Introduction to logarithmic geometry
Key words and phrases:
Logarithmic geometry, toroidal geometry, resolution of singularities1. 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 -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)
-
(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 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 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 be a regular scheme and a divisor.
(i) One says that is strictly (or simple) normal crossings or just snc at if locally one has that , where is a regular family of parameters at and we use notation .
(ii) One says that is normal crossings at if it is snc étale-locally at .
A divisor is called normal crossings or snc if this is so everywhere on .
Exercise 2.1.2.
(i) Show that the number in the definition is the number of branches of at (e.g. the number of irreducible formal components) and hence is an invariant of at , also called the multiplicity of at .
(ii) Define a stratification of by the multiplicity and show that each stratum is a regular locally closed subscheme.
(iii) Assume that is regular and is a reduced divisor with irreducible components . Show that is snc if and only if for each the scheme theoretic intersection 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 be a lattice and the dual lattice. This can be encoded by the non-degenerate pairing , but despite the symmetry, we will always view and as geometric spaces, while the elements of will be viewed as functions or .
Let be an -rational polyhedral cone, i.e. a cone given by finitely many conditions with . One associates to an affine toric -variety as follows. The dual cone is also rational and it follows easily that the monoid is finitely generated. Therefore, is an affine -algebra and is an affine variety. Note that is provided with the natural action of the torus whose lattice of characters is . Furthermore, contains an open orbit isomorphic to , and this is the source of the terminology. The information encoded in the pair , often called the combinatorial information, is equivalent to the information encoded in 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 can be reconstructed from the affine toric variety as follows: is the lattice of characters of and giving the -action on is equivalent to providing an -grading of , where -homogeneous elements are the equivariant ones with the character : they are acted on via the rule ; the cone is determined by the monoid , which is precisely the set of characters with a non-zero homogeneous component .
(ii) Show that the set of orbits of is in a natural one-to-one correspondence with the faces of . In particular, the open orbit corresponds to the vertex of .
Remark 2.2.3.
The affine theory can be globalized as follows. On the schematic part of the picture one glues affine toric varieties with the same torus along isomorphic open affine toric subvarieties. A resulting object is called a toric scheme with respect to the torus . On the combinatorial side one glues polyhedral cones along faces. The resulting objects are called cone complexes. Sometimes one only considers complexes embedded in and calls them polyhedral fans. If the union of all faces of a fan is the whole , 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 is called regular if its sharpening is a free monoid and hence . Show that this happens if and only if there exists a basis of such that each edge of contains some . Thus, 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 and the monoid generated by the elements of 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 is a pair with a -variety and an open subscheme such that étale locally possesses an étale morphism to a toric scheme such that is the preimage of the torus. Namely, there exists an étale covering and étale morphisms , called toroidal chart, such that is the preimage of the torus of . If, moreover, the covering can be chosen to be Zariski, say , then the toroidal variety is called simple (or without self-intersections).
(ii) A morphism of toroidal varieties is any morphism taking to .
Example 2.2.10.
If is regular, is a normal crossings divisor and , then is a toroidal variety modelled on regular simplices , and the free monoid is generated by the regular parameters which define the branches of (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 is a simple toroidal variety and set and (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 .
Exercise 2.2.12.
(i) Show that any local chart that maps to the closed stratum of some induces in isomorphism , in particular, depends only on and .
(ii) Conversely, any section of can be extended to a local chart on a neighborhood of . (Hint: in addition to one should choose a regular family of parameters on the stratum of through and lift them to elements of .)
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 . By we denote the subgroup of invertible elements and the sharpening of is . One says that is sharp if .
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 of monoids. The main examples of limits and colimits we will use are as follows:
Exercise 3.1.2.
(i) is just the usual product of sets with componentwise addition and it coincides with the coproduct, usually denoted .
(ii) For homomorphisms , the fiber product is the submonoid of given by .
(iii) Instead of kernels in the category of monoids one uses congruence relations, that is, equivalence relations which are also submonoids: if is a surjective homomorphism, then the induced equivalence relation is a submonoid, and conversely any congruence relation appears in this way.
(iv) For homomorphisms , the pushout is the quotient of by the minimal congruence relation such that for any .
3.1.3. Ideals
An ideal is a subset such that , where the convention is that is also an ideal (the analogue of the zero ideal of rings). An ideal is prime if implies that or . The set of all prime ideals is denoted and called the fan of . An ideal of the form is called principal.
Exercise 3.1.4.
Show that taking the preimage establishes a bijection between the ideals of and .
3.1.5. Fine monoids
We will usually work with fine monoids:
Definition 3.1.6.
(i) is finitely generated if there exists a surjective homomorphism .
(ii) is integral (or cancellative) if implies that .
(iii) 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 with a group, called the Grothendieck group of . (Hint: for example, one can bound the cardinality of because it is generated by the image of , and then general representability theorems do the job because the category of groups is complete.)
(ii) Construct explicitly as the quotient of by the following equivalence relation: if there exists such that .
(iii) Show that is integral if and only if the homomorphism is injective.
(iv) Define the integralization to be the image of in . Show that is the universal homomorphism from to an integral monoid and is the left adjoint functor to the embedding of the category of integral monoids into .
3.1.9. Fs monoids
An especially nice class of monoids is defined as follows:
Definition 3.1.10.
(i) An integral monoid is saturated if for each with for one has that .
(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 is equivalent to give a lattice and a rational polyhedral cone in such that .
(ii) Show that is naturally bijective to the set of faces of .
Exercise 3.1.12.
Let denote the category of saturated monoids. Show that the embedding possesses a left adjoint functor, which is called the saturation functor and denoted . Show that is just the saturation of in , that is the divisible hull of in .
3.2. Logarithmic structures
Definition 3.2.1.
Let be one of the following topologies – Zariski, étale or flat.
(i) A -prelogarithmic structure on a scheme is a sheaf of monoids on the site with a structure homomorphism of monoids . A homomorphism of prelog structures is a homomorphism of sheaves of monoids compatible with the structure homomorphisms.
(ii) A -logarithmic structure is a -prelogarithmic structure which induces an isomorphism , and hence also . The sharpening is called the characteristic monoid of .
(iii) The default topology in this definition is the étale topology, so usually it will not be mentioned. A log structure induces a Zariski log structure just by restricting. By a slight abuse of language we say that itself is Zariski if this restriction does not loose information, that is, for the morphism of sites .
Remark 3.2.2.
(i) The homomorphism is an analog of exponentiation, and this is one of the reasons to use additive monoids as the source. Traditionally it is denoted , say , but probably the exponential notation, such as instead of , is more suggestive. Informally, any such can be viewed as a branch of the logarithm of , 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 -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 .
(1) The largest log structure with an injective is , but it is not too useful. The main exception is when is a “small” scheme – a semi-local curve or the spectrum of a valuation ring, with the most useful case being when is a trait.
(2) A very important example of a log structure with an injective is as follows. Assume that is a closed subset and set where is the open immersion of the complement . This is the log structure of elements invertible outside of . Usually it is used when is the underlying closed set of a Cartier divisor; in this case is determined by the log structure and we call a divisorial log structure. In particular, the toroidal scheme structure can also be encoded in the log structure of -monomial elements, where .
(3) The other extreme case is provided by so-called hollow log schemes with for any . Usually they show up when one restricts a log structure on onto a closed subscheme such that is generically non-trivial on . Often this is the most economical way to encode certain information about the ambient scheme on . 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 can be canonically transformed into a log structure . The idea is that is a log structure if and only if 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 on let be the pushout of the diagram
Show that is a log structure, is the universal homomorphism of to a log structure and the functor 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 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 should then be applied.
Definition/Exercise 3.2.7.
Integralization and saturation of a log structure are defined by applying the corresponding functors to monoids of sections, sheafification and then applying the functor . 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 for a monoid . For shortness we will not distinguish the monoid and its sheafification.
(ii) A log structure is called quasi-coherent if étale-locally there exists a constant prelog structure such that . If, in addition, can be chosen finitely generated, is called coherent. We warn the reader that this notion is not related to coherence and quasi-coherence of -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 .
Finally, for a morphism one defines direct and inverse images of the log structures on and on :
Definition/Exercise 3.2.11.
(i) Show that is a log structure denoted (by a slight abuse of notation) .
(ii) The inverse image is the log structure associated to the prelog structure . Show that, as expected, is left adjoint to .
3.3. Logarithmic schemes
Now we can introduce the category of log schemes.
Definition 3.3.1.
(i) A log scheme is a tuple , where is a scheme called the underlying scheme and is a log structure on . Usually we will use looser notation when this cannot lead to a confusion, e.g. write instead of or just omit it, write instead of or even denote the underlying scheme by the same letter .
(ii) A log scheme is called quasi-coherent, coherent, integral, fine, fs, etc., if the log structure is quasi-coherent, coherent, integral, fine, fs, etc.
(iii) A morphism of log schemes consists of the underlying morphism of schemes and a homomorphism of log structures compatible with the structure homomorphisms and .
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 a morphism is uniquely determined by the underlying morphism of schemes, and extends to if the log structure on is “larger” than the image of the log structure on . If and are divisorial, then this happens if and only if .
(ii) In particular, providing a toroidal scheme with the divisorial log structure we obtain a fully faithful embedding of the category of toroidal schemes into the category of log schemes. Naturally, such will be called a toroidal log scheme.
(iii) The other extreme is when the log schemes are hollow: , . In this case, extends to via any homomorphism .
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 and in general:
Definition 3.3.5.
A morphism of log schemes is called strict if .
In particular, for a log scheme any morphism can be uniquely enhanced to a strict morphism of log schemes.
Remark 3.3.6.
Any morphism canonically factors into a composition such that is strict and . It is natural to view as a scheme-like morphism and as a morphism which only increases the log structure. However, this factorization is not especially useful. The reason is that 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 is a toroidal log scheme and is a toroidal subscheme.
(ii) The simplest and most important case is when is the affine line with marked origin (i.e. and , where we denote the generator of the monoid by to stress that it is mapped to by ) and is the origin. The induced log structure is and we call 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 .
Note also that any other point , has the trivial induced log structure since is mapped to and hence the functor shrinks to by sending to .
(iii) Similarly, for each the thick point can be provided with the log structure induced by . It is neither injective, nor hollow.
(iv) One can also consider other log points, for example, for a toric monoid the log structure induced by corresponds to the origin of the toric scheme .
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 consists of a ring , a monoid and a homomorphism .
(ii) The logarithmic spectrum of a log ring, usually denoted just by is the underlying scheme provided with the log structure induced by the prelog structure .
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 , 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 by the associated log scheme.
Definition/Exercise 3.3.12.
(i) For a monoid let denote the log scheme whose underlying scheme is and the log structure is induced by the homomorphism . If we work over a base scheme, for example, , then we will use the notation .
(ii) Let be a log scheme. Show that giving a global chart for the log structure is equivalent to giving a strict morphism of log schemes . Any such morphism is called a global chart of , 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 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 because various operations can produce non-sharp monoids. For example, removing the origin from , where , one obtains the scheme , which is also a chart of itself. On the other hand, its log structure is trivial, hence 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 is called neat at a geometric point if . In particular, is automatically sharp.
Example 3.3.15.
Let . Then the tautological chart 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 be a fine log scheme, let be a geometric point and let . Any neat chart at gives rise to a section of the sharpening homomorphism.
(i) Show, that conversely any section of induces a chart for a small enough étale neighborhood of and this chart is neat at .
(ii) Furthermore, show that such a section exists if and only if the sequence
splits, and this is automatic whenever has no torsion of order divisible by ([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 .)
(iv) Show that if the log structure is fs, then is automatically torsion free, and hence even Zariski log schemes possess neat charts locally at points of .
3.3.17. Monoidal morphisms
We say that a morphism of log schemes is monoidal if étale-locally on it is the base change of morphisms of the form . Informally speaking, 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 . They all be discussed later, and we start with the first two.
Exercise 3.3.18.
(i) Show that for any coherent log scheme there exists a universal morphism (resp. ) whose target is a fine (resp. fs) log scheme. In other words, the functor (resp. ) is left adjoint to the embedding of the category of fine (resp. fs) integral schemes into the category of coherent log schemes. In addition, is a closed immersion and is finite. (Hint: first, assuming that possesses a chart show that and are as required. In general, use local-étale charts and étale descent of finiteness.)
(ii) Show that for any strict morphism of coherent schemes one has that and .
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 be a finite diagram of log schemes. Show that exists and can be described as follows: and the log structure is the colimit of the pullbacks of to . (In particular, this involves the functor ; the shortest way is to pullback as prelog structures, then take the colimit, and then apply once.)
(ii) Show that if all are fine or fs, then there exists a limit in the same category, and it coincides with or , respectively.
Remark 3.3.21.
Often one mentions or in the notation of the limit, e.g. or , 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 is a scheme, , are homomorphisms of finitely generated monoids and , , , then , and .
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 with the log structure induced by , and let with the log structure induced by . Show that , in particular, the integralization functor cuts off the component from .
(ii) More generally, assume that are toric monoids such that . Then , that is, the morphism 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 at the origin. Clearly, is non-flat, the dimension of the fiber over the origin jumps, and this even gives rise to the new irreducible component in . Assume now that and are provided with the log structures generated by and . Then the component in the product acquires a non-cancellative log structure with monoid presented by generators and relation . This forces the integralization functor to remove the 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 with the log structure and with the log structure . We will later see that the Kummer cover is log étale when . Check that contains two components (diagonal and antidiagonal) intersecting over the origin and the characteristic at the intersection point is the non saturated monoid with generators subject to the relation . It is isomorphic to the submonoid of obtained by removing the element .
(ii) Show that is the normalization of and it is just the disjoit union of the two copies of – the diagonal and the antidiagonal. The characteristic of the two points over is – the sharpening of .
(iii) More generally, if is a scheme, is a Kummer extension of toric monoids, i.e. is the saturation of in , and , then is isomorphic to the split cover for . So, the log étale -Galois cover 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 possesses a natural stratification by .
Exercise 3.4.2.
(i) Let be the locus on which the rank of the characteristic monoid is at most . Show that these sets are closed and hence induce a stratification by the locally closed sets .
(ii) Refine this stratification to a log stratification of by (non-necessarily reduced) locally closed subschemes as follows: if the structure is Zariski at and , then the stratum at is given by the vanishing of , where is the maximal ideal of . Show that this is compatible with strict morphisms and hence extends to arbitrary log structures by étale descent. Finally, set
(iii) Show that, indeed, is the reduction of . Also, show that the log strata of the chart log schemes 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 is logarithmically regular if each locally closed subscheme is regular (in particular, reduced) and of codimension .
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 be a log regular log scheme with a geometric point over . Let . By a regular family of parameters at we mean a section and elements , such that restrict to a regular family of parameters of at . In particular, is the dimension of at and . We call regular parameters and the elements for will be called logarithmic parameters.
In view of Exercise 3.3.16(iv) such families of parameters exist, and if the log structure at is Zariski, one can even construct Zariski locally. As in the classical case, it follows from the definition that any regular family of parameters generates the maximal ideal at . Furthermore, parameters naturally give rise to very explicit étale and formal charts.
Exercise 3.4.8.
Assume that a log scheme is Zariski and log regular at a point with , and let and be a regular family of parameters at . If is of positive characteristic , let be a Cohen ring of , that is a DVR with maximal ideal and residue field . By Cohen’s theorem if is of equal characteristic at , then there exists a field of coefficients , while in the mixed characteristic case there exists a ring of coefficients . Prove the following theorem of Kato (see [Kat94, Theorem 3.2]), where denotes the formal completion of at the ideal :
(i) In the equal characteristic case the natural homomorphism , induced by and the parameters, is an isomorphism.
(ii) If , then the natural homomorphism induced by and the parameters is surjective with a principal kernel , where modulo .
Remark 3.4.9.
(i) The equal characteristic case naturally shows up in both cases, since one can take .
(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 are not related to parameters at . 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 . As in the case of usual varieties, a nice description (in terms of smoothness) is possible only for simple points , that is, points for which the extension is separable.
Exercise 3.4.11.
Assume that is a log variety over and is such that is separable and is log regular at . Let be a geometric point over .
(i) Prove that locally at there exist a chart with an étale , where and is the dimension of the logarithmic stratum at . Deduce that is toroidal at , 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 , which is neat at , is smooth at the image of .
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 be a morphism of log schemes. A chart of consists of charts and and a homomorphism such that the compositions and coincide. Equivalently, the chart is a commutative diagram
whose horizontal lines are charts. One says that the chart is modeled on .
It is easy to construct charts, morally, one just starts with a chart and enlarges it by adding enough elements of .
Exercise 4.1.2.
Let be a morphism of fine log schemes, a geometric point and a geometric point above . Prove that any étale-local fine chart at extends to a chart of étale-locally at . (Hint: start with any fine chart and take to be the image of .)
4.1.3. The standard splitting
If a log scheme is provided with a chart and is a homomorphism of monoids we will use the notation , which indicates that is obtained by a “base change of the monoidal structure”. A chart of induces a very useful splitting of into the composition , where the first morphism is strict, as is a chart of both the source and the target, and the second morphism is monoidal. In a sense, this decomposes 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 be a morphism of Zariski log schemes, , , and . The induced homomorphism has no kernel, since any non-unit of is mapped to a non-unit in and hence also to a non-unit in . Nevertheless, it might happen that is not injective, but . Here is an archetypical example:
Example 4.1.5.
Let with the log structure generated by and , and with a chart of the blowing up of at the origin with the log structure generated by and . Let be a point given by , . Then and because . The map sends to and generates the kernel of the map .
The above example motivates the following definition with the idea to provide with the minimal amount of units so that the homomorphism is injective.
Definition 4.1.6.
(i) The relative characteristic monoid of is . Note that it also coincides with , hence .
(ii) Assume that , , is a chart of a morphism of log schemes , and is a point with . The chart is called neat at if is injective and the induced homomorphism is an isomorphism.
Neat charts always exist in fppf topology, while in étale topology there might be an obstacle if has a -torsion for . 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 is a morphism of log schemes, a geometric point whose image in is . Assume that is a neat chart of a neighborhood of . If , then locally along there exists a neat chart of that extends . In particular, a neat chart exists when the order of torsion of is invertible in .
Remark 4.1.8.
The same argument proves that neat charts always exist fppf locally. Similarly, if the log structure at is Zariski and 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 given by a nilideal .
(ii) A morphism of log schemes is called formally log smooth (resp. formally log étale, resp. formally log unramified) if for any log thickening and compatible morphisms , étale locally on there exists (resp. there exists unique, resp. there exists at most one) lifting making the diagram commutative:
(iii) A morphism is log unramified if it is formally log unramified and is of finite type. A morphism is log smooth (resp. log étale) if it is formally log smooth (resp. log étale) and 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 be a homomorphism of log rings and and let be a -module. An -log derivation consists of an -derivation and a homomorphism such that and . The -module of all log derivations will be denoted .
Remark 4.2.5.
One should view as derivation of the branch of the logarithm of the monomial specified by .
4.2.6. Logarithmic differentials
Similarly to usual Kähler differentials, there always exists a universal log derivation whose target is the module of logarithmic differentials. This is not so standard, but we omit 1 in the notation because we will never consider modules of differential -forms with .
Exercise 4.2.7.
Prove that the functor is representable and describe the corresponding universal module as follows: is the quotient of by the submodule generated by the relations with , where we denote the image in also by .
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 -module for any morphism of log schemes, and it comes equipped with the universal log derivation . In particular, induces a homomorphism . 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 is a strict morphism of log schemes, then .
(ii) If is a ring, , and for , then
In particular, if , then is free and its rank is the rank of , and if , then is free and its rank is the -rank of . (Hint: show that .)
(iii) By a slight abuse of notation for a sharp monoid and a ring we denote by the homomorphism of monoids taking to 0 (but 0 goes to 1). Compute the differentials of a log point . More generally, show that for any ring and the hollow log scheme there is an isomorphism of -modules .
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 of fine log schemes and a geometric point the following conditions are equivalent:
-
(i)
is log smooth (reps. log étale) at ,
-
(ii)
Locally at there exists a chart , modelled on such that
-
(a)
The morphism is smooth (resp. étale)
-
(b)
and (resp. and ) are finite of order invertible in .
-
(a)
-
(iii)
The same condition as (ii) but with étale.
Exercise 4.2.11.
(i) Check that the condition on is equivalent to smoothness (resp. étaleness) of the morphism of diagonalizable groups and by the above criteria it is also equivalent to smoothness (resp. étaleness) of the map .
(ii) Deduce (iii) from (ii) by enlarging the chart to , where are regular parameters on the fiber of through . (Hint: for example, one can send the generators of to (i.e. one increases by adding the elements .)
(iii) Check that in condition (b) one can also achieve that (Hint: again, just increase 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 with a sharp fine are not log smooth over .
(ii) Take and consider the nodal curve with the log structure induced by . Show that the morphism taking to with is log smooth if and only if is invertible in . Show that if , then (unlike ) the module is invertible; in fact, it is generated by subject to the relation .
(iii) Give another proof of the results of (ii) by noting that the morphism is just the fiber over the origin of the morphism given by .
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 is a log regular log scheme and is log smooth, then is log regular.
4.3. Logarithmic blowings up
4.3.1. Log ideals
Definition 4.3.2.
Let be a fine log scheme. By a log ideal we mean a coherent sheaf of ideals , where coherence means that locally around any geometric point the ideal is generated by . A log ideal is called invertible if it is locally generated by a single element.
Remark 4.3.3.
Since monoids 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 be a log scheme and a log ideal on . A morphism is called the log blowing up of along and denoted if is the universal morphism of fine log schemes such that 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 , i.e. . In particular, applying this to one obtains that and hence 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 in a few steps.
(i) Prove that if and , then and the log structure on the -chart of the blowing up is given by (the submonoid of generated by and the elements of ). Furthermore, the chart of is the universal log scheme over such that the pullback of to is principal generated by .
(ii) Prove that if possesses a chart and is generated by an ideal , then . Deduce that log blowing up 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.)
Typically, one asks when a log blowing up coincides with the usual blowing up along the ideal generated by , because this seems to be the most adequate situation. Clearly, this happens if and only if the relevant toroidal blowings up of the charts are compatible with the chart maps . In particular, this is automatic if is flat or, at least, what is called -independent from , that is, . Using the latter criterion Nizioł proved the following claim, see [Niz06, Proposition 4.3]:
Theorem 4.3.8.
If is a log regular log scheme and is a log ideal, then the underlying scheme of the is the blowing up of along the induced ideal . In addition, the saturated log blowing up is log regular and the log structure is induced by the union of the preimage of the toroidal divisor on 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 is equicharacteristic. (Hint: formal completions of noetherian schemes are flat and hence compatible with blowings up. Formally locally looks as , hence the blowing up is described easily via the base change from the toric case via the flat homomorphism . 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 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 on the log point modeled on a fine sharp monoid is of dimension .
(ii) Let be the standard log point (i.e. in ) and
a log smooth -curve with mapped to . Let be the maximal ideal of . Show by a direct computation that is also log smooth over , the map is an isomorphism over the complement of the origin and the preimage of is a non-reduced double component with the nilpotent ideal and the log structure given by and mapping to .
(iii) Now, embed as the origin of and as the closed fiber of a log smooth (even semistable) -curve
with . Recall that is the pullback of the log scheme , whose underlying scheme is just . 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 is called Kummer if is of finite index and is the saturation of in . A log étale morphism of log schemes is called Kummer if the induced homomorphisms of monoids are Kummer. A Kummer étale cover 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 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: and, more generally, representable functors are sheaves (see [Niz08, Proposition 2.18 and Theorem 2.20]), etc. Ideals in are called Kummer ideals. Working with them provides a convenient formalism for extracting roots from monomials.
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 and such that 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 , apply log blowing up to the origin and another log blowing up to one of the two preimages of the origin with characteristic , obtaining a log étale morphism with the exceptional divisor consisting of two components . Then and 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 to be a cover, but for any the morphism 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 be a log étale morphism. Show that is a log étale cover if and only if for any log blowing up the morphism is surjective.
5. The stacks
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 -logarithmic structures on schemes over (the underlying scheme of) a base log scheme are classified by a stack . 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
5.1.1. The moduli definition of
To any log scheme Olsson assigns the category fibered in groupoids over the category of -schemes as follows: an object of is a logarithmic -scheme and a morphism is a strict morphism of logarithmic -schemes. The fiber functor just forgets the log structures.
Remark 5.1.2.
(i) If is an -scheme, then an object of is just a logarithmic -scheme whose underlying scheme is . Thus, the stack parameterizes the ways in which one can enhance -schemes with the structure of logarithmic -schemes. So, informally speaking, it parameterizes log structures over .
(ii) The association is naturally a functor from the category of schemes to the category of stacks.
5.1.3. Algebraicity
Olsson proved that the stack is in fact an Artin stack of locally finite type over . 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 -automorphisms of the log-structures log -schemes – naturally, they are extensions of diagonalizable groups by finite groups. The second property holds because is smooth and surjective. The latter will be discussed in §5.1.9 and then we will use it to construct very explicitly (in particular, the map from each factors through the quotient by the group of -automorphisms, which is the extension of the diagonalizable group by the finite group ).
5.1.4. The tautological log structure
By the definition of the stack 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 . In particular, itself is provided with a tautological log structure and for any the induced homomorphism is strict, that is, the log structure is induced from the tautological structure via the structure morphism . Similarly to Remark 5.1.2 this provides a formalization of the claim that is the universal log structure over . The following exercise essentially reduces to unravelling the definitions.
Exercise 5.1.5.
(i) Show that the log structure on induces a section of the structure morphism .
(ii) Let be an -morphisms of logarithmic -schemes and let and be the corresponding morphisms. Show that if and only if is strict.
(iii) For a morphism of log schemes the square
is Cartesian if and only if is strict.
5.1.6. The stacks
The stack 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 is a log scheme with a global chart and is a homomorphism of monoids. Note that the diagonalizable group acts on the -scheme , and hence also acts on the -scheme obtained by the base change. Let denote the quotient stack . 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 be a chart as above and let be a logarithmic -scheme.
(i) Show that flat locally on the -homomorphism lifts to a homomorphism and deduce that the functor of such liftings is a -torsor in the flat topology. Moreover, if is fs, then this functor is already an étale torsor. (Hint: use that the homomorphisms has a section flat locally, and this is even true étale locally if is fs and hence the groups are torsion free.)
(ii) Deduce that represents the functor on the category of log schemes over , while represents the functor . In particular, -homomorphisms are in a natural one-to-one correspondence with -homomorphisms . (Hint: the second claim is clear, to deduce the first one divide by the action of and use (i).)
Remark 5.1.8.
Keep the above notation and assume that is a chart that induces an isomorphism . There are many liftings of to a chart of , obtained by multiplying monomials by units, but this is precisely the ambiguity which is killed by dividing by . Thus can be viewed as the canonical -chart of determined only by the isomorphism . Its only ambiguity is the group of -automorphisms of .
5.1.9. A smooth presentation
Assume that possesses a global chart . Olsson proves that the natural morphism , where the union is over all homomorphisms from to a fine monoid, is strict, surjective and étale. In particular, is a smooth presentation of . In fact, this easily reduces to the fact that étale locally any logarithmic -scheme possesses an étale cover whose source possesses a chart and hence a strict morphism .
In general, possesses a global chart étale-locally, and the above construction is compatible with strict étale morphisms . So, a presentation of can be obtained from a presentation of its fine enough strict étale cover.
5.1.10. A groupoid presentation
The presentation already gives a non-bad approximation of the source, but clearly it factors through . Even the étale morphism is still not a monomorphism because -automorphisms of localizations of not necessarily come from , but it is easy to pin down the ambiguity – one needs to identify all localizations of in all possible ways, in particular, dividing by . Informally speaking, is obtained from the union of all charts by identifying all isomorphic open subcharts, in particular, dividing by automorphism: at first step this involves dividing by the groups , and then by identifying all localizations of and . In a sense, is nothing else but the universal -chart constructed purely combinatorially.
Now let us outline the construction. It is more convenient to work geometrically, when the contravariant functor is replaced by the functor from the category of affine Kato fans over 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 -fans with the morphisms being face embeddings, then the colimit exists as a Kato fan. Intuitively, it is a -“fan” which contains each as a face in a unique way. It is not so difficult to show that , in particular, the morphism is monoidal (in the stacky sense). Moreover, one can now give an explicit stacky presentation of . 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 -monoids by a join face we mean a -monoid and face embeddings with (in other words, we fix isomorphisms of and localizations of ). Show that the colimit of the diagram of all face embeddings of is a Kato fan, which we call the join of .
(ii) Construct a natural simplicial Kato fan and show that it is in fact a groupoid equivalent to a Kato fan which we denote . In other words, is a cover and its fiber powers are .
(iii) Show that is characterized by the following universal property: any -fan possesses a unique face embedding into .
(iv) Show that and deduce that is equivalent to the simplicial stack , which is, in fact, a groupoid.
5.2. Stacks 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 be a property of morphisms of schemes, for example, smooth, étale, flat, etc. A morphism of log schemes is said to be log (resp. weakly log ) if the associated morphism of stacks (resp. ) is .
Remark 5.2.2.
(i) Both definitions have some advantages. The morphism is “smaller” and easier to analyze; if is quasi-compact, then a morphism factors through an open substack of finitely presented over . On the other hand, when studying compositions it is certainly easier to work with the morphisms .
(ii) Despite the terminology, neither condition implies the other one. Olsson showed in [Ols03, Example 4.3] that if is “having geometrically connected fibers”, then log does not imply weakly log , but there is even a much more basic example: any morphism with a quasi-compact source is not weakly log surjective because is never quasi-compact, while is often log surjective, for example, when it is an isomorphism.
However, using that factors into the composition of an open immersion and we obtain that if is local on the source, then log implies weakly log .
(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.
Now we can also naturally interpret the notion of a log étale cover:
Exercise 5.2.4.
Prove that a morphism is a log étale cover if and only if 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 is stable under pullbacks, étale local on the source, and flat local on the base, then a morphism is log if and only if it is weakly log . This follows from a slightly surprising fact that can be obtained from its small piece by base change and flat descent.
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 also allow to interpret logarithmic derivations and differentials. In view of the fact that a morphism is log smooth if and only if the associated morphism is smooth, the following fact is very natural: and (see [ATW20, Lemma 2.4.4]).
Exercise 5.2.9.
(i) Assume that possesses a chart . Prove that . (Hint: compare the first fundamental sequences of log derivations associated with and of derivations associated with .)
(ii) Use the étale cover to deduce that, indeed, for any morphism of log schemes one has that and hence also .
5.2.10. Logarithmic fibers
Let be a morphism of log schemes. By the log fibers of we mean the connected components of the fibers of the induced morphism . For example, if 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 has a chart modeled on , then factors through the étale morphism hence the log fibers are nothing else but the fibers of the stacky chart .
Let us consider two general types of examples of opposite kind. If the homomorphisms are injective, then the log fibers have the most natural geometric interpretation:
Exercise 5.2.11.
Assume that has a chart modeled on with sharp and (such a chart exists étale-locally at with if is injective). Show that the log fibers of are the connected components of the log strata of the fibers of . 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 are nothing else but the points of (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, -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).