Point process

Last updated

In statistics and probability theory, a point process or point field is a collection of mathematical points randomly located on a mathematical space such as the real line or Euclidean space. [1] [2] Point processes can be used for spatial data analysis, [3] [4] which is of interest in such diverse disciplines as forestry, plant ecology, epidemiology, geography, seismology, materials science, astronomy, telecommunications, computational neuroscience, [5] economics [6] and others.

Contents

There are different mathematical interpretations of a point process, such as a random counting measure or a random set. [7] [8] Some authors regard a point process and stochastic process as two different objects such that a point process is a random object that arises from or is associated with a stochastic process, [9] [10] though it has been remarked that the difference between point processes and stochastic processes is not clear. [10] Others consider a point process as a stochastic process, where the process is indexed by sets of the underlying space [lower-alpha 1] on which it is defined, such as the real line or -dimensional Euclidean space. [13] [14] Other stochastic processes such as renewal and counting processes are studied in the theory of point processes. [15] [10] Sometimes the term "point process" is not preferred, as historically the word "process" denoted an evolution of some system in time, so point process is also called a random point field. [16]

Point processes on the real line form an important special case that is particularly amenable to study, [17] because the points are ordered in a natural way, and the whole point process can be described completely by the (random) intervals between the points. These point processes are frequently used as models for random events in time, such as the arrival of customers in a queue (queueing theory), of impulses in a neuron (computational neuroscience), particles in a Geiger counter, location of radio stations in a telecommunication network [18] or of searches on the world-wide web.

General point process theory

In mathematics, a point process is a random element whose values are "point patterns" on a set S. While in the exact mathematical definition a point pattern is specified as a locally finite counting measure, it is sufficient for more applied purposes to think of a point pattern as a countable subset of S that has no limit points.[ clarification needed ]

Definition

To define general point processes, we start with a probability space , and a measurable space where is a locally compact second countable Hausdorff space and is its Borel σ-algebra. Consider now an integer-valued locally finite kernel from into , that is, a mapping such that:

  1. For every , is a locally finite measure on .[ clarification needed ]
  2. For every , is a random variable over .

This kernel defines a random measure in the following way. We would like to think of as defining a mapping which maps to a measure (namely, ), where is the set of all locally finite measures on . Now, to make this mapping measurable, we need to define a -field over . This -field is constructed as the minimal algebra so that all evaluation maps of the form , where is relatively compact, are measurable. Equipped with this -field, then is a random element, where for every , is a locally finite measure over .

Now, by a point process on we simply mean an integer-valued random measure (or equivalently, integer-valued kernel) constructed as above. The most common example for the state space S is the Euclidean space Rn or a subset thereof, where a particularly interesting special case is given by the real half-line [0,∞). However, point processes are not limited to these examples and may among other things also be used if the points are themselves compact subsets of Rn, in which case ξ is usually referred to as a particle process.

Despite the name point process since S might not be a subset of the real line, as it might suggest that ξ is a stochastic process.

Representation

Every instance (or event) of a point process ξ can be represented as

where denotes the Dirac measure, n is an integer-valued random variable and are random elements of S. If 's are almost surely distinct (or equivalently, almost surely for all ), then the point process is known as simple .

Another different but useful representation of an event (an event in the event space, i.e. a series of points) is the counting notation, where each instance is represented as an function, a continuous function which takes integer values: :

which is the number of events in the observation interval . It is sometimes denoted by , and or mean .

Expectation measure

The expectation measure (also known as mean measure) of a point process ξ is a measure on S that assigns to every Borel subset B of S the expected number of points of ξ in B. That is,

Laplace functional

The Laplace functional of a point process N is a map from the set of all positive valued functions f on the state space of N, to defined as follows:

They play a similar role as the characteristic functions for random variable. One important theorem says that: two point processes have the same law if their Laplace functionals are equal.

Moment measure

The th power of a point process, is defined on the product space as follows :

By monotone class theorem, this uniquely defines the product measure on The expectation is called the th moment measure. The first moment measure is the mean measure.

Let . The joint intensities of a point process w.r.t. the Lebesgue measure are functions such that for any disjoint bounded Borel subsets

Joint intensities do not always exist for point processes. Given that moments of a random variable determine the random variable in many cases, a similar result is to be expected for joint intensities. Indeed, this has been shown in many cases. [2]

Stationarity

A point process is said to be stationary if has the same distribution as for all For a stationary point process, the mean measure for some constant and where stands for the Lebesgue measure. This is called the intensity of the point process. A stationary point process on has almost surely either 0 or an infinite number of points in total. For more on stationary point processes and random measure, refer to Chapter 12 of Daley & Vere-Jones. [2] Stationarity has been defined and studied for point processes in more general spaces than .

Examples of point processes

We shall see some examples of point processes in

Poisson point process

The simplest and most ubiquitous example of a point process is the Poisson point process, which is a spatial generalisation of the Poisson process. A Poisson (counting) process on the line can be characterised by two properties : the number of points (or events) in disjoint intervals are independent and have a Poisson distribution. A Poisson point process can also be defined using these two properties. Namely, we say that a point process is a Poisson point process if the following two conditions hold

1) are independent for disjoint subsets

2) For any bounded subset , has a Poisson distribution with parameter where denotes the Lebesgue measure.

The two conditions can be combined and written as follows : For any disjoint bounded subsets and non-negative integers we have that

The constant is called the intensity of the Poisson point process. Note that the Poisson point process is characterised by the single parameter It is a simple, stationary point process. To be more specific one calls the above point process a homogeneous Poisson point process. An inhomogeneous Poisson process is defined as above but by replacing with where is a non-negative function on

Cox point process

A Cox process (named after Sir David Cox) is a generalisation of the Poisson point process, in that we use random measures in place of . More formally, let be a random measure. A Cox point process driven by the random measure is the point process with the following two properties :

  1. Given , is Poisson distributed with parameter for any bounded subset
  2. For any finite collection of disjoint subsets and conditioned on we have that are independent.

It is easy to see that Poisson point process (homogeneous and inhomogeneous) follow as special cases of Cox point processes. The mean measure of a Cox point process is and thus in the special case of a Poisson point process, it is

For a Cox point process, is called the intensity measure. Further, if has a (random) density (Radon–Nikodym derivative) i.e.,

then is called the intensity field of the Cox point process. Stationarity of the intensity measures or intensity fields imply the stationarity of the corresponding Cox point processes.

There have been many specific classes of Cox point processes that have been studied in detail such as:

By Jensen's inequality, one can verify that Cox point processes satisfy the following inequality: for all bounded Borel subsets ,

where stands for a Poisson point process with intensity measure Thus points are distributed with greater variability in a Cox point process compared to a Poisson point process. This is sometimes called clustering or attractive property of the Cox point process.

Determinantal point processes

An important class of point processes, with applications to physics, random matrix theory, and combinatorics, is that of determinantal point processes. [25]

Hawkes (self-exciting) processes

A Hawkes process , also known as a self-exciting counting process, is a simple point process whose conditional intensity can be expressed as

where is a kernel function which expresses the positive influence of past events on the current value of the intensity process , is a possibly non-stationary function representing the expected, predictable, or deterministic part of the intensity, and is the time of occurrence of the i-th event of the process. [26]

Geometric processes

Given a sequence of non-negative random variables , if they are independent and the cdf of is given by for , where is a positive constant, then is called a geometric process (GP). [27]

The geometric process has several extensions, including the α- series process [28] and the doubly geometric process. [29]

Point processes on the real half-line

Historically the first point processes that were studied had the real half line R+ = [0,∞) as their state space, which in this context is usually interpreted as time. These studies were motivated by the wish to model telecommunication systems, [30] in which the points represented events in time, such as calls to a telephone exchange.

Point processes on R+ are typically described by giving the sequence of their (random) inter-event times (T1, T2, ...), from which the actual sequence (X1, X2, ...) of event times can be obtained as

If the inter-event times are independent and identically distributed, the point process obtained is called a renewal process.

Intensity of a point process

The intensityλ(t | Ht) of a point process on the real half-line with respect to a filtration Ht is defined as

Ht can denote the history of event-point times preceding time t but can also correspond to other filtrations (for example in the case of a Cox process).

In the -notation, this can be written in a more compact form:

The compensator of a point process, also known as the dual-predictable projection, is the integrated conditional intensity function defined by

Papangelou intensity function

The Papangelou intensity function of a point process in the -dimensional Euclidean space is defined as

where is the ball centered at of a radius , and denotes the information of the point process outside .

Likelihood function

The logarithmic likelihood of a parameterized simple point process conditional upon some observed data is written as

[31]

Point processes in spatial statistics

The analysis of point pattern data in a compact subset S of Rn is a major object of study within spatial statistics. Such data appear in a broad range of disciplines, [32] amongst which are

The need to use point processes to model these kinds of data lies in their inherent spatial structure. Accordingly, a first question of interest is often whether the given data exhibit complete spatial randomness (i.e. are a realization of a spatial Poisson process) as opposed to exhibiting either spatial aggregation or spatial inhibition.

In contrast, many datasets considered in classical multivariate statistics consist of independently generated datapoints that may be governed by one or several covariates (typically non-spatial).

Apart from the applications in spatial statistics, point processes are one of the fundamental objects in stochastic geometry. Research has also focussed extensively on various models built on point processes such as Voronoi tessellations, random geometric graphs, and Boolean models.

See also

Notes

  1. In the context of point processes, the term "state space" can mean the space on which the point process is defined such as the real line, [11] [12] which corresponds to the index set in stochastic process terminology.

Related Research Articles

<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 converts a function into a form that describes the frequencies 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.

In physics, a Langevin equation is a stochastic differential equation describing how a system evolves when subjected to a combination of deterministic and fluctuating ("random") forces. The dependent variables in a Langevin equation typically are collective (macroscopic) variables changing only slowly in comparison to the other (microscopic) variables of the system. The fast (microscopic) variables are responsible for the stochastic nature of the Langevin equation. One application is to Brownian motion, which models the fluctuating motion of a small particle in a fluid.

A directional derivative is a concept in multivariable calculus that measures the rate at which a function changes in a particular direction at a given point.

In mathematics, the Poisson summation formula is an equation that relates the Fourier series coefficients of the periodic summation of a function to values of the function's continuous Fourier transform. Consequently, the periodic summation of a function is completely defined by discrete samples of the original function's Fourier transform. And conversely, the periodic summation of a function's Fourier transform is completely defined by discrete samples of the original function. The Poisson summation formula was discovered by Siméon Denis Poisson and is sometimes called Poisson resummation.

In functional analysis, a branch of mathematics, the Borel functional calculus is a functional calculus, which has particularly broad scope. Thus for instance if T is an operator, applying the squaring function ss2 to T yields the operator T2. Using the functional calculus for larger classes of functions, we can for example define rigorously the "square root" of the (negative) Laplacian operator −Δ or the exponential

In differential geometry, a field in mathematics, a Poisson manifold is a smooth manifold endowed with a Poisson structure. The notion of Poisson manifold generalises that of symplectic manifold, which in turn generalises the phase space from Hamiltonian mechanics.

<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 probability theory and mathematical physics, a random matrix is a matrix-valued random variable—that is, a matrix in which some or all elements are random variables. Many important properties of physical systems can be represented mathematically as matrix problems. For example, the thermal conductivity of a lattice can be computed from the dynamical matrix of the particle-particle interactions within the lattice.

In probability theory, a Cox process, also known as a doubly stochastic Poisson process is a point process which is a generalization of a Poisson process where the intensity that varies across the underlying mathematical space is itself a stochastic process. The process is named after the statistician David Cox, who first published the model in 1955.

An -superprocess, , within mathematics probability theory is a stochastic process on that is usually constructed as a special limit of near-critical branching diffusions.

In actuarial science and applied probability, ruin theory uses mathematical models to describe an insurer's vulnerability to insolvency/ruin. In such models key quantities of interest are the probability of ruin, distribution of surplus immediately prior to ruin and deficit at time of ruin.

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.

In mathematical physics, the Berezin integral, named after Felix Berezin,, is a way to define integration for functions of Grassmann variables. It is not an integral in the Lebesgue sense; the word "integral" is used because the Berezin integral has properties analogous to the Lebesgue integral and because it extends the path integral in physics, where it is used as a sum over histories for fermions.

In probability theory, a Markov kernel is a map that in the general theory of Markov processes plays the role that the transition matrix does in the theory of Markov processes with a finite state space.

In mathematics, particularly linear algebra, the Schur–Horn theorem, named after Issai Schur and Alfred Horn, characterizes the diagonal of a Hermitian matrix with given eigenvalues. It has inspired investigations and substantial generalizations in the setting of symplectic geometry. A few important generalizations are Kostant's convexity theorem, Atiyah–Guillemin–Sternberg convexity theorem, Kirwan convexity theorem.

<span class="mw-page-title-main">Lie point symmetry</span>

Lie point symmetry is a concept in advanced mathematics. Towards the end of the nineteenth century, Sophus Lie introduced the notion of Lie group in order to study the solutions of ordinary differential equations (ODEs). He showed the following main property: the order of an ordinary differential equation can be reduced by one if it is invariant under one-parameter Lie group of point transformations. This observation unified and extended the available integration techniques. Lie devoted the remainder of his mathematical career to developing these continuous groups that have now an impact on many areas of mathematically based sciences. The applications of Lie groups to differential systems were mainly established by Lie and Emmy Noether, and then advocated by Élie Cartan.

<span class="mw-page-title-main">Poisson point process</span> Type of random mathematical object

In probability, statistics and related fields, a Poisson point process is a type of random mathematical object that consists of points randomly located on a mathematical space with the essential feature that the points occur independently of one another. The Poisson point process is also called a Poisson random measure, Poisson random point field or Poisson point field. When the process is defined on the real line, it is often called simply the Poisson process.

In mathematics, the Poisson boundary is a measure space associated to a random walk. It is an object designed to encode the asymptotic behaviour of the random walk, i.e. how trajectories diverge when the number of steps goes to infinity. Despite being called a boundary it is in general a purely measure-theoretical object and not a boundary in the topological sense. However, in the case where the random walk is on a topological space the Poisson boundary can be related to the Martin boundary, which is an analytic construction yielding a genuine topological boundary. Both boundaries are related to harmonic functions on the space via generalisations of the Poisson formula.

A mixed binomial process is a special point process in probability theory. They naturally arise from restrictions of (mixed) Poisson processes bounded intervals.

In the theory of stochastic processes, a ν-transform is an operation that transforms a measure or a point process into a different point process. Intuitively the ν-transform randomly relocates the points of the point process, with the type of relocation being dependent on the position of each point.

References

  1. Kallenberg, O. (1986). Random Measures, 4th edition. Academic Press, New York, London; Akademie-Verlag, Berlin. ISBN   0-12-394960-2, MR 854102.
  2. 1 2 3 Daley, D.J, Vere-Jones, D. (1988). An Introduction to the Theory of Point Processes. Springer, New York. ISBN   0-387-96666-8, MR 950166.
  3. Diggle, P. (2003). Statistical Analysis of Spatial Point Patterns, 2nd edition. Arnold, London. ISBN   0-340-74070-1.
  4. Baddeley, A. (2006). Spatial point processes and their applications. In A. Baddeley, I. Bárány, R. Schneider, and W. Weil, editors, Stochastic Geometry: Lectures given at the C.I.M.E. Summer School held in Martina Franca, Italy, September 13–18, 2004, Lecture Notes in Mathematics 1892, Springer. ISBN   3-540-38174-0, pp. 1–75
  5. Brown E. N., Kass R. E., Mitra P. P. (2004). "Multiple neural spike train data analysis: state-of-the-art and future challenges". Nature Neuroscience. 7 (5): 456–461. doi:10.1038/nn1228. PMID   15114358. S2CID   562815.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  6. Engle Robert F., Lunde Asger (2003). "Trades and Quotes: A Bivariate Point Process" (PDF). Journal of Financial Econometrics. 1 (2): 159–188. doi: 10.1093/jjfinec/nbg011 .
  7. Sung Nok Chiu; Dietrich Stoyan; Wilfrid S. Kendall; Joseph Mecke (27 June 2013). Stochastic Geometry and Its Applications. John Wiley & Sons. p. 108. ISBN   978-1-118-65825-3.
  8. Martin Haenggi (2013). Stochastic Geometry for Wireless Networks. Cambridge University Press. p. 10. ISBN   978-1-107-01469-5.
  9. D.J. Daley; D. Vere-Jones (10 April 2006). An Introduction to the Theory of Point Processes: Volume I: Elementary Theory and Methods. Springer Science & Business Media. p. 194. ISBN   978-0-387-21564-8.
  10. 1 2 3 Cox, D. R.; Isham, Valerie (1980). Point Processes. CRC Press. p. 3. ISBN   978-0-412-21910-8.
  11. J. F. C. Kingman (17 December 1992). Poisson Processes. Clarendon Press. p. 8. ISBN   978-0-19-159124-2.
  12. Jesper Moller; Rasmus Plenge Waagepetersen (25 September 2003). Statistical Inference and Simulation for Spatial Point Processes. CRC Press. p. 7. ISBN   978-0-203-49693-0.
  13. Samuel Karlin; Howard E. Taylor (2 December 2012). A First Course in Stochastic Processes. Academic Press. p. 31. ISBN   978-0-08-057041-9.
  14. Volker Schmidt (24 October 2014). Stochastic Geometry, Spatial Statistics and Random Fields: Models and Algorithms. Springer. p. 99. ISBN   978-3-319-10064-7.
  15. D.J. Daley; D. Vere-Jones (10 April 2006). An Introduction to the Theory of Point Processes: Volume I: Elementary Theory and Methods. Springer Science & Business Media. ISBN   978-0-387-21564-8.
  16. Sung Nok Chiu; Dietrich Stoyan; Wilfrid S. Kendall; Joseph Mecke (27 June 2013). Stochastic Geometry and Its Applications. John Wiley & Sons. p. 109. ISBN   978-1-118-65825-3.
  17. Last, G., Brandt, A. (1995).Marked point processes on the real line: The dynamic approach. Probability and its Applications. Springer, New York. ISBN   0-387-94547-4, MR 1353912
  18. Gilbert E.N. (1961). "Random plane networks". Journal of the Society for Industrial and Applied Mathematics. 9 (4): 533–543. doi:10.1137/0109045.
  19. Moller, J.; Syversveen, A. R.; Waagepetersen, R. P. (1998). "Log Gaussian Cox Processes". Scandinavian Journal of Statistics. 25 (3): 451. CiteSeerX   10.1.1.71.6732 . doi:10.1111/1467-9469.00115. S2CID   120543073.
  20. Moller, J. (2003) Shot noise Cox processes, Adv. Appl. Prob., 35.[ page needed ]
  21. Moller, J. and Torrisi, G.L. (2005) "Generalised Shot noise Cox processes", Adv. Appl. Prob., 37.
  22. Hellmund, G., Prokesova, M. and Vedel Jensen, E.B. (2008) "Lévy-based Cox point processes", Adv. Appl. Prob., 40. [ page needed ]
  23. Mccullagh,P. and Moller, J. (2006) "The permanental processes", Adv. Appl. Prob., 38.[ page needed ]
  24. Adams, R. P., Murray, I. MacKay, D. J. C. (2009) "Tractable inference in Poisson processes with Gaussian process intensities", Proceedings of the 26th International Conference on Machine Learning doi : 10.1145/1553374.1553376
  25. Hough, J. B., Krishnapur, M., Peres, Y., and Virág, B., Zeros of Gaussian analytic functions and determinantal point processes. University Lecture Series, 51. American Mathematical Society, Providence, RI, 2009.
  26. Patrick J. Laub, Young Lee, Thomas Taimre, The Elements of Hawkes Processes, Springer, 2022.
  27. Lin, Ye (Lam Yeh) (1988). "Geometric processes and replacement problem". Acta Mathematicae Applicatae Sinica. 4 (4): 366–377. doi:10.1007/BF02007241. S2CID   123338120.
  28. Braun, W. John; Li, Wei; Zhao, Yiqiang Q. (2005). "Properties of the geometric and related processes". Naval Research Logistics. 52 (7): 607–616. CiteSeerX   10.1.1.113.9550 . doi:10.1002/nav.20099. S2CID   7745023.
  29. Wu, Shaomin (2018). "Doubly geometric processes and applications" (PDF). Journal of the Operational Research Society. 69: 66–77. doi:10.1057/s41274-017-0217-4. S2CID   51889022.
  30. Palm, C. (1943). Intensitätsschwankungen im Fernsprechverkehr (German). Ericsson Technics no. 44, (1943). MR 11402
  31. Rubin, I. (Sep 1972). "Regular point processes and their detection". IEEE Transactions on Information Theory. 18 (5): 547–557. doi:10.1109/tit.1972.1054897.
  32. Baddeley, A., Gregori, P., Mateu, J., Stoica, R., and Stoyan, D., editors (2006). Case Studies in Spatial Point Pattern Modelling, Lecture Notes in Statistics No. 185. Springer, New York. ISBN   0-387-28311-0.