Stein's lemma

Theorem of probability theory

Stein's lemma, named in honor of Charles Stein, is a theorem of probability theory that is of interest primarily because of its applications to statistical inference — in particular, to James–Stein estimation and empirical Bayes methods — and its applications to portfolio choice theory.[1] The theorem gives a formula for the covariance of one random variable with the value of a function of another, when the two random variables are jointly normally distributed.

Note that the name "Stein's lemma" is also commonly used[2] to refer to a different result in the area of statistical hypothesis testing, which connects the error exponents in hypothesis testing with the Kullback–Leibler divergence. This result is also known as the Chernoff–Stein lemma[3] and is not related to the lemma discussed in this article.

Statement of the lemma

Suppose X is a normally distributed random variable with expectation μ and variance σ2. Further suppose g is a differentiable function for which the two expectations E(g(X) (X − μ)) and E(g ′(X)) both exist. (The existence of the expectation of any random variable is equivalent to the finiteness of the expectation of its absolute value.) Then

E ( g ( X ) ( X μ ) ) = σ 2 E ( g ( X ) ) . {\displaystyle E{\bigl (}g(X)(X-\mu ){\bigr )}=\sigma ^{2}E{\bigl (}g'(X){\bigr )}.}

In general, suppose X and Y are jointly normally distributed. Then

Cov ( g ( X ) , Y ) = Cov ( X , Y ) E ( g ( X ) ) . {\displaystyle \operatorname {Cov} (g(X),Y)=\operatorname {Cov} (X,Y)E(g'(X)).}

For a general multivariate Gaussian random vector ( X 1 , . . . , X n ) N ( μ , Σ ) {\displaystyle (X_{1},...,X_{n})\sim N(\mu ,\Sigma )} it follows that

E ( g ( X ) ( X μ ) ) = Σ E ( g ( X ) ) . {\displaystyle E{\bigl (}g(X)(X-\mu ){\bigr )}=\Sigma \cdot E{\bigl (}\nabla g(X){\bigr )}.}

Proof

The univariate probability density function for the univariate normal distribution with expectation 0 and variance 1 is

φ ( x ) = 1 2 π e x 2 / 2 {\displaystyle \varphi (x)={1 \over {\sqrt {2\pi }}}e^{-x^{2}/2}}

Since x exp ( x 2 / 2 ) d x = exp ( x 2 / 2 ) {\displaystyle \int x\exp(-x^{2}/2)\,dx=-\exp(-x^{2}/2)} we get from integration by parts:

E [ g ( X ) X ] = 1 2 π g ( x ) x exp ( x 2 / 2 ) d x = 1 2 π g ( x ) exp ( x 2 / 2 ) d x = E [ g ( X ) ] {\displaystyle E[g(X)X]={\frac {1}{\sqrt {2\pi }}}\int g(x)x\exp(-x^{2}/2)\,dx={\frac {1}{\sqrt {2\pi }}}\int g'(x)\exp(-x^{2}/2)\,dx=E[g'(X)]} .

The case of general variance σ 2 {\displaystyle \sigma ^{2}} follows by substitution.

More general statement

Isserlis' theorem is equivalently stated as

E ( X 1 f ( X 1 , , X n ) ) = i = 1 n Cov ( X 1 X i ) E ( X i f ( X 1 , , X n ) ) . {\displaystyle \operatorname {E} (X_{1}f(X_{1},\ldots ,X_{n}))=\sum _{i=1}^{n}\operatorname {Cov} (X_{1}X_{i})\operatorname {E} (\partial _{X_{i}}f(X_{1},\ldots ,X_{n})).}
where ( X 1 , X n ) {\displaystyle (X_{1},\dots X_{n})} is a zero-mean multivariate normal random vector.

Suppose X is in an exponential family, that is, X has the density

f η ( x ) = exp ( η T ( x ) Ψ ( η ) ) h ( x ) . {\displaystyle f_{\eta }(x)=\exp(\eta 'T(x)-\Psi (\eta ))h(x).}

Suppose this density has support ( a , b ) {\displaystyle (a,b)} where a , b {\displaystyle a,b} could be , {\displaystyle -\infty ,\infty } and as x a  or  b {\displaystyle x\rightarrow a{\text{ or }}b} , exp ( η T ( x ) ) h ( x ) g ( x ) 0 {\displaystyle \exp(\eta 'T(x))h(x)g(x)\rightarrow 0} where g {\displaystyle g} is any differentiable function such that E | g ( X ) | < {\displaystyle E|g'(X)|<\infty } or exp ( η T ( x ) ) h ( x ) 0 {\displaystyle \exp(\eta 'T(x))h(x)\rightarrow 0} if a , b {\displaystyle a,b} finite. Then

E [ ( h ( X ) h ( X ) + η i T i ( X ) ) g ( X ) ] = E [ g ( X ) ] . {\displaystyle E\left[\left({\frac {h'(X)}{h(X)}}+\sum \eta _{i}T_{i}'(X)\right)\cdot g(X)\right]=-E[g'(X)].}

The derivation is same as the special case, namely, integration by parts.

If we only know X {\displaystyle X} has support R {\displaystyle \mathbb {R} } , then it could be the case that E | g ( X ) | <  and  E | g ( X ) | < {\displaystyle E|g(X)|<\infty {\text{ and }}E|g'(X)|<\infty } but lim x f η ( x ) g ( x ) 0 {\displaystyle \lim _{x\rightarrow \infty }f_{\eta }(x)g(x)\not =0} . To see this, simply put g ( x ) = 1 {\displaystyle g(x)=1} and f η ( x ) {\displaystyle f_{\eta }(x)} with infinitely spikes towards infinity but still integrable. One such example could be adapted from f ( x ) = { 1 x [ n , n + 2 n ) 0 otherwise {\displaystyle f(x)={\begin{cases}1&x\in [n,n+2^{-n})\\0&{\text{otherwise}}\end{cases}}} so that f {\displaystyle f} is smooth.

Extensions to elliptically-contoured distributions also exist.[4][5][6]

See also

References

  1. ^ Ingersoll, J., Theory of Financial Decision Making, Rowman and Littlefield, 1987: 13-14.
  2. ^ Csiszár, Imre; Körner, János (2011). Information Theory: Coding Theorems for Discrete Memoryless Systems. Cambridge University Press. p. 14. ISBN 9781139499989.
  3. ^ Thomas M. Cover, Joy A. Thomas (2006). Elements of Information Theory. John Wiley & Sons, New York. ISBN 9781118585771.
  4. ^ Cellier, Dominique; Fourdrinier, Dominique; Robert, Christian (1989). "Robust shrinkage estimators of the location parameter for elliptically symmetric distributions". Journal of Multivariate Analysis. 29 (1): 39–52. doi:10.1016/0047-259X(89)90075-4.
  5. ^ Hamada, Mahmoud; Valdez, Emiliano A. (2008). "CAPM and option pricing with elliptically contoured distributions". The Journal of Risk & Insurance. 75 (2): 387–409. CiteSeerX 10.1.1.573.4715. doi:10.1111/j.1539-6975.2008.00265.x.
  6. ^ Landsman, Zinoviy; Nešlehová, Johanna (2008). "Stein's Lemma for elliptical random vectors". Journal of Multivariate Analysis. 99 (5): 912––927. doi:10.1016/j.jmva.2007.05.006.