Arenas F Alexandrov Spaces


Acta Math. Univ. Comenianae 17
Vol. LXVIII, 1(1999), pp. 17 25
ALEXANDROFF SPACES
F. G. ARENAS
Abstract. In this paper we mean by an Alexandroff space a topological space such
that every point has a minimal neighborhood. We do not assume that the space
is T0. There spaces were first introduced by P. Alexandroff in 1937 in [1] and have
become relevant for the study of digital topology. We make a systematic study of
them from several points of view, including quasi-uniform spaces.
1. Introduction
In this paper we mean by an Alexandroff space (or an space with the property
of Alexandroff) a topological space such that every point has a minimal neighbor-
hood, or equivalently, has unique minimal base. This is also equivalent to the fact
that the intersection of every family of open sets is open. The minimal neighbor-
hood is denoted by V (x) and is the intersection of all open sets containing x. We
do not assume that the space is T0.
Although they were first introduced by P. Alexandroff in 1937 in [1] with the
name of Diskrete Rume, these spaces have not between systematically stud-
ied, perhaps because there were no reasons to do it. In fact the only references
that the author was able to find before the eighties were two ([8] and [7]) that
are mainly concerned with finite spaces (the most important particular case, un-
doubtely) and make claims (easily proved) of the type this property also hold
for Alexandroff spaces.
In the eighties, the interest in Alexandroff spaces was a consequence of the
very important role of finite spaces in digital topology and the fact that these
spaces have all the properties of finite spaces relevant for such theory (see [6], [5]).
However nobody has intended (as far as I know) a systematic study of all topo-
logical properties of these spaces, independently the property has direct digital
applications or not. That is the main purpose of this paper.
Received July 18, 1996.
1980 Mathematics Subject Classification (1991 Revision). Primary 54F05; Secondary 54D10,
54E15, 54E05.
Key words and phrases. Alexandroff spaces, quasi-uniform spaces, homotopy, space of
functions.
The author is partially supported by DGES grant PB95-0737.
18 F. G. ARENAS
2. Topological Properties of Alexandroff Spaces
We begin with the following characterization of a minimal base.
Theorem 2.1. Let X be an Alexandroff space and U a family of open sets.
Then U is the minimal base for the topology of X if and only if:
1. U covers X.
2. If A, B " U there exists a subfamily {Ui : i " I} of U such that A )" B =

Ui.
i"I

3. If a subfamily {Ui : i " I} of U verifies Ui " U, then there exists
i"I

i0 " I such that Ui = Ui0.
i"I
We show now in the next result how minimal bases can be induced in a subspace
or in a product, so the Alexandroff property is hereditary and finitely productive.
Theorem 2.2. Let X and Y be Alexandroff spaces with minimal bases U and V.
Then
1. If X is a subspace of Y , then U = {V )" X : V " V}.
2. X Y is an Alexandroff space with minimal base U V = {U V : U "
U, V " V}.
These first two results are quoted without proof from [8]. Note that the property
of Alexandroff is not countably productive, since discrete spaces are Alexandroff
and any compact metric totally disconnected space is a subspace of a countable
product of finite discrete spaces (see [9, 29.15]) and the Cantor set is not Alexan-
droff, for example.
The following result comes from [1].
Theorem 2.3. Let X be an Alexandroff space. X is T0 if and only if V (x) =
V (y) implies x = y.
Note that if X is Alexandroff, X is T1 if and only if V (x) = {x}, and in
that case the space is discrete. The interest of this result (and the reason to
impose the hypothesis of being T0 in the rest of the paper) is that allows to
consider a functional equivalence between the categories of T0 Alexandroff spaces
and partially ordered sets (posets in the following):
Given a poset P we construct the T0-Alexandroff space X(P ) as the set P
with the topology generated by {]!, x] : x " X}, which is a T0-Alexandroff space
with V (x) = ]!, x]. Conversely, given a T0-Alexandroff space X, we construct
a poset P (X) as the set P with the order x d" y if and only if x " V (y). It
is straightforward that X(P (X)) = X and P (X(P )) = P and that under the
functors, continuous mappings become order preserving mappings and conversely
(see Section 4 of [7]). Note that the orders can be defined in the reversed way.
There is another way to assign a topological space to a poset, via the geometric
realization of a simplicial complex, as can be found in Section 9 of [2]. That is,
ALEXANDROFF SPACES 19
given a poset P , the order complex "(P ) of P is the simplicial complex whose
k-faces are k-chains in P . Conversely, given a simplicial complex K the face poset
P (K) is K ordered by inclusion. There is also a standard way to topologize a
simplicial complex called the geometric realization. We denote it by |K|.
In Section 2 of [7], the simplicial complex "(P (X)) is called barycentric sub-
division, so we are going to denote it by sd X; on the other hand it is customary
(see again Section 9 of [2]) to call sd K = "(P (K)) the barycentric subdivision of
the simplicial complex K. There is no confusion between the two notations.
Now, to relate topological properties of the space | sd (X)| (what is called the
geometric realization of X) with those of X, we quote with our notation Theo-
rems 2 and 3 of [7].
Theorem 2.4. There exists a weak homotopy equivalence fX : | sd (X)| X
defined as fX(u) = x0 where u is in the unique open simplex of | sd (X)| with
vertices {x0, . . . , xn} " sd (X) and x0 < < xn in P (X).
Each mapping Ć: X Y between T0-Alexandroff spaces induces a simplicial
mapping |Ć|: | sd (X)| | sd (Y )| such that Ć ć% fX = fY ć% |Ć|.
Theorem 2.5. There exists a weak homotopy equivalence gK: |K| X(P (K)),
defined as gK = fX(P (K)), since "(P (X(P (K)))) = sd K and | sd K| = |K|.
Each simplicial mapping  : K L between simplicial spaces induces a mapping
" : X(P (K)) X(P (L)) such that " ć% fK is homotopic to fL ć% ||.
The following result shows the behaviours of these functors.
Theorem 2.6. Let X and Y be T0-Alexandroff spaces and let P and Q be
posets.
n n
1. "(P (sd X)) = sd "(P (X)) and "(P (sd X)) = sd "(P (X)).
2. P ("i"IXi) = "i"IP (Xi).
3. P (X Y ) = P (X) P (Y ) ( is the direct product between the posets
P (X) and P (Y )).
Proof. Straightforward.
Note that X(P Q) is P Q topologized by V (x, y) = V (x) Q *" {x} V (y),
so is not X(P ) X(Q).
Now we shall study the connectivity properties of these spaces.
Theorem 2.7. Let X be a T0-Alexandroff space. The following statements are
equivalent.
1. X is path-connected.
2. X is connected.
3. X is chain-connected.
4. For every a, b " X, there exist a0, . . . , an+1 " X such that a0 = a, an+1 =
b and V (ai) )" V (aj) = " if |i - j| d" 1.

20 F. G. ARENAS
5. For every a, b " X, there exist a0, . . . , am+1 " X such that a0 = a, am+1 =
b and V (ai) )" V (aj) = " if |i - j| d" 1.

6. For every a, b " X, there exist a0, . . . , ak+1 " X such that a0 = a, ak+1 =
b and {ai} )" {aj} = " if |i - j| d" 1.

Proof. (1) ! (2) ! (3) are valid in any topological space, (4) is the form that
(3) has in these spaces and (4) ! (5) ! (6) ! (1) is straightforward. Note that
if we rewrite (4) by means of the poset P (X) we obtain [8, 3.5].
The following results are an account of point-set topological properties satisfied
by T0 (and non-T0) Alexandroff spaces.
Theorem 2.8. Let X be a T0-Alexandroff space.
1. X is locally path-connected.
2. X is first countable.
3. X is orthocompact.
4. X is paracompact if and only if every V (x) meets only a finite number of
V (y), so if X is paracompact, them X is locally finite (but not conversely).
5. X is second countable if and only if it is countable.
"
6. X is separable if and only if X = {xn}.
n=1
"
7. X is Lindelf if and only if X = V (xn).
n=1
8. There exists Lindelf T0-Alexandroff spaces that are not separable and
separable T0-Alexandroff spaces that are not Lindelf.
9. If X is finite, then X is compact.
10. If X is locally finite, then it is locally compact.
11. X is countable if and only if X is locally countable and Lindelf.
12. If X is locally finite, X is compact if and only if X is finite.
Proof. 1. Apply the preceding theorem to the minimal neighborhood.
2. Obvious,
3. The minimal base is an open interior/preserving refinement of every open
covering.
4. The minimal base is a refinement of every open cover, so the first assertion
is clear from the definition of paracompactness.
To prove assertion, suppose not, that is, there exists x " X such that V (x) is
infinite. Then V (y) " V (x) for every y " V (x), a contradiction. X = N with
V (n) = {1, . . . , n} is locally finite (and countable) but not paracompact, since
V (1) meets every V (n).
5. Countable and first countable, in any topological space, implies second count-
able. Since second countable mean that the minimal base is countable, so is the
space.
6. If D is a countable dense subset, D = {xn : n " N} and for every x " X,
V (x) )" D is nonempty, hence xn " V (x) for some n " N, what means x " {xn},
"
so X = {xn}. The converse is in the same way.
n=1
ALEXANDROFF SPACES 21
7. Apply that X is Lindelf to the covering {V (x) : x " X}. For the converse,
if U is an open covering, for each U " U there exists n " N such that xn " U.
Rename that U as Un and note that V (xn) " Un.
8. If X = R+ and V (0) = X, V (x) = {x} for every x > 0, we have that
0
{0} = {0} and {x} = {0, x} for every x > 0, so X is Lindelf but not separable.
On the other hand, if X = [0, 1[, the first uncountable ordinal and V (x) = {y "
"
X : y d" x}, X = {0}, so X is separable, but V (xn) = [0, Sup {xn : n " N}]
n=1
and Sup {xn : n " N} = ą < 1 (the countable supremum of countable ordinals is
countable).
9. Obvious.
10. Obvious.
11. From (7) X is a countable union of countable subsets.
12. Apply compactness to the covering {V (x) : x " X}.
Theorem 2.9. Let X be an Alexandroff space.
1. X is regular if and only if V (x) is closed for every x " X (hence X is
0-dimensional).
2. If X is regular and compact, then X is locally compact.
3. If X is regular and separable, then X is perfectly normal.
4. X is pseudo-metrizable if and only if V (x) is closed and finite for every
x " X.
Proof. 1. X is regular means X locally closed.
2. Obvious.
3. Note first that if X is regular, from (1) we have that {xn} " V (xn) and (6)
and (7) of 2.8 and the separability of X give X is Lindelf. Recall also that if X
is regular and Lindelf, it is also normal. As in (6) of 2.8, we can write any open
"
set A as A = {xn} (the fact that {xn} " A comes from regularity), so every
n=1
open set is F; this together with the normality gives perfect normality.
4. From the Nagata-Smirnov s pseudo-metrization theorem, X is semimetriz-
able if and only if X is regular and has a -locally finite base. Since a theorem of
Michael says that a space is paracompact if and only if every open covering has a
-locally finite refinement, with a reasoning similar to that of (3) of 2.8 we have
that X has a -locally finite base if and only if X is paracompact. But under the
additional condition of being V (x) closed for every x " X we have that paracom-
pact is equivalent to locally finite (V (y) )" V (X) = " means y " V (x) = V (x), so

the set of y s such that V (y) meets V (x) is the set of elements of V (x)).
Note that the only T0-Alexandroff pseudo-metrizable spaces are the discrete
ones (which are metrizable in fact).
22 F. G. ARENAS
3. Alexandroff Spaces and Spaces of Functions. Homotopy
Let X be a topological space and let Y be a T0-Alexandroff space. Denote by
C(X, Y ) the space of continuous mappings from X to Y with the compact open
topology. We partially order C(X, Y ) by f d" g if and only if f(x) d" g(x) for every
x " X. The following result can be proved as Proposition 9 is in [8].
Theorem 3.1. Let X be a topological space and let Y be a T0-Alexandroff space.
The intersection of all open sets in C(X, Y ) containing the map f is {g " C(X, Y ) :
g d" f}. That is, C(X, Y ) is T0-Alexandroff and the order in P (C(X, Y )) is just
the order defined above.
We also quote from [4] the following standard result:
Lemma 3.2. Let X and Y be topological spaces and suppose that X is locally
compact. Then:
1. If Ć: X I Y is continuous, so is Ć1 : I C(X, Y ) defined as
Ć1(t): X Y for every t " I, where Ć1(t)(x) = Ć(t, x) for every x " X.
2. If  : I C(X, Y ) is continuous, so is 1 : X I Y defined as
1(x, t) = (t)(x).
In particular, these conditions are satisfied if X is an Alexandroff and locally finite
space.
Thus one has:
Corollary 3.3. Let X and Y be Alexandroff spaces and suppose that X is
locally finite. Then
1. The homotopy classes of maps from X to Y are in one-to-one correspon-
dence with the connected components of C(X, Y ).
2. If f, g : X Y and f d" g, then f is homotopic to g by a homotopy which
keeps pointwise fixed the set {x " F : f(x) = g(x)}.
Now we are going to obtain a homotopy-type classification in a way similar to
that of Section 4 of [8] for finite spaces.
Definition 3.4. Let X be an Alexandroff space.
1. x " X is linear if when it exists y, z > x then z e" y.
2. x " X is colinear if when it exists y, z < x then z d" y.
3. X is said to be a core (with base point p " X) if it is T0 and there exists
no linear or colinear points (except possibly p).
4. A core of the space X (with base point p " X) is a subspace X1 of X
(with the same base point) such that X1 ((X1, p)) is a core and such that
X1 is a strong deformation retract of X.
ALEXANDROFF SPACES 23
The proof of Theorem 2 of [8] cannot be extended to Alexandroff spaces, even
if they are countable and locally finite, since the space N topologized by V (n) =
{0, . . . , n} is an Alexandroff space with P (N) the natural numbers with the usual
order, have core (take F = {0} and the obvious mapping verifies f d" indentity,
so is a strong deformation retract) but that core cannot be constructed as in [8],
since you may take F0 = N - {0} as the first step of induction ({0} is a linear
point) and you will never obtain a core.
On the other hand, X(Z, d"), where d" is the usual order in Z, is a countable
but not locally finite T0-Alexandroff space without core. So it arise the question
of a different proof of Theorem 2 of [8] for locally finite T0-Alexandroff spaces (or
a counterexample). Anyway, assuming there is a core, Theorems 3 and 4 can be
easily generalized.
Theorem 3.5. Let X (or (X, p)) be a locally finite core. Then any mapping
f : X X (preserving the base point) which is homotopic to the identity (relative
to base points) is the identity.
Proof. The same of Theorem 3 of [8]; local finiteness let us to make the same
introduction as in that proof and also let us to apply 3.3.
Theorem 3.6. Let X and Y be two locally finite Alexandroff spaces (with base
point p " X and q " Y ) and suppose they have cores X1 and Y1. Then X is
homotopy equivalent to Y if and only if X1 is homeomorphic to Y1 (relative to
base points).
Proof. The same as in Theorem 4 of [8].
Since contractible means homotopy equivalent to a point, it follows:
Corollary 3.7. Let X be a locally finite Alexandroff space. Then X is con-
tractible if and only if some point of X is a strong deformation retract of X.
Finally, the same proof given in [8] for Proposition 12 gives for Alexandroff
spaces the following:
Proposition 3.8. Let (X, p) be a core, x " X. Then:
1. x is less than two distinct maximal points, or
2. x is maximal, or
3. x is linear under a maximal point; hence x = p.
and
1. x is greater than two distinct minimal points, or
2. x is minimal, or
3. x is colinear over a minimal point; hence x = p.
24 F. G. ARENAS
4. Alexandroff Spaces and Quasi-Uniform Spaces
Another interesting way to handle Alexandroff spaces is the use of quasi-uni-
formities. As it is well-known, quasi-uniformities provide an useful tool to develop
properties of any topological space. In the case of Alexandroff spaces we obtain a
very simple characterization of these spaces on terms of bases of quasi-uniformities.
Theorem 4.1. Let X be a topological space. X is an Alexandroff space if and
only if there exists a subset A of X X containing the diagonal " such that {A}
is a base for a quasi-uniformity compatible with the topology of X.
Proof. According to [3], the whole topology is an interior preserving family
(since the intersection of every family of open sets is open), so the finest quasi-
uniformity associated to X is FT = {UC : C " T }, where UC = {(x, y) " X X :

x " X, y " C} = {x} W (x, C), where W (x, C) = C.
x"C"C x"X x"C"C
Since (x) " W (x, C) for every C " T , we have that {A} is a base for FT if
V
A = {x} V (x), which clearly contains the diagonal.
x"X
Conversely, from Proposition 1.4 of [3], if {A} is a base for a quasi-uniformity
compatible with the topology of X, then {A(x)} is a neighborhood base for x
in X, so every point of X has a minimal neighborhood and then X is Alexandroff.
In fact, since A(x) = {y " X : (x, y) " A} and clearly A can be written as

A = {x} A(x), we have that the finest quasi-uniformity associated to the
x"X
topological space associated to {A} is just the given quasi-uniformity.
Thus, the theorem essentially says that Alexandroff spaces are the only spaces
whose topology is determined by only one set.
From this theorem we can say that a quasi-uniform space is Alexandroff if it
has a base of the quasi-uniformity with only one member.
Recall now that given an uniformity U we can construct two uniformities U-1
and U". One can easily check that U-1 is the uniformity with {A-1} as a base
and U" is the uniformity with {A"} as a base, and A" is just the diagonal when
the space is T0 (sketch: (x, y) " A" if and only if (x, y) " A and (y, x) " A, hence
V (x) = V (y), so x = y).
So any uniform property that relates properties of U, U-1 and U" becomes easy
to handle, since the first two are generated by one set and the third is the discrete
one. As an example we have the following.
Corollary 4.2. Any Alexandroff uniform space is complete and bicomplete.
Proof. (X, U) is bicomplete if and only if (X, U") is complete, and the discrete
uniformity always is.
About completeness, recall that in this case, a filter F is U-Cauchy if there is
x " X: A(x) " F. Since A(x) = V (x) and V (x) is the minimal neighborhood
of x, the filter F converges to x, hence (X, U) is complete.
ALEXANDROFF SPACES 25
From 4.1, it is easy to associate a quasi-proximity to the topology of X.
Theorem 4.3. Let X be an Alexandroff space. Then AB if and only if
A )" B = " is a quasi-proximity compatible with the topology (so A B if and only

if A " B0 is a strong inclusion compatible with the topology).
Proof. The quasi-proximity associated to the finest quasi-uniformity compatible

with X is AB if and only if A B )" U = " for every {x} V (x) " U, that

x"X
is, AB if and only if there exists (x, y) " A B : y " V (x), what is equivalent to
A )" B = ".

From the above results we have the functors A and A-1 between the full subcat-
egory of Alexandroff topological spaces (AT ) and the full subcategory of Alexan-
droff quasi-uniform spaces (AU) defined as A(X, T ) = (X, U) and A(f) = f and
A-1(X, U) = (X, T ) and A-1(f) = f, where T and U are defined according to 4.1.
It is clear that if f is quasi-uniformly continuous from one quasi-uniform Alexan-
droff space into another one, it is continuous between the induced topological
spaces. Conversely, if f : X Y is continuous, then f-1(W (f(x))) is a a neigh-
borhood of x, so it contains V that is, {x}V (x) " f-1(f(x))f-1(W (f(x))),
(x),
hence {x} V (x) " {f-1(y)} f-1(W (y)) = (f f)-1( {y}
x"X y"Y y"Y
W (y)), that is AX " (f f)-1(AY ) and the quasi-uniformities are defined as those
subsets of the product containing AX and AY respectively. So it only remains to
apply the definition of quasi-uniformly continuous mapping.
So A and A-1 are both functors, each inverse of the other, and AU and AT
are equivalent categories.
References
1. Alexandroff P., Diskrete Rume, Mat. Sb. (N.S.) 2 (1937), 501 518.
2. Bjrner A., Topological methods, Chapter from Handbook of Combinatorics (Graham R.,
Grtschel and Lovasz L., eds.), North Holland, Amsterdam.
3. Fletcher P. and Lindgrem W. F., Quasi-uniform spaces, Marcel Dekker, New York, 1982.
4. Fox R. H., On topologies for function spaces, Bull. A.M.S. 51 (1945), 429 432.
5. Herman G. T., On topology as applied to image analysis, Comput. Vision, Graphics Image
Process 52 (1990), 409 415.
6. Kronheimer E. H., The topology of digital images, Top. and its Appl. 46 (1992), 279 303.
7. McCord M. C., Singular homology and homotopy groups of finite topological spaces, Duke
Math. Jour, 33 (1966), 465 474.
8. Stong R. E., Finite topological spaces, Trans. A.M.S. 123 (1966), 325 340.
9. Willard S., General Topology, Addison Wesley Publishibg Co., 1970.
F. G. Arenas, Area of Geometry and Topology, Faculty of Sciences, Univesidad de Almera, 04071
Almera, Spain; e-mail: farenas@ualm.es


Wyszukiwarka

Podobne podstrony:
2001 09 Andromeda9 Spaceship Adventures
MAKIJAŻ 261 ALEXANDRE HERCHCOVITCH
egypt alexandria the lighthouse e?
Readme For Making Bases and Arenas

więcej podobnych podstron