arXiv:1504.05938v1 [math.PR] 22 Apr 2015

NEW BERRY-ESSEEN AND WASSERSTEIN BOUNDS IN THE CLT FOR NON-RANDOMLY CENTERED RANDOM SUMS BY PROBABILISTIC METHODS CHRISTIAN DÖBLER Abstract. We prove abstract bounds on the Wasserstein and Kolmogorov distances between non-randomly centered random sums of real i.i.d. random variables with a finite third moment and the standard normal distribution. Except for the case of mean zero summands, these bounds involve a coupling of the summation index with its size biased distribution as was previously considered in [GR96] for the normal approximation of nonnegative random variables. When being specialized to concrete distributions of the index like the Binomial, Poisson and Hypergeometric distribution, our bounds turn out to be of the correct order of magnitude.

1. Introduction Let N, X1 , X2 , . . . be random variables on a common probability space such that the Xj , j ≥ 1, are real-valued and N assumes values in the nonnegative integers Z+ = {0, 1, . . . }. Then, the random variable (1)

S :=

N X

Xj

j=1

is called a random sum. Such random variables appear frequently in modern probabiliy theory, as many models for example from physics, finance, reliability and risk theory naturally lead to the consideration of such sums. Furthermore, sometimes a model, which looks quite different from (1) at the outset, may be transformed into a random sum and then general theory of such sums may be invoked to study the original model [GK96]. There already exists a huge body of literature about the asymptotic distributions of random sums. Their investigation evidently began with the work [Rob48] of Robbins, who assumes that the random variables X1 , X2 , . . . are i.i.d. with a finite second moment and that N also has a finite second moment. One of the results of [Rob48] is that under these assumptions asymptotic normality of the index N automatically implies asymptotic normality of the corresponding random sum. The book [GK96] gives a comprehensive description of the limiting behaviour of such random sums under the assumption that the random variables N, X1 , X2 , . . . are independent. In particular, one may ask under what conditions the sum S in (1) is asymptotically normal, where asymptotically refers to the fact that the random index N in fact usually depends on a parameter, which is send either to infinity or to zero. Once a CLT is known to hold, one might ask about the accuracy of the normal approximation to the distribution of the given random sum. It turns out that it Université du Luxembourg, Unité de Recherche en Mathématiques [email protected] Keywords: random sums, central limit theorem, Kolmogorov distance, zero bias couplings, size bias couplings.

2

CHRISTIAN DÖBLER

is generally much easier to derive rates of convergence for random sums of centered random variables, or, which amounts to the same thing, for random sums centered by random variables than for random sums of not necessarily centered random variables. In the centered case one might, for instance, first condition on the value of the index N, then use known error bounds for sums of a fixed number of independent random variables like the classical Berry-Esseen theorem and, finally, take expectation with respect to N. This technique is illustrated e.g. in the manuscript [Döb12] and also works for non-normal limiting distributions like the Laplace distribution. For this reason we will mainly be interested in deriving sharp rates of convergence for the case of non-centered summands, but will also consider the mean-zero case and hint at the relevant differences. Also, we will not assume from the outset that the index N has a certain fixed distribution like the Binomial or the Poisson, but will be interested in the general situation. For non-centered summands and general index N, the relevant literature on rates of convergence in the random sums CLT seems quite easy to survey. Under the same assumptions as in [Rob48] the paper [Eng83] gives an upper bound on the Kolmogorov distance between the distribution of the random sum and a suitable normal distribution, which is proved to be sharp in some sense. However, this bound is not very explicit as it contains the Kolmogorov distance of N to the normal distribution with the same mean and variance as N as one of the terms appearing in the bound, for instance. This might make the task of applying this result difficult for a concrete distribution of N. Furthermore, the method of proof cannot be easily adapted to probability metrics different from the Kolmogorov distance like e.g. the Wasserstein distance. In [Kor87] a bound on the Kolmogorov distance is given which improves upon the result of [Eng83] with respect to the constants appearing in the bound. However, the bound given in [Kor87] is no longer strong enough to assure the well-known asymptotic normality of Binomial and Poisson random sums, unless the summands are centered. To the best of our knowledge, the article [Sun14] is the only one, which gives bounds on the Wasserstein distance between random sums for general indices N and the standard normal distribution. However, as mentioned by the same author in [Sun14], the results of [Sun13] generally do not yield accurate bounds, unless the summands are centered. Indeed, the results from [Sun13] do not even yield convergence in distribution for Binomial random sums of non-centered summands. The main purpose of the present article is to combine Stein’s method of normal approximation with several modern probabilistic concepts like certain coupling constructions and conditional independence, to prove accurate abstract upper bounds on the distance between suitably standardized random sums of i.i.d. summands measured by two popular probability metrics, the Kolmogorov and Wasserstein distances. These upper bounds, in their most abstract forms (see Theorem 2.5 and Corollary 2.8 below), involve moments of the difference of a coupling of N with its size-biased distribution but reduce to very explicit expressions if either N has a concrete distribution like the Binomial, Poisson or dirac delta distribution, the summands Xj are centered, or, if the distribution of N is infinitely divisible. These special cases are extensively presented in order to illustrate the wide applicability and strength of our results. As indicated above, this seems to be the first work which gives Wasserstein bounds in the random sums CLT for general indices N, which reduce to bounds of optimal order, when specializing to concrete distributions like the Binomial and the

ERROR BOUNDS IN THE CLT FOR RANDOM SUMS

3

Poisson distributions. Using our abstract approach via size-bias couplings, we are also able to prove rates for Hypergeometric random sums. These do not seem to have been treated in the literature, yet. This is not a surprise, because the Hypergeometric distribution is conceptually more complicated than the Binomial or Poisson distribution, as it is neither a natural convolution of i.i.d. random variables nor infinitely divisible. It should be mentioned that Stein’s method and coupling techniques have previously been used to bound the error of exponential approximation [PR11] and approximation by the Laplace distribution [PR12] of certain random sums. The remainder of the article is structured as follows: In Section 2 we review the relevant probability distances, the size biased distribution and state our quantitative results on the normal approximation of random sums. Furthermore, we prove new identities for the distance of a nonnegative random variable to its size-biased distribution in three prominent distances and show that for some concrete distributions, natural couplings are L1 -optimal and, hence, yield the Wasserstein distance. In Section 3 we collect necessary facts from Stein’s method of normal approximation and introduce a variant of the zero-bias transformation, which we need for the proofs of our results. Then, in Section 4, the proof of our main theorems, Theorem 2.5 and Theorem 2.6 is given. Finally, Section 5 contains the proofs of some auxiliary results, needed for the proof in Section 4. 2. Main results Recall that for probability measures µ and ν on (R, B(R)), their Kolmogorov distance is defined by   dK (µ, ν) := sup µ (−∞, z] − µ (−∞, z] = kF − Gk∞ , z∈R

where F and G are the distribution functions corresponding to µ and µ, respectively. Also, if both µ and ν have finite first absolute moment, then one defines the Wasserstein distance between them via Z Z dW (µ, ν) := sup hdµ − hdν , h∈Lip(1)

where Lip(1) denotes the class of all Lipschitz-continous functions g on R with Lipschitz constant not greater than 1. In view of Lemma 2.1 below, we also introduce the total variation distance bewtween µ and ν by dT V (µ, ν) := sup µ(B) − ν(B) . B∈B(R)

If the real-valued random variables X and Y have distributions µ and ν, respectively,  then we simply write dK (X, Y ) for dK L(X), L(Y ) and similarly for the Wasserstein and total variation distances and also speak of the respective distance between the random variables X and Y . Before stating our results, we have to review the concept of the size-biased distribution corresponding to a distribution supported on [0, ∞). Thus, if X is a nonnegative random variable with 0 < E[X] < ∞, then a random variable X s is said to have the X-size biased distribution, if for all bounded and measurable functions h on [0, ∞) (2)

E[Xh(X)] = E[X]E[h(X s )] ,

4

CHRISTIAN DÖBLER

see, e.g. [GR96], [AG10] or [AGK13]. Equivalently, the distribution of X s has RadonNikodym derivative with respect to the distribution of X given by P (X s ∈ dx) x = , P (X ∈ dx) E[X]

which immediately implies both existence and uniqueness of the X-size biased distribution. Also note that (2) holds true for all measurable functions h for which E|Xh(X)| < ∞. In consequence, if X ∈ Lp (P ) for some 1 ≤ p < ∞, then X s ∈ Lp−1 (P ) and h i E X p   p−1 E Xs = . E[X] The following lemma, which seems to be new and might be of independent interest, gives identities for the distance of X to X s in the three metrics mentioned above. The proof is deferred to the end of this section. Lemma 2.1. Let X be a nonnegative random variable such that 0 < E[X] < ∞. Then, the following identities hold true:   E (X − E[X])1 {X>E[X]} (a) dK (X, X s ) = E[X] E|X − E[X]| (b) dT V (X, X s ) = 2E[X] Var(X) (c) If additionally E[X 2 ] < ∞, then dW (X, X s ) = . E[X] Remark 2.2. (a) It is well known (see e.g. [Dud02]) that the Wasserstein distance dW (X, Y ) between the real random variables X and Y has the dual representation (3)

dW (X, Y ) =

inf

ˆ Yˆ )∈π(X,Y ) (X,

ˆ − Yˆ | , E|X

ˆ Yˆ ) where π(X, Y ) is the collection of all couplings of X and Y , i.e. of all pairs (X, D D ˆ = of random variables on a joint probability space such that X X and Yˆ = Y . Also, the infimum in (3) is always attained, e.g. by the quantile transformation: If U is uniformly distributed on (0, 1) and if, for a distribution function F on R, we let F −1 (p) := inf{x ∈ R : F (x) ≥ p}, p ∈ (0, 1) ,

denote the corresponding generalized inverse of F , then F −1 (U) is a random variable with distribution function F . Thus, letting FX and FY denote the distribution functions of X and Y , respectively, it was proved e.g. in [Maj78] that Z 1 −1 −1 ˆ ˆ inf E|X − Y | = E|FX (U) − FY (U)| = |FX−1 (t) − FY−1 (t)|dt . ˆ Yˆ )∈π(X,Y ) (X,

0

Furthermore, it is not difficult to see that X s is always stochastically larger than ˆ Xˆs ) of X and X s such that Xˆs ≥ X ˆ X, implying that there is a coupling (X, (see [AG10] for details). In fact, this property is already achieved by the coupling via the quantile transformation. By the dual representation (3) and the fact that the coupling via the quantile transformation yields the minimum L1 distance in

ERROR BOUNDS IN THE CLT FOR RANDOM SUMS

5

ˆ Xˆs ) such that Xˆs ≥ X ˆ is optimal (3) we can conclude that every coupling (X, in this sense, since         ˆ = E Xˆs − E X ˆ = E F −1s (U) − E F −1 (U) E Xˆs − X X X = E FX−1s (U) − FX−1 (U) = dW (X, X s ) .

(b) Due to a result by Steutel [Ste73], the distribution of X is infinitely divisible, if and only if there exists a coupling (X, X s ) of X and X s such that X s − X is nonnegative and independent of X (see e.g. [AG10] for a nice exposition and a proof of this result). According to (a) such a coupling always achieves the minimum L1 -distance. example 2.3. (a) Let X ∼ Poisson(λ) have the Poisson distribution with paramter λ > 0. From the Stein characterization of Poisson(λ) (see [Che75]) it is known that E[Xf (X)] = λE[f (X + 1)] = E[X]E[f (X + 1)] for all bounded and measurable f . Hence, X + 1 has the X-size biased distribution. As X +1 ≥ X, by Remark 2.2 this coupling yields the minimum L1 -distance between X and X s , which is equal to 1 in this case. (b) Let n be a positive integer, p ∈ (0, 1] and let X1 , . . . , Xn be i.i.d. random variables such that X1 ∼ Bernoulli(p). Then, n X X := Xj ∼ Bin(n, p) j=1

has the Binomial distribution with parameters n and p. From the construction in [GR96] one easily sees that n X s Xj + 1 X := j=2

has the X-size biased distribution. As X s ≥ X, by Remark 2.2 this coupling yields the minimum L1 -distance between X and X s , which is equal to Var(X) dW (X, X s ) = E[1 − X1 ] = 1 − p = E[X] in accordance with Lemma 2.1. (c) Let n, r, s be positive integers such that n ≤ r + s and let X ∼ Hyp(n; r, s) have the Hypergeometric distribution with parameters n, r and s, i.e.  s  r P (X = k) =

Then, E[X] =

nr r+s

k

and, hence,

kP (X = k) P (X s = k) = = E[X] Thus, D

n−k  r+s n

s k r r k n−k  n r+s r+s n





,

=

k = 0, 1, . . . , n .

r−1 k−1



s n−1−(r−1)  r−1+s n−1



,

k = 1, 2, . . . , n .

X s = Y + 1 , where Y ∼ Hyp(n − 1; r − 1, s) . Imagaine un urn with r red and s silver balls. If we draw n times without replacement from this urn and denote by X the total number of drawn red balls,

6

CHRISTIAN DÖBLER

then X ∼ Hyp(n; r, s). For j = 1, . . . , n denote by Xj the P indicator of the event that a red ball is drawn at the j-th draw. Then, X = nj=1 Xj . Also, fix one of the red balls in the urn and, for j = 2, . . . , n, denote by Yj the indicator of the event that at the j-th draw this fixed red ball is drawn. Then, it is not difficult to see that n n X X Y := 1{X1 =1} Xj + 1{X1 =0} (Xj − Yj ) ∼ Hyp(n − 1; r − 1, s) j=2

j=2

and, hence,

s

X := Y + 1 = 1{X1 =1}

n X

Xj + 1{X1 =0}

j=2

n X j=2

(Xj − Yj ) + 1

n   X Yj = 1{X1 =1} X + 1{X1 =0} X + 1 − j=2

P has the X-size biased distribution. Note that since nj=2 Yj ≤ 1 we have n   X s Yj ≥ 0 , X − X = 1{X1 =0} 1 − j=2

and consequently, by Remark 2.2 (a), the coupling (X, X s ) is optimal in the L1 -sense and yields the Wasserstein distance between X and X s : Var(X) n r s r+s−n s(r + s − n) r+s−1 dW (X, X s ) = E X s − X = = r+s r+s . = nr E[X] (r + s)(r + s − 1) r+s

We now turn back to the asymptotic behaviour of random sums. We will rely on the following general assumptions and notation, which we adopt and extend from [Rob48]. Assumption 2.4. The random variables N, X1 , X2 , . . . are independent, X1 , X2 , . . . being i.i.d. and such that E|X1 |3 < ∞ and E[N 3 ] < ∞. Furthermore, we let α := E[N],

β 2 := E[N 2 ],

a := E[X1 ],

b2 := E[X12 ],

γ 2 := Var(N) = β 2 − α2 , c2 := Var(X1 ) = b2 − a2

δ 3 := E[N 3 ], 3 and ξ := E X1 − E[X1 ] .

By Wald’ s equation and the Blackwell-Girshick formula, from Assumption 2.4 we have (4)

µ := E[S] = αa and σ 2 := Var(S) = αc2 + a2 γ 2 .

The main purpose of this paper is to assess the accuracy of the standard normal approximation to the normalized version S−µ S − αa (5) W := =p σ αc2 + aγ 2

of S measured by the Kolmogorov and the Wasserstein distance, respectively. As can be seen from the paper [Rob48], under the general assumption that σ 2 = αc2 + a2 γ 2 → ∞ ,

there are four typical situations in which W is asymptotically normal, which we will now briefly review.

ERROR BOUNDS IN THE CLT FOR RANDOM SUMS

7

1) a = 0 6= c and γ = o(α) 2) c 6= 0 6= a and γ 2 = o(α): 3) c 6= 0 and N itself is asymptotically normal 4) c = 0 6= a and N itself is asymptotically normal We remark that 1) and 2) roughly mean that N tends to infinity in a certain sense, but such that it only fluctuates slightly around its mean α. For instance, this is the case, whenever N is roughly equal to α. Of course, 4) is quite uninteresting, here, since in this case S = aN a.s. Although we do not exclude such cases explicitly, we will taciturnly assume that c 6= 0 in what follows. Theorem 2.5. Let Assumption 2.4 hold, let W be given by (5) and let Z have the standard normal distribution Also, let (N, N s ) be a coupling of N and N s having the N-size biased distribution and define D := N s − N. Then, r q  2c2 bγ 2 3αξ αa2 2 dW (W, Z) ≤ Var E[D | N] + + σ3 σ3 σ2 π 2 2  α|a|b 2αa b  2 E 1 D + E[D 2 ] and + {D 0. Then, 1  2c2 3ξ |a|  dW (W, Z) ≤ √ and + 3 + b b λ b2 √ √  ξ 1 2π (3 2π + 4)ξ c3  7 √ 2 + 3 dK (W, Z) ≤ √ +1+ + + 4 8b3 b3 2 cb2 λ √  |a| |a| c2 c |a|( 2π + 4 + 8ξ) + √ + √ + 2 e−λ + e−λ/2 . + 8b b b c 2π b 2π Proof. We apply the result of Corollary 2.8. In this case, by Example 2.3 (a), we can choose D = 1, yielding that E[D 2 ] = 1 and

 Var E[D|N] = 0 .

Note that

  q   E 1{N ≥1} N −1/2 ≤ E 1{N ≥1} N −1

by Jensen’s ineqaulity. Also, using k + 1 ≤ 2k for all k ∈ N, we can bound ∞ ∞ ∞ X X X   λk λk+1 2 λk ≤ 2e−λ = e−λ E 1{N ≥1} N −1 = e−λ kk! (k + 1)k! λ (k + 1)! k=1 k=1 k=1 ∞

X λl 2 2 = e−λ ≤ . λ l! λ l=2

Hence, 

E 1{N ≥1} N Noting that

−1/2



√ 2 ≤√ . λ

α = γ 2 = λ and σ 2 = λ(a2 + c2 ) = λb2 , the result follows from Corollary 2.8.  Remark 2.14. The Berry-Esseen bound presented in Corollary 2.13 is of the same order of λ as the bound given in [KS12], which seems to be the best currently available, but has a worst constant. However, it should be mentioned that the bound in [KS12] was obtained using special properties of the Poisson distribution and does not seem likely to be easily transferable to other distributions of N. Corollary 2.15. In addition to the assumptions from Theorem 2.5 suppose that N ∼ Bin(n, p) has the Binomial distribution with parameters n ∈ N and p ∈ (0, 1].

12

CHRISTIAN DÖBLER

Then, 

1

 2 2 dW (W, Z) ≤ √ 3/2 2c b + |a|b (1 − p) + 3ξ np b2 − pa2 r  p 2 2 p 2 2 and a p b − pa 1 − p + π √ √  √ 1 ( 2π + 4)bc2 1 − p (3 2π + 4)ξ 3 dK (W, Z) ≤ √ + 3/2 c + 4 8 np b2 − pa2 √  √ |a|b2 1 − p |a|b2 2π(1 − p) + + 2 8  ξ p √  1 9√ 2  + 2 + 2 1 − p a p + 2|a|bξ +√ 2 c np b2 − pa2 p   2(1 − p)b 2b2 − a2 √ + c 2π n+1 |a|b c2 n 2 . (1 − p) + (1 − p) + 2 b − pa2 b2 − pa2 Remark 2.16. Bounds for binomial random sums have also been derived in [Sun14] using a technique developed in [Tih80]. Our bounds are of the same order (np)−1/2 of magnitude. Proof of Corollary 2.15. Here, we clearly have α = np ,

γ 2 = np(1 − p) and σ 2 = np(a2 (1 − p) + c2 ) .

Also, using the same coupling as in Example 2.3 (b) we have D ∼ Bernoulli(1 − p), E[D 2 ] = E[D] = 1 − p and E[D|N] = 1 −

N . n

This yields  1 p(1 − p) Var E[D|N] = 2 Var(N) = . n n We have D 2 = D and, by Cauchy-Schwartz, q  q    p  p  E D1{N ≥1} N −1/2 ≤ E[D 2 ] E 1{N ≥1} N −1 = 1 − p E 1{N ≥1} N −1 .

Using

      1 n 2 2 n+1 n+1 ≤ ≤ , k k n+1 k+1 n k+1

we have 

E 1{N ≥1} N

−1



1 ≤ k ≤ n,

   n  n X 2 X n+1 k 1 n k n−k p (1 − p)n−k p (1 − p) ≤ = k + 1 k k n k=1 k=1  n+1  X 2 2 n+1 l = p (1 − p)n+1−l ≤ . l np np l=2

ERROR BOUNDS IN THE CLT FOR RANDOM SUMS

Thus, 

E D1{N ≥1} N

−1/2

Also, we can bound h





p

2(p − 1) √ pn

√   2 −1/2 and E 1{N ≥1} N ≤√ . np

2 i   E E D N ≤ E D 4 = E[D] = 1 − p . 

13

2

Now, using a2 + c2 = b2 , the claim follows from Corollary 2.8.



Corollary 2.17. In addition to the assumptions from Theorem 2.5 suppose that N ∼ Hyp(n; r, s) has the Hypergeometric distribution with parameters n, r, s ∈ N such that n ≤ min{r, s}. Then,  nr −1/2  2b s r + s − n 3ξ |a|b2 s r + s − n  + 3 + 2 dW (W, Z) ≤ r+s c r+s r+s−1 c c r+s r+s−1 r 2 a 2p + 2 ε(n, r, s) and c π " √  nr −1/2 ( 2π + 4)b  s r + s − n 1/2 1+ dK (W, Z) ≤ r+s 4c r+s r+s−1 √ √  2π  |a|b2  3 2π 9 √ s(r + s − n) 5 ξ + + + 1 + 2+ 3 3 8 2 2 c 8 c (r + s)(r + s − 1) 1/2 #    √ b 2s(r + s − n) |a|bξ + |a|b2 c3 2π + 2 + √ c (r + s)(r + s − 1) c 2π 1/2  (s)n a2 p (s)n s(r + s − n) |a|b + + . ε(n, r, s) + 2 (r + s)n c2 c (r + s)n (r + s)(r + s − 1)

where ε(n, r, s) is defined in (7) below and (m)n = m(m − 1) · . . . · (m − n + 1) denotes the lower factorial. Proof. In this case, we clearly have nr s r+s−n nr , γ2 = and α= r+s r+s r+s r+s−1 s r + s − n nr  2 c + a2 . σ2 = r+s r+s r+s−1 Hence, nr s  nr  2 r  nr  2 c2 c + a2 = b − a2 . ≤ σ2 ≤ r+s r+s r+s r+s r+s We use the coupling constructed in Example 2.3 (c) but write N for X and N s for X s , here. Recall that we have n   X s D = N − N = 1{X1 =0} 1 − Yj ≥ 0 and D = D 2 . j=2

Furthermore, we know that

E[D] = E[D 2 ] = dW (N, N s ) =

Var(N) s(r + s − n) = . E[N] (r + s)(r + s − 1)

14

CHRISTIAN DÖBLER

Elementary combinatorics yield   E Yj X1 , . . . , Xn = r −1 1{Xj =1} .

Thus,

n  X  1 N E D X1 , . . . , Xn = 1{X1 =0} − 1{X1 =0} and 1{Xj =1} = 1{X1 =0} 1 − r r j=2       N N n − N E D N = 1 − P X1 = 0 N = 1 − r r n (r − N)(n − N)  N N  = 1− . = 1− nr r n Using a computer algebra system, one may check that    Var E D N = nrs − n3 rs − r 2 s + 5n2 r 2 s + 2n3 r 2 s − 8nr 3 s − 8n2 r 3 s + 2nrs5



− n3 r 3 s + 4r 4 s + 10nr 4 s + 3n2 r 4 s − 4r 5 s − 3nr 5 s + r 6 s + ns2

− n3 s2 − 2rs2 + 4n2 rs2 − 2n3 rs2 − 14nr 2s2 − 4n2 r 2 s2 + n3 r 2 s2 + 12r 3s2 + 20nr 3 s2 + 2n2 r 3 s2 − 14r 4 s2 − 7nr 4 s2 + 4r 5 s2 − s3

− n2 s3 + 2n3 s3 − 5nrs3 + 4n2 rs3 + n3 rs3 + 13r 2s3 + 8nr 2 s3

− 4n2 r 2 s3 − 18r 3 s3 − 3nr 3 s3 + 6r 4 s3 + ns4 − n3 s4 + 6rs4 − 4nrs4 − 2n2 rs4 − 10r 2 s4 + 3nr 2 s4 + 4r 3 s4 + s5 − 2ns5 + n2 s5 − 2rs5  −1 2 2 2 5 nr(r + s) (r + s − 1) (r + s − 2)(r + s − 3) +r s

(7)

=: ε(n, r, s) .

Also, by the conditional version of Jensen’s inequality h   i   s(r + s − n) 2 . E E D2 N ≤ E D 4 = E[D] = (r + s)(r + s − 1)

Using

   −1 X n s 1 r r+s k k n−k n k=1   −1 X  n+1  2 s r+1 r+s ≤ n+1−l l n r+1 l=2  −1   2 2(r + s + 1) r+1+s r+s r+s ≤ = ≤2 , n+1 n r+1 (n + 1)(r + 1) nr

  E N −1 1{N ≥1} =

we get 

E D1{N ≥1} N

−1/2







p

 q   E[D 2 ] E 1{N ≥1} N −1 ≤ 2

s(r + s − n) r+s (r + s)(r + s − 1) nr

√  s r + s − n 1/2 = 2 and nr r + s − 1 q     √  r + s 1/2 E 1{N ≥1} N −1/2 ≤ E 1{N ≥1} N −1 ≤ 2 . nr

1/2

ERROR BOUNDS IN THE CLT FOR RANDOM SUMS

15

Finally, we have P (N = 0) =

s n  r+s n



=

(s)n s(s − 1) · . . . · (s − n + 1) = . (r + s)(r + s − 1) · . . . · (r + s − n + 1) (r + s)n

Thus, the result follows from Corollary 2.8.  Remark 2.18. ways

(1) One can check that under the assumption n ≤ min{r, s} al-

As always

 min{r, s} . ε(n, r, s) = O n(r + s) 

min{r, s} r+s 1 ≤ = , n(r + s) nr E[N] we conclude that the bounds in Theorem 2.17 are of order E[N]−1/2 . (2) One typical situation, in which a CLT for Hypergeometric random sums holds, is when N, itself, is asymptotically normal. Using the the same coupling (N, N s ) as in the above proof and the rsults from [GR96], one obtains that under the condition max{r, s} −→ 0 (8) n min{r, s}

(9)

the index N is asymptotically normal. This condition is stricter than that r+s −→ 0 , E[N]−1 = nr which implies the random sums CLT. For instance, choosing r ∝ n1+ε ,

and s ∝ n1+κ

with ε, κ ≥ 0, then (8) holds, if and only if |ε − κ| < 1, whereas (9) is equivalent to κ − ε < 1 in this case. Proof of Lemma 2.1. Let h be a measurable function such that all the expected values in (2) exist. By (2) we have   1  (10) E[h(X s )] − E[h(X)] = E X − E[X] h(X) . E[X] It is well known that (11)

dT V (X, Y ) = sup E[h(X)] − E[h(Y )] , h∈H

where H is the class of all measurable functions on R such that khk∞ ≤ 1/2. If khk∞ ≤ 1/2, then   E|X − E[X]| 1  E X − E[X] h(X) ≤ E[X] 2E[X] Hence, from (11) and (10) we conclude that (12)

dT V (X, X s ) ≤

E|X − E[X]| . 2E[X]

16

CHRISTIAN DÖBLER

On the other hand, letting  1 h(x) := 1{x>E[X]} − 1{x≤E[X]} 2 in (10) we have h ∈ H and obtain E|X − E[X]| (13) E[h(X s )] − E[h(X)] = 2E[X]

proving (b). Note that we have   dK (X, X s ) = sup P (X s > t) − P (X > t) = sup P (X s > t) − P (X > t) t≥0 t≥0   (14) = sup E[gt (X s )] − E[gt (X)] , t≥0

where gt := 1(t,∞) . If 0 ≤ t < E[X] we obtain          E X − E[X] 1{X>t} = E X − E[X] 1{tE[X]}    ≤ E X − E[X] 1{X>E[X]} . (15) Also, if t ≥ E[X], then       (16) E X − E[X] 1{X>t} ≤ E X − E[X] 1{X>E[X]} .

Thus, by (10) from (14), (15) and (16) we conclude the claim of (a). Finally, if h is 1-Lipschitz continuous, then       E X − E[X] h(X) = E X − E[X] h(X) − h(E[X])   ≤ kh′ k∞ E |X − E[X]|2 = Var(X) . On the other hand, the function h(x) := x − E[X] is 1-Lipschitz and    E X − E[X] h(X) = Var(X) .

Thus, also (c) is proved.



3. Elements of Stein’s method In this section we review some well-known and also some recent results about Stein’s method of normal approximation. Our general reference for this topic is the book [CGS11]. Throughout, Z will denote a standard normal random variable. Stein’s method originated from Stein’s seminal observation (see [Ste72]) that a realvalued random variable X has the standard normal distribution, if and only if for all, say, Lipschitz-continuous functions f , the identity     (17) E f ′ (X) = E Xf (X) holds. For a given random variable W , which is supposed to be asymptotically normal, and a Borel-measurable test function h on R with E|h(Z)| < ∞ it was then Stein’s idea to solve the Stein equation (18)

f ′ (x) − xf (x) = h(x) − E[h(Z)]

and to use properties of the solution f and of W in order to bound the right hand side of      ′  E h(W ) − E h(Z) = E f (W ) − W f (W )

ERROR BOUNDS IN THE CLT FOR RANDOM SUMS

17

rather than bounding the left hand side directly. For h as above, by fh we denote the standard solution to the Stein equation (18) which is given by Z x  2 x2 /2 h(t) − E[h(Z)] e−t /2 dt fh (x) = e −∞ Z ∞  2 x2 /2 = −e (19) h(t) − E[h(Z)] e−t /2 dt . x

Note that generally fh is only differentiable and satisfies (18) at the continuity points of h. In order to be able to deal with distributions which might have point masses, if x ∈ R is a point at which fh is not differentiable, one defines (20)

fh′ (x) := xfh (x) + h(x) − E[h(Z)]

such that, by definition, fh satisfies (18) at each point x ∈ R. This gives a Borelmeasurable version of the derivative of fh in the Lebesgue sense. Properties of the solutions fh for various classes of test functions h have been studied. Since we are only interested in the Kolmogorov and Wasserstein distances, we either suppose that h is 1-Lipschitz or that h = hz = 1(−∞,z] for some z ∈ R. In the latter case we write fz for fhz . We need the following properties of the solutions fh . If h is 1-Lipschitz, then it is well known (see e.g. [CGS11]) that fh is continuously differentiable and that both fh and fh′ are Lipschitz-continuous with r 2 ′ (21) kfh k∞ ≤ 1 , kfh k∞ ≤ and kfh′′ k∞ ≤ 2 . π Here, for a function g on R, we denote by kg ′ k∞ := sup x6=y

|g(x) − g(y)| |x − y|

its minimum Lipschitz constant. Note that if g is absolutely continuous, then kgh′ k∞ coincides with the essential supremum norm of the derivative of g in the Lebesgue sense. Hence, the double use of the symbol k·k∞ does not cause any problems. For an absolutely continuous function g on R, a fixed choice of its derivative g ′ and for x, y ∈ R we let (22)

Rg (x, y) := g(x + y) − g(x) − g ′ (x)y

denote the remainder term of its first order Taylor expansion around x at the point x + y. If h is 1-Lipschitz, then we obtain for all x, y ∈ R that Rf (x, y) = fh (x + y) − fh (x) − fh′ (x)y ≤ y 2 . (23) h This follows from (21) via

Z x+y ′  fh (x + y) − fh (x) − fh′ (x)y = fh (t) − fh′ (x) dt x Z x+y y 2 kf ′′ k h ∞ ≤ kfh′′ k∞ |t − x|dt = ≤ y2 . 2 x

18

CHRISTIAN DÖBLER

For h = hz we list the following properties of fz : The function fz has the representation ( (1−Φ(z))Φ(x) , x≤z ϕ(x) (24) fz (x) = Φ(z)(1−Φ(x)) , x > z. ϕ(x) Here, Φ denotes the standard normal distribution function and ϕ := Φ′ the corresponding continuous density. It is easy to see from (24) that fz is infinitely often differentiable on R\{z}. Furthermore, it is well-known that fz is Lipschitz-continuous with Lipschitz constant 1 and that it satisfies √ 2π (25) 0 < fz (x) ≤ f0 (0) = , x, z ∈ R . 4 These properties already easily yield that for all x, u, v, z ∈ R √ !  (x + u)fz (x + u) − (x + v)fz (x + v) ≤ |x| + 2π |u| + |v| . (26) 4 Proofs of the above mentioned classic facts about the functions fz can again be found in [CGS11], for instance. As fz is not differentiable at z (the right and left derivatives do exist but are not equal) by the above convention we define (27)

fz′ (z) := zfz (z) + 1 − Φ(z)

such that f = fz satisfies (18) with h = hz for all x ∈ R. Furthermore, with this definition, for all x, z ∈ R we have (28)

|fz′ (x)| ≤ 1 .

The following quantitative version of the first order Taylor approximation of fz has recently been proved by Lachièze-Rey and Peccati [LRP15] and had already been used implicitly in [ET14]. Applying Convention (27), for all x, u, z ∈ R we have Rfz (x, u) = fz (x + u) − fz (x) − f ′ (x)u z √ !   2π u2 |x| + + |u| 1{x