• Keine Ergebnisse gefunden

Supplements to the Asymptotic Theory of

2.2 Least Favorable Radius

Consequentially,

lim sup

n→∞

ωminc,n = lim sup

n→∞

trAn

En|AnΛn−an| ≤ωcmin (2.1.143)

∗=v, k= 1 : The minimum bias given in Theorem1.3.9(b) reads ωv,nmin= Enn)+−1

(2.1.144) where

lim sup

n→∞

Enn)+≤lim sup

n→∞

Enn| ≤lim sup

n→∞

pEnn)2 (2.1.145)

=√

I <∞ (2.1.146)

Since Lnn)−→w L(Λ) by assumption (2.1.103), an application of Vitali’s theorem yields

n→∞lim Enn)+= E(Λ)+ (2.1.147)

i.e., (2.1.106) holds. ////

Remark 2.1.12 (a) As shown at the beginning of the proof of Theorem 2.1.11, conditions (2.1.102) and (2.1.103) imply the convergence of the corresponding Fisher information

Iθn= EθnΛθnΛτθ

n→EθΛθΛτθ=Iθ (2.1.148) (b)In view of the approximation used in the proof of Lemma 6.4.4 of Rieder (1994) it might even be possible to abandon assumption (2.1.102).

(c)By the convergences stated in the previous theorem we at once get E|˜ηθn|2= trAθn−rn2b2θ

n −→trAθ−r2b2θ= E|˜ηθ|2 (2.1.149) as n→ ∞. That is, the trace of the asymptotic covariance is continuous in θ∈Θ . (d) In view of Ruckdeschel (2005b) a generalization of Proposition 2.1.11 to arbitrary D∈Rp×k with rkD=p seems to be in reach. ////

2.2 Least Favorable Radius

Given a neighborhood radius, we can now determine the optimally robust ICs via the implicit equations stated in Subsections 1.3.3and 1.3.4. But, in most applica-tions the neighborhood radius is unknown or unknown except to belong to some radius interval, respectively. That is, the radius appears as a one-dimensional nui-sance parameter of the infinitesimal neighborhood models introduced in Section1.2.

InRieder et al.(2001) we numerically solve the implicit equations for location, scale and linear regression models and calculate the increase of the maximum asymptotic risk over the minimax asymptotic risk in case that the optimally robust estimator for the false radius is used. More precisely, we determine theinefficiency- the limit

36 Supplements to the Asymptotic Theory of Robustness

of the ratio of sample sizes such as to achieve the same accuracy asymptotically;

i.e., in case of the MSE, we consider

relMSE(˜ηr0, r) =maxMSE(˜ηr0, r)

maxMSE(˜ηr, r) (2.2.1)

where

maxMSE(˜ηr0, r) = E|η˜r0|2+r2ω(˜ηr0)2 ∗=c, v (2.2.2) That is, the maximum asymptotic MSE of the AL estimator with IC ˜ηr0 which is optimal for an infinitesimal neighborhood of radius r0 ∈ [0,∞] , is evaluated over an infinitesimal neighborhood of another radius r ∈ [0,∞] , and is related to the minimax asymptotic MSE for that radius r. Often, we also use the term subefficiency which means inefficiency minus 1 .

The trace of the asymptotic covariance of the MSE solution ˜ηr is increasing and the standardized asymptotic bias is decreasing in the radius r∈(0,∞) . The proof of the following lemma is based on arguments provided by P. Ruckdeschel.

Lemma 2.2.1 Let η˜r be the solution for radius r ∈ (0,∞) provided by Theo-rem1.3.11. Then, E|˜ηr|2 is increasing and ω(˜ηr) is decreasing in r.

Proof Given some b ∈ (0,∞) with b ≥ ωmin , the solutions ˜ηb to the corre-sponding Hampel type problems (1.3.7) are given by Theorems 1.3.7 and 1.3.9.

Obviously, g(b) = ω(˜ηb)2 = b2 is non-negative, convex and strictly increasing in b; confer also Lemma 1.3.4. Moreover, given t∈[0,1] and b0, b1 ∈[ωmin,∞) , define bt= (1−t)b0+tb1. Then, ηt= (1−t)˜ηb0+tη˜b1 ∈ΨD2 with |ηt| ≤bt and

E|˜ηbt|2≤E|ηt|2≤(1−t) E|˜ηb0|2+tE|˜ηb1|2 (2.2.3) by the convexity of | · |2; i.e., f(b) = E|˜ηb|2 is convex in b. In addition, f(b) is non-negative and strictly decreasing in b. Now, consider the corresponding MSE problem which reads

h(r, b) =f(b) +r2g(b) (2.2.4) and let br be the optimal clipping bound for radius r; i.e.,

h(r, br) = min{h(r, b)|b∈[ωmin ,∞)} (2.2.5) By the convexity of f, g and h in b, one-sided derivatives in b∈(ωmin,∞) exist and we denote them by f0, g0 and ∂2h, respectively. Then, by convexity of h

2h(r, br−0)≤0≤∂2h(r, br+ 0) (2.2.6) for br∈(ωmin,∞) which implies

−f0(br−0)

g0(br−0) ≥r2≥ −f0(br+ 0)

g0(br+ 0) (2.2.7)

where the convexity of f and g entails that −f0 and 1/g0 are non-negative and decreasing. Now, assume r1< r2 and ωmin≤br1 < br2. Then,

r22≤ −f0(br2−0)

g0(br2−0) ≤ −f0(br1+ 0)

g0(br1+ 0) ≤r21 (2.2.8)

2.2 Least Favorable Radius 37

which is a contradiction. Hence, br = ω(˜ηr) is decreasing and consequentially f(br) = E|η˜r|2 is increasing in r∈(0,∞) . ////

Remark 2.2.2 As the notation in the proof suggests, the preceding lemma holds more generally, namely for all loss functions of the form

h(α, x) =f(x) +αg(x) (2.2.9)

where α∈(0,∞) and f and g are non-negative, convex and decreasing, respec-tively increasing in x∈R. Hence, it for instance applies to the setup ofRuckdeschel and Rieder(2004); confer also Remark1.3.3. ////

The following lemma shows that the MSE–inefficiency curves attain two relative maxima at the boundaries if we fix some interval for the neighborhood radius r. Lemma 2.2.3 Let r∈[rl, ru] with 0< rl< ru<∞.

(a)Then,

sup

r∈[rl,ru]

relMSE(˜ηs, r)≤maxnE|˜ηs|2

E|˜ηrl|2, ω(˜ηs)2 ω(˜ηru)2

o ∀s∈[rl, ru] (2.2.10)

and there exists some r0∈[rl, ru] such that E|˜ηr0|2

E|˜ηrl|2 = ω(˜ηr0)2

ω(˜ηru)2 (2.2.11) Moreover, in case rl= 0 and ru=∞,

inf

s∈[0,∞] sup

r∈[0,∞]

relMSE(˜ηs, r) = E|˜ηr0|2

trDI−1Dτ = ω(˜ηr0)2

min )2 (2.2.12) (b)It holds,

sup

r∈[rl,ru]

relMSE(˜ηs, r) = relMSE(˜ηs, rl)∨relMSE(˜ηs, ru) (2.2.13) for all s∈[rl, ru]. Moreover, there exists some r0∈[rl, ru] such that

relMSE(˜ηr0, rl) = relMSE(˜ηr0, ru) (2.2.14) and

inf

s∈[rl,ru] sup

r∈[rl,ru]

relMSE(˜ηs, r) = relMSE(˜ηr0, rl) = relMSE(˜ηr0, ru) (2.2.15) Proof

(a)Let s, r∈[rl, ru] with 0< rl< ru<∞. If EEηηs|2

r|2ωω( ˜ηs)2

( ˜ηr)2, E|˜ηs|2+r2ω(˜ηs)2≤ ω(˜ηs)2

ω(˜ηr)2 E|˜ηr|2+r2ω(˜ηr)2

(2.2.16)

38 Supplements to the Asymptotic Theory of Robustness andRuckdeschel(2005b). Hence, the intermediate value theorem yields some r0∈ [rl, ru] such that equality holds in (2.2.11). In case rl= 0 and ru=∞, we obtain

2.2 Least Favorable Radius 39

This leads us to (r22−r12)b2s

E|η˜s|2+r12b2s−E|˜ηr2|2+r22b2r2−E|˜ηr1|2−r21b2r1 E|˜ηr1|2+r21b2r1

E|˜ηs|2+r21b2s

E|η˜r2|2+r22b2r2 ≥0 (2.2.27) which can be simplified to

E|˜ηs|2+r22b2s

E|η˜r2|2+r22b2r2 − E|˜ηs|2+r21b2s

E|˜ηr1|2+r21b2r1 ≥0 (2.2.28) i.e., relMSE(˜ηs, r2)≥relMSE(˜ηs, r1) . Analogously, we may verify relMSE(˜ηs, r) is decreasing for r < s. Hence, (2.2.13) holds. Moreover, we have

E|˜ηr2|2+s2b2r

2−E|˜ηr1|2−s2b2r

1

= (s2−r21)(b2r

2−b2r

1) +r21(b2r

2−b2r

1) + E|˜ηr2|2−E|˜ηr1|2 (2.2.29)

= (s2−r21)(b2r

2−b2r

1) + E|η˜r2|2+r21b2r

2−E|˜ηr1|2−r12b2r

1 ≥0 (2.2.30) That is,

relMSE(˜ηr1, s)≤relMSE(˜ηr2, s) (2.2.31) for s < r1< r2. Analogously, we may show

relMSE(˜ηr1, s)≤relMSE(˜ηr2, s) (2.2.32) for s > r1 > r2. Thus, using the fact relMSE(˜ηrl, rl) = 1 = relMSE(˜ηru, ru) and the continuity of relMSE(˜ηr, s) in r, which is entailed by the continuity of E|˜ηr|2 and br in r (cf. Proposition 2.1.9 and Ruckdeschel (2005b)), the shown monotonicity (2.2.31), respectively (2.2.32) together with the intermediate value

theorem implies (2.2.14) and (2.2.15). ////

Remark 2.2.4 (a)The proof of the preceding lemma is based on arguments pro-vided by P. Ruckdeschel.

(b)If ¯r= 0 (cf. Remark3.1.3(b)), respectively rl≥r¯, there is only one solu-tion ˜η on [0,∞] , respectively [rl, ru] ; confer Propositions2.1.3and 2.1.5. Hence, we have to consider bounded intervals in the previous lemma.

(c)The calculation of some radius r0 such that both boundary values are equal leads to an AL estimator that is radius–minimax; i.e., minimizes the maximum inefficiency over the given radius range. We call this radius r0 least favorable. In case the true radius is completely unknown, respectively unknown except to belong to some radius interval, we recommend to use this optimally robust estimator.

(d) Ruckdeschel and Rieder (2004) prove that Lemma 2.2.3 (a) holds more generally for a large class of optimization problems of form (1.3.10) where G in addition has to fulfill a certain homogeneity condition; confer Theorem 6.1 (a) (idid.). In particular, by parameterizing ˜η not by the radius r but by the optimal clipping bound b, they show, that the radius–minimax IC for completely unknown radius r is the same for all Lq-risks (q∈[1,∞) ); confer Theorem 6.1 (b) (idid.).

(e)InRieder et al.(2001) and also in this thesis we consider the cases that the radius r is completely unknown, respectively unknown up to a factor of 1/3 or 1/2 ;

40 Supplements to the Asymptotic Theory of Robustness

i.e., any r3 or r2 such that the true radius r certainly would stay within [13r3,3r3] or [12r2,2r2] , respectively. In a second step, we then determine least favorable values of r3 and r2 by maximizing the minimax subefficiencies over [13r3,3r3] and

[12r2,2r2] , respectively. ////

In the following remark we state the conclusions ofRieder et al.(2001), Section 1.6.

Remark 2.2.5 (a) The minimax subefficiency is small. Small in compari-son with the most robust estimators, and small for practical purposes. Consistent estimation of the radius from the data hence seems neither necessary nor worth-while – however, under the provision that the radius–minimax robust estimator is employed.

(b) The least favorable radii are small. This surprising fact seems to confirm Huber(1997), p 61, who distinguishes robustness from diagnostics by its purpose to safeguard against – as opposed to find and identify – deviations from the assumptions; in particular, to safeguard against deviations below or near the limits of detectability. LikeHuber(1997), the small least favorable radii we obtain might question the breakdown literature, which is concerned only with (stability under) large contamination and, at most, (efficiency under) zero contamination. ////