<<

Differential geometry

Mikhail G. Katz∗

M. Katz, Department of Mathematics, Bar Ilan Univer- sity, Ramat Gan 52900 Israel Email address: [email protected] Abstract. Lecture notes for course 88-826 in differential geom- etry on differentiable via coordinate charts and tran- sition functions, systoles, exterior differential complex, de Rham cohomology, Wirtinger inequality, Gromov’s systolic inequality for complex projective spaces. Contents

Chapter1. Differentiablemanifolds 9 1.1. Definitionofdifferentiablemanifold 9 1.2. Metrizability 11 1.3. Hierarchy of smoothness of M 11 1.4. Opensubmanifolds,Cartesianproducts 12 1.5. Circle, tori 13 1.6. Projective spaces 15 1.7. Derivations,Leibnizrule 17

Chapter 2. Derivations, tangent and cotangent bundles 19 2.1. Thespaceofderivations 19 2.2. Tangentbundleofasmoothmanifold 20 2.3. Vectorfieldasasectionofthebundle 22 2.4. Vectorfieldsincoordinates 22 2.5. Representingavectorbyapath 23 2.6. Vectorfieldsinpolarcoordinates 23 2.7. Source,sink,circulation 24 2.8. Dualityinlinearalgebra 25 2.9. Polar,cylindrical,andsphericalcoordinates 26 2.10. Cotangentspaceandcotangentbundle 28

Chapter3. Metricdifferentialgeometry 31 3.1. Isometries, constructing bilinear forms out of 1-forms 31 3.2. Riemannianmetric,firstfundamentalform 32 3.3. Metric as sum of squared 1-forms; element of length ds 33 3.4. Hyperbolic metric 34 3.5. Surface of revolution in R3 35 3.6. Coordinate change 37 3.7. Conformalequivalence 38 3.8. Lattices,uniformisationtheoremfortori 39 3.9. Isothermalcoordinates 41 3.10. Tori of revolution 42 3.11. Standard fundamental domain, conformal parameter τ 42 3.12. Conformalparameterofrectangularlattices 43 3 4 CONTENTS

3.13. Conformal parameter of standard tori of revolution 44

Chapter 4. Differential forms, and algebra 49 4.1. Differential 1-form as section of 49 4.2. Fromfunctiontodifferential1-form 50 4.3. Space Ω1(M) of differential 1-forms, exterior derivative 51 4.4. & exterior derivative; musical 51 4.5. Exterior product and algebra, alternating property 53 4.6. Exterioralgebraoveradim1vectorspace 55 4.7. Areasintheplane;signedarea 55 4.8. Algebraiccharacterisationofsignedarea 56 4.9. Vector product, triple product, and wedge product 57 4.10. Anticommutativityofthewedgeproduct 58 4.11. The k-exterior power; simple/decomposable multivs 58 4.12. Basisanddimensionofexterioralgebra 59

Chapter5. Exteriordifferentialcomplex 61 5.1.Rankofamultivector 61 5.2. Rankof2-multivectors;matrixofcoefficients 61 5.3. Constructionofthetensoralgebra 62 5.4. Constructionoftheexterioralgebra 64 5.5. Exteriorbundleandexteriorderivative 64 5.6. Exteriordifferentialcomplex 66 5.7. Antisymmetricmultilinearfunctions 67 5.8. Case of 2-forms 68

Chapter6. Normsonforms,Wirtingerinequality 71 6.1. Norm on 1-forms 71 6.2. Polarcoordinatesanddualnorms 72 6.3. Dual bases and dual lattices in a Euclidean space 73 6.4. Shortestnonzerovectorinalattice 73 6.5. Euclidean norm on k-multivectors and k-forms 74 6.6. Comass norm 74 6.7. Symplecticformfromacomplexviewpoint 75 6.8. Hermitian product 76 6.9. Orthogonaldiagonalisation 78 6.10. Wirtinger inequality 79 6.11. ProofofWirtingerinequality 80

Chapter 7. Complex projective spaces; de Rham cohomology 85 7.1. Celldecompositionofrealprojectivespace 85 7.2. Complex projective line CP1 86 7.3. Complex projective space CPn 87 CONTENTS 5

7.4. Unitary group, projective space as quotient by action 87 7.5. Hopf bundle 88 7.6. Flags and cell decomposition of CPn 88 7.7. Closure of cells produces compact submanifolds 89 7.8. g, α, and J inacomplexvectorspace 90 7.9. RelationtoHermitianproduct 91 7.10. Explicit formula for Fubini–Study 2-form on CP1 92 7.11. Exteriordifferentialcomplexrevisited 93 7.12. DeRhamcocyclesandcoboundaries 94 7.13. De Rham cohomology of M,Bettinumbers 94 Chapter8. DeRhamcohomologyandtopology 97 8.1. De Rham cohomology in dimensions 0 and n 97 8.2. deRhamcohomologyofacircle 98 8.3. Fubini–Study form and the area of CP1 99 8.4. 2-cohomologygroupofthetorus 100 8.5. 2-cohomologygroupofthesphere 102 8.6. Fubini–Study form and on CPn 104 8.7. Cup product, ring structure in de Rham cohomology 105 8.8. Cohomologyofcomplexprojectivespace 106 8.9. Abelianisationingrouptheory 107 8.10. The fundamental group π1(M) 108 Chapter9. Homologygroups;stablenorm 111 9.1. 1-cyclesonmanifolds 111 9.2. Orientationonamanifold 112 9.3. Interior product y ofaformbyavector 113 9.4. Induced orientation on the boundary ∂M 113 9.5. Surfacesandtheirboundaries 114

9.6. Restriction h⇂∂Σg and1-homologygroup 115 9.7. Lengthofcyclesand1-homologyclasses 116 9.8. Length in 1-homology group of orientable surfaces 117 9.9. Stable norm for higher-dimensional manifolds 118 9.10. Systoleandstable1-systole 119 9.11. 2-homologygroupand2-systole 120 9.12. Stablenormin2-homology 121 Chapter 10. Gromov’s stable systolic inequality for CPn 123 10.1. Pu’sinequalityforrealprojectiveplane 123 10.2. Integrationover1-cyclesandover2-cycles 124 10.3. Duality of homology and de Rham cohomology 126 10.4. VolumeformonRiemannianmanifolds 127 k k 10.5. Comass norm in Ω (M) and in HdR(M) 128 6 CONTENTS

10.6. Dualityofcomassandstablenorm 129 10.7. Fubini–Study metric on complex projective line 129 10.8. Fubini–Study metric on complex projective space 131 10.9. Gromov’s inequality for complex projective space 132 10.10. Homology class C and cohomology class ω 133 10.11. Comass and application of Wirtinger inequality 134 10.12. ProofofGromov’sinequality 134 Chapter11. Loewner’sinequality 137 11.1. Eisensteinintegers,Hermiteconstant 137 11.2. Standardfundamentaldomain 138 11.3. Loewner’storusinequality 139 11.4. Expectedvalueandvariance 140 11.5. Application of computational formula for variance 141 11.6. Flattorusaspencilofparallelgeodesics 142 11.7. ProofofLoewner’storusinequality 143 11.8. Boundary case of equality in Loewner’s inequality 144 11.9. Rectangularlattices 145 Chapter12. Pu’sinequalityandgeneralisations 147 12.1. StatementofPu’sinequality 147 12.2. Tangent map 148 12.3. Riemanniansubmersions 148 12.4. A Stiefel manifold 149 12.5. AmetricontheStiefelmanifold 150 12.6. Adoubleintegration 151 12.7. Aprobabilisticinterpretation 152 12.8. Fromthespheretotherealprojectiveplane 152 12.9. AgeneralisationofPu’sinequality 153 Chapter13. AlternativeproofofPu’sinequality 155 13.1. Hopf fibration h 155 13.2. HopffibrationisaRiemanniansubmersion 156 13.3. Hamiltonquaternions 156 13.4. Complex structures on the algebra H 157 13.5. Fromcomplexstructuretofibration 158 13.6. Lie groups 159 13.7. Lie group SO(3) as quotient of S3 160 13.8. Unitspheretangentbundle 161 13.9. 2-sphereasahomogeneousspace 161 13.10. Geodesicflowonthetangentbundle 162 13.11. Dual real projective plane and its double cover 163 13.12. Apairoffibrations 164 CONTENTS 7

13.13. Double fibration of T uS2, integral geometry on S2 165 Chapter 14. Gromov’s inequality for essential manifolds 169 14.1. Essentialmanifolds 169 14.2. Gromov’sinequalityforessentialmanifolds 169 14.3. Fillingradiusofaloopintheplane 169 14.4. The sup-norm 171 14.5. Riemannian manifolds as spaces with a distance function171 14.6. Kuratowskiembedding 171 14.7. Homologytheoryforgroups 171 14.8. Relativefillingradius 172 14.9. Absolutefillingradius 172 14.10. 172 Bibliography 173

CHAPTER 1

Differentiable manifolds

(1) Course site http://u.math.biu.ac.il/~katzmik/88-826.html (2) The final exam is 90% of the grade and the targilim 10%. (3) The first homework assignment is due on 7 april ’21.

1.1. Definition of differentiable manifold An n-dimensional manifold is a set M possessing certain addi- tional properties. A formal definition appears below as Definition 1.1.2. Here M is assumed to be covered by a collection of subsets (called coor- dinate charts or neighborhoods), typically denoted A or B, and having the following properties. For each coordinate neighborhood A M we have an injective map u: A Rn whose image ⊆ → u(A) Rn ⊆ is an open set in Rn. A coordinate chart is the pair (A, u). The maps are required to satisfy the following compatibility condition. Let u: A Rn, u =(ui) , (1.1.1) → i=1,...,n and similarly v : B Rn, v =(vα) (1.1.2) → α=1,...,n be a pair of coordinate charts. Whenever the overlap A B is nonempty, it has a nonempty image v(A B) in Euclidean space.∩ Both u(A) and u(A B), etc., are assumed∩ to be open subsets of Rn. ∩ 1 Definition 1.1.1. Let v− be the inverse map of the map v of (1.1.2). 1 n Thus v− is a map from the image (in R ) of the injective map v back to M. Restricting to the subset v(A B) Rn, we obtain a one-to-one map ∩ ⊆ 1 n φ = u v− : v(A B) R (1.1.3) ◦ ∩ → from an open set v(A B) Rn to Rn. ∩ ⊆1 n n Similarly, the map v u− from the open set u(A B) R to R is one-to-one. We can now◦ state the formal definition.∩ ⊆ 9 10 1.DIFFERENTIABLEMANIFOLDS

Definition 1.1.2. A smooth n-dimensional manifold M is a union

M = α I Aα, ∪ ∈ n where I is an index set, together with injective maps uα : Aα R , satisfying the following compatibility condition: the map φ of→ (1.1.3) is differentiable for all choices of coordinate neighborhoods A = Aα and B = A (where α, β I) as above. β ∈ 1 1 Definition 1.1.3. The map φ = u v− is called a transition map. The collection of coordinate charts as◦ above is called an atlas for the manifold M. Definition 1.1.4. A 2-dimensional manifold is called a surface. Note that we have not said anything yet about a topology on M. We define a topology on M as follows. Definition 1.1.5. The coordinate charts induce a topology on M by imposing the usual conditions: n 1 (1) If S R is an open set then v− (S) M is defined to be open;⊆2 ⊆ (2) arbitrary unions of open sets in M are open; (3) finite intersections of open sets are open. Remark 1.1.6. We will usually assume that M is connected. Given the manifold structure as above, connectedness of M is equivalent to path-connectedness3 of M. Identifying A B with a subset of Rn by means of the coordi- nates (ui), we can∩ think of the map v as given by n real-valued func- tions vα(u1,...,un), α =1,...,n. (1.1.4) These functions will be used to define the classes of smoothness of the manifold M in Section 1.3. Remark 1.1.7 (Dependent and independent variables). We present an alternative formulation in terms of dependent and independent variables and slightly more detailed notation. We retain the nota- tion u =(u1,...,un) and v =(v1,...,vn) for the coordinate variables in the respective Rn, and denote by : A Rn and : B Rn the U → V → 1funktsiat maavar or haatakat maavar 2Here v−1(S) is by definition the set of points x M such that v(x) S. If S happens to be disjoint from the image of v then by definition∈ v−1(S)= . Note∈ that the symbol v−1 is used here in a slightly different sense than in Definition∅ 1.1.1. 3kshir-mesila 1.3. HIERARCHY OF SMOOTHNESS OF MANIFOLD M 11

1 two charts. We let C = A B. The transition map φ = − is ∩n 1 V◦U defined in the set (C) R , and its inverse φ− is defined in (C). If we view u = (u1U,...,u⊆n) as the independent variables and v asV the dependent variables, then the dependence vα(u1,...,un) for each α = 1,...,n is expressed by the components of the transition map φ. In particular, we can write ∂φ ∂v1 ∂vn = ,..., ∂ui ∂ui ∂ui   for each i. If v is viewed as the independent variable and u as the depen- dent variable then the dependence u = u(v) is given by the transition 1 function ψ = φ− . In particular, one can write ∂ψ ∂u1 ∂un = ,..., ∂vα ∂vα ∂uα   for each index α.

1.2. Metrizability There are some non-Hausdorff examples that are pathological from the viewpoint of differential geometry, such as the following. Example 1.2.1. Let X be two copies of R glued along an open halfline of R. Then X satisfies the compatibility condition of Defini- tion 1.1.2. To rule out such examples, the simplest condition is that of metriz- ability: the topology generated by open balls of the metric (distance function) coincides with the topology underlying the differentiable struc- ture as in Definition 1.1.5. See e.g., Example 1.5.1 and Theorem 1.6.6. For connected manifolds, metrizability implies separability; see Gauld [4]. Examples of manifolds will be given in Sections 1.4 and 1.5.

1.3. Hierarchy of smoothness of manifold M The manifold condition stated in Definition 1.1.2 can be reformu- lated as the requirement that the n real-valued functions vα(u1,...,un) appearing in (1.1.4) are all smooth. Whatr is the precise meaning of smoothness? Definition 1.3.1. The usual hierarchy of smoothness (of func- k an tions), denoted C (or C∞, or C ), in Euclidean space generalizes to manifolds as follows. Here the conditions listed are assumed to be satisfied for all coordinate charts. 12 1.DIFFERENTIABLEMANIFOLDS

(1) For k = 1 a manifold M is C1 if and only if all n2 partial derivatives ∂vα , α =1,...,n, i =1,...,n ∂ui exist and are continuous. (2) The manifold M is C2 if all n3 second partial derivatives ∂2vα ∂ui∂uj exist and are continuous. (3) The manifold M is Ck if all the nk+1 partial derivatives ∂kvα ∂ui1 ∂uik exist and are continuous.··· (4) The manifold M is C∞ if for each k N, all partial derivatives ∈ ∂kvα ∂ui1 ∂uik exist. ··· (5) The manifold M is Can if for each k N, all partial derivatives ∈ ∂kvα ∂ui1 ∂uik are real analytic functions.··· The last condition is of course the strongest one of the five listed. 1.4. Open submanifolds, Cartesian products The notion of open and closed set in M is inherited from Euclidean space via the coordinate charts (see Definition 1.1.5). Definition 1.4.1. An open subset C M of a manifold M is itself a manifold, called an open submanifold⊆ of M. The differen- tiable structure on C is obtained by the restriction of the coordinate map u = (ui) of (A, u) for all charts (A, u) in M. The restriction will be denoted u⇂A C . ∩ Example 1.4.2. Let Matn,n(R) be the set of square matrices with real coefficients. This is linearly identified with Euclidean space of dimension n2, and is therefore a manifold. Theorem 1.4.3. Define a subset GL(n, R) Mat (R) by setting ⊆ n,n GL(n, R)= A Mat (R): det(A) =0 . { ∈ n,n 6 } Then GL(n, R) is an open submanifold. 1.5.CIRCLE,TORI 13

Proof. The determinant function is a polynomial in the entries xij of the matrix X. Therefore it is a continuous function of the entries, 2 which are the coordinates in Rn . Thus GL(n, R) is the inverse image of the open set R 0 under a continuous map, and is therefore an open set, hence a manifold\{ } with respect to the restricted atlas. 

Remark 1.4.4. The complement D of GL(n, R) in Matn,n(R) is the closed set consisting of matrices of zero determinant: D = Mat (R) GL(n, R)= A: det(A)=0 . n,n \ { } The set D for n 2 is not a manifold. ≥ The proof of the following theorem is straightforward. Theorem 1.4.5. Let M and N be two differentiable manifolds of dimensions m and n. Then the Cartesian product M N is a differ- entiable manifold of dimension m + n. The differentiable× structure is defined by coordinate neighborhoods of the form (A B,u v), where (A, u) is a coordinate chart on M, while (B,v) is× a coordinate× chart on N. Here the function u v on A B is defined by × × (u v)(x,y)=(u(x),v(y)) × for all x A, y B. ∈ ∈ 1.5. Circle, tori We give some additional examples of manifolds. Theorem 1.5.1. The circle S1 = (x,y) R2 : x2 + y2 = 1 is a differentiable manifold. { ∈ } Proof. Let us give an explicit atlas for the circle, in four steps. Step 1. Let A+ S1 be the open upper halfcircle ⊆ A+ = a =(x,y) S1 : y > 0 . { ∈ } Consider the coordinate chart (A+,u), namely u: A+ R, (1.5.1) → defined by setting u(x,y)= x (vertical projection to the x-axis). Step 2. Consider also the open lower halfcircle 1 A− = a =(x,y) S : y < 0 . { ∈ } It provides a coordinate chart (A−,u) where the coordinate u is defined by the same formula (1.5.1). Step 3. We define the right halfcircle B+ = a =(x,y) S1 : x> 0 , { ∈ } 14 1.DIFFERENTIABLEMANIFOLDS yielding a coordinate chart (B+,v) where v(x,y)= y, (1.5.2) and similarly for B−. Step 4. The theorem results from the Proposition 1.5.2 below on the transition functions φ. 

+ + Proposition 1.5.2. The four charts A , A−, B , and B− cover the circle S1 and define a differentiable structure on S1. Proof. Let us determine the transition functions. Step 1. The transition function between A+ and B+ is calculated as follows. Note that in the overlap A+ B+ one has both x> 0 and y > ∩ 1 0. Let us calculate the transition function u v− . It follows from (1.5.2) 1 1 ◦ 2 1 that the map v− sends the point y R to the point ( 1 y ,y) S , ∈ − ∈ and then the coordinate map u sends the point ( 1 y2,y) to its first −p coordinate 1 y2 R1. − ∈ 1 p Step 2. The composed map φ = u v− given by p ◦ φ(y)= 1 y2 (1.5.3) − is the transition function in this case.p 4 The function (1.5.3) is smooth for all y (0, 1). Similar remarks apply to the remaining transition functions.∈ It follows that the circle is an analytic manifold of dimen- sion 1, modulo checking the Hausdorff condition. Step 3. Finally we discuss the metrizability condition (see Sec- tion 1.2). We define a distance function on S1 by setting d(p,q) = arccos p,q (1.5.4) h i This gives a metric on S1 having all the required properties. It follows that S1 is metrizable.  Example 1.5.3. The torus T2 = S1 S1 is a 2-dimensional manifold by Theorem 1.4.5. Similarly, the n-torus× Tn = S1 S1 (product of n copies of the circle) is an n-dimensional manifold,×···× for all n 1. The circle itself is thought of as the 1-dimensional torus T1. ≥

4In more detail, let C = A+ B+ be the quarter circle in the first quadrant. Denote by p: C R the projection∩ to the x-axis, and by q : C R the projec- tion to the y-axis.→ Let φ = p q−1 be the transition function. If→y is viewed as ◦ the independent variable then x = φ(y) = 1 y2 and dx = φ′(y) = −2y . − dy √1−y2 When x is viewed as the independent variablep then similar formulas exist with φ−1 in place of φ. A similar picture emerges in the higher-dimensional case, where the dx dy ∂vα derivatives dy and dx are replaced by partial derivatives ∂ui , as in Section 1.3. 1.6.PROJECTIVESPACES 15

Example 1.5.4. The unit sphere Sn Rn+1 admits an atlas similar to the case of the circle. The distance function⊆ is defined by the same formula (1.5.4), establishing metrizability in this case. 1.6. Projective spaces Another example of a manifold that has fundamental importance in both differential and algebraic geometry is the projective space, defined as follows. In this section we will mainly consider the real case. For the complex case, see Corollary 1.6.7 and further links there. Remark 1.6.1. Complex projective spaces are dealt with in more detail in Section 7.3. To define the real projective space, let X = Rn+1 0 be the set of (n + 1)-tuples x =(x0,...,xn) distinct from the origin.\{ } Definition 1.6.2. An equivalence relation between elements x,y X is defined by setting x y if and only∼ if there is a ∈ t = 0 such that y = tx, i.e.,∼ 6 yi = txi, i =0,...,n. (1.6.1) Denote by [x] the equivalence class of x X. ∈ Definition 1.6.3. (x0,...,xn) are called the homogeneous coordi- nates of the point [x]. Lemma 1.6.4. For every x X, we have [x]=[ x]. ∈ − Proof. This is immediate from the choice t = 1 of the scalar.  − Definition 1.6.5. The real projective space, denoted RPn, is the collection of equivalence classes [x], i.e., RPn = [x]: x X . { ∈ } Theorem 1.6.6. The space RPn is a smooth n-dimensional mani- fold. Proof. To show that the set RPn is a manifold, we need to exhibit an atlas. Step 1. We define coordinate neighborhoods Ai, where i =0,...,n by setting i Ai = [x]: x =0 . (1.6.2) { 6 } n We will now define the coordinate pair (Ai,ui), where ui : Ai R , namely the corresponding coordinate chart. We set → x0 xi 1 xi+1 xn u (x)= ,..., − , ,..., . (1.6.3) i xi xi xi xi   16 1.DIFFERENTIABLEMANIFOLDS

Here formula (1.6.3) is valid since division by xi is allowed in the neigh- borhood Ai by condition (1.6.2). The coordinate chart ui is well-defined because if x y as in formula (1.6.1), then ui(x)= ui(y) by canceling out the scalar∼t in the numerator and denominator. Step 2. Let us calculate the transition maps. We let u = ui and v = uj. Assume for simplicity that i < j. We wish to calculate 1 5 the transition map φ = u v− associated with the overlap Ai Aj. Take a point ◦ ∩ 0 n 1 n z =(z ,...,z − ) R ∈ in the image of v. Since we are working with the condition xj = 0, we can rescale the homogeneous coordinates (x0,...,xn) (see Defini-6 j 1 tion 1.6.3) so that x = 1, so that we can represent v− (z)bythe(n+1)- tuple 1 0 j 1 j n 1 v− (z)=(z ,...,z − , 1,z ,...,z − ). (1.6.4)

Now we apply u = ui to (1.6.4), obtaining 0 i 1 i+1 j 1 j n 1 1 z z − z z − 1 z z − u v− (z)= ,..., , ,... , , ,..., (1.6.5) ◦ zi zi zi zi zi zi zi   since i < j. Step 3. All functions appearing in (1.6.5) are rational functions. Therefore the transition maps are smooth. Thus RPn is a differentiable manifold (in fact, analytic). Step 4. Let us check the metrizability condition (see Section 1.2). For unit vectors p,q X we set ∈ d(p,q) = arccos p,q . (1.6.6) |h i| Note that due to the use of the absolute value in (1.6.6), we have d( p,q)= d(p,q), which is consistent with the fact that [p]=[ p] (see Lemma− 1.6.4). For arbitrary p,q X we use the formula − ∈ p,q d(p,q) = arccos |h i|. (1.6.7) p q | || | Formula (1.6.7) provides a distance function on RPn with all the re- quired properties, showing that RPn is metrizable.  Note that p and p represent the same point in projective space. − Corollary 1.6.7. The complex projective space CPn is a smooth 2n- dimensional manifold.

5chafifa 1.7.DERIVATIONS,LEIBNIZRULE 17

Proof. We use C in place of R to define the coordinate neighbor- n hoods Ai as in formula (1.6.2), so that each Ai is a copy of C . The transition functions defined by (1.6.5) are rational functions which are therefore smooth in their domain of definition. To address the metriz- ability issue, we note that the space carries the natural Fubini–Study metric (8.6.1) dealt with in more detail in Section 8.6.  1.7. Derivations, Leibniz rule Let M be a differentiable manifold as defined in Sections 1.1 and 1.2. The , denoted TpM, at a point p M is intuitively the collection of all tangent vectors at the point p.6 ∈ In modern differential geometry, a tangent vector can be defined via derivations; see Definition 1.7.4. We first define the function ring Dp as follows. Definition 1.7.1. Let p M. Let ∈ D = f : f C∞ p { ∈ } be the ring of C∞ real-valued functions f defined in an (arbitrarily small) open neighborhood of p M. ∈ Definition 1.7.2. The ring operations in Dp are pointwise multi- plication fg and pointwise addition f + g, where we choose the inter- section of the domains of f and g as the domain of the new function (respectively sum or product). Thus, we set (fg)(x)= f(x) g(x) for all points x at which both functions are defined, and similarly for f+ g. Choose local coordinates (u1,...,un) near the point p M. Then a function f defined near p can be thought of as a function∈ of n vari- ables, f(u1,...,un). The following is proved in multivariate calculus. ∂ Theorem 1.7.3. A partial derivative ∂ui at the point p is a linear ∂ D R D form, or 1-form, denoted ∂ui : p on the space p, satisfying the Leibniz rule → ∂(fg) ∂f ∂g = g(p)+ f(p) (1.7.1) ∂ui ∂ui ∂ui p p p D for all f,g p. ∈ 6A preliminary notion of a tangent space, or plane, to a surface is developed in introductory courses based on a Euclidean embedding of the surface; see e.g., http://u.math.biu.ac.il/~katzmik/egreglong.pdf (course notes for the course 88-201). 18 1.DIFFERENTIABLEMANIFOLDS

Formula (1.7.1) can be written briefly as ∂ ∂ ∂ ∂ui (fg)= ∂ui (f) g + f ∂ui (g) keeping in mind that both sides are evaluated only at the point p (not in a neighborhood of the point). Formula (1.7.1) motivates the following more general definition of a derivation at p M. ∈ Definition 1.7.4. A derivation X at the point p M is a ∈ X : D R p → on the space Dp satisfying the Leibniz rule: X(fg)= X(f)g(p)+ f(p)X(g) (1.7.2) for all f,g D . ∈ p Note that the definition is coordinate-free. CHAPTER 2

Derivations, tangent and cotangent bundles

2.1. The space of derivations In Section 1.1 we defined the notion of a smooth manifold M. In Section 1.7 we gave a coordinate-free definition of a derivation X at a point p M, as a linear form on the space of smooth functions Dp such that ∈X satisfies the Leibniz rule at p. It turns out that the space of derivations is spanned by partial derivatives. Namely, we have the following theorem.

Theorem 2.1.1. Let M be an n-dimensional manifold. Let p M. Then the collection of all derivations at p is a of dimen-∈ sion n.

Proof in case n =1. We will prove the result in the case n = 1. For example, one could think of the 1-dimensional manifold M = R with the standard smooth structure. Thus we have a single coordi- nate u in a neighborhood of a point p M which can taken to be 0, i.e., p = 0. We argue in four steps as follows.∈

Step 1. Let X : Dp R be a derivation at p. To prove the theorem, → d it suffices to show that X is proportional to the derivative dx . Consider the constant function 1 Dp. Let us determine X(1). We have X(1) = X(1 1) = 2X(1) by the∈ Leibniz rule. Therefore X(1) = 0. Similarly for any· constant a we have X(a)= aX(1) = 0 by linearity of X.

Step 2. Now consider the monic polynomial u = u1 of degree 1, viewed as a linear function u Dp=0. We evaluate the derivation X at the element u D and set∈c = X(u). Thus c R. ∈ p ∈

Step 3. By the Taylor remainder formula, every function f Dp=0 can be written as follows: ∈

f(u)= a + bu + g(u)u, a,b R, ∈ 19 20 2. DERIVATIONS, TANGENT AND COTANGENT BUNDLES where g is smooth and g(0) = 0. Recall that f ′(0) = b. Therefore we have by linearity and Leibniz rule X(f)= X(a + bu + g(u)u) = bX(u)+ X(g)u(0) + g(0) c · = bc +0+0 d = c (f). du

Step 4. The formula d ( f D ) X(f)= c (f) ∀ ∈ p du d means that derivation X coincides with the derivation c du . Hence the d space of derivations is 1-dimensional and spanned by the derivation du , proving the theorem in the case n = 1. The case of general n is treated similarly using a Taylor formula with partial derivatives. 

Definition 2.1.2. The n-dimensional tangent space TpM to M at p is is the space of derivations at p.

2.2. of a smooth manifold Let M be a differentiable manifold. In Section 2.1 we defined the tangent space TpM at p M as the space of derivations at p. Next, we define the tangent bundle∈ of M. Definition 2.2.1. As a set, the tangent bundle, denoted TM, of an n-dimensional manifold M is the disjoint union of all tangent spaces TpM as p ranges through M, or in formulas:

TM = TpM. p M [∈ Definition 2.2.2 (Einstein summation convention). The rule is that whenever a product contains a symbol with a lower index and another symbol with the same upper index, take summation over this repeated index.

∂ Remark 2.2.3. The index i in ∂ui is considered to be a lower index since it occurs in the denominator. We will use such notation in the proof of the following theorem. 2.2. TANGENT BUNDLE OF A SMOOTH MANIFOLD 21

Theorem 2.2.4. The tangent bundle TM of a smooth n-dimen- sional manifold M has a natural structure of a smooth manifold of dimension 2n. Proof. We coordinatize TM locally using 2n coordinate functions as follows. By Theorem 2.1.1, a tangent vector v at a point p decom- i ∂ poses as v = v ∂ui (with respect to the Einstein summation conven- tion). We combine the coordinates (u1,...,un) in a coordinate neighbor- hood of M, together with the components vi of tangent vectors v ∂ ∂ ∂ i ∂∈ TpM, with respect to the basis ( ∂u1 , ∂u2 ,..., ∂un ), namely v = v ∂ui . The resulting string of 2n coordinates (u1,...,un,v1,...,vn) of the pair (p,v) parametrizes a neighborhood of TM. It can be checked that the transition functions are smooth, showing that TM is a (2n)- dimensional manifold.  Theorem 2.2.5. The tangent bundle TS1 of the circle S1 is a 2- dimensional manifold diffeomorphic to an (infinite) cylinder S1 R. × Proof. We represent the circle as the set of complex numbers of unit length: S1 = eiθ C. { }⊆ Recall that ei(θ+2πn) = eiθ for all n Z. We use the coordinate θ on the circle to express a tangent vector∈ at a point eiθ S1 as t d ∈ dθ where t R. Then the pair (eiθ,t) gives a parametrisation for the tangent bundle∈ of S1 by the Cartesian product S1 R, with eiθ S1 and t R. × ∈  ∈ Remark 2.2.6. The tangent bundle of S2 is not diffeomorphic to S2 R2; see Corollary 13.8.3. The Lie group S3 is parallelizable. The sphere× S7 is parallelizable. Interesting details on the 7-dimensional case can be found at https://mathoverflow.net/q/58131. Definition 2.2.7 (Canonical projection). Given the tangent bun- dle TM of a manifold M, let π : TM M, (p,v) p (2.2.1) M → 7→ be the canonical projection “forgetting” the tangent vector v and keep- ing only its initial point p. Note that there is no “canonical projection” to the second compo- nent v. 22 2. DERIVATIONS, TANGENT AND COTANGENT BUNDLES

2.3. Vector field as a section of the bundle Recall that we have a canonical projection π : TM M of (2.2.1). M → Definition 2.3.1. [Section] A vector field X on M is a section1 of the tangent bundle. Namely, a vector field is a map X : M TM satisfying the condition → π X = Id . M ◦ M Thus a vector field is a rule that assigns, to each point of M, a tan- gent vector at that point. We will express a vector field more concretely in terms of local coordinates in Section 2.4.

2.4. Vector fields in coordinates i Consider a coordinate chart (A, u) in M where u = (u )i=1,...,n. ∂ We have a basis ( ∂ui ) for TpM by Theorem 2.1.1. Thus an arbitrary vector X T M is a linear combination ∈ p ∂ Xi , ∂ui for appropriate coefficients Xi R depending on the point p. Here we use the Einstein summation convention.∈ ∂ Recall that the vectors ∂ui form a basis for the tangent space at every point of a coordinate neighborhood A M, by Theorem 2.1.1.  ⊆ Definition 2.4.1. A choice of component functions Xi(u1,...,un) in the neighborhood defines a vector field ∂ Xi(u1,...,un) ∂ui in the neighborhood A. Here the components Xi are required to be of an appropriate dif- ferentiability type. In more detail, we have the following definition.

Definition 2.4.2. (see [Boothby 1986, p. 117]) Let M be a C∞ manifold. A vector field X of class Cr on M is a map assigning to each i point p of M, a vector Xp TpM whose components (X ) in any local coordinate (A, u) are functions∈ of class Cr. Example 2.4.3. Let M be the Euclidean plane R2. Via obvious identifications, the Euclidean norm in the (x,y)-plane leads naturally to a Euclidean norm on the tangent space (i.e., tangent plane) at | | 1Chatach; mikta’ 2.6. VECTOR FIELDS IN POLAR COORDINATES 23

∂ ∂ every point with respect to which both tangent vectors ∂x and ∂y are orthogonal and have unit norm: ∂ ∂ = =1. ∂x ∂y

Note that each of ∂ and ∂ defines a global vector field on R2 (defined ∂x ∂y at every point of the plane). Any combination ∂ ∂ X = X1(x,y) + X2(x,y) ∂x ∂y is also a vector field in the plane, with norm X = (X1)2 +(X2)2 | | since the basis ( ∂ , ∂ ) is orthonormal. ∂x ∂y p 2.5. Representing a vector by a path ∂ Example 2.5.1. The derivation i.e., vector ∂x at a point p = (a,b) R2 can be represented by the path α(s)=(a + s,b), in the sense that∈ ∂ ( f D ) f = d f(α(s)), ∀ ∈ p ∂x ds s=0 ∂ and we have = α ′ = 1. | ∂x | | | ∂ Example 2.5.2. Similarly, the vector ∂y at a point p = (a,b) is represented by the path α(s)=(a,b + s).

2.6. Vector fields in polar coordinates Remark 2.6.1. Eventually we will develop the notion of a differ- ential k-form, generalizing the notion of a 1-form. The 1-forms, also known as covectors, are dual to vectors. We will first discuss some examples of vector fields to motivate the introduction of differential forms starting in Section 4.1. Interesting examples of vector fields are provided by polar coordi- nates (r, θ). These may be undefined at the origin, i.e., apriori only defined in the open submanifold R2 0 R2 . \{ }⊆ A point with polar coordinates (r, θ) has Cartesian coordinates given ∂ by (r cos θ,r sin θ). As in Section 2.5, the vector ∂θ at such a point is represented by the path α(θ)=(r cos θ,r sin θ) with derivative

α′(θ)=( r sin θ,r cos θ)= r( sin θ, cos θ) − − 24 2. DERIVATIONS, TANGENT AND COTANGENT BUNDLES and therefore α ′ = r. Hence at the point with polar coordinates (r, θ), we have | | ∂ = r. (2.6.1) ∂θ

2 For an alternative argument see the note.

1 ∂ Corollary 2.6.2. The rescaled vector r ∂θ is of norm 1.

2.7. Source, sink, circulation In this section we will describe some illustrative examples of vector fields.

∂ Example 2.7.1 (Zero of type source/sink). The vector field ∂r in ∂ ∂ ∂ the plane is undefined at the origin, but r ∂r = x ∂x + y ∂y has a smooth extension which is a vector field vanishing at the origin.

∂ Definition 2.7.2. The vector field in the plane defined by r ∂r is called a source while the opposite vector field X = r ∂ is called a − ∂r sink.3 Note that the integral curves of a source flow from the origin and away from it, whereas the integral curves of a sink flow into the ori- gin, and converge to it for large time. At a point p R2 with polar coordinates (r, θ) the sink is given by ∈ r ∂ if p =0 X(p)= − ∂r 6 (0 if p =0. Example 2.7.3 (Zero of type “circulation” in the plane). The vec- ∂ tor ∂θ in the plane, viewed as a tangent vector at a point at distance r from the origin, tends to zero as r tends to 0, as is evident from (2.6.1). Therefore the vector field defined by ∂ on R2 0 extends by continu- ∂θ \{ } ity to the point p = 0. Moreover it equals y ∂ +x ∂ and hence smooth. − ∂x ∂y 2An alternative argument can be given in terms of differentials. Since dr2 = dx2 + dy2 by Pythagoras, we have dr = 1 as well. Meanwhile θ = arctan y and 2 x 1 x| |xdy−ydx xdy−ydx therefore dθ = 1+(y/x)2 d(y/x)= y2+x2 x2 = r2 . Hence xdy ydx r 1 dθ = | − | = = . (2.6.2) | | r2 r2 r Thus rdθ is a unit covector. We therefore have an orthonormal basis (dr, rdθ) for ∂ the cotangent space. Since dθ ∂θ = 1, equation (2.6.2) implies (2.6.1). Therefore 1 ∂ is a unit vector. r ∂θ  3Makor and kior (with kaf) or rather bor. 2.8.DUALITYINLINEARALGEBRA 25

Thus we obtain a continuous vector field p X(p)= y ∂ +x ∂ on R2 7→ − ∂x ∂y which vanishes at the origin: ∂ if p =0 X(p)= ∂θ 6 (0 if p =0. Such a vector field is sometimes described as having circulation4 around the point 0. The integral curves of a circulation are circles around the origin. Remark 2.7.4. The zero of the vector field associated with small oscillations of the pendulum is of circulation type. These were studied in e.g., [Kanovei et al. 2016]. Next we discuss zeros of type circulation on the sphere. ∂ Lemma 2.7.5. The vector field ∂θ on the sphere has two zeros of circulation type, namely north and south poles. Proof. Spherical coordinates (ρ,θ,ϕ) in R3 restrict to the unit sphere S2 R3 to give coordinates (θ,ϕ) on S2. The north pole is defined by⊆ϕ = 0. At this point, the angle θ is undefined but the ∂ vector field ∂θ can be extended by continuity as in the plane (see Ex- ample 2.7.3), and we obtain a zero of circulation type going counter- ∂ clockwise. Similarly the south pole is defined by ϕ = π. Here ∂θ has a zero of circulation type but going clockwise with respect to the natural orientation on the 2-sphere. 

2.8. Duality in linear algebra Let V be a real vector space. We will assume all vector spaces to be finite dimensional unless stated otherwise. Example 2.8.1. Euclidean space Rn is a real vector space of di- mension n.

Example 2.8.2. The tangent plane TpM of a regular surface M (see Definition 1.1.4) at a point p M is a real vector space of dimension 2. ∈ Definition 2.8.3. A linear form, also called 1-form, φ on a vector space V is a linear functional from V to R.

Definition 2.8.4. The of V , denoted V ∗, is the space of all linear forms φ on V . Namely,

V ∗ = φ: φ is a 1-form on V . { } 4machzor, tzirkulatsia. 26 2. DERIVATIONS, TANGENT AND COTANGENT BUNDLES

Evaluating φ at an element x V produces a scalar φ(x) R. ∈ ∈ Definition 2.8.5. The natural pairing between V and V ∗ is a , : V V ∗ R, h i × → defined by setting x,y = y(x), for all x V and y V ∗. h i ∈ ∈ Theorem 2.8.6. If V admits a basis of vectors (xi), then V ∗ ad- mits a unique basis, called the dual basis (yj), satisfying x ,y = δ , (2.8.1) h i ji ij for all i, j =1,...,n, where δij is the Kronecker delta function. ∂ ∂ Example 2.8.7. The vectors ∂x and ∂y form a basis for the tangent plane T E of the Euclidean plane E at each point p E. The dual p ∈ space is denoted Tp∗E and called the cotangent plane.

∂ ∂ Definition 2.8.8. The basis dual to ∂x , ∂y is denoted (dx,dy). Thus (dx,dy) is a basis for the cotangent plane T ∗at every point p E. p ∈ Polar coordinates will be dealt with in detail in Subsection 2.9. They provide helpful examples of vectors and 1-forms, as follows. ∂ ∂ Example 2.8.9. In polar coordinates (r, θ), we have a basis ( ∂r , ∂θ ) for the tangent plane TpE of the Euclidean plane E at each point p ∂ ∂ ∈ E 0 . The dual space T ∗ has a basis denoted (dr, dθ) dual to , . \{ } p ∂r ∂θ Example 2.8.10. In polar coordinates, the 1-form  rdr occurs frequently in calculus. This 1-form vanishes at the origin (de- fined by the condition r = 0), and gets “bigger and bigger” as we get further away from the origin, as discussed in Section 2.9.

2.9. Polar, cylindrical, and spherical coordinates Polar coordinates5 (r, θ) satisfy r2 = x2 + y2 and x = r cos θ, y = r sin θ. In R2 0 , one way of defining the ranges for the variables is to require \ { } r> 0 and θ [0, 2π). ∈ It is shown in elementary calculus that the area of a region D in the plane in polar coordinates is calculated using the area element dA = rdrdθ.

5koordinatot koteviot 2.9. POLAR, CYLINDRICAL, AND SPHERICAL COORDINATES 27

Thus, the area is expressed by the following integral:

area(D)= dA = rdrdθ. ZD ZZ Cylindrical coordinates in Euclidean 3-space are studied in vector cal- culus. Definition 2.9.1. Cylindrical coordinates (koordinatot gliliot) (r,θ,z) are a natural extension of the polar coordinates (r, θ) in the plane. The volume of an open region D is calculated with respect to cylin- drical coordinates using the volume element dV = rdrdθdz. Thus the volume of D can be expressed as follows:

vol(D)= dV = rdrdθdz. ZD ZZZ Example 2.9.2. Find the volume of a right circular cone with height h and base a circle of radius b. Spherical coordinates6 (ρ,θ,ϕ) in Euclidean 3-space are studied in vector calculus. Definition 2.9.3. Spherical coordinates (ρ,θ,ϕ) are defined as fol- lows. The coordinate ρ is the distance from the point to the origin, satisfying ρ2 = x2 + y2 + z2, or ρ2 = r2 +z2, where r2 = x2 +y2. If we project the point orthogonally to the (x,y)-plane, the polar coordinates of its image, (r, θ), satisfy x = r cos θ and y = r sin θ. Definition 2.9.4. The coordinate ϕ of a point in R3 is the angle between the position vector of the point and the third basis vector e3 = (0, 0, 1)t in 3-space. Thus z = ρ cos ϕ while r = ρ sin ϕ.

6koordinatot kaduriot 28 2. DERIVATIONS, TANGENT AND COTANGENT BUNDLES

Remark 2.9.5. The ranges of the coordinates are often chosen as follows: 0 ρ, while 0 θ 2π, and 0 ϕ π ≤ ≤ ≤ ≤ ≤ (note the different upper bounds for θ and ϕ). Recall that the volume of a region D R3 is calculated using a volume element of the form ⊆ dV = ρ2 sin ϕ dρdθdϕ, so that the volume of a region D is

vol(D)= dV = ρ2 sin ϕ dρdθdϕ. ZD ZZZD Example 2.9.6. Calculate the volume of the spherical shell between spheres of radius α> 0 and β α. ≥ Remark 2.9.7. Consider a sphere Sρ of radius ρ = β. The area of a spherical region on Sρ is calculated using the area element 2 dASρ = β sin ϕ dθdϕ. Thus the area of a spherical region D S is given by the integral ⊆ β 2 area(D)= dASρ = β sin ϕ dθdϕ. ZD ZZ Example 2.9.8. Calculate the area of the spherical region on a sphere of radius β included in the first octant, (so that all three Carte- sian coordinates are positive).

2.10. Cotangent space and cotangent bundle Derivations were already discussed in Section 1.7. Recall that the tangent space T M at p M is the space of derivations at p. p ∈ Definition 2.10.1. The vector space dual to the tangent space Tp is called the cotangent space, and denoted Tp∗. Thus an element of a tangent space is a vector, while an element of a cotangent space is called a 1-form, or a covector.

Definition 2.10.2. As a set, the cotangent bundle, denoted T ∗M, of an n-dimensional manifold M is the disjoint union of all cotangent spaces Tp∗M as p ranges through M, or in formulas:

T ∗M = Tp∗M. p M [∈ 2.10. COTANGENT SPACE AND COTANGENT BUNDLE 29

∂ Definition 2.10.3. The basis dual to the basis ( ∂ui )i=1,...,n is de- noted (dui), i =1,...,n.

Theorem 2.10.4. The cotangent bundle T ∗M of a smooth n-dimen- sional manifold M is a smooth manifold of dimension 2n. i Thus each du is by definition a 1-form on Tp, or a cotangent vector ∂ (covector for short). We are therefore working with dual bases ∂ui for vectors, and (duj) for covectors. The pairing as in (2.8.1) gives  ∂ ∂ ,duj = duj = δj, (2.10.1) ∂ui ∂ui i     where δj is the Kronecker delta: δj =1 if i = j and δj =0 if i = j. i i i 6 Example 2.10.5. Examples of 1-forms in the plane are dx, dy, dr, rdr, dθ. In analogy with vector fields, we define a differential 1-form as a section of the cotangent bundle T ∗M; see Chapter 4.

CHAPTER 3

Metric differential geometry

In chapters 1 and 2, we dealt mainly with differentiable manifolds M without additional structure, other than the existence of a distance function to ensure metrizability; see Section 1.2. In the present chapter we will deal more systematically with lengths and distances on M defined via the first fundamental forms and Rie- mannian metrics. This prepares the ground for Gromov’s systolic inequality dealt with in Chapter 10, Loewner’s systolic inequality for the torus in Chapter 11, and Pu’s systolic inequality for the real projective plane in Chapter 12.

3.1. Isometries, constructing bilinear forms out of 1-forms Definition 3.1.1. A metric at a point p M is a symmetric, ∈ positive definite, bilinear form on the tangent space Tp. A smooth assignment of a form at every point of M results in a globally defined metric on M. A more detailed definition appears in Section 3.2. Definition 3.1.2. A differentiable manifold equipped with a metric is called a . A metric enables us to measure the length of paths in M. This is done by integrating the norm of the tangent vector of a parametrisation of the path. Definition 3.1.3. A map f between Riemannian manifolds M and N are called isometry if it preserves the length of all paths. Thus, if γ : [0, 1] M is a smooth path in M then γ = f γ , where absolute values→ denote the length. | | | ◦ | Definition 3.1.4. Manifolds M and N are called isometric if there exists a one-to-one, onto isometry between them. Proposition 3.1.5. To prescribe a metric it is sufficient to pre- scribe the corresponding quadratic form. 31 32 3.METRICDIFFERENTIALGEOMETRY

Proof. The polarisation formula allows one to reconstruct a sym- metric bilinear form B, from the quadratic form Q(v) = B(v,v), at least if the characteristic is not 2: 1 B(v,w)= (Q(v + w) Q(v w)). (3.1.1) 4 − − Thus the bilinear form can be recovered from the quadratic form.  One can construct quadratic forms out of the 1-forms dui on M. Namely a quadratic form is a linear combination of the rank-1 quadratic forms (dui)2. Proposition 3.1.6. A positive definite quadratic form (over the i 2 + field of scalars R) can be represented as a sum ai(du ) where ai R i ∂ ∈ and (du ) is the basis dual to ∂ui . Proof. The proof is immediate  from orthogonal diagonalisation of symmetric matrices.1 

Each bilinear form on the tangent space TpM is the polarisation of a suitable quadratic form (3.1.1).

3.2. Riemannian metric, first fundamental form Let M be a manifold, and p M a point on the manifold. Recall ∈ 2 that the tangent space Tp at a point p is the fiber (i.e., inverse image of a point) of the tangent bundle π : TM M. M → Definition 3.2.1. A Riemannian metric g on M is choice of a symmetric positive definite bilinear form on the tangent space Tp = TpM at p: g: T T R, p × p → defined for all p M and varying smoothly as a function of p. ∈ Example 3.2.2. The more naive viewpoint3 is as follows. Here we use the terminology first fundamental form instead of Riemannian metric. Consider an open set U R2, with coordinates (u1,u2). As- sume U is embedded in Euclidean⊆ 3-space R3 by means of a regular map x(u1,u2), x: U R3 . → 1 See http://u.math.biu.ac.il/~katzmik/88-201.html for details. 2siv 3Such a viewpoint is taken in the course 88201; see http://u.math.biu.ac. il/~katzmik/88-201.html 3.3. METRIC AS SUM OF SQUARED 1-FORMS; ELEMENT OF LENGTH ds 33

The image of x is a surface M R3. The tangent plane of the resulting surface M in R3 is spanned by⊆ vectors ∂x x = where i =1, 2. (3.2.1) i ∂ui The ambient space R3 is equipped with a standard inner product de- noted , R3 . The restriction of the inner product to the tangent plane of M givesh i a first fundamental form on M:

g: T T R, g(v,w)= v,w R3 . p × p → h i The first fundamental form is traditionally expressed by a matrix of coefficients called metric coefficients gij, with respect to coordi- nates (u1,u2) on the surface. 3 Definition 3.2.3. The metric coefficient gij of a surface M R parametrized by x is given by the inner product of the i-th and the⊆j-th vector of the basis of TpM: ∂x ∂x gij = i , j . ∂u ∂u 3  R In particular, the diagonal coefficient satisfies ∂x 2 g = , ii ∂ui

∂x namely it is the square length of the i-th basis vector i . ∂u 3.3. Metric as sum of squared 1-forms; element of length ds In this section, we go beyond the naive viewpoint summarized in Section 3.2 and adopt the modern viewpoint developed in the previous chapter. In the case of surfaces, at every point p = (u1,u2), we have the 1 2 metric coefficients gij = gij(u ,u ). Each metric coefficient is thus a scalar-valued function of two variables. Consider the case when the matrix (gij) is diagonal. This can al- ways be achieved at a point by a change of coordinates (see Propo- sition 3.1.6). Suppose this is true everywhere in a coordinate chart.4 Then in the notation developed in Section 3.1, we can write the Rie- mannian metric as follows: 1 2 1 2 1 2 2 2 g = g11(u ,u )(du ) + g22(u ,u )(du ) . (3.3.1) 4Such a property is weaker than having isothermal coordinates where one re- quires the matrix to be a scalar matrix at every point (possibly depending on the point); see Section 3.9. 34 3.METRICDIFFERENTIALGEOMETRY

Example 3.3.1. If the metric coefficients form an identity matrix, we obtain the standard flat metric

2 2 g = du1 + du2 . (3.3.2)

Here the interior superscript (inside  the parentheses) denotes an index, while exterior superscript denotes the squaring operation.

Sometimes it is convenient to simplify notation by replacing the coordinates (u1,u2) simply by (x,y). Then the standard flat met- ric (3.3.2) is g = dx2 + dy2. Often the metric is expressed in terms of the length element ds by writing

ds2 = dx2 + dy2. (3.3.3)

3.4. Hyperbolic metric

The hyperbolic metric ghyp in the upperhalf plane (x,y): y > 0 is the Riemannian metric expressed by the quadratic form{ } 1 g = (dx2 + dy2), where y > 0. (3.4.1) hyp y2

1 2 2 Thus its length element ds is ds = y dx + dy . Note that this ex- pression is undefined for y = 0. The hyperbolic metric in the upper p half plane is a complete metric.

Theorem 3.4.1. The Gaussian curvature K = K(x,y) of the met- ric (3.4.1) satisfies K = 1 at every point. − Proof. We use the formula for for the Gaussian curvature involv- 1 ∂2 ∂2 ing the Laplace-Beltrami operator ∆LB = f 2 ∂x2 + ∂y2 for the metric f 2(dx2 + dy2). Namely, we have  

K = ∆ ln f for the metric f 2(dx2 + dy2), (3.4.2) − LB 1 5 and use the fact that the second derivative of ln y is 2 .  − y It is a deep theorem that upperhalf plane with the hyperbolic metric does not admit a global isometric embedding in Euclidean space.

5 For details see http://u.math.biu.ac.il/~katzmik/egreglong.pdf. 3.5. SURFACE OF REVOLUTION IN R3 35

3.5. Surface of revolution in R3 In Section 3.3, we developed a formalism for describing a metric on an arbitrary surface, in coordinates (u1,u2). In the special case of a surface of revolution in R3, it is customary to use the notation u1 = θ and u2 = ϕ for the coordinates, as in formula (3.5.2) below. Then the Riemannian metric g of (3.3.1) can then be written as6

2 2 g = g11(θ,ϕ)(dθ) + g22(θ,ϕ)(dϕ) which can be abbreviated as

2 2 g = g11(θ,ϕ) dθ + g22(θ,ϕ) dϕ . The starting point in the construction of a surface of revolution in R3 is an embedded regular curve C in the xz-plane, called the generating curve. The curve C is parametrized by a pair of functions x = f(ϕ), z = g(ϕ). (3.5.1) We will assume that f(ϕ) > 0 for all ϕ. Definition 3.5.1. The surface of revolution (around the z-axis), defined by the curve C, is the surface parametrized as follows: x(θ,ϕ)=(f(ϕ)cos θ,f(ϕ)sin θ,g(ϕ)). (3.5.2) The condition f(ϕ) > 0 ensures that the resulting surface is em- bedded in R3. Remark 3.5.2. If the generating curve C R2 is a Jordan curve in the plane (and, as before, f(ϕ) > 0) then the⊆ resulting surface is an embedded torus; see Section 3.10. Definition 3.5.3. A pair of functions (f(ϕ),g(ϕ)) provides an arc- length parametrisation of the curve C if

2 2 ϕ, df + dg =1. (3.5.3) ∀ dϕ dϕ     Proposition 3.5.4. Let (f(ϕ),g(ϕ)) be an arclength parametri- sation of the generating curve. Then in the notation of Section 3.3, the metric g of the surface of revolution is given by the formula g = f 2(ϕ)dθ2 + dϕ2.

6We boldface the g of the metric mainly to distinguish it from the function g occurring below in (3.5.1). 36 3.METRICDIFFERENTIALGEOMETRY

Proof of Proposition 3.5.4. To calculate the first fundamen- tal form of the surface of revolution (3.5.2), note that the tangent vectors (see (3.2.1)) are x = ∂x =( f sin θ,f cos θ, 0)t, 1 ∂θ − while t ∂x df df dg x2 = ∂ϕ = dϕ cos θ, dϕ sin θ, dϕ . 2 2 2 2 2 Thus we have g11 = f sin θ + f cos θ = f and 2 2 2 2 df 2 2 dg df dg g22 = dϕ (cos θ + sin θ)+ dϕ = dϕ + dϕ . Furthermore,        df df g = f sin θ cos θ + f cos θ sin θ =0. 12 − dϕ dϕ Thus the Riemannian metric or the first fundamental form is f 2 0 2 2 (gij)= df dg . (3.5.4) 0 dϕ + dϕ !     If ϕ is an arclength parameter then g22 = 1 and the proposition follows from (3.5.3).  Using 1-forms in place of matrix notation, equation (3.5.4) can be reformulated as follows. Corollary 3.5.5. Assume (f(ϕ),g(ϕ)) is a regular parametrisa- tion of the generating curve. In the notation of Section 3.3, the Rie- mannian metric g of the surface of revolution is given by the formula 2 2 2 2 df dg 2 g = f dθ + dϕ + dϕ dϕ .      Example 3.5.6. The unit sphere is a surface of revolution whose generating curve C is the circle. Setting f(ϕ) = sin ϕ and g(ϕ) = cos ϕ, we obtain a parametrisation of the sphere S2 in spherical coordinates, with respect to which the metric takes the form 2 2 2 gS2 = sin ϕdθ + dϕ . (3.5.5) 2 In other words, g11 = sin ϕ and g22 = 1. The corresponding area form is therefore sin ϕdθdϕ. Remark 3.5.7. As in (3.5.5), in general the first fundamental form of a surface of revolution as parametrized above is diagonal but not nec- essarily scalar. In Section 3.9, we will perform an appropriate change of coordinates so as to express the metric of a surface of revolution by 3.6. COORDINATE CHANGE 37 a first fundamental form given by a scalar matrix (where the value of the scalar is a function of the point of the surface).

3.6. Coordinate change In this section we will compare different coordinate charts for a sur- face M R3. Given a metric in one coordinate chart, we would like to understand⊆ how the metric coefficients tranform under change of coor- dinates. Consider coordinate charts (A, (ui)) and (B, (vα)). Consider a change from a coordinate chart

(ui), i =1, 2, to the coordinate chart

(vα), α =1, 2.

In the overlap A B of the two domains, the coordinates can be ex- pressed in terms of∩ each other, e.g., v = v(u).

Definition 3.6.1. We will use the following notation: 1 2 gij = gij(u ,u ) for the metric coefficients of M with respect • to chart (ui); 1 2 g˜αβ =g ˜αβ(v ,v ) for the metric coefficients with respect to the • chart (vα).

We will use the Einstein summation convention; see Definition 2.2.2.

Proposition 3.6.2. Consider a surface M R3. Under a coordi- nate change, the metric coefficients transform as⊆ follows:

∂ui ∂uj g˜ = g . αβ ij ∂vα ∂vβ Proof. By assumption, the metric in a neighborhood of M is in- duced by a Euclidean embedding

x: U R3, → defined by x = x(u) = x(u1,u2). Changing coordinates, we obtain a new parametrisation y(v)= x(u(v)). 38 3.METRICDIFFERENTIALGEOMETRY

Applying chain rule, we obtain

∂y ∂y g˜ = , αβ ∂vα ∂vβ   ∂x ∂ui ∂x ∂uj = , ∂ui ∂vα ∂uj ∂vβ   ∂ui ∂uj ∂x ∂x = , ∂vα ∂vβ ∂ui ∂uj   ∂ui ∂uj = g , ij ∂vα ∂vβ completing the proof. 

3.7. Conformal equivalence i j i j Definition 3.7.1. Two metrics, g = gijdu du and h = hijdu du , on a surface M are called conformally equivalent, or conformal for short, if there exists a function f = f(u1,u2) > 0 such that

g = f 2h, in other words, 2 gij = f hij for all i,j. (3.7.1)

Definition 3.7.2. The function f above is called the conformal factor of the metric g with respect to the metric h.

Remark 3.7.3. Sometimes it is more convenient to refer, instead, to the function λ = f 2 as the conformal factor.

1 2 2 Example 3.7.4. The hyperbolic metric y2 (dx +dy ) of Section 3.4 is conformally equivalent to the flat one (dx2 + dy2), with conformal 1 factor f(x,y)= y . Lemma 3.7.5. Under a conformal change of metric with conformal factor f, the length of every vector at a point p =(u1,u2) is multiplied by the factor f(u1,u2).

i ∂ Proof. Consider a vector v = v ∂ui at a point p which is a unit vector for the metric h. Let us show that v is “stretched” by a factor of f(p). In other words, its length with respect to g equals f(p). Indeed, 3.8. LATTICES, UNIFORMISATION THEOREM FOR TORI 39 the new length of v is 1 ∂ ∂ 2 g(v,v)= g vi ,vj ∂ui ∂uj   p 1 ∂ ∂ 2 = g , vivj ∂ui ∂uj     i j = gijv v

2 i j = pf (p)hijv v

q i j = f(p) hijv v = f(p).p  Definition 3.7.6. An equivalence class of metrics on a surface M conformal to each other is called a conformal structure7 on M.

3.8. Lattices, uniformisation theorem for tori The following result is a consequence of the uniformisation theorem. Theorem 3.8.1. Locally, every metric on a surface can be written as f(x,y)2(dx2 + dy2) with respect to suitable coordinates (x,y). There is a global version of this theorem for tori, namely the uni- formisation theorem in the genus 1 case (i.e., for tori); see Theo- rem 3.8.8. Recall that by the Pythagorian theorem, the square-length of a vector in the plane equipped with the standard metric is the sum of the squares of its components, or ds2 = dx2 + dy2. (3.8.1) We will use the notation ds2 of (3.8.1) as shorthand for this standard flat metric, as in (3.4.2). Definition 3.8.2 (Lattice). A lattice L R2 a subgroup isomor- phic to Z2 which is not included in any line in⊆ R2.

2 Example 3.8.3. The Gaussian integers LG R is the lattice con- sisting of points with integer coordinates: ⊆ L = (m,n) R2 : m,n Z . G { ∈ ∈ } 7mivneh 40 3.METRICDIFFERENTIALGEOMETRY

Definition 3.8.4. Let L be a lattice in the plane. A function f(p) on R2 is called L-periodic if f(p + ℓ) = f(p) for all ℓ L and all p = (x,y) R2. ∈ ∈ Definition 3.8.5. A flat torus is a quotient R2 /L, where L is a lattice in the plane.

Remark 3.8.6. Note that the constant 1-forms dx and dy in the plane are in particular invariant under the translations R2 R2, p p + ℓ where ℓ L. This enables us to view dx and dy as→ 1-forms on7→ the torus R2 /L∈, as well.

2 Example 3.8.7. Let c,d R 0 . A rectangular lattice Lc,d R is defined by ∈ \{ } ⊆ L = cme + dne : n,m Z . c,d { 1 2 ∈ } One can also view Lc,d as a subset of C given by

L = cm + dni: n,m Z . c,d { ∈ } A more detailed version of the Uniformisation Theorem 3.8.1 in the case of tori characterizes Riemannian tori up to isometry (see Defini- tion 3.1.4) in terms of the corresponding lattice, as follows.

Theorem 3.8.8 (Uniformisation theorem for tori). For every met- ric g on the 2-torus T2, there exists a lattice L R2 and a positive L- periodic function f(x,y) on R2 such that the torus⊆ (T2, g) is isometric to R2/L,f 2ds2 , (3.8.2) where ds2 = dx2 + dy2 is the standard flat metric of R2.8

8More generally, there is also a global form of the theorem in arbitrary genus, which can be stated in several ways. One of such ways involves Gaussian curva- ture: Every metric on a connected (kashir) surface is conformally equivalent to a metric of constant Gaussian curvature. From the complex analytic viewpoint, the uniformisation theorem states that every Riemann surface is covered by ei- ther the sphere, the plane, or the upper halfplane. Thus no notion of curvature is needed for the statement of the uniformisation theorem. However, from the dif- ferential geometric point of view, what is relevant is that every conformal class of metrics contains a metric of constant Gaussian curvature. See [Ab81] for a lively account of the history of the uniformisation theorem. More information on the uni- formisation theorem and the Riemann mapping theorem can be found at https:// mathoverflow.net/q/10516. 3.9. ISOTHERMAL COORDINATES 41

3.9. Isothermal coordinates Definition 3.9.1. Coordinates (u1,u2) in a neighborhood U M ⊆ are called isothermal if the matrix (gij) of the first fundamental form (metric) of M, with respect to (u1,u2), is a scalar matrix at every point of U.

In terms of the metric coefficients gij, the isothermal condition is expressed by the pair of equations g11 = g22 and g12 = 0. The following lemma expresses the metric of a surface of revolution in isothermal coordinates. Theorem 3.9.2. Suppose (f(ϕ),g(ϕ)), where f(ϕ) > 0, is an arc- length parametrisation of the generating curve9 of a surface of revolu- tion in R3. Consider the change of variable 1 ψ = dϕ. (3.9.1) f(ϕ) Z With respect to the new parametrisation of the surface in terms of vari- ables (θ,ψ), the metric becomes f 2(ϕ(ψ))(dθ2 + dψ2), 2 so that the matrix of metric coefficients is the scalar matrix (f δij). A proof is given below following Corollary 3.9.3. Corollary 3.9.3. The change of variables (3.9.1) produces an ex- plicit conformal equivalence between the metric on the surface of revo- lution and the standard flat metric dθ2 + dψ2 in the (θ,ψ) coordinates. The existence of such a parametrisation is predicted by the uni- formisation theorem (see Theorem 3.8.8) in the case of a general regular surface. The advantage of formula (3.9.1) is its explicit nature. Proof of Theorem 3.9.2. Let ϕ = ϕ(ψ). We assume that the dependence of ϕ on ψ is monotone. Then there exists an inverse func- df df dϕ tion ψ = ψ(ϕ). By chain rule, dψ = dϕ dψ . Now consider again the first fundamental form of Proposition 3.5.5. To impose the condition

g11 = g22, (3.9.2) 2 2 2 df dg we need to solve the equation f = dψ + dψ , or     df 2 dg 2 dϕ 2 f 2 = + . dϕ dϕ dψ     !   9This curve does not have to be closed. We will specialize to Jordan curves in Section 3.10. 42 3.METRICDIFFERENTIALGEOMETRY

In the case when the generating curve is parametrized by arclength, we dϕ are therefore reduced to the equation f = dψ . Equivalently, we have dϕ ψ = . f(ϕ) Z Thus we obtain a parametrisation of the surface of revolution in coor- dinates (θ,ψ), such that the matrix of metric coefficients satisfies the relation (3.9.2) and is therefore a scalar matrix. 

3.10. Tori of revolution If the generating curve C R2 is a Jordan curve and f(ϕ) > 0 then the resulting surface is a torus⊆ of revolution. Assuming that ϕ is the arclength paramenter, the change of vari- dϕ ables ψ = f(ϕ) results in isothermal coordinates (θ,ψ) on the torus. The lattice L of such a torus in the (θ,ψ)-plane is rectangular (see R Example 3.8.7). The following lemma is immediate from the results of the previous section. Lemma 3.10.1. Suppose C is a Jordan curve in the plane. For the isothermal (θ,ψ)-parametrisation of the torus of revolution with gener- ating curve C, the variable θ ranges from 0 to 2π, while the variable ψ µ dϕ varies from 0 to 0 f(ϕ) where µ is the length of the curve C. The result ofR Section 3.9 has the following corollary. 2 3 Corollary 3.10.2. Consider a torus of revolution (T , grev) in R formed by rotating a Jordan curve of length µ > 0, with unit speed parametisation (f(ϕ),g(ϕ)) where ϕ [0,µ]. Then the torus is con- ∈2 formally equivalent to a flat torus R /Lc,d, defined by a rectangular lattice L R2, where c =2π and d = µ dϕ . The metric in coordi- c,d ⊆ 0 f(ϕ) nates (θ,ψ) is given by R 2 2 2 grev = f (ϕ(ψ))(dθ + dψ ), dϕ for the change of coordinate ψ = f(ϕ) . 3.11. Standard fundamentalR domain, conformal parameter τ In this section we undertake a more detailed study of flat tori or equivalently, lattices in the plane. Identifying R2 with C, we can think of a lattice L as a subset of C. Definition 3.11.1. We say that lattice L C is similar to lattice ⊆ L′ C if there exists a number λ C such that L = λL′. ⊆ ∈ 3.12. CONFORMAL PARAMETER OF RECTANGULAR LATTICES 43

Figure 3.11.1. Torus: lattice (left) and embedding (right)

Writing λ = reiθ we see that similar lattices differ by a rotation and multiplication by a real scalar. Similarly, we will say that the corresponding tori C/L and C/L′ are similar. The standard fundamental domain in the complex plane is impor- tant in many branches of mathematics, and is defined as follows.

Definition 3.11.2. The standard fundamental domain D0 C is the region ⊆ D = z = x + iy C: 1 x< 1 , y> 0, z 1 . (3.11.1) 0 { ∈ − 2 ≤ 2 | |≥ } Theorem 3.11.3. Every flat torus C/L is similar to the torus ob- tained as the quotient of C by the lattice spanned by the pair (1,τ) where τ D0 as in (3.11.1). The parameter τ is called the conformal parameter∈10 of the torus. A proof appears in Section 11.2 in the context of a proof of Loewner’s systolic inequality in Chapter 11. 3.12. Conformal parameter of rectangular lattices

Consider the rectangular lattice Lc,d C where c > 0 and d > 0 (see Example 3.8.7). Thus L = mc + ndi⊆ : m,n Z . c,d { ∈ } Lemma 3.12.1. The conformal parameter τ of the lattice Lc,d is c d τ(Lc,d)= i max d , c . 1 Proof. Suppose d c. Then we scale the lattice by a factor of c to obtain a similar lattice≥ ′ d Z L′ = L d = m + n i: m,n . 1, c c ∈  10Not to be confused with conformal factor. 44 3.METRICDIFFERENTIALGEOMETRY

d d Since c 1 we have c i D0 and therefore the conformal parameter ≥ d ∈ of L′ is τ = c i. 1 Now suppose c d. Then we scale the lattice by a factor of d ≥ π and also multiply by ei 2 (rotation by a right angle). We then obtain a similar lattice ′ c L′ = L c = m + n i: m,n Z . 1, d d { ∈ } In this case the conformal parameter is τ = c i D .  d ∈ 0 In Section 3.10 we studied the torus of revolution generated by a Jordan curve of length µ with arclength parametrisation (f(ϕ),g(ϕ)). If a lattice is rectangular, the conformal parameter τ is pure imaginary. We therefore obtain the following corollary. Corollary 3.12.2. The conformal parameter τ of a torus of rev- olution is pure imaginary:

τ = iy0 of absolute value y = max c , d 1, 0 d c ≥ µ dϕ where c = 2π and d = 0 f(ϕ) , where µ is the length of the generating Jordan curve. R

3.13. Conformal parameter of standard tori of revolution We will use the expression in terms of an integral obtained in Sec- tion 3.12 to compute the conformal parameter τ of the standard tori embedded in R3. Definition 3.13.1. Let a,b R such that 0

Theorem 3.13.2. The flat torus in the conformal class of the torus of revolution Ta,b is given by a rectangular lattice in the (θ,ψ) plane of the form Lc,d where c =2π and 2π d = . (a/b)2 1 − The conformal parameter τ ofp the torus Ta,b is 1 τ = i max , (a/b)2 1 . (a/b)2 1 − ! − p Proof. As in Corollaryp 3.12.2, we replace ϕ by ϕ(ψ) to produce isothermal coordinates (θ,ψ) for the standard torus of revolution parametrized as in (3.13.2), where dϕ dϕ ψ = = . f(ϕ) a + b cos ϕ Z Z b Therefore the flat metric is defined by a lattice in the (θ,ψ) plane with c =2π and µ=2πb dϕ d = . (3.13.3) a + b cos ϕ Z0 b ϕ To evaluate the integral (3.13.3), we change the variable to t = b and a let α = b . Then 2π dt d = . α + cos t Z0 the proof is now immediate from Lemma 3.13.3 below by a residue calculation. 

3.13.1. A residue calculation. The material in this subsection is op- tional. Lemma 3.13.3. Let α > 1. We have the following value of the definite integral: 2π dt 2π = . 2 0 α + cos t √α 1 Z − Proof. We will use the residue theorem for complex functions. First note that 2π dt 2π dt = α + cos t α + Re(eit) Z0 Z0 2π 2dt = . 2α + eit + e it Z0 − 46 3.METRICDIFFERENTIALGEOMETRY

it idz The change of variables z = e yields dt = −z . We thus obtain the following integral along the unit circle: 2idz 2idz − = − z(2α + z + z 1) z2 + 2αz + 1 I − I (3.13.4) 2idz = − , (z λ )(z λ ) I − 1 − 2 where λ = α + α2 1, λ = α α2 1. (3.13.5) 1 − − 2 − − − From (3.13.5) we obtain p p λ λ = 2 α2 1. (3.13.6) 1 − 2 − The root λ2 is outside the unit circle.p Thus z = λ1 is the only singu- larity of the meromorphic function under the integral sign in (3.13.4) in- side the unit circle. Hence we need to compute the residue at λ1 so as to apply the residue theorem. The residue at λ1 is obtained by multiplying through by (z λ1) and then evaluating at z = λ1. We therefore obtain − 2i 2i i from (3.13.6) Res 1 = − = − 2 = −2 . The integral is deter- λ λ1 λ2 2√α 1 √α 1 − − − mined by the residue theorem in terms of the residue at the pole z = λ1. Therefore the integral equals 2πi Res = 2π , proving the lemma and λ1 √α2 1 the theorem. − 

3.13.2. Function as a section of trivial bundle (E,B,π). This material is optional. Before dealing with general bundles, let us make some elementary re- marks about graphs of functions. Let M be a manifold. Consider a real- valued function f : M R. → Definition 3.13.4. The graph of f in M R is the collection of pairs × (p,y) M R: y = f(p) . { ∈ × } We view M R as a trivial bundle (i.e., product bundle) over M, de- × noted πM : π : M R M. M × → We can also consider the section s: M M R defined by the formula → × s(p)=(p,f(p)). Then the graph of f is the image of s in M R. The section thus defined clearly satisfies the identity × π s = Id , M ◦ M in other words πM (s(p)) = p for all p M. We will now develop the language of bundles and sections in the context∈ of the tangent and cotangent bundles of M. 3.13. CONFORMAL PARAMETER OF STANDARD TORI OF REVOLUTION 47

The total space of a bundle is typically denoted E, the base is denoted B, and the surjective bundle projection is denoted π: E

π  B The whole package is referred to as the triple (E,B,π). (3.13.7)

3.13.3. M¨obius strip as a first example of a nontrivial bundle. This material is optional. The tangent bundle of a circle is a cylinder S1 R (see Theorem 2.2.5). The cylinder can be thought of as a trivial bundle× over the circle. A first example of a non-trivial bundle is provided by the M¨obius strip.11 The latter can be thought of as a non-trivial bundle over the circle, constructed as follows. Consider the circle C R3 parametrized by ⊆ α(θ) = 10(cos θ, sin θ, 0), (3.13.8) and its unit normal vector n(θ) = (cos θ, sin θ, 0). The circle can be extended to a parametrized M¨obius strip M R3 as follows. We first construct the boundary of M by tracing the endpoints⊆ of a “rotating” interval at every point of the circle: µ(θ)= α(θ) cos θ n(θ)+sin θ e . (3.13.9) ± 2 2 3 The resulting set (the image of µ) is the boundary of the M¨obius strip. We now “fill in” the pair of points cos θ n(θ)+sin θ e and cos θ n(θ)+sin θ e 2 2 3 − 2 2 3 by the interval joining them:   s (cos θ ) n(θ) + (sin θ ) e , where s [ 1, 1]. (3.13.10) 2 2 3 ∈ − Remark 3.13.5. The interval is rotating at half the speed of the parametri- sation of the circle (3.13.8). The result is that by the time we complete a full rotation around the circle, the interval will only be rotated by π. We thus obtain a M¨obius band with parametrisation x(θ,s) defined as follows: θ θ x(θ,s)= α(θ)+ s cos 2 n(θ)+ sin 2 e3 , (3.13.11) where 0 θ 2π and 1 s 1.    ≤ ≤ − ≤ ≤ 11tabaat mobius, retzuat mobius. 48 3.METRICDIFFERENTIALGEOMETRY

Definition 3.13.6. The projection π : M C (3.13.12) M → of M to the circle C collapses each interval θ θ α(θ)+ s cos 2 n(θ)+ sin 2 e3 (for a fixed θ) to its midpoint x(θ, 0) = α(θ) C defined by s = 0.  ∈   Theorem 3.13.7. The map πM of (3.13.12) defines a non-trivial inter- val bundle over the circle. Proof. Consider the M¨obius band parametrized as in (3.13.11). Its boundary ∂M is connected. Meanwhile the boundary of the cylinder S1 I has two connected components, each homeomorphic to a circle. Therefore× the bundle πM of (3.13.12) is not equivalent to the trivial bundle.  3.13.4. Hairy ball theorem. This material is optional. In Section 3.13.3, we discussed the simplest example of a nontrivial bun- dle, given by the M¨obius band. Another example of a nontrivial bundle is provided by the following famous result. Theorem 3.13.8 (The hairy ball theorem). There is no nonvanishing continuous tangent vector field on a sphere S2. A proof can be obtained from Corollary 13.8.3. Thus, if f is a continuous function on S2 that assigns a vector in R3 to every point p on a sphere S2 R3, such that f(p) is always tangent to the sphere at p, then there is at least⊆ one p such that f(p) = 0. The theorem was first stated by Henri Poincar´ein the late 19th century. The theorem is famously stated as “you can’t comb a hairy ball flat”, or sometimes, “you can’t comb the hair of a coconut”. It was first proved in 1912 by Brouwer. The theorem has the following corollary. Corollary 3.13.9. The tangent bundle of the sphere S2 is a nontrivial bundle. Proof. If the bundle were equivalent to the trivial bundle with fiber R2, 2 then the constant vector field e1 (the first basis vector of R ) would map to an everywhere nonvanishing vector field on S2, contradicting the hairy ball theorem 3.13.8.  CHAPTER 4

Differential forms, exterior derivative and algebra

To express the de Rham theorem and related results for a differen- tiable manifold M, we need to develop the language of differential k- forms on M.

4.1. Differential 1-form as section of cotangent bundle In Section 2.2, we defined the tangent bundle of M. We now define the dual object, the cotangent bundle, following Section 2.10. Definition 4.1.1. The cotangent bundle of M is the collection of pairs T ∗M = (p, ω): p M, ω T ∗ ∈ ∈ p where Tp∗ is the vector space dual to Tp. Theorem 4.1.2. The cotangent bundle of M n is a 2n-dimensional manifold.

Proof. We have a natural projection πM : T ∗M M. Consider a coordinate chart (u1,...,un) in a neighborhood A → M. We then obtain a basis, denoted (du1,...,dun), for the cotangent⊆ space at ev- ery point in A. Every cotangent vector ωp at a point p with coor- 1 n i 1 dinates (u ,...,u ) is a linear combination ωp = ωidu = ω1du + 2 n ω2du + +ωndu . Hence we can coordinatize T ∗M locally by the 2n- ···1 n tuple (u ,...,u , ω1,...,ωn). 

With respect to this coordinate chart in T ∗M, the standard projec- tion π : T ∗M M “forgets” the last n coordinates: M → 1 n 1 n πM (u ,...,u , ω1,...,ωn)=(u ,...,u ). Definition 4.1.3. A differential 1-form ω on M is a section of the cotangent bundle of M depending smoothly on p, so that πM (ω)=IdM . Thus at every point p M, a differential form ω gives a linear form ω : T M R. In coordinates,∈ a differential 1-form ω is given by p p → 1 n i ω = ωi(u ,...,u )du , where each of the ωi is a real-valued function of n variables. 49 50 4. DIFFERENTIAL FORMS, EXTERIOR DERIVATIVE AND ALGEBRA

4.2. From function to differential 1-form Definition 4.2.1. Given a differentiable manifold M, we denote by C∞(M) the space of infinitely differentiable real-valued functions on M.

In Section 1.7 we defined the notion of a derivation on M.

Definition 4.2.2. Let M be a differentiable manifold. Given a smooth function f C∞(M), we define a differential 1-form, de- noted df, on M, as follows.∈ For each vector field X we set

df(X)= Xf (4.2.1) at every p M and Xp TpM, where Xf denotes the evaluation of the derivation∈ X T ∈at the function f thought of as an element p ∈ p of Dp. Remark 4.2.3. The differential 1-form df is related to the gradient of f; the precise relation to the gradient will be clarified in Corol- lary 4.4.8 below.

Formula (4.2.1) applies at every point of M. Denoting ω = df, we see that ω (X )= X f R p p p ∈ where Xp is the value of the vector field X at the point p. Theorem 4.2.4. In coordinates (u1,...,un) in a neighborhood A M, we have ⊆ ∂f df = dui, (4.2.2) ∂ui with Einstein summation convention.

The theorem is a restatement of chain rule in several variables. The usual Leibniz rule for derivatives then implies the following.

Corollary 4.2.5 (Leibniz rule in terms of differential forms). We have the following version of the Leibniz rule for functions on M:

f,g C∞(M), d(fg)= f dg + g df. (4.2.3) ∀ ∈ Example 4.2.6. Let u1 = x and u2 = y for simplicity. Let f(x,y)= x2+y2. Then the 1-form df takes the following form: df =2xdx+2ydy. 4.4. GRADIENT & EXTERIOR DERIVATIVE; MUSICAL ISOMORPHISMS 51

4.3. Space Ω1(M) of differential 1-forms, exterior derivative

Lew ω be a 1-form on M, i.e., smooth map from M to T ∗M satis- fying π ω = Id . M ◦ M Definition 4.3.1. We denote by Ω1(M)= ω : π ω = Id { M ◦ M } the set of all differential 1-forms ω on M. Note that Ω1(M) is an infinite-dimensional vector space. The space Ω1(M) is by definition the space of sections of the cotangent bundle of M. The 1-form df was defined in (4.2.1). Definition 4.3.2. The exterior derivative d on functions is the R- linear map 1 d: C∞(M) Ω (M), f df → 7→ 1 from the space C∞(M) of smooth functions on M to the space Ω (M) of differential 1-forms on M. Corollary 4.3.3. The exterior derivative d satisfies the Leibniz rule (4.2.3).

4.4. Gradient & exterior derivative; musical isomorphisms The exterior derivative is defined by the differentiable structure on the manifold M without recourse to metrics. Nevertheless it is easier to grasp the geometric significance of the exterior derivatives on functions by comparing df to the gradient of f. Note that the gradient is only defined in reference to a metric. Definition 4.4.1. Let V, , be an inner product space. Let f be a function on V . The gradienth i f of f is defined by setting ∇ f,X = Xf, (4.4.1) h∇ i for every vector X V . ∈ Remark 4.4.2. This definition is independent of coordinates chosen in V , since no coordinates were used in the definition (4.4.1). In other words, the gradient depends only on the inner product in V but not on the coordinates. With respect to an orthonormal basis the gradient of f is given by the list of partial derivatives of f. Definition 4.4.3 (Musical isomorphisms). Let V, , be an inner product space. We define a map h i  ♭ ♭: V V ∗, X X → 7→ 52 4. DIFFERENTIAL FORMS, EXTERIOR DERIVATIVE AND ALGEBRA

(bemol or flat) by setting Y V, X♭(Y )= X,Y . ∀ ∈ h i Theorem 4.4.4. Choose coordinates (ui) in a neighborhood of the origin in the inner product space V in such a way that the standard ∂ ∂ basis ( ∂u1 ,..., ∂un ) is an orthonormal basis at the origin. Then ∂ ♭ = dui. ∂ui   In particular, the map ♭ is an .

∂ j ∂ Proof. Let X = ∂ui . We write Y in coordinates as Y = y ∂uj . Then ∂ ∂ ∂ ∂ X,Y = ,yj = yj , = yjδ = yi. (4.4.2) h i ∂ui ∂uj ∂ui ∂uj ij     Thus X extracts the ith coordinate of Y via the inner product as in the formula (4.4.2), just as dui does. Hence X♭ = dui. Since the 1- i forms du , i =1,...,n form a basis for V ∗ by Theorem 2.1.1, the result follows.  Definition 4.4.5. Let V, , be an inner product space. The inverse of the isomorphism ♭ ofh Definitioni 4.4.3 is denoted  ♯ ♯: V ∗ V, ω ω , → 7→ for every 1-form ω. Corollary 4.4.6. In an inner product space, if the partial deriva- ∂ tives ∂ui form an orthonormal basis then ∂ i, (dui)♯ = . ∀ ∂ui Let M be a Riemannian manifold, i.e., a differentiable manifold equipped with a metric (first fundamental form), i.e., a bilinear form at each point of M; see Section 3.1. Definition 4.4.7 (Musical isomorphisms). Let M be a Riemann- ian manifold. We use the first fundamental form on Tp to define iso- morphisms ♭: T T ∗ and ♯: T ∗ T at every point p M. p → p p → p ∈ Then Corollary 4.4.6 can be formulated as follows. Recall that relative to an orthonormal basis ∂ , the gradient f of f is given ∂ui ∇ by the n-tuple ( ∂f ) where i =1,...,n. ∂ui  4.5. EXTERIOR PRODUCT AND ALGEBRA, ALTERNATING PROPERTY 53

Corollary 4.4.8. Let M be a Riemannian manifold. The exterior derivative of a function f at p M is related to the gradient f of f at p M as follows: ∈ ∇ ∈ f = ♯(df), df = ♭( f). (4.4.3) ∇ ∇ at every point p of the surface M. Proof. The proof is immediate by checking for each element in an orthonormal basis.  Remark 4.4.9. Formula (4.4.3) is basis-independent (even though the proof is carried out in an orthonormal basis), since both the gra- dient of f and df are defined in a way independent of the basis; see Remark 4.4.2. See also the note.1

4.5. Exterior product and algebra, alternating property The is sometimes referred to as the Grassman al- gebra.2 Generalizing the notion of differential 1-form to differential k-forms on a differentiable manifold M requires certain linear-algebraic prelim- inaries concerning the exterior algebra.

1Plan for building de Rham cohomology. This material is optional. The con- struction of de Rham cohomology of a differentiable manifold M involves several stages. It may be helpful to keep these stages in mind as we build up the relevant mathematical machinery step-by-step. We start with linear algebra and end with linear algebra. (1) Linear-algebraic stage: from a vector space V to its exterior alge- k bra V = k V . (2) Topological⊕ stage: from a smooth manifold M to its tangent bundle TM andV its cotangentV bundle T ∗M. (3) A vector field as a section of TM. (4) A section of T ∗M is a differential 1-form on M. k ∗ (5) The exterior bundle M = k (T M) is the bundle of exterior alge- bras parametrized by points⊕ of M. This is a finite-dimensional manifold. k V V (6) A section of (T ∗M) is a differential k-form on M. (7) The space Ωk(M) of sections of k(T ∗M) is an infinite-dimensional vec- V tor space: we are back to linear algebra. (8) The exterior derivative d turns ΩV∗(M) into a differential graded algebra. (9) De Rham cohomology groups are defined in terms of the differential graded algebra (Ω∗(M),d).

2See https://mathoverflow.net/questions/22247/ geometrical-meaning-of-grassmann-algebra for a motivating discussion. 54 4. DIFFERENTIAL FORMS, EXTERIOR DERIVATIVE AND ALGEBRA

The exterior product, or wedge product, of vectors is an algebraic construction generalizing certain features of the cross product to higher dimensions. Remark 4.5.1 (Basis-independent). In linear algebra, the exterior product provides a basis-independent manner for describing the deter- minant and the minors of a linear transformation. The exterior algebra over a vector space V is generated by an op- eration called exterior product. It is a key ingredient in the definition of the algebra of differential forms. Let V be a vector space over a field K. The exterior algebra over V is a certain unital associative3 algebra over the field K, that includes V itself as a subspace. Such an algebra is denoted by (V ). Remark 4.5.2 (Construction)^. We will begin defining the exterior algebra in Section 4.6 and provide examples. A construction of the exterior algebra appears in Definition 5.4.1. The multiplication operation in the algebra (V ), denoted , is also known as the wedge product, or the exterior product. ∧ V Definition 4.5.3. The wedge product is an associative and bilinear operation: : (V ) (V ) (V ), (α, β) α β, ∧ × → 7→ ∧ with the essential^ feature^ that it^ is anticommutative on V , meaning that v V, v v =0. (4.5.1) ∀ ∈ ∧ Remark 4.5.4. Property (4.5.1) implies in particular u v = v u (4.5.2) ∧ − ∧ for all u,v V , and ∈ v v v =0 (4.5.3) 1 ∧ 2 ∧···∧ k whenever v ,...,v V are linearly dependent. 1 k ∈ Remark 4.5.5. Unlike associativity and bilinearity which are re- quired for all elements of the algebra (V ), the three properties (4.5.1), (4.5.2), (4.5.3) are satisfied only on the elements of the subspace V (V ). The defining property (4.5.1)V and property (4.5.3) are equiva-⊆ lent; properties (4.5.1) and (4.5.2) are equivalent unless the character- isticV of K is two. 3chok kibutz 4.7.AREASINTHEPLANE;SIGNEDAREA 55

4.6. Exterior algebra over a dim 1 vector space We assume the existence of such an algebra and derive some of its properties. A construction will be provided in Section 5.4. Theorem 4.6.1. Let V be a 1-dimensional vector space over R. Let (V ) be the exterior algebra over V . Then (1) the algebra is 2-dimensional; V (2) the algebra is isomorphic to the subalgebra of the algebra of 2 2 matrices consisting of uppertriangular4 matrices with a pair× of identical eigenvalues: x y (V ) : x,y R ≃ 0 x ∈ ^    Proof. As a vector space, the algebra is the direct sum (V ) = R 1+ R n, where 1 denotes the identity matrix and n is the matrix · · V 0 1 n = . 0 0   The matrix n is nilpotent: n2 = 0, as required by (4.5.1). Here the vector space V is identified with the line R n (V ).  ⊆ Definition 4.6.2. One uses the notation 0(VV )= R 1 and 1(V )= V = R n, so that (V )= 0(V )+ 1(V ). V V 4.7.V AreasV in the plane;V signed area The purpose of this section is to motivate the skewness of the ex- terior product on vectors in V based on geometric considerations of areas. ∧ Example 4.7.1. The area of a parallelogram spanned by vectors v v1 w1 and w is the absolute value of the determinant of the matrix v2 w2 of the coordinates of the vectors.   In more detail, the Cartesian plane R2 is a vector space equipped t with a basis. The basis consists of a pair of unit vectors e1 = (1, 0) t and e2 = (0, 1) . Suppose that 1 2 1 2 v = v e1 + v e2, w = w e1 + w e2 are a pair of vectors in R2, written in components. There is a unique parallelogram having v and w as two of its sides. The area of this

4meshulashit 56 4. DIFFERENTIAL FORMS, EXTERIOR DERIVATIVE AND ALGEBRA parallelogram is given by the standard determinant formula: A = det v w 1 2 2 1 = v w v w . | − | Consider now the exterior product of v and w and exploit the properties stipulated above: v w =(v1e + v2e ) (w1e + w2e ) ∧ 1 2 ∧ 1 2 = v1w1e e + v1w2e e + v2w1e e + v2w2e e 1 ∧ 1 1 ∧ 2 2 ∧ 1 2 ∧ 2 =(v1w2 v2w1)e e , − 1 ∧ 2 where the first step uses the distributive law for the wedge product, and the last uses the fact that the wedge product is alternating. Remark 4.7.2 (Signed area). The coefficient in this last expression is precisely the determinant of the matrix [v w]. The fact that this may be positive or negative has the intuitive meaning that v and w may be oriented in a counterclockwise or clockwise sense as the vertices of the parallelogram they define. Such an area is called the signed area of the parallelogram: the absolute value of the signed area is the ordinary area, and the sign determines its orientation.

4.8. Algebraic characterisation of signed area If A(v,w) denotes the signed area of the parallelogram spanned by the pair of vectors v and w, then A must have the following properties: (1) A(av,bw) = abA(v,w) for any real numbers a and b, since rescaling either of the sides rescales the area by the same amount (and reversing the direction of one of the sides reverses the orientation of the parallelogram). (2) A(v,v) = 0, since the area of the degenerate parallelogram determined by v,v (i.e., a line segment) is zero. (3) A(w,v) = A(v,w), since interchanging the roles of v and w reverses the− orientation of the parallelogram. (4) A(v + aw,w) = A(v,w), since adding a multiple of w to v affects neither the base nor the height of the parallelogram and consequently preserves its area. (5) A(e1,e2) = 1, since the area of the unit square is one. With the exception of the last property, the wedge product satisfies the same formal properties as the area. In a certain sense, the wedge product generalizes the final property by allowing the area of a parallel- ogram to be compared to that of any “standard” chosen parallelogram. 4.9. VECTOR PRODUCT, TRIPLE PRODUCT, AND WEDGE PRODUCT 57

Remark 4.8.1. The exterior product in two-dimensions is a basis- independent formulation of area. This axiomatization of areas is due to Leopold Kronecker and Karl Weierstrass.

4.9. Vector product, triple product, and wedge product For a 3-dimensional vector space V over R, the product in the exterior algebra (V ) is closely related to the vector product and triple product.5 V 3 Example 4.9.1. Using the standard basis e1,e2,e3 for R , the wedge product of a pair of vectors { } 1 2 3 u = u e1 + u e2 + u e3 and 1 2 3 v = v e1 + v e2 + v e3 is u v =(u1v2 u2v1)(e e )+(u1v3 u3v1)(e e )+(u2v3 u3v2)(e e ), ∧ − 1∧ 2 − 1∧ 3 − 2∧ 3 where e1 e2,e1 e3,e2 e3 is the basis for the three-dimensional { 2 ∧ ∧ ∧ } space (R3) (in notation similar to that of Definition 4.6.2). This formula is similar to the usual definition of the vector product of vectors in threeV dimensions.

1 2 3 Example 4.9.2. Bringing in a third vector w = w e1 +w e2 +w e3, the wedge product of three vectors is u v w =(u1v2w3+u2v3w1+u3v1w2 u1v3w2 u2v1w3 u3v2w1)(e e e ), ∧ ∧ − − − 1∧ 2∧ 3 3 3 where e1 e2 e3 is the basis vector for the one-dimensional space (R ). Note that∧ the∧ coefficient is the usual triple product (u v) w. × · V The vector product and triple product in three dimensions each admit both geometric and algebraic interpretations. (1) (geometric interpretation) The vector product u v can be interpreted as a vector which is perpendicular to both× u and v and whose magnitude is equal to the area of the parallelogram determined by the two vectors. (2) (algebraic interpretation) The vector product can also be in- terpreted as the vector consisting of the minors of the matrix with columns u and v.

5machpela meurevet 58 4. DIFFERENTIAL FORMS, EXTERIOR DERIVATIVE AND ALGEBRA

Remark 4.9.3. The triple product of u, v, and w is geometri- cally a (signed) volume. It is also the determinant of the matrix with columns u, v, and w. The exterior product in three dimensions allows for similar inter- pretations. In fact, in the presence of a positively oriented orthonormal basis, the exterior product generalizes these notions to higher dimen- sions.

4.10. Anticommutativity of the wedge product Assume v v = 0 for all v V . ∧ ∈ Theorem 4.10.1. The wedge product is anticommutative on ele- ments of V (V ). ⊆ Proof. LetV x,y V . Then 0=(x + y) (x + y) = x x + x y + y x + y y = x∈ y + y x. Hence x y∧= y x. ∧ ∧ ∧ ∧ ∧ ∧ ∧ − ∧ Theorem 4.10.2. If x1,x2,...,xk are elements of V , and σ is a permutation of the integers (1, 2,...,k), then x x x = σ(1) ∧ σ(2) ∧···∧ σ(k) sgn(σ)x1 x2 xk, where sgn(σ) is the sign (plus or minus) of a permutation∧ σ∧···∧S . ∈ k A proof can be found in greater generality in Bourbaki (1989).

4.11. The k-exterior power; simple/decomposable multivs Definition 4.11.1. The k-th exterior power of V , denoted k(V ), is the vector subspace of (V ) spanned by elements of the form x1 V ∧ x2 xk, xi V,i =1, 2,...,k. ∧···∧ ∈ V Definition 4.11.2. If α k(V ), then α issaidtobea k-multivector. If, furthermore, α can be expressed∈ as a wedge product of k elements of V , then α is said to be decomposable,V or simple.

Although decomposable multivectors span k(V ), not every ele- ment of k(V ) is decomposable. V ExampleV 4.11.3. In R4, the following 2-multivector is not decom- posable: α = e e + e e . (4.11.1) 1 ∧ 2 3 ∧ 4 Here α is in fact a symplectic form, since α α = 0. ∧ 6 4.12. BASIS AND DIMENSION OF EXTERIOR ALGEBRA 59

4.12. Basis and dimension of exterior algebra

Theorem 4.12.1. If the dimension of V is n and e1,...,en is a basis of V , then the set e e e : 1 i

Proof. GivenV any wedge product of the form v1 vk, every ∧···∧ vector vj can be written as a linear combination of the basis vectors ei. Using the bilinearity of the wedge product, the expression v1 vk can be expanded to a linear combination of wedge products of s∧···∧uch basis vectors. Any wedge product in which the same basis vector appears more than once is zero. Any wedge product in which the basis vectors do not appear in the proper order can be reordered, changing the sign whenever two basis vectors change places.  In general, the resulting coefficients of the basis k-vectors can be computed as the minors of the matrix that describes the vectors vj in terms of the basis ei. By counting the basis elements, we obtain the following corollary. Corollary 4.12.2. The dimension of k(V ) is the binomial coef- n n! k ficient k = k!(n k)! . In particular, (V )=0 for k>n. − V Theorem  4.12.3. The dimensionV of (V ) equals 2n. Proof. Any element of the exteriorV algebra can be written as a sum of multivectors. Hence, as a vector space the exterior algebra is a direct sum (V )= 0(V )+ 1(V )+ 2(V )+ + n(V ) (where by 0 1 ··· convention (V ) = K and (V ) = V ), and therefore its dimension is equal to theV sum ofV the binomialV coefficientsV (as k runsV from 1 to n), which is 2n.V V 

CHAPTER 5

Exterior differential complex

5.1. Rank of a multivector In Section 4.11 we defined the k-th exterior power k(V ) of a vec- tor space V . Let α k(V ) be a k-multivector. Thus, α is a linear combination of a finite∈ number, say s, of decomposableV (simple) mul- tivectors: V

α = α(1) + α(2) + + α(s) (5.1.1) ··· where each α(i) is decomposable into a wedge product of the following form: α(i) = α(i) α(i), i =1, 2,...,s 1 ∧···∧ k where α(i) 1(V )= V for all i =1,...,s and j =1,...,k. j ∈ DefinitionV 5.1.1. The rank rank(α) of the multivector α is the minimal number s of decomposable multivectors in all possible expan- sions (5.1.1) of α.

5.2. Rank of 2-multivectors; matrix of coefficients

Let k = 2. Let (ei) be a basis for V . Then each multivector α 2 ∈ (V ) can be expressed uniquely as V α = a e e . (5.2.1) ij i ∧ j i

Proof. This follows from the fact that every antisymmetric real matrix can be orthogonally diagonalized into 2 by 2 blocks; see Theo- rem 6.9.1.  2 4 Example 5.2.3. The symplectic form α = e1 e2 + e3 e4 R of (4.11.1) has rank 2, whereas its matrix of coefficients∧ ∧ ∈ V 0 1 0 0 10 0 0 A = α −0 0 0 1  0 0 1 0  −  1  is of rank 4. Thus rank α = 2 rank Aα. For further details see Example 6.7.1. We mention the following theorem for general culture. Theorem 5.2.4. Over a field of characteristic 0, a 2-multivector α has rank p if and only if the p-fold product satisfies α α =0 ∧···∧ 6 p and | {z } α α =0. ∧···∧ p+1 5.3. Construction| {z of the} tensor algebra In Chapter 4, the exterior algebra was introduced axiomatically, i.e., characterized via its properties. In Section 5.4, we will provide a construction of the exterior algebra via tensor products. The tensor product V W of two vector spaces V and W over the field R is defined as follows.⊗ Consider the set of ordered pairs (v,w) in the Cartesian product V W . × Remark 5.3.1. For the purposes of the construction, regard the Cartesian product as a set rather than a vector space. Definition 5.3.2. A typical element of V W viewed as a set will × be denoted e(v,w) for the purposes of the construction that follows. Definition 5.3.3. The free vector space F = F (V W ) on V W is the vector space in which the elements e of V ×W are a basis.× (v,w) × Thus, F (V W ) is the collection of all finite linear combinations and can be written× as follows:

F (V W )= α e : α R,v V,w W . × i (vi,wi) i ∈ i ∈ i ∈ ( i ) X 5.3. CONSTRUCTION OF THE TENSOR ALGEBRA 63

The terms e(v,w) are by definition linearly independent in F (V W ) for distinct pairs (v,w) V W . The tensor product arises by× defining the following equivalence∈ × relations on the free vector space F (V W ): × e e + e (v1+v2,w) ∼ (v1,w) (v2,w) e(v,w1+w2) e(v,w1) + e(v,w2) ∼ (5.3.1) ce e (v,w) ∼ (cv,w) ce e (v,w) ∼ (v,cw) where v ,v V , w ,w W , and c K. 1 2 ∈ 1 2 ∈ ∈ Definition 5.3.4. Let S F (V W ) be the vector subspace generated by the four equivalence⊆ relations× (5.3.1); in other words, by all the differences

e(v1+v2,w) (e(v1,w) + e(v2,w)), e − (e + e ), (v,w1+w2) − (v,w1) (v,w2) ce(v,w) e(cv,w), ce − e . (v,w) − (v,cw) The equivalence relation among elements α, β F (V W ) is defined by ∼ ∈ × α β if and only if α β S. (5.3.2) ∼ − ∈ Definition 5.3.5. The tensor product of vector spaces V and W can be described in the following two equivalent ways: (1) the quotient space V W = F (V W )/S; (2) V W = α e ⊗ : α R×,v V,w W . ⊗ i i (vi,wi) i ∈ i ∈ i ∈ Here in (2), the expressionP i αie(vi,wi) denotes the equivalence class of the finite sum αie(v ,w ) relative to the equivalence relation (5.3.2). i i iP  Definition P5.3.6. The equivalence class of the element e(v,w) in V W will be denoted v w. ⊗ ⊗ Theorem 5.3.7. The dimensions of V and W multiply under ten- sor product: dim(V W ) = dim V dim W. ⊗ Proof. Let (e1,...,en) be a basis for V . Let(f1,...,fm) be a basis for W . Then the mn elements ei fj, where i =1,...,n, j =1,...,m form a basis for the space V W⊗.  ⊗ Remark 5.3.8. While bases are useful in computing dimensions, the advantage of the construction via the free vector space lies in its independence of the choice of basis. 64 5.EXTERIORDIFFERENTIALCOMPLEX

k ℓ (k+ℓ) Definition 5.3.9. The product V ⊗ V ⊗ V ⊗ is defined on × → representative vectors by sending the pair (v1 ... vk, v1 ... vℓ) to (v ... v v ... v ). ⊗ ⊗ ⊗ ⊗ 1 ⊗ ⊗ k ⊗ 1 ⊗ ⊗ ℓ 5.4. Construction of the exterior algebra We will now define the exterior powers of a vector space V in terms of the tensor products of Section 5.3. Definition 5.4.1. The exterior product V V of V by itself is the vector space obtained as the quotient of V V ∧by the vector subspace generated by all the sums v w + w v where⊗ v,w V . ⊗ ⊗ ∈ Definition 5.4.2. The image of v w under the surjective homo- morphism V V V V is denoted ⊗v w. ⊗ → ∧ ∧ Given a basis (ei) for V , the products ei ej, where i < j, form a 2 ∧ basis for the space V V = (V ). ∧ Definition 5.4.3. The thirdV exterior power 3 V is defined as the quotient of V V V by the subspace generated by differences of the form ⊗ ⊗ V u u u sgn(σ)u u u (5.4.1) 1 ⊗ 2 ⊗ 3 − σ(1) ⊗ σ(2) ⊗ σ(3) for all σ S . ∈ 3 The higher exterior powers k(V ) are defined similarly as quotients of V V ... V (k times) by imposing the anticommutation property as in⊗ (5.4.1).⊗ ⊗ V Definition 5.4.4. The wedge product : k V ℓ V k+ℓ V is induced from the tensor product kV ℓ∧V k+×ℓV by quotienting→ as above (choose representing k-fold⊗ and×⊗ℓ-fold→⊗V tensors,V multiplyV them in the tensor algebra, and take the equivalence class in k+ℓ).

5.5. Exterior bundle and exterior derivativeV Let p M be a point in a differentiable manifold M of dimension n. ∈

Definition 5.5.1. Associating to every cotangent space Tp∗ its ex- terior algebra (Tp∗), we obtain the exterior bundle

V M = (T ∗M), with a canonical projection^πM : ^M M with typical fiber (Tp∗) of dimension 2n. → V V 5.5. EXTERIOR BUNDLE AND EXTERIOR DERIVATIVE 65

As for the tangent and cotangent bundles, one can exhibit local charts and transition functions to demonstrate local triviality of the exterior bundle. k Similarly putting together the k-exterior powers Tp∗, we obtain the subbundle k M = k(T M) of the bundle (M). ∗ V DefinitionV5.5.2. AVdifferential k-form on aV manifold M is a sec- tion of the exterior bundle k M of k-multivectors built from elements k of T ∗M. The space of such sections is denoted Ω (M). V Recall that we have an exterior derivative d defined on smooth 1 n functions f C∞(M) locally in a coordinate chart (A, (u ,...,u )) by the formula∈ ∂f df = dui, (5.5.1) ∂ui with Einstein summation convention; see (4.2.2). We next define the exterior derivative on 1-forms. Definition 5.5.3. The exterior derivative, or differential, d: Ω1(M) Ω2(M) → is defined on a 1-form f(u1,...,un)du by setting d(fdu)= df du. (5.5.2) ∧ Remark 5.5.4. It is immediate from the definition that for each constant form dui one has d(dui)=0. Remark 5.5.5 (Issue of signs). If one views the form fdu as the product (du)f, one would need to introduce a sign in order to be com- patible with formula (5.5.2): d (du)f = du df − ∧ by the basic property of 1-forms: df du = du df. ∧ − ∧ A similar formula defines the exterior derivative for an arbitrary k- form. Thus, for k = 2 we have the following. Definition 5.5.6. For a 2-form fdu dv Ω2(M), one sets ∧ ∈ d(fdu dv)= df du dv Ω3(M). ∧ ∧ ∧ ∈ The general form of the Leibniz rule for differential forms appears in (5.6.1) in Section 5.6. 66 5.EXTERIORDIFFERENTIALCOMPLEX

5.6. Exterior differential complex We will define the de Rham cohomology of a differentiable mani- fold M of dimension n by means of the exterior differential complex. Theorem 5.6.1. The exterior derivative d gives rise to an exterior differential complex

d 1 d 2 d d n 0 C∞(M) Ω (M) Ω (M) ... Ω (M) 0, → → → → → → with the following properties: (1) one has d2 = d d =0 at each stage of the complex; (2) one has the following◦ form of the Leibniz superrule for d ap- plied to differential forms: d(α β)= dα β +( 1)p α dβ, (5.6.1) ∧ ∧ − ∧ where p is the degree of α. Remark 5.6.2. If the form α has even degree then the sign disap- pears and we obtain the naive Leibniz rule d(α β)= dα β + α dβ, regardless of the degree of β. ∧ ∧ ∧ Proof. Let us check the property d2 = 0 listed in item (1) as applied to a typical 1-form ω = fdu. Thus, ∂f dω = df du = dui du. ∧ ∂ui ∧ Then we exploit the equality of mixed second partial derivatives to write ∂f d(dω)= d dui du ∂ui ∧   ∂2f = duj dui du ∂ui∂uj ∧ ∧ ∂2f = duj dui du + dui duj du ∂ui∂uj ∧ ∧ ∧ ∧ i

We would like to define an antisymmetric multilinear map f = fy as in (5.7.1), associated with the multivector y. We define such a map by setting 1 k fy(x1,...,xk)= sign(σ) y (xσ(1)) ...y (xσ(k)). (5.7.2) σ S X∈ k In other words, i fy(x) = det y (xj) is a k k-determinant. The antisymmetric  property follows from a similar× property of the determinant under permutations of columns. 1 Note that we do not multiply by k! .  Corollary 5.7.5. The dimension of the space of antisymmetric k R n multilinear maps from V to is the binomial coefficient k .

Recall that V ∗ is the dual of V . We have a similar property  for exterior powers. k k Theorem 5.7.6. The space (V ∗) is naturally dual to (V ). k Proof. We view an elementV of (V ∗) as an antisymmetricV k- multilinear function φ. The duality is given on simple elements v1 k ∧ ... v of (V ) by the pairing V ∧ k φ, v ... v = φ(v ,...,v ), V h 1 ∧ ∧ ki 1 k and extended by linearity to all of k V . One readily checks the in- dependence of the resulting value of the particular representation of the simple multivector as a productV (see Proposition 5.8.2 for more details).  The case k = 2 will be examined in detail in the next section.

5.8. Case of 2-forms 1 2 Let y ,y V ∗ be 1-forms on V . Consider the decomposable (sim- ple) exterior 2-form∈

1 2 2 ω = y y (V ∗). ∧ ∈ Then ω defines an antisymmetric bilinear^ map f : V 2 R ω → defined as follows. Let u,v V . Following (5.7.2), we set ∈ y1(u) y1(v) f (u,v)= y1(u)y2(v) y1(v)y2(u) = det ω − y2(u) y2(v)   5.8. CASE OF 2-FORMS 69

i i where y (u) is the evaluation of the covector y V ∗ on the vector u V . ∈ ∈ Example 5.8.1. In the (x,y)-plane V = R2, consider the 2-form 2 ω = dx dy (V ∗). ∧ ∈ It defines a bilinear function fω : V V^ R whose geometric meaning is the signed area (see Section 4.7) of× the→ parallelogram spanned by the pair of vectors u,v.

Proposition 5.8.2. The value of fω on the pair (u,v) depends only on its image ξ = u v in 2(V ). ∧ Proof. The proof is immediateV from the interpretation of fω(u,v) as a generalized signed area of the parallelogram spanned by u and v. Namely, the antisymmetric form fω is proportional to the standard area 2 form ω = dx dy since (V ∗) is 1-dimensional, and dx dy calculates ∧ ∧ the signed area Au,v of the parallelogram spanned by u and v, where u v = A e1 e2 V2(R2).  ∧ u,v ∧ ∈ Remark 5.8.3. WeV will sometimes write ω(ξ) in place of fω(u,v), where ξ = u v 2(R2). Thus by definition, ∧ ∈ V ω(ξ)= fω(u,v).

CHAPTER 6

Norms on forms, Wirtinger inequality

6.1. Norm on 1-forms One major objective of our course is the proof of Gromov’s systolic inequality for complex projective space. To this end, we will need to study norms, determined by a Riemannian metric on a manifold M, on the de Rham cohomology groups of M. We start with a discussion of norms and their duals. Let V be a vector space. Given a norm (not necessarily of k k Euclidean type) on V , there is a natural norm on the dual space V ∗, defined as follows. We will denote the new norm ∗ for the purposes of this section. k k

Definition 6.1.1. The dual norm ∗ on V ∗ is defined for y V ∗ k k ∈ by setting y ∗ = sup y(x): x V, x 1 . k k ∈ k k≤ In other words, we calculate the dual norm of the covector y by maximizing its value over vectors x V of norm at most 1. ∈ Remark 6.1.2. The inequality x 1 can be replaced by equality in this definition: k k≤

y ∗ = sup y(x): x V, x =1 , k k { ∈ k k } i.e., the norm of y V ∗ can be calculated over the unit vectors x V . ∈ ∈ ∂ ∂ Example 6.1.3. If ∂x , ∂y is an orthonormal basis for the Eu- clidean norm in the plane V =R2, then the dual basis

(dx,dy)

∂ ∂ for V ∗ is an orthonormal basis. Indeed, dx( ∂x )=1, dx( ∂y )=0, and it follows that dx ∗ = 1 (we use a pair of single bars because in this case the norm is Euclidean).| |

Additional examples will appear in the next section. 71 72 6. NORMS ON FORMS, WIRTINGER INEQUALITY

6.2. Polar coordinates and dual norms The Euclidean metric in the plane V defines a norm on the tangent plane by the usual identification of a space and its tangent space at a point. Remark 6.2.1. When speaking of a norm in the cotangent plane we will always mean the norm dual to that on the tangent plane (in the sense of Section 6.1). Lemma 6.2.2. In polar coordinates (r, θ) in V 0 , we have \ { } xdy ydx ∗ = r, | − | and therefore x y dy dx ∗ =1. r − r

Proof. The proof is immediate from the fact that dx and dy are orthonormal. Here we use single bars because in this case the norms are Euclidean.  Note that θ is undefined at the origin. The pair (dr, dθ) is a basis for the cotangent plane at every point of V 0 . However, the basis is not orthonormal. We will specify the norm\ of { dθ} in Proposition 6.2.4. Lemma 6.2.3. We have an identity r2dθ = xdy ydx. − y Proof. We have tan θ = x . Equivalently x sin θ = y cos θ. Differ- entiating with respect to x, we obtain dθ dy dθ sin θ + x cos θ = cos θ y sin θ . dx dx − dx Multiplying by rdx we obtain ydx + x2dθ = xdy y2dθ. − Thus xdy ydx =(x2 + y2)dθ = r2dθ.  − Proposition 6.2.4. In the (r, θ) coordinates on V 0 , the cotan- gent plane at each point admits an orthonormal basis given\{ } by the pair of 1-forms (dr, rdθ). Proof. By Lemma 6.2.2 and lemma 6.2.3, we obtain 1 dθ ∗ = . | | r x y Thus we obtain a unit-norm 1-form rdθ = r dy r dx, proving the proposition. −  6.4. SHORTEST NONZERO VECTOR IN A LATTICE 73

6.3. Dual bases and dual lattices in a Euclidean space We now consider dual bases in Euclidean space. The discussion is similar to the situation with a pair of dual vector spaces treated in Section 2.8. Thus, let V = Rn be a Euclidean vector space equipped with an inner product denoted , . h i Definition 6.3.1. Given a basis (xi) for V , there exists a unique basis (yj) for V satisfying x ,y = δ , h i ji ij where δij is the Kronecker delta function. This basis will also be re- ferred to as the dual basis.

n n Definition 6.3.2. Given a lattice L R , the dual lattice L∗ R is the lattice ⊆ ⊆ n L∗ = y R : x L, x,y Z . { ∈ ∀ ∈ h i∈ } Thus the pairing between a point in a lattice and a point in its dual is by definition always an integer.

n Theorem 6.3.3. Consider a Z-basis (xi) for a lattice L R . n ⊆ Consider the basis (yj) dual to the basis (xi) in R . Then (yj) is a Z- basis for the dual lattice L∗.

6.4. Shortest nonzero vector in a lattice Definition 6.4.1. Given a lattice L Rn with its Euclidean norm , denote by λ (L) the least length of⊆ a nonzero vector in L: || 1 λ (L)= min v : v L 0 . 1 | | ∈ \ { } We now consider the case of dimension 1.

Proposition 6.4.2. Let L R be a lattice and L∗ R the lattice dual to L. Then ⊆ ⊆ 1 λ1(L∗)= . (6.4.1) λ1(L) Proof. Given a lattice L R, denote by α > 0 its generator, so that L = Z α. For an element y⊆ R to pair integrally with α, it must 1 ∈ be a multiple of β = α . Thus

L∗ = Z β

1 and λ1(L∗)= β = α .  74 6. NORMS ON FORMS, WIRTINGER INEQUALITY

Remark 6.4.3. Note that the relation (6.4.1) does not hold in gen- eral for dimension greater than 1. Already in dimension 2, the prod- uct λ1(L∗)λ1(L) can be greater than 1, as illustrated by the following result.

Example 6.4.4. For the Eisenstein lattice LE C spanned by the cube roots of unity, we obtain λ (L )λ (L )= 2 ⊆. 1 E∗ 1 E √3 Some optional related material appears in Section 6.11.2.

6.5. Euclidean norm on k-multivectors and k-forms We will now study norms on the space of multivectors. There are two distinct natural norms on k-multivectors: the Eucldean norm and the comass norm. Definition 6.5.1. The Euclidean norm on the space k(Rn) of k-multivectors is defined by declaring the basis| | V e e e : 1 i

Example 6.5.2. Each simple k-form ei1 ... eik has unit Euclidean norm: e ... e = 1, and they are all∧ mutually∧ perpendicular. | i1 ∧ ∧ ik | Example 6.5.3. For the symplectic form α = e1 e2 +...+e2n 1 e on V = R2n, we have α = √n. ∧ − ∧ 2n | | In the next section, we will define a different norm on k which is not Euclidean in general. V 6.6. Comass norm Let V be an inner product space, for instance Rn. Recall that we have the following result. k Lemma 6.6.1. A k-multivector y V ∗ can be viewed as a k- ∈k linear antisymmetric function fy on V via the formula V 1 k fy(x1,...,xk)= sgn(σ) y (xσ(1)) ...y (xσ(k)) σ S X∈ k as in (5.7.2). We will use this identification without designating a special symbol for the antisymmetric function. The comass norm is defined as follows. Recall that an exterior form is called simple or decomposable if it can be expressed as a wedge product of 1-forms; see Section 4.11. 6.7. SYMPLECTIC FORM FROM A COMPLEX VIEWPOINT 75

Definition 6.6.2. The comass norm of a k-linear function is its maximal value on a k-tuple of unit vectors in V .

k In formulas, the comass norm ω of a k-linear form ω (V ∗) is k k ∈ V ω = max ω(e ,...,e ): e V, e =1 for 1 i k . (6.6.1) k k 1 k i ∈ | i| ≤ ≤ Here e denotes the Euclidean norm of the vector e V . | | ∈ Example 6.6.3. For the symplectic form α = e1 e2 + e3 e4 4 ∧ ∧ on V = (R )∗, we have α = √2, while for comass we have α = 1 (and hence its comass norm| | is smaller than its Euclidean normk: kα < α ). This will be proved in the context of Wirtinger’s inequality;k k see |Section| 6.10. Lemma 6.6.4. The comass norm for a simple k-form coincides with the natural Euclidean norm on k-forms. Proof. A k-tuple of 1-forms span a k-dimensional subspace P V . The k-tuple can be replaced by an orthogonal k-tuple forming⊆ a basis for P . This can be done by means of a volume-preserving transformation, by applying the Gram-Schmidt process. 

k Lemma 6.6.5. Every form ω (V ∗) satisfies the inequality ω ω . If ω is simple then equality∈ is attained. k k≤ | | V k k Proof. The spaces (V ∗) and (V ) are dual by Theorem 5.7.4. Therefore the Euclidean norm ω of ω is defined by formula V | | V ω = max ω(ξ): ξ k(V ), ξ =1 (6.6.2) | | ∈ | | similar to the formula (6.6.1)n for comass,^ with the differenco e that the maximum in (6.6.2) is taken over all k-forms in k(V ) and not merely the simple (decomposable) ones. This proves the inequality.  V 6.7. Symplectic form from a complex viewpoint Example 6.7.1. Consider the 2-form dx dy. Its associated anti- symmetric bilinear function on R2 is represented∧ by the antisymmetric 0 1 matrix already discussed in Example 5.2.3. 1 0 −  This motivates the following definitions. Consider the ν-dimensional ν complex vector space C . Denote by Zj the j-th coordinate Z : Cν C, j → 76 6. NORMS ON FORMS, WIRTINGER INEQUALITY with the usual decomposition Zj = Xj +iYj into the real and imaginary parts. The complex conjugate Z¯ is by definition Z¯ = X iY , for j j j − j all j = 1,...,ν. Then the exterior product Zj Z¯j can be expressed as follows: ∧ 2 Z Z¯ =(X + iY ) (X iY )= 2iX Y = X Y . j ∧ j j j ∧ j − j − j ∧ j i j ∧ j Thus i Z Z¯ = X Y . (6.7.1) 2 j ∧ j j ∧ j Formula (6.7.1) implies the following.

Proposition 6.7.2 (Symplectic form). Let Z1,...,Zν be the coor- dinate functions in Cν. Then the standard symplectic 2-form (4.11.1), 2 ν denoted α ((C )∗), is given by F S ∈ V ν α = i Z Z¯ . (6.7.2) F S 2 j ∧ j j=1 X Remark 6.7.3. A more traditional way of writing the form is in terms of the real basis, by the expression ν α = X Y = X Y + ... + X Y . F S j ∧ j 1 ∧ 1 ν ∧ ν j=1 X The expression (6.7.2), emphasizing the complex structure, is useful for future applications, including the proof of Wirtinger’s inequality.

Example 6.7.4. The corresponding coefficient matrix Aα is a block 0 1 diagonal matrix with ν diagonal blocks of the form corre- 1 0 −  sponding to each of the complex coordinates Zj.

6.8. Hermitian product Let V be a ν-dimensional vector space over C. Let H = H(v,w) be a Hermitian product on V . Example 6.8.1. The standard Hermitian product on V = Cν is given by ν vjwj (6.8.1) j=1 X where v =(vj) and w =(wj). 6.8.HERMITIANPRODUCT 77

Remark 6.8.2. Here we adopt the convention that a Hermitian product H is complex linear in the second variable. There are varying conventions in textbooks regarding this issue. We follow Federer’s book [Fe69] which uses the convention (6.8.1). Lemma 6.8.3. We have H(v,w)= H(w,v). (6.8.2) Definition 6.8.4. We consider the real part v w = Re(H(v,w)) · of the Hermitian inner product, where v,w are viewed as vectors in R2n. Thus, v w is a scalar product on V viewed as a 2ν-dimensional real · vector space. Consider also the imaginary part αF S = αF S(v,w) of the Hermitian product, so that H(v,w)= v w + iα (v,w). (6.8.3) · F S Remark 6.8.5. The above decomposition is similar to the decom- position

zz¯ ′ =(x iy)(x′ + iy′)=(xx′ + yy′)+ i(xy′ x′y). − − Lemma 6.8.6. The imaginary part αF S of the Hermitian product is skew-symmetric and therefore can be viewed as an element αF S 2 ∈ (V ∗), the second exterior power of V ∗.

V Proof. We have αF S(w,v) = Im(H(w,v)) = Im H(v,w) by (6.8.2). Therefore α (w,v)= Im(H(v,w)) = α (v,w).   F S − − F S Lemma 6.8.7. The comass of the standard symplectic form αF S satisfies αF S =1, where the value of the comass is attained by α(v,w) if and onlyk if kiv = w.

Proof. By the previous Lemma 6.8.6, the 2-form αF S is antisym- metric. Thus it suffices to evaluate αF S on a 2-vector ξ = v w (see Remark 5.8.3), where v and w are orthonormal. Thus v w =∧ 0. We therefore have · H(v,w)= iαF S(v,w) (6.8.4) and α (v,w)= iH(v,w)= H(iv,w) by complex conjugate-linearity F S − in the first variable. Since αF S(ξ) is real, the pairing ξ,αF S = αF S(ξ) can be evaluated as follows using (6.8.4): h i α (ξ)= α (v,w)= H(iv,w) = Re(H(iv,w))=(iv) w 1 (6.8.5) F S F S · ≤ by the Cauchy-Schwarz inequality. Equality in the Cauchy-Schwarz inequality holds if and only if one has iv = w (up to sign).  78 6. NORMS ON FORMS, WIRTINGER INEQUALITY

6.9. Orthogonal diagonalisation In this section, we recall some standard material from linear algebra. Theorem 6.9.1. Every skew-symmetric real matrix can be orthog- onally diagonalized into 2 by 2 blocks as well as a possible block whose entries are identically zero. Alternatively, the theorem can be formulated as Theorem 6.9.3 be- low. Definition 6.9.2. An endomorphism f of Rn is anti-selfadjoint if ( x,y Rn) f(x),y = x,f(y) . ∀ ∈ h i −h i The following result is well known. Theorem 6.9.3. Given an anti-selfadjoint endomorphism f of Rn, n there exists an orthogonal decomposition of R into subspaces Vj in- variant under f, where dim V 2 for all j. j ≤ Recall (Section 5.1) that the rank of a 2-form α is the least possible number of simple (decomposable) 2-forms α(j) occurring in a presenta- (j) tion α = j α . CorollaryP 6.9.4. The rank of a 2-form equals half the rank of the matrix of the antisymmetric bilinear function. Proof. A 2-form is given by a skew-symmetric matrix A, which can be thought of as an anti-selfadjoint endomorphism fA. Step 1. Applying the diagonalisation theorem to fA, we obtain an orthonormal basis e1,...,en with respect to which the endomorphism decomposes into (1) µ nonvanishing blocks of size 2 by 2, and (2) an s s block on which the endomorphism vanishes, × where 2µ + s = n. Step 2. With respect to such a basis, the 2-form will have the following form. We have an orthonormal set ω ,...,ω V ∗ and 1 2µ ∈ nonnegative numbers λ1,...,λµ, so that µ

fA = λj (ω2j 1 ω2j) . (6.9.1) − ∧ j=1 X Therefore rank(A)=2µ whereas the rank(α)= µ.  Remark 6.9.5. With respect to the new basis, the nonzero part of 0 λ the matrix will consist of 2 by 2 blocks of the form j . λ 0 − j  Rn n Corollary 6.9.6. The rank of a 2-form on V = is at most 2 . 6.10. WIRTINGER INEQUALITY 79

6.10. Wirtinger inequality We exploit the material of Section 6.9 to prove Wirtinger’s inequal- ity for 2-forms. Then we proceed to the study of Riemannian metrics on the complex projective space. Recall that the standard symplectic form on V (where V is isomor- Cν 2 2 phic to ) is denoted αF S (V ∗) where (V ∗) is the space of all 2-linear antisymmetric functions∈ on V (see Theorem 5.7.6). The V V form αF S can be thought of as the imaginary part of the standard Hermitian product as in (6.8.3) and (6.10.1). Following H. Federer [Fe69, p. 40], we prove an optimal upper bound for the comass norm , cf. Definition 6.6.2, of the exterior powers of a 2-form. k k We will use the decomposition H(v,w)= v w + iα (v,w) (6.10.1) · F S of a Hermitian inner product into the sum of a scalar product and the symplectic form. µ Definition 6.10.1. We will use the notation α ∧ = α F S F S ∧···∧ αF S (µ times). µ 2µ Thus α∧ (V ∗). We will use the pairing , between dual ∈ h i spaces (V ) and (V ∗). V TheoremV 6.10.2V (Wirtinger inequality). Let V be a ν-dimensional complex vector space and α its standard symplectic form. Let µ 1. ≥ (1) If ξ 2µ V and ξ is simple, then ∈ µ V ξ,α ∧ µ! ξ . h F S i≤ | | (2) equality holds if and only if there exist elements v ,...,v V 1 µ ∈ such that ξ = v1 (iv1) vµ (ivµ). ∧ ∧···∧ ∧ µ ∧ (3) Consequently, the comass norm satisfies αF S = µ! for each µ ν. k k ≤ We will exploit the following combinatorial result in the proof of Wirtinger’s inequality. µ µ Proposition 6.10.3. In the polynomial j=1 λjxj with com- muting variables x , the coefficient of the monomial x x x is the j P 1 2 µ product λ λ µ!. ··· 1 ··· µ Proof. Consider the product of µ parentheses

(λ1x1 + λ2x2 + . . . + λµxµ)(λ1x1 + λ2x2 + . . . + λµxµ) (λ1x1 + λ2x2 + . . . + λµxµ) . · · · 80 6. NORMS ON FORMS, WIRTINGER INEQUALITY

This product can be analyzed combinatorially as follows: we have µ possibilities of choosing the variable x1 out of the µ parenthetical ex- pressions. Out of the remaining µ 1 parenthetical expressions, we − have as many possibilities for choosing x2, etc., resulting in µ! possibil- ities.  Remark 6.10.4. In the calculation of the comass of the powers of the symplectic form, we will apply this combinatorial fact to the commuting variables xj given by the 2-forms xj = ω2j 1 ω2j. − ∧ 6.11. Proof of Wirtinger inequality ν µ On the space V = C , we wish to study the form α∧ and its evaluation on simple (decomposable) 2µ-multivectors ξ. Remark 6.11.1. The main idea is that in real dimension 2µ, ev- ery 2-form splits into a sum of at most µ orthogonal simple (decom- posable) pieces. Let be the natural Euclidean norm in 2µ V . || Step 1. We can assume that ξ = 1. To prove Wirtinger’s inequal- | | V µ ity, we need to show that the pairing satisfies the bound ξ,αF S µ! for all such ξ. The case µ = 1 was treated in Lemmah 6.8.7 viai≤ the Cauchy–Schwarz inequality. Step 2. In the general case µ 1, we proceed as follows. Consider ≥ the 2µ-dimensional subspace T = Tξ V associated with the sim- ple 2µ-vector ξ. Here T is spanned by the⊆ vectors in the decomposition of ξ as a wedge product of 1-vectors (to be chosen more specifically later). Let f : T ֒ V be the inclusion map.1 → 2 Step 3. Recall that α (V ∗) is (identified with) a bilinear F S ∈ antisymmetric function on V . Consider the restriction of αF S to the 2 V 2 subspace T V , denoted ( f)αF S (T ∗). Let us show that restriction (to⊆T ) and power-raising∧ are∈ commuting operations. V Lemma 6.11.2. The restiction of α µ to T , denoted ( 2µf)α µ, F S ∧ F S coincides with the µ-th power of the restriction of αF S to T . µ µ 2µ ∧ 2 Proof. The values of both forms ( f)αF S and (( f)α)∧ on 2µ-tuples of vectors in T coincide. ∧ ∧  Step 4. We orthogonally diagonalize the anti-symmetric bilinear function ( 2f)α on T . Namely, we decompose ( 2f)α into 2 2 ∧ F S ∧ F S × 1hachala (not hachlala) 6.11. PROOF OF WIRTINGER INEQUALITY 81 diagonal blocks corresponding to 2-dimensional subspaces of T . Thus, we can choose dual orthonormal bases:

(1) (e1,...,e2µ) of T and 1 (2) (ω1,...,ω2µ) of (T ∗), and nonnegative numbers λ 0,...,λ 0, so that V 1 ≥ µ ≥ µ 2 ( f)αF S = λj (ω2j 1 ω2j) . (6.11.1) ∧ − ∧ j=1 X By Lemma 6.8.7 we have α = 1 and therefore k F Sk λj = αF S(e2j 1,e2j) αF S =1 (6.11.2) − ≤ k k for each j =1,...,µ. By the combinatorial Proposition 6.10.3 (with xj = ω2j 1 ω2j) and Lemma 6.11.2 we obtain − ∧ µ 2µf α ∧ = µ! λ ...λ ω ω . (6.11.3) ∧ F S 1 µ 1 ∧···∧ 2µ

Step 5. The simple multivector ξ decomposes as ξ = ǫe1 e2µ with ǫ = 1. Therefore by (6.11.3), ∧···∧ ± µ ξ,α ∧ = ǫµ! λ ...λ µ! (6.11.4) h F S i 1 µ ≤ since λ 1 from (6.11.2). j ≤ Step 6. Equality occurs in (6.11.4) if and only if ǫ = 1 and λj =1 for each j. Applying the proof of Lemma 6.8.7 based on Cauchy– Schwarz inequality, we conclude that e2j = ie2j 1, for each j =1,..., 2µ, completing the proof of Wirtinger inequality. − 6.11.1. Wirtinger inequality for an arbitrary 2-form. This ma- terial is optional. Recall from Section 3.1 that the polarisation formula allows one to reconstruct a symmetric bilinear form B, from the quadratic form Q(v)= B(v,v) (if the characteristic is not 2): 1 B(v,w)= (Q(v + w) Q(v w)). (6.11.5) 4 − − This will be exploited in the proof of Proposition 6.11.3 below. There is a useful generalisation of Wirtinger’s inequality for arbitrary 2- forms. The idea of the proof is still to use the Cauchy-Schwarz inequality, after adjusting the Hermitian product to be compatible with the 2-form under consideration in a suitable sense.

Proposition 6.11.3. Given an orthonormal basis (ω1,...,ω2µ) for the 1 µ space (C )∗, and nonzero real numbers λ1,...,λµ, the form µ V α = λj (ω2j 1 ω2j) (6.11.6) − ∧ Xj=1 82 6. NORMS ON FORMS, WIRTINGER INEQUALITY has comass α = max λ . k k j | j| Proof. We can assume without loss of generality that each λj is pos- itive. This can be attained in one of two ways. One can permute the coordinates, by applying the transposition flipping ω2j 1 and ω2j, so as to change the sign of λ . Alternatively, one can replace,− say, ω by ω , j 2j − 2j which similarly changes the sign of λj. We now set 1/2 1/2 ω2′ j 1 = λj ω2j 1, ω2′ j = λj ω2j. − − we can then write α = µ ω ω . We now exploit the polarisation j=1 2′ j 1 ∧ 2′ j formula (6.11.5). − P Remark 6.11.4. The polarisation works on a real slice; one then com- plexifies to go from a real dot product to a Hermitian product.

Consider the Hermitian inner product Hα on T obtained by polarizing the quadratic form Qα defined by

2 2 1/2 2 1/2 2 Qα = ω2′ j 1 + ω2′ j = λj ω2j 1 + λj ω2j , − − j j   X     X     where the squares are understood in the sense of the symmetric product (see (3.1.1)).

Lemma 6.11.5. The Hα-norm v α of each H-unit vector v is bounded 1/2 | | above by (maxj λj) : 1/2 v α max(λj) . | | ≤ j

Proof. The Hermitian inner product Hα is set up in such a way as to satisfy Hα(v,w) = iα(v,w), or α(v,w) = iHα(v,w) for each Hα- orthogonal pair v,w. Now let ζ be a unit 2-vector− such that α = α(ζ), (6.11.7) || || Consider its 2-plane T V . In the 2-plane T , the unit disc of the scalar ⊆ product defined by Hα is an ellipse with respect to the scalar product of H. We will exploit its (perpendicular) principal axes. Let v,w be an orthonor- mal pair proportional to the principal axes. We can then write ζ = v w. ∧ The H-orthonormal pair v,w is also Hα-orthogonal (though in general not H -orthonormal), as well. Then, as in (6.8.5), we have α(ζ) = iH (ζ) = α − α Hα(iv,w) iv α w α maxj λj by Lemma 6.11.5. This proves the lemma. ≤ | | | | ≤  The lemma generalizes the the µ-th powers of a 2-form, giving a gener- alisation of Wirtinger’s inequality as stated in the proposition.  Corollary 6.11.6. Every real antisymmetric 2-form A satisfies the co- mass bound Aµ µ! A µ. (6.11.8) k k≤ k k 6.11. PROOF OF WIRTINGER INEQUALITY 83

Proof. We may assume that A = 1. By Lemma 6.11.3, we have λj 1 for each j = 1,...,ν. k k ≤ An inspection of the proof Proposition 6.10.2 reveals that the orthogonal diagonalisation argument, cf. (6.11.2), applies to an arbitrary 2-form A with comass A = 1.  k k 2 2 Rn 6.11.2. (V ) is isomorphic to the standard model 0( ). This material is optional. Let 2(Rn) = Span e e : 1 ij By (6.11.9), we have e viwje viwje . Switching (v,w) ∼ i

CHAPTER 7

Complex projective spaces; de Rham cohomology

7.1. Cell decomposition of real projective space We start with some examples of decomposing n-dimensional mani- folds as union of cells ei of dimension i =0, 1,...,n. The discussion is mostly motivational. In particular, some gen- eral topology and material on CW-complexes will be assumed without further mention. Definition 7.1.1 (Cells). The n-cell, en, n 0 is by definition diffeomorphic to an open ball in Rn (or to Rn itself).≥ Example 7.1.2. The 0-cell e0 is a point. Definition 7.1.3. The n-sphere Sn, n 0, is the bounday of the (n + 1)-cell. ≥ Example 7.1.4. The 0-dimensional sphere S0 R1 consists of two points, 1. We can therefore represent it as a disjoint⊆ union ± S0 = e0 a(e0), ∪ where a(x)= x is the antipodal map on R. − Example 7.1.5. The 1-dimensional sphere S1 R2 can be parti- tioned into a union ⊆ S1 = e0 a(e0) e1 a(e1), ∪ ∪ ∪ where e0 = (1, 0) is a point and e1 is the open upper halfcircle and a(e1) its{ antipodal} image, where a: R2 R2 is the antipodal map. → Example 7.1.6. The 2-sphere S2 can be decomposed as S2 = e0 a(e0) e1 a(e1) e2 a(e2), ∪ ∪ ∪ ∪ ∪ where e2 is the open northern hemisphere of S2. Proposition 7.1.7. The n-sphere Sn can be decomposed as Sn = e0 a(e0) e1 a(e1) ... en a(en). ∪ ∪ ∪ ∪ ∪ ∪ 85 86 7. COMPLEX PROJECTIVE SPACES; DE RHAM COHOMOLOGY

In Section 1.6 we presented a detailed construction of the real pro- jective space RPn. For our present purposes, we view RPn is the quo- tient of Sn by the antipodal map. Definition 7.1.8. The real projective space is the space RPn = [x]: x Sn , where [x] is the equivalence class of x Sn relative to the {equivalence∈ } relation defined by x x for all x ∈ Sn. ∼ ∼− ∈ By Proposition 7.1.7 we obtain the following. Proposition 7.1.9. There is a decomposition into a disjoint union RPn = e0 e1 ... en. (7.1.1) ∪ ∪ ∪ Remark 7.1.10. The projective line RP1 = e0 e1 can be identified with a circle S1. ∪ Theorem 7.1.11. The real projective space RPn admits a cell de- composition with a single cell in each of the dimensions 0, 1, 2,...,n. In particular, we have

n n 1 n RP = RP − R , (7.1.2) ⊔ where denotes disjoint union. ⊔ In the context of the de Rham cohomology and Gromov’s inequality, we will specify a similar decomposition of the complex projective space in Sections 7.2 and 7.3.

7.2. Complex projective line CP1 In projective geometry, one starts with an arbitrary field K. The familiar construction of “adding a point at infinity” then produces the projective line ∞ KP1 = K ∪ {∞} over K. Example 7.2.1. When the field K = C is that of the complex numbers, we thereby obtain the 1-point compactification C of C, namely the Riemann sphere S2: ∪{∞} CP1 = C = S2. (7.2.1) ∪ {∞} The familiar stereographic projection provides an identification of the complement of a point in S2 and the complex numbers C. 7.4. UNITARY GROUP, PROJECTIVE SPACE AS QUOTIENT BY ACTION 87

7.3. Complex projective space CPn The real projective space was already defined in Section 1.6. The complex projective space is defined similarly; see Definition 7.3.3. Example 7.3.1. Generalizing (7.2.1), we will show in Theorem 7.6.2 that one has the decomposition n n n 1 CP = C CP − ⊔ n 1 similar to (7.1.2), where CP − is thought of as the hyperplane at infinity.1 Theorem 7.3.2. Complex projective n-space CPn is a complex man- ifold of real dimension 2n. Definition 7.3.3. A point in CPn is represented by n +1 complex coordinates as (z ,z ,...,z ) Cn+1, (z ,z ,...,z ) = (0, 0,..., 0) 0 1 n ∈ 0 1 n 6 where we identify the tuples differing by an overall rescaling: λ C 0 , (z ,z ,...,z ) (λz ,λz ,...,λz ). (7.3.1) ∀ ∈ \ { } 0 1 n ∼ 0 1 n These are homogeneous coordinates in the traditional sense of pro- jective geometry. Definition 7.3.4. The homogeneous coordinates of a point in CPn are denoted [z0,z1,...,zn] (using the traditional square brackets).

7.4. Unitary group, projective space as quotient by action We can think of both RPn and CPn as quotients by the action of a compact group as follows. Proposition 7.4.1 (Real projective space as a quotient). The real projective space can be thought of as an antipodal quotient of the sphere: RPn = Sn/G, where G = 1 is the cyclic group of order 2. {± } There is a similar presentation of complex projective space, where in place of the cyclic group G, we have the unitary group U(1); see Theorem 7.5.1. 1This point of view is explained (in the case n = 2) in more detail in the course 88537; see choveret at http://u.math.biu.ac.il/~katzmik/88-537.html for the Hartshorne reference. 88 7. COMPLEX PROJECTIVE SPACES; DE RHAM COHOMOLOGY

Definition 7.4.2. The unitary group U(n) is the group of com- plex n n matrices M satisfying the relation × t MM¯ = In. Example 7.4.3. The group U(1) is the circle of complex numbers of unit absolute value, S1 C. ⊆ In the next section, we will exploit the action of U(1) on Cn+1 by scalar multiplication as in (7.3.1). 7.5. Hopf bundle We consider the unit sphere S2n+1 Cn+1. ⊆ Theorem 7.5.1. The space CPn is a quotient space of S2n+1 under the action of U(1): CPn = S2n+1/U(1). Proof. Every complex line Cv in Cn+1 intersects the unit sphere in a circle. We restrict the vectors in the equivalence relation (7.3.1) to the unit sphere, and the scalars λ in (7.3.1) to the unit circle in C. We then identify points in an orbit under the natural action of U(1) to obtain CPn.  Definition 7.5.2 (Hopf fibration). The associated continuous map S2n+1 CPn → is a bundle with fiber S1, called the Hopf fibration. Example 7.5.3. For n = 1, this construction yields the classical Hopf bundle: S2 = S3/U(1), where S2 = CP1. 7.6. Flags and cell decomposition of CPn Recall that the real projective space has a cell decomposition with a cell in each dimension; see formula (7.1.1). An analogous decomposi- tion, but with cells in even dimensions, exists for the complex projective space. 2 n+1 Definition 7.6.1. A complete flag (Vj) in V = C is a choice 3 of a nested sequence of subspaces Vj of dimensions j =0, 1, 2,...,n +1, namely: 0 = V V V V = V. { } 0 ⊆ 1 ⊆ 2 ⊆···⊆ n+1 2degel 3mekanenet 7.7. CLOSURE OF CELLS PRODUCES COMPACT SUBMANIFOLDS 89

Theorem 7.6.2. A choice of a complete flag (Vj) determines a unique decomposition of CPn into cells in each even dimension be- tween 0 and 2n: CPn = e0 e2 e2n. (7.6.1) ∪ ∪···∪ Proof. The 2j-dimensional cell e2j CPn consists of points with ⊆ homogeneous coordinates [z0,z1,...,zn] contained in the j-th set-theore- tic difference: (z0,z1,...,zn) Vj Vj 1.  ∈ \ − For the standard complete flag (see below), we will represent a point 2j of e uniquely by the tuple (z0,z1,...,zn 1, 1). − Definition 7.6.3. The standard complete flag is the nested se- quence 1 2 n+1 0 = V0 C C C , j { } ⊆ ⊆ ⊆···⊆ where Vj = C is the subspace of points whose first j coordinates are arbitrary and the remaining n +1 j coordinates vanish. − With respect to the standard complete flag, we obtain e2j = [z ,z ,...,z , 0,..., 0]: z =0 . (7.6.2) 0 1 j j 6 Proposition 7.6.4 . The collection e2j of (7.6.2) of points with j coordinates in Vj Vj 1 is diffeomorphic to C . \ − Proof. The diffeomorphism is e2j Cj → z0 zj 1 j (z0,...,zj 1,zj, 0,..., 0) ,..., − C , − 7→ zj zj ∈   proving the proposition.  7.7. Closure of cells produces compact submanifolds Recall that we have a cell decomposition CPn = e0 e2 ... e2n, where for each j = 0, 1, 2,...,n, one has in homogeneous∪ ∪ coordi-∪ nates e2j = [z ,z ,...,z , 0,..., 0]: z =0 ; see (7.6.2). 0 1 j j 6 Theorem 7.7.1. The closure e2j CP n of each cell e2j is the complex projective subspace CPj CPn⊆: ⊆ e2j = CPj. j j j 1 Proof. Recall that we have a decomposition CP = C CP − . We need to show that an arbitrary point of the projective⊔ hyper- j 1 n j plane CP − CP can be approximated by points in C . A point j 1 ⊆ of CP − can be represented by a nonzero (n + 1)-tuple

(z0,...,zj 1, 0,..., 0), (7.7.1) − 90 7. COMPLEX PROJECTIVE SPACES; DE RHAM COHOMOLOGY where zj = ... = zn = 0. Such a point can be approximated by a point in the cell e2j with non-vanishing j-th coordinate, of the form

(z0,...,zj 1,ǫ, 0,..., 0), − for arbitrarily small ǫ. As ǫ tends to zero, we obtain that every point j 1 of CP − represented by (7.7.1) is in the closure of the 2j-cell. Thus the closure is precisely

2j j j 1 j e = C CP − = CP , ∪ completing the proof.  Remark 7.7.2. The submanifolds of CPn given by CPj for j = 0, 1, 2,...,n generate all of the homology groups of CPn; see Sec- tion 9.11.

7.8. g, α, and J in a complex vector space As motivation, we consider the trigonometric relation sin(θ) = cos( π θ). (7.8.1) 2 − The corresponding relation in terms of scalar products and determi- nants is the following. Example 7.8.1. Let v,w R2. If v,w are unit vectors forming an angle θ then ∈ (1) v w = cos θ, (2) sin· θ is the signed area det[v w] of the parallelogram spanned by v,w. π The rotation by 2 implied in formula (7.8.1) can be formalized as follows. Lemma 7.8.2. For any pair v,w R2 of (not necessarily unit) vectors, the 2 2 determinant satisfies∈ × det[v w]=(iv) w, (7.8.2) · π where i denotes the rotation by 2 resulting from the usual identifica- tion R2 = C. The relation (7.8.2) is thought of as analogous to the trigonometric relation (7.8.1). We have the following reformulation. Corollary 7.8.3. In C = R2, the area form and the metric are related by the formula

αF S(v,w)= g(iv,w), (7.8.3) 7.9.RELATIONTOHERMITIANPRODUCT 91 where αF S is the standard symplectic form dx dy (in this case the area form), and g is the standard metric dx2 + dy∧ 2 in C. Remark 7.8.4. The corollary illustrates the passage from a sym- metric form, g, to an antisymmetric form, α, using multiplication of one of the vectors by i. We will now deal with a generalisation of this phenomenon to higher dimensions. Definition 7.8.5. Let V be a complex vector space. The complex structure J on V is the endomorphism J : V V → defined by multiplication by the scalar i C, and satisfying J 2 = Id . ∈ − V Example 7.8.6. For V = Cµ, we have in coordinates

J(v)=(iv1,iv2,...,ivµ). 7.9. Relation to Hermitian product In Section 6.8 we saw that a Hermitian product H(v,w) on a com- plex vector space V decomposes as a sum

H(v,w)= g(v,w)+ iαF S(v,w), Cµ R2µ where g is a scalar product in = and αF S is the symplectic form.

Remark 7.9.1. Here g is symmetric in the two variables while αF S is antisymmetric.

Moreover, we have the following relation among J, g, and αF S. Proposition 7.9.2. One has the following relation on V :

αF S(v,w)= g(Jv,w). (7.9.1) Proof. We have g(iv,w) = Re(H(iv,w)) (by definition) = Re( iH(v,w)) (skew-linearity) − = Re( i(iα (v,w))) (imaginary part of g is zero) − F S = i2α (v,w) (α is real-valued) − F S = αF S(v,w), proving the proposition.  Remark 7.9.3. Our Proposition 7.9.2 will enable us, starting with a metric on a complex manifold M, to construct a differential 2-form on M in Section 7.10. When M is the complex projective space, the 2-form is the closed symplectic form called the Fubini–Study form. 92 7. COMPLEX PROJECTIVE SPACES; DE RHAM COHOMOLOGY

7.10. Explicit formula for Fubini–Study 2-form on CP1 We will now examine in detail the case of the complex projective line CP1, i.e. the 2-sphere. We obtain the differential 2-form on CP1 from the metric by formula (7.9.1). CP1 Definition 7.10.1. The Fubini–Study 2-form αF S on is the area form of the metric. Theorem 7.10.2. The round metric of the sphere can be expressed in stereographic coordinates as dx2 + dy2 g = (7.10.1) (1 + x2 + y2)2 in the complement of a point in S2. This normalisation results in a 1 metric of a round sphere (missing a point) of radius 2 and constant Gaussian curvature K = +4. Proof. The Gaussian curvature K of the metric (7.10.1) is given 1 by K = ∆ ln f where f(x,y) = 2 2 by formula (3.4.2). We − LB 1+x +y have ln f = ln(1 + x2 + y2). ∂ 2x − Then (ln f)= 2 2 and ∂x − 1+x +y ∂2 2(1 + x2 y2) (ln f)= − . (7.10.2) ∂x2 − (1 + x2 + y2)2 Similarly, ∂2 2(1 x2 + y2) (ln f)= − . (7.10.3) ∂y2 − (1 + x2 + y2)2 4 Adding (7.10.2) and (7.10.3), we obtain ∆ ln f = 2 2 2 2 = 4 LB − f (1+x +y ) − and therefore K = +4 at every point of the coordinate chart.  Corollary 7.10.3. In an affine neighborhood in CP1 with coordi- nates (x,y), we have the following formulas for the metric g and the corresponding Fubini–Study 2-form αF S (the area form of the metric) on the complex projective line: dx2 + dy2 g = (1 + x2 + y2)2 and dx dy α = ∧ (7.10.4) F S (1 + x2 + y2)2 Proof. This follows from formula (7.9.1).  7.11. EXTERIOR DIFFERENTIAL COMPLEX REVISITED 93 CP1 Note that both tensors g and αF S are globally defined on all of . The proof is immediate from (7.8.3); see also (10.7.2). 7.11. Exterior differential complex revisited We started dealing with the exterior differential complex in Sec- tion 5.6. Recall that on an n-dimensional differentiable manifold M, we have an exterior differential complex

d0 1 d1 2 d2 dn−1 n 0 C∞(M) Ω (M) Ω (M) ... Ω (M) 0. → −→ −→ −→ −→ →(7.11.1) i i Here each Ω (M) is the space of sections of the exterior bundle (T ∗M), while the differential d : Ωi Ωi+1 V i → raising the degree by 1, is defined as follows. For f C∞(M), we define the 1-form df Ω1(M) by setting ∈ ∈ ∂f df = d f = duj. 0 ∂uj Similarly, the differential d = d : Ω1(M) Ω2(M) 1 → is defined on a 1-form fdu by d(fdu)= df du, (7.11.2) ∧ and similarly for the higher differentials. The following proposition was already dealt with in Section 5.6. Proposition 7.11.1. The sequence of maps as in (7.11.1) is a dif- ferential complex, in the sense that d d =0, or more explicitly ◦ k, dk dk 1 =0. ∀ ◦ − Proof. This was already proved in Section 5.6.  The condition d d = 0 is sometimes written as d2 = 0 where d2 is understood as the composition◦ of two consecutive differentials.4

4The following calculation already appeared in Section 5.6. We check the condition d2 = 0 at the level of Ω1(M). Given a function f C∞(M), we ∈ ∂f i first apply the differential d = d0 to obtain a differential 1-form df = ∂ui du . 1 2 ∂f i Next, we apply the differential d = d1 : Ω (M) Ω (M) to the 1-form ∂ui du . → 2 ∂f i ∂f i ∂ f j i Using (7.11.2), we obtain d i du = d i du = i j du du , with a ∂u ∂u ∂u ∂u ∧ double summation with indices i and j running  from 1 to n independently of each other. Since dui dui = 0, we exploit the equality of mixed partials to 2 2 ∂f i ∧ ∂ f j i i j write d f = d i du = i j du du + du du = 0, since exte- ∂u i

We can restate the proposition as follows. Corollary 7.11.2. We have an inclusion k Image(dk 1) Ker(dk) Ω (M). − ⊆ ⊆ This leads to the definition of the de Rham cohomology of M in Section 7.12.

7.12. De Rham cocycles and coboundaries Definition 7.12.1. The group Zk (M) = Ker(d ) Ωk(M) dR k ⊆ is called the group of de Rham k-cocycles. Definition 7.12.2. A differential form ω is called closed if dω = 0. Thus, a de Rham k-cocycle is by definition a closed differential k-form. Definition 7.12.3. The group k k BdR(M) = Image(dk 1) Ω (M) − ⊆ is called the group of de Rham k-coboundaries. Definition 7.12.4. An exact differential form is a de Rham cobound- ary.5

7.13. De Rham cohomology of M, Betti numbers Definition 7.13.1. The k-th de Rham cohomology group of M, k denoted HdR(M), is by definition the group k k k HdR(M)= ZdR(M)/BdR(M) = Ker(dk)/Image(dk 1) − (cocycles modulo the coboundaries). Remark 7.13.2. As already mentioned in Remark 5.6.4, the dif- k ferential graded associative algebra (dga) Ω(M)= kΩ (M) can yield finer invariants called Massey products when the⊕ cup products in de Rham cohomology vanish.

k Even though the object HdR(M) is traditionally referred to as a group, it is in fact a real vector space.

1 2 ∂f i 1 2 1 2 ∂f i 1 have d(fdu du )= ∂ui du du . Therefore d (fdu du )= d ∂ui du du . ∧ ∧ 2 ∧ ∧ 2 1 2 ∂ f j i i j 1 Thus, d (fdu du )= i j du du + du du u = 0 as before.  ∧ i

Definition 7.13.3. The k-th Betti number bk(M) of M is by def- k inition the dimension of the real vector space HdR(M): k bk(M) = dim HdR(M).

CHAPTER 8

De Rham cohomology and topology

8.1. De Rham cohomology in dimensions 0 and n We will present several cases where the de Rham cohomology groups can be calculated explicitly in an elementary fashion.

Proposition 8.1.1. Every connected manifold M has unit 0-th Betti number b (M)=1, i.e. H0 (M) R. 0 dR ≃ 1 Proof. Consider the segment 0 C∞(M) Ω (M) of the exte- rior differential complex. → →

Step 1. The space of the 0-coboundaries is trivial. The space of 0- cocycles consists of all functions f C∞(M) satisfing df = 0. Thus at ∈ 0 every point, all the partial derivatives of a function f ZdR(M) must vanish. By the mean value theorem, such a function∈ must be locally constant.

Step 2. Since M is connected, the function f must also be globally constant.

Step 3. We therefore obtain an identification

0 0 Z (M)= H (M) R C∞(M) dR dR ≃ ⊆ of the 0-th de Rham cohomology group with the space of constant functions on M. 

Definition 8.1.2 (Closed, orientable). A manifold M is called closed if it is compact without boundary. M is called orientable if it admits an atlas where the determinant of the Jacobian matrix of each transition function is positive everywhere.

Theorem 8.1.3. Every closed connected orientable n-dimensional manifold M satisfies b (M)=1, i.e., Hn (M) R. n dR ≃ The proof of this result is more difficult than the previous one. We will establish it in some special cases in Sections 8.2 and 8.4. 97 98 8. DE RHAM COHOMOLOGY AND TOPOLOGY

8.2. de Rham cohomology of a circle 1 1 R Theorem 8.2.1. We have an isomorphism HdR(S ) , i.e., 1 ≃ b1(S )=1. Proof. Consider the circle M = S1 = eiθ : θ R . Thus M is a closed connected 1-dimensional manifold (in fact∈ a unique such manifold up to diffeomorphism).  Step 1. The standard 1-form dθ satisfies

dθ =2π, IM where M is the path integral over the circle. It follows by Stokes theorem that it is not exact.1 We will use integration to construct the H required isomorphism. Step 2. Recall that all 1-forms on a 1-dimensional manifold are closed. Consider the space of 1-forms Ω1(M). We define a homomor- phism Φ: Ω1(M) R → as follows. Let ω Ω1(M) be a 1-form. We define a real num- ber Φ(ω) R depending∈ on ω by setting ∈ Φ(ω)= ω. IM Step 3. Since dθ is nonvanishing at every point, we can express ω in terms of the standard 1-form dθ Ω1(M) by writing ω = f(θ)dθ, where f is a suitable single-valued continuous∈ function on the circle. Thus, 2π Φ(ω)= f(θ)dθ = F (2π) F (0), − Z0 where F : [0, 2π] R is an antiderivative for f, so that dF = f(θ)dθ. Note that in general→ F (0) = F (2π). 6 1 Step 4. We now show that the 1-form ω 2π Φ(ω)dθ is exact. Consider the function − 1 g(θ)= F (θ) Φ(ω)θ where 0 θ< 2π. − 2π ≤

1The form dθ is not exact even though it appears to be the d of θ. However, θ is multiple-valued, and a single-valued branch cannot be chosen in a continuous fashion over the entire circle. 8.3. FUBINI–STUDY FORM AND THE AREA OF CP1 99

The function g satisfies

1 g(2π)= F (2π) Φ(ω)2π = F (2π) F (2π)+ F (0) = F (0) = g(0). − 2π − Hence g descends to a smooth single-valued function on the circle. Therefore by definition, we have dg B1 (M). Note that ∈ dR 1 1 dg = fdθ Φ(ω)dθ = ω Φ(ω)dθ. − 2π − 2π Therefore the kernel kerΦ consists precisely of the de Rham 1-cobounda- ries, i.e., the exact forms.

Step 5. By the isomorphism theorem from basic group theory, the 1-cohomology group of the circle is isomorphic to R, where the 1 1 equivalence class [ω] HdR(M) of a form ω Ω (M) corresponds to the real number Φ(ω).∈ ∈ 

8.3. Fubini–Study form and the area of CP1 A specific representative of a non-trivial 2-dimensional cohomology 2 CP1 class in HdR( ), called the Fubini–Study form αF S, was defined in Section 7.10 as the bilinear antisymmetric function

αF S(u,v)= g(Ju,v) where g is a round metric on CP1 and J is the complex structure. In 1 an affine neighborhood α is given by 2 2 dx dy. The form α F S (1+r ) ∧ F S is a globally defined closed 2-form on CP1 (in this dimension it can be viewed as the area form of the metric). Its cohomology class

[α ] H2 (CP1) F S ∈ dR spans H2 (CP1) R.2 dR ≃ CP1 Theorem 8.3.1. We have area( )= CP1 αF S = π. R 2 In the sequel it will be important that the 2-form αFS has unit comass (see Section 6.6) at every point. A more general formula defines such a 2-form αFS on the CPn 2 CPn projective space (see Section 10.8). Its class [αFS] spans the group HdR( ). 100 8. DE RHAM COHOMOLOGY AND TOPOLOGY

Proof. We use polar coordinates and formula (7.10.4) as follows: CP1 area( )= αF S CP1 Z 1 = dx dy (1 + r2)2 ∧ ZZC rdrdθ = (1 + r2)2 ZZ ∞ rdr =2π (1 + r2)2 Z0 ∞ du = π u2 Z1 = π, where we used the u-substitution u =1+ r2.  Remark 8.3.2. The resulting value π for the area is consistent with the fact that we are dealing with a 2-sphere of Gaussian curvature 4, 1 π radius 2 , and Riemannian diameter 2 ; cf. Theorem 7.10.2. Remark 8.3.3. In Section 10.3 we will define an integer lattice L2 (CPn) in the de Rham cohomology of CPn. The element 1 α dR π F S ∈ H2 (CP1) is a generator of the integer lattice L2 (CP1) Z. dR dR ≃ 8.4. 2-cohomology group of the torus In this section we will compute the top-dimensional de Rham co- momology group of the 2-torus. In Section 8.5 we will do the same for the 2-sphere. 2 T2 R Theorem 8.4.1. We have an isomorphism HdR( ) , i.e., T2 ≃ b2( )=1. Proof. Let T2 = R2/ Z2 be the torus obtained as the quotient by the standard integer lattice in the (x,y)-plane. Let x and y be the standard coordinates on the torus, both ranging from 0 to 1. Note that T2 is obtained from the square [0, 1] [0, 1] by means of the familiar identifications on the boundary. Let α ×= dx dy be the area form of the torus. ∧ Step 1. Recall that on a 2-dimensional manifold all 2-forms are closed. Let Ω2(T2) be the space of 2-forms on the torus. Since the vector space 2(R2) is 1-dimensional and α is nonvanishing at every point, each 2-form can be written as f(x,y)α where f is a continuous function on theV torus. 8.4. 2-COHOMOLOGY GROUP OF THE TORUS 101

Step 2. Consider the map Φ: Ω2(T2) R given by the integral → over the torus, namely Φ(fα) = T2 fα. As in the case of the circle (see Section 8.2) this map will induce the required isomorphism. R Step 3. Suppose f(x,y) is a function satisfying Φ(fα) = 0. To prove the theorem, it suffices to show that the 2-form f(x,y)α is nec- essarily exact. Then the homomorphism Φ descends to the required isomorphism. Thus we need to find functions g(x,y) and h(x,y) on the torus such that f dx dy = d(gdx + hdy). (8.4.1) ∧ In fact h can be chosen to depend on the first variable only.3 1 Step 4. For each x, we consider the average a(x) = 0 f(x,t)dt. We define g by setting y R g(x,y)= f(x,t)dt + a(x)y. (8.4.2) − Z0 Let us show that the function g is periodic in both x and y and therefore is well defined on the torus. We have g(x, 1) = a(x)+ a(x)=0= g(x, 0). − Hence the function g is periodic in the variable y. Furthermore, 1 1 a(0) = f(0,t)dt = f(1,t)dt = a(1) Z0 Z0 since f is periodic in x. It follows that g is periodic in the x-direction, as well. Step 5. We need to choose a function h appropriately so as to satisfy equation (8.4.1) which shows that fdx dy is exact. We have ∧ ∂g ∂h ∂g ∂h d(gdx + hdy)= dy dx + dx dy = + dx dy ∂y ∧ ∂x ∧ − ∂y ∂x ∧   by antisymmetry of wedge product. Thus equation (8.4.1) is equivalent to the following PDE: ∂g ∂h + = f. (8.4.3) − ∂y ∂x We apply the fundamental theorem of calculus to the formula (8.4.2), to obtain ∂g = f + a(x). Then equation (8.4.3) becomes ∂h = a(x) ∂y − ∂x and we set x h(x)= a(s)ds. Z0 Note that h depends only on x.

3This feature will be useful when we develop a similar proof for S2 in Section 8.5. 102 8. DE RHAM COHOMOLOGY AND TOPOLOGY

Step 6. To show that h descends to a continuous function on the torus, note that 1 1 1 h(1) = a(s)ds = f(s,t)dtds =0 Z0 Z0 Z0 since f has total integral zero by hypothesis (see Step 3 above). Of course h(0) = 0 also, so h is a periodic function and therefore is well defined and continuous on the torus. This establishes the existence of the required functions g and h, proving the theorem. 

8.5. 2-cohomology group of the sphere 2 In Section 8.4, we showed that b2(T ) = 1. The same holds for the 2-sphere. Theorem 8.5.1. We have H2 (S2) R, i.e., b (S2)=1. dR ≃ 2 We use spherical coordinates (θ,φ), θ [0, 2π], φ [0,π], where z = cos φ, r = sin φ, x = r cos θ, y = r sin θ. The∈ form ∈ α = sin φdθ dφ (8.5.1) ∧ 2 is an area form for S . Assume that a function f(θ,φ) satisfies S2 fα = 0, i.e., R f sin φdθdφ =0. (8.5.2) 2 ZS We will solve the equation analogous to (8.4.1), making the necessary changes to allow for the factor sin φ in (8.5.1) as follows. We are looking for functions G(θ,φ) and H(φ) on the sphere such that f dθ sin φdφ = d(Hdθ + G sin φdφ). (8.5.3) ∧ Since sin φdφ is a well-defined global 1-form (vanishing at the poles), G can be chosen to be any smooth function. The function H needs to be chosen so as to compensate for the singularity of dθ at the poles. For each φ, we set 1 θ=2π A(φ)= f(θ,φ)dθ. 2π Zθ=0 Then A(0) equals the value of f at the north pole and A(π), at south pole. Note that φ=π A(φ)sin φ dφ =0 (8.5.4) Zφ=0 by (8.5.2). We define G by setting θ G(θ,φ)= f(t,φ)dt A(φ)θ. (8.5.5) − Z0 8.5. 2-COHOMOLOGY GROUP OF THE SPHERE 103

Then (1) G is 2π-periodic in θ; (2) G(θ,φ) defines a continuous function on S2. We will choose a function H(φ) appropriately so as to satisfy equa- tion (8.5.3). We have dH ∂G d(Hdθ + G sin φdφ)= dφ dθ + dθ sin φdφ dφ ∧ ∂θ ∧ dH ∂G = + sin φ dθ dφ − dφ ∂θ ∧   by the antisymmetry of wedge product. Thus equation (8.5.3) is equiv- alent to the following differential equation: dH ∂G + sin φ = f sin φ. (8.5.6) − dφ ∂θ We apply the fundamental theorem of calculus to (8.5.5), to obtain ∂G = f(θ,φ) A(φ). ∂θ − Then equation (8.5.6) becomes dH = A(φ)sin φ. dφ − We therefore set φ H(φ)= A(s)sin s ds. (8.5.7) − Z0 Note that H(0) = 0. Furthermore, it follows from (8.5.7) that H has a zero of order at least 2 at 0, compensating for the singularity of dθ at the north pole (note that sin2 φdθ is smooth everywhere). Meanwhile, H(π) = 0 from (8.5.4) since f has average zero with respect to the area form of S2 by hypothesis. Hence H(φ) defines a function on S2 which vanishes at both poles. When φ π we have → π H(φ)=+ A(s)sin s ds. Zφ Therefore H has a zero of order 2 at the south pole, as well. It follows that Hdθ is a well-defined global 1-form. Therefore Hdθ + G sin φdφ is a primitive for fα and therefore fα is exact. 104 8. DE RHAM COHOMOLOGY AND TOPOLOGY

8.6. Fubini–Study form and volume form on CPn CP1 The Fubini–Study metric g and the 2-form αF S on were defined dx2+dy2 dx dy in Section 7.10 by setting g = 2 and α = ∧ 2 . They satisfy (1+r2) F S (1+r2) αF S(u,v) = g(Ju,v). We would like to generalize the metric and the 2-form to higher dimensions. Theorem 8.6.1. Let p,q Cn+1 represent points p,¯ q¯ CPn in homogeneous coordinates. A metric∈ g on CPn is uniquely determined∈ by the distance function H(p,q) d(¯p, q¯) = arccos | |, (8.6.1) p q | || | where H is the Hermitian inner product in Cn+1. For more details see Section 10.7. Remark 8.6.2. The formula should be compared to the analogous formula (1.6.7) for the real projective space. We exploit the metric g on CPn determined by the distance function of formula (8.6.1), and the complex structure J in the tangent space at each point of CPn to define the Fubini–Study form as in Section 7.8. Definition 8.6.3. The Fubini–Study form α Ω2(CPn) is the F S ∈ differential 2-form αF S(v,w)= g(Jv,w).

We will provide an explicit formula for the Fubini–Study 2-form αF S on CPn in Section 10.8. Theorem 8.6.4. The wedge power

n ∧ αF S of the Fubini–Study form is a nonzero 2n-form at every point of CPn. Proof. It suffices to check such a relationship in the tangent space at a point. The Fubini–Study form at a point is the symplectic form n dxj dyj. Its top exterior power is j=1 ∧ P n! dx1 dy1 dxn dyn, (8.6.2) ∧ ∧···∧ ∧ as a special case of the calculation peformed in the context of the proof of Wirtinger’s inequality in Section 6.10. The form (8.6.2) is nonzero and spans the line 2n(Cn).  V 8.7. CUP PRODUCT, RING STRUCTURE IN DE RHAM COHOMOLOGY 105

8.7. Cup product, ring structure in de Rham cohomology De Rham cohomology of a manifold M possesses a natural product structure, described as follows. Recall the following:

(1) The exterior algebra (Tp∗) at every point p M possesses a wedge product. ∈ (2) The associated complexV of differential forms Ω(M) similarly possesses a product where two differential forms are multiplied pointwise using the wedge product in the exterior algebra. Definition 8.7.1. The cup-product of cohomology classes is the operation ∪ : Ha (M) Hb (M) Ha+b(M) (8.7.1) ∪ dR × dR → dR defined by the anticommuting wedge product at the level of the representing differential forms. ∧ The cup product is well-defined due to the following lemma. Lemma 8.7.2. The wedge product of a pair of closed differential forms is still closed. Proof. This is immediate from the Leibniz rule (5.6.1). Indeed, if forms α and β are both closed, then d(α β)= dα β α dβ = 0.  ∧ ∧ ± ∧ Theorem 8.7.3. The product in cohomology is independent of the choice of representative closed differential forms. Proof. Let α and β be closed forms, and γ an arbitrary form such that α and dγ are are in the same k. Then (α + dγ) β = α β + dγ β = α β + d(γ β) ∧ ∧ V∧ ∧ ∧ since β is closed. Therefore adding an exact form to α does not change a b the cohomology class [α β]. Thus, if [α] HdR(M)and[β] HdR(M), then we have a well-defined∧ class ∈ ∈ [α β] Ha+b(M), ∪ ∈ dR proving the theorem.  Definition 8.7.4. The de Rham cohomology ring is the graded ring4 Hk (M), k =0,...,n ⊕k dR where (1) the additive structure comes from the real vector space struc- ture on the individual groups, and

4algebra medureget 106 8. DE RHAM COHOMOLOGY AND TOPOLOGY

(2) the multiplicative structure is given by the cup product of (8.7.1). ∪ 8.8. Cohomology of complex projective space The following result is a generalisation of Corollary 8.5.1. Theorem 8.8.1. The Betti numbers of complex projective space CPn satisfy

n 1 foralleven k =0, 2,..., 2n bk(CP )= (0 forallother k. Remark 8.8.2. The geometric counterpart of this result is the de- composition of complex projective space with a single cell in each even dimension between 0 and 2n; see (7.6.1). Recall that the closure of these cells gives precisely the complex projective subspaces CPk. We have the following analogous fact for 2-dimensional homology (dealt with in Section 9.11) and higher homology of CPn (given here for general culture). Theorem 8.8.3. The complex submanifolds p CP1 CP2 3 n 1 n { } ⊆ ⊆ ⊆ CP CP − CP generate all of the homology of complex projective⊆ ··· space ⊆ CPn. ⊆ More precisely, the 2-homology class represented by the surface 1 n CP is a generator of H2(CP ; Z) which is isomorphic to Z. The 4- homology class represented by the 4-submanifold CP2 is a generator n of H4(CP ; Z) which is similarly isomorphic to Z for n 2, etc. Note that Theorem 8.8.1 can be refined as follows,≥ taking into ac- count the ring structure in cohomology with product operation de- scribed in Section 8.7. ∪ Theorem 8.8.4. The cohomology ring of CPn is the truncated 5 polynomial ring in a single 2-dimensional variable. Namely, the ring is generated by a single class ω H2 (CPn), with the unique relation ∈ dR (n+1) ω∪ =0, where ω can be taken to be the class of the Fubini–Study 2-form αF S Ω2(CPn). ∈ Corollary 8.8.5. The 2n-dimensional class n 2n n ω∪ H (CP ) ∈ dR spans the group H2n (CPn) R. dR ≃ 5Katum: kuf, tet, vav, memsofit 8.9. ABELIANISATION IN GROUP THEORY 107

8.9. Abelianisation in group theory In this section and the ones following, we provide a brief review of fundamental groups and homology groups.

Definition 8.9.1. Given a group π defined by generators (gi) and relations (rk), one usually writes π = g r . h i | ki Example 8.9.2. The group Z can be presented as follows:

Z = g ∅ . h 1 | i

Example 8.9.3. The cyclic group Zn = Z /n Z can be presented as follows: Z /n Z = g gn , h 1 | 1 i where the right-hand side is written multiplicatively. The multiplica- tive group can be thought of as the subgroup of S1 C consisting of 2πi ⊆ n the n-th roots of unity, with g1 = e . Definition 8.9.4. The commutator of two elements g,h π of a group π is the element ∈

1 1 [g,h]= ghg− h− π. ∈ Example 8.9.5. The group Z2 can be presented as follows in mul- tiplicative notation: Z2 = g ,g [g ,g ] . h 1 2 | 1 2 i By augmenting the set of relations of a group π by the commutation relations between each pair of generators, we obtain an abelian group πab called the abelianisation of π, as follows.

Definition 8.9.6. Given a group π = gi rk , its abelianisa- tion πab is the group h | i

πab = g r , [g ,g ] i, j i k i j ∀ D E obtained by adding a commutation relation for each pair of generators. ab Thus in the abelianisation π , the formulas [¯gi, g¯j] = e automat- ically hold, where e πab is the neutral element andg ¯ is the image ∈ i of gi under the quotient homomorphism. 108 8. DE RHAM COHOMOLOGY AND TOPOLOGY

8.10. The fundamental group π1(M) The fundamental group of a connected manifold M is the group defined as follows. Consider a pair (M,p) where p M is a fixed basepoint. ∈ Definition 8.10.1. A based loop in (M,p) is a map f : S1 M such that f(e0i)= p. → We use based loops to define equivalence using homotopy as in intro- 6 ductory algebraic topology to define the fundamental group π1(M,p). Proposition 8.10.2. A based loop defines a trivial class in the 1 fundamental group π1(M,p) if and only if the map S M can be “filled” by a disk D, i.e., extended to a map D M where→ ∂D = S1. → Proposition 8.10.3. The isomorphism class of the fundamental group π1(M,p) of a connected manifold M is independent of the choice of the basepoint in M. The basepoint is frequently suppressed from the notation for the fundamental group, and one writes simply π1(M). Theorem 8.10.4. We have the following important cases: (1) The fundamental group of Sn,n 2, is trivial. (2) The fundamental group of CPn,n≥ 1, is trivial. (3) The fundamental group of the real≥ projective space RPn,n 2 n ≥ is the finite cyclic group Z2: π1(RP )= Z2. (4) The fundamental group of the torus T2 is the infinite group Z2: 2 2 π1(T )= Z . The fundamental group of a surface of genus g 2 is not abelian. The abelianisation of the fundamental group of a surface≥ of genus g is the group Z2g; see Section 9.6.7 for details.

8.10.1. The second homotopy group π2(M). This material is op- tional. Similarly to the fundamental group π1(M,p), one defines the second 2 homotopy group π2(M,p) as the group generated by based maps S M modulo based homotopy connecting a pair of based maps. Similarly to→ the case of π1, a map belongs to a trivial class in π2(M) if it can be extended to a map of the 3-ball: B3 M. → Remark 8.10.5. Unlike the fundamental group, the second homotopy group π2(M) is always abelian. Example 8.10.6. The second homotopy group of the 2-sphere S2 = CP1 2 is infinite cyclic: π2(S )= Z.

6See e.g., Armstrong [Ar83]. 8.10. THE FUNDAMENTAL GROUP π1(M) 109

More generally, we have the following result. Theorem 8.10.7. The second homotopy group of the CPn is infinite n cyclic: π2(CP )= Z for each n = 1, 2, 3,... Remark 8.10.8 (Remark for general culture). The long exact sequence 2 (not defined in this course) of the Hopf fibration helps calculate π3(S ) Z. This group is generated by the Hopf map S3 S2 of Section 7.5. ≃ →

CHAPTER 9

Homology groups; stable norm

To formulate Gromov’s inequality for CPn we will need the notion of homology group Hk of a manifold M at least for k =1, 2. 9.1. 1-cycles on manifolds

The singular homology groups with integer coefficients, Hk(M; Z) for k =0, 1,... of M are abelian groups which are homotopy invariants of the manifold M. Developing the singular homology theory in arbi- trary dimension is time-consuming. The cases that we will be primarily interested in are

(1) the 1-dimensional homology group H1(M; Z) treated in Sec- tion 9.6, and (2) the 2-dimensional homology group H2(M; Z) treated in Sec- tion 9.11. In these cases, the homology groups can be characterized more easily without the general machinery of singular simplices and chains. We will therefore follow such an easier approach. Definition 9.1.1 (Circle with orientation). Let S1 C be the unit circle, which we think of as a 1-dimensional manifold with⊆ an orienta- tion given by a parametrisation in the counterclockwise direction. Remark 9.1.2. The existence of orientations on parametrized loops in C is familiar from the course in complex functions. See Sections 9.2 and 9.4 for a more general treatment of orientations and induced ori- entations. Definition 9.1.3. A1-cycle C on a manifold M is an integer linear combination C = nifi i X where n Z is called the multiplicity (ribui), while each i ∈ f : S1 M i → is a loop given by a smooth map from the circle to M, where each 1 loop fi is endowed with the orientation coming from S . 111 112 9. HOMOLOGY GROUPS; STABLE NORM

Definition 9.1.4. The space of 1-cycles on M is denoted Z1(M; Z).

9.2. Orientation on a manifold Let M be a connected n-dimensional manifold, so that for ev- ery p M, we have dim(Tp∗M) = n. The n-th exterior algebra at p ∈ n is 1-dimensional, i.e., dim (Tp∗M) = 1. Thus, any differential n- form η Ωn(M) can be written in a coordinate chart as ∈ V  η = f(u1,...,un) du1 dun ∧···∧ where f is a function defined in the coordinate patch. Definition 9.2.1. A differential n-form η Ωn(M) is called a non- degenerate on M if it is nonvanishing at every point,∈ i.e., f(u1,...,un) = 0. 6 Remark 9.2.2. A nondegenerate n-form may not necessarily exist globally on M. Thus, when M = RP2 such a form does not exist. Namely, every 2-form on RP2 necessarily vanishes at some point. Definition 9.2.3. Two nondegenerate n-forms η andη ˜ are equiv- alent if they differ by a positive function g = g(p) > 0 on M: η η˜ ( g > 0)( p M), η(p)= g(p)˜η(p). (9.2.1) ∼ ⇐⇒ ∃ ∀ ∈ Definition 9.2.4. An orientation on a connected manifold M is an equivalence class in the sense of (9.2.1) of a nondegenerate n-form η Ωn(M). ∈ Definition 9.2.5. On the circle S1, the standard orientation is represented by the nondegenerate differential 1-form dθ (or any positive multiple of it). The “opposite” orientation on the circle represented by dθ. This can be thought of as an arrow marked along the circle, indicat− ing the clockwise or counterclockwise direction. Example 9.2.6. The area form dx dy in the plane R2 or the torus T2 represents an orientation in the∧ plane or on the torus. The opposite orientation is represented by the form dx dy. − ∧ CP1 Example 9.2.7. The Fubini–Study differential 2-form αF S on defined by αF S(u,v)= g(Ju,v) represents an orientation. The opposite orientation is represented by α . − F S 9.4. INDUCED ORIENTATION ON THE BOUNDARY ∂M 113

9.3. Interior product y of a form by a vector To define an orientation on cycles in Section 9.4, we will exploit the interior product. Recall that an n-form η at a point p M can be ∈ thought of as an antisymmetric multilinear form on TpM. For example, for a 2-form η we can write η(v,w) R where v,w T M. ∈ ∈ p Definition 9.3.1 (Case n = 2). Let v TpM be a tangent vector, 2 ∈ 1 and η p M a 2-form. The interior product ιvη = vy η p M is a 1-form,∈ defined by setting ∈ V V u T M, (vy η)(u)= η(v,u). ∀ ∈ p We define the interior product similarly for n-forms. Definition 9.3.2. The interior product of an n-form η by a vector v is an (n 1)-form denoted ιv(η)= vy η, obtained by substituting v into the first− variable of η:

ιv(η)=(vy η)(x1,...,xn 1)= η(v,x1,...,xn 1). − − Example 9.3.3. Consider the area form η = dx dy in the plane. ∂ ∧ If v = ∂x then vy η = dy. Lemma 9.3.4. If w = ∂ then wy η = dx. ∂y − Proof. The 1-form wy η can be calculated as follows: ( ∂ y η)(u)= η( ∂ ,u)= η(x, ∂ ). ∂y ∂y − ∂y Thus ( ∂ y η)( ∂ )= 1, ( ∂ y η)( ∂ ) = 0. Hence the 1-form ∂ y η coin- ∂y ∂x − ∂y ∂y ∂y cides with dx.  − Example 9.3.5. In polar coordinates, since η = dr rdθ, we have ∂ y η = rdθ and ∂ yη = rdr. ∧ ∂r ∂θ − Remark 9.3.6. The interior product satisfies a Leibniz rule: ι (ω η)=(ι ω) η +( 1)deg(ω)ω (ι η) v ∧ v ∧ − ∧ v (see [14, p. 401]). We will not need this rule.

9.4. Induced orientation on the boundary ∂M We will now exploit the exterior product of Section 9.3 to define the induced orientation on the boundary1 of a manifold.

1Safa with a “sin”. 114 9. HOMOLOGY GROUPS; STABLE NORM

Definition 9.4.1. M is an n-manifold with boundary if we have a submanifold S M where S is (n 1)-dimensional (not necessarily connected), and⊆ an open neighborhood− of S in M is diffeomorphic to the cylinder S I where I = (0, 1] is a halfclosed interval, where S is identified with ×S 1 M. × { }⊆ 1 n 1 We use the notation ∂M for the boundary of M. Let(u ,...,u − ) be coordinates in a neighborhood of S, and let t I parametrize the ∂ ∈ interval, so that ∂t is an (inward-pointing) vector field transverse (i.e., not tangent) to− the boundary S. Definition 9.4.2 (Induced orientation). Let M be an oriented manifold with boundary S, with orientation represented by a nonde- n generate form orientM Ω (M). The induced orientation orientS on S ∈ ∂ is generated by the interior product ( ∂t )y orientM restricted to S, viewed as a form on S itself: − ∂ n 1 orient =( )y orient Ω − (S). S − ∂t M ∈ Remark 9.4.3 (Connected components). The equivalence class of ∂ the form ( ∂t )y orientM on S, relative to the equivalence relation (9.2.1), provides the− orientation on each connected component of S. 9.5. Surfaces and their boundaries

Definition 9.5.1. We denote by (Σg,∂Σg) a compact surface with boundary ∂Σg, where g is the genus of the surface. The following is a basic result in the topology of surfaces.

Lemma 9.5.2. The boundary ∂Σg is a disjoint union of finitely many circles. In Section 9.4 we defined the induced orientation on the boundary via the interior product denoted y; see Definition 9.4.2. We therefore obtain the following.

Corollary 9.5.3. Assume the surface Σg is oriented. Then a nonvanishing 2-form η Ω2(Σ ) representing the orientation on the ∈ g surface induces an orientation ( ∂ )y η on each boundary circle. − ∂t We thus obtain an orientation-preserving identification of each bound- 1 ary component of Σg with the standard unit circle S C with its counterclockwise orientation as in Definition 9.2.5. ⊆ 2 Example 9.5.4. Let Σ0 R be the unit disk together with the orientation represented by η =⊆dx dy. At points other than the origin we can represent η by η = rdr dθ∧ . The inward-pointing vector field ∧ h 9.6. RESTRICTION ⇂∂Σg AND 1-HOMOLOGY GROUP 115

1 ∂ along the boundary circle S = ∂Σ0 is ∂r . The induced orientation 1 ∂ − on S is ( ∂r )yη = rdθ. This is the clockwise orientation on the circle. − −

9.6. Restriction h⇂∂Σg and 1-homology group

Let M be a manifold (orientable or not). Let Σg be an oriented surface. Given a map h: Σ M, its restriction g →

h⇂∂Σg to the boundary produces a 1-cycle ∂Σg Z1(M; Z) in the sense of Definition 9.1.3, with orientation defined∈ as in Corollary 9.5.3. We define 1-boundaries as the 1-cycles that can be obtained by restriction, as follows. Definition 9.6.1. A 1-boundary in a manifold M is a 1-cycle n f Z (M; Z) i i ∈ 1 i X such that there exists a map of an oriented surface h: Σg M, for some genus g which may depend on the 1-cycle, whose restriction→ ⇂ to ⇂ the boundary satisfies h ∂Σg = nifi. i X Definition 9.6.2. The space of all 1-boundaries is denoted B (M; Z) Z (M; Z). 1 ⊆ 1 Definition 9.6.3. The 1-homology group of M with integer coef- ficients is the quotient group H1(M; Z) = Z1(M; Z)/B1(M; Z). Given a 1-cycle C Z (M; Z), its homology class is denoted [C] H (M; Z). ∈ 1 ∈ 1 The relation between the fundamental group and the 1-homology group is given by the following theorem. For the notion of abelianisa- tion see Definition 8.9.6.

Theorem 9.6.4. The 1-homology group H1(M; Z) is the abeliani- sation of the fundamental group π1(M): ab H1(M; Z)= π1(M) . Remark 9.6.5 (Free loops). A significant difference between the fundamental group and the first homology group is the following. While only based loops participate in the definition of the fundamental group, the definition of H1(M; Z) involves free (not based) loops. 116 9. HOMOLOGY GROUPS; STABLE NORM

The fundamental groups of the real projective plane RP2 and the 2-torus T2 are abelian, and therefore isomorphic to their first homology groups, yielding the following examples. Proposition 9.6.6. We have 2 (1) H1(RP ; Z) Z /2 Z, (2) H (T2; Z) ≃Z2. 1 ≃ We mention the following result for general culture. Theorem 9.6.7. The fundamental group of an orientable closed surface Σg of genus g is a group on 2g generators with a single relation r which is a product of g commutators: π (Σ ) a ,b ,...,a ,b r , r =[a ,b ] [a ,b ]. (9.6.1) 1 g ≃ 1 1 g g 1 1 ··· g g Corollary 9.6.8. Abelianizing the group (9.6.1), we obtain

H (Σ ; Z) Z2g . 1 g ≃ 9.7. Length of cycles and 1-homology classes Now let M be a Riemannian manifold. We will deal with norms in homology groups associated with Riemannian metrics on M. Assume that an n-dimensional manifold M has a Riemannian metric given in a coordinate patch by the metric coefficients (gij). We will denote by t a parameter on the circle S1. Definition 9.7.1. Given a smooth loop f : S1 M, its volume (length) with respect to the metric of M is →

i j vol(f)= gij(f(t)) α ′(t)α ′(t) dt Z q where αi are the components of the loop with respect to a suitable coordinate chart: f(t)=(α1(t),...,αn(t)), and αi′ are the derivatives with respect to t. As usual, if the loop travels through several charts of M, we work with partitions of M to define the global volume (length). ˜ Definition 9.7.2. The volume (length) of a 1-cycle C = i nifi, ni Z in M, is the combined length of all the individual loops, with∈ multiplicity : P | | vol(C˜)= n vol(f ). | i| i i X Since we take absolute values of all the integer coefficients, the volume is by definition nonnegative. 9.8. LENGTH IN 1-HOMOLOGY GROUP OF ORIENTABLE SURFACES 117

Definition 9.7.3. Let C H1(M; Z) be a 1-homology class. We define the volume of C as the∈ infimum of volumes of representative 1- cycles: vol(C) = inf vol(C˜): C˜ C , ∈ where the infimum is over all cycles C˜ = n f , n Z, representing  i i i i ∈ the class C H1(M; Z). ∈ P Example 9.7.4. In the case of the real projective plane, there is only one nontrivial first homology class. The volume of this class with respect to a given metric g on RP2 is then the least length of a non- contractible2 loop for the metric g. Proposition 9.7.5. For the standard metric of constant Gaussian curvature +1 on RP2, the volume (length) of the nontrivial class C 2 2 ∈ H1(RP ; Z) is vol(C)= π. The projection to RP of each longitude on the sphere under the double cover S2 RP2represents a minimizing closed geodesic in the class C. → The standard metric on RP2 satisfies the boundary case of equality in Pu’s inequality; see Section 10.1.

9.8. Length in 1-homology group of orientable surfaces For closed orientable surfaces, the volume of 1-homology classes is multiplicative, in the following sense.

Theorem 9.8.1. Let M = Σg be a closed orientable surface en- dowed with a metric. Let C H1(M; Z). Then for all j N, we have ∈ ∈ vol(jC)= j vol(C), (9.8.1) where jC denotes the class C + C + ... + C, with j summands. This is proved below. Remark 9.8.2. Another way of writing identity (9.8.1) for a ho- mology class C is as follows: 1 j N, vol(C)= vol(jC). (9.8.2) ∀ ∈ j A similar formula will define the stable norm in homology for higher- dimensional manifolds; see formula (9.9.1). We have the following elementary fact about the topology of a cylin- der.

2lo-kvitza, with kaf 118 9. HOMOLOGY GROUPS; STABLE NORM

Lemma 9.8.3. A loop γ going around a cylinder twice necessarily + has a point of self-intersection enabling a decomposition γ = γ γ− into a pair of noncontractible loops on the cylinder. ∪ Proof. We will give a proof in the case when the loop is the graph of a 4π-periodic function f(t). Consider the difference h(t) = f(t) f(t +2π). Clearly, h(t) = h(t +2π). Thus h takes both positive− and negative values (if h(t) is− positive, then h(t +2π) is negative). By the intermediate function theorem, the function h must have a zero t0. Then f(t0)= f(t0 +2π). Then the parameter value t0 produces a point of self-intersection of the loop. 

Proof of Theorem 9.8.1. Choose a representative loop C˜ C (for example, a minimizing closed geodesic). Choose the basepoint∈p ∈ M to be a point of C˜. Let N be the cover of M = Σg whose fundamental group is infinite cyclic and is generated by the class of C˜ in π1(M,p). Then N is a cylinder. The lift of C˜ to N represents a generator of H1(N; Z). To fix ideas, we let j = 2. Consider a minimizing loop γ representing a multiple class 2C. By Lemma 9.8.3, the loop γ will necessarily intersect itself in a suitable + point p N resulting in a decomposition γ = γ γ− at the level ∈ + ∪ of loops and decomposition [γ]=[γ ]+[γ−] at the level of homology classes. We take the shorter of the two loops γ±. It produces a minimizing loop in the class C, proving the identity (9.8.1) in the case j = 2. The general case follows similarly. 

9.9. Stable norm for higher-dimensional manifolds In Section 9.8, we studied the volume (length) of 1-homology classes of an orientable surface. The phenomenon of multiplicativity of vol- ume (length) expressed by identity (9.8.2) no longer holds in higher- dimensional manifolds. Namely, the volume (length) of a 1-homology class is not necessarily multiplicative. However, the limit as j exists and is called the stable norm. → ∞ Definition 9.9.1. Let M be a compact manifold of dimension n endowed with a metric. The associated stable norm of a class C k k ∈ H1(M; Z) is the limit 1 C = lim vol(jC). (9.9.1) j k k →∞ j 9.10.SYSTOLEANDSTABLE1-SYSTOLE 119

Remark 9.9.2. The existence of the limit is nontrivial and was proved by Federer in [Fe69]. Remark 9.9.3. It is obvious from the definition that one has C vol(C). k k≤ However, the inequality may be strict in general. By Theorem 9.8.1, for orientable 2-dimensional manifolds M we have C = vol(C). k k Proposition 9.9.4. The stable norm vanishes for a class of finite order.

Proof. If C H1(M; Z) is a class of finite order, one has finitely many possibilities∈ for vol(jC) as j varies. Let V be the maximal such volume. Then the expression 1 vol(jC) V in (9.9.1) tends to zero, j ≤ j proving that C = 0.  k k Example 9.9.5. Since the fundamental group of the real projective n space is the finite cyclic group Z2, the stable norm on H1(RP ; Z) Z2 vanishes. ≃ Similarly, if two classes differ by a class of finite order, their sta- ble norms coincide. Thus the stable norm passes to the quotient lat- tice L1(M) defined below. Definition 9.9.6. Let M be a compact manifold. The torsion sub- group of H (M; Z) will be denoted T (M) H (M; Z). The quotient 1 1 ⊆ 1 lattice L1(M) is the lattice

L1(M)= H1(M; Z)/T1(M). Recall that every finitely generated abelian group is a product Zb F where F is finite. The Betti numbers were defined in Defi- nition× 7.13.3. Proposition 9.9.7. Let M be a compact manifold. Then the lat- b1(M) tice L1(M) is isomorphic to Z , where b1 is the first Betti number of M.

9.10. Systole and stable 1-systole Let M be a non-simply-connected closed manifold endowed with a Riemannian metric g.

Definition 9.10.1. The 1-systole sys1(M, g) of a metric g on a surface M is the least g-length of a noncontractible loop on M. 120 9. HOMOLOGY GROUPS; STABLE NORM

We now consider the associated stable norm as in (9.9.1) in the k k homology group H1(M; Z). Recall that λ1(L) is the least length of a nonzero vector in a lattice L; see Definition 6.4.1.

Definition 9.10.2. The stable 1-systole of M, denoted stsys1(M), is the least norm of a 1-homology class of infinite order: stsys (M) = inf C : C H (M; Z) T (M) 1 k k ∈ 1 \ 1 = λ L (M), 1  1 k k Corollary 9.10.3. For an arbitrary metric g on the 2-torus T2, the 1-systole and the stable 1-systole coincide: T2 T2 sys1( ) = stsys1( ). This is immediate from Theorem 9.8.1. For an arbitrary orientable surface M, the group H1(M; Z) contains no torsion, and a similar re- sult holds for the homology systole (where unlike the homotopy systole of Definition 9.10.1, the infimum is taken only over homologically non- trivial loops).

9.11. 2-homology group and 2-systole Let M be a manifold. We will now define the 2-homology group in a way similar to the procedure for the 1-homology group defined in Section 9.6. An example of particular interest to us is the complex projective space M = CPn.

Definition 9.11.1. A 2-cycle in M is a linear combination nifi, where ni Z, of maps fi : Σg M of oriented closed surfaces of ∈ i → P arbitrary genus gi into M. Remark 9.11.2. A significant difference from the case of 1-homology group where we used a unique source manifold (namely the circle) is that we now allow arbitrary genus for the source surface. Thus, each fi is a map fi : Σgi M where Σgi is a closed surface of genus gi depend- ing on i. → A 1-boundary was defined in Section 9.1 as the boundary of a map from a surface to M. We define 2-boundaries in a similar way.

Definition 9.11.3. A 2-boundary in M is a 2-cycle i nifi real- izable as the boundary ∂N of a map F : N M from an oriented 3- manifold N into M: → P

nifi = F ⇂ ∂N . i X 9.12. STABLE NORM IN 2-HOMOLOGY 121

Remark 9.11.4. The orientation induced on the boundary sur- face ∂N was defined in Section 9.2 via the interior product y.

Definition 9.11.5. We let Z2(M; Z) and B2(M; Z) be the spaces of 2-cycles and 2-boundaries of M, respectively. Definition 9.11.6. The 2-homology group is the quotient group

H2(M; Z)= Z2(M; Z)/B2(M; Z). Theorem 9.11.7. We have the following important cases:

(1) For every orientable surface Σg of genus g 0, H2(Σg; Z) Z. n ≥ ≃ (2) We have H2(S ; Z) Z if n =2 and 0 otherwise. n ≃ (3) We have H2(CP ; Z) Z for each n 1. The group is generated by the class≃ of the 2-cycle given≥ by the inclusion CP1 CPn. → 9.12. Stable norm in 2-homology

If M is a manifold with a metric g, the stable norm in H2(M; Z) is defined in a way similar to the case k = 1, using thek k stabilisation formula (9.9.1), as follows. We first define the area of a map of a surface into M as fol- lows. Consider a domain U R2 with coordinates (u1,u2), and a ⊆ ∂x ∂x map x: U M. We form the tangent vectors x1 = ∂u1 and x2 = ∂u2 in T M, and→ calculate the area x x of the parallelogram in T M p | 1 ∧ 2| p spanned by x1 and x2 with respect to the metric g on M. Definition 9.12.1. The volume (area) of the map x: U M is the integral →

1 2 1 2 vol(x)= x1 x2 du du = det(gij) du du , (9.12.1) U | ∧ | U Z Z q where g = g(x ,x ) and det(g )= g g g2 . ij i j ij 11 22 − 12 The area of the parallelogram can be calculated by the usual for- mula as x1 x2 = x1 x2 sin α where α is the angle between the | ∧ | | || | 1 2 vectors. Here x = g(x ,x ), and cos α = g(x ,x ) where g is the | i| i i x1 x2 metric on M. As in the case of 1-homology, we define| | | | the volume of a p 2-homology class. Definition 9.12.2. Given a homology class C H (M; Z), we ∈ 2 define vol(C) as the infimum of volumes of representative 2-cycles C˜ C. ∈ The stable norm is then defined as follows. 122 9. HOMOLOGY GROUPS; STABLE NORM

Definition 9.12.3. Let C H2(M; Z). We define the stable norm in 2-dimensional homology by∈ a formula similar to (9.9.1): 1 C = lim vol(jC). j k k →∞ j The stable 2-systole is defined similarly to the case of stable 1- systole as follows. Definition 9.12.4. The stable 2-systole of M is stsys (M) = inf C : C H (M; Z) T (M) 2 k k ∈ 2 \ 2 = λ L (M), , 1  2 k k Z where L2(M) is the lattice given by the quotient of H2(M; ) by the torsion subgroup T2 H2(M; Z), and is the stable norm as in Definition 9.12.3. ⊆ k k The stable 2-systole is the geometric invariant appearing in Gro- mov’s stable systolic inequality for complex projective space (see Sec- tion 10.9). This inequality is a higher dimensional analogue of Pu’s systolic inequality for the real projective plane RP2. In Section 10.1 we review Pu’s inequality as motivation for Gromov’s inequality. CHAPTER 10

Gromov’s stable systolic inequality for CPn

Gromov’s stable systolic inequality1 concerns metrics on the com- plex projective space. As motivation, we will first consider an analogous inequality for metrics on the real projective plane.

10.1. Pu’s inequality for real projective plane Recall that the real projective plane RP2 is a smooth non-orientable surface. Among its elementary properties are the following. 2 2 (1) π1(RP )= H1(RP ; Z) Z2. (2) The orientable double cover≃ of RP2 is the sphere S2. 2 (3) Every metric g on RP naturally lifts to a Z2-invariant metric, denoted g˜, on the double cover S2. 2 2 (4) RP is the quotient of S by the action of Z2 where the non- trivial element a: S2 S2 is the antipodal map of S2, namely a(p)= p when S2 is→ thought of as the unit sphere in R3. − Thus we have a smooth map S2 RP2 which is 2-to-1. → RP2 Definition 10.1.1. The systole sys1( , g) can be thought of in two equivalent ways: (1) the least g-length of a loop representing the nontrivial homol- 2 ogy class in H1(RP ; Z); (2) the least g˜-length of a path joining points p and a(p) on S2 as p varies over S2, whereg ˜ is the metric on S2 obtained as the lift of g. Theorem 10.1.2 (P. M. Pu). Every metric g on the real projective plane RP2 satisfies the sharp 2 inequality π sys (g)2 area(g), 1 ≤ 2 where sys1 is the 1-systole. The boundary case of equality is attained precisely when g is of constant Gaussian curvature.

1Gromov has many other systolic inequalities, as well; see [9]. 2haduk 123 124 10. GROMOV’S STABLE SYSTOLIC INEQUALITY FOR CPn

This is proved in Chapter 12 and again in Chapter 13.3 Remark 10.1.3. There is an analogous optimal inequality called Loewner inequality for the torus; see Chapter 11. Our main interest in Pu’s inequality in this chapter is as motiva- tion for the complex case. In the complex case, we have an analogous inequality for the stable systole (see Definition 9.12.4): stsys (g)n n! vol(g), 2 ≤ for every metric g on CPn; see Section 10.9. The Fubini–Study metric satisfies the boundary case of equality. Remark 10.1.4. Is there a systolic analogue for CP2 of Pu’s in- 2 equality, of the form sys2(g) C vol(g)? It turns out that there are counterexamples to such an inequality;≤ see [12]. We will formulate Gromov’s stable systolic inequality for complex projective space in Section 10.9. We first review some material from calculus on manifolds. 10.2. Integration over 1-cycles and over 2-cycles This section is intended as motivation for the process of integration of closed forms over cycles in a manifold; see Section 10.3. Consider a circle Cr = z C: z = r of radius r> 0 centered at the origin of C. The counterclockwise{ ∈ | | orientation} turns the circle into a 1-cycle (see Section 9.2). Thus we can think of Cr as an element of the space of cycles in M = C 0 with integer coefficients: \ { } C Z (M; Z). r ∈ 1 In polar coordinates (r, θ) in M, the differential 1-form dθ is defined everywhere in M. We will denote its restriction to the circle Cr M by the same symbol dθ. Here θ is a multi-valued function on M⊆and also on the circle Cr.

Proposition 10.2.1. The 1-cycle Cr Z1(M) satisfies the relation dθ =2π, independent of r. ∈ Cr R Proof. We can parametrize the circle by the interval [0, 2π], for instance by means of the parametrisation reiθ, θ [0, 2π]. (10.2.1) ∈ 3There is a similarity between Pu’s inequality and the isoperimetric inequality for Jordan curves in the plane. Both inequalities relate a length and an area, and both are sharp (optimal), but the inequalitites comparing length and area go in opposite directions. 10.2. INTEGRATION OVER 1-CYCLES AND OVER 2-CYCLES 125

Thus we can think of θ as a single-valued function on the circle with a i0 i0 point removed: θ : Cr re R, omitting the point re R C which corresponds to\ the { values} → 0 or 2π of the parameter∈θ.4 ⊆We use the parametrisation (10.2.1) to perform the integration using the fundamental theorem of calculus:

2π dθ = θ 0 =2π, ZCr proving the proposition.   Another phenomenon related to integration of closed forms over cycles is Cauchy’s residue theorem. Proposition 10.2.2 (Residue calculation). The residue5 of the 1 function z at the origin is 1. Proof. We have dz ireiθdθ 2π = iθ = i dθ =2πi. z =r z re 0 I| | Z Z Hence the residue at the origin equals 1 by Cauchy’s theorem.  Remark 10.2.3. Cauchy’s theorem on the residues is a special case of the fact that the contour integral depends only on the homology class [Cr] of the loop Cr in the group H (C 0 ; Z). 1 \ { } This in turn is a special case of the general result on integration over cycles in manifolds; see Section 10.3. Remark 10.2.4. Similar remarks apply to integrating a 2-form over a 2-cycle in a manifold. Here 2-cycles are taken in the sense of the homology theory developed in Section 9.11. The area of a 2-cycle is calculated by means of formula (9.12.1) from classical differential geometry as in Section 9.12. Namely, the area of a 2-cycle F (x,y) with coordinates x = u1, y = u2, is expressed by a double integral:

area = F F dxdy | x ∧ y| ZU 4As already mentioned in note 1 in Section 8.2, the closed 1-form dθ, inspite of appearances, is not exact, i.e., it is not a 1-coboundary. This is because θ is not a true function on the circle but only a multi-valued one. Note that the 1-form rdθ is the volume (length) form of Cr. 5She’erit in complex function theory 126 10. GROMOV’S STABLE SYSTOLIC INEQUALITY FOR CPn

∂F ∂F where Fx = ∂x and Fy = ∂y . From the differential geometric view- point, a more correct form of the expression for the area is the double integral F F dx dy. U | x ∧ y| ∧ 10.3.R Duality of homology and de Rham cohomology In previous chapters, we have developed (1) a cohomology theory (de Rham cohomology), and (2) a homology theory (singular homology) of a manifold M. In this section we will begin to relate the two theories. Recall that Ωk(M) on M is the space of sections of the exterior k k bundle M = (T ∗M) of the manifold M. As illustrated in Sec- tions 10.2 and Remark 10.2.4, a differential k-form η Ωk(M) on M V V ∈ can be integrated over a k-cycle C Zk(M; Z) to obtain a real num- R ∈ ber C η . Integration thus defines a pairing, which we will denote by the symbol∈ , . We have defined k-cycles on a manifold M only in theR cases k =h 1 andi k = 2. These two values are meant throughout this section. We thus obtain the following lemma. Lemma 10.3.1. There exists a pairing between the space Ωk(M) of differential k-forms on M, and the space Zk(M; Z) of k-cycles in M: , : Ωk(M) Z (M; Z) R . h i × k → The relation between the two theories is given by the following theorem. Theorem 10.3.2. Let M be a compact manifold. Let η Ωk(M) ∈ be a closed differential form. Let C˜ Zk(M; Z) be a k-cycle. Then the ˜ ∈ value of η, C = C˜ η only depends on the following data: (1) the cohomology class ω =[η] Hk (M) and R ∈ dR (2) the homology class C =[C˜] H (M; Z). ∈ k Since the integral vanishes over any torsion class in Hk(M; Z), we obtain the following pairing. Definition 10.3.3. We have a pairing , : Hk (M) L (M) R, (10.3.1) h i dR × k → where Lk(M)= Hk(M; Z)/Tk(M). Duality between lattices will be understood in the sense of Defini- tion 6.3.2. Definition 10.3.4. Using the pairing (10.3.1), we define an integer k k k lattice LdR = LdR(M) in the de Rham cohomology HdR(M) in the following three equivalent ways. 10.4. VOLUME FORM ON RIEMANNIAN MANIFOLDS 127

k (1) LdR is the lattice dual to the homology lattice Lk(M) via the pairing (10.3.1). (2) Lk = ω Hk (M): η Z ( C˜ C L (M) ( η dR ∈ dR C˜ ∈ ∀ ∈ ∈ k ∀ ∈ ω) .  R (3) Choosing a basis (y ,...,y ) for L (M), we define the lat- 1 b k tice Lk by setting Lk = ω Hk (M): ω Z i =1,...,b . dR dR ∈ dR yi ∈ ∀ An integer class in de Rham cohomologyn is a classR contained in the o k integer lattice LdR(M). The fundamental result about de Rham cohomology is the following theorem. Theorem 10.3.5. Let M be a closed orientable n-manifold. Then the pairing , between L (M) and Lk (M) is non-degenerate. h i k dR Corollary 10.3.6. The lattice Lk(M) is naturally isomorphic to k ∗ k 6 the lattice LdR(M) , and LdR(M) is naturally isomorphic to (Lk(M))∗. 10.4. Volume form on Riemannian manifolds Definition 10.4.1. Let M be an oriented manifold of dimension n. A basis for Tp∗M is direct if it is compatible with the chosen orientation on M. Now suppose we are given a Riemannian metric on M. Recall the following. (1) in Section 6.6, we defined the comass norm of an alternat- ing (exterior) form; k k (2) in Section 9.2 we defined the notion of a nondegenerate top- dimensional differential form ζ Ωn(M). ∈ Definition 10.4.2. A volume form on a Riemannian manifold M is a nondegenerate form ζ Ωn(M) such that one of the following equivalent conditions is satisfied∈ at every point p M: ∈ (1) ζ = 1; k pk (2) with respect to any direct orthonormal basis ω1,...,ωn for Tp∗M, one has ζ = ω ω ; p 1 ∧···∧ n (3) in coordinates, ζ = det(g ) du1 ... dun. ij ∧ ∧ Definition 10.4.3. Thep volume of M is the integral vol(M)= ζ (10.4.1) ZM 6The integer lattice in de Rham cohomology can also be thought of as the image, under the de Rham homomorphism Hk(M; Z) Hk (M), of the integer → dR lattice in singular cohomology. Singular cohomology was not treated in these notes. 128 10. GROMOV’S STABLE SYSTOLIC INEQUALITY FOR CPn where ζ is the volume form of M. The volume form is commonly denoted dvolM . The following elementary lemma be useful in applications and par- ticularly in the proof of Gromov’s stable systolic inequality for complex projective space.

Lemma 10.4.4. For each nonnegative function h C∞(M) on the manifold, we have a bound ∈

hdvol (max ) vol(M), M ≤ h ZM where maxh is the maximal value of h on M. Example 10.4.5. In the case of the unit circle S1, the 1-form dθ has 1 unit pointwise norm, therefore S1 dθ = 2π = vol(S ). Note that [dθ] is not in L1 (S1) where L1 (S1) is the integer lattice. dR dR R Lemma 10.4.6. The form dθ H1 (S1) (10.4.2) 2π ∈ dR is a generator of the integer lattice L1 (S1) Z. dR ≃ Indeed, the form (10.4.2) integrates to 1 over S1. Let us now con- sider CP1.

Proposition 10.4.7. The Fubini–Study 2-form αF S divided by π represents a generator of the integer lattice in de Rham cohomology:

α FS L2 (CP1) Z . π ∈ dR ≃ Proof. The Fubini–Study  volume 2-form in an affine neighbor- 1 dx dy hood C CP can be written as α = 2∧ 2 2 . The form integrates ⊆ F S (1+x +y ) to π as shown in Theorem 8.3.1. 

k k 10.5. Comass norm in Ω (M) and in HdR(M) Let M be a Riemannian manifold. We denote by the comass k k norm in the exterior algebra (Tp∗M), as in Definition 6.6.2. Definition 10.5.1. The comassV norm of a differential k-form η Ωk(M) is the supremum of the pointwisek k∞ comass norms: ∈ η = sup ηp : p M . k k∞ k k ∈ Next, we define a notion of comass norm in de Rham cohomology, as follows. 10.7. FUBINI–STUDY METRIC ON COMPLEX PROJECTIVE LINE 129

Definition 10.5.2. Let ω Hk (M) be a class in de Rham coho- ∈ dR mology. The comass norm ω ∗ is the infimum of comass norms η of the representative differentialk k k-forms η ω, or equivalently k k∞ ∈ ω ∗ = inf sup ηp : p M,η ω , k k η p k k ∈ ∈ where denotes the pointwise comass norm. k k 10.6. Duality of comass and stable norm The following theorem (Theorem 10.6.1) is of central importance for the sequel. First we recall the following.

(1) Lk(M) is the lattice defined as the quotient of homology mod- ulo torsion. k (2) LdR(M) is the integer lattice in de Rham cohomology; see Definition 10.3.4. (3) The two lattices are dual; see Section 10.3. The following theorem and its corollary assert that the normed lattices are dual, as well. Duality of norms was defined in Section 6.1.

Theorem 10.6.1 (Federer; Gromov). The stable norm on Lk(M) k k k k and the comass norm ∗ on LdR(M) HdR(M) are dual to each other. k k ⊆ See Gromov [9, Section 4.34, p. 261]; Federer [Fe74, Section 4.10, p. 380] using the Hahn–Banach theorem. See also Pansu [15, Lemma 17]. Thus we obtain a strengthened version of Corollary 10.3.6.

Corollary 10.6.2. The normed lattice (Lk(M), ) and the normed k k k lattice (L (M), ∗) are dual to each other. dR k k 10.7. Fubini–Study metric on complex projective line The material in this section is a review of Section 8.3, with an eye to generalizing it to complex projective space in Section 10.8. The case of the complex projective line CP1 is special in that the Gaussian curvature K is constant: K = 4. The complex projective line CP1 = C = S2 carries a natural metric called the Fubini– ∪ {∞} Study metric gF S. The Fubini–Study 2-form was discussed in Sec- tion 7.10. Explicit formulas for the Fubini–Study metric on CPn will be given in Section 10.8. The distance function d(v,w), associated with the Fubini–Study metric, takes its simplest form in homogeneous 2 coordinates [Z0,Z1]. Given a pair of vectors v,w C , we have the cor- responding complex 1-dimensional subspaces Cv∈ C2 and Cw C2, thought of as points in the projective line. The distance⊆ between⊆ them 130 10. GROMOV’S STABLE SYSTOLIC INEQUALITY FOR CPn will be denoted, briefly, d([v], [w]). The distance is defined in terms of the standard Hermitian product H(v,w). Namely, the distance be- tween points [v], [w] in CP1 is defined by H(v,w) d([v], [w]) = arccos | |. (10.7.1) v w | || | For unit vectors, one has cos d([v], [w]) = H(v,w) = v1w1 + v2w2 , where the bar stands for complex conjugation.| | | | Remark 10.7.1. The distance vanishes (i.e. cos(d([v], [w])) = 1)if v and w differ by a complex scalar: v = λw. This is consistent with the fact that these vectors define the same point of the complex projective line. In an affine neighborhood Z = 0 in CP1, we introduce a coordi- 1 6 nate z = Z0/Z1 = x + iy. Theorem 10.7.2. With respect to coordinate z, the Fubini–Study 1 metric gF S on CP corresponding to the distance function (10.7.1) takes the form dz 2 dx2 + dy2 dr2 g = | | = = . (10.7.2) F S (1 + zz)2 (1 + r2)2 (1 + r2)2 Here r2 = zz = x2 + y2, where r is the first of the usual polar coordi- nates (r, θ). We recall some elementary global-geometric properties of the met- ric. Theorem 10.7.3. The maximal distance between a pair of points CP1 π of with respect to the distance (10.7.1) is 2 , i.e. its Riemannian π diameter is 2 . π Indeed, the arccosine of a positive value does not exceed 2 . Since the unit sphere metric has Riemannian diameter π, which is twice the value that appears in the theorem, we obtain the following corollary. Corollary 10.7.4. Complex projective line CP1 with the Fubini– 1 R3 Study metric gF S is isometric to the sphere of radius 2 in of total area π and Gaussian curvature K =4. Proof. As calculated in Theorem 7.10.2, the Gaussian curvature K of the Fubini–Study metric gF S is K = 4 at every point. The curvature can be calculated directly using the formula K = ∆LB log f where f 1 − is the conformal factor f = 1+r2 as in (10.7.2). Thus we obtain a metric on the 2-sphere CP1 of constant Gaussian curvature 4. By a general 10.8. FUBINI–STUDY METRIC ON COMPLEX PROJECTIVE SPACE 131 result in Riemannian geometry, such a metric is isometric to a concen- tric sphere in R3. The normalisation K = 4 forces the radius of the 1 sphere to be 2 . Hence the total area is π. Alternatively, this results from the Gauss–Bonnet theorem. 

10.8. Fubini–Study metric on complex projective space The Fubini–Study metric on the complex projective line was re- viewed in Section 10.7. The complex projective space CPn also car- ries a natural metric called the Fubini–Study metric. The distance function d([v], [w]), associated with the Fubini–Study metric, takes its simplest form in homogeneous coordinates [Z0,...,Zn].

Definition 10.8.1. In homogeneous coordinates [Z0,...,Zn], given n+1 H(v,w) C | | a pair of vectors v,w , we have cos d([v], [w]) = v w where ∈ | | | | n 7 H(v,w)= j=0 vjwj .

Definition P 10.8.2 . In an affine neighborhood defined by the con- dition Z =0 , the affine coordinates are z = Z /Z = x + iy . { 0 6 } j j 0 j j With respect to these coordinates, we consider the Hermitian prod- n uct Ha(v,w) = j=1 vjwj . Then the Fubini–Study metric takes the following form.8 P

Theorem 10.8.3. Let Cn CPn be an affine neighborhood. The ⊆ Cn Fubini–Study metric gF S(X,X) at a point u is given by the for- mula ∈ H (X,X)(1+ u 2) H (X,u) 2 g (X,X)= a | | −| a | . (10.8.1) F S 1+ u 2 | | CPn Definition 10.8.4. The Fubini–Study 2-form αF S on is the form αF S(v,w) = gF S(Jv,w) where J is the complex structure, i.e., multiplication by i = √ 1. −

7Lawson [13, p. 50] gives the following formula the the metric in homogeneous 2 2 |ZΛdZ| 2 2 2 2 coordinates: ds = 4 4 where ZΛdZ = Z dZ Z,dZ . 0 |Z| | | | | | | − |h i| 8For general culture we note the following result concerning sectional curvature, which is a generalisation of Gaussian curvature; see [Car92]. In the case n = 1, we have H(X,u) = X u and the formula reduces to (10.7.2). Unlike the 1- dimensional| case,| sectional| | | curvature| is not constant when n 2: theq sectional curvature K of the Fubini–Study metric on CPn satisfies 1 K≥ 4 at every point and in every 2-dimensional direction. ≤ ≤ 132 10. GROMOV’S STABLE SYSTOLIC INEQUALITY FOR CPn

One can write down an explicit formula for αF S, as well. Namely one can write down αF S in an affine neighborhood as follows:

dzj dzj ( zjdzj) zjdzj α = i j ∧ ∧ F S 1+ r2 − (1 + r2)2 P P P ! related to (10.8.1). See [GriH78].

Proposition 10.8.5. The form αF S has the following properties. (1) It is a globally defined closed 2-form on CPn and has unit co- mass at every point. 2 CPn R (2) Its cohomology class [αF S] spans the line HdR( ) . (3) The class 1 [α ] is a generator of L2 (CPn) Z. ≃ π F S dR ≃ 10.9. Gromov’s inequality for complex projective space Definition 10.9.1. A fundamental cohomology class for a closed orientable manifold M of dimension d is a generator of the integer lattice Ld (M) Hd (M) where Ld (M) Z. dR ⊆ dR dR ≃ Recall that the cohomology ring of CPn is a truncated polynomial ring (see Theorem 8.8.4). We will use the following related result which will be useful for the proof of Gromov’s inequality. 2 CPn Theorem 10.9.2. If ω is a generator of LdR( ) then the cup n n power ω∪ is a fundamental cohomology class of CP . We are now ready to state Gromov’s stable systolic inequality for complex projective space. Theorem 10.9.3 (M. Gromov). Every Riemannian metric g on complex projective space CPn satisfies the optimal inequality stsys (g)n n!vol (g). 2 ≤ 2n The Fubini–Study metric on CPn satisfies the boundary case of equality. The proof of Gromov’s inequality is based on the following main ingredients: (1) the duality between the stable norm and the comass; see The- orem 10.6.1; (2) the generalized version of the Wirtinger inequality given in Proposition 6.11.3, so as to obtain an optimal constant in the inequality. To clarify the nature of the proof, we will state the following slight generalisation of Gromov’s inequality. A similar proof works for the more general theorem. 10.10. HOMOLOGY CLASS C AND COHOMOLOGY CLASS ω 133

Theorem 10.9.4. Let M be a 2n-dimensional Riemannian mani- fold, satisfying the following conditions: 2 R (1) b2(M)=1, i.e., HdR(M) ; ≃2 n 2n (2) For a nonzero class ω H (M), we have ω∪ =0 in H (M). ∈ dR 6 dR Then every Riemannian metric g on M satisfies the inequality stsys (g)n n!vol (g). 2 ≤ 2n The theorem will be proved in Section 10.11.

10.10. Homology class C and cohomology class ω To prove Gromov’s theorem we proceed as follows; cf. [9, Theo- rem 4.36]. We let C H (CPn; Z)) Z (10.10.1) ∈ 2 ≃ be the generator in homology represented by the inclusion of the 2- sphere CP1 CPn with its natural orientation stemming from the complex structure.⊆ Let L2 (CPn) Z dR ≃ 2 CPn be the integer lattice in de Rham cohomology HdR( ). Corollary 10.3.6, 2 CPn Z the lattices LdR and H2( ; ) are dual. Definition 10.10.1. Let ω L2 (CPn) be the generator in coho- ∈ dR mology dual to the homology class C, so that C ω = 1, where integra- tion is understood in the sense of a representative cycle in the class C and representative 2-form in the class ω. R

αFS Here ω is represented by the form π where αF S is the Fubini– α Study form: ω = FS H2 (CPn). Recall that wedge product π ∈ dR ∧ i in Ω∗(M) descendsh to thei cup product in HdR∗ (M)= iHdR(M); see Section 8.7. ∪ ⊕ Recall that the cup power ωn of ω is a generator of the integer 2n CPn Z CPn lattice LdR( ) , i.e., a fundamental cohomology class for ; see Theorem 10.9.2.≃ We therefore obtain the following corollary. 2 CPn Corollary 10.10.2. Let ω be a generator of the integer lattice LdR( ) and let η ω be a closed differential 2-form. Then ∈ n η∧ =1. (10.10.2) CPn Z The proof of Gromov’s inequality will be completed in Sections 10.11 and 10.12. 134 10. GROMOV’S STABLE SYSTOLIC INEQUALITY FOR CPn

10.11. Comass and application of Wirtinger inequality Continuing with the proof of Gromov’s inequality, let g be an ar- bitrary metric on the 2n-dimensional manifold M (e.g, on CPn). The metric defines a norm on each tangent and cotangent space, as well as all the exterior powers. We can then calculate the pointwise co- mass ηp of η at a point p M. Recall from Definition 10.5.1 that the thek comassk η of a differential∈ k-form η Ω(M) is the supremum of the pointwisek comassk∞ norms of Definition 6.6.2.∈ Theorem 10.11.1. Let η be a differential 2-form. Then we have n the following inequality for the comass of the form η∧ : n n η∧ n!( η ) . k k∞ ≤ k k∞ Proof. By the Wirtinger inequality and Corollary 6.11.6, we ob- tain the following bound for the norm at p: n n n ηp∧ n!( ηp ) n!( η ) , (10.11.1) k k≤ k k ≤ k k∞ where is the comass norm on differential forms. Taking the supre- mum ofk k (10.11.1)∞ over all p M we obtain the result.  ∈ 10.12. Proof of Gromov’s inequality In this section, we will complete the proof of Theorem 10.9.4, namely the inequality stsys (g)n n!vol (g). 2 ≤ 2n Step 1. Let (M, g) be a Riemannian manifold as in the theorem. We consider the following data: 2 (1) the integer lattice LdR(M) in de Rham cohomology; 2 Z (2) the generator ω LdR(M) ; (3) a representative∈ closed differential≃ 2-form η ω, η Ω2(M); ∈ ∈ (4) the volume form d volM of M. n n Thus at every point p M, we have, up to sign, ηp∧ = ηp∧ d volM . By Theorem 10.9.2, we∈ have k k

n 1 η∧ ≤ ZM

n η∧ d volM ≤ M k k Z n n!( η ) vol2n(M, g). ≤ k k∞ The last step follows by Wirtinger’s inequality as in (10.11.1). We therefore obtain n ( η ω) 1 n!( η ) vol2n(M, g). (10.12.1) ∀ ∈ ≤ k k∞ 10.12. PROOF OF GROMOV’S INEQUALITY 135

2 This is true for every form η representing the generator ω LdR(M) 2 ∈ ⊆ HdR(M). Step 2. The infimum of the right-hand side in (10.12.1) over all η ∈ ω gives the following bound for the comass ω ∗ of the cohomology class ω: k k n 1 n!( ω ∗) vol (M, g) , (10.12.2) ≤ k k 2n where ∗ is the comass norm in cohomology. k k Step 3. Denote by the stable norm in homology. Consider k k the integer lattice L2(M) = H2(M; Z)/T (in the case of the complex projective space the torsion subgroup is trivial and therefore can be left out of the formula). Recall that the normed lattices (L2(M), ) 2 k k and (LdR(M), ∗) are dual to each other; see Theorem 10.6.1. Since b1(M)= k k 2 1, we have λ1(L2)λ1(LdR) = 1. It follows that the class C L2(M) of (10.10.1) satisfies ∈ 1 C = . (10.12.3) k k ω k k∗ Step 4. Using identity (10.12.3), we obtain stsys (g)n = C n n!vol (g), (10.12.4) 2 k k ≤ 2n completing the proof of Gromov’s inequality. Step 5. When M = CPn, equality is attained by the two-point homogeneous Fubini–Study metric. This is because of the following three equalities. (1) The standard CP1 CPn is calibrated by the Fubini–Study ⊆ αFS Kahler 2-form αF S, i.e., CP1 αF S = π, or CP1 π = 1. α n FS ∧ (2) CPn π = 1. R R (3) We have equality for α in the Wirtinger inequality at every R  F S point. In the special case n = 2 we obtain the following inequality analo- gous to Pu’s; see Section 10.1. Corollary 10.12.1. Every metric g on the complex projective plane satisfies the optimal inequality stsys (CP2, g)2 2vol (CP2, g). (10.12.5) 2 ≤ 4

CHAPTER 11

Loewner’s inequality

In this chapter we prove Loewner’s torus inequality. In Chapter 12 we will prove Pu’s inequality for the real projective plane. A different proof of Pu’s inequality via quaternions appears in Chapter 13.

11.1. Eisenstein integers, We first review some arithmetic preliminaries. The Eisenstein inte- gers were already mentioned in Example 6.4.4. 2 Definition 11.1.1. The lattice LE R = C of the Eisenstein in- tegers (also known as the hexagonal lattice)⊆ is the lattice in C spanned by the elements 1 and the sixth root of unity. Remark 11.1.2. To visualize the lattice, start with an equilateral C 1 √3 1 triangle in with vertices 0, 1, and 2 + i 2 , and construct a tiling of the plane by repeatedly reflecting in all sides of the triangle. The Eisenstein integers are by definition the set of vertices of the resulting tiling.

Definition 11.1.3. Let b N. The Hermite constant γb is defined in one of the following two equivalent∈ ways:

(1) γb is the square of the biggest λ1(L) among all lattices L such that vol(Rb /L)=1; (2) γb is defined by the formula

λ1(L) b √γb = sup : L (R , ) , (11.1.1) vol(Rb/L)1/b ⊆ | |   where the supremum is over all lattices L Rb with a Eu- clidean norm . ⊆ | | Remark 11.1.4 (Relation to packing). A lattice realizing the supre- mum in (11.1.1) may be thought of as the one realizing the densest 2 Rb 1 packing in when we place balls of radius 2 λ1(L) at the points of L.

1ritzuf 2ariza hachi tzfufa 137 138 11. LOEWNER’S INEQUALITY

For b = 2, we will show that the lattice of Eisenstein integers realizes the supremum in (11.1.1).

11.2. Standard fundamental domain Some of the material in this section already appeared in Section 3.11.

Definition 11.2.1. The standard fundamental domain D0 C is the domain ⊆ D = z C: z 1, 1 Re(z) < 1 , Im(z) > 0 (11.2.1) 0 ∈ | |≥ − 2 ≤ 2 It is a fundamental domain for the action of the group PSL(2, Z) in the upperhalf plane of C. Definition 11.2.2. Two lattices in C are called similar if they differ by a multiplicative scalar. Lemma 11.2.3. Given a lattice L˜ C, one can choose a Z-basis 1,τ for a similar lattice L in such a way⊆ that τ D . { } ∈ 0 Proof. Consider a lattice L˜ C = R2. ⊆ Step 1. Choose a “shortest” vector z L˜, i.e. we have z = λ1(L˜). 1 ∈ | | We replace L˜ by the lattice L = z− L˜. Then the complex number +1 ∈ C is a shortest element in the new lattice L, and λ1(L)=1. Step 2. We complete the element +1 L to a Z-basis ∈ 1,τ ′ (11.2.2) { } for L. By replacing τ ′ by τ ′ if necessary, we can ensure the condition of positive imaginary part.−

Step 3. The real part Re(τ ′) of τ ′ does not necessarily satisfy the condition defining the domain D0. Next, we adjust the basis by adding a suitable integer to τ ′, i.e., replacing it by τ ′ + n, so that the result τ L satisfies the condition 1 Re(τ) < 1 . ∈ − 2 ≤ 2 Step 4. Since τ L, we have τ λ1(L) = 1. Hence τ lies in D0 of Definition 11.2.1. ∈ | |≥  The following result is an essential part of the proof of Loewner’s torus inequality with isosystolic defect. Proposition 11.2.4. When b =2, we have the following value for the Hermite constant: γ = 2 = 1.1547 .... The corresponding best 2 √3 lattice is similar to the lattice of the Eisenstein integers. 11.3. LOEWNER’S TORUS INEQUALITY 139

Proof. Multiplying L by nonzero complex numbers does not change 2 λ1(L) the value of the scale-invariant quotient area(C/L) . Thus we may as- sume that λ (L) = 1, and moreover 1 L. We choose τ as in 1 ∈ Lemma 11.2.3. The imaginary part of τ satisfies Im(τ) √3 , with 2 2 i π ≥ equality only when τ = e 3 D0. The area of the parallelogram in C ∈ √3 spanned by and +1 and τ is the imaginary part Im(τ) 2 , proving the lemma. ≥ 

11.3. Loewner’s torus inequality Loewner’s torus inequality relates the total area to the systole, i.e., least length of a noncontractible loop on the torus (T2, g). Loewner’s inequality first appeared in [Pu52].

Theorem 11.3.1 (Loewner’s torus inequality). Every metric g on the torus satisfies

area(g) √3 sys (g)2 0. (11.3.1) − 2 1 ≥ Theorem 11.3.2. The boundary case of equality in (11.3.1) is at- tained if and only if the metric is similar to the flat metric obtained as the quotient of R2 by the lattice formed by the Eisenstein integers.

Recall the following. (1) The 1-systole of a Riemannian manifold M is the least length of a noncontractible loop on M; see Section 9.10. (2) The fundamental group of the torus T2 is abelian. (3) Due to the multiplicavity of the volume of 1-homology classes T2 T2 Z for orientable surfaces, we have sys1( , g)= λ1 H1 ( ; ), , where is the stable norm of the metric g. k k k k  In this sense, Gromov’s inequality (10.12.5) for the complex projective plane is a higher-dimensional analogue of Loewner’s inequality. We pursue an analogy with the isoperimetric inequality mentioned in footnote 3 in Section 10.1. Bonnesen’s inequality asserts the follow- ing strengthening of the isoperimetric inequality.

Theorem 11.3.3 (Bonnesen’s inequality). Consider a Jordan curve of length L in the plane. Let A be the area of the region bounded by the curve. Let R be the circumradius of the bounded region, and let r be its inradius. Then L2 4πA π2(R r)2. (11.3.2) − ≥ − 140 11. LOEWNER’S INEQUALITY

The inequality appeared in [Bo21]. See [BuraZ80, p. 3] for a proof. The error term π2(R r)2 on the right hand side of (11.3.2) is traditionally called the isoperimetric− defect term.3 Loewner’s torus inequality (11.3.1) can be strengthened by intro- ducing a “systolic defect” term `ala Bonnesen. To express such an improvement of Loewner’s inequality, we need to review the conformal representation theorem (uniformisation theorem) already discussed in in Section 3.8. Theorem 11.3.4 (Conformal representation theorem). Every met- ric g on the torus T2 is isometric to a metric of the form f 2(x,y)(dx2 + dy2), (11.3.3) 2 2 2 with respect to a unit area flat metric g0 = dx +dy on the torus R /L, where L R2 is a lattice. ⊆ See (3.8.2) for more details. The defect term in the strengthened Loewner inequality is the variance4 of the conformal factor f in (11.3.3), as in Theorem 11.4.3 below.

11.4. Expected value and variance 2 2 Consider a flat torus (T , g0)= R /L where L is a lattice and g0 = 2 2 2 dx + dy , such that area(R /L, g0) = 1. The torus can be thought of as the quotient of the plane by for the action of L on R2 by translations. Definition 11.4.1. An open domain D R2 is called an open fundamental domain for the torus if D maps injectively⊆ to T2 and the closure D¯ maps surjectively to T2. Such fundamental domains will be used in the proof of Loewner’s torus inequality with remainder term in Section 11.7. Definition 11.4.2. Let D be an open fundamental domain for the torus, and f a function on the torus. The mean, or expected value,5 of f is the quantity

m = E(f)= f dx dy T2 ∧ Z (11.4.1) = f(x,y) dx dy. ∧ ZD where dx dy is the standard area form of R2. ∧ 3she’erit izoperimetrit 4shonut 5tochelet 11.5. APPLICATION OF COMPUTATIONAL FORMULA FOR VARIANCE 141

We have the following strengthening of Loewner’s torus inequality. Theorem 11.4.3 ([Horowitz et al.]). Every metric g on the torus satisfies the inequality area(g) √3 sys(g)2 Var(f). (11.4.2) − 2 ≥ Here the remainder term on the right-hand side is the variance of the conformal factor f of the metric g = f 2(dx2+dy2) on the torus, relative 2 2 to the mesure induced by the unit area flat metric g0 = dx + dy in the conformal class of g.

11.5. Application of computational formula for variance The proof of inequalities with isosystolic defect relies upon the com- putational formula for the variance of a random variable6 in terms of expected values. Definition 11.5.1. Let µ be a probability measure on a space M, meaning that the total measure of M is 1. The computational formula for the variance of a random variable f on M is the formula E (f 2) (E (f))2 = Var(f). (11.5.1) µ − µ Here the variance is Var(f)= E (f m)2 , µ − where m = Eµ(f) is the expected value, i.e., the mean. In our differential geometric application, the random variable f is the conformal factor of the metric on the torus.

2 Definition 11.5.2. Given a lattice L of unit coarea, let g0 = dx + dy2 be a flat metric of unit area on the 2-torus T2 = R2/L. Denote the associated measure on T2 by µ, so that µ(T2) = 1. The measure coincides with the usual area for the kind of domains we are interested in. In other words, for every domain A T2, we ⊆ have µ(A)= A dx dy 0. Since µ is a probability measure, we can apply the computational∧ ≥ formula for the variance, (11.5.1), to µ. R Remark 11.5.3. Here f can be thought of either as a function on the torus or as a doubly periodic function on R2 (i.e., periodic with respect to translations by elements of the lattice L).

6mishtaneh mikri 142 11. LOEWNER’S INEQUALITY

2 Lemma 11.5.4. Consider a metric g = f g0 on the torus conformal to the flat metric g0 of unit area, where f > 0 is the conformal factor. Then we have

2 2 Eµ(f )= f dx dy = area(g). T2 ∧ Z Proof. Indeed, f 2dx dy is the area 2-form of the metric g.  ∧ Equation (11.5.1) therefore becomes area(g) (E (f))2 = Var(f). (11.5.2) − µ In Section 11.6, we will relate the expected value Eµ(f) to the systole of the metric g. Then we will then relate formula (11.5.2) to Loewner’s torus inequality in Section 11.7.

11.6. Flat torus as pencil of parallel geodesics 2 2 2 2 Consider the torus T = R /L˜ with the metric g0 = dx + dy of unit area. By Lemma 11.2.4, the lattice L˜ of deck transformations of the flat torus (T2, g ) admits a Z-basis similar to the basis (1,τ) C, 0 ⊆ where τ D0 as in formula (11.2.1). In other words, L˜ is similar to the lattice∈ L = Z 1+ Z τ C. ⊆ Definition 11.6.1. We set σ = Im τ L > 0, where Im(τ) is the imaginary part of the conformal parameterq τ C.  ∈ Lemma 11.6.2. We have σ2 √3 , with equality if and only if the ≥ 2 conformal parameter τ is the primitive cube root of unity 1 + i √3 . − 2 2 Proof. This is immediate from the geometry of the fundamental domain.  Lemma 11.6.3. The basis for the lattice L˜ (group of deck tranfor- mations of g0) can be taken to be 1 1 (σ− ,σ− τ), 1 where Im(σ− τ)= σ.

Proof. This is immediate from the fact that g0 is of unit area.  2 Corollary 11.6.4. The flat torus (T , g0) of unit area admits a fundamental domain B with the following properties: (1) B is a rectangle in coordinates x,y; 1 (2) B has base σ− and height σ; 11.7. PROOF OF LOEWNER’S TORUS INEQUALITY 143

(3) B is a ruled surface 7 by a pencil 8 of horizontal closed geodesics, denoted γy = γy(x); 1 (4) each geodesic γy is of length lengthg0 (γy)= σ− ; (5) the width of the pencil (i.e., height of B) is σ, i.e., y [0,σ]; ∈ (6) in the torus we have γσ = γ0. 11.7. Proof of Loewner’s torus inequality 2 Lemma 11.7.1. For the metric g = f g0 on the torus, we have the following lower bound for the lengths of closed geodesics γy: y [0,σ], length (γ ) sys (g). (11.7.1) ∀ ∈ g y ≥ 1 Proof. This is immediate from the fact that each of these closed geodesics is noncontractible9 in T2. 

We analyze the expected value term Eµ(f) = T2 fdx dy in for- mula (11.5.2). We will use Lemma 11.7.1 in the proof of the∧ following proposition. R Proposition 11.7.2. Let σ = Im τ(L). Then E (f) σ sys (g). µ ≥ 1 Proof. Let B be the fundamentalp domain of Corollary 11.6.4. By Fubini’s theorem, we pass to the iterated integral10 to obtain the fol- lowing lower bound for the expected value:

E (f)= f(x,y)dx dy µ ∧ ZB σ = f(x,y)dx dy Z0 Zγy ! σ = lengthg(γy)dy Z0 σ sys (g), ≥ 1 by inequality (11.7.1) of Lemma 11.7.1.  Loewner’s torus inequality now follows from the following more gen- eral result. Theorem 11.7.3. Every metric g in the conformal class with con- formal parameter τ D satisfies ∈ 0 area(g) Im(τ)sys (g)2 Var(f). (11.7.2) − 1 ≥ 7a ruled surface is mishtach yesharim 8aluma with alef 9lo-kvitza, with kaf 10Integral nishneh 144 11. LOEWNER’S INEQUALITY where f is the conformal factor with respect to the flat metric g0, 2 i.e., g = f g0. Proof. Recall that σ = Im(τ). By Proposition 11.7.2 we have an inequality p E (f) σ sys (g). µ ≥ 1 Substituting this inequality into the formula (11.5.2), we obtain the inequality area(g) σ2 sys (g)2 Var(f), proving Theorem 11.7.3.  − 1 ≥ 2 √3 Now by Lemma 11.6.2, we have σ 2 . Therefore we obtain in particular Loewner’s torus inequality with≥ isosystolic defect. Corollary 11.7.4. Every metric g on the torus satisfies area(g) √3 sys (g)2 Var(f). (11.7.3) − 2 1 ≥ We connect this discussion to the standard terminology of the ca- pacity of a cylinder (annulus). Definition 11.7.5. Consider the cylinder C (i.e., annulus) ob- tained by cutting the torus along the loop corresponding to the gen- erator 1 of the pair (1,τ). The capacity cap(C) of the cylinder is the inverse of the quantity Im(τ).

2 sys1 In this terminology, inequality (11.7.2) implies area cap(C) where “cap” is the capacity. Thus we obtain the following corollar≤ y. Corollary 11.7.6. Let R/Z be the circle of unit length. Suppose a surface M contains an annulus11 A M conformal to the cylin- R Z 2 1 ⊆ R Z der / (0,σ ) with capacity σ2 . Suppose the circle / 0 M is noncontractible.× Then every metric satisfies the inequality×{area(}⊆M) σ2 sys2(M), or equivalently sys2(M) cap(C) area(M). ≥ ≤ Proof. This results by applying the Cauchy–Schwarz inequality to the conformal factor f where the metric on the annulus is f 2(dx2 +dy2) where 0

11.8. Boundary case of equality in Loewner’s inequality Corollary 11.8.1. A metric satisfying the boundary case of equal- ity in Loewner’s torus inequality (11.3.1) is necessarily flat and similar to the quotient of R2 by the lattice of Eisenstein integers.

11tabaat (tet, bet, ayin, tav) 11.9.RECTANGULARLATTICES 145

Proof. If a metric f 2(dx2 + dy2) satisfies the boundary case of equality in (11.3.1), then the variance of the conformal factor f must vanish by (11.7.3). Hence f is a constant function. The proof is com- pleted by applying Lemma 11.2.4 on the Hermite constant in dimen- sion 2, taking into account the fact that the lattice LE of Eisenstein √3 integers is the only one satisfying the equality Im(τ)= 2 . 

11.9. Rectangular lattices Now suppose τ is pure imaginary, i.e. the lattice L is a rectangular lattice of coarea 1. Corollary 11.9.1. If τ is pure imaginary, then the metric g = 2 f g0 satisfies the inequality area(g) sys(g)2 Var(f). (11.9.1) − ≥ Proof. If τ is pure imaginary then σ = Im(τ) 1, and the inequality follows from (11.7.2). ≥  p Corollary 11.9.2. Every torus of revolution satisfies the inequal- ity (11.9.1). Proof. This is immediate from the fact that its lattice is rectan- gular by Corollary 3.10.2. 

11.9.1. Surface tension and shape of a water drop. This material is optional. We first make the following introductory remarks. Perhaps the most familiar physical manifestation of the 3-dimensional isoperimetric inequality is the shape of a drop of water. Namely, a drop will typically assume a symmetric round shape. Since the amount of water in a drop is fixed, surface tension forces the drop into a shape which minimizes the surface area of the drop, namely a round sphere. Thus the round shape of the drop is a consequence of the phenomenon of surface tension (metach panim). Mathematically, this phenomenon is expressed by the isoperimetric inequality in the plane. The solution to the isoperimetric problem in the plane is usually expressed in the form of an inequality that relates the length L of a closed curve and the area A of the planar region that it encloses. The isoperimetric inequality states that 4πA L2, and that the equality holds if and ≤ only if the curve is a round circle. The inequality is an upper bound for area in terms of length. It can be rewritten as follows: L2 4πA 0. − ≥ Recall the notion of central symmetry: a Euclidean polyhedron is called centrally symmetric if it is invariant under the “antipodal” map x x. Thus, in the plane central symmetry is 7→ − the rotation by 180 degrees. For example, an ellipse is centrally symmetric, as is any ellipsoid in 3-space.

11.9.2. Centrally symmetric polyhedron in 3-space. This material is optional. There is a geometric inequality that is in a sense dual to the isoperimetric inequality in the following sense. Both involve a length and an area. The isoperimetric inequality is an upper bound for area in terms of length. There is a geometric inequality which provides an upper bound for a certain length in terms of area. More precisely it can be described as follows. Any centrally symmetric convex body of surface area A can be squeezed through a noose of length √πA, with the tightest fit achieved by a sphere. This property is equivalent to a special case of Pu’s inequality (see below), one of the earliest systolic inequalities. 146 11. LOEWNER’S INEQUALITY

Example 11.9.3. An ellipsoid is an example of a convex centrally symmetric body in 3-space. It may be helpful to the reader to develop an intuition for the property mentioned above in the context of thinking about ellipsoidal examples. An alternative formulation is as follows. Every convex centrally symmetric body P in R3 admits a pair of opposite (antipodal) points and a path of length L joining them and lying on the boundary ∂P of P , satisfying L2 π area(∂P ). ≤ 4 11.9.3. Notion of systole. This material is optional and is a review of Section 9.10. The systole of a compact metric space X is a metric invariant of X, defined to be the least length of a noncontractible loop in X, denoted sys(X). When X is a graph, the invariant is usually referred to as the girth, ever since the 1947 article by W. Tutte [Tu47]. Possibly inspired by Tutte’s article, Loewner started thinking about systolic questions on surfaces in the late 1940s, resulting in a 1950 thesis by his student P.M. Pu [Pu52]. The actual term “systole” itself was not coined until a quarter century later, by Marcel Berger. This line of research was, apparently, given further impetus by a remark of Ren´eThom, in a conversation with Berger in the library of Strasbourg University during the 1961-62 academic year, shortly after the publication of the papers of R. Accola and C. Blatter. Referring to these systolic inequalities, Thom reportedly exclaimed: “Mais c’est fondamental!” [These results are of fundamental importance!] Subsequently, Berger popularized the subject in a series of articles and books, most recently in the march ’08 issue of the Notices of the American Mathematical Society [Ber08]. is a rapidly developing field, featuring a number of recent publications in leading journals. Recently, an intriguing link has emerged with the Lusternik- Schnirelmann category. The existence of such a link can be thought of as a theorem in “systolic topology”. In projective geometry, the real projective plane RP2 is defined as the collection of lines through the origin in R3. The distance function on RP2 is most readily understood from this point of view. Namely, the distance between two lines through the origin is by definition the angle between them, or more precisely the lesser of the two angles. This distance function corresponds to the metric of constant Gaussian curvature +1. Alternatively, RP2 can be defined as the surface obtained by identifying each pair of antipodal points on the 2-sphere. Other metrics on RP2 can be obtained by quotienting metrics on S2 embedded in 3-space in a centrally symmetric way, see the discussion in Section 11.9.2. Topologically, RP2 can be obtained from the M¨obius strip by attaching a disk along the boundary. Among closed surfaces, the real projective plane is the simplest non-orientable such surface. CHAPTER 12

Pu’s inequality and generalisations

The material in this chapter is joint work with Tahl Nowik [7]. A different proof of Pu’s inequality via quaternions appears in Chapter 13.

12.1. Statement of Pu’s inequality Pu’s inequality applies to arbitrary Riemannian metrics on the real projective plane RP2. A student of Charles Loewner’s, P.M. Pu proved it in a 1950 thesis (published in 1952 as [Pu52]). Pu’s inequality is analogous to Loewner’s torus inequality of Section 11.3. Theorem 12.1.1 (Pu). In the real projective plane, we have the inequality π sys (g)2 area(g) 1 ≤ 2 for every metric g on RP2. The case of equality is attained precisely by the metrics of constant Gaussian curvature on RP2. There is a vast generalisation of the inequalities of Pu and Loewner, due to M. Gromov, called Gromov’s systolic inequality for essential manifolds (different from Gromov’s inequality for CPn as in Theo- rem 10.9.3). This result involves a topological notion of essential man- ifold as in Section 14.1.1

1Loewner’s systolic inequality for the torus and Pu’s inequality for the real pro- jective plane were historically the first results in systolic geometry. Great stimulus was provided in 1983 by Gromov’s paper [5], and later by his book [9]. Our goal is to prove a strengthened version with a remainder term of Pu’s systolic inequal- ity sys2(g) π area(g) (for an arbitrary metric g on RP2), analogous to Bonnesen’s ≤ 2 inequality L2 4πA π2(R r)2, where L is the length of a Jordan curve in the plane, A is the− area of≥ the region− bounded by the curve, R is the circumradius and r is the inradius. Note that both the original proof in Pu ([Pu52], 1952) and the one given by Berger ([2], 1965, pp. 299–305) proceed by averaging the metric and showing that the averaging process decreases the area and increases the systole. Such an approach involves a 5-dimensional integration (instead of a 3-dimensional one given here) and makes it harder to obtain an explicit expression for a remain- der term. Analogous results for the torus were obtained in [Horowitz et al.] with generalisations in [1], [6], [8]. 147 148 12. PU’S INEQUALITY AND GENERALISATIONS

12.2. Tangent map To introduce the notion of a Riemannian submersion2 in Section 12.3, we will need the notion of the tangent map of a smooth map φ between manifolds. Proposition 12.2.1. Consider a smooth map φ: M N between differentiable manifolds. Then φ defines a natural map → dφ: TM TN → called the tangent map. Proof. We represent a tangent vector v T M by a path ∈ p c(t): I M, → such that c′(0) = v. The composite map σ = φ c: I N is a path ◦ → in N. Then the vector σ′(0) is the image of v under dφ:

dφ(v)= σ′(0), proving the proposition. 

12.3. Riemannian submersions In this section we will define the notion of a Riemannian submer- sion. Consider a map φ: M N between closed manifolds. We assume that dφ is onto. Then by the→ implicit function theorem, the fibers are smooth submanifolds. Let F M be the fiber over a point p N. ⊆ ∈ Proposition 12.3.1. The kernel of the tangent map dφ: TM TN is the subspace T F TM at every point x F . → x ⊆ ∈ Proof. A vector tangent to the fiber can be represented by a path lying entirely in the fiber. Therefore its image in the base manifold is a constant map.  Definition 12.3.2. the vertical space3 in TM is the subspace

ker(dφx)= TxF. Definition 12.3.3. Given a Riemannian metric on M, the hori- zontal space H T M is the orthogonal complement of the vertical x ⊆ x space TxF in TxM.

2Hatzafah according to Amit Solomon at http://ma.huji.ac.il/~amit/ hebrew_geometric_dictionary_2.pdf 3merchav anchi 12.4.ASTIEFELMANIFOLD 149

Here “H” stands for horizontal. Thus we have an orthogonal de- composition

TxM = TxF + Hx. Definition 12.3.4. A map φ: M N between Riemannian man- ifolds is a Riemannian submersion if→ at every point x M, the re- ∈ stricted map dφ: Hx Tφ(x)N is an isometry, i.e., preserves the length of vectors. →

12.4. A Stiefel manifold To prove Pu’s inequality, we will exploit a special closed 3-dimensional manifold M R3 R3 called Stiefel manifold. ⊆ × Definition 12.4.1. The Stiefel manifold M is

M = (v,w) R3 R3 : v v =1, w w =1, v w =0 (12.4.1) { ∈ × · · · } where v w is the scalar product on R3. · Lemma 12.4.2. The Stiefel manifold M is diffeomorphic to the Lie group SO(3, R) of orthogonal three by three matrices of determinant 1.

Proof. We think of SO(3, R) is the space of matrices of unit deter- minant, with orthonormal column vectors. We define a diffeomorphism

φ: M SO(3, R), (v,w) (v w n), → 7→ where n = v w is the vector product on R3. The map ×φ is injective because it has a left inverse given by sending an orthogonal matrix (abc) to the pair of vectors (a,b). Given a matrix P SO(3) with first two columns v and w, the only possibilities for P are∈ (v w v w) and (v w v w). Only the first possibility has positive determinant.× Hence P =(− v× w v w)= φ(v,w). Therefore φ is onto. × 

Proposition 12.4.3. Given a point (v,w) M, the tangent space ∈ T(v,w)M is

T M = (X,Y ) R3 R3 : X v =0, Y w =0, X w + Y v =0 . (v,w) { ∈ × · · · · } Proof. The tangent space is identified by differentiating the three defining equations of M from (12.4.1) along a path through (v,w) with initial tangent vector (X,Y ).  150 12. PU’S INEQUALITY AND GENERALISATIONS

12.5. A metric on the Stiefel manifold

We define a Riemannian metric gM on M as follows. Given a point (v,w) M, let n = v w. We declare the basis (0,n), (n, 0), (w, v) ∈ × − of T(v,w)M to be orthonormal. This metric is a modification of the metric restricted to M from R3 R3 = R6. Namely, with respect to the Euclidean metric on R6 the above× three vectors are orthogonal and the first two have length 1. However, the third vector has Euclidean length √2, whereas we need to define its length to be 1. Thus the metric on M is defined as follows. Let (v,w) M, and let n = v w. Let A T(v,w)M be the span of (0,n) and (∈n, 0), and let B T× M be spanned⊆ by (w, v). We ⊆ (v,w) − view the Euclidean metric g on R6 as a quadratic form.

Definition 12.5.1. The metric gM on M is obtained from the Euclidean metric g by setting 1 g = g⇂ + g⇂ . (12.5.1) M A 2 B Theorem 12.5.2. The vectors (n, 0), (0,n) and (w, v) form an − orthonormal basis for T(v,w)M relative to the metric gM .

Proof. Our choice of the metric gM ensures that (w, v) is a unit vector. −  Definition 12.5.3. The natural projections p,q : M S2 are given by p(v,w)= v and q(v,w)= w. → Each of the projections exhibits M as a circle bundle over S2.

Lemma 12.5.4. The maps p and q on (M, gM ) are Riemannian submersions, where the metric on S2 is restricted from R3. Proof. For the projection p, given (v,w) M, the vector (0,n) 1 ∈ as defined above is tangent to the fiber p− (v). Hence dp(0,n)=0. The other two vectors, (n, 0) and (w, v), are thus an orthonormal − basis for the horizontal subspace (see Section 12.3) of T(v,w)M normal to the fiber, and are mapped by dp to the orthonormal basis n,w 2 of TvS .  1 Remark 12.5.5. The projection p maps the fiber q− (w) onto a great circle C S2. ⊆ This map preserves length since the unit vector (n, 0), tangent to 1 2 the fiber q− (w)at(v,w), is mapped by dp to the unit vector n TvS . The same comments apply when the roles of p and q are reversed.∈ 12.6. A DOUBLE INTEGRATION 151

12.6. A double integration In the following proposition, integration takes place respectively over great circles C S2, over the fibers in M, over S2, and over M. The integration is always⊆ with respect to the volume element of the given Riemannian metric. Theorem 12.6.1 (Fubini’s theorem). Given a Riemannian submer- sion M N, the integral over M can be computed by performing two successive→ integrations: (1) integrating over each fiber; (2) integrating the result of (1) over the base N. Since p and q are Riemannian submersions by Lemma 12.5.4, we can use Fubini’s Theorem to integrate over M by integrating first over the fibers of either p or q, and then over S2; cf. [3, Lemma 4]. 1 Remark 12.6.2. By the remarks above, if C = p(q− (w)) and a 2 function f : S R is continuous, then −1 f p = f. → q (w) ◦ C Proposition 12.6.3. Given a functionR f : S2 R+R, we define the least mean m R by setting → ∈ m = min f : C S2 a great circle . ⊆ ZC  Then m2 1 2 f f 2, (12.6.1) π ≤ 4π 2 ≤ 2  ZS  ZS where equality in the second inequality occurs if and only if f is con- stant. Proof. Step 1. We use the fact that the Stiefel manifold M is the total space of a pair of Riemannian submersions p and q. We apply Fubini’s theorem twice to obtain 1 f = f p 2 2 2π −1 ◦ ZS ZS  Zp (v)  1 = f p 2π ◦ ZM 1 = f p 2π 2 −1 ◦ ZS  Zq (w)  1 m =2m, ≥ 2π 2 ZS proving the first inequality of (12.6.1). 152 12. PU’S INEQUALITY AND GENERALISATIONS

Step 2. By the Cauchy–Schwarz inequality, we have 2 1 f 4π f 2, S2 · ≤ S2  Z  Z proving the second inequality of (12.6.1). Step 3. Equality occurs if and only if f and 1 are linearly depen- dent, i.e., if and only if f is constant. 

2 1 2 We define the quantity Vf by setting Vf = S2 f 4π S2 f . Then Proposition 12.6.3 can be restated as follows. − R R  Corollary 12.6.4. Let f : S2 R+ be continuous. Then → 2 2 m f Vf 0, 2 − π ≥ ≥ ZS and Vf =0 if and only if f is constant. Proof. The proof is obtained from Proposition 12.6.3 by noting that a b c if and only if c a c b 0.  ≤ ≤ − ≥ − ≥ 12.7. A probabilistic interpretation

We can assign a probabilistic meaning to the term Vf as follows. 2 Let gcan be the canonical metric of curvature K =1 on S . Definition 12.7.1. Let µ be the probability measure induced by 1 the metric 4π gcan. A function f : S2 R+ is then thought of as a random variable → 1 with expectation Eµ(f)= 4π S2 f. Its variance is thus given by 2 2 R2 1 2 1 1 Varµ(f)= E(f ) E(f) = f f = Vf . − 4π 2 − 4π 2 4π ZS  ZS  The variance of a random variable f is non-negative, and it vanishes if and only if f is constant. This reproves the corresponding properties of Vf established above via the Cauchy–Schwarz inequality.

12.8. From the sphere to the real projective plane

Now let g0 be the metric of constant Gaussian curvature K = 1 on RP2. The orientable double cover ρ: (S2,g ) (RP2,g ) can → 0 is a local isometry. Each projective line C RP2 is the image under ρ of a great circle of S2. ⊆ 12.9. A GENERALISATION OF PU’S INEQUALITY 153

Proposition 12.8.1. Given a function f : RP2 R+, we de- fine m¯ R by setting → ∈ m¯ = min f : C RP2 a projective line . ⊆ ZC  Then 2¯m2 1 2 f f 2, π ≤ 2π RP2 ≤ RP2  Z  Z where equality in the second inequality occurs if and only if f is con- stant. Proof. We apply Proposition 12.6.3 to the composition f ρ. Note that ◦ f ρ =2 f and f ρ =2 f. −1 ◦ 2 ◦ RP2 Zρ (C) ZC ZS Z The condition for f to be constant holds since f is constant if and only if f ρ is constant.  ◦ RP2 ¯ 2 1 2 1 For we define Vf = RP2 f 2π RP2 f = 2 Vf ρ. We obtain the following restatement of Proposition− 12.8.1. ◦ R R  Corollary 12.8.2. Let f : RP2 R+ be a continuous function. Then → 2 2 2¯m f V¯f 0, RP2 − π ≥ ≥ Z where V¯f =0 if and only if f is constant. 1 RP2 Relative to the probability measure induced by 2π g0 on , we 1 1 ¯ have E(f)= 2π RP2 f, and therefore Var(f)= 2π Vf , providing a prob- abilistic meaning for the quantity V¯ , as before. R f 12.9. A generalisation of Pu’s inequality 2 For a metric g0 of constant Gaussian curvature +1 on S , we denote by dAg0 the area element of the metric. Theorem 12.9.1. Let g be a Riemannian metric on RP2. Let L be the shortest length of a noncontractible loop in (RP2,g). Let f : RP2 + 2 → R be such that g = f g0. Then 2L2 area(g) 2π Var(f), − π ≥ where the variance is with respect to the probability measure induced 1 2L2 by 2π g0. Furthermore, equality area(g)= π holds if and only if f is constant. 154 12. PU’S INEQUALITY AND GENERALISATIONS

Proof. Step 1. By the uniformization theorem, every metric g 2 2 on RP is of the form g = f g0 where g0 is a metric of constant Gaussian curvature +1 (unique up to isometry), and the function f : RP2 R+ is continuous. → 2 2 Step 2. The area of g is RP f dAg0 . The g-length of a projective line C is f. Let L be the shortest length of a noncontractible loop. C R Then L m¯ wherem ¯ is defined in Proposition 12.8.1, since a projective line in RP≤R2 is a noncontractible loop. 2 Step 3. Corollary 12.8.2 implies area(RP2,g) 2L V¯ 0. − π ≥ f ≥ Step 4. To characterize the boundary case of equality in Pu’s RP2 2L2 ¯ inequality, note that if area( ,g) = π then Vf = 0, which implies that f is constant, by Corollary 12.8.2. Conversely, if f is a constant c, then the only geodesics are the projective lines, and therefore L = cπ. 2 2L 2 RP2 Hence π =2πc = area( ).  CHAPTER 13

Alternative proof of Pu’s inequality

The alternative proof of Pu’s inequality exploits certain properties of circle fibrations of the 3-sphere, and relies on the following ingredi- ents: (1) Hopf fibration; (2) quaternions; (3) geodesic flow1 of a Riemannian manifold; (4) a pair of orthogonal fibrations of the 3-sphere; (5) a suitable integral-geometric identity.

13.1. Hopf fibration h To prove Pu’s inequality, we need to study the Hopf fibration of Section 7.5 more closely. The circle action in Cn restricts to the unit 2n 1 n sphere S − C , which therefore admits a fixed-point-free circle action. Namely,⊆ the circle S1 = eiθ : θ R C { ∈ }⊆ acts on a point (z ,...,z ) Cn by 1 n ∈ iθ iθ iθ e .(z1,...,zn)=(e z1,...,e zn). (13.1.1) Lemma 13.1.1. The action is an isometry with respect to the natural Euclidean metric. Proof. In real coordinates, the matrix of this action by eiθ = cos θ + i sin θ looks as follows for n = 2: cos θ sin θ 0 0 sin θ −cos θ 0 0  0 0 cos θ sin θ −  0 0 sin θ cos θ    This is clearly an isometry.   2n 1 1 The quotient manifold (space of orbits) S − /S is the the complex n 1 2 projective space CP − . For n = 2 we get the 2-sphere S .

1Zrima geodesit 155 156 13. ALTERNATIVE PROOF OF PU’S INEQUALITY

Definition 13.1.2. The quotient map 2n 1 n 1 h: S − CP − (13.1.2) → is called the Hopf fibration. Definition 13.1.3. It will be convenient in the sequel to denote this data by a two-arrow diagram of the following type:

1 2n 1 h n 1 S S − CP − , (13.1.3) −→ −→ where the first arrow denotes the inclusion of a fiber, while the second arrow denotes the Hopf fibration h.

13.2. Hopf fibration is a Riemannian submersion We specialize to the case n = 2. Proposition 13.2.1. Let S3 be the unit 3-sphere. The Hopf fibra- tion h: S3 S2 is a Riemannian submersion, for which the natural metric on the→ base S2 is a metric of constant Gaussian curvature +4 1 and radius 2 . We note the following. 1 π (1) The maximal distance between a pair of S orbits is 2 rather than π (see next item); (2) a pair of antipodal points of S3 lies in a common orbit and therefore descends to the same point of the quotient space S2; (3) a pair of points in S2 at maximal distance is defined by the orbits of a pair v,w S3 such that w is orthogonal to Cv; (4) for such a pair we have∈ H(v,w) = 0 where H is the Hermitian inner product in C2.

13.3. Hamilton quaternions Definition 13.3.1. The algebra H of the Hamilton quaternions is the real 4-dimensional vector space with real basis (1,i,j,k), so that H = R 1+ R i + R j + R k equipped with an associative,2 distributive, and non-commutative prod- uct operation. This operation has the following properties: (1) the center of H is R 1; (2) the operation satisfies the relations i2 = j2 = k2 = 1; (3) also the relations ij = ji = k, jk = kj = i, ki =− ik = j. − − − 2chok kibutz 13.4. COMPLEX STRUCTURES ON THE ALGEBRA H 157

The relation of item (3) to the vector product in R3 is mentioned in Proposition 13.4.2. Lemma 13.3.2. The algebra H has a natural structure of a complex vector space C2 via the identification R 1+ R i C. ≃ Proof. The complementary subspace R j +R k of R 1+R i can be thought of as follows: R j + R ij =(R 1+ R i)j using associativity and distributivity. We therefore obtain a natural isomorphism H C1+ Cj, ≃ showing that the pair of elements (1, j) is a complex basis for H.  Theorem 13.3.3. Nonzero quaternions form a group under quater- nionic multiplication. Proof. Given q = a+bi+cj +dk H, let N(q)= a2 +b2 +c2 +d2, andq ¯ = a bi cj dk. Then one checks∈ that qq¯ = N(q). Therefore q has a multiplicative− − − inverse

1 1 q− = q,¯ N(q) proving the theorem. 

13.4. Complex structures on the algebra H

Definition 13.4.1. A pure imaginary Hamilton quaternion q H0 in H = R 1+ R i + R j + R k, is a real linear combination of i, j, and∈ k. Thus H H is a real 3-dimensional subspace. 0 ⊆ Proposition 13.4.2. With respect to the natural Euclidean met- ric on H0, quaternion multiplication of a pair of orthogonal elements coincides with the vector product on R3. Proof. This is easily checked with respect to a standard basis such as i,j,k.  1 More generally for pure quaternions p,q one has p q = 2 (pq qp) (see https://en.wikipedia.org/wiki/Quaternion#Quaternion× −s_and_ the_space_geometry for a more detailed discussion.) Corollary 13.4.3. Right or left multiplication by a unit quater- nion q is an isometry of H equipped with the standard inner product of R4. 158 13. ALTERNATIVE PROOF OF PU’S INEQUALITY

Proof. The proof is analogous to that of Lemma 13.1.1, relying on Proposition 13.4.2.  This can also be used using N(qr) = N(q)N(r) which however also requires proof. Corollary 13.4.3 will be used in the proof of the following result, generalizing Lemma 13.3.2 on the natural complex structure on H. Proposition 13.4.4. A choice of a pure imaginary quaternion q ∈ H0 specifies a natural complex structure H C C = q + q⊥, C R R C H where q = 1+ q and q⊥ is its orthogonal complement in . Proof. Let q = bi + cj + dk with b,c,d R. We can assume without loss of generality that q = 1. ∈ | | Step 1. By anticommutation among i,j,k, the cross-terms cancel out and we obtain q2 = b2i2 + c2j2 + d2k2 + bc(ij + ji)+ bd(ik + ki)+ cd(jk + kj) = b2i2 + c2j2 + d2k2 = 1. − 2 Since q = 1, the subspace Cq = R 1+R q with the product operation restricted from− H is naturally isomorphic to the field C. Step 2. Let us show that the orthogonal complement of the sub- 4 space Cq H = R has a natural structure of a complex line for multiplication⊆ by the “imaginary unit” q. Let r H be a unit-norm quaternion orthogonal to C H. We ∈ 0 q ⊆ need to show that the quaternion qr is also orthogonal to Cq. We 1 multiply both sides of the dot product qr 1 by q− = q. By Corol- lary 13.4.3, we obtain · − 1 qr 1= r (q− 1) = r q =0 · · − · by hypothesis on r, and similarly qr q = r 1 = 0. Thus qr is orthogonal · · to Cq. Hence the orthogonal complement of Cq is the subspace Cqr, and we obtain the required orthogonal decomposition H = Cq + Cqr. 

13.5. From complex structure to fibration

Recall that we have a decomposition H = R +H0 where (1) R = R 1 is the center of H, and (2) H0 is the space of pure imaginary quaternions. 13.6. LIE GROUPS 159

S1 S2 ❆ > ❆❆ ⑥⑥ ❆❆ fr ⑥⑥ ❆❆ ⑥⑥ ❆ ⑥⑥ S3 > ❆ ⑥⑥ ❆❆ f ⑥⑥ ❆❆ q ⑥⑥ ❆❆ ⑥⑥ ❆ S1 S2

Figure 13.5.1. A pair of fibrations; cf. Figure 13.12.1

Consider the complex structure on H R4 defined by a pure quater- nion q H , as in Proposition 13.4.4.≃ As in formula (13.1.2), the ∈ 0 choice of a complex structure leads to a Hopf fibration fq of the unit 3- sphere S3 H: ⊆ f S1 S3 q S2, (13.5.1) −→ −→ where each fiber of f is a circle S1 S3 which is an orbit of the action q ⊆ by the unit circle in Cq, similar to (13.1.1). Here we use the notation of Definition 13.1.3, where the first arrow in (13.5.1) denotes the inclusion of a fiber, while the second arrow denotes the Hopf map.

Remark 13.5.1. Distinct pure imaginary quaternions q,r H0 define distinct complex structures and hence lead to distinct Hopf∈ fi- brations fq, fr. Such a pair of fibrations, illustrated in diagram of Figure 13.5.1 following the notation used in (13.1.3), will play a crucial role in the proof of Pu’s inequality in Section 13.13.

13.6. Lie groups A Lie group is simultaneously a group and a smooth manifold, in such a way that the two structures are compatible. More precisely, we have the following definition. Definition 13.6.1. A Lie group is a manifold G with an associa- tive3 operation µ: G G G and inverse ν : G G with the following properties: × → → (1) the operation defines the structure of a group on G, so that in particular there exists an element e G such that µ(e,x)= x for all x G, and furthermore µ(x,ν∈(x)) = e; (2) both µ and∈ ν are smooth maps.

3chok kibutz 160 13. ALTERNATIVE PROOF OF PU’S INEQUALITY

Corollary 13.6.2. The manifold S3 admits the structure of a Lie group. Proof. Nonzero quaternions form a group by Theorem 13.3.3. The multiplication restricts to the unit sphere S3 R4 H and defines a well-defined smooth product. The smooth inverse⊆ is≃ given by 1 3 the rule q− =q ¯ for each q S .  ∈

13.7. Lie group SO(3) as quotient of S3 As we mentioned in Section 13.6, the sphere S3 can be thought of as the Lie group formed of the unit (Hamilton) quaternions in H C2. In the sequel, an important role will be played by the Lie group SO≃ (3) of orthogonal 3 3 matrices of determinant 1. It turns out that the group ×

SO(3) = SO(H0) can be identified with the quotient S3/ 1 of the Lie group S3 by its center, as follows. {± } Recall that conjugation by a quaternion is an isometry of H0.

3 Proposition 13.7.1. Consider H0 = R . (1) We have a natural isomorphism of Lie groups φ: S3/ Z ≃ 2 −→ SO(3), q Cq where Cq is conjugation by q acting on H0. (2) The nontrivial7→ element in the kernel is the element 1 S3 H. − ∈ ⊆ Proof. Consider a unit quaternion q S3 H. The homomor- ∈ 3 ⊆ phism φ sends q to the isometry Cq of H0 R given by conjugation 3 3 ≃ 1 by q. Namely, Cq : R R is the map Cq(x) = q− xq. Since the element 1 is in the center→ of H, conjugation by 1 gives a trivial − − isometry C 1 = Id. Thus the homomorphism descends to the quo- tient S3/ −1 . {± } To get all the rotations in SO(3) around an axis q0 H0, we con- jugate by cos θ 1+sin θq , where θ R. ∈  0 ∈ We obtain the following corollary.

3 Corollary 13.7.2. The Hopf fibration fq of S defined by a pure quaternion q H descends to a fibration ∈ 0 SO(3) S2. (13.7.1) → This result will be exploited in Section 13.9. 13.9. 2-SPHERE AS A HOMOGENEOUS SPACE 161

13.8. Unit sphere tangent bundle Definition 13.8.1. Let M be a Riemannian manifold. The unit sphere tangent bundle, denoted T uM, or the unit tangent bundle for short, is the submanifold of TM consisting of elements of unit norm: T uM = v TM : v =1 . ∈ | | Proposition 13.8.2. The choice of a unit tangent vector v T uSn ∈ provides a natural identification Fv of the group SO(n+1) with the unit tangent bundle of M = Sn. Proof. Choose a fixed unit tangent vector v T u(Sn). We will construct a natural identification ∈ F : SO(n + 1) T u(Sn), g dg(v). v → 7→ Namely, the group SO(n + 1) acts naturally on the sphere Sn Rn+1 by matrix multiplication. To construct the required identification⊆ F , we send an element g SO(n + 1), acting on Sn, to the image of the vector v under the tangent∈ map (see Section 12.2) dg : TSn TSn. Namely, F (g)= dg(v) T u(Sn). →  v ∈ Corollary 13.8.3. We have the following three distinct ways of viewing the same Lie group: (1) the group SO(3) of orthogonal matrices; (2) the quotient S3/ 1 of the 3-sphere; (3) the unit tangent{± bundle} of the 2-sphere. The underlying smooth manifold can in fact be identified with the real projective space RP3, as well, as is obvious from item (2).

13.9. 2-sphere as a homogeneous space The circle SO(2) of rotations of the (x,y)-plane can be viewed as the subgroup of SO(3) stabilizing the north pole, as follows. Lemma 13.9.1. The stabilizer4 of the north pole (0, 0, 1) S2 is the subgroup of matrices of the form ∈ cos θ sin θ 0 sin θ −cos θ 0  0 0 1 which is isomorphic to SO(2).  Let H G be a Lie subgroup of G. ⊆ 4meyatzev 162 13. ALTERNATIVE PROOF OF PU’S INEQUALITY

Definition 13.9.2. A homogeneous space G/H is the space of or- bits of a Lie group G under the action by a subgroup H. Corollary 13.9.3. The sphere S2 is the homogeneous space of the Lie group SO(3), cf. (13.7.1), so that S2 = SO(3)/SO(2). (13.9.1) See e.g., [Ar83, p. 82]. Definition 13.9.4. Given a point σ S2 viewed as a homogeneous ∈ space via (13.9.1), we will denote by SO(2)σ the fiber over (i.e., the stabilizer of) σ. Proposition 13.9.5. The projection p: SO(3) S2 (13.9.2) → is a Riemannian submersion in both of the following cases: (1) for a metric of constant Gaussian curvature +4 on S2 and the metric on SO(3) is of constant sectional curvature 1 (i.e., antipodal quotient of the unit 3-sphere); (2) for a metric of curvature 1 on the base S2 and metric of cur- 1 vature 4 on the total space SO(3). Remark 13.9.6. For the purposes of proving Pu’s inequality, we will view fibration (13.9.2) as a fibration of the unit circle tangent bundle of the sphere: p: T uS2 S2 (13.9.3) → using the identification of T uS2 and SO(3) discussed in Section 13.7.

13.10. Geodesic flow on the tangent bundle Let M be an n-dimensional Riemannian manifold. We define a geodesic in a coordinate patch as follows. Definition 13.10.1. A geodesic α(t)=(α1(t), ,αn(t)) in M with initial point p = (p1,...,pn) and initial velocity···v = (v1,...,vn) is defined using the Γ symbols of the Riemannian metric as a curve satisfying the ordinary differential equations k′′ k i′ j′ α +Γij α α =0, k =1,...,n, with the initial conditions α(0) = p and α′(0) = v. Definition 13.10.2. The geodesic starting at point p M with initial velocity v T M is denoted γ(p,v,t). ∈ ∈ p 13.11. DUAL REAL PROJECTIVE PLANE AND ITS DOUBLE COVER 163

It is easy to show5 that a geodesic has constant speed. Therefore we can define the geodesic flow on the unit tangent bundle as follows. Definition 13.10.3. The geodesic flow on T uM is the map u u R T M T M, (t, (p,v)) γ(p,v,t),γ′(p,v,t) . × → 7→ It was proved in differentialit 1 that great circles are geode sics on the sphere. Therefore we have the following. Lemma 13.10.4. When M = S2 with the standard metric normal- ized to have K =1, the flow is periodic with period 2π, so that γ(p,v,t+ 2π)= γ(p,v,t) when v =1. | | Recalling that the unit tangent circle bundle of S2 can be identified with SO(3), we obtain the following theorem. Theorem 13.10.5. Consider the geodesic flow of M = S2. Then (1) the flow defines a circle action S1 SO(3) SO(3); (2) its homogeneous space (i.e., space× of orbits)→ is the space of oriented 6 great circles of S2. D Thus we have a fibration π : SO(3) . →D Corollary 13.10.6. The unit tangent bundle T uS2 admits the fol- lowing pair of circle fibrations p and π: (1) the fibration p over the base manifold S2, with typical fiber 2 SO(2)σ over each point σ S ; (2) the fibration π over the space∈ of orbits of the geodesic flow, where the typical fiber ν T uSD2 is parametrized by the closed ⊆ geodesic γ(p,v,t),γ′(p,v,t) .

13.11. Dual real projective plane and its double cover Let RP2 be the real projective plane. The orientable double cover of RP2 can be identified with S2. 2 Definition 13.11.1. Let RP ∗ the dual real projective plane, namely the space of projective lines RP1 of RP2. 2 Definition 13.11.2. Denote by the orientable double cover of RP ∗, identifiable with the space of orientedD great circles (i.e., with the 2- sphere). Here “D” is an allusion to both “double” and “dual.”

5 See differentsialit 1 at https://u.cs.biu.ac.il/~katzmik/88-201.html 6mekuvanim 164 13. ALTERNATIVE PROOF OF PU’S INEQUALITY

Remark 13.11.3. The following observations should be kept in mind. (1) The preimage of a real projective line under the double cover S2 RP2 is a great circle on the 2-sphere; (2) We→ will avoid using the notation S2 for so as not to con- fuse the base manifolds of two distinct fibrationsD p and π of Corollary 13.10.6; (3) We think of as the configuration space of oriented great circles on theD 2-sphere. Definition 13.11.4. Let ν represent a generic oriented great circle. We will denote by dν the∈D measure (i.e., the area 2-form) on . D Theorem 13.11.5. The unit tangent bundle T uS2 can be repre- sented as a subset of the Cartesian product S2 as follows: ×D T uS2 = (x,ν) S2 : x ν ∈ ×D ∈ Proof. An oriented (directed) line through a point x S2 defines a unique unit tangent vector at x, and vice versa. ∈  13.12. A pair of fibrations We consider again the fibration p: T uS2 S2 of formula (13.9.3). From the point of view of Theorem 13.11.5,→ this fibration sends (x,ν) to x: p(x,ν)= x. By Theorem 13.11.5, there is a second fibration π of T uS2, defined by the formula π(x,ν) = ν. Thus, the total space T uS2 admits another Riemannian submersion, denoted π, over the configuration space of oriented circles: π : T uS2 , (13.12.1) →D whose typical fiber ν is an orbit of the geodesic flow on T uS2 (see Section 13.10). Lemma 13.12.1. Each orbit ν T uS2 of the geodesic flow projects under p to a great circle on the sphere.⊆ Proof. A fiber of fibration π is the collection of unit vectors tan- gent to a given directed closed geodesic, i.e., great circle, on S2. This great circle is precisely the image of ν under the projection p.  Remark 13.12.2. While fibration π may seem very different from fibration p, the two are actually equivalent from the quaternionic point of view; cf. Sections 13.3 and 13.5. The diagram of Figure 13.12.1 illustrates the maps defined so far. 13.13. DOUBLE FIBRATION OF T uS2, INTEGRAL GEOMETRY ON S2 165

SO(2) ( ,dν) ■ 9 ■■ tt9D ■■ π tt ■■ ttt ■$ tt T uS2 ✉: ❏❏ ✉✉ ❏❏ p ✉✉ ❏❏ ✉✉ ❏❏ ✉✉ ❏❏% ✉✉ % ν / (S2,dσ) great circle

Figure 13.12.1. Integral geometry on the sphere; cf. Figure 13.5.1, (13.9.2) and (13.12.1)

13.13. Double fibration of T uS2, integral geometry on S2 In this section, we prove Pu’s inequality using integral geometry. The latter has its origin in results of P. Funk [Fu16] determining a symmetric function on the two-sphere from its great circle integrals; see [Hel99, Proposition 2.2, p. 59], as well as Preface therein. Let g0 be the standard metric of constant Gaussian curvature +1 2 2 on S . Consider a metric g = f g0, where f > 0 is a function on the 2-sphere.

Theorem 13.13.1. We have the following inequality:

1 area(S2, g) L2 , (13.13.1) ≥ π min

2 where L min is the least g-length of a great circle of S . In the boundary case of equality, the function f must be constant. 166 13. ALTERNATIVE PROOF OF PU’S INEQUALITY

Proof. Let dσ be the area element of g0. Applying the Cauchy– Schwarz inequality and Fubini’s theorem twice (to Riemannian sub- mersions p and π), we obtain7 area(S2, g)= f 2dσ 2 ZZS 1 2 fdσ (by Cauchy–Schwarz) ≥ 4π 2 ZZS  1 1 2 = f p dvol (by Fubini applied to p) 4π 2π u 2 ◦  ZZZT S  1 2 = 3 f p dvol 16π u 2 ◦ ZZZT S  1 2 = 3 dν f(t)dt (by Fubini applied to π) 16π ν ZZD  Z  1 2 = length (ν)dν 16π3 g ZZD  1 (4πL )2 ≥ 16π3 min 1 = L2 π min Based on inequality (13.13.1), one proves Pu’s inequality for the real projective plane as in Chapter 12. 

Proof of Pu’s inequality. By the uniformisation theorem for the sphere S2, every metric on S2 is conformal to the standard one. A metric g 2 2 2 on RP lifts to a centrally symmetric metric g˜ = f g0 on S whose conformal factor f is invariant under the antipodal map. Applying Theorem 13.13.1 to the metric g˜, we obtain a great circle of g˜-length L satisfying A 1 L2, ≥ π where A = area(S2, g˜). Note that sys (g) L and area(g)= A . Hence 1 ≤ 2 2 L2 πA π sys2(g) = area(g), 1 ≤ 4 ≤ 4 2 proving Pu’s inequality. See [Iv02] for an alternative proof. 

13.13.1. Pu’s inequality with isosystolic remainder term. This material is optional as it was already covered in Chapter 12. Consider 1 2 the unit-area metric g1 = 4π g0 on S , where g0 is the metric of Gaussian curvature K = 1. Thus, g is isometric to a 2-sphere of radius 1 and 1 √4π

7This part could be modified to incorporate the variance remainder term. 13.13. DOUBLE FIBRATION OF T uS2, INTEGRAL GEOMETRY ON S2 167 constant Gaussian curvature K = 4π. Then 2 2 2 g = f g0 = 4πf g1 = √4πf g1.

Then the area form of g1 defines a probability measure µ. We will denote by dσ the area form of g0.

Lemma 13.13.2. The integral S2 f dσ can be expressed in terms of the expected value of the random variable √4πf by the formula R

f dσ = √4π Eµ √4πf . S2 Z   √ √ √ 1 Proof. We have Eµ( 4πf)= S2 4πf d areag1 = S2 4πf 4π d areag0 = 1 2 fd area , proving the lemma.  √4π S g0 R R R 2 Lemma 13.13.3. The integral S2 f dσ can be expressed in terms of the expected value of the random variable (√4πf)2 = 4πf 2 by the formula R 2 2 f dσ = Eµ 4πf . 2 ZS 2 2  2 1 Proof. We have Eµ(4πf )= S2 4πf d areag1 = S2 4πf 4π d areag0 = 2 2 f d area , proving the lemma.  S g0 R R R Lemma 13.13.4. We have the following identity: 2 2 1 f dσ fdσ = Varµ √4πf S2 − 4π S2 ZZ ZZ    Proof. The computational formula for the variance of the random vari- 2 able √4πf gives E 4πf 2 E √4πf = Var √4πf .  µ − µ µ 2 Corollary 13.13.5 . Let g= f g0. Then we have the following strength- ened version of inequality (13.13.1): area(S2, g) 1 L2 Var √4πf . − π min ≥ µ Corollary 13.13.6. We have the following strengthened form of Pu’s inequality: 2 1 area(RP2, g) sys2(RP2, g) Var √4πf (13.13.2) − π 1 ≥ 2 µ   where √4πf is viewed as a random variable on the double cover of RP2. An immediate application is the characterisation of the boundary case of equality in Pu’s inequality. Namely if equality is attained in Pu’s inequality then (13.13.2) implies Varµ √4πf = 0 and therefore f is constant. 

CHAPTER 14

Gromov’s inequality for essential manifolds

14.1. Essential manifolds We deal with Gromov’s systolic inequality for essential manifolds. To give some examples, all surfaces are essential with the excep- tion of the sphere S2. Real projective spaces RPn are essential. If a manifold M admits a metric of negative curvature, then M is essential. More generally, to define the property of being essential for an n- dimensional closed manifold M, we need a minimum of homology the- ory.

14.2. Gromov’s inequality for essential manifolds One of the deepest results in the field of systolic geometry is Gro- mov’s inequality for the homotopy 1-systole of an essential n-manifold M: sys(M)n C vol(M), ≤ n where Cn is a universal constant only depending on the dimension of M. Here the systole sys(M) is by definition the least length of a noncontractible loop in M. Example 14.2.1. The inequality holds for all nonsimply connected surfaces, for real projective spaces, for all manifolds admitting a hy- perbolic metric. The proof involves a new invariant called the filling radius, intro- duced by Gromov, defined as follows.

14.3. Filling radius of a loop in the plane The filling radius of a simple loop C in the plane is defined as the largest radius, R> 0, of a circle that fits inside C, see Figure 14.2.1: Fillrad(C R2)= R. ⊆ There is a kind of dual point of view that allows one to generalize this notion in a fruitful way, as shown in [Gro83]. Namely, we consider 2 the ǫ-neighborhoods of the loop C, denoted UǫC R , see Figure 14.3.1. ⊆ 169 170 14. GROMOV’S INEQUALITY FOR ESSENTIAL MANIFOLDS

R

C

Figure 14.2.1. Largest inscribed circle has radius R

C

Figure 14.3.1. Neighborhoods of loop C

As ǫ> 0 increases, the ǫ-neighborhood UǫC swallows up more and more of the interior of the loop. The “last” point to be swallowed up is precisely the center of a largest inscribed circle. Therefore we can reformulate the above definition by setting

Fillrad(C R2) = inf ǫ> 0 loop C contracts to a point in U C . ⊆ { | ǫ }

To define an absolute filling radius in a situation where M is equipped with a Riemannian metric g, Gromov exploits an embedding due to C. Kuratowski. 14.7. HOMOLOGY THEORY FOR GROUPS 171

14.4. The sup-norm

On the space L∞(X) of bounded Borel functions f on a set X, one can consider the norm f = sup f(x) . k k x X | | ∈ This norm is called the sup-norm.

14.5. Riemannian manifolds as spaces with a distance function We consider a manifold M equipped with a distance function d(x,y), where x,y M. In particular, a Riemannian metric on M defines such a distance∈ function, by minimizing the length of all paths joining x to y.

14.6. Kuratowski embedding

One embeds M in the Banach space L∞(X) of bounded Borel functions on M, equipped with the sup norm . Namely, we map k k a point x M to the function f L∞(M) defined by the for- ∈ x ∈ mula fx(y) = d(x,y) for all y M, where d is the distance function defined by the metric. By the∈ triangle inequality we have d(x,y) = fx fy , and therefore the embedding is strongly isometric, in the pre- kcise− sensek that internal distance and ambient distance coincide. Such a strongly isometric embedding is impossible if the ambient space is a Hilbert space, even when M is the Riemannian circle (the distance between opposite points must be π, not 2!). We then set E = L∞(M) in the formula above.

14.7. Homology theory for groups Recall that the n-dimensional homology group is nontrivial. If M is orientable, then Hn(M; Z)= Z. If M is non-orientable, then Hn(N, Z2)= Z2. In either case, a generator of the homology group is denoted [M] and called the fundamental class. There is a parallel homology theory for groups. In particular, one can consider the homology of the fundamental group π = π1(M), namely Hn(π). One additional ingredient we need to know is the exis- tence of a natural homomorphism φ: H (M) H (π). n → n 172 14. GROMOV’S INEQUALITY FOR ESSENTIAL MANIFOLDS

Then M is called essential if its fundamental class [M] maps to a nonzero class in the homology of its fundamental group: φ([M]) =0 H (π). 6 ∈ n 14.8. Relative filling radius We first define a notion of a filling radius of M relative to an em- bedding. Given an embedding of M in Euclidean space E, let ǫ > 0, and denote by UǫM the neighborhood in E consisting of all points at distance at most ǫ from M. Let φ : H (M) H (E) ǫ n → n be the inclusion homomorphism induced by the inclusion of M in its ǫ-neighborhood UǫM in E. We set Fillrad(M E) = inf ǫ> 0 φ ([M])=0 H (U M) , ⊆ { | ǫ ∈ n ǫ } 14.9. Absolute filling radius We define the filling radius of M to be its relative filling radius relative to the canonical embedding in L∞(M), by setting

FillRad(M) = FillRad(M L∞(M)) . ⊆ Namely, Gromov proved a sharp inequality relating the systole and the filling radius: sysπ1 6 FillRad(M), valid for all essential manifolds M; as well as an inequality≤

1 FillRad C vol n (M), ≤ n n valid for all closed manifolds M. A summary of a proof, based on recent results in geometric measure theory by S. Wenger, building upon earlier work by L. Ambrosio and B. Kirchheim, appears in Section 12.2 of the book ”Systolic geometry and topology” referenced below. 14.10. Systolic freedom First examples of systolic freedom were given by Gromov in [Gro96, pp. 350–351]. The systolic freedom of Sn Sn for n 3 was proved by Katz in ([10], 1995). The systolic freedom× of S2 ≥S2 was proved by Katz and Suciu in ([12], 1999). For other results see× [BaK98], [Fr99]. Bibliography

[1] Bangert, V.; Croke, C.; Ivanov, S.; Katz, M. Boundary case of equal- ity in optimal Loewner-type inequalities. Transactions of the American Mathematical Society 359 (2007) (1), 1–17. [2] Berger, M. Lectures on geodesics in Riemannian geometry. Tata Insti- tute of Fundamental Research Lectures on Mathematics, No. 33 Tata Institute of Fundamental Research, Bombay 1965. [3] Csik´os, B.; Horv´ath, M. Harmonic manifolds and tubes. J. Geom. Anal. 28 (2018), no. 4, 3458–3476. [4] Gauld, D. Selections and metrisability of manifolds. Topology Appl. 160 (2013), no. 18, 2473–2481. [5] Gromov, M. Filling Riemannian manifolds. J. Differential Geom. 18 (1983), no. 1, 1–147. [6] Ivanov, S.; Katz, M. Generalized degree and optimal Loewner-type in- equalities. Israel J. Math. 141 (2004), 221–233. [7] Katz, M.; Nowik, T. A systolic inequality with remainder in the real projective plane. Open Mathematics (2020), to appear. [8] Katz, M.; Sabourau, S. Hyperelliptic surfaces are Loewner. Proceedings of the American Mathematical Society 134 (2006), no. 4, 1189–1195. [Ab81] Abikoff, W.: The uniformization theorem. Amer. Math. Monthly 88 (1981), no. 8, 574–592. [Ac60] Accola, R.: Differentials and extremal length on Riemann surfaces. Proc. Nat. Acad. Sci. USA 46 (1960), 540–543. [Am04] Ammann, B.: Dirac eigenvalue estimates on two-tori. J. Geom. Phys. 51 (2004), no. 3, 372–386. [Ar83] Armstrong, M.: Basic topology. Corrected reprint of the 1979 original. Undergraduate Texts in Math. Springer-Verlag, New York-Berlin, 1983. [Ar74] Arnold, V.: Mathematical methods of classical mechanics. Translated from the 1974 Russian original by K. Vogtmann and A. Weinstein. Cor- rected reprint of the second (1989) edition. Graduate Texts in Mathe- matics, 60. Springer-Verlag, New York, 199?. [Ba93] Babenko, I.: Asymptotic invariants of smooth manifolds. Russian Acad. Sci. Izv. Math. 41 (1993), 1–38. [Ba02a] Babenko, I. Forte souplesse intersystolique de vari´et´es ferm´ees et de poly`edres. Annales de l’Institut Fourier 52, 4 (2002), 1259-1284. [Ba02b] Babenko, I. Loewner’s conjecture, Besicovich’s example, and relative systolic geometry. [Russian]. Mat. Sbornik 193 (2002), 3-16. [Ba04] Babenko, I.: G´eom´etrie systolique des vari´et´es de groupe fondamen- tal Z2, S´emin. Th´eor. Spectr. G´eom. Grenoble 22 (2004), 25-52.

173 174 BIBLIOGRAPHY

[BB05] Babenko, I.; Balacheff, F.: G´eom´etrie systolique des sommes connexes et des revˆetements cycliques, Mathematische Annalen (to appear). [BaK98] Babenko, I.; Katz, M.: Systolic freedom of orientable manifolds, Annales Scientifiques de l’E.N.S. (Paris) 31 (1998), 787-809. [Bana93] Banaszczyk, W. New bounds in some transference theorems in the ge- ometry of numbers. Math. Ann. 296 (1993), 625–635. [Bana96] Banaszczyk, W. Inequalities for convex bodies and polar reciprocal lat- tices in Rn II: Application of K-convexity. Discrete Comput. Geom. 16 (1996), 305–311. [BCIK04] Bangert, V; Croke, C.; Ivanov, S.; Katz, M. Boundary case of equality in optimal Loewner-type inequalities. Trans. Amer. Math. Soc. 359 (2007), no. 1, 1–17. See https://arxiv.org/abs/math/0406008 [BCIK05] Bangert, V; Croke, C.; Ivanov, S.; Katz, M.: Filling area conjecture and ovalless real hyperelliptic surfaces, Geometric and Functional Analysis (GAFA) 15 (2005) no. 3. See arXiv:math.DG/0405583 [BK03] Bangert, V.; Katz, M.: Stable systolic inequalities and cohomology products, Comm. Pure Appl. Math. 56 (2003), 979–997. Available at arXiv:math.DG/0204181 [BK04] Bangert, V; Katz, M.: An optimal Loewner-type systolic inequality and harmonic one-forms of constant norm. Comm. Anal. Geom. 12 (2004), number 3, 703-732. See arXiv:math.DG/0304494 [Bar57] Barnes, E. S.: On a theorem of Voronoi. Proc. Cambridge Philos. Soc. 53 (1957), 537–539. [Bav86] Bavard, C.: In´egalit´eisosystolique pour la bouteille de Klein. Math. Ann. 274 (1986), no. 3, 439–441. [Bav88] Bavard, C.: In´egalit´es isosystoliques conformes pour la bouteille de Klein. Geom. Dedicata 27 (1988), 349–355. [Bav92] Bavard, C.: La systole des surfaces hyperelliptiques, Prepubl. Ec. Norm. Sup. Lyon 71 (1992). [Bav06] Bavard, C.: Une remarque sur la g´eom´etrie systolique de la bouteille de Klein. Arch. Math. (Basel) 87 (2006), no. 1, 72–74. [BeM] Berg´e, A.-M.; Martinet, J. Sur un probl`eme de dualit´eli´eaux sph`eres en g´eom´etrie des nombres. J. Number Theory 32 (1989), 14–42. [Ber72] Berger, M. A l’ombre de Loewner. Ann. Scient. Ec. Norm. Sup. (S´er. 4) 5 (1972), 241–260. [Ber80] Berger, M.: Une borne inf´erieure pour le volume d’une vari´et´erieman- nienne en fonction du rayon d’injectivit´e. Ann. Inst. Fourier 30 (1980), 259–265. [Ber93] Berger, M. Systoles et applications selon Gromov. S´eminaire N. Bour- baki, expos´e771, Ast´erisque 216 (1993), 279–310. [Ber08] Berger, M.: What is... a Systole? Notices Amer. Math. Soc. 55 (2008), no. 3, 374-376. [Ber76] Berstein, I.: On the Lusternik–Schnirelmann category of Grassmannians, Proc. Camb. Phil. Soc. 79 (1976), 129-134. [Bes87] Besse, A.: Einstein manifolds. Ergebnisse der Mathematik und ihrer Grenzgebiete (3), 10. Springer-Verlag, Berlin, 1987. [BesCG95] Besson, G.; Courtois, G.;, Gallot, S.: Volume et entropie minimale des espaces localement sym´etriques. Invent. Math. 103 (1991), 417–445. BIBLIOGRAPHY 175

[BesCG95] Besson, G.; Courtois, G.;, Gallot, S.: Hyperbolic manifolds, amalga- mated products and critical exponents. C. R. Math. Acad. Sci. Paris 336 (2003), no. 3, 257–261. [Bl61] Blatter, C. Uber¨ Extremall¨angen auf geschlossenen Fl¨achen. Comment. Math. Helv. 35 (1961), 153–168. [Bl61b] Blatter, C.: Zur Riemannschen Geometrie im Grossen auf dem M¨obius- band. Compositio Math. 15 (1961), 88–107. [Bo21] Bonnesen, T.: Sur une am´elioration de l’in´egalit´e isop´erimetrique du cercle et la d´emonstration d’une in´egalit´ede Minkowski. C. R. Acad. Sci. Paris 172 (1921), 1087-1089. [Boothby 1986] Boothby, W. An introduction to differentiable manifolds and Rie- mannian geometry. Second edition. Pure and Applied Mathematics, 120. Academic Press, Inc., Orlando, FL, 1986. [Bo94] Borzellino, J. E.: Pinching theorems for teardrops and footballs of rev- olution. Bull. Austral. Math. Soc. 49 (1994), no. 3, 353–364. [BG76] Bousfield, A.; Gugenheim, V.: On PL de Rham theory and rational homotopy type. Mem. Amer. Math. Soc. 8, 179, Providence, R.I., 1976. [BI94] Burago, D.; Ivanov, S.: Riemannian tori without conjugate points are flat. Geom. Funct. Anal. 4 (1994), no. 3, 259–269. [BI95] Burago, D.; Ivanov, S.: On asymptotic volume of tori. Geom. Funct. Anal. 5 (1995), no. 5, 800–808. [BuraZ80] Burago, Y.; Zalgaller, V.: Geometric inequalities. Grundlehren der Mathematischen Wissenschaften 285. Springer Series in Soviet Math- ematics. Springer-Verlag, Berlin, 1988 (original Russian edition: 1980). [BS94] Buser, P.; Sarnak, P.: On the period matrix of a Riemann surface of large genus. With an appendix by J. H. Conway and N. J. A. Sloane. Invent. Math. 117 (1994), no. 1, 27–56. [Cal96] Calabi, E.: Extremal isosystolic metrics for compact surfaces, Actes de la table ronde de geometrie differentielle, Sem. Congr. 1 (1996), Soc. Math. France, 167–204. [Ca76] Manfredo do Carmo, Differential geometry of curves and surfaces. Prentice-Hall, 1976. [Car92] do Carmo, M.: Riemannian geometry. Translated from the second Por- tuguese edition by Francis Flaherty. Mathematics: Theory & Applica- tions. Birkh¨auser Boston, Inc., Boston, MA, 1992. [Ca71] Cassels, J. W. S.: An introduction to the geometry of numbers. Sec- ond printing. Grundlehren der mathematischen Wissenschaften, 99. Springer-Verlag, Berlin-Heidelberg-New York, 1971. [Ch93] Chavel, I.: Riemannian geometry—a modern introduction. Cambridge Tracts in Mathematics, 108. Cambridge University Press, Cambridge, 1993. [CoS94] Conway, J.; Sloane, N.: On lattices equivalent to their duals. J. Number Theory 48 (1994), no. 3, 373–382. [ConS99] Conway, J. H.; Sloane, N. J. A.: Sphere packings, lattices and groups. Third edition. Springer, 1999. [CoT03] Cornea, O.; Lupton, G.; Oprea, J.; Tanr´e, D.: Lusternik-Schnirelmann category. Mathematical Surveys and Monographs, 103. American Math- ematical Society, Providence, RI, 2003. 176 BIBLIOGRAPHY

[Cr80] Croke, C.: Some isoperimetric inequalities and eigenvalue estimates. Ann. Sci. Ecole´ Norm. Sup. 13 (1980), 519–535. [Cr88] Croke, C.: An isoembolic pinching theorem. Invent. Math. 92 (1988), 385–387. [DaS98] Danilov, V. I.; Shokurov, V. V.: Algebraic curves, algebraic manifolds and schemes. Translated from the 1988 Russian original by D. Coray and V. N. Shokurov. Reprint of the original English edition from the series Encyclopaedia of Mathematical Sciences [Algebraic geometry. I, Encyclopaedia Math. Sci., 23, Springer, Berlin, 1994]. Springer-Verlag, Berlin, 1998. [DR03] Dranishnikov, A.; Rudyak, Y.: Examples of non-formal closed (k 1)- connected manifolds of dimensions 4k 1, Proc. Amer. Math.− Soc. 133 (2005), no. 5, 1557-1561. See arXiv:math.AT/0306299≥ − . [FK92] Farkas, H. M.; Kra, I.: Riemann surfaces. Second edition. Graduate Texts in Mathematics, 71. Springer-Verlag, New York, 1992. [Fe69] Federer, H.: Geometric Measure Theory. Grundlehren der mathematis- chen Wissenschaften, 153. Springer-Verlag, Berlin, 1969. [Fe74] Federer, H.: Real flat chains, cochains, and variational problems. Indiana Univ. Math. J. 24 (1974), 351–407. [FHL98] Felix, Y.; Halperin, S.; Lemaire, J.-M.: The rational LS category of products and of Poincar´eduality complexes. Topology 37 (1998), no. 4, 749–756. [Fer77] Ferrand, J.: Sur la regularite des applications conformes, C.R. Acad.Sci. Paris Ser.A-B 284 (1977), no.1, A77-A79. [Fr99] Freedman, M.: Z2-Systolic-Freedom. Geometry and Topology Mono- graphs 2, Proceedings of the Kirbyfest (J. Hass, M. Scharlemann eds.), 113–123, Geometry & Topology, Coventry, 1999. [Fu16] Funk, P.: Uber eine geometrische Anwendung der Abelschen Integral- gleichung. Math. Ann. 77 (1916), 129-135. [GaHL04] Gallot, S.; Hulin, D.; Lafontaine, J.: Riemannian geometry. Third edi- tion. Universitext. Springer-Verlag, Berlin, 2004. [Ga71] Ganea, T.: Some problems on numerical homotopy invariants. Sympo- sium on Algebraic Topology (Battelle Seattle Res. Center, Seattle Wash., 1971), pp. 23–30. Lecture Notes in Math., 249, Springer, Berlin, 1971. [GG92] G´omez-Larra˜naga, J.; Gonz´alez-Acu˜na, F.: Lusternik-Schnirel’mann category of 3-manifolds. Topology 31 (1992), no. 4, 791–800. [Gre67] Greenberg, M. Lectures on Algebraic Topology. W.A. Benjamin, Inc., New York, Amsterdam, 1967. [GHV76] Greub, W.; Halperin, S.; Vanstone, R.: Connections, curvature, and cohomology. Vol. III: Cohomology of principal bundles and homogeneous spaces, Pure and Applied Math., 47, Academic Press, New York-London, 1976. [GriH78] Griffiths, P.; Harris, J.: Principles of algebraic geometry. Pure and Ap- plied Mathematics. Wiley-Interscience [John Wiley & Sons], New York, 1978. [Gro81a] Gromov, M.: Structures m´etriques pour les vari´et´es riemanniennes. (French) [Metric structures for Riemann manifolds] Edited by J. La- fontaine and P. Pansu. Textes Math´ematiques, 1. CEDIC, Paris, 1981. BIBLIOGRAPHY 177

[Gro81b] Gromov, M.: Volume and bounded cohomology. Publ. IHES. 56 (1981), 213–307. [Gro83] Gromov, M.: Filling Riemannian manifolds, J. Diff. Geom. 18 (1983), 1-147. [Gro96] Gromov, M.: Systoles and intersystolic inequalities, Actes de la Table Ronde de G´eom´etrie Diff´erentielle (Luminy, 1992), 291–362, S´emin. Congr., 1, Soc. Math. France, Paris, 1996. www.emis.de/journals/SC/1996/1/ps/smf sem-cong 1 291-362.ps.gz [9] Gromov, M. Metric structures for Riemannian and non-Riemannian spaces. Based on the 1981 French original With appendices by M. Katz, P. Pansu and S. Semmes. Translated from the French by Sean Michael Bates. Progress in Mathematics, 152. Birkh¨auser Boston, Boston, MA, 1999. [GroH81] Gross, B. H.; Harris, J.: Real algebraic curves. Ann. Sci. Ecole Norm. Sup. (4) 14 (1981), no. 2, 157–182. [GruL87] Gruber, P.; Lekkerkerker, C.: Geometry of numbers. Second edition. North-Holland Mathematical Library, 37. North-Holland Publishing Co., Amsterdam, 1987. [Hat02] Hatcher, A.: Algebraic topology. Cambridge University Press, Cam- bridge, 2002. [He86] Hebda, J. The collars of a Riemannian manifold and stable isosystolic inequalities. Pacific J. Math. 121 (1986), 339-356. [Hel99] Helgason, S.: The Radon transform. 2nd edition. Progress in Mathemat- ics, 5. Birkh¨auser, Boston, Mass., 1999. [Hem76] Hempel, J.: 3-Manifolds. Ann. of Math. Studies 86, Princeton Univ. Press, Princeton, New Jersey 1976. [Hil03] Hillman, J. A.: Tits alternatives and low dimensional topology. J. Math. Soc. Japan 55 (2003), no. 2, 365–383. [Hi04] Hillman, J. A.: An indecomposable PD3-complex: II, Algebraic and Geometric Topology 4 (2004), 1103-1109. [Horowitz et al.] Horowitz, C.; Usadi Katz, K.; Katz, M. Loewner’s torus inequality with isosystolic defect. J. Geom. Anal. 19 (2009), no. 4, 796–808. [Iv02] Ivanov, S.: On two-dimensional minimal fillings, St. Petersbg. Math. J. 13 (2002), no. 1, 17–25. [IK04] Ivanov, S.; Katz, M.: Generalized degree and optimal Loewner-type in- equalities, Israel J. Math. 141 (2004), 221-233. arXiv:math.DG/0405019 [Iw02] Iwase, N.: A∞-method in Lusternik-Schnirelmann category. Topology 41 (2002), no. 4, 695–723. [Jo48] John, F. Extremum problems with inequalities as subsidiary conditions. Studies and Essays Presented to R. Courant on his 60th Birthday, Jan- uary 8, 1948, 187–204. Interscience Publishers, Inc., New York, N. Y., 1948. [Kanovei et al. 2016] Kanovei, V.; Katz, K.; Katz, M.; Nowik, T. “Small os- cillations of the pendulum, Euler’s method, and adequality.” Quan- tum Studies: Mathematics and Foundations 3 (2016), no. 3, 231-236. See http://dx.doi.org/10.1007/s40509-016-0074-x and https:// arxiv.org/abs/1604.06663 178 BIBLIOGRAPHY

[Kato83] Katok, A.: Entropy and closed geodesics, Ergo. Th. Dynam. Sys. 2 (1983), 339–365. [Kato92] Katok, S.: Fuchsian groups. Chicago Lectures in Mathematics. Univer- sity of Chicago Press, Chicago, IL, 1992. [KatH95] Katok, A.; Hasselblatt, B.: Introduction to the modern theory of dynam- ical systems. With a supplementary chapter by Katok and L. Mendoza. Encyclopedia of Mathematics and its Applications, 54. Cambridge Uni- versity Press, Cambridge, 1995. [10] Katz, M. Counterexamples to isosystolic inequalities. Geom. Dedicata 57 (1995), no. 2, 195–206. [Ka83] Katz, M.: The filling radius of two-point homogeneous spaces, J. Diff. Geom. 18 (1983), 505-511. [Ka88] Katz, M.: The first diameter of 3-manifolds of positive scalar curvature. Proc. Amer. Math. Soc. 104 (1988), no. 2, 591–595. [Ka02] Katz, M. Local calibration of mass and systolic geometry. Geom. Funct. Anal. (GAFA) 12, issue 3 (2002), 598-621. See arXiv:math.DG/0204182 [Ka03] Katz, M.: Four-manifold systoles and surjectivity of period map, Com- ment. Math. Helv. 78 (2003), 772-786. See arXiv:math.DG/0302306 [Ka07] Katz, M.: Systolic geometry and topology. With an appendix by Jake P. Solomon. Mathematical Surveys and Monographs, 137. American Math- ematical Society, Providence, RI, 2007. [11] Katz, M.; Leichtnam, E. “Commuting and noncommuting infinites- imals.” American Mathematical Monthly 120 (2013), no. 7, 631– 641. See http://dx.doi.org/10.4169/amer.math.monthly.120.07. 631 and https://arxiv.org/abs/1304.0583 [KL04] Katz, M.; Lescop, C.: Filling area conjecture, optimal systolic inequali- ties, and the fiber class in abelian covers. Proceedings of conference and workshop in memory of R. Brooks, held at the Technion, Israel Math- ematical Conference Proceedings (IMCP), Contemporary Math., Amer. Math. Soc., Providence, R.I. (to appear). See arXiv:math.DG/0412011 [KR04] Katz, M.; Rudyak, Y.: Lusternik-Schnirelmann category and of low dimensional manifolds. Communications on Pure and Applied Mathematics, to appear. See arXiv:math.DG/0410456 [KR05] Katz, M.; Rudyak, Y.: Bounding volume by systoles of 3-manifolds. See arXiv:math.DG/0504008 [KatzRS06] Katz, M.; Rudyak, Y.; Sabourau, S.: Systoles of 2-complexes, Reeb graph, and Grushko decomposition. International Math. Research No- tices 2006 (2006). Art. ID 54936, pp. 1–30. See arXiv:math.DG/0602009 [KS04] Katz, M.; Sabourau, S.: Hyperelliptic surfaces are Loewner, Proc. Amer. Math. Soc., to appear. See arXiv:math.DG/0407009 [KS05] Katz, M.; Sabourau, S.: Entropy of systolically extremal surfaces and asymptotic bounds, Ergodic Theory and Dynamical Systems, 25 (2005). See arXiv:math.DG/0410312 [KS06] Katz, M.; Sabourau, S.: An optimal systolic inequality for CAT(0) met- rics in genus two, preprint. See arXiv:math.DG/0501017 [KSV05] Katz, M.; Schaps, M.; Vishne, U.: Logarithmic growth of sys- tole of arithmetic Riemann surfaces along congruence subgroups. See arXiv:math.DG/0505007 BIBLIOGRAPHY 179

[12] Katz, Mikhail G.; Suciu, Alexander I. Volume of Riemannian manifolds, geometric inequalities, and homotopy theory. Tel Aviv Topology Con- ference: Rothenberg Festschrift (1998), 113–136, Contemp. Math., 231, Amer. Math. Soc., Providence, RI, 1999. [Kl82] Klein, F.: Uber¨ Riemanns Theorie der algebraischen Funktionen und ihrer Integrale. – Eine Erg¨anzung der gew¨ohnlichen Darstellungen. Leipzig: B.G. Teubner 1882. [Kli82] Klingernberg, W.: Riemannian geometry. [Ko87] Kodani, S.: On two-dimensional isosystolic inequalities, Kodai Math. J. 10 (1987), no. 3, 314–327. [Kon03] Kong, J.: Seshadri constants on Jacobian of curves. Trans. Amer. Math. Soc. 355 (2003), no. 8, 3175–3180. [Ku78] Kung, H. T.; Leiserson, C. E.: Systolic arrays (for VLSI). Sparse Matrix Proceedings 1978 (Sympos. Sparse Matrix Comput., Knoxville, Tenn., 1978), pp. 256–282, SIAM, Philadelphia, Pa., 1979. [LaLS90] Lagarias, J.C.; Lenstra, H.W., Jr.; Schnorr, C.P.: Bounds for Korkin- Zolotarev reduced bases and successive minima of a lattice and its recip- rocal lattice. Combinatorica 10 (1990), 343–358. [13] Lawson, H.B. The Riemannian geometry of holomorphic curves. Proc. Conf. on Holomorphic mappings and minimal surfaces. Bol. Soc. Brazil. Mat. 2, 45-62 (1971) [La75] Lawson, H.B. The stable homology of a flat torus. Math. Scand. 36 (1975), 49–73. [14] Lee, John M. Introduction to Riemannian manifolds. Second edition. Graduate Texts in Mathematics, 176. Springer, Cham, 2018. [Li69] Lichnerowicz, A.: Applications harmoniques dans un tore. C.R. Acad. Sci., S´er. A 269, 912–916 (1969). [Lo71] Loewner, C.: Theory of continuous groups. Notes by H. Flanders and M. Protter. Mathematicians of Our Time 1, The MIT Press, Cambridge, Mass.-London, 1971. [Ma69] Macbeath, A. M.: Generators of the linear fractional groups. 1969 Num- ber Theory (Proc. Sympos. Pure Math., Vol. XII, Houston, Tex., 1967) pp. 14–32 Amer. Math. Soc., Providence, R.I. [Ma79] Manning, A.: Topological entropy for geodesic flows, Ann. of Math. (2) 110 (1979), 567–573. [Ma69] May, J.: Matric Massey products. J. Algebra 12 (1969) 533–568. [MP77] Millman, R.; Parker, G. Elements of differential geometry. Prentice-Hall Inc., Englewood Cliffs, N. J., 1977. [MS86] Milman, V. D.; Schechtman, G.: Asymptotic theory of finite-dimensional normed spaces. With an appendix by M. Gromov. Lecture Notes in Math- ematics, 1200. Springer-Verlag, Berlin, 1986. [MilH73] Milnor, J.; Husemoller, D.: Symmetric bilinear forms. Springer, 1973. [Mi95] Miranda, R.: Algebraic curves and Riemann surfaces. Graduate Stud- ies in Mathematics, 5. American Mathematical Society, Providence, RI, 1995. [MiTW73] Misner, Charles W.; Thorne, Kip S.; Wheeler, John Archibald: Gravi- tation. W. H. Freeman and Co., San Francisco, Calif., 1973. 180 BIBLIOGRAPHY

[Mo86] Moser, J. Minimal solutions of variational problems on a torus. Ann. Inst. H. Poincar´e-Analyse non lin´eaire 3 (1986), 229–272. [MS92] Moser, J.; Struwe, M. On a Liouville-type theorem for linear and non- linear elliptic differential equations on a torus. Bol. Soc. Brasil. Mat. 23 (1992), 1–20. [NR02] Nabutovsky, A.; Rotman, R.: The length of the shortest closed geodesic on a 2-dimansional sphere, Int. Math. Res. Not. 2002:23 (2002), 1211- 1222. [OR01] Oprea, J.; Rudyak, Y.: Detecting elements and Lusternik-Schnirelmann category of 3-manifolds. Lusternik-Schnirelmann category and related topics (South Hadley, MA, 2001), 181–191, Contemp. Math., 316, Amer. Math. Soc., Providence, RI, 2002. arXiv:math.AT/0111181 [15] Pansu, P. Profil isop´erim´etrique, m´etriques p´eriodiques et formes d’´equilibre des cristaux. ESAIM Control Optim. Calc. Var. 4 (1999), 631–665. [Pa04] Parlier, H.: Fixed-point free involutions on Riemann surfaces. See arXiv:math.DG/0504109 [PP03] Paternain, G. P.; Petean, J.: Minimal entropy and collapsing with cur- vature bounded from below. Invent. Math. 151 (2003), no. 2, 415–450. [Pu52] Pu, P.M.: Some inequalities in certain nonorientable Riemannian mani- folds, Pacific J. Math. 2 (1952), 55–71. [Ri1854] B. Riemann, Hypoth`eses qui servent de fondement `a la g´eom´etrie, in Oeuvres math´ematiques de Riemann. J. Gabay, 1990. [RT90] Rosenbloom, M.; Tsfasman, M.: Multiplicative lattices in global fields. Invent. Math. 101 (1990), no. 3, 687–696. [Rudin 1976] Rudin, W. Principles of mathematical analysis. Third edition. Inter- national Series in Pure and Applied Mathematics. McGraw-Hill Book Co., New York-Auckland-D¨usseldorf, 1976. [Ru87] Rudin, W.: Real and complex analysis. Third edition. McGraw-Hill Book Co., New York, 1987. [Ru99] Rudyak, Y.: On category weight and its applications. Topology 38, 1, (1999), 37–55. [RT00] Rudyak, Y.; Tralle, A.: On Thom spaces, Massey products, and non- formal symplectic manifolds. Internat. Math. Res. Notices 10 (2000), 495–513. [Sa04] Sabourau, S.: Systoles des surfaces plates singuli`eres de genre deux, Math. Zeitschrift 247 (2004), no. 4, 693–709. [Sa05] Sabourau, S.: Entropy and systoles on surfaces, preprint. [Sa06] Sabourau, S.: Systolic volume and minimal entropy of aspherical mani- folds, preprint. [Sak88] Sakai, T.: A proof of the isosystolic inequality for the Klein bottle. Proc. Amer. Math. Soc. 104 (1988), no. 2, 589–590. [SS03] Schmidt, Thomas A.; Smith, Katherine M.: Galois orbits of principal congruence Hecke curves. J. London Math. Soc. (2) 67 (2003), no. 3, 673–685. [Sen91a] Senn, W. Strikte Konvexit¨at f¨ur Variationsprobleme auf dem n- dimensionalen Torus. manuscripta math. 71 (1991), 45–65. n BIBLIOGRAPHY 181

[Sen91b] Senn, W. Uber¨ Mosers regularisiertes Variationsproblem f¨ur mini- male Bl¨atterungen des n-dimensionalen Torus. J. Appl. Math. Physics (ZAMP) 42 (1991), 527–546. [Ser73] Serre, J.-P.: A course in arithmetic. Translated from the French. Gradu- ate Texts in Mathematics, No. 7. Springer-Verlag, New York-Heidelberg, 1973. [Sik05] Sikorav, J.: Bounds on primitives of differential forms and cofilling in- equalities. See arXiv:math.DG/0501089 [Sin78] Singhof, W.: Generalized higher order cohomology operations induced by the diagonal mapping. Math. Z. 162, 1, (1978), 7–26. [Sm62] Smale, S.: On the structure of 5-manifolds. Ann. of Math. (2) 75 (1962), 38–46. [Spivak 1979] Spivak, M. A comprehensive introduction to differential geometry. Vol. 4. Second edition. Publish or Perish, Inc., Wilmington, Del., 1979. [Tr93] Tralle, A.: On compact homogeneous spaces with nonvanishing Massey products, in Differential geometry and its applications (Opava, 1992), pp. 47–50, Math. Publ. 1, Silesian Univ. Opava, 1993. [Tu47] Tutte, W.: A family of cubical graphs. Proc. Cambridge Philos. Soc. 43 (1947), 459–474. [Wa83] Warner, F.W.: Foundations of Differentiable Manifolds and Lie Groups. Graduate Texts in Mathematics, 94. Springer-Verlag, New York- Heidelberg-Berlin, 1983. [We55] Weatherburn, C. E.: Differential geometry of three dimensions. Vol. 1. Cambridge University Press, 1955. [50] White, B.: Regularity of area-minimizing hypersurfaces at bound- aries with multiplicity. Seminar on minimal submanifolds (E. Bombieri ed.), 293-301. Annals of Mathematics Studies 103. Princeton University Press, Princeton N.J., 1983. [Wr74] Wright, Alden H.: Monotone mappings and degree one mappings be- tween PL manifolds. Geometric topology (Proc. Conf., Park City, Utah, 1974), pp. 441–459. Lecture Notes in Math., Vol. 438, Springer, Berlin, 1975. [Ya88] Yamaguchi, T.: Homotopy type finiteness theorems for certain precom- pact families of Riemannian manifolds, Proc. Amer. Math. Soc. 102 (1988), 660–666. [Zygmund 1945] Zygmund, A. Smooth functions. Duke Math. J. 12. 47–76.