Banach–Alaoglu theorem explained
In functional analysis and related branches of mathematics, the Banach–Alaoglu theorem (also known as Alaoglu's theorem) states that the closed unit ball of the dual space of a normed vector space is compact in the weak* topology.[1] A common proof identifies the unit ball with the weak-* topology as a closed subset of a product of compact sets with the product topology. As a consequence of Tychonoff's theorem, this product, and hence the unit ball within, is compact.
This theorem has applications in physics when one describes the set of states of an algebra of observables, namely that any state can be written as a convex linear combination of so-called pure states.
History
According to Lawrence Narici and Edward Beckenstein, the Alaoglu theorem is a “very important result—maybe most important fact about the weak-* topology—[that] echos throughout functional analysis.” In 1912, Helly proved that the unit ball of the continuous dual space of
is countably weak-* compact. In 1932,
Stefan Banach proved that the closed unit ball in the continuous dual space of any separable
normed space is sequentially weak-* compact (Banach only considered
sequential compactness). The proof for the general case was published in 1940 by the mathematician
Leonidas Alaoglu. According to Pietsch [2007], there are at least twelve mathematicians who can lay claim to this theorem or an important predecessor to it.
The Bourbaki–Alaoglu theorem is a generalization[2] [3] of the original theorem by Bourbaki to dual topologies on locally convex spaces. This theorem is also called the Banach–Alaoglu theorem or the weak-* compactness theorem and it is commonly called simply the Alaoglu theorem.
Statement
See also: Dual system and Polar set.
If
is a vector space over the field
then
will denote the algebraic dual space of
and these two spaces are henceforth associated with the
bilinear \left\langle ⋅ , ⋅ \right\rangle:X x X\#\toK
defined by
where the triple
\left\langleX,X\#,\left\langle ⋅ , ⋅ \right\rangle\right\rangle
forms a
dual system called the .
If
is a
topological vector space (TVS) then its continuous dual space will be denoted by
where
always holds. Denote the weak-* topology on
by
and denote the weak-* topology on
by
\sigma\left(X\prime,X\right).
The weak-* topology is also called the topology of pointwise convergence because given a map
and a
net of maps
the net
converges to
in this topology if and only if for every point
in the domain, the net of values
converges to the value
Proof involving duality theory
If
is a
normed vector space, then the polar of a neighborhood is closed and norm-bounded in the dual space. In particular, if
is the open (or closed) unit ball in
then the polar of
is the closed unit ball in the continuous dual space
of
(with the
usual dual norm). Consequently, this theorem can be specialized to:
When the continuous dual space
of
is an infinite dimensional normed space then it is for the closed unit ball in
to be a compact subset when
has its usual norm topology. This is because the unit ball in the norm topology is compact if and only if the space is finite-dimensional (cf.
F. Riesz theorem). This theorem is one example of the utility of having different topologies on the same vector space.
It should be cautioned that despite appearances, the Banach–Alaoglu theorem does imply that the weak-* topology is locally compact. This is because the closed unit ball is only a neighborhood of the origin in the strong topology, but is usually not a neighborhood of the origin in the weak-* topology, as it has empty interior in the weak* topology, unless the space is finite-dimensional. In fact, it is a result of Weil that all locally compact Hausdorff topological vector spaces must be finite-dimensional.
Elementary proof
The following elementary proof does not utilize duality theory and requires only basic concepts from set theory, topology, and functional analysis. What is needed from topology is a working knowledge of net convergence in topological spaces and familiarity with the fact that a linear functional is continuous if and only if it is bounded on a neighborhood of the origin (see the articles on continuous linear functionals and sublinear functionals for details). Also required is a proper understanding of the technical details of how the space
of all functions of the form
is identified as the
Cartesian product and the relationship between
pointwise convergence, the
product topology, and
subspace topologies they induce on subsets such as the algebraic dual space
and products of subspaces such as
An explanation of these details is now given for readers who are interested.
For every real
Br~\stackrel{\scriptscriptstyledef
}~ \ will denote the closed ball of radius
centered at
and
rU~\stackrel{\scriptscriptstyledef
}~ \ for any
Identification of functions with tuples
The Cartesian product is usually thought of as the set of all
-indexed
tuples
but, since tuples are technically just functions from an indexing set, it can also be identified with the space
of all functions having prototype
as is now described:
belonging to
is identified with its (
-indexed) ""
s\bull~\stackrel{\scriptscriptstyledef
}~ (s(x))_.
in
is identified with the function
defined by
s(x)~\stackrel{\scriptscriptstyledef
}~ s_x; this function's "tuple of values" is the original tuple
This is the reason why many authors write, often without comment, the equality and why the Cartesian product is sometimes taken as the definition of the set of maps
(or conversely). However, the Cartesian product, being the
(categorical) product in the
category of
sets (which is a type of
inverse limit), also comes equipped with associated maps that are known as its (coordinate) .
The at a given point
is the function
where under the above identification,
sends a function
to
Stated in words, for a point
and function
"plugging
into
" is the same as "plugging
into
".
In particular, suppose that
are non-negative real numbers. Then
\prodx
\subseteq\prodxK=KX,
where under the above identification of tuples with functions,
is the set of all functions
such that
for every
If a subset
partitions
into
then the linear bijection
canonically identifies these two Cartesian products; moreover, this map is a
homeomorphism when these products are endowed with their product topologies. In terms of function spaces, this bijection could be expressed as
Notation for nets and function composition with nets
in
is by definition a function
from a non-empty
directed set
Every
sequence in
which by definition is just a function of the form
is also a net. As with sequences, the value of a net
at an index
is denoted by
; however, for this proof, this value
may also be denoted by the usual function parentheses notation
Similarly for
function composition, if
is any function then the net (or sequence) that results from "plugging
into
" is just the function
although this is typically denoted by
\left(F\left(xi\right)\right)i
(or by
\left(F\left(xi\right)\right)
if
is a sequence). In the proofs below, this resulting net may be denoted by any of the following notations
depending on whichever notation is cleanest or most clearly communicates the intended information. In particular, if
is continuous and
in
then the conclusion commonly written as
\left(F\left(xi\right)\right)i\toF(x)
may instead be written as
F\left(x\bull\right)\toF(x)
or
Topology
The set is assumed to be endowed with the product topology. It is well known that the product topology is identical to the topology of pointwise convergence. This is because given
and a
net
where
and every
is an element of
then the net
converges in the product topology if and only if
for every
the net
\Pr{}z\left(\left(fi\right)i\right)\to\Pr{}z(f)
converges in
where because
and
this happens if and only if
for every
the net
\left(fi(z)\right)i\tof(z)
converges in
Thus
converges to
in the product topology if and only if it converges to
pointwise on
This proof will also use the fact that the topology of pointwise convergence is preserved when passing to topological subspaces. This means, for example, that if for every
is some
(topological) subspace of
then the topology of pointwise convergence (or equivalently, the product topology) on
is equal to the
subspace topology that the set
inherits from
And if
is closed in
for every
then
is a closed subset of
Characterization of
An important fact used by the proof is that for any real
where
denotes the
supremum and
f(U)~\stackrel{\scriptscriptstyledef
}~ \. As a side note, this characterization does not hold if the closed ball
is replaced with the open ball
(and replacing
with the strict inequality
will not change this; for counter-examples, consider
X~\stackrel{\scriptscriptstyledef
}~ \mathbb and the
identity map f~\stackrel{\scriptscriptstyledef
}~ \operatorname on
).
The essence of the Banach–Alaoglu theorem can be found in the next proposition, from which the Banach–Alaoglu theorem follows. Unlike the Banach–Alaoglu theorem, this proposition does require the vector space
to endowed with any topology.
Before proving the proposition above, it is first shown how the Banach–Alaoglu theorem follows from it (unlike the proposition, Banach–Alaoglu assumes that
is a
topological vector space (TVS) and that
is a neighborhood of the origin).
The conclusion that the set
=\left\{f\inKX:f(U)\subseteqB1\right\}
is closed can also be reached by applying the following more general result, this time proved using nets, to the special case
and
Observation: If
is any set and if
is a
closed subset of a topological space
then
UB~\stackrel{\scriptscriptstyledef
}~ \left\ is a closed subset of
in the topology of pointwise convergence.
Proof of observation: Let
and suppose that
is a net in
that converges pointwise to
It remains to show that
which by definition means
For any
because
\left(fi(u)\right)i\tof(u)
in
and every value
belongs to the closed (in
) subset
so too must this net's limit belong to this closed set; thus
which completes the proof.
Let
and suppose that
is a net in
the converges to
in
To conclude that
it must be shown that
is a linear functional. So let
be a scalar and let
For any
let
denote
Because
in
which has the topology of pointwise convergence,
in
for every
By using
in place of
it follows that each of the following nets of scalars converges in
Proof that
Let
be the "multiplication by
" map defined by
M(c)~\stackrel{\scriptscriptstyledef
}~ s c. Because
is continuous and
in
it follows that
M\left(f\bull(x)\right)\toM(f(x))
where the right hand side is
and the left hand side is
which proves that
Because also
and limits in
are unique, it follows that
as desired.
Proof that
Define a net
z\bull=\left(zi\right)i:I\toK x K
by letting
zi~\stackrel{\scriptscriptstyledef
}~ \left(f_i(x), f_i(y)\right) for every
Because
f\bull(x)=\left(fi(x)\right)i\tof(x)
and
f\bull(y)=\left(fi(y)\right)i\tof(y),
it follows that
in
Let
be the addition map defined by
A(x,y)~\stackrel{\scriptscriptstyledef
}~ x + y. The continuity of
implies that
A\left(z\bull\right)\toA(f(x),f(y))
in
where the right hand side is
and the left hand side is
which proves that
Because also
it follows that
as desired.
The lemma above actually also follows from its corollary below since
is a Hausdorff complete uniform space and any subset of such a space (in particular
) is closed if and only if it is complete.
Because the underlying field
is a complete Hausdorff locally convex topological vector space, the same is true of the
product space A closed subset of a complete space is complete, so by the lemma, the space
\left(X\#,\sigma\left(X\#,X\right)\right)
is complete.
The above elementary proof of the Banach–Alaoglu theorem actually shows that if
is any subset that satisfies
X=(0,infty)U~\stackrel{\scriptscriptstyledef
}~ \ (such as any
absorbing subset of
), then
U\#~\stackrel{\scriptscriptstyledef
}~ \left\ is a weak-* compact subset of
As a side note, with the help of the above elementary proof, it may be shown (see this footnote)[4]
Notes and References
- , Theorem 3.15.
- , Theorem (4) in §20.9.
- , Theorem 23.5.
- Bell. J.. Fremlin. David. A Geometric Form of the Axiom of Choice. Fundamenta Mathematicae. 1972. 77. 2. 167–170. 10.4064/fm-77-2-167-170. 26 Dec 2021.
- This tuple
m\bull~\stackrel{\scriptscriptstyledef
}~ \left(m_x\right)_ is the least element of
with respect to natural induced pointwise partial order defined by
if and only if
for every
Thus, every neighborhood
of the origin in
can be associated with this unique (minimum) function
For any
if
is such that
then
so that in particular,
and
for every
) that the unique least element of
with respect to
this may be used as an alternative definition of this (necessarily convex and balanced) set. The function m\bull~\stackrel{\scriptscriptstyledef
}~ \left(m_x\right)_ : X \to [0, \infty)</math> is a [[seminorm]] and it is unchanged if
is replaced by the convex balanced hull of
(because U\#=[\operatorname{cobal}U]\#
). Similarly, because U\circ=\left[\operatorname{cl}XU\right]\circ,
is also unchanged if
is replaced by its closure in
Sequential Banach–Alaoglu theorem
A special case of the Banach–Alaoglu theorem is the sequential version of the theorem, which asserts that the closed unit ball of the dual space of a separable normed vector space is sequentially compact in the weak-* topology. In fact, the weak* topology on the closed unit ball of the dual of a separable space is metrizable, and thus compactness and sequential compactness are equivalent.
Specifically, let
be a separable normed space and
the closed unit ball in
Since
is separable, let
be a countable dense subset. Then the following defines a metric, where for any
in which
denotes the duality pairing of
with
Sequential compactness of
in this metric can be shown by a diagonalization argument similar to the one employed in the proof of the Arzelà–Ascoli theorem.Due to the constructive nature of its proof (as opposed to the general case, which is based on the axiom of choice), the sequential Banach–Alaoglu theorem is often used in the field of partial differential equations to construct solutions to PDE or variational problems. For instance, if one wants to minimize a functional
on the dual of a separable normed vector space
one common strategy is to first construct a minimizing sequence
which approaches the infimum of
use the sequential Banach–Alaoglu theorem to extract a subsequence that converges in the weak* topology to a limit
and then establish that
is a minimizer of
The last step often requires
to obey a (sequential) lower semi-continuity property in the weak* topology.When
is the space of finite Radon measures on the real line (so that
is the space of continuous functions vanishing at infinity, by the Riesz representation theorem), the sequential Banach–Alaoglu theorem is equivalent to the Helly selection theorem.Consequences
Consequences for normed spaces
Assume that
is a normed space and endow its continuous dual space
with the usual dual norm. - The closed unit ball in
is weak-* compact. So if
is infinite dimensional then its closed unit ball is necessarily compact in the norm topology by F. Riesz's theorem (despite it being weak-* compact).
- A Banach space is reflexive if and only if its closed unit ball is
\sigma\left(X,X\prime\right)
-compact; this is known as James' theorem. - If
is a reflexive Banach space, then every bounded sequence in
has a weakly convergent subsequence. (This follows by applying the Banach–Alaoglu theorem to a weakly metrizable subspace of
; or, more succinctly, by applying the Eberlein–Šmulian theorem.) For example, suppose that
is the space Lp space
where
and let
satisfy
Let
be a bounded sequence of functions in
Then there exists a subsequence
and an
such thatThe corresponding result for
is not true, as
is not reflexive.
Consequences for Hilbert spaces
- In a Hilbert space, every bounded and closed set is weakly relatively compact, hence every bounded net has a weakly convergent subnet (Hilbert spaces are reflexive).
- As norm-closed, convex sets are weakly closed (Hahn–Banach theorem), norm-closures of convex bounded sets in Hilbert spaces or reflexive Banach spaces are weakly compact.
- Closed and bounded sets in
are precompact with respect to the weak operator topology (the weak operator topology is weaker than the ultraweak topology which is in turn the weak-* topology with respect to the predual of
the trace class operators). Hence bounded sequences of operators have a weak accumulation point. As a consequence,
has the Heine–Borel property, if equipped with either the weak operator or the ultraweak topology.
Relation to the axiom of choice and other statements
See also: Krein–Milman theorem#Relation to other statements.
The Banach–Alaoglu may be proven by using Tychonoff's theorem, which under the Zermelo–Fraenkel set theory (ZF) axiomatic framework is equivalent to the axiom of choice. Most mainstream functional analysis relies on ZF + the axiom of choice, which is often denoted by ZFC. However, the theorem does rely upon the axiom of choice in the separable case (see above): in this case there actually exists a constructive proof. In the general case of an arbitrary normed space, the ultrafilter Lemma, which is strictly weaker than the axiom of choice and equivalent to Tychonoff's theorem for compact spaces, suffices for the proof of the Banach–Alaoglu theorem, and is in fact equivalent to it.
The Banach–Alaoglu theorem is equivalent to the ultrafilter lemma, which implies the Hahn–Banach theorem for real vector spaces (HB) but is not equivalent to it (said differently, Banach–Alaoglu is also strictly stronger than HB). However, the Hahn–Banach theorem is equivalent to the following weak version of the Banach–Alaoglu theorem for normed space[4] in which the conclusion of compactness (in the weak-* topology of the closed unit ball of the dual space) is replaced with the conclusion of (also sometimes called);
Compactness implies convex compactness because a topological space is compact if and only if every family of closed subsets having the finite intersection property (FIP) has non-empty intersection. The definition of convex compactness is similar to this characterization of compact spaces in terms of the FIP, except that it only involves those closed subsets that are also convex (rather than all closed subsets).
Notes
Proofs
References
- Book: Meise. Reinhold. Vogt. Dietmar. Introduction to Functional Analysis. Clarendon Press. Oxford, England. 1997. 0-19-851485-9. Theorem 23.5. 264.
- See Theorem 3.15, p. 68.
Further reading
- For any non-empty subset
the equality
\cap\left\{Ba:a\inA\right\}=
holds (the intersection on the left is a closed, rather than open, disk − possibly of radius
− because it is an intersection of closed subsets of
and so must itself be closed). For every
let mx=inf\left\{Rx:R\bull\inTP\right\}
so that the previous set equality implies \cap\operatorname{Box}P=
\prodx
=\prodx
=\prodx
.
From P\subseteq\cap\operatorname{Box}P
it follows that
and \cap\operatorname{Box}P\in\operatorname{Box}P,
thereby making
the least element of
with respect to
(In fact, the family
is closed under (non-nullary) arbitrary intersections and also under finite unions of at least one set). The elementary proof showed that
and
are not empty and moreover, it also even showed that
has an element
that satisfies
for every
which implies that
for every
The inclusion P~\subseteq~\left(\cap\operatorname{Box}P\right)\capX\prime~\subseteq~\left(\cap\operatorname{Box}P\right)\capX\#
is immediate; to prove the reverse inclusion, let f\in\left(\cap\operatorname{Box}P\right)\capX\#.
By definition, f\inP~\stackrel{\scriptscriptstyledef
}~ U^ if and only if
so let
and it remains to show that
From f\in\cap\operatorname{Box}P=\prod
,
it follows that f(u)=\Pr{}u(f)\in\Pr{}u\left(\prodx
\right)=
,
which implies that
as desired.
that there exist
-indexed non-negative real numbers
such that where these real numbers
can also be chosen to be "minimal" in the following sense: using P~\stackrel{\scriptscriptstyledef
}~ U^ (so
as in the proof) and defining the notation \prod
~\stackrel{\scriptscriptstyledef
}~ \prod_ B_ for any R\bull=\left(Rx\right)x\in\RX,
if then
and for every
mx=inf\left\{Rx:R\bull\inTP\right\},
which shows that these numbers
are unique; indeed, this infimum formula can be used to define them. In fact, if
denotes the set of all such products of closed balls containing the polar set
then where denotes the intersection of all sets belonging to
This implies (among other things[4]