ApexMD

Power law - Wikipedia


Scale invarianceEdit

One attribute of power laws is their scale invariance. Given a relation f(x)=ax−k{\displaystyle f(x)=ax^{-k}} , scaling the argument x{\displaystyle x} by a constant factor c{\displaystyle c} causes only a proportionate scaling of the function itself. That is,

: f(cx)=a(cx)−k=c−kf(x)∝f(x),{\displaystyle f(cx)=a(cx)^{-k}=c^{-k}f(x)\propto f(x),!}

where ∝{\displaystyle \propto } denotes direct proportionality. That is, scaling by a constant c{\displaystyle c} simply multiplies the original power-law relation by the constant c−k{\displaystyle c^{-k}} . Thus, it follows that all power laws with a particular scaling exponent are equivalent up to constant factors, since each is simply a scaled version of the others. This behavior is what produces the linear relationship when logarithms are taken of both f(x){\displaystyle f(x)} and x{\displaystyle x} , and the straight-line on the log–log plot is often called the signature of a power law. With real data, such straightness is a necessary, but not sufficient, condition for the data following a power-law relation. In fact, there are many ways to generate finite amounts of data that mimic this signature behavior, but, in their asymptotic limit, are not true power laws (e.g., if the generating process of some data follows a Log-normal distribution).[[citation needed](https://en.m.wikipedia.org/wiki/Wikipedia:Citationneeded "Wikipedia:Citation needed")_] Thus, accurately fitting and validating power-law models is an active area of research in statistics; see below.

Lack of well-defined average valueEdit

A power-law x−k{\displaystyle x^{-k}} has a well-defined mean over x∈[1,∞){\displaystyle x\in [1,\infty )} only if k>2{\displaystyle k>2}2}" data-class="mwe-math-fallback-image-inline"> , and it has a finite variance only if k>3{\displaystyle k>3}3}" data-class="mwe-math-fallback-image-inline"> ; most identified power laws in nature have exponents such that the mean is well-defined but the variance is not, implying they are capable of black swan behavior.[2] This can be seen in the following thought experiment:[10] imagine a room with your friends and estimate the average monthly income in the room. Now imagine the world's richest person entering the room, with a monthly income of about 1 billion US$. What happens to the average income in the room? Income is distributed according to a power-law known as the Pareto distribution (for example, the net worth of Americans is distributed according to a power law with an exponent of 2).

On the one hand, this makes it incorrect to apply traditional statistics that are based on variance and standard deviation (such as regression analysis).[[citation needed](https://en.m.wikipedia.org/wiki/Wikipedia:Citationneeded "Wikipedia:Citation needed")] On the other hand, this also allows for cost-efficient interventions.[[10]](#citenote-CCSSCS9-10) For example, given that car exhaust is distributed according to a power-law among cars (very few cars contribute to most contamination) it would be sufficient to eliminate those very few cars from the road to reduce total exhaust substantially.[11]

The median does exist, however: for a power law xk, with exponent k>1{\displaystyle k>1} 1" data-class="mwe-math-fallback-image-inline"> , it takes the value 21/(k – 1)xmin, where xmin is the minimum value for which the power law holds.[12]

UniversalityEdit

The equivalence of power laws with a particular scaling exponent can have a deeper origin in the dynamical processes that generate the power-law relation. In physics, for example, phase transitions in thermodynamic systems are associated with the emergence of power-law distributions of certain quantities, whose exponents are referred to as the critical exponents of the system. Diverse systems with the same critical exponents—that is, which display identical scaling behaviour as they approach criticality—can be shown, via renormalization group theory, to share the same fundamental dynamics. For instance, the behavior of water and CO2 at their boiling points fall in the same universality class because they have identical critical exponents.[[citation needed](https://en.m.wikipedia.org/wiki/Wikipedia:Citationneeded "Wikipedia:Citation needed")][[clarification needed](https://en.m.wikipedia.org/wiki/Wikipedia:Pleaseclarify "Wikipedia:Please clarify")] In fact, almost all material phase transitions are described by a small set of universality classes. Similar observations have been made, though not as comprehensively, for various self-organized critical systems, where the critical point of the system is an attractor. Formally, this sharing of dynamics is referred to as universality, and systems with precisely the same critical exponents are said to belong to the same universality class.

Scientific interest in power-law relations stems partly from the ease with which certain general classes of mechanisms generate them. The demonstration of a power-law relation in some data can point to specific kinds of mechanisms that might underlie the natural phenomenon in question, and can indicate a deep connection with other, seemingly unrelated systems; see also universality above. The ubiquity of power-law relations in physics is partly due to dimensional constraints, while in complex systems, power laws are often thought to be signatures of hierarchy or of specific stochastic processes. A few notable examples of power laws are Pareto's law of income distribution, structural self-similarity of fractals, and scaling laws in biological systems. Research on the origins of power-law relations, and efforts to observe and validate them in the real world, is an active topic of research in many fields of science, including physics, computer science, linguistics, geophysics, neuroscience, sociology, economics and more.

However, much of the recent interest in power laws comes from the study of probability distributions: The distributions of a wide variety of quantities seem to follow the power-law form, at least in their upper tail (large events). The behavior of these large events connects these quantities to the study of theory of large deviations (also called extreme value theory), which considers the frequency of extremely rare events like stock market crashes and large natural disasters. It is primarily in the study of statistical distributions that the name "power law" is used.

In empirical contexts, an approximation to a power-law o(xk){\displaystyle o(x^{k})} often includes a deviation term ε{\displaystyle \varepsilon } , which can represent uncertainty in the observed values (perhaps measurement or sampling errors) or provide a simple way for observations to deviate from the power-law function (perhaps for stochastic reasons):

: y=axk+ε.{\displaystyle y=ax^{k}+\varepsilon .!}

Mathematically, a strict power law cannot be a probability distribution, but a distribution that is a truncated power function is possible: p(x)=Cx−α{\displaystyle p(x)=Cx^{-\alpha }} for x>xmin{\displaystyle x>x{\text{min}}} x\text{min}" data-class="mwe-math-fallback-image-inline"> where the exponent α{\displaystyle \alpha } (Greek letter alpha, not to be confused with scaling factor a{\displaystyle a} used above) is greater than 1 (otherwise the tail has infinite area), the minimum value xmin{\displaystyle x{\text{min}}} is needed otherwise the distribution has infinite area as _x approaches 0, and the constant C is a scaling factor to ensure that the total area is 1, as required by a probability distribution. More often one uses an asymptotic power law – one that is only true in the limit; see power-law probability distributions below for details. Typically the exponent falls in the range 2<α<3{\displaystyle 2<\alpha <3} , though not always.

ExamplesEdit

More than a hundred power-law distributions have been identified in physics (e.g. sandpile avalanches), biology (e.g. species extinction and body mass), and the social sciences (e.g. city sizes and income).[15] Among them are:

AstronomyEdit

CriminologyEdit

  • number of charges per criminal offender

[1]

PhysicsEdit

BiologyEdit

  • Kleiber's law relating animal metabolism to size, and allometric laws in general
  • The two-thirds power law, relating speed to curvature in the human motor system.
  • The Taylor's law relating mean population size and variance of populations sizes in ecology
  • Neuronal avalanches[4]
  • The species richness (number of species) in clades of freshwater fishes[17]
  • The Harlow Knapp effect, where a subset of the kinases found in the human body compose a majority of published research[18]

MeteorologyEdit

  • The size of rain-shower cells,[19] energy dissipation in cyclones,[20] and the diameters of dust devils on Earth and Mars [21]

General scienceEdit

MathematicsEdit

EconomicsEdit

  • Population sizes of cities in a region or urban network, Zipf's law.

  • Distribution of artists by the average price of their artworks.[28]

  • Distribution of income in a market economy.

  • Distribution of degrees in banking networks.

FinanceEdit

  • The mean absolute change of the logarithmic mid-prices[29]
  • Number of tick counts over time
  • Size of the maximum price move
  • Average waiting time of a directional change[30]
  • Average waiting time of an overshoot

VariantsEdit

Broken power lawEdit

Some models of the initial mass function use a broken power law; here Kroupa (2001) in red.

A broken power law is a piecewise function, consisting of two or more power laws, combined with a threshold. For example, with two power laws:[31]

: f(x)∝xα1{\displaystyle f(x)\propto x^{\alpha {1}}} for x<xth,{\displaystyle x<x{\text{th}},}: f(x)∝xthα1−α2xα2 for x>xth{\displaystyle f(x)\propto x{\text{th}}^{\alpha _{1}-\alpha _{2}}x^{\alpha _{2}}{\text{ for }}x>x{\text{th}}}x_\text{th}" data-class="mwe-math-fallback-image-inline"> .

Power law with exponential cutoffEdit

A power law with an exponential cutoff is simply a power law multiplied by an exponential function:[32]

: f(x)∝xαeβx.{\displaystyle f(x)\propto x^{\alpha }e^{\beta x}.}

Curved power lawEdit

: f(x)∝xα+βx{\displaystyle f(x)\propto x^{\alpha +\beta x}}[33]

In a looser sense, a power-law probability distribution is a distribution whose density function (or mass function in the discrete case) has the form, for large values of x{\displaystyle x} ,[34]

: P(X>x)∼L(x)x−(α−1){\displaystyle P(X>x)\sim L(x)x^{-(\alpha -1)}}x)\sim L(x)x^{-(\alpha -1)}}" data-class="mwe-math-fallback-image-inline">

where α>1{\displaystyle \alpha >1}1" data-class="mwe-math-fallback-image-inline"> , and L(x){\displaystyle L(x)} is a slowly varying function, which is any function that satisfies limx→∞L(rx)/L(x)=1{\displaystyle \lim {x\rightarrow \infty }L(r\,x)/L(x)=1} for any positive factor r{\displaystyle r} . This property of L(x){\displaystyle L(x)} follows directly from the requirement that p(x){\displaystyle p(x)} be asymptotically scale invariant; thus, the form of L(x){\displaystyle L(x)} only controls the shape and finite extent of the lower tail. For instance, if L(x){\displaystyle L(x)} is the constant function, then we have a power law that holds for all values of x{\displaystyle x} . In many cases, it is convenient to assume a lower bound xmin{\displaystyle x{\mathrm {min} }} from which the law holds. Combining these two cases, and where x{\displaystyle x} is a continuous variable, the power law has the form

: p(x)=α−1xmin(xxmin)−α,{\displaystyle p(x)={\frac {\alpha -1}{x{\min }}}\left({\frac {x}{x{\min }}}\right)^{-\alpha },}

where the pre-factor to α−1xmin{\displaystyle {\frac {\alpha -1}{x_{\min }}}} is the normalizing constant. We can now consider several properties of this distribution. For instance, its moments are given by

: ⟨xm⟩=∫xmin∞xmp(x)dx=α−1α−1−mxminm{\displaystyle \langle x^{m}\rangle =\int {x{\min }}^{\infty }x^{m}p(x)\,\mathrm {d} x={\frac {\alpha -1}{\alpha -1-m}}x_{\min }^{m}}

which is only well defined for m<α−1{\displaystyle m<\alpha -1} . That is, all moments m≥α−1{\displaystyle m\geq \alpha -1} diverge: when α≤2{\displaystyle \alpha \leq 2} , the average and all higher-order moments are infinite; when 2<α<3{\displaystyle 2<\alpha <3} , the mean exists, but the variance and higher-order moments are infinite, etc. For finite-size samples drawn from such distribution, this behavior implies that the central moment estimators (like the mean and the variance) for diverging moments will never converge – as more data is accumulated, they continue to grow. These power-law probability distributions are also called Pareto-type distributions, distributions with Pareto tails, or distributions with regularly varying tails.

A modification, which does not satisfy the general form above, with an exponential cutoff, is

: p(x)∝L(x)x−αe−λx.{\displaystyle p(x)\propto L(x)x^{-\alpha }\mathrm {e} ^{-\lambda x}.}

In this distribution, the exponential decay term e−λx{\displaystyle \mathrm {e} ^{-\lambda x}} eventually overwhelms the power-law behavior at very large values of x{\displaystyle x} . This distribution does not scale and is thus not asymptotically as a power law; however, it does approximately scale over a finite region before the cutoff. (Note that the pure form above is a subset of this family, with λ=0{\displaystyle \lambda =0} .) This distribution is a common alternative to the asymptotic power-law distribution because it naturally captures finite-size effects.

The Tweedie distributions are a family of statistical models characterized by closure under additive and reproductive convolution as well as under scale transformation. Consequently, these models all express a power-law relationship between the variance and the mean. These models have a fundamental role as foci of mathematical convergence similar to the role that the normal distribution has as a focus in the central limit theorem. This convergence effect explains why the variance-to-mean power law manifests so widely in natural processes, as with Taylor's law in ecology and with fluctuation scaling[35] in physics. It can also be shown that this variance-to-mean power law, when demonstrated by the method of expanding bins, implies the presence of 1/f noise and that 1/f noise can arise as a consequence of this Tweedie convergence effect.[36]

Graphical methods for identificationEdit

Although more sophisticated and robust methods have been proposed, the most frequently used graphical methods of identifying power-law probability distributions using random samples are Pareto quantile-quantile plots (or Pareto Q–Q plots),[[citation needed](https://en.m.wikipedia.org/wiki/Wikipedia:Citationneeded "Wikipedia:Citation needed")] mean residual life plots[[37]](#citenote-37)[38] and log–log plots. Another, more robust graphical method uses bundles of residual quantile functions.[39] (Please keep in mind that power-law distributions are also called Pareto-type distributions.) It is assumed here that a random sample is obtained from a probability distribution, and that we want to know if the tail of the distribution follows a power law (in other words, we want to know if the distribution has a "Pareto tail"). Here, the random sample is called "the data".

Pareto Q–Q plots compare the quantiles of the log-transformed data to the corresponding quantiles of an exponential distribution with mean 1 (or to the quantiles of a standard Pareto distribution) by plotting the former versus the latter. If the resultant scatterplot suggests that the plotted points " asymptotically converge" to a straight line, then a power-law distribution should be suspected. A limitation of Pareto Q–Q plots is that they behave poorly when the tail index α{\displaystyle \alpha } (also called Pareto index) is close to 0, because Pareto Q–Q plots are not designed to identify distributions with slowly varying tails.[39]

On the other hand, in its version for identifying power-law probability distributions, the mean residual life plot consists of first log-transforming the data, and then plotting the average of those log-transformed data that are higher than the i-th order statistic versus the i-th order statistic, for i = 1, ..., n, where n is the size of the random sample. If the resultant scatterplot suggests that the plotted points tend to "stabilize" about a horizontal straight line, then a power-law distribution should be suspected. Since the mean residual life plot is very sensitive to outliers (it is not robust), it usually produces plots that are difficult to interpret; for this reason, such plots are usually called Hill horror plots [40]

A straight line on a log–log plot is necessary but insufficient evidence for power-laws, the slope of the straight line corresponds to the power law exponent.

Log–log plots are an alternative way of graphically examining the tail of a distribution using a random sample. Caution has to be exercised however as a log–log plot is necessary but insufficient evidence for a power law relationship, as many non power-law distributions will appear as straight lines on a log–log plot.[41][42] This method consists of plotting the logarithm of an estimator of the probability that a particular number of the distribution occurs versus the logarithm of that particular number. Usually, this estimator is the proportion of times that the number occurs in the data set. If the points in the plot tend to "converge" to a straight line for large numbers in the x axis, then the researcher concludes that the distribution has a power-law tail. Examples of the application of these types of plot have been published.[43] A disadvantage of these plots is that, in order for them to provide reliable results, they require huge amounts of data. In addition, they are appropriate only for discrete (or grouped) data.

Another graphical method for the identification of power-law probability distributions using random samples has been proposed.[39] This methodology consists of plotting a bundle for the log-transformed sample. Originally proposed as a tool to explore the existence of moments and the moment generation function using random samples, the bundle methodology is based on residual quantile functions (RQFs), also called residual percentile functions,[44][45][46][47][48][49][50] which provide a full characterization of the tail behavior of many well-known probability distributions, including power-law distributions, distributions with other types of heavy tails, and even non-heavy-tailed distributions. Bundle plots do not have the disadvantages of Pareto Q–Q plots, mean residual life plots and log–log plots mentioned above (they are robust to outliers, allow visually identifying power laws with small values of α{\displaystyle \alpha } , and do not demand the collection of much data).[[citation needed](https://en.m.wikipedia.org/wiki/Wikipedia:Citationneeded "Wikipedia:Citation needed")_] In addition, other types of tail behavior can be identified using bundle plots.

Plotting power-law distributionsEdit

In general, power-law distributions are plotted on doubly logarithmic axes, which emphasizes the upper tail region. The most convenient way to do this is via the (complementary) cumulative distribution (cdf), P(x)=Pr(X>x){\displaystyle P(x)=\mathrm {Pr} (X>x)} x)" data-class="mwe-math-fallback-image-inline"> ,

: P(x)=Pr(X>x)=C∫x∞p(X)dX=α−1xmin−α+1∫x∞X−αdX=(xxmin)−α+1.{\displaystyle P(x)=\Pr(X>x)=C\int {x}^{\infty }p(X)\,\mathrm {d} X={\frac {\alpha -1}{x{\min }^{-\alpha +1}}}\int {x}^{\infty }X^{-\alpha }\,\mathrm {d} X=\left({\frac {x}{x{\min }}}\right)^{-\alpha +1}.} x) = C \intx^\infty p(X)\,\mathrm{d}X = \frac{\alpha-1}{x\min^{-\alpha+1}} \intx^\infty X^{-\alpha}\,\mathrm{d}X = \left(\frac{x}{x\min} \right)^{-\alpha+1}." data-class="mwe-math-fallback-image-inline">

Note that the cdf is also a power-law function, but with a smaller scaling exponent. For data, an equivalent form of the cdf is the rank-frequency approach, in which we first sort the n{\displaystyle n} observed values in ascending order, and plot them against the vector [1,n−1n,n−2n,…,1n]{\displaystyle \left[1,{\frac {n-1}{n}},{\frac {n-2}{n}},\dots ,{\frac {1}{n}}\right]} .

Although it can be convenient to log-bin the data, or otherwise smooth the probability density (mass) function directly, these methods introduce an implicit bias in the representation of the data, and thus should be avoided.[51][52] The cdf, on the other hand, is more robust to (but not without) such biases in the data and preserves the linear signature on doubly logarithmic axes. Though a cdf representation is favored over that of the pdf while fitting a power law to the data with the linear least square method, it is not devoid of mathematical inaccuracy. Thus, while estimating exponents of a power law distribution, maximum likelihood estimator is recommended.

Estimating the exponent from empirical dataEdit

There are many ways of estimating the value of the scaling exponent for a power-law tail, however not all of them yield unbiased and consistent answers. Some of the most reliable techniques are often based on the method of maximum likelihood. Alternative methods are often based on making a linear regression on either the log–log probability, the log–log cumulative distribution function, or on log-binned data, but these approaches should be avoided as they can all lead to highly biased estimates of the scaling exponent.

Maximum likelihoodEdit

For real-valued, independent and identically distributed data, we fit a power-law distribution of the form

: p(x)=α−1xmin(xxmin)−α{\displaystyle p(x)={\frac {\alpha -1}{x{\min }}}\left({\frac {x}{x{\min }}}\right)^{-\alpha }}

to the data x≥xmin{\displaystyle x\geq x{\min }} , where the coefficient α−1xmin{\displaystyle {\frac {\alpha -1}{x{\min }}}} is included to ensure that the distribution is normalized. Given a choice for xmin{\displaystyle x_{\min }} , the log likelihood function becomes:

: L(α)=log⁡∏i=1nα−1xmin(xixmin)−α{\displaystyle {\mathcal {L}}(\alpha )=\log \prod {i=1}^{n}{\frac {\alpha -1}{x{\min }}}\left({\frac {x{i}}{x{\min }}}\right)^{-\alpha }}

The maximum of this likelihood is found by differentiating with respect to parameter α{\displaystyle \alpha } , setting the result equal to zero. Upon rearrangement, this yields the estimator equation:

: α^=1+n[∑i=1nln⁡xixmin]−1{\displaystyle {\hat {\alpha }}=1+n\left[\sum {i=1}^{n}\ln {\frac {x{i}}{x_{\min }}}\right]^{-1}}

where {xi}{\displaystyle {x{i}}} are the n{\displaystyle n} data points xi≥xmin{\displaystyle x{i}\geq x{\min }} .[[2]](#citenote-Newman-2)[53] This estimator exhibits a small finite sample-size bias of order O(n−1){\displaystyle O(n^{-1})} , which is small when n > 100. Further, the standard error of the estimate is σ=α^−1n+O(n−1){\displaystyle \sigma ={\frac {{\hat {\alpha }}-1}{\sqrt {n}}}+O(n^{-1})} . This estimator is equivalent to the popular[[citation needed](https://en.m.wikipedia.org/wiki/Wikipedia:Citationneeded "Wikipedia:Citation needed")]Hill estimator from quantitative finance and extreme value theory.[[citation needed](https://en.m.wikipedia.org/wiki/Wikipedia:Citationneeded "Wikipedia:Citation needed")]

For a set of n integer-valued data points {xi}{\displaystyle {x{i}}} , again where each xi≥xmin{\displaystyle x{i}\geq x_{\min }} , the maximum likelihood exponent is the solution to the transcendental equation

: ζ′(α^,xmin)ζ(α^,xmin)=−1n∑i=1nln⁡xixmin{\displaystyle {\frac {\zeta '({\hat {\alpha }},x{\min })}{\zeta ({\hat {\alpha }},x{\min })}}=-{\frac {1}{n}}\sum {i=1}^{n}\ln {\frac {x{i}}{x_{\min }}}}

where ζ(α,xmin){\displaystyle \zeta (\alpha ,x_{\mathrm {min} })} is the incomplete zeta function. The uncertainty in this estimate follows the same formula as for the continuous equation. However, the two equations for α^{\displaystyle {\hat {\alpha }}} are not equivalent, and the continuous version should not be applied to discrete data, nor vice versa.

Further, both of these estimators require the choice of xmin{\displaystyle x{\min }} . For functions with a non-trivial L(x){\displaystyle L(x)} function, choosing xmin{\displaystyle x{\min }} too small produces a significant bias in α^{\displaystyle {\hat {\alpha }}} , while choosing it too large increases the uncertainty in α^{\displaystyle {\hat {\alpha }}} , and reduces the statistical power of our model. In general, the best choice of xmin{\displaystyle x_{\min }} depends strongly on the particular form of the lower tail, represented by L(x){\displaystyle L(x)} above.

More about these methods, and the conditions under which they can be used, can be found in . Further, this comprehensive review article provides usable code (Matlab, Python, R and C++) for estimation and testing routines for power-law distributions.

Kolmogorov–Smirnov estimationEdit

Another method for the estimation of the power-law exponent, which does not assume independent and identically distributed (iid) data, uses the minimization of the Kolmogorov–Smirnov statistic, D{\displaystyle D} , between the cumulative distribution functions of the data and the power law:

: α^=argminαDα{\displaystyle {\hat {\alpha }}={\underset {\alpha }{\operatorname {arg\,min} }}\,D_{\alpha }}

with

: Dα=maxx|Pemp(x)−Pα(x)|{\displaystyle D{\alpha }=\max _{x}|P{\mathrm {emp} }(x)-P_{\alpha }(x)|}

where Pemp(x){\displaystyle P{\mathrm {emp} }(x)} and Pα(x){\displaystyle P{\alpha }(x)} denote the cdfs of the data and the power law with exponent α{\displaystyle \alpha } , respectively. As this method does not assume iid data, it provides an alternative way to determine the power-law exponent for data sets in which the temporal correlation can not be ignored.[4]

Two-point fitting methodEdit

This criterion[[clarification needed](https://en.m.wikipedia.org/wiki/Wikipedia:Pleaseclarify "Wikipedia:Please clarify")] can be applied for the estimation of power-law exponent in the case of scale free distributions and provides a more convergent estimate than the maximum likelihood method.[[citation needed](https://en.m.wikipedia.org/wiki/Wikipedia:Citationneeded "Wikipedia:Citation needed")] It has been applied to study probability distributions of fracture apertures.[[citation needed](https://en.m.wikipedia.org/wiki/Wikipedia:Citationneeded "Wikipedia:Citation needed")] In some contexts the probability distribution is described, not by the cumulative distribution function, by the cumulative frequency of a property _X, defined as the number of elements per meter (or area unit, second etc.) for which X > x applies, where x is a variable real number. As an example,[[citation needed](https://en.m.wikipedia.org/wiki/Wikipedia:Citationneeded "Wikipedia:Citation needed")] the cumulative distribution of the fracture aperture, _X, for a sample of N elements is defined as 'the number of fractures per meter having aperture greater than x . Use of cumulative frequency has some advantages, e.g. it allows one to put on the same diagram data gathered from sample lines of different lengths at different scales (e.g. from outcrop and from microscope).

Notes

  1. [^](#cite_ref-1)Yaneer Bar-Yam. "Concepts: Power Law". New England Complex Systems Institute. Retrieved 18 August 2015.
  2. ^ abcNewman, M. E. J. (2005). "Power laws, Pareto distributions and Zipf's law". [Contemporary Physics](https://en.m.wikipedia.org/wiki/ContemporaryPhysics "Contemporary Physics"). 46 (5): 323–351. arXiv:cond-mat/0412004. Bibcode:2005ConPh..46..323N. doi:10.1080/00107510500052444._
  3. [^](#citeref-Humphries3-0)Humphries NE, Queiroz N, Dyer JR, Pade NG, Musyl MK, Schaefer KM, Fuller DW, Brunnschweiler JM, Doyle TK, Houghton JD, Hays GC, Jones CS, Noble LR, Wearmouth VJ, Southall EJ, Sims DW (2010). "Environmental context explains Lévy and Brownian movement patterns of marine predators" (PDF). _Nature. 465 (7301): 1066–1069. Bibcode:2010Natur.465.1066H. doi:10.1038/nature09116. PMID20531470._
  4. ^ abcKlaus A, Yu S, Plenz D (2011). Zochowski M (ed.). "Statistical Analyses Support Power Law Distributions Found in Neuronal Avalanches". _PLoS ONE. 6 (5): e19779. Bibcode:2011PLoSO...619779K. doi:10.1371/journal.pone.0019779. PMC3102672. PMID21720544._
  5. [^](#cite_ref-5)Albert, J. S.; Reis, R. E., eds. (2011). Historical Biogeography of Neotropical Freshwater Fishes. Berkeley: University of California Press.
  6. [^](#cite_ref-6)Cannavò, Flavio; Nunnari, Giuseppe (2016-03-01). "On a Possible Unified Scaling Law for Volcanic Eruption Durations". _Scientific Reports. 6: 22289. Bibcode:2016NatSR...622289C. doi:10.1038/srep22289. ISSN2045-2322. PMC4772095. PMID26926425._
  7. [^](#cite_ref-7)Stevens, S. S. (1957). "On the psychophysical law". _Psychological Review. 64: 153–181. doi:10.1037/h0046162._
  8. [^](#cite_ref-8)Staddon, J. E. R. (1978). "Theory of behavioral power functions". _Psychological Review. 85: 305–320. doi:10.1037/0033-295x.85.4.305. hdl:10161/6003._
  9. ^ ab 9na CEPAL Charlas Sobre Sistemas Complejos Sociales (CCSSCS): Leyes de potencias, https://www.youtube.com/watch?v=4uDSEs86xCI
  10. [^](#cite_ref-11) Malcolm Gladwell (2006), Million-Dollar Murray; "Archived copy". Archived from the original on 2015-03-18. Retrieved 2015-06-14.CS1 maint: archived copy as title (link)
  11. [^](#cite_ref-12)Newman, Mark EJ. "Power laws, Pareto distributions and Zipf's law." Contemporary physics 46.5 (2005): 323-351.
  12. [^](#cite_ref-15)Andriani, P.; McKelvey, B. (2007). "Beyond Gaussian averages: redirecting international business and management research toward extreme events and power laws". _Journal of International Business Studies. 38 (7): 1212–1230. doi:10.1057/palgrave.jibs.8400324._
  13. [^](#cite_ref-16)Bolmatov, D.; Brazhkin, V. V.; Trachenko, K. (2013). "Thermodynamic behaviour of supercritical matter". _Nature Communications. 4: 2331. arXiv:1303.3153. Bibcode:2013NatCo...4.2331B. doi:10.1038/ncomms3331. PMID23949085._
  14. [^](#cite_ref-17)Albert, J. S., H. J. Bart, & R. E. Reis (2011). "Species richness & cladal diversity". In Albert, J. S., & R. E. Reis (ed.). _Historical Biogeography of Neotropical Freshwater Fishes. Berkeley: University of California Press. pp. 89–104._CS1 maint: multiple names: authors list (link)
  15. [^](#cite_ref-18)Yu, Frank H.; Willson, Timothy; Frye, Stephen; Edwards, Aled; Bader, Gary D.; Isserlin, Ruth (2011-02-02). "The human genome and drug discovery after a decade. Roads (still) not taken". _Nature. 470 (7333): 163–5. arXiv:1102.0448v2. Bibcode:2011Natur.470..163E. doi:10.1038/470163a. PMID21307913._
  16. [^](#citeref-Machado19-0)Machado L, Rossow, WB (1993). "Structural characteristics and radial properties of tropical cloud clusters". _Monthly Weather Review. 121 (12): 3234–3260. doi:10.1175/1520-0493(1993)1213234:scarpo2.0.co;2._
  17. [^](#citeref-Corral20-0)Corral, A, Osso, A, Llebot, JE (2010). "Scaling of tropical cyclone dissipation". _Nature Physics. 6 (9): 693–696. arXiv:0910.0054. Bibcode:2010NatPh...6..693C. doi:10.1038/nphys1725._
  18. [^](#citeref-Lorenz21-0)Lorenz RD (2009). "Power Law of Dust Devil Diameters on Earth and Mars". _Icarus. 203 (2): 683–684. Bibcode:2009Icar..203..683L. doi:10.1016/j.icarus.2009.06.029._
  19. [^](#citeref-ReedHughes22-0)Reed, W. J.; Hughes, B. D. (2002). "From gene families and genera to incomes and internet file sizes: Why power laws are so common in nature" (PDF). _Phys Rev E. 66: 067103. doi:10.1103/physreve.66.067103._
  20. ^ abHilbert, Martin (2013). "Scale-free power-laws as interaction between progress and diffusion". _Complexity(Submitted manuscript). 19 (4): 56–65. Bibcode:2014Cmplx..19d..56H. doi:10.1002/cplx.21485._
  21. [^](#cite_ref-24)["Horton's Laws – Example"](http://www.engr.colostate.edu/~ramirez/ceold/classes/cive322-Ramirez/CE322Web/ExampleHortonhtml.htm). _www.engr.colostate.edu. Retrieved 2018-09-30._
  22. [^](#cite_ref-25)Li, W. (November 1999). "Random texts exhibit Zipf's-law-like word frequency distribution". _IEEE Transactions on Information Theory. 38 (6): 1842–1845. doi:10.1109/18.165464. ISSN0018-9448._
  23. [^](#cite_ref-26)Lewis Fry Richardson (1950). _The Statistics of Deadly Quarrels._
  24. [^](#cite_ref-27)Martin, Charles H.; Mahoney, Michael W. (2018-10-02). "Implicit Self-Regularization in Deep Neural Networks: Evidence from Random Matrix Theory and Implications for Learning". arXiv:1810.01075 [cs.LG].
  25. [^](#cite_ref-28)Etro, F.; Stepanova, E. (2018). "Power-laws in art". _Physica A: Statistical Mechanics and Its Applications. 506: 217–220. Bibcode:2018PhyA..506..217E. doi:10.1016/j.physa.2018.04.057._
  26. [^](#cite_ref-29)Müller, Ulrich A.; Dacorogna, Michel M.; Olsen, Richard B.; Pictet, Olivier V.; Schwarz, Matthias; Morgenegg, Claude (1990-12-01). "Statistical study of foreign exchange rates, empirical evidence of a price change scaling law, and intraday analysis". _Journal of Banking & Finance. 14 (6): 1189–1208. doi:10.1016/0378-4266(90)90009-Q. ISSN0378-4266._
  27. [^](#cite_ref-30)Glattfelder, J. B.; Dupuis, A.; Olsen, R. B. (2011-04-01). "Patterns in high-frequency FX data: discovery of 12 empirical scaling laws". _Quantitative Finance. 11 (4): 599–614. arXiv:0809.1040. doi:10.1080/14697688.2010.481632. ISSN1469-7688._
  28. [^](#cite_ref-31)Jóhannesson, Gudlaugur; Björnsson, Gunnlaugur; Gudmundsson, Einar H. (2006). "Afterglow Light Curves and Broken Power Laws: A Statistical Study". _The Astrophysical Journal. 640 (1): L5. arXiv:astro-ph/0602219. Bibcode:2006ApJ...640L...5J. doi:10.1086/503294._
  29. [^](#cite_ref-32)Clauset, Aaron (2009). "Power-Law Distributions in Empirical Data". _SIAM Review. 51 (4): 661–703. arXiv:0706.1062. Bibcode:2009SIAMR..51..661C. doi:10.1137/070710111._
  30. [^](#cite_ref-33)"Curved-power law". Archived from the original on 2016-02-08. Retrieved 2013-07-07.
  31. [^](#cite_ref-34) N. H. Bingham, C. M. Goldie, and J. L. Teugels, Regular variation. Cambridge University Press, 1989
  32. [^](#citeref-Kendal2011a35-0)Kendal, WS; Jørgensen, B (2011). "Taylor's power law and fluctuation scaling explained by a central-limit-like convergence". _Phys. Rev. E. 83 (6): 066115. Bibcode:2011PhRvE..83f6115K. doi:10.1103/physreve.83.066115. PMID21797449._
  33. [^](#citeref-Kendal2011b36-0)Kendal, WS; Jørgensen, BR (2011). "Tweedie convergence: a mathematical basis for Taylor's power law, 1/f noise and multifractality" (PDF). _Phys. Rev. E. 84 (6): 066120. Bibcode:2011PhRvE..84f6120K. doi:10.1103/physreve.84.066120. PMID22304168._
  34. [^](#cite_ref-37) Beirlant, J., Teugels, J. L., Vynckier, P. (1996a) Practical Analysis of Extreme Values, Leuven: Leuven University Press
  35. [^](#cite_ref-38) Coles, S. (2001) An introduction to statistical modeling of extreme values. Springer-Verlag, London.
  36. ^ abcdDiaz, F. J. (1999). "Identifying Tail Behavior by Means of Residual Quantile Functions". _Journal of Computational and Graphical Statistics. 8 (3): 493–509. doi:10.2307/1390871. JSTOR1390871._
  37. [^](#cite_ref-40)Resnick, S. I. (1997). "Heavy Tail Modeling and Teletraffic Data". _The Annals of Statistics. 25 (5): 1805–1869. doi:10.1214/aos/1069362376._
  38. [^](#cite_ref-41)"So You Think You Have a Power Law — Well Isn't That Special?". _bactra.org. Retrieved 27 March 2018._
  39. [^](#cite_ref-42)Clauset, Aaron; Shalizi, Cosma Rohilla; Newman, M. E. J. (4 November 2009). "Power-law distributions in empirical data". _SIAM Review. 51 (4): 661–703. arXiv:0706.1062. Bibcode:2009SIAMR..51..661C. doi:10.1137/070710111._
  40. [^](#cite_ref-43)Jeong, H; Tombor, B. Albert; Oltvai, Z.N.; Barabasi, A.-L. (2000). "The large-scale organization of metabolic networks". _Nature. 407 (6804): 651–654. arXiv:cond-mat/0010278. Bibcode:2000Natur.407..651J. doi:10.1038/35036627. PMID11034217._
  41. [^](#cite_ref-44)Arnold, B. C.; Brockett, P. L. (1983). "When does the βth percentile residual life function determine the distribution?". _Operations Research. 31 (2): 391–396. doi:10.1287/opre.31.2.391._
  42. [^](#cite_ref-45)Joe, H.; Proschan, F. (1984). "Percentile residual life functions". _Operations Research. 32 (3): 668–678. doi:10.1287/opre.32.3.668._
  43. [^](#cite_ref-46) Joe, H. (1985), "Characterizations of life distributions from percentile residual lifetimes", Ann. Inst. Statist. Math. 37, Part A, 165–172.
  44. [^](#cite_ref-47)Csorgo, S.; Viharos, L. (1992). "Confidence bands for percentile residual lifetimes" (PDF). _Journal of Statistical Planning and Inference. 30 (3): 327–337. doi:10.1016/0378-3758(92)90159-p. hdl:2027.42/30190._
  45. [^](#cite_ref-48)Schmittlein, D. C.; Morrison, D. G. (1981). "The median residual lifetime: A characterization theorem and an application". _Operations Research. 29 (2): 392–399. doi:10.1287/opre.29.2.392._
  46. [^](#cite_ref-49)Morrison, D. G.; Schmittlein, D. C. (1980). "Jobs, strikes, and wars: Probability models for duration". _Organizational Behavior and Human Performance. 25 (2): 224–251. doi:10.1016/0030-5073(80)90065-3._
  47. [^](#cite_ref-50)Gerchak, Y (1984). "Decreasing failure rates and related issues in the social sciences". _Operations Research. 32 (3): 537–546. doi:10.1287/opre.32.3.537._
  48. [^](#cite_ref-51)Bauke, H. (2007). "Parameter estimation for power-law distributions by maximum likelihood methods". _European Physical Journal B. 58 (2): 167–173. arXiv:0704.1867. Bibcode:2007EPJB...58..167B. doi:10.1140/epjb/e2007-00219-y._
  49. [^](#cite_ref-52)Clauset, A., Shalizi, C. R., Newman, M. E. J. (2009). "Power-Law Distributions in Empirical Data". _SIAM Review. 51 (4): 661–703. arXiv:0706.1062. Bibcode:2009SIAMR..51..661C. doi:10.1137/070710111._CS1 maint: multiple names: authors list (link)
  50. [^](#citeref-Hall53-0)Hall, P. (1982). "On Some Simple Estimates of an Exponent of Regular Variation". [Journal of the Royal Statistical Society, Series B](https://en.m.wikipedia.org/wiki/JournaloftheRoyalStatisticalSociety,SeriesB "Journal of the Royal Statistical Society, Series B"). 44 (1): 37–42. JSTOR2984706._
  51. ^ abStumpf, M.P.H. (2012). "Critical Truths about Power Laws". [Science](https://en.m.wikipedia.org/wiki/Science(journal) "Science (journal)"). 335 (6069): 665–666. Bibcode:2012Sci...335..665S. doi:10.1126/science.1216142. PMID22323807._

Bibliography


about 2 years ago

ApexMD