
Let W=X1+X2+⋯+XN be a random sum and Z be the standard normal random variable. In this paper, we investigated uniform and non-uniform bounds of the stop-loss distance, which measures the difference between two random variables, W and Z, using the expression |Ehk(W)−Ehk(Z)|, where hk(x)=(x−k)+ is a call function. In particular, we focused on the case that X1,X2,… are independent random variables, and N is a non-negative, integer-valued random variable independent of the Xj's. Our methods were Stein's method and the concentration inequality approach. The value Ehk(W)=E(W−k)+ represents the excess over a threshold and is relevant to applications in collateralized debt obligations (CDOs) and the collective risk model.
Citation: Punyapat Kammoo, Kritsana Neammanee, Kittipong Laipaporn. Bounds for the stop-loss distance of an independent random sum via Stein's method[J]. AIMS Mathematics, 2025, 10(6): 13082-13103. doi: 10.3934/math.2025587
[1] | Tatpon Siripraparat, Suporn Jongpreechaharn . The exponential non-uniform bound on the half-normal approximation for the number of returns to the origin. AIMS Mathematics, 2024, 9(7): 19031-19048. doi: 10.3934/math.2024926 |
[2] | Huiming Zhang, Hengzhen Huang . Concentration for multiplier empirical processes with dependent weights. AIMS Mathematics, 2023, 8(12): 28738-28752. doi: 10.3934/math.20231471 |
[3] | Amani Alahmadi, Abdelkader Benkhaled, Waleed Almutiry . On the effectiveness of the new estimators obtained from the Bayes estimator. AIMS Mathematics, 2025, 10(3): 5762-5784. doi: 10.3934/math.2025265 |
[4] | Baishuai Zuo, Chuancun Yin . Stein’s lemma for truncated generalized skew-elliptical random vectors. AIMS Mathematics, 2020, 5(4): 3423-3433. doi: 10.3934/math.2020221 |
[5] | Ran Tamir . Testing for correlation in Gaussian databases via local decision making. AIMS Mathematics, 2025, 10(4): 7721-7766. doi: 10.3934/math.2025355 |
[6] | Xueping Hu, Jingya Wang . A Berry-Ess$\acute{e}$n bound of wavelet estimation for a nonparametric regression model under linear process errors based on LNQD sequence. AIMS Mathematics, 2020, 5(6): 6985-6995. doi: 10.3934/math.2020448 |
[7] | Rasha Abd El-Wahab Attwa, Shimaa Wasfy Sadk, Hassan M. Aljohani . Investigation the generalized extreme value under liner distribution parameters for progressive type-Ⅱ censoring by using optimization algorithms. AIMS Mathematics, 2024, 9(6): 15276-15302. doi: 10.3934/math.2024742 |
[8] | Min-Ku Lee, Jeong-Hoon Kim . Closed-form approximate solutions for stop-loss and Russian options with multiscale stochastic volatility. AIMS Mathematics, 2023, 8(10): 25164-25194. doi: 10.3934/math.20231284 |
[9] | Emad R. Attia . On the upper bounds for the distance between zeros of solutions of a first-order linear neutral differential equation with several delays. AIMS Mathematics, 2024, 9(9): 23564-23583. doi: 10.3934/math.20241145 |
[10] | Haichao Yu, Yong Zhang . The law of iterated logarithm for a class of random variables satisfying Rosenthal type inequality. AIMS Mathematics, 2021, 6(10): 11076-11083. doi: 10.3934/math.2021642 |
Let W=X1+X2+⋯+XN be a random sum and Z be the standard normal random variable. In this paper, we investigated uniform and non-uniform bounds of the stop-loss distance, which measures the difference between two random variables, W and Z, using the expression |Ehk(W)−Ehk(Z)|, where hk(x)=(x−k)+ is a call function. In particular, we focused on the case that X1,X2,… are independent random variables, and N is a non-negative, integer-valued random variable independent of the Xj's. Our methods were Stein's method and the concentration inequality approach. The value Ehk(W)=E(W−k)+ represents the excess over a threshold and is relevant to applications in collateralized debt obligations (CDOs) and the collective risk model.
Let X1,X2,X3,… be a sequence of random variables and let N be a non-negative, integer-valued random variable independent of the Xj's. The summation, X1+X2+⋯+XN, is called a random sum. Random sums appear frequently in modern probability theory and are widely applied across various fields, including finance for risk assessment and portfolio optimization, telecommunications for modeling call arrivals, operations research for inventory management, and particularly in insurance contexts (see Chapter 17 in [1,2,3] for additional examples).
The approximation of random sums X1+X2+⋯+XN by the standard normal random variable Z was begun by Robbins [4] and Gnedenko & Korolev [5]. A widely used probability metric to compare such approximations is the Wasserstein distance. The Wasserstein distance between random variables W and Z, represented as dW(W,Z), is defined by
dW(W,Z)=suph∈H|Eh(W)−Eh(Z)|, |
where H is the class of all Lipschitz continuous functions h on R with a Lipschitz constant not greater than 1. Many authors have studied approximations of the random sums using this metric under various assumptions about the random variable N. In the absence of specific distributional assumptions about N, the results for Gaussian approximation are provided in works such as [6,7,8]. When N is treated as a known random variable, common distributional assumptions include Poisson (see [6,9]), binomial [10,11], negative binomial[9,12], or mixed Poisson [9].
For k>0, let hk(x)=(x−k)+ be a call function, where x+=max{0,x} and let W be any random variable. For an insurance context, if we consider W as the claim amount claimed by an insured and W as a retention, then (W−k)+ represents excess money paid by a reinsurer (see [13] for more details). The main concern for a reinsurer is determining the magnitude of excess losses, i.e., E(W−k)+. In this paper, we focus on a specialized metric known as the stop-loss distance. This distance is useful for modeling the excess over a set threshold, which is important for assessing potential losses. For any random variables W and Z, the stop-loss distance is defined for some k∈R+ as
d(k)SL(W,Z)=|E(W−k)+−E(Z−k)+| | (1.1) |
and the uniform version as
dSL(W,Z)=supk∈R+|E(W−k)+−E(Z−k)+|. | (1.2) |
The stop-loss distances can be classified into two distinct categories. For a fixed value of k in Eq (1.1), the resulting error bound typically depends on k, which is referred to as a non-uniform bound. In contrast, the error bound corresponding to Eq (1.2), which applies uniformly across all values of k∈R+, is called a uniform bound. In situations where the stop-loss threshold k is sufficiently large, non-uniform bounds provide a more precise approximation than uniform bounds.
We focus on the stop-loss distance because it is useful for estimating potential losses and can be applied in many different areas. In risk management and insurance, for example, it models excess losses in the collective risk model and default risks in financial products, such as collateralized debt obligations (CDOs). Its relevance also extends to finance and investment, where understanding the distribution of excess losses beyond a certain point is crucial for effective risk management. For instance, in an insurance context, if we consider W:=X1+X2+⋯+XN as the total claim amount with a retention k, where Xj denotes the claim amount from the jth contract and N represents the number of claims, then (W−k)+ represents excess money paid by a reinsurer. A key application of this setup is in modeling total claims from a portfolio of insurance contracts. There are two approaches to modeling W: the individual risk model and the collective risk model, which differ based on whether N is fixed or random. In the individual risk model, the total claims are represented as X1+X2+⋯+Xn, where n is the fixed number of contracts in the portfolio. However, in many real-world insurance products, such as accident insurance, the number of claims is not fixed and can vary unpredictably. This uncertainty leads to the collective risk model, where N is treated as a random variable representing the number of claims. In practice, obtaining the precise value of E(W−k)+ is extremely difficult. To address this, we estimate E(W−k)+ using stop-loss distances with suitable limiting distributions Z. Beyond applications in the insurance context, this concept is also applied to risk management for CDOs, as discussed further in Section 4.
Throughout this paper, we let X1,X2,… be independent random variables with a zero mean and finite absolute third moment, and let N be a non-negative, integer-valued random variable with finite variance that is independent of the Xj's.
We will consistently use the following notation for any random variable Xi: σ2i=Var(Xi)=EX2i and γi=E|Xi|3. For a positive integer n, let
s2n=n∑i=1σ2iandδn=1s2n√Es2Nn∑i=1γi. |
For a non-negative integer-valued random variable N, we similarly define the random variables
s2N=N∑i=1σ2iandδN=1s2N√Es2NN∑i=1γi. |
Additionally, we denote
Yi=Xi√Es2N,andWN=N∑i=1Yi. |
From these, we observe that EWN=0 and Var(WN)=1, since Var(N∑i=1Xi)=Es2N.
As previously discussed, numerous authors have explored approximations for the random sum X1+X2+⋯+XN with a particular emphasis on Gaussian approximations. In this paper, we aim to find both uniform and non-uniform bounds for the stop-loss distance under the assumption mentioned earlier. A recent result on the Wasserstein distance was presented by Döbler [7] in 2015, which is stated as follows:
Theorem 1.1. [7] Let X1,X2,… be identically distributed. Then
dW(WN,Z)≤2√Var(N)EN+3γ1σ31√EN. |
We observe that dSL(WN,Z)≤dW(WN,Z), since any call function is a Lipschitz continuous function. Using this inequality, Theorem 1.1 directly provides a uniform bound for dSL(WN,Z) in the case where the Xj's are independent and identically distributed (i.i.d.). However, we cannot apply this to obtain a non-uniform bound. In this paper, we provide uniform and non-uniform bounds for the stop-loss distance without requiring the Xj's to be identically distributed. A key tool in our approach, as well as in Döbler's work [7], is Stein's method. Additionally, Stein's method has been applied in non-i.i.d. settings in works such as [14,15]. Using Stein's method along with the concentration inequality approach outlined in Section 3, we derive the main results presented in the following theorems.
Theorem 1.2. (Uniform bound)
dSL(WN,Z)≤√2π(√Var(s2N)Es2N)+7.24E[δNsN]√Es2N+3E[δNs2N]Es2N. |
Theorem 1.3. (Non-uniform bound) For k≥3, we have
d(k)SL(WN,Z)≤11+k[1.38√Var(s2N)Es2N+25.047E[δNsN]√Es2N+1.725E[δNs2N]Es2N+1.68E[δNs4N](Es2N)2]. |
To gain insight into the behavior of the bounds, which we expect to converge to zero, we often examine a specific case where the Xi's are assumed to be identically distributed. Under this assumption, we can directly derive the following corollary.
Corollary 1.1. Let X1,X2,… be identically distributed. Then
(i)
dSL(WN,Z)≤√2π(√Var(N)EN)+10.24γ1σ31√EN,and |
(ii) for k≥3, we have
d(k)SL(WN,Z)≤11+k[1.38√Var(N)EN+26.772γ1σ31√EN+1.68γ1EN2σ31(EN)52]. |
To ensure the convergence of each bound, it is essential to verify that all terms involving the random variable N such as √Var(N)EN, 1√EN, and EN2(EN)52, approach zero. These terms represent different aspects of the variability and scale of N, and their convergence is crucial for controlling the bounds effectively. As stated in the above corollary, in the trivial case where N=n, it is evident that the bound converges to zero as n→∞. To explore other possible and more complex forms of N, we provide the following remark, which illustrates the behavior of the bound in such cases.
Remark 1.1. We present the result from Corollary 1.1 for the case where N follows specific, well-known distributions:
(i) For N∼Bin(n,p),n∈N,p∈(0,1), we have
dSL(WN,Z)≤1√np[√2(1−p)π+10.24γ1σ31]andd(k)SL(WN,Z)≤1(1+k)√np[1.38√1−p+26.772γ1σ31+1.68γ1((1−p)+np)npσ31],for k≥3. |
In this case, it is straightforward to observe that the bounds converge to zero as n→∞.
(ii) For N∼Poi(λ), λ>0,
dSL(WN,Z)≤1√λ[√2π+10.24γ1σ31]andd(k)SL(WN,Z)≤1(1+k)√λ[1.38+26.772γ1σ31+1.68γ1(λ+1)λσ31],for k≥3. |
As λ→∞, the bounds still converge to zero.
(iii) For N∼NB(r,p), r>0,p∈(0,1),
dSL(WN,Z)≤1√r(1−p)[√2π+10.24γ1√pσ31]andd(k)SL(WN,Z)≤1(1+k)√r(1−p)[1.38+26.772γ1√pσ31+1.68γ1√p(1+r(1−p))r(1−p)σ31],for k≥3. |
In the final case of the remark, it is clear that both bounds approach zero, depending on the size of r.
We observe that to apply the theorem most effectively, it is important to carefully select an appropriate form of N along with its parameters, as these choices significantly influence the efficiency and tightness of the resulting bounds.
In the following remark, we present specific examples and situations where Döbler's approach yields better bounds and, conversely, where our results show improvements over Döbler's.
Remark 1.2. Let X1,X2,… be independent and identically distributed random walks with probability 0.5, i.e., P(Xj=±1)=0.5, and let N be a non-negative, integer-valued random variable independent of the Xj's.
(i) For N∼Bin(n,0.5), we apply Theorem 1.1 to obtain that dSL(WN,Z)≤6.25√n while Corollary 1.1 provides dSL(WN,Z)≤15.28√n.
(ii) For N∼NB(r,0.001), using Theorem 1.1 and Corollary 1.1, we have, for r≥1, dSL(WN,Z)≤3.2√r and dSL(WN,Z)≤1.13√r, respectively.
In this work, we discuss Stein's method, which is an important tool for our study, in Section 2. Subsequently, the detailed proof of the main result is presented in Section 3. Finally, we demonstrate the applications of our results in specific areas, such as collective risk models and CDOs.
In this section, we introduce the foundational tools for our work, beginning with the ingenious approach developed by Charles Stein in 1972, commonly referred to as Stein's method [16]. Let Φ be the distribution function of the standard normal Z, and Cbd be the set of continuous and piecewise continuously differentiable functions f:R→R with E|f′(Z)|<∞. Stein's method begins with the Stein equation for normal approximation,
xf(x)−f′(x)=h(x)−Eh(Z) | (2.1) |
for a given function h and f∈Cbd. The solution of Eq (2.1) is
f(x)=ex22∞∫xe−t22[h(t)−Eh(Z)] dt, |
see [17, p. 15].
In this work, we apply the Stein equation (2.1) with h(x)=(x−k)+ and k>0. Then we have
xf(x)−f′(x)=(x−k)+−E(Z−k)+. | (2.2) |
The solution of Eq (2.2) is
fk(x)={√2πex22E(Z−k)+Φ(x),if x≤k,1−√2πex22(k+E(Z−k)+)Φ(−x),if x>k, |
and the expression for the first derivative, denoted as f′k, is as follows:
f′k(x)={E(Z−k)+(1+√2πxex22Φ(x)),if x<k,(k+E(Z−k)+)(1−√2πxex22Φ(−x)),if x>k. | (2.3) |
It is important to note that f′k is not defined at the point x=k due to the discontinuity of fk at this point. However, by using the solution at x=k in conjunction with Stein's equation (2.2), we can refine the expression for f′k(k) by
f′k(k)=E(Z−k)+(1+√2πkek22Φ(k)). |
From this fact and (2.3), we have
f′k(x)={E(Z−k)+(1+√2πxex22Φ(x)),if x≤k,(k+E(Z−k)+)(1−√2πxex22Φ(−x)),if x>k. |
A crucial component of Stein's method is identifying the properties of the solution fk. From Lemma 2.4 of Chen et al. [18] and some observations shown by Jongpreechaharn and Neammanee [19, p. 210–211], we have that
0≤f′k(x)≤√2π,for x∈R. | (2.4) |
For the non-uniform bound for f′k, we use some results from [19] to derive the bounds for f′k in the following proposition.
Proposition 2.1. For k≥3 and x∈R, 0≤f′k(x)≤1.381+k.
Proof. Using the fact that 0≤f′k(x)≤e−k22√2πk2+1k for all x∈R and k≥1 [19, p. 210–211], and
1k≤43(1+k)andek22≥1+k22≥1+k,for k≥3, |
we have
0≤f′k(x)≤1√2π(1+k)k2+43(1+k)≤1.381+k. |
Additional, the second derivative, denoted as f″k, is expressed as follows:
f″k(x)={E(Z−k)+[x+(√2πex22Φ(x))(x2+1)],if x<k,(k+E(Z−k)+)[x−(√2πex22Φ(−x))(x2+1)],if x>k. |
We notice that f″k does not exist at x=k. It is important to highlight that this prevents the direct application of the mean value theorem on any interval [a,b] where k∈(a,b), since the theorem requires differentiability over the entire open interval (a,b). The uniform bound for f″k was observed by [20, p. 3501], indicating that
|f″k(x)|≤2,for x∈R∖{k}. | (2.5) |
Beyond the uniform bound for f″k, Jongpreechaharn and Neammanee [20, p. 3502] utilized the result
E(Z−k)+≤e−k22√2πmin{1,1k2} | (2.6) |
in [21, p. 115] to show that
|f″k(x)|≤1.43e3k28 for x≤k2. | (2.7) |
We employ some ideas from the process of finding (2.7) to obtain the following results.
Proposition 2.2. For k≥3,
(i) |f″k(x)|≤0.231+k for x≤k−1,
(ii) |f″k(x)|≤1.12 for x∈R∖{k}.
Proof. (i) For x<0, by (2.6) and the fact that |f″k(x)|≤2E(Z−k)+ [20, p. 3502], we obtain
|f″k(x)|≤2e−k22k2√2π≤89(1+k)e92√2π≤0.0041+k, |
where we use 1k2≤49(1+k) in the second inequality. Otherwise, for 0≤x≤k−1, we first observe that
0≤f″k(x)=E(Z−k)+[x+√2πΦ(x)ex22(x2+1)]≤E(Z−k)+[(k−1)+√2πe(k−1)22((k−1)2+1)]=A+B, |
where
A=(k−1)E(Z−k)+andB=E(Z−k)+[√2πe(k−1)22((k−1)2+1)]. |
Using (2.6) along with the fact that k−1k2≤11+k, we obtain that
(k−1)E(Z−k)+≤(k−1)e−k22k2√2π≤1(1+k)e92√2π≤0.00451+k. |
To bound B, we divide into two cases. If 3≤k≤4, we can use (2.6) along with the facts that 1ek≤15(1+k) and (k−1)2k2≤916 to show that
E(Z−k)+[√2π(k−1)2e(k−1)22]≤e−k22k2√2π[√2π(k−1)2e(k−1)22]=(k−1)2√ek2ek<0.18551+k |
and
E(Z−k)+[√2πe(k−1)22]≤e−k22k2√2π[√2πe(k−1)22]=√ek2ek≤0.041+k. |
Thus, for 3≤k≤4, |f″k(x)|≤0.231+k. For k≥4, we follow the same argument, using the facts that 1ek≤0.11+k and (k−1)2k2≤1, to obtain that |f″k(x)|≤0.181+k.
(ii) We first observe that, for k−1<x<k, we make use of (2.6) to conclude that
0≤f″k(x)=E(Z−k)+[x+(√2πex22Φ(x))(x2+1)]≤e−k22k2√2π[k+(√2πek22Φ(k))(k2+1)]≤e−k22k√2π+1+1k2<1.12for k≥3. | (2.8) |
For the final case where x>k, we note that
f″k(x)=(k+E(Z−k)+)[x−(√2πex22Φ(−x))(x2+1)]. |
To bound this term, we employ the Gaussian tail bound (see [18, p. 16, 38] and [22, p. 252]):
xe−x22(x2+1)√2π≤Φ(−x)≤min{12,1x√2π}e−x222 for x>0, |
which gives us
x≤√2πex22Φ(−x)(x2+1)≤x+1x. |
Thus, for x>k, we have f″k(x)≤0 which leads to
|f″k(x)|=(k+E(Z−k)+)[(√2πex22Φ(−x))(x2+1)−x]≤(k+E(Z−k)+)1x≤1+e−k22k3√2π≤1.01for k≥3, | (2.9) |
where we use (2.6) prior to the last inequality. For x<k−1, (ii) follows directly from (i). Using this fact, along with (2.8) and (2.9), we conclude that (ii) holds.
We apply Stein's equation (2.2) to establish the following:
EWNfk(WN)−Ef′k(WN)=E(WN−k)+−E(Z−k)+. | (3.1) |
Thus, we can bound |EWNfk(WN)−Ef′(WN)| instead of |E(WN−k)+−E(Z−k)+|. To handle the term |EWNfk(WN)−Ef′(WN)|, we use the concentration inequality approach as a complementary tool alongside Stein's method. This technique was first applied by Chen [23] and has been frequently employed to find bounds in this field (see [18,20,22,24] for more examples). Furthermore, the concentration inequality approach has been applied in recent studies, including the work of [25] and Auld and Neammanee [26]. We recall that, for i=1,2,…,n, Yi=Xi√Es2N and, in context outside of the random sum framework, we denote
Wn=n∑i=1Yi,andW(i)n=Wn−Yi. |
From these notations, we can observe that
EW2n=n∑i=1EY2i=s2nEs2N,andn∑i=1E|Yi|3=δns2nEs2N. | (3.2) |
For i=1,2,…, let Ki(t)=E(YiI(0≤t≤Yi)−YiI(Yi≤t<0)). The K-function acts as a bridge that links the concentration inequality approach with Stein's method. It allows us to control and understand how well a test function approximates a distribution, especially in terms of its tail behavior and deviations from expected values. To illustrate its tail behavior more clearly, we provide examples in Figures 1 and 2.
Chen [23, p. 100] showed that for all real t,
∫∞−∞Ki(t)dt=EY2i and ∫∞−∞|t|Ki(t)dt=12E|Yi|3, |
which implies the following:
n∑i=1∫∞−∞Ki(t)dt=n∑i=1EY2i=s2nEs2N and n∑i=1∫∞−∞|t|Ki(t)dt=12n∑i=1E|Yi|3=δns2n2Es2N. | (3.3) |
Using these results, together with Lyapunov's inequality, we derive that
∫∞−∞Ki(t)E(|Yi|+|t|)dt=EY2iE|Yi|+12E|Yi|3≤32E|Yi|3. |
This leads to the following bound:
n∑i=1∫∞−∞Ki(t)E(|Yi|+|t|)dt≤32n∑i=1E|Yi|3=3δns2n2Es2N. | (3.4) |
In order to apply Stein's method, we encounter terms like E[Wnfk(Wn)] and Ef′k(Wn) that require careful handling. By the fact that Yi is independent of W(i)n and EYi=0, we have that
E[Wnfk(Wn)]=n∑i=1E∫∞−∞f′k(W(i)n+t)Ki(t)dt, | (3.5) |
and
Ef′k(Wn)=Es2Ns2nn∑i=1E∫∞−∞f′k(Wn)Ki(t)dt, | (3.6) |
see [18, p. 20]. These results lead to the key expression:
|E[Wnfk(Wn)]−Ef′k(Wn)|=|n∑i=1E∫∞−∞(f′k(W(i)n+t)−Es2Ns2nf′k(Wn))Ki(t)dt|, |
which plays a crucial role in our analysis. Beyond the importance of the K-function highlighted in its introduction, we will see that the K-function serves as a key tool in addressing the expression above, paving the way for the application of concentration inequalities. The use of these inequalities, however, will be demonstrated later in the proof of the main results. One famous concentration inequality theorem, which will be used in our work, is stated as follows:
Lemma 3.1. [18, p. 54] Let ξ1,ξ2,…,ξn be independent random variables with zero means, satisfying n∑j=1Var(ξj)=1. For all real a<b, and for every 1≤i≤n,
P(a≤n∑j=1,j≠iξj≤b)≤√2(b−a)+2(√2+1)n∑j=1E|ξj|3. |
We are now ready to prove the main results.
Proof of Theorem 1.2. From (3.3), (3.5), and (3.6), we have that
|E[Wnfk(Wn)]−Ef′k(Wn)|=|n∑i=1E∫∞−∞(f′k(W(i)n+t)−Es2Ns2nf′k(Wn))Ki(t)dt|≤n∑i=1E∫∞−∞|f′k(W(i)n+t)−f′k(Wn)|Ki(t)dt+n∑i=1E∫∞−∞|1−Es2Ns2n||f′k(Wn)|Ki(t)dt≤n∑i=1E∫∞−∞|f′k(W(i)n+t)−f′k(Wn)|Ki(t)dt+‖f′k‖|s2n−Es2Ns2n|n∑i=1E∫∞−∞Ki(t)dt=n∑i=1E∫∞−∞|f′k(W(i)n+t)−f′k(Wn)|Ki(t)dt+‖f′k‖|s2n−Es2Ns2n|s2nEs2N≤T+‖f′k‖|s2n−Es2NEs2N|, | (3.7) |
where ‖f′k‖=supx∈R|f′k(x)| and
T=n∑i=1E∫∞−∞|f′k(W(i)n+t)−f′k(Wn)|Ki(t)dt. |
To bound T, we note that
T=n∑i=1E∫∞−∞|f′k(W(i)n+t)−f′k(Wn)|Ki(t)I(Ai,t)dt+n∑i=1E∫∞−∞|f′k(W(i)n+t)−f′k(Wn)|Ki(t)I(Aci,t)dt:=T1+T2, | (3.8) |
where
Ai,t={W(i)n+Yi>k,W(i)n+t≤k}∪{W(i)n+Yi≤k,W(i)n+t>k}. |
To bound T1, we observe that
I(Ai,t)=I(k−Yi<W(i)n≤k−t)+I(k−t<W(i)n≤k−Yi)≤I(k−|t|−|Yi|<W(i)n≤k+|t|+|Yi|). | (3.9) |
Using Lemma 3.1, (3.2), and the fact that Var(Wn√Es2Nsn) = 1, we have
E[P(k−|t|−|Yi|<W(i)n≤k+|t|+|Yi|∣Yi)]=E[P(k−|t|−|Yi|√Var(Wn)<W(i)n√Var(Wn)≤k+|t|+|Yi|√Var(Wn)∣Yi)]=E[P((k−|t|−|Yi|)√Es2Nsn<W(i)n√Var(Wn)≤(k+|t|+|Yi|)√Es2Nsn∣Yi)]≤2√2(|t|+E|Yi|)√Es2Nsn+2(√2+1)(Es2N)32s3nn∑j=1E|Yj|3=2√2(|t|+E|Yi|)√Es2Nsn+2(√2+1)δn√Es2NsnEs2N. | (3.10) |
Applying this inequality along with (3.3), (3.4), (3.9), and (3.10), we obtain that
n∑i=1E∫∞−∞Ki(t)I(Ai,t)dt≤n∑i=1E∫∞−∞Ki(t)I(k−|t|−|Yi|<W(i)n≤k+|t|+|Yi|)dt=n∑i=1∫∞−∞E[E(Ki(t)I(k−|t|−|Yi|<W(i)n≤k+|t|+|Yi|)∣Yi)]dt=n∑i=1∫∞−∞Ki(t)E(P(k−|t|−|Yi|<W(i)n≤k+|t|+|Yi|∣Yi))dt≤n∑i=1∫∞−∞Ki(t)E[2√2(|t|+E|Yi|)√Es2Nsn+2(√2+1)δn√Es2NsnEs2N]dt≤n∑i=1[3√2E|Yi|3√Es2Nsn+2(√2+1)δnEY2i√Es2NsnEs2N]≤(5√2+2)δnsn√Es2N. |
Combining this fact and (2.4), we derive the bound for T1 as follows:
T1≤√2π((5√2+2)δnsn√Es2N)≤7.24δnsn√Es2N. | (3.11) |
To bound T2, we note that
I(W(i)n+t≤k,W(i)n+Yi≤k)+I(W(i)n+t>k,W(i)n+Yi>k)=I(Aci,t)≤1. |
Since f″k exists on (−∞,k) and (k,∞), we can apply the mean value theorem on Aci,t, (2.5) and (3.4) to obtain that
T2=n∑i=1E∫∞−∞|f′k(W(i)n+t)−f′k(Wn)|Ki(t)I(Aci,t)dt≤2n∑i=1E∫∞−∞(|t|+|Yi|)Ki(t)I(Aci,t)dt≤3δns2nEs2N. | (3.12) |
From (2.4), (3.1), (3.7), (3.8), (3.11), and (3.12), we have
|E(WN−k)+−E(Z−k)+|≤∞∑n=1P(N=n)|E[Wnfk(Wn)]−Ef′k(Wn)|≤∞∑n=1P(N=n)[7.24δnsn√Es2N+3δns2nEs2N+√2π|s2n−Es2NEs2N|]≤7.24E[δNsN]√Es2N+3E[δNs2N]Es2N+√2π(√Var(s2N)Es2N). |
Proof of Theorem 1.3. From the proof of Theorem 1.2, we start to bound T. Note that
T=n∑i=1E∫∞−∞|f′k(W(i)n+t)−f′k(Wn)|Ki(t)I(B1,i,t)dt+n∑i=1E∫∞−∞|f′k(W(i)n+t)−f′k(Wn)|Ki(t)I(B2,i,t)dt+n∑i=1E∫∞−∞|f′k(W(i)n+t)−f′k(Wn)|Ki(t)I(B3,i,t)dt:=S1+S2+S3, | (3.13) |
where
B1,i,t={W(i)n+Yi>k−1,W(i)n+t≤k−1}∪{W(i)n+Yi≤k−1,W(i)n+t>k−1}∪{k−1<Wn≤k,W(i)n+t>k}∪{Wn>k,k−1<W(i)n+t≤k},B2,i,t={Wn≤k−1,W(i)n+t≤k−1},andB3,i,t={k−1<Wn≤k,k−1<W(i)n+t≤k}∪{Wn>k,W(i)n+t>k}. |
To bound S1, we note from (3.9) that
I(B1,i,t)=I(k−1−Yi<W(i)n≤k−1−t)+I(k−1−t<W(i)n≤k−1−Yi)+I(k−1<Wn≤k,W(i)n+t>k)+I(Wn>k,k−1<W(i)n+t≤k)≤I(k−1−|t|−|Yi|<W(i)n≤k−1+|t|+|Yi|)+I(k−|t|−|Yi|<W(i)n≤k+|t|+|Yi|). |
By applying Lemma 3.1 in a similar fashion as (3.10), we derive the following inequality:
E[P(k−1−|t|−|Yi|<W(i)n≤k−1+|t|+|Yi|∣Yi)]=2√2(|t|+E|Yi|)√Es2Nsn+2(√2+1)δn√Es2NsnEs2N. |
Applying this inequality along with (3.3), (3.4), and (3.10), we obtain that
n∑i=1E∫∞−∞Ki(t)I(B1,i,t)dt=n∑i=1∫∞−∞Ki(t)[E(P(k−1−|t|−|Yi|<W(i)n≤k−1+|t|+|Yi|)∣Yi)+E(P(k−|t|−|Yi|<W(i)n≤k+|t|+|Yi|)∣Yi)]dt≤n∑i=1∫∞−∞Ki(t)E[4√2(|t|+|Yi|)√Es2Nsn+4(√2+1)δn√Es2NsnEs2N]dt=18.15δnsn√Es2N. |
Combining this fact and Proposition 2.1, we derive the bound for S1 as follows:
S1≤18.15δnsn‖f′k‖√Es2N≤11+k(25.047δnsn√Es2N). | (3.14) |
To bound S2, we use the mean value theorem, Proposition 2.2(i), and (3.4) to obtain that
S2=n∑i=1E∫∞−∞|f′k(W(i)n+t)−f′k(Wn)|Ki(t)I(W(i)n+t≤k−1,W(i)n+Yi≤k−1)dt≤n∑i=1E∫∞−∞|f″k(w)|(|t|+|Yi|)Ki(t)dt,for some w≤k−1≤0.231+k[n∑i=1E∫∞−∞(|t|+|Yi|)Ki(t)dt]≤0.231+k[3δns2n2Es2N]=0.345δns2n(1+k)Es2N. | (3.15) |
To bound S3, we note that
S3=n∑i=1E∫∞−∞|f′k(W(i)n+t)−f′k(Wn)|Ki(t)I(B3,i,t)dt=n∑i=1E∫∞−∞|f′k(W(i)n+t)−f′k(Wn)|Ki(t)I(B3,i,t)I(|Yi|≤1)dt+n∑i=1E∫∞−∞|f′k(W(i)n+t)−f′k(Wn)|Ki(t)I(B3,i,t)I(|Yi|>1)dt:=S3,1+S3,2. |
For S3,1, we note that
I(k−1<Wn≤k,k−1<W(i)n+t≤k)I(|Yi|≤1)≤I(Wn>k−1)I(|Yi|≤1)≤I(W(i)n>k−2)I(|Yi|≤1). |
Similarly,
I(Wn>k,W(i)n+t>k)I(|Yi|≤1)≤I(Wn>k)I(|Yi|≤1)≤I(W(i)n>k−2)I(|Yi|≤1). |
From these observations and since {k−1<Wn≤k,k−1<W(i)n+t≤k} and {Wn>k,W(i)n+t>k} are disjoint, we conclude that
I(B3,i,t)I(|Yi|≤1)≤I(W(i)n>k−2)I(|Yi|≤1). |
By applying the mean value theorem in conjunction with the above fact, Proposition 2.2(ii), and (3.2), we obtain
S3,1≤1.12n∑i=1E∫∞−∞(|Yi|+|t|)Ki(t)I(B3,i,t)I(|Yi|≤1)dt≤1.12n∑i=1E∫∞−∞(|Yi|+|t|)Ki(t)I(W(i)n>k−2)I(|Yi|≤1)dt≤1.12n∑i=1[3E|Yi|32P(W(i)n>k−2)]≤1.68n∑i=1[E|Yi|3(E(W(i)n)2(k−2)2)]≤1.68n∑i=1[E|Yi|3(s2n(1+k)Es2N)]=1.68δns4n(1+k)(Es2N)2, |
where we use the fact that for k≥3, 1(k−2)2≤11+k in the fifth inequality. To bound S3,2, we use Proposition 2.1, (3.2), and (3.3) to obtain
S3,2≤1.381+kn∑i=1E∫∞−∞Ki(t)I(|Yi|>1)dt≤1.381+kn∑i=1EY2iP(|Yi|>1)≤1.381+kn∑i=1E|Yi|3=1.38δns2n(1+k)Es2N. |
Then
S3≤1.68δns4n(1+k)(Es2N)2+1.38δns2n(1+k)Es2N. | (3.16) |
Combining (3.7) and (3.13)–(3.16), we have
|E[Wnfk(Wn)]−Ef′k(Wn)|≤11+k[25.047δnsn√Es2N+1.725δns2nEs2N+1.68δns4n(1+k)(Es2N)2+1.38|s2n−Es2NEs2N|]. |
Therefore, using this inequality along with (3.1), we conclude that
|E(WN−k)+−E(Z−k)+|≤11+k[25.047E[δNsN]√Es2N+1.725E[δNs2N]Es2N+1.68E[δNs4N](Es2N)2+1.38√Var(s2N)Es2N]. |
Since the Xj's are identically distributed, we have δn=γ1σ31√EN, which implies that
Es2N=σ21EN,δnsn√Es2N=γ1√nσ31EN,andδns2nEs2N=γ1nσ31(EN)32. |
Then
E[δNsN]√Es2N=E[δNs2N]Es2N=γ1σ31√EN. |
In conclusion, the proof can be finalized by noting that
E[δNs4N](Es2N)2=γ1EN2σ31(EN)52. |
A fascinating example of utilizing the call function has emerged in the pricing of CDO tranches. A CDO is a complex financial product that pools together various debt instruments, such as bonds and loans, and then divides the pool into different tranches with varying levels of risk and return. Each tranche is assigned a different payment priority and interest rate. Normally, the tranches primarily used in CDOs are typically known as senior, mezzanine, and junior. The senior tranche includes securities with high credit ratings and tends to be low risk, and therefore has lower returns. Investors have the option to invest in various tranches based on their preferences. In the conventional pricing model for a CDO with a total of n portfolios, each portfolio i is assumed to possess a recovery rate R>0. This recovery rate signifies the proportion of bad debt that can be recovered. We can obtain the percentage loss at the time T by the total loss on the portfolio,
L(T)=(1−R)nn∑i=1I{τi≤T}, |
where τi is the default time of the ith portfolio and IA is an indicator function of the set A.
For each CDO tranche, there exists a detachment point (a limit above which the tranche loss does not increase) and an attachment point (a limit below which the tranche bears none of the loss). Since the cash flow of a CDO tranche is driven by its loss, the pricing problem can be reduced to the problem of calculating the expectation of a call function, i.e., E(L(T)−l)+ where l is the attachment or the detachment point of the tranche (see [27,28,29,30] for more details).
The approximation of the stop-loss model for CDOs has been widely studied, as evidenced by [20,27,31,32]. To apply Theorem 1.3 effectively in the Gaussian approximation, we need to centralize the summand variables. Specifically, for each i=1,2,…,n, define
ξi=1n(1−R)I{τi≤T}andXi=ξi−μi, |
where μi=Eξi=(1−R)pin, pi=P(τi≤T). Then we let
Yi=XisnandWn=n∑i=1Yi=L(T)−μsn, |
where μ:=n∑i=1μi. Then we can apply Theorems 1.2 and 1.3 to obtain that
|E(Wn−k)+−E(Z−k)+|≤10.24δn |
and for k≥3,
|E(Wn−k)+−E(Z−k)+|≤28.452δn1+k, |
where
s2n=n∑i=1Var(Xi)=(1−Rn)2n∑i=1pi(1−pi)andδn=(1−Rnsn)3n∑i=1pi(1−pi)(1−2pi+2p2i). |
Based on the behavior of the bounds above, we observe that both bounds converge to zero at the rate O(δn)=O(1√n).
As previously discussed, the collective risk model represents total claims as ξ1+ξ2+⋯+ξN, where ξi's are independent and identically distributed, each ξi represents the claim amount from the ith claim, and N represents either the number of claims or the number of claim payments. Allow us to provide a more precise explanation of why we regard N as a random variable: When selling insurance that permits individuals to make multiple claims within a given period, such as accident insurance, we face uncertainty regarding the exact number of claims. To address this uncertainty, we use a random variable, N, to interpret it. Commonly used distributions for the number of claims include the binomial, Poisson, and negative binomial distributions.
Before applying our results, we need to denote that for i=1,2…,
Xi=ξi−μ,Yi=Xiσ√EN,andWN=N∑i=1Yi, |
where μ=Eξi and σ2=Var(ξi). To align our results more closely with realistic scenarios, we consider ξ as the claim amount in a health insurance contract. It is common to model the random index N, representing the number of claims, using a Poisson distribution, as it is well-suited for rare events (e.g., the likelihood of a single person frequently claiming health insurance is relatively low). Referring to Remark 1.1, we observe that if N∼Poi(λ), then
|E(WN−k)+−E(Z−k)+|≤1√λ[√2π+10.24γ1σ31] |
and for k≥3,
|E(WN−k)+−E(Z−k)+|≤1(1+k)√λ[1.38+28.46γ1σ31+1.68γ1λσ31]. |
As noted in Remark 1.1, these bounds converge to zero at a rate of O(1λ) when λ→∞.
In this work, we studied the uniform and non-uniform bounds for the stop-loss distance between a random sum W=X1+X2+⋯+XN and the standard normal random variable Z. The stop-loss distance, expressed as |Ehk(W)−Ehk(Z)|, with hk(x)=(x−k)+, serves as a crucial measure of the deviation between the two random variables.
Our approach combined Stein's method with concentration inequalities, leveraging their strengths to derive precise bounds. By focusing on the case where X1,X2,…,XN are independent random variables and N is a non-negative integer-valued random variable independent of the Xj's, we provided detailed insights into the behavior of the stop-loss distance under various conditions.
In particular, for the i.i.d. case, our results demonstrated improvements over Döbler's results in specific scenarios, as illustrated with examples in the paper. Furthermore, our investigation into non-uniform bounds represents a novel contribution, offering a detailed characterization of the stop-loss distance for random sums compared to the normal distribution.
Lastly, we explored practical applications of our results, emphasizing their relevance to financial and insurance contexts, such as collateralized debt obligations (CDOs) and the collective risk model. These applications highlight the utility of our findings in understanding and managing risk in real-world scenarios.
Punyapat Kammoo: Visualization, writing – original draft, funding acquisition, formal analysis, validation, writing – review & editing; Kritsana Neammanee: Conceptualization, methodology, project administration, visualization, writing – original draft, Formal analysis, Validation, Writing – review & editing; Kittipong Laipaporn: Conceptualization, methodology, project administration, funding acquisition, formal analysis, validation, writing – review & editing. All authors have read and approved the final version of the manuscript for publication.
The authors declare that they have not used Artificial Intelligence (AI) tools in the creation of this article.
This work was supported by the Development and Promotion of Science and Technology Talents Project (DPST).
All authors declare no conflicts of interest in this paper.
[1] | S. D. Promislow, Fundamentals of actuarial mathematics, 2 Eds., John Wiley & Sons, Ltd, 2006. http://dx.doi.org/10.1002/9781119971528 |
[2] |
N. Kolev, D. Paiva, Multinomial model for random sums, Insurance Math. Econom., 37 (2005), 494–504. http://dx.doi.org/10.1016/j.insmatheco.2005.05.005 doi: 10.1016/j.insmatheco.2005.05.005
![]() |
[3] |
N. Kolev, D. Paiva, Random sums of exchangeable variables and actuarial applications, Insurance Math. Econom., 42 (2008), 147–153. http://dx.doi.org/10.1016/j.insmatheco.2007.01.010 doi: 10.1016/j.insmatheco.2007.01.010
![]() |
[4] |
H. Robbins, The asymptotic distribution of the sum of a random number of random variables, Bull. Amer. Math. Soc., 54 (1948), 1151–1161. https://doi.org/10.1090/S0002-9904-1948-09142-X doi: 10.1090/S0002-9904-1948-09142-X
![]() |
[5] | B. V. Gnedenko, V. Y. Korolev, Random summation: Limit theorems and applications, Boca Raton: CRC Press, 1996. http://dx.doi.org/10.1201/9781003067894 |
[6] |
F. Daly, Gamma, Gaussian and Poisson approximations for random sums using size-biased and generalized zero-biased couplings, Scand. Actuar. J., 2022 (2022), 471–487. https://doi.org/10.1080/03461238.2021.1984293 doi: 10.1080/03461238.2021.1984293
![]() |
[7] | C. Döbler, New Berry-Esseen and Wasserstein bounds in the CLT for non-randomly centered random sums by probabilistic methods, ALEA, Lat. Am. J. Probab. Math. Stat., 12 (2015), 863–902. |
[8] |
J. K. Sunklodas, L1 bounds for asymptotic normality of random sums of independent random variables, Lith. Math. J., 53 (2013), 438–447. https://doi.org/10.1007/S10986-013-9220-X doi: 10.1007/S10986-013-9220-X
![]() |
[9] |
I. G. Shevtsova, Convergence rate estimates in the global CLT for compound mixed Poisson distributions, Theory Prob. Appl., 63 (2018), 72–93. https://doi.org/10.1137/S0040585X97T988927 doi: 10.1137/S0040585X97T988927
![]() |
[10] |
I. G. Shevtsova, A moment inequality with application to convergence rate estimates in the global CLT for Poisson-binomial random sums, Theory Prob. Appl., 62 (2018), 278–294. https://doi.org/10.1137/S0040585X97T988605 doi: 10.1137/S0040585X97T988605
![]() |
[11] |
J. K. Sunklodas, On the normal approximation of a binomial random sum, Lith. Math. J., 54 (2014), 356–365. http://dx.doi.org/10.1007/s10986-014-9248-6 doi: 10.1007/s10986-014-9248-6
![]() |
[12] |
J. K. Sunklodas, On the normal approximation of a negative binomial random sum, Lith. Math. J., 55 (2015), 150–158. https://doi.org/10.1007/s10986-015-9271-2 doi: 10.1007/s10986-015-9271-2
![]() |
[13] |
H. You, X. Zhou, The Pareto-optimal stop-loss reinsurance, Math. Probl. Eng., 2021 (2021), 2839726. https://doi.org/10.1155/2021/2839726 doi: 10.1155/2021/2839726
![]() |
[14] |
X. Fang, Y. Koike, High-dimensional central limit theorems by Stein's method, Ann. Appl. Probab., 31 (2021), 1660–1686. https://doi.org/10.1214/20-AAP1629 doi: 10.1214/20-AAP1629
![]() |
[15] |
S. Bhar, R. Mukherjee, P. Patil, Kac's central limit theorem by Stein's method, Stat. Probab. Lett., 219 (2025), 110329. https://doi.org/10.1016/j.spl.2024.110329 doi: 10.1016/j.spl.2024.110329
![]() |
[16] | C. Stein, A bound for the error in the normal approximation to the distribution of a sum of dependent random variables, In: Proceedings of the sixth Berkeley symposium on mathematical statistics and probability, 2 (1972), 583–602. |
[17] | C. Stein, Approximation computaional of expectations, In: Institute of mathematical statistics lecture notes, 7 (1986), 161–164. |
[18] | L. H. Y. Chen, L. Goldstein, Q. M. Shao, Normal approximation by Stein's method, 1 Eds., Heidelberg: Springer Berlin, 2010. https://doi.org/10.1007/978-3-642-15007-4 |
[19] | S. Jongpreechaharn, K. Neammanee, Non-uniform bound on normal approximation for call function of locally dependent random variables, Int. J. Math. Comput. Sci., 17 (2022), 207–216. |
[20] |
S. Jongpreechaharn, K. Neammanee, Normal approximation for call function via Stein's method, Comm. Statist. Theory Methods, 48 (2019), 3498–3517. https://doi.org/10.1080/03610926.2018.1476716 doi: 10.1080/03610926.2018.1476716
![]() |
[21] | S. Jongpreechaharn, K. Neammanee, A constant of approximation of the expectation of call function by the expectation of normal distributed random variable, In: Proceeding for 11th conference on science and technology for youth year 2016, 2016,112–121. |
[22] |
L. H. Y. Chen, Q. M. Shao, A non-uniform Berry–Esseen bound via Stein's method, Probab. Theory Relat. Fields, 120 (2001), 236–254. https://doi.org/10.1007/PL00008782 doi: 10.1007/PL00008782
![]() |
[23] | L. H. Y. Chen, Stein's method: Some perspectives with applications, In: Probability towards 2000, New York: Springer, 1998, 97–122. https://doi.org/10.1007/978-1-4612-2224-8_6 |
[24] | N. Chaidee, M. Tuntapthai, Berry-Esseen bounds for random sums of Non-i.i.d. random variables, Int. Math. Forum, 4 (2009), 1281–1288. |
[25] |
L. H. Y. Chen, Stein's method of normal approximation: Some recollections and reflections, Ann. Statist., 49 (2021), 1850–1863. https://doi.org/10.1214/21-aos2083 doi: 10.1214/21-aos2083
![]() |
[26] |
G. Auld, K. Neammanee, Explicit constants in the nonuniform local limit theorem for Poisson binomial random variables, J. Inequal. Appl., 2024 (2024), 67. https://doi.org/10.1186/s13660-024-03143-z doi: 10.1186/s13660-024-03143-z
![]() |
[27] |
N. E. Karoui, Y. Jiao, Stein's method and zero bias transformation for CDO tranche pricing, Finance Stoch., 13 (2009), 151–180. https://doi.org/10.1007/s00780-008-0084-6 doi: 10.1007/s00780-008-0084-6
![]() |
[28] |
P. Glasserman, S. Suchintabandid, Correlations for CDO pricing, J. Bank. Finance, 31 (2007), 1375–1398. https://doi.org/10.1016/j.jbankfin.2006.10.018 doi: 10.1016/j.jbankfin.2006.10.018
![]() |
[29] |
J. Hull, A. White, Valuation of a CDO and an nth to default CDS without Monte Carlo simulation, J. Deriv., 12 (2004), 8–23. http://dx.doi.org/10.3905/jod.2004.450964 doi: 10.3905/jod.2004.450964
![]() |
[30] |
N. E. Karoui, Y. Jiao, D. Kurtz, Gaussian and poisson approximation: Applications to CDOs tranche pricing, J. Comput. Finance, 12 (2008), 31–58. http://dx.doi.org/10.21314/JCF.2008.180 doi: 10.21314/JCF.2008.180
![]() |
[31] |
K. Neammanee, N. Yonghint, Poisson approximation for call function via Stein-Chen method, Bull. Malays. Math. Sci. Soc., 43 (2020), 1135–1152. https://doi.org/10.1007/s40840-019-00729-5 doi: 10.1007/s40840-019-00729-5
![]() |
[32] |
N. Yonghint, K. Neammanee, Refinement on Poisson approximation of CDOs, Sci. Asia, 47 (2021), 388–392. http://dx.doi.org/10.2306/scienceasia1513-1874.2021.041 doi: 10.2306/scienceasia1513-1874.2021.041
![]() |