• Keine Ergebnisse gefunden

Does polynomial parallel volume imply convexity?

N/A
N/A
Protected

Academic year: 2022

Aktie "Does polynomial parallel volume imply convexity?"

Copied!
11
0
0

Wird geladen.... (Jetzt Volltext ansehen)

Volltext

(1)

DOI: 10.1007/s00208-003-0497-7

Math. Ann. (2004)

Mathematische Annalen

Does polynomial parallel volume imply convexity?

Matthias Heveling·Daniel Hug·G ¨unter Last

Received: 1 April 2003 / Revised version: 16 July 2003 / Published online: 6 January 2004 – © Springer-Verlag 2004

Abstract. For a non-empty compact setARd,d2, andr0, letA⊕rdenote the set of points whose distance fromAisrat the most. It is well-known that the volume,Vd(A⊕r), ofA⊕r

is a polynomial of degreedin the parameterrifAis convex. We pursue the reverse question and ask whetherAis necessarily convex ifVd(A⊕r)is a polynomial inr. An affirmative answer is given in dimensiond=2, counterexamples are provided ford3. A positive resolution of the question in all dimensions is obtained if the assumption of a polynomial parallel volume is strengthened to the validity of a (polynomial) local Steiner formula.

Mathematics Subject Classification (2000): 52A38, 28A75, 52A22, 53C65

1. Introduction and results

Convex sets in a Euclidean space can be characterized in various ways within the class of closed sets. Among the classical characteristic properties are the exis- tence of supporting hyperplanes through each boundary point and the existence of a nearest point map. The present paper investigates an apparently new charac- terization of the convexity of a closed set, in terms of the volumes of its (local) parallel sets.

LetAdenote a non-empty compact subset ofRd,d ≥2, and letA⊕r be the parallel set of Aat distance r ≥ 0, i.e. the set of all pointsx ∈ Rd whose dis- tance fromAis at mostr. Further, letVd denote the volume function (Lebesgue measure). In the following, we consider the parallel volume,Vd(A⊕r), ofAas a function of the distance parameterr ≥0. IfAis convex, then

M. Heveling

Institut f¨ur Mathematische Stochastik und Mathematisches Institut II, Universit¨at Karlsruhe (TH), Englerstr. 2, D-76128 Karlsruhe, Germany

(e-mail:matthias.heveling@math.uni-karlsruhe.de) D. Hug

Mathematisches Institut, Albert-Ludwigs-Universit¨at Freiburg, Eckerstr. 1, D-79104 Freiburg, Germany (e-mail:daniel.hug@math.uni-freiburg.de)

G. Last

Institut f¨ur Mathematische Stochastik, Universit¨at Karlsruhe (TH), Englerstr. 2, D-76128 Karlsruhe, Germany (e-mail:g.last@math.uni-karlsruhe.de)

(2)

Vd(A⊕r)= d j=0

rd−jκd−jVj(A), (1.1)

whereκj is the (j-dimensional) volume of the Euclidean unit ball inRj and the coefficients V0(A), . . . , Vd(A) are the intrinsic volumes of the convex bodyA (see, e.g., [16, (4.2.27)]). It is well known thatVd−1(A)is half the surface area (if Ahas non-empty interior) andV0(A)=1. The Steiner formula (1.1) and its ram- ifications are of central importance in geometry. Applications of the fundamental relation (1.1) and of the functionalsVjdefined by it can be found in various other branches of mathematics including stochastic geometry [18], [17], statistics [9], [11], discrete mathematics [2], geometric functional analysis [7], and recently also in physics [12], [13].

It is natural to investigate conditions under which a Steiner formula of the form (1.1) can be obtained for sets which are not necessarily convex. Results in this direction, which are available in the literature, usually state a Steiner formula under regularity assumptions on the set and for a restricted range of the distance parameterr; see, for instance, the classical contributions by Hadwiger [8] and Federer [4]. In the present paper, we reverse this point of view and ask the fol- lowing question.

Does polynomial parallel volume imply convexity?

It is tempting to guess a positive answer to this question. Simple examples (see Section 4), however, show that, for all dimensions d ≥3, there are non-convex sets whose parallel volume is a polynomial. On the other hand, in dimensiond =2 we can answer our question in an affirmative way.

Theorem 1. LetA⊂R2be compact such that

V2(A⊕t)=V2(A)+c1t +c2t2, t ≥0, (1.2) for somec1, c2∈R. ThenAis convex.

Note that the seemingly more general assumption thatV2(A⊕t)is a polynomial of degreen ∈ Nin the parametert ≥ 0 immediately implies (1.2). Theorem 1 provides a characterization of convexity by a global property. A corresponding result is wrong in higher dimensions, but the conclusion of Theorem 1 remains true under a stronger (local) hypothesis. We prepare such a result, together with an extension, by introducing some more notation.

LetAdenote an arbitrary non-empty closed subset ofRd. The distanced(A, z) from a pointz ∈ Rd to a setA ⊂Rd is defined as inf{|y−z| : yA}, where

| · |denotes the Euclidean norm onRd and inf∅ := ∞. We putp(A, z) := y wheneveryis a uniquely determined point inAwithd(A, z)= |y−z|, and we call this point the metric projection ofzonA. If 0< d(A, z) <∞andp(A, z) is defined, then p(A, z) lies on the boundary∂A of A and we putu(A, z) :=

(3)

(zp(A, z))/d(A, z). The exoskeleton exo(A) of A consists of all points of Rd \A which do not admit a metric projection onA. This is a measurable set (see, e.g., Lemma 6.1 in [10]) satisfyingHd(exo(A))=0, whereHddenotes the d-dimensional Hausdorff measure. Clearly, ifAis convex, then exo(A)= ∅. We extend the definition ofp(A, z)∈Rdandu(A, z)Sd−1in a suitable and mea- surable way to allz∈Rd. HereSd−1denotes the unit sphere{z∈Rd :|z| =1}. Now letA ⊂ Rd be a non-empty closed convex set. Extending and unify- ing results by Aleksandrov [1], Fenchel & Jessen [6] and Federer [4], Schneider [14], [15] has shown thatAsatisfies a local Steiner formula. It states that, for any measurable bounded function f : Rd ×Sd−1 → Rwith compact support and t ≥0,

Rd\A f (p(A, z), u(A, z))1{d(A, z)≤t}dz

=

d−1

j=0

td−jκd−j

f (x, u)Cj(A;d(x, u)), (1.3) where dz denotes integration with respect to Lebesgue measure. The measures Cj(A; ·),j ∈ {0, . . . , d −1}, are the support measures (generalized curvature measures) of the convex setA, which simultaneously generalize the surface area measures and the curvature measures ofA. They are non-negative, locally finite and concentrated on∂A×Sd−1. We refer to [16] for further background infor- mation.

We now modify our initial question as follows:

Does the validity of a local Steiner formula imply convexity?

The positive answer is given in the following theorem.

Theorem 2. LetA ⊂ Rd be non-empty and closed. Assume that for any mea- surable bounded functionf : Rd×Sd−1 →Rwith compact support there are c1(f ), . . . , cd(f )∈Rsuch that, for allt >0,

Rd\Af (p(A, z), u(A, z))1{d(A, z)≤t}dz=d

j=1

cj(f )tj.

ThenAis convex.

Theorem 2 is an immediate consequence of a more general result which is stated as Theorem 3 below. Let again A ⊂ Rd be a closed set. The reach of A, denoted by reach(A), is defined as the supremum of allr > 0 such that the metric projection of zon Ais defined for allz ∈ Rd withd(A, z) < r, where sup∅:=0. We say thatAhas positive reach whenever reach(A) >0. Of course, a closed convex setAsatisfies reach(A)= ∞. An extension of the local Steiner formula (1.3) to sets with positive reach was established by Federer [4], a simpler

(4)

and more general approach was later developed by M. Z¨ahle [19]. To state such a local Steiner formula, letA⊂Rdsatisfy reach(A)r. Then, for any measurable bounded functionf :Rd×Sd−1→Rwith compact support andt(0, r),

Rd\Af (p(A, z), u(A, z))1{d(A, z)≤t}dz

=

d−1

j=0

td−jκd−j

f (x, u)Cj(A;d(x, u)).

HereCj(A; ·),j ∈ {0, . . . , d−1}, are locally finite signed measures on the Borel sets ofRd×Sd−1which extend the support measures of convex sets.

The following theorem provides a corresponding characterization of sets with positive reach which generalizes Theorem 2.

Theorem 3. LetA ⊂ Rd be non-empty and closed. Assume that for any mea- surable bounded functionf : Rd×Sd−1 →Rwith compact support there are c1(f ), . . . , cd(f )∈Rsuch that, for allt(0, r),

Rd\Af (p(A, z), u(A, z))1{d(A, z)≤t}dz=d

j=1

cj(f )tj. (1.4) Then reach(A)r.

For the proof of Theorem 1, we will compare the given compact set with its convex hull, provide a result concerning the differentiation of the parallel volumes of a compact set, and (in a sense) exploit the fact that among all rectifiable curves connecting two points precisely the segment has minimal length. The crucial tool for the proof of Theorem 3 is a Steiner-type formula for arbitrary closed subsets of Rd that has recently been developed in [10]. We give a brief introduction to such a formula at the beginning of Section 3.

In the following, the Euclidean ball inRdwith centera ∈Rdand radiusr ≥0 is denoted byBd(a, r). Thei-dimensional Hausdorff measure is denoted byHi. The interior of a setA⊂Rdis denoted by int(A).

2. Proof of Theorem 1

The two-dimensional special case of the following lemma will play a crucial role in the first part of the proof of Theorem 1. In the very special case of a compact convex set, a stronger assertion is well known.

Lemma 1. LetA⊂Rdbe compact. Then d

dtVd(A⊕t)=Hd−1(∂A⊕t) forH1-almost allt >0.

(5)

Proof. Putf (x):=d(A, x),x ∈Rd. By Lemma 3.2.34 in [5], we have Hd(A⊕t \A)=

t

0

Hd−1(f1({s}))ds. (2.1) It is easy to check that∂A⊕sf1({s})for alls >0. In general, this inclusion cannot be replaced by an equality. Instead, Hd−1(f1({s})\∂A⊕s) = 0 will be shown for H1-almost all s > 0. SinceHd(exo(A)) = 0, an application of the coarea formula (see [5]), applied to the distance functionf, shows that, for H1-almost alls >0,

Hd−1(exo(A)f1({s}))=0. (2.2) Lets >0 satisfy (2.2) and choose anyxf1({s})\exo(A). ThenxA⊕s, and we will show thatx /∈int(A⊕s). In fact, sincex /∈exo(A), there is a uniquely determined point yA such that Bd(x, s)A = {y}. Hence, for any zA\int(Bd(y, s)), we have|z−x|> s. This implies the existence of some >0 such thatz /Bd(x, s+)wheneverzA\int(Bd(y, s)). Thus we get

A⊂[Rd\Bd(x, s+)]∪[Bd(y, s)\int(Bd(x, s))]. (2.3) From (2.3) we infer thatx+λ|xy|1(xy) /A⊕s forλ(0, ).

So far we have shown thatf1({s})can be replaced by∂A⊕sunder the integral in (2.1). The assertion of the lemma then follows by a well known property of

absolutely continuous functions.

Proof of Theorem 1. Let A ⊂ R2 be a compact and non-empty set satisfying (1.2). We will show thatAis equal to its convex hullC :=conv(A). The proof is divided into three steps.

I. SinceCis convex,

V2(C⊕t)=V2(C)+d1t+κ2t2, t ≥0, (2.4) whered1is the boundary length ofC. Since there is someaAandAC, we get{a}⊕tA⊕tC⊕t, and hence

κ2t2V2(A⊕t)V2(C⊕t). (2.5) A comparison of (1.2), (2.4) and (2.5) first shows thatκ2c2κ2, thusκ2=c2, and consequentlyc1d1. Therefore, for allt >0,

d

dtV2(A⊕t)=c1+2κ2td1+2κ2t = d

dtV2(C⊕t). (2.6) Combining (2.6) and Lemma 1, we get, forH1-almost allt >0,

H1(∂A⊕t)H1(∂C⊕t). (2.7)

(6)

II. We fixt > diam(A)(the diameter ofA) such that (2.7) is satisfied. Then by Lemma 1 in [3] and after a translation, the setA⊕t is a star body (see [16]).

In particular, ϕ : S1∂A⊕t, uρ(A⊕t, u)uis a homeomorphism, where ρ(A⊕t,·) denotes the radial function of A⊕t. Since A⊕tC⊕t, we have that ρ(A⊕t, u)ρ(C⊕t, u)for alluS1. Now we define the spherical set:= {u∈ S1:ρ(A⊕t, u) < ρ(C⊕t, u)}.

Sinceis an open subset ofS1, there is an at most countable (possibly empty) familyS of mutually disjoint non-empty open spherical segmentsωS1such that =

ω∈Sω. Each support line of C⊕t contains a point of∂A⊕t∂C⊕t. Hence each open halfspace contains some u /. Therefore, for allωS, H1(ω) < πandωcan be represented as the open geodesic segment between two pointsαωandβωinS1. Letωdenote the closure ofω.

Letψ :S1∂C⊕tbe defined byψ(u):=ρ(C⊕t, u)u. Thenaω :=ϕ(αω)= ψ(αω)andbω :=ϕ(βω)=ψ(βω)forωS. Next we show thatIω :=ψ(ω)is a segment forωS. Choose an arbitrary pointγωω, putcω :=ψ(γω), and letHω

denote a support line ofC⊕tatcω. It is sufficient to show thataω, bωHω. Since C⊕t = conv(A⊕t)andcωHω, there are pointsaω, bω∂A⊕tHω∂C⊕t

such that cω(aω, bω), where aω = bω by the choice of cω. It follows that aω, bω/ ψ(ω)so thataω ∈ conv({0, aω, bω}). Sinceaω is a boundary point of C⊕t, it lies in [0, aω)∪[0, bω)or in [aω, bω]. The first case is clearly impossible as 0∈int(C⊕t). Henceaω ∈[aω, bω]⊂Hω, and by symmetrybωHω.

We putω :=ϕ(ω)and next prove that

H1(ω)H1(Iω) (2.8)

for allωS. LetHω denote the line which containsIω, and letJω denote the orthogonal projection of ω onto Hω. Sinceaω, bωJω and ω is (topologi- cally) connected, we get thatIωJω. Moreover, since the orthogonal projection ontoHωis a contraction, we obtain thatH1(ω)H1(Jω)H1(Iω), i.e. (2.8).

Inequality (2.7) implies that

0≤H1(∂C⊕t)H1(∂A⊕t)=

ω∈S

(H1(Iω)H1(ω)). (2.9)

From (2.8) and (2.9) we deduce that

H1(ω)=H1(Iω)

forωS. For anyωS, we havepω :=ϕ(γω) /Iω. Letωandωdenote the geodesic segments connectingαω, γωandγω, βωinS1. By the triangle inequality and by repeating the argument leading to (2.8), we then get

H1(Iω) <|pωaω| + |bωpω| ≤H1(ϕ(ω))+H1(ϕ(ω))=H1(ω), a contradiction. This implies thatS = ∅, and thereforeA⊕t =C⊕t is convex.

(7)

III. So far we have shown thatA⊕tis convex forH1-almost allt >diam(A). HenceA⊕r is convex for everyr >diam(A), and thusA⊕r =C⊕r. In particular, V2(A⊕r)=V2(C⊕r)forr > diam(A), and hence by (1.2) and the convexity of C,

V2(A)+c1r +κ2r2=V2(C)+d1r+κ2r2, (2.10) first for r > diam(A), but then also for any r ≥ 0. This in turn shows that V2(A⊕r)=V2(C⊕r)forr ≥0. SinceA⊕r is compact andC⊕r is the closure of its interior forr >0, we deduce thatA⊕r =C⊕r is convex for anyr > 0. This

implies the asserted convexity ofA.

In the second part of the proof of Theorem 1, one can use a global integralgeometric Crofton formula instead of arguing locally by distinguishing between different parts of the boundaries ofA⊕t andC⊕t.

3. Support measures and proof of Theorem 3

First, we recall the main result of [10]. LetA⊂ Rd be a non-empty closed set.

The normal bundle ofAis defined by

N(A):= {(p(A, z), u(A, z)):z /A∪exo(A)}.

It is a measurable subset of∂A×Sd−1. The reach functionδ(A,·):Rd×Sd−1→ [0,∞] ofAis defined by

δ(A, x, u):=inf{t ≥0 :x+tu∈exo(A)}, (x, u)N(A),

and δ(A, x, u) := 0 for(x, u) /N(A). Note that δ(A,·) > 0 on N(A). By Lemma 6.2 in [10], δ(A,·) is a measurable function. The reach function ofA localizes the notion of reach introduced in Section 1 (see also [4]). It is easy to check that reach(A)=inf{δ(A, x, u):(x, u)N(A)}. In particular,Ais convex if and only ifδ(A,·)≡ ∞onN(A).

In [10], the support measuresµ0(A; ·), . . . , µd−1(A; ·)of a non-empty closed setA⊂Rd have been introduced as real-valued functions, defined on all Borel subsets ofRd×Sd−1which are contained in

((Rd×Sd−1)\N(A))∪ {(x, u):xB, δ(A, x, u)s},

for somes > 0 and some compactB ⊂Rd. These signed measures vanish on each Borel subset of(Rd×Sd−1)\N(A). The total variation measure ofµi(A; ·) is denoted by|µi|(A; ·).

Putab:=min{a, b}fora, b∈R. Then Theorem 2.1 in [10] states that

N(A)

1{x ∈B}(δ(A, x, u)r)d−jj|(A;d(x, u)) <∞,

(8)

j = 0, . . . , d −1, for all compact setsB ⊂ Rd and allr > 0, and, for any measurable bounded functionf :Rd →Rwith compact support,

Rd\Af (z)dz=

d−1

i=0

(di)κd−i

0

N(A)

td−1−i1{t < δ(A, x, u)}

×f (x+tu)µi(A;d(x, u))dt. (3.1) The support measures are uniquely determined by the local Steiner-type formula (3.1).

Proof of Theorem 3. We fix someτ(0, r)and a compact setB ⊂Rd. Then we put

Nτ,B := {(x, u)∈N(A):δ(A, x, u)τ, xB}.

Letf : Rd ×Sd−1 →[0,1] be defined byf (x, u) :=1{(x, u) ∈ Nτ,B}. Then (1.4) and (3.1) imply that, for allt(0, r),

d j=1

cj(f )tj =

d−1

i=0

(di)κd−i

t

0

Nτ,B

sd−1−i

×1{s < δ(A, x, u)}µi(A;d(x, u))ds (3.2) for somec1(f ), . . . , cd(f ) ∈R. The function on the right-hand side of (3.2) is independent oft fort(τ, r). Hence the polynomial on the left-hand side of (3.2) must be zero, which implies that

Hd({zA⊕r \A:(p(A, z), u(A, z))Nτ,B})=0.

Since this is true for anyτ(0, r)and any compact setB ⊂Rd, we obtain that δ(A, p(A, z), u(A, z))rforHd-almost allzA⊕r\A. We will show that this implies that reach(A)r.

Let z ∈ Rd \A with r0 := d(A, z) < r. There is a sequence of points ziA⊕r \A, i ∈ N, converging toz asi → ∞for which the metric projec- tion pi := p(A, zi)A exists and δ(A, pi, ui)r withui := u(A, zi)Sd−1. Passing to a subsequence and changing notation (if necessary) we can assume that pipA and uiuSd−1. Chooser1(r0, r). Then Bd(pi+r1ui, r1)A= {pi}for alli ∈N; henceAdoes not intersect the inte- rior ofBd(p+r1u, r1). Clearly,zi =pi +d(A, zi)ui fori ∈ N, and therefore z=p+d(A, z)u=p+r0u. Moreover,Bd(p+r0u, r0)∩A= {p}, which shows thatpis the unique nearest point ofz inA. This proves the required assertion.

(9)

4. Examples

This section is devoted to the construction of non-convex, compact sets in Rd, d ≥ 3, whose parallel volumes are polynomials. For this purpose let L be a (d −2)-dimensional linear subspace ofRd, and letCLbe a(d−2)-dimen- sional compact convex set. Then fix a closed setAˆ ⊂ CwithLC ⊂ ˆA, where

LC denotes the boundary ofC relative toL. Then, fort ≥diam(C)/2, we will show that the setA:= ˆA⊕t ⊂Rd has polynomial parallel volume, i.e.

Vd(A⊕r)=d

k=0

ckrk, r ≥0,

whereck ∈Rfork =0, . . . , d. The setAis convex if and only ifAˆ is convex.

SinceA⊕r = ˆA⊕(t+r), we deduce that

Vd(A⊕r)=Vd(C⊕(t+r))Vd(C⊕(t+r)\ ˆA⊕(t+r)),

whereVd(C⊕(t+r))is a polynomial inr ≥0 due to the convexity ofC. As we will now see, the second term in the sum does not depend onr ≥ 0, but is in fact a constant.

For anyyL, Aˆ⊕(t+r)(y+L)=

y+BL

(t+r)2d(A, y)ˆ 2

,ifd(A, y)ˆ ≤t +r,

∅, otherwise,

whereBL(s)=Bd(0, s)Ldenotes the ball inLwith center 0 and radius s. This equation also holds withAˆ replaced byC. Moreover, sinceLC⊂ ˆA, we get thatd(A, y)ˆ =d(C, y)wheneveryL\C, and thus

Aˆ⊕(t+r)((L\C)×L)=C⊕(t+r)((L\C)×L). (4.1) Therefore,

Vd(C⊕(t+r)\ ˆA⊕(t+r))

=Vd(C⊕(t+r))Vd(Aˆ⊕(t+r))

=

C

L

1{y+zC⊕(t+r)} −1{y+z∈ ˆA⊕(t+r)} H2(dz)Hd−2(dy),

where Fubini’s theorem and (4.1) were applied in the last step. Since t ≥ diam(C)/2, and henced(A, y)ˆ ≤ t +r forr ≥ 0 andyC, we further de- duce that

Vd(C⊕(t+r)\ ˆA⊕(t+r))

=

C

H2(BL(t+r))H2

BL

(t+r)2d(A, y)ˆ 2

Hd−2(dy)

=π

L

d(A, y)ˆ 2Hd−2(dy),

(10)

which is independent ofr ≥ 0. Thus,Vd(A⊕r)is a polynomial inr ≥ 0 with non-negative coefficients.

Turning our attention to special cases, in dimensiond =3 for instance,Ccan be a segment of length 2 andAˆ the set consisting of the two endpoints ofC. For t =1, the setAis then the union of two touching balls with radius 1.

More generally, letAˆ ⊂Lbe a compact set whose convex hull is a ball inL with radiuss and takeA := ˆA⊕t for somets. (In dimensiond = 3,Aˆ can be any compact subset of a line.) Then the parallel volume ofAis a polynomial.

However, ifAˆ is non-convex, thenAis non-convex as well.

IfA ⊂ Rd is any one of the sets constructed above, ifn is an even natural number, and ifRd is identified with a d-dimensional linear subspace ofRd+n, then the parallel volume ofAinRd+nis again a polynomial.

References

1. Aleksandrov, A.D.: Zur Theorie der gemischten Volumina von konvexen K¨orpern, I. Verall- gemeinerung einiger Begriffe der Theorie der konvexen K¨orper (in Russian). Mat. Sbornik N.S. 2, 947–972 (1937)

2. Betke, U., Henk, M., Wills, J.M.: Finite and infinite packings. J. Reine Angew. Math. 453, 165–191 (1994)

3. Brown, M.: Sets of constant distance from a planar set. Michigan Math. J. 19, 321–323 (1972)

4. Federer, H.: Curvature measures. Trans. Amer. Math. Soc. 93, 418–491 (1959) 5. Federer, H.: Geometric Measure Theory. Springer, Berlin, 1969

6. Fenchel, W., Jessen, B.: Mengenfunktionen und konvexe K¨orper. Danske Vid. Selsk. Mat.- Fys. Medd. 16, 1–31 (1938)

7. Giannopoulos, A.A., Milman, V.D.: Euclidean structure in finite dimensional normed spaces, 707–779, W. B. Johnson, J. Lindenstrauss (Eds.), Handbook of the Geometry of Banach Spaces. Volume 1, Elsevier, Amsterdam, 2001

8. Hadwiger, H.: ¨Uber das Volumen der Parallelmengen, Mitt. Naturf. Ges. Bern 3, 121–125 (1946)

9. Hotelling, H.: Tubes and spheres inn-space, and a class of statistical problems. Amer. J.

Math. 61, 440–460 (1939)

10. Hug, D., Last, G., Weil, W.: A local Steiner–type formula for general closed sets and appli- cations. Math. Z. 246, 237–272 (2004)

11. Kuriki, S., Takemura, A.: Tail probabilities of the maxima of multilinear forms and their applications. Ann. Stat. 29, 328–371 (2001)

12. Mecke, K.R., Stoyan, D.: (Eds.), Statistical Physics and Spatial Statistics - The Art of Ana- lyzing and Modeling Spatial Structures and Pattern Formation. Lecture Notes in Physics 554, Springer-Verlag, Berlin, Heidelberg, New York, 2000

13. Mecke, K.R., Stoyan, D.: (Eds.), Morphology of Condensed Matter - Physics and Geome- try of Spatially Complex Systems. Lecture Notes in Physics 600, Springer-Verlag, Berlin, Heidelberg, New York, 2002

14. Schneider, R.: Bestimmung konvexer K¨orper durch Kr¨ummungsmaße. Comment. Math.

Helv. 54, 42–60 (1979)

15. Schneider, R.: Parallelmengen mit Vielfachheit und Steiner-Formeln. Geom. Dedicata 9, 111–127 (1980)

(11)

16. Schneider, R.: Convex Bodies: the Brunn-Minkowski Theory. Encyclopedia of Mathemat- ics and its Applications 44, Cambridge University Press, Cambridge, 1993

17. Schneider, R., Weil, W.: Stochastische Geometrie. Teubner Skripten zur Mathematischen Stochastik, Teubner, Stuttgart, 2000

18. Stoyan, D., Kendall, W.S., Mecke, J.: Stochastic Geometry and its Applications. Second Edition, Wiley, Chichester, 1995

19. Z¨ahle, M.: Integral and current representation of Federer’s curvature measures. Arch. Math.

46, 557–567 (1986)

Referenzen

ÄHNLICHE DOKUMENTE

Pbtscher (1983) used simultaneous Lagrange multiplier statistics in order to test the parameters of ARMA models; he proved the strong consistency of his procedure

PLASMA, the model equations are solved on an unstructured, triangular grid, which is dynamically adapted to the evolving atmospheric flow every time step.. The dynamic grid

New Thought, from the point of view of the West, has a further advantage over theosophy, an advantage of an empirical and accidental nature, but for this very reason

Methods of estimation of polynomial distributed lags in econo- metrics and procedures relating tree ring growth data to climatic and hydrologic data are shown to be equivalent to

As the proof of Corollary 6 (which showed Hölder calmness for systems of polynomi- als) is in the end based on Proposition 2 (Hörmander’s error bound), an analysis of the proof of

Our final example is a stationary lattice germ process M such that a germ-grain model with deterministic spherical grains has a spherical contact distribution function that is of

Theorem 2.0.2 not only says that every automorphism is tame, but also gives us a de- composition of the group as an amalgamated product of two subgroups.. In this section we

Open Access This article is licensed under a Creative Commons Attribution 4.0 Interna- tional License, which permits use, sharing, adaptation, distribution and reproduction in