Fact-checked by Grok 2 weeks ago

Residue theorem

The Residue Theorem, also known as Cauchy's Residue Theorem, is a fundamental result in that relates the value of a contour integral of a holomorphic inside and on a simple closed positively oriented except at finitely many isolated singularities inside the to the sum of the residues of the at those singularities. Specifically, if f(z) is analytic inside and on a simple closed C except for finitely many isolated singularities z_k in the interior, then \oint_C f(z) \, dz = 2\pi i \sum_k \operatorname{Res}(f, z_k), where \operatorname{Res}(f, z_k) denotes the residue at z_k. This theorem generalizes and provides a systematic way to compute that would otherwise be challenging, by focusing solely on local behavior at singularities rather than the entire . Named after the French mathematician , the theorem emerged from his pioneering work in the 1820s on complex integration, initially applied to rectangular contours before being extended to general closed paths. Cauchy's contributions built upon earlier work on complex integration by mathematicians such as Euler and Lagrange, but he formalized the connection to contour integrals in his 1825 memoir Mémoire sur les intégrales définies prises entre des limites imaginaires. The residue itself is defined as the a_{-1} in the expansion of f(z) around a , i.e., f(z) = \sum_{n=-\infty}^\infty a_n (z - z_k)^n, capturing the principal part's contribution to the integral. Beyond its theoretical elegance, the Residue Theorem is indispensable for practical computations, enabling the evaluation of real definite integrals (such as those involving rational functions or trigonometric forms) by deforming contours in the to enclose poles. For instance, integrals like \int_{-\infty}^\infty \frac{\sin x}{x} \, dx can be resolved using semicircular contours and residue calculations at relevant poles. Its applications extend to physics and engineering, including , , and solving partial differential equations via transforms, where simplifies otherwise intractable problems. The theorem's power lies in reducing global integral properties to local residue computations, often using limits like \operatorname{Res}(f, z_k) = \lim_{z \to z_k} (z - z_k) f(z) for simple poles.

Preliminaries

states that if a f is holomorphic throughout a simply connected D and \gamma is a simple closed contour within D, then the contour integral of f over \gamma vanishes: \int_{\gamma} f(z) \, dz = 0. /09%3A_Contour_Integration/9.02%3A_Cauchys_Integral_Theorem) A f: D \to \mathbb{C} is holomorphic on the D \subset \mathbb{C} if it is complex differentiable at every point in D, meaning the limit f'(z) = \lim_{h \to 0} \frac{f(z + h) - f(z)}{h} exists for all z \in D, where h is complex. This property implies that holomorphic functions satisfy the Cauchy-Riemann equations and are infinitely differentiable, enabling powerful analytic continuations. A domain D is simply connected if it is path-connected and every simple closed curve in D can be continuously contracted to a point within D, ensuring no "holes" that could enclose singularities. A closed contour \gamma in the is a piecewise smooth curve that is parametrized by a \gamma: [a, b] \to \mathbb{C} with \gamma(a) = \gamma(b), traversed in a specific (typically counterclockwise for positive )./08%3A_Complex_Representations_of_Functions/8.05%3A_Complex_Integration) This theorem, first proved by in his 1825 memoir on complex integration, forms a cornerstone of by establishing that integrals of holomorphic functions over closed paths depend only on the boundary behavior in simply connected regions. It laid the groundwork for evaluating real definite integrals via contour deformation and inspired subsequent developments in function theory. A basic proof proceeds in two steps: first, establish that every holomorphic function on a simply connected domain admits an antiderivative (primitive function F such that F' = f); second, note that the integral over any closed contour \gamma then equals F(\gamma(b)) - F(\gamma(a)) = 0 since \gamma(a) = \gamma(b). To prove the existence of the antiderivative without assuming continuity of f', Édouard Goursat's 1900 refinement (Cauchy-Goursat theorem) shows that if f is holomorphic inside and on a simple closed positively oriented contour \gamma, then \int_{\gamma} f(z) \, dz = 0, even without continuous differentiability. For the general simply connected case, cover D with a triangulation of triangles where the theorem applies locally, then sum the integrals over internal edges that cancel, yielding zero overall./09%3A_Contour_Integration/9.02%3A_Cauchys_Integral_Theorem) When no singularities are present within the , the residue theorem specializes to .

expresses the value of a at an interior point of a in terms of an over the contour itself. Suppose f is holomorphic in a simply connected domain containing the simple closed positively oriented contour \gamma and its interior, and let a be a point inside \gamma. Then, f(a) = \frac{1}{2\pi i} \oint_{\gamma} \frac{f(z)}{z - a} \, dz. This representation, first established by Augustin-Louis Cauchy, demonstrates that the function's value at a is uniquely determined by its boundary values on \gamma, highlighting the rigidity of holomorphic functions. The formula isolates the contribution from the point a by constructing an integrand that has a simple pole at a, with the residue at that pole effectively capturing f(a). This allows evaluation of f(a) without direct knowledge of the function inside \gamma, relying solely on contour integration, which underscores the power of complex integration for pointwise determination. In applications, it enables computation of function values via known integrals or vice versa, serving as a foundational tool in complex analysis. A generalization extends the formula to higher derivatives of f. For the n-th derivative at a, where n \geq 1, f^{(n)}(a) = \frac{n!}{2\pi i} \oint_{\gamma} \frac{f(z)}{(z - a)^{n+1}} \, dz. This follows by repeated differentiation under the integral sign, valid due to the uniform convergence of the resulting integrals on compact sets within the domain. Geometrically, the formula incorporates the winding number of \gamma around a, which measures how many times the contour encircles the point; for a simple closed curve with winding number 1, the prefactor \frac{1}{2\pi i} normalizes the integral to yield f(a) directly. This interpretation emphasizes the topological aspect of contour integration, where the "winding" encodes the encirclement contributing to the function's value. The formula acts as a precursor to residue computations at isolated singularities, such as poles.

Core Concepts

Laurent Series Expansion

The Laurent series provides a generalization of the expansion for functions of a that are holomorphic in an annular region surrounding an , allowing for both positive and negative powers of (z - a). Specifically, if f is holomorphic in the open annulus r < |z - a| < R for 0 ≤ r < R ≤ ∞, then there exists a unique series representation f(z) = \sum_{n=-\infty}^{\infty} c_n (z - a)^n that converges to f(z) in that annulus, where the coefficients c_n are numbers determined by the function f. The coefficients c_n in the Laurent series are computed using Cauchy's integral formula adapted to the annular domain. For any integer n, c_n = \frac{1}{2\pi i} \oint_{\gamma} \frac{f(\zeta)}{(\zeta - a)^{n+1}} \, d\zeta, where γ is a simple closed contour lying in the annulus r < |z - a| < R and oriented positively with respect to a, ensuring the integral captures the behavior of f in the region. This formula holds for all n ∈ ℤ, including negative values, distinguishing it from the Taylor series case where only non-negative powers appear. The Laurent series consists of two parts: the principal part, comprising the terms with negative exponents \sum_{n=1}^{\infty} c_{-n} (z - a)^{-n}, and the regular (or analytic) part, \sum_{n=0}^{\infty} c_n (z - a)^n. The principal part encapsulates the singular behavior of f at the point a, while the regular part is a standard power series holomorphic inside the outer radius R. The nature of the isolated singularity at z = a is classified based on the principal part: if it vanishes (i.e., c_n = 0 for all n < 0), the singularity is removable; if it is a finite sum (c_n = 0 for n < -m for some finite m), it is a pole of order m; and if it has infinitely many nonzero terms, the singularity is essential. Convergence of the Laurent series occurs uniformly on every compact subset of the annulus r < |z - a| < R, analogous to the uniform convergence of Taylor series in disks. The inner radius r limits the domain to exclude the singularity at a, while the outer radius R bounds the region of analyticity; outside this annulus, the series may diverge, reflecting potential other singularities of f. This expansion, first introduced by Pierre Alphonse Laurent in 1843, forms the foundational tool for analyzing isolated singularities in complex analysis.

Definition of Residue

In complex analysis, the residue of a meromorphic function f at an isolated singularity a \in \mathbb{C} is formally defined as the coefficient c_{-1} in its Laurent series expansion centered at a: f(z) = \sum_{n=-\infty}^{\infty} c_n (z - a)^n, \quad \operatorname{Res}(f, a) = c_{-1}. This coefficient captures the principal part of the singularity and is well-defined precisely because a is isolated, allowing the series to converge in a punctured disk $0 < |z - a| < R for some R > 0. Equivalently, the residue admits an integral representation: \operatorname{Res}(f, a) = \frac{1}{2\pi i} \oint_{\gamma} f(z) \, dz, where \gamma is any simple closed , positively oriented, that encloses a but no other of f. This formulation links the residue directly to and underscores its role in evaluating integrals around singularities. The residue quantifies the "strength" of the singularity at a, providing a single that encodes the most singular contribution to the function's behavior near that point, independent of higher- or lower-order terms in the expansion. By Cauchy's theorem on deformation of contours, the value of \operatorname{Res}(f, a) remains unchanged regardless of the specific \gamma chosen, provided it encircles only the at a. This integral form is a direct consequence of the residue theorem applied to a contour enclosing only the singularity at a.

Statement and Proof

Formal Statement

The residue theorem, also known as Cauchy's residue theorem, asserts that if a f is holomorphic in a D except for a finite number of isolated singularities at points a_1, a_2, \dots, a_n in D, and \gamma is a simple closed positively oriented in D that does not pass through any of the singularities, then the of f over \gamma equals $2\pi i times the sum of the residues of f at those singularities enclosed by \gamma. This requires that f has isolated singularities (such as poles or essential singularities) within the region bounded by \gamma, and \gamma is a Jordan curve oriented counterclockwise. The precise formula is \int_\gamma f(z) \, dz = 2\pi i \sum_{k=1}^n \operatorname{Res}(f, a_k), where the sum is over all singularities a_k inside \gamma. For more general closed contours, the theorem extends using the : if \gamma is a closed contour (not necessarily simple) and the singularities are isolated points not on \gamma, then \int_\gamma f(z) \, dz = 2\pi i \sum_k n(\gamma, a_k) \operatorname{Res}(f, a_k), where n(\gamma, a_k) denotes the winding number of \gamma about a_k. The theorem was originally generalized by Augustin-Louis Cauchy in the early 19th century as part of his foundational work on complex integration, with further refinements throughout the 19th century by subsequent mathematicians building on Cauchy's integral theorems.

Outline of Proof

The proof of the residue theorem proceeds by decomposing the function into a holomorphic part and the principal parts of its Laurent series expansions at each isolated singularity inside the contour, leveraging Cauchy's integral theorem to show that the integral of the holomorphic component vanishes./09%3A_Residue_Theorem/9.05%3A_Cauchy_Residue_Theorem) Specifically, for a function f that is holomorphic in a domain except at finitely many isolated singularities a_k enclosed by a simple closed contour \gamma, one constructs a function h(z) that subtracts the singular principal parts from f(z), rendering h holomorphic inside and on \gamma. By Cauchy's integral theorem, the contour integral \oint_\gamma h(z) \, dz = 0, so \oint_\gamma f(z) \, dz equals the sum of the integrals of the principal parts over small circles around each a_k. For a single at a, the principal part of the of f around a contributes to the integral over a small C_a enclosing a. The residue \operatorname{Res}(f, a), defined as the coefficient of (z - a)^{-1} in this expansion, satisfies \oint_{C_a} f(z) \, dz = 2\pi i \operatorname{Res}(f, a), which follows from applying to the term involving $1/(z - a) in the series, as the integrals of all other powers vanish. This extraction isolates the residue's contribution precisely. When multiple isolated singularities a_1, \dots, a_n lie inside \gamma, the proof extends by considering a collection of small non-overlapping circles C_k around each a_k and annular regions between them and \gamma. The over \gamma equals the of over the C_k (after cancellations on the connecting paths), yielding \oint_\gamma f(z) \, dz = 2\pi i \sum_k \operatorname{Res}(f, a_k)./09%3A_Residue_Theorem/9.05%3A_Cauchy_Residue_Theorem) This outline assumes the singularities are isolated and finite in number within the contour, as required for the Laurent expansions to apply locally. For meromorphic functions, where singularities are poles (non-essential), the theorem extends naturally, though essential singularities are handled similarly via their full .

Computing Residues

Residues at Simple Poles

A simple pole of a f(z) at an a occurs when the principal part of its expansion around a consists only of the single term \frac{c_{-1}}{z - a}, where c_{-1} \neq 0 is finite./09:_Residue_Theorem/9.04:_Residues) This residue c_{-1} is the coefficient of the (z - a)^{-1} term in the . The residue at a simple pole a can be computed using the limit formula: \operatorname{Res}(f, a) = \lim_{z \to a} (z - a) f(z). This follows directly from the form of the Laurent series, as the limit isolates the c_{-1} term./09:_Residue_Theorem/9.04:_Residues) For rational functions f(z) = \frac{p(z)}{q(z)}, where p and q are analytic at a, p(a) \neq 0, q(a) = 0, and q'(a) \neq 0 (confirming the simple pole), the residue simplifies to: \operatorname{Res}(f, a) = \frac{p(a)}{q'(a)}. This formula arises from applying L'Hôpital's rule to the limit expression or directly from the Laurent expansion. As an example, consider f(z) = \frac{1}{z^2 + 1} = \frac{1}{(z - i)(z + i)}, which has a simple pole at z = i. Here, p(z) = 1 and q(z) = z^2 + 1, so q'(z) = 2z and q'(i) = 2i. Thus, \operatorname{Res}(f, i) = \frac{1}{2i} = -\frac{i}{2}. Alternatively, using the limit: \operatorname{Res}(f, i) = \lim_{z \to i} (z - i) \frac{1}{(z - i)(z + i)} = \lim_{z \to i} \frac{1}{z + i} = \frac{1}{2i} = -\frac{i}{2}. This matches the formula./09:_Residue_Theorem/9.04:_Residues)

Residues at Higher-Order Poles

A pole of order m at a point z = a for a function f(z) occurs when (z - a)^m f(z) is holomorphic in a neighborhood of a and (z - a)^m f(z) \big|_{z=a} \neq 0, while lower powers of (z - a) times f(z) either fail to be holomorphic or vanish at a./09%3A_Residue_Theorem/9.04%3A_Residues) The residue of f(z) at such a is given by the \operatorname{Res}(f, a) = \frac{1}{(m-1)!} \lim_{z \to a} \frac{d^{m-1}}{dz^{m-1}} \left[ (z - a)^m f(z) \right]. This expression extracts the coefficient of (z - a)^{-1} in the expansion of f(z) around a./09%3A_Residue_Theorem/9.04%3A_Residues) For rational functions, residues at higher-order poles can be computed using this limit formula, often in conjunction with when the limit involves indeterminate forms after differentiation, or by , which expresses the function as a sum of terms including polynomials over powers of (z - a), where the residue is the coefficient of the $1/(z - a) term./09%3A_Residue_Theorem/9.04%3A_Residues) Consider the example f(z) = \frac{1}{(z-1)^3}, which has a of order 3 at z = 1. Here, (z-1)^3 f(z) = 1, so \operatorname{Res}(f, 1) = \frac{1}{2!} \lim_{z \to 1} \frac{d^2}{dz^2} {{grok:render&&&type=render_inline_citation&&&citation_id=1&&&citation_type=wikipedia}} = \frac{1}{2} \cdot 0 = 0. This result aligns with the f(z) = (z-1)^{-3}, which lacks a (z-1)^{-1} term.

Residues at Essential Singularities

An of a f(z) at an z_0 occurs when the principal part of its expansion about z_0 contains infinitely many nonzero terms. This contrasts with poles, where the principal part has only finitely many terms, allowing for limit-based computations. A classic example is the function f(z) = e^{1/z} at z = 0, where the is \sum_{n=0}^{\infty} \frac{1}{n!} z^{-n}, confirming the infinite negative powers. To compute the residue at an essential singularity, one must determine the c_{-1} of the (z - z_0)^{-1} term in the full expansion. This typically requires deriving the series explicitly, often through substitutions like w = 1/(z - z_0) to transform the function into a form amenable to known expansions, such as for exponentials or . In general, the is given by c_k = \frac{1}{2\pi i} \oint_C \frac{f(z)}{(z - z_0)^{k+1}} \, dz for k = -1, but practical evaluation relies on series manipulation rather than direct . For the example f(z) = e^{1/z} at z = 0, substitute w = 1/z to obtain e^w = \sum_{n=0}^{\infty} \frac{w^n}{n!} = \sum_{n=0}^{\infty} \frac{1}{n!} z^{-n}. The term for z^{-1} corresponds to n=1, yielding c_{-1} = \frac{1}{1!} = 1. Thus, \operatorname{Res}(e^{1/z}, 0) = 1. Unlike residues at poles, which can be found using finite-order limit formulas, residues at essential singularities lack a universal closed-form expression and demand case-by-case series analysis, often complicated by the infinite extent of the principal part. This necessitates tailored expansions for each function, such as Fourier or asymptotic series in specific contexts.

Residue at Infinity

The residue at infinity of a f on the extended is defined as \operatorname{Res}(f, \infty) = -\frac{1}{2\pi i} \oint_{\gamma} f(z) \, dz, where \gamma is a closed positively oriented of sufficiently large enclosing all singularities of f in the finite ./09:_Residue_Theorem/9.06:Residue_at%E2%88%9E) This definition arises from considering the integral over \gamma as capturing the "contribution" from the point at when the contour is oriented relative to the exterior region. An equivalent and often more practical is obtained via the w = 1/z, which maps the neighborhood of to the neighborhood of the in the w-plane. Under this , dz = -dw / w^2, and the residue at transforms to \operatorname{Res}(f, \infty) = -\operatorname{Res}_{w=0} \left( \frac{f(1/w)}{w^2} \right). /09:_Residue_Theorem/9.06:Residue_at%E2%88%9E) This allows by expanding the transformed in a around w = 0 and identifying the coefficient of $1/w. A key property is that the sum of the residues of f at all its singularities in the finite plane, together with the residue at infinity, equals zero: \sum_{\text{finite poles } a} \operatorname{Res}(f, a) + \operatorname{Res}(f, \infty) = 0. This follows from applying the residue theorem to a large contour \gamma, where the integral over \gamma vanishes as the radius tends to infinity if f behaves appropriately at infinity, implying the total residue sum is zero on the Riemann sphere. To compute the residue at infinity, one substitutes w = 1/z into f(z) to form g(w) = f(1/w)/w^2, finds the residue of g at w = 0 using standard methods (such as or pole formulas), and negates the result. For example, consider f(z) = 1/z^2. Then g(w) = [1/(1/w)^2]/w^2 = (w^2)/w^2 = 1, which has $1 + 0 \cdot w + \cdots around w = 0, so \operatorname{Res}(g, 0) = 0 and thus \operatorname{Res}(f, \infty) = -0 = 0./09:_Residue_Theorem/9.06:Residue_at%E2%88%9E) This concept extends the residue theorem to the extended and is particularly useful for analyzing contour integrals over large circles.

Applications

Evaluation of Contour Integrals

The residue theorem enables the evaluation of contour integrals over closed paths in the by relating them directly to the residues of the integrand at its isolated singularities within the . For a f(z) that is analytic inside and on a simple closed positively oriented \gamma, except at a finite number of isolated singularities, the theorem states that \int_\gamma f(z) \, dz = 2\pi i \sum \operatorname{Res}(f, z_k), where the sum is over all singularities z_k enclosed by \gamma. The general procedure involves first identifying all singularities of f(z) that lie inside \gamma, ensuring the function is meromorphic in that region. Next, compute the residue at each such singularity using established methods, such as the limit formula for simple poles or the expansion for higher-order poles. The residues are then summed, and the is obtained by multiplying the sum by $2\pi i. This approach simplifies computations that would otherwise require parametrizing the and integrating directly. The selection of the contour \gamma is essential and tailored to the function's and singularities. Circular , such as |z| = R for sufficiently large R, are commonly used for rational functions where the over the arc vanishes as R \to \infty, provided the degree of the denominator exceeds that of the numerator by at least two. Rectangular prove effective for functions exhibiting periodicity or in the imaginary direction, allowing the over opposite sides to cancel or simplify. Keyhole , which encircle the positive real axis while avoiding a branch cut, are particularly useful for functions involving logarithms or fractional powers. For functions with branch cuts, the contour must be designed to respect the branch structure, typically by indenting around the cut with small semicircles or choosing paths that enclose only the desired singularities without crossing the cut. This ensures the function remains single-valued along \gamma. A representative example is the evaluation of \int_{|z|=2} \frac{z^2 + 1}{z(z-1)} \, dz. The integrand has simple poles at z=0 and z=1, both inside the unit circle of radius 2 centered at the origin. The residue at z=0 is \lim_{z \to 0} z \cdot \frac{z^2 + 1}{z(z-1)} = \frac{1}{-1} = -1. The residue at z=1 is \lim_{z \to 1} (z-1) \cdot \frac{z^2 + 1}{z(z-1)} = \frac{2}{1} = 2. The sum of the residues is 1, so the integral equals $2\pi i \cdot 1 = 2\pi i.

Reduction to Real Integrals

One common application of the residue theorem involves evaluating real definite integrals over the infinite line \int_{-\infty}^{\infty} f(x) \, dx, where f(z) is a function analytic except for isolated singularities, often rational functions with the degree of the denominator exceeding that of the numerator by at least two. The technique extends the real integral to a complex contour integral over a semicircular path \gamma_R in the upper half-plane, consisting of the real interval [-R, R] and the semicircular arc \Gamma_R of radius R. By the residue theorem, \int_{\gamma_R} f(z) \, dz = 2\pi i \sum \Res(f, z_k), where the sum is over poles z_k inside \gamma_R. As R \to \infty, if the integral over \Gamma_R vanishes, then \int_{-\infty}^{\infty} f(x) \, dx = 2\pi i \sum \Res(f, z_k) for poles in the upper half-plane; the lower half-plane may be used analogously, with the sign adjusted for orientation. The vanishing of the arc integral \int_{\Gamma_R} f(z) \, dz \to 0 as R \to \infty requires that |f(z)| \to 0 uniformly for \arg z \in [0, \pi], typically ensured by the function's behavior at infinity, such as |f(z)| \leq M / |z|^{1+\epsilon} for some M > 0 and \epsilon > 0. For integrals involving oscillatory factors like e^{i a x} with a > 0, closing in the upper half-plane, Jordan's lemma provides stricter conditions: if |f(z)| \leq M / R on \Gamma_R for large R, then \left| \int_{\Gamma_R} f(z) e^{i a z} \, dz \right| \to 0 as R \to \infty, leveraging the exponential decay \operatorname{Im}(z) > 0 implies |e^{i a z}| = e^{-a \operatorname{Im}(z)} \to 0. This lemma is essential for Fourier-type integrals. A classic example is \int_{-\infty}^{\infty} \frac{dx}{x^2 + 1}, where f(z) = 1/(z^2 + 1) has simple poles at z = \pm i. Since the function is even and the poles are symmetric, close in the upper half-plane enclosing z = i, where \Res(f, i) = 1/(2i) = -i/2. The arc vanishes because |f(z)| \sim 1/|z|^2 \to 0. Thus, the integral equals $2\pi i \cdot (-i/2) = \pi. For odd integrands, the integral may vanish by symmetry, but the method still applies if the contour conditions hold. When poles lie on the real axis, the integral is interpreted as the , requiring an indented semicircular around the with \epsilon \to 0. The contribution from the indentation is -\pi i \Res(f, p) for a simple at real p (upper half-plane ), added to $2\pi i times interior residues, yielding the principal value plus this half-residue term. This handles cases like \int_{-\infty}^{\infty} \frac{dx}{x^2 (x-1)}.

Sums and Infinite Products

One prominent application of the residue theorem in evaluating infinite sums involves the \pi \cot(\pi z), which has simple poles at all s n \in \mathbb{Z} with residue 1 at each such point. For a function f(z) that is analytic at the integers and meromorphic elsewhere, consider the integral \oint_{C_N} \pi \cot(\pi z) f(z) \, dz over a large square C_N with vertices at (N + 1/2)(\pm 1 \pm i), where N is a positive , enclosing the poles at integers from -N to N and any poles of f inside. By the residue theorem, this integral equals $2\pi i times the sum of residues inside C_N, which includes \sum_{n=-N}^N f(n) from the poles of \cot(\pi z) and the residues of \pi \cot(\pi z) f(z) at the poles of f. Under suitable growth conditions on f, such as |f(z)| = O(1/|z|^{1+\epsilon}) for some \epsilon > 0 as |z| \to \infty in the strips parallel to the real axis, the integral over C_N vanishes as N \to \infty because |\cot(\pi z)| is bounded on the contour away from integers, and the length of C_N grows like N while |f(z)| \sim 1/N^{1+\epsilon}. Thus, the sum of all residues of \pi \cot(\pi z) f(z) is zero, yielding the cotangent summation formula: \sum_{n=-\infty}^\infty f(n) = -\sum_k \operatorname{Res}_{z=z_k} \left[ \pi \cot(\pi z) f(z) \right], where the sum on the right is over the poles z_k of f. This holds provided f satisfies the aforementioned decay condition to ensure convergence of the series and vanishing of the contour integral. A classic example is the evaluation of \sum_{n=1}^\infty \frac{1}{n^2} = \frac{\pi^2}{6}, a solution to the . Consider f(z) = \frac{1}{z^2}, which has a pole of order 2 at z=0. The residues at nonzero integers n are f(n) = \frac{1}{n^2}, and there are no other poles. The residue at z=0 is found from the \pi \cot(\pi z) = \frac{1}{z} - \frac{\pi^2 z}{3} + O(z^3), so \frac{\pi \cot(\pi z)}{z^2} = \frac{1}{z^3} - \frac{\pi^2}{3z} + O(z), giving \operatorname{Res}_{z=0} = -\frac{\pi^2}{3}. The contour integral vanishes under the decay condition, so \sum_{n \neq 0} \frac{1}{n^2} = \frac{\pi^2}{3}, and thus \sum_{n=1}^\infty \frac{1}{n^2} = \frac{\pi^2}{6}. The residue theorem also facilitates the evaluation of through . For an P(z) expressed as a Weierstrass product P(z) = z^m e^{g(z)} \prod_n (1 - z/a_n) e^{z/a_n + \cdots}, the \frac{P'(z)}{P(z)} = \frac{m}{z} + g'(z) + \sum_n \frac{1}{z - a_n} + \sum_n \left( \frac{1}{a_n} + \cdots \right) can be derived by considering integrals of \frac{P'(w)}{P(w)} \frac{1}{w - z} or using residue expansions akin to that of \pi \cot(\pi z), which itself arises from residues in the product formula for \sin(\pi z). Specifically, the partial fraction expansion \pi \cot(\pi z) = \frac{1}{z} + \sum_{n=1}^\infty \left( \frac{1}{z - n} + \frac{1}{z + n} \right) is obtained via residues of \pi \cot(\pi w) / (w - z)^2 or similar kernels, linking directly to the \sin(\pi z) = \pi z \prod_{n=1}^\infty (1 - z^2/n^2) by integrating or differentiating the expansion. This method extends to general products by summing residues to determine the exponents and factors.

Special Functions

The , originally defined by the infinite series \zeta(s) = \sum_{n=1}^{\infty} n^{-s} for \Re(s) > 1, admits an to the (with a simple pole at s=1) through representations involving contour integrals evaluated via the residue theorem. A key such representation employs the Hankel contour C, which starts at +\infty, encircles the origin counterclockwise while avoiding the positive real axis, and returns to +\infty, yielding \zeta(s) = \frac{1}{2\pi i} \int_C \frac{z^{s-1}}{e^z - 1} \, dz for \Re(s) > 0, where the branch is defined appropriately. This integral provides the meromorphic continuation, as the integrand's behavior on the contour ensures convergence, and residues are not directly computed here but underpin the deformation arguments for broader domains. The \zeta(s) = 2^s \pi^{s-1} \sin\left(\frac{\pi s}{2}\right) \Gamma(1-s) \zeta(1-s) is derived using residues at the poles of the in a related . Consider the \pi \cot(\pi z) \Gamma(1-z) \zeta(1-z); integrating this over a suitable indented enclosing the positive integers and deforming it leads to the equation via the residue theorem, where residues at z = n (integers) recover the series, and poles from \Gamma(1-z) at negative integers contribute to the . Specifically, the residues at the poles z = 1, 2, \dots of \cot(\pi z) sum to terms involving \zeta(1-z), while shifting the captures the gamma factor's poles, equating the two sides. This approach, sketched by Riemann in , highlights the residue theorem's role in linking the 's values across the critical line. Values of \zeta(2k) for positive integers k are computed explicitly using residues of \pi \cot(\pi z) / z^{2k}. The residue at z=0 of this function is -\frac{B_{2k}}{2k}, where B_{2k} are numbers, yielding the formula \zeta(2k) = (-1)^{k+1} \frac{B_{2k} (2\pi)^{2k}}{2 (2k)!}, obtained by applying the residue theorem to a large where the integral vanishes, leaving the sum of residues at integers equal to the residue at zero. The reflection formula for the , \Gamma(s) \Gamma(1-s) = \frac{\pi}{\sin(\pi s)}, is likewise established via residues. Consider the \pi \cot(\pi z) \Gamma(z) \Gamma(1-z); its residues at the poles z = n (non-positive integers from \Gamma(z)) and z = 1-n sum to zero over a vanishing contour integral, confirming the by evaluating the simple poles of \cot(\pi z). This formula, dating to Euler in the but rigorously proved with complex methods in the 19th, interconnects with via the functional equation's gamma factor. Eisenstein series, introduced by Gotthold Eisenstein in the mid-19th century as sums over lattice points, G_k(\tau) = \sum_{(m,n) \neq (0,0)} (m\tau + n)^{-k} for \Im(\tau) > 0 and even integer k \geq 4, are non-constant holomorphic modular forms whose properties involve residue computations. In the theory of modular forms, residues arise in the analytic continuation of non-holomorphic Eisenstein series E(z, s) = \sum_{\gamma \in \Gamma_\infty \backslash \Gamma} \Im(\gamma z)^s |cz + d|^{-2s}, where the residue at s = k/2 yields the holomorphic Eisenstein series G_k(\tau). These residues, computed via unfolding the sum over the modular group \Gamma = \mathrm{SL}_2(\mathbb{Z}), express G_k(\tau) in terms of lattice sums, with constant terms linked to zeta values like \zeta(1-k) = -\frac{B_k}{k}. Such computations, building on 19th-20th century developments by Eisenstein and later mathematicians like Hecke, underscore residues' utility in modular form decompositions and Fourier expansions.

References

  1. [1]
    Residue Theorem -- from Wolfram MathWorld
    This amazing theorem therefore says that the value of a contour integral for any contour in the complex plane depends only on the properties of a few very ...
  2. [2]
    [PDF] 18.04 S18 Topic 8: Residue Theorem - MIT OpenCourseWare
    In this section we'll explore calculating residues. We've seen enough already to know that this will be useful. We will see that even more clearly when we look ...
  3. [3]
    Applications of Contour Integration and the Residue Theorem - NHSJS
    Nov 30, 2024 · According to, Cauchy conjured and proved the first version of this theorem in 1822 for closed rectangular contours, and later expanded this ...
  4. [4]
    Historical synopsis of Cauchy residue theorem - NASA/ADS
    This paper is aimed to discuss the importance of Cauchy residue theorem through different aspects. Firstly, we will overview Cauchy residue theorem and its ...
  5. [5]
    The Cauchy residue trick: spectral analysis made “easy”
    Nov 7, 2020 · The Cauchy residue theorem can be used to compute integrals, by choosing the appropriate contour, looking for poles and computing the associated ...
  6. [6]
    Residue Theorem - GeeksforGeeks
    Jul 5, 2024 · Residue Theorem is a powerful tool in complex analysis for evaluating contour integrals. Residue Theorem states that if a function f(z) is analytic inside and ...
  7. [7]
    Cauchy-Goursat Theorem - Complex Analysis
    In 1825 the French mathematician Augustin-Louis Cauchy proved one of the most important theorems in complex analysis: (Cauchy's Theorem) Suppose that f is ...
  8. [8]
    Holomorphic Function | Brilliant Math & Science Wiki
    In complex analysis, a holomorphic function is a complex differentiable function. The condition of complex differentiability is very strong, and leads to an ...
  9. [9]
  10. [10]
    Simply connected definition - Math Insight
    A simply connected domain is a path-connected domain where one can continuously shrink any simple closed curve into a point while remaining in the domain.
  11. [11]
    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.
  12. [12]
    Cauchy integral theorem - Encyclopedia of Mathematics
    Jan 3, 2014 · This, essentially, was the original formulation of the theorem as proposed by A.L. Cauchy (1825) (see [Ca]); similar formulations may be found ...
  13. [13]
    What was the motivation for Cauchy's Integral Theorem?
    Jan 24, 2019 · The original motivation, and an inkling of the integral formula, came from Cauchy's attempts to compute improper real integrals.
  14. [14]
    [PDF] The Cauchy-Goursat Theorem - UCSB Math
    Combining this theorem with Theorem (§42), every function f that is analytic on a simply connected domain D must have an antiderivative on the domain D. • Given ...
  15. [15]
    Mémoire sur les intégrales définies, prises entre des limites ...
    Nov 26, 2009 · Mémoire sur les intégrales définies, prises entre des limites imaginaires. by: Cauchy, Augustin Louis, Baron, 1789-1857. Publication date: 1825.
  16. [16]
    Cauchy Integral Formula -- from Wolfram MathWorld
    Cauchy's integral formula states that f(z_0)=1/(2pii)∮_gamma(f(z)dz)/(z-z_0), (1) where the integral is a contour integral along the contour gamma enclosing ...
  17. [17]
    [PDF] Cauchy's Integral Formula - Trinity University
    prove the Cauchy integral formula. Daileda. Cauchy's Formula. Page 9. Strong ... (or winding number) of γ with respect to z0 is. I(γ;z0) = 1. 2πi ∫γ dz z ...
  18. [18]
    [PDF] Complex VARIABLES AND APPLICATIONS, EIGHTH EDITION
    JAMES WARD BROWN is Professor of Mathematics at The University of. Michigan– Dearborn. He earned his A.B. in physics from Harvard University and his. A.M. and ...
  19. [19]
    Complex Residue -- from Wolfram MathWorld
    The residues of a holomorphic function at its poles characterize a great deal of the structure of a function, appearing for example in the amazing residue ...
  20. [20]
    [PDF] The Residue Theorem and its consequences
    With Laurent series and the classification of singularities in hand, it is easy to prove the Residue Theorem. In addition to being a handy tool for ...
  21. [21]
    [PDF] A First Course in Complex Analysis - matthias beck
    The goal our book works toward is the Residue Theorem, including some nontraditional applications from both continuous and discrete mathematics. More than 250 ...
  22. [22]
    [PDF] Math 346 Lecture #30 11.7 The Residue Theorem
    To state the Residue Theorem we first need to understand isolated singularities of holomorphic functions and quantities called winding numbers. As always we let ...
  23. [23]
    [PDF] Section 6.71. Residues at Infinity
    Apr 7, 2018 · The residue of f at infinity is. Resz=∞f(z) = 1. 2πi ∫C0 f(z)dz. Note. The residue of f at infinity is defined in terms of parameter R0 which.
  24. [24]
    [PDF] 9 Definite integrals using the residue theorem - MIT OpenCourseWare
    The theorems in this section will guide us in choosing the closed contour described in the intro- duction. The first theorem is for functions that decay ...
  25. [25]
    [PDF] the residue theorem • Trig and indefinite integrals
    • Residue Theorem. Let D be a ... Thus contour integration is a powerful new tool to evaluate real integrals, although it is not a magic wand and cannot.
  26. [26]
    Jordan's Lemma -- from Wolfram MathWorld
    Jordan's lemma shows the value of the integral I=int_(-infty)^inftyf(x)e^(iax)dx along the infinite upper semicircle and with a>0 is 0 for nice functions.
  27. [27]
    [PDF] Mathematics 503 Complex Analysis Fall 2017 Using the residue ...
    ... residue theorem applies to give information about real integrals. The main observation is that when a function satisfying lim|z|→∞ zf(z) = 0 is integrated ...
  28. [28]
    [PDF] Definite Integrals by Contour Integration
    Definite integrals can be found using contour integrals by locating poles, finding residues, and applying the residue theorem. Choosing a suitable contour is ...
  29. [29]
    [PDF] The Calculus of Residues
    In this section we shall see how to use the residue theorem to to evaluate certain real integrals which were not possible using real integration techniques ...
  30. [30]
    [PDF] COMPLEX ANALYSIS: LECTURE 27 (27.0) Residue theorem - review.
    The residue theorem uses a holomorphic function, a contour, and points inside the contour to compute real integrals. It combines contour deformation and  ...
  31. [31]
    [PDF] The residue theorem and its applications
    This text contains some notes to a three hour lecture in complex analysis given at Caltech. The lectures start from scratch and contain an essentially ...Missing: textbook | Show results with:textbook
  32. [32]
    [PDF] Summation of Series via the Residue Theorem - UNCW
    The contour CN can be broken into four pieces, as noted in Figure. 2. | cot 7z| = \. \. \. \ ei7z + e−i7z.
  33. [33]
    [PDF] Numerous Proofs of ζ(2) =
    Theorem 11 (Residue theorem). Assume f is analytic except for singu- larities at zj , and that C is a closed, piecewise smooth curve. Then. 1. 2πi. Z. C f(z) dz ...
  34. [34]
    [PDF] On the Number of Prime Numbers less than a Given Quantity ...
    On the Number of Prime Numbers less than a. Given Quantity. (Ueber die Anzahl der Primzahlen unter einer gegebenen Grösse.) Bernhard Riemann. Translated by ...
  35. [35]
    [PDF] RIEMANN ZETA-FUNCTION
    The Riemann zeta-function {(.s) has its origin in the identity ... By the calculus of residues, the first term is equal to . {; 1. 1. 2 m 6 irifl ...
  36. [36]
    [PDF] Proof of Functional Equation by Contour Integral and Residues ...
    Derivation of Functional Equation. We now look at the functional equation for the Riemann zeta function. ζ(1 − z)=21−zπ−z cos.
  37. [37]
    [PDF] Math 3228 - Week 9 • The Riemann Zeta function - UCLA Mathematics
    We use the residue theorem. The function og. < <-Вis not analytic on the positive real axis, but is meromorphic everywhere else, with simple poles at integer ...
  38. [38]
    [PDF] Lecture-25 - IISc Math
    Theorem 0.3 (Euler reflection formula). The Gamma function satisfies the identity. Γ(s)Γ(1 - s) = π sinπs . Proof. By analytic continuation, it enough to ...
  39. [39]
    Examples of Modular Forms of Level 1 - William Stein
    In this chapter we study in detail the structure of level 1 modular forms, ie, modular forms on \SL_2(\Z)=\Gamma_0(1)=\Gamma_1(1).
  40. [40]
    [2105.05437] Residue of some Eisenstein series - arXiv
    May 12, 2021 · The purpose of this study is to provide concrete forms of the residue of E_0^{(m)}(z,s) at s=m/2.Missing: theorem | Show results with:theorem
  41. [41]
    [PDF] Modular forms on GL2 - UCSD Math
    We will prove a generalization of Theorem 3 for the Eisenstein series E(g,Φ,s). Let us first explain, however, that these functions E(g,Φ,s) do generalize ...
  42. [42]
    [PDF] modular forms and the four squares Theorem
    Essentially, all modular forms are given by Eisenstein series, which we will now define. Furthermore, the Eisenstein series of higher weights are generated ...