Markov chain

Source: Wikipedia, the free encyclopedia.
A diagram representing a two-state Markov process. The numbers are the probability of changing from one state to another state.

A Markov chain or Markov process is a

continuous-time process is called a continuous-time Markov chain (CTMC). It is named after the Russian mathematician Andrey Markov
.

Markov chains have many applications as statistical models of real-world processes,[1][4][5][6] such as studying cruise control systems in motor vehicles, queues or lines of customers arriving at an airport, currency exchange rates and animal population dynamics.[7]

Markov processes are the basis for general stochastic simulation methods known as Markov chain Monte Carlo, which are used for simulating sampling from complex probability distributions, and have found application in Bayesian statistics, thermodynamics, statistical mechanics, physics, chemistry, economics, finance, signal processing, information theory and speech processing.[7][8][9]

The adjectives Markovian and Markov are used to describe something that is related to a Markov process.[1][10][11]

Principles

Russian mathematician Andrey Markov

Definition

A Markov process is a stochastic process that satisfies the Markov property[1] (sometimes characterized as "memorylessness"). In simpler terms, it is a process for which predictions can be made regarding future outcomes based solely on its present state and—most importantly—such predictions are just as good as the ones that could be made knowing the process's full history.[12] In other words, conditional on the present state of the system, its future and past states are independent.

A Markov chain is a type of Markov process that has either a discrete

state space or a discrete index set (often representing time), but the precise definition of a Markov chain varies.[13] For example, it is common to define a Markov chain as a Markov process in either discrete or continuous time with a countable state space (thus regardless of the nature of time),[14][15][16][17] but it is also common to define a Markov chain as having discrete time in either countable or continuous state space (thus regardless of the state space).[13]

Types of Markov chains

The system's

state space
and time parameter index need to be specified. The following table gives an overview of the different instances of Markov processes for different levels of state space generality and for discrete time v. continuous time:

Countable state space Continuous or general state space
Discrete-time (discrete-time) Markov chain on a countable or finite state space Markov chain on a measurable state space (for example, Harris chain)
Continuous-time Continuous-time Markov process or Markov jump process Any continuous stochastic process with the Markov property (for example, the Wiener process)

Note that there is no definitive agreement in the literature on the use of some of the terms that signify special cases of Markov processes. Usually the term "Markov chain" is reserved for a process with a discrete set of times, that is, a discrete-time Markov chain (DTMC),[1][18] but a few authors use the term "Markov process" to refer to a continuous-time Markov chain (CTMC) without explicit mention.[19][20][21] In addition, there are other extensions of Markov processes that are referred to as such but do not necessarily fall within any of these four categories (see Markov model). Moreover, the time index need not necessarily be real-valued; like with the state space, there are conceivable processes that move through index sets with other mathematical constructs. Notice that the general state space continuous-time Markov chain is general to such a degree that it has no designated term.

While the time parameter is usually discrete, the

state space of a Markov chain does not have any generally agreed-on restrictions: the term may refer to a process on an arbitrary state space.[22] However, many applications of Markov chains employ finite or countably infinite state spaces, which have a more straightforward statistical analysis. Besides time-index and state-space parameters, there are many other variations, extensions and generalizations (see Variations
). For simplicity, most of this article concentrates on the discrete-time, discrete state-space case, unless mentioned otherwise.

Transitions

The changes of state of the system are called transitions.[1] The probabilities associated with various state changes are called transition probabilities. The process is characterized by a state space, a transition matrix describing the probabilities of particular transitions, and an initial state (or initial distribution) across the state space. By convention, we assume all possible states and transitions have been included in the definition of the process, so there is always a next state, and the process does not terminate.

A discrete-time random process involves a system which is in a certain state at each step, with the state changing randomly between steps.

natural numbers, and the random process is a mapping of these to states.[23] The Markov property states that the conditional probability distribution
for the system at the next step (and in fact at all future steps) depends only on the current state of the system, and not additionally on the state of the system at previous steps.

Since the system changes randomly, it is generally impossible to predict with certainty the state of a Markov chain at a given point in the future.[23] However, the statistical properties of the system's future can be predicted.[23] In many applications, it is these statistical properties that are important.

History

weak law of large numbers to hold.[1][31] In his first paper on Markov chains, published in 1906, Markov showed that under certain conditions the average outcomes of the Markov chain would converge to a fixed vector of values, so proving a weak law of large numbers without the independence assumption,[1][25][26][27] which had been commonly regarded as a requirement for such mathematical laws to hold.[27] Markov later used Markov chains to study the distribution of vowels in Eugene Onegin, written by Alexander Pushkin, and proved a central limit theorem for such chains.[1][25]

In 1912

Maurice Fréchet became interested in Markov chains, eventually resulting in him publishing in 1938 a detailed study on Markov chains.[25][33]

Andrey Kolmogorov developed in a 1931 paper a large part of the early theory of continuous-time Markov processes.[34][35] Kolmogorov was partly inspired by Louis Bachelier's 1900 work on fluctuations in the stock market as well as Norbert Wiener's work on Einstein's model of Brownian movement.[34][36] He introduced and studied a particular set of Markov processes known as diffusion processes, where he derived a set of differential equations describing the processes.[34][37] Independent of Kolmogorov's work, Sydney Chapman derived in a 1928 paper an equation, now called the Chapman–Kolmogorov equation, in a less mathematically rigorous way than Kolmogorov, while studying Brownian movement.[38] The differential equations are now called the Kolmogorov equations[39] or the Kolmogorov–Chapman equations.[40] Other mathematicians who contributed significantly to the foundations of Markov processes include William Feller, starting in 1930s, and then later Eugene Dynkin, starting in the 1950s.[35]

Examples

  • Poisson process,[28] which are considered the most important and central stochastic processes in the theory of stochastic processes.[46][47][48] These two processes are Markov processes in continuous time, while random walks on the integers and the gambler's ruin problem are examples of Markov processes in discrete time.[41][42]
  • A famous Markov chain is the so-called "drunkard's walk", a random walk on the number line where, at each step, the position may change by +1 or −1 with equal probability. From any position there are two possible transitions, to the next or previous integer. The transition probabilities depend only on the current position, not on the manner in which the position was reached. For example, the transition probabilities from 5 to 4 and 5 to 6 are both 0.5, and all other transition probabilities from 5 are 0. These probabilities are independent of whether the system was previously in 4 or 6.
  • A series of independent states (for example, a series of coin flips) satisfies the formal definition of a Markov chain. However, the theory is usually applied only when the probability distribution of the next state depends on the current one.

A non-Markov example

Suppose that there is a coin purse containing five quarters (each worth 25¢), five dimes (each worth 10¢), and five nickels (each worth 5¢), and one by one, coins are randomly drawn from the purse and are set on a table. If represents the total value of the coins set on the table after n draws, with , then the sequence is not a Markov process.

To see why this is the case, suppose that in the first six draws, all five nickels and a quarter are drawn. Thus . If we know not just , but the earlier values as well, then we can determine which coins have been drawn, and we know that the next coin will not be a nickel; so we can determine that with probability 1. But if we do not know the earlier values, then based only on the value we might guess that we had drawn four dimes and two nickels, in which case it would certainly be possible to draw another nickel next. Thus, our guesses about are impacted by our knowledge of values prior to .

However, it is possible to model this scenario as a Markov process. Instead of defining to represent the total value of the coins on the table, we could define to represent the count of the various coin types on the table. For instance, could be defined to represent the state where there is one quarter, zero dimes, and five nickels on the table after 6 one-by-one draws. This new model could be represented by possible states, where each state represents the number of coins of each type (from 0 to 5) that are on the table. (Not all of these states are reachable within 6 draws.) Suppose that the first draw results in state . The probability of achieving now depends on ; for example, the state is not possible. After the second draw, the third draw depends on which coins have so far been drawn, but no longer only on the coins that were drawn for the first state (since probabilistically important information has since been added to the scenario). In this way, the likelihood of the state depends exclusively on the outcome of the state.

Formal definition

Discrete-time Markov chain

A discrete-time Markov chain is a sequence of random variables X1, X2, X3, ... with the Markov property, namely that the probability of moving to the next state depends only on the present state and not on the previous states:

if both conditional probabilities are well defined, that is, if

The possible values of Xi form a countable set S called the state space of the chain.

Variations

  • Time-homogeneous Markov chains are processes where
    for all n. The probability of the transition is independent of n.
  • Stationary Markov chains are processes where
    for all n and k. Every stationary chain can be proved to be time-homogeneous by Bayes' rule.
    A necessary and sufficient condition for a time-homogeneous Markov chain to be stationary is that the distribution of is a stationary distribution of the Markov chain.
  • A Markov chain with memory (or a Markov chain of order m) where m is finite, is a process satisfying
    In other words, the future state depends on the past m states. It is possible to construct a chain from which has the 'classical' Markov property by taking as state space the ordered m-tuples of X values, i.e., .

Continuous-time Markov chain

A continuous-time Markov chain (Xt)t ≥ 0 is defined by a finite or countable state space S, a

transition rate matrix
Q with dimensions equal to that of the state space and initial probability distribution defined on the state space. For i ≠ j, the elements qij are non-negative and describe the rate of the process transitions from state i to state j. The elements qii are chosen such that each row of the transition rate matrix sums to zero, while the row-sums of a probability transition matrix in a (discrete) Markov chain are all equal to one.

There are three equivalent definitions of the process.[49]

Infinitesimal definition

The continuous time Markov chain is characterized by the transition rates, the derivatives with respect to time of the transition probabilities between states i and j.

Let be the random variable describing the state of the process at time t, and assume the process is in a state i at time t. Then, knowing , is independent of previous values , and as h → 0 for all j and for all t,

where is the
little-o notation
. The can be seen as measuring how quickly the transition from i to j happens.

Jump chain/holding time definition

Define a discrete-time Markov chain Yn to describe the nth jump of the process and variables S1, S2, S3, ... to describe holding times in each of the states where Si follows the exponential distribution with rate parameter −qYiYi.

Transition probability definition

For any value n = 0, 1, 2, 3, ... and times indexed up to this value of n: t0, t1, t2, ... and all states recorded at these times i0, i1, i2, i3, ... it holds that

where pij is the solution of the

first-order differential equation
)

with initial condition P(0) is the identity matrix.

Finite state space

If the state space is finite, the transition probability distribution can be represented by a matrix, called the transition matrix, with the (i, j)th element of P equal to

Since each row of P sums to one and all elements are non-negative, P is a

right stochastic matrix
.

Stationary distribution relation to eigenvectors and simplices

A stationary distribution π is a (row) vector, whose entries are non-negative and sum to 1, is unchanged by the operation of transition matrix P on it and so is defined by

By comparing this definition with that of an

eigenvector
we see that the two concepts are related and that

is a normalized () multiple of a left eigenvector e of the transition matrix P with an

eigenvalue
of 1. If there is more than one unit eigenvector then a weighted sum of the corresponding stationary states is also a stationary state. But for a Markov chain one is usually more interested in a stationary state that is the limit of the sequence of distributions for some initial distribution.

The values of a stationary distribution are associated with the state space of P and its eigenvectors have their relative proportions preserved. Since the components of π are positive and the constraint that their sum is unity can be rewritten as we see that the

simplex
.

Time-homogeneous Markov chain with a finite state space

If the Markov chain is time-homogeneous, then the transition matrix P is the same after each step, so the k-step transition probability can be computed as the k-th power of the transition matrix, Pk.

If the Markov chain is irreducible and aperiodic, then there is a unique stationary distribution π.[50] Additionally, in this case Pk converges to a rank-one matrix in which each row is the stationary distribution π:

where 1 is the column vector with all entries equal to 1. This is stated by the Perron–Frobenius theorem. If, by whatever means, is found, then the stationary distribution of the Markov chain in question can be easily determined for any starting distribution, as will be explained below.

For some stochastic matrices P, the limit does not exist while the stationary distribution does, as shown by this example:

(This example illustrates a periodic Markov chain.)

Because there are a number of different special cases to consider, the process of finding this limit if it exists can be a lengthy task. However, there are many techniques that can assist in finding this limit. Let P be an n×n matrix, and define

It is always true that

Subtracting Q from both sides and factoring then yields

where In is the identity matrix of size n, and 0n,n is the zero matrix of size n×n. Multiplying together stochastic matrices always yields another stochastic matrix, so Q must be a stochastic matrix (see the definition above). It is sometimes sufficient to use the matrix equation above and the fact that Q is a stochastic matrix to solve for Q. Including the fact that the sum of each the rows in P is 1, there are n+1 equations for determining n unknowns, so it is computationally easier if on the one hand one selects one row in Q and substitutes each of its elements by one, and on the other one substitutes the corresponding element (the one in the same column) in the vector 0, and next left-multiplies this latter vector by the inverse of transformed former matrix to find Q.

Here is one method for doing so: first, define the function f(A) to return the matrix A with its right-most column replaced with all 1's. If [f(PIn)]−1 exists then[51][50]

Explain: The original matrix equation is equivalent to a system of n×n linear equations in n×n variables. And there are n more linear equations from the fact that Q is a right stochastic matrix whose each row sums to 1. So it needs any n×n independent linear equations of the (n×n+n) equations to solve for the n×n variables. In this example, the n equations from “Q multiplied by the right-most column of (P-In)” have been replaced by the n stochastic ones.

One thing to notice is that if P has an element Pi,i on its main diagonal that is equal to 1 and the ith row or column is otherwise filled with 0's, then that row or column will remain unchanged in all of the subsequent powers Pk. Hence, the ith row or column of Q will have the 1 and the 0's in the same positions as in P.

Convergence speed to the stationary distribution

As stated earlier, from the equation (if exists) the stationary (or steady state) distribution π is a left eigenvector of row stochastic matrix P. Then assuming that P is diagonalizable or equivalently that P has n linearly independent eigenvectors, speed of convergence is elaborated as follows. (For non-diagonalizable, that is, defective matrices, one may start with the Jordan normal form of P and proceed with a bit more involved set of arguments in a similar way.[52]

Let U be the matrix of eigenvectors (each normalized to having an L2 norm equal to 1) where each column is a left eigenvector of P and let Σ be the diagonal matrix of left eigenvalues of P, that is, Σ = diag(λ1,λ2,λ3,...,λn). Then by

eigendecomposition

Let the eigenvalues be enumerated such that:

Since P is a row stochastic matrix, its largest left eigenvalue is 1. If there is a unique stationary distribution, then the largest eigenvalue and the corresponding eigenvector is unique too (because there is no other π which solves the stationary distribution equation above). Let ui be the i-th column of U matrix, that is, ui is the left eigenvector of P corresponding to λi. Also let x be a length n row vector that represents a valid probability distribution; since the eigenvectors ui span we can write

If we multiply x with P from right and continue this operation with the results, in the end we get the stationary distribution π. In other words, π = a1 u1xPP...P = xPk as k → ∞. That means

Since π is parallel to u1(normalized by L2 norm) and π(k) is a probability vector, π(k) approaches to a1 u1 = π as k → ∞ with a speed in the order of λ2/λ1 exponentially. This follows because hence λ2/λ1 is the dominant term. The smaller the ratio is, the faster the convergence is.[53] Random noise in the state distribution π can also speed up this convergence to the stationary distribution.[54]

General state space

Harris chains

Many results for Markov chains with finite state space can be generalized to chains with uncountable state space through Harris chains.

The use of Markov chains in Markov chain Monte Carlo methods covers cases where the process follows a continuous state space.

Locally interacting Markov chains

Considering a collection of Markov chains whose evolution takes in account the state of other Markov chains, is related to the notion of locally interacting Markov chains. This corresponds to the situation when the state space has a (Cartesian-) product form. See

stochastic cellular automata
(probabilistic cellular automata). See for instance Interaction of Markov Processes[55] or.[56]

Properties

Two states are said to communicate with each other if both are reachable from one another by a sequence of transitions that have positive probability. This is an equivalence relation which yields a set of communicating classes. A class is closed if the probability of leaving the class is zero. A Markov chain is irreducible if there is one communicating class, the state space.

A state i has period k if k is the greatest common divisor of the number of transitions by which i can be reached, starting from i. That is:

A state i is said to be transient if, starting from i, there is a non-zero probability that the chain will never return to i. It is called recurrent (or persistent) otherwise.[57] For a recurrent state i, the mean hitting time is defined as:

State i is positive recurrent if is finite and null recurrent otherwise. Periodicity, transience, recurrence and positive and null recurrence are class properties — that is, if one state has the property then all states in its communicating class have the property.[58]

A state i is called absorbing if there are no outgoing transitions from the state.

Irreducibility

Since periodicity is a class property, if a Markov chain is irreducible, then all its states have the same period. In particular, if one state is aperiodic, then the whole Markov chain is aperiodic.

If a finite Markov chain is irreducible, then all states are positive recurrent, and it has a unique stationary distribution given by .

Ergodicity

A state i is said to be ergodic if it is aperiodic and positive recurrent. In other words, a state i is ergodic if it is recurrent, has a period of 1, and has finite mean recurrence time.

If all states in an irreducible Markov chain are ergodic, then the chain is said to be ergodic. Equivalently, there exists some integer such that all entries of are positive.

It can be shown that a finite state irreducible Markov chain is ergodic if it has an aperiodic state. More generally, a Markov chain is ergodic if there is a number N such that any state can be reached from any other state in any number of steps less or equal to a number N. In case of a fully connected transition matrix, where all transitions have a non-zero probability, this condition is fulfilled with N = 1.

A Markov chain with more than one state and just one out-going transition per state is either not irreducible or not aperiodic, hence cannot be ergodic.

Terminology

Some authors call any irreducible, positive recurrent Markov chains ergodic, even periodic ones.[59] In fact, merely irreducible Markov chains correspond to ergodic processes, defined according to ergodic theory.[60]

Some authors call a matrix primitive iff there exists some integer such that all entries of are positive.[61] Some authors call it regular.[62]

Index of primitivity

The index of primitivity, or exponent, of a regular matrix, is the smallest such that all entries of are positive. The exponent is purely a graph-theoretic property, since it depends only on whether each entry of is zero or positive, and therefore can be found on a directed graph with as its adjacency matrix.

There are several combinatorial results about the exponent when there are finitely many states. Let be the number of states, then[63]

  • The exponent is . The only case where it is an equality is when the graph of goes like .
  • If has diagonal entries, then its exponent is .
  • If is symmetric, then has positive diagonal entries, which by previous proposition means its exponent is .
  • (Dulmage-Mendelsohn theorem) The exponent is where is the girth of the graph. It can be improved to , where is the diameter of the graph.[64]

Measure-preserving dynamical system

If a Markov chain has a stationary distribution, then it can be converted to a measure-preserving dynamical system: Let the probability space be , where is the set of all states for the Markov chain. Let the sigma-algebra on the probability space be generated by the cylinder sets. Let the probability measure be generated by the stationary distribution, and the Markov chain transition. Let be the shift operator: . Similarly we can construct such a dynamical system with instead.[65]

Since irreducible Markov chains with finite state spaces have a unique stationary distribution, the above construction is unambiguous for irreducible Markov chains.

In ergodic theory, a measure-preserving dynamical system is called "ergodic" iff any measurable subset such that implies or (up to a null set).

The terminology is inconsistent. Given a Markov chain with a stationary distribution that is strictly positive on all states, the Markov chain is irreducible iff its corresponding measure-preserving dynamical system is ergodic.[60]

Markovian representations

In some cases, apparently non-Markovian processes may still have Markovian representations, constructed by expanding the concept of the "current" and "future" states. For example, let X be a non-Markovian process. Then define a process Y, such that each state of Y represents a time-interval of states of X. Mathematically, this takes the form:

If Y has the Markov property, then it is a Markovian representation of X.

An example of a non-Markovian process with a Markovian representation is an autoregressive time series of order greater than one.[66]

Hitting times

The hitting time is the time, starting in a given set of states until the chain arrives in a given state or set of states. The distribution of such a time period has a phase type distribution. The simplest such distribution is that of a single exponentially distributed transition.

Expected hitting times

For a subset of states A ⊆ S, the vector kA of hitting times (where element represents the expected value, starting in state i that the chain enters one of the states in the set A) is the minimal non-negative solution to[67]

Time reversal

For a CTMC Xt, the time-reversed process is defined to be . By Kelly's lemma this process has the same stationary distribution as the forward process.

A chain is said to be reversible if the reversed process is the same as the forward process. Kolmogorov's criterion states that the necessary and sufficient condition for a process to be reversible is that the product of transition rates around a closed loop must be the same in both directions.

Embedded Markov chain

One method of finding the

ergodic continuous-time Markov chain, Q, is by first finding its embedded Markov chain (EMC). Strictly speaking, the EMC is a regular discrete-time Markov chain, sometimes referred to as a jump process. Each element of the one-step transition probability matrix of the EMC, S, is denoted by sij, and represents the conditional probability
of transitioning from state i into state j. These conditional probabilities may be found by

From this, S may be written as

where I is the identity matrix and diag(Q) is the diagonal matrix formed by selecting the main diagonal from the matrix Q and setting all other elements to zero.

To find the stationary probability distribution vector, we must next find such that

with being a row vector, such that all elements in are greater than 0 and = 1. From this, π may be found as

(S may be periodic, even if Q is not. Once π is found, it must be normalized to a unit vector.)

Another discrete-time process that may be derived from a continuous-time Markov chain is a δ-skeleton—the (discrete-time) Markov chain formed by observing X(t) at intervals of δ units of time. The random variables X(0), X(δ), X(2δ), ... give the sequence of states visited by the δ-skeleton.

Special types of Markov chains

Markov model

Markov models are used to model changing systems. There are 4 main types of models, that generalize Markov chains depending on whether every sequential state is observable or not, and whether the system is to be adjusted on the basis of observations made:

System state is fully observable System state is partially observable
System is autonomous Markov chain Hidden Markov model
System is controlled Markov decision process Partially observable Markov decision process

Bernoulli scheme

A Bernoulli scheme is a special case of a Markov chain where the transition probability matrix has identical rows, which means that the next state is independent of even the current state (in addition to being independent of the past states). A Bernoulli scheme with only two possible states is known as a Bernoulli process.

Note, however, by the

stationary stochastic process
is isomorphic to a Bernoulli scheme; the Markov chain is just one such example.

Subshift of finite type

When the Markov matrix is replaced by the

Applications

Research has reported the application and usefulness of Markov chains in a wide range of topics such as physics, chemistry, biology, medicine, music, game theory and sports.

Physics

Markovian systems appear extensively in thermodynamics and statistical mechanics, whenever probabilities are used to represent unknown or unmodelled details of the system, if it can be assumed that the dynamics are time-invariant, and that no relevant history need be considered which is not already included in the state description.[69][70] For example, a thermodynamic state operates under a probability distribution that is difficult or expensive to acquire. Therefore, Markov Chain Monte Carlo method can be used to draw samples randomly from a black-box to approximate the probability distribution of attributes over a range of objects.[70]

The paths, in the path integral formulation of quantum mechanics, are Markov chains.[71]

Markov chains are used in lattice QCD simulations.[72]

Chemistry

Michaelis-Menten kinetics. The enzyme (E) binds a substrate (S) and produces a product (P). Each reaction is a state transition in a Markov chain.

A reaction network is a chemical system involving multiple reactions and chemical species. The simplest stochastic models of such networks treat the system as a continuous time Markov chain with the state being the number of molecules of each species and with reactions modeled as possible transitions of the chain.[73]

Markov chains and continuous-time Markov processes are useful in chemistry when physical systems closely approximate the Markov property. For example, imagine a large number n of molecules in solution in state A, each of which can undergo a chemical reaction to state B with a certain average rate. Perhaps the molecule is an enzyme, and the states refer to how it is folded. The state of any single enzyme follows a Markov chain, and since the molecules are essentially independent of each other, the number of molecules in state A or B at a time is n times the probability a given molecule is in that state.

The classical model of enzyme activity, Michaelis–Menten kinetics, can be viewed as a Markov chain, where at each time step the reaction proceeds in some direction. While Michaelis-Menten is fairly straightforward, far more complicated reaction networks can also be modeled with Markov chains.[74]

An algorithm based on a Markov chain was also used to focus the fragment-based growth of chemicals in silico towards a desired class of compounds such as drugs or natural products.[75] As a molecule is grown, a fragment is selected from the nascent molecule as the "current" state. It is not aware of its past (that is, it is not aware of what is already bonded to it). It then transitions to the next state when a fragment is attached to it. The transition probabilities are trained on databases of authentic classes of compounds.[76]

Also, the growth (and composition) of copolymers may be modeled using Markov chains. Based on the reactivity ratios of the monomers that make up the growing polymer chain, the chain's composition may be calculated (for example, whether monomers tend to add in alternating fashion or in long runs of the same monomer). Due to steric effects, second-order Markov effects may also play a role in the growth of some polymer chains.

Similarly, it has been suggested that the crystallization and growth of some epitaxial superlattice oxide materials can be accurately described by Markov chains.[77]

Biology

Markov chains are used in various areas of biology. Notable examples include:

Testing

Several theorists have proposed the idea of the Markov chain statistical test (MCST), a method of conjoining Markov chains to form a "Markov blanket", arranging these chains in several recursive layers ("wafering") and producing more efficient test sets—samples—as a replacement for exhaustive testing. MCSTs also have uses in temporal state-based networks; Chilukuri et al.'s paper entitled "Temporal Uncertainty Reasoning Networks for Evidence Fusion with Applications to Object Detection and Tracking" (ScienceDirect) gives a background and case study for applying MCSTs to a wider range of applications.

Solar irradiance variability

Solar irradiance variability assessments are useful for solar power applications. Solar irradiance variability at any location over time is mainly a consequence of the deterministic variability of the sun's path across the sky dome and the variability in cloudiness. The variability of accessible solar irradiance on Earth's surface has been modeled using Markov chains,[80][81][82][83] also including modeling the two states of clear and cloudiness as a two-state Markov chain.[84][85]

Speech recognition

Hidden Markov models are the basis for most modern automatic speech recognition systems.

Information theory

Markov chains are used throughout information processing.

state estimation and pattern recognition. Markov chains also play an important role in reinforcement learning
.

Markov chains are also the basis for hidden Markov models, which are an important tool in such diverse fields as telephone networks (which use the Viterbi algorithm for error correction), speech recognition and bioinformatics (such as in rearrangements detection[86]).

The LZMA lossless data compression algorithm combines Markov chains with Lempel-Ziv compression to achieve very high compression ratios.

Queueing theory

Markov chains are the basis for the analytical treatment of queues (queueing theory). Agner Krarup Erlang initiated the subject in 1917.[87] This makes them critical for optimizing the performance of telecommunications networks, where messages must often compete for limited resources (such as bandwidth).[88]

Numerous queueing models use continuous-time Markov chains. For example, an

Poisson process
and describe job arrivals, while transitions from i to i – 1 (for i > 1) occur at rate μ (job service times are exponentially distributed) and describe completed services (departures) from the queue.

Internet applications

A state diagram that represents the PageRank algorithm with a transitional probability of M, or .

The PageRank of a webpage as used by Google is defined by a Markov chain.[89][90][91] It is the probability to be at page in the stationary distribution on the following Markov chain on all (known) webpages. If is the number of known webpages, and a page has links to it then it has transition probability for all pages that are linked to and for all pages that are not linked to. The parameter is taken to be about 0.15.[92]

Markov models have also been used to analyze web navigation behavior of users. A user's web link transition on a particular website can be modeled using first- or second-order Markov models and can be used to make predictions regarding future navigation and to personalize the web page for an individual user.

Statistics

Markov chain methods have also become very important for generating sequences of random numbers to accurately reflect very complicated desired probability distributions, via a process called

posterior distributions
to be simulated and their parameters found numerically.

Economics and finance

Markov chains are used in finance and economics to model a variety of different phenomena, including the distribution of income, the size distribution of firms, asset prices and market crashes.

Laurent E. Calvet and Adlai J. Fisher, which builds upon the convenience of earlier regime-switching models.[98][99]
It uses an arbitrarily large Markov chain to drive the level of volatility of asset returns.

Dynamic macroeconomics makes heavy use of Markov chains. An example is using Markov chains to exogenously model prices of equity (stock) in a

Credit rating agencies produce annual tables of the transition probabilities for bonds of different credit ratings.[101]

Social sciences

Markov chains are generally used in describing

democratic regime.[102]

Games

Markov chains can be used to model many games of chance.

", for example, are represented exactly by Markov chains. At each turn, the player starts in a given state (on a given square) and from there has fixed odds of moving to certain other states (squares).

Music

Markov chains are employed in algorithmic music composition, particularly in software such as Csound, Max, and SuperCollider. In a first-order chain, the states of the system become note or pitch values, and a probability vector for each note is constructed, completing a transition probability matrix (see below). An algorithm is constructed to produce output note values based on the transition matrix weightings, which could be MIDI note values, frequency (Hz), or any other desirable metric.[103]

1st-order matrix
Note A C E
A 0.1 0.6 0.3
C 0.25 0.05 0.7
E 0.7 0.3 0
2nd-order matrix
Notes A D G
AA 0.18 0.6 0.22
AD 0.5 0.5 0
AG 0.15 0.75 0.1
DD 0 0 1
DA 0.25 0 0.75
DG 0.9 0.1 0
GG 0.4 0.4 0.2
GA 0.5 0.25 0.25
GD 1 0 0

A second-order Markov chain can be introduced by considering the current state and also the previous state, as indicated in the second table. Higher, nth-order chains tend to "group" particular notes together, while 'breaking off' into other patterns and sequences occasionally. These higher-order chains tend to generate results with a sense of phrasal structure, rather than the 'aimless wandering' produced by a first-order system.[104]

Markov chains can be used structurally, as in Xenakis's Analogique A and B.[105] Markov chains are also used in systems which use a Markov model to react interactively to music input.[106]

Usually musical systems need to enforce specific control constraints on the finite-length sequences they generate, but control constraints are not compatible with Markov models, since they induce long-range dependencies that violate the Markov hypothesis of limited memory. In order to overcome this limitation, a new approach has been proposed.[107]

Baseball

Markov chain models have been used in advanced baseball analysis since 1960, although their use is still rare. Each half-inning of a baseball game fits the Markov chain state when the number of runners and outs are considered. During any at-bat, there are 24 possible combinations of number of outs and position of the runners. Mark Pankin shows that Markov chain models can be used to evaluate runs created for both individual players as well as a team.[108] He also discusses various kinds of strategies and play conditions: how Markov chain models have been used to analyze statistics for game situations such as

base stealing and differences when playing on grass vs. AstroTurf.[109]

Markov text generators

Markov processes can also be used to generate superficially real-looking text given a sample document. Markov processes are used in a variety of recreational "parody generator" software (see dissociated press, Jeff Harrison,[110] Mark V. Shaney,[111][112] and Academias Neutronium). Several open-source text generation libraries using Markov chains exist.

Probabilistic forecasting

Markov chains have been used for forecasting in several areas: for example, price trends,[113] wind power,[114] stochastic terrorism,[115][116] and solar irradiance.[117] The Markov chain forecasting models utilize a variety of settings, from discretizing the time series,[114] to hidden Markov models combined with wavelets,[113] and the Markov chain mixture distribution model (MCM).[117]

See also

Notes

  1. ^ .
  2. ^ "Markov chain | Definition of Markov chain in US English by Oxford Dictionaries". Oxford Dictionaries. Archived from the original on December 15, 2017. Retrieved 2017-12-14.
  3. ^ Definition at Brilliant.org "Brilliant Math and Science Wiki". Retrieved on 12 May 2019
  4. from the original on 23 March 2017.
  5. from the original on 23 March 2017.
  6. from the original on 23 March 2017.
  7. ^ from the original on 23 March 2017.
  8. from the original on 23 March 2017.
  9. from the original on 23 March 2017.
  10. ^ "Markovian". Oxford English Dictionary (Online ed.). Oxford University Press. (Subscription or participating institution membership required.)
  11. .
  12. .
  13. ^ from the original on 23 March 2017.
  14. from the original on 20 November 2017.
  15. from the original on 23 March 2017.
  16. from the original on 2017-03-23.
  17. from the original on 2017-03-23.
  18. (Table 6.1)
  19. (entry for "Markov chain")
  20. ^ Meyn, S. Sean P., and Richard L. Tweedie. (2009) Markov chains and stochastic stability. Cambridge University Press. (Preface, p. iii)
  21. ^ .
  22. .
  23. ^ .
  24. ^ from the original on 23 March 2017.
  25. ^ .
  26. ^ from the original on 2017-03-23.
  27. .
  28. .
  29. .
  30. .
  31. .
  32. ^ .
  33. ^ .
  34. from the original on 23 March 2017.
  35. from the original on 23 March 2017.
  36. .
  37. from the original on 23 March 2017.
  38. .
  39. ^ from the original on 23 March 2017.
  40. ^ from the original on 23 March 2017.
  41. .
  42. .
  43. from the original on 2017-03-23.
  44. from the original on 20 November 2017.
  45. ^ Joseph L. Doob (1990). Stochastipoic processes. Wiley. p. 46 and 47. Archived from the original on 2017-11-20.
  46. from the original on 20 November 2017.
  47. .
  48. ^ .
  49. ^ "Chapter 11 "Markov Chains"" (PDF). Archived (PDF) from the original on 2017-02-15. Retrieved 2017-06-02.
  50. CiteSeerX 10.1.1.28.6191
    .
  51. . Retrieved 2021-05-31.
  52. .
  53. .
  54. from the original on 2017-02-05. Retrieved 2016-03-04.
  55. .
  56. ^ Peres, Yuval. "Show that positive recurrence is a class property". Mathematics Stack Exchange. Retrieved 2024-02-01.
  57. .
  58. ^ a b Shalizi, Cosma (1 Dec 2023). "Ergodic Theory". bactra.org. Archived from the original on 31 Dec 2023. Retrieved 2024-02-01.
  59. .
  60. ^ "10.3: Regular Markov Chains". Mathematics LibreTexts. 2020-03-22. Retrieved 2024-02-01.
  61. .
  62. .
  63. .
  64. ^ Doblinger, G. (September 1998). "Smoothing of noisy AR signals using an adaptive Kalman filter" (PDF). 9th European Signal Processing Conference (EUSIPCO 1998): 781–784.
  65. .
  66. ^ a b c Matthew Nicol and Karl Petersen, (2009) "Ergodic Theory: Basic Examples and Constructions", Encyclopedia of Complexity and Systems Science, Springer https://doi.org/10.1007/978-0-387-30440-3_177
  67. ^ Fitzpatrick, Richard. "Thermodynamics and Statistical Mechanics" (PDF). Archived (PDF) from the original on 2016-11-30. Retrieved 2017-06-02.
  68. ^
    PMID 26968853
    .
  69. .
  70. from the original on 2017-08-01.
  71. .
  72. .
  73. .
  74. .
  75. .
  76. .
  77. .
  78. .
  79. .
  80. .
  81. .
  82. .
  83. .
  84. ^ O'Connor, John J.; Robertson, Edmund F., "Markov chain", MacTutor History of Mathematics Archive, University of St Andrews
  85. ^ S. P. Meyn, 2007. Control Techniques for Complex Networks Archived 2015-05-13 at the Wayback Machine, Cambridge University Press, 2007.
  86. ^ U.S. patent 6,285,999
  87. from the original on 23 March 2017.
  88. (PDF) from the original on 2017-09-21.
  89. .
  90. .
  91. ^ Simon, Herbert; C Bonini (1958). "The size distribution of business firms". Am. Econ. Rev. 42: 425–40.
  92. .
  93. ^ e.g.Fama, E (1965). "The behavior of stock market prices". Journal of Business. 38.
  94. JSTOR 1912559
    .
  95. .
  96. .
  97. ^ Brennan, Michael; Xiab, Yihong. "Stock Price Volatility and the Equity Premium" (PDF). Department of Finance, the Anderson School of Management, UCLA. Archived from the original (PDF) on 2008-12-28.
  98. ^ "A Markov Chain Example in Credit Risk Modelling" (PDF). Columbia University. Archived from the original (PDF) on March 24, 2016.
  99. PMID 22198760
    .
  100. .
  101. .
  102. ^ "Continuator". Archived from the original on July 13, 2012.
  103. ^ Pachet, F.; Roy, P.; Barbieri, G. (2011) "Finite-Length Markov Processes with Constraints" Archived 2012-04-14 at the Wayback Machine, Proceedings of the 22nd International Joint Conference on Artificial Intelligence, IJCAI, pages 635–642, Barcelona, Spain, July 2011
  104. ^ Pankin, Mark D. "MARKOV CHAIN MODELS: THEORETICAL BACKGROUND". Archived from the original on 2007-12-09. Retrieved 2007-11-26.
  105. ^ Pankin, Mark D. "BASEBALL AS A MARKOV CHAIN". Archived from the original on 2009-05-13. Retrieved 2009-04-24.
  106. ^ "Poet's Corner – Fieralingue". Archived from the original on December 6, 2010.
  107. ^ Kenner, Hugh; O'Rourke, Joseph (November 1984). "A Travesty Generator for Micros". BYTE. 9 (12): 129–131, 449–469.
  108. .
  109. ^ a b de Souza e Silva, E.G.; Legey, L.F.L.; de Souza e Silva, E.A. (2010). "Forecasting oil price trends using wavelets and hidden Markov models". Energy Economics. 32.
  110. ^ .
  111. . Retrieved 5 October 2023.
  112. ^ Woo, Gordon (December 2003). "Insuring Against Al-Quaeda" (PDF). Cambridge: National Bureau of Economic Research. Retrieved 26 March 2024.
  113. ^
    S2CID 146076100
    .

References

External links