Doob's martingale convergence theorems

In mathematics  specifically, in the theory of stochastic processes  Doob's martingale convergence theorems are a collection of results on the long-time limits of supermartingales, named after the American mathematician Joseph L. Doob.[1]

Statement of the theorems

In the following, (Ω, F, F, P), F = (Ft)t ≥ 0, will be a filtered probability space and N : [0, +∞) × Ω  R will be a right-continuous supermartingale with respect to the filtration F; in other words, for all 0  s  t < +∞,

Doob's first martingale convergence theorem

Doob's first martingale convergence theorem provides a sufficient condition for the random variables Nt to have a limit as t  +∞ in a pointwise sense, i.e. for each ω in the sample space Ω individually.

For t  0, let Nt = max(Nt, 0) and suppose that

Then the pointwise limit

exists and is finite for P-almost all ω  Ω.[2]

Doob's second martingale convergence theorem

It is important to note that the convergence in Doob's first martingale convergence theorem is pointwise, not uniform, and is unrelated to convergence in mean square, or indeed in any Lp space. In order to obtain convergence in L1 (i.e., convergence in mean), one requires uniform integrability of the random variables Nt. By Chebyshev's inequality, convergence in L1 implies convergence in probability and convergence in distribution.

The following are equivalent:

  • there exists an integrable random variable N  L1(Ω, P; R) such that Nt  N as t  +∞ both P-almost surely and in L1(Ω, P; R), i.e.

Corollary: convergence theorem for continuous martingales

Let M : [0, +∞) × Ω  R be a continuous martingale such that

for some p > 1. Then there exists a random variable M  Lp(Ω, P; R) such that Mt  M as t  +∞ both P-almost surely and in Lp(Ω, P; R).

Discrete-time results

Similar results can be obtained for discrete-time supermartingales and submartingales, the obvious difference being that no continuity assumptions are required. For example, the result above becomes

Let M : N × Ω  R be a discrete-time martingale such that

for some p > 1. Then there exists a random variable M  Lp(Ω, P; R) such that Mk  M as k  +∞ both P-almost surely and in Lp(Ω, P; R)

Convergence of conditional expectations: Lévy's zero–one law

Doob's martingale convergence theorems imply that conditional expectations also have a convergence property.

Let (Ω, F, P) be a probability space and let X be a random variable in L1. Let F = (Fk)kN be any filtration of F, and define F to be the minimal σ-algebra generated by (Fk)kN. Then

both P-almost surely and in L1.

This result is usually called Lévy's zero–one law or Levy's upwards theorem. The reason for the name is that if A is an event in F, then the theorem says that almost surely, i.e., the limit of the probabilities is 0 or 1. In plain language, if we are learning gradually all the information that determines the outcome of an event, then we will become gradually certain what the outcome will be. This sounds almost like a tautology, but the result is still non-trivial. For instance, it easily implies Kolmogorov's zero–one law, since it says that for any tail event A, we must have almost surely, hence .

Similarly we have the Levy's downwards theorem :

Let (Ω, F, P) be a probability space and let X be a random variable in L1. Let (Fk)kN be any decreasing sequence of sub-sigma algebras of F, and define F to be the intersection. Then

both P-almost surely and in L1.

Doob's upcrossing inequality

The following result, called Doob's upcrossing inequality or, sometimes, Doob's upcrossing lemma, is used in proving Doob's martingale convergence theorems.[2]
Hypothesis.
Let be a natural number. Let , for , be a martingale with respect to a filtration , for . Let , be two real numbers with .
Define the random variables , for , as follows: if and only if is the largest integer such that there exist integers , satisfying 1 ≤ < and, for , for each pair the inequalities and are satisfied. Each is called the number of upcrossings with respect to the interval for the martingale , .
Conclusion.

.[3][4]

See also

  • Backwards martingale convergence theorem[5]

References

  1. Doob, J. L. (1953). Stochastic Processes. New York: Wiley.
  2. 1 2 "Martingale Convergence Theorem" (PDF). Massachusetts Institute of Tecnnology, 6.265/15.070J Lecture 11-Additional Material, Advanced Stochastic Processes, Fall 2013, 10/9/2013.
  3. Bobrowski, Adam (2005). Functional Analysis for Probability and Stochastic Processe: An Introduction. Cambridge University Press. pp. 113–114. ISBN 9781139443883.
  4. Gushchin, A. A. (2014). "On pathwise counterparts of Doob's maximal inequalities". Proceedings of the Steklov Institute of Mathematics. 287 (287): 118–121. arXiv:1410.8264. doi:10.1134/S0081543814080070.
  5. Doob, Joseph L. (2012). Measure theory. Graduate Texts in Mathematics, Vol. 143. Springer. p. 197. ISBN 9781461208778.
  • Øksendal, Bernt K. (2003). Stochastic Differential Equations: An Introduction with Applications (Sixth ed.). Berlin: Springer. ISBN 3-540-04758-1. (See Appendix C)
  • Durrett, Rick (1996). Probability: theory and examples (Second ed.). Duxbury Press. ISBN 978-0-534-24318-0. ; Durrett, Rick (2010). 4th edition. ISBN 9781139491136.
This article is issued from Wikipedia. The text is licensed under Creative Commons - Attribution - Sharealike. Additional terms may apply for the media files.