Similarly the ring LogCH(X)\operatorname{LogCH}(X) is graded by codimension, and we write divLogCH(X)\operatorname{divLogCH}(X) for the subring generated in degree 1.

Lemma 4.20.

If TdivCH(X)T\subseteq\operatorname{divCH}(X) is any subring, then T𝗅𝗈𝗀divLogCH(X)T^{\mathsf{log}}\subseteq\operatorname{divLogCH}(X).

Proof.

We may assume XX is quasi-compact. Let tT𝗅𝗈𝗀LogCH(X)t\in T^{\mathsf{log}}\subseteq\operatorname{LogCH}(X), then there exists a simple blowup X~X\tilde{X}\to X on which tt is determined; write t~CH(X~)\tilde{t}\in\operatorname{CH}(\tilde{X}) for the determination. By theorem 3.8 there exists a polynomial pp in piecewise-linear functions on X~\tilde{X} such that ΦX~(p)=t~\Phi_{\tilde{X}}(p)=\tilde{t}, where ΦX~\Phi_{\tilde{X}} is the map as in eq. 3.3.12. Now ΦX~\Phi_{\tilde{X}} is a ring homomorphism and piecewise-linear functions map to classes of degree 1, so the result follows. ∎

Theorem 4.21 ([MPS21, Conjecture C]).

Suppose kk is a field of characteristic zero. Then 𝖫𝗈𝗀𝖣𝖱\mathsf{LogDR} lies in divLogCH\operatorname{divLogCH}.

Proof.

We know 𝖫𝗈𝗀𝖣𝖱[η]𝗅𝗈𝗀\mathsf{LogDR}\in{\mathbb{Q}}[\eta]^{\mathsf{log}}, and the class η\eta (defined in eq. 4.6.1) has degree 1, so the result follows from lemma 4.20. ∎

5 The double-double ramification cycle

5.1 Iterated double ramification cycles

Let rr be a positive integer, and let 𝔍𝔞𝔠r\mathfrak{Jac}^{r} be the fibre product of rr copies of 𝔍𝔞𝔠\mathfrak{Jac} over 𝔐\mathfrak{M}. This is smooth and log smooth, and comes with rr projection maps to 𝔍𝔞𝔠\mathfrak{Jac}. According to definition 2.10 we can pull back 𝖫𝗈𝗀𝖣𝖱\mathsf{LogDR} along each of the projection maps, yielding rr elements of LogCH(𝔍𝔞𝔠r)\operatorname{LogCH}(\mathfrak{Jac}^{r}). We define 𝖫𝗈𝗀𝖣𝖱r\mathsf{LogDR}_{r} to be the product of these elements in the ring LogCH(𝔍𝔞𝔠r)\operatorname{LogCH}(\mathfrak{Jac}^{r}).

We can also give a more direct construction of 𝖫𝗈𝗀𝖣𝖱r\mathsf{LogDR}_{r}. Write 𝖩r\mathsf{J}^{r} for the rr-fold fibre product of 𝖩\mathsf{J} with itself over 𝔐\mathfrak{M}, with ere_{r} the unit section. Then over the locus of smooth curves we have a tautological morphism σr:𝔍𝔞𝔠r𝖩r\sigma_{r}\colon\mathfrak{Jac}^{r}\to\mathsf{J}^{r}, we view it as a rational map σr:𝔍𝔞𝔠r𝖩r\sigma_{r}\colon\mathfrak{Jac}^{r}\dashrightarrow\mathsf{J}^{r}, and let 𝔍𝔞𝔠r{\mathfrak{Jac}}^{r\lozenge} be the universal σr\sigma_{r}-extending stack over 𝔍𝔞𝔠r\mathfrak{Jac}^{r}. The pullback σrer\sigma_{r}^{*}e_{r} is proper over 𝔐\mathfrak{M}, so we can apply the construction in section 3.5 to obtain a class [σrer]𝗅𝗈𝗀LogCH(𝔍𝔞𝔠r)[\sigma_{r}^{*}e_{r}]_{\mathsf{log}}\in\operatorname{LogCH}(\mathfrak{Jac}^{r}).

Lemma 5.1.

These two constructions of 𝖫𝗈𝗀𝖣𝖱r\mathsf{LogDR}_{r} coincide, i.e.

𝖫𝗈𝗀𝖣𝖱r=[σrer]𝗅𝗈𝗀.\mathsf{LogDR}_{r}=[\sigma_{r}^{*}e_{r}]_{\mathsf{log}}. (5.1.1)
Proof.

We begin by comparing 𝔍𝔞𝔠r{\mathfrak{Jac}}^{r\lozenge} with (𝔍𝔞𝔠)r({\mathfrak{Jac}}^{\lozenge})^{r}, where the latter denotes the rr-fold fibre product over 𝔐\mathfrak{M} in the category of fs log algebraic stacks. The composites 𝔍𝔞𝔠r𝖩r𝖩{\mathfrak{Jac}}^{r\lozenge}\to\mathsf{J}^{r}\to\mathsf{J} are σ\sigma-extending, hence the universal property furnishes rr maps 𝔍𝔞𝔠r𝔍𝔞𝔠{\mathfrak{Jac}}^{r\lozenge}\to{\mathfrak{Jac}}^{\lozenge}, hence a map

𝔍𝔞𝔠r(𝔍𝔞𝔠)r{\mathfrak{Jac}}^{r\lozenge}\to({\mathfrak{Jac}}^{\lozenge})^{r} (5.1.2)

to the fibre product. On the other hand, the fibre product (𝔍𝔞𝔠)r({\mathfrak{Jac}}^{\lozenge})^{r} is σr\sigma_{r}-extending, yielding an inverse to eq. 5.1.2.

The claimed equality of cycles is then immediate from the construction in section 3.5 and an application of [Ful84, example 6.5.2] (whose proof carries over to this setting essentially unchanged). ∎

Write 1,,r{\mathcal{L}}_{1},\dots,{\mathcal{L}}_{r} for the tautological line bundles on the universal curve over 𝔍𝔞𝔠r\mathfrak{Jac}^{r}, with corresponding classes

ηi=π(c1()2)CH(𝔍𝔞𝔠),\eta_{i}=\pi_{*}(c_{1}({\mathcal{L}})^{2})\in\operatorname{CH}(\mathfrak{Jac}), (5.1.3)

and let [ηr]{\mathbb{Q}}[\eta^{r}] denote the sub-{\mathbb{Q}}-algebra of CH(𝔍𝔞𝔠r)\operatorname{CH}(\mathfrak{Jac}^{r}) generated by these classes. From corollary 4.19 and the first construction of 𝖫𝗈𝗀𝖣𝖱r\mathsf{LogDR}_{r} we obtain

Lemma 5.2.
𝖫𝗈𝗀𝖣𝖱r[ηr]𝗅𝗈𝗀LogCH(𝔍𝔞𝔠r).\mathsf{LogDR}_{r}\in{\mathbb{Q}}[\eta^{r}]^{\mathsf{log}}\subseteq\operatorname{LogCH}(\mathfrak{Jac}^{r}). (5.1.4)

5.2 GLr()\operatorname{GL}_{r}({\mathbb{Z}})-invariance

Let G/SG/S be a commutative group scheme and MM an r×rr\times r matrix with integer coefficients. Writing G×SrG^{\times_{S}^{r}} for the fibre product of GG with itself rr times over SS, we write

[M]:G×SrG×Sr[M]\colon G^{\times_{S}^{r}}\to G^{\times_{S}^{r}} (5.2.1)

for the endomorphism induced by MM. If MGLr()M\in\operatorname{GL}_{r}({\mathbb{Z}}) then this is an automorphism.

Applying this to 𝔍𝔞𝔠\mathfrak{Jac} over 𝔐\mathfrak{M} with MGLr()M\in\operatorname{GL}_{r}({\mathbb{Z}}) yields an automorphism

[M]:𝔍𝔞𝔠r𝔍𝔞𝔠r,[M]\colon\mathfrak{Jac}^{r}\to\mathfrak{Jac}^{r}, (5.2.2)

and pulling back along the map yields an automorphism

[M]:LogCH(𝔍𝔞𝔠r)LogCH(𝔍𝔞𝔠r).[M]^{*}\colon\operatorname{LogCH}(\mathfrak{Jac}^{r})\to\operatorname{LogCH}(\mathfrak{Jac}^{r}). (5.2.3)
Theorem 5.3.

The map [M][M]^{*} of eq. 5.2.3 takes 𝖫𝗈𝗀𝖣𝖱r\mathsf{LogDR}_{r} to itself.

Proof.

For this we use the second construction of 𝖫𝗈𝗀𝖣𝖱r\mathsf{LogDR}_{r}, going via 𝔍𝔞𝔠r{\mathfrak{Jac}}^{r\lozenge}. We write σr:𝔍𝔞𝔠r𝖩r\sigma_{r}\colon\mathfrak{Jac}^{r}\dashrightarrow\mathsf{J}^{r}, and we write ee for the unit section of 𝖩r\mathsf{J}^{r}. We define 𝔍𝔞𝔠M\mathfrak{Jac}^{M\lozenge} to be the limit (in the fs category) of the solid diagram

𝔍𝔞𝔠r{\mathfrak{Jac}^{r\lozenge}}𝔍𝔞𝔠r{\mathfrak{Jac}^{r}}𝔍𝔞𝔠M{\mathfrak{Jac}^{M\lozenge}}𝔍𝔞𝔠r{\mathfrak{Jac}^{r\lozenge}}𝔍𝔞𝔠r{\mathfrak{Jac}^{r}}ν\scriptstyle{\nu}[M]\scriptstyle{[M]}s\scriptstyle{s}t\scriptstyle{t}ν\scriptstyle{\nu} (5.2.4)

(we can think of 𝔍𝔞𝔠M\mathfrak{Jac}^{M\lozenge} as the common refinement of 𝔍𝔞𝔠r\mathfrak{Jac}^{r\lozenge} with its translation along [M][M]). Now the composite νs\nu\circ s is σr\sigma_{r}-extending, as is the composite νt\nu\circ t, so we obtain a commutative diagram

𝔍𝔞𝔠r{\mathfrak{Jac}^{r}}𝖩r{\mathsf{J}^{r}}𝔍𝔞𝔠M{\mathfrak{Jac}^{M\lozenge}}𝔍𝔞𝔠r{\mathfrak{Jac}^{r}}𝖩r.{\mathsf{J}^{r}.}[M]\scriptstyle{[M]}[M]\scriptstyle{[M]}νs\scriptstyle{\nu\circ s}νt\scriptstyle{\nu\circ t}σs\scriptstyle{\sigma_{s}}σt\scriptstyle{\sigma_{t}} (5.2.5)

Now σse\sigma_{s}^{*}e is a cycle on 𝔍𝔞𝔠M\mathfrak{Jac}^{M\lozenge}, which can induce (following section 3.5) a logarithmic cycle on 𝔍𝔞𝔠r\mathfrak{Jac}^{r} in two ways; either via the map νs\nu\circ s or via the map νt\nu\circ t. Our notation is [σse]νs,𝗅𝗈𝗀[\sigma_{s}^{*}e]_{\nu\circ s,\mathsf{log}} for the former and [σse]νt,𝗅𝗈𝗀[\sigma_{s}^{*}e]_{\nu\circ t,\mathsf{log}} for the latter, and we define analogously [σte]νs,𝗅𝗈𝗀[\sigma_{t}^{*}e]_{\nu\circ s,\mathsf{log}} and [σte]νt,𝗅𝗈𝗀[\sigma_{t}^{*}e]_{\nu\circ t,\mathsf{log}}, all elements of LogCH(𝔍𝔞𝔠r)\operatorname{LogCH}(\mathfrak{Jac}_{r}). Applying lemma lemma 5.1 and commutativity of the diagram yields the relations

𝖫𝗈𝗀𝖣𝖱r=[σse]νs,𝗅𝗈𝗀=[σte]νt,𝗅𝗈𝗀,\mathsf{LogDR}_{r}=[\sigma_{s}^{*}e]_{\nu\circ s,\mathsf{log}}=[\sigma_{t}^{*}e]_{\nu\circ t,\mathsf{log}}, (5.2.6)
[M1]𝖫𝗈𝗀𝖣𝖱r=[σse]νt,𝗅𝗈𝗀and[M]𝖫𝗈𝗀𝖣𝖱r=[σte]νs,𝗅𝗈𝗀.[M^{-1}]^{*}\mathsf{LogDR}_{r}=[\sigma_{s}^{*}e]_{\nu\circ t,\mathsf{log}}\;\;\;\text{and}\;\;\;[M]^{*}\mathsf{LogDR}_{r}=[\sigma_{t}^{*}e]_{\nu\circ s,\mathsf{log}}. (5.2.7)

Finally, we note that [M]e=e[M]^{*}e=e and σt=[M]σs\sigma_{t}=[M]\circ\sigma_{s}, so that

[M]𝖫𝗈𝗀𝖣𝖱r=[σte]νs,𝗅𝗈𝗀=[σsMe]νs,𝗅𝗈𝗀=[σse]νs,𝗅𝗈𝗀=𝖫𝗈𝗀𝖣𝖱r.[M]^{*}\mathsf{LogDR}_{r}=[\sigma_{t}^{*}e]_{\nu\circ s,\mathsf{log}}=[\sigma_{s}^{*}M^{*}e]_{\nu\circ s,\mathsf{log}}=[\sigma_{s}^{*}e]_{\nu\circ s,\mathsf{log}}=\mathsf{LogDR}_{r}.\qed (5.2.8)
Remark 5.4.

One can alternatively prove this theorem by appealing to the invariance of 𝐃𝐢𝐯{\mathbf{Div}} (see section 4.2.1) under the action of MM.

5.3 On the moduli space of curves

Here we translate the above results into the setting of [HPS19]. We fix non-negative integers gg, nn and a positive integer rr. We choose rr line bundles 1,,r{\mathcal{L}}_{1},\dots,{\mathcal{L}}_{r} of total degree zero on the universal curve CC over ¯g,n\overline{{\mathcal{M}}}_{g,n}, of the form

i=ωki(j=1nai,jxj){\mathcal{L}}_{i}=\omega^{k_{i}}(-\sum_{j=1}^{n}a_{i,j}x_{j}) (5.3.1)

where ai,1,,ai,na_{i,1},\dots,a_{i,n} are integers summing to ki(2g2)k_{i}(2g-2). The tuple 1,,r{\mathcal{L}}_{1},\dots,{\mathcal{L}}_{r} defines a morphism

Ψ:¯g,n𝔍𝔞𝔠r,\Psi\colon\overline{{\mathcal{M}}}_{g,n}\to\mathfrak{Jac}^{r}, (5.3.2)

and we denote

𝖫𝗈𝗀𝖣𝖱(1,,r)=Ψ𝖫𝗈𝗀𝖣𝖱LogCH(¯g,n),\mathsf{LogDR}({\mathcal{L}}_{1},\dots,{\mathcal{L}}_{r})=\Psi^{*}\mathsf{LogDR}\in\operatorname{LogCH}(\overline{{\mathcal{M}}}_{g,n}), (5.3.3)

and

𝖣𝖱(1,,r)=ν𝖫𝗈𝗀𝖣𝖱(1,,r)CH(¯g,n).\mathsf{DR}({\mathcal{L}}_{1},\dots,{\mathcal{L}}_{r})=\nu_{*}\mathsf{LogDR}({\mathcal{L}}_{1},\dots,{\mathcal{L}}_{r})\in\operatorname{CH}(\overline{{\mathcal{M}}}_{g,n}). (5.3.4)
Theorem 5.5 (DDR is tautological).

Suppose kk has characteristic zero. The cycle 𝖣𝖱(1,,r)\mathsf{DR}({\mathcal{L}}_{1},\dots,{\mathcal{L}}_{r}) lies in the tautological subring of CH(¯g,n)\operatorname{CH}(\overline{{\mathcal{M}}}_{g,n}).

Proof.

We define the tautological ring of 𝔍𝔞𝔠r\mathfrak{Jac}^{r} to be the ring generated by the pullbacks of the tautological rings of the factors 𝔍𝔞𝔠\mathfrak{Jac}. Then the pullback along Ψ\Psi of the tautological ring of 𝔍𝔞𝔠r\mathfrak{Jac}^{r} coincides with the tautological ring TT of ¯g,n\overline{{\mathcal{M}}}_{g,n}, since each i{\mathcal{L}}_{i} is of the form eq. 5.3.1. Lemma 5.2 then implies that 𝖫𝗈𝗀𝖣𝖱(1,,r)\mathsf{LogDR}({\mathcal{L}}_{1},\dots,{\mathcal{L}}_{r}) lies in T𝗅𝗈𝗀T^{\mathsf{log}}. Now TT is tectonic by lemma 3.17, and so 𝖣𝖱(1,,r)T\mathsf{DR}({\mathcal{L}}_{1},\dots,{\mathcal{L}}_{r})\in T by definition 3.16. ∎

Note that the ring [η]CH(¯g,n){\mathbb{Q}}[\eta]\subseteq\operatorname{CH}(\overline{{\mathcal{M}}}_{g,n}) is not in general tectonic, so we really need to work with the tautological ring here.

Theorem 5.6 (GL()\operatorname{GL}({\mathbb{Z}})-invariance of DDR).

If MGLr()M\in\operatorname{GL}_{r}({\mathbb{Z}}) and

M[1,,r]=[1,,r],M[{\mathcal{L}}_{1},\dots,{\mathcal{L}}_{r}]=[{\mathcal{F}}_{1},\dots,{\mathcal{F}}_{r}],

then

𝖣𝖱(1,,r)=𝖣𝖱(1,,r).\mathsf{DR}({\mathcal{L}}_{1},\dots,{\mathcal{L}}_{r})=\mathsf{DR}({\mathcal{F}}_{1},\dots,{\mathcal{F}}_{r}). (5.3.5)
Proof.

Immediate from theorem 5.3. ∎

In the case r=2r=2 this recovers [HPS19, Theorem 1.2].

References

  • [ALT18] Karim Adiprasito, Gaku Liu, and Michael Temkin. Semistable reduction in characteristic 0. arXiv preprint arXiv:1810.03131, 2018.
  • [AW18] Dan Abramovich and Jonathan Wise. Birational invariance in logarithmic gromov–witten theory. Compositio Mathematica, 154(3):595–620, 2018.
  • [BHP+20] Younghan Bae, David Holmes, Rahul Pandharipande, Johannes Schmitt, and Rosa Schwarz. Pixton’s formula and Abel-Jacobi theory on the Picard stack. arXiv preprint arXiv:2004.08676, 2020.
  • [BR19] Alexandr Buryak and Paolo Rossi. Quadratic double ramification integrals and the noncommutative KdV hierarchy. arXiv preprint arXiv:1909.11617, 2019.
  • [BV12] Niels Borne and Angelo Vistoli. Parabolic sheaves on logarithmic schemes. Advances in Mathematics, 231(3-4):1327–1363, 2012.
  • [Dud15] Bashar Dudin. Compactified universal jacobian and the double ramification cycle. http://arxiv.org/abs/1505.02897, 2015.
  • [FP05] Carel Faber and Rahul Pandharipande. Relative maps and tautological classes. Journal of the European Mathematical Society (Print), 7(1):13–49, 2005.
  • [FP16] Gavril Farkas and Rahul Pandharipande. The moduli space of twisted canonical divisors (with an appendix by janda, pandharipande, pixton, and zvonkine). Journal of the Institute of Mathematics of Jussieu, pages 1–58, 2016.
  • [Ful84] William Fulton. Intersection theory. Springer, 1984.
  • [Gil12] William Gillam. Logarithmic stacks and minimality. International Journal of Mathematics, 23(07):1250069, 2012.
  • [GS13] Mark Gross and Bernd Siebert. Logarithmic gromov-witten invariants. Journal of the American Mathematical Society, 26(2):451–510, 2013.
  • [GZ14] Samuel Grushevsky and Dmitry Zakharov. The double ramification cycle and the theta divisor. Proc. Amer. Math. Soc., 142(no. 12):4053–4064, 2014.
  • [Hai13] Richard Hain. Normal functions and the geometry of moduli spaces of curves. In G. Farkas and I. Morrison, editors, Handbook of Moduli, Volume I. Advanced Lectures in Mathematics, Volume XXIV. International Press, Boston, 2013.
  • [HKP18] David Holmes, Jesse Kass, and Nicola Pagani. Extending the double ramification cycle using jacobians. European Journal of Mathematics, https://doi.org/10.1007/s40879-018-0256-7, 4(3):1087–1099, 2018.
  • [HMOP20] David Holmes, Samouil Molcho, Giulio Orecchia, and Thibault Poiret. Models of jacobians of curves. arXiv preprint arXiv:2007.10792, 2020.
  • [Hol19a] David Holmes. Extending the double ramification cycle by resolving the abel-jacobi map. J. Inst. Math. Jussieu, https://doi.org/10.1017/S1474748019000252, 2019.
  • [Hol19b] David Holmes. Néron models of jacobians over base schemes of dimension greater than 1. http://arxiv.org/abs/1402.0647, J. Reine Angew. Math., 747:109–145, 2019.
  • [HPS19] David Holmes, Aaron Pixton, and Johannes Schmitt. Multiplicativity of the double ramification cycle. Doc. Math., 24:545–562, 2019.
  • [HS19] David Holmes and Johannes Schmitt. Infinitesimal structure of the pluricanonical double ramification locus. arXiv preprint https://arxiv.org/abs/1909.11981, 2019.
  • [HW21] Leo Herr and Jonathan Wise. Costello’s pushforward formula: errata and generalization. 03 2021.
  • [IT14] Luc Illusie and Michael Temkin. Gabber’s modification theorem (absolute case). Travaux de Gabber sur l’uniformisation locale et la cohomologie étale des schémas quasi-excellents, 2014.
  • [JPPZ17] Felix Janda, Rahul Pandharipande, Aaron Pixton, and Dimitri Zvonkine. Double ramification cycles on the moduli spaces of curves. Publications mathématiques de l’IHÉS, 125(1):221–266, 2017.
  • [JPPZ18] Felix Janda, Rahul Pandharipande, Aaron Pixton, and Dimitri Zvonkine. Double ramification cycles with target varieties. arXiv preprint https://arxiv.org/abs/1812.10136, 2018.
  • [Kat89a] Kazuya Kato. Logarithmic degeneration and dieudonné theory. preprint, 1989.
  • [Kat89b] Kazuya Kato. Logarithmic structures of Fontaine-Illusie. In Algebraic analysis, geometry, and number theory (Baltimore, MD, 1988), pages 191–224. Johns Hopkins Univ. Press, Baltimore, MD, 1989.
  • [Kat00] Fumiharu Kato. Log smooth deformation and moduli of log smooth curves. Internat. J. Math., 11(2):215–232, 2000.
  • [Kre99] Andrew Kresch. Cycle groups for Artin stacks. Invent. Math., 138(3):495–536, 1999.
  • [MPS21] Samouil Molcho, Rahul Pandharipande, and Johannes Schmitt. The hodge bundle, the universal 0-section, and the log chow ring of the moduli space of curves. https://arxiv.org/pdf/2101.08824.pdf, 2021.
  • [MW20] Steffen Marcus and Jonathan Wise. Logarithmic compactification of the Abel-Jacobi section. Proceedings of the London Mathematical Society, 121(5):1207–1250, 2020.
  • [Niz06] Wiesława Nizioł. Toric singularities: log-blow-ups and global resolutions. Journal of algebraic geometry, 15(1):1–29, 2006.
  • [Ogu18] Arthur Ogus. Lectures on logarithmic algebraic geometry, volume 178. Cambridge University Press, 2018.
  • [Pay06] Sam Payne. Equivariant chow cohomology of toric varieties. Math. Res. Lett, 13(1):29–41, 2006.
  • [Sch16] Johannes Schmitt. Dimension theory of the moduli space of twisted kk-differentials. arXiv preprint arXiv:1607:08429, 2016.
  • [Sko12] J Skowera. Proper pushforward of integral cycles on algebraic stacks. preprint https://expz. github. io/files/stacks-pp. pdf, 2012.
  • [Sta13] The Stacks Project Authors. Stacks project. http://stacks.math.columbia.edu, 2013.
  • [Vis89] Angelo Vistoli. Intersection theory on algebraic stacks and on their moduli spaces. Invent. Math., 97(3):613–670, 1989.

David Holmes, Rosa Schwarz
Mathematisch Instituut, Universiteit Leiden, Postbus 9512, 2300 RA Leiden, Netherlands
E-mail addresses: [email protected], [email protected]