Graduate real analysis

David Bate david.bate@warwick.ac.uk Zeeman Building, University of Warwick, Coventry CV4 7AL
Autumn 2021

Part I General measure theory

1. Measures

We wish to assign a value to the size of subsets of some given space, such as the length, area or volume of subsets of .

Definition 1.1.

A measure on a set is a function

such that

  1. ;

  2. whenever ;

  3. whenever .

A function satisfying (2) is said to be monotonic and a function satisfying (3) is said to be countably sub-additive.

Definition 1.2.

Let be a measure on a set . A set is -measurable if, for every ,

(1.1)
Remark 1.3.
  1. Since a measure is countably sub-additive, it is sufficient to check the inequality in (1.1).

  2. In particular, it suffices to check (1.1) for with .

  3. If is -measurable then so is .

  4. If then is -measurable.

Definition 1.4.

If is a measure on a set and , the restriction of to is defined as

Lemma 1.5.

Let be a measure on a set and . Then is a measure on and any -measurable set is also -measurable.

Proof.

The fact that is a measure follows immediately from the fact that is a measure. If is -measurable, then for any ,

as required. ∎

Theorem 1.6.

Let be a measure on a set and let be the set of -measurable subsets of .

  1. If then and .

  2. is countably additive on . That is, if are disjoint then

  3. If then

  4. If and then

Proof.

We first prove (1) for finite unions and intersections. If then for every ,

by sub-additivity. Thus is -measurable and induction gives finite unions. Taking complements gives finite intersections.

To prove (2) note that the inequality is given by sub-additivity. For the other inequality, for each let be disjoint and for each let

which is measurable by (1). Note that

and that this union is disjoint. Therefore, since is -measurable,

since the are all disjoint. Therefore, by induction, for each . Finally, for each , since is monotonic,

and so letting gives (2).

(3) follows by applying (2) to the disjoint measurable sets .

(4) follows from (3) by setting , so that

and the increase. By sub-additivity,

by applying (1) for finite unions. Since , (4) follows.

Finally, to prove (1) for countable unions, for each let

an increasing sequence, and let with . Since the are -measurable,

using the fact that the are -measurable in the second equality. Taking complements shows that countable intersections of measurable sets are measurable. ∎

Definition 1.7.

A collection of subsets of a set is a -algebra if

  1. ;

  2. ;

  3. .

Theorem 1.6 shows that the set of -measurable sets is a -algebra.

For a set of subset of a set , the -algebra generated by is

By creftype 1.2, it is a -algebra.

The Borel -algebra of a topological space is the -algebra generated by the open (respectively closed) subsets of . It will be denoted by and its elements called the Borel subsets of .

A measure for which all Borel sets are measurable is a Borel measure. It is Borel regular if for every there exists a Borel with .

Theorem 1.8 (Carathéodory criterion).

Let be a metric space and a measure on which is additive on separated sets. That is, whenever with

we have

Then is a Borel measure.

Proof.

Let be closed and with . We need to show

For each let

and

Since is closed,

Moreover, the with odd are pairwise separated so

and so the sum is convergent. Similarly the sum over even indices is convergent and so

Therefore

using the additivity on separated sets for the equality and countable sub-additivity for the second inequality. ∎

Definition 1.9 (Carathéodory construction).

Let be a metric space, a set of subsets of and . For each and define

where the infimum is taken over all countable

such that

Finally, define .

For any , is a measure, as is . Theorem 1.8 shows that is a Borel measure on . Indeed, if then

If consists only of Borel sets then is Borel regular.

Remark 1.10.

The fact that whenever implies that

Definition 1.11.

We define some properties of a measure on a topological space .

  1. is locally finite if every point in has a neighbourhood of finite measure.

  2. is -finite if there exist measurable with and .

  3. is finite if .

  4. A Borel regular measure is a Radon measure if

    1. for all compact ,

    2. for all Borel .

    3. for all Borel .

Definition 1.12.

Let be a measure on a set . A property of points in holds almost everywhere (or -a.e.) if the set of points for which the property doesn’t hold has measure zero.

Definition 1.13.

Let be sets, be a measure on and let . The push forward of under , written is defined by

Definition 1.14.

The Lebesgue measure on , denoted , is defined using the Carathéodory construction with the set of cubes and . Its measurable sets are called the Lebesgue measurable subsets of .

The following lemma is very useful.

Lemma 1.15.

Let be a finite measure on a set and let be a set of -measurable subsets of . There exists disjoint such that any with

satisfies .

In particular, if each -measurable subset of of positive measure contains an element of of positive measure then we can decompose almost all of into countably many disjoint elements of .

Proof.

We find the by induction. First let be countable and disjoint such that

Now let be the set of all that are countable, disjoint and disjoint from . Let be such that

Inductively, given countable, disjoint such that each and are disjoint for , let be the set of all that are countable, disjoint and disjoint from . Let be such that

We claim that any with

satisfies . If not, let be such that . Then and

a contradiction. ∎

1.1. Exercises

Exercise 1.1.

Usually in measure theory, a measure is defined as a countably additive function defined on a -algebra. However, using our definition is simply a convenience rather than a restriction.

Indeed, suppose is a countably additive function defined on a -algebra of with . Show that it can be extended to the power set of by

and that any is -measurable. What about

Conversely, any measure is countably additive when restricted to any -algebra of measurable sets.

Exercise 1.2.

Let be a set of subsets of a set . Show that is a -algebra. Note that it is the smallest -algebra of containing .

Exercise 1.3.

Show that the following sets are Borel subsets of : , , the set of points in whose first decimal is even.

Let . Show that the set of points where is continuous is a Borel set. What about the set of points where is differentiable?

Exercise 1.4.

Let be a set and . The Dirac measure at is defined as if , otherwise. Show that is a measure on . What are its measurable sets?

Define the counting measure on to be the cardinality (finite or ) of any subset of . Show that this is a measure. What are its measurable sets?

Exercise 1.5.

For a metric space and the -dimensional Hausdorff measure on , denoted , is defined using the Carathéodory construction with the set of all sets and .

  1. Show that and are non-zero, translation invariant and -homogenous measures. That is, for any , and , and (and similarly for ).

  2. On show that there exists a such that .

  3. Let be an -Lipschitz function between two metric spaces. Show that for any and ,

  4. For any metric space , show that is the counting measure on .

  5. For , suppose that . Show that . Hence there exists a single for which for all and for all . This is called the Hausdorff dimension of , denoted .

Exercise 1.6.

The Cantor set is defined as follows. Let and for each let be obtained from deleting the “middle third” open interval from each of the intervals in . That is, ,

etc. Define . Note that is compact and hence Borel.

  1. Show that is uncountable.

  2. Let . Show that .

In particular, is an uncountable subset of with .

Exercise 1.7.

Give an examples of with for which is:

  1. finite,

  2. -finite but not finite,

  3. not -finite.

Exercise 1.8.

The fundamental properties of measures are those given in Theorem 1.6, in particular countable additivity. It is necessary for us to only require this to be true for measurable sets, as can be seen from the existence of non-measurable sets.

Define a Vitali set as follows. Consider the equivalence relation on defined by iff . By the density of in , each equivalence class intersects . Therefore, by the axiom of choice(!), we may construct a set consisting of exactly one member of each equivalence class.

Show:

  1. If are rational then and are disjoint.

  2. .

  3. Show that .

  4. Deduce that is not Lebesgue measurable.

Exercise 1.9.

Show that the two extensions given in creftype 1.1 may not agree. For example, after extending Lebesgue measure (restricted to the Borel sets), what are the values of a Vitali set?

Exercise 1.10.

Let be a finite Borel measure on a metric space . Prove that for every Borel ,

(1.2)

and

(1.3)

Property (1.2) is called inner regularity by closed sets and (1.3) is called outer regularity by open sets.

Hint: observe that it suffices to show that all Borel sets satisfy (1.2). Show that the set

is a -algebra that contains all closed subsets of .

Show that a -finite is inner regular by closed sets. Show that a -finite is outer regular by open sets if there exist open sets with for all and . Give an example of a -finite that is not outer regular by open sets.

Exercise 1.11.

Let be a complete and separable metric space. Show that any finite Borel measure on is a Radon measure.

Hint: a metric space is compact if and only if it is complete and totally bounded.

2. Integration

Definition 2.1.

Let be a measure on a set . A simple function is any function of the form

where each and the are disjoint -measurable sets. We treat .

Definition 2.2.

Let be a measure on a set and let . The (lower) integral of with respect to is

Definition 2.3.

Let be a measure on a set . A function is -measurable if is -measurable for every .

For measurable, let and (both -measurable), so that and . If one of and are finite, we say that is -integrable and we define the integral of with respect to to be

If only (respectively ) we write (respectively ).

Let be a topological space. A function is a Borel function if is a Borel set for every .

There are some simple properties of the integral to check, such as linearity and monotonicity. See creftype 2.3.

Linear combinations of measurable functions are measurable, as are limits of measurable functions. See creftype 2.2.

Theorem 2.4 (Fatou’s lemma).

Let be a measure on a set and -measurable. Then

Proof.

Let

be a simple function with

for each and each and let . For each , the sets

monotonically increase to as increases. Therefore

and hence

Since is arbitrary, the conclusion follows. ∎

Remark 2.5 (Reverse Fatou).

Suppose that there exists with and for all . Then

Indeed, this follows by applying Fatou’s lemma to .

Theorem 2.6 (Monotone convergence theorem).

Let be a measure on a set and -measurable. Suppose that for every and all , . Then

Proof.

The monotonicity of the integral gives whilst Fatou’s lemma gives . ∎

Theorem 2.7.

Let be a measure on and -measurable such that pointwise. Suppose that there exists measurable with such that for all . Then

Proof.

Observe that for all , and that . Then by the reverse Fatou lemma,

2.1. Exercises

Exercise 2.1.

For a measure on a set , let be measurable, respectively Borel. Show that the pre-image of any Borel is -measurable, respectively Borel. Compare this to the definition of a continuous function.

Exercise 2.2.

Let be a measure on and for each let be -measurable. Show that the functions

are -measurable.

Show that a linear combination of -measurable functions is -measurable. Show that a countable (pointwise) sum of -measurable functions is -measurable.

Exercise 2.3.

There are some simple properties of the integral to check:

  1. If -a.e. then

  2. The integral with respect to is a linear operator;

  3. If is -measurable then

  4. ;

  5. etc…

Exercise 2.4.

Show that is -measurable if and only if

for every and .

Exercise 2.5.

State and prove a reverse monotone convergence theorem for monotonically decreasing sequences of functions.

Exercise 2.6.

Show that the Fatou lemma is false if the functions are not uniformly bounded below.

Show that the reverse Fatou lemma is false if the sequence is not bounded above by an integrable .

Show that the monotone convergence theorem is false if the sequence does not monotonically increase.

Exercise 2.7.

Let be sets, a measure on and . Show that is a measure on . If are topological spaces and are Borel, show that is a Borel measure on .

3. Some standard theorems

Theorem 3.1 (Egorov’s theorem).

Let be a finite measure on a set and a sequence of -measurable functions such that pointwise -a.e. Then for every there exists a measurable with and uniformly on .

Proof.

Fix , and for each let

By assumption, the are measurable sets that monotonically decrease to a -null set as . Therefore, there exists such that . Let

Then and, for each and each , , so there exists such that

for all . That is, uniformly on , as required. ∎

Theorem 3.2 (Lusin’s theorem).

Let be a finite Borel measure on a metric space and let be -measurable. Then for every there exists a closed with such that is continuous.

Proof.

Fix and for each let

a collection of disjoint Borel sets which cover . Since , there exists such that

Since is Borel, for each there exists a closed with .

For a moment fix and let

Since is closed, the sets monotonically decrease to as , which is disjoint from . Therefore there exists such that

satisfies . Note that is closed.

Let and , a closed set. For any and and , . Therefore, if with , . Repeat this for each with and let , so that . Then is continuous on . ∎

Definition 3.3.

Let be measures on a set . We say is absolutely continuous with respect to , written , if for every , . We say that is singular with respect to , written if there exists with .

Let be a measure on a set and let . The set valued function

defines a measure on . Note also that . The Radon-Nikodym theorem provides the converse to this statement.

Theorem 3.4 (Radon–Nikodym).

Let be finite measures on a set such that . There exists a and -measurable such that

for all and -measurable . The function is called the Radon-Nikodym derivative of with respect to .

Proof.

Let be the set of all and -measurable such that

for all and -measurable . Note that and

(3.1)

Let

so that , and let be such that

Eq. 3.1 implies that we may suppose the monotonically increase. Let be the pointwise limit of the . Then is and -measurable and, by the monotone convergence theorem, and . Thus

(3.2)

We claim that satisfies the conclusion of the proposition. Indeed, suppose that is -measurable but

and let be such that

(3.3)

Let be the collection of all and -measurable such that

We claim that there exists a and -measurable of positive -measure such that each and -measurable of positive -measure is not contained in . Indeed, if not, then each of positive -measure contains an element of of positive -measure. Thus Lemma 1.15 gives a countable disjoint decomposition of -almost all of into elements of . Since this implies

contradicting (3.3).

Note that . Indeed, if is -measurable,

On the other hand, since ,

contradicting the definition of . ∎

Theorem 3.5 (Lebesgue decomposition theorem).

Let be finite measures on a set . There exists a -measurable with such that, for all , . That is, with and .

Proof.

Let be the set of all -measurable with . By Lemma 1.15, there exists such that each with

satisfies . Since , this is the required decomposition. ∎

3.1. Exercises

Exercise 3.1.

Let be a Borel measure on a metric space , -integrable and .

  1. Show that there exists a simple function such that

  2. If is positive show that we may require in the previous point.

  3. Show that if is finite, there exists with

  4. Show that the previous point may fail if is only -finite.

Exercise 3.2.

Prove the following variant of Lusin’s theorem for the case that is not finite but is -integrable: for every there exists a closed with

such that is continuous.

Exercise 3.3.

Let be measures on a set . Show that if is measurable then it is also -measurable.

Exercise 3.4.

Prove the Radon-Nikodym and Lebesgue decomposition theorems for -finite measures.

Exercise 3.5.

Let be finite measures on a set and suppose . For any -measurable show that

where is the Radon-Nikodym derivative of with respect to .

Exercise 3.6.

For a measure on a set , we say a function is -measurable if each component of is -measurable and define the integral of component by component.

An -dimensional vector valued measure is a function

for which there exists a measure on and a -measurable such that

for each .

  1. Show that if

    are two representations of a vector valued measure then (when restricted to the set of -measurable sets), and hence -a.e. We denote this unique measure by . It is called the total variation of .

  2. Show that the set of all vector valued measures form a normed vector space when equipped with

  3. Show that this space is complete.

A signed measure is a 1-dimensional vector valued measure.

Exercise 3.7.

Let be a -algebra on . The standard definition of a signed measure on is a countably additive function

The Hahn decomposition theorem states that there exist disjoint with such that:

  • For every with , and

  • For every with , .

That is, and are (positive) measures.

Use the Lebesgue decomposition and Radon–Nikodym theorems to show that the two definitions of a signed measure agree.

4. The Daniell integral

Let be a measure on a set and let be a set of real valued -measurable functions on . The formula

defines an operator on . Moreover, it has the following two properties:

  • is monotonic: for all with , ;

  • is continuous with respect to monotone convergence: if monotonically increase to then .

In the next theorem we see that these two properties completely characterise the Lebesgue integral. That is, we could equivalently develop a theory of integration (and hence measure) by beginning with operators on sets of functions.

Definition 4.1.

Let be a set. A lattice of functions on is a non-empty set of functions which satisfies the following conditions: for any and any , , , and all belong to and if then too. Note that any vector space of functions closed under is a lattice.

If is a lattice we let

We say a is

  1. linear if, for all and ,

  2. monotonic if, for all with ,

  3. continuous with respect to monotone convergence if, for all that monotonically increase to ,

  4. bounded if for every ,

A satisfying Items 3, 2 and 1 is called a monotone Daniell integral (it necessarily satisfies Item 4). A satisfying Items 3, 4 and 1 is called a Daniell integral.

Theorem 4.2.

Let be a lattice on and let be a monotone Daniell integral. Then there exists a measure on for which each is -measurable such that

(4.1)

for all .

Proof.

First note that, for any ,

For we say that a sequence suits if the monotonically increase and

Define

Then is a measure on . Indeed, and is monotonic. If for each and suit then

suits and

Next we show that each is -measurable. By creftype 2.4, it suffices to show, for every and that, for

we have

Suppose suit and let

Then for all and

Then suit and suit . Therefore

Finally we show that

(4.2)

First suppose that , suit and satisfies . Then monotonically increase to and so

Consequently,

(4.3)

Now fix and for each let . For a moment fix . Then, for each ,

(4.4)
(4.5)

and

(4.6)

Note that, for any , suits

and so

(4.7)

By Eqs. 4.6 and 4.4 and Eq. 4.5 respectively,

(4.8)

Finally by (4.3),

(4.9)

Combining Eqs. 4.8, 4.9 and 4.7 gives

Summing from 1 to gives

Since monotonically increases to ,

Therefore, since monotonically decreases to 0, increases to and so this gives (4.2).

Note that, if , then and so

Observation 4.3.

For any measure satisfying the conclusion of Theorem 4.2, the value of , for and , is uniquely determined by the values of on .

Proof.

For any and , observe that the functions

converge pointwise to the characteristic function of and are bounded above by . Therefore, by (4.1) and the dominated convergence theorem,

The Banach–Alaoglu theorem (see creftype 4.8) motivates us to consider representations of Borel measures by elements of the dual of a Banach space, namely of . To use Theorem 4.2, we must upgrade the pointwise convergence in the hypotheses to uniform convergence in . Recall that is the set of all continuous functions on with compact support, and that pointwise monotonic convergence in implies uniform convergence (see creftype 4.4). Therefore, for any monotonic , Theorem 4.2 produces a measure that represents . If is a metric space, then all compact sets are measurable with respect to . We next show, on locally compact spaces, how to obtain a Borel measure.

Lemma 4.4.

Let be a locally compact metric space and suppose that is a finite measure on . There exists a unique Radon measure on that agrees with on .

Proof.

Let be the set of open subsets of . For each define

Since is monotone, for any , , for the interior of .

Since is locally compact, for any and with , there exists with and

Consequently,

(4.10)

Also, since is a measure and is a metric space,

(4.11)

For each define

Note that both and are monotone and give value 0 to the empty set. To see that is a measure, let and let be open. We must show that

It suffices to show that

By (4.10), for any , there exists with compact closure such that

Since has compact closure, it is contained in the union of finitely many . Therefore, it suffices to show that is finitely sub-additive.

Let and be compact. Let and

and

Then are closed subsets of with and and . Since is finitely sub-additive,

and so

Therefore, is finitely sub-additive by induction. As shown above, this implies that is a measure.

To see that is a Borel measure, if are separated, then there exist separated open sets , . Since is finitely additive, and so is additive on separated sets. By construction, is Borel regular. Also, (4.10) shows that open sets are inner regular by compact sets. Combining this with outer regularity by open sets shows that is a Radon measure.

To see that agrees with on , note that for any and with ,

and so . For the other inequality,

by (4.11).

Finally, if are two Radon measures that agree with on , then for any open ,

Since are both Borel regular, they must agree. ∎

Theorem 4.5 (Riesz representation theorem).

Let be a locally compact metric space and let be monotone. Then there exists a unique Radon measure such that

(4.12)
Proof.

Observe that is a monotone Daniell integral on . By Theorem 4.2, there exists a measure on for which (4.12) holds with replaced by . By Lemma 4.4, there exists a Radon measure that agrees with on . For any , the value of

is determined by the value of on compact sets, and so

so that (4.12) holds. creftype 4.3 implies that, on a locally compact space, the measure of any compact set is uniquely determined by (4.12). Thus any Radon measure satisfying (4.12) is uniquely determined. ∎

The Riesz representation theorem allows us to identify the set of finite Radon measures on a locally compact space with the set of monotonic elements of . For a Banach space, a sequence weak* converges to if for every . In , this translates to the following.

Definition 4.6.

Let be a locally compact metric space. A sequence of finite Radon measures on weak* converges to a finite Radon measure if, for every ,

By the Banach-Alaoglu theorem (see creftype 4.8), the unit ball of is weak* compact. Since the weak* limit of a sequence of monotonic operators on is monotonic, we have the following.

Theorem 4.7.

Let be a locally compact metric space and a sequence of Radon measures with uniformly bounded total measures. There exists a finite Radon measure on and subsequence that weak* converges to .

4.1. Exercises

Exercise 4.1.

Let be a lattice of functions on . For any show that

Exercise 4.2.

Let be a lattice on and a monotone Daniell integral on . Give an example to show that there may be more than one measure satisfying (4.1) (recall creftype 1.9). Compare to creftype 4.3.

Exercise 4.3.

Let be a lattice on and a Daniell integral on . Define on by

and

  1. There show that are monotone Daniell integrals on .

  2. If with then and so

  3. Deduce that

Exercise 4.4.

Let be a compact metric space and suppose that monotonically increase to . Show that uniformly.

Exercise 4.5.

Let be the set of all such that exists and define by . Let be the set of all bounded . Equip and with the supremum norm.

  1. Observe that is linear and continuous on and hence can be extended by the Hahn-Banach theorem to a linear and continuous element of . (Such an extension is called a Banach limit.)

  2. Show that any finite Borel measure on is a convergent sum of Dirac masses.

  3. Hence show that there is no Borel measure on such that .

Exercise 4.6.

Adapt the previous exercise to show that the Riesz representation theorem is false in non locally compact metric spaces.

Exercise 4.7.

Let be a metric space. Show that any is a Daniell integral.

  1. Let let be obtained from creftype 4.3.

  2. Show that .

  3. By Theorem 4.5, any can be identified with two measures and and hence with a signed measure (recall creftype 3.6). Show that this identification is an isometric isomorphism.

Exercise 4.8.

Let be a separable Banach space and a countable dense subset of . Suppose that satisfy for some .

  1. Show that there exists a subsequence and a such that for each .

  2. Deduce that, for any , .

That is, closed and bounded subsets of are weak* compact.

5. Fubini’s theorem

Definition 5.1.

Let be measures on sets respectively and define the set of rectangles to be

The product measure on is defined by

where the infimum is taken over all countable collections of rectangles with

This is a measure, see creftype 5.1.

Lemma 5.2.

Let be sets and measures on respectively. Then is equivalently defined by the formula

where the infimum is taken over all disjoint countable collections of rectangles with

Proof.

If are -measurable and are -measurable then

is a decomposition into disjoint rectangles. Since and are and measurable respectively,

Thus the two formulae agree. ∎

Theorem 5.3 (Fubini’s theorem).

Let be sets and -finite measures on respectively.

  1. If is -measurable and is -measurable then is -measurable and

  2. If is -measurable then

    is -measurable for -a.e. , is -measurable;

    is -measurable for -a.e. , is -measurable; and

  3. If is -measurable or is integrable then

Proof.

Note that (3) follows from (2) by the monotone convergence theorem. We prove the theorem for finite .

To begin, let

Let be the set of such that is -measurable and for any define

Observe that is monotonic in .

If then is -measurable and . If with

a disjoint union, then

is a countable sum of -measurable functions and so . Moreover,

(5.1)

Thus, for any ,

(5.2)

To prove (1), let be -measurable and be -measurable. By definition, and, since is monotonic, whenever . Thus, by (5.2),

Let and . Observe

are disjoint members of . Therefore, by (5.1) and (5.2),

When taking the infimum over all such , the left hand side converges to , and so is -measurable. This also implies that all elements of are measurable.

Let and suppose that are such that . Since the intersection of any two rectangles is a rectangle,

for each . Moreover, monotonically decreases to . Let . Note that, for each , monotonically decreases to , so that . Since are finite, the monotone convergence theorem implies that and hence . Since , the monotonicity of implies .

To prove (2), if is -measurable, then . By (5.2) there exists with . That is, for -a.e. , and hence the first conclusion of (2). Moreover, , which concludes the proof. ∎

5.1. Exercises

Exercise 5.1.

Let be sets and measures on respectively. Prove that is a measure on .

Exercise 5.2.

Let be separable metric spaces. Show that

Exercise 5.3.

Let be separable metric spaces and finite Borel measures on respectively. Show that is the unique countably additive set function on satisfying for all .

Exercise 5.4.

Prove Fubini’s theorem and Exercises 5.2 and 5.3 for -finite measures .

Exercise 5.5.

Show that Theorem 5.3 (3) may fail if

  1. is measurable but not integrable; Hint: consider the counting measure on . Exploit the “identity”

  2. is measurable but is not -finite. Hint: consider , the counting measure on .

Prove Theorem 5.3 (3) for integrable, even if are not -finite.

Exercise 5.6.

Note in Theorem 5.3 (2) we must exclude a set of measure zero. Indeed, if is a Vitali set, note that is -measurable.

Exercise 5.7.

Let be a separable metric space and a Borel function. Prove that

Hint: consider

Exercise 5.8.

A measure on a metric space is uniformly distributed if there exists a such that for all and . Let be uniformly distributed Borel regular measures on a separable metric space (with functions and respectively). Let be open.

  1. Observe that, for any ,

    for every .

  2. Deduce that

  3. Deduce that

Deduce that for some .

6. Covering theorems

We will use to denote the closed ball in a metric space centred at with radius . Since the centre and radius of a ball are not uniquely defined by its elements, formally by a “ball” we mean a pair , but in practice we mean the set of its elements.

Lemma 6.1 (Vitali covering lemma).

Let be a metric space and an arbitrary collection of closed balls of uniformly bounded radii. There exists a disjoint sub-collection such that any intersects a ball with

In particular,

Here, denotes the ball with the same centre as and 5 times the radius.

Proof.

For each let

Since the balls in have uniformly bounded radii, the are empty for all , for some . Let be a maximal disjoint sub-collection of . That is, the elements of are disjoint elements of and if , there exists a with . (In general such a maximal collection exists by Zorn’s lemma. See also creftype 6.1.) Let be a maximal collection such that is a disjoint collection. Repeat this for each , obtaining a maximal collection such that is a disjoint collection, and set .

Now suppose that , say . Then by construction there exists for some with . In particular, .

The final statement of the lemma follows from the triangle inequality. ∎

Definition 6.2.

Let be a metric space and . A Vitali cover of is a collection of closed balls such that, for each and each , there exists a ball with and .

Proposition 6.3.

Let be a metric space, and suppose that is a Vitali cover of . Then there exists a disjoint such that, for every finite ,

In particular, if is countable (for example, if is separable), then

for each .

Proof.

Note that we may suppose consists of balls with uniformly bounded radii. Let be a disjoint sub-collection of obtained from Lemma 6.1. If is finite then

is closed. Therefore, if , since is a Vitali cover of , there exists with such that . However, must intersect some with , and so . That is, belongs to

as required. ∎

Definition 6.4.

A Borel measure on a metric space is a doubling measure if there exists a such that

for all balls .

Remark 6.5.

Note that, for any ,

Lebesgue measure is a doubling measure.

Theorem 6.6 (Vitali covering theorem).

Let be a doubling measure on a metric space and let be a Vitali cover of a set . There exists a countable disjoint such that

Proof.

First note that it suffices to prove the result for bounded, say is contained in some ball . We may also suppose that each is a subset of .

Let be a disjoint sub-collection of obtained from Proposition 6.3. Note that is countable. Indeed, for each , at most balls can satisfy .

Enumerate . Since the are disjoint subsets of ,

By the conclusion of Proposition 6.3,

for each . Since is doubling, for each and so

as required. ∎

Definition 6.7.

Let be a Borel measure on a metric space and -measurable. Suppose that for all and all .

Define the Hardy–Littlewood maximal function of by

By creftype 6.3, the maximal function is a Borel function.

Theorem 6.8 (Hardly–Littlewood maximal inequality).

Let be a doubling measure on a metric space . There exists a such that, for any and ,

(6.1)
Proof.

For , let

and, for each , let be those for which there exists such that

(6.2)

Similarly to creftypecap 6.3, each is a Borel set. Moreover, the monotonically increase to . For a moment fix . Let be the collection of balls with and that satisfy (6.2) and let satisfy the conclusion of Lemma 6.1. Then

In particular, (6.1) holds for . ∎

Theorem 6.9 (Lebesgue differentiation theorem).

Let be a doubling measure on a metric space and with . For -a.e. ,

as . Such an is called a Lebesgue point of .

Proof.

First note that the theorem is true if is continuous.

Fix and let be continuous with

(such a exists by creftype 3.1). Let

so that .

If

then by Theorem 6.8,

Moreover, if ,

for all . In particular, since is continuous at ,

Therefore, if ,

We are now done; repeat the above for a countable collection of . The corresponding monotonically decrease to a set of measure zero. The set of that does not belong to infinitely many of the has full measure, and for such an ,

Corollary 6.10.

Let be a doubling measure on a metric space and let be -measurable with . Then

equals 1 for -a.e.  and 0 for -a.e. . Such an for which the limit equals 1 is called a density point of .

6.1. Exercises

Exercise 6.1.

Let be a separable metric space. Show that for any collection of balls, there exists a maximal disjoint sub-collection.

Exercise 6.2.

Show that the covering Lemma may not be true if the radii are not uniformly bounded.

Exercise 6.3.

Let be a finite Borel measure on a metric space and let such that is a decreasing sequence. Let denote the open ball centred on with radius .

  1. Show that, for any ,

    decreases to the empty set.

  2. Deduce that is lower semi-continuous.

  3. Give an example to show that may not be continuous.

  4. Show that, for any ,

  5. Deduce that, for any , is a Borel function.

  6. Show that the Hardy–Littlewood maximal function is equivalently defined by taking the supremum over all rational .

  7. Deduce that the Hardy–Littlewood maximal function is a Borel function.

Exercise 6.4.

Let be a doubling measure on a metric space and let with . Suppose that there exists a -measurable with . Show that

for -a.e. .

Exercise 6.5.

Prove that on , for some . There are two ways to prove this (recall creftype 1.5 Items 1 and 2).

Let be two finite Borel measures on a set with and suppose that is doubling. Show that the Radon–Nikodym derivative of with respect to is given by

for -a.e. .

Exercise 6.6.

For , show that does not have finite integral.

7. Differentiability of Lipschitz functions

The regularity of a Lipschitz function is very interesting. Of course, Lipschitz functions are continuous, but they may not be differentiable everywhere. However, it is quite easy to convince yourself that they cannot be non-differentiable on quite a large set. The question to quantify how large the non-differentiability set of a Lipschitz function can be was one of the motivating questions of Lebesgue’s development of measure theory.

Definition 7.1.

Let . The total variation of , is defined by

where the supremum ranges over all .

If , is said to have bounded variation (BV).

Definition 7.2.

A function is absolutely continuous (AC) if for any there exists a such that, for any intervals with , we have .

Note that AC functions are BV, and Lipschitz functions are AC (see creftype 7.1). Also, if is BV then and are non-decreasing. If is AC then so are and , see creftype 7.2.

Theorem 7.3 (Lebesgue).

Let be absolutely continuous. Then is differentiable almost everywhere. Moreover, for any ,

Proof.

By creftype 7.2 it suffices to assume that is non-decreasing. In this case define a measure on using the Carathéodory construction with the set of compact intervals and . This defines a finite Borel measure such that for all intervals . Indeed, for any , we may cover by finitely many intervals of width , showing

The reverse inequality holds because is non-decreasing.

Note that . Indeed, given , let be given by the definition of being absolutely continuous. If , we may cover by countably many closed intervals such that . In particular and hence . Therefore,

with .

By the Lebesgue differentiation theorem, for any Lebesgue point of ,

Theorem 7.4 (Rademacher).

Any Lipschitz is differentiable almost everywhere.

Proof.

For notational simplicity, we prove the case .

For each , is a Lipschitz function and so is differentiable -a.e. That is, for every , exists for -a.e. . By Fubini’s theorem, exists -a.e. Similarly, exists almost everywhere too.

Fix . For and let

These are Borel sets. Further, for , if

then for sufficiently large . That is,

is a set of full measure.

Fix and . Let be a density point of . Let such that

for all . In particular, for every there exists with

Now let and . Set , and with

Also let lie on the same vertical line as such that have the same vertical component as do . Then, since ,

(7.1)

Since is Lipschitz,

(7.2)

Since ,

(7.3)

Since is Lipschitz,

(7.4)

By combining Eqs. 7.4, 7.3, 7.2 and 7.1,

This is true for all with and for any density point of the full measure set . That is, for -a.e . Taking a countable intersection over concludes the proof. ∎

7.1. Exercises

Exercise 7.1.

Prove that Lipschitz functions are AC and that AC functions are BV.

Exercise 7.2.

Let be BV. Show that and are non-decreasing. If is AC then show that and are AC.

Exercise 7.3.

Show that any monotonic is continuous except at countably many points.

Exercise 7.4.

In this exercise we will show that monotonic functions are differentiable almost everywhere.

Let be non-decreasing. For each let

Observe that the set of where is not differentiable at is the countable union, over , of the sets

We now fix .

  1. Let

    Note that satisfies the hypotheses of the Vitali covering theorem (recall Theorem 6.6). Let be a disjoint sub-cover obtained from the Vitali covering theorem with respect to Lebesgue measure and let . Prove that

    Note: this is the step where we require to be monotonic.

  2. Similarly, prove that .

  3. Deduce that is differentiable almost everywhere.

  4. Deduce that a BV function is differentiable almost everywhere.

However, BV functions do not satisfy the fundamental theorem of calculus:

Exercise 7.5.

Recall the definition of the Cantor set from creftype 1.6. Define the Cantor function as follows. For each , define by

Show that the converge uniformly on to a monotonic, continuous function . For each , show that .

Thus, is monotonic and hence BV, has derivative almost everywhere, but does not satisfy the fundamental theorem of calculus.

Exercise 7.6.

In lectures we proved that the derivative of any AC function is an absolutely continuous measure. Prove the converse: for any finite, absolutely continuous measure on , show that

defines an absolutely continuous function.

Up to now, we have considered points where functions are differentiable. We now consider points of non-differentiability (which are much more interesting).

Exercise 7.7.

Show that the Cantor function is not differentiable at any point of the Cantor set.

Exercise 7.8.

Let satisfy .

  1. For each , iteratively construct a countable collection of open intervals such that, for each ,

    • is contained in the union of ;

    • for every there exists with ;

    • for each ,

  2. Let

    the “limsup” of the ( is the set of points that are contained in infinitely many intervals from the ). In particular, .

    For each , let be the largest for which there exists with . Define if is even, otherwise. Finally, for each define

    Show that is Lipschitz, monotonic, and not differentiable at any point of . Hint: show that and for each .

Part II Some topics in Geometric Measure Theory

8. Hausdorff measure and densities

Recall the definition of Hausdorff measure from creftype 1.5.

We are interested in the measure , for some -measurable set with . In particular, we require some counterpart to the Lebesgue density theorem, but, of course, may not be locally finite.

Definition 8.1.

Let be a metric space, and . The upper and lower Hausdorff densities of are

and

Lemma 8.2.

Let be a metric space, and with . Then

for -a.e. .

Proof.

The set of with is a countable union countable of the sets

Thus, for the first inequality, it suffices to show that for all .

Fix . We may cover by sets such that, for each , , and

For each let and set . Then

Since is arbitrary and , this implies , as required.

For the second inequality, since is Borel regular (see creftype 8.2), it suffices to assume that is Borel. As before, given , it suffices to prove that

satisfies . Fix and let be open with

(which exists by the outer regularity of the measure ). Let be the collection of balls centred at a point of with such that and

(8.1)

This is a Vitali cover of . Let be obtained from Proposition 6.3.

Since , is separable (see creftype 8.4) and so is countable and the conclusion of Proposition 6.3 states that

for each . Since for each , the and may be used to estimate . For each we obtain

where the second inequality follows by (8.1). Since the are disjoint and , the second term converges to 0 as . Since the are subsets of we obtain

Since is arbitrary, this implies and hence , as required. ∎

Lemma 8.3.

Let be a metric space, and let be -measurable with . Then

for -a.e. .

Proof.

It suffices to show that, for , the set

satisfies . Fix . Since is -measurable, is Borel regular. Therefore, since , there exists an open with

For each and there exists a ball centred on with such that

By Lemma 6.1 there exists a disjoint collection of such balls such that

Since , is separable and each of these balls contains a point of , is countable. Therefore

Since are arbitrary, this completes the proof. ∎

8.1. Exercises

Exercise 8.1.

Let be a Vitali set as constructed in creftype 1.8.

  1. Show that, for any Borel , .

  2. Deduce that and hence, if is a Borel set with

    then .

  3. Hence show that .

Note however that we cannot deduce the value of from our construction in creftype 1.8. Indeed, for any , that construction may produce a .

Exercise 8.2.

Let be a metric space and .

  1. Show that is Borel regular. Hint: first show that in the definition of , we may take to be the collection of closed sets.

  2. We are usually interested in for some . Show that for any , is a Borel measure.

  3. Now assume that is -measurable with . Show that is Borel regular. Hint: show that there exist Borel sets with .

  4. Show that may not be Borel regular if is not measurable. Hint: consider creftype 8.1.

Exercise 8.3.

In this exercise we construct the four corner Cantor set. Let . Let be the “four corners” of of side length 1/4. That is

Inductively, is constructed by taking the four corners of side length of all the squares of . Finally let , a compact set. Show that .

Exercise 8.4.

For let be a metric space with . Show that is separable.

Exercise 8.5.

Show that Lemmas 8.3 and 8.2 may be false if has only -finite measure.

9. Rectifiable sets and approximate tangent planes

Rectifiable sets are the measure theoretic counterpart to manifolds.

Definition 9.1.

A -measurable set is -rectifiable if there exist Lipschitz such that

We will show that -rectifiable sets possess a unique approximate -dimensional tangent plane at almost every point.

Given , and define the cone around centred at with aperture as

Definition 9.2.

Let and . A is an approximate tangent plane to at if

(9.1)

and, for every ,

(9.2)

Rademacher’s theorem gives a candidate for the approximate tangent plane to a rectifiable set. There are three steps required to prove that the derivative is indeed an approximate tangent plane: show that the derivative has full rank at almost every point; prove the density condition (9.1); and show that the sets from other parametrisations of the rectifiable set do not destroy the approximation by a tangent plane at almost every point.

The second and third steps follow from the results of the previous section. For the first step we use the following.

Lemma 9.3 (Easy Sard’s theorem).

If is Lipschitz then

Proof.

Let . Fix , and let

For let

Then for sufficiently small ,

Since , the set on the right hand side can be covered by cubes of side length .

Since is covered by balls of the form , there exists a disjoint collection of balls such that is covered by the union of the . Then

By the previous argument, each factor of the right hand side is covered by cubes of side length . Thus

However, the are disjoint subsets of and so

Since is arbitrary, this implies that and hence . Taking a countable union over completes the proof. ∎

Lemma 9.4.

Let be Lipschitz and . Suppose that there exists a such that, for each ,

Then has a unique approximate tangent plane at almost every point.

Proof.

By Lemma 9.3, we may suppose for every . By Lemma 8.2, we may suppose for every . Fix and . There exists such that

for all . Moreover, if then

That is, if and with and ,

Therefore, is an approximate tangent plane to at .

This approximate tangent plane is unique at any density point of . Indeed, if , let and let be such that . Since and is a density point of , for sufficiently small there exists such that and

In particular, is not an approximate tangent to at . ∎

Theorem 9.5.

Let be -rectifiable with . Then for -a.e. , has a unique approximate tangent plane at .

Proof.

Let be one of the Lipschitz functions as in the definition of a rectifiable set and let . It suffices to prove that has a unique approximate tangent plane at for -a.e. .

By Lemma 9.3, we may suppose that for all . Fix such an and let . There exists such that

for all . In particular, by the triangle inequality,

Therefore, the sets

are Borel and monotonically increase to as . Therefore it suffices to prove the result for -a.e.  in some fixed . Cover by finitely many balls of radius . It suffices to prove the result for -a.e.  in some fixed .

However, satisfies the hypotheses of Lemma 9.4 and so has a unique approximate tangent at almost every point. To see that this tangent is a unique approximate tangent to at almost every point, we simply use Lemma 8.3: for -a.e. , . ∎

In Theorem 10.10 we will see that the converse to Theorem 9.5 holds.

For , write for the orthogonal projection onto and equip with the metric . We will consider on an element of .

Lemma 9.6.

Let be Lipschitz and let satisfy . For suppose that there exists an invertible linear such that, for all ,

Then for any with , .

Proof.

For any ,

and so, if ,

Thus has Lipschitz inverse on and hence by creftype 1.5. ∎

Corollary 9.7.

Let be -rectifiable with . Then there exists such that for all with .

Remark 9.8.

The set of that satisfy the conclusion of Corollary 9.7 is very large; try some examples in reasonable dimensions.

Proof.

Since is rectifiable with , there exists a Lipschitz with . In particular, satisfies . By Lemma 9.3, for -a.e. , is injective.

Fix For , the set of invertible with may be covered by countably many sets of diameter . Varying , we see that almost all of is covered by countably many sets of the form

Moreover, each of these sets may be covered by countably many sets of the form

Finally, these sets may be covered by countably many sets of diameter . Therefore, for each there exists and invertible such that, for all ,

and .

Since , there exists with . Then satisfies the hypotheses of Lemma 9.6 and so for all with . Let . Repeat this for each with . The set is compact and so we may suppose that . The only for which satisfy and hence for each . That is, as required. ∎

9.1. Exercises

Exercise 9.1.

Let be a metric space, and -Lipschitz. Define by

  1. Show that is an -Lipschitz extension of to . This is called the McShane–Whitney extension theorem

  2. If is -Lipschitz, show that there is a -Lipschitz extension of to .

  3. The following example shows that the vector valued extension cannot have the same Lipschitz constant in general: Let

    and define

    Show that is 1-Lipschitz but has no 1-Lipschitz extension to .

  4. However, the Kirszbraun extension theorem states that any Lipschitz map between any two Hilbert spaces may be extended whilst preserving the Lipschitz constant.

Exercise 9.2.
  1. Let be -rectifiable. Show that is -finite.

  2. Show that Theorem 9.5 may not be true if does not satisfy .

10. Purely unrectifiable sets

Definition 10.1.

A -measurable set is purely -unrectifiable if, for all -rectifiable , .

Lemma 10.2.

The four corner Cantor set is purely 1-unrectifiable.

Proof.

Observe that the coordinate projections of have measure zero (see creftype 10.1). If there existed a rectifiable with , then is a 1-rectifiable set of positive measure and hence, by Corollary 9.7, one of the coordinate projections must have positive measure, a contradiction.

For a second proof see creftype 10.2. ∎

Lemma 10.3.

Let be -measurable with . There exists a decomposition with -rectifiable and purely -unrectifiable.

Proof.

Let

Since , . Let be -rectifiable with . Then is -rectifiable and is contained in . Therefore

and so . Then is purely -unrectifiable. Indeed, if is Lipschitz, , is -measurable and so

We now state a very important theorem on the structure of purely unrectifiable sets. It requires the notion of a natural measure on that is invariant under the action of . There are several ways to construct this measure. The simplest is to consider as a (compact) metric space equipped with the metric

Then is given by (a scalar multiple of) . We will not discuss the specific details of this measure. When , we may identify with . In this case, is simply .

Theorem 10.4 (Besicovitch–Federer projection theorem).

Let be purely -unrectifiable with . Then, for -a.e. ,

Conversely, if is purely -unrectifiable with , for -a.e. ,

Remark 10.5.

The converse statement is given by Corollary 9.7.

We will prove the projection theorem for and , which was proved by Besicovitch. First we prove some preliminary geometric properties of purely unrectifiable sets.

Lemma 10.6.

Let , and . Suppose that, for every ,

Then is -rectifiable.

Proof.

Since may be divided into countably many sets of diameter at most , we may suppose . In this case, has Lipschitz inverse on . Indeed, if then and so

Therefore, is covered by a Lipschitz image of . ∎

Lemma 10.7.

Let be purely -unrectifiable, , and . If

(10.1)

for every and then

for every .

Remark 10.8.

Note that this is certainly not true for a rectifiable set ; the first cone may be empty for every .

Proof.

For a fixed , we may suppose . By Lemma 10.6, we may suppose that

for every . For every let

so that for all . Pick with and let be the cylinder

We claim that

(10.2)

(Draw a picture!) Suppose does not belong to the first set. Then

where the penultimate inequality follows because and . Therefore

That is, belongs to the first set in (10.2).

By (10.2) and (10.1),

We apply Lemma 6.1 to the balls

with . This gives countably many , for which these balls are disjoint, and

Therefore

But, the are disjoint subsets of and so the final sum is bounded above by . ∎

Corollary 10.9.

If is purely -unrectifiable with then for every , every and -a.e. ,

Proof.

For a fixed , this is immediate from the fact that almost everywhere. To obtain the conclusion for all , note that the conclusion is determined by a countable dense set of . ∎

Theorem 10.10.

Let satisfy and suppose that, for -a.e. , has a unique approximate tangent plane at . Then is -rectifiable.

Proof.

By Lemma 10.3, there exists a decomposition , where is -rectifiable and is purely -unrectifiable. We must show that . Note that, by applying Lemma 8.3 to , we see that the approximate tangent plane to at is also an approximate tangent plane to for -a.e. .

It suffices to show, for a fixed , that the set of whose approximate tangent plane lies in has measure zero. Suppose not. Then, for any , there exists an such that the set of those with

has positive measure. Fix an . Since , for every we have

Thus, for and ,

If , Corollary 10.9 implies , a contradiction. ∎

We now prove the Besicovitch projection theorem [1]

Theorem 10.11.

Let be purely 1-unrectifiable with . Then for -a.e. ,

We follow the presentation of Orponen [2].

From Corollary 10.9, we see that a purely unrectifiable set has many radiating out of almost every point in all directions at almost every point. We now precisely describe two ways in which this can occur.

Notation 10.12.

Let and . For let be the half line and for , let be the cone . For let be those for which

That is, contains another point of . Also let , the directions that contain other points of arbitrarily close to . For , we let be those for which .

For let be those for which there exists and an interval with and such that

That is, the density of in the cone at scale is very high, compared to the length of . Also let . For , we let be those for which .

The main step in proving Theorem 10.11 is the following.

Proposition 10.13.

Let be purely 1-unrectifiable with . For -a.e. , .

Before proving Proposition 10.13, we will demonstrate how it is used to prove Theorem 10.11.

Lemma 10.14 (Special case of the coarea formula).

For any and any compact ,

In particular, if then for any , .

Proof.

Since is compact,

is a Borel function. Indeed, if, for ,

then monotonically increases to as . Since is compact, the are lower semi-continuous. Therefore, by the monotone convergence theorem, it suffices to bound the integral of each .

Fix and cover by sets with such that

Note that

Therefore

as required. ∎

Lemma 10.15.

Let be -measurable with . Then for any , .

Proof.

Fix . For any and there exists an , and an interval with such that

Apply Lemma 6.1 to the intervals to obtain a disjoint collection such that

Therefore

where the final inequality follows from the disjointness of the sets , . Since this is true for all , . ∎

Proof of Theorem 10.11 using Proposition 10.13.

By the inner regularity of , it suffices to prove the result for compact . By definition, we have

Proposition 10.13 implies that the left hand expression has -measure zero and so Fubini’s theorem implies that, for -a.e. , . Therefore, by Lemmas 10.14 and 10.15, for -a.e. . ∎

Proof of Proposition 10.13.

Fix and which satisfies the conclusion of Corollary 10.9. That is,

(10.3)

for every interval . It suffices to show that . In fact, we will show that, for any ,

from which the result follows by the Lebesgue density theorem.

To this end, fix with . Then for all sufficiently small intervals with ,

(10.4)

Fix such an . By Eq. 10.3, there exists with

(10.5)

Note that (10.4) implies

(10.6)

We will prove that

(10.7)

Since is any sufficiently small interval containing , this implies

as required.

By (10.6) we may cover by disjoint intervals with

(indeed, the disjointness of the intervals is shown in creftype 10.6). By the definition of , we know that

(10.8)

Let be those with

(10.9)

Note that by (10.5) and (10.8),

That is, the cones associated to cover a large proportion of . Moreover, , because if then for some which satisfies the definition of .

Thus, to show (10.7), it would be enough to bound from below by a multiple of . But this is not necessarily true: the intervals that form could be extremely thin compared to . To accommodate this, we enlarge the intervals with as follows. For each enlarge until (10.9) becomes an equality or until intersects another . If the first possibility occurs then we still have . If the second possibility occurs then we merge the two intervals; since both sides of (10.9) are linear in , and the boundary of each contains no points of , (10.9) remains true after the merge. By the same reasoning, both sides of (10.9) are continuous under expanding and consequently one of the two possibilities must occur.

This results in a disjoint collection of intervals for which (10.9) is an equality. Moreover,

and, by construction, each . Therefore

10.1. Exercises

Exercise 10.1.

Prove that the coordinate projections of the four corner Cantor set have Lebesgue measure zero.

Exercise 10.2.

A second proof that the four corner Cantor set is purely 1 unrectifiable.

Suppose that satisfies .

  1. Prove that there exists that is a density point of such that .

  2. Therefore, for sufficiently small , is approximated by a line segment of length that is mostly contained in . Derive a contradiction.

Exercise 10.3.

Prove that the decomposition given in Lemma 10.3 is unique up to -null sets.

Exercise 10.4.

Think about Propositions 10.13 and 10.9 in regard to the four corner Cantor set.

Exercise 10.5.

Let be compact. Show that

are Borel subsets of .

Exercise 10.6.

Show that in the definitions of and we may suppose the covering intervals, respectively sets, may be chosen to be disjoint.

Exercise 10.7 (Open problem).

For , does there exist a compact purely 1-unrectifiable with that intersects every line in at most points?

References

  • [1] A. S. Besicovitch (1939) On the fundamental geometrical properties of linearly measurable plane sets of points (III). Math. Ann. 116 (1), pp. 349–357. External Links: ISSN 0025-5831, Document, Link, MathReview Entry Cited by: §10.
  • [2] T. Orponen (2018) Geometric measure theory. Online lecture notes. Cited by: §10.