Jump to content
 







Main menu
   


Navigation  



Main page
Contents
Current events
Random article
About Wikipedia
Contact us
Donate
 




Contribute  



Help
Learn to edit
Community portal
Recent changes
Upload file
 








Search  

































Create account

Log in
 









Create account
 Log in
 




Pages for logged out editors learn more  



Contributions
Talk
 



















Contents

   



(Top)
 


1 Statement  



1.1  Remark  







2 Proof  



2.1  Existence  





2.2  Uniqueness  







3 Corollary  



3.1  Proof  







4 Example  





5 Application  





6 Generalization  





7 Citations  





8 References  














Doob decomposition theorem






Deutsch
Español
 

Edit links
 









Article
Talk
 

















Read
Edit
View history
 








Tools
   


Actions  



Read
Edit
View history
 




General  



What links here
Related changes
Upload file
Special pages
Permanent link
Page information
Cite this page
Get shortened URL
Download QR code
Wikidata item
 




Print/export  



Download as PDF
Printable version
 
















Appearance
   

 






From Wikipedia, the free encyclopedia
 


In the theory of stochastic processesindiscrete time, a part of the mathematical theory of probability, the Doob decomposition theorem gives a unique decomposition of every adapted and integrable stochastic process as the sum of a martingale and a predictable process (or "drift") starting at zero. The theorem was proved by and is named for Joseph L. Doob.[1]

The analogous theorem in the continuous-time case is the Doob–Meyer decomposition theorem.

Statement

[edit]

Let be a probability space, I = {0, 1, 2, ..., N} with or a finite or countably infinite index set, afiltration of , and X = (Xn)nI an adapted stochastic process with E[|Xn|] < ∞ for all nI. Then there exist a martingale M = (Mn)nI and an integrable predictable process A = (An)nI starting with A0 = 0 such that Xn = Mn + An for every nI. Here predictable means that Anis-measurable for every nI \ {0}. This decomposition is almost surely unique.[2][3][4]

Remark

[edit]

The theorem is valid word for word also for stochastic processes X taking values in the d-dimensional Euclidean space or the complex vector space . This follows from the one-dimensional version by considering the components individually.

Proof

[edit]

Existence

[edit]

Using conditional expectations, define the processes A and M, for every nI, explicitly by

(1)

and

(2)

where the sums for n = 0 are empty and defined as zero. Here A adds up the expected increments of X, and M adds up the surprises, i.e., the part of every Xk that is not known one time step before. Due to these definitions, An+1 (ifn + 1 ∈ I) and Mn are Fn-measurable because the process X is adapted, E[|An|] < ∞ and E[|Mn|] < ∞ because the process X is integrable, and the decomposition Xn = Mn + An is valid for every nI. The martingale property

    a.s.

also follows from the above definition (2), for every nI \ {0}.

Uniqueness

[edit]

To prove uniqueness, let X = M' + A' be an additional decomposition. Then the process Y := MM' = A'A is a martingale, implying that

    a.s.,

and also predictable, implying that

    a.s.

for any nI \ {0}. Since Y0 = A'0A0 = 0 by the convention about the starting point of the predictable processes, this implies iteratively that Yn = 0 almost surely for all nI, hence the decomposition is almost surely unique.

Corollary

[edit]

A real-valued stochastic process X is a submartingale if and only if it has a Doob decomposition into a martingale M and an integrable predictable process A that is almost surely increasing.[5] It is a supermartingale, if and only if A is almost surely decreasing.

Proof

[edit]

IfX is a submartingale, then

    a.s.

for all kI \ {0}, which is equivalent to saying that every term in definition (1) of A is almost surely positive, hence A is almost surely increasing. The equivalence for supermartingales is proved similarly.

Example

[edit]

Let X = (Xn)n be a sequence in independent, integrable, real-valued random variables. They are adapted to the filtration generated by the sequence, i.e. Fn = σ(X0, . . . , Xn) for all n. By (1) and (2), the Doob decomposition is given by

and

If the random variables of the original sequence X have mean zero, this simplifies to

    and    

hence both processes are (possibly time-inhomogeneous) random walks. If the sequence X = (Xn)n consists of symmetric random variables taking the values +1 and −1, then X is bounded, but the martingale M and the predictable process A are unbounded simple random walks (and not uniformly integrable), and Doob's optional stopping theorem might not be applicable to the martingale M unless the stopping time has a finite expectation.

Application

[edit]

Inmathematical finance, the Doob decomposition theorem can be used to determine the largest optimal exercise time of an American option.[6][7] Let X = (X0, X1, . . . , XN) denote the non-negative, discounted payoffs of an American option in a N-period financial market model, adapted to a filtration (F0, F1, . . . , FN), and let denote an equivalent martingale measure. Let U = (U0, U1, . . . , UN) denote the Snell envelope of X with respect to . The Snell envelope is the smallest -supermartingale dominating X[8] and in a complete financial market it represents the minimal amount of capital necessary to hedge the American option up to maturity.[9] Let U = M + A denote the Doob decomposition with respect to of the Snell envelope U into a martingale M = (M0, M1, . . . , MN) and a decreasing predictable process A = (A0, A1, . . . , AN) with A0 = 0. Then the largest stopping time to exercise the American option in an optimal way[10][11]is

Since A is predictable, the event {τmax = n} = {An = 0, An+1 < 0} is in Fn for every n ∈ {0, 1, . . . , N − 1}, hence τmax is indeed a stopping time. It gives the last moment before the discounted value of the American option will drop in expectation; up to time τmax the discounted value process U is a martingale with respect to .

Generalization

[edit]

The Doob decomposition theorem can be generalized from probability spaces to σ-finite measure spaces.[12]

Citations

[edit]
  1. ^ Doob (1953), see (Doob 1990, pp. 296−298)
  • ^ Durrett (2010)
  • ^ (Föllmer & Schied 2011, Proposition 6.1)
  • ^ (Williams 1991, Section 12.11, part (a) of the Theorem)
  • ^ (Williams 1991, Section 12.11, part (b) of the Theorem)
  • ^ (Lamberton & Lapeyre 2008, Chapter 2: Optimal stopping problem and American options)
  • ^ (Föllmer & Schied 2011, Chapter 6: American contingent claims)
  • ^ (Föllmer & Schied 2011, Proposition 6.10)
  • ^ (Föllmer & Schied 2011, Theorem 6.11)
  • ^ (Lamberton & Lapeyre 2008, Proposition 2.3.2)
  • ^ (Föllmer & Schied 2011, Theorem 6.21)
  • ^ (Schilling 2005, Problem 23.11)
  • References

    [edit]
    Retrieved from "https://en.wikipedia.org/w/index.php?title=Doob_decomposition_theorem&oldid=1206877684"

    Categories: 
    Theorems regarding stochastic processes
    Martingale theory
    Hidden categories: 
    Articles with short description
    Short description matches Wikidata
    Articles containing proofs
     



    This page was last edited on 13 February 2024, at 10:28 (UTC).

    Text is available under the Creative Commons Attribution-ShareAlike License 4.0; additional terms may apply. By using this site, you agree to the Terms of Use and Privacy Policy. Wikipedia® is a registered trademark of the Wikimedia Foundation, Inc., a non-profit organization.



    Privacy policy

    About Wikipedia

    Disclaimers

    Contact Wikipedia

    Code of Conduct

    Developers

    Statistics

    Cookie statement

    Mobile view



    Wikimedia Foundation
    Powered by MediaWiki