Have you ever wondered as a kid as you watched someone use an Etch-a-Sketch, and were still adapting to using the knobs, how to draw a pen stroke curve? It turns out that you don’t really need to be an expert on the knobs, but with a lot of patience you can draw a pen stroke curve to any degree of accuracy just using the up-down, left-right in jagged steps, making shorter and shorter lines at each approximation to whatever line you’re trying to draw. Mathematically, we can say we are approximating a continuous curve on a finite interval with dyadic step functions. We will cover many topics, as I work my way through(doing exercises where needed) David Walnut’s book to try to understand uniform convergence of dyadic step functions to continuous functions on finite closed intervals, Generalized Fourier Series and orthonormal bases, and most importantly that the Haar series converges in
and interestingly enough also uniformly, the intricate proof of which is at the end. I use the Theorems and Definitions directly from the book, and maybe add in a little extra detail or two that helped me better understand the proofs or exercises. To begin we go over the Heine-Cantor Theorem, that a continuous function on a compact set is uniformly continuous, the proof uses metric spaces, but less generally, the result can be used to show that a real continuous function defined on a closed and bounded interval
,
, with
is also uniformly continuous.

If
is continuous, where
and
are metric spaces, and
is compact, then
is uniformly continuous, that is, for every
there exists a
such that for all
with
,
.
The proof of the theorem can be found on Wikipedia and I’ll rewrite it here in my own words.
The proof takes full advantage of the definition of compact in a metric space, that is if every open covering
of
has a finite subcovering, i.e., there is a finite collection
such that
.
First, by definition of continuous, is continuous at each point
, i.e., for every
, for every
there is a
such that if
, then
. Now letting
be arbitrary, since
is continuous at every
we can find a
(note here that the
is indexed by
to denote that it is specific to that
) such that if
is at most
apart from
,
, then
.
Note that is the ball centered at
of radius
and in particular these balls form an open covering for
,
, because
for every
by the definition of metric,
. Since
is compact, there exists a finite subcovering such that
.
Now we can pick the smallest because there are only finitely many of them, and denote it by
, Wikipedia asserts that
, but I think that should be obvious, you can’t pick out a negative number or zero out of a finite set of positive numbers. The emboldened
will be the
used in the definition of uniform convergence. It is important to note before proceeding that by the definition of a metric,
for all
, otherwise known as the triangle inequality. I used
in this case but the same applies for
because these properties hold for arbitrary metrics.
Given any such that
, since
,
must be an element of some
for some
. Thus
, and making
our center point between
and
, by the triangle inequality:
where the last inequality follows by the definition of . Note that both
, and by the definition of continuity and the way things were set up above for the deltas indexed by
,
. Now with
in the role of the center point, a second use of the triangle inequality gives
.
In particular, thanks to the Heine-Cantor theorem we know that a continuous real function on a closed and bounded interval is uniformly continuous. Thus, if
is continuous, then
is uniformly continuous by Heine-Cantor. This fact is particularly useful when trying to prove Lemma 5.23 in David Walnut’s Book
which is left as an exercise.
I will use/copy the definitions presented in the book for the infinity norm, or norm, and the
norm, here for reference and so that the uninitiated reader is not completely lost, these are the usual definitions.
A piecewise continuous function
defined on an interval
is bounded (or
) on
if there is a number
such that
for all
.
The -norm of a function
is defined by
A piecewise continuous function
defined on an interval
is square-integrable (or of class
or simply
) on
if the integral
is finite .
The -norm of a function
is defined by
I also need to show what a dyadic interval is, all they are is the partitioning of the real line into intervals of length a power of two, here’s the definition directly from the book:
For each pair of integers
, define the interval
by
The collection of all such intervals is called the collection of dyadic subintervals of
Given
, with either
or
, then either
or
In the latter two cases, the smaller interval is contained in either the right half or left half of the larger.
This is left as an exercise but the result should be intuitive. Given two intervals, one is the smaller one, and you can slide the smallest interval up and down the real line and it will fit snuggly inside the target zone of the larger interval if it so lands there. Either the smaller interval is outside the larger interval, they are disjoint, or it is contained inside the larger one. Since every finer partition divides every interval in two, and then the two halves in two again, and so on and so on, if a smaller interval is contained in a larger one it must be in one of the two halves, the right or the left.
Given a dyadic interval at scale
,
, we write
, where
and
are dyadic intervals at scale
, to denote the left half and right half of the interval
In fact,
and
has midpoint
, and the end points can be written as
and
I could also formally present the definition of a scale dyadic step function, as you can imagine it is a step function that is constant on the dyadic intervals:
A dyadic step function is a step function
with the property that for some
,
is constant on all dyadic intervals
,
. We say in this case that
is a scale
dyadic step function.
For any interval , a dyadic step function on
is a dyadic step function that is supported on
.
We can now proceed to proving the lemma.
Given
continuous on
, and
, there is a
, and a scale
dyadic step function
supported in
such that
, for all
; that is,
.
Since
is continuous, by Heine-Cantor,
is uniformly continuous, and so for all
there is a
such that for all
, if
, then
. Let
be such that
. If
and
are in the dyadic interval
, then
. Given a scale
dyadic step function
supported on
, that is
is constant on the intervals
for
, fix any point
in the dyadic interval
, define
for all
. By the uniform continuity of
, what happens on one dyadic interval happens for the rest. By above (that
is uniformly continuous), for every
, there is a
and a
such that
where for all
,
implies
, for
.
Since
is continuous, by Heine-Cantor,
is uniformly continuous, and so for all
there is a
such that for all
, if
, then
. Let
be such that
and fix
be the midpoint
of the dyadic interval
. If
is in the dyadic interval
, then
satisfies
. Given a scale
dyadic step function
supported on
, that is
is constant on the intervals
for
, define
for all
. By the uniform continuity of
, what happens on one dyadic interval happens on the rest. By above (that
uniformly continuous), for every
, there is a
such that
such that for all
,
implies
, for
.
Less formally, we can always find dyadic intervals of length less than , and given the uniform continuity of
, all points
inside these intervals satisfy
which implies
. Fixing any point
in the dyadic intervals
, and letting
(our dyadic step function supported on
) be equal to
on each respective interval, since
for all
,
but
for all
thus
.
Thus it is possible to approximate a continuous function on to any degree of accuracy with dyadic step functions supported on
. In fact, this is true for any closed and bounded interval
, but
will have some convenient properties as I will show shortly. We will develop a collection of special functions with special properties, but first the concept of orthogonality and correspondingly, orthonormality, must be defined. As stated in Walnut’s book:
A collection of functions
,
on an interval
is a (general) orthogonal system on
provided that
The collection is a (general) orthonormal system on
provided that it is an orthogonal system on
and
I will only consider real valued functions, thus there is no need for the notation denoting the complex conjugate. Part
of Remark 2.46 also states:
… we will use the inner product notation to represent the integrals in Definition 2.45. That is, we write for any functions , and
on
,
This means in particular that
I will make use of the same convenient notation as well. The question now arises about Generalized Fourier Series, which section 2.3.2 treats splendidly:
Given a function
,
on an interval
, and an orthonormal system
on
, the (generalized) Fourier coefficients,
of
with respect to
are defined by
The (generalized) Fourier series of with respect to
is
and then goes on to determine when the can be replaced with
. First two useful Lemmas are proved
Let
be an orthonormal system on an interval
. then for every
,
on
, and every
,
it is stated as well in the text, but, the proof takes full advantage of the orthonormality of
but by the orthonormality of ,
if
, and
if
, thus
and since we are only dealing with real valued functions,
, thus we have
and the desired result follows,
Let
be an orthonormal system on
. Then for every
,
on
, and every finite sequence of numbers
,
Let
be given, and let
be an orthonormal system. Then
again as in the previous lemma, by the orthonormality of
and adding and subtracting
by the previous lemma.
The two lemmas, especially 2.51, will now come in handy to show that if a generalized Fourier Series does converge to it’s infinite sum, then the generalized Fourier coefficients are precisely of the form
Given a collection of functions
,
on an interval
, the span of
, denoted
, is the collection of all finite linear combinations of the elements of
. In other words,
if and only if
for some finite sequence
.
Where the related notion of mean-square (or ) closure of
, denoted
is defined as,
if for every
, there is a function
such that
Let
be an orthonormal system on
. Then
is complete on
provided that every function
,
on
, is in
. A complete orthonormal system is called an orthonormal basis.
We will see that we can replace an arbitrary function with it’s Fourier series precisely when the collection of functions is complete. Theorem 2.57 contains several equivalent criteria to determine when an orthonormal system of functions is complete but I will only prove , that is,
is complete on
if and only if for every
,
on
,
in
on
, which follow by Theorem 2.55.
will also be useful later, that is,
is complete on
if and only if every function
on
is in
Let
be an orthonormal system on on interval
. Then a function
,
on
, is in
if and only if
in on
.
The statement
in
is equivalent to the statement of uniform convergence in
, for every
, there is an
such that
for every
and
, or more succinctly
but it turns out the expanded definition is more useful using the
definition of mean-square closure of since we need to find a linear combination
such that
, for every
. That is exactly it, for every
we can find an
such that
and thus
.
Supposing that
, if
, then by definition, there is a finite sequence
for some
, such that
Since ,
, therefore certainly
but the right hand side of the inequality is precisely the result of Lemma 2.51, and thus
David Walnut now leaves it as an exercise to prove that the result holds for all , this is not a difficult application of Lemma 2.50. To note that
is a decreasing sequence, remember by Lemma 2.50,
, where
is fixed and
, thus as
increases, we are subtracting by a nonnegative number each time, and so
from which it follows by taking the square root of both sides, since the arguments being squared are known to be positive and the square root function is an increasing function
and is a decreasing sequence. We can now replace
above with any
and the result would still hold, and so
.
By the definition of complete, is complete on
if every function
,
on
, is in
, by Theorem 2.55, a function
,
on
, is in
if and only if
, in
on
.
is complete if and only if for every function ,
on
,
.
The question is, is if we can find a collection of step-functions that is an orthonormal basis on ? As it will turn out, there will be many. Let’s begin by defining some functions
The dilation and translation operators can now be specified with their definitions:
Given
, the dilation operator,
, defined on functions
,
or
on
, is given by
Given , the translation operator,
, defined on functions
,
or
on
, is given by
Given the definition of , we can then consider all dilations of
and translations by
of
, for
, that is
The collection is known as the Haar system on
We will show the orthonormality of the system
The orthonormality of the Haar system is shown in Theorem 5.13
The Haar system on
is an orthonormal system on
If
is fixed, then orthonormality can be shown in a given scale
Since it is known by Lemma 5.2 that
If , then
because the functions are supported on disjoint intervals, hence,
If , then
To show orthonormality between scales, suppose that where we can assume that
with
Again Lemma 5.2 comes in handy, there are three possibilities
As before
because the functions are supported on disjoint intervals.
is the constant
on
, and so also on
, thus
is the constant
on
and so also on
, thus
Theorem 5.13 does most of the heavy lifting to prove that the system is orthonormal, all that is left to check is that
and
, but these are easy to see
and if , then
and clearly
, but if
, then
as well, and is an orthonormal system.
Back to our special interval we can always consider a subset of the latter collection, namely the Haar system on
, or
, which we know must be orthonormal since it is a subset of
. To show that the Haar system on
is complete the book makes use of the Splitting Lemma, which I quite like thus I will present it here
(The Splitting Lemma) Let
and let
be a scale
dyadic step function. Then
can be written as
where
has the form
for some coefficients and
is a scale
dyadic step function.
Since
is a scale
dyadic step function, it is constant on the intervals
Denote the value that
take on the interval
by
For each interval
define the scale
step function
on
by
and takes the average values of
on the left and right halves of
Letting
, since
is a scale
dyadic step function, so is
Fixing a dyadic interval
recall that
Then
Since is a scale
dyadic step function, it must be constant on the intervals
and since it’s integral on
is
, and the interval has left and right intervals
and
it must be a multiple of
on those intervals, and thus has the above form.
We can now prove that the Haar system on is a complete orthonormal system. It is convenient to use the equivalent statements of Theorem 2.57, in this case
, that is the orthonormal system is complete if and only if every
on
is in
Given
and
by Lemma 5.23 there exists
and a scale
dyadic step function on
such that
where the second to last inequality follows because by definition of the infinity norm, thus
All that is needed is to show that is in that span of the Haar system on
The Splitting Lemma can be used to write
as
where
has the form
where now since
is a scale
dyadic step function, the process can be repeated, and so on, until we get
and where
for some constants
and
is a scale
dyadic step function on
and must be constant on
and is thus a multiple of
It follows that
is in the span of the Haar system on
and
and so the Haar system on
is a complete orthonormal system.
By the completeness of the orthonormal system , we know that every function
that is
on
can be replaced with its infinite generalized Fourier series
in the sense on
, which is a very cool concept if you really think about it because of the counterintuitive notion of being able to build up any continuous function on
with discontinuous step functions that are equally spaced above and below the
-axis, and if we need to build a function that is always above the
-axis, that is when the
function comes to the rescue and bumps everything up!
In closing, one can show that the generalized Fourier series, or Haar series above is also uniformly convergent to , in addition to being convergent in
sense. Before proving this we’re going to need the Mean Value Theorem for integrals.
Let
be continuous, then there exists a
such that
Consider
where
, then by the first part of the Fundamental Theorem of Calculus,
is continuous on
and differentiable on
and
for all
and correspondingly the Mean Value Theorem applies on
, and there must exist some
such that
By the second part of the Fundamental Theorem of Calculus thus since
,
Lastly, some notation is needed, let’s denote the mean of a continuous function on by
In other words the Mean Value Theorem for Integrals implies there is a
such that
Also, the
function is handy in condensing notation, where
is a set
If
is a continuous function, then the Haar series
is uniformly convergent to
The argument will be by induction, first consider only
, then
and
, and
First note that
Now consider summing until , then
,
and together with the first term we get
I will include summing up to to illustrate the subtlety of the argument, and it will strengthen the perception of the apparent pattern.
Since and
,
, thus
Assuming for we have
then adding
to both sides yields
and
thus
and so the result holds for
By induction, for all
By the Mean Value Theorem for integrals, there is some such that
, thus we have a dyadic step function on the right hand side that is uniformly convergent to
by the first proof of Lemma 5.23.