Sobolev spaces on Euclidean space
\setkeysGrotunits=360
Contents:
- 1 Introduction
- 2 Definition of Sobolev spaces
- 3 Basic properties of Sobolev spaces
- 3.1 Banach and Hilbert space structure of Sobolev spaces
- 3.2 Sobolev spaces are the completion of the space of smooth functions in the Sobolev norm (the Meyers-Serrin theorem)
- 3.3 The dual space of a Sobolev space
- 3.4 Positive distributions can be represented by measures (the Riesz representation theorem)
- 4 Embedding and compactness theorems
- A Notation and basic definitions
- B Basic results from measure theory
- References
1 Introduction
The purpose of these notes is to outline the basic definitions and theorems for Sobolev spaces defined on open subsets of Euclidean space. Of course, there are already many good references on this topic, and, rather than duplicate this here, instead the goal is to give examples where possible to illustrate the theory, and to orient the reader towards the different approaches contained in the literature. In addition, there is an appendix containing some basic results from measure theory (again, this contains examples and references to some of the literature on the subject).
There are a number of more advanced topics that have been left to future versions of these notes; for example, complete proofs of the embedding and compactness theorems (as well as examples where embeddings don’t exist), the chain rule and the behaviour of weak derivatives under co-ordinate transformations. Good references for this material include the book [1] by Adams (a classic on the subject), and Ziemer’s book [16]. Sobolev spaces on manifolds and their use in gauge theory would also be good topics for an expanded version of these notes. Future versions of these notes will also contain more examples.
Notation. First-order partial derivatives are denoted or . Higher-order partial derivatives use the standard notation for multi-indices (see [4, Appendix A]): Given a multi-index we write for the order of , and define the order partial derivative by
2 Definition of Sobolev spaces
This section contains all of the necessary definitions needed to define Sobolev spaces on open subsets of . In order to get a feel for distributional derivatives and Sobolev spaces, some basic examples are given throughout the section.
The approach taken in these notes is to follow the historical definition of Sobolev spaces. First, in this section, we define the distributional and weak derivatives, and then define Sobolev spaces in terms of these. Later on, in Section 3.2, we prove the Meyers-Serrin theorem, which says that Sobolev spaces are the completion of the space of smooth functions in the Sobolev norm.
2.1 Distributions and test functions
Let be open and non-empty, and let denote the space of smooth complex-valued functions with compact support in .
Definition 2.1.
The space of test functions on , denoted , is the locally convex topological vector space (more precisely the LF-space) consisting of all the functions in , with the following notion of convergence: A sequence converges in to the function if and only if there is some fixed compact set such that the support of is in for all , and that uniformly for each .
Remark 2.2.
The notation is used to emphasise the topology on described above.
Note that the definition does not imply that the constants from the uniform convergence are independent of .
To see that is a locally convex topological vector space, indeed, let us construct a family of seminorms which induces the topology on . To this end denote first by the set of compact exhaustions of , which means the set of all families such that is compact, and for all . For every compact exhaustion and every pair and of sequences of natural numbers denote then by the map defined by
Note that, the sum in this formula is always finite since the support of is compact.
It is straightforward to check that is a seminorm on and that the family defines a locally convex topology on that exactly recaptures the convergence in as defined above (see [14, Chapter 13] for more details on the LF-space structure of ). We present this explicit description of a family of seminorms describing the locally convex topology on here, since we do not know of a reference to this in the literature.
Definition 2.3.
A distribution is a continuous complex-valued linear functional on the space of test functions . The space of distributions is the dual
Remark 2.4.
In the above definition, linearity simply means that if and , then
Continuity means that whenever a sequence converges in (in the sense of Definition 2.1) to , then as a sequence in .
The space is also equipped with a notion of convergence, defined as follows.
Definition 2.5.
A sequence converges in to if for every we have in .
Remark 2.6.
This is the usual notion of convergence in the topology on a dual space.
The following gives some examples of distributions (recall the definition of from Appendix A).
Example 2.7.
Given , the delta functional is the distribution
Clearly this is linear. To see that it is continuous, note that if in , then , and so .
The functional
is a distribution. (Note that since is continuous with compact support then it is also integrable.) Again, this is clearly linear. It is also continuous, since if in , then, by definition, there exists a fixed compact set such that . Therefore
and since uniformly on the compact set , then . Note that it is essential that has finite measure for this argument to work.
Given , let be the functional
(2.1) Note that Hölder’s inequality shows that , where denotes the (compact) support of . Therefore is always finite since implies that is finite, and so for all . As for the previous examples, clearly is linear, and it only remains to show that it is continuous. Note that if , then there is a fixed compact set with , and so
since and uniformly on . Therefore .
We will revisit this example later, since it appears in the definition of weak derivative in Section 2.2.
The last example above is an important one, it shows that there is a linear map given by (recall that all and spaces are defined to be equivalence classes of functions that are equal almost everywhere, and note that this map is well-defined on equivalence classes of functions in , since a.e. implies that for any test function ). In fact, since Hölder’s inequality shows that there is an inclusion for all , then there is also a map . The next theorem says that this map is injective.
Theorem 2.8.
Let be open, and let and be functions in . Suppose that the distributions and are equal, i.e. for all test functions . Then a.e. in .
Proof.
This is proved in [8, Theorem 6.5] using convolutions, however, for variety, here we give a slightly different proof. Firstly note that it is sufficient to prove the result for real-valued functions and , since we can take real and imaginary parts. Suppose that there exists a set whose Lebesgue measure is finite and non-zero, and which satisfies for all . Since the Lebesgue measure is Borel regular (see Lemmas B.20 and B.21) then we can assume without loss of generality that is compact. Define to be the subset such that , and again note that without loss of generality we can assume that is compact with non-zero measure. Define the constant
Now let be a collection of open sets such that
and for each , and
the Lebesgue measure of satisfies .
The existence of each is guaranteed since the Lebesgue measure is Borel regular. Now use Urysohn’s lemma (see Appendix A.3) to construct a smooth positive function such that for all , for all , and for all . Therefore
The last term in the above equation satisfies the estimate
and, since as , then
since the integral of a fixed measurable function is an absolutely continuous set function (see for example [15, Corollary 10.41]).
Therefore there exists an such that
which is a contradiction. Therefore almost everywhere. ∎
A consequence of this theorem is that the distribution associated to a function uniquely determines an equivalence class in . Therefore, the following definition makes sense.
Definition 2.9.
A distribution represents the function if
for all test functions . A function is represented by the distribution .
Theorem 2.8 shows that each distribution can represent at most one element of , i.e. the map given by is injective. The following example shows that not all distributions represent functions in , i.e. the map given by is not surjective.
Example 2.10.
Given any , let be the delta functional defined in Example 2.7. We claim that this does not represent any function in . To see this, suppose for contradiction that for some and every . Consider a sequence of bump functions such that for all satisfying we have
,
,
for all , and
.
Then, since has support in , dominated convergence shows that as , which contradicts for all .
Therefore the delta functional is an example of a distribution that cannot be represented by a function in . It can, however, be represented by a measure (see the measure in Example 3.28), and in Section 3.4 we will show that positive distributions can always be represented by measures (see Theorem 3.35).
2.2 Distributional derivatives and Sobolev spaces
Before defining Sobolev spaces, first we have to define the notion of the derivative of a distribution.
Definition 2.11.
Let be open, let , and let . The distributional derivative of is the distribution defined by
for all test functions . The distributional gradient, denoted , is the -tuple of distributions
If and both represent functions in (i.e. and for some ) then we say that is a weak derivative of , and write . In this case we say that the weak derivative of exists.
Remark 2.12.
Since the weak derivative is defined by the relation
then it is only defined up to equivalence almost everywhere.
The distributional derivative always exists for any multi-index , since the definition only involves differentiating test functions, which are smooth. Since partial derivatives of smooth functions commute, then distributional derivatives also commute, i.e.
As we will see in the examples below, the weak derivative does not always exist, and, in fact, may not even exist for any value of (in Example 2.19 we show that the step function is an example of such a function).
The following lemma shows that the weak derivative extends the notion of classical derivative of differentiable functions. It says that the distributional derivative of the distribution associated to a differentiable function is the distribution associated to the classical derivative of .
Lemma 2.13.
Let . Then for all we have
(2.2) |
Proof.
The proof simply involves applying the definitions and the integration by parts formula from Section A.2. ∎
Remark 2.14.
It is important to emphasise that is used to denote both the distributional derivative and the classical derivative in the statement of the lemma: is the distributional derivative of the distribution associated to the function , and is the distribution associated to the classical derivative .
It is an important exercise to think through the precise meaning of all of the statements above, to understand the distinction between a weak derivative and a distributional derivative, and to understand the meaning of each term in (2.2).
The next lemma shows that functions that are equal almost everywhere have the same distributional derivatives. As a consequence, when defining Sobolev spaces in Definition 2.16, we can define them as subsets of the and spaces (i.e. we consider equivalence classes of functions that are equal almost everywhere).
Lemma 2.15.
If almost everywhere, then as distributions.
Proof.
The proof is another straightforward application of the definition of distributional derivative. For any test function we have
Therefore, for all , and so as elements of . ∎
Now that we have developed the necessary machinery, we are ready to define Sobolev spaces.
Definition 2.16.
The Sobolev space is the space of equivalence classes of all functions such that the weak derivative exists and is in for all such that .
The Sobolev space is the space of all functions such that the weak derivative exists and is in for all such that .
The space has a norm given by
and we define to be the closure of the space in the topology induced by this norm.
For a compact subset we define the norm
where the weak derivatives are defined on .
Lemma 2.17.
The norm gives the structure of a normed linear space for .
Proof.
Recall that we have to check
The space has a unique element of zero norm, i.e. if and only if .
The norm is linear with respect to scalar multiplication, i.e. for all and .
The triangle inequality holds, i.e.
for all .
It is easy to check (2.2): since the result is true for , we have , and for all .
The weak derivative commutes with scalar multiplication, i.e. for all , and so we also have . Therefore (2.2) is satisfied by definition of the Sobolev norm.
The triangle inequality for follows from the definition of the Sobolev norm and the triangle inequality for (which is Minkowski’s inequality, see for example [15, Theorem 8.10]). ∎
It is worth recalling that can never be a normed linear space for , since the triangle inequality fails in this case. See for example the remark on p130 of [15], and also [15, Theorem 8.16]. For more discussion of spaces for , see [13, pp35-36].
Remark 2.18.
We will see later, in Section 3.1, that is a Banach space with this norm.
It is worth studying some examples of distributional and weak derivatives. The first example is the step function, for which the distributional derivative is the delta functional from Example 2.7. This is an important example, since it shows that the step function is not in any Sobolev space or for , because the delta functional cannot be represented by a function.
Example 2.19.
Let be the step function
Given a test function , consider the integral
(Recall that vanishes at infinity since it has compact support.) Therefore the distributional derivative of is the linear functional given by , i.e. is the delta functional . Example 2.10 shows that this cannot be represented by a function, and therefore the weak derivative of the step function does not exist, so the step function is not in or for any .
Example 2.20.
Let . To compute the weak derivative we first consider
(2.3) |
Let
and note that the previous calculation (2.3) shows that . Therefore the weak derivative of is the step function .
The next example generalises the method of the previous example to locally Lipschitz functions.
Example 2.21.
In this example we show that if is locally Lipschitz on then . Rademacher’s theorem shows that the partial derivatives of exist almost everywhere (see Corollary A.17), and the goal of this example is to show that these partial derivatives are equal almost everywhere to the weak derivative of in each co-ordinate direction.
For each compact set , let be the associated Lipschitz constant, i.e. for all we have
(2.4) |
(Note that this differs slightly from Definition A.14, however we can easily extend this to compact sets by taking an open cover of .)
Equation (2.4) implies that . Therefore the integral
is defined for any . To show that has a weak derivative, we need to show that there exists such that
for all test functions .
Let . Since is compact then there exists such that if then for all , and so is well-defined for small values of . Therefore
The next step involves using dominated convergence to interchange the order of integration and differentiation. Since this is a standard technique that is used in many examples then we include all of the details here. First note that since is smooth with compact support, then it is uniformly Lipschitz, and so the absolute value of the difference quotients is uniformly bounded by a constant (call it ) for . Since and the difference quotients have compact support, then
and so we can use dominated convergence to write
Changing variables, and recalling that the upper bound on was chosen so that for all , gives us
(2.5) |
(Even though may not be in for arbitrary , we do have for all . Since the support of is , then we can define
and therefore the integral in the above calculation makes sense.)
The quanitity is uniformly bounded for (since is locally Lipschitz), and so another application of dominated convergence gives us
(2.6) |
Rademacher’s theorem shows that for each , the partial derivative exists almost everywhere in , and, on the compact set it is bounded above by the Lipschitz constant . Let be a function defined on all of that is equal almost everywhere to . Therefore
and so we have shown that
Therefore the weak derivative exists and is equal almost everywhere to . Since almost everywhere on each compact set , then .
Remark 2.22.
The part of the above proof that requires the Lipschitz condition on is the application of dominated convergence in (2.6). The fact that the derivative of exists almost everywhere is not sufficient for a weak derivative to exist, for example, the derivative of the step function is zero almost everywhere, but we showed in Example 2.19 that the step function does not have a weak derivative. The reason is that (2.6) fails for the step function (the rest of the proof does go through for the step function).
3 Basic properties of Sobolev spaces
In this section we prove some basic results about Sobolev spaces. The results of Sections 3.1 and 3.3 describe basic functional analytic properties of Sobolev spaces, while Section 3.2 gives an alternative characterisation of Sobolev spaces as the completion of the space of smooth functions. Section 3.4 provides an answer to an earlier question by showing that, although distributions cannot always be represented by locally integrable functions, the positive distributions can always be represented by regular Borel measures.
3.1 Banach and Hilbert space structure of Sobolev spaces
It is well-known that (with the norm) is a Banach space, and that (with the inner product) is a Hilbert space. In a similar way, we can show that the Sobolev spaces have the structure of a Banach space, and that has the structure of a Hilbert space, and it is the goal of this section to give the details of this proof. This is a useful theorem, since it allows us to use theorems from functional analysis to study sequences of functions in Sobolev spaces.
Firstly, recall that the space , together with the norm, is complete when (see for example [8, Theorem 2.7] or [15, Theorem 8.14]). To extend this to the Sobolev space , we use an inductive argument. The proof of the following lemma gives the basic idea of this argument for .
Lemma 3.1.
Let be an open set and . Then the space is complete in the norm .
Proof.
Let be a Cauchy sequence in . Then, by definition of the Sobolev norm, , and so is also Cauchy in . Similarly, since (again this follows from the definition of the Sobolev norm), we have that is a Cauchy sequence in .
Since is complete, then there are functions such that
Hölder’s inequality shows that , and so each determines a distribution given by
for all test functions .
Another application of Hölder’s inequality gives the following estimate for any
where is the conjugate Hölder exponent of . (Note that the integral exists since is bounded, has compact support, and .) Since in then this shows that in .
The same argument with replaced by and replaced by shows that . We then have for every test function
Therefore, by Theorem 2.8, we have almost everywhere, where is the weak derivative, which exists since . Therefore, we have shown that in , and so is complete. ∎
Using this technique we can now prove the following theorem, which, together with Lemma 2.17, says that is a Banach space.
Theorem 3.2.
Let be open and . Then is complete in the norm for all . In particular, is a Banach space for all and .
Proof.
The proof uses induction on . The case follows from standard results about spaces (see for example [15, Theorem 8.14]). Suppose that is complete, and let be a Cauchy sequence in . Therefore the sequences and (for ) are Cauchy, and the completeness of shows that there exist functions such that
Note that the inductive hypothesis shows that in for all multi-indices such that , and so it only remains to show that for each .
As in the previous proof we can show that
and so for all test functions we have
and so Theorem 2.8 shows that almost everywhere. This, together with the previous statement that in for all multi-indices such that , shows that in for all such that .
Therefore, we have shown that there exists such that , and so is complete. ∎
In the case , the previous theorem, together with the following inner product, gives the structure of a Hilbert space.
Definition 3.3.
The inner product on is defined to be
(3.1) |
Remark 3.4.
The Sobolev norm on is the same as the norm induced by the inner product
Theorem 3.5.
is a Hilbert space.
3.2 Sobolev spaces are the completion of the space of smooth functions in the Sobolev norm (the Meyers-Serrin theorem)
In this section we prove the Meyers-Serrin theorem, which says that the Sobolev spaces defined in Section 2 are the completion of the space of smooth functions in the Sobolev norm. Therefore we now have two equivalent definitions of Sobolev spaces, which gives us a broader range of techniques to draw upon when proving theorems.
First recall the following well-known theorem that says that a normed linear space has a unique completion (see for example [11, Theorem I.3]).
Theorem 3.7.
If is a normed linear space, then there exists a unique complete normed linear space such that is isometric to a dense subset of .
Let be the space of -times differentiable functions . Since the weak derivative of a differentiable function is just the classical derivative (Lemma 2.13), then the weak derivatives of any exist up to order , and we can define the subspace
Let denote the completion of in the -norm. Since is complete by Theorem 3.2, and , then we have proved
Lemma 3.8.
For we have
It turns out that the converse is also true for , this is known as the Meyers-Serrin theorem, and the proof will occupy the rest of this section.
Example 3.9.
To see that the converse of the previous lemma can never be true for , in this example we show that . Consider first the case and , where the step function
is not in the completion of , since for any continuous function we have . To extend this example to for , simply consider the function
and note that is a step function. It is easy then to extend this idea to the case where the domain is an open subset of .
Next, we recall some basic facts needed in the proof of Theorem 3.15. The first is the existence of partitions of unity.
Theorem 3.10.
Let be an arbitrary subset of , and let be a collection of open sets in that cover . Then there exists a collection such that
For every and every , we have .
If then all but at most finitely many vanish identically on .
For every there exists such that .
For every we have (note that the sum makes sense because of the local finiteness condition (3.10)).
The collection is called a partition of unity of subordinate to .
Proof.
The case where is compact is given in [12, Theorem 2.13]. If is open, then for each define
and note that is compact and satisfies for each . Moreover, we can also write as the union of compact sets
Also, for notational convenience in what follows, define .
Given an open cover of , for each we can define an open cover of the compact set by
By the result for compact sets, for each there exists a partition of unity for the compact set that is subordinate to , and has finitely many elements. Moreover, since for each and , then for each . Therefore, since each satisfies for at most finitely many , then the sum
has at most finitely many terms for each , and also satisfies for each . Now define the collection of functions
This is now a partition of unity of subordinate to .
In the case where is an arbitrary subset of with an open cover , define the open set , note that is an open cover of , and apply the previous result to find a partition of unity of subordinate to . Since then is also a partition of unity of subordinate to . ∎
The second basic fact needed is the convergence of sequences of mollified functions. Let be a non-negative real-valued function in such that
if .
.
For example we can choose
where is chosen so that . The function is called a mollifier. For any , let , and define the mollification of to be the convolution
Lemma 3.11.
If then is smooth for all .
Since is smooth for all , then this follows from [15, Theorem 9.3].
Theorem 3.12.
Let be an open subset of , and let be an open subset with compact closure. If and , then
in .
Proof.
When this is a standard result for spaces (see for example [15, Theorem 9.6] for a proof). The general case follows by reducing to the case.
First we show that for any we have in the distributional sense on . To see this, let denote the zero extension of from to all of , and note that for any test function we have
(All of the derivatives above are taken with respect to the variable .)
Since for each , then the result for spaces shows that
This is true for all such that , and so converges to in the norm. ∎
Next, we introduce the notion of a nested open cover, which will be used in the sequel.
Definition 3.13.
Let be an open subset of . A nested open cover of is a collection of open sets such that
for all .
For all there exists such that .
Lemma 3.14.
Let be an open set in , and let be a nested open cover of . If satisfies for all , then .
Proof.
The inclusion induces an inclusion . Therefore the weak derivative of on is just the restriction of the weak derivative of on , since for all test functions we have
The dominated convergence theorem shows that for each , and as a consequence we have
Therefore . ∎
Now we are ready to prove that the space of smooth functions is dense in .
Theorem 3.15 (Meyers-Serrin).
Let be an open subset of , and let . Then for any , and for every , there exists such that .
Proof.
Fix . For each , define the open sets
Then is a nested open cover of , and, in particular, we can apply Lemma 3.14 (we will use this at the end of the proof). Moreover, each has compact closure in , and so Theorem 3.12 applies. Define , and note that is also an open cover of (although it is not nested).
Let be a partition of unity for subordinate to , and note that the local finiteness property of partitions of unity shows that for all , and we also have
for all .
From the definition of , if then has support in the set
Since , then by Theorem 3.12 we can find such that and
Define
On any compact subset , all by finitely many terms in the sum vanish, and so . Now note that if , then
and so for each
An application of Lemma 3.14 then shows that , as required. ∎
This theorem shows that . Combining this with Lemma 3.8 gives us the following corollary, which states that is the completion of the space of functions in the Sobolev norm .
Corollary 3.16.
Let be an open subset of , and let . Then
for any .
Remark 3.17.
The statement of the corollary above is that is the completion of the space of functions with respect to the -norm. Since and Theorem 3.15 is stated for smooth functions, then we also have that is the completion of the space of smooth functions in the Sobolev norm.
3.3 The dual space of a Sobolev space
In this section denotes an open subset of , , and denotes the conjugate exponent to , i.e. if and if .
First recall Theorem A.4, which says that the dual of is isomorphic to if . The proof of this theorem involves showing that for each linear functional there exists a function (unique up to equivalence in ) such that
for all . Moreover, as part of the construction, the proof also shows that . The converse is also true, so we have an isometric isomorphism .
The goal of this section is to provide a description of the dual space to the Sobolev space . It is important to point out that most of the hard work is done in proving the previous theorem for spaces, and that the proofs given below rely heavily on this construction. More details can be found in [1, Chapter 3]
Let denote the dual pairing
for and , and, for , let denote the product of copies of .
There is a map that takes a vector of functions to the linear functional . The first theorem below shows that this map is surjective, and therefore we can characterise elements of the dual in terms of elements of .
Theorem 3.18.
Given , let be the number of multi-indices such that . For every functional there exists such that for all we have
Moreover, if we define to be the set of all satisfying the previous equation, then
(3.2) |
and this infimum is attained by some .
Proof.
First note that, by the definition of , there exists a linear map
By the definition of the norms on and , the map is an isometry, and therefore is an isometric isomorphism onto its image.
Given define , a linear functional on the image of , by
Since is an isometric isomorphism, then
The Hahn-Banach theorem (see for example [11, p76]) shows that there is a norm-preserving extension of to all of , and, together with the characterisation of the dual of , this shows that there exists such that
for any . Moreover, we also have
Therefore, we have shown that for any there exists such that for all we have
Moreover, at each stage of the construction, we also showed that
∎
Unfortunately this map is not an isomorphism, since it may have a non-trivial kernel, as the next example shows.
Example 3.19.
Let be an open subset of , and let be a smooth function on with compact support. Then
(3.3) |
by the definition of weak derivative. Now consider the vector . The linear functional associated to this vector is
which is zero by (3.3). Therefore, for every non-zero smooth function with compact support contained in , the vector is a non-trivial element of the kernel of the map .
Remark 3.20.
More generally, if the functional is represented by a vector of smooth functions, i.e. , then we can write
Therefore , where . In particular, we see that is the zero functional if .
The next lemma shows that each element of the dual of a Sobolev space can be regarded as an extension of some distribution.
Lemma 3.21.
Let . Then there exists such that for all .
Proof.
Using the previous theorem, there exists such that
for every . Note that if , then
where, in the second last term, refers to the weak derivative of .
Define
Then we have shown that for all . ∎
The previous theorems give different characterisations of elements of the dual of : Theorem 3.18 shows that there is a surjective map , while Lemma 3.21 shows that the restriction of each linear functional to is a distribution. Therefore we have maps and .
Unfortunately, these results do not give a nice description of the kernel of the first map and the image of the second map. In addition, the second map may have a non-trivial kernel (see Remark 3.24). It turns out that has better properties with respect to the second map, and the next theorem describes the image of the subspace .
Theorem 3.22.
The dual space is isometrically isomorphic to the Banach space consisting of those distributions that satisfy
(3.4) |
for some , and whose norm is given by
(3.5) |
Proof.
Given , let be the space of distributions satisfying (3.4). Let . The goal of the proof is to show that has a unique extension to some , and, moreover, that this map is the inverse of the restriction map from the previous lemma.
Given , let be a sequence of test functions converging to in the -norm (note that this is not the same as convergence in the topology on the space of test functions). Such a sequence exists by the definition of . We claim that is a Cauchy sequence in , which is a consequence of the following calculation
which converges to zero, since is a Cauchy sequence in . Therefore exists, and we claim that the limit only depends on . To see this, consider another sequence of test functions converging to in the norm, and note that the same calculation as above shows that
which converges to zero as . Therefore, we can define
Clearly is linear, since both and the operation of taking the limit in are linear. To see that is bounded, we compute
and so .
Therefore, we have shown that has an extension to , and, moreover, this extension is unique since is dense in . More precisely, any other bounded linear functional that restricts to on must satisfy
By construction, for every test function , and so the map is the inverse of the restriction map from Lemma 3.21. To see that this is an isometry, note that Theorem 3.18 shows that the norm on given by (3.5) is the same as the norm on given by (3.2). Therefore is isometrically isomorphic to , which also implies that is a Banach space. ∎
Remark 3.23.
The space is a strict subset of , since there are many distributions that cannot be written as
for some . For example, the delta functional can never be written in this form, since Example 2.10 shows that it cannot be represented by a function.
Remark 3.24.
As part of the previous proof we showed that the restriction map
is injective. It is natural to ask whether these results can be extended to , however the previous proof will not work since it depends on the fact that, by definition, is the completion of in the norm (since the first step is to approximate an element of by a sequence of smooth functions with compact support).
One could still ask whether there is an alternative proof that works for , however it turns out that in general the answer is no, since the extension of a linear functional to a linear functional may be non-unique. When the domain is bounded and the boundary has good properties, then one can construct examples using the trace operator (see [4, Section 5.5] for the construction), which is zero on but non-zero in general. Therefore the restriction map has non-zero kernel, and so we cannot identify with a subspace of in this case. Note that the trace operator is zero precisely on the subspace (see [4, Theorem 2, Section 5.5] for more details).
3.4 Positive distributions can be represented by measures (the Riesz representation theorem)
Given the results of the previous section on the dual space of a Sobolev space, it is natural to ask whether there is a nice characterisation of in terms of familiar objects, and it is the goal of this section to answer this question for positive, real-valued distributions.
As we have seen from Theorem 2.8, there is an injective map . Unfortunately, as explained in Example 2.10, the set is too small to provide a unique representative for every distribution. In Theorem 3.35 we show that regular Borel measures are the right class of objects to represent distributions.
This theorem is also proved in [12, Theorem 2.14] (for the dual of the space of continuous functions with compact support) and [8, Theorem 6.22] (for the dual of the space of smooth functions with compact support). Both proofs follow a similar strategy, which involves first using the distribution to define an outer measure, and then showing that open sets are all measurable with respect to this outer measure. Rudin also considers the case of complex-valued distributions in [12, Theorem 6.19], and a more general proof (for the dual of the space ) is given in [5, Section 1.8]
Note that in [8] the proof only uses the Riemann integral, and in particular it does not involve Lebesgue measure. Since we are assuming the construction of Lebesgue measure (and a construction using outer measure is also given in Definition B.18), then we are free to use it here where it simplifies the proof.
For this entire section we use the following notation: let be an open subset of , let denote the collection of open subsets of , and let denote the Borel -algebra generated by the open subsets of .
Definition 3.25.
Let . The distribution is a positive distribution if for all such that for all .
In the following, let be an open set, and define to be the set of functions with and (note that Urysohn’s lemma shows that this set is nonempty if is nonempty).
Lemma 3.26.
Let be a positive distribution. Then the function defined by
(3.6) |
satisfies
if are open sets,
for every countable collection of open subsets .
Proof.
The first property follows from the fact that implies that .
To prove the second property, we first show that
for any open sets . Given any , let , and apply Lemma A.12 to show that there exist functions and such that , , and . Therefore
for all , and so . Induction then shows that for any
(3.7) |
and so it only remains to extend this to countable collections of open sets. To do this, note that any has compact support in , and so there exists a finite collection of sets (re-order so that these are ) such that . Equation (3.7) then gives us
which completes the proof. ∎
Now extend to a function on the set of all subsets of by
(3.8) |
Lemma 3.27.
The function is an outer measure on .
Proof.
Recall that we have to prove that each of the following conditions hold.
for all and ,
if , and
for any countable collection of sets .
The first two of the above properties follow easily from the respective definitions of and , and so it only remains to show countable subadditivity. For any , let be a collection of open subsets of such that (these sets exist since is defined using the infimum). Then
Since we can do this for any , then we have
as required. ∎
It is worth pausing at this stage to consider some examples.
Example 3.28.
Given , let , the delta functional. Then for any subset we have
Let with co-ordinates , and let be the distribution defined by integration on the subspace , i.e.
Then for any subset we have , where denotes the one-dimensional Lebesgue measure on .
Theorem B.16 shows that to construct a measure from we need to restrict to the -algebra of measurable subsets. The next lemma shows that, for the outer measure constructed above, this -algebra contains the Borel -algebra .
Lemma 3.29.
All open sets are measurable with respect to , i.e. for every set we have
Proof.
Since , then the inequality
follows from the previous lemma, and so it only remains to show the reverse inequality. First consider the case where is an open subset of . Given any open set and any , choose such that (such a exists since is defined using the supremum). Let . Then is open, and implies that .
Now choose such that (again, such a exists since is defined using the supremum). Since and , then and have disjoint support, and so
where the last step follows from Lemma 3.26 and the fact that . We can do this for any , and so for any open set .
Now consider the case where is an arbitrary subset of . Then for any open set containing and any open set we have from Lemma 3.27
Therefore
for every open set containing , and any open set . Therefore, since is defined using the infimum, then
which completes the proof. ∎
Therefore, by Theorem B.16, the function restricts to a measure (call it ) on the Borel sigma algebra . Note that this measure is given by (3.6) on open sets. The next two lemmas give a characterisation of on compact sets.
Lemma 3.30.
Given any compact set , and any such that on and on , we have .
Proof.
(See also [12, p43].) For all such that , let . Then each is open, and since on we have . Moreover, if then for all . Therefore (since is a positive distribution) and we have
for all such that . Therefore . ∎
Corollary 3.31.
If is compact, then is finite.
Lemma 3.32.
Let be a compact set. Then
(3.9) |
Proof.
Firstly note that compact sets are closed and therefore elements of the Borel -algebra. Given any , let be an open set such that and (the existence of follows from outer regularity of , which is a direct consequence of the definition of in (3.8)). Recall from Urysohn’s lemma (Theorem A.11) that there exists such that , for all , and on . Then , and so by (3.6). Therefore Lemma 3.30 implies that
We can do this for any and any compact set , therefore (3.9) holds for any compact set . ∎
Lemma 3.33.
Given any and any measurable set there exists an open set with and .
Proof.
If is finite then the result follows easily, since is measurable and is defined to be the infimum of for open.
If is infinite, then we first write the open set as the countable union of compact sets
(for example we could take each to be a closed ball), and note that
Each is a subset of a compact set, and therefore has finite measure, so we can find an open set such that and
Then is an open set containing , and
∎
We can now show that the measure is Borel regular (recall Definition B.8).
Lemma 3.34.
is a regular Borel measure on .
Proof.
Outer regularity of follows easily from the definition of , and therefore it only remains to show that it is inner regular, i.e. for any measurable set we have
(3.10) |
Given , outer regularity of shows that there exists an open set such that and . Then, since we also have , then
and so the previous lemma shows that there exists a closed set such that
Any closed set is the countable union of compact sets; for example we can take for each and write . For as above, let . If is infinite, then is infinite also. If is finite, then so is , therefore there exists such that implies that .
In both of these cases we see that can be approximated by the measure of compact sets contained in , which completes the proof of (3.10). ∎
We are now ready to prove the main theorem of this section.
Theorem 3.35.
Given a positive distribution there is a unique, positive, regular Borel measure on such that
for all compact , and
for all we have
(3.11)
Proof.
Given such a distribution , we have already constructed a positive regular Borel measure , which is defined on and is finite on compact sets, and so it only remains to show (3.11) for all .
First note that we can reduce to the case of , since both and the integral with respect to are linear, and any test function can be written for non-negative test functions (Lemma A.13).
For each , define compact sets (these are compact since is continuous with compact support), and define . Let be the characteristic function of . Then
Moreover, converges pointwise to , since for each . Since , and is compact, then we can construct a function in that dominates , and so the dominated convergence theorem shows that
Therefore it only remains to show that the integral of with respect to converges to . To see this, note that for each outer regularity of shows that we can choose to be an open set containing such that , and use Urysohn’s lemma (Theorem A.11) to find such that on and . Then by construction, we have
and therefore . By the definition of on open sets, we also have
This is true for all , and so
for all . Taking the limit as gives us .
Similarly, we can approximate from below by simple functions to obtain the opposite inequality. Since the idea is the same as above then we only sketch the details here.
For each , let , and let be the characteristic function of . Then
for all , and converges pointwise to since . Dominated convergence then shows that
For each , inner regularity of implies that we can find a compact set such that and . Then use Urysohn’s lemma to find such that on and . Then the same argument as before shows that
This is true for all , and so
Therefore , as required. ∎
Remark 3.36.
Recall that defines a distribution . Conversely, the Radon-Nikodym theorem and the Lebesgue decomposition (Theorems B.28 and B.30 respectively) show that the distribution can be represented by a function in if and only if the measure from Theorem 3.35 is absolutely continuous with respect to Lebesgue measure. We have already seen that there exist distributions that cannot be represented by functions in , for example the delta functional from Example 2.10. For these distributions, the measure constructed in Theorem 3.35 will not be absolutely continuous with respect to Lebesgue measure, i.e. it will have a non-trivial singular component with respect to the Lebesgue decomposition (B.7).
4 Embedding and compactness theorems
The goal of this section is to state the Sobolev Embedding Theorem and the Rellich-Kondrachov compactness theorem. For now, the proof has been postponed until a future version of these notes. An excellent source for the embedding and compactness theorems is [1], which also contains many examples that show the bounds from the theorems are sharp.
First, we have to define the class of domains under consideration. Given an open subset , let
Definition 4.1.
Let be an open subset of . We say that satisfies the cone condition if there exists a finite cone such that each is the vertex of a finite cone contained in and congruent to .
We say that satisfies the uniform cone condition if there exists a locally finite open cover of the boundary of and a corresponding sequence of finite cones, each congruent to some fixed finite cone , such that
there exists such that every has diameter less than ,
for some ,
for every , and
for some , every collection of of the sets has empty intersection.
Since is open, then continuously differentiable on does not imply bounded. For , define to be the space of functions in that are bounded and have bounded partial derivatives up to order.
This is a Banach space with norm
Recall that a linear map of normed linear spaces is an embedding if is bounded with respect to the norms on and . Since the elements of are equivalence classes of functions defined almost everywhere, then the meaning of an inclusion map from into is that each equivalence class in contains a function in .
The meaning of an inclusion map from into (where is the intersection of with a plane of dimension in ) is that each function in is the limit of a sequence of functions (see Section 3.2) and the restriction of these smooth functions to converges to a limit in . For the map to be well-defined then this limit needs to be independent of the original choice of sequence, however this is guaranteed if the norm on is bounded by a constant times the norm on (which always occurs in the cases considered below).
Theorem 4.2 (Sobolev Embedding Theorem).
Let be an open set satisfying the cone condition, and, for , let be the intersection of with a plane of dimension in . Let and be integers, and let . Then
If either , or and , then
and
If and , then
If and either , or and , then
Note that in each case, it is the quantity that determines the allowed embeddings. Increasing this quantity by either (a) giving up more derivatives, or (b) increasing the power , allows for “better” embeddings in the following sense: when then we get an embedding into the space of continuously differentiable functions (the first case above), and when then we get an embedding (the third case above). The same philosophy applies to the compactness theorem below, as well as the Sobolev multiplication theorem (which has been postponed until a future version of the notes).
Theorem 4.3 (Rellich-Kondrachov compactness theorem).
Let be an open set satisfying the cone condition, let be a bounded open subset of , and let be the intersection of with a -dimensional plane in . Let and be integers, and let . Then
If then the following embeddings are compact
If , then the following embeddings are compact
Appendix A Notation and basic definitions
A.1 spaces and spaces
The spaces and form the basis for the definition of the Sobolev spaces and in Definition 2.16, and so we review some of their basic properties here.
Definition A.1.
Let be an open set, let , and let denote the space of Lebesgue measurable functions on . Define
where if almost everywhere. When , define
where
is the essential supremum of on .
It is well-known that when the spaces are Banach spaces with the norm (see for example [12, Theorem 3.11] or [15, Theorem 8.14]).
Definition A.2.
Let . The conjugate exponent of is the real number such that
If then the conjugate exponent of is , and if then the conjugate exponent of is .
One of the most important inequalities for spaces is Hölder’s inequality. For a proof, see for example [12, Theorem 3.5]
Theorem A.3 (Hölder’s inequality).
Let be open, and let , , where and are conjugate exponents. Then
The following theorem characterises the dual space of . It is also well-known, for a proof see for example [2, Chapter IV], [8, Theorem 2.14], or [15, Theorem 10.44].
Theorem A.4.
Let be open, let , and let be the conjugate exponent of . Then
Remark A.5.
The isomorphism has an explicit form
It is not true that , since, for any , the Hahn-Banach theorem shows that the delta functional defined on extends to a bounded linear functional (call it ) on . A similar argument to Example 2.10 shows that cannot be represented by a function in , i.e. there is no such that
for all .
More generally, this theorem is true for any -finite measure space (see [15, pp182-185]). Since the proof uses the Radon-Nikodym theorem then the result may not be true if the measure is not -finite (see Appendix B for the relevant definitions and statements of the theorems). The following example illustrates this for a simple case.
Example A.6.
Let be the trivial -algebra on a space , and let be the measure , . Then the measurable functions are constants, and so we see that consists of only the zero function. Therefore the dual is , however , and so the dual of is not isomorphic to in this case.
Next, we define the space , which consists of locally integrable functions, in the sense that their integral is finite on compact sets.
Definition A.7.
Let be an open set in , and let be the set of Lebesgue measurable functions on . Then
Remark A.8.
One can easily extend this definition to arbitrary measure spaces that also have a topology (and hence a notion of compactness).
Clearly we have an inclusion . The following examples show that this is not surjective.
Example A.9.
. Let , and note that
where denotes the Lebesgue measure of . Therefore for any , even though .
. Let , and note that is bounded on any compact subset . Therefore , and so , but for any . (Note that we can extend this to any by choosing for some , or even a function that grows faster at the origin, such as .)
The spaces and for have radically different properties to those described above for other values of . These properties are discussed further in [13, pp35-36].
A.2 Integration by parts
Since we use integration by parts on open subsets of in Section 2.2, then we recall the formula here. See [4, Appendix C.1] for a more complete description.
Given an open set with a boundary, define
to be the outward pointing normal at each point of the boundary , and let denote the volume element on the boundary.
Theorem A.10 (Integration by parts).
Let be a bounded open subset of with a boundary, and let . Then for all we have
If has compact support in , then the boundary term disappears, and we have
A.3 The smooth Urysohn lemma and partitions of unity
The goal of this section is to give some consequences of the smooth Urysohn lemma and the existence of partitions of unity on open subsets of .
Theorem A.11.
Let be an open set, let be compact, and let be an open set such that . Then there exists a smooth function such that for all , on , and on . Moreover, there also exists such that for all , and on .
See [3, Theorem 2.6.1] for a proof.
As a consequence of Urysohn’s lemma, we have the following useful results.
Lemma A.12.
Let be open sets in , and let be compact. Then there exist non-negative functions which are smooth on and satisfy
for all ,
and .
Proof.
Urysohn’s lemma shows that there exists such that and on . Apply Urysohn’s lemma again to find such that and on a neighbourhood of the compact set . Let , and note that
,
,
on ,
and on .
Define , and note that
,
, and
on .
Therefore and satisfy the stated conditions. ∎
Any smooth function can be written as the difference of two non-negative continuous functions, just by taking the positive and negative parts of the original function. The next lemma shows that a smooth function can also be written as the difference of two non-negative smooth functions.
Lemma A.13.
Let . Then there exist functions such that and for all and for all .
Proof.
Using Urysohn’s lemma, construct a non-negative smooth function such that on , and . Then both and are non-negative smooth functions, and so we can define , . ∎
A.4 A corollary of Rademacher’s theorem
Rademacher’s theorem states that a Lipschitz function is differentiable almost everywhere (with respect to Lebesgue measure). This is used in Example 2.21 as part of the proof that a Lipschitz continuous function is in . The actual statement used in Example 2.21 is that the partial derivatives of exist almost everywhere (a slightly weaker statement than Rademacher’s theorem). The purpose of this section is to recall the basic definitions and state the theorem. A proof of Rademacher’s theorem can be found in [5].
First, recall the following definition.
Definition A.14.
Let be an open set. A function is locally Lipschitz continuous on if for every there exists a constant and a neighbourhood of such that the following inequality is satisfied
(A.1) |
If there exists a uniform constant such that for all then we say that is uniformly Lipschitz on . The smallest value of the constant is called the Lipschitz constant
(A.2) |
Theorem A.15.
Let be open, and let be locally Lipschitz on . Then is differentiable almost everywhere in (with respect to the Lebesgue measure on ).
For a proof, see [5, Section 3.1.2].
Remark A.16.
Uniformly Lipschitz implies absolutely continuous, and so we know that the theorem is true for functions with one-dimensional domains by general theory of absolutely continuous functions (see for example [15, Theorem 7.29]). This fact is used in the proof for , however some further analysis is also necessary (see [5, Section 3.1.2] for the details).
As a consequence of this, we have the following
Corollary A.17.
Let be an open subset of , and let be a locally Lipschitz function. Then the partial derivatives of exist almost everywhere on (with respect to the Lebesgue measure on ).
Appendix B Basic results from measure theory
Since Section 3.4 deals with measure, then, for completeness, here we review the basic definitions. In particular, this includes the definition of a complex measure. Since these notes assume knowledge of Lebesgue integration and the basic theorems associated to this (monotone convergence, dominated convergence, etc.) then this is not included here, the purpose is just to recall the important definitions that are used elsewhere in the notes. Of course, there are already good sources for this material such as [12], [15], [7], [10] (and many more!), so only the material relevant to the rest of the notes is covered in this section. Examples are included wherever possible in order to clarify the theory.
Since we want to deal with sets of infinite measure ( is the standard example), then first we have to define arithmetic in . This is an extension of the usual operations of addition and multiplication on , together with the following definitions.
Care must be taken when cancelling terms from an equation: implies only if , and only if . The consequence of these definitions is that the integral of any function over a set of measure zero will be zero, and the integral of the zero function over any set will also be zero.
Definition B.1.
A collection of subsets of a set is a -algebra on if all of the following hold.
.
If , then .
If for all , then .
Example B.2.
The set of all subsets of forms a -algebra.
is a -algebra, called the trivial -algebra on .
The set of all subsets of that are open is not a -algebra, since the complement of an open set is not necessarily open.
The set of all subsets of that are either open or closed is not a -algebra, since it is not closed under the operation of countable unions. For example, in the case
is neither open nor closed.
If is a -algebra on and , then the collection
is a -algebra on .
Since open and closed subsets of are of fundamental importance, then it would be useful to have a -algebra that contains all of these sets. The -algebra of all subsets of is too large for interesting measures to exist (see [10, Section 5] for more insight into why this is the case), so it would also be useful for this -algebra to have some minimality property, i.e. it is the “smallest” -algebra that contains all of the open and closed subsets of . The next theorem shows that such a -algebra exists.
Theorem B.3.
Let be a collection of subsets of a set . Then there exists a unique -algebra on , call it , such that
, and
any other -algebra containing satisfies (i.e. is the smallest -algebra containing ).
This -algebra is called the -algebra generated by .
Proof of Theorem B.3.
Consider the family of all -algebras on that contain . Since the set of all subsets of is a -algebra containing , then this family is non-empty. We claim that the intersection of all -algebras containing is also a -algebra, and the result will then follow, since such a -algebra clearly satisfies both of the conditions of the theorem.
Firstly note that the set is in every -algebra containing , and so also. If a subset is in , then it is in every -algebra containing , and so is in every -algebra containing , therefore also. Therefore it only remains to check that is closed under countable unions. To see this, let be a countable collection of sets in . Then for every -algebra containing , and so also. Therefore , and we have shown that is a -algebra. ∎
Definition B.4.
The Borel -algebra on is the smallest -algebra that contains the collection of open and closed subsets of . The sets in are called the Borel subsets of .
Definition B.5.
Let be a -algebra on a set . A positive measure on is a function such that
(B.1) |
for any disjoint collection . A function satisfying (B.1), but without the restriction that the range is , is called a countably additive set function.
A complex measure on is a countably additive function (see [12, Chapter 6] for more about complex measures).
Definition B.6.
A measure space consists of a set , a -algebra of subsets of , and a measure on .
A measure space is finite if is finite. A measure space is -finite if is the countable union of sets with finite for each .
Remark B.7.
A -finite measure is the countable sum of finite measures. To see this, let be -finite, with and finite for each . Define measures
and note that .
An important class of measures on are those defined on the Borel -algebra.
Definition B.8.
A Borel measure on is a measure defined on the Borel -algebra.
A Borel measure is inner regular (resp. outer regular) if for every we have
If a Borel measure is both inner and outer regular, then we say that is Borel regular.
Remark B.9.
Again, this notion can be extended to measures on a locally compact Hausdorff space (see [12]).
The definition of a Borel regular measure given above is equivalent to the requirement that every measurable set has the same measure as some Borel sets and . To see this, let be the intersection of open sets such that and , and let be the union of compact sets such that and .
A useful way to construct a measure with certain desired properties is to start with an outer measure. For example, Lebesgue measure and Hausdorff measure can both be constructed using outer measures (see also [12] for a construction of Lebesgue measure that doesn’t use outer measure), and the construction in Section 3.4 of a measure associated to a distribution also uses outer measure.
Definition B.10.
A function defined on the power set of a space is called an outer measure on if it satisfies all of the following.
, .
if .
for any countable collection of sets .
The point of studying outer measures is that it is easy to construct an outer measure with certain properties.
To get a feel for outer measure we recall here two main examples: Lebesgue outer measure and Hausdorff outer measure.
Example B.11 (Lebesgue outer measure).
The Lebesgue outer measure is defined on compact rectangular subsets
by
For an arbitrary subset , consider the collection of all countable covers of by rectangular sets , and define
(B.2) |
It is easy to check that this definition satisfies the conditions of an outer measure (see for example [15, Theorems 3.3 & 3.4]).
The next theorem is a useful characterisation of the Lebesgue outer measure.
Theorem B.12.
Let . Then for each , there exists an open set such that and .
In particular, we have
For a proof, see [15, Theorem 3.6].
The Hausdorff outer measure and the associated Hausdorff measure are useful for studying certain subsets of that have Lebesgue measure zero. For example, the -dimensional Hausdorff measure of a -dimensional ball in is non-trivial, even though the Lebesgue measure is zero. Furthermore, the Hausdorff measure can be used to distinguish fractal sets (sets of fractional Hausdorff dimension), and the study of the properties of Hausdorff measure is a major component of Geometric Measure Theory (see [5], [6], [9]).
Example B.13 (Hausdorff outer measure).
The diameter of a set is defined to be
Fix (not necessarily an integer), and let . Let denote the collection of countable covers of such that for each . Given , define
If , then , and so . Therefore
is called the -dimensional Hausdorff outer measure of . Again, it is easy to check that this is an outer measure (see for example [15, Theorem 11.12]).
The next definition and theorem show that each outer measure has an associated -algebra and that the restriction of the outer measure to this -algebra is a measure.
Definition B.14.
Let be an outer measure on . A subset is -measurable if and only if
(B.3) |
for every subset .
Remark B.15.
An equivalent definition is that is -measurable if and only if
(B.4) whenever and . To see that the first definition implies the second, given any sets and , let . Clearly (B.3) implies (B.4). Conversely, given any set , let and . Clearly these satisfy the requirements and , and we have . Again, it is clear that (B.4) implies (B.3).
Both of these definitions have the same basic idea: the -measurable subsets of are those for which is additive on arbitrary decompositions into disjoint subsets.
The next theorem justifies the use of the term “measurable” in the previous definition.
Theorem B.16 (Caratheodory).
Let be an outer measure on . Then the collection of -measurable subsets of forms a -algebra, and the restriction of to this -algebra is a measure.
For a proof, see for example [8, Theorem 1.15].
Remark B.17.
This theorem is used in Section 3.4 to construct the measure associated to a positive distribution.
Definition B.18.
Remark B.19.
Open and closed sets are Lebesgue measurable, and therefore the -algebra of Lebesgue measurable sets contains the Borel -algebra.
Using this definition of Lebesgue measure, together with Theorem B.12, we see that Lebesgue measure is Borel outer regular.
Lemma B.20.
For any Lebesgue measurable set we have
The proof follows by restricting the result of Theorem B.12 to the -algebra of Lebesgue-measurable sets. By taking complements, we also see that is Borel inner regular.
Lemma B.21.
For any Lebesgue measurable set we have
This is a consequence of [15, Lemma 3.22], which states that is measurable if and only if for all there exists a closed set such that . The lemma above then follows by taking a sequence of compact sets .
A natural question arising from Theorem B.3 is whether two measures that agree on the Borel subsets of also agree on the Borel -algebra (the minimal -algebra generated by the Borel subsets). This question is answered in more generality by the Caratheodory-Hahn Extension Theorem, for which we first need the following definitions.
Definition B.22.
An algebra of subsets of is a non-empty collection of subsets of that is closed under the operations of taking complements and finite unions.
Note that, as a consequence, an algebra is also closed under finite intersections, and therefore both and the empty set are both in . The difference between this definition and that of a -algebra is that a -algebra is also closed under countable unions. For example, the set of all open and closed subsets of is an algebra, but not a -algebra.
Definition B.23.
A measure on an algebra is a function such that , and
whenever is a countable collection of disjoint sets in whose union also belongs to .
Given a measure on an algebra , we can construct an outer measure on as follows. For each subset , let be the collection of countable covers of by sets in . Define
(B.5) |
Theorem B.24.
Let be a measure on an algebra , and let be as defined in (B.5). Then
is an outer measure,
for all , and
is -measurable for all .
For a proof, see [15, Theorems 11.18 and 11.19].
Definition B.25.
Let be a measure on an algebra . If is a measure on a -algebra containing , and for all , then we say that is an extension of the measure to the -algebra .
Theorem B.16 shows that the outer measure defined in (B.5) is a measure on some -algebra containing . The next theorem shows that this is the unique extension of to any -algebra contained in .
Theorem B.26 (Caratheodory-Hahn Extension Theorem).
Let be a measure on an algebra , let be the corresponding outer measure, and let be the -algebra of -measurable sets. Then the restriction of to is an extension of . Moreover, if is -finite with respect to , and if is any -algebra with , then is the only measure on that is an extension of .
For a proof see [15, Theorem 11.20].
Sobolev spaces are defined in terms of distributions, and in many of the examples from Sections 2 and 3 we consider distributions that are represented by a function , i.e. where
Many distributions cannot be represented by a function, for example the delta functional from Example 2.7. The main theorem of Section 3.4 shows that instead of using functions to represent distributions, the right class of objects to look at is the class of regular Borel measures (see Theorem 3.35). A natural question is to ask when a measure can be represented by a function, and, if not, then how can this failure be expressed in terms of properties of the measure. This is the content of the Lebesgue decomposition and Radon-Nikodym theorem.
Definition B.27.
Let and be measures on the same -algebra on a space . The measure is absolutely continuous with respect to if for every set with . The measure is singular with respect to if there is a set with , and for every such that .
In other words, if sets of -measure zero are also sets of -measure zero, then is absolutely continuous with respect to . If is supported on a set of -measure zero then it is singular with respect to .
Theorem B.28 (Radon-Nikodym theorem).
Let be a -finite measure space, and let be a measure on that is absolutely continuous with respect to . Then there exists such that
(B.6) |
for each .
Theorem B.29.
Let be a measure space, and let be a measure on that is singular with respect to . Then there exists a set with , and
for each .
Theorem B.30 (Lebesgue Decomposition).
Let be a -finite measure on a -algebra , and let be a finite measure on . Then there is a unique decomposition
(B.7) |
where and are measures on such that is absolutely continuous with respect to , and is singular with respect to .
See [12] or [15] for different proofs of the above statements. Note that Rudin in [12] considers the more general case of a complex measure.
The following simple example shows that -finiteness is a necessary condition in the Radon-Nikodym theorem. Another example using the counting measure is described in [12, pp123-124].
Example B.31.
Let be the trivial -algebra on a set , and let and be measures on with , , , and . Note that is absolutely continuous with respect to , and that is not a -finite measure on . Then the -measurable functions are the constants (since measurable implies that for all open sets ), and so for any non-zero measurable function. Since , then there cannot exist any measurable function such that , and therefore the Radon-Nikodym theorem does not hold in this case.
The next lemma is a consequence of the well-known Vitali covering lemma.
Lemma B.32.
Let be an open set. Then for all there exists a countable collection of disjoint closed balls in such that
for all , and
has Lebesgue measure zero.
See [5, Corollary 2, p28] for a proof.
References
- 1Robert A. Adams.Sobolev spaces.Academic Press [A subsidiary of Harcourt Brace Jovanovich, Publishers], New York-London, 1975.Pure and Applied Mathematics, Vol. 65.
- 2S. Banach.Theory of linear operations, volume 38 of North-Holland Mathematical Library.North-Holland Publishing Co., Amsterdam, 1987.Translated from the French by F. Jellett, With comments by A. Pełczyński and Cz. Bessaga.
- 3Lawrence Conlon.Differentiable manifolds: a first course.Birkhäuser Advanced Texts: Basler Lehrbücher. [Birkhäuser Advanced Texts: Basel Textbooks]. Birkhäuser Boston Inc., Boston, MA, 1993.
- 4Lawrence C. Evans.Partial differential equations, volume 19 of Graduate Studies in Mathematics.American Mathematical Society, Providence, RI, 1998.
- 5Lawrence C. Evans and Ronald F. Gariepy.Measure theory and fine properties of functions.Studies in Advanced Mathematics. CRC Press, Boca Raton, FL, 1992.
- 6Herbert Federer.Geometric measure theory.Die Grundlehren der mathematischen Wissenschaften, Band 153. Springer-Verlag New York Inc., New York, 1969.
- 7Paul R. Halmos.Measure Theory.D. Van Nostrand Company, Inc., New York, N. Y., 1950.
- 8Elliott H. Lieb and Michael Loss.Analysis, volume 14 of Graduate Studies in Mathematics.American Mathematical Society, Providence, RI, second edition, 2001.
- 9Frank Morgan.Geometric measure theory.Academic Press Inc., San Diego, CA, third edition, 2000.A beginner’s guide.
- 10John C. Oxtoby.Measure and category, volume 2 of Graduate Texts in Mathematics.Springer-Verlag, New York, second edition, 1980.A survey of the analogies between topological and measure spaces.
- 11Michael Reed and Barry Simon.Methods of modern mathematical physics. I.Academic Press Inc. [Harcourt Brace Jovanovich Publishers], New York, second edition, 1980.Functional analysis.
- 12Walter Rudin.Real and complex analysis.McGraw-Hill Book Co., New York, third edition, 1987.
- 13Walter Rudin.Functional analysis.International Series in Pure and Applied Mathematics. McGraw-Hill Inc., New York, second edition, 1991.
- 14François Trèves.Topological vector spaces, distributions and kernels.Academic Press, New York, 1967.
- 15Richard L. Wheeden and Antoni Zygmund.Measure and integral.Marcel Dekker Inc., New York, 1977.An introduction to real analysis, Pure and Applied Mathematics, Vol. 43.
- 16William P. Ziemer.Weakly differentiable functions, volume 120 of Graduate Texts in Mathematics.Springer-Verlag, New York, 1989.Sobolev spaces and functions of bounded variation.