You are currently browsing the tag archive for the ‘Hilbert spaces’ tag.
In this previous blog post I noted the following easy application of Cauchy-Schwarz:
Lemma 1 (Van der Corput inequality) Letbe unit vectors in a Hilbert space
. Then
Proof: The left-hand side may be written as for some unit complex numbers
. By Cauchy-Schwarz we have
As a corollary, correlation becomes transitive in a statistical sense (even though it is not transitive in an absolute sense):
Corollary 2 (Statistical transitivity of correlation) Letbe unit vectors in a Hilbert space
such that
for all
and some
. Then we have
for at least
of the pairs
.
Proof: From the lemma, we have
One drawback with this corollary is that it does not tell us which pairs correlate. In particular, if the vector
also correlates with a separate collection
of unit vectors, the pairs
for which
correlate may have no intersection whatsoever with the pairs in which
correlate (except of course on the diagonal
where they must correlate).
While working on an ongoing research project, I recently found that there is a very simple way to get around the latter problem by exploiting the tensor power trick:
Corollary 3 (Simultaneous statistical transitivity of correlation) Letbe unit vectors in a Hilbert space for
and
such that
for all
,
and some
. Then there are at least
pairs
such that
. In particular (by Cauchy-Schwarz) we have
for all
.
Proof: Apply Corollary 2 to the unit vectors and
,
in the tensor power Hilbert space
.
It is surprisingly difficult to obtain even a qualitative version of the above conclusion (namely, if correlates with all of the
, then there are many pairs
for which
correlates with
for all
simultaneously) without some version of the tensor power trick. For instance, even the powerful Szemerédi regularity lemma, when applied to the set of pairs
for which one has correlation of
,
for a single
, does not seem to be sufficient. However, there is a reformulation of the argument using the Schur product theorem as a substitute for (or really, a disguised version of) the tensor power trick. For simplicity of notation let us just work with real Hilbert spaces to illustrate the argument. We start with the identity
A separate application of tensor powers to amplify correlations was also noted in this previous blog post giving a cheap version of the Kabatjanskii-Levenstein bound, but this seems to not be directly related to this current application.
A (complex, semi-definite) inner product space is a complex vector space equipped with a sesquilinear form
which is conjugate symmetric, in the sense that
for all
, and non-negative in the sense that
for all
. By inspecting the non-negativity of
for complex numbers
, one obtains the Cauchy-Schwarz inequality
if one then defines , one then quickly concludes the triangle inequality
which then soon implies that is a semi-norm on
. If we make the additional assumption that the inner product
is positive definite, i.e. that
whenever
is non-zero, then this semi-norm becomes a norm. If
is complete with respect to the metric
induced by this norm, then
is called a Hilbert space.
The above material is extremely standard, and can be found in any graduate real analysis course; I myself covered it here. But what is perhaps less well known (except inside the fields of additive combinatorics and ergodic theory) is that the above theory of classical Hilbert spaces is just the first case of a hierarchy of higher order Hilbert spaces, in which the binary inner product is replaced with a
-ary inner product
that obeys an appropriate generalisation of the conjugate symmetry, sesquilinearity, and positive semi-definiteness axioms. Such inner products then obey a higher order Cauchy-Schwarz inequality, known as the Cauchy-Schwarz-Gowers inequality, and then also obey a triangle inequality and become semi-norms (or norms, if the inner product was non-degenerate). Examples of such norms and spaces include the Gowers uniformity norms
, the Gowers box norms
, and the Gowers-Host-Kra seminorms
; a more elementary example are the family of Lebesgue spaces
when the exponent is a power of two. They play a central role in modern additive combinatorics and to certain aspects of ergodic theory, particularly those relating to Szemerédi’s theorem (or its ergodic counterpart, the Furstenberg multiple recurrence theorem); they also arise in the regularity theory of hypergraphs (which is not unrelated to the other two topics).
A simple example to keep in mind here is the order two Hilbert space on a measure space
, where the inner product takes the form
In this brief note I would like to set out the abstract theory of such higher order Hilbert spaces. This is not new material, being already implicit in the breakthrough papers of Gowers and Host-Kra, but I just wanted to emphasise the fact that the material is abstract, and is not particularly tied to any explicit choice of norm so long as a certain axiom are satisfied. (Also, I wanted to write things down so that I would not have to reconstruct this formalism again in the future.) Unfortunately, the notation is quite heavy and the abstract axiom is a little strange; it may be that there is a better way to formulate things. In this particular case it does seem that a concrete approach is significantly clearer, but abstraction is at least possible.
Note: the discussion below is likely to be comprehensible only to readers who already have some exposure to the Gowers norms.
In the next few lectures, we will be studying four major classes of function spaces. In decreasing order of generality, these classes are the topological vector spaces, the normed vector spaces, the Banach spaces, and the Hilbert spaces. In order to motivate the discussion of the more general classes of spaces, we will first focus on the most special class – that of (real and complex) Hilbert spaces. These spaces can be viewed as generalisations of (real and complex) Euclidean spaces such as and
to infinite-dimensional settings, and indeed much of one’s Euclidean geometry intuition concerning lengths, angles, orthogonality, subspaces, etc. will transfer readily to arbitrary Hilbert spaces; in contrast, this intuition is not always accurate in the more general vector spaces mentioned above. In addition to Euclidean spaces, another fundamental example of Hilbert spaces comes from the Lebesgue spaces
of a measure space
. (There are of course many other Hilbert spaces of importance in complex analysis, harmonic analysis, and PDE, such as Hardy spaces
, Sobolev spaces
, and the space
of Hilbert-Schmidt operators, but we will not discuss those spaces much in this course. Complex Hilbert spaces also play a fundamental role in the foundations of quantum mechanics, being the natural space to hold all the possible states of a quantum system (possibly after projectivising the Hilbert space), but we will not discuss this subject here.)
Hilbert spaces are the natural abstract framework in which to study two important (and closely related) concepts: orthogonality and unitarity, allowing us to generalise familiar concepts and facts from Euclidean geometry such as the Cartesian coordinate system, rotations and reflections, and the Pythagorean theorem to Hilbert spaces. (For instance, the Fourier transform is a unitary transformation and can thus be viewed as a kind of generalised rotation.) Furthermore, the Hodge duality on Euclidean spaces has a partial analogue for Hilbert spaces, namely the Riesz representation theorem for Hilbert spaces, which makes the theory of duality and adjoints for Hilbert spaces especially simple (when compared with the more subtle theory of duality for, say, Banach spaces). Much later (next quarter, in fact), we will see that this duality allows us to extend the spectral theorem for self-adjoint matrices to that of self-adjoint operators on a Hilbert space.
These notes are only the most basic introduction to the theory of Hilbert spaces. In particular, the theory of linear transformations between two Hilbert spaces, which is perhaps the most important aspect of the subject, is not covered much at all here (but I hope to discuss it further in future lectures.)
We now begin our study of measure-preserving systems , i.e. a probability space
together with a probability space isomorphism
(thus
is invertible, with T and
both being measurable, and
for all
and all n). For various technical reasons it is convenient to restrict to the case when the
-algebra
is separable, i.e. countably generated. One reason for this is as follows:
Exercise 1. Let be a probability space with
separable. Then the Banach spaces
are separable (i.e. have a countable dense subset) for every
; in particular, the Hilbert space
is separable. Show that the claim can fail for
. (We allow the
spaces to be either real or complex valued, unless otherwise specified.)
Remark 1. In practice, the requirement that be separable is not particularly onerous. For instance, if one is studying the recurrence properties of a function
on a non-separable measure-preserving system
, one can restrict
to the separable sub-
-algebra
generated by the level sets
for integer n and rational q, thus passing to a separable measure-preserving system
on which f is still measurable. Thus we see that in many cases of interest, we can immediately reduce to the separable case. (In particular, for many of the theorems in this course, the hypothesis of separability can be dropped, though we won’t bother to specify for which ones this is the case.)
We are interested in the recurrence properties of sets or functions
. The simplest such recurrence theorem is
Theorem 1. (Poincaré recurrence theorem) Let
be a measure-preserving system, and let
be a set of positive measure. Then
. In particular,
has positive measure (and is thus non-empty) for infinitely many n.
(Compare with Theorem 1 of Lecture 3.)
Proof. For any integer , observe that
, and thus by Cauchy-Schwarz
(1)
The left-hand side of (1) can be rearranged as
(2)
On the other hand, . From this one easily obtains the asymptotic
(3)
where o(1) denotes an expression which goes to zero as N goes to infinity. Combining (1), (2), (3) and taking limits as we obtain
as desired.
Remark 2. In classical physics, the evolution of a physical system in a compact phase space is given by a (continuous-time) measure-preserving system (this is Hamilton’s equations of motion combined with Liouville’s theorem). The Poincaré recurrence theorem then has the following unintuitive consequence: every collection E of states of positive measure, no matter how small, must eventually return to overlap itself given sufficient time. For instance, if one were to burn a piece of paper in a closed system, then there exist arbitrarily small perturbations of the initial conditions such that, if one waits long enough, the piece of paper will eventually reassemble (modulo arbitrarily small error)! This seems to contradict the second law of thermodynamics, but the reason for the discrepancy is because the time required for the recurrence theorem to take effect is inversely proportional to the measure of the set E, which in physical situations is exponentially small in the number of degrees of freedom (which is already typically quite large, e.g. of the order of the Avogadro constant). This gives more than enough opportunity for Maxwell’s demon to come into play to reverse the increase of entropy. (This can be viewed as a manifestation of the curse of dimensionality.) The more sophisticated recurrence theorems we will see later have much poorer quantitative bounds still, so much so that they basically have no direct significance for any physical dynamical system with many relevant degrees of freedom.
Exercise 2. Prove the following generalisation of the Poincaré recurrence theorem: if is a measure-preserving system and
is non-negative, then
.
Exercise 3. Give examples to show that the quantity in the conclusion of Theorem 1 cannot be replaced by any larger quantity in general, regardless of the actual value of
. (Hint: use a Bernoulli system example.)
Exercise 4. Using the pigeonhole principle instead of the Cauchy-Schwarz inequality (and in particular, the statement that if , then the sets
cannot all be disjoint), prove the weaker statement that for any set E of positive measure in a measure-preserving system, the set
is non-empty for infinitely many n. (This exercise illustrates the general point that the Cauchy-Schwarz inequality can be viewed as a quantitative strengthening of the pigeonhole principle.)
For this lecture and the next we shall study several variants of the Poincaré recurrence theorem. We begin by looking at the mean ergodic theorem, which studies the limiting behaviour of the ergodic averages in various
spaces, and in particular in
.
Recent Comments