Having developed much technology for Lebesgue integration, let’s do a quick sanity check that we do, in fact, recover the usual Riemann integral.
Theorem 1. Let be a bounded measurable function. If
is Riemann-integrable, then
is Lebesgue-integrable, and
where the integral on the left-hand side denotes the Lebesgue integral (here denotes the Lebesgue measure), and the integral on the right-hand side denotes the usual Riemann integral.
Proof. Suppose for simplicity. We first note that all step functions
are simple functions. By the definition of the lower integral
and upper integral
,
Since is Riemann-integrable, both left-hand side and right-hand side equal to
, so that
Since the right-hand side is finite, so is the left-hand side, so that is Lebesgue-integrable.
For the general case, define . Now
is bounded, measurable, and Riemann-integrable. Since it is nonnegative, by the first result, it is Lebesgue-integrable. Therefore,
when restricted to
is also Lebesgue-integrable, and
Theorem 1 therefore tells us that the Lebesgue integral generalises the Riemann integrable, at least when is a bounded function. If
and measurable (but not necessarily integrable), more is true.
Let be a measure space.
Lemma 1. Let be measurable. The map
defined by
is a measure. Furthermore, implies that
.
Proof. For the empty set condition,
For the countable additivity condition, fix a pairwise-disjoint sequence . Define
. The sequence
of measurable functions monotonically increases to the measurable function
. By the monotone convergence theorem,
Therefore, is countably additive. Finally, suppose
. Fix any simple function
,
, where
. By the monotonicity of
,
. Hence,
Therefore,
as required.
Lemma 2. Let be a non-negative measurable function. Then
if and only if there exists some
with
such that
. In this case, we say that
–almost everywhere (abbreviated:
-a.e.).
Proof. For the direction , the proof in Lemma 1 yields
Hence,
For the direction , we will prove by contrapositive. Fix
with
and
. Then
Lemma 3. Let be integrable functions. Then
for any
if and only if there exists some
with
such that
. In this case, we say that
–a.e..
Proof. By linearity of the Lebesgue integral, it suffices to prove the case . For the direction
, for any
,
For the direction , define
and
. Then
are nonnegative functions and by Lemma 2,
Thus, there exists with
such that
. Similarly,
-a.e., and there exists
with
such that
. Observe that
Hence, writing ,
Therefore,
-a.e..
Lemma 4. For any non-negative measurable function , there exists a sequence
of simple functions
such that
monotonically.
Proof. For each , define for
Furthermore, define . Define the non-negative simple functions
by
Here are the functions for for illustration:
Each is made up of
pieces. By performing necessary real-analysis calculations, it’s not hard to verify that
monotonically as
.
Theorem 2 (Change-of-Variables). Let be a measurable space, and
be a measurable function that induces the pushforward measure
on
. Then for any integrable
,
Proof. We first prove the result for measurable . Firstly, suppose
is a simple function. We observe that
Therefore, , so that
Now suppose is nonnegative and measurable. Use Lemma 4 to find a sequence
of non-negative simple functions
that monotonically converge to
. It is obvious that
monotonically as well. By the monotone convergence theorem,
Finally suppose is integrable. Write
. Applying relevant linearity properties,
Let denote either of the measure spaces
or
.
Definition 2. Let be a measurable map. Suppose there exists a non-negative integrable function
such that its distribution
satisfies
If , then we call
a discrete random variable so that
and for any
,
If , then we call
a continuous random variable, and for any
,
A random variable is said to be continuous if there exists a non-negative integrable function
such that its distribution
satisfies
In these special cases, we call the probability density function of
, which is
-almost everywhere unique by Lemma 3, and define the cumulative distribution function by
.
We see therefore the unification that measure theory offers our study of probability theory—in fact we can just take these results for granted when eventually talking about (absolutely) continuous random variables.
Theorem 3. Suppose is a probability measure. Then for any continuous random variable
and continuous
,
Proof. We first claim that for any -integrable
and
,
It suffices to prove the case when is simple, and the rest follows by the monotone convergence theorem (for non-negative
) and linearity arguments on the decomposition
(for integrable
). To that end,
The result is then obvious.
Theorem 3 helps us recover the usual formula for expectation when we particularise to discrete or continuous random variables in the spirit of Definition 2:
- for discrete
,
,
- for continuous
,
.
Furthermore, by using rather than their specific implementations, our arguments remain valid when we particularise to higher-dimensional spaces like
so that we can discuss the distributions of combinations of random variables like
or even more specifically
.
We turn to cumulative distribution functions and continuous random variables next time.
—Joel Kindiak, 13 Jul 25, 2349H
Leave a comment