JustToThePoint English Website Version
JustToThePoint en español

Zeros of Analytic Functions and the Identity Theorem

Mathematics is not about numbers, equations, computations, or algorithms: it is about understanding, William Paul Thurston

Topology and Limits

Introduction

Definition. Complex sequence A sequence of complex numbers is a function $a: \mathbb{N} \to \mathbb{C}$. We usually denote it by $(a_n)_{n \in \mathbb{N}}$ or simply $(a_n)$, where $a_n := a(n)$. The value $a_1$ is called the first term of the sequence, $a_2$ the second term, and in general $a_n$ the n-th term of the sequence.

Definition. Convergent complex sequence. A complex sequence $(a_n)_{n \in \mathbb{N}}$ is said to converge to a complex number $L \in \mathbb{C}$ if for every $\varepsilon > 0$ there exists an integer $N \in \mathbb{N}$ such that for all $n \geq N$ one has $|a_n - L| < \varepsilon$. In this case we write $\lim_{n \to \infty} a_n = L$ or $a_n \to L$ as $n \to \infty$, and L is called the limit of the sequence $(a_n)_{n \in \mathbb{N}}$.

Definition. Cauchy sequence. A complex sequence $(a_n)_{n \in \mathbb{N}}$ is called a Cauchy sequence if for every $\varepsilon > 0$ there exists an integer $N \in \mathbb{N}$ such that for all $n, m \geq N$ one has $|a_n - a_m| < \varepsilon$.

Definition. Series and partial sums.Let $(a_n)_{n \in \mathbb{N}}$ be a complex sequence. For each n $\in \mathbb{N}$, the finite sum $s_n := a_1 + a_2 + \cdots + a_n = \sum_{k=1}^n a_k$ is called the n-th partial sum of the (infinite) series $\sum_{k=1}^\infin a_k$ which we also denote simply by $\sum a_n$ when the index is clear from the context.

Definition. Convergent series. The series $\sum_{n=1}^{\infty} a_n$ is said to converge to the sum $s \in \mathbb{C}$ if the sequence of partial sums $(s_n)_{n \in \mathbb{N}}$ defined by $s_n = a_1 + a_2 + \cdots + a_n = \sum_{k=1}^n a_k$ converges to s, that is, $\lim_{n \to \infty} s_n = s$. In this case we write $s := \sum_{n=1}^\infin a_n$. If the sequence $(s_n)_{n \in \mathbb{N}}$ does not converge, we say that the series $\sum_{n=1}^{\infty} a_n$ diverges (or does not converge).

Definition. A complex power series centered at 0 in the variable z is a series of the form $a_0 + a_1z + a_2z^2 + \cdots = \sum_{n=0}^\infty a_n z^n$ with coefficients $a_i \in \mathbb{C}$

Definition. A complex power series centered at a complex number $a \in \mathbb{C} $ is an infinite series of the form: $\sum_{n=0}^\infty a_n (z - a)^n,$ where each $a_n \in \mathbb{C}$ is a coefficient, z is a complex variable, and $(z - a)^n$ is the nth power about the center.

Theorem. Given a power series $\sum_{n=0}^\infty a_n z^n$, there exists a unique value R, $0 \le R \le \infin$ (called the radius of convergence) such that:

  1. For any z with |z| < R (inside the circle), the series $\sum_{n=0}^\infty a_n z^n$ converges absolutely (this is a “green light” zone).
  2. For any z with |z| > R, the series diverges (this is a “red light” zone).

    On the Circle (|z| = R), this theorem gives no information. This is the yellow light zone —the series could converge or diverge.

Differentiability of Power Series. If $f(z) = \sum_{n=0}^{\infty} a_nz^n$ for |z| < R (R > 0), then f is analytic on B(0; R) and $f'(z) = \sum_{n=1}^{\infty} na_nz^{n-1}$ for |z| < R.

Weierstrass M-test. Let $\{u_k(z)\}_{k=0}^\infty$ be a sequence of complex-valued functions defined on a set $\gamma^* \subseteq \mathbb{C}$. If there exists a sequence of non-negative real numbers $\{M_k\}_{k=0}^\infty$ such that:

  1. Bounding Condition: $|u_k(z)| \leq M_k$ for all $z \in \gamma^*$ and all $k \in \mathbb{N}$
  2. Convergence of Bound: The series $\sum_{k=0}^\infty M_k$ converges

Then, the original series $\sum_{k=0}^\infty u_k(z)$ converges uniformly on $\gamma^*$.

Coefficients of power series. Let f(z) = $\sum_{k=0}^\infty c_kz^k$ where this power series has radius of convergence R > 0. Then,the n-th coefficient of a power series $c_n$ can be extracted using the integral formula, $c_n = \frac{1}{2\pi i} \int_{C_r} \frac{f(z)}{z^{n+1}}dz, 0 \le r \lt R, n \ge 0$ where $C_r$ is a circle of radius r centered at 0 and oriented positively.

Taylor’s Theorem. If f is analytic on an open disk B(a; R) (a disk of radius R centered at a), then f(z) can be represented exactly by a unique power series within that disk: $f(z) = \sum_{n=0}^{\infty}a_n (z - a)^n, \forall z \in B(a; R)$

This theorem bridges the gap between differentiability and power series. It guarantees that if a function behaves well (it is analytic) in a disk, it must also be infinitely differentiable and expressed or representable by a power series (an infinite polynomial) within that disk.

Furthermore, there exist unique constants $a_n = \frac{f^{(n)}(a)}{n!} = \frac{1}{2\pi i}\int_{C_r} \frac{f(w)}{(w-a)^{n+1}}dw$ where $C_r$ is a circle of radius r < R centered at a and oriented in the counterclockwise direction (positively oriented).

Zeros of Analytic Functions and the Identity Theorem

Local Identity Theorem. Suppose that f is analytic on a disk B(a; r), r > 0 and a is a zero of f, i.e., f(a) = 0. Then, exactly one of the following two scenarios must be true:

  1. The “Flat” Scenario. f is identically zero on the whole disk, $f(z) = 0, \forall z \in B(a; r)$.
  2. The “Lonely” Scenario. The zero of f at a is isolated, i.e., there exists an $\varepsilon \gt 0$ such that the punctured disk B’(a; r) (the disk minus the center) containing no zeroes of f, $f(z) \ne 0$ for all $0 < |z-a| < \varepsilon$.

Zeros are isolated. If f is analytic on a disk B(a; R), not identically zero, and f(a) = 0, then there exists a small neighborhood around a where f(z) is never zero(except obviously at a itself).

Corollary. Let f be holomorphic in a domain $D \subseteq \mathbb{C}$. If a is a limit point of the zero set $Z(f)=\{ z\in D:f(z)=0\}$, then $f \equiv 0 \text{ in some neighborhood } B(a;r) \subseteq D.$

Proof.

Suppose $a \in D$ is a limit point of Z(f). Then every epsilon neighborhood contains a zero of f, for every $n \in \mathbb{N}$, the ball B(a; 1/n) contains some $z_n \in Z(f)$ with $z_n \neq a$.

Therefore, the lonely scenario cannot happen which implies the flat scenario, f ≡ 0 in B(a; r), the function must collapse entirely — it is identically zero in a neighborhood.

Theorem. Identity Theorem General Form. Let G be a region (an open, connected set) and suppose that f is analytic on G. Assume that the set of zeros Z(f) = {$z \in G: f(z) = 0$} has a limit point in G. Then, f is identically zero on G (i.e., f(z) = 0 for all $z \in G$).

It states that an analytic function cannot be zero on a cluster of points without being zero everywhere.

Proof.

Let E be the set of limits points of the zero set Z(f) that lie within G. It contains the “cluster points” of zeroes. $E = \{ a \in G : a \text{ is a limit point of } Z(f) \}$

By the theorem’s assumption, we know $E \ne \emptyset$. Let $a \in E$. By definition of a limit point, there exists a sequence of zeros $a_n$ converging to a, $\lim_{n \to \infin}a_n = a$. Since f is continuous (because it is analytic): $f(a) = f(\lim_{n \to \infty} a_n) =[\text{f is continuous}] \lim_{n \to \infty} f(a_n) = \lim_{n \to \infty} 0 = 0$. Thus, $f(a) = 0$, so $a \in Z(f)$. Every limit point of zeros is itself a zero.

Claim: E is open (it spreads locally).

$\forall a \in E \subseteq Z(f), a \in Z(f) \leadsto f(a) = 0$. Since a (a ∈ E) is a limit point of zeros, a is not an isolated zero. By the Local Identity Theorem, if a zero is not isolated, the function must be identically zero in some small neighborhood $B(a; r) \subseteq G$. If f(z) = 0 for all z in this disk, then every point inside this disk is a limit point of zeros (Consider any point $z_0$ inside this small disk $B(a; r)$. All the “close neighbors” of $z_0$, specifically those inside the disk, are also zeros, we can find a sequence of distinct zeros within the disk approaching $z_0$. This makes $z_0$ a limit point of zeros). Therefore, the entire (open) disk $B(a; r)$ is contained in E, hence E is an open set.

Claim: E is closed (it contains its own boundaries).

To show E is closed, we will show that its complement, $G \setminus E$, is open. Pick any point $a \in G \setminus E$, then there are two possibilities for a:

  1. If $f(a) \ne 0$, then by continuity of f, there is a small neighborhood around a where f is never zero, $\exist r_1 \gt 0, f(z) \ne 0, \forall z \in B(a; r_1)$. If there are no zeros in this neighborhood, there are certainly no limit points of zeros. Thus, this neighborhood is in $G \setminus E$.
  2. If f(a) = 0 but a is not a limit point, then a is an isolated zero, there exists a neighborhood where a is the only zero, $\exists r_2 \gt 0, f(z) \ne 0, \forall z \in B'(a; r_2)$. Since there are no other zeros in this disk, no point in this disk can be a limit point of zeros. Thus, this neighborhood is in $G \setminus E$.

In both cases, a has a neighborhood outside E. Thus $G \setminus E$ is open, which implies E is closed.

Recall. A connected set G cannot be split into two disjoint, non-empty open sets. Equivalently, the only subsets of G that are both open and closed are the empty set $\emptyset$ and the whole set G.

Finally, since E is non-empty, open, and closed set inside a connected set G, it must be G itself, E = G. This means every point in G is a limit point of zeros. For an analytic function (which is continuous), this implies f(z) = 0 for all $z \in G$.

Proposition. The Arithmetic of Zeroes. Let f and g be analytic on a disk B(a; r). Suppose both have a zero at a with orders $n_f$ and $n_g$ respectively. Then, the product function $(f \cdot g)(z)$ has a zero at a of order $n_f + n_g$.

You could think of analytic functions as “infinite polynomials.” If f(z) behaves like $z^2$ (order 2) and g(z) behaves like $z^3$ (order 3), then $f(z) \cdot g(z)$ behaves like $z^2 \cdot z^3 = z^5$ (order 2 + 3 = 5). This simple exponent rule works for all analytic functions.

Proof.

Recall that if a function f has a zero of order n at a, we can factor it perfectly: $f(z) = (z-a)^n h(z)$ where h(z) is analytic and $h(a) \ne 0$.

Since f and g have zero of order $n_f$ and $n_g$ respectively, there exists analytic functions h(z) and k(z) such that $h(a) \ne 0, k(a) \ne 0$ and $f(z) = (z-a)^{n_f} h(z), g(z) = (z-a)^{n_g} k(z)$.

Now, construct the product $(f \cdot g)(z), (f \cdot g)(z) = \left[ (z-a)^{n_f} h(z) \right] \cdot \left[ (z-a)^{n_g} k(z) \right] =\text{[Group the terms by type]} (z-a)^{n_f} (z-a)^{n_g} \cdot [h(z) k(z)] = (z-a)^{n_f + n_g} \cdot [h(z) k(z)]$.

Let $H(z) = h(z)k(z)$.

  1. Is H analytic? Yes, the product of two analytic functions is analytic.
  2. Is $H(a) \ne 0$? Since $h(a) \ne 0$ and $k(a) \ne 0$, their product is non-zero.
  3. We have successfully written the product in the structural form: $(f \cdot g)(z) = (z-a)^{N} H(z)$ where $N = n_f + n_g$ and $H(a) \ne 0$. Therefore, the order is $n_f + n_g$.

Uniqueness theorem. Let G be a region (open and connected). Suppose that f and g are analytic functions on G. If f(z) = g(z) for all points in a set S and S has a limit point in G, then then f and g are identical everywhere on G ($f \equiv g$).

Proof.

Consider the difference function $d(z) = f(z) - g(z)$. $\forall z \in S, d(z) = 0$. Therefore, the set of zeros $Z(d)$ has a limit point in G. By the Identity Theorem, $d(z)$ must be identically zero. If f(z) - g(z) = 0, then f(z) = g(z).

Counting Zeros (The Logarithmic Derivative).

It shows how we can use an integral to “count” the order (multiplicity) of a zero.

Suppose f is analytic on a disk B(a; r) and has a zero at a of order m. We assume f is not identically zero, so a is an isolated zero.

Using the Structural Definition of a zero, we can write: $f(z) = (z-a)^mh(z), \forall z \in B(a; r)$ where h is analytic (and therefore continuous) and $h(a) \ne 0$. Because $h(a) \neq 0$ and h is continuous, there is a small neighborhood $B(a; \varepsilon)$ where h(z) is never zero.

f’(z) =[Use the Product Rule on $f(z) = (z-a)^m h(z)$] $m(z-a)^{m-1}h(z) + (z-a)^mh'(z), \forall z \in B(a; r)$

We want to analyze the fraction $\frac{f'(z)}{f(z)}$. This is often called the Logarithmic Derivative.

For $z \ne a, z \in B(a; \varepsilon), \frac{f'(z)}{f(z)} = \frac{m}{z-a}+\frac{h'(z)}{h(z)}$

Now, we integrate this ratio around a small circle $C_{\varepsilon_0}$ centered at a with radius $\varepsilon_0 < \varepsilon$, oriented counterclockwise.

$\int_{C_{\varepsilon_0}}\frac{f'(z)}{f(z)} dz = \int_{C_{\varepsilon_0}}\frac{m}{z-a}dz + \int_{C_{\varepsilon_0}} \frac{h'(z)}{h(z)} dz$

$\int_{C_{\varepsilon_0}} \frac{h'(z)}{h(z)} dz$ vanishes because $\frac{h'(z)}{h(z)}$ is an analytic function (h is analytic, $h(z) \neq 0$ in this small circle, therefore the quotient is analytic inside and on the circle) on the contour $C_{\varepsilon_0}$. By Cauchy’s Theorem (integral of an analytic function on a closed loop), this integral is zero.

$\int_{C_{\varepsilon_0}}\frac{f'(z)}{f(z)} dz = m \cdot (2\pi i) \leadsto \frac{1}{2\pi i} \int_{C_{\varepsilon_0}}\frac{f'(z)}{f(z)} dz = m$

Cauchy integral formula: $f(z_{0})=\frac{1}{2\pi i}\oint_{C}\frac{f(z)}{z-z_{0}}dz, \oint_{C} \frac{1}{z-a} dz = 2\pi i$

$m = \frac{1}{2\pi i} \oint_{C_{\varepsilon_0}} \frac{f'(z)}{f(z)} dz$ tells us that integrating $\frac{f'}{f}$ around a zero effectively “scans” the point and “detects” the multiplicity; so this formula returns an integer equal to the order of that zero.

Bitcoin donation

JustToThePoint Copyright © 2011 - 2025 Anawim. ALL RIGHTS RESERVED. Bilingual e-books, articles, and videos to help your child and your entire family succeed, develop a healthy lifestyle, and have a lot of fun. Social Issues, Join us.

This website uses cookies to improve your navigation experience.
By continuing, you are consenting to our use of cookies, in accordance with our Cookies Policy and Website Terms and Conditions of use.