• Keine Ergebnisse gefunden

• a continuous mapX :Expcκ(VX)→ SX.

We callAXtheatom spaceofXand defineX = idAXand

X = XX : AX∪Expcκ(VX) → VX,

the comprehension map of X. The morphisms f : X → Y of Cm are continuous mapsVf : VXVY such that the image ofAf = Vf AXis a subset ofAY, the image ofSf = Vf SX is a subset ofSYand the diagram

Expcκ(VX)

X

Expcκ(Vf) //Expcκ(VY)

Y

SX Sf //SY

commutes. Note that as the hyperspace of a locally κ-compact Hausdorff space, Expcκ(VX)is also locallyκ-compact Hausdorff by Lemma 23. And by Lemma 30 it has weight6 λ.

We define the functor:CmCm as follows:

AX = AX SX = Expcκ(VX) Af = Af Sf = Expcκ(Vf) X = idAX X = Expcκ(VΣX)

Then ΣX : X → Xis a morphism and the following diagram commutes, which shows that X 7→ ΣXis a natural transformation fromto the identity:

X

ΣX

f //Y

ΣY

X f //Y

Thecomprehension maps categoryCmis the full subcategory ofCmgiven by those objectsX whose universe spaceVX(or equivalentlySX) isκ-compact. And thecategory of (comprehen-sion map) hyperuniversesCHypis the full subcategory ofCmgiven by those objectsXwhere ΣX is bijective and thus a homeomorphism. The restriction Cmis a functor from Cmto Cm. Finally, letPCm be the subcategory ofCm of only those objectsXwithκ-properX and only the κ-proper morphisms.

We define theκ-Alexandroff compactification functorωfromPCm toCmas follows:

AωX = AX Aωf = Af

SωX = ωSX Sωf = ωSf,

that is, Sωf(p) = p, and using the universal property of the κ-Alexandroff compactification, letωX :Expcκ(VωX) → SωXbe the unique map such that the diagram

Expcκ(VX)

X

ExpcκVX) //Expκ(VωX)

X

SX ιSX //SωX

2.6. INVERSE LIMIT MODELS 71

commutes and everything not in rng(ExpcκVX)) is mapped to p, that is, such that X = idAX∪ιSX defines a morphism ιX : X → ωX. This is possible by Lemma 29 because as SX is open inSωX, Expcκ(VX) is open in Expκ(VωX), and X is assumed to beκ-proper. For a κ-proper morphismf :X→ Y,ωfis in fact a morphism, because every cell in the diagram

Expκ(VωX) Expκ(Vωf) //

X

Expκ(VωY)

Y

Expcκ(VX)

Expcκ(VιX)

ii

x

Expcκ(Vf) //Expcκ(VY)

Expcκ(VιY) 55

Y

SX

X

tt

Sf //SY

Y

**SωX Sωf //SωY

is commutative, soωY◦Expκ(Vωf)◦Expcκ(VιX) = SωfωX◦Expcκ(VιX)and Expcκ(VιX) can be cancelled on the right because ifa ∈ Expκ(VωX) is not in the image of Expcκ(VιX), then it containsp, and in that case,p∈Expκ(Vωf)(a)and hence

ωY(Expκ(Vωf)(a)) = p = Sωf(p) = Sωf(SΣωX(a)).

The mapX7→ ιXis a natural transformation from the identity onPCmtoω.

Lemma 51. WheneverYis inCHyp,Xis inPCmXis surjective andf:X→ Y homeomor-phically embedsVXas an open subset inVY, then there is a unique morphismg : Y → ωX such thatg◦f = ιXandg(x) =pfor everyx ∈VY\rng(Vf).

Proof. Applying Lemma 29 to (Vf)−1, we obtain a map Vg : VYVωXwith the required properties, and we only have to show that it defines a homeomorphism. Firstly, leta =f(b).

Then by definition

ΣωX◦Expκ(Vg)(a) = ΣωX◦Expcκ(Vg◦f)(b) = ΣωX◦ExpcκX)(b) = ιX◦ΣX(b)

= Vg◦f◦ΣX(b) = g◦ΣY◦f(b) = Vg◦ΣY(a)

If, on the other hand, a ∈ Expκ(VY)\rng(f), then a contains an element not in rng(f) and hencep ∈ Expκ(Vg)(a)and ΣωX◦Expκ(Vg)(a) = p. Therefore it suffices to show that ΣY(a) ∈/ rng(f), because that entailsVg(ΣY(a)) = p, too. So assume ΣY(a) = f(x). Then ΣY(a) = f(ΣX(b))for someb, because ΣXis surjective. It follows that ΣY(a) = ΣY◦f(b) and thusa= f(b), contradicting our assumption.

Lemma 52. Inverse limits exist in the categoryCm. LethXi,fijibe an inverse system inCm and lethX,fi :X→ Xiibe its inverse limit inCm. Then:

hVX,Vfii= lim

TophVXi,Vfiji and hSX,Sfii= lim

TophSXi,Sfiji andX is the unique map such that for alli,Xi◦Exp(Vfi) =SfiX.

Proof. This is exactly dual to Lemma 46 about direct limits in Ex: We show that the object X with the morphismsfi defined by those topological limits is in fact an inverse limit of the given system inCm.

For each i, let hi : Z → Xi be a morphism and assume that for all j > i, hi = fij◦hj. We then defineVhusing the inverse limit property ofVX. This is the unique candidate for a suitable morphism h : Z → X, and we just have to show that it is in fact a morphism. For eachi, every path from Expcκ(VZ)toSXiis equal in the diagram

Expcκ(VX)

Expcκ(Vfi) **

X

Expcκ(VZ)

Expcκ(Vh)

oo

Expcκ(Vhi)

tt

Z

Expcκ(VXi)

Xi

SXi

SX

Sfi

33

Sh //SZ

Shi

kk

and thus SfiShZ = SfiX◦Expκ(Vh). Since hSX,Sfii is the inverse limit of the SXi, it follows thatShZ =X◦Expκ(Vh).

Also dually to the situation in the categoryEx, we recursively define functorsβ :CmCm and morphismsΣXα,β:βX→ αXfor ordinalsα 6β:

0X = X ΣXα,α = idαX 0f = f α+1X = αX ΣXα,β+1 = ΣXα,β◦ΣβX α+1f = αf For limit ordinalsγ, we define

hγX,ΣXα,γi = lim

CmhαX,ΣXα,βiα6β<γ

and iff :X → Y is a morphism, we use the inverse limit property ofγX: γfis defined as the unique morphism, such thatαf◦ΣXα,γ = ΣYα,γγffor allα < γ. ThenX 7→ ΣXα,βis a natural transformation fromβtoα, that is, for every morphismf : X→ Y, the following diagram commutes:

αX

αf

βX

ΣXα,β

oo

βf

αY βY

ΣYα,β

oo

The construction will stop atκX, becauseXκ+1,κ = κX will already be a homeomor-phism: Sinceκis regular, the systems

hVXα,VΣXα,βi and hExpκ(VXα),Expκ(VΣXα,β)i = hSXα+1,SΣXα+1,β+1i

2.6. INVERSE LIMIT MODELS 73

areκ-directed, so by Lemma 37,Xκ,κ+1is a homeomorphism from the hyperspace Expκ(VXκ) of the limit of the former to the limitSXκ of the latter.

We define =κ. Thenis a functor fromCmto the category of hyperuniversesCHyp.

Note that ifΣXis surjective,ΣX0,also is, so thatXis a quotient of the resulting hyperuniverse.

Finally, we show thatΣX0,∞is terminal among the morphisms from objects ofCHyptoX: Theorem 53. LetXbe an object ofCm,Y an object ofCHypandf :Y → X. Then there is a unique morphismg :Y → Xsuch thatf= ΣX0,∞◦g, namelyg =f◦(ΣY0,∞)−1.

X Y

f

oo

g

wwX

ΣX0,∞

gg

The functor :CmCHypis a right adjoint to the inclusion functor.

Proof. Again, this proof is mostly dual to that of Theorem 48, although the situation is a bit easier because instead ofγXwe can always useκalias∞. SinceΣ0,κ is a natural transforma-tion,

ΣX0,κ◦g = ΣX0,κκf◦(ΣY0,κ)−1 = f, soκf◦(ΣY0,κ)−1has the required property.

It remains to show that for every morhism g, the equation f = ΣX0,κ◦g implies g = κf◦ (ΣY0,κ)−1. By induction onα 6κ, we will show that the following diagram commutes:

αX κX

ΣXα,κ

oo

αY

αf

OO

ΣY0,α

//Y

g

OO

Then the caseα =κwill imply our claim, proving uniqueness:

κf = ΣXκ,κ◦g◦ΣY0,κ = g◦ΣY0,κ The caseα= 0 is just our assumptionf= ΣX0,κ◦g.

Now letα = β+1. Applying to the induction hypothesis forβ, we obtain the left cell in the diagram

αX κ+1X

ΣXα,κ+1

oo

ΣXκ,κ+1

//κX

αY

αf

OO

ΣY1,α

//Y

g

OO

ΣY0,1

//Y

g

OO

and the right cell commutes becauseΣis a natural transformation. SinceΣXκ,κ+1 = ΣκX is an isomorphism, this proves the caseα.

Finally, letαbe a limit ordinal and assume the induction hypothesis for allβ < α. Then every cell in the diagram

αX

ΣXβ,α **

κX

ΣXα,κ

oo

ΣXβ,κ

ttβX

βY

βf

OO

αY

ΣYβ,α 44

ΣY0,α

//

αf

OO

Y

ΣY0,β

jj

g

OO

commutes and hence

ΣXβ,ααf◦ ΣY0,α−1

= ΣXβ,α◦ΣXα,κ◦g : Y →βX for allβ < α. So by the inverse limit property ofαX, it follows that

αf◦ ΣY0,α−1

= ΣXα,κ◦g.

To prove thatκis a right adjoint, we show that the bijection

ΦY,X :hom(Y,X)→ hom(Y,κX), f7→ κf◦(ΣY0,κ)−1 has the required property:

Φ

Y,eXe(h1◦f◦h0) = κ(h1◦f◦h0)◦(ΣY0,κe )−1 = κh1κf◦κh0◦(ΣY0,κe )−1

= κh1κf◦(ΣY0,κ)−1◦h0 = κh1◦ΦY,X(f)◦h0

for all morphismsh0 :Ye→ Y,h1 :X→ Xeandf:Y → X.

A difference with the categoryExis that proving nontriviality ofXis a lot easier here: It suffices thatΣX is surjective andXis nontrivial. And in fact, everyXarises also from an objectYwith surjectiveΣY:

Proposition 54. Let Xbe an object of Cm. LetVY = rng(ΣX0,) and SY = SXVY. Then ΣY = ΣX (AX∪Expκ(VY))defines an objectY andXis isomorphic toY.

Proof. Givena∈ Expκ(VY), letb= (ΣX0,)−1[a]∈Expκ(VX). Then

ΣX(a) = ΣXX0,[b]) = ΣX0,1X1,κ+1(b)) = ΣX0,Xκ,κ+1(b)) ∈ rng(ΣX0,) ThereforeΣXactually maps Expκ(VY)toVYand the restriction defines an object ofCm.

2.6. INVERSE LIMIT MODELS 75

X0, also defines a surjective morphism h : X → Y, such that h◦f = ΣX0, for the canonical inclusionf :Y → X. By Theorem 53, there is a uniqueeh :X→ Ysuch that f◦ΣY0,◦eh =ΣX0,. But then

ΣX0,∞f◦eh = f◦ΣY0,∞◦eh = ΣX0,∞ = ΣX0,∞◦idX.

Since again by Theorem 53, idXis the unique morphismX→ Xwith that property, f◦he = idX. Analogously,

f◦ΣY0,◦he◦f = ΣX0,f = f◦ΣY0, = f◦ΣY0,◦idY

implieseh◦f = idY, becausef, being injective, cancels on the left, so Theorem 53 can be applied again.

Proposition 55. There exists aκ-hyperuniverse of weight>λ

Proof. Let AX = ∅,VX discrete with |VX| = λ. Then Expcκ(VX) is discrete, too: Everya ∈ Expcκ(VX)isκ-small and

{a} = a ∩ \

x∈a

♦{x}

is open. Sinceλ = λ, Expcκ(VX)also has sizeλand there is a bijectionX : Expcκ(VX) → VX(which in particular isκ-proper). ThenXis inPCmandωXis aκ-hyperuniverse.ΣωX0, is a surjection fromVωXontoVωX. ButVωXhasλisolated points and their preimages in VωXmust be disjoint open sets, soVωXhas at least weightλ.

Another very general example of a surjective mapΣXis the union map: Take anyκ-compact Hausdorff spaces Z 6= ∅ and AX, setSX = Expκ(Z). Then the mapX : Exp(VX) → SX defined as

X(b) = S

(b∩SX) ifb /∈ AX Z ifb∈ AX

is continuous, because the preimage of V is ♦V∪(V∪AX) and the preimage of ♦V is♦♦V∪AX for every proper open subsetV ⊂ Z. The set of singletons in SX is homeo-morphic toZ, and so is the set of singletons of singletons {{x}} ∈ SX withx ∈ Z, and so on. These subspaces also survive limit steps and it turns out thatXhas a closed subset of autosingletons homeomorphic toZ.8

A subsetA⊆VXis calledtransitiveif(ΣX)−1[A∩SX]⊆A.Aispersistentif it is transitive, open andΣXmaps(ΣX)−1[A]homeomorphically ontoA.

Lemma 56. LetXbe an object ofCmandA⊆VX.

1. If A is transitive in X, then f−1[A]is transitive in Y for every morphism f : Y → X.

Moreover,A∪(A∩AX)is transitive inXandιX[A]is transitive inωX.

8A similar hyperuniverse, constructed in a different way, is described extensively in [FH96b].

2. IfAis persistent inX, then(ΣX)−1[A]andA∪AXboth are persistent inX, andιX[A]

is persistent inωX.

3. If hXα|fα,βiα,β<γ is an inverse system in Cm, f0,α maps f−10,α[A] homeomorphically ontoAfor everyα < γandf−10,α[A]is persistent for eachα < γ, then if

hXγ|fα,γiα<γ= lim

CmhXα|fα,βiα,β<γ

is the inverse limit,f0,γ f−10,γ[A]is a homeomorphism andf−10,γ[A]is persistent.

Proof. (1): These claims are easily verified by direct calculation:

Y)−1[f−1[A]∩SY] = (ΣY)−1[f−1[SX∩A]] = (f)−1[(ΣX)−1[SX∩A]]

⊆ (f)−1[A] =(f−1[A])

X)−1[A] ⊆ ((ΣX)−1[A]) = ((ΣX)−1[A∩SX]∪(A∩AX))

⊆ (A∪(A∩AX))⊆(A∪AX)

ωX)−1X[A]∩SωX] = ιX[(ΣX)−1[A∩SX]⊆ιX[A] =(ιX[A])

Since (A∪(A∩AX))SX = A, the second calculation proves that A∪(A∩AX) is transitive.

(2): In the light of (1) and the fact that(ΣX)−1[A],A,A∩AXandιX[A]are open, we only have to worry about whether their preimages are mapped homeomorphically onto the sets in question. Since (ΣX)−1[A] is a subset of A∪(A∩AX) and X is a homeomorphism by definition, it suffices to consider the mapΣXX)−1[A]. By definition that equals ExpcκXX)−1[A]), and the exponential of a homeomorphism is a homeomorphism itself.

ForιX[A], the statement is even more trivial, becauseΣXandΣωXagree on Expcκ(VX)and in particular onA.

(3): From (1) we know thatf−10,γ[A]is transitive. It follows from the construction of the inverse limit in the category of topological spaces that f−10,γ Ais a homeomorphism because every f−10,α Ais. But then Expcκ(f−10,γ A)is a homeomorphism, too. Since(ΣX0)−1[A∩SX] ⊆ A, the left, top and bottom arrow in the diagram

Expcκ(VX0)

X0

Expcκ(VXγ)

Expcκ(f0,γ)

oo

SX0 SXγ

Sf0,γ

oo

mapA homeomorphically to its preimage, so the right one must be a homeomorphism be-tween these preimages, too, implying thatf−10,γ[A]is persistent.

Theorem 57. If X is an object of Cm and A ⊆ VX is persistent, then (ΣωX0,α)−1◦ιX[A] is persistent inαωX, andωX0,α)−1◦ιX Ais a homeomorphism for every ordinalα.

2.6. INVERSE LIMIT MODELS 77

In particular,(ΣωX0,∞)−1◦ιX[A]is persistent in the hyperuniverseωXand mapped homeo-morphically ontoA.

Proof. The proof is by induction on α and follows immediately from Lemma 56: The κ-compactification step at the beginning is (2) and the limit step is (3).

Proposition 58. LetXbe an object ofPCm with non-κ-compactVXand letΣXbe a homeo-morphism. TheneX = (ΣωX0,)−1◦ιXis a morphism which embedsVXas a dense open subset inVωX.

IfYis inCHypand ˜e :X→ YembedsVXas a dense open subset inVY, then there is a unique morphismg :Y → ωXsuch thateX = g◦˜e.

Proof. IfΣXis a homeomorphism, thenΣωXmaps rng(ιX) = (ΣωX)−1[rng(ιX)] homeomor-phically onto rng(ιX), which is dense and open. It follows inductively that everyΣωX0,αmaps the dense and open preimage of rng(ιX)homeomorphically onto rng(ιX). Thus the definition ofVeXmakes sense. VeXalso defines a morphism: Let a ∈ Expκ(X). Then since ExpκωX0,κ) also maps the preimage of rng(ExpcκX))homeomorphically onto rng(ExpcκX)),

SeXX = (SΣωX0,)−1XX = (SΣωX0,)−1ωX◦ExpcκX)

= (SΣωX0,κ)−1ωX◦ExpκωX0,κ)◦(ExpκωX0,κ))−1◦ExpcκX)

= (SΣωX0,κ)−1ωX0,κωX◦(ExpκωX0,κ))−1◦ExpcκX)

= ωX◦Expcκ(eX)

Given ˜e, by Lemma 51 there is an f : Y → ωX with f◦˜e = ιX, and since the image of ˜e is dense, this f is unique. Theorem 53 yields a unique g : Y → ωX with f = ΣX0,◦g.

Thus g is unique such that ΣX0,◦g◦e˜ = ιX. But that means that the image of g◦˜e is in (ΣωX0,∞)−1[rng(ιX)]and by composing with (ΣωX0,∞)−1 on the left, we obtain the equivalent equation eX =g◦e.˜

As in the case ofEx, letVVc = Vκwith the discrete topology,AVc = {∅}andVc the identity – which makes sense because Expcκ(VVc) = SVc. ThenVc = ωVc is a hyperuniverse in which the set eVc[VVc]of well-founded sets is dense and the homeomorphic image of VVc. By Proposition 58, it is the smallest hyperuniverse with that property. In fact, Vc turns out to beκ-ultrametrizable in a canonical way which shows that it is isomorphic to the structure originally described by R. J. Malitz in [Mal76] and E. Weydert in [Wey89]:

Theorem 59. Letx∼0 yfor allx,y∈ VVc. For everyα, letx∼α+1 ywhenever

∀˜x∈ (ΣVc)−1(x) ∃˜y∈(ΣVc)−1(y) ˜x∼α y˜ and vice versa.

At limit steps, take the intersection. Then the sequenceh∼α |α < κidefines aκ-ultrametric inducing the topology ofVVc.

Proof. Firstly we show that∼α actually is a superset of∼β for allα < β < κ. We do this by induction onβ. Assume x ∼β y. If βis a limit ordinal, thenx ∼α yfor everyα < βby definition. Now assumeβ =γ+1. Since for all ˜x ∈(ΣVc)−1(x)and ˜y∈ (ΣVc)−1(y), ˜x∼γ y˜ implies ˜x ∼δ y˜ for all δ < γ by the induction hypothesis, it follows thatx ∼δ+1 y. If γis a limit, this impliesx ∼γ y, because ∼γ is the intersection of these ∼δ. Otherwise, it follows from the case whereδis the immediate predecessor ofγ.

It also follows inductively that every∼α is an equivalence relation: Intersections of equiva-lences are equivaequiva-lences, the definition is symmetric, and whenever x ∼α+1 y ∼α+1 z, just choose a ˜ywith ˜x ∼α y, and a ˜˜ zwith ˜y ∼α ˜zfor every ˜x as in the definition – then by the induction hypothesis ˜x∼α ˜z, proving thatx∼α+1 z.

If all[x]αare open, then byκ-compactness, there are onlyκ-few of them. Hence every

[x]α+1 = [

˜ x∈x

[˜x]α ∩ \

˜ x∈x

♦[˜x]α is open, too. Ifαis a limit, every[x]αequalsT

β<α[x]β. Thus it follows inductively, that there areκ-few[x]αfor everyα < κ, and they are all open.

Next we show that ∼κ, which is the intersection of all∼α for α < κ, equals∼κ+1, that is, it is not a proper superset of it: Let x 6∼κ+1 y. Wlog assume that there is an ˜x ∈ (ΣVc)−1(x) such that there is no ˜y ∈ (ΣVc)−1(y) with ˜x ∼κ ˜y. Thus for every ˜y ∈ (ΣVc)−1(y), there is anαy˜ such that ˜xis not in[˜y]αy˜. Sinceyisκ-compact, there is a family of such[˜yi]αi with aκ-small index setI, which covers y. Letβ < κbe an upper bound to these αi. This means that ˜x∼βy˜ holds fornoy˜ ∈(ΣVc)−1(y). Thusx6∼β+1 yand in particularx6∼κ y.

Now let q : VVcVMbe the quotient map, where VM = VVc/ ∼κ. Then q◦ΣVc(a) = q◦ΣVc(b) iff ΣVc(a) ∼κ ΣVc(b), which is equivalent to ΣVc(a) ∼κ+1 ΣVc(b), which in turn means that a and b intersect the same equivalence classes [x]κ. Thus Expκ(q)(a) = Expκ(q)(b). We have shown thatΣVc factors through∼κin the sense that it induces a home-omorphismM from {∅}∪Expκ(VM) to VMand thus a quotient hyperunverse M, with a quotient morphismq :Vc → M.

On the other hand, it follows recursively from the definition thatxα ywheneverx6=yand x,y∈ eVc[Vα]. In particular,q◦eVc is still injective and embedsVc as an open subset inM.

Since Vc is minimal with that property, q must therefore be an isomorphism. That implies that∼κis the diagonal and thus the familyh∼α |α < κiis aκ-ultrametric onVVcitself. Since every[x]αis open, it induces the topology.

Thetree model, a hyperuniverse presented by E. Weydert in [Wey89], is isomorphic toX, whereAXandSXare one-point spaces.9 He conjectured that the isolated points are dense in VX(in his terms, that VX isperfect). Theorem 61 is a more general criterion for this property and proves Weydert’s conjecture.

9Asequential tree T is a tree of sequences closed with respect to restrictions. For sets of sequences A let Aα= {xα|x∈A}. Define recursively: Bα ={x∈Q

ξ<αP(Bξ) |∀ζ<ξ<α xζ =xξζ}. Thend(x,y) =min({κ} dom(x∆y))defines aκ-ultrametric onVY = Bκ. WithAY = {∅}, the mapΣY : Bκ AYExpκ(Bκ),ΣY(x) = {y|∀α yα∈x}is a homeomorphism. This is E. Weydert’stree model. But in factBα+1corresponds exactly to the hyperspace ofBα, and for limit ordinalsα,Bαis the inverse limit of its predecessors. Hence these sets correspond to the spacesVαXin a canonical way.

2.6. INVERSE LIMIT MODELS 79

Given an object Xof Cm, we call a point x ∈ VαX simpleif at least one of the following conditions is met:

• αis not a successor’s successor.

• x∈ AαX.

• α =β+2,x∈ SαX,ΣXβ,β+1 xis injective and everyy∈ xis simple.

Asimple sequenceis a sequencehxγ,γ ∈ [α,β)isuch that for everyγ,xγVγXis a simple point andΣXγ12(xγ2) = xγ1 wheneverγ2 > γ1.

Lemma 60.LetXbe an object ofCmsuch thatVXis discrete andΣXis surjective,α < β < κ, and lethxγ,γ∈[α,β)ibe a simple sequence. Then the sequence can be extended to[α,β+1) and ifβ >α+ω,xβis the unique point such thatΣXγ,β(xβ) = xγ for allγ ∈[α,β).

Proof. First of all, everyVαXwithα < κis discretely small and hence discrete. The proof goes by induction onβ.

Ifβis a limit ordinal, it follows from the definition of the inverse limit thatxβis unique (and it is simple by definition).

Next assume that β = δ+1. Choose any z ∈ (ΣXδ,β[{xδ}])−1. Thenxγ+1 = ΣXγ+1,β(z) = ΣXγ,δ[z] for every γ ∈ [α,δ). Thus for every y ∈ z, the sequence hΣγ,δ(y),γ ∈ [α,δ)i is a simple sequence. By the induction hypothesis, it has a simple extension wy. Then the set xβ = {wy | y∈z}is simple itself and since

ΣXγ+1,β(xβ) = ΣXγ,δ[z] = ΣXγ+1,β(z)

for everyγ ∈[α,δ), we haveΣXδ,β(xβ) = xδ(either by the inverse limit property ifδis a limit, or simply becauseδis of the formγ+1).

Ifβ > α+ω, the sequenceshΣγ,δ(y),γ ∈ [α,δ)i by the induction hypothesis have a unique extension. Hencewy =yfor ally∈zand thusz= xβ.

Theorem 61. LetXbe an object ofCmsuch thatVXis discrete. Then the isolated points are dense inVX.

Proof. Wlog assume thatΣX is surjective. The nonempty sets of the form(ΣXα,∞)−1[U]with limit ordinalsαconstitute an open base ofVX, so it suffices to show that each contains an isolated point. Givenxα ∈ U, recursively choose simple pointsxγ according to Lemma 60, such thathxγ|γ ∈ [α,κ)iis a simple sequence. Then there is exactly one pointxκVX such thatΣXγ,κ(xκ) = xγ for eachγ. And by Lemma 60,xκis in fact the only preimage of the pointxα+ω. But{xα+ω}is open, so{xκ}is open, too, and hencexκ is isolated.