Conditional expectation

Last updated

In probability theory, the conditional expectation, conditional expected value, or conditional mean of a random variable is its expected value evaluated with respect to the conditional probability distribution. If the random variable can take on only a finite number of values, the "conditions" are that the variable can only take on a subset of those values. More formally, in the case when the random variable is defined over a discrete probability space, the "conditions" are a partition of this probability space.

Contents

Depending on the context, the conditional expectation can be either a random variable or a function. The random variable is denoted analogously to conditional probability. The function form is either denoted or a separate function symbol such as is introduced with the meaning .

Examples

Example 1: Dice rolling

Consider the roll of a fair die and let A = 1 if the number is even (i.e., 2, 4, or 6) and A = 0 otherwise. Furthermore, let B = 1 if the number is prime (i.e., 2, 3, or 5) and B = 0 otherwise.

123456
A010101
B011010

The unconditional expectation of A is , but the expectation of A conditional on B = 1 (i.e., conditional on the die roll being 2, 3, or 5) is , and the expectation of A conditional on B = 0 (i.e., conditional on the die roll being 1, 4, or 6) is . Likewise, the expectation of B conditional on A = 1 is , and the expectation of B conditional on A = 0 is .

Example 2: Rainfall data

Suppose we have daily rainfall data (mm of rain each day) collected by a weather station on every day of the ten–year (3652-day) period from January 1, 1990, to December 31, 1999. The unconditional expectation of rainfall for an unspecified day is the average of the rainfall amounts for those 3652 days. The conditional expectation of rainfall for an otherwise unspecified day known to be (conditional on being) in the month of March, is the average of daily rainfall over all 310 days of the ten–year period that falls in March. And the conditional expectation of rainfall conditional on days dated March 2 is the average of the rainfall amounts that occurred on the ten days with that specific date.

History

The related concept of conditional probability dates back at least to Laplace, who calculated conditional distributions. It was Andrey Kolmogorov who, in 1933, formalized it using the Radon–Nikodym theorem. [1] In works of Paul Halmos [2] and Joseph L. Doob [3] from 1953, conditional expectation was generalized to its modern definition using sub-σ-algebras. [4]

Definitions

Conditioning on an event

If A is an event in with nonzero probability, and X is a discrete random variable, the conditional expectation of X given A is

where the sum is taken over all possible outcomes of X.

If , the conditional expectation is undefined due to the division by zero.

Discrete random variables

If X and Y are discrete random variables, the conditional expectation of X given Y is

where is the joint probability mass function of X and Y. The sum is taken over all possible outcomes of X.

Remark that as above the expression is undefined if .

Conditioning on a discrete random variable is the same as conditioning on the corresponding event:

where A is the set .

Continuous random variables

Let and be continuous random variables with joint density 's density and conditional density of given the event The conditional expectation of given is

When the denominator is zero, the expression is undefined.

Conditioning on a continuous random variable is not the same as conditioning on the event as it was in the discrete case. For a discussion, see Conditioning on an event of probability zero. Not respecting this distinction can lead to contradictory conclusions as illustrated by the Borel-Kolmogorov paradox.

L2 random variables

All random variables in this section are assumed to be in , that is square integrable. In its full generality, conditional expectation is developed without this assumption, see below under Conditional expectation with respect to a sub-σ-algebra. The theory is, however, considered more intuitive [5] and admits important generalizations. In the context of random variables, conditional expectation is also called regression.

In what follows let be a probability space, and in with mean and variance . The expectation minimizes the mean squared error:

.

The conditional expectation of X is defined analogously, except instead of a single number , the result will be a function . Let be a random vector. The conditional expectation is a measurable function such that

.

Note that unlike , the conditional expectation is not generally unique: there may be multiple minimizers of the mean squared error.

Uniqueness

Example 1: Consider the case where Y is the constant random variable that's always 1. Then the mean squared error is minimized by any function of the form

Example 2: Consider the case where Y is the 2-dimensional random vector . Then clearly

but in terms of functions it can be expressed as or or infinitely many other ways. In the context of linear regression, this lack of uniqueness is called multicollinearity.

Conditional expectation is unique up to a set of measure zero in . The measure used is the pushforward measure induced by Y.

In the first example, the pushforward measure is a Dirac distribution at 1. In the second it is concentrated on the "diagonal" , so that any set not intersecting it has measure 0.

Existence

The existence of a minimizer for is non-trivial. It can be shown that

is a closed subspace of the Hilbert space . [6] By the Hilbert projection theorem, the necessary and sufficient condition for to be a minimizer is that for all in M we have

.

In words, this equation says that the residual is orthogonal to the space M of all functions of Y. This orthogonality condition, applied to the indicator functions , is used below to extend conditional expectation to the case that X and Y are not necessarily in .

Connections to regression

The conditional expectation is often approximated in applied mathematics and statistics due to the difficulties in analytically calculating it, and for interpolation. [7]

The Hilbert subspace

defined above is replaced with subsets thereof by restricting the functional form of g, rather than allowing any measurable function. Examples of this are decision tree regression when g is required to be a simple function, linear regression when g is required to be affine, etc.

These generalizations of conditional expectation come at the cost of many of its properties no longer holding. For example, let M be the space of all linear functions of Y and let denote this generalized conditional expectation/ projection. If does not contain the constant functions, the tower property will not hold.

An important special case is when X and Y are jointly normally distributed. In this case it can be shown that the conditional expectation is equivalent to linear regression:

for coefficients described in Multivariate normal distribution#Conditional distributions.

Conditional expectation with respect to a sub-σ-algebra

Conditional expectation with respect to a s-algebra: in this example the probability space
(
O
,
F
,
P
)
{\displaystyle (\Omega ,{\mathcal {F}},P)}
is the [0,1] interval with the Lebesgue measure. We define the following s-algebras:
A
=
F
{\displaystyle {\mathcal {A}}={\mathcal {F}}}
;
B
{\displaystyle {\mathcal {B}}}
is the s-algebra generated by the intervals with end-points 0,
.mw-parser-output .frac{white-space:nowrap}.mw-parser-output .frac .num,.mw-parser-output .frac .den{font-size:80%;line-height:0;vertical-align:super}.mw-parser-output .frac .den{vertical-align:sub}.mw-parser-output .sr-only{border:0;clip:rect(0,0,0,0);clip-path:polygon(0px 0px,0px 0px,0px 0px);height:1px;margin:-1px;overflow:hidden;padding:0;position:absolute;width:1px}
1/4,
1/2,
3/4, 1; and
C
{\displaystyle {\mathcal {C}}}
is the s-algebra generated by the intervals with end-points 0,
1/2, 1. Here the conditional expectation is effectively the average over the minimal sets of the s-algebra. LokaleMittelwertbildung.svg
Conditional expectation with respect to a σ-algebra: in this example the probability space is the [0,1] interval with the Lebesgue measure. We define the following σ-algebras: ; is the σ-algebra generated by the intervals with end-points 0, 14, 12, 34, 1; and is the σ-algebra generated by the intervals with end-points 0, 12, 1. Here the conditional expectation is effectively the average over the minimal sets of the σ-algebra.

Consider the following:

Since is a sub -algebra of , the function is usually not -measurable, thus the existence of the integrals of the form , where and is the restriction of to , cannot be stated in general. However, the local averages can be recovered in with the help of the conditional expectation.

A conditional expectation of X given , denoted as , is any -measurable function which satisfies:

for each . [8]

As noted in the discussion, this condition is equivalent to saying that the residual is orthogonal to the indicator functions :

Existence

The existence of can be established by noting that for is a finite measure on that is absolutely continuous with respect to . If is the natural injection from to , then is the restriction of to and is the restriction of to . Furthermore, is absolutely continuous with respect to , because the condition

implies

Thus, we have

where the derivatives are Radon–Nikodym derivatives of measures.

Conditional expectation with respect to a random variable

Consider, in addition to the above,

  • A measurable space , and
  • A random variable .

The conditional expectation of X given Y is defined by applying the above construction on the σ-algebra generated by Y:

.

By the Doob-Dynkin lemma, there exists a function such that

.

Discussion

  • This is not a constructive definition; we are merely given the required property that a conditional expectation must satisfy.
    • The definition of may resemble that of for an event but these are very different objects. The former is a -measurable function , while the latter is an element of and for .
    • Uniqueness can be shown to be almost sure: that is, versions of the same conditional expectation will only differ on a set of probability zero.
  • The σ-algebra controls the "granularity" of the conditioning. A conditional expectation over a finer (larger) σ-algebra retains information about the probabilities of a larger class of events. A conditional expectation over a coarser (smaller) σ-algebra averages over more events.

Conditional probability

For a Borel subset B in , one can consider the collection of random variables

.

It can be shown that they form a Markov kernel, that is, for almost all , is a probability measure. [9]

The Law of the unconscious statistician is then

.

This shows that conditional expectations are, like their unconditional counterparts, integrations, against a conditional measure.

General Definition

In full generality, consider:

The conditional expectation of given is the up to a -nullset unique and integrable -valued -measurable random variable satisfying

for all . [10] [11]

In this setting the conditional expectation is sometimes also denoted in operator notation as .

Basic properties

All the following formulas are to be understood in an almost sure sense. The σ-algebra could be replaced by a random variable , i.e. .

Proof

Let . Then is independent of , so we get that

Thus the definition of conditional expectation is satisfied by the constant random variable , as desired.

Proof

For each we have , or equivalently

Since this is true for each , and both and are -measurable (the former property holds by definition; the latter property is key here), from this one can show

And this implies almost everywhere.

Proof

All random variables here are assumed without loss of generality to be non-negative. The general case can be treated with .

Fix and let . Then for any

Hence almost everywhere.

Any simple function is a finite linear combination of indicator functions. By linearity the above property holds for simple functions: if is a simple function then .

Now let be -measurable. Then there exists a sequence of simple functions converging monotonically (here meaning ) and pointwise to . Consequently, for , the sequence converges monotonically and pointwise to .

Also, since , the sequence converges monotonically and pointwise to

Combining the special case proved for simple functions, the definition of conditional expectation, and deploying the monotone convergence theorem:

This holds for all , whence almost everywhere.

See also

Probability laws

Notes

  1. Kolmogorov, Andrey (1933). Grundbegriffe der Wahrscheinlichkeitsrechnung (in German). Berlin: Julius Springer. p. 46.
  2. Oxtoby, J. C. (1953). "Review: Measure theory, by P. R. Halmos" (PDF). Bull. Amer. Math. Soc. 59 (1): 89–91. doi: 10.1090/s0002-9904-1953-09662-8 .
  3. J. L. Doob (1953). Stochastic Processes. John Wiley & Sons. ISBN   0-471-52369-0.
  4. Olav Kallenberg: Foundations of Modern Probability. 2. edition. Springer, New York 2002, ISBN   0-387-95313-2, p. 573.
  5. "probability - Intuition behind Conditional Expectation". Mathematics Stack Exchange.
  6. Brockwell, Peter J. (1991). Time series : theory and methods (2nd ed.). New York: Springer-Verlag. ISBN   978-1-4419-0320-4.
  7. Hastie, Trevor. The elements of statistical learning : data mining, inference, and prediction (PDF) (Second, corrected 7th printing ed.). New York. ISBN   978-0-387-84858-7.
  8. Billingsley, Patrick (1995). "Section 34. Conditional Expectation". Probability and Measure (3rd ed.). John Wiley & Sons. p. 445. ISBN   0-471-00710-2.
  9. Klenke, Achim. Probability theory : a comprehensive course (Second ed.). London. ISBN   978-1-4471-5361-0.
  10. Da Prato, Giuseppe; Zabczyk, Jerzy (2014). Stochastic Equations in Infinite Dimensions. Cambridge University Press. p. 26. doi:10.1017/CBO9781107295513. (Definition in separable Banach spaces)
  11. Hytönen, Tuomas; van Neerven, Jan; Veraar, Mark; Weis, Lutz (2016). Analysis in Banach Spaces, Volume I: Martingales and Littlewood-Paley Theory. Springer Cham. doi:10.1007/978-3-319-48520-1. (Definition in general Banach spaces)
  12. "Conditional expectation". www.statlect.com. Retrieved 2020-09-11.
  13. Kallenberg, Olav (2001). Foundations of Modern Probability (2nd ed.). York, PA, USA: Springer. p. 110. ISBN   0-387-95313-2.

Related Research Articles

<span class="mw-page-title-main">Expected value</span> Average value of a random variable

In probability theory, the expected value is a generalization of the weighted average. Informally, the expected value is the arithmetic mean of the possible values a random variable can take, weighted by the probability of those outcomes. Since it is obtained through arithmetic, the expected value sometimes may not even be included in the sample data set; it is not the value you would "expect" to get in reality.

<span class="mw-page-title-main">Random variable</span> Variable representing a random phenomenon

A random variable is a mathematical formalization of a quantity or object which depends on random events. The term 'random variable' can be misleading as its mathematical definition is not actually random nor a variable, but rather it is a function from possible outcomes in a sample space to a measurable space, often to the real numbers.

<span class="mw-page-title-main">Independence (probability theory)</span> When the occurrence of one event does not affect the likelihood of another

Independence is a fundamental notion in probability theory, as in statistics and the theory of stochastic processes. Two events are independent, statistically independent, or stochastically independent if, informally speaking, the occurrence of one does not affect the probability of occurrence of the other or, equivalently, does not affect the odds. Similarly, two random variables are independent if the realization of one does not affect the probability distribution of the other.

<span class="mw-page-title-main">Martingale (probability theory)</span> Model in probability theory

In probability theory, a martingale is a sequence of random variables for which, at a particular time, the conditional expectation of the next value in the sequence is equal to the present value, regardless of all prior values.

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.

<span class="mw-page-title-main">Jensen's inequality</span> Theorem of convex functions

In mathematics, Jensen's inequality, named after the Danish mathematician Johan Jensen, relates the value of a convex function of an integral to the integral of the convex function. It was proved by Jensen in 1906, building on an earlier proof of the same inequality for doubly-differentiable functions by Otto Hölder in 1889. Given its generality, the inequality appears in many forms depending on the context, some of which are presented below. In its simplest form the inequality states that the convex transformation of a mean is less than or equal to the mean applied after convex transformation; it is a simple corollary that the opposite is true of concave transformations.

<span class="mw-page-title-main">Markov property</span> Memoryless property of a stochastic process

In probability theory and statistics, the term Markov property refers to the memoryless property of a stochastic process, which means that its future evolution is independent of its history. It is named after the Russian mathematician Andrey Markov. The term strong Markov property is similar to the Markov property, except that the meaning of "present" is defined in terms of a random variable known as a stopping time.

The proposition in probability theory known as the law of total expectation, the law of iterated expectations (LIE), Adam's law, the tower rule, and the smoothing theorem, among other names, states that if is a random variable whose expected value is defined, and is any random variable on the same probability space, then

In probability theory and statistics, the conditional probability distribution is a probability distribution that describes the probability of an outcome given the occurrence of a particular event. Given two jointly distributed random variables and , the conditional probability distribution of given is the probability distribution of when is known to be a particular value; in some cases the conditional probabilities may be expressed as functions containing the unspecified value of as a parameter. When both and are categorical variables, a conditional probability table is typically used to represent the conditional probability. The conditional distribution contrasts with the marginal distribution of a random variable, which is its distribution without reference to the value of the other variable.

In mathematics, a π-system on a set is a collection of certain subsets of such that

This article discusses how information theory is related to measure theory.

In mathematics, uniform integrability is an important concept in real analysis, functional analysis and measure theory, and plays a vital role in the theory of martingales.

In probability theory, a random measure is a measure-valued random element. Random measures are for example used in the theory of random processes, where they form many important point processes such as Poisson point processes and Cox processes.

In probability theory, a standard probability space, also called Lebesgue–Rokhlin probability space or just Lebesgue space 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.

In mathematics – specifically, in the theory of stochastic processes – Doob's martingale convergence theorems are a collection of results on the limits of supermartingales, named after the American mathematician Joseph L. Doob. Informally, the martingale convergence theorem typically refers to the result that any supermartingale satisfying a certain boundedness condition must converge. One may think of supermartingales as the random variable analogues of non-increasing sequences; from this perspective, the martingale convergence theorem is a random variable analogue of the monotone convergence theorem, which states that any bounded monotone sequence converges. There are symmetric results for submartingales, which are analogous to non-decreasing sequences.

<span class="mw-page-title-main">Conditional mutual information</span> Information theory

In probability theory, particularly information theory, the conditional mutual information is, in its most basic form, the expected value of the mutual information of two random variables given the value of a third.

In probability theory, regular conditional probability is a concept that formalizes the notion of conditioning on the outcome of a random variable. The resulting conditional probability distribution is a parametrized family of probability measures called a Markov kernel.

In probability theory, a Markov kernel is a map that in the general theory of Markov processes plays the role that the transition matrix does in the theory of Markov processes with a finite state space.

In probability theory, the Doob–Dynkin lemma, named after Joseph L. Doob and Eugene Dynkin, characterizes the situation when one random variable is a function of another by the inclusion of the -algebras generated by the random variables. The usual statement of the lemma is formulated in terms of one random variable being measurable with respect to the -algebra generated by the other.

In probability theory, the g-expectation is a nonlinear expectation based on a backwards stochastic differential equation (BSDE) originally developed by Shige Peng.

References