Home Extremizers of the Alexandrov–Fenchel inequality within a new class of convex bodies
Article
Licensed
Unlicensed Requires Authentication

Extremizers of the Alexandrov–Fenchel inequality within a new class of convex bodies

  • Daniel Hug and Paul A. Reichert EMAIL logo
Published/Copyright: January 23, 2025
Become an author with De Gruyter Brill

Abstract

Mixed volumes in n-dimensional Euclidean space are functionals of n-tuples consisting of convex bodies K, L, C1, …, Cn–2. The Alexandrov–Fenchel inequalities are fundamental inequalities between mixed volumes of convex bodies, which cover as very special cases many important inequalities between basic geometric functionals. The problem of characterizing completely the equality cases in the Alexandrov–Fenchel inequality is wide open. Major recent progress was made by Yair Shenfeld and Ramon van Handel [25; 26], in particular they resolved the problem in the cases where K, L are general convex bodies and C1, …, Cn–2 are polytopes, zonoids or smooth bodies (under some dimensional restriction).

We introduce the class of polyoids, which includes polytopes, zonoids and triangle bodies, and characterize polyoids by using generating measures. Based on this characterization and Shenfeld and van Handel's contribution, we extend their result to a class of convex bodies containing all polyoids and smooth bodies (under some dimensional restriction). Our result is stated in terms of the support of the mixed area measure of the unit ball Bn and convex bodies C1, …, Cn–2. A geometric description of this support is provided in the accompanying work [12].

MSC 2010: 52A39; 52A20; 52A21; 52A40

Funding statement: D. Hug was supported by DFG research grant HU 1874/5-1 (SPP 2265).

Acknowledgements

The authors are grateful to Ramon van Handel for helpful comments on an earlier version of the manuscript.

  1. Communicated by: M. Henk

References

[1] A. D. Alexandrov, Selected works. Part I, volume 4 of Classics of Soviet Mathematics. Gordon and Breach Publishers, Amsterdam 1996. MR1629804 Zbl 0960.01035Search in Google Scholar

[2] C. Berg, Shephard’s approximation theorem for convex bodies and the Milman theorem. Math. Scand. 25 (1969), 19–24. MR261442 Zbl 0185.5020210.7146/math.scand.a-10936Search in Google Scholar

[3] P. Billingsley, Convergence of probability measures. Wiley-Interscience 1999. MR1700749 Zbl 0944.6000310.1002/9780470316962Search in Google Scholar

[4] S. H. Chan, I. Pak, Equality cases of the Alexandrov-Fenchel inequality are not in the polynomial hierarchy. In: STOC’24—Proceedings of the 56th Annual ACM Symposium on Theory of Computing, 875–883, ACM, New York 2024. MR4764866 Zbl arXiv:2309.0576410.1145/3618260.3649646Search in Google Scholar

[5] D. Cordero-Erausquin, B. Klartag, Q. Merigot, F. Santambrogio, One more proof of the Alexandrov-Fenchel inequality. C. R. Math. Acad. Sci. Paris 357 (2019), 676–680. MR4015332 Zbl 1428.5201310.1016/j.crma.2019.07.004Search in Google Scholar

[6] A. Deza, L. Pournin, Diameter, decomposability, and Minkowski sums of polytopes. Canad. Math. Bull. 62 (2019), 741–755. MR4028484 Zbl 1429.5201410.4153/S0008439518000668Search in Google Scholar

[7] P. Goodey, D. Hug, W. Weil, Kinematic formulas for area measures. Indiana Univ. Math. J. 66 (2017), 997–1018. MR3663334 Zbl 1376.5200510.1512/iumj.2017.66.6047Search in Google Scholar

[8] P. Goodey, M. Kiderlen, W. Weil, Section and projection means of convex bodies. Monatsh. Math. 126 (1998), 37–54. MR1633259 Zbl 0922.5200110.1007/BF01312454Search in Google Scholar

[9] P. Goodey, W. Weil, A uniqueness result for mean section bodies. Adv. Math. 229 (2012), 596–601. MR2854184 Zbl 1232.5200510.1016/j.aim.2011.09.009Search in Google Scholar

[10] P. Goodey, W. Weil, Sums of sections, surface area measures, and the general Minkowski problem. J. Differential Geom. 97 (2014), 477–514. MR3263512 Zbl 1301.5200910.4310/jdg/1406033977Search in Google Scholar

[11] E. Grinberg, G. Zhang, Convolutions, transforms, and convex bodies. Proc. London Math. Soc. (3) 78 (1999), 77–115. MR1658156 Zbl 0974.5200110.1112/S0024611599001653Search in Google Scholar

[12] D. Hug, P. A. Reichert, The support of mixed area measures involving a new class of convex bodies. J. Funct. Anal. 287 (2024), Paper No. 110622. MR4790895 Zbl arXiv:2309.1687210.1016/j.jfa.2024.110622Search in Google Scholar

[13] D. Hug, W. Weil, Lectures on convex geometry. Springer 2020. MR4180684 Zbl 1487.5200110.1007/978-3-030-50180-8Search in Google Scholar

[14] J. Lukeš, J. Malý, I. Netuka, J. Spurný, Integral representation theory, volume 35 of De Gruyter Studies in Mathematics. De Gruyter 2010. MR2589994 Zbl 1216.4600310.1515/9783110203219Search in Google Scholar

[15] Z. Y. Ma, Y. Shenfeld, The extremals of Stanley’s inequalities for partially ordered sets. Adv. Math. 436 (2024), Paper No. 109404, 72 pages. MR4669326 Zbl 0778163810.1016/j.aim.2023.109404Search in Google Scholar

[16] R. R. Phelps, Lectures on Choquet’s theorem. Springer 2001. MR1835574 Zbl 0997.4600510.1007/b76887Search in Google Scholar

[17] W. Ricker, A new class of convex bodies. In: Papers in algebra, analysis and statistics (Hobart, 1981), volume 9 of Contemp. Math., 333–340, Amer. Math. Soc. 1981. MR655993 Zbl 0485.5200510.1090/conm/009/655993Search in Google Scholar

[18] G. T. Sallee, On the indecomposability of the cone. J. London Math. Soc. (2) 9 (1974/75), 363–367. MR370356 Zbl 0294.5200410.1112/jlms/s2-9.2.363Search in Google Scholar

[19] R. Schneider, On the Aleksandrov-Fenchel inequality. In: Discrete geometry and convexity (New York, 1982), volume 440 of Ann. New York Acad. Sci., 132–141, New York Acad. Sci., New York 1985. MR809200 Zbl 0567.5200410.1111/j.1749-6632.1985.tb14547.xSearch in Google Scholar

[20] R. Schneider, On the Aleksandrov-Fenchel inequality involving zonoids. Geom. Dedicata 27 (1988), 113–126. MR950327 Zbl 0644.5200610.1007/BF00181617Search in Google Scholar

[21] R. Schneider, Simple valuations on convex bodies. Mathematika 43 (1996), 32–39. MR1401706 Zbl 0864.5200910.1112/S0025579300011578Search in Google Scholar

[22] R. Schneider, Convex bodies: the Brunn-Minkowski theory, volume 151 of Encyclopedia of Mathematics and its Applications. Cambridge Univ. Press 2014. MR3155183 Zbl 1287.52001Search in Google Scholar

[23] R. Schneider, W. Weil, Stochastic and integral geometry. Springer 2008. MR2455326 Zbl 1175.6000310.1007/978-3-540-78859-1Search in Google Scholar

[24] Y. Shenfeld, R. van Handel, Mixed volumes and the Bochner method. Proc. Amer. Math. Soc. 147 (2019), 5385–5402. MR4021097 Zbl 1457.5200810.1090/proc/14651Search in Google Scholar

[25] Y. Shenfeld, R. van Handel, The extremals of Minkowski’s quadratic inequality. Duke Math. J. 171 (2022), 957–1027. MR4393790 Zbl 1487.5201310.1215/00127094-2021-0033Search in Google Scholar

[26] Y. Shenfeld, R. van Handel, The extremals of the Alexandrov-Fenchel inequality for convex polytopes. Acta Math. 231 (2023), 89–204. MR4652411 Zbl 1529.0503210.4310/ACTA.2023.v231.n1.a3Search in Google Scholar

[27] V. S. Varadarajan, On the convergence of sample probability distributions. Sankhyā 19 (1958), 23–26. MR94839 Zbl 0082.34201Search in Google Scholar

[28] X. Wang, A remark on the Alexandrov-Fenchel inequality. J. Funct. Anal. 274 (2018), 2061–2088. MR3762095 Zbl 1387.3203010.1016/j.jfa.2018.01.016Search in Google Scholar

[29] W. Weil, Kontinuierliche Linearkombination von Strecken. Math. Z. 148 (1976), 71–84. MR400052 Zbl 0309.5200310.1007/BF01187872Search in Google Scholar

A A macroid that is not a polyoid

In this section, we construct an example of a macroid K that is not a polyoid, thereby showing that the class of macroids is strictly larger than the class of polyoids. The presentation is divided into four subsections. First, we provide some auxiliary results concerning the new concept of a zonotope kernel of a polytope, which will be applied in Subsection A.4 (in the proof of Lemma A.15). A specific construction of a convex body K which is a macroid but not a polyoid, as a series of suitable polytopes (Pi)i∈ℕ, is presented in Subsection A.2. In Section A.3 we show that if μ is any macroid-generating measure of K, then the support sets of μ-almost all polytopes P are scaled summands of the corresponding support sets of K. In the final subsection, we show that each of the polytopes Pi employed in the construction of K is a scaled summand of a polytope in the support of μ. From this fact we finally conclude that K is not a polyoid.

At the beginning of each subsection, further comments, motivation and insights should provide some guidance for the reader.

A.1 Zonotope kernels of polytopes

As a preparation for the main proof, we show that every polytope has a unique largest (centered) zonotope summand and that the induced mapping from polytopes to zonotopes is measurable.

Let K, L, M ∈ 𝓚n be convex bodies. If hKhL = hM, then KL := M is called the Minkowski difference of K and L.

Lemma A.1

Let K ∈ 𝓚n be a convex body and let e, f be two linearly independent segments that are summands of K. Then e + f is also a summand of K.

Proof

To show that e + f slides freely in K (see [22, Section 3.2] and in particular Theorem 3.2.2 there), it suffices to consider two-dimensional slices of K parallel to e + f. Hence we can reduce to the case that K is two-dimensional.

Let ± u be the normals of e and ± v the normals of f. As F(e, ± v) are trivial, F(Ke, ± v) are translates of F(K, ± v). So translates of f are not only contained in F(K, ± v) but also in F(Ke, ± v). Then [22, Theorem 3.2.11] yields that f is a summand of Ke. This completes the proof. □

Lemma A.2

The function ζ : 𝓟n → 𝓟n that maps a polytope to its unique largest (i.e. inclusion-maximal) zonotope summand, centrally symmetric around the origin, is well-defined. Every zonotope summand of P ∈ 𝓟n is a summand of ζ(P).

Proof

We show that every polytope P has a unique largest zonotope summand. Let 𝓩(P) denote the nonempty set of origin centered zonotope summands of P. First note that summands of polytopes are polytopes (see [22, p. 157]) and polytopes that are zonoids are zonotopes (see [22, Corollary 3.5.7]). Hence the set of all origin centered zonotopes that are summands of P equals the set of all origin centered zonoids that are summands of P. The latter set is compact as the intersection of a compact set (the set of centered zonoids having mean width less or equal the mean width of P) and a closed set (the set of summands of P). It follows that there is a Z ∈ 𝓩(P) of maximum mean width. This Z is inclusion-maximal in 𝓩(P).

Let Y ∈ 𝓩(P). Then there exist pairwise linearly independent x1, …, xk ∈ 𝕊n−1 and scalars y1, …, yk, z1, …, zk ≥ 0 such that

Y=i=1kyi[xi,xi],Z=i=1kzi[xi,xi].

Assume for a contradiction that Y is not a summand of Z. Up to reordering of the indices, it follows that y1 > z1. Then y1 [− x1, x1] is a summand of P, but, as Z is maximal, not a summand of

Pi=2kzi[xi,xi].

Let ∈ [k] be the largest index such that y1 [−x1, x1] is a summand of P~:=Pi=2zi[xi,xi]. Then l < k and z+1 [−x+1, x+1] is also a summand of . Now Lemma A.1 shows that y1 [−x1, x1] + z+1 [−x+1, x+1] is a summand of , but this contradicts the maximality of . Hence, every Y ∈ 𝓩(P) is a summand of Z, and there is only one maximal zonotope summand of P. □

Next, we aim to prove that ζ is measurable. We write h(K, u) = hK(u) for the support function of K ∈ 𝓚n evaluated at u ∈ 𝕊n–1. We write B(K, r) for a ball with center K and radius r ≥ 0 with respect to the Hausdorff metric d on the space 𝓚n of convex bodies.

Lemma A.3

Let X be a separable metric space and f : X → 𝓚n a function such that for any u ∈ 𝕊n−1 and λ ∈ ℝ,

Sf(u,λ):={xX:h(f(x),u)λ}

is closed. Then f is measurable.

Proof

Fix some countable and dense set Q ⊆ 𝕊n−1. Let K ∈ 𝓚n and r > 0. By continuity of hL for every L ∈ 𝓚n,

B(K,r)=uQh(,u)1([hK(u)r,hK(u)+r]).

Taking the preimage under f, we get

f1(B(K,r))=uQh(f(),u)1([hK(u)r,hK(u)+r]).

By hypothesis, h(f(⋅), u) is Borel measurable for every u ∈ 𝕊n−1. Since Q is countable, f–1(B(K, r)) is a Borel set. Because balls like B(K, r) generate the Borel σ-algebra of 𝓚n, this shows that f is measurable. □

Lemma A.4

ζ is a measurable function.

Proof

We apply Lemma A.3. Let u ∈ 𝕊n−1 and λ ∈ ℝ. It suffices to show that

Sζ(u,λ)={PPn:h(ζ(P),u)λ}

is closed. Let (Pi) be a sequence in Sζ(u, λ) that converges to P ∈ 𝓟n. Applying the Blaschke selection theorem to the bounded sequence of centered zonoids (ζ(Pi)), we find a subsequence (Qi) that converges to a centered zonoid Z. Then Z is a summand of P. Because summands of polytopes are polytopes and polytopes that are zonoids are zonotopes, it follows that Z is a zonotope. So Z is a summand of ζ(P) and, in particular,

h(ζ(P),u)h(Z,u)=limih(ζ(Qi),u)λ.

So PSζ(u, λ), proving that the latter set is closed. An application of Lemma A.3 concludes the proof. □

A.2 Admissible sequences of polytopes

Our strategy to construct a macroid that is not a polyoid is to consider a series of polytopes. Hence, the resulting counterexample has a discrete generating measure. In this section, we define the required properties of the series of polytopes.

Let K ⊆ ℝ3 be a convex body. A support set F(K, u) will be called a singleton or trivial if it is zero-dimensional, an edge if it is one-dimensional, and a facet if it is two-dimensional. It should be observed that unless K is a polytope, the current definition does not imply that the normal cone of K at a point in the relative interior of an edge is two-dimensional.

Definition A.5

Let (Pi) be a bounded sequence of (indecomposable) polytopes in ℝ3 with the following properties:

  1. All facets are triangles.

  2. For every i ∈ ℕ, Pi is a 3-dimensional (i + 3)-tope.

  3. For every i ∈ ℕ, no two edges of Pi have the same direction.

  4. If i, j ∈ ℕ are distinct and u is a facet normal of Pi, then F(Pj, u) is trivial.

  5. If , i, j ∈ ℕ are distinct and e, f, g are edges of P, Pi, Pj, then e + f + g is 3-dimensional. In particular, e + f is 2-dimensional.

  6. K:=i=1Pi is a well-defined convex body.

We call such a sequence admissible and K its associated body.

Remark A.6

If (Pi) is an admissible sequence of polytopes and K is its associated body, then K is a macroid. A suitable choice of a generating measure is given by

μ=i=1w(Pi)w(K)δw(K)w(Pi)Pi

which is a probability measure on 𝓟n with bounded support. Recall that in general μ is not uniquely determined.

Example A.7

Let (Pi)i∈ℕ be a bounded sequence of polytopes such that Pi is a 3-dimensional (i + 3)-tope having only triangular faces with no edge direction occurring twice. When we apply independent uniform random rotations to each of the Pi, we almost surely obtain an admissible sequence.

Remark A.8

Let Ki, i ∈ ℕ, and K be convex bodies in 𝓚n. Then K = i=1 Ki holds (where the convergence of the partial sums is meant with respect to the Hausdorff metric) if and only if hK = i=1 hKi (where the convergence holds pointwise, but then also uniformly on the unit sphere). If u ∈ 𝕊n–1, then the support sets of K and Ki, i ∈ ℕ, are related by F(K, u) = i=1 F(Ki, u).

Remark A.9

Let (Pi)i∈ℕ be an admissible sequence of polytopes and K its associated body. Items (iii) and (v) imply that if F(K, u) is an edge of K, then there is a unique i ∈ ℕ such that F(Pi, u) is an edge of Pi. In this situation, F(K, u) is a translate of F(Pi, u) and no other edge of any of the polytopes Pj, j ∈ ℕ, is parallel to F(K, u). From item (iv) we conclude that if F(K, u) is a triangular facet, then there is a unique i such that F(K, u) is a translate of F(Pi, u). See Lemma A.14 for further discussion.

A.3 Properties of the supporting polytopes

Throughout the section, let (Pi) be an admissible sequence, K the associated macroid and μ any macroid-generating measure of K. It is important that we do not impose any additional requirements on μ because the final goal is to show that no choice of μ is a polyoid-generating measure. To get a better understanding of the polytopes P in supp μ, we aim to show that almost all of them have strongly restrained support sets. In Lemma A.13, we will conclude that for each u ∈ 𝕊2, their support sets F(P, u) are summands of F(K, u), at least up to a scaling factor.

We begin with the simpler case, analyzing the possible generating measures of polytopes. Recall that every summand of a polytope is a polytope; see [22, p. 157]. For a polytope Q ∈ 𝓟n, we consider the convex cone

S(Q):={PPn:RPn,α>0:Q=αP+R}.

The elements of 𝓢(Q) are called scaled summands of Q.

Lemma A.10

Let Q ∈ 𝓟n be a polytope with macroid-generating measure ν on 𝓟n, that is,

hQ=hPν(dP).

Then supp ν ⊆ 𝓢(Q).

Proof

  1. Let β > 0 be a lower bound on the lengths of the edges of Q, and let P ∈ 𝓢(Q) be nontrivial. Then βdiamP P is a summand of Q, as we show first.

    Let F(P, u) be an edge. Since F(P, u) is a scaled summand of F(Q, u), the latter must have an edge e (which is also an edge of Q) homothetic to F(P, u). The length of F(P, u) is at most diam P and the length of e is at least β, so e contains a translate of βdiamP F(P, u). Hence [22, Theorem 3.2.11] implies that βdiamP P is a summand of Q.

  2. The set 𝓢(Q) is closed in 𝓟n, as we show next.

    Let (Pi) be a sequence in 𝓢(Q) converging to some P ∈ 𝓟n. If P is trivial, then P ∈ 𝓢(Q). Otherwise, there are a sequence of nontrivial polytopes (Pi) and a sequence of polytopes (Ri) such that Q=βdiamPiPi+Ri, and (Ri) must also converge to some R ∈ 𝓚n such that Q = βdiamP P + R. As R is a summand of P, it must be a polytope. So P ∈ 𝓢(Q).

  3. Assume for a contradiction that there is some L ∈ supp ν ∖ 𝓢(Q). Then d := d(L, 𝓢(Q)) > 0 and λ := ν(B(L, d/2)) > 0. Define convex bodies L′ and R by

    hL:=λ1B(L,d/2)hPν(dP)andhR:=B(L,d/2)hPν(dP)

    so that Q = λL′ + R. It follows that R is a polytope and L′ ∈ 𝓢(Q), and hence

    d=d(L,S(Q))d(L,L)d/2<d,

    which is a contradiction. □

Recall that we write span A for the linear subspace parallel to the affine hull of a given nonempty set A ⊆ ℝ3.

Lemma A.11

For every edge e of a polytope P in3, there are a normal v ∈ 𝕊2 of e and u ∈ ℚ3span {e} such that uv and e = F(P, v).

Proof

Let U be the relatively open normal cone of the edge e. Then span U is two-dimensional, and U is open in span U. Choose w ∈ 𝕊2U, so that span U = w. Let × denote the cross product. The continuous and surjective map ℝ3 → span U, xw × x, maps the dense set S := ℚ3span {e} ⊆ ℝ3 to the dense set {w} × S ⊆ span U. Because U ∖ {0} ⊆ span U is nonempty and open, there must be some ∈ (U ∖ {0}) ∩ ({w} × S). By construction, there is uS such that = w × u.

Now, v := ∥–1 ∈ 𝕊2 is an inner normal of e, i.e. e = F(P, v), and is orthogonal to uS = ℚ3span {e}. □

In the following, we write πuL for the orthogonal projection of L ∈ 𝓚3 to u for a vector u ∈ ℝ3 ∖ {0}. Moreover, S1 (M, ⋅) denotes the first area measure of a convex body Mu with respect to u as the ambient space.

Lemma A.12

There is a set 𝓜 ⊆ 𝓟3 of full μ-measure that satisfies the following property: If some P ∈ 𝓜 has an edge with direction v ∈ 𝕊2, then one of the Pi also has an edge in direction v.

Proof

We intend to use Lemma A.11. If u ∈ ℝ3 ∖ {0}, then

πuK=i=1πuPi,

and hence the weak continuity and the Minkowski linearity of the area measures imply that

S1(πuK,)=i=1S1(πuPi,).

So S1 (πu K, ⋅) is a discrete Borel measure (i.e., has countable support) in u ∩ 𝕊2. Denoting by

ωu:={vuS2:S1(πuK,{v})>0}

the set of its atoms, we obtain from special cases of [12, Theorem 2.23, Lemma 3.4] that

0=S1(πuK,ωu)=S1(πuP,ωu)μ(dP).

So the set

M1:=uQ3{0}{PP3:S1(πuP,ωu)=0}

has full μ-measure.

Since for each u ∈ ℚ3 ∖ {0} the set ωu is countable, the set of pairs

C:={(u,v)(Q3{0})×S2:vωu}

is countable. Using the notation from Remark 2.18, we define

M2:=(u,v)CF(,v)1(suppFv(μ)).

The set 𝓜2 has full μ-measure. To see this, it is sufficient to consider v ∈ 𝕊2 and P ∈ 𝓟3 with F(P, v) ∉ supp Fv(μ) and to show that P ∉ supp μ. By assumption, there is a neighbourhood U′ of F(P, v) with Fv(μ)(U′) = 0, hence μ({P′ : F(P′, v) ∈ U′}) = 0. Since {P′ : F(P′, v) ∈ U′} is a neighbourhood of P, it follows that P ∉ supp μ.

Furthermore, for all P ∈ 𝓜2 and (u, v) ∈ C, Lemma A.10 shows that F(P, v) is a scaled summand of F(K, v).

Now let 𝓜 := 𝓜1 ∩ 𝓜2. Assume P ∈ 𝓜 and that e is an edge of P. By Lemma A.11, there are u ∈ ℚ3span e and vu ∩ 𝕊2 such that e = F(P, v). In particular, F(πu P, v) is nontrivial and S1 (πu P, {v}) > 0. Since P ∈ 𝓜1, it follows that vωu and so (u, v) ∈ C. Since P ∈ 𝓜2, this implies that e = F(P, v) is a scaled summand of the nontrivial support set F(K, v), which is then either an edge or a parallelogram. If it is an edge, it has the same direction as e and also is an edge of one of the Pi and we are done. If it is a parallelogram, then one of the sides of the parallelogram must have the same orientation as e, and one of the Pi has an edge with this orientation. This concludes the proof. □

Lemma A.13

There is a set 𝓜′ of full μ-measure such that for each P ∈ 𝓜′ and for each u ∈ 𝕊2, F(P, u) is a scaled summand of F(K, u).

Proof

  1. Let P ∈ 𝓜, where 𝓜 is as in the statement of Lemma A.12. If F(P, u) is trivial, so is the claim. If F(P, u) is an edge, then by Lemma A.12, there is i ∈ ℕ such that F(P, u) is homothetic to the edge F(Pi, u) and hence a scaled summand of F(K, u).

  2. Now consider the case that P ∈ 𝓜 and F(P, u) is a facet. The edges of the polytopes Pi, i ∈ ℕ, together have only countably many directions. Denote the countable set of these directions by A ⊂ 𝕊2. The facet F(P, u) is incident to (at least) two edges with linearly independent directions v, wA that determine the facet normal u up to sign σ ∈ {−1, 1} via

    ϕ:{1,1}×{(a,b)A2:a±b}R3,(σ,u,v)σu×vu×v.

    So the facet normals of P are contained in the countable image of ϕ, which is independent of the choice of P.

    For each u ∈ 𝕊2, the set F(K, u) is a polytope. Consider the set of full μ-measure

    M3:=uimϕF(,u)1(suppFu(μ)).

    If P is also in 𝓜3, then by Lemma A.10 and u ∈ imϕ, the support set F(P, u) is a scaled summand of F(K, u).

    Now the assertion follows from (I) and (II) with 𝓜′ := 𝓜 ∩ 𝓜3. □

A.4 Recoverability of generating polytopes

Again, let (Pi) be an admissible sequence and K its associated macroid together with an arbitrary generating measure μ. We call the Pi, i ∈ ℕ, the generating polytopes of K.

Then main work in this section is devoted to showing that any generating measure μ is supported in the set of translates of finite positive Minkowski combinations of the generating polytopes Pi, i ∈ ℕ (see Lemma A.16). As a consequence, we obtain in Lemma A.17 that every generating polytope Pi of K is a scaled summand of some element of supp μ. The conclusion that K is not a polyoid then follows almost immediately. We begin with a lemma that is independent of μ.

Lemma A.14

Let e be an edge of F(K, u) for some u ∈ 𝕊2. Then there is a unique i ∈ ℕ such that F(Pi, u) has an edge homothetic to e, e is in fact a translate of F(Pi, u), and this edge is unique among the edges of Pi.

Proof

The uniqueness statements immediately follow from the properties of admissible sequences (Pi). Note that an edge of F(K, u) need not be an edge of K as defined here.

If F(K, u) is a singleton, it does not have any edges.

If F(K, u) is an edge, then there is i ∈ ℕ such that F(Pi, u) is a translate of F(K, u).

If F(K, u) is a triangle, then there is i ∈ ℕ such that F(Pi, u) is a translate of F(K, u). So a translate of e is an edge of F(Pi, u).

If F(K, u) is a parallelogram, then there are unique i, j ∈ ℕ with ij such that F(Pi, u) is a translate of an edge of Pi, F(Pj, u) is a translate of an edge of Pj and F(Pi, u) + F(Pj, u) is a translate of F(K, u). So e is either a translate of F(Pi, u) or of F(Pj, u). □

Lemma A.15

The polytopes with a nontrivial zonotope summand are contained in a μ-zero set 𝓝.

Proof

Recall the measurable function ζ from Lemmas A.2 and A.4. The macroid Z generated by ζ(μ), the image measure of μ under the map ζ, is a zonoid and summand of K, implying

S2(Z,)S2(K,)=n=1m=1S(Pn,Pm,).

Because the right-hand side is a discrete measure, so is S2(Z, ⋅). If we can show that Z has no facets, then S2(Z, ⋅) is a discrete measure without atoms, hence zero. Then Z is at most one-dimensional. If we can also show that Z has no edges, then Z must be trivial, and the set 𝓝 of polytopes P with nontrivial ζ(P) is a μ-zero set, proving the claim.

It remains to show that Z has no facets or edges. We aim at a contradiction and assume that F(Z, u) is a facet or an edge.

Since F(Z, u) is a summand of the polytope F(K, u), it has an edge e that is homothetic to an edge of F(K, u). Because Z is centrally symmetric around the origin, F(Z, -u) = -F(Z, u) also has an edge that is a translate of e, and therefore F(K, − u) has an edge that is homothetic to e. Hence, F(K, u) and F(K, − u) both contain an edge homothetic to e. By Lemma A.14, and especially the uniqueness statement, there is i ∈ ℕ such that F(Pi, u) and F(Pi, − u) intersect in the very same edge homothetic to e. But this contradicts Pi being 3-dimensional, and Z cannot have edges or facets. This completes the proof. □

Lemma A.16

The generating measure μ is supported in translates of pos ((Pi)i), the set of translates of finite positive (Minkowski) combinations of polytopes from the sequence (Pi)i∈ℕ.

Proof

By Lemmas A.13 and A.15, μ is supported in the polytopes P that have no nontrivial zonotope summand and such that for all u ∈ 𝕊2, the support set F(P, u) is a scaled summand of F(K, u). Let 𝓜4 be a set of such polytopes of full μ-measure, and let P ∈ 𝓜4. For the proof, we may assume that P is nontrivial.

The strategy of proof is to show that there exist nonnegative numbers αi(P), for i ∈ ℕ, at most finitely many of which are non-zero, such that := i=1 αi(P)Pi has the following property: All facets of P and are triangles or parallelograms, the triangular facets of P are translates of the triangular facets of , and vice versa, and a similar fact holds for the facets that are parallelograms. Once this is proved, it follows that P and are translates of each other.

  1. All facets of K are triangles and parallelograms. The only scaled summands of a triangle are homothets of that triangle; the only scaled summands of a parallelogram are (possibly degenerate) parallelograms with the same edge directions but possibly different proportions. So all facets of P are of this kind.

  2. Let u ∈ 𝕊2 be such that F(P, u) is a triangular facet. Then F(K, u) is homothetic to F(P, u), that is, there are unique αu > 0 and tu ∈ ℝn such that F(P, u) = αu F(K, u) + tu. Moreover, there are unique i = i(u) ∈ ℕ and tu ∈ ℝn such that F(K, u) is a translate of F(Pi, u) and F(P, u) = αu F(Pi, u) + tu .

    Also note that there are at most two triangular facets of P that have an edge parallel to a fixed direction; otherwise, there would be an i ∈ ℕ such that Pi also had more than two such facets, contradicting the hypothesis that Pi has at most one edge parallel to a given direction.

  3. We observe that dim P = 3. Recall that P is nontrivial. If dim P = 1, then P is a segment, which is a zonotope, a contradiction. If dim P = 2, then P is a triangle or a non-degenerate parallelogram, which is a zonotope. The latter is excluded. Hence P is a triangle with P = F(P, v) = F(P, − v) for a unit vector v. Let e be an edge of P. Then Pi(v) and Pi(−v) both contain an edge parallel to e. Hence, i := i(v) = i(v) and F(Pi, v) = F(Pi, − v) and thus dim Pi = 2, a contradiction.

  4. Let G be the graph with the edges of P as G-vertices, where two edges, i.e. G-vertices, are connected if and only if they are opposite edges in a parallelogram facet of P. Since every edge is only part of two facets, the maximum degree of a G-vertex, i.e. an edge of P, is two. The connected components of G are cycles or chains.

    Let us first make sure that no cycles can occur. Assume that the edge e of P with direction u is part of a cycle. Then πu P is a convex polygon and πu e is one of its vertices. The two edges incident to πu e are projections of parallelograms that connect e to the two neighbors of e in G, and an induction shows that all support sets of πu P either are projections of edges parallel to e or parallelograms connecting two such edges. For the sake of applying [22, Theorem 3.2.22], let F(e, v) be an edge of e. In this case, vu, and so e is a summand of F(P, v). Then [22, Theorem 3.2.22] guarantees that e is a summand of P, in contradiction to P having no nontrivial zonotope summand. Therefore, the connected component of any edge e of P is a chain e1 − … − ek of e-translates.

    The endpoints of this chain must be edges of two triangular facets of P. By (ii), there can be no other chain with edges parallel to e1, …, ek. So if f is an edge parallel to e, then f = ej for some j ∈ [k] and f is a translate of e. Moreover, for any edge e of P there are exactly two triangular facets of P with an edge parallel to e.

  5. Let u, v ∈ 𝕊2, and i := i(u) as in (ii), such that F(P, u) is a triangle and F(Pi, v) is a facet adjacent to F(Pi, u) via an edge e. By (iii), there is exactly one w ∈ 𝕊2 besides u such that F(P, w) is a triangle with an edge parallel to e. By (ii), F(Pi, w) ≠ F(Pi, u) is then also a triangle with an edge parallel to e. Because Pi contains no other edge parallel to e, it follows that v = w. So we have

    F(P,u)=αuF(Pi,u)+tu,F(P,v)=αvF(Pi,v)+tv.

    By (iii), all edges of P parallel to e are translates of each other, hence it follows that αu = αv.

  6. Let u, v ∈ 𝕊2, and i := i(u) as in (ii), such that F(P, u) and F(Pi, v) are triangles. Then the triangles F(Pi, u) and F(Pi, v) are connected via a chain of neighboring facets. Iteration of (v) shows that F(P, v) is a triangle and αu = αv. So αu only depends on P and i(u), and we set αi(u)}(P) := αu > 0. If i ∈ ℕ and P contains no triangular facet F(P, w) with i(w) = i, then we set αi(P) := 0.

    For each i ∈ ℕ, there are uncountably many edge normals of Pi but only countably many facet normals of K. Let u ∈ 𝕊2 be such that F(Pi, u) is an edge and F(K, u) is not a facet and in fact a translate of F(Pi, u). Then for each P ∈ 𝓜4, F(P, u) is a summand of F(K, u) and satisfies V(F(P, u)) = αi(P) V(F(Pi, u)). This shows that 𝓜4Pαi(P) is measurable,

    V(F(K,u))=M4V(P,u)μ(dP)=M4αi(P)μ(dP)V(F(Pi,u))

    and thus we get

    M4αi(P)μ(dP)=1. (16)
  7. Let := i=1 αi(P)Pi, involving only finitely many nonzero summands. Note that dim = 3, since αi(P) > 0 for some i ∈ ℕ. Every facet of P or is either triangular or a parallelogram. The preceding items show that the triangular facets of P are translates of the triangular facets of , and vice versa.

  8. It remains to consider the parallelogram facets. Let u ∈ 𝕊2. If F(K, u) is not a parallelogram, it is a singleton, an edge or a triangle. For all P ∈ 𝓜4, neither F(P, u) nor F(, u) is then a parallelogram.

  9. We consider the situation from (viii). From now on, we assume that F(K, u) is a parallelogram. We choose v, w ∈ 𝕊2u such that the edges of F(K, u) are F(F(K, u), ± v) and F(F(K, u), ± w). There are unique distinct i, j ∈ ℕ such that F(F(K, u), ± v) are translates of F(Pi, u) and F(F(K, u), ± w) are translates of F(Pj, u). Let P ∈ 𝓜4. Then F(, u) is a translate of

    αi(P)F(Pi,u)+αj(P)F(Pj,u),

    which might be a singleton, an edge or a parallelogram. On the other hand, F(P, u) is a translate of

    F(F(P,u),v)+F(F(P,u),w).
  10. We consider the situation from (ix) and aim to show that F(P, u) and F(, u) are translates of each other. In the current item, we show that the conclusion holds at least if P is taken from a subset of 𝓜4 of full measure. The argument will be completed in (xi).

    Let P ∈ 𝓜4. If F(F(P, u), v) is not a singleton, then it is an edge parallel to F(Pi, u). Hence it must be a translate of αi(P)F(Pi, u). In either case,

    V(F(F(P,u),v))αi(P)V(F(Pi,u)). (17)

    Relation (17) can be used to bound the integrand in

    V(F(Pi,u))=V(F(F(K,u),v))=M4V(F(F(P,u),v))μ(dP).

    But then (16) from (vi) implies that equality must hold in (17), for μ-almost all polytopes P ∈ 𝓜4. Hence, F(F(P, u), v) is a translate of αi(P) F(Pi, u), and a similar argument shows that F(F(P, u), w) is a translate of αj(P) F(Pj, u), for μ-almost all P ∈ 𝓜4. So there is a measurable set 𝓜5(u) ⊆ 𝓜4 of full μ-measure such that for all P ∈ 𝓜5(u), a translate of F(P, u) is

    αi(P)F(Pi,u)+αj(P)F(Pj,u),

    and hence also a translate of F(, u).

  11. Finally, set 𝓜5 := ⋂u 𝓜5(u), where we take the countable intersection over all normals of parallelogram facet normals of K. Then 𝓜5 is a measurable set of full μ-measure and for all P ∈ 𝓜5 and u ∈ 𝕊2, F(P, u) is a parallelogram if and only if F(, u) is, and in this case both are translates of each other: When F(K, u) is a parallelogram, it follows from (x) that F(P, u) and F(, u) are translates, and if it is not, neither of them is a parallelogram due to (viii).

The proof is concluded by an application of Minkowski’s uniqueness theorem for area measures of convex polytopes. □

Lemma A.17

Let (Pi) be an admissible sequence and K its associated body together with a macroid-generating measure μ. Then for all i ∈ ℕ there is P ∈ supp μ such that Pi is a scaled summand of P.

Proof

Let u be the normal of a (necessarily triangular) facet of Pi. Then F(K, u) is a translate of this triangular facet. Clearly, μ is concentrated on supp μ and, according to Lemma A.16, on the set of translates of all finite positive combinations of the Pj, j ∈ ℕ. By (10) we have

hF(K,u)=hF(P,u)μ(dP),

hence there is P ∈ pos((Pj)j) ∩ supp μ such that F(P, u) is nontrivial. This can only be the case if Pi is a scaled summand in the finite positive combination defining P, as F(Pj, u) is trivial for all ji. □

Theorem A.18

Let (Pi) be an admissible sequence and K its associated body. Then the macroid K is not a polyoid.

Proof

Assume for contradiction that K is a k-polyoid with a generating measure μ supported in the space of k-topes. Lemma A.17 shows that Pk+1 is a scaled summand of some P ∈ supp μ. But then P is not a k-tope (see, e.g., [6, Lemma 2.3]), in contradiction to the property supp μ Pk3 of μ. So K is not a k-polyoid for any k ∈ ℕ. □

Remark A.19

Recall that Example A.7 shows that such K really exists.

Received: 2023-10-19
Revised: 2024-06-30
Published Online: 2025-01-23
Published in Print: 2025-01-29

© 2025 Walter de Gruyter GmbH, Berlin/Boston

Downloaded on 8.10.2025 from https://www.degruyterbrill.com/document/doi/10.1515/advgeom-2024-0030/html
Scroll to top button