Home  

Random  

Nearby  



Log in  



Settings  



Donate  



About Wikipedia  

Disclaimers  



Wikipedia





Normal distribution





Article  

Talk  



Language  

Watch  

Edit  


(Redirected from Gaussian distribution)
 


Inprobability theory and statistics, a normal distributionorGaussian distribution is a type of continuous probability distribution for a real-valued random variable. The general form of its probability density functionis The parameter is the meanorexpectation of the distribution (and also its median and mode), while the parameter is the variance. The standard deviation of the distribution is . A random variable with a Gaussian distribution is said to be normally distributed, and is called a normal deviate.

Normal distribution
Probability density function
The red curve is the standard normal distribution.
Cumulative distribution function
Notation
Parameters = mean (location)
= variance (squared scale)
Support
PDF
CDF
Quantile
Mean
Median
Mode
Variance
MAD
Skewness
Excess kurtosis
Entropy
MGF
CF
Fisher information

Kullback–Leibler divergence
Expected shortfall [1]

Normal distributions are important in statistics and are often used in the natural and social sciences to represent real-valued random variables whose distributions are not known.[2][3] Their importance is partly due to the central limit theorem. It states that, under some conditions, the average of many samples (observations) of a random variable with finite mean and variance is itself a random variable—whose distribution converges to a normal distribution as the number of samples increases. Therefore, physical quantities that are expected to be the sum of many independent processes, such as measurement errors, often have distributions that are nearly normal.[4]

Moreover, Gaussian distributions have some unique properties that are valuable in analytic studies. For instance, any linear combination of a fixed collection of independent normal deviates is a normal deviate. Many results and methods, such as propagation of uncertainty and least squares[5] parameter fitting, can be derived analytically in explicit form when the relevant variables are normally distributed.

A normal distribution is sometimes informally called a bell curve.[6] However, many other distributions are bell-shaped (such as the Cauchy, Student's t, and logistic distributions). For other names, see Naming.

The univariate probability distribution is generalized for vectors in the multivariate normal distribution and for matrices in the matrix normal distribution.

Definitions

edit

Standard normal distribution

edit

The simplest case of a normal distribution is known as the standard normal distributionorunit normal distribution. This is a special case when   and  , and it is described by this probability density function (or density):   The variable   has a mean of 0 and a variance and standard deviation of 1. The density   has its peak  at  and inflection pointsat  and  .

Although the density above is most commonly known as the standard normal, a few authors have used that term to describe other versions of the normal distribution. Carl Friedrich Gauss, for example, once defined the standard normal as   which has a variance of 1/2, and Stephen Stigler[7] once defined the standard normal as   which has a simple functional form and a variance of  

General normal distribution

edit

Every normal distribution is a version of the standard normal distribution, whose domain has been stretched by a factor   (the standard deviation) and then translated by   (the mean value):

 

The probability density must be scaled by   so that the integral is still 1.

If  is a standard normal deviate, then   will have a normal distribution with expected value   and standard deviation  . This is equivalent to saying that the standard normal distribution   can be scaled/stretched by a factor of   and shifted by   to yield a different normal distribution, called  . Conversely, if   is a normal deviate with parameters   and  , then this   distribution can be re-scaled and shifted via the formula   to convert it to the standard normal distribution. This variate is also called the standardized form of  .

Notation

edit

The probability density of the standard Gaussian distribution (standard normal distribution, with zero mean and unit variance) is often denoted with the Greek letter   (phi).[8] The alternative form of the Greek letter phi,  , is also used quite often.

The normal distribution is often referred to as  or .[9] Thus when a random variable   is normally distributed with mean   and standard deviation  , one may write

 

Alternative parameterizations

edit

Some authors advocate using the precision   as the parameter defining the width of the distribution, instead of the standard deviation   or the variance  . The precision is normally defined as the reciprocal of the variance,  .[10] The formula for the distribution then becomes

 

This choice is claimed to have advantages in numerical computations when   is very close to zero, and simplifies formulas in some contexts, such as in the Bayesian inference of variables with multivariate normal distribution.

Alternatively, the reciprocal of the standard deviation   might be defined as the precision, in which case the expression of the normal distribution becomes

 

According to Stigler, this formulation is advantageous because of a much simpler and easier-to-remember formula, and simple approximate formulas for the quantiles of the distribution.

Normal distributions form an exponential family with natural parameters   and  , and natural statistics x and x2. The dual expectation parameters for normal distribution are η1 = μ and η2 = μ2 + σ2.

Cumulative distribution function

edit

The cumulative distribution function (CDF) of the standard normal distribution, usually denoted with the capital Greek letter   (phi), is the integral

 

Error Function

edit

The related error function   gives the probability of a random variable, with normal distribution of mean 0 and variance 1/2 falling in the range  . That is:

 

These integrals cannot be expressed in terms of elementary functions, and are often said to be special functions. However, many numerical approximations are known; see below for more.

The two functions are closely related, namely

 

For a generic normal distribution with density  , mean   and variance  , the cumulative distribution function is

 

The complement of the standard normal cumulative distribution function,  , is often called the Q-function, especially in engineering texts.[11][12] It gives the probability that the value of a standard normal random variable   will exceed  :  . Other definitions of the  -function, all of which are simple transformations of  , are also used occasionally.[13]

The graph of the standard normal cumulative distribution function   has 2-fold rotational symmetry around the point (0,1/2); that is,  . Its antiderivative (indefinite integral) can be expressed as follows:  

The cumulative distribution function of the standard normal distribution can be expanded by Integration by parts into a series:

 

where   denotes the double factorial.

Anasymptotic expansion of the cumulative distribution function for large x can also be derived using integration by parts. For more, see Error function#Asymptotic expansion.[14]

A quick approximation to the standard normal distribution's cumulative distribution function can be found by using a Taylor series approximation:

 

Recursive computation with Taylor series expansion

edit

The recursive nature of the  family of derivatives may be used to easily construct a rapidly converging Taylor series expansion using recursive entries about any point of known value of the distribution, :

 

where:

 

Using the Taylor series and Newton's method for the inverse function

edit

An application for the above Taylor series expansion is to use Newton's method to reverse the computation. That is, if we have a value for the cumulative distribution function,  , but do not know the x needed to obtain the  , we can use Newton's method to find x, and use the Taylor series expansion above to minimize the number of computations. Newton's method is ideal to solve this problem because the first derivative of  , which is an integral of the normal standard distribution, is the normal standard distribution, and is readily available to use in the Newton's method solution.

To solve, select a known approximate solution,  , to the desired  .   may be a value from a distribution table, or an intelligent estimate followed by a computation of   using any desired means to compute. Use this value of   and the Taylor series expansion above to minimize computations.

Repeat the following process until the difference between the computed   and the desired  , which we will call  , is below a chosen acceptably small error, such as 10−5, 10−15, etc.:

 

where

  is the   from a Taylor series solution using   and  

 

When the repeated computations converge to an error below the chosen acceptably small value, x will be the value needed to obtain a   of the desired value,  .

Standard deviation and coverage

edit
 
For the normal distribution, the values less than one standard deviation away from the mean account for 68.27% of the set; while two standard deviations from the mean account for 95.45%; and three standard deviations account for 99.73%.

About 68% of values drawn from a normal distribution are within one standard deviation σ away from the mean; about 95% of the values lie within two standard deviations; and about 99.7% are within three standard deviations.[6] This fact is known as the 68–95–99.7 (empirical) rule, or the 3-sigma rule.

More precisely, the probability that a normal deviate lies in the range between   and   is given by   To 12 significant digits, the values for   are:[citation needed]

        OEIS
1 0.682689492137 0.317310507863
3 .15148718753
OEISA178647
2 0.954499736104 0.045500263896
21 .9778945080
OEISA110894
3 0.997300203937 0.002699796063
370 .398347345
OEISA270712
4 0.999936657516 0.000063342484
15787 .1927673
5 0.999999426697 0.000000573303
1744277 .89362
6 0.999999998027 0.000000001973
506797345 .897

For large  , one can use the approximation  .

Quantile function

edit

The quantile function of a distribution is the inverse of the cumulative distribution function. The quantile function of the standard normal distribution is called the probit function, and can be expressed in terms of the inverse error function:   For a normal random variable with mean   and variance  , the quantile function is   The quantile   of the standard normal distribution is commonly denoted as  . These values are used in hypothesis testing, construction of confidence intervals and Q–Q plots. A normal random variable   will exceed   with probability  , and will lie outside the interval   with probability  . In particular, the quantile  is1.96; therefore a normal random variable will lie outside the interval   in only 5% of cases.

The following table gives the quantile   such that   will lie in the range   with a specified probability  . These values are useful to determine tolerance interval for sample averages and other statistical estimators with normal (orasymptotically normal) distributions.[15] The following table shows  , not   as defined above.

         
0.80 1.281551565545 0.999 3.290526731492
0.90 1.644853626951 0.9999 3.890591886413
0.95 1.959963984540 0.99999 4.417173413469
0.98 2.326347874041 0.999999 4.891638475699
0.99 2.575829303549 0.9999999 5.326723886384
0.995 2.807033768344 0.99999999 5.730728868236
0.998 3.090232306168 0.999999999 6.109410204869

For small  , the quantile function has the useful asymptotic expansion  [citation needed]

Properties

edit

The normal distribution is the only distribution whose cumulants beyond the first two (i.e., other than the mean and variance) are zero. It is also the continuous distribution with the maximum entropy for a specified mean and variance.[16][17] Geary has shown, assuming that the mean and variance are finite, that the normal distribution is the only distribution where the mean and variance calculated from a set of independent draws are independent of each other.[18][19]

The normal distribution is a subclass of the elliptical distributions. The normal distribution is symmetric about its mean, and is non-zero over the entire real line. As such it may not be a suitable model for variables that are inherently positive or strongly skewed, such as the weight of a person or the price of a share. Such variables may be better described by other distributions, such as the log-normal distribution or the Pareto distribution.

The value of the normal distribution is practically zero when the value   lies more than a few standard deviations away from the mean (e.g., a spread of three standard deviations covers all but 0.27% of the total distribution). Therefore, it may not be an appropriate model when one expects a significant fraction of outliers—values that lie many standard deviations away from the mean—and least squares and other statistical inference methods that are optimal for normally distributed variables often become highly unreliable when applied to such data. In those cases, a more heavy-tailed distribution should be assumed and the appropriate robust statistical inference methods applied.

The Gaussian distribution belongs to the family of stable distributions which are the attractors of sums of independent, identically distributed distributions whether or not the mean or variance is finite. Except for the Gaussian which is a limiting case, all stable distributions have heavy tails and infinite variance. It is one of the few distributions that are stable and that have probability density functions that can be expressed analytically, the others being the Cauchy distribution and the Lévy distribution.

Symmetries and derivatives

edit

The normal distribution with density   (mean   and variance  ) has the following properties:

Furthermore, the density   of the standard normal distribution (i.e.   and  ) also has the following properties:

Moments

edit

The plain and absolute moments of a variable   are the expected values of   and  , respectively. If the expected value  of  is zero, these parameters are called central moments; otherwise, these parameters are called non-central moments. Usually we are interested only in moments with integer order  .

If  has a normal distribution, the non-central moments exist and are finite for any   whose real part is greater than −1. For any non-negative integer  , the plain central moments are:[23]   Here   denotes the double factorial, that is, the product of all numbers from   to 1 that have the same parity as  

The central absolute moments coincide with plain moments for all even orders, but are nonzero for odd orders. For any non-negative integer  

  The last formula is valid also for any non-integer   When the mean   the plain and absolute moments can be expressed in terms of confluent hypergeometric functions   and  [24]

 

These expressions remain valid even if   is not an integer. See also generalized Hermite polynomials.

Order Non-central moment Central moment
1    
2    
3    
4    
5    
6    
7    
8    

The expectation of   conditioned on the event that   lies in an interval   is given by   where   and   respectively are the density and the cumulative distribution function of  . For   this is known as the inverse Mills ratio. Note that above, density  of  is used instead of standard normal density as in inverse Mills ratio, so here we have   instead of  .

Fourier transform and characteristic function

edit

The Fourier transform of a normal density   with mean   and variance  is[25]

 

where   is the imaginary unit. If the mean  , the first factor is 1, and the Fourier transform is, apart from a constant factor, a normal density on the frequency domain, with mean 0 and variance  . In particular, the standard normal distribution   is an eigenfunction of the Fourier transform.

In probability theory, the Fourier transform of the probability distribution of a real-valued random variable   is closely connected to the characteristic function   of that variable, which is defined as the expected valueof , as a function of the real variable   (the frequency parameter of the Fourier transform). This definition can be analytically extended to a complex-value variable  .[26] The relation between both is:  

Moment- and cumulant-generating functions

edit

The moment generating function of a real random variable   is the expected value of  , as a function of the real parameter  . For a normal distribution with density  , mean   and variance  , the moment generating function exists and is equal to

 

The cumulant generating function is the logarithm of the moment generating function, namely

 

Since this is a quadratic polynomial in  , only the first two cumulants are nonzero, namely the mean   and the variance  .

Some authors prefer to instead work with E[eitX] = eiμtσ2t2/2 and ln E[eitX] = iμt1/2σ2t2.

Stein operator and class

edit

Within Stein's method the Stein operator and class of a random variable   are   and   the class of all absolutely continuous functions  .

Zero-variance limit

edit

In the limit when   tends to zero, the probability density   eventually tends to zero at any  , but grows without limit if  , while its integral remains equal to 1. Therefore, the normal distribution cannot be defined as an ordinary function when  .

However, one can define the normal distribution with zero variance as a generalized function; specifically, as a Dirac delta function   translated by the mean  , that is   Its cumulative distribution function is then the Heaviside step function translated by the mean  , namely  

Maximum entropy

edit

Of all probability distributions over the reals with a specified finite mean   and finite variance  , the normal distribution   is the one with maximum entropy.[27] To see this, let   be a continuous random variable with probability density  . The entropy of   is defined as[28][29][30]  

where   is understood to be zero whenever  . This functional can be maximized, subject to the constraints that the distribution is properly normalized and has a specified mean and variance, by using variational calculus. A function with three Lagrange multipliers is defined:

 

At maximum entropy, a small variation   about   will produce a variation   about   which is equal to 0:

 

Since this must hold for any small  , the factor multiplying   must be zero, and solving for   yields:

 

The Lagrange constraints that   is properly normalized and has the specified mean and variance are satisfied if and only if  ,  , and   are chosen so that   The entropy of a normal distribution   is equal to   which is independent of the mean  .

Other properties

edit

  1. If the characteristic function   of some random variable   is of the form   in a neighborhood of zero, where   is a polynomial, then the Marcinkiewicz theorem (named after Józef Marcinkiewicz) asserts that   can be at most a quadratic polynomial, and therefore   is a normal random variable.[31] The consequence of this result is that the normal distribution is the only distribution with a finite number (two) of non-zero cumulants.
  2. If  and   are jointly normal and uncorrelated, then they are independent. The requirement that   and   should be jointly normal is essential; without it the property does not hold.[32][33][proof] For non-normal random variables uncorrelatedness does not imply independence.
  3. The Kullback–Leibler divergence of one normal distribution   from another   is given by:[34]   The Hellinger distance between the same distributions is equal to  
  4. The Fisher information matrix for a normal distribution w.r.t.   and   is diagonal and takes the form  
  5. The conjugate prior of the mean of a normal distribution is another normal distribution.[35] Specifically, if   are iid   and the prior is  , then the posterior distribution for the estimator of   will be  
  6. The family of normal distributions not only forms an exponential family (EF), but in fact forms a natural exponential family (NEF) with quadratic variance function (NEF-QVF). Many properties of normal distributions generalize to properties of NEF-QVF distributions, NEF distributions, or EF distributions generally. NEF-QVF distributions comprises 6 families, including Poisson, Gamma, binomial, and negative binomial distributions, while many of the common families studied in probability and statistics are NEF or EF.
  7. Ininformation geometry, the family of normal distributions forms a statistical manifold with constant curvature  . The same family is flat with respect to the (±1)-connections   and  .[36]
  8. If  are distributed according to  , then  . Note that there is no assumption of independence.[37]

edit

Central limit theorem

edit
 
As the number of discrete events increases, the function begins to resemble a normal distribution.
 
Comparison of probability density functions,   for the sum of   fair 6-sided dice to show their convergence to a normal distribution with increasing  , in accordance to the central limit theorem. In the bottom-right graph, smoothed profiles of the previous graphs are rescaled, superimposed and compared with a normal distribution (black curve).

The central limit theorem states that under certain (fairly common) conditions, the sum of many random variables will have an approximately normal distribution. More specifically, where   are independent and identically distributed random variables with the same arbitrary distribution, zero mean, and variance   and   is their mean scaled by     Then, as   increases, the probability distribution of   will tend to the normal distribution with zero mean and variance  .

The theorem can be extended to variables   that are not independent and/or not identically distributed if certain constraints are placed on the degree of dependence and the moments of the distributions.

Many test statistics, scores, and estimators encountered in practice contain sums of certain random variables in them, and even more estimators can be represented as sums of random variables through the use of influence functions. The central limit theorem implies that those statistical parameters will have asymptotically normal distributions.

The central limit theorem also implies that certain distributions can be approximated by the normal distribution, for example:

Whether these approximations are sufficiently accurate depends on the purpose for which they are needed, and the rate of convergence to the normal distribution. It is typically the case that such approximations are less accurate in the tails of the distribution.

A general upper bound for the approximation error in the central limit theorem is given by the Berry–Esseen theorem, improvements of the approximation are given by the Edgeworth expansions.

This theorem can also be used to justify modeling the sum of many uniform noise sources as Gaussian noise. See AWGN.

Operations and functions of normal variables

edit
 
a: Probability density of a function   of a normal variable   with   and  . b: Probability density of a function   of two normal variables   and  , where  ,  ,  ,  , and  . c: Heat map of the joint probability density of two functions of two correlated normal variables   and  , where  ,  ,  ,  , and  . d: Probability density of a function   of 4 iid standard normal variables. These are computed by the numerical method of ray-tracing.[39]

The probability density, cumulative distribution, and inverse cumulative distribution of any function of one or more independent or correlated normal variables can be computed with the numerical method of ray-tracing[39] (Matlab code). In the following sections we look at some special cases.

Operations on a single normal variable

edit

If  is distributed normally with mean   and variance  , then

Operations on two independent normal variables
edit
Operations on two independent standard normal variables
edit

If  and   are two independent standard normal random variables with mean 0 and variance 1, then

Operations on multiple independent normal variables

edit

Operations on multiple correlated normal variables

edit

Operations on the density function

edit

The split normal distribution is most directly defined in terms of joining scaled sections of the density functions of different normal distributions and rescaling the density to integrate to one. The truncated normal distribution results from rescaling a section of a single density function.

Infinite divisibility and Cramér's theorem

edit

For any positive integer  , any normal distribution with mean   and variance   is the distribution of the sum of   independent normal deviates, each with mean   and variance  . This property is called infinite divisibility.[45]

Conversely, if   and   are independent random variables and their sum   has a normal distribution, then both   and   must be normal deviates.[46]

This result is known as Cramér's decomposition theorem, and is equivalent to saying that the convolution of two distributions is normal if and only if both are normal. Cramér's theorem implies that a linear combination of independent non-Gaussian variables will never have an exactly normal distribution, although it may approach it arbitrarily closely.[31]

Bernstein's theorem

edit

Bernstein's theorem states that if   and   are independent and   and   are also independent, then both X and Y must necessarily have normal distributions.[47][48]

More generally, if   are independent random variables, then two distinct linear combinations   and  will be independent if and only if all   are normal and  , where   denotes the variance of  .[47]

Extensions

edit

The notion of normal distribution, being one of the most important distributions in probability theory, has been extended far beyond the standard framework of the univariate (that is one-dimensional) case (Case 1). All these extensions are also called normalorGaussian laws, so a certain ambiguity in names exists.

A random variable X has a two-piece normal distribution if it has a distribution

   

where μ is the mean and σ12 and σ22 are the variances of the distribution to the left and right of the mean respectively.

The mean, variance and third central moment of this distribution have been determined[49]

     

where E(X), V(X) and T(X) are the mean, variance, and third central moment respectively.

One of the main practical uses of the Gaussian law is to model the empirical distributions of many different random variables encountered in practice. In such case a possible extension would be a richer family of distributions, having more than two parameters and therefore being able to fit the empirical distribution more accurately. The examples of such extensions are:

Statistical inference

edit

Estimation of parameters

edit

It is often the case that we do not know the parameters of the normal distribution, but instead want to estimate them. That is, having a sample   from a normal   population we would like to learn the approximate values of parameters   and  . The standard approach to this problem is the maximum likelihood method, which requires maximization of the log-likelihood function:   Taking derivatives with respect to   and   and solving the resulting system of first order conditions yields the maximum likelihood estimates:  

Then   is as follows:

 

Sample mean

edit

Estimator   is called the sample mean, since it is the arithmetic mean of all observations. The statistic  iscomplete and sufficient for  , and therefore by the Lehmann–Scheffé theorem,   is the uniformly minimum variance unbiased (UMVU) estimator.[50] In finite samples it is distributed normally:   The variance of this estimator is equal to the μμ-element of the inverse Fisher information matrix  . This implies that the estimator is finite-sample efficient. Of practical importance is the fact that the standard errorof  is proportional to  , that is, if one wishes to decrease the standard error by a factor of 10, one must increase the number of points in the sample by a factor of 100. This fact is widely used in determining sample sizes for opinion polls and the number of trials in Monte Carlo simulations.

From the standpoint of the asymptotic theory,  isconsistent, that is, it converges in probabilityto as . The estimator is also asymptotically normal, which is a simple corollary of the fact that it is normal in finite samples:  

Sample variance

edit

The estimator   is called the sample variance, since it is the variance of the sample ( ). In practice, another estimator is often used instead of the  . This other estimator is denoted  , and is also called the sample variance, which represents a certain ambiguity in terminology; its square root   is called the sample standard deviation. The estimator   differs from   by having (n − 1) instead of n in the denominator (the so-called Bessel's correction):   The difference between   and   becomes negligibly small for large n's. In finite samples however, the motivation behind the use of   is that it is an unbiased estimator of the underlying parameter  , whereas   is biased. Also, by the Lehmann–Scheffé theorem the estimator   is uniformly minimum variance unbiased (UMVU),[50] which makes it the "best" estimator among all unbiased ones. However it can be shown that the biased estimator   is better than the   in terms of the mean squared error (MSE) criterion. In finite samples both   and   have scaled chi-squared distribution with (n − 1) degrees of freedom:   The first of these expressions shows that the variance of   is equal to  , which is slightly greater than the σσ-element of the inverse Fisher information matrix  . Thus,   is not an efficient estimator for  , and moreover, since   is UMVU, we can conclude that the finite-sample efficient estimator for   does not exist.

Applying the asymptotic theory, both estimators   and   are consistent, that is they converge in probability to   as the sample size  . The two estimators are also both asymptotically normal:   In particular, both estimators are asymptotically efficient for  .

Confidence intervals

edit

ByCochran's theorem, for normal distributions the sample mean   and the sample variance s2 are independent, which means there can be no gain in considering their joint distribution. There is also a converse theorem: if in a sample the sample mean and sample variance are independent, then the sample must have come from the normal distribution. The independence between   and s can be employed to construct the so-called t-statistic:   This quantity t has the Student's t-distribution with (n − 1) degrees of freedom, and it is an ancillary statistic (independent of the value of the parameters). Inverting the distribution of this t-statistics will allow us to construct the confidence interval for μ;[51] similarly, inverting the χ2 distribution of the statistic s2 will give us the confidence interval for σ2:[52]     where tk,p and χ 2
k,p
 
are the pthquantiles of the t- and χ2-distributions respectively. These confidence intervals are of the confidence level 1 − α, meaning that the true values μ and σ2 fall outside of these intervals with probability (orsignificance level) α. In practice people usually take α = 5%, resulting in the 95% confidence intervals.

Approximate formulas can be derived from the asymptotic distributions of   and s2:     The approximate formulas become valid for large values of n, and are more convenient for the manual calculation since the standard normal quantiles zα/2 do not depend on n. In particular, the most popular value of α = 5%, results in |z0.025| = 1.96.

Normality tests

edit

Normality tests assess the likelihood that the given data set {x1, ..., xn} comes from a normal distribution. Typically the null hypothesis H0 is that the observations are distributed normally with unspecified mean μ and variance σ2, versus the alternative Ha that the distribution is arbitrary. Many tests (over 40) have been devised for this problem. The more prominent of them are outlined below:

Diagnostic plots are more intuitively appealing but subjective at the same time, as they rely on informal human judgement to accept or reject the null hypothesis.

Goodness-of-fit tests:

Moment-based tests:

Tests based on the empirical distribution function:

Bayesian analysis of the normal distribution

edit

Bayesian analysis of normally distributed data is complicated by the many different possibilities that may be considered:

The formulas for the non-linear-regression cases are summarized in the conjugate prior article.

Sum of two quadratics

edit
Scalar form
edit

The following auxiliary formula is useful for simplifying the posterior update equations, which otherwise become fairly tedious.

 

This equation rewrites the sum of two quadratics in x by expanding the squares, grouping the terms in x, and completing the square. Note the following about the complex constant factors attached to some of the terms:

  1. The factor   has the form of a weighted averageofy and z.
  2.   This shows that this factor can be thought of as resulting from a situation where the reciprocals of quantities a and b add directly, so to combine a and b themselves, it is necessary to reciprocate, add, and reciprocate the result again to get back into the original units. This is exactly the sort of operation performed by the harmonic mean, so it is not surprising that   is one-half the harmonic meanofa and b.
Vector form
edit

A similar formula can be written for the sum of two vector quadratics: If x, y, z are vectors of length k, and A and B are symmetric, invertible matrices of size  , then

 

where

 

The form xA x is called a quadratic form and is a scalar:   In other words, it sums up all possible combinations of products of pairs of elements from x, with a separate coefficient for each. In addition, since  , only the sum   matters for any off-diagonal elements of A, and there is no loss of generality in assuming that Aissymmetric. Furthermore, if A is symmetric, then the form  

Sum of differences from the mean

edit

Another useful formula is as follows:   where  

With known variance

edit

For a set of i.i.d. normally distributed data points X of size n where each individual point x follows   with known variance σ2, the conjugate prior distribution is also normally distributed.

This can be shown more easily by rewriting the variance as the precision, i.e. using τ = 1/σ2. Then if   and   we proceed as follows.

First, the likelihood function is (using the formula above for the sum of differences from the mean):

 

Then, we proceed as follows:

 

In the above derivation, we used the formula above for the sum of two quadratics and eliminated all constant factors not involving μ. The result is the kernel of a normal distribution, with mean   and precision  , i.e.

 

This can be written as a set of Bayesian update equations for the posterior parameters in terms of the prior parameters:

 

That is, to combine n data points with total precision of (or equivalently, total variance of n/σ2) and mean of values  , derive a new total precision simply by adding the total precision of the data to the prior total precision, and form a new mean through a precision-weighted average, i.e. a weighted average of the data mean and the prior mean, each weighted by the associated total precision. This makes logical sense if the precision is thought of as indicating the certainty of the observations: In the distribution of the posterior mean, each of the input components is weighted by its certainty, and the certainty of this distribution is the sum of the individual certainties. (For the intuition of this, compare the expression "the whole is (or is not) greater than the sum of its parts". In addition, consider that the knowledge of the posterior comes from a combination of the knowledge of the prior and likelihood, so it makes sense that we are more certain of it than of either of its components.)

The above formula reveals why it is more convenient to do Bayesian analysisofconjugate priors for the normal distribution in terms of the precision. The posterior precision is simply the sum of the prior and likelihood precisions, and the posterior mean is computed through a precision-weighted average, as described above. The same formulas can be written in terms of variance by reciprocating all the precisions, yielding the more ugly formulas

 

With known mean

edit

For a set of i.i.d. normally distributed data points X of size n where each individual point x follows   with known mean μ, the conjugate prior of the variance has an inverse gamma distribution or a scaled inverse chi-squared distribution. The two are equivalent except for having different parameterizations. Although the inverse gamma is more commonly used, we use the scaled inverse chi-squared for the sake of convenience. The prior for σ2 is as follows:

 

The likelihood function from above, written in terms of the variance, is:

 

where

 

Then:

 

The above is also a scaled inverse chi-squared distribution where

 

or equivalently

 

Reparameterizing in terms of an inverse gamma distribution, the result is:

 

With unknown mean and unknown variance

edit

For a set of i.i.d. normally distributed data points X of size n where each individual point x follows   with unknown mean μ and unknown variance σ2, a combined (multivariate) conjugate prior is placed over the mean and variance, consisting of a normal-inverse-gamma distribution. Logically, this originates as follows:

  1. From the analysis of the case with unknown mean but known variance, we see that the update equations involve sufficient statistics computed from the data consisting of the mean of the data points and the total variance of the data points, computed in turn from the known variance divided by the number of data points.
  2. From the analysis of the case with unknown variance but known mean, we see that the update equations involve sufficient statistics over the data consisting of the number of data points and sum of squared deviations.
  3. Keep in mind that the posterior update values serve as the prior distribution when further data is handled. Thus, we should logically think of our priors in terms of the sufficient statistics just described, with the same semantics kept in mind as much as possible.
  4. To handle the case where both mean and variance are unknown, we could place independent priors over the mean and variance, with fixed estimates of the average mean, total variance, number of data points used to compute the variance prior, and sum of squared deviations. Note however that in reality, the total variance of the mean depends on the unknown variance, and the sum of squared deviations that goes into the variance prior (appears to) depend on the unknown mean. In practice, the latter dependence is relatively unimportant: Shifting the actual mean shifts the generated points by an equal amount, and on average the squared deviations will remain the same. This is not the case, however, with the total variance of the mean: As the unknown variance increases, the total variance of the mean will increase proportionately, and we would like to capture this dependence.
  5. This suggests that we create a conditional prior of the mean on the unknown variance, with a hyperparameter specifying the mean of the pseudo-observations associated with the prior, and another parameter specifying the number of pseudo-observations. This number serves as a scaling parameter on the variance, making it possible to control the overall variance of the mean relative to the actual variance parameter. The prior for the variance also has two hyperparameters, one specifying the sum of squared deviations of the pseudo-observations associated with the prior, and another specifying once again the number of pseudo-observations. Each of the priors has a hyperparameter specifying the number of pseudo-observations, and in each case this controls the relative variance of that prior. These are given as two separate hyperparameters so that the variance (aka the confidence) of the two priors can be controlled separately.
  6. This leads immediately to the normal-inverse-gamma distribution, which is the product of the two distributions just defined, with conjugate priors used (aninverse gamma distribution over the variance, and a normal distribution over the mean, conditional on the variance) and with the same four parameters just defined.

The priors are normally defined as follows:

 

The update equations can be derived, and look as follows:

 

The respective numbers of pseudo-observations add the number of actual observations to them. The new mean hyperparameter is once again a weighted average, this time weighted by the relative numbers of observations. Finally, the update for   is similar to the case with known mean, but in this case the sum of squared deviations is taken with respect to the observed data mean rather than the true mean, and as a result a new interaction term needs to be added to take care of the additional error source stemming from the deviation between prior and data mean.

Proof

The prior distributions are  

Therefore, the joint prior is

 

The likelihood function from the section above with known variance is:

 

Writing it in terms of variance rather than precision, we get:   where  

Therefore, the posterior is (dropping the hyperparameters as conditioning factors):  

In other words, the posterior distribution has the form of a product of a normal distribution over   times an inverse gamma distribution over  , with parameters that are the same as the update equations above.

Occurrence and applications

edit

The occurrence of normal distribution in practical problems can be loosely classified into four categories:

  1. Exactly normal distributions;
  2. Approximately normal laws, for example when such approximation is justified by the central limit theorem; and
  3. Distributions modeled as normal – the normal distribution being the distribution with maximum entropy for a given mean and variance.
  4. Regression problems – the normal distribution being found after systematic effects have been modeled sufficiently well.

Exact normality

edit
 
The ground state of a quantum harmonic oscillator has the Gaussian distribution.

Certain quantities in physics are distributed normally, as was first demonstrated by James Clerk Maxwell. Examples of such quantities are:

Approximate normality

edit

Approximately normal distributions occur in many situations, as explained by the central limit theorem. When the outcome is produced by many small effects acting additively and independently, its distribution will be close to normal. The normal approximation will not be valid if the effects act multiplicatively (instead of additively), or if there is a single external influence that has a considerably larger magnitude than the rest of the effects.

Assumed normality

edit
 
Histogram of sepal widths for Iris versicolor from Fisher's Iris flower data set, with superimposed best-fitting normal distribution

I can only recognize the occurrence of the normal curve – the Laplacian curve of errors – as a very abnormal phenomenon. It is roughly approximated to in certain distributions; for this reason, and on account for its beautiful simplicity, we may, perhaps, use it as a first approximation, particularly in theoretical investigations.

There are statistical methods to empirically test that assumption; see the above Normality tests section.

 
Fitted cumulative normal distribution to October rainfalls, see distribution fitting

Methodological problems and peer review

edit

John Ioannidis argued that using normally distributed standard deviations as standards for validating research findings leave falsifiable predictions about phenomena that are not normally distributed untested. This includes, for example, phenomena that only appear when all necessary conditions are present and one cannot be a substitute for another in an addition-like way and phenomena that are not randomly distributed. Ioannidis argues that standard deviation-centered validation gives a false appearance of validity to hypotheses and theories where some but not all falsifiable predictions are normally distributed since the portion of falsifiable predictions that there is evidence against may and in some cases are in the non-normally distributed parts of the range of falsifiable predictions, as well as baselessly dismissing hypotheses for which none of the falsifiable predictions are normally distributed as if were they unfalsifiable when in fact they do make falsifiable predictions. It is argued by Ioannidis that many cases of mutually exclusive theories being accepted as validated by research journals are caused by failure of the journals to take in empirical falsifications of non-normally distributed predictions, and not because mutually exclusive theories are true, which they cannot be, although two mutually exclusive theories can both be wrong and a third one correct.[56]

Computational methods

edit

Generating values from normal distribution

edit
 
The bean machine, a device invented by Francis Galton, can be called the first generator of normal random variables. This machine consists of a vertical board with interleaved rows of pins. Small balls are dropped from the top and then bounce randomly left or right as they hit the pins. The balls are collected into bins at the bottom and settle down into a pattern resembling the Gaussian curve.

In computer simulations, especially in applications of the Monte-Carlo method, it is often desirable to generate values that are normally distributed. The algorithms listed below all generate the standard normal deviates, since a N(μ, σ2) can be generated as X = μ + σZ, where Z is standard normal. All these algorithms rely on the availability of a random number generator U capable of producing uniform random variates.

Numerical approximations for the normal cumulative distribution function and normal quantile function

edit

The standard normal cumulative distribution function is widely used in scientific and statistical computing.

The values Φ(x) may be approximated very accurately by a variety of methods, such as numerical integration, Taylor series, asymptotic series and continued fractions. Different approximations are used depending on the desired level of accuracy.

 

Shore (1982) introduced simple approximations that may be incorporated in stochastic optimization models of engineering and operations research, like reliability engineering and inventory analysis. Denoting p = Φ(z), the simplest approximation for the quantile function is:  

This approximation delivers for z a maximum absolute error of 0.026 (for 0.5 ≤ p ≤ 0.9999, corresponding to 0 ≤ z ≤ 3.719). For p < 1/2 replace pby1 − p and change sign. Another approximation, somewhat less accurate, is the single-parameter approximation:  

The latter had served to derive a simple approximation for the loss integral of the normal distribution, defined by  

This approximation is particularly accurate for the right far-tail (maximum error of 10−3 for z≥1.4). Highly accurate approximations for the cumulative distribution function, based on Response Modeling Methodology (RMM, Shore, 2011, 2012), are shown in Shore (2005).

Some more approximations can be found at: Error function#Approximation with elementary functions. In particular, small relative error on the whole domain for the cumulative distribution function   and the quantile function   as well, is achieved via an explicitly invertible formula by Sergei Winitzki in 2008.

History

edit

Development

edit

Some authors[65][66] attribute the credit for the discovery of the normal distribution to de Moivre, who in 1738[note 2] published in the second edition of his The Doctrine of Chances the study of the coefficients in the binomial expansionof(a + b)n. De Moivre proved that the middle term in this expansion has the approximate magnitude of  , and that "If mor1/2n be a Quantity infinitely great, then the Logarithm of the Ratio, which a Term distant from the middle by the Interval , has to the middle Term, is  ."[67] Although this theorem can be interpreted as the first obscure expression for the normal probability law, Stigler points out that de Moivre himself did not interpret his results as anything more than the approximate rule for the binomial coefficients, and in particular de Moivre lacked the concept of the probability density function.[68]

 
Carl Friedrich Gauss discovered the normal distribution in 1809 as a way to rationalize the method of least squares.

In 1823 Gauss published his monograph "Theoria combinationis observationum erroribus minimis obnoxiae" where among other things he introduces several important statistical concepts, such as the method of least squares, the method of maximum likelihood, and the normal distribution. Gauss used M, M, M′′, ... to denote the measurements of some unknown quantity V, and sought the most probable estimator of that quantity: the one that maximizes the probability φ(MV) · φ(MV) · φ(M′′ − V) · ... of obtaining the observed experimental results. In his notation φΔ is the probability density function of the measurement errors of magnitude Δ. Not knowing what the function φ is, Gauss requires that his method should reduce to the well-known answer: the arithmetic mean of the measured values.[note 3] Starting from these principles, Gauss demonstrates that the only law that rationalizes the choice of arithmetic mean as an estimator of the location parameter, is the normal law of errors:[69]   where h is "the measure of the precision of the observations". Using this normal law as a generic model for errors in the experiments, Gauss formulates what is now known as the non-linear weighted least squares method.[70]

 
Pierre-Simon Laplace proved the central limit theorem in 1810, consolidating the importance of the normal distribution in statistics.

Although Gauss was the first to suggest the normal distribution law, Laplace made significant contributions.[note 4] It was Laplace who first posed the problem of aggregating several observations in 1774,[71] although his own solution led to the Laplacian distribution. It was Laplace who first calculated the value of the integral et2 dt = π in 1782, providing the normalization constant for the normal distribution.[72] Finally, it was Laplace who in 1810 proved and presented to the academy the fundamental central limit theorem, which emphasized the theoretical importance of the normal distribution.[73]

It is of interest to note that in 1809 an Irish-American mathematician Robert Adrain published two insightful but flawed derivations of the normal probability law, simultaneously and independently from Gauss.[74] His works remained largely unnoticed by the scientific community, until in 1871 they were exhumed by Abbe.[75]

In the middle of the 19th century Maxwell demonstrated that the normal distribution is not just a convenient mathematical tool, but may also occur in natural phenomena:[76] The number of particles whose velocity, resolved in a certain direction, lies between x and x + dxis 

Naming

edit

Today, the concept is usually known in English as the normal distributionorGaussian distribution. Other less common names include Gauss distribution, Laplace-Gauss distribution, the law of error, the law of facility of errors, Laplace's second law, and Gaussian law.

Gauss himself apparently coined the term with reference to the "normal equations" involved in its applications, with normal having its technical meaning of orthogonal rather than usual.[77] However, by the end of the 19th century some authors[note 5] had started using the name normal distribution, where the word "normal" was used as an adjective – the term now being seen as a reflection of the fact that this distribution was seen as typical, common – and thus normal. Peirce (one of those authors) once defined "normal" thus: "...the 'normal' is not the average (or any other kind of mean) of what actually occurs, but of what would, in the long run, occur under certain circumstances."[78] Around the turn of the 20th century Pearson popularized the term normal as a designation for this distribution.[79]

Many years ago I called the Laplace–Gaussian curve the normal curve, which name, while it avoids an international question of priority, has the disadvantage of leading people to believe that all other distributions of frequency are in one sense or another 'abnormal'.

Also, it was Pearson who first wrote the distribution in terms of the standard deviation σ as in modern notation. Soon after this, in year 1915, Fisher added the location parameter to the formula for normal distribution, expressing it in the way it is written nowadays:  

The term "standard normal", which denotes the normal distribution with zero mean and unit variance came into general use around the 1950s, appearing in the popular textbooks by P. G. Hoel (1947) Introduction to Mathematical Statistics and A. M. Mood (1950) Introduction to the Theory of Statistics.[80]

See also

edit

Notes

edit
  1. ^ For example, this algorithm is given in the article Bc programming language.
  • ^ De Moivre first published his findings in 1733, in a pamphlet Approximatio ad Summam Terminorum Binomii (a + b)n in Seriem Expansi that was designated for private circulation only. But it was not until the year 1738 that he made his results publicly available. The original pamphlet was reprinted several times, see for example Walker (1985).
  • ^ "It has been customary certainly to regard as an axiom the hypothesis that if any quantity has been determined by several direct observations, made under the same circumstances and with equal care, the arithmetical mean of the observed values affords the most probable value, if not rigorously, yet very nearly at least, so that it is always most safe to adhere to it." — Gauss (1809, section 177)
  • ^ "My custom of terming the curve the Gauss–Laplacian or normal curve saves us from proportioning the merit of discovery between the two great astronomer mathematicians." quote from Pearson (1905, p. 189)
  • ^ Besides those specifically referenced here, such use is encountered in the works of Peirce, Galton (Galton (1889, chapter V)) and Lexis (Lexis (1878), Rohrbasser & Véron (2003)) c. 1875.[citation needed]
  • References

    edit

    Citations

    edit
    1. ^ Norton, Matthew; Khokhlov, Valentyn; Uryasev, Stan (2019). "Calculating CVaR and bPOE for common probability distributions with application to portfolio optimization and density estimation" (PDF). Annals of Operations Research. 299 (1–2). Springer: 1281–1315. arXiv:1811.11301. doi:10.1007/s10479-019-03373-1. S2CID 254231768. Retrieved February 27, 2023.
  • ^ Normal Distribution, Gale Encyclopedia of Psychology
  • ^ Casella & Berger (2001, p. 102)
  • ^ Lyon, A. (2014). Why are Normal Distributions Normal?, The British Journal for the Philosophy of Science.
  • ^ Jorge, Nocedal; Stephan, J. Wright (2006). Numerical Optimization (2nd ed.). Springer. p. 249. ISBN 978-0387-30303-1.
  • ^ a b "Normal Distribution". www.mathsisfun.com. Retrieved August 15, 2020.
  • ^ Stigler (1982)
  • ^ Halperin, Hartley & Hoel (1965, item 7)
  • ^ McPherson (1990, p. 110)
  • ^ Bernardo & Smith (2000, p. 121)
  • ^ Scott, Clayton; Nowak, Robert (August 7, 2003). "The Q-function". Connexions.
  • ^ Barak, Ohad (April 6, 2006). "Q Function and Error Function" (PDF). Tel Aviv University. Archived from the original (PDF) on March 25, 2009.
  • ^ Weisstein, Eric W. "Normal Distribution Function". MathWorld.
  • ^ Abramowitz, Milton; Stegun, Irene Ann, eds. (1983) [June 1964]. "Chapter 26, eqn 26.2.12". Handbook of Mathematical Functions with Formulas, Graphs, and Mathematical Tables. Applied Mathematics Series. Vol. 55 (Ninth reprint with additional corrections of tenth original printing with corrections (December 1972); first ed.). Washington D.C.; New York: United States Department of Commerce, National Bureau of Standards; Dover Publications. p. 932. ISBN 978-0-486-61272-0. LCCN 64-60036. MR 0167642. LCCN 65-12253.
  • ^ Vaart, A. W. van der (October 13, 1998). Asymptotic Statistics. Cambridge University Press. doi:10.1017/cbo9780511802256. ISBN 978-0-511-80225-6.
  • ^ Cover, Thomas M.; Thomas, Joy A. (2006). Elements of Information Theory. John Wiley and Sons. p. 254. ISBN 9780471748816.
  • ^ Park, Sung Y.; Bera, Anil K. (2009). "Maximum Entropy Autoregressive Conditional Heteroskedasticity Model" (PDF). Journal of Econometrics. 150 (2): 219–230. CiteSeerX 10.1.1.511.9750. doi:10.1016/j.jeconom.2008.12.014. Archived from the original (PDF) on March 7, 2016. Retrieved June 2, 2011.
  • ^ Geary RC(1936) The distribution of the "Student's ratio for the non-normal samples". Supplement to the Journal of the Royal Statistical Society 3 (2): 178–184
  • ^ Lukacs, Eugene (March 1942). "A Characterization of the Normal Distribution". Annals of Mathematical Statistics. 13 (1): 91–93. doi:10.1214/AOMS/1177731647. ISSN 0003-4851. JSTOR 2236166. MR 0006626. Zbl 0060.28509. Wikidata Q55897617.
  • ^ a b c Patel & Read (1996, [2.1.4])
  • ^ Fan (1991, p. 1258)
  • ^ Patel & Read (1996, [2.1.8])
  • ^ Papoulis, Athanasios. Probability, Random Variables and Stochastic Processes (4th ed.). p. 148.
  • ^ Winkelbauer, Andreas (2012). "Moments and Absolute Moments of the Normal Distribution". arXiv:1209.4340 [math.ST].
  • ^ Bryc (1995, p. 23)
  • ^ Bryc (1995, p. 24)
  • ^ Cover & Thomas (2006, p. 254)
  • ^ Williams, David (2001). Weighing the odds : a course in probability and statistics (Reprinted. ed.). Cambridge [u.a.]: Cambridge Univ. Press. pp. 197–199. ISBN 978-0-521-00618-7.
  • ^ Smith, José M. Bernardo; Adrian F. M. (2000). Bayesian theory (Reprint ed.). Chichester [u.a.]: Wiley. pp. 209, 366. ISBN 978-0-471-49464-5.{{cite book}}: CS1 maint: multiple names: authors list (link)
  • ^ O'Hagan, A. (1994) Kendall's Advanced Theory of statistics, Vol 2B, Bayesian Inference, Edward Arnold. ISBN 0-340-52922-9 (Section 5.40)
  • ^ a b Bryc (1995, p. 35)
  • ^ UIUC, Lecture 21. The Multivariate Normal Distribution, 21.6:"Individually Gaussian Versus Jointly Gaussian".
  • ^ Edward L. Melnick and Aaron Tenenbein, "Misspecifications of the Normal Distribution", The American Statistician, volume 36, number 4 November 1982, pages 372–373
  • ^ "Kullback Leibler (KL) Distance of Two Normal (Gaussian) Probability Distributions". Allisons.org. December 5, 2007. Retrieved March 3, 2017.
  • ^ Jordan, Michael I. (February 8, 2010). "Stat260: Bayesian Modeling and Inference: The Conjugate Prior for the Normal Distribution" (PDF).
  • ^ Amari & Nagaoka (2000)
  • ^ "Expectation of the maximum of gaussian random variables". Mathematics Stack Exchange. Retrieved April 7, 2024.
  • ^ "Normal Approximation to Poisson Distribution". Stat.ucla.edu. Retrieved March 3, 2017.
  • ^ a b Das, Abhranil (2021). "A method to integrate and classify normal distributions". Journal of Vision. 21 (10): 1. arXiv:2012.14331. doi:10.1167/jov.21.10.1. PMC 8419883. PMID 34468706.
  • ^ Bryc (1995, p. 27)
  • ^ Weisstein, Eric W. "Normal Product Distribution". MathWorld. wolfram.com.
  • ^ Lukacs, Eugene (1942). "A Characterization of the Normal Distribution". The Annals of Mathematical Statistics. 13 (1): 91–3. doi:10.1214/aoms/1177731647. ISSN 0003-4851. JSTOR 2236166.
  • ^ Basu, D.; Laha, R. G. (1954). "On Some Characterizations of the Normal Distribution". Sankhyā. 13 (4): 359–62. ISSN 0036-4452. JSTOR 25048183.
  • ^ Lehmann, E. L. (1997). Testing Statistical Hypotheses (2nd ed.). Springer. p. 199. ISBN 978-0-387-94919-2.
  • ^ Patel & Read (1996, [2.3.6])
  • ^ Galambos & Simonelli (2004, Theorem 3.5)
  • ^ a b Lukacs & King (1954)
  • ^ Quine, M.P. (1993). "On three characterisations of the normal distribution". Probability and Mathematical Statistics. 14 (2): 257–263.
  • ^ John, S (1982). "The three parameter two-piece normal family of distributions and its fitting". Communications in Statistics – Theory and Methods. 11 (8): 879–885. doi:10.1080/03610928208828279.
  • ^ a b Krishnamoorthy (2006, p. 127)
  • ^ Krishnamoorthy (2006, p. 130)
  • ^ Krishnamoorthy (2006, p. 133)
  • ^ Huxley (1932)
  • ^ Jaynes, Edwin T. (2003). Probability Theory: The Logic of Science. Cambridge University Press. pp. 592–593. ISBN 9780521592710.
  • ^ Oosterbaan, Roland J. (1994). "Chapter 6: Frequency and Regression Analysis of Hydrologic Data" (PDF). In Ritzema, Henk P. (ed.). Drainage Principles and Applications, Publication 16 (second revised ed.). Wageningen, The Netherlands: International Institute for Land Reclamation and Improvement (ILRI). pp. 175–224. ISBN 978-90-70754-33-4.
  • ^ Why Most Published Research Findings Are False, John P. A. Ioannidis, 2005
  • ^ Wichura, Michael J. (1988). "Algorithm AS241: The Percentage Points of the Normal Distribution". Applied Statistics. 37 (3): 477–84. doi:10.2307/2347330. JSTOR 2347330.
  • ^ Johnson, Kotz & Balakrishnan (1995, Equation (26.48))
  • ^ Kinderman & Monahan (1977)
  • ^ Leva (1992)
  • ^ Marsaglia & Tsang (2000)
  • ^ Karney (2016)
  • ^ Monahan (1985, section 2)
  • ^ Wallace (1996)
  • ^ Johnson, Kotz & Balakrishnan (1994, p. 85)
  • ^ Le Cam & Lo Yang (2000, p. 74)
  • ^ De Moivre, Abraham (1733), Corollary I – see Walker (1985, p. 77)
  • ^ Stigler (1986, p. 76)
  • ^ Gauss (1809, section 177)
  • ^ Gauss (1809, section 179)
  • ^ Laplace (1774, Problem III)
  • ^ Pearson (1905, p. 189)
  • ^ Stigler (1986, p. 144)
  • ^ Stigler (1978, p. 243)
  • ^ Stigler (1978, p. 244)
  • ^ Maxwell (1860, p. 23)
  • ^ Jaynes, Edwin J.; Probability Theory: The Logic of Science, Ch. 7.
  • ^ Peirce, Charles S. (c. 1909 MS), Collected Papers v. 6, paragraph 327.
  • ^ Kruskal & Stigler (1997).
  • ^ "Earliest Uses... (Entry Standard Normal Curve)".
  • ^ Sun, Jingchao; Kong, Maiying; Pal, Subhadip (June 22, 2021). "The Modified-Half-Normal distribution: Properties and an efficient sampling scheme". Communications in Statistics – Theory and Methods. 52 (5): 1591–1613. doi:10.1080/03610926.2021.1934700. ISSN 0361-0926. S2CID 237919587.
  • Sources

    edit
  • Aldrich, John; Miller, Jeff. "Earliest Known Uses of Some of the Words of Mathematics". In particular, the entries for "bell-shaped and bell curve", "normal (distribution)", "Gaussian", and "Error, law of error, theory of errors, etc.".
  • Amari, Shun-ichi; Nagaoka, Hiroshi (2000). Methods of Information Geometry. Oxford University Press. ISBN 978-0-8218-0531-2.
  • Bernardo, José M.; Smith, Adrian F. M. (2000). Bayesian Theory. Wiley. ISBN 978-0-471-49464-5.
  • Bryc, Wlodzimierz (1995). The Normal Distribution: Characterizations with Applications. Springer-Verlag. ISBN 978-0-387-97990-8.
  • Casella, George; Berger, Roger L. (2001). Statistical Inference (2nd ed.). Duxbury. ISBN 978-0-534-24312-8.
  • Cody, William J. (1969). "Rational Chebyshev Approximations for the Error Function". Mathematics of Computation. 23 (107): 631–638. doi:10.1090/S0025-5718-1969-0247736-4.
  • Cover, Thomas M.; Thomas, Joy A. (2006). Elements of Information Theory. John Wiley and Sons.
  • Dia, Yaya D. (2023). "Approximate Incomplete Integrals, Application to Complementary Error Function". SSRN. doi:10.2139/ssrn.4487559. S2CID 259689086.
  • de Moivre, Abraham (1738). The Doctrine of Chances. American Mathematical Society. ISBN 978-0-8218-2103-9.
  • Fan, Jianqing (1991). "On the optimal rates of convergence for nonparametric deconvolution problems". The Annals of Statistics. 19 (3): 1257–1272. doi:10.1214/aos/1176348248. JSTOR 2241949.
  • Galton, Francis (1889). Natural Inheritance (PDF). London, UK: Richard Clay and Sons.
  • Galambos, Janos; Simonelli, Italo (2004). Products of Random Variables: Applications to Problems of Physics and to Arithmetical Functions. Marcel Dekker, Inc. ISBN 978-0-8247-5402-0.
  • Gauss, Carolo Friderico (1809). Theoria motvs corporvm coelestivm in sectionibvs conicis Solem ambientivm [Theory of the Motion of the Heavenly Bodies Moving about the Sun in Conic Sections] (in Latin). Hambvrgi, Svmtibvs F. Perthes et I. H. Besser. English translation.
  • Gould, Stephen Jay (1981). The Mismeasure of Man (first ed.). W. W. Norton. ISBN 978-0-393-01489-1.
  • Halperin, Max; Hartley, Herman O.; Hoel, Paul G. (1965). "Recommended Standards for Statistical Symbols and Notation. COPSS Committee on Symbols and Notation". The American Statistician. 19 (3): 12–14. doi:10.2307/2681417. JSTOR 2681417.
  • Hart, John F.; et al. (1968). Computer Approximations. New York, NY: John Wiley & Sons, Inc. ISBN 978-0-88275-642-4.
  • "Normal Distribution", Encyclopedia of Mathematics, EMS Press, 2001 [1994]
  • Herrnstein, Richard J.; Murray, Charles (1994). The Bell Curve: Intelligence and Class Structure in American Life. Free Press. ISBN 978-0-02-914673-6.
  • Huxley, Julian S. (1932). Problems of Relative Growth. London. ISBN 978-0-486-61114-3. OCLC 476909537.
  • Johnson, Norman L.; Kotz, Samuel; Balakrishnan, Narayanaswamy (1994). Continuous Univariate Distributions, Volume 1. Wiley. ISBN 978-0-471-58495-7.
  • Johnson, Norman L.; Kotz, Samuel; Balakrishnan, Narayanaswamy (1995). Continuous Univariate Distributions, Volume 2. Wiley. ISBN 978-0-471-58494-0.
  • Karney, C. F. F. (2016). "Sampling exactly from the normal distribution". ACM Transactions on Mathematical Software. 42 (1): 3:1–14. arXiv:1303.6257. doi:10.1145/2710016. S2CID 14252035.
  • Kinderman, Albert J.; Monahan, John F. (1977). "Computer Generation of Random Variables Using the Ratio of Uniform Deviates". ACM Transactions on Mathematical Software. 3 (3): 257–260. doi:10.1145/355744.355750. S2CID 12884505.
  • Krishnamoorthy, Kalimuthu (2006). Handbook of Statistical Distributions with Applications. Chapman & Hall/CRC. ISBN 978-1-58488-635-8.
  • Kruskal, William H.; Stigler, Stephen M. (1997). Spencer, Bruce D. (ed.). Normative Terminology: 'Normal' in Statistics and Elsewhere. Statistics and Public Policy. Oxford University Press. ISBN 978-0-19-852341-3.
  • Laplace, Pierre-Simon de (1774). "Mémoire sur la probabilité des causes par les événements". Mémoires de l'Académie Royale des Sciences de Paris (Savants étrangers), Tome 6: 621–656. Translated by Stephen M. Stigler in Statistical Science 1 (3), 1986: JSTOR 2245476.
  • Laplace, Pierre-Simon (1812). Théorie analytique des probabilités [Analytical theory of probabilities]. Paris, Ve. Courcier.
  • Le Cam, Lucien; Lo Yang, Grace (2000). Asymptotics in Statistics: Some Basic Concepts (second ed.). Springer. ISBN 978-0-387-95036-5.
  • Leva, Joseph L. (1992). "A fast normal random number generator" (PDF). ACM Transactions on Mathematical Software. 18 (4): 449–453. CiteSeerX 10.1.1.544.5806. doi:10.1145/138351.138364. S2CID 15802663. Archived from the original (PDF) on July 16, 2010.
  • Lexis, Wilhelm (1878). "Sur la durée normale de la vie humaine et sur la théorie de la stabilité des rapports statistiques". Annales de Démographie Internationale. II. Paris: 447–462.
  • Lukacs, Eugene; King, Edgar P. (1954). "A Property of Normal Distribution". The Annals of Mathematical Statistics. 25 (2): 389–394. doi:10.1214/aoms/1177728796. JSTOR 2236741.
  • McPherson, Glen (1990). Statistics in Scientific Investigation: Its Basis, Application and Interpretation. Springer-Verlag. ISBN 978-0-387-97137-7.
  • Marsaglia, George; Tsang, Wai Wan (2000). "The Ziggurat Method for Generating Random Variables". Journal of Statistical Software. 5 (8). doi:10.18637/jss.v005.i08.
  • Marsaglia, George (2004). "Evaluating the Normal Distribution". Journal of Statistical Software. 11 (4). doi:10.18637/jss.v011.i04.
  • Maxwell, James Clerk (1860). "V. Illustrations of the dynamical theory of gases. — Part I: On the motions and collisions of perfectly elastic spheres". Philosophical Magazine. Series 4. 19 (124): 19–32. doi:10.1080/14786446008642818.
  • Monahan, J. F. (1985). "Accuracy in random number generation". Mathematics of Computation. 45 (172): 559–568. doi:10.1090/S0025-5718-1985-0804945-X.
  • Patel, Jagdish K.; Read, Campbell B. (1996). Handbook of the Normal Distribution (2nd ed.). CRC Press. ISBN 978-0-8247-9342-5.
  • Pearson, Karl (1901). "On Lines and Planes of Closest Fit to Systems of Points in Space" (PDF). Philosophical Magazine. 6. 2 (11): 559–572. doi:10.1080/14786440109462720. S2CID 125037489.
  • Pearson, Karl (1905). "'Das Fehlergesetz und seine Verallgemeinerungen durch Fechner und Pearson'. A rejoinder". Biometrika. 4 (1): 169–212. doi:10.2307/2331536. JSTOR 2331536.
  • Pearson, Karl (1920). "Notes on the History of Correlation". Biometrika. 13 (1): 25–45. doi:10.1093/biomet/13.1.25. JSTOR 2331722.
  • Rohrbasser, Jean-Marc; Véron, Jacques (2003). "Wilhelm Lexis: The Normal Length of Life as an Expression of the "Nature of Things"". Population. 58 (3): 303–322. doi:10.3917/pope.303.0303.
  • Shore, H (1982). "Simple Approximations for the Inverse Cumulative Function, the Density Function and the Loss Integral of the Normal Distribution". Journal of the Royal Statistical Society. Series C (Applied Statistics). 31 (2): 108–114. doi:10.2307/2347972. JSTOR 2347972.
  • Shore, H (2005). "Accurate RMM-Based Approximations for the CDF of the Normal Distribution". Communications in Statistics – Theory and Methods. 34 (3): 507–513. doi:10.1081/sta-200052102. S2CID 122148043.
  • Shore, H (2011). "Response Modeling Methodology". WIREs Comput Stat. 3 (4): 357–372. doi:10.1002/wics.151. S2CID 62021374.
  • Shore, H (2012). "Estimating Response Modeling Methodology Models". WIREs Comput Stat. 4 (3): 323–333. doi:10.1002/wics.1199. S2CID 122366147.
  • Stigler, Stephen M. (1978). "Mathematical Statistics in the Early States". The Annals of Statistics. 6 (2): 239–265. doi:10.1214/aos/1176344123. JSTOR 2958876.
  • Stigler, Stephen M. (1982). "A Modest Proposal: A New Standard for the Normal". The American Statistician. 36 (2): 137–138. doi:10.2307/2684031. JSTOR 2684031.
  • Stigler, Stephen M. (1986). The History of Statistics: The Measurement of Uncertainty before 1900. Harvard University Press. ISBN 978-0-674-40340-6.
  • Stigler, Stephen M. (1999). Statistics on the Table. Harvard University Press. ISBN 978-0-674-83601-3.
  • Walker, Helen M. (1985). "De Moivre on the Law of Normal Probability" (PDF). In Smith, David Eugene (ed.). A Source Book in Mathematics. Dover. ISBN 978-0-486-64690-9.
  • Wallace, C. S. (1996). "Fast pseudo-random generators for normal and exponential variates". ACM Transactions on Mathematical Software. 22 (1): 119–127. doi:10.1145/225545.225554. S2CID 18514848.
  • Weisstein, Eric W. "Normal Distribution". MathWorld.
  • West, Graeme (2009). "Better Approximations to Cumulative Normal Functions" (PDF). Wilmott Magazine: 70–76. Archived from the original (PDF) on February 29, 2012.
  • Zelen, Marvin; Severo, Norman C. (1964). Probability Functions (chapter 26). Handbook of mathematical functions with formulas, graphs, and mathematical tables, by Abramowitz, M.; and Stegun, I. A.: National Bureau of Standards. New York, NY: Dover. ISBN 978-0-486-61272-0.
  • edit

    Retrieved from "https://en.wikipedia.org/w/index.php?title=Normal_distribution&oldid=1233145167"
     



    Last edited on 7 July 2024, at 14:18  





    Languages

     


    Alemannisch
    العربية
    Asturianu
    Azərbaycanca
    تۆرکجه
     / Bân-lâm-gú
    Беларуская
    Български
    Bosanski
    Català
    Чӑвашла
    Čeština
    Cymraeg
    Dansk
    Deutsch
    Eesti
    Ελληνικά
    Español
    Esperanto
    Euskara
    فارسی
    Français
    Gaeilge
    Galego

    Հայերեն
    ि
    Hrvatski
    Bahasa Indonesia
    Íslenska
    Italiano
    עברית

    Қазақша
    Latina
    Latviešu
    Lietuvių
    Lombard
    Magyar
    Македонски

    Bahasa Melayu
    Nederlands

    Nordfriisk
    Norsk bokmål
    Norsk nynorsk
    Piemontèis
    Polski
    Português
    Română
    Русский
    Shqip
    Simple English
    Slovenčina
    Slovenščina
    Српски / srpski
    Srpskohrvatski / српскохрватски
    Sunda
    Suomi
    Svenska
    Tagalog
    ி
    Татарча / tatarça

    Türkçe
    Українська
    اردو
    Tiếng Vit

    ייִדיש


     

    Wikipedia


    This page was last edited on 7 July 2024, at 14:18 (UTC).

    Content is available under CC BY-SA 4.0 unless otherwise noted.



    Privacy policy

    About Wikipedia

    Disclaimers

    Contact Wikipedia

    Code of Conduct

    Developers

    Statistics

    Cookie statement

    Terms of Use

    Desktop