2022 Book InteractionsWithLatticePolytop
2022 Book InteractionsWithLatticePolytop
Alexander M. Kasprzyk
Benjamin Nill Editors
Interactions
with Lattice
Polytopes
Magdeburg, Germany, September 2017
Springer Proceedings in Mathematics &
Statistics
Volume 386
This book series features volumes composed of selected contributions from
workshops and conferences in all areas of current research in mathematics and
statistics, including data science, operations research and optimization. In addition
to an overall evaluation of the interest, scientific quality, and timeliness of each
proposal at the hands of the publisher, individual contributions are all refereed to the
high quality standards of leading journals in the field. Thus, this series provides the
research community with well-edited, authoritative reports on developments in the
most exciting areas of mathematical and statistical research today.
Mathematics Subject Classification: 14M25, 52B20, 05E10, 06A07, 14J33, 14J45, 14M15, 14M27,
14N15, 52B12, 52B40, 53D05
This Springer imprint is published by the registered company Springer Nature Switzerland AG
The registered company address is: Gewerbestrasse 11, 6330 Cham, Switzerland
Preface
This volume contains original research and survey articles highlighting interdisci-
plinary connections between a diverse range of topics. The common points of interest
are lattice polytopes. Lattice polytopes are fundamental combinatorial objects—
convex polytopes whose vertices have integer coordinates—with many beautiful and
deep connections across modern mathematics. Topics considered include: algebraic
geometry, mirror symmetry, symplectic geometry, discrete geometry, the geometry
of numbers, and algebraic combinatorics.
The study of lattice polytopes continues to open up fertile and unforeseen inter-
actions. In order to enhance this exchange of ideas, the workshop Interactions with
Lattice Polytopes took place 14–16 September, 2017, at the Otto-von-Guericke-
Universität Magdeburg, Germany. There were 15 talks given by world-leading
experts from several different backgrounds, elaborating upon the theme of appli-
cations of lattice polytopes. Many of the presented results can be found in this
volume.
Contributions to this volume contain original as well as expository research arti-
cles that illustrate some of the varied topical approaches and settings where lattice
polytopes play an important role. This volume should be particularly beneficial to
researchers and graduate students interested in learning more about the multifaceted
use of lattice polytopes across a broad range of active research areas.
This book relies deeply on the enthusiasm and engagement of the diverse and
collegial lattice polytope community. We are extremely grateful to the contribu-
tors for their high-quality articles, and to the anonymous referees for their careful
work. We would like to express our gratitude to everyone involved for their
patience and assistance. We are also thankful for logistical and financial support
from: the Otto-von-Guericke-Universität Magdeburg; the Research Training Group
Mathematical Complexity Reduction, funded by the Deutsche Forschungsgemein-
schaft (DFG, German Research Foundation)—314838170, GRK 2297 MathCoRe;
v
vi Preface
and the Engineering and Physical Sciences Research Council (EPSRC) Fellowship
EP/N022513/1.
vii
viii Contents
ix
x Contributors
Gennadiy Averkov
1.1 Introduction
By |X | we denote the cardinality of a finite set X . Let N be the set of all positive
integers and let d ∈ N be the dimension. Elements of Zd are called integral points
or integral vectors. We call a polyhedron P ⊆ Rd integral if P is the convex hull
of P ∩ Zd . Let Aff(Zd ) be the group of affine transformations A : Rd → Rd satis-
fying A(Zd ) = Zd . We call elements of Aff(Zd ) affine unimodular transformations.
For a family X of subsets of Rd , we consider the family of equivalence classes
X/ Aff(Zd ) := A(X ) : A ∈ Aff(Zd ) : X ∈ X
G. Averkov (B)
Brandenburgische Technische Universität Cottbus-Senftenberg, Fakultät 1, Platz der Deutschen
Einheit 1, 03046 Cottbus, Germany
e-mail: [email protected]
1 1
κ(a) := κ(a1 , . . . , ad ) = + ··· + .
a1 ad
1 Difference Between Families of Weakly and Strongly … 3
κ(x1 , . . . , xd ) = 1
It can happen that some facets of a maximal lattice-free polyhedron are more than
just blocked. We introduce a respective notion. Recall that the integer hull K I of a
compact convex set K in Rd is defined by
K I := conv(K ∩ Zd ).
T (a) := conv{o, a1 e1 , . . . , ad ed }.
1 1 1
= + , (1.1)
t t + 1 t (t + 1)
1 1 1 1
= + + , (1.2)
t t + 2 t (t + 2) t (t + 2)
1 2 1
= + . (1.3)
t 3t 3t
1 Difference Between Families of Weakly and Strongly … 5
The map φd replaces the component ad by two other components based on (1.1),
while ξd replaces ad based on (1.3). The map ψd acts by replacing the component ad
based on (1.1) and then replacing the component ad + 1 based on (1.2). Identities
(1.1)–(1.3) imply
d−1
1 3t − 1
1d+1 = ai ei + λ ed + ed+1 + μ 3ted+1
a
i=1 i
2
holds for λ = 2
3t−1
and μ = t−1
t (3t−1)
, where
1 Difference Between Families of Weakly and Strongly … 7
d−1
1
+ λ + μ = 1.
a
i=1 i
For d ≥ 4, Nill and Ziegler [7] construct one vector a ∈ Rd>0 with T (a) I ∈ Ld \Md .
We generalize this construction and provide many further vectors a with the above
properties. We will also need to verify that for different choices of a, we get essentially
different polytopes T (a) I .
Lemma 11 Let P and Q be d-dimensional strongly blocked lattice-free polytopes
such that for their integral hulls the equality Q I = A(PI ) holds for some A ∈
Aff(Zd ). Then Q = A(P).
Proof Since A is an affine transformation, we have
Lemma 12 Let a, b ∈ Rd>0 be such that the equality T (b) = A(T (a)) holds for
some A ∈ Aff(Zd ). Then a and b coincide up to permutation of components.
where
η(x) := ξd+4 (ψd+1 (φd (x)))
Since all known elements of Ld are of the form PI , for some strongly blocked
lattice-free polytope P, we ask the following:
Question 14 Do there exist polytopes L ∈ Ld which cannot be represented as L =
PI for any strongly blocked lattice-free polytope P?
If there is a gap between the families Ld and the family
PI : P ⊆ Rd strongly blocked lattice-free polytope ,
then it would be interesting to understand how irregular the polytopes from this gap
can be. For example, one can ask the following:
Acknowledgements I would like to thank Christian Wagner for valuable comments and Christian
Elsholtz for pointing to [5, 8, 10].
References
1. Averkov, G., Krümpelmann, J., Weltge, S.: Notions of maximality for integral lattice-free
polyhedra: the case of dimension three. Math. Oper. Res. 42(4), 1035–1062 (2017)
2. Averkov, G., Wagner, C., Weismantel, R.: Maximal lattice-free polyhedra: finiteness and an
explicit description in dimension three. Math. Oper. Res. 36(4), 721–742 (2011)
3. Blanco, M., Haase, C., Hofmann, J., Santos, F.: The finiteness threshold width of lattice poly-
topes. Trans. Am. Math. Soc. Ser. B 8, 399–419 (2021)
4. Del Pia, A., Weismantel, R.: Relaxations of mixed integer sets from lattice-free polyhedra.
Ann. Oper. Res. 240(1), 95–117 (2016)
5. Elsholtz, C.: Egyptian fractions with odd denominators. Q. J. Math. 67(3), 425–430 (2016)
6. Konyagin, S.V.: Double exponential lower bound for the number of representations of unity
by Egyptian fractions. Math. Notes 95(1–2), 277–281 (2014). Translation of Mat. Zametki 95
(2014), no. 2, 312–316
7. Lagarias, J.C., Ziegler, G.M.: Bounds for lattice polytopes containing a fixed number of interior
points in a sublattice. Canad. J. Math. 43(5), 1022–1035 (1991)
8. Landau, E.: Über die Klassenzahl der binären quadratischen Formen von negativer Discrimi-
nante. Math. Ann. 56(4), 671–676 (1903)
9. Lovász, L.: Geometry of numbers and integer programming. In: Mathematical Programming,
Tokyo (1988). Mathematics Applied (Japanese Ser.), vol. 6, pp. 177–201. SCIPRESS, Tokyo
(1989)
10. Newman, M.: A bound for the number of conjugacy classes in a group. J. London Math. Soc.
43, 108–110 (1968)
11. Nill, B., Ziegler, G.M.: Projecting lattice polytopes without interior lattice points. Math. Oper.
Res. 36(3), 462–467 (2011)
10 G. Averkov
n 1
12. Sándor, C.: On the number of solutions of the Diophantine equation i=1 xi = 1. Period.
Math. Hungar. 47(1–2), 215–219 (2003)
13. Treutlein, J.: 3-dimensional lattice polytopes without interior lattice points. Ph.D. thesis, Eber-
hart Karls Universität Tübingen (2010)
Chapter 2
On the Fine Interior of
Three-Dimensional Canonical Fano
Polytopes
2.1 Introduction
V. Batyrev
Mathematisches Institut, Universität Tübingen, Auf der Morgenstelle 10, 72076 Tübingen,
Germany
e-mail: [email protected]
A. Kasprzyk
School of Mathematical Sciences, University of Nottingham, Nottingham NG7 2RD, UK
e-mail: [email protected]
K. Schaller (B)
Mathematisches Institut, Freie Universität Berlin, Arnimallee 3, 14195 Berlin, Germany
e-mail: [email protected]
One can show that only finitely many linear inequalities x, n ≥ ord (n) + 1
are necessary to define FI . Therefore, FI is a convex hull of finitely many rational
points p ∈ MQ . Moreover, any lattice point p ∈ ◦ ∩ M in the usual interior of
is contained in FI . Therefore, FI contains the convex hull of ∩ M, i.e., we get
the inclusion conv(◦ ∩ M) ⊆ FI . In particular, FI is non-empty if ◦ ∩ M is
non-empty. Moreover, for any lattice polytope of dimension d ≤ 2 one has the
equality conv(◦ ∩ M) = FI [3]. The Fine interior FI of a lattice polytope of
dimension d ≥ 3 may happen to be strictly larger than the convex hull conv(◦ ∩ M).
The simplest famous example of such a situation is due to M. Reid. Other similar
examples based on hollow 3-topes can be found in Sect. 2.7:
Consider the M-lattice 3-tope ⊆ MQ defined as the convex hull of 4 lattice points
where am ∈ C are sufficiently general complex numbers. The importance of the Fine
interior is explained by the following theorem [3, 15, 20]:
2 On the Fine Interior of Three-Dimensional … 13
and the equation h d−1 (OV ) = |◦ ∩ M|. The numbers h i (OV ) are birational
invariants of Z ; they do not depend on a smooth projective compactification V
of Z . In particular, the number |◦ ∩ M| is the geometric genus pg of the affine
hypersurface Z ⊆ Td .
Smooth projective compactifications of non-degenerate hypersurfaces in Td can
be obtained using the theory of toric varieties [14].
Let ⊆ MQ be a lattice d-tope. We consider the normal fan of in the dual
space NQ , i.e., := {σ θ | θ }, where σ θ is the cone generated by all inward-
pointing facet normals of facets containing the face θ of . One has dim(σ θ ) +
dim(θ ) = d for any face θ . We denote by X the normal projective toric variety
constructed via the normal fan . In particular, the above function ord : NQ → Q
is a piecewise linear function with respect to this fan defining an ample Cartier divisor
on X . In particular, the cones σ θ ∈ are defined as
σ θ = y ∈ NQ ord (y) = x, y for all x ∈ θ .
Remark 5 Using the normal fan , one can compute the fundamental group
π1 (V ) of a smooth projective birational model V of a non-degenerate affine hyper-
surface Z (given as in Theorem 3). The fundamental group π1 (V ) does not depend
on the choice of the smooth birational model and it is isomorphic to the quotient of
the lattice N modulo the sublattice N generated by all lattice points in (d − 1)-
dimensional cones σ θ of the normal fan [4].
Example 6 The minimal model S of a non-degenerate affine surface Z defined
by a Laurent polynomial with the Newton polytope from Example 2 is a Godeaux
surface. It is a surface of general type with pg = q = 0, K 2 = 1, and π1 (S ) ∼
=
Z/5Z.
Definition 7 A lattice d-tope is called canonical Fano d-tope if |◦ ∩ M| = 1.
Up to a shift by a lattice vector, we will assume without loss of generality that 0 ∈ M
is the single lattice point in the interior ◦ of the canonical Fano d-tope , i.e.,
◦ ∩ M = {0}.
14 V. Batyrev et al.
All canonical Fano 3-topes have been classified [16]. There exists exactly 674,688
canonical Fano 3-topes . The aim of this paper is to present computational results
of their Fine interiors FI and some related combinatorial invariants. These data
are important for computing minimal smooth projective surfaces S with pg = 1
and q = 0 which are birational to affine non-degenerate hypersurfaces Z ⊆ T3 ∼ =
(C× )3 .
The simplest description of the minimal surface S has been obtained when is
a reflexive 3-tope [5].
is a lattice polytope.
There exist 4,319 reflexive 3-topes, classified by Kreuzer and Skarke [17]. They
form a small subset in the list of all 674,688 canonical Fano 3-topes [16]. Reflex-
ive 4-topes are also classified by Kreuzer and Skarke [18]. There exist 473,800,776
reflexive 4-topes, but the complete list of all canonical Fano 4-topes is unknown and
expected to be much bigger.
If is a reflexive d-tope, then X is a Gorenstein toric Fano d-fold and the Zariski
closure Z in X is a Gorenstein Calabi-Yau (d − 1)-fold. If d = 3, then Z is a
K 3-surface with at worst finitely many Du Val singularities of type Ak . The mini-
mal surface S is a smooth K 3-surface which is obtained as the minimal (crepant)
desingularization of Z [5].
One motivation for the present paper is due to Corti and Golyshev, who have
found 9 interesting examples of canonical Fano 3-simplices such that the affine
surfaces Z are birational to elliptic surfaces of Kodaira dimension κ = 1 [11].
The computation of the Fine interior FI for all canonical Fano 3-topes ⊆ MQ
has shown that the dimension of the Fine interior FI has only three values: 0, 1, and 3.
It is rather surprising that there are no canonical Fano 3-topes with dim(FI ) = 2.
The condition dim(FI ) = 0 holds if and only if FI equals the lattice point 0 ∈
M. There exist exactly 665,599 canonical Fano 3-topes with FI = {0}, where 0 ∈
M is the only interior lattice point of . These polytopes are characterized in [3,
Proposition 3.4] by the condition that 0 ∈ N is an interior lattice point of the lattice 3-
tope
[∗ ] := conv(∗ ∩ N ).
Remark 9 If is a canonical Fano 3-tope, then FI = {0} if and only if the non-
degenerate affine surface Z is birational to a K 3-surface [3, Theorem 2.26].
The case dim(FI ) = 1 splits in two subcases. There exists exactly 20 canonical
Fano 3-topes such that 0 ∈ M is the midpoint of the Fine interior FI . Therefore,
we call this Fine interior symmetric. Canonical Fano 3-topes with 1-dimensional sym-
metric Fine interior are characterized by the condition that [∗ ] is a reflexive 2-tope.
2 On the Fine Interior of Three-Dimensional … 15
The Fine interior of the remaining 9,020 canonical Fano 3-topes with dim(FI ) = 1
contains 0 ∈ M as a vertex. Therefore, we call this Fine interior asymmetric. Canon-
ical Fano 3-topes with 1-dimensional asymmetric Fine interior are combinatorially
characterized by the condition that 0 ∈ N is contained in the relative interior of a
facet [∗ ] of the lattice 3-tope [∗ ]. The minimal surfaces S corresponding
to canonical Fano 3-topes with 1-dimensional Fine interior (symmetric and asym-
metric) are elliptic surfaces of Kodaira dimension κ = 1.
There exist exactly 49 canonical Fano 3-topes with dim(FI ) = 3. These poly-
topes are characterized by the condition that 0 ∈ N is a vertex of the lattice 3-
tope [∗ ]. The surfaces S corresponding to canonical Fano 3-topes with 3-
dimensional Fine interior FI are of general type (i.e., S has maximal Kodaira
dimension κ = dim(S ) = 2).
Remark 10 The Fine interior computations were done using
FI = x ∈ MQ x, n ≥ ord (n) + 1 ,
θ n∈H(σ θ )
where H(σ θ ) denotes the set of all irreducible elements in the monoid σ θ ∩ N .
It is the minimal generating set of the monoid σ θ ∩ N and is called Hilbert basis
of σ θ ∩ N .
In the next sections we consider examples and discuss additional properties of
canonical Fano 3-topes in dependence of their Fine interiors FI . All computations
were done using the Graded Ring Database [8], including the data of all 674,688
canonical Fano 3-topes, and Magma [7]. Therefore, all statements have been checked
by computer calculations. The canonical Fano 3-topes used as examples in this paper
appear with an ID that is the example’s ID in the Graded Ring Database.1
F I = {0}.
Proof If [∗ ] is reflexive, then = (∗ )∗ is contained in the dual reflexive poly-
tope [∗ ]∗ . Therefore, the Fine interior of is contained in the Fine interior of the
reflexive polytope [∗ ]∗ and ([∗ ]∗ )FI = {0}. Thus, FI = {0}.
1 http://www.grdb.co.uk/forms/toricf3c.
16 V. Batyrev et al.
The converse statement is not true in general for d ≥ 5, but there exist many
equivalent characterizations of reflexive and almost reflexive d-topes among canon-
ical Fano d-topes if d = 3 or d = 4.
Let us recall some combinatorial invariants of arbitrary lattice d-topes.
where |k ∩ M| denotes the number of lattice points in the k-th dilate k of .
where ψi () are non-negative integers for all 0 ≤ i ≤ d [22] such that ψ0 () = 1
d
and ψ1 () = | ∩ M| − d − 1. Moreover, i=0 ψi () = v(), where v() :=
d! · vol() denotes the normalized volume of .
One has the following characterization of reflexive d-topes:
is a rational function
where ϕ0 () = 0 and ϕ1 () = |◦ ∩ M|. Using Serre duality, one obtains [12,
Sect. 4, 5.11]
ϕi () = ψd+1−i () (1 ≤ i ≤ d + 1),
i.e., in particular
ψd () = ϕ1 () = |◦ ∩ M|
and
ψd−1 () = ϕ2 () = |2◦ ∩ M| − (d + 1)|◦ ∩ M|.
2 On the Fine Interior of Three-Dimensional … 17
Therefore, the lattice d-tope is a canonical Fano d-tope if and only if ψd () = 1.
Moreover,
ψd−1 () = |(2)◦ ∩ M| − (d + 1)
| ∩ M| = |(2)◦ ∩ M| .
[2◦ ] ⊆ [∗ ]∗ .
Computations showed that among all 665,599 canonical Fano 3-topes with FI =
{0} there exist exactly 211,941 canonical Fano 3-tops such that [2◦ ] is reflexive.
For the remaining canonical Fano 3-topes the lattice 3-topes [2◦ ] are larger
than , but are not equal to the reflexive hull [∗ ]∗ .
We note that the set of all reflexive 3-topes forms a rather small part of the set of all
canonical Fano 3-topes. The majority of the canonical Fano 3-topes belong to the
subset of almost reflexive 3-topes. The proof of the following statement is based on
the result of Skarke [21] and the explanations in the previous section.
1. FI = {0};
2. 0 ∈ N is an interior lattice point of [∗ ];
3. is contained in some reflexive 3-tope;
4. τ k () is the reflexive 3-tope [∗ ]∗ for some sufficiently large k (1 ≤ k ≤ 5);
5. the lattice 3-tope [2◦ ] has exactly one interior lattice point;
6. the non-degenerate affine hypersurface Z defined by a Laurent polynomial with
Newton polytope is birational to a smooth K 3-surface.
Computations show that there exist exactly 665,599 almost reflexive canonical
Fano 3-topes. The set of almost reflexive 3-topes includes all 4,319 reflexive 3-topes.
We have shown that for any almost reflexive 3-tope , the reflexive polytope ref :=
[∗ ]∗ is the smallest reflexive 3-tope containing . We call ref the reflexive hull
of . Thus we obtain a natural surjective map → ref from the set of almost
reflexive 3-topes to the set of reflexive 3-topes, which is the identity on the set of
reflexive 3-topes. The minimal surface S is a K 3-surface if and only if is an
almost reflexive 3-tope. If is an almost reflexive 3-tope, but not reflexive, then the
minimal surface S is a crepant desingularization of the Zariski closure of Z in the
Gorenstein toric Fano threefold X ref defined by the reflexive hull of .
A generalization of the reflexive hull of almost reflexive 3-topes for arbitrary
lattice d-topes with non-empty Fine interior can be obtained using the notion of the
support of the Fine interior FI .
v3
v3
v1
v1 v2
v2
v5
v4
v4
(a) ID 547386. (b) ID 547385.
Fig. 2.1 Canonical Fano 3-topes with FI ={0}. Shaded faces are occluded and the Fine
interior {0} is shown in grey with a double border. The whole polytope is the canonical hull
can as well as the reflexive hull ref and the grey coloured polytope is . a Reflexive polytope
= conv{v1 , v2 , v3 , v4 } with v1 = (1, 0, 0), v2 = (0, 1, 0), v3 = (0, 0, 1), v4 = (−1, −1, −1), and
ref = can = . All facets of have lattice distance 1 to the origin. b Almost reflexive polytope
= conv{v1 , v2 , v3 , v4 } with v1 = (1, 0, 0), v2 = (0, 1, 0), v3 = (0, 0, 1), v4 = (−1, −1, −2), and
ref = can = conv{v1 , v2 , v3 , v4 , v5 } with v5 = (0, 0, −1) reflexive. The dark grey coloured facet
of has lattice distance 2 and all other facets have lattice distance 1 to the origin
hull can equals the reflexive hull ref of the polytope , i.e., can = ref = [∗ ]∗ .
In particular, in this case can is always a lattice 3-tope.
There exists a smooth projective toric variety X defined by a fan whose 1-
dimensional cones are generated by all lattice vectors from the finite set supp(FI ).
Then the minimal surface S is a K 3-surface which is the Zariski closure of Z
in X [3].
v1 := (1, 0, 0), v2 := (0, 1, 0), v3 := (0, 0, 1), and v4 := (−1, −1, −1)
and
can = [∗ ]∗ = (∗ )∗ =
v1 := (1, 0, 0), v2 := (0, 1, 0), v3 := (0, 0, 1), and v4 := (−1, −1, −2)
and
can = [∗ ]∗ = conv{v1 , v2 , v3 , v4 , v5 }
with v5 := (0, 0, −1) because is almost reflexive, i.e., ref = can = reflex-
ive (Fig. 2.1b).
There exist exactly 9,020 canonical Fano 3-topes with 1-dimensional Fine interior
such that 0 ∈ N belongs to a facet [∗ ] of the lattice 3-tope [∗ ]. This class of
canonical Fano 3-topes is characterized by the property that the lattice 3-tope [2◦ ]
has exactly 2 interior lattice points.
The corresponding minimal surfaces S are simply connected (i.e., have triv-
ial fundamental group π1 (S )) elliptic surfaces of Kodaira dimension κ = 1. We
observed that the facet [∗ ] is a reflexive 2-tope corresponding to one of the
three types pictured in Fig. 2.2. All N -lattice points on the boundary of belong
to supp(FI ). It was checked that for all these 3-topes the canonical hull can is
again a lattice 3-tope. Moreover, the Fine interior FI is contained in the ray gener-
ated by the primitive lattice vector v ∈ M which is the primitive inward-pointing
facet normal of , i.e., x, y = 0 for all x ∈ FI , y ∈ . The lattice point 0 ∈ M is
a vertex of FI . More precisely, one has
where λ ∈ {1/2, 2/3}. The primitive lattice vector v is the unique interior lattice
point on a reflexive facet θ+ of of one of the three possible types pictured in
Fig. 2.2. These three reflexive polygons θ+ are characterized by the condition that the
dual reflexive polygons θ+∗ are obtained from θ+ (Fig. 2.3) by enlarging the lattice Z2
in the following ways: Z2 + Z(1/3, 2/3) (Fig. 2.3a), Z2 + Z(1/2, 0) (Fig. 2.3b), and
Z2 + Z(1/2, 1/2) (Fig. 2.3c). Moreover, the reflexive facet θ+ of is isomorphic to
the facet of [∗ ]. The projection M → M/Zv of or of θ+ along v is a reflexive
polygon of one of the three types pictured in Fig. 2.3, which is dual to θ+ and . The
lattice vector v defines a character of the 3-dimensional torus χ : T3 → C× . For
almost all α ∈ C× , the fiber χ −1 (α) is an affine elliptic curve defined by a Laurent
22 V. Batyrev et al.
Fig. 2.2 Reflexive Facets of Containing ±v . Three types of reflexive facets θ±
of containing ±v for all 9,020 + 20 canonical Fano 3-topes with dim(FI ) = 1. Ver-
tices are coloured black, boundary points that are not vertices grey, and the origin light grey.
a conv{(1, 0), (0, 1), (−1, −1)}. b conv{(1, 0), (−1, 1), (−1, −1)}. c conv{(±1, 0), (0, ±1)}
Fig. 2.3 Reflexive Projection Polytopes. Three types of reflexive polytopes obtained via a pro-
jection of along ±v for all 9,020 + 20 canonical Fano 3-topes with dim(FI ) = 1. Ver-
tices are coloured black, boundary points that are not vertices grey, and the origin light grey.
a conv{(−1, 2), (−1, −1), (2, −1)}. b conv{(−2, −1), (0, 1), (2, −1)}. c conv{(±1, ±1)}
polynomial with the reflexive Newton polytope ∗ ∼ = θ+∗ of one of the three types
pictured in Fig. 2.3 with the distribution shown in Table 2.1. So χ defines birationally
an elliptic fibration.
Table 2.1 Distribution of the Reflexive Facets of Containing ±v . Table contains: Type of
the reflexive facet θ± containing ±v , type of the dual reflexive facet θ±∗ , the enlarged lattice used
to obtain θ±∗ from θ± , the number of canonical Fano 3-topes asym := { | 1-dim. asym. FI }, and
the number of canonical Fano 3-topes sym := { | 1-dim. sym. FI } with respect to the facet type
of θ± pictured in Fig. 2.2
θ± θ±∗ Enlarged lattice #asym #sym
Figure 2.2a Figure 2.3a Z2 + 3,038 7
Z(1/3, 2/3)
Figure 2.2b Figure 2.3b Z2 + Z(1/2, 0) 4,663 9
Figure 2.2c Figure 2.3c Z2 + 1,319 4
Z(1/2, 1/2)
2 On the Fine Interior of Three-Dimensional … 23
v1 v2
v3
v5
v1
v2
v5
v3 v4 v4
Fig. 2.4 Canonical Fano 3-topes with Asymmetric Fine Interior of Dimension 1 Shaded
faces are occluded. The Fine interior and the origin are shown in grey, with a double border
around the origin. The facet θ+ is grey dotted. a The whole polytope is = conv{v1 , v2 , v3 , v4 }
with v1 = (2, 3, 8), v2 = (1, 0, 0), v3 = (0, 1, 0), and v4 = (−1, −1, −1). Moreover, FI =
conv{(0, 0, 0), (1/2, 1/2, 1)}, θ+ = conv{v1 , v2 , v3 }, and can = conv{v1 , v2 , v3 , v4 , v5 } with v5 =
(0, 1, 4). b The whole polytope is = conv{v1 , v2 , v3 , v4 } with v1 = (−1, 1, −2), v2 =
(1, −2, 3), v3 = (1, 0, 0), and v4 = (−2, 5, −3). Moreover, FI = conv{(0, 0, 0), (0, 2/3, 0)}
and θ+ = conv{v2 , v3 , v4 }, and can = conv{v1 , v2 , v3 , v4 , v5 } with v5 = (−2, 4, −3)
v1 := (2, 3, 8), v2 := (1, 0, 0), v3 := (0, 1, 0), and v4 := (−1, −1, −1)
where v = (1, 1, 2). One has v1 + 2v2 + v3 = 4v . Therefore, v is the interior lat-
tice point of the reflexive facet θ+ of with vertices v1 , v2 , v3 and the images v 1 , v 2 , v 3
of v1 , v2 , v3 in M/Zv are vertices of the dual reflexive triangle θ+∗ (Fig. 2.3b) satis-
fying the relation
v 1 + 2v 2 + v 3 = 0.
can = conv{v1 , v2 , v3 , v4 , v5 }
Table 2.2 9 Canonical Fano 3-topes with Asymmetric Fine Interior of Dimension 1. Table
contains: vertices vert() of , vertices vert(FI ) of the Fine interior FI , unique primitive lattice
point v ∈ θ+ in the reflexive facet θ+ , and weights (wi )0≤i≤3 of the weighted projective 3-
space P(w0 , . . . , w3 ) appearing in [11]
ID vert() vert(FI ) v (wi )0≤i≤3
547324 (2, 3, 8), (1, 0, 0), (0, 1, 0), (−1, −1, −1) 0, 1/2 · v (1, 1, 2) (1, 5, 6, 8)
547323 (−1, 1, −2), (1, −2, 3), (1, 0, 0), (−2, 5, −3) 0, 2/3 · v (0, 1, 0) (1, 4, 7, 9)
547311 (−1, 4, 2), (−1, −1, 0), (0, 0, −1), (2, 0, 1) 0, 2/3 · v (0, 1, 1) (2, 5, 8, 9)
547490 (1, 2, 4), (1, 0, 0), (1, −2, 3), (−1, 1, −2) 0, 1/2 · v (0, 1, 0) (1, 5, 8, 14)
547321 (1, −2, 3), (0, 1, 0), (1, 0, 0), (−6, 3, −8) 0, 1/2 · v (−1, 1, −2) (3, 7, 8, 10)
547305 (0, 1, 0), (1, 0, 0), (1, 2, 4), (−4, −6, −7) 0, 2/3 · v (−1, −1, −1) (4, 7, 9, 10)
547526 (1, 0, 0), (0, 1, 0), (−2, 1, 5), (2, −4, −9) 0, 2/3 · v (1, −1, −3) (5, 9, 8, 11)
547454 (2, 1, 7), (1, 0, 0), (0, 1, 0), (−2, −3, −3) 0, 1/2 · v (0, 0, 1) (3, 7, 8, 18)
547446 (0, 1, 1), (−6, 7, −15), (1, −2, 3), (1, 0, 0) 0, 1/2 · v (−1, 1, −2) (5, 8, 9, 22)
Table 2.3 9 Canonical Fano 3-topes with Asymmetric Fine Interior of Dimension 1. Table
contains: primitive inward-pointing facet normals (n i )1≤i≤4 of , vertices vert(θ+ ) of the reflexive
facet θ+ , and primitive inward-pointing facet normal n θ+ of the reflexive facet θ+
ID (n i )1≤i≤4 vert(θ+ ) n θ+
547324 (−2, −2, 1), (−1, −1, 3), (−1, 3, −1), (7, −3, −1) (2, 3, 8), (1, 0, 0), (0, 1, 0) (−2, −2, 1)
547323 (−3, −3, −2), (−1, 0, 1), (−1, 6, 4), (17, 3, −5) (1, −2, 3), (1, 0, 0), (−2, 5, −3) (−3, −3, −2)
547311 (−1, −1, 1), (−1, 2, 1), (1, 2, −5), (7, −2, 5) (−1, 4, 2), (−1, −1, 0), (2, 0, 1) (1, 2, −5)
547490 (−2, −2, 1), (−1, 0, 0), (−1, 6, 4), (23, 2, −8) (1, 2, 4), (1, 0, 0), (−1, 1, −2) (−2, −2, 1)
547321 (−3, −3, −2), (−2, −2, 1), (−1, 3, 2), (9, −5, −8) (0, 1, 0), (1, 0, 0), (−6, 3, −8) (−2, −2, 1)
547305 (−7, −7, 11), (−2, −2, 1), (−1, 2, −1), (7, −3, −1) (0, 1, 0), (1, 2, 4), (−4, −6, −7) (7, −3, −1)
547526 (−5, −5, −2), (−3, −3, 1), (−1, 2, −1), (25, −8, 10) (1, 0, 0), (0, 1, 0), (2, −4, −9) (−3, −3, 1)
547454 (−7, −7, 2), (−1, −1, 2), (−1, 1, 0), (7, −2, −2) (2, 1, 7), (0, 1, 0), (−2, −3, −3) (7, −2, −2)
547446 (−9, 21, 14), (−5, −3, −2), (−1, −1, 0), (9, 1, −3) (0, 1, 1), (−6, 7, −15), (1, −2, 3) (9, 1, −3)
v1 := (−1, 1, −2), v2 := (1, −2, 3), v3 := (1, 0, 0), and v4 := (−2, 5, −3)
(ID 547323, Fig. 2.4b, Tables 2.2 and 2.4). Then (Table 2.3)
where v = (0, 1, 0). One has v2 + v3 + v4 = 3v . Therefore, v is the interior lat-
tice point of the reflexive facet θ+ of with vertices v2 , v3 , v4 and the images v 2 , v 3 , v 4
of v2 , v3 , v4 in M/Zv are vertices of the dual reflexive triangle θ+∗ (Fig. 2.3a) satis-
fying the relation
v 2 + v 3 + v4 = 0.
Table 2.4 9 Canonical Fano 3-topes with Asymmetric Fine Interior of Dimension 1. Table contains: vertices vert( ) of the reflexive facet [∗ ],
support supp(FI ) of the Fine interior FI , and vertices vert(can ) of the canonical hull can
ID vert( ) supp(FI ) vert(can )
547324 (−1, 3, −1), (−1, −1, 1), (1, −1, 0) (−2, −2, 1), (−1, −1, 1), S1 vert(), (0, 1, 4)
547323 (−1, 0, 1), (−1, 0, 0), (2, 0, −1) (−3, −3, −2), (−1, 0, 0), S2 vert(), (−2, 4, −3)
547311 (−1, −1, 1), (0, 1, −1), (1, 0, 0) (−1, −1, 1), (−1, 0, 1), S3 vert(), (−1, 2, 0)
547490 (−1, 0, 0), (−1, 0, 1), (3, 0, −1) (−2, −2, 1), (−1, 0, 0), S4 vert(), (1, −1, 4)
547321 (−1, −1, 0), (−1, 3, 2), (1, −1, −1) (−2, −2, 1), (−1, −1, 0), (−1, 0, 0), (−1, 1, 1), (−1, 3, 2), vert(), (1, 0, 1), (0, −3, 4)
(0, −1, −1), (1, −1, −1)
547305 (−1, 2, −1), (1, −1, 0), (0, −1, 1) (−1, −1, 1), (−1, 0, 0), (−1, 2, −1), (0, −1, 1), (1, −1, 0), vert(), (0, −2, −3), (1, 2, 2)
2 On the Fine Interior of Three-Dimensional …
can = conv{v1 , v2 , v3 , v4 , v5 }
There exist exactly 20 canonical Fano 3-topes such that 0 is the center of 1-
dimensional Fine interior FI . In this case, S is an elliptic surface of Kodaira
dimension κ = 1 with non-trivial fundamental group π1 (S ) of order 2 or 3. Com-
putations show that one always has = can and
with λ = 1
2
if and only if |π1 (S )| = 2 and
with μ = 23 if and only if |π1 (S )| = 3. The primitive lattice vectors ±v are the
two unique interior lattice points in two reflexive facets θ± of one of the three
possible types pictured in Fig. 2.2. The reflexive facets θ± of are isomorphic to
the facet of [∗ ]. The projections M → M/Z(±v ) of or of θ± along ±v
reveal a reflexive polygon of one of the three types pictured in Fig. 2.3, which is
dual to θ± and . The lattice vector v defines a character of the 3-dimensional
torus χ : T3 → C× . For almost all α ∈ C× , the fiber χ −1 (α) is an affine elliptic
curve defined by a Laurent polynomial with the reflexive Newton polytope ∗ ∼ = θ±∗
of one of the three types pictured in Fig. 2.3 with the distribution shown in Table 2.1.
So χ defines birationally an elliptic fibration. The vertex sets of and these reflexive
facets are related via vert() = vert(θ+ ) ∪ vert(θ− ). Moreover, every edge of is
either an edge of θ+ or θ− of these two facets or it is parallel to v .
Example 32 Let ⊆ MQ be a canonical Fano 3-tope given as the convex hull
v4 v1
v2
v3
v2
v1
v3 v4
with λ = 21 , where v = (0, 0, 1). One has 2v1 + v3 + v4 = 4v and 2v1 + v2 +
v3 = 4(−v ). Therefore, v is the interior lattice point of the reflexive facet θ+ = θ134
of and −v is the interior lattice point of the reflexive facet θ− = θ123 of
(Fig. 2.2b). The images v1 , v 3 , v 4 of v1 , v3 , v4 in M/Zv and the images v 1 , v 2 , v 3
of v1 , v2 , v3 in M/Z(−v ) are vertices of the dual reflexive triangle θ±∗ (Fig. 2.3b)
satisfying the relation
2v 1 + v3 + v4 = 0
and
2v 1 + v 2 + v 3 = 0,
respectively.
To compute the canonical hull can of , we obtain supp(FI ) = {si | 1 ≤ i ≤ 6}
with s1 := (−1, −2, 2), s2 := (−1, 1, 0), s3 := (0, −1, 0), s4 := (1, −1, 0), s5 :=
(2, −1, 0), and s6 := (9, −2, −2), which leads to can = .
28 V. Batyrev et al.
Table 2.5 20 Canonical Fano 3-topes with Symmetric Fine Interior of Dimension 1. Table
contains: vertices vert() of
ID vert()
547393 (0, 1, 0), (2, 1, 1), (−2, −3, −5), (2, 1, 9)
547409 (−4, 2, 9), (1, 0, 0), (0, 1, 0), (7, −6, −18)
547461 (0, 1, 0), (2, 1, 1), (−2, −3, −5), (0, 1, 4)
544442 (1, 0, 0), (0, 1, 0), (3, −6, 8), (1, −4, 4), (−5, 6, −12)
544443 (−1, −2, 0), (3, −6, 8), (0, 1, 0), (1, 0, 0), (−3, 4, −8)
544651 (−4, 1, −3), (4, −2, 3), (0, 1, 0), (1, −2, 3), (−1, 1, −3)
544696 (5, −4, −15), (1, 0, 0), (0, 1, 0), (−4, 2, 9), (−3, 1, 6)
544700 (−2, −3, −3), (0, 1, 0), (1, 0, 0), (−1, −4, −6), (2, 5, 9)
544749 (−6, −5, −8), (0, 1, 0), (1, 0, 0), (−2, −1, 0), (3, 2, 4)
520925 (0, 1, 0), (0, 0, 1), (−2, −1, 0), (−2, 0, −1), (8, 2, 3), (−2, −3, −2)
520935 (3, 4, 6), (2, 1, 2), (−3, −2, −2), (1, 0, 0), (0, 1, 0), (−6, −5, −8)
522056 (−1, −1, 0), (0, 1, 0), (1, 0, 0), (−1, −1, −3), (−5, −3, −6), (6, 4, 9)
522059 (2, 5, 6), (−2, −3, −3), (0, 1, 0), (1, 0, 0), (−1, −4, −6), (0, 1, 3)
522087 (1, 0, −3), (1, 0, 0), (0, 1, 0), (−4, 2, 9), (−3, 1, 6), (5, −4, −12)
522682 (2, 1, 4), (−3, −2, −4), (−2, −3, −4), (1, 2, 4), (1, 0, 0), (0, 1, 0)
522684 (−2, −1, −4), (3, 2, 4), (−2, −1, 0), (1, 0, 0), (0, 1, 0), (−4, −3, −4)
526886 (−3, 4, −6), (1, 0, 0), (0, 1, 0), (3, −6, 8), (0, 1, −2), (2, −5, 6)
439403 (1, 2, 2), (−1, 0, 0), (−1, 1, −1), (1, 0, 0), (−1, −2, −2), (1, 1, 3), (1, −3, −1)
275525 (4, 1, 2), (0, 1, 0), (−2, −1, 0), (1, 1, 2), (−3, −1, −2), (−2, −1, −2), (1, 1, 0), (1, −1, 0)
275528 (−1, 0, −1), (−3, −2, 1), (−2, −1, 2), (0, −1, 0), (0, 1, 0), (1, 0, 1), (2, 1, −2), (3, 2, −1)
v1 := (−4, 2, 9), v2 := (1, 0, 0), v3 := (0, 1, 0), and v4 := (7, −6, −18)
and
v 1 + v 3 + v 4 = 0,
respectively.
Table 2.6 20 Canonical Fano 3-topes with Symmetric Fine Interior of Dimension 1. Table contains: vertices vert(FI ) of the Fine interior FI , unique
primitive lattice points ±v ∈ θ± in the reflexive facets θ± , vertices vert(θ± ) of the reflexive facets θ± , and support supp(FI ) of the Fine interior FI
(here: can = )
ID vert(FI ) ±v vert(θ± ) supp(FI )
547393 ±1/2 · v ±(0, 0, 1) (0, 1, 0), (2, 1, 1), (−2, −3, −5) (−1, −2, 2), (−1, 1, 0), (0, −1, 0), (1, −1, 0), (2, −1, 0), (9, −2, −2)
(0, 1, 0), (−2, −3, −5), (2, 1, 9)
547409 ±2/3 · v ±(1, −1, −3) (−4, 2, 9), (1, 0, 0), (0, 1, 0) (−3, −3, −1), (−1, −1, 0), (−1, 2, −1), (2, −1, 1), (15, −3, 7)
(−4, 2, 9), (0, 1, 0), (7, −6, −18)
547461 ±1/2 · v ±(0, 0, 1) (0, 1, 0), (2, 1, 1), (−2, −3, −5) (−3, 6, −2), (−1, −2, 2), (−1, 1, 0), (0, −1, 0), (1, −1, 0), (2, −1, 0)
(2, 1, 1), (−2, −3, −5), (0, 1, 4)
544442 ±1/2 · v ±(1, −1, 2) (0, 1, 0), (1, −4, 4), (−5, 6, −12) (−2, −2, −1), (−1, −1, 0), (−1, 1, 1), (1, −1, −1), (3, −1, −2), (10, −2, −5)
2 On the Fine Interior of Three-Dimensional …
v3
v4
v4
v1
v3 v2
v1
v2
Fig. 2.6 Canonical Fano 3-topes with Fine Interior of Dimension 3. Shaded
faces are occluded. The Fine interior and the origin are shown in grey with a
double border around the origin. a The whole polytope is = conv{v1 , v2 , v3 , v4 }
with v1 = (1, 0, 0), v2 = (−2, −4, −5), v3 = (1, 2, 4), and v4 = (1, 4, 2). More-
over, FI = conv{(0, 0, 0), (−1/2, −1, −3/2), (0, −1/3, −2/3), (0, 1/3, −1/3)}
and can = . b The whole polytope is = conv{v1 , v2 , v3 , v4 } with v1 =
(−3, −2, −2), v2 = (1, 0, 0), v3 = (1, 3, 1), and v4 = (1, 1, 3). Moreover, FI =
conv{(0, 0, 0), (−1, −1/2, −1/2), (0, 3/4, 1/4), (0, 1/4, 3/4)} and can =
(ID 547444, Fig. 2.6a, Tables 2.7, 2.8, and 2.9). The primitive inward-pointing facet
normals of the facets θ124 , θ234 , θ123 , and θ134 of this simplex are
n 1 := (−2, −1, 2), n 2 := (5, −1, −1), n 3 := (−1, 2, −1), and n 4 := (−1, 0, 0),
n 1 + n 2 + n 3 + 2n 4 = 0.
M := {(m 1 , m 2 ,m 3 , m 4 ) ∈ Z4 |
m 1 + m 2 + 2m 3 + 2m 4 = 8, 3m 2 + m 3 + 3m 4 ≡ 0 (mod 4)}
Table 2.7 49 Canonical Fano 3-topes with Fine Interior of Dimension 3 Table contains: ver-
tices vert() of
ID vert()
547444 (1, 0, 0), (−2, −4, −5), (1, 2, 4), (1, 4, 2)
547465 (−3, −2, −2), (1, 0, 0), (1, 3, 1), (1, 1, 3)
547524 (0, 2, 1), (−2, −3, −5), (2, 1, 1), (0, 0, 1)
547525 (0, 0, 1), (0, 1, 0), (2, 1, 1), (−2, −5, −7)
545317 (−3, 4, −6), (0, 1, 0), (1, 0, 0), (1, −2, 4), (3, −5, 6)
545932 (0, −1, −1), (1, −1, −3), (−2, 1, 5), (1, 0, 0), (1, 2, −2)
546013 (3, −5, 6), (1, −2, 4), (1, 0, 0), (−1, 1, −2), (−1, 3, −2)
546062 (0, 1, 3), (−2, 1, −1), (0, 1, 0), (1, 0, 0), (−1, −2, −2)
546070 (0, −2, −3), (0, 2, 1), (−2, −3, −5), (2, 1, 1), (0, 0, 1)
546205 (1, 2, −2), (−1, 0, 2), (1, 0, 0), (−2, 1, 5), (1, −1, −3)
546219 (1, 1, 1), (−3, −2, −2), (1, 0, 0), (1, 3, 1), (−1, −1, 1)
546663 (2, −3, −1), (1, 0, 0), (0, 1, 0), (0, 0, 1), (−2, −3, −3)
546862 (1, 0, 0), (0, 1, 0), (−2, 1, 5), (1, −1, −3), (1, 2, −2)
546863 (−1, −1, 1), (1, 3, 1), (0, 0, 1), (1, 0, 0), (−3, −2, −2)
547240 (−1, 1, −2), (0, 1, 0), (1, 0, 0), (1, −2, 4), (3, −5, 6)
547246 (0, −2, −3), (−2, −3, −5), (2, 1, 1), (0, 1, 0), (0, 0, 1)
532384 (1, −1, −3), (−2, 1, 5), (1, 0, 0), (1, −1, −2), (0, −1, −1), (1, 2, −2)
532606 (0, −1, 2), (−1, −1, 0), (0, 1, 0), (1, 0, 0), (2, 2, −3), (−2, 0, −3)
533513 (−1, 1, 2), (1, 0, 0), (0, 1, 0), (1, 1, 2), (−1, −2, −4), (−2, −3, −4)
534667 (1, 0, 3), (−1, −1, −1), (0, 1, 0), (1, 0, 0), (−1, −1, 0), (5, 2, 3)
534669 (1, 3, 0), (5, 3, 2), (−1, −1, −1), (0, 0, 1), (1, 0, 0), (−1, −1, 0)
534866 (−1, −1, −3), (1, 0, 0), (0, 1, 0), (1, 1, 1), (−1, −1, 0), (−3, −5, −3)
535952 (3, −5, 6), (1, −2, 4), (1, 0, 0), (0, 1, 0), (−1, 1, −2), (−1, 2, −2)
536013 (0, 1, 1), (0, 0, 1), (0, 1, 0), (2, 1, 1), (−2, −3, −5), (0, −2, −3)
536498 (1, 2, −2), (1, −1, −2), (1, 0, 0), (0, 1, 0), (−2, 1, 5), (1, −1, −3)
537834 (0, 0, 1), (1, 0, 0), (0, 1, 0), (−2, 1, 5), (1, −1, −3), (1, 2, −2)
538356 (−2, −3, −3), (−1, −3, −1), (0, 0, 1), (0, 1, 0), (1, 0, 0), (−1, −1, −3)
539063 (−1, 1, −1), (1, 1, 3), (−3, −2, −2), (1, 0, 0), (0, 1, 0), (1, 1, 2)
539304 (1, 0, 1), (−3, −1, −2), (1, 1, 2), (−2, −1, 0), (1, 0, 0), (1, 2, 0)
539313 (1, −1, −2), (1, 1, −1), (−1, 2, 2), (1, −1, −3), (−2, 1, 5), (1, 0, 0)
540602 (0, 0, 1), (1, 0, 0), (−2, 1, 5), (1, −1, −3), (−1, 2, 2), (1, 1, −1)
540663 (1, 0, 0), (0, 1, 0), (1, 1, 2), (−3, −1, −2), (1, 1, 1), (−3, −2, 0)
474457 (−1, 2, −3), (1, 0, 2), (0, 0, 1), (0, 1, 0), (1, 0, 0), (−1, −1, 0), (−3, −2, −3)
481575 (3, 2, 4), (−1, −1, −2), (−3, −1, −2), (−2, −1, 0), (0, 1, 0), (1, 0, 0), (0, 0, −1)
483109 (3, 0, 2), (1, −2, −2), (0, 0, −1), (−1, −1, 0), (1, 1, 1), (0, 1, 0), (−1, 0, 0)
490478 (1, −1, −2), (1, 1, −1), (−1, 2, 2), (1, −1, −3), (−2, 1, 5), (1, 0, 0), (−1, 0, 2)
490481 (−3, −2, 0), (−5, −3, −2), (1, 0, 0), (0, 1, 0), (1, 1, 2), (−1, −1, −1), (2, 1, 1)
490485 (−1, −1, 0), (1, 2, 0), (1, 0, 0), (−2, −1, 0), (1, 1, 2), (−3, −1, −2), (1, 0, 1)
490511 (1, 0, 0), (0, 1, 0), (−2, −1, 0), (1, 1, 2), (2, 1, 1), (1, 0, 1), (−5, −2, −4)
495687 (0, 0, −1), (1, 1, −1), (−1, 2, 2), (1, −1, −3), (−2, 1, 5), (1, 0, 0), (0, 0, 1)
499287 (1, 1, 1), (−1, −1, −3), (1, 0, 0), (0, 1, 0), (0, 0, 1), (−1, −3, −1), (−2, −3, −3)
499291 (−1, −1, −1), (−1, −1, −3), (1, 0, 0), (0, 1, 0), (0, 0, 1), (−1, −3, −1), (−2, −3, −3)
499470 (1, 0, 0), (0, 1, 0), (−2, −1, 0), (1, 1, 2), (0, 0, 1), (−5, −2, −4), (2, 1, 1)
501298 (3, −6, 8), (−1, 1, −2), (1, −2, 3), (0, 1, 0), (1, 0, 0), (0, 1, −1), (3, −5, 6)
501330 (1, 0, 0), (0, 1, 0), (−2, −1, 0), (1, 1, 2), (1, 1, 1), (0, 0, 1), (−5, −2, −4)
354912 (3, 1, 2), (1, 0, 0), (0, 1, 0), (−2, −1, 0), (1, 1, 2), (2, 1, 1), (1, 0, 1), (−5, −2, −4)
372528 (2, 1, 1), (−1, −1, −1), (1, 1, 2), (0, 1, 0), (1, 0, 0), (−5, −3, −2), (−3, −2, 0), (1, 1, 0)
372973 (−5, −2, −4), (1, 0, 1), (2, 1, 1), (1, 1, 2), (−2, −1, 0), (0, 1, 0), (1, 0, 0), (2, 1, 2)
388701 (1, 1, 1), (−2, −3, −3), (−1, −3, −1), (0, 0, 1), (0, 1, 0), (1, 0, 0), (−1, −1, −3), (−1, −1, −1)
34 V. Batyrev et al.
Table 2.8 49 Canonical Fano 3-topes with Fine Interior of Dimension 3. Table contains: ver-
tices vert(FI ) of the Fine interior FI
ID vert(FI )
547444 0, (−1/2, −1, −3/2), (0, −1/3, −2/3), (0, 1/3, −1/3)
547465 0, (−1, −1/2, −1/2), (0, 3/4, 1/4), (0, 1/4, 3/4)
547524 0, (0, 1/2, 0), (1/3, 1/3, 0), (−1/3, −1/3, −1)
547525 0, (0, 0, −1/2), (1/3, 0, −1/3), (−1/3, −1, −5/3)
545317 0, (1, −3/2, 2), (2/3, −2/3, 1), (1/2, −1/2, 1), (2/3, −1, 5/3)
545932 0, (−1/2, 1/2, 3/2), (0, 1/3, 2/3), (0, 2/3, 1/3)
546013 0, (1, −3/2, 2), (0, 1/2, 0), (1/2, −1/4, 1/2), (1/2, −3/4, 3/2)
546062 0, (−1/2, −1/2, −1/2), (−2/3, 0, −1/3), (−1/3, 0, 1/3)
546070 0, (0, 1/2, 0), (1/2, 1/4, 0), (0, −1/2, −1), (−1/2, −3/4, −3/2)
546205 0, (−1/2, 1/2, 3/2), (0, 1/3, 2/3), (0, 2/3, 1/3)
546219 0, (−1, −1/2, −1/2), (−1/3, 1/3, 0), (−2/3, −1/3, 0)
546663 0, (0, −1/2, 0), (1/3, −1, −1/3), (−1/3, −1, −2/3)
546862 0, (−1/2, 1/2, 3/2), (0, 1/3, 2/3), (0, 2/3, 1/3)
546863 0, (−1, −1/2, −1/2), (−1/3, 1/3, 0), (−2/3, −1/3, 0)
547240 0, (1, −3/2, 2), (2/3, −2/3, 1), (1/2, −1/2, 1), (2/3, −1, 5/3)
547246 0, (0, 0, −1/2), (1/3, 0, −1/3), (0, −1/2, −1), (−1/3, −2/3, −4/3)
532384 0, (−1/2, 1/2, 3/2), (0, 1/3, 2/3), (0, 2/3, 1/3)
532606 0, (0, 1/2, −1/2), (1/3, 2/3, −1), (−1/3, 1/3, −1)
533513 0, (−1/2, −1/2, −1), (−1/2, 0, 0), (−1/3, 0, −1/3), (−2/3, −2/3, −1)
534667 0, (1/2, 1/2, 1/2), (4/3, 2/3, 1), (2/3, 1/3, 1)
534669 0, (1/2, 1/2, 1/2), (4/3, 1, 2/3), (2/3, 1, 1/3)
534866 0, (0, −1/2, −1/2), (−1/3, −2/3, −1), (−2/3, −4/3, −1)
535952 0, (1, −3/2, 2), (2/3, −2/3, 1), (1/2, −1/2, 1), (2/3, −1, 5/3)
536013 0, (0, 0, −1/2), (1/3, 0, −1/3), (0, −1/2, −1), (−1/3, −2/3, −4/3)
536498 0, (−1/2, 1/2, 3/2), (0, 1/3, 2/3), (0, 2/3, 1/3)
537834 0, (−1/2, 1/2, 3/2), (0, 1/3, 2/3), (0, 2/3, 1/3)
538356 0, (0, −1/2, −1/2), (−1/3, −2/3, −1), (−1/3, −1, −2/3), (−1/2, −1, −1)
539063 0, (−1, −1/2, −1/2), (−2/3, 0, −1/3), (−1/3, 0, 1/3)
539304 0, (0, 1/2, 0), (−1/2, 0, 0), (0, 1/3, 1/3), (−2/3, 0, −1/3)
539313 0, (−1/2, 1/2, 3/2), (0, 1/3, 2/3), (0, 1/2, 1/2), (−1/3, 2/3, 1)
540602 0, (−1/2, 1/2, 3/2), (0, 1/3, 2/3), (0, 1/2, 1/2), (−1/3, 2/3, 1)
540663 0, (−1/2, 0, 0), (−1, −1/2, 0), (−1/3, 0, 1/3), (−1, −1/3, −1/3)
474457 0, (0, 0, −1/2), (−1/3, 1/3, −1), (−2/3, −1/3, −1)
481575 0, (−1/2, 0, 0), (1/2, 1/2, 1), (0, 1/3, 1/3), (−1/3, 0, 1/3)
483109 0, (0, −1/2, 0), (2/3, −1/3, 1/3), (1/3, −2/3, −1/3)
490478 0, (−1/2, 1/2, 3/2), (0, 1/3, 2/3), (0, 1/2, 1/2), (−1/3, 2/3, 1)
490481 0, (−1/2, 0, 0), (−1, −1/2, 0), (−1/3, 0, 1/3), (−4/3, −2/3, −1/3)
490485 0, (0, 1/2, 0), (−1/2, 0, 0), (0, 1/3, 1/3), (−2/3, 0, −1/3)
490511 0, (−3/2, −1/2, −1), (−1/2, 0, 0), (−2/3, 0, −1/3), (−1, −1/3, −1/3)
495687 0, (−1/2, 1/2, 3/2), (0, 1/3, 2/3), (0, 1/2, 1/2), (−1/3, 2/3, 1)
499287 0, (0, −1/2, −1/2), (−1/3, −2/3, −1), (−1/3, −1, −2/3), (−1/2, −1, −1)
499291 0, (0, −1/2, −1/2), (−1/3, −2/3, −1), (−1/3, −1, −2/3), (−1/2, −1, −1)
499470 0, (−3/2, −1/2, −1), (−1/2, 0, 0), (−2/3, 0, −1/3), (−1, −1/3, −1/3)
501298 0, (1/2, −1/2, 1), (2/3, −2/3, 1), (1, −3/2, 2), (1, −5/3, 7/3)
501330 0, (−3/2, −1/2, −1), (−1/2, 0, 0), (−2/3, 0, −1/3), (−1, −1/3, −1/3)
354912 0, (−3/2, −1/2, −1), (−1/2, 0, 0), (−2/3, 0, −1/3), (−1, −1/3, −1/3)
372528 0, (−1/2, 0, 0), (−1, −1/2, 0), (−1/3, 0, 1/3), (−4/3, −2/3, −1/3)
372973 0, (−3/2, −1/2, −1), (−1/2, 0, 0), (−2/3, 0, −1/3), (−1, −1/3, −1/3)
388701 0, (0, −1/2, −1/2), (−1/3, −2/3, −1), (−1/3, −1, −2/3), (−1/2, −1, −1)
Table 2.9 49 Canonical Fano 3-topes with Fine Interior of Dimension 3. Table contains: support supp(FI ) of the Fine interior FI , vertices vert(can ) of
the canonical hull can , and order of fundamental group |π1 (S )| of the minimal model S
ID supp(FI ) vert(can ) |π1 (S )|
547444 (−2, −1, 2), (−1, 0, 0), (−1, 2, −1), (1, 1, −1), (3, 0, −1), (5, −1, −1) vert() 1
547465 (−1, −1, 3), (−1, 0, 0), (−1, 0, 1), (−1, 0, 2), (−1, 1, 0), (−1, 1, 1), (−1, 2, 0), (−1, 3, −1), vert() 2
(2, −1, −1)
547524 (−1, −2, 2), (−1, 1, 0), (−1, 2, −1), (0, 0, −1), (0, 1, −1), (0, 2, −1), (1, 0, −1), (1, 1, −1), vert(), (0, −1, −1) 1
(2, 0, −1), (3, 0, −1)
2 On the Fine Interior of Three-Dimensional …
547525 (−1, −2, 2), (−1, 2, −1), (0, −1, 0), (0, 0, −1), (0, 1, −1), (1, −1, −1), (1, −1, 0), (1, 0, −1), vert(), (1, 1, 1), (−1, −2, −3) 1
(1, 1, −1), (2, −1, −1), (2, −1, 0), (2, 0, −1), (3, −1, −1), (3, −1, 0), (3, 0, −1), (4, −1, −1),
(4, 0, −1), (5, −1, −1), (6, −1, −1)
545317 (−2, −2, −1), (−1, −1, 0), (−1, 2, 2), (1, −1, −1), (1, 2, 1), (3, 2, 0) vert() 1
545932 (−2, −1, −1), (−1, 0, 0), (−1, 2, −1), (0, 1, 0), (1, 1, 0), (2, 0, 1), (3, 0, 1), (5, −1, 2) vert(), (1, −1, −2), (1, 0, −3) 1
546013 (−2, −2, −1), (−1, 0, 1), (−1, 2, 2), (0, 1, 1), (1, 0, 0), (1, 2, 1), (2, 1, 0), (3, 0, −1), (3, 2, 0) vert() 2
546062 (−1, −1, 0), (−1, −1, 1), (−1, −1, 2), (−1, 0, 0), (−1, 0, 1), (−1, 1, 0), (−1, 2, −1), (0, −1, 0), vert() 1
(2, 1, −1)
(continued)
35
36
(1, −1, −1), (1, −1, 0), (1, 0, −2), (1, 0, −1), (2, −1, −2), (2, −1, −1), (2, −1, 0)
535952 (−2, −2, −1), (−1, −1, 0), (−1, 0, 1), (−1, 2, 2), (0, 1, 1), (1, −1, −1), (1, 0, 0), (1, 2, 1), vert() 1
(2, 1, 0), (3, 0, −1), (3, 2, 0)
536013 (−1, −2, 2), (−1, 2, −1), (0, −1, 0), (0, 0, −1), (0, 1, −1), (0, 2, −1), (1, −1, 0), (1, 0, −1), vert() 1
(1, 1, −1), (2, −1, 0), (2, 0, −1), (3, 0, −1)
536498 (−2, −1, −1), (−1, 0, 0), (−1, 2, −1), (0, 1, 0), (1, 1, 0), (1, 2, 0), (2, 0, 1), (2, 3, 0), vert() 1
(3, 0, 1), (3, 1, 1), (4, 2, 1), (5, −1, 2), (5, 0, 2), (6, 1, 2), (7, −1, 3), (8, 0, 3), (10, −1, 4)
537834 (−2, −1, −1), (−1, 0, 0), (−1, 2, −1), (−1, 3, −1), (0, 1, 0), (0, 4, −1), (1, 1, 0), (1, 2, 0), vert() 1
(2, 0, 1), (2, 3, 0), (3, 0, 1), (3, 1, 1), (4, 2, 1), (5, −1, 2), (5, 0, 2), (6, 1, 2), (7, −1, 3),
(8, 0, 3), (10, −1, 4)
(continued)
37
38
495687 (−2, −1, −1), (−1, 0, 0), (−1, 2, −1), (−1, 3, −1), (0, 1, 0), (0, 4, −1), (1, −1, 1), (1, 1, 0), vert() 1
(1, 2, 0), (2, 0, 1), (2, 3, 0), (3, 0, 1), (3, 1, 1), (4, 2, 1)
499287 (−2, 1, 1), (−1, −1, 1), (−1, 0, 0), (−1, 1, −1), (0, −1, 0), (0, 0, −1), (1, −1, −1), (1, −1, 0), vert(), (−1, −1, −1) 1
(1, 0, −1), (2, −1, −1), (2, −1, 0), (2, 0, −1), (3, −1, −1)
499291 (−2, 1, 1), (−1, −1, −1), (−1, −1, 0), (−1, −1, 1), (−1, 0, −1), (−1, 0, 0), (−1, 1, −1), vert() 1
(0, −1, −1), (0, −1, 0), (0, 0, −1), (1, −1, −1), (1, −1, 0), (1, 0, −1), (2, −1, −1),
(2, −1, 0), (2, 0, −1), (3, −1, −1)
499470 (−1, −1, 2), (−1, 0, 1), (−1, 1, 0), (−1, 1, 1), (−1, 2, −1), (−1, 2, 0), (−1, 3, −1), (−1, 3, 0), vert() 1
(0, −1, 0), (0, 1, −1), (2, −2, −1)
(continued)
39
40
(ID 547465, Fig. 2.6b, Tables 2.7, 2.8, and 2.9). The primitive inward-pointing facet
normals of the facets θ123 , θ124 , θ234 , θ134 of this simplex are
n 1 := (−1, −1, 3), n 2 := (−1, 3, −1), n 3 := (−1, 0, 0), and n 4 := (2, −1, −1),
n 1 + n 2 + 2n 3 + 2n 4 = 0.
unimodular 1-simplex and the double unimodular 2-simplex have empty Fine inte-
rior. Treutlein has found 9 maximal exceptional hollow polytopes, which was not
an complete list. Averkov et al. [1, 2] have found the complete list consisting of 12
maximal exceptional hollow 3-topes i (1 ≤ i ≤ 12) (Table 2.10, Fig. 2.7). Com-
putations show that exactly 9 of 12 maximal exceptional hollow 3-topes i have
non-empty Fine interior iFI (Table 2.10). Moreover, no one of these 9 polytopes
contains a proper lattice 3-subpolytope with non-empty Fine interior. Thus, there
exist exactly 9 hollow 3-topes i with non-empty Fine interior iFI .
It is remarkable that all minimal surfaces Si corresponding to these 9 hollow 3-
topes i have non-trivial fundamental group π1 (S ) of order 2, 3, or 5 (Table 2.10).
There exist exactly 5 hollow 3-topes i with 0-dimensional Fine interior iFI = {R},
where R ∈ 21 M \ M is a rational point (Table 2.10). The normal fans i of these 5
hollow polytopes i define 5 toric Fano threefolds X i with at worst canonical
singularities (Table 2.11). These Fano threefolds can be obtained as quotients of
Gorenstein toric Fano threefolds X in the following 5 ways:
In addition, Table 2.12 contains the support supp(iFI ) of the Fine interior iFI and
the vertices of the canonical hull ican for all 9 hollow polytopes i with non-empty
Fine interior iFI .
6 (0, 0, 0), (2, 2, 0), (4, 0, 0), (2, −2, 0), (3, 1, 2) 2 0 1/2 · (5, 1, 2) 2
7 (0, 0, 0), (1, 1, 0), (2, −2, 0), (3, −1, 0), (1, −1, 2), (2, 0, 2) 2 0 1/2 · (3, −1, 2) 2
8 (0, 0, 0), (1, 1, 0), (1, −1, 0), (2, 0, 0), (1, −1, 2), (2, 0, 2), 2 0 1/2 · (3, −1, 2) 2
(2, −2, 2), (3, −1, 2)
9 (0, 0, 0), (3, 0, 0), (1, 3, 0), (2, 0, 3) 3 1 (4/3, 1, 1), (5/3, 1, 1) 3
10 (0, 0, 0), (1, 2, 0), (1, −1, 0), (3, 0, 0), (2, 1, 3) 3 1 (4/3, 2/3, 1), (5/3, 1/3, 1) 3
11 (0, 0, 0), (1, 1, 0), (3, 0, 0), (2, −1, 0), (4, 1, 3), (2, 2, 3) 3 1 (5/3, 2/3, 1), (7/3, 1/3, 1) 3
12 (−1, 0, 0), (0, 1, −2), (1, 2, 1), (2, −2, −1) 3 3 (1/5, 1/5, −2/5), (2/5, 2/5, −4/5), (3/5, 3/5, −1/5), 5
(4/5, −1/5, −3/5)
43
44 V. Batyrev et al.
P4 P4
P4
P2
P2
P1
P1
P3
P2
P3 P3 P1
P4 P4
P3
P5
P4
P1
P2 P1 P1
P3
P3 P2 P2
P4
P5
P6
P6
P5 P8
P7
P3
P2
P4 P3 P2
P1 P1 P4
P2 P3 P1
P5
P3
P6 P5
P1
P4
P2
P2 P2
P3 P3
P1 P1 P4 P4
(j) 10 . (k) 11 . (l) 12 .
Fig. 2.7 12 Maximal Hollow 3-topes. Shaded faces are occluded. The Fine interior is shown in
grey with double borders around its vertices
Table 2.11 5 Hollow 3-topes with 0-dimensional Fine Interior. Table contains: index i of the maximal hollow 3-tope i , rays of the normal fan i
corresponding to i , ID of the canonical Fano 3-tope such that i ∼ = , rays of the spanning fan , ID of the reflexive canonical Fano 3-tope
used to construct the Gorenstein toric Fano threefold X to obtain the toric Fano threefold X with at worst canonical singularities as a μ2 quotient, and
reference to the corresponding Gorenstein toric Fano threefold X including the precise μ2 -action on Sect. 2.7
i i (1) ID( ) (1) ID( ) X
4 (2, −1, −3), (0, 0, 1), (0, 1, 0), (−2, −1, −1) 547354 (−2, −3, −5), (2, 1, 1), (0, 1, 0), (0, 0, 1) 547363 (i)
2 On the Fine Interior of Three-Dimensional …
5 (1, −1, 0), (1, 1, −1), (−1, 1, 2), (−1, −1, −1) 547364 (0, 0, 1), (0, 2, 1), (2, 1, 0), (−2, −3, −2) 547367 (ii)
6 (0, 0, 1), (1, 1, −2), (1, −1, −1), (−1, −1, 0), (−1, 1, −1) 544353 (1, 0, 0), (0, 1, 0), (−2, −1, 0), (1, 1, 2), (−3, −1, −2) 544357 (iii)
7 (0, 0, 1), (−1, 1, −1), (1, −1, −1), (1, 1, 0), (−1, −1, 0) 544310 (−1, −1, −2), (1, 1, 2), (−2, −1, 0), (0, 1, 0), (1, 0, 0) 544342 (iv)
8 (−1, 1, 1), (0, 0, −1), (−1, −1, 0), (1, 1, 0), (1, −1, −1), (0, 0, 1) 520134 (1, 0, 0), (0, 1, 0), (1, 1, 2), (−1, 0, 0), (0, −1, 0), 520140 (v)
(−1, −1, −2)
45
46 V. Batyrev et al.
Table 2.12 9 Hollow 3-topes with Non-empty Fine Interior. Table contains: index i of the
maximal hollow 3-tope i , support supp(iFI ) of iFI , and vertices vert(ican ) of the canonical
hull ican
i supp(iFI ) vert(ican )
4 (−2, −1, −1), (0, −1, −2), (2, −1, −3), (0, 0, 1), (0, 0, −1), (0, 1, 0) vert(i )
5 (1, −1, 0), (1, 1, −1), (0, 0, 1), (0, 0, −1), (−1, −1, −1), (−1, 1, 2) vert(i )
6 (1, 1, −2), (1, −1, −1), (−1, −1, 0), (−1, 1, −1), (0, 0, 1), (0, 0, −1) vert(i )
7 (1, 1, 0), (1, −1, −1), (−1, −1, 0), (−1, 1, −1), (0, 0, 1), (0, 0, −1) vert(i )
8 (1, 1, 0), (1, −1, −1), (−1, −1, 0), (0, 0, 1), (0, 0, −1), (−1, 1, 1) vert(i )
9 (0, −1, −1), (0, 0, 1), (3, −1, −2), (0, 1, 0), (−3, −2, −1) vert(i )
10 (−1, 2, −1), (1, 1, −1), (−1, −1, 0), (2, −1, −1), (0, 0, 1) vert(i )
11 (1, −1, 0), (0, 0, 1), (−1, −2, 1), (−1, 1, 0), (1, 2, −2) vert(i )
12 (1, 1, 1), (1, −1, 0), (−2, −1, 1), (0, 1, −2) vert(i )
References
1. Averkov, G., Krümpelmann, J., Weltge, S.: Notions of maximality for integral lattice-free
polyhedra: the case of dimension three. Math. Oper. Res. 42(4), 1035–1062 (2017)
2. Averkov, G., Wagner, C., Weismantel, R.: Maximal lattice-free polyhedra: finiteness and an
explicit description in dimension three. Math. Oper. Res. 36(4), 721–742 (2011)
3. Batyrev, V.: The stringy Euler number of Calabi-Yau hypersurfaces in toric varieties and the
Mavlyutov duality. Pure Appl. Math. Q. 13(1), 1–47 (2017)
4. Batyrev, V., Kreuzer, M.: Integral cohomology and mirror symmetry for Calabi-Yau 3-folds.
In: Mirror symmetry. V, AMS/IP Studies in Advanced Mathematics, vol. 38, pp. 255–270.
American Mathematical Society, Providence, RI (2006)
5. Batyrev, V.V.: Dual polyhedra and mirror symmetry for Calabi-Yau hypersurfaces in toric
varieties. J. Algebraic Geom. 3(3), 493–535 (1994)
6. Beck, M., Robins, S.: Computing the continuous discretely. Undergraduate Texts in Mathe-
matics, 2nd edn. Springer, New York (2015). Integer-point enumeration in polyhedra, With
illustrations by David Austin
7. Bosma, W., Cannon, J., Playoust, C.: The Magma algebra system. I. The User Language, pp.
235–265 (1997). Computational algebra and number theory (London, 1993)
8. Brown, G., Kasprzyk, A.M.: Graded Ring Database. http://www.grdb.co.uk
9. Catanese, F.: Surfaces with K 2 = pg = 1 and their period mapping. In: Algebraic Geometry
(Proc. Summer Meeting, Univ. Copenhagen, Copenhagen, 1978). Lecture Notes in Mathemat-
ics, vol. 732, pp. 1–29. Springer, Berlin (1979)
10. Catanese, F., Debarre, O.: Surfaces with K 2 = 2, pg = 1, q = 0. J. Reine Angew. Math. 395,
1–55 (1989)
11. Corti, A., Golyshev, V.: Hypergeometric equations and weighted projective spaces. Sci. China
Math. 54(8), 1577–1590 (2011)
12. Danilov, V.I., Khovanskiı̆, A.G.: Newton polyhedra and an algorithm for calculating Hodge-
Deligne numbers. Izv. Akad. Nauk SSSR Ser. Mat. 50(5), 925–945 (1986)
13. Fine, J.: Resolution and completion of algebraic varieties. Ph.D. thesis, University of Warwick
(1983)
14. Hovanskiı̆, A.G.: Newton polyhedra, and the genus of complete intersections. Funktsional.
Anal. i Prilozhen. 12(1), 51–61 (1978)
15. Ishii, S.: The minimal model theorem for divisors of toric varieties. Tohoku Math. J. 51(2),
213–226 (1999)
16. Kasprzyk, A.M.: Canonical toric Fano threefolds. Canad. J. Math. 62(6), 1293–1309 (2010)
2 On the Fine Interior of Three-Dimensional … 47
17. Kreuzer, M., Skarke, H.: Classification of reflexive polyhedra in three dimensions. Adv. Theor.
Math. Phys. 2(4), 853–871 (1998)
18. Kreuzer, M., Skarke, H.: Complete classification of reflexive polyhedra in four dimensions.
Adv. Theor. Math. Phys. 4(6), 1209–1230 (2000)
19. Kynev, V.I.: An example of a simply connected surface of general type for which the local
Torelli theorem does not hold. C. R. Acad. Bulgare Sci. 30(3), 323–325 (1977)
20. Reid, M.: Young person’s guide to canonical singularities. In: Algebraic Geometry, Bowdoin,
1985 (Brunswick, Maine, 1985), Proceedings of Symposia in Pure Mathematics, vol. 46, pp.
345–414. American Mathematical Societ, Providence, RI (1987)
21. Skarke, H.: Weight systems for toric Calabi-Yau varieties and reflexivity of Newton polyhedra.
Modern Phys. Lett. A 11(20), 1637–1652 (1996)
22. Stanley, R.P.: Decompositions of rational convex polytopes. Ann. Discrete Math. 6, 333–342
(1980)
23. Todorov, A.N.: Surfaces of general type with pg = 1 and (K , K ) = 1. I. Ann. Sci. École Norm.
Sup. 13(1, 4), 1–21 (1980)
24. Todorov, A.N.: A construction of surfaces with pg = 1, q = 0 and 2 ≤ (K 2 ) ≤ 8. Counterex-
amples of the global Torelli theorem. Invent. Math. 63(2), 287–304 (1981)
25. Treutlein, J.: 3-dimensional lattice polytopes without interior lattice points. Ph.D. thesis, Eber-
hart Karls Universität Tübingen (2010)
Chapter 3
Lattice Distances in 3-Dimensional
Quantum Jumps
Mónica Blanco
Abstract For Q a lattice polytope and x ∈ / Q a lattice point, we say that (Q, x) is
a quantum jump if conv(Q ∪ {x}) contains exactly one more lattice point than Q.
Usually this can only happen when the lattice distance between x and Q is somehow
bounded. In this paper I collect several results and information on the bound for
that distance in 3-dimensional quantum jumps, and the consequences on the distance
between the boundary of a polytope and its interior lattice points.
3.1 Introduction
M. Blanco (B)
Department of Mathematics, Statistics and Computation, University of Cantabria, Avda. Los
Castros s/n, 39005 Santander, Cantabria, Spain
e-mail: [email protected]
Finally, for S ⊂ Rd , we denote by conv(S) and aff(S) the convex and affine hulls
of S. In particular, an affine subspace S ⊂ Rd is lattice if aff(S ∩ Zd ) = S.
Let us now introduce the main two definitions in this paper:
The name of quantum jump was first used by Bruns, Gubeladze, and Michałek [5].
Notice that they restrict the concept of quantum jump (Q, x) for when both Q
and conv(Q ∪ {x}) are full-dimensional and normal. Remember that a lattice d-
polytope Q is normal if, for all k ∈ N, every lattice point in k Q can be written as the
sum of k lattice points in Q.
Now, if we want to take a look at the distance of quantum jumps, we first need to
define the distance between a point and a polytope. Following Definition 1, they are
well and naturally defined the following distances:
Definition 3 1. Let Q ⊂ Rd be a lattice (d − 1)-polytope, and let x ∈ Zd \ aff(Q),
then
dist(x, Q) := dist(x, aff(Q)).
dist(1 , 2 ) := dist(H1 , H2 ),
where H1 , H2 are the unique pair of parallel lattice hyperplanes such that i ⊂ Hi .
4. Let s1 , s2 ⊂ R3 be lattice segments such that aff(s1 ∪ s2 ) = R3 , then
Q Q Q
x x x
Fig. 3.1 The three figures show different facets of a lattice polygon Q, the hyperplane they are
contained in, and the lattice point x. Only the facet in the middle figure is visible from x
Notice that the width and the distance are heavily related. In broad terms, the dis-
tance between two lower-dimensional objects R1 , R2 ⊂ Rd with aff(R1 ∪ R2 ) = Rd
is the width of conv(R1 ∪ R2 ) with respect to a specific functional that is determined
by the relative position between R1 and R2 . In general, if R := conv(R1 ∪ R2 ) ⊂ Rd
is not full-dimensional, the distance between R1 and R2 is measured in the lat-
tice aff(R) ∩ Zd ∼ = Zdim(R) . In the case of lattice segments, we call (lattice) length of
a segment the distance between its two endpoints (vertices). Notice that a lattice seg-
ment of length k has exactly k + 1 lattice points. We say that a segment is primitive
if it has length one.
Now, the distance that is not necessarily well-defined is the distance from a point
to a full-dimensional polytope. This notion will be written in terms of the distance
to the visible facets (see Fig. 3.1):
Definition 4 Let Q ⊂ Rd be a lattice d-polytope, F ⊂ Q a facet of Q and x ∈
Zd \ Q. F is visible from x if aff(F) strictly separates x from Q.
See Fig. 3.2 for a 2-dimensional example of the maximum and minimum distances
between a point and a polygon. Notice that Dx (Q) is the height of x over Q as in [5,
Definition 4.1].
Let us see what we know about the distance of quantum jumps in each dimension.
For this, we can also think of a quantum jump as follows: any d-dimensional quantum
jump is of the form (P v , v), for P ⊂ Rd a lattice d-polytope, v ∈ vert(P) a vertex of P
and P v := conv(P ∩ Zd \ {v}). Notice that the dimension of P v can be d or d − 1.
For example, the lattice distance in quantum jumps of dimension ≤ 2 is always
one:
52 M. Blanco
Lemma 6 Let P be a lattice polytope of dimension d ∈ {1, 2}, and let v be a vertex
of P.
If P v is of dimension d − 1, then dist(v, P v ) = 1 and, if P v is of dimension d, we
have dv (P v ) = Dv (P v ) = 1.
To understand the idea in general: let P v be d-dimensional. For any (d − 1)-
dimensional face of P v that is visible from v, chose S an empty (d − 1)-dimensional
simplex in it. Since (P v , v) is a quantum jump, so is (S, v), which implies that the
convex hull of S and v is an empty d-simplex. Remember that an empty simplex of
dimension d is a lattice d-polytope with d + 1 vertices and such that those vertices
are its only lattice points.
In the cases of d = 1, 2, any empty simplex has to be unimodular, hence the vertex-
facet distance (lattice distance between a vertex and the only facet that does not
contain it) is always 1. In dimension 3 things get more complicated since we have
empty tetrahedra of arbitrarily high volume, and hence arbitrarily high vertex-facet
distance (e.g. Reeve tetrahedra [8]). That is, quantum jumps between a unimodular
triangle and a lattice point that is at arbitrarily high lattice distance from it.
In Sect. 3.2 of this paper I put together some information on the lattice dis-
tance of 3-dimensional quantum jumps (Q, x) that derives partially from previous
research [2–4]. We distinguish when Q is 2 or 3-dimensional:
1. If Q is 2-dimensional (Sect. 3.2.1) it so happens that the classifications of lattice 3-
polytopes of size 5 and 6 [2, 3], together with a suitable classification of lattice
polygons, give all the information there is to know about the distance from Q
to x. It can be summarized as follows:
Theorem 7 (see Corollary 14) Let Q ⊂ R3 be a lattice polygon, and let x ∈ Z3 \
aff(Q) such that (Q, x) is a quantum jump. Then, the lattice distance from x to Q
is at most 3 unless Q is a lattice triangle of width one, in which case the distance is
unbounded.
2. As a direct consequence of the results of the previous section, in Sect. 3.2.2 we
have the following result on the distance of a quantum union of lattice segments:
Theorem 8 (see Corollary 16) Let s, t ⊂ R3 be lattice segments with aff(s ∪ t) =
R3 such that (s, t) is a quantum union. Then, the lattice distance from s to t is one,
unless both s and t are primitive, in which case it is unbounded.
3 Lattice Distances in 3-Dimensional Quantum Jumps 53
3. For the case of Q being 3-dimensional (Sect. 3.2.3), Example 18 shows that there
exist quantum jumps of this type at arbitrarily high distance. We also look at
all the lattice 3-polytopes P of size 11 and width > 1 (database of [4]), and for
each vertex v of P such that P v is 3-dimensional we compute the minimum and
maximum distances from v to P v . Looking at the numbers one can easily see that
there is no hope in trying to bound these distances, having very high numbers for
both the minimum and the maximum distances.
Finally, in Sect. 3.3 I use all the information gathered in the previous section to
study the distance between the boundary of a polytope and its interior. More precisely,
for P ⊂ R3 a lattice 3-polytope we look at the distance between a lattice point or
segment in ∂ P (the boundary of P) and the inner lattice polytope of P, which
is I P := conv{int(P) ∩ Z3 }. Notice that I P together with a point (or segment) of the
boundary is always a quantum jump (or union). The definition of inner polytope also
applies to rational polytopes.
We only look at inner polytopes I P of size ≥ 3 (see Remark 20), and we separate
cases according to its dimension:
1. I P has dimension 1, that is, it is a lattice segment of length ≥ 2. In this case we
look at the distance between I P and a segment in the boundary. By Corollary 16,
this distance must be one, leading to:
Theorem 9 (see Theorem 21) The projection of P in the direction of the segment I P
is a reflexive polygon (polygon with a unique interior lattice point).
2. I P has dimension 2. In Sect. 3.3.2 we prove a specific property that a polygon
has to satisfy in order to appear as the inner polygon of a 3-dimensional lattice
polytope (see Theorem 23). Together with the results of Corollary 14 we obtain:
Corollary 10 (see Corollary 24) For I P of dimension 2 and size ≥ 12, the distance
from any boundary point of P to I P is at most 1.
3. I P has dimension 3. Again we look at the classification of lattice 3-polytopes
of size ≤ 11 and width > 1 [4], take the polytopes with 3-dimensional inner
polytope, and look at the minimum and maximum distances from any vertex
to the inner polytope. In Sect. 3.3.3 we simply collect some information on the
numbers obtained, without exploring it further. This time the values look more
promising, since the largest value that appears is a maximum distance of 6, and
in very high proportion the maximum and minimum distances are 1.
For future work one could try and complete the results on distances between a
lattice point of the boundary of a polytope, and its 2 or 3-dimensional inner polytope.
For the inner polytope of dimension 2, it is left to explore the cases when I P has up
to 11 lattice points. This seems perfectly doable with the help of the classification
of polygons of Proposition 11, together with the results of Sect. 3.2.1. On the other
hand, for I P of dimension 3, one would have to identify in the used database all the
polytopes that yield maximum and minimum distances equal to 1 and try to derive
the properties they have as opposed to those that yield larger distances. One would
have then to try and extend this to lattice 3-polytopes of size larger than 11.
54 M. Blanco
Δ2 T1 T2
o o o
F1 (k) F2 (k)
o (k, 0) o (k, 0)
F3 (k) F4 (k , k)
o (k, 0) (−k , 0) o (k − k , 0)
Fig. 3.3 The complete classification of lattice polygons that do not contain a unit square
We first see at what distance can a lattice point be from a lattice polygon, so that
they form a quantum jump. For this, we first classify lattice polygons in a way that
is suitable distance-wise. In the following lemma, we call unit square any lattice
polygon unimodularly equivalent to [0, 1]2 .
Proposition 11 Let Q ⊂ R2 be a lattice polygon. Then Q either contains a unit
square or is equivalent to one of the following configurations:
1. 2 , the unimodular triangle;
2. T1 := conv{(1, 0), (0, 1), (−1, −1)}, the unique terminal triangle;
3. T2 := conv{(2, 0), (0, 1), (−1, −2)}, a clean triangle with three non-collinear
interior lattice points;
4. F1 (k) := conv{(0, 0), (0, 1), (k, 0)}, for k ≥ 2;
5. F2 (k) := conv{(0, 1), (0, −1), (k, 0)}, for k ≥ 2;
6. F3 (k) := conv{(−1, −1), (0, 1), (k, 0)}, for k ≥ 2; or
7. F4 (k , k) := conv{(0, 1), (0, −1), (−k , 0), (k − k , 0)}, for 0 < k < k.
See Fig. 3.3 for a depiction of the polygons of Proposition 11. For its proof, let us
first establish the following notation.
3 Lattice Distances in 3-Dimensional Quantum Jumps 55
T x
1. Suppose the three collinear points (−1, 0), (0, 0) and (1, 0) are in a facet of Q.
Then Q ⊂ {y ≥ 0}. Since Q does not contain a unit square, the points (−1, 1)
and (1, 1) cannot be in Q. Moreover, by Remark 12 this implies that no point in
the affine cones C(−1,1) (T ) and C(1,1) (T ) is in Q. See Fig. 3.4. That means that
the only lattice points that can lie in Q \ T are in the following sets:
A := (i, 0), i ∈ Z \ [−1, 1] , B := (0, j), j ∈ Z, j ≥ 2
Q can contain points of A or points of B, but in order for (1, 1) and (−1, 1) not
to be in Q, it cannot contain points of A and B at the same time. Adding points
of A to T gives rise to polygons of the type F1 (k), and adding points of B gives
rise to F2 (k).
2. If no facet of Q contains three collinear points, then the origin, which is in the
relative interior of the segment conv{(−1, 0), (1, 0)} ⊂ Q, must be an interior
point of Q. So Q must contain some lattice point in {y < 0} and, by Lemma 6,
56 M. Blanco
The points in A and B cannot be in Q at the same time, and each gives rise
to a configuration equivalent to T2 . The points in C or D cannot be in Q at
the same time as the points in A or B. If Q has points of D, we have configu-
rations F4 (k , k) and, if Q only has points of C we get configurations F2 (k).
c. p− = (0, −1). In this case, T = F4 (1, 2). After excluding the points in the
cones with apex in {−1, 1}2 , Q can have other lattice points in:
Q cannot have points of A and B at the same time, and adding to T points
of either A or B gives rise to configurations equivalent to F4 (k , k).
3 Lattice Distances in 3-Dimensional Quantum Jumps 57
y y y
(−1, 2) (0, s)
(−1, −2)
Fig. 3.6 The three possible polygons T in the proof of Proposition 11. In each figure, the dark
gray area is the polygon T (or equivalent). The light gray area is the union of the cones that do not
intersect Q. Black dots are lattice points of T , crosses are lattice points that cannot be in Q, and
white dots are the possible lattice points of Q \ T
Let us now take that classification and see what are the conditions on the coordi-
nates of a lattice point x ∈ Z3 so that a polygon Q and x ∈/ aff(Q) form a quantum
jump.
For the purpose of simplifying notation in Lemma 13 and its proof, let us denote by
Q (resp. R) both the lattice polygon in R2 and its embedding Q × {0} (resp. R × {0})
in R3 .
Proof Part 1 of the statement follows from the classification of empty tetrahedra [9],
which states that a lattice tetrahedron is empty if one of the three pairs of opposite
edges are at lattice distance one. It is also required that these opposite edges are
primitive segments (gcd condition in the statement).
In each of the cases 2–6, we choose a subpolygon R of Q of size 4 or 5:
2. R is the unit square in Q;
3. R := Q of size 4;
4. R := F1 (2) ⊆ Q of size 4;
5. R := F3 (2) ⊆ Q, of size 5;
58 M. Blanco
In terms of the distance from x to Q, which in Lemma 13 is the value |c|, we have
the following result.
Notice that the only cases when the distance is unbounded are Q ∼
= 2 or Q ∼
=
F1 (k), that is, when Q is a triangle of width one.
3 Lattice Distances in 3-Dimensional Quantum Jumps 59
The two cases where the distance is unbounded in the previous section also have in
common that all the lattice points are along two lattice segments. Let us think about
them as quantum unions of lattice segments.
Cases 1 and 4 of Lemma 13, reformulated in terms of the distance between seg-
ments that form a quantum union, are as follows:
For the case when Q ⊂ R3 is a lattice 3-polytope, remember that we defined the
distance from a point x ∈ Z3 \ Q to Q in terms of the distance to the facets of Q
that are visible from x (Definition 5). In particular, one can study the distance from x
to Q, for (Q, x) a quantum jump, by combining the results of the previous section
on the facets of Q that are visible from x.
3 23
so that, for all x ∈ RI , the facets of Q that are visible from x are exactly those of I.
Notice that the closures of the regions RI are rational polytopes or polyhedra. See
Fig. 3.7 for a 2-dimensional example of this subdivision of the space.
Let RI be one of those regions and suppose that x ∈ RI ∩ Z3 is such that (Q, x)
is a quantum jump. We can have three different types of situations.
1. If RI is bounded (or if RI ∩ Z3 is finite), the distance of x to Q is automatically
bounded.
2. If RI ∩ Z3 has infinitely many points (in particular RI is unbounded), and some
facet of I is not a triangle of width one, then the distance of x to Q is bounded
by the results of the previous section (Corollary 14).
3. Finally, if RI ∩ Z3 has infinitely many points and all the facets in I are triangles
of width one, the distance from x to Q may not be bounded. Notice that, even in
this last case the distance from x to Q could still be bounded by combining the
restrictions given for the coordinates of x as in parts 1 and 4 of Lemma 13 for all
the different facets of I.
For instance, we can find arbitrarily high distance in these types of quantum jumps.
We can also choose Q of size n, for any n ≥ 4, since the regions C, C x or C y have
infinitely many lattice points. See Fig. 3.8 for a depiction of the polyhedral regions C
and C y .
Even though it is clear that arbitrarily bad examples can occur, how often does
this happen? Are they rare or does the general picture look bad? For this, we look at
our database of lattice 3-polytopes of size ≤ 11 and width > 1 [4]. For each of those
polytopes P ⊂ R3 and for each v ∈ vert(P) such that P v is full dimensional, we look
at the distance in the quantum jump (P v , v). Notice that our database contains all the
information on the types of quantum jumps when P v is of size ≤ 10 and extends to
a polytope P of width > 1. That is, of size ≤ 10, we do not have the information on
polytopes of width one that extend to polytopes of width one, which are infinitely
many for each size (and no enumeration exists).
For each quantum jump (P v , v) we compute the values dv (P v ) and Dv (P v ) and
store the following vectors:
1. d P := dv (P v ) v∈vert(P),P v full-dimensional
2. D P := Dv (P v ) v∈vert(P),P v full-dimensional
We separate the 216, 453 polytopes of our database in three different groups. Notice
that the entries of each vector d P and D P are positive integers.
1. d P = (1, 1, . . . , 1) = D P . This is the best case scenario we can find, since every
vertex v of P, with P v full-dimensional, is at distance one from all the facets
of P v that are visible from v. However, only 5,796 polytopes (about 2.7%) fall
into this category.
2. d P = (1, 1, . . . , 1), D P
= (1, 1, . . . , 1). In this case, things are not as nice, but
we still have that every vertex v of P, with P v full-dimensional, is at distance one
62 M. Blanco
from at least one facet of P v that is visible from v. In this category we have 77,443
polytopes (∼35.8%).
3. d P , D P
= (1, 1, . . . , 1). This is the worst case we can have, in which some
vertex v of P, with P v full-dimensional, is at distance larger than one from all
the facets of P v that are visible form v. This is the case for most of the polytopes
in our database: 133,214 polytopes, or ∼61.5% of the total.
In terms of the magnitudes of the entries, we have that the largest entries in the
vectors d P and D P , for each n the size of P, are:
n 5 6 7 8 9 10 11
max dv (P v ) 5 7 13 19 25 31 37
max Dv (P v ) 7 13 19 25 31 37 43
Notice that the maximum values for dv (P v ) and Dv (P v ), for P of size n, are 6(n −
5) + 1 and 6(n − 4) + 1, respectively (for n
= 5 in the first case). This has to do with
the fact that, as h grows (see Example 18) we need more lattice points to construct a
polytope of width > 1 that yields a vertex at distance h.
The average values of the dv (I P ) and Dv (I P ) are, respectively, 1.42 and 3.35.
Remark 19 If we were to follow the lines of Sect. 3.2.1, we would want to have,
in this section, an irredundant list of lattice 3-polytopes Q, and the maximum and
minimum distances a point x can be from Q, for (Q, x) a quantum jump.
However, we need to consider that we have 216,453 polytopes and that, for each of
those polytopes P and each vertex v of P we have a different polytope P v . Organizing
the information on the distances with no redundancies among the P v does not seem
to be worth undertaking, in light of the distances that appear and the arguments made.
that there are only finitely many lattice 3-polytopes with 1 and 2 interior lattice points,
but the author does not believe it is worth exploring the more than 23 million such
polytopes.
So let P be such that I P has size at least three. Let S ⊂ ∂ P be a lattice point
or primitive segment in the boundary of P, we look at the distance between S
and I P , relying on the fact that (I P , S) is a quantum jump (or union). This hap-
pens because conv(I P ∪ S) \ S ⊂ int(P), and the only interior lattice points of P
are those of I P .
If I P is a lattice segment (see Fig. 3.9), and since I P has size at least 3, by Corol-
lary 16, the distance from I P to any lattice segment in the boundary must be one. A
consequence of this is the following result1 :
Proof Since k > 0, the projection π is well defined and unique, up to unimodular
transformation. Because the k + 1 collinear lattice points are in the interior of P,
their projection, i. e. the origin, is an interior point of π(P). Let e be an edge of π(P),
then there exists a lattice segment e in the boundary of P such that π(e ) = e. Take
the following polytope Re := conv(I P ∪ e ) ⊂ P. Since e ⊂ ∂ P and I P ⊂ int(P),
then Re cannot contain any extra lattice points. That is, it is the quantum union
of two lattice segments. By Corollary 16, and since I P is not primitive, the distance
between I P and e must be one. In the projection, this directly implies that the distance
from the edge e and the origin (the respective projections of the segments) is one.
Hence π(P) is reflexive.
1Discussed in the Oberwolfach mini-workshop Lattice polytopes: Methods, Advances and Appli-
cations, September 2017.
64 M. Blanco
This result can help, for example, in the full classification of lattice 3-polytopes P
with I P a lattice segment. The projection has 16 possibilities: the 16 reflexive poly-
gons. For one such Q fixed, all the lattice points in P must be in π −1 (Q).
For P having inner polytope I P of dimension 2 (see Fig. 3.10), our main result resides
in proving a specific property that a polygon must have so that it can actually appear
as the inner polytope of a lattice 3-polytope. For this, let us introduce the concept
of front:
Definition 22 Let Q ⊂ Z2 be a lattice polygon and let v be a vertex of Q. A front
of Q from v is a facet of the polygon Q v := conv(Q \ {v} ∩ Z2 ) that is visible from v.
See Fig. 3.11 for an example of the fronts of a polygon.
Theorem 23 If P is a lattice 3-polytope with I P of dimension 2, then the fronts of I P
have length ≤ 8.
Proof Let F ⊂ I P be the longest front of I P , of length > 0. We can assume with-
out loss of generality that I P ⊂ {z = 0}, ≥ 3, that v := (0, 1, 0) is a vertex of I P
Q Q
3 Lattice Distances in 3-Dimensional Quantum Jumps 65
r1 r2 ()
v0
v
T
x
o (, 0)
and F := conv{(0, 0, 0), (, 0, 0)} (that is, F is a front of I P from v). In particular, we
have that T := F1 () × {0} = conv(F ∪ {v}) ⊆ I P (for F1 () as in Proposition 11)
and that the only lattice points of I P in {y ≥ 0} are those of F and v. We need to
prove that ≤ 8.
The intersection of P with the plane {z = 0} is the rational polygon P0 := P ∩
{z = 0}. We have that I P = conv(relint(P0 ) ∩ Z3 ). That is, the inner polytope of P
coincides with the relative inner polygon of P0 . For now let us identify R2 × {0}
and R2 in the trivial way, so from now on we simply say interior of P0 for the
relative interior of it embedded in the space R3 .
The vertex v = (0, 1) of I P is an interior point of P0 , so for any line passing
through v there must be a vertex of P0 in each of the open halfspaces determined by
this line. In particular, there must be a vertex of P0 in the open halfspace {y > 1}. Let
us denote this vertex by v0 and consider the rational polygon T := conv(T ∪ {v0 }).
Since T ⊆ I P ⊂ int(P0 ) and v0 ∈ ∂ P0 , then T \ {v0 } ⊂ int(P0 ). That is, the only
lattice points of T \ {v0 } are those of T . In particular, (−1, 1), (1, 1) ∈ / T \ {v0 },
which implies that v0 ∈ / C(−1,1) (T ) ∪ C(1,1) (T ) (see Remark 12).
This in turn implies that v0 must lie in the open rational triangle R determined
by the hyperplanes {y = 1}, r1 := aff{(0, 0), (−1, 1)} = {x + y = 0} and r2 () :=
aff{(, 0), (1, 1)} = {x + ( − 1)y = } (see Fig. 3.12).
That is,
−
v0 ∈ R = int conv (−1, 1), (1, 1), , ,
−2 −2
Fig. 3.13 The intersection K 0 of K with the plane {z = 0}, and the projection of K under π
3 Lattice Distances in 3-Dimensional Quantum Jumps 67
of a lattice segment of length 2 and a primitive lattice segment uw. By Corollary 16,
the lattice distance between these two lattice segments must be one. Taking again
projection π , this is equivalent to the segment π(uw) being at distance one from the
origin. Which is impossible since (0, 1) is a lattice point strictly in between π(uw)
and the origin (see the picture in the right of Fig. 3.13). In particular, v is not a vertex
of K 0 and the picture on the left of Fig. 3.13 is not accurate.
That is, the oriented matroid of the six lattice points of K can be described as
follows:
1. four of them are vertices (o, (2, 0, 0), u and w);
2. one non-vertex point is in an edge ((1, 0, 0) = 21 ((0, 0, 0) + (2, 0, 0)));
3. the hyperplane containing the three collinear points and the other non-vertex point
({z = 0}) leaves the remaining two vertices (u and w) strictly in opposite sides of
it.
Notice that there are four different types (or orbits) of points: the endpoints of
the collinearity (o and (2, 0, 0)), the middle point of the collinearity ((1, 0, 0)),
the other two vertices (u and w), and the remaining non-vertex point (v). This
sixth point v has three different possibilities, in terms of the oriented matroid.
The three possibilities for v are: (I) it is in the relative interior of one of the
facets conv{o, u, w} or conv{(2, 0, 0), u, w}; (II) it is in the relative interior of the
triangle conv{(1, 0, 0), u, w}; or (III) it lies in the interior of one of the tetrahe-
dra conv{o, (1, 0, 0), u, w} or conv{(1, 0, 0), (2, 0, 0), u, w}. These three options
are shown in Fig. 3.14. For each of these three cases, the oriented matroid is fully
described. Without going into details of how the oriented matroids are represented and
classified in [3], one can derive that the oriented matroid of the three options (I), (II)
and (III) are, respectively, oriented matroids 3.6, 3.8 and 4.11 as encoded in [3,
Fig. 1].
In [3, Tables 8 and 9] we can see that the only lattice 3-polytopes of size 6,
width > 1 and with one of the three specified oriented matroids are B.7 (oriented
matroid 3.8), C.1 (oriented matroid 3.6), and F.13 to F.17 (oriented matroid 4.11).
The following 3 × 6 matrices have, as columns, the six lattice points of each of those
seven polytopes:
Fig. 3.14 The three possibilities (I), (II), and (III) in the proof of Theorem 23
68 M. Blanco
⎛ ⎞
0 1 0 −2 1 4
⎛ ⎞ F.13 ⎝ 0 1 0 −1 0 1 ⎠ ⎛ ⎞
0 1 0 −1 0 0 0 1 1 −2 0 2 0 1 −1 0 1 2
B.7 0 0 1 −1 0 2 ⎠
⎝ ⎛ ⎞ F.16 0 0 −2 0 3 6 ⎠
⎝
0 0 0 0 1 −1 0 0 −1 1 1 1 0 0 −1 1 1 1
F.14 0 0 −1 2 0 −2 ⎠
⎝
⎛ ⎞ 0 1 −1 1 0 −1 ⎛ ⎞
0 1 0 −1 0 −1 0 1 1 0 −1 −2
⎛ ⎞
C.1 0 0 1 −1 0 0 ⎠
⎝ 0 0 1 1 −1 −3 F.17 0 0 3 0 −2 −4 ⎠
⎝
000 0 1 2 F.15 0 0 0 2 −1 −4 ⎠
⎝ 0 0 1 1 −1 −3
0 1 0 1 −1 −3
That is, our polytope K must be equivalent to one of them, say K̃ , and let t :
R3 → R3 be any unimodular transformation that maps K to K̃ . Then t will send the
edge conv{(0, 0, 0), (2, 0, 0)} to the unique collinearity of three lattice points in K̃ ,
and v = (0, 1, 0) to the only non-vertex of the remaining lattice points.
Since unimodular transformations preserve distances, we have that {du , dw } =
{du , dw }, for du := dist(t (u), t (H )) and dw := dist(t (w), t (H )). Moreover, we can
assume without loss of generality that du ≤ dw . Then:
⎧
⎪
⎨(1, 1) if K ∼
= B.7, C.1, F.13, F.15
(du , dw ) = (1, 2) ∼ F.14, F.17
if K =
⎪
⎩
(1, 3) if K ∼
= F.16
Let us now see that the denominator of the rational coordinates of v0 can only
be 2, 3 or 4:
v0 = (1 − λ)u + λw, for some λ ∈ [0, 1]
zu du
0 = (1 − λ)z u + λz w =⇒ λ = =
zu − zw du + dw
(−3, 3)
R3
(−1, 1) (1, 1)
(−2, 2)
−5 5
,
3 3
R4 R5
−3 3
, −7 7
2 2 ,
R6 5 5 R7
−4 4
,
3 3 R8 −9 9
, R9
7 7
Fig. 3.15 The regions R3 to R9 , with the points of the lattices L 2 , L 3 and L 4 contained in them.
Large squares are points of L 2 \ Z2 , medium squares are the points of L 3 \ Z2 , and small squares
the points of L 4 \ L 2 . Black dots and crosses represent points of Z2 in ∂ R and R2 \ R , respectively
Remember also that v0 must lie in the open triangle R . To prove the statement of
the theorem it remains to see that the intersection of R with any of the lattices L 2 , L 3
2
or L 4 , for L i := 1i Z × {0} is empty for ≥ 9. This is true since R9 does not contain
any point of those lattices, and since R ⊆ R9 for ≥ 9. To help the reader visualize
this we have drawn in Fig. 3.15 all the regions R3 to R9 with the possible positions
for the point v0 .
70 M. Blanco
Proof The first part of the statement follows from Proposition 11 and Theorem 23,
considering that the longest fronts in F1 (k), F2 (k), F3 (k) and F4 (k , k) have length k.
The second part follows from Corollary 14.
Remark 25 In case 5, the distance of any boundary lattice point of P to I P will also
be bounded since there are only finitely many lattice 3-polytopes with those particular
polygons as inner polytopes. However, this bound can only be found globally, and
not locally, since the distance from a single lattice point to I P is a priori unbounded
(see Corollary 14).
Now there is only left the case where I P is 3-dimensional. In particular, these are
quantum jumps of the type considered in Sect. 3.2.3, and we can apply the results
of Sect. 3.2.1 as explained in Remark 17. However, notice that we cannot use the
results of Sect. 3.3.2, since they heavily rely on I P being 2-dimensional.
We do the same as we did in Sect. 3.2.3: we check our database of lattice 3-
polytopes of size ≤ 11, width > 1 and 3-dimensional inner polytope, of which
there are 15,763 polytopes [4]. In this case, since a polytope with interior lattice
points cannot have width one, we are not losing cases by having only polytopes
of width > 1, but we only have polytopes with at most 11 lattice points in total.
Since I P is 3-dimensional, it has at least size 4, and since vert(P) ⊂ P \ I P , then I P
has at most 7 lattice points. That is, our database contains the information on I P of
size k ∈ {4, 5, 6, 7}, with P of size n ∈ {k + 4, . . . , 11} and with n − k ≤ 7 lattice
points in the boundary. In particular, both P and I P are very clean (few points in the
boundary) among the polytopes being checked.
3 Lattice Distances in 3-Dimensional Quantum Jumps 71
n 8 9 10 11
max Dv (I P ) 3 4 5 6
max dv (I P ) 3 3 3 4
and the average values of the Dv (I P ) and dv (I P ) are, respectively, 1.12 and 1.02.
Remark 26 Following the reasonings of Remark 19, in this case we would want
to have a list of polytopes Q and the maximum distance we can have a point x
so that (Q, x) is a quantum jump and there exists a polytope P such that Q = I P
and x ∈ ∂ P. From our database we are only considering 15,763 polytopes and, for
each of those polytopes P, we have exactly one polytope I P .
Putting together all the equivalent inner polytopes, we find out that there are
only 39 equivalence classes of inner polytopes. Moreover, around 9,000 polytopes in
the database (more than half) have the unimodular tetrahedron as its inner polytope.
The maximum and minimum distances for I P of size k are as follows:
k 4 5 6 7
max Dv (I P ) 4 4 5 6
max dv (I P ) 4 3 2 2
Altogether, it seems that we could find manageable bounds for the distance in
quantum jumps (I P , v), although further work is required.
Acknowledgements I would like to thank the referee for the useful comments and suggestions.
72 M. Blanco
References
1. Balletti, G., Kasprzyk, A.M.: Three-dimensional lattice polytopes with two interior lattice points
(2016). arXiv:1612.08918 [math.CO]
2. Blanco, M., Santos, F.: Lattice 3-polytopes with few lattice points. SIAM J. Discrete Math.
30(2), 669–686 (2016)
3. Blanco, M., Santos, F.: Lattice 3-polytopes with six lattice points. SIAM J. Discrete Math. 30(2),
687–717 (2016)
4. Blanco, M., Santos, F.: Enumeration of lattice 3-polytopes by their number of lattice points.
Discrete Comput. Geom. 60(3), 756–800 (2018)
5. Bruns, W., Gubeladze, J., Michałek, M.: Quantum jumps of normal polytopes. Discrete Comput.
Geom. 56(1), 181–215 (2016)
6. De Loera, J.A., Rambau, J., Santos, F.: Triangulations. Algorithms and Computation in Mathe-
matics, vol. 25. Springer, Berlin (2010). Structures for algorithms and applications
7. Kasprzyk, A.M.: Canonical toric Fano threefolds. Canad. J. Math. 62(6), 1293–1309 (2010)
8. Reeve, J.E.: On the volume of lattice polyhedra. Proc. Lond. Math. Soc. 3(7), 378–395 (1957)
9. White, G.K.: Lattice tetrahedra. Can. J. Math. 16, 389–396 (1964)
Chapter 4
Flag Matroids: Algebra and Geometry
A. Cameron
Max Planck Institute for Mathematics in the Sciences, Inselstr. 22, 04103 Leipzig, Germany
e-mail: [email protected]
Department of Mathematics and Computer Science, Eindhoven University of Technology, PO
Box 513, 5600 Eindhoven, The Netherlands
R. Dinu · M. Michałek (B)
University of Konstanz, Fachbereich Mathematik und Statistik, Fach D 197, 78457 Konstanz,
Germany
e-mail: [email protected]
R. Dinu
e-mail: [email protected]
T. Seynnaeve
Mathematisches Institut, Universität Bern, Alpeneggstrasse 22, 3012 Bern, Switzerland
e-mail: [email protected]
4.1 Introduction
The aim of this article is to present beautiful interactions among matroids and alge-
braic varieties. Apart from discussing classical results, we focus on a special class
of polymatroids known as flag matroids. The ultimate result is a definition of a Tutte
polynomial for flag matroids. Our construction is geometric in nature and follows
the ideas of Fink and Speyer for ordinary matroids [16]. The audience we are aim-
ing at is the union of combinatorists, algebraists and algebraic geometers, not the
intersection.
Matroids are nowadays central objects in combinatorics. Just as groups abstract
the notion of symmetry, matroids abstract the notion of independence. The interplay
of matroids and geometry is in fact already a classical subject [21]. Just one of such
interactions (central for our article) is the following set of associations:
We finish the article with a few open questions. As the construction of the Tutte
polynomial we propose is quite involved it would be very nice to know more direct,
combinatorial properties and definitions.
Note 1 Ewill always denote a finite set of cardinality n. P(E) is the set of all subsets
of E, and Ek is the set of all subsets of E of cardinality k. We use [n] as a shorthand
notation for the set {1, 2, . . . , n}. We will denote the difference of two sets X and Y
by X − Y . This does not imply that Y ⊆ X . If Y is a singleton {e}, we write X − e
instead of X − {e}.
B1. B = ∅; and
B2. (basis exchange) if B1 , B2 ∈ B and e ∈ B1 − B2 , there exists f ∈ B2 − B1 such
that (B1 − e) ∪ f ∈ B.
A reader new to matroid theory should not be surprised by the borrowed termi-
nology from linear algebra: matroids were presented as a generalisation of linear
independence in vector spaces in the paper by Whitney [57] initiating matroid the-
ory. Matroids also have a lot in common with graphs, thus explaining even more
of the terminology used. For instance, very important matroid operations are that
of minors. These are analogous to the graph operations of the same names. As there,
deletion is very simple, while contraction requires a bit more work.
We will now give two examples of classes of matroids which show exactly the
relationship matroids have with linear algebra and graph theory. The first one plays
a central role in our article.
Remark 7 Our definition differs slightly from the one found in literature: typically
one identifies E with φ(E). Our definition does not require φ to be injective; we can
take the same vector several times. We also note that the matroid represented by φ :
E → V only depends on the underlying map φ : E → P(V ), assuming φ(E) ⊂
V \{0}.
8
7 9
1 2 3
{1, 2, 3}, {4, 5, 6}, {1, 5, 7}, {1, 6, 8}, {2, 4, 7}, {2, 6, 9}, {3, 4, 8}, {3, 5, 9}.
Further matroid definitions will be given later, but we have covered enough to give
the major object of our interest in this paper, namely the Tutte polynomial. This
is the most famous matroid (and graph) invariant, and, like matroids themselves,
has multiple definitions. These will be mentioned where relevant. Here, we give
the corank-nullity formula, two terms which will be defined below.
Definition 10 Let M = (E, r ) be a matroid with ground set E and rank function r :
P(E) → Z≥0 . The Tutte polynomial of M is
TM (x, y) = (x − 1)r (M)−r (S) (y − 1)|S|−r (S) .
S⊆E
The term r (M) − r (S) is called the corank, while the term |S| − r (S) is called
the nullity. Readers familiar with matroid theory should be careful not to confuse
a mention of corank with dual rank, given the usual naming convention of dual
objects. By identifying the rank function of a matroid with the connectivity function
78 A. Cameron et al.
of a graph in an appropriate way, one can pass between this formula and the original
formulation of the Tutte polynomial which was given for graphs.
Example 11 For the (matroid of the) complete graph K 4 , there are four subsets with
three elements of rank 2 and all the other subsets with three elements have rank 3.
In this case, the Tutte polynomial is
TM(K 4 ) (x, y) = x 3 + 3x 2 + 2x + 4x y + 2y + 3y 2 + y 3 .
Readers interested in seeing what the Tutte polynomial looks like for a range of
different classes of matroids should consult [36].
The prevalence of the Tutte polynomial in the literature is due to the wide range
of applications it has. The simplest of these occurs when we evaluate the polynomial
at certain points, these being called Tutte invariants. For instance, T (1, 1) gives the
number of bases in the matroid (or the number of spanning trees in a graph). In this
way we can also count the number of independent sets in a matroid or graph, and
the number of acyclic orientations of a graph, as well as some other such quantities.
Beyond numerics, the Tutte invariants also include other well-known polynomials,
appearing in graph theory (the chromatic polynomial, concerned with graph colour-
ings; see also Theorem 93) and network theory (the flow and reliability polynomials).
Extending to further disciplines, one can find multivariate versions of the Tutte poly-
nomial which specialise to the Potts model [55] from statistical physics and the Jones
polynomial [52] from knot theory. In this paper, we will be looking at the classical
Tutte polynomial from an algebraic point of view.
We noted that there are multiple definitions of the Tutte polynomial. One is both so
useful and attractive that we would be remiss to not include it. It states that, instead of
calculating the full sum above, we can instead simply form a recurrence over minors
of our matroid, which can lead to faster calculations. Note that a coloop is an element
of E which is in every basis of M, while a loop is an element which is in no basis.
The Tutte polynomial is in fact universal for such formulae: any formula for
matroids (or graphs) involving just deletions and contractions will be an evaluation
of the Tutte polynomial. There are numerous proofs of this in the literature, and also
extensions to related classes of objects. One such reference is [13, Sect. 4].
4 Flag Matroids: Algebra and Geometry 79
We will now give two more axiom systems for matroids. The first one, via base
polytopes, will play a fundamental role in this paper.
We first define what the base polytope of a matroid is: let B be the set of bases of a
matroid M = (E, r ). We work in the vector space R E = {(ri | i ∈ E)}, where ri ∈ R.
For a set U ⊆ E, eU ∈ R E is the indicator vector of U , that is, eU is the sum of the
unit vectors ei , for all i ∈ U . Note that e{i} = ei .
Note that this is always a lattice polytope. Its dimension is equal to |E| minus the
number of connected components of the matroid [15, Proposition 2.4]. We also note
that the vertices of P(M) correspond to the bases of M. In particular: given P(M) ⊂
R E , we can recover M.
The following theorem gives a characterisation of which lattice polytopes appear
as the base polytope of a matroid. It can be used as an axiom system to define
matroids:
Theorem 14 ([12], see also [21, Theorem 4.1]) A polytope P ⊂ R E is the base
polytope of matroid on E if and only if the following two conditions hold:
P1. every vertex of P is a 0, 1-vector; and
P2. every edge of P is parallel to ei − e j for some i, j ∈ E.
More generally the description of faces of matroid base polytopes is provided in [15,
30]. The base polytope is a face of the independent set polytope of M, which is the
convex hull of indicator vectors of the independent sets of M.
We move on to another axiom system: via Gale orderings. This definition is orginally
due to Gale [20]; our formulation is based on lecture notes by Reiner [45].
Definition 15 Let ω be a linear ordering
on E, which we will denote by ≤. Then
the dominance ordering ≤ω on Ek , also called Gale ordering, is defined as follows.
Let A, B ∈ Ek , where
A = {i 1 , . . . , i k }, i 1 < · · · < i k
and
B = { j1 , . . . , jk }, j1 < · · · < jk .
80 A. Cameron et al.
Then we set
A ≤ω B if and only if i 1 ≤ j1 , . . . , i k ≤ jk .
Theorem 16 ([20]) Let B ⊆ Ek . We have that B is the set of bases of a matroid if
and only if for every linear ordering ω on E, the collection B has a maximal element
under the Gale ordering ≤ω (i.e. there is a unique member A ∈ B such that B ≤ω A,
for all B ∈ B).
Then l is also a family of independent sets for a matroid, known as the union
of M1 , . . . , Mk . Further, the rank of any set A ⊂ E for the union matroid is given
by:
k
r (A) = min{|A\B| + ri (B)}.
B⊂A
i=1
The proof can be found e.g. in [43, 12.3.1]. The following corollary is essentially
due to Edmonds.
Corollary 18 Let M1 , . . . , Mk be matroids on a ground set E with rank functions
respectively r1 , . . . , rk . E can be partitioned into independent
k sets, one for each
matroid, if and only if for all subsets A ⊂ E we have |A| ≤ i=1 ri (A).
k
By assumption, for any B ⊂ E we have |E| − |B| + i=1 ri (B) ≥ |E|. Further,
equality holds for B = ∅. Hence, rU (E) = |E|. Thus, E is an independent set of U .
By definition it is a union of k independent sets, one in each of the Mi ’s.
Consider what happens if we drop one of the rank axioms, namely that which
states r (X ) ≤ |X |. What object do we get, and what relation does it have to matroids?
This object was originally studied by Edmonds [12] (although in a different guise,
see Definition 23), and dubbed a polymatroid. The class of polymatroids includes,
naturally, the class of matroids, and is greatly important in the field of combinatorial
optimisation.
Remark 20 As we assume that our rank function take only integral values, the object
we defined is sometimes referred to in the literature as a discrete polymatroid [27].
Analogously to the matroid case, we can give an axiom system for polymatroids in
terms of their bases.
Lemma 22 ([27, Theorem 2.3]) A nonempty finite set B ⊂ Z≥0
E
is the set of bases
of a polymatroid on E if and only if B satisfies
1. all u ∈ B have the same modulus (sum of entries); and
2. if u = (u 1 , . . . , u n ) and v = (v1 , . . . , vn ) belong to B with u i > vi then there
is j ∈ E with u j < v j such that u − ei + e j ∈ B.
Definition 23 Let M = (E, r ) be a polymatroid. Let I ⊆ Z≥0 E
be the set of indepen-
dent vectors, and B ⊆ Z≥0 be the set of bases. We have the independent set polytope,
E
As before, the base polytope is a face of the extended polymatroid. When the
polymatroid considered is a matroid, these definitions coincide exactly with those
from Sect. 4.2.3.
Theorem 14 generalises to the case of polymatroids, giving us another equivalent
definition of polymatroids in terms of their base polytopes:
Theorem 24 ([27, Theorem 3.4]) A polytope P ⊂ Rn is the base polytope of a
polymatroid on [n] if and only if the following two conditions hold:
1. every vertex of P has coordinates in Z≥0 ; and
2. every edge of P is parallel to ei − e j , for some i, j ∈ [n].
If M is a polymatroid, then the bases (resp. independent vectors) of M are precisely
the lattice points of P(M) (resp. E P(M)). The following proposition describes
which bases of M correspond to vertices of P(M).
Proposition 25 ([12], see also [27, Proposition 1.3]) Let M = ([n], r ) be a poly-
matroid and assign an ordering S to the ground set [n]. Let Si be the first i elements
according to this ordering. Every possible S corresponds (not necessarily uniquely)
to a vertex of P(M); x = x S , where x = (x1 , . . . xn ), and
xi = r (Si ) − r (Si−1 ).
where λi = d, μj = d and λi , μj ∈ Z≥0 .
By restricting the set E we may assume that all coordinates of p = ( p1 , . . . , pn )
are nonzero (i.e. pi ∈ {1, 2}), where we identify E with [n]. Dually, by contracting
the elements of E that belong to all bases corresponding to any ti and q j , we may
assume p = (1, . . . , 1).
We want to prove that the ground set E is covered by a basis of M1 and a basis
of M2 . Hence, by Corollary 18 it is sufficient to prove the following:
For any A ⊂ E we have |A| ≤ r M1 (A) + r M2 (A).
The Tutte polynomial for polymatroids is not nearly as well-studied as in the matroid
case. We gave two examples of where it was considered in certain classes of poly-
matroids. We will now go into detail about one suggestion how to construct Tutte
polynomial in full generality.
As mentioned, Cameron and Fink [8] form a polynomial having Tutte-like prop-
erties for polymatroids, which specialises to an evaluation of the Tutte polynomial
when applied only to matroids. This is a construction which takes a polytopal, lattice-
point-counting, approach as opposed to a straight combinatorial one. It is motivated
by an alternative definition of the Tutte polynomial to those we have discussed so
far.
A cocircuit is a minimal set among sets intersecting every basis. We will not be
using this notion again in the article.
We will denote the number of internally active elements with respect to B
with I (B) and the number of externally active elements by E(B). Then we have
the following result.
Theorem 28 ([53])
TM (x, y) = x I (B) y E(B) .
B∈B
u t
Q M (t, u) := #(P(M) + u + t∇) = ci j .
i, j
j i
Changing the basis of the vector space of rational polynomials gives the polynomial
Q M (x, y) = ci j (x − 1)i (y − 1) j
ij
4.4.2 Grassmannians
k
G(k, n) = G(k, V ) = {[v1 ∧ . . . ∧ vk ] ⊂ P( V )}.
Here, v1 , . . . , vk are the rows of the aforementioned matrix A, and thus a point
of G(k, n) is identified with the subspace v1 , . . . , vk . The embedding presented
above is known as the Plücker embedding and the Grassmannian is defined by
quadratic polynomials known as Plücker relations [34]. Explicitly, in coordinates,
the map associates to the matrix A the value of all k × k minors. We refer the readers
not familiar with algebraic geometry, and in particular Grassmannians, to a short
introduction in [39, Chap. 5].
The Plücker embedding may be identified with a very ample line bundle on G(k, n),
which we will denote by O(1). Other very ample line bundles on G(k, n) are the d-th
tensor powers O(d). They can be realised as a composition
of the Plücker embedding
with the d-th Veronese map P( k V ) → P(Symd k V ).
Remark 32 A reader not familiar at all with very ample line bundles may think
about them as maps into projective spaces. Let us present this with the example
of the projective space Pn (which also equals G(1, n + 1)). We have an identity
map Pn → Pn , which corresponds to O(1). The r -th Veronese map embeds Pn in a
larger projective space P( n )−1 by evaluating on a point all degree r monomials. The
n+r
P1 [x : y] → [x 2 : x y : y 2 ] ∈ P2 .
It will follow from Proposition 34 that the embedding of G(k, n) by O(d) spans
a projectivisation of an irreducible representation Vλ0 of G L n . The Young dia-
gram λ0 = (d, . . . , d) consists of k rows of length d.
Hence, flag varieties are in bijection with (nonempty) subsets of {1, . . . , n − 1},
while Grassmannians correspond to singletons.
From now on we will abbreviate the tuple (k1 , . . . , ks ) to k, and the flag vari-
ety Fl(k1 , . . . , ks ; n) to Fl(k, n). A point in Fl(k, n) can be represented by a full-
rank n × n-matrix A: the row span of the first di rows is Vi . (Although note that
only the first ks rows of the matrix are relevant.) As with Grassmannians, different
matrices can represent the same point in Fl(k, n). More precisely, if we partition the
rows of A into blocks of size k1 , k2 − k1 , . . . , n − ks , then we are allowed to do row
operations on A, with the restriction that to a certain row we can only add a multiple
of a row in the same block or a block above. Another way to think about this is the
following: let Pk ⊂ G L n (C) be the parabolic subgroup of all invertible matrices A
with Ai j = 0 if i ≤ kr < j, for some r . Then two n × n matrices represent the same
flag if and only if they are the same up to left multiplication with an element of Pk .
Hence Fl(k, n) can also be described as the quotient Pk \G L n (C) (a homogeneous
variety).
Just as for Grassmannians we may study different embeddings of flag varieties.
The natural one is given by the containment
k1
ks
k1
ks
G(k1 , V ) × · · · × G(ks , V ) ⊂ P( V ) × · · · × P( V ) ⊂ P( V ⊗ ··· ⊗ V ),
where the last map is the Segre embedding. The representation k1 V ⊗ · · · ⊗ ks V
in general is reducible—a precise decomposition is known by Pieri’s rule (or more
generally by the Littlewood-Richardson rule), see for example [19, Proposition
15.25]. As we will prove below, the flag variety spans an irreducible representation
with the corresponding Young diagram with s columns of lengths ks , ks−1 , . . . , k1
respectively.
Other embeddings can be obtained as follows. We replace the Segre map by the
Segre–Veronese, i.e. we first re-embed a Grassmannian with a Veronese map. Thus,
a flag variety with an embedding can be specified by a function:
f : {1, . . . , n − 1} → Z≥0 .
We now pass to the embedding. As f (1) = 2 we have to consider the second Veronese
map P(V ) → P(S 2 (V )) given by [v] → [v · v]. We obtain
2
Fl(1, 2) = {[v1 · v1 ] × [v1 ∧ v2 ] ∈ P(S 2 V ) × G(2, V )} ⊂ P(S 2 V ⊗ V ).
2
S2 V ⊗ V = S3,1 V ⊕ S2,1,1 V.
Hence, S3,1 (V ) corresponds to the Young diagram with the first row of length three
and the second row of length one. We note that this diagram indeed has 2 columns
of length 1, 1 column of length 2, and 0 columns of length 3.
The flag variety is always contained in the lexicographically-first (highest weight)
irreducible component—cf. Proposition 34 below; in our example this is S3,1 (V ). In
particular, we may realise the representation S3,1 (V ) as a linear span of the affine
cone over the flag variety:
2
2) = {(v1 · v1 ) ⊗ (v1 ∧ v2 ) : v1 , v2 ∈ V } ⊂ S 2 V ⊗
Fl(1, V.
For readers not familiar with the construction of Sλ V , this can be taken as a def-
inition. For a proof that this definition is equivalent to the usual construction, see
Proposition 34 below.
n-1
V1 ⊂ V2 ⊂ · · · ⊂ Vn−1 ⊂ Cn .
The complete flag variety maps to any other flag variety, simply by forgetting the
appropriate vector spaces. We note that all our constructions are explicit and only
use exterior (for Grassmannians), symmetric (for Veronese) and usual (for Segre)
tensor products, as in Example 33.
We are now ready to prove a special case of the Borel-Weil-Bott Theorem relating
representations and embeddings of flag varieties.
Proposition 34 (Borel-Weil) Any flag variety Fl(k1 , . . . , ks ; n) with an embedding
given by a function f spans the irreducible G L(V )-representation Sλ V , where the
Young diagram λ has f ( j) columns of length j.
Proof Let A be the matrix whose rows span the space corresponding to p. The
parameterisation of T p is given by:
k
φ : T → P( Cn ).
The coordinates of the ambient space are indexed by k-element subsets of the n
columnsof the matrix A. The Plücker coordinate indexed by I of φ(t1 , . . . , tn )
equals i∈I ti times the k × k minor of A determined by I , which we will denote
by det(A I ). In other words, the map φ in Plücker coordinates is given as follows:
φ(t1 , . . . , tn ) = (det(A I ) · ti ) I ∈([n]) .
k
i∈I
The I -th coordinate is nonzero if and only if I is a basis of M p . Hence, the ambient
space of T p has coordinates indexed by basis elements of M p . After restricting
to this ambient space and composing with the isomorphism inverting the nonzero
minors det(A I ), our map can be written as
φ(t1 , . . . , tn ) = ti .
i∈I I ∈P(M p )
We start by defining flag matroids in the way they are usually defined in the literature:
using Gale orderings.
Definition 41 Let 0 < k1 ≤ . . . ≤ ks < n be natural numbers. Let k = (k1 , . . . , ks ).
A flag F of rank k on E is an increasing sequence
F1 ⊆ F2 ⊂ · · · ⊆ Fs
of subsets of E such that |F i | = ki for all i. The set of all such flags will be denoted
by F Ek .
Let ω be a linear ordering on E. We can extend the Gale ordering ≤ω to flags:
Definition 44 Let M and N be matroids on the same ground set E. We say that N
is a quotient of M if one of the following equivalent statements holds:
1. every circuit of M is a union of circuits of N ;
2. if X ⊆ Y ⊆ E, then r M (Y ) − r M (X ) ≥ r N (Y ) − r N (X );
3. there exists a matroid R and a subset X of E(R) such that M = R\X and N =
R/ X ;
/ B, there is a basis B of N with B ⊆ B and such
4. for all bases B of M and all x ∈
that {y : (B − y) ∪ x ∈ B(N )} ⊆ {y : (B − y) ∪ x ∈ B(M)}.
For the equivalence of 1, 2 and 3, we refer to [43, Proposition 7.3.6]. Part 4 is left to
the reader.
Here are some basic properties of matroid quotients:
Proposition 45 Let N be a quotient of M.
1. Every basis of N is contained in a basis of M, and every basis of M contains a
basis of N .
2. rk(N ) ≤ rk(M) and in case of equality N = M.
The next result will be essential for defining representable flag matroids. It also
explains where the term “matroid quotient” comes from—below we think of W as a
vector space quotient of V .
4 Flag Matroids: Algebra and Geometry 95
Proposition 46 ([6, Proposition 7.4.8 (2)]) Let V and W be vector spaces and ψ :
E → V be a map. Furthermore, let f : V → W be a linear map. Consider the
matroid M represented by ψ, and the matroid N represented by f ◦ ψ. Then N is a
matroid quotient of M.
and let N be the rank-2 matroid on [8] whose bases are all 2-element subsets except
for {2, 6} and {3, 5}. It is easy to see that N is a representable matroid: just pick
six pairwise independent vectors in the plane, and map 2 and 6, as well as 3 and 5,
to the same vector. Now N is a matroid quotient of M, since the matroid R from
Example 8 satisfies M = R\9 and N = R/9. However, it is not possible to find
representations V (resp. W ) of M (resp. N ) such that there is a map f : V → W as
in Proposition 46. Roughly speaking, the problem is that the “big” matroid R is not
representable. For a more precise argument, see [3, Sect. 1.7.5].
Remark 51 Example 48 shows that it can happen that all constituents of a flag
matroid are representable matroids, but still the flag matroid is not representable
(because the matroid representations are “not compatible”).
Definition 53 The base polytope of a flag matroid F on [n] is the convex hull of
the set {e F | F ∈ F } ⊂ Rn .
Example 54 Let F be the rank (1, 2) flag matroid on [3] whose bases are 1 ⊆
12, 1 ⊆ 13, 2 ⊆ 12 and 3 ⊆ 13. Then its base polytope is the convex hull of the
points (2, 1, 0), (2, 0, 1), (1, 2, 0), (1, 0, 2). Its constituents are the uniform rank 1
matroid on [3], and the rank 2 matroid with bases 12 and 13. The base polytope of this
flag matroid is depicted in Fig. 4.4. F is a representable flag matroid: a representing
flag is for example e1 + e2 + e3 ⊂ e1 , e2 + e3 ⊂ C3 .
Theorem 56 ([3, Corollary 1.13.5]) The polytope of a flag matroid is the Minkowski
sum of the matroid polytopes of its constituent matroids.
4 Flag Matroids: Algebra and Geometry 97
(2,0,1)
(2,1,0) (1,2,0)
Thus, each flag matroid defines a polytope that is a base polytope of a polymatroid.
Remark 57 It follows from Theorem 26 and the previous theorem that the lat-
tice points of a flag matroid polytope correspond to tuples (not necessarily
flags) (B1 , . . . , Bs ), where Bi is a basis of Mi . For example point (1, 1, 1) in Fig. 4.4
corresponds to a basis of a polymatroid, but not to a flag, i.e. not to a basis of the flag
matroid.
Consider the flag variety Fl(k, n), as described in Sect. 4.4. The action of the
torus T = (C∗ )n on Cn induces an action of T on Fl(k, n). A point p ∈ Fl(k, n)
gives rise to a representable flag matroid M on [n], as in Definition 50. All points
in the orbit T p give rise to the same flag matroid. This last statement follows easily
from the analogous fact for matroids and the fact that a flag matroid is determined
by its constituent matroids. The analogue of Theorem 37 holds:
Theorem 58 The lattice polytope representing the toric variety T p is equal to the
flag matroid polytope of M.
Proof The proof is a straightforward generalisation of the proof of Theorem 37, with
the parameterisation of T p given by:
k1
ks
φ : T → P( Cn ) × · · · × P( Cn ).
that e∈E φ(e) = V . The representable polymatroid M(φ) is defined by the rank
function: for A ⊆ E, define r (A) as the dimension of φ(A) := a∈A φ(a).
Proof For this proof we will use the more intrinsic definition of representable
flag matroids and polymatroids using quotients. Let F be the flag matroid repre-
sented
by the flag of quotients Cn → V1 → · · · → Vr . Then the quotient Ckn →
V1 · · · Vr represents a polymatroid M on [n]. We need to argue that F and M
have the same base polytope.
Bases of F correspond to flags [n] F 1 ⊇ . . . ⊇ F r such that F i gives a basis
of Vi . On the other hand, choosing a basis of M corresponds to choosing for every i
a subset F i [n] such that F i gives a basis of Vi . From this it follows that every
vertex of P(F ) is a lattice point of P(M), and (using Remark 57) that every lattice
point of P(M) is a lattice point of P(F ). So P(M) = P(F ) as desired.
Remark 64 Thinking again about flag varieties in terms of subspaces, we have that
if F is represented by a flag V1 ⊆ . . . ⊆ Vr Cn , then M(F ) is represented by a
subspace V1 ⊕ · · · ⊕ Vr Cr n . Geometrically, this construction corresponds to an
algebraic map
Fl(k1 , . . . , kr ; n) → G( ki , r n).
i
Remark 66 In all the cases we study the ring K 0 (X ) is isomorphic to the cohomol-
ogy ring and to the Chow ring (after tensoring with Q). Note however that the map
from K 0 (X ) to the Chow ring is nontrivial and given by the Chern character.
Example 67 Consider the projective space Pn . The (rational) Chow ring is A(Pn ) =
Q[H ]/(H n+1 ). Here one should think about H as a hyperplane in Pn and H k as a codi-
mension k projective subspace. The most important n line bundle is O(1). The Chern
character ch : K 0 (Pn ) → A(Pn ) sends [O(1)] to i=0 H i /i!. Note that K 0 (Pn ) can
be written as Z[α]/(α n+1 ), where α = 1 − [O(−1)] is the class of the structure sheaf
of a hyperplane. As a special case, the K -theory of a point is Z.
character (also called Hilbert series) Hilb(W ) := c∈Char(T ) (dim Wc )c. We point
out that even for infinite-dimensional T -modules, Hilb(W ) makes sense as a formal
power series, as long as Wc is finite-dimensional for all c.
We finish this section by describing the relation between ordinary and T -
equivariant K -theory:
Theorem 68 ([37, Theorem 4.3]) Let X be a smooth projective variety with an
action of a torus T . Let S ⊆ T be a subtorus. Then the natural map
K T0 (X ) ⊗Z[Char(T )] Z[Char(S)] → K S0 (X )
K T0 (X ) ⊗Z[Char(T )] Z → K 0 (X )
is an isomorphism.
We note that the map Z[Char(T )] → Z above is given in coordinates by sending
each generator ti of T to 1.
Let X be a smooth projective variety over C, and T a torus acting on it. If X has only
finitely many torus-fixed points, we can use the method of equivariant localisation
to give an explicit combinatorial description of classes in K T0 (X ). Our exposition
here is largely based on the one in [16]. The following theorem is central to our
discussion.
Theorem 69 ([42, Theorem 3.2], [16, Theorem 2.5] and references therein) If X is
a smooth projective variety with a torus action, then the restriction map K T0 (X ) →
K T0 (X T ) is an injection.
From now on we will always assume that X has only finitely many torus-fixed points.
In this case K T0 (X T ) is simply the ring of functions from X T to Z[Char(T )]. In other
words, we can describe a class in K T0 (X ) just by giving a finite collection of Laurent
polynomials in Z[Char(T )].
Remark 70 In the literature, a variety X for which K T0 (X ) is a free Z[Char(T )]-
module, and has a Z[Char(T )]-basis that restricts to a Z-basis of K 0 (X ), is
called equivariantly formal. This notion was first introduced in [24]. In [2, Sect. 2.4],
it is noted that smooth projective varieties with finitely many T -fixed points are
equivariantly formal.
We now explicitly describe the class of a T -equivariant coherent sheaf on X . We will
do this under the following additional assumption (which is not essential but makes
notation easier and will hold for all varieties of interest):
102 A. Cameron et al.
Ai = C[x0 , . . . , xˆi , . . . , xn ]
Example 76 (Example 74 continued) Note that Pn has only finitely many one-
dimensional torus orbits: for every pair pi , p j of T -fixed points, there is a unique T -
orbit whose closure contains pi and p j . Furthermore, T acts on this orbit with
character t −1 d d −1
j ti . We see that ti ≡ t j mod 1 − t j ti , so that the class [O(d)] indeed
T
for x ∈ X T .
Describing the pushforward of [F]T ∈ K T0 (X ) is a bit more complicated. Sup-
pose that X and Y are contracting. For every point x ∈ X T (resp. y ∈ Y T ), we
pick as before an open neighbourhood Ux (resp. Vy ) on which T acts by charac-
ters χ1 (x), . . . , χr (x) (resp. η1 (y), . . . , ηs (y)). Then the pushforward of [F]T is
determined by the formula
(t0 − t1 , 0) ⊗ 1 = (1, 0) ⊗ (1 − 1) = 0
but this is wrong! Indeed, (1, 0) does not satisfy the condition from Theorem 75,
hence is not in K T0 (X ). In fact, one can check that (t0 − t1 , 0) is the equivariant class
of the torus-fixed point [1 : 0] ∈ P2 .
For more details about the topic of this subsection we refer to [9, Sect. 1.2] and [48,
convex polyhedral rational cone is a subset of Rn of the
Sect. 4.5]. Recall that a
form C = cone(S) := { u∈S λu u | λu ∈ R≥0 }, where S ⊂ Zn ⊂ Rn is finite. A cone
104 A. Cameron et al.
is called pointed if it does not contain a line. If C is a pointed rational cone, then
every one-dimensional face ρ contains a unique lattice point uρ that is closest to the
origin. It is not hard to see that M G(C) := {uρ | ρa one-dimensional face ofC} is
a minimal generating set of C. If the minimal generators are linearly independent
over R, we call C simplicial. If they are part of a Z-basis of Zn , we call C regular.
For a pointed cone C in Rn , we define its Hilbert series Hilb(C) by:
Hilb(C) := ta .
a∈C∩Zn
Thisis always a rational function [48, Theorem 4.5.11] whose denominator is equal
to u∈M G(C) (1 −tu ). If C is a regular cone, then its Hilbert series is easy to com-
pute: Hilb(C) = u∈M G(C) 1/(1 − tu ). If C is a simplicial cone, we can compute
Hilbert series as follows. First compute the finite set DC := {b ∈ C ∩ Z : b =
n
its
b∈M G(C) λu u | 0 ≤ λu < 1}. Then
⎛ ⎞
1
Hilb(C) = ⎝ tb ⎠ .
b∈DC u∈M G(C)
1 − tu
For a general rational polyhedral cone, we can compute its Hilbert series by
triangulating it.
In this subsection we compute the class in equivariant K -theory of a torus orbit closure
in a Grassmannian. We then note that this class only depends on the underlying
matroid, and give a combinatorial algorithm to get the class in K -theory directly
from the matroid. This algorithm can then be used as a definition to associate a class
in K -theory to an arbitrary (not necessarily representable) matroid. This was first
done by Fink and Speyer [16].
Let us first fix the following sign conventions. The torus T = (C∗ )n acts on Cn as
follows: t · (x1 , . . . , xn ) = (t1−1 x1 , . . . , tn−1 xn ). The action of T on G(k, n) is induced
from this action. Explicitly, if p ∈ G(k, n) has Plücker coordinates [PI ] I ∈([n]) , then t ·
k
p has Plücker coordinates [( i∈I ti−1 )PI ] I ∈([n]) .
k
We begin by describing the T -equivariant K -theory of the Grassmannian G(k, n)
using equivariant localisation.
The torus-fixed points of G(k, n) are easy to describe: for every size k sub-
set I ⊂ [n], we define the k-plane VI = span({ei | i ∈ I }) ⊂ Cn , and denote the
corresponding point in G(k, n) by p I. In Plücker coordinates, p I is given by PJ = 0
if J = I . It is easy to see that the nk points p I are precisely the torus-fixed points
of G(k, n).
4 Flag Matroids: Algebra and Geometry 105
We can also describe the one-dimensional torus orbits: there is a (unique) one-
dimensional torus orbit between p I and p J if and only if |I ∩ J | = k − 1. In this
case, we write I − J = {i}, J − I = { j}. If we identify the one-dimensional orbit
from p I to p J with A1 \0 in such a way that the origin corresponds to the torus-fixed
point p I (and so p J corresponds to the point at infinity), then T acts on the orbit with
character t −1
j ti .
Let us now check that the action of T is contracting. We fix a torus-fixed point p I ,
and consider the open neighbourhood U I given by PI = 1. Then U I ∼ = Ak(n−k) .
PI −i∪ j
For p ∈ U I , we will denote its coordinates with (u i, j )i∈I, j ∈I / , where u i, j = PI .
Then t · p has coordinates (t −1j ti u i, j )i∈I, j ∈I
/ . Thus, T acts on this space with char-
−1 a1
acters t j ti , where i ∈ I, j ∈
/ I . Identifying t1 · · · tn with (a1 , . . . , an ), all these
an
Example 78 We compute the class of O(1). The sheaf O(1) on G(k, n) was already
(nk)−1 via the Plücker embed-
k ∨ of O(1) on P
mentioned in Sect. 4.4.2: it is the pullback
ding. We can also describe O(1) as S , where S is the tautological vector bundle
on G(k, n) whose fiber over a point is the corresponding k-plane.
We can apply Theorem 68 to the result from Example 74 to replace the torus
action there with a different torus action, induced from the action on the Plücker
coordinates. By applying pullback formula (4.2), we find that the class [O(1)]T in
equivariant K -theory is the map
where we wrote I = {i 1 , . . . , i k }.
Since cone I (M) is the positive real span of all vectors e J − e I , where J ∈ B(M),
we find that Hilb(R M,I ) = Hilb(cone I (M)). So (4.4) can also be written as
[T p]T ( p I ) = Hilb(cone I (M)) (1 − ti−1 t j ). (4.5)
i∈I j ∈I
/
We note that (4.5) only depends on the matroid M and not on the chosen point p
or even the torus orbit T p. Moreover, the formulas make sense even for non-
representable matroids. Thus we can use them as a definition for the class in K -theory
for a matroid:
Definition 80 ([16]) For any rank k matroid M on [n], we define y(M) :
Gr (k, n)T → Z[Char(T )] by
y(M)( p I ) = Hilb(cone I (M)) (1 − ti−1 t j ).
i∈I j ∈I
/
Theorem 81 ([16, Proposition 3.3]) The class y(M) ∈ K T0 (Gr (k, n)T ) satisfies the
condition of Theorem 75, and hence defines a class in K T0 (Gr (k, n)).
In this section, we generalize the results from the previous section replacing
“matroids” by “flag matroids” and “Grassmannians” by “flag varieties”.
We first describe the equivariant K -theory of a flag variety Fl(k, n). The torus-
fixed points are given as follows: for every (set-theoretic) flag F = (F1 ⊆ . . . ⊆ Fs )
of rank k on [n], we define a (vector space) flag V1 ⊆ . . . ⊆ Vs by Vi = span({e j | j ∈
Fi }). We will denote the corresponding point in Fl(k, n) by p F . The Plücker coor-
dinates of p F are given by PS = 1 if S is a constituent of F and PS = 0 otherwise.
Here, the Plücker coordinates
of a point in Fl(k, n) are the ones induced from the
embedding Fl(k, n) → G(ki , n).
We can also describe the one-dimensional torus orbits: let p F be a torus-fixed
point. We define S(F) to be the set of all pairs (i, j) ∈ [n] × [n] for which there
exists an
such that i ∈ F
and j ∈ / F
. For every (i, j) ∈ S(F), we define a new
flag F = Fi→ j by switching the roles of i and j. More precisely: if i ∈ Fr but j ∈
/ Fr ,
then Fr := (Fr − i) ∪ j; in any other case Fr := Fr . Then there is a unique one-
dimensional torus orbit between p F and p F , and all one-dimensional torus orbits
arise in this way. The torus T acts on this orbit with character t −1
j ti .
Definition 83 We define cone F (F ) to be the cone conee F (P(F )), as in Definition 79.
As before, we find that Hilb(RF ,F ) = Hilb(cone F (F )). Hence, (4.6) can also be
written as
[T p]T ( p F ) = Hilb(cone F (F )) (1 − ti−1 t j ). (4.7)
(i, j)∈S(F)
As before, (4.7) only depends on the flag matroid F and not on the chosen point p
or even the torus orbit T p. Moreover the formulas make sense even for non-
representable flag matroids. Thus we can use them as a definition for the class
in K -theory for a flag matroid:
Definition 84 For any rank k flag matroid F on [n], we define y(F ) : Fl(k, n)T →
Z[Char(T )] by
y(F )( p F ) = Hilb(cone F (F )) (1 − ti−1 t j ).
(i, j)∈S(F)
that y(F )( p F ) = 1 − t1−1 t3 . We can do the same for the other torus-fixed points. The
result is summarised in Fig. 4.5.
where Y (M) is the class associated to the matroid M in the non-equivariant K-theory
of the Grassmannian.
In other words, the Tutte polynomial of a matroid can be viewed as a Fourier-Mukai
transform of its associated class in K -theory.
We can now generalize this construction to get a definition of the Tutte polynomial
of a flag matroid.
Fl(1, k, n − 1; n)
πk
Pn−1 × Pn−1
Let F be a flag matroid on [n] of rank k, and let Y (F ) ∈ K 0 (Fl(k; n)) be its class in
non-equivariant K -theory, as in Definition 84. Then the K -theoretic Tutte polynomial
of F is defined to be
We computed the Tutte polynomial for some small examples using Sage [51],
Macaulay2 [25], and Normaliz [5]. Our code is available at [47]. The program first
computes the equivariant class (π1(n−1) )∗ πd∗ (y(M) · [O(1)]) ∈ K T0 (Pn−1 × Pn−1 )
using equivariant localisation, and then computes the underlying non-equivariant
class.
Remark 89 After we finished the article Christopher Eur implemented the algorithm
in Macaulay2 [14].
Y ( p) = (1 − η−1 )X ( p).
Consider for example p = (e1 , e1 , e2 ). Then p ∈ Fl(1, 2; 3) has an open
neighbourhood where T acts by characters t2 t3−1 , t1 t3−1 , t1 t2−1 , while p ∈ P2 × P2
has an open neighbourhood where T acts by characters t2 t3−1 , t1 t3−1 , t1 t3−1 , t1 t2−1 . We
compute that
TF (x, y) = x 2 y 2 + x 2 y + x y 2 + x 2 + x y.
x 3 y 3 + 2x 3 y 2 + 2x 2 y 3 + 3x 3 y + 8x 2 y 2 + 3x y 3 + 4x 3 +
8x 2 y + 8x y 2 + 4y 3 + 2x 2 + 4x y + 2y 2 .
After the preprint of the article appeared Christopher Eur found an example of
a K -theoretic Tutte polynomial TU(1,3):5 (x, y) with a strictly negative coefficient.
Our definition of the Tutte polynomial of a flag matroid is admittedly quite involved.
It is natural to wonder whether there is an easier definition, avoiding geometry:
Problem 92 Is there a purely combinatorial description of the K -theoretic Tutte
polynomial of a flag matroid? In particular, can one obtain the K -theoretic Tutte
polynomial from the Tutte polynomials of the constituents?
For matroids, one can define the characteristic polynomial (also called chromatic
polynomial, as it generalises the chromatic polynomial of a graph) by
4 Flag Matroids: Algebra and Geometry 111
In 2015, Adiprasito, Huh and Katz proved the following conjecture by Rota-Heron-
Welsh:
Theorem 93 ([1]) Let wi (M) be the absolute value of the coefficient of λr (M)−i in
the characteristic polynomial of M. Then the sequence wi (M) is log-concave:
Since we now have a definition for the Tutte polynomial of a flag matroid, we can
define the characteristic polynomial of a rank k flag matroid F by
Acknowledgements We would like to thank Christopher Eur for useful comments and sharing
with us interesting examples. We thank Alastair Craw and the anonymous reviewers for their careful
reading and suggestions to improve the manuscript. The third author was supported by the Polish
National Science Centre grant no. 2015/19/D/ST1/01180.
References
1. Adiprasito, K., Huh, J., Katz, E.: Hodge theory for combinatorial geometries. Ann. Math.
188(2), 381–452 (2018)
2. Anderson, D.: Introduction to equivariant cohomology in algebraic geometry. In: Contributions
to Algebraic Geometry, EMS Series of Congress Reports, pp. 71–92. European Mathematical
Society, Zürich (2012)
3. Borovik, A.V., Gel’fand, I.M., White, N.: Coxeter Matroids, Progress in Mathematics, vol.
216. Birkhäuser Boston Inc, Boston, MA (2003)
4. Brändén, P.: Obstructions to determinantal representability. Adv. Math. 226(2), 1202–1212
(2011)
5. Bruns, W., Ichim, B.: Normaliz: algorithms for affine monoids and rational cones. J. Algebra
324(5), 1098–1113 (2010)
6. Brylawski, T.: Constructions. In: Theory of Matroids, The Encyclopedia of Mathematics and
its Applications, vol. 26, pp. 127–223. Cambridge University Press, Cambridge (1986)
7. Brylawski, T., Oxley, J.: The Tutte polynomial and its applications. In: Matroid Applications, the
Encyclopedia of Mathematics and Its Applications, vol. 40, pp. 123–225. Cambridge University
Press, Cambridge (1992)
8. Cameron, A., Fink, A.: The Tutte polynomial via lattice point counting (2018).
arXiv:1802.09859
9. Cox, D.A., Little, J.B., Schenck, H.K.: Toric Varieties, Graduate Studies in Mathematics, vol.
124. American Mathematical Society, Providence, RI (2011)
10. Craw, A.: Quiver flag varieties and multigraded linear series. Duke Math. J. 156(3), 469–500
(2011)
11. Dupont, C., Fink, A., Moci, L.: Universal Tutte characters via combinatorial coalgebras. Algebr.
Comb. 1(5), 603–651 (2018)
12. Edmonds, J.: Submodular functions, matroids, and certain polyhedra. In: Combinatorial Struc-
tures and their Applications (Proc. Calgary Internat. Conf., Calgary, Alta., 1969), pp. 69–87.
Gordon and Breach, New York (1970)
13. Ellis-Monaghan, J.A., Merino, C.: Graph polynomials and their applications I: the Tutte poly-
nomial. In: Structural Analysis of Complex Networks, pp. 219–255. Birkhäuser/Springer, New
York (2011)
14. Eur, C.: Webpage. https://github.com/chrisweur
15. Feichtner, E.M., Sturmfels, B.: Matroid polytopes, nested sets and Bergman fans. Port. Math.
(N.S.) 62(4), 437–468 (2005)
16. Fink, A., Speyer, D.E.: K -classes for matroids and equivariant localization. Duke Math. J.
161(14), 2699–2723 (2012)
17. Fulton, W.: Introduction to Toric Varieties, Annals of Mathematics Studies, vol. 131. Princeton
University Press, Princeton, NJ (1993). The William H. Roever Lectures in Geometry
18. Fulton, W.: Intersection theory, Ergebnisse der Mathematik und ihrer Grenzgebiete. 3. Folge.
A Series of Modern Surveys in Mathematics [Results in Mathematics and Related Areas. 3rd
Series. A Series of Modern Surveys in Mathematics], vol. 2, 2nd edn. Springer, Berlin (1998)
19. Fulton, W., Harris, J.: Representation Theory, Graduate Texts in Mathematics, vol. 129.
Springer, New York (1991). A first course, Readings in Mathematics
20. Gale, D.: Optimal assignments in an ordered set: an application of matroid theory. J. Comb.
Theory 4, 176–180 (1968)
4 Flag Matroids: Algebra and Geometry 113
21. Gel’fand, I.M., Goresky, R.M., MacPherson, R.D., Serganova, V.V.: Combinatorial geometries,
convex polyhedra, and Schubert cells. Adv. Math. 63(3), 301–316 (1987)
22. Gel’fand, I.M., Serganova, V.V.: Combinatorial geometries and the strata of a torus on homo-
geneous compact manifolds. Uspekhi Mat. Nauk 42(2(254)), 107–134, 287 (1987)
23. Gel’fand, I.M., Serganova, V.V.: On the general definition of a matroid and a greedoid. Dokl.
Akad. Nauk SSSR 292(1), 15–20 (1987)
24. Goresky, M., Kottwitz, R., MacPherson, R.: Equivariant cohomology, Koszul duality, and the
localization theorem. Invent. Math. 131(1), 25–83 (1998)
25. Grayson, D.R., Stillman, M.E.: Macaulay2, a software system for research in algebraic geom-
etry. http://www.math.uiuc.edu/Macaulay2/
26. Greene, C.: A multiple exchange property for bases. Proc. Am. Math. Soc. 39, 45–50 (1973)
27. Herzog, J., Hibi, T.: Discrete polymatroids. J. Algebr. Combin. 16(3), 239–268 (2002, 2003)
28. Huh, J.: Milnor numbers of projective hypersurfaces and the chromatic polynomial of graphs.
J. Am. Math. Soc. 25(3), 907–927 (2012)
29. Kálmán, T.: A version of Tutte’s polynomial for hypergraphs. Adv. Math. 244, 823–873 (2013)
30. Kim, S.: Flag enumerations of matroid base polytopes. J. Combin. Theory Ser. A 117(7),
928–942 (2010)
31. Lafforgue, L.: Chirurgie Des Grassmanniennes. CRM Monograph Series, vol. 19. American
Mathematical Society, Providence, RI (2003)
32. Lasoń, M.: On the toric ideals of matroids of a fixed rank. Selecta Math. (N.S.) 27(2), Paper
No. 18, 17 (2021)
33. Lasoń, M., Michałek, M.: On the toric ideal of a matroid. Adv. Math. 259, 1–12 (2014)
34. Manivel, L.: Symmetric functions, Schubert polynomials and degeneracy loci. SMF/AMS Texts
and Monographs, vol. 6. American Mathematical Society, Providence, RI; Société Mathéma-
tique de France, Paris (2001). Translated from the 1998 French original by John R. Swallow,
Cours Spécialisés [Specialized Courses], 3
35. McMullen, P.: Lattice invariant valuations on rational polytopes. Arch. Math. (Basel) 31(5),
509–516 (1978/79)
36. Merino, C., Ramírez-Ibáñez, M., Rodríguez-Sánchez, G.: The Tutte polynomial of some
matroids. Int. J. Comb. Art. ID 430859, 40 (2012)
37. Merkur’ev, A.S.: Comparison of the equivariant and the standard K -theory of algebraic vari-
eties. Algebra i Analiz 9(4), 175–214 (1997)
38. Michałek, M.: Selected topics on toric varieties. In: The 50th anniversary of Gröbner bases.
Adv. Stud. Pure Math. Math. Soc. Japan, Tokyo 77, 207–252 (2018)
39. Michałek, M., Sturmfels, B.: Invitation to nonlinear algebra, vol. 211. American Mathematical
Society (2021)
40. Miller, E., Sturmfels, B.: Combinatorial Commutative Algebra, Graduate Texts in Mathematics,
vol. 227. Springer, New York (2005)
41. Mnëv, N.E.: The universality theorems on the classification problem of configuration varieties
and convex polytopes varieties. In: Topology and geometry—Rohlin Seminar. Lecture Notes
in Mathematics, vol. 1346, pp. 527–543. Springer, Berlin (1988)
42. Nielsen, H.A.: Diagonalizably linearized coherent sheaves. Bull. Soc. Math. France 102, 85–97
(1974)
43. Oxley, J.: Matroid theory. Oxford Graduate Texts in Mathematics, vol. 21, 2nd edn. Oxford
University Press, Oxford (2011)
44. Oxley, J., Whittle, G.: A characterization of Tutte invariants of 2-polymatroids. J. Combin.
Theory Ser. B 59(2), 210–244 (1993)
45. Reiner, V.: Lectures on Matroids and Oriented Matroids (2005). https://www-users.math.umn.
edu/~reiner/Talks/Vienna05/Lectures.pdf
46. Schrijver, A.: Combinatorial optimization. Polyhedra and efficiency. Algorithms and Combi-
natorics, vol. C, 24. Springer, Berlin (2003). Disjoint paths, hypergraphs, Chapters 70–83
47. Seynnaeve, T.: Webpage. https://mathsites.unibe.ch/seynnaeve/index.html
48. Stanley, R.P.: Enumerative combinatorics. Cambridge Studies in Advanced Mathematics, vol. 1,
49. Cambridge University Press, Cambridge (1997). With a foreword by Gian-Carlo Rota,
Corrected reprint of the 1986 original
114 A. Cameron et al.
49. Sturmfels, B.: Gröbner Bases and Convex Polytopes. University Lecture Series, vol. 8. Amer-
ican Mathematical Society, Providence, RI (1996)
50. Sturmfels, B.: Algorithms in Invariant Theory. Texts and Monographs in Symbolic Computa-
tion, 2nd edn. Springer, Wien, NewYork, Vienna (2008)
51. The Sage Developers: SageMath, the Sage Mathematics Software System (Version 8.4.0)
(2018). http://www.sagemath.org
52. Thistlethwaite, M.B.: A spanning tree expansion of the Jones polynomial. Topology 26(3),
297–309 (1987)
53. Tutte, W.T.: A contribution to the theory of chromatic polynomials. Canad. J. Math. 6, 80–91
(1954)
54. Vezzosi, G., Vistoli, A.: Higher algebraic K -theory for actions of diagonalizable groups. Invent.
Math. 153(1), 1–44 (2003)
55. Welsh, D.J.A., Merino, C.: The Potts Model and the Tutte Polynomial. pp. 1127–1152 (2000).
Probabilistic techniques in equilibrium and nonequilibrium statistical physics
56. White, N.L.: The basis monomial ring of a matroid. Adv. Math. 24(3), 292–297 (1977)
57. Whitney, H.: On the abstract properties of linear dependence. Am. J. Math. 57(3), 509–533
(1935)
Chapter 5
Classification of Minimal Polygons with
Specified Singularity Content
Abstract It is known that there are only finitely many mutation-equivalence classes
with a given singularity content, and each of these equivalence classes contains
only finitely many minimal polygons. We describe an efficient algorithm to clas-
sify these minimal polygons. To illustrate this algorithm we compute all mutation-
equivalence classes of Fano polygons with basket of singularities given by (i) B =
{m 1 × 13 (1, 1), m 2 × 16 (1, 1)} and (ii) B = {m × 15 (1, 1)}.
5.1 Introduction
D. Cavey (B)
School of Mathematical Sciences, University of Nottingham, Nottingham NG7 2RD, UK
e-mail: [email protected]
E. Kutas
Mathematics Institute, University of Warwick, Coventry CV4 7AL, UK
e-mail: [email protected]
rial analogues in the Fano polygon P; examples include the singularities and the
anticanonical degree (−K X P )2 . Toric geometry can be further studied in [6, 8].
The dual lattice of N is M = Hom(N , Z) with the pairing ·, · : N × M → Z.
The lattice length of a line segment E ⊂ NR is given by the value |E ∩ N | − 1.
The lattice height of a line segment is given by the lattice distance from the origin:
that is, given the unique primitive inward pointing normal n E of E belonging to M,
the height is given by |v, n E |, for any v ∈ E.
The motivation for the paper comes from an attempt to classify Fano varieties using
mirror symmetry. An understanding of mirror symmetry can be gained from Coates–
Corti–Galkin–Golyshev–Kasprzyk [4]. Mirror symmetry suggests that classifying
Fano polytopes could help to classify Fano varieties.
More specifically we study Fano polytopes up to mutation-equivalence. A muta-
tion is a combinatorial operation on a Fano polygon P ⊂ NR introduced by Akhtar–
Coates–Galkin–Kasprzyk [2], and is described in Sect. 5.2. An important mutation
invariant of Fano polygons known as singularity content was introduced in [3] by
Akhtar–Kasprzyk. This is a combinatorial property of P that describes the singu-
larities of the corresponding Fano toric variety X P . In this setting Fano varieties
are considered up to qG-deformation, see Coates–Corti–Kasprzyk et al. and Kollár–
Shepherd-Barron [1, 12].
Definition 1 ([1]) A del Pezzo surface with cyclic quotient singularities that admits
a qG-deformation to a normal toric del Pezzo surface is said to be of class TG.
The reason we consider Fano polytopes and Fano varieties up to their respective
equivalence classes lies in the following conjecture.
Theorem 4 ([5, 11]) There are precisely 29 qG-deformation families of del Pezzo
surfaces with m ≥ 1 singular points of type 13 (1, 1). Precisely 26 of these are of class
TG, and furthermore are in bijective correspondence with 26 mutation-equivalence
classes of Fano polygons with singularity content (n, {m × 13 (1, 1)}), where m ≥ 1.
Within this context, the aim is to classify Fano polygons with a given singu-
larity content. Assuming Conjecture 2 holds, this is equivalent to a classification
of del Pezzo surfaces admitting a toric degeneration which have the singularities
5 Classification of Minimal Polygons with Specified Singularity Content 117
described by the specified singularity content. We will use our work to provide more
examples for understanding Conjecture 2. The main result of this paper is an efficient
algorithm (described in Sect. 5.4) to produce representations (called minimal poly-
gons) for the mutation-equivalence classes in a classification of Fano polygons with
a specified singularity content in the case where the maximal height of the edges of
each Fano polygon is given by an edge representing a non qG-smoothable singular-
ity. As a corollary to the algorithm the following classifications (derived in Sects. 5.5
and 5.6 respectively) have been completed:
Theorem 5 There are precisely 14 mutation-equivalence classes of Fano polygons
with singularity content (n, {m 1 × 13 (1, 1), m 2 × 16 (1, 1)}) with m 1 ≥ 0, m 2 > 0.
Theorem 6 There are precisely 12 mutation-equivalence classes of Fano polygons
with singularity content (n, {m × 15 (1, 1)}) with m > 0.
The reason for computing these classifications is that they are both in some sense
the simplest cases after the 13 (1, 1) classification of Theorem 4. A 16 (1, 1) singularity,
like a 13 (1, 1), is represented by an edge of a Fano polygon P of height 3. This
is the smallest possible height for an edge representing any singularity that is not
smoothable by a qG-deformation. Whereas 5 is the smallest value for r after 3 for
which a r1 (1, 1) singularity is not smoothable by a qG-deformation.
If one of the Fano polygons appearing in the classifications is (via mutation if
necessary) a triangle, then the corresponding toric variety will be the quotient of a
weighted projective space as described in [9]. In particular let ρ0 , ρ1 , ρ2 ∈ N be the
primitive generators of the rays for the fan of a Fano polygon P. Suppose that ρi
satisfy the equation
λ0 ρ0 + λ1 ρ1 + λ2 ρ2 = 0,
5.2.1 Mutations
By convention P + ∅ = ∅.
Let P ⊂ NR be a Fano polygon, and E be an edge of P. Consider the primitive
inward pointing normal n E ∈ M of this edge. This vector acts as a grading function
on P. For h ∈ Z, define
118 D. Cavey and E. Kutas
Note that ωh (P) may be empty (indeed it will be for infinitely many values of h)
and that ω−r E (P) = E, where r E is the height of E. Choose v E to be a primitive
vector of N such that n E (v E ) = 0. Note this is uniquely defined up to sign in the two
dimensional case. Set F = conv{0, v E }; a line of lattice length 1 and height 0, that
is parallel to E.
Definition 8 For all h < 0, suppose that there exists G h ⊂ NR such that
In the case ωh (P) = ∅ the inclusion holds taking G h = ∅. Then define the mutation
of P with respect to n E , F and G h to be
mut (n E ,F) (P) = conv Gh ∪ (ωh (P) + h F) ⊂ NR .
h<0 h≥0
Example 9 Consider the Fano polygon P = conv{(1, 0), (0, 1), (−5, −1)} corre-
sponding to the weighted projective space P(1, 1, 5). This polygon will be used
as a running
example throughout
the paper. Mutate P with respect to the edge
E = conv (1, 0), (0, 1) . The primitive inner pointing normal is given by
n E = (−1, −1) ∈ M. This describes
a grading on the points of N as shown in Fig. 5.1.
Set F = conv 0, (1, −1) , a primitive slice at height 0. Choose G −1 = {(0, 1)}
which satisfies the required inclusion:
(0, 1), (1, 0) ⊆ G −1 + F ⊆ conv (0, 1), (1, 0) .
-1
6 5 4 3 2 1 0
Q corresponds to the toric variety P(1, 5, 36). Informally the mutation subtracts
one copy of F from P along the edge E, and adds six copies of F at the opposite
vertex of P which is at (−5, −1). This is illustrated in Fig. 5.2.
Note mut (n E ,F) (P) is independent of the choice for G h . If there is no possible
choice of G h , then the mutation with respect to n E does not exist.
Lemma 10 ([11, Lemma 2.3]) Let E be an edge of a Fano polygon P with primitive
inner normal vector n E ∈ M. Then P admits a mutation with respect to n E if and
only if
|E ∩ N | − 1 ≥ r E .
Applying Lemma 10 to the polygon P in Example 9 gives that the edge conv{(0, 1),
(−5, −1)} does not admit a mutation: it has lattice length 1 and height 5.
There are a number of additional properties of mutations.
1. The choice of v E is not important: mut (n E ,F) (P) is isomorphic to mut (n E ,−F) (P)
via a GL(N )-equivalence.
2. Mutation is invertible: If Q = mut (n E ,F) (P), then P = mut (−n E ,F) (Q).
3. P is a Fano polytope if and only if mut (n E ,F) (P) is a Fano polytope
[2, Proposition 2].
Definition 11 Let P, Q ⊂ NR be two Fano polygons. Then P and Q are mutation-
equivalent if there exists a finite sequence of polygons P0 , P1 , . . . , Pn such that
P0 ∼
= P, Pn ∼ = Q and, Pi+1 = mut(ni ,Fi ) (Pi ) for some n i and Fi , for all
i ∈ {0, . . . , n − 1}.
120 D. Cavey and E. Kutas
p 1
= a1 −
q a2 − a −1 1
..
3
.
= [a1 , . . . , ak ].
k
KY = π ∗ K Z + di E i .
i=1
Let [a1 , . . . , akσ ] be the Hirzebruch–Jung continued fraction of R/(a − 1). Then
the values −ai are the self-intersection numbers of the exceptional divisors E i . Addi-
tionally define αi , βi for i ∈ {1, . . . , kσ } by:
α1 = βkσ = 1,
αi
= [ai−1 , . . . , a1 ], for i ∈ {2, . . . , kσ },
αi−1
βi
= [ai+1 , . . . , akσ ], for i ∈ {1, . . . , kσ − 1}.
βi+1
Note that the αi are in increasing order, and the βi are in decreasing order. The
discrepancy of E i is given by di = −1 + (αi + βi )/R. For further reading on minimal
resolutions, see Reid [15].
Proposition 19 ([3, Proposition 3.3, Corollary 3.5]) Let P be a Fano polygon and
let X P be the corresponding toric surface. Suppose P has singularity content (n, B).
Then
5 Classification of Minimal Polygons with Specified Singularity Content 123
(−K X P )2 = 12 − n − Aσ ,
σ ∈B
k σ 2 kσ −1
where Aσ = kσ + 1 − i=1 di ai + 2 i=1 di di+1 . Furthermore the anticanonical
Hilbert series of X P admits a decomposition
1 + (−K X P )2 − 2 t + t 2
Hilb(X P , −K X P ) = + Q σ (t),
(1 − t)3 σ ∈B
R−1
where Q R1 (1,a−1) (t) = 1
1−t R
(δai − δ0 )t i−1 is the Riemann–Roch contribution
i=1
j
coming from the singularity R1 (1, a − 1) and δ j = R1 ∈μ R ,=1 (1−)(1− a−1 ) are the
Dedekind sums.
Example 20 Recall from Example 16 that P = conv (0, 1), (1, 0), (−5, −1) has
1
singularity content 2, { 5 (1, 1)} . The Hirzebruch–Jung continued fraction of the
cyclic quotient singularity 15 (1, 1) is simply [5], so d1 = − 35 and A 15 (1,1) = 15 .
2
+t 3
Also Q 15 (1,1) = t−2t
5(1−t 5 )
. Therefore the anticanonical degree and Hilbert series of
X = P(1, 1, 5) are given by
49
(−K X )2 = ,
5
1 + 8t + 2t 3 − 2t 4 − 8t 6 − t 7
Hilb X, −K X = .
(1 − t 5 )(1 − t)3
1
(−K X P )2 = 12 − n − m,
5
Hilb(X P , −K X P ) =
−t 7 + (n − 10)t 6 + (m − 1)t 5 − 2mt 4 + 2mt 3 + (1 − m)t 2 + (10 − n)t + 1
.
(1 − t)3 (1 − t 5 )
minimal polygon from [11]. For a polygon P, the notation ∂ P denotes the boundary
of P.
Example 24 Considering P = conv{(0, 1), (1, 0), (−5, −1)}, calculate that:
v = (0, 1) + (1, 0) + (−5, −1) = (−4, 0).
v∈vert(P)
By the definition of Fano polygon, 0 ∈ int(P). Therefore the union of all cones
obtained from a Fano polygon P is equal to NR , so P has at least one special facet.
We use a result from [10] which is derived from a proof in [7].
where F (P) is the set of edges of P. Similarly define m B to be the maximum height
among the cones representing the R-singularities of P.
The classification of Fano polygons with a given basket of singularities B up to
mutation-equivalence is split into two cases:
1. m P = m B
2. m P > m B
The proof of [11, Theorem 6.3] efficiently tackles Case 2. Note the polygons
this proof outputs are not necessarily minimal. It remains to deal with Case 1. An
algorithm to compute this classification has been completed in [10]. However tack-
ling classifications beyond the case of polygons with only 13 (1, 1) R-singularities is
inefficient.
The main result of this paper is an efficient algorithm to tackle Case 1. The
algorithm is as follows: start with only a single vertex (a, l F ) = v ∈ F where F is
the special facet. We can assume that the other endpoint of F is (b, l F ), where b < a.
Since m B = m P we have a bound on the heights on all edges. Given a point v1 consider
the set of points
where E (v1 ,v2 ) is the line segment from v1 to v2 . Note S is a subset of a line L h . Given
vertices v1 , v2 , . . . vk consider the lines L 1 , . . . L m B then pick all the valid points u
on these lines that give us an edge E (vk ,u) that respects convexity and defines either
a T-singularity or a residual singularity of B. Lemma 25 gives a bound on how low
these lines can go. This often suffices as bound, however it is possible that either
1. there exists h such that L h is horizontal, or
2. there exists u ∈ L h such that the E (vk ,u) is horizontal.
126 D. Cavey and E. Kutas
Theorem 26 The algorithm gives a complete classification for Fano polygons with
a specific basket of singularities B.
Proof It suffices to show that at each stage there are only finitely many choices of
vertices, this is clear from the above description. There are also only finitely many
choices for inputs as up to a linear transformation 0 < a ≤ l f and l f ≤ m B , so there
are only finitely many choices of a.
k≥0
is an invariant under mutation. Hence two polygons with different periods cannot
be mutation-equivalent. We have computer code in Sage [16] that efficiently imple-
ments the algorithm.
5 Classification of Minimal Polygons with Specified Singularity Content 127
We apply our algorithm to classify all Fano polygons whose only R-singularities
are the cyclic quotient singularities 1
3
(1, 1) and 1
6
(1, 1). Set
B = {m 1 × 3 (1, 1), m 2 × 6 (1, 1)}, where m 1 ∈ Z≥0 and m 2 ∈ Z>0 . Here m 2 is
1 1
non-zero since a classification for Fano polygons with only 13 (1, 1) R-singularities
has been completed in [11].
In the 13 (1, 1) classification of [11], a bound on the number of R-singularities
is found by substituting the degree contribution A 13 (1,1) > 0 into an expression for
the anticanonical degree of the corresponding toric Fano variety from Proposition 19
since we know this value to be positive. However the degree contribution A 16 (1,1) is
128 D. Cavey and E. Kutas
negative and a similar method does not yield a bound. We show that this can be done
by a purely combinatorial argument instead.
Proof The result for m > 3 follows from the base case m = 3.
Let P be a polygon with B = {3 × 16 (1, 1)}. By a GL(N )-translation, assume that
one of the R-singularities is given by E 1 = conv{(−1, 3), (1, 3)}. By mutating with
respect to any T-singularity lying between E 1 and a second R-singularity, assume
this second R-singulary is adjacent to E 1 , given by an edge E 2 with endpoints (1, 3)
and (a, b). The primitive inner pointing normal of E 2 is given by
b − 3 1 − a
n E2 = , ∈M
g g
3a − b
h = −n E2 · (1, 3) = .
g
3a − b
= 3, b = 3a − 3 gcd(b − 3, 1 − a).
g
A similar argument shows that for a basket B = {m 1 × 13 (1, 1), m 2 × 16 (1, 1)} as
above, then m 1 + m 2 < 3.
Examples in this particular classification demonstrate a notion known as shattering
introduced by Wormleighton [17]. Let C1 = u, v, C2 = v, w be two cones in NR .
Suppose the vectors v − u, w − v are parallel. Then define the hyperplane sum of C1
and C2 to be given by C1 ∗ C2 = u, w.
Table 5.1 The polygons with singularity content of the form (n, {m 1 × 13 (1, 1), m 2 × 16 (1, 1)})
with m 2 = 0
# vert(P) n m1 m2 (−K X )2
1.1 (−1, 3), (1, 3), (0, −1) 2 0 1 32/3
1.2 (−1, 3), (1, 3), (1, 2), (0, −1) 3 0 1 29/3
1.3 (−1, 3), (1, 3), (1, 1), (0, −1) 4 0 1 26/3
1.4 (−1, 3), (1, 3), (1, 0), (0, −1) 5 0 1 23/3
1.5 (−1, 3), (1, 3), (1, 2), (0, −1), (−1, 0) 6 0 1 20/3
1.6 (−1, 3), (1, 3), (1, 2), (0, −1), (−1, −1) 7 0 1 17/3
1.7 (−1, 3), (1, 3), (1, 0), (0, −1), (−1, 0) 8 0 1 14/3
1.8 (−1, 3), (1, 3), (1, 0), (−1, −1) 8 0 1 14/3
1.9 (−1, 3), (1, 3), (1, 0), (0, −1), (−1, −1) 9 0 1 11/3
1.10 (−1, 3), (1, 3), (1, 2), (−1, −4) 10 0 1 8/3
1.11 (−1, 3), (1, 3), (1, −1), (−1, −3) 11 0 1 5/3
1.12 (−1, 3), (1, 3), (5, −1), (−5, −1) 12 0 1 2/3
1.13 (−1, 1), (1, 1), (5, −1), (−5, −1) 12 0 2 4/3
1.14 (−1, 3), (1, 3), (1, −1), (−1, −2) 9 1 1 2
Q σ1 + · · · + Q σn = 0,
lattice length(τ )
A σ1 + · · · + A σn = A τ = .
lattice height(τ )
Consider a T-cone C = cone (−2, 3), (1, 3) of height 3. By adding an additional
ray given by primitive generating vector (−1, 3) decompose C into two sub-cones C1
and C2 representing a 13 (1, 1) and a 16 (1, 1) R-singularity respectively. By Corollary 28
Knowing A 13 (1,1) = 5
3
and Q 31 (1,1) = − 3(1−t
t
3 ) , derive:
2 t
A 16 (1,1) = − , and Q 16 (1,1) = .
3 3(1 − t 3 )
Fig. 5.3 Minimal representatives of mutation-equivalence classes of Fano polygons with singularity
content (n, {m 1 × 13 (1, 1), m 2 × 16 (1, 1)}), where m 1 ≥ 0, m 2 > 0
y3 y2 y 1 1
f = x y 3 + 3x y 2 + ay 3 + 3x y + by 2 + + x + cy + 3 + 3 + + ,
x x x y x
3 2
y y y 1 1
g = x y 3 + 3x y 2 + dy 3 + 3x y + ey 2 + +x + fy +4 +6 +4 + .
x x x x xy
It is easy to see that these periods are not equal and hence the polygons cannot be
mutation-equivalent. All other Fano polygons in this classification have pairwise
distinct singularity content, hence are not mutation equivalent.
5 Classification of Minimal Polygons with Specified Singularity Content 131
We find all Fano polygons with singularity content of the form (n, {m × 15 (1, 1)})
with m > 0. Similarly to Sect. 5.5, bounds on n and m are required to ensure the
complete classification.
Proof Similarly to the proof of Lemma 27, assume the existence of a Fano poly-
gon P with three 15 (1, 1) singularities, and perform a combination of GL(N )-
translations and mutations so that one of the R-singularities is represented by
the edge E 1 = conv{(−3, 5), (−2, 5)}, and another by E 2 = conv{(−2, 5), (a, b)},
where (a, b) = (−3, 5). We show you can always mutate P so that the third R-
singularity is represented by E 3 = conv{(−3, 5), (c, d)}, where (c, d) = (−3, 5),
without disrupting the original two 15 (1, 1) singularities sitting adjacently.
Study the possible T-cones that when mutated with respect to would separate the
adjacent R-singularities. Calculate the line of points from (−2, 5) that give an edge
at height 5, since (a, b) must lie on this line in order for E 2 to define a 15 (1, 1) R-
singularity. Unlike Lemma 27, since we are only interested in 15 (1, 1) singularities,
assume that the inner pointing normal n E2 = (b − 5, −2 − a) is primitive. This line
of points on which (a, b) lies, provides a bound on where (c, d) can lie by convexity.
Convexity also determines that d ≤ 5. Furthermore since P is Fano, the origin (0, 0)
must lie in the interior which further bounds the region (c, d) lies in. Finally since
we are only interested in the case where the prospective T-singularity would disrupt
the adjacent R-cones when mutated with respect to we obtain a final bound on the
Table 5.2 The polygons with singuarity content of the form (n, {m × 15 (1, 1)}) with m > 0
# vert(P) n m (−K X )2
2.1 (−3, 5), (−2, 5), (1, −2) 2 1 49/5
2.2 (−3, 5), (−2, 5), (−1, 3), (1, −2) 3 1 44/5
2.3 (−3, 5), (−2, 5), (−1, 3), (1, −2), (−2, 3) 4 1 39/5
2.4 (−3, 5), (−2, 5), (−1, 3), (1, −2), (−1, 1) 5 1 34/5
2.5 (−3, 5), (−2, 5), (0, 1), (1, −2), (−1, 1) 6 1 29/5
2.6 (−3, 5), (−2, 5), (0, 1), (1, −2), (0, −1) 7 1 24/5
2.7 (−3, 5), (−2, 5), (1, −1), (0, −1) 7 1 24/5
2.8 (−3, 5), (−2, 5), (1, −1), (1, −2), (0, −1) 8 1 19/5
2.9 (−3, 5), (−2, 5), (1, −1), (1, −3) 9 1 14/5
2.10 (−3, 5), (−2, 5), (2, −3), (2, −5) 10 1 9/5
2.11 (−3, 5), (−2, 5), (4, −1), (−3, −1) 11 1 4/5
2.12 (−3, 5), (−2, 5), (3, −5), (2, −5) 10 2 8/5
132 D. Cavey and E. Kutas
region in which (c, d) can lie. It is then possible to exhaustively check that none of
the primitive lattice points in this region give the second vertex of a T-cone.
Hence assume that the three R-cones lie adjacently. Calculating the points (c, d)
so that E 3 is height 5 and comparing with the possible choices of (a, b), note that
there are no choices of (a, b) and (c, d) that maintain convexity.
Therefore
there can
be no minimal Fano polygon with residual basket given
by B = 3 × 15 (1, 1) with m P = 5.
The above method extends very nicely to a combinatorial proof that
Lemma 30 There exist no Fano polygons P ⊂ NR with m P = p and residual basket
given by B = {m × 1p (1, 1)}, where m ≥ 3, p is odd and p ≥ 3.
2.12
Fig. 5.4 Minimal representatives of mutation-equivalence classes of Fano polygons with singularity
content (n, {m × 15 (1, 1)}) where m > 0
5 Classification of Minimal Polygons with Specified Singularity Content 133
All other Fano polygons in the classification have pairwise distinct singularity content
and therefore belong to different mutation equivalence classes.
Acknowledgements We thank Alexander Kasprzyk for his guidance and many useful conversa-
tions. Much of the paper was written in during a visit of EK to Nottingham supported by EPRSC
Fellowship EP/NO22513/1. EK would also like to thank his doctoral advisor, Miles Reid.
References
1. Akhtar, M., Coates, T., Corti, A., Heuberger, L., Kasprzyk, A.M., Oneto, A., Petracci, A.,
Prince, T., Tveiten, K.: Mirror symmetry and the classification of orbifold del Pezzo surfaces.
Proc. Amer. Math. Soc. 144(2), 513–527 (2016)
2. Akhtar, M., Coates, T., Galkin, S., Kasprzyk, A.M.: Minkowski polynomials and mutations.
SIGMA Symmetry Integr. Geom. Methods Appl. 8, Paper 094, 17 (2012)
3. Akhtar, M., Kasprzyk, A.M.: Singularity content (2014). arXiv:1401.5458 [math.AG]
4. Coates, T., Corti, A., Galkin, S., Golyshev, V., Kasprzyk, A.M.: Mirror symmetry and Fano
manifolds. In: European Congress of Mathematics, pp. 285–300. European Mathematical Soci-
ety, Zürich (2013)
5. Corti, A., Heuberger, L.: Del Pezzo surfaces with 13 (1, 1) points. Manuscripta Math. 153(1–2),
71–118 (2017)
6. Cox, D.A., Little, J.B., Schenck, H.K.: Toric Varieties. Graduate Studies in Mathematics, vol.
124. American Mathematical Society, Providence (2011)
7. Dais, D.I., Nill, B.: A boundedness result for toric log del Pezzo surfaces. Arch. Math. (Basel)
91(6), 526–535 (2008)
8. Fulton, W.: Introduction to toric varieties. Annals of Mathematics Studies, vol. 131. Princeton
University Press, Princeton (1993). The William H. Roever Lectures in Geometry
9. Kasprzyk, A.M.: Bounds on fake weighted projective space. Kodai Math. J. 32(2), 197–208
(2009)
10. Kasprzyk, A.M., Kreuzer, M., Nill, B.: On the combinatorial classification of toric log del
Pezzo surfaces. LMS J. Comput. Math. 13, 33–46 (2010)
11. Kasprzyk, A.M., Nill, B., Prince, T.: Minimality and mutation-equivalence of polygons. Forum
Math. Sigma 5, e18, 48 (2017)
134 D. Cavey and E. Kutas
12. Kollár, J., Shepherd-Barron, N.I.: Threefolds and deformations of surface singularities. Invent.
Math. 91(2), 299–338 (1988)
13. Øbro, M.: An algorithm for the classification of smooth Fano polytopes (2007).
arXiv:0704.0049 [math.CO]
14. Reid, M.: Surface cyclic quotient singularities and Hirzebruch–Jung resolutions. http://
homepages.warwick.ac.uk/~masda/surf/more/cyclic.pdf
15. Reid, M.: Young person’s guide to canonical singularities. In: Algebraic Geometry, Bowdoin,
1985 (Brunswick, Maine, 1985), Proceedings of Symposia in Pure Mathematics, vol. 46, pp.
345–414. American Mathematical Society, Providence, RI (1987)
16. The Sage Developers: SageMath, the Sage Mathematics Software System (Version 8.4.0)
(2018). http://www.sagemath.org
17. Wormleighton, B.: Reconstruction of singularities on orbifold del Pezzo surfaces from their
Hilbert series. Comm. Algebra 48(1), 119–140 (2020)
Chapter 6
On the Topology of Fano Smoothings
Abstract Suppose that X is a Fano manifold that corresponds under Mirror Sym-
metry to a Laurent polynomial f , and that P is the Newton polytope of f . In this
setting it is expected that there is a family of algebraic varieties over the unit disc
with general fiber X and special fiber the toric variety defined by the spanning fan
of P. Building on recent work and conjectures by Corti–Hacking–Petracci, who con-
struct such families of varieties, we determine the topology of the general fiber from
combinatorial data on P. This provides evidence for the Corti–Hacking–Petracci
conjectures, and verifies that their construction is compatible with expectations from
Mirror Symmetry.
6.1 Introduction
There has been recent interest in the classification of Fano manifolds via Mirror
Symmetry [1, 7, 8]. For us, an n-dimensional Fano manifold X corresponds under
Mirror Symmetry to a Laurent polynomial f ∈ C[x1±1 , . . . , xn±1 ] if the regularized
quantum period of X , which is a generating function for certain genus-zero Gromov–
Witten invariants of X , coincides with classical period π f of f :
1 1 d x1 d xn
π f (t) = ··· .
(2π i)n S 1 ×···×S 1 1 − t f x1 xn
Yt Y
πt π (6.1)
Xη Xt XP
where the arrow A B means that A is the general fiber in a family over
with special fiber B. Here X η X t is Namikawa’s smoothing of Fano vari-
eties with ordinary double points. The Fano variety X η is our desired smoothing
of X P , and the diagram above allows us to compute its Betti numbers. The central
fiber Y is a toric variety, so we know its cohomology groups explicitly; we can com-
pute the Betti numbers of Yt by analysing the vanishing cycles of the degeneration
Yt Y and, since the left-hand part of the diagram is a conifold transition
from Yt to X η , this determines the Betti numbers of X η .
We begin by reviewing the cohomology of toric varieties and the vanishing cycle
exact sequence. We then explain in Sect. 6.3 how to compute the Betti numbers of the
smoothing X from the decomposition data, and give examples in Sect. 6.4. In Sect. 6.5
we prove that the Betti numbers of X depend on the decomposition data only via
6 On the Topology of Fano Smoothings 137
the Laurent polynomial f determined by those data, and in Sect. 6.6 we verify that
these Betti numbers of X coincide with the Betti numbers of the Fano manifold that
corresponds to f under Mirror Symmetry.
We will compute the Betti numbers of the fiber Y in diagram (6.1) using the fact that
it is a toric variety.1
Theorem 1 (cf. [16, Proposition 3.5.3]) Let be a complete fan in a three-
dimensional lattice and let X be the toric threefold defined by . Let di denote
the number of i-dimensional cones in , and let bi denote the ith Betti number
of X . Then:
d1 − d2 + d3 = 2 b2 = rk Pic(X )
b0 = b6 = 1 b3 = rk Pic(X ) − d2 + 2d1 − 3
b1 = b5 = 0 b4 = d1 − 3
We will compute the Betti numbers of the fiber Yt in diagram (6.1) by analysing the
vanishing cycles for the degeneration Yt Y . Consider a complex variety Y,
a disc ⊂ C, and a projective morphism f : Y → . Let ∗ = \ {0} be the
punctured disc, and
i0 j0
{0} − − ∗
→←
be the natural inclusions. Denote the fiber over t ∈ ∗ by Yt , and the fiber over 0 ∈
by Y . Choose a universal covering map p0 : ∗ → ∗ , and consider the diagram
Y
i
Y
j
Y\Y
p
Y \Y
j0 p0
{0}
i0
∗ ∗ .
Let S be a stratification for Y and suppose that F• ∈ DbS (Y). The nearby sheaf is
defined [14, Exposé I] to be the complex
ψ f F• = i ∗ R( j ◦ p)∗ ( j ◦ p)∗ F• .
i ∗ F• → ψ f F• .
The sheaf of vanishing cycles φ f F• is the cone on this map (ibid.), and there is a
distinguished triangle:
+1
i ∗ F• → ψ f F• → φ f F• −→ (6.2)
Hi (ψ f F• ) y ∼
= Hi (F f,y , F• )
Hi (φ f F• ) y ∼
= Hi+1 (B(y, ) ∩ Y, F f,y ; F• ).
where Hvi (Yt , Q) is the subspace in H i (Yt , Q) generated by vanishing cycles, that
is, cycles in the kernel of the natural map H i (Yt , Q) → H i (Y0 , Q).
Every three-dimensional Gorenstein toric Fano variety is the toric variety X P defined
by the spanning fan of a three-dimensional reflexive polytope P. This gives a one-to-
one correspondence between three-dimensional Gorenstein toric Fano varieties up to
isomorphism and three-dimensional reflexive polytopes up to GL(3, Z)-equivalence.
In general such a toric variety X P is singular. As mentioned in the Introduction, Corti–
Hacking–Petracci construct, starting from a three-dimensional reflexive polytope P
decorated with some additional data, a smoothing X → of X P . In this section we
describe their construction and the additional data required.
6 On the Topology of Fano Smoothings 139
F1 + e1 , F2 + e2 , . . . , Fk + ek
There is a whole body of theory here, which we will not discuss, concerning
polyhedral subdivisions which may be neither fine nor mixed: an introduction to
these topics can be found in the extremely beautiful book by De Loera et al. [11].
We will only consider regular fine mixed subdivisions.
For the rest of this section, fix a three-dimensional reflexive polytope P. Corti–
Hacking–Petracci consider the polytope P together with decomposition data. This
is, for each facet F of P:
A. a choice of admissible Minkowski decomposition F = F1 + · · · + Fk ;
B. a choice of regular fine mixed subdivision of F subordinate to (A);
satisfying a condition that we now describe. Note that by taking cones over the regular
fine mixed subdivisions of each facet, we obtain a complete fan that refines the
140 T. Coates et al.
spanning fan of P, and thus a toric crepant partial resolution Y → X P . Recall first
that irreducible toric curves in X P (respectively in Y ) correspond to two-dimensional
cones in the spanning fan of P (respectively in the fan ). Thus irreducible toric
curves in X P correspond to edges of P, and irreducible toric curves in Y correspond
to edges in the polyhedral subdivision of the boundary of P determined by the
decomposition data.
This amounts to the statement that each polygon in the regular fine mixed subdi-
vision (B) above is either a standard 2-simplex or a quadrilateral with sides of unit
length. The curves in the toric 1-skeleton of Y , therefore, either meet at 3-valent
vertices—which are the torus-fixed points on Y corresponding to the cones over
the 2-simplices—or at 4-valent vertices, which are the torus-fixed points on Y cor-
responding to the cones over the quadrilaterals. The 3-valent vertices are smooth
points on Y , and the 4-valent vertices are the qODPs.
Let e denote the irreducible toric curve in X P determined by the edge e of P,
and let e denote the set of irreducible toric curves in Y that map dominantly to e
under the partial resolution Y → X P . The fact that each polygon in the subdivision
(B) above is either a standard 2-simplex or a quadrilateral with sides of unit length
implies that | e | =
(e), the lattice length of the edge e.
Let denote the toric 1-skeleton of Y . Consider the partial normalisation
→
constructed by normalising each 4-valent vertex as shown in Fig. 6.1.
The partial normalisation
consists of rational curves that meet at either bivalent
or trivalent vertices. Let e denote an edge of P, and e denote the corresponding toric
curve in X P . We define two partitions of the set e , as follows. Let p ∈ X P be one of
the endpoints of e , and F ⊂ P the corresponding face of P. Consider the part
p
of the partially-normalised toric 1-skeleton
that lies over p; this consists of the
dual graph to the polyhedral subdivision (B) of F, partially normalised as described
in Fig. 6.1. The components of the partially-normalised dual graph define a partition
of
e . There is one such partition for each of the two endpoints p of e : let us denote
them by e and
e (the order will not matter). The condition that Corti–Hacking–
Petracci impose on their decomposition data is: for each edge e of P such that the
dual edge e has lattice length
(e ) equal to 1 and for each pair T ∈ , T
∈
, we
have |T ∩ T
| ≤ 1.
6 On the Topology of Fano Smoothings 141
Corti–Hacking–Petracci prove:
Theorem 8 Let P be a 3-dimensional reflexive polytope and let X P be the toric
Fano threefold associated to the spanning fan of P. Fix decomposition data for P as
above, and let π : Y → X P be the associated crepant toric partial resolution. Then:
1. Y is unobstructed and smoothable;
2. if Yt is a general smoothing of Y , then Yt is a weak Fano threefold and the
anticanonical morphism πt : Yt → X t , where X t = Proj R(Yt , −K Yt ), contracts
a finite number of disjoint nonsingular rational curves, each with normal bun-
dle O(−1) ⊕ O(−1);
3. X t is a Fano threefold with ordinary nodes as singularities and it is a deformation
of X P .
A theorem of Namikawa [18, Theorem 11] now implies that X t is smoothable. It
follows that X P is smoothable. Let X η denote a generic smoothing of X t . Our goal
is to compute the Betti numbers of X η .
We begin by analysing the topology of Yt . The vanishing cycles for the degen-
eration Yt Y are three-dimensional spheres. The sheaf of vanishing cycles φ f
from (6.2), which is concentrated at the nodes of Y , therefore has stalk at each node
equal to Q concentrated in degree three, and the vanishing cycle exact sequence (6.3)
gives
b0 (Yt ) = b0 (Y ) b4 (Yt ) = b2 (Y )
b1 (Yt ) = b1 (Y ) b5 (Yt ) = b1 (Y )
(6.5)
b2 (Yt ) = b2 (Y ) b6 (Yt ) = b0 (Y )
b3 (Yt ) = b3 (Y ) − b4 (Y ) + b2 (Y ) + k
b0 (X η ) = b0 (Yt ) = 1 b4 (X η ) = b2 (Yt ) − l
b1 (X η ) = b1 (Yt ) = 0 b5 (X η ) = b1 (Yt ) = 0
(6.6)
b2 (X η ) = b2 (Yt ) − l b6 (X η ) = b0 (Yt ) = 1
b3 (X η ) = b3 (Yt ) + 2m − 2l
the two partitions of e defined in the discussion around Fig. 6.1, and define n c,c
by starting with
(e) and subtracting one for each partition that has c and c
in the
same part. The set of exceptional curves for πt : Yt → X t is conjecturally indexed by
edges e of P such that
(e) ≥ 2 and pairs of distinct elements c, c
∈ e : it contains
precisely n c,c
curves in the homology class
Fv
v∈(c,c
)
and no others. Note that n c,c
here is non-negative; it can be zero. This conjecture
determines the subspace L ⊂ H2 (Yt ) spanned by exceptional curves, and thus deter-
mines l = dim L. For m, suppose that the elements of the two partitions of e have
sizes a1 , a2 , . . . and b1 , b2 , . . . respectively. Set
(e) ai bi
Ne =
(e ) − − . (6.7)
2 i
2 i
2
Then
m= Ne . (6.8)
e :
(e)≥2
2 The fiber Fv is homologous in Y to the sum of toric curves corresponding to 2-dimensional cones
in that contain the ray spanned by v and that lie entirely on one side of the hyperplane defined
by the edge e. The choice of side does not matter here, as the resulting sums are homologous.
6 On the Topology of Fano Smoothings 143
6.4 Examples
6.4.1 Cube
Consider the cube P centered at the origin with vertices (±1, ±1, ±1). This has six
non-simplicial facets and twelve edges of length two; thus the toric variety X P defined
by the spanning fan of P has six singular points and twelve curves of transverse A1
singularities. These are arranged as on the right-hand side of Fig. 6.2, with the singular
points at the vertices of the octahedron and the singular curves as the edges.
Each facet F of P is a square with side-length two; this has a unique admissible
Minkowski decomposition, as a Minkowski sum of four line segments, which in turn
leads to the unique fine mixed subdivision of F shown in Fig. 6.3.
The fan that defines the partial resolution Y of X P is therefore obtained by taking
cones over the polyhedral decomposition of the boundary of P shown in Fig. 6.4.
The variety Y contains 24 ordinary double points and 48 toric curves, arranged as on
the right-hand side of Fig. 6.4, with the singular points at the vertices and the toric
curves (along which Y is non-singular) as the edges. Furthermore each edge in the
dual polygon P has length 1, and each partition e of e is into singleton sets, so
the polyhedral decomposition satisfies the condition to be decomposition data.
Applying Theorem 1 gives
b0 (Y ) = 1 b4 (Y ) = 23
b1 (Y ) = 0 b5 (Y ) = 0
b2 (Y ) = 4 b6 (Y ) = 1
b3 (Y ) = 5
b0 (Yt ) = 1 b4 (Yt ) = 4
b1 (Yt ) = 0 b5 (Yt ) = 0
b2 (Yt ) = 4 b6 (Yt ) = 1.
b3 (Yt ) = 10
b0 (X η ) = 1 b4 (X η ) = 1
b1 (X η ) = 0 b5 (X η ) = 0
b2 (X η ) = 1 b6 (X η ) = 1
b3 (X η ) = 28
The only smooth Fano 3-fold with these Betti numbers is V8 . Thus X η is isomorphic
to V8 , which is consistent with the fact that the Minkowski polynomial
(0, 0, 1), (0, 1, −1), (1, 1, −1), (1, 0, −1), (0, −1, −1), (−1, −1, −1), (−1, 0, −1).
This polytope P is reflexive. It has six facets that are standard simplices, one non-
simplicial facet (a hexagon), and 12 edges of length 1. Thus the toric variety X P
defined by the spanning fan of P contains six non-singular toric points, a unique
singular point (at which the singularity is a cone over the del Pezzo surface of
degree 6), and 12 non-singular toric curves. These are arranged as on the right-hand
side of Fig. 6.6, with the non-singular toric points at the 3-valent vertices, the singular
point at the 6-valent vertex, and the toric curves as the edges.
The hexagonal facet F of P admits two Minkowski decompositions: as the sum of
three line segments, and as the sum of two triangles. Up to automorphism, these each
give rise to a unique fine mixed subdivision of F, as shown in Fig. 6.7. Consider first
the left-most fine mixed subdivision in Fig. 6.7. This leads to the polyhedral subdi-
vision of the boundary of P shown in Fig. 6.8. The fan given by taking cones over
Fig. 6.6 The polytope P and a schematic picture of the toric variety X P
146 T. Coates et al.
b0 (Y ) = 1 b4 (Y ) = 5
b1 (Y ) = 0 b5 (Y ) = 0
b2 (Y ) = 2 b6 (Y ) = 1
b3 (Y ) = 0
and since there are k = 3 quadrilaterals in the polyhedral subdivision of the boundary
of P, we have
b0 (Yt ) = 1 b4 (Yt ) = 2
b1 (Yt ) = 0 b5 (Yt ) = 0
b2 (Yt ) = 2 b6 (Yt ) = 1.
b3 (Yt ) = 0
The discussion in Sect. 6.3.1 implies that, assuming the conjectures of Corti–
Hacking–Petracci, there are no exceptional curves in Yt and so the smoothings X η
of X P and Yt of Y are isomorphic. Thus
6 On the Topology of Fano Smoothings 147
b0 (X η ) = 1 b4 (X η ) = 2
b1 (X η ) = 0 b5 (X η ) = 0
b2 (X η ) = 2 b6 (X η ) = 1.
b3 (X η ) = 0
These are the Betti numbers of the hypersurface W1,1 of bidegree (1, 1) in P2 × P2 ,
which is consistent with the fact that the Minkowski polynomial
(1 + x)(1 + y)(1 + x y)
f = +z
x yz
b0 (Y ) = 1 b4 (Y ) = 5
b1 (Y ) = 0 b5 (Y ) = 0
b2 (Y ) = 3 b6 (Y ) = 1
b3 (Y ) = 0
and since there are k = 2 quadrilaterals in the polyhedral subdivision of the boundary
of P, we have
148 T. Coates et al.
b0 (Yt ) = 1 b4 (Yt ) = 3
b1 (Yt ) = 0 b5 (Yt ) = 0
b2 (Yt ) = 3 b6 (Yt ) = 1.
b3 (Yt ) = 0
Once again, there are (conjecturally) no exceptional curves in Yt and so the smooth-
ings X η of X P and Yt of Y are isomorphic. Thus
b0 (X η ) = 1 b4 (X η ) = 3
b1 (X η ) = 0 b5 (X η ) = 0
b2 (X η ) = 3 b6 (X η ) = 1
b3 (X η ) = 0
These are the Betti numbers of P1 × P1 × P1 , which is consistent with the fact that
the Minkowski polynomial
(1 + x + x y)(1 + y + x y)
f = +z
x yz
Consider the three-dimensional reflexive polytope P, pictured in Fig. 6.10, with ver-
tices
(1, 0, 0), (0, 1, 0), (0, 0, 1), (−1, −1, −1), (−1, −1, 2).
This polytope has four facets that are standard simplices, two non-standard simplicial
facets, eight edges of length 1, and one edge of length 3. Thus the toric variety X P
Fig. 6.10 The polytope P and a schematic picture of the toric variety X P
6 On the Topology of Fano Smoothings 149
defined by the spanning fan of P contains four non-singular toric points, two orb-
ifold points, eight non-singular toric curves, and one toric curve with transverse A2
singularities. These are arranged as on the right-hand side of Fig. 6.10, with the toric
points at the vertices, the orbifold points indicated in red, and the toric curves as the
edges. The curve of A2 singularities is the edge between the two orbifold points.
Figure 6.11 shows the unique fine mixed subdivision of the boundary of P. This
defines decomposition data for P. Taking cones over the polyhedra in this decom-
position gives a fan that defines a toric resolution Y of X P . The variety Y is
smooth, with ten toric points and fifteen toric curves arranged as on the right-hand
side of Fig. 6.11: the toric points are the vertices and the toric curves are the edges.
Note the two toric surfaces in Y that map to the curve of singularities under the
resolution Y → X P .
Applying Theorem 1 gives
b0 (Y ) = 1 b4 (Y ) = 4
b1 (Y ) = 0 b5 (Y ) = 0
b2 (Y ) = 4 b6 (Y ) = 1
b3 (Y ) = 0
and since there are no quadrilaterals in the polyhedral subdivision, we find that the
Betti numbers of Yt coincide with those of Y . Computing the subspace L ⊂ H2 (Yt ) ∼
=
H2 (Y ) of classes of exceptional curves, as in the cube example, we find that l = 2
and that generators for L are as shown in Fig. 6.12. These generators are the fibers
of the toric surfaces in Y that resolve the curve of transverse A2 singularities.
From (6.7) and (6.8) we find that (conjecturally) there are m = 3 exceptional
curves in total, and therefore:
b0 (X η ) = 1 b4 (X η ) = 2
b1 (X η ) = 0 b5 (X η ) = 0
b2 (X η ) = 2 b6 (X η ) = 1.
b3 (X η ) = 2
150 T. Coates et al.
Fig. 6.12 Toric curves that generate the subspace of exceptional curves
These are the Betti numbers of the blow-up X of P3 in a plane cubic, which is
consistent with the fact that the Minkowski polynomial
(1 + z)3
f = +x+y+z
x yz
a. b2 (Y ) = b2 (Y
);
b. the Betti numbers of Yt and Yt
coincide;
c. the Betti numbers of X and X
coincide.
The key point is that the polyhedral decompositions of the boundary of P
that define Y and Y
differ by a sequence of the two types of moves depicted in
Figs. 6.13 and 6.14, or their inverses. The vertices of the outer pentagon in Fig. 6.13
are (0, 0), (1, 0), (α + 1, β), (β, α + 1), (0, 1), where α and β are positive coprime
integers, and the interior lattice point pictured is at (α, β); unless α = β = 1 then
there are other interior lattice points which are not pictured. The precise values of α
and β will not affect the analysis. The vertices of the outer quadrilateral in Fig. 6.14
are at (0, 0), (2, 0), (1, 1), and (0, 1).
Let us analyse how b2 changes under a Type I move. Suppose first that Y1 , Z , and Y2
are three-dimensional toric varieties defined by polyhedral decompositions of the
boundary of P that differ only as shown in Fig. 6.15.3 Then Y1 and Y2 are toric
partial resolutions of X P which differ by a Type I move, and there is a diagram
Y1 Y2
f1 f2
Z
3 The co-ordinates of the vertices and interior lattice point pictured in Fig. 6.15 are as in Fig. 6.13.
152 T. Coates et al.
1 2
Fig. 6.15 The fans for Y1 , Z , and Y2 differ only at the cones over these polygons
4 5
j1∗ j2∗
Pic(Z 0 )
We will identify Pic(Y1 ) and Pic(Y2 ) as subspaces of Pic(Z 0 ). To give a line bundle
on Z 0 is to give a piecewise-linear function on each maximal cone in the fan for Z 0 ,
subject to the constraint that these piecewise-linear functions agree along faces. Let
us write the values of such a piecewise linear function at the vertices of the polyhedral
decomposition that we are considering as in Fig. 6.16.
Since the fan for Z 0 does not include the cone over the pentagon pictured in
Fig. 6.16, that cone does not impose any relation between the values a1 , . . . , a5 .
(There may be other relations from the part of the fan not pictured, but these will
be the same for Z , Y1 , and Y2 .) This piecewise-linear function defines a line bundle
on Y1 if and only if it is piecewise-linear on the two cones pictured on the left-hand
side of Fig. 6.15, that is, if and only if a1 + a3 = a2 + a5 . So
Pic(Y1 ) = a1 + a3 = a2 + a5 } ⊂ Pic(Z 0 ).
These are the same subspace of Pic(Z 0 ); therefore Pic(Y1 ) and Pic(Y2 ) are canonically
isomorphic. Since all maximal cones in the fan for Y1 are full-dimensional, the Picard
group of Y1 is isomorphic to H 2 (Y1 ); the same statement is true for Z and for Y2 .
Thus b2 is invariant under Type I moves.
A result of Fulton and Sturmfels [12, Proposition 1.1] implies that b4 (Yi ) is equal to
three less than the number of vertices in the polyhedral decomposition that defines Yi ,
and so b4 (Y2 ) = b4 (Y1 ) + 1. Furthermore χ (Y2 ) = χ (Y1 ) + 1—here we used The-
orem 1—and so b3 (Y1 ) = b3 (Y2 ). In summary:
An essentially identical argument shows that the Betti numbers of Y are invariant
under Type II moves. We are now in a position to prove Theorem 9.
Since Y and Y
differ by a sequence of moves of Type I and II, and their inverses,
we have that b2 (Y
) = b2 (Y ). This is part (a) of the Theorem. Furthermore b3 (Y
) =
b3 (Y ), and if the sequence of moves connecting Y to Y
contains M moves of Type I
and N moves of (Type I)−1 then b4 (Y
) = b4 (Y ) + M − N , and the numbers k and k
of quadrilaterals in the polyhedral decompositions defining Y and Y
satisfy k
=
k + M − N . The vanishing cycle analysis (6.5) now implies (b).
To prove (c), it suffices to show that the quantities l and m occurring in Eq. (6.6)
are the same for Y and Y
. This is obvious for the number of nodes m: the conjectural
formula (6.8) for m depends only on the sizes of the partitions of e , which in turn
depends on the choices of Minkowski decomposition (A) but not on the fine mixed
subdivisions (B). It remains to show that the dimension l of the subspace L of H2 (Y )
spanned by the classes of exceptional curves in the smoothing Yt is the same as the
dimension l
of the subspace L
of H2 (Y
) spanned by the classes of exceptional
curves in the smoothing Yt
. Let us return to the situation considered in Sect. 6.5.1,
where Y1 and Y2 are three-dimensional toric varieties that differ by a Type I move.
We showed there that H 2 (Y1 ) and H 2 (Y2 ) are isomorphic, via the inclusions
154 T. Coates et al.
H 2 (Y1 ) H 2 (Y2 )
j1∗ j2∗
H 2 (Z 0 ).
Dualising gives
H2 (Y1 ) H2 (Y2 )
j1 ∗ j2 ∗
H2 (Z ).
0
and since the subspace of H2 (Yi ) spanned by exceptional curves is pushed forward
from H2 (Z 0 ) via ji ∗ it follows that the dimension l of this subspace is also invariant
under Type I moves. Repeating this analysis for Type II moves shows that l = l
, and
proves Theorem 9.
The computation of Betti numbers described in Sect. 6.3 can be automated. The key
ingredients are as follows.
1. Algorithms for computing with lattice polyhedra and their duals. There are sev-
eral robust and well-tested implementations here, including those in Magma [4],
Sage [20], and polymake [13].
2. The Kreuzer–Skarke classification [17] of three-dimensional reflexive polytopes.
3. Altmann’s determination [3] of all Minkowski summands of a given polytope.
4. The computation of fine mixed subdivisions (Definition 6), that is, the determi-
nation of all regular triangulations of a Cayley polytope. For this we use Jörg
Rambau’s TOPCOM package [19].
5. An HPC cluster. Some of the computations involved are quite challenging.
Full source code for these computations, written in Magma, can be found at [6].
This relies in an essential way on code from the Fanosearch project [9].
There are 4319 three-dimensional reflexive polytopes, which in total admit more
than a billion decomposition data. These decomposition data give rise to 3857 distinct
Minkowski polynomials,4 which together give mirrors to the 98 three-dimensional
4 The number of Minkowski polynomials here differs slightly from the count in [2], because there
the authors required Minkowski decompositions of facets to satisfy an additional lattice condition
(ibid., Definition 7) and here we do not regard GL(3, Z)-equivalent Minkowski polynomials as the
same.
6 On the Topology of Fano Smoothings 155
Fano manifolds with very ample5 anticanonical bundle. We analysed several mil-
lion decomposition data, including at least one decomposition for each of the 3857
Minkowski polynomials. In each case we found that
This provides evidence for the conjectural picture described in the introduction6 :
that if a Fano manifold X corresponds under Mirror Symmetry to a Laurent poly-
nomial f then there is a degeneration X → with general fiber X and special
fiber the toric variety defined by the spanning fan of the Newton polytope of f . It
also provides evidence for the conjectures of Corti–Hacking–Petracci described in
Sect. 6.3, on the number and homology class of the exceptional curves in their res-
olution πt : Y y → X t . If these conjectures are correct then, in view of Theorem 9,
3857 of these calculations give a computer-assisted rigorous proof of (∗ ).
Acknowledgements This project has received funding from the European Research Council (ERC)
under the European Union’s Horizon 2020 research and innovation programme (grant agreement
No. 682603), and from the EPSRC Programme Grant EP/N03189X/1, Classification, Computation,
and Construction: New Methods in Geometry. We thank Paul Hacking and Andrea Petracci for
many extremely useful conversations, and thank Andy Thomas, Matt Harvey, and the Imperial
College Research Computing Service team for invaluable technical assistance. TC thanks Alexander
Kasprzyk for a number of very useful conversations about computations.
References
1. Akhtar, M., Coates, T., Corti, A., Heuberger, L., Kasprzyk, A.M., Oneto, A., Petracci, A.,
Prince, T., Tveiten, K.: Mirror symmetry and the classification of orbifold del Pezzo surfaces.
Proc. Am. Math. Soc. 144(2), 513–527 (2016)
2. Akhtar, M., Coates, T., Galkin, S., Kasprzyk, A.M.: Minkowski polynomials and mutations.
SIGMA Symmetry Integrability Geom. Methods Appl. 8, Paper 094, 17 (2012)
3. Altmann, K.: The versal deformation of an isolated toric Gorenstein singularity. Invent. Math.
128(3), 443–479 (1997)
4. Bosma, W., Cannon, J., Playoust, C.: The Magma algebra system. I. The user language, pp.
235–265 (1997). Computational algebra and number theory (London, 1993)
5. Candelas, P., Green, P.S., Hübsch, T.: Rolling among Calabi-Yau vacua. Nucl. Phys. B 330(1),
49–102 (1990)
6. Coates, T., Corti, A., Da Silva Jr., G.: Code Repository (2019). https://bitbucket.org/fanosearch/
3d_fano_smoothing
7. Coates, T., Corti, A., Galkin, S., Golyshev, V., Kasprzyk, A.M.: Mirror symmetry and Fano
manifolds. In: European Congress of Mathematics, pp. 285–300. European Mathematical Soci-
ety, Zürich (2013)
5 The seven three-dimensional Fano manifolds without very ample canonical bundle are not expected
to admit Laurent polynomial mirrors with reflexive Newton polytopes, and so fall outside the range
of the Corti–Hacking–Petracci construction.
6 Betti numbers for three-dimensional Fano manifolds can be found in [15].
156 T. Coates et al.
8. Coates, T., Corti, A., Galkin, S., Kasprzyk, A.M.: Quantum periods for 3-dimensional Fano
manifolds. Geom. Topol. 20(1), 103–256 (2016)
9. Coates, T., Kasprzyk, A.M.: Code repository (2019). https://bitbucket.org/fanosearch/magma-
core
10. Corti, A., Hacking, P., Petracci, A.: Smoothing Toric Fano Threefolds (2021). In preparation
11. De Loera, J.A., Rambau, J., Santos, F.: Triangulations. Structures for algorithms and applica-
tions. Algorithms and Computation in Mathematics, vol. 25. Springer-verlag, Berlin (2010)
12. Fulton, W., Sturmfels, B.: Intersection theory on toric varieties. Topology 36(2), 335–353
(1997)
13. Gawrilow, E., Joswig, M.: Polymake: a framework for analyzing convex polytopes. In:
Polytopes—Combinatorics and Computation (Oberwolfach, 1997), DMV SEM, vol. 29, pp.
43–73. Birkhäuser, Basel (2000)
14. Grothendieck, A., Raynaud, M., Rim, D. (eds.): Groupes de monodromie en géométrie
algébrique. I. Lecture Notes in Mathematics, Vol. 288. Springer, Berlin, New York (1972).
Séminaire de Géométrie Algébrique du Bois-Marie 1967–1969 (SGA 7 I)
15. Iskovskikh, V.A., Prokhorov, Y.G.: Fano varieties. In: Algebraic geometry, V. Encylopaedia of
Mathematical Sciences, vol. 47, pp. 1–247. Springer, Berlin (1999)
16. Jordan, A.: Homology and cohomology of toric varieties. Ph.D. thesis, Universität Konstanz
(1997)
17. Kreuzer, M., Skarke, H.: Classification of reflexive polyhedra in three dimensions. Adv. Theor.
Math. Phys. 2(4), 853–871 (1998)
18. Namikawa, Y.: Smoothing Fano 3-folds. J. Algebr. Geom. 6(2), 307–324 (1997)
19. Rambau, J.: TOPCOM: triangulations of point configurations and oriented matroids. In: Math-
ematical Software (Beijing, 2002), pp. 330–340. World Scientific Publishing, River Edge, NJ
(2002)
20. The Sage Developers: SageMath, the Sage Mathematics Software System (Version 8.4.0)
(2018). http://www.sagemath.org
Chapter 7
Computing Seshadri Constants on
Smooth Toric Surfaces
7.1 Introduction
L·C
(X, L; x) = inf ,
C mult x (C)
where the infimum is taken over all irreducible curves C passing through x.
The motivation for studying Seshadri constants is that they measure the local positiv-
ity of L at x. This can be seen from the Seshadri criterion which says that L is ample
if and only if (X, L; x) > 0 for all points x ∈ X [11, Theorem 1.4.13]. These con-
stants were first introduced by Demailly in relation to the Nagata conjecture [1, 4].
In most cases the interest lies on the value of the Seshadri constant at a general point,
which is harder to compute with respect to specific points, like torus invariant points
for toric varieties. By semi-continuity the value of these constants at special points
can drop below that of general points. As a consequence the known results are often
either bounds or only valid for certain special points of certain classes of varieties,
see for example [1, 10, 15]. In this paper we approach the problem of computing
Seshadri constants at the general point on smooth polarized toric surfaces. We do so
by relating these constants to other invariants from local positivity and adjunction
theory that are easy to compute on smooth toric surfaces.
The methods used in this paper rely on the fact that a polarized toric variety (X, L)
corresponds to a convex lattice polytope PL . Under this correspondence the fixpoints
of the torus action on X correspond to the vertices of P. The Seshadri constant at these
fixpoints is equal to the lattice length of the shortest edge through the corresponding
vertex of PL , [1, 9], and it is thus easy to compute. It is known that the values
of the Seshadri constant at the general point and at the general point of all torus-
invariant subvarieties of X determine the Seshadri constant at every point of X [9,
Proposition 3.2]. For a general point there is however no exact description, even
though lately some very useful bound has been proven [9, 14].
In this note we related the Seshadri constant to the degree of jet separation, s(L, 1),
and the unnormalized spectral value, μ(L). This allows us to compute the Seshardi
constant at the general point for a large class of smooth toric surfaces. Unfortunately
not for all. For high values of either the degree of jet separation or unnormalized
spectral value the combinatorics quickly becomes intractable, see Example 40. We
will therefore concentrate on low values of these constants.
Definition 2 Let x be a point of a projective variety X with maximal ideal mx . The
degree of jet separation s(L, x) of a line bundle L at the point x ∈ X is the largest
integer k for which the natural map jxk : H 0 (X, L) → H 0 (X, L ⊗ O X /mk+1
x ) is onto.
Like Seshadri constants the degree of jet separation is a semi-continuous invariant.
In fact for smooth toric varieties s(L, x) obtains its minimum at some torus fixpoint
and its maximum at the general point [14]. We will use the notation s(L, 1) to denote
the degree of jet separation at a general point. Our first main result says that if s(L, 1)
is small on a smooth polarized toric surface, then it equals the Seshadri constant.
Theorem 3 Let (X, L) be smooth polarized toric surface with s(L, 1) ≤ 2, then
Observe that since any ample line bundle is big it holds that μ(L) < ∞. Rationality
of the nef and ample cone implies that μ(L) ∈ Q for toric varieties. By the toric
dictionary any big Q-Weil divisor L on a toric variety X of dimension n corresponds
to a n-dimensional polytope PL ⊂ Rn , whose vertices have rational coordinates. Any
such polytope has a unique minimal description as
PL = {x ∈ Rn : Ax ≥ b}
where b ∈ Qn and A is a matrix with integer coefficient with the entries in every row
being relatively prime. If PL is as above, then tK X + L corresponds to the polytope
PtK X +L = {x ∈ Rn : Ax ≥ b + t1},
where 1 = (1, . . . , 1)T . Here the polytope P ab K X +L is the polytope obtained by first
dilating the polytope PL by the factor b and then moving all supporting hyperplanes a
steps inwards. It follows that μ(L) is the maximum over all t such that PtK X +L is
non-empty. Following [5] we will call the polytope Pμ(L)−1 K X +L the core of PL and
denote it by core(PL ). Observe that, unless P is a point, the dimension of PL is
strictly larger than the dimension of its core. Thus if X is a toric surface, PL is a
polygon and core(PL ) is either a line segment or a point. Our second main result is
the following
Theorem 5 Let (X, L) be a smooth polarized toric surface associated to the poly-
tope PL . If core(PL ) is a line segment and μ(L)−1 < 3, then (X, L; 1) = 2μ(L)−1 .
We remark that, unlike the Seshadri constant at the general point, μ(L) is eas-
ily computable in the toric setting, see [5, Proposition 1.14]. It is worth pointing
out that for (P2 , OP2 (1)) it holds that core(PO(1) ) is a point, μ(O(1))−1 = 1/3,
while (P2 , OP2 (1); 1) = 1. Thus the assumption that core(P) is a line segment is
necessary in Theorem 5. We do not however have an example of a smooth polar-
ized toric surface (X, L) such that core(PL ) is a line segment, while (X, L; 1) =
2μ(L)−1 . Unfortunately it is apparent from the proof of Theorem 5 that the combina-
torics involved quickly grows, with increasing values of μ(L)−1 . This is further sup-
ported by Example 34 which shows that further complications will appear for higher
values of μ(L)−1 . Moreover Example 33 indicates that already the case μ(L)−1 = 3
160 S. Di Rocco and A. Lundman
will be more involved than the cases we consider. We leave it as an open question to
find an optimal bound on μ(L)−1 in Theorem 5.
Our next result makes use of Lemma 31, which says that if PL is a polygon
and core(P) is a line segment, then there exist two edges e and e
of PL which
are parallel to core(P). To state our next theorem we let K (P) be the linear space
parallel to the affine hull of core(P). Moreover we will let π : R2 → R2 /K (P)
and π ⊥ : R2 → R2 /K (P)⊥ be the natural projections, where K (P)⊥ is the orthog-
onal complement of K (P).
Theorem 6 Let (X, L) be a smooth polarized toric surface associated to the poly-
tope PL and assume that core(P) is a line segment. If π has a fiber of lattice length
at least 2μ(L)−1 that is a rational polytope or if there is a fiber of π ⊥ intersecting
both e and e
, then (X, L; 1) = 2μ(L)−1 .
Theorem 7 Let (X, L) be a smooth polarized toric surface such that core(P) is
a point, then X is given by a sequence of consecutive equivariant blow-ups of P2
or P1 × P1 .
This paper offers encouraging results for a class of smooth toric embeddings. Results
that open a series of future directions and related questions.
One natural question to ask is if the bounds appearing in Theorems 3 and 5 are
sharp. In the case of Theorem 3 this means asking for a minimal integer k such that
if s(L P , 1) ≤ k, then (X P , L P ; 1) = s(L P , 1). By Example 40 we know that k < 7,
but it would be interesting to find a sharp bound.
In relation to Theorem 5, we do not believe that the equality (X P , L P ; 1) =
2μ(L P )−1 will hold in more generality, whenever core(P) is a line segment. However
we are not aware of any counter example. Finding such an example would be enlight-
ening and give further insight into the relation between (X P , L P ; 1) and μ(L P ),
for such varieties.
Regarding Theorem 7, notice that as P2 and P1 × P1 are projective bundles Theo-
rem 3 implies that s(L, 1) = (X, L; 1) for these surfaces. However when consider-
ing equivariant blow-ups of these surfaces it is not clear how the relationship between
the Seshadri constant and degree of jet separation will change. It is worth pointing out
that, as shown in Example 40, it can happen that s(L, 1), (X, L; 1) and 2μ(L)−1
are all distinct if core(P) is a point. The surface appearing in Example 40 can be
7 Computing Seshadri Constants on Smooth Toric Surfaces 161
described both via six consecutive equivariant blow-ups of P2 and via five consec-
utive equivariant blow-ups of P1 × P1 . If instead X is obtained by a small number
of blow-ups of either P2 or P1 × P1 it seems reasonable to believe that (X, L; 1)
could be computed in terms of s(L, 1) or μ(L)−1 . For example the blow-up of P2 at
one fixpoint is the projective bundle F1 so s(L, 1) = (F1 , L; 1) for any ample line
bundle L on F1 . The main difficulty in proving such statements lies in the number
of different varieties achieved through consecutive equivariant blow-ups grows very
fast in the number of blow-ups performed. We leave the computation of the Seshadri
constant at the general point for these varieties as an open problem.
Another possible direction for future research is trying to achieve similar results
for higher dimensional smooth polarized toric varieties.
7.2 Background
The projectivization of the image P(Im( jxk )) is called the kth osculating space,
Tkx (X, L) of (X, L) at x. Moreover we will say that L is k-jet spanned at x if jxk is
onto and call L k-jet spanned if jxk is onto for all points x ∈ X .
We remark that the map jxk is givenmby sending a section s to its Taylor expan-
sion around x. Thus if H 0 (X, L) = i=1 Csi , then dim(Tkx (X, L)) = rk(Jxk ) − 1,
where Jxk is the matrix whose rows are given by the partial derivatives of (s0 , . . . , sm )
of order at most k evaluated at x. We will call Jxk the matrix of k-jets at x. Thus
given sufficient knowledge about the global sections of a line bundle L, it is straight
forward to check if L is k-jet spanned at a given point x. As in the introduction we
will let s(L, x) denote the degree of jet separation of L at x, which is defined as
the maximal k such that L is k-jet spanned at x. By counting the number of partial
derivatives of order at most k it follows from the above that s(L, x) satisfies the
relation
n + s(L, x)
≤ dim(H 0 (X, L)) < ∞. (7.1)
s(L, x)
Thus the the degree of jet separation, s(L, x) can be computed in finite time.
162 S. Di Rocco and A. Lundman
s(nL, x)
(X, L; x) = lim
n→∞ n
We have the following well-known corollary of Theorem 10, which we will use
repeatedly in this paper.
Corollary 11 Let X be a projective variety and L be a nef line bundle on X . For
any point x ∈ X it holds that
Proof If L is k-jet spanned line bundle, then nL is nk-jet spanned. Therefore it holds
that s(nL, x) ≥ ns(L; x) and the corollary follows by taking the limit in Theorem 10.
In this paper we assume that the reader is familiar with the basic properties of toric
varieties. We recommend [3] as a good general introduction. A key fact in toric
geometry is that any normal toric variety X corresponds to a polyhedral fan .
Moreover every torus invariant subvariety of codimension d on a toric variety X
corresponds to a d-dimensional cone in the fan . In particular any torus invariant
prime divisor Dρ corresponds to a 1-dimensional cone ρ, which is typically called
a ray of , and fixed points m σ correspond to the cones σ.
This interplay between toric and convex geometry is stronger when considering
polarized toric varieties (X, L). In fact any such variety (X, L) corresponds to a
convex lattice polytope PL . To see how this correspondence is defined let D be
a divisor having the property that O(D) = L. As the class group of a projective
toric variety
X is generated by the torus invariant prime divisors of X , it holds
that D = ρ∈(1) aρ Dρ , where (1) denotes the set of rays of and aρ ∈ Z. We
then define the polytope PL associated to (X, L) as
PL = {x ∈ Rn : x, ρ ≥ −aρ , ∀ρ ∈ (1)}.
Via this connection many concepts in algebraic geometry can be understood in con-
vex geometric terms and vice versa. One example is the following theorem. After
choosing a basis of the lattice Zn and hence choosing an affine patch Uσ correspond-
7 Computing Seshadri Constants on Smooth Toric Surfaces 163
ing to the vertex vσ placed at 0, then one has a convenient monomial basis for the
global sections:
Theorem 12 ([3, Theorem 4.3.3]) Let X be a projective toric variety and L a line
bundle on X, then
H 0 (X, L) ∼
= C < t1m 1 · · · tnm n > .
(m 1 ,...,m n )∈PL ∩Zn
where the monomial basis is express in terms of torus coordinates (t1 , · · · , tn ) in the
corresponding affine patch Uσ .
In particular every entry in the matrix of k-jets, Jxk is the evaluation of a monomial.
This leads to the following proposition.
Proposition 13 ([14]) Let X be a smooth polarized toric variety. Then s(L, 1) = k
if and only if there exist a polynomial of degree k + 1 vanishing on the lattice points
of PL and any other polynomial with the same property has at least degree k + 1.
We recall that a polarized toric variety (X, L) of dimension n is smooth if and only
if for every vertex v of PL the shortest lattice vectors along the edges through v form
a basis for Zn . We will call these vectors the edge-directions at v.
Next we introduce maps of fans which are the combinatorial equivalent of equiv-
ariant maps of toric varieties.
Theorem 16 ([6, Theorem VI.7.5]) Let X be a smooth toric surface. Then there
exist a chain of equivariant blow-downs of X to either a Hirzebruch surface or P2 .
We next focus on Seshadri constants on toric varieties.
Definition 17 Let P be a full-dimensional polytope. Then P is called a Cayley
polytope if the vertices of P are contained in two parallel hyperplanes H0 and H1 .
Moreover if the lattice distance between H0 and H1 is k, then P is said to be of
type [P0 ∗ P1 ]k , where P0 = P ∩ H0 and P1 = P ∩ H1 . Lastly if P is a Cayley
polytope, then we call the projection onto the linear space perpendicular to H0 and H1
the defining projection of P.
In particular P is a Cayley polytope of type [P0 ∗ P1 ]1 if and only if P has lattice
width 1. In this case the defining projection projects P onto a line segment of lattice
length 1.
The two following theorems relate Seshadri constants and Cayley polytopes.
Theorem 18 ([10, Theorem 1.3]) Let P be a lattice polytope associated to the
polarized toric variety (X P , L P ), then the following are equivalent:
1. (X P , L P ; 1) = 1;
2. P = [P0 ∗ P1 ]1 ;
3. for every point p ∈ X P there exist a curve C containing p, having the property
that (C, L|C ) ∼ = (P1 , OP1 (1)).
Theorem 19 ([12]) Let P be a smooth lattice polytope associated to the polar-
ized toric variety (X P , L P ) and let k be a positive integer. Then the following are
equivalent:
1. (X P , L P ; x) = k for all points x ∈ X P ;
2. s(L P , x) = k for all points x ∈ X P ;
3. P = [P0 ∗ P1 ]k and every edge of P has lattice length at least k.
The following theorem gives bounds on the Seshadri constant at a general point
and is the most important result in establishing the new results of this paper.
Theorem 20 ([9, Theorem 3.6]) Let M be a lattice and MR = M ⊗R R. Let fur-
ther P ⊂ MR be a smooth polygon and π : M → Z be a lattice projection. Then
Given a linear system |L| on projective variety X , a linear system of the form |K X +
sL| is called an adjoint system to L. In this paper we are interested in two invariants
associated to |K X + sL|, namely the nef value and the unnormalized spectral value.
Both the nef value and the unnormalized spectral value have convex geometric
interpretations for toric varieties. We briefly recall these interpretations.
Remark 23 Let P be a lattice polytope associated to the polarized toric vari-
ety (X P , L P ). For any s ∈ Q the polytope PsK X P +L P is denoted by P (s) . The nef
value is the minimum over all s such that the inner-normal fan of P is a refinement of
the inner-normal fan of P (1/s) . The unnormalized spectral value μ(L P ) of (X P , L P )
is the minimum over all s such that P (1/s) is non-empty. The polytope P (1/μ(L P )) ,
called the core of P, is a polytope of strictly smaller dimension than P and is denoted
by core(P). We remark that the core of a polytope and μ(L P )−1 are easily computable
using [5, Proposition 1.14]. These definitions are illustrated in Fig. 7.2, for further
details on these correspondences we refer to [5].
Definition 24 Let P ⊂ Rn be a polytope and let K (P) be the linear space parallel
to the affine span of the core of P. The projection π : Rn → Rn /K (P) is called
the natural projection associated to P.
Remark 25 If one restricts the attention to polarized toric surfaces, then there are
two possibilities, either core(P) is a point or a line segment. When the core of P is
a point the natural projection is the identity and when it is a line segment the natural
projection is the projection onto a line.
166 S. Di Rocco and A. Lundman
Fig. 7.2 A polytope P, P (1/τ (L P )) , core(P) = P (1/μ(L P )) and the natural projection
In the following section we prove Theorem 3. To make the exposition more accessible
we first prove Lemmas 27 and 28 which are key-steps in the proof of the theorem.
We further make the following notational definition.
Definition 26 Let P be a smooth polytope, then we say that P is canonically posi-
tioned if P has a vertex at the origin and an edge along each coordinate axis in the
positive direction.
Proof Since P has dimension 2 the assumption that P is a Cayley polytope means
that either P ∼= k 2 = conv((0, 0), (k, 0), (0, k)) or P is the convex hull of two line
segments. If P ∼ = k 2 then s(L P , 1) = (X P , L P ; 1) = k. If instead P = [l0 , l1 ]k ,
where l0 and l1 are line segments, then there are two cases: either k = s(L P , 1)
or k = s(L P , 1). If k = s(L P , 1), then projecting onto the defining projection of P
shows that (X P , L P ; 1) ≤ s(L P , 1) which by Corollary 11 implies s(L P , 1) =
(X P , L P ; 1). If instead k = s(L P , 1), we first observe that it must hold that k >
s(L P , 1), otherwise the defining projection of P would give that (X P , L P ; 1) <
s(L P , 1) by Theorem 20, which contradicts Corollary 11. Thus without loss of gen-
erality we can assume that k > s(L P , 1) and that P is canonically positioned with
the longest defining line segment along the x-axis. The assumption k > s(L P , 1)
implies that (0, s(L P , 1) + 1) ∈ P. There are now two cases:
1. (s(L P , 1) + 1, 0) ∈
/ P, in which case the projection onto the first coordinate axis
shows that (X P , L P ; 1) = s(L P , 1) by Theorem 20 and Corollary 11.
2. (s(L P , 1) + 1, 0) ∈ P, in which case (s(L P , 1) + 1) 2 ⊆ P by convexity. This
implies that P is (s(L P , 1) + 1)-jet spanned at the general point contradicting
the definition of s(L P , 1).
From the above we conclude that s(L P , 1) = (X P , L P ; 1).
7 Computing Seshadri Constants on Smooth Toric Surfaces 167
Proof By Lemma 27 we can without loss of generality assume that P is not a Cayley
polytope. Assume that P is canonically positioned, then (0, 0), (1, 0), (0, 1) ∈ P by
the smoothness assumption. Furthermore we claim that the points (2, 1) and (1, 2)
must lie in P. To see why this is the case note that if either of these points fail to be
in P, then by smoothness and convexity either P = 2 2 or all lattice points of P
lie on two parallel lines at lattice distance 1 apart. In the former case s(L P , 1) =
(X P , L P ; 1) = 2 and the latter is a contradiction to s(L P , 1) = 2. From convexity
it now follows that (1, 1) ∈ P. Thus the indicated lattice points in Fig. 7.3 lie in P.
In Fig. 7.3 we also introduce four lines and enumerate some open regions delimited
by these lines.
We now observe that if either the horizontal, vertical or both diagonal lines are
supporting lines of P, then there exists a projection showing that (X P , L P ; 1) ≤ 2
by Theorem 20. Thus without loss of generality we can assume that P contains at
least one point on the other side of the horizontal and vertical line and one point
outside the diagonal strip. Those constrains can be satisfied in a few different ways
and we will end the proof by considering all possibilities.
(1) If P contains a lattice point in region 4, then convexity implies that the
points (2, 2) and (3, 2) lie in P. Furthermore if there is a point in region 4, then
the line L 2 cannot be a supporting line of P. By the smoothness assumption at the
vertex of P on the x-axis, different from the origin, it then follows that (3, 1) ∈ P.
The matrix of 3-jets evaluated at 1, for the 9 lattice points we now know lie in P,
consists of 9 independent rows and a row of zeros corresponding to the derivative ∂∂y 3 .
3
that L P is 3-jet spanned at the general point if it contains a point in region 4, which
is a contradiction. We remark that this argument also takes care of the case when
168 S. Di Rocco and A. Lundman
there is a point of P on the line L 1 to the right of (3, 2). Since in that case either L 1
is a supporting line of P or P contains a point whose y-coordinate is at least 3. In
the former case (X P , L P ; 1) ≤ 2 and there is nothing to prove while in the latter
case L P is generically 3-jet spanned. Moreover by symmetry P cannot have a point
in region 2 and can be assumed not to have a point on L 2 above (2, 3).
(2) Assume that P contains a lattice point in region 5. We claim first that without
loss of generality we can assume that (3, 1) ∈ P. To see this note simply that other-
wise (3, 0) ∈ P but (3, 1) ∈ / P, which implies that P must be a Cayley polytope. By
smoothness and convexity it then follows that (3, 0) ∈ P or (3, 2) ∈ P. If (3, 2) ∈ P
we can apply the same argument as in the case when P contains a point in region 4.
Thus we can without loss of generality assume that (3, 0) ∈ P but that (3, 2) ∈ / P.
Next we note that we can assume that (2, 2) ∈ P. To see this note that otherwise one
readily checks that, (1, 2) has to be a vertex of P. Thus L 1 is a supporting line of P
and the projection onto the y-axis shows that (X P , L P ; 1) = 2. Thus we can assume
that (2, 2) ∈ P. Let J13 be the matrix of 3-jets at the general point for the convex hull
of the points we now can assume lie in P. Then J13 has nine linearly independent
rows and one row of zeros corresponding to the derivative ∂∂y 3 . Thus there are two
3
possibilities: either P contains a point above L 1 and L P is 3-jet spanned at the general
point or the projection onto the y-axis shows that (X P , L P ; 1) = 2. In conclusion
if P contains a point in region 5 and s(L P , 1) = 2, then (X P , L P ; 1) = 2. The case
when P contains a point in region 1, follows by symmetry.
By the above we can assume that all lattice points of P lie in the convex hull of L 2
and L 3 , except possibly the points (2, 0) and (0, 2). Thus if (X P , L P ; 1) = 2, then P
must contain a point in the region 3 or on one of the lines L 2 and L 3 further away from
the origin than (2, 1) or (2, 1) respectively. However convexity and smoothness then
show that neither (2, 0) nor (0, 2) can be a point in P, since if for example (2, 0) ∈ P
then (3, 1) ∈ P, which is a contradiction. Thus P is contained in the convex hull of L 2
and L 3 and the diagonal projection shows that (X P , L P ; 1) = 2.
Proof If |P ∩ Zn | < 20, then s(L P , 1) ≤ 2, by the bound (7.1) and Theorem 3
implies the claim. If instead |P ∩ Zn | = 20, then either s(L P , 1) ≤ 2 or s(L P , 1) = 3
and P = 3 2 by the main theorem in [7]. However 3 2 corresponds to (P2 , O(3)),
so the corollary follows since s(O(3), 1) = (P2 , O(3); 1) = 3.
We claim that s(L P , 1) = 1 while (X P , L P , 1) = 2. To prove the claim note first
that s(L P , 1) = 1 since f (x, y) = x(x − 1) + y(y − 1) − x y is a degree 2 polyno-
mial vanishing at the black points. To show that (X P , L P , 1) ≥ 2 it is by Theo-
rem 10 enough to show that s(2L P , 1) = s(2L P
, 1) = 4, where P
is the complete
embedding corresponding to the polytope conv(P ∩ M). This follows from the fact
that 2P
∩ M = 2P ∩ M. Which in turn follows from the following set of inclusions
of sets.
2P ∩ M ⊇ P ∩ M + P ∩ M = P
∩ M + P
∩ M = 2P
∩ M ⊇ 2P ∩ M
Here the first and last inclusion follows by definition, while the second equality
follows by the projective normality of P
. Checking the first equality can be done
directly since a priori a lattice point p = (x, y) + (z, w) ∈ P
∩ M + P
∩ M can
fail to be in P ∩ M + P ∩ M if and only if (x, y) = (1, 1) or (z, w) = (1, 1). By
symmetry assume (x, y) = (1, 1). If (z, w) ∈ {(0, 0), (1, 1), (2, 1)}, then (x, y) −
(0, 1) ∈ P ∩ M and (z, w) + (0, 1) ∈ P ∩ M hence
If (z, w) ∈ {(0, 1), (1, 0)}, then (x, y) − (1, 1) ∈ P ∩ M and (z, w) + (1, 1) ∈ P ∩
M. If instead (z, w) = (1, 2), then (x, y) − (1, 0) ∈ P ∩ M and (z, w) + (1, 0) ∈
P ∩ M. Finally if (z, w) = (2, 2), then (x, y) + (1, 0) ∈ P ∩ M and (z, w) − (1, 0)
∈ P ∩ M. We conclude on the one hand that 2P ∩ M = 2P
∩ M, i.e. that (X P ,
L P , 1) ≥ 2. On the other hand it is clear that s(nL P , 1) ≤ s(nL P
, 1), since the global
sections of L P is a subspace of the global sections of L P
. Thus (X P , L P , 1) =
(X P , L P
, 1) = 2 by Theorem 10.
In the current section we prove Theorems 5 and 6. We start by showing the following
lemma.
Lemma 31 Let P be a polygon. If core(P) is a line segment, then
Assume that P is canonically positioned with the edge e of Lemma 31 along the x-
axis. By Theorem 20 it is enough to show the existence of a projection π such
that |π(P)| ≥ 2μ(L P )−1 and a fiber F of π that satisfies |F| ≥ 2μ(L P )−1 and
corresponds to a rational polytope. Choosing π to be the natural projection proves
the first part of the theorem since e and e
are at distance 2μ(P)−1 apart.
It remains to consider the case when there is a fiber of the projection onto the x-
axis containing both a point in e and a point in e
. Let π be the projection onto
the x-axis. The assumption is equivalent to requiring that there exists a fiber of π
with length exactly 2μ(P)−1 . Since e and e
are lattice line segments, the fiber can
be taken to be a lattice polytope. It remains to show that |π(P)| ≥ 2μ(P)−1 .
Define now e = (e1 , e2 , . . . , e p ) to be the sequence of distinct edges of maximal
length such that:
1. e1 is the edge sharing a vertex with e different from the origin;
2. ei shares a vertex with ei−1 ;
3. e p = e
.
Let furthermore K be the sequence defined by letting ki be the slope of the edge ei
and ki = ∞ if ei is a vertical edge. From convexity we first observe that the
sequence K can be:
1. positive and strictly monotonically increasing, not including ∞;
2. positive and strictly monotonically increasing and then negative and strictly
monotonically increasing.
7 Computing Seshadri Constants on Smooth Toric Surfaces 171
2 2
1 1
In the former case we will say that K has one phase and in the latter that K has two
phases (see Fig. 7.4). We will call the subsequence of K that is strictly positive and
monotonically increasing the first phase of K and the subsequence which is negative
and strictly monotonically increasing the second phase of K.
Consider first the case when K has one phase and let e p be the last edge in e.
By the smoothness assumption e p must pass through a lattice point on the line y =
2μ(L P )−1 − 1 and end at a vertex of the edge e
. As e
lies on the line y = μ(L P )−1
we can therefore conclude that e p has slope at most 1. Because K is strictly increasing
the same holds for all edges in e. Thus |π(e)| is larger or equal to the distance between
the x-axis and the line y = 2μ(L P )−1 , i.e. |π(P)| ≥ 2μ(L P )−1 .
We now turn to the case when K has two phases. We claim that by the smoothness
of P these two phases of the sequence K are separated by ∞, corresponding to the
slope of a vertical line. To see this let (−a, −b) and (x, y) be the edge-direction
of two adjacent edges in e, where a, b, y > 0 while x ≤ 0. Thus (−a, −b) is the
direction vector of an edge of the first phase of K and (x, y) is a connecting edge
which does not belong to this phase. Then the smoothness condition at the vertex of
intersection between these edges says that
−a x
= bx − ay = ±1.
−b y
Example 33 Consider the smooth polygon P depicted in Fig. 7.5, which has the
property that core(P) is a line segment and μ(L P )−1 = 3.
Here the natural projection is the projection on the y-axis. The longest fibers of
that projection are the fibers between the two dashed lines. The length of any such
172 S. Di Rocco and A. Lundman
core( )
fiber is 11/2. However the point 3 under the projection onto the x-axis has a fiber
containing a point in e and a point in e
. Thus Theorem 6 shows that (X P , L P ; 1) =
2μ(L P )−1 = 6.
However it then follows that f has slope one and that core(P) is a point (or empty)
which is a contradiction.
Thus we can assume that f has a vertex at (1, 0). Then f can be assumed to
pass through the point (3, 1) since otherwise either the fiber of 1 has length at
least 2μ(L P )−1 or we get a contradiction to that core(P) is a line segment. In this
case let g be the edge sharing a vertex with f , other than e. By convexity the slope
of g is greater than that of f . However there must be an edge passing through a
lattice point on the line y = 2μ(L P )−1 − 1 that also shares a vertex with e
. Thus
by smoothness the only possibilities are g = e
or that g has slope 1 or 2/3. As a
consequence either f has length at least 2 of g has slope 2/3, since otherwise core(P)
would be a point (or empty). If f has length 2 the point (5, 2) lies in P. Therefore
the fiber of 2 under the projection onto the y-axis has length at least 2μ(L P )−1 ,
since (1, 2) ∈ P. If instead g has slope 2/3, then (6, 3) ∈ P. However by con-
vexity and smoothness either (2, 3) ∈ P or there is an edge h of slope 1 passing
through the lattice points (0, 1) and (1, 2). In the former case the fiber of 3 under
the projection onto the y-axis has length at least 4. In the latter case by moving the
edges e, e
, f and h we see that P (2) is a point which is a contradiction. This shows
that (X P , L P ; 1) = 2μ(L P )−1 = 4 in this case.
It remains to consider the case μ(L P )−1 = 5/2. Arguing in the same way as above
we see that we can assume that f has a vertex at either (1, 0) or (2, 0) or (3, 0). If f
has a vertex at (3, 0), then we can assume that f pass through the point (4, 1).
Since otherwise the fiber of 1 under the projection onto the y-axis has length at
least 2μ(L P )−1 or P (5/2) would be empty. Thus f can be assumed to have slope 1.
Furthermore f can be assumed to have length 1, since otherwise either (0, 2) ∈ P and
the fiber of 2 under the projection onto the y-axis has length 2μ(L)−1 or (0, 2) ∈ / P
−1
and P μ(L P ) is empty. Let g be the edge different than e sharing an edge with f .
By smoothness and convexity the edge-direction of g is (0, 1), (1, 2) or (2, 3), since
there must be an edge through a lattice point on the line y = 2μ(L P )−1 − 1 sharing
a vertex with e
. If g has edge-direction (1, 2), then the edge sharing a vertex with g
different from f must be vertical. Therefore if g has edge-direction (0, 1) or (1, 2),
we get that P (5/2) is a point or empty by moving e, e
and the vertical edges, yielding a
contradiction. Assume instead that g has edge-direction (2, 3). Then the edge sharing
the vertex (6, 4) with g must have slope 1 by smoothness and convexity, i.e. the same
slope as f which is a contradiction.
If instead f has a vertex at (2, 0), then f can be assume to pass through the
point (4, 1). Again because otherwise P (5/2) would be empty or the fiber of 1 would
have length at least 2μ(L P )−1 . If f has length at least 2, then (6, 2) ∈ P and the
fiber of 2 under the projection on the y-axis has length at least 2μ(L P )−1 . Thus we
can assume that f ends at (4, 1). Let g be the edge sharing an edge with f at (4, 1)
and let (a, b) be the edge-direction of g. Note first that b = 1, 2 since there must be
an edge of P passing through a lattice point on the line y = 2μ(L P )−1 that shares a
vertex with e
having a different slope than f .
If b = 1 then g has slope 1. In this case (1, 3) ∈ P, since otherwise P (5/2) is a
point or empty. Now either g has length at least 2 or g ends at (5, 2). In the former
case the fiber of 3 under the projection onto the y-axis has length 2μ(L P )−1 . In
174 S. Di Rocco and A. Lundman
the latter case let h be the edges sharing a vertex with g different from f . As there
must be an edge i passing through a lattice point on the line y = 2μ(L P )−1 − 1
sharing a vertex with e
, it follows from the smoothness assumption that h has edge-
direction (1, 2) and i is vertical. In this case the point (1, 4) ∈ P, since otherwise
there is an edge of P passing through the lattice points (1, 3) and (2, 4) and P (5/2)
is empty. But as (6, 4) ∈ P the fiber of 4 under the projection onto the y-axis has a
length 2μ(L P )−1 proving that (X P , L P ; 1) = 2μ(L P )−1 .
If instead b = 2, then smoothness and convexity implies that a = 3, i.e. (7, 3) ∈
P. Consider now the edge h different than f which shares a vertex with g. As there
must be an edge i through a lattice point on the line y = 2μ(L P )−1 − 1 sharing
a vertex with e
, the edge-directions of h are on the form (a, 1), with a ∈ Z. By
convexity and smoothness it then follows that a = 1 so that h has slope 1. We now
claim that (2, 3) ∈ P which proves that the fiber of 3 under the projection onto
the y-axis has length at least 2μ(L P )−1 . To see that (2, 3) ∈ P note first that by
smoothness the edge j different from e sharing a vertex with the edge along the y-
axis must contain a point on the line x = 1. If (2, 3) ∈ / P, then j must have slope 1
and end at (1, 2). However that leads to a contradiction since then P (5/2) is a point
(or empty), as can be seen by moving the edges e, e
, h and j.
Lastly we consider the case when (1, 0) is a vertex of P. In this final case we can
assume that the edge f must pass through either (3, 1) or (4, 1). Like before this is
because otherwise we either get a contradiction to that core(P) is a line segment or
the fiber of 1 under the projection onto the y-axis has length at least 2μ(L P )−1 . If f
passes through (4, 1), then either f has length at least 2 or f ends at (4, 1). By the
same argument as above the next edge g has edge-direction (2, 1), (5, 2) or (8, 3).
Thus regardless if f ends at (4, 1) or not it holds that (6, 2) ∈ P. Since (1, 2) ∈ P
it then follows that the fiber of 2 under the projection onto the y-axis has length at
least 2μ(L P )−1 .
It remains to check the case when f passes through (3, 1). In this case the
−1
point (1, 3) ∈ P, since otherwise P (μ(L P ) ) is empty. There are then two possi-
bilities either f has length at least 2 or it ends at (3, 1). It is easy to check that, like
before, the edge-directions of the next edge g is (5, 3), (3, 2) or (1, 1) by smoothness
and convexity. If f has length 1 and g has direction (1, 1) then one gets a contradic-
tion to that core(P) is a line segment. In all other cases (6, 3) ∈ P so the fiber of 3
under the projection onto the y-axis has length at least 2μ(L P )−1 . This concludes
the case μ(L P )−1 = 5/2 (Figs. 7.6, 7.7, 7.8 and 7.9).
Example 34 Let P be the smooth polygon depicted in Fig. 7.10. It can readily be
checked that core(P) is a line segment and that μ(L P )−1 = 4. By direct computation
one can also show that s(L, 1) = 8, which implies that (X P , L P ; 1) = 8. However
the lattice length of a fiber is bounded from above by the maximal lattice length of a
line segment contained in P. It is tedious but straight forward to check that the lattice
length of any line segment in P is strictly less than 8. Thus there exists no projection
of P such that the lower bound in Theorem 20 suffices to compute (X P , L P ; 1).
7 Computing Seshadri Constants on Smooth Toric Surfaces 175
core( )
Remark 35 We are not aware of any smooth polygon P such that the core of P is
a line segment but 2μ(P)−1 = (X P , L P ; 1). However Example 34 illustrates that
if μ(L P )−1 ≥ 4, then the methods used to prove Theorem 5 will not suffice to show
that (X P , L P ; 1) = 2μ(L P )−1 .
Thus X P is a Fano variety since L P was assumed to be ample and μ(L P ) > 0.
7 Computing Seshadri Constants on Smooth Toric Surfaces 177
Proof By definition τ (L P )−1 K X + L P is a line bundle which is nef but not ample.
It follows that (X P , τ (L P )−1 K X P + L P ) corresponds to a polytope P1 whose inner-
normal fan is a coarsening of the inner-normal fan of X , since every normal of an
edge in P1 also is a normal of an edge in P. Let (X 1 , L1 ) be the polarized toric variety
and ample line bundle associated to P1 . Because the fan of X 1 is a coarsening of the
fan of X we get a toric morphism φ : X P → X 1 which contracts some torus invariant
curves on X P . Observe now that if a < τ (L P ), then (X, a −1 K X P + L P ) will yield
the same polygon as (X 1 , (a −1 − τ (L P )−1 )K X 1 + L1 ). To see this let E be a torus
invariant divisor contracted by φ, then (τ (L P )−1 K X P + L P ) · E = 0 and K X P · E <
0. Thus
1 1 1 1
KX P + LP · E = − KX P · E + K X P + L P · E < 0.
a a τ (L P ) τ (L P )
XP X1 ... Xr (7.2)
Here X r has the property that μ(Lr ) = τ (Lr ). The argument for replacing K X P
with K X 1 shows that once the edge corresponding to a prime divisor has disappeared
it will not reappear later in the sequence. Thus all maps corresponds to refinements
of the associated fans. Furthermore the sequence is finite since its length is bounded
from above by the Picard number of X minus one. By the argument in the proof of
Proposition 36 it now follows that X r is a Fano variety.
The following example due to Atsushi Ito shows that (X P , L P ; 1) may fail to be
an integer for smooth polarized toric surfaces. When this happens no combination
of Corollary 11 and Theorem 20 can be used to compute it.
Example 40 Let P be the smooth polygon depicted in Fig. 7.11. It is straight forward
to check that the core of P is a point, s(L P , 1) = 7 and μ(L P )−1 = 4. However
as shown by Ito [8], (X P , L P ; 1) = 15/2. Moreover this example is part of an
infinite family of smooth polarized toric surfaces having the property that the Seshadri
constant is an odd multiple of 1/2. Furthermore for any n Ito has constructed an
infinite family of smooth polarized toric n-folds such that (X P , L P ; 1) is an odd
multiple of 1/2.
Acknowledgements We are grateful to Atsushi Ito for sharing Example 40 with us and letting us
use it in the paper. The authors’ research was supported by the Swedish Research Council, grant
number 2014-4763. The research for this paper was supported by a VR grant [NT:2014-4763].
References
1. Bauer, T., Di Rocco, S., Harbourne, B., Kapustka, M., Knutsen, A., Syzdek, W., Szemberg,
T.: A primer on Seshadri constants. In: Interactions of Classical and Numerical Algebraic
Geometry. Contemporary Mathematics, vol. 496, pp. 33–70. American Mathematical Society,
Providence (2009)
2. Beltrametti, M.C., Sommese, A.J.: The Adjunction Theory of Complex Projective Varieties.
De Gruyter Expositions in Mathematics, vol. 16. Walter de Gruyter & Co., Berlin (1995)
7 Computing Seshadri Constants on Smooth Toric Surfaces 179
3. Cox, D.A., Little, J.B., Schenck, H.K.: Toric Varieties. Graduate Studies in Mathematics, vol.
124. American Mathematical Society, Providence (2011)
4. Demailly, J.P.: Singular Hermitian metrics on positive line bundles. In: Complex Algebraic
Varieties (Bayreuth, 1990). Lecture Notes in Mathematics, vol. 1507, pp. 87–104. Springer,
Berlin (1992)
5. Di Rocco, S., Haase, C., Nill, B., Paffenholz, A.: Polyhedral adjunction theory. Algebra Number
Theory 7(10), 2417–2446 (2013)
6. Ewald, G.: Combinatorial Convexity and Algebraic Geometry. Graduate Texts in Mathematics,
vol. 168. Springer, New York (1996)
7. Fulton, W., Kleiman, S., Piene, R., Tai, H.: Some intrinsic and extrinsic characterizations of
the projective space. Bull. Soc. Math. France 113(2), 205–210 (1985)
8. Ito, A.: Personal communication
9. Ito, A.: Seshadri constants via toric degenerations. J. Reine Angew. Math. 695, 151–174 (2014)
10. Ito, A.: Algebro-geometric characterization of Cayley polytopes. Adv. Math. 270, 598–608
(2015)
11. Lazarsfeld, R.: Positivity in algebraic geometry. I. Ergebnisse der Mathematik und ihrer Gren-
zgebiete. 3. Folge. A Series of Modern Surveys in Mathematics [Results in Mathematics and
Related Areas. 3rd Series. A Series of Modern Surveys in Mathematics], vol. 48. Springer,
Berlin (2004). Classical setting: line bundles and linear series
12. Lundman, A.: Local positivity of line bundles on smooth toric varieties and Cayley polytopes.
J. Symbolic Comput. 74, 109–124 (2016)
13. Nill, B.: Gorenstein toric Fano varieties. Manuscripta Math. 116(2), 183–210 (2005)
14. Perkinson, D.: Inflections of toric varieties. pp. 483–515 (2000). Dedicated to William Fulton
on the occasion of his 60th birthday
15. Szemberg, T.: An effective and sharp lower bound on Seshadri constants on surfaces with
Picard number 1. J. Algebra 319(8), 3112–3119 (2008)
Chapter 8
The Characterisation Problem of
Ehrhart Polynomials of Lattice Polytopes
Akihiro Higashitani
Abstract One of the most important invariants of a lattice polytope is the Ehrhart
polynomial. The problem of which polynomials can be Ehrhart polynomials of lattice
polytopes is now well-studied. In this survey paper, after recalling the fundamental
properties of the Ehrhart polynomials of lattice polytopes, we survey the results
on this problem including recent developments. We discuss the characterisation of
Ehrhart polynomials in several particular cases: small dimensions; small volumes;
palindromic; small degrees. We also suggest some possible further questions.
8.1 Introduction
We say that a convex polytope P ⊂ Rd is a lattice polytope if all of its vertices belong
to the standard lattice Zd . Given a lattice polytope P ⊂ Rd of dimension d, we can
associate the enumerative function n → (n P ∩ Zd ), which counts the number of
lattice points contained in the nth dilation of P. Ehrhart [11] showed that there
exists a polynomial E P (n) in n of degree d, where d is the dimension of P, such
that E P (n) = (n P ∩ Zd ) for any positive integer n. We call the polynomial E P (n)
the Ehrhart polynomial of P. The Ehrhart polynomial E P (n) has the following
properties:
1. the constant term of the polynomial E P (n) is always equal to 1;
2. the leading term of the polynomial is equal to the Euclidean volume of P.
Moreover, E P (n) also satisfies the Ehrhart-Macdonald reciprocity [27]:
A. Higashitani (B)
Department of Pure and Applied Mathematics, Graduate School of Information Science
and Technology, Osaka University, Osaka 565-0871, Japan
e-mail: [email protected]
where P ◦ denotes the relative interior of P. Namely, the number of lattice points
contained in the nth dilation of the interior of P can be expressed by using the Ehrhart
polynomial of P.
When formally expressed as a rational function, the generating function of E P (n)
is known to be
∞ ∗ i
i≥0 h i t
E P (n)t =
n
,
n=0
(1 − t)d+1
and it is known that h i∗ = 0 for i ≥ d + 1 and that each h i∗ is an integer [30]. We call
the integer sequence
h ∗ (P) = (h ∗0 , h ∗1 , . . . , h ∗d )
d
h ∗P (t) = h i∗ t i
i=0
d
n+d −i
E P (n) = h i∗ . (8.2)
i=0
d
∞ d
h ∗d−i t i+1
(n P ◦ ∩ Zd )t n = i=0
. (8.3)
n=1
(1 − t)d+1
In particular, deg(P) ≤ d.
The following is one of the most important unsolved problems in Ehrhart theory:
Problem 1 Characterise the sequences (or polynomials) that are the h ∗ -vectors
(or h ∗ -polynomials) of lattice polytopes.
This problem has two steps: to prove the necessity and the sufficiency. More pre-
cisely, one step is to show some conditions for an integer sequence (h ∗0 , h ∗1 , . . . , h ∗d )
to be the h ∗ -vector of some lattice polytope (e.g., h i∗ ≥ h ∗1 for each 1 ≤ i ≤ d − 1
if h ∗d > 0, h ∗1 ≥ h ∗d and so on), and the other step is to construct a concrete lattice
polytope whose h ∗ -vector coincides with a desired integer sequence.
We see that it is quite difficult to solve this problem in general. Furthermore, it
might be believed that there exists no nice solution to this problem. Thus, in this
survey paper, we will restrict to some particular cases. More concretely, we discuss
the following cases:
1. small dimensions, i.e.,dd is small—see Sect. 8.3;
2. small volumes, i.e., i=0 h i∗ is small—see Sect. 8.4;
3. palindromic, i.e., h i∗ = h ∗deg(P)−i for any i—see Sect. 8.5;
4. small degrees, i.e., deg(P) is small—see Sect. 8.6.
8.2 Preliminaries
In this section, we present the basic concepts that we will use in this paper.
1. For two lattice polytopes P, P
⊂ Rd , we say that P and P
are unimodularly
equivalent if there exist f ∈ GLd (Z) and u ∈ Zd such that P
= f (P) + u.
184 A. Higashitani
This new lattice polytope is said to be a lattice pyramid over P. It is not so hard
to see that for a lattice polytope P, we have [7, Theorem 2.4]
∞
h ∗P (t)
E Pyr(P) (n)t n = .
n=0
(1 − t)d+2
In particular, we have
We recall the finite abelian group associated with a lattice simplex and discuss
some properties on a lattice simplex in terms of this group. Let ⊂ Rd be a lattice
simplex of dimension d with its vertices v1 , v2 , . . . , vd+1 ∈ Zd . Let
d+1
d+1
= (x1 , x2 , . . . , xd+1 ) ∈ [0, 1)d+1 : xi vi ∈ Zd , xi ∈ Z
i=1 i=1
F (d) → A(d); →
In particular,
We see that a lattice simplex of dimension d is not a lattice pyramid if and only if
for each i ∈ [d + 1], there is (x1 , . . . , xd+1 ) ∈ such that xi = 0.
Furthermore, we have
Fig. 8.1 Construction of lattice polygons whose h ∗ -vectors are one of 1, 2 or 3 in Theorem 3
will appear in Sect. 8.5 and the result in the case h ∗3 = 0 will appear in Sect. 8.6. The
characterisation in the case h ∗3 = 1 can be essentially obtained from [23] and the
case h ∗3 = 2 is from [3].
Moreover, by the series of the papers [8–10] by Blanco–Santos, an algorithm
which provides a complete list of all lattice polytopes of dimension 3 up to unimodular
equivalence has been given. This enables us to get the complete list of h ∗ -vectors
of lattice polytopes of dimension 3 containing only a few lattice points. In [10],
the list of lattice polytopes P ⊂ R3 of dimension 3 with (P ∩ Z3 ) ≤ 11 is given
up to unimodular equivalence. As a by-product, we also know the characterisation
of h ∗ -vectors (h ∗0 , h ∗1 , h ∗2 , h ∗3 ) with h ∗1 ≤ 7.
The contribution of the papers [8–10] naturally suggests the following problem:
Problem 4 Characterise the possible h ∗ -vectors of lattice polytopes of dimension 3.
In fact, Balletti [1, Conjecture 8.7] predicts some necessary inequalities for h ∗ -
vectors (h ∗0 , h ∗1 , h ∗2 , h ∗3 ) of lattice polytopes of dimension 3.
Next, we consider the case of small normalized volumes. Namely, we discuss whether
a given integer sequence (h ∗0 , h ∗1 , . . . , h ∗d ) is a possible h ∗ -vector of some lattice
d
polytope of dimension d for the case i=0 h i∗ is small.
8 The Characterisation Problem of Ehrhart Polynomials of Lattice Polytopes 187
d
In the case i=0 h i∗ = 1, since each h i∗ is nonnegative and h ∗0 = 1, we conclude
that h 0 = 1 and h 1 = · · · = h ∗d = 0. This is the h ∗ -vector of a unit simplex of dimen-
∗ ∗
sion d, where a unit simplex of dimension d is the convex hull of the origin of Rd
d of R
d
and the unit vectors .
∗
For the case i=0 h i ≥ 2, we recall two well-known inequalities on h ∗ -vectors.
Let P ⊂ Rd be a lattice polytope of dimension d with degree s. One inequality is
This appears in the work of Hibi [15]. Note that these inequalities (8.6) and (8.7) are
vastly generalized by Stapledon [32, 33]. All inequalities in (8.6) hold with equality
for any i if and only if h i∗ = h ∗s−i for i = 0, 1, . . . , s. This is equivalent to the lattice
polytope being Gorenstein. See Sect. 8.5. All inequalities in (8.7) hold with equality
∗
for any i if and only if h i+1 = h ∗d−i for i = 0, 1, . . . , d − 1. This is called shifted
symmetric, introduced in [19]. A lattice polytope P has a shifted symmetric h ∗ -vector
if and only if P is a lattice simplex all of whose facets have normalized volume 1
(see [19, Theorem 2.1]). d
A characterisation in the cases of i=0 h i∗ = 2 and 3 is given as follows:
Theorem 5 ([17, Theorem 0.1]) Given a sequence (h ∗0 , h ∗1 , . . . , h ∗d ) of nonnegative
d
integers, where h ∗0 = 1 and i=0 h i∗ ≤ 3, there exists a lattice polytope of dimen-
sion d whose h ∗ -vector coincides with (h ∗0 , h ∗1 , . . . , h ∗d ) if and only if (h ∗0 , h ∗1 , . . . , h ∗d )
satisfies the inequalities (8.6) and (8.7).
The “only if” part directly follows from the works by Stanley and Hibi. The
essential part of this theorem is the “if” part.
dThe ∗following example shows that Theorem 5 is no longer true for the case
i=0 h i = 4.
d
h i∗ t i = 1 + t i1 + · · · + t im−1 ,
i=0
This remark implies that for the complete characterisations with larger normalized
volumes, we have to see the h ∗ -vectors of not only lattice simplices but also non-
simplices. Hence, for the further investigation, it is natural to think of the h ∗ -vectors
of lattice simplices as a first step.
Theorem 11 ([20, Theorem 1.1]) Let P be a lattice simplex dof dimension d and
let h ∗ (P) = (h ∗0 , h ∗1 , . . . , h ∗d ) be its h ∗ -vector. Suppose that i=0 h i∗ = p is an odd
d
prime. Let i 1 , . . . , i p−1 be the positive integers such that i=0 h i∗ t i = 1 + t i1 + · · · +
t i p−1 , where 1 ≤ i 1 ≤ · · · ≤ i p−1 ≤ d.Then:
8 The Characterisation Problem of Ehrhart Polynomials of Lattice Polytopes 189
for k,
≥ 1 with k +
≤ 6.
d
Remark 14 (see [20, Sect. 5.1]) Theorems 12 and 13 say that when i=0 h i∗ is 5
or 7, the necessary conditions in Theorem 11 are also sufficient for lattice sim-
plices. However, this is not true for all prime numbers. In fact, since the volume
of a lattice polytope containing a unique lattice point in its interior has an upper
bound (see, e.g., [26]), if p is a sufficiently large prime number, then the integer
sequence (1, 1, p − 3, 1) cannot be the h ∗ -vector of any lattice simplex of dimen-
sion 3, although (1, 1, p − 3, 1) satisfies all the conditions in Theorem 11.
Recently, a complete characterisation of the possible h ∗ -vectors (h ∗0 , h ∗1 , . . . , h ∗d )
d
of lattice polytopes with i=0 h i∗ = 5 has been given.
Theorem 15 ([35, Theorem 0.4]) Let 1 + t i1 + t i2 + t i3 + t i4 be a polynomial with 1
≤ i 1 ≤ · · · ≤ i 4 ≤ d. Then there exists a lattice polytope of dimension d whose h ∗ -
polynomial equals 1 + t i1 + t i2 + t i3 + t i4 if and only if
1. (i 1 , i 2 , i 3 , i 4 ) satisfies the conditions in Theorem 12; or
2. 1 + t i1 + t i2 + t i3 + t i4 = 1 + 3t + t 2 or 1 + t + 3t 2 or 1 + t + t 2 + 2t 3 .
Note that the polynomials 1 + 3t + t 2 and 1 + t + 3t 2 and 1 + t + t 2 + 2t 3 are
the h ∗ -polynomials of some lattice polytopes respectively (see [35, Example 2.4] for
a more detailed construction of those lattice polytopes), but these are never the h ∗ -
polynomials of lattice simplices.
The proof of Theorem 15 relies on the following theorem, which is a recent, major
contribution to Ehrhart theory:
Theorem 16 ([22, Theorem 1.3]) If a lattice polytope P is spanning then the h ∗ -
vector of P has no gap, i.e., h i∗ > 0 for any i = 0, 1, . . . , deg(P).
Here, we say that a lattice polytope P ⊂ Rd is spanning if the affine lattice generated
by P ∩ Zd , denoted by P , is equal to Zd , Note that the converse of Theorem 16 is not
true in general. In fact, the lattice simplex P = conv({e1 , e2 , ±(e1 + e2 + 2e3 )}) ⊂
R3 has the h ∗ -vector (1, 1, 2, 0), while this simplex is not spanning since e3 ∈ Z3 is
not contained in P .
The key lemma for the proof of Theorem 15 is the following:
190 A. Higashitani
Lemma 17 ([35, Theorem 1.1]) Let P be a lattice polytope which is not a simplex.
Suppose that vol(P) is prime. Then P is spanning.
Note that this lemma is not true for non-prime case. For example, let
P = conv({0, e1 , e2 , e1 + e2 , e3 , e1 + e2 + e3 + 2e4 }) ⊂ R4 .
1 + 4t, 1 + 3t + t 2 , 1 + t + 3t 2 , 1 + 2t + 2t 2 , 1 + t + 2t 2 + t 3 , 1 + t + t 2 + 2t 3 , 1 + t + t 2 + t 3 + t 4 .
1 + 2t + 3t 2 + t 3 , 1 + t + 3t 2 + 2t 3 , 1 + t + 2t 2 + 3t 3 , 1 + t + t 2 + 4t 3 ,
1 + t + t 2 + 3t 3 + t 4 , 1 + t + 3t 2 + t 3 + t 4 , 1 + t + t 2 + t 3 + 2t 4 + t 5 .
8.5 Palindromic
s
We say that a polynomial i=0 ai t i of degree s is palindromic if ai = as−i for i =
∗
0, 1, . . . , s. Next, we consider the case of palindromic hs -polynomials. Namely,
we discuss whether a given palindromic polynomial i=0 h i∗ t i of degree s is a
possible h ∗ -polynomial of some lattice polytope with degree s.
8 The Characterisation Problem of Ehrhart Polynomials of Lattice Polytopes 191
is also a lattice polytope, where ·, · denotes the usual inner product of Rd . Reflexivity
of lattice polytopes is characterised as follows:
Proposition 20 (cf. [5, 14]) Let P be a lattice polytope of dimension d, let
(h ∗0 , h ∗1 , . . . , h ∗d ) be its h ∗ -vector and let E P (n) = ad n d + ad−1 n d−1 + · · · + 1 be
its Ehrhart polynomial. Then the following four conditions are equivalent:
1. P is a reflexive polytope;
2. deg(P) = d and h ∗P (t) is palindromic, i.e., h i∗ = h ∗d−i for i = 0, 1, . . . , d;
3. the functional equation E P (n) = (−1)d E P (−n − 1) holds;
4. we have dad = 2ad−1 .
We also recall that a lattice polytope P ⊂ Rd is called Gorenstein of index
if
P is unimodularly equivalent to a reflexive polytope. Since (
P ∩ Zd ) = 1,
we see that deg(P) = dim P + 1 −
by (8.4). The following easily follows from
Proposition 20.
Proposition 21 Let P be a lattice polytope of dimension d with its h ∗ -vector
(h ∗0 , h ∗1 , . . . , h ∗d ). Then the following three conditions are equivalent:
1. P is a Gorenstein polytope of index
;
2. deg(P) = s and h ∗P (t) is palindromic, i.e., h i∗ = h ∗s−i for i = 0, 1, . . . , s, where
s = d + 1 −
;
3. the functional equation E P (n) = (−1)d E P (−n −
) holds.
Reflexive polytopes together with Gorenstein polytopes are particularly impor-
tant in many areas, e.g., commutative algebra, toric geometry and mirror symmetry
(e.g., [4, 5, 31] and so on). Hence, many researchers are quite interested in reflexive
polytopes and they have been intensively studied. On the characterisation of palin-
dromic h ∗ -polynomials, i.e., the h ∗ -polynomials of Gorenstein polytopes, the case
of 2-dimensional reflexive polytopes is well known:
Proposition 22 (See e.g. [28]) Given an integer a, the sequence (1, a, 1) is the h ∗ -
vector of a reflexive polytope of dimension 2 if and only if a ∈ {1, 2, . . . , 7}.
The case of Gorenstein polytopes of degree 2 is also known, which is highly
non-trivial:
Theorem 23 ([4, Theorem 2.10]) Given an integer a, the polynomial 1 + at + t 2 is
the h ∗ -polynomial of a Gorenstein polytope of degree 2 if and only if a ∈ {0, 1, . . . , 7}.
The following is one natural analogue of Theorem 23.
Theorem 24 ([21, Corollary 1.1]) Given an integer a, the polynomial 1 + at k + t 2k
is the h ∗ -polynomial of a Gorenstein polytope of degree 2k with k ≥ 2 if and only if
one of the following conditions holds:
192 A. Higashitani
1. a ∈ {0, 1, 2, 4, 6};
2. there exist
≥ 4 and m ≥ 1 such that a = 2
− 2 and k = 2
−3 m;
3. there exist
≥ 3 and m ≥ 1 such that a = 3
− 2 and k = 3
−2 m.
The case of 3-dimensional reflexive polytopes is also known:
Proposition 25 ([24]) Given an integer a, the integer sequence (1, a, a, 1) is the h ∗ -
vector of a reflexive polytope of dimension 3 if and only if a ∈ {1, 2, . . . , 35} \
{33, 34}.
Similar to this theorem, the necessary and sufficient condition for the integer
sequence (1, a, b, a, 1) to be the h ∗ -vector of a reflexive polytope of dimension 4 is
also known by [25]. (There are 14, 373 possible h ∗ -vectors.)
Since we already know the possible h ∗ -vectors of reflexive polytopes of dimen-
sion 3, the following is a natural problem to try as a next step:
Problem 26 Characterise the h ∗ -polynomials of Gorenstein polytopes of degree 3.
Next, weconsider the case of small degrees. Namely, we discuss whether a given poly-
nomial i=0s
h i∗ t i is a possible h ∗ -polynomial of some lattice polytope of degree s
when s is small.
In general, we see the following:
d
Proposition 27 If a given polynomial i=0 h i∗ t i is the h ∗ -polynomial of some lattice
polytope of dimension d, then it is also the h ∗ -polynomial of some lattice polytope
of dimension d + m for any m ∈ Z≥0 .
In the case degree at most 2, we see that there is a difference between these two
sets. More precisely, the following is known. It looks quite similar to Theorem 3 but
slightly different.
Theorem 28 ([34, Theorem 2]) Let P be a lattice polytope of degree at most 2. Then
its h ∗ -polynomial 1 + h ∗1 t + h ∗2 t 2 satisfies one of the following conditions:
1. h ∗2 = 0;
2. h ∗1 ≤ 3h ∗2 + 3 and h ∗2 > 0; or
3. (h ∗1 , h ∗2 ) = (7, 1).
8 The Characterisation Problem of Ehrhart Polynomials of Lattice Polytopes 193
Note that the conditions in Theorem 28 are also sufficient. Namely, it is proved
in [12, Proposition 1.10] that given nonnegative integers (h ∗1 , h ∗2 ) satisfying one of
the conditions in Theorem 28, there exists a lattice polytope P such that h ∗P (t) = 1 +
h ∗1 t + h ∗2 t 2 . Those polytopes can be constructed by lattice polytopes of dimension
at most 3. Hence, the possible h ∗ -polynomials of degree at most 2 are completely
characterised.
In Theorem 28, the condition “h ∗2 ≤ h ∗1 ” is missing, while b ≤ a appears in
Theorem 3. The inequality b ≤ a comes from h ∗1 ≥ h ∗d (see Introduction) and b ≤ a
is a particular condition for d = 2. Hence, Theorem 3 essentially characterises the
possible h ∗ -polynomials of lattice polytopes of not only dimension 2 but also degree
at most 2.
The following is natural to try as a further problem in addition to Problem 4.
Problem 29 Characterise the h ∗ -polynomials of lattice polytopes of degree ≤ 3.
Acknowledgements The author would like to be grateful to Christopher Borger, Johannes Hof-
scheier, Alexander Kasprzyk and Benjamin Nill for the organization of and the invitation to the
wonderful conference “Interactions with Lattice Polytopes”, held in September 14–16 2017 at
Magdeburg. The author would also like to thank the anonymous referee for a lot of helpful com-
ments.
References
1. Balletti, G.: Enumeration of lattice polytopes by their volume. Discrete Comput. Geom. 65(4),
1087–1122 (2021)
2. Balletti, G., Higashitani, A.: Universal inequalities in Ehrhart theory. Israel J. Math. 227(2),
843–859 (2018)
3. Balletti, G., Kasprzyk, A.M.: Three-dimensional lattice polytopes with two interior lattice
points (2016). arXiv:1612.08918 [math.CO]
4. Batyrev, V., Juny, D.: Classification of Gorenstein toric del Pezzo varieties in arbitrary dimen-
sion. Mosc. Math. J. 10(2), 285–316, 478 (2010)
5. Batyrev, V.V.: Dual polyhedra and mirror symmetry for Calabi-Yau hypersurfaces in toric
varieties. J. Algebraic Geom. 3(3), 493–535 (1994)
6. Batyrev, V.V., Hofscheier, J.: Lattice polytopes, finite abelian subgroups in SL(n, C) and coding
theory (2013). arXiv:1309.5312 [math.CO]
7. Beck, M., Robins, S.: Computing the continuous discretely, 2nd edn. Undergraduate Texts
in Mathematics. Springer, New York (2015). Integer-point enumeration in polyhedra, With
illustrations by David Austin
8. Blanco, M., Santos, F.: Lattice 3-polytopes with few lattice points. SIAM J. Discrete Math.
30(2), 669–686 (2016)
9. Blanco, M., Santos, F.: Lattice 3-polytopes with six lattice points. SIAM J. Discrete Math.
30(2), 687–717 (2016)
10. Blanco, M., Santos, F.: Enumeration of lattice 3-polytopes by their number of lattice points.
Discrete Comput. Geom. 60(3), 756–800 (2018)
11. Ehrhart, E.: Sur les polyèdres rationnels homothétiques à n dimensions. C. R. Acad. Sci. Paris
254, 616–618 (1962)
12. Henk, M., Tagami, M.: Lower bounds on the coefficients of Ehrhart polynomials. Eur. J.
Combin. 30(1), 70–83 (2009)
13. Hibi, T.: Algebraic Combinatorics on Convex Polytopes. Carslaw Publications, Glebe (1992)
14. Hibi, T.: Dual polytopes of rational convex polytopes. Combinatorica 12(2), 237–240 (1992)
15. Hibi, T.: A lower bound theorem for Ehrhart polynomials of convex polytopes. Adv. Math.
105(2), 162–165 (1994)
16. Hibi, T., Higashitani, A., Li, N.: Hermite normal forms and δ-vectors. J. Combin. Theory Ser.
A 119(6), 1158–1173 (2012)
17. Hibi, T., Higashitani, A., Nagazawa, Y.: Ehrhart polynomials of convex polytopes with small
volumes. Eur. J. Combin. 32(2), 226–232 (2011)
18. Hibi, T., Higashitani, A., Tsuchiya, A., Yoshida, K.: Ehrhart polynomials with negative coef-
ficients. Graphs Combin. 35(1), 363–371 (2019)
19. Higashitani, A.: Shifted symmetric δ-vectors of convex polytopes. Discrete Math. 310(21),
2925–2934 (2010)
20. Higashitani, A.: Ehrhart polynomials of integral simplices with prime volumes. Integers 14,
Paper No. A45, 15 (2014)
21. Higashitani, A., Nill, B., Tsuchiya, A.: Gorenstein polytopes with trinomial h ∗ -polynomials.
Beitr. Algebra Geom. 62(3), 667–685 (2021)
8 The Characterisation Problem of Ehrhart Polynomials of Lattice Polytopes 195
22. Hofscheier, J., Katthän, L., Nill, B.: Ehrhart theory of spanning lattice polytopes. Int. Math.
Res. Not. IMRN 19, 5947–5973 (2018)
23. Kasprzyk, A.M.: Canonical toric Fano threefolds. Canad. J. Math. 62(6), 1293–1309 (2010)
24. Kreuzer, M., Skarke, H.: Classification of reflexive polyhedra in three dimensions. Adv. Theor.
Math. Phys. 2(4), 853–871 (1998)
25. Kreuzer, M., Skarke, H.: Complete classification of reflexive polyhedra in four dimensions.
Adv. Theor. Math. Phys. 4(6), 1209–1230 (2000)
26. Lagarias, J.C., Ziegler, G.M.: Bounds for lattice polytopes containing a fixed number of interior
points in a sublattice. Canad. J. Math. 43(5), 1022–1035 (1991)
27. Macdonald, I.G.: Polynomials associated with finite cell-complexes. J. Lond. Math. Soc. 2(4),
181–192 (1971)
28. Poonen, B., Rodriguez-Villegas, F.: Lattice polygons and the number 12. Amer. Math. Monthly
107(3), 238–250 (2000)
29. Scott, P.R.: On convex lattice polygons. Bull. Austral. Math. Soc. 15(3), 395–399 (1976)
30. Stanley, R.P.: Decompositions of rational convex polytopes. Ann. Discrete Math. 6, 333–342
(1980)
31. Stanley, R.P.: On the Hilbert function of a graded Cohen-Macaulay domain. J. Pure Appl.
Algebra 73(3), 307–314 (1991)
32. Stapledon, A.: Inequalities and Ehrhart δ-vectors. Trans. Amer. Math. Soc. 361(10), 5615–5626
(2009)
33. Stapledon, A.: Additive number theory and inequalities in Ehrhart theory. Int. Math. Res. Not.
IMRN 5, 1497–1540 (2016)
34. Treutlein, J.: Lattice polytopes of degree 2. J. Combin. Theory Ser. A 117(3), 354–360 (2010)
35. Tsuchiya, A.: Ehrhart polynomials of lattice polytopes with normalized volumes 5. J. Comb.
10(2), 283–290 (2019)
Chapter 9
The Ring of Conditions
for Horospherical Homogeneous Spaces
Johannes Hofscheier
Abstract These are notes of a five talks lecture series during the “Graduate Summer
School in Algebraic Group Actions”, at McMaster University, June 11th–15th, 2018.
The aim of this lecture series is to introduce the ring of conditions of a spherical homo-
geneous space with a special emphasis on the horospherical case, i.e., homogeneous
spaces with respect to a connected complex reductive group which are torus bundles
over a flag variety. In these notes, we start with an example from enumerative geom-
etry which naturally yields first instances of spherical varieties. We continue with
a recollection of the necessary background on reductive groups needed throughout
the rest of the manuscript. After that we introduce spherical varieties: we discuss the
Luna–Vust theory of spherical embeddings and explain the complete combinatorial
description of horospherical varieties (an important subfamily of spherical varieties).
We conclude with the definition of the ring of conditions of spherical homogeneous
spaces and give an explicit description for the horospherical case.
9.1 Motivation
J. Hofscheier (B)
School of Mathematical Sciences, University of Nottingham, Nottingham NG7 2RD, UK
e-mail: [email protected]
Problem 1 How many lines in 3-space C3 intersect 4 given lines in general position?
Recall the general trick how to rephrase an affine geometric question into a linear
one: Suppose X is an affine geometric object in Cn . Introduce one further dimension
and consider the linear span of X regarded as a subset of the affine hyperplane {xn+1 =
1} where xn+1 denotes the additional coordinate.
Hence, the above question reduces to a problem in the Grassmannian Gr(2, 4) (2-
planes in C4 ). This algebraic variety admits a transitive action by GL4 (the general lin-
ear group of invertible 4 × 4-matrices with complex entries). Indeed, let e1 , . . . , e4 be
the standard basis in C4 and consider the natural action of GL4 on C4 . Clearly GL4 acts
transitively on planes in C4 and the stabilizer P of the coordinate plane span{e1 , e2 }
is given by
AC
P= : A, B ∈ GL2 , C ∈ Mat(2 × 2) .
0 B
So we have to understand the conditions σ1 · σ2 and σ3 · σ4 . By using some
heuristics, Schubert came to the conclusion that “perturbations” of the condi-
tion σ1 · · · σ4 do not change the answer (the conservation of number principle),
i.e., we are allowed to move the lines i . In particular, we may assume that 1 , 2
lie on a plane, and so do 3 and 4 (see Fig. 9.1). From that it straightforwardly fol-
lows that a line intersects both 1 and 2 if and only if it is either contained in the
9 The Ring of Conditions for Horospherical Homogeneous Spaces 199
Fig. 9.1 The conservation of number principle implies that we are allowed to move the lines i
where a is the intersection point of 1 and 2 and is the plane spanned by 1 and 2
and similarly for 3 , 4 .
Thus, we get
Clearly there is exactly one line passing through both a and a and there is exactly
one line contained in both and . On the other hand, as a is not contained in ,
the condition σa · σ is not satisfied by any line, and similarly for σa · σ .
We obtain
Of course, we haven’t given a precise explanation yet and in his fifteenth problem
Hilbert asked for a rigorous foundation of Schubert Calculus. Our goal will be to
understand De Concini’s and Procesi’s solution to Hilbert’s problem. For that, we
also need to understand spherical geometry, a topic which is exciting in its own right.
200 J. Hofscheier
The classical books by Borel, Humphreys and Springer [2, 10, 25] are excellent refer-
ences for what follows. A more modern and accessible book is [18]. For convenience,
we work over the field of complex numbers C.
An algebraic variety G is called an algebraic group if G is a group and the
maps G × G → G, (g, h) → gh and G → G, g → g −1 are morphisms of algebraic
varieties. The Lie algebra of G, usually denoted by g, is the tangent space Te G at
the identity element e ∈ G equipped with a binary operation [·, ·] called the Lie
bracket. Important examples of algebraic groups are GLn (=the general linear
group of invertible n × n-matrices with complex entries), SLn (=the special linear
group of n × n-matrices with complex entries and determinant 1), abelian varieties
(=complete connected algebraic groups) or elliptic curves (=1-dimensional abelian
varieties). We will work with linear algebraic groups, i.e., Zariski closed subgroups
of GLn . If G ⊆ GLn is a linear algebraic group, then Te G = g ⊆ gln = Te GLn =
{(n × n) − matrices}, and the Lie bracket can be defined as the commutator of matri-
ces
[A, B] := AB − B A.
From now on all algebraic groups are assumed to be linear, unless stated otherwise.
G= Bw B
w∈W
In particular,
G/B = Bw B/B.
w∈W
9 The Ring of Conditions for Horospherical Homogeneous Spaces 203
G/P = Bw P/P
w∈W/W P
The closure of B-orbits in G/P are the Schubert varieties (denoted by X (w)).
They play an important role in the study of G/P. The dimension of X (w) equals
the length l(w) of w (i.e., the minimal number of factors needed to write w as a
product of simple reflections). In particular, there exists a unique element w0 of
maximal length in W/W P .
Example 23 Let G = SLn and T be the maximal torus of diagonal matrices con-
tained in the Borel subgroup B of upper-triangular matrices. The Lie algebra g = sln
(i.e., the tangent space Te SLn equipped with the Lie bracket [·, ·]) is the set of
traceless matrices in Mat(n × n) equipped with the commutator bracket [A, B] =
AB − B A. Furthermore, the Lie algebra t of T coincides with the subspace of
diagonal matrices in sln . Observe that the Lie bracket induces a map ad : t →
End(g); A → [A, ·] which is a representation of Lie algebras, i.e., ad([A, B]) =
ad(A) ad(B) − ad(B) ad(A) for any A, B ∈ t (check this!). Let ε1 , . . . , εn be the
linear forms in t∗ induced by the diagonal entries, i.e., εi (diag(t1 , . . . , tn )) = ti and
set εi j = εi − ε j . It is straightforward to show that the Lie algebra decomposes into
eigenspaces as follows
where R + = {εi j : i < j} and this set is called the set of positive roots. The set
of simple roots S = {αi := εi,i+1 : i = 1, . . . , n − 1} (cf. Fig. 9.2) forms a basis of t∗
(check this!) and induces an isomorphism t∗ ∼ = {(x1 , . . . , xn ) ∈ Rn : x1 + . . . + xn =
0} ⊆ R . To any simple root αi one associates a reflection si , namely the linear
n
where gα denotes the linear subspace of eigenvectors of weight α, i.e., the set of
vectors x ∈ g such that [h, x] = α(h)x for all h ∈ t. The Lie algebra b of B can be
written as
b=t⊕ gα
α∈R+
9 The Ring of Conditions for Horospherical Homogeneous Spaces 205
for some subset R + ⊆ R, called the set of positive roots. There exists a unique basis S
contained in R + such that all positive roots are linear combinations of elements in S
with nonnegative integer coefficients. The elements of S are called simple roots.
The following fundamental theorem on parabolic subgroups can be found in any
introductory text on algebraic groups. Recall the definition of the Weyl group: W =
NG (T )/T . Let R be the set of roots and let S be the set of simple roots induced by
the choice of the Borel subgroup B.
Theorem 24 The assignment I → PI = w∈W I Bw B induces a bijection between
subsets of the set of simple roots S and parabolic subgroups of G which contain B
(here W I denotes the group generated by the simple reflections sα for α ∈ I ).
Definition 35 The combinatorial objects needed in the Luna–Vust theory are listed
in Table 9.2. The rank of M is also called the rank of the spherical homogeneous
space G/H , i.e., rk(G/H ) = rk(M). Let N := Hom(M, Z) be the dual lattice
of M and note that we have a dual pairing
·, · : N × M → Z. Furthermore,
set MQ := M ⊗ Q and NQ = Hom(M, Q). Recall, that in our context a valuation
is a map ν : C(G/H )∗ = C(G/H ) \ {0} → Q which satisfies:
1. ν( f 1 + f 2 ) ≥ min{ν( f 1 ), ν( f 2 )} whenever f 1 , f 2 , f 1 + f 2 ∈ C(G/H )∗ ;
2. ν( f 1 f 2 ) = ν( f 1 ) + ν( f 2 ) for all f 1 , f 2 ∈ C(G/H )∗ ; and
3. ν(C∗ ) = 0.
A valuation ν is called G-invariant if ν(g · f ) = ν( f ) for all g ∈ G and f ∈
C(G/H )∗ .
Exercise 38 Find the “Luna–Vust data” for the spherical homogeneous spaces from
Exercise 33. In particular, you should see the following phenomena:
1. SL2 /T : the map ρ : D → N might not be injective;
2. SL2 /N : the image of a color ρ(D) might not be primitive in N;
3. SL2 /B: the image of a color might even be zero, i.e., ρ(D) = 0.
Example 39 The “Luna–Vust data” of the toric case is listed in Table 9.3.
Example 40 Consider the natural action of SL2 on C2 . Let B be the Borel subgroup
of upper triangular matrices, T the maximal torus of diagonal matrices and U the
unipotent radical of B. Denote the standard basis of C2 by e1 , e2 . Then SL2 /U ∼
=
SL2 ·e1 = C2 \ {0} and B · e2 = C × C∗ is the open B-orbit. The rational functions
on SL2 /U are given by C(x, y) = C(A2 ). It is straightforward to verify that
C(SL2 /U )(B) ∼
= {cy k : c ∈ C, k ∈ Z}.
9 The Ring of Conditions for Horospherical Homogeneous Spaces 209
The classification of spherical homogeneous spaces G/H turns out to be quite hard.
Luna’s brilliant insight in spherical varieties made it possible to find such a description
from Wasserman’s list of certain spherical varieties of rank 2 [28]. Inspired by it,
Luna [16] formulated a conjectural description and proved it for spherical varieties
of type A. Only recently Luna’s Conjecture was proven in general with the combined
efforts of several researchers [3, 6, 15]. Unfortunately, time does not permit to give
more details on this exciting topic, instead we will see a complete answer for an
interesting special case.
Lemma 49 Let the torus S act freely on the normal irreducible variety E with good
geometric quotient p : E → E/S. Then for each y ∈ E/S there exists an affine open
neighbourhood U ⊆ E/S of x such that the diagram in Fig. 9.6 commutes and the
upper left isomorphism is S-equivariant.
S × G/H → G/H ; ( p H, x H ) → x p −1 H .
It is straightforward to check that S acts freely on G/H , and thus the result follows
by Lemma 49.
(2) ⇒ (1): Suppose that the fibers of the torus bundle p : G/H → G/P are
isomorphic to the algebraic torus T . As the two actions by G and T commute, the
morphisms ϕt : G/H → G/H ; x H → t · x H for t ∈ T are G-equivariant automor-
phisms. It follows by [26, Proposition 1.8] that we may consider T as a subgroup
of NG (H )/H . Set N := NG (H ). Let T be the preimage of T under the natural pro-
jection map N → N /H . Note that N → N /H is a morphism of algebraic groups
and that T is a closed subgroup of G. Since p −1 (x P) ∼
= T for any x P ∈ G/P, it fol-
lows that a conjugate of P is contained in T, and thus it contains a maximal unipotent
subgroup U . As the natural projection morphism of algebraic groups T → T maps
unipotent elements on unipotent elements, it follows that U is in its kernel which
implies that U ⊆ H .
(1) ⇔ (3) straightforwardly follows from Proposition 47 (3).
Exercise 52 Show that SL2 /Uk is indeed a torus bundle over SL2 /B.
Exercise 53 Use the Luna–Vust theory to classify all spherical embeddings of
SL2 /U where U ⊆ SL2 is a maximal unipotent subgroup. Draw the correspond-
ing colored fans. Hint: You should find 6.
Proposition 54 ([20, Proposition 1.6]) The assignment which associates to a
horospherical subgroup H ⊆ G the pair (M, I ) (see Proposition 47) induces a
bijection between horospherical subgroups of G and pairs (M, J ) where J ⊆ S
and M ⊆ X(T ) is a sublattice such that
α̌, χ = 0 for any α ∈ J and all χ ∈ M.
The horospherical
subgroup associated to a pair (M, J ) as in Proposition 54 is
given by H = χ∈M ker χ where M ⊆ X(PJ ).
Exercise 55 Use the combinatorial description of horospherical subgroups to clas-
sify those contained in SL2 .
A colored fan is called toroidal if F = ∅ for any (C, F ) ∈ . Observe that
in the horospherical case toroidal fans coincide with fans in the toric sense. In this
special case, we have the following explicit construction of horospherical toroidal
varieties:
Proposition 56 ([20, Examples 1.13 (2)]) If H ⊆ G is a horospherical subgroup
containing U and is a toroidal fan, then the corresponding spherical embed-
ding is G-equivariantly isomorphic to G × P X where X denotes the toric variety
corresponding to the fan (with acting torus P/H where P = NG (H )).
214 J. Hofscheier
A good reference for the ring of conditions is the classical paper by De Concini and
Procesi [7].
Let G be a connected complex algebraic group and H ⊆ G a closed subgroup
(not necessarily spherical). Consider the homogeneous space G/H .
Recall that two subvarieties X, Y ⊆ G/H are said to intersect properly if either
X ∩ Y = ∅ or each irreducible component of the intersection X ∩ Y has dimen-
sion dim(X ) + dim(Y ) − dim(G/H ). They are said to intersect transversally if
the intersection X ∩ Y is smooth and has pure dimension dim(X ) + dim(Y ) −
dim(G/H ).
Led by this observation, De Concini and Procesi showed the remarkable fact that
the intersection product of Definition 60 is well-defined on spherical homogeneous
spaces. Let C be the set of smooth (or complete) spherical embeddings G/H
→
X . This set C admits the partial ordering defined such that a spherical embed-
ding G/H
→ X 1 is greater than G/H
→ X 2 if there exists an equivariant mor-
phism X 1 → X 2 . De Concini’s and Procesi’s idea is to show that for any X ∈ C
and any algebraic cycle Y ⊆ G/H , there is an X ∈ C with X ≤ X such that the
closure Y of Y in X intersects the boundary of the open G-orbit in X properly.
The existence of such a “good compactification” ensures that if one considers the
embedding X then we may always assume (up to generic translations by G) that the
intersection with Y takes place in the open G-orbit G/H . To get an isomorphism of
rings, we have to consider “good compactifications” of all algebraic cycles at once.
Theorem 63 ([7, Sect. 6.3]) The intersection product from Definition 60 is well-
defined on a spherical homogeneous space G/H and there is a canonical isomor-
phism of graded rings
where the limit is taken over complete (or equivalently smooth) spherical embed-
dings G/H
→ X .
Exercise 65 Explicitly compute the ring of conditions for some spherical homoge-
neous spaces SL2 /H where H ⊆ SL2 is a spherical subgroup. (Hint: In this case,
the rank of the spherical homogeneous space is bounded by 1, and thus there are
only finitely many spherical embeddings, so that we can straightforwardly compute
the direct limit of cohomology rings.)
Exercise 66 Use the ring of conditions of Gr(2, 4) to solve the “4-lines problem”.
where c1 (δ(m)) ∈ H 2 (G/P, Z) denotes the first Chern class of the line bundle δ(m).
9 The Ring of Conditions for Horospherical Homogeneous Spaces 217
Note that the first ideal in the sum of ideals in Theorem 67 corresponds to the Stanley-
Reisner ideal of the toric variety. The challenge is to find a good description of the
ring in Theorem 67 as we want to take the direct limit over all smooth projective
toroidal fans .
The following approach is inspired by [5]. To keep notation simple, set MQ =
M ⊗ Q and NQ = HomZ (M, Q). Let be a smooth projective toroidal fan in NQ .
A map f : NQ → Q is piecewise polynomial if for any σ ∈ , the map f |σ : σ → Q
extends to a polynomial function on the linear space spanQ {σ }, i.e., a piece-
wise polynomial function f on is a collection of compatible polynomial func-
tions f σ : σ → Q. In particular, such a function is continuous. We denote by R
the set of all piecewise polynomial functions on which is a ring under pointwise
addition and multiplication. Let S ∗ (MQ ) be the symmetric algebra of the Q-vector
space MQ . Recall that S ∗ (MQ ) can be naturally identified with the polynomial
functions on NQ . Note that R is a positively graded Q-algebra with graded subal-
gebra S ∗ (MQ ). Indeed, any piecewise polynomial function f = ( f σ )σ ∈ uniquely
decomposes into a sum of homogeneous piecewise polynomial functions.
Exercise 68 Let be a smooth projective toroidal fan in NQ . Show that for any
ray ρ there is a piecewise linear function ϕρ on which vanishes on all the other rays
and satisfies ϕ(u ρ ) = 1 where u ρ is the primitive ray generator in N of the ray ρ.
Let us write (1) for the set of rays of a fan and u ρ for the primitive generator
in N of the ray ρ ∈ (1).
Lemma 69 If is a smooth projective toroidal fan, then {ϕρ : ρ ∈ (1)} (where ϕρ
is defined in Exercise 68) forms a basis of R
1
the space of piecewise linear functions
on .
Exercise 70 Let be a smooth projective toroidal fan. Show that R is isomorphic
to the Stanley-Reisner algebra R , i.e., the quotient ring of Q[Tρ : ρ ∈ (1)] by the
k
relations i=1 Tρi = 0 whenever ρ1 , . . . , ρk are distinct rays which do not generate a
k
cone of . (Hint: Clearly i=1 ϕρi = 0 whenever ρ1 , . . . , ρk do not generate a cone
of . Therefore, there is a unique algebra homomorphism from R to R , which
sends Tρ to ϕρ . Show that this map is an isomorphism.)
We can now reformulate Theorem 67:
Proposition 71 Let X be a smooth projective toroidal horospherical variety
defined by an (uncolored) fan . Then the cohomology ring H ∗ (X , Q) is iso-
morphic as an H ∗ (G/P, Q)-algebra to the quotient of H ∗ (G/P, Q) ⊗ R by the
ideal
c1 (δ(m)) ⊗ 1 −
u ρ , m1 ⊗ ϕρ : m ∈ M =
ρ∈(1)
c1 (δ(m)) ⊗ 1 − 1 ⊗
·, m : m ∈ M ,
Proof The statement is a reformulation of Theorem 67 except the equality of the two
ideals which remains to be shown. Recall from Lemma 69 that the set of piecewise
linear functions {ϕρ : ρ ∈ (1)} (where ϕρ are defined in Exercise 68) forms a basis
of R1
. Then the (piecewise) linear function
·, m!
for m ∈ M can be expressed as a
linear combination in this basis, namely
·, m = ρ∈(1)
u ρ , mϕρ .
where the limit is taken over all smooth projective toroidal embeddings of G/H
which is a directed set. Indeed, for any two smooth projective toroidal embeddings
with corresponding fans 1 , 2 , we can find a third smooth projective toroidal fan
which refines both fans 1 and 2 . We introduce the relation whenever
refines . Suppose , so that we obtain an equivariant map X → X .
Our goal is to understand how the representation of cohomology rings given in
Proposition 71 behaves under this map. By Proposition 71, the cohomology rings
(as A-algebras) are generated by classes of divisors, so that the map corresponding
to X → X is given by pulling back divisors which in turn induces the natural inclu-
sion R ⊆ R . Let I :=
c1 (δ(m)) ⊗ 1 − 1 ⊗
·, m : m ∈ M ⊆ A ⊗ R . Simi-
larly define I in A ⊗ R . As I ⊆ I , we obtain the natural map μ, : (A ⊗
R )/I → (A ⊗ R )/I . Then ((A ⊗ R )/I , μ, ) is the direct system yield-
ing the direct limit lim H ∗ (X , Q). Moreover, we obtain two more direct systems,
−→
namely (I , I ⊆ I ) and (A ⊗ R , A ⊗ R ⊆ A ⊗ R ) (for ). Indeed,
we obtain a direct system of exact sequences:
0 → I → A ⊗ R → (A ⊗ R )/I → 0.
The statement follows by the fact that taking direct limits in the category of modules
is an exact functor, lim A ⊗ R = A ⊗ R, and lim I = I , where I denotes the ideal
−→ −
→
in the statement.
9 The Ring of Conditions for Horospherical Homogeneous Spaces 219
Acknowledgements We want to express our gratitude to Megumi Harada and Adam Van Tuyl
for organizing the exciting “Graduate Summer School in Algebraic Group Actions” at McMaster
University and their great support. We also want to thank the participants of the summer school for
their constructive feedback. Finally, we are grateful to the anonymous referee for several valuable
suggestions and remarks.
References
1. Akhiezer, D.N.: Actions with a finite number of orbits. Funktsional. Anal. i Prilozhen. 19(1),
1–5, 96 (1985)
2. Borel, A.: Linear algebraic groups. Graduate Texts in Mathematics, vol. 126, 2nd edn. Springer,
New York (1991)
3. Bravi, P., Pezzini, G.: Primitive wonderful varieties. Math. Z. 282(3–4), 1067–1096 (2016)
4. Brion, M.: Vers une généralisation des espaces symétriques. J. Algebra 134(1), 115–143 (1990)
5. Brion, M.: Piecewise polynomial functions, convex polytopes and enumerative geometry. In:
Parameter spaces (Warsaw, 1994), Banach Center Publications, vol. 36, pp. 25–44. Polish
Academy of Sciences Institute of Mathematics, Warsaw (1996)
6. Cupit-Foutou, S.: Wonderful varieties: a geometrical realization (2009). arXiv:0907.2852
7. De Concini, C., Procesi, C.: Complete symmetric varieties. II. Intersection theory. In: Algebraic
Groups and Related Topics (Kyoto/Nagoya, 1983), Advanced Studies in Pure Mathematics,
vol. 6, pp. 481–513. North-Holland, Amsterdam (1985)
8. Gagliardi, G., Hofscheier, J.: Gorenstein spherical Fano varieties. Geom. Dedicata 178, 111–
133 (2015)
9. Hartshorne, R.: Algebraic Geometry. Springer, New York, Heidelberg (1977). Graduate Texts
in Mathematics, No. 52
10. Humphreys, J.E.: Linear Algebraic Groups. Springer, New York, Heidelberg (1975). Graduate
Texts in Mathematics, No. 21
11. Kiritchenko, V.: Geometry of spherical varieties (2009). Course notes. https://users.mccme.ru/
valya/sph1.pdf
12. Kleiman, S.L.: The transversality of a general translate. Compositio Math. 28, 287–297 (1974)
13. Knop, F.: The Luna-Vust theory of spherical embeddings. In: Proceedings of the Hyderabad
Conference on Algebraic Groups (Hyderabad, 1989), pp. 225–249. Manoj Prakashan, Madras
(1991)
14. Knop, F.: Localization of spherical varieties. Algebra Number Theory 8(3), 703–728 (2014)
15. Losev, I.V.: Uniqueness property for spherical homogeneous spaces. Duke Math. J. 147(2),
315–343 (2009)
16. Luna, D.: Variétés sphériques de type A. Publ. Math. Inst. Hautes Études Sci. 94, 161–226
(2001)
17. Luna, D., Vust, T.: Plongements d’espaces homogènes. Comment. Math. Helv. 58(2), 186–245
(1983)
18. Malle, G., Testerman, D.: Linear algebraic groups and finite groups of Lie type. In: Cambridge
Studies in Advanced Mathematics, vol. 133. Cambridge University Press, Cambridge (2011)
19. Pasquier, B.: Fano horospherical varieties. Ph.D. thesis, Université Joseph-Fourier-Grenoble I
(2006)
20. Pasquier, B.: Variétés horosphériques de Fano. Bull. Soc. Math. France 136(2), 195–225 (2008)
21. Sankaran, P., Uma, V.: Cohomology of toric bundles. Comment. Math. Helv. 78(3), 540–554
(2003)
22. Schubert, H.: Kalkül der abzählenden Geometrie. Springer, Berlin-New York: Reprint of the
1879 original. With an introduction by Steven L, Kleiman (1979)
23. Shafarevich, I.R.: Basic algebraic geometry. 1, Russian edn. Springer, Heidelberg (2013).
Varieties in projective space
220 J. Hofscheier
24. Shafarevich, I.R.: Basic Algebraic Geometry. 2, 3rd edn. Springer, Heidelberg (2013). Schemes
and complex manifolds, Translated from the 2007 third Russian edition by Miles Reid
25. Springer, T.A.: Linear Algebraic Groups, 2nd edn. Modern Birkhäuser Classics. Birkhäuser
Boston Inc, Boston, MA (2009)
26. Timashev, D.A.: Homogeneous spaces and equivariant embeddings. Encyclopaedia of Mathe-
matical Sciences, vol. 138. Springer, Heidelberg (2011). Invariant Theory and Algebraic Trans-
formation Groups, 8
27. Vakil, R.: The rising sea: foundations of algebraic geometry (2017). Course notes.
http://math.stanford.edu/∼vakil/216blog/FOAGnov1817public.pdf
28. Wasserman, B.: Wonderful varieties of rank two. Transform. Groups 1(4), 375–403 (1996)
Chapter 10
Linear Recursions for Integer Point
Transforms
Katharina Jochemko
10.1 Introduction
where xm denotes x1m 1 · · · xnm n for all m ∈ Zn . In this note we study sequences
{σk P (x)}k≥0 of integer point transforms of integer dilates of polytopes P and rel-
atives. We prove the following linear recursion.
Theorem 1 Let Q be a polytope in Rn and let P be a lattice polytope with vertex
set vert(P) = {v1 , . . . , vr }. Then the sequence {σk P+Q (x)}k≥0 satisfies the linear
K. Jochemko (B)
Department of Mathematics, KTH Royal Institute of Technology, SE–100 44 Stockholm, Sweden
e-mail: [email protected]
© Springer Nature Switzerland AG 2022 221
A. M. Kasprzyk and B. Nill (eds.), Interactions with Lattice Polytopes,
Springer Proceedings in Mathematics & Statistics 386,
https://doi.org/10.1007/978-3-030-98327-7_10
222 K. Jochemko
recursion
σ(k+r )P+Q (x) = (−1)1+|I | x i∈I vi
σ(k+r −|I |)P+Q (x)
∅= I ⊆[r ]
In particular, the recursion only depends on the vertices of P. This improves results
by Alexandersson [2] where it was assumed that P has the integer decomposition
property and Q = {0}.
Employing classical results from valuation theory, in Sect. 10.2 we first prove a
recursion for indicator functions of dilated polytopes. Then, in Sect. 10.3, we apply
these results to integer point transforms and prove Theorem 1. We recover Brion’s
Theorem in Sect. 10.4 and by applying our results to Schur polynomials we disprove
a conjecture of Alexandersson [1] in Sect. 10.5.
In this section we prove a linear recursion for indicator functions of integer dilates of
a polytope P. Let P denote the set of polytopes in Rn and let G be an abelian group.
A valuation is a map ϕ : P → G such that ϕ(∅) = 0 and
for all P, Q ∈ P such that also P ∪ Q ∈ P. The volume, the number of lattice points
inside a polytope and the integer point transform are examples of valuations. It was
shown by Volland [19] that every valuation satisfies the inclusion-exclusion property.
That is, for polytopes P, P1 , . . . , Pr such that P = P1 ∪ · · · ∪ Pr
ϕ(P) = (−1)|I |+1 ϕ(PI ) ,
∅= I ⊆[r ]
PI := i∈I Pi . Stronger even, it follows from a result of Groemer
where [9], that
if αi 1 Pi = 0 for polytopes P1 , . . . , Pm and some α1 , . . . , αm ∈ Z then i αi ϕ(Pi )
= 0 where1 P denotes the indicator function for every polytope P. A function of
the form αi 1 Pi is called a polytopal simple function. By Groemer’s result, every
valuation uniquely defines a homomorphism from the abelian group of polytopal
simple functions to G, that is, every polytope can be identified with its indicator
10 Linear Recursions for Integer Point Transforms 223
function. For valuations on lattice polytopes this was proved by McMullen [12]. It
is well-known that for every affine linear map T : Rn → Rm
1 P → 1T (P) (10.1)
defines a valuation (see, e.g., [4, Chap. 8]). Using this push forward map we obtain
the following recursion on indicator functions.
Theorem 2 Let P be a polytope in Rn with vertex set vert(P) = {v1 , . . . , vr }. Then
1(k+r )P = (−1)1+|I | 1 Q k,r
I
∅= I ⊆[r ]
I = (k + r − |I |)P +
for all k ≥ 0 where Q k,r i∈I vi .
Proof We first assume that P is the (d − 1)-dimensional standard simplex d−1 =
{x ∈ Rd : x1 + · · · + xd = 1, x1 , . . . , xd ≥ 0}. Its (k + d)th dilate is given by
(k + d)d−1 = {x ∈ Rd : x1 + · · · + xd = d + k, x1 , . . . , xd ≥ 0}
Proof (2nd proof of Theorem 2) The proof follows from Theorem 3 by expanding
Eq. (10.2) and multiplying both sides with 1k P .
I = (k + r − |I |)P +
for all k ≥ 0 where Q k,r i∈I vi .
d
ak = c j ak− j
j=1
d all k ≥d−
for d. The corresponding characteristic polynomial χc is defined as X d −
j=1 c j X
j
∈ C(x1 , . . . , xn )[X ]. The polynomial χc is called minimal if for every
vector c = (c1 , . . . , cd ) corresponding to a linear recursion of a we have χc |χc .
where the last equation follows by observing that σ P+v (x) = xv σ P (x) for all v ∈ Zn .
We observe that χ P;Q is the characteristic polynomial of this linear recursion.
Now let Q be a lattice polytope and suppose that χ P;Q is not minimal. Then, for
some vertex u of P, {σk P+Q (x)}k≥0 satisfies a linear recursion with characteristic
polynomial v∈vert(P)\{u} (X − xv ). That is
10 Linear Recursions for Integer Point Transforms 225
|vert(P)|−1
σ(k+r )P+Q (x) + (−1) j e j ({xv : v ∈ vert(P) \ {u}})σ(k+r − j)P+Q (x) = 0
j=1
a linear functional such that (u) > ( p) for all p = u in P and (v) > (q) for
j
all q = v in Q. Then ((k + r − j) p + q + l=1 vl )=(k + r − j)( p) + (q) +
j
l=1 (vl ) < (k + r − j)(u)
+ (v) + j(u) = ((k + r )u + v) for all p ∈ P and q ∈ Q and the conclusion fol-
lows.
Every linear map f : Rn → Rl with the property that f (Zn ) ⊆ Zl induces an
algebra homomorphism
f¯ : C x1±1 , . . . , xn±1 → C x1±1 , . . . , xl±1
xm → x f (m)
The following two examples show that the minimality of a characteristic polyno-
mial is not necessarily preserved under affine transformations or taking Minkowski
sums.
In this section we provide a proof of Brion’s Theorem using the recursion given
in Theorem 1. For a polytope P ⊆ Rn and a vertex v of P the tangent cone Kv is
defined as {v + w : v + εw ∈ P for 0 < ε 1}. If the polytope P has rational edge
directions, in particular, if it is a lattice polytope, then the integer point transform
of Kv is a rational function.
Theorem 9 (Brion’s Theorem [6]) Let P be a lattice polytope with vertex set
vert(P). Then
σ P (x) = σKv (x)
v∈vert(P)
as rational functions.
The following is an immediate consequence of [5, Lemma 13.5.]
A further ingredient for our proof of Brion’s Theorem is the following well-known
result (see, e.g., [17, Chap. 5]).
Lemma 11 Let {an }n∈N be a sequence of elements of a field K that satisfy a linear
recursion of order d with characteristic polynomial
d
(X − ri ).
i=1
d
an = αi rin , for all n ∈ N.
i=1
10 Linear Recursions for Integer Point Transforms 227
for all k ≥ 0. Our goal is to show that cw · xw = σKw (x) as rational functions for
all w ∈ vert(P), or, equivalently, that cw equals the integer point transform of the
0 of the translated polytope P − w. Equation (10.3) is
tangent cone K̃0 of the vertex
equivalent to σk(P−w) (x) = v∈vert(P) cv xk(v−w) . As k goes to infinity σk(P−w) (x) con-
verges absolutely to σK̃0 (x) on WK̃0 = {x ∈ Cn : |xv−w | < 1 for all
v ∈ vert(P) \ {w}} by Lemma 10. On the other hand, v∈vert(P) cv xk(v−w) converges
to cw . Thus σK̃0 (x) and cw coincide on WK̃0 and are therefore the same as rational
functions.
χ (X ) = X − xw(T ) .
T ∈Tλn /μμ
∞
Then, for sufficiently large r , {sκ+lμ/λ+lν (x)}l=r satisfies a linear recursion with min-
imal polynomial
χ (X ) = X − xw .
w∈W
for all l ≥ r . The claim now follows from Proposition 5 since the weight function w
is linear.
Since typically there are more lattice points in GLλ /μμ than vertices, Corollary 14
shows that the characteristic polynomial given in Theorem 12 is in general not mini-
mal. The next example shows that also the polynomial given in Conjecture 13 is not
minimal in general, thus refuting it.
Example 15 Let n = 3, λ = (5, 3, 1) and μ = (3, 0, 0). Consider the skew Young
tableau T and its corresponding Gelfand–Tsetlin pattern p depicted in Fig. 10.2.
Then
w(T ) = w( p) = (4, 2, 0) (3, 2, 1) = λ − μ .
2 3 1 3 5
0 1 4
2 3 3 0 0 3
0 0 3
3
Fig. 10.2 The skew Young tableau T in Example 15 and its corresponding Gelfand–Tsetlin pat-
tern p
230 K. Jochemko
From the face structure studied in [10, 13] it follows that the coordinates of any
vertex of GLλ /μμ are in the set {0, 1, 3, 5}. Let x = {xi, j } be a Gelfand–Tsetlin pattern
that is a vertex of GLλ/μμ . Then x4,1 , x4,2 and x3,1 are 0. Furthermore, x2,1 ∈ {0, 1}.
If x2,1 = 0, then the sum of entries of the first row of x is odd and the sum of entries
of the second is even, therefore w(x)1 is odd and w(x) = (4, 2, 0). On the other hand,
if x2,1 = 1, then x3,2 ∈ {1, 3} and in that case w(x)2 is odd and again w(x) = (4, 2, 0).
In summary, (4, 2, 0) ∈ W is not the weight of a vertex of GLλ /μμ and therefore
X − xw X − xw(v) .
w∈W v∈V
Therefore, by Corollary 14, w∈W (X − xw ) cannot be the minimal polynomial.
Remark 16 To verify the counterexample given in Example 15, one can also use [2,
Proposition 6].
Acknowledgements The author would like to thank Per Alexandersson and Raman Sanyal for
inspiring and fruitful discussions and many helpful comments. The author was partially supported
by a Hilda Geiringer Scholarship at the Berlin Mathematical School, the Knut and Alice Wallen-
berg Foundation and a Microsoft Research Fellowship of the Simons Institute for the Theory of
Computing.
References
1. Alexandersson, P.: Stretched skew Schur polynomials are recurrent. J. Combin. Theory Ser. A
122, 1–8 (2014)
2. Alexandersson, P.: Polynomials defined by tableaux and linear recurrences. Electron. J. Com-
bin. 23(1), Paper 1.47, 24 (2016)
3. Ardila, F., Bliem, T., Salazar, D.: Gelfand-Tsetlin polytopes and Feigin-Fourier-Littelmann-
Vinberg polytopes as marked poset polytopes. J. Combin. Theory Ser. A 118(8), 2454–2462
(2011)
4. Barvinok, A.: A course in convexity, Graduate Studies in Mathematics, vol. 54. American
Mathematical Society, Providence (2002)
5. Barvinok, A.: Integer points in polyhedra. Zurich Lectures in Advanced Mathematics. European
Mathematical Society (EMS), Zürich (2008)
6. Brion, M.: Points entiers dans les polyèdres convexes. Ann. Sci. École Norm. Sup. (4) 21(4),
653–663 (1988)
7. Ehrhart, E.: Sur les polyèdres rationnels homothétiques à n dimensions. C. R. Acad. Sci. Paris
254, 616–618 (1962)
8. Fang, X., Fourier, G., Pegel, C.: The Minkowski property and reflexivity of marked poset
polytopes. Electron. J. Combin. 27(1), Paper No. 1.27, 19 (2020)
9. Groemer, H.: On the extension of additive functionals on classes of convex sets. Pacific J. Math.
75(2), 397–410 (1978)
10. Jochemko, K., Sanyal, R.: Arithmetic of marked order polytopes, monotone triangle reciprocity,
and partial colorings. SIAM J. Discrete Math. 28(3), 1540–1558 (2014)
11. Lawrence, J.: Three rings of polyhedral simple functions. J. Res. Natl. Inst. Stand. Technol.
111(2), 127–134 (2006)
12. McMullen, P.: Valuations on lattice polytopes. Adv. Math. 220(1), 303–323 (2009)
10 Linear Recursions for Integer Point Transforms 231
13. Pegel, C.: The face structure and geometry of marked order polyhedra. Order 35(3), 467–488
(2018)
14. Pukhlikov, A.V., Khovanskiı̆, A.G.: Finitely additive measures of virtual polyhedra. Algebra i
Analiz 4(2), 161–185 (1992)
15. Sam, S.V.: A bijective proof for a theorem of Ehrhart. Amer. Math. Monthly 116(8), 688–701
(2009)
16. Schneider, R.: Convex bodies: the Brunn-Minkowski theory. Encyclopedia of Mathematics and
its Applications, vol. 151, expanded edn. Cambridge University Press, Cambridge (2014)
17. Spiegel, M.R.: Calculus of Finite Differences and Difference Equations. Schaum’s Outline
Series. McGraw-Hill, New York (1971)
18. Stanley, R.P.: Enumerative combinatorics, vol. 2. Cambridge Studies in Advanced Mathematics,
vol. 62. Cambridge University Press, Cambridge (1999). With a foreword by Gian-Carlo Rota
and appendix 1 by Sergey Fomin
19. Volland, W.: Ein Fortsetzungssatz für additive Eipolyederfunktionale im euklidischen Raum.
Arch. Math. (Basel) 8, 144–149 (1957)
Chapter 11
Schubert Calculus on Newton–Okounkov
Polytopes
11.1 Introduction
that has Poincaré duality. The polytope rings were originally used to give a convenient
functorial description of the cohomology rings of smooth toric varieties. In this
case, P is always a simple lattice polytope, that is, all vertices of P belong to Zd ⊂ Rd ,
V. Kiritchenko (B)
Laboratory of Algebraic Geometry and Faculty of Mathematics, National Research University
Higher School of Economics, Usacheva str. 6, 119048 Moscow, Russia
Institute for Information Transmission Problems, Moscow, Russia
e-mail: [email protected]
M. Padalko
SimilarWeb, Tel Aviv, Israel
and only d edges meet at every vertex of P. In [10], Kaveh noted that polytope
rings can also be used for a partial description of the cohomology rings of spherical
varieties. In this case, P is still a lattice polytope but not necessarily simple.
For simple polytopes, every face ⊂ P can be naturally identified with an ele-
ment x ∈ R P so that
x x = x∩
for any two transverse faces and . This is no longer true for non-simple polytopes,
that is, individual faces of P do not have natural counterparts in R P . However, it is
still possible to identify every element of R P with a linear combination of faces
of P so that the product in the polytope ring corresponds to the intersection of faces.
In [14], the first author, Smirnov and Timorin developed a general framework for
such calculus on polytopes, and studied its applications to Schubert calculus on
Gelfand–Zetlin polytopes in type A. In this paper, we mainly consider applications
to Schubert calculus in types B and C.
Representation theory of classical groups is a source of several interesting fami-
lies of lattice convex polytopes. For S L n (C) (type A), there is a well-known family
of Gelfand–Zetlin (GZ) polytopes G Z λ . Here λ := (λ1 , . . . , λn ) ∈ Zn runs through
dominant weights of S L n (C), that is, λ1 ≥ λ2 ≥ · · · ≥ λn . Originally, GZ polytopes
were constructed using representation theory, namely, lattice points in the polytope
G Z λ parameterize the vectors in a special basis in the irreducible representation Vλ
of S L n (C) with the highest weight λ (see [21] for a survey on GZ bases). In con-
vex geometric terms, the GZ polytope G Z λ ⊂ Rd , where d := n(n−1) 2
, is defined as
the set of all points (z 11 , z 21 , . . . , z n−1
1
; z 12 , . . . , z n−2
2
; . . . ; z 1n−1 ) ∈ Rd that satisfy the
following interlacing inequalities:
λ1 λ2 λ3 ... λn
z 11 z 21 ... 1
z n−1
z 12 ... 2
z n−2
.. .. (G Z A )
. .
z 1n−2 z 2n−2
z 1n−1
11.2 Preliminaries
In this section, we recall the definitions of polytope rings, GZ polytopes and flag
varieties in types B and C. We discuss the relationship between the polytope rings
of GZ polytopes and cohomology rings of flag varieties. We also define Newton–
Okounkov polytopes of flag varieties.
P1 + P2 = {x1 + x2 ∈ Rd | x1 ∈ P1 , x2 ∈ P2 }
11 Schubert Calculus on Newton–Okounkov Polytopes 237
It is not hard to check that this semigroup has cancelation property. We can also
multiply polytopes in S P by positive real numbers using dilation:
λP = {λx | x ∈ P}, λ ≥ 0.
Hence, we can embed the semigroup of convex polytopes into its Grothendieck
group V P , which is a real vector space. The elements of V P are called virtual polytopes
analogous to P.
On the vector space V P , there is a homogeneous polynomial vol P of degree d,
called the volume polynomial. It is uniquely characterized by the property that its
value vol P (Q) on any convex polytope Q ∈ S P is equal to the volume of Q.
Let P be a lattice in V P generated by some lattice polytopes (with respect to L)
analogous to P (we do not assume that P contains all lattice polytopes analogous
to P). The symmetric algebra Sym( P ) of P can be thought of as the ring of differ-
ential operators with constant integer coefficients acting on R[V P ], the space of all
polynomials on V P . If D ∈ Sym( P ) and ϕ ∈ R[V P ], then we write Dϕ ∈ R[V P ] for
the result of this action. Define A P as the homogeneous ideal in Sym( P ) consisting
of all differential operators D such that D vol P = 0. Set R P = Sym( P )/A P . This
ring is called the polytope ring associated with the polytope P and the lattice P .
λ1 λ2 λ3 ... λn 0
x11 x21 ... xn1
y11 y21 . . . yn−1
1
0
x12 ... 2
xn−1
y12 . . . yn−2
2
0 G ZC
.. .. ..
. . .
x1n−1 x2n−1
n−1
y1 0
x1n
Again, every coordinate in this table is bounded from above by its upper left neighbor
and bounded from below by its upper right neighbor (the table encodes 2d inequal-
ities). We regard SG Z λ as a lattice polytope with respect to the standard lattice Zd .
Roughly speaking, SG Z λ is the polytope defined using half of the GZ pattern (G Z A )
for S L 2n (C).
1
vol SG Z (λ1 , λ2 ) = λ1 λ2 (λ1 − λ2 )(λ1 + λ2 ).
6
This volume times 4! is equal to the degree degλ (Sp4 (C)/B) of the isotropic flag
variety.
Remark 4 Family of odd orthogonal GZ polytopes (as defined in [1, 20]) consists of
two subfamilies parameterized by integer and half-integer λ. The group S O2n+1 (C)
is not simply connected, and half-integer weights correspond to the characters of
the maximal torus in the universal cover Spin(2n + 1). If we define the polytope
ring R SG Z using the first subfamily we get a subring of H ∗ (S O2n+1 /B, Z) generated
by the first Chern classes of line bundles L λ corresponding to the characters λ of the
maximal torus in S O2n+1 (C).
Example 5 The polytope OG Z λ ⊂ R4 for Sp4 (C) is given by the same 8 inequali-
ties as in Example 3. However, its volume polynomial is computed using a different
volume form chosen so that the covolume of L B is 1. Since Z4 ⊂ L B has index 4,
we get vol OG Z = 4 vol SG Z .
There is an exceptional isomorphism Sp4 (C)/±1 S O5 (C). In particular, flag
varieties in types B2 and C2 are the same. This isomorphism takes the dominant
weight λ = (λ1 , λ2 ) of Sp4 (C) to the dominant weight λ = (λ1 + λ2 )/2, (λ1 −
λ2 )/2) of S O5 (C). This agrees with the identity vol(SG Z λ ) = vol(OG Zλ ).
Example 6 In type A, we can identify the open Schubert cell C with an affine
space Cd (for d = n(n − 1)/2) by choosing for every flag M • a basis v1 ,…, vn in Cn
of the form:
v1 = en + x1n−1 en−1 + · · · + x11 e1 ,
λ = (1, . . . , 1, 0 . . . , 0),
k n−k
then Vλ∗ can be identified with the subspace of C(x ij )i+ j<n spanned by the minors of
the n × k matrix formed by the first k columns of the matrix (F F L V ). These minors
are exactly the Plücker coordinates of the Grassmannian G(k, n) in the Plücker
embedding. The map X → H 0 (L λ , X )∗ is the composition of the projection X →
11 Schubert Calculus on Newton–Okounkov Polytopes 241
G(k, n) (obtained by forgetting all subspaces in the flag M • except for the V k ) and
the Plücker embedding of G(k, n).
To assign the Newton–Okounkov convex body to V we need an extra ingredient.
Choose a translation-invariant total order on the lattice Zd (e.g., we can take the
lexicographic order). Consider a map
v : C(x1 , . . . , xd ) \ {0} → Zd ,
By V k we denote the subspace spanned by the kth powers of the functions from V .
Example 9 Using coordinates of Example 6 we can define the valuation v as fol-
lows. Order the coefficients (x ij )i+ j<n of the matrix (F F L V ) by starting from col-
umn (n − 1) and going from top to bottom in every column and from right to left
along columns. Then v (X, Vλ∗ ) coincides with the Feigin–Fourier–Littelmann–
Vinberg polytope F F L V (λ) [17]. Moreover, the inclusion F F L V (λ) ⊂ v (X, Vλ∗ )
follows from a straightforward computation of the valuation v on the minors of the
matrix (F F L V ) (see [17, Example 2.9] for more details).
Different valuations might yield different Newton–Okounkov convex bodies. In
particular, GZ polytopes can also be obtained as Newton–Okounkov polytopes of
flag varieties [6, 11]. Okounkov made the first explicit computation of this kind,
namely, he exhibited symplectic GZ polytopes as Newton–Okounkov polytopes of
the isotropic flag varieties [22].
Example 10 Figure 11.2 illustrates the idea of mitosis in the simplest example. The
trapezoid and rectangle on the left picture have the same number of lattice points
with given sum of coordinates. The same is true for the right picture. However, the
trapezoid on the right picture becomes a virtual polytope (in particular, lattice points
marked with circles have to be counted with the zero coefficient) while the rectangle
remains a true polytope. There is a price to pay: the left vertical edge of the trapezoid
corresponds to two edges of the rectangle (that is, a single edge of the trapezoid has
the same number of lattice points as two edges of the rectangle). In short, mitosis
preserves positivity at the cost of more involved combinatorics.
Consider a vector space with the direct sum decomposition
Rd = Rd1 ⊕ · · · ⊕ Rdr ,
and choose coordinates x = (x11 , . . . , xd11 ; . . . ; x1r , . . . , xdrr ) with respect to this
decomposition. Let C ⊂ Rd be a convex polyhedral cone with the vertex at the ori-
gin 0. Assume that C is given by inequalities either of type x ij ≤ ax ij where a > 0
and i = i or of type 0 ≤ x ij . In what follows, we use the bijective correspondence
between facets of C and inequalities, namely, every inequality x ij ≤ ax ij defines the
facet H (i, j; i , j ) given by the equation x ij = ax ij , and every inequality 0 ≤ x ij
defines the facet H (0, 0; i, j) given by the equation x ij = 0.
In addition, assume that C does not contain any rays parallel to the x ij -axis
unless j = 1. Then the geometric mitosis of [15, Sect. 5.1] can be defined on faces
of C. Below we describe the resulting mitosis operations M1 ,…, Mr from a combi-
natorial viewpoint.
Let be a face of the cone C of codimension . The ith mitosis operation Mi
applied to will produce a collection Mi () (possibly empty) of faces of C. Choose
a minimal subset of facets H1 ,…, H of C such that = H1 ∩ · · · ∩ H . If none
of these facets coincides with H ( p, q; i, di ) for some p and q, then Mi () = ∅.
Otherwise, let s be the smallest number such that the subset {H1 , . . . , H } contains
facets of type H (·, ·; i, j) for all j = s, s + 1, …, di . For brevity, we label these
facets by H + (i, s), H + (i, s + 1), …, H + (i, di ). For every j = s + 1, s + 2,…, di ,
we now label by H+ (i, j) the facet of type H (i, j; ·, ·). If there are two such
11 Schubert Calculus on Newton–Okounkov Polytopes 243
p
facets H (i, j; p, q) and H (i, j; p , q ), and xq ≤ xq everywhere on then we
p
where the set {H1 ( j), . . . , H −1 ( j)} is obtained from the set {H1 , . . . , H } by the
following rule. First, remove the facet H + (i, j). Second, for every k ∈ Ji () such
that k > j replace the facet H + (i, k) by the facet H+ (i, k). Note that dim j =
dim + 1.
Definition 11 The ith mitosis operation Mi sends to
Mi () = { j | j ∈ Ji ()}.
Let C be the vertex cone of the GZ polytope in type A for the vertex a =
(λ2 , . . . , λn ; λ3 , . . . , λn ; . . . ; λn ) (see table (G Z A )). After an affine change of coor-
dinates x = z − a the inequalities that define C can be written as follows:
{0} = + + −→ + −→ + −→
M1 M2 M1
=C
+ +
244 V. Kiritchenko and M. Padalko
⎧ ⎫
⎨ ⎬
{0} −→ + + −→ +
M2 M1 M2
, −→ C
⎩ +⎭
For arbitrary n, the mitosis operations M1 ,…, Mn−1 encoded by tables coincide
with Knutson–Miller mitosis on pipe dreams [18] after reflecting tables in a vertical
line. Instead of the vertex cone C we could take the GZ polytope in type A and
consider mitosis operations on faces that contain the vertex a (so called Kogan
faces). Geometric meaning of the resulting collections of faces is described in [14,
Theorem 5.1, Corollary 5.3]. In particular, the following analog of Kushnirenko’s
theorem holds.
Recall that Schubert subvarieties X w are labeled by the elements of the Weyl
group of G, namely, X w is the closure of the B-orbit Bw B/B, where w is an element
of the Weyl group of G. The Weyl group of G = S L n (C) is the symmetric group Sn .
By s1 ,…, sn−1 we denote the elementary transpositions.
Theorem 12 ([14, Theorem 5.4]) Let X w ⊂ S L n (C)/B be the Schubert subvariety
corresponding to a permutation w ∈ Sn . Let w = s j1 . . . s j be a reduced decompo-
sition of a permutation w ∈ Sn such that ( j1 , . . . , j ) is a subword of (1; 2, 1; 3, 2,
1; . . . ; n − 1, . . . , 1). Let Sw ⊂ G Z λ be the set of all faces produced from the ver-
tex a ∈ G Z λ by applying successively the operations Mn− j ,…, Mn− j1 :
Then
degλ (X w ) = ! Vol().
∈Sw
This implies that the Schubert cycle [X w ] (that is, the cohomology class of X w
in H ∗ (S L n (C)/B, Z)) in the polytope ring RG Z H ∗ (S L n (C)/B, Z) is represented
by the sum of faces in Sw .
Fig. 11.3 Faces M1 (a), M2 (a), M2 M1 (a), M1 M2 (a), M1 M2 M1 (a) of the GZ polytope in
type A, n = 3
x21
x21 ≤ x12 + λ2 , 0 ≤ x22 ≤ λ2 , x22 ≤
2
(these polytopes can also be realized as Newton–Okounkov polytopes of the isotropic
flag variety Sp4 /B [16, Proposition 4.1]). The vertex cone C of the vertex 0 is given by
4 homogeneous inequalities: 0 ≤ x11 , 0 ≤ 2x22 ≤ x21 ≤ 2x12 . It is convenient to encode
a face of C by a (2n − 1) × n table (skew pipe dream) for n = 2 filled with +
as follows (see Sect. 11.3.3 for the general definition of skew pipe dreams). The
table contains + in cell (3 − i, i) (for i = 1, 2) iff ⊂ H (0, 0; i, i), + in cell (2, 2)
iff ⊂ H (2, 2; 1, 2) and + in cell (3, 2) iff ⊂ H (1, 2; 2, 1). There are two mitosis
operations M1 and M2 .
+ + + +
M1 M2 M1 M2
{0} = + + −→ + −→ + −→ −→ =C
+ +
⎧ ⎫
+ ⎨ + +⎬
M2 M1 M2
{0} −→ + + −→ , + −→
⎩ ⎭
+
⎧ ⎫
⎨ ⎬
M2 M1
−→ + , , + −→ C
⎩ ⎭
+
Sw = M j1 · · · M j (0).
Then
degλ (X w ) = ! Vol().
∈Sw
This example can be extended to DDO polytopes for Sp2n . For n = 3 and the
DDO polytope for (s3 s2 s1 )3 (where s3 is the simple reflection with respect to the
longer root) this was done in [23]. The corresponding family of DDO polytopes
in R9 = R3 ⊕ R3 ⊕ R3 is given by inequalities:
1 2 1 1 1 2
0 ≤ x11 ; 0 ≤ x21 ≤ x12 ; 0 ≤ x33 ≤ x ≤ x3 ≤ x2 ≤ x13 ;
2 3 2 2
1 2 1
x3 ≤ x23 ≤ x22 .
2 2
An analog of Proposition 14 follows easily from [16, Corollary 3.6]. However, com-
binatorics of mitosis becomes more involved as analogs of pipe dreams in this case
have a loop.
Recently, Fujita identified DDO polytopes with certain Nakashima–Zelevinsky
polyhedral realizations of crystal bases [5, Theorem 4.1]. In particular, there are
explicit inequalities for these polytopes in types A, B, C, D and G 2 [5, Example 4.3].
In type A, they coincide with the GZ polytope and in type C2−3 with the polytopes
described in this section. It would be interesting to apply geometric mitosis to these
polytopes in the other cases.
(sn sn−1 . . . s2 s1 s2 . . . sn−1 sn ) . . . (s2 s1 s2 )(s1 ) of the longest element in the Weyl group
(here s1 corresponds to the longer root). The corresponding string polytope coincides
with the symplectic GZ polytope after a unimodular change of coordinates [20, Sect.
6]. The cone C is simplicial and is given by d = n 2 inequalities:
for all i = 1,…, n. We define symplectic mitosis as the geometric mitosis associ-
ated with the cone C. Combinatorics of the symplectic mitosis is quite simple and
described in detail in [16, Sect. 5.2] using skew pipe dreams. However, arguments
of [16, Corollary 3.6] do not directly yield presentations for Schubert cycles since
the symplectic GZ polytope does not satisfy the necessary conditions. Still computa-
tions for n = 2, 3 suggest that the collections of faces of the symplectic GZ polytope
obtained using symplectic mitosis do represent the corresponding Schubert cycles
in the polytope ring R SG Z . Below we describe a bijection between faces of C and
faces of SG Z λ that we used.
Let v be the vertex of SG Z λ given by equations λs = x ij = ylk for all triples λs , x ij
and ylk such that s = i + j − 1 = k + l. We now define a bijection between those
facets of Pλ that contain v and skew pipe dreams of size n with exactly one +. Recall
that a skew pipe dream of size n is a (2n − 1) × n table whose cells are either empty or
filled with +. Only cells (i, j) with n − j < i < n + j are allowed to have + (see [16,
Sect. 5.2] for more details on skew pipe dreams). Put yi0 := λi for i = 1,…, n. The
facet given by equation x ij = y i−1 j corresponds to the skew pipe dream with + in
cell (i + j − 1, n − i + 1). The facet given by equation y ij = x ij+1 corresponds to
the skew pipe dream with + in cell (2n − i − j + 1, n − i + 1). In what follows,
we denote by H(k,l) the facet whose skew pipe dream under this correspondence
contains + in cell (k, l).
This correspondence between facets and skew pipe dreams with a single + extends
to all faces of the symplectic GZ polytope that contain the vertex v. Namely, the
face Hk1 ,l1 ∩ · · · ∩ Hks ,ls obtained as the intersection of s facets corresponds to the
skew pipe dream that has + precisely in cells (k1 , l1 ),…, (ks , ls ). In particular, the
vertex v corresponds to the skew pipe dream D0 that has + in all (fillable) cells. In
what follows, we denote by FD the face corresponding to a skew pipe dream D.
We now formulate a conjecture. Let w be an element of the Weyl group
of G = Sp2n . Choose a reduced decomposition w = s j1 . . . s j such that it is a sub-
word of (sn sn−1 . . . s2 s1 s2 . . . sn−1 sn ) . . . (s2 s1 s2 )(s1 ).
This conjecture is verified in the case n = 2 and for certain w in the case n = 3 [24].
Note that the bijection between faces of SG Z λ that contain the vertex v and faces
of the string cone C does not come from the unimodular change of coordinates that
identifies the string polytope and the symplectic GZ polytope. There are might be
piecewise linear transformations (such as the ones used in [17, Sect. 5.2]) that yield
scissor congruence of unions of faces of SG Z λ and faces of another polytope for
which geometric meaning of symplectic mitosis is more transparent.
Note that the Weyl groups of Sp2n (C) and S O2n+1 (C) are the same. Since the GZ
polytopes for both groups differ only by lattices symplectic mitosis is also a natu-
ral tool for finding presentations of Schubert cycles by faces of OG Z λ in type B.
However, coefficients will be rational rather than integer (with powers of 2 in denom-
inator) because the torsion index of S O2n+1 (C) is a power of 2. Note also that the
volumes of faces of both SG Z λ and OG Z λ should be computed with respect to their
lattices. The difference is already visible in the case n = 2 (see Example 5).
Acknowledgements The study has been partially funded by the Russian Academic Excellence
Project ‘5–100’.
References
1. Berenstein, A.D., Zelevinsky, A.V.: Tensor product multiplicities and convex polytopes in
partition space. J. Geom. Phys. 5(3), 453–472 (1988)
2. Brion, M.: Lectures on the geometry of flag varieties. In: Topics in Cohomological Studies of
Algebraic Varieties, Trends in Mathematics, pp. 33–85. Birkhäuser, Basel (2005)
3. Fang, X., Fourier, G., Littelmann, P.: Essential bases and toric degenerations arising from
birational sequences. Adv. Math. 312, 107–149 (2017)
4. Feigin, E., Fourier, G., Littelmann, P.: Favourable modules: filtrations, polytopes, Newton-
Okounkov bodies and flat degenerations. Transform. Groups 22(2), 321–352 (2017)
5. Fujita, N.: Polyhedral realizations of crystal bases and convex-geometric Demazure operators.
Selecta Math. (N.S.) 25(5), Paper No. 74 (2019)
6. Fujita, N., Oya, H.: A comparison of Newton-Okounkov polytopes of Schubert varieties. J.
Lond. Math. Soc. (2) 96(1), 201–227 (2017)
7. Harada, M., Yang, J.J.: Newton-Okounkov bodies of Bott-Samelson varieties and Grossberg-
Karshon twisted cubes. Michigan Math. J. 65(2), 413–440 (2016)
8. Harada, M., Yang, J.J.: Singular string polytopes and functorial resolutions from Newton-
Okounkov bodies. Illinois J. Math. 62(1–4), 271–292 (2018)
9. Ilyukhina, M.: Schubert Calculus and Geometry of a String Polytope for the Group sp4 . Moscow
State University, Diploma (2012)
10. Kaveh, K.: Note on cohomology rings of spherical varieties and volume polynomial. J. Lie
Theory 21(2), 263–283 (2011)
11. Kaveh, K.: Crystal bases and Newton-Okounkov bodies. Duke Math. J. 164(13), 2461–2506
(2015)
11 Schubert Calculus on Newton–Okounkov Polytopes 249
12. Kaveh, K., Khovanskii, A.G.: Newton-Okounkov bodies, semigroups of integral points, graded
algebras and intersection theory. Ann. of Math. (2) 176(2), 925–978 (2012)
13. Kaveh, K., Villella, E.: On a notion of anticanonical class for families of convex polytopes
(2018). arXiv:1802.06674 [math.AG]
14. Kirichenko, V.A., Smirnov, E.Y., Timorin, V.A.: Schubert calculus and Gelfand-Tsetlin poly-
topes. Uspekhi Mat. Nauk 67(4(406)), 89–128 (2012)
15. Kiritchenko, V.: Divided difference operators on polytopes. In: Schubert calculus—Osaka
2012, Advanced Studies in Pure Mathematics, vol. 71, pp. 161–184. Mathematics Society
Japan, [Tokyo] (2016)
16. Kiritchenko, V.: Geometric mitosis. Math. Res. Lett. 23(4), 1071–1097 (2016)
17. Kiritchenko, V.: Newton-Okounkov polytopes of flag varieties. Transform. Groups 22(2), 387–
402 (2017)
18. Knutson, A., Miller, E.: Gröbner geometry of Schubert polynomials. Ann. of Math. (2) 161(3),
1245–1318 (2005)
19. Lazarsfeld, R., Mustaţă, M.: Convex bodies associated to linear series. Ann. Sci. Éc. Norm.
Supér. (4) 42(5), 783–835 (2009)
20. Littelmann, P.: Cones, crystals, and patterns. Transform. Groups 3(2), 145–179 (1998)
21. Molev, A.I.: Gelfand-Tsetlin bases for classical Lie algebras. In: Handbook of Algebra. vol. 4,
Handbook of Algebra, vol. 4, pp. 109–170. Elsevier/North-Holland, Amsterdam (2006)
22. Okounkov, A.: Multiplicities and Newton Polytopes. In: Kirillov’s seminar on representation
theory, American Mathematical Society Translations: Series 2, vol. 181, pp. 231–244. American
Mathematical Society, Providence, RI (1998)
23. Padalko, M.: Mitosis for symplectic flag varieties. B.Sc. Thesis, National Research University
Higher School of Economics (2015)
24. Padalko, M.: Schubert calculus and symplectic Gelfand–Zetlin polytopes. M.Sc. Thesis,
National Research University Higher School of Economics (2017)
25. Pukhlikov, A.V., Khovanskiı̆, A.G.: The Riemann-Roch theorem for integrals and sums of
quasipolynomials on virtual polytopes. Algebra i Analiz 4(4), 188–216 (1992)
26. Totaro, B.: The torsion index of the spin groups. Duke Math. J. 129(2), 249–290 (2005)
Chapter 12
An Eisenbud–Goto-Type Upper Bound
for the Castelnuovo–Mumford Regularity
of Fake Weighted Projective Spaces
Bach Le Tran
Abstract We will give an upper bound for the k-normality of very ample lattice
simplices, and then give an Eisenbud–Goto-type bound for some special classes of
projective toric varieties.
12.1 Introduction
The study of the Castelnuovo–Mumford regularity for projective varieties has been
greatly motivated by the Eisenbud–Goto conjecture [7] which asks for any irreducible
and reduced variety X , is it always the case that
The Eisenbud–Goto conjecture is known to be true for some particular cases. For
example, it holds for smooth surfaces in characteristic zero [13], connected reduced
curves [8], etc. Inspired by the conjecture, there are also many attempts to give an
upper bound for the Castelnuovo–Mumford regularity for various types of algebraic
and geometric structures [5, 12, 15, 20].
For toric geometry, suppose that (X, L) is a polarized projective toric varieties such
that L is very ample. Then there is a corresponding
very ample lattice polytope P :=
PL associated to L such that (X, L) = m∈P∩M C · χ m [4, Sect. 5.4]. Therefore,
by studying the k-normality of P (cf. Definition 2), we can obtain the k-normality
and also the regularity of the original variety (X, L). For the purpose of this article,
B. L. Tran (B)
School of Mathematics, University of Edinburgh, James Clerk Maxwell Building, Peter Guthrie
Tait Road, Edinburgh EH9 3FD, UK
e-mail: [email protected]
we will focus on the case that X is a fake weighted projective d-space and PL a d-
simplex.
For any fake weighted projective d-space X embedded in Pr via a very ample line
bundle, Ogata [17] gives an upper bound for its k-normality:
dim X
k X ≤ dim X + − 1.
2
In this article, we will improve Ogata’s bound by giving a new upper bound for
the k-normality of very ample lattice simplices and show that
dim X
reg(X ) ≤ deg(X ) − codim(X ) + . (12.1)
2
Hi (X, F (k − i)) = 0
for all i > 0. The regularity of F , denoted by reg(F ), is the minimum number k such
that F is k-regular. We say that X is k-regular if the ideal sheaf I X of X is k-regular
and use reg(X ) to denote the regularity of X (or of I X ).
As the main object of the article is to find an upper bound for k-normality of very
ample lattice simplices, it is important for us to revisit the definition of k-normality
of lattice polytopes.
Definition 2 A lattice polytope P is k-normal if the map
12 An Eisenbud–Goto-Type Upper Bound for the Castelnuovo–Mumford … 253
P ∩ M + ·
· · + P ∩ M → k P ∩ M
k times
From this, we obtain a combinatorial form of the Eisenbud–Goto conjecture: for very
ample lattice polytope P ⊂ MR , is it always true that
max{deg(P), k P } ≤ Vol(P) − |P ∩ M| + d + 1?
The first inequality was confirmed to be true recently [11, Proposition 2.2]; namely,
Therefore, in order to verify the Eisenbud–Goto conjecture for the polarized toric
variety (X, L), it suffices to check if
k P ≤ Vol(P) − |P ∩ M| + d + 1. (12.4)
We now recall some basic facts about Ehrhart theory of polytopes and the definition
of their degree.
Let P be a lattice polytope of dimension d. We define ehr P (k) = |k P ∩ M|, the
number of lattice points in k P. Then from Ehrhart’s theory [6, 19],
∞
h ∗P (t)
Ehr P (t) = ehr P (k)t k =
k=0
(1 − t)d+1
for some polynomial h ∗P ∈ Z≥0 [t] of degree less than or equal to d. Let h ∗P (t) =
d ∗ i
i=0 h i t . We have
254 B. Le Tran
d
h ∗0 = 1, h ∗1 = |P ∩ M| − d − 1, h ∗d = |P 0 ∩ M|, and h i∗ = Vol(P).
i=0
The following lemma by Ogata is crucial to the main result of this article:
Lemma 4 ([17, Lemma 2.1]) Let P = conv(v0 , . . . , vd ) be a very ample lattice n-
simplex. Suppose that k ≥ 1 is an integer and x ∈ k P ∩ M. For any i = 0, . . . , d,
we have
2k−1
x + (k − 1)vi = uj
j=1
for some u j ∈ P ∩ M.
Using the ideas in [17, Lemma 2.5], we generalize the above lemma as follows.
Lemma 5 Suppose that P = conv(v0 , . . . , vd ) is a very ample
d-simplex. Let k ∈
d
N≥1 . Then for any x ∈ k P ∩ M, a0 , . . . , ad ∈ Z≥0 such that i=0 ai = k − 1, we
have
d
2k−1
ai vi + x = ui
i=0 i=1
for some u i ∈ P ∩ M.
Proof We will use induction in this proof. The case k = 1 is trivial. Suppose that
the lemma holds for k = s − 1. We will now
show that it holds for k = s; i.e., for
d
any x ∈ s P ∩ M, a1 , . . . , ad ∈ Z≥0 such that i=0 ai = s − 1, we have
d
2s−1
ai vi + x = ui (12.5)
i=0 i=1
2s−1
(s − 1)v0 + x = wi
i=1
12 An Eisenbud–Goto-Type Upper Bound for the Castelnuovo–Mumford … 255
2s−1
for some wi ∈ P ∩ M. Since v0 = 0, we can write x = i=1 wi . If wi + w j ∈ P ∩
M for any i = j, then we can let ti = w
2i−1 + w2i for i = 1, . . . , s − 1 and have x =
s
t1 + · · · + ts−1 + w2s−1 , which lies in i=1 P ∩ M. Therefore,
d
d
s−1
ai vi + x = ai vi + ti + w2s−1 ,
i=0 i=0 i=1
d
d
ai vi + x = v0 + x + (a0 − 1)v0 + ai vi
i=0 i=1
d
= w1 + w2 + y + (a0 − 1)v0 + ai vi
i=1
2(s−1)−1
= w1 + w2 + wi
.
i=0
(k + 1)P ∩ M = P ∩ M + k P ∩ M.
We also define ν P to be the smallest positive integer such that for any k ≥ ν P ,
(k + 1)P ∩ M = V + k P ∩ M.
k P ≤ ν P + d P − 1.
256 B. Le Tran
d
for some x ∈ d P P ∩ M, u i ∈ P ∩ M, i=0 ai = k − ν P . By assumption, k − ν P ≥
d P − 1, so it follows from Lemma 5 that
d d P +k−ν
P
x+ ai vi = u i
(12.7)
i=0 i=1
ν
P −d P d P +k−ν
P
p= ui + u i
.
i=1 i=1
Remark 8 This bound is stronger than [17, Proposition 2.4] since ν P ≤ d [21,
Proposition 2.2] and d P ≤ d/2 [17, Proposition 2.2].
V + k P ∩ M (k + 1)P ∩ M.
d
Indeed, any x ∈ (k + 1)P ∩ M can be written as x = i=0 ai vi such that ai ≥ 0
d
d
and i=0 ai = k + 1. If ai < 1 for all i, then d > k and the point i=0 (1 − ai )vi
is an interior lattice point of (d − k)P, a contradiction since d − k ≤ d − deg(P).
Hence, ai ≥ 1 for some i, say a0 ≥ 1. Then
12 An Eisenbud–Goto-Type Upper Bound for the Castelnuovo–Mumford … 257
d
d
x = v0 + (a0 − 1)v0 + ai vi = v0 + (a0 − 1)v0 + ai vi ∈ V + k P ∩ M.
i=1 i=1
k P ≤ d P + ν P − 1 ≤ d P + deg(P) − 1
≤ d P + Vol(P) − |P ∩ M| + d.
Remark 11 We show some cases that the result of Proposition 10 is stronger
than [17, Proposition 2.4]:
1. Vol(P) ≤ |P ∩ M| + 2. In this case,
d d
Vol(P) − |P ∩ M| + d + ≤d+ − 2.
2 2
We will show in next section that this is the only case that we still need to consider
in order to verify the Eisenbud–Goto conjecture for very ample simplices.
Example 13 Consider P = 2d for d ≥ 4, where d is the standard d-simplex.
Then deg(P) = 2 and by Proposition 7,
258 B. Le Tran
d d
k P ≤ dP + 1 ≤ +1< + d − 1.
2 2
Proof Let P be the corresponding polytope of the embedding. From (12.2), (12.3),
and Proposition 10, it follows that
d d
reg(X ) ≤ Vol(P) − |P ∩ M| + d + + 1 = deg(X ) − codim(X ) + .
2 2
In this section, we will improve the bound of k-normality for non-hollow very ample
simplices.
We now show that the inequality (12.4) holds for non-hollow very ample simplices.
Proposition 16 Let P ⊆ MR be a non-hollow very ample lattice d-simplex. Then
k P ≤ Vol(P) − |P ∩ M| + d + 1.
d
d
p= λi vi , λi = d P
i=0 i=0
and
12 An Eisenbud–Goto-Type Upper Bound for the Castelnuovo–Mumford … 259
d
d
u= λi∗ vi , λi∗ = 1,
i=0 i=0
respectively. It follows from the condition of p that λi < 1 for all 0 ≤ i ≤ d and
there exists 0 ≤ i ≤ d such that λi < λi∗ , say i = 0. By Lemma 4,
d
p + (d P − 1)v1 = ai vi + bu
i=0
d
d
for some ai , b ∈ Z≥0 such that b + ai = 2d P − 1. Replacing p by λi vi
d i=0 i=0
and u by i=0 λi∗ vi yields
λ0 = a0 + bλ∗0
λ1 + d P − 1 = a1 + bλ∗1
λ2 = a2 + bλ∗2
..
.
λd = ad + bλ∗d .
h ∗0 = 1
h ∗1 = |P ∩ M| − d − 1 ≥ 2
h ∗d = |P 0 ∩ M| ≥ 2.
Vol(P) − |P ∩ M| + d + 1 = h ∗0 + h ∗2 + · · · + h ∗d ≥ 1 + 2(d − 1) = 2d − 1.
Corollary 19 The Eisenbud–Goto conjecture holds for any projective toric variety
corresponding to a very ample Fano simplex.
Finally, we would love to see a classification of hollow very ample lattice simplices.
For dimension 2, Rabinotwiz [18, Theorem 1] showed that any such simplex is
unimodularly equivalent to either T p,1 := conv(0, ( p, 0), (0, 1)) for some p ∈ N
or T2,2 = conv(0, (2, 0), (0, 2)). Now we will show a way to obtain some hollow
very ample simplices in any dimension with arbitrary volume.
We recall the definition of lattice pyramids as in [16]:
Definition 21 Let B ⊆ Rk be a lattice polytope with respect to Zk . Then conv(0, B ×
{1}) ⊆ Rk+1 is a lattice polytope with respect to Zk+1 , called the (1-fold) standard
pyramid over B. Recursively, we define for l ∈ N≥1 in this way the l-fold standard
pyramid over B. As a convention, the 0-fold standard pyramid over B is B itself.
Proposition 22 Let P be a lattice polytope. Then the 1-fold pyramid over P is very
ample if and only if P is normal.
12 An Eisenbud–Goto-Type Upper Bound for the Castelnuovo–Mumford … 261
Proof Let Q = conv(0, P × {1}) be the 1-fold pyramid over P. Then it is easy to
see that if P is normal then so is Q. Now suppose that Q is very ample. We have k Q ≥
k P [21, Lemma 4.2.2] and each lattice point of k Q Q ∩ M sits in (t P ∩ M) × {t} for
some 0 ≤ t ≤ k Q . In particular, suppose that (x, t) ∈ (t P ∩ M) × {t} ⊆ k Q Q ∩ M.
Then
t
(x, t) = (u i , 1) + (k Q − t)0
i=1
t
for some u i ∈ P ∩ M. It follows that x = i=1 u i . Hence, P is t-normal for all k Q ≥
t ≥ 1. Since k Q ≥ k P , it follows that P is normal. The conclusion follows.
From Proposition 22, if we take any (d − 2)-fold pyramid over either T p,1 with p ∈
Z≥1 or T2,2 , which are all normal, then we obtain a hollow normal (hence very
ample) d-simplex with normalized volume p. The Eisenbud–Goto conjecture holds
for these.
Example 23 We give here an example to demonstrate the case that if Q is very
ample but not normal then the 1-fold pyramid over Q is not very ample. Let Q be
the convex polytope given by taking s = 4 in Example 20. Then Q is very ample;
however, the 1-fold pyramid of Q, which is given by the convex hull of
⎛ ⎞
0 0 1 0 0 1 0 11
⎜0 0 0 1 0 0 1 1 1⎟
⎜ ⎟,
⎝0 0 0 0 1 1 1 4 5⎠
0 1 1 1 1 1 1 11
Acknowledgements We would like to thank Milena Hering for reading the drafts of this article
and for some valuable suggestions.
References
1. Batyrev, V., Nill, B.: Multiples of lattice polytopes without interior lattice points. Mosc. Math.
J. 7(2), 195–207, 349 (2007)
2. Beck, M., Delgado, J., Gubeladze, J., Michał ek, M.: Very ample and Koszul segmental fibra-
tions. J. Algebraic Combin. 42(1), 165–182 (2015)
3. Bruns, W., Gubeladze, J.: Polytopes, Rings, and K -Theory. Springer Monographs in Mathe-
matics. Springer, Dordrecht (2009)
4. Cox, D.A., Little, J.B., Schenck, H.K.: Toric Varieties. Graduate Studies in Mathematics, vol.
124. American Mathematical Society, Providence (2011)
5. Derksen, H., Sidman, J.: A sharp bound for the Castelnuovo-Mumford regularity of subspace
arrangements. Adv. Math. 172(2), 151–157 (2002)
6. Ehrhart, E.: Sur les polyèdres rationnels homothétiques à n dimensions. C. R. Acad. Sci. Paris
254, 616–618 (1962)
262 B. Le Tran
7. Eisenbud, D., Goto, S.: Linear free resolutions and minimal multiplicity. J. Algebra 88(1),
89–133 (1984)
8. Giaimo, D.: On the Castelnuovo-Mumford regularity of connected curves. Trans. Amer. Math.
Soc. 358(1), 267–284 (2006)
9. Hering, M.S.: Syzygies of toric varieties. ProQuest LLC, Ann Arbor, MI (2006). Thesis
(Ph.D.)–University of Michigan
10. Hibi, T.: A lower bound theorem for Ehrhart polynomials of convex polytopes. Adv. Math.
105(2), 162–165 (1994)
11. Hofscheier, J., Katthän, L., Nill, B.: Ehrhart theory of spanning lattice polytopes. Int. Math.
Res. Not. IMRN 19, 5947–5973 (2018)
12. Kwak, S.: Castelnuovo regularity for smooth subvarieties of dimensions 3 and 4. J. Algebraic
Geom. 7(1), 195–206 (1998)
13. Lazarsfeld, R.: A sharp Castelnuovo bound for smooth surfaces. Duke Math. J. 55(2), 423–429
(1987)
14. McCullough, J., Peeva, I.: Counterexamples to the Eisenbud-Goto regularity conjecture. J.
Amer. Math. Soc. 31(2), 473–496 (2018)
15. Miyazaki, C.: Sharp bounds on Castelnuovo-Mumford regularity. Trans. Amer. Math. Soc.
352(4), 1675–1686 (2000)
16. Nill, B.: Lattice polytopes having h ∗ -polynomials with given degree and linear coefficient. Eur.
J. Combin. 29(7), 1596–1602 (2008)
17. Ogata, S.: k-normality of weighted projective spaces. Kodai Math. J. 28(3), 519–524 (2005)
18. Rabinowitz, S.: A census of convex lattice polygons with at most one interior lattice point. Ars
Combin. 28, 83–96 (1989)
19. Stanley, R.P.: Decompositions of rational convex polytopes. Ann. Discrete Math. 6, 333–342
(1980)
20. Sturmfels, B.: Equations defining toric varieties. In: Algebraic Geometry—Santa Cruz 1995,
Proceedings of Symposia in Pure Mathematics, vol. 62, pp. 437–449. American Mathematical
Society, Providence (1997)
21. Tran, B.L.: On k-normality and regularity of normal projective toric varieties. Ph.D. thesis,
University of Edinburgh (2018)
Chapter 13
Toric Degenerations in Symplectic
Geometry
Milena Pabiniak
13.1 Introduction
Manifolds and algebraic varieties equipped with a group action are usually better
understood as a presence of an action is a sign of certain symmetries. In particular,
toric varieties form a very well understood class of varieties. These are varieties which
contain an algebraic torus TCn := (C∗ )n as a dense open subset and are equipped with
an action of TCn which extends the usual action of TCn on itself. For more about toric
varieties see, for example, [5, 12]. To understand a given projective variety X one
can try to “degenerate” it to a toric one, i.e., form a family of varieties with generic
member X and one special member some toric variety X 0 . The varieties X and X 0
are closely related and thus one can obtain information about X by studying X 0 .
Moreover, such a degeneration gives a map from X to X 0 which, in certain situations,
preserves some special structures X and X 0 might be equipped with (for example: a
symplectic structure).
One can use the method of toric degenerations to solve problems in symplectic
geometry. In this work we discuss the following two applications:
M. Pabiniak (B)
Mathematisches Institut, Universität zu Köln, Weyertal 86-90, 50931 Köln, Germany
e-mail: [email protected]
1. calculating lower bounds for the Gromov width, i.e. trying to find the largest ball
which can be symplectically embedded into a given symplectic manifold;
2. constructing symplectomorphisms needed for a cohomological rigidity problem
for symplectic toric manifolds. This problem is about checking whether any two
symplectic toric manifolds with isomorphic integral cohomology rings (via an
isomorphism preserving the class of symplectic form) are symplectomorphic.
Recall that an 2n-dimensional symplectic manifold (M, ω) equipped with an
effective Hamiltonian action of an n-dimensional torus T = (S 1 )n is called a symplec-
tic toric manifold. The action being Hamiltonian means that there exists a moment
map1 μ : M → Rn . Such a manifold can be given a complex structure interacting
well with the symplectic one so that one calls ω a Kähler form and (M, ω) a Kähler
manifold. In particular, symplectic toric manifolds are toric varieties in the sense of
algebraic geometry. A theorem of Delzant states that we have a bijection2
1 A moment map is a T -invariant map μ : M → Lie(T )∗ = ∼ Rn such that for every X ∈ Lie(T ) it
holds that ι X ω = dμ where X denotes the vector field on M induced by X and μ X : M → R is
X
defined by μ X ( p) = μ( p), X ..
2 Recall that a polytope in Rn is called rational if the directions of its edges are in Zn . It is called
smooth if for every vertex the primitive vectors in the directions of edges meeting at that vertex
form a Z-basis for Zn .
13 Toric Degenerations in Symplectic Geometry 265
dimension n, equipped with a very ample line bundle L, with some fixed Her-
mitian structure. Let L := H0 (X, L) denote the vector space of holomorphic sec-
tions, L : X → P(L ∗ ) the Kodaira embedding and ω = ∗L (ω F S ) the pull back of
the Fubini–Study form, i.e., of the standard symplectic structure on complex projec-
tive spaces. Then (X, ω) is a Kähler manifold. With this data they construct (under
certain assumptions) not only a flat family π : X → C but also a Kähler structure ω̃
on (the smooth part of) X so that (π −1 (1), ω̃|π −1 (1) ) is symplectomorphic to (X, ω).
Moreover, the special fiber X 0 = π −1 (0) obtains a Kähler form, the restriction of ω̃,
defined on its smooth part U0 := (X 0 )smooth , and thus it also obtains a divisor. If X 0 is
normal, then the polytope associated to X 0 and this divisor by the usual procedure of
toric algebraic geometry (see, for example, [5, Chap. 4]) is the closure of the moment
image of the (non-compact) symplectic toric manifold (U0 , ω̃|U0 ). As we will see, this
polytope can be computed by analyzing the behaviour of the holomorphic sections
of L. Here are more details of this procedure.
Denote by L m the image of span f 1 · . . . · f m ; f i ∈ L in H0 (X, L⊗m ) and by
R = C[X ] = ⊕m≥0 L m the homogeneous coordinate ring of X with respect to the
embedding L . An important ingredient of the construction is a choice of a val-
uation with one dimensional leaves, ν : C(X ) \ {0} → Zn , from the ring C(X ) of
rational functions on X . A precise definition of a general valuation can be found, for
example, in [16, Definition 3.1]. In this paper we only use valuations induced by a
flag of subvarieties and a special case of these, called lowest/highest term valuations
associated to a coordinate system.
Example 4 (Lowest/highest term valuations [16, Example 3.2]) Fix a smooth
point p ∈ X and let (u 1 , . . . , u n ) be a system of coordinates in a neighborhood
of p, meaning that u 1 , . . . , u n are regular functions at p, vanishing at p, and such
that their differentials du 1 , . . . , du n are linearly independent at p. Then any reg-
ular function at p can be represented as a power series α∈Zn≥0 cα u α . Here by u α ,
with α = (α1 , . . . , αn ) ∈ Zn≥0 , we mean u α1 1 · . . . · u αn n . Choose and fix a total order >
on Zn respecting the addition, for example the lexicographic order. Define a map ν
from the set of functions regular at p to Zn by
ν cα u α = min{α; cα = 0},
α∈Zn≥0
{ p} = Yn ⊂ . . . ⊂ Y0 = X,
that u k|Yk−1 is a rational function on Yk−1 which is not identically zero and which
has a zero of first order on Yk . Then the lowest term valuation with respect to
the lexicographic order can alternatively be described in the following way: for
any f ∈ C(X ), f = 0, the valuation v( f ) = (k1 , . . . , kn ) where k1 is the order of
vanishing of f on Y1 , k2 is the order of vanishing of f 1 := (u −k
1
1
f )|Y1 on Y2 , etc.
Given such X , L, and ν we form a semigroup S = S(ν, L), in the following way.
Fix a non-zero element h ∈ L and use it to identify L with a subspace of C(X ) by
mapping f ∈ L to f / h ∈ C(X ). Similarly identify L m with a subspace of C(X ) by
sending f ∈ L m to f / h m ∈ C(X ). As any valuation satisfies that ν( f g) = ν( f ) +
ν(g), the set
S = S(ν, L) = {(m, ν( f / h m )) | f ∈ L m \ {0} }
m≥0
is a semigroup with identity (i.e. a monoid). If S is finitely generated, one can con-
struct a toric degeneration whose special fiber is a toric variety Proj C[S] (which is
normal if S is saturated). Moreover we obtain an Okounkov body
=
(S) = conv {x/m | (m, x) ∈ S} ⊂ Rn .
m>0
4 Recall that for a graded algebra A = ⊕∞ j=0 A j the set Proj A is the set of homogeneous prime ideals
in A that do not contain all of A+ := ⊕∞ j=1 A j . The topology on Proj A is defined by setting the
closed sets to be V (I ) := {J ; J ⊂ I is a homogeneous prime ideal of A not containing all of A+ },
for some homogeneous ideal I of A. For more details see, for example [17, II.2], [9, III.2], and [5,
Chap. 7].
268 M. Pabiniak
of the set of the positive real numbers a such that the ball of capacity a (radius πa ),
a
n
Ba2n = B 2n = (x1 , y1 , . . . , xn , yn ) ∈ R2n π (xi2 + yi2 ) < a ⊂ (R2n , ωst ),
π
i=1
can be symplectically embedded in (X, ω). Here ωst = j d x j ∧ dy j denotes the
standard symplectic form on R2n . This question was motivated by the Gromov non-
squeezing theorem which states that a ball B 2n (r ) ⊂ (R2n , ωst ) cannot be symplec-
tically embedded into B 2 (R) × R2n−2 ⊂ (R2n , ωst ) unless r ≤ R.
J -holomorphic curves give obstructions to ball embeddings, while Hamiltonian
torus actions can lead to constructions of such embeddings (by extending a Darboux
chart using the flow of the vector field induced by the action).
This is why toric degenerations provide a useful tool for finding lower bounds on
the Gromov width. Given a toric degeneration of (X, ω), as described in Theorem 6,
one can use the toric action on X 0 to construct embeddings of balls into a smooth
symplectic toric manifold (U0 , ω̃|U0 ), where U0 = (X 0 )smooth . Postcomposing such
embedding with the symplectomorphism φ −1 produces a symplectic embedding
into (X, ω).
Moreover, many embeddings of balls into symplectic toric manifolds can be read
off from the associated (by the Delzant classification theorem) polytope. Identify
the dual of the Lie algebra of the compact torus T with Euclidean space using the
convention that S 1 = R/Z, i.e. the lattice of t∗ is mapped to Zdim T ⊂ Rdim T . With
this convention, the moment map for the standard (S 1 )n action on (R2n , ωst ) maps Ba2n
onto an n-dimensional simplex of size a, closed on n sides
⎧ ⎫
⎨
n ⎬
Sn (a) := (x1 , . . . , xn ) ∈ Rn | 0 ≤ x j < a, xj < a .
⎩ ⎭
j=1
Moreover, if the moment image contains an open simplex of size a, then for any ε > 0
a ball of capacity a − ε can be embedded into the given symplectic toric manifold:
see [28, Lemma 5.3.1] and, independently, [26, Propositions 2.1 and 2.4].
13 Toric Degenerations in Symplectic Geometry 269
Proposition 7 ([24, Proposition 1.3] and [26, Proposition 2.5]) For any connected,
proper (not necessarily compact) symplectic toric manifold U of dimension 2n, with
a momentum map μ, the Gromov width of U is at least
sup{a > 0 | ∃ ∈ GL(n, Z), x ∈ Rn , such that (int Sn (a)) + x ⊂ μ(U )}.
The appearance of
and x comes from the facts that the identification t∗ ∼ = Rdim T
depends on a splitting of T into (dim T ) circles, and that a translation of a moment
map also provides a moment map.
The above results lead to the following method for finding lower bounds on the
Gromov width.
Corollary 8 Let X be a smooth projective variety of complex dimension n, L an
ample line bundle on X , and ω = ∗L (ω F S ) ∈ H2 (X, Z) an integral Kähler form
obtained using the Kodaira embedding L : X → P(L ∗ ). Suppose that there exists
a valuation ν giving a finitely generated and saturated semigroup S = S(ν, L). Let
Proof By the result of [16] cited here as Theorem 6, there exists a toric degenera-
tion of (X, ω) to a normal toric variety X 0 = Proj C[S], and a surjective continuous
map φ : X → X 0 whose appropriate restriction is a symplectomorphism. The sub-
set U := φ −1 (U0 ) of X inherits a toric action whose moment image contains int
,
the interior of
(recall that a moment map sends singular points of a toric variety to
the boundary of the moment polytope). The corollary follows from Proposition 7.
In fact one does not need S to be saturated. The same corollary holds even if X 0
is not a normal toric variety. This is because a normalization map for X 0 induces a
biholomorphism between (X 0 )smooth and an appropriate subset of the normalization
of X 0 .
It is, however, necessary that S is finitely generated for a toric degeneration to
exist. Otherwise one can still form a family of manifolds, but one cannot guarantee
that this family is flat, and thus X and X 0 are no longer so strongly related. As we
already mentioned, Kaveh in [20] observed that such a (not necessarily flat) family,
with X 0 = (C∗ )n , still provides information about the Gromov width of (X, ω).
To state this result we need additional notation. In the notation of Sect. 13.2, for
any m ∈ Z>0 let
1
Am := {ν( f / h m ) | f ∈ L m \ {0} } ⊂ Zn ,
m = conv(Am ).
m
Note that
= ∪m>0
m . Fix m and let r = rm denote the number of elements
in Am = {β1 , . . . , βr }. From these data we form a symplectic form, ωm , on (C∗ )n
using a standard procedure: ωm is the pull back of the Fubini–Study form on CPr −1 via
the map
m : (C∗ )n → CPr −1 , u → (u β1 c1 , . . . , u βr cr ), where c = [(c1 , . . . , cr )] is
270 M. Pabiniak
some element in CPr −1 with all ci = 0. (In [20] the elements ci come from coeffi-
cients of leading terms of elements in appropriately chosen basis of L m . One also
needs that the differences of elements in Am span Zn , which, by [20, Remark 5.6],
is always true for lowest term valuations.)
Kaveh proved that:
1. for every m > 0 there exists an open subset U ⊂ X such that (U, ω) is symplec-
tomorphic to ((C∗ )n , m1 ωm ) [20, Theorem 10.5];
2. the Gromov width of ((C∗ )n , m1 ωm ) is at least Rm , where Rm is the size of the
largest open simplex that fits in the interior of
m = m1 conv (Am ) [20, Corollary
12.3].
This leads to the following corollary.
Corollary 9 ([20, Corollary 12.4]) Let X be a smooth projective variety of dimen-
sion n, L an ample line bundle on X , and ω = ∗L (ω F S ) ∈ H 2 (X, Z) an integral
Kähler form. Let ν be a lowest term valuation on C(X ), with values in Zn , and
the associated Okounkov body. The Gromov width of (X, ω) is at least R, where R
is the size of the largest open simplex that fits in the interior of
.
The methods for finding the Gromov width described in Corollaries 8 and 9 have
been used in [11, 14] for coadjoint orbits of compact Lie groups.
Recall that given a compact Lie group K each orbit O ⊂ k∗ := (Lie K )∗ of the
coadjoint action of K on k∗ is naturally a symplectic manifold. Namely it can be
equipped with the Kostant–Kirillov–Souriau symplectic form ω K K S defined by:
ωξK K S (X # , Y # ) = ξ, [X, Y ], ξ ∈ O ⊂ k∗ , X, Y ∈ k,
5 During the work on the project [18], about complex Grassmannians, Karshon and Tolman looked
at various examples of other coadjoint orbits and got the impression that the above value might be
the Gromov width of all coadjoint orbits. They never formulated this expectation formally as their
13 Toric Degenerations in Symplectic Geometry 271
For example, as {eii − e j j ; i = j} forms a root system for the unitary group U(n, C),
the Gromov width of its coadjoint orbit Oλ passing through a point
λ = diag(λ1 , . . . , λn ) ∈ u(n)∗ ,
The first usage of toric degenerations in Gromov width problems appeared in [14],
where the generic orbits of the symplectic group Sp(n) = U(n, H) are considered.
Then it was used in [11] to prove that the formula (13.1) is a lower bound for the
Gromov width of any coadjoint orbit of any compact connected simple Lie group K ,
passing through a point in the Weyl chamber, integral up to scaling, i.e. to prove
Theorem 1.
The rationality assumption comes from the fact that the toric degeneration method
can be applied only to the orbits passing through an integral point λ of a positive Weyl
conjecture, but they shared this idea with other mathematicians in private communications. This is
how this value became to be known as the expected Gromov width for coadjoint orbits.
6 A coadjoint orbit through a point λ in the interior of a chosen positive Weyl chamber is called
indecomposable in [29] if there exists a simple positive root α such that for any positive root α
there exists a positive integer k such that λ, α = kλ, α.
7 The result about SO(2n + 1, C) holds only for orbits satisfying a mild technical condition: the
point λ of intersection of the orbit and a chosen positive Weyl chamber should not belong to a
certain subset of one wall of the chamber; see [26] for more details. In particular, all generic orbits
satisfy this condition.
272 M. Pabiniak
where Y j denotes the Schubert variety corresponding to element siα j+1 · . . . · siα N of
the Weyl group. We denote by νw0 the highest term valuation associated with this
flag of subvarieties.
13 Toric Degenerations in Symplectic Geometry 273
The following section is based on a project joint with Sue Tolman [27].
274 M. Pabiniak
1 1
It was observed by Hirzebruch that H−A and H− A are diffeomorphic if and only
if A =∼A mod 2. Moreover, the symplectic toric manifolds (H−A , ωλ ) and (H− A,
ωλ ) are (not equivariantly) symplectomorphic if and only if A ∼ =A mod 2 and the
widths and the areas of the associated polytopes agree, i.e. λ1 =
λ1 and λ2 − 21 Aλ1 =
λ2 − 21 Aλ1 . For example, the manifolds presented on Fig. 13.1 are symplectomor-
phic. The cohomology ring can be presented as
with vertices in Zn ) and there exists a very ample line bundle L over X P inducing ω P .
In this situation a basis of the space of holomorphic sections of L can be identified
with the integral points of P, ([6], see also [15]). Without loss of generality we can
assume that P in a neighborhood of some vertex looks like (R≥0 )n in a neighborhood
of the origin in Rn . Then we can identify L = H0 (X P , L) with a subset of the ring of
rational functions, C(X P ), as described on Sect. 13.2, using the section corresponding
to the origin as the fixed element h:
f
f → .
section corresponding to the origin
Note 14 For simplicity of notation, given a valuation ν we will write ν(L) to denote
also gives a coordinate system. Let ν be the associated lowest term valuation (as in
Example 4). The image ν(L) can be obtain by using a “sliding” operator F−ek +cel ,
defined as follows. For each affine line in Rn in the direction of −ek + cel , with P ∩
∩ Zn = ∅, translate the set {P ∩ ∩ Zn } by a(−ek + cel ) with a ≥ 0 maximal non-
negative number for which a(−ek + cel ) + {P ∩ ∩ Zn } ⊂ (R≥0 )n .
Lemma 15 One obtains ν(L) by sliding the integral points of P in the direc-
tion −ek + cel , inside (R≥0 )n , i.e.
Instead of the proof, which can be found in [27], we give the following example
which ilustrates the main idea behind the proof.
Example 16 Let (X P , ω P ) be the symplectic toric manifold corresponding to the
polytope P = conv {(0, 0), (1, 0), (1, 3), (0, 3)} ⊂ R2 . That is, X P is diffeomorphic
to CP1 × CP1 with product symplectic structure (with different rescaling of the
Fubini–Study symplectic form on each factor). Let ν be the lowest term valuation
associated to the coordinate system
u 1 = f 1 − f 22 , u 2 = f 2 .
13 Toric Degenerations in Symplectic Geometry 277
Line {(0, 2) + t (1, −2); t ∈ R} intersects P in two integral points: (1, 0) and (0, 2).
The corresponding functions are f 1 and f 22 , and one can easily calculate that
Similarly, using the integral points on the line {(0, 3) + t (1, −2); t ∈ R} we obtain
More generally, if the integral points (a, b), (a, b) + (1, −2), . . . , (a, b) + m(1, −2)
are in P (implying that b − 2m > 0), then one can use the corresponding functions to
construct functions with valuations (0, b + 2a), (0, b + 2a) + (1, −2), . . . , (0, b +
2a) + m(1, −2) = (m, 2a + b − 2m). Precisely, for any l = 0, . . . , m
l
a+ j b−2 j
f 1a f 2b−2l ( f 1 − f 22 )l = (−1)l− j f 1 f 2 and
j=0
This proves that ν(L) ⊃ F(−1,2) (P ∩ Z2 ). By [16, Proposition 3.4] the cardinality
of ν(L) is the dimension of L, that is, the number of integral points in P. Therefore
ν(L) = F(−1,2) (P ∩ Z2 ).
S = (cone ({1} ×
)) ∩ (Z × Zn ).
∀ m ≥ 1 ν(L m ) = m
∩ Zn = conv(ν(L m )) ∩ Zn .
that is, X P is diffeomorphic to CP1 × CP1 as in the previous example, but the
symplectic form is different. As before, let ν be the lowest term valuation associated
to the coordinate system
u 1 = f 1 − f 22 , u 2 = f 2 .
Then
ν(L) = F(−1,2) (P ∩ Z2 )
= {(0, j); j = 0, . . . , 6} ∪ {(1, 0), (1, 2)} conv (ν(L)) ∩ Z2 .
Let us analyse why in the above example the point (1, 1) is missing. Observe that
the parallel lines 1 := {(0, 2) + t (−1, 2); t ∈ R}, 2 := {(0, 3) + t (−1, 2); t ∈ R}
and 3 := {(0, 4) + t (−1, 2); t ∈ R} intersect P at intervals of the same length but
with, respectively, 2, 1 and 2 integral points. Therefore the intersections of 1 , 2
13 Toric Degenerations in Symplectic Geometry 279
Assume that
λ2 − cλ1 > 0.
(conv F(−1,c) (
∩ Z2 )) ∩ Z2 = F(−1,c) (
∩ Z2 ).
=
(A, λ) = p ∈ Rn p, e j ≥ 0 and p, e j + Aij ei ≤ λ j ∀ 1 ≤ j ≤ n ,
i
with [ωλ ] = i λi xi ∈ H∗ (M A ; Z) ⊗Z R. If all coefficients λi are integral then [ωλ ]
is an integral symplectic. Note that this particular presentation of H∗ (M A ; Z) depends
8In the standard action of (S 1 )n on (CP1 )n each S 1 in (S 1 )n acts on the respective copy of CP1
by eit · [(z 0 , z 1 )] = [(z 0 , eit z 1 )].
13 Toric Degenerations in Symplectic Geometry 281
on
A. (The element x j is the Poincaré dual to the preimage of facet
(A, λ) ∩
{ p, e j + i Aij ei = λ j }.)
We say that a Bott manifold is Q-trivial if H∗ (M; Q) H∗ ((CP1 )n ; Q). For
example, observe that all Hirzebruch surfaces are Q-trivial Bott manifolds.
Recall that we want to prove Theorem 2 which says that for Q-trivial Bott mani-
folds (N , ω N ) and (M, ω M ), and any ring isomorphism F : H∗ (M; Z) → H∗ (N ; Z),
with F([ω M ]) = [ω N ], there exists a symplectomorphism f : (N , ω N ) → (M, ω M )
inducing F. The key ingredient of the proof of Theorem 2 is the following construc-
tion of symplectomorphisms, which uses toric degenerations.
Proof (sketch) Without loss of generality we can assume that the polytope
(A, λ)
associated to (A, λ) is normal, that is, any integral point of m
(A, λ) can be
expressed as a sum of m integral points of
(A, λ). Indeed, if
(A, λ) is not a normal
polytope, replace (M, ω) and ( M, ω) by (M, (n − 1) ω) and ( M, (n − 1) ω). This
dilates the corresponding polytopes by (n − 1). For any integral polytope P ⊂ Rn
its dialate m P with m ≥ n − 1 is normal (see, for example, [5, Theorem 2.2.12]).
Obviously if (M, (n − 1) ω) and ( M, (n − 1) ω) are symplectomorphic, then so
are (M, ω) and ( M, ω). As usually, let L denote the very ample line bundle over M
corresponding to ω and L the space of its holomorphic sections. Note that normality
implies that L m can be identified with H 0 (M, L⊗m ) because a basis for both of these
vector spaces is given by the integral points m
(A, λ) ∩ Zn .
Also without loss of generality we can assume that A k ≥ Ak . Let c = 1 (Ak +
2
A ) ≥ 0. We will work with a lowest term valuation ν associated to the following
k
coordinate system
From Lemma 15 and the normality assumption, for all m ≥ 1 we have that
For that computation one uses relations between A, λ, A and λ which are implied
by the facts that
(A, λ) and
( A, λ) are combinatorially hypercubes, and by the
existence of the isomorphism described in the statement of the proposition. In partic-
ular, these relations also allow to generalize Corollary 19 (precisely: to show that the
equivalent of condition λ2 − cλ1 > 0 holds). Therefore the semigroup S associated
to the valuation ν of (M, ω) is exactly S = (cone
( A, λ)) ∩ (Z × Zn ). Then the
claim follows from Proposition 17.
Using Proposition 20 we show below (Corollary 23) that each Q-trivial Bott manifold
is associated to a matrix A of a particularly easy form. To explain this idea we need
few more definitions. Recall the presentation of the cohomology of symplectic Bott
manifold M A given in (13.2). We define the following special elements
1
αk = − Akj x j ∈ H∗ (M A ; Z), yk = xk − αk ∈ H∗ (M A ; Q)
j
2
for all k. We say xk is of even (odd) exceptional type if αk = cyl for some l, where c is
an even (respectively, odd) integer. In “coordinates”, this means that Akj = 0 for j < l
and Akj = 21 Alk Alj for j > l. Note that if xk is even (resp. odd) exceptional, say αk =
myl , then one can construct an isomorphism of Proposition 20 from H∗ (M A ; Z)
to H∗ (M A; Z) for some A with Alk equal to 0 (resp. −1). For example if xk is of
even exceptional type, i.e. αk = 2myl for some m and l, implying that Alk = −2m
and Akj = −m Alj for j = l, then one should put A lk = 0, A
ij = Aij for all i and
all j = l, and Al = Al + m Ak for all i = k. Therefore, consecutive applications of
i i i
the above proposition lead to simplifying the description of a given Bott manifold.
Corollary 21 Any symplectic toric Bott manifold, with integral symplectic form is
symplectomorphic to one for which Alk = 0 (resp. Alk = −1) whenever xk has even
(resp. odd) exceptional type and αk = myl .
In the case of Q-trivial Bott manifolds all xi have exceptional type, [3, Proposition
3.1]. Therefore, any Q-trivial symplectic toric Bott manifold with integral symplectic
form must be a product of the following standard models of Q-trivial Bott manifolds.
Example 22 (Q-trivial Bott manifold) Take n ∈ Z>0 . Let Ain = −1 for all 1 ≤
i < n, and Aij = 0 otherwise. For such upper triangular matrix A = [Aij ] and
any λ ∈ (R>0 )n , the polytope
(A, λ) is combinatorially a hypercube, thus it defines
a symplectic toric Bott manifold, which we will denote by H = H(λ1 , . . . , λn ).
Observe that
H∗ (H; Z) = Z[x1 , . . . , xn ]/ x12 − x1 xn , . . . , xn−1
2
− xn−1 xn , xn2 .
m
More generally, any partition of n, i=1 li = n together with λ ∈ (R>0 )n , define
a Q-trivial Bott manifold
The above standard model is easy enough, so that one can understand all possible
ring isomorphisms between cohomology rings and prove that they are induced by
maps on manifolds.
m m
Lemma 24 Fix n ∈ Z>0 . Let i=1 li = i=1
li = n be partitions of n, and let λ,
λ∈
(R>0 )n . Consider symplectic Bott manifolds
(M, ω) = H(λ1 , . . . , λn ) and ( M, ω) = H(
λ1 , . . . ,
λn ).
The Q-triviality assumption implies that there are exactly 2n primitive classes
in H2 (M; Z) which square to 0. A short computation shows that these are ±z 1 , . . . ,
±z n , where z n = xn and z i = 2xi − xn for all i < n. Similarly for M. As the coho-
mology of a symplectic toric manifold is generated in degree 2, any ring isomorphism
between H∗ (M; Z) and H∗ ( M; Z) restricts to a bijection on the set of such elements,
that is, there exists = (1 , . . . , n ) ∈ {−1, 1}n and a permutation σ ∈ Sn such
that F(z j ) = jz σ ( j) . Moreover, presenting [ω] (resp. [ ω]) in R-basis {z 1 , . . . , z n }
of H∗ (M; Z) ⊗Z R (resp. { z 1 , . . . ,
z n }) and recalling that the isomorphism F is to
map [ω] to [ ω], one can deduce that F acts by a permutation: F(z j ) = z σ ( j) for some
permutation σ ∈ Sn with σ (n) = n, and that λ j = λσ ( j) . Moreover F takes xi to xσ (i)
and it holds that Aij = A σ (i) for all i, j. If ∈ GL(n, Z) denotes the unimodular
σ ( j)
matrix taking ei to eσ (i) , then T (
( A, λ)) =
(A, λ); Therefore, by the Delzant
theorem, the manifolds (M, ω) and ( M, ω) are (equivariantly) symplectomorphic,
by some symplectomorphism f . Moreover, as T maps the facet{ p, eσ ( j) =
0} ∩
( A, λ) to the facet { p, e j = 0} ∩
(A, λ), and { p, eσ ( j) + i A i ei =
σ ( j)
i
λσ ( j) } ∩
( A, λ) to { p, e j + i A j ei = λ j } ∩
(A, λ), the map H∗ ( f ) induced
284 M. Pabiniak
Proof (Proof of Theorem 2) Let (M, ω), ( M, ω) be two Q-trivial Bott manifolds
with symplectic forms integral up to scaling and let F : H∗ (M; Z) → H∗ ( M; Z)
be a ring isomorphism such that F[ω] = [ ω]. Rescaling the symplectic forms if
necessary we can assume that both ω and ω are integral. As the cohomology of a
symplectic toric manifold is generated in degree 2, the isomorphism F must map
Z). Using (13.2) we see that dim H2 (M; Z) = 1 dim M, and
H2 (M; Z) to H2 ( M; 2
2 Therefore dim M = dim M. We will denote
similarly dim H ( M; Z) = 21 dim M.
this dimension by 2n. By Corollary 23 and the assumption that the symplectic forms
are integral we have that
m m
for some i=1 li = i=1
li = n partitions of n, and some λ,
λ ∈ (Z>0 )n . Now
Lemma 24 gives that there exist a symplectomorphism f from ( M, ω) to (M, ω) so
that H∗ ( f ) = F.
Acknowledgements First of all, the author would like to thank her collaborators: Xin Fang, Iva
Halacheva, Peter Littelmann and Sue Tolman. Results contained in this manuscript were obtained
in collaboration with the above mathematicians [11, 14, 27], and therefore all of them could also
be considered as the authors of this paper. The author also thanks the organisers of the workshops
“Interactions with Lattice Polytopes” for giving her the opportunity to participate and present her
results at these workshops. The author is supported by the DFG (Die Deutsche Forschungsgemein-
schaft) grant CRC/TRR 191 “Symplectic Structures in Geometry, Algebra and Dynamics”.
13 Toric Degenerations in Symplectic Geometry 285
References
1. Anderson, D.: Okounkov bodies and toric degenerations. Math. Ann. 356(3), 1183–1202 (2013)
2. Caviedes Castro, A.: Upper bound for the Gromov width of coadjoint orbits of compact Lie
groups. J. Lie Theory 26(3), 821–860 (2016)
3. Choi, S., Masuda, M.: Classification of Q-trivial Bott manifolds. J. Symplectic Geom. 10(3),
447–461 (2012)
4. Choi, S., Masuda, M., Suh, D.Y.: Rigidity problems in toric topology: a survey. Tr. Mat.
Inst. Steklova 275(Klassicheskaya i Sovremennaya Matematika v Pole Deyatel’nosti Borisa
Nikolaevicha Delone), 188–201 (2011)
5. Cox, D.A., Little, J.B., Schenck, H.K.: Toric varieties, Graduate Studies in Mathematics, vol.
124. American Mathematical Society, Providence, RI (2011)
6. Danilov, V.I.: The geometry of toric varieties. Uspekhi Mat. Nauk 33(2(200)), 85–134, 247
(1978)
7. da Silva, A.C.: Lectures on symplectic geometry. Lecture Notes in Mathematics, vol. 1764.
Springer, Berlin (2001)
8. Eisenbud, D.: Commutative algebra. Graduate Texts in Mathematics, vol. 150. Springer, New
York (1995). With a view toward algebraic geometry
9. Eisenbud, D., Harris, J.: The geometry of schemes. In: Graduate Texts in Mathematics, vol.
197. Springer, New York (2000)
10. Fang, X., Fourier, G., Littelmann, P.: Essential bases and toric degenerations arising from
birational sequences. Adv. Math. 312, 107–149 (2017)
11. Fang, X., Littelmann, P., Pabiniak, M.: Simplices in Newton-Okounkov bodies and the Gromov
width of coadjoint orbits. Bull. Lond. Math. Soc. 50(2), 202–218 (2018)
12. Fulton, W.: Introduction to toric varieties. Annals of Mathematics Studies, vol. 131. Princeton
University Press, Princeton, NJ (1993). The William H. Roever Lectures in Geometry
13. Grossberg, M., Karshon, Y.: Bott towers, complete integrability, and the extended character of
representations. Duke Math. J. 76(1), 23–58 (1994)
14. Halacheva, I., Pabiniak, M.: The Gromov width of coadjoint orbits of the symplectic group.
Pacific J. Math. 295(2), 403–420 (2018)
15. Hamilton, M.D.: The quantization of a toric manifold is given by the integer lattice points in
the moment polytope. In: Toric topology, Contemporary Mathematics, vol. 460, pp. 131–140.
American Mathematical Society, Providence, RI (2008)
16. Harada, M., Kaveh, K.: Integrable systems, toric degenerations and Okounkov bodies. Invent.
Math. 202(3), 927–985 (2015)
17. Hartshorne, R.: Algebraic Geometry. Springer, New York, Heidelberg (1977). Graduate Texts
in Mathematics, No. 52
18. Karshon, Y., Tolman, S.: The Gromov width of complex Grassmannians. Algebr. Geom. Topol.
5, 911–922 (2005)
19. Kaveh, K.: Crystal bases and Newton-Okounkov bodies. Duke Math. J. 164(13), 2461–2506
(2015)
20. Kaveh, K.: Toric degenerations and symplectic geometry of smooth projective varieties. J.
Lond. Math. Soc. (2) 99(2), 377–402 (2019)
21. Lane, J.: A Completely Integrable System on g2 Coadjoint Orbits (2016). arXiv:1605.01676
22. Littelmann, P.: Cones, crystals, and patterns. Transform. Groups 3(2), 145–179 (1998)
23. Lu, G.: Gromov-Witten invariants and pseudo symplectic capacities. Israel J. Math. 156, 1–63
(2006)
24. Lu, G.: Symplectic capacities of toric manifolds and related results. Nagoya Math. J. 181,
149–184 (2006)
25. Nishinou, T., Nohara, Y., Ueda, K.: Toric degenerations of Gelfand-Cetlin systems and potential
functions. Adv. Math. 224(2), 648–706 (2010)
26. Pabiniak, M.: Gromov width of non-regular coadjoint orbits of U (n), S O(2n) and S O(2n + 1).
Math. Res. Lett. 21(1), 187–205 (2014)
286 M. Pabiniak
27. Pabiniak, M., Tolman, S.: Symplectic cohomological rigidity via toric degnerations. In prepa-
ration
28. Schlenk, F.: Embedding problems in symplectic geometry. In: De Gruyter Expositions in Math-
ematics, vol. 40. Walter de Gruyter GmbH & Co. KG, Berlin (2005)
29. Zoghi, M.: The Gromov Width of Coadjoint Orbits of Compact Lie Groups. ProQuest LLC,
Ann Arbor, MI (2010). Thesis (Ph.D.)–University of Toronto (Canada)
Chapter 14
On Deformations of Toric Fano Varieties
Andrea Petracci
Abstract In this note we collect some results on the deformation theory of toric
Fano varieties.
14.1 Introduction
A Fano variety is a normal projective variety X over C such that its anticanonical
divisor −K X is Q-Cartier and ample. Fano varieties constitute the basic building
blocks of algebraic varieties, according to the Minimal Model Program. The geometry
of Fano varieties is a well studied area. In particular, moduli (and consequently
deformations) of Fano varieties constitute a very interesting and important topic in
algebraic geometry, e.g. [21, 62, 69].
Here we will concentrate on deformations and smoothings of toric Fano varieties.
These varieties occupy a prominent role in Mirror Symmetry, a large part of which
is based on the phenomenon of toric degeneration as in [17, 18, 30, 43].
Toric Fano varieties correspond to certain polytopes which are called Fano poly-
topes. The goal of this note is to present some combinatorial criteria on Fano poly-
topes which can detect whether the corresponding toric Fano variety is smoothable,
i.e. can be deformed to a smooth (Fano) variety.
Special attention is given to toric Fano threefolds with Gorenstein singularities.
These varieties correspond to the 4319 reflexive polytopes of dimension 3, which
were classified by Kreuzer and Skarke [66]. In this case, thanks to the use of the soft-
ware Magma [22], we were able to produce a lot of examples for the combinatorial
criteria discussed in this note.
A. Petracci (B)
Dipartimento di Matematica, Università di Bologna, Piazza di Porta San Donato 5,
40126 Bologna, Italy
e-mail: [email protected]
14.1.1 Outline
We work over C, but everything will hold over a field of characteristic zero with
appropriate modifications.
In Sects. 14.3 and 14.4 we assume that the reader is familiar with the basic notions
of toric geometry, which can be found in [34, 41]. All toric varieties considered here
are normal. A lattice is a finitely generated free abelian group. The letters N , N , Ñ
stand for lattices and M, M, M̃ for their duals, e.g. M = HomZ (N , Z); the duality
pairing M × N → Z and its extension MR × NR → R are denoted by ·, ·.
In a real vector space of finite dimension a polytope is the convex hull of finitely
many points, or equivalently a compact subset which is the intersection of finitely
many closed halfspaces. We refer the reader to the book [99] for the geometry of
polytopes.
14.2 Deformations
Let (Comp) be the category of noetherian complete local C-algebras with residue
field C. For every R ∈ (Comp) we denote by m R the maximal ideal of R. Let (Art)
be the subcategory of (Comp) whose objects are artinian, i.e. local finite C-algebras.
14 On Deformations of Toric Fano Varieties 289
A functor of Artin rings is a functor F from the category (Art) to the category of
sets such that F(C) is the set with one element. We will only consider functor of
Artin rings which satisfy some additional properties: Schlessinger’s axioms (H1)
and (H2) [91] and Fantechi–Manetti condition (L) [37, (2.9)]. We will not specify
these conditions here, but we refer the reader to [37, Sect. 2] for a quick introduction.
Precise formulations and additional details about the notions we introduce below
can be found in any reference about deformation theory, e.g. [13, 36, 50, 70, 91, 92,
95, 97].
A natural transformation (or briefly map) of functors φ : F → G is called smooth
if the lifting property in Grothendieck’s definition of formally smooth morphisms
holds, i.e. for every local surjection A A in (Art) the natural map F(A ) →
F(A) ×G(A) G(A ) is surjective; in particular, if φ is smooth then φ(A) : F(A) →
G(A) is surjective for all A ∈ (Art). A functor F is called smooth if the map from F
to the trivial functor is smooth.
For a functor F, the set F(C[t]/(t 2 )) has a natural structure of a C-vector space,
denoted by TF and called the tangent space of F. One can prove that F is the trivial
functor if and only if TF = 0. If φ : F → G is a map, then the function φ(C[t]/(t 2 ))
is linear and denoted by Tφ : TF → TG.
If R ∈ (Comp) one can consider the functor h R = Hom(·, R) prorepresented
by R. A map h R → F is equivalent to a pro-object of F on R = lim R/mn+1 R , i.e.
←−
an element of the set lim F(R/m R ). A hull for a functor F is a ring R ∈ (Comp)
n+1
←−
together with a smooth morphism φ : h R → F such that Tφ is bijective. A hull exists
if and only if TF has finite dimension. If a hull exists, it is unique. Provided that TF
has finite dimension r , then F is smooth if and only if the hull of F is isomorphic
to C[[t1 , . . . , tr ]].
For a functor F, consider the set E made up of pairs (π, ξ ), where π : A →
A is a surjection in (Art) such that m A · (ker π ) = 0 and ξ ∈ F(A). A C-vector
space
V is called an obstruction space for F if there exists a function ω : E →
(π,ξ )∈E ker π ⊗C V such that the two following conditions are satisfied:
called a formal deformation of X over R. If R is a hull for Def X , then the cor-
responding formal deformation of X over R is called the miniversal deformation
of X . We say that X is rigid if all deformations of X are trivial. If X is reduced,
then the tangent space of Def X is Ext 1 ( X , O X ); in this case X is rigid if and only
if Ext 1 ( X , O X ) = 0. If X is either normal or reduced and local complete intersec-
tion (l.c.i. for short), then Ext 2 ( X , O X ) is an obstruction space for Def X . If X is
smooth, then Hi (X, TX ) = Exti ( X , O X ) for all i ≥ 0. In particular, if X is smooth
and affine then it is rigid.
Proposition 1 If X is a smooth Fano variety, then Hi (X, TX ) = 0 for each i ≥ 2. In
particular, the infinitesimal deformations of X are unobstructed, i.e. Def X is smooth.
Proof Let n be the dimension of X . Since the anticanonical line bundle ω∨X is ample,
by Kodaira–Nakano vanishing we have Hi (X, n−1 X ⊗ ω∨X ) = 0 whenever i + n −
1 > n, i.e. i ≥ 2. We conclude because the tangent sheaf TX is isomorphic to n−1
X ⊗
ω∨X .
Let X be a scheme of finite type over C and let Def ltX be the subfunctor of Def X
made up of the locally trivial deformations of X . The tangent space of Def ltX
is H1 (X, TX ) and H2 (X, TX ) is an obstruction space for Def ltX .
Proposition 2 Let X be a reduced scheme of finite type over C such that X is either
l.c.i. or normal. If H0 (X, Ext 1 ( X , O X )) = 0, then all deformations of X are locally
trivial, i.e. Def ltX = Def X .
Proof The local-to-global spectral sequence for Ext gives the following exact
sequence.
The vanishing of H0 (Ext 1 ( X , O X )) implies that the inclusion φ : Def ltX → Def X
induces an isomorphism on tangent spaces and an injection on obstruction spaces.
Therefore φ is smooth, and consequently surjective.
In particular, all deformations of a smooth scheme are locally trivial.
Let X be a reduced scheme of finite type over C with isolated singularities.
For each singular point x ∈ X , let Ux be an affine open neighbourhood of x such
that Ux \ {x} is smooth. Then define
X :=
Def loc Def Ux .
x∈Sing(X )
Proposition 3 Let X be a reduced scheme of finite type over C with isolated sin-
gularities. Assume that X is either l.c.i. or normal. If H2 (X, TX ) = 0 then there
are no local-to-global obstructions for the infinitesimal deformations of X , i.e. the
map Def X → Def locX is smooth.
Proof We consider the local-to-global spectral sequence for Ext • ( X , O X ). The sec-
p,q
ond page is given by E 2 = H p (Ext q ( X , O X )). Since X has isolated singularities,
the sheaves Ext ( X , O X ) are supported on isolated points for q ≥ 1; in particu-
q
p,q
lar they do not have higher cohomology. This means that E 2 is supported on the
lines p = 0 and q = 0. Therefore, in E 2 the only non-zero differential is
d2 : H0 (Ext 1 ( X , O X )) −→ H2 (TX ).
We obtain that the bottom left corner of the third page E 3 is the following.
H3 (TX ) 0 0 0
cokerd2 0 0 0
H1 (TX ) 0 0 0
H0 (TX ) ker d2 H0 (Ext 2 ( X , O X )) H0 (Ext 3 ( X , O X ))
d3 : H0 (Ext 2 ( X , O X )) −→ H3 (TX ).
cokerd3 0 0 0
cokerd2 0 0 0
H1 (TX ) 0 0 0
H0 (TX ) ker d2 ker d3 H0 (Ext 3 ( X , O X ))
From the fourth page on, the pieces of total degree ≤ 3 do not change any more.
Therefore we have two short exact sequences:
d2
0 −→ H1 (TX ) −→ Ext 1 ( X , O X ) −→ H0 (Ext 1 ( X , O X )) −→
d2
−→ H2 (TX ) −→ Ext 2 ( X , O X ) −→ H0 (Ext 2 ( X , O X ))
292 A. Petracci
So far we did not use the assumption H2 (TX ) = 0. From this vanishing, via the long
exact sequence above we deduce that the map Def X → Def loc X induces a surjection
on tangent spaces and an injection on obstruction spaces.
14.2.2 Smoothings
Here we discuss smoothability conditions for schemes of finite type over C. We will
only consider the case of equidimensional schemes and we will refer the reader to [50,
Sect. 29] for a more general treatment, which uses the Lichtenbaum–Schlessinger
functors.
If X is a proper scheme over C, a smoothing of X is a proper flat morphism X → B
such that B is an integral scheme of finite type over C of positive dimension and
there exists a closed point b0 ∈ B such that the fibre over b0 is X and all the other
fibres are smooth. By restricting to a curve in B and normalising it, we may require
that the base B is a smooth affine curve and that the maximal ideal corresponding
to b0 is principal. We say that X is smoothable if it admits a smoothing.
For every n ≥ 0, set Sn := Spec C[t]/(t n+1 ). If X is a scheme of finite type over C
with pure dimension d, then a formal smoothing of X is a formal deformation {X n →
Sn }n of X over C[[t]] such that there exists m such that t m is in the dth Fitting ideal
of X m /Sm . We refer the reader to [35, Sect. 20.2] for the definition and the properties
of Fitting ideals. We say that X is formally smoothable if it admits a formal smoothing.
It is clear that if X is formally smoothable, then every open subscheme of X is
formally smoothable.
Fittd ( X n /Sn ) = (Fittd ( X n+1 /Sn+1 ) + t n+1 O X n+1 )/t n+1 O X n+1 .
Therefore if t n ∈ Fittd ( X n /Sn ) then t n ∈ Fittd ( X n+1 /Sn+1 ) + t n+1 O X n+1 , hence t n+1 ∈
tFittd ( X n+1 /Sn+1 ) ⊆ Fittd ( X n+1 /Sn+1 ) as t n+2 = 0 in O X n+1 .
Then:
1. if π is a smoothing of X , then ξ is a formal smoothing of X ;
2. if ξ is a formal smoothing of X , then there exists an open neighbourhood B of b0
in B such that X × B B → B is a smoothing of X .
14 On Deformations of Toric Fano Varieties 293
Proof We may assume that X is connected. Therefore X has pure dimension, say d.
(1) This follows immediately from Lemma 5 and from the fact that if X is formally
smoothable then every open subscheme of X is formally smoothable.
(2) Set d := dim X . Let ξ = {X n → Sn }n be a formal smoothing of X , where Sn
is Spec C[t]/(t n+1 ) as usual. Let m be such that t m is in the dth Fitting ideal of X m /Sm .
As X is proper over C, the tangent space of Def X has finite dimension, there-
fore Def X has a hull R ∈ (Comp). Let η = {ηn : Yn → Spec R/mn+1 R }n be the
miniversal deformation of X . By [95, Proposition 6.51] or [92, Theorem 2.5.13],
from H2 (O X ) = 0 we deduce that η is effective, i.e. there exists a projective flat
morphism X → Spec R whose m R -adic completion is η.
294 A. Petracci
By a theorem of Artin [11, Theorem 1.6] (see also [50, Theorem 21.3]), the mor-
phism X → Spec R is algebraizable in the following sense: there exist a scheme Z
of finite type over C, a closed point z 0 ∈ Z , and a proper flat morphism X → Z ,
with fibre X over z 0 , such that R is the completion
O Z ,z0 of the local ring of Z at z 0
and X is isomorphic, as R-schemes, to X × Z Spec R. In particular, the miniversal
deformation η is the collection {X × Z Spec O Z ,z0 /mn+1
z 0 → Spec O Z ,z 0 /mz 0 }n . The
n+1
Yn X X
ηn
Spec R/mn+1
R Spec R = Spec
O Z ,z0 Z
ϕ:
O Z ,z0 = R −→ C[[t]]
ϕ :
O Z ,z0 = R −→
O B,b0 = C[[t]]
The following theorem ensures that a projective scheme with formally smoothable
isolated singularities is smoothable, provided that some local and cohomological
conditions hold.
We now see some conditions that imply that a scheme is not smoothable.
14.2.3 Invariants
In this section we will consider deformations of toric singularities, that is affine toric
varieties. We refer the reader to [34, 41] for an introduction to toric geometry.
If X is an affine toric variety of dimension 2, then X is a cyclic quotient surface
singularity. There is extensive literature about deformations of this kind of singular-
ities, e.g. [19, 27, 64, 88, 93, 94]. In particular, it is known that every affine toric
variety of dimension 2 is smoothable [12].
The study of the deformation theory of affine toric varieties of dimension at least 3
has been initiated by Altmann [4–8]. For example, he computed the tangent space of
the deformation functor of an affine toric variety. We will not write down the explicit
14 On Deformations of Toric Fano Varieties 297
F0 + F1 + · · · + Fr := {v0 + v1 + · · · + vr | v0 ∈ F0 , v1 ∈ F1 , . . . , vr ∈ Fr }.
If F is a lattice polygon, then the affine toric threefold U F has the following
properties:
1. U F has, at most, an isolated singularity if and only if the edges of F are unitary,
i.e. have lattice length 1;
2. U F is Q-factorial if and only if F is a triangle.
Now we provide some examples of lattice polygons and their corresponding toric
Gorenstein affine threefolds.
Theorem 14 (Altmann [7]) Let F be a lattice polygon with unitary edges and let U F
be the corresponding isolated Gorenstein toric singularity of dimension 3. Let R be
the hull of Def U F . Then there exists a one-to-one correspondence between minimal
primes of R and maximal Minkowski decompositions of F. Moreover, if a minimal
prime p ⊂ R corresponds to the maximal Minkowski decomposition F = F0 + F1 +
· · · + Fr , then r = dim R/p.
14 On Deformations of Toric Fano Varieties 299
Corollary 15 Let F be a lattice polygon with unitary edges and let U F be the
associated isolated Gorenstein toric singularity of dimension 3. Then Def U F has an
artinian hull if and only if F is Minkowski indecomposable.
Here we study an explicit example of what has been considered in Sect. 14.3.1. In
the lattice N = Z2 consider the pentagon
1 1 0 −1 0
F = conv , , , , ⊆ N R, (14.2)
0 1 1 0 −1
which is depicted on the left of Fig. 14.2. The toric variety associated to the face
fan of F is the smooth del Pezzo surface of degree 7, which is denoted by dP7 and
is the blow up of P2 in 2 distinct points. The anticanonical map of dP7 is a closed
embedding into P7 .
Now we put the pentagon F at height 1 in the lattice N = N ⊕ Z and we consider
the cone over it:
⎧⎛ ⎞ ⎛ ⎞ ⎛ ⎞ ⎛ ⎞ ⎛ ⎞⎫
⎨ 1 1 0 −1 0 ⎬
σ F = cone ⎝0 ⎠ , ⎝ 1⎠ , ⎝ 1 ⎠ , ⎝ 0 ⎠ , ⎝ −1 ⎠ ⊆ N R ⊕ R.
⎩ ⎭
1 1 1 1 1
The affine toric variety U F = Spec C[σ F∨ ∩ (M ⊕ Z)] is the affine cone over the anti-
canonical embedding of dP7 and has an isolated Gorenstein canonical non-terminal
singularity at the vertex of the cone.
Altmann [7, (9.1)] shows that the hull of Def U F is C[[t1 , t2 ]]/(t12 , t1 t2 ), which is a
line with an embedded point. The reduction of the miniversal deformation, i.e. the
base change to the reduction of the hull, is induced by the unique maximal Minkowski
decomposition of the pentagon F in the following way.
0 −1 0 0 1
F = conv , , + conv , , (14.3)
0 0 −1 0 1
which is illustrated in Fig. 14.2. Following [5, (3.4)], in the lattice Ñ = N ⊕ Ze1 ⊕
Ze2 we construct the cone
⎧⎛ ⎞ ⎛ ⎞ ⎛ ⎞ ⎛ ⎞ ⎛ ⎞⎫
⎪
⎪ 0 −1 0 0 1 ⎪
⎨⎜ ⎟ ⎜ ⎟ ⎜ ⎟ ⎜ ⎟ ⎜ ⎟⎪ ⎬
0 0 −1 0
⎟ , ⎜ ⎟ , ⎜ ⎟ , ⎜ ⎟ , ⎜1⎟ ⊆ ÑR .
σ̃ = cone ⎜ ⎝ 1 ⎠ ⎝ 1 ⎠ ⎝ 1 ⎠ ⎝ 0 ⎠ ⎝0 ⎠⎪
⎪
⎪ ⎪
⎩ ⎭
0 0 0 1 1
Notice that the the first three rays of σ̃ come from the vertices of the first summand
of F in (14.3), whereas the last two rays of σ̃ come from the vertices of the second
summand of F in (14.3). Let Ũ = Spec C[σ̃ ∨ ∩ M̃] be the affine toric variety asso-
ciated to the cone σ̃ , where M̃ denotes the dual of Ñ . One can prove that Ũ has only
an isolated terminal Gorenstein singularity. Let f 1 and f 2 be the regular functions
on Ũ associated to the characters (0, 0, 1, 0) ∈ M̃ and (0, 0, 0, 1) ∈ M̃, respectively.
The variety U F is the zero locus of the function f 1 − f 2 , i.e. we have a cartesian
diagram
UF Ũ (14.4)
π
Spec C A1C
Proposition 16 Let F be the pentagon defined in (14.2) and let U F be the cor-
responding Gorenstein toric threefold singularity. Then the collection of the base
change of π in (14.4) via Spec C[t]/(t n+1 ) → Spec C[t] = A1C for all n is a formal
smoothing of U F . In particular, U F is formally smoothable.
Proof We want to study the closed fibres of π . The fibre over the origin of A1C
is U F . Let us fix λ ∈ C \ {0} and we consider the fibre π −1 (λ) of π over the closed
point (t − λ) of A1C corresponding to λ. We consider the subcone τ1 (resp. τ2 ) of σ̃
that is generated by the first three (resp. last two) rays of σ̃ . We consider the affine
toric variety W j = Spec C[τ j∨ ∩ M̃], for j = 1, 2. We have that W1 and W2 are open
subschemes of Ũ .
14 On Deformations of Toric Fano Varieties 301
We have that W1 is the open subset of Ũ where the function f 2 does not vanish,
i.e. W1 = { f 2 = 0} ⊆ Ũ , and analogously W2 = { f 1 = 0} ⊆ Ũ . It is clear that there
is an isomorphism W1 A3 × Gm with respect to which the function f 1 |W1 becomes
a projection onto a A1 -factor in A3 and the function f 2 |W1 becomes the projection onto
the Gm -factor. There is also an isomorphism W2 A2 × G2m with respect to which
the function f 1 |W2 becomes a projection onto a Gm -factor and the function f 2 |W2
becomes a projection onto an A1 -factor.
Now π −1 (λ) ∩ W1 = { f 1 = f 2 + λ} ∩ W1 is isomorphic to A2 × Gm and π −1 (λ)
∩ W2 = { f 2 = f 1 − λ} ∩ W2 is isomorphic to A1 × G2m . Since λ = 0, it is clear
that π −1 (λ) ⊆ W1 ∪ W2 . Therefore we have proved that π −1 (λ) is smooth.
One can also show that the generic fibre of π is smooth over the generic point
of A1C . In order to prove this, it is enough to base change to the spectrum of the
field C(t) of rational function of A1C and pursue a similar argument, which deals
with toric varieties over the field C(t).
In particular, π is flat of relative dimension 3 and has Cohen–Macaulay fibres. As
in the proof of Lemma 5, from the fact that all non-special fibres of π are smooth we
can deduce that π induces a formal smoothing of U F .
If P is a Fano polytope, then X P is a Fano variety. All toric Fano varieties arise
in this way from a Fano polytope [34, Sect. 8.3]. The variety X P is Gorenstein,
i.e. its (anti)canonical divisor is Cartier, if and only if P is reflexive, i.e. the facets
of P lie on hyperplanes with height 1 with respect to the origin. The maximal toric
affine charts of X P (or equivalently the torus-fixed points of X P ) are in one-to-one
correspondence with the facets of P. If n is the dimension of P, for every 0 ≤ k ≤ n
there is a one-to-one correspondence between the k-dimensional torus-orbits of X P
and the (n − k − 1)-dimensional faces of P.
Fano polytopes of small dimension with specific properties have been classi-
fied [15, 16, 57–59, 65–67, 77, 78, 89, 90, 98]. We refer the reader to [60] for a
survey on the classification of Fano polytopes.
302 A. Petracci
Proof Let U be the affine toric open subscheme of X associated to the cone
spanned by the face F. The condition (1) means that U is smooth in codimension
2. The condition (2) means that U is Q-factorial in codimension 3. Therefore U is
rigid by Proposition 11. The condition (3) implies that U is singular. Therefore, by
Corollary 9, X is not smoothable.
Proof The proof is very similar to the proof of Theorem 18. Let U be the affine
toric open subscheme of X associated to the cone spanned by F. The conditions (1)
and (3) means that U has an isolated singularity. Since P is reflexive, U is Gorenstein.
By Corollary 15, from (2) we deduce that Def U has an artinian hull. Therefore, by
Proposition 8, U is not formally smoothable. By Corollary 9, X is not smoothable.
14 On Deformations of Toric Fano Varieties 303
14.4.3 Rigidity
Here we will see that if a toric Fano variety has very mild singularities then it is rigid.
Proof Set n = dim X . Consider the smooth locus j : U → X . Let D be the toric
boundary of X . The sheaves TX and ( j∗ Un−1 ⊗ O X (D))∨∨ are reflexive on X
and their restrictions to U coincide, because U is smooth and TU is isomorphic
to Un−1 ⊗ ωU∨ . Therefore, since the complement of U has codimension at least 2, by
[49, Proposition 1.6] we have that TX is isomorphic to ( j∗ Un−1 ⊗ O X (D))∨∨ . Since D
is ample, we conclude by Bott–Steenbrink–Danilov vanishing [34, Theorem 9.3.1]
(see also [24, 40, 75]).
This result was originally proved by Bien and Brion [20]. Later de Fernex and
Hacon [38] proved the rigidity of Q-factorial terminal toric Fano varieties. The
following theorem, due to Totaro, is the most general rigidity theorem for toric
Fano varieties of which we are aware.
Theorem 22 (Totaro [96, Theorem 5.1]) A Fano toric variety which is smooth in
codimension 2 and Q-factorial in codimension 3 is rigid.
In Sects. 14.4.4 and 14.4.5 we will study deformations of toric Fanos with isolated
singularities and of dimension 2 or 3.
A del Pezzo surface is a Fano variety of dimension 2. A toric del Pezzo surface is
associated to a Fano polygon, which is a Fano polytope of dimension 2.
304 A. Petracci
Proof Let X be an arbitrary toric del Pezzo surface. It is well known that X
is a normal Cohen–Macaulay projective variety. By Demazure vanishing
[34, Theorem 2.9.3], H2 (O X ) = 0. By Lemma 20, H2 (TX ) = 0. Since X is normal
and of dimension 2, X has isolated singularities. By Theorem 7 it is enough to check
that the singularities of X are formally smoothable.
The singularities of X are cyclic quotient surface singularities. This kind of sin-
gularities is always smoothable; indeed, it is enough to pick the Artin component of
the base of the miniversal deformation [12].
Remark 25 When the canonical divisor of a normal variety X is not Cartier, flat
deformations of X are too wild for hoping to study moduli of varieties. For a
normal Q-Gorenstein non-Gorenstein variety X one should consider a subfunctor
of Def X which is made up of the deformations of X in which the canonical divi-
sor deforms well. This is the theory of Q-Gorenstein deformations, developed by
Kollár–Shepherd-Barron [64] (see also [1, 9, 45, 68]).
In the context of Q-Gorenstein deformations the analogous statement of
Theorem 24 is false: there exist non-Gorenstein toric del Pezzo surfaces which can-
not be deformed via Q-Gorenstein deformations to a smooth del Pezzo surface, e.g.
the weighted projective space P(1, 1, 3). Nonetheless, it is true that for Q-Gorenstein
deformations of del Pezzo surfaces there are no local-to-global obstructions [2,
Lemma 6]. Therefore, a del Pezzo surface is Q-Gorenstein smoothable if and only
if its singularities are Q-Gorenstein smoothable.
Since the main focus of this note is the study of deformations of Gorenstein toric
Fano threefolds, we will omit to discuss the theory of Q-Gorenstein deformations.
We refer the reader to [46, 82] for the study of toric del Pezzo surfaces which have Q-
Gorenstein smoothings.
Proof By Example 13 we have that the singularities of X are at most ordinary double
points (i.e. nodes). These singularities are formally smoothable. By Theorem 26 we
conclude.
The proof of this corollary is essentially a specific case of [39, Sect. 4.a]. The
corollary could have been deduced also from a more general result by Namikawa
according to which every Fano threefold with Gorenstein terminal singularities is
smoothable [76]. The smooth Fano threefolds which are the smoothings of the toric
Fano threefold appearing in Corollary 27 have been studied by Galkin [42].
For d ∈ {6, 7}, let dPd be the smooth del Pezzo surface of degree d; it is toric. The
complete anticanonical linear system on dPd induces a closed embedding dPd → Pd .
We consider the projective cone C(dPd ) ⊆ Pd+1 over this embedding; we have
that C(dPd ) is a toric Fano threefold with a Gorenstein canonical non-terminal iso-
lated singularity. In Sect. 14.4.6 we will see that C(dP7 ) is smoothable. In [80] it is
shown that C(dP6 ) has two smoothings (see also [56, Example 3.3]).
Here we study the deformations of an explicit toric Fano threefold with an isolated
Gorenstein non-terminal singularity.
Fix the lattice N = Z2 . Consider the pentagon F ⊆ N R defined in (14.2), imagine
to put it into the plane N R × {1} in N R ⊕ R R3 , and create the pyramid over it
with apex at the point (0, 0, −1): this is the polytope
⎧⎛ ⎞ ⎛ ⎞ ⎛ ⎞ ⎛ ⎞ ⎛ ⎞ ⎛ ⎞⎫
⎨ 1 1 0 −1 0 0 ⎬
P = conv ⎝0⎠ , ⎝1⎠ , ⎝1⎠ , ⎝ 0 ⎠ , ⎝−1⎠ , ⎝ 0 ⎠ (14.5)
⎩ ⎭
1 1 1 1 1 −1
in the lattice N ⊕ Z and is depicted in Fig. 14.3. It is clear that P is a Fano polytope.
Let X be the toric variety associated to the spanning fan of P. Then X is the
projective cone over the anticanonical embedding of the smooth del Pezzo surface
of degree 7. The affine toric variety U F considered in Sect. 14.3.2 is the affine open
toric subscheme of X associated to the pentagonal facet F of P. We have that X is a
Fano threefold with an isolated non-terminal canonical Gorenstein singularity at the
vertex of the cone.
Proposition 28 Let X be the toric Fano threefold associated to the polytope P in
(14.5), i.e. X is the projective cone over the anticanonical embedding of the smooth
del Pezzo surface of degree 7. Then X is smoothable and can be deformed to the
smooth Fano threefold P(OP2 ⊕ OP2 (1)).
Proof By Proposition 16, X has an isolated singularity which is formally smoothable.
By Theorem 26 we know that X is smoothable. We need to know to which smooth
Fano threefold X can be deformed.
From toric geometry [34, Theorem 13.4.3], we have that the anticanonical
degree (−K X )3 is the normalised volume of the polar polytope of P, which is 56 in
this case. Since X has Gorenstein canonical singularities, by Proposition 10 we have
that the anticanonical degree is preserved in the smoothing. By inspecting the list
of smooth Fano threefolds (see [53, 54, 72–74] or [55, Sect. 12]), there is a unique
smooth Fano threefold of anticanonical degree 56, namely P(OP2 ⊕ OP2 (1)).
In addition to the result of Proposition 19, here we present another obstruction for
the smoothability of a toric Fano threefold with Gorenstein singularities.
Theorem 29 ([79]) Let N be a lattice of rank 3, let M = HomZ (N , Z), let ·, · : M
× N → Z be the duality pairing, let P be a reflexive polytope in N , and let X be the
toric Fano threefold associated to the spanning fan of P. Assume that there are two
adjacent facets F0 and F1 of P such that:
1. both F0 and F1 are An -triangles for some integer n ≥ 1 (see the definition in
Example 13);
2. F0 ∩ F1 is a segment with n + 2 lattice points;
3. w1 , v0 = 0, where w1 ∈ M is such that F1 ⊆ {v ∈ NR | w1 , v = 1} and v0 ∈
N is the vertex of F0 which does not lie on the segment F0 ∩ F1 .
Then X is not smoothable.
With the terminology of [79], the two triangles F0 and F1 are called “two adjacent
almost-flat An -triangles”.
Proof (Sketch of the proof of Theorem 29) We refer the reader to [79] for all the
details missing here. Let Ui be the toric open affine subscheme of X associated to
the facet Fi , for each i = 0, 1. Set U = U0 ∪ U1 .
14 On Deformations of Toric Fano Varieties 307
One can show that U admits an An -bundle structure over P1 . More precisely, one
can construct a toric morphism π : U → P1 such that, for each i = 0, 1, if Vi denotes
the ith standard affine chart of P1 then π −1 (Vi ) = Ui and the restriction π |Ui : Ui →
Vi is the projection Spec C[x, y, z, w]/(x y − z n+1 ) → Spec C[w]. This An -bundle
may be non trivial, depending on the relative position of the two triangles F0 and F1 .
Set d = w1 , v0 . By [79, Proposition 3.5] there exists an isomorphism of coherent
sheaves on P1 :
π∗ ExtO1 U (U , OU ) OP1 (− jd − j).
2≤ j≤n+1
Since d = 0, the sheaf on the right is a direct sum of negative line bundles on P1 ,
hence we have H0 (U, ExtO1 U (U , OU )) = 0. By Corollary 9, X is not smoothable.
With the same technique of the theorem above one can also construct some rigid
toric Fano threefolds with only c A1 -singularities (see [79, Theorem 1.2]). This refutes
a conjecture of Prokhorov [85] according to which every Fano threefold with com-
pound Du Val singularities is smoothable.
Here we briefly collect some other results on deformations and smoothings of toric
Fano varieties. Most of these results have been motivated by Mirror Symmetry for
Fano varieties (see [2, 3, 26, 30, 31, 61, 86, 87]).
By analysing cluster transformations of tori, Akhtar–Coates–Galkin–Kasprzyk [3]
have introduced the notion of mutation of Fano polytopes. A mutation is a combi-
natorial procedure that, under certain conditions, transforms a Fano polytope P into
another Fano polytope P . Ilten [51] has proved that mutations of Fano polytopes
induce deformations of the corresponding toric Fano varieties; more precisely, if P
and P are related via a mutation, then he has constructed a flat family over P1 such
that the fibre over 0 is X P and the fibre over ∞ is X P .
Ilten, Lewis and Przyjalkowski [52] have constructed toric degenerations of
smooth Fano threefolds with Picard rank 1.
Christophersen and Ilten [29] have constructed degenerations of smooth Fano
threefolds of low degree to certain unobstructed Fano Stanley–Reisner schemes.
Since these unobstructed Fano Stanley–Reisner schemes are also degenerations of
singular toric Fano varieties, this implies the following result.
variety X P into a bigger toric variety Y . Often X P is a complete intersection in the Cox
coordinates of Y , therefore it is easy to construct embedded deformations of X P in Y .
In many cases this produces smoothings of X P . For instance, Cavey and Prince [25]
have successfully applied the scaffolding method to construct deformations of toric
del Pezzo surfaces to del Pezzo surfaces with a single k1 (1, 1) singularity.
Moreover, Prince [83] has found necessary and sufficient conditions in order to
have that the ambient toric variety Y is smooth: this is the notion of cracked polytope.
He has also found a sufficient condition for a smoothing of X P to exist inside Y . Via
the scaffolding method and cracked polytopes, in [84] he constructs a degeneration
of each smooth Fano threefold with very ample anticanonical bundle and Picard
rank ≥ 2 to a Gorenstein toric Fano threefold.
Let S denote the set of all reflexive polytopes of dimension 3, i.e. the set of positive
integers not greater than 4319. We have:
Below we write down the elements of most of the sets mentioned above.
Ssmooth = {1, 5, 6, 7, 8, 25, 26, 27, 28, 29, 30, 31, 82, 83, 84, 85, 219, 220}
Sisol = {3, 4, 11, 12, 17, 21, 22, 23, 24, 42, 48, 49, 50, 51, 54, 68, 69, 70, 71, 72,
73, 74, 75, 76, 77, 78, 79, 80, 81, 155, 156, 158, 159, 160, 167, 168, 170, 177, 187,
188, 198, 199, 200, 201, 202, 203, 204, 205, 206, 207, 208, 209, 210, 211, 212, 213,
214, 215, 216, 217, 218, 360, 363, 364, 365, 366, 376, 377, 378, 380, 385, 403, 410,
411, 412, 413, 414, 415, 416, 417, 418, 419, 420, 421, 422, 423, 424, 425, 426, 427,
686, 688, 689, 692, 693, 694, 695, 696, 707, 710, 725, 729, 730, 731, 732, 733, 734,
735, 736, 737, 738, 739, 740, 741, 1085, 1086, 1087, 1091, 1092, 1093, 1109, 1110,
1111, 1112, 1113, 1114, 1517, 1518, 1519, 1524, 1528, 1529, 1530, 1941, 1943,
2355, 2356}
Snodes = {4, 21, 22, 23, 24, 68, 69, 70, 71, 72, 73, 74, 75, 76, 77, 78, 79, 80, 81,
198, 199, 200, 201, 202, 203, 204, 205, 206, 207, 208, 209, 210, 211, 212, 213, 214,
215, 216, 217, 218, 410, 411, 412, 413, 414, 415, 416, 417, 418, 419, 420, 421, 422,
423, 424, 425, 426, 427, 729, 730, 731, 732, 733, 734, 735, 736, 737, 738, 739, 740,
741, 1109, 1110, 1111, 1112, 1113, 1114, 1528, 1529, 1530, 1943, 2356}
Slow = {1946, 2711, 2756, 2817, 3043, 3051, 3053, 3079, 3314, 3319, 3329,
3331, 3349, 3350, 3390, 3393, 3406, 3416, 3447, 3452, 3453, 3505, 3573, 3620,
3625, 3626, 3667, 3683, 3702, 3727, 3728, 3731, 3733, 3735, 3736, 3738, 3739,
3740, 3756, 3760, 3762, 3777, 3790, 3791, 3792, 3795, 3796, 3844, 3845, 3846,
3848, 3853, 3857, 3868, 3869, 3874, 3875, 3879, 3901, 3903, 3922, 3923, 3927,
3928, 3933, 3936, 3937, 3938, 3946, 3962, 3964, 3965, 3966, 3967, 3981, 3983,
3984, 3985, 3991, 3995, 4003, 4004, 4005, 4006, 4007, 4022, 4023, 4024, 4027,
4031, 4032, 4041, 4042, 4043, 4044, 4056, 4058, 4059, 4060, 4070, 4074, 4075,
4076, 4080, 4088, 4092, 4094, 4095, 4102, 4104, 4117, 4118, 4119, 4122, 4124,
4131, 4132, 4133, 4134, 4135, 4143, 4144, 4145, 4149, 4159, 4160, 4161, 4167,
4168, 4169, 4170, 4179, 4180, 4181, 4182, 4183, 4184, 4186, 4190, 4191, 4194,
4200, 4202, 4203, 4205, 4206, 4214, 4215, 4216, 4217, 4218, 4219, 4220, 4225,
4228, 4229, 4231, 4232, 4233, 4235, 4236, 4238, 4239, 4241, 4244, 4245, 4246,
4247, 4249, 4250, 4251, 4252, 4254, 4255, 4256, 4258, 4260, 4261, 4263, 4267,
4268, 4269, 4270, 4272, 4273, 4275, 4278, 4280, 4281, 4282, 4284, 4285, 4286,
4287, 4288, 4290, 4291, 4292, 4293, 4294, 4295, 4297, 4298, 4299, 4300, 4301,
4303, 4304, 4307, 4308, 4309, 4310, 4311, 4312, 4313, 4314, 4315, 4317, 4318,
4319}
Sindec = {3, 12, 17, 32, 38, 48, 49, 51, 54, 88, 91, 94, 98, 99, 100, 101, 102, 103,
105, 115, 119, 121, 134, 137, 138, 141, 142, 155, 158, 159, 170, 188, 228, 235, 239,
242, 243, 247, 248, 252, 254, 256, 260, 262, 265, 271, 278, 293, 294, 298, 299, 301,
317, 318, 330, 351, 353, 360, 378, 380, 438, 439, 440, 443, 445, 455, 468, 480, 491,
310 A. Petracci
492, 493, 497, 501, 502, 515, 525, 526, 529, 530, 532, 539, 541, 543, 546, 550, 553,
562, 570, 575, 604, 608, 609, 614, 620, 645, 650, 660, 663, 688, 744, 752, 753, 754,
756, 760, 774, 775, 776, 780, 784, 790, 791, 792, 800, 834, 841, 844, 845, 852, 856,
859, 864, 866, 887, 900, 908, 912, 914, 923, 935, 963, 979, 990, 991, 1012, 1019,
1020, 1130, 1151, 1154, 1183, 1199, 1204, 1205, 1208, 1215, 1218, 1220, 1261,
1275, 1277, 1283, 1299, 1302, 1309, 1311, 1352, 1370, 1384, 1397, 1547, 1585,
1598, 1631, 1636, 1638, 1679, 1683, 1687, 1693, 1728, 1750, 1751, 1777, 1791,
1992, 2014, 2046, 2047, 2050, 2051, 2080, 2081, 2084, 2096, 2124, 2129, 2379,
2404, 2425, 2427, 2455, 2456, 2716, 2750, 2751, 2755}
Saft = {15, 16, 36, 41, 45, 53, 58, 59, 61, 65, 66, 102, 105, 110, 111, 112, 113,
116, 117, 124, 125, 128, 135, 141, 142, 144, 146, 147, 148, 149, 152, 162, 172, 179,
183, 189, 192, 193, 197, 230, 236, 244, 248, 261, 268, 271, 272, 277, 278, 279, 280,
281, 282, 286, 288, 290, 292, 302, 310, 324, 325, 327, 331, 332, 333, 334, 335, 337,
340, 343, 347, 349, 351, 355, 356, 358, 361, 362, 386, 399, 400, 407, 443, 445, 448,
452, 453, 456, 457, 463, 467, 487, 490, 496, 497, 499, 501, 502, 505, 507, 508, 509,
511, 512, 516, 523, 540, 545, 550, 563, 569, 577, 579, 581, 582, 583, 594, 599, 600,
601, 605, 606, 617, 629, 633, 658, 670, 671, 672, 674, 679, 682, 687, 705, 760, 764,
770, 771, 780, 781, 786, 787, 792, 797, 799, 809, 811, 812, 815, 816, 824, 859, 865,
868, 873, 875, 878, 883, 884, 889, 891, 892, 893, 894, 895, 902, 905, 929, 956, 960,
965, 987, 1003, 1004, 1006, 1011, 1021, 1038, 1045, 1051, 1156, 1160, 1168, 1175,
1177, 1199, 1203, 1209, 1216, 1217, 1225, 1232, 1234, 1251, 1252, 1253, 1255,
1256, 1260, 1262, 1265, 1275, 1286, 1287, 1293, 1300, 1305, 1308, 1324, 1327,
1351, 1371, 1383, 1398, 1533, 1545, 1550, 1551, 1554, 1561, 1579, 1589, 1613,
1614, 1615, 1620, 1637, 1638, 1656, 1665, 1666, 1671, 1686, 1690, 1693, 1697,
1711, 1747, 1748, 1760, 1763, 1989, 2000, 2001, 2027, 2045, 2051, 2052, 2068,
2071, 2072, 2076, 2084, 2096, 2098, 2102, 2379, 2380, 2385, 2403, 2405, 2423,
2424, 2425, 2427, 2738, 2777, 2778, 2792, 3047, 3057, 3063, 3064}
We have |Sindec ∪ Saft | = 442. Therefore there exist at least 442 non-smoothable
toric Fano threefolds with Gorenstein singularities.
References
1. Abramovich, D., Hassett, B.: Stable varieties with a twist. In: Classification of algebraic vari-
eties, EMS Series of Congress Reports, pp. 1–38. European Mathematical Society, Zürich
(2011)
2. Akhtar, M., Coates, T., Corti, A., Heuberger, L., Kasprzyk, A.M., Oneto, A., Petracci, A.,
Prince, T., Tveiten, K.: Mirror symmetry and the classification of orbifold del Pezzo surfaces.
Proc. Amer. Math. Soc. 144(2), 513–527 (2016)
14 On Deformations of Toric Fano Varieties 311
3. Akhtar, M., Coates, T., Galkin, S., Kasprzyk, A.M.: Minkowski polynomials and mutations.
SIGMA Symmetry Integr. Geom. Methods Appl. 8, Paper 094, 17 (2012)
4. Altmann, K.: Computation of the vector space T 1 for affine toric varieties. J. Pure Appl. Algebra
95(3), 239–259 (1994)
5. Altmann, K.: Minkowski sums and homogeneous deformations of toric varieties. Tohoku Math.
J. (2) 47(2), 151–184 (1995)
6. Altmann, K.: Infinitesimal deformations and obstructions for toric singularities. J. Pure Appl.
Algebra 119(3), 211–235 (1997)
7. Altmann, K.: The versal deformation of an isolated toric Gorenstein singularity. Invent. Math.
128(3), 443–479 (1997)
8. Altmann, K.: One parameter families containing three-dimensional toric-Gorenstein singular-
ities. In: Explicit birational geometry of 3-folds. London Mathematical Society Lecture Note
series, vol. 281, pp. 21–50. Cambridge University Press, Cambridge (2000)
9. Altmann, K., Kollár, J.: The dualizing sheaf on first-order deformations of toric surface singu-
larities. J. Reine Angew. Math. 753, 137–158 (2019)
10. Artin, M.: Algebraic approximation of structures over complete local rings. Inst. Hautes Études
Sci. Publ. Math. 36, 23–58 (1969)
11. Artin, M.: Algebraization of formal moduli. I. In: Global Analysis (Papers in Honor of K.
Kodaira), pp. 21–71. University Tokyo Press, Tokyo (1969)
12. Artin, M.: Algebraic construction of Brieskorn’s resolutions. J. Algebra 29, 330–348 (1974)
13. Artin, M.: Lectures on Deformations of Singularities. Tata Institute of Fundamental Research,
Bombay (1976)
14. The Stacks project (2018). https://stacks.math.columbia.edu
15. Batyrev, V.V.: Toric Fano threefolds. Izv. Akad. Nauk SSSR Ser. Mat. 45(4), 704–717, 927
(1981)
16. Batyrev, V.V.: On the classification of toric Fano 4-folds. pp. 1021–1050 (1999). Algebraic
geometry, 9
17. Batyrev, V.V.: Toric degenerations of Fano varieties and constructing mirror manifolds. In: The
Fano Conference, pp. 109–122. University Torino, Turin (2004)
18. Batyrev, V.V., Ciocan-Fontanine, I., Kim, B., van Straten, D.: Mirror symmetry and toric degen-
erations of partial flag manifolds. Acta Math. 184(1), 1–39 (2000)
19. Behnke, K., Riemenschneider, O.: Quotient surface singularities and their deformations. In:
Singularity Theory (Trieste, 1991), pp. 1–54. World Scientific Publishing, River Edge (1995)
20. Bien, F., Brion, M.: Automorphisms and local rigidity of regular varieties. Compositio Math.
104(1), 1–26 (1996)
21. Birkar, C.: Singularities of linear systems and boundedness of Fano varieties. Ann. of Math.
(2) 193(2), 347–405 (2021)
22. Bosma, W., Cannon, J., Playoust, C.: The Magma algebra system. I. The user language. pp.
235–265 (1997). Computational algebra and number theory (London, 1993)
23. Brown, G., Kasprzyk, A.M.: Graded Ring Database. http://www.grdb.co.uk
24. Buch, A., Thomsen, J.F., Lauritzen, N., Mehta, V.: The Frobenius morphism on a toric variety.
Tohoku Math. J. (2) 49(3), 355–366 (1997)
25. Cavey, D., Prince, T.: Del Pezzo surfaces with a single 1/k(1, 1) singularity. J. Math. Soc.
Japan 72(2), 465–505 (2020)
26. Cheltsov, I., Katzarkov, L., Przyjalkowski, V.: Birational geometry via moduli spaces. In: Bira-
tional Geometry, Rational Curves, and Arithmetic, Simons Symposia, pp. 93–132. Springer,
Cham (2013)
27. Christophersen, J.A.: On the components and discriminant of the versal base space of cyclic
quotient singularities. In: Singularity theory and its applications, Part I (Coventry, 1988/1989).
Lecture Notes in Math., vol. 1462, pp. 81–92. Springer, Berlin (1991)
28. Christophersen, J.A., Ilten, N.: Hilbert schemes and toric degenerations for low degree Fano
threefolds. J. Reine Angew. Math. 717, 77–100 (2016)
29. Christophersen, J.A., Ilten, N.O.: Degenerations to unobstructed Fano Stanley-Reisner
schemes. Math. Z. 278(1–2), 131–148 (2014)
312 A. Petracci
30. Coates, T., Corti, A., Galkin, S., Golyshev, V., Kasprzyk, A.M.: Mirror symmetry and Fano
manifolds. In: European Congress of Mathematics, pp. 285–300. European Mathematical Soci-
ety, Zürich (2013)
31. Coates, T., Corti, A., Galkin, S., Kasprzyk, A.M.: Quantum periods for 3-dimensional Fano
manifolds. Geom. Topol. 20(1), 103–256 (2016)
32. Coates, T., Kasprzyk, A., Prince, T.: Laurent inversion. Pure Appl. Math. Q. 15(4), 1135–1179
(2019)
33. Coates, T., Kasprzyk, A.M.: Code repository (2019). https://bitbucket.org/fanosearch/magma-
core
34. Cox, D.A., Little, J.B., Schenck, H.K.: Toric varieties. Graduate Studies in Mathematics, vol.
124. American Mathematical Society, Providence (2011)
35. Eisenbud, D.: Commutative algebra. Graduate Texts in Mathematics, vol. 150. Springer, New
York (1995). With a view toward algebraic geometry
36. Fantechi, B., Göttsche, L.: Local properties and Hilbert schemes of points. In: Fundamental
Algebraic Geometry. Mathematical Surveys and Monographs, vol. 123, pp. 139–178. American
Mathematical Society, Providence (2005)
37. Fantechi, B., Manetti, M.: Obstruction calculus for functors of Artin rings. I. J. Algebra 202(2),
541–576 (1998)
38. de Fernex, T., Hacon, C.D.: Deformations of canonical pairs and Fano varieties. J. Reine Angew.
Math. 651, 97–126 (2011)
39. Friedman, R.: Simultaneous resolution of threefold double points. Math. Ann. 274(4), 671–689
(1986)
40. Fujino, O.: Multiplication maps and vanishing theorems for toric varieties. Math. Z. 257(3),
631–641 (2007)
41. Fulton, W.: Introduction to toric varieties. Annals of Mathematics Studies, vol. 131. Princeton
University Press, Princeton (1993). The William H. Roever Lectures in Geometry
42. Galkin, S.: Small toric degenerations of Fano threefolds (2018). arXiv:1809.02705 [math.AG]
43. Gross, M., Siebert, B.: From real affine geometry to complex geometry. Ann. of Math. (2)
174(3), 1301–1428 (2011)
44. Grothendieck, A.: Éléments de géométrie algébrique. IV. Étude locale des schémas et des
morphismes de schémas. III. Inst. Hautes Études Sci. Publ. Math. (28), 255 (1966)
45. Hacking, P.: Compact moduli of plane curves. Duke Math. J. 124(2), 213–257 (2004)
46. Hacking, P., Prokhorov, Y.: Smoothable del Pezzo surfaces with quotient singularities. Compos.
Math. 146(1), 169–192 (2010)
47. Hartshorne, R.: Residues and duality. Lecture notes of a seminar on the work of A.
Grothendieck, given at Harvard 1963/64. With an appendix by P. Deligne. Lecture Notes in
Mathematics, No. 20. Springer, Berlin (1966)
48. Hartshorne, R.: Algebraic Geometry. Springer, New York (1977). Graduate Texts in Mathe-
matics, No. 52
49. Hartshorne, R.: Stable reflexive sheaves. Math. Ann. 254(2), 121–176 (1980)
50. Hartshorne, R.: Deformation theory. Graduate Texts in Mathematics, vol. 257. Springer, New
York (2010)
51. Ilten, N.O.: Mutations of Laurent polynomials and flat families with toric fibers. SIGMA
Symmetry Integr. Geom. Methods Appl. 8, Paper 047, 7 (2012)
52. Ilten, N.O., Lewis, J., Przyjalkowski, V.: Toric degenerations of Fano threefolds giving weak
Landau-Ginzburg models. J. Algebra 374, 104–121 (2013)
53. Iskovskih, V.A.: Fano threefolds. I. Izv. Akad. Nauk SSSR Ser. Mat. 41(3), 516–562, 717
(1977)
54. Iskovskih, V.A.: Fano threefolds. II. Izv. Akad. Nauk SSSR Ser. Mat. 42(3), 506–549 (1978)
55. Iskovskikh, V.A., Prokhorov, Y.G.: Fano varieties. In: Algebraic Geometry, V, Encylopaedia
of Mathematical Sciences, vol. 47, pp. 1–247. Springer, Berlin (1999)
56. Jahnke, P., Radloff, I.: Terminal Fano threefolds and their smoothings. Math. Z. 269(3–4),
1129–1136 (2011)
14 On Deformations of Toric Fano Varieties 313
57. Kasprzyk, A.M.: Toric Fano three-folds with terminal singularities. Tohoku Math. J. (2) 58(1),
101–121 (2006)
58. Kasprzyk, A.M.: Canonical toric Fano threefolds. Canad. J. Math. 62(6), 1293–1309 (2010)
59. Kasprzyk, A.M., Kreuzer, M., Nill, B.: On the combinatorial classification of toric log del
Pezzo surfaces. LMS J. Comput. Math. 13, 33–46 (2010)
60. Kasprzyk, A.M., Nill, B.: Fano polytopes. In: Strings. Gauge Fields, and the Geometry Behind,
pp. 349–364. World Scientific Publishing, Hackensack (2013)
61. Katzarkov, L., Przyjalkowski, V.: Landau-Ginzburg models—old and new. In: Proceedings of
the Gökova Geometry-Topology Conference 2011, pp. 97–124. International Press, Somerville,
MA (2012)
62. Kollár, J., Miyaoka, Y., Mori, S.: Rational connectedness and boundedness of Fano manifolds.
J. Differ. Geom. 36(3), 765–779 (1992)
63. Kollár, J., Mori, S.: Birational geometry of algebraic varieties. Cambridge Tracts in Mathemat-
ics, vol. 134. Cambridge University Press, Cambridge (1998). With the collaboration of C. H.
Clemens and A. Corti, Translated from the 1998 Japanese original
64. Kollár, J., Shepherd-Barron, N.I.: Threefolds and deformations of surface singularities. Invent.
Math. 91(2), 299–338 (1988)
65. Kreuzer, M., Nill, B.: Classification of toric Fano 5-folds. Adv. Geom. 9(1), 85–97 (2009)
66. Kreuzer, M., Skarke, H.: Classification of reflexive polyhedra in three dimensions. Adv. Theor.
Math. Phys. 2(4), 853–871 (1998)
67. Kreuzer, M., Skarke, H.: Complete classification of reflexive polyhedra in four dimensions.
Adv. Theor. Math. Phys. 4(6), 1209–1230 (2000)
68. Lee, Y., Nakayama, N.: Grothendieck duality and Q-Gorenstein morphisms. Publ. Res. Inst.
Math. Sci. 54(3), 517–648 (2018)
69. Li, C., Wang, X., Xu, C.: On the proper moduli spaces of smoothable Kähler-Einstein Fano
varieties. Duke Math. J. 168(8), 1387–1459 (2019)
70. Manetti, M.: Differential graded Lie algebras and formal deformation theory. In: Algebraic
Geometry—Seattle 2005. Part 2. Proceedings of Symposia in Pure Mathematics, vol. 80, pp.
785–810. American Mathematical Society, Providence (2009)
71. Mavlyutov, A.R.: Deformations of toric varieties via Minkowski sum decompositions of poly-
hedral complexes (2009). arXiv:0902.0967 [math.AG]
72. Mori, S., Mukai, S.: Classification of Fano 3-folds with B2 ≥ 2. Manuscripta Math. 36(2),
147–162 (1981/82)
73. Mori, S., Mukai, S.: On Fano 3-folds with B2 ≥ 2. In: Algebraic Varieties and Analytic Varieties
(Tokyo, 1981). Advanced Studies in Pure Mathematics, vol. 1, pp. 101–129. North-Holland,
Amsterdam (1983)
74. Mori, S., Mukai, S.: Erratum: “Classification of Fano 3-folds with B2 ≥ 2” [Manuscripta Math.
36(2), 147–162 (1981/82); MR0641971 (83f:14032)]. Manuscripta Math. 110(3), 407 (2003)
75. Mustaţă, M.: Vanishing theorems on toric varieties. Tohoku Math. J. (2) 54(3), 451–470 (2002)
76. Namikawa, Y.: Smoothing Fano 3-folds. J. Algebraic Geom. 6(2), 307–324 (1997)
77. Nill, B., Øbro, M.: Q-factorial Gorenstein toric Fano varieties with large Picard number. Tohoku
Math. J. (2) 62(1), 1–15 (2010)
78. Øbro, M.: An algorithm for the classification of smooth Fano polytopes (2007).
arXiv:0704.0049 [math.CO]
79. Petracci, A.: Some examples of non-smoothable Gorenstein Fano toric threefolds. Math. Z.
295(1–2), 751–760 (2020)
80. Petracci, A.: An example of mirror symmetry for Fano threefolds. In: Birational geometry and
moduli spaces. Springer INdAM Serirs, vol. 39, pp. 173–188. Springer, Cham ([2020] © 2020)
81. Petracci, A.: Homogeneous deformations of toric pairs. Manuscripta Math. 166(1–2), 37–72
(2021)
82. Prince, T.: Smoothing toric Fano surfaces using the Gross-Siebert algorithm. Proc. Lond. Math.
Soc. (3) 117(3), 617–660 (2018)
83. Prince, T.: Cracked polytopes and Fano toric complete intersections. Manuscripta Math. 163(1–
2), 165–183 (2020)
314 A. Petracci
84. Prince, T.: From cracked polytopes to Fano threefolds. Manuscripta Math. 164(1–2), 267–320
(2021)
85. Prokhorov, Y.G.: The degree of Fano threefolds with canonical Gorenstein singularities. Mat.
Sb. 196(1), 81–122 (2005)
86. Przhiyalkovskiı̆, V.V.: Weak Landau-Ginzburg models of smooth Fano threefolds. Izv. Ross.
Akad. Nauk Ser. Mat. 77(4), 135–160 (2013)
87. Przyjalkowski, V.: On Landau-Ginzburg models for Fano varieties. Commun. Number Theory
Phys. 1(4), 713–728 (2007)
88. Riemenschneider, O.: Deformationen von Quotientensingularitäten (nach zyklischen Grup-
pen). Math. Ann. 209, 211–248 (1974)
89. Sato, H.: Toward the classification of higher-dimensional toric Fano varieties. Tohoku Math.
J. (2) 52(3), 383–413 (2000)
90. Sato, H.: Smooth toric Fano five-folds of index two. Proc. Japan Acad. Ser. A Math. Sci. 82(7),
106–110 (2006)
91. Schlessinger, M.: Functors of Artin rings. Trans. Amer. Math. Soc. 130, 208–222 (1968)
92. Sernesi, E.: Deformations of algebraic schemes, Grundlehren der Mathematischen Wis-
senschaften [Fundamental Principles of Mathematical Sciences], vol. 334. Springer, Berlin
(2006)
93. Stevens, J.: On the versal deformation of cyclic quotient singularities. In: Singularity Theory
and Its Applications, Part I (Coventry, 1988/1989). Lecture Notes in Mathematics, vol. 1462,
pp. 302–319. Springer, Berlin (1991)
94. Stevens, J.: The versal deformation of cyclic quotient singularities. In: Deformations of Sur-
face Singularities. Bolyai Society Mathematical Studies, vol. 23, pp. 163–201. János Bolyai
Mathematical Society, Budapest (2013)
95. Talpo, M., Vistoli, A.: Deformation theory from the point of view of fibered categories. In:
Handbook of Moduli. Vol. III, Advanced Lectures in Mathematics (ALM), vol. 26, pp. 281–397.
International Press, Somerville (2013)
96. Totaro, B.: Jumping of the nef cone for Fano varieties. J. Algebraic Geom. 21(2), 375–396
(2012)
97. Vistoli, A.: The deformation theory of local complete intersections (1997).
arXiv:alg-geom/9703008
98. Watanabe, K., Watanabe, M.: The classification of Fano 3-folds with torus embeddings. Tokyo
J. Math. 5(1), 37–48 (1982)
99. Ziegler, G.M.: Lectures on Polytopes. Graduate Texts in Mathematics, vol. 152. Springer, New
York (1995)
Chapter 15
Polygons of Finite Mutation Type
Thomas Prince
Abstract We classify Fano polygons with finite mutation class. This classification
exploits a correspondence between Fano polygons and cluster algebras, refining the
notion of singularity content due to Akhtar and Kasprzyk. We also introduce examples
of cluster algebras associated to Fano polytopes in dimensions greater than two.
15.1 Introduction
T. Prince (B)
Mathematical Institute, University of Oxford, Woodstock Road, Oxford OX2 6GG, UK
e-mail: [email protected]
content of a gauge theory arising on a stack of D3-branes probing the toric Calabi–
Yau singularity, (see for example [1, 5, 10, 15, 21, 22, 29] for a selection of the
literature on this subject). The construction of a quiver (and cluster algebra) from a
polygon has also been used by Gross–Hacking–Keel [20] in the study of associated
log Calabi–Yau varieties, and to study the derived category of the toric variety, or the
associated local toric Calabi–Yau as pursued, for example, in [7, 23, 24, 31, 32]. In
each setting the basic construction is the same, and we recall the version relevant to
our applications in Sect. 15.3.
Our main result, Theorem 30, is a classification of the mutation classes of polygons
which contain only finitely many polygons. This parallels a finite type result of
Mandel [30], for rank two cluster varieties. In particular we see that finite mutation
classes of polygons fall into four types An1 , for n ∈ Z≥0 , A2 , A3 , and D4 .
There is a close connection between mutation classes of Fano polygons and Q-
Gorenstein deformations of the corresponding toric varieties which is described in
detail in [2]. Following these ideas we predict the existence of a finite type param-
eter space for these deformations, together with a boundary stratification such that
each 0-stratum corresponds to a polygon in the given mutation class, and the 1-strata
corresponds to the mutation families constructed by Ilten [26].
While our main result applies in dimension two, we note that polytope mutation
is defined in all dimensions, and the construction of a quiver and cluster algebra
we provide applies to ‘compatible collection’ of mutations in any dimension, see
Definition 16. This definition is, unfortunately, less well behaved in dimensions
greater than two, but we provide an example indicating that polytope mutation can
detect known examples of cluster structures appearing on linear sections of Grass-
mannians of planes. We expect this to extend to a wide variety of other cluster
structures found in Fano manifolds and their mirror manifolds.
We devote this section to fixing the various conventions and notation, as well as
recalling the basic definitions. We recall the definition of cluster algebra, and in
order to address both geometric and combinatorial applications we shall adapt our
treatment from the work of Fomin–Zelevinsky [13], and the work of Fock–Goncharov
[11] and Gross–Hacking–Keel [20]. We first fix the following data:
1. N , a fixed lattice with skew-symmetric form {−, −} : N × N → Z;
2. a saturated sublattice Nu f ⊆ N , the unfrozen sublattice;
3. an index set I , |I | = rk(N ) together with a subset Iu f ⊆ I such that |Iu f | =
rk(Nu f ). For later convenience we set n := |Iu f |.
Remark 1 The requirement that the form is integral is not necessary, but is suffi-
ciently general for our applications and simplifies the exposition considerably.
Definition 2 A (labelled) seed is a pair s = (E, C), where:
1. E is a basis of N indexed by I , such that the subset indexed by Iu f is a basis
of Nu f ;
15 Polygons of Finite Mutation Type 317
Recall that the matrix B := (bkl )k,l∈Iu f is typically referred to as the exchange
matrix of the seed.
Definition 5 A cluster algebra is the subalgebra of F generated by the union of all
clusters obtained by mutation from a given seed.
Any skew-symmetric n × n matrix B determines a skew-symmetric form on a
(based) lattice Zn . Set N = Zn , I = Iu f = {1, . . . , n}, E to be the standard basis
on Zn , and let C = {x1 , . . . , xn }. We let A(B) denote the cluster algebra associated
to the seed (E, C).
Definition 5 is really a special case of the definition of a cluster algebra, a class
referred to as the skew-symmetric cluster algebras of geometric type. In the general
case the form {−, −} need only be skew-symmetrizable. One consequence of the
skew-symmetry of the form {−, −} is the identification of each exchange matrix
with an (unfrozen) quiver. One may assign this quiver in the obvious way, assigning
a vertex vi to each element i ∈ Iu f , and bi j arrows vi → v j , oriented according to
the sign of bi j . We may also add ‘frozen’ vertices vi for each element of i ∈ I \Iu f ,
with arrows introduced between frozen and unfrozen vertices similarly. Equivalently
we may consider the quiver associated to the extended exchange matrix, but we do
not make further use of this terminology. There is a well-known notion of quiver
mutation, going back to Fomin–Zelevinsky [13], generalising the reflection functors
of Bernstein–Gelfand–Ponomarev [6]. Mutating a seed in a skew-symmetric cluster
algebra induces a corresponding mutation of the associated quiver.
Definition 6 Given a quiver Q and an element i ∈ Iu f , the mutation of Q at vi is
the quiver mut(Q, v) obtained from Q by:
318 T. Prince
Xs = T M As = TN .
The dual pair of bases for the respective lattices define identifications of these tori
with split tori,
Xs −→ G|Im | , As −→ G|Im | .
Letting s denote the kth mutation of s, we associate birational maps μk : Xs Xs
and μk : As As to each seed, defined by setting
Theorem 9 ([9, Theorem 6.1]) Given a quiver Q with finite mutation class, its
adjacency matrix bi j is the adjacency matrix of a triangulation of a bordered surface
or is mutation equivalent to one of eleven exceptional types.
algebra). We recall that, working over C, if M is the lattice dual to N , the torus TM
is defined to be Spec(C[N ]).
Definition 11 Given an element w ∈ M, the weight vector, and f ∈ Ann(w), the fac-
tor, define a birational map φw, f : TM TM sending
z n
→ z n (1 + z f )w,n .
Given a Laurent polynomial W ∈ C[N ] such that φw, f (W ) ∈ C[N ] say that W is
mutable with weight vector w and factor f .
Definition 12 (Cf. [3, pg. 12]) Fix a Fano polytope P ⊂ NQ and its dual P ⊂
MQ , a weight vector w ∈ M, and factor f ∈ Ann(w). Define a piecewise linear
map Tw, f : MQ → MQ by setting
Tw, f : m
→ m + max(0, m, f )w.
If Tw, f (P ) is a convex polytope then we say P admits the mutation (w, f ) and
that P mutates to (Tw, f (P )) .
Remark 13 This definition of mutation is really a ‘dual characterisation’ of [3,
Definition 5], which encodes how the Newton polytope of a Laurent polynomial
changes under algebraic mutation.
Remark 14 In [3] the authors show that the result of applying a mutation to a Fano
polytope produces another Fano polytope, so the last dualization in Definition 12 is
well-defined.
Proposition 15 Given w ∈ M, f ∈ Ann(w) and a mutable Laurent polynomial W ∈
C[N ] we have the following identity;
Newt φw, fW = Tw, f Newt(W ) .
μk
As Aμk (s) (15.2)
p p
TM TM ,
φ(wk , fk )
E := {(w1 , f 1 ), (w2 , f 2 )}
where,
We recall that there is an A2 cluster structure on the co-ordinate ring of the Grass-
mannian, and a toric degeneration of Gr(2, 5) for each cluster chart in the dual
Grassmannian [33]. We expect that cluster structures in the mirror to a Fano variety
to be detected by such compatible collections of mutations.
Note that the polytopes we show in Fig. 15.1 are not dual to Fano polytopes.
However, recalling that B5 has Fano index 2, we can obtain a reflexive polytope by
dilating each of the polytopes shown in Fig. 15.1 by a factor of two, and translating.
In the two dimensional case, we can canonically define a maximal set of compat-
ible mutations, making use of the notion of singularity content [4].
Definition 26 (Cf. [27, Sect. 1.2]) Given a Fano polygon P ⊂ NQ with singularity
content (n, B) and m := |B| + n, we define:
1. an index set I of size m containing a subset Iu f of size n, together with functions
φu f : Iu f → {edges of P} φ f : I \Iu f → B
such that fibres φu−1f (E) contain m E := (E)/r E elements, where (E) is the
lattice length of the edge E, and r E is the Gorenstein (or local) index of the cone
over E, while the map φ f is a bijection;
2. a lattice map ρ : Zm → M sending each basis element to the primitive, inward-
pointing normal to the edge of P defined by the cone given by the specified
functions φu f and φ;
3. a form {ei , e j } := ρ(ei ) ∧ ρ(e j ). Note that this is an integral skew-symmetric
form.
324 T. Prince
The value m E appears in the definition of the singularity content of a two dimen-
sional cone; and is equal to the maximal number of T -cones of Gorenstein index r E
which fit inside the cone on E.
By [27, Proposition 3.17] the construction of a quiver from mutation data provided
by Definition 26 intertwines polygon and quiver mutations. We let (E P , C P ) denote
the seed associated to a Fano polygon, where E P is the standard basis ei of Zm , and C P
is the standard transcendence basis of the field of rational functions in n variables
over Q(xi : i ∈ I \Iu f ). We let Q P denote the unfrozen quiver associated to (E P , C P ).
We say a Fano polygon is of finite mutation type if it is mutation equivalent to only
finitely many Fano polygons.
Conjecture 27 The cluster algebra C P associated to a Fano polygon P, together
with a bijection between the set of frozen variables and B, is a complete mutation
invariant of the Fano polygon P.
Example 28 Consider the Fano polygon P for P2 (this is depicted in Fig. 15.2).
Computing the determinant of the inward-pointing normals we obtain the quiver Q P
•
3 3
• •
3
The mutations of this quiver are well-known, and the triple (3a, 3b, 3c) of non-zero
entries of the exchange matrix satisfy the Markov equation a 2 + b2 + c2 = 3abc.
Indeed, as the polygon P is mutated the corresponding toric surfaces are P(a 2 , b2 , c2 )
for the same triples (a, b, c). We see that in this case the mutations of the quivers
exactly capture the mutations of the polygon.
Example 29 Consider the toric surface (using the notation for these surfaces appear-
ing in [8]), X 5,5/3 associated with the Fano polygon shown in Fig. 15.3. The unfrozen
quiver associated to this surface is simply the A2 quiver:
• •
We now make use of the classification of finite type and finite mutation type cluster
algebras to establish the following result.
Theorem 30 P is of finite mutation type if and only if Q P is mutation equivalent to
a quiver of type (A1 )n , A2 , A3 , or D4 .
Remark 31 The types referred to in Theorem 30 may also be referred to as
type In , I I , I I I , and I V respectively; in analogy with Kodaira’s monodromy matri-
ces. The relationship between these matrices, log Calabi–Yau manifolds, and mon-
odromy in certain integral affine manifolds is explored by Mandel in [30].
Remark 32 We remark that all the cases which appear in Theorem 30 do occur as
(unfrozen) quivers associated to polygons. Several examples can be found in [8, p.
42] and are tabulated below.
Quiver Q P Polygon Surface
∅ = A01 9 X 6,2
A31 11 X 4,7/3
A61 12 B2,8/3
A2 7 X 5,5/3
A3 17 X 3,4
D4 5 X 4,4/3
We first make two straightforward observations. First we note that the cluster
algebra C P induces a sequence of surjections:
326 T. Prince
{Clusters of C P } (15.3)
The first vertical arrow follows from the fact that algebraic mutations determine
combinatorial mutations, the second from Lemma 20. For example, using this tower
of surjections in the case of a type A2 cluster algebra, we can immediately state the
following result.
Proposition 33 If a Fano polygon P has singularity content (2, B) and the primitive
inward-pointing normal vectors of the two edges corresponding to the unfrozen
variables of C P form a basis of the lattice M, then the mutation-equivalence class
of P has at most five members.
Proposition 24 implies that the mutation class of P has at most five elements. Note
that we do not have a non-trivial lower bound: there is only one polygon in mutation
equivalent to the polygon described in Example 29 up to GL(2, Z) equivalence. Next
observe that the sequence of surjections shown in (15.3) immediately implies that
Lemma 34 Given a Fano polygon P of finite mutation type, Q P does not contain a
Kronecker subquiver
k
Q k := { v1 v2 },
Remark 35 This result is expected from results on the corresponding cluster alge-
bra. The Kronecker quiver defines a rank 2 cluster algebra which is known not to be
of finite type when k > 1. Given that P is the Newton polygon of a superpotential
which is itself a combination of cluster monomials, we expect the polygon P to grow
as we mutate.
(h 1 , h 2 ) (h 1 , h 2 ) (h 1 , h 2 ).
ρ (u i1 ) − ρ (u i2 ) = (ρ (E i ))e3−i
,
where signs and orientations are chosen such that (E) is always positive. Study-
ing Fig. 15.4 note that h 1 + h 1 ≥ (ρ (E 2 )), however—by the calculation above—
(ρ (E 2 )) = k(E 2 ). Moreover, we have that (E 2 ) ≥ h 2 , since the Fano polygon P
admits a mutation along this edge. Hence we observe that
h 1 ≥ kh 2 − h 1 h 2 ≥ kh 1 − h 2 .
328 T. Prince
Consider the case k ≥ 3, and assume without loss of generality that h 2 ≥ h 1 . We have
that h 1 ≥ 3h 2 − h 1 ≥ 2h 2 ≥ 2h 1 . Thus in this case the values in the pair (h 1 , h 2 )
grow (at least) exponentially with mutation, and in particular take infinitely many
values.
Next consider the case k = 2. The inequalities above become,
h 1 ≥ 2h 2 − h 1 h 2 ≥ 2h 1 − h 2 ,
Remark 36 Proposition 34 implies all the quivers that we consider from now on
are directed graphs. Hence we refer to vertices as adjacent if they are adjacent in the
underlying graph.
Proof (Proof of Theorem 30) By Lemma 37, if Q P is not connected, Q P ∼ = An1 for
some n. Similarly, if Q P is of type A or D, then it must be one of A2 , A3 or D4 .
Thus we only need to show that there is no Fano polygon P of finite mutation type
such that C P is not of finite-type. However C P is of finite mutation type, and we
use the classification described in Theorems 9 and 10, following [9, 12]. In fact,
using Lemma 37, none of the eleven exceptional types can occur as Q P for a Fano
polygon P. Hence we can restrict to quivers which admit a block decomposition and
work case-by-case.
We claim that every quiver Q P associated to a Fano polygon P which admits a
block decomposition is either mutation equivalent to an orientation of a simply-laced
Dynkin diagram or to a quiver which contains a subquiver Q k for k > 1. We assume
for contradiction that Q P is the quiver associated to a Fano polygon P of finite-type
which is not mutation equivalent to a simply laced Dynkin diagram.
15 Polygons of Finite Mutation Type 329
Block V: First observe that, since only one vertex of the block V is an outlet, the V
block quiver is a subquiver of any quiver which contains the V block in its decom-
position. However this quiver mutates to a quiver with a Q 2 subquiver as shown in
Fig. 15.6. Therefore block V never appears in a decomposition of a quiver Q P . For
later use we shall fix the following intermediate quiver, V , shown in Fig. 15.7.
Blocks IIIa and IIIb: Assume there is a type III block (a or b) connected to a quiver Q
at a vertex v. If there is a vertex v of Q such that v and v are not adjacent, the quiver
violates Lemma 37. In particular the vertex set of Q must be the vertex set of a single
block. In particular, using the previous part, Q has at most four vertices. Case by
case study shows that only the A3 and D4 types appear.
Block IV: Consider the case of a decomposition only using type IV blocks. Note that
the type IV block is itself of type D4. Consider attaching two type IV blocks. If the
blocks are attached at a single outlet the resulting quiver contradicts Lemma 37. In
fact it is easy to see that it is impossible to add additional type IV blocks to meet
this condition. If both pairs of outlets are matched there are two possible quivers
depending on the relative orientations of the arrow between the outlets, one orienta-
330 T. Prince
For decompositions of Q P with type I and II blocks we divide the proof into cases
indexed by the number of type II blocks. For a single type II block, we can attach a
type I block to two outlets and in this way reduce to the type III case. Attaching each
type I block to a type II block in at most one outlet, we use the fact that every new
vertex must be adjacent to at least two of the vertices of the type II block. Thus we
can obtain only two undirected graphs—the underlying graph of a type IV block or
an orientation of a tetrahedron, these cases can easily be eliminated. For example,
there is no orientation of the tetrahedron making every cycle oriented; hence after a
single mutation we obtain a quiver violating Lemma 34.
Consider the case of a pair of type II blocks. If these have disjoint vertex sets, each
outlet of a type II block cannot be adjacent to two of the outlets of the other type II
block. Thus we must cancel the arrow between these two outlets with a type I block.
However this creates a pair of 1-valent non-outlet vertices which can be eliminated
similarly to the type III case. At the other extreme, if we attach along all three outlets,
we produce two easy cases. Attaching along a pair of outlets we generate either a Q 2
subquiver or a 4-cycle. Considering the 4-cycle with two outlets v1 and v2 (on non-
adjacent corners) to meet the conditions of Lemma 37 any vertex adjacent to one
of v1 or v2 must be adjacent to the other. Moreover, if the resulting quiver contains
an arrow between v1 and v2 , a mutation at one of the non-outlet vertices gives a Q 2
subquiver. Given a vertex v adjacent to v1 and v2 , if this defines a path between them,
mutating at this node and a non-outlet in the four cycle produces a Q 2 subquiver. If v
does not lie on a path between v1 and v2 then mutating at both outlets produces a Q 2
subquiver.
Attaching the type II blocks at a single outlet, the four arrows incident to this
vertex are now fixed, so any new vertex must be adjacent to each of the remaining
four outlets by Lemma 37. However this cannot be achieved with type I blocks.
Attaching more than two type II blocks together, we can eliminate the case where
two are connected to form a 4-cycle as above. Since we can easily eliminate the case
that two type II blocks meet in three outlets, we assume that each type II block meets
every other in at most one outlet. Some pair of type II blocks must be attached in
an outlet (otherwise we can argue as in the case of type II block separated by type I
blocks). Thus, since every new vertex must be adjacent to all four outlets formed by
attaching two type II blocks, all possible quivers can be represented as an octahedron
with some orientation, see Fig. 15.9.
Considering an orientation of the octahedron; if any triangular face does not form
a cycle we can mutate to form a Q 2 subquiver. Assuming every triangle is a cycle,
and possibly mutating, the vertices adjacent to the ‘top’ of the octahedron form a
type V block subquiver. Following the same reasoning as for the type V block case
(although note that the type V block is not part of a block decomposition here) these
cases can be eliminated.
332 T. Prince
Acknowledgements We thank Alexander Kasprzyk for his insights on polytope mutation, and
our many conversations about quivers. The author is supported by a Fellowship by Examination at
Magdalen College, Oxford. This work was undertaken while the author was a graduate student at
Imperial College London.
References
1. Aharony, O., Hanany, A.: Branes, superpotentials and superconformal fixed points. Nucl. Phys.
B 504(1–2), 239–271 (1997)
2. Akhtar, M., Coates, T., Corti, A., Heuberger, L., Kasprzyk, A.M., Oneto, A., Petracci, A.,
Prince, T., Tveiten, K.: Mirror symmetry and the classification of orbifold del Pezzo surfaces.
Proc. Am. Math. Soc. 144(2), 513–527 (2016)
3. Akhtar, M., Coates, T., Galkin, S., Kasprzyk, A.M.: Minkowski polynomials and mutations.
SIGMA Symmetry Integrability Geom. Methods Appl. 8, 094, 17 (2012)
4. Akhtar, M., Kasprzyk, A.M.: Singularity Content (2014). arXiv:1401.5458
5. Bergman, A., Proudfoot, N.: Moduli spaces for D-branes at the tip of a cone. J. High Energy
Phys. (3), 073, 9 (2006)
6. Bernšteı̆n, I.N., Gel fand, I.M., Ponomarev, V.A.: Coxeter functors, and Gabriel’s theorem.
Uspehi Mat. Nauk 28(2(170)), 19–33 (1973)
7. Bridgeland, T., Stern, D.: Helices on del Pezzo surfaces and tilting Calabi-Yau algebras. Adv.
Math. 224(4), 1672–1716 (2010)
8. Corti, A., Heuberger, L.: Del Pezzo surfaces with 13 (1, 1) points. Manuscripta Math. 153(1–2),
71–118 (2017)
9. Felikson, A., Shapiro, M., Tumarkin, P.: Skew-symmetric cluster algebras of finite mutation
type. J. Eur. Math. Soc. (JEMS) 14(4), 1135–1180 (2012)
10. Feng, B., Hanany, A., He, Y.H.: Phase structure of D-brane gauge theories and toric duality. J.
High Energy Phys. (8), 40, 25 (2001)
11. Fock, V.V., Goncharov, A.B.: Cluster ensembles, quantization and the dilogarithm. II. The
intertwiner. In: Algebra, arithmetic, and geometry: in honor of Yu. I. Manin, vol. I, Progress in
Mathematics, vol. 269, pp. 655–673. Birkhäuser Boston, Inc., Boston, MA (2009)
12. Fomin, S., Shapiro, M., Thurston, D.: Cluster algebras and triangulated surfaces. I. Cluster
complexes. Acta Math. 201(1), 83–146 (2008)
13. Fomin, S., Zelevinsky, A.: Cluster algebras. I. Foundations. J. Am. Math. Soc. 15(2), 497–529
(2002)
15 Polygons of Finite Mutation Type 333
14. Fomin, S., Zelevinsky, A.: Cluster algebras. II. Finite type classification. Invent. Math. 154(1),
63–121 (2003)
15. Franco, S., Hanany, A., Martelli, D., Sparks, J., Vegh, D., Wecht, B.: Gauge theories from toric
geometry and brane tilings. J. High Energy Phys. (1), 128, 40 (2006)
16. Galkin, S., Usnich, A.: Mutations of Potentials (2010). Preprint IPMU 10-0100
17. Givental, A.: A mirror theorem for toric complete intersections. In: Topological Field Theory,
Primitive Forms and Related Topics (Kyoto, 1996), Progress in Mathematics, vol. 160, pp.
141–175. Birkhäuser Boston, Boston, MA (1998)
18. Givental, A.B.: Homological geometry and mirror symmetry. In: Proceedings of the Interna-
tional Congress of Mathematicians, vol. 1, 2 (Zürich, 1994), pp. 472–480. Birkhäuser, Basel
(1995)
19. Givental, A.B.: Equivariant Gromov-Witten invariants. Internat. Math. Res. Not. 13, 613–663
(1996)
20. Gross, M., Hacking, P., Keel, S.: Birational geometry of cluster algebras. Algebr. Geom. 2(2),
137–175 (2015)
21. Hanany, A., Kazakopoulos, P., Wecht, B.: A new infinite class of quiver gauge theories. J. High
Energy Phys. (8), 054, 30 (2005)
22. Hanany, A., Vegh, D.: Quivers, tilings, branes and rhombi. J. High Energy Phys. (10), 029, 35
(2007)
23. Herzog, C.P.: Seiberg duality is an exceptional mutation. J. High Energy Phys. (8), 064, 31
(2004)
24. Hille, L., Perling, M.: Exceptional sequences of invertible sheaves on rational surfaces. Compos.
Math. 147(4), 1230–1280 (2011)
25. Hori, K., Katz, S., Klemm, A., Pandharipande, R., Thomas, R., Vafa, C., Vakil, R., Zaslow,
E.: Mirror symmetry, Clay Mathematics Monographs, vol. 1. American Mathematical Society,
Providence, RI; Clay Mathematics Institute, Cambridge, MA (2003). With a preface by Vafa
26. Ilten, N.O.: Mutations of Laurent polynomials and flat families with toric fibers. SIGMA
Symmetry Integrability Geom. Methods Appl. 8, 047, 7 (2012)
27. Kasprzyk, A.M., Nill, B., Prince, T.: Minimality and mutation-equivalence of polygons. Forum
Math. Sigma 5, e18, 48 (2017)
28. Kontsevich, M.: Lectures at ENS Paris (1998). Set of notes taken by J. Bellaiche, J.-F. Dat, I.
Martin, G. Rachinet and H. Randriambololona
29. Leung, N.C., Vafa, C.: Branes and toric geometry. Adv. Theor. Math. Phys. 2(1), 91–118 (1998)
30. Mandel, T.: Classification of rank 2 cluster varieties. SIGMA Symmetry Integrability Geom.
Methods Appl. 15, 042, 32 (2019)
31. Mukhopadhyay, S., Ray, K.: Seiberg duality as derived equivalence for some quiver gauge
theories. J. High Energy Phys. (2), 070, 22 (2004)
32. Perling, M.: Examples for exceptional sequences of invertible sheaves on rational surfaces. In:
Geometric methods in representation theory. II, Sémin. Congr., vol. 24, pp. 371–392. Society
Mathematics France, Paris (2012)
33. Rietsch, K., Williams, L.: Newton-Okounkov bodies, cluster duality, and mirror symmetry for
Grassmannians. Duke Math. J. 168(18), 3437–3527 (2019)
Chapter 16
Orbit Spaces of Maximal Torus Actions
on Oriented Grassmannians of Planes
Hendrik Süß
16.1 Introduction
the natural action of a maximal torus in SOn . Our main result determines the orbit
space of this action.
n/2−2
S n/2−1 ∗ PC .
H. Süß (B)
Friedrich-Schiller-Universität Jena, Institut für Mathematik, Ernst-Abbe-Platz 2, 07743 Jena,
Germany
e-mail: [email protected]
For smooth varieties with a torus action of complexity 1 we derive the following
general results on the structure of their orbit spaces.
Note that results comparable to Theorem 2 have been proved by Ayzenberg [3]
and Cherepanov [8], Theorem 3 has been proved independently, but using similar
methods, by Karshon and Tolman [15]. Their work covers the more general setting
of symplectic manifolds with Hamiltonian torus actions. In their paper they also
prove Theorem 1 for the cases of complexity 1, i.e. for n = 5, 6.
Our main tool is Geometric Invariant Theory (GIT) and its symplectic coun-
terpart in combination with the Kempf–Ness Theorem. This approach suggest to
stratify the manifold and eventually the orbit space via a polyhedral subdivision of
the momentum polytope, which encodes the variation of GIT quotients. In general
these stratifications can become arbitrarily complicated. However, in the cases con-
sidered in this paper they turn out to be almost trivial allowing us to derive concrete
results about the orbits spaces.
In Sect. 16.2 we fix our setting for compact torus actions induced by algebraic
torus actions on complex varieties and recall crucial results from Geometric Invariant
Theory. Moreover, we derive first results on the structure of orbits spaces in suitable
situations. We then apply these to the special cases of oriented Grassmannians of
planes in Sect. 16.3 and TC -varieties of complexity 1 in Sect. 16.4.
In order to distinguish between the algebraic and the topological category, we are
going to denote isomorphism of algebraic varieties by ∼ = and homeomorphisms of
topological spaces by ≈.
t.(z 0 : . . . : z N ) = (t u 0 z 1 : . . . : t u N z N ),
(u j )1 (u j )k
where t u j := t1 · · · tk . Then a moment map of this action is given by
j |z j |2 u j
ν: PCN → R ; (z 0 : . . . z N ) →
k
.
j |z j |2
16 Orbit Spaces of Maximal Torus Actions on Oriented Grassmannians … 337
For an embedded projective variety X ⊂ PCN , which is invariant under under this torus
action, a moment map of the induced torus action on X is given by the restriction μ =
ν| X . The moment image P = μ(X ) is known to be a convex polytope [2, 11]. We
start with some notions known as variation of GIT quotients with [9, 12, 16] being the
most relevant references. For a point x ∈ X the moment image (x) = μ(TC .x) ⊂ P
of its orbit closure is again a polytope and the orbit TC .x is mapped to the relative
interior ◦ (x) ⊂ (x).
For a point u ∈ P we define
Hence, X ss (u) consists of those points in X whose orbit closures intersect μ−1 (u)
and X ps (u) consists of those points whose orbits intersect μ−1 (u). Equivalenty X ps (u)
is the union of closed TC -orbits in X ss (u).
Now for every u ∈ P we may consider
λ(u) = (x); λ◦ (u) = ◦ (x)
x,u∈(x) x,u∈◦ (x)
Since only finitely many polytopes occur as moment images of orbit closures their
intersections are again polytopes. We denote the set of all these polytopes λ(u) by .
This set is partially ordered by the face relation ≺. The polytopes λ ∈ form a
polyhedral subdivision of P and one obtains a stratification of P via their relative
interiors.
P= λ◦ .
λ∈
get P = μ(P2 ) = [−1, 1] ⊂ R. The orbits can be described as follows. We have the
fixed points (1 : 0 : 0), (0 : 1 : 0) and (0 : 0 : 1) with moment images 1, −1 and 0,
respectively. The moment images of the other TC -orbits are
Hence, in this case is obtained by subdividing the interval [−1, 1] at the point 0,
or more formally = {[−1, 0], [0, 1], {−1}, {0}, {1}}.
By the Kempf–Ness Theorem for rational values of u the definition of X ss (u)
coincides with the semi-stable locus of Mumford’s Geometric Invariant Theory.
Hence, there exists a categorical quotient morphism qλ : X λss → Yλ = X λss //TC where
ps
Yλ is an orbit space for the TC -action on X λ and the corresponding quotient map is
ps
given by the restriction of qλ to X λ . The occurring quotients Yu have the expected
dimension for u ∈ P ◦ , but can be lower-dimensional for elements u ∈ ∂ P. By [17,
Lemma 7.2] for u ∈ λ◦ every TC -orbit in X λ intersects μ−1 (u) in exactly one T -orbit.
ps
Moreover, the GIT quotients X ss (u)//TC are again toric varieties corresponding
to the polytope F −1 (u) ∩ Q, see [13, Proposition 3.5].
Already in [10] it has been observed that orbit space of the T -action on X can be
constructed out of the inverse system of GIT quotients.
Theorem 7 ([10, Sect. 5]) We have
X/T ≈ λ × Yλ /∼,
λ∈
We easily derive the following result, which turns out to be a little bit handier in
some situations.
Corollary 8 Assume that we have a compact topological space Y and with proper
surjective maps rλ : Y → Yλ being compatible with the inverse system above. Then
we have the following homeomorphism.
X/T ≈ (P × Y )/∼r .
for u ∈ λ◦ .
Proof There is a canonical map
P ×Y → λ × Yλ /∼; (u, y) → [(u, rλ(u) (y)]
λ∈
This map is surjective and continuous and identifies exactly those pairs which are
equivalent under ∼r . The quotient (P × Y )/∼r is compact as (P × Y ) is and by
Theorem 7 the codomain of the map is homeomorphic to X/T , which is a Hausdorff
space. Hence, the induced continuous bijection (P × Y )/∼r → X/T is a homeo-
morphism.
Remark 9 In [6, 7] such a Y is called a universal parameter space for the TC -obits.
In algebraic geometry a natural choice for such a dominating algebraic object Y
would be the inverse limit of the Yλ or the Chow quotient of X by TC , which can be
identified with a distinguished irreducible component of this inverse limit.
If the structure of the inverse system {Yλ }λ∈ of GIT quotients is complicated
Corollary 8 might not give much concrete information about the orbit space X/T .
However, in certain situations this structure turns out to be almost trivial allowing us
to effectively calculate the orbit space.
Definition 10 We say the TC -action on X ⊂ PCN has an almost trivial variation of
GIT if for λ ⊂ ∂ P the quotients Yλ are all isomorphic to some Y .
Example 11 If the torus action has complexity one than the quotients Yλ are smooth
algebraic curves or just a point, where the latter happens at most over the boundary
of P. The only contraction morphisms here are isomorphisms or the contraction of
a curve to a point. Hence, the definition is automatically fulfilled.
Proposition 12 Consider a TC -action on X with almost trivial variation of GIT and
only finitely many lower-dimensional TC -orbits. Then X/T is homeomorphic to the
topological join S k−1 ∗ Y .
340 H. Süß
X/T ≈ (P × Y )/∼∂ ,
Lemma 13 Consider the closed unit disc D k and the unit sphere S k−1 . Then for any
compact topological manifold Y we have
S k−1 ∗ Y ≈ (D k × Y )/∼∂ ,
Proof Recall that the join S k−1 ∗ Y is defined as (S k−1 × Y × [0, 1])/∼, with
the equivalence relation being generated by (s, y, 0) ∼ (s , y, 0) and (s, y, 1) ∼
(s, y , 1). Now the homeomorphism is given by
Hence,
G+
R (2, n) = {(v1 , v2 ) ∈ R
n×2
| vi , v j = δi j }/ SO2 .
: Rn×2 → Cn ; (v1 , v2 ) → w = v1 + i · v2
induces an embedding ¯ : G+
R (2, n)
→ PC . This is well-defined as ((v
n
1 , v2 )Q φ ) =
e · (v1 , v2 ). Moreover the condition v1 , v2 = 0 is equivalent to ( j w2j ) = 0
iφ
and |v1 |2 /|v2 |2 = 1 is equivalent to ( j w2j ) = 0. Hence, the image of the embed-
ding in PnC is cut out by the equation j w2j = 0. A change of coordinates
or
k
z n2 + z 2 j−1 z 2 j = 0, (16.2)
j=1
respectively. Now in these coordinates one easily checks that for an oriented
plane E ∈ G +R (2, n) with
342 H. Süß
¯
(E) = (z 1 : . . . : z n )
we have
iφ1
¯
(diag(Q φ1 , . . . , Q φk )E) = e z 1 : e
−iφ1
z 2 : . . . : eiφk z 2k−1 : e−iφ1 z 2k
in the case n = 2k + 1. Let e j denote the jth canonical basis vector of Zk . Then the
action of T = (S 1 )k above is induced by an algebraic torus action of TC = (C∗ )k
with weights
k
μ(z 1 : . . . : z n ) = n (|z 2 j−1 |2 − |z 2 j |2 )e j . (16.3)
i=1 |z i |
2
j=1
The moment image of X is the cross-polytope given as the convex hull of the
weights P = βk = conv(±e1 , . . . , ±ek ). The fixed point (1 : 0 : . . . : 0) is mapped
to e1 , (0 : 1 : . . . : 0) to −e1 and similarly for the other coordinates.
Remark 15 The proper faces of the cross-polytope P are exactly the convex hulls of
subsets of {±e1 , . . . , ±ek } where for every j at most one of e j and −e j is contained.
or equivalently
16 Orbit Spaces of Maximal Torus Actions on Oriented Grassmannians … 343
k
k
|z 2 j−1 |2 − |z 2 j |2
= |z 2 j−1 |2 + |z 2 j |2 .
j=1 j=1
q : Pn−1
C Pk−2
C , (z 1 : . . . : z n ) → (z 3 z 4 : . . . : z n−1 z n ).
This map is easily seen to be invariant under the TC -action. It is well-defined on the
locus of points where at least one of the products z 2 j−1 z 2 j for j = 2, . . . , k does not
vanish. For a point z ∈ X this is equivalent to the fact that μ(z) ∈ P ◦ .
|z 1 | − |z 2 | |z | − |z |
= 1 2 ,
N N
Note, that by (16.4) we have |s1 |−2 ≤ NN ≤ |s1 |2 . Now, from (z 1 , z 2 ) = (s1 z 1 , s1−1 z 2 )
we obtain
|z 1 |2 − |z 2 |2 |s1 |2 |z 1 |2 − |s1 |−2 |z 2 |2
= ,
N N
which implies (N /N − |s1 |2 )|z 1 |2 = (N /N − |s1 |−2 )|z 2 |2 . For sign reasons this is
only possible if |s1 |2 = N /N = 1. Now, it follows from (16.4) that |s1 | = · · · =
|sk | = 1 and s ∈ T .
Since X ss (u) consists exactly of the orbits whose closures intersect μ−1 (u) it
follows also that q| X ss (u) is a good quotient in the sense of Geometric Invariant
Theory. Hence, it coincides with the GIT quotient.
We conclude this section by studying the moment images (x) of TC -orbit clo-
sures and the induced subdivision of P from Sect. 16.2. For n = 2k + 1 the convex
hull of every subset of vertices of P occurs as a moment image. For n = 2k such con-
vex hulls are moment images if an only if they are faces of P or contain at least two
pairs of opposite vertices {ei , −ei }, {e j , −e j }. Indeed, for every such polytope a
corresponding TC -orbit is given by TC · (z 1 : z 2 : . . . : z n ) with z 2 j−1 = 0 ⇔ e j ∈
and z 2 j = 0 ⇔ −ei ∈ for j = 1, . . . , k. Note, that for n = 2k + 1 such (z 1 : z 2 :
. . . : z n ) fulfilling (16.2) always exist, but for n = 2k there is obviously no non-trivial
solution of (16.1) where all but one monomial vanish. In both cases, with the excep-
tion of k = 2, the induced subdivision of P is the same and coincides with the stellar
subdivision of P obtained by starring in the origin.
Remark 18 In [7] Buchstaber and Terzić introduced the notion of (2n, k)-manifold.
It’s relatively straightforward to check that for n = 2k, the axioms of this notion are
indeed fulfilled for the associated effective torus action by TC /±1. However, in
the odd case on a generic point of the hyperplane section [z n = 0] we have finite
stabilisers of order 2, which violates the conditions for a (2n, k)-manifold.
Remark 19 Note that G + R (2, 6) can be identified with G C (2, 4) as both are given by
the smooth quadric hypersurface in P5C . Hence, for this case we just rediscover the
results of [5]. Combinatorially this fact is reflected by coincidence of the moment
polytopes, i.e. the cross-polytope β3 and the hypersimplex 4,2 .
16 Orbit Spaces of Maximal Torus Actions on Oriented Grassmannians … 345
If the complexity of the torus action is 1 the possible GIT quotients Yλ are either single
points or isomorphic to a fixed algebraic curve Y . Hence, when applying Corollary 8
to this situation the maps pλ : Y → Yλ are either isomorphisms or contractions to a
point. Our main aim in this section is to prove that in this situation the resulting orbits
spaces are topological manifolds with boundary and even spheres if the number of
lower dimensional TC -orbits is finite.
Remark 20 In the toric case the orbit space can be identified with the moment poly-
tope. In particular, it is also a topological manifold with boundary. Hence, Theorem 2
can be seen as generalisation of this fact. On the other hand, this phenomenon is very
special to complexity 0 and 1. In higher dimensions this will almost never be the
case. For example for a smooth projective variety Y the join S n ∗ Y , which occurs
as an orbit space in the situation of Proposition 12, is a topological manifold if and
only if Y ∼
= P1C .
Proof The moment image of PdC is given by the convex hull of the weights u 0 , . . . , u d .
The weights u 0 , . . . , u d are necessarily affinely dependent in Rd−1 . On the other hand
they span Rd−1 as an affine space due to the effectiveness of the torus action. Hence,
d d
there is a non-trivial choice of α j ∈ Z, such that 0 = i=0 αi u i and 0 = i=0 αi
and the coefficients are unique up to simultaneous scaling.
Set K = {i ∈ {0, . . . , d} | αi = 0}. Then P is obtained as the join Q ∗ of
the lower-dimensional polytopes Q = conv{u i }i∈K and = conv{u i }i ∈K / of dimen-
sions m := (#K − 2) and n := (d − #K ), respectively. Here, we allow that = ∅
and use the non-standard convention Q ∗ ∅ := Q. Note, that is a simplex (or
empty).
Hence, a u ∈ has a unique representation as u= j∈K λ j u j with λ j ≥ 0
and j λ j = 1. For u ∈ Q such a representation u = j ∈K / λ j u j is unique if and
only if u ∈ ∂ Q. It follows that u ∈ P = Q ∗ has a unique such representation if
and only if u ∈ ∂ Q ∗ .
Now,Yλ(u) = μ−1 (u)/T is a point whenever u ∈ P has a unique representation
as u = j λ j u j and μ−1 (u)/T ≈ P1C otherwise. This is just a special case of Exam-
ple 6, when F : Rd → Rk is given by F(ei ) = u i for i = 1, . . . , d. Then theintersec-
tion of F −1 (u) and the standard simplex consists nof all linear combinations
n
i=1 λi ei
with non-negative coefficients, such that u = i=1 λi u i and λi = 1. The result is
a point if the linear combination is unique or a line segment if not. The correspond-
ing toric varieties are a single point and P1C , respectively. Alternatively, it not hard to
show that the non-trivial quotient morphisms X λss → Yλ = P1 are all restrictions of
the rational map
346 H. Süß
αi
P P ,
d 1
(z 0 : . . . : z d ) → z 0 i
: z iαi .
i
Note, that the disc D m+n can be identified with the hemisphere via projection
and D m+n ∗ {pt} with the corresponding halfdisc. Now, by choosing H to be an
arbitrary halfspace it follows from Lemma 14 that the orbit space is a hemisphere.
For = ∅ and P = Q we see directly (P, ∂ Q ∗ ) = (Q, ∂ Q) ≈ (D d−1 , S d−2 )
and we obtain PdC /T ≈ S d+1 from invoking Lemma 14 again, this time with H =
Rd−1 .
Proof (Proof of Theorem 2) We first consider the situation of a complexity-one
torus action on the affine space Cd . Such an action is linearisable by [4]. We may
equivariantly compactify the TC -action on Cd to a TC -action on PdC . Then Cd /T is an
open subset of PdC /T . Hence, the claim follows from the observation in Proposition 21
that the orbit space PdC /T is a manifold with boundary.
To deduce the general case we consider the two situations from Lemma 23. If
contractions to a point do not occur the equivalence relation ∼r is trivial and by
Corollary 8 we have X/T ≈ P × Y which is a product of topological manifolds
with boundary.
In the second situation, we have Y ∼ = P1 . Then by [1, Theorem 5] we have an
equivariant open cover of X by copies of Cn and the result follows directly from the
consideration above.
Remark 22 To deduce the general case from the case X = Cd in the proof of
Theorem 2 we may alternatively consider the induced TC -action on the tangent space
at a fixed point and reduce everything to this situation by applying Luna’s Slice
Theorem.
: P → DivR P1C
from the polytope to the vector space of R-divisors on P1 . By [21, Sect. 3.1] we
have Yλ(u) = {pt} if and only deg (u) = 0. Now, one checks that in every case
the subset {u ∈ ∂ P | deg (u) > 0} ⊂ ∂ P consists of the interior of k disjoint facts
of P. Applying Proposition 25 gives the desired result.
Remark 27 Note, that Q is the smooth quadric and by Sect. 16.3 coincides with
G+R (2, 5). Moreover, 2.32 is the variety of complete flags in C . Hence, we recover
3
Acknowledgements This research was supported by the program Interdisciplinary Research con-
ducted jointly by the Skolkovo Institute of Science and Technology and the Interdisciplinary Sci-
entific Center J.-V. Poncelet. In particular, I am grateful for the warm hospitality offered by Center
Poncelet. This work was also partially supported by the grant 346300 for IMPAN from the Simons
Foundation and the matching 2015–2019 Polish MNiSW fund. I would like to thank Anton Ayzen-
berg, Victor Buchstaber, Alexander Perepechko and Nigel Ray for stimulating discussions on the
subject of this paper. Finally, I want to thank Yael Karshon and Susan Tolman for making their
draft [15] available to me.
References
1. Arzhantsev, I., Perepechko, A., Süß, H.: Infinite transitivity on universal torsors. J. Lond. Math.
Soc. (2) 89(3), 762–778 (2014)
2. Atiyah, M.F.: Convexity and commuting Hamiltonians. Bull. Lond. Math. Soc. 14(1), 1–15
(1982)
3. Aı̆zenberg, A.A.: Torus actions of complexity 1 and their local properties. Tr. Mat. Inst. Steklova
302(Topologiya i Fizika), 23–40 (2018). English version published in Proc. Steklov Inst. Math.
302(1), 16–32 (2018)
4. Biał ynicki Birula, A.: Remarks on the action of an algebraic torus on k n . Bull. Acad. Polon.
Sci. Sér. Sci. Math. Astronom. Phys. 14, 177–181 (1966)
5. Buchstaber, V.M., Terzić, S.: Topology and geometry of the canonical action of T 4 on the
complex Grassmannian G 4,2 and the complex projective space CP5 . Mosc. Math. J. 16(2),
237–273 (2016)
6. Buchstaber, V.M., Terzić, S.: Toric topology of the complex Grassmann manifolds. Mosc.
Math. J. 19(3), 397–463 (2019)
7. Bukhshtaber, V.M., Terzich, S.: The foundations of (2n, k)-manifolds. Mat. Sb. 210(4), 41–86
(2019)
8. Cherepanov, V.V.: Orbit spaces of torus actions on Hessenberg varieties (2019).
arXiv:1905.02294 [math.AT]
9. Dolgachev, I.V., Hu, Y.: Variation of geometric invariant theory quotients. Inst. Hautes Études
Sci. Publ. Math. (87), 5–56 (1998). With an appendix by Nicolas Ressayre
10. Goresky, M., MacPherson, R.: On the topology of algebraic torus actions. In: Algebraic groups
Utrecht 1986, Lecture Notes in Mathematics, vol. 1271, pp. 73–90. Springer, Berlin (1987)
11. Guillemin, V., Sternberg, S.: Convexity properties of the moment mapping. Invent. Math. 67(3),
491–513 (1982)
12. Hu, Y.: The geometry and topology of quotient varieties of torus actions. Duke Math. J. 68(1),
151–184 (1992)
13. Kapranov, M.M., Sturmfels, B., Zelevinsky, A.V.: Quotients of toric varieties. Math. Ann.
290(4), 643–655 (1991)
14. Karshon, Y., Tolman, S.: Centered complexity one Hamiltonian torus actions. Trans. Amer.
Math. Soc. 353(12), 4831–4861 (2001)
15. Karshon, Y., Tolman, S.: Topology of complexity one quotients. Pacific J. Math. 308(2), 333–
346 (2020)
16. Kirwan, F.: Momentum maps and reduction in algebraic geometry. pp. 135–171 (1998). Sym-
plectic geometry
17. Kirwan, F.C.: Cohomology of quotients in symplectic and algebraic geometry, Mathematical
Notes, vol. 31. Princeton University Press, Princeton (1984)
18. Kobayashi, S., Nomizu, K.: Foundations of differential geometry. Vol. II. Interscience Tracts
in Pure and Applied Mathematics, No. 15 Vol. II. Interscience Publishers John Wiley & Sons,
Inc., New York (1969)
16 Orbit Spaces of Maximal Torus Actions on Oriented Grassmannians … 349
19. Mori, S., Mukai, S.: Classification of Fano 3-folds with B2 ≥ 2. Manuscripta Math. 36(2),
147–162 (1981/82)
20. Przhiyalkovskiı̆, V.V., Cheltsov, I.A., Shramov, K.A.: Fano threefolds with infinite automor-
phism groups. Izv. Ross. Akad. Nauk Ser. Mat. 83(4), 226–280 (2019)
21. Süss, H.: Fano threefolds with 2-torus action: a picture book. Doc. Math. 19, 905–940 (2014)
Chapter 17
The Reflexive Dimension
of (0, 1)-Polytopes
Akiyoshi Tsuchiya
Abstract Haase and Melnikov showed that every lattice polytope is unimodularly
equivalent to a face of some reflexive polytope. The reflexive dimension of a lattice
polytope P is the minimal d so that P is unimodularly equivalent to a face of some d-
dimensional reflexive polytope. Computing the reflexive dimension of a lattice poly-
tope is a hard problem in general. In this survey, we discuss the reflexive dimension
of a (0, 1)-polytope. In particular, virtue of the algebraic technique on Gröbner bases
and a linear algebraic technique, many families of reflexive polytopes arising from
several classes of (0, 1)-polytopes are presented, and we see that the (0, 1)-polytopes
are unimodularly equivalent to facets of some reflexive polytopes.
17.1 Introduction
The reflexive polytope is one of the keywords belonging to the current trends in the
research of convex polytopes. In fact, many authors have studied reflexive polytopes
from the viewpoints of combinatorics, commutative algebra and algebraic geometry.
Hence, finding new classes of reflexive polytopes is an important problem.
A lattice polytope is a convex polytope all of whose vertices have integer coor-
dinates. Two lattice polytopes P ⊆ Rd and P ⊆ Rd are said to be unimodularly
equivalent if there exists an affine map from the affine span aff(P) of P to the
affine span aff(P ) of P that maps Zd ∩ aff(P) bijectively onto Zd ∩ aff(P ) and
that maps P to P . Note that every lattice polytope is unimodularly equivalent to a
full-dimensional one. A lattice polytope P ⊂ Rd of dimension d is called reflexive
if the origin of Rd is the unique lattice point belonging to the interior of P and its
A. Tsuchiya (B)
Graduate School of Mathematical Sciences, University of Tokyo, 3-8-1 Komaba, Meguro-ku,
Tokyo 153-8914, Japan
e-mail: [email protected]
dual polytope
P∨ := {y ∈ Rd | x, y ≤ 1 for all x ∈ P}
is also a lattice polytope, where x, y is the usual inner product of Rd . It is known
that reflexive polytopes correspond to Gorenstein toric Fano varieties, and they are
related to mirror symmetry, e.g., [1, 3]. In each dimension, there exist only finitely
many reflexive polytopes up to unimodular equivalence [15]. Moreover, Haase and
Melnikov [4] showed that every lattice polytope is unimodularly equivalent to a face
of some reflexive polytope. From this result they defined the reflexive dimension of
a lattice polytope.
We immediately know the following proposition from the fact that there are only
finitely many reflexive polytopes up to unimodular equivalence.
Our interest is to classify lattice polytopes whose reflexive dimensions are a given
integer. We remark that classifying lattice polytopes whose reflexive dimensions
are equal to their dimensions is equivalent to classifying reflexive polytopes. In
particular, it is known that there is one reflexive polytope in dimension one, there
are 16 in dimension two, 4,319 in dimension three and 473,800,776 in dimension
four according to computations by Kreuzer and Skarke [14]. As a next step, we focus
on lattice polytopes whose reflexive dimensions are equal to their dimensions plus
one. Namely, we consider the following question.
Question 3 For which lattice polytope P, does it follow that rdim(P) = dim(P) +
1?
Question 4 For every (0, 1)-polytope P, does it follow that rdim(P) = dim(P) + 1?
Equivalently, is every (0, 1)-polytope a facet of some reflexive polytope?
In order to show that this second Question 4 has a positive answer for some
class of (0, 1)-polytopes, we give higher-dimensional construction of lattice poly-
topes. Given two lattice polytopes P ⊂ Rd and Q ⊂ Rd , we set the lattice polytope
(P, Q) ⊂ Rd+1 with
If P = Q, then we will write (P) := (P, P). We remark that the origin of Rd+1
is always a relative interior lattice point of (P). Assume that P is full-dimensional.
Then (P) is also full-dimensional. In particular, P × {1} is a facet of (P). Hence
for a lattice polytope P ⊂ Rd of dimension d, if (P) is reflexive, then we know
that P is unimodularly equivalent to a facet of some reflexive polytope.
In this survey, by using this construction, we will show that for several classes
of (0, 1)-polytopes, Question 4 has a positive answer. In particular, we will present
many large families of reflexive polytopes arising from well-known classes of (0, 1)-
polytopes.
This survey is organized as follows: In Sect. 17.2, we will introduce an algebraic
technique to show that a given lattice polytope is reflexive. In particular, we will recall
basic materials and notation on toric ideals. In Sect. 17.3, we will give two families
of reflexive polytopes arising from the order polytopes and the chain polytopes of
finite partially ordered sets. In Sect. 17.4, we will give a family of reflexive polytopes
arising from the stable set polytopes of perfect graphs. Finally, in Sect. 17.5, we will
introduce a linear algebraic technique to show that a lattice polytope is reflexive, and
we will give a family of reflexive polytopes arising from the edge polytopes of finite
simple graphs.
Hence the initial ideal in<rev (IP ) of the toric ideal IP with respect to the reverse lexi-
cographic order <rev on the polynomial ring K [x] induced by the ordering xd+2 <rev
xd+1 <rev · · · <rev x1 is squarefree. Therefore, by Lemma 5, P is a reflexive polytope
with a regular unimodular triangulation.
In this section, we give two families of reflexive polytopes with regular unimodular
triangulations arising from order polytopes and chain polytopes of finite partially
ordered sets. In particular, we show that Question 4 has a positive answer for order
and for chain polytopes.
17 The Reflexive Dimension of (0, 1)-Polytopes 355
First, we recall some terminologies of finite partially ordered sets and introduce
two lattice polytopes arising from finite partially ordered sets. Let P denote a finite
partially ordered set (poset, for short) on the ground set [d] := {1, . . . , d}. A subset I
of [d] is called a poset ideal of P if i ∈ I and j ∈ P together with j ≤ i in P,
then j ∈ I . Note that the empty set ∅ and [d] are poset ideals of P. Let J(P)
denote the set of poset ideals of P. A subset A of [d] is called an antichain of P if
i and j belonging to A with i = j are incomparable. In particular, the empty set ∅
and each 1-element subsets { j} are antichains of P. Let A(P) denote the set of
antichains of P. For a poset ideal I of P, we write max(I ) for the set of maximal
elements of I . In particular, max(I ) is an antichain. A linear extension of P is a
permutation σ = i 1 i 2 · · · i d of [d] which satisfies a < b if i a < i b in P.
Stanley [26] introduced two classes of lattice polytopes arising from finite posets,
which are called order polytopes and chain polytopes. The order polytope O P of P
is defined to be the convex polytope consisting of those (x1 , . . . , xd ) ∈ Rd such that:
1. 0 ≤ xi ≤ 1 for 1 ≤ i ≤ d;
2. xi ≥ x j if i ≤ j in P.
The chain polytope C P is defined to be the convex polytope consisting of those
(x1 , . . . , xd ) ∈ Rd such that:
1. xi ≥ 0 for 1 ≤ i ≤ d;
2. xi1 + · · · + xik ≤ 1 for every maximal chain i 1 < · · · < i k of P.
For each subset I ⊂ [d], we define the (0, 1)-vectors ρ(I ) := i∈I ei . In particu-
lar ρ(∅) is the origin 0 of Rd . Both order polytopes and chain polytopes are (0, 1)-
polytopes of dimension d. In fact, in [26, Corollary 1.3 and Theorem 2.2], it is shown
that
Moreover, both order polytopes and chain polytopes are compressed, hence, they
possess the integer decomposition property. However, the class of order polytopes is
different from the class of chain polytopes [6, Example 3.5 and Corollary 3.9].
Now, we consider the two lattice polytopes (O P ) and (C P ) for a finite poset P
on [d]. In particular, we will see that the toric ideals of these lattice polytopes are
squarefree with respect to some monomial orders on K [x]. Let
denote the polynomial rings over K , and define the surjective ring homomorphisms
πO and πC by the following:
356 A. Tsuchiya
πO (x I ) = tρ(I ∪{d+1}) s,
πO (y J ) = t−ρ(J ∪{d+1}) s, and
πO (z) = s;
Here I, J ∈ J(P). Then the toric ideal I(O P ) (resp. I(C P ) is the kernel of πO
(resp. πC ).
Next, we introduce monomial orders <O and <C , and GO and GC which are the
set of binomials. Let <O denote a reverse lexicographic order on K [O] satisfying:
1. z <O y J <O x I ;
2. x I <O x I if I ⊂ I ;
3. y J <O y J if J ⊂ J ;
and GO ⊂ K [O] the set of binomials:
x I x I − x I ∪I x I ∩I ,
y J y J − y J ∪J y J ∩J ,
x I y J − x I \{i} y J \{i} ,
x∅ y∅ − z 2 .
Here
1. I and I (resp. J and J ) are poset ideals of P which are incomparable in J(P);
17 The Reflexive Dimension of (0, 1)-Polytopes 357
Proposition 7 ([11, Propositions 2.2 and 2.4]) Let the notation be as above. Then GO
(resp. GC ) is a Gröbner basis of I(O P ) (resp. I(C P ) ) with respect to <O (resp. <C ).
Theorem 8 ([11, Theorem 1.3]) Let P be a finite poset on [d]. Then each of (O P )
and (C P ) is a reflexive polytope with a regular unimodular triangulation.
Hence we have:
Corollary 9 Let P be a finite poset on [d]. Then one has
rdim(O P ) = rdim(C P ) = d + 1.
Remark 10 In [11], larger families of reflexive polytopes are given. In fact, for two
finite posets P and Q with |P| = |Q| = d, the three lattice polytopes (O P , O Q ),
(O P , C Q ) and (C P , C Q ) are studied. By the same technique, we know that (O P ,
O Q ) is a reflexive polytope (with a regular unimodular triangulation) if and only if P
and Q have a common linear extension, and (O P , C Q ) and (C P , C Q ) are always
reflexive polytopes with regular unimodular triangulations [11, Theorem 1.3].
If P = Q, then we will write (P) := (P, P). The two lattice polytopes (P, Q)
and (P, Q) often have the same properties. In fact, (O P , O Q ) is a reflexive polytope
(with a regular unimodular triangulation) if and only if P and Q have a common
linear extension [7, Corollary 2.2]. Moreover, (O P , C Q ) and (C P , C Q ) are always
reflexive polytopes with regular unimodular triangulations; see [10, Corollary 1.2]
and [9, Corollary 1.3]. In [9, 29], combinatorial properties of these polytopes, for
example, their volumes, are studied.
First, we recall what perfect graphs are and introduce the stable set polytopes of
finite simple graphs. Let G be a finite simple graph on the vertex set [d] and E(G)
the set of edges of G. (A finite graph G is called simple if G possesses no loop and
no multiple edge.) A subset W ⊂ [d] is called stable if, for all i and j belonging
to W with i = j, one has {i, j} ∈ / E(G). We remark that a stable set is often called
an independent set. A clique of G is a subset W ⊂ [d] which is a stable set of
the complementary graph G of G. The clique number ω(G) of G is the maximal
cardinality of a clique of G. The chromatic number χ (G) of G is the smallest
integer t ≥ 1 for which there exist stable set W1 , . . . , Wt of G with [d] = W1 ∪
· · · ∪ Wt . In general, it follows that ω(G) ≤ χ (G). A finite simple graph G is said to
be perfect [2] if, for any induced subgraph H of G including G itself, one has ω(H ) =
χ (G). Perfect graphs include many important classes of graphs, for example, chordal
graphs and comparability graphs. Moreover, it is known that the complementary
graph of a perfect graph is perfect. This characterization of perfect graphs is called
the perfect graph theorem. Recently, a stronger characterization of perfect graphs,
which is called the strong perfect graph theorem, is known. An odd hole is an induced
odd cycle of length ≥ 5 and an odd antihole is the complementary graph of an odd
hole.
Next, we introduce the stable set polytopes of finite simple graphs. Let S(G)
denote the set of stable sets of G. One has ∅ ∈ S(G) and {i} ∈ S(G) for each i ∈ [d].
The stable set polytope QG ⊂ Rd of G is the (0, 1)-polytope which is the convex hull
of {ρ(W ) : W ∈ S(G)} in Rd . Then the dimension QG is equal to d. It is known that
every chain polytope is a stable set polytope. In fact, let P be a finite poset on [d].
Its comparability graph G P is the finite simple graph on [d] such that {i, j} ∈ E(G P )
if and only if i < j or j < i in P. Then a stable set of G P corresponds to an antichain
of P. Moreover, one has C P = QG P . Since every comparability graph is perfect, the
class of the chain polytopes is contained in the class of the stable set polytopes of
perfect graphs. We see a characterization of perfect graphs in terms of the stable set
polytopes.
Proposition 13 ([19, Example 1.3 (c)]) Let G be a finite simple graph on [d].
Then G is perfect if and only if QG is compressed.
Now, we consider the lattice polytope (QG ) for a perfect graph G on [d]. In
particular, we see that the toric ideal of this lattice polytope is squarefree with respect
to some monomial order on K [x]. Let
denote the polynomial ring over K and define the surjective ring homomorphism
πQ : K [Q] → K [(QG )] by:
17 The Reflexive Dimension of (0, 1)-Polytopes 359
πQ (x S ) = tρ(S∪{d+1}) s,
πQ (yT ) = t−ρ(T ∪{d+1}) s, and
πQ (z) = s.
By combining Lemma 5 and Proposition 14, we can show that (QG ) is a reflexive
polytope with a regular unimodular triangulation. Moreover, we can give a polyhedral
characterization of perfect graphs.
Theorem 15 ([12, Theorem 1.1 (b)]) Let G be a finite simple graph on [d]. Then
the following arguments are equivalent:
1. G is perfect;
2. (QG ) is a reflexive polytope with a regular unimodular triangulation;
3. (QG ) has a regular unimodular triangulation.
Remark 17 In [12], a lager family of reflexive polytopes is given. In fact, for two
finite simple graphs G 1 and G 2 on [d], the lattice polytope (QG 1 , QG 2 ) is studied.
By the same technique, it follows from [12, Theorem 1.1 (b)] that the following
arguments are equivalent:
1. G 1 and G 2 are perfect;
2. (QG 1 , QG 2 ) is a reflexive polytope with a regular unimodular triangulation;
3. (QG 1 , QG 2 ) has a regular unimodular triangulation.
360 A. Tsuchiya
Remark 18 In [21], the lattice polytope (QG 1 , QG 2 ) for two finite simple graphs G 1 ,
G 2 on [d] is studied. The two lattice polytopes (QG 1 , QG 2 ) and (QG 1 , QG 2 ) have
a very similar but different property. In fact, it follows from [21, Theorem 2.8] that
the following arguments are equivalent:
1. G 1 and G 2 are perfect;
2. (QG 1 , QG 2 ) is a reflexive polytope with a regular unimodular triangulation;
3. (QG 1 , QG 2 ) is a reflexive polytope.
Remark 19 In [13], the lattice polytopes (QG , O P ) and (QG , O P ) for a finite
simple graph G on [d] and a finite poset P on [d] are studied. In fact, it follows
from [13, Theorem 1.2] that the following arguments are equivalent:
1. G is perfect;
2. (QG , O P ) has a regular unimodular triangulation;
3. (QG , O P ) is a reflexive polytope.
In particular, if G is perfect, then each of (QG , O P ) and (QG , O P ) is a reflexive
polytope with a regular unimodular triangulation.
Then for any subset I = {i 1 , . . . , i d+1 } ⊂ [n] with det(X I ) = 0, it follows that for
some integer 0 ≤ s ≤ d,
⎛ ⎞⎫
1 ⎬
⎜ .. ⎟ s
⎜ .
⎜
⎜ 1
0 ⎟⎭
⎟
⎟
XI ∼ ⎜
⎜
⎟ .
⎟
⎜ 2 ⎟
⎜ ⎟
⎝ 0 ..
.
2
⎠
By combining Lemma 20 and Proposition 21, we can obtain the following theo-
rem.
The class of lattice polytopes which satisfy the condition of Theorem 22 contains
a well-known large family of (0, 1)-polytopes, namely the edge polytopes of finite
simple graphs. Let G be a simple graph on [d]. The edge polytope PG ⊂ Rd of G
is the convex hull of all vectors ei + e j such that {i, j} ∈ E(G). This means that
the edge polytope of PG of G is the convex hull of all row vectors of the incidence
matrix A G of G, where A G is the matrix in {0, 1} E(G)×[d] with
362 A. Tsuchiya
1 if v ∈ e,
ae,v =
0 otherwise.
Corollary 23 Let G be a finite simple graph on [d]. Then one has rdim(PG ) =
dim(PG ) + 1.
In general, an edge polytope does not possess the integer decomposition property.
In fact, it is known when the edge polytope of a connected finite simple graph
possesses the integer decomposition property.
Then G satisfies the condition of Proposition 24. Hence PG possesses the integer
decomposition property. However, (PG ) does not possess the integer decomposi-
tion property.
Theorem 26 ([16, Theorem 3.2]) Let G be a connected finite simple graph on [d].
Then (PG ) possesses the integer decomposition property if and only if G does not
contain two disjoint odd cycles.
17 The Reflexive Dimension of (0, 1)-Polytopes 363
Remark 28 In [20], the lattice polytope (PG ) for a connected finite simple graph G
is studied. In fact, if G does not contain two disjoint odd cycles, then (PG ) is a
reflexive polytope which possesses the integer decomposition property. However,
in general, (PG ) is not always reflexive. For instance, if G is the connected finite
simple graph which appears in Example 25, then (PG ) is not reflexive and does not
possess the integer decomposition property.
Acknowledgements The author is very grateful to the anonymous referee for his or her insightful
report that led to significant improvements of the form of the paper. This manuscript was pre-
pared as a contribution to the conference proceedings of the Interactions with Lattice Polytopes at
Otto-von-Guericke-Universität Magdeburg on September 14th – 16th, 2017. The author thanks the
organizers, Christopher Borger, Alexander Kasprzyk, Benjamin Nill and Johannes Hofscheier, for
their support during this conference. The author was partially supported by Grant-in-Aid for JSPS
Fellows 16J01549.
References
1. Batyrev, V.V.: Dual polyhedra and mirror symmetry for Calabi-Yau hypersurfaces in toric
varieties. J. Algebraic Geom. 3(3), 493–535 (1994)
2. Chudnovsky, M., Robertson, N., Seymour, P., Thomas, R.: The strong perfect graph theorem.
Ann. of Math. (2) 164(1), 51–229 (2006)
3. Cox, D.A., Little, J.B., Schenck, H.K.: Toric varieties, Graduate Studies in Mathematics, vol.
124. American Mathematical Society, Providence (2011)
4. Haase, C., Melnikov, I.V.: The reflexive dimension of a lattice polytope. Ann. Comb. 10(2),
211–217 (2006)
5. Hibi, T.: A quick introduction to Gröbner bases. In: Gröbner bases, pp. 1–54. Springer, Tokyo
(2013)
6. Hibi, T., Li, N., Li, T.X., Mu, L.L., Tsuchiya, A.: Order-chain polytopes. Ars Math. Contemp.
16(2), 299–317 (2019)
7. Hibi, T., Matsuda, K.: Quadratic Gröbner bases of twinned order polytopes. Eur. J. Combin.
54, 187–192 (2016)
8. Hibi, T., Matsuda, K., Ohsugi, H., Shibata, K.: Centrally symmetric configurations of order
polytopes. J. Algebra 443, 469–478 (2015)
9. Hibi, T., Matsuda, K., Tsuchiya, A.: Gorenstein Fano polytopes arising from order polytopes
and chain polytopes (2015). arXiv:1507.03221 [math.CO]
10. Hibi, T., Matsuda, K., Tsuchiya, A.: Quadratic Gröbner bases arising from partially ordered
sets. Math. Scand. 121(1), 19–25 (2017)
11. Hibi, T., Tsuchiya, A.: Facets and volume of Gorenstein Fano polytopes. Math. Nachr. 290(16),
2619–2628 (2017)
12. Hibi, T., Tsuchiya, A.: Reflexive polytopes arising from perfect graphs. J. Combin. Theory Ser.
A 157, 233–246 (2018)
13. Hibi, T., Tsuchiya, A.: Reflexive polytopes arising from partially ordered sets and perfect
graphs. J. Algebraic Combin. 49(1), 69–81 (2019)
364 A. Tsuchiya
14. Kreuzer, M., Skarke, H.: Complete classification of reflexive polyhedra in four dimensions.
Adv. Theor. Math. Phys. 4(6), 1209–1230 (2000)
15. Lagarias, J.C., Ziegler, G.M.: Bounds for lattice polytopes containing a fixed number of interior
points in a sublattice. Canad. J. Math. 43(5), 1022–1035 (1991)
16. Nagaoka, T., Tsuchiya, A.: Reflexive polytopes arising from edge polytopes. Linear Algebra
Appl. 557, 438–454 (2018)
17. Ohsugi, H., Herzog, J., Hibi, T.: Combinatorial pure subrings. Osaka J. Math. 37(3), 745–757
(2000)
18. Ohsugi, H., Hibi, T.: Normal polytopes arising from finite graphs. J. Algebra 207(2), 409–426
(1998)
19. Ohsugi, H., Hibi, T.: Convex polytopes all of whose reverse lexicographic initial ideals are
squarefree. Proc. Amer. Math. Soc. 129(9), 2541–2546 (2001)
20. Ohsugi, H., Hibi, T.: Centrally symmetric configurations of integer matrices. Nagoya Math. J.
216, 153–170 (2014)
21. Ohsugi, H., Hibi, T.: Reverse lexicographic squarefree initial ideals and Gorenstein Fano poly-
topes. J. Commut. Algebra 10(2), 171–186 (2018)
22. Ohsugi, H., Tsuchiya, A.: Enriched chain polytopes. Israel J. Math. 237(1), 485–500 (2020)
23. Ohsugi, H., Tsuchiya, A.: Reflexive polytopes arising from bipartite graphs with γ -positivity
associated to interior polynomials. Selecta Math. (N.S.) 26(4), Paper No. 59, 22 (2020)
24. Ohsugi, H., Tsuchiya, A.: Enriched order polytopes and enriched Hibi rings. Eur. J. Math. 7(1),
48–68 (2021)
25. Schrijver, A.: Theory of linear and integer programming. Wiley-Interscience Series in Discrete
Mathematics. Wiley, Chichester (1986). A Wiley-Interscience Publication
26. Stanley, R.P.: Two poset polytopes. Discrete Comput. Geom. 1(1), 9–23 (1986)
27. Sturmfels, B.: Gröbner bases and convex polytopes. University Lecture Series, vol. 8. American
Mathematical Society, Providence (1996)
28. Sullivant, S.: Compressed polytopes and statistical disclosure limitation. Tohoku Math. J. (2)
58(3), 433–445 (2006)
29. Tsuchiya, A.: Volume, facets and dual polytopes of twinned chain polytopes. Ann. Comb.
22(4), 875–884 (2018)