Lattice Polytopes
Lattice Polytopes
2.4.3 Volumes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
2.5 Software . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
2.6 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
3 Ehrhart Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61
3.1 Motivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62
3.1.1 Examples of Ehrhart polynomials . . . . . . . . . . . . . . 63
3.2 Generating Functions for Lattice Points . . . . . . . . . . . . . . . 65
3.3 Ehrhart’s theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71
3.4 Stanley’s theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74
3.4.1 Half-Open Decompositions of Cones . . . . . . . . . . . . 74
3.4.2 The integer point generating function of half-open
cones . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77
3.4.3 Stanley’s theorem and the h∗ -polynomial of a
lattice polytope . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77
3.4.4 Where does the h∗ -notation come from? . . . . . . . . . 79
3.5 Reciprocity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80
3.5.1 Stanley reciprocity for cones . . . . . . . . . . . . . . . . . . . 81
3.5.2 Ehrhart-Macdonald reciprocity for lattice polytopes 83
3.6 Properties of the h∗ -polynomial . . . . . . . . . . . . . . . . . . . . . . 85
3.6.1 Degree and codegree of lattice polytopes . . . . . . . . 85
3.6.2 Ehrhart polynomials of lattice polygons . . . . . . . . . 88
3.6.3 Polytopes with Small Degree . . . . . . . . . . . . . . . . . . . 89
3.7 Brion’s theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89
3.8 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93
4 Geometry of Numbers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99
4.1 Minkowski’s Theorems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 100
4.2 Coverings and Packings . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 104
4.3 Flatness Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108
4.4 Finiteness of lattice polytopes with few interior lattice
points . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109
4.4.1 Finiteness of barycentric coordinates of lattice
simplices . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 110
4.4.2 Coefficient of asymmetry . . . . . . . . . . . . . . . . . . . . . . 113
4.4.3 Bounding the volume . . . . . . . . . . . . . . . . . . . . . . . . . 114
4.5 Lower Bounds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 115
4.6 Empty lattice simplices . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 117
4.7 Lattice polytopes without interior lattice points . . . . . . . . 122
4.8 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 126
Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 229
Before reading on, we would like to invite the reader to pull out a sheet
of graph paper and play with lattice polygons, making one’s first own
discoveries and building a feeling and an appreceation for the objects and
the questions treated in this book.
When we say lattice polygon, we mean a closed convex polygon with
all vertices at crossing points of your graph paper as in Figure 1.1 (and
Fig. 1.1: A lattice polygon not as in Figs. 1.2 and 1.3).
For such a polygon, we can record the number b of graph-paper-
crossing-points on the boundary, the number i of graph-paper-crossing-
points in the interior, and the enclosed area a measured in units of
graph-paper-squares. For example, our first polygon above has (b, i, a) =
(6, 5, 7).
Your task is now to play with these figures and find out what triples
(b, i, a) you can achieve. That is, if I say (3, 0, 1/2), you draw the picture in
Fig. 1.2: non-lattice Figure 1.4, and if I say (9, 1, 4.5), you draw the picture in Figure 1.5. The
reader is invited to try realizing the cases (5, 2, 4), (18, 0, 9), (3, 17, 17.5)
and (11, 2, 6.5).
In this section, we formally introduce lattice polygons, which are the key
player of this chapter, find a famous connection between its lattice points
Fig. 1.3: non-convex and its volume, and develop the appropriate notion of when to consider
two such polygons the same.
Definition 1.1 A lattice polygon is the convex hull in R2 of finitely
many points in Z2 .
Here, the word lattice refers to Z2 whose elements we call lattice points.
In other words, in dimension 2, lattice points are simply the points on the
grid given by the vectors with all coordinates integral. A lattice polygon
Fig. 1.4: (b, i, a) = (3, 0, 1/2) is the smallest convex set that contains a given finite set of lattice points.
Restricting the vertices to lattice points is quite restrictive. In particular,
a lattice polygon cannot be arbitrarily small. We bound the volume with
the next proposition. We will see that even more is true. Any polygon
with this volume will be a triangle, and any such triangles are equivalent
to each other in some sense we develop below.
Proposition 1.2 (Pick’s Theorem) Any lattice triangle with only
three lattice points (which must be its vertices) has area 1/2.
Fig. 1.5: (b, i, a) = (9, 1, 4.5)
This Proposition may seem unspectacular, at first. But it is remarkable
in several ways. First, the corresponding statement is plain wrong in
z = λv + µw = bλcv + bµcw + u
where
but also
u = z − bλcv − bµcw ∈ Z2 .
{λv + µw : λ, µ ∈ Z} = Z2 .
The goal of this section is to prove an elegant formula for the computation
of the area of a lattice polygon just by counting lattice points. We have
seen this miracle already for triangles (Pick’s Theorem (Proposition 1.2))
and we will now generalize it to arbitrary lattice polygons. We thus find
a first relation among the parameters b, i, a from Section 1.1.
Keep in mind that the Reeve simplices (1.1) exclude a straightforward
generalization to higher dimensions. To find some generalization is the
Fig. 1.8: Reeve’s Tetrahedron
topic of Chapter 3 on Ehrhart Theory.
Theorem 1.8 (Pick’s Formula) Let P be a lattice polygon with i in-
terior lattice points, b lattice points on the boundary, and (Euclidean)
area a. Then
b
a = i+ −1.
2
This shows, for instance, that the triple (5, 2, 4) from Section 1.1 can
never come from a lattice polygon.
i = i1 + i2 + ie , b = b1 + b2 − 2ie − 2 ,
so
1
a = a1 + a2 = i1 + i2 + (b1 + b2 ) − 2
2
1 b
= i − ie + (b + 2ie + 2) − 2 = i + − 1 .
2 2
If b = 3 and i ≥ 1 then we can split P into three pieces Q1 , Q2 , and Q3
by coning over some interior point of P . See Figure 1.9.
Again, all three pieces have fewer lattice points than P , so we know
Pick’s Formula for those by our induction hypothesis. A similar compu-
tation as the one above shows that Pick’s Formula also holds for P (see
Exercise 1.10). u
t
The reader is invited to prove that the same relation is true for
non-convex lattice polygons in Exercise 1.13. One can also prove that this
theorem is equivalent to the Euler relation (see Exercise 1.11). Note that
the same induction shows that any lattice polygon can be subdivided into Fig. 1.10: Not allowed in a triangula-
tion
triangles which are isomorphic to ∆2 (called unimodular triangles). As
the Reeve simplex shows, this is not true in higher dimensions. Questions Exercise 1.10
about the existence of such unimodular triangulations will be discussed
in the last Chapter 8 of this book. See also Exercise 1.12. Exercise 1.11
The idea of subdiving a convex object into triangles, or simplices Exercise 1.12
in dimensions 3 and above, is quite influential. We will devote a whole
chapter to this (Chapter 8).
Definition 1.9 (Triangulation) Let P be a lattice polygon. A (lattice)
triangulation of P is a collection S of lattice triangles such that
(1) Any two triangles ∆d , simplex0 ∈ S intersect in an edge of both, and
(2) the union (as point sets) of all triangles is P .
Figure 1.10 shows two configurations of triangles that are not allowed in
a triangulation. In the first, the intersection is not a full edhe in both, in
the second, the triangles intersect in more than an edge. A proper lattice
triangulation is shown in Figure 1.11.
In fact, Pick’s Formula (Theorem 1.8) also holds in a non-convex
setting. We can generalize it to any (non-convex) lattice polygon with
the property, that any vertex is incident to exactly two edges and no
Exercise 1.13 edges intersect. You will prove this in Exercise 1.13.
q0 = 3 2 ≤ p ≤ p0 . (1.2)
Fig. 1.13: The bot- The polygon P intersects the bottom and top edge of the rectangle in
tom and top edge of P
edges of length q and q 0 , see Figure 1.12.
At most 2(p − 1) boundary lattice points of P are not on the two
horizontal edges of R, so
b ≤ q + 1 + p − 1 + q 0 + 1 + p − 1 = q + q 0 + 2p. (1.3)
Subdividing the convex hull of the top edge and the bottom edge into
Fig. 1.14: Bounding the volume
two triangles as in Figure 1.14 we get
1 1 p
a ≥ pq + q 0 p = (q + q 0 ). (1.4)
2 2 2
a ≥ b/2 . (1.5)
We have p ≥ 4, so that the left hand side is bounded by 16. This shows
b ≤ a + 4. u
t
Note that for polygons that are not lattice polygons, there is no such
upper bound on their areas. Figure 1.17 shows why.
Scott’s theorem defines a polyhedral set L given by the inequalities
i ≥ 1 a ≥ 3/2 a ≤ 2(i − 1)
such that any pair (a, i) coming from a lattice polygon is either (a, i) =
Fig. 1.17: An arbitrarily big rational (9/2, 1) or inside L. What about the converse? Is every point
triangle with one interior lattice point
1
(a, i) ∈ Z×Z∩L
2
the volume an number of interior lattice points of some lattice polygon?
This is easier to answer when we move from the pair (a, i) used in Scott’s
Theorem to the pair (b, i) using Pick’s Theorem. With this transformation
our inequalities read
i ≥ 1 b ≥ 3 b ≤ 2(i + 3) .
We can indeed realize all of these pairs of integers as the number of lattice
points in the interior and the boundary of a lattice polygon. Figure 1.18
Exercise 1.15 shows the construction.
Exercise 1.16
Exercise 1.17
— 14 — Haase, Nill, Paffenholz: Lattice Polytopes
Chapter 1. An invitation to lattice polytopes (draft of June 28, 2021)
1.5 Dilations
1.6 Problems
included on page 8
1.4.
included on pa
1.5.
included on pa
included on page 9
1.9. Show that Rd (m) are d-dimensional simplices of volume m/d! with
d + 1 lattice points.
Hint: projection map.
included on page 11
1.10. Finish the induction in the proof of Pick’s Formula (Theorem 1.8)
for the missing case b = 3 and i ≥ 1.
included on page 11
1.11. Euler’s Formula states that a finite planar graph with v nodes, e
edges and f bounded faces satisfies
v − e + f = 1
1.12.
included on page 12
1.14. Show that a polygon with volume 9/2, one interior lattice points, 9
boundary lattice points, and whose vertices are on parallel lines at
distance 3 must be the simplex 3 ∆2 .
1.15. Describe as precisely as you can which pairs (b, i) can be realized
for lattice polygons.
Hint: Consider long and flat quadrilaterals in which the interior
points are lined up on a straight line. We will study this in
detail in Chapter 3
included on page 14
1.17. Following up on Exercise 1.16 one can plot (b(P ), i(P )) for all
lattice triangles P in a given range, see Figure 1.20. Prove that the
region at the bottom, denoted by σ1◦ , is given by Scott’s theorem
(the special point (9, 1) is not visible in the very dense plot). Can
you also describe the other prominent regions σi◦ (for i ≥ 1)?
included on page 15
***1.18.
2000
σ6o
1500
σ5o
i (interior points)
σ4o
1000
σ3o
500 σ2o
σ1o
0
0 500 1000 1500 2000
b (boundary points)
You have seen in the previous chapter how polygons and integer points
interact nicely and produce some nice and useful classification results.
We will show that all these results can, with appropriate modifications,
Lecture Notes Lattice Polytopes (draft of June 28, 2021)
2.1 Polyhedra
Polyhedral cones are the intersection of a finite set of linear half spaces.
Generalizing to intersections of affine half spaces leads to polyhedra. We
are mainly interested in the subset of bounded polyhedra, the polytopes.
Specializing further, we will deal with integral polytopes.
In the second part of this chapter we link integral polytopes to
lattices, which are discrete subgroups of the additive group Rd . This gives
a connection to commutative algebra by interpreting a point v ∈ Zd as
the exponent vector of a monomial in d variables.
where λx = 0 for all but finitely x ∈ X. The linear hull or linear span
lin(X ) of X is the set of all linear combinations of X,
( )
X λx ∈ R
lin(X ) := λx x : .
and λx = 0 for all but finitely many x
x∈X
The linear span of X is the smallest linear space containing X and the
common intersection of all linear spaces containing X. Similarly, the affine
hull of X is the smallest affine space containing X and the intersection
of all affine spaces containing X. For a matrix A ∈ Rd×n with column
vectors a1 , . . . , an we also write
Y − t := {y − t : y ∈ Y }
For any affine space A = aff X we can consider its translation by a vector
x ∈ A. This is a linear space. The dimension of A is the dimension of
A − x. Hence, any point in the affine hull of X can be written as an affine
combination of at most d + 1 points in X.
Definition 2.2 A linear combination is conic if all coefficients are non-
negative, and it is convex if it is conic and affine. The set of all conic
combinations of a set X is the cone over X, denoted by cone(X ). The
set of all convex combinations of X is the convex hull conv(X ). X is a
cone if X = cone(X ) and X is a convex set if X = conv(X ).
A polyhedral cone is the cone of a finite subset of Rd . A polytope is
the convex hull of finitely many points in Rd .
The dimension of a cone is the dimension of its linear span. The
dimension of a polytope is the dimension of the affine space it spans.
See Figure 2.1 for an example. Again, we sometimes write cone(A) and
conv(A) for the conic or convex hull of the set of column vectors of a
matrix A ∈ Rd×n . We are mostly interested in cones and convex sets
defined by a finite set X. Clearly, if the dimension of a polytope is less
than the dimension of the ambient space, then we can restrict to that
affine space. Hence, we may assume that the dimension of our polytopes
coincides with the dimension of the space (we will see later that it will
be useful to also consider lower dimensional polytopes, though).
Fig. 2.1: Cone (blue) and Polytope
For polytopes in dimension 2 we have already seen the polygons in (red) for the point set of the red
the previous chapter. Those are all 2-dimensional polytopes. points.
We need at least d + 1 affinely independent points in Rd to affinely
span Rd , so any full-dimensional polytope has at least d + 1 points in its
defining set. Any polytope defined by precisely d + 1 affinely independent
points is called simplex. Any two simplices can be identified via a bijective
depends on x. The followinf theorem, whose proof ist left as Exercise 2.1,
makes this precise.
Theorem 2.4 (Carathéodory’s Theorem) Let X ⊆ Rd , C = cone(X ),
and y ∈ C. Then there are x1 , x2 , . . . , xd ∈ X and λ1 , λ2 , . . . , λd ≥ 0 such
that
d
X
y = λ i xi .
i=1
H := {x | h a, x i ≤ β}
H − := {x | h a, x i ≤ β} .
P = { x | Ax ≤ b }
xi ≥ 0 xi ≤ 1 for 1 ≤ i ≤ d .
x1 ≥ 0 x1 − x2 ≤ 2 x1 + 3x2 ≤ 10 3x1 − x2 ≥ 0
3x1 − x2 ≥ 0
define a polygon with vertices
x1 − x2 ≤ 2 " # " # " # " #
0 2 4 1
x1 + 3x2 ≤ 10 .
0 0 2 3
x1 ≥ 0 see Figure 2.5.
Ha−i ,0
\
P = (2.1)
Fig. 2.5: The poly-
gon of Example 2.7(2) for some a1 , a2 , . . . , ak ∈ (Rd )? .
We have already defined a cone over a set X as the set of all conic
combinations in the previous section. We will see below that this and the
newly defined notion of a polyhedral cone coincide if X is a finite set,
i.e. any polyedral cone can equally be described as the cone over some
suitably chosen finite set X, and any cone over a finite set is polyhedral.
We will not encounter non-polyhedral cones, that is, cones defined
as the set of conic combinations over an infinite set X, in this book.
Therefore, we will often omit the word polyhedral and just speak of cones
in the text, and only stress this restriction in definitions and theorems.
The dimension of such a polyhedron defined by half spaces is again
defined as the dimension of its affine hull. We sometimes use the notion d-
polytope for a d-dimensional polyhedron. A polyhedron is full dimensional
if dim P = d.
Ha−i ,0
\ \
rec P = and lineal P = Hai ,0 .
Example 2.10
so if P := {x | h a1 , x i ≤ β1 , . . . , h am , x i ≤ βm } ⊆ Rd with ai ∈ (Rd )? ,
βi ∈ R for i ∈ [m] then
X + Y := {x + y | x ∈ X, y ∈ Y } .
3x1 − x2 ≥ 3 x1 ≥ 1 x1 + x2 ≥ 3 x1 − 3x2 ≤ −1 .
v2 The first is called the interior or V-description, The second is the exterior
r2 or H-description. Both are important in polytope theory, as some things
v3 are easy to describe in one and may be difficult to define in the other.
Observe that the empty set is also a face of P . For any face F we have
F ∩ P = aff F ∩ P ,
max{h a, x i | x ∈ P }
f (P ) := (f0 (P ), . . . , fd−1 (P )) ,
and its vertices are 1/bj aj for each facet defining inequality h aj , x i ≤ bj
of P . u
t
You will show in Exercise 2.7 that dualizing twice gives back the original
Exercise 2.7 polytope.
Corollary 2.17 Let P ⊆ Rd be a d-dimensional polytope with 0 ∈
int P . Then we have a bijective correspondence between k-faces of P
and (d − 1 − k )-faces of P ? for 0 ≤ k ≤ d − 1 and if a k-face F of P is
contained in a (k + 1)-face G of P , then the face corresponding to G in
P ? is contained in the face corresponding to F in P ? . u
t
so ϕP is again a polyhedron.
ϕP = Q ψQ = P .
Example 2.21
f (C) = (f−1 , f0 , . . . , fd )
χ ( C ) + χ ( C0 ) = χ ( C ∪ C0 ) − χ ( C ∩ C0 ) . (2.4)
Example 2.32
conv((wi , vi ) | 1 ≤ i ≤ m) ,
where the lower hull is the polyhedral complex of those facets whose normal
has negative first coordinate.
Given a set of points V := {v1 , . . . , vm } and a weight vector w ∈ Rm
we denote by Sw (V ) the regular subdivision obtained as the lower hull of
lift(w ) := conv((wi , vi ) | 1 ≤ i ≤ m).
You will show in Exercise 2.11 that all subdivisions of a polygon using only
Exercise 2.11 the vertices of the polygon are regular. WISHLIST: convex piecewise
linear function Ψw (x) = min{h : (h, x) ∈ lift(w )}
Definition 2.34 (polyhedral sphere, polyhedral ball)
(w1 , v1 ), . . . , (wd , vd )
2.3 Lattices
We introduce the central tool for this book. It will link our geometric
objects, the polytopes, to algebraic objects, namely toric ideals and toric
varieties.
Throughout this section, V will be a finite-dimensional real vector
space equipped with the topology induced by a norm k . k and with a
translation invariant volume form. Fig. 2.9: A non-regular subdivision
Lattices can be defined in two different (but equivalent) ways. On the one
hand as the integral generation of a linearly independent set of vectors,
on the other hand as a discrete abelian subgroup of the vector space.
We will start with the latter characterization of a lattice, This is often
very useful to describe lattices without the explicit choice of a basis. We
will deduce the other representation in a sequence of propositions that
introduce some interesting structure ofr lattices.
Recall that a subset Λ ⊆ V is an additive subgroup of V if
(1) 0∈Λ
(2) x+y ∈ Λ for any x, y ∈ Λ
(3) −x ∈ Λ for any x ∈ Λ .
Λ2;3 := x ∈ Z2 : x1 + x2 ≡ 0 mod 3
is a lattice.
(4) Let Λ be a lattice and L ⊂ V be a proper linear subspace of V . Then
Λ ∩ L is a lattice in L.
Exercise 2.17
For a set B = {b1 , . . . , bd } ⊂ V of linearly independent vectors we define
the subgroup
( d )
X
Λ(B) : = λi bi | λi ∈ Z, 1 ≤ i ≤ d
i=1
generated by B is a lattice.
For the proof we need some prerequisites. The following lemma is imme-
diate from the definition.
Lemma 2.47 If K ⊂ V is bounded, then K ∩ Λ is finite. u
t
In the situation of the proposition, we will often write Λ/U for π (Λ).
Proof. (1) π (Λ) is the image of a group under a homomorphism. Hence,
it is a subgroup of V /U . The hard part of the proposition is to prove
that π (Λ) is discrete in V /U .
The space U is Λ-rational. So we can choose a vector space basis
{v1 , . . . , vr } ⊂ Λ ∩ U of U . We can extend this basis to a vector space
on lin Λ and
d
! 0
X
λi vi +U := max ({|λi | : i = r + 1, . . . , d})
i=1
on lin Λ/U . Denote the unit ball of lin Λ by W . By Lemma 2.47, the
set W ∩ Λ is finite. Set
Exercise 2.19
That means, for the proof that the last two quantities agree, we can
choose bases as in Theorem 2.53. Then the change of bases matrix is
diagonal with determinant k1 · . . . · kr , while the set Π(B0 ) ∩ Λ consists
P
of the points i li bi for 0 ≤ li ≤ ki − 1. u
t
Exercise 2.22
In dimensions d ≥ 2 there are infinitely many unimodular matrices. Exercise 2.23
Hence, there are also infinitely many different bases of a lattice. In
Exercise 2.24
Section 6.4 we deal with the problem of finding bases of a lattice with
Exercise 2.25
some nice properties. We will e.g. construct bases with “short” vectors.
Exercise 2.26
Definition 2.55 (dual lattice) Let Λ ⊂ V be a lattice with lin Λ = V . Exercise 2.27
Then set Exercise 2.28
Exercise 2.29
Λ? := {α ∈ V ? | α(a) ∈ Z for all a ∈ Λ}
d(x, y ) := kx − yk
and d(x, S) := inf (d(x, z ))
z∈S
for any x, y ∈ Rd , S ⊆ Rd .
Br (Π) := {x | d(x, Π) ≤ r} .
minimal distance. We will show that these choices satisfy the requirements
of the proposition.
Let w ∈ Λ − V and y ∈ V . By definition of V there are coefficients
λ1 , . . . , λk ∈ R such that
k
X
y= λi vi .
i=1
Xk k
X
Set z := bλci vi , and z 0 := {λ}i vi ,
i=1 i=1
Then z, w − z ∈ Λ and z 0 = y − z ∈ Π. Further, w − z 6∈ V . Hence,
d(y, w ) = d(y − z, w − z ) ≥ d(w − z, Π) ≥ d(v, Π) = d(v, x) . u
t
We obtain a second proof that any lattice has a basis. For this, let
v1 , . . . , vd ∈ Λ be any linearly independent set in Λ. We consider the
chain of subspaces
L0 := {0} and Li := lin (v1 , . . . , vi )
L0 ( L1 ( . . . ( Ld .
By Lemma 2.57 we can find b1 ∈ L1 closest to 0. Let w be any other
lattice vector in L1 . Then there is λ such that w = λb1 , and
0 ≤ kw − bλcb1 k < 1 .
By our choice of b1 now λ ∈ Z.
Now assume by induction that we habe a lattice basis of Lk−1 . Choose
any bk ∈ Łk \ Lk−1 closest to Lk−1 via Lemma 2.57. Let w ∈ Λ ∩ Lk .
Then there are µi , ηi ∈ R for 1 ≤ i ≤ k such that
X X
bk := µ i vi and w := η i vi .
Potentially flipping bk we can assume than µk > 0. We can find some
` ∈ Z such that 0 ≤ ηk0 := ηk − `µk < µk . Then
d(w − `bk , Lk−1 ) = d(ηk0 vk , Lk−1 ) < d(µk vk , Lk−1 ) = d(bk , Lk−1 ) .
Hence, by our choice of bk we conclude that w − `bk ∈ Lk−1 − capΛ, so
that ` ∈ Z. As we already know a lattice basis b1 , . . . , bk−1 of Lk−1 we
ontain an integral representation of w in b1 , . . . , bk . Hence, wen we habe
reached Ld , then we have constructed a lattice basis of Λ. This reproves
Theorem 2.46.
So far, we have considered lattices in linear spaces. We can shift all
definitions to affine spaces.
Definition 2.58 (affine lattice) Let Λ be a subset of an affine space
A. Λ is an affine lattice if for some x ∈ Λ the set Λ − x is a lattice. A
subset B ⊆ Λ is an affine lattice basis of Λ if B − x is a lattice basis of
Λ − b. An affine lattice isomorphism is a map on Λ that comes from a
lattice isomoprhism on Λ − x.
So far, all our considerations about lattices did not depend on a particular
basis and a representation of transformations in coordinates w.r.t. to such
a basis. However, sometimes, in particular for explicit computations in
examples, it is more convenient to consider lattices and transformations
in a given basis. We now reconsider some notions in the presence of a
basis and introduce the Hermite and Smith normal form. Those allow us
to compute bases and reprove Theorem 2.53.
Lemma 2.59 Let B and B0 be bases of the lattices Λ and Λ0 respectively.
Then a linear map T : lin Λ → lin Λ0 is unimodular if and only if the
matrix representation A of T with respect to the bases B and B0 is
integral and satisfies | det A| = 1.
Proof. The matrix A has only integral entries if and only if T(Λ) ⊆ Λ0 .
Similarly, if T is unimodular, then the inverse transformation exists,
and its matrix A−1 also has integral entries. Thus, det A and det A−1 are
integers with product 1.
Conversely, if A is integral with | det A| = 1, then, by Cramer’s rule
A−1 exists and is integral. u
t
detΛ Λ0 := | det A| .
Then c11 > 0 as A has full row rank. Further, if c12 6= 0, then we can
subtract the second from the first column and reorder the columns if
necessary to obtain a smaller total sum c. Hence, c12 = c13 = . . . =
c1,m−k = 0. The column operations on C clearly extend to A without
affecting B and M , so we can apply them to A to obtain a matrix
B 0 0
A = m c11 0 ,
M 0 c01 C 0
(1) swap a column with a non-zero entry in the first position to the front,
possibly multiply by −1 to make it positive
(2) for any column cj with non-zero first entry c1j one computes the
greatest common divisor g of c11 and c1j and two integers x, y such
that g = xc11 + y + c1j . This can be done with the extended Euclidean
algorithm. Now we replace the first column c1 by xc1 + ycj and
the column cj by 1/g (c1j c1 − c11 cj ). Note that in the second linear
combinations the coeffcients 1/gc1j and 1/gc11 are both integral.
Remark 2.68 We can use the Hermite normal form to efficiently per-
form various tasks on lattices. For this, Let B and B 0 be matrices whose
columns generate lattices Λ := Λ(B ) and Λ0 := Λ(B 0 ).
(1) The first d colunms of the Hermite normal form of B give a basis of
the lattice Λ.
(2) The lattices Λ and Λ0 are equal if and only if the Hermite normal
forms of B and B 0 coincide.
(3) lattice0 is a sublattice of Λ if and only if the Hermite normal forms of
B and the matrix obtained by adding the columns of B 0 to B coincide.
For the Hermite normal form we have used elementary column trans-
formations, which we can realize by multiplication with a unimodular
matrix from the right. Clearly, we can study the same transformations
also for the rows of a matrix, and we can realize them by multiplications
with a unimodular matrix from the left. This leads to another important
normal form of a matrix, which we explain with the next theorem.
Theorem 2.69 (Smith normal form) Let A ∈ Zd×m be a matrix
of full row rank. Then there are unimodular matrices L ∈ Zd×d and
R ∈ Zm×m such that S = (sij )1≤i≤d,1≤j≤m := LAR satisfies
(1) sij = 0 for i 6= j,
(2) sii > 0 for 1 ≤ i ≤ d, and
(3) si−1,i−1 divides sii for 2 ≤ i ≤ d.
The matrix S is unique, the companion matrices L and R are not.
The last statement about the non-uniqueness of L and R follows from
the obsevation that there are unimodular matrices that commute with
S. When actually computing smith normal forms with their companions,
this fact can be used for an attempt to keep entries in L and R small.
" #
S 0
A = (2.7)
0C
You will use the Smith normal form to reprove Theorem 2.53 using bases
of the lattices in a representation w.r.t. to a basis of the vector space in
Exercise 2.35. Exercise 2.35
In this section, we will give a short discussion about lattices and metric
geometry (mainly following [4]). This is the first point in the book where
Λ really is meant to be a (non-standard) lattice in Rd . An interesting
geometric application can be found in the next section.
Usually, when dealing with lattice polytopes we start with an abstract
∼ Zd and associate an abstract vector space Λ ⊗Z R =
lattice Λ = ∼ Rd with
the volume form which evaluates as 1/d! on a fundamental domain of
Λ. In particular, note that the ’length’ of a vector is not well-defined. In
general, we define the dual lattice as
Λ? := HomZ (Λ, Z)
(Λ ⊗Z R)? := HomR (Λ ⊗Z R, R) .
Note that the dual lattice naturally sits inside of the dual vector space.
While these definitions are abstract, they stress the point that in general
it is not necessary and often misleading to identify dual spaces or lattices.
In contrast, in lattice theory the viewpoint is opposite to ours. The
starting point is an euclidean vector space, say, Rd with the usual scalar
∼ ( Rd ) ? ,
Rd = x → h ·, x i.
Λ? = {x ∈ Rd : h x, y i ∈ Z ∀ y ∈ Λ} ⊆ Rd .
k
X
x = (bηi c + {ηi }) yi ,
i=1
so that
k
X k
X
x− bηi cyi = {ηi }yi .
i=1 i=1
The left side of this equation is a lattice point. Hence, also the right
side is a lattice point. But
k
X
h := {ηi }yi ∈ Π ,
i=1
n o
y not a sum of
Let K := y ∈ C ∩ Zm | y 6= 0, two other integral vectors in C .
Then K ⊆ H, so K is finite.
Assume that K is not a Hilbert basis. Then there is x ∈ C such that
x 6∈ Z≥0 K. Choose x such that bt x is as small as possible.
Since x 6∈ K, there must be are x1 , x2 ∈ C such that x = x1 + x2 .
But
bt x 1 ≥ 0 , bt x2 ≥ 0, bt x ≥ 0 and bt x = bt x1 + bt x2 ,
so bt x 1 ≤ bt x , bt x 2 < b t x .
From the above proof it follows that for a simplicial cone all Hilbert
basis elements ecept for the generators of the cone are contained in the
fundamental parallelepiped of the cone. Computing these points in the
parallelepiped can be done by computing the Smith normal form. This,
together with the generators of the cone is only a generating set G for
the integer points in the cone. So we need to check for all of the (finitely
many points) whether it is a sum of two other elements in G and in this
case remove it from G.
For non-simplicial cones we can obtain a Hilbert basis in three steps:
ht(x) := ct x
2.4.1 Equivalence
0
Definition 2.74 Two lattice polytopes P ⊂ Rd and P 0 ⊂ Rd (with
0
respect to lattices Λ ⊂ Rd and Λ0 ⊂ Rd ) are isomorphic or unimodularly
equivalent, if there is an affine lattice isomorphism of the ambient lattices
Λ ∩ aff (P ) → Λ0 ∩ aff (P 0 ) mapping the vertices of P onto the vertices of
P 0.
Recall from Definition 2.51 that a lattice isomorphism is just an iso-
morphism of abelian groups. Moreover, an affine lattice isomorphism is
an isomorphism of affine lattices. Here, note that an affine lattice does
not need to have an origin (e.g., consider the set of lattice points in
a hyperplane). However, if we fix some lattice point to be the origin,
an affine lattice isomorphism can be defined as a lattice isomorphism
followed by a translation, i.e., x 7→ T x + b where T : Λ → Λ0 is a (linear)
lattice isomorphism and b ∈ Λ0 .
Pyr(P ) := conv({0} × P , e0 ) ,
BiPyr(P ) := conv(P , x, y ) ,
where 0d and0e are the zero vectors in dimension d and e. This is clearly
again a lattice polytope. Note that a pyramid is a special case of this,
where we take Q to be a single point.
Further constructions, e.g. Cayley polytopes and Lawrence prisms
will be discussed at the relevant places in the next chapters.
Lattice polytopes also play an important role in various branches of
mathematics. We give a few examples.
(1) In enumerative combinatorics one can study the order polytope
(2) cut polytopes or traveling salesperson polytopes in combinatorial opti-
mization
(3) hypersimplex
(4) Birkhoff polytope and the permutation polytopes
2.4.3 Volumes
d! vol(P ) ∈ Z≥1 .
nvol(P ) := d! vol(P ) .
Remark 2.81 Note that it makes sense to extend the previous defini-
tion also to low-dimensional lattice polytopes by considering them as
full-dimensional polytopes with respect to their ambient lattice. Hence,
nvol(P ) ≥ 1 for any lattice polytope.
We will first prove Theorem 2.82 for simplices. We need the following
useful observation to extend this result to arbitrary polytopes. The proof
is left as Exercise 2.38. The centroid of a simplex with vertices v0 , . . . , vd
1 Pd
is d+ 1 i = 0 vi .
S ⊆ P ⊆ (−d)(S − x) + x ,
Proof (of Theorem 2.82). We can assume that one vertex of P is the
origin 0. First, let P = conv(0, v1 . . . , vd ) be a simplex. Let V ∈ Zd×d
be the matrix whose columns are the coordinate vectors of v1 , . . . , vd .
By the Hermite normal form (Theorem 2.65) (where we transpose the
left and right side of the equation) there exists U ∈ Gld (Z) such that
U V = H and H is an upper triangular matrix with non-negative integer
entries such that in each column the maximal element is on the diagonal.
We denote the columns of the right matrix by h1 , . . . , hd . Therefore,
U defines a unimodular transformation mapping P to
where the last inclusion follows, as det H ≥ hii for all 1 ≤ i ≤ d and
0 ≤ hij ≤ max hii . This proves the claim for simplices, as nvol(P ) =
nvol(Q) = det H.
In general, there exists a lattice d-simplex S ⊆ P as in Lemma 2.84.
Then the previous part of the proof shows that there exists a unimodular
transformation ϕ : Zd → Zd such that
2.5 Software
We can du actual computations with polytopes, cones and fans using the
software framework polymake.
polytope> $c=cube(3);
polytope> print $c->VERTICES;
1 0 0 0
1 1 0 0
1 0 1 0
1 0 0 1
1 1 1 0
1 1 0 1
1 0 1 1
1 1 1 1
2.6 Problems
included on pa
2.1. Prove Carathéodory’s Theorem (Theorem 2.4).
included on page 27
2.10. Prove that the tangent cone of a face of a polytope is precisely the
intersection of the half spaces defining F .
included on page 32
2.11. Show that any subdivision S of a polygon P such that V(S) = V(P )
is regular.
included on page 33
2.15. Show that the definition of a lattice in Definition 2.37 does not
depend on the norm chosen to define the balls.
2.16. Show that in Definition 2.37 we can choose the same ε for all x ∈ Λ.
and (x + Λ) ∩ (y + Λ) = ∅ for x, y ∈ Λ, x 6= y.
included on page 38
included on pa
included on page 39
Show that
2.35. Reprove Theorem 2.53 using the Smith normal form (Theo-
rem 2.69).
and
where
d
1 X
x := vi
d+1
i=0
is the centroid of S.
Hint: Choose S := conv(v0 , . . . , vd ) ⊆ P with maximal volume in
P.
For any 0 ≤ i ≤ d let Hi be the facet hyperplane of the
facet of S not containing vi , ri := d(vi , Hi ) and Ri := {x :
d(x, Hi ) ≤ ri }.
Show that P ⊆ Ri for 0 ≤ i ≤ d.
Td
Express i=0 Ri , (−d)(S − x) + x, and (d + 2)(S − x) − vx
in barycentric coordinates with respect to v0 , . . . , vd and
compare.
included on page 9
2.41. Wie kann man die affine Hülle mithilfe von ‘Affin-Kombinationen’
beschreiben? kleinster ist!
included on page 9
2.43. Check that Weyl-Minkowski for cones implies that for polytopes
included on page 9
2.44. Man erinnere sich, wieso die Determinante einer linearen Abbildung
unabhängig von der Basiswahl ist. Nun beweise für ein Gitter
Λ ⊂ Rd , dass wenn T eine lineare Abbildung von Rd nach Rd
ist, so dass die Einschränkung TΛ : Λ → Λ wohldefiniert ist und
surjektiv, dann ist TΛ bijektiv. (Tipp: wieso ist T bijektiv?) Zeigen
Sie allgemeiner(?), dass ein Homomorphismus von Λ → Λ surjektiv
ist g.d.w. bijektiv.
included on page 9
2.47. Man mache sich an einem Beispiel plausibel (oder beweise für
i = 1), dass für gegebenes i ∈ {1, . . . , n} und für ganzzahlige n × n-
Matrizen der ggT aller Determinanten von i × i-Untermatrizen bei
Multiplikation mit unimodularen Matrizen invariant bleibt. Wieso
impliziert dies die Eindeutigkeit der Smith-Normalform?
included on page 9
ein Isomorphismus.
included on page 9
2.54. Zeige, dass unter einer affin-linearen Abbildung die konvexe Hülle
des Bildes einer Menge gleich dem Bild der konvexen Hülle ist.
included on page 9
2.56. Finde das normalisierte Volumen der konvexen Hülle von (2, 0, 4),
(1, 1, 0), (0, 2, −2).
included on page 9
2.57. Ist der Endlichkeitssatz für Gittersimplizes scharf? (Offen: was für
Gitterpolytope?)
ehrP (k ) := |k · P ∩ Zd | ,
3.1 Motivation
for 0 ≤ j ≤ n − 1.
Volumes The most important natural invariant of a convex body is
its volume. Computing the volume of a convex body is in general a
complicated problem. Counting lattice points in multiples of a polytope
is directly related to it. Let P ⊂ Rd be a convex body. As illustrated
in Figure 3.1, we can approximate the volume by counting the volume
of little cubes centered at the more and more refined lattice Zd /k (for
k → ∞).
Z
1 1
vol(P ) = dx = lim d |P ∩ (Zd /k )| = lim d |kP ∩ Zd |
P k→∞ k k→∞ k
1
= lim d ehrP (k )
Fig. 3.1: Approximating a convex k→∞ k
body by smaller and smaller cubes (3.1)
kS := {kx : x ∈ S} .
k 7−→ kS ∩ Zd .
Figure 3.2 shows the interval I = [0, 32 ] and its second and third dilation.
If the boundary points a and b are integral and a ≤ b, then we can
simplify the formula. In this case also all multiples of a and b are integral,
and we can omit the floor and ceiling operations to obtain
ehrL (k ) = k (b − a) + 1 .
∆d := conv(0, e1 , . . . , ed )
introduced in Definition 2.76. See Figure 3.3 for the lattice points in a
multiple of this ismplex.
Proposition 3.2 Let ∆d be the d-dimensional standard simplex. Then
d+k (d + k ) · (d + k − 1) · . . . · (k + 1)
ehr∆d (k ) = = .
d d!
Observe that this is a polynomial in the variable k of degree d with
leading coefficient 1/d! .
· · | · · · ||· ←→ x = (2, 3, 0)
This yields a bijection between the sequences and lattice points with
P
non-negative coordinates and with xi ≤ k. u
t
Another simple, but very important example is the unit cube defined in
Example 2.7(1). The k-th dilate of the cube is kCd = k · [0, 1]d = [0, k ]d .
Hence, the Ehrhart counting function is given by
ehrCd (k ) = (k + 1)d .
This works well in this small example, but consider the structurally
similar example P10002 := [0, 10002]. Here, our plain list
0, 1, 2, 3, 4, 5, 6, 7, 8, 9, 10, 11, 12, 13, 14, 15, 16, 17, 18, 19, 20, 21, 22, . . .
easily exceeds a line, and also the length of this book. To get a compact
encoding of the points we need a better idea.
Here is one that might look really strange at first, but will prove to
be very powerful. We can replace each point k ∈ P3 with its monomial tk .
With this we define a polynomial that contains precisely the monomial
corresponding to points in our polytope:
3
X
1+t+t +t 2 3
= ti ,
i=0
1 − t4
GP3 (t) := .
1−t
whose expansion is again our polynomial. Again, doing the same for
P10002 does not really make this notation more complicated:
1 − t10003
G[0,10002] (t) := .
1−t
We will see that this idea of using a geometric series to specify the
lattice points in a polyhedron is both sufficiently flexible to work for all
polyhedra, and efficient enough that we can use it to really study the
structure of the set of lattice points.
Her comes another surprising and powerful property of our last
observation. If we try to do write down the lattice points of the unbounded
polyhedron P∞ := [0, ∞), then our first two approaches obviously become
infeasible. However, the third works and turns out to be even shorter
1 If the reader feels slightly wary about and more appealing1 ! As a geometric series we can consisely describe all
what happens at t = 1, be assured that lattice points in P∞ via the monomials in
in our approach here we will not deal
with any analytic convergence issues 1
and will not evaluate at certain values. G[0,∞) (t) := .
1−t
As this extended example suggests, the generating function we used to
encode the lattice points will indeed provide a powerful bookkeeping tool
for counting and enumerating lattice points in polytopes.
It will soon become apparent it is indeed quite useful and natural to
encode lattice points not only in polytopes, but more generally in any
bounded or unbounded subset of Rd , as in the last example of a ray in
R1 . You should keep this in mind for the following considerations.
In the above example of the one-dimensional cone x ≥ 0 ⊆ R we
have seen that we can use rational functions in one variable t to describe
the infinite series of all monomials corresponding to the lattice points in
the cone. We now want to formalize this idea, and directly generalize it
to arbitrary dimensions. Let | be some ground field (you can just think
of | = C, if you like). We assign the monomial
L = |[t±1 ±1
1 , . . . , td ] .
Moreover, note that the sum of monomials for the cone x ≥ 0 is infinite.
Since we do not care about convergence, we will actually consider our
sums not as Laurant polynomials, but as series in a subset of the L-module
L := |t±1
b ±1
1 , . . . , td
t21 t22
+ t2 + t1 t2 + t21 t2
+ t1 + t21 + t31
Fig. 3.5: The polygon of Example 3.3.
+ t21 /t2 .
Remark 3.5 (Warning) When dealing with formal power series and
rational series, one has to be very careful in order not to make a mistake.
Therefore, we would like to give an example here (just for one variable)
that justifies this caution: Consider the following expression of formal
Laurent series
b R (t) (1 − t) = (· · · + t−2 + t−1 + 1 + t + t2 + · · · )(1 − t)
G
= (· · · + t−2 + t−1 + 1 + t + t2 + · · · )
− (· · · + t−1 + 1 + t + t2 + t3 · · · )
= 0
Clearly, we deduce
Actually, not all Laurent series appear as a generating series for lattice
points in polyhedra. The ones we will encounter have a nice additional
structure that we will work out with the next definitions and theorems.
Definition 3.6 (summable Laurent series) A Laurent series G b∈b L
is summable if there is a Laurent polynomial g ∈ L such that the series
gGb is a Laurent polynomial.
Example 3.8 Before we continue we want to work out some simple, but
quite important examples of summable series coming from polyhedra.
(1) Let us first consider the polyhedron P∞ = [0, ∞) that we introduced
above. The integer point series is
X
b P (t) =
G ∞ ta = 1 + t + t2 + t3 + · · · .
a∈Z≥0
= 1 + t + s + t2 + s2 + ts + t3 + · · · .
g (t, s) := (1 − t)(1 − s)
to obtain
X X
g (t, s) · G
b C (t, s) = (1 − t) · ta · ( 1 − s ) · sb = 1.
a∈Z≥0 b∈Z≥0
Φ : Lsum −→ R := |(t1 , . . . , td ) ,
mapping G b = f in b
b to f /g if g G L. We will abbreviate this also by writing
b → Φ f
G 7
g
The proof of this proposition is left as Exercise 3.5. Exercise 3.5
GS (t) := Φ(G
b S (t)) .
Proof. Let Z≥0 V stand for the set of Z≥0 -linear combinations of V . By
replacing the Laurent monomial tei by tai in (3.3), we get
d
Y X d
Y X
(1 − tai ) · tz = (1 − tai ) · (ta1 )n1 · · · (tad )nd = 1.
i=1 z∈Z≥0 V i=1 (n1 ,...,nd )∈Zd≥0
Our integer point generating series contain a monomial for every lattice
point in a set. If we have the series for two sets S, S 0 , then we can obtain
the series for the union S ∪ S 0 by adding the two series and subtracting all
lattice points that we encoded in both series, i.e. the generating series for
the lattice points in the intersection S ∩ S 0 . This principle clearly extends
to the union of any finite number of sets. We can compute the generating
series from the generating series of the sets and all partial intersections
if we keep track of the multiplicities a partial intersection appears in
the total sum. This is called the principle of inclusion-exclusion. You
will study this in more detail in Exercise 3.9. Triangulating a rational
polyhedral cone into rational simplicial cones (see Section 2.2)) and using
P2
inclusion-exclusion (see also Figure 3.6) yields the following general result.
Q
Corollary 3.13 The integer point generating series of a rational poly-
hedral cone is summable. P1
We usually write a vector x ∈ Rd+1 with indices starting from 0 and use
x0 for the special coordinate. See Figure 3.7 for the cone over a triangle.
In our setting the especially convenient property of this representation
of our polytope is the fact that we can recover all dilates of P from C (P ).
More precisely, for any k ≥ 0 we get the k-th dilate of P by intersecting
C (P ) with the hyperplane x0 = k, and the lattice points in kP by
intersecting with {k} × Zd . Hence,
X X
G
b
C (P ) (t, 1, . . . , 1) = kP ∩ Zd tk = 1 + ehrP (t) tk .
Fig. 3.7
k≥0 k≥1
Now, let us compute the Ehrhart generating function for lattice simplices.
We have now collected all necessary tools and definitions to prove Fig. 3.9: Inclusion-Exculsion on the
boundary complex
Ehrhart’s Theorem (Theorem 3.14).
Proof (of Ehrhart’s Theorem (Theorem 3.14)). Combining Proposi-
tion 3.19 with Proposition 3.18 we get that the Ehrhart counting function
of an n-dimensional lattice simplex in Rd uniquely extends to a polyno-
mial function of degree at most n.
For a general polytope P we triangulate it into maximal-dimensional
simplices Fi and consider the triangulation of C (P ) into the associated
simplicial cones C (Fi ). Then we apply inclusion-exclusion (e.g. Exer-
cise 3.9). Finally, we use (3.1). u
t
Exercise 3.9
ehrP (0) = 0P ∩ Zd = 1.
Fig. 3.10: Making a cone half open. C ξ := {y ∈ C : y + εξ ∈ C for all ε > 0 small enough} .
The right face and the origin are
not part of the half open cone.
We say ξ ∈ Rd is generic with respect to C (respectively, a triangulation
T of C) if ξ is not in the linear hull of a (d − 1)-dimensional face of C
(respectively, any simplicial (d − 1)-cone in T).
C −ξ = int C .
Proof. The fact that the translates by Λ-vectors are pairwise disjoint
follows from the uniqueness in Lemma 3.26. From the existence part we
see that Rd is covered by all Λ-translates of Πξ (D ). It remains to observe
that for w ∈ Λ
w + Πξ (D ) for w ∈ Z V
≥0
D ξ ∩ w + Πξ ( D ) =
∅ else,
Let us compute the integer point generating series and generating function
for half-open cones.
In particular, both series are summable, and (3.5) also holds on the level
of rational functions.
Then
R := conv(0, e1 , e2 , e1 + e2 + 13e3 ) .
1 + t + t2 3t
= +1.
(1 − t)2 (1 − t)2
We also see that there is exactly one choice for the multiplicity of the
origin (in this case zero) so that the h? -polynomial has degree ≤ 1 which
we need if we want the counting function to agree with the evaluation of
a polynomial. We will identify this choice with the Euler characteristic in
Remark 3.46 below.
Exercise 3.25
Finally, let us note the following theorem proved by Stanley in [53]. A
completely different proof appears in (Beck, Sottile [9]). The reader can
try to give a proof using the methods developed above in Exercise 3.26.
Theorem 3.37 (Stanley’s Monotonicity Theorem) Let P and Q
be two lattice polytopes such that P ⊆ Q, d = dim Q and let h?P and h?Q
be their h? -polynomials. Then h?P ,i ≤ h?Q,i for all 0 ≤ i ≤ d.
Exercise 3.26
The origin for the funky notation h? is its close connection to the
h-vector from enumerative combinatorics. Suppose C is a simplicial
3.5 Reciprocity
The interior of L = [a, b] is int L = (a, b). For integers a, b we can count
the lattice points inside int L:
ehrint L (k ) = k (b − a) − 1 .
This just means that we only want to count lattice points that satisfy
the inequalities xi ≥ 1 for 1 ≤ i ≤ d, and whose coordinates sum up to
at most k − 1. Hence, we want to count lattice points in the set defined
by the inequalities
d
X
xi ≥ 1 and xi ≤ k − 1 .
i=1
and this simplex clearly contains the same number of lattice points. We
have computed this number in Proposition 3.2, so
k−1
ehrint ∆d (k ) = .
d
We can make the same observation as for the interval: The lattice points
in the interior of the k-th dilation of the simplex are (up to a sign) the
evaluation at −k of the Ehrhart polynomial!
Let us check one more example, before we attempt to prove our
observation. Consider the standard unit cube Cd . Counting the interior
points in this case is rather simple. We obtain
and again, the number of lattice points in the interior is given by the
Ehrhart polynomial evaluated at negative values.
X ts(S ) GΠ−ξ (S ) 1t
= Q v
S∈T [d] v∈V (S ) (1 − t )
GΠ−ξ (S ) 1t
X
d
= (−1) Q 1
S∈T [d] v∈V (S ) (1 − tv )
d
X 1 d 1
= (−1) GS −ξ = (−1) Gint C . u
t
t t
S∈T [d]
It is important to note that Stanley’s theorem is clearly wrong on the
level of Laurent series!
The proof needs a little fact about the map Φ that maps summable
Laurent series to rational functions.
X k − 1 −k X k − 1 −k
= (−1)m t = (−1)m t
m m
k≥1 k≥m+1
X k + m
m −(m+1)
= (−1) t t−k
m
k≥0
1 (−1)m
= (−1)m t−(m+1) 1 m+1 =
(1 − t) tm+1 (1 − 1t )m+1
Comparing coefficients of these two Laurent series gives the desired result.
u
t
As an immediate application we can compute the Euler characteristic.
Proposition 3.45 (Euler-Characteristic) Let S be a subdivision of
the rational polytope P ⊂ Rd into rational polytopes. Then
X
(−1)dim F = 1 .
∅6=F ∈S
Remark 3.46 The same argument shows that the constant coefficient
of the counting polynomial of a complex of lattice polytopes equals the
Euler-characteristic of the complex: ehrC (0) = χ(C). This resolves the
riddle raised in Remarks 3.20 and 3.36.
codeg(P ) := d + 1 − deg(P ).
Proof. This follows from Lemma 3.49 and the Ehrhart-Macdonald Reci-
procity (Theorem 3.43). u
t
Exercise 3.27
The proof is left as Exercise 3.28. Applied to our situation this has the
following immediate consequence. Exercise 3.28
In particular, this implies that for deg P = 0 only the vertices are lattice
points, and for deg P = 1 the only lattice points that are not vertices
are on the edges of P .
Let X = {x1 , . . . xk } ⊆ P ∩ Λ be a set of k lattice points that is not
entirely contained in a proper face of P . Then x := x1 + · · · + xk is an
interior lattice point of kP . This proves the next propisition.
where e1 , . . . , ed+1 is the standard basis of Rd+1 . See Figure 3.14 for an
example. In Exercise 3.30 you will show the following proposition.
Definition 3.56 (Lawrence Prism) A Lawrence prism with heights Exercise 3.30
Exercise 3.31
h1 , . . . , hd ≥ 0 and h1 + · · · + hd ≥ 2
is the polytope
0, e1 , . . . , ed−1 ,
Law(h1 , . . . , hd ) := conv .
e1 + h1 ed , . . . , ed−1 + hd−1 ed , hd ed
u1 + u2 = u3 + u4 .
Proof.
Proof. Let us first show that these conditions are necessary. Note that h?2
is the number of interior lattice points i, while h?1 = b + i − 3, where b is the
number of boundary lattice points. Moreover, vol(P ) = nvolZd (P )/2 =
(1 + h?1 + h?2 )/2. Hence, Scott’s theorem tells us that, if i ≥ 1 and P 6= ∼
3∆2 , then h1 ≤ 3h2 + 3. Finally, if i ≥ 1, then h2 = i ≤ h1 = b + i − 3,
? ? ? ?
since b ≥ 3.
(0, 0) (2, 0) It suffices to realize lattice polygons satisfying each of these conditions.
For i = 0, any b ≥ 3 can be realized by lattice polygons of the form
Fig. 3.16: Lattice polygons re-
alizing 4 ≤ b ≤ 2i + 6 as depicted in Figure 3.15. In fact, it is not difficult to show that these
In fact, as Exercise 3.33 shows these are precisely the lattice polygons
without interior lattice points.
Let i ≥ 1. The condition h?2 ≤ h?1 ≤ 3h?2 + 3 is equivalent to 3 ≤ b ≤
2i + 6. The case b = 3 is easy to realize, so let b ≥ 4. Then any of these
cases is realized by Figure 3.16. u
t
All possible pairs (h?1 , h?2 ) are depicted in Figure 3.17. Let us now deduce
all Ehrhart polynomials c2 t2 + c1 t + 1 of lattice polygons. By Pick’s
Theorem 1.8 c2 equals the area of P , and by Proposition 3.51, c1 is
half the number of boundary lattice points of P . The following theorem
characterizes all pairs (c1 , c2 ) that correspond to an Ehrhart polynomial
of a polygon.
Corollary 3.62 A polynomial c2 t2 + c1 t + 1 with c1 , c2 ∈ 1/2Z and
c1 ≥ 3/2 defines the Ehrhart polynomial of a lattice polygon P if and only
if one of the following three conditions is satisfied:
Fig. 3.17: (h?2 , h?1 ) of lattice polygons
(1) c1 − c2 = 1. Then P has no interior lattice points.
(2) c1 = c2 = 9/2. Then P is 3∆2 .
(3) c1 ≤ c2/2 + 2. Then P has interior lattice points.
Exercise 3.34
The goal of this final section is the celebrated Theorem of Brion. It relates
for any lattice d-polytope the integer point generating functions of all
vertex cones of P to the integer point generating function of the polytope.
Let P be a rational d-dimensional polytope and F a face of P . Recall
the tangent cone of F in P from Definition 2.29:
TF P := {v ∈ Rd : ∃ w ∈ F , ε > 0 : w + ε(v − w ) ∈ P } .
But then, this strict inequality is true for every point of the form (1 −
λ)v + λw for λ ∈ [0, 1). So none of them can belong to P . That is, F is
visible from v.
If, conversely, v ∈ TF P , we know that there are w0 ∈ F and an ε > 0
so that v 0 := w0 + ε(v − w0 ) ∈ P . Further, taking a smaller ε if necessary,
we can assume that w00 := w + ε(w − w0 ) ∈ F , because w ∈ relint F . But
then the point
ε2 v + (1 − ε2 )w = (1 − ε)w00 + εv 0
= Φ(G
b R+P (t)) = Φ(G
b R (t)) + Φ(G
b P (t))
This can only hold if Φ(Gb R (t)) = 0, i.e. if Φ maps the infinite series
k
P
k∈Z t to 0. The following proposition shows that this indeed holds in
general for cones with nontrivial lineality space.
v 6= 0 implies Φ(G
b C ) = 0. u
t
We can apply the observation of this proposition to obtain a very simple
formula for the integer point generating function of a polytope.
Theorem 3.67 (Brion’s Theorem) Let P be a rational d-polytope.
Then
X
GP ( t ) = GTv P (t) .
v vertex of P
3.8 Problems
included on page 63
3.1. Determine a formula for the volume of a 3-dimensional lattice
polytope using the number of lattice points in the k-multiple for
k = 1, 2 and 3.
included on page 63
3.3. Show
∞
X k+d k 1
x = .
d (1 − x )d+1
k =0
Φ:b
L −→ R := |(x1 , . . . , xd ) ,
mapping G b = f in b
b to f /g if g G L.
included on page 69
included on pa
1
cd−1 = vol(∂P ).
2
Here, vol(∂P ) denotes the surface area of P , namely,
X
vol(∂P ) := vol(F ),
F ∈F(P )
where F(P ) is the set of facets of P and vol(F ) denotes the (non-
normalized) volume with respect to the lattice aff (F ) ∩ Zd . For
√
instance, note that vol(conv((1, 0), (0, 1))) equals 1 and not 2.
Hence,
(1) Argue that its vertices are its only lattice points. (White
proved a converse: every lattice tetrahedron with only four
lattice points is unimodularly equivalent to a ∆pq .)
(2) Compute the Ehrhart polynomial and the h? -polynomial of
∆pq .
(3) For which parameters are ∆pq and ∆p0 q0 unimodularly equiva-
lent?
included on page 74
3.20. Check carefully and rigorously the last identity in the proof of
Proposition 3.27.
included on page 76
3.23. Consider the polygon with vertices (0, 0), (1, 0), (0, 2), (2, 4). Com-
pute, using a half-open decomposition GC (P ) (t1 , t2 , t3 ).
included on page 78
included on pa
3.34. Are
(1) 1 + 8t + t2
(2) 1 + 9t + t2
(3) 1 + 2t + 2t2
(4) 1 + t + 2t2
3.35. Compute the Ehrhart generating function of P = [0, 1]2 using the
Brion’s Theorem (Theorem 3.67).
and verify that both rational functions coincide (you may want to
use a computer for this).
While most of the theory treats general convex bodies, in these notes
we will focus on those tools which we need to prove results that apply
only to lattice polytopes.
Minkowski’s two theorems are the basis of this whole branch of discrete
mathematics. Both essentially tell us something about generators of
the lattice and prove that we can find such generators with a bounded
Euclidean length. The theorems, of which we will only prove the first, are
however not constructive. We will remedy this in Chapter 6. Throughout,
Λ ⊂ Rd is a lattice of rank d (the reader may think of Zd ).
A set K ⊆ Rd is centrally symmetric if x ∈ K implies −x ∈ K.
Definition 4.1 A subset K ⊆ Rn is a convex body if K is bounded and
convex. The set of convex bodies in Rd is deneted by C. The subset of
centrally symmetric convex bodies is C0 .
Note that the definition of the term convex body varies in the literature.
The following theorem establishes a fundamental correspondence between
lattice points in a centrally symmetric convex body and its volume.
Theorem 4.2 (van Der Corput, 1935) Let K ⊂ Rd be a centrally
symmetric convex set. Then
vol(K ) ≤ 2d |K ∩ Λ| det Λ .
Sx := {y ∈ Π | x + y ∈ S} = Π ∩ (S − x) x+S
Now, we can easily prove van der Corput’s Theorem (Theorem 4.2).
Proof (of van der Corput’s Theorem (Theorem 4.2)). We will give an
indirect proof. Let us assume that y
for a positive integer m. Our goal is to show that there exist m distinct (b) And the shifted
pairs of non-zero lattice points ±x1 , . . . , ±xm in K. Together with the intersections with the fun-
damental parallelepiped
origin this will give 2m + 1 lattice points in K.
K
Let T := 12 K. Then vol T = vol 2d
> m det Λ. Hence, by Gen- Fig. 4.2: Illustrating the proof of
eralized Blichfeldt’s theorem (Lemma 4.4), there are m + 1 distinct Generalized Blichfeldt’s theorem
(Lemma 4.4)
points p1 , . . . , pm+1 ∈ T such that pi − pj ∈ Λ for all i, j. Choose
xi := pi − pm+1 for i = 1, . . . , m as the desired lattice points. Note
that xi = pi + (−pm+1 ) ∈ T + T = K.
Let K be compact and vol K = 2d det Λ. Since K is compact, for
each x ∈ 2K\K there exists 0 < x < 1 such that x 6∈ (1 + x )K.
Then
λ1 ≤ λ2 ≤ · · · ≤ λd .
kxkK := max ( µx ∈ K ) ,
µ
But wil − wil is a lattice vector, so wik = wil for sufficiently large k, l.
Hence, wik = w for sufficiently large k, and w is a lattice vector. u
t
Remark 4.9 The vectors found in the previous proposition need not be
a basis of the lattice Λ. For an example, the lattice polytope
in the lattice Z3 is centrally symmetric and its lattice points are the
vertices and the origin. Hence, the successive minima are λ1 = λ2 =
λ3 = 1, but no subset of the vertices is a lattice basis of Z3 .
Br (z ) := {x ∈ Rd | kx − zk < r}
to determine for which radii these translates are pairwise disjoint or cover
the whole space, and relations between these two. We start with the first
and introduce the packing radius of a lattice, which is the largers redius
of a ball such that any two translates to a lattice point either coincide or
are disjoint.
Definition 4.14 (packing radius) Let Λ be a lattice in Rd . The pack-
ing radius is
i.e. the largest r > 0 such that the open balls of radius r around any two
distinct lattice points do not intersect.
Exercise 4.4
See Figure 4.4 for some examples. You will prove the folloing lemma in
Exercise 4.4.
Proposition 4.15 Let Λ ⊆ Rd be a lattice and v a shortest non-zero
lattice vector in Λ. Then %(Λ) = 12 kvk. u
t
Exercise 4.5
Recall that the dual of a lattice Λ is defined to be the set of all linear
functionals that map lattice points to integers. This is itself a lattice Λ?
?
in Rd .
i.e. the largest possible distance between any point in Rd and its nearest
lattice point.
See Figure 4.5 for an example. The reader should convince her- or himself
that the covering radius is indeed well-defined and finite. In particular, the
maximum is attained for some point x ∈ Rd (by a standard compactness
argument).
Lemma 4.18 Let Λ be a lattice with successive minima λ1 , . . . , λd and
linearly independent vectors v1 , . . . , vd such that λi = kvi k for 1 ≤ i ≤ d.
Then Fig.
of the
1
µ(Λ) ≥ kvi k for 1 ≤ i ≤ d .
2 Exercise 4.7
Proof. Let u = 1/2vd . Assume there is w ∈ Λ such that d(u, w ) < 1/2kvd k.
Then
4 µ ( Λ ) · % ( Λ? ) ≥ 1
The vectors v1 , . . . , vd are a basis, so for at least one i we have |vi (u)| ≥ 1.
Hence, for that i
kvi k · kuk ≥ 1 ,
The following theorem is the key ingredient for the flatness theorem that
we will prove in Section 4.3.
4 µ(Λ) · %(Λ? ) ≤ d /2 .
3
% ( Γ ? ) ≥ % ( Λ? ) . (4.3)
ku − yk ≤ µ(Γ ) .
Consider the line π −1 (u). Any two neighboring lattice points of Λ on this
line have distance kvk. Hence, we can pick a point w ∈ Λ ∩ π −1 (u) such
that
1
d(x, w + (y − u)) ≤ kvk .
2
Using the right angled triangle x, w, w + (y − u) we compute
µ ( Λ ) 2 · % ( Λ ? ) 2 ≤ µ ( Γ ) 2 · % ( Λ? ) 2 + % ( Λ ) 2 · % ( Λ? ) 2
≤ µ ( Γ ) 2 · % ( Γ ? ) 2 + % ( Λ ) 2 · % ( Λ? ) 2
1
≤ (d − 1)3 + d2
16
≤ d3 ,
where the second inequality follows from (4.3), the third from Proposi-
tion 4.16 and the fourth by induction. This proves the theorem. u
t
You will show in Exercise 4.8 that for full-dimensional convex bodies the
infimum is actually a minimum, and in Exercise 4.9 that the width of
convex bodies with dimension less than the ambient dimension is actually
Exercise 4.8 0. Recall that an ellipsoid is the image of a ball (in some norm)
Exercise 4.9 under an affine linear map. See Definition A.2 for a full definition and
the whole Appendix A for properties. Our approach to bound the lattice
width of empty convex bodies will proceed in three steps. We first prove
it for balls, then extend to ellipsoids and finally use Theorem A.5 to
approximate an arbitrary convex body with ellipsoids from the interior
and the exterior. The following lemma does the first two steps.
Lemma 4.22 Let Λ be a lattice, v ∈ Λ? a shortest non-zero lattice vector
and E an ellipsoid such that E ∩ Λ = ∅. Then widthv (E ) ≤ d3/2 .
Proof. We prove this first for the case that E is a ball. In this case we
know by Proposition 4.15 that kvk = 2 %(Λ? ). Let r be the radius of the
ball. Then r ≤ µ(Λ). Now
We can extend our bound for the width of a convex body from balls and
ellipsoids to general convex bodies with empty interior, albeit only with
a weaker right hand side. The key observation for this is Theorem A.5,
which tells us the we can estimate any convex body from the interior and
exterior with a suitably chosen ellipsoid.
Theorem 4.23 Let K ⊂ Rd be a convex body with K ∩ Λ = ∅. Then
5
widthΛ (K ) ≤ d 2 .
widthv (K ) ≤ d widthv (E ) ≤ d · d /2 = d /2 .
3 5
u
t
Remark 4.24 In fact, the bound of the previous theorem can be strength-
ened to be of order d3/2 , so that
3
widthΛ (K ) ≤ O d 2 .
Note that the upper bound only depends on the dimension and not on
the given lattice. It is unknown and an active subject of current research,
whether the sharp bound is actually of the form O (d).
polytopes. The reason for this is one of the arguably most important
finiteness result about lattice polytopes. We prove here a qualitative
version, following and extending an idea of Borisov & Borisov [14] (see
also [13, Theorem 4.1]).
∆d := {[x] ∈ T d : x ∈ ◦∆d } ⊂ T d .
Mid := {y ∈ ∆d : |hyi ∩ ∆d | ≤ i} ⊂ T d
hyi = {[(0, 0)], [(1/2, 1/3)], [(0, 2/3)], [(1/2, 0)], [(0, 1/3)], [(1/2, 2/3)]}.
Then
β (int (S ) ∩ Λ) ⊆ Mid .
Look at Figure 4.10 to get a better intuition for this definition. Note
that there are two kinds of elements of the group T d : the rational points
have finite order, and the irrational points have infinite order.
Lemma 4.30 For x ∈ {0}r × Rd−r (with 0 ≤ r ≤ d) and ε > 0 there is
Fig. 4.10: Illustrating the definition of
a positive integer k and z ∈ {0}r × Rd−r with [z ] = [kx] and ||z|| < ε. our metric
We assume that Mid (thus, Mˆid ) is not finite. Then there exists an
accumulation point x∗ ∈ ∆d of Mˆid . Changing the coordinate system
on T d if necessary, we may assume that x∗1 = . . . = x∗r = 0, and
x∗r +1 , . . . , x∗d , 1 − dj=1 x∗j > 0. In particular,
P
x∗ ∈ {0}r × Rd−r
>0 .
We can assume this by permuting just the coordinates, since Mid can be
Exercise 4.10 characterized in a symmetric way, see Exercise 4.10.
Let us choose ε > 0 so that Bε (x∗ ) ∩ (Rr>0 × Rd−r ) ⊂ ◦∆d d , see
Figure 4.12.
Now, Lemma 4.30 implies the existence of a positive integer k and
and z ∗ ∈ {0}r × Rd−r so that [kx∗ ] = [z ∗ ] and kz ∗ k < ε/2i. With x∗
being an accumulation point of Mˆid , there is an x ∈ Mˆid such that
where all these (i + 1) elements are pairwise different. This would show
[x] 6∈ Mid , which is a contradiction.
First, we observe that
É wj ∈ Rr>0 × Rd−r .
É ||wj − x∗ || < j 2i
ε
+ (jk + 1) 2(ikε+1) ≤ ε.
wj = x + j (z ∗ + k (x − x∗ )).
Now that we have braved the technical core in the proof of Theorem 4.25,
the machinery will give us a volume bound V (d, i) in terms of a parameter
ε(d, i) that we introduce below. It will depend only on the dimension
and the number of interior lattice points.
The smallest barycentric coordinate of a point inside a simplex is
a measure for how far in the interior of the simplex the point sits. To
measure the same thing for points in more general polytopes (or convex
bodies), we use the convex-geometric notion of coefficient of asymmetry.
Definition 4.31 Let K ⊂ Rd be a d-dimensional convex body, w ∈
int K, then
w + λη
max{λ > 0 : w + λη ∈ K} λ0
ca(K; w ) := sup
η∈Rd \{0} max{λ > 0 : w − λη ∈ K} K w
λ
is the coefficient of asymmetry.
We have ca(K; w ) ≥ 1. Note that ca(K; w ) = 1 if and only if K is
centrally symmetric with respect to w. So, the closer ca(K; w ) is to 1 (a) An example with ca(K; w ) = 1
the more w lies in the ’center’ of K (the converse may not be true). See
Figure 4.13 for two examples. We now extend Proposition 4.28 from λ0
w
simplices to polytopes.
K λ
Definition 4.32 (Minimal barycentric coordinates) For positive in-
tegers d and i, the minimal barycentric coordinate of any interior lattice
point in any lattice d-simplex with precisely i interior lattice points will be
denoted by sbc(d, i). This definition yields a well-defined positive number
(b) An example with ca(K; w ) 1
sbc(d, i) because of Corollary 4.29, a consequence of the main result of
the previous section. Fig. 4.13: Coefficient of Asymmetry
You will prove some simple properties in Exercise 4.11. Here is a simple
Exercise 4.11
fact that you will prove in Exercise 4.12.
Lemma 4.33 For d ≥ 1 and i ≥ 1 we have sbc(d + 1, i) ≤ sbc(d, i)/2.
u
t
Exercise 4.12
Exercise 4.13 1
ca(S; w ) = − 1.
β0
Exercise 4.14 You will give a proof of this in Exercise 4.14.
Exercise 4.15
Proof (of Proposition 4.34). Let v be a vertex of P as in (4.4). Let
c := ca(P ; w ), and denote the opposite point by v 0 := w − 1c (v − w ).
There is a face F of P which contains v 0 in its relative interior. In a
lattice triangulation of F there must be a lattice simplex S 0 which contains
v 0 in its relative interior. Hence, the lattice simplex S := conv(v, S 0 ) of
dimension 1 ≤ d0 ≤ d contains w in its relative interior, and the segment
conv(v, v 0 ) certifies ca(S; w ) ≥ c. Furthermore, every relative interior
point of S is an interior point of P , so S contains j interior lattice points
Fig. 4.15: ca(P ; w ) ≤ for 1 ≤ j ≤ i. The statement follows now from the previous corollary. u t
ca(S; w ) for some simplex S
We are finally in the position to finish the proof of Theorem 4.25. For this
we need the following observation. You will prove this in Exercise 4.17.
Lemma 4.37 Let K be a d-dimensional convex body with 0 ∈ int K. Set
c := ca(K; 0). Then − 1c K ⊆ K.
Exercise 4.17
The bound is tight, as you will show in Exercise 4.18. This implies Exercise 4.18
Theorem 4.25 with
d
1
V (d, i) = (i + 1)2d−1 −1 .
sbc(d, i)
Let us finish by presenting without proof the currently best and most
general result. The following theorem is a combination of results obtained
by Hensley [27], Lagarias and Ziegler [37] and Pikhurko [45].
Theorem 4.39 (Hensley; Lagarias & Ziegler; Pikhurko) Let P ⊆
Rd be a d-dimensional lattice polytope such that Il (P ) 6= ∅. Then
2d+1
vol(P ) ≤ (8dl )d (8l + 7)d·2 | int P ∩ lZd |
We can make the bound V (d, i) of the previous theorem more precise.
Similar to the volume we can also look at the total number of lattice
points of a lattice polytope when the number of interior lattice points is
fixed. For this, let L(d, i) be the maximal number of lattice points of a
d-dimensional lattice polytope with exeactly i interior lattice points.
For integers m1 , . . . , md we consider the following simplex
X xi
S(dm1 ,...,md ) := x ∈ Rd | xi ≥ 0 , ≤1 .
mi
This is a lattice simplex with vertices 1 and mi ei for the unit basis vectors
ei . Hence, it is a lattice simplex. See Figure 4.16 for an example.
We compute its volume and number of lattice points for a particular
choice of the parameters mi . We define a sequence (aj )j ≥ 1 via Fig. 4.16: S(d2,3,12)
j−1
Y
a1 : = 2 and aj := al + 1 for j ≥ 2
l =1
We estimate the number of lattice points. For this, let R be the ridge of
Exercise 4.20 Sid defined by x1 = x2 = 0.
Exercise 4.21 It follows from Exercise 4.20 that L(S + t) ≤ L(S ) with equality if
Exercise 4.22 t ∈ Zd . By Exercise 4.21 know that
Z
vol(R) = |(R + t) ∩ Zd−2 | dt
Cd−2
Z
= L(R + t) dt ≤ L(R) ≤ L(Sid ) .
Cd−2
Thus
d(d − 1)
L(Sid ≥ vol(R) ≥ vol(Sid ) (4.6)
2·3
Proof. For the special cases in dimensions 3 and 4 in (4.7) see Exer-
cise 4.23.
We know that the volume of Sid is
i+1
vol(Sid ) = (ad − 1)2
d!
by (4.5). Using Lemma 4.40(2) we estimate
d−1−a
ad − 1 ≥ 22
from which the lower bound on the voume follows. The bound for the
lattice points now follows from this and (4.6). u
t
Exercise 4.23
Every lattice polytope has a triangulation which uses all the lattice points.
Its simplices have the property that the only lattice points they contain
are their vertices.
Definition 4.42 (empty) A lattice polytope P is empty if the vertices
of P are the only lattice points in P .
Empty simplices in this sense can be regarded as the fundamental building
blocks of lattice polytope theory. Because of its number theoretic nature,
the study of these objects is in general very hard. In this section, we will
present what we know (in low dimension) and what we do not yet know
(in higher dimensions).
Let us start with the simplest cases d = 1 and d = 2. There is
(up to equivalence) only one empty segment, [0, 1]. By Pick’s Formula
(Theorem 1.8), an empty triangle must be unimodular as well. The
only other empty polygon is the unit square [0, 1]2 up to unimodular
equivalence (Exercise 4.24). The situation is significantly more subtle in
dimension ≥ 3. The Reeve simplices Rd (m) that we have seen in (1.1)
are examples of empty lattice simplices in any dimension ≥ 3 of arbitrary
large normalized volume m. Exercise 4.24
Surprisingly, in dimension 3 an astonishing and completely not obvi-
ous result still holds.
Theorem 4.43 (Howe (in Scarf 1985)) If P ⊂ R3 is an empty lat-
tice polytope in the lattice Z3 , then widthZ3 (P ) = 1.
Here widthZ3 (P ) is the lattice width defined in Definition 4.21. We will
only prove the theorem for the crucial case of an empty tetrahedron. Here
the previous result takes the following form.
Note that the first coordinate is always 0 or 1 on the vertices of Tpq , hence,
the lattice width is one. By now, there are several proofs known (White
1964 [60], Morrison & Stevens 1984 [40], Howe (in Scarf 1985) [47], Sebö
2 The history of these proofs is 1999 [52])2 Before we prove these two theorems we need the following
interesting useful observation by Scarf which is true in arbitrary dimensions, not
just in dimension three. For this, let us define the following function. For
a ∈ Zd with D := dj=1 aj − 1 > 0 define
P
d
X aj h
f : {1, . . . , D − 1} → Z, h 7→ .
D
j =1
h D−h aj if D divides aj h
ai + ai =
D D aj + 1 if D does not divide aj h
Proof (of the Theorem of White (Theorem 4.44)). Our proof is based
on an argument by Scarf, and proceeds in two steps.
(1) We show first that if T is empty, then T is, for some a1 , a2 , a3 ≥ 0,
equivalent to
1 0 0 a1
conv 0 1 0 a2
0 0 0 a3
(2) Then we use Scarf’s criterion (Lemma 4.45) together with the lower
bound a1 , a2 , a3 ≥ 2 to obtain a contradiction.
Step 1 (Standard form): We can assume that
0 v11 v12 v13
T = conv 0 0 v22 v23
0 0 0 v33
(cf. Definition 2.64, note that here we look at the transposed form). The
triangle spanned by the first three vertices is empty and hence unimodular
by Pick’s Formula (Theorem 1.8). This implies v11 = v22 = 1, v12 = 0.
Now,
0 0 1 0 v13 0 1 0 a1
T =∼ T + 0 = conv 0 0 1 v23 =
∼
conv 0 0 1 a2
1 1 1 1 v33 + 1 1 0 0 a3
u
t
Exercise 4.28
Exercise 4.29 The reader may wonder whether the previous result also holds in
higher dimension. Unfortunately, it fails already in dimension 4, as we
will show now. For this, let us define a lattice simplex S (a, b, c, d) ⊂ R4
as the convex hull of the columns of the following matrix:
1000a
0 1 0 0 b
0 0 1 0 c
0001d
The following theorem states that this is the only way to get hollow
polytopes – up to finitely many exceptions.
Theorem 4.47 There are only finitely many (equivalence classes of)
hollow lattice d-polytopes which do not allow an integral projection onto
a hollow lattice (d − 1)-polytope.
Step 2:
Lemma 4.50 Let P ⊂ Rd be a lattice polytope. If |P ∩ Zd | > N d , then
P contains a segment of length N .
We can now easily deduce the flatness theorem for hollow lattice polytopes.
For this we simply note that under lattice projection the width cannot
decrease (Exercise 4.31). Exercise 4.31
4.8 Problems
Exercise 4.32
4.1. Let K ⊆ Rd be a centrally symmetric convex body with int (K ) ∩
included on page 100
Zd = {0} and nvolZd (K ) = 2d . Show that K is a polytope and
each facet of K contains at least one lattice point in its relative
interior.
Hint: For any lattice point x choose a half space Hx containing both
x and K. Let Sx := Hx ∩ −Hx . Now consider the intersection
of all Sx and prove that this satisfies the assumptions of
Minkowski’s First Theorem (Corollary 4.3).
included on page 100
kxk∞ ≤ (det Λ) /d ], .
1
4.7. Prove
√
(1) µ(Zd ) = d/2
(2) µ(D3 ) = 1
√
(3) µ(Dn ) = n/2 for n ≥ 4
(4) µ(E8 ) = 1
included on page 108
4.8. Show that in the definition of lattice width we can replace the
infimum with a minimum for a full-dimensional convex body K.
Thus, the width is strictly positive.
4.9. Show that the lattice width of a low dimensional convex body is 0.
K 0 ⊆ K ⊆ µK 0 + v
ca(K; w ) ≤ µ ca(K 0 ; w ) + µ − 1 .
included on page 114
4.20. Let Sid be the Sylvester simplex, l its number of lattice points and
x ∈ R. Show that, if the translate Sid has l lattices points then
x ∈ Zd .
Hint: Show that if one of the facets parallel to a coordinate hyper-
plane does not contain a lattice point then you can translate the
simplex so that it does, while the number of lattice points does
not decrease.
included on page 116
4.21. Let P be a lattice polytope and Cd the unit cube. Show that
Z
vol(P ) = |(P + t) ∩ Zn | dt .
Cd
included on page 116
4.22. recursively we define
n−1
Y
t1 := 2, tn : = 1 + tj for n ≥ 2.
j =1
Show that
(1) tn = t2n−1 − tn−1 + 1 for n ≥ 2.
Pn−1 1 1
(2) j =1 tj = 1 − tn −1 for n ≥ 2.
n−1 n−2
(3) 22 ≥ tn ≥ 22 .
4.22. let d, i ∈ Z≥1 . Define the d-dimensional lattice simplex Sd,i mit
vertices v0 , . . . , vd :
Show that
4.24. Show that ∆2 and [0, 1]2 are up to unimodular equivalence the
only empty lattice polygons.
included on page 119
4.28. Show, using the table in the proof of the Theorem of White (Theo-
rem 4.44) that T is not empty for (a1 , a2 , a3 ) = (3, 4, 5).
included on page 122
4.29. Use the proof of the Theorem of White (Theorem 4.44) to show
the following: Let P be an empty lattice tetrahedron with one
vertex in the origin. Then the subgroup (or sublattice) of Z3 that
is spanned by the vertices of P is a cyclic quotiont group.
included on page 122
4.31. Prove that projection cannot decrease the width of a convex body.
included on page 126
= (d+3 2) + (d+3 1)
Pd 2
4.32. Show that i=0 k
Hint: You should try to do this via Ehrhart Theory. Consider a
(d − 1)-fold pyramid over a square.
5.1.1 Introduction
(1) The sum over the coefficients of its h? -polynomial equals the (nor-
malized) volume of P . Since, by Theorem 3.32, all the coefficients
are non-negative integers, the question about the number of lattice
polytopes with the same h? -polynomial is essentially equivalent to
the problem of finding all lattice polytopes with given normalized
volume and degree (of the h? -polynomial).
(2) The simplest h? -polynomial consists only of the constant coefficient
1. In this case, P has normalized volume 1, so it is a unimodular
simplex (of arbitrary dimension).
(3) More generally, lattice pyramids with base P have the same h? -
Fig. 5.2: Unimodular simplices polynomial as P , see Exercise 3.30. Therefore, it makes sense to
in dimensions 0, 1, 2 and 3 consider lattice polytopes ‘modulo’ lattice pyramid constructions.
For instance, one may think of unimodular simplices as (successive)
lattice pyramids over a point, see Figure 5.2. In this respect, the
degree deg(P ) seems to be a more natural invariant to consider than
the dimension of a lattice polytope.
(4) It is also insightful to recall that not only the volume but any
coefficient of the h? -polynomial behaves monotonically with respect
to inclusions: Q ⊆ P =⇒ h?i (Q) ≤ h?i (P ) for any i. In particular,
deg(Q) ≤ deg(P ).
Here is the main result of this section. Let us define the function
f (c, s) := c(2s + 1) + 4s − 2 .
d > f (|V(P )| − d − 1, s)
supp m := {i : λi 6= 0}
The proof of the following observation is left as Exercise 5.1. Exercise 5.1
| supp P | ≤ 4s − 1 .
Let us first bound the support of one lattice point in Π(P ). {1} × P
Lemma 5.4 Let m ∈ Π(P ) ∩ Zd+1 . Then
supp m ≤ 2s . m?
Proof. We define
X X
m? := ( 1 − λ i ) vi = vi − m .
i∈supp m i∈supp m Fig. 5.3: The central symmetry
m ↔ m? about the ‘green’ center of
Π(P )
Here is another proof, which uses the notion of the codegree (see Defini-
tion 3.47 and Corollary 3.48).
In particular,
c ≤ a. (5.1)
Let us write
d
X
mk = λki vi .
i=0
|A| ≤ 2 deg(C ) + 2 .
|A+ | ≥ codeg(C )
= dim(C ) + 1 − deg(C )
= (|A+ | + |A− | − 2) + 1 − deg(C ) .
This implies |A− | ≤ deg(C ) + 1. The same bound also holds for |A+ |. u
t
Proof (Proof of Theorem 5.1 in the general case). Let c := |V(P ) − d − 1|.
We prove the statement by induction on c. The case c = 0 was dealt with
in the previous section. So, we may assume c > 0, and d > f (c, s) =
c(2s + 1) + 4s − 1. Since P is not a simplex, there exists a vertex v such
Fig. 5.6: The induction step
that P 0 := conv (V(P )\{v}) is d-dimensional (see Exercise 5.3). Note
that |V(P 0 )| − dim(P 0 ) − 1 < c by construction, and deg(P 0 ) ≤ s by the
monotonicity theorem. Hence, the induction hypothesis yields that P 0
is a (possibly successive) lattice pyramid over a lattice polytope B 0 of
dimension
5.2.1 Cayley-Polytopes
x0 + · · · + xt = u
Proof. (1) ⇒ (2): The definition of a Cayley polytope implies the existence
of a surjective lattice homomorphism Zd → Z t , mapping the vertices of
P onto {0, e1 , . . . , et }.
(2) ⇒ (1): This follows also directly from the definition.
(1) ⇒ (3): In Definition 5.8, let b1 , . . . , bd , e1 , . . . , et be the stan-
dard basis of Zd , and b∗1 , . . . , b∗d , e∗1 , . . . , e∗t its dual lattice basis. As
ϕ : Zd → Zt+1
m 7→ (h x0 , m i, . . . , h xt , m i)
codeg(P0 ∗ · · · ∗ Pt ) ≥ t + 1 .
The following result (which we state without proof) shows that any
lattice polytope in high dimensions of small degree decomposes into
small-dimensional lattice polytopes.
Theorem 5.10 Any lattice polytope of degree s is a Cayley polytope of
length ≥ d + 1 − (s2 + 19s − 4)/2 (equivalently, P is a Cayley polytope
of lattice polytopes in dimension ≤ (s2 + 19s − 4)/2).
It is conjectured that (s2 + 19s − 4)/2 can be replaced by a linear function
in s. Here is an application in Ehrhart theory.
Theorem 5.11 There exist only finitely many h∗ -polynomials of lattice
polytopes of degree s and with given leading coefficient h∗s .
This result follows immediately from the following statement combined
with Theorem 5.10. Here, let V (q, k ) be given as in Theorem 4.25.
nvolZd (P ) ≤ N ! N N V (N , h∗s )N .
π −1 (λ) ∩ P = λ × (λ0 P0 + · · · + λc Pc ) .
Let ωij be the width of Pj with respect to the i-th coordinate on Rq , the
difference between the maximum and the minimum of the i-th coordinates
of points in Pj . By Theorem 2.82 there is a choice of coordinates on Rq
such that λ0 P0 + · · · + λc Pc is contained in the standard cube [0, C ]q
with the side length C being equal to q times the normalized volume of
λ0 P0 + · · · + λc Pc . Now, we may choose by Theorem 4.25
C = q V (q, k ).
Since widths are additive and each λj is a positive integer, it follows that
ωi0 + · · · + ωic ≤ C, for 1 ≤ i ≤ q.
Now P projects onto S, so we can express the normalized volume of
P as an integral
Z
vol π −1 (λ) ∩ P dλ,
nvolZd (P ) = d! ·
S
where the sum is over (j1 , . . . , jq ) ∈ {0, . . . , c}q . Now it follows from
Hölder’s inequality that the integral over S of the monomial λj1 · · · λjq
is bounded above by the integral of λq1 , and a straightforward induction
shows that
Z
λq1 dλ = q!/d!.
S
The sum on the right hand side may be written as qi=1 (ωi0 + · · · + ωic ),
Q
∼ conv(w1 , . . . , wl )
where l is the number of vertices of P such that P =
and the following property holds:
∼ Q ⇐⇒ N(P ) = N(Q).
P =
Hence, any total ordering on the set of integral matrices induces a total
ordering on all unimodular equivalence classes of d-dimensional lattice
polytopes.
We will describe the various ingredients in this construction step by
step. They will also play an important role in the classification algorithm
of reflexive polytopes which is described in § 7.5.2.
1 We remark that there are several choices of such an implementation, and we do
not claim that ours is precisely the one used in PALP.
H = {x ∈ Rd : h ηF , x i = b}
∼Q
P = =⇒ VPM(P ) = VPM(Q) .
It is important to observe that the converse does not hold. The two
non-isomorphic lattice polygons in Figure 5.7 have the same vertex-
pairing-matrix
300
0 3 0
003
We can find an ordering of the set of vertices and facets of P such that
Of course, there may be several choices how to do so. Let us make this
precise. We define Ord(P ) as the set of bijections σ : {1, . . . , l} → V(P )
such that there exists some bijection π : {1, . . . , k} → F (P ) with
We will need the Hermite normal form theorem for integral matrices
in its full generality. For the proof we refer to Exercise ??? .
Example 5.15 Let us consider the triangle P on the left of Figure 5.8.
Here,
600
VPM(P ) = 0 3 0
002
Therefore,
!
1 1 −1
Hσ =
0 3 −2
Example 5.16 Let us consider the triangle P on the right of Figure 5.8.
Here,
300
VPM(P ) = 0 3 0
001
and
!
0 0 1
Aσ1 =
2 −1 2
Therefore,
!
1 −2 0
Hσ1 =
0 0 1
and
!
2 −1 0
Hσ2 =
0 0 1
In order to get the desired affine normal form we apply (5.14) to the
d × l-matrix
A0 : = 0 v 1 − v 0 · · · v l − v 0 .
Example 5.17 Let us continue with Example 5.15. For the unique σ ∈
Ord(P ) we have !
0 030
( Aσ ) =
002
This is already in Hermite normal form, hence
!
0 030
Hσ =
002
See [23]
5.3 Problems
Exercise 5.5
5.1. Prove Lemma 5.3.
included on page 133
5.3. Prove that for any d-dimensional polytope P which is not a simplex
there exists a vertex such that the convex hull of the other vertices
is full-dimensional.
included on page 141
5.4. pb:affine-to-cone
included on page 145
In Section 4.1 we have shown that any lattice has a non-zero vector of
√
length at most d det Λ. However, the proof was not constructive and we
have left open the question how one can actually compute such a vector in
polynomial time. We will solve this problem in Section 6.3. This algorithm
hinges on the fact that we need to compute a reduced basis efficiently. We
will give an algorithm for one special reduced basis, the δ-reduced basis
using the LLL-method in Section 6.4. This will finally give the complete
algorithm for counting lattice points. Using this basis algorithm we can
also solve the problem of computing a rational generating function for
Lecture Notes Lattice Polytopes (draft of June 28, 2021)
Smith normal forms using Kannan and Bachem [29]. See also Gerald
Jäger, A new algorithm for computing the Smith normal form and its
implementation on parallel machines
Index(C ) := #(Π(v1 , . . . , vd ) ∩ Zd )
= | det(v1 , . . . , vd )|
= vol Π(v1 , . . . , vd )
vol(K ) = 2d .
and the right hand side is strictly less than Index C if Index C ≥ 2. As
the index is an integral number we see that the index actually drops by
at least one.
We define a corresponding sign function to make a signed subdivision
of C with the cones Cj . For 1 ≤ j ≤ d let
0 if dim Cj < d
εj := 1 if det(v1 , . . . , vd ) · det(v1 , . . . , vj−1 , w, vj +1 , . . . , vd ) > 0
−1 otherwise.
and we need to check for which n this number drops below 2 (recall that
the index is integral, so it must be 1). We take the binary logarithm twice
to solve this expression for n:
d−1 n d−1 n
lg2 lg2 (Index C ) d = lg2 lg2 (Index C )
d
d−1
= n lg2 + lg2 lg2 (Index C )
d
d
= −n lg2 + lg2 lg2 (Index C ) .
d−1
(6.2)
Hence, if
lg2 lg2 (Index C )
n > = O (d lg2 lg2 Index C ) ,
d
lg2 d−1
i.e.
Index D = 1 .
2
= (lg2 Index C )M d lg2 d
2
= (lg2 Index C )O (d lg2 d)
.
X tai
GP (t) := εi ,
(1 − tvi1 ) · · · (1 − tvisi )
i∈I
The map
d
Y
λ 7−→ h m ( λ ) , vj i
j =1
Now consider (we look at evaluation at 1, the case for (t, 1) is similar)
lim GP (x(r )) .
r→0
Let
αi := h m, ai i νij := h m, vij i .
Then
X eαi r
GP (x(r )) = ε i Q si uij r ) ,
i∈I j =1 (1 − e
and the summands are all rational functions in one variable r which
are defined for all r > 0. We want to compute the constant term of the
Laurent expansion of all summands at r = 0. Now consider a single such
fraction. We can transform it to obtain
si
e αi r 1 αi r Y r
Qs i = e . (6.3)
j =1 (1 − e
uij r ) r s i 1 − euij r
j =1
Let ci be the coefficient of rsi (note that (6.3) has an additional factor of
1
rsi , so for this product ci is the constant coefficient). We sum them up
with the given signes to obtain
X
c := εi ci .
i∈I
one may obtain a closed formula for the evaluation of GC (1), see e.g. [18,
Thm. 7.2.1]. However, this requires us to evaluate the Todd polynomials.
You can use GP (t) also to solve linear programs. If you want to
maximize over a functional c ∈ Zd , then you can just substitute
t = (z c1 , z c2 , . . . , z cd ). The highest degree of a monomial in the result is
the optimal solution.
kuk ≤ γλ1 .
In fact, our algorithm for (SVP) is based on a solution for the approximate
problem by constructing a reduced basis. We show that the coefficients of
a shortest vector in such a basis are bounded by a constant depending
on the dimension only. Hence, we can solve the shortest vector problem
by enumerating over all possible coefficients (in fixed dimension).
In this section we explore the difficulties in the computation of a short
basis vector, introduce reduced bases and show how such bases bound
the size of coefficients for a shortest vector. This proves that, in fixed
dimension and given a reduced basis, we can solve (SVP) in polynomial
time.
wk := vk − πk−1 (vk )
k−1
X h vk , w j i
= vk − λjk wj with λjk := .
kwj k2
j =1
Fig. 6.3: Orthogonal projec-
tion of a lattice generator See also Figure 6.3. This is the Gram-Schmidt-orthogonalization of the
lattice basis v1 , . . . , vd . So in particular we have
! !
2 1
Example 6.4 Let v1 := , v2 : = . This is a basis of R2 with
1 2
Gram-Schmidt-orthogonalization
!
−3/5
w1 = v1 and w2 = = −4/5v1 + v2 ,
6/5
v2
see Figure 6.5. w2
v1 = w 1
We can write the original lattice basis in terms of the Gram-Schmidt
basis as Fig. 6.5: The lattice basis of Exam-
ple 6.4
k−1
X
vk = w k + λjk wj for 1 ≤ k ≤ d . (6.4)
j =1
The Gram-Schmidt vectors give a lower bound for the length of a shortest
vector of the lattice.
Proposition 6.5 Let Λ be a d-dimensional lattice with basis v1 , v2 , . . . , vd
and Gram-Schmidt-orthogonalization w1 , w2 , . . . , wd . Then
We have seen in the previous section that reduced bases allow the compu-
tation of a shortest vector, and thus count lattice points in a polytope, in
polynomial time. Yet, so far we don’t even know that such bases exist for
our lattice Λ. We address this question in the following by introducing
a very special type of reduced basis together with a polynomial time
algorithm first described by Arjen Lenstra, Hendrik Lenstra and László
Lovász in 1982 [38]. The algorithm is known as as the LLL-algorithm in
honour of the three authors.
Then λ13 violates condition (1) of the definition, and v1 and v2 violate
condition (2), as d(v1 , V0 )2 = 5 and d(v1 , V0 )2 = 9/5 but 3/4 · 5 > 9/5.
However, the basis
0 2 1
v10 := 1 v20 := 1 v30 := −2 ,
1 0 2
We need some preparations for the proof of this theorem. Consider the
representation of our given basis in (6.4). Assume that for some pair of
indices i < k the absolute value of the coefficient λik is larger than 1/2.
Then there is a unique µik ∈ R and aik ∈ Z such that
We set
This leaves the subspaces Vj invariant and v10 , . . . , vj0 is still a basis of
the lattice Λj for 1 ≤ j ≤ d. The new basis has the same Gram-Schmidt
k−1
X
vk0 = wk + λjk wj − aik vi .
j =1
So we can write
i
X k−1
X
vk0 = wk + (λjk − aik ηj )wj + λjk wj ,
j =1 j =i+1
So the coefficient λ0ik := λik − ηi is weakly reduced, and the new coeffi-
cients for vk0 are
We want to use this to make all coefficients weakly reduced, that is, to
make the lattice basis weakly reduced. However, we have to be careful,
as making λ0ik weakly reduced also affects λjk for j < i. So, for fixed
k we apply this procedure to all λik with i in decreasing order. Doing
this for all 1 ≤ k ≤ d gives us a weakly reduced basis. This procedure is
formalized in Algorithm 6.2.
There are (d2) coefficients, and reducing λik with the above procedure
we have to touch at most i ≤ d other coefficients, so this process terminates
after at most O (d3 ) steps. We summarize this in the following proposition.
Proposition 6.13 Any lattice Λ has a weakly reduced basis. More pre-
cisely, we can transform any basis into a weakly reduced one in O (d3 )
steps using Algorithm 6.2. u
t
We repeat these two steps until we reach a reduced basis. This is formalized
in Algorithm 6.3.
while
wi0 := wi for i 6= j, j + 1
j−1
X
wj0 := vj +1 − λi,j +1 wi = vj +1 − πj−1 (vj +1 )
i=1
j−1
X h vj , wj0 i
wj0 +1 := vj − λij wi − wj0
kwj0 k2
i=1
h vj , wj0 i
= vj − πj−1 (vj ) − wj0
kwj0 k2
for some β < δ < 1. Hence, 1/β kwj0 +1 k2 = kwj +1 k2 as for all other i we
have wi0 = wi and
d
Y d
Y
det Λ = kwi k = kwi0 k
i=1 i=1
is invariant. This implies that
√
det Λ0j < δ det Λj and det Λ0i = det Λi for i 6= j ,
so also
√
D (v10 , . . . , vd0 ) = δD (v1 , . . . , vd ) < D (v1 , . . . , vd ) . (6.7)
Hence, in each iteration, D (v1 , . . . , vd ) drops be at least a factor of
√
δ < 1, so Algorithm 6.3 runs in polynomial time if D (v1 , . . . , vd ) has a
polynomial lower bound in terms of the dimension. We provide such a
bound with the following lemma.
Lemma 6.14 Let Λ ⊂ Rd be a lattice with lattice basis v1 , . . . , vd and
first successive minimum λ1 := minu∈Λ\{0} (kuk) introduced in Defini-
tion 4.6. Then
d
Y
j − /2 .
d(d+1)/2 j
D (v1 , . . . , vd ) ≥ λ1
j =1
for coefficients −1/2 ≤ λkj ≤ 1/2. Taking the norm and using that the
scalar product of any two of the wi ’s is 0 implies
j−1 j−1
X 1X
kvj k2 = kwj k2 + λ2jk kwk k2 ≤ kwj k2 + kwk k2 .
4
k =1 k =1
so
1
kwj +1 k2 + kwj k ≥ δ · kwj k2
4
The second inequality follows. u
t
1
We define α := .
δ− 14
so
d−1
kv1 k ≤ α 2 λ−1.
This proves the first item. Taking the product of (6.9) for all j gives
d
d(d−1) d(d−1)
(det Λ)2 .
Y
kv1 k2d ≤ αi−1 kwi k2 = α 2 kw1 k2 · · · kwd k2 = α 2
i=1
u
t
Proof. We compute
j−1 j−1
!
1X 1 X j−k
kvj k 2
≤ kwj k + 2
kwk k2 ≤ kwj k2 1+ 2 ≤ 2j−1 kwj k2
4 4
k =1 k =1
and
d
Y d
Y
kvj k ≤ 2j−1 kwj k2
j =1 j =1
1 d
Y 1 d
= 22 2 j = 1d kwj k ≤ 2 2 2 det Λ .
u
t
d−i
d−1 3
|ηi | ≤ 2 2 α ≤ 3d α for 1 ≤ i ≤ d .
2
D : = D ( v1 , . . . , vd ) ≥ 1 . (6.10)
√
By (6.7), the value of D drops by at least a factor of δ each time we
swap a pair of basis vectors that violate condition Definition 6.10(1), so
we need at most k = d log 11/√δ log De swaps.
Now kwj k ≤ kvj k for 1 ≤ j ≤ d, so we can bound the initial value of
D by
d(d+1)/2
D ≤ max kvj k ,
j
so the right hand side of (6.11) is polynomial in I for δ < 1. More precisely,
we habe that the number of swaps we have to perform in the worst case
satisfies
k = O d2 (log d + s)
Remark 6.20 The additional assumption that all lattice basis vectors
all have integral entries can be removed. We just scale the initial basis
with the common denominator of all entries of the basis vectors.
Remark 6.21 In the form we have discussed the LLL algorithm needs
at most O d5 (log d + s) steps on numbers of size O (d + s). This cor-
Then
\
C = Di∨
i∈I
for δI ∈ {0, ±1}. Observe that any nontrivial union of two dual cones
Di∨ and Dj∨ for i 6= j contains a line, as there is a linear hyperplane H
spanned by some vector u in primal space (the dual of the dual) that
intersects the cones at most in their boundary and has the two cones on
different sides. Then Ru ∈ Dj∨ ∪ Dj∨ . Hence, these serices vanish if we
pass to the generating functions via our map Φ, so that
X
GC (x) = GD ∨ ( x ) ,
i
i∈I
where we have lost all contributions except for the dual maximal cones.
The following observation shows that this is a decomposition into cones
of index 1. You will prove this in Exercise 6.2.
max(h c, x i)
subject to Ax ≤ b
x ∈ Zd .
{x : h c, x i = β} .
for some β ∈ Z.
This may destroy reducedness of the basis, but does not affect the
orthogonality defect. We distinguish two cases
(1) if kvd k ≤ d1 , then we consider the representation
d
X
z = λ i vi
i=1
Hence u ∈ Bd ∩Λ.
(2) Otherwise we have kvd k > d1 . Let
H := lin(v1 , . . . , vd−1 ).
Then H + vd = H + wd , so
[ [
Λ ⊆ H + βvd = H + βwd .
β∈Z β∈Z
Now
normvd d(d−1)
≤ M ≤ 2 2 ,
kwd k
so
d(d−1) 1 − d(d−1)
kwd k ≥ 2− 2 kvd k ≥ 2 2 .
d
Further
d(d−1)
So H + βvd ∩ Λ 6= ∅ for at most 2d2 2 different β’s. Let c :=
wd
kwd k2
. Any u ∈ Λ can be written as u = µwd + h for h ∈ H and
µ ∈ Z. This implies
h c, u i = µ ∈ Z .
d
so that c ∈ Λ? and kck ≤ 1
kwd k
≤ 2O (2 ) . u
t
a + E ⊆ K ⊆ a + dE .
However, the proof of this theorem given in the appendix was not con-
structive. Note first, that for us it is sufficient to find an approximation
of E where we relax the scaling factor on the right hand side. It suffices
to find E such that
a + E ⊆ K ⊆ a + 2dE .
a + E ⊆ K ⊆ a + 2dE . u
t
where the first d comes from the flatness theorem and the rest from the
bound given for ellipsoids. This implies
d
Y k (k−1) 3
T (d) ≤ k·2 2 + 1 ≤ 2O (d ) . u
t
k =1
6.8 Software
6.9 Problems
included on page 168
6.1. Show that we can use a half open decomposition in Barvinok’s
algorithm.
included on page 169
Exercise 7.1 Check Exercises 7.1 to 7.4 for some examples and properties.
Exercise 7.2 It is well-known that the vertices of P ? correspond one-to-one to
Exercise 7.3 facets of P . More precisely, the vertices are the unique inner facet normals
Exercise 7.4
evaluating as −1 on facets. Recall from Chapter 2 that for full-dimensional
lattice polytopes there is a canonical choice for the facet normals as a
primitive lattice vector in the dual lattice. We will in the following
alsways assume that normal vectors are primitive dual lattice vectors.
The most important result is the duality theorem (which holds more
generally for convex bodies containing the 0 in their interior):
P ?? = P
h ηF , x i = −cF ∀x ∈ F
h ηF , x i ≥ −cF ∀x ∈ P
In this case, cF is called (signed) integral distance of the origin from the
facet F . The origin is in the polytope if cF ≥ 0.
Recall that the lattice distance of a lattice point v from a lattice hyperplane
H is given by
|h u, x i − h u, v i| (7.1)
int P ∩ Λ = {w}
no lattice points
Proof. Let F ∈ F (P ). By definition, no lattice point lies strictly be-
cF + 1 cF
tween the hyperplanes aff (F ) and its parallel hyperplane through w.
See Figure 7.1. Therefore, conv(w, F ) ∩ Λ = {w} ∪ (F ∩ Λ). Since
S
P = F ∈F (P ) conv(w, F ), the statement follows. u
t Fig. 7.1: Proof of Proposition 7.3.
É P reflexive,
É P ∗ lattice polytope
É P ∗ reflexive. u
t Fig. 7.2: (Non-)reflexive polygons
This result is illustrated in Proposition 7.1. Coming back to Proposi-
tion 7.3, the reader will prove in Exercise 7.6 that any lattice polygon with
one interior lattice points is also a reflexive polygon (see Exercise 7.14 for
the complete list). However, this is not true in dimension 3 and higher
for cFi ∈ Z≤−1 . If ck ≤ −2, we define P 0 as the set of (x, xd+1 ) ∈ Rd+1
satisfying
É h ηFi , x i ≥ cFi for i = 1, . . . , k − 1
É h ηFk , x i − xd+1 ≥ cFk + 1
É xd+1 ≥ −1
Fig. 7.3: P and P 0 Then f (P 0 ) < f (P ). In particular, iterating this construction yields after
finitely many steps a lattice polytope P 00 such that f (P 00 ) = 0, i.e., P 00
is reflexive. u
t
w + w 0 = a(τ )v .
Proof. Let {v1? , v2? } be the basis dual to the basis {v1 , v2 }. Then the Fig. 7.5: The dual edge has length
vertices of P ? dual to the edges at v2 containing v1 and v3 respectively 2−a
In light of this lemma, Theorem 7.7 follows from the following theorem.
Theorem 7.10 Let Σ be a complete unimodular fan in R2 . Then
X
(3 − a(τ )) = 12 , (7.2)
τ ∈Σ[1]
Proof (of Theorem 7.7). Let P be a reflexive polygon with vertex set V
and boundary points B. Then B is the set of generators of a uniomdular Fig. 7.6: Smooth blow-up of a 2-
dimensional fan
complet fan, and we get
X X X
12 = 3 − a(u) = 1 + 2 − a(u) = |∂ P ∩ Λ| + |∂ P ? ∩ Λ? |
u∈B u∈B u∈V
u
t
The defining property of such a smooth blow-up is the fact that the
new ray % has ray parameter a(%) = 1. As the reader can verify in
Exercise 7.9, these steps preserve the validity of equation (7.2). You will
classify 2-dimensional fans which are minimal with respect to blow ups in
Exercise 7.9 Exercise 7.10. You will consider the 3-dimensional setting in Exercise 7.11.
Exercise 7.10
Lemma 7.12 If the complete unimodular fan Σ in R2 satisfies (7.2),
Exercise 7.11
and Σ0 is a smooth blow-up of Σ, then Σ0 also satisfies (7.2).
rays of Σ in the interior, then the convex hull of the primitive generators
has a vertex in the interior of σ. The corresponding ray falls into case (2)
of Lemma 7.14 and hence must have parameter = 1. u
t
Putting it all together, Lemma 7.12 and Theorem 7.13 imply Theo-
rem 7.10.
This result was first proved by Dimitrios Dais as follows. By [8], a general
anticanonical hypersurface Z in the toric variety associated with P
must be a 2-dimensional Calabi–Yau, i.e., a K3-surface which has Euler
characteristic χ(Z ) = 24. By [17], the above sum computes χ(Z ). For
about a decade, this remained the only proof (apart from exhaustion).
We will provide an elementary proof in the present section.
Sadly, the story does not continue in dimensions ≥ 4. But in dimension
three, we can carry out a similar program as we did in dimension two.
First, we describe parameters for unimodular fans which will replace dual
edge lengths (Exercise 7.12). Exercise 7.12
As in dimension two, let {v1? , v2? , v3? } be the basis dual to the basis
{v1 , v2 , v3 }. Then the vertices of P ? dual to the facets of P containing
{v1 , v2 , v3 } and {v1 , v2 , v30 } are v1? + v2? + v3? and v1? + v2? + (a(τ , v1 ) +
a(τ , v2 ) − 1)v3? , respectively. u
t
We could, again, prove this theorem using a walk in the space of fans.
The invariance under smooth blow-ups is elementary. But connectivity
of the space of fans is a deep theorem, way beyond the scope of these
notes. Luckily, one can deduce the 24 from the 12 by double counting.
Lemma 7.21 Let Σ be a complete unimodular fan in R3 , and let % ∈
Σ[1] with primitive generator v. Then the projection π : R3 → R3 /Rv
maps star(%; Σ) to a complete unimodular fan Σ /%. If τ ∈ star(%; Σ) is
a 2-cone with primitive generators v, v 0 , then the corresponding ray π (τ )
of Σ /% has parameter a(π (τ )) = a(τ , v 0 ).
Proof (Theorem 7.20). Let us first collect what we need. Our fan Σ
gives rise to a triangulation of the 2-sphere with vertex set Σ[1], edge set
Σ[2], and triangle set Σ[3]. As such, we have 3| Σ[1]| − | Σ[2]| = 6 from
Euler’s formula and double counting of edge-triangle-incidences. Also, we
P
have v∈Σ[1] deg v = 2| Σ[2]|, where deg v denotes the number of edges
containing the vertex v. Armed with these formulas we compute
X
(2 − a(τ , v1 ) − a(τ , v2 ))
τ ∈Σ[2]
with primitive
generators v1 ,v2
X X
= (1 − a(τ , w ))
cone(v )∈Σ[1] τ ∈Σ[2]
τ =cone(v,w )
X X
= ((3 − a(τ , w )) − 2)
cone(v )∈Σ[1] τ ∈Σ[2]
τ =cone(v,w )
X
= 12 − 2 deg v = 12| Σ[1]| − 4| Σ[2]| = 24 .
cone(v )∈Σ[1]
Here we have used Theorem 7.10 for the quotient fans Σ /% in the third
equality. u
t
Example 7.22
Fig. 7.12: The hexagon
d = 2: |V(P )| ≤ 6, only attained by the reflexive hexagon H , see Figure 7.12.
d = 3: |V(P )| ≤ 14, only attained by the polytope in Figure 7.11
d = 4: |V(P )| ≤ 36, only attained by H × H .
This question is still wide open. It has been shown to hold for simple
centrally symmetric reflexive polytopes, since this class of reflexive poly-
topes can be completely classified. In the following we will present some
of the techniques used to prove this.
Example 7.24 42
(1) Let P1 = P2 = [−1, 1] be a segment of length 2 around the origin.
Then P1 ◦ P2 is the convex hull of ±e1 , ±e2 . See Figure 7.13.
(2) For P1 = H and P2 = [−1, 1] the free sum P1 ◦ P2 is BiPyr(H ).
See Figure 7.14.
You will examine the relation between products and sums in Exercise 7.13.
Fig. 7.13: The free sum of
It follows directly from the definition that the free-sum construction is
P1 = P2 = [−1, 1]
the dual operation to products:
Exercise 7.13 While the first result follows directly from the construction, the second
one is not obvious. We will leave this as an exercise. We will prove it in
the last chapter (Corollary 8.27). Here is our main theorem. Its proof
will occupy the remainder of this section.
Theorem 7.26 Let P be reflexive and simplicial. Then |V(P )| ≤ 3d,
∼ H ◦...◦H .
and equality holds only if d is even and P = | {z }
d
2
Proof. Assume (1) and (2) do not hold and (3) is wrong. Then duality
yields that there exists a facet F ∈ F (P ) such that −1 > hηF , x + yi ∈ Z.
Hence, −2 ≥ hηF , x + yi = hηF , xi + hηF , yi where hηF , xi, hηF , yi ≥ 1.
This would imply x, y ∈ F , a contradiction.
Now, let F ∈ F (P ) such that −1 = hηF , x + yi = hηF , xi + hηF , yi.
Since hηF , xi, hηF , yi ∈ Z≥−1 , we get either x ∈ F and hηF , yi = 0
or y ∈ F , hηF , xi = 0. Let us assume the first case. We consider the
x + by
pair x + y, y. If x + y ∼ y, we are done. So assume not. Then we get
(x + y ) + y ∈ ∂P ∩ Λ. Hence, since h ηF , y i = 0, we still have x + 2y ∈ F .
Now, we consider the pair x + 2y, y. Since |P ∩ Λ| < ∞, we cannot repeat ..
this argument ad infimum, so there has to exist some b ∈ Z≥1 such that .
x + by ∼ y. u
t
You can use this in Exercise 7.14 to construct the complete list of 16 y x + 2y
reflexive polygons.
Note that even if x, y are vertices, z does not have to be a vertex x+y
again, if the dimension of P is larger than two. This result has many F
applications. As an immediate result we deduce the following constraints x
on the combinatorics of a simplicial reflexive polytope (Exercise 7.15).
−1 0
Corollary 7.29 The diameter of the vertex-edge graph of a simplicial
reflexive polytope is at most three. Exercise 7.14
Exercise 7.15
7.2.4 Vertices between parallel facets
In this section, we use the partial addition of lattice points to deduce the
precise form of vertices that lie between two parallel facets of a simplicial
reflexive polytope.
(1) For 1 ≤ d
v ∈ Fi ⇐⇒ v = mi ⇐⇒ hb∗i , vi < 0.
ηFi = ηF + αi b∗i .
It suffices to check this equality for the vertices of Fi , where the left side
always evaluates to −1: j 6= i: hηF , bj i + αi hb∗i , bj i = −1,
hηF , mi i + αi hb∗i , mi i = −1. Now, we can prove (1) and (2).
lattice basis. u
t
Combining this lemma with the addition property for the lattice points
in the dual reflexive polytope yields our desired result.
Proof. We use the notation in the proof of the previous lemma. Let
(1)
Then i ∈ I =⇒ v = mi .
−F ∈ F (P ) =⇒ Fi ∩ F = ∅
=⇒ ηFi 6∼ η−F = −ηF
ηFi − ηF ∈ ∂P ∗ ∩ Λ∗
Addition
=⇒
=⇒ −1 ≤ hαi∗ b∗i , ±b∗i i = ±αi ∈ Z
αi >0
=⇒ αi = 1
=⇒ hb∗i , vi = hηFi − ηF , vi = hηFi , vi − hηF , vi
| {z } | {z }
=−1 =0
=⇒ hb∗i , vi = −1
Using the same argument for −F shows that hb∗j , vi > 0 =⇒ hb∗j , vi = 1.
u
t
We can now prove Theorem 7.26. The key idea is the following notion
(due to Øbro).
Definition 7.32 (Special Facet) Let F ∈ F (P ) such that v∈V(P ) v ∈
P
HP (F , i) := {v ∈ V(P ) : h ηF , v i = i} ∀i ∈ Z≥−1
Clearly,
|HP (F , 0)| = d .
|HP (F , 1)| ≤ d.
HP (F , 0) = {m1 , . . . , md },
σ : {1, . . . , d} → {1, . . . , d}
b2 k 7→ jk
b1
is a fixed-point free (σ (i) 6= i) involution (σ 2 = 1, jjk = k). We may
assume that this permutation is of the form
b1 − b2
∼ H ◦···◦H .
P = | {z }
d
2
hηF , mi i = 0
Hence, Lemma 7.30(2) finishes the argument. This proves Theorem 7.26.
CP := cone({1} × P ) ⊆ Rd+1
In particular,
Hence, x − w ∈ (k − r )P ∩ Λ, as desired.
(2) ⇒ (3): We want to show that h uF , w i = 1 for all facets F of P .
For this let us define C 0 := cone(w, F 0 ). Choose a lattice basis b1 , . . . , bd+1
such that b1 , . . . , bd ∈ cone(F ) and bd+1 ∈
/ cone(F ). By translating bd+1
Pd
with a multiple of i=1 bd we can assume that bd+1 ∈ C 0 .
Therefore, bd+1 ∈ int CP ∩ Λ̄. In particular, there exists some k ≥ r
such that bd+1 ∈ int({k} × kP ) ∩ Λ̄. By assumption, there exists some
m ∈ (k − r )P ∩ Λ such that bd+1 = w0 + m0 and
h b∗d+1 , F 0 i = 0 = h uF , F 0 i
for the dual lattice vector to bd+1 . As b∗d+1 is primitive, we see b∗d+1 = uF .
Therefore,
1 = h w0 , u0F i = −kβF + h uF , w i .
CP
Hence, kP is reflexive (w.r.t. w).
Let us show the additional last statement in the theorem. Assume (0, 1)
k > r, then P
This is a contradiction. u
t
(1) See Figure 7.17. CP∨ is not a Gorenstein cone, ⇒ P is not a Goren-
stein polytope. r = codeg P = 1, int P ∩ Λ = 2 > 1; h∗0 = 1, h∗1 = 2. Fig. 7.18: The unit square
while
X X
tr b
EhrP (t) = ehrP (k ) tk +r = ehrP (k − r ) tk .
k≥0 k≥r
P ∨ := {x ∈ CP? : huC ? , xi = 1}
P
= conv((−cF , ηF ) : F ∈ F (P ))
Note that this duality is quite subtle. For instance, for r > 1, P ∨ does
not lie in the hyperplane Rd × 1. Thus, it is not intrinsically embedded
in Rd . It is merely given as a d-dimensional polytope in Rd+1 . Moreover,
except for codegree 1, P ∨ is not isomorphic to (rP − w )∗ , as one might
guess at first. For instance, in Example 7.36(2) with r = 2, P ∨ is just
isomorphic to P . See also Exercise 7.21. Exercise 7.21
Let us consider the case of a reflexive polytope P , say, 0 ∈ int(P ).
Then we recover the duality of reflexive polytopes:
The proof of Theorem 7.40 follows now from applying Proposition 5.12
with N = 2s − 1.
Øbro [43]
7.6 Problems
included on page 176
7.1. Show that for ∅ 6= A ⊂ Rd (equipped with some scalar product)
we always have ((A◦ )◦ )◦ = A.
7.7. Take [−1, 1]d , remove one vertex and take the convex hull of the
remaining lattice points. Show that it still contains precisely one
interior lattice point. Is this a reflexive polytope?
included on page 178
7.15. Prove Corollary 7.29, i.e. show that any pair of vertices of a
simplicial reflexive polytope can be connected by at most three
edges.
included on pa
7.17. Show explicitely that for P = conv(0, 3) the dual cone CP∗ is not
Gorenstein.
included on page 191
7.21. Find the dual Gorenstein polytope for the Gorenstein polytope
P = [0, 1]2 and show that it is isomorphic to P .
É unimodular cover
É integer Carathéodory + BGHMW
É IDP
É vectors u, v, w; covectors a, b, c; scalars α, β, γ
É \psubdiv S
É \regsubdiv{\vw}{V} Sw (V )
É \reglift{\vw}{V} lift(w )
É \regfunction{\vw}{V} Ψw
É \pull{\psubdiv}{\vv} pull(S; v )
Here, the lower hull is the polyhedral complex of those facets whose
normal has negative first coordinate. The faces of Sω (A ) are the domains
of linearity of the function Ψω : P → R given by
v 7→ min h : (v, h) ∈ lift(ω ) .
h aF , v i + ωv ≥ αF
h aF , a i + ω (a) − αF > 0
λ h aF 0 , u i + ωu0 − αF 0 + µ h aG , u i + ωu0 − αG = 0 .
λ h aF 0 , v i + ωv0 − αF 0 + µ h aG , v i + ωv0 − αG ≥ 0
Exercise 8.1 The proof of this lemma ist left as Exercise 8.1.
Proof. The dicing cells have width one with respect to all their facets by
construction. Thus, any pulling refinement of the canonical subdivision
will be unimodular. u
t
The goal of this section is to prove the following theorem of Bruns and
Römer.
Theorem 8.18 The h? -vector of a Gorenstein polytope with a regular
unimodular triangulation is unimodal. That is, 1 = h?0 ≤ . . . ≤ h?bs/2c ≥
. . . ≥ h?s .
This theorem and its proof are due to Bruns and Roemer [16]. For
general Gorenstein polytopes the theorem fails, as shown by Payne and
Mustata [41]. However, it is still open whether the following property
might suffice.
Definition 8.19 A lattice d-polytope P ⊆ Rd is said to possess the
integer decomposition property ( IDP) if for every k ∈ Z≥2 and for
every lattice point u ∈ kP ∩ Zd there exist v1 , . . . , vk ∈ P ∩ Zd such that
u = v1 + · · · + vk .
The reflexive cube [−1.1]d has also the origin as a special simplex.
Fig. 8.6: Two special simplices in the
(3) Figure 8.7 shows a special simplex in the tetrahedron with vertices 0, unit cube
e1 , e2 and 2e3 . Note that also the tetrahedron itself is special.
(4) Figure 8.8 shows two special simplices in the bipyramid over a trian-
gle.
(5) A special simplex in the Birkoff polytope for (n × n) − matrices is,
for instance, the simplex spanned by the vertices corresponding to the
permutations matrices for the permutations
The punchline in the proof of Theorem 8.18 will be that we project the
Fig. 8.8: Special simplices in a polytope along a special simplex, and obtain a reflexive polytope with
triangle bipyramid. Note that
there may be more than one such. the same h∗ -vector which inherits a regular unimodular triangulation
from P . The following definition describes a subcomplex of P which will
project bijectively onto the boundary of that reflexive polytope.
Fig. 8.9
Definition 8.24 Let S = conv(v1 , . . . , vr ) ⊆ {1} × P be a special sim-
plex. Denote by Γ (P , S ) the subcomplex of ∂P generated by faces of the
form F1 ∩ . . . ∩ Fr where Fi is a facet of CP with vi 6∈ Fi for i = 1, . . . , r.
cone(F ∪ S ).
Conversely, suppose x ∈ {1} × P so that ω (x) = ri=1 h ai , x i. Set
P
Pr
xΓ := x − j =1 h aj , x ivj . For any a0i ∈ Ai we have
h a0i , xΓ i = h a0i , x i − h ai , x i ≥ 0
v1 , . . . , vr , w1 , . . . , ws
λi = h ai , x i ∈ Z .
Thus,
X
y = x − λi vi ∈ T ∩ Zd+1
i
P
and, as T is unimodular, y = j µj wj for integral µj , 1 ≤ j ≤ s.
(3): Every vertex u is either in S or in Γ as otherwise all facets containing
u would also contain vi for some fixed i. However, all facets containing
vi describe the tangent cone Tvi P which has only one vertex.
(4): This follows immediately from the previous considerations.
(5): Suppose ω 0 ∈ RA induces a triangulation of P which restricts to T
along Γ (P , S ). Then the weights ω + εω 0 for ε > 0 small enough will induce
the triangulation T ? S: by the pertubation Lemma 8.5, the resulting
subdivision is a refinement of Γ (P , S ) ? S (induced by ω) so that every
conv(F ∪ S ) = F ? S is subdivided according to ω 0 . Every subdivision of
this join is the join of S with its restriction to F (cf. Exercise 8.4). u t
Exercise 8.4
8.5 Dilations
One of the first theorems about unimodular triangulations was proved
in the early days of toric geometry by Knudsen, Mumford, and Water-
Theorem 8.28 ([30]) There is a factor c = c(P ) ∈ Z>0 such that the
dilation c · P admits a regular unimodular triangulation.
Proof (Proof of Theorem 8.28). The theorem is true for lattice polyhedral
complexes: every cell F is a lattice polytope in its own lattice ΛF , and
these lattices are compatible along intersections. In fact, the additional
flexibility offered by this structure is used in the proof. Every triangulation
of P carries two distinguished lattice structures: the one given by the
embedding P ⊂ Rd on the one hand, and the one which declares every
simplex to be unimodular on the other.
Starting from a full triangulation of P , the proof proceeds by in-
duction on the maximal normalized volume V of a cell. If V is a prime
number, the different cells of volume V do not interfere. They can be
subdivided independently. But if V is composite, then this very fact
is used to interpolate between the unimodular lattice structure and a
multiple of the given one. The two cases of the induction step are treated
in Lemmas 8.29 and 8.33 below. The proofs occupy the remainder of this
section. u
t
Lemma 8.29 Let V be a composite integer, and suppose that for every
lattice simplicial complex S all whose cells have volume less than V there
is a factor c ∈ Z>0 such that cS has a unimodular triangulation.
Then the same is true for all lattice simplicial complexes all whose
cells have volume no more than V .
Corollary:
Lemma 8.33 Let V be a prime number, and suppose that for every
lattice simplicial complex S all whose cells have volume less than V there
is a factor c ∈ Z>0 such that cS has a unimodular triangulation.
Then the same is true for all lattice simplicial complexes all whose
cells have volume no more than V .
8.6 Problems
included on page 206
8.1. Prove Lemma 8.15
included on pa
8.6. Use the methods of proof of Theorem 8.18 to show that the pro-
jection of a Gorenstein polytope of codegree r along a special
simplex of dimension r − 1 yields a reflexive polytope with the
same h? -polynomial.
A.2 Ellipsoids
E : = E (T , t) : = T (B ) + t
See Figure A.1 for an example. We can write the ellipsoid explicitely as
n o
E = x ∈ Rd | h T −1 (x − t), T −1 (x − t) i ≤ 1
n o
= x ∈ Rd | h Q ( x − t ) , x − t i ≤ 1
(T , a) 7−→ | det T |
L := conv(Bd ∪{z}) ⊆ K .
(u − ε)2 v2
1 = 2
+ 2 (A.6)
a b
Now we want to determine the particular tangent to the ellipsoid that
also passes through me1 and touches the unit ball, i.e. the boundary
segment of L added in the convex hull of Bd with me1 . This line touches
the unit ball in a point (p, q ) and can thus also be written as
1 u − ε b2
−√ = − 2 .
m2 − 1 a v
Squaring and first using (A.6) and then the first equation of (A.8) we
obtain
−1
(u − ε)2 2 (u − ε)2
1
= b 1−
m2 − 1 a4 a2
−1
a2
1 2
= b 1 −
(m − ε)2 (m − ε)2
(m − ε)2 − (1 + ε)2
b2 = .
m2 − 1
Now let us return to the ellipsoid Fd in dimension d. Its volume is
Now
(d−1)/2
(m − ε)2 − (1 + ε)2
d−1
ab = (1 + ε)
m2 − 1
A.3 Problems
included on page 219
A.1. Let Cd be the cube defined by |xi | ≤ 1. Prove that the maximum
volume ellipsoid is the unit ball.
included on page 219
4.11. (1) Let A, B and C be the vertices of the triangle realising sbc(2, i)
for an interior point L and assume the smallest coefficient is
at vertex C. We can transform the triangle so that A is the
origin and B is on the x-axis and C = (c1 , c2 ) is in the positive
orthant. Then smallestbary (2, i) = c12 , so that sbc(2, i) is
realized by a triangle with maximal height and i interior lattice
points.
This is obtained if B = (2, 0) and c1 = 0. In this case c2 =
2i + 2, so sbc(2, i) = 2i1+2 .
(2)
(3)
[11] Ulrich Betke and Peter McMullen. Lattice points in lattice polytopes.
Monatsh. Math., 99:4 (1985), pp. 253–265. doi: 10 . 1007 / BF01312545
(cit. on p. 80)
[12] Louis J. Billera and Carl W. Lee. Sufficiency of McMullen’s conditions
for f -vectors of simplicial polytopes. Bull. Amer. Math. Soc. (N.S.), 2:1
(1980), pp. 181–185 (cit. on p. 29)
[13] Alexandr A. Borisov. Convex lattice polytopes and cones with few lat-
tice points inside, from a birational geometry viewpoint. arXiv : math /
0001109[math.AG]. 2000 (cit. on p. 109)
[14] Alexandr A. Borisov and Lev A. Borisov. Singular toric Fano varieties. Mat.
Sb., 183:2 (1992), pp. 134–141. doi: 10.1070/SM1993v075n01ABEH003385.
url: http://dx.doi.org/10.1070/SM1993v075n01ABEH003385 (cit. on
p. 109)
[15] Winfried Bruns and Joseph Gubeladze. Polytopes, Rings, and K-Theory.
Monographs in Mathematics. XIV, 461 p. 52 illus. Springer-Verlag, 2009
(cit. on p. 211)
[16] Winfried Bruns and Tim Römer. h-vectors of Gorenstein polytopes. J.
Combin. Theory Ser. A, 114:1 (2007), pp. 65–76. doi: 10.1016/j.jcta.
2006.03.003. url: http://dx.doi.org/10.1016/j.jcta.2006.03.003
(cit. on p. 206)
[17] Vladimir I. Danilov and Askol’d G. Khovanskii. Newton polyhedra and an
algorithm for computing Hodge–Deligne numbers. Math. USSR Izvestiya,
29:2 (1987), pp. 279–298 (cit. on p. 181)
[18] Jesús A. De Loera, Raymond Hemmecke, and Matthias Köppe. Algebraic
and geometric ideas in the theory of discrete optimization. MOS-SIAM
Series on Optimization (vol. 14). Society for Industrial and Applied Math-
ematics (SIAM) (Philadelphia, PA), 2013, xx+322 pages (cit. on pp. 75,
155)
[19] Jesus deLoera, Francisco Santos, and Jörg Rambau. Triangulations. Algo-
rithms and Computation in Mathematics (vol. 25). Springer, 2010 (cit. on
pp. 32, 200)
[20] Jan Draisma, Tyrrell B. McAllister, and Benjamin Nill. Lattice width
directions and Minkowski’s 3d -theorem (Jan. 2009). eprint: 0901.1375
(cit. on p. 104)
[21] Robert M. Erdahl and Sergej S. Ryshkov. On lattice dicing. English. Eur.
J. Comb., 15:5 (1994), pp. 459–481. doi: 10.1006/eujc.1994.1049 (cit. on
p. 205)
[22] Israel M. Gel0 fand, Michael M. Kapranov, and Andrei V. Zelevinsky. Dis-
criminants, resultants, and multidimensional determinants. Mathematics:
Theory & Applications. Birkhäuser Boston Inc. (Boston, MA), 1994, x+523
pages. doi: 10.1007/978-0-8176-4771-1 (cit. on p. 200)
[23] Roland Grinis and Alexander Kasprzyk. Normal forms of convex lattice
polytopes. Jan. 2013. arXiv: 1301.6641 [math.CO] (cit. on p. 145)
[24] Martin Grötschel, László Lovász, and Alexander Schrijver. Geometric
algorithms and combinatorial optimization. Second. Algorithms and Com-
binatorics (vol. 2). Springer-Verlag (Berlin), 1993, xii+362 pages (cit. on
p. 102)
[25] Christian Haase and Günter M. Ziegler. On the maximal width of empty
lattice simplices. Eur. J. Comb., 21:1 (2000), pp. 111–119 (cit. on p. 122)
[26] Martin Henk. Successive minima and lattice points. Rend. Circ. Mat.
Palermo (2) Suppl.,: 70, part I (2002). IV International Conference in
“Stochastic Geometry, Convex Bodies, Empirical Measures & Applications
to Engineering Science”, Vol. I (Tropea, 2001), pp. 377–384. eprint: math.
MG/0204158 (cit. on p. 103)
[27] Douglas Hensley. Lattice vertex polytopes with interior lattice points. Pacific
J. Math., 105:1 (1983), pp. 183–191. doi: 10.2140/pjm.1983.105.183
(cit. on p. 115)
[28] Lutz Hille and Harald Skarke. Reflexive polytopes in dimension 2 and
certain relations in SL2 (Z). English. J. Algebra Appl., 1:2 (2002), pp. 159–
173. doi: 10.1142/S0219498802000124 (cit. on p. 179)
[29] Ravindran Kannan and Achim Bachem. Polynomial algorithms for com-
puting the Smith and Hermite normal forms of an integer matrix. SIAM
J. Comput., 8:4 (1979), pp. 499–507. doi: 10.1137/0208040. url: https:
//doi.org/10.1137/0208040 (cit. on p. 148)
[30] George R. Kempf, Finn F. Knudsen, David Mumford, and Bernard Saint–
Donat. Toroidal Embeddings I. Lecture Notes in Mathematics (vol. 339).
Springer–Verlag, 1973 (cit. on p. 211)
[31] Matthias Köppe and Sven Verdoolaege. Computing Parametric Rational
Generating Functions with a Primal Barvinok Algorithm. Electronic journal
of Combinatorics, 15 (2008) (cit. on p. 74)
[32] Maximilian Kreuzer and Harald Skarke. Classification of reflexive polyhedra
in three dimensions. Adv. Theor. Math. Phys., 2:4 (1998), pp. 853–871
(cit. on p. 140)
[33] Maximilian Kreuzer and Harald Skarke. Complete classification of reflex-
ive polyhedra in four dimensions. Adv. Theor. Math. Phys., 4:6 (2000),
pp. 1209–1230 (cit. on pp. 140, 194)
[34] Maximilian Kreuzer and Harald Skarke. PALP: a package for analysing
lattice polytopes with applications to toric geometry. Comput. Phys. Comm.,
157:1 (2004), pp. 87–106. doi: 10.1016/S0010-4655(03)00491-0. url:
http://dx.doi.org/10.1016/S0010-4655(03)00491-0 (cit. on p. 140)
[35] J. C. Lagarias. Knapsack public key cryptosystems and Diophantine approx-
imation (extended abstract). In: Advances in cryptology (Santa Barbara,
Calif., 1983). Plenum, New York, 1984, pp. 3–23 (cit. on p. 168)
[36] J. C. Lagarias, H. W. Lenstra Jr., and C.-P. Schnorr. Korkin-Zolotarev bases
and successive minima of a lattice and its reciprocal lattice. Combinatorica,
10:4 (1990), pp. 333–348. doi: 10.1007/BF02128669. url: https://doi.
org/10.1007/BF02128669 (cit. on p. 107)
[37] Jeffrey C. Lagarias and Günter M. Ziegler. Bounds for lattice polytopes
containing a fixed number of interior points in a sublattice. English. Can.
J. Math., 43:5 (1991), pp. 1022–1035 (cit. on p. 115)
[38] A. K. Lenstra, H. W. Lenstra Jr., and L. Lovász. Factoring polynomials
with rational coefficients. Math. Ann., 261:4 (1982), pp. 515–534. doi:
10.1007/BF01457454. url: http://dx.doi.org/10.1007/BF01457454
(cit. on pp. 159, 167, 168)
[39] P. McMullen. The numbers of faces of simplicial polytopes. Israel J. Math.,
9 (1971), pp. 559–570. doi: 10.1007/BF02771471. url: http://dx.doi.
org/10.1007/BF02771471 (cit. on p. 29)
— 229 —
Lecture Notes Lattice Polytopes (draft of June 28, 2021)
— 235 —