17 Series of Functions


After looking at Sequences of Functions, it is natural to look at series of functions. The convergence of a series is equivalent to the convergence of its sequence of partial sums. This allows to immediately extend some definitions and theorems to series of functions. This material is from Section 4.6 in the textbook.

Definition: Let $D\subseteq \R$, and let $\{f_k\}_{k=1}^\infty$ be a sequence of functions defined on $D$. We say that the series of functions $\sum_{k=1}^\infty f_k$ converges pointwise to a function $f$ on the domain $D$ if the sequence of partial sums $\big\{\sum_{k=1}^n f_k\big\}_{n=1}^\infty$ converges pointwise to $f$ on $D$. We say that the series of functions $\sum_{k=1}^\infty f_k$ converges uniformly to a function $f$ on the domain $D$ if the sequence of partial sums $\big\{\sum_{k=1}^n f_k\big\}_{n=1}^\infty$ converges uniformly to $f$ on $D$.

Theorem: (a) If each function $f_k$ is continuous on the interval $I$, and the series $\sum_{k=1}^\infty f_k$ converges uniformly to $f$ on $I$, then $f$ is continuous on $I$.

(b) If each function $f_k$ is integrable on $[a,b]$, and the series $\sum_{k=1}^\infty f_k$ converges uniformly to $f$ on $[a,b]$, then $f$ is integrable on $[a,b]$, and $\int_a^b f = \sum_{k=1}^\infty \big(\int_a^b f_k\big)$.

(c) If each function $f_k$ is differentiable on $[a,b]$, the series $\sum_{k=1}^\infty f_k$ converges (pointwise) to $f$ on $[a,b]$, and the series $\sum_{k=1}^\infty f_k'$ converges uniformly to $g$ on $[a,b]$, then $f$ is differentiable and $f' = g$; that is, $\frac{d}{dx}\big(\sum_{k=1}^\infty f_k(x)\big) = \sum_{k=1}^\infty\big(\frac{d}{dx}f_k(x)\big)$ on $[a,b].$

We also get a new theorem that is useful in proving uniform convergence:

Theorem (Weierstrass M-test): Let $\{f_k\}$ be a sequence of functions defined on a domain $D\subseteq\R$. Let $\{M_k\}_{k=1}^\infty$ be a sequence of nonnegative numbers. Suppose that for all $k\in\N$ and all $x\in D$, $f_k(x)<M_k$. If $\sum_{k=1}^\infty$ converges, then $\sum_{k=1}^\infty f_k$ converges uniformly on $D$.


For us, the most important kind of series of functions is power series. A power series is like an infinite polynomial. That is, it is a sum of powers of $x$ or, more generally, powers of $(x-a)$, multiplied by constant coefficients.

Definition: Let $a\in\R$, and let $\{a_k\}_{k=0}^\infty$ be a sequence of numbers. Then $\sum_{k=0}^\infty a_k(x-a)^k$ is a power series centered at $a$.

Note that a power series is a sum of functions, where each function is a monomial $f_k(x) = a_k(x-a)^k$. Also note that power series most commonly start at $k=0$, rather than $k=1$, but in practice, of course, any starting point for the sum is allowed.

Since monomials are continuous, differentiable, and integrable, many of the hypotheses of the theorems about series of functions are fulfilled automatically for power series. The major remaining question is about uniform convergence, and the result here is remarkable: If a power series $\sum_{k=0}^\infty a_k(x-a)^k$ converges at any point $x_o\ne a$, then it converges (pointwise) on the entire open interval $|x-a|<r$, where $r=|x_o-a|$. Furthermore, for any $s$ satisfying $0<s<|x_o-a|$, it converges uniformly on the closed interval $|x-a|\le s$. Similarly, if the series diverges at some point $x_1$, then it diverges for any $z$ such that $|z-a|>|x_1-a|$. (This follows immediately from the previous result, since if the series were to converge at such a $z$, then it would also converge at $x_1$.)

This allows us to define a radius of convergence for any power series. Let $\sum_{k=0}^\infty a_k (x-a)^k$ be a power series centered at 0.

Given a power series $\sum_{n=0}^\infty a_n(x-a)^n$, we can differenetiate it term by term to get the derivative power series $\sum_{n=1}^\infty na_n(x-a)^{n-1}$. Again remarkably, this derivative series has the same radius of convergence as the original series. If the series converges pointwise to a function $f(x) = \sum_{n=0}^\infty a_n(x-a)^n$ for $|x-a|<R$, then by the theorems about series of functions, $f'(x)$ exists and $f'(x)=\sum_{n=1}^\infty na_n(x-a)^{n-1}$ for $|x-a|<R.$

We can also form the antiderivative series $\sum_{n=0}^\infty \frac{a_n}{n+1}(x-a)^{n+1}$. Again, this series has the same radius of convergence as the original series, and if $f(x) = \sum_{n=0}^\infty a_n(x-a)^n$ for $|x-a|<R$, then $\sum_{n=0}^\infty \frac{a_n}{n+1}(x-a)^{n+1}$ converges to an antiderivative of $f$ for $|x-a|<R$. In terms of indefinite integrals, we have that $\int_a^b \big(\sum_{n=0}^\infty a_n(x-a)^n\big)\,dx = C + \sum_{n=0}^\infty \frac{a_n}{n+1}(x-a)^{n+1}$ on the interval $|x-a|<R.$


If $\sum_{n=0}^\infty a_n(x-a)^n$ converges to $f(x)$ for $|x-a|<R$, where $R>0$, then $f(x)$ is infinitely differentiable at $a$, and $f^{(n)}(a) = n!\cdot a_n$. This means that $$f(x) = \sum_{n=0}^\infty \frac{f^{(n)}(a)}{n!}(x-a)^n$$ This is like an infinite Taylor polynomial for $f$, and it is called the Taylor Series for $f$ at $a.$

It is important to understand that if we start with a function $g$ that is infinitely differentiable at $a$, and form the series $\sum_{n=0}^\infty \frac{g^{(n)}(a)}{n!}(x-a)^n$, it is not necessarily true that this series converges to $g(x)$ for any $x$ other than $a.$ This is so even if the series has a non-zero radius of convergence. The series might converge to a completely different function than $g.$

Suppose that $f$ is a function on an open interval $I$, and for each $a\in I$, $f$ is given by a power series on some open interval containing $a$. (Note that in this case, that series must be the Taylor series for $f$ at $a$, and $f$ must be infinitely differentiable at $a$.) Then $f$ is said to be real analytic on the interval $I$. Being real analytic is a much stronger condition than simply being infinitely differentiable.


(back to contents)