A basic type of problem that occurs throughout mathematics is the lifting problem: given some space that “sits above” some other “base” space due to a projection map
, and some map
from a third space
into the base space
, find a “lift”
of
to
, that is to say a map
such that
. In many applications we would like to have
preserve many of the properties of
(e.g., continuity, differentiability, linearity, etc.).
Of course, if the projection map is not surjective, one would not expect the lifting problem to be solvable in general, as the map
to be lifted could simply take values outside of the range of
. So it is natural to impose the requirement that
be surjective, giving the following commutative diagram to complete:
If no further requirements are placed on the lift , then the axiom of choice is precisely the assertion that the lifting problem is always solvable (once we require
to be surjective). Indeed, the axiom of choice lets us select a preimage
in the fiber of each point
, and one can lift any
by setting
. Conversely, to build a choice function for a surjective map
, it suffices to lift the identity map
to
.
Of course, the maps provided by the axiom of choice are famously pathological, being almost certain to be discontinuous, non-measurable, etc.. So now suppose that all spaces involved are topological spaces, and all maps involved are required to be continuous. Then the lifting problem is not always solvable. For instance, we have a continuous projection from
to
, but the identity map
cannot be lifted continuously up to
, because
is contractable and
is not.
However, if is a discrete space (every set is open), then the axiom of choice lets us solve the continuous lifting problem from
for any continuous surjection
, simply because every map from
to
is continuous. Conversely, the discrete spaces are the only ones with this property: if
is a topological space which is not discrete, then if one lets
be the same space
equipped with the discrete topology, then the only way one can continuously lift the identity map
through the “projection map”
(that maps each point to itself) is if
is itself discrete.
These discrete spaces are the projective objects in the category of topological spaces, since in this category the concept of an epimorphism agrees with that of a surjective continuous map. Thus can be viewed as the unique (up to isomorphism) projective object in this category that has a bijective continuous map to
.
Now let us narrow the category of topological spaces to the category of compact Hausdorff (CH) spaces. Here things should be better behaved; for instance, it is a standard fact in this category that continuous bijections are homeomorphisms, and it is still the case that the epimorphisms are the continuous surjections. So we have a usable notion of a projective object in this category: CH spaces such that any continuous map
into another CH space can be lifted via any surjective continuous map
to another CH space.
By the previous discussion, discrete CH spaces will be projective, but this is an extremely restrictive set of examples, since of course compact discrete spaces must be finite. Are there any others? The answer was worked out by Gleason:
Proposition 1 A compact Hausdorff space
is projective if and only if it is extremally disconnected, i.e., the closure of every open set is again open.
Proof: We begin with the “only if” direction. Let was projective, and let
be an open subset of
. Then the closure
and complement
are both closed, hence compact, subsets of
, so the disjoint union
is another CH space, which has an obvious surjective continuous projection map
to
formed by gluing the two inclusion maps together. As
is projective, the identity map
must then lift to a continuous map
. One easily checks that
has to map
to the first component
of the disjoint union, and
ot the second component; hence
, and so
is open, giving extremal disconnectedness.
Conversely, suppose that is extremally disconnected, that
is a continuous surjection of CH spaces, and
is continuous. We wish to lift
to a continuous map
.
We first observe that it suffices to solve the lifting problem for the identity map , that is to say we can assume without loss of generality that
and
is the identity. Indeed, for general maps
, one can introduce the pullback space
which is clearly a CH space that has a continuous surjection . Any continuous lift of the identity map
to
, when projected onto
, will give a desired lift
.
So now we are trying to lift the identity map via a continuous surjection
. Let us call this surjection
minimally surjective if no restriction
of
to a proper closed subset
of
remains surjective. An easy application of Zorn’s lemma shows that every continuous surjection
can be restricted to a minimally surjective continuous map
. Thus, without loss of generality, we may assume that
is minimally surjective.
The key claim now is that every minimally surjective map into an extremally disconnected space is in fact a bijection. Indeed, suppose for contradiction that there were two distinct points
in
that mapped to the same point
under
. By taking contrapositives of the minimal surjectivity property, we see that every open neighborhood of
must contain at least one fiber
of
, and by shrinking this neighborhood one can ensure the base point is arbitrarily close to
. Thus, every open neighborhood of
must intersect every open neighborhood of
, contradicting the Hausdorff property.
It is well known that continuous bijections between CH spaces must be homeomorphisms (they map compact sets to compact sets, hence must be open maps). So is a homeomorphism, and one can lift the identity map to the inverse map
.
Remark 2 The property of being “minimally surjective” sounds like it should have a purely category-theoretic definition, but I was unable to match this concept to a standard term in category theory (something along the lines of a “minimal epimorphism”, I would imagine).
In view of this proposition, it is now natural to look for extremally disconnected CH spaces (also known as Stonean spaces). The discrete CH spaces are one class of such spaces, but they are all finite. Unfortunately, these are the only “small” examples:
Lemma 3 Any first countable extremally disconnected CH space
is discrete.
Proof: If such a space were not discrete, one could find a sequence
in
converging to a limit
such that
for all
. One can sparsify the elements
to all be distinct, and from the Hausdorff property one can construct neighbourhoods
of each
that avoid
, and are disjoint from each other. Then
and then
are disjoint open sets that both have
as an adherent point, which is inconsistent with extremal disconnectedness: the closure of
contains
but is disjoint from
, so cannot be open.
Thus for instance there are no extremally disconnected compact metric spaces, other than the finite spaces; for instance, the Cantor space is not extremally disconnected, even though it is totally disconnected (which one can easily see to be a property implied by extremal disconnectedness). On the other hand, once we leave the first-countable world, we have plenty of such spaces:
Lemma 4 Let
be a complete Boolean algebra. Then the Stone dual
of
(i.e., the space of boolean homomorphisms
) is an extremally disconnected CH space.
Proof: The CH properties are standard. The elements of
give a basis of the topology given by the clopen sets
. Because the Boolean algebra is complete, we see that the closure of the open set
for any family
of sets is simply the clopen set
, which obviously open, giving extremal disconnectedness.
Remark 5 In fact, every extremally disconnected CH space
is homeomorphic to a Stone dual of a complete Boolean algebra (and specifically, the clopen algebra of
); see Gleason’s paper.
Corollary 6 Every CH space
is the surjective continuous image of an extremally disconnected CH space.
Proof: Take the Stone-Čech compactification of
equipped with the discrete topology, or equivalently the Stone dual of the power set
(i.e., the ultrafilters on
). By the previous lemma, this is an extremally disconnected CH space. Because every ultrafilter on a CH space has a unique limit, we have a canonical map from
to
, which one can easily check to be continuous and surjective.
Remark 7 In fact, to each CH space
one can associate an extremally disconnected CH space
with a minimally surjective continuous map
. The construction is the same, but instead of working with the entire power set
, one works with the smaller (but still complete) Boolean algebra of domains – closed subsets of
which are the closure of their interior, ordered by inclusion. This
is unique up to homoeomorphism, and is thus a canonical choice of extremally disconnected space to project onto
. See the paper of Gleason for details.
Several facts in analysis concerning CH spaces can be made easier to prove by utilizing Corollary 6 and working first in extremally disconnected spaces, where some things become simpler. My vague understanding is that this is highly compatible with the modern perspective of condensed mathematics, although I am not an expert in this area. Here, I will just give a classic example of this philosophy, due to Garling and presented in this paper of Hartig:
Theorem 8 (Riesz representation theorem) Let
be a CH space, and let
be a bounded linear functional. Then there is a (unique) Radon measure
on
(on the Baire
-algebra, generated by
) such
for all
.
Uniqueness of the measure is relatively straightforward; the difficult task is existence, and most known proofs are somewhat complicated. But one can observe that the theorem “pushes forward” under surjective maps:
Proposition 9 Suppose
is a continuous surjection between CH spaces. If the Riesz representation theorem is true for
, then it is also true for
.
Proof: As is surjective, the pullback map
is an isometry, hence every bounded linear functional on
can be viewed as a bounded linear functional on a subspace of
, and hence by the Hahn–Banach theorem it extends to a bounded linear functional on
. By the Riesz representation theorem on
, this latter functional can be represented as an integral against a Radon measure
on
. One can then check that the pushforward measure
is then a Radon measure on
, and gives the desired representation of the bounded linear functional on
.
In view of this proposition and Corollary 6, it suffices to prove the Riesz representation theorem for extremally disconnected CH spaces. But this is easy:
Proposition 10 The Riesz representation theorem is true for extremally disconnected CH spaces.
Proof: The Baire -algebra is generated by the Boolean algebra of clopen sets. A functional
induces a finitely additive measure
on this algebra by the formula
. This is in fact a premeasure, because by compactness the only way to partition a clopen set into countably many clopen sets is to have only finitely many of the latter sets non-empty. By the Carathéodory extension theorem,
then extends to a Baire measure, which one can check to be a Radon measure that represents
(the finite linear combinations of indicators of clopen sets are dense in
).