Expected value






In probability theory, the expected value of a random variable, intuitively, is the long-run average value of repetitions of the experiment it represents. For example, the expected value in rolling a six-sided die is 3.5, because the average of all the numbers that come up in an extremely large number of rolls is close to 3.5. Less roughly, the law of large numbers states that the arithmetic mean of the values almost surely converges to the expected value as the number of repetitions approaches infinity. The expected value is also known as the expectation, mathematical expectation, EV, average, mean value, mean, or first moment.


More practically, the expected value of a discrete random variable is the probability-weighted average of all possible values. In other words, each possible value the random variable can assume is multiplied by its probability of occurring, and the resulting products are summed to produce the expected value. The same principle applies to an absolutely continuous random variable, except that an integral of the variable with respect to its probability density replaces the sum. The formal definition subsumes both of these and also works for distributions which are neither discrete nor absolutely continuous; the expected value of a random variable is the integral of the random variable with respect to its probability measure.[1][2]


The expected value does not exist for random variables having some distributions with large "tails", such as the Cauchy distribution.[3] For random variables such as these, the long-tails of the distribution prevent the sum or integral from converging.


The expected value is a key aspect of how one characterizes a probability distribution; it is one type of location parameter. By contrast, the variance is a measure of dispersion of the possible values of the random variable around the expected value. The variance itself is defined in terms of two expectations: it is the expected value of the squared deviation of the variable's value from the variable's expected value.


The expected value plays important roles in a variety of contexts. In regression analysis, one desires a formula in terms of observed data that will give a "good" estimate of the parameter giving the effect of some explanatory variable upon a dependent variable. The formula will give different estimates using different samples of data, so the estimate it gives is itself a random variable. A formula is typically considered good in this context if it is an unbiased estimator— that is if the expected value of the estimate (the average value it would give over an arbitrarily large number of separate samples) can be shown to equal the true value of the desired parameter.


In decision theory, and in particular in choice under uncertainty, an agent is described as making an optimal choice in the context of incomplete information. For risk neutral agents, the choice involves using the expected values of uncertain quantities, while for risk averse agents it involves maximizing the expected value of some objective function such as a von Neumann–Morgenstern utility function. One example of using expected value in reaching optimal decisions is the Gordon–Loeb model of information security investment. According to the model, one can conclude that the amount a firm spends to protect information should generally be only a small fraction of the expected loss (i.e., the expected value of the loss resulting from a cyber or information security breach).[4]




Contents






  • 1 Definition


    • 1.1 Finite case


      • 1.1.1 Examples




    • 1.2 Countably infinite case


      • 1.2.1 Example




    • 1.3 Absolutely continuous case


    • 1.4 General case




  • 2 Basic properties


    • 2.1 E⁡[1A]=P⁡(A){displaystyle operatorname {E} [{mathbf {1} }_{A}]=operatorname {P} (A)}{displaystyle operatorname {E} [{mathbf {1} }_{A}]=operatorname {P} (A)}


    • 2.2 If X=Y{displaystyle X=Y}X=Y (a.s.) then E⁡[X]=E⁡[Y]{displaystyle operatorname {E} [X]=operatorname {E} [Y]}{displaystyle operatorname {E} [X]=operatorname {E} [Y]}


    • 2.3 Expected value of a constant


    • 2.4 Linearity


    • 2.5 E⁡[X]{displaystyle operatorname {E} [X]}operatorname {E} [X] is finite if and only if E⁡|X|{displaystyle operatorname {E} |X|}{displaystyle operatorname {E} |X|} is


    • 2.6 If X≥0{displaystyle Xgeq 0}{displaystyle Xgeq 0} (a.s.) then E⁡[X]≥0{displaystyle operatorname {E} [X]geq 0}{displaystyle operatorname {E} [X]geq 0}


    • 2.7 Monotonicity


    • 2.8 If |X|≤Y{displaystyle |X|leq Y}{displaystyle |X|leq Y} (a.s.) and E⁡[Y]{displaystyle operatorname {E} [Y]}{displaystyle operatorname {E} [Y]} is finite then so is E⁡[X]{displaystyle operatorname {E} [X]}operatorname {E} [X]


    • 2.9 If E⁡|Xβ|<∞{displaystyle operatorname {E} |X^{beta }|<infty }{displaystyle operatorname {E} |X^{beta }|<infty } and 0<α{displaystyle 0<alpha <beta }{displaystyle 0<alpha <beta } then E⁡|Xα|<∞{displaystyle operatorname {E} |X^{alpha }|<infty }{displaystyle operatorname {E} |X^{alpha }|<infty }


      • 2.9.1 Counterexample for infinite measure




    • 2.10 Extremal property


    • 2.11 Non-degeneracy


    • 2.12 If E⁡[X]<+∞{displaystyle operatorname {E} [X]<+infty }{displaystyle operatorname {E} [X]<+infty } then X<+∞{displaystyle X<+infty }{displaystyle X<+infty } (a.s.)


      • 2.12.1 Corollary: if E⁡[X]>−{displaystyle operatorname {E} [X]>-infty }{displaystyle operatorname {E} [X]>-infty } then X>−{displaystyle X>-infty }{displaystyle X>-infty } (a.s.)


      • 2.12.2 Corollary: if E⁡|X|<∞{displaystyle operatorname {E} |X|<infty }{displaystyle operatorname {E} |X|<infty } then X≠±{displaystyle Xneq pm infty }{displaystyle Xneq pm infty } (a.s.)




    • 2.13 |E⁡[X]|≤E⁡|X|{displaystyle |operatorname {E} [X]|leq operatorname {E} |X|}{displaystyle |operatorname {E} [X]|leq operatorname {E} |X|}


    • 2.14 Non-multiplicativity


    • 2.15 Counterexample: E⁡[Xi]↛E⁡[X]{displaystyle operatorname {E} [X_{i}]not to operatorname {E} [X]}{displaystyle operatorname {E} [X_{i}]not to operatorname {E} [X]} despite Xi→X{displaystyle X_{i}to X}{displaystyle X_{i}to X} pointwise


    • 2.16 Countable non-additivity


    • 2.17 Countable additivity for non-negative random variables




  • 3 E⁡[XY]=E⁡[X]E⁡[Y]{displaystyle operatorname {E} [XY]=operatorname {E} [X]operatorname {E} [Y]}{displaystyle operatorname {E} [XY]=operatorname {E} [X]operatorname {E} [Y]} for independent X{displaystyle X}X and Y{displaystyle Y}Y


  • 4 Inequalities


    • 4.1 Cauchy–Bunyakovsky–Schwarz inequality


    • 4.2 Markov's inequality


    • 4.3 Bienaymé-Chebyshev inequality


    • 4.4 Jensen's inequality


    • 4.5 Lyapunov’s inequality


    • 4.6 Hölder’s inequality


    • 4.7 Minkowski inequality




  • 5 Taking limits under the E{displaystyle operatorname {E} }operatorname {E} sign


    • 5.1 Monotone convergence theorem


    • 5.2 Fatou's lemma


    • 5.3 Dominated convergence theorem




  • 6 Relationship with characteristic function


  • 7 Uses and applications


  • 8 The law of the unconscious statistician


  • 9 Alternative formula for expected value


    • 9.1 Formula for non-negative random variables


      • 9.1.1 Finite and countably infinite case


        • 9.1.1.1 Example




      • 9.1.2 General case




    • 9.2 Formula for non-positive random variables


    • 9.3 General case




  • 10 History


  • 11 See also


  • 12 Notes


  • 13 Literature





Definition



Finite case


Let X{displaystyle X}X be a random variable with a finite number of finite outcomes x1{displaystyle x_{1}}x_{1}, x2{displaystyle x_{2}}x_{2}, ..., xk{displaystyle x_{k}}x_{k} occurring with probabilities p1{displaystyle p_{1}}p_{1}, p2{displaystyle p_{2}}p_{2}, ..., pk{displaystyle p_{k}}p_{k}, respectively. The expectation of X{displaystyle X}X is defined as



E⁡[X]=∑i=1kxipi=x1p1+x2p2+⋯+xkpk{displaystyle operatorname {E} [X]=sum _{i=1}^{k}x_{i},p_{i}=x_{1}p_{1}+x_{2}p_{2}+cdots +x_{k}p_{k}}{displaystyle operatorname {E} [X]=sum _{i=1}^{k}x_{i},p_{i}=x_{1}p_{1}+x_{2}p_{2}+cdots +x_{k}p_{k}}.

Since all probabilities pi{displaystyle p_{i}}p_{i} add up to 1 (p1+p2+…+pk=1{displaystyle p_{1}+p_{2}+ldots +p_{k}=1}{displaystyle p_{1}+p_{2}+ldots +p_{k}=1}), the expected value is the weighted average, with pi{displaystyle p_{i}}p_{i}’s being the weights.


If all outcomes xi{displaystyle x_{i}}x_{i} are equiprobable (that is, p1=p2=…=pk{displaystyle p_{1}=p_{2}=ldots =p_{k}}{displaystyle p_{1}=p_{2}=ldots =p_{k}}), then the weighted average turns into the simple average. This is intuitive: the expected value of a random variable is the average of all values it can take; thus the expected value is what one expects to happen on average. If the outcomes xi{displaystyle x_{i}}x_{i} 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 the others. The intuition however remains the same: the expected value of X{displaystyle X}X is what one expects to happen on average.




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.



Examples


  • Let X{displaystyle X}X represent the outcome of a roll of a fair six-sided die. More specifically, X{displaystyle X}X will be the number of pips showing on the top face of the die after the toss. The possible values for X{displaystyle X}X are 1, 2, 3, 4, 5, and 6, all equally likely (each having the probability of 1/6). The expectation of X{displaystyle X}X is

E⁡[X]=1⋅16+2⋅16+3⋅16+4⋅16+5⋅16+6⋅16=3.5.{displaystyle operatorname {E} [X]=1cdot {frac {1}{6}}+2cdot {frac {1}{6}}+3cdot {frac {1}{6}}+4cdot {frac {1}{6}}+5cdot {frac {1}{6}}+6cdot {frac {1}{6}}=3.5.}{displaystyle operatorname {E} [X]=1cdot {frac {1}{6}}+2cdot {frac {1}{6}}+3cdot {frac {1}{6}}+4cdot {frac {1}{6}}+5cdot {frac {1}{6}}+6cdot {frac {1}{6}}=3.5.}

If one rolls the die n{displaystyle n}n times and computes the average (arithmetic mean) of the results, then as n{displaystyle n}n grows, the average will almost surely converge to the expected value, a fact known as the strong law of large numbers. One example sequence of ten rolls of the die is 2, 3, 1, 2, 5, 6, 2, 2, 2, 6, which has the average of 3.1, with the distance of 0.4 from the expected value of 3.5. The convergence is relatively slow: the probability that the average falls within the range 3.5 ± 0.1 is 21.6% for ten rolls, 46.1% for a hundred rolls and 93.7% for a thousand rolls. See the figure for an illustration of the averages of longer sequences of rolls of the die and how they converge to the expected value of 3.5. More generally, the rate of convergence can be roughly quantified by e.g. Chebyshev's inequality and the Berry–Esseen theorem.

  • 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 X{displaystyle X}X represents the (monetary) outcome of a $1 bet on a single number ("straight up" bet). If the bet wins (which happens with probability 1/38 in American roulette), the payoff is $35; otherwise the player loses the bet. The expected profit from such a bet will be

E⁡[gain from $1 bet]=−$1⋅3738+$35⋅138=−$0.0526.{displaystyle operatorname {E} [,{text{gain from }}$1{text{ bet}},]=-$1cdot {frac {37}{38}}+$35cdot {frac {1}{38}}=-$0.0526.}{displaystyle operatorname {E} [,{text{gain from }}$1{text{ bet}},]=-$1cdot {frac {37}{38}}+$35cdot {frac {1}{38}}=-$0.0526.}

That is, the bet of $1 stands to lose $0.0526, so its expected value is -$0.0526.


Countably infinite case


Let X{displaystyle X}X be a random variable with a countable set of finite outcomes x1{displaystyle x_{1}}x_{1}, x2{displaystyle x_{2}}x_{2}, ..., occurring with probabilities p1{displaystyle p_{1}}p_{1}, p2{displaystyle p_{2}}p_{2}, ..., respectively,
such that the infinite sum i=1∞|xi|pi{displaystyle textstyle sum _{i=1}^{infty }|x_{i}|,p_{i}}{displaystyle textstyle sum _{i=1}^{infty }|x_{i}|,p_{i}} converges. The expected value of X{displaystyle X}X is defined as the series



E⁡[X]=∑i=1∞xipi{displaystyle operatorname {E} [X]=sum _{i=1}^{infty }x_{i},p_{i}}{displaystyle operatorname {E} [X]=sum _{i=1}^{infty }x_{i},p_{i}}.

Remark 1. Observe that |E⁡[X]|≤i=1∞|xi|pi<∞{displaystyle textstyle {Bigl |}operatorname {E} [X]{Bigr |}leq sum _{i=1}^{infty }|x_{i}|,p_{i}<infty }{displaystyle textstyle {Bigl |}operatorname {E} [X]{Bigr |}leq sum _{i=1}^{infty }|x_{i}|,p_{i}<infty }.


Remark 2. Due to absolute convergence, the expected value does not depend on the order in which the outcomes are presented. By contrast, a conditionally convergent series can be made to converge or diverge arbitrarily, via the Riemann rearrangement theorem.



Example


  • Suppose xi=i{displaystyle x_{i}=i}{displaystyle x_{i}=i} and pi=ki2i,{displaystyle p_{i}={frac {k}{i2^{i}}},}{displaystyle p_{i}={frac {k}{i2^{i}}},} for i=1,2,3,…{displaystyle i=1,2,3,ldots }{displaystyle i=1,2,3,ldots }, where k=1ln⁡2{displaystyle k={frac {1}{ln 2}}}{displaystyle k={frac {1}{ln 2}}} (with ln{displaystyle ln }ln being the natural logarithm) is the scale factor such that the probabilities sum to 1. Then


E⁡[X]=1(k2)+2(k8)+3(k24)+⋯=k2+k4+k8+⋯=k.{displaystyle operatorname {E} [X]=1left({frac {k}{2}}right)+2left({frac {k}{8}}right)+3left({frac {k}{24}}right)+dots ={frac {k}{2}}+{frac {k}{4}}+{frac {k}{8}}+dots =k.}{displaystyle operatorname {E} [X]=1left({frac {k}{2}}right)+2left({frac {k}{8}}right)+3left({frac {k}{24}}right)+dots ={frac {k}{2}}+{frac {k}{4}}+{frac {k}{8}}+dots =k.}

Since this series converges absolutely, the expected value of X{displaystyle X}X is k{displaystyle k}k.


  • For an example that is not absolutely convergent, suppose random variable X{displaystyle X}X takes values 1, −2, 3, −4, ..., with respective probabilities c12,c22,c32,c42{displaystyle {frac {c}{1^{2}}},{frac {c}{2^{2}}},{frac {c}{3^{2}}},{frac {c}{4^{2}}}}{displaystyle {frac {c}{1^{2}}},{frac {c}{2^{2}}},{frac {c}{3^{2}}},{frac {c}{4^{2}}}}, ..., where c=6π2{displaystyle c={frac {6}{pi ^{2}}}}{displaystyle c={frac {6}{pi ^{2}}}} is a normalizing constant that ensures the probabilities sum up to one. Then the infinite sum


i=1∞xipi=c(1−12+13−14+⋯){displaystyle sum _{i=1}^{infty }x_{i},p_{i}=c,{bigg (}1-{frac {1}{2}}+{frac {1}{3}}-{frac {1}{4}}+dotsb {bigg )}}sum _{i=1}^{infty }x_{i},p_{i}=c,{bigg (}1-{frac {1}{2}}+{frac {1}{3}}-{frac {1}{4}}+dotsb {bigg )}

converges and its sum is equal to 6ln⁡2≈0.421383{displaystyle {frac {6ln 2}{pi ^{2}}}approx 0.421383}{displaystyle {frac {6ln 2}{pi ^{2}}}approx 0.421383}. However it would be incorrect to claim that the expected value of X{displaystyle X}X is equal to this number—in fact E⁡[X]{displaystyle operatorname {E} [X]}operatorname {E} [X] does not exist (finite or infinite), as this series does not converge absolutely (see Alternating harmonic series).


  • An example that diverges arises in the context of the St. Petersburg paradox. Let xi=2i{displaystyle x_{i}=2^{i}}{displaystyle x_{i}=2^{i}} and pi=12i{displaystyle p_{i}={frac {1}{2^{i}}}}{displaystyle p_{i}={frac {1}{2^{i}}}} for i=1,2,3,…{displaystyle i=1,2,3,ldots }{displaystyle i=1,2,3,ldots }. The expected value calculation gives


i=1∞xipi=2⋅12+4⋅14+8⋅18+16⋅116+⋯=1+1+1+1+⋯.{displaystyle sum _{i=1}^{infty }x_{i},p_{i}=2cdot {frac {1}{2}}+4cdot {frac {1}{4}}+8cdot {frac {1}{8}}+16cdot {frac {1}{16}}+cdots =1+1+1+1+cdots ,.}{displaystyle sum _{i=1}^{infty }x_{i},p_{i}=2cdot {frac {1}{2}}+4cdot {frac {1}{4}}+8cdot {frac {1}{8}}+16cdot {frac {1}{16}}+cdots =1+1+1+1+cdots ,.}

Since this does not converge but instead keeps growing, the expected value is infinite.



Absolutely continuous case


If X{displaystyle X}X is a random variable whose cumulative distribution function admits a density f(x){displaystyle f(x)}f(x), then the expected value is defined as the following Lebesgue integral:


E⁡[X]=∫Rxf(x)dx.{displaystyle operatorname {E} [X]=int _{mathbb {R} }xf(x),dx.}{displaystyle operatorname {E} [X]=int _{mathbb {R} }xf(x),dx.}

Remark. From computational perspective, the integral in the definition of E⁡[X]{displaystyle operatorname {E} [X]}operatorname {E} [X] may often be treated as an improper Riemann integral +∞xf(x)dx.{displaystyle textstyle int _{-infty }^{+infty }xf(x),dx.}{displaystyle textstyle int _{-infty }^{+infty }xf(x),dx.} Specifically, if the function xf(x){displaystyle xf(x)}{displaystyle xf(x)} is Riemann-integrable on every finite interval [a,b]{displaystyle [a,b]}[a,b], and


min((−1)⋅(R)∫0xf(x)dx, (R)∫0+∞xf(x)dx)<∞,{displaystyle min left((-1)cdot {hbox{(R)}}int _{-infty }^{0}xf(x),dx, {hbox{(R)}}int _{0}^{+infty }xf(x),dxright)<infty ,}{displaystyle min left((-1)cdot {hbox{(R)}}int _{-infty }^{0}xf(x),dx, {hbox{(R)}}int _{0}^{+infty }xf(x),dxright)<infty ,}

then the values (whether finite or infinite) of both integrals agree.



General case


In general, if X{displaystyle X}X is a random variable defined on a probability space ,P){displaystyle (Omega ,Sigma ,operatorname {P} )}{displaystyle (Omega ,Sigma ,operatorname {P} )}, then the expected value of X{displaystyle X}X, denoted by E⁡[X]{displaystyle operatorname {E} [X]}operatorname {E} [X], X⟩{displaystyle langle Xrangle }{displaystyle langle Xrangle }, or {displaystyle {bar {X}}}{bar {X}}, is defined as the Lebesgue integral


E⁡[X]=∫ΩX(ω)dP⁡).{displaystyle operatorname {E} [X]=int _{Omega }X(omega ),doperatorname {P} (omega ).}{displaystyle operatorname {E} [X]=int _{Omega }X(omega ),doperatorname {P} (omega ).}

Remark 1. If X+(ω)=max(X(ω),0){displaystyle X_{+}(omega )=max(X(omega ),0)}{displaystyle X_{+}(omega )=max(X(omega ),0)} and X−)=−min(X(ω),0){displaystyle X_{-}(omega )=-min(X(omega ),0)}{displaystyle X_{-}(omega )=-min(X(omega ),0)}, then X=X+−X−.{displaystyle X=X_{+}-X_{-}.}{displaystyle X=X_{+}-X_{-}.} The functions X+{displaystyle X_{+}}X_{+} and X−{displaystyle X_{-}}{displaystyle X_{-}} can be shown to be measurable (hence, random variables), and, by definition of Lebesgue integral,


E⁡[X]=∫ΩX(ω)dP⁡)=∫ΩX+(ω)dP⁡)−ΩX−)dP⁡)=E⁡[X+]−E⁡[X−],{displaystyle {begin{aligned}operatorname {E} [X]&=int _{Omega }X(omega ),doperatorname {P} (omega )\&=int _{Omega }X_{+}(omega ),doperatorname {P} (omega )-int _{Omega }X_{-}(omega ),doperatorname {P} (omega )\&=operatorname {E} [X_{+}]-operatorname {E} [X_{-}],end{aligned}}}{displaystyle {begin{aligned}operatorname {E} [X]&=int _{Omega }X(omega ),doperatorname {P} (omega )\&=int _{Omega }X_{+}(omega ),doperatorname {P} (omega )-int _{Omega }X_{-}(omega ),doperatorname {P} (omega )\&=operatorname {E} [X_{+}]-operatorname {E} [X_{-}],end{aligned}}}

where E⁡[X+]{displaystyle operatorname {E} [X_{+}]}{displaystyle operatorname {E} [X_{+}]} and E⁡[X−]{displaystyle operatorname {E} [X_{-}]}{displaystyle operatorname {E} [X_{-}]} are non-negative and possibly infinite.


The following scenarios are possible:




  • E⁡[X]{displaystyle operatorname {E} [X]}operatorname {E} [X] is finite, i.e. max(E⁡[X+],E⁡[X−])<∞;{displaystyle max(operatorname {E} [X_{+}],operatorname {E} [X_{-}])<infty ;}{displaystyle max(operatorname {E} [X_{+}],operatorname {E} [X_{-}])<infty ;}


  • E⁡[X]{displaystyle operatorname {E} [X]}operatorname {E} [X] is infinite, i.e. max(E⁡[X+],E⁡[X−])=∞{displaystyle max(operatorname {E} [X_{+}],operatorname {E} [X_{-}])=infty }{displaystyle max(operatorname {E} [X_{+}],operatorname {E} [X_{-}])=infty } and min(E⁡[X+],E⁡[X−])<∞;{displaystyle min(operatorname {E} [X_{+}],operatorname {E} [X_{-}])<infty ;}{displaystyle min(operatorname {E} [X_{+}],operatorname {E} [X_{-}])<infty ;}


  • E⁡[X]{displaystyle operatorname {E} [X]}operatorname {E} [X] is neither finite nor infinite, i.e. E⁡[X+]=E⁡[X−]=∞.{displaystyle operatorname {E} [X_{+}]=operatorname {E} [X_{-}]=infty .}{displaystyle operatorname {E} [X_{+}]=operatorname {E} [X_{-}]=infty .}


Remark 2. If FX(x)=P⁡(X≤x){displaystyle F_{X}(x)=operatorname {P} (Xleq x)}{displaystyle F_{X}(x)=operatorname {P} (Xleq x)} is the cumulative distribution function of X{displaystyle X}X, then


E⁡[X]=∫+∞xdFX(x),{displaystyle operatorname {E} [X]=int _{-infty }^{+infty }x,dF_{X}(x),}{displaystyle operatorname {E} [X]=int _{-infty }^{+infty }x,dF_{X}(x),}

where the integral is interpreted in the sense of Lebesgue–Stieltjes.


Remark 3. An example of a distribution for which there is no expected value is Cauchy distribution.


Remark 4. For multidimensional random variables, their expected value is defined per component, i.e.


E⁡[(X1,…,Xn)]=(E⁡[X1],…,E⁡[Xn]){displaystyle operatorname {E} [(X_{1},ldots ,X_{n})]=(operatorname {E} [X_{1}],ldots ,operatorname {E} [X_{n}])}{displaystyle operatorname {E} [(X_{1},ldots ,X_{n})]=(operatorname {E} [X_{1}],ldots ,operatorname {E} [X_{n}])}

and, for a random matrix X{displaystyle X}X with elements Xij{displaystyle X_{ij}}X_{ij},



(E⁡[X])ij=E⁡[Xij]{displaystyle (operatorname {E} [X])_{ij}=operatorname {E} [X_{ij}]}{displaystyle (operatorname {E} [X])_{ij}=operatorname {E} [X_{ij}]}.


Basic properties


The properties below replicate or follow immediately from those of Lebesgue integral.



E⁡[1A]=P⁡(A){displaystyle operatorname {E} [{mathbf {1} }_{A}]=operatorname {P} (A)}{displaystyle operatorname {E} [{mathbf {1} }_{A}]=operatorname {P} (A)}


If A{displaystyle A}A is an event, then E⁡[1A]=P⁡(A),{displaystyle operatorname {E} [{mathbf {1} }_{A}]=operatorname {P} (A),}{displaystyle operatorname {E} [{mathbf {1} }_{A}]=operatorname {P} (A),} where 1A{displaystyle {mathbf {1} }_{A}}{displaystyle {mathbf {1} }_{A}} is the indicator function of the set A{displaystyle A}A.


Proof. By definition of Lebesgue integral of the simple function 1A=1A(ω){displaystyle {mathbf {1} }_{A}={mathbf {1} }_{A}(omega )}{displaystyle {mathbf {1} }_{A}={mathbf {1} }_{A}(omega )},



E⁡[1A]=1⋅P⁡(A)+0⋅P⁡A)=P⁡(A){displaystyle operatorname {E} [{mathbf {1} }_{A}]=1cdot operatorname {P} (A)+0cdot operatorname {P} (Omega setminus A)=operatorname {P} (A)}{displaystyle operatorname {E} [{mathbf {1} }_{A}]=1cdot operatorname {P} (A)+0cdot operatorname {P} (Omega setminus A)=operatorname {P} (A)}.


If X=Y{displaystyle X=Y}X=Y (a.s.) then E⁡[X]=E⁡[Y]{displaystyle operatorname {E} [X]=operatorname {E} [Y]}{displaystyle operatorname {E} [X]=operatorname {E} [Y]}


The statement follows from the definition of Lebesgue integral if we notice that X+=Y+{displaystyle X_{+}=Y_{+}}{displaystyle X_{+}=Y_{+}} (a.s.), X−=Y−{displaystyle X_{-}=Y_{-}}{displaystyle X_{-}=Y_{-}} (a.s.), and that changing a simple random variable on a set of probability zero does not alter the expected value.



Expected value of a constant


If X{displaystyle X}X is a random variable, and X=c{displaystyle X=c}{displaystyle X=c} (a.s.), where c∈[−,+∞]{displaystyle cin [-infty ,+infty ]}{displaystyle cin [-infty ,+infty ]}, then E⁡[X]=c{displaystyle operatorname {E} [X]=c}{displaystyle operatorname {E} [X]=c}. In particular, for an arbitrary random variable X{displaystyle X}X, E⁡[E⁡[X]]=E⁡[X]{displaystyle operatorname {E} [operatorname {E} [X]]=operatorname {E} [X]}{displaystyle operatorname {E} [operatorname {E} [X]]=operatorname {E} [X]}.







Linearity


The expected value operator (or expectation operator) E⁡[⋅]{displaystyle operatorname {E} [cdot ]}operatorname {E}[cdot ] is linear in the sense that


E⁡[X+Y]=E⁡[X]+E⁡[Y],E⁡[aX]=aE⁡[X],{displaystyle {begin{aligned}operatorname {E} [X+Y]&=operatorname {E} [X]+operatorname {E} [Y],\[6pt]operatorname {E} [aX]&=aoperatorname {E} [X],end{aligned}}}{displaystyle {begin{aligned}operatorname {E} [X+Y]&=operatorname {E} [X]+operatorname {E} [Y],\[6pt]operatorname {E} [aX]&=aoperatorname {E} [X],end{aligned}}}

where X{displaystyle X}X and Y{displaystyle Y}Y are (arbitrary) random variables, and a{displaystyle a}a is a scalar.


More rigorously, let X{displaystyle X}X and Y{displaystyle Y}Y be random variables whose expected values are defined (different from {displaystyle infty -infty }infty -infty ).


  • If E⁡[X]+E⁡[Y]{displaystyle operatorname {E} [X]+operatorname {E} [Y]}{displaystyle operatorname {E} [X]+operatorname {E} [Y]} is also defined (i.e. differs from {displaystyle infty -infty }infty -infty ), then

E⁡[X+Y]=E⁡[X]+E⁡[Y].{displaystyle operatorname {E} [X+Y]=operatorname {E} [X]+operatorname {E} [Y].}{displaystyle operatorname {E} [X+Y]=operatorname {E} [X]+operatorname {E} [Y].}

  • Let E⁡[X]{displaystyle operatorname {E} [X]}{displaystyle operatorname {E} [X]} be finite, and a∈R{displaystyle ain mathbb {R} }ain mathbb{R} be a finite scalar. Then E⁡[aX]=aE⁡[X].{displaystyle operatorname {E} [aX]=aoperatorname {E} [X].}{displaystyle operatorname {E} [aX]=aoperatorname {E} [X].}






E⁡[X]{displaystyle operatorname {E} [X]}operatorname {E} [X] is finite if and only if E⁡|X|{displaystyle operatorname {E} |X|}{displaystyle operatorname {E} |X|} is


The following statements regarding a random variable X{displaystyle X}X are equivalent:




  • E⁡[X]{displaystyle operatorname {E} [X]}operatorname {E} [X] is finite.

  • Both E⁡[X+]{displaystyle operatorname {E} [X_{+}]}{displaystyle operatorname {E} [X_{+}]} and E⁡[X−]{displaystyle operatorname {E} [X_{-}]}{displaystyle operatorname {E} [X_{-}]} are finite.


  • E⁡|X|{displaystyle operatorname {E} |X|}{displaystyle operatorname {E} |X|} is finite.


Sketch of proof. Indeed, |X|=X++X−{displaystyle |X|=X_{+}+X_{-}}{displaystyle |X|=X_{+}+X_{-}}. By linearity, E⁡|X|=E⁡[X+]+E⁡[X−]{displaystyle operatorname {E} |X|=operatorname {E} [X_{+}]+operatorname {E} [X_{-}]}{displaystyle operatorname {E} |X|=operatorname {E} [X_{+}]+operatorname {E} [X_{-}]}. The above equivalency relies on the definition of Lebesgue integral and measurability of X{displaystyle X}X.


Remark. For the reasons above, the expressions "X{displaystyle X}X is integrable" and "the expected value of X{displaystyle X}X is finite" are used interchangeably when speaking of a random variable throughout this article.



If X≥0{displaystyle Xgeq 0}{displaystyle Xgeq 0} (a.s.) then E⁡[X]≥0{displaystyle operatorname {E} [X]geq 0}{displaystyle operatorname {E} [X]geq 0}







Monotonicity


If X≤Y{displaystyle Xleq Y}{displaystyle Xleq Y} (a.s.), and both E⁡[X]{displaystyle operatorname {E} [X]}operatorname {E} [X] and E⁡[Y]{displaystyle operatorname {E} [Y]}{displaystyle operatorname {E} [Y]} exist, then E⁡[X]≤E⁡[Y]{displaystyle operatorname {E} [X]leq operatorname {E} [Y]}{displaystyle operatorname {E} [X]leq operatorname {E} [Y]}.


Remark. E⁡[X]{displaystyle operatorname {E} [X]}operatorname {E} [X] and E⁡[Y]{displaystyle operatorname {E} [Y]}{displaystyle operatorname {E} [Y]} exist in the sense that min(E⁡[X+],E⁡[X−])<∞{displaystyle min(operatorname {E} [X_{+}],operatorname {E} [X_{-}])<infty }{displaystyle min(operatorname {E} [X_{+}],operatorname {E} [X_{-}])<infty } and min(E⁡[Y+],E⁡[Y−])<∞.{displaystyle min(operatorname {E} [Y_{+}],operatorname {E} [Y_{-}])<infty .}{displaystyle min(operatorname {E} [Y_{+}],operatorname {E} [Y_{-}])<infty .}


Proof follows from the linearity and the previous property if we set Z=Y−X{displaystyle Z=Y-X}{displaystyle Z=Y-X} and notice that Z≥0{displaystyle Zgeq 0}{displaystyle Zgeq 0} (a.s.).



If |X|≤Y{displaystyle |X|leq Y}{displaystyle |X|leq Y} (a.s.) and E⁡[Y]{displaystyle operatorname {E} [Y]}{displaystyle operatorname {E} [Y]} is finite then so is E⁡[X]{displaystyle operatorname {E} [X]}operatorname {E} [X]


Let X{displaystyle X}X and Y{displaystyle Y}Y be random variables such that |X|≤Y{displaystyle |X|leq Y}{displaystyle |X|leq Y} (a.s.) and E⁡[Y]<∞{displaystyle operatorname {E} [Y]<infty }{displaystyle operatorname {E} [Y]<infty }. Then E⁡[X]≠±{displaystyle operatorname {E} [X]neq pm infty }{displaystyle operatorname {E} [X]neq pm infty }.


Proof. Due to non-negativity of |X|{displaystyle |X|}|X|, E⁡|X|{displaystyle operatorname {E} |X|}{displaystyle operatorname {E} |X|} exists, finite or infinite. By monotonicity, E⁡|X|≤E⁡[Y]<∞{displaystyle operatorname {E} |X|leq operatorname {E} [Y]<infty }{displaystyle operatorname {E} |X|leq operatorname {E} [Y]<infty }, so E⁡|X|{displaystyle operatorname {E} |X|}{displaystyle operatorname {E} |X|} is finite which, as we saw earlier, is equivalent to E⁡[X]{displaystyle operatorname {E} [X]}operatorname {E} [X] being finite.



If E⁡|Xβ|<∞{displaystyle operatorname {E} |X^{beta }|<infty }{displaystyle operatorname {E} |X^{beta }|<infty } and 0<α{displaystyle 0<alpha <beta }{displaystyle 0<alpha <beta } then E⁡|Xα|<∞{displaystyle operatorname {E} |X^{alpha }|<infty }{displaystyle operatorname {E} |X^{alpha }|<infty }


The proposition below will be used to prove the extremal property of
E⁡[X]{displaystyle operatorname {E} [X]}operatorname {E} [X] later on.


Proposition. If X{displaystyle X}X is a random variable, then so is {displaystyle X^{alpha }}{displaystyle X^{alpha }}, for every α>0{displaystyle alpha >0}alpha >0. If, in addition, E⁡|Xβ|<∞{displaystyle operatorname {E} |X^{beta }|<infty }{displaystyle operatorname {E} |X^{beta }|<infty } and
0<α{displaystyle 0<alpha <beta }{displaystyle 0<alpha <beta }, then E⁡|Xα|<∞{displaystyle operatorname {E} |X^{alpha }|<infty }{displaystyle operatorname {E} |X^{alpha }|<infty }.







Counterexample for infinite measure


The requirement that P⁡)<∞{displaystyle operatorname {P} (Omega )<infty }{displaystyle operatorname {P} (Omega )<infty } is essential. By way of counterexample, consider the measurable space


([1,+∞),BR[1,+∞),λ),{displaystyle ([1,+infty ),{mathcal {B}}_{mathbb {R} _{[1,+infty )}},lambda ),}{displaystyle ([1,+infty ),{mathcal {B}}_{mathbb {R} _{[1,+infty )}},lambda ),}

where BR[1,+∞){displaystyle {mathcal {B}}_{mathbb {R} _{[1,+infty )}}}{displaystyle {mathcal {B}}_{mathbb {R} _{[1,+infty )}}} is the Borel σ{displaystyle sigma }sigma -algebra on the interval [1,+∞),{displaystyle [1,+infty ),}{displaystyle [1,+infty ),} and λ{displaystyle lambda }lambda is the linear Lebesgue measure. The reader can prove that [1,+∞)1xdx=∞,{displaystyle textstyle int _{[1,+infty )}{frac {1}{x}},dx=infty ,}{displaystyle textstyle int _{[1,+infty )}{frac {1}{x}},dx=infty ,} even though [1,+∞)1x2dx=1.{displaystyle textstyle int _{[1,+infty )}{frac {1}{x^{2}}},dx=1.}{displaystyle textstyle int _{[1,+infty )}{frac {1}{x^{2}}},dx=1.} (Sketch of proof: S1xdx{displaystyle textstyle int _{S}{frac {1}{x}},dx}{displaystyle textstyle int _{S}{frac {1}{x}},dx} and S1x2dx{displaystyle textstyle int _{S}{frac {1}{x^{2}}},dx}{displaystyle textstyle int _{S}{frac {1}{x^{2}}},dx} define a measure μ{displaystyle mu }mu on [1,+∞)=∪n=1∞[1,n].{displaystyle textstyle [1,+infty )=cup _{n=1}^{infty }[1,n].}{displaystyle textstyle [1,+infty )=cup _{n=1}^{infty }[1,n].} Use "continuity from below" w.r. to μ{displaystyle mu }mu and reduce to Riemann integral on each finite subinterval [1,n]{displaystyle [1,n]}[1,n]).



Extremal property


Recall, as we proved early on, that if X{displaystyle X}X is a random variable, then so is X2{displaystyle X^{2}}X^{2}.


Proposition (extremal property of E⁡[X]){displaystyle operatorname {E} [X])}{displaystyle operatorname {E} [X])}). Let X{displaystyle X}X be a random variable, and E⁡[X2]<∞{displaystyle operatorname {E} [X^{2}]<infty }{displaystyle operatorname {E} [X^{2}]<infty }. Then E⁡[X]{displaystyle operatorname {E} [X]}{displaystyle operatorname {E} [X]} and Var⁡[X]{displaystyle operatorname {Var} [X]}{displaystyle operatorname {Var} [X]} are finite, and E⁡[X]{displaystyle operatorname {E} [X]}{displaystyle operatorname {E} [X]} is the best least squares approximation for X{displaystyle X}X among constants. Specifically,



  • for every c∈R{displaystyle cin mathbb {R} }{displaystyle cin mathbb {R} }, E⁡[X−c]2≥Var⁡[X];{displaystyle textstyle operatorname {E} [X-c]^{2}geq operatorname {Var} [X];}{displaystyle textstyle operatorname {E} [X-c]^{2}geq operatorname {Var} [X];}

  • equality holds if and only if c=E⁡[X].{displaystyle c=operatorname {E} [X].}{displaystyle c=operatorname {E} [X].}


(Var⁡[X]{displaystyle operatorname {Var} [X]}{displaystyle operatorname {Var} [X]} denotes the variance of X{displaystyle X}X).


Remark (intuitive interpretation of extremal property). In intuitive terms, the extremal property says that if one is asked to predict the outcome of a trial of a random variable X{displaystyle X}X, then E⁡[X]{displaystyle operatorname {E} [X]}operatorname {E} [X], in some practically useful sense, is one's best bet if no advance information about the outcome is available. If, on the other hand, one does have some advance knowledge F{displaystyle {mathcal {F}}}{mathcal {F}} regarding the outcome, then — again, in some practically useful sense — one's bet may be improved upon by using conditional expectations E⁡[X∣F]{displaystyle operatorname {E} [Xmid {mathcal {F}}]}{displaystyle operatorname {E} [Xmid {mathcal {F}}]} (of which E⁡[X]{displaystyle operatorname {E} [X]}operatorname {E} [X] is a special case) rather than E⁡[X]{displaystyle operatorname {E} [X]}operatorname {E} [X].


Proof of proposition. By the above properties, both E⁡[X]{displaystyle operatorname {E} [X]}operatorname {E} [X] and
Var⁡[X]=E⁡[X2]−E2⁡[X]{displaystyle operatorname {Var} [X]=operatorname {E} [X^{2}]-operatorname {E} ^{2}[X]}{displaystyle operatorname {Var} [X]=operatorname {E} [X^{2}]-operatorname {E} ^{2}[X]} are finite, and


E⁡[X−c]2=E⁡[X2−2cX+c2]=E⁡[X2]−2cE⁡[X]+c2=(c−E⁡[X])2+E⁡[X2]−E2⁡[X]=(c−E⁡[X])2+Var⁡[X],{displaystyle {begin{aligned}operatorname {E} [X-c]^{2}&=operatorname {E} [X^{2}-2cX+c^{2}]\[6pt]&=operatorname {E} [X^{2}]-2coperatorname {E} [X]+c^{2}\[6pt]&=(c-operatorname {E} [X])^{2}+operatorname {E} [X^{2}]-operatorname {E} ^{2}[X]\[6pt]&=(c-operatorname {E} [X])^{2}+operatorname {Var} [X],end{aligned}}}{displaystyle {begin{aligned}operatorname {E} [X-c]^{2}&=operatorname {E} [X^{2}-2cX+c^{2}]\[6pt]&=operatorname {E} [X^{2}]-2coperatorname {E} [X]+c^{2}\[6pt]&=(c-operatorname {E} [X])^{2}+operatorname {E} [X^{2}]-operatorname {E} ^{2}[X]\[6pt]&=(c-operatorname {E} [X])^{2}+operatorname {Var} [X],end{aligned}}}

whence the extremal property follows.



Non-degeneracy


If E⁡|X|=0{displaystyle operatorname {E} |X|=0}{displaystyle operatorname {E} |X|=0}, then X=0{displaystyle X=0}{displaystyle X=0} (a.s.).







If E⁡[X]<+∞{displaystyle operatorname {E} [X]<+infty }{displaystyle operatorname {E} [X]<+infty } then X<+∞{displaystyle X<+infty }{displaystyle X<+infty } (a.s.)







Corollary: if E⁡[X]>−{displaystyle operatorname {E} [X]>-infty }{displaystyle operatorname {E} [X]>-infty } then X>−{displaystyle X>-infty }{displaystyle X>-infty } (a.s.)



Corollary: if E⁡|X|<∞{displaystyle operatorname {E} |X|<infty }{displaystyle operatorname {E} |X|<infty } then X≠±{displaystyle Xneq pm infty }{displaystyle Xneq pm infty } (a.s.)



|E⁡[X]|≤E⁡|X|{displaystyle |operatorname {E} [X]|leq operatorname {E} |X|}{displaystyle |operatorname {E} [X]|leq operatorname {E} |X|}


For an arbitrary random variable X{displaystyle X}X, |E⁡[X]|≤E⁡|X|{displaystyle |operatorname {E} [X]|leq operatorname {E} |X|}{displaystyle |operatorname {E} [X]|leq operatorname {E} |X|}.


Proof. By definition of Lebesgue integral,


|E⁡[X]|=|E⁡[X+]−E⁡[X−]|≤|E⁡[X+]|+|E⁡[X−]|=E⁡[X+]+E⁡[X−]=E⁡[X++X−]=E⁡|X|.{displaystyle {begin{aligned}|operatorname {E} [X]|&={Bigl |}operatorname {E} [X_{+}]-operatorname {E} [X_{-}]{Bigr |}leq {Bigl |}operatorname {E} [X_{+}]{Bigr |}+{Bigl |}operatorname {E} [X_{-}]{Bigr |}\[5pt]&=operatorname {E} [X_{+}]+operatorname {E} [X_{-}]=operatorname {E} [X_{+}+X_{-}]\[5pt]&=operatorname {E} |X|.end{aligned}}}{displaystyle {begin{aligned}|operatorname {E} [X]|&={Bigl |}operatorname {E} [X_{+}]-operatorname {E} [X_{-}]{Bigr |}leq {Bigl |}operatorname {E} [X_{+}]{Bigr |}+{Bigl |}operatorname {E} [X_{-}]{Bigr |}\[5pt]&=operatorname {E} [X_{+}]+operatorname {E} [X_{-}]=operatorname {E} [X_{+}+X_{-}]\[5pt]&=operatorname {E} |X|.end{aligned}}}

Note that this result can also be proved based on Jensen's inequality.



Non-multiplicativity


In general, the expected value operator is not multiplicative, i.e. E⁡[XY]{displaystyle operatorname {E} [XY]}{displaystyle operatorname {E} [XY]} is not necessarily equal to E⁡[X]⋅E⁡[Y]{displaystyle operatorname {E} [X]cdot operatorname {E} [Y]}{displaystyle operatorname {E} [X]cdot operatorname {E} [Y]}. Indeed, let X{displaystyle X}X assume the values of 1 and -1 with probability 0.5 each. Then



E2⁡[X]=(12⋅(−1)+12⋅1)2=0{displaystyle operatorname {E^{2}} [X]=left({frac {1}{2}}cdot (-1)+{frac {1}{2}}cdot 1right)^{2}=0}{displaystyle operatorname {E^{2}} [X]=left({frac {1}{2}}cdot (-1)+{frac {1}{2}}cdot 1right)^{2}=0},

and



E⁡[X2]=12⋅(−1)2+12⋅12=1{displaystyle operatorname {E} [X^{2}]={frac {1}{2}}cdot (-1)^{2}+{frac {1}{2}}cdot 1^{2}=1}{displaystyle operatorname {E} [X^{2}]={frac {1}{2}}cdot (-1)^{2}+{frac {1}{2}}cdot 1^{2}=1},

so E⁡[X2]≠E2⁡[X]{displaystyle operatorname {E} [X^{2}]neq operatorname {E^{2}} [X]}{displaystyle operatorname {E} [X^{2}]neq operatorname {E^{2}} [X]}.


The amount by which the multiplicativity fails is called the covariance:


Cov⁡(X,Y)=E⁡[XY]−E⁡[X]E⁡[Y].{displaystyle operatorname {Cov} (X,Y)=operatorname {E} [XY]-operatorname {E} [X]operatorname {E} [Y].}operatorname {Cov} (X,Y)=operatorname {E} [XY]-operatorname {E} [X]operatorname {E} [Y].

If, however, the random variables X∈1,F1,P1){displaystyle Xin (Omega _{1},{mathcal {F}}_{1},operatorname {P} _{1})}{displaystyle Xin (Omega _{1},{mathcal {F}}_{1},operatorname {P} _{1})} and Y∈2,F2,P2){displaystyle Yin (Omega _{2},{mathcal {F}}_{2},operatorname {P} _{2})}{displaystyle Yin (Omega _{2},{mathcal {F}}_{2},operatorname {P} _{2})} are independent, then E⁡[XY]=E⁡[X]E⁡[Y]{displaystyle operatorname {E} [XY]=operatorname {E} [X]operatorname {E} [Y]}{displaystyle operatorname {E} [XY]=operatorname {E} [X]operatorname {E} [Y]}, and Cov⁡(X,Y)=0{displaystyle operatorname {Cov} (X,Y)=0}{displaystyle operatorname {Cov} (X,Y)=0}.



Counterexample: E⁡[Xi]↛E⁡[X]{displaystyle operatorname {E} [X_{i}]not to operatorname {E} [X]}{displaystyle operatorname {E} [X_{i}]not to operatorname {E} [X]} despite Xi→X{displaystyle X_{i}to X}{displaystyle X_{i}to X} pointwise


Let ([0,1],B[0,1],P){displaystyle left([0,1],{mathcal {B}}_{[0,1]},{mathrm {P} }right)}{displaystyle left([0,1],{mathcal {B}}_{[0,1]},{mathrm {P} }right)} be the probability space, where B[0,1]{displaystyle {mathcal {B}}_{[0,1]}}{displaystyle {mathcal {B}}_{[0,1]}} is the Borel σ{displaystyle sigma }sigma -algebra on [0,1]{displaystyle [0,1]}[0,1] and P{displaystyle {mathrm {P} }}{displaystyle {mathrm {P} }} the linear Lebesgue measure. For i≥1,{displaystyle igeq 1,}{displaystyle igeq 1,} define a sequence of random variables


Xi=i⋅1[0,1i]{displaystyle X_{i}=icdot {mathbf {1} }_{left[0,{frac {1}{i}}right]}}{displaystyle X_{i}=icdot {mathbf {1} }_{left[0,{frac {1}{i}}right]}}

and a random variable


X={+∞if x=00otherwise.{displaystyle X={begin{cases}+infty &{text{if}} x=0\0&{text{otherwise.}}end{cases}}}{displaystyle X={begin{cases}+infty &{text{if}} x=0\0&{text{otherwise.}}end{cases}}}

on [0,1]{displaystyle [0,1]}[0,1], with 1S{displaystyle {mathbf {1} }_{S}}{displaystyle {mathbf {1} }_{S}} being the indicator function of the set S⊆[0,1]{displaystyle Ssubseteq [0,1]}{displaystyle Ssubseteq [0,1]}.


For every x∈[0,1],{displaystyle xin [0,1],}{displaystyle xin [0,1],} as i→+∞,{displaystyle ito +infty ,}{displaystyle ito +infty ,} Xi(x)→X(x),{displaystyle X_{i}(x)to X(x),}{displaystyle X_{i}(x)to X(x),} and


E⁡[Xi]=i⋅P([0,1i])=i⋅1i=1,{displaystyle operatorname {E} [X_{i}]=icdot {mathrm {P} }left(left[0,{frac {1}{i}}right]right)=icdot {dfrac {1}{i}}=1,}{displaystyle operatorname {E} [X_{i}]=icdot {mathrm {P} }left(left[0,{frac {1}{i}}right]right)=icdot {dfrac {1}{i}}=1,}

so limi→E⁡[Xi]=1.{displaystyle lim _{ito infty }operatorname {E} [X_{i}]=1.}{displaystyle lim _{ito infty }operatorname {E} [X_{i}]=1.} On the other hand, P⁡({0})=0,{displaystyle mathop {mathrm {P} } ({0})=0,}{displaystyle mathop {mathrm {P} } ({0})=0,} and hence E⁡[X]=0.{displaystyle operatorname {E} left[Xright]=0.}{displaystyle operatorname {E} left[Xright]=0.}



Countable non-additivity


In general, the expected value operator is not σ{displaystyle sigma }sigma -additive, i.e.


E⁡[∑i=0∞Xi]≠i=0∞E⁡[Xi].{displaystyle operatorname {E} left[sum _{i=0}^{infty }X_{i}right]neq sum _{i=0}^{infty }operatorname {E} [X_{i}].}{displaystyle operatorname {E} left[sum _{i=0}^{infty }X_{i}right]neq sum _{i=0}^{infty }operatorname {E} [X_{i}].}

By way of counterexample, let ([0,1],B[0,1],P){displaystyle left([0,1],{mathcal {B}}_{[0,1]},{mathrm {P} }right)}{displaystyle left([0,1],{mathcal {B}}_{[0,1]},{mathrm {P} }right)} be the probability space, where B[0,1]{displaystyle {mathcal {B}}_{[0,1]}}{displaystyle {mathcal {B}}_{[0,1]}} is the Borel σ{displaystyle sigma }sigma -algebra on [0,1]{displaystyle [0,1]}[0,1] and P{displaystyle {mathrm {P} }}{displaystyle {mathrm {P} }} the linear Lebesgue measure. Define a sequence of random variables Xi=(i+1)⋅1[0,1i+1]−i⋅1[0,1i]{displaystyle textstyle X_{i}=(i+1)cdot {mathbf {1} }_{left[0,{frac {1}{i+1}}right]}-icdot {mathbf {1} }_{left[0,{frac {1}{i}}right]}}{displaystyle textstyle X_{i}=(i+1)cdot {mathbf {1} }_{left[0,{frac {1}{i+1}}right]}-icdot {mathbf {1} }_{left[0,{frac {1}{i}}right]}} on [0,1]{displaystyle [0,1]}[0,1], with 1S{displaystyle {mathbf {1} }_{S}}{displaystyle {mathbf {1} }_{S}} being the indicator function of the set S⊆[0,1]{displaystyle Ssubseteq [0,1]}{displaystyle Ssubseteq [0,1]}. For the pointwise sums, we have



i=0nXi=(n+1)⋅1[0,1n+1],{displaystyle sum _{i=0}^{n}X_{i}=(n+1)cdot {mathbf {1} }_{left[0,{frac {1}{n+1}}right]},}{displaystyle sum _{i=0}^{n}X_{i}=(n+1)cdot {mathbf {1} }_{left[0,{frac {1}{n+1}}right]},}

i=0∞Xi(x)={+∞if x=00otherwise.{displaystyle sum _{i=0}^{infty }X_{i}(x)={begin{cases}+infty &{text{if}} x=0\0&{text{otherwise.}}end{cases}}}{displaystyle sum _{i=0}^{infty }X_{i}(x)={begin{cases}+infty &{text{if}} x=0\0&{text{otherwise.}}end{cases}}}


By finite additivity,


i=0∞E⁡[Xi]=limn→i=0nE⁡[Xi]=limn→E⁡[∑i=0nXi]=1.{displaystyle sum _{i=0}^{infty }operatorname {E} [X_{i}]=lim _{nto infty }sum _{i=0}^{n}operatorname {E} [X_{i}]=lim _{nto infty }operatorname {E} left[sum _{i=0}^{n}X_{i}right]=1.}{displaystyle sum _{i=0}^{infty }operatorname {E} [X_{i}]=lim _{nto infty }sum _{i=0}^{n}operatorname {E} [X_{i}]=lim _{nto infty }operatorname {E} left[sum _{i=0}^{n}X_{i}right]=1.}

On the other hand, P⁡({0})=0,{displaystyle mathop {mathrm {P} } ({0})=0,}{displaystyle mathop {mathrm {P} } ({0})=0,} and hence


E⁡[∑i=0∞Xi]=0≠1=∑i=0∞E⁡[Xi].{displaystyle operatorname {E} left[sum _{i=0}^{infty }X_{i}right]=0neq 1=sum _{i=0}^{infty }operatorname {E} [X_{i}].}{displaystyle operatorname {E} left[sum _{i=0}^{infty }X_{i}right]=0neq 1=sum _{i=0}^{infty }operatorname {E} [X_{i}].}


Countable additivity for non-negative random variables


Let {Xi}i=0∞{displaystyle {X_{i}}_{i=0}^{infty }}{displaystyle {X_{i}}_{i=0}^{infty }} be non-negative random variables. It follows from monotone convergence theorem that


E⁡[∑i=0∞Xi]=∑i=0∞E⁡[Xi].{displaystyle operatorname {E} left[sum _{i=0}^{infty }X_{i}right]=sum _{i=0}^{infty }operatorname {E} [X_{i}].}{displaystyle operatorname {E} left[sum _{i=0}^{infty }X_{i}right]=sum _{i=0}^{infty }operatorname {E} [X_{i}].}


E⁡[XY]=E⁡[X]E⁡[Y]{displaystyle operatorname {E} [XY]=operatorname {E} [X]operatorname {E} [Y]}{displaystyle operatorname {E} [XY]=operatorname {E} [X]operatorname {E} [Y]} for independent X{displaystyle X}X and Y{displaystyle Y}Y


Let X{displaystyle X}X and Y{displaystyle Y}Y be independent random variables with finite expectations E⁡[X]{displaystyle operatorname {E} [X]}operatorname {E} [X] and E⁡[Y]{displaystyle operatorname {E} [Y]}{displaystyle operatorname {E} [Y]}. Then E⁡[XY]=E⁡[X]E⁡[Y]{displaystyle operatorname {E} [XY]=operatorname {E} [X]operatorname {E} [Y]}{displaystyle operatorname {E} [XY]=operatorname {E} [X]operatorname {E} [Y]}.







Inequalities



Cauchy–Bunyakovsky–Schwarz inequality



The Cauchy–Bunyakovsky–Schwarz inequality states that


(E⁡[XY])2≤E⁡[X2]⋅E⁡[Y2].{displaystyle (operatorname {E} [XY])^{2}leq operatorname {E} [X^{2}]cdot operatorname {E} [Y^{2}].}{displaystyle (operatorname {E} [XY])^{2}leq operatorname {E} [X^{2}]cdot operatorname {E} [Y^{2}].}


Markov's inequality



For a nonnegative random variable X{displaystyle X}X and a>0{displaystyle a>0}a>0, the Markov's inequality states that


P⁡(X≥a)≤E⁡[X]a.{displaystyle operatorname {P} (Xgeq a)leq {frac {operatorname {E} [X]}{a}}.}{displaystyle operatorname {P} (Xgeq a)leq {frac {operatorname {E} [X]}{a}}.}


Bienaymé-Chebyshev inequality



Let X{displaystyle X}X be an arbitrary random variable with finite expected value E⁡[X]{displaystyle operatorname {E} [X]}operatorname {E} [X] and finite variance Var⁡[X]≠0{displaystyle operatorname {Var} [X]neq 0}{displaystyle operatorname {Var} [X]neq 0}. The Bienaymé-Chebyshev inequality states that, for any real number k>0{displaystyle k>0}k>0,


P⁡(|X−E⁡[X]|≥kVar⁡[X])≤1k2.{displaystyle operatorname {P} {Bigl (}{Bigl |}X-operatorname {E} [X]{Bigr |}geq k{sqrt {operatorname {Var} [X]}}{Bigr )}leq {frac {1}{k^{2}}}.}{displaystyle operatorname {P} {Bigl (}{Bigl |}X-operatorname {E} [X]{Bigr |}geq k{sqrt {operatorname {Var} [X]}}{Bigr )}leq {frac {1}{k^{2}}}.}


Jensen's inequality



Let f:R→R{displaystyle f:{mathbb {R} }to {mathbb {R} }}{displaystyle f:{mathbb {R} }to {mathbb {R} }} be a Borel convex function and X{displaystyle X}X a random variable such that E⁡|X|<∞{displaystyle operatorname {E} |X|<infty }{displaystyle operatorname {E} |X|<infty }. Jensen's inequality states that


f(E⁡(X))≤E⁡(f(X)).{displaystyle f(operatorname {E} (X))leq operatorname {E} (f(X)).}{displaystyle f(operatorname {E} (X))leq operatorname {E} (f(X)).}

Remark 1. The expected value E⁡(f(X)){displaystyle operatorname {E} (f(X))}{displaystyle operatorname {E} (f(X))} is well-defined even if X{displaystyle X}X is allowed to assume infinite values. Indeed, E⁡|X|<∞{displaystyle operatorname {E} |X|<infty }{displaystyle operatorname {E} |X|<infty } implies that X≠±{displaystyle Xneq pm infty }{displaystyle Xneq pm infty } (a.s.), so the random variable f(X(ω)){displaystyle f(X(omega ))}{displaystyle f(X(omega ))} is defined almost sure, and therefore there is enough information to compute E⁡(f(X)).{displaystyle operatorname {E} (f(X)).}{displaystyle operatorname {E} (f(X)).}


Remark 2. Jensen's inequality implies that |E⁡[X]|≤E⁡|X|{displaystyle |operatorname {E} [X]|leq operatorname {E} |X|}{displaystyle |operatorname {E} [X]|leq operatorname {E} |X|} since the absolute value function is convex.



Lyapunov’s inequality


Let 0<s<t{displaystyle 0<s<t}{displaystyle 0<s<t}. Lyapunov's inequality states that


(E⁡|X|s)1/s≤(E⁡|X|t)1/t.{displaystyle {Bigl (}operatorname {E} |X|^{s}{Bigr )}^{1/s}leq left(operatorname {E} |X|^{t}right)^{1/t}.}{displaystyle {Bigl (}operatorname {E} |X|^{s}{Bigr )}^{1/s}leq left(operatorname {E} |X|^{t}right)^{1/t}.}

Proof. Applying Jensen's inequality to |X|s{displaystyle |X|^{s}}{displaystyle |X|^{s}} and g(x)=|x|t/s{displaystyle g(x)=|x|^{t/s}}{displaystyle g(x)=|x|^{t/s}}, obtain
|E⁡|Xs||t/s≤E⁡|Xs|t/s=E⁡|X|t{displaystyle {Bigl |}operatorname {E} |X^{s}|{Bigr |}^{t/s}leq operatorname {E} |X^{s}|^{t/s}=operatorname {E} |X|^{t}}{displaystyle {Bigl |}operatorname {E} |X^{s}|{Bigr |}^{t/s}leq operatorname {E} |X^{s}|^{t/s}=operatorname {E} |X|^{t}}. Taking the t{displaystyle t}tth
root of each side completes the proof.


Corollary.


E⁡|X|≤(E⁡|X|2)1/2≤(E⁡|X|n)1/n≤{displaystyle operatorname {E} |X|leq {Bigl (}operatorname {E} |X|^{2}{Bigr )}^{1/2}leq ldots leq {Bigl (}operatorname {E} |X|^{n}{Bigr )}^{1/n}leq ldots }{displaystyle operatorname {E} |X|leq {Bigl (}operatorname {E} |X|^{2}{Bigr )}^{1/2}leq ldots leq {Bigl (}operatorname {E} |X|^{n}{Bigr )}^{1/n}leq ldots }


Hölder’s inequality



Let p{displaystyle p}p and q{displaystyle q}q satisfy 1≤p≤{displaystyle 1leq pleq infty }{displaystyle 1leq pleq infty }, 1≤q≤{displaystyle 1leq qleq infty }{displaystyle 1leq qleq infty }, and 1/p+1/q=1{displaystyle 1/p+1/q=1}1/p+1/q=1. The Hölder's inequality states that


E⁡|XY|≤(E⁡|X|p)1/p(E⁡|Y|q)1/q.{displaystyle operatorname {E} |XY|leq (operatorname {E} |X|^{p})^{1/p}(operatorname {E} |Y|^{q})^{1/q}.}{displaystyle operatorname {E} |XY|leq (operatorname {E} |X|^{p})^{1/p}(operatorname {E} |Y|^{q})^{1/q}.}


Minkowski inequality



Let p{displaystyle p}p be an integer satisfying 1≤p≤{displaystyle 1leq pleq infty }{displaystyle 1leq pleq infty }. Let, in addition, E⁡|X|p<∞{displaystyle operatorname {E} |X|^{p}<infty }{displaystyle operatorname {E} |X|^{p}<infty } and E⁡|Y|p<∞{displaystyle operatorname {E} |Y|^{p}<infty }{displaystyle operatorname {E} |Y|^{p}<infty }. Then, according to the Minkowski inequality, E⁡|X+Y|p<∞{displaystyle operatorname {E} |X+Y|^{p}<infty }{displaystyle operatorname {E} |X+Y|^{p}<infty } and


(E⁡|X+Y|p)1/p≤(E⁡|X|p)1/p+(E⁡|Y|p)1/p.{displaystyle {Bigl (}operatorname {E} |X+Y|^{p}{Bigr )}^{1/p}leq {Bigl (}operatorname {E} |X|^{p}{Bigr )}^{1/p}+{Bigl (}operatorname {E} |Y|^{p}{Bigr )}^{1/p}.}{displaystyle {Bigl (}operatorname {E} |X+Y|^{p}{Bigr )}^{1/p}leq {Bigl (}operatorname {E} |X|^{p}{Bigr )}^{1/p}+{Bigl (}operatorname {E} |Y|^{p}{Bigr )}^{1/p}.}


Taking limits under the E{displaystyle operatorname {E} }operatorname {E} sign



Monotone convergence theorem



Let the sequence of random variables {Xn}{displaystyle {X_{n}}}{X_{n}} and the random variables X{displaystyle X}X and Y{displaystyle Y}Y be defined on the same probability space ,P).{displaystyle (Omega ,Sigma ,operatorname {P} ).}{displaystyle (Omega ,Sigma ,operatorname {P} ).} Suppose that



  • all the expected values E⁡[Xn],{displaystyle operatorname {E} [X_{n}],}{displaystyle operatorname {E} [X_{n}],} E⁡[X],{displaystyle operatorname {E} [X],}{displaystyle operatorname {E} [X],} and E⁡[Y]{displaystyle operatorname {E} [Y]}{displaystyle operatorname {E} [Y]} are defined (differ from {displaystyle infty -infty }infty -infty );

  • E⁡[Y]>−;{displaystyle operatorname {E} [Y]>-infty ;}{displaystyle operatorname {E} [Y]>-infty ;}

  • for every n,{displaystyle n,}{displaystyle n,}


Y≤Xn≤Xn+1≤+∞(a.s.);{displaystyle -infty leq Yleq X_{n}leq X_{n+1}leq +infty quad {hbox{(a.s.)}};}{displaystyle -infty leq Yleq X_{n}leq X_{n+1}leq +infty quad {hbox{(a.s.)}};}


  • X{displaystyle X}X is the pointwise limit of {Xn}{displaystyle {X_{n}}}{displaystyle {X_{n}}} (a.s.), i.e. X(ω)=limnXn(ω){displaystyle X(omega )=lim nolimits _{n}X_{n}(omega )}{displaystyle X(omega )=lim nolimits _{n}X_{n}(omega )} (a.s.).

The monotone convergence theorem states that


limnE⁡[Xn]=E⁡[X].{displaystyle lim _{n}operatorname {E} [X_{n}]=operatorname {E} [X].}{displaystyle lim _{n}operatorname {E} [X_{n}]=operatorname {E} [X].}







Fatou's lemma



Let the sequence of random variables {Xn}{displaystyle {X_{n}}}{X_{n}} and the random variable Y{displaystyle Y}Y be defined on the same probability space ,P).{displaystyle (Omega ,Sigma ,operatorname {P} ).}{displaystyle (Omega ,Sigma ,operatorname {P} ).} Suppose that



  • all the expected values E⁡[Xn],{displaystyle operatorname {E} [X_{n}],}{displaystyle operatorname {E} [X_{n}],} E⁡[lim infnXn],{displaystyle textstyle operatorname {E} [liminf _{n}X_{n}],}{displaystyle textstyle operatorname {E} [liminf _{n}X_{n}],} and E⁡[Y]{displaystyle operatorname {E} [Y]}{displaystyle operatorname {E} [Y]} are defined (differ from {displaystyle infty -infty }infty -infty );

  • E⁡[Y]>−;{displaystyle operatorname {E} [Y]>-infty ;}{displaystyle operatorname {E} [Y]>-infty ;}


  • Y≤Xn≤+∞{displaystyle -infty leq Yleq X_{n}leq +infty }{displaystyle -infty leq Yleq X_{n}leq +infty } (a.s.), for every n.{displaystyle n.}n.


Fatou's lemma states that


E⁡[lim infnXn]≤lim infnE⁡[Xn].{displaystyle operatorname {E} [liminf _{n}X_{n}]leq liminf _{n}operatorname {E} [X_{n}].}{displaystyle operatorname {E} [liminf _{n}X_{n}]leq liminf _{n}operatorname {E} [X_{n}].}

(Note that lim infnXn{displaystyle textstyle liminf _{n}X_{n}}{displaystyle textstyle liminf _{n}X_{n}} is a random variable, for every n,{displaystyle n,}{displaystyle n,} by the properties of limit inferior).






Corollary. Let




  • Xn→X{displaystyle X_{n}to X}{displaystyle X_{n}to X} pointwise (a.s.);


  • E⁡[Xn]≤C,{displaystyle operatorname {E} [X_{n}]leq C,}{displaystyle operatorname {E} [X_{n}]leq C,} for some constant C{displaystyle C}C (independent from n{displaystyle n}n);

  • E⁡[Y]>−;{displaystyle operatorname {E} [Y]>-infty ;}{displaystyle operatorname {E} [Y]>-infty ;}


  • Y≤Xn≤+∞{displaystyle -infty leq Yleq X_{n}leq +infty }{displaystyle -infty leq Yleq X_{n}leq +infty } (a.s.), for every n.{displaystyle n.}n.


Then E⁡[X]≤C.{displaystyle operatorname {E} [X]leq C.}{displaystyle operatorname {E} [X]leq C.}


Proof is by observing that X=lim infnXn{displaystyle textstyle X=liminf _{n}X_{n}}{displaystyle textstyle X=liminf _{n}X_{n}} (a.s.) and applying Fatou's lemma.



Dominated convergence theorem



Let {Xn}n{displaystyle {X_{n}}_{n}}{displaystyle {X_{n}}_{n}} be a sequence of random variables. If Xn→X{displaystyle X_{n}to X}{displaystyle X_{n}to X} pointwise (a.s.), |Xn|≤Y≤+∞{displaystyle |X_{n}|leq Yleq +infty }{displaystyle |X_{n}|leq Yleq +infty } (a.s.), and E⁡[Y]<∞{displaystyle operatorname {E} [Y]<infty }{displaystyle operatorname {E} [Y]<infty }. Then, according to the dominated convergence theorem,



  • the function X{displaystyle X}X is measurable (hence a random variable);


  • E⁡|X|<∞{displaystyle operatorname {E} |X|<infty }{displaystyle operatorname {E} |X|<infty };

  • all the expected values E⁡[Xn]{displaystyle operatorname {E} [X_{n}]}{displaystyle operatorname {E} [X_{n}]} and E⁡[X]{displaystyle operatorname {E} [X]}{displaystyle operatorname {E} [X]} are defined (do not have the form {displaystyle infty -infty }{displaystyle infty -infty });


  • limnE⁡[Xn]=E⁡[X]{displaystyle lim _{n}operatorname {E} [X_{n}]=operatorname {E} [X]}{displaystyle lim _{n}operatorname {E} [X_{n}]=operatorname {E} [X]} (both sides may be infinite);


  • limnE⁡|Xn−X|=0{displaystyle lim _{n}operatorname {E} |X_{n}-X|=0}{displaystyle lim _{n}operatorname {E} |X_{n}-X|=0}.



Relationship with characteristic function


The probability density function fX{displaystyle f_{X}}f_X of a scalar random variable X{displaystyle X}X is related to its characteristic function φX{displaystyle varphi _{X}}{displaystyle varphi _{X}} by the inversion formula:


fX(x)=12πRe−itxφX(t)dt.{displaystyle f_{X}(x)={frac {1}{2pi }}int _{mathbb {R} }e^{-itx}varphi _{X}(t),dt.}{displaystyle f_{X}(x)={frac {1}{2pi }}int _{mathbb {R} }e^{-itx}varphi _{X}(t),dt.}

For the expected value of g(X){displaystyle g(X)}g(X) (where g:R→R{displaystyle g:{mathbb {R} }to {mathbb {R} }}{displaystyle g:{mathbb {R} }to {mathbb {R} }} is a Borel function), we can use this inversion formula to obtain


E⁡[g(X)]=12πRg(x)[∫Re−itxφX(t)dt]dx.{displaystyle operatorname {E} [g(X)]={frac {1}{2pi }}int _{mathbb {R} }g(x)left[{int _{mathbb {R} }e^{-itx}varphi _{X}(t),dt}right]dx.}{displaystyle operatorname {E} [g(X)]={frac {1}{2pi }}int _{mathbb {R} }g(x)left[{int _{mathbb {R} }e^{-itx}varphi _{X}(t),dt}right]dx.}

If E⁡[g(X)]{displaystyle operatorname {E} [g(X)]}{displaystyle operatorname {E} [g(X)]} is finite, changing the order of integration, we get, in accordance with Fubini-Tonelli theorem,


E⁡[g(X)]=12πRG(t)φX(t)dt,{displaystyle operatorname {E} [g(X)]={frac {1}{2pi }}int _{mathbb {R} }G(t)varphi _{X}(t),dt,}{displaystyle operatorname {E} [g(X)]={frac {1}{2pi }}int _{mathbb {R} }G(t)varphi _{X}(t),dt,}

where


G(t)=∫Rg(x)e−itxdx{displaystyle G(t)=int _{mathbb {R} }g(x)e^{-itx},dx}{displaystyle G(t)=int _{mathbb {R} }g(x)e^{-itx},dx}

is the Fourier transform of g(x).{displaystyle g(x).}g(x). The expression for E⁡[g(X)]{displaystyle operatorname {E} [g(X)]}{displaystyle operatorname {E} [g(X)]} also follows directly from Plancherel theorem.



Uses and applications


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. P⁡(X∈A)=E⁡[1A]{displaystyle operatorname {P} ({Xin {mathcal {A}}})=operatorname {E} [{mathbf {1} }_{mathcal {A}}]}{displaystyle operatorname {P} ({Xin {mathcal {A}}})=operatorname {E} [{mathbf {1} }_{mathcal {A}}]}, where 1A{displaystyle {mathbf {1} }_{mathcal {A}}}{displaystyle {mathbf {1} }_{mathcal {A}}} is the indicator function of the set A{displaystyle {mathcal {A}}}{mathcal {A}}.




The mass of probability distribution is balanced at the expected value, here a Beta(α,β) distribution with expected value α/(α+β).


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


Var⁡(X)=E⁡[X2]−(E⁡[X])2.{displaystyle operatorname {Var} (X)=operatorname {E} [X^{2}]-(operatorname {E} [X])^{2}.}operatorname {Var} (X)=operatorname {E} [X^{2}]-(operatorname {E} [X])^{2}.

A very important application of the expectation value is in the field of quantum mechanics. The expectation value of a quantum mechanical operator A^{displaystyle {hat {A}}}{hat {A}} operating on a quantum state vector {displaystyle |psi rangle }|psi rangle is written as A^=⟨ψ|A|ψ{displaystyle langle {hat {A}}rangle =langle psi |A|psi rangle }langle {hat {A}}rangle =langle psi |A|psi rangle . The uncertainty in A^{displaystyle {hat {A}}}{hat {A}} can be calculated using the formula A)2=⟨A^2⟩A^2{displaystyle (Delta A)^{2}=langle {hat {A}}^{2}rangle -langle {hat {A}}rangle ^{2}}(Delta A)^{2}=langle {hat {A}}^{2}rangle -langle {hat {A}}rangle ^{2}.



The law of the unconscious statistician



The expected value of a measurable function of X{displaystyle X}X, g(X){displaystyle g(X)}g(X), given that X{displaystyle X}X has a probability density function f(x){displaystyle f(x)}f(x), is given by the inner product of f{displaystyle f}f and g{displaystyle g}g:


E⁡[g(X)]=∫Rg(x)f(x)dx.{displaystyle operatorname {E} [g(X)]=int _{mathbb {R} }g(x)f(x),dx.}{displaystyle operatorname {E} [g(X)]=int _{mathbb {R} }g(x)f(x),dx.}

This formula also holds in multidimensional case, when g{displaystyle g}g is a function of several random variables, and f{displaystyle f}f is their joint density.[5][6]



Alternative formula for expected value



Formula for non-negative random variables



Finite and countably infinite case


For a non-negative integer-valued random variable X:Ω{0,1,2,3,…}∪{+∞},{displaystyle X:Omega to {0,1,2,3,ldots }cup {+infty },}{displaystyle X:Omega to {0,1,2,3,ldots }cup {+infty },}


E⁡[X]=∑i=1∞P⁡(X≥i).{displaystyle operatorname {E} [X]=sum _{i=1}^{infty }operatorname {P} (Xgeq i).}{displaystyle operatorname {E} [X]=sum _{i=1}^{infty }operatorname {P} (Xgeq i).}






General case


If X:Ω[0,+∞]{displaystyle X:Omega to [0,+infty ]}{displaystyle X:Omega to [0,+infty ]} is a non-negative random variable, then


E⁡[X]=∫[0,+∞)P⁡(X≥x)dx=∫[0,+∞)P⁡(X>x)dx,{displaystyle operatorname {E} [X]=int limits _{[0,+infty )}operatorname {P} (Xgeq x),dx=int limits _{[0,+infty )}operatorname {P} (X>x),dx,}{displaystyle operatorname {E} [X]=int limits _{[0,+infty )}operatorname {P} (Xgeq x),dx=int limits _{[0,+infty )}operatorname {P} (X>x),dx,}

and


E⁡[X]=(R)∫0∞P⁡(X≥x)dx=(R)∫0∞P⁡(X>x)dx,{displaystyle operatorname {E} [X]={hbox{(R)}}int limits _{0}^{infty }operatorname {P} (Xgeq x),dx={hbox{(R)}}int limits _{0}^{infty }operatorname {P} (X>x),dx,}{displaystyle operatorname {E} [X]={hbox{(R)}}int limits _{0}^{infty }operatorname {P} (Xgeq x),dx={hbox{(R)}}int limits _{0}^{infty }operatorname {P} (X>x),dx,}

where (R)∫0∞{displaystyle {hbox{(R)}}textstyle int _{0}^{infty }}{displaystyle {hbox{(R)}}textstyle int _{0}^{infty }} denotes improper Riemann integral.







Formula for non-positive random variables


If X:Ω[−,0]{displaystyle X:Omega to [-infty ,0]}{displaystyle X:Omega to [-infty ,0]} is a non-positive random variable, then


E⁡[X]=−(−,0]P⁡(X≤x)dx=−(−,0]P⁡(X<x)dx,{displaystyle operatorname {E} [X]=-int limits _{(-infty ,0]}operatorname {P} (Xleq x),dx=-int limits _{(-infty ,0]}operatorname {P} (X<x),dx,}{displaystyle operatorname {E} [X]=-int limits _{(-infty ,0]}operatorname {P} (Xleq x),dx=-int limits _{(-infty ,0]}operatorname {P} (X<x),dx,}

and


E⁡[X]=−(R)∫0P⁡(X≤x)dx=−(R)∫0P⁡(X<x)dx,{displaystyle operatorname {E} [X]=-{hbox{(R)}}int limits _{-infty }^{0}operatorname {P} (Xleq x),dx=-{hbox{(R)}}int limits _{-infty }^{0}operatorname {P} (X<x),dx,}{displaystyle operatorname {E} [X]=-{hbox{(R)}}int limits _{-infty }^{0}operatorname {P} (Xleq x),dx=-{hbox{(R)}}int limits _{-infty }^{0}operatorname {P} (X<x),dx,}

where (R)∫0{displaystyle {hbox{(R)}}textstyle int _{-infty }^{0}}{displaystyle {hbox{(R)}}textstyle int _{-infty }^{0}} denotes improper Riemann integral.


This formula follows from that for the non-negative case applied to X.{displaystyle -X.}{displaystyle -X.}


If, in addition, X{displaystyle X}X is integer-valued, i.e. X:Ω{…,−3,−2,−1,0}∪{−}{displaystyle X:Omega to {ldots ,-3,-2,-1,0}cup {-infty }}{displaystyle X:Omega to {ldots ,-3,-2,-1,0}cup {-infty }}, then


E⁡[X]=−i=−1−P⁡(X≤i).{displaystyle operatorname {E} [X]=-sum _{i=-1}^{-infty }operatorname {P} (Xleq i).}{displaystyle operatorname {E} [X]=-sum _{i=-1}^{-infty }operatorname {P} (Xleq i).}


General case


If X{displaystyle X}X can be both positive and negative, then E⁡[X]=E⁡[X+]−E⁡[X−]{displaystyle operatorname {E} [X]=operatorname {E} [X_{+}]-operatorname {E} [X_{-}]}{displaystyle operatorname {E} [X]=operatorname {E} [X_{+}]-operatorname {E} [X_{-}]},
and the above results may be applied to X+{displaystyle X_{+}}X_{+} and X−{displaystyle X_{-}}{displaystyle X_{-}} separately.



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's properly finished. This problem had been debated for centuries, and many conflicting proposals and solutions had been suggested over the years, when it was posed in 1654 to Blaise Pascal by French writer and amateur mathematician Chevalier de Méré. 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 a now 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 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]


Three years later, in 1657, a Dutch mathematician Christiaan Huygens, who had just visited Paris, published a treatise (see Huygens (1657)) "De ratiociniis in ludo aleæ" on probability theory. In this book he considered the problem of points and presented a solution based on the same principle as the solutions of Pascal and Fermat. Huygens also 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). In this sense this book can be seen as the first successful attempt at laying down the foundations of the theory of probability.


In the foreword to his book, 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." (cited by Edwards (2002)). Thus, Huygens learned about de Méré's Problem in 1655 during his visit to France; later on in 1656 from his correspondence with Carcavi he learned that his method was essentially the same as Pascal's; so that before his book went to press in 1657 he knew about Pascal's priority in this subject.


Neither Pascal nor Huygens used the term "expectation" in its modern sense. In particular, Huygens writes: "That my Chance or Expectation to win any thing is worth just such a Sum, as wou'd procure me 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:


.mw-parser-output .templatequote{overflow:hidden;margin:1em 0;padding:0 40px}.mw-parser-output .templatequote .templatequotecite{line-height:1.5em;text-align:left;padding-left:1.6em;margin-top:0}

… 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.


The use of the letter E to denote expected value goes back to W.A. Whitworth in 1901,[8] who used a script E. The symbol has become popular since for English writers it meant "Expectation", for Germans "Erwartungswert", for Spanish "Esperanza matemática" and for French "Espérance mathématique".[9]



See also



  • Center of mass

  • Central tendency


  • Chebyshev's inequality (an inequality on location and scale parameters)

  • Conditional expectation

  • Expected value is also a key concept in economics, finance, and many other subjects

  • The general term expectation

  • 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)


  • Wald's equation for calculating the expected value of a random number of random variables



Notes





  1. ^ Sheldon M Ross (2007). "§2.4 Expectation of a random variable". Introduction to probability models (9th ed.). Academic Press. p. 38 ff. ISBN 0-12-598062-0..mw-parser-output cite.citation{font-style:inherit}.mw-parser-output q{quotes:"""""""'""'"}.mw-parser-output code.cs1-code{color:inherit;background:inherit;border:inherit;padding:inherit}.mw-parser-output .cs1-lock-free a{background:url("//upload.wikimedia.org/wikipedia/commons/thumb/6/65/Lock-green.svg/9px-Lock-green.svg.png")no-repeat;background-position:right .1em center}.mw-parser-output .cs1-lock-limited a,.mw-parser-output .cs1-lock-registration a{background:url("//upload.wikimedia.org/wikipedia/commons/thumb/d/d6/Lock-gray-alt-2.svg/9px-Lock-gray-alt-2.svg.png")no-repeat;background-position:right .1em center}.mw-parser-output .cs1-lock-subscription a{background:url("//upload.wikimedia.org/wikipedia/commons/thumb/a/aa/Lock-red-alt-2.svg/9px-Lock-red-alt-2.svg.png")no-repeat;background-position:right .1em center}.mw-parser-output .cs1-subscription,.mw-parser-output .cs1-registration{color:#555}.mw-parser-output .cs1-subscription span,.mw-parser-output .cs1-registration span{border-bottom:1px dotted;cursor:help}.mw-parser-output .cs1-hidden-error{display:none;font-size:100%}.mw-parser-output .cs1-visible-error{font-size:100%}.mw-parser-output .cs1-subscription,.mw-parser-output .cs1-registration,.mw-parser-output .cs1-format{font-size:95%}.mw-parser-output .cs1-kern-left,.mw-parser-output .cs1-kern-wl-left{padding-left:0.2em}.mw-parser-output .cs1-kern-right,.mw-parser-output .cs1-kern-wl-right{padding-right:0.2em}


  2. ^ Richard W Hamming (1991). "§2.5 Random variables, mean and the expected value". The art of probability for scientists and engineers. Addison–Wesley. p. 64 ff. ISBN 0-201-40686-1.


  3. ^ Richard W Hamming (1991). "Example 8.7–1 The Cauchy distribution". The art of probability for scientists and engineers. 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!


  4. ^ Gordon, Lawrence; Loeb, Martin (November 2002). "The Economics of Information Security Investment". ACM Transactions on Information and System Security. 5 (4): 438–457. doi:10.1145/581271.581274.


  5. ^ Expectation Value, retrieved August 8, 2017


  6. ^ Papoulis, A. (1984), Probability, Random Variables, and Stochastic Processes, New York: McGraw–Hill, pp. 139–152


  7. ^ "Ore, Pascal and the Invention of Probability Theory". The American Mathematical Monthly. 67 (5): 409–419. 1960. doi:10.2307/2309286.


  8. ^ Whitworth, W.A. (1901) Choice and Chance with One Thousand Exercises. Fifth edition. Deighton Bell, Cambridge. [Reprinted by Hafner Publishing Co., New York, 1959.]


  9. ^ "Earliest uses of symbols in probability and statistics".




Literature


.mw-parser-output .refbegin{font-size:90%;margin-bottom:0.5em}.mw-parser-output .refbegin-hanging-indents>ul{list-style-type:none;margin-left:0}.mw-parser-output .refbegin-hanging-indents>ul>li,.mw-parser-output .refbegin-hanging-indents>dl>dd{margin-left:0;padding-left:3.2em;text-indent:-3.2em;list-style:none}.mw-parser-output .refbegin-100{font-size:100%}



  • 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 (English translation, published in 1714:).










Popular posts from this blog

Lambaréné

Chris Pine

Kashihara Line