In probability theory, a standard probability space, also called Lebesgue - Rokhlin probability space or just Lebesgue space (the latter term is ambiguous) is a probability space satisfying certain assumptions introduced by Vladimir Rokhlin in 1940. Informally, it is a probability space consisting of an interval and/or a finite or countable number of atoms.
The theory of standard probability spaces was started by von Neumann in 1932 and shaped by Vladimir Rokhlin in 1940. Rokhlin showed that the unit interval endowed with the Lebesgue measure has important advantages over general probability spaces, yet can be effectively substituted for many of these in probability theory. The dimension of the unit interval is not an obstacle, as was clear already to Norbert Wiener. He constructed the Wiener process (also called Brownian motion) in the form of a measurable map from the unit interval to the space of continuous functions.
The theory of standard probability spaces was started by von Neumann in 1932[1] and shaped by Vladimir Rokhlin in 1940.[2] For modernized presentations see,, and .
Nowadays standard probability spaces may be (and often are) treated in the framework of descriptive set theory, via standard Borel spaces, see for example . This approach is based on the isomorphism theorem for standard Borel spaces . An alternate approach of Rokhlin, based on measure theory, neglects null sets, in contrast to descriptive set theory.Standard probability spaces are used routinely in ergodic theory.[3] [4]
One of several well-known equivalent definitions of the standardness is given below, after some preparations. All probability spaces are assumed to be complete.
An isomorphism between two probability spaces
style(\Omega1,l{F}1,P1)
style(\Omega2,l{F}2,P2)
stylef:\Omega1\to\Omega2
stylef
stylef-1
Two probability spaces are isomorphic if there exists an isomorphism between them.
Two probability spaces
style(\Omega1,l{F}1,P1)
style(\Omega2,l{F}2,P2)
style\operatorname{mod}0
styleA1\subset\Omega1
styleA2\subset\Omega2
style\Omega1\setminusA1
style\Omega2\setminusA2
A probability space is standard, if it is isomorphic
style\operatorname{mod}0
See,, and . See also, and . In the measure is assumed finite, not necessarily probabilistic. In atoms are not allowed.
The space of all functions
stylef:R\toR
styleRR
styleR
styleR
style\gamma=N(0,1)
style(R,\gamma)R
style(R,\gamma)
style\gammaR
styleRR
style\gammaR
However, the integral of a white noise function from 0 to 1 should be a random variable distributed N(0, 1). In contrast, the integral (from 0 to 1) of
stylef\instyle(R,\gamma)R
Let
styleZ\subset(0,1)
styleZ
stylem
styleZ
stylem(Z\capA)=\operatorname{mes}(A)
styleA\subset(0,1)
style\operatorname{mes}
style(Z,m)
style\operatorname{mod}0
style((0,1),\operatorname{mes})
style(Z,m)
style((0,1),\operatorname{mes})
However, it is not. A random variable
styleX
styleX(\omega)=\omega
style(0,1)
styleX=x
stylex
style((0,1),\operatorname{mes})
style(Z,m)
stylex\notinZ
A perforated circle is constructed similarly. Its events and random variables are the same as on the usual circle. The group of rotations acts on them naturally. However, it fails to act on the perforated circle.
See also .
Let
styleZ\subset(0,1)
style(A\capZ)\cup(B\setminusZ),
styleA
styleB
stylel{F};
styleZ.
\displaystylem((A\capZ)\cup(B\setminusZ))=p\operatorname{mes}(A)+(1-p)\operatorname{mes}(B)
stylem
style((0,1),l{F})
stylep\in[0,1]
stylep=0.5.
However, it is the perforated interval in disguise. The map
f(x)=\begin{cases} 0.5x&forx\inZ,\\ 0.5+0.5x&forx\in(0,1)\setminusZ \end{cases}
is an isomorphism between
style((0,1),l{F},m)
\displaystyleZ1=\{0.5x:x\inZ\}\cup\{0.5+0.5x:x\in(0,1)\setminusZ\},
See also .
Standardness of a given probability space
style(\Omega,l{F},P)
stylef
style(\Omega,l{F},P)
style(X,\Sigma).
style(X,\Sigma)
stylef
style(X,\Sigma)
stylef
style(\Omega,l{F},P).
stylef:\Omega\toR,
stylef:\Omega\toRn,
stylef:\Omega\toRinfty,
style(A1,A2,...)
stylef:\Omega\to\{0,1\}infty.
Two conditions will be imposed on
stylef
stylef
The probability space
style(\Omega,l{F},P)
A measurable function
stylef:\Omega\toR
f*P
style\mu
styleR,
\displaystyle\mu(B)=(f*P)(B)=P(f-1(B))
styleB\subsetR.
f
stylef(\Omega)
\displaystyle\mu*(f(\Omega))=infB\mu(B)=infBP(f-1(B))=P(\Omega)=1,
stylef(\Omega)
style\mu.
A measurable function
stylef:\Omega\toR
stylel{F}
P
stylef-1(B),
styleB\subsetR
Caution. The following condition is not sufficient for
stylef
styleA\inl{F}
styleB\subsetR
styleP(An{\Delta}f-1(B))=0.
style\Delta
Theorem. Let a measurable function
stylef:\Omega\toR
\mu(stylef(\Omega))=1
stylef(\Omega)
(\Omega,l{F},P)
See also .
The same theorem holds for any
Rn
R
f:\Omega\toRn
X1,...,Xn:\Omega\toR,
f
l{F}
X1,...,Xn.
The theorem still holds for the space
Rinfty
f:\Omega\toRinfty
X1,X2,...:\Omega\toR,
f
l{F}
X1,X2,....
In particular, if the random variables
Xn
f:\Omega\to\{0,1\}infty
A1,A2,\ldots\inl{F}.
f
l{F}
A1,A2,....
In the pioneering work sequences
A1,A2,\ldots
f
(\Omega,l{F},P)
f(\Omega)
\mu,
The four cases treated above are mutually equivalent, and can be united, since the measurable spaces
R,
Rn,
Rinfty
\{0,1\}infty
Existence of an injective measurable function from
style(\Omega,l{F},P)
style(X,\Sigma)
style(X,\Sigma).
style(X,\Sigma)=\{0,1\}infty
Existence of a generating measurable function from
style(\Omega,l{F},P)
style(X,\Sigma)
style(X,\Sigma).
style(X,\Sigma)=\{0,1\}infty
Probability space | Countably separated | Countably generated | Standard | ||
---|---|---|---|---|---|
Interval with Lebesgue measure | |||||
Naive white noise | |||||
Perforated interval |
Every injective measurable function from a standard probability space to a standard measurable space is generating. See,, . This property does not hold for the non-standard probability space dealt with in the subsection "A superfluous measurable set" above.
Caution. The property of being countably generated is invariant under mod 0 isomorphisms, but the property of being countably separated is not. In fact, a standard probability space
style(\Omega,l{F},P)
style\Omega
Let
style(\Omega,l{F},P)
style\Omega
Definition.
style(\Omega,l{F},P)
See and . "Absolutely measurable" means: measurable in every countably separated, countably generated probability space containing it.
Definition.
style(\Omega,l{F},P)
See . "Perfect" means that for every measurable function from
style(\Omega,l{F},P)
R
stylel{F}
R
Definition.
style(\Omega,l{F},P)
style\tau
style\Omega
style(\Omega,\tau)
stylel{F}
style\tau
style\varepsilon>0
styleK
style(\Omega,\tau)
styleP(K)\ge1-\varepsilon.
See .
Every probability distribution on the space
styleRn
The same holds on every Polish space, see,,, and .
For example, the Wiener measure turns the Polish space
styleC[0,infty)
style[0,infty)\toR,
Another example: for every sequence of random variables, their joint distribution turns the Polish space
styleRinfty
(Thus, the idea of dimension, very natural for topological spaces, is utterly inappropriate for standard probability spaces.)
The product of two standard probability spaces is a standard probability space.
The same holds for the product of countably many spaces, see,, and .
A measurable subset of a standard probability space is a standard probability space. It is assumed that the set is not a null set, and is endowed with the conditional measure. See and .
Every probability measure on a standard Borel space turns it into a standard probability space.
In the discrete setup, the conditional probability is another probability measure, and the conditional expectation may be treated as the (usual) expectation with respect to the conditional measure, see conditional expectation. In the non-discrete setup, conditioning is often treated indirectly, since the condition may have probability 0, see conditional expectation. As a result, a number of well-known facts have special 'conditional' counterparts. For example: linearity of the expectation; Jensen's inequality (see conditional expectation); Hölder's inequality; the monotone convergence theorem, etc.
Given a random variable
styleY
style(\Omega,l{F},P)
stylePy
style\omega\in\Omega
styleY(\omega)=y
style(\Omega,l{F},P)
The conditional Jensen's inequality is just the (usual) Jensen's inequality applied to the conditional measure. The same holds for many other facts.
Given two probability spaces
style(\Omega1,l{F}1,P1)
style(\Omega2,l{F}2,P2)
stylef:\Omega1\to\Omega2
stylef(\Omega1)
style\Omega2
styleP2(f(\Omega1))
stylef(\Omega1)
style(\Omega1,l{F}1,P1)
style(\Omega2,l{F}2,P2)
styleP2(f(\Omega1))=1
stylef
styleA\inl{F}1
stylef(A)\inl{F}2
styleP2(f(A))=P1(A)
stylef-1
"There is a coherent way to ignore the sets of measure 0 in a measure space" . Striving to get rid of null sets, mathematicians often use equivalence classes of measurable sets or functions. Equivalence classes of measurable subsets of a probability space form a normed complete Boolean algebra called the measure algebra (or metric structure). Every measure preserving map
stylef:\Omega1\to\Omega2
styleF
styleF(B)=f-1(B)
styleB\inl{F}2
It may seem that every homomorphism of measure algebras has to correspond to some measure preserving map, but it is not so. However, for standard probability spaces each
styleF
stylef