2. Compactification of C; Holomorphic functions and Cauchy–Riemann equations  PDF TEX

Riemann sphere

Compact sets

  • Let \((X, d)\) be a metric space and \(\Omega\subseteq X\). An open covering of \(\Omega\) is a family of open sets \((U_{\alpha})_{\alpha\in A}\) (\(A\) is not necessarily countable) such that \[\Omega \subseteq \bigcup_{\alpha\in A} U_{\alpha}.\]

    Definition.

    A set \(\Omega\) in a metric space is said to be compact if and only if every open covering of \(\Omega\) has a finite subcovering.

    Theorem.

    A set \(\Omega\) in a metric space is compact if and only if every sequence \((z_{n})_{n\in\mathbb N} \subseteq \Omega\) has a subsequence that converges to a point in \(\Omega\).

    Theorem.

    A set \(\Omega\subseteq \mathbb C\) is compact if and only if it is closed and bounded.

Compactification of \(\mathbb C\)

  • By adjoining a point \(\infty\), which we call the point at \(\infty\), to \(\mathbb{C}\) we obtain the extended complex plane \[\mathbb{C}_{\infty}=\mathbb{C} \cup\{\infty\} .\]

  • Next, we introduce the Riemann sphere \[\mathbb S=\left\{\left(x_{1}, x_{2}, x_{3}\right) \in \mathbb{R}^{3} : x_{1}^{2}+x_{2}^{2}+x_{3}^{2}=1\right\}\] and \[f: \mathbb S \rightarrow \mathbb{C}_{\infty}\] given by \[f\left(\left(x_{1}, x_{2}, x_{3}\right)\right)=\frac{x_{1}+i x_{2}}{1-x_{3}} \quad \text { if } \quad \left(x_{1}, x_{2}, x_{3}\right) \neq(0,0,1),\] and \[f((0,0,1))=\infty.\]

  • We write \(N=(0,0,1)\) and we call \(N\) the north pole of \(\mathbb S\). Further the function \(f\) is called stereographic projection.

Stereographic projection is one-to-one

  • First, we show that \(f\) is one-one. Let \((x_{1}, x_{2}, x_{3})\) and \((y_{1}, y_{2}, y_{3})\) be distinct elements of \(\mathbb S\) such that \(f\left(\left(x_{1}, x_{2}, x_{3}\right)\right)=f\left((y_{1}, y_{2}, y_{3})\right)\).

  • Then \((x_{1}, x_{2}, x_{3})\) and \((y_{1}, y_{2}, y_{3})\) are different from \(N\), and therefore \[\frac{x_{1}+i x_{2}}{1-x_{3}}=\frac{y_{1}+i y_{2}}{1-y_{3}}=z \in \mathbb{C}.\]

  • Thus \[|z|^{2}=\frac{x_{1}^{2}+x_{2}^{2}}{\left(1-x_{3}\right)^{2}}=\frac{y_{1}^2+y_{2}^2}{\left(1-y_{3}\right)^{2}} .\]

  • Since \(\left(x_{1}, x_{2}, x_{3}\right)\) and \((y_{1}, y_{2}, y_{3})\) are in \(\mathbb S\), we have \[|z|^{2}=\frac{1-x_{3}^{2}}{\left(1-x_{3}\right)^{2}}=\frac{1-y_{3}^{2}}{\left(1-y_{3}\right)^{2}}\quad \iff\quad |z|^{2}=\frac{1+x_{3}}{1-x_{3}}=\frac{1+y_{3}}{1-y_{3}}.\]

  • Therefore \[\frac{|z|^{2}-1}{|z|^{2}+1}=x_{3}, \quad \frac{|z|^{2}-1}{|z|^{2}+1}=y_{3}.\]

  • Thus \(x_{3}=y_{3}\) and hence \(x_{1}=y_{1},\) and \(x_{2}=y_{2}\).

Stereographic projection is onto

  • Next, we show that \(f\) is onto. If \(z=\infty\), we take \(N \in \mathbb S\) so that \(f(N)=\infty\).

  • Thus, we may suppose that \(z \in \mathbb{C}\). We take \[x_{1}=\frac{z+\overline{z}}{|z|^{2}+1}, \quad x_{2}=\frac{z-\overline{z}}{i\left(|z|^{2}+1\right)}, \quad x_{3}=\frac{|z|^{2}-1}{|z|^{2}+1}.\]

  • Then \[\begin{aligned} x_{1}^{2}+x_{2}^{2}+x_{3}^{2}&=\frac{(z+\overline{z})^{2}-(z-\overline{z})^{2}+\left(|z|^{2}-1\right)^{2}}{\left(|z|^{2}+1\right)^{2}}\\ &=\frac{4|z|^{2}+\left(|z|^{2}-1\right)^{2}}{\left(|z|^{2}+1\right)^{2}}=1 \end{aligned}\]

    implying \(\left(x_{1}, x_{2}, x_{3}\right) \in \mathbb S\). Further, we check that \(f\left(\left(x_{1}, x_{2}, x_{3}\right)\right)=z\).

Geometric interpretation of the stereographic projection

  • Let \(z=x+i y\) be a point in the complex plane and we identify it by \((x, y, 0)\) in \(\mathbb{R}^{3}\). We consider the line joining \(z\) to \(N\) in \(\mathbb{R}^{3}\).

  • Its parametric representation is \[t N+(1-t) z,\quad \text{ for } \quad -\infty<t<\infty .\]

  • The points on this line are \[((1-t) x,(1-t) y, t),\quad \text{ for } \quad -\infty<t<\infty .\]

  • Thus, the line intersects \(\mathbb S\) if and only if \[(1-t)^{2} x^{2}+(1-t)^{2} y^{2}+t^{2}=1,\] which we re-write as \[(1-t)|z|^{2}=(1-t) x^{2}+(1-t) y^{2}=1+t .\]

  • Hence \[t=\frac{|z|^{2}-1}{|z|^{2}+1} .\]

  • Thus, the line intersects \(\mathbb S\) at \(Z\) given by \[\begin{aligned} Z=&\Bigg(\left(1-\frac{|z|^{2}-1}{|z|^{2}+1}\right) x,\left(1-\frac{|z|^{2}-1}{|z|^{2}+1}\right) y, \frac{|z|^{2}-1}{|z|^{2}+1}\Bigg)\\ & =\left(\frac{2 x}{|z|^{2}+1}, \frac{2 y}{|z|^{2}+1}, \frac{|z|^{2}-1}{|z|^{2}+1}\right) \\ & =\left(\frac{z+\bar{z}}{|z|^{2}+1}, \frac{z-\bar{z}}{i\left(|z|^{2}+1\right)}, \frac{|z|^{2}-1}{|z|^{2}+1}\right) . \end{aligned}\]

  • We summarize the above procedure as follows:

    • Let \(z=(x, y)\) be a point in the complex plane.

    • The line passing through \((x, y, 0)\) and \((0,0,1)\) intersects \(\mathbb S\) exactly at one point \(Z=\left(x_{1}, x_{2}, x_{3}\right)\) given above and \(f\left(\left(x_{1}, x_{2}, x_{3}\right)\right)=z\).

    • We call \(\left(x_{1}, x_{2}, x_{3}\right)\) the spherical coordinates of \(z\) and \((0,0,1)\) the spherical coordinates of \(\infty\).

Distance between points in the extended complex plane

  • Let \(z \in \mathbb{C}_{\infty}\) and \(z^{\prime} \in \mathbb{C}_{\infty}\). Let \((x_{1}, x_{2}, x_{3})\) and \((x_{1}^{\prime}, x_{2}^{\prime}, x_{3}^{\prime})\) be spherical coordinates of \(z\) and \(z^{\prime}\), respectively. Namely, \[x_{1}=\frac{z+\overline{z}}{|z|^{2}+1}, \quad x_{2}=\frac{z-\overline{z}}{i\left(|z|^{2}+1\right)}, \quad x_{3}=\frac{|z|^{2}-1}{|z|^{2}+1},\] and \[x_{1}'=\frac{z'+\overline{z}'}{|z'|^{2}+1}, \quad x_{2}'=\frac{z'-\overline{z}'}{i\left(|z'|^{2}+1\right)}, \quad x_{3}'=\frac{|z'|^{2}-1}{|z'|^{2}+1}.\]

  • Then we define \[\begin{aligned} d\left(z, z^{\prime}\right)=\sqrt{\left(x_{1}-x_{1}^{\prime}\right)^{2}+\left(x_{2}-x_{2}^{\prime}\right)^{2} +\left(x_{3}-x_{3}^{\prime}\right)^{2}} . \end{aligned}\]

  • Thus

    \[\begin{aligned} d\left(z, z^{\prime}\right)^{2}&=\left(x_{1}-x_{1}^{\prime}\right)^{2}+\left(x_{2}-x_{2}^{\prime}\right)^{2}+\left(x_{3}-x_{3}^{\prime}\right)^{2}\\ &=2-2\left(x_{1} x_{1}^{\prime}+x_{2} x_{2}^{\prime}+x_{3} x_{3}^{\prime}\right). \end{aligned}\]

  • Now we note \[\begin{aligned} &x_{1} x_{1}^{\prime}+x_{2} x_{2}^{\prime}+x_{3} x_{3}^{\prime} \\ &=\frac{(z+\overline{z})\left(z^{\prime}+\overline{z}^{\prime}\right)}{\left(|z|^{2}+1\right)\left(\left|z^{\prime}\right|^{2}+1\right)}-\frac{(z-\overline{z})\left(z^{\prime}-\overline{z}^{\prime}\right)}{\left(|z|^{2}+1\right)\left(\left|z^{\prime}\right|^{2}+1\right)} +\frac{\left(|z|^{2}-1\right)\left(\left|z^{\prime}\right|^{2}-1\right)}{\left(|z|^{2}+1\right)\left(\left|z^{\prime}\right|^{2}+1\right)} \\ & =\frac{\left(|z|^{2}-1\right)\left(\left|z^{\prime}\right|^{2}-1\right)+2 z \overline{z}^{\prime}+2 \overline{z} z^{\prime}}{\left(|z|^{2}+1\right)\left(\left|z^{\prime}\right|^{2}+1\right)} \\ & =\frac{\left(|z|^{2}+1\right)\left(\left|z^{\prime}\right|^{2}+1\right)-2\left(|z|^{2}+1\right)-2\left(\left|z^{\prime}\right|^{2}+1\right)+4+2 z \overline{z}^{\prime}+2 \overline{z} z^{\prime}}{\left(|z|^{2}+1\right)\left(\left|z^{\prime}\right|^{2}+1\right)} \end{aligned}\]

  • Further \[\begin{aligned} &2\left(|z|^{2}+1\right)+2\left(\left|z^{\prime}\right|^{2}+1\right)-4-2 z \overline{z}^{\prime}-2 \overline{z} z^{\prime}\\ &=2\left(|z|^{2}+\left|z^{\prime}\right|^{2}-z \overline{z}^{\prime}-\overline{z} z^{\prime}\right)=2\left|z-z^{\prime}\right|^{2} . \end{aligned}\]

  • Hence \[2\left(x_{1} x_{1}^{\prime}+x_{2} x_{2}^{\prime}+x_{3} x_{3}^{\prime}\right)=2-\frac{4\left|z-z^{\prime}\right|^{2}}{\left(|z|^{2}+1\right)\left(\left|z^{\prime}\right|^{2}+1\right)}.\]

  • Since \(d\left(z, z^{\prime}\right)^{2}=2-2\left(x_{1} x_{1}^{\prime}+x_{2} x_{2}^{\prime}+x_{3} x_{3}^{\prime}\right)\), we obtain \[d\left(z, z^{\prime}\right)^2=\frac{4\left|z-z^{\prime}\right|^{2}}{\left(|z|^{2}+1\right)\left(\left|z^{\prime}\right|^{2}+1\right)}\]

  • Hence \[d\left(z, z^{\prime}\right)=\frac{2\left|z-z^{\prime}\right|}{\sqrt{\left(|z|^{2}+1\right)\left(\left|z^{\prime}\right|^{2}+1\right)}}\]

  • If \(z^{\prime}=\infty\). Then \(x_{1}^{\prime}=0, x_{2}^{\prime}=0\) and \(x_{3}^{\prime}=1\) and \[d(z, \infty)^{2}=2-2 x_{3}=2-2\left(\frac{|z|^{2}-1}{|z|^{2}+1}\right)=\frac{4}{|z|^{2}+1} .\]

  • Hence \[d(z, \infty)=\frac{2}{\sqrt{|z|^{2}+1}}.\]

  • Thus \(\mathbb{C}_{\infty}\) is a metric space.

  • The above metric \(d\) induces the following topology on \(\mathbb{C}_{\infty}\):

    • Let \(E \subseteq \mathbb{C}_{\infty}\).

    • If \(\infty \notin E\), then E is open in \(\mathbb{C}_{\infty}\) if and only if it is open in \(\mathbb{C}\).

    • If \(\infty \in E\), then \(E\) is open in \(\mathbb{C}_{\infty}\) if and only its complement in \(\mathbb{C}_{\infty}\) is a closed and bounded (i.e. compact) subset of \(\mathbb{C}\).

  • Finally, we check that \(f\) and \(f^{-1}\) are continuous to conclude the following result.

Theorem.

The extended complex plane \(\mathbb C_\infty\) is a metric space homeomorphic to the Riemann sphere \(\mathbb S\). In particular, \(\mathbb C_\infty\) is compact.

Complex functions

Functions on the complex plane

  • Let \(E\subseteq\mathbb C\) be a set of complex numbers. A function \(f\) defined on \(E\) is a rule that assigns to each \(z\) in \(E\) a unique complex number \(w\).

  • The number \(w\) is called the value of \(f\) at \(z\) and is denoted by \(f(z)\). The set \(E\) is called the domain of definition of \(f\).

  • Suppose \(w = u + iv\) is the value of a function \(f\) at \(z = x + iy\), meaning \[u + iv = f(x + iy).\]

  • The real numbers \(u\) and \(v\) depend on the real variables \(x\) and \(y\).

  • Consequently, \(f(z)\) can be expressed as a pair of real-valued functions of the real variables \(x\) and \(y\) as follows: \[f(z) = u(x,y) + iv(x,y).\]

  • The functions \(u\) and \(v\) are usually denoted by \(\operatorname{Re} f(z)\) and \(\operatorname{Im} f(z)\) respectively.

Examples

  • Consider the function \(f(z) = z^2\). Then \(f(x+iy) = x^2 - y^2 + i2xy\). Therefore, we have \(\operatorname{Re} f(z) = x^2 - y^2\) and \(\operatorname{Im} f(z) = 2xy\). The domain of \(f\) is the entire complex plane.

  • Let us consider \(f(z) = z^{1/n}\), where we assign to each complex number its \(n\)-th root. Then \(f\) is not a function in the sense of the definition above, since for each \(z \neq 0\), there are \(n\) roots of \(z\).

  • Let us express \(z\) as \(z = |z|(\cos\theta + i \sin\theta)\), where \(\theta \in (-\pi, \pi]\) is the argument of \(z\). Then the roots of \(z\) are given by \[z_k = \sqrt[n]{|z|} \left( \cos\frac{\theta + 2k\pi}{n} + i \sin\frac{\theta + 2k\pi}{n} \right), \quad k = 0, 1,\ldots, n-1.\]

  • Clearly, for each \(z \in \mathbb{C}\), we have \(n\) values of \(z^{1/n}\) determined by \(k\). In this case, we say that the \(n\)-th root has \(n\) branches. We can make \(z^{1/n}\) a single-valued function by choosing a specific \(k\).

  • A branch corresponding to \(k = 0\) is called the principal branch. Each branch is a single-valued function.

Continuity of complex functions

Definition.

Let \(E\subseteq \mathbb C\) and let \(f:E\to \mathbb C\) be a function. We say that \(f\) is continuous at the point \(z_{0} \in E\) if for every \(\varepsilon>0\) there exists \(\delta>0\) such that whenever \(z \in E\) and \(\left|z-z_{0}\right|<\delta\) then \[\left|f(z)-f\left(z_{0}\right)\right|<\varepsilon.\]

An equivalent definition is that for every sequence \((z_n)_{n\in\mathbb N} \subset E\) such that \(\lim_{n\to \infty} z_{n}=z_{0}\), then \[\lim_{n\to \infty} f\left(z_{n}\right)=f\left(z_{0}\right).\]

  • The function \(f\) is said to be continuous on \(E\) if it is continuous at every point of \(E\).

  • Sums and products of continuous functions are also continuous.

  • The function \(f\) is continuous if and only if its real and imaginary parts are continuous.

Properties of continuous functions

  • Since the convergence notions for the complex numbers and points in \(\mathbb{R}^2\) are the same, the function \(f\) of the complex argument \(z = x + i y\) is continuous if and only if it is continuous when viewed as a function of the two real variables \(x\) and \(y\).

  • By the triangle inequality, it is immediate that if \(f\) is continuous, then the real-valued function defined by \(z \mapsto |f(z)|\) is continuous.

  • We say that \(f\) attains a maximum at the point \(z_0 \in E\) if \[|f(z)| \leq \left| f(z_0) \right| \quad \text{for all } \quad z \in E,\] with the inequality reversed for the definition of a minimum.

    Theorem.

    A continuous complex function on a compact set \(E\) is bounded and attains a maximum and minimum on \(E\).

Examples

  • The function \(f(z) = z^2\) is continuous on \(\mathbb{C}\), which can be easily shown using the definition of continuity.

  • The principal branch of the square root function \(z^{1/2}\) is continuous on \(\mathbb{C} \setminus (-\infty, 0]\).

  • Recall that \[z_k = \sqrt{|z|} \left( \cos\frac{\theta + 2k\pi}{2} + i \sin\frac{\theta + 2k\pi}{2} \right), \quad k = 0, 1.\]

  • This exclusion is necessary because for \(z = x + iy\) where \(x \leq 0\) and \(y \to 0\) we may observe the following:

    • If \(y \to 0^+\), then \(\arg(z) \to \pi\), leading to \(z_0 \to i \sqrt{|x|}\).

    • If \(y \to 0^-\), then \(\arg(z) \to -\pi\), leading to \(z_0 \to -i \sqrt{|x|}\).

  • We see that the principal branch \(z^{1/2}\) exhibits discontinuity depending on the direction of approach along the imaginary axis.

Complex Differentiation

Complex Differentiation

Definition.

Suppose \(f\) is a complex function defined in open set \(\Omega\subseteq\mathbb C\). If \(z_{0} \in \Omega\) and \[\begin{aligned} \lim _{z \rightarrow z_{0}} \frac{f(z)-f\left(z_{0}\right)}{z-z_{0}} \end{aligned}\] exists, we denote this limit by \(f^{\prime}\left(z_{0}\right)\) and call it the derivative of \(f\) at \(z_{0}\).

  • If the derivative \(f^{\prime}\left(z_{0}\right)\) exists for every \(z_{0} \in \Omega\), we say that \(f\) is holomorphic (or analytic) in \(\Omega\).

  • The class of all holomorphic functions in \(\Omega\) will be denoted by \(H(\Omega)\)

  • To be more explicit, the derivative \(f^{\prime}\left(z_{0}\right)\) exists if to every \(\varepsilon>0\) there exists a \(\delta>0\) such that

    \[\left|\frac{f(z)-f\left(z_{0}\right)}{z-z_{0}}-f^{\prime}\left(z_{0}\right)\right|<\varepsilon \quad \text { for all } \quad z \in D^{\prime}\left(z_{0},\delta\right).\]

  • A function holomorphic in the whole \(\mathbb{C}\) is called an entire function.

Properties of holomorphic functions

  • If \(f \in H(\Omega)\) and \(g \in H(\Omega)\), then \(f+g \in H(\Omega)\) and \(fg \in H(\Omega)\), so that \(H(\Omega)\) is a ring.

  • Moreover, if \(f \in H(\Omega)\) and \(g \in H(\Omega)\) and additionally \(g\) is never \(0\) on \(\Omega\), then \(f/g \in H(\Omega)\).

  • The usual differentiation rules apply: for any \(f \in H(\Omega)\) and \(g \in H(\Omega)\) and \(\alpha, \beta\in\mathbb C\) we have \[\begin{aligned} (\alpha f+\beta g)'&=\alpha f'+\beta g',\\ (fg)'&=f'g+fg', \end{aligned}\]

  • If \(f \in H(\Omega)\) and \(g \in H(\Omega)\) and additionally \(g\) is never \(0\) on \(\Omega\), then \[\begin{aligned} \bigg(\frac{f}{g}\bigg)'=\frac{f'g-fg'}{g^2}. \end{aligned}\]

  • More interestingly, superpositions of holomorphic functions are also holomorphic.

Lemma.

If \(f \in H(\Omega)\), \(f(\Omega) \subseteq \Omega_{1}\), \(g \in H(\Omega_{1})\), and \(h = g \circ f\), then \(h \in H(\Omega)\), and \(h'\) can be computed using the chain rule: \[h'(z_{0}) = g'(f(z_{0})) \cdot f'(z_{0}), \quad \text{ for } \quad z_{0} \in \Omega.\]

Proof: To prove this, fix \(z_{0} \in \Omega\), and put \(w_{0}=f\left(z_{0}\right)\). Then \[\begin{gathered} f(z)-f\left(z_{0}\right)=\left[f^{\prime}\left(z_{0}\right)+\varepsilon(z)\right]\left(z-z_{0}\right),\\ g(w)-g\left(w_{0}\right)=\left[g^{\prime}\left(w_{0}\right)+\eta(w)\right]\left(w-w_{0}\right), \end{gathered}\] where \(\lim_{z \to z_{0}}\varepsilon(z) = 0\) and \(\lim_{w \to w_{0}}\eta(w) = 0\). Put \(w=f(z)\), and substitute it to the last equation. If \(z \neq z_{0}\), we obtain \[\frac{h(z)-h\left(z_{0}\right)}{z-z_{0}}=\left[g^{\prime}\left(f\left(z_{0}\right)\right)+\eta(f(z))\right]\left[f^{\prime}\left(z_{0}\right)+\varepsilon(z)\right].\] The differentiability of \(f\) forces \(f\) to be continuous at \(z_{0}\). Hence the conclusion follows from the last equation.$$\tag*{$\blacksquare$}$$

Theorem.

Let \(g\) be analytic on the open set \(\Omega_{1}\), and let \(f\) be a continuous complex-valued function on the open set \(\Omega\). Assume that

  • \(f(\Omega) \subseteq \Omega_{1}\),

  • \(g^{\prime}\) is never 0 ,

  • \(g(f(z))=z\) for all \(z \in \Omega\) (thus \(f\) is one-to-one).

Then \(f\) is analytic on \(\Omega\) and \(f^{\prime}=1 /\left(g^{\prime} \circ f\right)\).

Proof: Let \(z_{0} \in \Omega\), and let \((z_{n})_{n\in\mathbb N}\subseteq \Omega \backslash\left\{z_{0}\right\}\) with \(\lim_{n\to \infty}z_{n} =z_{0}\). Then \[\frac{f\left(z_{n}\right)-f\left(z_{0}\right)}{z_{n}-z_{0}}=\frac{f\left(z_{n}\right)-f\left(z_{0}\right)}{g\left(f\left(z_{n}\right)\right)-g\left(f\left(z_{0}\right)\right)}=\left[\frac{g\left(f\left(z_{n}\right)\right)-g\left(f\left(z_{0}\right)\right)}{f\left(z_{n}\right)-f\left(z_{0}\right)}\right]^{-1}\] (Note that \(f\left(z_{n}\right) \neq f\left(z_{0}\right)\) since \(f\) is 1-1 and \(z_{n} \neq z_{0}\).) By continuity of \(f\) at \(z_{0}\), the expression in brackets approaches \(g^{\prime}\left(f\left(z_{0}\right)\right)\) as \(n \rightarrow \infty\). Since \(g^{\prime}\left(f\left(z_{0}\right)\right) \neq 0\), the result follows. $$\tag*{$\blacksquare$}$$

Examples

  • The function \(f(z) = z\) is holomorphic on any open set in \(\mathbb{C}\), and \(f^{\prime}(z) = 1\). In fact, any polynomial \[p(z) = a_{0} + a_{1} z + \cdots + a_{n} z^{n}\] is holomorphic in the entire complex plane and \[p^{\prime}(z) = a_{1} + 2a_{2} z + \cdots + n a_{n} z^{n-1}.\]

  • The function \(\frac{1}{z}\) is holomorphic on any open set in \(\mathbb{C}\) that does not contain the origin, and \(f^{\prime}(z) = -\frac{1}{z^{2}}\).

  • The function \(f(z) = \overline{z}\) is not holomorphic. Indeed, we have \[\frac{f\left(z_{0} + h\right) - f\left(z_{0}\right)}{h} = \frac{\overline{h}}{h}\] which has no limit as \(h \rightarrow 0\), as one can see by first taking \(h\) real and then \(h\) purely imaginary.

Cauchy–Riemann equations

Real differentiability

Definition.

Suppose that \(E\subseteq \mathbb R^n\) is open \(F:E\to \mathbb R^m\), and \(x \in E\). If there exists a linear transformation \(A: \mathbb R^n\to \mathbb R^m\) such that \[\lim_{h\to 0}\frac{|F(x+h)-F(x)-Ah|}{|h|}=0, \tag{*}\] then we say that \(F\) is differentiable at \(x\), and we write \[A=F'(x).\] If \(F\) is differentiable at every \(x\in E\), we say that \(F\) is differentiable in \(E\).

  • The relation (*) can be rewritten in the form \[F(x+h)-F(x)= F'(x)h +r(h), \quad \text{ and } \quad \lim_{h\to 0}\frac{|r(h)|}{|h|}=0. \tag{**}\]

  • If \(F=(F_1,\ldots, F_m):E\to\mathbb R^m\) and \(F'(x_0)\) exists at \(x_0\in E\), then \[F'(x_0)= \begin{bmatrix} \frac{\partial F_1}{\partial x_1}(x_0) & \ldots & \frac{\partial F_1}{\partial x_n}(x_0)\\ \vdots & \ddots & \vdots\\ \frac{\partial F_m}{\partial x_1}(x_0) & \ldots & \frac{\partial F_m}{\partial x_n}(x_0) \end{bmatrix}.\]

  • We know that if all the partial derivatives \(\frac{\partial F_1}{\partial x_1},\ldots, \frac{\partial F_i}{\partial x_j},\ldots, \frac{\partial F_m}{\partial x_n}\) of \(F\) exist in a neighborhood of \(x_0\in E\) and they are continues at the point \(x_0\), then the function \(F\) is differentiable at \(x_0\).

  • The converse is not true. Consider the example \[F(x, y)= \begin{cases} \frac{y^3}{x^2+y^2} & \text{ if } (x, y)\neq(0,0),\\ 0& \text{ if } (x, y)=(0,0). \end{cases}\] The function \(F\) is not differentiable at the point \((0, 0)\), even though the partial derivatives \(\frac{\partial F}{\partial x}\) and \(\frac{\partial F}{\partial y}\) exist at this point. However, these partial derivatives are not continuous at the origin.

  • Recall that the function \(f(z) = \overline{z}\) is not holomorphic. Indeed, we have \[\frac{f\left(z_{0} + h\right) - f\left(z_{0}\right)}{h} = \frac{\overline{h}}{h},\] which has no limit as \(h \rightarrow 0\), as one can see by first taking \(h\) real and then \(h\) purely imaginary.

  • If the function \(f(z)=\overline{z}\) is considered as a real variable function then it corresponds to the map \(F(x, y) =(x,-y)\), which is indefinitely differentiable in the real sense.

  • This example illustrates that the existence of the real derivative need not guarantee that \(f\) is holomorphic.

  • This example leads us to associate more generally to each complex valued function \[f=u+i v,\] the mapping from \(F: E\to \mathbb{R}^{2}\), given by \[F(x, y)=(u(x, y), v(x, y)).\]

\(F(x, y)=(u(x, y), v(x, y))\)

  • If \(m=n=2\) and \((x_0, y_0)\in E\) is fixed and \(h=(h_1, h_2)\) and \(|h|\to0\), then the relation (**) can be rewritten in the form \[\begin{aligned} u(x_0+h_1, y_0+h_2)-u(x_0, y_0)=& \frac{\partial u}{\partial x}(x_0, y_0)h_1 +\frac{\partial u}{\partial y}(x_0, y_0)h_2\\ &+r_1(h_1, h_2), \end{aligned}\] and \[\begin{aligned} v(x_0+h_1, y_0+h_2)-v(x_0, y_0)=& \frac{\partial v}{\partial x}(x_0, y_0)h_1 +\frac{\partial v}{\partial y}(x_0, y_0)h_2\\ &+r_2(h_1, h_2), \end{aligned}\] where \(r(h_1, h_2)=(r_1(h_1, h_2), r_2(h_1, h_2))\) and \[\lim_{h\to 0}\frac{|r_1(h_1, h_2)|}{|h|}=0, \quad \text{ and } \quad \lim_{h\to 0}\frac{|r_2(h_1, h_2)|}{|h|}=0.\]

Cauchy–Riemann equations

Theorem.

Let \(\Omega\subseteq \mathbb C\) be open and \(f: \Omega\to \mathbb C\) be holomorphic. Then \[\frac{\partial f}{\partial x}=\frac{1}{i}\frac{\partial f}{\partial y}, \tag{A}\] where \(\partial/\partial x\) and \(\partial/\partial y\) denote the usual partial derivatives in the \(x\) and \(y\) variables respectively. If \(f=u+iv\) for some real valued functions \(u, v:\Omega\to \mathbb C\), then we have \[\frac{\partial u}{\partial x}=\frac{\partial v}{\partial y} \quad \text{ and } \quad \frac{\partial u}{\partial y}=-\frac{\partial v}{\partial x}. \tag{B}\] These relations are called the Cauchy–Riemann equations.

Proof: Fix \(z_0\in\Omega\). Since \(f: \Omega\to \mathbb C\) is holomorphic, we have \[\lim_{z\to z_0}\frac{f(z)-f(z_0)}{z-z_0}=f'(z_0). \tag{C}\]

  • To find relation (A), observe that \(z_0=(x_0, y_0)\) can be approached from different directions.

  • Firstly, with \(z\) of the form \((x_0+h)+iy_0\) in (C), observe that \[f'(z_0)=\lim_{h\to 0}\frac{f(x_0+h, y_0)-f(x_0, y_0)}{h}=\frac{\partial f}{\partial x}.\]

  • Secondly, with \(z\) of the form \(x_0+i(y_0+h)\) in (C), observe that \[f'(z_0)=\lim_{h\to 0}\frac{f(x_0, y_0+h)-f(x_0, y_0)}{ih}=\frac{1}{i}\frac{\partial f}{\partial y}.\]

  • Thus we obtain equation (A): \[\frac{\partial f}{\partial x}=\frac{1}{i}\frac{\partial f}{\partial y}.\]

  • If \(f=u+iv\) for some real valued functions \(u, v:\Omega\to \mathbb C\), then \[\frac{\partial u}{\partial x}+i\frac{\partial v}{\partial x}=-i\frac{\partial u}{\partial y}+\frac{\partial v}{\partial y}.\]

  • This implies equations in (B): \[\frac{\partial u}{\partial x}=\frac{\partial v}{\partial y} \quad \text{ and } \quad \frac{\partial u}{\partial y}=-\frac{\partial v}{\partial x}.\]

  • This completes the proof.$$\tag*{$\blacksquare$}$$

Cauchy–Riemann equations and holomorphicity

Theorem.

Suppose \(f=u+i v\) is a complex-valued function defined on an open set \(\Omega\). If \(u\) and \(v\) are differentiable in the real sense and satisfy the Cauchy–Riemann equations on \(\Omega\), then \(f\) is holomorphic on \(\Omega\).

Proof: Fix \(z_0=x_0+iy_0\in\Omega\) and let \(h=h_1+ih_2\). Recall that \[\begin{aligned} u(x_0+h_1, y_0+h_2)-u(x_0, y_0)= \frac{\partial u}{\partial x}(x_0, y_0)h_1 +\frac{\partial u}{\partial y}(x_0, y_0)h_2+r_1(h_1, h_2), \end{aligned}\] and \[\begin{aligned} v(x_0+h_1, y_0+h_2)-v(x_0, y_0)= \frac{\partial v}{\partial x}(x_0, y_0)h_1 +\frac{\partial v}{\partial y}(x_0, y_0)h_2+r_2(h_1, h_2), \end{aligned}\] where \[\lim_{h\to 0}\frac{|r_1(h_1, h_2)|}{|h|}=0, \quad \text{ and } \quad \lim_{h\to 0}\frac{|r_2(h_1, h_2)|}{|h|}=0.\]

  • Thus \[\begin{aligned} f(z_0+h)-f(z_0)=&\frac{\partial u}{\partial x}(x_0, y_0)h_1 +\frac{\partial u}{\partial y}(x_0, y_0)h_2\\ &+i\bigg( \frac{\partial v}{\partial x}(x_0, y_0)h_1 +\frac{\partial v}{\partial y}(x_0, y_0)h_2\bigg)+r(h), \end{aligned}\] where \(r(h)=r_1(h_1, h_2)+ir_2(h_1, h_2)\). Clearly \(\lim_{h\to 0}\frac{|r(h)|}{|h|}=0\).

  • By the Cauchy–Riemann equations \(\frac{\partial u}{\partial x}=\frac{\partial v}{\partial y}\), \(\frac{\partial u}{\partial y}=-\frac{\partial v}{\partial x}\), we find that \[f(z_0+h)-f(z_0)=\bigg(\frac{\partial u}{\partial x}(x_0, y_0)-i\frac{\partial u}{\partial y}(x_0, y_0)\bigg)h+r(h).\]

  • Then \[f'(z_0)=\frac{\partial u}{\partial x}(x_0, y_0)-i\frac{\partial u}{\partial y}(x_0, y_0). \quad \tag*{$\blacksquare$}\]

Example

  • Note that the hypothesis of real differentiability at the point \(z_0\) is essential and cannot be dispensed with.

  • For example, consider the function \[f ( x , y ) = \sqrt{| x y |},\] regarded as a complex function with imaginary part identically zero.

    • Then it has both partial derivatives at \(( x_0 , y_0 ) = ( 0 , 0 )\).

    • It moreover satisfies the Cauchy–Riemann equations at that point.

    • But it is not differentiable in the real sense, and so the first condition, that of real differentiability, is not met.

  • Therefore, this function is not complex differentiable.

Top