Product integral

Last updated

A product integral is any product-based counterpart of the usual sum-based integral of calculus. The first product integral ( Type I below) was developed by the mathematician Vito Volterra in 1887 to solve systems of linear differential equations. [1] [2] Other examples of product integrals are the geometric integral ( Type II below), the bigeometric integral ( Type III below), and some other integrals of non-Newtonian calculus. [3] [4] [5]

Contents

Product integrals have found use in areas from epidemiology (the Kaplan–Meier estimator) to stochastic population dynamics using multiplication integrals (multigrals), analysis and quantum mechanics. The geometric integral, together with the geometric derivative, is useful in image analysis [6] and in the study of growth/decay phenomena (e.g., in economic growth, bacterial growth, and radioactive decay). [7] [8] The bigeometric integral, together with the bigeometric derivative, is useful in some applications of fractals, [9] [10] [11] [12] and in the theory of elasticity in economics. [3] [5] [13]

This article adopts the "product" notation for product integration instead of the "integral" (usually modified by a superimposed "times" symbol or letter P) favoured by Volterra and others. An arbitrary classification of types is also adopted to impose some order in the field.

Basic definitions

The classical Riemann integral of a function can be defined by the relation

where the limit is taken over all partitions of the interval whose norms approach zero.

Roughly speaking, product integrals are similar, but take the limit of a product instead of the limit of a sum. They can be thought of as "continuous" versions of "discrete" products.

The most popular product integrals are the following:

Type I: Volterra integral

The type I product integral corresponds to Volterra's original definition. [2] [14] [15] The following relationship exists for scalar functions :

which is not a multiplicative operator. (So the concepts of product integral and multiplicative integral are not the same).

The Volterra product integral is most useful when applied to matrix-valued functions or functions with values in a Banach algebra, where the last equality is no longer true (see the references below).

When applied to scalars belonging to a non-commutative field, to matrixes, and to operators, i.e. to mathematical objects that don't commute, the Volterra integral splits in two definitions [16]

Left Product integral

With the notation of left products (i.e. normal products applied from left)

Right Product Integral

With the notation of right products (i.e. applied from right)

Where is the identity matrix and D is a partition of the interval [a,b] in the Riemann sense, i.e. the limit is over the maximum interval in the partition. Note how in this case time ordering comes evident in the definitions.

For scalar functions, the derivative in the Volterra system is the logarithmic derivative, and so the Volterra system is not a multiplicative calculus and is not a non-Newtonian calculus. [2]

Type II: geometric integral

which is called the geometric integral and is a multiplicative operator.

This definition of the product integral is the continuous analog of the discrete product operator

(with ) and the multiplicative analog to the (normal/standard/additive) integral

(with ):

additive multiplicative
discrete
continuous

It is very useful in stochastics, where the log-likelihood (i.e. the logarithm of a product integral of independent random variables) equals the integral of the logarithm of these (infinitesimally many) random variables:

Type III: bigeometric integral

where r = ln a, and s = ln b.

The type III product integral is called the bigeometric integral and is a multiplicative operator.

Results

Basic results

The following results are for the type II product integral (the geometric integral). Other types produce other results.

The geometric integral (type II above) plays a central role in the geometric calculus, [3] [4] [17] which is a multiplicative calculus. The inverse of the geometric integral, which is the geometric derivative, denoted , is defined using the following relationship:

Thus, the following can be concluded:

The fundamental theorem
Product rule
Quotient rule
Law of large numbers

where X is a random variable with probability distribution F(x).

Compare with the standard law of large numbers:

Lebesgue-type product-integrals

Just like the Lebesgue version of (classical) integrals, one can compute product integrals by approximating them with the product integrals of simple functions. Each type of product integral has a different form for simple functions.

Type I: Volterra integral

Because simple functions generalize step functions, in what follows we will only consider the special case of simple functions that are step functions. This will also make it easier to compare the Lebesgue definition with the Riemann definition.

Given a step function with corresponding partition and a tagged partition

one approximation of the "Riemann definition" of the type I product integral is given by [18]

The (type I) product integral was defined to be, roughly speaking, the limit of these products by Ludwig Schlesinger in a 1931 article.[ which? ]

Another approximation of the "Riemann definition" of the type I product integral is defined as

When is a constant function, the limit of the first type of approximation is equal to the second type of approximation. [19] Notice that in general, for a step function, the value of the second type of approximation doesn't depend on the partition, as long as the partition is a refinement of the partition defining the step function, whereas the value of the first type of approximation does depend on the fineness of the partition, even when it is a refinement of the partition defining the step function.

It turns out that [20] that for any product-integrable function , the limit of the first type of approximation equals the limit of the second type of approximation. Since, for step functions, the value of the second type of approximation doesn't depend on the fineness of the partition for partitions "fine enough", it makes sense to define [21] the "Lebesgue (type I) product integral" of a step function as

where is a tagged partition, and again is the partition corresponding to the step function . (In contrast, the corresponding quantity would not be unambiguously defined using the first type of approximation.)

This generalizes to arbitrary measure spaces readily. If is a measure space with measure , then for any product-integrable simple function (i.e. a conical combination of the indicator functions for some disjoint measurable sets ), its type I product integral is defined to be

since is the value of at any point of . In the special case where , is Lebesgue measure, and all of the measurable sets are intervals, one can verify that this is equal to the definition given above for that special case. Analogous to the theory of Lebesgue (classical) integrals, the Volterra product integral of any product-integrable function can be written as the limit of an increasing sequence of Volterra product integrals of product-integrable simple functions.

Taking logarithms of both sides of the above definition, one gets that for any product-integrable simple function :

where we used the definition of integral for simple functions. Moreover, because continuous functions like can be interchanged with limits, and the product integral of any product-integrable function is equal to the limit of product integrals of simple functions, it follows that the relationship

holds generally for any product-integrable . This clearly generalizes the property mentioned above.

The Volterra product integral is multiplicative as a set function, [22] which can be shown using the above property. More specifically, given a product-integrable function one can define a set function by defining, for every measurable set ,

where denotes the indicator function of . Then for any two disjoint measurable sets one has

This property can be contrasted with measures, which are additive set functions.

However the Volterra product integral is not multiplicative as a functional. Given two product-integrable functions , and a measurable set , it is generally the case that

Type II: geometric integral

If is a measure space with measure , then for any product-integrable simple function (i.e. a conical combination of the indicator functions for some disjoint measurable sets ), its type II product integral is defined to be

This can be seen to generalize the definition given above.

Taking logarithms of both sides, we see that for any product-integrable simple function :

where we have used the definition of the Lebesgue integral for simple functions. This observation, analogous to the one already made above, allows one to entirely reduce the "Lebesgue theory of geometric integrals" to the Lebesgue theory of (classical) integrals. In other words, because continuous functions like and can be interchanged with limits, and the product integral of any product-integrable function is equal to the limit of some increasing sequence of product integrals of simple functions, it follows that the relationship

holds generally for any product-integrable . This generalizes the property of geometric integrals mentioned above.

See also

Related Research Articles

<span class="mw-page-title-main">Feynman diagram</span> Pictorial representation of the behavior of subatomic particles

In theoretical physics, a Feynman diagram is a pictorial representation of the mathematical expressions describing the behavior and interaction of subatomic particles. The scheme is named after American physicist Richard Feynman, who introduced the diagrams in 1948. The interaction of subatomic particles can be complex and difficult to understand; Feynman diagrams give a simple visualization of what would otherwise be an arcane and abstract formula. According to David Kaiser, "Since the middle of the 20th century, theoretical physicists have increasingly turned to this tool to help them undertake critical calculations. Feynman diagrams have revolutionized nearly every aspect of theoretical physics." While the diagrams are applied primarily to quantum field theory, they can also be used in other fields, such as solid-state theory. Frank Wilczek wrote that the calculations that won him the 2004 Nobel Prize in Physics "would have been literally unthinkable without Feynman diagrams, as would [Wilczek's] calculations that established a route to production and observation of the Higgs particle."

<span class="mw-page-title-main">Integral</span> Operation in mathematical calculus

In mathematics, an integral assigns numbers to functions in a way that describes displacement, area, volume, and other concepts that arise by combining infinitesimal data. The process of finding integrals is called integration. Along with differentiation, integration is a fundamental, essential operation of calculus, and serves as a tool to solve problems in mathematics and physics involving the area of an arbitrary shape, the length of a curve, and the volume of a solid, among others.

In mathematical analysis, the Haar measure assigns an "invariant volume" to subsets of locally compact topological groups, consequently defining an integral for functions on those groups.

<span class="mw-page-title-main">Log-normal distribution</span> Probability distribution

In probability theory, a log-normaldistribution is a continuous probability distribution of a random variable whose logarithm is normally distributed. Thus, if the random variable X is log-normally distributed, then Y = ln(X) has a normal distribution. Equivalently, if Y has a normal distribution, then the exponential function of Y, X = exp(Y), has a log-normal distribution. A random variable which is log-normally distributed takes only positive real values. It is a convenient and useful model for measurements in exact and engineering sciences, as well as medicine, economics and other topics.

In calculus, and more generally in mathematical analysis, integration by parts or partial integration is a process that finds the integral of a product of functions in terms of the integral of the product of their derivative and antiderivative. It is frequently used to transform the antiderivative of a product of functions into an antiderivative for which a solution can be more easily found. The rule can be thought of as an integral version of the product rule of differentiation.

In calculus, the power rule is used to differentiate functions of the form , whenever is a real number. Since differentiation is a linear operation on the space of differentiable functions, polynomials can also be differentiated using this rule. The power rule underlies the Taylor series as it relates a power series with a function's derivatives.

In mathematics, a Gaussian function, often simply referred to as a Gaussian, is a function of the base form

In mathematics, Fatou's lemma establishes an inequality relating the Lebesgue integral of the limit inferior of a sequence of functions to the limit inferior of integrals of these functions. The lemma is named after Pierre Fatou.

<span class="mw-page-title-main">Path integral formulation</span> Formulation of quantum mechanics

The path integral formulation is a description in quantum mechanics that generalizes the action principle of classical mechanics. It replaces the classical notion of a single, unique classical trajectory for a system with a sum, or functional integral, over an infinity of quantum-mechanically possible trajectories to compute a quantum amplitude.

In mathematics, particularly in functional analysis, a projection-valued measure (PVM) is a function defined on certain subsets of a fixed set and whose values are self-adjoint projections on a fixed Hilbert space. Projection-valued measures are formally similar to real-valued measures, except that their values are self-adjoint projections rather than real numbers. As in the case of ordinary measures, it is possible to integrate complex-valued functions with respect to a PVM; the result of such an integration is a linear operator on the given Hilbert space.

<span class="mw-page-title-main">Stochastic differential equation</span> Differential equations involving stochastic processes

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 are used to model various phenomena such as stock prices or physical systems subject to thermal fluctuations. Typically, SDEs contain a variable which represents random white noise calculated as the derivative of Brownian motion or the Wiener process. However, other types of random behaviour are possible, such as jump processes. Random differential equations are conjugate to stochastic differential equations.

In statistics and information theory, a maximum entropy probability distribution has entropy that is at least as great as that of all other members of a specified class of probability distributions. According to the principle of maximum entropy, if nothing is known about a distribution except that it belongs to a certain class, then the distribution with the largest entropy should be chosen as the least-informative default. The motivation is twofold: first, maximizing entropy minimizes the amount of prior information built into the distribution; second, many physical systems tend to move towards maximal entropy configurations over time.

von Mises distribution Probability distribution on the circle

In probability theory and directional statistics, the von Mises distribution is a continuous probability distribution on the circle. It is a close approximation to the wrapped normal distribution, which is the circular analogue of the normal distribution. A freely diffusing angle on a circle is a wrapped normally distributed random variable with an unwrapped variance that grows linearly in time. On the other hand, the von Mises distribution is the stationary distribution of a drift and diffusion process on the circle in a harmonic potential, i.e. with a preferred orientation. The von Mises distribution is the maximum entropy distribution for circular data when the real and imaginary parts of the first circular moment are specified. The von Mises distribution is a special case of the von Mises–Fisher distribution on the N-dimensional sphere.

In physics, Fujikawa's method is a way of deriving the chiral anomaly in quantum field theory. It uses the correspondence between functional determinants and the partition function, effectively making use of the Atiyah–Singer index theorem.

The concept of an abstract Wiener space is a mathematical construction developed by Leonard Gross to understand the structure of Gaussian measures on infinite-dimensional spaces. The construction emphasizes the fundamental role played by the Cameron–Martin space. The classical Wiener space is the prototypical example.

In mathematical analysis, and especially in real, harmonic analysis and functional analysis, an Orlicz space is a type of function space which generalizes the Lp spaces. Like the Lp spaces, they are Banach spaces. The spaces are named for Władysław Orlicz, who was the first to define them in 1932.

<span class="mw-page-title-main">Lebesgue integration</span> Method of integration

In mathematics, the integral of a non-negative function of a single variable can be regarded, in the simplest case, as the area between the graph of that function and the x-axis. The Lebesgue integral, named after French mathematician Henri Lebesgue, extends the integral to a larger class of functions. It also extends the domains on which these functions can be defined.

A product distribution is a probability distribution constructed as the distribution of the product of random variables having two other known distributions. Given two statistically independent random variables X and Y, the distribution of the random variable Z that is formed as the product is a product distribution.

In mathematical analysis, Krein's condition provides a necessary and sufficient condition for exponential sums

In mathematics, integrals of inverse functions can be computed by means of a formula that expresses the antiderivatives of the inverse of a continuous and invertible function , in terms of and an antiderivative of . This formula was published in 1905 by Charles-Ange Laisant.

References

  1. V. Volterra, B. Hostinský, Opérations Infinitésimales Linéaires, Gauthier-Villars, Paris (1938).
  2. 1 2 3 A. Slavík, Product integration, its history and applications, ISBN   80-7378-006-2, Matfyzpress, Prague, 2007.
  3. 1 2 3 M. Grossman, R. Katz, Non-Newtonian Calculus, ISBN   0-912938-01-3, Lee Press, 1972.
  4. 1 2 Michael Grossman. The First Nonlinear System of Differential And Integral Calculus, ISBN   0977117006, 1979.
  5. 1 2 Michael Grossman. Bigeometric Calculus: A System with a Scale-Free Derivative, ISBN   0977117030, 1983.
  6. Luc Florack and Hans van Assen."Multiplicative calculus in biomedical image analysis", Journal of Mathematical Imaging and Vision, doi : 10.1007/s10851-011-0275-1, 2011.
  7. Diana Andrada Filip and Cyrille Piatecki. "An overview on non-Newtonian calculus and its potential applications to economics", Applied Mathematics – A Journal of Chinese Universities, Volume 28, China Society for Industrial and Applied Mathematics, Springer, 2014.
  8. Agamirza E. Bashirov, Emine Misirli, Yucel Tandogdu, and Ali Ozyapici."On modelling with multiplicative differential equations", Applied Mathematics – A Journal of Chinese Universities, Volume 26, Number 4, pages 425–428, doi : 10.1007/s11766-011-2767-6, Springer, 2011.
  9. Marek Rybaczuk."Critical growth of fractal patterns in biological systems", Acta of Bioengineering and Biomechanics, Volume 1, Number 1, Wroclaw University of Technology, 1999.
  10. Marek Rybaczuk, Alicja Kedzia and Witold Zielinski (2001) "The concept of physical and fractal dimension II. The differential calculus in dimensional spaces", Chaos, Solitons, & FractalsVolume 12, Issue 13, October 2001, pages 2537–2552.
  11. Aniszewska, Dorota (October 2007). "Multiplicative Runge–Kutta methods". Nonlinear Dynamics . 50 (1–2): 265–272. doi:10.1007/s11071-006-9156-3. S2CID   120404112.
  12. Dorota Aniszewska and Marek Rybaczuk (2005) "Analysis of the multiplicative Lorenz system", Chaos, Solitons & Fractals Volume 25, Issue 1, July 2005, pages 79–90.
  13. Fernando Córdova-Lepe. "The multiplicative derivative as a measure of elasticity in economics", TMAT Revista Latinoamericana de Ciencias e Ingeniería, Volume 2, Number 3, 2006.
  14. Dollard, J. D.; Friedman, C. N. (1979). Product integration with applications to differential equations. Addison Wesley. ISBN   0-201-13509-4.
  15. Gantmacher, F. R. (1959). The Theory of Matrices. Vol. 1 and 2.
  16. Cherednikov, Igor Olegovich; Mertens, Tom; Van der Veken, Frederik (2 December 2019). Wilson Lines in Quantum Field Theory. ISBN   9783110651690.
  17. A. E. Bashirov, E. M. Kurpınar, A. Özyapıcı. Multiplicative calculus and its applications, Journal of Mathematical Analysis and Applications, 2008.
  18. A. Slavík, Product integration, its history and applications, p. 65. Matfyzpress, Prague, 2007. ISBN   80-7378-006-2.
  19. A. Slavík, Product integration, its history and applications, p. 71. Matfyzpress, Prague, 2007. ISBN   80-7378-006-2.
  20. A. Slavík, Product integration, its history and applications, p. 72. Matfyzpress, Prague, 2007. ISBN   80-7378-006-2.
  21. A. Slavík, Product integration, its history and applications, p. 80. Matfyzpress, Prague, 2007. ISBN   80-7378-006-2
  22. Gill, Richard D., Soren Johansen. "A Survey of Product Integration with a View Toward Application in Survival Analysis". The Annals of Statistics 18, no. 4 (December 1990): 1501—555, p. 1503.