You are on page 1of 15

Expected value

In probability theory, the expected value of a random variable , often denoted , , or , is a


generalization of the weighted average, and is intuitively the arithmetic mean of a large number of
independent realizations of . The expectation operator is also commonly stylized as or .[1][2][3]
The expected value is also known as the expectation, mathematical expectation, mean, average, or first
moment. Expected value is a key concept in economics, finance, and many other subjects.

By definition, the expected value of a constant random variable is .[4] The expected value of a
random variable with equiprobable outcomes is defined as the arithmetic mean of the
terms If some of the probabilities of an individual outcome are unequal, then the
expected value is defined to be the probability-weighted average of the s, that is, the sum of the
products .[5] The expected value of a general random variable is defined as an integral.

Contents
History
Etymology
Notations
Definition
Finite case
Countably infinite case
Absolutely continuous case
General case
Basic properties
Uses and applications
Interchanging limits and expectation
Inequalities
Expected values of common distributions[17]
Relationship with characteristic function
See also
References
Literature

History
The idea of the expected value originated in the middle of the 17th century from the study of the so-called
problem of points, which seeks to divide the stakes in a fair way between two players, who have to end
their game before it is properly finished.[6] This problem had been debated for centuries. Many conflicting
proposals and solutions had been suggested over the years when it was posed to Blaise Pascal by French
writer and amateur mathematician Chevalier de Méré in 1654. Méré claimed that this problem couldn't be
solved and that it showed just how flawed mathematics was when it came to its application to the real
world. Pascal, being a mathematician, was provoked and determined to solve the problem once and for all.

He began to discuss the problem in the famous series of letters to Pierre de Fermat. Soon enough, they both
independently came up with a solution. They solved the problem in different computational ways, but their
results were identical because their computations were based on the same fundamental principle. The
principle is that the value of a future gain should be directly proportional to the chance of getting it. This
principle seemed to have come naturally to both of them. They were very pleased by the fact that they had
found essentially the same solution, and this in turn made them absolutely convinced that they had solved
the problem conclusively; however, they did not publish their findings. They only informed a small circle of
mutual scientific friends in Paris about it.[7]

In Dutch mathematician Christiaan Huygen's book, he considered the problem of points, and presented a
solution based on the same principle as the solutions of Pascal and Fermat. Huygens published his treatise
in 1657, (see Huygens (1657)) "De ratiociniis in ludo aleæ" on probability theory just after visiting Paris.
The book extended the concept of expectation by adding rules for how to calculate expectations in more
complicated situations than the original problem (e.g., for three or more players), and can be seen as the first
successful attempt at laying down the foundations of the theory of probability.

In the foreword to his treatise, Huygens wrote:

It should be said, also, that for some time some of the best mathematicians of France have
occupied themselves with this kind of calculus so that no one should attribute to me the honour
of the first invention. This does not belong to me. But these savants, although they put each
other to the test by proposing to each other many questions difficult to solve, have hidden their
methods. I have had therefore to examine and go deeply for myself into this matter by
beginning with the elements, and it is impossible for me for this reason to affirm that I have
even started from the same principle. But finally I have found that my answers in many cases
do not differ from theirs.

— Edwards (2002)

During his visit to France in 1655, Huygens learned about de Méré's Problem. From his correspondence
with Carcavine a year later (in 1656), he realized his method was essentially the same as Pascal's.
Therefore, he knew about Pascal's priority in this subject before his book went to press in 1657.[8]

In the mid-nineteenth century, Pafnuty Chebyshev became the first person to think systematically in terms
of the expectations of random variables.[9]

Etymology

Neither Pascal nor Huygens used the term "expectation" in its modern sense. In particular, Huygens
writes:[10]

That any one Chance or Expectation to win any thing is worth just such a Sum, as wou'd
procure in the same Chance and Expectation at a fair Lay. ... If I expect a or b, and have an
equal chance of gaining them, my Expectation is worth (a+b)/2.
More than a hundred years later, in 1814, Pierre-Simon Laplace published his tract "Théorie analytique des
probabilités", where the concept of expected value was defined explicitly:[11]

… this advantage in the theory of chance is the product of the sum hoped for by the probability
of obtaining it; it is the partial sum which ought to result when we do not wish to run the risks
of the event in supposing that the division is made proportional to the probabilities. This
division is the only equitable one when all strange circumstances are eliminated; because an
equal degree of probability gives an equal right for the sum hoped for. We will call this
advantage mathematical hope.

Notations
The use of the letter to denote expected value goes back to W. A. Whitworth in 1901.[12] The symbol
has become popular since then for English writers. In German, stands for "Erwartungswert", in Spanish
for "Esperanza matemática", and in French for "Espérance mathématique".[13]

When E is used to denote expected value, authors use a variety of notation: the expectation operator can be
stylized as (upright), (italic), or (in blackboard bold), while round brackets ( ), square brackets
( ), or no brackets ( ) are all used.

Another popular notation is , whereas is commonly used in physics, and in Russian-


language literature.

Definition

Finite case

Let be a (discrete) random variable with a finite number of finite outcomes occurring
with probabilities respectively. The expectation of is defined as [5]

Since the expected value is the weighted sum of the values, with the
probabilities as the weights.

If all outcomes are equiprobable (that is, ), then the weighted average turns into
the simple average. On the other hand, if the outcomes are not equiprobable, then the simple average
must be replaced with the weighted average, which takes into account the fact that some outcomes are more
likely than others.

Examples
Let represent the outcome of a roll of a fair six-sided die. More specifically, will be the
number of pips showing on the top face of the die after the toss. The possible values for
are 1, 2, 3, 4, 5, and 6, all of which are equally likely with a probability of 16 . The expectation
of is
An illustration of the convergence of sequence averages of rolls of a
die to the expected value of 3.5 as the number of rolls (trials) grows.

If one rolls the die times and computes the average (arithmetic mean) of the results, then
as grows, the average will almost surely converge to the expected value, a fact known
as the strong law of large numbers.

The roulette game consists of a small ball and a wheel with 38 numbered pockets around
the edge. As the wheel is spun, the ball bounces around randomly until it settles down in
one of the pockets. Suppose random variable represents the (monetary) outcome of a $1
1
bet on a single number ("straight up" bet). If the bet wins (which happens with probability 38
in American roulette), the payoff is $35; otherwise the player loses the bet. The expected
profit from such a bet will be

That is, the bet of $1 stands to lose , so its expected value is

Countably infinite case

Intuitively, the expectation of a random variable taking values in a countable set of outcomes is defined
analogously as the weighted sum of the outcome values, where the weights correspond to the probabilities
of realizing that value. However, convergence issues associated with the infinite sum necessitate a more
careful definition. A rigorous definition first defines expectation of a non-negative random variable, and
then adapts it to general random variables.

Let be a non-negative random variable with a countable set of outcomes occurring with
probabilities respectively. Analogous to the discrete case, the expected value of is then
defined as the series
Note that since , the infinite sum is well-defined and does not depend on the order in which it is
computed. Unlike the finite case, the expectation here can be equal to infinity, if the infinite sum above
increases without bound.

For a general (not necessarily non-negative) random variable with a countable number of outcomes, set
and . By definition,

Like with non-negative random variables, can, once again, be finite or infinite. The third option here
is that is no longer guaranteed to be well defined at all. The latter happens whenever
.

Examples

Suppose and for , where (with being the natural

logarithm) is the scale factor such that the probabilities sum to 1. Then, using the direct
definition for non-negative random variables, we have

An example where the expectation is infinite arises in the context of the St. Petersburg
paradox. Let and for . Once again, since the random variable

is non-negative, the expected value calculation gives

For an example where the expectation is not well-defined, suppose the random variable
takes values with respective probabilities , ..., where

is a normalizing constant that ensures the probabilities sum up to one.

Then it follows that takes value with probability for


and takes value with remaining probability. Similarly, takes value with probability
for and takes value with remaining probability. Using the definition
for non-negative random variables, one can show that both and
(see Harmonic series). Hence, the expectation of is not well-defined.

Absolutely continuous case

If is a random variable with a probability density function of , then the expected value is defined as
the Lebesgue integral
where the values on both sides are well defined or not well defined simultaneously.

Example. A random variable that has the Cauchy distribution[14] has a density function, but the expected
value is undefined since the distribution has large "tails".

General case

In general, if is a random variable defined on a probability space , then the expected value of
, denoted by , is defined as the Lebesgue integral

For multidimensional random variables, their expected value is defined per component. That is,

and, for a random matrix with elements ,

Basic properties
The basic properties below (and their names in bold) replicate or follow immediately from those of
Lebesgue integral. Note that the letters "a.s." stand for "almost surely"—a central property of the Lebesgue
integral. Basically, one says that an inequality like is true almost surely, when the probability
measure attributes zero-mass to the complementary event .

For a general random variable , define as before and


, and note that , with both and
nonnegative, then:

Let denote the indicator function of an event , then

Formulas in terms of CDF: If is the cumulative distribution function of the probability


measure and is a random variable, then

where the values on both sides are well defined or not well defined simultaneously, and the
integral is taken in the sense of Lebesgue-Stieltjes. Here, is the extended
real line.
Additionally,

with the integrals taken in the sense of Lebesgue.


The proof of the second formula follows.

Proof

For an arbitrary

The last equality holds because the inequality where implies that
and hence Conversely, if then and

The integrand in the expression for is non-negative, so Tonelli's theorem applies,


and the order of integration may be switched:

Arguing as above,

and

Recalling that completes the proof.


Non-negativity: If (a.s.), then .
Linearity of expectation:[4]The expected value operator (or expectation operator) is
linear in the sense that, for any random variables and , and a constant ,

whenever the right-hand side is well-defined. This means that the expected value of the
sum of any finite number of random variables is the sum of the expected values of the
individual random variables, and the expected value scales linearly with a multiplicative
constant. Symbolically, for random variables and constants , we have
.

Monotonicity: If (a.s.), and both and exist, then .


Proof follows from the linearity and the non-negativity property for , since
(a.s.).
Non-multiplicativity: In general, the expected value is not multiplicative, i.e. is not
necessarily equal to . If and are independent, then one can show that
. If the random variables are dependent, then generally
, although in special cases of dependency the equality may hold.
Law of the unconscious statistician: The expected value of a measurable function of ,
, given that has a probability density function , is given by the inner product of
and : [4]

This formula also holds in multidimensional case, when is a function of several random
variables, and is their joint density.[4][15]
Non-degeneracy: If , then (a.s.).
For a random variable with well-defined expectation: .
The following statements regarding a random variable are equivalent:
exists and is finite.
Both and are finite.
is finite.

For the reasons above, the expressions " is integrable" and "the expected value of is
finite" are used interchangeably throughout this article.

If then (a.s.). Similarly, if then (a.s.).


If and then
If (a.s.), then . In other words, if X and Y are random variables that take
different values with probability zero, then the expectation of X will equal the expectation of
Y.
If (a.s.) for some constant , then . In particular, for a random
variable with well-defined expectation, . A well defined expectation
implies that there is one number, or rather, one constant that defines the expected value.
Thus follows that the expectation of this constant is just the original expected value.
For a non-negative integer-valued random variable

Proof

If then On the other hand,

so the series on the right diverges to and the equality holds.

If then

Define the infinite upper-triangular matrix

The double series is the sum of 's elements if summation is done


row by row. Since every summand is non-negative, the series either converges absolutely or
diverges to In both cases, changing summation order does not affect the sum.
Changing summation order, from row-by-row to column-by-column, gives us
Uses and applications
The expectation of a random variable plays an important role in a variety of contexts. For example, in
decision theory, an agent making an optimal choice in the context of incomplete information is often
assumed to maximize the expected value of their utility function. For a different example, in statistics,
where one seeks estimates for unknown parameters based on available data, the estimate itself is a random
variable. In such settings, a desirable criterion for a "good" estimator is that it is unbiased; that is, the
expected value of the estimate is equal to the true value of the underlying parameter.

It is possible to construct an expected value equal to the probability of an event, by taking the expectation
of an indicator function that is one if the event has occurred and zero otherwise. This relationship can be
used to translate properties of expected values into properties of probabilities, e.g. using the law of large
numbers to justify estimating probabilities by frequencies.

The expected values of the powers of X are called the moments of X; the moments about the mean of X are
expected values of powers of X − E[X]. The moments of some random variables can be used to specify
their distributions, via their moment generating functions.

To empirically estimate the expected value of a random variable, one repeatedly measures observations of
the variable and computes the arithmetic mean of the results. If the expected value exists, this procedure
estimates the true expected value in an unbiased manner and has the property of minimizing the sum of the
squares of the residuals (the sum of the squared differences between the observations and the estimate). The
law of large numbers demonstrates (under fairly mild conditions) that, as the size of the sample gets larger,
the variance of this estimate gets smaller.

This property is often exploited in a wide variety of applications, including general problems of statistical
estimation and machine learning, to estimate (probabilistic) quantities of interest via Monte Carlo methods,
since most quantities of interest can be written in terms of expectation, e.g. , where
is the indicator function of the set .

In classical mechanics, the center of mass is an analogous concept


to expectation. For example, suppose X is a discrete random
variable with values xi and corresponding probabilities pi. Now
consider a weightless rod on which are placed weights, at locations
xi along the rod and having masses pi (whose sum is one). The
point at which the rod balances is E[X].

Expected values can also be used to compute the variance, by


means of the computational formula for the variance
The mass of probability distribution
is balanced at the expected value,
here a Beta(α,β) distribution with
expected value α/(α+β).
A very important application of the expectation value is in the field
of quantum mechanics. The expectation value of a quantum
mechanical operator operating on a quantum state vector is written as . The
uncertainty in can be calculated using the formula .

Interchanging limits and expectation


In general, it is not the case that despite pointwise. Thus, one cannot
interchange limits and expectation, without additional conditions on the random variables. To see this, let
be a random variable distributed uniformly on . For define a sequence of random variables

with being the indicator function of the event . Then, it follows that (a.s). But,
for each . Hence,

Analogously, for general sequence of random variables , the expected value operator is not
-additive, i.e.

An example is easily obtained by setting and for , where is as in


the previous example.

A number of convergence results specify exact conditions which allow one to interchange limits and
expectations, as specified below.

Monotone convergence theorem: Let be a sequence of random variables, with


(a.s) for each . Furthermore, let pointwise. Then, the
monotone convergence theorem states that

Using the monotone convergence theorem, one can show that expectation indeed satisfies
countable additivity for non-negative random variables. In particular, let be non-
negative random variables. It follows from monotone convergence theorem that

Fatou's lemma: Let be a sequence of non-negative random variables.


Fatou's lemma states that

Corollary. Let with for all . If (a.s), then


Proof is by observing that (a.s.) and applying Fatou's lemma.
Dominated convergence theorem: Let be a sequence of random variables. If
pointwise (a.s.), (a.s.), and . Then, according to the
dominated convergence theorem,
;

Uniform integrability: In some cases, the equality holds when the


sequence is uniformly integrable.
Inequalities

There are a number of inequalities involving the expected values of functions of random variables. The
following list includes some of the more basic ones.

Markov's inequality: For a nonnegative random variable and , Markov's inequality


states that

Bienaymé-Chebyshev inequality: Let be an arbitrary random variable with finite expected


value and finite variance . The Bienaymé-Chebyshev inequality states that,
for any real number ,

Jensen's inequality: Let be a Borel convex function and a random variable


such that . Then

The right-hand side is well defined even if assumes non-finite values. Indeed, as noted
above, the finiteness of implies that is finite a.s.; thus is defined a.s..
Lyapunov's inequality:[16] Let . Lyapunov's inequality states that

Proof. Applying Jensen's inequality to and , obtain


. Taking the root of each side completes the proof.
Cauchy–Bunyakovsky–Schwarz inequality: The Cauchy–Bunyakovsky–Schwarz inequality
states that

Hölder's inequality: Let and satisfy , , and . The


Hölder's inequality states that

Minkowski inequality: Let be a positive real number satisfying . Let, in addition,


and . Then, according to the Minkowski inequality,
and

Expected values of common distributions [17]


Distribution Notation Mean E(X)

Bernoulli

Binomial

Poisson

Geometric

Uniform

Exponential

Normal

Standard Normal

Pareto for ; for

Cauchy undefined

Relationship with characteristic function


The probability density function of a scalar random variable is related to its characteristic function
by the inversion formula:

For the expected value of (where is a Borel function), we can use this inversion formula
to obtain

If is finite, changing the order of integration, we get, in accordance with Fubini–Tonelli theorem,

where

is the Fourier transform of The expression for also follows directly from Plancherel
theorem.

See also
Center of mass
Central tendency
Chebyshev's inequality (an inequality on location and scale parameters)
Conditional expectation
Expectation (the general term)
Expectation value (quantum mechanics)
Law of total expectation—the expected value of the conditional expected value of X given Y
is the same as the expected value of X.
Moment (mathematics)
Nonlinear expectation (a generalization of the expected value)
Sample mean
Population mean
Wald's equation—an equation for calculating the expected value of a random number of
random variables

References
1. "Expectation | Mean | Average" (https://www.probabilitycourse.com/chapter3/3_2_2_expecta
tion.php). www.probabilitycourse.com. Retrieved 2020-09-11.
2. Hansen, Bruce. "PROBABILITY AND STATISTICS FOR ECONOMISTS" (https://ssc.wisc.e
du/~bhansen/probability/Probability.pdf) (PDF). Retrieved 2021-07-20.
3. Wasserman, Larry (December 2010). All of Statistics: a concise course in statistical
inference. Springer texts in statistics. p. 47. ISBN 9781441923226.
4. Weisstein, Eric W. "Expectation Value" (https://mathworld.wolfram.com/ExpectationValue.ht
ml). mathworld.wolfram.com. Retrieved 2020-09-11.
5. "Expected Value | Brilliant Math & Science Wiki" (https://brilliant.org/wiki/expected-value/).
brilliant.org. Retrieved 2020-08-21.
6. History of Probability and Statistics and Their Applications before 1750. Wiley Series in
Probability and Statistics. 1990. doi:10.1002/0471725161 (https://doi.org/10.1002%2F04717
25161). ISBN 9780471725169.
7. Ore, Oystein (1960). "Ore, Pascal and the Invention of Probability Theory". The American
Mathematical Monthly. 67 (5): 409–419. doi:10.2307/2309286 (https://doi.org/10.2307%2F23
09286). JSTOR 2309286 (https://www.jstor.org/stable/2309286).
8. Mckay, Cain (2019). Probability and Statistics. p. 257. ISBN 9781839473302.
9. George Mackey (July 1980). "HARMONIC ANALYSIS AS THE EXPLOITATION OF
SYMMETRY - A HISTORICAL SURVEY". Bulletin of the American Mathematical Society.
New Series. 3 (1): 549.
10. Huygens, Christian. "The Value of Chances in Games of Fortune. English Translation" (http
s://math.dartmouth.edu/~doyle/docs/huygens/huygens.pdf) (PDF).
11. Laplace, Pierre Simon, marquis de, 1749-1827. (1952) [1951]. A philosophical essay on
probabilities. Dover Publications. OCLC 475539 (https://www.worldcat.org/oclc/475539).
12. Whitworth, W.A. (1901) Choice and Chance with One Thousand Exercises. Fifth edition.
Deighton Bell, Cambridge. [Reprinted by Hafner Publishing Co., New York, 1959.]
13. "Earliest uses of symbols in probability and statistics" (http://jeff560.tripod.com/stat.html).
14. Richard W Hamming (1991). "Example 8.7–1 The Cauchy distribution". The art of probability
for scientists and engineers (https://books.google.com/books?id=jX_F-77TA3gC&q=Cauch
y). Addison-Wesley. p. 290 ff. ISBN 0-201-40686-1. "Sampling from the Cauchy distribution
and averaging gets you nowhere — one sample has the same distribution as the average of
1000 samples!"
15. Papoulis, A. (1984), Probability, Random Variables, and Stochastic Processes, New York:
McGraw–Hill, pp. 139–152
16. Agahi, Hamzeh; Mohammadpour, Adel; Mesiar, Radko (November 2015). "Generalizations
of some probability inequalities and $L^{p}$ convergence of random variables for any
monotone measure" (https://doi.org/10.1214%2F14-BJPS251). Brazilian Journal of
Probability and Statistics. 29 (4): 878–896. doi:10.1214/14-BJPS251 (https://doi.org/10.121
4%2F14-BJPS251). ISSN 0103-0752 (https://www.worldcat.org/issn/0103-0752).
17. Selvamuthu, D (2018). Introduction to Statistical Methods, Design of Experiments and
Statistical Quality Control. Springer.

Literature
Edwards, A.W.F (2002). Pascal's arithmetical triangle: the story of a mathematical idea
(2nd ed.). JHU Press. ISBN 0-8018-6946-3.
Huygens, Christiaan (1657). De ratiociniis in ludo aleæ (http://www.york.ac.uk/depts/maths/h
iststat/huygens.pdf) (English translation, published in 1714).

Retrieved from "https://en.wikipedia.org/w/index.php?title=Expected_value&oldid=1062533414"

This page was last edited on 29 December 2021, at 01:35 (UTC).

Text is available under the Creative Commons Attribution-ShareAlike License; 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.

You might also like