• Aucun résultat trouvé

Liftings of diagrams of semilattices by diagrams of dimension groups

N/A
N/A
Protected

Academic year: 2021

Partager "Liftings of diagrams of semilattices by diagrams of dimension groups"

Copied!
28
0
0

Texte intégral

(1)

HAL Id: hal-00004021

https://hal.archives-ouvertes.fr/hal-00004021

Submitted on 21 Jan 2005

HAL

is a multi-disciplinary open access archive for the deposit and dissemination of sci- entific research documents, whether they are pub- lished or not. The documents may come from teaching and research institutions in France or abroad, or from public or private research centers.

L’archive ouverte pluridisciplinaire

HAL, est

destinée au dépôt et à la diffusion de documents scientifiques de niveau recherche, publiés ou non, émanant des établissements d’enseignement et de recherche français ou étrangers, des laboratoires publics ou privés.

Liftings of diagrams of semilattices by diagrams of dimension groups

Jiri Tuma, Friedrich Wehrung

To cite this version:

Jiri Tuma, Friedrich Wehrung. Liftings of diagrams of semilattices by diagrams of dimension groups.

Proceedings of the London Mathematical Society, London Mathematical Society, 2003, 87 (3), pp.1-28.

�10.1112/S0024611502013941�. �hal-00004021�

(2)

ccsd-00004021, version 1 - 21 Jan 2005

BY DIAGRAMS OF DIMENSION GROUPS

JI ˇR´I T˚UMA AND FRIEDRICH WEHRUNG

Introduction

There are various ways to obtain distributive semilattices from other mathemat- ical objects. Two of them are the following; we refer to Section 1 for more precise definitions. Adimension groupis a directed, unperforated partially ordered abelian group with the interpolation property, see also K. R. Goodearl [6]. With a dimen- sion group G we can associate its semilattice of compact ( = finitely generated) ideals IdcG. Because of the interpolation property the positive coneG+ has the refinement property, thus the compact ideal semilattice IdcGis distributive. A ring R is locally matricial over a field K, if it is isomorphic to a direct limit of finite products of full matricial ringsMn(K). IfRis locally matricial, or, more generally, regular (in von Neumann’s sense), then its semilattice IdcR of compact ideals is distributive, see, for example, K. R. Goodearl [5].

These two different contexts are related as follows. With a locally matricial ringR, we can associate its (partially ordered)Grothendieck groupK0(R). It turns out thatK0(R) is a dimension group, and the following relation holds:

IdcR∼= Idc(K0(R)), (0.1)

see [5, Corollary 15.21]. Therefore, the compact ideal semilattice of a locally ma- tricial ring is the compact ideal lattice of a dimension group. By the results of K. R. Goodearl and D. E. Handelman [7], the converse holds for dimension groups of cardinality at mostℵ1, but it fails for larger dimension groups, see F. Wehrung [17, 18]. Nevertheless this makes it worthwhile to study the functor that with a dimension group G associates its semilattice of compact ideals IdcG. For a sur- vey paper on this and related issues we refer the reader to K. R. Goodearl and F.

Wehrung [8].

The central open problem that motivates the present paper is Problem 10.1 of [8], that asks whether any distributive h∨,0i-semilattice of cardinality at most ℵ1 is isomorphic to the compact ideal semilattice of some dimension group. Let us call this the compact ideal representation problem. By [7], this is equivalent to asking whether any distributive h∨,0i-semilattice of cardinality at most ℵ1 is isomorphic to the compact ideal semilattice of some locally matricial ring. Although we do not solve this problem here, we settle related issues with some interest of their own,

2000Mathematics Subject Classification. 06A12, 06C20, 06F20, 15A03, 15A24, 15A48, 16E20, 16E50, 19A49, 19K14.

Key words and phrases. Distributive semilattice, dimension group, locally matricial algebra, compact ideal, direct limit, flat, generic.

The authors were supported by the Barrande program and by the institutional grant CEZ:J13/98:113200007a. The first author was also partially supported by GA CR 201/99 and by GA UK 162/1999.

1

(3)

and, hopefully, that will provide a stepping stone towards a solution for the general problem.

We first recall the answers to some related questions. As a corollary of the main result of G. M. Bergman [2] and a theorem of K. R. Goodearl [5], the following result is obtained, see also [8] for different proofs.

Theorem 0.1. Every countable distributive h∨,0i-semilattice S is isomorphic to IdcGfor some countable dimension group G.

Another result of [8] is the following.

Theorem 0.2. Every distributive lattice L with zero is isomorphic to IdcG for some dimension groupG.

On the other hand, P. R˚uˇziˇcka [14] constructs a distributiveh∨,0i-semilattice of cardinalityℵ2 that is not isomorphic to IdcGfor any dimension group G.

These results leave open the case of distributive h∨,0i-semilattices of cardinal- ity ℵ1. By P. Pudl´ak [13], every distributive h∨,0i-semilattice is isomorphic to the direct union of its finite distributive h∨,0i-subsemilattices. A variant of this result, stating that every distributive h∨,0i-semilattice is a direct limit of finite Booleanh∨,0i-semilattices, is proved in [8]. Thus it is natural to try an inductive approach to the compact ideal representation problem for distributive semilattices of cardinality ℵ1: find asimultaneous representation of a suitable limit system of finite Boolean semilattices that converges to a given distributive semilattice S of cardinalityℵ1. By A. P. Huhn [10], one can assume that the limit system is indexed by an infinite lattice in which every principal ideal is a finitedismantlable lattice.

After preliminary results, most notably about the existence of the so-calledλ- generic maps in Sections 4 and 5, we prove in Section 6 our main results:

Theorem 6.4. Every finite dismantlable diagram of finite Boolean h∨,0i-semi- lattices has a simultaneous representation with respect to the functor Idc.

Theorem 7.1. For any countable distributiveh∨,0i-semilattice S and any count- able dimension vector space H, every h∨,0i-homomorphism f: S →IdcH can be lifted by a positive homomorphism f:G→H for some dimension vector space G.

We also state consequences of Theorem 7.1, as well in ring theory as in lattice theory, see Corollaries 7.4 and 7.5.

The remaining sections contain various counterexamples to other strategies for a positive solution of the compact ideal representation problem for distributive semilattices of cardinalityℵ1.

1. Basic concepts

For basic concepts about partially ordered abelian groups, we refer the reader to K. R. Goodearl [6]. We recall some of the definitions here. For a partially ordered abelian groupG, we denote by G+ ={x∈G|0≤x} the positive cone of G, and we say thatG isdirected, if G=G++ (−G+). We putG++ =G+\ {0}, and we noteN=Z++. A subgroupH ofGisconvex, if 0≤x≤aimplies thatx∈H, for alla ∈H and x∈G. We say that H is anideal of G, if it is a directed, convex subgroup of G, and a compact ideal of G, if it is an ideal generated by a finite subset ofG.

(4)

As every ideal ofGis directed, any ideal (resp., compact ideal) ofGis generated (as an ideal) by a subset of G+ (resp., finite subset of G+). In the latter case, one may replace a finite subset {a1, . . . , an} of G+ by the singleton {a}, where a=Pn

i=1ai, so the compact ideals ofGare exactly the subsets ofGof the form G(a) =

x∈G| ∃n∈Z+ such that −na≤x≤na , fora∈G+. An order-unit ofGis an elementa∈G+ such thatG(a) =G. We denote by IdG the lattice of ideals ofG, and by IdcGtheh∨,0i-semilattice of compact ideals ofG.

Observe that IdGis an algebraic lattice and that IdcGis its h∨,0i-semilattice of compact elements, see G. Gr¨atzer [9] for unexplained terminology. For elements a, b ∈G+, we write a ∝b, if there exists n∈ Nsuch that a ≤nb, and a ≍b, if a∝b∝a. Hence

IdcG=

G(a)|a∈G+ ,

with containment and equality among theG(a)-s determined by G(a)⊆G(b) iffa∝b and G(a) =G(b) iffa≍b,

for all a, b ∈G+. Hence, IdcGis isomorphic to ∇(G+), themaximal semilattice quotient ofG+, see [8].

The assignmentG7→IdcGcan be naturally extended to afunctor, by defining, for a positive homomorphism (that is, an order-preserving group homomorphism) f:G→Hof partially ordered abelian groups and a compact idealIofG, (Idcf)(I) as the compact ideal of H generated by the image f[I] of I under f. Hence, (Idcf)(G(a)) =H(f(a)), for all a∈G+, and it is an easy exercise to verify that the functor Idc thus definedpreserves direct limits.

Most of the partially ordered abelian groups we shall deal with will be equipped with an additional structure of vector space, always over the field Q of rational numbers. Any partially ordered vector space E has the additional property that mx ≥ 0 implies that x ≥ 0, for all m ∈ N and all x ∈ E, we say that E is unperforated, see [6]. We observe that all group homomorphisms between vector spaces preserve the vector space structure.

A refinement monoid (see A. Tarski [15], F. Wehrung [16]) is a commutative monoidM such that for any positive integersm,nand elementsa0, . . . ,am−1,b0, . . . , bn−1 ofM such that P

i<mai =P

j<nbj, there are elements ci,j (fori < m and j < n) such that ai = P

j<nci,j, for all i < m, and bj = P

i<mci,j, for all j < n.

A partially ordered abelian groupGis an interpolation group (see [6]) if for all positive integers m, n and elements a0, . . . , am−1, b0, . . . , bn−1 of G such that ai ≤bj, for all i < m and j < n, there exists x ∈G such that ai ≤x ≤bj, for all i < mand j < n. Equivalently, the positive cone G+ is a refinement monoid (see [6, Proposition 2.1]). In addition, we say that G is a dimension group (see [6]) if G is directed and unperforated. If Gis an interpolation group, then IdcG is a refinement semilattice. Refinement semilattices are usually called distributive semilattices, see [9].

We can prove right away the following very elementary lifting result:

Proposition 1.1. For every h∨,0i-semilattice S, there exists a partially ordered vector spaceE such that IdcE∼=S.

Proof. We first observe that S can be h∨,0i-embedded into some powerset semi- lattice P(X), for a set X. For example, take X = S and use the embedding

(5)

that with anys∈ S associates j(s) ={x∈S|sx}. We then letF denote the vector space of all maps f:X → Q with finite range. For any f ∈ F, we put suppf ={x∈X |f(x)6= 0}, and we put

M ={f ∈F|f(x)≥0, for allx∈X and suppf ∈S}.

Then M is an additive submonoid of F (in particular, it is cancellative), M ∩ (−M) ={0}, andM is closed under multiplication by nonnegative scalars. Hence M is the positive cone of a structure of directed partially ordered vector space on E=M+ (−M). Letπ:M →S be the map defined by the ruleπ(f) = suppf, for allf ∈M. Thenπ(f)≤π(g) (resp., π(f) =π(g)) ifff ∝g(resp.,f ≍g) inE, for allf, g ∈M, whence π induces an isomorphismϕfrom IdcE ontoS by the rule

ϕ(E(f)) = suppf, for allf ∈M.

Unfortunately, even if S is distributive, the partially ordered vector space E constructed above may not have interpolation.

Throughout the paper we shall often formulate our results in basic categorical language. Among the main categories we shall be working with are the following:

— The category E of pairs A = (A0,1A), where A0 is a partially ordered vector space (that we will subsequently identify with A) and 1A is an order-unit ofA0. For objectsAandB ofE, ahomomorphism, orpositive homomorphism, from A to B is a positive homomorphism f:A0 → B0. Observe that we do not require that f(1A) = 1B. We say that a homo- morphism f:A →B isnormalized, iff(1A) = 1B. In particular, we will slightly abuse terminology by using isomorphisms only in the ‘normalized’

sense, that is, an isomorphismf:A→Bsatisfies the equalityf(1A) = 1B.

— The full subcategoryEdofEconsisting of alldimension vector spacesinE.

— The categorySofh∨,0i-semilattices andh∨,0i-homomorphisms.

— The full subcategorySfbofSwhose objects are the finite Boolean semilat- tices.

In particular, Idc defines, by restriction, a functor fromEtoSthat preserves direct limits.

Every partially ordered set (P,≤) can be viewed as a category with objects x ∈ P and a unique morphism εx,y: x → y whenever x ≤ y in P. Let E be a subcategory of E. We say that a diagram Φ : (P,≤) → S of semilattices (i.e., a functor fromP, viewed as a category, toS) has alifting, with respect to the functor Idc, to the category E, if there is a diagram Ψ : (P,≤) → E such that the two functors Φ and Idc◦Ψ are naturally equivalent, that is, if there are isomorphisms ιx: Φ(x)→Idc(Ψ(x)), forx∈P, such that the following diagram commutes,

Idc(Ψ(x))Idc(Ψ(εx,y))

//Idc(Ψ(y))

Φ(x) ιx

OO

Φ(εx,y) //Φ(y) ιy

OO

for all x≤y in P. We will usually omit the phrasewith respect to Idc, since lifts with respect to other functors will not be investigated.

For a setI, we denote byP(I) the powerset ofI, and, for anyi∈I, we put Pi(I) ={X ∈P(I)|i∈X}, Pi(I) ={X∈P(I)|i /∈X}.

(6)

2. Pseudo-simplicial spaces

For a nonempty, finite setX, letQX denote the partially ordered vector space with underlying vector spaceQX and partial ordering defined by

f ≤g⇔ either f =gor f(x)< g(x), for allx∈X,

endowed with the canonical order-unit 1X, defined as the constant function onX with value 1. ThenQX is a dimension vector space. It issimple, that is, IdcQX∼= 2. We denote by (1Xx)x∈X the canonical basis of QX, so that 1X = P

x∈X1Xx. We observe that while 1X ∈ (QX)++, the vectors 1Xx, for x ∈ X, do not belong to (QX)++ unless X is a singleton. Nevertheless they keep a certain positivity character, captured by the following definition:

Definition 2.1. For a partially ordered abelian groupG, we define a partial pre- ordering≤arch onGby the rule

x≤archy⇐⇒ ∃u∈Gsuch thatnx≤ny+u, for alln∈Z+, (2.1) for all x, y ∈ G. Then we define the partially preordered abelian group Garch = (G,≤arch), thearchimedean quotient ofG.

In particular, for any finite setX, the archimedean quotient of QX isQX, and 0<arch1Xx, for allx∈X.

Definition 2.2. An objectAofE is

(i) simple pseudo-simplicial, if A∼=QX for some nonempty finite setX, (ii) pseudo-simplicial, ifA =L

i<nAi for some natural number n and some simple pseudo-simplicial spacesAi, fori < n.

The formula in (ii) above means that A = L

i<nAi as vector spaces, 1A = P

i<n1Ai, and the canonical bijection fromQ

i<nAi to Ais an isomorphism in E (that is, for the vector space structure as well as the—componentwise—partial order structure).

In particular, simplicial vector spaces, that is, spaces of the form Qn with the positive cone (Qn)+ consisting of all vectors with non-negative coordinates, are particular cases of pseudo-simplicial spaces. Observe that for any pseudo-simplicial spaceE, the spaceEarch is simplicial.

For a simple pseudo-simplicial spaceA, thecanonical basis ofAis defined as the setT of minimal elementstofA++arch such thatt∧(1A−t) = 0, where∧denotes the meet operation in the lattice-ordered groupAarch. HenceT is a (finite) vector space basis ofA and the map from QT to Athat with any sequence (xt)t∈T associates P

t∈Txtt is an isomorphism fromQT ontoA. It sends 1Tt to t, for allt∈T, and 1T to 1A.

We denote by Eps the full subcategory of E whose objects are the pseudo-sim- plicial spaces.

We state without proof the following lemma, that summarizes some elementary properties of pseudo-simplicial spaces.

Lemma 2.3. Letmbe a natural number, letA0, . . . ,Am−1 be simple pseudo-sim- plicial spaces. Put A =L

i<mAi, and identify Ai with its canonical image in A, for alli < m. Then the following assertions hold.

(i) The ideals ofA are exactly the subsets of the formL

i∈IAi for a subsetI of {0,1, . . . , m−1}.

(7)

(ii) The simple ideals ofA are exactly theAi, for i < m.

(iii) IdcA∼=2m, a finite Boolean semilattice.

For a partially ordered vector spaceE, the two binary relations∝and≍onE+ can be refined as follows. For a positive, rational number λ and a, b ∈ E+, we introduce the following notations:

a∝λb, ifa≤λb;

a≍λb, ifa∝λb andb∝λa;

a∝archλ b, ifa∝λ+εb, for all rationalε >0;

a≍archλ b, ifa≍λ+εb, for all rationalε >0.

Hencea∝archλ b(resp.,a≍archλ b) inE implies thata∝λb(resp.,a≍λb) inEarch; the converse holds ifE issimple.

3. Basic facts about refinement We start with an easy and useful lemma.

Lemma 3.1. LetM be a refinement monoid, letIandTi,i∈I, be finite nonempty sets, let a,ai,j (for i∈I andj ∈Ti)be elements of M such that

a= X

j∈Ti

ai,j, for alli∈I.

Then there are elementsxϕ (for ϕ∈T =Q

i∈ITi)ofM such that a=X

ϕ∈T

xϕ and ai,j = X

ϕ∈T ϕ(i)=j

xϕ, for all i∈I and j∈Ti. (3.1)

Proof. An easy induction on the cardinality of I. There is nothing to prove if

|I| = 1. Now assume that the claim holds for any setI of cardinalityn ≥1 and I =I∪ {k} 6=I. By the induction hypothesis on I we get elements xϕ ∈M for ϕ∈T such that

a= X

ϕ∈T

xϕ and ai,j = X

ϕ∈T ϕ(i)=j

xϕ, for all i∈I and j ∈Ti.

Since also a = P

l∈Tkak,l and M is a refinement monoid, we get that there are elementsxϕ,l∈M, forϕ∈T andl∈Tk, such that

xϕ= X

l∈Tk

xϕ,l and ak,l= X

ϕ∈T

xϕ,l

for anyϕ∈T andl∈Tk. Thus a= X

l∈Tk

ak,l=X

l∈Tk

X

ϕ∈T

xϕ,l.

SinceT×Tk=Q

i∈I∪{k}Ti, the indices (ϕ, l) are in one-to-one correspondence with functions ψ∈T×Tk. Moreover, for everyi∈I andj∈Ti we get

ai,j = X

ϕ∈T ϕ(i)=j

xϕ= X

ϕ∈T ϕ(i)=j

X

l∈Tk

xϕ,l= X

ψ∈T×Tk ψ(i)=j

xψ.

(8)

Finally, for everyl∈Tk we get ak,l=X

ϕ∈T

xϕ,l= X

ψ∈T×Tk ψ(k)=l

xψ.

Lemma 3.2. Let λ≥1 be a rational number, letE be a dimension vector space, let I be a finite set, letai (for i∈ I) be elements of E+. Then the following are equivalent:

(i) aiλaj, for alli,j ∈I.

(ii) There are elementsbX (for X ∈P(I))of E+ such that ai= X

X∈P i(I)

bX+λ· X

X∈Pi(I)

bX, for all i∈I.

Proof. (ii)⇒(i) is trivial.

Now assume that (i) holds. By applying interpolation to the system of inequal- itiesλ−1ai ≤aj, for alli,j∈I, we obtaina∈Esuch that λ−1ai≤a≤aj, for all i,j ∈I. Hence,a≤ai≤λa, for all i∈I, so that there areai,a′′i ∈E+ such that ai=a+aiand (λ−1)a=ai+a′′i, for alli∈I. By Lemma 3.1 applied toM =E+ andJ ={0,1}, there are elements bX ofE+, forX ∈P(I), such that

ai= (λ−1) X

X∈Pi(I)

bX, a′′i = (λ−1) X

X∈Pi(I)

bX, for alli∈I;

a= X

X∈P(I)

bX.

Therefore, for alli∈I,

ai= X

X∈P(I)

bX+ (λ−1) X

X∈Pi(I)

bX

= X

X∈P i(I)

bX+λ· X

X∈Pi(I)

bX.

4. Flat and generic homomorphisms with simple target

Definition 4.1. Let m be a natural number, let A0, . . . , Am−1, B be simple objects ofEd, letf: L

i<mAi →B be a positive homomorphism, let λ≥1 be a rational number. We say thatf is

(i) λ-flat, iff(1Ai)≍archλ f(1Aj), for alli,j < msuch thatf(1Ai),f(1Aj)6= 0.

(ii) λ-generic, if for any simple pseudo-simplicial space C and for any λ-flat homomorphismg: L

i<mAi→C such that

f[Ai]6={0} ⇔g[Ai]6={0}, for alli < m, there exists a homomorphismh:B →Csuch thatg=h◦f.

The following simple result summarizes some of the basic properties of flat and generic homomorphisms:

Lemma 4.2. Let m be a natural number, let A be a pseudo-simplicial space, let A0, . . . ,Am−1,B,Gbe simpleobjects ofEd. Letα,β,λbe rational numbers such that 1≤α≤β and1≤λ.

(i) Every homomorphism f: L

i<mAi →B isλ-flat for someλ≥1.

(9)

(ii) Every α-flat homomorphism from a pseudo-simplicial space to a simple pseudo-simplicial space is β-flat.

(iii) Everyβ-generic homomorphism from a pseudo-simplicial space to a simple pseudo-simplicial space is α-generic.

(iv) Suppose that G is pseudo-simplicial, let f: A → G and h: A → B be homomorphisms, and letg: IdcG→IdcBbe ah∨,0i-homomorphism such that Idch=g◦Idcf. Iff isλ-generic andhisλ-flat, then there exists a homomorphism g:G→B such that g= Idcg andh=g◦f.

The situation of (iv) above may be summarized by the following commutative diagrams:

IdcB B

IdcG gvvvvvv;;

vv

v G

g ??







IdcA Idcf

ccHHHHHHHHH

Idch

OO

A f

__??

????

??

h

OO

Proof. To prove (i) observe that if f(1Ai) 6= 0, then it is an order-unit in B.

Thus for anyi,j < msuch that bothf(1Ai) andf(1Aj) are nonzero, there exists λi,j∈Q++such thatf(1Ai)≤λi,jf(1Aj). Now setλany positive rational number larger than 1 and allλi,j.

Claims (ii) and (iii) are straightforward.

To prove (iv) observe first thatg has a lifting, that is, a homomorphismg:G→ B such that Idcg = g. Indeed, if g = 0, take g = 0, while if g is the unique isomorphism from IdcG to IdcB, any nonzero positive homomorphism from G toB is a lifting ofg, for example, if G=Qmand ifb∈B++, the mapg:G→B defined by the ruleg(x0, . . . , xm−1) = P

i<mxi

b, for allx0, . . . ,xm−1∈Q.

Now, for (iv) above, if h = 0, then any lifting g of g satisfies the required conditions. So suppose that h 6= 0. Then, by the λ-genericity of f and the λ- flatness of h, there exists a homomorphism g:G → B such that h = g◦f. In particular,g6= 0. Similarly, from Idch=g◦Idcf and Idch6= 0 follows thatg6= 0.

Therefore, since bothGand B are simple,g lifts g.

We shall now define the canonical generic maps. We are given a pseudo-simplic- ial spaceAand ah∨,0i-homomorphismf: IdcA→2. We shall construct a simple pseudo-simplicial space G= Gen(A,f) and a liftingf:A→ Goff with respect to the Idc functor.

PutA=L

i<mAi, for a natural numbermand simple pseudo-simplicial spaces A0, . . . ,Am−1. We shall define a pseudo-simplicial spaceG= Gen(A,f) and, for every rational number λ ≥ 1, a lifting fλ: A → G of f with respect to the Idc

functor.

For all i < m, let Ti denote the canonical basis of Ai (see Section 2). Hence the Ti, for i < m, are mutually disjoint finite sets and S

i<mTi is a vector space basis ofA. Let I={i < m|f(Ai) = 1}(observe thatAi is a compact ideal of A, thus it belongs to the domain off, see Lemma 2.3). Furthermore, we put

T =Y

i∈I

Ti, F =P(I)×T, and G=QF.

(10)

Observe that bothT andF are finite, nonempty sets. For a rational numberλ≥1, we define a linear mapfλ:A→QF by its action on the elements ofS

i<mTi. For i < mandt∈Ti, we definefλ(t) = 0 ifi /∈I, while, ifi∈I, we put

fλ(t) = X

(X,ϕ)∈F i /∈X, ϕ(i)=t

1F(X,ϕ)+λ· X

(X,ϕ)∈F i∈X, ϕ(i)=t

1F(X,ϕ). (4.1)

For alli < m, letfi,λdenote the restriction offλ toAi.

Lemma 4.3. The map fi,λ is a positive homomorphism from Ai to QF, for all i < m, and it is nonzero iff i∈I.

Proof. Fori /∈I, fi,λ = 0. Suppose now that i∈ I. It is trivial that fi,λ defines a positive homomorphism from (Ai)arch ∼= QTi to QF. To prove that it is also a nonzero positive homomorphism from Ai to QF, it is sufficient to prove that fi,λ(1Ai) is an order-unit ofQF. By observing that 1Ai =P

t∈Tit and by using (4.1), we can compute:

fi,λ(1Ai) = X

(X,ϕ)∈F i /∈X

1F(X,ϕ)+λ· X

(X,ϕ)∈F i∈X

1F(X,ϕ).

Since all the components of this vector relatively to all 1F(X,ϕ), for (X, ϕ)∈F, are positive,fi,λ(1Ti) is indeed an order-unit ofQF. It follows from Lemma 4.3 thatfλis a positive homomorphism fromAtoQF. It follows immediately from Lemma 4.3 thatfλ:A→QF is a lifting of the semilattice mapf: IdcA→2. Indeed, if we take the canonical isomorphism from IdcQF to2, the following diagram

2 can. //IdcQF

IdcA f

OO

Idcfλ

::u

uu uu uu uu

is commutative, sincefλ[Ai]6={0}ifff(Ai) = 1, for alli < m.

Lemma 4.4. The homomorphismfλ isλ-flat.

Proof. For all i ∈ I, fλ(1Ai) = fi,λ(1Ai) is a linear combination of the vectors 1F(X,ϕ), for (X, ϕ)∈ F, with coefficients either 1 or λ(see (4.1)). The conclusion

follows.

We observe here the relevance of the definition of flatness by using≍archinstead of≍: indeed, the vectors 1F(X,ϕ), for (X, ϕ)∈F, do not belong to the positive cone ofQF except ifF is a singleton.

Now we come to the main result of this section:

Lemma 4.5. The homomorphismfλ isλ-generic.

Proof. LetC be a simple pseudo-simplicial space, letg: L

i<mAi→Cbe aλ-flat homomorphism such thatg[Ai]6={0}ifffλ[Ai]6={0}(that is,i∈I), for alli < m.

Set againAi =QTi for every i < m. We define elements ai (for i < m) anda(t)

(fori < mandt∈Ti) ofC by

ai=g(1Ai), a(t)=g(t).

(11)

Observe that ai ∈ C+ while we can only say that a(t) ∈ Carch+ . Moreover, ai = P

t∈Tia(t). From the fact that g is λ-flat follows that aiarchλ aj in C, for all i, j∈I, hence, by Lemma 3.2 applied toCarch (that has interpolation, becauseC is pseudo-simplicial), there are elementsbX (forX ∈P(I)) ofCarch+ such that

ai= X

X∈P i(I)

bX+λ· X

X∈Pi(I)

bX, for alli∈I. (4.2)

Now, for fixedi∈I, we apply refinement to the equality X

t∈Ti

a(t)= X

X∈P i(I)

bX+λ· X

X∈Pi(I)

bX,

which holds inCarch+ . We obtain elementscX,i,t(for (X, i)∈P(I)×I and t∈Ti) ofCarch+ such that

a(t)= X

X∈P i(I)

cX,i,t+λ· X

X∈Pi(I)

cX,i,t (for all i∈I andt∈Ti) (4.3) bX =X

t∈Ti

cX,i,t (for all X ∈P(I)). (4.4)

Now we apply Lemma 3.1 to the system of equations (4.4) in Carch+ . We find elementsd(X,ϕ)(for (X, ϕ)∈F) ofCarch+ such that

bX= X

ϕ∈T

d(X,ϕ) (for allX∈P(I)) (4.5)

cX,i,t= X

ϕ∈T ϕ(i)=t

d(X,ϕ) (for allX∈P(I), i∈I, t∈Ti). (4.6)

We define a linear maph: QF →C by the rule

h(1F(X,ϕ)) =d(X,ϕ), for all (X, ϕ)∈F. (4.7) Hencehis a positive homomorphism fromQF toCarch. To verify thathis a positive homomorphism fromQF toC, it suffices to verify thath(1F)∈C+. This is trivial forI=∅(thenQF =Q). Suppose thatI6=∅. We compute:

h(1F) = X

(X,ϕ)∈F

h(1F(X,ϕ)) = X

(X,ϕ)∈F

d(X,ϕ)

= X

X∈P(I)

bX (by (4.5)).

Hence, it follows from (4.2) that h(1F)≍ai in Carch for anyi∈I. But ai∈C+, thush(1F)∈C+.

It remains to prove that h◦fλ(t) =g(t), for all i < mand allt ∈Ti. This is obvious ifi /∈I, in which case both sides of the equality are zero. So suppose that

(12)

ibelongs toI. We compute:

h◦fλ(t) = X

(X,ϕ)∈F i /∈X, ϕ(i)=t

d(X,ϕ)+λ· X

(X,ϕ)∈F i∈X, ϕ(i)=t

d(X,ϕ) (by (4.1) and (4.7))

= X

X∈Pi(I)

cX,i,t+λ· X

X∈Pi(I)

cX,i,t (by (4.6))

=a(t) (by (4.3))

=g(t) (by the definition ofa(t)),

which concludes the proof.

5. Flat and generic homomorphisms with arbitrary target Definition 5.1. Let n be a natural number, let B0, . . . , Bn−1 be simple pseu- do-simplicial spaces, letAbe a pseudo-simplicial space, let f:A→L

j<nBj be a homomorphism. Write

f(x) = (f0(x), . . . , fn−1(x)), for allx∈A,

withfj:A→Bj, for allj < n. For any rational numberλ≥1, we say thatf is (i) λ-flat, iff0, . . . ,fn−1 areλ-flat.

(ii) λ-generic, iff0, . . . ,fn−1 areλ-generic.

Because of Lemma 2.3, the decompositionB =L

j<nBiofB into simple ideals is unique, hence the definition above is well-stated.

Part of Lemma 4.2 has a straightforward analogue for these more general defi- nitions of flat and generic maps:

Lemma 5.2. Let αandβ be rational numbers such that1≤α≤β.

(i) Every homomorphism f:A→L

j<nBj isλ-flat for someλ≥1.

(ii) Every α-flat homomorphism isβ-flat.

(iii) Every β-generic homomorphism is α-generic.

The definition of canonical generic homomorphisms of Section 4 extends easily to this new context, as follows.

LetAbe a pseudo-simplicial space, letnbe a natural number, letf: IdcA→2n be ah∨,0i-homomorphism. So there are uniqueh∨,0i-homomorphismsfj: IdcA→ 2, forj < n, such that

f(X) = (f0(X), . . . ,fn−1(X)), for allX∈IdcA.

We put Gj = Gen(A,fj), for all j < n, and G = L

j<nGj. For any rational number λ≥ 1, we define the canonical λ-generic homomorphism fλ:A →G by the rule

fλ(x) = (f0,λ(x), . . . , fn−1,λ(x)), for allx∈A,

wherefj,λ:A→Gj is the canonicalλ-generic map associated withA andfj, for all j < n. It is trivial, by Lemmas 4.4 and 4.5 and by Definition 5.1, that fλ is bothλ-flat andλ-generic.

Now define a mapι:2n→IdcGby

ι(i0, i1, . . . , in−1) =M

j∈J

Gj, (5.1)

(13)

where we putJ={j < n|ij = 1}. Since the diagram 2 can. //IdcGj

IdcA fj

OO

Idcfj,λ

::v

vv vv vv vv

is commutative for anyj < n, we get that also the diagram 2n ι //IdcG

IdcA f

OO

Idcfλ

;;v

vv vv vv vv

is commutative. Thus we have proved the following lemma.

Lemma 5.3. The identityι◦f = Idcfλ holds. Thus the mapfλ liftsf. The main result of this section is the following analogue of Lemma 4.2(iv).

Lemma 5.4. LetA,B,Gbe pseudo-simplicial spaces, letf:A→Gandh:A→B be positive homomorphisms, and let g: IdcG →IdcB be a h∨,0i-homomorphism such that Idch=g◦Idcf. Letm, pbe the natural numbers such thatIdcA∼=2m andIdcG∼=2p, and put q= min

2m−1, p if m,p >0, andq= 1otherwise. Let λ≥1 be a rational number. Iff isqλ-generic and hisλ-flat, then there exists a homomorphism g:G→B such thatg= Idcg andh=g◦f.

Proof. First of all, by decomposing B as the direct sum of its simple ideals (see Lemma 2.3), it is easy to reduce the problem to the case whereB issimple. Fur- thermore, the problem is trivial if either A or G is trivial, so suppose from now on that bothm and pare nonzero. Let A=L

i<mAi and G=L

j<pGj be the decompositions ofA andGas the direct sums of their simple ideals. We write

f(x) = (f0(x), . . . , fp−1(x)), for allx∈A,

with homomorphisms fj:A → Gj, for j < p. Moreover, by using the natural identification of IdcGwithQ

j<pIdcGj, we can decomposegas g(Y0, . . . , Yp−1) = _

j<p

gj(Yj), for all (Y0, . . . , Yp−1)∈Y

j<p

IdcGj, withh∨,0i-homomorphismsgj: IdcGj→IdcB, forj < p. Hence Idch=W

j<phj, where we puthj=gj◦Idcfj, for allj < p. Sohj is ah∨,0i-homomorphism from IdcAto IdcB.

We define subsetsH,Hj (forj < p) ofmas follows:

H={i < m|h[Ai]6={0}}, Hj={i < m|hj(Ai)6={0}}. Put H = {Hj |j < p}. Hence H = S

j<pHj = SH. Furthermore, put ni =

| {K∈H|i∈K} |, for alli < m. We make the following obvious observation:

ni= 0, for alli /∈H, and 1≤ni≤q, for alli∈H. (5.2) For any K ∈ H, we define a homomorphism h(K):A → B by its restriction to eachAi, fori < m. First, we define the restriction ofh(K)toAias equal to zero if

(14)

i /∈K. Ifi∈K, we put

h(K)(x) = 1 ni

h(x), for allx∈Ai. (5.3) Hence,h(K)[Ai]6={0} iffi∈K, hence

Idch(Hj)=hj, for allj < p. (5.4) Furthermore, for all i∈H and allx∈Ai,

X

K∈H

h(K)(x) = X

K∈H i∈K

1 ni

h(x) =h(x),

whence we obtain the equality X

K∈H

h(K)=h. (5.5)

Next, we putpj=| {j< p|Hj =Hj} |andhj= 1 pj

h(Hj), for allj < p. Hencepj

is a positive integer and, by (5.4), the following holds:

Idchj=hj, for allj < p. (5.6) Moreover,P

j<phj=P

K∈Hh(K)=h(the last equality follows from (5.5)). So we have obtained

X

j<p

hj=h. (5.7)

Now we prove the following crucial claim:

Claim 1. The homomorphism hj isqλ-flat, for allj < p.

Proof of Claim. By the definition ofhj, it suffices to prove thath(K)isqλ-flat, for allK∈H. For all i∈K,h(K)(1Ai) = 1

nih(1Ai), so, for alli,j∈K, h(K)(1Aj) = 1

nj

h(1Aj)≤arch λ nj

h(1Ai) (because hisλ-flat)

=niλ nj

h(K)(1Ai)

≤niλh(K)(1Ai)

≤qλh(K)(1Ai) (by (5.2)),

which concludes the proof of the claim. Claim 1.

Sincef isqλ-generic, it follows from Lemma 4.2 that for allj < p, there exists a homomorphismgj:Gj →B such that Idcgj =gj and gj◦fj =hj. We define g:G→Bby the rule

g(y0, . . . , yp−1) =X

j<p

gj(yj), for all (y0, . . . , yp−1)∈B.

Then Idcg=g and, by (5.7),g◦f =h.

(15)

Example 5.5. The following example shows that one cannot replaceqby 1 in the statement of Lemma 5.4. Let h: Q2 →Q be defined by the ruleh(x, y) =x+y (for all x, y ∈ Q), and let f: Q2 → A⊕B be any 1-generic lifting of the map f:22 →22 defined by the rulef(x,y) = (x∨y,y), withA and B being simple pseudo-simplicial spaces. Let g:22 → 2be defined by the rule g(x,y) = x∨y.

By identifying Idc(A⊕B) with22 and IdcQwith 2, the mapg defines a h∨,0i- homomorphism from Idc(A⊕B) to IdcQ, and Idch =g◦Idcf. Obviously,h is 1-flat. However, we prove that there exists no lifting g of g such that h=g◦f. The situation is summarized on the diagrams below:

IdcQ Q

Idc(A⊕B)

gooooooo77 oo

oo

A⊕B

g u::

u u u u

Idc(Q⊕Q) Idcf

ggOOO

OOOOOOOO

Idch

OO

Q⊕Q f

ddIIIIIIIIII

h

OO

There area,c∈A++ andb∈B++ such that

f(x, y) = (xa+yc, yb), for allx, y∈Q.

From the fact thatf is 1-flat follows thata≍arch1 c in A. Furthermore, there are positive, nonzero homomorphismsu:A→Qandv: B→Qsuch that

g(x, y) =u(x) +v(y), for allx∈Aandy∈B.

From a ≍arch1 c follows that u(a) ≍arch1 u(c), thus, sinceu(a),u(c) ∈Q+, u(a) = u(c). However, by applying h= g◦f to (1,0) and (0,1) respectively, we obtain thatu(a) = 1 andu(c) +v(b) = 1, hence 0< u(c)< u(a) =u(c), a contradiction.

We end this section with an easy lifting result:

Lemma 5.6. Let E be a pseudo-simplicial space, let m be a natural number, let f:2m→IdcE be ah∨,0i-homomorphism. Then there exists a positive homomor- phism f:Qm → E such that, if ι: 2m → IdcQm denotes the canonical isomor- phism, the equality f = Idcf◦ι holds.

Proof. Let E = L

j<nEj be the decomposition of E as a direct sum of simple pseudo-simplicial spaces. Denote by (ei)i<m (resp., (ei)i<m) the canonical basis of 2m (resp., Qm). For all i < m, there exists a subset Ji of {0,1, . . . , n−1}

such that f(ei) = L

j∈JiEj. Let f: Qm → E be the linear map defined by f(ei) =P

j∈Ji1Ej, for alli < m. Then, for alli < m, (Idcf ◦ι)(ei) = (Idcf)(Qei) =E(f(ei)) =M

j∈Ji

Ei=f(ei),

whence Idcf◦ι=f.

6. Pseudo-simplicial liftings of dismantlable diagrams

The existence of λ-generic mappings enables us to construct liftings of a large class of finite diagrams with respect to the functor Idc. The following definition is due to K. A. Baker, P. C. Fishburn, and F. S. Roberts [1].

(16)

Definition 6.1. A finite latticeLof cardinalitynis calleddismantlable if there is a chain L1 ⊂L2 ⊂ · · · ⊂Ln =L of sublattices ofL such that |Li| =i for every i= 1, 2, . . . ,n.

We shall use an extension of Definition 6.1 to partially ordered sets. We shall consider∅as a partially ordered set. For a finite partially ordered setP, an element xofP isdoubly-irreducible, ifxhas at most one upper cover and at most one lower cover. We denote by IrrP the set of all doubly-irreducible elements of P. For a subsetQofP, letQ⋖P be the statement thatQ=P\ {x} for somex∈IrrP. Definition 6.2. A finite partially ordered setP of cardinalitynisdismantlable, if there exists a chain∅=P0⋖P1⋖· · ·⋖Pn =P of subsets ofP.

Of course, a finite lattice is dismantlable iff it is dismantlable as a partially ordered set. On the other hand, every finite bounded dismantlable partially ordered set is a lattice as observed in D. Kelly and I. Rival [12]. Furthermore, a nonempty partially ordered set P is dismantlable iff there existsx∈IrrP such thatP \ {x}

is dismantlable.

There is a large supply of finite dismantlable lattices. For example, the following result was proved in [1]; another proof can be found in D. Kelly and I. Rival [12].

Theorem 6.3. Every finite planar lattice is dismantlable.

There are also non-planar modular dismantlable lattices, like the one diagrammed on Figure 1. On the other hand, every distributive dismantlable lattice is planar, see D. Kelly and I. Rival [11].

Figure 1. A dismantlable, modular, non-planar lattice The main result of this section is the following.

Theorem 6.4. Every diagramΦ :P→Sfbof finite Boolean semilattices indexed by a finite dismantlable partially ordered setP has a pseudo-simplicial liftingΨ :P → Eps.

Proof. We shall proceed by induction on the cardinality ofP. We may also assume without loss of generality that for everyx∈P we have Φ(x) =2q for someq∈Z+. IfP =∅then the result is trivial.

Now assume thatP is nonempty. So there existsx∈IrrPsuch thatQ=P\ {x}

is dismantlable. By the induction hypothesis, the restriction Φof Φ toQhas a pseu- do-simplicial lifting Ψ:Q→Eps. So there are isomorphismsιt: Φ(t)→Idc(t)),

(17)

fort∈Q, such that the following holds:

ιt◦Φ(εs,t) = Idcs,t))◦ιs, for alls≤tin Q. (6.1) Letn∈Z+ such that Φ(x) =2n.

Suppose first thatxis comparable to no element ofQ. Then, in order to lift Φ, it is sufficient to adjoin one lifting of Φ(x) to the diagram Ψ; take Ψ(x) =Qn.

So, from now on, suppose thatxcan be compared to at least one element ofQ.

We denote byu(resp.,v) the unique lower cover (resp., upper cover) ofxin P if it exists. By assumption, eitheruorv exists.

We start with the case wherexis a minimal element of P. Thenv exists, and, by Lemma 5.6, there exists a positive homomorphismf:Qn→Ψ(v) such that, if ιx: 2n →IdcQn denotes the canonical isomorphism, the equality

ιv◦Φ(εx,v) = Idcf◦ιx (6.2) holds. We define Ψ(x) =Qn, and, for everyt∈Qsuch thatx < t, that is,v≤t, we put Ψ(εx,t) = Ψv,t)◦f. It follows from (6.1) that the right half of the following diagram commutes,

Idc(Ψ(x)) = Idc(Qn) Idcf

//Idc(v))Idcv,t))

//Idc(t))

Φ(x) =2n ιx

OO

Φ(εx,v) //Φ(v) ιv

OO

Φ(εv,t) //Φ(t) ιt

OO

while the commutation of the left half of this diagram follows from (6.2). This settles the case wherexis a minimal element ofP.

So, suppose, from now on, that x is not minimal in P, so that u exists. Let m ∈ Z+ such that Φ(u) = 2m. Observe that f = Φ(εu,x)◦ι−1u is a h∨,0i-ho- momorphism from Idc(u)) to 2n. Put G = Gen(Ψ(u),f) (see Section 5), let ιx: 2n →IdcGbe the isomorphism given by (5.1). For any rational numberµ≥1, denote byfµ: Ψ(u)→Gthe canonicalµ-generic lifting off, see Section 5. Hence, by Lemma 5.3, the following diagram commutes,

2n ιx

//IdcG

Idc(u)) f

OO

Idcfµ

99s

ss ss ss ss s

which means that the following equality holds:

Idcfµ◦ιux◦Φ(εu,x). (6.3) In case x is maximal in P, set µ = 1. In case x is not maximal in P, there exists, by Lemma 5.2(i) a rational numberλ≥1 such that Ψu,v) is λ-flat. Put q= min

2m−1, n if m, n >0,q= 1 otherwise, and put µ=qλ. In both cases, we define Ψ(x) =Gand Ψ(εu,x) =fµ, the canonicalµ-generic lifting off. Hence for everyt ≤uwe have to set Ψ(εt,x) =fµ◦Ψt,u), and we need to check that

(18)

the following diagram commutes:

Idc(t))Idct,u))

//Idc(u)) Idcfµ

//Idc(Ψ(x)) = IdcG

Φ(t) ιt

OO

Φ(εt,u) //Φ(u) ιu

OO

Φ(εu,x) //Φ(x) ιx

OO

The commutativity of the left half of this diagram follows from (6.1), while the commutativity of the right half follows from (6.3). In particular, this completes the proof of the induction step ifxis the largest element ofP.

Ifxhas a (unique) upper cover v we continue as follows. We first observe that g =ιv◦Φ(εx,v)◦ι−1x is a h∨,0i-homomorphism from IdcG to Idc(v)). Since Ψu,v) is λ-flat and Ψu,x) =fµ isqλ-generic, Lemma 5.4 can be applied to the two following commutative diagrams:

Ψ(v) IdcΨ(v)

Ψ(x) =G IdcG

ee g

KKKKKK KKKK

Ψ(u) Ψu,v)

OO

fµ

99s

ss ss ss ss

Idc(u)) Idcu,v))

OO

ιx◦f

99s

ss ss ss ss s

Thus there exists a positive homomorphismg:G→Ψ(v) such that Idcg=g and g◦fµ = Ψu,v). We define Ψ(εx,v) = g, thus for every t ≥ v we have to set Ψ(εx,t) = Ψv,t)◦g, and we need to check that the following diagram commutes:

Idc(u)) Idcfµ

//IdcG g= Idcg

//Idc(v))Idcv,t))

//Idc(t))

Φ(u) ιu

OO

Φ(εu,x) //Φ(x) ιx

OO

Φ(εx,v) //Φ(v) ιv

OO

Φ(εv,t) //Φ(t) ιt

OO

The commutativity of the right third of this diagram follows from (6.1), the com- mutativity of the left third follows from (6.3), and the commutativity of the middle third of this diagram follows from the fact that Idcg =g and the definition ofg.

This completes the verification that the extension Ψ of Ψ thus defined is as re-

quired.

7. Liftingh∨,0i-homomorphisms between countable h∨,0i-semilattices The main result of this section is a wide generalization of Lemma 5.6:

Theorem 7.1. LetSbe a countable distributiveh∨,0i-semilattice, letH be a count- able dimension vector space, let f:S → IdcH be a h∨,0i-homomorphism. Then f can be lifted, that is, there are a countable dimension vector space G, a pos- itive homomorphism f:G → H, and an isomorphism α:S → IdcG such that f = (Idcf)◦α.

(19)

Proof. It follows from E. G. Effros, D. E. Handelman, and C.-L. Shen [3] (see also [6, Theorem 3.19]) that H is the direct limit of a sequence of simplicial (partially ordered) groups, that is, groups of the formZn, for n∈Z+. Since H is a vector space, the isomorphismH ∼=H⊗Q implies thatH is, in fact, the direct limit of a sequence of simplicial vector spaces: more specifically, H = lim−→i∈Z+Hi, where Hi = Qni for some ni ∈ Z+, with transition maps ti:Hi → Hi+1 and limiting mapsti:Hi→H, both being positive homomorphisms, for alli∈Z+.

Furthermore, it follows from [8, Corollary 6.7(i)] thatS is the direct limit of a countable sequence of finite Booleanh∨,0i-semilattices, say,S= lim−→i∈Z+2mi, with mi ∈Z+, transition maps si: 2mi →2mi+1 and limiting mapssi:2mi →S, both being h∨,0i-homomorphisms, for all i ∈Z+. The situation may be visualized by the following commutative diagrams:

2mi+1 si+1

!!D

DD DD DD

DD Hi+1

ti+1

!!D

DD DD DD D

2mi si

OO

si //S Hi

ti

OO

ti //H

For all i ∈ Z+, the h∨,0i-homomorphismf ◦si:2mi → IdcH ∼= lim−→j∈Z+IdcHj

factors through IdcHj for somej∈Z+. Hence, by possibly replacing the sequence (Hj)j∈Z+ by some subsequence, we can suppose that j =i, allowing to factor f through ah∨,0i-homomorphismfi:2mi →IdcHi. The situation may be visualized by the following commutative diagram, where the two parallel horizontal sequences are direct limits:

IdcH0

Idct0

//

Idct0

IdcH1

Idct1

//

Idct1

$$_

_

_ //IdcH

2m0 f0

OO

s0 //

s0

??2m1 s1 //

f1

OO

s1

;;_

_

_ //S

f

OO (7.1)

Now we construct, fori∈Z+, a certain pseudo-simplicial spaceGi, together with an isomorphismαi: 2mi→IdcGi and a positive homomorphismfi:Gi→Hisuch that the following equality holds:

fi= (Idcfi)◦αi. (7.2)

For i = 0, we pick G0, f0, and α0 satisfying (7.2); their existence is ensured by Lemma 5.6. Now suppose Gi, fi, and αi constructed satisfying (7.2) above, we construct Gi+1, fi+1, and αi+1. First, by Lemma 5.2(i), there exists a rational number λ ≥ 1 such that the homomorphism ti◦fi: Gi → Hi+1 is λ-flat. We put q = min

2mi−1, ni+1 if mi, ni+1 > 0 and q = 1 otherwise. Furthermore, we put Gi+1 = Gen(Gi,si ◦α−1i ), and we let si: Gi → Gi+1 be the canonical

(20)

qλ-generic lifting ofsi◦α−1i . In particular, Lemma 5.3 yields us an isomorphism αi+1:2mi+1→IdcGi+1 such that the following equality holds:

Idcsii+1◦si◦α−1i . (7.3) It follows from Lemma 5.4 that there exists a positive homomorphismfi+1:Gi+1→ Hi+1such that Idcfi+1=fi+1◦α−1i+1 andfi+1◦si=ti◦fi. The situation may be visualized by the following commutative diagrams:

Hi+1 IdcHi+1

Gi+1

fi+1

ccGG

GGGG GG

IdcGi+1

fi+1◦α−1i+1

ffLLL

LLLLLLL

Gi

ti◦fi

OO

si

;;w

ww ww ww ww

IdcGi

Idc(ti◦fi)

OO

αi+1◦si◦α−1i

88r

rr rr rr rr r

Hence the following diagrams are commutative:

Hi

ti

//Hi+1 IdcGi

Idcsi

//IdcGi+1 IdcGi+1

Idcfi+1

&&

LL LL LL LL LL

Gi

fi

OO

si //Gi+1

fi+1

OO

2mi αi

OO

si //2mi+1 αi+1

OO

2mi+1 αi+1

OO

fi+1//IdcHi+1

(7.4) We letGbe the direct limit of the sequence (Gi)i∈Z+, with transition mapssi:Gi→ Gi+1 and limiting mapssi:Gi→G, for alli∈Z+. HenceGis a dimension vector space. Furthermore, since the functor Idc preserves direct limits, the commutativ- ity of the left diagram and the middle diagram in (7.4) imply the existence of a positive homomorphism f: G→H and an isomorphism α:S →IdcGsuch that the following infinite diagrams are commutative:

H0

t0

//H1

t1

// ___ //H

G0

f0

OO

s0 //G1

f1

OO

s1 // _ _ _ //G f

OO (7.5)

and

IdcG0

Idcs0

//IdcG1

Idcs1

// ___ //IdcG

2m0 α0

OO

s0 //2m1 α1

OO

s1 // _ _ _ //S α

OO (7.6)

(For sake of clarity, we do not represent on these diagrams the transition maps corresponding to the direct limits in the rows of (7.5), (7.6).) To verify the equality (Idcf)◦α = f, it suffices to verify that (Idcfi)◦αi =fi holds, for all i ∈ Z+,

which is exactly (7.2).

In particular, it follows from Theorem 7.1 that every countable distributive h∨,0i-semilattice is isomorphic to IdcG for some dimension vector space G, but

Références

Documents relatifs

This property, which is a particular instance of a very general notion due to Malcev, was introduced and studied in the group-theoretic setting by Gordon- Vershik [GV97] and St¨

It is still an open problem whether every diagram D ~ of finite h∨, 0i-semilattices and h∨, 0i-homomorphisms, indexed by a finite lattice, can be lifted, with respect to the Con

Then, Section 4 presents an abstract syntax for ADs, an operational semantics based on finite automata, a denota- tional semantics based on execution traces, and semantic

We present here a modification of the method of return word and derivative sequences introduced in [Du1]. This method will be used in the next section to prove that every

No functorial solution of the Congruence Lattice Problem We shall give in this section a very elementary diagram of finite Boolean semi- lattices and {∨, 0}-homomorphisms, that

By the second author’s results in [12, 13], any variety satisfying the conclusion of Problem 2 (provided there is any) has the property that every diagram of finite distributive h∨,

T˚ uma and Wehrung prove in [34] that there exists a diagram of finite Boolean semilattices, indexed by a finite partially ordered set, that cannot be lifted, with respect to the Con

— There is no exchange ring (thus, no von Neumann regular ring and no C*-algebra of real rank zero) with finite stable rank whose semilattice of finitely generated,