Home Recurrence relations for the joint distribution of the sum and maximum of independent random variables
Article
Licensed
Unlicensed Requires Authentication

Recurrence relations for the joint distribution of the sum and maximum of independent random variables

  • Christos N. Efrem ORCID logo EMAIL logo
Published/Copyright: August 3, 2024

Abstract

In this paper, the joint distribution of the sum and maximum of independent, not necessarily identically distributed, nonnegative random variables is studied for two cases: (i) continuous and (ii) discrete random variables. First, a recursive formula of the joint cumulative distribution function (CDF) is derived in both cases. Then recurrence relations of the joint probability density function (PDF) and the joint probability mass function (PMF) are given in the former and the latter case, respectively. Interestingly, there is a fundamental difference between the joint PDF and PMF. The proofs are simple and mainly based on the following tools from calculus and discrete mathematics: differentiation under the integral sign (also known as Leibniz’s integral rule), the law of total probability, and mathematical induction. In addition, this work generalizes previous results in the literature, and finally presents several extensions of the methodology.

A Partial differentiation under the integral sign

In this part of the paper, we provide a generalization of the Leibniz integral rule for partial differentiation of integrals with variable limits. For this purpose, the following two lemmas will be instrumental.

Lemma A.1 (Leibniz’s integral rule with constant limits of integration, [16, p. 276, Theorem 7.1]).

Let f ( x , t ) , t f ( x , t ) be defined and continuous on X × T , where X [ x 1 , x 2 ] and T [ t 1 , t 2 ] are subsets of R with - < x 1 x 2 < and - < t 1 < t 2 < . Moreover, let a , b X be two constants. Then, for all t T ,

(A.1) d d t ( a b f ( x , t ) d x ) = a b t f ( x , t ) d x .

Lemma A.2 (Leibniz’s integral rule with variable limits of integration).

Let f ( x , t ) and t f ( x , t ) be defined and continuous on X × T , where X [ x 1 , x 2 ] and T [ t 1 , t 2 ] are subsets of R with - < x 1 x 2 < and - < t 1 < t 2 < . In addition, let a , b : T X be differentiable functions on T . Then, for all t T ,

(A.2) d d t ( a ( t ) b ( t ) f ( x , t ) d x ) = f ( x , t ) | x = b ( t ) d d t b ( t ) - f ( x , t ) | x = a ( t ) d d t a ( t ) + a ( t ) b ( t ) t f ( x , t ) d x .

Proof.

This is a consequence of the fundamental theorem of calculus, the basic form of Leibniz’s integral rule (Lemma A.1), and the chain rule for multivariate functions. Let us define the function I ( u , v , t ) v u f ( x , t ) d x for all ( u , v , t ) 𝒳 2 × 𝒯 . From the fundamental theorem of calculus, we have

u I ( u , v , t ) = f ( u , t ) and v I ( u , v , t ) = - f ( v , t ) ,

while t I ( u , v , t ) = v u t f ( x , t ) d x due to Lemma A.1. Now, by applying the chain rule, we obtain

d d t ( a ( t ) b ( t ) f ( x , t ) d x ) = d d t I ( b ( t ) , a ( t ) , t ) = u I ( u , v , t ) | u = b ( t ) , v = a ( t ) d d t b ( t )
+ v I ( u , v , t ) | u = b ( t ) , v = a ( t ) d d t a ( t ) + t I ( u , v , t ) | u = b ( t ) , v = a ( t ) d d t t
= f ( u , t ) | u = b ( t ) d d t b ( t ) - f ( v , t ) | v = a ( t ) d d t a ( t ) + a ( t ) b ( t ) t f ( x , t ) d x

for all t 𝒯 . ∎

Proposition A.3.

Let h ( x , y , z ) , z h ( x , y , z ) and 2 y z h ( x , y , z ) be defined and continuous on S X × Y × Z , where X [ x 1 , x 2 ] , Y [ y 1 , y 2 ] , and Z [ z 1 , z 2 ] are subsets of R with - < x 1 x 2 < , - < y 1 < y 2 < , and - < z 1 < z 2 < . Furthermore, suppose that y h ( x , y , z ) exists on S , and a , b : Z X are differentiable functions on Z . Then, for all ( y , z ) Y × Z ,

(A.3)

2 y z ( a ( z ) b ( z ) h ( x , y , z ) d x ) = y h ( x , y , z ) | x = b ( z ) d d z b ( z ) - y h ( x , y , z ) | x = a ( z ) d d z a ( z )
+ a ( z ) b ( z ) 2 y z h ( x , y , z ) d x .

Proof.

Firstly, we will compute the partial derivative of a ( z ) b ( z ) h ( x , y , z ) d x with respect to the variable z. By virtue of Lemma A.2 (extended to functions of three variables), we obtain

z ( a ( z ) b ( z ) h ( x , y , z ) d x ) = h ( x , y , z ) | x = b ( z ) d d z b ( z ) - h ( x , y , z ) | x = a ( z ) d d z a ( z ) + a ( z ) b ( z ) z h ( x , y , z ) d x

for all ( y , z ) 𝒴 × 𝒵 . Secondly, by differentiating the previous equation with respect to the variable y, the second-order mixed partial derivative of a ( z ) b ( z ) h ( x , y , z ) d x is given by

2 y z ( a ( z ) b ( z ) h ( x , y , z ) d x ) = y h ( x , y , z ) | x = b ( z ) d d z b ( z ) - y h ( x , y , z ) | x = a ( z ) d d z a ( z )
+ y ( a ( z ) b ( z ) z h ( x , y , z ) d x ) .

Finally, we conclude the proof by applying Lemma A.1 (extended to functions of three variables) to the last integral. ∎

Remark A.4.

Proposition A.3 remains valid if we replace the assumption that 2 y z h ( x , y , z ) is continuous on 𝒮 with the assumption that a ( z ) b ( z ) 2 y z h ( x , y , z ) d x converges uniformly on 𝒴 × 𝒵 . This follows from the fact that Lemma A.1 still holds when the continuity of t f ( x , t ) on 𝒳 × 𝒯 is replaced by the uniform convergence of a b t f ( x , t ) d x on 𝒯 ; see [25] and [24, p. 260].

Next, we give modified versions of Lemmas A.1, A.2 and Proposition A.3 by relaxing the continuity assumptions (some results now hold almost everywhere). Note that the new conditions are weaker than continuity, because if a real-valued function is continuous on a compact set, then it is bounded and so integrable; in addition, there always exist constant functions ϑ ( x ) , φ ( x , y ) and ϕ ( x ) satisfying the assumptions in the following results.

Lemma A.5 (Leibniz’s integral rule with constant limits of integration – modified version [11, p. 56, Theorem 2.27]).

Let f ( x , t ) and t f ( x , t ) be defined on X × T , where X [ x 1 , x 2 ] and T [ t 1 , t 2 ] are subsets of R with - < x 1 x 2 < and - < t 1 < t 2 < . Let a , b X be two constants. Moreover, assume that:

  1. the function f ( x , t ) is integrable, i.e., a b | f ( x , t ) | d x < for all t 𝒯 ,

  2. there exists a function ϑ ( x ) such that | t f ( x , t ) | ϑ ( x ) for all ( x , t ) 𝒳 × 𝒯 , and a b ϑ ( x ) d x < .

Then (A.1) holds for all t T .

Lemma A.6 (Leibniz’s integral rule with variable limits of integration – modified version).

Let f ( x , t ) and t f ( x , t ) be defined on X × T , where X [ x 1 , x 2 ] and T [ t 1 , t 2 ] are subsets of R with - < x 1 x 2 < and - < t 1 < t 2 < . Let a , b : T X be differentiable functions on T . In addition, assume that:

  1. the function f ( x , t ) is integrable, i.e., a ( t ) b ( t ) | f ( x , t ) | d x < for all t 𝒯 ,

  2. there exists a function ϑ ( x ) such that | t f ( x , t ) | ϑ ( x ) for all ( x , t ) 𝒳 × 𝒯 , and a ( t ) b ( t ) ϑ ( x ) d x < for all t 𝒯 .

Then (A.2) holds for almost all t T .

Proof.

This result follows from similar steps as in the proof of Lemma A.2, but exploiting the fundamental theorem of calculus for Lebesgue integrals [11, pp. 105–106, Corollary 3.33 and Theorem 3.35] (instead of the classical one) as well as Lemma A.5 (instead of Lemma A.1). In particular, the fundamental theorem of calculus for Lebesgue integrals implies that u I ( u , v , t ) = a.e. f ( u , t ) and v I ( u , v , t ) = a.e. - f ( v , t ) . Therefore, (A.2) is true for almost every t 𝒯 . ∎

Proposition A.7.

Let h ( x , y , z ) , z h ( x , y , z ) , and 2 y z h ( x , y , z ) be defined on S X × Y × Z , where X [ x 1 , x 2 ] , Y [ y 1 , y 2 ] , and Z [ z 1 , z 2 ] are subsets of R with - < x 1 x 2 < , - < y 1 < y 2 < , and - < z 1 < z 2 < . Suppose that y h ( x , y , z ) exists on S , and a , b : Z X are differentiable functions on Z . Furthermore, assume that:

  1. the function h ( x , y , z ) is integrable, i.e., a ( z ) b ( z ) | h ( x , y , z ) | d x < for all ( y , z ) 𝒴 × 𝒵 ,

  2. there exists a function φ ( x , y ) such that | z h ( x , y , z ) | φ ( x , y ) for all ( x , y , z ) 𝒮 , and a ( z ) b ( z ) φ ( x , y ) d x < for all ( y , z ) 𝒴 × 𝒵 ,

  3. there is a function ϕ ( x ) such that | 2 y z h ( x , y , z ) | ϕ ( x ) for all ( x , y , z ) 𝒮 , and a ( z ) b ( z ) ϕ ( x ) d x < for all z 𝒵 .

Then (A.3) holds for almost all ( y , z ) Y × Z .

Proof.

In the proof of Proposition A.3, we make the following substitutions: Lemma A.1 Lemma A.5 and Lemma A.2 Lemma A.6, while the two equations with derivatives are valid almost everywhere now. Observe that the second assumption implies that a ( z ) b ( z ) | z h ( x , y , z ) | d x < for all ( y , z ) 𝒴 × 𝒵 , which is required to apply Lemma A.5 in the final step. ∎

B Dirac delta function: Definition and properties

Loosely speaking, the Dirac delta function/distribution δ ( x ) can be regarded as a function defined on that is zero everywhere except at x = 0 , where it is infinite, and also satisfies the following properties: δ ( - x ) = δ ( x ) 0 , - + δ ( t ) d t = - ε + ε δ ( t ) d t = 1 , - + f ( t ) δ ( x - t ) d t = x - ε x + ε f ( t ) δ ( x - t ) d t = f ( x ) , and f ( x ) δ ( x - y ) = f ( y ) δ ( x - y ) for all x , y , ε + and for all functions f ( x ) that are bounded and piecewise continuous on . Note that δ ( x ) is not a function in the conventional sense, because no function has these properties. Nevertheless, the Dirac delta function can be rigorously defined as a generalized function or distribution. Strictly speaking, the Dirac delta function δ ( x ) is the limit (in the sense of distributions) of a Dirac sequence { δ k ( x ) } k , that is, δ ( x ) = lim k δ k ( x ) .

Definition B.1 ([16, p. 284]).

A Dirac sequence is a sequence of functions { δ k ( x ) } k that satisfy three properties:

  1. δ k ( - x ) = δ k ( x ) 0 for all k and for all x ,

  2. δ k ( x ) is continuous on and - + δ k ( x ) d x = 1 for all k ,

  3. for any ϵ , ζ > 0 , there exists K such that - - ζ δ k ( x ) d x + ζ + δ k ( x ) d x < ϵ for all integers k K .

In simple words, (P3) means that the area under the curve y = δ k ( x ) is concentrated around x = 0 for sufficiently large k. For example, the sequence of zero-mean normal/Gaussian distributions δ k ( x ) = k π e - ( k x ) 2 is a Dirac sequence.

The following proposition is an important result in analysis about the convolution of a function with a Dirac sequence; the convolution of two functions f ( x ) and g ( x ) is defined by f * g - + f ( t ) g ( x - t ) d t and is commutative, that is, f * g = g * f .

Proposition B.2 ([16, p. 285, Theorem 1.1]).

Suppose that f ( x ) is a bounded and piecewise-continuous function on R , D is a compact subset of R on which f ( x ) is continuous, and { δ k ( x ) } k N is a Dirac sequence. Then the sequence of functions { f k ( x ) } k N , where f k ( x ) f * δ k = - + f ( t ) δ k ( x - t ) d t , converges to f ( x ) uniformly on D .

Therefore, if f ( x ) is bounded and piecewise continuous on , then - + f ( t ) δ ( x - t ) d t = f ( x ) rigorously means that: for each Dirac sequence { δ k ( x ) } k , lim k - + f ( t ) δ k ( x - t ) d t = f ( x ) uniformly on every compact subset of where f ( x ) is continuous.

Acknowledgements

The author thanks the Editor-in-Chief for handling the review process of the paper, and the anonymous referees for their useful comments and suggestions.

References

[1] M.-S. Alouini, A. Abdi and M. Kaveh, Sum of gamma variates and performance of wireless communication systems over Nakagami-fading channels, IEEE Trans. Vehicular Technol. 50 (2001), no. 6, 1471–1480. 10.1109/25.966578Search in Google Scholar

[2] C. W. Anderson and K. F. Turkman, The joint limiting distribution of sums and maxima of stationary sequences, J. Appl. Probab. 28 (1991), no. 1, 33–44. 10.2307/3214738Search in Google Scholar

[3] C. W. Anderson and K. F. Turkman, Limiting joint distributions of sums and maxima in a statistical context, Theory Probab. Appl. 37 (1992), no. 2, 314–316. 10.1137/1137063Search in Google Scholar

[4] M. Arendarczyk, T. J. Kozubowski and A. K. Panorska, The joint distribution of the sum and maximum of dependent Pareto risks, J. Multivariate Anal. 167 (2018), 136–156. 10.1016/j.jmva.2018.04.002Search in Google Scholar

[5] M. Arendarczyk, T. J. Kozubowski and A. K. Panorska, The joint distribution of the sum and the maximum of heterogeneous exponential random variables, Statist. Probab. Lett. 139 (2018), 10–19. 10.1016/j.spl.2018.03.013Search in Google Scholar

[6] N. Balakrishnan, Recurrence relations for order statistics from n independent and non-identically distributed random variables, Ann. Inst. Statist. Math. 40 (1988), no. 2, 273–277. 10.1007/BF00052344Search in Google Scholar

[7] N. Balakrishnan, S. M. Bendre and H. J. Malik, General relations and identities for order statistics from nonindependent nonidentical variables, Ann. Inst. Statist. Math. 44 (1992), no. 1, 177–183. 10.1007/BF00048680Search in Google Scholar

[8] K. Butler and M. A. Stephens, The distribution of a sum of independent binomial random variables, Methodol. Comput. Appl. Probab. 19 (2017), no. 2, 557–571. 10.1007/s11009-016-9533-4Search in Google Scholar

[9] T. L. Chow and J. L. Teugels, The sum and the maximum of i.i.d. random variables, Proceedings of the Second Prague Symposium on Asymptotic Statistics, North-Holland, Amsterdam (1979), 81–92. Search in Google Scholar

[10] H. A. David and H. N. Nagaraja, Order Statistics, Wiley Ser. Probab. Stat., John Wiley & Sons, Hoboken, 2003. 10.1002/0471722162Search in Google Scholar

[11] G. B. Folland, Real Analysis: Modern Techniques and Their Applications, 2nd ed., Pure Appl. Math., John Wiley & Sons, New York, 1999. Search in Google Scholar

[12] H.-C. Ho and T. Hsing, On the asymptotic joint distribution of the sum and maximum of stationary normal random variables, J. Appl. Probab. 33 (1996), no. 1, 138–145. 10.2307/3215271Search in Google Scholar

[13] T. Hsing, A note on the asymptotic independence of the sum and maximum of strongly mixing stationary random variables, Ann. Probab. 23 (1995), no. 2, 938–947. 10.1214/aop/1176988296Search in Google Scholar

[14] T. Hsing, On the asymptotic independence of the sum and rare values of weakly dependent stationary random variables, Stochastic Process. Appl. 60 (1995), no. 1, 49–63. 10.1016/0304-4149(95)00054-2Search in Google Scholar

[15] H. V. Khuong and H.-Y. Kong, General expression for pdf of a sum of independent exponential random variables, IEEE Commun. Lett. 10 (2006), no. 3, 159–161. 10.1109/LCOMM.2006.1603370Search in Google Scholar

[16] S. Lang, Undergraduate Analysis, Undergrad. Texts Math., Springer, New York, 1997. 10.1007/978-1-4757-2698-5Search in Google Scholar

[17] Y. Lazar and B. Almutairi, Dirac distributions related to sums of independent nonidentically uniform random variables, Braz. J. Probab. Stat. 35 (2021), no. 3, 435–441. 10.1214/20-BJPS484Search in Google Scholar

[18] W. P. McCormick and Y. Qi, Asymptotic distribution for the sum and maximum of Gaussian processes, J. Appl. Probab. 37 (2000), no. 4, 958–971. 10.1239/jap/1014843076Search in Google Scholar

[19] M. Morrison and F. Tobias, Some statistical characteristics of a peak to average ratio, Technometrics 7 (1965), 379–385. 10.1080/00401706.1965.10490270Search in Google Scholar

[20] P. G. Moschopoulos, The distribution of the sum of independent gamma random variables, Ann. Inst. Statist. Math. 37 (1985), no. 3, 541–544. 10.1007/BF02481123Search in Google Scholar

[21] E. G. Olds, A note on the convolution of uniform distributions, Ann. Math. Statistics 23 (1952), 282–285. 10.1214/aoms/1177729446Search in Google Scholar

[22] Z. Peng and S. Nadarajah, On the joint limiting distribution of sums and maxima of stationary normal sequences, Theory Probab. Appl. 47 (2002), no. 4, 706–709. 10.1137/S0040585X97980142Search in Google Scholar

[23] F. Qeadan, T. J. Kozubowski and A. K. Panorska, The joint distribution of the sum and the maximum of IID exponential random variables, Comm. Statist. Theory Methods 41 (2012), no. 3, 544–569. 10.1080/03610926.2010.529524Search in Google Scholar

[24] K. Rogers, Advanced Calculus, Merrill, Columbus, 1976. Search in Google Scholar

[25] E. Talvila, Necessary and sufficient conditions for differentiating under the integral sign, Amer. Math. Monthly 108 (2001), no. 6, 544–548. 10.1080/00029890.2001.11919782Search in Google Scholar

[26] J. A. Woodward and C. G. S. Palmer, On the exact convolution of discrete random variables, Appl. Math. Comput. 83 (1997), no. 1, 69–77. 10.1016/S0096-3003(96)00047-1Search in Google Scholar

Received: 2024-01-08
Revised: 2024-06-15
Accepted: 2024-06-20
Published Online: 2024-08-03
Published in Print: 2025-06-01

© 2024 Walter de Gruyter GmbH, Berlin/Boston

Downloaded on 17.9.2025 from https://www.degruyterbrill.com/document/doi/10.1515/jaa-2024-0004/html
Scroll to top button