Some Banach Spaces
To complete what we were saying about the spaces, we need to show that they’re complete. As it turns out, we can adapt the proof that mean convergence is complete, but we will take a somewhat different approach. It suffices to show that for any sequence of functions
in
so that the series of
-norms converges
the series of functions converges to some function .
For finite , Minkowski’s inequality allows us to conclude that
The monotone convergence theorem now tells us that the limiting function
is defined a.e., and that . The dominated convergence theorem can now verify that the partial sums of the series are
-convergent to
:
In the case , we can write
. Then
except on some set
of measure zero. The union of all the
must also be negligible, and so we can throw it all out and just have
. Now the series of the
converges by assumption, and thus the series of the
must converge to some function bounded by the sum of the
(except on the union of the
).
The Supremum Metric
We can actually extend what we’ve been doing with Hölder’s inequality and Minkowski’s inequality a little further. Given a metric space , we’ve already discussed the idea of an “essentially bounded” function — one for which there is some real constant
so that
for almost all
. We will write
for the collection of essentially bounded functions on the measure space. It should be clear that these form a vector space.
We also discussed the “essential supremum” of an essentially bounded function. We’ll now write this as
, suggesting that it’s a norm. And it’s clear that
, and that
if and only if
almost everywhere. Verifying the triangle identity is exactly Minkowski’s inequality.
And, indeed, we know that and
a.e., so
a.e., so whatever the least such essential upper bound is smaller still. That is,
.
Now for Hölder’s inequality. For this purpose we consider , and thus
, which means that
and
are Hölder-conjugates. Thus our assertion is that if
is integrable and
is essentially bounded, then
is integrable and
. Indeed, we know that
, and so
— both inequalities holding almost everywhere. From this, we conclude that
as we asserted. From now on, we’ll allow (and
) whenever we’re talking about a Hölder-conjugate pair or
-space.
What’s Happening?
You might be expecting another “Sunday Sample” here today, as there have been for the last 187 weeks, but not today. And you’ll find that they don’t even exist here anymore!
What happened? They’ve all been moved (along with this week’s installment) to drmathochist.wordpress.com, which will be my place for more personal posts. I’ll also relax the filter there, so Dr. Friedman can quit complaining about self-censorship (and I can post Cee Lo Green’s new track!)
At the same time, I’ve just opened up The Unapologetic Programmer for my more computer-programming related posts. It’ll be much less organized than here, I’m sure, but I know there’s a bunch of people into that sort of thing here.
Minkowski’s Inequality
We continue our project to show that the spaces are actually Banach spaces with Minkowski’s inequality. This will allow us to conclude that
is a normed vector space. It states that if
and
are both in
, then their sum
is in
, and we have the inequality
We start by considering Hölder’s inequality in a toy space I’ll whip up right now. Take two isolated points, and let each one have measure ; the whole space of both points has measure
. A function is just an assignment of a pair of real values
, and integration just means adding them together. Hölder’s inequality for this space tells us that
where and
are Hölder-conjugate to each other. We can set
,
, and
and use this inequality to find
Dividing out and raising both sides to the
th power, we conclude that
. Thus if both
and
are integrable, then so is
. Thus
must be in
.
Now we calculate
Dividing out by we find that
This lets us conclude that is a vector space. But we can also verify the triangle identity now. Indeed, if
,
, and
are all in
, then Minkowski’s inequality shows us that
which is exactly the triangle inequality we want. Thus is a norm, and
is a normed vector space.
Hölder’s Inequality
We’ve seen the space of integrable functions on a measure space , which we called
or
. We’ve seen that this gives us a complete normed vector space — a Banach space. This is what we’d like to generalize.
Given a real number , we define the space
or
to be the collection of all measurable functions
for which
is integrable. As in the case of
, we identify two functions if they’re equal
-almost everywhere.
It will turn out that these are Banach spaces. We define the norm
and we write to define a metric. This is clearly non-negative, and we see that
if and only if
-a.e., just as before. It’s also clear that
. What we need to work to check is the triangle inequality. It’s also not quite so apparent a problem, but we actually don’t know yet that this is a vector space at all! That is, how do we know that
is integrable if
and
are?
As a first step in this direction, we prove Hölder’s inequality: if and
are real numbers greater than
such that
, and if
and
, then the product
and
. To see this, we will use the function
defined for all positive real numbers by
Differentiating, we see that , so the only (positive) critical point of
is
. Since the limit as
approaches
and
are both positive infinite,
must be a local minimum. That is
For any two real numbers and
, we can consider the value
and it follows that
and thus
which is clearly also true even if we allow or
to be zero. This is known as “Young’s inequality”.
Okay, so now we can turn to the theorem itself. If either or
, the inequality clearly holds. Otherwise, we define
we can plug these into the above inequality to find
Since the measurability of and
implies that of
, and the right hand side of this inequality is integrable, we conclude that
is integrable. If we integrate, we find
and Hölder’s inequality follows.
The condition relating and
is very common in this discussion, so we will say that such a pair of real numbers are “Hölder conjugates” of each other. Given
, the Hölder conjugate
is uniquely defined by
, which is a strictly decreasing function sending
to itself (with order reversed, of course). The fact that this function has a (unique) fixed point at
will be important. In particular, we will see that this norm is associated with an inner product on
, and that Hölder’s inequality actually implies the Cauchy-Schwarz inequality!
The Measure Algebra of the Unit Interval
Let be the unit interval
, let
be the class of Borel sets on
, and let
be Lebesgue measure. If
is a sequence of partitions of the maximal element
of the measure algebra
into intervals, and if the limit of the sequence of norms
is zero, then
is dense.
If is a positive number, then we can find some
so that
. If
is a subinterval of
, then let
be the unique interval containing the left endpoint of
. If
also contains the right endpoint, then we can stop. Otherwise, let
be the next interval of
to the right of
, and keep going (at most a finite number of steps) until we get to an interval
containing the right endpoint of
. The union
of all the
can overshoot
by at most
on the left and the same amount on the right, and so
. Thus any interval can be approximated arbitrarily closely by some partition in the sequence. The general result follows, because we can always find a finite collection of intervals whose union is arbitrarily close to any given Borel set.
Now, say that every separable, non-atomic, normalized measure algebra is isomorphic to the measure ring
.
Since the metric space has a diameter of
— no two sets can differ by more than
, and
— we can find a dense sequence
in the space. For each
we can consider sets of the form
where either or
. The collection of all such sets for a given
is a partition
. It should be clear that the sequence
is decreasing, and the fact that
is dense implies that the sequence of partitions is also dense. We thus conclude that
.
The first partition has two elements
and
. We can define
and
, so that
and
. This gives us a partition
of
that reflects the structure of
. Similarly, we can carve up each of these intervals so that the resulting partition
reflects the structure of
. And we can continue, each time subdividing
into smaller intervals so that the next partition
reflects the structure of
. Since this correspondence preserves measures, it follows that
, and the above result then shows that the sequence
is dense.
Now, we can extend from partition elements occurring in
to finite unions of such elements by sending such a finite union to the finite union of corresponding elements of
. This gives an isometry from a dense subset of
to a dense subset of
. Thus we can uniquely extend it to an isometry
. Since
preserves unions and differences, and since these operations are uniformly continuous, it follows that
gives us an isomorphism of measure algebras.
Partitions in Measure Algebras
Let be a totally finite measure algebra, and write
for the maximal element. Without loss of generality, we can assume that
is normalized so that
.
We define a “partition” of an element
to be a finite set of “disjoint” elements of
whose “union” is
. Remember, of course, that the elements of
are not (necessarily) sets, so the set language is suggestive, but not necessarily literal. That is, if
then
and
The “norm” of a partition
is the maximum of the numbers
. If
is a partition of
and if
is any element of
below
, then
is a partition of
.
If and
are partitions, then we write
if each element in
is contained in an element of
. We say that a sequence of partitions is “decreasing” if
for each
. A sequence of partitions is “dense” if for every
and every positive number
there is some
and an element
so that
, and
is exactly the union of some elements in
. That is, we can use the elements in a fine enough partition in the sequence to approximate any element of
as closely as we want.
Now, if is a totally finite, non-atomic measure algebra, and if
is a dense, decreasing sequence of partitions of
, then
. Indeed, the sequence of norms
is monotonic and bounded in the interval
, and so it must have a limit. We will assume that this limit is some positive number
, and find a contradiction.
So if then at least one of the
must be big enough that
for all
. Otherwise the sequence of norms would descend below
and that couldn’t be the limit. Let
be just such an element, and consider the sequence
of partitions of
. The same argument is just as true, and we find another element
from the partition
, and so on.
Now, let be the intersection of the sequence
. By assumption, each of the
has
, and so
as well. Since
is non-atomic,
can’t be an atom, and so there must be an
with
. This element must be either contained in or disjoint from each element of each partition
.
We can take smaller than either
or
. Now no set made up of the union of any elements of any partition
can have a distance less than
from
. This shows that the sequence of partitions cannot be dense, which is the contradiction we were looking for. Thus the limit of the sequence of norms is zero.
Stone Spaces
The Stone space functor we’ve been working with sends Boolean algebras to topological spaces. Specifically, it sends them to compact Hausdorff spaces. There’s another functor floating around, of course, though it might not be the one you expect.
The clue is in our extended result. Given a topological space we define
to be the Boolean algebra of all clopen subsets. This functor is contravariant — given a continuous map
, we get a homomorphism of Boolean algebras
sending the clopen set
to its preimage
. It’s straightforward to see that this preimage is clopen. Another surprise is that this is known as the “Stone functor”, not to be confused with the Stone space functor
.
So what happens when we put these two functors together? If we start with a Boolean algebra and build its Stone space
, then the Stone functor applied to this space gives us a Boolean algebra
. This is, by construction, isomorphic to
itself. Thus the category
is contravariantly equivalent to some subcategory
of
. But which compact Hausdorff spaces arise as the Stone spaces of Boolean algebras?
Look at the other composite; starting with a topological space , we find the Boolean algebra
of its clopen subsets, and then the Stone space
of this Boolean algebra. We also get a function
. For each point
we define the Boolean algebra homomorphism
that sends a clopen set
to
if and only if
. We can see that this is a continuous map by checking that the preimage of any basic set is open. Indeed, a basic set of
is
for some clopen set
. That is,
. Which functions of the form
are in
? Exactly those for which
. Since
is clopen, this preimage is open.
Two points and
are sent to the same function
if and only if every clopen set containing
also contains
, and vice versa. That is,
and
must be in the same connected component. Indeed, if they were in different connected components, then there would be some clopen containing one but not the other. Conversely, if there is a clopen that contains one but not the other they can’t be in the same connected component. Thus this map
collapses all the connected components of
into points of
.
If this map is a homeomorphism, then no two points of
are in the same connected component. Thus each singleton
is a connected component, and we call the space “totally disconnected”. Clearly, such a space is in the image of the Stone space functor. On the other hand, if
, then
, and so this is both a necessary and a sufficient condition. Thus the “Stone spaces” form the full subcategory of
, consisting of the totally disconnected compact Hausdorff spaces. Stone’s representation theorem shows us that this category is equivalent to the dual of the category of Boolean algebras.
As a side note: I’d intended to cover the Stone-Čech compactification, but none of the references I have at hand actually cover the details. There’s a certain level below which everyone seems to simply assert certain facts and take them as given, and I can’t seem to reconstruct them myself.
Stone’s Representation Theorem III
We conclude our coverage of Stone’s representation theorem with a version for Boolean -algebra. Each such algebra
is isomorphic to a
-algebra of subsets of some space, modulo a
-ideal.
We start as we did for any Boolean algebra, by using the map sending
to the Boolean algebra of clopen subsets of the Stone space
. This algebra is, of course, our identified base. Let
be the
-ring generated by this base.
When we first dealt with measure rings, we quotiented out by the ideal of negligible sets. We don’t have a measure on , so we don’t have measurable sets, but we do have something almost as good. Just as when we discussed Baire spaces, we can use nowhere-denseness as a topological stand-in for negligibility. In fact, we’ll define a “meager” set to be any countable union of nowhere-dense sets, and we’ll let
be the collection of all meager sets in
. It is straightforward to verify that
is a
-ideal in
.
A quick side note: classically, meager sets were said to be “of the first category”. Any other sets were said to be “of the second category”. This is the root of the term “category” in the Baire category theorem.
Now if is a sequence of clopen sets, then we can find a unique preimage
. Since
is a
-algebra, we can take the countable union of these elements, and then apply
to the result. That is, we can define the set
Now, of course, we can also take the countable union of the sets themselves. If were an isomorphism of
-algebras, then this union would be exactly
itself; but we aren’t usually so lucky. Instead, I say that the difference
is nowhere-dense, and thus countable unions of clopen sets are clopen modulo meager sets.
Indeed, the countable union of the (open) sets is open, and so its complement is closed. The difference above is the intersection of the (closed) set
with the (closed) complement of the union, and is thus closed. So if it were dense on some nonempty open set
it would have to actually contain
. But since
is open it must be the union of some collection of basic clopen sets, and we can take
to be one of these sets. That is,
and
for each
. Since these relations only involve clopen sets, we find that
and
. But there’s no way this can happen if
is the union of the
!
So the map takes elements of
to clopen sets in
, and then on to their equivalence classes in
, and this map commutes with countable unions. All that remains to show that this is an isomorphism
is to show that no two clopen sets in
represent the same equivalence class. That is, if
and
are distinct clopen sets, then
cannot be meager. Equivalently, no nonempty clopen set is meager.
But this is just a consequence of the Baire category theorem, here in its formulation for compact Hausdorff spaces. Indeed, if a clopen set in any Baire space — and a compact Hausdorff space is Baire — were the countable union of closed nowhere-dense sets, then its interior would be empty. But since it’s open its interior is itself, and thus is would have to be empty. Thus if and
are clopen sets representing the same equivalence class modulo
then their (clopen) symmetric difference
is meager, and thus empty. That is,
.
Stone’s Representation Theorem II
We can extend yesterday’s result in the case that is a Boolean algebra. Now as a ring,
has a unit. We adjust our definition of the Stone space to define
That is, we insist that the ring homomorphisms preserve the identity, sending the top element of to
. This doesn’t really change anything we said yesterday, and it all goes through as before.
What is new is that the image of — the identified base for the topology on
— consists of all the subsets of
which are clopen — both open and closed. That is, elements of
correspond to unions of connected components of
.
First, we must show that is closed, since we already know that it’s open by definition. I say that the complement
is actually the open basic set
. Indeed,
, and we calculate
Thus each set is the complement of an open set, and is thus closed as well.
It also happens that our base is closed under finite unions. Indeed, we use DeMorgan’s laws and calculate
And from there we can extend to any finite unions we want.
Now I say that if a base of clopen sets in a compact space is closed under finite unions, then it contains every clopen set in the space. Indeed, such a clopen set can be written as a union of sets in the base since it’s open. This union gives an open covering of the set. Since the set is closed, it is compact. And so the open covering we just found has a finite subcover. That is, we can write our clopen set as a finite union of basic sets, and so it is itself in the base by assumption.
Thus in our particular case, our base of consists of all the clopen sets in the Stone space, as we asserted!