Maximum spacing estimation

Last updated
The maximum spacing method tries to find a distribution function such that the spacings, D(i), are all approximately of the same length. This is done by maximizing their geometric mean. Spacings.svg
The maximum spacing method tries to find a distribution function such that the spacings, D(i), are all approximately of the same length. This is done by maximizing their geometric mean.

In statistics, maximum spacing estimation (MSE or MSP), or maximum product of spacing estimation (MPS), is a method for estimating the parameters of a univariate statistical model. [1] The method requires maximization of the geometric mean of spacings in the data, which are the differences between the values of the cumulative distribution function at neighbouring data points.

Contents

The concept underlying the method is based on the probability integral transform, in that a set of independent random samples derived from any random variable should on average be uniformly distributed with respect to the cumulative distribution function of the random variable. The MPS method chooses the parameter values that make the observed data as uniform as possible, according to a specific quantitative measure of uniformity.

One of the most common methods for estimating the parameters of a distribution from data, the method of maximum likelihood (MLE), can break down in various cases, such as involving certain mixtures of continuous distributions. [2] In these cases the method of maximum spacing estimation may be successful.

Apart from its use in pure mathematics and statistics, the trial applications of the method have been reported using data from fields such as hydrology, [3] econometrics, [4] magnetic resonance imaging, [5] and others. [6]

History and usage

The MSE method was derived independently by Russel Cheng and Nik Amin at the University of Wales Institute of Science and Technology, and Bo Ranneby at the Swedish University of Agricultural Sciences. [2] The authors explained that due to the probability integral transform at the true parameter, the “spacing” between each observation should be uniformly distributed. This would imply that the difference between the values of the cumulative distribution function at consecutive observations should be equal. This is the case that maximizes the geometric mean of such spacings, so solving for the parameters that maximize the geometric mean would achieve the “best” fit as defined this way. Ranneby (1984) justified the method by demonstrating that it is an estimator of the Kullback–Leibler divergence, similar to maximum likelihood estimation, but with more robust properties for some classes of problems.

There are certain distributions, especially those with three or more parameters, whose likelihoods may become infinite along certain paths in the parameter space. Using maximum likelihood to estimate these parameters often breaks down, with one parameter tending to the specific value that causes the likelihood to be infinite, rendering the other parameters inconsistent. The method of maximum spacings, however, being dependent on the difference between points on the cumulative distribution function and not individual likelihood points, does not have this issue, and will return valid results over a much wider array of distributions. [1]

The distributions that tend to have likelihood issues are often those used to model physical phenomena. Hall & al. (2004) seek to analyze flood alleviation methods, which requires accurate models of river flood effects. The distributions that better model these effects are all three-parameter models, which suffer from the infinite likelihood issue described above, leading to Hall's investigation of the maximum spacing procedure. Wong & Li (2006), when comparing the method to maximum likelihood, use various data sets ranging from a set on the oldest ages at death in Sweden between 1905 and 1958 to a set containing annual maximum wind speeds.

Definition

Given an iid random sample {x1, ..., xn} of size n from a univariate distribution with continuous cumulative distribution function F(x;θ0), where θ0 ∈ Θ is an unknown parameter to be estimated, let {x(1), ..., x(n)} be the corresponding ordered sample, that is the result of sorting of all observations from smallest to largest. For convenience also denote x(0) = −∞ and x(n+1) = +∞.

Define the spacings as the “gaps” between the values of the distribution function at adjacent ordered points: [7]

Then the maximum spacing estimator of θ0 is defined as a value that maximizes the logarithm of the geometric mean of sample spacings:

By the inequality of arithmetic and geometric means, function Sn(θ) is bounded from above by −ln(n+1), and thus the maximum has to exist at least in the supremum sense.

Note that some authors define the function Sn(θ) somewhat differently. In particular, Ranneby (1984) multiplies each Di by a factor of (n+1), whereas Cheng & Stephens (1989) omit the 1n+1 factor in front of the sum and add the “−” sign in order to turn the maximization into minimization. As these are constants with respect to θ, the modifications do not alter the location of the maximum of the function Sn.

Examples

This section presents two examples of calculating the maximum spacing estimator.

Example 1

Plots of the log value of l for the simplistic example under both likelihood and spacing estimation. The values for which both likelihood and spacing are maximized, the maximum likelihood and maximum spacing estimates, are identified. Spacing Estimation plot for MSE example.svg
Plots of the log value of λ for the simplistic example under both likelihood and spacing estimation. The values for which both likelihood and spacing are maximized, the maximum likelihood and maximum spacing estimates, are identified.

Suppose two values x(1) = 2, x(2) = 4 were sampled from the exponential distribution F(x;λ) = 1 − e, x ≥ 0 with unknown parameter λ > 0. In order to construct the MSE we have to first find the spacings:

iF(x(i))F(x(i−1))Di = F(x(i)) − F(x(i−1))
11 − e−2λ01 − e−2λ
21 − e−4λ1 − e−2λe−2λ − e−4λ
311 − e−4λe−4λ

The process continues by finding the λ that maximizes the geometric mean of the “difference” column. Using the convention that ignores taking the (n+1)st root, this turns into the maximization of the following product: (1 − e−2λ) · (e−2λ − e−4λ) · (e−4λ). Letting μ = e−2λ, the problem becomes finding the maximum of μ5−2μ4+μ3. Differentiating, the μ has to satisfy 5μ4−8μ3+3μ2 = 0. This equation has roots 0, 0.6, and 1. As μ is actually e−2λ, it has to be greater than zero but less than one. Therefore, the only acceptable solution is

which corresponds to an exponential distribution with a mean of 1λ ≈ 3.915. For comparison, the maximum likelihood estimate of λ is the inverse of the sample mean, 3, so λMLE = ⅓ ≈ 0.333.

Example 2

Suppose {x(1), ..., x(n)} is the ordered sample from a uniform distribution U(a,b) with unknown endpoints a and b. The cumulative distribution function is F(x;a,b) = (xa)/(ba) when x∈[a,b]. Therefore, individual spacings are given by

Calculating the geometric mean and then taking the logarithm, statistic Sn will be equal to

Here only three terms depend on the parameters a and b. Differentiating with respect to those parameters and solving the resulting linear system, the maximum spacing estimates will be

These are known to be the uniformly minimum variance unbiased (UMVU) estimators for the continuous uniform distribution. [1] In comparison, the maximum likelihood estimates for this problem and are biased and have higher mean-squared error.

Properties

Consistency and efficiency

J-shaped density.svg
Density
J-shaped density - corresponding distribution.svg
Distribution
Plot of a “J-shaped” density function and its corresponding distribution. A shifted Weibull with a scale parameter of 15, a shape parameter of 0.5, and a location parameter of 10. The density asymptotically approaches infinity as x approaches 10, rendering the estimates of the other parameters inconsistent. Note that there is no inflection point in the graph of the distribution.

The maximum spacing estimator is a consistent estimator in that it converges in probability to the true value of the parameter, θ0, as the sample size increases to infinity. [2] The consistency of maximum spacing estimation holds under much more general conditions than for maximum likelihood estimators. In particular, in cases where the underlying distribution is J-shaped, maximum likelihood will fail where MSE succeeds. [1] An example of a J-shaped density is the Weibull distribution, specifically a shifted Weibull, with a shape parameter less than 1. The density will tend to infinity as x approaches the location parameter rendering estimates of the other parameters inconsistent.

Maximum spacing estimators are also at least as asymptotically efficient as maximum likelihood estimators, where the latter exist. However, MSEs may exist in cases where MLEs do not. [1]

Sensitivity

Maximum spacing estimators are sensitive to closely spaced observations, and especially ties. [8] Given

we get

When the ties are due to multiple observations, the repeated spacings (those that would otherwise be zero) should be replaced by the corresponding likelihood. [1] That is, one should substitute for , as

since .

When ties are due to rounding error, Cheng & Stephens (1989) suggest another method to remove the effects. [note 1] Given r tied observations from xi to xi+r−1, let δ represent the round-off error. All of the true values should then fall in the range . The corresponding points on the distribution should now fall between and . Cheng and Stephens suggest assuming that the rounded values are uniformly spaced in this interval, by defining

The MSE method is also sensitive to secondary clustering. [8] One example of this phenomenon is when a set of observations is thought to come from a single normal distribution, but in fact comes from a mixture normals with different means. A second example is when the data is thought to come from an exponential distribution, but actually comes from a gamma distribution. In the latter case, smaller spacings may occur in the lower tail. A high value of M(θ) would indicate this secondary clustering effect, and suggesting a closer look at the data is required. [8]

Moran test

The statistic Sn(θ) is also a form of Moran or Moran-Darling statistic, M(θ), which can be used to test goodness of fit. [note 2] It has been shown that the statistic, when defined as

is asymptotically normal, and that a chi-squared approximation exists for small samples. [8] In the case where we know the true parameter , Cheng & Stephens (1989) show that the statistic has a normal distribution with

where γ is the Euler–Mascheroni constant which is approximately 0.57722. [note 3]

The distribution can also be approximated by that of , where

,

in which

and where follows a chi-squared distribution with degrees of freedom. Therefore, to test the hypothesis that a random sample of values comes from the distribution , the statistic can be calculated. Then should be rejected with significance if the value is greater than the critical value of the appropriate chi-squared distribution. [8]

Where θ0 is being estimated by , Cheng & Stephens (1989) showed that has the same asymptotic mean and variance as in the known case. However, the test statistic to be used requires the addition of a bias correction term and is:

where is the number of parameters in the estimate.

Generalized maximum spacing

Alternate measures and spacings

Ranneby & Ekström (1997) generalized the MSE method to approximate other measures besides the Kullback–Leibler measure. Ekström (1997) further expanded the method to investigate properties of estimators using higher order spacings, where an m-order spacing would be defined as .

Multivariate distributions

Ranneby & al. (2005) discuss extended maximum spacing methods to the multivariate case. As there is no natural order for , they discuss two alternative approaches: a geometric approach based on Dirichlet cells and a probabilistic approach based on a “nearest neighbor ball” metric.

See also

Notes

  1. There appear to be some minor typographical errors in the paper. For example, in section 4.2, equation (4.1), the rounding replacement for , should not have the log term. In section 1, equation (1.2), is defined to be the spacing itself, and is the negative sum of the logs of . If is logged at this step, the result is always  0, as the difference between two adjacent points on a cumulative distribution is always ≤ 1, and strictly < 1 unless there are only two points at the bookends. Also, in section 4.3, on page 392, calculation shows that it is the variance which has MPS estimate of 6.87, not the standard deviation . – Editor
  2. The literature refers to related statistics as Moran or Moran-Darling statistics. For example, Cheng & Stephens (1989) analyze the form where is defined as above. Wong & Li (2006) use the same form as well. However, Beirlant & al. (2001) uses the form , with the additional factor of inside the logged summation. The extra factors will make a difference in terms of the expected mean and variance of the statistic. For consistency, this article will continue to use the Cheng & Amin/Wong & Li form. -- Editor
  3. Wong & Li (2006) leave out the Euler–Mascheroni constant from their description. -- Editor

Related Research Articles

The likelihood function is the joint probability of observed data viewed as a function of the parameters of a statistical model.

In statistics, maximum likelihood estimation (MLE) is a method of estimating the parameters of an assumed probability distribution, given some observed data. This is achieved by maximizing a likelihood function so that, under the assumed statistical model, the observed data is most probable. The point in the parameter space that maximizes the likelihood function is called the maximum likelihood estimate. The logic of maximum likelihood is both intuitive and flexible, and as such the method has become a dominant means of statistical inference.

<span class="mw-page-title-main">Gamma distribution</span> Probability distribution

In probability theory and statistics, the gamma distribution is a two-parameter family of continuous probability distributions. The exponential distribution, Erlang distribution, and chi-squared distribution are special cases of the gamma distribution. There are two equivalent parameterizations in common use:

  1. With a shape parameter k and a scale parameter θ
  2. With a shape parameter and an inverse scale parameter , called a rate parameter.
<span class="mw-page-title-main">Cramér–Rao bound</span> Lower bound on variance of an estimator

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

In mathematical statistics, the Fisher information is a way of measuring the amount of information that an observable random variable X carries about an unknown parameter θ of a distribution that models X. Formally, it is the variance of the score, or the expected value of the observed information.

Estimation theory is a branch of statistics that deals with estimating the values of parameters based on measured empirical data that has a random component. The parameters describe an underlying physical setting in such a way that their value affects the distribution of the measured data. An estimator attempts to approximate the unknown parameters using the measurements. In estimation theory, two approaches are generally considered:

In Bayesian statistics, a maximum a posteriori probability (MAP) estimate is an estimate of an unknown quantity, that equals the mode of the posterior distribution. The MAP can be used to obtain a point estimate of an unobserved quantity on the basis of empirical data. It is closely related to the method of maximum likelihood (ML) estimation, but employs an augmented optimization objective which incorporates a prior distribution over the quantity one wants to estimate. MAP estimation can therefore be seen as a regularization of maximum likelihood estimation.

In econometrics and statistics, the generalized method of moments (GMM) is a generic method for estimating parameters in statistical models. Usually it is applied in the context of semiparametric models, where the parameter of interest is finite-dimensional, whereas the full shape of the data's distribution function may not be known, and therefore maximum likelihood estimation is not applicable.

In statistics, the method of moments is a method of estimation of population parameters. The same principle is used to derive higher moments like skewness and kurtosis.

In statistics, the Bayesian information criterion (BIC) or Schwarz information criterion is a criterion for model selection among a finite set of models; models with lower BIC are generally preferred. It is based, in part, on the likelihood function and it is closely related to the Akaike information criterion (AIC).

The James–Stein estimator is a biased estimator of the mean, , of (possibly) correlated Gaussian distributed random variables with unknown means .

In statistics, M-estimators are a broad class of extremum estimators for which the objective function is a sample average. Both non-linear least squares and maximum likelihood estimation are special cases of M-estimators. The definition of M-estimators was motivated by robust statistics, which contributed new types of M-estimators. However, M-estimators are not inherently robust, as is clear from the fact that they include maximum likelihood estimators, which are in general not robust. The statistical procedure of evaluating an M-estimator on a data set is called M-estimation.

In estimation theory and decision theory, a Bayes estimator or a Bayes action is an estimator or decision rule that minimizes the posterior expected value of a loss function. Equivalently, it maximizes the posterior expectation of a utility function. An alternative way of formulating an estimator within Bayesian statistics is maximum a posteriori estimation.

In statistics, the bias of an estimator is the difference between this estimator's expected value and the true value of the parameter being estimated. An estimator or decision rule with zero bias is called unbiased. In statistics, "bias" is an objective property of an estimator. Bias is a distinct concept from consistency: consistent estimators converge in probability to the true value of the parameter, but may be biased or unbiased; see bias versus consistency for more.

Location estimation in wireless sensor networks is the problem of estimating the location of an object from a set of noisy measurements. These measurements are acquired in a distributed manner by a set of sensors.

Minimum-distance estimation (MDE) is a conceptual method for fitting a statistical model to data, usually the empirical distribution. Often-used estimators such as ordinary least squares can be thought of as special cases of minimum-distance estimation.

<span class="mw-page-title-main">Wrapped Cauchy distribution</span>

In probability theory and directional statistics, a wrapped Cauchy distribution is a wrapped probability distribution that results from the "wrapping" of the Cauchy distribution around the unit circle. The Cauchy distribution is sometimes known as a Lorentzian distribution, and the wrapped Cauchy distribution may sometimes be referred to as a wrapped Lorentzian distribution.

In probability theory and statistics, empirical likelihood (EL) is a nonparametric method for estimating the parameters of statistical models. It requires fewer assumptions about the error distribution while retaining some of the merits in likelihood-based inference. The estimation method requires that the data are independent and identically distributed (iid). It performs well even when the distribution is asymmetric or censored. EL methods can also handle constraints and prior information on parameters. Art Owen pioneered work in this area with his 1988 paper.

<span class="mw-page-title-main">Hermite distribution</span> Statistical probability Distribution for discrete event counts

In probability theory and statistics, the Hermite distribution, named after Charles Hermite, is a discrete probability distribution used to model count data with more than one parameter. This distribution is flexible in terms of its ability to allow a moderate over-dispersion in the data.

In statistics, the variance function is a smooth function that depicts the variance of a random quantity as a function of its mean. The variance function is a measure of heteroscedasticity and plays a large role in many settings of statistical modelling. It is a main ingredient in the generalized linear model framework and a tool used in non-parametric regression, semiparametric regression and functional data analysis. In parametric modeling, variance functions take on a parametric form and explicitly describe the relationship between the variance and the mean of a random quantity. In a non-parametric setting, the variance function is assumed to be a smooth function.

References

Citations

Works cited

  • Anatolyev, Stanislav; Kosenok, Grigory (2005). "An alternative to maximum likelihood based on spacings" (PDF). Econometric Theory. 21 (2): 472–476. CiteSeerX   10.1.1.494.7340 . doi:10.1017/S0266466605050255. S2CID   123004317 . Retrieved 2009-01-21.
  • Beirlant, J.; Dudewicz, E.J.; Györfi, L.; van der Meulen, E.C. (1997). "Nonparametric entropy estimation: an overview" (PDF). International Journal of Mathematical and Statistical Sciences. 6 (1): 17–40. ISSN   1055-7490. Archived from the original (PDF) on May 5, 2005. Retrieved 2008-12-31.Note: linked paper is an updated 2001 version.
  • Cheng, R.C.H.; Amin, N.A.K. (1983). "Estimating parameters in continuous univariate distributions with a shifted origin". Journal of the Royal Statistical Society, Series B. 45 (3): 394–403. doi:10.1111/j.2517-6161.1983.tb01268.x. ISSN   0035-9246. JSTOR   2345411.
  • Cheng, R.C.H; Stephens, M. A. (1989). "A goodness-of-fit test using Moran's statistic with estimated parameters". Biometrika. 76 (2): 386–392. doi:10.1093/biomet/76.2.385.
  • Ekström, Magnus (1997). "Generalized maximum spacing estimates". University of Umeå, Department of Mathematics. 6. ISSN   0345-3928. Archived from the original on February 14, 2007. Retrieved 2008-12-30.
  • Hall, M.J.; van den Boogaard, H.F.P.; Fernando, R.C.; Mynett, A.E. (2004). "The construction of confidence intervals for frequency analysis using resampling techniques". Hydrology and Earth System Sciences. 8 (2): 235–246. doi: 10.5194/hess-8-235-2004 . ISSN   1027-5606.
  • Pieciak, Tomasz (2014). The maximum spacing noise estimation in single-coil background MRI data (PDF). IEEE International Conference on Image Processing. Paris. pp. 1743–1747. Retrieved 2015-07-07.
  • Pyke, Ronald (1965). "Spacings". Journal of the Royal Statistical Society, Series B. 27 (3): 395–449. doi:10.1111/j.2517-6161.1965.tb00602.x. ISSN   0035-9246. JSTOR   2345793.
  • Ranneby, Bo (1984). "The maximum spacing method. An estimation method related to the maximum likelihood method". Scandinavian Journal of Statistics. 11 (2): 93–112. ISSN   0303-6898. JSTOR   4615946.
  • Ranneby, Bo; Ekström, Magnus (1997). "Maximum spacing estimates based on different metrics". University of Umeå, Department of Mathematics. 5. ISSN   0345-3928. Archived from the original on February 14, 2007. Retrieved 2008-12-30.
  • Ranneby, Bo; Jammalamadakab, S. Rao; Teterukovskiy, Alex (2005). "The maximum spacing estimation for multivariate observations" (PDF). Journal of Statistical Planning and Inference. 129 (1–2): 427–446. doi:10.1016/j.jspi.2004.06.059 . Retrieved 2008-12-31.
  • Wong, T.S.T; Li, W.K. (2006). "A note on the estimation of extreme value distributions using maximum product of spacings". Time series and related topics: in memory of Ching-Zong Wei. Institute of Mathematical Statistics Lecture Notes – Monograph Series. Beachwood, Ohio: Institute of Mathematical Statistic. pp. 272–283. arXiv: math/0702830v1 . doi:10.1214/074921706000001102. ISBN   978-0-940600-68-3. S2CID   88516426.