When the world says, ‘Give up,’ hope whispers, ‘Try it one more time,’ Lyndon B. Johnson

If a function f(z) = u(x, y) + iv(x, y) is complex-differentiable at a point z₀ =x₀ +iy₀, then its real and imaginary parts, u(x,y) and v(x,y), must satisfy the Cauchy-Riemann equations at that point: $\frac{\partial u}{\partial x}(x_0, y_0) = \frac{\partial v}{\partial y}(x_0, y_0), \frac{\partial u}{\partial y}(x_0, y_0) = -\frac{\partial v}{\partial x}(x_0, y_0)$.
Proposition. Let f(z) = u(x, y) + iv(x, y) for z = x + iy ∈ D where D ⊆ ℂ is an open set. Assume that u and v have continuous first partial derivatives throughout D and they satisfy the Cauchy-Riemann equations at a point z ∈ D. Then, f'(z) exists, i.e., f is differentiable at z.
Definition. A complex function f is said to be analytic (holomorphic) at a point z₀ if it satisfies any of the following equivalent conditions:
Here the limit must be the same regardless of the direction from which h approaches zero in the complex plane.
The complex exponential function, denoted as eᶻ or exp(z), is one of the most important functions in all of mathematics. It is the unique entire function that extends the real exponential function eˣ to the complex plane. This function is not only entire but also periodic.
To define the complex exponential function, we seek a function f: ℂ→ℂ that satisfies two key properties:
Let z = x + iy, where x, y ∈ ℝ. f(z) = f(x + iy) =[1] f(x)·f(iy) =[2] eˣ·f(iy)
Writing f(iy) = A(y) + iB(y), f(z) = eˣ·A(y) + ieˣ·B(y), where A(y) and B(y) are real-valued functions of y.
For f(z) to be a well-behaved function, we want it to be differentiable everywhere, which means it must satisfy the Cauchy-Riemann equations. The real and imaginary parts of f(z) are: u(x, y) = eˣ·A(y), v(x, y) = eˣ·B(y). The partial derivatives are uₓ = eˣ·A(y), uy = eˣ·A’(y), vₓ = eˣ·B(y), and vy = eˣ·B’(y).
Applying the Cauchy-Riemann equations, uₓ = vy and uy = - vₓ, we get: eˣ·A(y) = eˣ·B’(y), then, A(y) = B’(y). Similarly, eˣ·A’(y) = - eˣ·B(y), then A’(y) = -B(y).
From these two equations, we can derive a second-order differential equation for B(y): B’’(y) = -B(y) ↭ B’’(y) + B(y) = 0.
This linear, homogeneous, constant-coefficient ODE has characteristic equation r² + 1 = 0 ⇒ r = ±i. Complex Conjugate Roots: If the roots are complex conjugates a ± bi, the general solution is $y(t) = e^{at}(αcos(bt) + βsin(bt))$, where ‘α’ and ‘β’ are real numbers.
The general solution to our equation (a = 0, b = 1) is B(y) = αcos(y) + βsin(y) for some constants α, β ∈ ℝ. Using A(y) = B’(y), we find A(y) = -αsin(y) + βcos(y).
Now, we use the initial condition from the real exponential. Since f(0) = e⁰ = 1 =[f(z) = eˣ·A(y) + ieˣ·B(y)] e⁰A(0) + ie⁰B(0) ⇒ A(0) = 1, B(0) = 0.
From A(0) = 1 ⇒[-αsin(y) + βcos(y)] -αsin(0) + βcos(0) = 1 ⇒ β = 1.
From B(0) = 0 ⇒[B(y) = αcos(y) + βsin(y)] αcos(0) + βsin(0) = 0 ⇒ α = 0. Thus, we have uniquely determined the functions A(y) and B(y): B(y) = sin(y), A(y) = cos(y). Substituting these back into the expression for f(z), we arrive at the definition of the complex exponential function, f(z) = eˣ·cos(y) + ieˣ·sin(y) = eˣ(cos(y) + isin(y)).
To determine where this series converges, we use the ratio test. Let the power series be $\sum_{n=0}^∞ aₙ(x-c)ⁿ$. The ratio test involves computing: $L = \lim_{n \to ∞} |\frac{aₙ₊₁}{aₙ}|$. If L < 1, the series converges.
$L = \lim_{n \to ∞} |\frac{\frac{1}{(n+1)!}}{\frac{1}{n!}}| = \lim_{n \to ∞} |\frac{n!}{(n+1)!}| = \lim_{n \to ∞} \frac{1}{n+1} = 0 $ < 1, the series converges absolutely for any fixed z ∈ ℂ, the radius of convergence is R = 1/L = 1/0 = ∞. Because R = ∞, the series converges for all $z \in \mathbb{C}$ making eᶻ entire (analytic everywhere in the complex plane).
$e^{z_1}⋅e^{z_2} =[\text{By definition}] e^{x_1}(cos(y_1) + isin(y_1))⋅e^{x_2}(cos(y_2) + isin(y_2)) = e^{(x_1+x_2)}[(cos(y_1)cos(y_2)-sin(y_1)sin(y_2) + i(sin(y_1)cos(y_2)+cos(y_1)sin(y_2)))] =$
Using the angle addition formulas for cosine and sine: $ = e^{x_1+x_2}(cos(y_1+y_2)+isin(y_1+y_2)) = e^{(x_1+x_2)+i(y_1+y_2)} = e^{z_1+z_2}$
If we write α in polar form as $α = e^{i\theta}$ where R = |α| and θ = arg(α), then the solutions for eᶻ = α are given by: eˣ = R ⇒ x = ln(R) and y = θ + 2πk, k ∈ ℤ. Therefore, the solutions are ln(|α|) + i(arg(α) + 2πk), for any integer k ∈ ℤ.
Example: eᶻ = 2 + 2i, 2 + 2i = $2\sqrt{2}(\frac{1}{\sqrt{2}}+\frac{i}{\sqrt{2}}) = 2\sqrt{2}(cos(\frac{\pi}{4}+isin(\frac{\pi}{4}))) = 2\sqrt{2}(cos(2n\pi+\frac{\pi}{4})+isin(2n\pi+\frac{\pi}{4}))$ where n ∈ ℤ
take eˣ $=2\sqrt{2}, \text{i.e.,} x = ln(2\sqrt{2}), y = 2n\pi+\frac{\pi}{4}$
Then, eᶻ = $e^{ln(2\sqrt{2})+i(2n\pi+\frac{\pi}{4})}$ = 2 + 2i where n ∈ ℤ. Therefore, the equation eᶻ = 2 + 2i has infinitely many solutions.
If f is an analytic function with real and imaginary parts f(z) = eᶻ = u(x, y) + iv(x, y) = eˣcos(y) + ieˣsin(y), then $f'(z) = (eᶻ)' = \frac{\partial u}{\partial x} + i\frac{\partial v}{\partial x} = eˣcos(y) + ieˣsin(y) = eᶻ$
An alternative proof is as follows.
Theorem. If a function f(z) can be represented by a power series, f(z) = $\sum_{n=0}^∞a_n(z-z_0)^n$, that converges in a disk ∣z − z₀∣ < R (with R > 0), then f is analytic within that disk, and its derivative is found by differentiating the power series term by term: f’(z) = $\sum_{n=1}^∞na_n(z-z_0)^{n-1}$ (we are using the power rule for differentiation on each term, $\frac{d}{dz}a_n(z-z_0)^n = na_n(z-z_0)^{n-1}$ and the term for n = 0 is a constant, a₀, so its derivative is zero).
In our case, $e^z = \sum_{n=0}^∞ \frac{z^n}{n!}$ center at z₀ = 0 with $a_n = \frac{1}{n!}$ and R = = ∞. Thus, term-by-term differentiation is valid for all z ∈ ℂ.
Differentiate the series term by term: $\frac{d}{dz}e^z = \frac{d}{dz}(\sum_{n=0}^∞ \frac{z^n}{n!}) = \sum_{n=0}^∞ \frac{d}{dz}(\frac{z^n}{n!})$
For n = 0, $\frac{d}{dz}(\frac{z^0}{0!}) = \frac{d}{dz}(1) = 0$. For n ≥ 1, $\frac{d}{dz}(\frac{z^n}{n!}) = \frac{nz^{n-1}}{n!} = \frac{z^{n-1}}{(n-1)!}$. The series becomes: $\sum_{n=0}^∞ \frac{d}{dz}(\frac{z^n}{n!}) = \sum_{n=1}^∞ \frac{z^{n-1}}{(n-1)!} = [\text{Reindexing the Series, k = n-1}] \sum_{k=0}^∞ \frac{z^{k}}{k!} = e^z$ ∎
The derivative of eᶻ is identical to itself. Since eᶻ is entire, repeated differentiation yields: $\frac{d^2}{dz^2}eᶻ = eᶻ, \frac{d^3}{dz^3}eᶻ = eᶻ, \dots$ confirming eᶻ is “smooth” (C∞) and infinitely differentiable in ℂ, meaning having derivatives of all orders that behave nicely.
The given differential equation is a classic example of a separable differential equation. The equation can be rewritten as $\frac{df}{f} = dz$. Integrating both sides gives: $\int \frac{df}{f} = \int dz \leadsto ln|f| = z + C$
To solve for f, we exponentiate both sides: $|f| = e^{z + C} = e^{z}e^{C}$. Let’s rename the constant $e^{C}$ as A. Since $e^{C}$ is always positive, A can be any non-zero real number. The general solution is: f(z) = Aeᶻ.
Now, we use the initial condition f(0)=1 to find the value of A: f(0) = Ae⁰ = A·1 = A, hence A = 1. Plugging this value back into the general solution gives us the specific solution: f(z) = Aeᶻ = 1eᶻ = eᶻ. This is the unique solution to the initial value problem.
This formula establishes a profound connection between the exponential function and the trigonometric functions. It shows that the complex exponential function can be viewed as a generalization of the unit circle in the complex plane. As the real variable y increases, the point eⁱʸ traces out the unit circle ∣z∣=1 in a counterclockwise direction, starting from eⁱ⁰ = 1.
This provides a powerful and compact way to represent complex numbers in polar form. Any complex number z = x + iy can be written as z = r(cosθ + isinθ), where r = ∣z∣ and θ = arg(z). Using Euler’s formula, this become $z = re^{i\theta}$. This exponential form is extremely useful for performing operations like multiplication, division, and finding powers and roots of complex numbers, e.g., $z_1·z_2 = (r_1e^{i\theta_1})(r_2e^{i\theta_2}) = r_1r_2e^{i(\theta_1+\theta_1)}$ (to multiply two complex numbers, we multiply their moduli and add their arguments.)