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

Introduction to logarithmic geometry

Michael Temkin Einstein Institute of Mathematics
The Hebrew University of Jerusalem
Edmond J. Safra Campus, Giv’at Ram, Jerusalem, 91904, Israel
[email protected]
Key words and phrases:
Logarithmic geometry, toroidal geometry, resolution of singularities
This research is supported by BSF grants 2014365 and 2018193, ERC Consolidator Grant 770922 - BirNonArchGeom.

1. Introduction

These notes will substitute a chapter in a book on recent advances in resolution of singularities based on a series of minicourses given at an Oberwolfach seminar. It is based on a minicourse on logarithmic resolution of singularities given by the author, and it provides an extended version of its first part devoted to introduction to logarithmic geometry with a view towards applications to resolution. I do not aim to build a theory with proofs (and this is impossible in a 3-4 lecture long course). The goal is to make the reader familiar with basic definitions, constructions, techniques and results of logarithmic geometry. I formulate most of the results as “Exercises” and try to keep them at a reasonable level of difficulty. References to the literature are also provided. At the first reading of the material it may be worth just to read the formulations and hints or comments about main ideas of the arguments, without trying to solve them or read proofs in the cited papers.

1.1. History and motivation

1.1.1. The discovery

Logarithmic structures and schemes were discovered by J.-M. Fontaine and L. Illusie on Sunday, July 17, 1988 during a discussion in a train on their travel to Oberwolfach workshop ”Aritmetische Algebraische Geometrie”. In fact, the discussion was in the continuation of an IHES seminar that had taken place in the spring, and the construction was motivated by the necessity of finding a suitable framework in which an analogue of Steenbrink’s limiting Hodge structure for a semistable reduction over a complex disc could be defined in mixed characteristic in order to make sense of the CstC_{\rm st}-conjecture of Fontaine and Jannsen.

During the workshop Illusie prepared a short summary of the discussion and showed it to K. Kato, who was very enthusiastic about the new notion and very quickly wrote the first paper, where these notions were introduced: “Logarithmic structures of Fontaine-Illusie”. The new theory turned out to be extremely useful because of the following features:

  • (1)

    It provides a more general notion of smoothness, which allows to work with many classical non-smooth objects similarly to the smooth ones. In particular, it conceptually adjusts various cohomology theories to this generalized context.

  • (2)

    It conceptually treats various notions of boundaries, such as normal crossings divisors, and it often provides a functorial way to compactify various moduli spaces – smooth objects often degenerate to logarithmically smooth (but non-smooth) objects over the boundary.

  • (3)

    It provides a conceptual way to bookkeep information on closed subspaces and fibers, in particular, leading to better solutions of deformation problems.

Off course these three large classes of properties are tightly connected and often show up altogether.

1.1.2. Precursors

In fact, log geometry had numerous precursors, which it absorbed and generalized. Without pretending to provide a full list, here are a few most important ones, which will be discussed in §2 in more detail:

  • (1)

    Normal crossings divisors, especially, when viewed as a boundary used to compactify a smooth variety, correspond to log structures. In fact, a smooth variety with a normal crossing divisor is nothing else but a log smooth variety, which is smooth.

  • (2)

    Toroidal geometry, which was introduced in [KKMSD73] to prove semistable reduction theorem, is, in fact, the theory of log smooth log varieties. Toroidal morphisms between toroidal varieties are nothing else but log smooth morphisms.

  • (3)

    Deligne’s generalized divisors and logarithmic structures of Deligne-Faltings, [Kat89, Complement 1]. See also [Ill15, Section 2], where Deligne’s letter and its influence on the discovery of log schemes is described.

  • (4)

    Logarithmic differentials, logarithmic versions of various complexes, etc., which were defined ad hoc, obtain a conceptual interpretation in log geometry.

  • (5)

    Semistable morphisms are log smooth, so semistable reduction theorem literally becomes a desingularization theorem in log geometry. Furthermore, the snc divisor sitting in the closed fiber of a semistable family is log smooth over the log point – a more exotic object, which bookkeeps the log structure purely algebraically. In a sense, a log point is a logarithmic analogue of usual non-reduced (or fat) points in the theory of schemes.

In fact, the role of log geometry in some other classical problems is being gradually clarified even nowadays. For example, it was clear that it is involved in resolution of singularities, at least through the exceptional divisor, but a careful study of this question not only shed a new light on known methods, but also led to discovery of a new generation of methods, which will be discussed in another chapter. Also, I think that the role of log geometry in compactifying various moduli spaces is not fully exploited yet, and it will increase in future research. This is tangential to the material we want to cover, so we will only discuss a couple of such examples in the sequel and refer the reader to [ACG+13], where the theory of log schemes is described from the point of view of applications to the theory of moduli spaces. Our main motivation is application to resolution of singularities in the chapter on logarithmic and relative resolution of singularities.

1.2. Structure of the chapter

1.2.1. Overview

The chapter starts with Section 2 on precursors of logarithmic geometry: we discuss the log structure encoded by snc divisors, which is probably the first time log structures implicitly showed up in mathematics, and then recall the more general theory of toroidal varieties. Log schemes are introduced in §3. We first study necessary properties of monoids and introduce log structures, and in the end of the section we discuss log regular log schemes. Various properties of morphisms: charts, log smoothness and log étaleness, log differentials and log blowings up are reviewed in §4. Finally, in Section 5 we discuss Olsson’s stacks 𝐋𝐨𝐠X{\mathbf{Log}}_{X} and the technique of reducing log geometry to geometry of stacks. This is the most technically demanding section, and the only one in which stacks are used. It will be used in the construction of the relative desingularization functor, but not in the more basic case of the absolute logarithmic desingularization.

1.2.2. References and sources

The single reference with all foundations worked out in detail is the recent book of Arthur Ogus [Ogu18]. Originally, logarithmic geometry was established by Kazuya Kato in [Kat89], and the theory of log regular log schemes, their desingularization and log blowings up was developed in [Kat94] and [Niz06]. Stacks 𝐋𝐨𝐠X{\mathbf{Log}}_{X} were introduced and their relation to logarithmic properties was studied by Martin Olsson in [Ols03], and some further results were obtained in [MT21].

1.2.3. Conventions

Often we write “log” instead of “logarithmic”.

1.2.4. Acknowledgments

I am very grateful to Luc Illusie for telling the story of discovery of log schemes and for reading the notes and making many helpful comments.

2. Precursors

In this section we discuss some situations, where log structures are implicit actors. Later on they will serve as a source of examples and illustrations.

2.1. Normal crossings divisors

Definition 2.1.1.

Let XX be a regular scheme and DXD\hookrightarrow X a divisor.

(i) One says that DD is strictly (or simple) normal crossings or just snc at xXx\in X if locally one has that D=V(t1ts)D=V(t_{1}\ldots t_{s}), where t1,,tnt_{1},\dots,t_{n} is a regular family of parameters at xx and we use notation V(f1,,fn)=SpecX(𝒪X/(f1,,fn))V(f_{1},\dots,f_{n})={\rm Spec}_{X}({\mathcal{O}}_{X}/(f_{1},\dots,f_{n})).

(ii) One says that DD is normal crossings at xx if it is snc étale-locally at xx.

A divisor is called normal crossings or snc if this is so everywhere on XX.

Exercise 2.1.2.

(i) Show that the number s=s(x)s=s(x) in the definition is the number of branches of DD at ss (e.g. the number of irreducible formal components) and hence is an invariant of DD at xx, also called the multiplicity of DD at xx.

(ii) Define a stratification of DD by the multiplicity and show that each stratum D(s)D(s) is a regular locally closed subscheme.

(iii) Assume that XX is regular and D=iIDiD=\cup_{i\in I}D_{i} is a reduced divisor with irreducible components DiD_{i}. Show that DD is snc if and only if for each JIJ\subseteq I the scheme theoretic intersection DJ=jJDjD_{J}=\cap_{j\in J}D_{j} is regular.

Remark 2.1.3.

Local computations with nc (resp. snc) divisors are done using étale (resp. Zariski) neighborhood, where it is given by the vanishing of the product of a subset of a family of regular parameters.

2.2. Toroidal schemes

2.2.1. Toric schemes

Let MM be a lattice and N=Hom(M,)N={\rm Hom}(M,{\mathbb{Z}}) the dual lattice. This can be encoded by the non-degenerate pairing M×NM\times N\to{\mathbb{Z}}, but despite the symmetry, we will always view NN and N=NN_{\mathbb{R}}=N\otimes{\mathbb{R}} as geometric spaces, while the elements of MM will be viewed as functions NN\to{\mathbb{Z}} or NN_{\mathbb{R}}\to{\mathbb{R}}.

Let σN\sigma\in N_{\mathbb{R}} be an MM-rational polyhedral cone, i.e. a cone given by finitely many conditions mi(x)0m_{i}(x)\geq 0 with miMm_{i}\in M. One associates to σ\sigma an affine toric kk-variety 𝐓σ{\bf T}_{\sigma} as follows. The dual cone σ={zM|z(σ)0}\sigma^{\vee}=\{z\in M_{\mathbb{R}}|\ z(\sigma)\geq 0\} is also rational and it follows easily that the monoid Mσ:=σMM_{\sigma}:=\sigma^{\vee}\cap M is finitely generated. Therefore, Aσ:=k[Mσ]A_{\sigma}:=k[M_{\sigma}] is an affine kk-algebra and 𝐓σ:=Spec(Aσ){\bf T}_{\sigma}:={\rm Spec}(A_{\sigma}) is an affine variety. Note that 𝐓σ{\bf T}_{\sigma} is provided with the natural action of the torus 𝐓0=𝐃M=Spec(k[M]){\bf T}_{0}={\bf D}_{M}={\rm Spec}(k[M]) whose lattice of characters is MM. Furthermore, 𝐓σ{\bf T}_{\sigma} contains an open orbit isomorphic to 𝐓0{\bf T}_{0}, and this is the source of the terminology. The information encoded in the pair (M,σ)(M,\sigma), often called the combinatorial information, is equivalent to the information encoded in 𝐓σ{\bf T}_{\sigma} with the torus action, and there is a very tight and natural relation between the combinatorial and geometric pictures:

Exercise 2.2.2.

(i) Show that (M,σ)(M,\sigma) can be reconstructed from the affine toric variety 𝐓σ{\bf T}_{\sigma} as follows: MM is the lattice of characters of 𝐓0{\bf T}_{0} and giving the 𝐓0{\bf T}_{0}-action on Spec(A){\rm Spec}(A) is equivalent to providing an MM-grading of AA, where mm-homogeneous elements are the equivariant ones with the character mm: they are acted on via the rule t(f)=m(t)ft(f)=m(t)f; the cone σ\sigma is determined by the monoid MσM_{\sigma}, which is precisely the set of characters mMm\in M with a non-zero homogeneous component (Aσ)m(A_{\sigma})_{m}.

(ii) Show that the set of orbits of 𝐓σ{\bf T}_{\sigma} is in a natural one-to-one correspondence with the faces of σ\sigma. In particular, the open orbit 𝐓0{\bf T}_{0} corresponds to the vertex 0 of σ\sigma.

Remark 2.2.3.

The affine theory can be globalized as follows. On the schematic part of the picture one glues affine toric varieties TiT_{i} with the same torus 𝐃M{\bf D}_{M} along isomorphic open affine toric subvarieties. A resulting object is called a toric scheme with respect to the torus DMD_{M}. On the combinatorial side one glues polyhedral cones σiN\sigma_{i}\subset N_{\mathbb{R}} along faces. The resulting objects are called cone complexes. Sometimes one only considers complexes embedded in NN_{\mathbb{R}} and calls them polyhedral fans. If the union of all faces of a fan is the whole NN_{\mathbb{R}}, the fan is called a subdivision. In fact, any separated toric variety corresponds to a fan and a proper toric variety corresponds to a subdivision. We do not go into details and recommend a standard literature, e.g. [Ful93].

2.2.4. Monoidal resolution of singularities

In the geometry of polyhedral complexes regular simplicial cones play the role of non-singular points.

Definition/Exercise 2.2.5.

(i) A polyhedral cone σ\sigma is called regular if its sharpening M¯σ=Mσ/(Mσ)×{\overline{M}}_{\sigma}=M_{\sigma}/(M_{\sigma})^{\times} is a free monoid r{\mathbb{N}}^{r} and hence Mσ~rsM_{\sigma}\widetilde{\to}{\mathbb{N}}^{r}\oplus{\mathbb{Z}}^{s}. Show that this happens if and only if there exists a basis e1,,ene_{1},\dots,e_{n} of NN such that each edge of σ\sigma contains some eie_{i}. Thus, σ\sigma is a simplicial cone of a very special form.

(ii) A polyhedral complex is called regular if all its cones are.

A combinatorial or monoidal resolution of singularities is the following result:

Theorem 2.2.6.

Any polyhedral complex possesses a regular subdivision.

Remark 2.2.7.

We will not prove this theorem, but only make a couple of remarks. One can construct such subdivision by applying functorial (hence equivariant) resolution of singularities to toric varieties and translating it to the combinatorial language. Clearly, this is a too complicated solution that does not admit a simple combinatorial interpretation. A construction of a simple canonical solution of this problem was missing in the literature until very recently, see [Wł20, Theorem 4.6.1], but various non-canonical solutions were well known. First, the claim easily reduces to the case of a single cone. Second, using the barycentric subdivision one reduces to the case of a simplicial cone. Then one defines some invariants of the singularity (essentially, they measure the discrepancy between σN\sigma\cap N and the monoid generated by the elements of NN lying on the edges) and finds simplicial subdivisions that decrease it.

2.2.8. Toroidal embeddings

Toroidal varieties (or schemes) étale-locally (or formally locally) are modelled on toric schemes. This turns out to be sufficient to extend various toric constructions, such as toric blowings up and toric resolution of singularities, to much wider context. The miracle enabling this is that these constructions pull back to the same operation on the toroidal scheme independently of the choice of a toric chart.

For simplicity we will only consider the case of varieties. There is no torus action anymore, but it turns out that a large portion of the structure can be encoded just in the open orbit:

Definition 2.2.9.

(i) A toroidal variety (or a toroidal embedding) over a field kk is a pair (X,U)(X,U) with XX a kk-variety and UXU\hookrightarrow X an open subscheme such that étale locally XX possesses an étale morphism to a toric scheme such that UU is the preimage of the torus. Namely, there exists an étale covering iXiX\coprod_{i}X_{i}\to X and étale morphisms hi:Xi𝐓σih_{i}\colon X_{i}\to{\bf T}_{\sigma_{i}}, called toroidal chart, such that U×XXiU\times_{X}X_{i} is the preimage of the torus of 𝐓σi{\bf T}_{\sigma_{i}}. If, moreover, the covering can be chosen to be Zariski, say X=iXiX=\cup_{i}X_{i}, then the toroidal variety is called simple (or without self-intersections).

(ii) A morphism of toroidal varieties (Y,V)(X,U)(Y,V)\to(X,U) is any morphism YXY\to X taking VV to UU.

Example 2.2.10.

If XX is regular, DD is a normal crossings divisor and U=XDU=X\setminus D, then (X,U)(X,U) is a toroidal variety modelled on regular simplices σ\sigma, and the free monoid MσM_{\sigma} is generated by the regular parameters which define the branches of DD (on an appropriate étale neighborhood).

2.2.11. (Non-)uniqueness of charts

A very natural question is to what extent the charts are unique. We will show below that the toric monoid is essentially unique, and the chart is unique up to units. For simplicity, assume that i:UXi\colon U\hookrightarrow X is a simple toroidal variety and set =i𝒪U×𝒪X{\mathcal{M}}=i_{*}{\mathcal{O}}_{U}^{\times}\cap{\mathcal{O}}_{X} and ¯=/𝒪X×{\overline{\mathcal{M}}}={\mathcal{M}}/{\mathcal{O}}_{X}^{\times} (in the general case one would have to use étale sheaves, as we will do in the section about log schemes). The latter sheaf is the sheaf of toroidal Cartier divisors, i.e. divisors supported on D=XUD=X\setminus U.

Exercise 2.2.12.

(i) Show that any local chart that maps xx to the closed stratum of some 𝐓σ{\bf T}_{\sigma} induces in isomorphism Mσ=¯xM_{\sigma}={\overline{\mathcal{M}}}_{x}, in particular, 𝐓σ{\bf T}_{\sigma} depends only on (X,U)(X,U) and xx.

(ii) Conversely, any section s:¯xx𝒪X,xs\colon{\overline{\mathcal{M}}}_{x}\to{\mathcal{M}}_{x}\subset{\mathcal{O}}_{X,x} of x¯x{\mathcal{M}}_{x}\twoheadrightarrow{\overline{\mathcal{M}}}_{x} can be extended to a local chart on a neighborhood of xx. (Hint: in addition to ss one should choose a regular family of parameters on the stratum of D=XUD=X\setminus U through xx and lift them to elements of 𝒪X,x{\mathcal{O}}_{X,x}.)

3. Logarithmic structures and schemes

In this section we introduce the category of log schemes and study its basic properties.

3.1. Monoids

Unless said to the contrary, by a monoid we always mean a commutative additively written monoid M=(M,+,0)M=(M,+,0). By M×M^{\times} we denote the subgroup of invertible elements and the sharpening of MM is M¯=M/M×{\overline{M}}=M/M^{\times}. One says that MM is sharp if M=M¯M={\overline{M}}.

3.1.1. Basic constructions

All categories of algebraic objects, such as groups, rings, commutative rings, etc. are complete and cocomplete – possess all small limits and colimits. Furthermore, limits are compatible with set-theoretic limits, and colimits are obtained using generators and relations. In particular, this is true for the category Mon{\rm Mon} of monoids. The main examples of limits and colimits we will use are as follows:

Exercise 3.1.2.

(i) M×NM\times N is just the usual product of sets with componentwise addition and it coincides with the coproduct, usually denoted MNM\oplus N.

(ii) For homomorphisms f:MLf\colon M\to L, g:NLg\colon N\to L the fiber product M×LNM\times_{L}N is the submonoid of M×NM\times N given by f(m)=g(n)f(m)=g(n).

(iii) Instead of kernels in the category of monoids one uses congruence relations, that is, equivalence relations RM×MR\subseteq M\times M which are also submonoids: if MNM\to N is a surjective homomorphism, then the induced equivalence relation RM×MR\subseteq M\times M is a submonoid, and conversely any congruence relation appears in this way.

(iv) For homomorphisms f:LMf\colon L\to M, g:LNg\colon L\to N the pushout MLNM\oplus_{L}N is the quotient of MNM\oplus N by the minimal congruence relation such that f(l)g(l)f(l)\sim g(l) for any lLl\in L.

3.1.3. Ideals

An ideal IMI\subseteq M is a subset such that I+M=II+M=I, where the convention is that I=I=\emptyset is also an ideal (the analogue of the zero ideal of rings). An ideal II is prime if x+yIx+y\in I implies that xIx\in I or yIy\in I. The set of all prime ideals is denoted Spec(M){\rm Spec}(M) and called the fan of MM. An ideal of the form (a)=a+M(a)=a+M is called principal.

Exercise 3.1.4.

Show that taking the preimage establishes a bijection between the ideals of MM and M¯{\overline{M}}.

3.1.5. Fine monoids

We will usually work with fine monoids:

Definition 3.1.6.

(i) MM is finitely generated if there exists a surjective homomorphism lM{\mathbb{N}}^{l}\to M.

(ii) MM is integral (or cancellative) if n+m=n+mn+m=n+m^{\prime} implies that m=mm=m^{\prime}.

(iii) MM is fine if it is integral and finitely generated.

3.1.7. The Grothendieck group

Recall that there is a canonical way to turn monoid into a group:

Definition/Exercise 3.1.8.

(i) Show that there is a universal homomorphism MMgpM\to M^{\rm gp} with MgpM^{\rm gp} a group, called the Grothendieck group of MM. (Hint: for example, one can bound the cardinality of MgpM^{\rm gp} because it is generated by the image of MM, and then general representability theorems do the job because the category of groups is complete.)

(ii) Construct MgpM^{\rm gp} explicitly as the quotient of M2M^{2} by the following equivalence relation: (m,n)~(m,n)(m,n)\widetilde{\to}(m^{\prime},n^{\prime}) if there exists lMl\in M such that l+m+n=l+m+nl+m+n^{\prime}=l+m^{\prime}+n.

(iii) Show that MM is integral if and only if the homomorphism MMgpM\to M^{\rm gp} is injective.

(iv) Define the integralization MintM^{\rm int} to be the image of MM in MgpM^{\rm gp}. Show that MMintM\to M^{\rm int} is the universal homomorphism from MM to an integral monoid and MMintM\mapsto M^{\rm int} is the left adjoint functor to the embedding of the category of integral monoids Monint{\rm Mon}^{\rm int} into Mon{\rm Mon}.

3.1.9. Fs monoids

An especially nice class of monoids is defined as follows:

Definition 3.1.10.

(i) An integral monoid MM is saturated if for each mMgpm\in M^{\rm gp} with amMam\in M for m>0m\in{\mathbb{Z}}_{>0} one has that mMm\in M.

(ii) A fine saturated monoid is called fs.

(iii) A sharp fs monoid is called toric.

Exercise 3.1.11.

(i) Show that to give a toric monoid MM is equivalent to give a lattice MgpM^{\rm gp} and a rational polyhedral cone σ\sigma in MgpM^{\rm gp}\otimes{\mathbb{R}} such that M=MgpσM=M^{\rm gp}\cap\sigma.

(ii) Show that Spec(M){\rm Spec}(M) is naturally bijective to the set of faces of σ\sigma.

Exercise 3.1.12.

Let Monsat{\rm Mon}^{\rm sat} denote the category of saturated monoids. Show that the embedding MonsatMon{\rm Mon}^{\rm sat}\hookrightarrow{\rm Mon} possesses a left adjoint functor, which is called the saturation functor and denoted MMsatM\mapsto M^{\rm sat}. Show that MsatM^{\rm sat} is just the saturation of MintM^{\rm int} in MgpM^{\rm gp}, that is the divisible hull of MintM^{\rm int} in MgpM^{\rm gp}.

3.2. Logarithmic structures

Definition 3.2.1.

Let τ\tau be one of the following topologies – Zariski, étale or flat.

(i) A τ\tau-prelogarithmic structure on a scheme XX is a sheaf of monoids {\mathcal{M}} on the site XτX_{\tau} with a structure homomorphism of monoids u:(𝒪X,)u:{\mathcal{M}}\to({\mathcal{O}}_{X},\cdot). A homomorphism of prelog structures {\mathcal{M}}\to{\mathcal{M}}^{\prime} is a homomorphism of sheaves of monoids compatible with the structure homomorphisms.

(ii) A τ\tau-logarithmic structure is a τ\tau-prelogarithmic structure which induces an isomorphism u1(𝒪Xτ×)~𝒪Xτ×u^{-1}({\mathcal{O}}_{X_{\tau}}^{\times})\widetilde{\to}{\mathcal{O}}_{X_{\tau}}^{\times}, and hence also ×~𝒪Xτ×{\mathcal{M}}^{\times}\widetilde{\to}{\mathcal{O}}_{X_{\tau}}^{\times}. The sharpening ¯=/𝒪Xτ×{\overline{\mathcal{M}}}={\mathcal{M}}/{\mathcal{O}}_{X_{\tau}}^{\times} is called the characteristic monoid of {\mathcal{M}}.

(iii) The default topology in this definition is the étale topology, so usually it will not be mentioned. A log structure {\mathcal{M}} induces a Zariski log structure Zar{\mathcal{M}}_{\rm Zar} just by restricting. By a slight abuse of language we say that {\mathcal{M}} itself is Zariski if this restriction does not loose information, that is, =ε(Zar){\mathcal{M}}=\varepsilon^{*}({\mathcal{M}}_{\rm Zar}) for the morphism of sites ε:XetXZar\varepsilon\colon X_{\rm et}\to X_{\rm Zar}.

Remark 3.2.2.

(i) The homomorphism uu is an analog of exponentiation, and this is one of the reasons to use additive monoids as the source. Traditionally it is denoted α{\alpha}, say α(m)=x{\alpha}(m)=x, but probably the exponential notation, such as x=umx=u^{m} instead of u(m)u(m), is more suggestive. Informally, any such mm can be viewed as a branch of the logarithm of xx, so the log structure can be viewed as fixing a monoid of branches of logarithms.

(ii) The étale topology is used instead of the Zariski topology first of all in order to adequately treat toroidal embeddings, which are not strict. As a rule, this might pose mild technical inconveniences, which can be bypassed. For example, Kato restricted the generality to Zariski log structures in [Kat94], but these results were generalized to the general case in [Niz06].

(iii) Fppf log structures are sometimes needed to bypass positive characteristic problems, usually by extracting appropriate pp-th roots. They are rarely used and will not show up in these notes.

To get an initial feeling let us consider some examples.

Example 3.2.3.

(0) The minimal or trivial log structure is just =𝒪Xet×{\mathcal{M}}={\mathcal{O}}_{X_{\rm et}}^{\times}.

(1) The largest log structure with an injective uu is =𝒪Xet{\mathcal{M}}={\mathcal{O}}_{X_{\rm et}}, but it is not too useful. The main exception is when XX is a “small” scheme – a semi-local curve or the spectrum of a valuation ring, with the most useful case being when XX is a trait.

(2) A very important example of a log structure with an injective u:𝒪Xu\colon{\mathcal{M}}\to{\mathcal{O}}_{X} is as follows. Assume that DXD\hookrightarrow X is a closed subset and set (logD)=𝒪Xeti𝒪Uet×{\mathcal{M}}({\rm log}D)={\mathcal{O}}_{X_{\rm et}}\cap i_{*}{\mathcal{O}}^{\times}_{U_{\rm et}} where i:UXi\colon U\hookrightarrow X is the open immersion of the complement U=XDU=X\setminus D. This is the log structure of elements invertible outside of DD. Usually it is used when DD is the underlying closed set of a Cartier divisor; in this case DD is determined by the log structure and we call (logD){\mathcal{M}}({\rm log}D) a divisorial log structure. In particular, the toroidal scheme structure (X,U)(X,U) can also be encoded in the log structure of DD-monomial elements, where D=XUD=X\setminus U.

(3) The other extreme case is provided by so-called hollow log schemes with um=0u^{m}=0 for any m×m\in{\mathcal{M}}\setminus{\mathcal{M}}^{\times}. Usually they show up when one restricts a log structure on XX onto a closed subscheme ZZ such that X{\mathcal{M}}_{X} is generically non-trivial on ZZ. Often this is the most economical way to encode certain information about the ambient scheme XX on ZZ. In particular, this is useful in deformation theory. In fact, it was such kind of an example, and not toroidal schemes, which led Fontaine and Illusie to introduce log schemes.

3.2.4. Associated log structure

Any prelog structure {\mathcal{M}} can be canonically transformed into a log structure a{\mathcal{M}}^{a}. The idea is that {\mathcal{M}} is a log structure if and only if u1(𝒪Xet×)~𝒪Xet×u^{-1}({\mathcal{O}}_{X_{\rm et}}^{\times})\widetilde{\to}{\mathcal{O}}_{X_{\rm et}}^{\times} so we should force this map to be an isomorphism. This works for any topology, so for shortness we only consider the étale one.

Exercise 3.2.5.

Given a prelog structure {\mathcal{M}} on XX let a{\mathcal{M}}^{a} be the pushout of the diagram

u1(𝒪Xet)×𝒪Xet×.{\mathcal{M}}\hookleftarrow u^{-1}({\mathcal{O}}_{X_{\rm et}})^{\times}\to{\mathcal{O}}_{X_{\rm et}}^{\times}.

Show that a{\mathcal{M}}^{a} is a log structure, a{\mathcal{M}}\to{\mathcal{M}}^{a} is the universal homomorphism of {\mathcal{M}} to a log structure and the functor a{\mathcal{M}}\mapsto{\mathcal{M}}^{a} is left adjoint to the embedding of the category of log structures into the category of prelog structures.

Remark 3.2.6.

One can view the functor a{\mathcal{M}}\mapsto{\mathcal{M}}^{a} as an analogue of sheafification. Various operations on sheaves are defined by applying sheafification to the analogous operation in the category of presheaves (the sheafification is not needed for right exact functors, but is usually needed for other functors). In the same vein, various naive operations on log structures result in a prelog structure only (one works with sheaves, so the usual sheafification is used), and the functor a{\mathcal{M}}\mapsto{\mathcal{M}}^{a} should then be applied.

Definition/Exercise 3.2.7.

Integralization int{\mathcal{M}}^{\rm int} and saturation sat{\mathcal{M}}^{\rm sat} of a log structure {\mathcal{M}} are defined by applying the corresponding functors to monoids of sections, sheafification and then applying the functor aa. Check that the resulting logarithmic structure is indeed saturated or integral.

3.2.8. Coherent log structures

One reason to consider prelog structures was already discussed – they form a simpler category and various operations on log structures often have prelog structures as intermediate results. Another reason is that prelog structures are used as charts (see §3.3.11 below), usually of finite type, for log structures, which are often very large because of the invertible part.

Definition 3.2.9.

(i) A prelog structure is called constant if it is the sheafification of a homomorphism PΓ(𝒪X)P\to\Gamma({\mathcal{O}}_{X}) for a monoid PP. For shortness we will not distinguish the monoid PP and its sheafification.

(ii) A log structure {\mathcal{M}} is called quasi-coherent if étale-locally there exists a constant prelog structure PP such that =Pa{\mathcal{M}}=P^{a}. If, in addition, PP can be chosen finitely generated, {\mathcal{M}} is called coherent. We warn the reader that this notion is not related to coherence and quasi-coherence of 𝒪X{\mathcal{O}}_{X}-modules.

(iii) A coherent and integral (resp. saturated) logarithmic structure is called fine (resp. fs).

The following result can be found, for example, in [Ogu18, Corollary II.2.3.6]

Exercise 3.2.10.

Show that a log structure is fine (resp. saturated) if and only if étale locally it possesses a fine (resp. fs) chart P𝒪XP\to{\mathcal{O}}_{X}.

Finally, for a morphism f:YXf\colon Y\to X one defines direct and inverse images of the log structures {\mathcal{M}} on XX and 𝒩{\mathcal{N}} on YY:

Definition/Exercise 3.2.11.

(i) Show that f(𝒩)×f(𝒪Y)𝒪Xf_{*}({\mathcal{N}})\times_{f_{*}({\mathcal{O}}_{Y})}{\mathcal{O}}_{X} is a log structure denoted (by a slight abuse of notation) f(𝒩)f_{*}({\mathcal{N}}).

(ii) The inverse image f()f^{*}({\mathcal{M}}) is the log structure associated to the prelog structure f1()f1(𝒪X)𝒪Yf^{-1}({\mathcal{M}})\to f^{-1}({\mathcal{O}}_{X})\to{\mathcal{O}}_{Y}. Show that, as expected, ff^{*} is left adjoint to ff_{*}.

3.3. Logarithmic schemes

Now we can introduce the category of log schemes.

Definition 3.3.1.

(i) A log scheme XX is a tuple (X¯,X,uX)({\underline{X}},{\mathcal{M}}_{X},u_{X}), where X¯{\underline{X}} is a scheme called the underlying scheme and uX:X𝒪X¯u_{X}\colon{\mathcal{M}}_{X}\to{\mathcal{O}}_{\underline{X}} is a log structure on XX. Usually we will use looser notation when this cannot lead to a confusion, e.g. write uu instead of uXu_{X} or just omit it, write 𝒪X{\mathcal{O}}_{X} instead of 𝒪X¯{\mathcal{O}}_{\underline{X}} or even denote the underlying scheme by the same letter XX.

(ii) A log scheme XX is called quasi-coherent, coherent, integral, fine, fs, etc., if the log structure X{\mathcal{M}}_{X} is quasi-coherent, coherent, integral, fine, fs, etc.

(iii) A morphism of log schemes f:YXf\colon Y\to X consists of the underlying morphism f¯:Y¯X¯{\underline{f}}\colon{\underline{Y}}\to{\underline{X}} of schemes and a homomorphism of log structures fXYf^{*}{\mathcal{M}}_{X}\to{\mathcal{M}}_{Y} compatible with the structure homomorphisms uXu_{X} and uYu_{Y}.

Remark 3.3.2.

(i) We warn the reader that the word “integral” becomes slightly overused, because the notion of integral schemes means something completely different in the theory of schemes. So one should use it carefully, to avoid misunderstandings. Typically, if needed, one stresses that the underlying scheme is integral.

(ii) Already in [Kat89] Kato noticed that non-integral log schemes are too pathological and mainly restricted consideration to the category of fine log schemes. Although some aspects of a more general theory were developed quite systematically in [Ogu18], most of studies are done in the generality of fine log schemes, and this indeed seems to be the best choice. The second popular choice, which is often used once one works with log blowings up, is to work with the subcategory of fs log schemes. One benefit of working with these categories is that integral or saturated fiber products often reveal a nicer behaviour. In particular, log blowings up are log étale monomorphisms, see Exercise 4.3.7 and this fact essentially uses integralization.

Remark 3.3.3.

(i) In case of a log structure with an injective Y𝒪Y{\mathcal{M}}_{Y}\hookrightarrow{\mathcal{O}}_{Y} a morphism f:YXf\colon Y\to X is uniquely determined by the underlying morphism of schemes, and f¯{\underline{f}} extends to ff if the log structure on YY is “larger” than the image of the log structure on XX. If Y=(logE){\mathcal{M}}_{Y}={\mathcal{M}}({\rm log}E) and X=(logD){\mathcal{M}}_{X}={\mathcal{M}}({\rm log}D) are divisorial, then this happens if and only if f1(D)Ef^{-1}(D)\subseteq E.

(ii) In particular, providing a toroidal scheme (X,U)(X,U) with the divisorial log structure X=(log(XU)){\mathcal{M}}_{X}={\mathcal{M}}({\rm log}(X\setminus U)) we obtain a fully faithful embedding of the category of toroidal schemes into the category of log schemes. Naturally, such (X,X)(X,{\mathcal{M}}_{X}) will be called a toroidal log scheme.

(iii) The other extreme is when the log schemes are hollow: uX=0u_{X}=0, uY=0u_{Y}=0. In this case, f¯{\underline{f}} extends to ff via any homomorphism f(X)Yf^{*}({\mathcal{M}}_{X})\to{\mathcal{M}}_{Y}.

3.3.4. Strict morphisms

A very important class of morphisms are those that “minimally modify the log structure”. In the case of divisorial log structures this just means that E=f1(D)E=f^{-1}(D) and in general:

Definition 3.3.5.

A morphism f:YXf\colon Y\to X of log schemes is called strict if f(X)=Yf^{*}({\mathcal{M}}_{X})={\mathcal{M}}_{Y}.

In particular, for a log scheme XX any morphism Y¯X¯{\underline{Y}}\to{\underline{X}} can be uniquely enhanced to a strict morphism of log schemes.

Remark 3.3.6.

Any morphism YXY\to X canonically factors into a composition YgZhXY\stackrel{{\scriptstyle g}}{{\to}}Z\stackrel{{\scriptstyle h}}{{\to}}X such that hh is strict and Y¯=Z¯{\underline{Y}}={\underline{Z}}. It is natural to view hh as a scheme-like morphism and gg as a morphism which only increases the log structure. However, this factorization is not especially useful. The reason is that gg is not a “monoidal-like” morphism, see §4.1.3 below.

Example 3.3.7.

(i) In the category of log schemes strict closed immersions play the role of usual closed immersions in the category of schemes. Often strict closed immersions have hollow sources. A typical example is when XX is a toroidal log scheme and Y¯X¯{\underline{Y}}\hookrightarrow{\underline{X}} is a toroidal subscheme.

(ii) The simplest and most important case is when XX is the affine line with marked origin (i.e. X¯=Spec(k[t]){\underline{X}}={\rm Spec}(k[t]) and X=(logt)a{\mathcal{M}}_{X}=({\mathbb{N}}\log t)^{a}, where we denote the generator of the monoid by logt\log t to stress that it is mapped to tt by uXu_{X}) and y¯=Spec(k[t]/(t)){\underline{y}}={\rm Spec}(k[t]/(t)) is the origin. The induced log structure is y=k×logt{\mathcal{M}}_{y}=k^{\times}\oplus{\mathbb{N}}\log t and we call (y,y)(y,{\mathcal{M}}_{y}) the standard log point. It is an analogue of points with non-reduced scheme structure in the usual algebraic geometry, in particular, we will later see that it is not log smooth and has non-trivial log differentials coming from the log direction logt\log t.

Note also that any other point x=Spec(k[t]/(ta))x={\rm Spec}(k[t]/(t-a)), ak×a\in k^{\times} has the trivial induced log structure since logt{\mathbb{N}}\log t is mapped to 𝒪x×{\mathcal{O}}_{x}^{\times} and hence the functor a{\mathcal{M}}\mapsto{\mathcal{M}}^{a} shrinks k×logtk^{\times}\oplus{\mathbb{N}}\log t to x=k×{\mathcal{M}}_{x}=k^{\times} by sending logt\log t to aa.

(iii) Similarly, for each n1n\geq 1 the thick point Y¯n=Spec(k[t]/tn+1){\underline{Y}}_{n}={\rm Spec}(k[t]/t^{n+1}) can be provided with the log structure induced by logt{\mathbb{N}}\log t. It is neither injective, nor hollow.

(iv) One can also consider other log points, for example, for a toric monoid PP the log structure induced by P0kP\stackrel{{\scriptstyle 0}}{{\to}}k corresponds to the origin of the toric scheme Spec(k[P]){\rm Spec}(k[P]).

3.3.8. Log rings

For a local work with log schemes it is often convenient to use the following logarithmic spectrum construction.

Definition 3.3.9.

(i) A log ring (A,P,u)(A,P,u) consists of a ring AA, a monoid PP and a homomorphism u:P(A,)u\colon P\to(A,\cdot).

(ii) The logarithmic spectrum of a log ring, usually denoted just by X=Spec(PuA)X={\rm Spec}(P\stackrel{{\scriptstyle u}}{{\to}}A) is the underlying scheme X¯=Spec(A){\underline{X}}={\rm Spec}(A) provided with the log structure induced by the prelog structure uu.

Remark 3.3.10.

(i) In a sense this is a prelog ring, but the notion of a log ring does not make too much sense - even if P=Γ(X)P=\Gamma({\mathcal{M}}_{X}), this condition will be lost after localizations.

(ii) A log scheme is Zariski if and only if it is a spectrum of log rings Zariski-locally. For general log schemes this is only true étale-locally, so in this aspect they are analogous to algebraic spaces.

3.3.11. Charts

The notion of charts of log structures can be adopted to the category of log schemes once one replaces a monoid PP by the associated log scheme.

Definition/Exercise 3.3.12.

(i) For a monoid PP let 𝐀P{\bf A}_{P} denote the log scheme whose underlying scheme is Spec([P]){\rm Spec}({\mathbb{Z}}[P]) and the log structure is induced by the homomorphism P[P]P\hookrightarrow{\mathbb{Z}}[P]. If we work over a base scheme, for example, S=Spec(k)S={\rm Spec}(k), then we will use the notation 𝐀S,P=S[P]:=S×[P]{\bf A}_{S,P}=S[P]:=S\times{\mathbb{Z}}[P].

(ii) Let XX be a log scheme. Show that giving a global chart ϕ:P𝒪X\phi\colon P\to{\mathcal{O}}_{X} for the log structure X{\mathcal{M}}_{X} is equivalent to giving a strict morphism of log schemes f:X𝐀Pf\colon X\to{\bf A}_{P}. Any such morphism ff is called a global chart of XX, and we will not make a real distinction between two presentations of a chart. A chart is called fine, saturated, sharp, etc. if the monoid PP is fine, saturated, sharp, etc. In particular, check that a coherent (resp. fine, resp. fs) log scheme is a log scheme which étale-locally possesses a finitely generated (resp. fine, resp. fs) chart.

Remark 3.3.13.

It is important to consider charts with non-sharp PP because various operations can produce non-sharp monoids. For example, removing the origin from S[]=Spec(k[t])S[{\mathbb{N}}]={\rm Spec}(k[t]), where S=Spec(k)S={\rm Spec}(k), one obtains the scheme S[]=Spec(k[t±1])S[{\mathbb{Z}}]={\rm Spec}(k[t^{\pm 1}]), which is also a chart of itself. On the other hand, its log structure is trivial, hence S[]SS[{\mathbb{Z}}]\to S is also a chart.

Nevertheless, when working locally at a point one might want to take a smallest possible chart and often this is possible. The follwoing notion is due to Kato:

Definition 3.3.14.

A chart P𝒪XP\to{\mathcal{O}}_{X} is called neat at a geometric point x¯X{\overline{x}}\to X if P~¯x¯P\widetilde{\to}{\overline{\mathcal{M}}}_{\overline{x}}. In particular, PP is automatically sharp.

Example 3.3.15.

Let X=𝐀k,=Spec(k[t])X={\bf A}_{k,{\mathbb{N}}}={\rm Spec}(k[t]). Then the tautological chart logt𝒪X{\mathbb{N}}\log t\to{\mathcal{O}}_{X} is neat at the origin, but not at the other points, where a smaller chart (the trivial one) exists.

The following example demonstrates one of rare aspects, in which fppf fine log schemes behave nicer than their étale analogues. However, even this is only needed when the log schemes are only fine but not fs. The following results can be found, for example, in [Ogu18, §II.2.3].

Exercise 3.3.16.

Let XX be a fine log scheme, let x¯X¯{\overline{x}}\to{\underline{X}} be a geometric point and let P=¯X,x¯P={\overline{\mathcal{M}}}_{X,{\overline{x}}}. Any neat chart at x¯{\overline{x}} gives rise to a section PX,x¯P\to{\mathcal{M}}_{X,{\overline{x}}} of the sharpening homomorphism.

(i) Show, that conversely any section s:PX,x¯s\colon P\to{\mathcal{M}}_{X,{\overline{x}}} of X,x¯P{\mathcal{M}}_{X,{\overline{x}}}\to P induces a chart for a small enough étale neighborhood of x¯{\overline{x}} and this chart is neat at x¯{\overline{x}}.

(ii) Furthermore, show that such a section exists if and only if the sequence

1𝒪Xet,x¯×X,x¯gpPgp01\to{\mathcal{O}}^{\times}_{X_{\rm et},{\overline{x}}}\to{\mathcal{M}}^{\rm gp}_{X,{\overline{x}}}\to P^{\rm gp}\to 0

splits, and this is automatic whenever PgpP^{\rm gp} has no torsion of order divisible by p=char(k(x¯))p={\rm char}(k({\overline{x}})) ([Ogu18, Proposition II.2.3.7]).

(iii) Show that any fine fppf log scheme admits neat charts fppf-locally. (Hint: using the fppf topology one can also extract roots of order pp.)

(iv) Show that if the log structure is fs, then PgpP^{\rm gp} is automatically torsion free, and hence even Zariski log schemes possess neat charts locally at points of X¯{\underline{X}}.

3.3.17. Monoidal morphisms

We say that a morphism of log schemes YXY\to X is monoidal if étale-locally on XX it is the base change of morphisms of the form 𝐀Q𝐀P{\bf A}_{Q}\to{\bf A}_{P}. Informally speaking, YY is obtained by first changing the monoidal structure and then adjusting the underlying scheme in the minimal needed way. Also, such morphisms can be viewed as base changes of morphisms of fans of monoids, e.g. see the informal notation [Kat94, Definition 9.10]. Main examples of such morphisms that are integralization, saturation, log blowings up and the morphism 𝐋𝐨𝐠XX{\mathbf{Log}}_{X}\to X. They all be discussed later, and we start with the first two.

Exercise 3.3.18.

(i) Show that for any coherent log scheme XX there exists a universal morphism XXintX\to X^{\rm int} (resp. XXsatX\to X^{\rm sat}) whose target is a fine (resp. fs) log scheme. In other words, the functor XXintX\mapsto X^{\rm int} (resp. XXsatX\mapsto X^{\rm sat}) is left adjoint to the embedding of the category of fine (resp. fs) integral schemes into the category of coherent log schemes. In addition, XXintX\to X^{\rm int} is a closed immersion and XXsatX\to X^{\rm sat} is finite. (Hint: first, assuming that XX possesses a chart XSpec([P])X\to{\rm Spec}({\mathbb{Z}}[P]) show that Xint=X[P][Pint]X^{\rm int}=X\otimes_{{\mathbb{Z}}[P]}{\mathbb{Z}}[P^{\rm int}] and Xsat=X[P][Psat]X^{\rm sat}=X\otimes_{{\mathbb{Z}}[P]}{\mathbb{Z}}[P^{\rm sat}] are as required. In general, use local-étale charts and étale descent of finiteness.)

(ii) Show that for any strict morphism of coherent schemes YXY\to X one has that Yint=Xint×XYY^{\rm int}=X^{\rm int}\times_{X}Y and Ysat=Xsat×XYY^{\rm sat}=X^{\rm sat}\times_{X}Y.

3.3.19. Fiber products

Similarly to the category of schemes, the category of log schemes and its subcategories possess all finite limits, of which we will use only fiber products.

Exercise 3.3.20.

(i) Let {Xi}iI\{X_{i}\}_{i\in I} be a finite diagram of log schemes. Show that X=limIXiX=\lim_{I}X_{i} exists and can be described as follows: X¯=limIX¯i{\underline{X}}=\lim_{I}{\underline{X}}_{i} and the log structure X{\mathcal{M}}_{X} is the colimit of the pullbacks of Xi{\mathcal{M}}_{X_{i}} to X{\mathcal{M}}_{X}. (In particular, this involves the functor aa; the shortest way is to pullback Xi{\mathcal{M}}_{X_{i}} as prelog structures, then take the colimit, and then apply aa once.)

(ii) Show that if all XiX_{i} are fine or fs, then there exists a limit in the same category, and it coincides with XintX^{\rm int} or XsatX^{\rm sat}, respectively.

Remark 3.3.21.

Often one mentions int{\rm int} or sat{\rm sat} in the notation of the limit, e.g. Y×XsatZY\times_{X}^{\rm sat}Z or (Y×XZ)int(Y\times_{X}Z)^{\rm int}, to avoid confusions. Sometimes, when only fs (resp. fine) log schemes are considered, this superscript can be omitted.

To feel how this works, we start with the following almost tautological fact.

Exercise 3.3.22.

If SS is a scheme, PQP\to Q, PRP\to R are homomorphisms of finitely generated monoids and X=S[P]X=S[P], Y=S[Q]Y=S[Q], Z=S[R]Z=S[R], then Y×XZ=S[QPR]Y\times_{X}Z=S[Q\oplus_{P}R], (Y×XZ)int=S[(QPR)int](Y\times_{X}Z)^{\rm int}=S[(Q\oplus_{P}R)^{\rm int}] and (Y×XZ)sat=S[(QPR)sat](Y\times_{X}Z)^{\rm sat}=S[(Q\oplus_{P}R)^{\rm sat}].

Now we can consider an archetypical example of a subtle behaviour of fine log schemes, as opposed to coherent log schemes or schemes.

Exercise 3.3.23.

(i) Let X=Spec(k[x,y])X={\rm Spec}(k[x,y]) with the log structure induced by log(x)log(y){\mathbb{N}}{\rm log}(x)\oplus{\mathbb{N}}{\rm log}(y), and let Y=Spec(k[x,yx])Y={\rm Spec}(k[x,\frac{y}{x}]) with the log structure induced by log(x)(log(y)log(x)){\mathbb{N}}{\rm log}(x)\oplus{\mathbb{N}}({\rm log}(y)-{\rm log}(x)). Show that Y=Y×XintYY=Y\times^{\rm int}_{X}Y, in particular, the integralization functor cuts off the 𝔸2{\mathbb{A}}^{2} component from Y¯×X¯Y¯{\underline{Y}}\times_{\underline{X}}{\underline{Y}}.

(ii) More generally, assume that PQP\subsetneq Q are toric monoids such that Pgp=QgpP^{\rm gp}=Q^{\rm gp}. Then 𝐀Q=𝐀Q×𝐀Pint𝐀Q{\bf A}_{Q}={\bf A}_{Q}\times^{\rm int}_{{\bf A}_{P}}{\bf A}_{Q}, that is, the morphism 𝐀Q𝐀P{\bf A}_{Q}\to{\bf A}_{P} is a monomorphism in the category of fine log schemes, but not in the category of coherent log schemes.

Remark 3.3.24.

Usual blowing up have some nasty properties. For concreteness, consider the blowup up chart f:Y¯=Spec(k[x,y=yx])X¯=Spec(k[x,y])f\colon{\underline{Y}}={\rm Spec}(k[x,y^{\prime}=\frac{y}{x}])\to{\underline{X}}={\rm Spec}(k[x,y]) at the origin. Clearly, ff is non-flat, the dimension of the fiber over the origin jumps, and this even gives rise to the new irreducible component Spec(k[y1,y2]){\rm Spec}(k[y^{\prime}_{1},y^{\prime}_{2}]) in Y¯×X¯Y¯{\underline{Y}}\times_{\underline{X}}{\underline{Y}}. Assume now that X¯{\underline{X}} and Y¯{\underline{Y}} are provided with the log structures generated by log(x),log(y)\log(x),\log(y) and log(x),log(y)\log(x),\log(y^{\prime}). Then the 𝔸2{\mathbb{A}}^{2} component in the product acquires a non-cancellative log structure with monoid presented by generators log(x),log(y1),log(y2)\log(x),\log(y^{\prime}_{1}),\log(y^{\prime}_{2}) and relation log(x)+log(y1)=log(y)=log(x)+log(y2)\log(x)+\log(y^{\prime}_{1})=\log(y)=\log(x)+\log(y^{\prime}_{2}). This forces the integralization functor to remove the 𝔸2{\mathbb{A}}^{2} component and only its diagonal, which is the intersection with the other component, is left. In fact, we will later see that many similar morphisms (log blowings up and their charts) are monomorphisms in the fine category.

And here is an archetypical example of a property of the saturated category. It explains why one usually restricts the setting to fs log schemes when studying Kummer covers.

Exercise 3.3.25.

(i) Let X=Spec(k[x])X={\rm Spec}(k[x]) with the log structure log(x){\mathbb{N}}\log(x) and Y=Spec(k[x1/2])Y={\rm Spec}(k[x^{1/2}]) with the log structure 12log(x)\frac{1}{2}{\mathbb{N}}\log(x). We will later see that the Kummer cover YXY\to X is log étale when char(k)2{\rm char}(k)\neq 2. Check that Z=Y×XY=(Y×XY)intZ=Y\times_{X}Y=(Y\times_{X}Y)^{\rm int} contains two components (diagonal and antidiagonal) intersecting over the origin and the characteristic ¯z{\overline{\mathcal{M}}}_{z} at the intersection point zz is the non saturated monoid 1212\frac{1}{2}{\mathbb{N}}\oplus_{\mathbb{N}}\frac{1}{2}{\mathbb{N}} with generators (12,0),(0,12)(\frac{1}{2},0),(0,\frac{1}{2}) subject to the relation (1,0)=(0,1)(1,0)=(0,1). It is isomorphic to the submonoid of /2{\mathbb{Z}}/2{\mathbb{Z}}\oplus{\mathbb{N}} obtained by removing the element (1,0)(1,0).

(ii) Show that ZsatZ^{\rm sat} is the normalization of ZZ and it is just the disjoit union of the two copies of YY – the diagonal and the antidiagonal. The characteristic of the two points over zz is {\mathbb{N}} – the sharpening of ¯zsat~/2{\overline{\mathcal{M}}}_{z}^{\rm sat}\widetilde{\to}{\mathbb{Z}}/2{\mathbb{Z}}\oplus{\mathbb{N}}.

(iii) More generally, if SS is a scheme, PQP\subseteq Q is a Kummer extension of toric monoids, i.e. QQ is the saturation of PP in QgpQ^{\rm gp}, and X=S[P],Y=S[Q]X=S[P],Y=S[Q], then Z=Y×XsatYZ=Y\times_{X}^{\rm sat}Y is isomorphic to the split cover Y×GY\times G for G=Qgp/PgpG=Q^{\rm gp}/P^{\rm gp}. So, the log étale GG-Galois cover S[Q]S[P]S[Q]\to S[P] behaves similarly to étale covers only in the category of fs log schemes.

3.4. Logarithmic regularity

In this section we only consider fs log schemes.

3.4.1. The logarithmic stratification

Each log scheme XX possesses a natural stratification by rk(¯x)=dim(¯xgp){\rm rk}({\overline{\mathcal{M}}}_{x})={\rm dim}_{\mathbb{Q}}({\overline{\mathcal{M}}}_{x}^{\rm gp}\otimes{\mathbb{Q}}).

Exercise 3.4.2.

(i) Let CdX¯C_{\geq d}\subseteq{\underline{X}} be the locus on which the rank of the characteristic monoid is at most dd. Show that these sets are closed and hence induce a stratification by the locally closed sets Cd=CdC>dC_{d}=C_{\geq d}\setminus C_{>d}.

(ii) Refine this stratification to a log stratification X¯d{\underline{X}}_{d} of XX by (non-necessarily reduced) locally closed subschemes as follows: if the structure is Zariski at xx and d=rk(¯x)d={\rm rk}({\overline{\mathcal{M}}}_{x}), then the stratum X¯d{\underline{X}}_{\geq d} at xx is given by the vanishing of ux+u^{{\mathcal{M}}_{x}^{+}}, where x+{\mathcal{M}}_{x}^{+} is the maximal ideal of x{\mathcal{M}}_{x}. Show that this is compatible with strict morphisms and hence extends to arbitrary log structures by étale descent. Finally, set Xd=XdX>dX_{d}=X_{\geq d}\setminus X_{>d}

(iii) Show that, indeed, CdC_{\geq d} is the reduction of XdX_{\geq d}. Also, show that the log strata of the chart log schemes 𝐀P{\bf A}_{P} are reduced.

3.4.3. Logarithmic regularity

The following definition is a far-reaching generalization of the classical fact recalled in Exercise 2.1.2(iii).

Definition 3.4.4.

A locally noetherian fs logarithmic scheme XX is logarithmically regular if each locally closed subscheme X¯r=X¯rX¯r+1{\underline{X}}_{r}={\underline{X}}_{\geq r}\setminus{\underline{X}}_{\geq r+1} is regular (in particular, reduced) and of codimension rr.

Remark 3.4.5.

The original definition by Kato only considered fs log schemes. Most of results about log regularity can be extended to fine log schemes, and this was worked out by Gabber, but we will not touch this direction in the notes.

3.4.6. Log parameters

As in the case of regular schemes, when working with log regular log schemes it is very convenient to use local parameters. The classical notion is generalized as follows:

Definition 3.4.7.

Let XX be a log regular log scheme with a geometric point x¯X{\overline{x}}\to X over xXx\in X. Let r=rk(¯x¯)r={\rm rk}({\overline{\mathcal{M}}}_{\overline{x}}). By a regular family of parameters at x¯{\overline{x}} we mean a section s:¯x¯x¯s\colon{\overline{\mathcal{M}}}_{\overline{x}}\to{\mathcal{M}}_{\overline{x}} and elements t1,,td𝒪xt_{1},\dots,t_{d}\in{\mathcal{O}}_{x}, such that t1,,tdt_{1},\dots,t_{d} restrict to a regular family of parameters of X¯r{\underline{X}}_{r} at xx. In particular, dd is the dimension of X¯r{\underline{X}}_{r} at xx and r+d=dimx(X)r+d={\rm dim}_{x}(X). We call tit_{i} regular parameters and the elements us(m)u^{s(m)} for 0m¯x0\neq m\in{\overline{\mathcal{M}}}_{x} will be called logarithmic parameters.

In view of Exercise 3.3.16(iv) such families of parameters exist, and if the log structure at xx is Zariski, one can even construct ss Zariski locally. As in the classical case, it follows from the definition that any regular family of parameters generates the maximal ideal at x¯{\overline{x}}. Furthermore, parameters naturally give rise to very explicit étale and formal charts.

Exercise 3.4.8.

Assume that a log scheme XX is Zariski and log regular at a point xXx\in X with P=¯xP={\overline{\mathcal{M}}}_{x}, and let s:Pxs\colon P\to{\mathcal{M}}_{x} and t1,,td𝒪xt_{1},\dots,t_{d}\in{\mathcal{O}}_{x} be a regular family of parameters at xx. If k=k(x)k=k(x) is of positive characteristic pp, let CkC_{k} be a Cohen ring of kk, that is a DVR with maximal ideal (p)(p) and residue field kk. By Cohen’s theorem if XX is of equal characteristic at pp, then there exists a field of coefficients i:k𝒪^xi\colon k\hookrightarrow{\widehat{\mathcal{O}}}_{x}, while in the mixed characteristic case there exists a ring of coefficients i:Ck𝒪^xi\colon C_{k}\hookrightarrow{\widehat{\mathcal{O}}}_{x}. Prove the following theorem of Kato (see [Kat94, Theorem 3.2]), where APA\llbracket P\rrbracket denotes the formal completion of A[P]A[P] at the ideal A[P+]A[P^{+}]:

(i) In the equal characteristic case the natural homomorphism kPt1,,td𝒪^xk\llbracket P\rrbracket\llbracket t_{1},\dots,t_{d}\rrbracket\to{\widehat{\mathcal{O}}}_{x}, induced by ii and the parameters, is an isomorphism.

(ii) If char(k)>0{\rm char}(k)>0, then the natural homomorphism CkPt1,,td𝒪^xC_{k}\llbracket P\rrbracket\llbracket t_{1},\dots,t_{d}\rrbracket\to{\widehat{\mathcal{O}}}_{x} induced by ii and the parameters is surjective with a principal kernel (θ)(\theta), where θp\theta\equiv p modulo (P+,t1,,td,p)2(P^{+},t_{1},\dots,t_{d},p)^{2}.

Remark 3.4.9.

(i) The equal characteristic pp case naturally shows up in both cases, since one can take θ=p\theta=p.

(ii) In the equal characteristic case Kato’s theorem tells that log regularity is the same as being formally-locally isomorphic to a toric variety. So, the theory of log regular schemes can be viewed as a generalization of toroidal geometry to the mixed characteristic case.

(iii) A relatively difficult theorem asserts that log regularity is preserved by localizations. Kato’s original proof is incomplete, but Gabber later provided missing arguments. The source of the difficulty is clear – regular parameters at generizations of xx are not related to parameters at xx. In fact, this is completely parallel to the situation with usual regularity. However, in the classical case there is a conceptual proof which uses Serre’s cohomological criterion of regularity, and no logarithmic analogue was found so far.

3.4.10. Log regularity and toroidal varieties

Finally, let us describe log regular log varieties over a field kk. As in the case of usual varieties, a nice description (in terms of smoothness) is possible only for simple points xx, that is, points for which the extension k(x)/kk(x)/k is separable.

Exercise 3.4.11.

Assume that XX is a log variety over S=Spec(k)S={\rm Spec}(k) and xXx\in X is such that k(x)/kk(x)/k is separable and XX is log regular at xx. Let x¯X{\overline{x}}\to X be a geometric point over xx.

(i) Prove that locally at x¯{\overline{x}} there exist a chart f:US[Pd]f\colon U\to S[P\oplus{\mathbb{Z}}^{d}] with an étale ff, where P=¯x¯P={\overline{\mathcal{M}}}_{\overline{x}} and dd is the dimension of the logarithmic stratum at xx. Deduce that XX is toroidal at x¯{\overline{x}}, and thus being log regular and toroidal at a simple point are equivalent. (Hint: use parameters and work with étale topology instead of the formal one.)

(ii) Prove that any chart f:US[P]f\colon U\to S[P], which is neat at x¯{\overline{x}}, is smooth at the image of x¯{\overline{x}}.

4. Morphisms of logarithmic schemes

Our next goal is to study morphisms of log schemes in more detail.

4.1. Charts

Recall that charts of log schemes were defined in Definition 3.3.12. Naturally, by a chart of a morphism of log schemes one means compatible charts of the source and the target:

Definition 4.1.1.

Let f:YXf\colon Y\to X be a morphism of log schemes. A chart of ff consists of charts Y𝐀QY\to{\bf A}_{Q} and X𝐀PX\to{\bf A}_{P} and a homomorphism ϕ:PQ\phi\colon P\to Q such that the compositions PΓ(X)Γ(Y)P\to\Gamma({\mathcal{M}}_{X})\to\Gamma({\mathcal{M}}_{Y}) and PQΓ(Y)P\to Q\to\Gamma({\mathcal{M}}_{Y}) coincide. Equivalently, the chart is a commutative diagram

Y\textstyle{Y\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}f\scriptstyle{f}𝐀Q\textstyle{{\bf A}_{Q}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝐀ϕ\scriptstyle{{\bf A}_{\phi}}X\textstyle{X\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝐀P\textstyle{{\bf A}_{P}}

whose horizontal lines are charts. One says that the chart is modeled on ϕ\phi.

It is easy to construct charts, morally, one just starts with a chart PXP\to X and enlarges it by adding enough elements of Y{\mathcal{M}}_{Y}.

Exercise 4.1.2.

Let f:YXf\colon Y\to X be a morphism of fine log schemes, x¯X{\overline{x}}\to X a geometric point and y¯Y{\overline{y}}\to Y a geometric point above x¯{\overline{x}}. Prove that any étale-local fine chart Px¯P\to{\mathcal{M}}_{\overline{x}} at x¯{\overline{x}} extends to a chart of ff étale-locally at y¯{\overline{y}}. (Hint: start with any fine chart Qy¯Q^{\prime}\to{\mathcal{M}}_{\overline{y}} and take QQ to be the image of PQy¯P\oplus Q^{\prime}\to{\mathcal{M}}_{\overline{y}}.)

4.1.3. The standard splitting

If a log scheme XX is provided with a chart X𝐀PX\to{\bf A}_{P} and PQP\to Q is a homomorphism of monoids we will use the notation XP[Q]=X×𝐀P𝐀QX_{P}[Q]=X\times_{{\bf A}_{P}}{\bf A}_{Q}, which indicates that XP[Q]X_{P}[Q] is obtained by a “base change of the monoidal structure”. A chart of ff induces a very useful splitting of ff into the composition YXP[Q]XY\to X_{P}[Q]\to X, where the first morphism is strict, as 𝐀Q{\bf A}_{Q} is a chart of both the source and the target, and the second morphism is monoidal. In a sense, this decomposes ff into a composition of a scheme-like morphism and a monoidal-like morphism, and such a splitting is ubiquitous in log geometry. However, it is non-canonical and exists only étale-locally.

4.1.4. Neat charts

As with the neat absolute charts, to efficiently work with charts of morphisms one would like to construct minimal charts, or even a minimal chart extending a given neat chart of the target. It turns out that in general one cannot construct charts modeled on homomorphisms of sharp monoids, as we are going to demonstrate. Let f:YXf\colon Y\to X be a morphism of Zariski log schemes, yYy\in Y, x=f(y)x=f(y), Q=¯yQ={\overline{\mathcal{M}}}_{y} and P=¯xP={\overline{\mathcal{M}}}_{x}. The induced homomorphism ϕ:PQ\phi\colon P\to Q has no kernel, since any non-unit of x{\mathcal{M}}_{x} is mapped to a non-unit in 𝒪x{\mathcal{O}}_{x} and hence also to a non-unit in 𝒪y{\mathcal{O}}_{y}. Nevertheless, it might happen that ϕ\phi is not injective, but Ker(ϕgp)P=0{\rm Ker}(\phi^{\rm gp})\cap P=0. Here is an archetypical example:

Example 4.1.5.

Let X=Spec(k[s,t])X={\rm Spec}(k[s,t]) with the log structure generated by ss and tt, and Y=Spec(k[s,u])Y={\rm Spec}(k[s,u]) with u=t/su=t/s a chart of the blowing up of XX at the origin xXx\in X with the log structure generated by ss and uu. Let yYy\in Y be a point given by s=0s=0, u=ak×u=a\in k^{\times}. Then P=¯x=log(s)log(t)P={\overline{\mathcal{M}}}_{x}={\mathbb{N}}\log(s)\oplus{\mathbb{N}}\log(t) and Q=¯y=log(s)Q={\overline{\mathcal{M}}}_{y}={\mathbb{N}}\log(s) because u𝒪y×u\in{\mathcal{O}}_{y}^{\times}. The map PQP\to Q sends log(t)\log(t) to log(s)\log(s) and log(u)=log(t)log(s)\log(u)=\log(t)-\log(s) generates the kernel of the map PgpQgpP^{\rm gp}\to Q^{\rm gp}.

The above example motivates the following definition with the idea to provide QQ with the minimal amount of units so that the homomorphism PQP\to Q is injective.

Definition 4.1.6.

(i) The relative characteristic monoid of f:YXf\colon Y\to X is Y/X=Coker(f(X)Y){\mathcal{M}}_{Y/X}={\rm Coker}(f^{*}({\mathcal{M}}_{X})\to{\mathcal{M}}_{Y}). Note that it also coincides with Coker(f1(¯X)¯Y){\rm Coker}(f^{-1}({\overline{\mathcal{M}}}_{X})\to{\overline{\mathcal{M}}}_{Y}), hence Y/X=¯Y/X{\mathcal{M}}_{Y/X}={\overline{\mathcal{M}}}_{Y/X}.

(ii) Assume that Y𝐀QY\to{\bf A}_{Q}, X𝐀PX\to{\bf A}_{P}, 𝐀ϕ{\bf A}_{\phi} is a chart of a morphism of log schemes f:YXf\colon Y\to X, and yYy\in Y is a point with x=f(y)x=f(y). The chart is called neat at yy if ϕ\phi is injective and the induced homomorphism Coker(ϕgp)Y/X,ygp{\rm Coker}(\phi^{\rm gp})\to{\mathcal{M}}^{\rm gp}_{Y/X,y} is an isomorphism.

Neat charts always exist in fppf topology, while in étale topology there might be an obstacle if Coker(ϕgp){\rm Coker}(\phi^{\rm gp}) has a pp-torsion for p=char(k(y))p={\rm char}(k(y)). Proof of the following result, which can be found in [Ogu18, Theorem II.2.4.4], involves a bit of diagram chasing with monoids and their Grothendieck groups.

Exercise 4.1.7.

Assume that f:YXf\colon Y\to X is a morphism of log schemes, y¯Y{\overline{y}}\to Y a geometric point whose image in XX is xXx\in X. Assume that h:U𝐀Ph\colon U\to{\bf A}_{P} is a neat chart of a neighborhood of xx. If Ext1(Y/X,y¯gp,𝒪y¯×)=0{\rm Ext}^{1}({\mathcal{M}}^{\rm gp}_{Y/X,{\overline{y}}},{\mathcal{O}}_{\overline{y}}^{\times})=0, then locally along y¯{\overline{y}} there exists a neat chart of ff that extends hh. In particular, a neat chart exists when the order of torsion of Coker(Y/X,y¯gp){\rm Coker}({\mathcal{M}}^{\rm gp}_{Y/X,{\overline{y}}}) is invertible in k(y¯)k({\overline{y}}).

Remark 4.1.8.

The same argument proves that neat charts always exist fppf locally. Similarly, if the log structure at yy is Zariski and Y/X,y¯gp{\mathcal{M}}^{\rm gp}_{Y/X,{\overline{y}}} is torsion free, a neat chart exists Zariski locally, but if the relative characteristic monoid contains a torsion, one might need to extract roots of some units, ending up with an étale-local or fppf-local chart – depending on the torsion.

4.2. Logarithmic smoothness

4.2.1. Log thickenings

The following definition is a direct extension of its scheme-theoretic analogue.

Definition 4.2.2.

(i) A log thickening is a strict closed immersion STS\hookrightarrow T given by a nilideal 𝒪T{\mathcal{I}}\subset{\mathcal{O}}_{T}.

(ii) A morphism of log schemes f:YXf\colon Y\to X is called formally log smooth (resp. formally log étale, resp. formally log unramified) if for any log thickening and compatible morphisms i:SYi\colon S\to Y, TXT\to X étale locally on TT there exists (resp. there exists unique, resp. there exists at most one) lifting TXT\to X making the diagram commutative:

S\textstyle{S\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}i\scriptstyle{i}Y\textstyle{Y\ignorespaces\ignorespaces\ignorespaces\ignorespaces}f\scriptstyle{f}T\textstyle{T\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}X\textstyle{X}

(iii) A morphism ff is log unramified if it is formally log unramified and f¯{\underline{f}} is of finite type. A morphism ff is log smooth (resp. log étale) if it is formally log smooth (resp. log étale) and f¯{\underline{f}} is finitely presented.

Similarly to the theory of schemes one can study these notions most effectively by use of log differentials.

4.2.3. Logarithmic derivations

Log geometry has a version of the theory of derivations and differentials. The idea is to extend the usual theory by adding logarithmic differentials of monomials.

Definition 4.2.4.

Let ABA\to B be a homomorphism of log rings A=(PA¯)A=(P\to{\underline{A}}) and B=(QB¯)B=(Q\to{\underline{B}}) and let NN be a B¯{\underline{B}}-module. An AA-log derivation (d,δ):BN(d,\delta)\colon B\to N consists of an A¯{\underline{A}}-derivation B¯N{\underline{B}}\to N and a homomorphism δ:QN\delta\colon Q\to N such that δ(P)=0\delta(P)=0 and δ(q)=d(uq)uq\delta(q)=\frac{d(u^{q})}{u^{q}}. The B¯{\underline{B}}-module of all log derivations will be denoted DerB/A(E){\rm Der}_{B/A}(E).

Remark 4.2.5.

One should view δ\delta as derivation of the branch of the logarithm of the monomial uqu^{q} specified by qq.

4.2.6. Logarithmic differentials

Similarly to usual Kähler differentials, there always exists a universal log derivation whose target is the module ΩB/A1\Omega^{1}_{B/A} of logarithmic differentials. This is not so standard, but we omit 1 in the notation because we will never consider modules of differential pp-forms with p>1p>1.

Exercise 4.2.7.

Prove that the functor DerB/A(){\rm Der}_{B/A}(\cdot) is representable and describe the corresponding universal module as follows: ΩB/A1\Omega^{1}_{B/A} is the quotient of ΩB¯/A¯1(B(Qgp/Pgp))\Omega^{1}_{{\underline{B}}/{\underline{A}}}\oplus(B\otimes(Q^{\rm gp}/P^{\rm gp})) by the submodule generated by the relations (duq,uqq)(du^{q},-u^{q}\otimes q) with qQq\in Q, where we denote the image in Qgp/PgpQ^{\rm gp}/P^{\rm gp} also by qq.

We will not go into details, but most of classical results about Kähler differentials, such as base change, compatibility with localizations and fundamental sequences, extend to log rings and log differentials. In addition, the module of differentials of any strict étale homomorphism vanishes, hence by use of étale descent the definition can be globalized to the definition of an 𝒪Y{\mathcal{O}}_{Y}-module ΩY/X1\Omega^{1}_{Y/X} for any morphism YXY\to X of log schemes, and it comes equipped with the universal log derivation (d,δ):(X𝒪X)ΩY/X1(d,\delta)\colon({\mathcal{M}}_{X}\to{\mathcal{O}}_{X})\to\Omega^{1}_{Y/X}. In particular, dd induces a homomorphism ΩY¯/X¯1ΩY/X1\Omega^{1}_{{\underline{Y}}/{\underline{X}}}\to\Omega^{1}_{Y/X}. As usual in log geometry, to get some feeling of a new notion one should look at the two polar cases, and the case of strict morphisms reduces to the usual scheme theory.

Example/Exercise 4.2.8.

Use the previous exercise to prove that:

(i) If YXY\to X is a strict morphism of log schemes, then ΩY¯/X¯1=ΩY/X1\Omega^{1}_{{\underline{Y}}/{\underline{X}}}=\Omega^{1}_{Y/X}.

(ii) If RR is a ring, Y=𝐀R,PY={\bf A}_{R,P}, Y=𝐀R,QY={\bf A}_{R,Q} and f=𝐀R,ϕf={\bf A}_{R,\phi} for ϕ:PQ\phi\colon P\to Q, then

ΩY/X1=Coker(ϕgp)R[Q].\Omega^{1}_{Y/X}={\rm Coker}(\phi^{\rm gp})\otimes R[Q].

In particular, if R{\mathbb{Q}}\subseteq R, then ΩY/X1\Omega^{1}_{Y/X} is free and its rank is the rank of Coker(ϕgp){\rm Coker}(\phi^{\rm gp}), and if 𝐅pR{\bf F}_{p}\subseteq R, then ΩY/X1\Omega^{1}_{Y/X} is free and its rank is the pp-rank of Coker(ϕgp){\rm Coker}(\phi^{\rm gp}). (Hint: show that DerR[Q]/R[P](E)=Hom(Coker(ϕgp),E){\rm Der}_{R[Q]/R[P]}(E)={\rm Hom}({\rm Coker}(\phi^{\rm gp}),E).)

(iii) By a slight abuse of notation for a sharp monoid PP and a ring RR we denote by P0RP\stackrel{{\scriptstyle 0}}{{\to}}R the homomorphism of monoids taking P+P^{+} to 0 (but 0 goes to 1). Compute the differentials of a log point Spec(P0k){\rm Spec}(P\stackrel{{\scriptstyle 0}}{{\to}}k). More generally, show that for any ring RR and the hollow log scheme X=Spec(P0R)X={\rm Spec}(P\stackrel{{\scriptstyle 0}}{{\to}}R) there is an isomorphism of RR-modules ΩX/R1=PgpR\Omega^{1}_{X/R}=P^{\rm gp}\otimes R.

4.2.9. Chart criterion

The main theorem of Kato about logarithmic smoothness gives the following criterion for log smoothness in terms of charts, see [Kat89, Theorem 3.5]:

Theorem 4.2.10.

For a morphism f:YXf\colon Y\to X of fine log schemes and a geometric point y¯Y¯{\overline{y}}\to{\underline{Y}} the following conditions are equivalent:

  • (i)

    ff is log smooth (reps. log étale) at y¯{\overline{y}},

  • (ii)

    Locally at y¯{\overline{y}} there exists a chart V𝐀QV\to{\bf A}_{Q}, U𝐀PU\to{\bf A}_{P} modelled on ϕ:PQ\phi\colon P\to Q such that

    • (a)

      The morphism VUP[Q]V\to U_{P}[Q] is smooth (resp. étale)

    • (b)

      Ker(ϕgp){\rm Ker}(\phi^{\rm gp}) and Coker(ϕgp)tor{\rm Coker}(\phi^{\rm gp})_{\rm tor} (resp. Ker(ϕgp){\rm Ker}(\phi^{\rm gp}) and Coker(ϕgp){\rm Coker}(\phi^{\rm gp})) are finite of order invertible in k(y¯)k({\overline{y}}).

  • (iii)

    The same condition as (ii) but with VUP[Q]V\to U_{P}[Q] étale.

Exercise 4.2.11.

(i) Check that the condition on ϕ\phi is equivalent to smoothness (resp. étaleness) of the morphism of diagonalizable groups Dk(y¯),QgpDk(y¯),PgpD_{k({\overline{y}}),Q^{\rm gp}}\to D_{k({\overline{y}}),P^{\rm gp}} and by the above criteria it is also equivalent to smoothness (resp. étaleness) of the map 𝐀k,Q𝐀k,P{\bf A}_{k,Q}\to{\bf A}_{k,P}.

(ii) Deduce (iii) from (ii) by enlarging the chart to QnQ\oplus{\mathbb{Z}}^{n}, where t1,,tnt_{1},\dots,t_{n} are regular parameters on the fiber of VUP[Q]V\to U_{P}[Q] through y¯{\overline{y}}. (Hint: for example, one can send the generators of n{\mathbb{Z}}^{n} to 1+ti1+t_{i} (i.e. one increases QQ by adding the elements log(1+ti)\log(1+t_{i}).)

(iii) Check that in condition (b) one can also achieve that Ker(ϕ)=0{\rm Ker}(\phi)=0 (Hint: again, just increase QQ accordingly.)

Here as a very typical example involving log points. We will return to this setting also in example 4.3.10.

Example 4.2.12.

(i) Show that log points Spec(P0k){\rm Spec}(P\stackrel{{\scriptstyle 0}}{{\to}}k) with a sharp fine P0P\neq 0 are not log smooth over kk.

(ii) Take P=log(t)P={\mathbb{N}}\log(t) and consider the nodal curve C=Spec(k[x,y])C={\rm Spec}(k[x,y]) with the log structure induced by Q=log(x)log(y)Q={\mathbb{N}}\log(x)\oplus{\mathbb{N}}\log(y). Show that the morphism CPC\to P taking log(t)\log(t) to alog(x)+blog(y)a\log(x)+b\log(y) with a,ba,b\in{\mathbb{N}} is log smooth if and only if (a,b)(a,b) is invertible in kk. Show that if (a,b)k×(a,b)\in k^{\times}, then (unlike ΩC¯/k1\Omega^{1}_{\underline{C}/k}) the module ΩC/P1\Omega^{1}_{C/P} is invertible; in fact, it is generated by δ(x),δ(y)\delta(x),\delta(y) subject to the relation aδ(x)+bδ(y)=0a\delta(x)+b\delta(y)=0.

(iii) Give another proof of the results of (ii) by noting that the morphism CPC\to P is just the fiber over the origin of the morphism 𝐀Q𝐀P{\bf A}_{Q}\to{\bf A}_{P} given by t=xaybt=x^{a}y^{b}.

An important theorem of Kato states that log regularity is preserved by log smooth morphisms, see [Kat94, Theorem 8.2]:

Theorem 4.2.13.

If XX is a log regular log scheme and YXY\to X is log smooth, then YY is log regular.

4.3. Logarithmic blowings up

4.3.1. Log ideals

Definition 4.3.2.

Let XX be a fine log scheme. By a log ideal we mean a coherent sheaf of ideals JXJ\subseteq{\mathcal{M}}_{X}, where coherence means that locally around any geometric point x¯{\overline{x}} the ideal is generated by Jx¯J_{\overline{x}}. A log ideal is called invertible if it is locally generated by a single element.

Remark 4.3.3.

Since monoids X(U){\mathcal{M}}_{X}(U) are integral there is no need to impose a condition analogous to being a non-zero divisor in a ring, and invertible ideals are preserved by arbitrary pullbacks, unlike the theory of schemes.

4.3.4. Log blowings up

Log blowings up are defined by the same universal property as usual blowings up, but with log ideals used instead of ideals.

Definition 4.3.5.

Let XX be a log scheme and JJ a log ideal on XX. A morphism f:YXf\colon Y\to X is called the log blowing up of XX along JJ and denoted LogBlJ(X)XLogBl_{J}(X)\to X if ff is the universal morphism of fine log schemes such that f1(J)f^{-1}(J) is invertible. The saturated log blowing up is defined by the same property but in the category of fs log schemes. Thus, it is nothing else but the saturation of the log blowing up.

Exercise 4.3.6.

Log blowings up are preserved by any base changes h:XXh\colon X^{\prime}\to X, i.e. Y=LogBlh1J(X)=Y×XXY^{\prime}=LogBl_{h^{-1}J}(X^{\prime})=Y\times_{X}X^{\prime}. In particular, applying this to YY one obtains that Y×XY=YY\times_{X}Y=Y and hence YXY\to X is a monomorphism.

Now we are going to prove that log blowings up exist and are, in fact, monoidal morphisms, as one might expect, since they realize a monoidal construction.

Exercise 4.3.7.

We will construct Y=LogBlJ(X)Y=LogBl_{J}(X) in a few steps.

(i) Prove that if X=𝐀PX={\bf A}_{P} and I=[J]I={\mathbb{Z}}[J], then Y¯=BlI(X){\underline{Y}}={\rm Bl}_{I}(X) and the log structure on the ss-chart Y¯s=Spec([P][Is]){\underline{Y}}_{s}={\rm Spec}({\mathbb{Z}}[P][\frac{I}{s}]) of the blowing up is given by P[Js]P[J-s] (the submonoid of PgpP^{\rm gp} generated by PP and the elements of JsJ-s). Furthermore, the chart YsY_{s} of YY is the universal log scheme over XX such that the pullback of JJ to YsY_{s} is principal generated by ss.

(ii) Prove that if XX possesses a chart X𝐀PX\to{\bf A}_{P} and JJ is generated by an ideal J0PJ_{0}\subseteq P, then LogBlJ(X)=LogBlJ0(𝐀P)×𝐀PXLogBl_{J}(X)=LogBl_{J_{0}}({\bf A}_{P})\times_{{\bf A}_{P}}X. Deduce that log blowing up YXY\to X always exists and is a proper monoidal morphism. (Hint: use part (i), Exercise 4.3.6 for the first claim and then apply étale descent.)

(iii) Prove that any log blowing up is a log étale morphism. (Hint: by base change and étale descent this reduces to the case described in (i) and then Kato’s criterion does the job.)

(iv) Finally, reinterpret examples we have seen: using Exercise 3.3.23(ii) prove directly that Y=Y×XYY=Y\times_{X}Y, and explain which chart of a log blowing up is considered in Remark 3.3.24.

Typically, one asks when a log blowing up coincides with the usual blowing up along the ideal uJ𝒪Xu^{J}{\mathcal{O}}_{X} generated by JJ, because this seems to be the most adequate situation. Clearly, this happens if and only if the relevant toroidal blowings up of the charts 𝐀P{\bf A}_{P} are compatible with the chart maps h:U𝐀Ph\colon U\to{\bf A}_{P}. In particular, this is automatic if hh is flat or, at least, what is called Tor{\rm Tor}-independent from JJ, that is, Tor1[P](𝒪X,[P]/Jn[P]])=0{\rm Tor}_{1}^{{\mathbb{Z}}[P]}({\mathcal{O}}_{X},{\mathbb{Z}}[P]/J^{n}{\mathbb{Z}}[P]])=0. Using the latter criterion Nizioł proved the following claim, see [Niz06, Proposition 4.3]:

Theorem 4.3.8.

If XX is a log regular log scheme and JJ is a log ideal, then the underlying scheme of the LogBlJ(X)LogBl_{J}(X) is the blowing up of XX along the induced ideal uJ𝒪Xu^{J}{\mathcal{O}}_{X}. In addition, the saturated log blowing up LogBlJ(X)satLogBl_{J}(X)^{\rm sat} is log regular and the log structure is induced by the union of the preimage of the toroidal divisor on XX and the exceptional divisor.

In fact the approach with Tor functors is only needed in the more difficult case of mixed characteristics.

Exercise 4.3.9.

Prove Theorem 4.3.8 when XX is equicharacteristic. (Hint: formal completions of noetherian schemes are flat and hence compatible with blowings up. Formally locally XX looks as Spec(kPt1,,td){\rm Spec}(k\llbracket P\rrbracket\llbracket t_{1},\dots,t_{d}\rrbracket), hence the blowing up is described easily via the base change from the toric case via the flat homomorphism k[P]kPt1,,tdk[P]\hookrightarrow k\llbracket P\rrbracket\llbracket t_{1},\dots,t_{d}\rrbracket. The log regularity follows from Theorem 4.2.13.)

Now let us discuss a few cases, that are simpler to compute but are often viewed as pathological and not worth consideration. In particular, one can easily have that the proper morphism LogBlJ(X)XLogBl_{J}(X)\to X is not birational. Using such morphisms becomes critical if one wants to study relative desingularization over singular bases (e.g. the log point or thick log point with a non-reduced scheme structure), but this direction has not studied yet in the literature, and we will only mention it in a couple of remarks later.

Example/Exercise 4.3.10.

(i) Show that the log blowing up of the ideal P+P^{+} on the log point Spec(P0k){\rm Spec}(P\stackrel{{\scriptstyle 0}}{{\to}}k) modeled on a fine sharp monoid PP is of dimension rk(P)1{\rm rk}(P)-1.

(ii) Let s=Spec(log(t)0k)s={\rm Spec}({\mathbb{N}}\log(t)\stackrel{{\scriptstyle 0}}{{\to}}k) be the standard log point (i.e. t=0t=0 in kk) and

Cs=Spec(log(x)log(y)k[x,y]/(xy))C_{s}={\rm Spec}({\mathbb{N}}\log(x)\oplus{\mathbb{N}}\log(y)\to k[x,y]/(xy))

a log smooth ss-curve with log(t)\log(t) mapped to log(x)+log(y)\log(x)+\log(y). Let JJ be the maximal ideal of log(x)log(y){\mathbb{N}}\log(x)\oplus{\mathbb{N}}\log(y). Show by a direct computation that Xs=LogBlJ(Cs)X_{s}=LogBl_{J}(C_{s}) is also log smooth over ss, the map XsCsX_{s}\to C_{s} is an isomorphism over the complement of the origin OCsO\in C_{s} and the preimage EE of OO is a non-reduced double 𝐏k1{\bf P}^{1}_{k} component with the nilpotent ideal (ε)(\varepsilon) and the log structure given by log(ε){\mathbb{N}}\log(\varepsilon) and log(t)\log(t) mapping to 2log(z)2\log(z).

(iii) Now, embed ss as the origin of S=Spec(log(t)k[t])S={\rm Spec}({\mathbb{N}}\log(t)\to k[t]) and CsC_{s} as the closed fiber of a log smooth (even semistable) SS-curve

C=Spec(log(x)log(y)k[x,y])C={\rm Spec}({\mathbb{N}}\log(x)\oplus{\mathbb{N}}\log(y)\to k[x,y])

with t=xyt=xy. Recall that LogBlJ(Cs)LogBl_{J}(C_{s}) is the pullback of the log scheme LogBlJ(C)LogBl_{J}(C), whose underlying scheme is just BlO(C)Bl_{O}(C). Use this to conceptually explain the results of (ii), in particular, the reason why the new component is doubled (has a non-reduced structure).

4.4. Logarithmic étaleness

In this section we restrict to the fs setting, in which Kummer covers are usually studied. We just give definitions, check simplest properties and mention various directions studied in the literature.

4.4.1. Kummer étale morphisms

A homomorphism of toric monoids ϕ:PQ\phi\colon P\to Q is called Kummer if PgpQgpP^{\rm gp}\subseteq Q^{\rm gp} is of finite index and QQ is the saturation of PP in QgpQ^{\rm gp}. A log étale morphism of log schemes YXY\to X is called Kummer if the induced homomorphisms of monoids ¯x¯y{\overline{\mathcal{M}}}_{x}\to{\overline{\mathcal{M}}}_{y} are Kummer. A Kummer étale cover YXY\to X is a surjective Kummer étale morphism.

Exercise 4.4.2.

Check that, indeed, this notion of a covering defines a Grothendieck topology called Kummer étale of ket{\rm ket} topology. (Hint: this mainly reduces to the check that Kummer étale covers are preserved by base changes.)

Remark 4.4.3.

(i) Kummer étale topology is the closest analogue of the étale topology in the setting of fs log schemes. For example, see Exercise 3.3.25. An important fact is that the theory of descent works pretty similarly to the case of étale (or flat) topology: 𝒪Xket{\mathcal{O}}_{X_{\rm ket}} and, more generally, representable functors are sheaves (see [Niz08, Proposition 2.18 and Theorem 2.20]), etc. Ideals in 𝒪Xket{\mathcal{O}}_{X_{\rm ket}} are called Kummer ideals. Working with them provides a convenient formalism for extracting roots from monomials.

(ii) Theories of Kummer étale and general log étale cohomologies are now developed rather deeply, see [Nak97] and [Nak17]. The first one is simpler, but in order to have some fundamental theorems in full generality one has to work with the whole log étale site.

4.4.4. Log étale site

For the sake of completeness, let us discuss how one defines the notion of a log étale covering in general. Our motivation is just to see a few more examples from log geometry, and we will not discuss the log étale cohomology theories.

Exercise 4.4.5.

Construct an example of surjective log étale morphisms YXY\to X and ZXZ\to X such that Y×XZXY\times_{X}Z\to X is not surjective. In particular, a base change does not have to be surjective, and hence can even be empty. (Hint: for example, one can take the plane with the monoid 2{\mathbb{N}}^{2}, apply log blowing up to the origin and another log blowing up to one of the two preimages of the origin with characteristic 2{\mathbb{N}}^{2}, obtaining a log étale morphism XXX^{\prime}\to X with the exceptional divisor consisting of two components E=E1E2E=E_{1}\cup E_{2}. Then Y=XE1Y=X^{\prime}\setminus E_{1} and Z=XE2Z=X^{\prime}\setminus E_{2} do the job. It is also instructive to consider a purely combinatorial (or toric) description of this example.)

The above example shows that one should be careful with the notion of surjectivity. Naturally, we would like to declare any log blowing up YXY\to X to be a cover, but for any YYY^{\prime}\subsetneq Y the morphism YXY^{\prime}\to X should not be a cover. So, one defines the log étale topology to be the topology generated by Kummer étale covers and log blowings up.

Exercise 4.4.6.

Let f:YXf\colon Y\to X be a log étale morphism. Show that ff is a log étale cover if and only if for any log blowing up XXX^{\prime}\to X the morphism Y×XXXY\times_{X}X^{\prime}\to X^{\prime} is surjective.

5. The stacks LogSLog_{S}

This section is devoted to a very important technique in logarithmic geometry, which was introduced by Olsson in [Ols03] (with a strong influence of ideas of Luc Illusie). It turns out that SS-logarithmic structures on schemes TT over (the underlying scheme of) a base log scheme SS are classified by a stack 𝐋𝐨𝐠S{\mathbf{Log}}_{S}. Working with such stacks allows to interpret various logarithmic constructions and notions in terms of usual algebraic geometry of schemes and stacks. In particular, some results can be deduced from the non-logarithmic analogs on the nose.

5.1. Constructions of 𝐋𝐨𝐠S{\mathbf{Log}}_{S}

5.1.1. The moduli definition of 𝐋𝐨𝐠S{\mathbf{Log}}_{S}

To any log scheme S=(S¯,S)S=({\underline{S}},{\mathcal{M}}_{S}) Olsson assigns the category 𝐋𝐨𝐠S{\mathbf{Log}}_{S} fibered in groupoids over the category of S¯{\underline{S}}-schemes as follows: an object of 𝐋𝐨𝐠S{\mathbf{Log}}_{S} is a logarithmic SS-scheme XX and a morphism is a strict morphism YXY\to X of logarithmic SS-schemes. The fiber functor just forgets the log structures.

Remark 5.1.2.

(i) If TT is an S¯{\underline{S}}-scheme, then an object of 𝐋𝐨𝐠S(T){\mathbf{Log}}_{S}(T) is just a logarithmic SS-scheme whose underlying scheme is TT. Thus, the stack 𝐋𝐨𝐠S{\mathbf{Log}}_{S} parameterizes the ways in which one can enhance S¯{\underline{S}}-schemes with the structure of logarithmic SS-schemes. So, informally speaking, it parameterizes log structures over SS.

(ii) The association S𝐋𝐨𝐠SS\mapsto{\mathbf{Log}}_{S} is naturally a functor from the category of schemes to the category of stacks.

5.1.3. Algebraicity

Olsson proved that the stack 𝐋𝐨𝐠S{\mathbf{Log}}_{S} is in fact an Artin stack of locally finite type over SS. The proof goes by checking the usual properties – representability of the diagonal and existence of a smooth presentation. The first property reduces to a simple study of the group of S{\mathcal{M}}_{S}-automorphisms of the log-structures log SS-schemes – naturally, they are extensions of diagonalizable groups by finite groups. The second property holds because PQSP[Q]𝐋𝐨𝐠S\coprod_{P\to Q}S_{P}[Q]\to{\mathbf{Log}}_{S} is smooth and surjective. The latter will be discussed in §5.1.9 and then we will use it to construct 𝐋𝐨𝐠S{\mathbf{Log}}_{S} very explicitly (in particular, the map from each SP[Q]S_{P}[Q] factors through the quotient by the group of SS-automorphisms, which is the extension of the diagonalizable group 𝐃Qgp/Pgp{\bf D}_{Q^{\rm gp}/P^{\rm gp}} by the finite group AutP(Q){\rm Aut}_{P}(Q)).

5.1.4. The tautological log structure

By the definition of the stack 𝐋𝐨𝐠S{\mathbf{Log}}_{S} any scheme over it is provided with a canonical log structure, and by descent one immediately obtains that the same is true for stacks over 𝐋𝐨𝐠S{\mathbf{Log}}_{S}. In particular, 𝐋𝐨𝐠S{\mathbf{Log}}_{S} itself is provided with a tautological log structure {\mathcal{M}} and for any x𝐋𝐨𝐠S(T)x\in{\mathbf{Log}}_{S}(T) the induced homomorphism (T,x)(𝐋𝐨𝐠S,)(T,{\mathcal{M}}_{x})\to({\mathbf{Log}}_{S},{\mathcal{M}}) is strict, that is, the log structure x{\mathcal{M}}_{x} is induced from the tautological structure {\mathcal{M}} via the structure morphism TS¯T\to{\underline{S}}. Similarly to Remark 5.1.2 this provides a formalization of the claim that {\mathcal{M}} is the universal log structure over SS. The following exercise essentially reduces to unravelling the definitions.

Exercise 5.1.5.

(i) Show that the log structure on SS induces a section S𝐋𝐨𝐠SS\hookrightarrow{\mathbf{Log}}_{S} of the structure morphism 𝐋𝐨𝐠SS{\mathbf{Log}}_{S}\to S.

(ii) Let f:YXf\colon Y\to X be an SS-morphisms of logarithmic SS-schemes and let x:X𝐋𝐨𝐠Sx\colon X\to{\mathbf{Log}}_{S} and y:Y𝐋𝐨𝐠Sy\colon Y\to{\mathbf{Log}}_{S} be the corresponding morphisms. Show that y=xfy=x\circ f if and only if ff is strict.

(iii) For a morphism f:YXf\colon Y\to X of log schemes the square

𝐋𝐨𝐠Y\textstyle{{\mathbf{Log}}_{Y}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝐋𝐨𝐠f\scriptstyle{{\mathbf{Log}}_{f}}𝐋𝐨𝐠X\textstyle{{\mathbf{Log}}_{X}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Y\textstyle{Y\ignorespaces\ignorespaces\ignorespaces\ignorespaces}f\scriptstyle{f}X\textstyle{X}

is Cartesian if and only if ff is strict.

5.1.6. The stacks 𝒳P[Q]{\mathcal{X}}_{P}[Q]

The stack 𝐋𝐨𝐠X{\mathbf{Log}}_{X} is huge, but it is a rather simple object that can be described by charts very explicitly. A first approximation for this is the following construction due to Olsson. Assume that XX is a log scheme with a global chart X𝐀PX\to{\bf A}_{P} and PQP\to Q is a homomorphism of monoids. Note that the diagonalizable group 𝐃Qgp/Pgp=Spec([Qgp/Pgp]){\bf D}_{Q^{\rm gp}/P^{\rm gp}}={\rm Spec}({\mathbb{Z}}[Q^{\rm gp}/P^{\rm gp}]) acts on the 𝐀P{\bf A}_{P}-scheme 𝐀Q{\bf A}_{Q}, and hence also acts on the XX-scheme XP[Q]X_{P}[Q] obtained by the base change. Let 𝒳P[Q]{\mathcal{X}}_{P}[Q] denote the quotient stack [XP[Q]/𝐃Qgp/Pgp][X_{P}[Q]/{\bf D}_{Q^{\rm gp}/P^{\rm gp}}]. The importance of these stacks introduced by Olsson becomes clear from the universal property they satisfy, which we are going to establish now. A short proof can be found in [MT21, Lemma 2.2.4].

Exercise 5.1.7.

Let X𝐀PX\to{\bf A}_{P} be a chart as above and let YY be a logarithmic XX-scheme.

(i) Show that flat locally on YY the PP-homomorphism Q¯YQ\to{\overline{\mathcal{M}}}_{Y} lifts to a homomorphism QYQ\to{\mathcal{M}}_{Y} and deduce that the functor of such liftings is a DQgp/PgpD_{Q^{\rm gp}/P^{\rm gp}}-torsor in the flat topology. Moreover, if YY is fs, then this functor is already an étale torsor. (Hint: use that the homomorphisms Ygp¯Ygp{\mathcal{M}}^{\rm gp}_{Y}\twoheadrightarrow{\overline{\mathcal{M}}}^{\rm gp}_{Y} has a section flat locally, and this is even true étale locally if YY is fs and hence the groups ¯Y,y¯gp{\overline{\mathcal{M}}}^{\rm gp}_{Y,{\overline{y}}} are torsion free.)

(ii) Deduce that 𝒳P[Q]{\mathcal{X}}_{P}[Q] represents the functor HomP(Q,¯Y){\rm Hom}_{P}(Q,{\overline{\mathcal{M}}}_{Y}) on the category of log schemes over XX, while XP[Q]X_{P}[Q] represents the functor HomP(Q,Y){\rm Hom}_{P}(Q,{\mathcal{M}}_{Y}). In particular, XX-homomorphisms Y𝒳P[Q]Y\to{\mathcal{X}}_{P}[Q] are in a natural one-to-one correspondence with PP-homomorphisms Q¯YQ\to{\overline{\mathcal{M}}}_{Y}. (Hint: the second claim is clear, to deduce the first one divide by the action of DQgp/Pgp=𝒳P[Qgp]D_{Q^{\rm gp}/P^{\rm gp}}={\mathcal{X}}_{P}[Q^{\rm gp}] and use (i).)

Remark 5.1.8.

Keep the above notation and assume that QYQ\to{\mathcal{M}}_{Y} is a chart that induces an isomorphism ϕ:Q=Γ(Y,¯Y)\phi\colon Q=\Gamma(Y,{\overline{\mathcal{M}}}_{Y}). There are many liftings of ϕ\phi to a chart of YY, obtained by multiplying monomials by units, but this is precisely the ambiguity which is killed by dividing by 𝐃Qgp/Pgp{\bf D}_{Q^{\rm gp}/P^{\rm gp}}. Thus Y𝒳P[Q]Y\to{\mathcal{X}}_{P}[Q] can be viewed as the canonical XX-chart of YY determined only by the isomorphism ϕ\phi. Its only ambiguity is the group AutP(Q){\rm Aut}_{P}(Q) of PP-automorphisms of QQ.

5.1.9. A smooth presentation

Assume that XX possesses a global chart PXP\to{\mathcal{M}}_{X}. Olsson proves that the natural morphism PQ𝒳P[Q]𝐋𝐨𝐠X\coprod_{P\to Q}{\mathcal{X}}_{P}[Q]\to{\mathbf{Log}}_{X}, where the union is over all homomorphisms from PP to a fine monoid, is strict, surjective and étale. In particular, PQXP[Q]𝐋𝐨𝐠X\coprod_{P\to Q}X_{P}[Q]\to{\mathbf{Log}}_{X} is a smooth presentation of 𝐋𝐨𝐠X{\mathbf{Log}}_{X}. In fact, this easily reduces to the fact that étale locally any logarithmic XX-scheme YY possesses an étale cover YYY^{\prime}\to Y whose source possesses a chart YXP[Q]Y^{\prime}\to X_{P}[Q] and hence a strict morphism Y𝒳P[Q]Y^{\prime}\to{\mathcal{X}}_{P}[Q].

In general, XX possesses a global chart étale-locally, and the above construction is compatible with strict étale morphisms XXX^{\prime}\to X. So, a presentation of 𝐋𝐨𝐠X{\mathbf{Log}}_{X} can be obtained from a presentation of its fine enough strict étale cover.

5.1.10. A groupoid presentation

The presentation f:PQ𝒳P[Q]𝐋𝐨𝐠Xf\colon\coprod_{P\to Q}{\mathcal{X}}_{P}[Q]\to{\mathbf{Log}}_{X} already gives a non-bad approximation of the source, but clearly it factors through f:PQ𝒳P[Q]/AutP(Q)f^{\prime}\colon\coprod_{P\to Q}{\mathcal{X}}_{P}[Q]/{\rm Aut}_{P}(Q). Even the étale morphism ff^{\prime} is still not a monomorphism because PP-automorphisms of localizations of QQ not necessarily come from AutP(Q){\rm Aut}_{P}(Q), but it is easy to pin down the ambiguity – one needs to identify all localizations of QQ in all possible ways, in particular, dividing 𝒳P[Q]{\mathcal{X}}_{P}[Q] by AutP(Q){\rm Aut}_{P}(Q). Informally speaking, 𝐋𝐨𝐠S{\mathbf{Log}}_{S} is obtained from the union of all charts PQXP[Q]\coprod_{P\to Q}X_{P}[Q] by identifying all isomorphic open subcharts, in particular, dividing by automorphism: at first step this involves dividing by the groups 𝐃Qgp/Pgp{\bf D}_{Q^{\rm gp}/P^{\rm gp}}, and then by identifying all localizations of QiQ_{i} and QjQ_{j}. In a sense, 𝐋𝐨𝐠X{\mathbf{Log}}_{X} is nothing else but the universal XX-chart constructed purely combinatorially.

Now let us outline the construction. It is more convenient to work geometrically, when the contravariant functor Q𝒳P[Q]Q\mapsto{\mathcal{X}}_{P}[Q] is replaced by the functor Spec(Q)𝒳P[Q]{\rm Spec}(Q)\to{\mathcal{X}}_{P}[Q] from the category of affine Kato fans over Spec(P){\rm Spec}(P) because the latter globalizes in the obvious way. Moreover, one can naturally define a wider category of Kato stacks, and this functor extends to Kato stacks by (an appropriate) descent. Consider the diagram of all affine PP-fans Spec(Q){\rm Spec}(Q) with the morphisms being face embeddings, then the colimit P{\mathcal{L}}_{P} exists as a Kato fan. Intuitively, it is a PP-“fan” which contains each Spec(Q){\rm Spec}(Q) as a face in a unique way. It is not so difficult to show that 𝐋𝐨𝐠X=𝒳P[P]{\mathbf{Log}}_{X}={\mathcal{X}}_{P}[{\mathcal{L}}_{P}], in particular, the morphism 𝐋𝐨𝐠XX{\mathbf{Log}}_{X}\to X is monoidal (in the stacky sense). Moreover, one can now give an explicit stacky presentation of 𝐋𝐨𝐠X{\mathbf{Log}}_{X}. We outline the main results in a (difficult) exercise below and refer to [MT21, Sections 2,3] for detailed arguments.

Exercise 5.1.11.

(i) Given PP-monoids Q0,,QnQ_{0},\dots,Q_{n} by a join face we mean a PP-monoid RR and face embeddings Spec(R)Spec(Qi){\rm Spec}(R)\hookrightarrow{\rm Spec}(Q_{i}) with 0in0\leq i\leq n (in other words, we fix isomorphisms of RR and localizations of QiQ_{i}). Show that the colimit J(Q0,,Qn)J(Q_{0},\dots,Q_{n}) of the diagram of all face embeddings of Q0,,QnQ_{0},\dots,Q_{n} is a Kato fan, which we call the join of Q0,,QnQ_{0},\dots,Q_{n}.

(ii) Construct a natural simplicial Kato fan n=Q0,,QnJ(Q0,,Qn){\mathcal{L}}_{n}=\coprod_{Q_{0},\dots,Q_{n}}J(Q_{0},\dots,Q_{n}) and show that it is in fact a groupoid equivalent to a Kato fan which we denote P{\mathcal{L}}_{P}. In other words, 1=PQSpec(Q)P{\mathcal{L}}_{1}=\coprod_{P\to Q}{\rm Spec}(Q)\to{\mathcal{L}}_{P} is a cover and its fiber powers are n{\mathcal{L}}_{n}.

(iii) Show that P{\mathcal{L}}_{P} is characterized by the following universal property: any PP-fan Spec(Q){\rm Spec}(Q) possesses a unique face embedding into {\mathcal{L}}.

(iv) Show that 𝒳P[P]=𝐋𝐨𝐠X{\mathcal{X}}_{P}[{\mathcal{L}}_{P}]={\mathbf{Log}}_{X} and deduce that 𝐋𝐨𝐠X{\mathbf{Log}}_{X} is equivalent to the simplicial stack 𝒳n=Q0,,Qn𝒳P[J(Q0,,Qn)]{\mathcal{X}}_{n}=\coprod_{Q_{0},\dots,Q_{n}}{\mathcal{X}}_{P}[J(Q_{0},\dots,Q_{n})], which is, in fact, a groupoid.

5.2. Stacks 𝐋𝐨𝐠{\mathbf{Log}} and logarithmic properties

Now, we will show how one can systematically interpret various logarithmic properties of morphisms of log schemes. Initially such notions as log smoothness, log flatness, log étaleness, etc. were defined in a rather ad hoc manner. Then in [Ols03] Olsson found a very general way to unify these definitions.

Definition 5.2.1.

Let 𝒫{\mathcal{P}} be a property of morphisms of schemes, for example, smooth, étale, flat, etc. A morphism of log schemes f:YXf\colon Y\to X is said to be log 𝒫{\mathcal{P}} (resp. weakly log 𝒫{\mathcal{P}}) if the associated morphism of stacks 𝐋𝐨𝐠f:𝐋𝐨𝐠Y𝐋𝐨𝐠X{\mathbf{Log}}_{f}\colon{\mathbf{Log}}_{Y}\to{\mathbf{Log}}_{X} (resp. Y𝐋𝐨𝐠XY\to{\mathbf{Log}}_{X}) is 𝒫{\mathcal{P}}.

Remark 5.2.2.

(i) Both definitions have some advantages. The morphism Y𝐋𝐨𝐠XY\to{\mathbf{Log}}_{X} is “smaller” and easier to analyze; if YY is quasi-compact, then a morphism YXY\to X factors through an open substack of 𝐋𝐨𝐠X{\mathbf{Log}}_{X} finitely presented over XX. On the other hand, when studying compositions it is certainly easier to work with the morphisms 𝐋𝐨𝐠Y𝐋𝐨𝐠X{\mathbf{Log}}_{Y}\to{\mathbf{Log}}_{X}.

(ii) Despite the terminology, neither condition implies the other one. Olsson showed in [Ols03, Example 4.3] that if 𝒫{\mathcal{P}} is “having geometrically connected fibers”, then log 𝒫{\mathcal{P}} does not imply weakly log 𝒫{\mathcal{P}}, but there is even a much more basic example: any morphism f:YXf\colon Y\to X with a quasi-compact source is not weakly log surjective because 𝐋𝐨𝐠X{\mathbf{Log}}_{X} is never quasi-compact, while ff is often log surjective, for example, when it is an isomorphism.

However, using that Y𝐋𝐨𝐠XY\to{\mathbf{Log}}_{X} factors into the composition of an open immersion Y𝐋𝐨𝐠YY\hookrightarrow{\mathbf{Log}}_{Y} and 𝐋𝐨𝐠f{\mathbf{Log}}_{f} we obtain that if 𝒫{\mathcal{P}} is local on the source, then log 𝒫{\mathcal{P}} implies weakly log 𝒫{\mathcal{P}}.

(iii) For the properties of being log smooth, log étale and log flat Olsson showed that the original definitions given by Kato are equivalent to the new ones and also equivalent to the corresponding weak logarithmic properties.

Exercise 5.2.3.

Use Olsson’s definition to reprove Kato’s criterion of log smoothness from Theorem 4.2.10. Most probably, this will lead you to a conceptual explanation of the fact that condition 4.2.10(ii)(b) on ϕ:PQ\phi\colon P\to Q is equivalent to smoothness of the morphism 𝐃k(y¯),Q𝐃k(y¯),P{\bf D}_{k({\overline{y}}),Q}\to{\bf D}_{k({\overline{y}}),P}.

Now we can also naturally interpret the notion of a log étale cover:

Exercise 5.2.4.

Prove that a morphism f:YXf\colon Y\to X is a log étale cover if and only if 𝐋𝐨𝐠f:𝐋𝐨𝐠Y𝐋𝐨𝐠X{\mathbf{Log}}_{f}\colon{\mathbf{Log}}_{Y}\to{\mathbf{Log}}_{X} is an étale cover. In other words, a log étale morphism is a cover if and only if it is log surjective.

5.2.5. Equivalence of the conditions

A general result about equivalence of the two definitions was obtained in [MT21, Theorem 4.3.1]: if 𝒫{\mathcal{P}} is stable under pullbacks, étale local on the source, and flat local on the base, then a morphism is log 𝒫{\mathcal{P}} if and only if it is weakly log 𝒫{\mathcal{P}}. This follows from a slightly surprising fact that 𝐋𝐨𝐠Y𝐋𝐨𝐠X{\mathbf{Log}}_{Y}\to{\mathbf{Log}}_{X} can be obtained from its small piece Y𝐋𝐨𝐠XY\to{\mathbf{Log}}_{X} by base change and flat descent.

Exercise 5.2.6.

(i) Assume that YXY\to X has a global chart 𝐀Q𝐀P{\bf A}_{Q}\to{\bf A}_{P}. Show that for any homomorphism QRQ\to R both squares in the following diagram are Cartesian:

(ii) Using the presentation of the stacks 𝐋𝐨𝐠{\mathbf{Log}} constructed in §5.1.10, étale descent and claim (i) deduce the assertion of [MT21, Theorem 4.3.1].

5.2.7. Log regularity

The above equivalence result applies, in particular, to the following properties: smoothness, étaleness, flatness and regularity. In fact, log regularity was introduced in [MT21] and the equivalence with weak log regularity was used to establish its basic properties, e.g. a chart criterion analogous to Kato’s chart criterion of log smoothness. In fact, this was the original motivation of the research of [MT21], which, in its turn, was motivated by relative resolution of singularities.

5.2.8. Logarithmic differentials

The stacks 𝐋𝐨𝐠{\mathbf{Log}} also allow to interpret logarithmic derivations and differentials. In view of the fact that a morphism f:YXf\colon Y\to X is log smooth if and only if the associated morphism h:Y𝐋𝐨𝐠Xh\colon Y\to{\mathbf{Log}}_{X} is smooth, the following fact is very natural: ΩY/X1=ΩY¯/𝐋𝐨𝐠X1\Omega^{1}_{Y/X}=\Omega^{1}_{{\underline{Y}}/{\mathbf{Log}}_{X}} and DerY/X=DerY¯/𝐋𝐨𝐠X{\rm Der}_{Y/X}={\rm Der}_{{\underline{Y}}/{\mathbf{Log}}_{X}} (see [ATW20, Lemma 2.4.4]).

Exercise 5.2.9.

(i) Assume that YXY\to X possesses a chart 𝐀Q𝐀P{\bf A}_{Q}\to{\bf A}_{P}. Prove that DerY/X=DerY¯/𝒳P[Q]{\rm Der}_{Y/X}={\rm Der}_{{\underline{Y}}/{\mathcal{X}}_{P}[Q]}. (Hint: compare the first fundamental sequences of log derivations associated with YXP[Q]XY\to X_{P}[Q]\to X and of derivations associated with YXP[Q]𝒳P[Q]Y\to X_{P}[Q]\to{\mathcal{X}}_{P}[Q].)

(ii) Use the étale cover Q𝒳P[Q]𝐋𝐨𝐠X\coprod_{Q}{\mathcal{X}}_{P}[Q]\to{\mathbf{Log}}_{X} to deduce that, indeed, for any morphism of log schemes YXY\to X one has that DerY/X=DerY¯/𝐋𝐨𝐠X{\rm Der}_{Y/X}={\rm Der}_{{\underline{Y}}/{\mathbf{Log}}_{X}} and hence also ΩY/X1=ΩY¯/𝐋𝐨𝐠X1\Omega^{1}_{Y/X}=\Omega^{1}_{{\underline{Y}}/{\mathbf{Log}}_{X}}.

5.2.10. Logarithmic fibers

Let f:YXf\colon Y\to X be a morphism of log schemes. By the log fibers of ff we mean the connected components of the fibers of the induced morphism h:Y𝐋𝐨𝐠Xh\colon Y\to{\mathbf{Log}}_{X}. For example, if ff is log smooth or log regular, then the log fibers are smooth or regular, respectively. It is easy to compute the log fibers étale-locally: if ff has a chart modeled on 𝐀Q𝐀P{\bf A}_{Q}\to{\bf A}_{P}, then hh factors through the étale morphism 𝒳P[Q]𝐋𝐨𝐠X{\mathcal{X}}_{P}[Q]\to{\mathbf{Log}}_{X} hence the log fibers are nothing else but the fibers of the stacky chart Y𝒳P[Q]Y\to{\mathcal{X}}_{P}[Q].

Let us consider two general types of examples of opposite kind. If the homomorphisms ¯x¯y{\overline{\mathcal{M}}}_{x}\to{\overline{\mathcal{M}}}_{y} are injective, then the log fibers have the most natural geometric interpretation:

Exercise 5.2.11.

Assume that f:YXf\colon Y\to X has a chart modeled on 𝐀Q𝐀P{\bf A}_{Q}\to{\bf A}_{P} with sharp PP and QQ (such a chart exists étale-locally at yYy\in Y with x=f(y)x=f(y) if ¯x¯y{\overline{\mathcal{M}}}_{x}\to{\overline{\mathcal{M}}}_{y} is injective). Show that the log fibers of ff are the connected components of the log strata of the fibers of ff. In particular, log fibers of a toroidal variety over a field are just the connected components of its log stratification.

Such a description certainly cannot work for log blowings up, which might have non-discrete fibers but are log étale.

Exercise 5.2.12.

Show that the log fibers of any log blowing up YXY\to X are nothing else but the points of yy (as one would expect in the case of a monomorphism).

In general, one can locally factor a morphism into a composition of a sharp morphism and a log blowing up, so log fibers admit a sort of a mixed description, but we do not discuss this here and refer the interested reader to [ATW20, §2.2].

References

  • [ACG+13] Dan Abramovich, Qile Chen, Danny Gillam, Yuhao Huang, Martin Olsson, Matthew Satriano, and Shenghao Sun, Logarithmic geometry and moduli, Handbook of moduli. Vol. I, Adv. Lect. Math. (ALM), vol. 24, Int. Press, Somerville, MA, 2013, pp. 1–61. MR 3184161
  • [ATW20] Dan Abramovich, Michael Temkin, and Jarosław Włodarczyk, Relative desingularization and principalization of ideals, arXiv e-prints (2020), arXiv:2003.03659.
  • [Ful93] William Fulton, Introduction to toric varieties, Annals of Mathematics Studies, vol. 131, Princeton University Press, Princeton, NJ, 1993, The William H. Roever Lectures in Geometry. MR 1234037
  • [Ill15] Luc Illusie, From Pierre Deligne’s secret garden: looking back at some of his letters, Jpn. J. Math. 10 (2015), no. 2, 237–248. MR 3392531
  • [Kat89] Kazuya Kato, Logarithmic structures of Fontaine-Illusie, Algebraic analysis, geometry, and number theory (Baltimore, MD, 1988), Johns Hopkins Univ. Press, Baltimore, MD, 1989, pp. 191–224. MR 1463703 (99b:14020)
  • [Kat94] by same author, Toric singularities, Amer. J. Math. 116 (1994), no. 5, 1073–1099. MR 1296725 (95g:14056)
  • [KKMSD73] George Kempf, Finn Faye Knudsen, David Mumford, and Bernard Saint-Donat, Toroidal embeddings. I, Lecture Notes in Mathematics, Vol. 339, Springer-Verlag, Berlin, 1973. MR 0335518 (49 #299)
  • [MT21] Sam Molcho and Michael Temkin, Logarithmically regular morphisms, Math. Ann. 379 (2021), no. 1-2, 325–346. MR 4211089
  • [Nak97] Chikara Nakayama, Logarithmic étale cohomology, Math. Ann. 308 (1997), no. 3, 365–404. MR 1457738
  • [Nak17] by same author, Logarithmic étale cohomology, II, Adv. Math. 314 (2017), 663–725. MR 3658728
  • [Niz06] Wiesława Nizioł, Toric singularities: log-blow-ups and global resolutions, J. Algebraic Geom. 15 (2006), no. 1, 1–29. MR 2177194 (2006i:14015)
  • [Niz08] by same author, KK-theory of log-schemes. I, Doc. Math. 13 (2008), 505–551. MR 2452875
  • [Ogu18] Arthur Ogus, Lectures on logarithmic algebraic geometry, Cambridge Studies in Advanced Mathematics, vol. 178, Cambridge University Press, Cambridge, 2018. MR 3838359
  • [Ols03] Martin C. Olsson, Logarithmic geometry and algebraic stacks, Ann. Sci. École Norm. Sup. (4) 36 (2003), no. 5, 747–791. MR 2032986
  • [Wł20] Jarosław Włodarczyk, Functorial resolution except for toroidal locus. Toroidal compactification, arXiv e-prints (2020).