Cauchy's integral formula

Last updated

In mathematics, Cauchy's integral formula, named after Augustin-Louis Cauchy, is a central statement in complex analysis. It expresses the fact that a holomorphic function defined on a disk is completely determined by its values on the boundary of the disk, and it provides integral formulas for all derivatives of a holomorphic function. Cauchy's formula shows that, in complex analysis, "differentiation is equivalent to integration": complex differentiation, like integration, behaves well under uniform limits – a result that does not hold in real analysis.

Contents

Theorem

Let U be an open subset of the complex plane C, and suppose the closed disk D defined as

is completely contained in U. Let f : UC be a holomorphic function, and let γ be the circle, oriented counterclockwise, forming the boundary of D. Then for every a in the interior of D,

The proof of this statement uses the Cauchy integral theorem and like that theorem, it only requires f to be complex differentiable. Since can be expanded as a power series in the variable

it follows that holomorphic functions are analytic, i.e. they can be expanded as convergent power series. In particular f is actually infinitely differentiable, with

This formula is sometimes referred to as Cauchy's differentiation formula.

The theorem stated above can be generalized. The circle γ can be replaced by any closed rectifiable curve in U which has winding number one about a. Moreover, as for the Cauchy integral theorem, it is sufficient to require that f be holomorphic in the open region enclosed by the path and continuous on its closure.

Note that not every continuous function on the boundary can be used to produce a function inside the boundary that fits the given boundary function. For instance, if we put the function f(z) = 1/z, defined for |z| = 1, into the Cauchy integral formula, we get zero for all points inside the circle. In fact, giving just the real part on the boundary of a holomorphic function is enough to determine the function up to an imaginary constant — there is only one imaginary part on the boundary that corresponds to the given real part, up to addition of a constant. We can use a combination of a Möbius transformation and the Stieltjes inversion formula to construct the holomorphic function from the real part on the boundary. For example, the function f(z) = iiz has real part Re f(z) = Im z. On the unit circle this can be written i/ziz/2. Using the Möbius transformation and the Stieltjes formula we construct the function inside the circle. The i/z term makes no contribution, and we find the function iz. This has the correct real part on the boundary, and also gives us the corresponding imaginary part, but off by a constant, namely i.

Proof sketch

By using the Cauchy integral theorem, one can show that the integral over C (or the closed rectifiable curve) is equal to the same integral taken over an arbitrarily small circle around a. Since f(z) is continuous, we can choose a circle small enough on which f(z) is arbitrarily close to f(a). On the other hand, the integral

over any circle C centered at a. This can be calculated directly via a parametrization (integration by substitution) z(t) = a + εeit where 0 ≤ t ≤ 2π and ε is the radius of the circle.

Letting ε → 0 gives the desired estimate

Example

Surface of the real part of the function g(z) =
z/z + 2z + 2 and its singularities, with the contours described in the text. ComplexResiduesExample.png
Surface of the real part of the function g(z) = z/z + 2z + 2 and its singularities, with the contours described in the text.

Let

and let C be the contour described by |z| = 2 (the circle of radius 2).

To find the integral of g(z) around the contour C, we need to know the singularities of g(z). Observe that we can rewrite g as follows:

where z1 = − 1 + i and z2 = − 1 − i.

Thus, g has poles at z1 and z2. The moduli of these points are less than 2 and thus lie inside the contour. This integral can be split into two smaller integrals by Cauchy–Goursat theorem; that is, we can express the integral around the contour as the sum of the integral around z1 and z2 where the contour is a small circle around each pole. Call these contours C1 around z1 and C2 around z2.

Now, each of these smaller integrals can be evaluated by the Cauchy integral formula, but they first must be rewritten to apply the theorem. For the integral around C1, define f1 as f1(z) = (zz1)g(z). This is analytic (since the contour does not contain the other singularity). We can simplify f1 to be:

and now

Since the Cauchy integral formula says that:

we can evaluate the integral as follows:

Doing likewise for the other contour:

we evaluate

The integral around the original contour C then is the sum of these two integrals:

An elementary trick using partial fraction decomposition:

Consequences

The integral formula has broad applications. First, it implies that a function which is holomorphic in an open set is in fact infinitely differentiable there. Furthermore, it is an analytic function, meaning that it can be represented as a power series. The proof of this uses the dominated convergence theorem and the geometric series applied to

The formula is also used to prove the residue theorem, which is a result for meromorphic functions, and a related result, the argument principle. It is known from Morera's theorem that the uniform limit of holomorphic functions is holomorphic. This can also be deduced from Cauchy's integral formula: indeed the formula also holds in the limit and the integrand, and hence the integral, can be expanded as a power series. In addition the Cauchy formulas for the higher order derivatives show that all these derivatives also converge uniformly.

The analog of the Cauchy integral formula in real analysis is the Poisson integral formula for harmonic functions; many of the results for holomorphic functions carry over to this setting. No such results, however, are valid for more general classes of differentiable or real analytic functions. For instance, the existence of the first derivative of a real function need not imply the existence of higher order derivatives, nor in particular the analyticity of the function. Likewise, the uniform limit of a sequence of (real) differentiable functions may fail to be differentiable, or may be differentiable but with a derivative which is not the limit of the derivatives of the members of the sequence.

Another consequence is that if f(z) = Σ anzn is holomorphic in |z| < R and 0 < r < R then the coefficients an satisfy Cauchy's inequality [1]

From Cauchy's inequality, one can easily deduce that every bounded entire function must be constant (which is Liouville's theorem).

The formula can also be used to derive Gauss's Mean-Value Theorem, which states [2]

In other words, the average value of f over the circle centered at z with radius r is f(z). This can be calculated directly via a parametrization of the circle.

Generalizations

Smooth functions

A version of Cauchy's integral formula is the Cauchy–Pompeiu formula, [3] and holds for smooth functions as well, as it is based on Stokes' theorem. Let D be a disc in C and suppose that f is a complex-valued C1 function on the closure of D. Then [4] [5] [6]

One may use this representation formula to solve the inhomogeneous Cauchy–Riemann equations in D. Indeed, if φ is a function in D, then a particular solution f of the equation is a holomorphic function outside the support of μ. Moreover, if in an open set D,

for some φCk(D) (where k  1), then f(ζ, ζ) is also in Ck(D) and satisfies the equation

The first conclusion is, succinctly, that the convolution μk(z) of a compactly supported measure with the Cauchy kernel

is a holomorphic function off the support of μ. Here p.v. denotes the principal value. The second conclusion asserts that the Cauchy kernel is a fundamental solution of the Cauchy–Riemann equations. Note that for smooth complex-valued functions f of compact support on C the generalized Cauchy integral formula simplifies to

and is a restatement of the fact that, considered as a distribution, z)−1 is a fundamental solution of the Cauchy–Riemann operator /. [7]

The generalized Cauchy integral formula can be deduced for any bounded open region X with C1 boundary X from this result and the formula for the distributional derivative of the characteristic function χX of X:

where the distribution on the right hand side denotes contour integration along X. [8]

Proof

For calculate:

then traverse in the anti-clockwise direction. Fix a point and let denote arc length on measured from anti-clockwise. Then, if is the length of is a parametrization of . The derivative is a unit tangent to and is the unit outward normal on . We are lined up for use of the divergence theorem: put so that and we get

Hence we proved .

Now we can deduce the generalized Cauchy integral formula:

Proof

Since and since this distribution is locally in of the form "distribution times C function", so we may apply the Leibniz rule to calculate its derivatives:

Using that z)−1 is a fundamental solution of the Cauchy–Riemann operator /, we get :

Applying to :

where is used in the last line.

Rearranging, we get

as desired.

Several variables

In several complex variables, the Cauchy integral formula can be generalized to polydiscs. [9] Let D be the polydisc given as the Cartesian product of n open discs D1, ..., Dn:

Suppose that f is a holomorphic function in D continuous on the closure of D. Then

where ζ = (ζ1,...,ζn) ∈ D.

In real algebras

The Cauchy integral formula is generalizable to real vector spaces of two or more dimensions. The insight into this property comes from geometric algebra, where objects beyond scalars and vectors (such as planar bivectors and volumetric trivectors) are considered, and a proper generalization of Stokes' theorem.

Geometric calculus defines a derivative operator ∇ = êii under its geometric product — that is, for a k-vector field ψ(r), the derivative ψ generally contains terms of grade k + 1 and k − 1. For example, a vector field (k = 1) generally has in its derivative a scalar part, the divergence (k = 0), and a bivector part, the curl (k = 2). This particular derivative operator has a Green's function:

where Sn is the surface area of a unit n-ball in the space (that is, S2 = 2π, the circumference of a circle with radius 1, and S3 = 4π, the surface area of a sphere with radius 1). By definition of a Green's function,

It is this useful property that can be used, in conjunction with the generalized Stokes theorem:

where, for an n-dimensional vector space, dS is an (n − 1)-vector and dV is an n-vector. The function f(r) can, in principle, be composed of any combination of multivectors. The proof of Cauchy's integral theorem for higher dimensional spaces relies on the using the generalized Stokes theorem on the quantity G(r, r′) f(r′) and use of the product rule:

When f = 0, f(r) is called a monogenic function, the generalization of holomorphic functions to higher-dimensional spaces — indeed, it can be shown that the Cauchy–Riemann condition is just the two-dimensional expression of the monogenic condition. When that condition is met, the second term in the right-hand integral vanishes, leaving only

where in is that algebra's unit n-vector, the pseudoscalar. The result is

Thus, as in the two-dimensional (complex analysis) case, the value of an analytic (monogenic) function at a point can be found by an integral over the surface surrounding the point, and this is valid not only for scalar functions but vector and general multivector functions as well.

See also

Notes

  1. Titchmarsh 1939 , p. 84
  2. "Gauss's Mean-Value Theorem". Wolfram Alpha Site.
  3. Pompeiu 1905
  4. "§2. Complex 2-Forms: Cauchy-Pompeiu's Formula" (PDF).
  5. Hörmander 1966 , Theorem 1.2.1
  6. "Theorem 4.1.1 (Cauchy–Pompeiu)" (PDF).
  7. Hörmander 1983 , pp. 63, 81
  8. Hörmander 1983 , pp. 62–63
  9. Hörmander 1966 , Theorem 2.2.1

Related Research Articles

<span class="mw-page-title-main">Laplace's equation</span> Second-order partial differential equation

In mathematics and physics, Laplace's equation is a second-order partial differential equation named after Pierre-Simon Laplace, who first studied its properties. This is often written as

<span class="mw-page-title-main">Dirac delta function</span> Generalized function whose value is zero everywhere except at zero

In mathematical analysis, the Dirac delta function, also known as the unit impulse, is a generalized function on the real numbers, whose value is zero everywhere except at zero, and whose integral over the entire real line is equal to one. Since there is no function having this property, to model the delta "function" rigorously involves the use of limits or, as is common in mathematics, measure theory and the theory of distributions.

<span class="mw-page-title-main">Fourier series</span> Decomposition of periodic functions into sums of simpler sinusoidal forms

A Fourier series is an expansion of a periodic function into a sum of trigonometric functions. The Fourier series is an example of a trigonometric series, but not all trigonometric series are Fourier series. By expressing a function as a sum of sines and cosines, many problems involving the function become easier to analyze because trigonometric functions are well understood. For example, Fourier series were first used by Joseph Fourier to find solutions to the heat equation. This application is possible because the derivatives of trigonometric functions fall into simple patterns. Fourier series cannot be used to approximate arbitrary functions, because most functions have infinitely many terms in their Fourier series, and the series do not always converge. Well-behaved functions, for example smooth functions, have Fourier series that converge to the original function. The coefficients of the Fourier series are determined by integrals of the function multiplied by trigonometric functions, described in Common forms of the Fourier series below.

<span class="mw-page-title-main">Cauchy's integral theorem</span> Theorem in complex analysis

In mathematics, the Cauchy integral theorem in complex analysis, named after Augustin-Louis Cauchy, is an important statement about line integrals for holomorphic functions in the complex plane. Essentially, it says that if is holomorphic in a simply connected domain Ω, then for any simply closed contour in Ω, that contour integral is zero.

<span class="mw-page-title-main">Residue (complex analysis)</span> Attribute of a mathematical function

In mathematics, more specifically complex analysis, the residue is a complex number proportional to the contour integral of a meromorphic function along a path enclosing one of its singularities. Residues can be computed quite easily and, once known, allow the determination of general contour integrals via the residue theorem.

<span class="mw-page-title-main">Residue theorem</span> Concept of complex analysis

In complex analysis, the residue theorem, sometimes called Cauchy's residue theorem, is a powerful tool to evaluate line integrals of analytic functions over closed curves; it can often be used to compute real integrals and infinite series as well. It generalizes the Cauchy integral theorem and Cauchy's integral formula. The residue theorem should not be confused with special cases of the generalized Stokes' theorem; however, the latter can be used as an ingredient of its proof.

In vector calculus, Green's theorem relates a line integral around a simple closed curve C to a double integral over the plane region D bounded by C. It is the two-dimensional special case of Stokes' theorem.

In complex analysis, Liouville's theorem, named after Joseph Liouville, states that every bounded entire function must be constant. That is, every holomorphic function for which there exists a positive number such that for all is constant. Equivalently, non-constant holomorphic functions on have unbounded images.

<span class="mw-page-title-main">Radon transform</span> Integral transform

In mathematics, the Radon transform is the integral transform which takes a function f defined on the plane to a function Rf defined on the (two-dimensional) space of lines in the plane, whose value at a particular line is equal to the line integral of the function over that line. The transform was introduced in 1917 by Johann Radon, who also provided a formula for the inverse transform. Radon further included formulas for the transform in three dimensions, in which the integral is taken over planes. It was later generalized to higher-dimensional Euclidean spaces and more broadly in the context of integral geometry. The complex analogue of the Radon transform is known as the Penrose transform. The Radon transform is widely applicable to tomography, the creation of an image from the projection data associated with cross-sectional scans of an object.

In the mathematical field of complex analysis, contour integration is a method of evaluating certain integrals along paths in the complex plane.

In physics and mathematics, the Helmholtz decomposition theorem or the fundamental theorem of vector calculus states that any sufficiently smooth, rapidly decaying vector field in three dimensions can be resolved into the sum of an irrotational (curl-free) vector field and a solenoidal (divergence-free) vector field. This is named after Hermann von Helmholtz.

<span class="mw-page-title-main">Argument principle</span> Theorem in complex analysis

In complex analysis, the argument principle is a theorem relating the difference between the number of zeros and poles of a meromorphic function to a contour integral of the function's logarithmic derivative.

In calculus, the Leibniz integral rule for differentiation under the integral sign states that for an integral of the form

<span class="mw-page-title-main">Multiple integral</span> Generalization of definite integrals to functions of multiple variables

In mathematics (specifically multivariable calculus), a multiple integral is a definite integral of a function of several real variables, for instance, f(x, y) or f(x, y, z). Physical (natural philosophy) interpretation: S any surface, V any volume, etc.. Incl. variable to time, position, etc.

<span class="mw-page-title-main">Characteristic function (probability theory)</span> Fourier transform of the probability density function

In probability theory and statistics, the characteristic function of any real-valued random variable completely defines its probability distribution. If a random variable admits a probability density function, then the characteristic function is the Fourier transform of the probability density function. Thus it provides an alternative route to analytical results compared with working directly with probability density functions or cumulative distribution functions. There are particularly simple results for the characteristic functions of distributions defined by the weighted sums of random variables.

The gradient theorem, also known as the fundamental theorem of calculus for line integrals, says that a line integral through a gradient field can be evaluated by evaluating the original scalar field at the endpoints of the curve. The theorem is a generalization of the second fundamental theorem of calculus to any curve in a plane or space rather than just the real line.

The Kutta–Joukowski theorem is a fundamental theorem in aerodynamics used for the calculation of lift of an airfoil translating in a uniform fluid at a constant speed so large that the flow seen in the body-fixed frame is steady and unseparated. The theorem relates the lift generated by an airfoil to the speed of the airfoil through the fluid, the density of the fluid and the circulation around the airfoil. The circulation is defined as the line integral around a closed loop enclosing the airfoil of the component of the velocity of the fluid tangent to the loop. It is named after Martin Kutta and Nikolai Zhukovsky who first developed its key ideas in the early 20th century. Kutta–Joukowski theorem is an inviscid theory, but it is a good approximation for real viscous flow in typical aerodynamic applications.

In mathematics, a line integral is an integral where the function to be integrated is evaluated along a curve. The terms path integral, curve integral, and curvilinear integral are also used; contour integral is used as well, although that is typically reserved for line integrals in the complex plane.

In probability theory and directional statistics, a wrapped probability distribution is a continuous probability distribution that describes data points that lie on a unit n-sphere. In one dimension, a wrapped distribution consists of points on the unit circle. If is a random variate in the interval with probability density function (PDF) , then is a circular variable distributed according to the wrapped distribution and is an angular variable in the interval distributed according to the wrapped distribution .

In mathematics, singular integral operators on closed curves arise in problems in analysis, in particular complex analysis and harmonic analysis. The two main singular integral operators, the Hilbert transform and the Cauchy transform, can be defined for any smooth Jordan curve in the complex plane and are related by a simple algebraic formula. In the special case of Fourier series for the unit circle, the operators become the classical Cauchy transform, the orthogonal projection onto Hardy space, and the Hilbert transform a real orthogonal linear complex structure. In general the Cauchy transform is a non-self-adjoint idempotent and the Hilbert transform a non-orthogonal complex structure. The range of the Cauchy transform is the Hardy space of the bounded region enclosed by the Jordan curve. The theory for the original curve can be deduced from that of the unit circle, where, because of rotational symmetry, both operators are classical singular integral operators of convolution type. The Hilbert transform satisfies the jump relations of Plemelj and Sokhotski, which express the original function as the difference between the boundary values of holomorphic functions on the region and its complement. Singular integral operators have been studied on various classes of functions, including Hölder spaces, Lp spaces and Sobolev spaces. In the case of L2 spaces—the case treated in detail below—other operators associated with the closed curve, such as the Szegő projection onto Hardy space and the Neumann–Poincaré operator, can be expressed in terms of the Cauchy transform and its adjoint.

References