1. Introduction
Many questions remain open in the determination of probability measures with given margins and other constraints, since the seminal work of [Reference Strassen11]. One of the challenging and recently active questions is, for n given distributions $F_1,\ldots, F_n$ on ${\mathbb{R}}$ , to determine all possible distributions of $S_n = X_1 + \dots + X_n$ , where $X_1,\ldots,X_n$ are random variables with respective distributions $F_1,\ldots,F_n$ in an atomless probability space $(\Omega,\mathcal A,{\mathbb{P}})$ . Formally, denote the set of possible distributions of the sum by
where $X \sim F$ means that the cumulative distribution function (cdf) of a random variable X is F.
The question of characterizing $\mathcal D_n$ , although simply stated, is a challenging open question. Generally, it is not easy to determine whether a given distribution G is in $\mathcal D_n$ , although many moment inequalities can be used as necessary conditions. In the recent literature, some papers partially address the question of $\mathcal D_n$ and provide sufficient conditions for $G\in \mathcal D_n$ ; see [Reference Bernard, Jiang and Wang2], [Reference Mao and Wang4], and [Reference Wang and Wang13]. A particular question is whether a point-mass belongs to $\mathcal D_n$ , which is referred to as the problem of joint mixability [Reference Wang, Peng and Yang14], and has found many applications in optimization and risk management.
However, there are no known results on the characterization of $\mathcal D_n$ , except for the trivial case where each of $F_1,\ldots,F_n$ is a Bernoulli distribution in low dimension. Perhaps the attempt to characterize $\mathcal D_n$ for generic $F_1,\ldots,F_n$ is too ambitious. In this paper, we focus on the very special case where $F_1,\ldots,F_n$ are standard uniform distributions $\rm U[0,1]$ . This problem might look naive at first glance, but with the technical challenges we shall see in this paper, the characterization of $\mathcal D_n$ is highly non-trivial even for uniform distributions.
As a well-known fact, for any random variables $X_1,\ldots,X_n$ , their sum is dominated in convex order by $X_1^c+\dots+X_n^c$ , where $X^c_i$ is identically distributed as $X_i$ , $i=1,\ldots,n$ , and $X^c_1,\ldots,X^c_n$ are comonotonic. Hence, the set $\mathcal D_n(F_1,\ldots,F_n) $ is contained in the set $\mathcal C_n$ of distributions that are dominated in convex order by the distribution of the comonotonic sum (for the precise definitions, see Section 2). These two sets are asymptotically equivalent after normalization, as shown in [Reference Mao and Wang4]. One naturally wonders whether they coincide for a finite n.
The main results of this paper can be summarized below in two distinct cases. Recall that $F_1,\ldots,F_n$ are standard uniform distributions $\rm U[0,1]$ . For dimension $n=2$ , the sets $\mathcal D_n$ and $\mathcal C_n$ are not equivalent, and a complete determination of $\mathcal D_2$ for uniform margins is still unclear. In Section 3 we provide four results on equivalent conditions for various types of distributions to be in ${\mathcal D}_n$ , including unimodal, bi-atomic, and tri-atomic distributions, and distributions dominating a proportion of a uniform one. In Section 4 we are able to analytically characterize the set $\mathcal D_n$ for dimension $n\ge 3$ by showing that $\mathcal D_n=\mathcal C_n$ . This result came as a pleasant surprise to us, since it is well known that the dependence structure gets much more complicated as the dimension grows. As far as we are aware, this result is the first full characterization of $\mathcal D_n$ for any type of continuous marginal distribution.
For applications of the problem of possible distributions of the sum with specified marginal distributions, we refer to [Reference Bernard, Bondarenko and Vanduffel1], [Reference Embrechts, Puccetti and Rüschendorf3], [Reference McNeil, Frey and Embrechts5], and [Reference Rüschendorf9]. A particular problem on the sum of standard uniform random variables is the aggregation of p-values from multiple statistical tests, which are uniform by definition under the null hypothesis. These p-values, as obtained from different tests, typically have an unspecified dependence structure, and hence it is important to understand the possible distributions of the aggregated p-value; see [Reference Vovk and Wang12].
We remark that the determination of whether $\mathcal D_n=\mathcal C_n$ for general margins is unclear. Note that the determination of $\mathcal D_n=\mathcal C_n$ for a given tuple of margins requires more than the determination of joint mixability of the margins, and the latter is known to be an open question in general. We conjecture that for general margins on bounded intervals with sufficient smoothness, there is a dimension n above which $\mathcal D_n$ equals $\mathcal C_n$ , but this is out of reach by current techniques, as our proofs rely heavily on the specific form of uniform distributions.
2. Preliminaries and notation
In this paper, for any (cumulative) distribution function F, we let $F^{-1}(t) = \inf\{x \colon F(x) \geq t\}$ , $t \in (0, 1]$ , denote the quantile function of F. Let $\mathcal X$ denote the set of integrable random variables and let $\mathcal F$ be the set of distributions with finite mean. The terms ‘distributions’ and ‘distribution functions’ are treated as identical in this paper. For $F \in \mathcal F$ , ${Supp}(F)$ is the essential support of the distribution measure induced by F, which will be referred to as the support of F. For any distributions F, $G \in \mathcal F$ , we let $F\oplus G$ denote the distribution with quantile function $F^{-1}(t) + G^{-1}(t)$ , $t\in [0,1]$ . For a distribution F, $\mu(F)$ denotes the expectation of F. Throughout, ${\lceil x \rceil}$ and ${\lfloor x \rfloor}$ represent the ceiling and the floor of $x\in {\mathbb{R}}$ , respectively. A density function f is unimodal if there exists $a\in {\mathbb{R}}$ such that f is increasing on ${(-\infty, a] }$ and decreasing on $[a,\infty)$ .
The set $\mathcal D_n$ is related to the notion of convex order. A distribution $F\in \mathcal F$ is smaller than $G\in\mathcal F$ in convex order, denoted by $F \preceq_{cx} G$ , if
provided that both expectations exist (finite or infinite). Standard references for convex order are [Reference Müller and Stoyan6] and [Reference Shaked and Shanthikumar10]. For a given distribution $F \in \mathcal F$ , let $\mathcal {C}(F)$ denote the set of all distributions dominated by F in convex order, that is,
Checking whether a distribution G is in $\mathcal C(F)$ can be conveniently done using an equivalent condition (e.g. [10, Theorem 3.A.1]) is
As the focus of this paper is the distribution of the sum of uniform random variables, we use the following simplified notation. For $n\in {\mathbb{N}}$ and $x\in {\mathbb{R}}_+$ , write
where ${\textrm{U}}[a,b]$ stands for the uniform distribution over an interval $[a,b]\subset {\mathbb{R}}$ .
Below we list some basic properties of the sets $\mathcal D_n(\cdot)$ and $\mathcal C(\cdot)$ ; they also hold for $\mathcal D^{\textrm{U}}_n$ and $\mathcal C^{\textrm{U}}_x$ . First, note that if all distributions $F_1,\ldots,F_n$ are shifted by some constants or scaled by the same positive constant, then the elements in $\mathcal D_n(F_1,\ldots,F_n)$ are also simply shifted or scaled. Moreover, $\mathcal D_n(F_1,\ldots,F_n)$ is symmetric in the distributions $F_1,\ldots,F_n$ . These facts allow us to conveniently exchange the order of the distributions $F_1,\ldots,F_n$ and normalize these distributions by shifts and a common scale. For given distributions $F_1,\ldots,F_n \in\mathcal F$ , the distribution $F_1\oplus \dots \oplus F_n$ is the maximum in convex order in the set $\mathcal D_n(F_1,\ldots,F_n)$ . This fact is summarized in the following lemma.
Lemma 1. For $F_1,\ldots,F_n\in\mathcal F$ , $\mathcal D_n(F_1,\ldots,F_n)\subset \mathcal C(F_1\oplus \dots \oplus F_n)$ .
Lemma 1 can be equivalently stated as follows. If $X_1\sim F_1,\ldots, X_n \sim F_n$ and F is the distribution of $X_1+\dots+X_n$ , then $F\preceq_{cx} F_1 \oplus \dots \oplus F_n$ . For a history of this result, see, for instance, [Reference Puccetti and Wang7]. In particular, if $F \in \mathcal D_n(F_1,\ldots,F_n)$ , then the mean of F is fixed and equal to the sum of the means of $F_1,\ldots,F_n$ .
In view of Lemma 1, it would be natural to investigate when the two sets coincide, that is, $\mathcal D_n (F_1,\ldots,F_n) = \mathcal C(F_1\oplus\dots\oplus F_n)$ . Note that for a given $G\in \mathcal F$ , the determination of $G\in \mathcal C(F)$ can be analytically checked with its equivalent condition (1). Hence, if the above two sets coincide, then we have an analytical characterization of $\mathcal D_n (F_1,\ldots,F_n)$ . In the case of uniform distributions, one wonders whether ${\mathcal D}^{\textrm{U}}_n = {\mathcal C}^{\textrm{U}}_n$ , noting that ${\mathcal D}^{\textrm{U}}_n \subset {\mathcal C}^{\textrm{U}}_n$ always holds. Unfortunately, as shown in [Reference Mao and Wang4] by a counter-example, in the simple case $n=2$ , ${\mathcal D}_2^{\textrm{U}}$ is an essential subset of ${\mathcal C}^{\textrm{U}}_2$ ; see Theorem 3 in Section 3 for distributions in ${\mathcal C}^{\textrm{U}}_2$ but not in ${\mathcal D}_2^{\textrm{U}}$ .
Some basic properties of the set ${\mathcal D}_n(\cdot)$ are given in the following lemma.
Lemma 2. For any $F_1,\ldots, F_n\in \mathcal F$ , the set ${\mathcal D}_n(F_1,\ldots,F_n)$ is non-empty, convex, and closed with respect to weak convergence.
Another simple fact is that a distribution in ${\mathcal D}_n(F_1,\ldots, F_n)$ has to have the correct support generated by the marginal distributions. That is, for any $F_1,\ldots, F_n\in \mathcal F$ and $G\in {\mathcal D}_n(F_1,\ldots,F_n)$ , we have $\operatorname{Supp}(G)\subset \sum_{i=1}^n \operatorname{Supp}(F_i)$ . For more properties on the sets ${\mathcal D}_n(\cdot)$ and $\mathcal C(\cdot)$ , see [Reference Mao and Wang4].
A useful concept for our analysis of $\mathcal D_n(\cdot)$ is the joint mixability introduced in [Reference Wang, Peng and Yang14]. An n-tuple of distributions ${(F_1,\ldots,F_n)\in \mathcal F^n}$ is said to be jointly mixable (JM), if $ \mathcal D_n(F_1,\ldots,F_n)$ contains a point-mass $\delta_K$ , $K\in {\mathbb{R}}$ . This point-mass is unique if the distributions $F_1,\ldots,F_n$ have bounded supports. If $(F_1,\ldots,F_n)$ is JM and $F_1=\dots=F_n=F$ , then we say that F is n-completely mixable (n-CM).
3. Sums of two standard uniform random variables
In this section we look at the sum of two ${\textrm{U}}[0, 1]$ random variables. Unfortunately, as explained above, a full characterization of ${\mathcal D}_2^{\textrm{U}}$ appears difficult to obtain. We provide characterization results for four different types of distributions to be in ${\mathcal D}_2^{\textrm{U}}$ . We first present the main results on ${\mathcal D}_2^{\textrm{U}}$ in Section 3.1. Their proofs will be given in Section 3.2. Before presenting our main findings, we summarize some existing results on ${\mathcal D}_2^{\textrm{U}}$ . These facts can be derived from existing results on joint mixability in [Reference Wang and Wang13].
Proposition 1. We have the following.
${\mathcal D}_2^{U}\subsetneq {\mathcal C}^{U}_2$ .
Let F be any distribution with a monotone density function on $\operatorname{Supp}(F)$ . Then $F\in {\mathcal D}_2^{\textrm{U}}$ if and only if $\operatorname{Supp}(F)\subset [0,2]$ and F has mean 1.
Let F be any distribution with a unimodal and symmetric density function on $\operatorname{Supp}(F)$ . Then $F\in {\mathcal D}_2^{\textrm{U}}$ if and only if $\operatorname{Supp}(F)\subset [0,2]$ and F has mean 1.
Remark 1. For a uniform distribution on an interval of length a, ${\textrm{U}}[1-\frac{a}{2},1+\frac{a}{2}]\in\mathcal D_2^{\textrm{U}}$ if and only if $a\in [0,2]$ , a special case of (ii) and (iii) of Proposition 1. The case ${\textrm{U}}[\frac{1}{2},\frac{3}{2}]\in \mathcal D_2^{\textrm{U}}$ is shown in [Reference Rüschendorf8], and the general case $a\in [0,2]$ is shown in [Reference Wang and Wang13].
3.1. Main results
As a first new result in this paper, we show that the class of distributions with a unimodal density with the correct mean is contained in ${\mathcal D}_2^{\textrm{U}}$ .
Theorem 1. Let F be a distribution with a unimodal density on [0, 2] and mean 1. Then $F\in {\mathcal D}_2^{\textrm{U}}$ .
Both the two previous results in Proposition 1(ii,iii) are special cases of Theorem 1; thus Theorem 1 generalizes the existing results derived from [Reference Wang and Wang13].
The second result of the paper concerns the class of distributions which dominate a proportion of a uniform distribution.
Theorem 2. Let F be a distribution supported in $[a,a+b]$ with mean 1 and density function f. If there exists $h>0$ such that $f>3b/(4h)$ on $[1-h,1+h]$ , then $F\in { {\mathcal D}}_2^{\textrm{U}} $ .
In Theorem 2, the condition that the density of F dominates $3b/2$ times that of ${\textrm{U}}[1-h, 1+h]$ immediately implies the following admissible ranges of a, b, and h: $a\geq 1/3$ , $b\leq 2/3$ , and $h\leq 1/3$ . Hence, $\textrm{Supp}(F)\subset [0,2]$ , which is obviously necessary for $F\in {\mathcal D}_2^{\textrm{U}}(1, 1)$ . Theorem 2 also immediately implies the following fact: for a distribution F with mean 0 and bounded support, if F has a positive density $f > \varepsilon$ in a neighbourhood of 0 for some $ \varepsilon> 0$ , then $F \in {\mathcal D}_2(\textrm U[-m,m],\textrm U[-m,m])$ for sufficiently large $m > 0$ .
In addition to the two results on continuous distributions, we analyse discrete distributions. We shall obtain two results, one characterizing bi-atomic distributions in ${\mathcal D}_2^{\textrm{U}} $ , and one characterizing equidistant tri-atomic distributions in ${\mathcal D}_2^{\textrm{U}} $ .
Theorem 3. Let F be a bi-atomic distribution with mean 1 supported on $\{a,a + b\}$ with $b>0$ . Then $F \in {\mathcal D}^{\textrm{U}}_2 $ if and only if $1/b \in {\mathbb{N}}$ .
For a given $a\in [0,1)$ and $a+b\in (1,2]$ , there is a unique distribution on $\{a,a+b\}$ with mean 1. Hence, all the bi-atomic distributions that belong to ${\mathcal D}^{\textrm{U}}_2 $ have the corresponding distribution measures
Note that many bi-atomic distributions supported on $\{a,a+b\}$ are in ${\mathcal C}^{\textrm{U}}_2$ but not in ${\mathcal D}_2^{\textrm{U}} $ , as long as $1/b\not\in {\mathbb{N}}$ . For example, one can choose a bi-atomic distribution F with equal probability on $\{1-1/\pi,1+ 1/\pi\}$ , and easily see that $F\in{\mathcal C}^{\textrm{U}}_2$ , whereas from Theorem 3 we find that F is not in ${\mathcal D}_2^{\textrm{U}} $ . Thus, Theorem 3 implies ${\mathcal D}_2^{U} \subsetneq {\mathcal C}^{U}_2$ , a fact noted in [Reference Mao and Wang4].
For a tri-atomic distribution F, write $F = (f_1, f_2, f_3)$ , where $f_1, f_2, f_3$ are the probability masses of F. Note that, given three points, the set of tri-atomic distributions with mean 1 has one degree of freedom. For tractability, we study the case of F having an equidistant support in the form of $\{a-b,\,a,\,a+b\}$ for some $b>0$ . We only consider the case $b\leq a \leq 1$ since the case $a > 1$ is symmetric.
To state our characterization of tri-atomic distribution in $\mathcal D_2^{\textrm U}$ , we introduce the following notation. For $x > 0$ , define a measure of non-integrity:
Obviously $\lceil x \rfloor=0 \Leftrightarrow x \in {\mathbb{N}}$ .
Theorem 4. Suppose that $F = (f_1, f_2, f_3)$ is a tri-atomic distribution with mean 1 supported in $\{a - b, a, a + b\}$ and $0<b\leq a \leq1$ . Then $F \in \mathcal D_2^{\textrm U} $ if and only if one of the following three cases holds.
$a=1$ and $f_2\geq \lceil \frac{1}{(2b)}\rfloor$ .
$a<1$ and $\frac{1}{(2b)}\in {\mathbb{N}}$ .
$a<1$ , $\frac{1}{(2b)}-\frac{1}{2}\in {\mathbb{N}}$ and $f_2\geq a+b-\frac{1}{2}$ .
The corresponding distributions in Theorem 4 are summarized below. Write $c = a/b +1- 1/(2b)$ . Here $\operatorname{cx} (x, y)$ stands for the convex set generalized by some vectors x,y.
${(f_1,f_2,f_3)\in \operatorname{cx} \{(0,1,0), \frac{1}{2}(1-\lceil \frac{1}{2b} \rfloor, 2\lceil \frac{1}{2b} \rfloor, 1- \lceil \frac{1}{2b} \rfloor)\}}$ .
${(f_1,f_2,f_3)\in {cx} \{(0,c,1-c), \frac{1}{2}(c, 0, 2-c)\}}$ .
${(f_1,f_2,f_3)\in {cx} \{(0,c,1-c), \frac{1}{2}(c(1-b), 2bc, 2 - c(1-b))\}}$ .
3.2. Proofs of the main results
To prove Theorem 1, we need the following lemma. To state it, we introduce the notion of special simple unimodal functions. A function h is called a special simple unimodal (SSU) function on $[a,a+n)$ , $a\in{\mathbb{R}}$ , $n\in{\mathbb{N}}$ , if it is unimodal, and h(x) is a constant on $[a+k-1, a+k)$ for each $k=1,\ldots,n$ .
Lemma 3. Let F be a distribution function with density function f and support $[a,a+n)$ , $a\in{\mathbb{R}}_-$ , $n\in{\mathbb{N}}$ . Suppose that F has mean 0 and f is a SSU function on $[a,a+n)$ . Then, for $c\geq \max\{a+n,-a\}/2$ , we have $F\in {\mathcal D}_2({\textrm{U}}[-c,c],{\textrm{U}}[-c,c])$ .
Proof. We show the result by induction. For $n=1$ , then F is the uniform distribution on $[a,a+1)$ with mean 0. This means $a=-0.5$ , that is, F is the distribution of ${\textrm{U}}[-0.5,0.5]$ , and $c\geq 0.5$ . By [13, Theorem 3.1], we know that ${\textrm{U}}[-0.5,0.5]$ , ${\textrm{U}}[-c,c]$ , and ${\textrm{U}}[-c,c]$ are jointly mixable as the mean inequality $-0.5-2c+2c\leq 0 \leq 0.5+2c-2c$ is satisfied. This means $F\in {\mathcal D}_2({\textrm{U}}[-c,c],{\textrm{U}}[-c,c])$ .
Next, we assume that the result holds for $n\leq k$ and show that it holds for $n=k+1$ . Without loss of generality, we assume $a+n>-a$ . Otherwise consider the distribution of $X^*=-X$ with $X\sim F$ . Define a distribution H with density function $h\colon [a,a+n)\to{\mathbb{R}}_+$ defined as
It can be easily verified that it has mean 0 and $\alpha\geq 1/n\geq \beta>0$ as $2a+n>0$ . That is, h is a decreasing density function on $[a,a+n)$ with mean 0. Then, by [13, Theorem 3.2], H, ${\textrm{U}}[-c,c]$ and ${\textrm{U}}[-c,c])$ are jointly mixable. Hence, we have $H\in {\mathcal D}_2({\textrm{U}}[-c,c],{\textrm{U}}[-c,c])$ .
Denote $a_i = f(a+i)$ , $i=0,\ldots, n-1$ . Since f is unimodal, without loss of generality, assume that $a_1\leq \dots\leq a_k\geq \dots \geq a_n$ for some $k\in \{0,\ldots, n-1\}$ . Let
By contradiction, it immediately follows from $\sum_{i=1}^na_i=1$ and $\alpha+(n-1)\beta=1$ that $\lambda \leq 1$ . If $\lambda=1$ , then $a_1=\alpha$ and $\alpha_i=\beta$ , $i=2,\ldots,n$ , that is, $F=H\in {\mathcal D}_2({\textrm{U}}[-c,c],{\textrm{U}}[-c,c])$ , which shows the statement in the lemma. Next, we consider the case $\lambda<1$ . We first assert that the sequence
is unimodal such that either $b_1$ or $b_n$ is 0. To see it, we only need to show it is unimodal by the definition of $\lambda$ . We consider the following two cases.
If $k\geq 2$ , then $a_1\leq a_2$ and hence $b_1=a_1-\lambda\alpha \leq a_2 -\lambda\beta =b_2$ as $\alpha\geq \beta$ . Also, note that $b_2\leq \dots\leq b_k\geq \dots \geq b_n$ . Hence, the sequence $\{b_i,\,i=1,\ldots,n\}$ is unimodal.
If $k=1$ , then $a_1\geq a_2\geq \dots\geq a_n$ , and hence $b_2\geq \dots\geq b_n$ . No matter whether $b_1\geq b_2$ or $b_1\leq b_2$ , we have that $\{b_i,\,i=1,\ldots,n\}$ is unimodal.
Define
It is easy to check that $F_0$ is a distribution function with a density function $f_0$ taking value $b_i/(1-\lambda)$ on the interval $[a,a+i)$ , $i=1,\ldots,n$ . By the above observations on the sequence $\{b_i,\,i=1,\ldots,n\}$ , we know that $F_0$ is a distribution with mean 0 and with support on a subset of $[a,a+n-1)$ or $[a+1, a+n)$ . Then, by induction, we have $F_0\in {\mathcal D}_2({\textrm{U}}[-c,c], {\textrm{U}}[-c,c])$ . Since $F = (1 -\lambda)F_0 + \lambda H$ , and ${\mathcal D}_2$ is a convex set, we have $F\in {\mathcal D}_2({\textrm{U}}[-c,c], {\textrm{U}}[-c,c])$ .
Now we are ready to prove Theorem.1.
Proof of Theorem 1. Let f denote the density function of F, which is unimodal. Then there exists $x_0\in [0,2]$ such that $f(x_0)=\max\{ f(x),x\in[0,2]\}$ . Let X and Y be two independent random variables such that $X\sim F$ and $Y\sim {\textrm{U}}[0,1]$ , and define $X_m=\lfloor m(X-1)\rfloor +Y$ , $m\in{\mathbb{N}}$ . Then the density function of $X_m$ , denoted by $h_m$ , is given by
Let $k=\lfloor mx_0\rfloor$ . Then we have
and, since f(x) is unimodal on $[k/m,(k+1)/m]$ ,
That is, the sequence $\{p_k,\,k=1,\ldots,2m\}$ is unimodal. It can be verified that
which lies in $[-0.5,0.5]$ . Then the random variable $Y_m\,{:\!=}\,X_m-\mu(X_m)$ has mean 0 and takes values in a subset of $[-m-0.5,m+0.5]$ . By Lemma 3, the distribution of $Y_m$ belongs to ${\mathcal D}_2({\textrm{U}}[-c,c],{\textrm{U}}[-c,c])$ with $c=m/2+0.25$ . Then obviously, the distribution of $Y_m/(2m\,{+}\,0.5)+1$ belongs to ${\mathcal D}_2^{{\textrm{U}}} $ . Note that $Y_m/(2m+0.5)+1$ converges to X in the $L^\infty$ -norm as $m\to\infty$ . Hence, by the closure of ${\mathcal D}_2^{{\textrm{U}}} $ with respect to the $L^\infty$ -norm, we have $F\in{\mathcal D}_2^{{\textrm{U}}} $ . This completes the proof.
To prove Theorems 2–4, we need the following lemma.
Lemma 4. Let Z be a random variable with distribution F supported in $\{b-k,\, b-k+1,\ldots,b\}$ , $k\in{\mathbb{N}}$ , $b\in{\mathbb{R}}$ , satisfying
Then the following statements hold.
If $F\in {\mathcal D}_2({\textrm{U}}[0,T],{\textrm{U}}[-T,0])$ , then at least one of b and T is an integer.
If $k=1$ , then $F\in {\mathcal D}_2({\textrm{U}}[0,T],{\textrm{U}}[-T,0])$ if and only if $b\in (0,1)$ and $T\in{\mathbb{N}}$ .
If $k=2$ , then $F\in {\mathcal D}_2({\textrm{U}}[0,T],{\textrm{U}}[-T,0])$ if and only if one of the following three cases holds:
$b=1$ , $p_2=p_0=(1-p_1)/2$ , $p_1\geq r/T$ with $T=2m\pm r$ with $m\in{\mathbb{N}}$ and $r\in (0,1)$ , $b\in (0,1)\cup (1,2)$ , T is even, $p_1\in (0,\min\{b,2-b\})$ , $p_2=(b-p_1)/2$ , and $p_0=1-(b+p_1)/2$ ,
$b\in (0,1)\cup (1,2)$ , T is odd, $p_1\in (b^*/T,b^*)$ with $b^*=\min\{b,2-b\}$ , $p_2=(b-p_1)/2$ , and $p_0=1-(b+p_1)/2$ .
Proof. First we introduce the notation. For any random variable X and any set $L\subset {\mathbb{R}}$ , define random events
and a function
Note that $Z \equiv b \ \text{mod}\, 1$ , ${\mathbb{E}}[Z]=0$ and hence $b-k\leq 0\leq b$ . If $F\in {\mathcal D}_2({\textrm{U}}[0,T],{\textrm{U}}[-T,0])$ , then there exist two random variables $X\sim {\textrm{U}}[0,T]$ and $Y\sim {\textrm{U}}[-T,0]$ such that $Z=X+Y$ a.s. Since $Z=X+Y \equiv b \ \text{mod}\, 1$ a.s., we have $A_X(L)=A_Y(b-L)$ a.s. for any $L\subset {\mathbb{R}}$ , where $b-L=\{b-x\colon x\in L\}$ . We first show that at least one of b and T is an integer. To this end, assume that T is not an integer. Then there exists $\ell \in {\mathbb{N}}$ such that $T=\ell+t$ with $t\in(0,1)$ . Note that $X\sim {\textrm{U}}[0,T]$ . We have
By the definition of $A_X(L)$ and $Y\stackrel{\rm d}= -X$ , we have $A_Y([x-\delta,x+\delta]) =A_{-X}([x-\delta,x+\delta])= A_{X}([-x-\delta,-x+\delta])$ a.s., and thus $g_Y(x)=g_X(-x)$ for $x\in{\mathbb{R}}$ . Also, note that $A_X(L)=A_Y(b-L)$ a.s., which implies $g_X(x)=g_Y(b-x)$ . Therefore, we have
Note that there exists $x\in{\mathbb{R}}$ such that $x\ \text{mod}\, 1\in (0,t)$ and $x-b\ \text{mod}\, 1\in (t,1)$ , which contradicts the formula of $g_X$ . Hence, b must be an integer.
Next we consider the two cases $k=1$ and $k=2$ . For $k=1$ , note that for any $b\in{\mathbb{N}}$ , $\mu(F)$ could not be 0. Hence, we only need to show $F\in {\mathcal D}_2({\textrm{U}}[0,T],{\textrm{U}}[-T,0])$ when $b\in (0,1)$ and $T\in {\mathbb{N}}$ . By $\mu(F)=0$ , we have $p_1=b$ and $p_0=1-b$ . Let $X\sim {\textrm{U}}[0,T]$ and define a random variable Y such that
Then it is easy to see that $Y\sim {\textrm{U}}[-T,0]$ and $X+Y$ has the distribution F. Thus, we have $F\in\mathcal D_2({\textrm{U}}[0,T], {\textrm{U}}[-T,0])$ .
If $k=2$ , by the necessity condition that at least one of b and T is an integer, we consider the following three cases. Without loss of generality, assume $p_i={\mathbb{P}}(Z=b-i)>0$ , $i=0,1,2.$
(a) If b is an integer, for the mean-constraint to be satisfied, we have $b=1$ and
Let $T= 2m\pm r$ with $r\in [0,1]$ and $m\in{\mathbb{N}}$ . We only need to find the smallest value of $p_1$ such that $F\in {\mathcal D}_2({\textrm{U}}[0,T],{\textrm{U}}[-T,0])$ as $\delta_0\in {\mathcal D}_2({\textrm{U}}[0,T],{\textrm{U}}[-T,0])$ and ${\mathcal D}_2({\textrm{U}}[0,T],{\textrm{U}}[-T,0])$ is closed under mixture, where $\delta_0$ is the point-mass at 0. Assume that there exist $X\sim {\textrm{U}}[0,T]$ and $Y\sim {\textrm{U}}[-T,0]$ such that $Z=X+Y$ a.s. If $T=2m+r$ , then since b is an integer, we have
We let
where $A_1=\cup_{k=1}^{m} (r+2k-3,2k-1)$ and $A_2= \cup_{k=1}^{m} (r+2k-1,2k)$ . Then $[X+Y|X\in A_1\cup A_2]=[X+Y|A_X((r,1))]$ takes values on $\{-1,1\}$ . On the other hand, note that
Then, on the set $A_X((0,r))$ , $X+Y$ could not only take values on $\{-1,1\}$ . There is at least one $k\in\{0,\ldots,2m\}$ such that $X+Y=0$ on $\{X\in (k,k+r)\}$ . Hence, $p_1\geq (2m+1)r/(2m+1)T=r/T$ . We next show that $r/T$ can be attained by $p_1$ for $F\in {\mathcal D}_2({\textrm{U}}[0,T],{\textrm{U}}[-T,0])$ . It suffices to let $[Y|X\in (2m,2m+r)]=-X$ a.s., and for $k=0,\ldots,m-1$ ,
In this case, we have $p_1=r/T.$ By symmetry, we can also get the same result if $T=2m-r$ . To see this, note that almost surely
We let
where $A_1=\cup_{k=0}^{m-1} (2k,2k+1-r) $ and $A_2= \cup_{k=0}^{m-1} (2k+1,2k+2-r)$ . Then $[X+Y|X\in A_1\cup A_2]=[X+Y|A_X((0,1-r))]$ takes values on $\{-1,1\}$ . On the other hand, note that almost surely
Then, on the set $A_X((1-r,1))$ , $X+Y$ could not only take values on $\{-1,1\}$ . There is at least one $k\in\{1,\ldots,2m-1\}$ such that $X+Y=0$ on $\{X\in (k-r,k)\}$ . Hence, $p_1\geq (2m\,{-}\,1)r/(2m-1)T=r/T$ . We next show that $r/T$ can be attained by $p_1$ for $F\in {\mathcal D}_2({\textrm{U}}[0,T], {\textrm{U}}[-T,0])$ . It suffices to let $[Y|X\in (2m-1-r,2m-1)]=-X$ a.s., and for $k=1,\ldots, m-1$ ,
In this case, we have $p_1=r/T.$ Hence, when b is an integer we have $Z\in {\mathcal D}_2({\textrm{U}}[0,T],{\textrm{U}}[-T,0])$ if and only if $b=1$ and ${\mathbb{P}}(Z=-1)={\mathbb{P}}(Z=1)\leq 1/2-r/(2T)$ .
(b) If b is not an integer and T is an even integer, without loss of generality, assume $b\in(0,1)$ as the case $b\in (1,2)$ can be discussed similarly by considering the symmetric distribution on $\{-b,1-b,2-b\}$ and noting that ${\mathcal D}_2({\textrm{U}}[0,T],{\textrm{U}}[-T,0])$ is closed under symmetric transform. To make sure ${\mathbb{E}}[Z]=0$ , we have $p_0-p_2=1-b$ . By part (i), we know that the distribution, denoted by $F_1$ , on $\{b-1,b\}$ with mean 0 belongs to ${\mathcal D}_2({\textrm{U}}[0,T],{\textrm{U}}[-T,0])$ , which is also closed under mixture. Hence, $F\in {\mathcal D}_2({\textrm{U}}[0,T],{\textrm{U}}[-T,0])$ implies
where the second equality follows from $p_0-p_2=1-b$ . Therefore, we only need to find the smallest and largest values of $p_1$ such that $p_0+p_1+p_2=1$ , $p_0-p_2=1-b$ and $F\in {\mathcal D}_2({\textrm{U}}[0,T],{\textrm{U}}[-T,0])$ .
Note that $F\in {\mathcal D}_2({\textrm{U}}[0,T],{\textrm{U}}[-T,0])$ is equivalent to the existence of $X\sim {\textrm{U}}[0,T] $ and $Y\sim{\textrm{U}}[-T,0]$ such that $Z=X+Y$ a.s. Then, by $A_X(L)=A_Y(b-L)$ a.s., for any $L\subset {\mathbb{R}}$ , we have
It is easy to verify that ${\mathbb{E}}[X+Y| A_X((b,1)) ]=b$ . Note that the $X+Y$ takes values on $\{b-2,b-1,b\}$ . Then $[X+Y|A_X((b,1))]=b$ a.s. This implies that, on $A_X((b,1))$ , $X+Y$ is not equal to $b-1$ a.s., which in turn implies that b is an upper bound of $p_1$ . On the other hand, note that
Letting $[Y| A_X((0,b))] = b-X-1$ a.s. and $[Y| A_X((b,1))] = b-X$ a.s. yields that b is the largest value of $p_1$ .
To find the smallest value of $p_1$ , we consider the two cases that T is even and odd. If T is even, then $T=2m$ for some $m\in{\mathbb{N}}$ . On the set $A_X(0,b)$ , for $k=0,1,\ldots,m-1$ , we let
In this case $p_1$ is zero, which is the smallest possible value of $p_1$ .
(c) If b is not an integer and T is an odd integer, then, as for part (b), we only need to find the largest and smallest values of $p_1$ for the case $b\in (0,1)$ . The largest value of $p_1$ is b. To find the smallest value, note that $T=2m+1$ for some $m\in {\mathbb{N}}$ . On the set $A_X(0,b)$ , we let $[Y|X\in (2m,b+2m)] = b-X$ a.s., and for $k=0,1,\ldots,m-1$ ,
In this case $p_1=b/T$ , which is the smallest possible value of $p_1$ . This is due to $ [X+Y|X\in (2m,b+2m)] >-1>b-2$ as $\{X\in (2m,b+2m)\}\subset A_Y(0,b)$ and $[Y|A_Y((0,b))]\geq -2m-1$ a.s.
Combining the above three cases, we complete the proof for the case of $k=2$ .
Proof of Theorem 2. Note that $3b/4h \times 2h =3b/2 \leq1$ , hence $b\leq 2/3$ . Therefore, $[a,a\,{+}\,b]\subset [0,2]$ . Denote $\lambda=3b/2\geq 3h$ and let H be the distribution function of ${\textrm{U}}[1-h,1+h]$ . Define $G = (F - \lambda H)/(1-\lambda)$ . By the assumption on F, G is also a distribution with positive density and mean 1. There exists a sequence of distribution functions $\{G_n\}_{n\geq 2}$ with finite support on $[a,a+b]$ and mean 1 which converges to G in distribution as $n\to\infty$ . Define $F_n = \lambda H+(1-\lambda)G_n$ , $n\geq 2$ . Then $F_n$ converges to F in distribution as $n\to\infty$ . Note that ${ {\mathcal D}}_2^{\textrm{U}} $ is closed with respect to weak convergence by Lemma 2. We only need to show $F_n\in { {\mathcal D}}_2^{\textrm{U}} $ for $n\geq 2$ . Without loss of generality, assume
We show it by induction on n. Note that $a\leq 1-h\leq 1+h\leq b$ . For $n=2$ , without loss of generality, let $c,d>0$ be such that $x_1=1-c$ and $x_2=1+d>0$ , $3(c+d)/2\leq\lambda$ , and then
Since $ 1 \geq \lambda \geq c + d + h $ , we have
Then there exists some integer $k \geq 1 $ such that
That is, $ 1- k (c+d) \geq h $ and $ k(c+d) \geq 1-\lambda .$ Denote
Note that the length of the support of $ {{\textrm{U}}} [0,k(c+d) ] $ ( ${{\textrm{U}}}[ 1-k(c+d), 1 ] $ ) is a multiple of ${(1+d)-(1-c)}$ . By Lemma 4(i), we have $G_2\in \mathcal D_2 ( {{\textrm{U}}} [0,k(c+d) ] , {{\textrm{U}}}[ 1-k(c+d), 1 ] ) $ . Similarly, we have $ H \in \mathcal D_2 ( {{\textrm{U}}}[k(c+d) ,1], {{\textrm{U}}}[ 0 , 1-k(c+d)]) $ . Also, by Theorem 3.1, we know $H \in \mathcal D_2^{\textrm{U}} $ . It follows that
Suppose that $F_n=( 1- \lambda ) G_n + \lambda H \in \mathcal D_2^{\textrm{U}}$ for $n\leq k$ , and we aim to show that $F_{k+1}= (1- \lambda) G_{k+1} + \lambda H \in \mathcal D_2^{\textrm{U}} $ . Let $G_{k+1,1}$ be defined by (2) with $c=1-x_1>0$ and $d=x_{k+1}-1>0$ . That is,
Let $\alpha \,{:\!=}\, \min\{p_1(c+d)/d,p_n(c+d)/c\}$ . Then $G_{2\colon k+1} \,{:\!=}\, (G_{k+1} - \alpha G_{k+1,1})/(1-\alpha)$ is a distribution function with support on $\{x_1,\ldots,x_{k}\}$ or $\{x_2,\ldots,x_{k+1}\}$ and mean 1. Then we have
By induction, we have ${( 1- \lambda ) G_{k+1,1} + \lambda H \in \mathcal D_2^{\textrm{U}}}$ and ${( 1- \lambda ) G_{2\colon k+1}+ \lambda H\in \mathcal D_2^{\textrm{U}}}$ . Then, by the convexity of $\mathcal D_2^{\textrm{U}}$ from Lemma 2, we have $F_{k+1} \in \mathcal D_2^{\textrm{U}}$ . Thus, we complete the proof.
Proof of Theorems 3 and 4. We only give the proof of Theorem 4 as Theorem 3 can be proved similarly based on Lemma 4(ii). Note that the cumulative distribution function of a random variable Z belongs to $\mathcal D_2^{\textrm U}$ if and only if the cumulative distribution function of $T(Z-1)$ belongs to $\mathcal D_2({\textrm U}[0,T], {\textrm U}[-T,0])$ for $T\geq 0$ . Let Z be a random variable with distribution F and let $T\,{:\!=}\,1/b$ and $c \,{:\!=}\, (a+b-1)/b$ , that is, $b=1/T$ and $a=c/T+1-1/T$ . Then $Z_T\,{:\!=}\,T(Z-1)$ is a random variable satisfying
By Lemma 4(iii), we know that the cumulative distribution function of $Z_T$ belongs to $\mathcal D_2^{\textrm U}$ if and only if one of the three cases of Lemma 4(iii) holds by replacing b with c and T with $1/b$ . That is,
$a+b-1/b=1$ , $f_2\geq rb$ with $1/b=2m\pm r$ with $m\in{\mathbb{N}}$ and $r\in (-1,1)$ ,
${(a+b-1)/b<1}$ , $1/b$ is even, $f_2\in (0, (a+b-1)/b)$ , $f_3=(c-f_2)/2$ and $f_1=1-((a+b-1)/b+f_2)/2$ ,
${(a+b-1)/b<1}$ , $1/b$ is odd, $f_2\in (a+b-1,(a+b-1)/b)$ , $f_3=(c-f_2)/2$ and $f_1=1-(c+f_2)/2$ .
Note that under the constraint $0<b\leq a\leq 1$ , we have $c\leq 1$ ; $c=1$ is equivalent to $a=1$ ; $rb=\lceil 1/(2b) \rfloor$ ; $1/b$ is even if and only if $1/({2b})\in {\mathbb{N}}$ ; $1/b$ is odd if and only if $1/({2b}) -1 /2 \in {\mathbb{N}}$ . Also, by ${\mathbb{E}}[Z]=1$ , we have $f_3=f_1+(1-a)/b$ and thus $f_2=1-2f_1-(1-a)/b$ , which implies that $f_2\in (0, (a+b-1)/b)$ always holds. Hence, the statement in Theorem 3.1 holds.
4. Sums of three or more standard uniform random variables
In this section we aim to show that for $n \geq 3$ the two sets $\mathcal D_n^{\textrm{U}}$ and $\mathcal C^{\textrm{U}}_n$ are identical, in sharp contrast to the case of $n=2$ analysed in Section 3. We start with the bi-atomic distribution. Let F be the distribution function of a random variable X such that ${\mathbb{P}}(X=a)=p$ and ${\mathbb{P}}(X=b)=1-p$ with ${\mathbb{E}}[X]=1/2$ , $a<b$ and $0<p<1$ , and let $T_{n} (F)$ be the distribution function of nX. That is,
where $\delta_x$ denotes the point-mass at $x\in {\mathbb{R}}$ .
Lemma 5. Let F be a bi-atomic distribution on $\{a,b\}$ with $a<b$ . Then the following statements are equivalent.
$F{\preceq_{cx}} {{U}}[0,1]$ .
$b-a\leq 1/2$ and $\mu(F)=1/2$ .
$T_{n} (F) \in \mathcal D_n^{{{\textrm{U}}}}(1,\ldots,1)$ for $n\geq 3$ .
Proof. It is easy to verify (iii) $\Rightarrow$ (i). It suffices to show (i) $\Rightarrow$ (ii) $\Rightarrow$ (iii). Let X be a random variable having distribution F and ${\mathbb{P}}(X=a)=p=1-{\mathbb{P}}(X=b)$ . Note that under the constraint of (i), we must have $a,b\in (0,1)$ and $\mu(F)=1/2$ , which implies $0<a<1/2<b<1$ ; under the constraint of (ii), by $\mu(F)=1/2$ , we have $a<1/2<b$ . Then, by $b-a\leq 1/2$ , we also have $a>0$ and $b<1$ . Thus, in the following proof, we always assume $0<a<1/2<b<1$ .
Without loss of generality, we also assume $ 0 < a \leq 1-b \leq 1/2$ by symmetry. Otherwise, consider another random variable $X^*=1-X$ with distribution $F^*$ , and it suffices to note that $\mu(F^*)=1/2$ , $b^*-a^*=b-a$ , and $F{\preceq_{cx}} {{U}}[0,1]$ is equivalent to $F^*{\preceq_{cx}} {{U}}[0,1]$ .
(i) $\Rightarrow$ (ii) By $F{\preceq_{cx}} {{U}}[0,1]$ , we have ${\mathbb{E}}[X]=1/2$ and $ {\mathbb{E}}[(X-t)_+] \leq {(1-t)^2}/{2}, $ for $t\in [0,1]$ , that is,
To show (ii), we only need to show $b-a\leq 1/2$ , that is, the largest possible value of $b-a$ is $1/2$ . To do this, we fix the value of p and denote $\ell_1(t)\,{:\!=}\, (b-t)(1-p)$ , $t\in [a,b]$ and $\ell_2(t)\,{:\!=}\, {(1-t)^2}/{2}$ , $t\in [0,1]$ . Note that $\ell_1(t)$ is a linear function with fixed and constant derivative equal to $p-1$ and lies below the quadratic curve $\ell_2(t)$ for $t\in [a,b]$ . Also note that as a decreases or b increases, $\ell_1$ moves upwards. Hence, when $b-a$ attains its largest possible value, the line $\ell_1$ and the quadratic curve $\ell_2$ are tangent, and the tangent point satisfies the equation $\ell_2'(t)=t-1=p-1$ , that is, the tangent point is $t=p$ . Then, by $\ell_1(p)=\ell_2(p)$ , we get $b=1+p/2$ , which is the largest possible value of b. Then, by ${\mathbb{E}}[X]=1/2$ , we can get $a=p/2$ , which is the smallest possible value of a. Therefore, the largest possible value of $b-a$ is $1/2$ .
(ii) $\Rightarrow$ (iii) Let $G_{x,y}$ be the distribution of $ {{\textrm{U}}}[x,y]$ . It suffices to show that $G_{0,1}= (1\,{-}\,p) H_1 + p H_2$ such that both $H_1$ and $H_2$ are n-CM, $\mu(H_1)=a$ and $\mu(H_2)=b$ . We first define two distributions $H_1^*$ and $H_2^*$ for the following three cases, with $1-b\geq a$ in mind.
(a) If $a \geq 1/n$ , define $H_1^* = H_2^* = G_{0,1}$ , that is, the distribution function of ${{\textrm{U}}}[0,1]$ .
(b) If $1- b \geq 1/n \geq a $ , define
By ${\mathbb{E}}[X]=1/2$ , we have
where the inequality follows from $b-a\leq 1/2$ . Hence, $H_2^*$ is a distribution with positive density function. It is obvious that $ \mu(H_1^*) = {na}/{2} \geq a $ and thus $\mu (H_2^*) \leq b$ as $p\, \mu (H_1^*) + (1\,{-}\,p) \mu(H_2^*) = 1/2$ and $pa+(1-p)b=1/2$ .
(c) If $1-b< 1/n$ , then we have $n=3$ as $b-a\leq 1/2$ and $a\leq 1-b$ . Define
Note that
Hence, both $H_1^*$ and $H_2^*$ are distribution functions with positive densities. It is easy to calculate that
Note that
where the first inequality follows from $p>3b-2$ and $2(b-a)\leq 1$ , $A\stackrel{\rm sgn}=B$ represents that A and B have the same sign, and $\stackrel{\rm sgn}=$ is due to $1-b< 1/n$ and $n=3$ by the observations at the beginning of (c). Hence, we have $\mu(H_1^*)\geq a$ and similarly $\mu(H_2^*)\leq b$ .
In each of the above three cases, we have $G_{0,1}= p H_1^* + (1-p)H_2^*$ , $H_1^*$ has a decreasing density on [0,na] with $\mu(H_1^*) \geq a$ , and $H_2^*$ has an increasing density on $[1-n(1-b),1]$ with $\mu(H_2^*) \leq b$ .
On the other hand, let $H_1^0= G_{0,p}$ and $H_2^0 = G_{p,1}$ . Then it is obvious that $G_{0,1}=pH_1^0 + (1\,{-}\,p)H_2^0$ , $\mu(H_1^0) \leq a $ and $\mu(H_2^0)\geq b$ . Hence, we can find some $\alpha \in [0,1]$ such that $ \alpha \mu(H_1^*) + (1\,{-}\,\alpha) \mu(H_1^0) = a$ . Then the distribution $H_1\,{:\!=}\, \alpha H_1^* + (1-\alpha)H_1^0 $ is supported in [0,na] with decreasing density and $\mu(H_1)=a$ . By [13, Theorem 3.2], G is n-CM, that is, $\delta _{na} \in \mathcal D_n( G,\ldots,G )$ . Similarly, $H_2 = \alpha H_2^* + (1-\alpha)H_2^0 $ is supported in $[1-n(1-b),1]$ with an increasing density and $\mu(H_2)=b$ , which implies that $H_2$ is also n-CM. That is, $ \delta_ {nb} \in \mathcal D_n( H_2,\ldots,H_2)$ . By Lemma 2, we have
Thus, we complete the proof.
Now we are ready to show our main result in dimension $n\geq 3$ . It turns out that for standard uniform distributions, the two sets $\mathcal D_n(F_1,\ldots, F_n) $ and $ \mathcal C(F_1\oplus\dots\oplus F_n)$ in Lemma 2 coincide. The following theorem is, to the best of our knowledge, the first analytical characterization of $\mathcal D_n$ for continuous marginal distributions.
Theorem 5. For $n \geq3$ , we have $\mathcal D_n^{\textrm{U}} = \mathcal C^{\textrm{U}}_n$ .
Proof. As $\mathcal D_n^{\textrm{U}} \subseteq \mathcal C^{\textrm{U}}_n$ , it suffices to show $ \mathcal C^{\textrm{U}}_n\subseteq \mathcal D_n^{\textrm{U}},$ that is, for any distribution function F, $F{\preceq_{cx}} {{U}}[0,1]$ implies $\mathcal D_n^{\textrm{U}}(1/n)$ . Here and throughout the proof, we use the notation $ \mathcal D_n^{\textrm{U}}(x) = \mathcal D_n({\textrm{U}}[0,x],\ldots,{\textrm{U}}[0,x])$ for $x\in{\mathbb{R}}.$ For a distribution F, denote
where X is a random variable having the distribution function F. We first consider the special case that F is a distribution function of a discrete random variable $(a_1,p_1;\ldots;a_m,p_m)$ with $0\leq a_1<\dots<a_m\leq 1$ and $G_{x,y}$ is the distribution function of ${{\textrm{U}}}[x,y]$ , $x<y$ . By Lemma 4, we know the result holds for $m=2$ . Next, we show that it holds for general $m\geq 2$ by induction.
For general $m\geq 2$ , by $F\preceq_{cx} G_{0,1}$ , we have $W_F(t)\leq 0$ for $t\in[0,1]$ , that is, for $k=2,\ldots,m$ , we have
Next, we consider two cases.
(a) If there exists $t\in [a_{k-1}, a_{k})$ such that the equality of (4) holds, then $p_1+\dots+p_{k-1}=t$ . To see this, denote $\ell_1(t)\,{:\!=}\, (a_k-t)p_k+\dots+(a_m-t)p_m$ , and $\ell_2(t)={(1-t)^2}/{2}$ , $t\in [a_{k-1},a_k]$ . Note that $\ell_1$ is a linear function which lies below the decreasing quadratic curve $\ell_2$ . Thus, $\ell_1$ and the quadratic curve $\ell_2$ are tangent at the point t and the tangent point t satisfies the equation $t-1= -p_k-\dots-p_m$ , that is, $p_1+\dots+p_{k-1}=t$ . Let $X_1$ and $X_2$ be two random variables satisfying ${\mathbb{P}}(X_1=a_i)=p_i/t$ , $i=1,\ldots,k-1$ , and ${\mathbb{P}}(X_2=a_i)=p_i/(1-t)$ , $i=k,\ldots,m$ . Let $F_1$ and $F_2$ denote the distributions of $X_1$ and $X_2$ , respectively. It is easy to verify that $F=tF_1+(1-t)F_2$ , $F_1\preceq_{cx} G_{0,t}$ and $F_2\preceq_{cx} G_{t,1}$ . Then, by induction, we have $F_1\in \mathcal D_n^{\textrm{U}}(t/n)$ and $F_2\in \mathcal D_n({\textrm{U}}[t/n,1/n],\ldots,{\textrm{U}}[t/n,1/n])$ . That is, there exist $X_{11},\ldots,X_{1n}\sim {\textrm{U}}[0,t/n]$ and $X_{21},\ldots,X_{2n}\sim {\textrm{U}}[t/n,1/n]$ such that
Without loss of generality, assume that $X_{11},\ldots,X_{1n}$ are independent of $X_{21},\ldots,X_{2n}$ . Let A be a random event independent of $X_{11},\ldots,X_{1n}$ , $X_{21},\ldots,X_{2n}$ such that ${\mathbb{P}}(A)=p_1+\dots+p_{k-1}=t$ . Define
It is obvious that $Y_i\sim {\textrm{U}}[0,1]$ , $i=1,\ldots,n$ , and $Y_1+\dots+Y_n\sim F$ . This means that $F\in \mathcal D_n^{\textrm{U}}(1/n,\ldots, 1/n)$ .
(b) If the inequality of (4) is strict for every $t\in (0,1)$ , define two functions G and H by
where $\delta_{a}$ denotes the degenerate distribution at point a,
It is easy to verify that G and H are two distribution functions satisfying $ F = \theta G + (1-\theta) H $ . We also assert that $ H {\preceq_{cx}} F$ . To see it, let $X\sim F,$ $Y \sim G ,$ and $Z \sim H $ . Then, for any convex function $\phi$ , we have
It then follows that
Also, note that $F = \theta G + (1-\theta) H$ which implies $ {\mathbb{E}}[\phi(X)] = \theta {\mathbb{E}}[\phi(Y)] + (1- \theta ) {\mathbb{E}}[\phi(Z)]$ . Combined with $ {\mathbb{E}}[\phi(X)] \leq {\mathbb{E}}[\phi(Y)]$ , we have ${\mathbb{E}}[\phi(Z)] \leq {\mathbb{E}}[\phi(X)],$ that is, $ H {\preceq_{cx}} F$ . This implies $H{\preceq_{cx}} {{U}}[0,1]$ , and note that the support of distribution H has at most $m-1$ points. By induction hypothesis, we have $H \in \mathcal D_n^{\textrm{U}}(1/n)$ .
On the other hand, if $ a_m - a_1 \leq \frac{1}{2} $ , we know $ G \preceq_{cx} {{U}}[0,1] $ , $G \in \mathcal D_n^{\textrm{U}}(1/n,\ldots, 1/n)$ . Then $F = \theta G + (1-\theta) H \in \mathcal D_n^{\textrm{U}}(1/n)$ . Otherwise, if $ a_m - a_1 > \frac{1}{2} $ , then $ G \preceq_{cx} {{U}}[0,1]$ , that is, $ W_G(t_0) > 0 $ for some $t_0\in(0,1)$ . Define a function
which is a continuous function satisfying $\alpha (0+) = \alpha (1- ) = - 1 $ as $ W_F(t) = W_G(t) = - {t^2}/{2}$ for $t\in [0,a_1] $ and $W_F(t) = W_G(t) = -{(1-t)^2}/{2}$ for $t\in [a_m,1] $ . Hence, $\alpha(t) $ takes its maximum value at some $t\in (0,1)$ and the maximum value is positive. Without loss of generality, assume $\alpha_0 = \alpha (t_1) > 0 $ is its maximum value. Then we have
Define a distribution $F_0 \,{:\!=}\, (G + \alpha_0 F )/({1+ \alpha_0})$ . Then, by (5), we have $F_0 {\preceq_{cx}} {{U}}[0,1]$ and $W_{F_0} (t_1 ) = 0 $ . By case (a), we have $F_0 \in \mathcal D_n^{\textrm{U}}(1/n)$ .
Note that $ G = (1+\alpha_0) F_0 - \alpha_0 F $ and $F = \theta G + (1-\theta) H $ , which implies $ F = (1+\alpha_0) \theta F_0 - \alpha_0 \theta F + (1-\theta) H, $ that is,
Then, by Lemma 2, we have $F \in \mathcal D_n^{\textrm{U}}(1/n)$ .
If F is a general distribution function such that $F\preceq_{cx} {U}[0,1]$ , then ${Supp(F)\subset [0,1]}$ and it has no mass on 0 and 1. For any $n\in{\mathbb{N}}$ , define $F_n$ as the distribution function of $X_n$
where X is a random variable having distribution function F. Then $F_n$ converges to F in weak convergence as $n\to\infty$ , and $F_n\preceq_{cx} {U}[0,1]$ . By the above proof for discrete distributions with finite support, we have $F_n\in \mathcal D_n^{\textrm{U}}(1/n)$ for each $n\in{\mathbb{N}}$ . Then, by Lemma 2, we have $F\in \mathcal D_n^{\textrm{U}}(1/n)$ . Thus, we complete the proof.
For any random variable $X\sim F$ with mean 0, we have $F_a \preceq_{cx} F$ for any $a\in [0,1]$ , where $F_a$ is the distribution of aX. Hence, we immediately get the following corollary. Note that this corollary, although it looks simple, does not seem to allow for an elementary proof without using Theorem 5.
Corollary 1. For $n \geq 3$ , if $F\in\mathcal D_n ({{\textrm{U}}}[-1,1],\ldots,{{\textrm{U}}}[-1,1])$ , then so is $F_a$ for all $a\in [0,1]$ .
5. An application
In risk management, one often needs to optimize a statistical functional, mapping $\mathcal F$ to ${\mathbb{R}}$ (such as a risk measure), over the set of $\mathcal D_n(F_1,\ldots,F_n)$ , and this type of problem is called risk aggregation with dependence uncertainty (see e.g. [Reference Bernard, Jiang and Wang2] and [Reference Embrechts, Puccetti and Rüschendorf3]). These problems are typically quite difficult to solve in general, as the set $\mathcal D_n(F_1,\ldots,F_n)$ is a complicated object. For uniform marginal distributions, using results in this paper (in particular, Theorem 5), we are able to translate many optimization problems on $\mathcal D_n^{\textrm{U}}$ to $\mathcal C^{\textrm{U}}_n$ for $n\geq 3$ , which is a convenient object to work with.
We study an application of the problem of minimizing or maximizing, for a given interval A, the value of ${\mathbb{P}}(S\in A)$ , where S is the sum of n standard uniform random variables. A special case of the problem concerning bounds on ${\mathbb{P}}(S\leq x)$ for $x\in {\mathbb{R}}$ , i.e. bounds on F(x) for $F\in \mathcal D^{\textrm U} _n $ , is studied in [Reference Rüschendorf8]. Using Theorem 5, we are able to solve the problem of ${\mathbb{P}}(S\in A)$ completely.
Proposition 2. For $n\geq 3$ , and $0\leq a\leq a+b\leq n$ , we have
and
where $F_S$ stands for the cdf of S.
Proof. As Theorem 5 gives $ \mathcal D^{\textrm U} _n=\mathcal C^{\textrm{U}}_n$ , it suffices to look at the optimization problems for $ \mathcal C^{\textrm{U}}_n$ . For $0\leq u\leq v\leq n$ , let $\mathcal A_{u,v}$ be the sigma field generated by $\{U\leq u\}$ and $\{U\leq v\}$ . $S={\mathbb{E}}[U|\mathcal A_{u,v}]$ is tri-atomically distributed with distribution measure
Note that $F_S\preceq_{cx} F_U$ because S is a conditional expectation of U, and thus $F_S \in \mathcal C^{\textrm{U}}_n= \mathcal D^{\textrm U} _n$ .
We first verify the following inequality (indeed, it is [8, Theorem 1]). For any S which is the sum of n standard uniform random variables and $x\in {\mathbb{R}}$ , we have
Equation (9) can be shown using the equivalent condition of convex order (see [10, Theorem 3.A.5]). Using $F_S\preceq_{cx} F_U$ ,
Therefore, $F^{-1}_S(\alpha)\geq \frac{n\alpha}{2}$ , $\alpha \in (0,1)$ , and equivalently $F_S(x)\leq 2x/n$ , $x\in {\mathbb{R}}$ . The other inequality in (9) is symmetric by noting that $n-S$ is still the sum of n standard uniform random variables and ${\mathbb{P}}(S\geq x)={\mathbb{P}}(n-S\leq n-x)$ .
We now analyse the problem of the minimum in (6).
Suppose $b\leq n/2$ . Since $(n+u)/2-u/2 =n/2>b$ , we can find $u\in [0,n]$ such that $u/2\leq a$ and ${(n+u)/2\geq b+a}$ . By letting $S={\mathbb{E}}[U|\mathcal A_{u,u}]$ and using (8), we have ${\mathbb{P}}(S=u/2) = u/n$ and ${\mathbb{P}}(S=(n+u)/2)=(n-u)/n$ . In this case, ${\mathbb{P}}(S\in (a,a+b))=0$ ; thus (6) holds.
Suppose $b>n/2$ , which implies $a<n/2$ and $a+b>n/2$ . Let S be given by ${\mathbb{E}}[U|\mathcal A_{u,v}]$ where $u=2a$ and $v=2(a+b)-n$ . Note that ${\mathbb{P}}(S\in (a,a+b))= \frac{v-u}{n} = \frac{2b}n-1.$ This shows the ‘ $\leq $ ’ direction of (6). On the other hand, by (9), for any S which is the sum of n standard uniform random variables, ${\mathbb{P}}(S\leq a)\leq 2a/n$ and ${\mathbb{P}}(S\geq a+b)\leq 2(n-a-b)/n.$ Thus
This shows the ‘ $\geq$ ’ direction of (6).
Next, we analyse the problem of the maximum in (7).
If $a+b> n/2$ and $a< n/2$ , then $n/2\in [a,a+b]$ . Taking $S={\mathbb{E}}[U]=n/2$ gives ${\mathbb{P}}(S\in [a,a+b])=1$ ; thus (7) holds.
Suppose $a+b\leq n/2$ . By (9), for any S which is the sum of n standard uniform random variables, ${\mathbb{P}}(S\in [a,a+b])\leq {\mathbb{P}}(S\leq a+b) \leq 2(a+b)/n$ . To see that such a bound is attainable, take $S={\mathbb{E}}[U|\mathcal A_{u,u}]$ where $u=2(a+b)$ . Then, by (8), we have
Therefore, (7) holds.
Suppose $a\geq n/2$ . Similarly to the above case, by (9), for any S which is the sum of n standard uniform random variables, ${\mathbb{P}}(S\in [a,a+b])\leq {\mathbb{P}}(S\geq a) \leq 2(n-a)/n$ . To see that such a bound is attainable, take $S={\mathbb{E}}[U|\mathcal A_{u,u}]$ where $u=2a-n$ . Then, by (8), we have
Therefore, (7) holds.
Remark 2. Based on the proof of Proposition 5, we can identify some minimizing distributions for (6) and some maximizing distributions for (7). For $0\leq u\leq v\leq n$ , write the distribution
There are a few cases. For the minimum in (6):
If $b\leq n/2$ , then $F_{u,u}$ attains (6) for $u \in [2a+2b-2n, 2a]$ .
If $b> n/2$ , then $F_{u,v} $ attains (6) where $u=2a$ and $v=2(a+b)-n$ .
For the maximum in (7):
If $a+b> n/2$ and $a< n/2$ , then $\delta_{n/2}$ attains (7).
If $a+b\leq n/2$ , then $F_{u,u}$ attains (7) where $u=2(a+b)$ .
If $a\geq n/2$ , then $F_{u,u}$ attains (7) where $u=2a-n$ .
Acknowledgements
The authors thank the Editor, an Associate Editor, and an anonymous referee for careful reading of our paper and very helpful comments which improved the quality of the paper. T. Mao was supported by the NNSF of China (grants 71671176, 71871208, 11501575). B. Wang acknowledges financial support from the NNSF of China (grant 11501017). R. Wang acknowledges financial support from the Natural Sciences and Engineering Research Council of Canada (RGPIN-2018-03823, RGPAS-2018-522590), and the Center of Actuarial Excellence Research Grant from the Society of Actuaries.