Integral equation

Last updated

In mathematics, integral equations are equations in which an unknown function appears under an integral sign. [1] In mathematical notation, integral equations may thus be expressed as being of the form:

Contents

where is an integral operator acting on u. [1] Hence, integral equations may be viewed as the analog to differential equations where instead of the equation involving derivatives, the equation contains integrals. [1] A direct comparison can be seen with the mathematical form of the general integral equation above with the general form of a differential equation which may be expressed as follows:

where may be viewed as a differential operator of order i. [1] Due to this close connection between differential and integral equations, one can often convert between the two. [1] For example, one method of solving a boundary value problem is by converting the differential equation with its boundary conditions into an integral equation and solving the integral equation. [1] In addition, because one can convert between the two, differential equations in physics such as Maxwell's equations often have an analog integral and differential form. [2] See also, for example, Green's function and Fredholm theory.

Classification and overview

Various classification methods for integral equations exist. A few standard classifications include distinctions between linear and nonlinear; homogenous and inhomogeneous; Fredholm and Volterra; first order, second order, and third order; and singular and regular integral equations. [1] These distinctions usually rest on some fundamental property such as the consideration of the linearity of the equation or the homogeneity of the equation. [1] These comments are made concrete through the following definitions and examples:

Linearity

Linear: An integral equation is linear if the unknown function u(x) and its integrals appear linear in the equation. [1] Hence, an example of a linear equation would be: [1]

As a note on naming convention: i) u(x) is called the unknown function, ii) f(x) is called a known function, iii) K(x,t) is a function of two variables and often called the Kernel function, and iv) λ is an unknown factor or parameter, which plays the same role as the eigenvalue in linear algebra. [1] Nonlinear: An integral equation is nonlinear if the unknown function u(x) or any of its integrals appear nonlinear in the equation. [1] Hence, examples of nonlinear equations would be the equation above if we replaced u(t) with , such as:

Certain kinds of nonlinear integral equations have specific names. [3] A selection of such equations are: [3]

More information on the Hammerstein equation and different versions of the Hammerstein equation can be found in the Hammerstein section below.

Location of the unknown equation

First kind: An integral equation is called an integral equation of the first kind if the unknown function appears only under the integral sign. [3] An example would be: . [3]

Second kind: An integral equation is called an integral equation of the second kind if the unknown function also appears outside the integral. [3]

Third kind: An integral equation is called an integral equation of the third kind if it is a linear Integral equation of the following form: [3]

where g(t) vanishes at least once in the interval [a,b] [4] [5] or where g(t) vanishes at a finite number of points in (a,b). [6]

Limits of Integration

Fredholm: An integral equation is called a Fredholm integral equation if both of the limits of integration in all integrals are fixed and constant. [1] An example would be that the integral is taken over a fixed subset of . [3] Hence, the following two examples are Fredholm equations: [1]

Note that we can express integral equations such as those above also using integral operator notation. [7] For example, we can define the Fredholm integral operator as:

Hence, the above Fredholm equation of the second kind may be written compactly as: [7]

Volterra: An integral equation is called a Volterra integral equation if at least one of the limits of integration is a variable. [1] Hence, the integral is taken over a domain varying with the variable of integration. [3] Examples of Volterra equations would be: [1]

As with Fredholm equations, we can again adopt operator notation. Thus, we can define the linear Volterra integral operator , as follows: [3]

where and K(t,s) is called the kernel and must be continuous on the interval . [3] Hence, the Volterra integral equation of the first kind may be written as: [3]

with . In addition, a linear Volterra integral equation of the second kind for an unknown function and a given continuous function on the interval where :

Volterra-Fredholm: In higher dimensions, integral equations such as Fredholm-Volterra integral equations (VFIE) exist. [3] A VFIE has the form:

with and being a closed bounded region in with piecewise smooth boundary. [3] The Fredholm-Volterra Integral Operator is defined as: [3]

Note that while throughout this article, the bounds of the integral are usually written as intervals, this need not be the case. [7] In general, integral equations don't always need to be defined over an interval , but could also be defined over a curve or surface. [7]

Homogeneity

Homogenous: An integral equation is called homogeneous if the known function is identically zero. [1]

Inhomogenous: An integral equation is called inhomogeneous if the known function is nonzero. [1]

Regularity

Regular: An integral equation is called regular if the integrals used are all proper integrals. [7]

Singular or weakly singular: An integral equation is called singular or weakly singular if the integral is an improper integral. [7] This could be either because at least one of the limits of integration is infinite or the kernel becomes unbounded, meaning infinite, on at least one point in the interval or domain over which is being integrated. [1]

Examples include: [1]

These two integral equations are the Fourier transform and the Laplace transform of u(x), respectively, with both being Fredholm equations of the first kind with kernel and , respectively. [1] Another example of a singular integral equation in which the kernel becomes unbounded is: [1]

This equation is a special form of the more general weakly singular Volterra integral equation of the first kind, called Abel's integral equation: [7]

Strongly singular: An integral equation is called strongly singular if the integral is defined by a special regularisation, for example, by the Cauchy principal value. [7]

Integro-differential equations

An Integro-differential equation, as the name suggests, combines differential and integral operators into one equation. [1] There are many version including the Volterra integro-differential equation and delay type equations as defined below. [3] For example, using the Volterra operator as defined above, the Volterra integro-differential equation may be written as: [3]

For delay problems, we can define the delay integral operator as: [3]

where the delay integro-differential equation may be expressed as: [3]

Volterra integral equations

Uniqueness and existence theorems in 1D

The solution to a linear Volterra integral equation of the first kind, given by the equation:

can be described by the following uniqueness and existence theorem. [3] Recall that the Volterra integral operator , can be defined as follows: [3]

where and K(t,s) is called the kernel and must be continuous on the interval . [3]

Theorem  Assume that satisfies and for some Then for any with the integral equation above has a unique solution in .

The solution to a linear Volterra integral equation of the second kind, given by the equation: [3]

can be described by the following uniqueness and existence theorem. [3]

Theorem  Let and let denote the resolvent Kernel associated with . Then, for any , the second-kind Volterra integral equation has a unique solution and this solution is given by: .

Volterra integral equations in

A Volterra Integral equation of the second kind can be expressed as follows: [3]

where , , and . [3] This integral equation has a unique solution given by: [3]

where is the resolvent kernel of K. [3]

Uniqueness and existence theorems of Fredhom-Volterra equations

As defined above, a VFIE has the form:

with and being a closed bounded region in with piecewise smooth boundary. [3] The Fredholm-Volterrra Integral Operator is defined as: [3]

In the case where the Kernel K may be written as , K is called the positive memory kernel. [3] With this in mind, we can now introduce the following theorem: [3]

Theorem  If the linear VFIE given by: with satisfies the following conditions:

  • , and
  • where and

Then the VFIE has a unique solution given by where is called the Resolvent Kernel and is given by the limit of the Neumann series for the Kernel and solves the resolvent equations:

Special Volterra equations

A special type of Volterra equation which is used in various applications is defined as follows: [3]

where , the function g(t) is continuous on the interval , and the Volterra integral operator is given by:

with . [3]

Converting IVP to integral equations

In the following section, we give an example of how to convert an initial value problem (IVP) into an integral equation. There are multiple motivations for doing so, among them being that integral equations can often be more readily solvable and are more suitable for proving existence and uniqueness theorems. [7]

The following example was provided by Wazwaz on pages 1 and 2 in his book. [1] We examine the IVP given by the equation:

and the initial condition:

If we integrate both sides of the equation, we get:

and by the fundamental theorem of calculus, we obtain:

Rearranging the equation above, we get the integral equation:

which is a Volterra integral equation of the form:

where K(x,t) is called the kernel and equal to 2t, and f(x)=1. [1]

Power series solution for integral equations

In many cases, if the Kernel of the integral equation is of the form K(xt) and the Mellin transform of K(t) exists, we can find the solution of the integral equation

in the form of a power series

where

are the Z-transform of the function g(s), and M(n + 1) is the Mellin transform of the Kernel.

Numerical solution

It is worth noting that integral equations often do not have an analytical solution, and must be solved numerically. An example of this is evaluating the electric-field integral equation (EFIE) or magnetic-field integral equation (MFIE) over an arbitrarily shaped object in an electromagnetic scattering problem.

One method to solve numerically requires discretizing variables and replacing integral by a quadrature rule

Then we have a system with n equations and n variables. By solving it we get the value of the n variables

Integral equations as a generalization of eigenvalue equations

Certain homogeneous linear integral equations can be viewed as the continuum limit of eigenvalue equations. Using index notation, an eigenvalue equation can be written as

where M = [Mi,j] is a matrix, v is one of its eigenvectors, and λ is the associated eigenvalue.

Taking the continuum limit, i.e., replacing the discrete indices i and j with continuous variables x and y, yields

where the sum over j has been replaced by an integral over y and the matrix M and the vector v have been replaced by the kernelK(x, y) and the eigenfunction φ(y). (The limits on the integral are fixed, analogously to the limits on the sum over j.) This gives a linear homogeneous Fredholm equation of the second type.

In general, K(x, y) can be a distribution, rather than a function in the strict sense. If the distribution K has support only at the point x = y, then the integral equation reduces to a differential eigenfunction equation.

In general, Volterra and Fredholm integral equations can arise from a single differential equation, depending on which sort of conditions are applied at the boundary of the domain of its solution.

Wiener–Hopf integral equations

Originally, such equations were studied in connection with problems in radiative transfer, and more recently, they have been related to the solution of boundary integral equations for planar problems in which the boundary is only piecewise smooth.

Hammerstein equations

A Hammerstein equation is a nonlinear first-kind Volterra integral equation of the form: [3]

Under certain regularity conditions, the equation is equivalent to the implicit Volterra integral equation of the second-kind: [3]

where:

The equation may however also be expressed in operator form which motivates the definition of the following operator called the nonlinear Volterra-Hammerstein operator: [3]

Here is a smooth function while the kernel K may be continuous, i.e. bounded, or weakly singular. [3] The corresponding second-kind Volterra integral equation called the Volterra-Hammerstein Integral Equation of the second kind, or simply Hammerstein equation for short, can be expressed as: [3]

In certain applications, the nonlinearity of the function G may be treated as being only semi-linear in the form of: [3]

In this case, we the following semi-linear Volterra integral equation: [3]

In this form, we can state an existence and uniqueness theorem for the semi-linear Hammerstein integral equation. [3]

Theorem  Suppose that the semi-linear Hammerstein equation has a unique solution and be a Lipschitz continuous function. Then the solution of this equation may be written in the form: where denotes the unique solution of the linear part of the equation above and is given by: with denoting the resolvent kernel.

We can also write the Hammerstein equation using a different operator called the Niemytzki operator, or substitution operator, defined as follows: [3]

More about this can be found on page 75 of this book. [3]

Applications

Integral equations are important in many applications. Problems in which integral equations are encountered include radiative transfer, and the oscillation of a string, membrane, or axle. Oscillation problems may also be solved as differential equations.

See also

Bibliography

Related Research Articles

In mathematics, an operator is generally a mapping or function that acts on elements of a space to produce elements of another space. There is no general definition of an operator, but the term is often used in place of function when the domain is a set of functions or other structured objects. Also, the domain of an operator is often difficult to characterize explicitly, and may be extended so as to act on related objects. See Operator (physics) for other examples.

<span class="mw-page-title-main">Wave equation</span> Differential equation important in physics

The wave equation is a second-order linear partial differential equation for the description of waves or standing wave fields such as mechanical waves or electromagnetic waves. It arises in fields like acoustics, electromagnetism, and fluid dynamics.

<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 transform</span> Mathematical transform that expresses a function of time as a function of frequency

In physics, engineering and mathematics, the Fourier transform (FT) is an integral transform that takes as input a function and outputs another function that describes the extent to which various frequencies are present in the original function. The output of the transform is a complex-valued function of frequency. The term Fourier transform refers to both this complex-valued function and the mathematical operation. When a distinction needs to be made the Fourier transform is sometimes called the frequency domain representation of the original function. The Fourier transform is analogous to decomposing the sound of a musical chord into the intensities of its constituent pitches.

<span class="mw-page-title-main">Tautochrone curve</span> Concept in geometry

A tautochrone curve or isochrone curve is the curve for which the time taken by an object sliding without friction in uniform gravity to its lowest point is independent of its starting point on the curve. The curve is a cycloid, and the time is equal to π times the square root of the radius over the acceleration of gravity. The tautochrone curve is related to the brachistochrone curve, which is also a cycloid.

<span class="mw-page-title-main">Differential operator</span> Typically linear operator defined in terms of differentiation of functions

In mathematics, a differential operator is an operator defined as a function of the differentiation operator. It is helpful, as a matter of notation first, to consider differentiation as an abstract operation that accepts a function and returns another function.

In mathematics, the Fredholm integral equation is an integral equation whose solution gives rise to Fredholm theory, the study of Fredholm kernels and Fredholm operators. The integral equation was studied by Ivar Fredholm. A useful method to solve such equations, the Adomian decomposition method, is due to George Adomian.

In mathematics, the Poincaré lemma gives a sufficient condition for a closed differential form to be exact. Precisely, it states that every closed p-form on an open ball in Rn is exact for p with 1 ≤ pn. The lemma was introduced by Henri Poincaré in 1886.

In mathematics, the method of characteristics is a technique for solving partial differential equations. Typically, it applies to first-order equations, although more generally the method of characteristics is valid for any hyperbolic partial differential equation. The method is to reduce a partial differential equation to a family of ordinary differential equations along which the solution can be integrated from some initial data given on a suitable hypersurface.

In probability theory and related fields, Malliavin calculus is a set of mathematical techniques and ideas that extend the mathematical field of calculus of variations from deterministic functions to stochastic processes. In particular, it allows the computation of derivatives of random variables. Malliavin calculus is also called the stochastic calculus of variations. P. Malliavin first initiated the calculus on infinite dimensional space. Then, the significant contributors such as S. Kusuoka, D. Stroock, J-M. Bismut, Shinzo Watanabe, I. Shigekawa, and so on finally completed the foundations.

In physics, the Hamilton–Jacobi equation, named after William Rowan Hamilton and Carl Gustav Jacob Jacobi, is an alternative formulation of classical mechanics, equivalent to other formulations such as Newton's laws of motion, Lagrangian mechanics and Hamiltonian mechanics.

A stochastic differential equation (SDE) is a differential equation in which one or more of the terms is a stochastic process, resulting in a solution which is also a stochastic process. SDEs have many applications throughout pure mathematics and are used to model various behaviours of stochastic models such as stock prices, random growth models or physical systems that are subjected to thermal fluctuations.

<span class="mw-page-title-main">Linear time-invariant system</span> Mathematical model which is both linear and time-invariant

In system analysis, among other fields of study, a linear time-invariant (LTI) system is a system that produces an output signal from any input signal subject to the constraints of linearity and time-invariance; these terms are briefly defined below. These properties apply (exactly or approximately) to many important physical systems, in which case the response y(t) of the system to an arbitrary input x(t) can be found directly using convolution: y(t) = (xh)(t) where h(t) is called the system's impulse response and ∗ represents convolution (not to be confused with multiplication). What's more, there are systematic methods for solving any such system (determining h(t)), whereas systems not meeting both properties are generally more difficult (or impossible) to solve analytically. A good example of an LTI system is any electrical circuit consisting of resistors, capacitors, inductors and linear amplifiers.

In differential geometry, a spray is a vector field H on the tangent bundle TM that encodes a quasilinear second order system of ordinary differential equations on the base manifold M. Usually a spray is required to be homogeneous in the sense that its integral curves t→ΦHt(ξ)∈TM obey the rule ΦHt(λξ)=ΦHλt(ξ) in positive re-parameterizations. If this requirement is dropped, H is called a semi-spray.

In mathematics, the Volterra integral equations are a special type of integral equations. They are divided into two groups referred to as the first and the second kind.

In mathematics, Harnack's inequality is an inequality relating the values of a positive harmonic function at two points, introduced by A. Harnack. Harnack's inequality is used to prove Harnack's theorem about the convergence of sequences of harmonic functions. J. Serrin, and J. Moser generalized Harnack's inequality to solutions of elliptic or parabolic partial differential equations. Such results can be used to show the interior regularity of weak solutions.

In mathematics, and more specifically in partial differential equations, Duhamel's principle is a general method for obtaining solutions to inhomogeneous linear evolution equations like the heat equation, wave equation, and vibrating plate equation. It is named after Jean-Marie Duhamel who first applied the principle to the inhomogeneous heat equation that models, for instance, the distribution of heat in a thin plate which is heated from beneath. For linear evolution equations without spatial dependency, such as a harmonic oscillator, Duhamel's principle reduces to the method of variation of parameters technique for solving linear inhomogeneous ordinary differential equations. It is also an indispensable tool in the study of nonlinear partial differential equations such as the Navier–Stokes equations and nonlinear Schrödinger equation where one treats the nonlinearity as an inhomogeneity.

In mathematics, the spectral theory of ordinary differential equations is the part of spectral theory concerned with the determination of the spectrum and eigenfunction expansion associated with a linear ordinary differential equation. In his dissertation, Hermann Weyl generalized the classical Sturm–Liouville theory on a finite closed interval to second order differential operators with singularities at the endpoints of the interval, possibly semi-infinite or infinite. Unlike the classical case, the spectrum may no longer consist of just a countable set of eigenvalues, but may also contain a continuous part. In this case the eigenfunction expansion involves an integral over the continuous part with respect to a spectral measure, given by the Titchmarsh–Kodaira formula. The theory was put in its final simplified form for singular differential equations of even degree by Kodaira and others, using von Neumann's spectral theorem. It has had important applications in quantum mechanics, operator theory and harmonic analysis on semisimple Lie groups.

Quantum stochastic calculus is a generalization of stochastic calculus to noncommuting variables. The tools provided by quantum stochastic calculus are of great use for modeling the random evolution of systems undergoing measurement, as in quantum trajectories. Just as the Lindblad master equation provides a quantum generalization to the Fokker–Planck equation, quantum stochastic calculus allows for the derivation of quantum stochastic differential equations (QSDE) that are analogous to classical Langevin equations.

The separation principle is one of the fundamental principles of stochastic control theory, which states that the problems of optimal control and state estimation can be decoupled under certain conditions. In its most basic formulation it deals with a linear stochastic system

References

  1. 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 Wazwaz, Abdul-Majid (2005). A First Course in Integral Equation. World Scientific.
  2. admin (2022-09-10). "Maxwell's Equations: Derivation in Integral and Differential form". Ox Science. Retrieved 2022-12-10.
  3. 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 Brunner, Hermann (2004). Collocation Methods for Volterra Integral and Related Functional Differential Equations. Cambridge University Press.
  4. Bart, G. R.; Warnock, R. L. (November 1973). "Linear Integral Equations of the Third Kind". SIAM Journal on Mathematical Analysis. 4 (4): 609–622. doi:10.1137/0504053. ISSN   0036-1410.
  5. Shulaia, D. (2017-12-01). "Integral equations of the third kind for the case of piecewise monotone coefficients". Transactions of A. Razmadze Mathematical Institute. 171 (3): 396–410. doi: 10.1016/j.trmi.2017.05.002 . ISSN   2346-8092.
  6. Sukavanam, N. (1984-05-01). "A Fredholm-type theory for third-kind linear integral equations". Journal of Mathematical Analysis and Applications. 100 (2): 478–485. doi: 10.1016/0022-247X(84)90096-9 . ISSN   0022-247X.
  7. 1 2 3 4 5 6 7 8 9 10 Hackbusch, Wolfgang (1995). Integral Equations Theory and Numerical Treatment. Birkhauser.
  8. "Lecture Notes on Risk Theory" (PDF). 2010.
  9. Sachs, E. W.; Strauss, A. K. (2008-11-01). "Efficient solution of a partial integro-differential equation in finance". Applied Numerical Mathematics. 58 (11): 1687–1703. doi:10.1016/j.apnum.2007.11.002. ISSN   0168-9274.
  10. Feller, Willy (1941). "On the Integral Equation of Renewal Theory". The Annals of Mathematical Statistics. 12 (3): 243–267. ISSN   0003-4851.
  11. Daddi-Moussa-Ider, A.; Vilfan, A.; Golestanian, R. (6 April 2022). "Diffusiophoretic propulsion of an isotropic active colloidal particle near a finite-sized disk embedded in a planar fluid–fluid interface". Journal of Fluid Mechanics. 940: A12. arXiv: 2109.14437 . doi:10.1017/jfm.2022.232.
  12. Daddi-Moussa-Ider, A.; Lisicki, M.; Löwen, H.; Menzel, A. M. (5 February 2020). "Dynamics of a microswimmer–microplatelet composite". Physics of Fluids. 32 (2): 021902. arXiv: 2001.06646 . doi:10.1063/1.5142054.
  13. Donal., Agarwal, Ravi P. O'Regan (2000). Integral and integrodifferential equations : theory, method and applications. Gordon and Breach Science Publishers. ISBN   90-5699-221-X. OCLC   44617552.{{cite book}}: CS1 maint: multiple names: authors list (link)
  14. Burton, T.A. (2005). Volterra Integral and Differential Equations. Elsevier.
  15. "Chapter 7 It Mod 02-14-05 - Ira A. Fulton College of Engineering" (PDF).
  16. Corduneanu, C. (2008). Integral Equations and Applications. Cambridge University Press.
  17. Hochstadt, Harry (1989). Integral Equations. Wiley-Interscience/John Wiley & Sons.
  18. "Integral Equation".
  19. "Integral equation - Encyclopedia of Mathematics". encyclopediaofmath.org. Retrieved 2022-11-14.
  20. Jerri, Abdul J. Introduction to integral equations with applications. ISBN   0-9673301-1-4. OCLC   852490911.
  21. Pipkin, A.C. (1991). A Course on Integral Equations. Springer-Verlag.
  22. Polëiìanin, A.D. (2008). Handbook of Integral Equation. Chapman & Hall/CRC.

Further reading