11.07.2014 Views

Some aspects of toric topology - Library(ISI Kolkata) - Indian ...

Some aspects of toric topology - Library(ISI Kolkata) - Indian ...

Some aspects of toric topology - Library(ISI Kolkata) - Indian ...

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

<strong>Some</strong> <strong>aspects</strong> <strong>of</strong> <strong>toric</strong> <strong>topology</strong><br />

Soumen Sarkar<br />

<strong>Indian</strong> Statistical Institute, <strong>Kolkata</strong><br />

January,2011


<strong>Some</strong> <strong>aspects</strong> <strong>of</strong> <strong>toric</strong> <strong>topology</strong><br />

Soumen Sarkar<br />

Thesis submitted to the <strong>Indian</strong> Statistical Institute<br />

in partial fulfilment <strong>of</strong> the requirements<br />

for the award <strong>of</strong> the degree <strong>of</strong><br />

Doctor <strong>of</strong> Philosophy.<br />

January, 2011<br />

<strong>Indian</strong> Statistical Institute<br />

203, B.T. Road, <strong>Kolkata</strong>, India.


To my family


Acknowledgements<br />

Having written the last word <strong>of</strong> my thesis, I would like to take a few minutes to<br />

express my sincere thanks to those who have shaped my experience in graduate school.<br />

First and foremost, it is great pleasure to express my heart-felt gratitude to my<br />

supervisor Pr<strong>of</strong>. Mainak Poddar, without his hands-on guidance this thesis would not<br />

be possible. I thank him for accepting me as his Ph.D student and his involvement<br />

in my research. His patience, guidance and encouragement and lots <strong>of</strong> ideas increased<br />

my interest to do more research. Throughout this graduate course he provided friendly<br />

company and fruitful advices <strong>of</strong> mathematical and non-mathematical nature. He has<br />

given more <strong>of</strong> his time than I had any right to expect. To have been able to work so<br />

closely with such a motivating, caring and enthusiastic mathematician has been a great<br />

experience. I also thank him for allowing me to include the portion <strong>of</strong> our joint work in<br />

this thesis.<br />

I would like to express my deep indebtedness to Pr<strong>of</strong>. Goutam Mukherjee, Pr<strong>of</strong>.<br />

Parameswaran Sankaran and Pr<strong>of</strong>. Andrew Baker for their many helpful suggestion for<br />

improvement in my research.<br />

I am very grateful to Pr<strong>of</strong>. Amiya mukherjee, Pr<strong>of</strong>. Amartya Dutta, Pr<strong>of</strong>.S.C.Bagchi,<br />

Pr<strong>of</strong>. R. V. Gurjar, Pr<strong>of</strong>. T.N. Venkataramana, Pr<strong>of</strong>.S.M.Srivastava, Pr<strong>of</strong>. Debasis<br />

Goswami, Pr<strong>of</strong>. Mahuya Dutta and Pr<strong>of</strong>. Arup Bose for teaching me various courses.<br />

I would like to thank my teachers <strong>of</strong> R.K.Mission Vidyamandira, Belur (Howrah)<br />

who sparked my interest in mathematics during my college days.<br />

I express my gratitude to my teachers <strong>of</strong> C.U. (<strong>Kolkata</strong>) for their encouragements.<br />

I would be remiss not to acknowledge the support and co-operation <strong>of</strong> my friends<br />

Debasis, Arup, Swarup, Rabeyadi, Paruida, Ashisda, Prosenjitda, Santanuda, Subhojit,<br />

Radhe, Biswarup, Santanu, Debargha, Ronnie, Shameek and many others.<br />

I thank all members <strong>of</strong> Stat-Math unit who create pleasant workable environment.<br />

I wish to thank <strong>Indian</strong> Statistical Institute (<strong>Kolkata</strong>), TIFR (Mumbai) and National<br />

Board for Higher Mathematics (India) for the generous financial support.<br />

Finally I would like to thank my family members, my sister and her family for their<br />

unquestioned support in good and difficult times. A very special thank goes out to my<br />

wife, Moumita for her blind affection and invaluable support. Without her support I<br />

would have never been accomplished what I did.


Contents<br />

0 Introduction 1<br />

1 Quasi<strong>toric</strong> manifolds 5<br />

1.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5<br />

1.2 Definition and examples . . . . . . . . . . . . . . . . . . . . . . . . . . . 5<br />

1.3 Invariant closed submanifolds . . . . . . . . . . . . . . . . . . . . . . . . 14<br />

1.4 Face vectors and face ring <strong>of</strong> polytopes . . . . . . . . . . . . . . . . . . . 15<br />

1.5 Homology groups <strong>of</strong> quasi<strong>toric</strong> manifolds . . . . . . . . . . . . . . . . . . 16<br />

1.6 Orientation <strong>of</strong> quasi<strong>toric</strong> manifolds . . . . . . . . . . . . . . . . . . . . . 17<br />

1.7 Equivariant connected sums . . . . . . . . . . . . . . . . . . . . . . . . . 18<br />

1.8 Cohomology ring . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21<br />

1.9 Smooth and stable complex structure . . . . . . . . . . . . . . . . . . . . 26<br />

1.10 Chern classes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31<br />

2 Small covers and orbifolds 33<br />

2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33<br />

2.2 Small covers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33<br />

2.3 Classical effective orbifolds . . . . . . . . . . . . . . . . . . . . . . . . . . 37<br />

2.4 Tangent bundle <strong>of</strong> orbifolds . . . . . . . . . . . . . . . . . . . . . . . . . 39<br />

2.5 Orbifold fundamental group . . . . . . . . . . . . . . . . . . . . . . . . . 44<br />

3 Quasi<strong>toric</strong> orbifolds 47<br />

3.1 Definition by construction and orbifold structure . . . . . . . . . . . . . 47<br />

3.2 Axiomatic definition <strong>of</strong> quasi<strong>toric</strong> orbifolds . . . . . . . . . . . . . . . . 50<br />

3.3 Characteristic subspaces . . . . . . . . . . . . . . . . . . . . . . . . . . . 52<br />

3.4 Orbifold fundamental group . . . . . . . . . . . . . . . . . . . . . . . . . 53<br />

3.5 q-Cellular homology groups . . . . . . . . . . . . . . . . . . . . . . . . . 57<br />

3.6 Rational homology <strong>of</strong> quasi<strong>toric</strong> orbifolds . . . . . . . . . . . . . . . . . 58<br />

3.7 Gysin sequence for q-sphere bundle . . . . . . . . . . . . . . . . . . . . . 59<br />

3.8 Cohomology ring <strong>of</strong> quasi<strong>toric</strong> orbifolds . . . . . . . . . . . . . . . . . . 60<br />

v


3.9 Stable almost complex structure . . . . . . . . . . . . . . . . . . . . . . 63<br />

3.10 Line bundles and cohomology . . . . . . . . . . . . . . . . . . . . . . . . 67<br />

3.11 Chern numbers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68<br />

3.12 Chen-Ruan cohomology groups . . . . . . . . . . . . . . . . . . . . . . . 69<br />

4 Small orbifolds over simple polytopes 71<br />

4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71<br />

4.2 Definition and orbifold structure . . . . . . . . . . . . . . . . . . . . . . 71<br />

4.3 Orbifold fundamental group . . . . . . . . . . . . . . . . . . . . . . . . . 78<br />

4.4 Homology and Euler characteristic . . . . . . . . . . . . . . . . . . . . . 83<br />

4.5 Cohomology ring <strong>of</strong> small orbifolds . . . . . . . . . . . . . . . . . . . . . 88<br />

5 T 2 -cobordism <strong>of</strong> quasi<strong>toric</strong> 4-manifolds 95<br />

5.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95<br />

5.2 Edge-Simple Polytopes . . . . . . . . . . . . . . . . . . . . . . . . . . . 95<br />

5.3 Manifolds with quasi<strong>toric</strong> boundary . . . . . . . . . . . . . . . . . . . . 97<br />

5.4 Manifolds with small cover boundary . . . . . . . . . . . . . . . . . . . . 100<br />

5.5 <strong>Some</strong> observations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101<br />

5.6 Orientability <strong>of</strong> W (Q P , λ) . . . . . . . . . . . . . . . . . . . . . . . . . . 102<br />

5.7 Torus cobordism <strong>of</strong> quasi<strong>toric</strong> manifolds . . . . . . . . . . . . . . . . . . 104<br />

6 Oriented cobordism <strong>of</strong> CP 2k+1 113<br />

6.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113<br />

6.2 <strong>Some</strong> manifolds with quasi<strong>toric</strong> boundary . . . . . . . . . . . . . . . . . 113<br />

6.3 Orientability <strong>of</strong> W (△ n Q<br />

, η) . . . . . . . . . . . . . . . . . . . . . . . . . . 118<br />

6.4 Oriented cobordism <strong>of</strong> CP 2k+1 . . . . . . . . . . . . . . . . . . . . . . . . 119<br />

Bibliography 123


Chapter 0<br />

Introduction<br />

The main goal <strong>of</strong> this thesis is to study the <strong>topology</strong> <strong>of</strong> torus actions on manifolds<br />

and orbifolds. In algebraic geometry actions <strong>of</strong> the torus (C ∗ ) n on algebraic varieties<br />

with nice properties produce bridges between geometry and combina<strong>toric</strong>s see [Dan78],<br />

[Oda88] and [Ful93]. We see a similar bridge called moment map for Hamiltonian action<br />

<strong>of</strong> compact torus on symplectic manifolds see [Aud91] and [Gui94]. In particular whenever<br />

the manifold is compact the image <strong>of</strong> moment map is a simple polytope, the orbit<br />

space <strong>of</strong> the action. A topological counterpart called quasi<strong>toric</strong> manifolds, a class <strong>of</strong><br />

topological manifolds with compact torus action having combinatorial orbit space, were<br />

introduced by Davis and Januskiewicz in [DJ91]. A class <strong>of</strong> examples <strong>of</strong> quasi<strong>toric</strong> manifolds<br />

are nonsingular projective <strong>toric</strong> varieties, introduced by M. Demazure [Dem70].<br />

There are many properties <strong>of</strong> quasi<strong>toric</strong> manifolds akin to that <strong>of</strong> nonsingular complete<br />

<strong>toric</strong> varieties. The combinatorial formula for the cohomology ring <strong>of</strong> a nonsingular complete<br />

<strong>toric</strong> variety is analogous to the formula for quasi<strong>toric</strong> manifolds. Their K-theories<br />

described by P. Sankaran and V. Uma in [SU03] and [SU07] are also similar. The survey<br />

[BP02] is a good reference for many interesting developments and applications <strong>of</strong><br />

quasi<strong>toric</strong> manifolds.<br />

Inspired by the work <strong>of</strong> [DJ91] we generalize these quasi<strong>toric</strong> manifolds to quasi<strong>toric</strong><br />

orbifolds with compact torus action. We have studied structures and topological invariants<br />

<strong>of</strong> quasi<strong>toric</strong> orbifolds. In addition, we have introduced a class <strong>of</strong> n-dimensional<br />

orbifolds with Z n−1<br />

2 -action with nice combinatorial description. We have also given two<br />

applications <strong>of</strong> quasi<strong>toric</strong> manifolds to cobordism theory. This section briefly introduces<br />

the main ingredients <strong>of</strong> this thesis. We will meet all in much greater detail in the<br />

following chapters.<br />

We recall the definitions and topological invariants namely homology groups, cohomology<br />

rings and Chern classes <strong>of</strong> quasi<strong>toric</strong> manifolds in Chapter 1. A quasi<strong>toric</strong><br />

manifold M 2n is an even dimensional smooth manifold with a locally standard action<br />

<strong>of</strong> the compact torus T n = U(1) n such that the orbit space has the structure <strong>of</strong> an<br />

1


Chapter 0: Introduction 2<br />

n-dimensional simple polytope. Cohomology rings <strong>of</strong> these manifolds can be computed<br />

using equivariant cohomology [DJ91]. As a special case, one obtain the cohomology<br />

rings <strong>of</strong> nonsingular projective <strong>toric</strong> varieties without recourse to algebraic geometry.<br />

Buchstaber and Ray [BR01] showed the existence <strong>of</strong> smooth and stable almost complex<br />

structure on quasi<strong>toric</strong> manifolds. We present a different pro<strong>of</strong> <strong>of</strong> this following [Poddf].<br />

In Chapter 2 we recall the definitions <strong>of</strong> small covers and orbifolds. The category<br />

<strong>of</strong> small covers was introduced by Davis and Januszkiewicz [DJ91]. Following them we<br />

discuss some basic theory about small covers. The remaining sections <strong>of</strong> this chapter<br />

describe the definition, tangent bundle and orbifold fundamental group <strong>of</strong> orbifolds<br />

following [ALR07]. Orbifolds were introduced by Satake [Sat57], who called them V -<br />

manifolds. Precisely, orbifolds are singular spaces that locally look like the quotient <strong>of</strong><br />

an open subset <strong>of</strong> Euclidean space by an action <strong>of</strong> a finite group.<br />

In Chapter 3 we study topological invariants and stable almost complex structure<br />

on quasi<strong>toric</strong> orbifolds. We discuss equivalent definitions <strong>of</strong> quasi<strong>toric</strong> orbifolds, one is<br />

an axiomatic definition <strong>of</strong> a quasi<strong>toric</strong> orbifold via locally standard action and the other<br />

is constructive. Our constructive definition uses the combinatorial model (Q, N, {λ i }),<br />

where Q is a simple polytope, N is a free Z-module <strong>of</strong> finite rank, λ i is an assignment<br />

<strong>of</strong> a vector in N to each facet F i <strong>of</strong> Q satisfying certain conditions.<br />

Let ̂N be the<br />

submodule <strong>of</strong> N generated by the vectors λ i . We construct orbifold universal cover and<br />

compute the orbifold fundamental group <strong>of</strong> quasi<strong>toric</strong> orbifold.<br />

Theorem 0.0.1 (Theorem 3.2, [PS10]). The orbifold fundamental group π orb<br />

1 (X ) <strong>of</strong> the<br />

quasi<strong>toric</strong> orbifold X is isomorphic to N/ ̂N.<br />

We compute the homology <strong>of</strong> quasi<strong>toric</strong> orbifolds with coefficients in Q. For this<br />

we need to generalize the notion <strong>of</strong> CW -complex a little. We introduce the notion <strong>of</strong><br />

q-CW complex where an open cell is the quotient <strong>of</strong> an open disk by action <strong>of</strong> a finite<br />

group. We compute the rational cohomology ring <strong>of</strong> a quasi<strong>toric</strong> orbifold and show that<br />

it is isomorphic to a quotient <strong>of</strong> the Stanley-Reisner face ring <strong>of</strong> the base polytope Q,<br />

Theorem 3.8.2.<br />

We show the existence <strong>of</strong> a stable almost complex structure on a quasi<strong>toric</strong> orbifold<br />

corresponding to any given omniorientation. The universal orbifold cover <strong>of</strong> the quasi<strong>toric</strong><br />

orbifold is used here. As in the manifold case, we show that the cohomology<br />

ring is generated by the first Chern classes <strong>of</strong> some complex rank one orbifold vector<br />

bundles, canonically associated to facets <strong>of</strong> Q, see Section 3.10. We compute the top<br />

Chern number <strong>of</strong> an omnioriented quasi<strong>toric</strong> orbifold. We give a necessary condition for<br />

existence <strong>of</strong> torus invariant almost complex structure. Whether this condition is also<br />

sufficient remains open. Finally we compute the Chen-Ruan cohomology groups <strong>of</strong> an<br />

almost complex quasi<strong>toric</strong> orbifold.<br />

In Chapter 4 we introduce some n-dimensional smooth orbifolds on which there


3<br />

is Z n−1<br />

2 -action having a simple polytope as the orbit space. We call these orbifolds<br />

small orbifolds. The classification problem <strong>of</strong> small orbifolds remains open. We show<br />

that the orbifold universal cover <strong>of</strong> n-dimensional (n > 2) small orbifold is R n . We<br />

also show the space Z, constructed in Lemma 4.4 <strong>of</strong> [DJ91] associated to a simple n-<br />

polytope Q (n > 2), is diffeomorphic to R n if there is an s-characteristic function on Q<br />

(Theorem 3.3, [Sar10b]). The converse is an interesting open question. One application<br />

<strong>of</strong> this result is the following. The space Z corresponding to the n-simplex (n > 2) is<br />

homeomorphic to the n-sphere. So there does not exist any s-characteristic function <strong>of</strong><br />

n-simplex. Consequently there does not exist any small orbifold with the n-dimensional<br />

simplex as orbit space when n > 2. We calculate the orbifold fundamental group <strong>of</strong><br />

n-dimensional small orbifolds. We compute the homology groups <strong>of</strong> small orbifolds in<br />

terms <strong>of</strong> h-vector (see [DJ91]) <strong>of</strong> the polytope. When Q is an even dimensional simple<br />

polytope then small orbifolds over Q are orientable. We compute the cohomology rings<br />

<strong>of</strong> even dimensional small orbifolds.<br />

In Chapter 5 we compute the T 2 -cobordism group in the following category: the<br />

objects are all 4-dimensional quasi<strong>toric</strong> manifolds and morphisms are T 2 equivariant<br />

maps between quasi<strong>toric</strong> 4-manifolds. In this chapter we introduce a particular type <strong>of</strong><br />

polytope, which we call edge-simple polytope.<br />

Definition 0.0.2. An n-dimensional convex polytope P is called an n-dimensional edgesimple<br />

polytope if each edge <strong>of</strong> P is the intersection <strong>of</strong> exactly (n − 1) codimension one<br />

faces <strong>of</strong> P .<br />

The study <strong>of</strong> <strong>topology</strong> and combina<strong>toric</strong>s <strong>of</strong> these polytopes would be interesting.<br />

However we have not dealt with these questions in this thesis.<br />

We introduce the notion <strong>of</strong> isotropy function on the set <strong>of</strong> facets F(P ) <strong>of</strong> an n-<br />

dimensional edge-simple polytope P .<br />

Definition 0.0.3. A function λ: F(P ) → Z n−1 is called an isotropy function <strong>of</strong> the<br />

edge-simple polytope P if the set <strong>of</strong> vectors {λ(F i1 ), . . . , λ(F in−1 )} form a basis <strong>of</strong> Z n−1<br />

whenever the intersection <strong>of</strong> the facets {F i1 , . . . , F in−1 } is an edge <strong>of</strong> P .<br />

Deleting a suitable neighborhood <strong>of</strong> each vertex <strong>of</strong> P we get a simple n-polytope P Q .<br />

We construct an (2n − 1)-manifold with quasi<strong>toric</strong> boundary from simple n-polytope<br />

P Q and an isotropy function λ <strong>of</strong> P . There is a natural T n−1 -action on these manifolds<br />

with quasi<strong>toric</strong> boundary having P Q as the orbit space. We show that these manifolds<br />

with quasi<strong>toric</strong> boundary are orientable and compute their Euler characteristic.<br />

Now consider n = 2. Let π : M → Q be a quasi<strong>toric</strong> 4-manifold with the characteristic<br />

pair (Q, η). Here Q is a simple 2-polytope and η is the assignment, called<br />

characteristic function <strong>of</strong> M, <strong>of</strong> isotropy group to each facet <strong>of</strong> Q. Suppose the number<br />

<strong>of</strong> codimension one faces <strong>of</strong> Q is m. We construct an edge-simple 3-polytope P E such


Chapter 0: Introduction 4<br />

that P E has exactly one vertex which is the intersection <strong>of</strong> m codimension one faces<br />

and other vertices <strong>of</strong> P E are intersection <strong>of</strong> 3 codimension one faces. Extending the<br />

characteristic function η we define an isotropy function λ <strong>of</strong> P E . The pair (P E , λ) helps<br />

to construct an oriented T 2 manifold W with boundary<br />

∂W = M + k 1 CP 2 + k 2 CP 2<br />

where k 1 , k 2 are some integers. Thus we obtain the following lemma and theorem.<br />

Lemma 0.0.4 (Lemma 6.1, [Sar10c]). The T 2 -cobordism class <strong>of</strong> a Hirzebruch surface<br />

is trivial. In particular, oriented cobordism class <strong>of</strong> a Hirzebruch surface is also trivial.<br />

Lemma 0.0.5. Any 4-dimensional quasi<strong>toric</strong> manifold is equivariantly cobordant to the<br />

disjoint union ⊔ l 1 CP2 for some l, where the T 2 -action on different copies <strong>of</strong> CP 2 may<br />

be distinct.<br />

Theorem 0.0.6. The set {[CP 2 ξ ′] : [ξ] eq ∈ SL(2, Z)/ ∼ eq } is a set <strong>of</strong> generators <strong>of</strong> the<br />

oriented torus cobordism group CG 2 .<br />

In Chapter 6, we give a new pro<strong>of</strong> <strong>of</strong> the fact that the oriented cobordism class <strong>of</strong><br />

CP 2k+1 is trivial for each k ≥ 0 (Theorem 5.1 in [Sar10a]). The strategy <strong>of</strong> our pro<strong>of</strong><br />

is to first construct an odd dimensional compact orientable manifold with boundary,<br />

where the boundary is a disjoint union <strong>of</strong> three quasi<strong>toric</strong> manifolds. This involves<br />

an adaptation <strong>of</strong> the usual combinatorial method for constructing quasi<strong>toric</strong> manifolds.<br />

Moreover the combinatorial data is carefully chosen so that exactly one <strong>of</strong> the boundary<br />

components is CP 2k+1 while the other two components are identifiable by an orientation<br />

reversing homeomorphism.


Chapter 1<br />

Quasi<strong>toric</strong> manifolds<br />

1.1 Introduction<br />

In these chapter we recall the definitions and topological invariants <strong>of</strong> quasi<strong>toric</strong> manifolds.<br />

We follow [DJ91] for basic definitions, examples and a topological classification.<br />

We compute the homology group <strong>of</strong> quasi<strong>toric</strong> manifolds following [DJ91]. The next<br />

two sections deal with the study <strong>of</strong> orientability <strong>of</strong> quasi<strong>toric</strong> manifolds and the concept<br />

<strong>of</strong> connected sum <strong>of</strong> quasi<strong>toric</strong> manifolds respectively. The computation <strong>of</strong> the cohomology<br />

ring <strong>of</strong> quasi<strong>toric</strong> manifolds are explained following the work [DJ91]. The work<br />

<strong>of</strong> Buchstaber and Ray [BR01] on the existence <strong>of</strong> smooth and stable almost complex<br />

structure are the object <strong>of</strong> Section 1.9. Following this up one can show the existence <strong>of</strong><br />

Chern classes as an application <strong>of</strong> the stable almost complex structure and describe the<br />

formulae for top Chern number.<br />

1.2 Definition and examples<br />

Quasi<strong>toric</strong> manifolds are essentially an even dimensional manifolds M 2n . The torus<br />

T n = U(1) n action on M 2n must be effective and locally resemble the standard action<br />

<strong>of</strong> T n on C n up to an automorphism <strong>of</strong> T n .<br />

Definition 1.2.1. An n-dimensional simple polytope in R n is a convex polytope where<br />

exactly n bounding hyperplanes meet at each vertex. The codimension one faces <strong>of</strong><br />

convex polytope are called facets.<br />

The ready examples <strong>of</strong> simple polytopes are simplices and cubes. Through out the<br />

thesis Q stands for simple polytope.<br />

Definition 1.2.2. An action <strong>of</strong> T n on a 2n-dimensional manifold M 2n is said to be<br />

locally standard if every point y ∈ M has a T n -stable open neighborhood U y and a home-<br />

5


Chapter 1: Quasi<strong>toric</strong> manifolds 6<br />

omorphism ψ : U y → V , where V is a T n -stable open subset <strong>of</strong> C n and an isomorphism<br />

δ y : T n → T n such that ψ(t · x) = δ y (t) · ψ(x) for all (t, x) ∈ T n × U y .<br />

Definition 1.2.3. A T n -manifold M 2n is called quasi<strong>toric</strong> manifold over a simple polytope<br />

Q if the following conditions are satisfied:<br />

1. the T n action is locally standard,<br />

2. there is a projection map p : M 2n → Q constant on T n orbits which maps every<br />

l-dimensional orbit to a point in the relative interior <strong>of</strong> a l-dimensional face <strong>of</strong> Q.<br />

Definition 1.2.4. A quasi<strong>toric</strong> manifold having a smooth structure such that the action<br />

<strong>of</strong> torus is smoothly locally standard is called smooth quasi<strong>toric</strong> manifold.<br />

Example 1.2.5. Consider the complex projective space CP 2 ∼ = (C 3 − {0})/C ∗ ∼ = S 5 /S 1 .<br />

Denote the coordinates on C 3 by (z 1 , z 2 , z 3 ). So we may represent S 5 by the set<br />

{|z 1 | 2 + |z 2 | 2 + |z 3 | 2 = 1}<br />

and S 1 by the subgroup {α · I 3 : |α| = 1} <strong>of</strong> U(1) 3 where I 3 is a rank 3 identity matrix.<br />

The points <strong>of</strong> T 2 ∼ = U(1) 3 /S 1 and CP 2 can be identified to the class [t 1 : t 2 : t 3 ] and<br />

[z 1 : z 2 : z 3 ] respectively. The natural action <strong>of</strong> T 2 on CP 2 and C 2 are the following<br />

maps T 2 × CP 2 → CP 2 and T 2 × C 2 → C 2 , defined by<br />

([t 1 : t 2 : t 3 ], [z 1 : z 2 : z 3 ]) → [t 1 z 1 : t 2 z 2 : t 3 z 3 ]<br />

and ([t 1 : t 2 : t 3 ], (z ′ 1 , z′ 2 )) → (t 1t 3 −1 z ′ 1 , t 2t 3 −1 z ′ 2 ) (1.2.1)<br />

respectively. Clearly [1 : 0 : 0], [0 : 1 : 0] and [0 : 0 : 1] are the only fixed points <strong>of</strong> T 2<br />

action on CP 2 . Let U j = {[z 1 : z 2 : z 3 ] ∈ CP 2 : z j ≠ 0} for j = 1, 2, 3 and<br />

ψ 1 : U 1 → C 2 , ψ 2 : U 2 → C 2 , ψ 3 : U 3 → C 2 (1.2.2)<br />

be the maps defined by<br />

ψ 1 ([z 1 : z 2 : z 3 ]) = (z 2 /z 1 , z 3 /z 1 ), ψ 2 ([z 1 : z 2 : z 3 ]) = (z 1 /z 2 , z 3 /z 2 )<br />

and ψ 3 ([z 1 : z 2 : z 3 ]) = (z 1 /z 3 , z 2 /z 3 )<br />

(1.2.3)<br />

respectively. The subsets U 1 , U 2 and U 3 are T 2 -stable covering open subsets <strong>of</strong> CP 2 . The<br />

maps ψ 1 , ψ 2 ψ 3 are homeomorphisms. The map ψ 1 satisfy the following relation.<br />

ψ 1 ([t 1 : t 2 : t 3 ]· [z 1 : z 2 : z 3 ]) = ((t 2 t −1<br />

1 z 2)/z 1 , (t 3 t −1<br />

1 z 3)/z 1 ) = [t 2 , t 3 , t 1 ]· (z 2 /z 1 , z 3 /z 1 )<br />

Let δ 1 , δ 2 and δ 3 are automorphism <strong>of</strong> T 2 defined by<br />

δ 1 ([t 1 : t 2 : t 3 ]) = [t 2 , t 3 , t 1 ], δ 2 ([t 1 : t 2 : t 3 ]) = [t 1 , t 3 , t 2 ] and δ 3 ([t 1 : t 2 : t 3 ]) = [t 1 , t 2 , t 3 ]


7 1.2 Definition and examples<br />

respectively. Hence we get the following relation for all [t 1 : t 2 : t 3 ] ∈ T 2 ,<br />

ψ 1 ([t 1 : t 2 : t 3 ]· [z 1 : z 2 : z 3 ]) = δ 1 ([t 1 : t 2 : t 3 ])· ψ 1 ([z 1 : z 2 : z 3 ]), ∀ [z 1 : z 2 : z 3 ] ∈ U 1 ,<br />

So ψ 1 is a δ 1 -equivariant map. Similarly we can show that ψ i is a δ i -equivariant map<br />

for i = 2, 3. Hence the natural T 2 action on CP 2 is a locally standard action. The orbit<br />

space can be identified to the triangle △ 2 = {(|z 1 | 2 , |z 2 | 2 ) ∈ R 2 : |z 1 | 2 + |z 2 | 2 ≤ 1}. In<br />

fact the projective space CP n is a quasi<strong>toric</strong> manifold over the n-dimensional simplex<br />

△ n for each n.<br />

Example 1.2.6. In example 1.2.5 we show that the projective space CP 1 is a quasi<strong>toric</strong><br />

manifold over the interval I 1 = {x ∈ R : 0 ≤ x ≤ 1}. Then the space (CP 1 ) n is a<br />

quasi<strong>toric</strong> manifold over the standard cube I n . Here the torus T n action on (CP 1 ) n is<br />

the diagonal action.<br />

By the definition 1.2.3 a zero dimensional orbit maps to a vertex <strong>of</strong> Q. Since the<br />

T n action is locally standard, the fixed point sets are parameterized by the vertex set<br />

<strong>of</strong> Q. Let F be a codimension k face <strong>of</strong> Q. We denote its relative interior by int(F ).<br />

The space p −1 (int(F )) is a trivial T n−k -bundle over p −1 (int(F )). The isotropy group<br />

T n x := {t ∈ T n : tx = x} (1.2.4)<br />

at each point x ∈ p −1 (int(F )) is locally constant on p −1 (int(F )). Since p −1 (int(F )) is<br />

product <strong>of</strong> two connected sets and so,<br />

T n x = T n y for all points x, y ∈ p −1 (int(F )). (1.2.5)<br />

These common isotropy group is denoted by T n F . The group Tn F is isomorphic to Tk .<br />

Hence the orbit <strong>of</strong> each point over the relative interior <strong>of</strong> codimension-k face is the<br />

(n − k)-dimensional subtorus <strong>of</strong> T n .<br />

Let F 1 , . . . , F m be the facets <strong>of</strong> Q, denoted by F(Q). So the isotropy subgroup <strong>of</strong> the<br />

preimage p −1 (int(F j )) is the 1-dimensional subgroup T n F j<br />

<strong>of</strong> T n . Each F j is also a simple<br />

polytope. Let M 2(n−1)<br />

j<br />

be the T n F j<br />

fixed subset <strong>of</strong> M 2n . Locally standardness <strong>of</strong> the<br />

manifold M 2n imply that the space p −1 (F j ) = M 2(n−1)<br />

j<br />

is a T n−1 ∼ = T n /T n F j<br />

manifold.<br />

The T n−1 action on M 2(n−1)<br />

j<br />

is locally standard and restriction <strong>of</strong> p on M 2(n−1)<br />

j<br />

satisfy<br />

is a 2(n − 1)-dimensional quasi<strong>toric</strong> manifold<br />

the condition (2) <strong>of</strong> 1.2.3. Hence M 2(n−1)<br />

j<br />

over F j , called the characteristic submanifold <strong>of</strong> M 2n corresponding to F j . The isotropy<br />

subgroup T n F j<br />

may be identified to the elements <strong>of</strong> T n as<br />

T n F j<br />

= {(e 2πiλ 1jr , . . . , e 2πiλ njr ) ∈ T n ; ∀ r ∈ R} (1.2.6)


Chapter 1: Quasi<strong>toric</strong> manifolds 8<br />

for some primitive vector λ j = (λ 1j , . . . , λ nj ) ∈ Z n . The correspondence T n F j<br />

vector λ j is one-to-one upto a sign. Hence we can define a function<br />

to the<br />

λ : F(Q) → Z n by λ(F j ) = λ j , (1.2.7)<br />

called the characteristic function <strong>of</strong> M 2n . The vectors λ j ’s are called the characteristic<br />

vector corresponding to F j .<br />

Since Q is a simple polytope, the codimension-k face F is the intersection <strong>of</strong> unique<br />

collection <strong>of</strong> k facets F j1 , . . . , F jk . Then the isotropy group T n F is Tn F j1<br />

× . . . × T n F jk<br />

. The<br />

group T n F is a direct summand <strong>of</strong> Tn , since comparing the action <strong>of</strong> T n on M 2n to the<br />

standard action <strong>of</strong> T n on C n we can conclude that the span <strong>of</strong> λ j1 , . . . , λ jk in Z n is a<br />

k-dimensional direct summand <strong>of</strong> Z n . In particular, when unique collection <strong>of</strong> n facets<br />

F j1 , . . . , F jn meet at a vertex <strong>of</strong> Q the corresponding characteristic vectors λ j1 , . . . , λ jn<br />

form a basis <strong>of</strong> Z n .<br />

Example 1.2.7. Suppose in example 1.2.5 the projection p : CP 2 → △ 2 maps<br />

[1, 0, 0] → (0, 0) = O, [0, 1, 0] → (1, 0) = A and [0, 0, 1] → (0, 1) = B.<br />

Hence the characteristic submanifolds are p −1 (OA) = {[z 1 : z 2 : 0] ∈ CP 2 }, p −1 (OB) =<br />

{[z 1 : 0 : z 3 ] ∈ CP 2 } and p −1 (AB) = {[0 : z 2 : z 3 ] ∈ CP 2 }. Suppose<br />

a = [1/ √ 2 : 1/ √ 2 : 0], b = [1/ √ 2 : 0 : 1/ √ 2] and c = [0 : 1/ √ 2 : 1/ √ 2].<br />

Then T 2 a = T 2 OA , T2 b<br />

= T 2 OB and T2 c = T 2 AB . Suppose the isomorphism T2 ∼ =<br />

U(1) 3 /S 1 is given by (t 1 , t 2 ) → [t 1 : t 2 : 1]. We can show that<br />

T 2 a = {(e 2πir , e 2πir ) ∈ T 2 : r ∈ R}, T 2 b = {(e2πi0 , e 2πir ) ∈ T 2 : r ∈ R}<br />

and T 2 c = {(e 2πir , e 2πi0 ) ∈ T 2 : r ∈ R}.<br />

Hence the characteristic function <strong>of</strong> CP 2 associated to the T 2 action <strong>of</strong> example 1.2.5 is<br />

the function λ : F(△ 2 ) → Z 2 such that λ(OA) = (1, 1), λ(OB) = (0, 1), λ(AB) = (1, 0).<br />

We give the definition <strong>of</strong> characteristic model combinatorially followed by the<br />

construction <strong>of</strong> quasi<strong>toric</strong> manifold from this characteristic model. Let Q be an n-<br />

dimensional simple polytope and F(Q) be the set <strong>of</strong> facets <strong>of</strong> Q.<br />

Definition 1.2.8. A function ξ : F(Q) → Z n is called characteristic function if the span<br />

<strong>of</strong> ξ(F j1 ), . . . , ξ(F jk ) is a k-dimensional direct summand <strong>of</strong> Z n whenever the intersection<br />

F j1 ∩ . . . ∩ F jk is nonempty.<br />

The vectors ξ j := ξ(F j ) are called characteristic vectors and the pair (Q, ξ) is called<br />

a characteristic model.


9 1.2 Definition and examples<br />

The orbit space <strong>of</strong> the standard T n action on C n is the positive octant R n ≥0 . Consider<br />

the canonical projection p c : T n × R n ≥0 → Cn given by<br />

(t 1 , . . . , t n ) = t × (x 1 , . . . , x n ) = x ↦−→ (t 1 x 1 , . . . t n x n ) = tx. (1.2.8)<br />

Let p c (t, x) = ˜x. The fiber over ˜x ∈ R n <strong>of</strong> the projection p c is the isotropy group <strong>of</strong> ˜x<br />

<strong>of</strong> the T n action on C n . Denote this isotropy group by T n x. So we can identify C n with<br />

the quotient space T n × R n ≥ / ∼ 0, where the equivalence relation ∼ 0 is defined by<br />

(t, x) ∼ 0 (s, y) if and only if x = y and ts −1 ∈ T n x. (1.2.9)<br />

Let for each vertex v ∈ Q, C v = {p ∈ F : F is a face <strong>of</strong> Q not containing v}. Let<br />

U v be the open subset <strong>of</strong> Q complement to the set C v . Hence U v is diffeomorphic to<br />

R n ≥0<br />

as manifold with corner. We consider an identifications on faces <strong>of</strong> the product<br />

T n ×U v similarly to the standard equivalence relation ∼ 0 for each vertex v. Gluing them<br />

naturally one can reconstruct a quasi<strong>toric</strong> manifold from any characteristic model.<br />

Theorem 1.2.9 (Subsection 1.5, [DJ91]). A quasi<strong>toric</strong> manifold can be constructed<br />

from a characteristic model.<br />

Pro<strong>of</strong>. Let Q be a simple n-polytope and (Q, ξ) be a characteristic model. A<br />

codimension-k face F <strong>of</strong> the polytope Q is the intersection F j1 ∩ . . . ∩ F jk <strong>of</strong> unique<br />

collection <strong>of</strong> k facets F j1 , . . . , F jk <strong>of</strong> Q. Let Z(F ) be the submodule <strong>of</strong> Z n generated by<br />

the characteristic vectors ξ(F j1 ), . . . , ξ(F jk ). The module Z(F ) is a direct summand <strong>of</strong><br />

Z n . Therefore the torus T F := (Z(F ) ⊗ Z R)/Z(F ) is a direct summand <strong>of</strong> T n . Define<br />

Z(Q) = (0) and T Q to be the trivial subgroup <strong>of</strong> T n . If p ∈ Q, p belongs to relative<br />

interior <strong>of</strong> a unique face F <strong>of</strong> Q.<br />

Define an equivalence relation ∼ on the product T n × Q by<br />

(t, p) ∼ (s, q) if and only if p = q and s −1 t ∈ T F (1.2.10)<br />

where F is the unique face containing p in its relative interior. Let<br />

M(Q, ξ) := (T n × Q)/ ∼<br />

be the quotient space. The natural left action <strong>of</strong> T n on T n × Q is given by<br />

t × (s × p) ↦→ ts × p for all t, s ∈ T n and p ∈ Q. (1.2.11)<br />

This action induces a natural action <strong>of</strong> T n on M(Q, ξ). Then M(Q, ξ) is a T n -space.


Chapter 1: Quasi<strong>toric</strong> manifolds 10<br />

The projection onto the second factor <strong>of</strong> T n × Q descends to the quotient map<br />

p : M(Q, ξ) → Q defined by p([t, p]) = p. (1.2.12)<br />

So the orbit space <strong>of</strong> this T n action on M(Q, ξ) is the polytope Q. We show that<br />

the space M(Q, ξ) has the structure <strong>of</strong> a quasi<strong>toric</strong> manifold. The explanations are<br />

discussed in the following paragraph.<br />

Consider an open neighborhood U v <strong>of</strong> the vertices v <strong>of</strong> Q where U v is defined in the<br />

paragraph before the statement <strong>of</strong> the theorem 1.2.9. Let<br />

M v (Q, ξ) := p −1 (U v ) = (T n × U v )/ ∼ . (1.2.13)<br />

Let the facets F j1 , . . . , F jn meet at the vertex v. So the facets <strong>of</strong> U v are F j1 , . . . , F jn .<br />

Suppose φ : U v → R n ≥0 is a diffeomorphism such that φ sends the facets F j 1<br />

, . . . , F jn to<br />

the facets<br />

{x 1 = 0} ∩ R n ≥0, . . . , {x n = 0} ∩ R n ≥0 <strong>of</strong> R n ≥0<br />

respectively. Where x j = 0 is the j-th coordinate hyperplane in R n for j = 1, . . . , n.<br />

Let δ v be the automorphism <strong>of</strong> T n corresponding to the automorphism <strong>of</strong> Z n obtained<br />

by sending the basis vectors ξ j1 , . . . , ξ jn to the standard basis vectors e 1 , . . . , e n<br />

<strong>of</strong> Z n respectively. Since the quotient maps p and p c <strong>of</strong> equation 1.2.8 and 1.2.12 respectively<br />

are continuous surjections and φ is a diffeomorphism, the following commutative<br />

diagram ensure us that the lower horizontal map φ v is a homeomorphism.<br />

(T n × U v )<br />

⏐<br />

p↓<br />

δ v ×φ<br />

−−−−→ (T n × R n 0<br />

⏐ )<br />

p ⏐↓<br />

c<br />

(1.2.14)<br />

M v (Q, ξ)<br />

φ v<br />

−−−−→<br />

C<br />

n<br />

Again the commutativity <strong>of</strong> the following diagram show that the space M v (Q, ξ) is<br />

T n -stable and T n action on M v (Q, ξ) satisfy the relation φ v (t · z) = δ v (t) · φ v (z).<br />

T n × (T n × U v )<br />

⏐<br />

Id×p↓<br />

T n × M v (Q, ξ)<br />

⏐<br />

a ⏐↓<br />

v<br />

Id×δ v ×φ<br />

−−−−−−→ T n × (T n × R n 0<br />

⏐ )<br />

Id×p ⏐↓<br />

c<br />

Id×φ v<br />

−−−−→<br />

T n × C n<br />

a s<br />

⏐ ⏐↓<br />

(1.2.15)<br />

M v (Q, ξ)<br />

φ v<br />

−−−−→<br />

C<br />

n<br />

Where a v is the restriction <strong>of</strong> the T n action on M(Q, ξ) to the subset M v (Q, ξ) and a s<br />

is the standard T n action on C n . Since {U v } v∈V (Q) is an open covering <strong>of</strong> Q, the T n


11 1.2 Definition and examples<br />

stable subsets {M v (Q, ξ)} is an open covering <strong>of</strong> M(Q, ξ). From the diagram 1.2.15 we<br />

get that every l-dimensional orbit in M v (Q, ξ) maps to a point in the relative interior<br />

<strong>of</strong> a l-dimensional face <strong>of</strong> U v . Hence (Q, ξ) is a 2n-dimensional quasi<strong>toric</strong> manifold.<br />

Definition 1.2.10. The manifold M(Q, ξ) is called the quasi<strong>toric</strong> manifold derived from<br />

the characteristic model (Q, ξ).<br />

Remark 1.2.11. The signs <strong>of</strong> the characteristic vectors {ξ j } in the characteristic model<br />

do not affect the groups T F . Hence they do not change the homeomorphism class <strong>of</strong><br />

M(Q, ξ). However the signs <strong>of</strong> the characteristic vectors ξ j affect the action <strong>of</strong> T n . If<br />

the polytope <strong>of</strong> a characteristic model is replaced by a diffeomorphic polytope, we derive<br />

the same quasi<strong>toric</strong> manifold modulo an equivariant homeomorphism.<br />

In the following examples we use two notions, namely orientation and connected<br />

sum <strong>of</strong> quasi<strong>toric</strong> manifolds which we discuss in the sections 1.6 and 1.7.<br />

Example 1.2.12. Let Q be a triangle △ 2 in R 2 . The possible characteristic maps<br />

are indicated by the following Figure 1.1. The quasi<strong>toric</strong> manifold corresponding to the<br />

(0, 1)<br />

(1,1)<br />

(0, 1)<br />

(1, −1)<br />

(1, 0)<br />

(1, 0)<br />

Figure 1.1: The characteristic models corresponding to a triangle.<br />

first characteristic model is CP 2 with the usual T 2 action. The orientation on CP 2 is<br />

the standard orientation. The second correspond to the same T 2 action with the reverse<br />

orientation on CP 2 , we denote it by CP 2 . The similar considerations can apply whenever<br />

Q is an n-dimensional simplex △ n . So the quasi<strong>toric</strong> manifold over △ n is either CP n<br />

or CP n .<br />

Example 1.2.13. Suppose Q is combinatorially a square in R 2 . In this case there are<br />

many possible characteristic maps. <strong>Some</strong> examples are given by the Figure 1.2.<br />

The first characteristic pairs may construct an infinite family <strong>of</strong> 4-dimensional quasi<strong>toric</strong><br />

manifolds, denote them by Mk 4 for each k ∈ Z. The manifolds {M k 4 : k ∈ Z} are<br />

equivariantly distinct. Let L(k) be the complex line bundle over CP 1 with the first Chern<br />

class k. The complex manifold CP(L(k) ⊕ C) is the Hirzebruch surface for the integer<br />

k, where CP(·) denotes the projectivisation <strong>of</strong> a complex bundle. So each Hirzebruch<br />

surface is the total space <strong>of</strong> the bundle CP(L(k) ⊕ C) → CP 1 with fiber CP 1 . In [Oda88]<br />

the author shows that with the natural action <strong>of</strong> T 2 on P(L(k) ⊕ C) it is equivariantly


Chapter 1: Quasi<strong>toric</strong> manifolds 12<br />

(1, k)<br />

(1, −2)<br />

(0, 1)<br />

(0, 1)<br />

(0, 1)<br />

(−1, 1)<br />

(1, 0)<br />

(1, 0)<br />

Figure 1.2: <strong>Some</strong> characteristic models corresponding to a square.<br />

homeomorphic to Mk<br />

4 for each k. That is, with respect to the T2 -action, Hirzebruch<br />

surfaces are quasi<strong>toric</strong> manifolds where the orbit space is a combinatorial square and<br />

the corresponding characteristic map is described in Figure 1.2.<br />

Note that the second combinatorial model gives the quasi<strong>toric</strong> manifold CP 2 # CP 2 ,<br />

the equivariant connected sum <strong>of</strong> CP 2 .<br />

Remark 1.2.14. Orlik and Raymond ( [OR70], p. 553) showed that any 4-dimensional<br />

quasi<strong>toric</strong> manifold M 4 over 2-dimensional simple polytope is an equivariant connected<br />

sum <strong>of</strong> some copies <strong>of</strong> CP 2 , CP 2 and M 4 k<br />

for some k ∈ Z.<br />

This classification result is used in Chapter 5. In Chapter 2 we show that there do<br />

not exist any combinatorial model corresponding to some simple polytope.<br />

Lemma 1.2.15 ( [Dav78]). Suppose p : M 2n → Q is a smooth quasi<strong>toric</strong> manifold.<br />

Then there exists a continuous section s : Q → M 2n .<br />

Pro<strong>of</strong>. We give an outline <strong>of</strong> the pro<strong>of</strong>. Suppose E is a smooth vector bundle <strong>of</strong> rank<br />

k over a manifold M ′ . Let E 0 denote the complement <strong>of</strong> the zero section. The positive<br />

real numbers R >0 act on E and E 0 by fiberwise scalar multiplication. Consider the R >0<br />

action on the product E 0 ×[0, ∞) defined by r(x, u) = (xr −1 , ru). So the quotient space<br />

C + E := E 0 × R>0 [0, ∞)<br />

is a smooth bundle over M ′ with fibers homeomorphic to S k−1 × [0, ∞). Denote the<br />

image <strong>of</strong> (x, u) in C + E by [x, u]. The boundary <strong>of</strong> C + E is the sphere bundle<br />

E 0 × R>0 {0} ∼ = E 0 /R >0 .<br />

There is a canonical projection map pr : C + E → E defined by pr[x, u] = ux which<br />

is a diffeomorphism away from the zero section. Replacing E by C + E is called a real<br />

blow-up. The projection pr may be called a real blow-down.<br />

We can stratify M 2n = ⊔ M(F ◦ , ξ) where F varies over all faces <strong>of</strong> Q.<br />

Then a<br />

neighborhood <strong>of</strong> M(F ◦ , ξ) in M 2n is diffeomorphic to a R 2k -bundle over M(F ◦ , ξ), where


13 1.2 Definition and examples<br />

k is the codimension <strong>of</strong> F . We start with the minimal strata (vertices) progressively<br />

blow-up M 2n along the strata <strong>of</strong> increasing dimension to finally get a smooth manifold<br />

with boundary ̂M.<br />

The precise description may be found in Chapter 4 <strong>of</strong> [Dav78].<br />

One can show that ̂M is homeomorphic to T n × Q. The canonical blow-downs pr are<br />

combined to give a continuous map µ : ̂M → M 2n . Now choose a continuous section<br />

ŝ : Q → ̂M and compose with µ to get a continuous section s : Q → M 2n .<br />

Remark 1.2.16. We will show in Section 1.9 that every quasi<strong>toric</strong> manifold has a<br />

smooth structure. However the above lemma remains valid even if we drop the smoothness<br />

assumption.<br />

Corollary 1.2.17 (Lemma 1.8, [DJ91]). Suppose M 2n is a quasi<strong>toric</strong> manifold with<br />

characteristic model (Q, ξ).<br />

Suppose M(Q, ξ) is a quasi<strong>toric</strong> manifold derived from<br />

(Q, ξ). Then M(Q, ξ) is equivariantly homeomorphic to M 2n .<br />

Pro<strong>of</strong>. Let s : Q → M 2n be a continuous section. Consider the composition map<br />

g : T n × Q Id×s −−−→ T n × M 2n ·<br />

−→ M 2n , (1.2.16)<br />

where · is the T n action on M 2n . From the locally standardness <strong>of</strong> T n action it is clear<br />

that the fiber g −1 (x) <strong>of</strong> each x ∈ M 2n is the isotropy group <strong>of</strong> x. Hence the map g<br />

factors through the maps p and f in following commutative diagram.<br />

p<br />

T n × Q −−−−→ (T n × Q)/ ∼<br />

⏐<br />

⏐<br />

g↓<br />

f↓<br />

(1.2.17)<br />

M 2n M 2n<br />

Clearly the map f is a bijection. The diagram gives that f is an equivariant map. Again<br />

the locally standard property <strong>of</strong> T n action ensure that f is a homeomorphism.<br />

Definition 1.2.18. Let M1 2n and M2 2n be quasi<strong>toric</strong> manifolds whose associated base<br />

polytope is Q. Let δ be an automorphism <strong>of</strong> T n . A map f : M1 2n → M2 2n is called a<br />

δ-equivariant homeomorphism if f is a homeomorphism and satisfies f(t·x) = δ(t)·f(x)<br />

for all (x, t) ∈ M 2n<br />

1 × T n .<br />

Two δ-equivariant homeomorphisms f : M1 2n → M2 2n and g : M1 2n → M2 2n are<br />

said to be equivalent if there exist equivariant homeomorphisms h j : Mj<br />

2n → Mj<br />

2n , for<br />

j = 1, 2, such that the following diagram is commutative.<br />

M 2n<br />

1<br />

h 1<br />

⏐ ⏐↓<br />

M 2n<br />

1<br />

f<br />

−−−−→ M2<br />

2n<br />

⏐ ⏐↓<br />

h 2<br />

g<br />

−−−−→ M 2n<br />

2<br />

(1.2.18)


Chapter 1: Quasi<strong>toric</strong> manifolds 14<br />

For the automorphism δ we also define the δ-translation <strong>of</strong> a characteristic model<br />

(Q, ξ) to be the pair (Q, δ ′ ◦ ξ), where δ ′ is an automorphism <strong>of</strong> Z n induced by the<br />

automorphism δ. We can determine quasi<strong>toric</strong> manifolds over a fixed polytope up to<br />

δ-equivariant homeomorphism class using the following lemma.<br />

Lemma 1.2.19 (Proposition 2.6 <strong>of</strong> [BR01]). For any automorphism δ <strong>of</strong> T n , the assignment<br />

<strong>of</strong> characteristic models defines a bijection between equivalence classes <strong>of</strong> δ-<br />

equivariant homeomorphisms <strong>of</strong> quasi<strong>toric</strong> manifolds over Q and δ-translations <strong>of</strong> characteristic<br />

model (Q, ξ).<br />

Pro<strong>of</strong>. First we show that the inverse assignment is given by consisting the quasi<strong>toric</strong><br />

manifolds derived from characteristic models (Q, ξ) and (Q, δ ′ ◦ξ). To each δ-translation<br />

(Q, ξ) → (Q, δ ′ ◦ ξ) we associate the δ-equivariant diffeomorphism<br />

(δ × Id) ∼ : (T n × Q)/ ∼→ (T n × Q)/ ∼ δ ,<br />

where<br />

(t, q) ∼ δ (u, q) if and only if tu −1 ∈ δ ′ (ξ)(T F ). (1.2.19)<br />

Here F is the unique face containing p in its relative interior.<br />

It is clear from the<br />

definitions 1.2.10 and 1.2.19 that (δ×Id) ∼ descends to the original δ-translation (Q, ξ) →<br />

(Q, δ ′ (ξ)) <strong>of</strong> characteristic models.<br />

Conversely, let f : M1 2n → M2 2n be a δ-equivariant homeomorphism <strong>of</strong> quasi<strong>toric</strong><br />

manifolds over Q. This diffeomorphism descends to a δ-translation <strong>of</strong> characteristic<br />

models. Let (δ × Id) ∼ be the δ-equivariant homeomorphism derived from the corresponding<br />

δ-translation <strong>of</strong> characteristic models. The preferred section s 1 : Q → M 2n<br />

1<br />

automatically extends to an equivariant homeomorphism S 1 : (T n × Q)/ ∼→ M 2n<br />

1 . Let<br />

s 2 = f ◦ s 1 . Since f is a δ-equivariant homeomorphism, s 2 : Q → M 2n<br />

2 is a section. The<br />

section s 2 extends to an eqivariant homeomorphism S 2 : (T n × Q)/ ∼ δ → M 2n<br />

2 . Thus<br />

f ◦ S 1 = S 2 ◦ (δ × Id) ∼ . Hence f and (δ × Id) ∼ are equivalent, as required.<br />

1.3 Invariant closed submanifolds<br />

Corresponding to the faces <strong>of</strong> the polytope Q there are certain T n -invariant submanifolds<br />

<strong>of</strong> M 2n . If F is a face <strong>of</strong> Q <strong>of</strong> codimension-k, then define M(F, ξ) := p −1 (F ). Define<br />

Z ⊥ (F ) = Z n /Z(F ). Let ϱ F : Z n → Z ⊥ (F ) be the projection homomorphism. Let<br />

J(F ) ⊂ F(Q) be the index set <strong>of</strong> facets <strong>of</strong> Q, other than F in case k = 1, that intersect<br />

F . Observe that J(F ) indexes the set <strong>of</strong> facets <strong>of</strong> the (n − k)-dimensional polytope F .<br />

Let {H j : j ∈ J(F )} be the set <strong>of</strong> all facets <strong>of</strong> F . So H j = F ∩ F j for some facets F j <strong>of</strong><br />

Q, j ∈ J(F ). If ξ is the characteristic function <strong>of</strong> M 2n over Q, the assignment<br />

ξ F (H j ) := ϱ F ◦ ξ(F j ), j ∈ J(F ), (1.3.1)


15 1.4 Face vectors and face ring <strong>of</strong> polytopes<br />

defines the characteristic function ξ F <strong>of</strong> the quasi<strong>toric</strong> manifold M(F, ξ F ). With subspace<br />

<strong>topology</strong> on M(F, ξ), M(F, ξ) is equivariantly homeomorphic to M(F, ξ F ). Hence<br />

M(F, ξ) is a quasi<strong>toric</strong> manifold <strong>of</strong> dimension 2n − 2k. One can show that if M ′ is an<br />

invariant closed submanifold <strong>of</strong> M 2n then M ′ = M(F, ξ) for some face F <strong>of</strong> Q.<br />

Definition 1.3.1. When F is a facet <strong>of</strong> Q, the space M(F, ξ F ) is called a characteristic<br />

submanifold <strong>of</strong> M 2n corresponding to F .<br />

1.4 Face vectors and face ring <strong>of</strong> polytopes<br />

The face vector or f-vector is an important concept in the combina<strong>toric</strong>s <strong>of</strong> polytopes.<br />

Let L be a simplicial n-polytope and f j be the number <strong>of</strong> j-dimensional faces <strong>of</strong> L. The<br />

integer vector f(L) = (f 0 , . . . , f n−1 ) is called the f-vector <strong>of</strong> the simplicial polytope L.<br />

Let h i be the coefficients <strong>of</strong> t n−i in the polynomial<br />

(t − 1) n + Σ n−1<br />

0 f i (t − 1) n−1−i . (1.4.1)<br />

The vector h(L) = (h 0 , . . . , h n ) is called h-vector <strong>of</strong> L. Obviously h 0 = 1, and Σ n 0 h i =<br />

f n−1 . The f-vector and h-vector <strong>of</strong> a simple n-polytope Q is the f-vector and h-vector<br />

<strong>of</strong> its dual simplicial polytope Q ∗ respectively, that is<br />

f(Q) = f(Q ∗ ) and h(Q) = h(Q ∗ ).<br />

Hence for a simple n-polytope Q,<br />

f(Q) = (f 0 , . . . , f n−1 ), (1.4.2)<br />

where f j is the number <strong>of</strong> codimension (j + 1) faces <strong>of</strong> Q. Then h n = 1 and Σ n h 1 i is the<br />

number <strong>of</strong> vertices <strong>of</strong> Q. The face vectors are a combinatorial invariant <strong>of</strong> polytopes,<br />

that is it depends only on the face poset <strong>of</strong> the polytope.<br />

Let w 1 , . . . , w m be the vertices <strong>of</strong> a simplicial complex L. Let R be a commutative<br />

ring with unity. Consider the polynomial ring R[w 1 , . . . , w m ] where the w i ’s are indeterminates.<br />

Let I L be the homogeneous ideal generated by all monomials <strong>of</strong> the form<br />

w i1 . . . w il such that the set <strong>of</strong> vertices {w i1 , . . . , w il } does not span a simplex in L.<br />

Definition 1.4.1. The face ring or Stanley-Reisner ring <strong>of</strong> L with coefficients in R is<br />

the quotient ring R[w 1 , . . . , w m ]/I L , denoted by SR(L, R).<br />

We define the face ring <strong>of</strong> a polytope Q. Let L be the simplicial complex dual to Q.<br />

Definition 1.4.2. The face ring or Stanley-Reisner ring SR(Q, R) <strong>of</strong> Q over R is the<br />

ring SR(L, R). The face ring is graded by declaring the degree <strong>of</strong> each v i .


Chapter 1: Quasi<strong>toric</strong> manifolds 16<br />

1.5 Homology groups <strong>of</strong> quasi<strong>toric</strong> manifolds<br />

Following [DJ91] we compute the homology groups <strong>of</strong> quasi<strong>toric</strong> manifolds with coefficients<br />

in Z. The combinatorial type <strong>of</strong> the orbit space (essentially a simple convex<br />

polytope) <strong>of</strong> quasi<strong>toric</strong> manifold makes the computation <strong>of</strong> the singular homology groups<br />

easier. First we decompose the polytope into a disjoint union <strong>of</strong> relative open subsets<br />

such that this collections correspond to the set <strong>of</strong> vertices <strong>of</strong> polytope bijectively. Corresponding<br />

to each <strong>of</strong> these relative open subsets we get an even dimensional cell. These<br />

cells give a perfect CW -complex structure on the manifold. We decompose the polytope<br />

using the notion <strong>of</strong> index <strong>of</strong> a vertex that will describe the degree <strong>of</strong> each cell. Though<br />

there is no canonical choice <strong>of</strong> index, the homology group <strong>of</strong> any degree can be computed<br />

up to isomorphism.<br />

Let M 2n −→ p Q be a quasi<strong>toric</strong> manifold <strong>of</strong> dimension 2n. Suppose Q is a simple<br />

polytope in R n . Choose a vector z ∈ R n which is not perpendicular to any line joining<br />

two vertices <strong>of</strong> Q. Let ζ : R n → R be the linear functional defined by<br />

ζ(x) := 〈x, z〉 for all x ∈ R n . (1.5.1)<br />

Where 〈·, ·〉 is the standard inner product in R n . Choice <strong>of</strong> z distinguishes the vertices<br />

linearly according to ascending value <strong>of</strong> ζ. Since ζ is linear, we make the 1-skeleton <strong>of</strong><br />

Q into a directed graph by orienting each edge such that ζ increases along it. For each<br />

vertex v <strong>of</strong> Q the number <strong>of</strong> incident edges that point towards v is called its index,<br />

denoted by f(v).<br />

Let F v denote the smallest face <strong>of</strong> Q which contains the inward pointing edges<br />

incident to v. Since Q is simple polytope and ζ is a linear functional distinguishing<br />

the vertices <strong>of</strong> Q, such a face F v exist uniquely corresponding to each vertex v. Then<br />

dim F v = f(v) and if F ′ is a face <strong>of</strong> Q with top vertex v then F ′ is a face <strong>of</strong> F v . By top<br />

vertex we mean that f(v) > f(u) for all vertices u other than v. Let ̂F v be the relative<br />

open subset F v obtain by deleting all faces <strong>of</strong> F v not containing v. So Q = ⊔ v ̂Fv , where<br />

v run over the vertices <strong>of</strong> Q. The space ̂F v is diffeomorphic to the positive octant R f(v)<br />

≥0 .<br />

<strong>Some</strong> combinatorial arguments show that the number <strong>of</strong> vertices v with f(v) = j<br />

is h j for j = 0, . . . , n. For each vertex v put e v = p −1 ( ̂F v ). By the locally standard<br />

property at the fixed point corresponding to the vertex v we can show<br />

e v<br />

∼ = (TFv × T f(v) × ̂F v )/ ∼<br />

∼ = (T f(v) × R f(v)<br />

≥0 )/ ∼ 0<br />

∼= D 2f(v) , (1.5.2)<br />

where D 2f(v) is an open disk in R 2f(v) . Hence e v is a 2f(v)-dimensional cell in M 2n .<br />

So M 2n can be given the structure <strong>of</strong> a CW -complex structure as follows. Define


17 1.6 Orientation <strong>of</strong> quasi<strong>toric</strong> manifolds<br />

the 2k-skeleton<br />

M 2k :=<br />

∪<br />

M(F v , ξ) for 0 ≤ k ≤ n. (1.5.3)<br />

f(v)=k<br />

Define M 2k+1 = M 2k for 0 ≤ k ≤ n − 1, and M 2n = M 2n . The 2k-skeleton M 2k can<br />

be obtained from M 2k−1 by attaching those cells e v for which f(v) = k. The attaching<br />

maps are to be described.<br />

Define Z ⊥ (F ) := Z n /Z(F ). Then<br />

T(F ) ⊥ := Z ⊥ (F ) ⊗ R/Z ⊥ (F ) ∼ = T n /T(F ). (1.5.4)<br />

Let ∼ be the equivalence relation such that M(F v , ξ) = F v × T(F v ) ⊥ / ∼. The disk<br />

D 2f(v) ) can be identified with T(F v ) ⊥ × F v /≈ where<br />

(t, p) ≈ (s, q) if p = q and s −1 t ∈ T(F ′ )/T(F v ) (1.5.5)<br />

where F ′ is the minimal face <strong>of</strong> F v containing p whose top vertex is v. The attaching<br />

map S 2f(v)−1 → M 2f(v)−1 is the natural quotient map from<br />

(F v − ̂F v ) × T(F v ) ⊥ /≈ → (F v − ̂F v ) × T(F v ) ⊥ /∼. (1.5.6)<br />

Hence M 2n is a CW -complex with no odd dimensional cells and with f −1 (k) = h k<br />

number <strong>of</strong> 2k dimensional cells. Hence by cellular homology theory<br />

{ ⊕<br />

h k<br />

Z if 0 ≤ p ≤ n and p is even,<br />

H p (M 2n ; Z) =<br />

0 otherwise.<br />

(1.5.7)<br />

1.6 Orientation <strong>of</strong> quasi<strong>toric</strong> manifolds<br />

Let M 2n −→ p Q be a 2n-dimensional quasi<strong>toric</strong> manifold. From the previous section 1.5<br />

we get that the top homology group <strong>of</strong> M 2n is Z. So the manifold M 2n is orientable.<br />

In fact a choice <strong>of</strong> orientation on T n and Q gives an orientation on M 2n . We fix the<br />

standard orientation on T n . Hence an orientation on Q ⊂ R n determines an orientation<br />

on M 2n .<br />

Suppose the manifold M 2n has a smooth structure. Clearly the isotropy group <strong>of</strong> a<br />

characteristic submanifold is a circle subgroup <strong>of</strong> T n . So there is a natural S 1 action on<br />

the normal bundle <strong>of</strong> that characteristic submanifold. Thus the normal bundle has a<br />

complex structure and consequently an orientation. Whenever the sign <strong>of</strong> the characteristic<br />

vector <strong>of</strong> a facet is reverse, we get the opposite orientation on the normal bundle.<br />

An orientation on the normal bundle together with an orientation on M 2n induces an<br />

orientation on the characteristic submanifold. A structure called omniorientation provides<br />

a combinatorial description for a stable complex structure on quasi<strong>toric</strong> manifold.


Chapter 1: Quasi<strong>toric</strong> manifolds 18<br />

Definition 1.6.1. An omniorientation on a quasi<strong>toric</strong> manifold M 2n is a choice <strong>of</strong><br />

orientation for M 2n as well as for each characteristic submanifold <strong>of</strong> M 2n .<br />

If the polytope Q has m facets, there are 2 m+1 possible omniorientations for M 2n .<br />

From the above discussion we get the following remark.<br />

Remark 1.6.2. A choice <strong>of</strong> omniorientation is equivalent to a choice <strong>of</strong> orientation for<br />

Q and a choice <strong>of</strong> sign for each characteristic vector.<br />

We will apply the same terminology in the case <strong>of</strong> quasi<strong>toric</strong> orbifolds in Chapter 3.<br />

Example 1.6.3. Consider a triangle whose edges in counterclockwise order have characteristic<br />

vectors (1, 0), (0, 1) and (1, 1). The corresponding manifold is CP 2 .<br />

Now consider the orientation reversing map on R 2 that maps (1, 0) ↦→ (0, 1) and<br />

(0, 1) ↦→ (1, −1). Then (1, 1) ↦→ (1, 0). Rotating the triangle observe that CP 2 has characteristic<br />

vectors (1, 0), (0, 1), (1, −1). There are other choices <strong>of</strong> characteristic vectors<br />

for CP 2 such as those given by applying an orientation reversing automorphism <strong>of</strong> R 2<br />

to the standard one.<br />

(0, 1)<br />

(1,1)<br />

(0, 1)<br />

(1, −1)<br />

(1, 0)<br />

(1, 0)<br />

Figure 1.3: Omniorientation <strong>of</strong> CP 2 and CP 2 .<br />

1.7 Equivariant connected sums<br />

Following [BR01] we define connected sum <strong>of</strong> polytopes to perform the equivariant<br />

connected sums <strong>of</strong> quasi<strong>toric</strong> manifolds. Let Q 1 , Q 2 be two n-dimensional polytopes in<br />

R n . Consider the polyhedral template<br />

Γ = {(x 1 , x 2 , . . . , x n ) ∈ R n : 0 ≤ x 2 , . . . , x n and x 2 + . . . + x n ≤ 1}.<br />

Let<br />

and<br />

G j = {(x 1 , x 2 , . . . , x n ) ∈ Γ : x j = 0} for 2 ≤ j ≤ n,<br />

G 1 = {(x 1 , x 2 , . . . , x n ) ∈ Γ : x 2 + . . . + x n = 1}.


19 1.7 Equivariant connected sums<br />

So G 1 , G 2 , . . . , G n are the facets <strong>of</strong> Γ.<br />

The sets Γ, G 1 , . . . , G n are divided into two<br />

halves, namely positive and negative halves, determined by the sign <strong>of</strong> the coordinate<br />

x 1 . Let v and w are two distinct vertices <strong>of</strong> Q 1 and Q 2 respectively. Considering the<br />

local orientation at v, w ∈ R n , we order the facets <strong>of</strong> Q 1 meeting at v as F j ′ and the<br />

facets <strong>of</strong> Q 2 meeting at w as F j<br />

′′ for 1 ≤ j ≤ n. Let C v and C w are union <strong>of</strong> facets not<br />

containing the vertices v and w <strong>of</strong> Q 1 and Q 2 respectively. Suppose φ Q1 is a projective<br />

transformation which maps v to x 1 = +∞ and embeds Q 1 into Γ such that following<br />

two conditions are satisfied;<br />

1. The hyperplane defining F ′ j is identified with the hyperplane defining G j, for each<br />

1 ≤ j ≤ n.<br />

2. The images <strong>of</strong> the hyperplanes defining C v under the map φ Q1 belong to the<br />

negative half <strong>of</strong> Γ.<br />

Suppose v 1 , v 2 , . . . , v n are vertices <strong>of</strong> C v such that there is an edge joining v and v j for<br />

for each 1 ≤ j ≤ n. We may define the map φ Q1<br />

in the following way. Let<br />

△ n−1 = {(−1, x 2 , . . . , x n ) ∈ R n+1 : 0 ≤ x 2 , . . . , x n and x 2 + . . . + x n ≤ 1}.<br />

Let φ ′ Q 1<br />

be an affine equivalence mapping which sends v and v 1 , v 2 , . . . , v n to (1, 0, . . . , 0)<br />

and vertices <strong>of</strong> △ n−1 respectively. Consider the map A defined by A(x) = x/(1 − x 1 ).<br />

Then the composition φ Q1 = A ◦ φ ′ Q 1<br />

<strong>of</strong> maps φ ′ Q 1<br />

and A is the required projective<br />

transformation. Similarly we can choose φ Q2 such that it sends w to x 1 = −∞ and<br />

identifies the hyperplanes defining F j<br />

′′ and G j in such a way that the images <strong>of</strong> the<br />

hyperplanes defining C w belong to the positive half <strong>of</strong> Γ. We define the connected sum<br />

Q 1 # v;w Q 2 <strong>of</strong> Q 1 at v and Q 2 at w to be the n-dimensional simple polytope determined<br />

by all the hyperplanes in φ Q1 (C v ) and φ Q2 (C w ) together with G j for 1 ≤ j ≤ n. The<br />

connected sum is defined only up to combinatorial equivalence. Different choices for the<br />

vertices v and w or the orderings for F j ′ and F<br />

′′<br />

j may affect the combinatorial type <strong>of</strong><br />

resulting polytope. When the choices are clear, we use the abbreviation Q 1 #Q 2 .<br />

Let M(Q 1 , ξ) and M(Q 2 , µ) be two quasi<strong>toric</strong> manifolds over Q 1 and Q 2 with fixed<br />

points x and y corresponding to the vertices v ∈ Q 1 and w ∈ Q 2 respectively. We state<br />

the following Lemma and Theorem <strong>of</strong> Buchstaber and Ray without the pro<strong>of</strong>.<br />

Lemma 1.7.1 (Lemma 6.7, [BR01]). Up to δ-translation, we may assume that ξ identifies<br />

T Fj with the j-th coordinate subtorus T j , for each 1 ≤ j ≤ n.<br />

Applying the previous lemma to both ξ and µ we can define a characteristic function<br />

ξ µ <strong>of</strong> Q 1 #Q 2


Chapter 1: Quasi<strong>toric</strong> manifolds 20<br />

⎧<br />

⎪⎨ ξ(F ) if F is a facet in C v<br />

ξ µ = T j for F = G j and 1 ≤ j ≤ n<br />

⎪⎩<br />

µ(F ) if F is a facet in C w .<br />

(1.7.1)<br />

Theorem 1.7.2 (Theorem 6.9, [BR01]). The quasi<strong>toric</strong> manifold M(Q 1 #Q 2 , ξ µ ) is<br />

equivariantly homeomorphic to the connected sum <strong>of</strong> M(Q 1 , ξ) at x and M(Q 2 , µ) at y.<br />

Example 1.7.3. Let the triangles Q 1 = ABC and Q 2 = DEF be the orbit space <strong>of</strong><br />

quasi<strong>toric</strong> manifolds CP 2 and CP 2 respectively. Suppose the characteristic vectors along<br />

AB, BC, CA, DE, EF and F D are (1, 0), (0, 1), (1, 1), (0, 1), (1, −1) and (1, 0) respectively.<br />

To perform an equivariant connected sum we fix the vertices B and D <strong>of</strong> Q 1 and<br />

Q 2 respectively. Cut <strong>of</strong>f the corners at these vertices by open halves H 1 ′ and H′ 2 (see the<br />

Figure 1.4). Straighten out the remaining portions <strong>of</strong> the lines AB ′ , CB ′′ to make them<br />

perpendicular to AC. We do the same for the lines ED ′ , F D ′′ . Now identify B ′ B ′′<br />

with D ′ D ′′ such that A, B ′ = D ′ , E lie on a line (say AE) and C, B ′′ = D ′′ , F lie on<br />

the line (say CF ). So we get a quadrilateral Q = AEF C. Let us retain (1, 0), (1, 1)<br />

and (0, 1) as characteristic vectors for AE, AC and CF respectively. That means the<br />

characteristic vector (0, 1) <strong>of</strong> DE is mapped to (1, 0) and the characteristic vector (1, 0)<br />

<strong>of</strong> F D is mapped to (0, 1). These determine an orientation reversing isomorphism <strong>of</strong><br />

R 2 . Using this we get that the characteristic vector (1, −1) <strong>of</strong> EF should transform to<br />

the characteristic vector (−1, 1) for EF in Q. Thus the quadrilateral AEF C with characteristic<br />

vectors (1, 0), (1, −1), (0, 1), (1, 1) as in the Figure 1.4 represents CP 2 #CP 2 .<br />

(1, 1)<br />

C<br />

A<br />

H ′ 1<br />

H ′ 2<br />

F<br />

C (0, 1)<br />

F<br />

B ′′<br />

(1, −1)<br />

(0, 1) (1, 0)<br />

D ′′<br />

# = (1, 1) (−1, 1)<br />

B D<br />

B ′ D ′<br />

(1, 0)<br />

(0, 1)<br />

E<br />

A (1, 0) E<br />

P 1 P2<br />

P<br />

Figure 1.4: Equivariant connected sum for CP 2 and CP 2 .<br />

It is well known that CP 2 #CP 2 does not have any almost complex structure. This<br />

example then shows that not every quasi<strong>toric</strong> manifold is a <strong>toric</strong> variety. Note that the<br />

quasi<strong>toric</strong> manifold M1 4 , defined in example 1.2.13, is the equivariant connected sum<br />

CP 2 # CP 2 .<br />

Considering the omniorientation we can construct the omnioriented connected sum<br />

<strong>of</strong> quasi<strong>toric</strong> manifolds similarly. The Figure 1.5 describe how to perform an omniori-


21 1.8 Cohomology ring<br />

(0, 1)<br />

(0, 1)<br />

(0, 1)<br />

(−1, −1) #<br />

(−1, −1) = (−1, −1)<br />

(−1, −1)<br />

(1, 0)<br />

(1, 0)<br />

(1, 0)<br />

Figure 1.5: Omnioriented connected sum for CP 2 and CP 2 .<br />

ented connected sum <strong>of</strong> CP 2 and CP 2 .<br />

1.8 Cohomology ring<br />

We compute the cohomology ring <strong>of</strong> quasi<strong>toric</strong> manifold M 2n over simple n-polytope Q<br />

following [DJ91]. The main idea is to make use <strong>of</strong> the equivariant cohomology <strong>of</strong> M 2n<br />

with respect to the T n action. Consider the Borel space<br />

BQ := ET n × T n M 2n<br />

<strong>of</strong> the T n action on M 2n . Where ET n is a contractible space and T n action on ET n<br />

is free. Let 2 be the degree <strong>of</strong> each w i ∈ SR(Q, Z). All homology and cohomology<br />

modules in this section will have coefficients in Z. We show that the cohomology ring<br />

H ∗ (BQ, Z) is isomorphic to the face ring SR(Q, Z).<br />

We study the Leray spectral sequence <strong>of</strong> the fiber bundle M 2n ↩→ BQ −→ BT n .<br />

The spectral sequence degenerates at E 2 . Knowledge <strong>of</strong> the cohomology ring <strong>of</strong> BQ and<br />

BT n , together with the cohomology group <strong>of</strong> M 2n is sufficient to deduce the cohomology<br />

ring <strong>of</strong> M 2n . In our approach we use the localization principle <strong>of</strong> Atiyah-Bott [AB84].<br />

We show that M 2n is the union <strong>of</strong> 2n-dimensional disks centered around fixed points<br />

<strong>of</strong> T n action in the following. Let L be the simplicial complex associated to the boundary<br />

<strong>of</strong> the dual polytope <strong>of</strong> Q. Then there is a bijective correspondence between (n − 1)-<br />

dimensional faces <strong>of</strong> L and the vertices <strong>of</strong> Q. Also Q is the cone on the barycentric<br />

subdivision <strong>of</strong> L. So Q can be written as the union <strong>of</strong> cubes Q v where v varies over the<br />

vertices <strong>of</strong> Q. Recall the projection map p : M 2n → Q. Define<br />

M v := p −1 (Q v ).<br />

Then M 2n = ∪ v M v. The space BQ has a corresponding decomposition as follows. We<br />

regard the k-cube [0, 1] k as the orbit space <strong>of</strong> standard k-dimensional torus action on<br />

the 2k-disk<br />

D 2k = {(z 1 , · · · , z k ) ∈ C k : |z i | ≤ 1}. (1.8.1)


Chapter 1: Quasi<strong>toric</strong> manifolds 22<br />

Define<br />

BQ v = ET n × T n M v ≃ ET n × T n D 2n . (1.8.2)<br />

Then BQ = ∪ v BQ v. Let b p : BQ → BT n be the Borel map which is a fibration with<br />

fiber M 2n . So we have a homomorphism b p ∗ : H ∗ (BT n , Z) → H ∗ (BQ, Z) induced by<br />

b p .<br />

Theorem 1.8.1 (Theorem 4.8, [DJ91]). The homomorphism b p ∗ is a surjection and<br />

induces an isomorphism <strong>of</strong> graded rings H ∗ (BQ, Z) ∼ = SR(Q, Z).<br />

Pro<strong>of</strong>. Let F i1 , . . . , F in be the facets meeting at a vertex v <strong>of</strong> the polytope Q. Then<br />

BQ v = ET n × T n (D 2n ) is a D 2n fiber bundle over BT n . The associated complex vector<br />

bundle is γ v : ET n × T n C n → BT n . Regard BT n as the product <strong>of</strong> n copies <strong>of</strong> BU(1).<br />

Let p j be the projection from ∏ n<br />

j=1<br />

BU(1) to the j-th coordinate. Denote the universal<br />

complex line bundle over BU(1) by γ 1,∞ . Since the action <strong>of</strong> T n on C n is diagonal,<br />

γ v = ⊕ j p∗ j (γ 1,∞). That is p ∗ j (γ 1,∞) corresponds to j-th coordinate line in C n . So<br />

without confusion, we may set<br />

c 1 (p ∗ j(γ 1,∞ )) = w ij ∈ H 2 (BT n ; Z).<br />

Note that H ∗ (BT n , Z) = Z[w i1 , . . . , w in ]. Since D 2n is contractible, H ∗ (BQ v ; Z) =<br />

H ∗ (BT n ; Z) = Z[w i1 , · · · , w in ]. We compute H ∗ (BQ, Z) by gluing the spaces BQ v with<br />

the Mayer-Vietoris argument. We need the cohomology <strong>of</strong><br />

BQ Sv = ET n × T n ∂(D 2n ) = ET n × T n S 2n−1 .<br />

The fiber bundle b s : BQ Sv → BT n can be identified with the sphere bundle <strong>of</strong> the<br />

vector bundle γ v . By the Whitney product formula we get that c n (γ v ) = w i1 · · · w in .<br />

Hence e := w i1 · · · w in is the Euler class <strong>of</strong> the sphere bundle b s .<br />

Now consider the Gysin exact sequence for sphere bundles (see [MS74])<br />

· · · → H ∗ (BQ Sv ) → H ∗ (BT n )<br />

∪e<br />

−−−−→ H ∗+2n (BT n )<br />

b ∗ s<br />

−−−−→ H ∗+2n (BQ Sv ) → · · ·<br />

(1.8.3)<br />

Since the map ∪e is an injection in equation 1.8.3, by exactness <strong>of</strong> this sequence the<br />

map b ∗ s is a surjection and one get the following commutative diagram<br />

0 → H ∗ (BT n )<br />

⏐<br />

id↓<br />

∪e<br />

−−−−→ H ∗+2n (BT n )<br />

⏐<br />

id↓<br />

p ∗ s<br />

−−−−→ H ∗+2n (BQ Sv ) → 0<br />

(1.8.4)<br />

Z[w i1 , · · · , w in ]<br />

w i1 ...w i n<br />

−−−−−→ Z[w i1 , . . . , w in ]<br />

Hence from diagram (1.8.4) H ∗ (BQ Sv ) = Z[w i1 , · · · , w in ]/(w i1 . . . w in ). Now applying<br />

the Mayer-Vietoris argument for cohomology the theorem can be obtained.


23 1.8 Cohomology ring<br />

Let h N : N → N be the Hilbert function <strong>of</strong> the face ring SR(Q, Z). Stanley (Proposition<br />

3.2, [Sta75]) shows that<br />

⎧<br />

⎪⎨ Σ l−1<br />

j=0 f i l−1 C j if k = 2l<br />

h N (k) = 1 if k = 1<br />

⎪⎩<br />

0 if k is odd<br />

(1.8.5)<br />

Here f j ’s and h j ’s are f-vectors and h-vectors <strong>of</strong> Q respectively defined in Section 1.4.<br />

So the Poincaré series <strong>of</strong> SR(Q, Z) is Σ ∞ k=0 h N(k). Using the relations between f-vectors<br />

and h-vectors <strong>of</strong> Q see [BP02], one can show that there is an identity <strong>of</strong> formal power<br />

series<br />

(1 − t 2 ) n (Σ ∞ k=0 h N(k)) = Σ n 0 h j t 2j . (1.8.6)<br />

Theorem 1.8.2 (Corollary 4.13, [DJ91]). Let ι : M 2n → BQ be an inclusion <strong>of</strong> the<br />

fiber. The induced map ι ∗ : H ∗ (BQ, Z) → H ∗ (M 2n , Z) is a surjection.<br />

Pro<strong>of</strong>. Consider the Leray-Serre spectral sequence <strong>of</strong> the fibration p : BQ → BT n with<br />

fiber M 2n . The E 2 -terms <strong>of</strong> this sequence is E p,q<br />

2 = H p (BT n ; H q (M 2n )). Since BT n<br />

is simply connected we have E p,q<br />

2 = H p (BT n ) ⊗ H q (M 2n ). The Poincaré series <strong>of</strong> the<br />

E r -terms is by definition<br />

∑<br />

( ∑<br />

dimEr p,q )t k . (1.8.7)<br />

k<br />

p+q=k<br />

The Poincaré series <strong>of</strong> H ∗ (BT n ) is 1/(1 − t 2 ) n . The Poinearé series <strong>of</strong> H ∗ (M 2n ) is<br />

h(t) := h 0 + h 1 t 2 + . . . + h n t 2n .<br />

Hence, the Poinearé series <strong>of</strong> E 2 is h(t)/(1 − t 2 ) n . It turns out from Theorem 1.8.1 and<br />

equation 1.8.6 that the Poincaré series <strong>of</strong> E 2 equals to the Poincaré series <strong>of</strong> H ∗ (BQ) or<br />

E ∞ . However since E ∞ is an iterated subquotient <strong>of</strong> E 2 and they have the same Poincaré<br />

series, we get the following equality; E p,q<br />

2 = E∞ p,q . Hence it follows that H ∗ (BQ, Z) ∼ =<br />

H ∗ (BT n , Z) ⊗ H ∗ (M 2n , Z) as Z-modules. Thus the map ι ∗ : H ∗ (BQ) → H ∗ (M 2n ) is a<br />

surjection.<br />

Recall that F(Q) is the set <strong>of</strong> facets <strong>of</strong> the polytope Q. Let m be the cardinality <strong>of</strong><br />

F(Q). Consider the standard local model (R m ≥0 , ɛ) for Cm , where ɛ corresponds to the<br />

assignment <strong>of</strong> standard basis elements <strong>of</strong> Z m to the facets <strong>of</strong> R m ≥0 . Let p s : C m → R m ≥0<br />

be the projection map. Embed the polytope Q in R m ≥0 by the map d F : Q → R m where<br />

the i-th coordinate <strong>of</strong> d F (q) is the Euclidean distance d(q, F i ) <strong>of</strong> q from the hyperplane<br />

<strong>of</strong> the i-th facet F i ∈ F(Q) in R n . Define the moment angle complex Z(Q) as follows.<br />

Z(Q) := p −1<br />

s (d F (Q)). (1.8.8)


Chapter 1: Quasi<strong>toric</strong> manifolds 24<br />

Let Λ : Z m → Z n be the map <strong>of</strong> Z-modules which maps the standard generator e i <strong>of</strong><br />

Z m to the characteristic vector ξ i . Let K denote the kernel <strong>of</strong> this map. The sequence<br />

0 −→ K −→ Z m Λ −→ Z n −→ 0 (1.8.9)<br />

splits and we can write Z m = K ⊕ Z n . The torus T K := (K ⊗ R)/K is a subtorus <strong>of</strong><br />

T m and we have a split exact sequence<br />

1 −→ T K −→ T m Λ −→<br />

∗<br />

T n −→ 1 (1.8.10)<br />

Denote the Z-module Z m by N. For any face F <strong>of</strong> Q let N(F ) be the submodule <strong>of</strong><br />

N generated by the basis vectors e i such that d F (F ) intersects the i-th facet <strong>of</strong> R m ≥0 , that<br />

is the coordinate hyperplane {x i = 0}. Note that image <strong>of</strong> N(F ) under Λ is precisely<br />

Z(F ), so that the preimage Λ −1 (Z(F )) = K · N(F ). Consider the exact sequence<br />

0 −→ K · N(F )<br />

N(F )<br />

−→<br />

N<br />

N(F )<br />

Λ<br />

−→<br />

Zn<br />

−→ 0 (1.8.11)<br />

Z(F )<br />

Since the characteristic vectors corresponding to the facets whose intersection is F<br />

are linearly independent, it follows from the definition <strong>of</strong> K and Λ that K ∩N(F ) = {0}.<br />

Hence by the second isomorphism theorem we have a canonical isomorphism<br />

K · N(F )<br />

N(F )<br />

∼= K (1.8.12)<br />

So the equation 1.8.11 yields<br />

0 −→ K −→ N<br />

N(F )<br />

Λ<br />

−→<br />

We obtain the following split exact sequence <strong>of</strong> tori<br />

Zn<br />

−→ 0 (1.8.13)<br />

Z(F )<br />

0 −→ T K −→ T(F ; N) ⊥ −→ T(F ; Z n ) ⊥ −→ 0 (1.8.14)<br />

where T(F ; Z n ) ⊥ is the fiber <strong>of</strong> p : M 2n → Q and T(F ; N) ⊥ is the fiber <strong>of</strong> p s : Z(Q) → Q<br />

over any point in the relative interior <strong>of</strong> the arbitrary face F . It follows that M 2n is a<br />

quotient <strong>of</strong> Z(Q) by the above action <strong>of</strong> T K . This action <strong>of</strong> T K is same as the restriction<br />

<strong>of</strong> its action on C m as a subtorus <strong>of</strong> T m .<br />

Denote the standard basis <strong>of</strong> N by {e j : 1 ≤ j ≤ m}. The dual N ∗ := Hom Z (N, Z)<br />

<strong>of</strong> N is the character group <strong>of</strong> T m . That is any character is uniquely represented by<br />

∑<br />

aj e ∗ j where {e∗ j : 1 ≤ j ≤ m} denote the basis <strong>of</strong> L∗ dual to {e j } and a j ∈ Z. Denote<br />

the irreducible representation <strong>of</strong> T m corresponding to e ∗ j by C(e∗ j ). So the irreducible<br />

representation <strong>of</strong> T m corresponding to the character ∑ a j e ∗ j is C(∑ a j e ∗ j ) := ⊗ j C(e∗ j )a j<br />

.


25 1.8 Cohomology ring<br />

Define a line bundle ν(F i ) := Z(Q) × TK C(e ∗ i ) corresponding to each facet F i <strong>of</strong> Q. The<br />

following Lemma’s are proved in [DJ91]. We prove these Lemma’s in a different way<br />

following [Poddf].<br />

Lemma 1.8.3. c 1 (ν(F i )) = ι ∗ (w i ).<br />

Pro<strong>of</strong>. Note that<br />

ET m × T m (Z(Q) × C(e ∗ i )) = ET m × T n (Z(Q) × TK C(e ∗ i )) ≃ ET n × T n ν(F i ).<br />

Let x ∈ M(F i , ξ) be a fixed point <strong>of</strong> T n action on M 2n . Then as T n -representations,<br />

ν(F i )| x = C(ξi ∗). The line bundle ETn × T n C(ξi ∗) over BTn is equal to pri ∗(γ 1,∞). Let<br />

ῑ : x ↩→ M 2n be the inclusion map. So there is an associated umkehrungs homomorphism<br />

in equivariant cohomology ῑ ∗ : H ∗ (BT n ) → H ∗ (BQ). Let b e : BQ → BT n is the<br />

equivariant version to the collapsing map M 2n ↦→ {x}. The map ῑ ∗ can be identified with<br />

the map b ∗ e : H ∗ (BT n ) → H ∗ (BQ). In Theorem 1.8.1 we have identified ῑ ∗ c 1 (pri ∗γ 1,∞)<br />

with w i ∈ H ∗ (BQ). We consider the inclusion ι : M 2n ↩→ BQ as a fiber <strong>of</strong> b e . Then we<br />

have the following commutative diagram<br />

ν(F i ) −−−−→ ET n × T n ν(F i ) −−−−→ ET n × T n C(ξi ∗ )<br />

⏐<br />

⏐<br />

⏐<br />

↓<br />

↓<br />

↓<br />

N<br />

ι<br />

−−−−→<br />

BQ<br />

κ<br />

b e<br />

−−−−→ BT n .<br />

(1.8.15)<br />

Thus c 1 (ν(F i ) = ι ∗ b e ∗ c 1 (pr ∗ i (γ 1,∞)) = ι ∗ (w i ).<br />

Lemma 1.8.4. The line bundle ⊗ i ν(F i) a i<br />

(a 1 , . . . , a m ) belongs to the row space <strong>of</strong> the matrix <strong>of</strong> Λ.<br />

over M 2n is trivial if and only if the vector<br />

Pro<strong>of</strong>. Since ν(F i ) = Z(Q) × TK C(e ∗ i ), ⊗ i ν(F i) a i<br />

= Z(Q) × TK C( ∑ a i e ∗ i ). This line<br />

bundle is trivial if and only if the character ∑ a i e ∗ i restricts to the trivial character on<br />

T K . This holds if and only if ∑ a i e ∗ i (u) = 0 for all u ∈ K = Ker(Λ). This is equivalent<br />

to saying that (a 1 , . . . , a m ) is a linear combination <strong>of</strong> rows <strong>of</strong> Λ.<br />

Corollary 1.8.5. ∑ m<br />

i=1 a iι ∗ (w i ) = 0 whenever (a 1 , . . . , a m ) is in the row space <strong>of</strong> Λ.<br />

The above calculations in terms <strong>of</strong> characters imply that b e ∗ : H 2 (BT n ) → H 2 (BQ)<br />

can be identified with the map Λ ∗ : (Z n ) ∗ → (Z m ) ∗ , where Λ ∗ denotes the dual <strong>of</strong> the<br />

characteristic map Λ. So we have the following short exact sequence,<br />

b e<br />

∗<br />

0 → H 2 (BT n ) −−−−→ H 2 (BQ)<br />

∥<br />

∥<br />

(Z n ) ∗ Λ ∗<br />

−−−−→ (Z m ) ∗ .<br />

ι ∗<br />

−−−−→ H 2 (N) → 0


Chapter 1: Quasi<strong>toric</strong> manifolds 26<br />

Let J be the homogeneous ideal in Z[w 1 , . . . , w m ] generated by the polynomials<br />

{ξ j := ∑ m<br />

i=1 a j i<br />

w i |1 ≤ j ≤ n}, where (a j1 , . . . , a jm ) denotes the j-th row <strong>of</strong> Λ. The<br />

element ξ j can be identified with the image <strong>of</strong> the j-th generator <strong>of</strong> H 2 (BT n ) under Λ ∗ .<br />

Let J be the image <strong>of</strong> J in SR(Q, Z). Since ι ∗ : SR(Q, Z) → H ∗ (M 2n , Z) is onto and<br />

J is belongs to its kernel, ι ∗ induces a surjection SR(Q, Z)/J → H ∗ (M 2n , Z).<br />

Theorem 1.8.6 (Theorem 4.14, [DJ91]). Let M 2n be the quasi<strong>toric</strong> manifold associated<br />

to the characteristic model (Q, ξ). Then H ∗ (M 2n ; Z) is the quotient <strong>of</strong> the face ring<br />

SR(Q, Z) by J . That is H ∗ (M 2n ; Z) = Z[w 1 , . . . , w m ]/(I + J ).<br />

Pro<strong>of</strong>. We know that the cohomology ring H ∗ (BT n , Z) is a polynomial ring on n generators.<br />

The face ring H ∗ (BQ, Z) ∼ = H ∗ (BT n ) ⊗ H ∗ (M 2n ) as Z-modules. Also the map<br />

b e ∗ : H ∗ (BT n , Z) → H ∗ (BQ, Z) is an injection and J is identified with the image <strong>of</strong> b e ∗ .<br />

Thus H ∗ (M 2n ) = H ∗ (BQ, Z)/J = Z[w 1 , . . . , w m ]/(I + J ), where I is an ideal defined<br />

in definition 1.4.2.<br />

1.9 Smooth and stable complex structure<br />

In Section 6 <strong>of</strong> [BR01] the authors describe the existence <strong>of</strong> smooth equivariant connected<br />

sum operation for quasi<strong>toric</strong> manifolds. In this section we follow the paper [BR01]<br />

and the lecture note [Poddf]. We will realize the quasi<strong>toric</strong> manifold M 2n as the quotient<br />

<strong>of</strong> an open subset <strong>of</strong> C m . We follow the notation <strong>of</strong> Section 1.8. Identify R m with<br />

the space <strong>of</strong> functions R F(Q) . Consider the thickening W ⊂ R m ≥0 <strong>of</strong> the image d F(Q),<br />

defined by<br />

W = {f : F(Q) → R ≥0 |f −1 (0) ∈ L F (Q)} (1.9.1)<br />

where L F (Q) denotes the face lattice <strong>of</strong> Q. Let V Q be the n-dimensional linear subspace<br />

<strong>of</strong> R m parallel to d F (Q) and VQ ⊥ be its orthogonal complement in Rm . The group<br />

G := exp(VQ ⊥) acts naturally on Rm and W by coordinatewise multiplication. We want<br />

to produce a thickening Q ⊂ W <strong>of</strong> d F (Q) which will be close to a G principal bundle<br />

over d F (Q).<br />

Lemma 1.9.1 (Lemma 5.1, [Poddf]). The tangent spaces to the orbits <strong>of</strong> G-action on<br />

R m form an integrable distribution.<br />

Pro<strong>of</strong>. Let (y 1 , . . . , y m ) be the standard coordinates on R m . Let ∗ be the binary<br />

operation <strong>of</strong> coordinatewise multiplication <strong>of</strong> two vectors in R m . Let the vectors<br />

c j = (c j1 , . . . , c jm ), 1 ≤ j ≤ m − n, form a orthogonal basis for the subspace VQ ⊥ <strong>of</strong> Rm .<br />

Suppose q is any fixed point in R m with coordinate vector y(q) = (y 1 (q), . . . , y m (q)).<br />

Then the vectors y(q) ∗ c j , 1 ≤ j ≤ m − n, span the tangent space to the orbit through<br />

q <strong>of</strong> G-action. Clearly the Lie bracket <strong>of</strong> any two vector fields y(q) ∗ c j = ∑ y i (q)c ji<br />

and y(q) ∗ c k = ∑ ∂<br />

y i c ki is zero. Thus the distribution is integrable.<br />

∂y i<br />

∂<br />

∂y i


27 1.9 Smooth and stable complex structure<br />

Definition 1.9.2. We will denote the distribution consisting <strong>of</strong> tangent vectors to orbits<br />

<strong>of</strong> G-action on R m by χ.<br />

Consider the following decomposition <strong>of</strong> the space W as in [BR01]. Observe that<br />

each point q ∈ Q determines a function d F (q) : F(Q) → R, where d(q, F i ) is the<br />

Euclidean distance between p and the hyperplane containing the facet F i ∈ F(Q) for<br />

i = 1, . . . , m. These maps produce an embedding d F (Q) ⊂ R m <strong>of</strong> Q. For any subset<br />

G(Q) ⊂ F(Q), we realize R G(Q) as a subspace <strong>of</strong> R F(Q) by choosing those coordinates y i<br />

to be zero for which F i ∈ F(Q) − G(Q). Let F be a face (may be empty) <strong>of</strong> Q. Denote<br />

the set <strong>of</strong> facets <strong>of</strong> Q that contain F by F F . Let G F = F(Q) − F F . Then W is the<br />

union <strong>of</strong> open cones ∪ R G F<br />

>0 .<br />

Lemma 1.9.3 (Lemma 5.2, [Poddf]). For any face F <strong>of</strong> Q the orbits <strong>of</strong> G-action define<br />

a foliation on R G F<br />

>0 .<br />

Pro<strong>of</strong>. Clearly the open cone R G F<br />

>0 is an invariant subset under the G-action. Let q ∈ RG F<br />

>0<br />

be any point. Let B be a matrix whose row vectors c j form a basis for V ⊥ Q . Consider<br />

the matrix B(q) whose rows are y(q) ∗ c j . Then the row vectors <strong>of</strong> B(q) span the vector<br />

space χ(q). Denote the columns <strong>of</strong> the matrix B by β i , 1 ≤ i ≤ m. Then the i-th<br />

column <strong>of</strong> B(q) is y i (q)β i . This Lemma is clear if F is an empty face.<br />

Now suppose F is a vertex v <strong>of</strong> Q. Let d F (v) = q. Then exactly n coordinates <strong>of</strong><br />

y(q) are zero. Without loss <strong>of</strong> generality we may assume that the facets F 1 , . . . , F n <strong>of</strong><br />

Q meet at v. So we get that y 1 (q) = . . . = y n (q) = 0. Note that the vectors u 1 , . . . , u n<br />

tangent to the edges <strong>of</strong> d F (Q) meeting at q form a basis <strong>of</strong> V Q . By our assumption each<br />

vector u i has 0 in the first n positions except for the i-th position. A vector z belongs<br />

to V ⊥ Q<br />

if and only if it satisfies the following system <strong>of</strong> linear equations;<br />

z · u j = 0, for each 1 ≤ j ≤ n. (1.9.2)<br />

Let e 1 , . . . , e m be the standard basis <strong>of</strong> R m . Solving the above system we get a basis<br />

c 1 , . . . , c m−n <strong>of</strong> V ⊥ Q where c k has the form c k1 e 1 + . . . + c kn e n + e k+n for 1 ≤ k ≤ m − n.<br />

Hence we can assume that β n+1 , . . . , β m are linearly independent. Since the coordinates<br />

{y i (q) : n + 1 ≤ i ≤ m} are each positive we get y n+1 (q)β n+1 , . . . , y m (q)β m<br />

are linearly independent. Hence B(q) has the rank m − n. Therefore the integrable<br />

distribution χ has constant rank and corresponds to a foliation.<br />

The argument for faces <strong>of</strong> higher dimension is similar and follows from the zero<br />

dimensional case.<br />

Lemma 1.9.4 (Lemma 5.3, [Poddf]). The integrable distribution χ forms a foliation<br />

on a neighborhood ˜W <strong>of</strong> W in R m .<br />

Pro<strong>of</strong>. It is enough to show that the distribution χ has constant rank m − n in a neighborhood<br />

<strong>of</strong> each point q in R G F<br />

>0 where F Q is a face <strong>of</strong> Q. Without loss <strong>of</strong> generality


Chapter 1: Quasi<strong>toric</strong> manifolds 28<br />

we may assume that y i (q) > 0 for all i ≥ n + 1 and that β n+1 , . . . , β m are linearly<br />

independent. Let U be a small open ball in R m around q. Let s be any point in U.<br />

Then we may assume that y i (s) > 0 for all i ≥ n + 1. Hence y n+1 (s)β n+1 , . . . , y m (s)β m<br />

are linearly independent. Hence the vector space χ(s) has the rank m − n.<br />

Lemma 1.9.5 (Lemma 5.4, [Poddf]). The orbits <strong>of</strong> G-action are transverse to d F (Q).<br />

Pro<strong>of</strong>. Let q be a point in d F (Q) with coordinate vector y(q). Then a tangent vector<br />

to the orbit <strong>of</strong> G-action through q has the form c ∗ y(q) where c = (c 1 , . . . , c m ) ∈ V ⊥ Q .<br />

The inner product 〈c, c ∗ y(q)〉 = ∑ c 2 i y i(q). This is a strictly positive quantity. Thus<br />

the set χ(q) ∩ V Q is singleton. Since χ(q) and V Q have complementary dimensions, they<br />

are transversal.<br />

Definition 1.9.6. Define Q to be the union <strong>of</strong> all G orbits that pass through d F (Q).<br />

Lemma 1.9.7 (Lemma 5.5, [Poddf]). The space Q is an open subset <strong>of</strong> R m ≥0 .<br />

Pro<strong>of</strong>. Let y ∈ Q and X y denote the G orbit through y. By definition <strong>of</strong> Q, X y meets<br />

d F (Q) at some point y 1 . Let α be any path in X y from y 1 to y. Let T be a transversal<br />

to the foliation χ on ˜W at y. Let ˜Q := ˜W ∩ (V Q + y(q)) where q is any point on d F (Q).<br />

Since ˜Q is transversal to χ at y 1 , there exists small open set U 1 ⊂ ˜Q around y 1 which<br />

maps diffeomorphically onto a small open set U ⊂ T around y via the holonomy <strong>of</strong> the<br />

foliation χ along α. Note that each leaf <strong>of</strong> χ that intersects W lies completely in W.<br />

Hence the holonomy <strong>of</strong> χ along α maps U 1 ∩ d F (Q) onto U ∩ W. Thus there exists a<br />

small foliation chart X ⊂ ˜W around y such that every plaque in X ∩ W lies in a leaf<br />

that hits d F (Q). Therefore the neighborhood X ∩ W <strong>of</strong> y in W is contained in Q. Thus<br />

Q is open in W. Since W is open in R m ≥0 , so is the space Q.<br />

Let O m be the origin <strong>of</strong> R m . Let Q ′ be a small tubular neighborhood <strong>of</strong> d F (Q) in<br />

Q such that:<br />

1. Q ′ is diffeomorphic to the product <strong>of</strong> Q × S where S is an open neighborhood <strong>of</strong><br />

the identity in G,<br />

2. Q ′ is bounded and the Euclidean distance from Q ′ to O m is positive.<br />

Being a foliation local triviality is provided and global triviality follows from contractibility<br />

<strong>of</strong> d F (Q). Denote the restriction <strong>of</strong> the foliation χ on Q ′ by the same.<br />

Consider the group Υ = G × T K . So Υ is a subgroup <strong>of</strong> (C ∗ ) m . Let Z ′ := p −1<br />

s (Q ′ ).<br />

The map p s : Z ′ → Q ′ is smooth and transversal to χ. Also T K acts on Z ′ freely. These<br />

two fact induce a foliation, say χ ′ , on Z ′ . Let M be the leaf space <strong>of</strong> this foliation. In<br />

fact this is a fiber bundle over M. From the choice <strong>of</strong> Q ′ we get that each fiber <strong>of</strong> Z ′<br />

over M is diffeomorphic to S × T K where S is an open neighborhood <strong>of</strong> the identity in


29 1.9 Smooth and stable complex structure<br />

G. We use Lie groupoid theory to show a smooth structure on M. Now we give the<br />

definition <strong>of</strong> Lie groupoid following [ALR07].<br />

Definition 1.9.8. A Lie groupoid G consists <strong>of</strong> two smooth Hausdorff manifolds, one<br />

is a set <strong>of</strong> objects G 0 and another is a set <strong>of</strong> invertible arrows G 1 , together with the<br />

following smooth maps where the maps s, t are submersions.<br />

1. The source map s : G 1 → G 0 which assigns each arrow g to its source s(g).<br />

2. The target map t : G 1 → G 0 which assigns each arrow g to its target t(g). For an<br />

arrow g ∈ G 1 , we write g : y → x to indicate that s(g) = y and t(g) = x.<br />

3. The composition map c : G 1 × (s,t) G 1 → G 1 , where<br />

G 1 × (s,t) G 1 = {(g, h) ∈ G 1 × G 1 : s(h) = t(g)},<br />

is defined in the following. If g : x → y and h : y → z are two arrows then we<br />

can define their composition arrow hg : x → z. That is the composition map c is<br />

defined by c(g, h) = hg and is required to be associative.<br />

4. The identity map i : G 0 → G 1 such that si(x) = x = ti(x) and gi(x) = g = i(y)g<br />

for all x, y ∈ G 0 and g ∈ G 1 with g : x → y.<br />

5. An inverse map u : G 1 → G 1 , written by u(g) = g −1 , is a two-sided inverse for the<br />

composition. That is, if g : x → y then g −1 : y → x such that g −1 g = i(x) and<br />

gg −1 = i(y).<br />

Example 1.9.9. Let M be a connected manifold. Then the fundamental groupoid Π(M)<br />

<strong>of</strong> M is the groupoid with the space <strong>of</strong> objects Π(M) 0 = M and each homotopy class g<br />

<strong>of</strong> paths from x to y is an arrow g : x → y.<br />

Example 1.9.10. Let K be a Lie group which acts smoothly on a manifold M from the<br />

left. Define a Lie groupoid K ⋉ M by setting (K ⋉ M) 0 = M and (K ⋉ M) 1 = K × M,<br />

where the source map s : K × M → M is the projection onto the second factor and the<br />

target map t : K × M → M is the group action. Let g : x 1 → k 1 x 1 and h : x 2 → k 2 x 2<br />

be two arrows in (K ⋉ M) 1 such that k 1 x 1 = x 2 . The composition map is defined by<br />

hg : x 1 → (k 2 k 1 )x 1 . The identity map i : M → K × M is defined by i(x) = (1 K , x).<br />

This groupoid is called the translation groupoid associated to the group action. The unit<br />

groupoid is the translation groupoid for the action <strong>of</strong> the trivial group. Also by taking<br />

M to be a point we can view any Lie group K as a Lie groupoid with a single object.<br />

Definition 1.9.11. Let G be a Lie groupoid with the set <strong>of</strong> objects G 0 and the set <strong>of</strong><br />

arrows G 1 . For a point x ∈ G 0 , the set <strong>of</strong> all arrows g : x → x is a Lie group. Denote<br />

it by G x , called the isotropy group at x. The set ts −1 (x) <strong>of</strong> targets <strong>of</strong> arrows out <strong>of</strong> x


Chapter 1: Quasi<strong>toric</strong> manifolds 30<br />

is called the orbit <strong>of</strong> x. The orbit space |G| <strong>of</strong> G is the quotient space <strong>of</strong> G 0 under the<br />

equivalence relation x ∼ y if and only if x and y are in the same orbit. Conversely, G<br />

is called a groupoid presentation <strong>of</strong> |G|.<br />

Definition 1.9.12. A Lie groupoid G is called proper if the map (s, t) : G 1 → G 0 × G 0<br />

is a proper map.<br />

Note that in a proper Lie groupoid G, every isotropy group is compact (see Proposition<br />

1.37 <strong>of</strong> [ALR07]).<br />

Theorem 1.9.13 (Theorem 5.6, [Poddf]). The leaf space M <strong>of</strong> the foliation χ ′ can be<br />

identified with the quasi<strong>toric</strong> manifold M 2n . Thus M 2n has a smooth structure.<br />

Pro<strong>of</strong>. Let G be the Lie groupoid with the set <strong>of</strong> objects G 0 = Z ′ and the set <strong>of</strong> arrows<br />

G 1 = {(z, g) : z ∈ Z ′ , g ∈ Υ, z · g ∈ Z ′ }. Since the group G acts on R m coordinatewise<br />

and T K is compact, this is a proper Lie groupoid. Clearly isotropy group <strong>of</strong> each<br />

x ∈ G 0 is trivial. So the set ts −1 (x), the orbit <strong>of</strong> x in G, is diffeomorphic to S × T K<br />

for each x ∈ G 0 . That is ts −1 (x) the leaf <strong>of</strong> the foliation χ ′ on G 0 . Thus M is the<br />

orbit space <strong>of</strong> G. So there exists an embedded submanifold U x transversal to ts −1 (x)<br />

at x and a neighborhood S u ⊂ S <strong>of</strong> identity in G such that the map U x × S u → G 0<br />

given by the action is diffeomorphic onto its image. So the space M has a natural<br />

smooth structure (see Chapter 5, [MM03]).<br />

Topologically M can be identified with<br />

the quotient Z(P )/T K that is M 2n . As the T m ⊂ (C ∗ ) m action on Z ′ is smooth, the<br />

induced T n action on M is smooth. Hence M is a smooth quasi<strong>toric</strong> manifold with<br />

same characteristic pair. Therefore by the classification result 1.2.17 it is equivariantly<br />

homeomorphic to M 2n .<br />

The following Theorem was first proved by Buschtabar and Ray in [BR01]. We give<br />

a different pro<strong>of</strong> following the lecture note [Poddf].<br />

Theorem 1.9.14. A quasi<strong>toric</strong> manifold M 2n has a stable complex structure.<br />

Pro<strong>of</strong>. Let T Z ′ be the tangent bundle <strong>of</strong> Z ′ . Υ acts naturally on T Z ′ . Define the space<br />

D := T Z ′ /Υ, so that a point <strong>of</strong> D is a field <strong>of</strong> tangent vectors to Z ′ , defined along<br />

one <strong>of</strong> its fibers, and invariant under Υ. Let U denote the subbundle <strong>of</strong> T Z ′ formed by<br />

vectors tangential to the fibers <strong>of</strong> Z ′ over M. Then Υ acts on U and define R = U/Υ.<br />

The arguments <strong>of</strong> Atiyah [Ati57] apply, with complex analytic replaced by smooth.<br />

Therefore D has a natural vector bundle structure over M 2n and R is a subbundle <strong>of</strong> D.<br />

Moreover the following sequence <strong>of</strong> vector bundles, where T M 2n denotes the tangent<br />

bundle <strong>of</strong> M 2n , is exact.<br />

0 → R → D → T M 2n → 0 (1.9.3)


31 1.10 Chern classes<br />

The standard complex structure on C m restricts to a complex structure J on T Z ′ .<br />

The action <strong>of</strong> Υ commutes with J. Thus D inherits a complex structure, which we<br />

again denote by J.<br />

Let Υ denote the Lie algebra <strong>of</strong> Υ. Since every fiber <strong>of</strong> Z ′ can be identified as a Lie<br />

groupoid to a neighborhood <strong>of</strong> the identity in Υ, and Υ is commutative, following [Ati57]<br />

we obtain that R is isomorphic to Υ × M 2n . Thus we get a complex structure on<br />

T M 2n ⊕ (R 2m−2n × M 2n ), i.e. a stable complex structure on M 2n .<br />

Remark 1.9.15. Note that each omniorientation determines a stable complex structure<br />

by the above procedure. That is there is a canonical choice <strong>of</strong> stable complex structure<br />

only if omniorientation is fixed.<br />

The space T Z ′ splits naturally into a direct sum <strong>of</strong> m complex line bundles corresponding<br />

to the complex coordinate directions. These directions correspond to the<br />

facets <strong>of</strong> Q. We get a corresponding splitting D = ⊕ ν(F i ).<br />

1.10 Chern classes<br />

The existence <strong>of</strong> stable almost complex structure on M 2n implies that Chern classes<br />

can be defined. These classes depend on the choice <strong>of</strong> omniorientation on M 2n .<br />

Definition 1.10.1. The total Chern class <strong>of</strong> an omnioriented quasi<strong>toric</strong> manifold M 2n<br />

is defined to be the total Chern class <strong>of</strong> its stable tangent bundle D, c(T M 2n ) := c(D).<br />

Since D = ⊕ ν(F i ), by the Whitney product formula we obtain<br />

m∏<br />

m∏<br />

c(T M 2n ) = (1 + c 1 (ν(F i ))) = (1 + w i ). (1.10.1)<br />

i=1<br />

i=1<br />

Note that in our notation, w i depends on the characteristic vectors λ i and if we want<br />

to compare different omniorientations the signs <strong>of</strong> w i s have to be adjusted with respect<br />

to some fixed choice.<br />

From the combinatorial information <strong>of</strong> the combinatorial model we can calculate<br />

the Chern numbers <strong>of</strong> M 2n by localization. In a different approach, Panov [Pan01]<br />

uses results from index theory to give a beautiful formula for the χ y -genus <strong>of</strong> M 2n . To<br />

present the formula we need to first introduce some notation.<br />

Suppose v is a vertex <strong>of</strong> simple polytope Q which is the intersection <strong>of</strong> n facets<br />

F i1 , . . . , F in . To each facet F ik assign the unique edge E k such that E k ∩ F ik = v. Let<br />

e k be a vector along E k with origin v. Order the facets at v such that e 1 , . . . , e n form<br />

a positively oriented basis <strong>of</strong> R n . The characteristic vectors associated to these facets<br />

are also ordered accordingly. Adopting this convention on ordering for each vertex, we<br />

make the following definition.


Chapter 1: Quasi<strong>toric</strong> manifolds 32<br />

Definition 1.10.2. The sign sign(v) <strong>of</strong> a vertex v = F i1 ∩ . . . ∩ F in<br />

determinant <strong>of</strong> the n × n matrix Λ (v) := [λ i1 · · · λ in ].<br />

is defined to be the<br />

Let E be an edge <strong>of</strong> Q. Recall the module Z(E) from Theorem 1.2.9. Let α be a<br />

generator <strong>of</strong> the module Z(E) ⊥ := {v ∈ Z n | 〈v, w〉 = 0 ∀ w ∈ Z(E)}. We refer to α as<br />

the edge vector corresponding to E. It is determined up to choice <strong>of</strong> sign. For a given<br />

omniorientation, the sign <strong>of</strong> α may be locally fixed at a vertex v = F i1 ∩ . . . ∩ F in <strong>of</strong><br />

E = F i2 ∩ . . . ∩ F in by requiring that 〈α, λ i1 〉 > 0. Then each vertex has n well defined<br />

edge vectors α 1 , . . . , α n (ordered according to the convention discussed above).<br />

Definition 1.10.3. Let η ∈ Z n be a primitive vector such that 〈α, η〉 ̸= 0 for any edge<br />

vector α <strong>of</strong> Q. Then define the index <strong>of</strong> a vertex v = F i1 ∩ . . . ∩ F in <strong>of</strong> Q with respect to<br />

η to be<br />

ind η (v) := #{1 ≤ k ≤ n | 〈α k , η〉 < 0}.<br />

Theorem 1.10.4 (Theorem 3.1, [Pan01]). Let M 2n be an omnioriented quasi<strong>toric</strong> manifold.<br />

For any primitive vector η ∈ Z n such that 〈α k , η〉 ̸= 0 for any edge vector α k <strong>of</strong><br />

Q, the χ y -genus <strong>of</strong> M 2n may be calculated as<br />

χ y (X) = ∑ v<br />

(−y) ind η(v) sign(v).<br />

For the values 1, 0, 1 <strong>of</strong> y the χ y -genus specializes to the top Chern number, the Todd<br />

genus and signature or L-genus <strong>of</strong> M 2n respectively. Thus we readily obtain formulae<br />

for these important invariants from the above theorem. For instance, formula for the<br />

top Chern number <strong>of</strong> quasi<strong>toric</strong> manifold M 2n over Q is,<br />

c n (M 2n ) = ∑ v<br />

sign(v). (1.10.2)


Chapter 2<br />

Small covers and orbifolds<br />

2.1 Introduction<br />

The category <strong>of</strong> small covers were introduced by Davis and Januszkiewicz [DJ91]. Following<br />

the paper [DJ91] we discuss some basic theory about small covers.<br />

Orbifolds were introduced by Satake [Sat57], who called them V -manifolds. Orbifolds<br />

are singular spaces that are locally look like as a quotient <strong>of</strong> an open subset <strong>of</strong><br />

Euclidean space by an action <strong>of</strong> a finite group. Following [ALR07], we provide a definition<br />

<strong>of</strong> effective orbifolds. We recall the tangent bundle and Orbifold fundamental<br />

group <strong>of</strong> an orbifold.<br />

2.2 Small covers<br />

Small covers are real analog <strong>of</strong> quasi<strong>toric</strong> manifolds. Let N s be an n-dimensional manifold<br />

and ρ : Z n 2 × Rn → R n be the standard action.<br />

Definition 2.2.1. An action η : Z n 2 × N s → N s is said to be locally standard if the<br />

followings hold.<br />

1. Every point y ∈ N s has a Z n 2 -stable open neighborhood U y, that is η(Z n 2 ×U y) = U y .<br />

2. There exists a homeomorphism ψ : U y → V , where V is a Z n 2 -stable (that is<br />

ρ(Z n 2 × V ) = V ) open subset <strong>of</strong> Rn .<br />

3. There exists an isomorphism δ y : Z n 2 → Zn 2 such that ψ(η(t, x)) = ρ(δ y(t), ψ(x))<br />

for all (t, x) ∈ Z n 2 × U y.<br />

Definition 2.2.2. A closed n-dimensional manifold N s is said to be a small cover if<br />

there is an effective Z n 2 -action on N s such that<br />

1. the action is a locally standard action,<br />

33


Chapter 2: Small covers and orbifolds 34<br />

2. the orbit space <strong>of</strong> the action is a simple polytope.<br />

Example 2.2.3. Suppose T 1 = {(x, y) ∈ R 2 : |x| 2 + |y| 2 = 1} is the unit circle in R 2 .<br />

Let l be a line in R 2 passing through the origin. The group Z 2 acts on T 1 by a reflection<br />

along the line l. Denote this action by<br />

ρ l : Z 2 × T 1 → T 1 . (2.2.1)<br />

Clearly the orbit space is an interval and the action is locally standard.<br />

Example 2.2.4. Consider the n-fold product T n = (T 1 ) n ⊂ (R 2 ) n . An action <strong>of</strong> Z n 2<br />

defined on T n by<br />

ρ n ((g 1 , g 2 , . . . , g n ), (t 1 , t 2 , . . . , t n )) = (ρ l1 (g 1 , t 1 ), . . . , ρ ln (g n , t n )), (2.2.2)<br />

where l i is a line belongs to the i-th component <strong>of</strong> (R 2 ) n , for i = 1, 2, . . . , n. The action<br />

ρ n is locally standard and the orbit space is the standard n-cube in R n . So T n is a small<br />

cover over the n-cube.<br />

Example 2.2.5. The natural action <strong>of</strong> Z n 2 defined on the real projective space RPn by<br />

(g 1 , . . . , g n ) · [x 0 , x 1 , . . . , x n ] → [x 0 , g 1 x 1 , . . . , g n x n ] (2.2.3)<br />

is locally standard and the orbit space is diffeomorphic as manifold with corners to the<br />

standard n-simplex. Hence RP n is a small cover over the n-simplex △ n .<br />

Remark 2.2.6. We can define an equivariant connected sum <strong>of</strong> small covers following<br />

the Section 1.7 <strong>of</strong> Chapter 1. The equivariant connected sum <strong>of</strong> n-dimensional<br />

finitely many small covers is also a small cover. For example, the connected sum<br />

RP n #RP n # . . . #RP n <strong>of</strong> k copies <strong>of</strong> RP n is a small cover over the connected sum<br />

△ n # △ n # . . . #△ n (k times) <strong>of</strong> △ n .<br />

Let F(Q) = {F 1 , . . . , F m } be the set <strong>of</strong> facets <strong>of</strong> a simple n-polytope Q. We denote<br />

the underlying additive group <strong>of</strong> the vector space F n 2 by Zn 2 .<br />

Definition 2.2.7. The function β : F(Q) → F n 2 is called Z 2-characteristic function on<br />

Q if the span <strong>of</strong> {β(F j1 ), . . . , β(F jl )} is an l-dimensional subspace <strong>of</strong> F n 2 whenever the<br />

intersection <strong>of</strong> the facets F j1 , . . . , F jl is nonempty.<br />

The vectors β(F j ) = β j are called Z 2 -characteristic vectors and the pair (Q, β) is<br />

called Z 2 -characteristic pair.<br />

We show that associated to a small cover there exists a Z 2 -characteristic pair. Then<br />

we construct a small cover from a Z 2 -characteristic pair.


35 2.2 Small covers<br />

Let ς : N s → Q be a small cover over n-polytope Q. By the locally standard<br />

property <strong>of</strong> the Z n 2 -action on N s we can show that each N Fj = ς −1 (F j ) is a connected<br />

(n − 1)-dimensional submanifold <strong>of</strong> N s for j = 1, . . . , m. The submanifold N Fj is fixed<br />

pointwise by the subgroup G j ( ∼ = Z 2 ) <strong>of</strong> (Z 2 ) n . So we can correspond each facet F j to<br />

the subgroup G j . Let β j ∈ (Z 2 ) n be the nonidentity element <strong>of</strong> G j . Hence we can define<br />

a function<br />

β : F(Q) → F n 2 by β(F j ) = β j . (2.2.4)<br />

If the intersection <strong>of</strong> the facets F j1 , . . . , F jk is nonempty then F = F j1 ∩ . . . ∩ F jk is a<br />

codimension-k face <strong>of</strong> Q. Then the isotropy group G F <strong>of</strong> the submanifold ς −1 (F ) ⊂ N s<br />

is the subgroup <strong>of</strong> Z n generated by β j1 , . . . , β jk . Let v be a vertex <strong>of</strong> F and y = ς −1 (v).<br />

Comparing the action <strong>of</strong> Z n 2 on a Zn 2 -stable neighborhood <strong>of</strong> y in N s to the standard<br />

action we get that G F is a k-dimensional subspace <strong>of</strong> F n 2 . So we can assign a unique<br />

subgroup G F to each face F <strong>of</strong> Q. Hence (Q, β) is a Z 2 -characteristic pair.<br />

Let (Q, β) be a Z 2 -characteristic function. Let G F be the subgroup <strong>of</strong> Z n 2 generated<br />

by {β j1 , . . . , β jl } whenever F = F j1 ∩ . . . ∩ F jl . Define an equivalence relation ∼ z on<br />

Z n 2 × Q by (t, p) ∼ z (s, q) if p = q and s − t ∈ G F (2.2.5)<br />

where F ⊂ Q is the unique face whose relative interior contains p. Let<br />

N(Q, β) = (Z n 2 × Q)/ ∼ z<br />

be the quotient space. Following the pro<strong>of</strong> <strong>of</strong> the theorem 1.2.9 we can show that the<br />

quotient space N(Q, β) is a manifold. The action <strong>of</strong> Z n 2 by the left translations descends<br />

to a locally standard Z n 2 -action on N(Q, β). The projection onto the second factor <strong>of</strong><br />

Z n 2 × Q descends to a projection ς β : N(Q, β) → Q. Hence N(Q, β) is an n-dimensional<br />

small cover over Q.<br />

Theorem 2.2.8 (Proposition 1.8, [DJ91]). Let ς : N s → Q be a small cover over Q<br />

and the function β : F(Q) → F n 2 defined in 2.2.4 be its Z 2-characteristic function. Let<br />

ς β : N(Q, β) → Q be the constructed small cover from the pair (Q, β). Then there exists<br />

an equivariant homeomorphism from N s to N(Q, β) covering the identity over Q. Hence<br />

small cover is determined up to equivalence over Q by its Z 2 -characteristic function.<br />

Remark 2.2.9. The constructive definition <strong>of</strong> small cover give an idea to introduce the<br />

notion <strong>of</strong> small orbifolds, see Chapter 4.<br />

Example 2.2.10. Let Q 2 be the standard 2 simplex in R 2 . The only possible Z 2 -<br />

characteristic vectors are described by the Figure 2.1. The product Z 2 2 × Q2 is 4 copies<br />

<strong>of</strong> Q 2 . Identifying the faces <strong>of</strong> Z 2 2 × Q2 according to the equivalence relation ∼ z we can<br />

show that the small cover N(Q 2 , β) is the real projective space RP 2 . By theorem 2.2.8,


Chapter 2: Small covers and orbifolds 36<br />

(1, 0)<br />

β 1<br />

β 3<br />

P<br />

(1, 1)<br />

β 2<br />

(0, 1)<br />

Figure 2.1: The Z 2 -characteristic function corresponding to a triangle.<br />

we can show that the small cover N(△ n , β) corresponding to the Z 2 -characteristic pair<br />

(△ n , β) is equivariantly homeomorphic to RP n .<br />

Remark 2.2.11. Suppose λ is the characteristic function <strong>of</strong> a 2n-dimensional quasi<strong>toric</strong><br />

manifold N = N(Q, λ). Consider the involution ¯τ on T n ×Q defined by (t, p) → (t −1 , p).<br />

The fixed point set <strong>of</strong> ¯τ is Z n 2 × Q. The involution ¯τ descends to the involution τ on<br />

N(Q, λ) with the fixed point set homeomorphic to N(Q, β), where β : F(Q) → F n 2 is<br />

the mod 2 reduction <strong>of</strong> λ : F(Q) → Z n . For example, the fixed point set <strong>of</strong> the complex<br />

conjugation on the complex projective space CP n is the real projective space RP n .<br />

Following [DJ91], we give some examples <strong>of</strong> simple polytopes on which there exist<br />

no Z 2 -characteristic function.<br />

Example 2.2.12. For each integer n and k ≥ (n + 1), cyclic polytope is defined as<br />

the convex hull <strong>of</strong> k distinct points on the moment curve ℘ : R → R n<br />

defined by<br />

℘(t) = (t, t 2 , . . . , t n ) ∈ R n . We denote this cyclic polytope by Ck n . The Vandermonde<br />

determinant identity gives that no (n + 1) vertices <strong>of</strong> Ck n lie on a common affine hyperplane.<br />

Hence Ck n is a simplicial n-polytope with k vertices.<br />

Let n ≥ 4 and k ≥ 2 n . Let Q n k be the dual polytope <strong>of</strong> Cn k<br />

. Since n ≥ 4, the 1-skeleton<br />

<strong>of</strong> C n k is a complete graph. So for any two facets F i, F j ∈ F(Q n k ), the intersection F i ∩F j<br />

is a nonempty face <strong>of</strong> Q n k . Suppose there exist a Z 2-characteristic function<br />

β : F(Q n k ) → Fn 2 .<br />

So β(F i ) and β(F j ) are distinct nonzero vectors in F n k<br />

. This contradicts to the definition<br />

<strong>of</strong> Z 2 -characteristic function. Hence, there can be no such function β when k > 2 n .<br />

Therefore there does not exists a Z 2 -characteristic function on the set <strong>of</strong> facets <strong>of</strong> Q n k .<br />

Hence by remark 2.2.11, when n ≥ 4 and k ≥ 2 n the polytope Q n k<br />

cannot be the orbit<br />

space <strong>of</strong> a quasi<strong>toric</strong> manifold.


37 2.3 Classical effective orbifolds<br />

2.3 Classical effective orbifolds<br />

Following [ALR07] we give the definition <strong>of</strong> classical effective orbifolds.<br />

Hausdorff topological space.<br />

Let Y be a<br />

Definition 2.3.1. An n-dimensional orbifold chart on an open subset V ⊆ Y is given<br />

by a triple (Ṽ , H, ζ) where<br />

1. Ṽ is a connected open subset <strong>of</strong> R n ,<br />

2. H is a finite subgroup <strong>of</strong> smooth automorphisms <strong>of</strong> Ṽ ,<br />

3. ζ is a map from<br />

Ṽ to Y such that ζ is H-invariant map and induces a homeomorphism<br />

from Ṽ /H onto V .<br />

Definition 2.3.2. An embedding ξ : (Ṽ , H, ζ) → (Ũ, G, ϕ) between two orbifold charts<br />

is a smooth embedding ξ : Ṽ → Ũ <strong>of</strong> manifolds with ζ ◦ ϕ = ξ.<br />

Definition 2.3.3. Two orbifold charts (Ṽ , H, ζ) on V = ζ(Ṽ ) ⊆ Y and (Ũ, G, ϕ) on<br />

U = ϕ(Ũ) ⊆ Y with a point x ∈ V ∩ U are locally compatible if there exists an open<br />

neighborhood W ⊆ V ∩ U <strong>of</strong> x and an orbifold chart (˜W , K, µ) on W such that there<br />

are smooth embeddings (˜W , K, µ) → (Ṽ , H, ζ) and (˜W , K, µ) → (Ũ, G, ϕ).<br />

Definition 2.3.4. A smooth orbifold atlas on Y is a family V = {(Ṽ , H, ζ)} <strong>of</strong> locally<br />

compatible orbifold charts such that {ϕ(Ṽ ) : Ṽ ∈ V} is an open cover <strong>of</strong> Y .<br />

Definition 2.3.5. An atlas V is a refinement <strong>of</strong> an atlas U if for any chart (Ṽ , H, ζ) ∈ V<br />

there exists an embedding ξ : (Ṽ , H, ζ) → (Ũ, G, ϕ) into some chart (Ũ, G, ϕ) ∈ U.<br />

Two orbifold atlases are said to be equivalent if they have a common refinement.<br />

Denote the equivalence class <strong>of</strong> an atlas V by [V].<br />

Definition 2.3.6. Let Y be a paracompact Hausdorff space equipped with an equivalence<br />

class [V] <strong>of</strong> n-dimensional smooth orbifold atlases. The pair (Y, V), denoted by Y, is<br />

called an effective smooth orbifold <strong>of</strong> dimension n.<br />

Throughout this section we assume that all orbifolds are effective. We enlist some<br />

observations about the definition.<br />

Observation 2.3.7.<br />

1. For each orbifold chart (Ṽ , H, ζ) the group H is acting<br />

smoothly and effectively on Ṽ . In particular H acts freely on a dense open subset<br />

<strong>of</strong> Ṽ . With this property Y is called a reduced orbifold.<br />

2. A linear chart is the triple (R n , H, ζ), where H is a finite subgroup <strong>of</strong> O(n). The<br />

group H acts on R n via an orthogonal representation. Since smooth actions are<br />

locally smooth, any orbifold has an atlas consisting <strong>of</strong> linear charts.


Chapter 2: Small covers and orbifolds 38<br />

3. Given two embeddings <strong>of</strong> orbifold charts ξ 1 , ξ 2 : (Ṽ , H, ζ) ↩→ (Ũ, G, ϕ), there exists<br />

a unique g ∈ G such that ξ 2 = g ◦ ξ 1 . The pro<strong>of</strong> follows from Lemma 2.11<br />

in [MM03].<br />

As a consequence, an embedding <strong>of</strong> orbifold charts ξ : (Ṽ , H, ζ) ↩→ (Ũ, G, ϕ) induces<br />

a monomorphism ξ HG : H → G <strong>of</strong> groups.<br />

4. Given an embedding ξ : (Ṽ , H, ζ) ↩→ (Ũ, G, ϕ), if there exists g ∈ G such that<br />

ξ(Ṽ ) ∩ g ◦ ξ(Ṽ ) ≠ φ then g ∈ Im(ξ HG) and so ξ(Ṽ ) = g ◦ ξ(Ṽ ).<br />

5. If (Ṽ , H, ζ) and (Ũ, G, ϕ) are two charts for the same orbifold structure on V ⊆ Y<br />

and if<br />

Ṽ is simply connected, then there exists an embedding ξ : (Ṽ , H, ζ) ↩→<br />

(Ũ, G, ϕ) whenever ζ(Ṽ ) ⊂ ϕ(Ũ) ⊆ V .<br />

6. Every orbifold atlas for Y is contained in a unique maximal atlas.<br />

7. If the finite group actions on all the charts are free, then Y is locally Euclidean,<br />

hence a manifold.<br />

Definition 2.3.8. Let Y = (Y, V) be an orbifold and y ∈ Y . Let (Ṽ , H, ζ) be an orbifold<br />

chart so that y = ϕ(x) ∈ ϕ(Ṽ ) ⊂ Y . The local group at y is defined by the group<br />

H y = {h ∈ H : h · x = x}.<br />

The group H y is uniquely determined up to conjugacy. We use the notion <strong>of</strong> local<br />

group to define the singular set <strong>of</strong> the orbifold Y in the following definition.<br />

Definition 2.3.9. A point y ∈ Y is called a smooth point if the group H y is trivial,<br />

otherwise y is called singular point. The set <strong>of</strong> singular points <strong>of</strong> an orbifold Y = (Y, V)<br />

is called its singular set, denoted by ΣY. That is,<br />

ΣY = {y ∈ Y : H y ≠ 1}.<br />

Definition 2.3.10. Let H × M → M be an smooth and effective action <strong>of</strong> a finite<br />

group H on a smooth manifold M. The associated orbifold Y = (M/H, V) is called an<br />

effective global quotient, where V is constructed from a manifold atlas using the locally<br />

smooth structure.<br />

Example 2.3.11. Let H be a finite subgroup <strong>of</strong> GL n (C) and let Y = C n /H. This<br />

is a singular complex manifold called a quotient singularity. Y has the structure <strong>of</strong><br />

an algebraic variety, arising from the algebra <strong>of</strong> H-invariant polynomials on C n . If<br />

H ⊂ SL n (C), the quotient C n /H is called Gorenstein.


39 2.4 Tangent bundle <strong>of</strong> orbifolds<br />

Example 2.3.12. Consider<br />

S 2n+1 = {(z 0 , . . . , z n ) ∈ C n+1 : Σ i |z i | 2 = 1}, (2.3.1)<br />

the circle S 1 ∋ a act on S 2n+1 by<br />

a(z 0 , . . . , z n ) = (a a 0<br />

z 0 , . . . , a an z n ),<br />

where the a i ’s are coprime integers. The quotient space<br />

W P(a 0 , . . . , a n ) = S 2n+1 /S 1<br />

has an orbifold structure, denoted by WP(a 0 , . . . , a n ). This orbifold is called a weighted<br />

projective space. The orbifold WP(1, a) is the famous teardrop. It is well known that<br />

these orbifolds are non-global quotient orbifold. We will show in Chapter 3 that teardrops<br />

are quasi<strong>toric</strong> orbifolds.<br />

Example 2.3.13. Orbifold Riemann surfaces are generalization from the teardrop.<br />

These are a fundamental class <strong>of</strong> examples in orbifold theory. We need to specify the<br />

isolated singular points and the order <strong>of</strong> the local group at each one. Let y i is a singular<br />

point with order n i . Then the local chart at y i is (D 2 , Z ni , ζ yi ) where D 2 is an open ball<br />

centered at origin and ζ yi<br />

Z ni .<br />

is the orbit map <strong>of</strong> the action e · z = ez for a generator e <strong>of</strong><br />

Let Σ be an orbifold Riemann surface <strong>of</strong> genus g and k singular points. Thurston<br />

[Thu3m] has shown that Σ is a global quotient if and only if g + 2k ≥ 3 or g = 0 and<br />

k = 2 with n 1 = n 2 . An orbifold Riemann surface can be expressed as a orbit space<br />

M 3 /T 1 for some 3-dimensional Seifert fiber manifold M 3 with an effective action <strong>of</strong> T 1 .<br />

Next we define the notion <strong>of</strong> smooth maps between orbifolds.<br />

Definition 2.3.14. Let Y = (Y, V) and W = (W, U) be two orbifolds. A map f : Y →<br />

W is called smooth if for any point y ∈ Y there are charts (Ṽ , H, ζ) containing y and<br />

(Ũ, G, ϕ) containing f(y), such that f maps V = ζ(Ṽ ) into U = ϕ(Ũ) and f can be<br />

lifted to a smooth map ˜f : Ṽ → Ũ with ϕ ◦ ˜f = f ◦ ζ.<br />

Using this we can define the notion <strong>of</strong> diffeomorphism <strong>of</strong> orbifolds.<br />

Definition 2.3.15. Two orbifolds Y and W are diffeomorphic if there are smooth maps<br />

<strong>of</strong> orbifolds f : Y → W and g : W → Y with f ◦ g = 1 W and g ◦ f = 1 Y .<br />

2.4 Tangent bundle <strong>of</strong> orbifolds<br />

In this section we define the tangent <strong>of</strong> an orbifold and related notions following [ALR07].<br />

We show the identifications <strong>of</strong> orbifold charts to yield the original orbifold.


Chapter 2: Small covers and orbifolds 40<br />

Let Y = (Y, V) be an orbifold. Let (Ṽ , H, ζ) and (Ũ, G, ϕ) be two orbifold charts<br />

with y ∈ ζ(Ṽ ) ∩ ϕ(Ũ). So by definition there is a chart (˜W , K, µ) and embeddings<br />

ξ 1 : (˜W , K, µ) ↩→ (Ṽ , H, ζ) and ξ 2 : (˜W , K, µ) ↩→ (Ũ, G, ϕ).<br />

These two embeddings give rise the following equivariant diffeomorphisms<br />

ξ −1<br />

1 : ξ 1 (˜W ) → ˜W and ξ 2 : ˜W → ξ 2 (˜W ). (2.4.1)<br />

between K-spaces. Here K acts on ξ 1 (˜W ) and ξ 2 (˜W ) via the subgroups ξ 1KG (K) ⊆ G<br />

and ξ 2KH (K) ⊆ H respectively.<br />

diffeomorphism<br />

The composition <strong>of</strong> these maps give an equivariant<br />

ξ 2 ◦ ξ −1<br />

1 : ξ 1 (˜W ) → ξ 2 (˜W ). (2.4.2)<br />

between K-spaces. Hence we can glue the sets Ṽ /H and Ũ/G according to the equivalence<br />

relation ∼ t defined by<br />

ζ(ṽ) ∼ t ϕ(ũ) if ξ 2 ◦ ξ1 −1 (ṽ) = ũ. (2.4.3)<br />

Let<br />

Ŷ = ⊔ Ṽ ∈V(Ṽ /H)/ ∼ t (2.4.4)<br />

be the space obtained by gluing the sets {Ṽ /H : (Ṽ , H, ζ) ∈ V}. So we get a homeomorphism<br />

Φ : Ŷ → Y induced by the maps ζ : Ṽ → Y .<br />

We may consider the function ξ 2 ◦ ξ1 −1 as a transition function. Suppose there exist<br />

another two embeddings<br />

ξ 1 ′ : (˜W , K, µ) ↩→ (Ṽ , H, ζ), ξ′ 2 : (˜W , K, µ) ↩→ (Ũ, G, ϕ).<br />

We have observed in 3 <strong>of</strong> 2.3.7 that there exist unique h ∈ H and g ∈ G such that<br />

ξ ′ 1 = h ◦ ξ 1 and ξ ′ 2 = g ◦ ξ 2 .<br />

Hence g ◦ (ξ 2 ◦ ξ −1<br />

1 ) ◦ h−1 is the resulting transition function.<br />

Using this explicit computations, we construct the tangent bundle <strong>of</strong> the orbifold Y .<br />

Given an orbifold chart (Ṽ , H, ζ), we consider the tangent bundle T Ṽ <strong>of</strong> the manifold<br />

Ṽ . By the observation 1 <strong>of</strong> 2.3.7 the group H acts smoothly on Ṽ . Hence H also acts<br />

smoothly on T Ṽ . Suppose (ṽ, α) is an element <strong>of</strong> T Ṽ , then h ∈ H acts by<br />

h · (ṽ, α) = (h · ṽ, Dhṽ(α)). (2.4.5)


41 2.4 Tangent bundle <strong>of</strong> orbifolds<br />

Let T V = T Ṽ /H and ζ T V : T Ṽ → T V be the orbit map. So the triple (T Ṽ , H, ζ T U)<br />

is an orbifold chart on T U. The projection map T Ṽ → Ṽ is an equivariant map. This<br />

map induces a natural projection p V : T V → V . We describe the fibers <strong>of</strong> the map p V .<br />

If ỹ ∈ Ṽ and y = ζ(ỹ) ∈ V , then<br />

p −1<br />

V<br />

(y) = {H(z, v) : z = ỹ} ⊂ T V. (2.4.6)<br />

We claim that p −1<br />

V<br />

(y) is homeomorphic to the orbit space TỹŨ/H y <strong>of</strong> H y action on<br />

TỹŨ, where H y denotes the isotropy group <strong>of</strong> the H-action on Ṽ at ỹ. Define the map<br />

Observe that<br />

f y : p −1<br />

V (y) → T ỹṼ /H y by f y (H(ỹ, α)) = H y α. (2.4.7)<br />

H(ỹ, α 1 ) = H(ỹ, α 2 ) if and only if there exists a h ∈ H such that h · (ỹ, α 1 ) = (ỹ, α 2 ).<br />

Again<br />

h · (ỹ, α 1 ) = (ỹ, α 2 ) if and only if h ∈ H y and Dỹh(α 1 ) = α 2 .<br />

This is equivalent to the assertion that H y α 1 = H y α 2 . So f y is both well defined and<br />

injective. From the following commutative diagram it is clear that f y is a surjection<br />

and continuous map.<br />

T<br />

⏐ Ṽ −−−−→ Ṽ<br />

ζ ⏐↓ ⏐<br />

T V ζ↓<br />

T V<br />

(2.4.8)<br />

p<br />

−−−−→<br />

V<br />

V<br />

Hence we established our claim. The fiber p −1<br />

V (y) is a quotient <strong>of</strong> the form Rn /H 0 ,<br />

where H 0 ⊂ GL n (R) is a finite group. So the fiber p −1 (y) may not be a vector space.<br />

Hence we have constructed a bundle-like object T V over V .<br />

V<br />

Definition 2.4.1. The map p V : T V → V is called an orbifold tangent bundle associated<br />

to the orbifold chart (Ṽ , H, ζ).<br />

It may be clear how to construct the tangent bundle on an orbifold Y = (Y, Ṽ ). Let<br />

T V = {(T Ṽ , H, ζ T V ) : (Ṽ , H, ζ) ∈ V}.<br />

We need to identify the orbifold tangent bundles T V → V associated to the charts<br />

(V, H, ζ). We observe that the gluing maps ξ 12 = ξ 2 ◦ ξ −1<br />

1 for orbifold charts <strong>of</strong> V in<br />

equation 2.4.2 are smooth. We may use the equivariant differential<br />

Dξ 12 : T ξ 1 (˜W ) → T ξ 2 (˜W ) (2.4.9)


Chapter 2: Small covers and orbifolds 42<br />

as the transition functions to identify the bundles T V → V and T U → U. Let T Y be<br />

the resulting identification space<br />

( ⊔ Ṽ ∈V<br />

T V )/ ∼ T . (2.4.10)<br />

We consider the minimal <strong>topology</strong> on T Y such that each inclusion T V → T Y is homeomorphic<br />

onto an open subset <strong>of</strong> T Y . With this <strong>topology</strong>, T Y has an orbifold structure.<br />

T V is a family <strong>of</strong> locally compatible orbifold charts such that the collection<br />

{ζ T V (T Ṽ ) : (Ṽ , H, ζ) ∈ V} is an open cover <strong>of</strong> T Y . So the family T U is an orbifold atlas<br />

on T Y . By the identification ∼ T , there exists a continuous surjection p Y : T Y → Y<br />

such that p Y | T V = p V . Let T Y = (T Y, T V).<br />

Definition 2.4.2. The triple (T Y, p Y , Y) is called the orbifold tangent bundle <strong>of</strong> the<br />

orbifold Y.<br />

We summarize the above computations in the next proposition.<br />

Proposition 2.4.3 ( [ALR07], Proposition 1.21). The tangent bundle <strong>of</strong> an n-<br />

dimensional orbifold Y has the structure <strong>of</strong> a 2n-dimensional orbifold. Also the map p Y<br />

is a smooth map <strong>of</strong> orbifolds with fibers TỹṼ /H y .<br />

Now we define the frame bundle <strong>of</strong> an orbifold Y. Note that for each local chart<br />

(Ṽ , H, ζ) we can define an H-invariant inner product on T Ṽ . Let O(T ỹṼ ) be the<br />

orthogonal transformations <strong>of</strong> TỹṼ . We can construct the frame manifold<br />

F r(Ṽ ) = {(ỹ, A) : A ∈ O(T ỹṼ )} (2.4.11)<br />

and the induced left H-action on F r(Ṽ ) is given by<br />

h · (ỹ, A) = (hỹ, DhỹA). (2.4.12)<br />

Since the H-action on<br />

Ṽ is an effective action, the H-action on the frame manifold<br />

F r(Ṽ ) is free. So the quotient space F r(Ṽ )/H is a smooth manifold. Denote the orbit<br />

<strong>of</strong> (ỹ, A) by [ỹ, A]. There is a right O(n) action on F r(Ṽ )/H induced from the natural<br />

translation action on F r(Ṽ ), given by<br />

[ỹ, A] · B = [ỹ, AB] for B ∈ O(n). (2.4.13)<br />

Observe that this action is transitive on the fibers. Indeed, [ỹ, A] = [ỹ, I] · A. The<br />

isotropy subgroup for this orbit consists <strong>of</strong> those orthogonal matrices A such that<br />

(ỹ, A) = (hỹ, DhỹI)


43 2.4 Tangent bundle <strong>of</strong> orbifolds<br />

for some h ∈ H. This is equivalent to say that h ∈ H y and A = Dhỹ. So the differential<br />

establishes an injection H y → O(TỹṼ ). Hence we conclude that H y is the isotropy<br />

subgroup <strong>of</strong> the O(n)-action on F r(Ṽ )/H. The fiber is the associated homogeneous<br />

space O(n)/H y .<br />

Consider the orbit space <strong>of</strong> the action 2.4.13 on F r(Ṽ )/H. Clearly this orbit space<br />

is homeomorphic to V . So we obtain the natural projection F r(Ṽ )/H → V . Let F r(Y)<br />

be the space obtained by identifying the local charts F r(Ṽ )/H → UV using the O(n)-<br />

transition functions obtained from the tangent bundle <strong>of</strong> Y. Let p F r : F r(Y) → Y be<br />

the induced continuous surjection.<br />

Definition 2.4.4. The triple (F r(Y), p F r , Y) is called the frame bundle <strong>of</strong> an orbifold<br />

Y = (Y, V).<br />

This frame bundle has some useful properties, which we summarize below.<br />

Theorem 2.4.5 ( [ALR07], Theorem 1.23). For a given orbifold Y, its frame bundle<br />

F r(Y) is a smooth manifold with a smooth, effective and almost free O(n)-action. The<br />

orbifold Y is naturally isomorphic to the resulting quotient orbifold F r(Y)/O(n).<br />

The following is a very important consequence <strong>of</strong> the theorem 2.4.5.<br />

Corollary 2.4.6 ( [ALR07], Corollary 1.24). Every smooth effective n-dimensional<br />

orbifold Y is diffeomorphic to a quotient orbifold for a smooth, effective and almost free<br />

O(n)-action on a smooth manifold M.<br />

Definition 2.4.7. Let Y = (Y, V) denote an orbifold with tangent bundle (T Y, p Y , Y).<br />

1. A non-degenerate symmetric 2-tensor <strong>of</strong> S 2 (T Y) is called a Riemannian metric<br />

on Y.<br />

2. An almost complex structure on Y is an endomorphism J : T Y → T Y such that<br />

J 2 = −Id.<br />

3. A stable almost complex structure on Y is an endomorphism<br />

J : T Y ⊕ (Y × R k ) → T Y ⊕ (Y × R k )<br />

such that J 2 = −Id for some positive integer k.<br />

4. We call Y a complex orbifold if all the defining maps are holomorphic.<br />

Using the frame bundle <strong>of</strong> an orbifold, we see that techniques applicable to quotient<br />

spaces <strong>of</strong> almost free smooth action <strong>of</strong> a compact Lie group will yield results about<br />

orbifolds. For example, we have the following proposition. The pro<strong>of</strong> <strong>of</strong> this proposition<br />

can be found in [AP93].


Chapter 2: Small covers and orbifolds 44<br />

Proposition 2.4.8 ( [ALR07], Proposition 1.28). If a compact, connected Lie group<br />

G acts smoothly and almost freely on an orientable, connected, compact manifold M,<br />

then H ∗ (M/G; Q) is a Poincaré duality algebra. Hence, if Y is a compact, connected,<br />

orientable orbifold, then H ∗ (Y ; Q) will satisfy Poincaré duality.<br />

2.5 Orbifold fundamental group<br />

The goal <strong>of</strong> this section is to provide an idea how one can compute the orbifold fundamental<br />

group <strong>of</strong> an effective smooth orbifold.<br />

A covering orbifold or orbifold cover <strong>of</strong> an n-dimensional orbifold Y is a smooth map<br />

<strong>of</strong> orbifolds p : X → Y whose associated continuous map (also denoted by p) p : X → Y<br />

between underlying spaces satisfies the following condition:<br />

Each point y ∈ Y has a neighborhood V ∼ = Ṽ /H with Ṽ homeomorphic to a<br />

connected open set in R n , for which each component U i <strong>of</strong> p −1 (V ) is homeomorphic<br />

to Ṽ /H i for some subgroup H i ⊂ H such that the natural map p i : Ṽ /H i → Ṽ /H<br />

corresponds to the restriction <strong>of</strong> p on U i .<br />

Definition 2.5.1. Given an orbifold cover p : X → Y, a diffeomorphism h : X → X is<br />

called a deck transformation if p ◦ h = p.<br />

Definition 2.5.2. An orbifold cover p : X → Y is called a universal orbifold cover <strong>of</strong><br />

Y if given any orbifold cover p 1 : W → Y, there exists an orbifold cover p 2 : X → W<br />

such that p = p 1 ◦ p 2 .<br />

Every orbifold has a universal orbifold cover which is unique up to diffeomorphism,<br />

see [Thu3m].<br />

The corresponding group <strong>of</strong> deck transformations is called the<br />

orbifold fundamental group <strong>of</strong> Y and denoted by π orb<br />

1 (Y).<br />

Example 2.5.3 (Hurwitz cover). Suppose that p : Σ 1 → Σ 2 is a holomorphic map<br />

between two compact orbifold Riemann surfaces Σ 1 , Σ 2 . Usually, p is not a covering<br />

map. Instead, it ramifies in finitely many points z 1 , . . . , z k ∈ Σ 2 . Hence the restriction<br />

map, namely,<br />

p : Σ 1 − {∪ i p −1 (z i )} → Σ 2 − {z 1 , . . . , z k } (2.5.1)<br />

is a manifold covering map. Suppose that the preimage <strong>of</strong> z i is y i1 , . . . , y iji . Let n il be<br />

the ramification order at y il . That is, under suitable coordinate system near y il , the<br />

map p can be written as z → z n i l .<br />

We assign an orbifold structure on Σ 1 and Σ 2 as follows. We first assign an orbifold<br />

structure at y ip with order n ip . Let n i be the largest common factor <strong>of</strong> the n ip ’s. Then<br />

we assign an orbifold structure at z i with order n i . One readily verifies that under these<br />

assignments, p : Σ 1 → Σ 2 becomes an orbifold cover. The map p : Σ 1 → Σ 2 is referred<br />

to as a Hurwitz cover.


45 2.5 Orbifold fundamental group<br />

Example 2.5.4. If Y = M/H is a global quotient and ˜M → M is a universal cover,<br />

then ˜M → M → Y is the orbifold universal cover <strong>of</strong> Y. This gives an extension <strong>of</strong><br />

groups<br />

1 → π 1 (M) → π orb<br />

1 (Y) → H → 1. (2.5.2)<br />

This implies that an orbifold Y can not be a global quotient if π orb<br />

1 (Y) is trivial, unless<br />

Y is itself a manifold.<br />

Definition 2.5.5. An orbifold is a good orbifold if its orbifold universal cover is smooth<br />

manifold.<br />

Observation 2.5.6. The following observation is very useful in computations <strong>of</strong> orbifold<br />

fundamental groups. Suppose that p : X → Y is an orbifold universal cover. Then the<br />

restriction<br />

p : X − p −1 (ΣX ) → Y − ΣY (2.5.3)<br />

is a manifold cover. The covering group H <strong>of</strong> this cover is the orbifold fundamental<br />

group π1 orb (Y). The sets ΣX and ΣY are the singular points <strong>of</strong> X and Y respectively.<br />

Therefore, Y = X /H and there is a surjective homomorphism<br />

q p : π 1 (Y − ΣY) → H. (2.5.4)<br />

In general, there is no reason to expect that q p will be an isomorphism. However, to<br />

compute the group π orb<br />

1 (Y), we can start with the group π 1(Y − ΣY) and then specify<br />

the additional relations if required.<br />

Example 2.5.7. Consider the orbifold Riemann surface Σ g <strong>of</strong> genus g with k orbifold<br />

points O k = {x 1 , . . . , x k } <strong>of</strong> orders n 1 , . . . , n k . Then, according to [Sco83], (p. 424) a<br />

presentation for its orbifold fundamental group is given by<br />

π orb<br />

1 (Σ g ) = {α 1 , β 1 , . . . , α k , β k , σ 1 , . . . , σ k : σ 1 . . . σ k Π g 1 [α i, β i ] = 1, σ n i<br />

i<br />

= 1}, (2.5.5)<br />

where α i and β i are the generators <strong>of</strong> π 1 (Σ g ) and σ i are the generators <strong>of</strong> Σ g − O k<br />

represented by a loop around each orbifold point. Note that π orb<br />

1 (Σ g) is obtained from<br />

π 1 (Σ g − O k ) by introducing the relations σ m i<br />

= 1.<br />

Consider the special case when Σ = ˜Σ/H, where H is a finite group <strong>of</strong> automorphisms<br />

<strong>of</strong> ˜Σ. In this case, the orbifold fundamental group is isomorphic to π 1 (EH× H ˜Σ),<br />

which in turn fits into a group extension<br />

1 → π 1 (˜Σ) → π orb<br />

1 (Σ) → H → 1. (2.5.6)<br />

In other words, the orbifold fundamental group is a virtual surface group. This will be<br />

true for any good orbifold Riemann surface.


Chapter 2: Small covers and orbifolds 46


Chapter 3<br />

Quasi<strong>toric</strong> orbifolds<br />

In this chapter we study structures and invariants <strong>of</strong> quasi<strong>toric</strong> orbifolds. In particular,<br />

we discuss the constructive and axiomatic definitions <strong>of</strong> quasi<strong>toric</strong> orbifolds. We<br />

compute the orbifold fundamental group <strong>of</strong> these orbifolds. We determine whether any<br />

quasi<strong>toric</strong> orbifold can be the quotient <strong>of</strong> a smooth manifold by a finite group action<br />

or not. To calculate the rational homology groups <strong>of</strong> quasi<strong>toric</strong> orbifolds we need to<br />

generalize the usual CW -complex little bit. The cohomology ring <strong>of</strong> a quasi<strong>toric</strong> orbifold<br />

with coefficient in Q is computed in this chapter. We prove existence <strong>of</strong> stable<br />

almost complex structure and describe the Chen-Ruan cohomology groups <strong>of</strong> an almost<br />

complex quasi<strong>toric</strong> orbifold.<br />

3.1 Definition by construction and orbifold structure<br />

For any Z-module N denote N ⊗ Z R by N R . Let N be a free Z-module <strong>of</strong> rank n.<br />

The quotient T N = N R /N is a compact n-dimensional torus. Suppose N ′ is a free<br />

submodule <strong>of</strong> N <strong>of</strong> rank n ′ . Let T N ′ denote the torus N<br />

R ′ /N ′ . Let j : N<br />

R ′ → N R and<br />

j ∗ : T N ′ → N R /N ′ be the natural inclusions. The inclusion i : N ′ → N induces a<br />

homomorphism<br />

i ∗ : N R /N ′ → N R /N = T N defined by i ∗ (a + N ′ ) = a + N<br />

on cosets. Denote the composition i ∗ ◦ j ∗ : T N ′ → T N by ξ N ′ and also denote the<br />

image <strong>of</strong> ξ N ′ by Im(ξ N ′). Ker(i ∗ ) ≃ N/N ′ . If n ′ = n, then j ∗ is identity and i ∗ is<br />

a surjection. In this case ξ N ′ : T N ′ → T N is a surjection group homomorphism with<br />

kernel G N ′ = N/N ′ , a finite abelian group.<br />

A 2n-dimensional quasi<strong>toric</strong> orbifold may be constructed from the following data: a<br />

simple polytope Q <strong>of</strong> dimension n with set <strong>of</strong> facets F(Q) = {F i : i ∈ {1, . . . , m} = I},<br />

a free Z-module N <strong>of</strong> rank n and a dicharacteristic function, defined below.<br />

47


Chapter 3: Quasi<strong>toric</strong> orbifolds 48<br />

Definition 3.1.1. Let there exists an assignment <strong>of</strong> a vector λ i in N to each facet F i<br />

<strong>of</strong> Q such that whenever F i1 ∩ . . . ∩ F ik ≠ ∅ the corresponding vectors λ i1 , . . . , λ ik are<br />

linearly independent over Z. The function λ : F(Q) → N defined by λ(F i ) = λ i is called<br />

a dicharacteristic function <strong>of</strong> Q.<br />

These data will be referred to as a combinatorial model and abbreviated as<br />

(Q, N, {λ i }). The vector λ i is called the dicharacteristic vector corresponding to the<br />

i-th facet.<br />

Example 3.1.2. The quasi<strong>toric</strong> orbifolds associated to the first and second combinatorial<br />

model has 1 and 3 singular points respectively.<br />

(−3, −5)<br />

(1, 1)<br />

(2, 1)<br />

(0, 1)<br />

(1, 4)<br />

(1, 2)<br />

(1, 0)<br />

Figure 3.1: <strong>Some</strong> dicharacteristic function <strong>of</strong> triangle and receptacle.<br />

We give the constructive definition <strong>of</strong> quasi<strong>toric</strong> orbifolds below. Each face F <strong>of</strong><br />

Q <strong>of</strong> codimension k ≥ 1 is the intersection <strong>of</strong> a unique set <strong>of</strong> k facets F i1 , . . . , F ik .<br />

Let I(F ) = {i 1 , . . . , i k } ⊂ I. Let N(F ) denote the submodule <strong>of</strong> N generated by the<br />

characteristic vectors {λ j : j ∈ I(F )}. So T N(F ) = N(F ) R /N(F ) is a torus <strong>of</strong> dimension<br />

k. We will adopt the convention that T N(Q) = 1.<br />

Define an equivalence relation ∼ on the product Q × T N by<br />

(p, t) ∼ (q, s) if p = q and s −1 t ∈ Im(ξ N(F ) ), (3.1.1)<br />

where F is the unique face whose relative interior contains p. Let X = Q × T N / ∼<br />

be the quotient space. Let q : Q × T N → X denote the quotient map. Then X is a<br />

T N -space and let π : X → Q defined by π([p, t] ∼ ) = p be the associated map to the<br />

orbit space Q. The space X has the structure <strong>of</strong> an orbifold, which we explain next.<br />

Pick open neighborhoods U v <strong>of</strong> the vertices v <strong>of</strong> Q such that U v is the complement<br />

in Q <strong>of</strong> all facets that do not contain v. Let<br />

X v = π −1 (U v ) = U v × T N / ∼ .<br />

For a face F <strong>of</strong> Q containing v the inclusion {λ i : i ∈ I(F )} in {λ i : i ∈ I(v)} induces an<br />

inclusion <strong>of</strong> N(F ) in N(v) whose image will be denoted by N(v, F ). Since {λ i : i ∈ I(F )}<br />

extends to a basis {λ i : i ∈ I(v)} <strong>of</strong> N(v), the natural map from the torus<br />

T N(v,F ) = N(v, F ) R /N(v, F ) to T N(v) = N(v) R /N(v)


49 3.1 Definition by construction and orbifold structure<br />

defined by a + N(v, F ) ↦→ a + N(v) is an injection. We will identify its image with<br />

T N(v,F ) . Denote the canonical isomorphism T N(F ) → T N(v,F ) by i(v, F ).<br />

Define an equivalence relation ∼ v on U v × T N(v) by<br />

(p, t) ∼ v (q, s) if p = q and s −1 t ∈ T N(v,F )<br />

where F is the face whose relative interior contains p. Then W v = U v × T N(v) / ∼ v<br />

is θ-equivariantly diffeomorphic to an open ball in C n where θ : T N(v) → U(1) n is an<br />

isomorphism, see [DJ91]. Note that the map ξ N(F ) factors as ξ N(F ) = ξ N(v) ◦ i(v, F ).<br />

Since i(v, F ) is an isomorphism, t ∈ T N(v,F ) if and only if ξ N(v) (t) ∈ Imξ N(F ) . Hence<br />

the map ξ N(v) : T N(v) → T N induces a map<br />

ξ v : W v → X v defined by ξ v ([(p, t)] ∼ v<br />

) = [(p, ξ N(v) (t))] ∼<br />

on equivalence classes.The group G v = N/N(v), the kernel <strong>of</strong> ξ N(v) , is a finite subgroup<br />

<strong>of</strong> T N(v) and therefore has a natural smooth, free action on T N(v) induced by the group<br />

operation. This induces smooth action <strong>of</strong> G v on W v . This action is not free in general.<br />

Since T N<br />

∼ = TN(v) /G v , X v is homeomorphic to the quotient space W v /G v . So<br />

(W v , G v , ξ v ) is an orbifold chart on X v . To show the compatibility <strong>of</strong> these charts as v<br />

varies, we introduce some additional charts.<br />

For any proper face E <strong>of</strong> dimension k ≥ 1 define U E = ∩ U v , where the intersection<br />

is over all vertices v that belong to E. Let X E = π −1 (U E ). For a face F containing E<br />

there is an injective homomorphism T N(F ) → T N(E) whose image we denote by T N(E,F ) .<br />

Let<br />

N ∗ (E) = (N(E) ⊗ Z Q) ∩ N and G E = N ∗ (E)/N(E). (3.1.2)<br />

G E is a finite group. Let ξ ∗,E : T N(E) → T N ∗ (E) be the natural homomorphism. The<br />

map ξ ∗,E has kernel G E . Denote the quotient N/N ∗ (E) by N ⊥ (E). It is a free Z-<br />

module and N ∼ = N ∗ (E) ⊕ N ⊥ (E).<br />

Fixing a choice <strong>of</strong> this isomorphism (or fixing<br />

an inner product on N) we may regard N ⊥ (E) as a submodule <strong>of</strong> N. Consequently<br />

T N = T N ∗ (E) × T N ⊥ (E).<br />

Define an equivalence relation ∼ E on U E × T N(E) × T N ⊥ (E) by<br />

(p 1 , t 1 , s 1 ) ∼ E (p 2 , t 2 , s 2 ) if p 1 = p 2 , s 1 = s 2 and t −1<br />

2 t 1 ∈ T N(E,F )<br />

where F is the face whose relative interior contains p 1 . Let<br />

W E = U E × T N(E) × T N ⊥ (E)/ ∼ E .<br />

It is diffeomorphic to C n−k × (C ∗ ) k . There is a natural map ξ E : W E → X E induced<br />

by ξ ∗,E : T N(E) → T N ∗ (E) and the identity maps on U E and T N ⊥ (E). The triple


Chapter 3: Quasi<strong>toric</strong> orbifolds 50<br />

(W E , G E , ξ E ) is an orbifold chart on X E .<br />

Given E, fix a vertex v <strong>of</strong> Q contained in E. N(v) = N(E) ⊕ N ′ where N ′ is the<br />

free submodule <strong>of</strong> N(v) generated by the dicharacteristic vectors {λ j : j ∈ I(v)−I(E)}.<br />

Consequently T N(v) = T N(E) × T N ′. We can, without loss <strong>of</strong> generality, assume that<br />

N ′ ⊂ N ⊥ (E). Thus we have a covering homomorphism T N ′ → T N ⊥ (E). For a point<br />

x = [p, t, s] ∈ X E , choose a small neighborhood B <strong>of</strong> s in T N ⊥ (E) such that B lifts to<br />

T N ′. Choose any such lift and denote it by l : B → T N ′. Let<br />

W x = U E × T N(E) × B/ ∼ E .<br />

So (W x , G E , ξ E ) is an orbifold chart on a neighborhood <strong>of</strong> x, and it is induced by<br />

(W E , G E , ξ E ). The natural map W x ↩→ W v induced by the map l and the identification<br />

T N(v) = T N(E) ⊕ T N ′, and the natural injective homomorphism G E ↩→ G v induce an<br />

embedding <strong>of</strong> orbifold charts<br />

(W x , G E , ξ E ) → (W v , G v , ξ v ).<br />

The existence <strong>of</strong> these embeddings shows that the charts {(W v , G v , ξ v ) : v vertex <strong>of</strong> Q}<br />

are compatible and form part <strong>of</strong> a maximal 2n-dimensional orbifold atlas A for X. We<br />

denote the pair {X, A} by X . We say that X is the quasi<strong>toric</strong> orbifold associated to<br />

the combinatorial model (Q, N, {λ i }).<br />

Remark 3.1.3. Note that the orbifold X is a reduced orbifold. Also note that changing<br />

the sign <strong>of</strong> a dicharacteristic vector gives rise to a diffeomorphic orbifold.<br />

In the case <strong>of</strong> a quasi<strong>toric</strong> orbifold X , for any x ∈ X, π(x) belongs to the relative<br />

interior <strong>of</strong> a uniquely determined face E x <strong>of</strong> Q. The isotropy group G x = G E x (see<br />

(3.1.2)). We adopt the convention that G Q = 1.<br />

Definition 3.1.4. A quasi<strong>toric</strong> orbifold is called primitive if all its dicharacteristic<br />

vectors are primitive.<br />

Note that in a primitive quasi<strong>toric</strong> orbifold the local group actions are devoid <strong>of</strong><br />

complex reflections (that is, maps which have one as an eigenvalue with multiplicity<br />

n−1) and the classification theorem <strong>of</strong> [Pri67] for germs <strong>of</strong> complex orbifold singularities<br />

applies.<br />

3.2 Axiomatic definition <strong>of</strong> quasi<strong>toric</strong> orbifolds<br />

Analyzing the structure <strong>of</strong> the quasi<strong>toric</strong> orbifold associated to a combinatorial model,<br />

we make the following axiomatic definition. This is a generalization <strong>of</strong> the axiomatic definition<br />

<strong>of</strong> a quasi<strong>toric</strong> manifold using the notion <strong>of</strong> locally standard action, as mentioned<br />

in the introduction.


51 3.2 Axiomatic definition <strong>of</strong> quasi<strong>toric</strong> orbifolds<br />

Definition 3.2.1. A 2n-dimensional quasi<strong>toric</strong> orbifold Y is an orbifold whose underlying<br />

topological space Y has a T N action, where N is a fixed free Z-module <strong>of</strong> rank<br />

n, such that the orbit space is (diffeomorphic to) a simple n-dimensional polytope Q.<br />

Denote the projection map from Y to Q by π : Y → Q. Furthermore every point x ∈ Y<br />

has<br />

A 1 . a T N -invariant neighborhood V ,<br />

A 2 . an associated free Z-module N ′ <strong>of</strong> rank n with an isomorphism θ : T N ′<br />

→ U(1) n<br />

and an injective module homomorphism ι : N ′ → N which induces a surjective<br />

covering homomorphism ξ N ′ : T N ′ → T N ,<br />

A 3 . an orbifold chart (W, G, ξ) over V where W is θ-equivariantly diffeomorphic to<br />

an open set in C n , G = Kerξ N ′<br />

and ξ : W → V is an equivariant map i.e.<br />

ξ(t · y) = ξ N ′(t) · ξ(y) inducing a homeomorphism between W/G and V .<br />

It is obvious that a quasi<strong>toric</strong> orbifold defined constructively from a combinatorial<br />

model satisfies the axiomatic definition. We now demonstrate that a quasi<strong>toric</strong> orbifold<br />

defined axiomatically is associated to a combinatorial model.<br />

Take any facet F i <strong>of</strong> Q and let Fi<br />

0 be its relative interior. By the characterization <strong>of</strong><br />

local charts in A 3 ), the isotropy group <strong>of</strong> the T N action at any point x in π −1 (Fi 0)<br />

is a<br />

locally constant circle subgroup <strong>of</strong> T N . It is the image under ξ N ′ <strong>of</strong> a circle subgroup <strong>of</strong><br />

T N ′. Thus it determines a locally constant vector, up to choice <strong>of</strong> sign, λ i in N. Since<br />

π −1 (F 0<br />

i ) is connected, we get a characteristic vector λ i, unique up to sign, for each facet<br />

<strong>of</strong> Q. That the characteristic vectors corresponding to all facets <strong>of</strong> Q which meet at<br />

a vertex are linearly independent follows from the fact that their preimages under the<br />

appropriate ι form a basis <strong>of</strong> N ′ . Thus we recover a combinatorial model (Q, N, {λ i })<br />

starting from Y.<br />

Definition 3.2.2. Call the triple (Q, N, {λ i }) a combinatorial model <strong>of</strong> Y.<br />

Remark 3.2.3. Similarly to the quasi<strong>toric</strong> manifolds case we can prove that a quasi<strong>toric</strong><br />

orbifold has a smooth structure. In [GP09], the authors give an explicit smooth orbifold<br />

charts for 4-dimensional quasi<strong>toric</strong> orbifolds.<br />

Lemma 3.2.4. Let X be the quasi<strong>toric</strong> orbifold obtained from the combinatorial model<br />

(Q, N, {λ i }) <strong>of</strong> Y by the construction 3.1. Then X and Y are diffeomorphic orbifolds.<br />

Pro<strong>of</strong>. The hard part is to show the existence <strong>of</strong> T N -equivariant a continuous map from<br />

X → Y . This can be done following Lemma 1.2.15 and Corollary 1.2.17. The idea is to<br />

stratify Y according to normal orbit type, see Davis [Dav78]. Here we need to use the<br />

fact that the orbifold Y being reduced, is the quotient <strong>of</strong> a compact smooth manifold


Chapter 3: Quasi<strong>toric</strong> orbifolds 52<br />

by the foliated action <strong>of</strong> a compact Lie group. Then one can blow up (see [Dav78]) the<br />

singular strata <strong>of</strong> Y to get a manifold Ŷ equivariantly diffeomorphic to T N × Q.<br />

One has to modify the arguments <strong>of</strong> Davis slightly in the orbifold case. The important<br />

thing is that by the differentiable slice theorem each singular stratum has a<br />

neighborhood diffeomorphic to its orbifold normal bundle, and is thus equipped with a<br />

fiberwise linear structure so that the constructions <strong>of</strong> Davis go through. Finally there<br />

is a collapsing map Ŷ → Y and by composition with the above diffeomorphism a map<br />

T N × Q → Y . It is easily checked that this map induces a continuous equivariant map<br />

X → Y .<br />

Definition 3.2.5. Let X 1 and X 2 be quasi<strong>toric</strong> orbifolds whose associated base polytope<br />

Q and free Z-module N are identical. Let θ be an automorphism <strong>of</strong> T N . A map f :<br />

X 1<br />

→ X 2 <strong>of</strong> quasi<strong>toric</strong> orbifolds is called a θ-equivariant diffeomorphism if f is an<br />

diffeomorphism <strong>of</strong> orbifolds and the induced map on underlying spaces f : X 1 → X 2<br />

satisfies f(t · x) = θ(t) · f(x) for all x ∈ X 1 and t ∈ T N .<br />

Definition 3.2.6. Two θ-equivariant diffeomorphisms f, g : X 1 → X 2 are said to be<br />

equivalent if there exists equivariant diffeomorphisms h i : X i → X i , for i = 1, 2, such<br />

that g ◦ h 1 = h 2 ◦ f.<br />

We also define, for θ as above, the θ-translation <strong>of</strong> a combinatorial model (Q, N, {λ i })<br />

to be the combinatorial model (Q, N, {θ ′ (λ i )}), where θ ′ is an automorphism <strong>of</strong> N induced<br />

by θ.<br />

The following lemma classifies quasi<strong>toric</strong> orbifolds over a fixed polytope up to θ-<br />

equivariant diffeomorphism.<br />

Lemma 3.2.7. For any automorphism θ <strong>of</strong> T N , the assignment <strong>of</strong> combinatorial model<br />

defines a bijection between equivalence classes <strong>of</strong> θ-equivariant diffeomorphisms <strong>of</strong> quasi<strong>toric</strong><br />

orbifolds and θ-translations <strong>of</strong> combinatorial models.<br />

Pro<strong>of</strong>. Pro<strong>of</strong> is similar to the Lemma 1.2.19, which we discuss in details. Note that<br />

the existence <strong>of</strong> a preferred section s : Q → Y for an axiomatically defined quasi<strong>toric</strong><br />

orbifold Y follows from the blow up construction in the pro<strong>of</strong> <strong>of</strong> Lemma 3.2.4.<br />

3.3 Characteristic subspaces<br />

Of special importance are certain T N -invariant subspaces <strong>of</strong> X corresponding to the<br />

faces <strong>of</strong> the polytope Q. If F is a face <strong>of</strong> Q <strong>of</strong> codimension k, then define X(F ) :=<br />

π −1 (F ). With subspace <strong>topology</strong>, X(F ) is a quasi<strong>toric</strong> orbifold <strong>of</strong> dimension 2n − 2k.<br />

Recall that<br />

N ∗ (F ) = (N(F ) ⊗ Z Q) ∩ N and N ⊥ (F ) = N/N ∗ (F ).


53 3.4 Orbifold fundamental group<br />

Let ϱ F : N → N ⊥ (F ) be the projection homomorphism. Let J(F ) ⊂ I be the index set<br />

<strong>of</strong> facets <strong>of</strong> Q, other than F in case k = 1, that intersect F . Note that J(F ) indexes the<br />

set <strong>of</strong> facets <strong>of</strong> the n − k dimensional polytope F . The combinatorial model for X(F )<br />

is given by (F, N ⊥ (F ), ϱ F ◦ λ |J(F ) )}).<br />

Definition 3.3.1. X(F ) is called a characteristic subspace <strong>of</strong> X, if F is a facet <strong>of</strong> Q.<br />

Example 3.3.2. Let Q 1 be the 1-polytope with vertices v 1 , v 2 and N = Z. Let λ 1 =<br />

λ(v 1 ) = 1 and λ 2 = λ(v 1 ) = a ∈ Z − {−1, 0, 1}. So (Q 1 , N, {λ i }) is a combinatorial<br />

model. The quasi<strong>toric</strong> orbifold associated to this model is the weighted projective space<br />

WP(1, a).<br />

3.4 Orbifold fundamental group<br />

We first give a canonical construction <strong>of</strong> a quasi<strong>toric</strong> orbifold cover O for any given<br />

quasi<strong>toric</strong> orbifold X . We will prove later that O is the universal orbifold cover <strong>of</strong> X .<br />

Let ̂N be the submodule <strong>of</strong> N generated by the characteristic vectors <strong>of</strong> X .<br />

Definition 3.4.1. Let ̂λ i denote the characteristic vector λ i as an element <strong>of</strong> ̂N. Let<br />

O be the quasi<strong>toric</strong> orbifold associated to the combinatorial model (Q, ̂N, { ̂λ i )}. Denote<br />

the corresponding equivalence relation by ̂∼ so that the underlying topological space <strong>of</strong><br />

O is O = Q × T ̂N/̂∼. Denote the quotient map Q × T ̂N<br />

→ O by ̂π.<br />

Proposition 3.4.2. The quasi<strong>toric</strong> orbifold O is an orbifold cover <strong>of</strong> the quasi<strong>toric</strong><br />

orbifold X with deck group N/ ̂N.<br />

Pro<strong>of</strong>. The inclusion ι : ̂N ↩→ N induces a surjective group homomorphism<br />

ι ∗ : T ̂N<br />

= ( ̂N ⊗ R)/ ̂N → T N = (N ⊗ R)/N<br />

with kernel N/ ̂N. In fact for any face F <strong>of</strong> Q we have commuting diagram<br />

ξ ̂N(F )<br />

T ̂N(F )<br />

−−−−→ T<br />

⏐ ⏐<br />

̂N<br />

ι ⏐↓<br />

0 ι ⏐↓<br />

∗<br />

(3.4.1)<br />

ξ N(F )<br />

T N(F ) −−−−→ T N<br />

where ̂N(F ) is N(F ) viewed as a sublattice <strong>of</strong> ̂N and ι 0<br />

ι. Thus there is an induced surjective map<br />

is an isomorphism induced by<br />

ι 1 : T ̂N/Im(ξ ̂N(F )<br />

) → T N /Im(ξ N(F ) ). (3.4.2)


Chapter 3: Quasi<strong>toric</strong> orbifolds 54<br />

We obtain a torus equivariant map f : O → X defined fiberwise by 3.4.2, that is,<br />

for any point q ∈ Q belonging to the relative interior <strong>of</strong> the face F , the restriction <strong>of</strong><br />

f : ̂π −1 (q) → π −1 (q) matches ι 1 .<br />

The map f lifts to a smooth map <strong>of</strong> orbifolds f : O → X . Consider orbifold charts<br />

on X and O corresponding to vertex v. Identifying ̂N(v) and ̂N(v, F ) with N(v) and<br />

N(v, F ) respectively, we note that<br />

Ŵ v = U v × T ̂N(v)<br />

/̂∼ v may be identified with W v = U v × T N(v) /∼ v .<br />

Hence O v = W v /Ĝv and f : O v → X v is given by the projection W v /Ĝv → W v /G v<br />

where Ĝv = ̂N/N(v) is a subgroup <strong>of</strong> G v = N/N(v). So f : O → X is in fact an orbifold<br />

covering. The deck group for this covering is clearly N/ ̂N.<br />

Theorem 3.4.3. The quasi<strong>toric</strong> orbifold O is the orbifold universal cover <strong>of</strong> the quasi<strong>toric</strong><br />

orbifold X .<br />

N/ ̂N.<br />

The orbifold fundamental group π orb<br />

1 (X ) <strong>of</strong> X is isomorphic to<br />

Pro<strong>of</strong>. Let Σ denote the singular loci <strong>of</strong> X (refer to definition 2.3.9). The set Σ has real<br />

codimension at least 2 in X. Note that π(Σ) is a union <strong>of</strong> faces <strong>of</strong> Q. Let Q Σ<br />

= Q−π(Σ).<br />

Observe that<br />

X − Σ = π −1 (Q Σ<br />

) = (Q Σ<br />

× T N )/ ∼ .<br />

Since Q Σ<br />

is contractible, π 1 (Q Σ<br />

× T N ) ∼ = π 1 (T N ) ∼ = N. When we take quotient <strong>of</strong><br />

Q Σ<br />

× T N by the equivalence relation ∼, certain elements <strong>of</strong> this fundamental group are<br />

killed. Precisely, if Q Σ<br />

contains a point p which belongs to the intersection <strong>of</strong> certain<br />

facets F 1 , . . . , F k <strong>of</strong> Q, then the elements λ 1 , . . . , λ k <strong>of</strong> N given by the corresponding<br />

characteristic vectors map to the identity element <strong>of</strong> π 1 (X − Σ).<br />

Let I(Σ) be the collection <strong>of</strong> facets <strong>of</strong> Q that have nonempty intersection with Q Σ<br />

.<br />

Let N(Σ) be the submodule generated by those λ i for which i ∈ I(Σ).<br />

argument above suggests that<br />

π 1 (X − Σ) = N/N(Σ).<br />

Then the<br />

Indeed, this can be established easily by systematic use <strong>of</strong> the Seifert-Van Kampen<br />

theorem.<br />

It is instructive to first do the pro<strong>of</strong> in the case X is primitive (see Definition 3.1.4).<br />

Here G Fi = 1 (see 3.1.2) for each facet F i . Hence I(Σ) = I and N Σ = ̂N. Therefore<br />

π 1 (X − Σ) = N/ ̂N. Hence by Proposition 3.4.2,<br />

f 0 : O − f −1 (Σ) → X − Σ,<br />

where f 0 is the restriction <strong>of</strong> f, is the universal covering.


55 3.4 Orbifold fundamental group<br />

Now if p : W → X is any orbifold cover then the induced map<br />

p 0 : W − p −1 (Σ) → X − Σ<br />

is a manifold cover. Since p −1 (Σ) has real codimension at least two in W , W − p −1 (Σ)<br />

is connected. Hence f 0 factors through p 0 and f factors through p. Let<br />

g 0 : O − f −1 (Σ) → W − p −1 (Σ)<br />

be the covering map such that f 0 = p 0 ◦ g 0 . We’ll show that g 0 can be extend to an<br />

orbifold covering g : O → W such that f = p ◦ g. Locally the orbifold cover f : O → X<br />

is given by<br />

f : O v = W v /Ĝv → W v /G v = X v<br />

for each vertex v ∈ Q. Also we have the commutative diagram<br />

O v − f −1 Σ<br />

⏐<br />

f ⏐↓<br />

v<br />

g v<br />

−−−−→ p −1 (X v ) − p −1 (Σ)<br />

⏐<br />

p v<br />

⏐↓<br />

X v − Σ<br />

X v − Σ<br />

where f v , g v , p v are restrictions <strong>of</strong> f, g, p respectively. The orbifold O is compact,<br />

Hausdorff, second countable topological space, so by Urysohn metrization theorem O is<br />

a metric space with metric d o and the <strong>topology</strong> induced by d o is the <strong>topology</strong> <strong>of</strong> O.<br />

Let G 1 = π 1 (W − p −1 (Σ)). Define<br />

d 2 : W − p −1 (Σ) × W − p −1 (Σ) → [0, ∞) by d 2 (y 1 , y 2 ) = d 1 (G 1 x 1 , G 1 x 2 ),<br />

where x i ∈ O and g 0 (x i ) = y i , i = 1, 2. Then d 2 is a metric on W − p −1 (Σ). The<br />

<strong>topology</strong> induced by d 2 on W − p −1 (Σ) is quotient <strong>topology</strong> which is subspace <strong>topology</strong><br />

on W − p −1 (Σ)(⊆ W ). Suppose (˜W , ˜d 2 ) is metric completion <strong>of</strong> (W − p −1 (Σ), d 2 ). So<br />

W − p −1 (Σ) is a dense open subset <strong>of</strong> ˜W . Since W − p −1 (Σ) is a dense open subset <strong>of</strong><br />

compact spaces W , W is compact subset <strong>of</strong> ˜W and the <strong>topology</strong> induced by ˜d 2 on W<br />

is the <strong>topology</strong> <strong>of</strong> W .<br />

Now suppose x ∈ O v ∩ f −1 (Σ). From the construction <strong>of</strong> O there exist a Cauchy<br />

sequence {x n } in O v − f −1 (Σ) ⊂ O converging to x. Since g 0 is a covering map<br />

d 2 (g 0 (x l ), g 0 (x m )) = d 2 (G 1 x l , G 1 x m ) ≤ d 1 (x l , x m )<br />

Hence {g 0 (x n )} is a Cauchy sequence in W − p −1 (Σ) and converge to y ∈ W . Define<br />

g(x) = y. So g 0 can be extend to g : O → W .<br />

p(y) = p ◦ g(x)


Chapter 3: Quasi<strong>toric</strong> orbifolds 56<br />

= p ◦ lit n→∞<br />

g(x n )<br />

= lit n→∞<br />

p ◦ g(x n ) ; since p is continuous map<br />

= lit n→∞<br />

f(x n ) = f(x).<br />

Hence f factors through p. From the above commutative diagram it is clear that<br />

g : O → W is an orbifold cover.<br />

For the general case we will use an argument which is similar to that <strong>of</strong> Scott [Sco83]<br />

for orbifold Riemann surfaces.<br />

Proposition 13.2.4 <strong>of</strong> Thurston [Thu3m].<br />

The underlying idea also appeared in remarks after<br />

The group N/ ̂N is naturally a quotient <strong>of</strong> π 1 (X − Σ) = N/N(Σ) and the corresponding<br />

projection homomorphism has kernel K = ̂N/N(Σ). Consider the manifold<br />

covering f 0 : f −1 (X − Σ) → X − Σ obtained by restricting the map f : O → X. Note<br />

that<br />

π 1 (f −1 (X − Σ)) = K<br />

and the deck group <strong>of</strong> f 0 is N/ ̂N. Let W be any orbifold covering <strong>of</strong> X with projection<br />

map p. Then W 0 = W − p −1 (Σ) is a covering <strong>of</strong> X − Σ in the usual sense. We claim<br />

that π 1 (W 0 ) contains K as a subgroup.<br />

Let ¯λ i denote the image <strong>of</strong> λ i in N/N(Σ). Obviously {¯λ i , : i ∈ I − I(Σ)} generate<br />

K. Physically such a ¯λ i can be represented by the conjugate <strong>of</strong> a small loop c i in X − Σ<br />

going around some point x i ∈ π −1 (F ◦<br />

i ) once in a plane transversal to π−1 (F i ), where<br />

F ◦<br />

i denotes the relative interior <strong>of</strong> the facet F i .<br />

The point x i has a neighborhood U in X homeomorphic to C n−1 × (C/G Fi ). Therefore<br />

a connected component V <strong>of</strong> the preimage p −1 (U) ⊂ W is homeomorphic to<br />

C n−1 × (C/G ′ F i<br />

) where G ′ F i<br />

is a subgroup <strong>of</strong> G Fi . We may assume, without loss <strong>of</strong><br />

generality, that<br />

c i lies in the plane {0} × C/G Fi .<br />

By the definition <strong>of</strong> G Fi , ¯λ i is trivial in G Fi and hence in G ′ F i<br />

. Identifying G Fi with<br />

the deck group <strong>of</strong> the covering C ∗ → C ∗ /G Fi , we infer that c i lifts to a loop in C ∗ and<br />

consequently in C ∗ /G ′ F i<br />

. Hence c i lifts to a loop in V − p −1 (Σ). Thus each generator<br />

and therefore every element <strong>of</strong> K is represented by a loop in W 0 .<br />

homomorphism<br />

K → π 1 (W 0 ).<br />

This induces a<br />

This homomorphism is injective since K is a subgroup <strong>of</strong> the fundamental group <strong>of</strong> the<br />

space X − Σ which has W 0 as a cover.<br />

For any orbifold covering W <strong>of</strong> X , the associated covering W 0 <strong>of</strong> X − Σ admits a<br />

covering by f −1 (X −Σ) ⊂ O since π 1 (f −1 (X −Σ)) = K is a normal subgroup <strong>of</strong> π 1 (W 0 ).<br />

Thus O is an orbifold cover <strong>of</strong> W. Hence O is the universal orbifold cover <strong>of</strong> X and H<br />

is the orbifold fundamental group <strong>of</strong> X .


57 3.5 q-Cellular homology groups<br />

Remark 3.4.4. Note that the orbifold fundamental group <strong>of</strong> a quasi<strong>toric</strong> orbifold is<br />

always a finite group. It follows that a quasi<strong>toric</strong> orbifold is a global quotient if and only<br />

if its orbifold universal cover is a smooth manifold. Therefore Theorem 3.4.3 yields a<br />

rather easy method for determining if a quasi<strong>toric</strong> orbifold is a global quotient or not.<br />

For example, by Theorem 3.4.3 π1 orb (WP(1, a)) = 1 which imply that WP(1, a) is not a<br />

global quotient.<br />

Example 3.4.5. If<br />

̂N = N, then X is not a global quotient unless X is a manifold.<br />

For instance, let Q be a 2-dimensional simplex with characteristic vectors<br />

(1, 1), (1, −1), (−1, 0) and let X be the quasi<strong>toric</strong> orbifold corresponding to this model.<br />

Then N = ̂N, but X has an orbifold singularity at π −1 (v) where v = F 1 ∩ F 2 . Therefore<br />

X is not a global quotient.<br />

3.5 q-Cellular homology groups<br />

We introduce the notion <strong>of</strong> q-CW complex where an open cell is the quotient <strong>of</strong> an open<br />

disk by action <strong>of</strong> a finite group. Otherwise the construction mirrors the construction <strong>of</strong><br />

usual CW complex given in Hatcher [Hat02]. We show that our q-cellular homology <strong>of</strong><br />

a q-CW complex is isomorphic to its singular homology with coefficients in Q.<br />

Definition 3.5.1. Let G be a finite group acting linearly, preserving orientation, on an<br />

n-dimensional disk D n centered at the origin. Such an action preserves S n−1 = ∂D n .<br />

We call the quotient D n /G an n-dimensional q-disk. Call S n−1 /G a q-sphere.<br />

An n-dimensional q-cell e n G = en (G)/G is defined to be a copy <strong>of</strong> D n /G where e n (G)<br />

is G-equivariantly homeomorphic to D n . We will denote the boundary <strong>of</strong> e n (G) by S n−1<br />

without confusion.<br />

Start with a discrete set X 0 , where points are regarded as 0-dimensional q-cells. Inductively,<br />

form the n-dimensional q-skeleton X n from X n−1 by attaching n-dimensional<br />

q-cells e n G α<br />

via continuous maps<br />

φ α : S n−1<br />

α /G α → X n−1 .<br />

This means that X n is the quotient space <strong>of</strong> the disjoint union X n−1 ⊔ α e n G α<br />

<strong>of</strong> X n−1<br />

with a finite collection <strong>of</strong> n-dimensional q-disks e n α(G α )/G α under the identification<br />

x ∼ n φ α (x) for x ∈ S n−1<br />

α /G α .<br />

Assume X = X n for some finite n. The <strong>topology</strong> <strong>of</strong> X is the quotient <strong>topology</strong> built<br />

inductively. We call a space X constructed in this way a finite q-CW complex.<br />

By Proposition 2.22 and Corollary 2.25 <strong>of</strong> [Hat02],


Chapter 3: Quasi<strong>toric</strong> orbifolds 58<br />

Note that<br />

H p ((X n , X n−1 ); Q) = ⊕ α<br />

˜H p ( Dn α/G α<br />

S n−1 α /G α<br />

; Q) =<br />

{<br />

˜H p ( Dn α/G α<br />

S n−1 α /G α<br />

; Q) (3.5.1)<br />

H p−1 (S n−1<br />

α /G α ; Q) if p ≥ 2<br />

0 otherwise<br />

(3.5.2)<br />

Lemma 3.5.2. Let D n /G be a q-disk. Then S n−1 /G is a Q-homology sphere.<br />

Pro<strong>of</strong>. S n−1 admits a simplicial G-complex structure. Apply Theorem 2.4 <strong>of</strong> Bredon<br />

[Bre72] and Poincaré duality for orbifolds.<br />

Lemma 3.5.3. If X is a q-CW complex, then<br />

{<br />

1. H p ((X n , X n−1 ); Q) =<br />

0 for p ≠ n<br />

⊕ Q i∈I n for p = n<br />

where I n is the set <strong>of</strong> n-dimensional q-cells in X.<br />

2. H p (X n ; Q) = 0 for p > n. In particular, H p (X; Q) = 0 for p > dim(X).<br />

3. The inclusion i : X n ↩→ X induces an isomorphism i ∗ : H p (X n ; Q) → H p (X; Q) if<br />

p < n.<br />

Pro<strong>of</strong>. Pro<strong>of</strong> is similar to the pro<strong>of</strong> <strong>of</strong> Lemma 2.3.4 <strong>of</strong> [Hat02]. The key ingredient is<br />

Lemma 3.5.2.<br />

Using Lemma 3.5.3 we can define q-cellular chain complex (H p (X p , X p−1 ), d p ) and q-<br />

cellular groups Hp<br />

q-CW (X; Q) <strong>of</strong> X in the same way as cellular chain complex is defined<br />

in [Hat02], page 139.<br />

Theorem 3.5.4. H q-CW<br />

p (X; Q) ∼ = H p (X; Q) for all p.<br />

Pro<strong>of</strong>. Pro<strong>of</strong> is similar to the pro<strong>of</strong> <strong>of</strong> Theorem 2.35 <strong>of</strong> [Hat02].<br />

3.6 Rational homology <strong>of</strong> quasi<strong>toric</strong> orbifolds<br />

Following Goresky [Gor78] one may obtain a CW structure on a quasi<strong>toric</strong> orbifold.<br />

However it is too complicated for easy computation <strong>of</strong> homology. We now follow the<br />

main ideas <strong>of</strong> the computation for the manifold case as in Section 1.5 to compute the<br />

rational homology groups <strong>of</strong> quasi<strong>toric</strong> orbifold X over Q. First we construct a q-CW<br />

structure on X. We adhere the notations <strong>of</strong> Section 1.5.<br />

Let v ∈ Q be a vertex <strong>of</strong> index f(v) = k. We put e v = π −1 ( ̂F v ). Then e v is a<br />

contractible subspace <strong>of</strong> X(F v ) homeomorphic to the quotient <strong>of</strong> an open disk D 2f(v) in<br />

R 2f(v) by a finite group G(v) determined by the orbifold structure on X(F v ) described


59 3.7 Gysin sequence for q-sphere bundle<br />

in Section 3.3. ̂Fv is homeomorphic to the intersection <strong>of</strong> the unit disk in R f(v) with<br />

R f(v)<br />

≥0<br />

. Since the action <strong>of</strong> the group G(v) is obtained from a combinatorial model, see<br />

Section 3.3, e v is a 2f(v)-dimensional q-cell.<br />

X can be given the structure <strong>of</strong> a q-CW complex as follows. Define the k-skeleton<br />

X 2k := ∪ f(v)=k X(F v) for 0 ≤ k ≤ n. X 2k+1 = X 2k for 0 ≤ k ≤ n − 1 and X 2n = X.<br />

X 2k can be obtained from X 2k−1 by attaching those q-cells e v for which f(v) = k. The<br />

attaching maps are to be described. Let ∼ be the equivalence relation such that<br />

X(F v ) = F v × T N ⊥ (F v)/∼.<br />

The q-disk D 2f(v) /G(v) can be identified with F v × T N ⊥ (F v )/≈ where<br />

(p, t) ≈ (q, s) if p = q ∈ F ′ for some face F ′ whose top vertex is v and (p, t) ∼ (q, s).<br />

The attaching map φ v : S 2f(v)−1 /G(v) → X 2f(v)−1 is the natural quotient map<br />

(F v − ̂F v ) × T N ⊥ (F v)/≈ → (F v − ̂F v ) × T N ⊥ (F v)/∼.<br />

So X is a q-CW complex with no odd dimensional cells and with f −1 (k) = h k number<br />

<strong>of</strong> 2k-dimensional q-cells. Hence by q-cellular homology theory<br />

H q-CW<br />

p (X; Q) =<br />

{ ⊕<br />

f −1 (p/2) Q<br />

0 otherwise<br />

if p ≤ n and p is even<br />

(3.6.1)<br />

Hence by Theorem 3.5.4<br />

H p (X; Q) =<br />

{ ⊕<br />

h (p/2)<br />

Q if p ≤ n and p is even<br />

0 otherwise<br />

(3.6.2)<br />

3.7 Gysin sequence for q-sphere bundle<br />

Let ρ : E → B be a rank n vector bundle with paracompact base space B. Restricting<br />

ρ to the space E 0 <strong>of</strong> nonzero vectors in E, we obtain an associated projection map ρ 0 :<br />

E 0 → B. Fix a finite group G and a representation <strong>of</strong> G on R n . Such a representation<br />

induces a fiberwise linear action <strong>of</strong> G on E and E 0 . Consider the two fiber bundles<br />

ρ G : E/G → B and ρ G 0 : E 0 /G → B.<br />

There exist natural fiber bundle maps f 1 : E → E/G and f 2 : E 0 → E 0 /G. These<br />

induce isomorphisms<br />

f1 ∗ : H p (E/G) → H p (E) and f2 ∗ : H p (E 0 /G) → H p (E)


Chapter 3: Quasi<strong>toric</strong> orbifolds 60<br />

for each p. The second isomorphism is obtained by applying Theorem 2.4 <strong>of</strong> [Bre72]<br />

fiberwise and then using Kunneth formula, Mayer-Vietoris sequence and a direct limit<br />

argument as in the pro<strong>of</strong> <strong>of</strong> Thom isomorphism in [MS74]. The commuting diagram<br />

i<br />

E<br />

1<br />

j 1<br />

0 −−−−→ E −−−−→ (E, E0 )<br />

f 2<br />

⏐ ⏐↓<br />

f 1<br />

⏐ ⏐↓<br />

⏐ ⏐↓<br />

f 3<br />

E 0 /G i 2<br />

−−−−→ E/G j 2<br />

−−−−→ (E/G, E 0 /G)<br />

induces a commuting diagram <strong>of</strong> two exact rows<br />

· · · → H p−1 (E 0 )<br />

↑<br />

⏐<br />

δ ∗ 1<br />

−−−−→ H p (E, E 0 )<br />

↑<br />

⏐<br />

f2<br />

∗ ⏐ f3<br />

∗<br />

j ∗ 1<br />

−−−−→<br />

H p (E)<br />

↑<br />

⏐<br />

⏐f ∗ 1<br />

i ∗ 1<br />

−−−−→ H p (E 0 ) → · · ·<br />

↑<br />

⏐<br />

⏐f ∗ 2<br />

· · · → H p−1 (E 0 /G)<br />

δ ∗ 2<br />

−−−−→ H p (E/G, E 0 /G)<br />

j ∗ 2<br />

−−−−→ H p (E/G)<br />

i ∗ 2<br />

−−−−→ H p (E 0 /G) → · · ·<br />

By the five lemma f ∗ 3 is an isomorphism. Using the Thom isomorphism ∪u : Hp−n (E) →<br />

H p (E, E 0 ) we get the isomorphism<br />

∪u G : H p−n (E/G) → H p (E/G, E 0 /G) where ∪ u G = f ∗ 3 −1 ◦ ∪u ◦ f ∗ 1 .<br />

Substituting the isomorphic module H p−n (E/G) in place <strong>of</strong> H p (E/G, E 0 /G) in the<br />

second row <strong>of</strong> the above diagram, we obtain an exact sequence<br />

· · · → H p−n (E/G)<br />

g<br />

−−−−→ H p (E/G) → H p (E 0 /G) → H p−n+1 (E/G) → · · ·<br />

where g = j ∗ 2 ◦ ∪u G. The pull back <strong>of</strong> cohomology class u G |(E/G) in H n (B) by the<br />

zero section <strong>of</strong> ρ G will be called the Euler class e <strong>of</strong> ρ G . Now substitute the isomorphic<br />

cohomology ring H ∗ (B) in place <strong>of</strong> H ∗ (E/G) in the above sequence. This yields the<br />

Gysin exact sequence for the q-sphere bundle ρ G 0 : E 0/G → B<br />

· · · → H p−n (B; Q)<br />

∪e<br />

−−−−→ H p (B; Q) → H p (E 0 /G; Q) → H p−n+1 (B; Q) → · · · (3.7.1)<br />

Remark 3.7.1. Euler classes <strong>of</strong> ρ : E → B and ρ G : E/G → B are the same since f ∗ 1<br />

is an isomorphism.<br />

3.8 Cohomology ring <strong>of</strong> quasi<strong>toric</strong> orbifolds<br />

Again we will modify some technical details but retain the broad framework <strong>of</strong> the argument<br />

in [DJ91] to get the anticipated answer. All homology and cohomology modules<br />

in this section will have coefficients in Q. Let L be the simplicial complex associated<br />

to the boundary <strong>of</strong> the dual polytope <strong>of</strong> Q. Then Q is the cone on the barycentric


61 3.8 Cohomology ring <strong>of</strong> quasi<strong>toric</strong> orbifolds<br />

subdivision <strong>of</strong> L. Q can be split into cubes Q σ where σ varies over (n − 1)-dimensional<br />

faces <strong>of</strong> L. These correspond bijectively to vertices <strong>of</strong> Q. We regard the k-cube as the<br />

orbit space <strong>of</strong> standard k-dimensional torus action on the 2k-disk<br />

D 2k = {(z 1 , · · · , z k ) ∈ C k : |z i | ≤ 1} (3.8.1)<br />

Define<br />

BQ σ = ET N × TN ((Q σ × T N )/ ∼) ≃ ET N × TN (D 2n /G σ ),<br />

where G σ = G vσ , v σ being the vertex in Q dual to σ. If σ 1 is another (n − 1) simplex<br />

in L such that σ ∩ σ 1 is an (n − 2) simplex then BQ σ and BQ σ1 are glued along the<br />

common part <strong>of</strong> the boundaries <strong>of</strong> Q σ and Q σ1 . In this way BQ σ fit together to yield<br />

BQ = ET N × TN X.<br />

Let p : BQ → BT N be the Borel map which is a fibration with fiber X. The fibration<br />

p : BQ → BT N induces a homomorphism<br />

p ∗ : H ∗ (BT N ; Q) → H ∗ (BQ; Q).<br />

The face ring SR(Q, Q) is graded by declaring the degree <strong>of</strong> each w i to be 2. The<br />

following result resembles Theorem 4.8 <strong>of</strong> [DJ91].<br />

Theorem 3.8.1. Let Q be an n-polytope and SR(Q, Q) be the face ring <strong>of</strong> Q. The map<br />

p ∗ : H ∗ (BT N ; Q) → H ∗ (BQ; Q) is surjective and induces an isomorphism <strong>of</strong> graded<br />

rings H ∗ (BQ; Q) ∼ = SR(Q, Q).<br />

Pro<strong>of</strong>. Suppose σ is an (n − 1)-simplex in L with vertices w 1 , . . . , w n . Note that there<br />

is a one-to-one correspondence between facets <strong>of</strong> Q meeting at v σ and vertices <strong>of</strong> σ. Let<br />

Q σ be the corresponding n-cube in Q. Then BQ σ = ET N × TN (D 2n /G σ ) is a D 2n /G σ<br />

fiber bundle over BT N . Hence<br />

ET N × TN (S 2n−1 /G σ ) → BT N<br />

give the associated q-sphere bundle p σ : BQ ∂σ → BT N . Also consider the disk bundle<br />

r : ET N × TN D 2n → BT N .<br />

It is bundle homotopic to the complex vector bundle<br />

r ′ : ET N × TN C n → BT N .<br />

Since T N acts diagonally on C n , the last bundle is the sum <strong>of</strong> line bundles L 1 ⊕ · · · ⊕ L n


Chapter 3: Quasi<strong>toric</strong> orbifolds 62<br />

where L j corresponds to j-th coordinate direction in C n and hence to w j . Without<br />

confusion, we set<br />

c 1 (L i ) = w i ∈ H 2 (BT N ; Q).<br />

By the Whitney product formula c n (r ′ ) = w 1 · · · w n . Hence from Section 3.7 the Euler<br />

class <strong>of</strong> the q-sphere bundle p σ is e = w 1 · · · w n .<br />

Now consider the Gysin exact sequence for q-sphere bundles<br />

· · · → H ∗ (BQ ∂σ ) → H ∗ (BT N )<br />

→ H ∗+2n (BT N ) → · · ·<br />

∪e<br />

−−−−→ H ∗+2n (BT N )<br />

p ∗ σ<br />

−−−−→ H ∗+2n (BQ ∂σ )<br />

(3.8.2)<br />

Since the map ∪e is injective, by exactness p ∗ σ is surjective and we get the following<br />

diagram<br />

0 → H ∗ (BT N )<br />

⏐<br />

id↓<br />

Q[w 1 , · · · , w n ]<br />

∪e<br />

−−−−→ H ∗+2n (BT N )<br />

⏐<br />

id↓<br />

w 1 ...w n<br />

−−−−→ Q[w 1 , . . . , w n ].<br />

p ∗ σ<br />

−−−−→ H ∗+2n (BQ ∂σ ) → 0<br />

(3.8.3)<br />

Hence from diagram (3.8.3) H ∗ (BQ ∂σ ) = Q[w 1 , · · · , w n ]/(w 1 . . . w n ). Since D 2n /G σ<br />

is contractible, H ∗ (BQ σ ; Q) = H ∗ (BT N ; Q) = Q[w 1 , · · · , w n ]. Using induction on the<br />

dimension <strong>of</strong> L and an application <strong>of</strong> the Mayer-Vietoris sequence we get the conclusion<br />

<strong>of</strong> the Theorem.<br />

Let j : X → BQ be inclusion <strong>of</strong> the fiber. Consider the Serre spectral sequence <strong>of</strong><br />

the fibration p : BQ → BT N with fiber X. It has E 2 -term<br />

E p,q<br />

2 = H p (BT N ; H q (X)) = H p (BT N ) ⊗ H q (X).<br />

Using the formula for Poincaré series <strong>of</strong> X it can be proved that this spectral sequence<br />

degenerates, E p,q<br />

2 = E∞ p,q (see Theorem 1.8.2). So it follows that H ∗ (BQ, Q) ∼ =<br />

H ∗ (BT n , Q) ⊗ H ∗ (M 2n , Q) as Q-modules. Hence j ∗ : H ∗ (BQ, Q) → H ∗ (X, Q) is surjective.<br />

We have natural identifications H 2 (BQ) = Q m and H 2 (BT N ) = Q n . Here Q m is<br />

regarded as the Q vector space with basis corresponding to the set <strong>of</strong> codimension one<br />

faces <strong>of</strong> Q. The map<br />

p ∗ : H 2 (BQ) → H 2 (BT N )<br />

is naturally identified with the characteristic map Λ λ : Q m → Q n that sends w i , the<br />

i-th standard basis vector <strong>of</strong> Q m , to λ i . The map p ∗ : H 2 (BT N ) → H 2 (BQ) is then<br />

identified with the dual map Λ ∗ λ : (Qn ) ∗ → (Q m ) ∗ . Regarding the map Λ λ as an n × m


63 3.9 Stable almost complex structure<br />

matrix λ ij , the matrix for Λ ∗ λ<br />

is the transpose. Column vectors <strong>of</strong> Λ∗ λ<br />

regarded as linear combinations <strong>of</strong> w 1 , . . . , w m . Define<br />

can then be<br />

We have a short exact sequence<br />

λ i = λ i1 w 1 + · · · + λ im w m . (3.8.4)<br />

0 → H 2 (BT N )<br />

∥<br />

p ∗<br />

−−−−→ H 2 (BQ)<br />

∥<br />

j ∗<br />

−−−−→ H 2 (X) → 0<br />

(Q n ) ∗ Λ ∗ λ<br />

−−−−→ (Q m ) ∗ .<br />

Let J be the homogeneous ideal in Q[w 1 , . . . , w m ] generated by the λ i and let J be<br />

its image in SR(Q, Q). Since j ∗ : SR(Q, Q) → H ∗ (X) is onto and J is in its kernel, j ∗<br />

induces a surjection SR(Q, Q)/J → H ∗ (X).<br />

Theorem 3.8.2. Let X be the quasi<strong>toric</strong> orbifold associated to the combinatorial model<br />

(Q, N, λ). Then H ∗ (X; Q) is the quotient <strong>of</strong> the face ring <strong>of</strong> Q by J ; i.e., H ∗ (X; Q) =<br />

Q[w 1 , . . . , w m ]/(I + J ).<br />

Pro<strong>of</strong>. We know that H ∗ (BT N ) is a polynomial ring on n generators, and H ∗ (BQ) is<br />

the face ring. Since the spectral sequence degenerates, H ∗ (BQ) ≃ H ∗ (BT N ) ⊗ H ∗ (X).<br />

Furthermore, p ∗ : H ∗ (BT N ) → H ∗ (BQ) is injective and J is identified with the image<br />

<strong>of</strong> p ∗ . Thus H ∗ (X) = H ∗ (BQ)/J = Q[w 1 , . . . , w m ]/(I + J ).<br />

3.9 Stable almost complex structure<br />

Buchstaber and Ray [BR01] have shown the existence <strong>of</strong> a stable almost complex structure<br />

on omnioriented quasi<strong>toric</strong> manifolds. We generalize their result to omnioriented<br />

quasi<strong>toric</strong> orbifolds (see Section 1.6 for definition). Let m be the cardinality <strong>of</strong> I, the<br />

set <strong>of</strong> facets <strong>of</strong> the polytope Q. We will realize the orbifold X as the quotient <strong>of</strong> the<br />

action <strong>of</strong> an appropriate subgroup <strong>of</strong> (C ∗ ) m on an open set <strong>of</strong> C m . Consider the natural<br />

combinatorial model (R m ≥0 , L ∼ = Z m , {e i }) for C m , where e i is the i-th standard vector<br />

<strong>of</strong> Z m . Let<br />

π s : C m → R m ≥0<br />

be the projection map corresponding to taking modulus coordinatewise. Embed the<br />

polytope Q in R m ≥0 by the map<br />

d F : Q → R m<br />

where the i-th coordinate <strong>of</strong> d F (p) is the Euclidean distance (d(p, F i )) <strong>of</strong> p from the<br />

hyperplane <strong>of</strong> the i-th facet F i in R n . Consider the thickening W R (Q) ⊂ R m ≥0 <strong>of</strong> d F(Q),


Chapter 3: Quasi<strong>toric</strong> orbifolds 64<br />

defined by<br />

W R (Q) = {f : I → R ≥0 |f −1 (0) ∈ L F (Q)} (3.9.1)<br />

where L F (Q) denotes the face lattice <strong>of</strong> Q.<br />

Denote the n-dimensional linear subspace <strong>of</strong> R m parallel to d F (Q) by V Q and its<br />

orthogonal complement by V ⊥ Q .<br />

As a manifold with corners, W R (Q) is canonically<br />

diffeomorphic to the Cartesian product d F (Q) × exp(VQ ⊥ ) (see [BR01], Proposition 3.4).<br />

Define the spaces W (Q) and Z(Q) as follows.<br />

W (Q) := π −1<br />

s (W R (Q)), Z(Q) := π −1<br />

s (d F (Q)). (3.9.2)<br />

W (Q) is an open subset <strong>of</strong> C m and there is a canonical diffeomorphism<br />

W (Q) ∼ = Z(Q) × exp(V ⊥ Q ). (3.9.3)<br />

Let Λ : L → N be the map <strong>of</strong> Z-modules which maps the standard generator e i <strong>of</strong><br />

L to the dicharacteristic vector λ i . Let K denote the kernel <strong>of</strong> this map. Recall the<br />

submodule ̂N <strong>of</strong> N generated by the dicharacteristic vectors and the orbifold universal<br />

cover O from Section 3.4. Since the Z-modules L and ̂N are free, the sequence<br />

0 −→ K −→ L Λ −→ ̂N −→ 0 (3.9.4)<br />

splits and we may write L = K ⊕ ̂N. Hence K R ∩ L = K and applying the second isomorphism<br />

theorem for groups we can consider the torus T K := K R /K to be a subgroup<br />

<strong>of</strong> T L . In fact we get a split exact sequence<br />

Λ<br />

1 −→ T K −→ T<br />

∗<br />

L −→ T ̂N<br />

−→ 1 (3.9.5)<br />

For any face F <strong>of</strong> Q let L(F ) be the sublattice <strong>of</strong> L generated by the basis vectors<br />

e i such that d F (F ) intersects the i-th facet <strong>of</strong> R m ≥0 , that is the coordinate hyperplane<br />

{x i = 0}. Note that image <strong>of</strong> L(F ) under Λ is precisely ̂N(F ), so that the preimage<br />

Λ −1 ( ̂N(F )) = K · L(F ). Consider the exact sequence<br />

0 −→ K · L(F )<br />

L(F )<br />

−→<br />

L<br />

L(F )<br />

Λ<br />

−→<br />

̂N −→ 0. (3.9.6)<br />

̂N(F )<br />

Since the dicharacteristic vectors corresponding to the facets whose intersection is F<br />

are linearly independent, it follows from the definition <strong>of</strong> K and Λ that K ∩L(F ) = {0}.<br />

Hence by the second isomorphism theorem we have a canonical isomorphism<br />

K · L(F )<br />

L(F )<br />

∼= K. (3.9.7)


65 3.9 Stable almost complex structure<br />

So 3.9.6 yields<br />

0 −→ K −→ L<br />

L(F )<br />

Λ<br />

−→<br />

̂N −→ 0. (3.9.8)<br />

̂N(F )<br />

In general<br />

̂N<br />

̂N(F ) is not a free Z-module. Let ̂N ′ (F ) = ( ̂N(F ) ⊗ Z Q) ∩ ̂N. Define<br />

Λ ′ = Λ ◦ φ (3.9.9)<br />

where φ is the canonical projection<br />

φ :<br />

̂N<br />

̂N(F ) −→<br />

̂N<br />

̂N ′ (F ) . (3.9.10)<br />

Since<br />

̂N<br />

̂N ′ (F )<br />

is free, the following exact sequence splits<br />

0 −→ Λ−1 ( ̂N ′ (F ))<br />

L(F )<br />

−→<br />

L<br />

L(F )<br />

Λ ′<br />

−→<br />

̂N −→ 0. (3.9.11)<br />

̂N ′ (F )<br />

Denoting the modules in 3.9.11 by K, L and N respectively we obtain a split exact<br />

sequence <strong>of</strong> tori<br />

θ<br />

0 −→ T<br />

1 Λ<br />

K<br />

−→ ′ ∗ TL −→ T N<br />

−→ 0. (3.9.12)<br />

Note that K is a submodule <strong>of</strong> same rank <strong>of</strong> the free module K and there is a natural<br />

exact sequence<br />

0 −→ ̂N ′ (F )<br />

̂N(F ) −→ T K<br />

θ 2<br />

−→ TK −→ 0. (3.9.13)<br />

The composition<br />

θ 1 ◦ θ 2 : T K −→ T¯L (3.9.14)<br />

defines a natural action <strong>of</strong> T K on T L<br />

with isotropy ĜF = ̂N ′ (F )/ ̂N(F ) and quotient<br />

T N<br />

.<br />

Since T N<br />

is the fiber <strong>of</strong> ̂π : O → Q and T L<br />

is the fiber <strong>of</strong> π s : Z(Q) → Q over any<br />

point in the relative interior <strong>of</strong> the arbitrary face F , it follows O is quotient <strong>of</strong> Z(Q)<br />

by the above action <strong>of</strong> T K . This action <strong>of</strong> T K is same as the restriction <strong>of</strong> its action on<br />

C m as a subgroup <strong>of</strong> T L and hence (C ∗ ) m . By 3.9.3 it follows that O is the quotient <strong>of</strong><br />

the open set W (Q) in C m by the action <strong>of</strong> the subgroup T K × exp(VQ ⊥) <strong>of</strong> (C∗ ) m ,<br />

O =<br />

W (Q)<br />

T K × exp(VQ ⊥ (3.9.15)<br />

).<br />

The induced action <strong>of</strong> Ĥ := T K × exp(VQ ⊥ ) on the real tangent bundle T W (Q) <strong>of</strong>


Chapter 3: Quasi<strong>toric</strong> orbifolds 66<br />

W (Q) commutes with the almost complex structure<br />

J : T W (Q) → T W (Q)<br />

obtained by restriction <strong>of</strong> the standard almost complex structure on T C m . Therefore<br />

the quotient W <strong>of</strong> T W (Q) by Ĥ has the structure <strong>of</strong> an almost complex orbibundle (or<br />

orbifold vector bundle) over O. Moreover this quotient splits, by an Atiyah sequence<br />

( [Ati57]), as the direct sum <strong>of</strong> a trivial rank 2(n−m) real bundle ĥ over O corresponding<br />

to the Lie algebra <strong>of</strong> Ĥ and the orbifold tangent bundle T O <strong>of</strong> O. The existence <strong>of</strong> a<br />

stable almost complex structure on T O is thus established.<br />

The tangent bundle T C m splits naturally into a direct sum <strong>of</strong> m complex line bundles<br />

corresponding to the complex coordinate directions which <strong>of</strong> course correspond to the<br />

facets <strong>of</strong> Q. We get a corresponding splitting<br />

T W (Q) = ⊕C F .<br />

The bundles C F are invariant under J as well Ĥ. Therefore the quotient <strong>of</strong> C F by Ĥ is<br />

a complex orbibundle ̂ν(F ) <strong>of</strong> rank one on O and<br />

Ŵ = ⊕̂ν(F ).<br />

It is not hard to see that the natural action <strong>of</strong> T ̂N<br />

on Ŵ commutes with the almost<br />

complex structure on it. The quotient<br />

W := Ŵ/(N/ ̂N)<br />

is an orbibundle on X with an induced almost complex structure since (N/ ̂N) is a<br />

subgroup <strong>of</strong> T ̂N.<br />

Furthermore T X is the quotient <strong>of</strong> T O by N/ ̂N. Therefore<br />

W = T X ⊕ h<br />

where h is the quotient <strong>of</strong> ĥ by N/ ̂N. Since the action <strong>of</strong> T ̂N<br />

and hence N/ ̂N on ĥ is<br />

trivial, h is a trivial vector bundle on X. Hence the almost complex structure on W<br />

induces a stable almost complex structure on T X . We also have a decomposition<br />

W = ⊕ν(F )<br />

where the orbifold line bundle ν(F ) := ̂ν(F )/(N/ ̂N).


67 3.10 Line bundles and cohomology<br />

3.10 Line bundles and cohomology<br />

Recall the manifold Z(Q) <strong>of</strong> dimension m + n defined in equation 3.9.2. Let B L Q =<br />

ET L × TL Z(Q). Since O = Z(Q)/T K ,<br />

B L Q = ET L × TL Z(Q) = ET L × TK Z(Q)/(T L /T K ) = ET L × (Z(Q)/T K )/(T ̂N)<br />

≃<br />

ET ̂N<br />

× T ̂N<br />

O = ET ̂N<br />

× TN O/(N/ ̂N) ≃ ET N × TN X = BQ.<br />

Let w 1 , . . . , w m be the generators <strong>of</strong> H 2 (BQ) as in Section 3.8 and let F i denote the<br />

facet <strong>of</strong> Q corresponding to w i . Let<br />

α i : T L → T 1<br />

be the projection onto the i-th factor and C(α i ) denote the corresponding 1-dimensional<br />

representation space <strong>of</strong> T L . Define<br />

L i = ET L × TL ˜Li ,<br />

where ˜L i = C(α i ) × Z(Q) is the trivial equivariant line bundle over Z(Q). Then L i is<br />

an orbifold line bundle over BQ. Let c 1 (L i ) be the first Chern class <strong>of</strong> L i in H 2 (BQ; Q).<br />

We will show that c 1 (L i ) = w i .<br />

Since the i-th factor <strong>of</strong> T L acts freely on Z(Q) − πs<br />

−1 (F i ), the restriction <strong>of</strong> L i to<br />

BQ − BF i is trivial. Consider the following commutative diagram<br />

ι ∗ (L i ) −−−−→ L i<br />

⏐<br />

⏐<br />

↓<br />

↓<br />

(BQ − BF i )<br />

ι<br />

−−−−→ BQ<br />

where ι is inclusion map. By naturality c 1 (ι ∗ (L i )) = ι ∗ (c 1 (L i )). Since the bundle ι ∗ (L i )<br />

over BQ − BF i is trivial ι ∗ (c 1 (L i )) = c 1 (ι ∗ (L i )) = 0. It is easy to show that<br />

B(Q − F i ) = ET L × TL (π −1<br />

s (Q − F i )) ≃ BQ − BF i .<br />

From the pro<strong>of</strong> <strong>of</strong> Theorem 3.8.1 it is evident that H ∗ (BQ − BF i ; Q) ∼ = SR(Q − F i , Q).<br />

Hence H 2 (BQ − BF i , Q) = ⊕ j≠i Qw j. The map<br />

ι ∗ : H 2 (BQ; Q) → H 2 (BQ − BF i ; Q)<br />

is a surjective homomorphism with kernel Qw i implying c 1 (L i ) ∈ Qw i . Naturality axiom<br />

ensures, as follows, that c 1 (L i ) is nonzero, so that we can identify c 1 (L i ) with w i .<br />

Let F be an edge in F i . Then<br />

BF := ET L × TL (π −1<br />

s (F )) ≃ ET N × TN (π −1 (F )) = (ET N × TF π −1 (F ))/(T N /T F )


Chapter 3: Quasi<strong>toric</strong> orbifolds 68<br />

= (ET N × (π −1 (F )/T F ))/(T N /T F ) ≃ E(T N /T F ) × TN /T F<br />

π −1 (F ) ≃ ES 1 × S 1 S 2 ,<br />

where T F is the isotropy subgroup <strong>of</strong> F in T N and action <strong>of</strong> S 1 on S 2 is corresponding<br />

action <strong>of</strong> T N /T F on π −1 (F ). Let L i (F ) is the pullback <strong>of</strong> orbibundle L i . Using Thom<br />

isomorphism and cohomology exact sequence obtained from<br />

BF<br />

s<br />

−−−−→ L i (F ) → (L i (F ), BF )<br />

where s is zero section <strong>of</strong> L i bundle, we can show c 1 (L i (F )) is nonzero. Since c 1 (L i (F ))<br />

is pullback <strong>of</strong> c 1 (L i ), c 1 (L i ) is nonzero. Hence c 1 (L i ) = w i . Note that if F i is the facet<br />

<strong>of</strong> Q corresponding to L i ,<br />

L i = ET L × TL ˜Li = ET ̂N<br />

× T ̂N<br />

(˜L i /T K ) = ET ̂N<br />

× T ̂N<br />

̂ν(F i ) = ET ̂N<br />

× TN ̂ν(F i )/(N/ ̂N)<br />

≃ ET N × TN ν(F i ). Let<br />

j : ν(F i ) ↩→ L i<br />

be the inclusion <strong>of</strong> fiber covering j : X ↩→ BQ. Then j ∗ (L i ) = ν(F i ). Hence<br />

c 1 (ν(F i )) = j ∗ c 1 (L i ) = j ∗ w i .<br />

Hence by Theorem 3.8.2 the first Chern classes <strong>of</strong> the bundles ν(F i ) generate the cohomology<br />

ring <strong>of</strong> X. We also obtain the formula for the total Chern class <strong>of</strong> T X with the<br />

stable almost complex structure determined by the given dicharacteristic.<br />

m∏<br />

c(T ) = (1 + c 1 (ν(F i ))) (3.10.1)<br />

3.11 Chern numbers<br />

i=1<br />

Chern numbers <strong>of</strong> an omnioriented quasi<strong>toric</strong> orbifold, with the induced stable almost<br />

complex structure, can be computed using standard localization formulae, given for<br />

instance in Chapter 9 <strong>of</strong> [CK99]. The fixed points <strong>of</strong> the T N action correspond to the<br />

vertices <strong>of</strong> Q. While computing the numerator contributions at a vertex, one needs to<br />

recall that T N action on the bundle h is trivial. We will give a formula for the top Chern<br />

number below. In the manifold case similar formula was obtained by Panov in [Pan01].<br />

In principle any Hirzebruch genus associated to a series may be computed similarly.<br />

Fix an orientation for X by choosing orientations for Q ⊂ R n and T N . We order the<br />

facets or equivalently the dicharacteristic vectors at each vertex in a compatible manner<br />

as follows.<br />

Suppose the vertex v <strong>of</strong> Q is the intersection <strong>of</strong> facets F i1 , . . . , F in . To each <strong>of</strong> these


69 3.12 Chen-Ruan cohomology groups<br />

facets F ik<br />

assign the unique edge E k <strong>of</strong> Q such that<br />

F ik ∩ E k = v.<br />

Let ê k be a vector along E k with origin at v. Then ê 1 , . . . , ê n is a basis <strong>of</strong> R n which is oriented<br />

depending on the ordering <strong>of</strong> the facets. We will assume the ordering F i1 , . . . , F in<br />

to be such that ê 1 , . . . , ê n is positively oriented.<br />

For each vertex v, let Λ (v) be the matrix<br />

Λ (v) = [λ i1 . . . λ in ]<br />

whose columns are ordered as described above. Let σ(v) := detΛ (v) . Then we obtain<br />

the following formula for the top Chern number,<br />

c n (X ) = Σ v<br />

1<br />

σ(v) . (3.11.1)<br />

Remark 3.11.1. If the stable almost complex structure <strong>of</strong> an omnioriented quasi<strong>toric</strong><br />

orbifold admits a reduction to an almost complex structure, then σ(v) is positive for each<br />

vertex v. This follows from comparing orientations, taking X to be oriented according to<br />

the almost complex structure. The converse is true in the case <strong>of</strong> quasi<strong>toric</strong> manifolds,<br />

see subsection 5.4.2 <strong>of</strong> [BP02]. The orbifold case remains unsolved at the moment.<br />

3.12 Chen-Ruan cohomology groups<br />

We refer the reader to [CR04, ALR07] for definition and motivation <strong>of</strong> the Chen-Ruan<br />

cohomology groups <strong>of</strong> an almost complex orbifold. They may roughly be thought <strong>of</strong> as a<br />

receptacle for a suitable Chern character from orbifold or equivariant K-theory. Briefly,<br />

the Chen-Ruan cohomology with coefficients in Q is the direct sum <strong>of</strong> the cohomology<br />

<strong>of</strong> the underlying space and the cohomology <strong>of</strong> certain subspaces <strong>of</strong> it called twisted<br />

sectors which are counted with multiplicities and rational degree shifts depending on<br />

the orbifold structure. The verification <strong>of</strong> the statements below is straightforward.<br />

For an almost complex quasi<strong>toric</strong> orbifold X , each twisted sector is a T N -invariant<br />

subspace X(F ) as described in Section 3.3. The contribution <strong>of</strong> X(F ) is counted with<br />

multiplicity one less than the order <strong>of</strong> the group G F , corresponding to the nontrivial<br />

elements <strong>of</strong> G F . However the degree shift <strong>of</strong> these contributions depend on the particular<br />

element <strong>of</strong> G F to which the twisted sector corresponds. If<br />

g = (a + N(F )) ∈ G F<br />

where a ∈ N ∗ (F ), then the degree shift 2ι(g) can be calculated as follows.


Chapter 3: Quasi<strong>toric</strong> orbifolds 70<br />

as<br />

Suppose λ 1 , . . . , λ k is the defining basis <strong>of</strong> N(F ). Then a can be uniquely expressed<br />

k∑<br />

a = q i λ i<br />

i=1<br />

where each q i is a rational number in [0, 1), and ι(g) = ∑ k<br />

i=1 q i. Note that the rational<br />

homology and hence rational cohomology <strong>of</strong> X(F ) can be computed using its<br />

combinatorial model given in Section 3.3.


Chapter 4<br />

Small orbifolds over simple<br />

polytopes<br />

4.1 Introduction<br />

In this chapter we introduce some n-dimensional orbifolds on which there is a natural<br />

Z n−1<br />

2 action having a simple polytope as the orbit space. We call these orbifolds small<br />

orbifolds. Small orbifolds are closely related to the notion <strong>of</strong> small covers.<br />

We give the precise definition <strong>of</strong> small orbifold and show that they are smooth.<br />

We calculate the orbifold fundamental group <strong>of</strong> small orbifolds.<br />

We show that the<br />

universal orbifold cover <strong>of</strong> an n-dimensional (n > 2) small orbifold is diffeomorphic to<br />

R n . Theorem 4.3.4 shows that the space Z, constructed in Lemma 4.4 <strong>of</strong> [DJ91], is<br />

diffeomorphic to R n if there is an s-characteristic function (definition 4.2.1) <strong>of</strong> simple<br />

n-polytope. We compute the singular homology groups <strong>of</strong> small orbifold with integer<br />

coefficients. We establish a relation between the modulo 2 Betti numbers <strong>of</strong> a small<br />

orbifold and h-vector <strong>of</strong> the polytope. In the last section we discuss intersection theory<br />

<strong>of</strong> small orbifold and rewrite the Poincaré duality theorem for even dimensional small<br />

orbifold. We compute the singular cohomology groups and cohomology ring <strong>of</strong> even<br />

dimensional small orbifold.<br />

4.2 Definition and orbifold structure<br />

Let Q be a simple polytope <strong>of</strong> dimension n. Let F(Q) = {F i , i = 1, 2, . . . , m} be the set<br />

<strong>of</strong> facets <strong>of</strong> Q. Let V (Q) be the set <strong>of</strong> vertices <strong>of</strong> Q. We denote the underlying additive<br />

group <strong>of</strong> the vector space F n−1<br />

2 by Z n−1<br />

2 .<br />

Definition 4.2.1. A function ϑ : F(Q) → Z2 n−1 is called an s-characteristic function<br />

<strong>of</strong> the polytope Q if the facets F i1 , F i2 , . . . , F in intersect at a vertex <strong>of</strong> Q then the set<br />

71


Chapter 4: Small orbifolds over simple polytopes 72<br />

{ϑ i1 , ϑ i2 , . . . , ϑ ik−1 , ̂ϑ ik , ϑ ik+1 , . . . , ϑ in }, where ϑ i := ϑ(F i ), is a basis <strong>of</strong> F n−1<br />

2 over F 2 for<br />

each k, 1 ≤ k ≤ n. We call the pair (Q, ϑ) an s-characteristic pair.<br />

Here the symbol ̂ represents the omission <strong>of</strong> corresponding entry. We give examples<br />

<strong>of</strong> s-characteristic function in examples 4.2.4 and 4.2.5.<br />

Now we give the constructive definition <strong>of</strong> small orbifold using the s-characteristic<br />

pair (Q, ϑ). Let F be a face <strong>of</strong> the simple polytope Q <strong>of</strong> codimension k ≥ 1. Then<br />

F = F i1 ∩ F i2 ∩ . . . ∩ F ik ,<br />

for some facets F ij ∈ F(Q) containing F . Let G F be the subspace <strong>of</strong> F2 n−1 spanned by<br />

{ϑ i1 , ϑ i2 , . . . , ϑ ik }. Without any confusion we denote the underlying additive group <strong>of</strong><br />

the subspace G F by G F . By the definition <strong>of</strong> ϑ, G v = Z n−1<br />

2 for each v ∈ V (Q). So<br />

the s-characteristic function ϑ determines a unique subgroup <strong>of</strong> Z n−1<br />

2 associated to each<br />

face <strong>of</strong> the polytope Q. Note that if k < n then G F<br />

∼ = Z<br />

k<br />

2 . The subgroup G F <strong>of</strong> Z2<br />

n−1<br />

is a direct summand.<br />

Each point p <strong>of</strong> Q belongs to relative interior <strong>of</strong> a unique face F (p) <strong>of</strong> Q. Define an<br />

equivalence relation ∼ s on Z n−1<br />

2 × Q by<br />

(a, p) ∼ s (b, q) if p = q and b − a ∈ G F (p) . (4.2.1)<br />

Let X(Q, ϑ) = (Z n−1<br />

2 × Q)/ ∼ s be the quotient space. Then X(Q, ϑ) is a Z n−1<br />

2 -space<br />

with the orbit map<br />

π : X(Q, ϑ) → Q defined by π([a, p] ∼ s<br />

) = p. (4.2.2)<br />

Let ˆπ : (Z n−1<br />

2 × Q) → X(Q, ϑ) be the quotient map. Let B n be the open ball <strong>of</strong> radius<br />

1 in R n .<br />

We claim that the space X(Q, ϑ) has a smooth orbifold structure. To prove our claim<br />

we construct a smooth orbifold atlas. We show that for each vertex v <strong>of</strong> Q there exists<br />

an orbifold chart (B n , Z 2 , φ v ) <strong>of</strong> X(Q, ϑ) where φ v (B n ) is an open subset X v (Q, ϑ) <strong>of</strong><br />

X(Q, ϑ) and {X v (Q, ϑ) : v ∈ V (Q)} cover X(Q, ϑ). To show the compatibility <strong>of</strong> these<br />

charts as v varies over V (Q), we introduce some additional orbifold charts to make this<br />

collection an orbifold atlas.<br />

Let v ∈ V (Q) and U v be the open subset <strong>of</strong> Q obtained by deleting all faces <strong>of</strong> Q<br />

not containing v. Let<br />

X v (Q, ϑ) := π −1 (U v ) = (Z n−1<br />

2 × U v )/ ∼ s .


73 4.2 Definition and orbifold structure<br />

The subset U v is diffeomorphic as manifold with corners to<br />

B n 1 = {x = (x 1 , x 2 , . . . , x n ) ∈ R n ≥0 : Σ n 1 x j < 1}. (4.2.3)<br />

Let f v : B n 1 → U v be a diffeomorphism. Suppose the facets<br />

{x 1 = 0} ∩ B n 1 , {x 2 = 0} ∩ B n 1 , . . . , {x n = 0} ∩ B n 1<br />

<strong>of</strong> B n 1 map to the facets F i 1<br />

, F i2 , . . . , F in <strong>of</strong> U v respectively under the diffeomorphism<br />

f v . Then F i1 ∩ F i2 ∩ . . . ∩ F in = v. Define an equivalence relation ∼ 0 on the product<br />

Z n−1<br />

2 × B1 n by (a, x) ∼ 0 (b, y) if x = y and b − a ∈ G F (fv(x)). (4.2.4)<br />

Let Y 0 = (Z n−1<br />

2 × B n 1 )/ ∼ 0 be the quotient space with the orbit map π 0 : Y 0 → B n 1 . Let<br />

ˆπ 0 : Z n−1<br />

2 × B n 1 → Y 0 be the quotient map. The diffeomorphism<br />

id × f v : Z n−1<br />

2 × B n 1 → Z n−1<br />

2 × U v<br />

descends to the following commutative diagram.<br />

Z n−1<br />

2 × B1<br />

n ⏐ ⏐↓<br />

ˆπ 0<br />

id×f v<br />

−−−−→ Z<br />

n−1<br />

2 × U v<br />

⏐ ⏐↓<br />

ˆπ v<br />

(4.2.5)<br />

Y 0<br />

ˆfv<br />

−−−−→ Xv (Q, ϑ).<br />

Here ˆπ v is the map ˆπ restricted to Z n−1<br />

2 × U v . It is easy to observe that the map<br />

ˆf v is a bijection. Since the maps ˆπ v and ˆπ 0 are continuous and the map id × f v is a<br />

diffeomorphism, the map ˆf v is a homeomorphism.<br />

Let u ∈ [0, 1) and H u be the hyperplane {Σ n 1 x j = u} in R n . Then Q 0 = H 0 ∩ B n 1 is<br />

the origin <strong>of</strong> R n and<br />

Q u = H u ∩ B n 1<br />

is an (n − 1)-simplex for each u ∈ (0, 1). When u ∈ (0, 1), the facets <strong>of</strong> Q u are<br />

{F uj := {x j = 0} ∩ Q u ; j = 1, 2, . . . , n}.<br />

The map<br />

ϑ u : {F uj : j = 1, . . . , n} → Z n−1<br />

2 defined by ϑ u (F uj ) = ϑ ij (4.2.6)<br />

satisfies the following condition.<br />

If F u is the intersection <strong>of</strong> unique l (0 ≤ l ≤ n − 1) facets F uj1 , . . . , F ujl <strong>of</strong> Q u<br />

then the vectors ϑ u (F uj1 ), . . . , ϑ u (F ujl ) are linearly independent vectors <strong>of</strong> F n−1<br />

2 .


Chapter 4: Small orbifolds over simple polytopes 74<br />

Hence ϑ u is a Z 2 -characteristic function (see definition 2.2.7) <strong>of</strong> a small cover over the<br />

polytope Q u . Since Q u is an (n − 1)-simplex, the small cover corresponding to the Z 2 -<br />

characteristic pair (Q u , ϑ u ) is equivariantly homeomorphic to the real projective space<br />

RP n−1 , see Chapter 2. Here consider RP n−1 as the identification space<br />

{B n−1 /{x = −x} : x ∈ ∂B n−1 }.<br />

So at each point (u, 0, . . . , 0) ∈ B n 1<br />

− {0} we get an equivariant homeomorphism<br />

(Z n−1<br />

2 × Q u )/ ∼ 0<br />

∼ = RP n−1 , (4.2.7)<br />

which sends the fixed point [a, u] ∼ 0<br />

to the origin <strong>of</strong> B n−1 . It is clear from the definition<br />

<strong>of</strong> the equivalence relation ∼ 0 that at (0, . . . , 0) ∈ B n 1 , (Zn−1 2 × Q 0 )/ ∼ 0 is a point.<br />

Hence Y 0 is equivariantly homeomorphic to the open cone<br />

(RP n−1 × [0, 1))/RP n−1 × {0}<br />

on real projective space RP n−1 . Consider the following map<br />

S n−1 × [0, 1) → B n define by ((x 1 , x 2 , . . . , x n ), r) → (rx 1 , rx 2 , . . . , rx n ).<br />

This map induces a homeomorphism f : B n → (S n−1 ×[0, 1))/S n−1 ×{0}. The covering<br />

map S n−1 → RP n−1 induces a projection map<br />

φ 0 : (S n−1 × [0, 1))/S n−1 × {0} → (RP n−1 × [0, 1))/RP n−1 × {0}.<br />

Observe that this projection map φ 0 is nothing but the orbit map q <strong>of</strong> the antipodal<br />

action <strong>of</strong> Z 2 on B n . In other words the following diagram is commutative.<br />

B n<br />

q<br />

⏐<br />

↓<br />

f<br />

−−−−→ (S n−1 × [0, 1))/S n−1 × {0}<br />

⏐<br />

φ ⏐↓<br />

0<br />

(4.2.8)<br />

B n /Z 2<br />

ˆf<br />

−−−−→ (RP n−1 × [0, 1))/RP n−1 × {0}<br />

Since the map φ 0 is induced from the antipodal action on S n−1 the commutativity <strong>of</strong><br />

the diagram ensure that the map ˆf is a homeomorphism. Let φ v be the composition <strong>of</strong><br />

the following maps.<br />

B n<br />

q<br />

−−−−→ B n /Z 2<br />

ˆf<br />

−−−−→ (RP n−1 × [0, 1))/RP n−1 × {0} ∼ = Y 0<br />

ˆ fv<br />

−−−−→ Xv (Q, ϑ).<br />

Hence (B n , Z 2 , φ v ) is an orbifold chart <strong>of</strong> X v (Q, ϑ) corresponding to the vertex v <strong>of</strong> the<br />

polytope.


75 4.2 Definition and orbifold structure<br />

Now we introduce some additional orbifold charts corresponding to each face F <strong>of</strong><br />

codimension-k (0 < k < n) and the interior <strong>of</strong> polytope Q. Let<br />

U F = ∩ U v ,<br />

where the intersection is over all vertices v <strong>of</strong> F . Let X F (Q, ϑ) := π −1 (U F ). Fix a<br />

vertex v <strong>of</strong> F . Consider the diffeomorphism f v : B n 1 → U v. Observe that U F can be<br />

obtained from U v by deleting unique n − k facets <strong>of</strong> U v . Let F l1 , . . . , F ln−k be the facets<br />

<strong>of</strong> U v such that<br />

U F = U v − {F l1 ∪ . . . ∪ F ln−k },<br />

where {l 1 , . . . , l n−k } ⊂ {1, 2, . . . , n}. Let B n F = f −1<br />

v (U F ). Let {x l1 = 0}, . . . , {x ln−k = 0}<br />

be the coordinate hyperplanes in R n such that<br />

f v ({x l1 = 0} ∩ B n 1 ) = F l1 , . . . , f v ({x ln−k = 0} ∩ B n 1 ) = F ln−k .<br />

So B n F = Bn 1 − {{x l 1<br />

= 0} ∪ . . . ∪ {x ln−k = 0}}. Then ˆf v (π −1<br />

0 (B F )) = X F (Q, ϑ).<br />

Let u ∈ (0, 1) and Q ′ u = Q u − {x l1 = 0}. Since (Q u , ϑ u ) is a Z 2 -characteristic pair,<br />

there exist an equivariant homeomorphism from (Z n−1<br />

2 × Q ′ u)/ ∼ 0 to B n−1 ⊂ R n−1 such<br />

that (Z n−1<br />

2 × F uj )/ ∼ 0 maps to a coordinate hyperplane H j := {x ij = 0} ∩ B n−1 , for<br />

j ∈ {{1, 2, . . . , n} − l 1 }. Clearly H i ≠ H j for i ≠ j.<br />

Let Q ′′ u = Q ′ u − {{x l2 = 0} ∪ . . . ∪ {x ln−k = 0}}. Then<br />

(Z n−1<br />

2 × Q ′′ u)/ ∼ 0<br />

∼ = B n−1 − {H l2 ∪ . . . ∪ H ln−k } and B n F ∼ = (0, 1) × Q ′′ .<br />

So π −1<br />

0 (Bn F ) = (Zn−1 2 × B n F )/ ∼ 0 is homeomorphic to<br />

(0, 1) × {(Z n−1<br />

2 × Q ′′ u)/ ∼ 0 } ∼ = (0, 1) × {B n−1 − {H l2 ∪ . . . ∪ H ln−k }}.<br />

By our assumption<br />

(0, 1) × {B n−1 − {H l2 ∪ . . . ∪ H ln−k }} ↩→ (RP n−1 × [0, 1))/RP n−1 × {0}.<br />

So there exist two open subsets D F , D<br />

F ′ <strong>of</strong> Bn such that D<br />

F ′ = −D F and the following<br />

restrictions are homeomorphism.<br />

1. φ 0 ◦ f| DF : D F → (0, 1) × {B n−1 − {H l2 ∪ . . . ∪ H ln−k }}.<br />

2. φ 0 ◦ f| D ′<br />

F<br />

: D<br />

F ′ → (0, 1) × {Bn−1 − {H l2 ∪ . . . ∪ H ln−k }}.<br />

Hence the restriction φ v | DF<br />

: D F → X F (Q, ϑ) is homeomorphism. Clearly<br />

D F<br />

∼ = {{B n ∩ {x n > 0}} − ∪ (n−k−1)<br />

j=1,x lj ≠x n<br />

{x lj = 0}}. (4.2.9)


Chapter 4: Small orbifolds over simple polytopes 76<br />

The set D F is homeomorphic to an open ball in R n if k = n − 1. When k = n − 1,<br />

F is an edge <strong>of</strong> the polytope Q. Let E(Q) be the set <strong>of</strong> edges <strong>of</strong> the polytope Q and<br />

e ∈ E(Q). Let φ ev = φ v | De : D e → X e (Q, ϑ), where v ∈ V (e). Hence (D e , {0}, φ ev ) is<br />

an orbifold chart on X e (Q, ϑ) for each e ∈ E(Q) and v ∈ V (e).<br />

The set D F is disjoint union <strong>of</strong> open sets {A F (i) : i = 1, . . . , 2(n − k − 1)} in R n<br />

whenever 0 < k < n − 1. Here all A F (i) are homeomorphic to an open ball in R n . Let<br />

φ Fv(i) = φ v | AF (i)<br />

: A F (i) → X F (Q, ϑ) (4.2.10)<br />

be the restriction <strong>of</strong> the map φ v to the domain A F (i) , where v ∈ V (F ). So for each<br />

(i, v) ∈ {1, 2, . . . , 2(n − k − 1)} × V (F ), the triple (A F (i) , {0}, φ Fv (i)) is an orbifold chart<br />

on the image <strong>of</strong> φ Fv (i) in X F (Q, ϑ) ⊆ X(Q, ϑ).<br />

Let Q 0 be the interior <strong>of</strong> Q and X Q (Q, ϑ) = π −1 (Q 0 ). Hence<br />

D Q := {{B n ∩ {x n > 0}} − ∪ n−1<br />

j=1 {x j = 0}}<br />

is homeomorphic to X Q (Q, ϑ) under the restriction <strong>of</strong> φ v on D Q . The set D Q is disjoint<br />

union <strong>of</strong> connected open sets {B j : j = 1, . . . , 2(n − 1)} in R n where each B j is<br />

homeomorphic to the open ball B n . Let<br />

φ Qv (j) = φ v | Bj : B j → X Q (Q, ϑ) (4.2.11)<br />

be the restriction <strong>of</strong> the map φ v to the domain B j . Hence for (j, v) ∈ {1, . . . , 2(n−1)}×<br />

V (Q), (B j , {0}, φ Qv(j)) is an orbifold chart on the image <strong>of</strong> φ Qv(j) in X Q (Q, ϑ). Let<br />

U ′ = {(B n , Z 2 , φ v )} ∪ {(D e , {0}, φ ev )} ∪ {(A F (i) , {0}, φ Fv(i))} ∪ {(B j , {0}, φ Qv(j))}<br />

(4.2.12)<br />

where v ∈ V (Q), e ∈ E(Q), F run over the faces <strong>of</strong> codimension k (0 < k < n − 1),<br />

i = 1, . . . , 2(n − k − 1) and j = 1, . . . , 2(n − 1).<br />

From the description <strong>of</strong> orbifold charts corresponding to each faces and interior<br />

<strong>of</strong> polytope it is clear that the collection U ′ is an orbifold atlas on X(Q, ϑ). Clearly<br />

the inclusions D e ↩→ B n , A F (i) ↩→ B n and B j ↩→ B n induce the following smooth<br />

embeddings respectively:<br />

(D e , {0}, φ ev ) ↩→ (B n , Z 2 , φ v ), (A F (i) , {0}, φ Fv (i)) ↩→ (B n , Z 2 , φ v )<br />

and (B j , {0}, φ Qv (j)) ↩→ (B n , Z 2 , φ v ).<br />

So U ′ is a part <strong>of</strong> a maximal atlas U for X(Q, ϑ). Thus X (Q, ϑ) = (X, U) is a smooth<br />

n-dimensional orbifold.<br />

Definition 4.2.2. We call the smooth orbifold X (Q, ϑ) small orbifold corresponding to


77 4.2 Definition and orbifold structure<br />

the s-characteristic pair (Q, ϑ).<br />

Remark 4.2.3. 1. The small orbifold X (Q, ϑ) is reduced, that is, the group in each<br />

chart has an effective action. Singular set <strong>of</strong> the orbifold X (Q, ϑ) is<br />

ΣX (Q, ϑ) = {[t, v] ∼ s<br />

∈ X(Q, ϑ) : v ∈ V (Q)}.<br />

We call an element <strong>of</strong> ΣX (Q, ϑ) an orbifold point <strong>of</strong> X(Q, ϑ).<br />

2. We can not define an s-characteristic function for an arbitrary polytope. Later we<br />

will see some examples.<br />

3. The small orbifold X(Q, ϑ) is compact and connected.<br />

Example 4.2.4. Let Q 2 be a simple 2-polytope in R 2 . Define<br />

ϑ : F(Q 2 ) → Z 2 by ϑ(F ) = 1, ∀F ∈ F(Q 2 ). (4.2.13)<br />

So ϑ is the s-characteristic function <strong>of</strong> Q 2 . The resulting quotient space X(Q 2 , ϑ) is<br />

homeomorphic to the sphere S 2 . These are the only cases where the identification space<br />

is a manifold.<br />

(0, 1)<br />

v 8<br />

v 7<br />

v 4<br />

v 6<br />

v 3<br />

(1, 0)<br />

(1, 1)<br />

v 5<br />

v 1 v 2<br />

(1, 1)<br />

(1, 0) (0, 1)<br />

Figure 4.1: An s-characteristic function <strong>of</strong> I 3 .<br />

Example 4.2.5. Let I 3 = {(x, y, z) ∈ R 3 : 0 ≤ x, y, z ≤ 1} be the standard cube in R 3 .<br />

Let v 1 , . . . , v 8 be the vertices <strong>of</strong> I 3 , see Figure 4.1. So the facets <strong>of</strong> I 3 are the following<br />

squares<br />

F 1 = v 1 v 2 v 3 v 4 , F 2 = v 1 v 2 v 6 v 5 , F 3 = v 1 v 5 v 8 v 4 , F 4 = v 2 v 6 v 3 v 7 ,<br />

F 5 = v 4 v 3 v 7 v 8 and F 6 = v 5 v 6 v 7 v 8 .<br />

Define ϑ : F(I 3 ) → Z 2 2 by<br />

ϑ(F 1 ) = ϑ(F 6 ) = (1, 0), ϑ(F 2 ) = ϑ(F 5 ) = (0, 1), ϑ(F 3 ) = ϑ(F 4 ) = (1, 1).


Chapter 4: Small orbifolds over simple polytopes 78<br />

Hence ϑ is an s-characteristic function <strong>of</strong> I 3 . Then<br />

G F1 = G F6 = {(0, 0), (1, 0)}, G F2 = G F5 = {(0, 0), (0, 1)}, G F3 = G F4 = {(0, 0), (1, 1)}.<br />

For other proper face F <strong>of</strong> I 3 , G F = Z 2 2 . Hence X (I3 , ϑ) is a 3-dimensional small<br />

orbifold.<br />

Observation 4.2.6. Let F be a codimension-k (0 < k < n − 1) face <strong>of</strong> Q. Then F is<br />

a simple polytope <strong>of</strong> dimension n − k. Let F(F ) = {F j ′ 1<br />

, . . . , F j ′ l<br />

} be the set <strong>of</strong> facets <strong>of</strong><br />

F . So there exist unique facets F j1 , . . . , F jl <strong>of</strong> Q such that<br />

F j1 ∩ F = F ′ j 1<br />

, . . . , F jl ∩ F = F ′ j l<br />

.<br />

Fix an isomorphism b from the quotient field F n−1<br />

2 /G F to F n−1−k<br />

2 . Define a function<br />

ϑ ′ : F(F ) → Z n−1−k<br />

2 by ϑ ′ (F ′ j m<br />

) = b(ϑ jm + G F ).<br />

Observe that the function ϑ ′ is an s-characteristic function <strong>of</strong> F . Let ∼ ′ s be the restriction<br />

<strong>of</strong> ∼ s on Z n−1−k<br />

2 × F . So X (F, ϑ ′ ) is an (n − k)-dimensional smooth small orbifold<br />

associated to the s-characteristic pair (F, ϑ ′ ). The orbifold X (F, ϑ ′ ) is a suborbifold <strong>of</strong><br />

X (Q, ϑ). We have shown that for each edge e <strong>of</strong> Q, the set X e (Q, ϑ) is homeomorphic<br />

to the open ball B n . Let e ′ be an edge <strong>of</strong> F and U ′ e ′ = U e ′ ∩ F . Hence<br />

is homeomorphic to the open ball B n−k .<br />

W e ′ = (Z n−1−k<br />

2 × U ′ e ′)/ ∼′ s= (Z n−1<br />

2 × U ′ e ′)/ ∼ s<br />

4.3 Orbifold fundamental group<br />

Let X (Q, ϑ) be a small orbifold over simple n-polytope Q. The set <strong>of</strong> smooth points<br />

M(Q, ϑ) := X(Q, ϑ) − ΣX (Q, ϑ)<br />

<strong>of</strong> small orbifold X (Q, ϑ) is an n-dimensional manifold. For each v ∈ V (Q) we have<br />

X v (Q, ϑ) − [0, v] ∼ s ∼ = RP n−1 × I 0 .<br />

The sphere S n−1 is the double sheeted universal cover <strong>of</strong> RP n−1 . So the universal cover<br />

<strong>of</strong> X v (Q, ϑ) − [0, v] ∼s is S n−1 × I 0 ∼ = B n − 0. Actually the map φ v : B n → X v (Q, ϑ) is<br />

the orbifold universal covering. Let e be an edge containing the vertex v <strong>of</strong> Q. Define<br />

ē := e ∩ U v .<br />

Identifying the faces containing the edge ē <strong>of</strong> U v according to the equivalence relation


79 4.3 Orbifold fundamental group<br />

∼ s we get the quotient space Xē(U v , ϑ) homeomorphic to<br />

Be n := {(x 1 , x 2 , . . . , x n ) ∈ B n : x n ≥ 0}.<br />

The set X v (Q, ϑ) is obtained from Xē(U v , ϑ) by identifying the antipodal points <strong>of</strong><br />

the boundary <strong>of</strong> Xē(U v , ϑ) around the fixed point [0, v] ∼ s<br />

. Identifying two copies <strong>of</strong><br />

Xē(U v , ϑ) along their boundary via the antipodal map on the boundary we get a space<br />

homeomorphic to B n .<br />

Doing these identification associated to the orbifold points we obtain that the universal<br />

cover <strong>of</strong> M(Q, ϑ) is homeomorphic to R n − N for some infinite subset N <strong>of</strong> Z n<br />

where N depends on the polytope Q in R n . Let<br />

ζ : R n − N → M(Q, ϑ) (4.3.1)<br />

be the universal covering map. This map locally resemble the chart maps.<br />

The chart maps φ v : B n → X v (Q, ϑ) are uniformly continuous on a compact neighborhood<br />

<strong>of</strong> 0 ∈ B n and Q is an n-polytope in R n . So for each x ∈ N there exists a<br />

neighborhood V x ⊂ R n <strong>of</strong> x such that the restriction <strong>of</strong> the universal covering map ζ<br />

on V x − x is uniformly continuous. Hence the map ζ has a unique extension, say ˆζ, on<br />

their metric completion. The metric completion <strong>of</strong> R n − N and M(Q, ϑ) are R n and<br />

X(Q, ϑ) respectively. The map ˆζ sends N onto V (Q).<br />

We show the map ˆζ is an orbifold covering. Let ϱ : Z → X (Q, ϑ) be an orbifold<br />

cover. Then the restriction ϱ : Z −ΣZ → M(Q, ϑ) is an honest cover. Hence there exist<br />

a covering map ζ ϱ : R n − N → Z − ΣZ so that the following diagram is commutative.<br />

R n − N<br />

⏐<br />

ζ↓<br />

M(Q, ϑ)<br />

ζ ϱ<br />

−−−−→ Z − ΣZ<br />

⏐<br />

ϱ↓<br />

id<br />

−−−−→ M(Q, ϑ)<br />

(4.3.2)<br />

Since the map ζ is locally uniformly continuous and the maps ζ ϱ , ϱ are continuous,<br />

all the maps in the diagram 4.3.2 can be extended to their metric completion. That is<br />

we get a commutative diagram <strong>of</strong> orbifold coverings.<br />

R n<br />

ˆζ<br />

⏐<br />

↓<br />

X (Q, ϑ)<br />

ˆζϱ<br />

−−−−→<br />

Z<br />

⏐<br />

↓<br />

id<br />

−−−−→ X (Q, ϑ)<br />

ˆϱ<br />

(4.3.3)<br />

Hence ˆζ : R n → X (Q, ϑ) is an orbifold universal cover <strong>of</strong> X (Q, ϑ). Since the orbit map<br />

<strong>of</strong> antipodal action is smooth, the map ˆζ is a smooth map. Thus we get the following.


Chapter 4: Small orbifolds over simple polytopes 80<br />

Theorem 4.3.1. The universal orbifold cover <strong>of</strong> an n-dimensional small orbifold is<br />

diffeomorphic to R n .<br />

v 3<br />

v 2<br />

v 1<br />

v 4<br />

x<br />

y<br />

−x<br />

v 3<br />

v 7<br />

v v 7<br />

8 v 2<br />

−y<br />

v 3 v 3<br />

v 6<br />

v 5 v 6<br />

v 4<br />

v 2<br />

v 2<br />

v 1<br />

Figure 4.2: Identification <strong>of</strong> faces containing the edge v 5 v 8 <strong>of</strong> I 3 .<br />

Example 4.3.2. Recall the small orbifold X(I 3 , ϑ) <strong>of</strong> example 4.2.5. The set <strong>of</strong> smooth<br />

points<br />

M(I 3 , ϑ) := X(I 3 , ϑ) − ΣX (I 3 , ϑ)<br />

is a 3-dimensional manifold. The universal cover <strong>of</strong> M(I 3 , ϑ) is homeomorphic to R 3 −<br />

Z 3 . To show this we need to observe how the faces <strong>of</strong> Z 2 2 × I3 are identified by the<br />

equivalence relation ∼ s (see equation 4.2.1) on Z 2 2 × I3 . For each v ∈ V (I 3 )<br />

X v (I 3 , ϑ) − [a, v] ∼s ∼ = RP 2 × I 0 .<br />

The sphere S 2 is the double sheeted universal cover <strong>of</strong> RP 2 . So the universal cover <strong>of</strong><br />

X v (I 3 , ϑ) − [a, v] ∼ s<br />

is S 2 × I 0 ∼ = B 3 − 0. Hence the identification <strong>of</strong> faces around each<br />

vertex <strong>of</strong> I 3 tells us that the universal cover <strong>of</strong> M(I 3 , ϑ) is R 3 − Z 3 . We illustrate the<br />

identification <strong>of</strong> faces by the Figure 4.2, where x ∼ s −x on the upper face and y ∼ s −y<br />

on the lower face in that figure.<br />

We use the observation 2.5.6 to compute the orbifold fundamental group <strong>of</strong> X (Q, ϑ).<br />

Let {β 1 , β 2 , . . . , β m } be the standard basis <strong>of</strong> Z m 2 . Define a map β : F(Q) → Zm 2 by<br />

β(F j ) = β j . For each face F = F j1 ∩ F j2 ∩ . . . ∩ F jl , let H F be the subgroup <strong>of</strong> Z m 2<br />

generated by β j1 , β j2 , . . . , β jl . Define an equivalence relation ∼ β on Z m 2 × Q by<br />

(s, p) ∼ β (t, q) if and only if p = q and t − s ∈ H F<br />

where F ⊂ Q is the unique face whose relative interior contains p. So the quotient space<br />

N(Q, β) = (Z m 2 × Q)/ ∼ β is an n-dimensional smooth manifold. N(Q, β) is a Z m 2 -space


81 4.3 Orbifold fundamental group<br />

with the orbit map<br />

π u : N(Q, β) → Q defined by π u ([s, p] ∼ β<br />

) = p.<br />

We show Q has a smooth orbifold structure. Recall the open subset U v <strong>of</strong> Q associated<br />

to each vertex v ∈ V (Q). Note that open sets {U v : v ∈ V (Q)} cover Q. Let d be<br />

the Euclidean distance in R n . Let F i1 , F i2 , . . . , F in be the facets <strong>of</strong> Q such that v is the<br />

intersection <strong>of</strong> F i1 , F i2 , . . . , F in . For each p ∈ U v , let<br />

x j (p) = d(p, F ij ), for all j = 1, 2, . . . , n.<br />

Let B n v = {(x 1 (p), . . . , x n (p)) ∈ R n ≥0 : p ∈ U v}. So the map<br />

f : U v → B n v defined by p → (x 1 (p), . . . , x n (p))<br />

gives a diffeomorphism from U v to B n v . Consider the standard action <strong>of</strong> Z n 2 on Rn with<br />

the orbit map<br />

ξ : R n → R n ≥0.<br />

Then ξ −1 (Bv n ) is diffeomorphic to B n . Hence (ξ −1 (Bv n ), f −1 ◦ξ, Z n 2 ) is a smooth orbifold<br />

chart on U v . To show the compatibility <strong>of</strong> these charts as v varies over V (Q), we<br />

can introduce some additional smooth orbifold charts to make this collection a smooth<br />

orbifold atlas. From the definition <strong>of</strong> ∼ s it is clear that π : X(Q, ϑ) → Q is a smooth<br />

orbifold covering.<br />

Definition 4.3.3. Let L be the simplicial complex dual to Q. The right-angled Coxeter<br />

group Γ associated to Q is the group with one generator for each element <strong>of</strong> V (L) and<br />

relations between generators are the following; a 2 = 1 for all a ∈ V (L), (ab) 2 = 1 if<br />

{a, b} ∈ E(L).<br />

For each p ∈ Q, let F (p) ⊂ Q be the unique face containing p in its relative interior.<br />

Let F (p) = F j1 ∩. . .∩F jl . Let a j1 , . . . , a jl be the vertices <strong>of</strong> L dual to F j1 , . . . , F jl respectively.<br />

Let Γ F (p) be the subgroup generated by a j1 , . . . , a jl<br />

relation ∼ c on Γ × Q by<br />

(g, p) ∼ c (h, q) if p = q and h −1 g ∈ Γ F (p) .<br />

<strong>of</strong> Γ. Define an equivalence<br />

Let Y = (Γ × Q)/ ∼ c be the quotient space. We follow this construction from [DJ91].<br />

So Y is a Γ-space with the orbit map<br />

ξ Γ : Y → Q defined by ξ Γ ([g, p] ∼ c<br />

) = p. (4.3.4)<br />

Then Y is an n-dimensional manifold and ξ Γ is an orbifold covering. Since each facet


Chapter 4: Small orbifolds over simple polytopes 82<br />

is connected, whenever two generators <strong>of</strong> Γ commute the intersection <strong>of</strong> corresponding<br />

facets <strong>of</strong> Q is nonempty. From Theorems 10.1 and 13.5 <strong>of</strong> [Dav83], we get that Y is<br />

simply connected. Hence ξ Γ is a universal orbifold covering and the orbifold fundamental<br />

group <strong>of</strong> Q is Γ.<br />

Let H be the kernel <strong>of</strong> abelianization map Γ → Γ ab . The group H acts on Y freely<br />

and properly discontinuously. So the orbit space Y/H is a manifold. The space Y/H is<br />

called the universal abelian cover <strong>of</strong> Q. Note that N(Q, β) = Y/H. Let<br />

be the corresponding orbit map.<br />

ξ β : Y → N(Q, β) (4.3.5)<br />

Define a function ¯ϑ : Z m 2 → Zn−1 2 by ¯ϑ(β j ) = ϑ(F j ) = ϑ j on the basis <strong>of</strong> Z m 2 . So ¯ϑ<br />

is a linear surjection. ¯ϑ induces a surjection<br />

˜ϑ : N(Q, β) → X(Q, ϑ) defined by ˜ϑ([s, p] ∼ β<br />

) = [s, p] ∼c . (4.3.6)<br />

That is the following diagram commutes.<br />

N(Q, β)<br />

⏐ ⏐↓<br />

ˆπ u<br />

˜ϑ<br />

−−−−→ X(Q, ϑ)<br />

⏐ ⏐↓ ˆπ<br />

(4.3.7)<br />

Q<br />

id<br />

−−−−→<br />

Q<br />

From this commutative diagram we get ˜ϑ is a smooth orbifold covering <strong>of</strong> X(Q, ϑ).<br />

Hence the composition map<br />

˜ϑ ◦ ξ β : Y → X(Q, ϑ)<br />

is a smooth universal orbifold covering. From [Thu3m] and Theorem 4.3.1 we obtain<br />

the following necessary condition for existence <strong>of</strong> an s-characteristic function.<br />

Theorem 4.3.4. Let ϑ : F(Q) → Z n−1<br />

2 be an s-characteristic function <strong>of</strong> the n-polytope<br />

Q (n > 2). Then the space Y is diffeomorphic to R n .<br />

Note that when Q is an n-simplex, Y is homeomorphic to the n-dimensional sphere<br />

S n . So by this theorem there does not exist an s-characteristic function <strong>of</strong> n-simplex.<br />

Consequently there does not exist any small orbifold with the n-dimensional simplex as<br />

orbit space when n > 2.<br />

Let ξ ϑ be the following composition map<br />

Γ → Γ ab<br />

¯ϑ<br />

−−−−→ Z n−1<br />

2 . (4.3.8)<br />

Clearly Ker(ξ ϑ ), kernel <strong>of</strong> ξ ϑ , acts on Y with the orbit map ˜ϑ ◦ ξ β . Now using the<br />

observation 2.5.6, we get the following corollary.


83 4.4 Homology and Euler characteristic<br />

Corollary 4.3.5. The orbifold fundamental group <strong>of</strong> X(Q, ϑ) is ker(ξ ϑ ) which is a<br />

normal subgroup <strong>of</strong> the right-angled Coxeter group associated to the polytope Q.<br />

4.4 Homology and Euler characteristic<br />

To calculate the singular homology groups <strong>of</strong> small orbifold X(Q, ϑ) we will construct<br />

a CW -structure on these orbifolds and describe how the cells are attached. We redefine<br />

index function precisely.<br />

functional<br />

Realize Q as a convex polytope in R n and choose a linear<br />

φ : R n → R (4.4.1)<br />

which distinguishes the vertices <strong>of</strong> Q, as in the pro<strong>of</strong> <strong>of</strong> Theorem 3.1 in [DJ91]. The<br />

vertices are linearly ordered according to ascending value <strong>of</strong> φ. We make the 1-skeleton<br />

<strong>of</strong> Q into a directed graph by orienting each edge such that φ increases along edges. For<br />

each vertex <strong>of</strong> Q define its index ind Q (v), as the number <strong>of</strong> incident edges that point<br />

towards v.<br />

Definition 4.4.1. A subset ˆQ ⊆ Q <strong>of</strong> dimension k is called a proper subcomplex <strong>of</strong> Q<br />

if ˆQ is connected and ˆQ is the union <strong>of</strong> some k-dimensional faces <strong>of</strong> Q.<br />

In particular each face <strong>of</strong> Q is a proper subcomplex <strong>of</strong> Q. The 1-skeleton <strong>of</strong> a proper<br />

subcomplex ˆQ is a subcomplex <strong>of</strong> the 1-skeleton <strong>of</strong> Q. The restriction <strong>of</strong> φ on the 1-<br />

skeleton <strong>of</strong> ˆQ makes it a directed graph. We define index ind ˆQ(v) <strong>of</strong> each vertex v <strong>of</strong> ˆQ<br />

as the number <strong>of</strong> incident edges in ˆQ that point towards v. Let V ( ˆQ) and F( ˆQ) denote<br />

the set <strong>of</strong> vertices and the set <strong>of</strong> faces <strong>of</strong> ˆQ respectively. We construct a CW -structure<br />

on X(Q, ϑ) in the following. Let<br />

I Q = {(u, e u ) ∈ V (Q) × E(Q) : ind Q (u) = n and e u is the edge joining the vertices<br />

u, x u such that φ(u) > φ(x u ) > φ(x) for all x ∈ V (Q) − {u, x u }}.<br />

Let U eu = U u ∩ U xu and ˆQ n = Q. Then W eu = (Z2 n−1 × U eu )/ ∼ s is homeomorphic to<br />

the n-dimensional open ball B n ⊂ R n . Let<br />

ˆQ n−1 = Q − U eu . (4.4.2)<br />

Then ˆQ n−1 is the union <strong>of</strong> facets not containing the edge e u <strong>of</strong> Q.<br />

So ˆQ n−1 is an<br />

(n − 1)-dimensional proper subcomplex <strong>of</strong> Q and V (Q) = V ( ˆQ n−1 ). Let v ∈ V ( ˆQ n−1 )<br />

with ind ˆQn−1(v) = n − 1. Let Fv<br />

n−1 ∈ F( ˆQ n−1 ) be the smallest face which contains the<br />

inward pointing edges incident to v in ˆQ n−1 . If v 1 , v 2 are two vertices <strong>of</strong> ˆQn−1 with


Chapter 4: Small orbifolds over simple polytopes 84<br />

ind ˆQn−1(v 1 ) = n − 1 = ind ˆQn−1(v 2 ) then F n−1<br />

v 1<br />

≠ Fv n−1<br />

2<br />

. Let<br />

I ˆQn−1 = {(v, e v ) ∈ V (Q) × E(Q) : ind ˆQn−1(v) = n − 1 and e v is the edge joining the<br />

vertices v, y v ∈ V (Fv<br />

n−1 ) such that φ(v) > φ(y v ) > φ(y) for all y ∈ V (Fv n−1 ) − {v, y v }}.<br />

Let<br />

U ev = U v ∩ U yv ∩ F n−1<br />

v<br />

for each (v, e v ) ∈ I ˆQn−1.<br />

From the observation 4.2.6, W ev = (Z n−1<br />

2 × U ev )/ ∼ s is homeomorphic to the (n − 1)-<br />

dimensional open ball B n−1 ⊂ R n−1 . Let<br />

ˆQ n−2 = Q − {{<br />

∪<br />

U eu } ∪ {<br />

∪<br />

U ev }}. (4.4.3)<br />

(u,e u )∈I ˆQn<br />

(v,e v )∈I ˆQn−1<br />

So ˆQ n−2 is an (n − 2)-dimensional proper subcomplex <strong>of</strong> Q and V (Q) = V ( ˆQ n−2 ). Let<br />

w ∈ V ( ˆQ n−2 ) with ind ˆQn−2(w) = n−2. Let Fw<br />

n−2 ∈ F( ˆQ n−2 ) be the smallest face which<br />

contains the inward pointing edges incident to w in ˆQ n−2 . If w 1 , w 2 are two vertices <strong>of</strong><br />

ˆQ n−2 with ind ˆQn−2(w 1 ) = n − 1 = ind ˆQn−2(w 2 ) then Fw n−2<br />

1<br />

≠ Fw n−2<br />

2<br />

. Let<br />

I ˆQn−2 = {(w, e w ) ∈ V (Q) × E(Q) : ind ˆQn−2(w) = n − 2 and e w is the edge joining the<br />

vertices w, z w ∈ F n−2<br />

w<br />

such that φ(w) > φ(z w ) > φ(z) for all z ∈ V (F n−2<br />

w ) − {w, z w }}.<br />

Let<br />

U ew = U w ∩ U zw ∩ F n−2<br />

w<br />

for each (w, e w ) ∈ I ˆQn−2.<br />

From the observation 4.2.6, W ew = (Z n−1<br />

2 × U ew )/ ∼ s is homeomorphic to the (n − 2)-<br />

dimensional open ball B n−2 ⊂ R n−2 .<br />

Continuing this process we observe that ˆQ 1 ( ∼ = (Z2 n−1 × ˆQ 1 )/ ∼ s ) is a maximal<br />

tree <strong>of</strong> the 1-skeleton <strong>of</strong> Q and ˆQ 0 = V (Q). Hence relative interior <strong>of</strong> each edge <strong>of</strong><br />

(Z n−1<br />

2 × ˆQ 1 )/ ∼ s is homeomorphic to the 1-dimensional ball in R. So corresponding to<br />

each edge <strong>of</strong> polytope Q, we construct a cell <strong>of</strong> dimension ≥ 1 <strong>of</strong> X(Q, ϑ).<br />

Recall the h-vectors h i <strong>of</strong> the simple polytope Q. The integer h n−i is the number<br />

<strong>of</strong> vertices v ∈ V (Q) with ind Q (v) = i. The Dehn-Sommerville relation is<br />

h i = h n−i ∀ i = 0, 1, . . . , n,<br />

see Theorem 1.20 <strong>of</strong> [BP02]. Hence the number <strong>of</strong> k-dimensional cells in X(Q, ϑ) is<br />

|I ˆQk| = Σ n k h i. (4.4.4)<br />

We describe the attaching map for a k-dimensional cell. Here k-dimensional cells


85 4.4 Homology and Euler characteristic<br />

are<br />

{W ev : (v, e v ) ∈ I ˆQk}.<br />

Let (v, e v ) ∈ I ˆQk. Let Fv k ∈ F( ˆQ k ) be the smallest face containing the inward pointing<br />

edges to v in ˆQ k . Define an equivalence relation ∼ ev on Z2 n−1 × Fv k by<br />

(t, p) ∼ ev (s, q) if p = q ∈ F ′ and s − t ∈ G F ′ (4.4.5)<br />

where F ′ ∈ F(Fv k ) is a face containing the edge e v . The quotient space (Z2 n−1 ×Fv k )/ ∼ ev<br />

is homeomorphic to the closure <strong>of</strong> open ball B k ⊂ R k . The attaching map φ F k v<br />

is the<br />

natural quotient map<br />

φ F k v<br />

: S k−1 ∼ = (Z<br />

n−1<br />

2 × (F k v − U ev )/ ∼ ev → (Z n−1<br />

2 × (F k v − U ev )/ ∼ s . (4.4.6)<br />

Let X k =<br />

k∪<br />

i=1<br />

∪<br />

(v,e v)∈I ˆQi<br />

W ev . Then X k is the k-th skeleton <strong>of</strong> X(Q, ϑ) and<br />

X(Q, ϑ) =<br />

n∪<br />

X k .<br />

So we get a CW -complex structure on X(Q, ϑ) with Σ n k h i cells in dimension k, 0 ≤ k ≤<br />

n. Since singular homology and cellular homology are isomorphic, we compute cellular<br />

homology <strong>of</strong> X(Q, ϑ). To calculate cellular homology we compute the boundary map<br />

<strong>of</strong> the cellular chain complex for X(Q, ϑ). To compute the boundary map we need to<br />

k=1<br />

compute the degree <strong>of</strong> the following composition map β wev<br />

S k−1<br />

φ F k v<br />

−−−−→ (Z<br />

n−1<br />

2 × (F k v − U ev )/ ∼ s<br />

q<br />

−−−−→ X k−1<br />

X k−2<br />

=<br />

∨<br />

(w,e w )∈I ˆQk−1<br />

S k−1<br />

w<br />

q w<br />

−−−−→ S<br />

k−1<br />

w<br />

(4.4.7)<br />

where Fv k is a face <strong>of</strong> ˆQ k <strong>of</strong> dimension k (k ≥ 2), Sw<br />

k−1 ∼= S k−1 and q, q w are the quotient<br />

maps. Clearly the above composition map β wev is either surjection or constant up to<br />

homotopy. When the map is constant the degree <strong>of</strong> the composition β wev is zero. We<br />

calculate the degree <strong>of</strong> the composition when it is surjection.<br />

Let (w, e w ) ∈ I ˆQk−1 be such that β wev is a surjection. Let z w be the vertex <strong>of</strong> the<br />

edge e w other than w. Let Fw<br />

k−1 ∈ F( ˆQ k−1 ) be the smallest face which contains the<br />

inward pointing edges to w in ˆQ k−1 . Let<br />

U ew = U w ∩ U zw ∩ F k−1<br />

w .<br />

So U ew<br />

is an open subset <strong>of</strong> F k−1<br />

w and U ew contains the relative interior <strong>of</strong> e w . The face


Chapter 4: Small orbifolds over simple polytopes 86<br />

F k−1<br />

w ⊂ F k v − U ev is a facet <strong>of</strong> F k v . Note that<br />

W ew = (Z n−1<br />

2 × U ew )/ ∼ s = S k−1<br />

w − {X k−2 /X k−2 }.<br />

The quotient group G F<br />

k−1<br />

w /G F k v is isomorphic to Z 2. Hence from equations 4.4.5 and<br />

4.4.6 we get that (β wev ) −1 (W ew ) has two components Y 1 and Y 2 in S k−1 . The restrictions<br />

(β wev ) |Y 1 and (β wev ) |Y 2, on Y 1 and Y 2 respectively, give homeomorphism to W ew .<br />

Let y v be the vertex <strong>of</strong> the edge e v other than v. Observe that<br />

(Z n−1<br />

2 × (F k v − {U ev ∪ {v, y v }})/ ∼ ev<br />

∼ = I 0 × S k−2 .<br />

Hence from the definition <strong>of</strong> equivalence relation ∼ s , it is clear that Y 2 is the image<br />

<strong>of</strong> Y 1 under the map (possibly up to homotopy)<br />

(id × a) : I 0 × S k−2 → I 0 × S k−2 defined by (id × a)(r, x) = (r, −x), (4.4.8)<br />

where I 0 = (0, 1) ⊂ R. The degree <strong>of</strong> (id × a) is (−1) k−1 . Hence the degree <strong>of</strong> the<br />

composition map β wev<br />

is<br />

⎧<br />

⎪⎨ 2 if k ≥ 2 is odd and β av is a surjection<br />

d vw = deg(β wev ) = 0 if k ≥ 2 is even and β av is a surjection<br />

⎪⎩<br />

0 if β av is constant.<br />

Hence the cellular chain complex <strong>of</strong> the constructed CW -complex <strong>of</strong> X(Q, ϑ) is<br />

0 → Z<br />

d n<br />

−−−−→ ⊕|I ˆQn−1 |Z → · · ·<br />

d 3<br />

−−−−→ ⊕|I ˆQ2<br />

|Z<br />

d 2<br />

−−−−→ ⊕|I ˆQ1<br />

|Z<br />

(4.4.9)<br />

d 1<br />

−−−−→ ⊕|I ˆQ0<br />

|Z → 0<br />

(4.4.10)<br />

where d k is the boundary map <strong>of</strong> the cellular chain complex. If k ≥ 2 the formula <strong>of</strong> d k<br />

is<br />

d k (W ev ) =<br />

∑<br />

(w,e w )∈I ˆQk−1<br />

d vw W ew , (4.4.11)<br />

where (v, e v ) ∈ I ˆQk<br />

and d vw is the degree <strong>of</strong> the composition map β wev . Hence the map<br />

d k is represented by the following matrix with entries<br />

{d vw : (v, e v ) ∈ I ˆQk, (w, e w ) ∈ I ˆQk−1}. (4.4.12)<br />

So the map d k is the zero matrix for all even k. When k ≥ 2 is odd, the map d k is an<br />

injection and the image <strong>of</strong> the map d k is the submodule generated by<br />

{<br />

∑<br />

(w,e w)∈I ˆQk−1<br />

d vw W ew : (v, e v ) ∈ I ˆQk}. (4.4.13)


87 4.4 Homology and Euler characteristic<br />

Hence the quotient module (⊕ |I ˆQk−1 | Z)/Imd k is isomorphic to<br />

( ⊕ h k<br />

Z) ⊕ ( ⊕<br />

Z 2 ).<br />

Σ n k+1 h i<br />

The 1-skeleton X 1 is a tree with Σ n 0 h i vertices and Σ n 1 h i edges. The boundary map<br />

d 1 is an injection and image <strong>of</strong> d 1 is Σ n 1 h i dimensional direct summand <strong>of</strong> ⊕ |I ˆQ0<br />

|Z over<br />

Z. Hence (⊕ |I ˆQ0|Z)/d 1 (⊕ |I ˆQ1<br />

|Z) is isomorphic to Z. From the previous calculation we<br />

have proved the following theorem.<br />

Theorem 4.4.2. The singular homology groups <strong>of</strong> the small orbifold X(Q, ϑ) with coefficients<br />

in Z is<br />

⎧<br />

⎪⎨ Z<br />

if k = 0 and if k = n even<br />

H k (X(Q, ϑ), Z) = (⊕ hk Z) ⊕ (⊕ Σ n<br />

k+1 h i<br />

Z 2 ) if k is even, 0 < k < n<br />

⎪⎩<br />

0 otherwise.<br />

Remark 4.4.3. If Q is an even dimensional simple polytope then the small orbifold<br />

over Q is orientable.<br />

Corollary 4.4.4. The singular homology groups <strong>of</strong> the orbifold X(Q, ϑ) with coefficients<br />

in Q is<br />

⎧<br />

⎪⎨ Q if k = 0 and if k = n even<br />

H k (X(Q, ϑ), Q) = ⊕ hk Q if k is even, 0 < k < n<br />

⎪⎩<br />

0 otherwise.<br />

With coefficients in Z 2 the cellular chain complex 4.4.10 is<br />

0 → Z 2<br />

0<br />

−−−−→ ⊕ |I ˆQn−1 |Z 2<br />

0<br />

−−−−→ · · ·<br />

0<br />

−−−−→ ⊕ |IQ 1 |Z 2<br />

d 1<br />

−−−−→ ⊕|I ˆQ0<br />

|Z 2<br />

0<br />

−−−−→ 0<br />

(4.4.14)<br />

Where d 1 is an injection. Hence (⊕ |I ˆQ0<br />

|Z 2 )/d 1 (⊕ |I ˆQ1<br />

|Z 2 ) is isomorphic to Z 2 . So we get<br />

the following corollary.<br />

Corollary 4.4.5. The singular homology groups <strong>of</strong> the orbifold X(Q, ϑ) with coefficients<br />

in Z 2 is<br />

⎧<br />

⎪⎨ Z 2 if k = 0 and if k = n<br />

H k (X(Q, ϑ), Z 2 ) = ⊕ Σ n<br />

⎪⎩<br />

k h i<br />

Z 2 if 1 < k < n<br />

0 if k = 1.<br />

Remark 4.4.6. The k-th modulo 2 Betti number b k (X(Q, ϑ)) <strong>of</strong> small orbifold X(Q, ϑ)<br />

is zero when k = 1 and b k (X(Q, ϑ)) = Σ n k h i if 1 < k ≤ n and b 0 (X(Q, ϑ)) = h 0 = 1.<br />

Hence modulo 2 Euler characteristic <strong>of</strong> X(Q, ϑ) is<br />

X(X(Q, ϑ)) = h 0 + Σ n k=2 (−1)k Σ n k h i = Σ [n/2]<br />

0 h 2i . (4.4.15)


Chapter 4: Small orbifolds over simple polytopes 88<br />

Observe that b k (X(Q, ϑ)) ≠ b n−k (X(Q, ϑ)) if 1 k < n. Hence the Poincaré Duality<br />

for small orbifolds is not true with coefficients in Z 2 .<br />

4.5 Cohomology ring <strong>of</strong> small orbifolds<br />

We have shown that the even dimensional small orbifolds are compact, connected, orientable.<br />

Let X (Q, ϑ) be an even dimensional small orbifold over the polytope Q. Hence<br />

by the Proposition 2.4.8 we get that the cohomology ring <strong>of</strong> X (Q, ϑ) satisfy the Poincaré<br />

duality with coefficients in rationals.<br />

We rewrite Poincaré duality for small orbifolds using the intersection theory. The<br />

purpose is to show the cup product in cohomology ring is Poincaré dual to intersection,<br />

see equation 4.5.9. The pro<strong>of</strong> is akin to the pro<strong>of</strong> <strong>of</strong> Poincaré duality for oriented<br />

closed manifolds proved in [GH78]. To show these we construct a q-CW complex<br />

structure on X(Q, ϑ). The q-CW complex structure on a Hausdorff topological space<br />

is discussed in Section 3.5. Similarly to the q-cellular homology case, we can show that<br />

q-cellular cohomology <strong>of</strong> a q-CW complex is isomorphic to its singular cohomology with<br />

coefficients in rationals.<br />

Let Q be an n-dimensional simple polytope where n is even and π : X(Q, ϑ) → Q<br />

be a small orbifold over Q. Let Q ′ be the second barycentric subdivision <strong>of</strong> Q. Let<br />

{ηα k : α ∈ Λ(k) and k = 0, 1, . . . , n} (4.5.1)<br />

be the simplices in Q ′ . Here k is the dimension <strong>of</strong> ηα k and Λ(k) is an index set. Let<br />

(ηα) k 0 be the relative interior <strong>of</strong> k-dimensional simplex ηα.<br />

k<br />

Definition 4.5.1. A subset Y ⊆ X(Q, ϑ) is said to be relatively open subset <strong>of</strong> dimension<br />

k if for each point y ∈ Y there exist an orbifold chart (Ũ, G, ψ) such that ψ(V ) ∋ y<br />

is an open subset <strong>of</strong> Y , for some k-dimensional submanifold V <strong>of</strong> Ũ.<br />

Then (π −1 )(ηα) k 0 is disjoint union <strong>of</strong> the following relatively open subsets<br />

{(σα k i<br />

) 0 ⊂ X(Q, ϑ) : i = 1, . . . , α(k)}<br />

for some natural number α(k). Here σα k i<br />

is the closure <strong>of</strong> (σα k i<br />

) 0 in X(Q, ϑ). The<br />

restriction <strong>of</strong> π on σα k i<br />

is a homeomorphism onto the simplex ηα k for i = 1, . . . , α(k).<br />

Then the collection<br />

{σα k i<br />

: i = 1, . . . , α(k) and α ∈ Λ(k) and k = 0, 1, . . . , n} (4.5.2)<br />

gives a simplicial decomposition <strong>of</strong> the small orbifold X(Q, ϑ). So<br />

K = {σα k i<br />

, ∂} αi ,k (4.5.3)


89 4.5 Cohomology ring <strong>of</strong> small orbifolds<br />

is a simplicial complex <strong>of</strong> X(Q, ϑ).<br />

Definition 4.5.2. The transversality <strong>of</strong> two relatively open subsets U and V <strong>of</strong> X(Q, ϑ)<br />

at p ∈ U ∩ V is defined as follows:<br />

1. If p is a smooth point <strong>of</strong> X(Q, ϑ), we say U intersect V transversely at p whenever<br />

T p (U) + T p (V ) = T p (X(Q, ϑ)).<br />

2. If p is an orbifold point <strong>of</strong> X(Q, ϑ) there exist an orbifold chart (B n , Z 2 , φ v ) such<br />

that φ v (0) = p. We say U intersect V transversely at p whenever T 0 (φ −1<br />

v (U)) +<br />

T 0 (φ −1<br />

v (V )) = T 0 (B n ).<br />

Let σ k 1<br />

α i<br />

and ρ k 2<br />

β j<br />

be two simplices <strong>of</strong> dimension k 1 and k 2 respectively in the simplicial<br />

complex K <strong>of</strong> X(Q, ϑ).<br />

Definition 4.5.3. We say σ k 1<br />

α i<br />

and ρ k 2<br />

β j<br />

intersect transversely at p ∈ σ k 1<br />

α i<br />

∩ ρ k 2<br />

β j<br />

if there<br />

exist two relatively open subsets U ⊂ X(Q, ϑ) and V ⊂ X(Q, ϑ) containing σ k 1<br />

α i<br />

and ρ k 2<br />

β j<br />

respectively such that dim(U) = k 1 , dim(V ) = k 2 and U intersect V transversely at p.<br />

Let U and V be two complementary dimensional relatively open subset <strong>of</strong> X(Q, ϑ)<br />

that intersect transversely at p ∈ U ∩ V .<br />

Definition 4.5.4. Define the intersection index <strong>of</strong> U and V at p to be 1 if there exist<br />

oriented bases {ξ 1 , . . . , ξ k1 } and {η 1 , . . . , η k2 } for T p (U) (T 0 (φ −1<br />

v (U))) and T p (V )<br />

(T 0 (φ −1<br />

v (V ))) respectively such that {ξ 1 , . . . , ξ k1 , η 1 , . . . , η k2 } is an oriented basis for<br />

T p X(Q, ϑ)(T 0 B n ) whenever p is smooth (respectively orbifold) point <strong>of</strong> X(Q, ϑ). Otherwise<br />

the intersection index <strong>of</strong> U and V at p is −1.<br />

Since antipodal action on B n (as n is even) is orientation preserving there is no<br />

ambiguity in the above definition. Let<br />

A = Σn αi σ k 1<br />

α i<br />

and B = Σm βj ρ k 2<br />

β j<br />

be two cycles <strong>of</strong> the simplicial complex K such that n = k 1 + k 2 and they intersect<br />

transversely.<br />

Definition 4.5.5. Define the intersection number <strong>of</strong> A and B is the sum <strong>of</strong> the intersection<br />

indixes (counted with multiplicity) at their intersection points.<br />

The number is finite since A and B are closed subsets <strong>of</strong> compact space X(Q, ϑ).<br />

We show that the intersection number depends only on the homology class <strong>of</strong> the cycle.<br />

Let σ k 1<br />

α i<br />

and ρ k 2<br />

β j<br />

be two simplices in K with k 1 + k 2 = n. From the construction <strong>of</strong> the<br />

simplicial complex K we make some observations.<br />

Observation 4.5.6. 1. σ k 1<br />

α i<br />

and ρ k 2<br />

β j<br />

can not contain different orbifold points whenever<br />

their intersection is nonempty.


Chapter 4: Small orbifolds over simple polytopes 90<br />

2. Each σ k 1<br />

α i<br />

and ρ k 2<br />

β j<br />

can contain at most one orbifold point.<br />

3. If σ k 1<br />

α i<br />

and ρ k 2<br />

β j<br />

contain an orbifold point or not, whenever their intersection is<br />

nonempty, we can find a Z 2 -invariant smooth homotopy<br />

G : [0, 1] × X v (Q, ϑ) → X v (Q, ϑ)<br />

fixing the orbifold point <strong>of</strong> X v (Q, ϑ) such that G(0 × U k 1<br />

α i<br />

) and G(1 × V k 2<br />

β j<br />

) intersect<br />

transversely where U k 1<br />

α i<br />

and V k 2<br />

β j<br />

containing σ k 1<br />

α i<br />

and ρ k 2<br />

β j<br />

respectively are suitable<br />

relatively open subsets <strong>of</strong> X v (Q, ϑ) and dimU k 1<br />

α i<br />

= k 1 , dimV k 2<br />

β j<br />

= k 2 .<br />

Let σ k 1<br />

α 0<br />

+ . . . + σ k 1<br />

α k1<br />

be the boundary <strong>of</strong> (k 1 + 1)-simplex σ k 1+1<br />

α . The observations<br />

4.5.6 also hold for the simplices σ k 1+1<br />

α and ρ k 2<br />

β j<br />

although k 1 + 1 + k 2 = n + 1. If G ′ is the<br />

smooth homotopy and G ′ (0 × U k 1+1<br />

α ) ∩ G ′ (1 × V k 2<br />

β j<br />

) is nonempty then the subset<br />

G ′ (0 × U k 1+1<br />

α ) ∩ G ′ (1 × V k 2<br />

β j<br />

)<br />

<strong>of</strong> X(Q, ϑ) is a collection <strong>of</strong> piecewise smooth curves. After lifting a curve to an orbifold<br />

chart (if necessary), using the similar arguments as in [GH78] we can show that intersection<br />

number <strong>of</strong> σ k 1<br />

α 0<br />

+ . . . + σ k 1<br />

α k1<br />

and ρ k 2<br />

β j<br />

is zero. Integrating these computation to<br />

the boundary A = Σn αi σ k 1<br />

α i<br />

and the cycle B = Σm βj ρ k 2<br />

β j<br />

we ensure that the intersection<br />

number <strong>of</strong> A and B is zero.<br />

Let K ′ = {τ k α i<br />

, ∂} be the first barycentric subdivision <strong>of</strong> the complex K. Now we<br />

construct the dual q-cell decomposition <strong>of</strong> the complex K. For each vertex σα 0 i<br />

in the<br />

complex K, let<br />

∗ σα 0 i<br />

=<br />

∪<br />

τβ n j<br />

σα 0 ∈τ n i β j<br />

(4.5.4)<br />

be the n-dimensional q-cell which is the union <strong>of</strong> the n-simplices τβ n j<br />

∈ K ′ containing<br />

σα 0 i<br />

as a vertex. Then for each k-simplex σα k i<br />

in the decomposition K, let<br />

∗ σ k α i<br />

=<br />

∩<br />

σ 0 β j<br />

∈τ n α i<br />

∗σ 0 β j<br />

(4.5.5)<br />

be the intersection <strong>of</strong> the n-dimensional q-cells associated to the k + 1 vertices <strong>of</strong> σ k α i<br />

.<br />

The q-cells {∆ n−k<br />

α i<br />

= ∗σα k i<br />

} give a q-cell decomposition <strong>of</strong> X(Q, ϑ), called the dual q-cell<br />

decomposition <strong>of</strong> K. So the dual q-cell decomposition {∆α<br />

n−k } is a q-CW structure on<br />

X(Q, ϑ).<br />

From the description <strong>of</strong> dual q-cells it is clear that ∆α n−k<br />

i<br />

intersects σα k i<br />

transversely<br />

when dimension <strong>of</strong> σα k i<br />

is greater than zero. ∆ n α i<br />

is a quotient space <strong>of</strong> the antipodal<br />

action on a symmetric convex polyhedral centered at origin in R n . Since the antipodal


91 4.5 Cohomology ring <strong>of</strong> small orbifolds<br />

action on R n (n even) preserve orientation <strong>of</strong> R n , we can define the intersection number<br />

<strong>of</strong> σ 0 α i<br />

and ∆ n α i<br />

to be 1. We consider the orientation on the dual q-cell {∆ n α i<br />

} such that<br />

the intersection number <strong>of</strong> σα k i<br />

and ∆ n−k<br />

α i<br />

is 1.<br />

Using the same argument as Grifiths and Harris have made in the pro<strong>of</strong> <strong>of</strong> Poincaré<br />

duality theorem in [GH78], we can prove the following relation between boundary operator<br />

∂ on the cell complex {σ k α i<br />

} and coboundary operator δ on the dual q-cell complex<br />

{∆ n−k<br />

α i<br />

} when dimension <strong>of</strong> σ k α i<br />

is greater than one,<br />

δ({∆ n−k<br />

α i<br />

}) = (−1) n−k+1 ∗ (∂σ k α i<br />

). (4.5.6)<br />

Let {σ k α i<br />

} =< x, y >∈ K be a one simplex with the vertices x, y. The orientation on<br />

{σ k α i<br />

} comes from the orientation <strong>of</strong> X(Q, ϑ). Since we are considering q-cell structure<br />

on X(Q, ϑ), define δ({∆ n−1<br />

α i<br />

}) = ∗σy 0 −∗σx. 0 So we get a map σα k i<br />

→ ∆ n−k<br />

α i<br />

which induces<br />

an isomorphism<br />

ξ k ′ : H k(X(Q, ϑ), Q) → H n−k<br />

q-CW<br />

(X(Q, ϑ), Q), (4.5.7)<br />

where H n−k<br />

q-CW<br />

(X(Q, ϑ), Q) is n − k th q-cellular cohomology group. Hence we have the<br />

following theorem for even dimensional small orbifold.<br />

Theorem 4.5.7 (Poincaré duality). Let X(Q, ϑ) be an even dimensional small orbifold.<br />

The intersection pairing<br />

H k (X(Q, ϑ), Q) × H n−k (X(Q, ϑ), Q) → Q<br />

is nonsingular; that is, any linear functional H n−k (X(Q, ϑ), Q) → Q is expressible as<br />

the intersection with some class Θ ∈ H k (X(Q, ϑ), Q). There is an isomorphism ξ ′ k from<br />

H k (X(Q, ϑ), Q) to H n−k (X(Q, ϑ), Q).<br />

Using this Poincaré duality theorem for even dimensional small orbifold we can<br />

calculate the cohomology groups <strong>of</strong> small orbifold X(Q, ϑ).<br />

Theorem 4.5.8. The singular cohomology groups <strong>of</strong> the even dimensional small orbifold<br />

X(Q, ϑ) with coefficients in Q is<br />

⎧<br />

⎪⎨ Q if k = 0 and if k = n even<br />

H k (X(Q, ϑ), Q) = ⊕ hk Q if k is even, 0 < k < n<br />

⎪⎩<br />

0 otherwise.<br />

We can also define a product µ k1 k 2<br />

at orbifold points. The product<br />

similarly as in [GH78] but some care is needed<br />

µ k1 k 2<br />

: H n−k1 (X(Q, ϑ), Q) × H n−k2 (X(Q, ϑ), Q) → H n−k1 −k 2<br />

(X(Q, ϑ), Q) (4.5.8)


Chapter 4: Small orbifolds over simple polytopes 92<br />

on the homology <strong>of</strong> X(Q, ϑ) in arbitrary dimensions satisfying the following commutative<br />

diagram.<br />

H n−k1 (X(Q, ϑ), Q) × H n−k2 (X(Q, ϑ), Q)<br />

⏐<br />

ξ n−k1 ×ξ ⏐↓<br />

n−k2<br />

H k 1<br />

(X(Q, ϑ), Q) × H k 2<br />

(X(Q, ϑ), Q)<br />

µ k1 k<br />

−−−−→<br />

2<br />

Hn−k1 −k 2<br />

(X(Q, ϑ), Q)<br />

⏐<br />

ξ ⏐↓<br />

n−k1 −k 2<br />

u<br />

−−−−→ H k 1+k 2<br />

(X(Q, ϑ), Q)<br />

(4.5.9)<br />

where the lower horizontal map u is the cup product in cohomology ring.<br />

We write some observations about the transversality <strong>of</strong> faces <strong>of</strong> an n-dimensional<br />

polytope Q (n even). Let F and F ′ be two faces <strong>of</strong> Q. F and F ′ intersect transversely<br />

if codim(F ∩ F ′ ) = codimF + codimF ′ . Since Q is simple polytope, the following two<br />

properties are satisfied.<br />

Property 1. Let F be a 2k-dimensional face <strong>of</strong> Q and u be a vertex <strong>of</strong> F . Then there is<br />

a unique (n − 2k)-dimensional face F ′ <strong>of</strong> Q such that F and F ′ meet at u transversely.<br />

Property 2. Let F be a face <strong>of</strong> codimension 2k. Then there is k many distinct faces<br />

<strong>of</strong> codimension two such that they intersect transversely at each point <strong>of</strong> F .<br />

Lemma 4.5.9. Let π : X(Q, ϑ) → Q be an even dimensional small orbifold and<br />

X(F, ϑ ′ ) = π −1 (F ) for each face F <strong>of</strong> Q. Then<br />

1. For each 2k-dimensional face F <strong>of</strong> Q, the homology class represented by X(F, ϑ ′ ),<br />

denoted by [X(F, ϑ ′ )], is not zero in H ∗ (X(Q, ϑ), Q).<br />

2. The cohomology ring H ∗ (X(Q, ϑ), Q) is generated by 2-dimensional classes.<br />

Pro<strong>of</strong>. The space X(F, ϑ ′ ) is a 2k-dimensional suborbifold <strong>of</strong> X(Q, ϑ), for each 2kdimensional<br />

(0 ≤ 2k ≤ n) face F <strong>of</strong> Q. By Corollary 4.4.4 we get that the homology<br />

in degree 2k <strong>of</strong> X(Q, ϑ) is generated by the classes <strong>of</strong> form [X(F, ϑ ′ )], where F is a<br />

2k-dimensional face.<br />

By equation 4.5.9, the dual <strong>of</strong> X(F ∩ F ′ , ϑ ′ ) is the cup product <strong>of</strong> the dual <strong>of</strong><br />

[X(F, ϑ ′ )] and the dual <strong>of</strong> [X(F ′ , ϑ ′ )], if F and F ′ intersect transversely and otherwise<br />

the dual <strong>of</strong> X(F ∩ F ′ , ϑ ′ ) is zero.<br />

The property 1 tells that there is an (n − 2k)-dimensional face F ′ which intersects<br />

F transversely at a vertex <strong>of</strong> Q. Since the homology classes [X(F, ϑ ′ )] and [X(F ′ , ϑ ′ )]<br />

are dual in intersection pairing <strong>of</strong> Poincaré duality, they are both nonzero. This proves<br />

(1) <strong>of</strong> the above Lemma.<br />

In theorem 4.5.8 we show the odd dimensional cohomology group is zero. The<br />

cohomology in degree 2k is generated by Poincaré duals <strong>of</strong> classes <strong>of</strong> the form [X(F, ϑ ′ )],<br />

codimF = 2k. By property 2, F is the transverse intersection <strong>of</strong> distinct faces <strong>of</strong><br />

codimension two. Hence, the Poincaré dual <strong>of</strong> [X(F, ϑ ′ )] is the product <strong>of</strong> cohomology<br />

classes <strong>of</strong> dimension 2. This proves (2) <strong>of</strong> the above Lemma.


93 4.5 Cohomology ring <strong>of</strong> small orbifolds<br />

Recall the index function ind Q from Section 4.4. Let ˆF v ∈ F(Q) be the smallest<br />

face containing the inward pointing edges incident to the vertex v <strong>of</strong> Q. Let w be the<br />

Poincaré dual <strong>of</strong> class <strong>of</strong> the form [X( ˆF v , ϑ ′ )], also denoted by [v]. Let {v 1 , v 2 , . . . , v r }<br />

be the set <strong>of</strong> vertices <strong>of</strong> Q such that ind Q (v i ) = n − 2. We show that {w 1 , w 2 , . . . , w r }<br />

is a minimal generating set <strong>of</strong> H ∗ (X(Q, ϑ), Q).<br />

Let A j = {v ∈ V (Q) : ind Q (v) = j. Let U ˆFv<br />

be the open subset <strong>of</strong> ˆFv obtained<br />

by deleting all faces <strong>of</strong> ˆFv not containing the vertex v. From Section 5.3 it is clear<br />

that π −1 (U ˆFv<br />

) is homeomorphic to the orbit space B j /Z 2 , where Z 2 action on B j is<br />

antipodal. So π −1 (U ˆFv<br />

) is j-dimensional q-cell in X(Q, ϑ). Clearly<br />

X(Q, ϑ) =<br />

∪<br />

v∈V (Q)<br />

π −1 (U ˆFv<br />

).<br />

This gives a q-CW structure on X(Q, ϑ). From Theorem 1.20 <strong>of</strong> [BP02], we get the<br />

number <strong>of</strong> j-dimensional cells is h n−j , cardinality <strong>of</strong> A j . So the corresponding q-cellular<br />

chain complex gives that {[v] : v ∈ A j } is a basis <strong>of</strong> H j (X(Q, ϑ), Q) if j is even. Theorem<br />

4.5.8 tells that {w = ξ j ([v]) : v ∈ A j } is a basis <strong>of</strong> H n−j (X(Q, ϑ), Q) if j is even.<br />

Let F be a codimension 2k face <strong>of</strong> Q with top vertex v <strong>of</strong> index n − 2k. By property<br />

2 F is unique intersection <strong>of</strong> k many distinct codimension 2 faces ˆF vi1 , . . . , ˆF vik<br />

with top vertices v i1 , . . . , v ik ∈ {v 1 , v 2 , . . . , v r } respectively. Hence w i1 . . . w ik = w in<br />

H ∗ (X(Q, ϑ), Q). Consider the polynomial ring Q[w 1 , w 2 , . . . , w r ]. Let the map<br />

µ ni1 ...n il<br />

: H ni1 (X(Q, ϑ), Q) × · · · × H nil (X(Q, ϑ), Q) → H n−ni1 −···−n il<br />

(X(Q, ϑ), Q)<br />

(4.5.10)<br />

be defined by the repeated application <strong>of</strong> the product map µ ni1 n i2<br />

. Let I be the ideal<br />

<strong>of</strong> Q[w 1 , w 2 , . . . , w r ] generated by the following elements<br />

⎧<br />

⎪⎨<br />

S =<br />

⎪⎩<br />

w i1 w i2 . . . w il if n i1 , . . . , n il are even and µ ni1 ...n il<br />

(v i1 , v i2 , . . . , v il ) = 0<br />

in H n−{ni1 +···+n il }(X(Q, ϑ), Q)<br />

∏ l1<br />

1<br />

w ik − ∏ l 2<br />

1<br />

w jl if µ ni1 ...n il1<br />

(v i1 , v i2 , . . . , v il1 ) = µ nj1 ...n jl2<br />

(v j1 , v j2 , . . . , v jl2 ) in<br />

H n−{ni1 +···+n il }(X(Q, ϑ), Q) for some n i1 , . . . , n il1 , n j1 , . . . ,<br />

n jl2 such that n i1 + . . . + n il1 = n j1 + . . . + n jl2<br />

(4.5.11)<br />

The Poincaré Duality theorem and intersection theory ensure that the relations among<br />

w i ’s are exactly as described above. Hence we have the following theorem.<br />

Theorem 4.5.10. The cohomology ring <strong>of</strong> an even dimensional small orbifold X(Q, ϑ)<br />

over the simple polytope Q is isomorphic to the quotient ring Q[w 1 , w 2 , . . . , w r ]/I.


Chapter 4: Small orbifolds over simple polytopes 94


Chapter 5<br />

T 2 -cobordism <strong>of</strong> quasi<strong>toric</strong><br />

4-manifolds<br />

5.1 Introduction<br />

In this chapter we introduce the notion <strong>of</strong> edge-simple polytope. We give the brief<br />

definition <strong>of</strong> some manifolds with quasi<strong>toric</strong> and small cover boundary in a constructive<br />

way. There is a natural torus action on these manifolds with quasi<strong>toric</strong> boundary having<br />

a simple convex polytope as the orbit space. Interestingly, we show that such a manifold<br />

with quasi<strong>toric</strong> boundary could be viewed as the quotient space <strong>of</strong> a quasi<strong>toric</strong> manifold<br />

corresponding to a certain circle action on it. We show these manifolds with quasi<strong>toric</strong><br />

boundary are orientable and compute their Euler characteristic.<br />

We consider the following category: the objects are all quasi<strong>toric</strong> manifolds and<br />

morphisms are torus equivariant maps between quasi<strong>toric</strong> manifolds. We compute the<br />

T 2 -cobordism group <strong>of</strong> 4-dimensional manifolds in this category. We show that the T 2 -<br />

cobordism group <strong>of</strong> 4-dimensional quasi<strong>toric</strong> manifolds is generated by the T 2 -cobordism<br />

classes <strong>of</strong> the complex projective space CP 2 , see Theorem 5.7.7. We also show that T 2 -<br />

cobordism class <strong>of</strong> a Hirzebruch surface is trivial, see Lemma 5.7.3. The main tool is<br />

the theory <strong>of</strong> quasi<strong>toric</strong> manifolds.<br />

5.2 Edge-Simple Polytopes<br />

In this section we introduce a particular type <strong>of</strong> polytopes, which we call an edge-simple<br />

polytopes.<br />

Definition 5.2.1. An n-dimensional convex polytope P is called an n-dimensional edgesimple<br />

polytope if each edge <strong>of</strong> P is the intersection <strong>of</strong> exactly (n − 1) facets <strong>of</strong> P .<br />

95


Chapter 5: T 2 -cobordism <strong>of</strong> quasi<strong>toric</strong> 4-manifolds 96<br />

Example 5.2.2. 1. An n-dimensional simple polytope is an n-dimensional edgesimple<br />

polytope.<br />

2. The following convex polytopes are edge-simple polytopes <strong>of</strong> dimension 3.<br />

3. The dual polytope <strong>of</strong> a 3-dimensional simple polytope is a 3-dimensional edgesimple<br />

polytope. This result is not true for higher dimensional polytopes, that is if<br />

P is a simple polytope <strong>of</strong> dimension n ≥ 4 the dual polytope <strong>of</strong> P may not be an<br />

edge-simple polytope. For example the dual <strong>of</strong> the 4-dimensional standard cube in<br />

R 4 is not an edge-simple polytope.<br />

Proposition 5.2.3. (a) If P is a 2-dimensional simple polytope then the suspension<br />

SP on P is an edge-simple polytope and SP is not a simple polytope.<br />

(b) If P is an n-dimensional simple polytope then the cone CP on P is an (n + 1)-<br />

dimensional edge-simple polytope.<br />

Pro<strong>of</strong>. (a) Let P be a 2-dimensional simple polytope with m vertices {v i : i ∈ I =<br />

{1, 2, . . . , m}} and m edges {e i : i ∈ I}. Let a and b be the other two vertices <strong>of</strong> SP .<br />

Then facets <strong>of</strong> SP are the cone (Ce i ) x on e i at x = a, b. Edges <strong>of</strong> SP are {xv i : x = a, b<br />

and i ∈ I} ∪ {e i : i ∈ I}. The edge xv i is the intersection <strong>of</strong> (Ce i1 ) x and (Ce i2 ) x if<br />

v i = e i1 ∩e i2 for x = a, b and e i = (Ce i ) a ∩(Ce i ) b . Hence SP is an edge-simple polytope.<br />

Each vertex v i <strong>of</strong> P is the intersection <strong>of</strong> 4 facets <strong>of</strong> SP . So SP is not a simple polytope.<br />

(b) Let P be an n-dimensional simple polytope in R n × 0 ⊆ R n+1 with m facets<br />

{F i : i ∈ I = {1, 2, . . . , m}} and k vertices {v 1 , v 2 , . . . , v k }. Assume that the cone are<br />

taken at a fixed point a in R n+1 −R n lying above the centroid <strong>of</strong> P . Then facets <strong>of</strong> CP are<br />

{(CF i ): i = 1, 2, . . . , m} ∪ {P }. Edges <strong>of</strong> CP are {av i = C({v i }): i = 1, 2, . . . , k} ∪ {e l :<br />

e l is an edge <strong>of</strong> P }. Since P is a simple polytope, each vertex v i <strong>of</strong> P is the intersection<br />

<strong>of</strong> exactly n facets <strong>of</strong> P , namely {v i } = ∩ n j=1 F i j<br />

and each edge e l is the intersection<br />

<strong>of</strong> unique collection <strong>of</strong> (n − 1) facets {F l1 , . . . , F ln−1 }. Then C{v i } = ∩ n j=1 CF i j<br />

and<br />

e l = P ∩ CF l1 ∩ CF l2 ∩ . . . ∩ CF ln−1 . That is C{v i } and {e l } are the intersection <strong>of</strong><br />

exactly n facets <strong>of</strong> CP . Hence CP is an (n + 1)-dimensional edge-simple polytope.<br />

Cut <strong>of</strong>f a neighborhood <strong>of</strong> each vertex v i , i = 1, 2, . . . , k <strong>of</strong> an n-dimensional edgesimple<br />

polytope P ⊂ R n by an affine hyperplane H i , i = 1, 2, . . . , k in R n such that<br />

H i ∩ H j ∩ P are empty sets for i ≠ j. Then the remaining subset <strong>of</strong> the polytope P is a


97 5.3 Manifolds with quasi<strong>toric</strong> boundary<br />

simple polytope <strong>of</strong> dimension n, denote it by Q P . Suppose P Hi = H i ∩ P = H i ∩ Q P for<br />

i = 1, 2, . . . k. Then P Hi is a facet <strong>of</strong> Q P called the facet corresponding to the vertex v i<br />

for each i = 1, . . . , k. Since each vertex <strong>of</strong> P Hi is an interior point <strong>of</strong> an edge <strong>of</strong> P and<br />

P is an edge-simple polytope, P Hi is an (n − 1)-dimensional simple polytope for each<br />

i = 1, 2, . . . , k.<br />

Lemma 5.2.4. Let F be a codimension l < n face <strong>of</strong> P . Then F is the intersection <strong>of</strong><br />

unique set <strong>of</strong> l facets <strong>of</strong> P .<br />

Pro<strong>of</strong>. The intersection F ∩Q P is a codimension l face <strong>of</strong> Q P not contained in ∪ k i=0 {P H i<br />

}.<br />

Since Q P is a simple polytope, F ∩ Q P = ∩ l j=1 F ′<br />

i j<br />

for some facets {F ′<br />

i 1<br />

, . . . , F ′<br />

i 1<br />

} <strong>of</strong> Q P .<br />

Let F ij be the unique facet <strong>of</strong> P such that F ′<br />

i j<br />

⊆ F ij . Then F = ∩ l 1 F i j<br />

. Hence each face<br />

<strong>of</strong> P <strong>of</strong> codimension l < n is the intersection <strong>of</strong> unique set <strong>of</strong> l facets <strong>of</strong> P .<br />

Remark 5.2.5. If v i is the intersection <strong>of</strong> facets {F i1 , . . . , F il } <strong>of</strong> P for some positive<br />

integer l, the facets <strong>of</strong> P Hi are {P Hi ∩ F i1 , . . . , P Hi ∩ F il }.<br />

5.3 Manifolds with quasi<strong>toric</strong> boundary<br />

Definition 5.3.1. A manifold with quasi<strong>toric</strong> boundary is a manifold with boundary<br />

where the boundary is a disjoint union <strong>of</strong> some quasi<strong>toric</strong> manifolds.<br />

Let P be an edge-simple polytope <strong>of</strong> dimension n with m facets F 1 , . . . , F m and k<br />

vertices v 1 , . . . , v k . Let e be an edge <strong>of</strong> P . Then e is the intersection <strong>of</strong> unique collection<br />

<strong>of</strong> (n − 1) facets {F ij : j = 1, . . . , (n − 1)}. Suppose F(P ) = {F 1 , . . . , F m }.<br />

Definition 5.3.2. The function λ: F(P ) → Z n−1 is called an isotropy function <strong>of</strong> the<br />

edge-simple polytope P if the set <strong>of</strong> vectors {λ(F i1 ), . . . , λ(F in−1 )} form a basis <strong>of</strong> Z n−1<br />

whenever the intersection <strong>of</strong> the facets {F i1 , . . . , F in−1 } is an edge <strong>of</strong> P .<br />

The vectors λ i := λ(F i ) are called isotropy vectors and the pair (P, λ) is called an<br />

isotropy pair.<br />

We define some isotropy functions <strong>of</strong> the edge-simple polytopes I 3 and P 0 in examples<br />

5.3.5 and 5.3.6 respectively.<br />

Remark 5.3.3. It may not possible to define an isotropy function on the set <strong>of</strong> facets <strong>of</strong><br />

arbitrary edge-simple polytopes. For example there does not exist an isotropy function<br />

<strong>of</strong> the standard n-simplex △ n for each n ≥ 3.<br />

We construct a manifold with quasi<strong>toric</strong> boundary from the isotropy pair (P, λ). Let<br />

F be a face <strong>of</strong> P <strong>of</strong> codimension l < n. Then F is the intersection <strong>of</strong> a unique collection<br />

<strong>of</strong> l facets F i1 , F i2 , . . . , F il <strong>of</strong> P . Let T F be the torus subgroup <strong>of</strong> T n−1 corresponding to<br />

the submodule generated by λ i1 , λ i2 , . . . , λ il in Z n−1 . Assume T v = T n−1 for each vertex


Chapter 5: T 2 -cobordism <strong>of</strong> quasi<strong>toric</strong> 4-manifolds 98<br />

v <strong>of</strong> P and T P = 1. We define an equivalence relation ∼ b on the product T n−1 × P as<br />

follows.<br />

(t, p) ∼ b (u, q) if and only if p = q and tu −1 ∈ T F (5.3.1)<br />

where F ⊆ P is the unique face containing p in its relative interior. We denote the<br />

quotient space (T n−1 × P )/ ∼ b by X(P, λ). The space X(P, λ) is not a manifold except<br />

when P is a 2-dimensional polytope. If P is 2-dimensional polytope the space X(P, λ)<br />

is homeomorphic to the 3-dimensional sphere.<br />

But whenever n > 2 we can construct a manifold with boundary from the space<br />

X(P, λ). We restrict the equivalence relation ∼ b on the product (T n−1 × Q P ) where<br />

Q P ⊂ P is a simple polytope as constructed in Section 5.2 corresponding to the edgesimple<br />

polytope P . Let<br />

W (Q P , λ) = (T n−1 × Q P )/ ∼ b ⊂ X(P, λ) (5.3.2)<br />

be the quotient space. The natural action <strong>of</strong> T n−1 on W (Q P , λ) is induced by the group<br />

operation in T n−1 .<br />

Theorem 5.3.4. The space W (Q P , λ) is a manifold with boundary. The boundary is a<br />

disjoint union <strong>of</strong> quasi<strong>toric</strong> manifolds.<br />

For each edge e <strong>of</strong> P , e ′ = e ∩ Q P is an edge <strong>of</strong> the simple polytope Q P . Let U e ′<br />

the open subset <strong>of</strong> Q P obtained by deleting all facets <strong>of</strong> Q P that does not contain e ′ as<br />

an edge. Then the set U e ′ is diffeomorphic to I 0 × R n−1<br />

0 where I 0 is the open interval<br />

(0, 1) in R. The facets <strong>of</strong> I 0 × R n−1<br />

0<br />

are I 0 × {x 1 = 0}, . . . , I 0 × {x n−1 = 0} where<br />

{x j = 0, j = 1, 2, . . . , n − 1} are the coordinate hyperplanes in R n−1 . Let F ′<br />

i 1<br />

, . . . , F ′<br />

i n−1<br />

be the facets <strong>of</strong> Q P such that ∩ n−1<br />

j=1 F ′<br />

i j<br />

= e ′ . Suppose the diffeomorphism<br />

g: U e ′ → I 0 × R n−1<br />

0<br />

sends F i ′<br />

j<br />

∩ U e ′ to I 0 × {x j = 0} for all j = 1, 2, . . . , n − 1. Define an isotropy function λ e<br />

on the set <strong>of</strong> all facets <strong>of</strong> I 0 ×R n−1<br />

0 by λ e (I 0 ×{x j = 0}) = λ ij for all j = 1, 2, . . . , n − 1.<br />

We define an equivalence relation ∼ e on (T n−1 × I 0 × R n−1 ) as follows.<br />

0<br />

(t, b, x) ∼ e (u, c, y) if and only if (b, x) = (c, y) and tu −1 ∈ T g(F ) . (5.3.3)<br />

where g(F ) ⊆ I 0 × R n−1<br />

0 is the unique face containing (b, x) in its relative interior, for<br />

a unique face F <strong>of</strong> U e ′ and T g(F ) = T F . So for each a ∈ I 0 the restriction <strong>of</strong> λ e on<br />

{({a} × {x j = 0}) : j = 1, 2, . . . , n − 1} define a characteristic function (see definition<br />

1.2.8) on the set <strong>of</strong> facets <strong>of</strong> {a} × R n−1<br />

0 . From the constructive definition <strong>of</strong> quasi<strong>toric</strong><br />

manifold given in [DJ91] it is clear that the quotient space {a} × (T n−1 × R n−1<br />

0 )/ ∼ e is<br />

be


99 5.3 Manifolds with quasi<strong>toric</strong> boundary<br />

diffeomorphic to {a} × R 2(n−1) . Hence the quotient space<br />

(T n−1 × I 0 × R n−1<br />

0 )/ ∼ e = I 0 × (T n−1 × R0<br />

n−1 )/ ∼ e<br />

∼= I 0 × R 2(n−1) .<br />

Since the quotient maps<br />

π : (T n−1 × U e ′) → (T n−1 × U e ′)/ ∼ b<br />

and<br />

π e : (T n−1 × I 0 × R n−1<br />

0<br />

) → (Tn−1 × I 0 × R n−1 )/ ∼ e<br />

0<br />

are continuous surjection and g is a diffeomorphism, the following commutative diagram<br />

ensure that the lower horizontal map g e is a homeomorphism.<br />

(T n−1 × U e ′)<br />

⏐<br />

π↓<br />

id×g<br />

−−−−→<br />

(T n−1 × I 0 × R0 n−1<br />

⏐ )<br />

π ⏐↓<br />

e<br />

(T n−1 g e<br />

× U e ′)/ ∼ b −−−−→ (T n−1 × I 0 × R n−1 )/ ∼ e<br />

0<br />

(5.3.4)<br />

∼=<br />

−−−−→ I 0 × R 2(n−1)<br />

Let v ′ 1 and v′ 2 be the vertices <strong>of</strong> the edge e′ <strong>of</strong> Q P . Suppose H 1 ∩ e ′ = {v ′ 1 } and<br />

H 2 ∩ e ′ = {v ′ 2 }, where H 1 and H 2 are affine hyperplanes as considered in Section 5.2<br />

corresponding to the vertices v 1 and v 2 <strong>of</strong> e respectively. Let U v ′<br />

1<br />

and U v ′<br />

2<br />

be the open<br />

subset <strong>of</strong> Q P obtained by deleting all facets <strong>of</strong> Q P not containing v 1 ′ and v′ 2 respectively.<br />

Hence there exist diffeomorphism g 1 : U v ′<br />

1<br />

→ [0, 1) × R n−1<br />

0 and g 2 : U v ′<br />

2<br />

→ [0, 1) × R n−1<br />

0<br />

satisfying the same property as the map g. We get the following commutative diagram<br />

and homeomorphisms g j e for j = 1, 2.<br />

(T n−1 × U v ′<br />

j<br />

)<br />

⏐<br />

π↓<br />

id×g j<br />

−−−−→<br />

(T n−1 × [0, 1) × R n−1<br />

0<br />

⏐ )<br />

π ⏐↓<br />

e<br />

(T n−1 g<br />

× U v ′<br />

j<br />

)/ ∼ j e<br />

b −−−−→ (T n−1 × [0, 1) × R n−1 )/ ∼ e<br />

0<br />

(5.3.5)<br />

∼=<br />

−−−−→ [0, 1) × R 2(n−1)<br />

Hence each point <strong>of</strong> (T n−1 ×Q P )/ ∼ b has a neighborhood homeomorphic to an open<br />

subset <strong>of</strong> [0, 1) × R 2(n−1) . So W (Q P , λ) is a manifold with boundary. From the above<br />

discussion the interior <strong>of</strong> W (Q P , λ) is<br />

∪ e ′ (T n−1 × U e ′)/ ∼ b = W (Q P , λ) {(T n−1 × ⊔ k i=1P Hi )/ ∼ b }<br />

and the boundary is ⊔ k i=1 {(Tn−1 × P Hi )/ ∼ b }. Let F (H) ij be a facet <strong>of</strong> P Hi . So there<br />

exists a unique facet F j <strong>of</strong> P such that F (H) ij = F j ∩ Q P ∩ H i . The restriction <strong>of</strong> the<br />

function λ on the set <strong>of</strong> facets <strong>of</strong> P Hi (namely λ(F (H) ij ) = λ j ) give a characteristic


Chapter 5: T 2 -cobordism <strong>of</strong> quasi<strong>toric</strong> 4-manifolds 100<br />

function <strong>of</strong> a quasi<strong>toric</strong> manifold over P Hi . Hence restricting the equivalence relation ∼ b<br />

on (T n−1 × P Hi ) we get that the quotient space W i = (T n−1 × P Hi )/ ∼ b is a quasi<strong>toric</strong><br />

manifold over P Hi . The boundary ∂W (Q P , λ) is the disjoint union ⊔ k i=1 W i, where W i is<br />

a quasi<strong>toric</strong> manifold. So W (Q P , λ) is a manifold with quasi<strong>toric</strong> boundary. In Section<br />

5.6 we have shown that these manifolds with quasi<strong>toric</strong> boundary are orientable.<br />

Example 5.3.5. An isotropy function <strong>of</strong> the standard cube I 3 is described in the following<br />

Figure 5.1. Here simple convex polytopes P H1 , . . . , P H8 are triangles. The restriction<br />

<strong>of</strong> the isotropy function on P Hi gives that the space (T 2 ×P Hi )/ ∼ b is the complex projective<br />

space either CP 2 or CP 2 . Since antipodal map in R 3 is an orientation reversing map<br />

we can show that the disjoint union ⊔ 4 i=1 CP2 ⊔ 4 i=1 CP2 is the boundary <strong>of</strong> (T 2 ×Q I 3)/ ∼ b .<br />

P H8<br />

(0, 1)<br />

(1, 0)<br />

P H7<br />

v 8<br />

(1, 1)<br />

(0, 1)<br />

v 7<br />

P H4<br />

(1, 0)<br />

v 4 v 3<br />

v 5<br />

(1, 1)<br />

v 6<br />

v 1 v2<br />

(1, 1)<br />

(1, 0) (0, 1)<br />

(I 3 , λ)<br />

P H5<br />

P H3<br />

P H6<br />

(1, 1)<br />

P H1<br />

P H2<br />

(0, 1)<br />

(1, 0)<br />

(Q I 3, λ)<br />

Figure 5.1: An isotropy function λ <strong>of</strong> the edge-simple polytope I 3<br />

Example 5.3.6. In the following Figure 5.2 we define an isotropy function <strong>of</strong> the<br />

edge-simple polytope P 0 . Here simple polytopes P H1 , P H2 , P H3 , P H4 are triangles and<br />

the simple polytope P H5 is a rectangle. The restriction <strong>of</strong> the isotropy function on P Hi<br />

gives that the space (T 2 × P Hi )/ ∼ b is either CP 2 or CP 2 for each i ∈ {1, 2, 3, 4} and<br />

(T 2 × P H5 )/ ∼ b is CP 1 × CP 1 . Hence the space ⊔ 2 i=1 ⊔ CP2 ⊔ CP 2 ⊔ (CP 1 × CP 1 ) is the<br />

boundary <strong>of</strong> (T 2 × Q P0 )/ ∼ b , see Section 5.7.<br />

5.4 Manifolds with small cover boundary<br />

Definition 5.4.1. A manifold with small cover boundary is a manifold with boundary<br />

where the boundary is a disjoint union <strong>of</strong> some small covers.


101 5.5 <strong>Some</strong> observations<br />

(0, 1)<br />

P H4<br />

(1, 0) P H5<br />

(0, 1)<br />

v 5<br />

v 4 v 3<br />

P H3<br />

v 1 v P H1 2<br />

(1, 0)<br />

P H2<br />

(1, 1)<br />

P 0<br />

Q P0<br />

Figure 5.2: An isotropy function λ <strong>of</strong> the edge-simple polytope P 0<br />

Definition 5.4.2. The function λ s : F(P ) → F n−1<br />

2 is called a F 2 -isotropy function <strong>of</strong><br />

the edge-simple polytope P if the set <strong>of</strong> vectors {λ s (F i1 ), . . . , λ s (F in−1 )} form a basis <strong>of</strong><br />

F n−1<br />

2 whenever the intersection <strong>of</strong> the facets {F i1 , . . . , F in−1 } is an edge <strong>of</strong> P .<br />

The vectors λ s i := λs (F i ) are called F 2 -isotropy vectors and the pair (P, λ s ) is called<br />

F 2 -isotropy pair.<br />

We can construct a manifold with small cover boundary from the pair (P, λ s ). Assign<br />

each face F to the subgroup G F <strong>of</strong> Z n−1<br />

2 determined by the vectors λ s i 1<br />

, . . . , λ s i l<br />

where<br />

F is the intersection <strong>of</strong> the facets F i1 , . . . , F il . Let ∼ s be an equivalence relation on<br />

(Z n−1<br />

2 × P ) defined by the following.<br />

(t, p) ∼ s (u, q) if and only if p = q and t − u ∈ G F (5.4.1)<br />

where F ⊆ P is the unique face containing p in its relative interior. Consider the<br />

restriction <strong>of</strong> ∼ s on (Z n−1<br />

2 × Q P ). The quotient space (Z2 n−1 × Q P )/ ∼ s ⊂ (Z n−1<br />

2 ×<br />

P )/ ∼ s , denoted by S(Q P , λ s ), is a manifold with boundary. This can be shown by the<br />

same arguments given in the Section 5.3. The boundary <strong>of</strong> this manifold is {(Z2 n−1 ×<br />

⊔ k i=1 P H i<br />

)/ ∼ s } = ⊔ k i=1 {(Zn−1 2 × P Hi )/ ∼ s }. Clearly the restriction <strong>of</strong> the Z 2 -isotropy<br />

function λ s on the set <strong>of</strong> facets <strong>of</strong> P Hi gives the Z 2 -characteristic function <strong>of</strong> a small<br />

cover over P Hi . So (Z n−1<br />

2 × P Hi )/ ∼ s is a small cover for each i = 0, . . . , k. Hence<br />

S(Q P , λ s ) is a manifold with small cover boundary.<br />

5.5 <strong>Some</strong> observations<br />

The set <strong>of</strong> facets <strong>of</strong> the simple convex polytope Q P are<br />

F(Q P ) = {P Hj : j = 1, 2, . . . , k} ∪ {F ′<br />

i : i = 1, 2, . . . , m},


Chapter 5: T 2 -cobordism <strong>of</strong> quasi<strong>toric</strong> 4-manifolds 102<br />

where F i ′ = F i ∩ Q P for a unique facets F i <strong>of</strong> P . Define the function ξ : F(Q P ) → Z n<br />

as follows.<br />

{<br />

(0, . . . , 0, 1) ∈ Z n if F = P Hj and j ∈ {1, . . . , k}<br />

ξ(F ) =<br />

λ i ∈ Z n−1 × {0} ⊂ Z n<br />

(5.5.1)<br />

if F = F i and i ∈ {1, 2, . . . , m}.<br />

So the function ξ satisfies the condition for the characteristic function (see definition<br />

1.2.8) <strong>of</strong> a quasi<strong>toric</strong> manifold over the n-dimensional simple convex polytope Q P . Hence<br />

from the characteristic model (Q P , ξ) we can construct the quasi<strong>toric</strong> manifold M(Q P , ξ)<br />

over Q P , see [DJ91]. There is a natural T n action on M(Q P , ξ). Let T H be the circle<br />

subgroup <strong>of</strong> T n determined by the submodule {0} × {0} × . . . × {0} × Z <strong>of</strong> Z n . Hence<br />

W (Q P , λ) is the orbit space <strong>of</strong> the circle T H action on M(Q P , ξ). The orbit map<br />

g H : M(Q P , ξ) → W (Q P , λ) is not a fiber bundle map.<br />

Remark 5.5.1. The manifold S(Q p , λ s ) with small cover boundary constructed in Section<br />

5.4 is the orbit space <strong>of</strong> Z 2 action on a small cover.<br />

5.6 Orientability <strong>of</strong> W (Q P , λ)<br />

Suppose W = W (Q P , λ). The boundary ∂W has a collar neighborhood in W . Hence by<br />

the Proposition 2.22 <strong>of</strong> [Hat02] we get H i (W, ∂W ) = ˜H i (W/∂W ) for all i. We show the<br />

space W/∂W has a CW -structure. Actually we show that corresponding to each edge<br />

<strong>of</strong> P there exist an odd-dimensional cell <strong>of</strong> W/∂W . Realize Q P as a simple polytope in<br />

R n and choose a linear functional φ : R n → R which distinguishes the vertices <strong>of</strong> Q P ,<br />

as in the pro<strong>of</strong> <strong>of</strong> Theorem 3.1 in [DJ91]. The vertices are linearly ordered according to<br />

ascending value <strong>of</strong> φ. We make the 1-skeleton <strong>of</strong> Q P into a directed graph by orienting<br />

each edge such that φ increases along edges. For each vertex v <strong>of</strong> Q P define its index,<br />

ind(v), as the number <strong>of</strong> incident edges that point towards v. Suppose V(Q P ) is the set<br />

<strong>of</strong> vertices and E(Q P ) is the set <strong>of</strong> edges <strong>of</strong> Q P . For each j ∈ {1, 2, . . . , n}, let<br />

I j = {(v, e v ) ∈ V(Q P ) × E(Q P ) : ind(v) = j and e v is the incident edge that points<br />

towards v such that e v = e ∩ Q P for an edge e <strong>of</strong> P }.<br />

Suppose (v, e v ) ∈ I j . Let F ev denote the smallest face <strong>of</strong> Q P which contains the inward<br />

pointing edges incident to v. Then F ev is a unique face not contained in any P Hi . Let<br />

U ev be the open subset <strong>of</strong> F ev obtain by deleting all faces <strong>of</strong> F ev not containing the edge<br />

e v . The restriction <strong>of</strong> the equivalence relation ∼ b on (T n−1 ×U ev ) gives that the quotient<br />

space (T n−1 × U ev )/ ∼ b is homeomorphic to the open disk B 2j−1 . Hence the quotient<br />

space (W/∂W ) has a CW -complex structure with odd dimensional cells and one zero


103 5.6 Orientability <strong>of</strong> W (Q P , λ)<br />

dimensional cell only. The number <strong>of</strong> (2j − 1)-dimensional cell is |I j |, the cardinality <strong>of</strong><br />

I j for j = 1, 2, . . . , n. So we get the following theorem.<br />

⎧ ⊕<br />

Z if i = 2j − 1 and j ∈ {1, . . . , n}<br />

⎪⎨<br />

|I j |<br />

Theorem 5.6.1. H i (W, ∂W ) =<br />

Z if i = 0<br />

⎪⎩<br />

0 otherwise<br />

When j = n the cardinality <strong>of</strong> I j is one. So H 2n−1 (W, ∂W ) = Z. Hence W is an<br />

oriented manifold with boundary.<br />

Example 5.6.2. We adhere the notations <strong>of</strong> example 5.3.6. Here I 3 = {(v 14 , e v14 )},<br />

I 2 = {(v 8 , e v8 ), (v 13 , e v13 ), (v 15 , e v15 )} and I 1 = {(v 3 , e v3 ), (v 6 , e v6 ), (v 9 , e v9 )}, see Figure<br />

5.3. The face F ev13 corresponding to the point (v 13 , e v13 ) is v 0 v 3 v 5 v 13 v 12 v 1 . Thus we<br />

can give a CW -structure <strong>of</strong> W (Q P0 , λ)/∂W (Q P0 , λ) with one 0-cell, two 1-cells, three<br />

3-cells and one 5-cell.<br />

P H5<br />

v 15<br />

v 14<br />

e v15 v 13<br />

P v 12<br />

e v14<br />

H4<br />

v e v12 11 v 7<br />

v e e v13 v8<br />

10<br />

v v 8<br />

9 v 6 P H3<br />

e v9<br />

v 1<br />

v 5 e v6<br />

v 2<br />

e v3<br />

v 4<br />

P H1 v 0 v 3<br />

P H2<br />

Q P0<br />

Figure 5.3: The index function <strong>of</strong> Q P0 .<br />

In [DJ91] the authors showed that the odd dimensional homology <strong>of</strong> quasi<strong>toric</strong> manifolds<br />

are zero. So H 2i−1 (∂W ) = 0 for all i. Hence we get the following exact sequences<br />

for the collared pair (W, ∂W ).


Chapter 5: T 2 -cobordism <strong>of</strong> quasi<strong>toric</strong> 4-manifolds 104<br />

0 → H 2n−1 (W )<br />

.<br />

0 → H 3 (W )<br />

0 → H 1 (W )<br />

j ∗<br />

−−−−→ H2n−1 (W, ∂W )<br />

j ∗<br />

−−−−→ H3 (W, ∂W )<br />

j ∗<br />

−−−−→ H1 (W, ∂W )<br />

.<br />

∂<br />

−−−−→ H 2n−2 (∂W )<br />

∂<br />

−−−−→ H 2 (∂W )<br />

∂<br />

−−−−→ H 0 (∂W )<br />

.<br />

i ∗<br />

−−−−→ H2n−2 (W ) → 0<br />

i ∗<br />

−−−−→ H2 (W ) → 0<br />

i ∗<br />

−−−−→<br />

.<br />

H0 (W ) ↠ Z<br />

(5.6.1)<br />

Where Z ∼ = H 0 (W, ∂W ). Let (h i0 , . . . , h in−1 ) be the h-vector <strong>of</strong> P Hi , for i =<br />

1, 2, . . . , k. The definition <strong>of</strong> h-vector <strong>of</strong> simple polytope is given in Section 1.4 <strong>of</strong><br />

Chapter 1. Hence the Euler characteristic <strong>of</strong> the manifold W with quasi<strong>toric</strong> boundary<br />

is Σ k i=1 Σn−1 j=0 h i j<br />

− Σ n−1<br />

j=1 |I j|.<br />

Fix the standard orientation on T n−1 . Let I n = {(v ′ , e v ′)}. Then the (2n − 1)-<br />

dimensional cell (T n−1 × U ev ′ )/ ∼⊂ W represents a fundamental class <strong>of</strong> W/∂W with<br />

coefficient in Z. Thus an orientation <strong>of</strong> U ev ′ (hence <strong>of</strong> Q P ) determines an orientation <strong>of</strong><br />

W . Note that an orientation <strong>of</strong> Q P is induced by orienting the ambient space R n .<br />

So the boundary orientation on P Hi induced from the orientation <strong>of</strong> Q P gives the<br />

orientation on the quasi<strong>toric</strong> manifold W i ⊂ ∂W . In the next section we consider the<br />

orientation <strong>of</strong> Q’s and Q P ’s induced from the standard orientation <strong>of</strong> R n and R n+1<br />

respectively.<br />

5.7 Torus cobordism <strong>of</strong> quasi<strong>toric</strong> manifolds<br />

Let C be the following category: the objects are all quasi<strong>toric</strong> manifolds and morphisms<br />

are torus equivariant maps between quasi<strong>toric</strong> manifolds. We are considering torus<br />

cobordism in this category only. Quasi<strong>toric</strong> manifolds are orientable manifolds, see the<br />

Section 1.5 <strong>of</strong> Chapter 1.<br />

Definition 5.7.1. Two 2n-dimensional quasi<strong>toric</strong> manifolds M 1 and M 2 are said to<br />

be T n -cobordant if there exist an oriented T n manifold W with boundary ∂W such that<br />

∂W is T n equivariantly homeomorphic to M 1 ⊔ (−M 2 ) under an orientation preserving<br />

homeomorphism. Here −M 2 represents the reverse orientation <strong>of</strong> M 2 .<br />

We denote the T n -cobordism class <strong>of</strong> quasi<strong>toric</strong> 2n-manifold M by [M].<br />

Definition 5.7.2. The n-th torus cobordism group is the group <strong>of</strong> all cobordism classes<br />

<strong>of</strong> 2n-dimensional quasi<strong>toric</strong> manifolds with the operation <strong>of</strong> disjoint union. We denote<br />

this group by CG n .<br />

Let M → Q be a 4-dimensional quasi<strong>toric</strong> manifold over the square Q with the<br />

characteristic function ξ : F(Q) → Z 2 . Consider that the orientation on M comes from<br />

the standard orientation on T 2 and Q ⊂ R 2 . We construct an oriented T 2 manifold W


105 5.7 Torus cobordism <strong>of</strong> quasi<strong>toric</strong> manifolds<br />

with boundary ∂W , where ∂W is equivariantly homeomorphic to −M ⊔⊔ k1 CP 2 ⊔⊔ k2 CP 2<br />

for some integer k 1 , k 2 . To show this we construct a 3-dimensional edge-simple polytope<br />

P E such that P E has exactly one vertex O which is the intersection <strong>of</strong> 4 facets with<br />

P E ∩ H O = Q and other vertices <strong>of</strong> P E are intersection <strong>of</strong> 3 facets. We define an isotropy<br />

function λ, extending the characteristic function ξ <strong>of</strong> M, from the set <strong>of</strong> facets <strong>of</strong> P E<br />

to Z 2 . Then W (Q PE , λ) is the required oriented T 2 manifold with quasi<strong>toric</strong> boundary.<br />

We have done an explicit calculation in the following.<br />

Let Q = ABCD be a rectangle (see Figure 5.4) belongs to {(x, y, z) ∈ R 3 ≥0 : x + y +<br />

z = 1}. Let ξ : {AB, BC, CD, DA} → Z 2 be the characteristic function for a quasi<strong>toric</strong><br />

manifold M over ABCD such that the characteristic vectors are<br />

ξ(AB) = ξ 1 , ξ(BC) = ξ 2 , ξ(CD) = ξ 3 and ξ(DA) = ξ 4 .<br />

We may assume that ξ 1 = (0, 1) and ξ 2 = (1, 0). From the classification results given<br />

in Section 1.2, it is enough to consider the following cases only.<br />

ξ 3 = (0, 1) and ξ 4 = (1, 0) (5.7.1)<br />

ξ 3 = (0, 1) and ξ 4 = (1, k), k = 1 or − 1 (5.7.2)<br />

ξ 3 = (0, 1) and ξ 4 = (1, k), k ∈ Z − {−1, 0, 1} (5.7.3)<br />

ξ 3 = (−1, 1) and ξ 4 = (1, −2) (5.7.4)<br />

For the case 5.7.1: In this case the edge-simple polytope ˜P 1 , given in Figure 5.4,<br />

is the required edge-simple polytope. The isotropy vectors <strong>of</strong> ˜P 1 are given by<br />

λ(OGH) = ξ 1 , λ(OHI) = ξ 2 , λ(OIJ) = ξ 3 , λ(OGJ) = ξ 4 and λ(GHIJ) = ξ 1 + ξ 2 .<br />

So we get an oriented T 2 manifold W (Q ˜P1<br />

, λ) with quasi<strong>toric</strong> boundary where the<br />

boundary is the quasi<strong>toric</strong> manifold −M ⊔ ⊔ k1 CP 2 ⊔ ⊔ k2 CP 2 for some integers k 1 , k 2 .<br />

Note that orientation on ˜P 1 ⊂ R 3 ≥0 comes from the standard orientation <strong>of</strong> R3 . Let A ′<br />

and B ′ be the midpoints <strong>of</strong> GJ and HI respectively. Let H be the plane passing through<br />

O, A ′ and B ′ in R 3 . Since a reflection in R 3 is an orientation reversing homeomorphism,<br />

it is easy to observe that the reflection on H induces the following orientation reversing<br />

equivariant homeomorphisms.<br />

(T 2 × ˜P 1I )/ ∼→ (T 2 × ˜P 1H )/ ∼ and (T 2 × ˜P 1J )/ ∼→ (T 2 × ˜P 1G )/ ∼ .


Chapter 5: T 2 -cobordism <strong>of</strong> quasi<strong>toric</strong> 4-manifolds 106<br />

So k 1 = k 2 . Since [CP 2 ] = −[CP 2 ], [M] = 0[CP 2 ].<br />

For the case 5.7.2: In this case |det(ξ 2 , ξ 4 )| = 1. Let O be the origin <strong>of</strong> R 3 . Let<br />

C Q be the open cone on rectangle ABCD at the origin O. Let G, H, I, J be points on<br />

extended OA, OB, OC, OD respectively. Let E and F be two points in the interior <strong>of</strong><br />

the open cones on AB and CD at O respectively such that |OG| < |OE|, |OH| < |OE|<br />

and |OI| < |OF |, |OJ| < |OF |. Then the convex polytope P 1 ⊂ C Q on the set <strong>of</strong><br />

vertices {O, G, E, H, I, F, J} is an edge-simple polytope (see Figure 5.4) <strong>of</strong> dimension<br />

3. Define a function, denote by λ, on the set <strong>of</strong> facets <strong>of</strong> P 1 by<br />

λ(OGEH) = ξ 1 , λ(OHI) = ξ 2 , λ(OJF I) = ξ 3 , λ(OJG) = ξ 4 ,<br />

λ(HIF E) = ξ 4 and λ(GJF E) = ξ 2 .<br />

(5.7.5)<br />

Hence λ is an isotropy function on the edge-simple polytope P 1 . The boundary <strong>of</strong><br />

the oriented T 2 manifold W (Q P1 , λ) is the quasi<strong>toric</strong> manifold M ⊔ ⊔ k1 CP 2 ⊔ ⊔ k2 CP 2<br />

for some integers k 1 , k 2 .<br />

Similarly to the previous case we can show that a suitable<br />

reflection induces the following orientation reversing equivariant homeomorphisms.<br />

(T 2 × P 1H )/ ∼→ (T 2 × P 1I )/ ∼, (T 2 × P 1E )/ ∼→ (T 2 × P 1J )/ ∼<br />

and<br />

(T 2 × P 1G )/ ∼→ (T 2 × P 1J )/ ∼ .<br />

So k 1 = k 2 . Hence [M] = 0[CP 2 ].<br />

E<br />

F<br />

H<br />

G<br />

J G<br />

A<br />

I<br />

D<br />

λ 4<br />

H<br />

B<br />

C<br />

A<br />

λ 1<br />

λ 3<br />

B<br />

λ 2<br />

O<br />

O<br />

D<br />

C<br />

J<br />

I<br />

H 1<br />

E 1<br />

G 1<br />

J 1<br />

A<br />

I 1 1<br />

D 1<br />

B 1<br />

C 1<br />

F 1<br />

Figure 5.4: The edge-simple polytope P 1 , ˜P 1 and the convex polytope P ′ 1 respectively.<br />

For the case 5.7.3: Suppose det(ξ 2 , ξ 4 ) = k > 1. Define a function λ (1) on the set<br />

<strong>of</strong> facets <strong>of</strong> P 1 except GEF J by<br />

λ (1) (OGEH) = ξ 1 , λ (1) (OHI) = ξ 2 , λ (1) (OIF J) = ξ 3 , λ (1) (OGJ) = ξ 4 ,<br />

and λ (1) (EHIF ) = ξ 2 + ξ 1 .<br />

(5.7.6)


107 5.7 Torus cobordism <strong>of</strong> quasi<strong>toric</strong> manifolds<br />

E 1<br />

O<br />

B<br />

C<br />

H<br />

I<br />

A<br />

D<br />

E<br />

F<br />

H 1<br />

G 1<br />

H 1 E 1<br />

G<br />

H<br />

G 1<br />

J 1<br />

J<br />

O<br />

J 1<br />

F 1<br />

I<br />

F 1<br />

1<br />

I<br />

I 1<br />

Figure 5.5: The edge-simple polytope P 2 .<br />

So the function λ (1) satisfies the condition <strong>of</strong> an isotropy function <strong>of</strong> the edge-simple<br />

polytope P 1 along each edge except the edges <strong>of</strong> the rectangle GEF J. The restriction<br />

<strong>of</strong> the function λ (1) on the edges GE, EF, F J, GJ <strong>of</strong> the rectangle GEF J gives the<br />

following equations,<br />

|det[λ (1) (GE), λ (1) (EF )]| = 1, |det[λ (1) (EF ), λ (1) (F J)]| = 1,<br />

|det[λ (1) (F J), λ (1) (GJ)]| = 1, |det[λ (1) (GJ), λ (1) (GE)]| = 1<br />

and det[λ (1) (EF ), λ (1) (GJ)] = k − 1 < k.<br />

(5.7.7)<br />

Let P 1 ′ be a 3-dimensional convex polytope as in the Figure 5.4. Identifying the facet<br />

GEF J <strong>of</strong> P 1 and A 1 B 1 C 1 D 1 <strong>of</strong> P 1 ′ through a suitable diffeomorphism <strong>of</strong> manifold with<br />

corners such that the vertices G, E, F, J maps to the vertices A 1 , B 1 , C 1 , D 1 respectively,<br />

we can form a new convex polytope P 2 , see Figure 5.5. After the identification following<br />

holds.<br />

1. The facet <strong>of</strong> P 1 containing GE and the facet <strong>of</strong> P 1 ′ containing A 1B 1 make the facet<br />

OHH 1 E 1 G 1 <strong>of</strong> P 2 .<br />

2. The facet <strong>of</strong> P 1 containing EF and the facet <strong>of</strong> P 1 ′ containing B 1C 1 make the facet<br />

HH 1 I 1 I <strong>of</strong> P 2 .<br />

3. The facet <strong>of</strong> P 1 containing F J and the facet <strong>of</strong> P 1 ′ containing C 1D 1 make the facet<br />

OII 1 F 1 J 1 <strong>of</strong> P 2 .<br />

4. The facet <strong>of</strong> P 1 containing JG and the facet <strong>of</strong> P 1 ′ containing D 1A 1 make the facet<br />

OJ 1 G 1 <strong>of</strong> P 2 .<br />

The polytope P 2 is an edge-simple polytope. We define a function λ (2) on the set <strong>of</strong><br />

facets <strong>of</strong> P 2 except G 1 E 1 F 1 J 1 by


Chapter 5: T 2 -cobordism <strong>of</strong> quasi<strong>toric</strong> 4-manifolds 108<br />

λ (2) (OHH 1 E 1 G 1 ) = ξ 1 , λ (2) (OIH) = ξ 2 , λ (2) (OII 1 F 1 J 1 ) = ξ 3 ,<br />

λ (2) (OJ 1 G 1 ) = ξ 4 , λ (2) (HH 1 I 1 I) = ξ 2 + ξ 1<br />

and λ (2) (H 1 I 1 F 1 E 1 ) = ξ 2 + 2ξ 1 .<br />

(5.7.8)<br />

So the function λ (2) satisfies the condition <strong>of</strong> an isotropy function <strong>of</strong> the edge-simple<br />

polytope P 2 along each edge except the edges <strong>of</strong> the rectangle G 1 E 1 F 1 J 1 . The restriction<br />

<strong>of</strong> the function λ (2) on the edges namely G 1 E 1 , E 1 F 1 , F 1 J 1 , G 1 J 1 <strong>of</strong> the rectangle<br />

G 1 E 1 F 1 J 1 gives the following equations,<br />

|det[λ 2 (G 1 E 1 ), λ 2 (E 1 F 1 )]| = 1, |det[λ 2 (E 1 F 1 ), λ 2 (F 1 J 1 )]| = 1,<br />

|det[λ 2 (F 1 J 1 ), λ 2 (G 1 J 1 )]| = 1, |det[λ 2 (G 1 J 1 ), λ 2 (G 1 E 1 )]| = 1<br />

and det[λ 2 (E 1 F 1 ), λ 2 (G 1 J 1 )] = k − 2 < k − 1.<br />

(5.7.9)<br />

Proceeding in this way, at k-th step we construct an edge-simple polytope P k with<br />

the function λ (k) , extending the function λ (k−1) , on the set <strong>of</strong> facets <strong>of</strong> P k such that<br />

λ (k) (H k−2 H k−1 I k−1 I k−2 ) = ξ 2 + (k − 1)ξ 1 = λ (k−1) (H k−2 I k−2 F k−2 E k−2 ),<br />

λ (k) (OG k−1 J k−1 ) = ξ 4 = λ (k−1) (OG k−2 J k−2 ),<br />

λ (k) (H k−1 I k−1 F k−1 E k−1 ) = ξ 4 and λ (k) (G k−1 E k−1 F k−1 J k−1 ) = ξ 2 + (k − 1)ξ 1 .<br />

(5.7.10)<br />

Observe that the function λ := λ (k) is an isotropy function <strong>of</strong> the edge-simple polytope<br />

P k . So we get an oriented T 2 -manifold with boundary W (Q Pk , λ) where the boundary<br />

is the quasi<strong>toric</strong> manifold M ⊔ ⊔ k1 CP 2 ⊔ ⊔ k2 CP 2 for some integers k 1 , k 2 .Similarly<br />

to the previous cases we can construct the following orientation reversing equivariant<br />

homeomorphisms.<br />

(T 2 × P kH )/ ∼→ (T 2 × P kI )/ ∼, (T 2 × P kGk−1 )/ ∼→ (T 2 × P kJk−1 )/ ∼,<br />

(T 2 × P kEk−1 )/ ∼→ (T 2 × P kFk−1 )/ ∼ and (T 2 × P kHi )/ ∼→ (T 2 × P kIi )/ ∼<br />

for i = 1, . . . , k − 1. So k 1 = k 2 . Hence [M] = 0[CP 2 ]. If k < −1, similarly we can show<br />

[M] = 0[CP 2 ].<br />

From the calculations for the cases 5.7.1, 5.7.2 and 5.7.3 we get the following lemma.<br />

Lemma 5.7.3. The T 2 -cobordism class <strong>of</strong> a Hirzebruch surface is trivial. In particular,<br />

oriented cobordism class <strong>of</strong> a Hirzebruch surface is also trivial.<br />

For the case 5.7.4: In this case |det[ξ 1 , ξ 3 ]| = 1. Following case 5.7.2, we can<br />

construct an edge simple polytope P ′′ and an isotropy function λ over this edge-simple<br />

polytope, see Figure 5.6. Hence we can construct an oriented T 2 manifold with quasi<strong>toric</strong><br />

boundary W (Q P ′′, λ) where the boundary is −M ⊔ ⊔ k1 CP 2 ⊔ ⊔ k2 CP 2 for some integers


109 5.7 Torus cobordism <strong>of</strong> quasi<strong>toric</strong> manifolds<br />

E<br />

λ 1<br />

F<br />

λ 3<br />

G<br />

J<br />

H<br />

D<br />

λ 3<br />

A<br />

B<br />

λ 4<br />

λ 2<br />

λ 1<br />

I<br />

C<br />

O<br />

P ′′<br />

Figure 5.6: The edge-simple polytope P ′′ and an isotropy function λ associated to the<br />

case 5.7.4.<br />

k 1 , k 2 . We may assume that ’the angles between the planes OHI and HIF E’ and ’the<br />

angles between the planes EF JG and HIF E’ are equal. Clearly a suitable reflection<br />

induces the following orientation reversing equivariant homeomorphisms.<br />

(T 2 × P ′′<br />

H)/ ∼→ (T 2 × P ′′<br />

E)/ ∼ and (T 2 × P ′′<br />

I )/ ∼→ (T 2 × P ′′<br />

F )/ ∼ . (5.7.11)<br />

Let CP 2 J = (T2 × P<br />

J ′′)/<br />

∼ and CP2 G = (T2 × P<br />

G ′′ )/ ∼. Observe that the characteristic<br />

functions <strong>of</strong> the triangles P<br />

J ′′ and P<br />

′′<br />

G<br />

are differ by a non-trivial automorphism <strong>of</strong> T2<br />

(or Z 2 ). So CP 2 J and CP2 G are complex projective space CP2 with two non-equivariant<br />

T 2 -actions. Hence [M] = [CP 2 J ] + [CP2 G ].<br />

To compute the group CG 2 we use the induction on the number <strong>of</strong> facets <strong>of</strong> 2-<br />

dimensional simple convex polytope in R 2 . We rewrite the pro<strong>of</strong> <strong>of</strong> well-known following<br />

lemma briefly.<br />

Lemma 5.7.4. The equivariant connected sum <strong>of</strong> two quasi<strong>toric</strong> manifolds is equivariant<br />

cobordant to the disjoint union <strong>of</strong> these two quasi<strong>toric</strong> manifolds.<br />

Pro<strong>of</strong>. Let M 1 and M 2 be two quasi<strong>toric</strong> manifolds <strong>of</strong> dimension 2n. Then W 1 :=<br />

[0, 1] × M 1 and W 2 := [0, 1] × M 2 are oriented T n -manifolds with boundary such that<br />

∂W 1 = 0 × (−M 1 ) ⊔ 1 × M 1 and ∂W 2 = 0 × (−M 2 ) ⊔ 1 × M 2 .<br />

Let x 1 ∈ M 1 and x 2 ∈ M 2 be two fixed points. Let U 1 ⊂ W 1 and U 2 ⊂ W 2 be<br />

two T n invariant open neighborhoods <strong>of</strong> 1 × x 1 and 1 × x 2 respectively.<br />

Identifying<br />

∂U 1 ⊂ (W 1 − U 1 ) and ∂U 2 ⊂ (W 2 − U 2 ) via an orientation reversing equivariant map<br />

we get the lemma.


Chapter 5: T 2 -cobordism <strong>of</strong> quasi<strong>toric</strong> 4-manifolds 110<br />

Now consider the case <strong>of</strong> a quasi<strong>toric</strong> manifold M over a convex 2-polytope P with<br />

m facets, where m > 4. By the classification result <strong>of</strong> 4-dimensional quasi<strong>toric</strong> manifold<br />

which is discussed in Section 1.2, M is one <strong>of</strong> the following equivariant connected sum.<br />

M = N 1 #CP 2 (5.7.12)<br />

M = N 2 #CP 2 (5.7.13)<br />

M = N 3 #M 4 k (5.7.14)<br />

The quasi<strong>toric</strong> manifolds N 1 , N 2 and N 3 are associated to the 2-polytopes Q 1 , Q 2 and<br />

Q 3 respectively. The number <strong>of</strong> facets <strong>of</strong> Q 1 , Q 2 and Q 3 are m − 1, m − 1 and m − 2<br />

respectively. The quasi<strong>toric</strong> manifold Mk<br />

4 is defined in Section 1.2 <strong>of</strong> Chapter 1. In<br />

previous calculations we have shown that [Mk 4] = 0[CP2 ]. So by the Lemma 5.7.4 we<br />

get either [M] = [N 1 ] + [CP 2 ] or [M] = [N 2 ] − [CP 2 ] or [M] = [N 3 ]. Thus using the<br />

induction on m, the number <strong>of</strong> facets <strong>of</strong> Q, we can prove the following.<br />

Lemma 5.7.5. Any 4-dimensional quasi<strong>toric</strong> manifold is equivariantly cobordant to the<br />

disjoint union ⊔ l 1 CP2 for some l, where the T 2 -action on different copies <strong>of</strong> CP 2 may<br />

be distinct.<br />

We classify the equivariant cobordism classes <strong>of</strong> all T 2 -actions on CP 2 . Let Q be<br />

a triangle and {F 1 , F 2 , F 3 } be the edges (facets) <strong>of</strong> Q. Let ξ : {F 1 , F 2 , F 3 } → Z 2 be a<br />

characteristic function such that ξ(F 1 ) = (a 1 , b 1 ) and ξ(F 2 ) = (a 2 , b 2 ). Because <strong>of</strong> the<br />

Corollary 1.2.17, we may assume that<br />

det(ξ(F 1 ), ξ(F 2 )) = |(a 1 , b 1 ; a 2 , b 2 )| = 1<br />

where (a 1 , b 1 ; a 2 , b 2 ) is the 2 × 2 matrix in SL(2, Z) with row vectors ξ(F 1 ) and ξ(F 2 ).<br />

We denote this matrix by ξ also. Then either ξ(F 3 ) = (a 1 + a 2 , b 1 + b 2 ), ξ(F 3 ) =<br />

−(a 1 + a 2 , b 1 + b 2 ), ξ(F 3 ) = −(a 1 − a 2 , b 1 − b 2 ) or ξ(F 3 ) = (a 1 − a 2 , b 1 − b 2 ). Let ξ ′ and<br />

ξ ′′ be two characteristic function defined respectively by,<br />

ξ ′ (F 1 ) = (a 1 , b 1 ), ξ ′ (F 2 ) = (a 2 , b 2 ), ξ ′ (F 3 ) = (a 1 + a 2 , b 1 + b 2 )<br />

and<br />

ξ ′′ (F 1 ) = (a 1 , b 1 ), ξ ′′ (F 2 ) = (a 2 , b 2 ), ξ ′′ (F 3 ) = (a 1 − a 2 , b 1 − b 2 ).<br />

Denote the quasi<strong>toric</strong> manifolds associated to the pairs (Q, ξ ′ ) and (Q, ξ ′′ ) by CP 2 ξ ′<br />

CP 2 ξ ′′<br />

respectively. The quasi<strong>toric</strong> manifolds associated to other posible characteristic<br />

function are equivariantly homeomorphic to either CP 2 ξ ′ or CP2 ξ ′′. Define an equivalence<br />

and


111 5.7 Torus cobordism <strong>of</strong> quasi<strong>toric</strong> manifolds<br />

relation ∼ eq on SL(2, Z) by<br />

(a 1 , b 1 ; a 2 , b 2 ) ∼ eq (−a 1 , −b 1 ; −a 2 , −b 2 ).<br />

Denote the equivalence class <strong>of</strong> ξ ∈ SL(2, Z) by [ξ] eq . Using Corollary 1.2.17 we get the<br />

following classification.<br />

Lemma 5.7.6. A T 2 -actions on CP 2 is equivariantly homeomorphic to either CP 2 ξ ′<br />

CP 2 ξ ′′ for a unique [ξ] eq ∈ SL(2, Z)/ ∼ eq .<br />

or<br />

Note that the natural T 2 -actions on CP 2 ξ ′ and CP2 ξ ′′ are same. Consider the linear<br />

map L ξ : Z 2 → Z 2 , defined by L ξ (1, 0) = (a 1 , b 1 ), L ξ (0, 1) = (a 2 , b 2 ). The map L ξ<br />

induces orientation preserving homeomorphisms CP 2 s → CP 2 ξ ′ and CP2 s → CP 2 ξ ′′. Thus<br />

[CP 2 ξ ′] = −[CP2 ξ ′′]. Observe that if [ξ 1] eq ≠ [ξ 2 ] eq then the corresponding characteristic<br />

functions are differ by δ ∗ , for some non trivial auto morphism δ : T 2 → T 2 . So [CP 2 ξ ′ 1] ≠<br />

[CP 2 ξ ′ 2]. Hence we get the following.<br />

Theorem 5.7.7. The set {[CP 2 ξ ′] : [ξ] eq ∈ SL(2, Z)/ ∼ eq } is a set <strong>of</strong> generators <strong>of</strong> the<br />

oriented torus cobordism group CG 2 .


Chapter 5: T 2 -cobordism <strong>of</strong> quasi<strong>toric</strong> 4-manifolds 112


Chapter 6<br />

Oriented cobordism <strong>of</strong> CP 2k+1<br />

6.1 Introduction<br />

In this chapter we have given a new construction <strong>of</strong> oriented manifold with the boundary<br />

CP 2k+1 for each k ≥ 0. The main tool is the theory <strong>of</strong> quasi<strong>toric</strong> manifolds. The strategy<br />

<strong>of</strong> our pro<strong>of</strong> is to first construct some compact orientable manifolds with quasi<strong>toric</strong><br />

boundary. Then identifying suitable boundary components using certain homeomorphisms<br />

we obtain oriented manifold with the boundary CP 2k+1 for each k ≥ 0, see<br />

Theorem 6.4.1.<br />

6.2 <strong>Some</strong> manifolds with quasi<strong>toric</strong> boundary<br />

Set n = 2(k + 1). Corresponding to each even k ≥ 0 we construct a manifold with<br />

quasi<strong>toric</strong> boundary. Let {A j : j = 0, . . . , n} be the standard basis <strong>of</strong> R n+1 . Let<br />

△ n = {(x 0 , x 1 , . . . , x n ) ∈ R n+1 : x j ≥ 0 and Σ n 0 x j = 1}. (6.2.1)<br />

Then △ n is an n-dimensional simplex with vertices {A j : j = 0, . . . , n} in R n+1 . Define<br />

△ n−1<br />

j<br />

= {(x 0 , x 1 , . . . , x n ) ∈ △ n : x j = 0}. (6.2.2)<br />

So △ n−1<br />

j<br />

is the facet <strong>of</strong> △ n not containing the vertex A j . Let F be the largest face <strong>of</strong><br />

△ n not containing the vertices A j1 , . . . , A jl . Then<br />

F = ∩ l i=1△ n−1<br />

j i<br />

= {(x 0 , x 1 , . . . , x n ) ∈ △ n : x ji = 0, i = 1, . . . , l}. (6.2.3)<br />

Define a function η : {△ n−1<br />

j<br />

: j = 0, . . . , n} −→ Z n−1 as follows.<br />

113


Chapter 6: Oriented cobordism <strong>of</strong> CP 2k+1 114<br />

(0, . . . , 0, 1, 0, . . . , 0) if 0 ≤ j <<br />

⎧⎪ n 2<br />

− 1, here 1 is in the (j + 1)-th place<br />

⎨<br />

n−j ) = (1, . . . , 1, 1, 0, . . . , 0) if j = n 2 − 1, here 1 occurs up to n 2<br />

-th place<br />

(0, . . . , 0, 1, 0, . . . , 0) if n 2<br />

≤ j < n, here 1 is in the j-th place<br />

⎪ ⎩<br />

(0, . . . , 0, 1, 1, . . . , 1) if j = n, here 0 occurs up to ( n 2<br />

− 1)-th place<br />

(6.2.4)<br />

η(△ n−1<br />

Define<br />

η j := η(△ n−1<br />

n−j<br />

), for all j = 0, 1, . . . , n. (6.2.5)<br />

Example 6.2.1. For n = 4, let △ 4 be the 4-simplex in R 5 with vertices A 0 , A 1 , A 2 , A 3 ,<br />

A 4 (see Figure 6.1). Define a function η from the set <strong>of</strong> facets <strong>of</strong> △ 4 to Z 3 by,<br />

⎧<br />

⎪⎩<br />

(1, 0, 0) if j = 0<br />

⎪⎨ (1, 1, 0) if j = 1<br />

η(△ 3 4−j) = (0, 1, 0) if j = 2<br />

(0, 0, 1) if j = 3<br />

(0, 1, 1) if j = 4<br />

(6.2.6)<br />

A 4<br />

A 0<br />

A 1 A 2<br />

A 3<br />

△ n−1 n<br />

2<br />

Figure 6.1: The 4-simplex △ 4<br />

Suppose the faces F ′ and F ′′ <strong>of</strong> △ n are the intersection <strong>of</strong> facets {△ n−1<br />

n , △ n−1<br />

} and {△ n−1 n , △ n−1 n<br />

−1,<br />

. . . , △n−1 0 } respectively. Then<br />

2 2<br />

n−1 , . . . ,<br />

F ′ = {(x 0 , x 1 , . . . , x n ) ∈ △ n : x n<br />

2 = 0, . . . , x n = 0}, (6.2.7)<br />

F ′′ = {(x 0 , x 1 , . . . , x n ) ∈ △ n : x 0 = 0, . . . , x n<br />

2<br />

= 0}. (6.2.8)<br />

Hence dim(F ′ ) = dim(F ′′ ) = n 2 − 1 ≥ 1. The set <strong>of</strong> vectors {η 0, . . . , η n } and<br />

2<br />

{η n , . . . , η n} are linearly dependent sets in Z n−1 . But the submodules generated by<br />

2<br />

the vectors {η 0 , . . . , ̂η j , . . . , η n } and {η n , η n<br />

2 2 2 +1 , . . . , ̂η l , . . . , η n } are n 2<br />

-dimensional direct<br />

summands <strong>of</strong> Z n−1 for each j = 0, . . . , n 2 and l = n 2<br />

, . . . , n respectively. Here the symbol<br />

̂ represents the omission <strong>of</strong> the corresponding entry.


115 6.2 <strong>Some</strong> manifolds with quasi<strong>toric</strong> boundary<br />

Suppose e is an edge <strong>of</strong> △ n not contained in F ′ ∪ F ′′ . Then e = ∩ n−1<br />

j=1 △n−1 n−l j<br />

for some<br />

{l j : j = 1, . . . , n − 1} ⊂ {0, 1, . . . , n}. Observe that {η 0 , . . . , η n<br />

2 } {η l 1<br />

, . . . , η ln−1 } and<br />

{η n<br />

2 , η ( n+2<br />

2 ), . . . , η n} {η l1 , . . . , η ln−1 }. Hence the set <strong>of</strong> vectors {η l1 , . . . , η ln−1 } form a<br />

basis <strong>of</strong> Z n−1 .<br />

Let r 1 , r 2 be two positive real numbers such that r 1 < r 2 and r 1 + 2r 2 < 1. Consider<br />

the following affine hyperplanes in R n+1 .<br />

H 1 = {(x 0 , x 1 , . . . , x n ) ∈ R n+1 : x n<br />

2 + · · · + x n = r 2 }. (6.2.9)<br />

H 2 = {(x 0 , x 1 , . . . , x n ) ∈ R n+1 : x 0 + · · · + x n<br />

2 = r 2}. (6.2.10)<br />

H 3 = {(x 0 , x 1 , . . . , x n ) ∈ R n+1 : x n<br />

2 = 1 − r 1}. (6.2.11)<br />

H = {(x 0 , x 1 , . . . , x n ) ∈ R n+1 : x 0 + · · · + x n = 1}. (6.2.12)<br />

Then △ n ⊂ H and the intersections △ n ∩H 1 ∩H 2 , △ n ∩H 1 ∩H 3 , △ n ∩H 3 ∩H 2 are empty.<br />

A 1<br />

A 0<br />

H 3<br />

H 2<br />

H 1<br />

A 2 A 3<br />

A 4<br />

Figure 6.2: The 4-simplex △ 4 and the affine hyperplanes H 1 , H 2 and H 3 .<br />

We cut <strong>of</strong>f an open neighborhood <strong>of</strong> faces F ′ , F ′′ and {A n } by affine hyperplanes H 1∩H,<br />

2<br />

H 2 ∩ H and H 3 ∩ H respectively in H. Let H j ′ be the closed half space associated to<br />

the affine hyperplane H j such that the interior <strong>of</strong> half spaces H 1 ′ , H′ 2 , H′ 3 do not contain<br />

the faces F ′ , F ′′ , {A n } respectively. We illustrate such hyperplanes for the case n = 4<br />

2<br />

in Figure 6.2. Define<br />

△ n Q = △ n ∩ H ′ 1 ∩ H ′ 2 ∩ H ′ 3, P 1 = △ n ∩ H 1 , P 2 = △ n ∩ H 2 and P 3 = △ n ∩ H 3 . (6.2.13)<br />

The convex polytope △ n Q<br />

is a simple convex polytope <strong>of</strong> dimension n and the polytopes<br />

P 1 , P 2 and P 3 are also facets <strong>of</strong> △ n Q . The polytopes P 1, P 2 and P 3 are given by


Chapter 6: Oriented cobordism <strong>of</strong> CP 2k+1 116<br />

the following equations.<br />

P 1 = {(x 0 , x 1 , . . . , x n ) ∈ △ n : x 0 +· · ·+x n<br />

2 −1 = 1−r 2 and x n<br />

2 +· · ·+x n = r 2 }. (6.2.14)<br />

P 2 = {(x 0 , x 1 , . . . , x n ) ∈ △ n : x 0 +· · ·+x n<br />

2 = r 2 and x n<br />

2 +1 +· · ·+x n = 1−r 2 }. (6.2.15)<br />

P 3 = {(x 0 , x 1 , . . . , x n ) ∈ △ n : x n<br />

2 = 1 − r 1 and x 0 + · · · + ̂x n<br />

2 + · · · + x n = r 1 }. (6.2.16)<br />

By equations 6.2.14 and 6.2.15, the convex polytopes P 1 and P 2 are diffeomorphic<br />

to the product △ n 2 −1 × △ n 2 . From equation 6.2.16 P 3 is diffeomorphic to the simplex<br />

△ n−1 . The facets <strong>of</strong> P 1 , P 2 and P 3 are given by the following equations respectively.<br />

△ n−1<br />

j<br />

∩P 1 = {(x 0 , x 1 , . . . , x n ) ∈ P 1 : x j = 0} for all j ∈ {0, . . . , n}. (6.2.17)<br />

△ n−1<br />

j<br />

∩P 2 = {(x 0 , x 1 , . . . , x n ) ∈ P 2 : x j = 0} for all j ∈ {0, . . . , n}. (6.2.18)<br />

△ n−1<br />

j<br />

∩P 3 = {(x 0 , x 1 , . . . , x n ) ∈ P 3 : x j = 0} for all j ∈ {0, . . . , ̂n , . . . , n}. (6.2.19)<br />

2<br />

Now we want to construct (2n − 1)-dimensional manifold with quasi<strong>toric</strong> boundary.<br />

Let F be a face <strong>of</strong> △ n <strong>of</strong> codimension l. Then<br />

F = △ n−1<br />

n−j 1<br />

∩ . . . ∩ △ n−1<br />

n−j l<br />

for a unique {j 1 , . . . , j l } ⊂ {0, 1, . . . , n}. Suppose T F be the torus subgroup <strong>of</strong> T n−1<br />

determined by the submodule generated by {η j1 , . . . , η jl } in Z n−1 . Assume T △ n = {1}.<br />

We define an equivalence relation ∼ η on the product T n−1 × △ n as follows,<br />

(s, p) ∼ η (t, q) if and only if p = q and ts −1 ∈ T F (6.2.20)<br />

where F ⊂ △ n is the unique face containing the point p in its relative interior. Restrict<br />

the equivalence relation ∼ η on T n−1 × △ n Q . Define<br />

W (△ n Q, η) := (T n−1 × △ n Q)/ ∼ η<br />

to be the quotient space. So W (△ n Q , η) is a Tn−1 -space. Let<br />

p : W (△ n Q, η) → △ n Q,<br />

defined by p([s, p]) = p, be the corresponding orbit map.<br />

Let η 1 , η 2 and η 3 be the restriction <strong>of</strong> the function η on the set <strong>of</strong> facets <strong>of</strong> P 1 , P 2


117 6.2 <strong>Some</strong> manifolds with quasi<strong>toric</strong> boundary<br />

and P 3 respectively. Define<br />

η i j := η i (△ n−1<br />

j<br />

∩ P i ) =<br />

{<br />

η j if i = 1, 2 and j ∈ {0, 1, . . . , n}<br />

η j if i = 3 and j ∈ {0, 1, . . . , ṋ<br />

2 , . . . , n}. (6.2.21)<br />

Let v be a vertex <strong>of</strong> P i . So v belongs to the relative interior <strong>of</strong> a unique edge e v <strong>of</strong><br />

△ n not contained in F ′ ∪ F ′′ . If<br />

e v = ∩ n−1<br />

j=1 △n−1 n−l j<br />

for some {l j : j = 1, . . . , n − 1} ⊂ {0, 1, . . . , n}, the vectors {η l1 , . . . , η ln−1 } form a basis<br />

<strong>of</strong> Z n−1 . So<br />

v = ∩ n−1<br />

j=1 (△n−1 n−l j<br />

∩ P i )<br />

and the vectors {η i l 1<br />

, . . . , η i l n−1<br />

} form a basis <strong>of</strong> Z n−1 .<br />

So η i defines the characteristic<br />

function <strong>of</strong> a quasi<strong>toric</strong> manifold M(P i , η i ) over P i . Hence from the definition <strong>of</strong><br />

equivalence relation ∼ η we get that<br />

M(P i , η i ) = (T n−1 × P i )/ ∼ η for i = 1, 2, 3. (6.2.22)<br />

Let U i be the open subset <strong>of</strong> △ n Q obtained by deleting all faces F <strong>of</strong> △n Q<br />

such that<br />

the intersection F ∩P i is empty. Then △ n Q = U 1 ∪U 2 ∪U 3 . The space U i is diffeomorphic<br />

as manifold with corners to [0, 1) × P i . Let<br />

f i : U i → [0, 1) × P i<br />

be a diffeomorphism. From the definition <strong>of</strong> η and ∼ η we get the following homeomorphisms<br />

(T n−1 × f −1<br />

i<br />

({a} × P i ))/ ∼ ∼ = {a} × M(P i , η i ) for all a ∈ [0, 1). (6.2.23)<br />

Hence the space p −1 (U i ) is homeomorphic to<br />

(T n−1 × f −1<br />

i<br />

([0, 1) × P i ))/ ∼ η<br />

∼ = [0, 1) × M(Pi , η i ).<br />

Since W (△ n Q , η) = p−1 (U 1 ) ∪ p −1 (U 2 ) ∪ p −1 (U 3 ), the space W (△ n Q<br />

, η) is a manifold with<br />

quasi<strong>toric</strong> boundary. The intersections P 1 ∩ P 2 , P 2 ∩ P 3 and P 1 ∩ P 3 are empty. Hence<br />

the boundary<br />

∂W (△ n Q, η) = M(P 1 , η 1 ) ⊔ M(P 2 , η 2 ) ⊔ M(P 3 , η 3 ).


Chapter 6: Oriented cobordism <strong>of</strong> CP 2k+1 118<br />

6.3 Orientability <strong>of</strong> W (△ n Q , η)<br />

Fix the standard orientation on T n−1 . Then the boundary orientations on P 1 , P 2 and P 3<br />

induced from the orientation <strong>of</strong> △ n Q give the orientations <strong>of</strong> M(P 1, η 1 ), M(P 2 , η 2 ) and<br />

M(P 3 , η 3 ) respectively.<br />

Let W := W (△ n Q<br />

, η). The boundary ∂W has a collar neighborhood in W . Hence by<br />

the proposition 2.22 <strong>of</strong> [Hat02],<br />

H i (W, ∂W ) = ˜H i (W/∂W ) for all i.<br />

We show the space W/∂W has a CW -structure. Assuming △ n Q ⊂ Rn , we choose a linear<br />

functional<br />

ζ : R n → R<br />

which distinguishes the vertices <strong>of</strong> △ n Q<br />

, as in the Section 1.5 <strong>of</strong> Chapter 1. The vertices<br />

are linearly ordered according to ascending value <strong>of</strong> ζ. We make the 1-skeleton <strong>of</strong> △ n Q<br />

into a directed graph by orienting each edge such that ζ increases along edges. For<br />

each vertex v <strong>of</strong> △ n Q<br />

define its index ind(v) as the number <strong>of</strong> incident edges that point<br />

towards v. Suppose V (△ n Q ) is the set <strong>of</strong> vertices and E(△n Q ) is the set <strong>of</strong> edges <strong>of</strong> △n Q .<br />

For each j ∈ {1, 2, . . . , n}, let<br />

I j = {(v, e v ) ∈ V (△ n Q) × E(△ n Q) : ind(v) = j and e v is the incident edge that<br />

points towards v such that e v = e ∩ △ n Q for an edge e <strong>of</strong> △ n }.<br />

Suppose (v, e v ) ∈ I j and F v ⊂ △ n Q<br />

is the smallest face containing the inward pointing<br />

edges incident to v in △ n Q . Then ind(v) = dim(F v). Let U ev be the open subset <strong>of</strong> F v<br />

obtained by deleting all faces <strong>of</strong> F v not containing the edge e v . The restriction <strong>of</strong> the<br />

equivalence relation ∼ η on (T n−1 × U ev ) gives that the quotient space (T n−1 × U ev )/ ∼ η<br />

is homeomorphic to the open disk B 2j−1 ⊂ R 2j−1 .<br />

Hence the quotient space (W/∂W ) has a CW -complex structure with odd dimensional<br />

cells and one zero dimensional cell only. The number <strong>of</strong> (2j − 1)-dimensional cell<br />

is |I j |, the cardinality <strong>of</strong> I j for j = 1, 2, . . . , n. So we get the following theorem.<br />

⎧ ⊕<br />

Z if i = 2j − 1 and j ∈ {1, . . . , n}<br />

⎪⎨<br />

|I j |<br />

Theorem 6.3.1. H i (W, ∂W ) =<br />

Z if i = 0<br />

⎪⎩<br />

0 otherwise<br />

When j = n the cardinality <strong>of</strong> I j is one. So H 2n−1 (W, ∂W ) = Z. Hence W is<br />

an oriented manifold with quasi<strong>toric</strong> boundary. From the definition 6.2.20 we get that<br />

the boundary orientation on M(P i , η i ) is same as the orientation on M(P i , η i ) as the<br />

quasi<strong>toric</strong> manifold, for all i = 1, 2, 3.


119 6.4 Oriented cobordism <strong>of</strong> CP 2k+1<br />

6.4 Oriented cobordism <strong>of</strong> CP 2k+1<br />

We show the quasi<strong>toric</strong> manifolds M(P 1 , η 1 ) and M(P 2 , η 2 ) are equivariantly homeomorphic<br />

up to an automorphism <strong>of</strong> T n−1 . Consider the permutation<br />

defined by<br />

ρ : {0, 1, . . . , n} → {0, 1, . . . , n}<br />

⎧<br />

n − 1 − j if 0 ≤ j < n 2<br />

⎪⎨<br />

− 1 and n 2 < j < n<br />

n if j = n<br />

ρ(j) =<br />

2 − 1<br />

n<br />

2<br />

if j = n (6.4.1)<br />

2<br />

⎪⎩ n<br />

2 − 1 if j = n.<br />

So ρ is an even or odd permutation if n = 4l or n = 4l + 2 respectively. Define a linear<br />

automorphism Φ on R n+1 by<br />

Φ(x 0 , . . . , x j , . . . , x n ) = (x ρ(0) , . . . , x ρ(j) , . . . , x ρ(n) ). (6.4.2)<br />

Hence Φ is an orientation preserving or reversing diffeomorphism if n = 4l or n = 4l + 2<br />

respectively. From equations 6.2.14 and 6.2.15 it is clear that Φ maps P 1 diffeomorphically<br />

onto P 2 . We denote the restriction <strong>of</strong> Φ on the faces <strong>of</strong> P 1 by Φ. Also from the<br />

equations 6.2.17 and 6.2.18 we get that Φ maps the facet △ n−1<br />

j<br />

∩ P 1 <strong>of</strong> P 1 diffeomorphically<br />

onto the facet △ n−1<br />

ρ(j) ∩ P 2 <strong>of</strong> P 2 . So<br />

Φ(△ n−1<br />

j<br />

∩ P 1 ) = △ n−1<br />

ρ(j) ∩ P 2, for all j = 0, . . . , n. (6.4.3)<br />

Let α 1 , . . . , α n−1 be the standard basis <strong>of</strong> Z n−1 over Z. Let δ ′ be the linear automorphism<br />

<strong>of</strong> Z n−1 defined by<br />

δ ′ (α i ) = α n−i ∀ i = 1, . . . , (n − 1). (6.4.4)<br />

Hence<br />

δ ′ (η i ) = η ρ(i) and δ ′ (η ρ(i) ) = η i for i = 0, 1, . . . , n. (6.4.5)<br />

Let δ be the automorphism <strong>of</strong> T n−1 induced by δ ′ . Hence the automorphism δ is<br />

orientation reversing if n = 4l and it is orientation preserving if 4l + 2.<br />

equations 6.2.21, 6.4.3 and 6.4.5 we get that the following commutative diagram.<br />

Φ<br />

F(P 1 ) −−−−→ F(P 2 )<br />

⏐ ⏐↓ ⏐<br />

η 1 η 2↓<br />

Z n−1<br />

δ<br />

−−−−→ Z n−1 .<br />

From the


Chapter 6: Oriented cobordism <strong>of</strong> CP 2k+1 120<br />

So the diffeomorphism<br />

δ × Φ : T n−1 × P 1 → T n−1 × P 2<br />

induces a δ-equivariant orientation reversing homeomorphism<br />

g n : M(P 1 , η 1 ) → M(P 2 , η 2 ).<br />

From the definition 6.2.21 <strong>of</strong> the characteristic function η 3 we get that the quasi<strong>toric</strong><br />

manifold M(P 3 , η 3 ) is equivariantly homeomorphic to CP n−1 if n = 4l + 2 and CP n−1<br />

if n = 4l.<br />

Define an equivalence relation ∼ n on W (△ n Q<br />

, η) by<br />

x ∼ n y if and only if x ∈ M(P 1 , η 1 ) and y = g n (x). (6.4.6)<br />

So the quotient space W (△ n Q , η)/ ∼ n is an oriented manifold with boundary. The<br />

boundary <strong>of</strong> these manifold is CP n−1 if n = 4l + 2 and the boundary is CP n−1 if n = 4l.<br />

So the quotient space W (△ n Q , η)/ ∼ n is an oriented manifold with boundary and the<br />

boundary is CP n−1 . Hence we have proved the following theorem.<br />

Theorem 6.4.1. The complex projective space CP 2k+1 is boundary <strong>of</strong> an oriented manifold,<br />

for all k ≥ 0.<br />

Example 6.4.2. We adhere to definition and notations given in the example 6.2.1.<br />

The faces A 0 A 1 and A 3 A 4 are the intersection <strong>of</strong> facets {△ 3 4 , △3 3 , △3 2 } and {△3 2 , △3 1 , △3 0 }<br />

respectively <strong>of</strong> △ 4 .<br />

Here the polytopes P 1 , P 2 are prism and P 3 is 3-simplex, see the Figure 6.3. The<br />

restriction <strong>of</strong> η (namely η 1 , η 2 and η 3 ) on the facets <strong>of</strong> P 1 , P 2 and P 3 are given in<br />

following Figure 6.4.<br />

P 1<br />

P 3<br />

P 2<br />

Figure 6.3: The simple convex polytope △ 4 Q with the facets P 1, P 2 and P 3 .


121 6.4 Oriented cobordism <strong>of</strong> CP 2k+1<br />

(1, 0, 0)<br />

(0, 1, 0)<br />

(0, 0, 1)<br />

(0, 1, 1) (1, 1, 0)<br />

(0, 1, 1)<br />

(1, 0, 0) (0, 0, 1)<br />

(1, 0, 0)<br />

(1, 1, 0)<br />

(1, 1, 0)<br />

(P 1 , η 1 )<br />

(0, 1, 0)<br />

(0, 1, 1)<br />

(P 2 , η 2 )<br />

(0, 0, 1)<br />

(P 3 , η 3 )<br />

Figure 6.4: The characteristic functions η 1 , η 2 and η 3 <strong>of</strong> P 1 , P 2 and P 3 respectively.<br />

Let δ ′ be the automorphism <strong>of</strong> Z 3 defined by<br />

δ ′ (1, 0, 0) = (0, 0, 1), δ ′ (0, 1, 0) = (0, 1, 0) and δ ′ (0, 0, 1) = (1, 0, 0).<br />

Clearly the combinatorial pairs (P 1 , η 1 ) and (P 2 , η 2 ) give two δ-equivariantly homeomorphic<br />

quasi<strong>toric</strong> manifolds, namely M(P 1 , η 1 ) and M(P 2 , η 2 ) respectively. The combinatorial<br />

pair (P 3 , η 3 ) gives the quasi<strong>toric</strong> manifold CP 3 over P 3 . So the boundary <strong>of</strong><br />

W (△ 4 Q , η) is M(P 1, η 1 ) ⊔ M(P 2 , η 2 ) ⊔ CP 3 .<br />

Hence after identifying M(P 1 , η 1 ) and M(P 2 , η 2 ) via an orientation reversing homeomorphism,<br />

we get an oriented manifold with boundary and the boundary is CP 3 .<br />

Now we briefly give the previous pro<strong>of</strong> <strong>of</strong> Theorem 6.4.1 following a notes by Andrew<br />

J. Baker. Consider the unit sphere S 4k+3 ⊂ H k+1 where H denotes the quaternions.<br />

This is acted on freely by the unit quaternions S 3 ⊂ H and its subgroup <strong>of</strong> unit complex<br />

numbers S 1 ⊂ C ⊂ H. Note that the conjugation action <strong>of</strong> S 3 on H restricted to the<br />

pure quaternions gives a realization <strong>of</strong> S 3 as Spin(3) acting on R 3 via the natural map<br />

to SO(3). Furthermore, S 1 ⊂ S 3 is identifies with Spin(2) and we have<br />

Spin(3)/Spin(2) ∼ = SO(3)/SO(2) ∼ = S 2 . (6.4.7)<br />

Also<br />

HP k = S 4k+3 /S 3 ; CP 2k+1 = S 4k+3 /S 1 ; (6.4.8)<br />

and the natural map CP 2k+1 → HP k can be identified with the sphere bundle <strong>of</strong><br />

S 4k+3 × Spin(3) R 3 → HP k .


Chapter 6: Oriented cobordism <strong>of</strong> CP 2k+1 122<br />

Thus we have a commutative diagram<br />

CP 2k+1<br />

↩→<br />

−−−−→ S 4k+3 × Spin(3) D 3<br />

⏐<br />

⏐<br />

↓<br />

↓<br />

(6.4.9)<br />

HP k HP k<br />

in which CP 2 k + 1 identifies with the boundary <strong>of</strong> S 4k+3 × Spin(3) D 3 . It is easy to show<br />

that cohomology groups <strong>of</strong> W (△ n Q , η)/ ∼ n (n = 2(k + 1)) and S 4k+3 × Spin(3) D 3 are<br />

different for all k > 0, . So we get two different construction <strong>of</strong> manifold with the<br />

boundary CP 2k+1 .<br />

Remark 6.4.3. We can give a nice CW -structure on W (△ n Q , η)/ ∼ n with one cell in<br />

dimension greater than zero and two cells in dimension zero from the combinatorial<br />

information. So the Theorem 6.4.1 may be helpful in some computations.


Bibliography<br />

[AB84]<br />

M. F. Atiyah and R. Bott. The moment map and equivariant cohomology.<br />

Topology, 23(1):1–28, 1984.<br />

[ALR07] A. Adem, J. Leida, and Y. Ruan. Orbifolds and stringy <strong>topology</strong>, volume 171 <strong>of</strong><br />

Cambridge Tracts in Mathematics. Cambridge University Press, Cambridge,<br />

2007.<br />

[AP93]<br />

[Ati57]<br />

C. Allday and V. Puppe. Cohomological methods in transformation groups,<br />

volume 32 <strong>of</strong> Cambridge Studies in Advanced Mathematics. Cambridge University<br />

Press, Cambridge, 1993.<br />

M. F. Atiyah. Complex analytic connections in fibre bundles. Trans. Amer.<br />

Math. Soc., 85:181–207, 1957.<br />

[Aud91] M. Audin. The <strong>topology</strong> <strong>of</strong> torus actions on symplectic manifolds, volume 93<br />

<strong>of</strong> Progress in Mathematics. Birkhäuser Verlag, Basel, 1991. Translated from<br />

the French by the author.<br />

[BP02]<br />

[BR01]<br />

[Bre72]<br />

V. M. Buchstaber and T. E. Panov. Torus actions and their applications in<br />

<strong>topology</strong> and combina<strong>toric</strong>s, volume 24 <strong>of</strong> University Lecture Series. American<br />

Mathematical Society, Providence, RI, 2002.<br />

V. M. Buchstaber and N. Ray. Tangential structures on <strong>toric</strong> manifolds, and<br />

connected sums <strong>of</strong> polytopes. Internat. Math. Res. Notices, (4):193–219, 2001.<br />

G. E. Bredon. Introduction to compact transformation groups. Academic Press,<br />

New York, 1972. Pure and Applied Mathematics, Vol. 46.<br />

[CK99] D. A. Cox and S. Katz. Mirror symmetry and algebraic geometry, volume 68<br />

<strong>of</strong> Mathematical Surveys and Monographs. American Mathematical Society,<br />

Providence, RI, 1999.<br />

[CR04]<br />

W. Chen and Y. Ruan. A new cohomology theory <strong>of</strong> orbifold. Comm. Math.<br />

Phys., 248(1):1–31, 2004.<br />

123


BIBLIOGRAPHY 124<br />

[Dan78] V. I. Danilov. The geometry <strong>of</strong> <strong>toric</strong> varieties. Uspekhi Mat. Nauk,<br />

33(2(200)):85–134, 247, 1978.<br />

[Dav78] M. Davis. Smooth G-manifolds as collections <strong>of</strong> fiber bundles. Pacific J. Math.,<br />

77(2):315–363, 1978.<br />

[Dav83] M. W. Davis. Groups generated by reflections and aspherical manifolds not<br />

covered by Euclidean space. Ann. <strong>of</strong> Math. (2), 117(2):293–324, 1983.<br />

[Dem70] M. Demazure. Sous-groupes algébriques de rang maximum du groupe de Cremona.<br />

Ann. Sci. École Norm. Sup. (4), 3:507–588, 1970.<br />

[DJ91]<br />

[Ful93]<br />

[GH78]<br />

[Gor78]<br />

[GP09]<br />

M. W. Davis and T. Januszkiewicz. Convex polytopes, Coxeter orbifolds and<br />

torus actions. Duke Math. J., 62(2):417–451, 1991.<br />

W. Fulton. Introduction to <strong>toric</strong> varieties, volume 131 <strong>of</strong> Annals <strong>of</strong> Mathematics<br />

Studies. Princeton University Press, Princeton, NJ, 1993. The William H.<br />

Roever Lectures in Geometry.<br />

P. Griffiths and J. Harris. Principles <strong>of</strong> algebraic geometry. Wiley-Interscience<br />

[John Wiley & Sons], New York, 1978. Pure and Applied Mathematics.<br />

R. Mark Goresky. Triangulation <strong>of</strong> stratified objects. Proc. Amer. Math. Soc.,<br />

72(1):193–200, 1978.<br />

S. Ganguli and M. Poddar. Blowdowns and McKay correspondence on four<br />

dimensional quasi<strong>toric</strong> orbifolds. ArXiv:0911.0766, November 2009.<br />

[Gui94] V. Guillemin. Moment maps and combinatorial invariants <strong>of</strong> Hamiltonian<br />

T n -spaces, volume 122 <strong>of</strong> Progress in Mathematics. Birkhäuser Boston Inc.,<br />

Boston, MA, 1994.<br />

[Hat02] A. Hatcher. Algebraic <strong>topology</strong>. Cambridge University Press, Cambridge, 2002.<br />

[MM03] I. Moerdijk and J. Mrčun. Introduction to foliations and Lie groupoids, volume<br />

91 <strong>of</strong> Cambridge Studies in Advanced Mathematics. Cambridge University<br />

Press, Cambridge, 2003.<br />

[MS74]<br />

J. W. Milnor and J. D. Stasheff. Characteristic classes. Princeton University<br />

Press, Princeton, N. J., 1974. Annals <strong>of</strong> Mathematics Studies, No. 76.<br />

[Oda88] T. Oda. Convex bodies and algebraic geometry, volume 15 <strong>of</strong> Ergebnisse der<br />

Mathematik und ihrer Grenzgebiete (3) [Results in Mathematics and Related<br />

Areas (3)]. Springer-Verlag, Berlin, 1988. An introduction to the theory <strong>of</strong><br />

<strong>toric</strong> varieties, Translated from the Japanese.


125 BIBLIOGRAPHY<br />

[OR70]<br />

[Pan01]<br />

P. Orlik and F. Raymond. Actions <strong>of</strong> the torus on 4-manifolds. I. Trans. Amer.<br />

Math. Soc., 152:531–559, 1970.<br />

T. E. Panov. Hirzebruch genera <strong>of</strong> manifolds with torus action. Izv. Ross.<br />

Akad. Nauk Ser. Mat., 65(3):123–138, 2001.<br />

[Poddf] M. Poddar. Notes on quasi<strong>toric</strong> manifolds. February 2010,<br />

http://bprim.org/atm/2010/topo/poddarnotes.pdf.<br />

[Pri67]<br />

D. Prill. Local classification <strong>of</strong> quotients <strong>of</strong> complex manifolds by discontinuous<br />

groups. Duke Math. J., 34:375–386, 1967.<br />

[PS10] M. Poddar and S. Sarkar. On Quasi<strong>toric</strong> Orbifolds. Osaka J. Math.,<br />

47(4):1055−1076, 2010.<br />

[Sar10a] S. Sarkar. Oriented Cobordism <strong>of</strong> CP 2k+1 . ArXiv:1011.0554v3, November<br />

2010.<br />

[Sar10b] S. Sarkar. <strong>Some</strong> small orbifolds over polytopes. ArXiv:0912.0777v4, November<br />

2010.<br />

[Sar10c] S. Sarkar. T 2 -cobordism <strong>of</strong> Quasi<strong>toric</strong> 4-Manifolds. ArXiv:1009.0342v2,<br />

September 2010.<br />

[Sat57]<br />

I. Satake. The Gauss-Bonnet theorem for V -manifolds. J. Math. Soc. Japan,<br />

9:464–492, 1957.<br />

[Sco83] P. Scott. The geometries <strong>of</strong> 3-manifolds. Bull. London Math. Soc., 15(5):401–<br />

487, 1983.<br />

[Sta75]<br />

R. P. Stanley. The upper bound conjecture and Cohen-Macaulay rings. Studies<br />

in Appl. Math., 54(2):135–142, 1975.<br />

[SU03] P. Sankaran and V. Uma. Cohomology <strong>of</strong> <strong>toric</strong> bundles. Comment. Math.<br />

Helv., 78(3):540–554, 2003.<br />

[SU07]<br />

P. Sankaran and V. Uma. K-theory <strong>of</strong> quasi-<strong>toric</strong> manifolds. Osaka J. Math.,<br />

44(1):71–89, 2007.<br />

[Thu3m] W. P. Thurston. The geometry and <strong>topology</strong> <strong>of</strong> three-manifolds.<br />

lecture notes. http://library.msri.org/books/gt3m.<br />

Princeton

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!