RIEMANN-ROCH FOR ††thanks: Research supported by The Simons Foudation
Abstract
We prove a Riemann-Roch theorem of an entirely novel nature for divisors on the Arakelov compactification of the algebraic spectrum of the integers. This result relies on the introduction of three key concepts: the cohomologies (attached to a divisor), their integer dimension, and Serre duality. These notions directly extend their classical counterparts for function fields. The Riemann-Roch formula equates the (integer valued) Euler characteristic of a divisor with a slight modification of the traditional expression in terms of the sum of the degree of the divisor and the logarithm of 2. Both the definitions of the cohomologies and of their dimensions rely on a universal arithmetic theory over the sphere spectrum that we had previously introduced using Segal’s Gamma rings. By adopting this new perspective we can parallel Weil’s adelic proof of the Riemann-Roch formula for function fields including the use of Pontryagin duality.
keywords:
Riemann-Roch; Arakelov compactification; Adeles; Segal’s Gamma-ringkeywords:
[2010 Mathematics subject classification]14C40; 14G40; 14H05; 11R56; 13F35; 18N60; 19D55A. Connes and C. Consani
1 Introduction
The development of the theory of curves from complex Riemann surfaces to algebraic curves over arbitrary fields, led F. K. Schmidt, in the 1930s, to the first proof of the Riemann-Roch theorem for function fields over finite fields. In nowadays terminology, that result involves the integer dimensions of the two cohomologies and of a divisor on the algebraic curve associated to the function field, as vector spaces over the (finite) field of constants. The analogy between function fields in one variable over fields of positive characteristic and number fields suggests, as pointed out by A. Weil [13], to investigate the existence of a Riemann-Roch formula for number fields. Both the notions of a divisor with its degree, and the finite set underlying are easy to define. These ideas led S. Lang [9] to an asymptotic formula that relates for , and , when . Another attempt to a Riemann-Roch formula for a number field was promoted two decades ago by G. van der Geer and R. Shoof [11] and more recently, in the work of J.B. Bost [1]. In this approach one starts from the functional equation in terms of the theta function and rewrites that equation as a Riemann-Roch formula for the log-theta number, which is thus promoted to the status of a dimension. This view of the log-theta number as a dimension remains virtual for the obvious reason that it outputs real numbers rather than integers. An interpretation of these numbers as Murray-von Neumann dimensions could greatly improve their status as dimensions, but this step has not been achieved so far.
The most fundamental number field is the field of rational numbers that governs elementary arithmetics through its ring of integers. In this paper we prove a Riemann-Roch theorem of an entirely novel nature for this basic arithmetic structure. This result relies on the introduction of three key concepts: the two cohomologies (attached to a divisor ), their (integer) dimension, and Serre duality. These three notions directly extend their classical analogues for function fields.
Our main result is the following theorem.
Theorem 1.1.
Let be an Arakelov divisor on . Then
(1.1) |
Here, denotes the odd function on that agrees with the ceiling function on positive reals, and is the characteristic function of the exceptional set of finite Lebesgue measure which is the union of the intervals :
In this formula, the neperian logarithm that is traditionally used to define the degree of a divisor , is replaced by the logarithm in base .111All throughout the paper, to avoid confusion, we shall nevertheless use the neperian logarithm This alteration is equivalent to the division by i.e. .
Let us now comment on the key role played by the number . In the proof of Theorem 1.1, the ring
appears in a new light as a ring of polynomials with coefficients in the absolute base . At an elementary level, it is not difficult to check that any integer can be written uniquely as a finite sum of powers of with coefficients in . The deep reason behind this fact is that, for (and only for this rational prime) the Witt vectors with only finitely many non-zero components
form a subring inside the ring of -adic integers, which turns out to be isomorphic to . In particular, the formula for the addition of integers written as polynomials coincides with the formula for the addition of Witt vectors over .
Next, we describe the general formalism that allows us to give a precise mathematical meaning both to the base and (to) the two cohomologies with their integer-valued dimension. The basic step underlying this formalism is taken by going beyond the traditional use of abelian groups and rings in algebra using Segal’s -rings. In [2, 5] we developed the fundamentals of a geometric theory where the initial ring of the category of rings is replaced by the sphere spectrum . A similar change of structures is familiar in homotopy theory, but in our work we insist on staying at a concrete computational level in the definition of the basic objects, while the link with the -spaces of homotopy theory [6] enters only when doing homological algebra. The extension of the category of abelian groups is obtained by embedding faithfully and fully this category in the category of pointed covariant functors from finite pointed sets to pointed sets. Given an abelian group one assigns to a finite pointed set the pointed set of -valued divisors on . These divisors push forward by summing over the pre-image of a point. In the traditional notion of ring becomes that of Segal’s -ring, and the (classical) base is replaced by the -ring here understood in its most elementary form of identity functor (from finite pointed sets to pointed sets). The base is in fact the most elementary categorical form of the sphere spectrum in homotopy theory [6].
In analogy with the theory of algebraic curves over finite fields, we
have (previously) defined the structure sheaf of the one-point compactification of the algebraic spectrum of the integers by suitably extending, at the archimedean place, the structure sheaf of as a subsheaf of the constant sheaf . The global sections of this sheaf determine the -algebra . In [3] we have generalized the notion of homology for simplicial complexes, when group coefficients are replaced by -modules. In [4] we implemented homological algebra in this context by showing an extension of the Dold-Kan correspondence which associates a -space to a short complex of -modules.
Let us now introduce the cohomologies and their dimensions.
An Arakelov divisor on determines a compact subset of the adeles of . This is the product, indexed by the places of , of the abelian group for each finite prime , the abelian groups , and the interval . This last component is implemented by a sub-module of the Eilenberg-MacLane -module which functorially encodes the additive structure of the group . The morphisms of abelian groups used by Weil in the development of the adelic geometry of function fields still retain a meaning in this absolute set-up and they are viewed as morphisms of -modules. We focus, in particular, on the morphism
(1.2) |
where is embedded diagonally in the adeles. The -space associated by the Dold-Kan correspondence to (1.2) ( is viewed here as a short complex of -modules: see appendix B) provides the absolute incarnation of the Riemann-Roch problem for the divisor . By implementing linear equivalence i.e. the multiplicative action of on , one is reduced to consider only the case , for which is the -module that associates to a finite pointed set the -valued divisors , , fulfilling the condition ( euclidean absolute value). By definition, is a covariant functor from the small category (a skeleton of the category of finite pointed sets) to pointed sets. It keeps track of the partially defined addition in , so that for , the sum is meaningful provided . The dimension is defined to be the smallest number of linear generators of . More precisely, a subset linearly generates if and only if for every there exist coefficients such that , , and (see section 3). The -module only depends upon the integer part of and Proposition 3.3 determines its dimension to be222For the values , every element of is uniquely written in terms of the generating set (Proposition 3.3 ).
(1.3) |
The cokernel of in (1.2), that is , is defined in terms of the -space . The lack of transitivity of the homotopy relation in non-Kan complexes is dealt with using the classical notion of tolerance relation [12]. Thus is the couple of the -module and a suitable tolerance relation on it (see Proposition 2.2 and appendix A). This relation is defined by pairs whose difference belongs to the image of . More specifically, defines an object of the category of tolerance -modules (see appendix A). Its dimension is defined to be the smallest number of linear generators where a subset linearly generates if and only if for every there exists coefficients such that the pair belongs to the relation (see section 4).
An attentive reader should readily recognize that our construction strictly parallels Weil’s adelic proof of the Riemann-Roch formula for function fields and, in particular, that our notion of dimension is a straigthforward generalization of the classical dimension for vector spaces. For an archimedean divisor , the dimension of is the same as the dimension of the pair , where the relation on is given by
(1.4) |
for the translation invariant Riemannian metric of length on . In Proposition 4.1 we show that this dimension is the integer
(1.5) |
Theorem 1.1 follows from (1.3) when , since in this case . The case is proved using (1.5). The symmetry of the term under the replacement of by the divisor , with , is the numerical evidence of a geometric duality holding on the curve over (see section 5). This is precisely stated by the following (see Theorem 5.3)
Theorem 1.2.
Let be an Arakelov divisor on and , with . Then there is an isomorphism of -modules.
(1.6) |
where coincides with , playing the role of the dualizing module in Pontryagin duality.
The isomorphism is the analogue of Grothendieck’s version of Serre’s duality [7] (12.15.1955). The duality (1.6) derives from Pontryagin’s duality holding for tolerant -modules (Proposition 5.2), once again in analogy with the corresponding result proven in [13] for curves over fields of positive characteristic. The divisor plays the role of the canonical divisor.
The paper is organized as follows, in section 2 we adapt to the adelic framework the construction of the -space naturally associated to an Arakelov divisor . This gives, in Proposition 2.2, the cohomologies and . We compute the dimension of in section 3 (Theorem 3.4), and of in section 4 which concludes with the proof of the main Theorem 4.3. Section 5 is devoted to Serre and Pontrjagin dualities. In appendix A we develop the needed generalities on tolerance -modules and in appendix B the technical details of the construction, through the Dold-Kan correspondence, of the -space .
2 The cohomology
We recall, from [4], the construction of the -space naturally associated to an Arakelov divisor on . This space is the absolute homological incarnation of the Riemann-Roch problem for the divisor . We formulate this construction in terms of the adeles of .
Let be an Arakelov divisor on . This is a formal finite sum with and , where are rational primes in and the symbol stands for the restriction of the euclidean absolute value . Let denote the full set of places of . To is naturally associated the following idele of
The equality defines a map such that , , and , for . To the divisor corresponds the compact subset of the adeles of
By definition with
In particular, the product is a compact abelian group. The archimedean component, on the other hand, i.e. the real interval , is not, and for this reason we implement the functorial viewpoint to suitably understand . Indeed, to we associate the covariant functor
from the opposite of the Segal category (see e.g. [6] Chpt. 2 and [2]) to pointed sets, which maps a finite pointed set to the pointed set of -valued divisors on vanishing on the base point and whose total mass is bounded by . Covariant functors and their natural transformations determine the category of -sets (aka -modules). In particular, the Eilenberg-MacLane functor encodes an abelian group as the covariant functor that associates to the pointed set of -valued divisors on vanishing on the base point. Functoriality holds by taking sums on the inverse images of a point. The functor encodes the addition on . In general, let with and
(2.1) |
Given an -module and elements , , one writes
(2.2) |
For the key point is that this general construction respects the bound on the total mass while encoding the addition. One easily sees that is an -module (an “(-algebra” as in [6] 2.1.5), where , is the spherical monoid algebra of the multiplicative monoid . In particular, the inclusion functor determines as a sub -module of , where is the identity functor.
Given two -modules , , one defines their product as the functor
with the base point of taken to be . For abelian groups one has a canonical isomorphism . In particular, the morphism of addition in adeles , (using the diagonal embedding of in ) determines a morphism of -modules . Next proposition shows how this functorial construction is well combined with the adelic formalism.
Proposition 2.1.
-
(i)
The -module
determines, canonically, a sub -module .
-
(ii)
The restriction of to defines a morphism of -modules
(2.3)
Proof 1.
With , let be the inclusion of abelian groups. Then is the product of and the inclusion .
By composition one obtains .
The morphism in (2.3) is obtained by restricting to the (adelic) divisor , where is a morphism of abelian groups. The theory developed in [4] associates to a -space , by implementing the Dold-Kan correspondence in the special case of a short complex of abelian groups and its restriction to a sub--module. The -space supplies the absolute cohomology of the divisor . We recall this construction in appendix B. Moreover, we review how the generalized homotopy , where is the category of tolerance relations defined in A.1, is meaningful in this context supplying, with and , the (tolerant) -modules descriptions of resp. the kernel and the cokernel of . In other words, one obtains the following result
Proposition 2.2.
-
(i)
The kernel of is the -module:
(2.4) -
(ii)
The cokernel of is the -module endowed with the relations
(2.5) -
(iii)
Both and only depend, up to isomorphism, on the linear equivalence class of the divisor .
Proof 2.
We refer to appendix B.1.
3 The dimension of
In this section we consider, for each integer , the -module
and evaluate its dimension. It follows from the -module structure, that (2.2) holds for a sum of the form
By applying Definition A.2, a subset generates333For simplicity we use from now on the term “generates” for “linearly generates” if and only if for every element there exists coefficients , , such that and . The dimension is the minimal cardinality of a generating set.
The explicit computation of follows from the next two statements.
Lemma 3.1.
For , let . Consider the subset . Then is a generating set and the map
(3.1) |
is bijective. Moreover , for all .
Proof 3.
First of all, note that for all one has
This shows that for any one has , and, as in (2.2), . We prove the bijectivity of by induction on . It is clear for . For , and one writes the set as the disjoint union of three subsets as follows
This shows that is bijective. Next, by assuming to have shown the bijectivity up to , we deduce it for . Indeed, with let . The induction hypothesis ensures that every element of the set can be uniquely written as a sum . We divide the set into three disjoint parts of the form
One has and thus the union of these three disjoint sets is . Moreover one has with , and similarly . Thus, given one determines uniquely the coefficients such that . Indeed, one uses the above partition in three intervals to determine the coefficient and then one applies the induction hypothesis to determine the others.
Remark 3.2.
-
1.
The conceptual explanation of Lemma 3.1 derives from the following peculiar property of the Teichmüller lift . One has and extends to a canonical map from Witt vectors to
This map coincides with in (3.1) and one can compute the partially defined sums (2.2) in using the addition in . The prime is the only prime such that the subring coincides with the ring of Witt vectors with finitely many non-zero components. When one has , while for the integer is the Witt vector whose components are all equal to . In general the Witt vectors with finitely many non-zero components do not even form a subgroup of the additive group of Witt vectors.
-
2.
The Teichmüller lift , as a morphism of multiplicative pointed monoids, induces a morphism of -algebras.
-
3.
The ordering of the natural numbers encoded by coincides with the lexicographic ordering of the coefficients
We recall that the ceiling function associates to a positive real number the smallest integer .
Proposition 3.3.
-
(i)
For any integer one has
(3.2) -
(ii)
For , there exists a generating subset of cardinality , with , such that the following map surjects onto
Proof 4.
Set . Since the cardinality of is and the cardinality of is , it follows that if is a generating set one must have . Thus one has , and since is an integer one gets . This shows that . It remains to prove that for each one can find a generating set with . One uses Lemma 3.1, and distinguishes several cases starting with the easiest one. Set and let be the subset
(3.3) |
We list the first elements of as follows
First, assume that . Then let be the largest integer such that . If , then Lemma 3.1 provides one with a generating set of cardinality . Assume now that . Then, the map for covers the set , where . Thus by adjoining to the element one covers . More precisely, let , then we show that fulfills the conditions in . One has since the additional element does not belong to as, by hypothesis, . Since , . The sum of elements in is , so the sum of elements in is equal to . Next, we show that is surjective. Its image contains the three sets , , and . Using Lemma 3.1 one obtains, with
One has . This inequality prevents the three subsets from being disjoint, thus the upper limit of is greater or equal to the lower limit of the interval . Thus is surjective.
The next case is for of the form , for some ( in (3.3)). We assume that (this excludes the cases for , and for ). Let
(This choice avoids the repetition of the element while keeping the sum of elements of equal to ). We show that fulfills the conditions in . We have since . Moreover , thus . The sum of terms in is equal to one plus the sum of the powers of up to and thus is equal to . Let us now show that the map is surjective. By construction is the union of nine intervals obtained, with , by adding to the set the terms in with coefficients in . With the term one covers the set , where as the union of the three sets
One has , and thus it remains to cover the two elements and their opposite. This is done using the set
whose upper limit is , and appealing to the fact that since this interval contains .
Finally, we consider the case of the form , with . Here, the additional element does belong to so the proof of the first case considered above does not apply since appears twice in . We replace this double occurrence of by the two distinct elements . This gives
By construction . Furthermore one has: , and since one gets . The sum of elements of is the same as the sum of elements of plus which is equal to . It remains to show that, for such , is surjective. We first deal with the set and show that the map , for this set, surjects onto the interval , where . Lemma 3.1 shows that the image of is the set , with . To show the surjectivity of it is enough to show that the following intervals cover :
The upper limit of is , its lower limit is equal to the upper limit of . The lower limit of is which is the upper limit of . Moreover, the length of is so that two translates of of the form , necessarily overlap. This shows that the map surjects onto the interval . Thus is the symmetric enlargement of the interval obtained by adding to the upper limit. Then, as in the proof of Lemma 3.1, one obtains by induction that when one adjoins the higher powers , where , one achieves the enlargement of the interval obtained by adding to the upper limit. This shows that the map for is surjective. In fact the above arguments prove and , except that one has to take care of the special cases . For , is a generating set but the sum of its terms is . For , is a generating set but the sum of its terms is .
We are now ready to determine the dimension of the -module , for any Arakelov divisor . It is a non-decreasing function of but the formula for this dimension drops by on an exceptional set . It is the open set defined as the union of open intervals as follows
(3.4) |
It has finite Lebesgue measure
where is the Pochhammer symbol at .
Theorem 3.4.
Let be an Arakelov divisor on . If , then
(3.5) |
where is the characteristic function of the open set .
Proof 5.
Let and , then . By Proposition 3.3 one gets
Let us compare this formula with (3.5). Assume first that , thus there exists such that . Then, and with , one has . Thus and . Moreover, , where , hence one has , so that (3.5) holds since . Now, assume that , equivalently that for some integer (since by hypothesis). Then . One has
In this case and thus . This implies that and hence that . In this case (3.5) holds since .
4 The dimension of
Let be an abelian group endowed with a translation invariant metric . For , we shall refer to as the associated object of the category as in Proposition A.3. The -module structure determines the addition in , as well as the action of on it, where for , is the additive inverse of . The metric determines the tolerance relation on . Next result computes the dimension (Definition A.2) of the tolerant -module .
Proposition 4.1.
Let , and where is the abelian group endowed with the canonical metric of length . Then
(4.1) |
Proof 6.
For , any element of is at distance from , thus one can take as generating set since, by convention, . Thus . Next, we assume . Let be a generating set and let . One easily sees that there are at most elements of the form . The subsets cover , and since each of them has measure one gets the inequality . Thus . For one has and the subset generates, thus . When is an integer, one has . Let . The minimal distance between two elements of is the distance between and which is . Let us show that is a generating set. By Lemma 3.1 any integer in the interval , for can be written as , with . One then gets
Let , then and there exists an integer such that . Hence . This proves that is a generating set (see Definition A.2) and one derives . Assume now that , where is an integer. For any generating set of cardinality one has so that . The subset fulfills the first condition of Definition A.2 since the minimal distance between two elements of is which is larger than . As shown above, the subset is generating for and a fortiori for (as by assumption ). Thus one obtains and (4.1) is proven.
Remark 4.2.
We extend the ceiling function to negative values of the variable as follows
(4.2) |
We can now state and prove the absolute Riemann-Roch theorem for over . With the exceptional set defined in (3.4) one has
Theorem 4.3.
Let be an Arakelov divisor on . Then
(4.3) |
Proof 7.
By appealing to the invariance under linear equivalence (Proposition 2.2 ), one may assume that . Then it follows from Proposition A.5 that for . Assume first that , then by Proposition 4.1: , thus for (4.3) follows from (3.5). Let us now assume that . Then since when . Moreover since is lower bounded by . Thus (4.3) follows from (4.1) which gives, using (4.2)
This ends the proof of (4.3).
5 Duality
In this section we prove an absolute analogue of Serre’s duality, namely the following isomorphism of -modules, for any divisor on :
Here, the divisor plays the role of the canonical divisor. The choice of the tolerant -module as dualizing module is motivated by Pontryagin duality (see 5.2). One has . This equality, in fact, also holds for the tolerant -module for : the specific choice is dictated by the invariance of the Riemann-Roch formula (4.3) when one switches and and replaces by (ignoring the exceptional set ).
5.1
We start with the following general statement.
Lemma 5.1.
Let . The -algebra structure of induces an isomorphism of -modules
(5.1) |
Proof 8.
One starts by defining the morphism of -modules
(5.2) |
Precisely, the multiplication in the -algebra determines natural maps
inducing, for a fixed pointed set , the map .
The morphism is defined as the natural transformation of functors which associates to , the map defined as the restriction of . This restriction is meaningful in view of [2] (Proposition 6.1).
Next, we show that is an isomorphism. First, we determine the -module . For integers we let as in (2.1)
Let . By construction, the natural transformation reads, for each , as a map and by naturality we have
(5.3) |
Since an element is determined by its components , , with , (5.3) shows that is uniquely determined by the map acting componentwise, i.e.
(5.4) |
Moreover, the map fulfills
(5.5) |
as one sees using the naturality of for the map , i.e. using
By (5.5) one has , for any and . Thus the ‘germ of map’ uniquely extends to a map defined by for any such that . Moreover, again by (5.5), the map is additive and since , it is also continuous and hence determined by the multiplication by a real number . One has and hence . Thus one gets , . By (5.4) one obtains , . This shows that for , is bijective. The next step is to determine for . Let . One has and an element is determined by its components for , such that . In particular, for each , the , , are the components of . This implies by applying the naturality of , that
(5.6) |
As shown above the map fulfills
(5.7) |
and it extends to an additive map which is continuous since maps the interval inside . Thus there exists real numbers , , such that . One has and thus . Finally, using (5.6) it follows that . This proves that as in (5.2) is an isomorphism.
5.2 Pontryagin duality
In order to formulate Pontryagin duality in this context we consider, for , the -module , where is the abelian group endowed with its canonical metric of length . For a metric abelian group , we denote by the abelian group of continuous characters, i.e. of continuous group homomorphisms , where is endowed with the topology associated to the metric . We retain the notations of section A. Next statement is motivated by Lemma 5.1.
Proposition 5.2.
Let be an abelian group endowed with a translation invariant metric .
-
(i)
For , the -module is isomorphic to the sub -module of which, on the set , is given by k-tuples , of continuous characters such that, with , and for all finite collections , fulfill
(5.8) -
(ii)
Let be a surjective morphism of abelian groups and let be the pullback as in Proposition A.4. One has the following canonical isomorphism
(5.9)
Proof 9.
Let . By applying the forgetful functor which associates to the set (see Proposition A.2), one obtains an element . Since the Eilenberg-MacLane functor determines a full and faithful embedding of the category of abelian groups inside the category of -modules, there exists a unique group homomorphism such that . The condition that preserves the relation on the set means that for any , such that one has . By translation invariance of the metrics this condition is equivalent to
(5.10) |
This shows that consists exactly of the group homomorphisms fulfilling (5.8). Specializing (5.10) to the case where all , , one obtains the implication , and hence that is uniformly continuous. Let . The object of is , where the metric on the product group is defined by
Replacing with in the first part of the proof one obtains that , where , is a -tuple of characters of fulfilling (5.8). It follows that .
Let . As in the proof of , there exists a group homomorphism such that . Moreover preserves the relation for any , and this implies
In particular, taking all one obtains , and hence . This implies that there exists a group homomorphism such that .
We can now state and prove Serre’s duality.
Theorem 5.3.
Let be an Arakelov divisor on . There is a canonical isomorphism of -modules
(5.11) |
where is the divisor .
Proof 10.
By Proposition A.5, with , one has and by Proposition 5.2, , one gets the isomorphism
In fact, we can assume that , with so that . Then we apply Proposition 5.2 , with and . One has and the characters are given by multiplication by , i.e. , . Next, we need to determine the -submodule of which, on the set , is given by -tuples , of characters such that (5.8) holds. This means, using the distance on , that
(5.12) |
The distance is given, for any in the class of by the distance between and the closed subset . Thus for any integer one has: . Assume that , then (5.12) follows since
Conversely, assume (5.12). Then repeating times the same , gives
Taking large enough and one obtains , and hence . This proves that the -submodule of determined by (5.8) is equal to which gives (5.11).
Appendix A Tolerance -modules
The construction of the category of -spaces (see appendix B) can be broadly generalized by considering in place of the category of simplicial pointed sets any pointed category with initial and final object . In this way, one obtains a category of pointed covariant functor . We shall apply this formal construction to the category of tolerance relations and introduce the notion of tolerant -modules which plays a central role in the development of the absolute Riemann-Roch problem. We start with the following general fact
Lemma A.1.
Let be a pointed category with initial and final object . Then is naturally enriched in -modules. More precisely, the following formula endows the internal with a structure of -module defined by
(A.1) |
Proof 11.
Let . For every object of the morphism gives, by functoriality of , a morphism . These morphisms define a natural transformation of functors , and one obtains the functoriality on the right hand side of (A.1) using the left composition
A.1 The category
A tolerance relation on a set is a reflexive and symmetric relation on . Equivalently, is a subset which is symmetric and containing the diagonal. We shall denote by the category of tolerance relations . Morphisms in are defined by
We denote the pointed category under the object endowed with the trivial relation. One has the following
Definition A.1.
A tolerant -module is a pointed covariant functor . We denote by the category of tolerant -modules.
Next statement is an easy but useful fact
Proposition A.2.
-
(i)
The functor which endows a set with the diagonal relation, embeds the category of sets as a full subcategory of , and consequently the category of -modules as a full subcategory of the category .
-
(ii)
The forgetful functor is the right adjoint of the inclusion in .
A.2 The tolerant -module
A relevant example of tolerant -module is given by the following construction. Let be an additive abelian group. A translation invariant metric on is a metric on that satisfies , so that the triangle inequality can be read as . This fact implies that the inequality
(A.2) |
holds for any finite index set and maps .
Proposition A.3.
Let be an abelian group endowed with a translation invariant metric and let . The following relations turn the Eilenberg?MacLane -module into a tolerant -module :
(A.3) |
Proof 12.
For any , the map fulfills . Indeed, this follows from (A.2) applied to the finite sets which label pairs of elements of .
Let be the abelian group endowed with the canonical metric of length . We shall denote by the tolerant -module , ().
A.3 The dimension of a tolerant -module
In this part we introduce a notion of dimension for a tolerant -module that naturally generalizes, in the absolute context, the definition of dimension of a vector space.
Definition A.2.
Let be a tolerant -module. A subset generates if the following two conditions hold
-
1.
For , with
-
2.
For every there exists , and such that in the sense of (2.2), and .
The dimension is defined as the minimal cardinality of a generating set .
To familiarize with this notion we prove the following
Proposition A.4.
Let be a morphism of abelian groups and a tolerant -module.
-
(i)
Consider the relation . Then the pair is a tolerant module.
-
(ii)
If is surjective: .
Proof 13.
Since is a tolerant -module, for any one has . Also
which shows that is a tolerant module.
Let be a generating set for , and let be a lift of , with . Let us show that is a generating set for . Let , then there exists coefficients , , such that . It follows that using the lifts of one has , hence is a generating set for . Thus . Conversely, let be a generating set for and . Then condition 1. of Definition A.2 for implies the same condition for , thus one has . Let and with . Then there exists coefficients , , such that . This implies
so that is a generating set for .
Next, we apply this functorial machinery to the geometry of . We retain the notations of section 2.
Proposition A.5.
Let be an Arakelov divisor on and let be the projection of the adeles on their archimedean component modulo the lattice
(A.4) |
-
(i)
Let be the metric on induced by the standard metric on and set . Then one has .
-
(ii)
.
Proof 14.
Let , , be the embedding of finite adeles in adeles. Using the ultrametric property of the local norms at the finite places one sees that is a compact subgroup. Set : one has since all the adeles in have archimedean component equal to . Thus, the restriction of the morphism of addition, , , to determines an isomorphism of with the subgroup of . Note, in particular, that is closed in , since is discrete (hence closed) and is compact. In the following, we identify (set-theoretically) with the product endowed with the two projection morphisms and . The subgroup is dense and the subgroup is open. Hence . Thus the projection induces the isomorphism of groups , where . The kernel of the composite is the group of pairs such that . Such pairs are determined by the value of and thus
(A.5) |
By Proposition 2.2, is the tolerant -module where the relations are given by (2.5), i.e.
and where is as in (2.3). By construction one has . After quotienting both sides of (2.3) by , the map becomes
(A.6) |
where is the lattice (A.5). Thus one obtains that the relation is equal to the inverse image by the map of the relation (A.3).
Follows from Proposition A.4 .
Appendix B The -space
Let be a morphism of abelian groups. To one associates the following (short) complex of abelian groups indexed in non-negative degrees
(B.1) |
The Dold-Kan correspondence associates to the simplicial abelian group (see [8] III.2, Proposition 2.2)
(B.2) |
where denotes the simplicial category. The direct sum in (B.2) repeats the term of the chain complex as many times as the number of elements of the set of surjective morphisms . For the short complex , the allowed values of are . Therefore the set is reduced to , i.e. to a single point. A morphism is characterized by which is an hereditary subset of . It follows that the vertices in are labelled by the
(B.3) |
For each integer the finite set of surjective elements excludes and , thus the set has elements. This gives the identification
(B.4) |
We refer to [8] (III.2, pp. 160-161) for a detailed description of the simplicial structure, namely the definition for each of a map of sets . Next, we introduce some notations.
We identify the opposite of the simplicial category with (the skeleton of) the category of finite intervals. An interval is a totally ordered set with the smallest element distinct from the largest one. A morphism of intervals is a non-decreasing map that preserves the smallest and largest elements. The canonical contravariant functor which identifies the opposite category of with (described by intervals as above), maps the finite ordinal object in to the interval .
We denote by the category of pairs of pointed sets , with . The morphisms are maps of pairs of pointed sets. We let be the functor of collapsing to the base point .
Let be the functor that replaces an interval with the pair , where is the set pointed by its smallest element, and is the subset formed by the smallest and largest elements of .
Finally, we denote by (see section A) the functor that associates to an object of the covariant functor
The following formula defines a covariant functor that associates to a pair of pointed sets an abelian group directly related to the morphism
(B.5) |
On morphisms in with , the functor acts as follows
(B.6) | ||||
By [4] (Proposition 4.5), the Dold-Kan correspondence for the short complex (B.1), i.e. the simplicial abelian group in (B.4) is canonically isomorphic to the composite functor , with as in (B.5). By composing with the Eilenberg-MacLane functor one obtains a covariant functor to the category of -modules, which is naturally isomorphic to the functor , where is the forgetful functor (see op.cit. Lemma 4.6). Moreover, as shown in op.cit. (Theorem 4.7) the -space associated by the Dold-Kan correspondence to the complex is canonically isomorphic to the functor
(B.7) |
Notice that the above construction involves the morphism only through the composite functor . This latter functor is still meaningful when one restricts to a sub--module of the -module and it is given by
(B.8) |
This provides the following construction
Definition B.1.
Let be a morphism of abelian groups and a sub -module of the -module . We denote by the -space obtained as a sub-functor of (B.7)
(B.9) |
When evaluated on the object of , the -space defines a sub-simplicial set of the Kan simplicial set in (B.4). However it is not in general a Kan simplicial set, thus one needs to exert care when considering its homotopy. We refer to [3] 2.1, for the definition of the homotopies used there. The relation of homotopy between -simplices is defined as follows
(B.10) |
In general, the relation in (B.10) fails to be an equivalence relation. In place of the quotient we consider pairs of sets and relations. We define to be the set of spherical elements in (i.e. of -simplices , with ), endowed with the relation . Then we have the following result (we refer to section A for the notion of tolerant module)
Proposition B.1.
Let be a morphism of abelian groups and let be a sub--module of the Eilenberg-MacLane -module (where ).
-
(i)
The homotopy is the sub--module of
(B.11) -
(ii)
The homotopy is the tolerant -module endowed with the relations
(B.12)
Proof 15.
By construction, is the composite of the functors of (B.8) and . One has where with base point . Thus and by (B.8)
One has , where , with base point . Thus one has and , while and by (B.8) it follows that
(B.13) |
The boundaries are obtained as in [4] Proposition 4.11
(B.14) |
The spherical condition on means that and . Thus the solutions correspond to as in (B.11). One shows as in [4] Proposition 4.11 , that the relation of homotopy is the identity.
The relation on elements of is defined as follows
(B.15) |
Note that this relation is a tolerance relation i.e. is symmetric since is a sub -module of the -module , so that
The above general construction applies to the geometric adelic context and by implementing the action of on adeles, we obtain the following variant of Proposition 4.9 in [4].
Proposition B.2.
Let be an Arakelov divisor on . Let , and , . Let be the sub -module of as in Proposition 2.1. Then the functor
(B.16) |
defines a -space that depends only on the linear equivalence class of .
B.1 Proof of Proposition 2.2
With the notations of section A and Proposition B.2, one has . Proposition B.1, , gives
An element is a -tuple with for all , and , where . The condition means for all , so that is uniquely determined by the -tuple . Moreover, the allowed -tuples are those for which and for all . One has
This proves of Proposition 2.2. The statement follows from Proposition B.1 . Finally, follows from Proposition B.2.
Acknowledgments
The second author is partially supported by the Simons Foundation collaboration grant n. 691493 and thanks IHES for the hospitality, where part of this research was done.
References
- [1] J. B. Bost, Theta invariants of Euclidean lattices and infinite-dimensional Hermitian vector bundles over arithmetic curves. Progress in Mathematics, 334.
- [2] A. Connes, C. Consani, Absolute algebra and Segal’s Gamma sets, J. Number Theory 162 (2016), 518–551.
- [3] A. Connes, C. Consani, and the Gromov norm, Theory and Applications of Categories, 35, No. 6, (2020), pp. 155–178.
- [4] A. Connes, C. Consani, Segal’s Gamma rings and universal arithmetic, The Quarterly Journal of Mathematics, 72, 1-2, (2021), pp. 1–29.
- [5] A. Connes, C. Consani, On Absolute Algebraic Geometry, the affine case, Advances in Mathematics, 390, Paper No. 107909 (2021), 44 pp.
- [6] Dundas, Bjorn Ian; Goodwillie, Thomas G.; McCarthy, Randy The local structure of algebraic K-theory, Algebra and Applications, 18. Springer-Verlag London, Ltd., London, 2013.
- [7] Correspondance Grothendieck-Serre. Edited by Pierre Colmez and Jean-Pierre Serre. Documents Mathématiques (Paris), 2. Société Mathématique de France, Paris, 2001. xii+288 pp.
- [8] Goerss, Paul G.; Jardine, John F. Simplicial homotopy theory. Progress in Mathematics, 174. Birkhauser Verlag, Basel, 1999.
- [9] S. Lang, Algebraic Number Theory. Addison Wesley, 1970.
- [10] J. Neukirch, Algebraic number theory. Translated from the 1992 German original and with a note by Norbert Schappacher. With a foreword by G. Harder. Grundlehren der Mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences], 322. Springer-Verlag, Berlin, 1999.
- [11] G. van der Geer, R. Schoof, Effectivity of Arakelov divisors and the theta divisor of a number field, Selecta Math. (N.S.) 6 (2000), no. 4, 377–398.
- [12] H. Poincaré, Science and Hypothesis (with a preface by J.Larmor ed.) (1905) New York: 3 East 14th Street: The Walter Scott Publishing Co., Ltd. pp. 22–23.
- [13] A. Weil Sur l’analogie entre les corps de nombres algébriques et les corps de fonctions algébriques, Oeuvres scientifiques/Collected papers I. 1926–1951. Springer, Heidelberg, 2014.
- [14] A. Weil Basic number theory. Reprint of the second edition (1973). Classics in Mathematics. Springer-Verlag, Berlin, 1995
- [15] http://solbakkn.com/math/triadic-nums.htm