Kingman's subadditive ergodic theorem

Last updated

In mathematics, Kingman's subadditive ergodic theorem is one of several ergodic theorems. It can be seen as a generalization of Birkhoff's ergodic theorem. [1] Intuitively, the subadditive ergodic theorem is a kind of random variable version of Fekete's lemma (hence the name ergodic). [2] As a result, it can be rephrased in the language of probability, e.g. using a sequence of random variables and expected values. The theorem is named after John Kingman.

Contents

Statement of theorem

Let be a measure-preserving transformation on the probability space , and let be a sequence of functions such that (subadditivity relation). Then

for -a.e. x, where g(x) is T-invariant.

In particular, if T is ergodic, then g(x) is a constant.

Equivalent statement

Given a family of real random variables , with , such that they are subadditive in the sense thatThen there exists a random variable such that , is invariant with respect to , and a.s..

They are equivalent by setting

Proof

Proof due to (J. Michael Steele, 1989). [3]

Subadditivity by partition

Fix some . By subadditivity, for any

We can picture this as starting with the set , and then removing its length tail.

Repeating this construction until the set is all gone, we have a one-to-one correspondence between upper bounds of and partitions of .

Specifically, let be a partition of , then we have

Constructing g

Let , then it is -invariant.

By subadditivity,

Taking the limit, we have We can visualize as hill-climbing on the graph of . If actually causes a nontrivial amount of hill-climbing, then we would get a spatial contraction, and so does not preserve measure. Therefore a.e.

Let , then and since both sides have the same measure, by squeezing, they are equal a.e..

That is, , a.e..

Now apply this for all rational .

Reducing to the case of gₙ ≤ 0


By subadditivity, using the partition of into singletons. Now, construct the sequence which satisfies for all .

By the special case, converges a.e. to a -invariant function.

By Birkhoff's pointwise ergodic theorem, the running average converges a.e. to a -invariant function. Therefore, their sum does as well.

Bounding the truncation

Fix arbitrary , and construct the truncated function, still -invariant: With these, it suffices to prove an a.e. upper boundsince it would allow us to take the limit , then the limit , giving us a.e.

And by squeezing, we have converging a.e. to . Define two families of sets, one shrinking to the empty set, and one growing to the full set. For each "length" , defineSince , the family shrinks to the empty set.


Fix . Fix . Fix . The ordering of these qualifiers is vitally important, because we will be removing the qualifiers one by one in the reverse order.

To prove the a.e. upper bound, we must use the subadditivity, which means we must construct a partition of the set . We do this inductively:

Take the smallest not already in a partition.

If , then for some . Take one such – the choice does not matter.

If , then we cut out . Call these partitions “type 1”. Else, we cut out . Call these partitions “type 2”.

Else, we cut out . Call these partitions “type 3”.

Now convert this partition into an inequality: where are the heads of the partitions, and are the lengths.

Since all , we can remove the other kinds of partitions: By construction, each , thus Now it would be tempting to continue with , but unfortunately , so the direction is the exact opposite. We must lower bound the sum .

The number of type 3 elements is equal toIf a number is of type 2, then it must be inside the last elements of . Thus the number of type 2 elements is at most . Together, we have the lower bound:

Peeling off the first qualifier

Remove the qualifier by taking the limit.

By Birkhoff's pointwise ergodic theorem, there exists an a.e. pointwise limit satisfying
At the limit, we find that for a.e. ,

Peeling off the second qualifier

Remove the qualifier by taking the limit.

Since we have and as , we can apply the same argument used for proving Markov's inequality, to obtain
for a.e. .


In detail, the argument is as follows: since , and , we know that for any small , all large enough satisfies everywhere except on a set of size . Thus,with probability . Now take both .

Applications

Taking recovers Birkhoff's pointwise ergodic theorem.

Taking all constant functions, we recover the Fekete's subadditive lemma.

Kingman's subadditive ergodic theorem can be used to prove statements about Lyapunov exponents. It also has applications to percolations and longest increasing subsequence. [4]

Longest increasing subsequence

To study the longest increasing subsequence of a random permutation , we generate it in an equivalent way. A random permutation on is equivalently generated by uniformly sampling points in a square, then find the longest increasing subsequence of that.

Now, define the Poisson point process with density 1 on , and define the random variables to be the length of the longest increasing subsequence in the square . Define the measure-preserving transform by shifting the plane by , then chopping off the parts that have fallen out of .

The process is subadditive, that is, . To see this, notice that the right side constructs an increasing subsequence first in the square , then in the square , and finally concatenate them together. This produces an increasing subsequence in , but not necessarily the longest one.

Also, is ergodic, so by Kingman's theorem, converges to a constant almost surely. Since at the limit, there are points in the square, we have converging to a constant almost surely.

Related Research Articles

In mathematics, the prime number theorem (PNT) describes the asymptotic distribution of the prime numbers among the positive integers. It formalizes the intuitive idea that primes become less common as they become larger by precisely quantifying the rate at which this occurs. The theorem was proved independently by Jacques Hadamard and Charles Jean de la Vallée Poussin in 1896 using ideas introduced by Bernhard Riemann.

In mathematics, the branch of real analysis studies the behavior of real numbers, sequences and series of real numbers, and real functions. Some particular properties of real-valued sequences and functions that real analysis studies include convergence, limits, continuity, smoothness, differentiability and integrability.

<span class="mw-page-title-main">Uniform convergence</span> Mode of convergence of a function sequence

In the mathematical field of analysis, uniform convergence is a mode of convergence of functions stronger than pointwise convergence. A sequence of functions converges uniformly to a limiting function on a set as the function domain if, given any arbitrarily small positive number , a number can be found such that each of the functions differs from by no more than at every pointin. Described in an informal way, if converges to uniformly, then how quickly the functions approach is "uniform" throughout in the following sense: in order to guarantee that differs from by less than a chosen distance , we only need to make sure that is larger than or equal to a certain , which we can find without knowing the value of in advance. In other words, there exists a number that could depend on but is independent of , such that choosing will ensure that for all . In contrast, pointwise convergence of to merely guarantees that for any given in advance, we can find such that, for that particular, falls within of whenever .

In mathematics, an infinite series of numbers is said to converge absolutely if the sum of the absolute values of the summands is finite. More precisely, a real or complex series is said to converge absolutely if for some real number Similarly, an improper integral of a function, is said to converge absolutely if the integral of the absolute value of the integrand is finite—that is, if A convergent series that is not absolutely convergent is called conditionally convergent.

In the mathematical field of real analysis, the monotone convergence theorem is any of a number of related theorems proving the good convergence behaviour of monotonic sequences, i.e. sequences that are non-increasing, or non-decreasing. In its simplest form, it says that a non-decreasing bounded-above sequence of real numbers converges to its smallest upper bound, its supremum. Likewise, a non-increasing bounded-below sequence converges to its largest lower bound, its infimum. In particular, infinite sums of non-negative numbers converge to the supremum of the partial sums if and only if the partial sums are bounded.

In probability theory, Chebyshev's inequality provides an upper bound on the probability of deviation of a random variable from its mean. More specifically, the probability that a random variable deviates from its mean by more than is at most , where is any positive constant and is the standard deviation.

<span class="mw-page-title-main">Law of large numbers</span> Averages of repeated trials converge to the expected value

In probability theory, the law of large numbers (LLN) is a mathematical law that states that the average of the results obtained from a large number of independent random samples converges to the true value, if it exists. More formally, the LLN states that given a sample of independent and identically distributed values, the sample mean converges to the true mean.

In information theory, the asymptotic equipartition property (AEP) is a general property of the output samples of a stochastic source. It is fundamental to the concept of typical set used in theories of data compression.

<span class="mw-page-title-main">Limit of a sequence</span> Value to which tends an infinite sequence

In mathematics, the limit of a sequence is the value that the terms of a sequence "tend to", and is often denoted using the symbol. If such a limit exists and is finite, the sequence is called convergent. A sequence that does not converge is said to be divergent. The limit of a sequence is said to be the fundamental notion on which the whole of mathematical analysis ultimately rests.

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.

In mathematics, subadditivity is a property of a function that states, roughly, that evaluating the function for the sum of two elements of the domain always returns something less than or equal to the sum of the function's values at each element. There are numerous examples of subadditive functions in various areas of mathematics, particularly norms and square roots. Additive maps are special cases of subadditive functions.

In mathematics, the Riesz–Thorin theorem, often referred to as the Riesz–Thorin interpolation theorem or the Riesz–Thorin convexity theorem, is a result about interpolation of operators. It is named after Marcel Riesz and his student G. Olof Thorin.

In mathematics, the Stolz–Cesàro theorem is a criterion for proving the convergence of a sequence. It is named after mathematicians Otto Stolz and Ernesto Cesàro, who stated and proved it for the first time.

In mathematics, Fejér's theorem, named after Hungarian mathematician Lipót Fejér, states the following:

In the mathematical discipline of graph theory, the expander walk sampling theorem intuitively states that sampling vertices in an expander graph by doing relatively short random walk can simulate sampling the vertices independently from a uniform distribution. The earliest version of this theorem is due to Ajtai, Komlós & Szemerédi (1987), and the more general version is typically attributed to Gillman (1998).

In mathematics, the Brunn–Minkowski theorem is an inequality relating the volumes of compact subsets of Euclidean space. The original version of the Brunn–Minkowski theorem applied to convex sets; the generalization to compact nonconvex sets stated here is due to Lazar Lyusternik (1935).

In multivariable calculus, an iterated limit is a limit of a sequence or a limit of a function in the form

In mathematics, the Rokhlin lemma, or Kakutani–Rokhlin lemma is an important result in ergodic theory. It states that an aperiodic measure preserving dynamical system can be decomposed to an arbitrary high tower of measurable sets and a remainder of arbitrarily small measure. It was proven by Vladimir Abramovich Rokhlin and independently by Shizuo Kakutani. The lemma is used extensively in ergodic theory, for example in Ornstein theory and has many generalizations.

In the branch of mathematics known as integration theory, the McShane integral, created by Edward J. McShane, is a modification of the Henstock-Kurzweil integral. The McShane integral is equivalent to the Lebesgue integral.

In calculus, the Abel–Dini–Pringsheim theorem is a convergence test which constructs from a divergent series a series that diverges more slowly, and from convergent series one that converges more slowly. Consequently, for every convergence test based on a particular series there is a series about which the test is inconclusive. For example, the Raabe test is essentially a comparison test based on the family of series whose th term is and is therefore inconclusive about the series of terms which diverges more slowly than the harmonic series.

References

  1. S. Lalley, Kingman's subadditive ergodic theorem lecture notes, http://galton.uchicago.edu/~lalley/Courses/Graz/Kingman.pdf
  2. Chen. "Subadditive ergodic theorems" (PDF). New York University.
  3. Steele, J. Michael (1989). "Kingman's subadditive ergodic theorem" (PDF). Annales de l'I.H.P. Probabilités et statistiques. 25 (1): 93–98. ISSN   1778-7017.
  4. Pitman, Lecture 12: Subadditive ergodic theory, http://www.stat.berkeley.edu/~pitman/s205s03/lecture12.pdf