BRS-inequality

Last updated

BRS-inequality is the short name for Bruss-Robertson-Steele inequality. This inequality gives a convenient upper bound for the expected maximum number of non-negative random variables one can sum up without exceeding a given upper bound .

Contents

For example, suppose 100 random variables are all uniformly distributed on , not necessarily independent, and let , say. Let be the maximum number of one can select in such that their sum does not exceed . is a random variable, so what can one say about bounds for its expectation? How would an upper bound for behave, if one changes the size of the sample and keeps fixed, or alternatively, if one keeps fixed but varies ? What can one say about , if the uniform distribution is replaced by another continuous distribution? In all generality, what can one say if each may have its own continuous distribution function ?

General problem

Let be a sequence of non-negative random variables (possibly dependent) that are jointly continuously distributed. For and let be the maximum number of observations among that one can sum up without exceeding .

Now, to obtain one may think of looking at the list of all observations, first select the smallest one, then add the second smallest, then the third and so on, as long as the accumulated sum does not exceed . Hence can be defined in terms of the increasing order statistics of , denoted by , namely by

What is the best possible general upper bound for if one requires only the continuity of the joint distribution of all variables? And then, how to compute this bound?

Identically distributed random variables.

Theorem 1 Let be identically distributed non-negative random variables with absolutely continuous distribution function . Then

(2)

where solves the so-called BRS-equation

. (3)

As an example, here are the answers for the questions posed at the beginning. Thus let all be uniformly distributed on . Then on , and hence on . The BRS-equation becomes

The solution is , and thus from the inequality (2)

. (4)

Since one always has , this bound becomes trivial for .

For the example questions with this yields . As one sees from (4), doubling the sample size and keeping fixed, or vice versa, yield for the uniform distribution in the non-trivial case the same upper bound.

Generalised BRS-inequality

Theorem 2. Let be positive random variables that are jointly distributed such that has an absolutely continuous distribution function . If is defined as before, then

, (5)

where is the unique solution of the BRS-equation

(6)

Clearly, if all random variables have the same marginal distribution , then (6) recaptures (3), and (5) recaptures (2). Again it should be pointed out that no independence hypothesis whatsoever is needed.

Refinements of the BRS-inequality

Depending on the type of the distributions , refinements of Theorem 2 can be of true interest. We just mention one of them.

Let be the random set of those variables one can sum up to yield the maximum random number , that is,

,

and let denote the sum of these variables. The so-called residual is by definition always non-negative, and one has:

Theorem 3. Let be jointly continuously distributed with marginal distribution functions , and let be the solution of (6). Then

. (7)

The improvement in (7) compared with (5) therefore consists of

.

The expected residual in the numerator is typically difficult to compute or estimate, but there exist nice exceptions. For example, if all are independent exponential random variables, then the memoryless property implies (if s is exceeded) the distributional symmetry of the residual and the overshoot over . For fixed one can then show that :. This improvement fluctuates around , and the convergence to , (simulations) seems quick.

Source

The first version of the BRS-inequality (Theorem 1) was proved in Lemma 4.1 of F. Thomas Bruss and James B. Robertson (1991). This paper proves moreover that the upper bound is asymptotically tight if the random variables are independent of each other. The generalisation to arbitrary continuous distributions (Theorem 2) is due to J. Michael Steele (2016). Theorem 3 and other refinements of the BRS-inequality are more recent and proved in Bruss (2021).

Applications

The BRS-inequality is a versatile tool since it applies to many types of problems, and since the computation of the BRS-equation is often not very involved. Also, and in particular, one notes that the maximum number always dominates the maximum number of selections under any additional constraint, such as e.g. for online selections without recall. Examples studied in Steele (2016) and Bruss (2021) touch many applications, including comparisons between i.i.d. sequences and non-i.i.d. sequences, problems of condensing point processes, “awkward” processes, selection algorithms, knapsack problems, Borel-Cantelli-type problems, the Bruss-Duerinckx theorem, and online Tiling strategies.

Related Research Articles

In probability theory, the expected value is a generalization of the weighted average. Informally, the expected value is the arithmetic mean of a large number of independently selected outcomes of a random variable.

In probability theory, the central limit theorem (CLT) establishes that, in many situations, when independent random variables are summed up, their properly normalized sum tends toward a normal distribution even if the original variables themselves are not normally distributed.

In probability theory, Chebyshev's inequality guarantees that, for a wide class of probability distributions, no more than a certain fraction of values can be more than a certain distance from the mean. Specifically, no more than 1/k2 of the distribution's values can be k or more standard deviations away from the mean. The rule is often called Chebyshev's theorem, about the range of standard deviations around the mean, in statistics. The inequality has great utility because it can be applied to any probability distribution in which the mean and variance are defined. For example, it can be used to prove the weak law of large numbers.

In probability theory and statistics, the moment-generating function of a real-valued random variable is an alternative specification of its probability distribution. Thus, it provides the basis of 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 moment-generating functions of distributions defined by the weighted sums of random variables. However, not all random variables have moment-generating functions.

Jensens inequality 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. 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 probability theory, the Azuma–Hoeffding inequality gives a concentration result for the values of martingales that have bounded differences.

In probability theory, the central limit theorem states that, under certain circumstances, the probability distribution of the scaled mean of a random sample converges to a normal distribution as the sample size increases to infinity. Under stronger assumptions, the Berry–Esseen theorem, or Berry–Esseen inequality, gives a more quantitative result, because it also specifies the rate at which this convergence takes place by giving a bound on the maximal error of approximation between the normal distribution and the true distribution of the scaled sample mean. The approximation is measured by the Kolmogorov–Smirnov distance. In the case of independent samples, the convergence rate is n−1/2, where n is the sample size, and the constant is estimated in terms of the third absolute normalized moments.

In probability theory, the Chernoff bound gives exponentially decreasing bounds on tail distributions of sums of independent random variables. Despite being named after Herman Chernoff, the author of the paper it first appeared in, the result is due to Herman Rubin. It is a sharper bound than the known first- or second-moment-based tail bounds such as Markov's inequality or Chebyshev's inequality, which only yield power-law bounds on tail decay. However, the Chernoff bound requires that the variates be independent – a condition that neither Markov's inequality nor Chebyshev's inequality require, although Chebyshev's inequality does require the variates to be pairwise independent.

In probability theory, Hoeffding's inequality provides an upper bound on the probability that the sum of bounded independent random variables deviates from its expected value by more than a certain amount. Hoeffding's inequality was proven by Wassily Hoeffding in 1963.

In information theory, Shannon's source coding theorem establishes the limits to possible data compression, and the operational meaning of the Shannon entropy.

In probability theory and statistics, the Rademacher distribution is a discrete probability distribution where a random variate X has a 50% chance of being +1 and a 50% chance of being -1.

In statistics, an exchangeable sequence of random variables is a sequence X1X2X3, ... whose joint probability distribution does not change when the positions in the sequence in which finitely many of them appear are altered. Thus, for example the sequences

Dvoretzky–Kiefer–Wolfowitz inequality Statistical inequality

In the theory of probability and statistics, the Dvoretzky–Kiefer–Wolfowitz–Massart inequality bounds how close an empirically determined distribution function will be to the distribution function from which the empirical samples are drawn. It is named after Aryeh Dvoretzky, Jack Kiefer, and Jacob Wolfowitz, who in 1956 proved the inequality

In probability theory, Etemadi's inequality is a so-called "maximal inequality", an inequality that gives a bound on the probability that the partial sums of a finite collection of independent random variables exceed some specified bound. The result is due to Nasrollah Etemadi.

Stein's method is a general method in probability theory to obtain bounds on the distance between two probability distributions with respect to a probability metric. It was introduced by Charles Stein, who first published it in 1972, to obtain a bound between the distribution of a sum of -dependent sequence of random variables and a standard normal distribution in the Kolmogorov (uniform) metric and hence to prove not only a central limit theorem, but also bounds on the rates of convergence for the given metric.

In probability theory, the optional stopping theorem says that, under certain conditions, the expected value of a martingale at a stopping time is equal to its initial expected value. Since martingales can be used to model the wealth of a gambler participating in a fair game, the optional stopping theorem says that, on average, nothing can be gained by stopping play based on the information obtainable so far. Certain conditions are necessary for this result to hold true. In particular, the theorem applies to doubling strategies.

In mathematics, the second moment method is a technique used in probability theory and analysis to show that a random variable has positive probability of being positive. More generally, the "moment method" consists of bounding the probability that a random variable fluctuates far from its mean, by using its moments.

In probability theory, concentration inequalities provide bounds on how a random variable deviates from some value. The law of large numbers of classical probability theory states that sums of independent random variables are, under very mild conditions, close to their expectation with a large probability. Such sums are the most basic examples of random variables concentrated around their mean. Recent results show that such behavior is shared by other functions of independent random variables.

For certain applications in linear algebra, it is useful to know properties of the probability distribution of the largest eigenvalue of a finite sum of random matrices. Suppose is a finite sequence of random matrices. Analogous to the well-known Chernoff bound for sums of scalars, a bound on the following is sought for a given parameter t:

In probability theory, Eaton's inequality is a bound on the largest values of a linear combination of bounded random variables. This inequality was described in 1974 by Morris L. Eaton.

References

Bruss F. T. and Robertson J. B. (1991) ’Wald’s Lemma’ for Sums of Order Statistics of i.i.d. Random Variables, Adv. Appl. Probab., Vol. 23, 612-623.

Bruss F. T. and Duerinckx M. (2015), Resource dependent branching processes and the envelope of societie, Ann. of Appl. Probab., Vol. 25 (1), 324-372.

Steele J.M. (2016), The Bruss-Robertson Inequality: Elaborations, Extensions, and Applications, Math. Applicanda, Vol. 44, No 1, 3-16.

Bruss F. T. (2021),The BRS-inequality and its applications, Probab. Surveys, Vol.18, 44-76.