Sobolev spaces in one dimension and absolutely continuous functions

Jordan Bell
October 21, 2015

1 Locally integrable functions and distributions

Let λ be Lebesgue measure on . We denote by loc1(λ) the collection of Borel measurable functions f: such that for each compact subset K of ,

NK(f)=K|f|𝑑λ=1K|f|𝑑λ<.

We denote by Lloc1(λ) the collection of equivalence classes of elements of loc1(λ) where fg when f=g almost everywhere.

Write B(x,r)={y:|y-x|<r}=(x-r,x+r). For floc1(λ) and x, we say that x is a Lebesgue point of f if

limr01λ(B(x,r))B(x,r)|f(y)-f(x)|𝑑λ(y)=0.

It is immediate that if f is continuous at x then x is a Lebesgue point of f. The Lebesgue differentiation theorem11 1 Walter Rudin, Real and Complex Analysis, third ed., p. 138, Theorem 7.7. states that for floc1(λ), almost every x is a Lebesgue point of f. A sequence of Borel sets En is said to shrink nicely to x if there is some α>0 and a sequence rn0 such that EnB(x,rn) and λ(En)αλ(B(x,rn)). The sequence B(x,n-1)=(x-n-1,x+n-1) shrinks nicely to x, the sequence [x,x+n-1] shrinks nicely to x, and the sequence [x-n-1,x] shrinks nicely to x. It is proved that if floc1(λ) and for each x, En(x) is a sequence that shrinks nicely to x, then

f(x)=limn1λ(En)En(x)f𝑑λ

at each Lebesgue point of f.22 2 Walter Rudin, Real and Complex Analysis, third ed., p. 140, Theorem 7.10.

For a nonempty open set Ω in , we denote by Cck(Ω) the collection of Ck functions ϕ: such that

suppϕ={x:ϕ(x)0}¯

is compact and is contained in Ω. We write 𝒟(Ω)=Cc(Ω), whose elements are called called test functions. The following statement is called the fundamental lemma of the calculus of variations or the Du Bois-Reymond Lemma.33 3 Lars Hörmander, The Analysis of Linear Partial Differential Operators I, second ed., p. 15, Theorem 1.2.5.

Theorem 1.

If fLloc1(λ) and Rfϕ𝑑λ=0 for all ϕD(R), then f=0 almost everywhere.

Proof.

There is some η𝒟(-1,1) with η𝑑λ=1. We can explicitly write this out:

η(x)={c-1exp(1x2-1)|x|<10|x|1,

where

c=-11exp(1y2-1)𝑑λ(y)=0.443994.

For x a Lebesgue point of f and for 0<r<1,

f(x) =f(x)η(y)𝑑λ(y)
=f(x)1rη(yr)𝑑λ(y)
=f(x)1rη(x-yr)𝑑λ(y)
=1r(f(x)-f(y))η(x-yr)𝑑λ(y)+1rf(y)η(x-yr)𝑑λ(y)
=1r(f(x)-f(y))η(x-yr)𝑑λ(y)
=1r(x-r,x+r)(f(x)-f(y))η(x-yr)𝑑λ(y).

Then

|f(x)|η1r(x-r,x+r)|f(y)-f(x)|𝑑λ(y)0,r0,

meaning that f(x)=0. This is true for almost all x, showing that f=0 almost everywhere. ∎

For floc1(λ), define Λf:𝒟() by

Λf(ϕ)=fϕ𝑑λ.

𝒟() is a locally convex space, and one proves that Λf is continuous and thus belongs to the dual space 𝒟(), whose elements are called distributions.44 4 Walter Rudin, Functional Analysis, second ed., p. 157, §6.11. We say that a distribution Λ is induced by floc1(λ) if Λ=Λf. For Λ𝒟(), we define DΛ:𝒟() by

(DΛ)(ϕ)=-Λ(ϕ).

It is proved that DΛ𝒟().55 5 Walter Rudin, Functional Analysis, second ed., p. 158, §6.12.

Let f,gloc1(λ). If DΛf=Λg, we call g a distributional derivative of f. In other words, for floc1(λ) to have a distributional derivative means that there is some gloc1(λ) such that for all ϕ𝒟(),

-fϕ𝑑λ=gϕ𝑑λ.

If g1,g2loc1(λ) are distributional derivatives of f then (g1-g2)ϕ𝑑λ=0 for all ϕ𝒟(), which by Theorem 1 implies that g1=g2 almost everywhere. It follows that if f has a distributional derivative then the distributional derivative is unique in Lloc1(λ), and is denoted DfLloc1(λ):

-fϕ𝑑λ=(Df)ϕ𝑑λ,ϕ𝒟().

2 The Sobolev space H1()

We denote by 2(λ) the collection of Borel measurable functions f: such that |f|2𝑑λ<, and we denote by L2(λ) the collection of equivalence classes of elements of 2(λ) where fg when f=g almost everywhere, and write

f,gL2=fg𝑑λ.

It is a fact that L2(λ) is a Hilbert space.

We define the Sobolev space H1() to be the set of fL2(λ) that have a distributional derivative that satisfies DfL2(λ). We remark that the elements of H1() are equivalence classes of elements of 2(λ). We define

f,gH1=f,gL2+Df,DgL2.

Let f,gH1() and let ϕ𝒟(). Because f,g have distributional derivatives Df,Dg,

-(f+g)ϕ𝑑λ =-fϕ𝑑λ-gϕ𝑑λ
=Dfϕ𝑑λ+Dgϕ𝑑λ
=(Df+Dg)ϕ𝑑λ.

This means that f+g has a distributional derivative, D(f+g)=Df+Dg. Thus H1() is a linear space. If f,fH1=0 then |f|2𝑑λ=0, which implies that f=0 as an element of L2(λ). Therefore ,H1 is an inner product on H1().

If fn is a Cauchy sequence in H1(), then fn is a Cauchy sequence in L2(λ) and Dfn is a Cauchy sequence in L2(λ), and hence these sequences have limits f,gL2(λ). For ϕ𝒟(),

-fϕ𝑑λ =-limnfnϕ𝑑λ
=limn(Dfn)ϕ𝑑λ
=gϕ𝑑λ.

This means that f has distributional derivative, Df=g. Because f,DfL2(λ) it is the case that fH1(). Furthermore,

fn-fH12=fn-fL22+Dfn-DfL22=fn-fL22+Dfn-gL220,

meaning that fnf in H1(), which shows that H1() is a Hilbert space.

3 Absolutely continuous functions

We prove a lemma that gives conditions under which a function, for which integration by parts needs not make sense, is equal to a particular constant almost everywhere.66 6 Haim Brezis, Functional Analysis, Sobolev Spaces and Partial Differential Equations, p. 204, Lemma 8.1.

Lemma 2.

If fLloc1(λ) and

fϕ𝑑λ=0,ϕ𝒟(),

then there is some cR such that f=c almost everywhere.

Proof.

Fix η𝒟() with η𝑑λ=1. Let w𝒟() and define

h=w-ηw𝑑λ,

which belongs to 𝒟() and satisfies h𝑑λ=0. Define ϕ: by

ϕ(x)=-xh𝑑λ.

Using ϕ(x)=h(x) for all x and ϕ(x)h𝑑λ=0 as x, check that ϕ𝒟(). Then by hypothesis, fϕ𝑑λ=0, i.e.

0 =fh𝑑λ
=(fw-fηw𝑑λ)𝑑λ
=(f-fη𝑑λ)w𝑑λ.

Because this is true for all w𝒟(), by Theorem 1 we get that f=fη𝑑λ almost everywhere. ∎

Lemma 3.

Let gLloc1(λ), let aR, and define f:RR by

f(x)=axg(y)𝑑λ(y).

Then

fϕ𝑑λ=-gϕ𝑑λ

for all ϕD(R).

Proof.

Using Fubini’s theorem,

f(x)ϕ(x)𝑑λ(x) =--a(xag(y)𝑑λ(y))ϕ(x)𝑑λ(x)
+a(axg(y)𝑑λ(y))ϕ(x)𝑑λ(x)
=--a(-yϕ(x)𝑑λ(x))g(y)𝑑λ(y)
+a(yϕ(x)𝑑λ(x))g(y)𝑑λ(y)
=--aϕ(y)g(y)𝑑λ(y)-aϕ(y)g(y)𝑑λ(y)
=-g(y)ϕ(y)𝑑λ(y).

For real numbers a,b with a<b, we say that a function f:[a,b] is absolutely continuous if for all ϵ>0 there is some δ>0 such that whenever (a1,b1),,(an,bn) are disjoint intervals each contained in [a,b] with (bk-ak)<δ it holds that |f(bk)-f(ak)|<ϵ. We say that a function f: is locally absolutely continuous if for each nonempty compact interval [a,b], the restriction of f to [a,b] is absolutely continuous. We denote the collection of locally absolutely continuous by ACloc().

Let fH1(), let a, and define h: by

h(x)=axDf𝑑λ.

By Lemma 3 and by the definition of a distributional derivative,

hϕ𝑑λ=-(Df)ϕ𝑑λ=fϕ𝑑λ,ϕ𝒟().

Hence (f-h)ϕ𝑑λ=0 for all ϕ𝒟(), which by Lemma 2 implies that there is some c such that f-h=c almost everywhere. Let f~=c+h. On the one hand, the fact that DfLloc1(λ) implies that hACloc() and so f~ACloc(). On the other hand, f~=f almost everywhere. Furthermore, because f~ is locally absolutely continuous, integration by parts yields

f~ϕ𝑑λ=-f~ϕ𝑑λ,

and by definition of a distributional derivative,

f~ϕ𝑑λ=-(Df~)ϕ𝑑λ.

Therefore by Theorem 1, f~=Df~ almost everywhere. But the fact that f~=f almost everywhere implies that Df~=Df almost everywhere, so f~=Df almost everywhere. In particular, f~L2(λ).

Theorem 4.

For fH1(R), there is a function f~ACloc(R) such that f~=f almost everywhere and f~=Df almost everywhere. The function f~ is 12-Hölder continuous.

Proof.

For x,y,77 7 cf. Giovanni Leoni, A First Course in Sobolev Spaces, p. 222, Theorem 7.13.

f~(x)-f~(y)=yxf~𝑑λ,

and using the Cauchy-Schwarz inequality,

|f~(x)-f~(y)| yx|f~|𝑑λ
|x-y|1/2(yx|f~|2𝑑λ)1/2
DfL2|x-y|1/2.