Law of large numbers

Last updated
An illustration of the law of large numbers using a particular run of rolls of a single die. As the number of rolls in this run increases, the average of the values of all the results approaches 3.5. Although each run would show a distinctive shape over a small number of throws (at the left), over a large number of rolls (to the right) the shapes would be extremely similar. Lawoflargenumbers.svg
An illustration of the law of large numbers using a particular run of rolls of a single die. As the number of rolls in this run increases, the average of the values of all the results approaches 3.5. Although each run would show a distinctive shape over a small number of throws (at the left), over a large number of rolls (to the right) the shapes would be extremely similar.

In probability theory, the law of large numbers (LLN) is a mathematical theorem that states that the average of the results obtained from a large number of independent and identical random samples converges to the true value, if it exists. [1] More formally, the LLN states that given a sample of independent and identically distributed values, the sample mean converges to the true mean.

Contents

The LLN is important because it guarantees stable long-term results for the averages of some random events. [1] [2] For example, while a casino may lose money in a single spin of the roulette wheel, its earnings will tend towards a predictable percentage over a large number of spins. Any winning streak by a player will eventually be overcome by the parameters of the game. Importantly, the law applies (as the name indicates) only when a large number of observations are considered. There is no principle that a small number of observations will coincide with the expected value or that a streak of one value will immediately be "balanced" by the others (see the gambler's fallacy).

The LLN only applies to the average of the results obtained from repeated trials and claims that this average converges to the expected value; it does not claim that the sum of n results gets close to the expected value times n as n increases.

Throughout its history, many mathematicians have refined this law. Today, the LLN is used in many fields including statistics, probability theory, economics, and insurance. [3]

Examples

For example, a single roll of a fair, six-sided die produces one of the numbers 1, 2, 3, 4, 5, or 6, each with equal probability. Therefore, the expected value of the average of the rolls is:

According to the law of large numbers, if a large number of six-sided dice are rolled, the average of their values (sometimes called the sample mean) will approach 3.5, with the precision increasing as more dice are rolled.

It follows from the law of large numbers that the empirical probability of success in a series of Bernoulli trials will converge to the theoretical probability. For a Bernoulli random variable, the expected value is the theoretical probability of success, and the average of n such variables (assuming they are independent and identically distributed (i.i.d.)) is precisely the relative frequency.

A graphical example of the law of large numbers used for two dice rolls. The sum of the two dice fluctuates in the first few rolls, but as the number of rolls increases, the expected value of the sum of the two dice approaches 7. Law of Large Numbers Example Graph.jpg
A graphical example of the law of large numbers used for two dice rolls. The sum of the two dice fluctuates in the first few rolls, but as the number of rolls increases, the expected value of the sum of the two dice approaches 7.

For example, a fair coin toss is a Bernoulli trial. When a fair coin is flipped once, the theoretical probability that the outcome will be heads is equal to 12. Therefore, according to the law of large numbers, the proportion of heads in a "large" number of coin flips "should be" roughly 12. In particular, the proportion of heads after n flips will almost surely converge to 12 as n approaches infinity.

Although the proportion of heads (and tails) approaches 12, almost surely the absolute difference in the number of heads and tails will become large as the number of flips becomes large. That is, the probability that the absolute difference is a small number approaches zero as the number of flips becomes large. Also, almost surely the ratio of the absolute difference to the number of flips will approach zero. Intuitively, the expected difference grows, but at a slower rate than the number of flips.

Another good example of the LLN is the Monte Carlo method. These methods are a broad class of computational algorithms that rely on repeated random sampling to obtain numerical results. The larger the number of repetitions, the better the approximation tends to be. The reason that this method is important is mainly that, sometimes, it is difficult or impossible to use other approaches. [4]

Limitation

The average of the results obtained from a large number of trials may fail to converge in some cases. For instance, the average of n results taken from the Cauchy distribution or some Pareto distributions (α<1) will not converge as n becomes larger; the reason is heavy tails. [5] The Cauchy distribution and the Pareto distribution represent two cases: the Cauchy distribution does not have an expectation, [6] whereas the expectation of the Pareto distribution (α<1) is infinite. [7] One way to generate the Cauchy-distributed example is where the random numbers equal the tangent of an angle uniformly distributed between −90° and +90°. [8] The median is zero, but the expected value does not exist, and indeed the average of n such variables have the same distribution as one such variable. It does not converge in probability toward zero (or any other value) as n goes to infinity.

And if the trials embed a selection bias, typical in human economic/rational behaviour, the law of large numbers does not help in solving the bias. Even if the number of trials is increased the selection bias remains.

History

Diffusion is an example of the law of large numbers. Initially, there are solute molecules on the left side of a barrier (magenta line) and none on the right. The barrier is removed, and the solute diffuses to fill the whole container.
.mw-parser-output .plainlist ol,.mw-parser-output .plainlist ul{line-height:inherit;list-style:none;margin:0;padding:0}.mw-parser-output .plainlist ol li,.mw-parser-output .plainlist ul li{margin-bottom:0}
Top: With a single molecule, the motion appears to be quite random.
Middle: With more molecules, there is clearly a trend where the solute fills the container more and more uniformly, but there are also random fluctuations.
Bottom: With an enormous number of solute molecules (too many to see), the randomness is essentially gone: The solute appears to move smoothly and systematically from high-concentration areas to low-concentration areas. In realistic situations, chemists can describe diffusion as a deterministic macroscopic phenomenon (see Fick's laws), despite its underlying random nature. DiffusionMicroMacro.gif
Diffusion is an example of the law of large numbers. Initially, there are solute molecules on the left side of a barrier (magenta line) and none on the right. The barrier is removed, and the solute diffuses to fill the whole container.
  • Top: With a single molecule, the motion appears to be quite random.
  • Middle: With more molecules, there is clearly a trend where the solute fills the container more and more uniformly, but there are also random fluctuations.
  • Bottom: With an enormous number of solute molecules (too many to see), the randomness is essentially gone: The solute appears to move smoothly and systematically from high-concentration areas to low-concentration areas. In realistic situations, chemists can describe diffusion as a deterministic macroscopic phenomenon (see Fick's laws), despite its underlying random nature.

The Italian mathematician Gerolamo Cardano (1501–1576) stated without proof that the accuracies of empirical statistics tend to improve with the number of trials. [9] [3] This was then formalized as a law of large numbers. A special form of the LLN (for a binary random variable) was first proved by Jacob Bernoulli. [10] [3] It took him over 20 years to develop a sufficiently rigorous mathematical proof which was published in his Ars Conjectandi (The Art of Conjecturing) in 1713. He named this his "Golden Theorem" but it became generally known as "Bernoulli's theorem". This should not be confused with Bernoulli's principle, named after Jacob Bernoulli's nephew Daniel Bernoulli. In 1837, S. D. Poisson further described it under the name "la loi des grands nombres" ("the law of large numbers"). [11] [12] [3] Thereafter, it was known under both names, but the "law of large numbers" is most frequently used.

After Bernoulli and Poisson published their efforts, other mathematicians also contributed to refinement of the law, including Chebyshev, [13] Markov, Borel, Cantelli, Kolmogorov and Khinchin. [3] Markov showed that the law can apply to a random variable that does not have a finite variance under some other weaker assumption, and Khinchin showed in 1929 that if the series consists of independent identically distributed random variables, it suffices that the expected value exists for the weak law of large numbers to be true. [14] [15] These further studies have given rise to two prominent forms of the LLN. One is called the "weak" law and the other the "strong" law, in reference to two different modes of convergence of the cumulative sample means to the expected value; in particular, as explained below, the strong form implies the weak. [14]

Forms

There are two different versions of the law of large numbers that are described below. They are called thestrong law of large numbers and the weak law of large numbers. [16] [1] Stated for the case where X1, X2, ... is an infinite sequence of independent and identically distributed (i.i.d.) Lebesgue integrable random variables with expected value E(X1) = E(X2) = ... = µ, both versions of the law state that the sample average

converges to the expected value:

(Lebesgue integrability of Xj means that the expected value E(Xj) exists according to Lebesgue integration and is finite. It does not mean that the associated probability measure is absolutely continuous with respect to Lebesgue measure.)

Introductory probability texts often additionally assume identical finite variance (for all ) and no correlation between random variables. In that case, the variance of the average of n random variables is

which can be used to shorten and simplify the proofs. This assumption of finite variance is not necessary. Large or infinite variance will make the convergence slower, but the LLN holds anyway. [17]

Mutual independence of the random variables can be replaced by pairwise independence [18] or exchangeability [19] in both versions of the law.

The difference between the strong and the weak version is concerned with the mode of convergence being asserted. For interpretation of these modes, see Convergence of random variables.

Weak law

Blank300.png
Lawoflargenumbersanimation2.gif
Blank300.png
Simulation illustrating the law of large numbers. Each frame, a coin that is red on one side and blue on the other is flipped, and a dot is added in the corresponding column. A pie chart shows the proportion of red and blue so far. Notice that while the proportion varies significantly at first, it approaches 50% as the number of trials increases.

The weak law of large numbers (also called Khinchin's law) states that given a collection of iid samples from a random variable with finite mean, the sample mean converges in probability to the expected value [20]

That is, for any positive number ε,

Interpreting this result, the weak law states that for any nonzero margin specified (ε), no matter how small, with a sufficiently large sample there will be a very high probability that the average of the observations will be close to the expected value; that is, within the margin.

As mentioned earlier, the weak law applies in the case of i.i.d. random variables, but it also applies in some other cases. For example, the variance may be different for each random variable in the series, keeping the expected value constant. If the variances are bounded, then the law applies, as shown by Chebyshev as early as 1867. (If the expected values change during the series, then we can simply apply the law to the average deviation from the respective expected values. The law then states that this converges in probability to zero.) In fact, Chebyshev's proof works so long as the variance of the average of the first n values goes to zero as n goes to infinity. [15] As an example, assume that each random variable in the series follows a Gaussian distribution (normal distribution) with mean zero, but with variance equal to , which is not bounded. At each stage, the average will be normally distributed (as the average of a set of normally distributed variables). The variance of the sum is equal to the sum of the variances, which is asymptotic to . The variance of the average is therefore asymptotic to and goes to zero.

There are also examples of the weak law applying even though the expected value does not exist.

Strong law

The strong law of large numbers (also called Kolmogorov's law) states that the sample average converges almost surely to the expected value [21]

That is,

What this means is that the probability that, as the number of trials n goes to infinity, the average of the observations converges to the expected value, is equal to one. The modern proof of the strong law is more complex than that of the weak law, and relies on passing to an appropriate subsequence. [17]

The strong law of large numbers can itself be seen as a special case of the pointwise ergodic theorem. This view justifies the intuitive interpretation of the expected value (for Lebesgue integration only) of a random variable when sampled repeatedly as the "long-term average".

Law 3 is called the strong law because random variables which converge strongly (almost surely) are guaranteed to converge weakly (in probability). However the weak law is known to hold in certain conditions where the strong law does not hold and then the convergence is only weak (in probability). See differences between the weak law and the strong law.

The strong law applies to independent identically distributed random variables having an expected value (like the weak law). This was proved by Kolmogorov in 1930. It can also apply in other cases. Kolmogorov also showed, in 1933, that if the variables are independent and identically distributed, then for the average to converge almost surely on something (this can be considered another statement of the strong law), it is necessary that they have an expected value (and then of course the average will converge almost surely on that). [22]

If the summands are independent but not identically distributed, then

provided that each Xk has a finite second moment and

This statement is known as Kolmogorov's strong law, see e.g. Sen & Singer (1993 , Theorem 2.3.10).

Differences between the weak law and the strong law

The weak law states that for a specified large n, the average is likely to be near μ. [23] Thus, it leaves open the possibility that happens an infinite number of times, although at infrequent intervals. (Not necessarily for all n).

The strong law shows that this almost surely will not occur. It does not imply that with probability 1, we have that for any ε > 0 the inequality holds for all large enough n, since the convergence is not necessarily uniform on the set where it holds. [24]

The strong law does not hold in the following cases, but the weak law does. [25] [26]

  1. Let X be an exponentially distributed random variable with parameter 1. The random variable has no expected value according to Lebesgue integration, but using conditional convergence and interpreting the integral as a Dirichlet integral, which is an improper Riemann integral, we can say:
  2. Let X be a geometrically distributed random variable with probability 0.5. The random variable does not have an expected value in the conventional sense because the infinite series is not absolutely convergent, but using conditional convergence, we can say:
  3. If the cumulative distribution function of a random variable is
    then it has no expected value, but the weak law is true. [27] [28]
  4. Let Xk be plus or minus (starting at sufficiently large k so that the denominator is positive) with probability 12 for each. [22] The variance of Xk is then Kolmogorov's strong law does not apply because the partial sum in his criterion up to k = n is asymptotic to and this is unbounded. If we replace the random variables with Gaussian variables having the same variances, namely , then the average at any point will also be normally distributed. The width of the distribution of the average will tend toward zero (standard deviation asymptotic to ), but for a given ε, there is probability which does not go to zero with n, while the average sometime after the nth trial will come back up to ε. Since the width of the distribution of the average is not zero, it must have a positive lower bound p(ε), which means there is a probability of at least p(ε) that the average will attain ε after n trials. It will happen with probability p(ε)/2 before some m which depends on n. But even after m, there is still a probability of at least p(ε) that it will happen. (This seems to indicate that p(ε)=1 and the average will attain ε an infinite number of times.)

Uniform laws of large numbers

There are extensions of the law of large numbers to collections of estimators, where the convergence is uniform over the collection; thus the name uniform law of large numbers.

Suppose f(x,θ) is some function defined for θ ∈ Θ, and continuous in θ. Then for any fixed θ, the sequence {f(X1,θ), f(X2,θ), ...} will be a sequence of independent and identically distributed random variables, such that the sample mean of this sequence converges in probability to E[f(X,θ)]. This is the pointwise (in θ) convergence.

A particular example of a uniform law of large numbers states the conditions under which the convergence happens uniformly in θ. If [29] [30]

  1. Θ is compact,
  2. f(x,θ) is continuous at each θ ∈ Θ for almost all xs, and measurable function of x at each θ.
  3. there exists a dominating function d(x) such that E[d(X)] < ∞, and

Then E[f(X,θ)] is continuous in θ, and

This result is useful to derive consistency of a large class of estimators (see Extremum estimator).

Borel's law of large numbers

Borel's law of large numbers, named after Émile Borel, states that if an experiment is repeated a large number of times, independently under identical conditions, then the proportion of times that any specified event is expected to occur approximately equals the probability of the event's occurrence on any particular trial; the larger the number of repetitions, the better the approximation tends to be. More precisely, if E denotes the event in question, p its probability of occurrence, and Nn(E) the number of times E occurs in the first n trials, then with probability one, [31]

This theorem makes rigorous the intuitive notion of probability as the expected long-run relative frequency of an event's occurrence. It is a special case of any of several more general laws of large numbers in probability theory.

Chebyshev's inequality . Let X be a random variable with finite expected value μ and finite non-zero variance σ2. Then for any real number k > 0,

Proof of the weak law

Given X1, X2, ... an infinite sequence of i.i.d. random variables with finite expected value , we are interested in the convergence of the sample average

The weak law of large numbers states:

Proof using Chebyshev's inequality assuming finite variance

This proof uses the assumption of finite variance (for all ). The independence of the random variables implies no correlation between them, and we have that

The common mean μ of the sequence is the mean of the sample average:

Using Chebyshev's inequality on results in

This may be used to obtain the following:

As n approaches infinity, the expression approaches 1. And by definition of convergence in probability, we have obtained

Proof using convergence of characteristic functions

By Taylor's theorem for complex functions, the characteristic function of any random variable, X, with finite mean μ, can be written as

All X1, X2, ... have the same characteristic function, so we will simply denote this φX.

Among the basic properties of characteristic functions there are

if X and Y are independent.

These rules can be used to calculate the characteristic function of in terms of φX:

The limit eitμ is the characteristic function of the constant random variable μ, and hence by the Lévy continuity theorem, converges in distribution to μ:

μ is a constant, which implies that convergence in distribution to μ and convergence in probability to μ are equivalent (see Convergence of random variables.) Therefore,

This shows that the sample mean converges in probability to the derivative of the characteristic function at the origin, as long as the latter exists.

Proof of the strong law

We give a relatively simple proof of the strong law under the assumptions that the are iid, , , and .

Let us first note that without loss of generality we can assume that by centering. In this case, the strong law says that

or

It is equivalent to show that

Note that

and thus to prove the strong law we need to show that for every , we have

Define the events , and if we can show that

then the Borel-Cantelli Lemma implies the result. So let us estimate .

We compute

We first claim that every term of the form where all subscripts are distinct, must have zero expectation. This is because by independence, and the last term is zero --- and similarly for the other terms. Therefore the only terms in the sum with nonzero expectation are and . Since the are identically distributed, all of these are the same, and moreover .

There are terms of the form and terms of the form , and so

Note that the right-hand side is a quadratic polynomial in , and as such there exists a such that for sufficiently large. By Markov,

for sufficiently large, and therefore this series is summable. Since this holds for any , we have established the Strong LLN.


Another proof can be found in [32]

For a proof without the added assumption of a finite fourth moment, see Section 22 of. [33]

Consequences

The law of large numbers provides an expectation of an unknown distribution from a realization of the sequence, but also any feature of the probability distribution. [1] By applying Borel's law of large numbers, one could easily obtain the probability mass function. For each event in the objective probability mass function, one could approximate the probability of the event's occurrence with the proportion of times that any specified event occurs. The larger the number of repetitions, the better the approximation. As for the continuous case: , for small positive h. Thus, for large n:

With this method, one can cover the whole x-axis with a grid (with grid size 2h) and obtain a bar graph which is called a histogram.

Applications

One application of the LLN is the use of an important method of approximation, the Monte Carlo Method. [3] This method uses a random sampling of numbers to approximate numerical results. The algorithm to compute an integral of f(x) on an interval [a,b] is as follows: [3]

  1. Simulate uniform random variables X1, X2, ..., Xn which can be done using a software, and use a random number table that gives U1, U2, ..., Un independent and identically distributed (i.i.d.) random variables on [0,1]. Then let Xi = a+(b - a)Ui for i= 1, 2, ..., n. Then X1, X2, ..., Xn are independent and identically distributed uniform random variables on [a, b].
  2. Evaluate f(X1), f(X2), ..., f(Xn)
  3. Take the average of f(X1), f(X2), ..., f(Xn) by computing and then by the Strong Law of Large Numbers, this converges to = =

We can find the integral of on [-1,2]. Using traditional methods to compute this integral is very difficult, so the Monte Carlo Method can be used here. [3] Using the above algorithm, we get

= 0.905 when n=25

and

= 1.028 when n=250

We observe that as n increases, the numerical value also increases. When we get the actual results for the integral we get

= 1.000194

By using the LLN, the approximation of the integral was more accurate and was closer to its true value. [3]

Another example is the integration of f(x) = on [0,1]. [34] Using the Monte Carlo Method and the LLN, we can see that as the number of samples increases, the numerical value gets closer to 0.4180233. [34]

See also

Notes

  1. 1 2 3 4 Dekking, Michel (2005). A Modern Introduction to Probability and Statistics . Springer. pp.  181–190. ISBN   9781852338961.
  2. Yao, Kai; Gao, Jinwu (2016). "Law of Large Numbers for Uncertain Random Variables". IEEE Transactions on Fuzzy Systems. 24 (3): 615–621. doi:10.1109/TFUZZ.2015.2466080. ISSN   1063-6706. S2CID   2238905.
  3. 1 2 3 4 5 6 7 8 9 Sedor, Kelly. "The Law of Large Numbers and its Applications" (PDF).
  4. Kroese, Dirk P.; Brereton, Tim; Taimre, Thomas; Botev, Zdravko I. (2014). "Why the Monte Carlo method is so important today". Wiley Interdisciplinary Reviews: Computational Statistics. 6 (6): 386–392. doi:10.1002/wics.1314. S2CID   18521840.
  5. Dekking, Michel, ed. (2005). A modern introduction to probability and statistics: understandig why and how. Springer texts in statistics. London [Heidelberg]: Springer. p. 187. ISBN   978-1-85233-896-1.
  6. Dekking, Michel (2005). A Modern Introduction to Probability and Statistics . Springer. pp.  92. ISBN   9781852338961.
  7. Dekking, Michel (2005). A Modern Introduction to Probability and Statistics . Springer. pp.  63. ISBN   9781852338961.
  8. Pitman, E. J. G.; Williams, E. J. (1967). "Cauchy-Distributed Functions of Cauchy Variates". The Annals of Mathematical Statistics. 38 (3): 916–918. doi: 10.1214/aoms/1177698885 . ISSN   0003-4851. JSTOR   2239008.
  9. Mlodinow, L. (2008). The Drunkard's Walk. New York: Random House. p. 50.
  10. Bernoulli, Jakob (1713). "4". Ars Conjectandi: Usum & Applicationem Praecedentis Doctrinae in Civilibus, Moralibus & Oeconomicis (in Latin). Translated by Sheynin, Oscar.
  11. Poisson names the "law of large numbers" (la loi des grands nombres) in: Poisson, S. D. (1837). Probabilité des jugements en matière criminelle et en matière civile, précédées des règles générales du calcul des probabilitiés (in French). Paris, France: Bachelier. p.  7. He attempts a two-part proof of the law on pp. 139–143 and pp. 277 ff.
  12. Hacking, Ian (1983). "19th-century Cracks in the Concept of Determinism". Journal of the History of Ideas. 44 (3): 455–475. doi:10.2307/2709176. JSTOR   2709176.
  13. Tchebichef, P. (1846). "Démonstration élémentaire d'une proposition générale de la théorie des probabilités". Journal für die reine und angewandte Mathematik (in French). 1846 (33): 259–267. doi:10.1515/crll.1846.33.259. S2CID   120850863.
  14. 1 2 Seneta 2013.
  15. 1 2 Yuri Prohorov. "Law of large numbers". Encyclopedia of Mathematics. EMS Press.
  16. Bhattacharya, Rabi; Lin, Lizhen; Patrangenaru, Victor (2016). A Course in Mathematical Statistics and Large Sample Theory. Springer Texts in Statistics. New York, NY: Springer New York. doi:10.1007/978-1-4939-4032-5. ISBN   978-1-4939-4030-1.
  17. 1 2 "The strong law of large numbers – What's new". Terrytao.wordpress.com. 19 June 2008. Retrieved 2012-06-09.
  18. Etemadi, N. Z. (1981). "An elementary proof of the strong law of large numbers". Wahrscheinlichkeitstheorie Verw Gebiete. 55 (1): 119–122. doi: 10.1007/BF01013465 . S2CID   122166046.
  19. Kingman, J. F. C. (April 1978). "Uses of Exchangeability". The Annals of Probability. 6 (2). doi: 10.1214/aop/1176995566 . ISSN   0091-1798.
  20. Loève 1977 , Chapter 1.4, p. 14
  21. Loève 1977 , Chapter 17.3, p. 251
  22. 1 2 Yuri Prokhorov. "Strong law of large numbers". Encyclopedia of Mathematics.
  23. "What Is the Law of Large Numbers? (Definition) | Built In". builtin.com. Retrieved 2023-10-20.
  24. Ross (2009)
  25. Lehmann, Erich L.; Romano, Joseph P. (2006-03-30). Weak law converges to constant. Springer. ISBN   9780387276052.
  26. Dguvl Hun Hong; Sung Ho Lee (1998). "A Note on the Weak Law of Large Numbers for Exchangeable Random Variables" (PDF). Communications of the Korean Mathematical Society. 13 (2): 385–391. Archived from the original (PDF) on 2016-07-01. Retrieved 2014-06-28.
  27. Mukherjee, Sayan. "Law of large numbers" (PDF). Archived from the original (PDF) on 2013-03-09. Retrieved 2014-06-28.
  28. J. Geyer, Charles. "Law of large numbers" (PDF).
  29. Newey & McFadden 1994 , Lemma 2.4
  30. Jennrich, Robert I. (1969). "Asymptotic Properties of Non-Linear Least Squares Estimators". The Annals of Mathematical Statistics. 40 (2): 633–643. doi: 10.1214/aoms/1177697731 .
  31. Wen, Liu (1991). "An Analytic Technique to Prove Borel's Strong Law of Large Numbers". The American Mathematical Monthly. 98 (2): 146–148. doi:10.2307/2323947. JSTOR   2323947.
  32. Etemadi, Nasrollah (1981). "An elementary proof of the strong law of large numbers". Zeitschrift f{\"u}r Wahrscheinlichkeitstheorie und verwandte Gebiete. Springer. 55: 119–122. doi: 10.1007/BF01013465 . S2CID   122166046.
  33. Billingsley, Patrick (1979). Probability and Measure.
  34. 1 2 Reiter, Detlev (2008), Fehske, H.; Schneider, R.; Weiße, A. (eds.), "The Monte Carlo Method, an Introduction", Computational Many-Particle Physics, Lecture Notes in Physics, Berlin, Heidelberg: Springer Berlin Heidelberg, vol. 739, pp. 63–78, doi:10.1007/978-3-540-74686-7_3, ISBN   978-3-540-74685-0 , retrieved 2023-12-08

Related Research Articles

<span class="mw-page-title-main">Expected value</span> Average value of a random variable

In probability theory, the expected value is a generalization of the weighted average. Informally, the expected value is the arithmetic mean of the possible values a random variable can take, weighted by the probability of those outcomes. Since it is obtained through arithmetic, the expected value sometimes may not even be included in the sample data set; it is not the value you would "expect" to get in reality.

<span class="mw-page-title-main">Probability theory</span> Branch of mathematics concerning probability

Probability theory or probability calculus is the branch of mathematics concerned with probability. Although there are several different probability interpretations, probability theory treats the concept in a rigorous mathematical manner by expressing it through a set of axioms. Typically these axioms formalise probability in terms of a probability space, which assigns a measure taking values between 0 and 1, termed the probability measure, to a set of outcomes called the sample space. Any specified subset of the sample space is called an event.

<span class="mw-page-title-main">Variance</span> Statistical measure of how far values spread from their average

In probability theory and statistics, variance is the expected value of the squared deviation from the mean of a random variable. The standard deviation (SD) is obtained as the square root of the variance. Variance is a measure of dispersion, meaning it is a measure of how far a set of numbers is spread out from their average value. It is the second central moment of a distribution, and the covariance of the random variable with itself, and it is often represented by , , , , or .

In probability theory, the central limit theorem (CLT) states that, under appropriate conditions, the distribution of a normalized version of the sample mean converges to a standard normal distribution. This holds even if the original variables themselves are not normally distributed. There are several versions of the CLT, each applying in the context of different conditions.

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

In probability theory and statistics, the exponential distribution or negative exponential distribution is the probability distribution of the distance between events in a Poisson point process, i.e., a process in which events occur continuously and independently at a constant average rate; the distance parameter could be any meaningful mono-dimensional measure of the process, such as time between production errors, or length along a roll of fabric in the weaving manufacturing process. It is a particular case of the gamma distribution. It is the continuous analogue of the geometric distribution, and it has the key property of being memoryless. In addition to being used for the analysis of Poisson point processes it is found in various other contexts.

In probability theory, there exist several different notions of convergence of sequences of random variables. The different notions of convergence capture different properties about the sequence, with some notions of convergence being stronger than others. For example, convergence in distribution tells us about the limit distribution of a sequence of random variables. This is a weaker notion than convergence in probability, which tells us about the value a random variable will take, rather than just the distribution.

In probability theory, Markov's inequality gives an upper bound on the probability that a non-negative random variable is greater than or equal to some positive constant. Markov's inequality is tight in the sense that for each chosen positive constant, there exists a random variable such that the inequality is in fact an equality.

In information theory, the asymptotic equipartition property (AEP) is a general property of the output samples of a stochastic source. It is fundamental to the concept of typical set used in theories of data compression.

<span class="mw-page-title-main">Jensen's inequality</span> Theorem of convex functions

In mathematics, Jensen's inequality, named after the Danish mathematician Johan Jensen, relates the value of a convex function of an integral to the integral of the convex function. It was proved by Jensen in 1906, building on an earlier proof of the same inequality for doubly-differentiable functions by Otto Hölder in 1889. Given its generality, the inequality appears in many forms depending on the context, some of which are presented below. In its simplest form the inequality states that the convex transformation of a mean is less than or equal to the mean applied after convex transformation; it is a simple corollary that the opposite is true of concave transformations.

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

<span class="mw-page-title-main">Stable distribution</span> Distribution of variables which satisfies a stability property under linear combinations

In probability theory, a distribution is said to be stable if a linear combination of two independent random variables with this distribution has the same distribution, up to location and scale parameters. A random variable is said to be stable if its distribution is stable. The stable distribution family is also sometimes referred to as the Lévy alpha-stable distribution, after Paul Lévy, the first mathematician to have studied it.

<span class="mw-page-title-main">Directional statistics</span>

Directional statistics is the subdiscipline of statistics that deals with directions, axes or rotations in Rn. More generally, directional statistics deals with observations on compact Riemannian manifolds including the Stiefel manifold.

<span class="mw-page-title-main">Consistent estimator</span> Statistical estimator converging in probability to a true parameter as sample size increases

In statistics, a consistent estimator or asymptotically consistent estimator is an estimator—a rule for computing estimates of a parameter θ0—having the property that as the number of data points used increases indefinitely, the resulting sequence of estimates converges in probability to θ0. This means that the distributions of the estimates become more and more concentrated near the true value of the parameter being estimated, so that the probability of the estimator being arbitrarily close to θ0 converges to one.

von Mises distribution Probability distribution on the circle

In probability theory and directional statistics, the von Mises distribution is a continuous probability distribution on the circle. It is a close approximation to the wrapped normal distribution, which is the circular analogue of the normal distribution. A freely diffusing angle on a circle is a wrapped normally distributed random variable with an unwrapped variance that grows linearly in time. On the other hand, the von Mises distribution is the stationary distribution of a drift and diffusion process on the circle in a harmonic potential, i.e. with a preferred orientation. The von Mises distribution is the maximum entropy distribution for circular data when the real and imaginary parts of the first circular moment are specified. The von Mises distribution is a special case of the von Mises–Fisher distribution on the N-dimensional sphere.

<span class="mw-page-title-main">Empirical distribution function</span> Distribution function associated with the empirical measure of a sample

In statistics, an empirical distribution function is the distribution function associated with the empirical measure of a sample. This cumulative distribution function is a step function that jumps up by 1/n at each of the n data points. Its value at any specified value of the measured variable is the fraction of observations of the measured variable that are less than or equal to the specified value.

<span class="mw-page-title-main">Characteristic function (probability theory)</span> Fourier transform of the probability density function

In probability theory and statistics, the characteristic function of any real-valued random variable completely defines its probability distribution. If a random variable admits a probability density function, then the characteristic function is the Fourier transform of the probability density function. Thus it provides an alternative route to analytical results compared with working directly with probability density functions or cumulative distribution functions. There are particularly simple results for the characteristic functions of distributions defined by the weighted sums of random variables.

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

Stochastic approximation methods are a family of iterative methods typically used for root-finding problems or for optimization problems. The recursive update rules of stochastic approximation methods can be used, among other things, for solving linear systems when the collected data is corrupted by noise, or for approximating extreme values of functions which cannot be computed directly, but only estimated via noisy observations.

In probability and statistics, the Tweedie distributions are a family of probability distributions which include the purely continuous normal, gamma and inverse Gaussian distributions, the purely discrete scaled Poisson distribution, and the class of compound Poisson–gamma distributions which have positive mass at zero, but are otherwise continuous. Tweedie distributions are a special case of exponential dispersion models and are often used as distributions for generalized linear models.

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

In probability theory and directional statistics, a wrapped normal distribution is a wrapped probability distribution that results from the "wrapping" of the normal distribution around the unit circle. It finds application in the theory of Brownian motion and is a solution to the heat equation for periodic boundary conditions. It is closely approximated by the von Mises distribution, which, due to its mathematical simplicity and tractability, is the most commonly used distribution in directional statistics.

References