In estimation theory and statistics, the Cramér–Rao bound (CRB) relates to estimation of a deterministic (fixed, though unknown) parameter. The result is named in honor of Harald Cramér and C. R. Rao,[1][2][3] but has also been derived independently by Maurice Fréchet,[4] Georges Darmois,[5] and by Alexander Aitken and Harold Silverstone.[6][7] It is also known as Fréchet-Cramér–Rao or Fréchet-Darmois-Cramér-Rao lower bound. It states that the precision of any unbiased estimator is at most the Fisher information; or (equivalently) the reciprocal of the Fisher information is a lower bound on its variance.

Illustration of the Cramer-Rao bound: there is no unbiased estimator which is able to estimate the (2-dimensional) parameter with less variance than the Cramer-Rao bound, illustrated as standard deviation ellipse.

An unbiased estimator that achieves this bound is said to be (fully) efficient. Such a solution achieves the lowest possible mean squared error among all unbiased methods, and is, therefore, the minimum variance unbiased (MVU) estimator. However, in some cases, no unbiased technique exists which achieves the bound. This may occur either if for any unbiased estimator, there exists another with a strictly smaller variance, or if an MVU estimator exists, but its variance is strictly greater than the inverse of the Fisher information.

The Cramér–Rao bound can also be used to bound the variance of biased estimators of given bias. In some cases, a biased approach can result in both a variance and a mean squared error that are below the unbiased Cramér–Rao lower bound; see estimator bias.

Significant progress over the Cramér–Rao lower bound was proposed by A. Bhattacharyya through a series of works, called Bhattacharyya Bound.[8][9][10][11]

Statement edit

The Cramér–Rao bound is stated in this section for several increasingly general cases, beginning with the case in which the parameter is a scalar and its estimator is unbiased. All versions of the bound require certain regularity conditions, which hold for most well-behaved distributions. These conditions are listed later in this section.

Scalar unbiased case edit

Suppose   is an unknown deterministic parameter that is to be estimated from   independent observations (measurements) of  , each from a distribution according to some probability density function  . The variance of any unbiased estimator   of   is then bounded[12] by the reciprocal of the Fisher information  :

 

where the Fisher information   is defined by

 

and   is the natural logarithm of the likelihood function for a single sample   and   denotes the expected value with respect to the density   of  . If not indicated, in what follows, the expectation is taken with respect to  .

If   is twice differentiable and certain regularity conditions hold, then the Fisher information can also be defined as follows:[13]

 

The efficiency of an unbiased estimator   measures how close this estimator's variance comes to this lower bound; estimator efficiency is defined as

 

or the minimum possible variance for an unbiased estimator divided by its actual variance. The Cramér–Rao lower bound thus gives

 .

General scalar case edit

A more general form of the bound can be obtained by considering a biased estimator  , whose expectation is not   but a function of this parameter, say,  . Hence   is not generally equal to 0. In this case, the bound is given by

 

where   is the derivative of   (by  ), and   is the Fisher information defined above.

Bound on the variance of biased estimators edit

Apart from being a bound on estimators of functions of the parameter, this approach can be used to derive a bound on the variance of biased estimators with a given bias, as follows.[14] Consider an estimator   with bias  , and let  . By the result above, any unbiased estimator whose expectation is   has variance greater than or equal to  . Thus, any estimator   whose bias is given by a function   satisfies[15]

 

The unbiased version of the bound is a special case of this result, with  .

It's trivial to have a small variance − an "estimator" that is constant has a variance of zero. But from the above equation, we find that the mean squared error of a biased estimator is bounded by

 

using the standard decomposition of the MSE. Note, however, that if   this bound might be less than the unbiased Cramér–Rao bound  . For instance, in the example of estimating variance below,  .

Multivariate case edit

Extending the Cramér–Rao bound to multiple parameters, define a parameter column vector

 

with probability density function   which satisfies the two regularity conditions below.

The Fisher information matrix is a   matrix with element   defined as

 

Let   be an estimator of any vector function of parameters,  , and denote its expectation vector   by  . The Cramér–Rao bound then states that the covariance matrix of   satisfies

 ,
 

where

  • The matrix inequality   is understood to mean that the matrix   is positive semidefinite, and
  •   is the Jacobian matrix whose   element is given by  .


If   is an unbiased estimator of   (i.e.,  ), then the Cramér–Rao bound reduces to

 

If it is inconvenient to compute the inverse of the Fisher information matrix, then one can simply take the reciprocal of the corresponding diagonal element to find a (possibly loose) lower bound.[16]

 

Regularity conditions edit

The bound relies on two weak regularity conditions on the probability density function,  , and the estimator  :

  • The Fisher information is always defined; equivalently, for all   such that  ,
     
    exists, and is finite.
  • The operations of integration with respect to   and differentiation with respect to   can be interchanged in the expectation of  ; that is,
     
    whenever the right-hand side is finite.
    This condition can often be confirmed by using the fact that integration and differentiation can be swapped when either of the following cases hold:
    1. The function   has bounded support in  , and the bounds do not depend on  ;
    2. The function   has infinite support, is continuously differentiable, and the integral converges uniformly for all  .

Proof edit

Proof for the general case based on the Chapman–Robbins bound edit

Proof based on.[17]

Proof

First equation:

Let   be an infinitesimal, then for any  , plugging   in, we have  

Plugging this into multivariate Chapman–Robbins bound gives  .

Second equation:

It suffices to prove this for scalar case, with   taking values in   . Because for general   , we can take any  , then defining  , the scalar case gives

 
This holds for all  , so we can conclude
 
The scalar case states that   with  .

Let   be an infinitesimal, then for any  , taking   in the single-variate Chapman–Robbins bound gives  .

By linear algebra,   for any positive-definite matrix  , thus we obtain

 

A standalone proof for the general scalar case edit

For the general scalar case:

Assume that   is an estimator with expectation   (based on the observations  ), i.e. that  . The goal is to prove that, for all  ,

 

Let   be a random variable with probability density function  . Here   is a statistic, which is used as an estimator for  . Define   as the score:

 

where the chain rule is used in the final equality above. Then the expectation of  , written  , is zero. This is because:

 

where the integral and partial derivative have been interchanged (justified by the second regularity condition).


If we consider the covariance   of   and  , we have  , because  . Expanding this expression we have

 

again because the integration and differentiation operations commute (second condition).

The Cauchy–Schwarz inequality shows that

 

therefore

 

which proves the proposition.

Examples edit

Multivariate normal distribution edit

For the case of a d-variate normal distribution

 

the Fisher information matrix has elements[18]

 

where "tr" is the trace.

For example, let   be a sample of   independent observations with unknown mean   and known variance   .

 

Then the Fisher information is a scalar given by

 

and so the Cramér–Rao bound is

 

Normal variance with known mean edit

Suppose X is a normally distributed random variable with known mean   and unknown variance  . Consider the following statistic:

 

Then T is unbiased for  , as  . What is the variance of T?

 

(the second equality follows directly from the definition of variance). The first term is the fourth moment about the mean and has value  ; the second is the square of the variance, or  . Thus

 

Now, what is the Fisher information in the sample? Recall that the score   is defined as

 

where   is the likelihood function. Thus in this case,

 
 

where the second equality is from elementary calculus. Thus, the information in a single observation is just minus the expectation of the derivative of  , or

 

Thus the information in a sample of   independent observations is just   times this, or  

The Cramér–Rao bound states that

 

In this case, the inequality is saturated (equality is achieved), showing that the estimator is efficient.

However, we can achieve a lower mean squared error using a biased estimator. The estimator

 

obviously has a smaller variance, which is in fact

 

Its bias is

 

so its mean squared error is

 

which is less than what unbiased estimators can achieve according to the Cramér–Rao bound.

When the mean is not known, the minimum mean squared error estimate of the variance of a sample from Gaussian distribution is achieved by dividing by  , rather than   or  .

See also edit

References and notes edit

  1. ^ Cramér, Harald (1946). Mathematical Methods of Statistics. Princeton, NJ: Princeton Univ. Press. ISBN 0-691-08004-6. OCLC 185436716.
  2. ^ Rao, Calyampudi Radakrishna (1945). "Information and the accuracy attainable in the estimation of statistical parameters". Bulletin of the Calcutta Mathematical Society. 37. Calcutta Mathematical Society: 81–89. MR 0015748.
  3. ^ Rao, Calyampudi Radakrishna (1994). S. Das Gupta (ed.). Selected Papers of C. R. Rao. New York: Wiley. ISBN 978-0-470-22091-7. OCLC 174244259.
  4. ^ Fréchet, Maurice (1943). "Sur l'extension de certaines évaluations statistiques au cas de petits échantillons". Rev. Inst. Int. Statist. 11 (3/4): 182–205. doi:10.2307/1401114. JSTOR 1401114.
  5. ^ Darmois, Georges (1945). "Sur les limites de la dispersion de certaines estimations". Rev. Int. Inst. Statist. 13 (1/4): 9–15. doi:10.2307/1400974. JSTOR 1400974.
  6. ^ Aitken, A. C.; Silverstone, H. (1942). "XV.—On the Estimation of Statistical Parameters". Proceedings of the Royal Society of Edinburgh Section A: Mathematics. 61 (2): 186–194. doi:10.1017/S008045410000618X. ISSN 2053-5902. S2CID 124029876.
  7. ^ Shenton, L. R. (1970). "The so-called Cramer–Rao inequality". The American Statistician. 24 (2): 36. JSTOR 2681931.
  8. ^ Dodge, Yadolah (2003). The Oxford Dictionary of Statistical Terms. Oxford University Press. ISBN 978-0-19-920613-1.
  9. ^ Bhattacharyya, A. (1946). "On Some Analogues of the Amount of Information and Their Use in Statistical Estimation". Sankhyā. 8 (1): 1–14. JSTOR 25047921. MR 0020242.
  10. ^ Bhattacharyya, A. (1947). "On Some Analogues of the Amount of Information and Their Use in Statistical Estimation (Contd.)". Sankhyā. 8 (3): 201–218. JSTOR 25047948. MR 0023503.
  11. ^ Bhattacharyya, A. (1948). "On Some Analogues of the Amount of Information and Their Use in Statistical Estimation (Concluded)". Sankhyā. 8 (4): 315–328. JSTOR 25047897. MR 0026302.
  12. ^ Nielsen, Frank (2013). "Cramér-Rao Lower Bound and Information Geometry". Connected at Infinity II. Texts and Readings in Mathematics. Vol. 67. Hindustan Book Agency, Gurgaon. p. 18-37. arXiv:1301.3578. doi:10.1007/978-93-86279-56-9_2. ISBN 978-93-80250-51-9. S2CID 16759683.
  13. ^ Suba Rao. "Lectures on statistical inference" (PDF). Archived from the original (PDF) on 2020-09-26. Retrieved 2020-05-24.
  14. ^ "Cramér Rao Lower Bound - Navipedia". gssc.esa.int.
  15. ^ "Cramér-Rao Bound".
  16. ^ For the Bayesian case, see eqn. (11) of Bobrovsky; Mayer-Wolf; Zakai (1987). "Some classes of global Cramer–Rao bounds". Ann. Stat. 15 (4): 1421–38. doi:10.1214/aos/1176350602.
  17. ^ Polyanskiy, Yury (2017). "Lecture notes on information theory, chapter 29, ECE563 (UIUC)" (PDF). Lecture notes on information theory. Archived (PDF) from the original on 2022-05-24. Retrieved 2022-05-24.
  18. ^ Kay, S. M. (1993). Fundamentals of Statistical Signal Processing: Estimation Theory. Prentice Hall. p. 47. ISBN 0-13-042268-1.

Further reading edit

  • Amemiya, Takeshi (1985). Advanced Econometrics. Cambridge: Harvard University Press. pp. 14–17. ISBN 0-674-00560-0.
  • Bos, Adriaan van den (2007). Parameter Estimation for Scientists and Engineers. Hoboken: John Wiley & Sons. pp. 45–98. ISBN 978-0-470-14781-8.
  • Kay, Steven M. (1993). Fundamentals of Statistical Signal Processing, Volume I: Estimation Theory. Prentice Hall. ISBN 0-13-345711-7.. Chapter 3.
  • Shao, Jun (1998). Mathematical Statistics. New York: Springer. ISBN 0-387-98674-X.. Section 3.1.3.
  • Posterior uncertainty, asymptotic law and Cramér-Rao bound, Structural Control and Health Monitoring 25(1851):e2113 DOI: 10.1002/stc.2113

External links edit

  • FandPLimitTool a GUI-based software to calculate the Fisher information and Cramér-Rao lower bound with application to single-molecule microscopy.