A master equation for power laws (2024)

As a library, NLM provides access to scientific literature. Inclusion in an NLM database does not imply endorsem*nt of, or agreement with, the contents by NLM or the National Institutes of Health.
Learn more: PMC Disclaimer | PMC Copyright Notice

A master equation for power laws (1)

The Royal Society PublishingRoyal Society Open ScienceAboutBrowse by SubjectAlertsEditorial Board

R Soc Open Sci. 2022 Dec; 9(12): 220531.

Published online December 7, 2022. doi:10.1098/rsos.220531

PMCID: PMC9727680

PMID: 36483760

Sabin Roman, Conceptualization, Formal analysis, Investigation, Methodology, Project administration, Visualization, Writing – original draft, Writing – review & editingA master equation for power laws (2)1,2 and Francesco Bertolotti, Data curation, Validation, Writing – original draft, Writing – review & editing3

Author information Article notes Copyright and License information PMC Disclaimer

Associated Data

Data Availability Statement

Abstract

We propose a new mechanism for generating power laws. Starting from a random walk, we first outline a simple derivation of the Fokker–Planck equation. By analogy, starting from a certain Markov chain, we derive a master equation for power laws that describes how the number of cascades changes over time (cascades are consecutive transitions that end when the initial state is reached). The partial differential equation has a closed form solution which gives an explicit dependence of the number of cascades on their size and on time. Furthermore, the power law solution has a natural cut-off, a feature often seen in empirical data. This is due to the finite size a cascade can have in a finite time horizon. The derivation of the equation provides a justification for an exponent equal to 2, which agrees well with several empirical distributions, including Richardson’s Law on the size and frequency of deadly conflicts. Nevertheless, the equation can be solved for any exponent value. In addition, we propose an urn model where the number of consecutive ball extractions follows a power law. In all cases, the power law is manifest over the entire range of cascade sizes, as shown through log–log plots in the frequency and rank distributions.

Keywords: power laws, Markov chains, urn model

1.  Introduction

A power law is a nonlinear relationship between two quantities x and y that can be modelled generically by the following formula: y = axk, where k and a are constants, respectively, the exponent of the power law, and the width of the scaling relationship. The power law is scale-invariant relationship f(xc) ∝ f(x) that holds for any value of c. Graphically, it implies that the curve describing the relationship between x and y maintains its shape under any possible dilatation. Moreover, power laws can be represented as straight lines on a log–log plot, called the signature of the power law.

The signature can be employed to analyse empirical data by comparing the distribution of the data on a log–log plot with the best fitting power law. Probably, this representation was first introduced by J. C. Wills in 1922 to plot the distribution of the number of species in a genus [1]. Many lognormal relationships appear as power laws when plotted for small ranges. As a consequence, to assert that a relationship between two variables is a power law, it should hold for at least two orders of magnitude [2].

In the next section, we enumerate some representative examples of power law distributions from physics, biology, ecology and social sciences. Furthermore, we list the most well-known mechanisms for generating power laws. In the third section, we proceed to propose a Markov process model for generating power laws. First, we outline a derivation of the Fokker–Planck equation by starting from a random walk and taking a continuum limit from a discrete state space. This is a well-known approach for deriving the Fokker–Planck equation. We provide our own treatment of this procedure by changing the difference operator Δ acting on discrete functions to the differential operator d that acts on real variables. We then propose a Markov chain that displays a power law in its long-term equilibrium distribution. Based on this and in analogy with the Fokker–Planck equation derivation, we obtain a master equation for power laws that we simplify to an analytically tractable form. Surprisingly, the solution can be written in closed form. Afterwards, in the fourth section, we propose an urn model where the number of cascades (consecutive balls drawn) scales according to a power law. This improves on an urn model by Brunk [3] that displays a power law relationship only in a limited range. Finally, we summarize our results in the Conclusion.

2.  Background

The main challenge regarding power law distributions is understanding their origin. Broadly speaking, there are two ways power laws can emerge: probability transformations and generative processes. Probability transformations derive power law distributions from other distributions. Generative processes are algorithms that create new distributions. Here, we provide more details on generative processes, because our own work is of this nature. To the best of our knowledge, there does not exist a comprehensive review of all the power law distribution generative models (even if [1] finds a large set), and it is not in the scope of this paper to provide one.

There are three main strategies to obtain a power law distribution by transforming an existing probability distribution: combining exponential distributions, inverting quantities or looking at the extreme values of distributions.

In the first case, the transformation consists of substituting the variable of an exponential distribution with another exponential distribution. The resulting distribution follows a power law, in which parameters depend on the original exponential distribution. The distribution of sizes of populations that grow exponentially (i.e. with infinite carrying capacity) and can suddenly become extinct at any time step with the same probability follows a power law that can be derived from this model [4].

The second case requires the inversion of the quantities of the probability distribution. If these quantities pass through zero, the transformation results in a power law distribution with exponent equal to 2, resulting from the derivation of an inverted variable. The most notable example of a power law distribution obtained in this way occurs in the paramagnetic phase of the Ising model [5,6].

The third way originates from the extreme value theorems which provide results on the asymptotic behaviour of the extreme realizations. The Pickands–Balkema–de Haan theorem [7,8] states that the conditional distribution of a random variable above a certain threshold tends to a Pareto distribution when the threshold tends to infinity. Therefore, it provides a rationale for the widespread observation of Pareto or power law behaviour, since it is the limiting behaviour of large events for a whole class of probability distributions [9]. Analogously, the Fisher–Tippett–Gnedenko theorem [10,11] describes the possible distributions that the maximum value can have; this can shed light in cases where a simple power law does not appear from the application of the Pickands–Balkema–de Haan theorem [12].

In addition, we identified in the literature at least four notable generative processes for power law distributions: phase transitions (along with self-organized criticality), random walks, sample-space reduction processes and the Yule processes.

The idea underlying phase transitions and critical phenomena is simple but powerful. In systems governed by a single length scale, the scale can diverge in some given conditions, giving birth to scale-free systems in which quantities are distributed as a power law. The precise point at which the length scale diverges is called a critical continuous point or phase transition. Nevertheless, it is unlikely that the parameters that regulate the phase transition of a real-world system happen to fall on that specific value. So, the existence of critical phenomena is not enough to explain the presence of power law distributions in many natural and social systems. But some systems appear to self-organize to lay close to critical points, independently from the initial conditions. This phenomenon, called self-organized criticality, can be exemplified by the classic sandpile model [13], which generates avalanches whose sizes are distributed according to a power law. Other examples of how self-organized criticality generates power law distributions includes earthquakes [14], neuronal avalanches [15] and forest fires [1,16]. However, the emergence and applicability of self-organized criticality is debatable [17].

Random walks are a succession of randomly generated steps on a given space [18]. The statistical properties of random walks tend toward universal distributions when the steps are independent of each other, and their number grows unbounded [19]. As a consequence, they have the ability to generate power law distributions even when the underlying rules are very simple. One of the most famous examples (and probably the simplest) relates to the first-return time. It can be proved that, under specific conditions, the distributions of the first-return times follows a power law [1,19].

The sample-space reduction process models history-dependent systems, such as the formation of sentences, whose number of possible different meanings is progressively reduced the more words are added to it. Corominas-Murtra et al. demonstrate that a power law distribution can be generated from such a process [20]. More specifically, Zipf’s Law necessarily emerges from the reduction over time of the number of possible states in which a system can be, as a consequence of symmetry breaking in random sampling processes [20] or from dependency structures in component systems [21]. The reduction space process explains some well-known domain-specific power law generation models, such as preferential attachment.

The Yule process is a generative model developed by G. Udny Yule to explain the power law distribution of the number of species in taxonomic groups [22]. It derives from simple rules. Firstly, starting from a pool of groups, the probability for a group to increase by one is proportional to the number of its elements. Secondly, at any speciation event, there is a possibility of generating a new species belonging to a brand-new taxonomic group. The model is a simplification of reality since it ignores extinction events. Nevertheless, it has been customized to explain the emergence of power laws in other systems such as city sizes [23] or paper citations [24].

Finally, other generative processes include highly optimized tolerance [25,26], the coherent noise model of biological extinction [27], the repeated fragmentation model of fixed length elements [1], the dynamic of times between records in a random process [28], the Hawkes processes [29] and out-from-criticality feedback [30]. A master equation for power laws has been previously obtained in the literature [3135] but only considers stationary distributions or transient dynamics, while our results give a non-stationary solution [36]. Furthermore, our choice of transition rates has not been considered before.

To the best of our knowledge, the above examples and mechanisms for generating power laws do not generally address the question: how do power laws change (or are preserved) over time? Answering this question requires a way to specify the time evolution of the power law distribution. Furthermore, in many cases, empirical power laws have a cut-off point above which the relationship no longer holds [1], and this has not been captured by generative models, which are realistic only in specific value ranges. The next section provides a mechanism by which an equation dictating the time evolution of power laws can be obtained. In addition, the solution displays a natural cut-off beyond a certain time horizon.

3.  Markov process models

Below we present an outline for a well-known derivation of the Fokker–Planck equation starting from a discrete Markov chain and taking a continuum limit. By analogy, we introduce a discrete Markov chain that has a power law stationary distribution and then we derive a master equation for power laws.

3.1. Fokker–Planck equation

The transition probabilities for a one-dimensional random walk are

Pi,i+1=pandPi,i1=1p=q.}

3.1

The probability for x jumps to the right and nx jumps to the left is

P(x,n)=(nx)pxqnx12πnpqexp[(xnp)22npq],

3.2

where in the limit of large enough n we can approximate the binomial distribution with a Gaussian. Let i, s be arbitrary states. For a continuous-time Markov chain we write pi(t) = Psi(0; t). For the Markov chain in figure 1a, we can write down the following equation for pi(t):

dpi(t)dt=di1pi1+bi+1pi+1dipibipi=(ai+1pi+1aipi)+di1pi12dipi+di+1pi+1,

3.3

where bi, di are the rates at which the process leaves state i to state i − 1, respectively to state i + 1. Furthermore, we define ai = dibi. Using the forward and backward difference operators Δxn = xn+1xn and xn=xnxn1, the above equation can be written

dpi(t)dt=Δ(aipi(t))+Δ(dipi(t))p(x,t)t=x(a(x)p(x,t))+2x2(d(x)p(x,t)),}

3.4

which in the continuous limit gives the Fokker–Planck equation. If a(x) = 0 and d(x) = D where D is constant, then the solution with initial condition p(0, x) = δ(xx0) is given by

p(x,t)=14πDtexp[(xx0)24Dt].

3.5

A master equation for power laws (3)

Discrete-state Markov chains for: (a) a continuous-time random walk, with multi-step transition probabilities satisfying equation (3.3), (b) a continuous-time process, with multi-step transition probabilities satisfying equation (3.8). The stationary distribution of the embedded Markov chain is a power law, as shown in derivation (3.7).

3.2. Power law equation

We consider the Markov chain with the following transition matrix:

Pi,i+1=exp(ki+1)andPi,0=1exp(ki+1).}

3.6

The dominant contribution to the long-term equilibrium distribution is given by

π(x)ΠixPi,i+1=exp(ki1i)exp(klnx)=xk.

3.7

In figure 2a, we compare numerical solutions for the equilibrium distribution with the approximation (3.7), showing that the power law approximation is highly accurate. The Markov chain given by (3.6) is the embedded chain of the continuous-time Markov process in figure 1b. Using the same notation as in the previous section, the probability pi(t) satisfies the following differential equation:

dpi(t)gdt=ai1pi1aipibipi.

3.8

A master equation for power laws (4)

Log–log plots of (a) the stationary distribution (dotted black line) of the Markov chain with transition probabilities (3.6) for k = 2, 3, along with a linear fit (solid grey line), and (b) solution (3.12) (dotted black line) to equation (3.9) for b/a = 2, 3 up to times t = 102, 104, along with linear fits (solid grey line).

We can write (3.8) using the backward difference operator. If we let the space index go from discrete to continuous, such as in the previous section where ix, then equation (3.8) becomes

dpi(t)dt=(aipi(t))bipi(t)p(x,t)t=x(a(x)p(x,t))b(x)p(x),}

3.9

which gives

p(x,t)t=a(x)p(x,t)x(a(x)+b(x))p(x),

3.10

where a(x) = a ek/x, b(x) = b (1 − ek/x) and a, b are constants. By Taylor expanding up to 1/x, we get a(x) ≃ a − (ak/x) and a′(x) + b(x) ≃ (bk/x). Without loss of generality, we can relabel bk as b and the equation reduces to

p(x,t)t=ap(x,t)xbxp(x,t).

3.11

The term containing a/x can be neglected because the probability p(x, t) depends on a power of x (see below), and so each term on the right-hand side is of (3.11) is of the same order in x.

Let c(x, t) be the unnormalized solution of equation (3.11). The solution c(x, t) has a closed-form expression for x > 0 and x < at,

c(x,t)=c0(atx1)b/a,

3.12

where c(x, 0) = c0. In figure 2b, we plot the solution (3.12) for different values of b/a and see a power law behaviour over a range of scales that increases over time. The tail of the solution falls off abruptly just as we see in many real-world examples of power law distributions [1].

Equation (3.11) posits that cascades (consecutive transitions) occur over time in inverse proportion to their size. Furthermore, there is a constant rate a at which they can increase in size. For a finite time horizon t, there is a maximum size that a cascade can reach, namely at, as can be seen in the solution (3.12).

We can get additional insights from equation (3.10) if we consider the second order Taylor expansion of a′(x) + b(x) ≃ (bk/x) + ((2akbk2)/2x2). If k = 2 a/b then the 1/x2 term cancels, and this implies an exponent of 2 in the solution (3.12). This is not necessary in deriving equation (3.11) but can serve as a rationale for the experimental observation of power laws with exponent 2.

4.  Urn model

Building on the insights of the previous section, we propose an urn model that displays power law behaviour. We can approximate the probability (3.8) to move from i to i + 1 state as

exp(ki+1)1ki+1

4.1

for large enough i. Hence, if an adequate value of k is chosen, then we can expect to see a power law emerging even in a simple model.

Brunk [3] proposes a model claiming to exhibit cascades that occur with frequencies inversely proportional to the cascade size, and provides graphs showing that a power law distribution emerges. We attempted to reproduce the results of the paper, but we obtained an exponential distribution instead of a power law. A power law relationship is seen to occur in a narrow range of scales that the graphs in the article use (figure 3a). Furthermore, we can analytically prove that the distribution that emerges is an exponential distribution and not a power law. The article [3] has been cited in recent years, and we consider it important to highlight the limitations of the model.

A master equation for power laws (5)

Log–log plots of the rank distribution of (a) cascade sizes (black dots) in Brunk’s model [3], showing a narrow range where a power law is observed, along with a linear fit (solid grey line), and (b) cascade sizes (black dots) in the model we propose with exponent k = 1.5, along with a linear fit (solid grey line). The line of best fit has slope 0.49, in good agreement with the theoretical value of k − 1 = 0.5.

Assume the urn contains N balls in total, with K balls black and the rest white. The model has the following steps:

  • (i)

    A ball is drawn, if it is black then the next ball is drawn, and the process is repeated until a white ball is drawn. This constitutes a cascade within the model. Black balls are drawn without replacement.

  • (ii)

    Each time a white ball is drawn the number of black balls is increased by a fraction g, while the number of white balls stay the same. If a white ball is drawn, then we repeat from step (i).

The probability that a cascade of size n occurs, i.e. n black balls are drawn followed by a white ball, is

P(S=n)=KNK1N1Kn+1Nn+1NKNn=K!N!(Nn)!(Kn)!NKNnKK+1/2NN+1/2(Nn)Nn+1/2(Kn)Kn+1/2Nn(Kn)Nn=(1nN)N+1/2(1nK)(K+1/2)(KnNn)n(1KnNn).

4.2

If KN then the first two term in the last row of (4.2) cancel to a large extent. If we take the logarithm of (4.2), we get

logP(S=n)nlog(KnNn)+log(1KnNn)nlog(f)+log(1f),

4.3

where f = (Kn)/(Nn) ≃ K/N is approximately the fraction of black balls in the urn. Equation (4.3) shows that probability of a cascade is exponentially distributed, at least as long as nN. Furthermore, as K increases then f → 1, which implies an uniform distribution.

We propose a different urn model. Let N0, K0 and W be the initial number of balls, respectively, black and white ones. The procedure of the model is as follows:

  • (i)

    Black balls are drawn with replacement. The number of black balls is increased by W/k after each draw. We continue to draw until a white ball is drawn.

  • (ii)

    The process repeats with resetting the number of balls to their original values N = N0, K = K0.

The rank distribution of the cascade sizes is plotted in figure 3b for k = 1.5. As we can see, the graph is very well fit by a straight line, indicating that a genuine power law emerges across the entire range of cascade sizes.

At the ith extraction we have the total number of balls Ni, of black balls Ki and white balls W. Then

Ni=Ki+W.

4.4

The urn model is designed such that at each extraction the ratio of Ki and Ni is approximately given by

KiNi=ek/(i+1).

4.5

Using the previous two equations we obtain that

Ki=ek/(i+1)1ek/(i+1)W(i+1k1)W.

4.6

The number of black balls is increased after each draw by

Ki+1KiWk.

4.7

If a constant C number of black balls is added after each draw, then k = W/C.

5.  Discussion

The relevance of this study relies upon two aspects: the increasing importance of understanding statistical laws in complex systems due to the availability of larger datasets [37], and the presence of power law distributions in various application domains [4,38,39]. The applicability of the models proposed in §3.2 is restricted to phenomena where time plays a key role in increasing cascade sizes. The more time passes from the initial observation, the greater the chances of seeing a larger cascade (which would place it at the right-end tail of the distribution at the time of observation). These phenomena can appear in a wide variety of disciplinary fields.

The frequency versus the amplitude of earthquakes [1] and avalanches [40,41] follows a power law, as well as the distribution of the peak gamma-ray intensity of solar flares [42], and the size of moon craters per surface area [43]. The largest observed cascade in these examples depends on time, with more extreme examples being discovered the longer the observation record is.

The social sciences also offer numerous examples of power law distributions [44] that naturally evolve in time (but whose time dependence has not necessarily been quantified), such as the range of time between two deaths in serial killers’ behavioural patterns [45], in narrative structure [46] or in the budget distribution of movies [47]. In economics, they mainly derive from aggregation and rich-get-richer processes. The existence of power laws in the distribution of wealth and income are known at least from the nineteenth century [48]. Notably, the work of Pareto was the first one to discover the presence of a power law distribution in a social system [38]. Power law distributions also appear in markets. For example, the frequency of firm size, measured by the number of employees, is distributed according to a power law [49]. Specifically, the exponent of this power law is approximately 1.06, which suggests there exists a ‘Zip’s Law’ for firms [38].

Western popular music markets present power law distributions in the lifetime of albums [50]. Stock markets exhibited power law distributions since their very beginning [51]. It has been shown that the number of trades per day, the size of price movements when a large volume of shares are bought or sold, and the number of shares traded per time period are power law distributed with exponents 3, 0.5 and 1.5, respectively [5254]. Inter-trading times also exhibit scaling properties consistent with power laws [55].

Another field in which power law distributions are pervasive is urban studies, where there are many scaling relationships between quantifiable city properties [56], such as between the number of gas stations and population levels [57], and the distribution of the population of cities in a given geographical area [58]. This latter distribution appears in different geographical areas [59] and depends on the granularity of the sampling [60]. Power law distributions also characterize digital infrastructure. The network analysis of Internet topology shows that at the end of the 1990s the degrees of the nodes were already power law distributed [61,62]. Also, the number of connections to a server in a single day, by a specific subsection of Internet users, follows a power law [63].

Furthermore, the academic system generates power laws. As first noted by Price, the number of citations of papers follows a power law distribution [64]. Similar features in academic systems can be observed in recent times, even in specific disciplinary areas. For instance, the cumulative number of citations over time for papers dealing with protein kinases is distributed as a power law. This is a consequence of a phenomenon called the Harlow–Knapp (H-K) effect, which is the propensity of the biomedical and pharmaceutical research communities to concentrate their research on a tiny fraction of the proteome [65,66].

In history, the well-known Richardson’s Law states that the frequency of wars’ sizes (measured in casualties) are power law distributed. This law is considered one of the few robust statistical regularities in studies of political conflict [67]. Nevertheless, some recent studies seem to cast doubt on Richardson’s Law, or at least suggest having a greater caution regarding it [68]. Another example from history regards the Roman Empire [69], for which the survival time of its emperors is distributed according to a power law [70].

Certain power law relationships do not dependent on the observation period and the proposed models would not apply. For example, in physics, the well-known Stefan–Boltzmann Law describes a power law relationship between the total amount of energy radiated per time from a black body due to its temperature [71,72]. Furthermore, in particle physics and astrophysics there are numerous examples of power laws, from the Tully–Fisher relationship between a galaxy’s luminosity and its rate of rotation [73], to the proportionality of the spin and mass of hadrons [74] and the density of eigenvalues of the Dirac operator in certain theories [75,76]. Other examples come from biology, where many works describe a scaling relationship between the basal metabolic rate and the body mass of animals [7779], such as the well-known Kleiber’s Law, which indicates that the basal metabolism of mammals increases according to m3/4 where m is the overall body mass [80].

Another well-known example (with no time dependence) is Zipf’s Law, which states that the rank frequency distribution of words in any sufficiently long text is distributed as a power law [81]. This law is independent of the language and holds even for artificial languages such as Esperanto [82]. While Zipf investigates specifically the distribution of words, a generalization of this law was later developed by Mandelbrot [83].

We end the discussion with a cautionary note on the discovery of power law in data and models. As is well noted in the literature [1], linear relationships in log–log plots in the frequency distributions of observations is insufficient to robustly establish that a power law holds. A more reliable test is to check if the linear relationship holds in log–log plots of rank distributions. As we have shown, the model proposed by Brunk [3] does not show power law behaviour. Another example where a power law is claimed to be observed is in the emergence of a scale-free network in systems with bounded rationality [84]. As other work has shown [85], the topology is not scale-free but rather of a core–periphery type.

6.  Conclusion

In this paper, we have proposed a new mechanism to generate power laws based on Markov models. By analogy with the derivation of the Fokker–Planck equation, we have obtained a master equation for power laws. The result is supported by the fact that the underlying embedded Markov chain has an equilibrium distribution that is a power law and because a simplified version of the equation admits a closed form solution that is a power law.

We proposed three models: a discrete-time and discrete-space Markov chain with transition probabilities (3.6), a continuous-time and discrete-space Markov chain shown in figure 2b and a Markov process (continuum in both time and space), with master equation given by (3.9). A simplified form of (3.9) is equation (3.11), which is valid in the limit of large cascade sizes. Further considerations indicate that an exponent of k = 2 gives additional cancellations independent of cascade size. This value of k is consistent with exponents observed in Richardson’s Law [68] and in the net worth of Americans [1]. Nevertheless, equation (3.11) has a closed-form solution for any exponent value. In addition, the equation provides the time dynamics for the power law relationship along with a natural cut-off size depending on the time horizon.

The stationary solutions for the general equation are given by a first-order differential equation, whose solutions have been explored elsewhere [32]. However, a stationary solution to the power law equation might not always be well defined, as our time-dependent solution (3.12) for the simplified equation (3.11) shows (the time derivative is non-zero). The insight we get from the exact time-dependent solution (3.5) of the Fokker-Planck equation is that the standard deviation of spatial displacements is proportional to the square root of time, which is characteristic of Brownian motion. Similarly, the time-dependent solution (3.12) gives the largest cascade size we can expect to see is linearly proportional to time.

Based on the insights from the Markov process, we propose a simple urn model that illustrates power law behaviour over the entire range of cascade sizes. The model only considers a constant addition of balls over time and despite the simplicity of the mechanism, a robust power law is observed to emerge. Our results are in contrast to a prior model by Brunk [3] which we prove does not show a genuine power law distribution. Finally, we discuss some possible applications of the proposed model to phenomena where power laws are observed and where the duration of the period of observation is important. As far as we know, all these contributions are novel to the literature.

Data accessibility

This article has no additional data.

Authors' contributions

S.R.: conceptualization, formal analysis, investigation, methodology, project administration, visualization, writing—original draft, writing—review and editing; F.B.: data curation, validation, writing—original draft, writing—review and editing.

All authors gave final approval for publication and agreed to be held accountable for the work performed therein.

Conflict of interest declaration

We declare we have no competing interests.

Funding

This work was supported by Grantham Foundation for the Protection of the Environment. We also thank the World University Network (WUN) for early support of this work in 2014.

References

1. Newman ME. 2005. Power laws, Pareto distributions and Zipf’s law. Contemp. Phys.46, 323-351. ( 10.1080/00107510500052444) [CrossRef] [Google Scholar]

2. Stumpf MP, Porter MA. 2012. Critical truths about power laws. Science335, 665-666. ( 10.1126/science.1216142) [PubMed] [CrossRef] [Google Scholar]

3. Brunk GG. 2000. Understanding self-organized criticality as a statistical process. Complexity5, 26-33. ( 10.1002/(SICI)1099-0526(200001/02)5:3<26::AID-CPLX6>3.0.CO;2-C) [CrossRef] [Google Scholar]

4. Reed WJ. 2001. The Pareto, Zipf and other power laws. Econ. Lett.74, 15-19. ( 10.1016/S0165-1765(01)00524-9) [CrossRef] [Google Scholar]

5. Jan N, Moseley L, Ray T, Stauffer D. 1999. Is the fossil record indicative of a critical system?Adv. Complex Syst.2, 137-141. ( 10.1142/S0219525999000084) [CrossRef] [Google Scholar]

6. Krasnytska M, Berche B, Holovatch Y, Kenna R. 2021. Generalized Ising model on a scale-free network: an interplay of power laws. Entropy23, 1175. ( 10.3390/e23091175) [PMC free article] [PubMed] [CrossRef] [Google Scholar]

7. Balkema AA, de Haan L. 1994. Residual life time at great age. Ann. Probab.2, 792-804. ( 10.1214/aop/1176996548) [CrossRef] [Google Scholar]

8. Pickands J. 1975. Statistical inference using extreme order statistics. Ann. Stat.3, 119-131. ( 10.1214/aos/1176343003) [CrossRef] [Google Scholar]

9. Alfarano S, Lux T. 2010. Extreme value theory as a theoretical background for power law behavior. Kiel Working Paper no. 1648, pp. 1–10.

10. Fisher RA, Tippett LH. 1928. Limiting forms of the frequency distribution of the largest or smallest member of a sample. Math. Proc. Camb. Philos. Soc.24, 180-190. ( 10.1017/S0305004100015681) [CrossRef] [Google Scholar]

11. Gnedenko B. 1943. Sur la distribution limite du terme maximum d’une serie aleatoire. Ann. Math.44, 423-453. ( 10.2307/1968974) [CrossRef] [Google Scholar]

12. Sornette D. 2006. Critical phenomena in natural sciences. Chaos, fractals, selforganization and disorder: concepts and tools. Berlin, Germany: Springer. [Google Scholar]

13. Bak P, Tang C, Wiesenfeld K. 1987. Self-organized criticality: an explanation of the 1/f noise. Phys. Rev. Lett.59, 381-384. ( 10.1103/PhysRevLett.59.381) [PubMed] [CrossRef] [Google Scholar]

14. Olami Z, Feder HJS, Christensen K. 1992. Self-organized criticality in a continuous, nonconservative cellular automaton modeling earthquakes. Phys. Rev. Lett.68, 1244-1247. ( 10.1103/PhysRevLett.68.1244) [PubMed] [CrossRef] [Google Scholar]

15. Beggs JM, Plenz D. 2003. Neuronal avalanches in neocortical circuits. J. Neurosci.23, 111 67-111 77. ( 10.1523/JNEUROSCI.23-35-11167.2003) [PMC free article] [PubMed] [CrossRef] [Google Scholar]

16. McKenzie D, Kennedy MC. 2012. Power laws reveal phase transitions in landscape controls of fire regimes. Nat. Commun.3, 726. ( 10.1038/ncomms1731) [PubMed] [CrossRef] [Google Scholar]

17. Frigg R. 2003. Self-organised criticality—what it is and what it isn’t. Stud. Hist. Philos. Sci. A34, 613-632. ( 10.1016/S0039-3681(03)00046-3) [CrossRef] [Google Scholar]

18. Hughes BD. 1995. Random walks and random environments: random walks, vol. 1. Oxford, UK: Clarendon Press. [Google Scholar]

19. Vignat C, Plastino A. 2006. Power-law random walks. Phys. Rev. E—Stat. Nonlinear, Soft Matter Phys.74, 051124. ( 10.1103/PhysRevE.74.051124) [PubMed] [CrossRef] [Google Scholar]

20. Corominas-Murtra B, Hanel R, Thurner S. 2015. Understanding scaling through history-dependent processes with collapsing sample space. Proc. Natl Acad. Sci. USA112, 5348-5353. ( 10.1073/pnas.1420946112) [PMC free article] [PubMed] [CrossRef] [Google Scholar]

21. Mazzolini A, Grilli J, De Lazzari E, Osella M, Lagomarsino MC, Gherardi M. 2018. Zipf and Heaps laws from dependency structures in component systems. Phys. Rev. E98, 012315. ( 10.1103/PhysRevE.98.012315) [PubMed] [CrossRef] [Google Scholar]

22. Yule GU. 1925. A mathematical theory of evolution based on the conclusions of Dr. J. C. Willis, F.R.S. J. R. Stat. Soc.88, 433-436. ( 10.2307/2341419) [CrossRef] [Google Scholar]

23. Simon HA. 1955. On a class of skew distribution functions. Biometrika42, 425-440. ( 10.1093/biomet/42.3-4.425) [CrossRef] [Google Scholar]

24. Krapivsky PL, Redner S, Leyvraz F. 2000. Connectivity of growing random networks. Phys. Rev. Lett.85, 4629. ( 10.1103/PhysRevLett.85.4629) [PubMed] [CrossRef] [Google Scholar]

25. Carlson JM, Doyle J. 1999. Highly optimized tolerance: a mechanism for power laws in designed systems. Phys. Rev. E—Stat. Phys. Plasmas Fluids Relat. Interdiscip. Topics60, 1412. [PubMed] [Google Scholar]

26. Carlson JM, Doyle J. 2000. Highly optimized tolerance: robustness and design in complex systems. Phys. Rev. Lett.84, 2529. ( 10.1103/PhysRevLett.84.2529) [PubMed] [CrossRef] [Google Scholar]

27. Sneppen K, Newman ME. 1997. Coherent noise, scale invariance and intermittency in large systems. Physica D110, 209-222. ( 10.1016/S0167-2789(97)00128-0) [CrossRef] [Google Scholar]

28. Sibani P, Littlewood PB. 1993. Slow dynamics from noise adaptation. Phys. Rev. Lett.71, 1482. ( 10.1103/PhysRevLett.71.1482) [PubMed] [CrossRef] [Google Scholar]

29. Kanazawa K, Sornette D. 2020. Nonuniversal power law distribution of intensities of the self-excited Hawkes process: a field-theoretical approach. Phys. Rev. Lett.2, 137-141. ( 10.1103/PhysRevLett.125.138301) [PubMed] [CrossRef] [Google Scholar]

30. Schulman LS. 2021. Apparent power laws can occur without criticality. Entropy23, 1486. ( 10.3390/e23111486) [PMC free article] [PubMed] [CrossRef] [Google Scholar]

31. Biró TS, Néda Z. 2017. Dynamical stationarity as a result of sustained random growth. Physical Review E95, 032130. ( 10.1103/PhysRevE.95.032130) [PubMed] [CrossRef] [Google Scholar]

32. Biró T.S, Néda Z.. 2018. Unidirectional random growth with resetting. Phys. A: Stat. Mech. Appl.499, 335-361. ( 10.1016/j.physa.2018.02.078) [CrossRef] [Google Scholar]

33. Biró TS, Néda Z, Telcs A. 2019. Entropic divergence and entropy related to nonlinear master equations. Entropy21, 993. ( 10.3390/e21100993) [CrossRef] [Google Scholar]

34. Gere I, Kelemen S, Tóth G, Biró TS, Néda Z. 2021. Wealth distribution in modern societies: collected data and a master equation approach. Phys. A: Stat. Mech. Appl.581, 126194. ( 10.1016/j.physa.2021.126194) [CrossRef] [Google Scholar]

35. Biró TS, Csillag L, Néda Z. 2021. Transient dynamics in the random growth and reset model. Entropy23, 306. ( 10.3390/e23030306) [PMC free article] [PubMed] [CrossRef] [Google Scholar]

36. Inácio I, Velhinho J. 2022. Comments on mathematical aspects of the Biró–Néda model. Mathematics10, 644. ( 10.3390/math10040644) [CrossRef] [Google Scholar]

37. Gerlach M, Altmann EG. 2019. Testing statistical laws in complex systems. Phys. Rev. Lett.122, 168301. ( 10.1103/PhysRevLett.122.168301) [PubMed] [CrossRef] [Google Scholar]

38. Gabaix X. 2016. Power laws in economics: an introduction. J. Econ. Perspect.30, 185-206. ( 10.1257/jep.30.1.185) [CrossRef] [Google Scholar]

39. Pinto CM, Mendes Lopes A, Machado JA. 2012. A review of power laws in real life phenomena. Commun. Nonlinear Sci. Numer. Simul.17, 3558-3578. ( 10.1016/j.cnsns.2012.01.013) [CrossRef] [Google Scholar]

40. Naisbitt A, Forster R, Birkeland K, Harrison W. 2008. Avalanche frequency and magnitude: using power law exponents to investigate snow avalanche size proportions through space and time. In Int. Snow Science Workshop,Whistler, British Columbia, Canada. [Google Scholar]

41. Polizzi S, Pérez-Reche FJ, Arneodo A, Argoul F. 2021. Power-law and log-normal avalanche size statistics in random growth processes. Phys. Rev. E104, L052101. ( 10.1103/PhysRevE.104.L052101) [PubMed] [CrossRef] [Google Scholar]

42. Lu ET, Hamilton RJ. 1991. Avalanches and the distribution of solar flares. Astrophys. J.380, L89-L92. ( 10.1086/186180) [CrossRef] [Google Scholar]

43. Neukum G, Ivanov B. 1994. Crater size distributions and impact probabilities on Earth from lunar, terrestrial planeta, and asteroid cratering data. In Hazards due to comets and asteroids, pp. 359-416. Tucson, AZ: University of Arizona Press. [Google Scholar]

44. Turcotte DL, Rundle JB. 2002. Self-organized complexity in the physical, biological, and social sciences. Proc. Natl Acad. Sci. USA99, 2463-2465. ( 10.1073/pnas.012579399) [PMC free article] [PubMed] [CrossRef] [Google Scholar]

45. Simkin MV, Roychowdhury VP. 2014. Stochastic modeling of a serial killer. J. Theor. Biol.355, 111-116. ( 10.1016/j.jtbi.2014.03.039) [PMC free article] [PubMed] [CrossRef] [Google Scholar]

46. Gessey-Jones T, Connaughton C, Dunbar R, Kenna R, MacCarron P, O’Conchobhair C, Yose J. 2020. Narrative structure of A Song of Ice and Fire creates a fictional world with realistic measures of social complexity. Proc. Natl Acad. Sci. USA117, 28 582-28 588. ( 10.1073/pnas.2006465117) [PMC free article] [PubMed] [CrossRef] [Google Scholar]

47. Bertolotti F, Roman S. 2022. Risk sensitive scheduling strategies of production studios on the US movie market: an agent-based simulation. Intell. Artif.16, 81-92. [Google Scholar]

48. Pareto V. 1896. Écrits sur la courbe de la répartition de la richesse. Geneva, Switzerland: Librairie Droz. [Google Scholar]

49. Axtell RL. 2001. Zipf distribution of U.S. firm sizes. Science293, 1818-1820. ( 10.1126/science.1062081) [PubMed] [CrossRef] [Google Scholar]

50. Schneider L, Gros C. 2019. Five decades of US, UK, German and Dutch music charts show that cultural processes are accelerating. R. Soc. Open Sci.6, 190944. ( 10.1098/rsos.190944) [PMC free article] [PubMed] [CrossRef] [Google Scholar]

51. Ribeiro AS, Pinheiro FL, Santos FC, Polónia A, Pacheco JM. 2018. Structural and temporal patterns of the first global trading market. R. Soc. Open Sci.5, 180577. ( 10.1098/rsos.180577) [PMC free article] [PubMed] [CrossRef] [Google Scholar]

52. Bouchaud JP, Farmer JD, Lillo F. 2009. How markets slowly digest changes in supply and demand. In Handbook of financial markets: dynamics and evolution, pp. 57-160. Amsterdam, The Netherlands: North Holland. [Google Scholar]

53. Kyle AS, Obizhaeva AA. 2013. Market microstructure invariance: theory and empirical tests. See http://www.smallake.kr/wp-content/uploads/2016/03/Kyle_Obizhaeva_INVARIANCE_79_BERKELEY.pdf.

54. Plerou V, Gopikrishnan P, Stanley HE. 2005. Quantifying fluctuations in market liquidity: analysis of the bid-ask spread. Phys. Rev. E—Stat. Nonlinear Soft Matter Phys.71, 046131. ( 10.1103/PhysRevE.71.046131) [PubMed] [CrossRef] [Google Scholar]

55. Ivanov PC, Yuen A, Podobnik B, Lee Y. 2004. Common scaling patterns in intertrade times of US stocks. Phys. Rev. E69, 056107. ( 10.1103/PhysRevE.69.056107) [PubMed] [CrossRef] [Google Scholar]

56. Wu L, Gong C, Yan X. 2019. Taylor’s power law and its decomposition in urban facilities. R. Soc. Open Sci.6, 180770. ( 10.1098/rsos.180770) [PMC free article] [PubMed] [CrossRef] [Google Scholar]

57. Kühnert C, Helbing D, West GB. 2006. Scaling laws in urban supply networks. Physica A363, 96-103. ( 10.1016/j.physa.2006.01.058) [CrossRef] [Google Scholar]

58. Gabaix X, Ioannides YM. 2004. The evolution of city size distributions. In Handbook of regional and urban economics, vol. 4, pp. 2341-2378. Amsterdam, The Netherlands: North Holland. [Google Scholar]

59. Rozenfeld HD, Rybski D, Gabaix X, Makse HA. 2011. The area and population of cities: new insights from a different perspective on cities. Am. Econ. Rev.101, 2205-25. ( 10.1257/aer.101.5.2205) [CrossRef] [Google Scholar]

60. Mori T, Smith TE, Hsu WT. 2020. Common power laws for cities and spatial fractal structures. Proc. Natl Acad. Sci. USA117, 6469-6475. ( 10.1073/pnas.1913014117) [PMC free article] [PubMed] [CrossRef] [Google Scholar]

61. Barabási AL, Albert R. 1999. Emergence of scaling in random networks. Science286, 509-512. [PubMed] [Google Scholar]

62. Faloutsos M, Faloutsos P, Faioutsos C. 1999. On power-law relationships of the internet topology. Comput. Commun. Rev.29, 251-262. ( 10.1145/316194.316229) [CrossRef] [Google Scholar]

63. Adamic L, Huberman BA. 2005. The nature of markets in the world wide web. SSRN Electron. J. ( 10.2139/ssrn.166108) [CrossRef] [Google Scholar]

64. De Solla Price DJ. 1965. Networks of scientific papers. Science149, 510-515. ( 10.1126/science.149.3683.510) [PubMed] [CrossRef] [Google Scholar]

65. Fedorov O, Müller S, Knapp S. 2010. The (un)targeted cancer kinome. Nat. Chem. Biol.6, 166-169. ( 10.1038/nchembio.297) [PubMed] [CrossRef] [Google Scholar]

66. Grueneberg DA, et al.2008. Kinase requirements in human cells: I. Comparing kinase requirements across various cell types. Proc. Natl Acad. Sci. USA105, 16 472-16 477. ( 10.1073/pnas.0808019105) [PMC free article] [PubMed] [CrossRef] [Google Scholar]

67. Clauset A, Wiegel FW. 2010. A generalized aggregation-disintegration model for the frequency of severe terrorist attacks. J. Conflict Resolut.54, 179-197. ( 10.1177/0022002709352452) [CrossRef] [Google Scholar]

68. Zwetsloot R. 2018. Testing Richardson’s law: a (cautionary) note on power laws in violence data. SSRN Electron. J. ( 10.2139/ssrn.3112804) [CrossRef] [Google Scholar]

69. Roman S, Palmer E. 2019. The growth and decline of the Western Roman Empire: quantifying the dynamics of army size, territory, and coinage. Cliodynamics10, 1-23. [Google Scholar]

70. Ramos PL, Costa LF, Louzada F, Rodrigues FA. 2021. Power laws in the Roman Empire: a survival analysis. R. Soc. Open Sci.8, 210850. ( 10.1098/rsos.210850) [PMC free article] [PubMed] [CrossRef] [Google Scholar]

71. Boltzmann L. 1884. Ableitung des Stefan’schen Gesetzes, betreffend die Abhängigkeit der Wärmestrahlung von der Temperatur aus der electromagnetischen Lichttheorie. Ann. Phys. Chem.22, 291-294. [Google Scholar]

72. Stefan J. 1879. Uber die beziehung zwischen der wärmestrahlung und der temperatur. Wien. Ber. II79, 391-428. [Google Scholar]

73. Tully RB, Fisher JR. 1977. A new method of determining distances to galaxies. Astron. Astrophys.54, 661-673. [Google Scholar]

74. Chew GF, Frautschi SC. 1962. Regge trajectories and the principle of maximum strength for strong interactions. Phys. Rev. Lett.8, 41-44. [Google Scholar]

75. Cheng A, Hasenfratz A, Petropoulos G, Schaich D. 2013. Scale-dependent mass anomalous dimension from Dirac eigenmodes. J. High Energy Phys.2013, 1-23. ( 10.1007/JHEP07(2013)061) [CrossRef] [Google Scholar]

76. Del Debbio L, Lucini B, Pica C, Patella A, Rago A, Roman S. 2013. Large-volume results in SU (2) with adjoint fermions. (https://arxiv.org/abs/1311.5597)

77. Benedict FG. 1938. Vital energetics: a study in comparative basal metabolism. Washington, DC: Carnegie Institute Washington Publication. [Google Scholar]

78. Brody S, Lardy HA. 1946. Bioenergetics and growth. J. Phys. Chem.50, 168-169. ( 10.1021/j150446a008) [CrossRef] [Google Scholar]

79. Kleiber M. 1961. The fire of life: an introduction to animal energetics. New York, NY: John Wiley & Sons, Inc. [Google Scholar]

80. Niklas KJ, Kutschera U. 2015. Kleiber’s law: how the Fire of Life ignited debate, fueled theory, and neglected plants as model organisms. Plant Signal. Behav.10, e1036216. ( 10.1080/15592324.2015.1036216) [PMC free article] [PubMed] [CrossRef] [Google Scholar]

81. Zipf GK. 1949. Human behavior and the principle of least effort. Cambridge, MA: Addison-Wesley. [Google Scholar]

82. Manaris B, Pellicoro L, Pothering G, Hodges H. 2006. Investigating Esperanto’s statistical proportions relative to other languages using neural networks and Zipf’s law. In AIA Int. Conf. on Artificial Intelligence and Applications, Innsbruck, Austria, 13–16 February.

83. Mandelbrot B. 1968. Information theory and psycholinguistics: a theory of word frequencies. In Readings in mathematical social science, pp. 350–368. Cambridge, MA: MIT Press. [Google Scholar]

84. Kasthurirathna D, Piraveenan M. 2015. Emergence of scale-free characteristics in socio-ecological systems with bounded rationality. Sci. Rep.5, 1-16. ( 10.1038/srep10448) [PMC free article] [PubMed] [CrossRef] [Google Scholar]

85. Roman S, Brede M. 2017. Topology-dependent rationality and quantal response equilibria in structured populations. Phys. Rev. E95, 052310. ( 10.1103/PhysRevE.95.052310) [PubMed] [CrossRef] [Google Scholar]

Articles from Royal Society Open Science are provided here courtesy of The Royal Society

A master equation for power laws (2024)

FAQs

A master equation for power laws? ›

A power law is a nonlinear relationship between two quantities x and y that can be modelled generically by the following formula: y = axk, where k and a are constants, respectively, the exponent of the power law, and the width of the scaling relationship.

What is the power law in math? ›

A power law is a relationship in which a relative change in one quantity gives rise to a proportional relative change in the other quantity, independent of the initial size of those quantities. An example is the area of a square region in terms of the length of its side.

What is the signature of the power law? ›

A power law is a functional relation of the form f(x) = a xn. f(x) varies as the nth power of x. A signature of power law relations is a straight line on a log-log plot of f(x) vs x. This is a consequence of one of the well-known characteristics of power laws: scale invariance.

Is a power law exponential? ›

Yes, even a single exponential, because abx is the same as cekx with a=c and b=ek. But "power law" usually means a multiple of x−r, not bx. That can't be written as a sum of exponential decays, but can be written as a weighted average (= integral) of exponentials e−xt using the Gamma function: x−r=1Γ(r)∫∞0tr−1e−xtdt.

How do you calculate power law distribution? ›

A power law distribution has the form Y = k Xα, where:
  1. X and Y are variables of interest,
  2. α is the law's exponent,
  3. k is a constant.

What is the Yule process power law? ›

The Yule process generates the power law about the number of species within genera in biology. The Simon process generates the power laws about the frequency of use of a word in a text, the population of cities, and so on (see the list in the Introduction for details).

What is the formula for the power law? ›

A power law is a nonlinear relationship between two quantities x and y that can be modelled generically by the following formula: y = axk, where k and a are constants, respectively, the exponent of the power law, and the width of the scaling relationship.

What is the mathematical formula for power? ›

Power (physics)
Power
Common symbolsP
SI unitwatt (W)
In SI base unitskg⋅m2⋅s3
Derivations from other quantitiesP = E/t P = F·v P = V·I P = τ·ω
1 more row

What is the formula for power law transformation? ›

Power - Law transformations

Power Law Transformation is of two types of transformation nth power transformation and nth root transformation. Formula: s = cr ^ γ

What is the broken power law model? ›

A broken power law is a piecewise function given by a sequence of conjoined power laws ( ) where each section has its own power (index) and is defined by bounding "breaks".

What is power law Peter Thiel? ›

In a power-law distribution, the largest entity is typically bigger, more valuable, or more powerful than all others combined. The second-largest is likewise bigger than the total of all those after it, and so on. Also, in a power-law distribution, the top 20% of the entities typically hold 80% of the value or power.

What is the Pareto principle of power law? ›

Mathematically, the 80/20 rule is roughly described by a power law distribution (also known as a Pareto distribution) for a particular set of parameters. Many natural phenomena are distributed according to power law statistics. It is an adage of business management that "80% of sales come from 20% of clients."

Is Pareto the same as power law? ›

Since power-law cumulative distributions imply a power-law form for p(x), 'Zipf's law' and 'Pareto distribution' are effectively synonymous with 'power-law distribution'.

What is double power law? ›

(a) Double power law is a power law with two different exponents. (b) Pareto II (or Lomax) distribution behaves as a constant function for small values of x, and as a power law for large x, with smooth transition.

What is the alpha in the power law? ›

The alpha would be your degrees of freedom + 1, so 4.36 in your case. So the exponent in the power law (your alpha) is n+1.

What is the equation for electrical power law? ›

The equation for electrical power is P=IV. Where P is electrical Power, I is current, and V is voltage. If you multiply power by time this will give you the total energy, which is generally given in kilowatt-hours. The power equation and the Ohm's law can be combined to produce P=V^2/R and P=I^2R.

What is the theory of power law? ›

In statistics, a power law is a functional relationship between two quantities, where a relative change in one quantity results in a relative change in the other quantity proportional to a power of the change, independent of the initial size of those quantities: one quantity varies as a power of another.

Why is it called power law? ›

In the specific case of "power law" the term "power law" is a term in science. It refers to a connection between two phenomena that have a relationship that goes as a power, y∝xα.

What is a power law distribution in simple terms? ›

A power law distribution has the property that large numbers are rare, but smaller numbers are more common. So it is more common for a person to make a small amount of money versus a large amount of money. The Pareto distribution is a continuous power law distribution that is based on the observations that Pareto made.

What is the law of power? ›

Law 1: Never outshine the master Law 2: Never put too much trust in friends; learn how to use enemies Law 3: Conceal your intentions Law 4: Always say less than necessary. The text is bold and elegant, laid out in black and red throughout and replete with fables and unique word sculptures.

What is the power rule formula simple? ›

The power rule for the derivative of a power function is d d x ( a x n ) = n a x n − 1 . Simply put, the derivative of a power function involves bringing the exponent value down and multiplying it to the function and then subtracting one from the exponent.

What is the power law in basic calculus? ›

Power law for limits states that the limit of the nth power of a function equals the nth power of the limit of the function. Root law for limits states that the limit of the nth root of a function equals the nth root of the limit of the function.

What is the power rule in algebra? ›

Power Rules

Power of a power rule: If an expression of a base raised to a power is being raised to another power, multiply the exponents and keep the base the same.

Top Articles
Latest Posts
Article information

Author: Patricia Veum II

Last Updated:

Views: 5411

Rating: 4.3 / 5 (44 voted)

Reviews: 91% of readers found this page helpful

Author information

Name: Patricia Veum II

Birthday: 1994-12-16

Address: 2064 Little Summit, Goldieton, MS 97651-0862

Phone: +6873952696715

Job: Principal Officer

Hobby: Rafting, Cabaret, Candle making, Jigsaw puzzles, Inline skating, Magic, Graffiti

Introduction: My name is Patricia Veum II, I am a vast, combative, smiling, famous, inexpensive, zealous, sparkling person who loves writing and wants to share my knowledge and understanding with you.