I. McMullen’s decomposition and the Steiner formula The Steiner formula describes how the volume of a convex body behaves when the convex body is enlarged in every direction by a given radius r. In particular, the dependence on r is polynomial. It was first described by Jakob Steiner [5] in 1840 for dimensions 2 and 3. The Steiner formula can be deduced easily from a much more general result on the homogeneous decomposition of translation invariant valuations, i.e. maps that behave similar to volume and surface . The existence of such a decomposition was conjectured by McMullen and then independently proved by McMullen [1], by Meier [2] and by Spiegel [4].

Remarks: • Spiegel’s proof is probably the easiest to understand and uses mostly discrete , see also Schneider’s book [3]. • Since this topic will introduce mixed volumes, it should be one of the first in the seminar. References: [1] P. McMullen, Valuations and Euler-type relations on certain classes of convex polytopes, Proc. London Math. Soc. (3) 35 (1977), no. 1, 113–135. [2] Christoph Meier, Multilinearität bei Polyederaddition, Arch. Math. (Basel) 29 (1977), no. 2, 210–217 (German). [3] Rolf Schneider, Convex bodies: the Brunn-Minkowski theory, Second expanded edition, Encyclo- pedia of Mathematics and its Applications, vol. 151, Cambridge University Press, Cambridge, 2014. [4] Wolfgang Spiegel, Ein Beitrag über additive, translationsinvariante, stetige Eikörperfunktionale, Geometriae Dedicata 7 (1978), no. 1, 9–19 (German). [5] Jakob Steiner, Über parallele Flächen, Bericht über die zur Bekanntmachung geeigneten Ver- handlungen der Königlich Preußischen Akademie der Wissenschaften zu Berlin (1840), 114– 118, available at http://bibliothek.bbaw.de/bbaw/bibliothek-digital/digitalequellen/ schriften/anzeige/index_html?band=08-verh/1840&seite:int=114 (German).

II. Spherical harmonics and the cosine transform The classical theory of studies periodic functions on R or, equivalently, functions on S1. Spherical harmonics can be thought of as a generalization of this theory; it is used to study functions on Sn−1. The cosine transform of a signed Borel ρ on the Sn−1 is defined by Z Cρ = |u · v| dρ. Sn−1 Since this transform is closely related to the projection body and (generalized) zonoids, i.e. Minkowski sums of line segments, it has many uses in convex geometry.

1 For an integrable function f on the sphere, Cf is defined by considering f as a density. More precisely, Z Cf = |u · v|f(u) du. Sn−1 ∞ ∞ One important property of the cosine transform is that C : Ce → Ce is a con- ∞ tinuous bijection, where Ce is the set of infinitely differentiable even functions on the sphere. This fact can be proved using the Funk-Hecke theorem on spherical harmonics. As a consequence, generalized zonoids are dense in the set of all convex bodies.

Remarks: • The books of Gardner [1] and Schneider [3] contain good short introductions to the theory of spherical harmonics and also feature applications to convex geometry. • An in depth study of spherical harmonics can be found in Groemer’s book [2]. References: [1] Richard J. Gardner, Geometric tomography, 2nd ed. Encyclopedia of Mathematics and its Applications, vol. 58, Cambridge University Press, Cambridge, 2006. [2] H. Groemer, Geometric applications of Fourier series and spherical harmonics, Encyclopedia of Mathematics and its Applications, vol. 61, Cambridge University Press, Cambridge, 1996. [3] Rolf Schneider, Convex bodies: the Brunn-Minkowski theory, Second expanded edition, Encyclo- pedia of Mathematics and its Applications, vol. 151, Cambridge University Press, Cambridge, 2014.

III. The volume characterization theorem Given a map φ that assigns to each convex body a real number, what natural geometric conditions does one have to impose on φ to make sure that it is (a multiple of) volume? As was proved by Klain [2] in 1995, the following list is sufficient: • φ(K ∪ L) + φ(K ∩ L) = φ(K) + φ(L) whenever K, L and K ∪ L are convex bodies, i.e. φ is a valuation • φ is translation invariant • φ(−K) = φ(K) for all convex bodies K, i.e. φ is even • φ vanishes on lower dimensional convex bodies • φ is continuous in the Hausdorff Furthermore, this list is as short as possible in the sense that there are additional maps if we omit any one of these conditions. As a simple consequence, one recovers a famous theorem by Hadwiger [1] from 1957 which classifies all continuous rigid motion invariant valuations. Combining Klain’s result with the McMullen decomposition, one obtains a structure theory for even continuous translation invariant valuations, see Klain [3].

2 Remarks: • The books of Klain and Rota [4] and of Schneider [5] are also good resources. • The proof of the volume characterization theorem consists mostly of dissec- tion theory, but critically uses a result from harmonic analysis on the cosine transform at one point. References: [1] , Vorlesungen über Inhalt, Oberfläche und Isoperimetrie, Springer-Verlag, Berlin-Göttingen-Heidelberg, 1957 (German). [2] Daniel A. Klain, A short proof of Hadwiger’s characterization theorem, Mathematika 42 (1995), no. 2, 329–339. [3] Daniel A. Klain, Even valuations on convex bodies, Trans. Amer. Math. Soc. 352 (2000), no. 1, 71–93. [4] Daniel A. Klain and Gian-Carlo Rota, Introduction to geometric probability, Lezioni Lincee. [Lincei Lectures], Cambridge University Press, Cambridge, 1997. [5] Rolf Schneider, Convex bodies: the Brunn-Minkowski theory, Second expanded edition, Encyclo- pedia of Mathematics and its Applications, vol. 151, Cambridge University Press, Cambridge, 2014.

IV. Weyl’s tube formula Similarly to the Steiner Formula, Weyl’s tube formula [2] describes how to compute the volume of a (small) tube of radius r around an embedded submanifold M of n R . Again, the expression is a polynomial in r whose coefficients depend on M. Remarkably, these coefficients turn out to be intrinsic, i.e. they depend only on the (induced) Riemannian metric and not on the particular embedding.

Remarks: • There is a whole book on this formula and its generalizations by Gray [1]. • Finding the formula for the volume of a tube is a geometric calculation that does not hold many surprises. The hard part is to express the coefficients in terms of the Riemannian curvature tensor, for which Weyl uses his theory of invariants [3]. References: [1] Alfred Gray, Tubes, 2nd ed. Progress in Mathematics, vol. 221, Birkhäuser Verlag, Basel, 2004. With a preface by Vicente Miquel. [2] Hermann Weyl, On the Volume of Tubes, Amer. J. Math. 61 (1939), no. 2, 461–472. [3] Hermann Weyl, The Classical Groups. Their Invariants and Representations, Princeton Uni- versity Press, Princeton, N.J. 1939.

V. The Shephard Problem Given two n-dimensional symmetric convex bodies K and L, does the following implication hold?

⊥ ⊥ n−1 voln−1(K|u ) ≤ voln−1(L|u ) ∀u ∈ S ⇒ voln(K) ≤ voln(L)

3 Here, K|u⊥ denotes the orthogonal projection of K onto u⊥. This question was first asked by Shephard [4] in 1964. In dimension n = 2, it is not hard to see that the left hand side implies K ⊆ L and therefore also the right hand side. However, the implication is false in higher dimensions. Counterexamples were given independently by Petty [2] and by Schneider [3]. On the other hand, if L is an element of a special subclass of convex bodies, the implication is true for all convex bodies K.

Remarks: • A good overview is contained in Gardner’s book [1]. • The construction of the counter example uses the cosine transform and requires some harmonic analysis. References: [1] Richard J. Gardner, Geometric tomography, 2nd ed. Encyclopedia of Mathematics and its Applications, vol. 58, Cambridge University Press, Cambridge, 2006. [2] C. M. Petty, Projection bodies, Proc. Colloquium on Convexity (Copenhagen, 1965), Koben- havns Univ. Mat. Inst., Copenhagen, 1967, pp. 234–241. [3] Rolf Schneider, Zu einem Problem von Shephard über die Projektionen konvexer Körper, Math. Z. 101 (1967), 71–82 (German). [4] G. C. Shephard, Shadow systems of convex sets, Israel J. Math. 2 (1964), 229–236.

VI. Equidissectability of polytopes 3 The volume of a pyramid in R can be calculated easily as the area of the base times the height divided by 3. However, is there an elementary way to derive this formula that does not rely on approximation arguments, but instead uses only cutting and pasting? In dimension 2 this is easily possible. A triangle can be cut into pieces (by straight lines) in such a way that gluing those pieces together (after moving them around) gives a . The same can be done for any polytope, i.e. the convex hull of finitely many points, in dimension 2. Gauß bickered that there was no known proof of this type for the volume of a pyramid, which led Hilbert to pose this question as the third problem in his now famous list [3]. Shortly after, Dehn gave a negative answer [2]. One can define maps, known as Dehn invariants ∆f , 3 on polytopes in R that are invariant under cutting and pasting operations. Since for a regular pyramid there is a Dehn invariant which takes a different value for a cube of the same volume, one body cannot be build from the other by cutting and pasting. Much later it was shown by Sydler [4] that Dehn’s conditions for equidissectability are not only necessary but also sufficient.

Remarks: • Boltianski˘ıwrote a good book [1] on Hilbert’s third problem. • The proof of Dehn’s result uses only elementary discrete geometry. Sydler’s result on the other hand is obtained using more algebraic techniques and is quite sophisticated.

4 • Solving the problem of equidissectability in dimensions ≥ 5 is still an open problem. References: [1] Vladimir G. Boltianski˘ı, Hilbert’s third problem, V. H. Winston & Sons, Washington, D.C.; Halsted Press [John Wiley & Sons], New York-Toronto-London, 1978. Translated from the Russian by Richard A. Silverman; With a foreword by Albert B. J. Novikoff; Scripta Series in Mathematics. [2] M. Dehn, Ueber den Rauminhalt, Math. Ann. 55 (1901), no. 3, 465–478 (German). [3] David Hilbert, Mathematische Probleme, Nachrichten von der Gesellschaft der Wissenschaften zu Göttingen, Mathematisch-Physikalische Klasse (1900), 253–297, available at http://www. digizeitschriften.de/dms/resolveppn/?PPN=GDZPPN002498863 (German). [4] J.-P. Sydler, Conditions nécessaires et suffisantes pour l’équivalence des polyèdres de l’espace euclidien à trois dimensions, Comment. Math. Helv. 40 (1965), 43–80 (French). VII. The Brunn-Minkowski Given two convex bodies K and L, the Brunn-Minkowski inequality asserts that 1/n 1/n 1/n voln(K + L) ≥ voln(K) + voln(L) . Here, K + L := {x + y : x ∈ K, y ∈ L}. For n-dimensional convex bodies, equality holds if and only if K and L are homothetic. For dimensions ≤ 3 this result was first proved by Brunn in the late 19th century. Then, in the early 20th century, Minkowski gave a proof for all dimensions. The isoperimetric inequality and other delicate results are immediate consequences of the Brunn-Minkowski inequality. The Brunn-Minkowski inequality is not restricted to convex bodies, in fact it holds for all bounded (Borel) measurable sets K and L. It is also the geometric core of an analytic inequality, called the Prékopa-Leindler inequality, see e.g. [1].

Remarks: • A good overview and further references can be found in Schneider’s book [2], which contains three different proofs each using different techniques. References: [1] R. J. Gardner, The Brunn-Minkowski inequality, Bull. Amer. Math. Soc. (N.S.) 39 (2002), no. 3, 355–405. [2] Rolf Schneider, Convex bodies: the Brunn-Minkowski theory, Second expanded edition, Encyclo- pedia of Mathematics and its Applications, vol. 151, Cambridge University Press, Cambridge, 2014. VIII. The Petty projection inequality n For a convex body K in R , there exists a unique symmetric convex body ΠK such that the width of ΠK in direction u is equal to twice the area of K projected onto u⊥ for all u ∈ Sn−1. The convex body ΠK is called the projection body of K. Petty’s projection inequality [4], see also [2] and [3], says that  n ∗ n−1 κn voln(Π K) voln(K) ≤ . κn−1

5 n ∗ Here, κn denotes the n-dimensional volume of the unit ball in R and Π K the polar body of ΠK. For an n-dimensional convex body K, equality is attained if and only if K is an ellipsoid. This inequality can be regarded as a stronger version of the classical isoperimetric inequality. In fact, rewriting it gives the following chain of inequalities:

n+1 n+1 Z −n −1 n n−1 n κn  ⊥ n n κn · V (K) ≤ n · voln−1 K|u du ≤ S(K) κn−1 Sn−1 Moreover, the quantities in the Petty projection inequality are all SL(n) invariant, whereas the is only SO(n) invariant.

Remarks: • Gardner’s book [1] contains a good overview of this and related isoperimetric inequalities. • The proof uses a geometric technique called Steiner symmetrization. References: [1] Richard J. Gardner, Geometric tomography, 2nd ed. Encyclopedia of Mathematics and its Applications, vol. 58, Cambridge University Press, Cambridge, 2006. [2] Erwin Lutwak, Mixed projection inequalities, Trans. Amer. Math. Soc. 287 (1985), no. 1, 91– 105. [3] Erwin Lutwak, Selected affine isoperimetric inequalities, Handbook of convex geometry, Vol. A, B, North-Holland, Amsterdam, 1993, pp. 151–176. [4] C. M. Petty, Isoperimetric problems, Proceedings of the Conference on Convexity and Com- binatorial Geometry (Univ. Oklahoma, Norman, Okla. 1971), Dept. Math., Univ. Oklahoma, Norman, Okla. 1971, pp. 26–41. IX. The boundary of a convex body The boundary of convex bodies can show various different behaviours. For example, polytopes have a combinatorial structure associated with their boundaries. On the other hand, if a (full-dimensional) convex body has only regular boundary points, i.e. points with unique outer normal, it is actually a C1-manifold. Although the set of singular boundary points can lie dense, it always has (n − 1)-dimensional Hausdorff measure 0. Rather surprisingly, it is even possible to define principal curvatures almost everywhere on the boundary for all convex bodies.

Remarks: • This is a broad topic with very different theorems to choose from. • The books by Gruber [1] and Schneider [2] are good resources. References: [1] Peter M. Gruber, Convex and discrete geometry, Grundlehren der Mathematischen Wis- senschaften [Fundamental Principles of Mathematical Sciences], vol. 336, Springer, Berlin, 2007. [2] Rolf Schneider, Convex bodies: the Brunn-Minkowski theory, Second expanded edition, Encyclo- pedia of Mathematics and its Applications, vol. 151, Cambridge University Press, Cambridge, 2014.

6 X. The affine surface area The affine surface area is a notion from affine differential geometry. It can be defined for all convex bodies by Z 1/(n+1) n−1 Ω(K) = Hn−1(K, x) dH (x), bd K

where Hn−1(K, x) is the product of the principal curvatures of K at x, called Gauß-Kronecker curvature, which exists for almost all boundary points. Apart from behaving similarly to regular surface area, it has the remarkable property of being invariant under all volume preserving linear transformations. Furthermore it satisfies the so called affine isoperimetric inequality,

n+1 2 n+1 Ω(K) ≤ κnn Vn(K),

which was first proved by Lutwak [1].

Remarks: • Schneider writes about the long history of affine surface area and its various definitions in his book [2] References: [1] Erwin Lutwak, Extended affine surface area, Adv. Math. 85 (1991), no. 1, 39–68. [2] Rolf Schneider, Convex bodies: the Brunn-Minkowski theory, Second expanded edition, Encyclo- pedia of Mathematics and its Applications, vol. 151, Cambridge University Press, Cambridge, 2014.

7