Normal distribution

Probability density function
Probability density function for the normal distribution
The red line is the standard normal distribution
Cumulative distribution function
Cumulative distribution function for the normal distribution
Colors match the image above
notation: \mathcal{N}(\mu,\,\sigma^2)
parameters: μR — mean (location)
σ2 ≥ 0 — variance (squared scale)
support: xR   if σ2 > 0
x = μ   if σ2 = 0
pdf: \frac{1}{\sqrt{2\pi\sigma^2}}\; e^{ -\frac{(x-\mu)^2}{2\sigma^2} }
cdf: \frac12\Big[1 + \operatorname{erf}\Big( \frac{x-\mu}{\sqrt{2\sigma^2}}\Big)\Big]
mean: μ
median: μ
mode: μ
variance: σ2
skewness: 0
ex.kurtosis: 0
entropy: \tfrac12 \ln(2 \pi e \, \sigma^2)
mgf: e^{ \mu t + \tfrac{1}{2}\sigma^2t^2 }
cf: e^{ i\mu t - \tfrac{1}{2}\sigma^2 t^2 }
Fisher information: \begin{pmatrix}1/\sigma^2&0\\0&1/(2\sigma^4)\end{pmatrix}

In probability theory and statistics, the normal distribution or Gaussian distribution is a continuous probability distribution that often gives a good description of data that cluster around the mean. The graph of the associated probability density function is bell-shaped, with a peak at the mean, and is known as the Gaussian function or bell curve.

The normal distribution is often used to describe, at least approximately, any variable that tends to cluster around the mean. For example, the heights of adult males in the United States are roughly normally distributed, with a mean of about 70 inches (1.8 m). Most men have a height close to the mean, though a small number of outliers have a height significantly above or below the mean. A histogram of male heights will appear similar to a bell curve, with the correspondence becoming closer if more data are used.

By the central limit theorem, under certain conditions the sum of a number of random variables with finite means and variances approaches a normal distribution as the number of variables increases. For this reason, the normal distribution is commonly encountered in practice, and is used throughout statistics, natural sciences, and social sciences[1] as a simple model for complex phenomena. For example, the observational error in an experiment is usually assumed to follow a normal distribution, and the propagation of uncertainty is computed using this assumption.

The Gaussian distribution was named after Carl Friedrich Gauss, who introduced it in 1809 as a way of rationalizing the method of least squares. One year later Laplace proved the first version of the central limit theorem, demonstrating that the normal distribution occurs as a limiting distribution of arithmetic means of independent, identically distributed random variables with finite second moment. For this reason the normal distribution is sometimes called Laplacian, especially in French-speaking countries.

Contents

Definition

The simplest case of a normal distribution is known as the standard normal distribution, described by the probability density function


    \phi(x) = \tfrac{1}{\sqrt{2\pi}}\, e^{- \frac{\scriptscriptstyle 1}{\scriptscriptstyle 2} x^2},

The constant \scriptstyle\ 1/\sqrt{2\pi} in this expression ensures that the total area under the curve ϕ(x) is equal to one,[proof] and 12 in the exponent makes the  width” of the curve (measured as half of the distance between the inflection points of the curve) also equal to one. It is traditional[2] in statistics to denote this function with the Greek letter ϕ (phi), whereas density functions for all other distributions are usually denoted with letters ƒ or p. The alternative glyph φ is also used quite often, however within this article we reserve  φ” to denote characteristic functions.

More generally, a normal distribution results from exponentiating a quadratic function (just as an exponential distribution results from exponentiating a linear function):


    f(x) = e^{a x^2 + b x + c}. \,

This yields the classic  bell curve” shape (provided that a < 0 so that the quadratic function is concave). Notice that f(x) > 0 everywhere. One can adjust a to control the  width” of the bell, then adjust b to move the central peak of the bell along the x-axis, and finally adjust c to control the  height” of the bell. For f(x) to be a true probability density function over R, one must choose c such that \scriptstyle\int_{-\infty}^\infty f(x)\,dx\ =\ 1 (which is only possible when a < 0).

Rather than using a, b, and c, it is far more common to describe a normal distribution by its mean μ = −b/(2a) and variance σ2 = −1/(2a). Changing to these new parameters allows us to rewrite the probability density function in a convenient standard form,


    f(x) = \tfrac{1}{\sqrt{2\pi\sigma^2}}\, e^{\frac{-(x-\mu)^2}{2\sigma^2}}
         = \tfrac{1}{\sigma}\, \phi\!\left(\tfrac{x-\mu}{\sigma}\right).

Notice that for a standard normal distribution, μ = 0 and σ2 = 1. The last part of the equation above shows that any other normal distribution can be regarded as a version of the standard normal distribution that has been stretched horizontally by a factor σ and then translated rightward by a distance μ. Thus, μ specifies the position of the bell curve’s central peak, and σ specifies the  width” of the bell curve.

The parameter μ is at the same time the mean, the median and the mode of the normal distribution. The parameter σ2 is called the variance; as for any random variable, it describes how concentrated the distribution is around its mean. The square root of σ2 is called the standard deviation and is the width of the density function.

The normal distribution is usually denoted by N(μ, σ2).[3] Commonly the letter N is written in calligraphic font (typed as \mathcal{N} in LaTeX). Thus when a random variable X is distributed normally with mean μ and variance σ2, we write


    X\ \sim\ \mathcal{N}(\mu,\,\sigma^2). \,

Alternative formulations

Some authors[4] instead of σ2 use its reciprocal τ = σ−2, which is called the precision. This parameterization has an advantage in numerical applications where σ2 is very close to zero and is more convenient to work with in analysis as τ is a natural parameter of the normal distribution. Another advantage of using this parameterization is in the study of conditional distributions in multivariate normal case.

The question which normal distribution should be called the  standard” one is also answered differently by various authors. Starting from the works of Gauss the standard normal was considered to be the one with variance σ2 = 1/2:


    f(x) = \tfrac{1}{\sqrt\pi}\,e^{-x^2}

Stigler (1982) goes even further and suggests the standard normal with variance σ2 = 1/(2π):


    f(x) = e^{-\pi x^2}

According to the author, this formulation is advantageous because of a much simpler and easier-to-remember formula, the fact that the pdf has unit height at zero, and simple approximate formulas for the quantiles of the distribution.

Characterization

In the previous section the normal distribution was defined by specifying its probability density function. However there are other ways to characterize a probability distribution. They include: the cumulative distribution function, the moments, the cumulants, the characteristic function, the moment-generating function, etc.

Probability density function

The probability density function (pdf) of a random variable describes the relative frequencies of different values for that random variable. The pdf of the normal distribution is given by the formula explained in detail in the previous section:


    f(x;\,\mu,\sigma^2) = \frac{1}{\sqrt{2\pi\sigma^2}} \, e^{-(x-\mu)^2\!/(2\sigma^2)}
                        = \frac{1}{\sigma} \,\phi\!\left(\frac{x-\mu}{\sigma}\right),
    \qquad x\in\mathbb{R}.

This is a proper function only when the variance σ2 is not equal to zero. In that case this is a continuous smooth function, defined on the entire real line, and which is called the  Gaussian function”.

When σ2 = 0, the density function doesn’t exist. However we can consider a generalized function that would behave in a manner similar to the regular density function (in the sense that it defines a measure on the real line, and it can be plugged in into an integral in order to calculate expected values of different quantities):


    f(x;\,\mu,0) = \delta(x-\mu).

This is the Dirac delta function, it is equal to infinity at x = μ and is zero elsewhere.

Properties:

Cumulative distribution function

The cumulative distribution function (cdf) describes probabilities for a random variable to fall in the intervals of the form (−∞, x]. The cdf of the standard normal distribution is denoted with the capital Greek letter Φ (phi), and can be computed as an integral of the probability density function:


    \Phi(x) = \frac{1}{\sqrt{2\pi}} \int_{-\infty}^x e^{-t^2/2} \, dt
            = \frac12\Big[\, 1 + \operatorname{erf}\Big(\frac{x}{\sqrt{2}}\Big)\,\Big],\quad x\in\mathbb{R}.

This integral can only be expressed in terms of a special function erf, called the error function. The numerical methods for calculation of the standard normal cdf are discussed below. For a generic normal random variable with mean μ and variance σ2 > 0 the cdf will be equal to


  F(x;\,\mu,\sigma^2)
    = \Phi\Big(\frac{x-\mu}{\sigma}\Big)
    = \frac12\Big[\, 1 + \operatorname{erf}\Big(\frac{x-\mu}{\sigma\sqrt{2}}\Big)\,\Big],\quad x\in\mathbb{R}.

For a normal distribution with zero variance, the cdf is the Heaviside step function:


    F(x;\,\mu,0) = \mathbf{1}\{x\geq\mu\}\,.

The complement of the standard normal cdf, Q(x) = 1 − Φ(x), is referred to as the Q-function, especially in engineering texts.[6][7] This represents the tail probability of the Gaussian distribution, that is the probability that a standard normal random variable X is greater than the number x. Other definitions of the Q-function, all of which are simple transformations of Φ, are also used occasionally.[8]

Properties:

Quantile function

The inverse of the standard normal cdf, called the quantile function or probit function, is expressed in terms of the inverse error function:


    \Phi^{-1}(p) \equiv z_p = \sqrt2\;\operatorname{erf}^{-1}(2p - 1), \quad p\in(0,1).

Quantiles of the standard normal distribution are commonly denoted as zp. The quantile zp represents such a value that a standard normal random variable X has the probability of exactly p to fall inside the (−∞, zp] interval. The quantiles are used in hypothesis testing, construction of confidence intervals and Q-Q plots. The most  famous” normal quantile is 1.96 = z0.975. A standard normal random variable is greater than 1.96 in absolute value in only 5% of cases.

For a normal random variable with mean μ and variance σ2, the quantile function is


    F^{-1}(p;\,\mu,\sigma^2)
      = \mu + \sigma\Phi^{-1}(p)
      = \mu + \sigma\sqrt2\,\operatorname{erf}^{-1}(2p - 1), \quad p\in(0,1).

Characteristic function and moment generating function

The characteristic function φX(t) of a random variable X is defined as the expected value of eitX, where i is the imaginary unit, and t ∈ R is the argument of the characteristic function. Thus the characteristic function is the Fourier transform of the density ϕ(x). For a normally distributed X with mean μ and variance σ2, the characteristic function is [9]


    \varphi(t;\,\mu,\sigma^2) = \int_{-\infty}^\infty\! e^{itx}\tfrac{1}{\sqrt{2\pi\sigma^2}}e^{-\frac12 (x-\mu)^2/\sigma^2} dx = e^{i\mu t - \frac12 \sigma^2t^2}.

The moment generating function is defined as the expected value of etX. For a normal distribution, the moment generating function exists and is equal to


    M(t;\, \mu,\sigma^2) = \operatorname{E}[e^{tX}] = \varphi(-it;\, \mu,\sigma^2) = e^{ \mu t + \frac12 \sigma^2 t^2 }.

The cumulant generating function is the logarithm of the moment generating function:


    g(t;\,\mu,\sigma^2) = \ln M(t;\,\mu,\sigma^2) = \mu t + \tfrac{1}{2} \sigma^2 t^2.

Since this is a quadratic polynomial in t, only the first two cumulants are nonzero.

Moments

The normal distribution has moments of all orders. That is, for a normally distributed X with mean μ and variance σ 2, the expectation E|X|p exists and is finite for all p such that Re[p] > −1. Usually we are interested only in moments of integer orders: p = 1, 2, 3, ….

Order Raw moment Central moment Cumulant
1 μ 0 μ
2 μ2 + σ2 σ 2 σ 2
3 μ3 + 3μσ2 0 0
4 μ4 + 6μ2σ2 + 3σ4 3σ 4 0
5 μ5 + 10μ3σ2 + 15μσ4 0 0
6 μ6 + 15μ4σ2 + 45μ2σ4 + 15σ6 15σ 6 0
7 μ7 + 21μ5σ2 + 105μ3σ4 + 105μσ6 0 0
8 μ8 + 28μ6σ2 + 210μ4σ4 + 420μ2σ6 + 105σ8 105σ 8 0

Properties

Standardizing normal random variables

As a consequence of property 1, it is possible to relate all normal random variables to the standard normal. For example if X is normal with mean μ and variance σ2, then


    Z = \frac{X - \mu}{\sigma}

has mean zero and unit variance, that is Z has the standard normal distribution. Conversely, having a standard normal random variable Z we can always construct another normal random variable with specific mean μ and variance σ2:


    X = \sigma Z + \mu. \,

This  standardizing” transformation is convenient as it allows one to compute the pdf and especially the cdf of a normal distribution having the table of pdf and cdf values for the standard normal. They will be related via


    F_X(x) = \Phi\bigg(\frac{x-\mu}{\sigma}\bigg), \quad
    f_X(x) = \frac{1}{\sigma}\,\phi\bigg(\frac{x-\mu}{\sigma}\bigg).

Standard deviation and confidence intervals

Dark blue is less than one standard deviation from the mean. For the normal distribution, this accounts for about 68% of the set (dark blue), while two standard deviations from the mean (medium and dark blue) account for about 95%, and three standard deviations (light, medium, and dark blue) account for about 99.7%.

About 68% of values drawn from a normal distribution are within one standard deviation σ > 0 away from the mean μ; about 95% of the values are within two standard deviations and about 99.7% lie within three standard deviations. This is known as the 68-95-99.7 rule, or the empirical rule, or the 3-sigma rule.

To be more precise, the area under the bell curve between μ −  and μ +  in terms of the cumulative normal distribution function is given by

\begin{align}
    & F(\mu+n\sigma;\,\mu,\sigma^2) - F(\mu-n\sigma;\,\mu,\sigma^2) \\
    &\quad = \Phi(n)-\Phi(-n) = 2\Phi(n)-1=\mathrm{erf}\left(\frac{n}{\sqrt{2}}\right),\end{align}

where erf is the error function. To 12 decimal places, the values for the 1-, 2-, up to 6-sigma points are:

\scriptstyle\; n\; \;\scriptstyle\mathrm{erf}\left(\frac{n}{\sqrt{2}}\right)\; i.e. 1 minus ... or 1 in ...
1 0.682689492137 0.317310507863 3.15148718753
2 0.954499736104 0.045500263896 21.9778945081
3 0.997300203937 0.002699796063 370.398347380
4 0.999936657516 0.000063342484 15,787.192684
5 0.999999426697 0.000000573303 1,744,278.331
6 0.999999998027 0.000000001973 506,842,375.7

The next table gives the reverse relation of sigma multiples corresponding to a few often used values for the area under the bell curve. These values are useful to determine (asymptotic) confidence intervals of the specified levels based on normally distributed (or asymptotically normal) estimators:

\scriptstyle\;\mathrm{erf}\left(\frac{n}{\sqrt{2}}\right)\; \scriptstyle\;n\;
0.80 1.281551565545
0.90 1.644853626951
0.95 1.959963984540
0.98 2.326347874041
0.99 2.575829303549
0.995 2.807033768344
0.998 3.090232306168
0.999 3.290526731492
0.9999 3.890591886413
0.99999 4.417173413469

where the value on the left of the table is the proportion of values that will fall within a given interval and n is a multiple of the standard deviation that specifies the width of the interval.

Central limit theorem

The theorem states that under certain, fairly common conditions, the sum of a large number of random variables will have an approximately normal distribution. For example if (x1, …, xn) is a sequence of iid random variables, each having mean μ and variance σ2 but otherwise distributions of xi’s can be arbitrary, then the central limit theorem states that


    \sqrt{n}\bigg( \frac{1}{n}\sum_{i=1}^n x_i - \mu \bigg)\ \xrightarrow{d}\ \mathcal{N}(0,\,\sigma^2).

The theorem will hold even if the summands xi are not iid, although some constraints on the degree of dependence and the growth rate of moments still have to be imposed.

The importance of the central limit theorem cannot be overemphasized. A great number of test statistics, scores, and estimators encountered in practice contain sums of certain random variables in them, even more estimators can be represented as sums of random variables through the use of influence functions — all of these quantities are governed by the central limit theorem and will have asymptotically normal distribution as a result.

As the number of discrete events increases, the function begins to resemble a normal distribution

Another practical consequence of the central limit theorem is that certain other 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.

Miscellaneous

  1. The family of normal distributions is closed under linear transformations. That is, if X is normally distributed with mean μ and variance σ2, then a linear transform aX + b (for some real numbers a ≠ 0 and b) is also normally distributed:
    
    aX + b\ \sim\ \mathcal{N}(a\mu+b,\, a^2\sigma^2).
    Also if X1, X2 are two independent normal random variables, with means μ1, μ2 and standard deviations σ1, σ2, then their linear combination will also be normally distributed: [proof]
    
    aX_1 + bX_2 \ \sim\ \mathcal{N}(a\mu_1+b\mu_2,\, a^2\!\sigma_1^2 + b^2\sigma_2^2)
  2. The converse of (1) is also true: if X1 and X2 are independent and their sum X1 + X2 is distributed normally, then both X1 and X2 must also be normal. This is known as Cramér’s theorem.
  3. It is sometimes mistakenly believed that if two normal random variables are uncorrelated then they are also independent. That is false[proof]. The correct statement is that if the two random variables are jointly normal and uncorrelated, only then they are independent.
  4. Normal distribution is infinitely divisible: for a normally distributed X with mean μ and variance σ2 we can find n independent random variables {X1, …, Xn} each distributed normally with means μ/n and variances σ2/n such that
    
    X_1 + X_2 + \cdots + X_n \ \sim\ \mathcal{N}(\mu, \sigma^2)
  5. Normal distribution is stable (with exponent α = 2): if X1, X2 are two independent N(μ, σ2) random variables and a, b are arbitrary real numbers, then
    
    aX_1 + bX_2 \ \sim\ \sqrt{a^2+b^2}\cdot X_3\ +\ (a+b-\sqrt{a^2+b^2})\mu,
    where X3 is also N(μ, σ2). This relationship directly follows from property (1).
  6. The Kullback–Leibler divergence between two normal distributions X1N(μ1, σ21 )and X2N(μ2, σ22 )is given by:[10]
    
    D_\mathrm{KL}( X_1 \,\|\, X_2 ) = \frac{(\mu_1 - \mu_2)^2}{2\sigma_2^2} \,+\, \frac12\Bigg(\, \frac{\sigma_1^2}{\sigma_2^2} - 1 - \ln\frac{\sigma_1^2}{\sigma_2^2} \,\Bigg)\ .
    The Hellinger distance between the same distributions is equal to
    
    H^2(X_1,X_2) = 1 \,-\, \sqrt{\frac{2\sigma_1\sigma_2}{\sigma_1^2+\sigma_2^2}} \;
                           e^{-\frac{1}{4}\frac{(\mu_1-\mu_2)^2}{\sigma_1^2+\sigma_2^2}}\ .
  7. The Fisher information matrix for normal distribution is diagonal and takes form
    
    \mathcal I = \begin{pmatrix} \frac{1}{\sigma^2} & 0 \\ 0 & \frac{1}{2\sigma^4} \end{pmatrix}
  8. Normal distributions belongs to an exponential family with natural parameters  \scriptstyle\theta_1=\frac{\mu}{\sigma^2} and \scriptstyle\theta_2=\frac{-1}{2\sigma^2}, and natural statistics x and x2. The dual, expectation parameters for normal distribution are η1 = μ and η2 = μ2 + σ2.
  9. Of all probability distributions over the reals with mean μ and variance σ2, the normal distribution N(μ, σ2) is the one with the maximum entropy.
  10. The family of normal distributions forms a manifold with constant curvature −1. The same family is flat with respect to the (±1)-connections ∇(e) and ∇(m).[11]

Related distributions

Extensions

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. All these extensions are also called normal or Gaussian laws, so a certain ambiguity in names exists.

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:

Normality tests

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. A great number of tests (over 40) have been devised for this problem, the more prominent of them are outlined below:

Estimation of parameters

It is often the case that we don’t know the parameters of the normal distribution, but instead want to estimate them. That is, having a sample (x1, …, xn) from a normal N(μ, σ2) population we would like to learn the approximate values of parameters μ and σ2. The standard approach to this problem is the maximum likelihood method, which requires maximization of the log-likelihood function:


   \ln\mathcal{L}(\mu,\sigma^2)
     = \sum_{i=1}^n \ln f(x_i;\,\mu,\sigma^2)
     = -\frac{n}{2}\ln(2\pi) - \frac{n}{2}\ln\sigma^2 - \frac{1}{2\sigma^2}\sum_{i=1}^n (x_i-\mu)^2.

Taking derivatives with respect to μ and σ2 and solving the resulting system of first order conditions yields the maximum likelihood estimates:


    \hat{\mu} = \overline{x} \equiv \frac{1}{n}\sum_{i=1}^n x_i, \qquad
    \hat{\sigma}^2 = \frac{1}{n} \sum_{i=1}^n (x_i - \overline{x})^2.

Estimator \scriptstyle\hat\mu is called the sample mean, since it is the arithmetic mean of all observations. The statistic \scriptstyle\overline{x} is complete and sufficient for μ, and therefore by the Lehmann–Scheffé theorem, \scriptstyle\hat\mu is the uniformly minimum variance unbiased (UMVU) estimator. In finite samples it is distributed normally:


    \hat\mu \ \sim\ \mathcal{N}(\mu,\,\,\sigma^2\!\!\;/n).

The variance of this estimator is equal to the μμ-element of the inverse Fisher information matrix \scriptstyle\mathcal{I}^{-1}. This implies that the estimator is finite-sample efficient. Of practical importance is the fact that the standard error of \scriptstyle\hat\mu is proportional to \scriptstyle1/\sqrt{n}, 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, \scriptstyle\hat\mu is consistent, that is, it converges in probability to μ as n → ∞. The estimator is also asymptotically normal, which is a simple corollary of the fact that it is normal in finite samples:


    \sqrt{n}(\hat\mu-\mu) \ \xrightarrow{d}\ \mathcal{N}(0,\,\sigma^2).

The estimator \scriptstyle\hat\sigma^2 is called the sample variance, since it is the variance of the sample (x1, …, xn). In practice, another estimator is often used instead of the \scriptstyle\hat\sigma^2. This other estimator is denoted s2, and is also called the sample variance, which represents a certain ambiguity in terminology; its square root s is called the sample standard deviation. The estimator s2 differs from \scriptstyle\hat\sigma^2 by having (n − 1) instead of n in the denominator (the so called Bessel’s correction):


    s^2 = \frac{n}{n-1}\,\hat\sigma^2 = \frac{1}{n-1} \sum_{i=1}^n (x_i - \overline{x})^2.

The difference between s2 and \scriptstyle\hat\sigma^2 becomes negligibly small for large n’s. In finite samples however, the motivation behind the use of s2 is that it is an unbiased estimator of the underlying parameter σ2, whereas \scriptstyle\hat\sigma^2 is biased. Also, by the Lehmann–Scheffé theorem the estimator s2 is uniformly minimum variance unbiased (UMVU), which makes it the  best” estimator among all unbiased ones. However it can be shown that the biased estimator \scriptstyle\hat\sigma^2 is  better” than the s2 in terms of the mean squared error (MSE) criterion. In finite samples both s2 and \scriptstyle\hat\sigma^2 have scaled chi-squared distribution with (n − 1) degrees of freedom:


    s^2 \ \sim\ \tfrac{\sigma^2}{n-1} \cdot \chi^2_{n-1}, \qquad
    \hat\sigma^2 \ \sim\ \tfrac{\sigma^2}{n} \cdot \chi^2_{n-1}\ .

The first of these expressions shows that the variance of s2 is equal to 2σ4/(n−1), which is slightly greater than the σσ-element of the inverse Fisher information matrix \scriptstyle\mathcal{I}^{-1}. Thus, s2 is not an efficient estimator for σ2, and moreover, since s2 is UMVU, we can conclude that the finite-sample efficient estimator for σ2 does not exist.

Applying the asymptotic theory, both estimators s2 and \scriptstyle\hat\sigma^2 are consistent, that is they converge in probability to σ2 as the sample size n → ∞. The two estimators are also both asymptotically normal:


    \sqrt{n}(\hat\sigma^2 - \sigma^2) \simeq
    \sqrt{n}(s^2-\sigma^2)\ \xrightarrow{d}\ \mathcal{N}(0,\,2\sigma^4).

In particular, both estimators are asymptotically efficient for σ2.

By Cochran’s theorem, for normal distribution the sample mean \scriptstyle\hat\mu and the sample variance s2 are independent, which means there can be no gain in considering their joint distribution. There is also a reverse 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 \scriptstyle\hat\mu and s can be employed to construct the so-called t-statistic:


    t = \frac{\hat\mu-\mu}{s/\sqrt{n}} = \frac{\overline{x}-\mu}{\sqrt{\frac{1}{n(n-1)}\sum(x_i-\overline{x})^2}}\ \sim\ t_{n-1}

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; similarly, inverting the χ2 distribution of the statistic s2 will give us the confidence interval for σ2:

\begin{align}
    & \mu \in \Big[\, \hat\mu + t_{n-1,\alpha/2}\,  \tfrac{1}{\sqrt{n}}s,\ \ 
                      \hat\mu + t_{n-1,1-\alpha/2}\,\tfrac{1}{\sqrt{n}}s \,\Big] \approx
              \Big[\, \hat\mu - |z_{\alpha/2}|\tfrac{1}{\sqrt n}s,\ \ 
                      \hat\mu + |z_{\alpha/2}|\tfrac{1}{\sqrt n}s \,\Big], \\
    & \sigma^2 \in \bigg[\, \frac{(n-1)s^2}{\chi^2_{n-1,1-\alpha/2}},\ \ 
                            \frac{(n-1)s^2}{\chi^2_{n-1,\alpha/2}} \,\bigg] \approx
                   \Big[\, s^2 - |z_{\alpha/2}|\tfrac{\sqrt{2}}{\sqrt{n}}s^2,\ \ 
                           s^2 + |z_{\alpha/2}|\tfrac{\sqrt{2}}{\sqrt{n}}s^2 \,\Big],
  \end{align}

where tk,p and χk,p2 are the pth quantiles of the t- and χ2-distributions respectively. These confidence intervals are of the level 1 − α, meaning that the true values μ and σ2 fall outside of these intervals with probability α. In practice people usually take α = 5%, resulting in the 95% confidence intervals. The approximate formulas in the display above were derived from the asymptotic distributions of \scriptstyle\hat\mu 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.

Occurrence

The occurrence of normal distribution in practical problems can be loosely classified into three 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 one of the simplest and most convenient to use, frequently researchers are tempted to assume that certain quantity is distributed normally, without justifying such assumption rigorously. In fact, the maturity of a scientific field can be judged by the prevalence of the normality assumption in its methods.

Exact normality

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

Approximately normal distributions occur in many situations, as explained by the central limit theorem. When the outcome is produced by a large number of 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 which has a considerably larger magnitude than the rest of the effects.

Assumed normality

Fisher iris versicolor sepalwidth.svg
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. — Pearson (1901)

There are statistical methods to empirically test that assumption, see the #Normality tests section.

Generating values from normal distribution

The bean machine, a device invented by sir 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.

For computer simulations, especially in applications of Monte-Carlo method, it is often useful to generate values that have a normal distribution. All algorithms described here are concerned with generating the standard normal, since a N(μ, σ2) can be generated as X = μ + σZ, where Z is standard normal. The algorithms rely on the availability of a random number generator capable of producing random values distributed uniformly.

Numerical approximations for the normal cdf

The standard normal cdf 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.

History

Some authors[17][18] attribute at least partially the credit for the discovery of the normal distribution to de Moivre, who in 1738 published in the second edition of his  The Doctrine of Chances[19] [note 2] the study of the coefficients in the binomial expansion of (a + b)n. De Moivre proved that the middle term in this expansion has the approximate magnitude of \scriptstyle 2/\sqrt{2\pi n}, and that  If m or ½n 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 \scriptstyle -\frac{2\ell\ell}{n}.” 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.[20]

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

In 1809 Gauss published the monograph  Theoria motus corporum coelestium in sectionibus conicis solem ambientium” where among other things he introduces and describes several important statistical concepts, such as the method of least squares, the method of maximum likelihood, and the normal distribution. If the notation M, M′, M′′, … is used for the measurements of some unknown quantity V, Gauss seeks the  most probable” estimator: the one which maximizes the probability φ(M−V) · φ(M′−V) · φ(M′′−V) · … of obtaining the observed experimental results. In his notation φΔ is the probability law 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. Starting from these principles, Gauss demonstrates that the only law which rationalizes the choice of arithmetic mean as an estimator of the location parameter, is the normal law of errors:[21]


    \varphi\Delta = \tfrac{h}{\surd\pi}\, e^{-hh\Delta\Delta}\ ,

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 (NWLS) method.[22]

Marquis de 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, the merit of the contributions of Laplace cannot be underestimated.[note 3] It was Laplace who first posed the problem of aggregating several observations in 1774,[23] although his own solution led to the Laplacian distribution. It was Laplace who first calculated the value of the integral ∫ et ²dt = √π in 1782, providing the normalization constant for the normal distribution.[24] 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.[25]

It is of interest to note that in 1809 an American mathematician Adrain published two derivations of the normal probability law, simultaneously and independently from Gauss.[26] His works remained unnoticed until 1871 when they were rediscovered by Abbe,[27] mainly because the scientific community was virtually non-existent in the United States at that time.

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:[28] “The number of particles whose velocity, resolved in a certain direction, lies between x and x+dx is

 \mathrm{N}\frac{1}{\alpha\sqrt\pi}e^{-\frac{x^2}{\alpha^2}}dx

Since its introduction, the normal distribution has been known by many different names: the law of error, the law of facility of errors, Laplace’s second law, Gaussian law, etc. By the end of the 19th century some authors[note 4] start to occasionally use the name normal distribution, where the word “normal” is used as an adjective — the term was derived from the fact that this distribution was seen as typical, common, normal. Around the turn of the 20th century Pearson popularizes the term normal as a designation for this distribution.[29]

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.’ — Pearson (1920)

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:

 df = \frac{1}{\sigma\sqrt{2\pi}}e^{-\frac{(x-m)^2}{2\sigma^2}}dx

The term “standard normal” which denotes the normal distribution with zero mean and unit variance came into general use around 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”.[30]

See also

Notes

  1. For example, this algorithm is given in the article Bc programming language.
  2. 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 Helen M. Walker (1985).
  3.   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 (1904)
  4. Such use is encountered in the works of Peirce, Galton and Lexis approximately around 1875.

References

  1. Gale Encyclopedia of Psychology — Normal Distribution
  2. Halperin & et al. (1965, item 7)
  3. McPherson (1990) page 110
  4. Bernardo & Smith (2000)
  5. Patel & Read (1996, [2.1.8])
  6. Scott, Clayton; Robert Nowak (August 7, 2003). "The Q-function". Connexions. http://cnx.org/content/m11537/1.2/. 
  7. Barak, Ohad (April 6, 2006). "Q function and error function". Tel Aviv University. http://www.eng.tau.ac.il/~jo/academic/Q.pdf. 
  8. Weisstein, Eric W., "Normal Distribution Function" from MathWorld.
  9. Sanders, Mathijs A.. "Characteristic function of the univariate normal distribution". http://www.planetmathematics.com/CharNormal.pdf. Retrieved 2009-03-06. 
  10. http://www.allisons.org/ll/MML/KL/Normal/
  11. Amari & Nagaoka (2000)
  12. Mathworld entry for Normal Product Distribution
  13. Huxley (1932)
  14. Whittaker, E. T.; Robinson, G. (1967). The Calculus of Observations: A Treatise on Numerical Mathematics. New York: Dover. p. 179. 
  15. Johnson et al. (1995, Equation (26.48))
  16. Kinderman & Monahan (1976)
  17. Johnson et al. (1994, page 85)
  18. Le Cam (2000, p. 74)
  19. De Moivre (1738)
  20. Stigler (1986, p. 76)
  21. Gauss (1809, section 177)
  22. Gauss (1809, section 179)
  23. Laplace (1774, Problem III)
  24. Pearson (1905, p. 189)
  25. Stigler (1986, p. 144)
  26. Stigler (1978, p. 243)
  27. Stigler (1978, p. 244)
  28. Maxwell (1860), p. 23
  29. Kruskal & Stigler (1997)
  30. "Earliest uses… (entry STANDARD NORMAL CURVE)". http://jeff560.tripod.com/s.html. 

Literature