real-analysis-4

Real Analysis (4)

Sequences and Series of Functions [1]

Uniform Convergence of a Seuqence of Functions

Definition 6.2.1: Pointwise Convergence

For each \(n \in \mathbb{N}\), let \(f_n\) be a function defined on a set \(A \subseteq \mathbb{R}\). The sequence \((f_n)\) of functions converges pointwise on \(A\) to a function \(f\) if, for all \(x \in A\), the sequence of real numbers \(f_{n} (x)\) converges to \(f(x)\).

In this case, we write \(f_n \rightarrow f\), \(\lim_{n \rightarrow \infty} f_n = f\) or \(\lim_{n \rightarrow \infty} f_n(x) = f(x)\).

Definition 6.2.1B: Pointwise Convergence

Let \((f_n)\) be a sequence of functions defined on a set \(A \subseteq \mathbb{R}\). Then, \((f_n)\) convergence pointwise on \(A\) to a limit function \(f\) defined on \(A\) if, for every \(\epsilon > 0\) and \(x \in A\), there exists an \(N \in \mathbb{N}\) (perhaps depends on \(x\)) such that \(|f_{n} (x) - f(x) | < \epsilon\) whenever \(n \geq N\).

Definition 6.2.3: Uniform Convergence (Stronger than Pointwise)

Let \((f_n)\) be a sequence of functions defined on a set \(A \subseteq \mathbb{R}\). Then, \((f_n)\) convergence uniformly on \(A\) to a limit function \(f\) defined on \(A\) if, for every \(\epsilon > 0\), there exists an \(N \in \mathbb{N}\) such that \(|f_{n} (x) - f(x) | < \epsilon\) whenever \(n \geq N\) and \(x \in A\).

Similar to uniform continuous, this is a stronger, which involves finding \(N \in \mathbb{N}\) for a given \(\epsilon > 0\) for all \(x \in A\).

Graphically speaking, the uniform convergence of \(f_n\) to a limit \(f\) on a set \(A\) can be seen as there exists a point in the sequence after which each \(f_n\) is completely contained in the \(\epsilon\)-strip.

Theorem 6.2.5: Cauchy Criterion for Uniform Convergence

A sequence of functions \((f_n)\) defined on a set \(A \subseteq \mathbb{R}\) converges uniformly on \(A\) if and only if for every \(\epsilon > 0\) there exists an \(N \in \mathbb{N}\) such that \(|f_n(x) - f_m(x)| < \epsilon\) whenever \(m, n \geq N\) for all \(x \in A\).

Theorem 6.2.6: Continuous Limit Theorem

Let \((f_n)\) be a sequence of functions defined on \(A \subseteq \mathbb{R}\) that converges uniformly on \(A\) to a function \(f\). If each \(f_n\) is continuous at \(c \in A\), then \(f\) is continuous at \(c\).

Uniform Convergence and Differentiation

Theorem 6.3.1: Differentiable Limit Theorem

Let \(f_n \rightarrow f\) pointwise on the closed interval \([a, b]\), and assume that each \(f_n\) is differentiable. If \((f^{\prime}(c))\) converges uniformly on \([a, b]\) to a function \(g\), then the function \(f\) is differentiable and \(f^{\prime} = g\).

Theorem 6.3.2

Let \((f_n)\) be a sequence of differentiable functions defined on the closed interval \([a, b]\), and assume \((f^{\prime}_n)\) converges uniformly on \([a, b]\). If there exists a point \(x_0 \in [a, b]\), where \(f_{n} (x_0)\) is convergent, then \((f_n)\) converges uniformly on \([a, b]\).

Theorem 6.3.3: Combine Theorem 6.3.1 and 6.3.2

Let \((f_n)\) be a sequence of differentiable functions defined on the closed interval \([a, b]\), and assume \((f^{\prime}_n)\) converges uniformly on \([a, b]\) to a function \(g\) on \([a, b]\). If there exists a point \(x_0 \in [a, b]\), where \(f_{n} (x_0)\) is convergent, then \((f_n)\) converges uniformly on \([a, b]\). Moreover, the limit function \(f = \lim_{n\rightarrow \infty} f_n\) is differentiable and satisfies \(f^{\prime} = g\).

Series of Functions

Definition 6.4.1: Pointwise Convergence and Uniform Convergence of Series of Functions

For each \(n \in \mathbb{N}\), let \(f_n\) and \(f\) be functions defined on a set \(A \subseteq \mathbb{R}\). The infinite series:

\[\sum^{\infty}_{n=1} f_n = f_1 + f_2 + .... \]

converges pointwise on \(A\) to \(f\) if the sequence \(s_k\) of partial sums defined by:

\[s_k = f_1 + f_2 + ... + f_k\]

converges pointwise to \(f\). The series converges uniformly on \(A\) to \(f\) if the sequence \(s_k\) converges uniformly on \(A\) to \(f\). In either case, we can write:

\[f = \sum^{\infty}_{n=1} f_n\]

or

\[f(x) = \sum^{\infty}_{n=1} f_n(x)\]

Theorem 6.4.2: Term-by-Term Continuity Theorem

Let \(f_n\) be continuous functions defined on a set \(A \subseteq \mathbb{R}\), and assume \(\sum^{\infty}_{m=1} f_n\) converges uniformly on \(A\) to a function \(f\). Then \(f\) is continuous on \(A\).

Theorem 6.4.3: Term-by-Term Differentiability Theorem

Let \(f_n\) be differentiable functions defined on an interval \(A\), and assume \(\sum^{\infty}_{n=1} f^{\prime}_n\) converges uniformly to a limit function \(g\) on \(A\). If there exists a point \(x_0 \in [a, b]\) where \(\sum^{\infty}_{n=1} f_n (x_0)\) converges, then the series \(\sum^{\infty}_{n=1} f_n\) converges uniformly to a differentiable function \(f\) satisfying \(f^{\prime} = g\) on \(A\).

Theorem 6.4.4: Cauchy Criterion for Uniform Convergence of Series

A series \(\sum^{\infty}_{n=1} f_n\) converges uniformly on \(A \subseteq \mathbb{R}\) if and only if for every \(\epsilon > 0\) there exists an \(N \in \mathbb{N}\) s.t

\[|f_{m+1} (x) + f_{m+2} (x) + ... + f_{n} (x) | < \epsilon\]

whenever \(n > m \geq N\) and \(x \in A\)

Corollary 6.4.5: Weierstrass M-Test

For each \(n \in \mathbb{N}\), let \(f_n\) be a function defined on a set \(A \subseteq \mathbb{R}\), and let \(M_n > 0\) be a real number satisfying:

\[|f_n(x)| \leq M_n\]

for all \(x \in A\). If \(\sum^{\infty}_{n=1} M_n\) converges, then \(\sum^{\infty}_{n=1} f_n\) converges uniformly on \(A\).

Power Series

\[f(x) = \sum^{\infty}_{n=0} a_n x^n = a_0 + a_1x + a_2x^2 + a_3 x^3 + ...\]

Theorem 6.5.1

If a power series \(\sum^{\infty}_{n=0} a_n x^n\) converges at some point \(x_0 \in \mathbb{R}\), then it converges absolutely for any \(x\) satisfying \(|x| < |x_0|\).

Theorem 6.5.2

If a power series \(\sum^{\infty}_{n=0} a_n x^n\) converges absolutely at a point \(x_0\), then it converges uniformly on the closed interval \([-c, c]\), where \(c=|x_0|\).

Lemma 6.5.3: Abel's Lemma

Let \(b_n\) satisfy \(b_1 \geq b_2 \geq b_3 \geq .... \geq 0\), and let \(\sum^{\infty}_{n=1} a_n\) be a series for which the partial sums are bounded. In other words, assume there exists \(A > 0\) s.t

\[|a_1 + a_2 + ... + a_n| \leq A\]

for all \(n \in \mathbb{N}\). Then for all \(n \in \mathbb{N}\),

$$|a_1b_1 + a_2b_2 + .... + a_nb_n|

Theorem 6.5.4: Abel's Theorem

Let \(g(x) = \sum^{\infty}_{n=0} a_nx^n\) be a power series that converges at the point \(x = R > 0\). Then the series converges uniformly on the interval \([0, R]\). A similar result holds if the series converges at \(x = -R\).

Theorem 6.5.5

If a power series converges pointwise on the set \(A \subseteq \mathbb{R}\), then it converges uniformly on any compact set \(K \subseteq A\).

Theorem 6.5.6

If \(\sum^{\infty}_{n=0}a_n x^n\) converges for all \(x \in (-R, R)\), then the differentiated series \(\sum^{\infty}_{n=1} na_n x^{n-1}\) converges at each \(x\in (-R, R)\) as well. Consequently, the convergence is unifrom on compact sets contained in \((-R, R)\).

Theorem 6.5.7

Assume

\[f(x) = \sum^{\infty}_{n=0} a_n x^n\]

converges on an interval \(A \subseteq \mathbb{R}\). The function \(f\) is continuous on \(A\) and differentiable on any open interval \((-R, R) \subseteq A\). The derivative is given by:

\[f^{\prime} (x) = \sum^{\infty}_{n=1} na_nx^{n-1}\]

Moreover, \(f\) is infinitely differentiable on \((-R, R)\), and the successive derivatives can be obtained via term-by-term differentiation of the appropriate series.

Taylor Series

Given an infinitely differentiable function \(f\) defined on some interval centered at zero, the idea is to assume that \(f\) has a power series expansion and deduce what the coefficients must be. (Not every infinitely differentiable function can be represented by its taylor series)

Theorem 6.6.2: Taylor's Formula Centered at Zero (Maclaurin Series)

Let

\[f(x) = a_0 + a_1x + a_2x^2 + ...\]

be defined on some nontrivial interval centered at zero. Then:

\[a_n = \frac{f^{(n)}(0)}{n!}\]

Where \(f^{(n)}(0)\) is the \(n\)th derivative of \(f\) evaluate at \(0\).

This formula assumes that the function \(f\) has a power series representation.

Theorem 6.6.3: Lagrange's Remainder Theorem

Let \(f\) be differentiable \(N + 1\) times on \((-R, R)\), define \(a_n = \frac{f^{n} (0)}{n!}\) for \(n=0, 1, ... , N\), and let:

\[S_N(x) = a_0 + a_1 x + a_2 x^2 + ... + a_N x^N\]

Given \(x \neq 0\) in \((-R, R)\), there exists a point \(c\) satisfying \(|c| < |x|\) where the error function \(E_N(x) = f(x) - S_N (x)\) satisfies:

\[E_N(x) = \frac{f^{(N+1)} (c)}{(N+1)!} x^{N+1}\]

Definition 6.6.7: Taylor Series Centered at \(a \neq 0\)

If \(f\) is defined in some neighborhood of \(a \in \mathbb{R}\) and infinitely differentiable at \(a\), then the Taylor series expansion around \(a\) takes the form:

\[\sum^{\infty}_{n=0}c_n (x - a)^n\]

where \(c_n = \frac{f^{(n)} (a)}{n!}\)

The Weierstrass Approximation Theorem

Theorem 6.7.1: Weierstrass Approximation Theorem

Let \(f: [a, b] \rightarrow \mathbb{R}\) be continuous. Given \(\epsilon > 0\), there exists a polynomial \(p(x)\) satisfying:

\[|f(x) - p(x) | < \epsilon\]

for all \(x \in [a, b]\).

In other words, every continuous function on a closed interval can be uniformly approximated by a polynomial.

Definition 6.7.2: Polygonal

A continuous function \(\phi: [a, b] \rightarrow \mathbb{R}\) is polygonal if there is a partition:

\[a = x_0 < x_1 < x_2 < .... < x_n = b\]

of \([a, b]\) such that \(\phi\) is linear on each subinterval \([x_{i-1}, x_i]]\) where \(i=1, ..., n\)

Theorem 6.7.3

Let \(f: [a, b] \rightarrow \mathbb{R}\) be continuous. Given \(\epsilon > 0\), there exists a polygonal function \(\phi\) satisfying:

\[|f(x) - \phi(x) | < \epsilon\]

The Riemann Integral

The Definition of the Riemann Integral

Throughout this section, it is assumed that we are working with a bounded function \(f\) on a closed interval \([a, b]\), meaning that there exists an \(M > 0\) s.t \(|f(x)| \leq M, \; \forall x \in [a, b]\)

Definition 7.2.1: Partition, Lower Sum, Upper Sum

A partition \(P\) of \([a, b]\) is a finite set of points from \([a, b]\) that includes both \(a\) and \(b\). A partition \(P = \{x_0, ...., x_n\}\) is listed in increasing order:

\[a=x_0 < x_1 < x_2 < ... < x_n = b\]

For each subinterval \([x_{k-1}, x_k]\) of \(P\), let:

\[m_k = \inf\{f(x): x \in [x_{k-1}, x_k]\}\]

and

\[M_k = \sup\{f(x): x\in[x_{k-1}, x_k]\}\]

Since the function is bounded, the supreme and infimum always exists.

The lower sum of \(f\) w.r.t \(P\) is given by:

\[L(f, P) = \sum^{n}_{k=1} m_k (x_k - x_{k-1})\]

The upper sum of \(f\) w.r.t \(P\) is given by:

\[U(f, P) = \sum^{n}_{k=1} M_k (x_k - x_{k-1})\]

For a particular partition \(P\), it is clear that:

\[U(f, P) \geq L(f, P)\]

Definition 7.2.2: Refinement

A partition \(Q\) is a refinement of a partition \(P\) if \(Q\) contains all of the points of \(P\) if \(Q\) contains all of the points of \(P\); that is, if \(P \in Q\).

Lemma 7.2.3

If \(P \subseteq Q\), then \(L(f, P) \leq L(f, Q)\), and \(U(f, P) \geq U(f, Q)\).

Lemma 7.2.4

If \(P_1, P_2\) are any two partitions of \([a, b]\), then \(L(f, P_1) \leq U(f, P_2)\).

We can think of upper sum as an overestimate for the value of the integral and a lower sum as an underestimate. As the partitions get more refined, the upper sums get potentially smaller while the lower sums get potentially larger.

Definition 7.2.5: Upper Integral, Lower Integral

Let \(\mathbf{P}\) be the collection of all possible partitions of the interval \([a, b]\). The upper integral of \(f\) is defined to be:

\[U(f) = \inf\{U(f, P): P \in \mathbf{P}\}\]

In a similar way, define the lower integral of \(f\) by:

\[L(f) = \sup\{L(f, P): P\in\mathbf{P}\}\]

Lemma 7.2.6

For any bounded function \(f\) on \([a, b]\), it is always the case that \(U(f) \geq L(f)\).

Definition 7.2.7: Riemann Integrability

A bounded function \(f\) defined on the interval \([a, b]\) is Riemann-integrable if \(U(f) = L(f)\). In this case, we define \(\int^{b}_{a} f\) or \(\int^{a}_{b} f(x) dx\) to be this common value, namely:

\[\int^b_{a} f = U(f) = L(f)\]

Criteria for Integrability

Theorem 7.2.8: Integrability Criterion

A bounded function \(f\) is integrable on \([a, b]\) if and only if, for every \(\epsilon > 0\), there exists a partition \(P_{\epsilon}\) of \([a, b]\) s.t

\[U(f, P_\epsilon) - L(f, P_{\epsilon}) < \epsilon\]

Theorem 7.2.9

If \(f\) is continuous on \([a, b]\), then it is integrable.

Theorem 7.3.2

If \(f: [a, b] \rightarrow \mathbb{R}\) is bounded, and \(f\) is integrable on \([c, b]\) for all \(c \in (a, b)\), then \(f\) is integrable on \([a, b]\). An analogous result holds at the other end point.

Properties of the Integral

Theorem 7.4.1

Assume \(f: [a, b] \rightarrow \mathbb{R}\) is bounded, and let \(c \in (a, b)\). Then, \(f\) is integrable on \([a, b]\) if and only if \(f\) is integrable on \([a, c]\) and \([c, b]\). In this case, we have:

\[\int^{b}_{a} f = \int^{c}_{a} f + \int^{b}_{c} f\]

Theorem 7.4.2

Assume \(f, g\) are integrable functions on the interval \([a, b]\).

  1. The function \(f + g\) is integrable on \([a, b]\) with \(\int^{b}_{a} (f + g) = \int^{b}_{a} f + \int^{b}_{a} g\).
  2. For \(k \in \mathbb{R}\), the function \(kf\) is integrable with \(\int^{b}_{a} kf = k \int^{b}_{a} f\).
  3. If \(m \leq f(x) \leq M\) on \([a, b]\), then \(m(b - a) \leq \int^{b}_{a} f \leq M (b - a)\).
  4. If \(f(x) \leq g(x)\) on \([a, b]\), then \(\int^{b}_{a} f \leq \int^b_a g\)
  5. The function \(|f|\) is integrable and \(|\int^{b}_{a} f| \leq \int^{b}_{a} |f|\)

Definition 7.4.3

If \(f\) is integrable on the interval \([a, b]\), define:

\[\int^a_{b} f = - \int^b_a f\]

Also, for \(c \in [a, b]\), definte:

\[\int^c_c f = 0\]

Theorem 7.4.4: Integrable Limit Theorem

Assume that \(f_n \rightarrow f\) uniformly on \([a, b]\) and that each \(f_n\) is integrable. Then \(f\) is integrable and:

\[\lim_{n \rightarrow \infty} f_n = \int^b_{a} f\]

The Fundamental Theorem of Calculus

Theorem 7.5.1: Fundamental Theorem of Calculus

  1. If \(f: [a, b] \rightarrow \mathbb{R}\) is integrable, and \(\mathbb{F}: [a, b] \rightarrow \mathbb{R}\) satisfies \(F^{\prime} (x) = f(x) \;\; \forall x \in [a, b]\) satisfies \(F^{\prime} (x) = f(x), \; \forall x \in [a, b]\), then:

    \[\int^b_a f = F(b) - F(a)\]

  2. Let \(g: [a, b] \rightarrow \mathbb{R}\) be integrable, and for \(x \in [a, b]\), define:

    \[G(x) = \int^x_a g\]

    Then \(G\) is continuous on \([a, b]\). If \(g\) is continuous at some point \(c \in [a, b]\), then \(G\) is differentiable at \(c\) and \(G^{\prime}(c) = g(c)\)

    \[G^{\prime}(x) = \frac{d}{dx} \int^x_a g = g(x)\]

Definition 7.6.1: Measure Zero

A set \(A \subseteq \mathbb{R}\) has measure zero if, for all \(\epsilon > 0\), there exists a countable collection of open intervals \(O_n\) with the property that \(A\) is contained in the union of all of the intervals \(O_n\) and the sum of the lengths of all of the intervals is less than or equal to \(\epsilon\). More precisely, if \(|O_n|\) refers to the length of the interval \(O_n\), then we have:

\[A \subseteq \bigcup^{\infty}_{n=1} O_n\]

and

\[\sum^{\infty}_{n=1} | O_n | \leq \epsilon\]

Consider a finite set \(A = \{a_1, ..., a_n\}\). To show that \(A\) has measure zero, let \(\epsilon > 0\) be arbitrary. For each \(1 \leq n < N\), construct the interval: \[G_n = (a_n - \frac{\epsilon}{2N}, a_n + \frac{\epsilon}{2N})\]

Clearly, \(A\) is contained in the union of these intervals, and \[\sum^N_{n=1} | G_n | = \sum^N_{n=1}\frac{\epsilon}{N} = \epsilon\]

Definition 7.6.3: \(\alpha\)-Continuity

Let \(f\) be defined on \([a, b]\), and let \(\alpha > 0\). The function \(f\) is \(\alpha\)-continuous at \(x \in [a, b]\) if there exists \(\delta > 0\) s.t for all \(y, z \in (x - \delta, x + \delta)\), it follows that \(|f(y) - f(z)| < \alpha\).

Let \(f\) be a bounded function on \([a, b]\), for each \(\alpha > 0\), defined \(D^{\alpha}\) to be the set of points in \([a, b]\), where the function \(f\) fails to be \(\alpha\)-continuous;

Definition 7.6.4: Uniformly \(\alpha\)-continuous

A function \(f: A \rightarrow \mathbb{R}\) is uniformly \(\alpha\)-continuous on \(A\), if there exists a \(\delta > 0\) s.t whenever \(x\) and \(y\) are points in \(A\) satisfying \(|x - y| < \delta\), it follows that \(|f(x) - f(y)| < \alpha\)..

Theorem 7.6.5 Lebesgue's Theorem

Let \(f\) be a bounded function defined on the interval \([a, b]\). Then, \(f\) is Riemann-integrable if and only if the set of points where \(f\) is not continuous has measure zero.

Definition 8.4.3: Improper Riemann Integrals

Assume \(f\) is defined on \([a, \infty)\) and integrable on every interval of the form \([a, b]\). Then define \(\int^{\infty}_{a} f\) to be:

\[\lim_{b \rightarrow \infty} \int^{b}_{a} f\]

provided the limit exists, in this case we say that the improper integral \(\int^{\infty}_a f\) converges.

Interchange of Limits [2]

Theorem 8.2.4

Let \((f_n)\) be a sequence of functions be defined on \([a, b] \in \mathbb{R}\) and suppose that \((f_n)\) converges uniformly on \([a, b]\) to \(f\). Then \(f \in \mathbb{R}\) and:

\[\int^{a}_{b} f = \lim_{n \rightarrow \infty} \int^b_a f_n\]

We can switch the integral and the limit.

Theorem 8.2.5: Bounded Convergence Theorem

Let \((f_n)\) be a sequence defined on \(\mathbb{R}[a, b]\) that converges on \([a, b]\) to a function \(f \in \mathbb{R} [a, b]\). Suppose also that there exists a \(B > 0\) s.t \(|f_n(x)| \leq B, \; \forall x \in [a, b]\), \(n \in \mathbb{N}\). Then:

\[\int^{a}_{b} f = \lim_{n \rightarrow \infty} \int^b_a f_n\]

Theorem 8.2.6: Dini's Theorem

Suppose that \((f_n)\) is a monotone sequence of continuous functions on \(I:=[a, b]\) that converges on \(I\) to a continuous function \(f\). Then the convergence of the sequence is uniform.

The Exponential Function

Theorem 8.3.1:

There exists a function \(E: \mathbb{R} \rightarrow \mathbb{R}\) s.t:

  1. \(E^{\prime} (x) = E(x)\)
  2. \(E(0) = 1\)
Partial Proof of Theorem 8.3.1

We define a sequence \((E_n)\) of continuous functions as follows:

  1. \(E_{1} (x) := 1 + x\)
  2. \(E_{n + 1} (x) := 1 + \int^x_0 E_n(t) dt\)

for all \(n \in \mathbb{N}, x \in \mathbb{R}\), we can also show that:

\[E_{n} (x) = 1 + \frac{x}{1!} + \frac{x^2}{2!} + ... + \frac{x^n}{n!}\]

We can show that the limiting function \(E = \lim_{n\rightarrow \infty}\) exists and satisfies all two points of theorem 8.3.1.

Corollary 8.3.2

The function \(E\) has a derivative of every order and \(E^(n) (x) = E(x), \; \forall n \in \mathbb{N}, x \in \mathbb{R}\).

Corollary 8.3.3

If \(x > 0\), then \(1 + x < E(x)\)

Theorem 8.3.4

The function \(E: \mathbb{R} \rightarrow \mathbb{R}\) that satisfies theorem 8.3.1 (i) and (ii) is unique.

Definition 8.3.5

The unique function \(E: \mathbb{R} \rightarrow \mathbb{R}\), such that \(E^{\prime} (x) = E(x)\) for all \(x \in \mathbb{R}\) and \(E(0) = 1\) is called the exponential function. The number \(e := E(1)\) is called Euler's number:

\[\exp(x) := E(x)\]

or

\[e^x := E(x)\]

for \(x \in \mathbb{R}\)

Theorem 8.3.6

The exponential function satisfies the following properties:

  1. \(E(x) \neq 0, \; \forall x \in \mathbb{R}\)
  2. \(E(x + y) = E(x)E(y), \; \forall x, y \in \mathbb{R}\)
  3. \(E(r) = e^r, \;\forall r \in \mathbb{Q}\)

Theorem 8.3.7

The exponential function \(E\) is strictly increasing on \(\mathbb{R}\) and has range equal to \(y \in \mathbb{R}: y > 0\). Further, we have:

  1. \(\lim_{x \rightarrow \infty} E(x) = 0\)
  2. \(\lim_{x \rightarrow \infty} E(x) = \infty\)

The Logarithm Function

Definition 8.3.8: Logarithm Function

The function inverse to \(E: \mathbb{R} \rightarrow \mathbb{R}\) is called the logarithm (natural logarithm) denoted by \(L\) or by \(\ln\).

Since they are inverse, we have:

\[(L \circ E) (x) = x, \; \forall x \in \mathbb{R}\]

and

\[(E \circ L) (y) = y, \; \forall y \in \mathbb{R}, y > 0\]

Theorem 8.3.9

The logarithm is a strictly increasing function \(L\) with doman \(x \in \mathbb{R}: x > 0\) and range \(\mathbb{R}\). The derivative of \(L\) is given by:

  1. \(L^{\prime} (x) = \frac{1}{x}\)
  2. \(L(xy) = L(x) L(y)\)
  3. \(L(1) = 0, L(e) = 1\)
  4. \(L(x^r) = rL(x), \; \forall r \in \mathbb{Q}\)
  5. \(\lim_{x \rightarrow 0^+} L(x) = -\infty\) and \(\lim_{x \rightarrow \infty} L(x) = \infty\)

Power Function

Definition 8.3.10: Power Function

If \(\alpha \in \mathbb{R}\) and \(x > 0\), the number \(x^\alpha\) is defined to be

\[x^\alpha := e^{\alpha \ln x} = E(\alpha L(x))\]

The function \(x \rightarrow x^\alpha\) for \(x > 0\) is called the power function with exponent \(\alpha\).

Theorem 8.3.11

If \(\alpha \in \mathbb{R}\) and \(x, y\) belong to \((0, \infty)\), then:

  1. \(1^\alpha = 1\)
  2. \(x^\alpha > 0\)
  3. \((xy)^\alpha = x^\alpha y^\alpha\)
  4. \((\frac{x}{y})^{\alpha} = \frac{x^\alpha}{y^\alpha}\)

Theorem 8.3.12

If \(\alpha, \beta \in \mathbb{R}\) and \(x \in (0, \infty)\), then:

  1. \(x^{\alpha + \beta} = x^{\alpha} + x^{\beta}\)
  2. \((x^\alpha)^\beta = x^{\alpha\beta} = (x^{\beta})^\alpha\)
  3. \(x^{-\alpha} = \frac{1}{x^\alpha}\)
  4. If \(\alpha < \beta\), then \(x^\alpha < x^\beta\) for \(x > 1\)

Theorem 8.3.13

Let \(\alpha \in \mathbb{R}\). Then the function \(x \rightarrow x^\alpha\) on \((0, \infty)\) to \(\mathbb{R}\) is continuous and differentiable, and

\[D x^\alpha = \alpha x^{\alpha - 1}\]

If \(\alpha > 0\) the power function is strictly increasing, if \(\alpha < 0\), the power function is strictly decreasing.

Function \(\log_a\)

if \(a > 0\), \(a \neq 1\), it is sometimes useful to define the function \(\log_a\)

Definition 8.3.14

Let \(a > 0, a \neq 1\). We define:

\[\log_a (x) := \frac{\ln x}{\ln a}\]

for \(x \in (0, \infty)\)

This is called logarithm of \(x\) to the base \(a\).

The Trigonometric Function

Theorem 8.4.1

There exist functions \(C: \mathbb{R} \rightarrow \mathbb{R}\) and \(S: \mathbb{R} \rightarrow \mathbb{R}\) such that:

  1. \(C^{\prime\prime} (x) = - C(x)\) and \(S^{\prime\prime} (x) = - S(x)\) for all \(S: \mathbb{R} \rightarrow \mathbb{R}\).
  2. \(C(0) = 1, C^{\prime} (0)\), and \(S(0) = 0, S^{\prime} (0) = 1\)
Partial Proof of Theorem 8.4.1

Define the sequences \((C_n), (S_n)\) of continuous functions inductively as:

  1. \(C_1 (x) := 1, \; S_1 (x) := x\)
  2. \(S_n(x) := \int^x_0 C_n(t) dt\)
  3. \(C_{n+1} (x) := 1 - \int^x_0 S_n(t) dt\)

for all \(n \in \mathbb{N}, x \in \mathbb{R}\). By induction, we have:

\[C_{n+1} (x) = 1 - \frac{x^2}{2!} + \frac{x^4}{4!} - ... + (-1)^n \frac{x^{2n}}{(2n)!}\] \[S_{n+1} (x) = x - \frac{x^3}{3!} + \frac{x^5}{t!} - .... + (-1)^n \frac{x^{2n+1}}{(2n+1)!}\]

Then the sequences converges uniformly to some function \(C, S\) that satisfies all the properties of theorem 8.4.1.

Corollary 8.4.2

If \(C, S\) are the functions in theorem 8.4.1, then:

  1. \(C^{\prime} (x) = -S(x)\) and \(S^{\prime} (x) = C(x)\) for all \(x \in \mathbb{R}\).

Moreover, these functions have derivatives of all orders.

Corollary 8.4.3

The functions \(C, S\) satisfy the pythagorean identity:

  1. \((C(x))^2 + (S(x))^2 = 1\)

Theorem 8.4.4

The function \(C, S\) satisfying properties of theorem 8.4.1 are unique.

Definition 8.4.5

The unique functions \(C: \mathbb{R} \rightarrow \mathbb{R}\), \(S: \mathbb{R} \rightarrow \mathbb{R}\) satisfying properties in theorem 8.4.1 are called cosine function and sine function:

\[\cos x := C(x)\] \[\sin x := S(x)\]

Theorem 8.4.6

If \(f: \mathbb{R} \rightarrow \mathbb{R}\) is such that

\[f^{\prime\prime} (x) = -f(x), \;\; \forall x \in \mathbb{R}\]

then there exist real numbers \(\alpha, \beta\) s.t

\[f(x) = \alpha C(x) + \beta S(x)\]

Theorem 8.4.7:

The function \(C\) is even and \(S\) is odd in the sense that:

  1. \(C(-x) = C(x)\) and \(S(-x) = - S(x)\)

If \(x, y \in \mathbb{R}\), then we have the addition formula:

\[C(x + y) = C(x) C(y) - S(x)S(y)\] \[S(x + y) = S(x)C(y) + C(x)S(y)\]

Ref

[1] Understanding Analysis by Stephen Abbott

[2] Introduction to Real Analysis by G.Bartle, R.Sherbert