Chapter 8 Radon-Nikodym Theorem

8.1 Signed measures

We introduce the notion of signed measures which will be useful in the proof of the Radon-Nikodym theorem.

Definition 8.1 (Finite signed measure) A function \(\mu\) from a \(\sigma\)-algebra \(\mathcal{E}\) to \(\mathbb{R}\) is a finite signed measure if

  • \(\mu(\emptyset) = 0\),
  • If \((A_n)_{n \geq 1}\) is a sequence of disjoint sets then \(\mu (\bigcup_n A_n) = \sum_n \mu(A_n)\)

Example 8.1 If \((E, \mathcal{E}, \mu)\) is a measure space and \(f \in L^1(E)\) then \(\nu\) defined by \(\nu(A) = \mu(f1_A)\) is a signed measure.

We want to show two decomposition theorems which basically allow us to reduce the situation back to measures. First we need some more definitions and a useful Lemma.

Definition 8.2 If \((E, \mathcal{E})\) is a measurable space and \(\nu\) is a finite signed measure then we call \(A\) a positive set if for every \(B \in \mathcal{E}\) with \(B \subseteq A\) then \(\nu(B) \geq 0\). The negative sets are defined analogously.

Lemma 8.1 Suppose that \(\nu\) is a finite signed measure on \((E, \mathcal{E})\) and suppose \(A \in \mathcal{E}\) with \(\nu(A) <0\) then there exists a negative set \(B\) with \(B \subseteq A\) and \(\nu(B) \leq \nu(A)\).

Proof. We will produce this set \(A\) by an itterative process, define

\[ \delta_1 = \sup \{ \nu(C) \,:\, C \subseteq A\}, \]

then since \(\emptyset \subseteq A\) we have that \(\delta_1 \geq 0\). If \(\delta_1 = 0\) then we have a negative set so are done. If not we can find a set \(C_1 \subseteq A\) with

\[ \nu(C_1) \geq \min\{\delta_1/2 ,1\}. \]

(We take the minimum here because we don’t know that \(\delta_1\) is finite.) Now we will define a sequence of \(\delta_n\) and \(C_n\) by setting

\[ \delta_n = \sup\{ \nu(C) \,:\, C \subseteq (A \setminus \bigcup_{i=1}^{n-1}C_i)\} \]

and \(C_n \subseteq A\) a set so that \(C_n \cap \left( \bigcup_{i=1}^{n-1}C)i\right) = \emptyset\) and

\[ \nu(C_n) \geq \min\{ \delta_n /2 ,1\}. \]

Now let \(C_\infty = \bigcup_n C_n\) and \(B = A \setminus C_\infty\). We now need to check that \(B\) has the required properties,

\[ \nu(A) = \nu(C_\infty) + \nu(B) \geq \nu(B), \]

as \(\nu(C_\infty) \geq 0\) by construction. As \(\nu\) is a finite measure we must have \(\nu(C_\infty) < \infty\) and as the \(C_n\) are constructed to be disjoint this means we must have \(\lim_n \nu(C_n) = 0\). Therefore \(\lim_n \delta_n = 0\). If \(D \subseteq B\) then we must have that \(\nu(D) \leq \delta_n\) for every \(n\), therefore \(\nu(D) \leq 0\).

Now we are able to state and prove our two decomposition theorems.

Theorem 8.1 (Hahn Decomposition theorem) Let \((E, \mathcal{E})\) be a measure space and \(\nu\) a finite signed measure. Then there exists a positive set \(P\) and a negative set \(N\) for \(\nu\) such that \(E = P \cup N\).

Proof. Let \[L = \inf\{ \nu(A)\,:\, \mbox{$A$ is a negative set for $\nu$}\}\] then \(L\) is finite as otherwise we could construct a set with measure \(-\infty\). Then let \(A_n\) be a negative set with \(\nu(A_n) \leq L+1/n\) then let \(N = \bigcup_n A_n\).

We can check that \(N\) is a negative set and that \(\nu(N) = L\). If \(A \subseteq N\) then let \(B_n = A_n \setminus \bigcup_{k=1}^{n-1}A_k\) then \(A = \bigcup_n (A \cap B_n)\) and \(A \cap B_n \subseteq A_n\) so \(\nu(A \cap B_n) \leq 0\) and \(\nu(A) = \sum_n \nu(A \cap B_n) \leq 0\). Now since \(N\) is a negative set \(\nu(N \setminus A_n) \leq 0\), therefore \(\nu(N) = \nu(N \setminus A_n) + \nu(A_n) \leq \nu(A_n) \leq L +1/n\). This is true for any \(n\) so \(\nu(N) \leq L\) and since \(L\) is defined to be the infinmum over \(\nu(A)\) for all negative sets \(A\), we will have \(\nu(N) \geq L\), therefore \(\nu(N) = L\).

Let \(P = N^c\) we want to check that \(P\) is a positive set. Suppose there exists a set \(A \subseteq P\) with \(\nu(A) < 0\), then by our lemma there exists a negative set \(B \subseteq A\) with \(\nu(B) \leq \nu(A)<0\). Then \(N \cup B\) is a negative set and \(N\) and \(B\) are disjoint so \(\nu(N \cup B) = \nu(N) + \nu(B) < \nu(N)\) which contradicts the fact that \[\nu(N) = L = \inf\{ \nu(A)\,:\, \mbox{$A$ is a negative set for $\nu$}\}\] so we are done.

Theorem 8.2 (Jordan decompostion theorem) Every finite signed measure is the difference of two positive measures. Precisely, if \((E, \mathcal{E})\) is a measure space and \(\nu\) is a signed measure then there exist measures \(\nu_+\) and \(\nu_-\) such that for every \(A \in \mathcal{E}\) we have \(\nu(A) = \nu_+(A)- \nu_-(A)\).

Proof. Take some Hahn decomposition \((P, N)\) then let \(\nu_+(A) = \nu(A \cap P)\), as \(A \cap P \subseteq P\) then \(\nu(A \cap P) \geq 0\). Similarly let \(\nu_-(A) = -\nu(A \cap N)\). By additivity of \(\nu\) we have that \(\nu(A) = \nu_+(A) - \nu_-(A)\). Countable additivity of \(\nu_+\) and \(\nu_-\) follow immediately from countable additivity of \(\nu\).

Now we notice that if \(B \subseteq A\) then

\[ \nu(B) = \nu_+(B) - \nu_-(B) \leq \nu_+(B) \leq \nu_+(A) \]

and \(\nu_+(A) = \nu(A \cap P)\) therefore we have that

\[ \nu_+(A) = \sup\{\nu(B) \,:\, B \subseteq A, B \in \mathcal{E}\} \]

in the same way

\[ \nu_-(A) = \sup\{ - \nu(B) \,:\, B \subseteq A, B \in \mathcal{E}\}. \]

This shows that the values of \(\nu_+, \nu_-\) do not depend on the particular choice of Hahn decomposition.

8.2 Absolute Continuity

We now move on to the main focus of this section, the Radon-Nikodym theorem. In order to understand the theorem we need a definition.

Definition 8.3 Let \((E, \mathcal{E})\) be a measurable space and \(\mu\) and \(\nu\) be two measures then we say that \(\nu\) is absolutely continuous with respect to \(\mu\) of \(\nu \ll \mu\) if for every \(A \in \mathcal{E}\) with \(\mu(A) = 0\) we also have \(\nu(A) = 0\).

We can characterise absolute continuity

Lemma 8.2 Suppose that \((E, \mathcal{E})\) is a measurable space and \(\mu\) a measure, \(\nu\) a finite measure then \(\nu \ll \mu\) if and only if for earch \(\epsilon >0\) there exists a \(\delta>0\) such that \(\mu(A) < \delta\) implies that \(\nu(A) < \epsilon\).

Proof. First let us suppose there exists such at \(\delta\) for each \(\epsilon\), then if \(\mu(A) = 0\) we have that \(\mu(A)< \delta\) for every \(\delta\) so we must have \(\nu(A) < \epsilon\) for every \(\epsilon\) so \(\nu(A) =0\).

Now let us suppose that \(\nu \ll \mu\). We prove the result by contradiction. Suppose there exists an \(\epsilon\) such that for every \(\delta\) there exists a set \(A\) with \(\mu(A)< \delta\) but \(\nu(A) >\epsilon\). Then we can find a sequence of sets \(A_k\) such that \(\mu(A_k) < 2^{-k}\) but \(\nu(A_k) \geq \epsilon\). By the first Borel-Cantelli lemma we have that

\[ \mu \left( \bigcap_n \bigcup_{m \geq n} A_m \right) = 0. \]

We also have that \(\nu(\bigcup_{m \geq n} A_m) \geq \nu(A_n) \geq \epsilon\) and

\[\nu \left( \bigcap_n \bigcup_{m \geq n} A_m\right) = \lim_n \nu \left( \bigcup_{m \geq n} A_m \right) \geq \epsilon. \]

This show gives us a set with \(\mu(B) = 0\) but \(\nu(B) > 0\) which contradicts \(\nu \ll \mu\).

Now we can prove the main theorem for this section.

Theorem 8.3 (Radon-Nikodym Theorem) Let \((E, \mathcal{E})\) be a measure space and let \(\mu, \nu\) be two finite measures with \(\nu \ll \mu\). Then there exists a measurable function \(g: E \rightarrow [0, \infty)\) such that \(\nu(A) = \mu(g1_A)\). The function \(g\) is unique up to identifying almost everywhere equal functions. We write \(g = \mathrm{d}\nu/\mathrm{d}\mu\) and call it the Radon-Nikodym derivative of \(\nu\) with respect to \(\mu\).

Proof. Let us define the set \(\mathcal{F}\) which is the set of all measurable functions, \(f\), with \(\mu(f1_A) \leq \nu(A)\) for every \(A \in \mathcal{E}\). The idea is that \(\mathcal{F}\) contains a function \(g\) which achieves \(\mu(g) = \sup_{f \in \mathcal{F}} \mu(f)\).

As a first step we show that \(f_1 \vee f_2 = \max\{ f_1, f_2\} \in \mathcal{F}\) when \(f_1, f_2 \in \mathcal{F}\). Let us take any \(A \in \mathcal{E}\) then let \(A_1 = A \cap \{ f_1 \geq f_2\}\) and \(A_2 = A \cap \{ f_1 < f_2\}\). Then

\[ \mu(f_1 \vee f_2 1_A) = \mu(f_1 \vee f_2 1_{A_1}) + \mu(f_1 \vee f_2 1_{A_2}) = \mu(f_1 1_{A_1}) + \mu(f_2 1_{A_2}) \leq \nu(A_1) + \nu(A_2) = \nu(A). \]

Therefore \(f_1 \vee f_2 \in \mathcal{F}\).

Now take a sequence \(f_n\) such that \(\mu(f_n) \geq \sup_{f \in \mathcal{F}} \mu(f) - 1/n\). Then let \(g_n = f_1 \vee f_2 \vee \dots \vee f_n\), so that the sequence of function \(g_n\) is increasing and \(\mu(g_n) \geq \sup_{f \in \mathcal{F}} \mu(f) - 1/n\). Then as \(g_n\) is increasing it has a limit \(g\) and the monotone convergence theorem shows that

\[ \mu(g1_A) = \lim_n \mu(g_n1_A) \leq \nu(A). \]

So \(g \in \mathcal{F}\).

Now we can define another positive measure \(\nu_0(A) = \nu(A) - \mu(g1_A)\). We want to show that \(\nu_0 =0\) and will do this by contradiction. Suppose that there exists \(A \in \mathcal{E}\) such that \(\nu_0(A)>0\) then by monotonicity we will have \(\nu_0(E) >0\) and since \(\mu\) is a finite measure there exists a number \(\epsilon >0\) such that \(\nu_0(E) > \epsilon \mu(E)\). Now \(\nu_0 - \epsilon \mu\) is a finite signed measure. Let \((P, N)\) be a Hahn decomposition for this signed measure. Then \((\nu_0 - \epsilon \mu)(A \cap P) \geq 0\) so \(\nu_0(A\cap P) \geq \epsilon \mu(A \cap P)\). Hence we have

\[\begin{align*} \nu(A) &= \mu(g1_A) + \nu_0(A) \geq \mu(g1_A) + \nu_0(A \cap P) \\ & \geq \mu(g1_A) + \epsilon \mu(A \cap P) = \mu(1_A(g +\epsilon 1_{P})). \end{align*}\]

We also have that \(\mu(P) >0\) as if \(\mu(P) = 0\) then we would have \(\nu_0(P)=0\) as \(\nu_0 \ll \nu \ll \mu\), and this would mean

\[ (\nu_0 - \epsilon \mu)(E) = (\nu_0 - \epsilon \mu)(N) \leq 0, \]

which would contradict \(\nu_0(E) > \epsilon \mu(E)\). Therefore, \(g+ \epsilon 1_P\) belongs to \(\mathcal{F}\) but \(\mu(g+1_P) > \mu(g)\) which contradicts the fact that \(g\) achieves \(\mu(g) = \sup_{f \in \mathcal{F}} \mu(f)\). Hence \(\nu(A) = \mu(g1_A)\).

Now we turn to uniqueness suppose that we have two positive functions \(g,h\) such that \(\nu(A) = \mu(g1_A) = \mu(h1_A)\) for every \(A\), then as \(\nu\) is finite \(g\) and \(h\) are integrable so \(g-h\) is integrable and \(\mu((g-h)1_A) = 0\) for every \(A\). As \(g-h\) is measurable then \(\{x \in E\,:\, g-h \geq 0\}\) is a measurable set so \(\mu((g-h)1_{\{x \in E\,:\, g-h \geq 0\}}) = 0\). This shows that \((g-h)1_{\{x \in E\,:\, g-h \geq 0\}} = 0\) almost everywhere. In the same way \((g-h)1_{\{ x \in E\,:\, g-h \leq 0\}} =0\) almost everywhere. Therefore \(g=h\) \(\mu\)-almost everywhere.

8.3 Duality in \(L^p\) spaces - NONEXAMINABLE

The goal of this section is to prove that if \(1/p+1/q =1\) then the dual space of \(L^p(E)\) is isomorphic to the space \(L^q(E)\). First let us define a dual space.

Definition 8.4 Let \(\mathcal{V}\) be a Banach space (a complete, normed vector space) then the dual space of \(\mathcal{V}\) is written \(\mathcal{V}'\) and is the space of all bounded linear operators from \(\mathcal{V}\) to \(\mathbb{R}\). We recall that we call an operator \(K\) on \(\mathcal{V}\) bounded if \(|K(v)| \leq C\|v\|\) for all \(v \in \mathcal{V}\). We can define a norm on \(\mathcal{V}'\) by \(\|K\| = sup_{\|v\| =1}|K(v)|\).

The first thing to note is that if \(g \in L^q(E)\) then we can define a bounded linear operator on \(L^p(E)\) by \(K_g(f) = \mu(fg)\). This is bounded by Hölder’s inequality \(|\mu(fg)| \leq \mu(|fg|) = \|fg\|_1 \leq \|f\|_p \|g\|_p\). It is also linear thanks to the linearity of the integral. Therfore we can produce a map from \(L^q(E) \rightarrow (L^p(E))'\) by \(g \mapsto K_g\).

Theorem 8.4 Let \((E, \mathcal{E}, \mu)\) be a finite measure space and \(p \in (1, \infty)\). The dual space of \(L^p(E)\) is \(L^q(E)\) where \(1/p +1/q = 1\). Furthermore the map defined by \(g \mapsto K_g\) is an isometry.

Remark. This result also holds for arbitrary measure spaces (without the finite assumption). Extending to \(\sigma\)-finite measure spaces is relatively straightforward and then to any measure space is more complicated.

Proof. Remark: This result is similar in spirit to the Riesz representation result that was a non-examinable topic in week 6.

First we note that the map \(g \mapsto K_g\) is linear and \(\|K_g\|_{(L^p)'} \leq \|g\|_q\). Therefore the map is injective we want to show that \(\|K_g\| = \|g\|\) and that it is surjective.

First for the fact that \(\|K_g\| = \|g\|\) we look at the function \(f(x) = sgn(g)|g(x)|^{q-1}\) then \(\mu(|f|^p) = \mu(|g|^q) < \infty\). Therefore we can look at the action of \(K_g\) on \(f\) and we have \(K_g(f) = \mu(|g|^q)\) so we know that \(\|K_g\| \geq K_g(f)/\|f\|_p = \mu(|g|^q)/\mu(|g|^q)^{1/p} = \mu(|g|^q)^{1-1/p} = \|g\|_q\). Therefore \(g \mapsto K_g\) preserves norms.

Now we want to show that this map is surjective. Let us take \(K\) an arbitrary element of \((L^p(E))'\). Since \(\mu(E) < \infty\), \(1_A \in L^p(E)\) for every \(A \in \mathcal{E}\) so we can define a function on \(\mathcal{E}\) by \(k(A)=K(1_A)\). We want to show that \(k\) is a signed measure. \(k(\emptyset)=K(0)=0\) and let \(A_1,A_2,\dots\) be a sequence of disjoint measurable sets. Then \(1_{\bigcup_{j=1}^n A_j} = \sum_{j=1}^n 1_{A_j}\) then \(k(\bigcup_{j=1}^n A_j) = K(1_{\bigcup_{j=1}^n A_j}) = K( \sum_{j=1}^n 1_{A_j}) = \sum_{j=1}^n K(1_{A_j}) = \sum_{j=1}^n k(A_j)\). We also have that \(\| 1_{\bigcup_j A_j} - 1_{\bigcup_{j=1}^n A_j} \|_p \rightarrow 0\) as \(n \rightarrow 0\). Therefore, as \(K\) is a continuous map on \(L^p\) we have \(K(1_{\bigcup_n A_n}) = \sum_n K(1_{A_n})\), so \(k(\bigcup_n A_n) = \sum_n k(A_n)\) Therefore \(k\) is indeed a signed measure. By the Hahn decomposition and the Jordan decomposition we can write \(k = k_+ - k_-\) and there exists \(P \cup N\) a Hahn decomposition with \(k\) being positive on \(P\) and negative on \(N\).

Next we want to show that \(k_+ \ll \mu\) and \(k_- \ll \mu\). If \(A \in \mathcal{E}\) is such that \(\mu(A) = 0\) then \(\mu(A \cap P)=0\) and \(\mu(A \cap N) = 0\) and as \(1_{A \cap P}= 0 \mu\)-a.e. we have \(K(1_{A \cap P}) = K(0) = 0\) and \(K(1_{A \cap N}) = K(0) = 0\). Therefore \(k_+(A) = 0\) and \(k_-(A) = 0\).

Then by the Radon-Nikodym theorem there exists functions \(g_+\) and \(g_-\) such that \(k_+(A) = \mu(g_+1_A)\) and \(k_-(A) = - \mu(g_- 1_A)\). Now let \(g = g_+ - g_-\) we want to show that \(g \in L^q\) and that \(K = K_g\). This is complicated!

Let us define \(E_n\) by \(E_n = \{ x \,:\, |g(x)| \leq n\}\) then \(g1_{E_n}\) is bounded and so in \(L^q\) as \(\mu\) is finite. Then define a linear functional on \(L^p\) by \(K_n(f) = \mu(fg1_{E_n})\) and another by \(\tilde{K}_n (f) = K(f1_{E_n})\). Then if \(A\) is a measurable set we have \(K_n (1_A) = \tilde{K}_n (1_A)\), by linearity if \(h\) is a simple function then \(K_n(h) = \tilde{K}_n(h)\).

We showed in Assignment 4 that given a function \(f \in L^p, \epsilon >0\) there exists a simple function \(h\) with \(\|f-h\|_p \leq \epsilon\). Then we have that \[ |K_n(f) - \tilde{K}_n(f)| \leq |K_n(f) - K_n(h)| + |\tilde{K}_n(f) - \tilde{K}_n (h)| \leq \|K\| \|f-h\|_p + \|g1_{E_n}\|_q \|f-h\|_p \leq (\|K\| + \|g\|_q) \epsilon.\] Since \(\epsilon\) is arbitrary this shows that \(K_n(f) = \tilde{K}_n(f)\). \(\tilde{K}_n(f) = K_{g1_{E_n}}(f)\) so by our isometry we have \(\|\tilde{K}_n\| = \|g1_{E_n}\|_q\). We also have that \(\|\tilde{K}_n\| \leq \|K\|\) as \(\| \tilde{K}_n\| = \sup_{\|f\|_p =1} K_n(f) = \sup_{\|f\|_p = 1} K(f1_{E_n}) \leq \sup_{\|f\|_p=1} K(f) = \|K\|\). Therefore \(\|g1_{E_n}\|_q \leq \|K\|\) therefore \(\|g1_{E_n}\|_p^p = \int |g|^p 1_{E_n} \mu(\mathrm{d}x)\) and then by monotone convergence we get that \(\|g\|_q = \lim_n \|g 1_{E_n}\|_q \leq \|K\|\). Therefore, \(g \in L^q\). Then by exactly the same argument with which we showed \(K_n = \tilde{K}_n\) we have that \(K = K_g\). This concludes the proof in the finite case.

Extending to the sigma finite case works very similarly to previous results.