• 沒有找到結果。

1 The Order Statistics

N/A
N/A
Protected

Academic year: 2022

Share "1 The Order Statistics"

Copied!
30
0
0

加載中.... (立即查看全文)

全文

(1)

Chapter 2. Order Statistics

1 The Order Statistics

For a sample of independent observations X1, X2, . . . , Xn on a distribution F , the ordered sample values

X(1)≤ X(2)≤ · · · ≤ X(n), or, in more explicit notation,

X(1:n)≤ X(2:n)≤ · · · ≤ X(n:n),

are called the order statistics. If F is continuous, then with probability 1 the order statistics of the sample take distinct values (and conversely).

There is an alternative way to visualize order statistics that, although it does not necessarily yield simple expressions for the joint density, does allow simple derivation of many important properties of order statistics. It can be called the quantile function representation.

The quantile function (or inverse distribution function, if you wish) is defined by

F−1(y) = inf{x : F (x) ≥ y}. (1)

Now it is well known that if U is a Uniform(0,1) random variable, then F−1(U ) has distri- bution function F . Moreover, if we envision U1, . . . , Un as being iid Uniform(0,1) random variables and X1, . . . , Xn as being iid random variables with common distribution F , then

(X(1), . . . , X(n))= (Fd −1(U(1)), . . . , F−1(U(n))), (2) where = is to be read as “has the same distribution as.”d

1.1 The Quantiles and Sample Quantiles

Let F be a distribution function (continuous from the right, as usual). The proof of F is right continuous can be obtained from the following fact:

F (x + hn) − F (x) = P (x < X ≤ x + hn),

where {hn} is a sequence of real numbers such that 0 < hn ↓ 0 as n → ∞. It follows from the continuity property of probabaility (P (limnAn) = limnP (An) if lim An exists.) that

n→∞lim[F (x + hn) − F (x)] = 0,

and hence that F is right-continuous. Let D be the set of all discontinuity points of F and n be a positive integer. Set

Dn=



x ∈ D : P (X = x) ≥ 1 n

 .

(2)

Since F (∞)−F (−∞) = 1, the number of elements in Dncannot exceed n. Clearly D = ∪nDn, and it follows that D is countable. Or, the set of discontinuity points of a distribution function F is countable. We then conclude that every distribution function F admits the decomposition

F (x) = αFd(x) + (1 − α)Fc(x), (0 ≤ α ≤ 1),

where Fd and Fc are both continuous function such that Fd is a step function and Fc is continuous. Moreover, the above decomposition is unique.

Let λ denote the Lebesgue measure on B, the σ-field of Borel sets in R. It follows from the Lebesgue decomposition theorem that we can write Fc(x) = βFs(x) + (1 − β)Fac(x) where 0 ≤ β ≤ 1, Fs is singular with respect to λ, and Facis absolutely continuous with respect to λ. On the other hand, the Radon-Nikodym theorem implies that there exists a nonnegative Borel-measurable function on R such that

Fac(x) = Z x

−∞

f dλ,

where f is called the Radon-Nikodym derivative. This says that every distribution function F admits a unique decomposition

F (x) = α1Fd(x) + α2Fs(x) + α3Fac(x), (x ∈ R), where αi≥ 0 and P3

i=1αi= 1.

For 0 < p < 1, the pth quantile or fractile of F is defined as ξ(p) = F−1(p) = inf{x : F (x) ≥ p}.

This definition is motivated by the following observation:

• If F is continuous and strictly increasing, F−1 is defined by F−1(y) = x when y = F (x).

• If F has a discontinuity at x0, suppose that F (x0−) < y < F (x0) = F (x0+). In this case, although there exists no x for which y = F (x), F−1(y) is defined to be equal to x0.

• Now consider the case that F is not strictly increasing. Suppose that

F (x)









< y for x < a

= y for a ≤ x ≤ b

> y for x > b

Then any value a ≤ x ≤ b could be chosen for x = F−1(y). The convention in this case is to define F−1(y) = a.

(3)

Now we prove that if U is uniformly distributed over the interval (0, 1), then X = FX−1(U ) has cumulative distribution function FX(x). The proof is straightforward:

P (X ≤ x) = P [FX−1(U ) ≤ x] = P [U ≤ FX(x)] = FX(x).

Note that discontinuities of F become converted into flat stretches of F−1 and flat stretches of F into discontinuities of F−1.

In particular, ξ1/2= F−1(1/2) is called the median of F . Note that ξp satisfies F (ξ(p)−) ≤ p ≤ F (ξ(p)).

The function F−1(t), 0 < t < 1, is called the inverse function of F . The following proposition, giving useful properties of F and F−1, is easily checked.

Lemma 1 Let F be a distribution function. The function F−1(t), 0 < t < 1, is nondecreasing and left-continuous, and satisfies

(i) F−1(F (x)) ≤ x, −∞ < x < ∞, (ii) F (F−1(t)) ≥ t, 0 < t < 1.

Hence

(iii) F (x) ≥ t if and only if x ≥ F−1(t).

Corresponding to a sample {X1, X2, . . . , Xn} of observations on F , the sample pth quantile is defined as the pth quantile of the sample distribution function Fn, that is, as Fn−1(p). Regarding the sample pth quantile as an estimator of ξp, we denote it by ˆξpn, or simply by ˆξpwhen convenient.

Since the order stastistics is equivalent to the sample distribution function Fn, its role is fundamental even if not always explicit. Thus, for example, the sample mean may be regarded as the mean of the order statistics, and the sample pth quantile may be expressed as

ξˆpn=

Xn,np if np is an integer Xn,[np]+1 if np is not an integer.

1.2 Functions of Order Statistics

Here we consider statistics which may be expressed as functions of order statistics. A variety of short-cut procedures for quick estimates of location or scale parameters, or for quick tests of related hypotheses, are provided in the form of linear functions of order statistics, that is statistics of the form

n

X

i=1

cniX(i:n).

(4)

We term such statistics “L-estimates.” For example, the sample range X(n:n)− X(1:n)belongs to this class. Another example is given by the α-trimmed mean.

1 n − 2[nα]

n−[nα]

X

i=[nα]+1

X(i:n),

which is a popular competitor of ¯X for robust estimation of location. The asymptotic dis- tribution theory of L-statistics takes quite different forms, depending on the character of the coefficients {cni}.

The representations of ¯X and ˆξpn in terms of order statistics are a bit artificial. On the other hand, for many useful statistics, the most natural and efficient representations are in terms of order statistics. Examples are the extreme values X1:n and Xn:n and the sample range Xn:n− X1:n.

1.3 General Properties

Theorem 1 (1) P (X(k)≤ x) =Pn

i=kC(n, i)[F (x)]i[1 − F (x)]n−i for −∞ < x < ∞.

(2) The density of X(k) is given by nC(n − 1, k − 1)Fk−1(x)[1 − F (x)]n−kf (x).

(3) The joint density of X(k1) and X(k2) is given by n!

(k1− 1)!(k2− k1− 1)!(n − k2)![F (x(k1))]k1−1[F (x(k2)) − F (x(k1))]k2−k1−1 [1 − F (x(k2))]n−k2f (x(k1))f (x(k2))

for k1< k2 and x(k1)< x(k2).

(4) The joint pdf of all the order statistics is n!f (z1)f (z2) · · · f (zn) for −∞ < z1 < · · · <

zn< ∞.

(5) Define V = F (X). Then V is uniformly distributed over (0, 1).

Proof. (1) The event {X(k)≤ x} occurs if and only if at least k out of X1, X2, . . . , Xn are less than or equal to x.

(2) The density of X(k) is given by nC(n − 1, k − 1)Fk−1(x)[1 − F (x)]n−kf (x). It can be shown by the fact that

d dp

n

X

i=k

C(n, i)pi(1 − p)n−i= nC(n − 1, k − 1)pk−1(1 − p)n−k.

Heuristically, k − 1 smallest observations are ≤ x and n − k largest are > x. X(k) falls into a small interval of length dx about x is f (x)dx.

(5)

1.4 Conditional Distribution of Order Statistics

In the following two theorems, we relate the conditional distribution of order statistics (con- ditioned on another order statistic) to the distribution of order statistics from a population whose distribution is a truncated form of the original population distribution function F (x).

Theorem 2 Let X1, X2, . . . , Xn be a random sample from an absolutely continuous popula- tion with cdf F (x) and density function f (x), and let X(1:n)≤ X(2:n)≤ · · · ≤ X(n:n) denote the order statistics obtained from this sample. Then the conditional distribution of X(j:n), given that X(i:n)= xi for i < j, is the same as the distribution of the (j − i)th order statistic obtained from a sample of size n − i from a population whose distribution is simply F (x) truncated on the left at xi.

Proof. From the marginal density function of X(i:n) and the joint density function of X(i:n) and X(j:n), we have the conditional density function of X(j:n), given that X(i:n) = xi, as

fX(j:n)(xj|X(i:n)= xi) = fX(i:n),X(j:n)(xi, xj)/fX(i:n)(xi)

= (n − i)!

(j − i − 1)!(n − j)!

 F (xj) − F (xi) 1 − F (xi)

j−i−1

× 1 − F (xj) 1 − F (xi)

n−j

f (xj) 1 − F (xi).

Here i < j ≤ n and xi ≤ xj < ∞. The result follows easily by realizing that {F (xj) − F (xi)}/{1 − F (xi)} and f (xj)/{1 − F (xi)} are the cdf and density function of the population whose distribution is obtained by truncating the distribution F (x) on the left at xi.

Theorem 3 Let X1, X2, . . . , Xn be a random sample from an absolutely continuous popula- tion with cdf F (x) and density function f (x), and let X(1:n)≤ X(2:n)≤ · · · ≤ X(n:n) denote the order statistics obtained from this sample. Then the conditional distribution of X(i:n), given that X(j:n) = xj for j > i, is the same as the distribution of the ith order statistic in a sample of size j − 1 from a population whose distribution is simply F (x) truncated on the right at xj.

Proof. From the marginal density function of X(i:n) and the joint density function of X(i:n) and X(j:n), we have the conditional density function of X(i:n), given that X(j:n)= xj, as

fX(i:n)(xi|X(j:n)= xj) = fX(i:n),X(j:n)(xi, xj)/fX(j:n)(xj)

= (j − 1)!

(i − 1)!(j − i − 1)!

 F (xi) F (xj)

i−1

× F (xj) − F (xi) F (xj)

j−i−1

f (xi) F (xj).

(6)

Here 1 ≤ i < j and −∞ < xi ≤ xj. The proof is completed by noting that F (xi)/F (xj) and f (xi)/F (xj) are the cdf and density function of the population whose distribution is obtained by truncating the distribution F (x) on the right at xj

1.5 Computer Simulation of Order Statistics

In this section, we will discuss some methods of simulating order statistics from a distribution F (x). First of all, it should be mentioned that a straightforward way of simulating order statistics is to generate a pseudorandom sample from the distribution F (x) and then sort the sample through an efficient algorithm like quick-sort. This general method (being time- consuming and expensive) may be avoided in many instances by making use of some of the distributional properties to be established now.

For example, if we wish to generate the complete sample (x(1), . . . , x(n)) or even a Type II censored sample (x(1), . . . , x(r)) from the standard exponential distribution. This may be done simply by generating a pseudorandom sample y1, . . . , yrfrom the standard exponential distribution first, and then setting

x(i)=

i

X

j=1

yj/(n − j + 1), i = 1, 2, . . . , r.

The reason is as follows:

Theorem 4 Let X(1)≤ X(2) ≤ · · · ≤ X(n)be the order statistics from the standard exponen- tial distribution. Then, the random variables Z1, Z2, . . . , Zn, where

Zi= (n − i + 1)(X(i)− X(i−1)),

with X(0)≡ 0, are statistically independent and also have standard exponential distributions.

Proof. Note that the joint density function of X(1), X(2), . . . , X(n)is

f1,2,...,n:n(x1, x2, . . . , xn) = n! exp −

n

X

i=1

xi

!

, 0 ≤ x1< x2< · · · < xn< ∞.

Now let us consider the transformation

Z1= nX(1), Z2= (n − 1)(X(2)− X(1)), . . . , Zn= X(n)− X(n−1), or the equivalent transformation

X(1)= Z1/n, X(2)= Z1 n + Z2

n − 1, . . . , X(n)= Z1 n + Z2

n − 1+ · · · + Zn. After noting the Jacobian of this transformation is 1/n! and that Pn

i=1xi = Pn

i=1zi, we immediately obtain the joint density function of Z1, Z2, . . . , Zn to be

fZ1,Z2,...,Zn(z1, z2, . . . , zn) = exp −

n

X

i=1

zi

!

, 0 ≤ z1, . . . , zn< ∞.

(7)

If we wish to generate order statistics from the Uniform(0, 1) distribution, we may use the following two Theorems and avoid sorting once again. For example, if we only need the ith order statistic u(i), it may simply be generated as a pseudorandom observation from Beta(i, n − i + 1) distribution.

Theorem 5 For the Uniform(0, 1) distribution, the random variables V1 = U(i)/U(j) and V2= U(j), 1 ≤ i < j ≤ n, are statistically independent, with V1 and V2 having Beta(i, j − i) and Beta(j, n − j + 1) distributions, respectively.

Proof. From Theorem 1(3), we have the joint density function of U(i)and U(j)(1 ≤ i < j ≤ n) to be

fi,j:n(ui, uj) = n!

(i − 1)!(j − i − 1)!(n − j)!ui−1i (uj− ui)j−i−1(1 − uj)n−j, 0 < ui < uj< 1.

Now upon makin the transformation V1 = U(i)/U(j) and V2 = U(j) and noting that the Jacobian of this transformation is v2, we derive the joint density function of V1and V2to be

fV1,V2(v1, v2) = (j − 1)!

(i − 1)!(j − i − 1)!v1i−1(1 − v1)j−i−1

= n!

(j − 1)!(n − j)!v2j−1(1 − v2)n−j,

0 < v1 < 1, 0 < v2 < 1. From the above equation it is clear that the random variables V1

and V2 are statistically independent, and also that they are distributed as Beta(i, j − i) and Beta(j, n − j + 1), respectively.

Theorem 6 For the Uniform(0, 1) distribution, the random variables

V1=U(1)

U(2), V2= U(2) U(3)

2

, · · · , V(n−1) = U(n−1) U(n)

n−1

and Vn= U(n)n are all independent Uniform(0, 1) random variables.

Proof. Let X(1) < X(2) < · · · < X(n) denote the order statistics from the standard expo- nential distribution. Then upon making use of the facts that X = − log U has a standard exponential distribution and that − log u is a monotonically decreasing function in u, we immediately have X(i)

= − log Ud (n−i+1). The above equation yields

Vi=

 U(i) U(i+1)

I

=d  e−X(n−i+1) e−X(n−i)



= exp[−i(X(n−i+1)− X(n−i))]= exp(−Yd n−i+1) upon using the above theorem, where Yi,are independent standard exponential random vari- ables.

The just-described methods of simulating uniform order statistics may also be used easily to generate order statistics from any known distribution F (x) for which F−1(·) is relatively easy to compute. We may simply obtain the order statistics x(1), . . . , x(n)from the required distribution F (·) by setting x(i)= F−1(u(i)).

(8)

2 Large Sample Properties of Sample Quantile

2.1 An Elementary Proof

Consider the sample pth quantile, ˆξpn, which is X([np]) or X([np]+1) depending on whether np is an integer (here [np] denotes the integer part of np). For simplicity, we discuss the properties of X([np]) where p ∈ (0, 1) and n is large. This will in turn inform us of the properties of ˆξpn.

We first consider the case that X is uniformly distributed over [0, 1]. Let U([np])denote the sample pth quantile. If i = [np], we have

nC(n − 1, i − 1) = n!

(i − 1)!(n − i)! = Γ(n + 1)

Γ(i)Γ(n − i + 1) = B(i, n − i + 1).

Elementary computations beginning with Theorem 1(2) yields U([np])∼ Beta(i0, n + 1 − i0) where i0= [np]. Then U(i0)∼ Beta(i0, n − i0+ 1). Note that

EU([np]) = [np]

n + 1 → p nCov U(np1), U(np2)

 = n[np1](n + 1 − [np2])

(n + 1)2(n + 2) → p1(1 − p2).

Use these facts and Chebyschev inequality, we can show easily that U([np]) → p with rateP n−1/2. This generates the question whether we can claim that

ξˆpn

→ ξP p.

Recall that U = F (X). If F is absolutely continuous with finite positive density f at ξp, it is expected that the above claim holds.

Recall that U([np])→ p with rate nP −1/2. The next question would be what the distri- bution of √

n(U([np])− p) is? Note that U([np])is a Beta([np], n − [np] + 1) random variable.

Thus, it can be expressed as

U[np] =

Pi0

i=1Vi

Pi0

i=1Vi+Pn+1 i=i0+1Vi

,

where the Vi’s are iid Exp(1) random variables. Observe that

√n

P[np]

i=1Vi

P[np]

i=1Vi+Pn+1 i=[np]+1Vi

− p

!

=

1 n

n

(1 − p) P[np]

i=1Vi− [np]

− p Pn+1

[np]+1Vi− (n − [np] + 1)

+ [(1 − p)[np] − p(n − [np] + 1)]o Pn+1

i=1 Vi/n and (√

n)−1{(1 − p)[np] − p(n − [np] + 1)} → 0. Since E(Vi) = 1 and V ar(Vi) = 1, from the central limit theorem it follows that

Pi0

i=1Vi− i0

√i0

→ N (0, 1)d

(9)

and

Pn+1

i=i0+1Vi− (n − i0+ 1)

√n − i0+ 1

→ N (0, 1) .d

Consequently,

Pi0

i=1Vi− i0

√n

→ N (0, p)d

and

Pn+1

i=i0+1Vi− (n − i0+ 1)

√n

→ N (0, 1 − p) .d

SincePi0

i=1Vi andPn

i=i0+1Vi are independent for all n, (1 − p)

P[np]

i=1Vi− [np]

√n − p

Pn+1

[np]+1Vi− (n − [np] + 1)

√n

→ N (0, p(1 − p)).d

Now, using the weak law of large numbers, we have 1

n + 1

n+1

X

i=1

Vi

→ 1.P

Hence, by Slutsky’s Theorem,

√n U([np])− p d

→ N (0, p(1 − p)) .

For an arbitrary cdf F , we have X([np])= F−1(U([np])). Upon expanding F−1(U([np])) in a Taylor-series around the point E(U([np])) = [np]/(n + 1), we get

X([np])= Fd −1(p) + (U([np])− p){f (F−1(Dn))}−1

where the random variable Dn is betwen U([np])and p. This can be rearranged as

√nX([np])− F−1(p) d

=√

n U([np])− p {f (F−1(Dn)}−1.

When f is continuous at F−1(p), it follows that as n → ∞, f (F−1(Dn))→ f (FP −1(p)). Use the delta method, it yields

√n X([np])− F−1(p) d

→ N



0, p(1 − p) [f (F−1(p))]2

 .

2.2 A Probability Inequality for | ˆ ξ

p

− ξ

p

|

We shall use the following result of Hoeffding (1963) to show that P (| ˆξpn− ξp| > ) → 0 exponentially fast.

Lemma 2 Let Y1, . . . , Yn be independent random variables satisfying P (a ≤ Yi ≤ b) = 1, each i, where a < b. Then, for t > 0,

P

n

X

i=1

Yi

n

X

i=1

E(Yi) ≥ nt

!

≤ exp−2nt2/(b − a)2 .

(10)

Remark. Suppose that Y1, Y2, . . . , Yn are independent and identically distributed random variables. Use Bery-Esseen theorem, we have

P

n

X

i=1

Yi

n

X

i=1

E(Yi) ≥ nt

!

= Φ(tp

V ar(Y1)/√

n) + Error.

Here

|Error| ≤ c

√n

E|Y1− EY1|3 [V ar(Y1)]3/2 .

Theorem 7 Let 0 < p < 1. Suppose that ξp is the unique solution x of F (x−) ≤ p ≤ F (x).

Then, for every  > 0, P

| ˆξpn− ξp| > 

≤ 2 exp −2nδ2 , all n,

where δ= min{F (ξp+ ) − p, p − F (ξp− )}.

Proof. Let  > 0. Write P

| ˆξpn− ξp| > 

= P ˆξpn> ξp+ 

+ P ˆξpn< ξp−  . By Lemma ??,

P ˆξpn> ξp+ 

= P (p > Fnp+ ))

= P

n

X

i=1

I(Xi> ξp+ ) > n(1 − p)

!

= P

n

X

i=1

Vi

n

X

i=1

E(Vi) > nδ1

! ,

where Vi = I(Xi> ξp+ ) and δ1= F (ξp+ ) − p. Likewise, P ˆξpn< ξp− 

= P (p > Fnp− ))

= P

n

X

i=1

Wi

n

X

i=1

E(Wi) > nδ2

! ,

where Wi = I(Xi< ξp− ) and δ2 = p − F (ξp− ). Therefore, utilizing Hoeffding’s lemma, we have

P ˆξpn> ξp+ 

≤ exp −2nδ12 and

P ˆξpn< ξp− 

≤ exp −2nδ22 . Putting δ= min{δ1, δ2}, the proof is completed.

(11)

2.3 Asymptotic Normality

Theorem 8 Let 0 < p < 1. Suppose that F possesses a density f in a neighborhood of ξp

and f is positive and continuous at ξp, then

√n( ˆξpn− ξp)→ Nd



0,p(1 − p) [f (ξp)]2

 .

Proof. We only consider p = 1/2. Note that ξp is the unique median since f (ξp) > 0.

First, we consider that n is odd (i.e., n = 2m − 1).

P

√ n



X(m)− F−1 1 2



≤ t



= P

√

nX(m)≤ t

F−1 1 2



= 0



= P



X(m)≤ t/√ n

F−1 1 2



= 0

 . Let Sn be the number of X’s that exceed t/√

n. Then X(m)≤ t

√n if and only if Sn≤ m − 1 = n − 1 2 .

Or, Sn is a Bin(n, 1 − F (F−1(1/2) + tn−1/2)) random variable. Set F−1(1/2) = 0 and pn= 1 − F (n1/2t). Note that

P

√ n



X(m)− F−1 1 2



≤ t



= P



Sn ≤n − 1 2



= P Sn− npn

pnpn(1 − pn) ≤

1

2(n − 1) − npn

pnpn(1 − pn)

! .

Recall that pn depends on n. Write Sn− npn

pnpn(1 − pn) =

n

X

i=1

Yi− pn

pnpn(1 − pn) =

n

X

i=1

Yin.

Again, they can be expressed as a double array with Yi∼ Bin(1, pn) Now utilize the Berry-Esseen Theorem to have

P



Sn≤ n − 1 2



− Φ

"1

2(n − 1) − npn pnpn(1 − pn)

#

→ 0

as n → ∞. Writing

1

2(n − 1) − npn

pnpn(1 − pn) ≈

√n 12− pn

 1/2

=

√n −12+ F (n−1/2t)

1/2 = 2tF (n−1/2t) − F (0)

n−1/2t → 2tf (0).

Thus,

Φ

"1

2(n − 1) − npn

pnpn(1 − pn)

#

≈ Φ (2f (0) · t) or

√n X(m)− F−1(1/2) d

→ N



0, 1

4f2(F−1(1/2))

 .

(12)

When n is even (i.e., n = 2m), both P [√

n(X(m)− F−1(1/2)) ≤ t] and P [√

n(X(m+1)− F−1(1/2)) ≤ t] tend to Φ 2f (F−1(1/2)) · t.

We just prove asymptotic normality of ˆξp in the case that F possesses derivative at the point ξp. So far we have discussed in detail the asymptotic normality of a single quantile.

This discussion extends in a natural manner to the asymptotic joint normality of a fixed number of quantiles. This is made precise in the following result.

Theorem 9 Let 0 < p1 < · · · < pk < 1. Suppose that F has a density f in neighborhoods of ξp1, . . . , ξpk and that f is positive and continuous at ξp1, . . . , ξpk. Then ( ˆξp1, . . . , ˆξpk) is asymptotically normal with mean vector (ξp1, . . . , ξpk) and covariance σij/n, where

σij = pi(1 − pj)

f (ξpi)f (ξpj) for 1 ≤ i ≤ j ≤ k and σij = σji for i > j.

Suppose that we have a sample of size n from a normal distribution N (µ, σ2). Let mn

represent the median of this sample. Then because f (µ) = (√

2πσ)−1,

√n(mn− µ)→ N (0, (1/4)/fd 2(µ)) = N (0, πσ2/2).

Compare mnwith ¯Xn as an estimator of µ. We conclude immediately that ¯Xnis better than mn since the latter has a much larger variance. Now consider the above problem again with Cauchy distribution C(µ, σ) with density function

f (x) = 1 πσ

1

1 + [(x − µ)/σ]2. What is your conclusion? (Exercise)

2.4 A Measure of Dispersion Based on Quantiles

The joint normality of a fixed number of central order statistics can be used to construct simultaneous confidence regions for two or more population quantiles. As an illustration, we now consider the semi-interquantile range, R = 123/4− ξ1/4). (Note that the parameter σ in C(µ, σ) is the semi-interquantile range.) It is used as an alternative to σ to measure the spread of the data. A natural estimate of R is ˆRn= 12( ˆξ3/4− ˆξ1/4). Theorem 4 gives the joint distribution of ˆξ1/4and ˆξ3/4. We can use the following result, due to Cramer and Wold (1936), which reduces the convergence of multivariate distribution functions to the convergence of univariate distribution functions.

Theorem 10 In Rk, the random vectors Xn converge in distribution to the random vector X if and only if each linear combination of the components of Xn converges in distribution to the same linear combination of the components of X.

(13)

By Theorem 9 and the Cramer-Wold device, we have R ∼ ANˆ

 R, 1

64n

 3

[f (ξ1/4)]2 − 2

f (ξ1/4)f (ξ3/4)+ 3 [f (ξ3/4)]2



.

For F = N (µ, σ2), we have

R ∼ ANˆ



0.6745σ,(0.7867)2σ2 n

 .

2.5 Confidence Intervals for Population Quantiles

Assume that F posseses a density f in a neighborhood of ξpand f is positive and continuous at ξp. For simplicity, consider p = 1/2. Then

√n ˆξ1/2− ξ1/2 d

→ N



0, 1

4f21/2)

 .

Therefore, we can derive a confidence interval of ξ1/2 if either f (ξ1/2) is known or a good estimator of f (ξ1/2) is available. A natural question to ask then is how do we estimate f (ξ1/2).

Here we propose two estimates. The first estimate is

number of observatiobns falling in ξ1/2− hn, ξ1/2+ hn 2hn

which is motivated by

F ξ1/2+ hn − F ξ1/2− hn

 2hn

≈ f (ξ1/2).

The second one is

X[n/2+`]− X[n/2−`]

2`/n where ` = O(nd) for 0 < d < 1.

2.6 Distribution-Free Confidence Interval

When F is absolutely continuous, F (F−1(p)) = p and, hence, we have P (X(i:n)≤ F−1(p)) = P (F (X(i:n)) ≤ p) = P (U(i:n) ≤ p)

=

n

X

r=i

C(n, r)pr(1 − p)n−r. (3)

Now for i < j, consider

P (X(i:n)≤ F−1(p)) = P (X(i:n)≤ F−1(p), X(j:n)< F−1(p)) +P (X(i:n)≤ F−1(p), X(j:n)≥ F−1(p))

= P (X(j:n)< F−1(p)) + P (X(i:n)≤ F−1(p) ≤ X(j:n)).

Since X(j:n)is absolutely continuous, this equation can be written as P X(i:n) ≤ F−1(p) ≤ X(j:n)

= P X(i:n)≤ F−1(p) − P X(j:n)≤ F−1(p)

=

j−1

X

r=i

C(n, r)pr(1 − p)n−r, (4)

(14)

where the last equality follows from (??). Thus, we have a confidence interval [X(i:n), X(j:n)] for F−1(p) whose confidence coefficient α(i, j) given by (??), is free of F and can be read from the table of binomial probabilities. If p and the desired confidence level α0 are specified, we choose i and j so that α(i, j) exceeds α0. Because of the fact that α(i, j) is a step function, usually the interval we obtain tends to be conservative. Further, the choice of i and j is not unique, and the choice which makes (j − i) small appear reasonable. For a given n and p, the binomial pmf C(n, r)pr(1 − p)n−r increases as r increases up to around [np], and then decreases. So if we want to make (j − i) small, we have to start with i and j close to [np]

and gradually increase (j − i) until α(i, j) exceeds α0.

2.7 Q-Q Plot

Wilk and Gnanadesikan (1968) proposed a graphical, rather informal, method of testing the goodness-of-fit of a hypothesized distribution to given data. It essentially plots the quantile function of one cdf against that of another cdf. When the latter cdf is the empirical cdf defined below, order statistics come into the picture. The empirical cdf, to be denoted by Fn(x) for all real x, represents the proportion of sample values that do not exceed x. It has jumps of magnitude 1/n at X(i:n), 1 ≤ i ≤ n. Thus, the order statistics represent the values taken by Fn−1(p), the sample quantile function.

The Q-Q plot is the graphical representation of the points (F−1(p), X(i:n)), where population quantiles are recorded along the horizontal axis and the sample quantiles on the vertical axis. If the sample is in fact from F , we expect the Q-Q plot to be close to a straight line. If not, the plot may show nonlinearity at the upper or lower ends, which may be an indication of the presence of outliers. If the nonlinearity shows up at other points as well, one could question the validity of the assumption that the parent cdf is F .

3 Empirical Distribution Function

Consider an i.i.d. sequence {Xi} with distribution function F . For each sample of size n, {X1, . . . , Xn}, a corresponding sample (empirical) distribution function Fn is constructed by placing at each observation Xi a mass 1/n. Thus Fn may be represented as

Fn(x) = 1 n

n

X

i=1

I(Xi≤ x), − ∞ < x < ∞.

(The definition of F defined on Rk is completely analogous.)

For each fixed sample, {X1, . . . , Xn}, Fn(·) is a distribution function, considered as a function of x. On the other hand, for each fixed value of x, Fn(x) is a random variable, considered as a function of the sample. In a view encompassing both features, Fn(·) is a

(15)

random distribution function and thus may be treated as a particular stochastic process (a random element of a suitable space.)

Note that the exact distribution of nFn(x) is simply binomial(n, F (x)). It follows immediately that

Theorem 11 1. E[Fn(x)] = F (x).

2. V ar{Fn(x)} = F (x)[1−F (x)]

n → 0, as n → ∞.

3. For each fixed x, −∞ < x < ∞, Fn(x) is AN



F (x),F (x)[1 − F (x)]

n

 .

The sample distribution function is utilized in statistical inference in several ways.

Firstly, its most direct application is for estimation of the population distribution function F . Besides pointwise estimation of F (x), each x, it is also of interest to characterize globally the estimation of F by Fn. For each fixed x, the strong law of large numbers implies that

Fn(x)a.s.→ F (x).

We now describe the Glivenko-Cantelli Theorem which ensures that the ecdf converges uni- formly almost surely to the true distribution function.

Theorem 12

P {sup

x

|Fn(x) − F (x)| → 0} = 1.

Proof. Let  > 0. Find an integer k > 1/ and numbers

−∞ = x0< x1≤ x2≤ · · · ≤ xk−1< xk= ∞,

such that F (xj) ≤ j/k ≤ F (xj) for j = 1, . . . , k − 1. Note that if xj−1< xj, then F (xj ) − F (xj−1) ≤ . From the strong law of large numbers,

Fn(xj)a.s.→ F (xj) and Fn(xj)a.s.→ F (xj) for j = 1, . . . , k − 1. Hence,

4n= max(|Fn(xj) − F (xj)|, |Fn(xj) − F (xj)|, j = 1, . . . , k − 1)a.s.→ 0.

Let x be arbitrary and find j such that xj−1< x ≤ xj. Then,

Fn(x) − F (x) ≤ Fn(xj) − F (xj−1) ≤ Fn(xj) − F (xj) + , and

Fn(x) − F (x) ≥ Fn(xj−1) − F (xj) ≥ Fn(xj−1) − F (xj−1) − .

This implies that

sup

x

|Fn(x) − F (x)| ≤ 4n+ a.s.→ .

Since this holds for all  > 0, the theorem follows.

(16)

3.1 Kolmogorov-Smirnov Test

To this effect, a very useful measure of closeness of Fn to F is the Kolmogorov-Smirnov distance

Dn = sup

−∞<x<∞

|Fn(x) − F0(x)|.

A related problem is to express confidence bands for F (x), −∞ < x < ∞. Thus, for selected functions a(x) and b(x), it is of interest to compute probabilities of the form

P (Fn(x) − a(x) ≤ F (x) ≤ Fn(x) + b(x), −∞ < x < ∞).

The general problem is quite difficult. However, in the simplest case, namely a(x) = b(x) = d, the problem reduces to computation of P (Dn< d).

For the case of F 1-dimensional, an exponential-type probability inequality for Dn was established by Dvoretzky, Kiefer, and Wolfowitz (1956).

Theorem 13 The distribution of Dn under H0 is the same for all continuous distribution.

Proof. For simplicity we give the proof for F0 strictly increasing. Then F0has inverse F0−1 and as u ranges over (0, 1), F0−1(u) ranges over a;; the possible values of X. Thus

Dn = sup

0<u<1

|Fn(F0−1(u))F0(F0−1(u))| = sup

0<u<1

|Fn(F0−1(u)) − u|.

Next note that

Fn(F0−1(u)) = [number of Xi≤ F0−1(u)]/n = [number of F0(Xi) ≤ u]/n.

Let Ui = F0(Xi)|. Then U1, . . . , Un are a sample from U N IF (0, 1), since P [F0(Xi) ≤ u] = P [Xi ≤ F0−1(u)] = F0(F0−1(u)) = u, 0 < u < 1.

Thus,

Dn= sup

0<u<1

|Fn(u) − u|

where Fn(u) is the empirical distribution of the uniform sample U1, . . . , Un and the distribu- tion of Dn does not depend on F0.

We now give an important fact which is used in Donsker (1952) to give a rigorous proof of the Kolmogorov-Smirnov Theorems.

Theorem 14 The distribution of the order statistic (Y(1), . . . , Y(n)) of n iid random variables Y1, Y2, . . . from the uniform distribution on [0, 1] can also be obtained as the distribution of the ratios

 S1 Sn+1

, S2 Sn+1

, · · · , Sn Sn+1

 ,

where Sk= T1+ · · · + Tk, k ≥ 1, and T1, T2, . . . is an iid sequence of (mean 1) exponentially distributed random variables.

(17)

Intuitively, if the Tiare regarded as the successive times between occurrence of some phenom- ena, then Sn+1 is the time to the (n + 1)st occurrence and, in units of Sn+1, the occurrence times should be randomly distributed because of lack of memory and independence properties.

Recall that Dn = sup0<u<1|Fn(u) − u| where Fn(u) is the empirical distribution of the uniform sample U1, . . . , Un. We then have

Dn = √

n sup

0<u<1

|Fn(u) − u| =√ n max

k≤n

Y(k)−k n

=d √ n max

k≤n

Sk

Sn+1

−k n

= n

Sn+1

max

k≤n

Sk− k

√n −k n

Sn+1− n

√n

Theorem 15 Let F be defined on R. There exists a finite positive constant C (not depending on F ) such that

P (Dn< d) ≤ C exp(−2nd2), d > 0, for all n = 1, 2, . . ..

Moreover,

Theorem 16 Let F be 1-dimensional and continuous.. Then

n→∞lim P (n1/2Dn < d) = 1 − 2

X

j=1

(−1)j+1exp(−2j2d2), d > 0,

for all n = 1, 2, . . ..

Secondly, we consider goodness of fit test statistics based on the sample distribution function.

The null hypotheseis in the simple case is H0 : F = F0, where F0 is specified. A useful procedure is the Kolmogorov-Smirnov test statistic

4n= sup

−∞<x<∞

|Fn(x) − F0(x)|,

which reduces to Dn under the null hypothesis. More broadly, a class of such statistics is obtained by introducing weight functions:

sup

−∞<x<∞

|w(x)[Fn(x) − F0(x)]|.

Another important class of statistics is based on the Cramer-von Mises test statistic Cn= n

Z

−∞

[Fn(x) − F0(x)]2dF0(x)

and takes the general form nR w(F0(x))[Fn(x) − F0(x)]2dF0(x). For example, for w(t) = [t(1−t)]−1, each duscrepancy Fn(x)−F (x) becomes weighted by the reciprocal of its standard deviation (under H0), yielding the Anderson-Darling statistic.

(18)

3.2 Stieltjes Integral

If [a, b] is a compact interval, a set of points {x0, x1, . . . , xn}, satisfying the inequalities a = x0< x1< · · · < xn= b,

is called a partition of [a, b]. Write 4fk = f (xk) − f (xk−1) for k = 1, 2 . . . , n. If there exists a positive number M such that

n

X

k=1

| 4 fk| ≤ M

for all partitions of [a, b], then f is said to be of bounded variation on [a, b]. Let F (x) be a function of bounded variation and continuous from the left such as a distribution function.

Given a finite interval (a, b) and a function f (x) we can form the sum Jn=

n

X

i=1

f (x0i)[F (xi) − F (xi−1)]

for a division of (a, b) by points xi such that a < x1 < · · · < xn < b and arbitrary x0i ∈ (xi−1, xi). It may be noted that in the Riemann integration a similar sum is considered with the length of the interval (xi− xi−1) instead of F (xi) − F (xi−1). If J = limn→∞Jn as the length of each interval → 0, then J is called the Stieltjes integral of f (x) with respect to F (x) and is denoted by

J = Z b

a

f (x)dF (x).

The improper integral is defined by lim

a→−∞,b→∞

Z b a

f (x)dF (x) = Z

f (x)dF (x).

One point of departure from Riemann integration is that it is necessary to specify whether the end points are included in the integration or not. From the definition it is easily shown that

Z b a

f (x)dF (x) = Z b

a+0

f (x)dF (x) + f (a)[F (a+0) − F (a)]

where a+0 indicates that the end point a is not included. If F (x) jumps at a, then Z b

a

f (x)dF (x) − Z b

a+0

f (x)dF (x) + f (a)[F (a+0) − F (a)],

so that the integral taken over an interval that reduces to zero need not be zero. We shall follow the convention that the lower end point is always included but not the upper end point.

With this convention, we see that Rb

adF (x) = F (b) − F (a). If there exists a function p(x) such that F (x) =Rx

−∞p(x)dx, the Stieltjes integral reduces to a Riemann integral Z

f (x)dF (x) = Z

f (x)p(x)dx.

Theorem 17 Let α be a step function defined on [a, b] with jumps αk at xk. ThenRb

a f (x)dα(x) = Pn

k=1f (xkk.

(19)

4 Sample Density Functions

Recall that F0 = f . This suggests that we can use the derivative of Fn to estimate f . In particular, we consider

fn(x) = Fn(x + bn) − Fn(x − bn) 2bn

. Observe that 2nbnfn(x) ∼ Bin(n, F (x + bn) − F (x − bn)) and we have

Efn(x) = 1 2nbn

n [F (x + bn) − F (x − bn)]

≈ 1

2nbn

n · 2bnf (x) if bn→ 0, var(fn(x)) = 1

(2nbn)2n [F (x + bn) − F (x − bn)] [1 − F (x + bn) + F (x − bn)]

≈ 1

4nb2n2bnf (x) = f (x)

2nbn if nbn → ∞.

A natural question to ask is to find an optimal choice of bn. To do so, we need to find the magnitude of bias, Efn(x) − f (x).

Since the above estimate can be viewed as the widely used kernel estimate with kernel W (z) = 1/2 if |z| ≤ 1 and = 0 otherwise, we will find the the magnitude of bias for the following kernel estimate of f (x) instead.

fn(x) = 1 nbn

n

X

i=1

W x − Xi

bn

 ,

where W is an integrable nonnegative weight function. Typically, W are chosen to be a density function, R tW (t)dt = 0, and R t2W (t)dt = α 6= 0. We have

Efn(x0) = 1 bn

EW x0− X bn



= 1

bn

Z

W x0− x bn



f (x)dx = Z

W (t)f (x0− bnt)dt

= Z

W (t)h

f (x0) − bntf0(x0− θtbnt)i dt

= f (x0) − bn Z

tW (t)f0(x0− θtbnt)dt.

WhenR tW (t)dt 6= 0, we have

Efn(x0) = f (x0) − bnf0(x0) Z

tW (t)dt + o(bn).

WhenR tW (t)dt = 0 and R t2W (t)dt 6= 0, we have Efn(x0) =

Z W (t)



f (x0) − bntf0(x0) +b2n

2 t2f(x0− θtbnt)

 dt

= f (x0) +b2n 2 f(x0)

Z

t2W (t)dt + o(b2n).

Therefore, bn = O(n−1/3) when R tW (t)dt 6= 0 (i.e., Assume that f0 exists.), and bn = O(n−1/5) whenR tW (t)dt = 0, R t2W (t)dt 6= 0 (i.e., Assume that fexists.)

(20)

5 Applications of Order Statistics

The following is a brief list of settings in which order stastistics might have a significant role.

1. Robust Location Estimates. Suppose that n independent measurements are available, and we wish to estimate their assumed common mean. It has long been recognized that the sample mean, though attractive from many viewpoints, suffers from an extreme sensitivity to outliers and model violations. Estimates based on the median or the average of central order statistics are less sensitive to model assumptions. A particularly well-known application of this observation is the accepted practice of using trimmed means (ignoring highest and lowest scores) in evaluating Olympic figure skating performances.

2. Detection of Outliers. If one is confronted with a set of measurements and is concerned with determining whether some have been incorrectly made or reported, attention naturally focuses on certain order statistics of the sample. Usually the largest one or two and/or the smallest one or two are deemed most likely to be outliers. Typically we ask questions like the following: If the observations really were iid, what is the probability that the latgest order statistic would be as large as the suspiciously large value we have observed?

3. Censored Sampling. Considser life-testing experiments, in which a fixed number n of items are placed on test and the experiment is terminated as soon as a prescribed number r have failed. The observed lifetimes are thus X1:n ≤ · · · ≤ Xr:n, whereas the lifetimes Xr+1:n≤ · · · ≤ Xn:n remain unobserved.

4. Waiting for the Big One. Disastrous floods and destructive earthquakes re- cur throughout history. Dam construction has long focused on so called 100-year flood. Presumably the dams are built big enough and strong enough to handle any water flow to be encountered except for a level expected to occur only once every 100 years. Architects in California are particularly concerned with construc- tion designed to withstand “the big one,” presumably an earthquake of enormous strength, perhaps a “100-year quake.” Whether one agrees or not with the 100- year diaster philosophy, it is obvious that designers of dams and skycrapers, and even doghouses, should be concerned with the distribution of large order statistics from a possibly dependent, possibly not identically distributed sequence.

After the disastrous flood of February 1st, 1953, in which the sea-dikes broke in several parts of the Netherlands and nearly two thousand people were killed, the Dutch government appointed a committee (so-called Delta-committee) to recom- mend on an appropriate level for the dikes (called Delta-level since) since no specific

(21)

statistical study had been done to fix a safer level for the sea-dikes before 1953.

The Dutch government set as the standard for the sea-dikes that at any time in a given year the sea level exceeds the level of the dikes with probability 1/10, 000. A statistical group from the Mathematical Centre in Amsterdam headed by D. van Dantzig showed that high tides occurring during certain dangerous windstorms (to ensure independence) within the dangerous wintermonths December, January and February (for homogenity) follow closely an exponential distribution if the smaller high tides are neglected.

If we model the annual maximum flood by a random variable Z, the Dutch gov- ernment wanted to determine therefore the (1 − q)-quantile

F−1(1 − q) = inf{t ∈ R : F (t) ≥ 1 − q}

of Z, where F denotes the distribution of Z and q has the value 10−4.

5. Strength of Materials. The adage that a chain is no longer than its weakest link underlies much of the theory of strength of materials, whether they are threads, sheets, or blocks. By considering failure potential in infinitesimally small sections of the material, on quickly is led to strength distributions associated with limits of distributions of sample minima. Of course, if we stick to the finite chain with n links, its strength would be the minimum of the strengths of its n component links, again an order statistic.

6. Reliability. The example of a cord composed of n threads can be extended to lead us to reliability applications of order statistics. It may be that failiure of one thread will cause the cord to break (the weakest link), but more likely the cord will function as long as k (a number less than n) of the threads remains unbroken.

as such it is an example of a k out of n system commonly discussed in reliability settings. With regard to tire failure, the automobile is often an example of a 4 out of 5 system (remember the spare). Borrowing on terminology from electrical systems, the n out of n system is also known as a series system. Any component failure is disastrous. The 1 out of n system is known as a parallel system; it will function as long as any of the components survives. The life of the k out of n system is clearly Xn−k+1:n, the (n − k + 1)st largest of the component lifetimes, or, equivalently, the time until less than k components are functioning. But, in fact, they can be regarded as perhaps complicated hierarchies of parallel and series subsystems, and the study of system lifetime will necessarily involve distributions of order statistics.

7. Quality Control. Take a comfortable chair and watch the daily production of Snickers candy bars pass by on the conveyor belt. Each candy bar should weigh

參考文獻

相關文件

On the other hand Chandra and Taniguchi (2001) constructed the optimal estimating function esti- mator (G estimator) for ARCH model based on Godambes optimal estimating function

Understanding and inferring information, ideas, feelings and opinions in a range of texts with some degree of complexity, using and integrating a small range of reading

Now, nearly all of the current flows through wire S since it has a much lower resistance than the light bulb. The light bulb does not glow because the current flowing through it

(a) The principal of a school shall nominate such number of teachers of the school for registration as teacher manager or alternate teacher manager of the school as may be provided

Students in this Learning Unit should recognise the concepts of sample statistics and population parameters and their relationships:. Population Parameter

Numerical results are reported for some convex second-order cone programs (SOCPs) by solving the unconstrained minimization reformulation of the KKT optimality conditions,

It is well known that second-order cone programming can be regarded as a special case of positive semidefinite programming by using the arrow matrix.. This paper further studies

Abstract We investigate some properties related to the generalized Newton method for the Fischer-Burmeister (FB) function over second-order cones, which allows us to reformulate