Sufficient statistic
   HOME

TheInfoList



OR:

In
statistics Statistics (from German: '' Statistik'', "description of a state, a country") is the discipline that concerns the collection, organization, analysis, interpretation, and presentation of data. In applying statistics to a scientific, indust ...
, a
statistic A statistic (singular) or sample statistic is any quantity computed from values in a sample which is considered for a statistical purpose. Statistical purposes include estimating a population parameter, describing a sample, or evaluating a hypo ...
is ''sufficient'' with respect to a
statistical model A statistical model is a mathematical model that embodies a set of statistical assumptions concerning the generation of sample data (and similar data from a larger population). A statistical model represents, often in considerably idealized form ...
and its associated unknown
parameter A parameter (), generally, is any characteristic that can help in defining or classifying a particular system (meaning an event, project, object, situation, etc.). That is, a parameter is an element of a system that is useful, or critical, when ...
if "no other statistic that can be calculated from the same
sample Sample or samples may refer to: Base meaning * Sample (statistics), a subset of a population – complete data set * Sample (signal), a digital discrete sample of a continuous analog signal * Sample (material), a specimen or small quantity of ...
provides any additional information as to the value of the parameter". In particular, a statistic is sufficient for a
family Family (from la, familia) is a group of people related either by consanguinity (by recognized birth) or affinity (by marriage or other relationship). The purpose of the family is to maintain the well-being of its members and of society. Idea ...
of probability distributions if the sample from which it is calculated gives no additional information than the statistic, as to which of those probability distributions is the sampling distribution. A related concept is that of linear sufficiency, which is weaker than ''sufficiency'' but can be applied in some cases where there is no sufficient statistic, although it is restricted to linear estimators. The Kolmogorov structure function deals with individual finite data; the related notion there is the algorithmic sufficient statistic. The concept is due to
Sir Ronald Fisher Sir Ronald Aylmer Fisher (17 February 1890 – 29 July 1962) was a British polymath who was active as a mathematician, statistician, biologist, geneticist, and academic. For his work in statistics, he has been described as "a genius who a ...
in 1920. Stephen Stigler noted in 1973 that the concept of sufficiency had fallen out of favor in descriptive statistics because of the strong dependence on an assumption of the distributional form (see Pitman–Koopman–Darmois theorem below), but remained very important in theoretical work.


Background

Roughly, given a set \mathbf of
independent identically distributed In probability theory and statistics, a collection of random variables is independent and identically distributed if each random variable has the same probability distribution as the others and all are mutually independent. This property is us ...
data conditioned on an unknown parameter \theta, a sufficient statistic is a function T(\mathbf) whose value contains all the information needed to compute any estimate of the parameter (e.g. a
maximum likelihood In statistics, maximum likelihood estimation (MLE) is a method of estimating the parameters of an assumed probability distribution, given some observed data. This is achieved by maximizing a likelihood function so that, under the assumed stat ...
estimate). Due to the factorization theorem ( see below), for a sufficient statistic T(\mathbf), the probability density can be written as f_(x) = h(x) \, g(\theta, T(x)). From this factorization, it can easily be seen that the maximum likelihood estimate of \theta will interact with \mathbf only through T(\mathbf). Typically, the sufficient statistic is a simple function of the data, e.g. the sum of all the data points. More generally, the "unknown parameter" may represent a vector of unknown quantities or may represent everything about the model that is unknown or not fully specified. In such a case, the sufficient statistic may be a set of functions, called a ''jointly sufficient statistic''. Typically, there are as many functions as there are parameters. For example, for a
Gaussian distribution In statistics, a normal distribution or Gaussian distribution is a type of continuous probability distribution for a real-valued random variable. The general form of its probability density function is : f(x) = \frac e^ The parameter \mu ...
with unknown
mean There are several kinds of mean in mathematics, especially in statistics. Each mean serves to summarize a given group of data, often to better understand the overall value ( magnitude and sign) of a given data set. For a data set, the '' ar ...
and variance, the jointly sufficient statistic, from which maximum likelihood estimates of both parameters can be estimated, consists of two functions, the sum of all data points and the sum of all squared data points (or equivalently, the sample mean and
sample variance In probability theory and statistics, variance is the expectation of the squared deviation of a random variable from its population mean or sample mean. Variance is a measure of dispersion, meaning it is a measure of how far a set of numbe ...
). The concept is equivalent to the statement that,
conditional Conditional (if then) may refer to: *Causal conditional, if X then Y, where X is a cause of Y *Conditional probability, the probability of an event A given that another event B has occurred *Conditional proof, in logic: a proof that asserts a co ...
on the value of a sufficient statistic for a parameter, the joint probability distribution of the data does not depend on that parameter. Both the statistic and the underlying parameter can be vectors.


Mathematical definition

A statistic ''t'' = ''T''(''X'') is sufficient for underlying parameter ''θ'' precisely if the
conditional probability distribution In probability theory and statistics, given two jointly distributed random variables X and Y, the conditional probability distribution of Y given X is the probability distribution of Y when X is known to be a particular value; in some cases the ...
of the data ''X'', given the statistic ''t'' = ''T''(''X''), does not depend on the parameter ''θ''. Alternatively, one can say the statistic ''T''(''X'') is sufficient for ''θ'' if its
mutual information In probability theory and information theory, the mutual information (MI) of two random variables is a measure of the mutual dependence between the two variables. More specifically, it quantifies the " amount of information" (in units such ...
with ''θ'' equals the mutual information between ''X'' and ''θ''. In other words, the data processing inequality becomes an equality: :I\bigl(\theta ; T(X)\bigr) = I(\theta ; X)


Example

As an example, the sample mean is sufficient for the mean (''μ'') of a
normal distribution In statistics, a normal distribution or Gaussian distribution is a type of continuous probability distribution for a real-valued random variable. The general form of its probability density function is : f(x) = \frac e^ The parameter \mu ...
with known variance. Once the sample mean is known, no further information about ''μ'' can be obtained from the sample itself. On the other hand, for an arbitrary distribution the median is not sufficient for the mean: even if the median of the sample is known, knowing the sample itself would provide further information about the population mean. For example, if the observations that are less than the median are only slightly less, but observations exceeding the median exceed it by a large amount, then this would have a bearing on one's inference about the population mean.


Fisher–Neyman factorization theorem

'' Fisher's factorization theorem'' or ''factorization criterion'' provides a convenient characterization of a sufficient statistic. If the
probability density function In probability theory, a probability density function (PDF), or density of a continuous random variable, is a function whose value at any given sample (or point) in the sample space (the set of possible values taken by the random variable) ca ...
is ƒ''θ''(''x''), then ''T'' is sufficient for ''θ''
if and only if In logic and related fields such as mathematics and philosophy, "if and only if" (shortened as "iff") is a biconditional logical connective between statements, where either both statements are true or both are false. The connective is bic ...
nonnegative functions ''g'' and ''h'' can be found such that : f_\theta(x)=h(x) \, g_\theta(T(x)), i.e. the density ƒ can be factored into a product such that one factor, ''h'', does not depend on ''θ'' and the other factor, which does depend on ''θ'', depends on ''x'' only through ''T''(''x''). A general proof of this was given by Halmos and Savage and the theorem is sometimes referred to as the Halmos-Savage factorization theorem. The proofs below handle special cases, but an alternative general proof along the same lines can be given. It is easy to see that if ''F''(''t'') is a one-to-one function and ''T'' is a sufficient statistic, then ''F''(''T'') is a sufficient statistic. In particular we can multiply a sufficient statistic by a nonzero constant and get another sufficient statistic.


Likelihood principle interpretation

An implication of the theorem is that when using likelihood-based inference, two sets of data yielding the same value for the sufficient statistic ''T''(''X'') will always yield the same inferences about ''θ''. By the factorization criterion, the likelihood's dependence on ''θ'' is only in conjunction with ''T''(''X''). As this is the same in both cases, the dependence on ''θ'' will be the same as well, leading to identical inferences.


Proof

Due to Hogg and Craig. Let X_1, X_2, \ldots, X_n, denote a random sample from a distribution having the pdf ''f''(''x'', ''θ'') for ''ι'' < ''θ'' < ''δ''. Let ''Y''1 = ''u''1(''X''1, ''X''2, ..., ''X''''n'') be a statistic whose pdf is ''g''1(''y''1; ''θ''). What we want to prove is that ''Y''1 = ''u''1(''X''1, ''X''2, ..., ''X''''n'') is a sufficient statistic for ''θ'' if and only if, for some function ''H'', : \prod_^n f(x_i; \theta) = g_1 \left _1 (x_1, x_2, \dots, x_n); \theta \rightH(x_1, x_2, \dots, x_n). First, suppose that : \prod_^n f(x_i; \theta) = g_1 \left _1 (x_1, x_2, \dots, x_n); \theta \rightH(x_1, x_2, \dots, x_n). We shall make the transformation ''y''''i'' = ''u''i(''x''1, ''x''2, ..., ''x''''n''), for ''i'' = 1, ..., ''n'', having inverse functions ''x''''i'' = ''w''''i''(''y''1, ''y''2, ..., ''y''''n''), for ''i'' = 1, ..., ''n'', and
Jacobian In mathematics, a Jacobian, named for Carl Gustav Jacob Jacobi, may refer to: * Jacobian matrix and determinant * Jacobian elliptic functions * Jacobian variety *Intermediate Jacobian In mathematics, the intermediate Jacobian of a compact Kähle ...
J = \left _i/y_j \right. Thus, : \prod_^n f \left w_i(y_1, y_2, \dots, y_n); \theta \right = , J, g_1 (y_1; \theta) H \left w_1(y_1, y_2, \dots, y_n), \dots, w_n(y_1, y_2, \dots, y_n) \right The left-hand member is the joint pdf ''g''(''y''1, ''y''2, ..., ''y''''n''; θ) of ''Y''1 = ''u''1(''X''1, ..., ''X''''n''), ..., ''Y''''n'' = ''u''''n''(''X''1, ..., ''X''''n''). In the right-hand member, g_1(y_1;\theta) is the pdf of Y_1, so that H w_1, \dots , w_n, J, is the quotient of g(y_1,\dots,y_n;\theta) and g_1(y_1;\theta); that is, it is the conditional pdf h(y_2, \dots, y_n \mid y_1; \theta) of Y_2,\dots,Y_n given Y_1=y_1. But H(x_1,x_2,\dots,x_n), and thus H\left _1(y_1,\dots,y_n), \dots, w_n(y_1, \dots, y_n))\right/math>, was given not to depend upon \theta. Since \theta was not introduced in the transformation and accordingly not in the Jacobian J, it follows that h(y_2, \dots, y_n \mid y_1; \theta) does not depend upon \theta and that Y_1 is a sufficient statistics for \theta. The converse is proven by taking: :g(y_1,\dots,y_n;\theta)=g_1(y_1; \theta) h(y_2, \dots, y_n \mid y_1), where h(y_2, \dots, y_n \mid y_1) does not depend upon \theta because Y_2 ... Y_n depend only upon X_1 ... X_n, which are independent on \Theta when conditioned by Y_1, a sufficient statistics by hypothesis. Now divide both members by the absolute value of the non-vanishing Jacobian J, and replace y_1, \dots, y_n by the functions u_1(x_1, \dots, x_n), \dots, u_n(x_1,\dots, x_n) in x_1,\dots, x_n. This yields :\frac=g_1\left _1(x_1,\dots,x_n); \theta\right\frac where J^* is the Jacobian with y_1,\dots,y_n replaced by their value in terms x_1, \dots, x_n. The left-hand member is necessarily the joint pdf f(x_1;\theta)\cdots f(x_n;\theta) of X_1,\dots,X_n. Since h(y_2,\dots,y_n\mid y_1), and thus h(u_2,\dots,u_n\mid u_1), does not depend upon \theta, then :H(x_1,\dots,x_n)=\frac is a function that does not depend upon \theta.


Another proof

A simpler more illustrative proof is as follows, although it applies only in the discrete case. We use the shorthand notation to denote the joint probability density of (X, T(X)) by f_\theta(x,t). Since T is a function of X, we have f_\theta(x,t) = f_\theta(x), as long as t = T(x) and zero otherwise. Therefore: : \begin f_\theta(x) & = f_\theta(x,t) \\ pt& = f_\theta (x\mid t) f_\theta(t) \\ pt& = f(x\mid t) f_\theta(t) \end with the last equality being true by the definition of sufficient statistics. Thus f_\theta(x)=a(x) b_\theta(t) with a(x) = f_(x) and b_\theta(t) = f_\theta(t). Conversely, if f_\theta(x)=a(x) b_\theta(t), we have : \begin f_\theta(t) & = \sum _ f_\theta(x, t) \\ pt& = \sum _ f_\theta(x) \\ pt& = \sum _ a(x) b_\theta(t) \\ pt& = \left( \sum _ a(x) \right) b_\theta(t). \end With the first equality by the definition of pdf for multiple variables, the second by the remark above, the third by hypothesis, and the fourth because the summation is not over t. Let f_(x) denote the conditional probability density of X given T(X). Then we can derive an explicit expression for this: : \begin f_(x) & = \frac \\ pt& = \frac \\ pt& = \frac \\ pt& = \frac. \end With the first equality by definition of conditional probability density, the second by the remark above, the third by the equality proven above, and the fourth by simplification. This expression does not depend on \theta and thus T is a sufficient statistic.


Minimal sufficiency

A sufficient statistic is minimal sufficient if it can be represented as a function of any other sufficient statistic. In other words, ''S''(''X'') is minimal sufficient if and only if #''S''(''X'') is sufficient, and #if ''T''(''X'') is sufficient, then there exists a function ''f'' such that ''S''(''X'') = ''f''(''T''(''X'')). Intuitively, a minimal sufficient statistic ''most efficiently'' captures all possible information about the parameter ''θ''. A useful characterization of minimal sufficiency is that when the density ''f''θ exists, ''S''(''X'') is minimal sufficient if and only if :\frac is independent of ''θ'' :\Longleftrightarrow ''S''(''x'') = ''S''(''y'') This follows as a consequence from Fisher's factorization theorem stated above. A case in which there is no minimal sufficient statistic was shown by Bahadur, 1954. However, under mild conditions, a minimal sufficient statistic does always exist. In particular, in Euclidean space, these conditions always hold if the random variables (associated with P_\theta ) are all discrete or are all continuous. If there exists a minimal sufficient statistic, and this is usually the case, then every
complete Complete may refer to: Logic * Completeness (logic) * Completeness of a theory, the property of a theory that every formula in the theory's language or its negation is provable Mathematics * The completeness of the real numbers, which implies t ...
sufficient statistic is necessarily minimal sufficient(note that this statement does not exclude a pathological case in which a complete sufficient exists while there is no minimal sufficient statistic). While it is hard to find cases in which a minimal sufficient statistic does not exist, it is not so hard to find cases in which there is no complete statistic. The collection of likelihood ratios \left\ for i = 1, ..., k, is a minimal sufficient statistic if the parameter space is discrete \left\.


Examples


Bernoulli distribution

If ''X''1, ...., ''X''''n'' are independent Bernoulli-distributed random variables with expected value ''p'', then the sum ''T''(''X'') = ''X''1 + ... + ''X''''n'' is a sufficient statistic for ''p'' (here 'success' corresponds to ''X''''i'' = 1 and 'failure' to ''X''''i'' = 0; so ''T'' is the total number of successes) This is seen by considering the joint probability distribution: : \Pr\=\Pr\. Because the observations are independent, this can be written as : p^(1-p)^ p^(1-p)^\cdots p^(1-p)^ and, collecting powers of ''p'' and 1 − ''p'', gives : p^(1-p)^=p^(1-p)^ which satisfies the factorization criterion, with ''h''(''x'') = 1 being just a constant. Note the crucial feature: the unknown parameter ''p'' interacts with the data ''x'' only via the statistic ''T''(''x'') = Σ ''x''''i''. As a concrete application, this gives a procedure for distinguishing a fair coin from a biased coin.


Uniform distribution

If ''X''1, ...., ''X''''n'' are independent and uniformly distributed on the interval ,''θ'' then ''T''(''X'') = max(''X''1, ..., ''X''''n'') is sufficient for θ — the
sample maximum In statistics, the sample maximum and sample minimum, also called the largest observation and smallest observation, are the values of the greatest and least elements of a sample. They are basic summary statistics, used in descriptive statistics ...
is a sufficient statistic for the population maximum. To see this, consider the joint
probability density function In probability theory, a probability density function (PDF), or density of a continuous random variable, is a function whose value at any given sample (or point) in the sample space (the set of possible values taken by the random variable) ca ...
of ''X''  (''X''1,...,''X''''n''). Because the observations are independent, the pdf can be written as a product of individual densities :\begin f_(x_1,\ldots,x_n) &= \frac\mathbf_ \cdots \frac\mathbf_ \\ pt &= \frac \mathbf_\mathbf_ \end where 1 is the indicator function. Thus the density takes form required by the Fisher–Neyman factorization theorem, where ''h''(''x'') = 1, and the rest of the expression is a function of only ''θ'' and ''T''(''x'') = max. In fact, the minimum-variance unbiased estimator (MVUE) for ''θ'' is : \fracT(X). This is the sample maximum, scaled to correct for the
bias Bias is a disproportionate weight ''in favor of'' or ''against'' an idea or thing, usually in a way that is closed-minded, prejudicial, or unfair. Biases can be innate or learned. People may develop biases for or against an individual, a group ...
, and is MVUE by the Lehmann–Scheffé theorem. Unscaled sample maximum ''T''(''X'') is the maximum likelihood estimator for ''θ''.


Uniform distribution (with two parameters)

If X_1,...,X_n are independent and uniformly distributed on the interval alpha, \beta/math> (where \alpha and \beta are unknown parameters), then T(X_1^n)=\left(\min_X_i,\max_X_i\right) is a two-dimensional sufficient statistic for (\alpha\, , \, \beta). To see this, consider the joint
probability density function In probability theory, a probability density function (PDF), or density of a continuous random variable, is a function whose value at any given sample (or point) in the sample space (the set of possible values taken by the random variable) ca ...
of X_1^n=(X_1,\ldots,X_n). Because the observations are independent, the pdf can be written as a product of individual densities, i.e. :\begin f_(x_1^n) &= \prod_^n \left(\right) \mathbf_ = \left(\right)^n \mathbf_ \\ &= \left(\right)^n \mathbf_ \mathbf_. \end The joint density of the sample takes the form required by the Fisher–Neyman factorization theorem, by letting :\begin h(x_1^n)= 1, \quad g_(x_1^n)= \left(\right)^n \mathbf_ \mathbf_. \end Since h(x_1^n) does not depend on the parameter (\alpha, \beta) and g_(x_1^n) depends only on x_1^n through the function T(X_1^n)= \left(\min_X_i,\max_X_i\right), the Fisher–Neyman factorization theorem implies T(X_1^n) = \left(\min_X_i,\max_X_i\right) is a sufficient statistic for (\alpha\, , \, \beta).


Poisson distribution

If ''X''1, ...., ''X''''n'' are independent and have a Poisson distribution with parameter ''λ'', then the sum ''T''(''X'') = ''X''1 + ... + ''X''''n'' is a sufficient statistic for ''λ''. To see this, consider the joint probability distribution: : \Pr(X=x)=P(X_1=x_1,X_2=x_2,\ldots,X_n=x_n). Because the observations are independent, this can be written as : \cdot \cdots which may be written as : e^ \lambda^ \cdot which shows that the factorization criterion is satisfied, where ''h''(''x'') is the reciprocal of the product of the factorials. Note the parameter λ interacts with the data only through its sum ''T''(''X'').


Normal distribution

If X_1,\ldots,X_n are independent and normally distributed with expected value \theta (a parameter) and known finite variance \sigma^2, then :T(X_1^n)=\overline=\frac1n\sum_^nX_i is a sufficient statistic for \theta. To see this, consider the joint
probability density function In probability theory, a probability density function (PDF), or density of a continuous random variable, is a function whose value at any given sample (or point) in the sample space (the set of possible values taken by the random variable) ca ...
of X_1^n=(X_1,\dots,X_n). Because the observations are independent, the pdf can be written as a product of individual densities, i.e. :\begin f_(x_1^n) & = \prod_^n \frac \exp \left (-\frac \right ) \\ pt &= (2\pi\sigma^2)^ \exp \left ( -\sum_^n \frac \right ) \\ pt & = (2\pi\sigma^2)^ \exp \left (-\sum_^n \frac \right ) \\ pt & = (2\pi\sigma^2)^ \exp \left( - \left(\sum_^n(x_i-\overline)^2 + \sum_^n(\theta-\overline)^2 -2\sum_^n(x_i-\overline)(\theta-\overline)\right) \right) \\ pt &= (2\pi\sigma^2)^ \exp \left( - \left (\sum_^n(x_i-\overline)^2 + n(\theta-\overline)^2 \right ) \right ) && \sum_^n(x_i-\overline)(\theta-\overline)=0 \\ pt &= (2\pi\sigma^2)^ \exp \left( - \sum_^n (x_i-\overline)^2 \right ) \exp \left (-\frac (\theta-\overline)^2 \right ) \end The joint density of the sample takes the form required by the Fisher–Neyman factorization theorem, by letting :\begin h(x_1^n) &= (2\pi\sigma^2)^ \exp \left( - \sum_^n (x_i-\overline)^2 \right ) \\ ptg_\theta(x_1^n) &= \exp \left (-\frac (\theta-\overline)^2 \right ) \end Since h(x_1^n) does not depend on the parameter \theta and g_(x_1^n) depends only on x_1^n through the function :T(X_1^n)=\overline=\frac1n\sum_^nX_i, the Fisher–Neyman factorization theorem implies T(X_1^n) is a sufficient statistic for \theta. If \sigma^2 is unknown and since s^2 = \frac \sum_^n \left(x_i - \overline \right)^2 , the above likelihood can be rewritten as :\begin f_(x_1^n)= (2\pi\sigma^2)^ \exp \left( -\fracs^2 \right) \exp \left (-\frac (\theta-\overline)^2 \right ) . \end The Fisher–Neyman factorization theorem still holds and implies that (\overline,s^2) is a joint sufficient statistic for ( \theta , \sigma^2) .


Exponential distribution

If X_1,\dots,X_n are independent and exponentially distributed with expected value ''θ'' (an unknown real-valued positive parameter), then T(X_1^n)=\sum_^nX_i is a sufficient statistic for θ. To see this, consider the joint
probability density function In probability theory, a probability density function (PDF), or density of a continuous random variable, is a function whose value at any given sample (or point) in the sample space (the set of possible values taken by the random variable) ca ...
of X_1^n=(X_1,\dots,X_n). Because the observations are independent, the pdf can be written as a product of individual densities, i.e. :\begin f_(x_1^n) &= \prod_^n \, e^ = \, e^. \end The joint density of the sample takes the form required by the Fisher–Neyman factorization theorem, by letting :\begin h(x_1^n)= 1,\,\,\, g_(x_1^n)= \, e^. \end Since h(x_1^n) does not depend on the parameter \theta and g_(x_1^n) depends only on x_1^n through the function T(X_1^n)=\sum_^nX_i the Fisher–Neyman factorization theorem implies T(X_1^n)=\sum_^nX_i is a sufficient statistic for \theta.


Gamma distribution

If X_1,\dots,X_n are independent and distributed as a \Gamma(\alpha \, , \, \beta) , where \alpha and \beta are unknown parameters of a
Gamma distribution In probability theory and statistics, the gamma distribution is a two- parameter family of continuous probability distributions. The exponential distribution, Erlang distribution, and chi-square distribution are special cases of the gamma di ...
, then T(X_1^n) = \left( \prod_^n , \sum_^n X_i \right) is a two-dimensional sufficient statistic for (\alpha, \beta). To see this, consider the joint
probability density function In probability theory, a probability density function (PDF), or density of a continuous random variable, is a function whose value at any given sample (or point) in the sample space (the set of possible values taken by the random variable) ca ...
of X_1^n=(X_1,\dots,X_n). Because the observations are independent, the pdf can be written as a product of individual densities, i.e. :\begin f_(x_1^n) &= \prod_^n \left(\right) x_i^ e^ \\ pt &= \left(\right)^n \left(\prod_^n x_i\right)^ e^. \end The joint density of the sample takes the form required by the Fisher–Neyman factorization theorem, by letting :\begin h(x_1^n)= 1,\,\,\, g_(x_1^n)= \left(\right)^n \left(\prod_^n x_i\right)^ e^. \end Since h(x_1^n) does not depend on the parameter (\alpha\, , \, \beta) and g_(x_1^n) depends only on x_1^n through the function T(x_1^n)= \left( \prod_^n x_i, \sum_^n x_i \right), the Fisher–Neyman factorization theorem implies T(X_1^n)= \left( \prod_^n X_i, \sum_^n X_i \right) is a sufficient statistic for (\alpha\, , \, \beta).


Rao–Blackwell theorem

Sufficiency finds a useful application in the
Rao–Blackwell theorem In statistics, the Rao–Blackwell theorem, sometimes referred to as the Rao–Blackwell–Kolmogorov theorem, is a result which characterizes the transformation of an arbitrarily crude estimator into an estimator that is optimal by the mean-squ ...
, which states that if ''g''(''X'') is any kind of estimator of ''θ'', then typically the conditional expectation of ''g''(''X'') given sufficient statistic ''T''(''X'') is a better (in the sense of having lower variance) estimator of ''θ'', and is never worse. Sometimes one can very easily construct a very crude estimator ''g''(''X''), and then evaluate that conditional expected value to get an estimator that is in various senses optimal.


Exponential family

According to the Pitman–Koopman–Darmois theorem, among families of probability distributions whose domain does not vary with the parameter being estimated, only in exponential families is there a sufficient statistic whose dimension remains bounded as sample size increases. Intuitively, this states that nonexponential families of distributions on the real line require
nonparametric statistics Nonparametric statistics is the branch of statistics that is not based solely on parametrized families of probability distributions (common examples of parameters are the mean and variance). Nonparametric statistics is based on either being distri ...
to fully capture the information in the data. Less tersely, suppose X_n, n = 1, 2, 3, \dots are
independent identically distributed In probability theory and statistics, a collection of random variables is independent and identically distributed if each random variable has the same probability distribution as the others and all are mutually independent. This property is us ...
real random variables whose distribution is known to be in some family of probability distributions, parametrized by \theta, satisfying certain technical regularity conditions, then that family is an ''exponential'' family if and only if there is a \R^m-valued sufficient statistic T(X_1, \dots, X_n) whose number of scalar components m does not increase as the sample size ''n'' increases. This theorem shows that the existence of a finite-dimensional, real-vector-valued sufficient statistics sharply restricts the possible forms of a family of distributions on the real line. When the parameters or the random variables are no longer real-valued, the situation is more complex.


Other types of sufficiency


Bayesian sufficiency

An alternative formulation of the condition that a statistic be sufficient, set in a Bayesian context, involves the posterior distributions obtained by using the full data-set and by using only a statistic. Thus the requirement is that, for almost every ''x'', :\Pr(\theta\mid X=x) = \Pr(\theta\mid T(X)=t(x)). More generally, without assuming a parametric model, we can say that the statistics ''T'' is ''predictive sufficient'' if :\Pr(X'=x'\mid X=x) = \Pr(X'=x'\mid T(X)=t(x)). It turns out that this "Bayesian sufficiency" is a consequence of the formulation above, however they are not directly equivalent in the infinite-dimensional case. A range of theoretical results for sufficiency in a Bayesian context is available.


Linear sufficiency

A concept called "linear sufficiency" can be formulated in a Bayesian context, and more generally. First define the best linear predictor of a vector ''Y'' based on ''X'' as \hat E \mid X/math>. Then a linear statistic ''T''(''x'') is linear sufficient if :\hat E theta\mid X \hat E theta\mid T(X).


See also

* Completeness of a statistic * Basu's theorem on independence of complete sufficient and ancillary statistics * Lehmann–Scheffé theorem: a complete sufficient estimator is the best estimator of its expectation *
Rao–Blackwell theorem In statistics, the Rao–Blackwell theorem, sometimes referred to as the Rao–Blackwell–Kolmogorov theorem, is a result which characterizes the transformation of an arbitrarily crude estimator into an estimator that is optimal by the mean-squ ...
* Chentsov's theorem * Sufficient dimension reduction * Ancillary statistic


Notes


References

* * *Dodge, Y. (2003) ''The Oxford Dictionary of Statistical Terms'', OUP. {{DEFAULTSORT:Sufficient Statistic Statistical theory Statistical principles Articles containing proofs factorization