Fact-checked by Grok 2 weeks ago

Contour integration

Contour integration is a fundamental technique in complex analysis for evaluating line integrals of analytic functions along oriented paths, or contours, in the complex plane, where a contour C is typically parametrized by z(t) for t in some interval, yielding \int_C f(z) \, dz = \int_a^b f(z(t)) z'(t) \, dt. It exploits the unique properties of holomorphic (analytic) functions, such as their satisfaction of the Cauchy-Riemann equations, to simplify computations that are intractable using real-variable methods alone. Developed primarily by Augustin-Louis Cauchy in the 1820s, this approach revolutionized integration theory by linking real and complex domains through theorems that relate contour integrals to residues at singularities. At its core, contour integration relies on Cauchy's theorem, which states that if a function f(z) is analytic inside and on a simple closed contour C bounding a simply connected region, then \int_C f(z) \, dz = 0. This allows contours to be deformed without altering the integral value, provided no singularities are crossed, enabling the enclosure of poles for evaluation via the residue theorem: \int_C f(z) \, dz = 2\pi i \sum \operatorname{Res}(f, z_k), where the sum is over residues at poles z_k inside C (for counterclockwise orientation). A key extension is Cauchy's integral formula, which expresses the value of an analytic function at a point z_0 inside C as f(z_0) = \frac{1}{2\pi i} \int_C \frac{f(z)}{z - z_0} \, dz, highlighting how contour integrals encode function values and derivatives. The power of contour integration shines in its applications to real definite integrals, often by closing a real-axis path with a semicircular arc in the complex plane and applying the residue theorem after ensuring the arc contribution vanishes (e.g., via Jordan's lemma for exponentials like e^{iz}). Notable examples include evaluating \int_{-\infty}^{\infty} \frac{dx}{x^2 + a^2} = \frac{\pi}{a} for a > 0 using a semicircle in the upper half-plane enclosing the pole at ia, or \int_0^{2\pi} \frac{d\theta}{1 + a \cos \theta} = \frac{2\pi}{\sqrt{1 - a^2}} for |a| < 1 via the unit circle substitution z = e^{i\theta}. These methods extend to Fourier transforms, branch-cut integrals like \int_0^{\infty} \frac{dx}{x^3 + 1} = \frac{2\pi}{3\sqrt{3}}, and physical problems in quantum mechanics and electromagnetism, where complex contours model wave propagation and potential fields.

Curves and Paths in the Complex Plane

Directed Smooth Curves

In complex analysis, a directed smooth curve, also known as a contour or path, is a continuous function \gamma: [a, b] \to \mathbb{C} that is piecewise continuously differentiable, meaning it is differentiable on a finite number of subintervals of [a, b] where the derivative \gamma' is continuous and non-zero on each subinterval, ensuring differentiability almost everywhere to avoid singularities in the path. This structure allows the curve to be composed of smooth segments joined at finitely many points, such as corners, while maintaining overall continuity, which is essential for defining paths suitable for integration without introducing discontinuities. The parametrization of such a curve is given by \gamma(t) = x(t) + i y(t), where x(t) and y(t) are real-valued functions that are piecewise continuously differentiable on [a, b], with the derivative \gamma'(t) = x'(t) + i y'(t) satisfying |\gamma'(t)| \neq 0 almost everywhere to ensure the curve does not have stationary points. The arc length element along the curve is then ds = |\gamma'(t)| \, dt, which measures the infinitesimal displacement in the complex plane and provides a natural way to quantify the curve's length as \int_a^b |\gamma'(t)| \, dt. The orientation of the directed smooth curve is determined by the direction of traversal as t increases from a to b, assigning a consistent "forward" direction to the path. The reversal of this orientation, denoted by -\gamma, is the curve traversed in the opposite direction, defined by -\gamma(t) = \gamma(a + b - t) for t \in [a, b], which effectively negates the direction while preserving the image of the curve in the complex plane. Closed contours represent special cases of these directed smooth curves where \gamma(a) = \gamma(b), but their topological properties are addressed separately.

Closed Contours and Jordan Curves

A closed contour in the complex plane is a directed smooth curve \gamma: [a, b] \to \mathbb{C} such that \gamma(a) = \gamma(b), forming a loop that returns to its starting point. These contours are constructed from directed smooth curves, where the endpoint of one segment coincides with the starting point of the next, ensuring continuity along the path. The orientation of a closed contour is typically positive when traversed counterclockwise, which aligns with the standard convention for enclosing regions in the plane. A simple closed contour is a special case where the curve does not intersect itself except at the initial and final point, maintaining a non-self-overlapping structure throughout its parametrization. This property ensures that the contour bounds a well-defined region without complications from crossings. The Jordan curve theorem states that any simple closed contour in the complex plane divides \mathbb{C} into two disjoint connected components: a bounded interior region and an unbounded exterior region, both open sets with the contour as their common boundary. For a positively oriented simple closed contour (counterclockwise traversal), the interior is the region enclosed by the curve, which is crucial for topological considerations in complex analysis, such as determining the domain for applying theorems like Cauchy's. For practical purposes in integration, closed contours are often required to be rectifiable, meaning they have finite total arc length given by L = \int_a^b |\gamma'(t)| \, dt < \infty, where \gamma' is continuous and the integral converges. This finite length condition guarantees that the contour is of bounded variation, allowing for the well-defined existence of integrals along the path without divergence issues. Rectifiability is essential for simple closed contours as well, ensuring the Jordan decomposition holds in a measurable sense.

Definition and Properties of Contour Integrals

Formal Definition for Continuous Functions

In complex analysis, the contour integral of a continuous function f along a directed smooth curve \gamma in the complex plane is formally defined using a parametrization of the curve. Suppose \gamma: [a, b] \to \mathbb{C} is a smooth parametrization of the curve with \gamma'(t) \neq 0 for t \in [a, b], and f is continuous on the image of \gamma. Then, the contour integral is given by \int_\gamma f(z) \, dz = \int_a^b f(\gamma(t)) \gamma'(t) \, dt, where the right-hand side is the standard Riemann integral over the real interval [a, b]. The existence of this integral is guaranteed when f is continuous on a compact curve \gamma, as continuity on a compact set ensures f is bounded and uniformly continuous, allowing the Riemann integral to converge. For piecewise smooth curves, the integral is defined by summing the integrals over each smooth segment. The contour integral satisfies linearity: for complex constants \alpha, \beta and continuous functions f, g on \gamma, \int_\gamma (\alpha f(z) + \beta g(z)) \, dz = \alpha \int_\gamma f(z) \, dz + \beta \int_\gamma g(z) \, dz. This follows directly from the linearity of the Riemann integral. Additional properties include additivity over concatenated curves: if \gamma = \gamma_1 + \gamma_2, where \gamma_1 and \gamma_2 share an endpoint, then \int_\gamma f(z) \, dz = \int_{\gamma_1} f(z) \, dz + \int_{\gamma_2} f(z) \, dz. The reversal property states that \int_{-\gamma} f(z) \, dz = -\int_\gamma f(z) \, dz, where -\gamma traverses the curve in the opposite direction. Furthermore, if |f(z)| \leq M on \gamma and the length of \gamma is L, then the ML-inequality provides the bound \left| \int_\gamma f(z) \, dz \right| \leq M L.

Generalization from Riemann Integrals

Contour integration extends the concept of Riemann integrals from the real line to curves in the complex plane, providing a framework for evaluating integrals of complex-valued functions along directed paths. In the real case, for a real-valued function f defined on a curve \gamma: [a, b] \to \mathbb{R}^2, the line integral with respect to arc length is given by \int_\gamma f \, ds = \int_a^b f(\gamma(t)) \|\gamma'(t)\| \, dt, which measures the accumulation of f weighted by the infinitesimal arc length ds = \|\gamma'(t)\| \, dt. This formulation relies on the scalar nature of real differentials and preserves orientation through the absolute value of the speed. In contrast, the complex contour integral \int_\gamma f(z) \, dz for a complex-valued function f along a parametrized curve \gamma: [a, b] \to \mathbb{C} is defined as \int_\gamma f(z) \, dz = \int_a^b f(\gamma(t)) \gamma'(t) \, dt, omitting the modulus and thus incorporating the direction and magnitude of the complex displacement dz = \gamma'(t) \, dt. This generalization treats dz as a complex differential dx + i \, dy, enabling the integral to capture vector-like behavior in the plane without the scalar restriction of arc length. Writing f(z) = u(x, y) + i v(x, y) where z = x + i y, the contour integral decomposes into real line integrals: \int_\gamma f(z) \, dz = \int_\gamma (u \, dx - v \, dy) + i \int_\gamma (v \, dx + u \, dy), linking it directly to the real and imaginary components of the path. This vector form highlights how contour integration unifies separate real integrals into a single complex expression, facilitating connections to theorems like Green's theorem in vector calculus. The inclusion of dz in the integrand motivates the study of contour integrals by allowing tests for holomorphicity: a continuous function f is holomorphic in a domain if \int_\gamma f(z) \, dz = 0 for every closed contour \gamma in that domain (Morera's theorem), a property that, together with Cauchy's theorem, distinguishes complex from real integration. This criterion underpins Cauchy's theorem and enables powerful analytic continuations. The foundations of contour integration were developed by Augustin-Louis Cauchy in the 1820s as part of his pioneering work on complex function theory, including studies on residues and definite integrals that extended real analysis to the complex plane.

Direct Computation of Contour Integrals

Parametrization Techniques

To compute a contour integral \int_\gamma f(z) \, dz directly, where f is a continuous complex function and \gamma is a directed smooth curve, one parametrizes the curve as \gamma: [a, b] \to \mathbb{C} with \gamma(a) and \gamma(b) as the endpoints, and then evaluates the real integral \int_a^b f(\gamma(t)) \gamma'(t) \, dt, assuming \gamma is differentiable with continuous derivative. This approach reduces the complex line integral to a standard Riemann integral over the real parameter interval, leveraging real analysis techniques for evaluation. Common parametrizations simplify this process for basic curves. For a straight line segment from z_1 to z_2, use \gamma(t) = z_1 + t(z_2 - z_1) for t \in [0, 1], yielding \gamma'(t) = z_2 - z_1. For the unit circle centered at the origin, traversed counterclockwise, parametrize as \gamma(t) = e^{it} for t \in [0, 2\pi], so \gamma'(t) = i e^{it}. These choices ensure the parametrization is smooth and oriented correctly, facilitating substitution into the integral formula. For more intricate contours composed of multiple segments, such as polygonal paths or combinations of lines and arcs, decompose the contour into piecewise smooth parts \gamma = \gamma_1 + \gamma_2 + \cdots + \gamma_n, where each \gamma_k is parametrized separately over its interval, and sum the resulting integrals \int_\gamma f(z) \, dz = \sum_k \int_{\gamma_k} f(z) \, dz, provided continuity holds at the junctions \gamma_k(b_k) = \gamma_{k+1}(a_{k+1}). This additive property allows handling of non-smooth but piecewise smooth curves common in applications. Direct parametrization is practical only for relatively simple functions f and curves \gamma, as the resulting real integral may require advanced techniques like substitution or partial fractions; for complex topologies or non-elementary f, the method becomes computationally tedious and inefficient.

Example: Integral Over a Straight Line and Circle

To illustrate direct computation of contour integrals via parametrization, consider the integral of the function f(z) = z along the straight line segment from z = 0 to z = 1 + i. Parametrize the path \gamma as \gamma(t) = t + i t for t \in [0, 1], so \gamma'(t) = 1 + i. The integral becomes \int_{\gamma} z \, dz = \int_0^1 (t + i t)(1 + i) \, dt = (1 + i)^2 \int_0^1 t \, dt = 2i \cdot \frac{1}{2} = i. \tag{1} This result matches the antiderivative evaluation \frac{z^2}{2} \big|_0^{1+i} = i, confirming the computation. Next, evaluate the integral of f(z) = z^2 over the unit circle |z| = 1, traversed counterclockwise. Parametrize the contour as \gamma(t) = e^{i t} for t \in [0, 2\pi], so \gamma'(t) = i e^{i t}. The integral is \int_{|z|=1} z^2 \, dz = \int_0^{2\pi} (e^{i t})^2 \cdot i e^{i t} \, dt = i \int_0^{2\pi} e^{i 3 t} \, dt = i \left[ \frac{e^{i 3 t}}{i 3} \right]_0^{2\pi} = \frac{1}{3} (e^{i 6 \pi} - 1) = 0, \tag{2} since e^{i 6 \pi} = 1. This zero result holds because z^2 is entire (analytic everywhere), and the integral of an analytic function over a closed contour vanishes, as will be formalized later in Cauchy's theorem. For numerical approximation after parametrization, the midpoint rule can be applied to the resulting real integral, treating the complex-valued integrand componentwise. For the straight-line example in (1), divide [0, 1] into n subintervals of width h = 1/n, with midpoints t_k = (k - 1/2) h for k = 1, \dots, n. The approximation is \int_0^1 (t + i t)(1 + i) \, dt \approx h \sum_{k=1}^n [(t_k + i t_k)(1 + i)], which converges to i as n \to \infty. This method extends the standard midpoint rule for real integrals to the parametrized form.

Fundamental Theorems in Contour Integration

Cauchy's Theorem

Cauchy's theorem is a foundational result in complex analysis that asserts the vanishing of contour integrals of holomorphic functions over closed paths. Specifically, if f is holomorphic in a simply connected domain \Omega and \gamma is a simple closed contour in \Omega such that the interior of \gamma is also contained in \Omega, then \int_\gamma f(z) \, dz = 0. This holds under the condition that f is analytic throughout the domain enclosing \gamma and its interior, ensuring no singularities disrupt the integration. The proof relies on Goursat's strengthened version of the theorem, which eliminates the need for continuous differentiability on the boundary. Goursat's theorem states that if f is complex differentiable at every point in a triangular domain U with boundary T \subset U, then \int_T f(z) \, dz = 0. The proof for triangles involves partitioning the domain into smaller sub-triangles and using estimates based on the differentiability condition f(z) = f(z_0) + f'(z_0)(z - z_0) + o(|z - z_0|), showing the integral over the boundary tends to zero as the partition refines. Extending this to simply connected domains, a primitive F(z) = \int_{z_0}^z f(\zeta) \, d\zeta exists, satisfying F'(z) = f(z). For a closed contour \gamma, the integral becomes F(\text{end}) - F(\text{start}) = 0 since the path returns to the starting point. An important extension is Morera's theorem, which serves as a converse under milder assumptions. If f: \Omega \to \mathbb{C} is continuous on a region \Omega and \int_\gamma f(z) \, dz = 0 for every simple closed rectifiable curve \gamma in \Omega, then f is holomorphic on \Omega. The proof constructs a primitive via line integrals from a fixed point and verifies differentiability, leveraging the zero-integral condition to confirm analyticity. This theorem is particularly useful for proving holomorphicity of functions defined by integrals or series.

Cauchy Integral Formula and Derivatives

The Cauchy integral formula expresses the value of a holomorphic function at an interior point of a contour in terms of an integral over the contour itself. Specifically, if f is holomorphic on a domain containing a simple closed positively oriented contour \gamma and its interior, and a is a point inside \gamma, then f(a) = \frac{1}{2\pi i} \oint_\gamma \frac{f(z)}{z - a} \, dz. This formula, a direct consequence of Cauchy's theorem, allows evaluation of holomorphic functions using boundary integrals. To derive the formula, consider the function g(z) = \frac{f(z)}{z - a} for z \neq a. This function is holomorphic in the region inside \gamma except at a, where it has an isolated singularity. Let \epsilon > 0 be small enough that the circle |\zeta - a| = \epsilon lies inside \gamma, and apply Cauchy's theorem to the annular region between \gamma and this small circle, yielding \oint_\gamma g(z) \, dz = \oint_{|\zeta - a| = \epsilon} g(\zeta) \, d\zeta. As \epsilon \to 0, the integral over the small circle approaches f(a) \cdot 2\pi i, since g(\zeta) \approx \frac{f(a)}{\zeta - a} near a, and thus \oint_\gamma \frac{f(z)}{z - a} \, dz = 2\pi i f(a). The formula extends to higher derivatives by differentiating under the integral sign, justified by the uniform convergence of the differentiated integrand on \gamma. For the first derivative, f'(a) = \frac{1}{2\pi i} \oint_\gamma \frac{f(z)}{(z - a)^2} \, dz. In general, for the n-th derivative where n \geq 0, f^{(n)}(a) = \frac{n!}{2\pi i} \oint_\gamma \frac{f(z)}{(z - a)^{n+1}} \, dz. This generalization follows from repeated differentiation, preserving holomorphy inside \gamma. Cauchy estimates provide bounds on these derivatives, quantifying how rapidly holomorphic functions can grow. If |f(z)| \leq M on \gamma and r is the distance from a to \gamma, then |f^{(n)}(a)| \leq \frac{n! \, M}{r^n}. This inequality, derived via the ML-inequality on the contour integral for f^{(n)}(a), implies that derivatives are controlled by the maximum modulus on the boundary and the distance to it.

Residues and the Residue Theorem

Computing Residues

In complex analysis, the residue of a function f(z) at an isolated singularity z_0 is defined as the coefficient a_{-1} in its Laurent series expansion \sum_{n=-\infty}^{\infty} a_n (z - z_0)^n around z_0, or equivalently, \operatorname{Res}(f, z_0) = \frac{1}{2\pi i} \oint_{\gamma} f(z) \, dz, where \gamma is any simple closed positively oriented contour encircling z_0 once and lying in a region where f is analytic except at z_0. This integral representation isolates the contribution from the (z - z_0)^{-1} term in the series. For a simple pole at z_0 (a pole of order 1), the residue can be computed directly as \operatorname{Res}(f, z_0) = \lim_{z \to z_0} (z - z_0) f(z), assuming the limit exists and is finite. This formula arises from the Laurent series, where the principal part is a_{-1}/(z - z_0), and multiplying by (z - z_0) yields a_{-1}. For a function f(z) = g(z)/(z - z_0) with g analytic and g(z_0) \neq 0, this simplifies to g(z_0), which aligns with the Cauchy integral formula as a special case where the residue of f(z)/(z - z_0) at z_0 equals f(z_0). For a pole of order m > 1 at z_0, the residue is the coefficient a_{-1} from the Laurent series, given by \operatorname{Res}(f, z_0) = \frac{1}{(m-1)!} \lim_{z \to z_0} \frac{d^{m-1}}{dz^{m-1}} \left[ (z - z_0)^m f(z) \right]. Equivalently, if f(z) = g(z)/(z - z_0)^m with g analytic and g(z_0) \neq 0, then \operatorname{Res}(f, z_0) = \frac{g^{(m-1)}(z_0)}{(m-1)!}, obtained by expanding g(z) in its Taylor series around z_0 and identifying the term that produces the (z - z_0)^{-1} contribution. Essential singularities require extracting a_{-1} from the full Laurent series, as no finite-order formula applies due to the infinite principal part. For example, the function f(z) = e^{1/z} has an essential singularity at z = 0, with Laurent series \sum_{n=0}^{\infty} \frac{1}{n!} z^{-n}; the coefficient of z^{-1} is \frac{1}{1!} = 1, so \operatorname{Res}(f, 0) = 1. The residue at infinity, \operatorname{Res}(f, \infty), for a function analytic at infinity (except possibly at finite singularities), is defined as \operatorname{Res}(f, \infty) = -\frac{1}{2\pi i} \oint_{\gamma} f(z) \, dz, where \gamma is a large positively oriented circle enclosing all finite singularities. It can be computed via the substitution w = 1/z, yielding \operatorname{Res}(f, \infty) = -\operatorname{Res}\left( \frac{f(1/w)}{w^2}, 0 \right), which transforms the expansion at infinity into a Laurent series at w = 0.

Statement and Application of the Residue Theorem

The residue theorem, also known as Cauchy's residue theorem, provides a fundamental method for evaluating contour integrals in the complex plane. Suppose f(z) is analytic in a simply connected domain D except for a finite number of isolated singularities inside a simple closed positively oriented contour \gamma that lies in D and does not pass through any singularities. Then, \oint_\gamma f(z) \, dz = 2\pi i \sum_k \operatorname{Res}(f, z_k), where the sum is taken over all singularities z_k enclosed by \gamma, and \operatorname{Res}(f, z_k) denotes the residue of f at z_k. To apply the residue theorem for computing a contour integral, first identify all isolated singularities of f(z) lying inside the contour \gamma, assuming \gamma is traversed counterclockwise. Next, compute the residue at each such singularity using techniques such as Laurent series expansions or specific formulas for simple or higher-order poles. The integral is then obtained by summing these residues and multiplying the total by $2\pi i. This approach simplifies evaluation compared to direct parametrization, especially for functions with known singularities. When a singularity lies on the contour \gamma, the standard residue theorem does not apply directly, as the function is not analytic on \gamma. In such cases, modify the contour by introducing a small semicircular indentation around the pole, typically of radius \epsilon \to 0, to exclude it from the interior. The integral over the indented contour equals $2\pi i times the sum of residues inside, and the contribution from the semicircular indentation approaches \pm \pi i times the residue at the pole, depending on the orientation (counterclockwise yields +\pi i \operatorname{Res}; clockwise yields -\pi i \operatorname{Res}). The original integral is recovered as the Cauchy principal value along the unindented path plus the half-residue contribution from the indentation. Beyond integral evaluation, the residue theorem finds applications in stability analysis, particularly in control theory, where the locations and signs of residues of characteristic functions help determine system stability by assessing pole placements in the complex plane. For instance, residues are used to quantify sensitivities and participations in oscillatory stability studies of dynamic systems.

Applications to Evaluate Real Integrals

Trigonometric Integrals via Semicircular Contours

One common application of contour integration involves evaluating real integrals of the form \int_{-\infty}^{\infty} f(x) \cos(kx) \, dx or \int_{-\infty}^{\infty} f(x) \sin(kx) \, dx, where f(x) is a rational function that decays sufficiently fast at infinity and k > 0. These trigonometric integrals can be addressed by considering the complex function g(z) = f(z) e^{i k z} and integrating over a semicircular contour in the upper half-plane, consisting of the real axis from -R to R and a semicircular arc \Gamma_R of radius R. As R \to \infty, the integral over the real axis approaches the desired integral (or its imaginary/real part), provided the contribution from \Gamma_R vanishes. This approach leverages the residue theorem, where the contour integral equals $2\pi i times the sum of residues of g(z) at poles in the upper half-plane. A key tool for ensuring the arc integral over \Gamma_R tends to zero is Jordan's lemma. For k > 0, on the semicircle z = R e^{i\theta} with $0 \leq \theta \leq \pi, the term e^{i k z} = e^{i k R (\cos\theta + i \sin\theta)} = e^{i k R \cos\theta} e^{-k R \sin\theta} decays exponentially for \theta \in (0, \pi) since \sin\theta > 0. If |f(z)| \leq M / R for large R on \Gamma_R, the integral's magnitude is bounded by \pi M e^{-k R \delta} for some \delta > 0, which approaches zero as R \to \infty. Thus, \int_{-\infty}^{\infty} g(x) \, dx = 2\pi i \sum \operatorname{Res}(g(z)) over upper half-plane poles, and the original trigonometric integral is the real or imaginary part accordingly. The procedure typically involves: (1) expressing the trigonometric function via Euler's formula, e.g., \cos(kx) = \operatorname{Re}(e^{i k x}); (2) selecting the upper half-plane contour for k > 0 to exploit decay; (3) identifying simple poles of f(z) in \operatorname{Im}(z) > 0 and computing residues of g(z), often \operatorname{Res}(g, z_0) = f(z_0) e^{i k z_0} / f'(z_0) for simple poles; (4) applying the residue theorem and taking the limit R \to \infty; (5) extracting the real or imaginary part. For instance, consider \int_{-\infty}^{\infty} \frac{\cos x}{x^2 + 1} \, dx. Here, f(z) = 1/(z^2 + 1), poles at z = \pm i, with z = i in the upper half-plane. The residue at z = i is \frac{e^{i \cdot i}}{2i} = \frac{e^{-1}}{2i}, so the contour integral is $2\pi i \cdot \frac{e^{-1}}{2i} = \pi e^{-1}. Since the imaginary part (sine) integrates to zero by oddness, the cosine integral equals \pi e^{-1}. Semicircular contours are particularly suited for non-periodic trigonometric integrals over infinite domains, but contour methods also extend to periodic trigonometric integrals like \int_0^{2\pi} R(\sin \theta, \cos \theta) \, d\theta, where R is rational, using the unit circle contour instead. For example, to evaluate \int_0^{2\pi} \frac{d\theta}{a + b \cos \theta} with a > b > 0, substitute z = e^{i\theta}, so d\theta = dz / (i z) and \cos \theta = (z + 1/z)/2, transforming the integral to \oint_{|z|=1} \frac{2 dz}{i z (2a + b (z + 1/z))} = \frac{2}{b i} \oint \frac{dz}{z^2 + (2a/b) z + 1}. The poles are roots of z^2 + (2a/b) z + 1 = 0, z = \frac{-a \pm \sqrt{a^2 - b^2}}{b}, with the root inside the unit circle being the one with smaller magnitude. The residue at that pole yields the integral value \frac{2\pi}{\sqrt{a^2 - b^2}} via the residue theorem.

Integrals Involving Branch Cuts

In contour integration, multivalued functions such as the complex logarithm \log z require careful handling due to their branch points, typically at z = 0 and z = \infty, where the function fails to be single-valued. To define a single-valued branch, a branch cut is introduced, often along the positive real axis, rendering the function analytic in the cut plane \mathbb{C} \setminus [0, \infty). The argument of z is then restricted to (0, 2\pi), ensuring \log z = \ln |z| + i \arg z is holomorphic in this domain./01%3A_Complex_Algebra_and_the_Complex_Plane/1.11%3A_The_Function_log(z)) Integrals involving such functions, like those with fractional powers z^{a-1} for non-integer a, necessitate contours that respect the branch cut, either avoiding it or encircling it to capture the discontinuity. A common approach is the keyhole contour, which consists of a large circle of radius R, a small circle of radius \epsilon around the origin, and two nearly overlapping line segments parallel to the positive real axis, just above and below the cut, forming a narrow "keyhole" shape that avoids crossing the branch point at zero. As R \to \infty and \epsilon \to 0, the integrals over the circular arcs vanish under suitable convergence conditions, leaving the contributions from the line segments related by the phase shift across the cut. The residue theorem applies to the enclosed poles, provided they lie away from the cut. A representative example is the evaluation of \int_0^\infty \frac{x^{a-1}}{1 + x} \, dx for $0 < a < 1. Consider f(z) = \frac{z^{a-1}}{1 + z} with the branch cut along [0, \infty) and \arg z \in (0, 2\pi), so z^{a-1} = \exp((a-1) \log z). The keyhole contour \Gamma encloses the simple pole at z = -1, where the residue is \operatorname{Res}_{z=-1} f(z) = (-1)^{a-1} = e^{i\pi(a-1)} = -e^{i\pi a}. Thus, \oint_\Gamma f(z) \, dz = 2\pi i (-e^{i\pi a}). The upper line segment contributes I = \int_0^\infty \frac{x^{a-1}}{1 + x} \, dx, while the lower contributes -e^{2\pi i (a-1)} I = -e^{2\pi i a} I (noting the exponent is a-1). Combining yields (1 - e^{2\pi i a}) I = -2\pi i e^{i\pi a}, so I = \frac{\pi}{\sin(\pi a)}. This result, known as a special case of the beta function reflection formula, highlights how the branch cut induces the phase factor enabling real integral evaluation. Another specialized contour is the Hankel contour, used for integrals involving branch cuts along the negative real axis, such as in the representation of the Gamma function \Gamma(z). The contour begins at −∞ along the upper side of the cut (where \arg t = \pi), encircles the origin counterclockwise in a small circle of radius \epsilon, and returns to −∞ along the lower side (\arg t = -\pi), with the branch cut for t^{z-1} along (-\infty, 0]. The integral is \Gamma(z) = \frac{1}{2i \sin(\pi z)} \int_H e^t t^{z-1} \, dt, valid for \operatorname{Re}(z) > 0 and entire in z except at non-positive integers, where the sine factor cancels poles. As \epsilon \to 0, the small arc contribution vanishes, and for positive real z, the contour collapses to the standard Euler integral. This representation is particularly useful for analytic continuation and asymptotic analysis. In practice, contours involving branch cuts can often be deformed to simpler paths, provided the deformation avoids crossing singularities or the cut itself, preserving the integral value by Cauchy's theorem in regions of analyticity. For instance, the position of the branch cut may be shifted (e.g., from the positive to negative real axis) if the integrand remains analytic in the deformed region and encloses the same residues, simplifying computations while accounting for any induced phase changes. Such deformations are essential for handling complicated cut configurations in advanced applications./10%3A_Definite_Integrals_Using_the_Residue_Theorem/10.04%3A_Integrands_with_branch_cuts)

Special Examples and Advanced Techniques

Cauchy Distribution and Probability Densities

The Cauchy distribution is a continuous probability distribution with probability density function f(x) = \frac{1}{\pi (1 + x^2)}, \quad x \in \mathbb{R}. This form ensures it is non-negative and symmetric about zero. To confirm normalization, evaluate \int_{-\infty}^{\infty} f(x) \, dx = \frac{1}{\pi} \int_{-\infty}^{\infty} \frac{dx}{1 + x^2} = 1. The key integral I = \int_{-\infty}^{\infty} \frac{dx}{1 + x^2} is computed via contour integration in the complex plane by considering the function g(z) = \frac{1}{1 + z^2} over a semicircular contour \Gamma_R in the upper half-plane, consisting of the real interval [-R, R] and the semicircular arc C_R of radius R. The integrand g(z) has simple poles where $1 + z^2 = 0, at z = i and z = -i; only z = i lies inside \Gamma_R for large R > 1. The residue at z = i is computed as \operatorname{Res}(g, i) = \lim_{z \to i} (z - i) \frac{1}{(z - i)(z + i)} = \frac{1}{2i}. By the residue theorem, \oint_{\Gamma_R} g(z) \, dz = 2\pi i \cdot \frac{1}{2i} = \pi. As R \to \infty, the integral over C_R vanishes because |g(z)| \sim 1/|z|^2 on the arc, making the length times maximum value tend to zero. Thus, I = \pi, confirming \int_{-\infty}^{\infty} f(x) \, dx = 1. The moments of the Cauchy distribution are defined as \mu_n = \int_{-\infty}^{\infty} x^n f(x) \, dx for n \geq 1. Unlike the normalization case (n = 0), these integrals do not converge to finite values due to the heavy tails of the distribution, where f(x) \sim 1/(\pi x^2) as |x| \to \infty. Specifically, |\mu_n| \leq \int_{-\infty}^{\infty} |x|^n f(x) \, dx \sim \int_1^{\infty} x^{n-2} \, dx, which diverges for all n \geq 1. Contour integration illustrates this non-convergence: to evaluate \int_{-\infty}^{\infty} \frac{x^n}{1 + x^2} \, dx (up to the $1/\pi factor), consider h(z) = \frac{z^n}{1 + z^2} over the same upper-half-plane semicircle \Gamma_R. The poles remain at z = \pm i, off the real axis, so no indentation is needed. However, on the arc C_R, |h(z)| \sim |z|^{n-2}; the arc integral has magnitude bounded by \pi R \cdot R^{n-2} = \pi R^{n-1}, which tends to infinity as R \to \infty for n \geq 1, rather than vanishing. Thus, the residue theorem cannot be applied to extract the real-line integral, confirming the lack of finite moments. For odd n, the integrand is odd, so the Cauchy principal value exists and equals zero by symmetry, but absolute convergence fails, precluding a well-defined moment. A key application of contour integration to the Cauchy distribution is deriving its characteristic function \phi(t) = \mathbb{E}[e^{i t X}] = \int_{-\infty}^{\infty} e^{i t x} f(x) \, dx = \frac{1}{\pi} \int_{-\infty}^{\infty} \frac{e^{i t x}}{1 + x^2} \, dx, which equals e^{-|t|}. For t > 0, evaluate \int_{-\infty}^{\infty} \frac{e^{i t x}}{1 + x^2} \, dx using the upper-half-plane contour for k(z) = \frac{e^{i t z}}{1 + z^2}, as \operatorname{Im}(i t z) = -t \operatorname{Im}(z) < 0 for \operatorname{Im}(z) > 0, ensuring the arc integral vanishes by the Jordan lemma. The residue at z = i is \frac{e^{i t (i)}}{2i} = \frac{e^{-t}}{2i}, so the contour integral is $2\pi i \cdot \frac{e^{-t}}{2i} = \pi e^{-t}, yielding \phi(t) = e^{-t}. For t < 0, close in the lower half-plane (where the exponential decays for \operatorname{Im}(z) < 0), enclosing the pole at z = -i with residue \frac{e^{i t (-i)}}{-2i} = \frac{e^{t}}{-2i}. The contour is clockwise (negatively oriented), so the integral is -2\pi i \cdot \frac{e^{t}}{-2i} = \pi e^{t}. Since t < 0, e^{t} = e^{-|t|}, yielding \phi(t) = e^{-|t|} overall. This non-differentiable form at t = 0 further underscores the absence of moments.

Logarithmic Integrals and Residue at Infinity

Logarithmic integrals often arise in applications requiring the evaluation of real definite integrals from 0 to ∞ involving the natural logarithm, which introduces multivalued behavior in the complex plane. To handle this, contour integration employs a principal branch of the logarithm with a branch cut typically along the positive real axis, allowing the integral to be related to the residues of the complex function \frac{(\Log z)^k}{p(z)}, where k is a positive integer and p(z) is a rational function with appropriate poles. A representative example is the evaluation of \int_0^\infty \frac{(\log x)^2}{1 + x^2} \, dx. This can be computed using a semicircular contour in the upper half-plane, with the branch cut along the positive real axis and small indentations around the origin to avoid the branch point. The integrand \frac{(\Log z)^2}{1 + z^2} has poles at z = i in the upper half-plane and z = -i in the lower half-plane. The contour integral over this path yields $2\pi i times the residue at z = i, accounting for the contribution from the large arc vanishing as the radius tends to infinity due to the behavior of (\Log z)^2 / z^2. The real-axis contribution, adjusted for the branch, gives the desired integral as \frac{\pi^3}{8}. Alternatively, differentiation under the integral sign can be used: consider I(a) = \int_0^\infty \frac{x^a}{1 + x^2} \, dx, which evaluates to \frac{\pi}{2 \sin(\pi (a+1)/2)} for -1 < \Re(a) < 1 via contour methods, and then I''(0) = \int_0^\infty \frac{(\log x)^2}{1 + x^2} \, dx = \frac{\pi^3}{8}. This approach leverages the parameter differentiation to introduce the logarithmic factors while relying on residue calculus for the parameterized integral. For more involved cases like the square of the logarithm with indentations at both 0 and ∞, a keyhole contour modified with small circles around these points is employed. The function \frac{(\Log z)^2}{1 + z^2} is integrated over this contour, capturing residues at the poles z = \pm i and accounting for the branch jump across the cut. The contributions from the indentations at 0 and ∞ provide boundary terms, while the large and small arcs vanish appropriately, leading to the real integral via the residue sum. This method highlights the interplay between branch structure and pole locations for precise evaluation. The residue at infinity extends the residue theorem to the extended complex plane, defined as \operatorname{Res}_{z=\infty} f(z) = -\operatorname{Res}_{w=0} \left[ \frac{1}{w^2} f\left(\frac{1}{w}\right) \right], where w = 1/z. For a function f(z) analytic at infinity except possibly at finite poles, the sum of all residues, including at infinity, is zero. Consequently, for a large positively oriented contour \gamma enclosing all finite poles, \int_\gamma f(z) \, dz = -2\pi i \operatorname{Res}_{z=\infty} f(z). This is particularly useful for integrals over the entire real line when the arc contributions vanish, allowing evaluation via the single residue at infinity rather than summing multiple finite residues. In the context of principal value integrals involving logarithms, such as variants of \mathrm{P.V.} \int_{-\infty}^\infty \frac{\log |x|}{1 + x^2} \, dx, contours closing in the upper or lower half-plane are selected based on the imaginary part of the logarithm to ensure convergence. For the principal branch, closing in the upper half-plane captures the residue at z = i, while the lower half-plane uses z = -i, with the principal value arising from symmetric indentations around singularities on the real axis. The integral equals zero because \int_0^\infty \frac{\ln x}{1 + x^2} \, dx = 0, as the substitution x = 1/u yields I = -I, implying I = 0.

Multivariable and Parametric Extensions

Surface Integrals in Complex Variables

In several complex variables, the concept of contour integration generalizes to integrals of differential forms over surfaces embedded in ℂⁿ, where surfaces are typically real 2-dimensional submanifolds. For a 2-form ω defined on an open set in ℂ², the surface integral ∫_S ω over a parametrized oriented surface S ⊂ ℂ², with parametrization φ: U → S where U ⊂ ℝ² is an open set, is computed as ∫_U φ*ω, the integral of the pullback form over U with the induced orientation. This construction preserves the complex structure and allows for the analysis of holomorphic properties along the surface. A key relation is provided by the generalized Stokes' theorem, which states that for a compact oriented surface S with boundary ∂S and a smooth 1-form α on a neighborhood of S, ∫S dα = ∫{∂S} α, where d denotes the exterior derivative. In the complex setting, this theorem applies to forms of various types, including (2,0)-forms like those involving dz and dw, and facilitates the evaluation of integrals by reducing them to lower-dimensional boundary contributions. The theorem holds in the broader framework of oriented manifolds and is essential for deriving integral representations in higher dimensions. Consider the example of integrating the basic holomorphic (2,0)-form ω = dz ∧ dw over the distinguished boundary surface of the unit polydisk D = {(z, w) ∈ ℂ² : |z| < 1, |w| < 1}, which is the 2-dimensional torus Γ = { (z, w) : |z| = 1, |w| = 1 }. Parametrizing Γ by θ, ϕ ∈ [0, 2π) with z = e^{iθ}, w = e^{iϕ}, the pullback yields dz ∧ dw = -e^{i(θ + ϕ)} dθ ∧ dϕ, and the integral evaluates to zero due to the vanishing of the Fourier coefficients over full periods. However, this form relates to volume computations via Stokes' theorem: since dz ∧ dw = d(z dw), the integral over D (interpreted through boundary evaluation) connects to the complex volume measure, yielding the Euclidean volume of the polydisk as π² when paired with appropriate normalization factors for the full (2,2)-form. For holomorphic 2-forms of the type ω = f(z, w) dz ∧ dw, where f is holomorphic on a domain containing S, the surface integral ties directly to multivariable extensions of Cauchy's theorem. Specifically, if S is the boundary of a suitable domain and f satisfies the Cauchy-Riemann equations in several variables, the integral ∫_S ω can be linked to the Cauchy integral formula: for a point (z₀, w₀) interior to the domain bounded by S, f(z₀, w₀) = \frac{1}{(2\pi i)^2} \int_S \frac{f(\zeta, \eta)}{(\zeta - z_0)(\eta - w_0)} d\zeta \wedge d\eta, generalizing the one-variable case to higher dimensions via wedge products of 1-forms. This formula holds for polydiscs and more general pseudoconvex domains, emphasizing the role of holomorphic forms in representation theory. Applications of these surface integrals appear prominently in the study of several complex variables, particularly in Hartogs' extension theorem. This theorem asserts that if Ω ⊂ ℂⁿ (n ≥ 2) is open and K ⊂ Ω is compact with connected complement, then any holomorphic function on Ω \ K extends holomorphically to all of Ω. The proof relies on solving the inhomogeneous Cauchy-Riemann equation \bar{\partial} u = g using integral operators over surfaces like boundaries of pseudoconvex domains, where surface integrals of (0,1)-forms dual to holomorphic (n,0)-forms provide the necessary extensions, highlighting the absence of isolated singularities in higher dimensions unlike the one-variable case.

Integral Representations of Functions

Contour integrals provide powerful representations for special functions, enabling their extension to broader domains in the complex plane and yielding insights into their asymptotic behavior. These representations often involve carefully chosen contours that avoid singularities or branch cuts, leveraging the residue theorem to express the function in terms of integrals over paths where the integrand is analytic. Such formulations are essential for functions initially defined by series or real integrals, allowing analytic continuation and approximations for large parameters. One prominent example is the inverse Laplace transform, which recovers the original function from its transform via a contour integral along the Bromwich path—a vertical line parallel to the imaginary axis in the right half-plane, positioned to the right of all singularities. The formula is f(t) = \frac{1}{2\pi i} \int_{\gamma - i\infty}^{\gamma + i\infty} e^{st} F(s) \, ds, where F(s) is the Laplace transform of f(t), and \gamma > 0 ensures convergence. This integral can be evaluated using residues by closing the contour appropriately, depending on the sign of t. The reciprocal of the Gamma function admits a contour integral representation via the Hankel contour, which starts at −∞ along the real axis, encircles the origin counterclockwise in a small circle, and returns to −∞. The representation is \frac{1}{\Gamma(z)} = \frac{1}{2\pi i} \int_H e^{t} t^{-z} \, dt, valid for all complex z except non-positive integers, where t^{-z} takes its principal value with the branch cut along the negative real axis; this form provides analytic continuation of the Gamma function. For the Riemann zeta function, a contour integral provides an analytic continuation beyond \operatorname{Re}(s) > 1, incorporating the pole at s = 1. The representation is \zeta(s) = \frac{1}{s-1} + \frac{1}{2\pi i} \int_C \frac{(-t)^{s-1}}{e^t - 1} \frac{dt}{t}, where C is a contour that starts at +\infty, loops around the origin counterclockwise, and returns to +\infty, avoiding the branch cut along the positive real axis; this form facilitates the functional equation \zeta(s) = 2^s \pi^{s-1} \sin\left(\frac{\pi s}{2}\right) \Gamma(1-s) \zeta(1-s) through contour manipulation and residue computation. These integral representations enable key applications, such as analytic continuation by deforming contours to regions where series diverge, thus defining functions meromorphically in the complex plane. Additionally, they support asymptotic expansions via methods like the saddle-point approximation or steepest descent, where the contour is shifted to pass through dominant saddle points, yielding series approximations for large |z| or other parameters; for instance, the Stirling series for \log \Gamma(z) emerges from the Hankel representation.

References

  1. [1]
    [PDF] 3 Contour integrals and Cauchy's Theorem
    3.1 Line integrals of complex functions. Our goal here will be to discuss integration of complex functions f(z) = u + iv, with particular regard to analytic ...
  2. [2]
    [PDF] Residue theorems and their applications : computing integrals once ...
    May 3, 2013 · Augustin-Luis Cauchy (1789-1857) discovered and rediscovered countless amazing results in the area of complex analysis along with constructing ...
  3. [3]
    [PDF] 221A Lecture Notes on Contour Integrals
    Contour integrals are very useful tools to evaluate integrals. For example, there are many functions whose indefinite integrals can't be written in terms.
  4. [4]
    [PDF] Definite Integrals by Contour Integration
    We simply have to locate the poles inside the contour, find the residues at these poles, and then apply the residue theorem. The more subtle part of the job is ...
  5. [5]
    [PDF] Complex Variables Lecture Notes
    Jun 2, 2021 · in their definition we consider continuous curves, or smooth curves ... (b) Curve is piecewise smooth if γ(t) is piecewise continuously ...
  6. [6]
    [PDF] The Complex Plane
    A curve is called piecewise smooth if each piece of the curve γk|[ak,bk] is smooth for all k = 1,2,··· , n. If γ(t), a ≤ t ≤ b, is a curve, then the curve ...
  7. [7]
    [PDF] Complex Varables Lecture Notes for Math 122A
    Jul 27, 2011 · By a smooth directed curve C we mean either a smooth arc or a smooth closed curve with a choice of direction. Definition 3. A smooth contour ...
  8. [8]
    [PDF] 14.1 Line integral in the complex plane. 14.2 Cauchy's ... - NTNU
    Oct 18, 2018 · ▷ If ˜C is the same curve as C with the reverse orientation (and end points ze and z0 ) then /˜C f (z)dz = −/. C f (z)dz. Yurii Lyubarskii ...
  9. [9]
    Definition:Contour/Closed/Complex Plane - ProofWiki
    Dec 15, 2024 · Also known as ... A closed contour is called a loop in some texts. Some texts define a contour to be what Pr∞fWiki refers to as a closed contour.
  10. [10]
    [PDF] Section 4.39. Contours
    Jan 15, 2020 · A contour is a set of smooth arcs joined end to end. A simple closed contour has the same initial and final values.
  11. [11]
    Jordan Curve Theorem - ProofWiki
    ### Statement of Jordan Curve Theorem
  12. [12]
    [PDF] Contour Integrals I textbook sections 18.1-18.2 - Penn Math
    Mar 22, 2012 · Definition. Given a complex function f (z) continuous around* a piecewise-smooth contour C, we define the contour integral. Z. C f (z)dz = lim.
  13. [13]
    [PDF] Introduction to Complex Analysis Michael Taylor
    Chapter 2. Going deeper – the Cauchy integral theorem and consequences. 5. The Cauchy integral theorem and the Cauchy integral formula.
  14. [14]
    246A, Notes 2: complex integration - Terence Tao - WordPress.com
    Sep 27, 2016 · Remark 4 The above exercise allows one to view the space of curves up to equivalence as a category, with the points in the complex plane being ...
  15. [15]
    [PDF] Contents 3 Complex Integration - Evan Dummit
    evaluate a complex line integral. We may therefore refer to such a contour integral merely by describing the curve ( contour ) and function to be integrated on ...
  16. [16]
    [PDF] 9. Contour Integration
    Contour integration is a powerful technique, based on complex analysis, that allows us to solve certain integrals that are otherwise hard or impossible to solve ...
  17. [17]
    Augustin-Louis Cauchy (1789 - 1857) - Biography - MacTutor
    Cauchy was the first to make a rigorous study of the conditions for convergence of infinite series in addition to his rigorous definition of an integral. His ...
  18. [18]
    [PDF] 18.04 Complex analysis with applications - MIT Mathematics
    Complex analysis is a beautiful, tightly integrated subject. It revolves around complex analytic functions. These are functions that have a complex ...Missing: directed | Show results with:directed
  19. [19]
  20. [20]
    [PDF] NST Part IB Complex Methods - DAMTP
    Jan 15, 2021 · The integral (E.11) then becomes a contour integral along the unit circle and we parametrize the contour γ according to γ : θ 7→ z = eiθ.
  21. [21]
    [PDF] Ahlfors, Complex Analysis
    Complex Analysis has successfully maintained its place as the standard elementary text on functions of one complex variable. There is, never-.
  22. [22]
    [PDF] 2 Cauchy's Theorem and Its - Princeton University
    From there, we can prove the existence of primitives in the interior of some simple regions, and therefore prove Cauchy's theorem in that setting.
  23. [23]
    [PDF] Cauchy's Theorem(s) - John McCuan
    Mar 10, 2022 · The proof of Cauchy's theorem in this case is basically about creating a primitive g by integrating along particular simple paths that can be “ ...
  24. [24]
    [PDF] MORERA'S THEOREM Let Ω be a region. Recall some ideas
    Statement. Morera's theorem is a partial converse of Cauchy's theorem, as follows. Theorem 1.1 (Morera). Let Ω be a region, and let f : Ω −→ C be continuous.
  25. [25]
    [PDF] 18.04 S18 Topic 4: Cauchy's integral formula - MIT OpenCourseWare
    After some more examples we will prove the theorems. After that we will see some remarkable consequences that follow fairly directly from the Cauchy's formula.
  26. [26]
    [PDF] Cauchy's Integral Formula - Trinity University
    We've have proven: Theorem 1 (Cauchy's Theorem for a Disk). Let z0 ∈ C and r > 0. Suppose f (z) is analytic on the disk. D = {z : |z − z0| < r}. Then:.Missing: derivation | Show results with:derivation
  27. [27]
    [PDF] Cauchy's Integral Formula and Power and Laurent Series Expansion
    Cauchy Integral Formula. Goursat's technique removed the additional con- dition of the continuity of the derivative of the holomorphic function from.Missing: derivation | Show results with:derivation
  28. [28]
    [PDF] 18.04 S18 Topic 8: Residue Theorem - MIT OpenCourseWare
    Simple poles occur frequently enough that we'll study computing their residues in some detail. Here are a number of ways to spot a simple pole and compute its ...
  29. [29]
    [PDF] The Basel Problem, The Point at Infinity - UC Berkeley math
    Nov 17, 2015 · So, we multiply everything by 1/z2 to shift the coefficients by two places, and get Res(z−2f(1/z),0) = c−1 = -Res(f,c), as desired. The way to ...
  30. [30]
  31. [31]
    [PDF] 9 Definite integrals using the residue theorem - MIT OpenCourseWare
    To verify the inversion formula we consider the cases > 0 and < 0 separately. For > 0 we use the standard contour. Re(z). Im(z) x1.
  32. [32]
    (PDF) On Sensitivities, Residues and Participations. Applications to ...
    Aug 9, 2025 · The research described in this work has been aimed at the development and application of a promising new framework for analyzing and reducing ...
  33. [33]
    [PDF] Complex Analysis Lecture Notes - UC Davis Math
    Jun 15, 2021 · Solving polynomial equations: historically, this was the motivation for introducing complex numbers by Cardano, who published the famous formula ...
  34. [34]
    [PDF] MATH 311: COMPLEX ANALYSIS — CONTOUR INTEGRALS ...
    To evaluate this real integral using the residue calculus, define the complex function. f(z) = eiz z . This function is meromorphic on C, with its only pole ...
  35. [35]
    [PDF] keyhole contour - DPMMS
    Evaluation of a definite integral using a keyhole contour. In this example, we evaluate, for 0 <α< 1, the integral. I(α) = ∫ ∞. 0 xα x(x + 1) dx. We use a ' ...
  36. [36]
    [PDF] The Hankel representation of Γ(z). - DAMTP
    There are no singularities within the Hankel contour: no branch point and therefore no branch cut. Thus J(z) vanishes. The zero of J(z) cancels the zero of ...
  37. [37]
    [PDF] Cauchy Distribution
    The variance of a standard Cauchy distribution is where which approaches an indefinite value. Similarly, moments of the Cauchy distribution do not exist.
  38. [38]
    [PDF] Lecture 26: Characteristic Functions - EE@IITM
    Indeed, it requires considering two separate contour integrals for t > 0 and t < 0, and invoking Cauchy's residue theorem to evaluate the contour integrals.
  39. [39]
    Help with integrating $\int_0^{\infty} \frac{(\log x)^2}{x^2 + 1 ...
    Sep 6, 2012 · Since this has a pole at t=πi/2, a Laurent series expansion about this point gives the residue as iπ2/8, so the contour integral equals −π3/4. I ...
  40. [40]
    [PDF] Simple evaluation of one of Malmstén's integrals - arXiv
    Sep 25, 2017 · Recently, Blagouchine used contour integration methods to evaluate a family of logarithmic integrals that contains this integral. We evaluate ...
  41. [41]
    Contour integral with a logarithm squared: $\int_0^\infty \frac{\log^2 x ...
    Dec 27, 2015 · We have ∫∞1log2(x)(1+x)2dx=∫01log2(1/x)(1+1/x)2(−dxx2)=∫10log2(x)(1+x)2dx. Hence, our integral becomes I=∫∞0log2(x)(1+x)2dx=2∫10log ...<|separator|>
  42. [42]
    [PDF] Tasty Bits of Several Complex Variables
    May 20, 2025 · ... complex variable. It follows from Green's theorem (Stokes' theorem in two dimensions). You can look forward to for a proof of a more general ...
  43. [43]
    [PDF] Chapter 2 Elementary properties of holomorphic functions in several ...
    Oct 2, 2020 · For functions that are holomorphic in a polydisc we have a direct generalization of the Cauchy integral formula. Cauchy Integral Formula 2.2.1.
  44. [44]
    Cauchy-type integrals in several complex variables
    Jun 9, 2013 · The purpose of this survey is to study Cauchy-type integrals in several complex variables and to announce new results concerning these operators ...
  45. [45]
    A New Look at the Hartogs Extension Phenomenon
    Jan 6, 2020 · We give a new proof of the classical Hartogs extension phenomenon in several complex variables. This proof works equally well for meromorphic functions.<|separator|>
  46. [46]
    Bromwich Integral -- from Wolfram MathWorld
    The inverse of the Laplace transform, given by F(t)=1/(2pii)int_(gamma-iinfty)^(gamma+iinfty)e^(st)f(s)ds, where gamma is a vertical contour
  47. [47]
    5.9 Integral Representations ‣ Properties ‣ Chapter 5 Gamma ...
    Hankel's Loop Integral where the contour begins at − ∞ , circles the origin once in the positive direction, and returns to − ∞ . t − z has its principal value.
  48. [48]
    Sinc Function -- from Wolfram MathWorld
    The sinc function sinc(x), also called the sampling function, is a function that arises frequently in signal processing and the theory of Fourier transforms.
  49. [49]
    [PDF] Introduction to Analytic Number Theory The Riemann zeta function ...
    The Riemann zeta function and its functional equation. (and a review of the Gamma function and Poisson summation). Recall Euler's identity: [ζ(s) :=] ∞. X n=1.
  50. [50]
    [PDF] Chapter 3 Asymptotic Expansion of Integrals - UC Davis Mathematics
    They are the most basic functions that exhibit a transition from oscillatory to exponential behavior, and because of this they arise in many applications (for.