Mathematical Analysis

Leture 14

7.3 Basic properties

It is useful to define the integral $\ds\int_a^b f$ even if $a \not\lt b$. Suppose $b \lt a$ and $f \in \mathcal R[b,a]$, then define \begin{equation*} \int_a^b f \coloneqq - \int_b^a f . \end{equation*} For any function $f$, define $\ds \int_a^a f \coloneqq 0 . $



7.3 Basic properties

If the variable $x$ already has another meaning, then we may simply use a different letter. For example, \begin{equation*} \int_a^b f(s)\,ds \coloneqq \int_a^b f(x)\,dx . \end{equation*}






7.3 Basic properties

Theorem 7.3.1. (Linearity) Let $f$ and $g$ be in $\mathcal R[a,b]$ and $\alpha \in \R$.

  1. $\alpha \,f \in\mathcal R[a,b]$ and \begin{equation*} \int_a^b \alpha \,f(x) ~dx = \alpha \int_a^b f(x) ~dx . \end{equation*}
  2. $f+g \in\mathcal R[a,b]$ and \begin{equation*} \int_a^b \bigl( f(x)+g(x) \bigr) ~dx = \int_a^b f(x) ~dx + \int_a^b g(x) ~dx . \end{equation*}

Proof of Theorem 7.3.1, item 1

Consider $\alpha \geq 0$. Let $P$ be a partition of $[a,b]$, and $m_i \coloneqq \inf \bigl\{ f(x) : x \in [x_{i-1},x_i] \bigr\}$. As $\alpha \geq 0$,

$ \ds \inf \bigl\{ \alpha f(x) : x \in [x_{i-1},x_i] \bigr\}\qquad \qquad\qquad\qquad $

$= \alpha \inf \bigl\{ f(x) : x \in [x_{i-1},x_i] \bigr\} $ $ = \alpha m_i . $

Then $\, L(P,\alpha f) $ $\ds = \sum_{i=1}^n \alpha m_i \Delta x_i $ $ \ds = \alpha \sum_{i=1}^n m_i \Delta x_i $ $ \ds = \alpha L(P,f). $

Similarly $\, U(P,\alpha f) $ $ \ds = \alpha U(P,f). \qquad $



Proof of Theorem 7.3.1, item 1

Again, as $\alpha \geq 0$,

$ \ds \underline{\int_a^b} \alpha f(x)\,dx $ $\ds = \sup \, \bigl\{ L(P,\alpha f) : P \text{ a partition of } [a,b] \bigr\} $

$ \ds = \sup \, \bigl\{ \alpha L(P,f) : P \text{ a partition of } [a,b] \bigr\} $

$\ds = \alpha \, \sup \, \bigl\{ L(P,f) : P \text{ a partition of } [a,b] \bigr\} $

$\ds = \alpha \underline{\int_a^b} f(x)\,dx . \qquad\qquad\qquad\qquad\quad $

Similarly, we have $ \; \ds \overline{\int_a^b} \alpha f(x)\,dx = \alpha \overline{\int_a^b} f(x)\,dx . $



Proof of Theorem 7.3.1, item 1

Therefore, for $\alpha\geq 0$, $\alpha f\in \mathcal R[a,b]$ and \[ \int_a^b \alpha f(x)\,dx = \alpha \int_a^b f(x)\,dx . \]

To finish the proof of the first item (for $\alpha \lt 0$), we need to show that $-f$ is Riemann integrable and \[\int_a^b - f(x)\,dx = - \int_a^b f(x)\,dx.\] Try it! 📝 The proof of item 2 is also left as an Exercise 📝.



7.3 Basic properties

Theorem 7.3.2. Let $f \colon [a,b] \to \R$ and $g \colon [a,b] \to \R$ be bounded functions. Then

$\ds \overline{\int_a^b} (f+g) \leq \overline{\int_a^b}f+\overline{\int_a^b}g$

and $\ds\underline{\int_a^b} (f+g) \geq \underline{\int_a^b}f+\underline{\int_a^b}g.$




7.3 Basic properties

Theorem 7.3.3. (Monotonicity) Let $f \colon [a,b] \to \R$ and $g \colon [a,b] \to \R$ be bounded, and $f(x) \leq g(x)$ for all $x \in [a,b]$. Then

$\ds \underline{\int_a^b} f \leq \underline{\int_a^b} g\;\;$ and $\;\;\ds\overline{\int_a^b} f \leq \overline{\int_a^b} g . $





7.3.1 Continuous functions

Theorem 7.3.4. If $f \colon [a,b] \to \R$ is a continuous function, then $f \in \mathcal R[a,b]$.


7.3.1 Continuous functions

Theorem 7.3.4. If $f \colon [a,b] \to \R$ is a continuous function, then $f \in \mathcal R[a,b]$.

Proof. The first crucial observation is that because $f$ is continuous on a closed bounded interval, it is bounded and uniformly continuous. This means that given $\epsilon > 0$, there exists a $\delta > 0$ such that $$\abs{x-y} \lt \delta \Ra \abs{f(x)-f(y)} \lt \frac{\epsilon}{b-a}.$$ Now, let $P$ be a partition of $[a,b]$ where

$\Delta x_k = x_{k} - x_{k-1}\lt \delta\;$ for every subinterval of $P$.



7.3.1 Continuous functions

Theorem 7.3.4. If $f \colon [a,b] \to \R$ is a continuous function, then $f \in \mathcal R[a,b]$.

Proof. Given a particular subinterval $[x_{k-1},x_k]$ of $P$, we know from the Extreme Value Theorem (Thm 5.2.2) that the supremum $M_k = f(z_k)$ is attained for some $z_k \in [x_{k-1},x_k].$ And also the infimum $m_k$ is attained at some point $y_k \in [x_{k-1},x_k].$






7.3.1 Continuous functions

Theorem 7.3.4. If $f \colon [a,b] \to \R$ is a continuous function, then $f \in \mathcal R[a,b]$.

Proof. Given a particular subinterval $[x_{k-1},x_k]$ of P, we know from the Extreme Value Theorem (Thm 5.2.2) that the supremum $M_k = f(z_k)$ is attained for some $z_k \in [x_{k-1},x_k].$ And also the infimum $m_k$ is attained at some point $y_k \in [x_{k-1},x_k].$




7.3.1 Continuous functions

Theorem 7.3.4. If $f \colon [a,b] \to \R$ is a continuous function, then $f \in \mathcal R[a,b]$.

Proof. Given a particular subinterval $[x_{k-1},x_k]$ of P, we know from the Extreme Value Theorem (Thm 5.2.2) that the supremum $M_k = f(z_k)$ is attained for some $z_k \in [x_{k-1},x_k].$ And also the infimum $m_k$ is attained at some point $y_k \in [x_{k-1},x_k].$

But this means that $\abs{z_k-y_k}\lt \delta,$ so

$M_k-m_k $ $= f(z_k)-f(y_k)\quad$

$ \ds \lt \frac{\epsilon}{b-a} .$



7.3.1 Continuous functions

Theorem 7.3.4. If $f \colon [a,b] \to \R$ is a continuous function, then $f \in \mathcal R[a,b]$.

Proof. 👉 $\,M_k-m_k\lt \dfrac{\epsilon}{b-a} $. Consider now the difference

$\ds \overline{\int_a^b} f - \underline{\int_a^b} f $ $\leq U(P,f) - L(P,f) \qquad \qquad \qquad \qquad \quad$

$\ds = \left( \sum_{k=1}^n M_k \Delta x_k \right) - \left( \sum_{k=1}^n m_k \Delta x_k \right) $ $\ds = \sum_{k=1}^n (M_k-m_k) \Delta x_k$

$\ds \lt \frac{\epsilon}{b-a} \sum_{k=1}^n \Delta x_k$ $\ds = \frac{\epsilon}{b-a} (b-a) $ $\ds = \epsilon . \qquad \qquad \;\;\quad $






7.3.1 Continuous functions

Theorem 7.3.4. If $f \colon [a,b] \to \R$ is a continuous function, then $f \in \mathcal R[a,b]$.

Proof. Thus

$\ds \overline{\int_a^b} f - \underline{\int_a^b} f $ $\leq \epsilon .$

Since $\epsilon\gt 0 $ was arbitrary, \[ \overline{\int_a^b} f = \underline{\int_a^b} f \] and $f$ is Riemann integrable on $[a,b].$ $\;\bs$






7.3.1 Continuous functions

Bounded functions with finitely many discontinuities

Lemma 7.3.5. Let $f \colon [a,b] \to \R$ be a bounded function, $\{ a_n \}_{n=1}^\infty$ and $\{b_n \}_{n=1}^\infty$ be sequences such that $a \lt a_n \lt b_n \lt b$ for all $n$, with $\lim_{n\to\infty} a_n = a$ and $\lim_{n\to\infty} b_n = b$. Suppose $f \in \mathcal R[a_n,b_n]$ for all $n$. Then $f \in \mathcal R[a,b]$ and \begin{equation*} \int_a^b f = \lim_{n \to \infty} \int_{a_n}^{b_n} f . \end{equation*}



7.3.1 Continuous functions

Bounded functions with finitely many discontinuities

We say a function $f \colon [a,b] \to \R$ has finitely many discontinuities if there exists a finite set $$S = \{ x_1, x_2, \ldots, x_n \} \subset [a,b],$$ and $f$ is continuous at all points of $[a,b] \setminus S$.

Theorem 7.3.6. Let $f \colon [a,b] \to \R$ be a bounded function with finitely many discontinuities. Then $f \in \mathcal R[a,b]$.




7.3.1 Continuous functions

Bounded functions with finitely many discontinuities

Theorem 7.3.6. Let $f \colon [a,b] \to \R$ be a bounded function with finitely many discontinuities. Then $f \in \mathcal R[a,b]$.


7.3.1 Continuous functions

An example of a bounded function with a dense set of discontinuities

Thomae's function (Popcorn function): Riemann integrable.



7.3.1 Continuous functions

Another example of a bounded function with a dense set of discontinuities

Riemanns's example: $ \;\displaystyle f(x):=\sum_{n=1}^{\infty}\frac{(nx)}{n^2} $ where $(nx)= nx - \text{round} (nx) $.



7.3.2 More on integrable functions

Theorem 7.3.7. Let $f \colon [a,b] \to \R$ be Riemann integrable. Let $g \colon [a,b] \to \R$ be such that $f(x) = g(x)$ for all $x \in [a,b] \setminus S$, where $S$ is a finite set. Then $g$ is Riemann integrable and \begin{equation*} \int_a^b g = \int_a^b f. \end{equation*}





7.3.2 More on integrable functions

Theorem 7.3.8. Let $f:[a,b]\ra \R$ be a monotone function. Then $f\in \mathcal R[a,b]$.


The Integral: A Crux for Analysis

There are many theories of the integral. The Riemann integral is relatively simple, and it does the job nicely.

It is certainly the most common and widely-used integral in the world today.

Mathematicians, physicists, engineers, and many others use the Riemann integral regularly.




The Integral: A Crux for Analysis

However, there exist examples of continuous function possessing a bounded derivative which is not Riemann integrable.

This example was given by the italian mathematician Vito Volterra in 1881.


The Integral: A Crux for Analysis

This example was surprising since it shows that within the context of Riemann's theory of integration the fundamental operations of differentiation and integration are not entirely reversible.

Thus, the process of differentiation could produce bounded functions $F'$ which fails to be integrable in Riemann's sense. In this case, the formula: \[ \int_{a}^{b}F'(x)~dx=F(b)-F(a) \] provided by the fundamental theorem of integral calculus, turns out to be meaningless.


The Integral: A Crux for Analysis

On the other hand, more advanced studies in analysis require an integral with better convergence properties. For example, if $f_n$ is a sequence of functions on a bounded interval $[a, b],$ then we would like to know that \[ \lim_{n\ra \infty} \int_a^b f_n = \int_a^b \lim_{n\ra \infty} f_n \qquad (*) \]

For the Riemann integral, the standard condition to guarantee (*) is uniform convergence of the sequence of functions $\{f_n \}$.



Garden of integrals

  1. The Riemann integral is based on subdividing the domain of the function (by way of a partition).
  2. The Lebesgue integral is based on subdividing the range of the function (you need measure theory).
  3. The Riemann-Stieltjes integral where the partition of a closed interval $I$ is determined with a monotonically increasing function defined on $I$.
  4. The Daniel integral which is equivalent to the Lebesgue integral but avoids the necessity of first developing measure theory.
  5. The Henstock-Kurzweil integral where a partition is measured in a more general way using a function called gauge.


For more details consult:

Credits