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

{Frontmatter}

RIEMANN-ROCH FOR Spec¯{\overline{{\rm Spec\,}{\mathbb{Z}}}}thanks: Research supported by The Simons Foudation

ALAIN CONNES    CATERINA CONSANI alain@connes.org \orgnameCollège de France \orgaddress\cityParis, \stateF-75005 France \orgnameIHES \orgaddress\cityBures-sur-Yvette \state91440 France cconsan1@jhu.edu \orgdivDepartment of Mathematics \orgnameThe Johns Hopkins University, \orgaddress\cityBaltimore \state21218 USA
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-ring
keywords:
[2010 Mathematics subject classification]14C40; 14G40; 14H05; 11R56; 13F35; 18N60; 19D55
\authormark

A. 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 H0(D)H^{0}(D) and H1(D)H^{1}(D) of a divisor DD 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 DD with its degree, and the finite set underlying H0(D)H^{0}(D) are easy to define. These ideas led S. Lang [9] to an asymptotic formula that relates for {\mathbb{Q}}, log#H0(D)\log\#H^{0}(D) and degD+log2\deg D+\log 2, when degD\deg D\to\infty. 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 {\mathbb{Q}} of rational numbers that governs elementary arithmetics through its ring {\mathbb{Z}} 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 H(D)H^{\bullet}(D) (attached to a divisor DD), 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 DD be an Arakelov divisor on Spec¯{\overline{{\rm Spec\,}{\mathbb{Z}}}}. Then

dim𝕊[±1]H0(D)dim𝕊[±1]H1(D)=degD+log2𝟏L.\dim_{{{{\mathbb{S}}}[\pm 1]}}H^{0}(D)-\dim_{{{{\mathbb{S}}}[\pm 1]}}H^{1}(D)=\bigg{\lceil}\deg^{\prime}D+\log^{\prime}2\bigg{\rceil}^{\prime}-\mathbf{1}_{L}. (1.1)

Here, x\lceil x\rceil^{\prime} denotes the odd function on {\mathbb{R}} that agrees with the ceiling function on positive reals, and 𝟏L\mathbf{1}_{L} is the characteristic function of the exceptional set of finite Lebesgue measure which is the union of the intervals (log3k2,log3k+12)(\log^{\prime}\frac{3^{k}}{2},\log^{\prime}\frac{3^{k}+1}{2}):

In this formula, the neperian logarithm that is traditionally used to define the degree of a divisor D=jaj{pj}+a{}D=\sum_{j}a_{j}\{p_{j}\}+a\{\infty\}, is replaced by the logarithm in base 33.111All throughout the paper, to avoid confusion, we shall nevertheless use the neperian logarithm This alteration is equivalent to the division by log3\log 3 i.e. deg(D):=deg(D)/log3\deg^{\prime}(D):=\deg(D)/\log 3. Let us now comment on the key role played by the number 33. In the proof of Theorem 1.1, the ring {\mathbb{Z}} appears in a new light as a ring of polynomials aj3j\sum a_{j}3^{j} with coefficients in the absolute base 𝕊[±1]={0,±1}{{{\mathbb{S}}}[\pm 1]}=\{0,\pm 1\}. At an elementary level, it is not difficult to check that any integer nn\in{\mathbb{Z}} can be written uniquely as a finite sum of powers of 33 with coefficients in {0,±1}\{0,\pm 1\}. The deep reason behind this fact is that, for p=3p=3 (and only for this rational prime) the Witt vectors with only finitely many non-zero components form a subring inside the ring of pp-adic integers, which turns out to be isomorphic to {\mathbb{Z}}. In particular, the formula for the addition of integers written as polynomials aj3j\sum a_{j}3^{j} coincides with the formula for the addition of Witt vectors over 𝔽3{\mathbb{F}}_{3}.
Next, we describe the general formalism that allows us to give a precise mathematical meaning both to the base 𝕊[±1]={0,±1}{{{\mathbb{S}}}[\pm 1]}=\{0,\pm 1\} and (to) the two cohomologies H(D)H^{\bullet}(D) 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 Γ\Gamma-rings. In [2, 5] we developed the fundamentals of a geometric theory where the initial ring {\mathbb{Z}} of the category of rings is replaced by the sphere spectrum 𝕊{{\mathbb{S}}}. 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 Γ\Gamma-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 𝕊Mod{{{\mathbb{S}}}-{\rm{Mod}}} of pointed covariant functors from finite pointed sets to pointed sets. Given an abelian group HH one assigns to a finite pointed set XX the pointed set of HH-valued divisors on XX. These divisors push forward by summing over the pre-image of a point. In 𝕊Mod{{{\mathbb{S}}}-{\rm{Mod}}} the traditional notion of ring becomes that of Segal’s Γ\Gamma-ring, and the (classical) base {\mathbb{Z}} is replaced by the Γ\Gamma-ring 𝕊{{\mathbb{S}}} here understood in its most elementary form of identity functor (from finite pointed sets to pointed sets). The base 𝕊{{\mathbb{S}}} 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 Spec¯{\overline{{\rm Spec\,}{\mathbb{Z}}}} of the algebraic spectrum of the integers by suitably extending, at the archimedean place, the structure sheaf of Spec{{\rm Spec\,}{\mathbb{Z}}} as a subsheaf of the constant sheaf {\mathbb{Q}}. The global sections of this sheaf determine the 𝕊{{\mathbb{S}}}-algebra 𝕊[±1]{{{\mathbb{S}}}[\pm 1]}. In [3] we have generalized the notion of homology for simplicial complexes, when group coefficients are replaced by 𝕊{{\mathbb{S}}}-modules. In [4] we implemented homological algebra in this context by showing an extension of the Dold-Kan correspondence which associates a Γ\Gamma-space to a short complex of 𝕊{{\mathbb{S}}}-modules.
Let us now introduce the cohomologies and their dimensions. An Arakelov divisor D=jaj{pj}+a{}D=\sum_{j}a_{j}\{p_{j}\}+a\{\infty\} on Spec¯{\overline{{\rm Spec\,}{\mathbb{Z}}}} determines a compact subset 𝒪(D)𝔸{\mathcal{O}}(D)\subset{\mathbb{A}}_{\mathbb{Q}} of the adeles of {\mathbb{Q}}. This is the product, indexed by the places of {\mathbb{Q}}, of the abelian group pp{\mathbb{Z}}_{p}\subset{\mathbb{Q}}_{p} for each finite prime p{pj}jp\notin\{p_{j}\}_{j}, the abelian groups pjajpjp_{j}^{-a_{j}}{\mathbb{Z}}_{p_{j}}, and the interval [ea,ea][-e^{a},e^{a}]\subset{\mathbb{R}}. This last component is implemented by a sub-module of the Eilenberg-MacLane 𝕊[±1]{{{\mathbb{S}}}[\pm 1]}-module HH{\mathbb{R}} which functorially encodes the additive structure of the group {\mathbb{R}}. 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 𝕊[±1]{{{\mathbb{S}}}[\pm 1]}-modules. We focus, in particular, on the morphism

ψ:×𝒪(D)𝔸ψ(q,a)=q+aq,a𝒪(D),\psi:{\mathbb{Q}}\times{\mathcal{O}}(D)\to{\mathbb{A}}_{\mathbb{Q}}\quad\psi(q,a)=q+a\quad\forall q\in{\mathbb{Q}},~a\in{\mathcal{O}}(D), (1.2)

where {\mathbb{Q}} is embedded diagonally in the adeles. The Γ\Gamma-space 𝐇(D){\bf H}(D) associated by the Dold-Kan correspondence to (1.2) (ψ\psi is viewed here as a short complex of 𝕊[±1]{{{\mathbb{S}}}[\pm 1]}-modules: see appendix B) provides the absolute incarnation of the Riemann-Roch problem for the divisor DD. By implementing linear equivalence i.e. the multiplicative action of ×{\mathbb{Q}}^{\times} on 𝔸{\mathbb{A}}_{\mathbb{Q}}, one is reduced to consider only the case D=a{}D=a\{\infty\}, for which ker(ψ)\ker(\psi) is the 𝕊[±1]{{{\mathbb{S}}}[\pm 1]}-module H0(D)=HeaH^{0}(D)=\|H{\mathbb{Z}}\|_{e^{a}} that associates to a finite pointed set XX the {\mathbb{Z}}-valued divisors inixi\sum_{i}n_{i}x_{i}, xiXx_{i}\in X, fulfilling the condition i|ni|ea\sum_{i}|n_{i}|\leq e^{a} (||=|\cdot|= euclidean absolute value). By definition, H0(D)H^{0}(D) is a covariant functor Γo𝔖𝔢𝔱𝔰\Gamma^{o}\longrightarrow\mathfrak{Sets}_{*} from the small category Γo\Gamma^{o} (a skeleton of the category of finite pointed sets) to pointed sets. It keeps track of the partially defined addition in I=Hea(1+)=[ea,ea]I=\|H{\mathbb{Z}}\|_{e^{a}}(1_{+})=[-e^{a},e^{a}]\cap{\mathbb{Z}}, so that for niIn_{i}\in I, the sum ni\sum n_{i} is meaningful provided i|ni|ea\sum_{i}|n_{i}|\leq e^{a}. The dimension dim𝕊[±1]H0(D)\dim_{{{{\mathbb{S}}}[\pm 1]}}H^{0}(D) is defined to be the smallest number of linear generators of II. More precisely, a subset FIF\subset I linearly generates if and only if for every mIm\in I there exist coefficients α(f){1,0,1}=𝕊[±1](1+)\alpha(f)\in\{-1,0,1\}={{{\mathbb{S}}}[\pm 1]}(1_{+}) such that m=α(f)fm=\sum\alpha(f)f, fFf\in F, and |α(f)f|ea\sum|\alpha(f)f|\leq e^{a} (see section 3). The 𝕊[±1]{{{\mathbb{S}}}[\pm 1]}-module H0(D)H^{0}(D) only depends upon the integer part n=ean=\lfloor e^{a}\rfloor of eae^{a} and Proposition 3.3 determines its dimension to be222For the values n=4,13,40,,(3k1)/2n=4,13,40,\ldots,(3^{k}-1)/2, every element of II is uniquely written in terms of the generating set F={3i0i<k}F=\{3^{i}\mid 0\leq i<k\} (Proposition 3.3 (iii)(iii)).

dim𝕊[±1](Hn)=log(2n+1)log3.\dim_{{{{\mathbb{S}}}[\pm 1]}}(\|H{\mathbb{Z}}\|_{n})=\bigg{\lceil}\frac{\log(2n+1)}{\log 3}\bigg{\rceil}. (1.3)

The cokernel of ψ\psi in (1.2), that is H1(D)H^{1}(D), is defined in terms of the Γ\Gamma-space 𝐇(D){\bf H}(D). The lack of transitivity of the homotopy relation in non-Kan complexes is dealt with using the classical notion of tolerance relation [12]. Thus H1(D)H^{1}(D) is the couple of the 𝕊[±1]{{{\mathbb{S}}}[\pm 1]}-module H𝔸H{\mathbb{A}}_{\mathbb{Q}} and a suitable tolerance relation {\mathcal{R}} on it (see Proposition 2.2 and appendix A). This relation is defined by pairs whose difference belongs to the image of ψ\psi. More specifically, H1(D)H^{1}(D) defines an object of the category Γ𝒯\Gamma{\mathcal{T}}_{*} of tolerance 𝕊{{\mathbb{S}}}-modules (see appendix A). Its dimension dim𝕊[±1]H1(D)\dim_{{{{\mathbb{S}}}[\pm 1]}}H^{1}(D) is defined to be the smallest number of linear generators where a subset FH𝔸(1+)=𝔸F\subset H{\mathbb{A}}_{\mathbb{Q}}(1_{+})={\mathbb{A}}_{\mathbb{Q}} linearly generates if and only if for every xH𝔸(1+)x\in H{\mathbb{A}}_{\mathbb{Q}}(1_{+}) there exists coefficients α(f){1,0,1}\alpha(f)\in\{-1,0,1\} such that the pair (x,α(f)f)(x,\sum\alpha(f)f) belongs to the relation {\mathcal{R}} (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 D=a{}D=a\{\infty\}, the dimension of H1(D)H^{1}(D) is the same as the dimension of the pair (H(/),)(H({\mathbb{R}}/{\mathbb{Z}}),{\mathcal{R}}), where the relation {\mathcal{R}} on H(/)(1+)=/H({\mathbb{R}}/{\mathbb{Z}})(1_{+})={\mathbb{R}}/{\mathbb{Z}} is given by

(x,y)d(x,y)ea,(x,y)\in{\mathcal{R}}\iff d(x,y)\leq e^{a}, (1.4)

for dd the translation invariant Riemannian metric of length 11 on /{\mathbb{R}}/{\mathbb{Z}}. In Proposition 4.1 we show that this dimension is the integer

dim𝕊[±1](H(/),)=alog2log3.\dim_{{{{\mathbb{S}}}[\pm 1]}}(H({\mathbb{R}}/{\mathbb{Z}}),{\mathcal{R}})=\bigg{\lceil}\frac{-a-\log 2}{\log 3}\bigg{\rceil}. (1.5)

Theorem 1.1 follows from (1.3) when degDlog2\deg D\geq-\log 2, since in this case dim𝕊[±1]H1(D)=0\dim_{{{{\mathbb{S}}}[\pm 1]}}H^{1}(D)=0. The case degD<log2\deg D<-\log 2 is proved using (1.5). The symmetry of the term degD+log2log3\lceil\frac{\deg D+\log 2}{\log 3}\rceil^{\prime} under the replacement of DD by the divisor KDK-D, with K:=2{2}K:=-2\{2\}, is the numerical evidence of a geometric duality holding on the curve Spec¯{\overline{{\rm Spec\,}{\mathbb{Z}}}} over 𝕊[±1]{{{\mathbb{S}}}[\pm 1]} (see section 5). This is precisely stated by the following (see Theorem 5.3)

Theorem 1.2.

Let DD be an Arakelov divisor on Spec¯{\overline{{\rm Spec\,}{\mathbb{Z}}}} and K=2{2}K=-2\{2\}, with degK=2log2\deg K=-2\log 2. Then there is an isomorphism of 𝕊[±1]{{{\mathbb{S}}}[\pm 1]}-modules.

H0(KD)Hom¯Γ𝒯(H1(D),U(1)14),H^{0}(K-D)\simeq{\underline{{\rm{Hom}}}}_{\Gamma{\mathcal{T}}_{*}}(H^{1}(D),U(1)_{\frac{1}{4}}), (1.6)

where U(1)14U(1)_{\frac{1}{4}} coincides with H1(K)H^{1}(K), playing the role of the dualizing module in Pontryagin duality.

The isomorphism H0(KD)Hom¯Γ𝒯(H1(D),H1(K))H^{0}(K-D)\simeq{\underline{{\rm{Hom}}}}_{\Gamma{\mathcal{T}}_{*}}(H^{1}(D),H^{1}(K)) 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 𝕊[±1]{{{\mathbb{S}}}[\pm 1]}-modules (Proposition 5.2), once again in analogy with the corresponding result proven in [13] for curves over fields of positive characteristic. The divisor KK 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 Γ\Gamma-space 𝐇(D){\bf H}(D) naturally associated to an Arakelov divisor DD. This gives, in Proposition 2.2, the cohomologies H0(D)H^{0}(D) and H1(D)H^{1}(D). We compute the dimension of H0(D)H^{0}(D) in section 3 (Theorem 3.4), and of H1(D)H^{1}(D) 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 𝕊{{\mathbb{S}}}-modules and in appendix B the technical details of the construction, through the Dold-Kan correspondence, of the Γ\Gamma-space 𝐇(D){\bf H}(D).

2 The cohomology H(D)H^{\bullet}(D)

We recall, from [4], the construction of the Γ\Gamma-space 𝐇(D){\bf H}(D) naturally associated to an Arakelov divisor DD on Spec¯{\overline{{\rm Spec\,}{\mathbb{Z}}}}. This space is the absolute homological incarnation of the Riemann-Roch problem for the divisor DD. We formulate this construction in terms of the adeles of {\mathbb{Q}}.

Let D=jaj{pj}+a{}D=\sum_{j}a_{j}\{p_{j}\}+a\{\infty\} be an Arakelov divisor on Spec¯{\overline{{\rm Spec\,}{\mathbb{Z}}}}. This is a formal finite sum with aja_{j}\in{\mathbb{Z}} and aa\in{\mathbb{R}}, where pjp_{j} are rational primes in {\mathbb{Z}} and the symbol \infty stands for the restriction ||:|\cdot|:{\mathbb{Q}}\stackrel{{\scriptstyle}}{{\to}}{\mathbb{R}} of the euclidean absolute value |||\cdot|_{\infty}. Let Σ\Sigma_{\mathbb{Q}} denote the full set of places ν\nu of {\mathbb{Q}}. To DD is naturally associated the following idele exp(D)\exp(D) of {\mathbb{Q}}

exp(D)ν:={pjajif ν=νj1ννj,jeaif ν=.\exp(D)_{\nu}:=\begin{cases}p_{j}^{-a_{j}}&\text{if $\nu=\nu_{j}$}\\ 1&\forall\nu\neq\nu_{j},\,\forall j\\ e^{a}&\text{if $\nu=\infty$}.\end{cases}

The equality exp(D)(ν)=|exp(D)ν|ν\exp(D)(\nu)=|\exp(D)_{\nu}|_{\nu} defines a map exp(D):Σ+\exp(D):\Sigma_{\mathbb{Q}}\to{\mathbb{R}}_{+}^{*} such that exp(D)(ν)mod(ν)\exp(D)(\nu)\in{\rm mod}({\mathbb{Q}}_{\nu}), νΣ\forall\nu\in\Sigma_{\mathbb{Q}}, and exp(D)(ν)=1\exp(D)(\nu)=1, for ν,νj,j\nu\neq\infty,\nu_{j},\forall j. To the divisor DD corresponds the compact subset of the adeles of {\mathbb{Q}}

𝒪(D):={(aν)𝔸|aν|νexp(D)(ν),νΣ}𝔸.{\mathcal{O}}(D):=\big{\{}(a_{\nu})\in{\mathbb{A}}_{\mathbb{Q}}\mid|a_{\nu}|_{\nu}\leq\exp(D)(\nu),\forall\nu\in\Sigma_{\mathbb{Q}}\big{\}}\subset{\mathbb{A}}_{\mathbb{Q}}.

By definition 𝒪(D)=𝒪(D)f×𝒪(D)=ν𝒪(D)ν{\mathcal{O}}(D)={\mathcal{O}}(D)_{f}\times{\mathcal{O}}(D)_{\infty}=\prod_{\nu}{\mathcal{O}}(D)_{\nu} with

𝒪(D)ν={pjajpjif ν=νjνννj,j,ν<[ea,ea]if ν=.{\mathcal{O}}(D)_{\nu}=\begin{cases}p_{j}^{-a_{j}}{\mathbb{Z}}_{p_{j}}&\text{if $\nu=\nu_{j}$}\\ {\mathbb{Z}}_{\nu}&\forall\nu\neq\nu_{j},\,\forall j,~\nu<\infty\\ [-e^{a},e^{a}]&\text{if $\nu=\infty$}.\end{cases}

In particular, the product 𝒪(D)f=ν𝒪(D)ν{\mathcal{O}}(D)_{f}=\prod_{\nu\neq\infty}{\mathcal{O}}(D)_{\nu} is a compact abelian group. The archimedean component, on the other hand, i.e. the real interval [ea,ea][-e^{a},e^{a}], is not, and for this reason we implement the functorial viewpoint to suitably understand 𝒪(D){\mathcal{O}}(D). Indeed, to 𝒪(D){\mathcal{O}}(D)_{\infty} we associate the covariant functor

Hea:Γo𝔖𝔢𝔱𝔰Hea(F):={ϕH(F)F{}|ϕ(x)|ea}\|H{\mathbb{R}}\|_{e^{a}}:\Gamma^{o}\longrightarrow\mathfrak{Sets}_{*}\qquad\|H{\mathbb{R}}\|_{e^{a}}(F):=\bigg{\{}\phi\in H{\mathbb{R}}(F)\mid\sum_{F\setminus\{*\}}|\phi(x)|\leq e^{a}\bigg{\}}

from the opposite of the Segal category (see e.g. [6] Chpt. 2 and [2]) to pointed sets, which maps a finite pointed set FF to the pointed set of {\mathbb{R}}-valued divisors on FF vanishing on the base point and whose total mass F{}|ϕ(x)|\sum_{F\setminus\{*\}}|\phi(x)| is bounded by eae^{a}. Covariant functors Γo𝔖𝔢𝔱𝔰\Gamma^{o}\longrightarrow\mathfrak{Sets}_{*} and their natural transformations determine the category Γ𝔖𝔢𝔱𝔰\Gamma{\mathfrak{Sets}_{*}} of Γ\Gamma-sets (aka 𝕊{{\mathbb{S}}}-modules). In particular, the Eilenberg-MacLane functor HH encodes an abelian group AA as the covariant functor HA:Γo𝔖𝔢𝔱𝔰HA:\Gamma^{o}\longrightarrow\mathfrak{Sets}_{*} that associates to FF the pointed set of AA-valued divisors on FF vanishing on the base point. Functoriality holds by taking sums on the inverse images of a point. The functor HAHA encodes the addition on A=HA(1+)A=HA(1_{+}). In general, let σHomΓo(k+,1+)\sigma\in{\rm{Hom}}_{\Gamma^{o}}(k_{+},1_{+}) with σ()=1\sigma(\ell)=1 \forall\ell\neq* and

δ(j,k)HomΓo(k+,1+),δ(j,k)():={1if =jif j.\delta(j,k)\in{\rm{Hom}}_{\Gamma^{o}}(k_{+},1_{+}),\quad\delta(j,k)(\ell):=\begin{cases}1&\text{if $\ell=j$}\\ *&\text{if $\ell\neq j$}.\end{cases} (2.1)

Given an 𝕊{{\mathbb{S}}}-module {\mathcal{F}} and elements x,xj(1+)x,x_{j}\in{\mathcal{F}}(1_{+}), j=1,,kj=1,\ldots,k, one writes

x=jxjz(k+)s.t.(σ)(z)=x,(δ(j,k))(z)=xj,j.x=\sum_{j}x_{j}\iff\exists z\in{\mathcal{F}}(k_{+})~\text{s.t.}~{\mathcal{F}}(\sigma)(z)=x,~{\mathcal{F}}(\delta(j,k))(z)=x_{j},\ \forall j. (2.2)

For A=A={\mathbb{R}} the key point is that this general construction respects the bound on the total mass while encoding the addition. One easily sees that Hea\|H{\mathbb{R}}\|_{e^{a}} is an 𝕊[±1]{{{\mathbb{S}}}[\pm 1]}-module (an “(𝕊[±1]){{{\mathbb{S}}}[\pm 1]}\wedge-)-algebra” as in [6] 2.1.5), where 𝕊[±1]:Γo𝔖𝔢𝔱𝔰{{{\mathbb{S}}}[\pm 1]}:\Gamma^{o}\longrightarrow\mathfrak{Sets}_{*}, 𝕊[±1](F):={1,0,1}F{{{\mathbb{S}}}[\pm 1]}(F):=\{-1,0,1\}\wedge F is the spherical monoid algebra of the multiplicative monoid {±1}\{\pm 1\}. In particular, the inclusion functor HeaH\|H{\mathbb{R}}\|_{e^{a}}\longrightarrow H{\mathbb{R}} determines Hea\|H{\mathbb{R}}\|_{e^{a}} as a sub 𝕊[±1]{{{\mathbb{S}}}[\pm 1]}-module of HH{\mathbb{R}}, where 𝕊:Γo𝔖𝔢𝔱𝔰{{\mathbb{S}}}:\Gamma^{o}\longrightarrow\mathfrak{Sets}_{*} is the identity functor.

Given two 𝕊{{\mathbb{S}}}-modules j{\mathcal{F}}_{j}, j=1,2j=1,2, one defines their product as the functor

1×2:Γo𝔖𝔢𝔱𝔰,(1×2)(F)=1(F)×2(F){\mathcal{F}}_{1}\times{\mathcal{F}}_{2}:\Gamma^{o}\longrightarrow\mathfrak{Sets}_{*},\qquad({\mathcal{F}}_{1}\times{\mathcal{F}}_{2})(F)={\mathcal{F}}_{1}(F)\times{\mathcal{F}}_{2}(F)

with the base point of 1(F)×2(F){\mathcal{F}}_{1}(F)\times{\mathcal{F}}_{2}(F) taken to be (,)(*,*). For abelian groups A,BA,B one has a canonical isomorphism H(A×B)HA×HBH(A\times B)\simeq HA\times HB. In particular, the morphism of addition in adeles α:×𝔸𝔸\alpha:{\mathbb{Q}}\times{\mathbb{A}}_{\mathbb{Q}}\to{\mathbb{A}}_{\mathbb{Q}}, α(q,a)=q+a\alpha(q,a)=q+a (using the diagonal embedding of {\mathbb{Q}} in 𝔸{\mathbb{A}}_{\mathbb{Q}}) determines a morphism of 𝕊[±1]{{{\mathbb{S}}}[\pm 1]}-modules Hα:H×H𝔸H𝔸H\alpha:H{\mathbb{Q}}\times H{\mathbb{A}}_{\mathbb{Q}}\longrightarrow H{\mathbb{A}}_{\mathbb{Q}}. Next proposition shows how this functorial construction is well combined with the adelic formalism.

Proposition 2.1.
  1. (i)

    The 𝕊[±1]{{{\mathbb{S}}}[\pm 1]}-module

    H𝒪(D):=H(𝒪(D)f)×HeaH{\mathcal{O}}(D):=H({\mathcal{O}}(D)_{f})\times\|H{\mathbb{R}}\|_{e^{a}}

    determines, canonically, a sub 𝕊[±1]{{{\mathbb{S}}}[\pm 1]}-module ι:H𝒪(D)H𝔸\iota:H{\mathcal{O}}(D)\longrightarrow H{\mathbb{A}}_{\mathbb{Q}}.

  2. (ii)

    The restriction of HαH\alpha to H×H𝒪(D)H{\mathbb{Q}}\times H{\mathcal{O}}(D) defines a morphism of 𝕊[±1]{{{\mathbb{S}}}[\pm 1]}-modules

    ψ:H×H𝒪(D)H𝔸.\psi:H{\mathbb{Q}}\times H{\mathcal{O}}(D)\longrightarrow H{\mathbb{A}}_{\mathbb{Q}}. (2.3)
Proof 1.

(i)(i) With 𝔸=𝔸f×{\mathbb{A}}_{\mathbb{Q}}={\mathbb{A}}_{f}\times{\mathbb{R}}, let ιf:𝒪(D)f𝔸f\iota_{f}:{\mathcal{O}}(D)_{f}\to{\mathbb{A}}_{f} be the inclusion of abelian groups. Then ι:H𝒪(D)H𝔸\iota:H{\mathcal{O}}(D)\to H{\mathbb{A}}_{\mathbb{Q}} is the product of HιfH\iota_{f} and the inclusion HeaH\|H{\mathbb{R}}\|_{e^{a}}\to H{\mathbb{R}}.
(ii)(ii) By composition one obtains ψ=Hα(Hid×ι)\psi=H\alpha\circ(H{\rm id}_{\mathbb{Q}}\times\iota).

The morphism ψ\psi in (2.3) is obtained by restricting HαH\alpha to the (adelic) divisor DD, where α\alpha is a morphism of abelian groups. The theory developed in [4] associates to ψ\psi a Γ\Gamma-space 𝐇(D){\bf H}(D), by implementing the Dold-Kan correspondence in the special case of a short complex of abelian groups and its restriction to a sub-𝕊{{\mathbb{S}}}-module. The Γ\Gamma-space 𝐇(D){\bf H}(D) supplies the absolute cohomology of the divisor DD. We recall this construction in appendix B. Moreover, we review how the generalized homotopy π𝒯(𝐇(D))\pi^{{\mathcal{T}}}_{\ast}({\bf H}(D)), where 𝒯{\mathcal{T}} is the category of tolerance relations defined in A.1, is meaningful in this context supplying, with π1𝒯(𝐇(D))\pi^{{\mathcal{T}}}_{1}({\bf H}(D)) and π0𝒯(𝐇(D))\pi^{{\mathcal{T}}}_{0}({\bf H}(D)), the (tolerant) 𝕊{{\mathbb{S}}}-modules descriptions of resp. the kernel and the cokernel of ψ\psi. In other words, one obtains the following result

Proposition 2.2.
  1. (i)

    The kernel of ψ\psi is the 𝕊[±1]{{{\mathbb{S}}}[\pm 1]}-module:

    H0(D)=(Hid×ι)1(Hker(α))ι1H=H0(Spec,𝒪(D)f)ea.H^{0}(D)=(H{\rm id}_{\mathbb{Q}}\times\iota)^{-1}(H\ker(\alpha))\simeq\iota^{-1}H{\mathbb{Q}}=\|H^{0}({{\rm Spec\,}{\mathbb{Z}}},{\mathcal{O}}(D)_{f})\|_{e^{a}}. (2.4)
  2. (ii)

    The cokernel of ψ\psi is the 𝕊[±1]{{{\mathbb{S}}}[\pm 1]}-module H1(D)=(H𝔸,)H^{1}(D)=(H{\mathbb{A}}_{\mathbb{Q}},{\mathcal{R}}) endowed with the relations

    k:={(x,y)H𝔸(k+)×H𝔸(k+)xyImageψ(k+)}.{\mathcal{R}}_{k}:=\{(x,y)\in H{\mathbb{A}}_{\mathbb{Q}}(k_{+})\times H{\mathbb{A}}_{\mathbb{Q}}(k_{+})\mid x-y\in{\rm Image}\,\psi(k_{+})\}. (2.5)
  3. (iii)

    Both H0(D)H^{0}(D) and H1(D)H^{1}(D) only depend, up to isomorphism, on the linear equivalence class of the divisor DD.

Proof 2.

We refer to appendix B.1.

3 The dimension of H0(D)H^{0}(D)

In this section we consider, for each integer n>0n>0, the 𝕊[±1]{{{\mathbb{S}}}[\pm 1]}-module Hn\|H{\mathbb{Z}}\|_{n}

Hn:Γo𝔖𝔢𝔱𝔰Hn(F):={ϕH(F)F{}|ϕ(x)|n}\|H{\mathbb{Z}}\|_{n}:\Gamma^{o}\longrightarrow\mathfrak{Sets}_{*}\qquad\|H{\mathbb{Z}}\|_{n}(F):=\{\phi\in H{\mathbb{Z}}(F)\mid\sum_{F\setminus\{*\}}|\phi(x)|\leq n\}

and evaluate its dimension. It follows from the 𝕊[±1]{{{\mathbb{S}}}[\pm 1]}-module structure, that (2.2) holds for a sum of the form

Fαjj[n,n]=Hn(1+)αj{1,0,1}s.t.F|αjj|n.\sum_{F}\alpha_{j}j\in[-n,n]\cap{\mathbb{Z}}=\|H{\mathbb{Z}}\|_{n}(1_{+})\quad\forall\alpha_{j}\in\{-1,0,1\}~~\text{s.t.}~\sum_{F}|\alpha_{j}j|\leq n.

By applying Definition A.2, a subset F[n,n]=Hn(1+)F\subset[-n,n]\cap{\mathbb{Z}}=\|H{\mathbb{Z}}\|_{n}(1_{+}) generates333For simplicity we use from now on the term “generates” for “linearly generates” Hn(1+)\|H{\mathbb{Z}}\|_{n}(1_{+}) if and only if for every element xHn(1+)x\in\|H{\mathbb{Z}}\|_{n}(1_{+}) there exists coefficients αj{1,0,1}\alpha_{j}\in\{-1,0,1\}, jFj\in F, such that x=Fαjjx=\sum_{F}\alpha_{j}j and F|αjj|n\sum_{F}|\alpha_{j}j|\leq n. The dimension dim𝕊[±1]Hn\dim_{{{{\mathbb{S}}}[\pm 1]}}\|H{\mathbb{Z}}\|_{n} is the minimal cardinality of a generating set.

The explicit computation of dim𝕊[±1]Hn\dim_{{{{\mathbb{S}}}[\pm 1]}}\|H{\mathbb{Z}}\|_{n} follows from the next two statements.

Lemma 3.1.

For kk\in{\mathbb{N}}, let n=(3k1)/2n=(3^{k}-1)/2. Consider the subset F:={3i,0ik1}[n,n]F:=\{3^{i},0\leq i\leq k-1\}\subset[-n,n]. Then FF is a generating set and the map

θ:{1,0,1}×k[n,n],θ((αi)):=i=0k1αi3i\theta:\{-1,0,1\}^{\times k}\to[-n,n]\cap{\mathbb{Z}},\qquad\theta((\alpha_{i})):=\sum_{i=0}^{k-1}\alpha_{i}3^{i} (3.1)

is bijective. Moreover i=0k1|αi3i|n\sum_{i=0}^{k-1}|\alpha_{i}3^{i}|\leq n, for all αi{1,0,1}×k\alpha_{i}\in\{-1,0,1\}^{\times k}.

Proof 3.

First of all, note that for all α{1,0,1}×k\alpha\in\{-1,0,1\}^{\times k} one has

i=0k1|αi3i|i=0k13i=n.\sum_{i=0}^{k-1}|\alpha_{i}3^{i}|\leq\sum_{i=0}^{k-1}3^{i}=n.

This shows that for any α{1,0,1}×k\alpha\in\{-1,0,1\}^{\times k} one has z=(αi3i)Hn(k+)z=(\alpha_{i}3^{i})\in\|H{\mathbb{Z}}\|_{n}(k_{+}), and, as in (2.2), Hn(σ)(z)=θ((αi))\|H{\mathbb{Z}}\|_{n}(\sigma)(z)=\theta((\alpha_{i})). We prove the bijectivity of θ\theta by induction on kk. It is clear for k=1k=1. For k=2k=2, n=4n=4 and one writes the set [4,4][-4,4]\cap{\mathbb{Z}} as the disjoint union of three subsets as follows

[4,4]\displaystyle[-4,4]\cap{\mathbb{Z}} ={4,3,2}{1,0,1}{2,3,4}=\displaystyle=\{-4,-3,-2\}\cup\{-1,0,1\}\cup\{2,3,4\}=
=(3+{1,0,1}){1,0,1}(3+{1,0,1}).\displaystyle=(-3+\{-1,0,1\})\cup\{-1,0,1\}\cup(3+\{-1,0,1\}).

This shows that θ\theta is bijective. Next, by assuming to have shown the bijectivity up to k1k-1, we deduce it for kk. Indeed, with n=(3k1)/2n=(3^{k}-1)/2 let m=(3k11)/2m=(3^{k-1}-1)/2. The induction hypothesis ensures that every element of the set [m,m][-m,m]\cap{\mathbb{Z}} can be uniquely written as a sum i=0k2αi3i\sum_{i=0}^{k-2}\alpha_{i}3^{i}. We divide the set [n,n][-n,n]\cap{\mathbb{Z}} into three disjoint parts of the form

J1:=[n,n+2m],J0:=[m,m],J1:=[n2m,n].J_{-1}:=[-n,-n+2m]\cap{\mathbb{Z}},\quad J_{0}:=[-m,m]\cap{\mathbb{Z}},\quad J_{1}:=[n-2m,n]\cap{\mathbb{Z}}.

One has n2m=m+1n-2m=m+1 and thus the union of these three disjoint sets is [n,n][-n,n]\cap{\mathbb{Z}}. Moreover one has J1=(nm)+J0J_{1}=(n-m)+J_{0} with nm=3k1n-m=3^{k-1}, and similarly J1=J03k1J_{-1}=J_{0}-3^{k-1}. Thus, given x[n,n]x\in[-n,n]\cap{\mathbb{Z}} one determines uniquely the coefficients αi{1,0,1}\alpha_{i}\in\{-1,0,1\} such that θ((αi))=x\theta((\alpha_{i}))=x. Indeed, one uses the above partition in three intervals to determine the coefficient αk1\alpha_{k-1} and then one applies the induction hypothesis to determine the others.

Remark 3.2.
  1. 1.

    The conceptual explanation of Lemma 3.1 derives from the following peculiar property of the Teichmüller lift τ:𝔽33\tau:{\mathbb{F}}_{3}\to{\mathbb{Z}}_{3}. One has τ(𝔽3)={1,0,1}3\tau({\mathbb{F}}_{3})=\{-1,0,1\}\subset{\mathbb{Z}}\subset{\mathbb{Z}}_{3} and τ\tau extends to a canonical map τ~:Wk(𝔽3)\tilde{\tau}:W_{k}({\mathbb{F}}_{3})\to{\mathbb{Z}} from Witt vectors to {\mathbb{Z}}

    ξ=(ξj)Wk(𝔽3)τ~(ξ):=0k1τ(ξj)3j.\xi=(\xi_{j})\in W_{k}({\mathbb{F}}_{3})\mapsto\tilde{\tau}(\xi):=\sum_{0}^{k-1}\tau(\xi_{j})3^{j}\in{\mathbb{Z}}.

    This map coincides with θ\theta in (3.1) and one can compute the partially defined sums (2.2) in Hn\|H{\mathbb{Z}}\|_{n} using the addition in Wk(𝔽3)W_{k}({\mathbb{F}}_{3}). The prime p=3p=3 is the only prime such that the subring p=W(𝔽p){\mathbb{Z}}\subset{\mathbb{Z}}_{p}=W({\mathbb{F}}_{p}) coincides with the ring of Witt vectors with finitely many non-zero components. When p>3p>3 one has τ(𝔽p)\tau({\mathbb{F}}_{p})\not\subset{\mathbb{Z}}, while for p=2p=2 the integer 12-1\in{\mathbb{Z}}\subset{\mathbb{Z}}_{2} is the Witt vector whose components are all equal to 11. 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. 2.

    The Teichmüller lift τ\tau, as a morphism of multiplicative pointed monoids, induces a morphism 𝕊[±1]H{{{\mathbb{S}}}[\pm 1]}\to H{\mathbb{Z}} of 𝕊{{\mathbb{S}}}-algebras.

  3. 3.

    The ordering of the natural numbers encoded by θ\theta coincides with the lexicographic ordering of the coefficients (αi){1,0,1}×k(\alpha_{i})\in\{-1,0,1\}^{\times k}

    θ((αi))<θ((βi))js.t.αj<βjandα=β>j.\theta((\alpha_{i}))<\theta((\beta_{i}))\iff\exists j~\text{s.t.}~\alpha_{j}<\beta_{j}~~\text{and}~~\alpha_{\ell}=\beta_{\ell}\quad\forall\ell>j.

We recall that the ceiling function x\lceil x\rceil associates to a positive real number xx the smallest integer nxn\geq x.

Proposition 3.3.
  1. (i)

    For any integer n0n\geq 0 one has

    dim𝕊[±1]Hn=log(2n+1)log3.\dim_{{{{\mathbb{S}}}[\pm 1]}}\|H{\mathbb{Z}}\|_{n}=\bigg{\lceil}\frac{\log(2n+1)}{\log 3}\bigg{\rceil}. (3.2)
  2. (ii)

    For n{2,5}n\notin\{2,5\}, there exists a generating subset F{1,,n}F\subset\{1,\ldots,n\} of cardinality #F=log(2n+1)log3\#F=\lceil\frac{\log(2n+1)}{\log 3}\rceil, with jFj=n\sum_{j\in F}j=n, such that the following map surjects onto [n,n][-n,n]\cap{\mathbb{Z}}

    θ:{1,0,1}×(#F)[n,n],θ((αi)):=jFαjj.\theta:\{-1,0,1\}^{\times(\#F)}\to[-n,n]\cap{\mathbb{Z}},\qquad\theta((\alpha_{i})):=\sum_{j\in F}\alpha_{j}j.
Proof 4.

(i)(i) Set κ(n):=dim𝕊[±1]Hn\kappa(n):=\dim_{{{{\mathbb{S}}}[\pm 1]}}\|H{\mathbb{Z}}\|_{n}. Since the cardinality of [n,n][-n,n]\cap{\mathbb{Z}} is 2n+12n+1 and the cardinality of {1,0,1}×#F\{-1,0,1\}^{\times\#F} is 3#F3^{\#F}, it follows that if FF is a generating set one must have 3#F2n+13^{\#F}\geq 2n+1. Thus one has #Flog(2n+1)log3\#F\geq\frac{\log(2n+1)}{\log 3}, and since #F\#F is an integer one gets #Flog(2n+1)log3\#F\geq\lceil\frac{\log(2n+1)}{\log 3}\rceil. This shows that κ(n)log(2n+1)log3\kappa(n)\geq\lceil\frac{\log(2n+1)}{\log 3}\rceil. It remains to prove that for each n0n\geq 0 one can find a generating set FF with #F=log(2n+1)log3\#F=\lceil\frac{\log(2n+1)}{\log 3}\rceil. One uses Lemma 3.1, and distinguishes several cases starting with the easiest one. Set F(k):={3i,0ik1}F(k):=\{3^{i},0\leq i\leq k-1\} and let EE\subset{\mathbb{N}} be the subset

E={3+12(3m1)m,<m}.E=\bigg{\{}3^{\ell}+\frac{1}{2}\left(3^{m}-1\right)\mid m\in{\mathbb{N}},\ell<m\bigg{\}}. (3.3)

We list the first elements of EE as follows

2,5,7,14,16,22,41,43,49,67,122,124,130,148,202,365,367,373,391,445,6072,5,7,14,16,22,41,43,49,67,122,124,130,148,202,365,367,373,391,445,607\ldots

First, assume that nEn\notin E. Then let mm be the largest integer such that 3m2n+13^{m}\leq 2n+1. If 3m=2n+13^{m}=2n+1, then Lemma 3.1 provides one with a generating set of cardinality m=log(2n+1)log3m=\frac{\log(2n+1)}{\log 3}. Assume now that 3m<2n+13^{m}<2n+1. Then, the map θ\theta for F(m)F(m) covers the set [q,q][-q,q]\cap{\mathbb{Z}}, where q=12(3m1)q=\frac{1}{2}\left(3^{m}-1\right). Thus by adjoining to F(m)F(m) the element nqn-q one covers [n,n][-n,n]\cap{\mathbb{Z}}. More precisely, let F=F(m){n12(3m1)}F=F(m)\cup\{n-\frac{1}{2}\left(3^{m}-1\right)\}, then we show that FF fulfills the conditions in (ii)(ii). One has #F=#F(m)+1\#F=\#F(m)+1 since the additional element n12(3m1)n-\frac{1}{2}\left(3^{m}-1\right) does not belong to F(m)F(m) as, by hypothesis, nEn\notin E. Since #F(m)=m\#F(m)=m, #F=m+1=log(2n+1)log3\#F=m+1=\lceil\frac{\log(2n+1)}{\log 3}\rceil. The sum of elements in F(m)F(m) is 12(3m1)\frac{1}{2}\left(3^{m}-1\right), so the sum of elements in FF is equal to nn. Next, we show that θ\theta is surjective. Its image contains the three sets θ(F(m))\theta(F(m)), θ(F(m))+n12(3m1)\theta(F(m))+n-\frac{1}{2}\left(3^{m}-1\right), and θ(F(m))(n12(3m1))\theta(F(m))-(n-\frac{1}{2}\left(3^{m}-1\right)). Using Lemma 3.1 one obtains, with q=12(3m1)q=\frac{1}{2}\left(3^{m}-1\right)

θF(m)=[q,q],θF(m)+nq=[n2q,n],\displaystyle\theta F(m)=[-q,q]\cap{\mathbb{Z}},\quad\theta F(m)+n-q=[n-2q,n]\cap{\mathbb{Z}},
θF(m)(nq)=[n,n+2q].\displaystyle\theta F(m)-(n-q)=[-n,-n+2q]\cap{\mathbb{Z}}.

One has 3(2q+1)=3m+1>2n+13(2q+1)=3^{m+1}>2n+1. This inequality prevents the three subsets from being disjoint, thus the upper limit qq of θF(m)\theta F(m) is greater or equal to the lower limit n2qn-2q of the interval θF(m)+nq\theta F(m)+n-q. Thus θ\theta is surjective.
The next case is for nEn\in E of the form n=1+12(3m1)n=1+\frac{1}{2}\left(3^{m}-1\right), for some mm (=0\ell=0 in (3.3)). We assume that m>2m>2 (this excludes the cases n=2n=2 for m=1m=1, and n=5n=5 for m=2m=2). Let

F:={3k0km2}{2,3m11}.F:=\{3^{k}\mid 0\leq k\leq m-2\}\cup\{2,3^{m-1}-1\}.

(This choice avoids the repetition of the element 1=nq1=n-q while keeping the sum of elements of FF equal to nn). We show that FF fulfills the conditions in (ii)(ii). We have #F=m1+2=m+1\#F=m-1+2=m+1 since 3m11{1,2}3^{m-1}-1\notin\{1,2\}. Moreover 2n+1=3m+22n+1=3^{m}+2, thus #F=m+1=log(2n+1)log3\#F=m+1=\lceil\frac{\log(2n+1)}{\log 3}\rceil. The sum of terms in FF is equal to one plus the sum of the powers of 33 up to 3m13^{m-1} and thus is equal to nn. Let us now show that the map θ\theta is surjective. By construction θ(F)\theta(F) is the union of nine intervals obtained, with q=12(3m11)q^{\prime}=\frac{1}{2}(3^{m-1}-1), by adding to the set [q,q][-q^{\prime},q^{\prime}]\cap{\mathbb{Z}} the terms in {2,3m11}\{2,3^{m-1}-1\} with coefficients in {1,0,1}\{-1,0,1\}. With the term 3m113^{m-1}-1 one covers the set [q+1,q1][-q+1,q-1]\cap{\mathbb{Z}}, where q=12(3m1)q=\frac{1}{2}\left(3^{m}-1\right) as the union of the three sets

[q,q],([q,q]+3m11),([q,q](3m11)).[-q^{\prime},q^{\prime}]\cap{\mathbb{Z}},\ \ ([-q^{\prime},q^{\prime}]+3^{m-1}-1)\cap{\mathbb{Z}},\ \ ([-q^{\prime},q^{\prime}]-(3^{m-1}-1))\cap{\mathbb{Z}}.

One has q+3m11=n2q^{\prime}+3^{m-1}-1=n-2, and thus it remains to cover the two elements n1,nn-1,n and their opposite. This is done using the set ([q,q]+3m11+2)([-q^{\prime},q^{\prime}]+3^{m-1}-1+2)\cap{\mathbb{Z}} whose upper limit is nn, and appealing to the fact that since q>1q^{\prime}>1 this interval contains n1n-1.
Finally, we consider the case nEn\in E of the form n=3+12(3m1)n=3^{\ell}+\frac{1}{2}\left(3^{m}-1\right), with m,0<<mm\in{\mathbb{N}},~0<\ell<m. Here, the additional element n12(3m1)n-\frac{1}{2}\left(3^{m}-1\right) does belong to F(m)F(m) so the proof of the first case considered above does not apply since 33^{\ell} appears twice in F(m){n12(3m1)}F(m)\cup\{n-\frac{1}{2}\left(3^{m}-1\right)\}. We replace this double occurrence of 33^{\ell} by the two distinct elements 3±13^{\ell}\pm 1. This gives

F=(F(m){3}){31,3+1}.F=(F(m)\setminus\{3^{\ell}\})\cup\{3^{\ell}-1,3^{\ell}+1\}.

By construction #F(m)+1=m+1\#F(m)+1=m+1. Furthermore one has: n=3+12(3m1)n=3^{\ell}+\frac{1}{2}\left(3^{m}-1\right), 2n+1=3m+2×32n+1=3^{m}+2\times 3^{\ell} and since 0<<m0<\ell<m one gets #F=m+1=log(2n+1)log3\#F=m+1=\lceil\frac{\log(2n+1)}{\log 3}\rceil. The sum of elements of FF is the same as the sum of elements of F(m)F(m) plus 33^{\ell} which is equal to nn. It remains to show that, for such FF, θ\theta is surjective. We first deal with the set G():=F(){31,3+1}G(\ell):=F(\ell)\cup\{3^{\ell}-1,3^{\ell}+1\} and show that the map θG()\theta_{G(\ell)}, for this set, surjects onto the interval [t(),t()][-t(\ell),t(\ell)]\cap{\mathbb{Z}}, where t()=G()j=3+12(3+11)t(\ell)=\sum_{G(\ell)}j=3^{\ell}+\frac{1}{2}(3^{\ell+1}-1). Lemma 3.1 shows that the image of θF()\theta_{F(\ell)} is the set JJ\cap{\mathbb{Z}}, with J=[12(31),12(31)]J=[-\frac{1}{2}(3^{\ell}-1),\frac{1}{2}(3^{\ell}-1)]. To show the surjectivity of θG()\theta_{G(\ell)} it is enough to show that the following intervals cover [0,t()][0,t(\ell)]:

J,J+31,J+3+1,J+2×3.J,\quad J+3^{\ell}-1,\quad J+3^{\ell}+1,\quad J+2\times 3^{\ell}.

The upper limit of J+2×3J+2\times 3^{\ell} is 12(31)+2×3=t()\frac{1}{2}(3^{\ell}-1)+2\times 3^{\ell}=t(\ell), its lower limit is equal to the upper limit of J+3+1J+3^{\ell}+1. The lower limit of J+31J+3^{\ell}-1 is 3112(31)=12(31)3^{\ell}-1-\frac{1}{2}(3^{\ell}-1)=\frac{1}{2}(3^{\ell}-1) which is the upper limit of JJ. Moreover, the length of JJ is 3123^{\ell}-1\geq 2 so that two translates of JJ of the form J+aJ+a, J+a+2J+a+2 necessarily overlap. This shows that the map θG()\theta_{G(\ell)} surjects onto the interval [t(),t()][-t(\ell),t(\ell)]\cap{\mathbb{Z}}. Thus θG()\theta_{G(\ell)} is the symmetric enlargement of the interval θ(F(+1))\theta(F(\ell+1)) obtained by adding 33^{\ell} to the upper limit. Then, as in the proof of Lemma 3.1, one obtains by induction that when one adjoins the higher powers 33^{\ell^{\prime}}, where <<m\ell<\ell^{\prime}<m, one achieves the enlargement of the interval θ(F(m))\theta(F(m)) obtained by adding 33^{\ell} to the upper limit. This shows that the map θ\theta for FF is surjective. In fact the above arguments prove (ii)(ii) and (i)(i), except that one has to take care of the special cases n=2,5n=2,5. For n=2n=2, {1,2}\{1,2\} is a generating set but the sum of its terms is >2>2. For n=5n=5, {1,2,3}\{1,2,3\} is a generating set but the sum of its terms is >5>5.

We are now ready to determine the dimension of the 𝕊[±1]{{{\mathbb{S}}}[\pm 1]}-module H0(D)H^{0}(D), for any Arakelov divisor DD. It is a non-decreasing function of degD\deg D but the formula for this dimension drops by 11 on an exceptional set LL. It is the open set defined as the union of open intervals as follows

L+log2=k(klog3,klog3+ϵk),ϵk:=log(1+3k).L+\log 2=\bigcup_{k\in{\mathbb{N}}}(k\log 3,k\log 3+\epsilon_{k}),\ \ \epsilon_{k}:=\log(1+3^{-k}). (3.4)

It has finite Lebesgue measure

|L|=k0ϵk=log(1;13)=1.14099|L|=\sum_{k\geq 0}\epsilon_{k}=\log\left(-1;\frac{1}{3}\right)_{\infty}=1.14099\ldots

where (1;13)=n=0(1+3n)\left(-1;\frac{1}{3}\right)_{\infty}=\prod_{n=0}^{\infty}(1+3^{-n}) is the 1-1 Pochhammer symbol at 13\frac{1}{3}.

Theorem 3.4.

Let DD be an Arakelov divisor on Spec¯{\overline{{\rm Spec\,}{\mathbb{Z}}}}. If degDlog2\deg D\geq-\log 2, then

dim𝕊[±1]H0(D)=degD+log2log3𝟏L\dim_{{{{\mathbb{S}}}[\pm 1]}}H^{0}(D)=\bigg{\lceil}\frac{\deg D+\log 2}{\log 3}\bigg{\rceil}-{\mathbf{1}}_{L} (3.5)

where 𝟏L{\mathbf{1}}_{L} is the characteristic function of the open set LL.

Proof 5.

Let a=degDa=\deg D and n=exp(deg(D)n=\lfloor\exp(\deg(D)\rfloor, then H0(D)=HnH^{0}(D)=\|H{\mathbb{Z}}\|_{n}. By Proposition 3.3 one gets

dim𝕊[±1]H0(D)=log(2n+1)log3,n:=exp(a).\dim_{{{{\mathbb{S}}}[\pm 1]}}H^{0}(D)=\bigg{\lceil}\frac{\log(2n+1)}{\log 3}\bigg{\rceil},\qquad n:=\lfloor\exp(a)\rfloor.

Let us compare this formula with (3.5). Assume first that aLa\in L, thus there exists kk\in{\mathbb{N}} such that a+log2(klog3,klog3+ϵk)a+\log 2\in(k\log 3,k\log 3+\epsilon_{k}). Then, 2expa(3k,3k+1)2\exp a\in(3^{k},3^{k}+1) and with n=expan=\lfloor\exp a\rfloor, one has n=12(3k1)n=\frac{1}{2}(3^{k}-1). Thus 2n+1=3k2n+1=3^{k} and log(2n+1)log3=k\lceil\frac{\log(2n+1)}{\log 3}\rceil=k. Moreover, a+log2log3(k,k+ϵklog3)\frac{a+\log 2}{\log 3}\in(k,k+\frac{\epsilon_{k}}{\log 3}), where 0<ϵklog3<10<\frac{\epsilon_{k}}{\log 3}<1, hence one has a+log2log3=k+1\lceil\frac{a+\log 2}{\log 3}\rceil=k+1, so that (3.5) holds since 𝟏L(a)=1{\mathbf{1}}_{L}(a)=1. Now, assume that aLa\notin L, equivalently that a+log2log3[k+ϵklog3,k+1]\frac{a+\log 2}{\log 3}\in[k+\frac{\epsilon_{k}}{\log 3},k+1] for some integer k0k\geq 0 (since alog2a\geq-\log 2 by hypothesis). Then a+log2log3=k+1\lceil\frac{a+\log 2}{\log 3}\rceil=k+1. One has

2expa[exp(klog3)(1+3k),exp((k+1)log3)]=[3k+1,3k+1].2\exp a\in[\exp(k\log 3)(1+3^{-k}),\exp((k+1)\log 3)]=[3^{k}+1,3^{k+1}].

In this case n=expa[12(3k+1),12(3k+11)]n=\lfloor\exp a\rfloor\in[\frac{1}{2}(3^{k}+1),\frac{1}{2}(3^{k+1}-1)] and thus 2n+1[3k+2,3k+1]2n+1\in[3^{k}+2,3^{k+1}]. This implies that log(2n+1)log3(k,k+1]\frac{\log(2n+1)}{\log 3}\in(k,k+1] and hence that log(2n+1)log3=k+1\lceil\frac{\log(2n+1)}{\log 3}\rceil=k+1. In this case (3.5) holds since 𝟏L(a)=0{\mathbf{1}}_{L}(a)=0.

4 The dimension of H1(D)H^{1}(D)

Let (A,d)(A,d) be an abelian group endowed with a translation invariant metric dd. For λ>0\lambda\in{\mathbb{R}}_{>0}, we shall refer to (A,d)λ(A,d)_{\lambda} as the associated object of the category Γ𝒯\Gamma{\mathcal{T}}_{*} as in Proposition A.3. The 𝕊{{\mathbb{S}}}-module structure determines the addition in A=(A,d)λ(1+)A=(A,d)_{\lambda}(1_{+}), as well as the action of 𝕊[±1]{{{\mathbb{S}}}[\pm 1]} on it, where for xA=(A,d)λ(1+)x\in A=(A,d)_{\lambda}(1_{+}), x-x is the additive inverse of xx. The metric dd determines the tolerance relation {\mathcal{R}} on A=(A,d)λ(1+)A=(A,d)_{\lambda}(1_{+}). Next result computes the dimension (Definition A.2) of the tolerant 𝕊[±1]{{{\mathbb{S}}}[\pm 1]}-module (/,d)λ({\mathbb{R}}/{\mathbb{Z}},d)_{\lambda}.

Proposition 4.1.

Let λ>0\lambda\in{\mathbb{R}}_{>0}, and U(1)λ=(U(1),d)λU(1)_{\lambda}=(U(1),d)_{\lambda} where U(1)U(1) is the abelian group /{\mathbb{R}}/{\mathbb{Z}} endowed with the canonical metric dd of length 11. Then

dim𝕊[±1]U(1)λ={logλlog2log3if λ<120if λ12.\dim_{{{{\mathbb{S}}}[\pm 1]}}U(1)_{\lambda}=\begin{cases}\bigg{\lceil}\frac{-\log\lambda-\log 2}{\log 3}\bigg{\rceil}&\text{if $\lambda<\frac{1}{2}$}\\ 0&\text{if $\lambda\geq\frac{1}{2}$.}\end{cases} (4.1)
Proof 6.

For λ12\lambda\geq\frac{1}{2}, any element of U(1)λ=(/,d)λU(1)_{\lambda}=({\mathbb{R}}/{\mathbb{Z}},d)_{\lambda} is at distance λ\leq\lambda from 0, thus one can take F=F=\emptyset as generating set since, by convention, =0\sum_{\emptyset}=0. Thus dim𝕊[±1]U(1)λ=0\dim_{{{{\mathbb{S}}}[\pm 1]}}U(1)_{\lambda}=0. Next, we assume λ<12\lambda<\frac{1}{2}. Let FU(1)F\subset U(1) be a generating set and let k=#Fk=\#F. One easily sees that there are at most 3k3^{k} elements of the form Fαjj\sum_{F}\alpha_{j}j. The subsets {xU(1)d(x,αjj)λ}\{x\in U(1)\mid d(x,\alpha_{j}j)\leq\lambda\} cover U(1)U(1), and since each of them has measure 2λ2\lambda one gets the inequality 2λ3k12\lambda\cdot 3^{k}\geq 1. Thus klogλlog2log3k\geq\frac{-\log\lambda-\log 2}{\log 3}. For λ=16\lambda=\frac{1}{6} one has k1k\geq 1 and the subset F={13}F=\{\frac{1}{3}\} generates, thus dim𝕊[±1]U(1)λ=1\dim_{{{{\mathbb{S}}}[\pm 1]}}U(1)_{\lambda}=1. When logλlog2log3=m\frac{-\log\lambda-\log 2}{\log 3}=m is an integer, one has λ=123m\lambda=\frac{1}{2}3^{-m}. Let F(m)={13,,13m}F(m)=\{\frac{1}{3},\ldots,\frac{1}{3^{m}}\}. The minimal distance between two elements of F(m)F(m) is the distance between 3m3^{-m} and 3m+1=33m3^{-m+1}=3\cdot 3^{-m} which is 23m=4λ2\cdot 3^{-m}=4\lambda. Let us show that F(m)F(m) is a generating set. By Lemma 3.1 any integer qq in the interval [N,N][-N,N], for N=12(3m1)N=\frac{1}{2}(3^{m}-1) can be written as q=i=0m1αi3iq=\sum_{i=0}^{m-1}\alpha_{i}3^{i}, with αi{1,0,1}\alpha_{i}\in\{-1,0,1\}. One then gets

q3m=i=0m1αi3im=j=1mαmj3j.q\cdot 3^{-m}=\sum_{i=0}^{m-1}\alpha_{i}3^{i-m}=\sum_{j=1}^{m}\alpha_{m-j}3^{-j}.

Let x[12,12]x\in[-\frac{1}{2},\frac{1}{2}], then 3mx[3m2,3m2]3^{m}x\in[-\frac{3^{m}}{2},\frac{3^{m}}{2}] and there exists an integer q[N,N]q\in[-N,N] such that |3mxq|12|3^{m}x-q|\leq\frac{1}{2}. Hence d(x,q3m)λd(x,q\cdot 3^{-m})\leq\lambda. This proves that F(m)F(m) is a generating set (see Definition A.2) and one derives dim𝕊[±1]U(1)λ=m\dim_{{{{\mathbb{S}}}[\pm 1]}}U(1)_{\lambda}=m. Assume now that logλlog2log3(m,m+1)\frac{-\log\lambda-\log 2}{\log 3}\in(m,m+1), where mm is an integer. For any generating set FF of cardinality kk one has klogλlog2log3>mk\geq\frac{-\log\lambda-\log 2}{\log 3}>m so that km+1k\geq m+1. The subset F(m+1)={13,,13m+1}F(m+1)=\{\frac{1}{3},\ldots,\frac{1}{3^{m+1}}\} fulfills the first condition of Definition A.2 since the minimal distance between two elements of F(m+1)F(m+1) is 23m12\cdot 3^{-m-1} which is larger than λ<123m\lambda<\frac{1}{2}3^{-m}. As shown above, the subset F(m+1)F(m+1) is generating for λ=123m1\lambda=\frac{1}{2}3^{-m-1} and a fortiori for λ>123m1\lambda>\frac{1}{2}3^{-m-1} (as by assumption logλlog2log3<m+1\frac{-\log\lambda-\log 2}{\log 3}<m+1). Thus one obtains dim𝕊[±1]U(1)λ=m+1\dim_{{{{\mathbb{S}}}[\pm 1]}}U(1)_{\lambda}=m+1 and (4.1) is proven.

Remark 4.2.

The proof of Proposition 4.1 relies on a definite advantage of the triadic expansion of numbers [15]: truncating a number is identical to rounding it. This property does not hold in the decimal system where rounding a number requires the knowledge of the next digit.

We extend the ceiling function to negative values of the variable as follows

x={xifx0xifx0.\lceil x\rceil^{\prime}=\begin{cases}\lceil x\rceil\ \text{if}\ x\geq 0\\ -\lceil-x\rceil\ \text{if}\ x\leq 0.\end{cases} (4.2)

We can now state and prove the absolute Riemann-Roch theorem for Spec¯{\overline{{\rm Spec\,}{\mathbb{Z}}}} over 𝕊[±1]{{{\mathbb{S}}}[\pm 1]}. With LL the exceptional set defined in (3.4) one has

Theorem 4.3.

Let DD be an Arakelov divisor on Spec¯{\overline{{\rm Spec\,}{\mathbb{Z}}}}. Then

dim𝕊[±1]H0(D)dim𝕊[±1]H1(D)=degD+log2log3𝟏L.\dim_{{{{\mathbb{S}}}[\pm 1]}}H^{0}(D)-\dim_{{{{\mathbb{S}}}[\pm 1]}}H^{1}(D)=\bigg{\lceil}\frac{\deg D+\log 2}{\log 3}\bigg{\rceil}^{\prime}-{\mathbf{1}}_{L}. (4.3)
Proof 7.

By appealing to the invariance under linear equivalence (Proposition 2.2 (iii)(iii)), one may assume that D=a{}D=a\{\infty\}. Then it follows from Proposition A.5 (ii)(ii) that dim𝕊[±1]H1(D)=dim𝕊[±1]U(1)λ\dim_{{{{\mathbb{S}}}[\pm 1]}}H^{1}(D)=\dim_{{{{\mathbb{S}}}[\pm 1]}}U(1)_{\lambda} for λ=exp(deg(D))\lambda=\exp(\deg(D)). Assume first that exp(deg(D))12\exp(\deg(D))\geq\frac{1}{2}, then by Proposition 4.1: dim𝕊[±1]H1(D)=0\dim_{{{{\mathbb{S}}}[\pm 1]}}H^{1}(D)=0, thus for deg(D)log2\deg(D)\geq-\log 2 (4.3) follows from (3.5). Let us now assume that deg(D)<log2\deg(D)<-\log 2. Then dim𝕊[±1]H0(D)=0\dim_{{{{\mathbb{S}}}[\pm 1]}}H^{0}(D)=0 since H0(D)={}H^{0}(D)=\{\ast\} when degD<0\deg D<0. Moreover degDL\deg D\notin L since LL is lower bounded by log2-\log 2. Thus (4.3) follows from (4.1) which gives, using (4.2)

dim𝕊[±1]H1(D)=dim𝕊[±1]U(1)λ=logλlog2log3=degD+log2log3.-\dim_{{{{\mathbb{S}}}[\pm 1]}}H^{1}(D)=-\dim_{{{{\mathbb{S}}}[\pm 1]}}U(1)_{\lambda}=-\bigg{\lceil}\frac{-\log\lambda-\log 2}{\log 3}\bigg{\rceil}=\bigg{\lceil}\frac{\deg D+\log 2}{\log 3}\bigg{\rceil}^{\prime}.

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 𝕊[±1]{{{\mathbb{S}}}[\pm 1]}-modules, for any divisor DD on Spec¯{\overline{{\rm Spec\,}{\mathbb{Z}}}}:

H0(D)Hom¯Γ𝒯(H1(KD),U(1)14).H^{0}(D)\simeq{\underline{{\rm{Hom}}}}_{\Gamma{\mathcal{T}}_{*}}(H^{1}(K-D),U(1)_{\frac{1}{4}}).

Here, the divisor K=2{2}K=-2\{2\} plays the role of the canonical divisor. The choice of the tolerant 𝕊[±1]{{{\mathbb{S}}}[\pm 1]}-module (U(1),d)14(U(1),d)_{\frac{1}{4}} as dualizing module is motivated by Pontryagin duality (see 5.2). One has dim𝕊[±1](U(1)14)=1\dim_{{{{\mathbb{S}}}[\pm 1]}}(U(1)_{\frac{1}{4}})=1. This equality, in fact, also holds for the tolerant 𝕊[±1]{{{\mathbb{S}}}[\pm 1]}-module U(1)λU(1)_{\lambda} for 16λ<12\frac{1}{6}\leq\lambda<\frac{1}{2} : the specific choice λ=14\lambda=\frac{1}{4} is dictated by the invariance of the Riemann-Roch formula (4.3) when one switches H0H^{0} and H1H^{1} and replaces DD by KDK-D (ignoring the exceptional set LL).

5.1 Hom¯𝕊(Hλ,Hμ){\underline{{\rm{Hom}}}}_{{\mathbb{S}}}(\|H{\mathbb{R}}\|_{\lambda},\|H{\mathbb{R}}\|_{\mu})

We start with the following general statement.

Lemma 5.1.

Let λ,μ>0\lambda,\mu\in{\mathbb{R}}_{>0}. The 𝕊{{\mathbb{S}}}-algebra structure of HH{\mathbb{R}} induces an isomorphism of 𝕊{{\mathbb{S}}}-modules

Hμ/λHom¯𝕊(Hλ,Hμ).\|H{\mathbb{R}}\|_{\mu/\lambda}\simeq{\underline{{\rm{Hom}}}}_{{\mathbb{S}}}(\|H{\mathbb{R}}\|_{\lambda},\|H{\mathbb{R}}\|_{\mu}). (5.1)
Proof 8.

One starts by defining the morphism of 𝕊{{\mathbb{S}}}-modules

η:Hμ/λHom¯𝕊(Hλ,Hμ).\eta:\|H{\mathbb{R}}\|_{\mu/\lambda}\to{\underline{{\rm{Hom}}}}_{{\mathbb{S}}}(\|H{\mathbb{R}}\|_{\lambda},\|H{\mathbb{R}}\|_{\mu}). (5.2)

Precisely, the multiplication in the 𝕊{{\mathbb{S}}}-algebra HH{\mathbb{R}} determines natural maps

m:H(X)H(Y)H(XY)m:H{\mathbb{R}}(X)\wedge H{\mathbb{R}}(Y)\to H{\mathbb{R}}(X\wedge Y)

inducing, for a fixed pointed set XX, the map mX:H(X)Hom𝕊(H,H(X))m_{X}:H{\mathbb{R}}(X)\to{\rm{Hom}}_{{\mathbb{S}}}(H{\mathbb{R}},H{\mathbb{R}}(X\wedge-)). The morphism η\eta is defined as the natural transformation of functors which associates to X=k+X=k_{+}, the map ηX:Hμ/λ(X)Hom𝕊(Hλ,Hμ(X))\eta_{X}:\|H{\mathbb{R}}\|_{\mu/\lambda}(X)\to{\rm{Hom}}_{{\mathbb{S}}}(\|H{\mathbb{R}}\|_{\lambda},\|H{\mathbb{R}}\|_{\mu}(X\wedge-)) defined as the restriction of mXm_{X}. This restriction is meaningful in view of [2] (Proposition 6.1).
Next, we show that η\eta is an isomorphism. First, we determine the 𝕊{{\mathbb{S}}}-module Hom𝕊(Hλ,Hμ){\rm{Hom}}_{{\mathbb{S}}}(\|H{\mathbb{R}}\|_{\lambda},\|H{\mathbb{R}}\|_{\mu}). For integers 1jk1\leq j\leq k we let as in (2.1)

δ(j,k)HomΓo(k+,1+),δ(j,k)():={1if =jif j.\delta(j,k)\in{\rm{Hom}}_{\Gamma^{o}}(k_{+},1_{+}),\quad\delta(j,k)(\ell):=\begin{cases}1&\text{if $\ell=j$}\\ *&\text{if $\ell\neq j$}.\end{cases}

Let ϕHom𝕊(Hλ,Hμ)\phi\in{\rm{Hom}}_{{\mathbb{S}}}(\|H{\mathbb{R}}\|_{\lambda},\|H{\mathbb{R}}\|_{\mu}). By construction, the natural transformation ϕ\phi reads, for each kk, as a map ϕ(k+):Hλ(k+)Hμ(k+)\phi(k_{+}):\|H{\mathbb{R}}\|_{\lambda}(k_{+})\to\|H{\mathbb{R}}\|_{\mu}(k_{+}) and by naturality we have

Hμ(δ(j,k))ϕ(k+)=ϕ(1+)Hλ(δ(j,k)).\|H{\mathbb{R}}\|_{\mu}(\delta(j,k))\circ\phi(k_{+})=\phi(1_{+})\circ\|H{\mathbb{R}}\|_{\lambda}(\delta(j,k)). (5.3)

Since an element yHμ(k+)y\in\|H{\mathbb{R}}\|_{\mu}(k_{+}) is determined by its components yjy_{j}\in{\mathbb{R}}, 1jk1\leq j\leq k, with yj=Hμ(δ(j,k))(y)y_{j}=\|H{\mathbb{R}}\|_{\mu}(\delta(j,k))(y), (5.3) shows that ϕ(k+)\phi(k_{+}) is uniquely determined by the map ϕ(1+)\phi(1_{+}) acting componentwise, i.e.

ϕ(k+)((xj))=(ϕ(1+)(xj)).\phi(k_{+})((x_{j}))=\left(\phi(1_{+})(x_{j})\right). (5.4)

Moreover, the map f=ϕ(1+):[λ,λ][μ,μ]f=\phi(1_{+}):[-\lambda,\lambda]\to[-\mu,\mu] fulfills

f(x+y)=f(x)+f(y)x,ys.t.|x|+|y|λ,f(x+y)=f(x)+f(y)\quad\forall x,y\ s.t.\ |x|+|y|\leq\lambda, (5.5)

as one sees using the naturality of ϕ\phi for the map σHomΓo(2+,1+)\sigma\in{\rm{Hom}}_{\Gamma^{o}}(2_{+},1_{+}), i.e. using

(Hμ(σ)ϕ(2+))(x,y)=(ϕ(1+)Hλ(σ))(x,y)f(x)+f(y)=f(x+y).(\|H{\mathbb{R}}\|_{\mu}(\sigma)\circ\phi(2_{+}))(x,y)=(\phi(1_{+})\circ\|H{\mathbb{R}}\|_{\lambda}(\sigma))(x,y)\Longrightarrow f(x)+f(y)=f(x+y).

By (5.5) one has f(x2n)=2nf(x)f(x2^{-n})=2^{-n}f(x), for any x[λ,λ]x\in[-\lambda,\lambda] and n0n\geq 0. Thus the ‘germ of map’ ff uniquely extends to a map f~:\tilde{f}:{\mathbb{R}}\to{\mathbb{R}} defined by f~(x):=2nf(x2n)\tilde{f}(x):=2^{n}f(x2^{-n}) for any nn such that |x2n|λ|x2^{-n}|\leq\lambda. Moreover, again by (5.5), the map f~\tilde{f} is additive and since f~([2nλ,2nλ])[2nμ,2nμ]\tilde{f}([-2^{-n}\lambda,2^{-n}\lambda])\subset[-2^{-n}\mu,2^{-n}\mu], it is also continuous and hence determined by the multiplication by a real number rr. One has |rλ|μ|r\lambda|\leq\mu and hence rHμ/λ(1+)r\in\|H{\mathbb{R}}\|_{\mu/\lambda}(1_{+}). Thus one gets ϕ(1+)(x)=rx\phi(1_{+})(x)=rx, x[λ,λ]\forall x\in[-\lambda,\lambda]. By (5.4) one obtains ϕ(k+)((xj))=(rxj)\phi(k_{+})((x_{j}))=(rx_{j}), (xj)Hλ(k+)\forall(x_{j})\in\|H{\mathbb{R}}\|_{\lambda}(k_{+}). This shows that for X=1+X=1_{+}, ηX:Hμ/λ(X)Hom𝕊(Hλ,Hμ(X))\eta_{X}:\|H{\mathbb{R}}\|_{\mu/\lambda}(X)\to{\rm{Hom}}_{{\mathbb{S}}}(\|H{\mathbb{R}}\|_{\lambda},\|H{\mathbb{R}}\|_{\mu}(X\wedge-)) is bijective. The next step is to determine Hom𝕊(Hλ,Hμ(X)){\rm{Hom}}_{{\mathbb{S}}}(\|H{\mathbb{R}}\|_{\lambda},\|H{\mathbb{R}}\|_{\mu}(X\wedge-)) for X=+X=\ell_{+}. Let ϕHom𝕊(Hλ,Hμ(+))\phi\in{\rm{Hom}}_{{\mathbb{S}}}(\|H{\mathbb{R}}\|_{\lambda},\|H{\mathbb{R}}\|_{\mu}(\ell_{+}\wedge-)). One has +k+=(k)+\ell_{+}\wedge k_{+}=(\ell k)_{+} and an element zHμ(+k+)z\in\|H{\mathbb{R}}\|_{\mu}(\ell_{+}\wedge k_{+}) is determined by its components z(i,j)z(i,j)\in{\mathbb{R}} for 1i1\leq i\leq\ell, 1jk1\leq j\leq k such that |z(i,j)|μ\sum|z(i,j)|\leq\mu. In particular, for each jj, the z(i,j)z(i,j), 1i1\leq i\leq\ell, are the components of zj=Hμ(idδ(j,k)))(z)Hμ(+)z_{j}=\|H{\mathbb{R}}\|_{\mu}({\rm id}\wedge\delta(j,k)))(z)\in\|H{\mathbb{R}}\|_{\mu}(\ell_{+}). This implies by applying the naturality of ϕ\phi, that

ϕ(k+)(z)=(ϕ(1+)(zj)).\phi(k_{+})(z)=\left(\phi(1_{+})(z_{j})\right). (5.6)

As shown above the map f=ϕ(1+):[λ,λ]Hμ(+)f=\phi(1_{+}):[-\lambda,\lambda]\to\|H{\mathbb{R}}\|_{\mu}(\ell_{+}) fulfills

f(x+y)=f(x)+f(y)x,ys.t.|x|+|y|λf(x+y)=f(x)+f(y)\quad\forall x,y\ s.t.\ |x|+|y|\leq\lambda (5.7)

and it extends to an additive map f~:\tilde{f}:{\mathbb{R}}\to{\mathbb{R}}^{\ell} which is continuous since f~\tilde{f} maps the interval [λ,λ][-\lambda,\lambda] inside Hμ(+)\|H{\mathbb{R}}\|_{\mu}(\ell_{+}). Thus there exists real numbers rir_{i}, 1i1\leq i\leq\ell, such that f~(x)=(rix)\tilde{f}(x)=(r_{i}x) x\forall x\in{\mathbb{R}}. One has |riλ|μ\sum|r_{i}\lambda|\leq\mu and thus r=(ri)Hμ/λ(+)r=(r_{i})\in\|H{\mathbb{R}}\|_{\mu/\lambda}(\ell_{+}). Finally, using (5.6) it follows that ϕ=ηX(r)\phi=\eta_{X}(r). This proves that η\eta as in (5.2) is an isomorphism.

5.2 Pontryagin duality

In order to formulate Pontryagin duality in this context we consider, for λ>0\lambda>0, the 𝕊[±1]{{{\mathbb{S}}}[\pm 1]}-module U(1)λ=(U(1),d)λU(1)_{\lambda}=(U(1),d)_{\lambda}, where U(1)U(1) is the abelian group /{\mathbb{R}}/{\mathbb{Z}} endowed with its canonical metric dd of length 11. For a metric abelian group (A,d)(A,d), we denote by A^\widehat{A} the abelian group of continuous characters, i.e. of continuous group homomorphisms AU(1)A\to U(1), where AA is endowed with the topology associated to the metric dd. We retain the notations of section A. Next statement is motivated by Lemma 5.1.

Proposition 5.2.

Let (A,d)(A,d) be an abelian group endowed with a translation invariant metric dd.

  1. (i)

    For λ,μ>0\lambda,\mu\in{\mathbb{R}}_{>0}, the 𝕊[±1]{{{\mathbb{S}}}[\pm 1]}-module Hom¯Γ𝒯((A,d)λ,U(1)μ){\underline{{\rm{Hom}}}}_{\Gamma{\mathcal{T}}_{*}}((A,d)_{\lambda},U(1)_{\mu}) is isomorphic to the sub 𝕊[±1]{{{\mathbb{S}}}[\pm 1]}-module of HA^H\widehat{A} which, on the set k+k_{+}, is given by k-tuples (χj)(\chi_{j}), 1jk1\leq j\leq k of continuous characters χjA^\chi_{j}\in\widehat{A} such that, with |x|:=d(x,0)|x|:=d(x,0), and for all finite collections {xi}A\{x_{i}\}\subset A, fulfill

    i|xi|λi,j|χj(xi)|μ.\sum_{i}|x_{i}|\leq\lambda~\Rightarrow~\sum_{i,j}|\chi_{j}(x_{i})|\leq\mu. (5.8)
  2. (ii)

    Let p:AAp:A^{\prime}\to A be a surjective morphism of abelian groups and let p(A,d)λp^{*}(A,d)_{\lambda} be the pullback as in Proposition A.4. One has the following canonical isomorphism

    Hom¯Γ𝒯(p(A,d)λ,U(1)μ)Hom¯Γ𝒯((A,d)λ,U(1)μ).{\underline{{\rm{Hom}}}}_{\Gamma{\mathcal{T}}_{*}}(p^{*}(A,d)_{\lambda},U(1)_{\mu})\simeq{\underline{{\rm{Hom}}}}_{\Gamma{\mathcal{T}}_{*}}((A,d)_{\lambda},U(1)_{\mu}). (5.9)
Proof 9.

(i)(i) Let ϕHom¯Γ𝒯((A,d)λ,U(1)μ)(1+)=HomΓ𝒯((A,d)λ,U(1)μ)\phi\in{\underline{{\rm{Hom}}}}_{\Gamma{\mathcal{T}}_{*}}((A,d)_{\lambda},U(1)_{\mu})(1_{+})={\rm{Hom}}_{\Gamma{\mathcal{T}}_{*}}((A,d)_{\lambda},U(1)_{\mu}). By applying the forgetful functor :𝒯𝔖𝔢𝔱𝔰{\mathcal{F}}:{\mathcal{T}}\longrightarrow\mathfrak{Sets} which associates to (X,)(X,{\mathcal{R}}) the set XX (see Proposition A.2), one obtains an element (ϕ)Hom𝕊(HA,HU(1)){\mathcal{F}}(\phi)\in{\rm{Hom}}_{{\mathbb{S}}}(HA,HU(1)). Since the Eilenberg-MacLane functor HH determines a full and faithful embedding of the category of abelian groups inside the category of 𝕊{{\mathbb{S}}}-modules, there exists a unique group homomorphism χ:AU(1)\chi:A\to U(1) such that H(χ)=(ϕ)H(\chi)={\mathcal{F}}(\phi). The condition that ϕ\phi preserves the relation k{\mathcal{R}}_{k} on the set k+k_{+} means that for any xi,yiAx_{i},y_{i}\in A, 1ik1\leq i\leq k such that d(xi,yi)λ\sum d(x_{i},y_{i})\leq\lambda one has d(χ(xi),χ(yi))μ\sum d(\chi(x_{i}),\chi(y_{i}))\leq\mu. By translation invariance of the metrics this condition is equivalent to

i|xi|λi|χ(xi)|μ.\sum_{i}|x_{i}|\leq\lambda~\Rightarrow~\sum_{i}|\chi(x_{i})|\leq\mu. (5.10)

This shows that HomΓ𝒯((A,d)λ,U(1)μ){\rm{Hom}}_{\Gamma{\mathcal{T}}_{*}}((A,d)_{\lambda},U(1)_{\mu}) consists exactly of the group homomorphisms χ:AU(1)\chi:A\to U(1) fulfilling (5.8). Specializing (5.10) to the case where all xi=xx_{i}=x, i=1,,ni=1,\ldots,n, one obtains the implication |x|λ/n|χ(x)|μ/n|x|\leq\lambda/n\Rightarrow|\chi(x)|\leq\mu/n, and hence that χ\chi is uniformly continuous. Let ϕHom¯Γ𝒯((A,d)λ,U(1)μ)(k+)=HomΓ𝒯((A,d)λ,U(1)μ(k+))\phi\in{\underline{{\rm{Hom}}}}_{\Gamma{\mathcal{T}}_{*}}((A,d)_{\lambda},U(1)_{\mu})(k_{+})={\rm{Hom}}_{\Gamma{\mathcal{T}}_{*}}((A,d)_{\lambda},U(1)_{\mu}(k_{+}\wedge-)). The object U(1)μ(k+)U(1)_{\mu}(k_{+}\wedge-) of Γ𝒯\Gamma{\mathcal{T}}_{*} is (U(1)×k,dk)μ(U(1)^{\times k},d_{k})_{\mu}, where the metric dkd_{k} on the product group (U(1)×k(U(1)^{\times k} is defined by

dk((xi),(yi)):=1kd(xi,yi)xi,yiU(1).d_{k}((x_{i}),(y_{i})):=\sum_{1}^{k}d(x_{i},y_{i})\quad\forall x_{i},y_{i}\in U(1).

Replacing U(1)μU(1)_{\mu} with (U(1)×k,dk)μ(U(1)^{\times k},d_{k})_{\mu} in the first part of the proof one obtains that (ϕ)Hom𝕊(HA,HU(1)×k)=H((χj)){\mathcal{F}}(\phi)\in{\rm{Hom}}_{{\mathbb{S}}}(HA,HU(1)^{\times k})=H((\chi_{j})), where (χj)(\chi_{j}), 1jk1\leq j\leq k is a kk-tuple of characters of AA fulfilling (5.8). It follows that χjA^\chi_{j}\in\widehat{A}.
(ii)(ii) Let ϕHom¯Γ𝒯(p(A,d)λ,U(1)μ)(1+)=HomΓ𝒯(p(A,d)λ,U(1)μ)\phi\in{\underline{{\rm{Hom}}}}_{\Gamma{\mathcal{T}}_{*}}(p^{*}(A,d)_{\lambda},U(1)_{\mu})(1_{+})={\rm{Hom}}_{\Gamma{\mathcal{T}}_{*}}(p^{*}(A,d)_{\lambda},U(1)_{\mu}). As in the proof of (i)(i), there exists a group homomorphism χ:AU(1)\chi^{\prime}:A^{\prime}\to U(1) such that H(χ)=(ϕ)H(\chi^{\prime})={\mathcal{F}}(\phi). Moreover χ\chi^{\prime} preserves the relation k{\mathcal{R}}_{k} for any kk, and this implies

i=1k|p(xi)|λi=1k|χ(xi)|μ,(xi)(A)×k.\sum_{i=1}^{k}|p(x^{\prime}_{i})|\leq\lambda~\Rightarrow~\sum_{i=1}^{k}|\chi^{\prime}(x^{\prime}_{i})|\leq\mu,\qquad\forall(x^{\prime}_{i})\in(A^{\prime})^{\times k}.

In particular, taking all xi=xker(p)x^{\prime}_{i}=x^{\prime}\in\ker(p) one obtains |χ(x)|μ/k|\chi^{\prime}(x^{\prime})|\leq\mu/k k\forall k, and hence χ(ker(p))={1}\chi^{\prime}(\ker(p))=\{1\}. This implies that there exists a group homomorphism χ:AU(1)\chi:A\to U(1) such that χ=χp\chi^{\prime}=\chi\circ p.

We can now state and prove Serre’s duality.

Theorem 5.3.

Let D=jaj{pj}+a{}D=\sum_{j}a_{j}\{p_{j}\}+a\{\infty\} be an Arakelov divisor on Spec¯{\overline{{\rm Spec\,}{\mathbb{Z}}}}. There is a canonical isomorphism of 𝕊[±1]{{{\mathbb{S}}}[\pm 1]}-modules

H0(KD)Hom¯Γ𝒯(H1(D),U(1)14),H^{0}(K-D)\simeq{\underline{{\rm{Hom}}}}_{\Gamma{\mathcal{T}}_{*}}(H^{1}(D),U(1)_{\frac{1}{4}}), (5.11)

where KK is the divisor K=2{2}K=-2\{2\}.

Proof 10.

By Proposition A.5, with λ=expa\lambda=\exp a, one has H1(D)=π((/L,d)λ)H^{1}(D)=\pi^{*}(({\mathbb{R}}/L,d)_{\lambda}) and by Proposition 5.2, (ii)(ii), one gets the isomorphism

Hom¯Γ𝒯(H1(D),U(1)14)Hom¯Γ𝒯((/L,d)λ,U(1)14).{\underline{{\rm{Hom}}}}_{\Gamma{\mathcal{T}}_{*}}(H^{1}(D),U(1)_{\frac{1}{4}})\simeq{\underline{{\rm{Hom}}}}_{\Gamma{\mathcal{T}}_{*}}(({\mathbb{R}}/L,d)_{\lambda},U(1)_{\frac{1}{4}}).

In fact, we can assume that D=a{}D=a\{\infty\}, with a=degDa=\deg D so that L=L={\mathbb{Z}}. Then we apply Proposition 5.2 (i)(i), with A=/A={\mathbb{R}}/{\mathbb{Z}} and μ=14\mu=\frac{1}{4}. One has A^=\widehat{A}={\mathbb{Z}} and the characters χnA^\chi_{n}\in\widehat{A} are given by multiplication by nn, i.e. χn(s):=ns/\chi_{n}(s):=ns\in{\mathbb{R}}/{\mathbb{Z}}, s/\forall s\in{\mathbb{R}}/{\mathbb{Z}}. Next, we need to determine the 𝕊[±1]{{{\mathbb{S}}}[\pm 1]}-submodule of HA^=HH\widehat{A}=H{\mathbb{Z}} which, on the set k+k_{+}, is given by kk-tuples (nj)(n_{j}), 1jk1\leq j\leq k of characters njn_{j}\in{\mathbb{Z}} such that (5.8) holds. This means, using the distance dd on /{\mathbb{R}}/{\mathbb{Z}}, that

id(xi,0)λi,jd(njxi,0)14.\sum_{i}d(x_{i},0)\leq\lambda~\Rightarrow~\sum_{i,j}d(n_{j}x_{i},0)\leq\frac{1}{4}. (5.12)

The distance d(x,0)d(x,0) is given, for any xx^{\prime}\in{\mathbb{R}} in the class of xx by the distance between xx^{\prime} and the closed subset {\mathbb{Z}}\subset{\mathbb{R}}. Thus for any integer nn one has: d(nx,0)|n|d(x,0)d(nx,0)\leq|n|d(x,0). Assume that j|nj|14λ\sum_{j}|n_{j}|\leq\frac{1}{4\lambda}, then (5.12) follows since

i,jd(njxi,0)i,j|nj|d(xi,0)14λid(xi,0).\sum_{i,j}d(n_{j}x_{i},0)\leq\sum_{i,j}|n_{j}|d(x_{i},0)\leq\frac{1}{4\lambda}\sum_{i}d(x_{i},0).

Conversely, assume (5.12). Then repeating mm times the same xx, gives

d(x,0)λmjd(njx,0)14m.d(x,0)\leq\frac{\lambda}{m}~\Rightarrow~\sum_{j}d(n_{j}x,0)\leq\frac{1}{4m}.

Taking mm large enough and x=λmx=\frac{\lambda}{m} one obtains j|nj|x14m\sum_{j}|n_{j}|x\leq\frac{1}{4m}, and hence j|nj|14λ\sum_{j}|n_{j}|\leq\frac{1}{4\lambda}. This proves that the 𝕊[±1]{{{\mathbb{S}}}[\pm 1]}-submodule of HA^=HH\widehat{A}=H{\mathbb{Z}} determined by (5.8) is equal to H14λ\|H{\mathbb{Z}}\|_{\frac{1}{4\lambda}} which gives (5.11).

Appendix A Tolerance 𝕊{{\mathbb{S}}}-modules

The construction of the category Γ𝒮\Gamma{\mathcal{S}}_{*} of Γ\Gamma-spaces (see appendix B) can be broadly generalized by considering in place of the category 𝒮{\mathcal{S}}_{*} of simplicial pointed sets any pointed category 𝒞{\mathcal{C}} with initial and final object *. In this way, one obtains a category Γ𝒞\Gamma{\mathcal{C}} of pointed covariant functor Γo𝒞{\Gamma^{o}}\longrightarrow{\mathcal{C}}. We shall apply this formal construction to the category of tolerance relations and introduce the notion of tolerant 𝕊[±1]{{{\mathbb{S}}}[\pm 1]}-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 𝒞{\mathcal{C}} be a pointed category with initial and final object *. Then Γ𝒞\Gamma{\mathcal{C}} is naturally enriched in 𝕊{{\mathbb{S}}}-modules. More precisely, the following formula endows the internal Hom¯Γ𝒞(A,B){\underline{{\rm{Hom}}}}_{\Gamma{\mathcal{C}}}(A,B) with a structure of 𝕊{{\mathbb{S}}}-module defined by

Hom¯Γ𝒞(A,B)(k+):=HomΓ𝒞(A,B(k+))k.{\underline{{\rm{Hom}}}}_{\Gamma{\mathcal{C}}}(A,B)(k_{+}):={\rm{Hom}}_{\Gamma{\mathcal{C}}}(A,B(k_{+}\wedge-))\qquad k\in{\mathbb{N}}. (A.1)
Proof 11.

Let ϕHomΓo(k+,+)\phi\in{\rm{Hom}}_{\Gamma^{o}}(k_{+},\ell_{+}). For every object FF of Γo{\Gamma^{o}} the morphism ϕidHomΓo(k+F,+F)\phi\wedge{\rm id}\in{\rm{Hom}}_{\Gamma^{o}}(k_{+}\wedge F,\ell_{+}\wedge F) gives, by functoriality of B:Γo𝒞B:{\Gamma^{o}}\longrightarrow{\mathcal{C}}, a morphism B(ϕid)Hom𝒞(B(k+F),B(+F))B(\phi\wedge{\rm id})\in{\rm{Hom}}_{\mathcal{C}}(B(k_{+}\wedge F),B(\ell_{+}\wedge F)). These morphisms define a natural transformation of functors B(ϕ)HomΓ𝒞(B(k+),B(+))B(\phi\wedge-)\in{\rm{Hom}}_{\Gamma{\mathcal{C}}}(B(k_{+}\wedge-),B(\ell_{+}\wedge-)), and one obtains the functoriality on the right hand side of (A.1) using the left composition

ψHomΓ𝒞(A,B(k+))B(ϕ)ψHomΓ𝒞(A,B(+)).\psi\in{\rm{Hom}}_{\Gamma{\mathcal{C}}}(A,B(k_{+}\wedge-))\mapsto B(\phi\wedge-)\circ\psi\in{\rm{Hom}}_{\Gamma{\mathcal{C}}}(A,B(\ell_{+}\wedge-)).

A.1 The category Γ𝒯\Gamma{\mathcal{T}}_{*}

A tolerance relation {\mathcal{R}} on a set XX is a reflexive and symmetric relation on XX. Equivalently, {\mathcal{R}} is a subset X×X{\mathcal{R}}\subset X\times X which is symmetric and containing the diagonal. We shall denote by 𝒯{\mathcal{T}} the category of tolerance relations (X,)(X,{\mathcal{R}}). Morphisms in 𝒯{\mathcal{T}} are defined by

Hom𝒯((X,),(X,)):={ϕ:XX,ϕ()}.{\rm{Hom}}_{\mathcal{T}}((X,{\mathcal{R}}),(X^{\prime},{\mathcal{R}}^{\prime})):=\{\phi:X\to X^{\prime},\ \phi({\mathcal{R}})\subset{\mathcal{R}}^{\prime}\}.

We denote 𝒯{\mathcal{T}}_{*} the pointed category under the object {}\{\ast\} endowed with the trivial relation. One has the following

Definition A.1.

A tolerant 𝕊{{\mathbb{S}}}-module is a pointed covariant functor Γo𝒯{\Gamma^{o}}\longrightarrow{\mathcal{T}}_{*}. We denote by Γ𝒯\Gamma{\mathcal{T}}_{*} the category of tolerant 𝕊{{\mathbb{S}}}-modules.

Next statement is an easy but useful fact

Proposition A.2.
  1. (i)

    The functor 𝔖𝔢𝔱𝔰𝒯\mathfrak{Sets}\longrightarrow{\mathcal{T}} which endows a set with the diagonal relation, embeds the category of sets as a full subcategory of 𝒯{\mathcal{T}}, and consequently the category of 𝕊{{\mathbb{S}}}-modules as a full subcategory of the category Γ𝒯\Gamma{\mathcal{T}}_{*}.

  2. (ii)

    The forgetful functor is the right adjoint of the inclusion in (i)(i).

A.2 The tolerant 𝕊{{\mathbb{S}}}-module (A,d)λ(A,d)_{\lambda}

A relevant example of tolerant 𝕊{{\mathbb{S}}}-module is given by the following construction. Let AA be an additive abelian group. A translation invariant metric dd on AA is a metric on AA that satisfies d(x,y)=d(xy,0)d(x,y)=d(x-y,0), so that the triangle inequality can be read as d(x+y,0)d(x,0)+d(y,0)d(x+y,0)\leq d(x,0)+d(y,0) x,yA\forall x,y\in A. This fact implies that the inequality

d(Ixi,Iyi)Id(xi,yi)d\bigg{(}\sum_{I}x_{i},\sum_{I}y_{i}\bigg{)}\leq\sum_{I}d(x_{i},y_{i}) (A.2)

holds for any finite index set II and maps x,y:IAx,y:I\to A.

Proposition A.3.

Let (A,d)(A,d) be an abelian group endowed with a translation invariant metric dd and let λ>0\lambda\in{\mathbb{R}}_{>0}. The following relations turn the Eilenberg?MacLane 𝕊{{\mathbb{S}}}-module HAHA into a tolerant 𝕊[±1]{{{\mathbb{S}}}[\pm 1]}-module (A,d)λ:Γo𝒯(A,d)_{\lambda}:{\Gamma^{o}}\longrightarrow{\mathcal{T}}_{*}:

k={((xi),(yi))A×k×A×kI=k+d(xi,yi)λ}k+,k.\mathcal{R}_{k}=\{((x_{i}),(y_{i}))\in A^{\times k}\times A^{\times k}\mid\sum_{I=k_{+}}d(x_{i},y_{i})\leq\lambda\}\qquad\forall k_{+},~k\in{\mathbb{N}}. (A.3)
Proof 12.

For any ϕHomΓo(k+,+)\phi\in{\rm{Hom}}_{\Gamma^{o}}(k_{+},\ell_{+}), the map HA(ϕ):HA(k+)HA(+)HA(\phi):HA(k_{+})\to HA(\ell_{+}) fulfills HA(ϕ)×2(k)HA(\phi)^{\times 2}({\mathcal{R}}_{k})\subset{\mathcal{R}}_{\ell}. Indeed, this follows from (A.2) applied to the finite sets Ij={ik+ϕ(i)=j}I_{j}=\{i\in k_{+}\mid\phi(i)=j\} which label pairs of elements of AA.

Let U(1)U(1) be the abelian group /{\mathbb{R}}/{\mathbb{Z}} endowed with the canonical metric dd of length 11. We shall denote by U(1)λU(1)_{\lambda} the tolerant 𝕊{{\mathbb{S}}}-module (U(1),d)λ(U(1),d)_{\lambda}, (λ>0\lambda\in{\mathbb{R}}_{>0}).

A.3 The dimension of a tolerant 𝕊[±1]{{{\mathbb{S}}}[\pm 1]}-module

In this part we introduce a notion of dimension for a tolerant 𝕊[±1]{{{\mathbb{S}}}[\pm 1]}-module that naturally generalizes, in the absolute context, the definition of dimension of a vector space.

Definition A.2.

Let (E,)(E,{\mathcal{R}}) be a tolerant 𝕊[±1]{{{\mathbb{S}}}[\pm 1]}-module. A subset FE(1+)F\subset E(1_{+}) generates E(1+)E(1_{+}) if the following two conditions hold

  1. 1.

    For x,yFx,y\in F, with xy(x,y)x\neq y~\Longrightarrow~(x,y)\notin{\mathcal{R}}

  2. 2.

    For every xE(1+)x\in E(1_{+}) there exists αj{1,0,1}\alpha_{j}\in\{-1,0,1\}, jFj\in F and yE(1+)y\in E(1_{+}) such that y=FαjjE(1+)y=\sum_{F}\alpha_{j}j\in E(1_{+}) in the sense of (2.2), and (x,y)(x,y)\in{\mathcal{R}}.

The dimension dim𝕊[±1](E,)\dim_{{{{\mathbb{S}}}[\pm 1]}}(E,{\mathcal{R}}) is defined as the minimal cardinality of a generating set FF.

To familiarize with this notion we prove the following

Proposition A.4.

Let p:ABp:A\to B be a morphism of abelian groups and (HB,)(HB,{\mathcal{R}}) a tolerant 𝕊[±1]{{{\mathbb{S}}}[\pm 1]}-module.

  1. (i)

    Consider the relation p(k):={(x,y)HA(k+)×HA(k+)(Hp(x),Hp(y))k}p^{*}({\mathcal{R}}_{k}):=\{(x,y)\in HA(k_{+})\times HA(k_{+})\mid(Hp(x),Hp(y))\in{\mathcal{R}}_{k}\}. Then the pair p(HB,):=(HA,p())p^{*}(HB,{\mathcal{R}}):=(HA,p^{*}({\mathcal{R}})) is a tolerant 𝕊[±1]{{{\mathbb{S}}}[\pm 1]} module.

  2. (ii)

    If pp is surjective: dim𝕊[±1](p(HB,))=dim𝕊[±1](HB,)\dim_{{{{\mathbb{S}}}[\pm 1]}}(p^{*}(HB,{\mathcal{R}}))=\dim_{{{{\mathbb{S}}}[\pm 1]}}(HB,{\mathcal{R}}).

Proof 13.

(i)(i) Since (HB,)(HB,{\mathcal{R}}) is a tolerant 𝕊[±1]{{{\mathbb{S}}}[\pm 1]}-module, for any ϕHomΓo(k+,+)\phi\in{\rm{Hom}}_{\Gamma^{o}}(k_{+},\ell_{+}) one has HB(ϕ)×2(k)HB(\phi)^{\times 2}({\mathcal{R}}_{k})\subset{\mathcal{R}}_{\ell}. Also

HA(ϕ)×2(pk)pϕHomΓo(k+,+),HA(\phi)^{\times 2}(p^{*}{\mathcal{R}}_{k})\subset p^{*}{\mathcal{R}}_{\ell}\qquad\forall\phi\in{\rm{Hom}}_{\Gamma^{o}}(k_{+},\ell_{+}),

which shows that p(HB,)=(HA,p())p^{*}(HB,{\mathcal{R}})=(HA,p^{*}({\mathcal{R}})) is a tolerant 𝕊[±1]{{{\mathbb{S}}}[\pm 1]} module.
(ii)(ii) Let FHB(1+)=BF\subset HB(1_{+})=B be a generating set for (HB,)(HB,{\mathcal{R}}), and let FHA(1+)=AF^{\prime}\subset HA(1_{+})=A be a lift of FF, with #F=#F\#F^{\prime}=\#F. Let us show that FF^{\prime} is a generating set for p(HB,)=(HA,p())p^{*}(HB,{\mathcal{R}})=(HA,p^{*}({\mathcal{R}})). Let aHA(1+)=Aa\in HA(1_{+})=A, then there exists coefficients αj{1,0,1}\alpha_{j}\in\{-1,0,1\}, jFj\in F, such that (p(a),Fαjj)(p(a),\sum_{F}\alpha_{j}j)\in{\mathcal{R}}. It follows that using the lifts jFj^{\prime}\in F^{\prime} of jFj\in F one has (a,Fαjj)p()(a,\sum_{F^{\prime}}\alpha_{j}j^{\prime})\in p^{*}({\mathcal{R}}), hence FF^{\prime} is a generating set for p(HB,)p^{*}(HB,{\mathcal{R}}). Thus dim𝕊[±1](p(HB,))dim𝕊[±1](HB,)\dim_{{{{\mathbb{S}}}[\pm 1]}}(p^{*}(HB,{\mathcal{R}}))\leq\dim_{{{{\mathbb{S}}}[\pm 1]}}(HB,{\mathcal{R}}). Conversely, let FAF^{\prime}\subset A be a generating set for p(HB,)p^{*}(HB,{\mathcal{R}}) and F:=p(F)F:=p(F^{\prime}). Then condition 1. of Definition A.2 for FF^{\prime} implies the same condition for FF, thus one has #F=#F\#F^{\prime}=\#F. Let bHB(1+)=Bb\in HB(1_{+})=B and aAa\in A with p(a)=bp(a)=b. Then there exists coefficients αj{1,0,1}\alpha_{j}\in\{-1,0,1\}, jFj\in F^{\prime}, such that (a,Fαjj)p()(a,\sum_{F^{\prime}}\alpha_{j}j^{\prime})\in p^{*}({\mathcal{R}}). This implies (b,αjj)(b,\sum\alpha_{j}j)\in{\mathcal{R}} so that FF is a generating set for (HB,)(HB,{\mathcal{R}}).

Next, we apply this functorial machinery to the geometry of Spec¯{\overline{{\rm Spec\,}{\mathbb{Z}}}}. We retain the notations of section 2.

Proposition A.5.

Let D=jaj{pj}+a{}D=\sum_{j}a_{j}\{p_{j}\}+a\{\infty\} be an Arakelov divisor on Spec¯{\overline{{\rm Spec\,}{\mathbb{Z}}}} and let π:𝔸/L\pi:{\mathbb{A}}_{\mathbb{Q}}\to{\mathbb{R}}/L be the projection of the adeles on their archimedean component modulo the lattice

L=H0(Spec,𝒪(D)f):={q|q|νexp(D)(ν),ν}.L=H^{0}({\rm Spec\,}{\mathbb{Z}},{\mathcal{O}}(D)_{f}):=\{q\in{\mathbb{Q}}\mid|q|_{\nu}\leq\exp(D)(\nu),~\forall\nu\neq\infty\}. (A.4)
  1. (i)

    Let dd be the metric on /L{\mathbb{R}}/L induced by the standard metric on {\mathbb{R}} and set λ=expa\lambda=\exp a. Then one has H1(D)=π((/L,d)λ)H^{1}(D)=\pi^{*}(({\mathbb{R}}/L,d)_{\lambda}).

  2. (ii)

    dim𝕊[±1]H1(D)=dim𝕊[±1](/L,d)λ\dim_{{{{\mathbb{S}}}[\pm 1]}}H^{1}(D)=\dim_{{{{\mathbb{S}}}[\pm 1]}}({\mathbb{R}}/L,d)_{\lambda}.

Proof 14.

(i)(i) Let j:𝔸f𝔸j:{\mathbb{A}}_{\mathbb{Q}}^{f}\to{\mathbb{A}}_{\mathbb{Q}}, j(x):=(x,0)j(x):=(x,0), be the embedding of finite adeles in adeles. Using the ultrametric property of the local norms at the finite places one sees that j(𝒪(D)f)=𝒪(D)j(𝔸f)𝔸j({\mathcal{O}}(D)_{f})={\mathcal{O}}(D)\cap j({\mathbb{A}}_{\mathbb{Q}}^{f})\subset{\mathbb{A}}_{\mathbb{Q}} is a compact subgroup. Set G=×j(𝒪(D)f)G={\mathbb{Q}}\times j({\mathcal{O}}(D)_{f}): one has j(𝒪(D)f)={0}j({\mathcal{O}}(D)_{f})\cap{\mathbb{Q}}=\{0\} since all the adeles in j(𝒪(D)f)j({\mathcal{O}}(D)_{f}) have archimedean component equal to 0. Thus, the restriction of the morphism of addition, α:×𝔸𝔸\alpha:{\mathbb{Q}}\times{\mathbb{A}}_{\mathbb{Q}}\to{\mathbb{A}}_{\mathbb{Q}}, α(q,a)=q+a\alpha(q,a)=q+a, to GG determines an isomorphism of GG with the subgroup α(G)=+j(𝒪(D)f)\alpha(G)={\mathbb{Q}}+j({\mathcal{O}}(D)_{f}) of 𝔸{\mathbb{A}}_{\mathbb{Q}}. Note, in particular, that α(G)\alpha(G) is closed in 𝔸{\mathbb{A}}_{\mathbb{Q}}, since {\mathbb{Q}} is discrete (hence closed) and j(𝒪(D)f)j({\mathcal{O}}(D)_{f}) is compact. In the following, we identify (set-theoretically) 𝔸{\mathbb{A}}_{\mathbb{Q}} with the product 𝔸f×{\mathbb{A}}^{f}_{\mathbb{Q}}\times{\mathbb{R}} endowed with the two projection morphisms pf:𝔸𝔸fp_{f}:{\mathbb{A}}_{\mathbb{Q}}\to{\mathbb{A}}^{f}_{\mathbb{Q}} and p:𝔸p_{\infty}:{\mathbb{A}}_{\mathbb{Q}}\to{\mathbb{R}}. The subgroup pf()𝔸fp_{f}({\mathbb{Q}})\subset{\mathbb{A}}^{f}_{\mathbb{Q}} is dense and the subgroup pf(j(𝒪(D)f))=𝒪(D)f𝔸fp_{f}(j({\mathcal{O}}(D)_{f}))={\mathcal{O}}(D)_{f}\subset{\mathbb{A}}^{f}_{\mathbb{Q}} is open. Hence pf(α(G))=𝔸fp_{f}(\alpha(G))={\mathbb{A}}^{f}_{\mathbb{Q}}. Thus the projection pp_{\infty} induces the isomorphism of groups p:𝔸/α(G)/Lp:{\mathbb{A}}_{\mathbb{Q}}/\alpha(G)\stackrel{{\scriptstyle\sim}}{{\to}}{\mathbb{R}}/L, where L=p(ker(pfα))L=p_{\infty}(\ker(p_{f}\circ\alpha)). The kernel of the composite pfα:G𝔸fp_{f}\circ\alpha:G\to{\mathbb{A}}_{\mathbb{Q}}^{f} is the group of pairs (q,a)×j(𝒪(D)f)(q,a)\in{\mathbb{Q}}\times j({\mathcal{O}}(D)_{f}) such that pf(q)+pf(a)=0p_{f}(q)+p_{f}(a)=0. Such pairs are determined by the value of qq and thus

L=p(H0(Spec,𝒪(D)f))={q|q|νexp(D)(ν),ν}.L=p_{\infty}(H^{0}({\rm Spec\,}{\mathbb{Z}},{\mathcal{O}}(D)_{f}))=\{q\in{\mathbb{Q}}\mid|q|_{\nu}\leq\exp(D)(\nu),\quad\forall\nu\neq\infty\}. (A.5)

By Proposition 2.2, H1(D)H^{1}(D) is the tolerant 𝕊[±1]{{{\mathbb{S}}}[\pm 1]}-module H1(D)=(H𝔸,)H^{1}(D)=(H{\mathbb{A}}_{\mathbb{Q}},{\mathcal{R}}) where the relations are given by (2.5), i.e.

k:={(x,y)H𝔸(k+)×H𝔸(k+)xyImageψ(k+)},{\mathcal{R}}_{k}:=\{(x,y)\in H{\mathbb{A}}_{\mathbb{Q}}(k_{+})\times H{\mathbb{A}}_{\mathbb{Q}}(k_{+})\mid x-y\in{\rm Image}\,\psi(k_{+})\},

and where ψ\psi is as in (2.3). By construction one has ×𝒪(D)G×[ea,ea]{\mathbb{Q}}\times{\mathcal{O}}(D)\simeq G\times[-e^{a},e^{a}]. After quotienting both sides of (2.3) by G=×j(𝒪(D)f)G={\mathbb{Q}}\times j({\mathcal{O}}(D)_{f}), the map ψ\psi becomes

ϕ:HeaH(/L),ϕ(x)=x+L/Lx[ea,ea],\phi:\|H{\mathbb{R}}\|_{e^{a}}\to H({\mathbb{R}}/L),\quad\phi(x)=x+L\in{\mathbb{R}}/L\quad\forall x\in[-e^{a},e^{a}]\subset{\mathbb{R}}, (A.6)

where LL\subset{\mathbb{Q}}\subset{\mathbb{R}} is the lattice (A.5). Thus one obtains that the relation {\mathcal{R}} is equal to the inverse image by the map π:𝔸/L\pi:{\mathbb{A}}_{\mathbb{Q}}\to{\mathbb{R}}/L of the relation (A.3).
(ii)(ii) Follows from Proposition A.4 (ii)(ii).

Appendix B The Γ\Gamma-space 𝐇(D){\bf H}(D)

Let ϕ:AB\phi:A\to B be a morphism of abelian groups. To ϕ\phi one associates the following (short) complex ={Cn,ϕn}\mathfrak{C}=\{C_{n},\phi_{n}\} of abelian groups indexed in non-negative degrees

={C1ϕ1C0};C0=B,C1=A,ϕ1(x):=ϕ(x).\mathfrak{C}=\{C_{1}\stackrel{{\scriptstyle\phi_{1}}}{{\to}}C_{0}\};\qquad C_{0}=B,\ C_{1}=A,\quad\phi_{1}(x):=\phi(x). (B.1)

The Dold-Kan correspondence associates to \mathfrak{C} the simplicial abelian group (see [8] III.2, Proposition 2.2)

𝒜n=(n,k)Ck,(n,k):={σHomΔ([n],[k])σ([n])=[k]}{\mathcal{A}}_{n}=\bigoplus_{{\mathcal{F}}(n,k)}C_{k},\quad{\mathcal{F}}(n,k):=\{\sigma\in{\rm{Hom}}_{\Delta}([n],[k])\mid\sigma([n])=[k]\} (B.2)

where Δ\Delta denotes the simplicial category. The direct sum in (B.2) repeats the term CkC_{k} of the chain complex as many times as the number of elements of the set (n,k){\mathcal{F}}(n,k) of surjective morphisms σHomΔ([n],[k])\sigma\in{\rm{Hom}}_{\Delta}([n],[k]). For the short complex \mathfrak{C}, the allowed values of kk are k=0,1k=0,1. Therefore the set (n,0){\mathcal{F}}(n,0) is reduced to HomΔ([n],[0])=:Δn0{\rm{Hom}}_{\Delta}([n],[0])=:\Delta_{n}^{0}, i.e. to a single point. A morphism ξHomΔ([n],[1])=:Δn1\xi\in{\rm{Hom}}_{\Delta}([n],[1])=:\Delta^{1}_{n} is characterized by ξ1({1})\xi^{-1}(\{1\}) which is an hereditary subset of [n][n]. It follows that the vertices in Δn1\Delta^{1}_{n} are labelled by the

ξj,0jn+1,ξj(k)=1kj.\xi_{j},~0\leq j\leq n+1,\quad\xi_{j}(k)=1\iff k\geq j. (B.3)

For each integer n0n\geq 0 the finite set F(n):=(n,1)F(n):={\mathcal{F}}(n,1) of surjective elements ξΔn1\xi\in\Delta^{1}_{n} excludes ξ0\xi_{0} and ξn+1\xi_{n+1}, thus the set F(n)={ξj, 1jn}F(n)=\{\xi_{j},\,1\leq j\leq n\} has nn elements. This gives the identification

𝒜n=HB(1+)HA(F(n)+){\mathcal{A}}_{n}=HB(1_{+})\oplus HA(F(n)_{+}) (B.4)

We refer to [8] (III.2, pp. 160-161) for a detailed description of the simplicial structure, namely the definition for each θHomΔ([m],[n])\theta\in{\rm{Hom}}_{\Delta}([m],[n]) of a map of sets 𝒜(θ):𝒜n𝒜m{\mathcal{A}}(\theta):{\mathcal{A}}_{n}\to{\mathcal{A}}_{m}. Next, we introduce some notations.

We identify the opposite Δo{\Delta^{o}} 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 ΔΔo\Delta\longrightarrow{\Delta^{o}} which identifies the opposite category of Δ\Delta with Δo{\Delta^{o}} (described by intervals as above), maps the finite ordinal object [n]={0,1,,n}[n]=\{0,1,\ldots,n\} in Δ\Delta to the interval [n]={0,,n+1}[n]^{*}=\{0,\ldots,n+1\}.

We denote by 𝔖𝔢𝔱𝔰2,{\mathfrak{Sets}_{2,*}} the category of pairs of pointed sets (X,Y)(X,Y), with XYX\supset Y. The morphisms are maps of pairs of pointed sets. We let c:𝔖𝔢𝔱𝔰2,𝔖𝔢𝔱𝔰c:{\mathfrak{Sets}_{2,*}}\longrightarrow{\mathfrak{Sets}_{*}} be the functor (X,Y)X/Y(X,Y)\to X/Y of collapsing YY to the base point \ast.

Let :Δo𝔖𝔢𝔱𝔰2,\partial:{\Delta^{o}}\longrightarrow{\mathfrak{Sets}_{2,*}} be the functor that replaces an interval II with the pair I=(X,Y)\partial I=(X,Y), where XX is the set II pointed by its smallest element, and YXY\subset X is the subset formed by the smallest and largest elements of II.

Finally, we denote by γ:𝔖𝔢𝔱𝔰2,Γ𝔖𝔢𝔱𝔰2,\gamma:{\mathfrak{Sets}_{2,*}}\longrightarrow\Gamma{\mathfrak{Sets}_{2,*}} (see section A) the functor that associates to an object (X,Y)(X,Y) of 𝔖𝔢𝔱𝔰2,{\mathfrak{Sets}_{2,*}} the covariant functor

γ(X,Y):Γo𝔖𝔢𝔱𝔰2,,k+:={0,,k}(Xk+,Yk+).\gamma(X,Y):{\Gamma^{o}}\longrightarrow{\mathfrak{Sets}_{2,*}},\quad k_{+}:=\{0,\ldots,k\}\mapsto(X\wedge k_{+},Y\wedge k_{+}).

The following formula defines a covariant functor that associates to a pair of pointed sets an abelian group directly related to the morphism ϕ:AB\phi:A\to B

Hϕ:𝔖𝔢𝔱𝔰2,𝔄𝔟,Hϕ(X,Y):=HB(Y)×HA(X/Y).H_{\phi}:{\mathfrak{Sets}_{2,*}}\longrightarrow\mathfrak{Ab},\qquad H_{\phi}(X,Y):=HB(Y)\times HA(X/Y). (B.5)

On morphisms f:(X,Y)(X,Y)f:(X,Y)\to(X^{\prime},Y^{\prime}) in 𝔖𝔢𝔱𝔰2,{\mathfrak{Sets}_{2,*}} with α=(αY,αX/Y)Hϕ(X,Y)\alpha=(\alpha_{Y},\alpha_{X/Y})\in H_{\phi}(X,Y), the functor HϕH_{\phi} acts as follows

Hϕ(f)(α)\displaystyle H_{\phi}(f)(\alpha) =(αY,HA(f)(αX/Y))\displaystyle=\left(\alpha_{Y^{\prime}},HA(f)(\alpha_{X/Y})\right) (B.6)
αY(y)\displaystyle\alpha_{Y^{\prime}}(y^{\prime}) :=HB(f)(αY)(y)+xXY,f(x)=yϕ(αX/Y(x)),y.\displaystyle:=HB(f)(\alpha_{Y})(y^{\prime})+\sum_{x\in X\setminus Y,f(x)=y^{\prime}}\phi(\alpha_{X/Y}(x)),\quad y^{\prime}\neq*.

By [4] (Proposition 4.5), the Dold-Kan correspondence for the short complex (B.1), i.e. the simplicial abelian group 𝒜n\mathcal{A}_{n} in (B.4) is canonically isomorphic to the composite functor Hϕ:Δo𝔄𝔟H_{\phi}\circ\partial:{\Delta^{o}}\longrightarrow\mathfrak{Ab}, with HϕH_{\phi} as in (B.5). By composing HϕH_{\phi} with the Eilenberg-MacLane functor HH one obtains a covariant functor HHϕ=HHϕ:𝔖𝔢𝔱𝔰2,𝕊ModHH_{\phi}=H\circ H_{\phi}:{\mathfrak{Sets}_{2,*}}\longrightarrow{{{\mathbb{S}}}-{\rm{Mod}}} to the category of 𝕊{{\mathbb{S}}}-modules, which is naturally isomorphic to the functor ((UHϕ)×id)γ((U\circ H_{\phi})\times{\rm id})\circ\gamma, where U:𝔄𝔟𝔖𝔢𝔱𝔰U:\mathfrak{Ab}\longrightarrow{\mathfrak{Sets}_{*}} is the forgetful functor (see op.cit. Lemma 4.6). Moreover, as shown in op.cit. (Theorem 4.7) the Γ\Gamma-space associated by the Dold-Kan correspondence to the complex \mathfrak{C} is canonically isomorphic to the functor

((UHϕ)×id)γ:Δo𝕊Mod.((U\circ H_{\phi})\times{\rm id})\circ\gamma\circ\partial:{\Delta^{o}}\longrightarrow{{{\mathbb{S}}}-{\rm{Mod}}}. (B.7)

Notice that the above construction involves the morphism ϕ:AB\phi:A\to B only through the composite functor UHϕ:𝔖𝔢𝔱𝔰2,𝔖𝔢𝔱𝔰U\circ H_{\phi}:{\mathfrak{Sets}_{2,*}}\longrightarrow{\mathfrak{Sets}_{*}}. This latter functor is still meaningful when one restricts ϕ\phi to a sub-𝕊{{\mathbb{S}}}-module EE of the 𝕊{{\mathbb{S}}}-module HAHA and it is given by

Hϕ|E:𝔖𝔢𝔱𝔰2,𝔖𝔢𝔱𝔰,Hϕ(X,Y):=HB(Y)×E(X/Y).H_{\phi}|_{E}:{\mathfrak{Sets}_{2,*}}\longrightarrow{\mathfrak{Sets}_{*}},\qquad H_{\phi}(X,Y):=HB(Y)\times E(X/Y). (B.8)

This provides the following construction

Definition B.1.

Let ϕ:AB\phi:A\to B be a morphism of abelian groups and EE a sub 𝕊{{\mathbb{S}}}-module of the 𝕊{{\mathbb{S}}}-module HAHA. We denote by Γ(ϕ|E)\Gamma(\phi|_{E}) the Γ\Gamma-space obtained as a sub-functor of (B.7)

Γ(ϕ|E):=(Hϕ|E×id)γ:Δo𝕊Mod.\Gamma(\phi|_{E}):=(H_{\phi}|_{E}\times{\rm id})\circ\gamma\circ\partial:{\Delta^{o}}\longrightarrow{{{\mathbb{S}}}-{\rm{Mod}}}. (B.9)

When evaluated on the object 1+={0,1}1_{+}=\{0,1\} of Γo{\Gamma^{o}}, the Γ\Gamma-space Γ(ϕ|E)\Gamma(\phi|_{E}) defines a sub-simplicial set of the Kan simplicial set 𝒜n\mathcal{A}_{n} 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 {\mathcal{R}} of homotopy between nn-simplices (x,y)Xn×Xn(x,y)\in X_{n}\times X_{n} is defined as follows

(x,y)jx=jyj&zs.t.jz=sn1jxj<n,nz=x,n+1z=y.(x,y)\in{\mathcal{R}}\iff\partial_{j}x=\partial_{j}y\,\forall j\,\&\,\exists z~\text{s.t.}~\partial_{j}z=s_{n-1}\partial_{j}x\,\forall j<n,\ \partial_{n}z=x,\,\partial_{n+1}z=y. (B.10)

In general, the relation {\mathcal{R}} in (B.10) fails to be an equivalence relation. In place of the quotient we consider pairs of sets and relations. We define πn𝒯\pi^{{\mathcal{T}}}_{n} to be the set of spherical elements in XnX_{n} (i.e. of nn-simplices xx, with jx=\partial_{j}x=* j\forall j), endowed with the relation {\mathcal{R}}. Then we have the following result (we refer to section A for the notion of tolerant module)

Proposition B.1.

Let ϕ:AB\phi:A\to B be a morphism of abelian groups and let EE be a sub-𝕊[±1]{{{\mathbb{S}}}[\pm 1]}-module of the Eilenberg-MacLane 𝕊[±1]{{{\mathbb{S}}}[\pm 1]}-module HAHA (where (±1)x:=±x(\pm 1)x:=\pm x).

  1. (i)

    The homotopy π1𝒯(Γ(ϕ|E))\pi^{{\mathcal{T}}}_{1}(\Gamma(\phi|_{E})) is the sub-𝕊[±1]{{{\mathbb{S}}}[\pm 1]}-module ker(Hϕ|E)\ker(H\phi|_{E}) of EE

    ker(Hϕ|E)(k+)={xE(k+)Hϕ(x)=}.\ker(H\phi|_{E})(k_{+})=\{x\in E(k_{+})\mid H\phi(x)=*\}. (B.11)
  2. (ii)

    The homotopy π0𝒯(Γ(ϕ|E))\pi^{{\mathcal{T}}}_{0}(\Gamma(\phi|_{E})) is the tolerant 𝕊[±1]{{{\mathbb{S}}}[\pm 1]}-module HBHB endowed with the relations

    k={(x,y)HB(k+)×HB(k+)xyHϕ(E(k+))}k.{\mathcal{R}}_{k}=\{(x,y)\in HB(k_{+})\times HB(k_{+})\mid x-y\in H\phi(E(k_{+}))\}\quad k\in{\mathbb{N}}. (B.12)
Proof 15.

By construction, Γ(ϕ|E)\Gamma(\phi|_{E}) is the composite of the functors Hϕ|E:𝔖𝔢𝔱𝔰2,𝔖𝔢𝔱𝔰H_{\phi}|_{E}:{\mathfrak{Sets}_{2,*}}\longrightarrow{\mathfrak{Sets}_{*}} of (B.8) and γ:Δo𝔖𝔢𝔱𝔰2,\gamma\circ\partial:{\Delta^{o}}\longrightarrow{\mathfrak{Sets}_{2,*}}. One has [0]=(X,Y)\partial[0]^{*}=(X,Y) where X=Y={0,1}X=Y=\{0,1\} with base point 0. Thus γ[0](k+)=(Xk+,Yk+)=(k+,k+)\gamma\circ\partial[0]^{*}(k_{+})=(X\wedge k_{+},Y\wedge k_{+})=(k_{+},k_{+}) and by (B.8)

Hϕ|E(γ[0])(k+)=Hϕ|E(Xk+,Yk+)=HB(k+).H_{\phi}|_{E}(\gamma\circ\partial[0]^{*})(k_{+})=H_{\phi}|_{E}(X\wedge k_{+},Y\wedge k_{+})=HB(k_{+}).

One has [1]=(X,Y)\partial[1]^{*}=(X,Y), where X={0,1,2}X=\{0,1,2\}, Y={0,2}Y=\{0,2\} with base point 0. Thus one has γ[1](k+)=(Xk+,Yk+)\gamma\circ\partial[1]^{*}(k_{+})=(X\wedge k_{+},Y\wedge k_{+}) and Yk+=k+Y\wedge k_{+}=k_{+}, while (Xk+)/(Yk+)={0,1}k+=k+(X\wedge k_{+})/(Y\wedge k_{+})=\{0,1\}\wedge k_{+}=k_{+} and by (B.8) it follows that

Hϕ|E(γ[1])(k+)=E(k+)×HB(k+).H_{\phi}|_{E}(\gamma\circ\partial[1]^{*})(k_{+})=E(k_{+})\times HB(k_{+}). (B.13)

The boundaries j:Hϕ|E(γ[1])Hϕ|E(γ[0])\partial_{j}:H_{\phi}|_{E}(\gamma\circ\partial[1]^{*})\to H_{\phi}|_{E}(\gamma\circ\partial[0]^{*}) are obtained as in [4] Proposition 4.11

0(ψ)=ψ2,1(ψ)=Hϕ(ψ1)+ψ2,ψ=(ψ1,ψ2)Hϕ|E(γ[1]).\partial_{0}(\psi)=\psi_{2},\ \ \partial_{1}(\psi)=H\phi(\psi_{1})+\psi_{2},\quad\forall\psi=(\psi_{1},\psi_{2})\in H_{\phi}|_{E}(\gamma\circ\partial[1]^{*}). (B.14)

(i)(i) The spherical condition j(ψ)=\partial_{j}(\psi)=* on ψHϕ|E(γ[1])(k+)\psi\in H_{\phi}|_{E}(\gamma\circ\partial[1]^{*})(k_{+}) means that ψ2=0\psi_{2}=0 and Hϕ(ψ1)=0H\phi(\psi_{1})=0. Thus the solutions correspond to ker(Hϕ|E)(k+)\ker(H\phi|_{E})(k_{+}) as in (B.11). One shows as in [4] Proposition 4.11 (iii)(iii), that the relation of homotopy is the identity.
(ii)(ii) The relation k{\mathcal{R}}_{k} on elements of Hϕ|E(γ[0])(k+)H_{\phi}|_{E}(\gamma\circ\partial[0]^{*})(k_{+}) is defined as follows

(α,β)kψHϕ|E(γ[1])(k+)s.t.0(ψ)=α&1(ψ)=β.(\alpha,\beta)\in{\mathcal{R}}_{k}\iff\exists\psi\in H_{\phi}|_{E}(\gamma\circ\partial[1]^{*})(k_{+})~\text{s.t.}~\partial_{0}(\psi)=\alpha\ \&\ \partial_{1}(\psi)=\beta. (B.15)

By (B.14) it coincides with (B.12).

Note that this relation is a tolerance relation i.e. is symmetric since EE is a sub 𝕊[±1]{{{\mathbb{S}}}[\pm 1]}-module of the 𝕊[±1]{{{\mathbb{S}}}[\pm 1]}-module HAHA, so that

xyHϕ(E(k+))yxHϕ(E(k+)).x-y\in H\phi(E(k_{+}))\iff y-x\in H\phi(E(k_{+})).

The above general construction applies to the geometric adelic context and by implementing the action of ×{\mathbb{Q}}^{\times} on adeles, we obtain the following variant of Proposition 4.9 in [4].

Proposition B.2.

Let DD be an Arakelov divisor on Spec¯{\overline{{\rm Spec\,}{\mathbb{Z}}}}. Let A=×𝔸A={\mathbb{Q}}\times{\mathbb{A}}_{\mathbb{Q}}, B=𝔸B={\mathbb{A}}_{\mathbb{Q}} and α:AB\alpha:A\to B, α(q,a)=q+a\alpha(q,a)=q+a. Let E=H×H𝒪(D)E=H{\mathbb{Q}}\times H{\mathcal{O}}(D) be the sub 𝕊[±1]{{{\mathbb{S}}}[\pm 1]}-module of HAHA as in Proposition 2.1. Then the functor

𝐇(D):=Γ(α|E):Δo𝕊Mod{\bf H}(D):=\Gamma(\alpha|_{E}):{\Delta^{o}}\longrightarrow{{{\mathbb{S}}}-{\rm{Mod}}} (B.16)

defines a Γ\Gamma-space that depends only on the linear equivalence class of DD.

B.1 Proof of Proposition 2.2

With the notations of section A and Proposition B.2, one has H0(D):=π1𝒯(𝐇(D))=π1𝒯(Γ(α|E))H^{0}(D):=\pi^{{\mathcal{T}}}_{1}({\bf H}(D))=\pi^{{\mathcal{T}}}_{1}(\Gamma(\alpha|_{E})). Proposition B.1, (i)(i), gives

H0(D)(k+)={xE(k+)Hα(x)=}.H^{0}(D)(k_{+})=\{x\in E(k_{+})\mid H\alpha(x)=*\}.

An element xE(k+)x\in E(k_{+}) is a kk-tuple (qj,aj)(q_{j},a_{j}) with qjq_{j}\in{\mathbb{Q}} for all jj, and (aj)H𝒪(D)(k+)(a_{j})\in H{\mathcal{O}}(D)(k_{+}), where H𝒪(D):=H𝒪(D)f×HeaH{\mathcal{O}}(D):=H{\mathcal{O}}(D)_{f}\times\|H{\mathbb{R}}\|_{e^{a}}. The condition Hα(x)=H\alpha(x)=* means qj=ajq_{j}=-a_{j} for all jj, so that xx is uniquely determined by the kk-tuple (qj)(q_{j}). Moreover, the allowed kk-tuples are those for which |qj|ea\sum|q_{j}|\leq e^{a} and qj𝒪(D)fq_{j}\in{\mathcal{O}}(D)_{f} for all jj. One has

𝒪(D)f=H0(Spec,𝒪(D)f).{\mathbb{Q}}\cap{\mathcal{O}}(D)_{f}=H^{0}({{\rm Spec\,}{\mathbb{Z}}},{\mathcal{O}}(D)_{f}).

This proves (i)(i) of Proposition 2.2. The statement (ii)(ii) follows from Proposition B.1 (ii)(ii). Finally, (iii)(iii) follows from Proposition B.2.

{Backmatter}

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, Spec¯{\overline{{\rm Spec\,}{\mathbb{Z}}}} 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