Exam 6 (43–49)¶
2000 Nov¶
Instructions Do as many problems as you can. Complete solutions to five problems would be considered a good performance.
State the inverse function theorem. 1
Suppose \(L: ℝ^3 → ℝ^3\) is an invertible linear map and that \(g: ℝ^3 → ℝ^3\) has continuous first order partial derivatives and satisfies \(\|g(x)\| ≤ C\|x\|^2\) for some constant \(C\) and all \(x ∈ ℝ^3\). Here \(\|x\|\) denotes the usual Euclidean norm on \(ℝ^3\). Prove that \(f(x) = L(x) + g(x)\) is locally invertible near 0.
Let \(f\) be a differentiable real valued function on the interval \((0,1)\), and suppose the derivative of \(f\) is bounded on this interval.
Prove the existence of the limit \(L = \lim_{x → 0^+} f(x)\).
Let \(f\) and \(g\) be Lebesgue integrable functions on \([0,1]\), and let \(F\) and \(G\) be the integrals
Use Fubini’s theorem and/or Tonelli’s theorem to prove that
Other approaches to this problem are possible, but credit will be given only to solutions based on these theorems.
Let \((X, 𝔐, μ)\) be a finite measure space and suppose \(ν\) is a finite measure on \((X, 𝔐)\) that is absolutely continuous with respect to \(μ\). Prove that the norm of the Radon-Nikodym derivative \(f = \left[\frac{dν}{dμ}\right]\) is the same in \(L_∞ (μ)\) as it is in \(L_∞(ν)\).
Suppose that \(\{f_n\}\) is a sequence of (Lebesgue) measurable functions on \([0,1]\) such that \(\lim\limits_{n→∞} ∫_0^1 |f_n| \, dx = 0\) and there is an integrable function \(g\) on \([0,1]\) such that \(|f_n|^2 ≤ g\), for each \(n\). Prove that \(\lim\limits_{n→∞}∫_0^1 |f_n|^2 \, dx =0\).
Denote by \(\mathcal{P}_e\) the family of all even polynomials. Thus a polynomial \(p\) belongs to \(\mathcal{P}_e\) if and only if \(p(x) = \frac{p(x) + p(-x)}{2}\) for all \(x\). Determine, with proof, the closure of \(\mathcal{P}_e\) in \(L_1[-1,1]\). You may use without proof the fact that continuous functions on \([-1,1]\) are dense in \(L_1[-1,1]\).
Suppose that \(f\) is real valued and integrable with respect to Lebesgue measure \(m\) on and that there are real numbers \(a<b\) such that
for all open sets \(U\) in \(ℝ\). Prove that \(a ≤ f(x) ≤ b\) a.e.
Solution to Problem 43
A statement of the inverse function theorem appears in the appendix of theorems.
First note that \(L\) and \(g\) both have continuous first order partial derivatives; i.e., \(L, g ∈ C^1(ℝ^3)\). Therefore, the derivative of \(f = L + g\),
\[f'(x) \triangleq J_f(x) \triangleq \left(\frac{\partial f_i}{\partial x_j}\right)_{i,j=1}^3\]exists. Furthermore, \(J_f(x)\) is continuous in a neighborhood of the zero vector, because this is true of the partials of \(g(x)\), and the partials of \(L(x)\) are the constant matrix \(L\). Therefore, \(f ∈ C^1(ℝ^3)\). By the IFT, then, we need only show that \(f'(0)\) is invertible. Since \(f'(x) = L + g'(x)\), we must show \(f'(0) = L + g'(0)\) is invertible. Consider the matrix \(g'(0) = J_g(0)\). We claim, \(J_g(0)=0\). Indeed, if \(x_1, x_2, x_3\) are the elementary unit vectors (also known as i, j, k), then the elements of \(J_g(0)\) are
(21)¶\[\frac{∂ g_i}{∂ x_j}(0) = \lim_{h→0} \frac{g_i(0 + h x_j) - g_i(0)}{h} = \lim_{h→0} \frac{g_i(h x_j)}{h}.\]The second equality follows by the hypothesis that \(g\) is continuous and satisfies \(\|g(x)\| ≤ C\|x\|^2\), which implies that \(g(0) =0\). Finally, to show that (21) is zero, consider
\[|g_i(h x_j)| ≤ \|g(h x_j)\| ≤ C \|h x_j\|^2 = C|h|^2,\]which implies
\[\frac{|g_i(h x_j)|}{|h|} ≤ C \frac{|h x_j|^2}{|h|} = C|h| → 0, \text{ as } h→0.\]This proves that \(f'(0) = L\), which is invertible by assumption, so the inverse function theorem implies that \(f(x)\) is locally invertible near \(0\).
☐
Solution to Problem 51
First consider the sequence \(\{x_n\} := \{\frac{1}{n}\}\) for \(n = 2, 3, 4, \dots\).
Claim 1. \(\{f(x_n)\}\) is a Cauchy sequence.
Proof. Observe
\[\left| f\bigl( \frac{1}{n+j} \bigr)- f\bigl( \frac{1}{n} \bigr) \right| = \left| \frac{ f( \frac{1}{n+j} ) - f( \frac{1}{n} ) } { \frac{1}{n+j} - \frac{1}{n} } \right| \left| \frac{1}{n+j} - \frac{1}{n} \right|,\]and
\[\frac{1}{n+j} - \frac{1}{n} = \frac{n - (n + j)}{n(n+j)} = \frac{-j}{n+j}.\]By the mean value theorem, there exists \(c ∈ \left[\frac{1}{n+j},\frac{1}{n}\right]\) such that
\[\left| f'(c)\right| = \frac{\left| f(\frac{1}{n+j})- f(\frac{1}{n})\right|}{\left| \frac{1}{n+j}- \frac{1}{n} \right|}.\]Therefore,
\[\left| f\left(\frac{1}{n+j}\right)- f\left(\frac{1}{n}\right)\right| ≤ M \left| \frac{1}{n+j}- \frac{1}{n} \right| ≤ M/n,\]so given \(ε > 0\) we have \(|f(1/m)- f(1/n)| < ε\) for all \(m,n ≥ N > M / ε\).
This proves that \(\{f(\frac{1}{n})\}\) is a Cauchy sequence, and it follows that \(\lim_{n→ ∞} f(\frac{1}{n}) = ℓ\) exists in \(ℝ\), since \(ℝ\) is a complete metric space.
☐
Claim 2. \(\lim\limits_{x↘0} f(x) = ℓ\).
Proof. Fix \(ε>0\). We show \(∃ δ>0\) such that \(0 < x < δ\) implies \(|f(x) - ℓ| < ε\).
Given \(0<a<b<1\) there exists \(c∈ [0,1]\) such that \(f(b) - f(a) = f'(c) (b-a)\), and \(|f'(c)|≤ M\) by assumption.
Therefore, \(|f(b) - f(a)| ≤ M |b-a|\) for all \(0<a<b<1\).
Let \(N> 0\) be chosen so large that \(N > ε/(2M)\) and \(n≥ N \, ⟹ \, |f(1/n) - ℓ| < ε/2\).
Then for all \(0< x < 1/n\) we have
\[|f(x) - ℓ| ≤ |f(x) - f(1/n)| + |f(1/n) - f(ℓ)| ≤ M/n + ε/2 < M\frac{ε}{2M} + \frac{ε}{2} = ε.\]☐
☐
Solution to Problem 52
The left-hand side is
Consider \(f̃(t,x):= f(t)\) and \(g̃(t,x):= g(x)\). Then,
are measurable subsets of \([0,1] × [0,1]:= T × X\), so \(f̃\) and \(g̃\) are measurable on \(T× X\).
Therefore, \(f̃ g̃ = f g\) is measurable and we can apply Fubini’s theorem which asserts the following:
then \(f̃ g̃ ∈ L_1([0,1] × [0,1])\) and
So we calculate,
since \(f, g ∈ L_1\) by assumption. Therefore, the hypotheses of Fubini’s theorem are satisfied and we have
as desired.
☐
Solution to Problem 53
It seems an implicit assumption here is that the measures are positive, so we make that assumption in what follows.
Define (as usual) the ∞-norm relative to \(μ\) of a (real- or complex-valued) function \(f\) on \(X\) as follows:
where \(ℝ^∗ = ℝ ∪ \{-∞, ∞\}\) and \(\inf ∅ = ∞\).
Since there are two measures in context, let us denote the norm defined in (22) by \(\|f\|_μ\), and let \(\|f\|_ν\) denote the ∞-norm relative to \(ν\).
The claim to prove is \(\left\|\frac{dν}{dμ}\right\|_μ = \left\|\frac{dν}{dμ}\right\|_ν\).
Here is the proof.
Let \(f := \frac{dν}{dμ}\) so that for each \(E ∈ 𝔐\) we have \(ν E = ∫_E f\, dμ\).
Observe that \(f ≥ 0\), since we assumed \(μ, ν\) positive.
For each \(a ∈ ℝ^∗\), let \(E_a := \{x ∈ X ∣ f(x) > a\}\).
Then \(μ E_a = 0\) only if \(ν E_a = 0\), since \(ν ≪ μ\).
Therefore, \(\{a ∈ ℝ^∗ ∣ μ E_a = 0\} ⊆ \{a ∈ ℝ^∗ ∣ ν E_a = 0\}\), so
Suppose \(\|f\|_ν < \|f\|_μ\). Then there exists \(a ∈ ℝ\) such that \(ν E_a = 0\) and \(μ E_a > 0\), and
a contradiction. Therefore, equality must hold in (23).
☐
Solution to Problem 54
Fix an arbitrarily small \(ε > 0\) and let \(m\) denote Lebesgue measure on \(ℝ\).
Since \(g ∈ L_1[0,1]\), there exists \(A ⊆ [0,1]\) such that \(g\) is bounded on \(A\) and \(∫_{[0,1]-A} g \, dm < ε\).
This and the assumption \(g ≥ |f_n|^2 ≥ 0\) imply that there exists \(M < ∞\) such that \(0≤ g(x) ≤ M\) for all \(x ∈ A\).
Define,
and observe that \(A = A_n^+ ∪ A_n^-\), a disjoint union. Thus, for every \(n∈ ℕ\),
On \(A_n^+ := A ∩ \{|f_n|>1\}\) we have \(|f_n| ≤ |f_n|^2 ≤ g ≤ M\), so
On \(A_n^-:= A ∩ \{|f_n|≤1\}\) we have \(|f_n|^2 ≤ |f_n|\), so
Thus, for every \(n∈ ℕ\),
By assumption \(\lim\limits_{n→∞}∫_{[0,1]}|f_n| \, dm= 0\), so the right-hand side approaches \(∫_{[0,1] - A}g \, dm\) as \(n→∞\) and we conclude that
Since \(ε>0\) was arbitrary, this completes the proof.
☐
Solution to Problem 55
Define the operators \(E, O: L_1[-1,1] → L_1[-1,1]\) as follows: for \(f∈ L_1[-1,1]\),
Denote by \(Λ\) the collection of a.e.-even functions in \(L_1[-1,1]\); that is, \(g ∈ Λ\) iff \(g ∈ L_1[-1,1]\) and \(g(x) = (g(x) + g(-x))/2\) holds for almost every \(x∈ [-1,1]\).
Claim 1. \(E\) is continuous (i.e., bounded).
Proof. Evidently,
\[\begin{split}\|E f\|_1 &= ∫_{-1}^1 |E f(x)|\, dx = ∫_{-1}^1 \left| \frac{f(x) + f(-x)}{2} \right|\, dx \\ &≤ \frac{1}{2} \left( ∫_{-1}^1 |f(x)|\, dx + ∫_{-1}^1 |f(-x)|\, dx\right) = \|f\|_1.\end{split}\]☐
Claim 2. If \(p_n ∈ \mathcal{P}_e\), then \(E p_n = p_n\). (obvious)
Since continuous functions on [-1,1] are dense in L₁[-1,1], for each \(g∈ L_1[-1,1]\) there is a sequence \(\{f_n\}⊆ \overline{C[-1,1]}\) such that \(f_n → g\) in \(L_1[-1,1]\).
Fix \(ε>0\) and let \(N\) be a natural number large enough so that \(\|f_N - g\|_1 < ε/2\).
There exists \(\{p_n\}⊆ \mathcal P\) such that \(\|p_n - f\|_∞ → 0\) as \(n→∞\), so
for large enough \(n ∈ ℕ\).
Since \(E\) is continuous, \(E p_n → E g\).
Claim 3. \(E p_n ∈ \mathcal P_e\). (obvious)
Claim 4. If \(g ∈ L_1[-1,1]\) and \(g\) is even, then \(E g = g\). (obvious)
Claim 5. If \(g ∈ L_1[-1,1]\) and \(g\) is even, then \(O g = 0\) a.e. (why?)
Claim 6. \(O(\bar{\mathcal P}_e) = 0\). (why?)
Claim 7. \(\bar{\mathcal P}_e = \{g ∈ L_1[-1,1]: g \text{ even a.e.}\} = Λ\).
Proof. The inclusion \(⊆\) follows from Claims 5 and 6.
To prove \(⊇\), fix \(g∈ Λ\) and observe that there exists \(\{p_n\} ⊆ \mathcal P_e\) such that \(p_n \xrightarrow{L_1} g\). Therefore, \(E p_n → E g\) by continuity of \(E\), and \(E g = g\) a.e. since \(g\) is even a.e.
Since \(E p_n\) is an even polynomial, we have \(g = \lim\limits_{n→ ∞} E p_n\), so \(g ∈ \bar{\mathcal P}_e\).
☐
Solution to Problem 56
Suppose on the contrary that there exists a measurable set \(E\) of positive measure such that \(f(x) > b\) for all \(x ∈ E\).
Define \(E_n := \{x∈ ℝ ∣ f(x) > b + \frac{1}{n}\}\). Then \(E = ⋃_{n=1}^∞ E_n\) and there exists \(N ∈ ℕ\) such that \(m (E_N) > 0\).
Since \(f ∈ L_1(m)\), for all \(ε>0\) there exists \(δ>0\) such that
If \(ε_0 = 1/(2N)\), and if \(δ_0>0\) is chosen appropriately, then
whenever and \(m(S) < δ_0\).
Now, for all \(δ_1> 0\) there exists an open set \(U\) containing \(E_N\) such that,
From the first of these inequalities along with (24) it follows that
From the second inequality in (25) follows the second inequality in the sequence of calculations below.
Combining (26) and (27) we have, for arbitrary \(δ_1>0\),
We will show that an appropriate choice of \(δ_1>0\) yields
From this and (28) will follow,
contradicting the hypothesis that \(∫_U f \, dm ≤ b⋅m(U)\) holds for every open set \(U\) in \(ℝ\).
Calculate,
We want,
equivalently, \(m(U)/N - (b - 1)⋅ δ_1/N > 1/(2N)\).
Equivalently, we must prove any one (hence all) of the following (equivalent) conditions:
Thus if we choose \(δ_1\) so that it satisfies
then our goal (29) will be satisfied and the desired contradiction will obtain.
A similar argument can be used to show that the existence of a nonnegligible set \(E\) on which \(f < a\) will lead to a contradiction. Therefore, \(a ≤ f(x) ≤ b\) holds for almost every \(x ∈ ℝ\).
☐
Footnotes
- 1
Historically speaking, questions involving the inverse function theorem do not seem very popular. This is one of the few exams on which it appears.
Please email comments, suggestions, and corrections to williamdemeo@gmail.com