My inaugural blog post, hurray!! This post is based on a mandatory assignment in a course in complex dynamics held at the university of Oslo 2018. For sources we mainly the book of Walters [1] and handouts. Some of the more basic results will, due to time constraints, be left unproven, but can all be found in [1].
– Introduction –
Lastly, the theorem of Misiurewicz, Przytycki and Gromov, which relates the entropy of a holomorphic map on the

– The three entropies –
We give the definitions in chronological order.
– Measure theoretic entropy –
Let be a measurable space, and
a measurable map. The collection of all probability measures will be denoted
, and the T-invariant probability measures, that is, measures
for which
for all measurable
, will be denoted
. A measurable partition is defined as a collection
of disjoint measurable sets that cover
. We define the join of two measurable partitions as
The entropy of a measurable partition is defined as the (generalized) sum
with the conventions that if the union of all sets of measure zero in
have non-zero measure, and we let
if
.
Now the entropy of T with respect to a T-invariant probability measure and a partition
with
is defined as
It should be noted that some authors define a measurable partition as a collection of measurable sets whose intersection have measure zero and which cover the space almost everywhere. Personally, I prefer the definition of a measurable partition to be independent of choice of measure. It is also customary to remove sets of measure zero in stead (as an alternative to the convention we adopt here), but this again requires a choice of measure, which is unfortunate.
The map sending an invariant probability measure is called the entropy map. It is affine, but in general not continuous.
The measure theoretic entropy of a measurable map


We also will need the following definition in the proof of the Variational principle below
Let




– Topological Entropy –
Analogously, we may define a notion of entropy in the category of compact topological spaces, which we will call the topological entropy. It was first given in the article of Adler et. al. [2]. For this we need the following preliminary definitions (here will denote a compact space).
For any open cover of
the quantity
is the minimal cardinality of any subcover of
. We define the entropy of the cover
to be the number
The join of two open covers of
is defined as the open cover
The topological entropy of a continuous map , with respect to the open cover
of
is defined as the number
The entropy of a continuous map


Though we will not delve into this, it should be mentioned that there are more definitions of topological entropy in the literature for general topological spaces which all reduce to the above definition if the space is compact.
– Metric Entropy –
Lastly we have an entropy which (to the best of my knowledge) was first introduced in [3] by Bowen. It makes explicit use of a metric, hence the name. Some authors call it the topological entropy, since, as we will see, they are actually equivalent on compact metric spaces.
Let be a compact metric space, and
a continuous map. To define the metric entropy, we need the following preliminary definitions
We define the metric on
by
The metrics and
are (strongly) equivalent Why?






A collection of points is said to be
-separated if
By compactness this set must be finite.
We let denote the maximum carnality all
-separated subsets of
. This number is necessarily finite
Why?








Define the quantity to be
The metric entropy is defined as the limit
is is well defined by the following lemma
Lemma 1 If then
, hence
is monotone non-increasing as
and bounded below by
(by inspection).
















– Alternative Defintions –
We can also treat the topological and measure theoretic entropies as limits over appropriate directed index sets. Let’s see how this plays out in the case of topological entropy, bearing in mind that the same could just as well be done with the measure theoretic entropy.
With a compact Hausdorff space, the family
can be made into a directed set with respect to the order relation of refinement, that is if
is a refinement of
. It is not hard to check
is a partial ordering on the collection of open covers, and for any open cover
and
we have
The same can be done for the measure theoretic entropy with the order relation if for all
there is some
such that
. Again
so in both these cases,
determines a directed set.
Now we can define the topological entropy and measure theoretic entropy of a map by
The etropies defined above admit a concise definition through nets, given by


so is a monotone net and hence must converge in
to it’s supremum.
The situation for is completely analogous, using Proposition 3 (3) in place of Proposition 6 (2).
Note that for compact metric spaces, we know that the above defined net of open covers has a cofinal subsequence, that is, a subnet which is also a sequence, and since all subnets of a convergent net converge to the same limit, we may as well take the limit over this sequence.
To construct the sequence we need the Lebesgue number lemma, which asserts that for any open cover of a compact metric space there exists a number
such that any disc of radius
is completely contained in some element in the cover. Let
be the open cover of all balls of radius
. Now it follows that for any open cover
of
with a Lebesgue number
, we have
for all
such that
, hence the sequence is cofinal.
– Basic Properties (Measure Theoretic Entropy)-
Here we list some of the basic properties of the entropy. First and foremost let’s check that the definition given above makes sense.
exists, is finite, and is equal to
With we will show that the sequence
is decreasing as
increases. This suffices to prove the theorem since the sequence is bounded below by 0. Noting that
, we may, for some fixed
, partition the index set into steps of
, that is, we may write
where
. Now
The result now follows by noting that as ,
.
Proposition 2
-
for every
-
- By definition we know that
but note these summands are zero off of the diagonal of
, hence we get that
Consequently
and the claim follows.
- Though it is not generally true that
equals
, this will not mater in the limit case. Let
be a measurable partition of
(of finite entropy), then
Taking supremums over all partitions, we get
The reverse inequality follows directly from the inequality
- By definition we know that
Other noteworthy properties of the measure theoretic entropy are listed in chapter 4.5 in [1]. We state them here without proof. Let be finite entropy measurable partitions of
.
Proposition 3
-
-
-
- if
, then
- (continuity)
, where
is the metric on the space of finite partitions defined by
-
-
if
is finite, where
denotes the number of elements in
.
The following theorem is used repeatedly, is easy to state, and has an elegant one-line proof, so there is no good reason not to include it here.
Let







The entropy map is an affine map, ie. it maps convex combinations to convex combinations.
It is not too hard to check that by concavity of the function we get the inequality
By log-sum arithmetics, one can also produce the inequality
From this we get that
which, as yields
Deviating a bit from Walter’s proof, we may take the limit over the net of all measurable partitions (of finite entropy) in the above equality, which produces the equality
As the next proposition shows, the above proposition can be extended to arbitrary convex combinations of ergodic probability measures, which are defined as measures for which
. It relies on the ergodic decomposition theorem which I hope to cover in a subsequent post.
If


where , then
Consult Theorem 4.11 of [1] for the proof of the following important result,
Entropy is conjugacy invariant (hence an isomorphism invariant in the category of measure spaces).
– Basic Properties (Topological Entropy )-
In this section are compact metrizable space, the maps in question will always be continuous and
will denote an open cover of
. First we collect some basic results, whose proofs are all more or less straightforward. Recall that the order relation of refinement, which we denoted
is used to determine a partial order on the collection of open covers.
Proposition 6
and
, then
.
, then
and so
and so
- For any continuous
, we have
, and so
.
- For an homeomorphism
we have
, and so
- Let
, then
for some
and
hence there are
and
such that
and
, So
. The claim follows.
- If
, that is
refines
, then we can for any subcover
construct a subcover of
with (at most) the same cardinality,
, where
. It thus follows that
Taking log on both sides shows
- If
and
are minimal covers of
, then
is an open cover in
with (at most)
elements. The claim follows.
- Since
is a refinement of
, this follows from (2)
- This follows by substituting
with
in (4).
The first thing we should check is that the definition of the topological entropy given above is well defined,
If


exists and is finite.
where the first and second inequalitites follow from Proposition 6 (3) and (4) respectively.
Writing for
we see that
so we may repeat the proof of Proposition 1.
We have the following collection of basic properties relating to the topological entropy.
Proposition 8
- Entropy is a topological invariant.
for all
.
- Assume
is a homeomorphism of compact spaces,
is continuous, and
an open cover of
, then
since
is a homeomorphism it induces an isomorphism of measure algebras, so taking supremums over all cover covers
concludes the proof.
- we have
so we get
. Conversely, since
Proposition 6 (2) yields
so the reverse inequality also holds.
– Basic Properties (Metric Entropy)-
In this section is a compact metric space.
Some properties of Metric entropy
- If
is an isometry, then
- If two metrics
and
on
are equivalent, then
- If
is a compact dynamical system, such that
is equicontinuous,
Two metrics and
are said to be equivalent (denoted
) if they induce the same topology. On a compact metric space
, this is equivalent to the existence of two fixed real numbers
such that
for all
. On a general metric space,
and
are allowed to depend
or
.
- If
is an isometry, then
, so
is constant as
varies, and has been shown previously to be finite, hence
.
- With
, we noted in the definition that
is monotone non-increasing in
. If
, then there are positive real numbers
, such that
This immediately yields
, so
andLetting
, since
converges, we see that the we get equality, and conclude that
.
- If
is equicontinuous, then there exists a
such that
for all
, if
. It follows that for a fixed
and any
,
is less than the number of
balls needed to cover
. Hence
for all
.
– Relations between the definitions –
In the literature the topological and metric entropies are used interchangeably, the reason for which is the following theorem.
If








Step 1: First let’s produce the following inequality
Letting , that is
for some
. Note that we have
since
It follows that if , then
(since )
hence . So if
is any
-separated it can contain at most one point in each element of a subcover
. This proves the inequality and consequently we get that
Step 2 : If is a Lebesgue number of
, we will now prove that
First note that if is a Lebesgue number for
with respect to the metric
, then
is the Lebesgue number for
with respect to the metric
. To see this, let
be a
-disc in the metric
centred at some arbitrary
. If
then
from which we get
So, for any we have that
is contained in some ball of radius
, hence also contained in some
(since
is the Lebesgue number of
). Define
by
. It’s now easy to check that
.
We can construct a cover of
-neighbourhoods (in the metric
) inductively such that each ball is centered around points which are separated by a distance greater than
, and the have radius (in the metric
) less than
. This is a refinement of
and hence
and the center points form a collection of
-separated points. The claim now follows, and we get the inequality
Letting be the (cofinal) sequence of open covers of
defined earlier we have seen that
. But as
both
and
go to zero, hence the above inequalities become and equalities and we get
The variational principle relates the topological/metric entropy on a compact metric space with the measure theoretic entropy with respect to regular Borel measures. Explicitly we have
If










This follows, since, with , we get
, either way
and (by convention)
.
By Proposition 3 (7) we have that
Inserting this into the above equation, we get the inequality
Now , hence
From Proposition 3 (6) we get
Now define the partition . Note that
is also an open cover of
! For each
there are at most
distinct elements from
contained in
. To see this, let
, and assume
is contained in
. It follows that
which gives a total of (at most) possible combinations. Hence
where denotes the cardinality of the collections. Since
is a minimal cover, that is, it has no proper subcovers, one can verify that
is also minimal, hence we have
. Inserting this into the definitions, we get
Substituting with
we get that
which, since
is arbitrary, shows that
for all continuous maps on
. This concludes the first part of the proof. Now we will show that we can find a
, with
arbitrarily close to
(the metric entropy) which has been shown to be equal to
. To this end, fix
, let
be a
-separated set of maximal cardinality (that is
) and define
By compactness of , we may find a subsequences, indexed by
such that
, and
converges in the vague (or weak*)-topology on
. By definition of vague convergence we see that since all
are
invariant, it follows that
so . We will show that
which, since was arbitrary, shows that we may approximate
from below by
where
is a regular Borel measures. For this we will need the following lemma
Lemma 1
- If
and
there exists a
such that
(that is, the measure of the boundary of the
-ball centered at
is zero)
- if
there is a finite partition
of
such that
and
for each
.
- If all members of
have boundary measure zero, then all members of
have boundary measure zero.
Proof
so uncountably many
‘s satisfy this property.
- We know from (1) there is an open cover
of
of balls of radius
. Define the partition
and
. One can check that
does the job.
- Let
. Then we have
Why do we care about sets with boundary measure zero? The reason is that if is a sequence of probability measures which converges to
in the vague topology, and
has
for all
, , then
As a consequence for any partition consisting of sets with zero boundary measure, we have
(1)
Employing this lemma, let be a measurable partition of
such that
and
. We have
(2)
since each contains at most one
.
Here things get messy, but the idea is to split into more manageable pieces. Fixing
, with
and let
be given by
(i.e. the smallest integer greater than
). Now we decompose the set
where . By construction we have
. From this we deduce that
since, even though we have repeated some indices more than once on the right hand side, we exploit that for any measurable partition we have
, hence the repetition does not affect the join.
Now
where the last two inequalities follow by Proposition 3 (2) and (7) respectively.
If we sum over all j’s in the interval we end up with
Divide both sides by , and take the limit as
gives
We used that
Now dividing by and taking limits as
on the right hand side we finally get the following inequality
– The Degree formula –
This section is devoted the proof a theorem due to Misiurewicz, Przytycki and Gromov, which relates the topological entropy of a holomorphic map on the complex -sphere to the degree of the map. Here is the first result which is a shameless copy of the proof given Theorem 8.3.1 in [4].
Theorem 5 (Misiurewicz-Przytycki)
Let be a smooth compact orientable manifold and
a continuously differentiable map, then
.






and the conclusion follows, since was arbitrary.
Let (the (total) derivative of
), fix any
and define
(this will make sense shortly). Define the compact set
and, using the inverse function theorem, cover by open sets where
is injective. Let
be a Lebesgue number of this covering. That is, in every
-disk of
the function
is injective. Define
Note that by our choice of we have that
for all
, since
hence is strictly contractive on
, so
, where
is a normalized volume form. Since the critical values have measure zero (Sard’s theorem) there exists a regular value
, that is values with preimages of cardinality
. For
define the set
Now we construct the -separated set by the following induction process
We will show is the set we are looking for.
is
-separated: Let
, and assume for contradiction that
. Assume first that
and
are distinct points in
. By construction then we must have that
, but since
, and
is injective on all
-discs in
, we must have
, which is impossible since
and
are in the same fiber (namely
). Hence
. Continuing the process inductively we eventually get
.
To show is large enough, first note that
since
, so we know that for any
we must have
There are hence at least numbers
such that
. For each such number there are, by construction,
distinct elements in
in the same fiber as
. Each of these numbers have at least one distinct preimage in
. This shows that
and concludes the proof.
Next we specialise to the case where the manifold is the complex n-sphere.
Let

Denote by , and let
and
be the metrics on
defined by
These metrics induce the same topology, since . Note that the metric
defined in section “metric entropy” coincides with the metric
, by this we mean that
where and
are in
.
Next we will need a result from geometry, which says that for any we have a lower bound on the volume of the disc
which is independent of both
and
. That is
where depends on
but not on
or
, and
is the (normalized) volume form associated with the Riemann metric
. Be warned that this result requires some sort of compatibility between the metric and the complex structure, which
does not satisfy.
Note that, since , we have
Let
be a maximal
-separated set in
, then we know the collection
do not intersect, and hence neither will
. So we have
Now where
is a volume form on
. By inspection we have
(3)
where the last inequality follows from algebraic black magic.
Inserting this into the entropy formula, we get
Let

– Examples –
To wrap things up, here are some useful examples to keep in mind. Any isometry or contractive map of a metric space has zero entropy, both topological, metric and measure theoretic (the measures on topological spaces are always assumed to be regular Borel measures). This is easiest verified for the metric entropy, since the metric
, we have that
is constant as
increases. The result extends to the measure theoretic entropy and topological entropy by applying Theorem 1 and Theorem 2 respectively.
Previously it was shown that if is a continuous map on a compact metric space
, and the family
is (uniformly) equicontinuous, then the
has zero entropy.
In both the above examples the Fatou set the of the family is whole domain. By this observation, at least we can deduce that if the entropy of a holomorphic map on a compact metrizable space has non-zero entropy, then it’s Julia set must be non-empty. This doesn’t really give us that much in the case of
since for any holomorphic map
with
we know that
, and for
, we have
.
In general not that much can be said about the entropy given the size and shape of its Julia set. Take for instance the maps defined by
(see [5] problem 7-g) and
. It is know that the Julia set of
is the entire sphere, and the Julia set of
is the unit circle centered at zero. But the entropy is
.