+ All Categories
Home > Documents > Analytic and Differential geometry Mikhail G. Katz∗

Analytic and Differential geometry Mikhail G. Katz∗

Date post: 01-Mar-2022
Category:
Upload: others
View: 2 times
Download: 0 times
Share this document with a friend
229
Analytic and Differential geometry Mikhail G. Katz Department of Mathematics, Bar Ilan University, Ra- mat Gan 52900 Israel E-mail address : [email protected]
Transcript
Page 1: Analytic and Differential geometry Mikhail G. Katz∗

Analytic and Differential geometry

Mikhail G. Katz∗

Department of Mathematics, Bar Ilan University, Ra-

mat Gan 52900 Israel

E-mail address : [email protected]

Page 2: Analytic and Differential geometry Mikhail G. Katz∗

∗Supported by the Israel Science Foundation (grants no. 620/00-10.0and 84/03).

Abstract. We start with analytic geometry and the theory ofconic sections. Then we treat the classical topics in differentialgeometry such as the geodesic equation and Gaussian curvature.Then we prove Gauss’s theorema egregium and introduce the ab-stract viewpoint of modern differential geometry.

Page 3: Analytic and Differential geometry Mikhail G. Katz∗

Contents

Chapter 1. Analytic geometry 91.1. Course notes 91.2. Circle, isoperimetric inequality 91.3. Sphere 101.4. Relativity theory 111.5. Linear algebra, index notation 111.6. Einstein summation convention 121.7. Quadratic forms, polarisation formula 121.8. Matrix as a linear map 131.9. Symmetrisation and skew-symmetrisation 141.10. Matrix multiplication in index notation 141.11. Two types of indices: summation index and free index 151.12. The inverse matrix, Kronecker delta 16

Chapter 2. Eigenvalues of symmetric matrices, conic sections 192.1. Eigenvalues, symmetry 192.2. Finding an eigenvector of a symmetric matrix 202.3. Trace of product in index notation 222.4. Inner product spaces and self-adjoint operators 232.5. Orthogonal diagonalisation of symmetric matrices 232.6. Classification of conic sections: diagonalisation 242.7. Classification of conics: trichotomy, nondegeneracy 252.8. Characterisation of parabolas 27

Chapter 3. Quadratic surfaces 293.1. Exercise on index notation 293.2. Summary: classification of conics 293.3. Quadratic surfaces 303.4. Examples of quadratic surfaces 313.5. Hessian, minima, maxima, saddle points 323.6. Parametric representation of a curve 343.7. Implicit representation of a curve 343.8. Implicit function theorem 353.9. Unit speed parametrisation 36

3

Page 4: Analytic and Differential geometry Mikhail G. Katz∗

4 CONTENTS

3.10. Geodesic curvature 363.11. Shape operator for curves 37

Chapter 4. Curvature of curves 394.1. Osculating circle of a curve 394.2. Center of curvature, radius of curvature 404.3. Flat Laplacian; Bateman–Reiss operator 404.4. Geodesic curvature for an implicit curve 414.5. Curvature of conic sections 424.6. Curvature of graph of function 43

Chapter 5. From arclength to surfaces 455.1. Existence of arclength 455.2. Arnold’s observation on folding a page 465.3. Regular surface; Jacobian 475.4. Coefficients of first fundamental form of a surface 485.5. Metric coefficients in spherical coordinates 485.6. Tangent plane 49

Chapter 6. First fundamental form, Gamma symbols 516.1. First fundamental form 516.2. Plane and cylinder 516.3. Surfaces of revolution 526.4. From tractrix to pseudosphere 546.5. Chain rule in two variables 546.6. Measuring length of curves on surfaces 546.7. Normal vector, symbols Γkij of a surface 556.8. Gamma symbols 566.9. Basic properties of the Gamma symbols 566.10. Intrinsic nature of the Gamma symbols 576.11. Gamma Symbols for a surface of revolution 58

Chapter 7. From Clairaut’s relation to geodesic equation 617.1. Spherical coordinates 617.2. Great circles, Clairaut’s relation 627.3. Some basic lemmas 637.4. Clairaut’s relation 647.5. Spherical sine law 647.6. Longitudes 667.7. Proof of Clairaut’s relation 677.8. Differential equation of great circle 687.9. Metrics conformal to the flat metric and their Γkij 697.10. Gravitation, ants, and geodesics on a surface 70

Page 5: Analytic and Differential geometry Mikhail G. Katz∗

CONTENTS 5

7.11. Geodesic equation 71

Chapter 8. Geodesic equation, Clairaut 738.1. Equivalence of definitions 738.2. Geodesics on a surface of revolution 748.3. Integration in polar, cylindrical, spherical coordinates 768.4. Measuring area on surfaces 788.5. Directional derivative 788.6. Extending v.f. along surface to an open set in R3 798.7. Differentiating n and N 808.8. Hessian of a function at a critical point 81

Chapter 9. Weingarten map 839.1. Definition of the Weingarten map 839.2. Properties of Weingarten map 849.3. Relation to the Hessian of a function 859.4. Weingarten map of sphere and cylinder 869.5. Coefficients Lij of Weingarten map 879.6. Gaussian curvature 879.7. Second fundamental form 889.8. Geodesics and second fundamental form 899.9. Three formulas for Gaussian curvature 90

Chapter 10. Principal curvatures, minimal surfaces, thm egregium 9510.1. Principal curvatures 9510.2. Mean curvature, minimal surfaces 9610.3. Minimal surfaces in isothermal coordinates 9710.4. Minimality and harmonic functions 9910.5. Introduction to theorema egregium; intrinsic vs extrinsic 10010.6. Riemann’s formula 10110.7. Preliminaries to the theorema egregium 10210.8. An identity involving the Γkij and the Lij 10310.9. The theorema egregium of Gauss 104

Chapter 11. Signed curvature of plane curves 10711.1. Curvature with respect to an arbitrary parameter 10711.2. Jordan curves in the plane 10811.3. Gauss map 10811.4. Curvature expressed in terms of θ(s) 10911.5. Convex Jordan curves 11011.6. Total curvature of a convex Jordan curve 11111.7. Signed curvature kα of Jordan curves in the plane 113

Chapter 12. Laplacian formula K, Hyperbolic metric, Gauss–Bonnet115

Page 6: Analytic and Differential geometry Mikhail G. Katz∗

6 CONTENTS

12.1. Gauss map for nonconvex curves 11512.2. Mean versus Gaussian curvature 11712.3. The Laplace-Beltrami formula for K 11712.4. Hyperbolic metric in upperhalf plane 11912.5. Gaussian curvature as product of signed curvatures 12012.6. Area elements of the surface and of the sphere 12212.7. Gauss map for surfaces in R3 12212.8. Sphere S2 parametrized by Gauss map of M 12312.9. Comparison of two parametrisations 12412.10. Gauss–Bonnet theorem 12412.11. Euler characteristic 126

Chapter 13. Duality in linear algebra, calculus, diff. geometry 12913.1. Duality in linear algebra; 1-forms 12913.2. Dual vector space 13013.3. Derivations 13113.4. Characterisation of derivations 13213.5. Tangent space and cotangent space 13313.6. Constructing bilinear forms out of 1-forms 13413.7. First fundamental form 13513.8. Dual bases in differential geometry 13513.9. Surfaces of revolution in isothermal coordinates 13713.10. More on dual bases 138

Chapter 14. A hyperreal view 14114.1. Successive extensions N, Z, Q, R, ∗R 14114.2. Motivating discussion for infinitesimals 14214.3. Introduction to the transfer principle 14314.4. Transfer principle 14614.5. Orders of magnitude 14714.6. Standard part principle 14914.7. Differentiation 149

Chapter 15. Lattices and tori 15315.1. Circle via the exponential map 15315.2. Lattice, fundamental domain 15415.3. Fundamental domain 15415.4. Lattices in the plane 15515.5. Successive minima of a lattice 15615.6. Gram matrix 15715.7. Sphere and torus as topological surfaces 15715.8. Standard fundamental domain 15815.9. Conformal parameter τ of a lattice 159

Page 7: Analytic and Differential geometry Mikhail G. Katz∗

CONTENTS 7

15.10. Conformal parameter τ of tori of revolution 16115.11. θ-loops and φ-loops on tori of revolution 16215.12. Tori generated by round circles 16215.13. Conformal parameter of tori of revolution, residues 163

Chapter 16. Systole 16516.1. Definition of systole 16516.2. Loewner’s torus inequality 16916.3. Loewner’s inequality with defect 17016.4. Computational formula for the variance 17116.5. An application of the computational formula 17116.6. Conformal invariant σ 17216.7. Fundamental domain and Loewner’s torus inequality 17216.8. Boundary case of equality 173

Chapter 17. Manifolds and global geometry 17517.1. Global geometry of surfaces 17517.2. Definition of manifold 17617.3. Sphere as a manifold 17717.4. Dual bases 17717.5. Jacobian matrix 17817.6. Area of a surface, independence of partition 17817.7. Conformal equivalence 18017.8. Geodesic equation 18117.9. Closed geodesic 18217.10. Existence of closed geodesic 18217.11. Surfaces of constant curvature 18317.12. Real projective plane 18317.13. Simple loops for surfaces of positive curvature 18417.14. Successive minima 18517.15. Flat surfaces 18517.16. Hyperbolic surfaces 18617.17. Hyperbolic plane 18717.18. Loops, simply connected spaces 18817.19. Orientation on loops and surfaces 18917.20. Cycles and boundaries 18917.21. First singular homology group 19117.22. Stable norm in 1-dimensional homology 19117.23. The degree of a map 19217.24. Degree of normal map of an imbedded surface 19317.25. Euler characteristic of an orientable surface 19417.26. Gauss-Bonnet theorem 194

Page 8: Analytic and Differential geometry Mikhail G. Katz∗

8 CONTENTS

17.27. Change of metric exploiting Gaussian curvature 19517.28. Gauss map 19517.29. An identity 19617.30. Stable systole 19717.31. Free loops, based loops, and fundamental group 19717.32. Fundamental groups of surfaces 198

Chapter 18. Pu’s inequality 20118.1. Hopf fibration h 20118.2. Tangent map 20118.3. Riemannian submersion 20218.4. Riemannian submersions 20218.5. Hamilton quaternions 20318.6. Complex structures on the algebra H 204

Chapter 19. Approach using energy-area identity 20519.1. An integral-geometric identity 20519.2. Two proofs of the Loewner inequality 20619.3. Hopf fibration and the Hamilton quaternions 20819.4. Double fibration of SO(3) and integral geometry on S2 20919.5. Proof of Pu’s inequality 21119.6. A table of optimal systolic ratios of surfaces 211

Chapter 20. A primer on surfaces 21320.1. Hyperelliptic involution 21320.2. Hyperelliptic surfaces 21520.3. Ovalless surfaces 21520.4. Katok’s entropy inequality 217

Bibliography 221

Page 9: Analytic and Differential geometry Mikhail G. Katz∗

CHAPTER 1

Analytic geometry

These notes are based on a course taught at Bar Ilan University;see http://u.cs.biu.ac.il/~katzmik/88-526.html

After dealing with classical geometric preliminaries including thetheorema egregium of Gauss, we present new geometric inequalities onRiemann surfaces, as well as their higher dimensional generalisations.

We will first review some familiar objects from classical geometryand try to point out the connection with important themes in modernmathematics.

1.1. Course notes

Remark 1.1.1. There are various chovrot in hebrew circulating butthey contain serious errors.

Remark 1.1.2. In some cases there is a direct relationship betweenthe errors contained in the chovrot and mistakes made on the exam.

Remark 1.1.3. I strongly recommend following these online classnotes. If there are any difficulties with the English I would be happyto provide a clarification.

1.2. Circle, isoperimetric inequality

The familiar unit circle S1 in the plane, defined to be the locus ofthe equation

x2 + y2 = 1

in the (x, y)-plane. The circle solves the isoperimetic problem in theplane. Namely, consider simple (non-self-intersecting) closed curves ofequal perimeter, for instance a polygon.1

Among all such curves, the circle is the curve that encloses thelargest area. In other words, the circle satisfies the boundary case ofequality in the following inequality, known as the isoperimetric inequal-ity. Let L be the length of the Jordan curve and A the area of the finiteregion bounded by the curve.

1metzula

9

Page 10: Analytic and Differential geometry Mikhail G. Katz∗

10 1. ANALYTIC GEOMETRY

Theorem 1.2.1 (Isoperimetric inequality). Every Jordan curve inthe plane satisfies the inequality

(L

)2

− A

π≥ 0,

with equality if and only if the curve is a round circle. 2

1.3. Sphere

The familiar 2-sphere S2 is a surface that can be defined as thecollection of unit vectors in 3-sphace:

S2 ={(x, y, z) ∈ R3 | x2 + y2 + z2 = 1

}.

Definition 1.3.1. A great circle is the intersection of the spherewith a plane passing through the origin.

The equator is an example of a great circle.

Definition 1.3.2. The great circle distance d(p, q) on S2 is the dis-tance measured along the arcs of great circles connecting the points p, q ∈S2:

d(p, q) = arccos(p · q),where p · q is the usual dot product in Euclidean space.

Namely, the distance between a pair of points p, q ∈ S2 is the lengthof the smaller of the two arcs of the great circle passing through p and q.

2The round circle is the subject of Gromov’s filling area conjecture.

Definition 1.2.2. The Riemannian circle of length 2π is a great circle of theunit sphere, equipped with the great-circle distance.

The emphasis is on the fact that the distance is measured along arcs ratherthan chords (straight line intervals).

For all the apparent simplicity of the the Riemannian circle, it turns out thatit is the subject of a still-unsolved conjecture of Gromov’s, namely the filling areaconjecture.

A surface with a single boundary circle will be called a filling of that circle.We now consider fillings of the Riemannian circles such that the ambient distancedoes not diminish the great-circle distance (in particular, filling by the unit disk isnot allowed).

Conjecture 1.2.3 (M. Gromov). Among all fillings of the Riemannian circlesby a surface, the hemisphere is the one of least area.

Page 11: Analytic and Differential geometry Mikhail G. Katz∗

1.5. LINEAR ALGEBRA, INDEX NOTATION 11

1.4. Relativity theory

In relativity theory, one uses a framework similar to classical differ-ential geometry, with a technical difference having to do with the basicquadratic form being used. Nonetheless, some of the key concepts, suchas geodesic and curvature, are common to both approaches.

In the first approximation, one can think of relativity theory as thestudy of 4-manifolds with a choice of a “light cone”3 at every point.

Einstein gave a strong impetus to the development of differentialgeometry, as a tool in studying relativity. We will systematically useEinstein’s summation convention; see Section 1.6.

1.5. Linear algebra, index notation

Let Rn denote the Euclidean n-space. Its vectors will be denoted

v, w ∈ Rn.

Definition 1.5.1. A vectors in Rn is a column vectors v. To obtaina row vector we take the transpose, vt.

Let B ∈ Mn,n(R) be an n by n matrix. There are two ways ofviewing such a matrix, either as a linear map or as a bilinear form(cf. Remarks 1.5.2 and 1.8). Developing suitable notation to capturethis distinction helps simplify differential-geometric formulas down toreadable size, and also to motivate the important distinction betweena vector and a covector.

Remark 1.5.2 (Matrix as a bilinear form B(v, w)). Consider abilinear form

B(v, w) : Rn × Rn → R, (1.5.1)

sending the pair of vectors (v, w) to the real number vtBw. Here vt isthe row vector given by transpose of v. We write

B = (bij)i=1,...,n; j=1,...,n

so that bij is an entry while (bij), with parentheses, denotes the matrix.

Example 1.5.3. In the 2 by 2 case, we have

v =

(v1

v2

), w =

(w1

w2

), B =

(b11 b12b21 b22

).

Then

Bw =

(b11 b12b21 b22

)(w1

w2

)=

(b11w

1 + b12w2

b21w1 + b22w

2

).

3konus ha’or

Page 12: Analytic and Differential geometry Mikhail G. Katz∗

12 1. ANALYTIC GEOMETRY

Now the transpose vt = (v1 v2) is a row vector, so

B(v, w) = vtBw

= (v1 v2)

(b11w

1 + b12w2

b21w1 + b22w

2

)

= b11v1w1 + b12v

1w2 + b21v2w1 + b22v

2w2,

and therefore vtBw =∑2

i=1

∑2j=1 bijv

iwj.

We would like to simplify this notation by suppressing the summa-tion symbols “Σ”, as in Section 1.6.

1.6. Einstein summation convention

The following useful notational device was originally introduced byAlbert Einstein.

Definition 1.6.1. The rule is that whenever a product contains asymbol with a lower index and another symbol with the same upperindex, take summation over this repeated index.

Example 1.6.2. Using this notation, the bilinear form (1.5.1) de-fined by the matrix B can be written as follows:

B(v, w) = bijviwj,

with implied summation over both indices.

When we wish to consider a specific term such as ai times vi ratherthan the sum over a dummy index i, we use the underline notation asfollows.

Definition 1.6.3. The expression

aivi

denotes a specific term with index value i (rather than summation overa dummy index i).

1.7. Quadratic forms, polarisation formula

Definition 1.7.1. Let B be a symmetric matrix. The associ-ated quadratic form is a quadratic form associated with a bilinearform B(v, w) by the following rule:

Q(v) = B(v, v).

Let v = viei where {ei}i=1,...,n = {e1, . . . , en} is the standard basisof Rn.

Page 13: Analytic and Differential geometry Mikhail G. Katz∗

1.8. MATRIX AS A LINEAR MAP 13

Lemma 1.7.2. The associated quadratic form satisfies

Q(v) = bijvivj.

Proof. To compute Q(v), we must introduce an extra index j :

Q(v) = B(v, v) = B(viei, vjej) = B(ei, ej)v

ivj = bijvivj, (1.7.1)

proving the lemma. �

Definition 1.7.3. The polarisation formula is the following for-mula:

B(v, w) =1

4(Q(v + w)−Q(v − w)).

The polarisation formula allows one to reconstruct the bilinear formfrom the quadratic form, whenever the characteristic of the field is not 2(as is the case for R). For an application, see Section 13.6.

1.8. Matrix as a linear map

Given a real matrix B, consider the associated map

BR : Rn → Rn, v 7→ Bv.

In order to distinguish this case from the case of the bilinear form, wewill raise the first index: write B as

B = (bij)i=1,...,n; j=1,...,n

where it is important to s t a g g e r the indices as follows.

Definition 1.8.1. Staggering the indices meaning that we do notput j under i as in

bij ,

but, rather, leave a blank space (in the place were j used to be), as in

bij .

Consider vectors v = (vj)j=1,...,n and w = (wi)i=1,...,n. Then theequation w = Bv can be written as a system of n scalar equations,

wi = bijvj for i = 1, . . . , n

using the Einstein summation convention (here the repeated index is j).

Definition 1.8.2. The formula for the trace Tr(B) = b11 + b22 +· · ·+ bnn in Einstein notation becomes

Tr(B) = bii

(here the repeated index is i).

Page 14: Analytic and Differential geometry Mikhail G. Katz∗

14 1. ANALYTIC GEOMETRY

Remark 1.8.3. If we wish to specify the i-th diagonal coefficientof our matrix B we use the underline notation

bii

as in Definition 1.6.3.

1.9. Symmetrisation and skew-symmetrisation

For the purposes of dealing with symmetrisation and antisymmetri-sation of a matrix, it is convenient to return to the framework of bilinearforms (i.e., both indices are lower indices).

Definition 1.9.1. Let B = (bij). Its symmetric part S is by defi-nition

S =1

2(B +Bt) =

(12(bij + bji)

)i=1,...,nj=1,...,n

,

while the skew-symmetric part A is

A =1

2(B − Bt) =

(12(bij − bji)

)i=1,...,nj=1,...,n

.

Note that we have B = S + A.Another useful notation is that of symmetrisation

b{ij} =1

2(bij + bji) (1.9.1)

and antisymmetrisation

b[ij] =1

2(bij − bji). (1.9.2)

Lemma 1.9.2. A matrix B = (bij) is symmetric if and only if forall indices i and j one has b[ij] = 0.

Proof. We have b[ij] = 12(bij − bji) = 0 since symmetry of B

means bij = bji. �

1.10. Matrix multiplication in index notation

The usual way to define matrix multiplication is as follows. Atriple of matrices A = (aij), B = (bij), and C = (cij) satisfy theproduct relation C = AB if, introducing an additional dummy index k

(cf. formula (1.7.1)), we have the relation cij =∑

k

aikbkj .

Page 15: Analytic and Differential geometry Mikhail G. Katz∗

1.11. TWO TYPES OF INDICES: SUMMATION INDEX AND FREE INDEX 15

Example 1.10.1 (Skew-symmetrisation of matrix product). Bycommutativity of multiplication of real numbers, we have

aikbkj = bkjaik.

Then the coefficients c[ij] of the skew-symmetrisation of the matrix C =AB satisfy

c[ij] =∑

k

bk[jai]k.

Here by definition

bk[jai]k =1

2(bkjaik − bkiajk).

This notational device will be particularly useful in writing downthe theorema egregium (see Section 10.9). Given below are a few ex-amples:

• See Section 6.8, where we will use formulas of type

gmjΓmik + gmiΓ

mjk = 2gm{jΓ

mi}k;

• See Section 9.6 for Li[jLkl];

• See Section 10.8 for Γki[jΓnl]m.

Example 1.10.2. The index notation we have described reflectsthe fact that the natural products of matrices are the ones whichcorrespond to composition of maps. Thus, if A = (aij), B = (bij),

and C = (cij) then the product relation C = AB corresponding to thecomposition of linear maps CR = AR ◦BR simplifies to the relation

cij = aikbkj for all i, j. (1.10.1)

1.11. Two types of indices: summation index and free index

In expressions of type (1.10.1) it is important to distinguish betweentwo types of indices: a free index or an summation index.

Definition 1.11.1. An index appearing both as a subscript and asuperscript is called an summation index. The remaining indices arecalled free.

Thus in formula 1.10.1 the index k is a summation index whereasindices i and j are free indices.

Page 16: Analytic and Differential geometry Mikhail G. Katz∗

16 1. ANALYTIC GEOMETRY

1.12. The inverse matrix, Kronecker delta

Consider a matrix B = (bij).

Definition 1.12.1. The inverse matrix B−1 is the matrix

B−1 = (bij)i=1,...,n; j=1,...,n

where both indices have been raised.

Then the equation B−1B = I becomes

bikbkj = δij (1.12.1)

in Einstein notation with repeated index k.

Remark 1.12.2. In (1.12.1) the index k is a summation indexwhereas i and j are free indices.

Here δ is the Kronecker delta defined as follows.

Definition 1.12.3. The expression

δij =

{1 if i = j0 if i 6= j

is called the Kronecker delta function.

Example 1.12.4. The identity endomorphism I = (δij) by defi-

nition satisfies AI = A = IA for all endomorphisms A = (aij), orequivalently

aijδjk = aik = δija

jk, (1.12.2)

using the Einstein summation convention.

Remark 1.12.5. In expression 1.12.2 the index j is a summationindex whereas indices i and k are free indices.

Example 1.12.6. Let δij be the Kronecker delta function on Rn,where i, j = 1, . . . , n, viewed as a linear transformation Rn → Rn.

(1) Evaluate the expression δijδjk paying attention to which the

summation indices are;(2) Evaluate the expression δijδ

ji paying attention to which the

summation indices are.

Definition 1.12.7. Given a pair of vectors v = viei and w = wjejin R3, their vector product is a vector v ×w ∈ R3 satisfying one of thefollowing two equivalent conditions:

Page 17: Analytic and Differential geometry Mikhail G. Katz∗

1.12. THE INVERSE MATRIX, KRONECKER DELTA 17

(1) we have v × w = det

e1 e2 e3v1 v2 v3

w1 w2 w3

, in other words,

v × w = (v2w3 − v3w2)e1 − (v1w3 − v3w1)e2 + (v1w2 − v2w1)e3

= 2(v[2w3]e1 − v[1w3]e2 + v[1w2]e3

).

(2) the vector v × w is perpendicular to both v and w, of lengthequal to the area of the parallelogram spanned by the two vec-tors, and furthermore satisfying the right hand rule,4 mean-ing that the 3 by 3 matrix formed by the three vectors v, w,and v × w has positive determinant.

Remark 1.12.8. We have an identity

a× (b× c) = (a · c)b− (a · b)c (1.12.3)

for every triple of vectors a, b, c in R3. Note that the expression (1.12.3)vanishes if a is perpendicular to both b and c.

4Klal yad yamin

Page 18: Analytic and Differential geometry Mikhail G. Katz∗
Page 19: Analytic and Differential geometry Mikhail G. Katz∗

CHAPTER 2

Eigenvalues of symmetric matrices, conic sections

2.1. Eigenvalues, symmetry

Properly understanding surface theory and related key conceptssuch as theWeingarten map (see Section 9.1) depends on linear-algebraicbackground related to diagonalisation of symmetric matrices or, moregenerally, selfadjoint endomorphisms.

In general, a real matrix may have no real eigenvector or eigenvalue.

Example 2.1.1. The matrix of a 90 degree rotation in the plane,(0 −11 0

)

does not have a real eigenvector for obvious geometric reasons (nodirection is preserved but rather rotated).

In this section we prove the existence of a real eigenvector (andhence, a real eigenvalue) for a real symmetric matrix.

This fact has important ramifications in surface theory, since thevarious notions of curvature of a surface are defined in terms of theeigenvalues of a certain selfadjoint operator (see sections 8.5 and 10.1)which is an invariant formulation of the notion of a symmetric matrix.Let

I =

1 0

. . .0 1

be the (n, n) identity matrix. Thus

I = (δij) i=1,...,nj=1,...,n

where δij is the Kronecker delta. Let B be an (n, n)-matrix.

Definition 2.1.2. A real number λ is called an eigenvalue of B if

det(B − λI) = 0.

Theorem 2.1.3. If λ ∈ R is an eigenvalue of B, then there is avector v ∈ Rn, v 6= 0, such that

Bv = λv. (2.1.1)

19

Page 20: Analytic and Differential geometry Mikhail G. Katz∗

20 2. EIGENVALUES OF SYMMETRIC MATRICES, CONIC SECTIONS

The proof is one of the first results in linear algebra and is notreproduced here.

Definition 2.1.4. A nonzero vector satisfying (2.1.1) is called aneigenvector belonging to λ.

Definition 2.1.5. The Euclidean inner product of vectors v, w isdefined by

〈v, w〉 = v1w1 + · · ·+ vnwn =n∑

i=1

viwi.

Recall that all of our vectors are column vectors.

Lemma 2.1.6. The inner product can be expressed in terms of ma-trix multiplication in the following fashion:

〈v, w〉 = vt w.

Recall the following theorem from basic linear algebra.

Theorem 2.1.7. The transpose has the following property:

(AB)t = BtAt.

Definition 2.1.8. A square matrix A is called symmetric of At =A.

Lemma 2.1.9. A real matrix B is symmetric if and only if forall v, w ∈ Rn, one has

〈Bv,w〉 = 〈v, Bw〉.Proof. We have

〈Bv,w〉 = (Bv)tw = vtBtw = 〈v, Btw〉 = 〈v, Bw〉,proving the =⇒ direction. Conversely, if v = ri and w = ej then〈Bv,w〉 = bij while 〈v, Bw〉 = bji, proving the ⇐= direction. �

2.2. Finding an eigenvector of a symmetric matrix

Theorem 2.2.1. Every real symmetric matrix possesses a real eigen-vector.

We will give two proofs of this important theorem. The first proofis simpler, more algebraic, and passes via complexification. The secondproof is more geometric.

Page 21: Analytic and Differential geometry Mikhail G. Katz∗

2.2. FINDING AN EIGENVECTOR OF A SYMMETRIC MATRIX 21

First proof. Let n ≥ 1, and let B be an n × n real symmetricmatrix, B ∈Mn,n(R). As such, it defines a linear map

BR : Rn → Rn

sending a vector v ∈ Rn to the vector Bv. We now use the fieldextension R ⊆ C and view B as a complex matrix

B ∈Mn,n(C).

Then the matrix B defines a complex linear map

BC : Cn → Cn

sending v ∈ Cn to Bv. The characteristic polynomial of the endomor-phism BC is a polynomial of positive degree n > 0 and therefore hasa root λ ∈ C by the fundamental theorem of algebra. Let v ∈ Cn bean associated eigenvector. Let 〈 , 〉 be the standard Hermitian innerproduct in Cn. Recall that the Hermitian inner product is linear in onevariable and skew-linear in the other. Thus, we have1

〈z, w〉 =n∑

i

ziwi

Then 〈Bv, v〉 = 〈v, Bv〉 since B = Bt is real symmetric. Hence

〈λv, v〉 = 〈v, λv〉 = λ〈v, v〉Since the Hermitian inner product is skew-symmetric in the secondvariable. Therefore λ〈v, v〉 = λ〈v, v〉 so that λ = λ, i.e., λ is a realeigenvalue, as required. �

The second proof is somewhat longer but has the advantage of beingmore geometric, as well as more concrete in the choice of the vector inquestion.

We present the alternative proof here.

Second Proof. Let S ⊂ Rn be the unit sphere S = {v ∈ Rn |‖v‖ = 1}. Given a symmetric matrix B, define a function f : S → Ras the restriction to S of the function, also denoted f , given by f(v) =〈v, Bv〉. Let v0 be a maximum of f restricted to S. Let V ⊥

0 ⊆ Rn bethe orthogonal complement of the line spanned by v0. Let w ∈ V ⊥

0 .Consider the curve v0+ tw, t ≥ 0 (see also a different choice of curve in

Remark 2.2.2 below). Thend

dt

∣∣∣∣t=0

f(v0+tw) = 0 since v0 is a maximum

1Alternatively, some texts adopt the convention 〈z, w〉 =∑ni=1 ziwi.

Page 22: Analytic and Differential geometry Mikhail G. Katz∗

22 2. EIGENVALUES OF SYMMETRIC MATRICES, CONIC SECTIONS

and w is tangent to the sphere. Now

d

dt

∣∣∣∣∣0

f(v0 + tw) =d

dt

∣∣∣∣0

〈v0 + tw,B(v0 + tw)〉

=d

dt

∣∣∣∣0

(〈v0, Bv〉+ t〈v0, Bw〉+ t〈w,Bv0〉+ t2〈w,Bv〉)

= 〈v0, Bw〉+ 〈w,Bv0〉= 〈Bw, v0〉+ 〈Bv0, w〉= 〈Btv0, w〉+ 〈Bv0, w〉= 〈(Bt +B)v0, w〉= 2〈Bv0, w〉 by symmetry of B.

Thus 〈Bv0, w〉 = 0 for all w ∈ V ⊥0 . Hence Bv0 is proportional to v0

and so v0 is an eigenvector of B. �

Remark 2.2.2. Our calculation used the curve v0+tw which, whiletangent to S at v0 (see section 7.2), does not lie on S. If one prefers,one can use instead the curve (cos t)v0 + (sin t)w lying on S. Then

d

dt

∣∣∣∣∣t=0

〈(cos t)v0+(sin t)w,B((cos t)v0+(sin t)w)〉 = · · · = 〈(Bt+B)v0, w〉

and one argues by symmetry as before.

2.3. Trace of product in index notation

First we reinforce the material on index notation and Einstein sum-mation convention; see Section 1.6. The following result is importantin its own right. We reproduce it here because its proof is a goodillustration of the uses of the Einstein index notation.

Theorem 2.3.1. Let A and B be square n × n matrices. Thentr(AB) = tr(BA).

Proof. Let A = (aij) and B = (bij). Then

tr(AB) = tr(aikbkj) = aikb

ki

by definition of trace (see Definition 1.8.2). Meanwhile,

tr(BA) = tr(bkiaij) = bkia

ik = aikb

ki = tr(AB),

proving the theorem. �

Exercise 2.3.2. A matrix A is called idempotent if A2 = A. Writethe idempotency condition in indices with Einstein summation conven-tion (without Σ’s).

Page 23: Analytic and Differential geometry Mikhail G. Katz∗

2.5. ORTHOGONAL DIAGONALISATION OF SYMMETRIC MATRICES 23

Exercise 2.3.3. Matrices A and B are similar if there exists aninvertible matrix P such that AP − PB = 0. Write the similaritycondition in indices, as the vanishing of each (i, j)th coefficient of thedifference AP − PB.

2.4. Inner product spaces and self-adjoint operators

In Section 2.2 we worked with real matrices and showed that thesymmetry of a matrix guarantees the existence of a real eigenvector.In a more general situation where a basis is not available, a similarstatement holds for a special type of endomorphism of a real vectorspace.

Definition 2.4.1. Let (V, 〈 , 〉) be a real inner product space. Anendomorphism B : V → V is called selfadjoint if one has

〈Bv,w〉 = 〈v, Bw〉 ∀v, w ∈ V. (2.4.1)

Corollary 2.4.2. Every selfadjoint endomorphism of a real innerproduct space admits a real eigenvector.

Proof. The selfadjointness was the relevant property in the proofof Theorem 2.2.1. �

2.5. Orthogonal diagonalisation of symmetric matrices

Lemma 2.5.1. Let E : V → V be a selfadjoint endomorphism ofan inner product space V and let V1 ⊆ V be an E-invariant subspace.Then the orthogonal complement of V1 in V is also E-invariant.

Proof. Let v1 ∈ V1 and let w be orthogonal to V1. Then

〈E(w), v1〉 = 〈w,Ev1〉 = 〈w, λ1v1〉 = λ1〈w, v1〉 = 0

and therefore E(w) is also orthogonal to v1. �

Theorem 2.5.2. Every symmetric matrix can be orthogonally di-agonalized.

Proof. A symmetric n × n matrix S defines a selfadjoint endo-morphism SR of the real inner product space V = Rn given by

SR : V → V, v 7→ Sv.

By Corollary 2.4.2, every selfadjoint endomorphism has a real eigen-vector v1 ∈ V , which we can assume to be a unit vector:

|v1| = 1.

Let λ1 ∈ R be its eigenvalue. Now we let V1 = V and set

V2 ⊆ V1

Page 24: Analytic and Differential geometry Mikhail G. Katz∗

24 2. EIGENVALUES OF SYMMETRIC MATRICES, CONIC SECTIONS

be the orthogonal complement of the line Rv1 ⊆ V1. Thus we have anorthogonal decomposition

V1 = Rv1 + V2.

by Lemma 2.5.1, V2 is invariant under SR. The restriction of SR to V2is still selfadjoint by inheriting the property (2.4.1). Namely, sinceproperty (2.4.1) holds for all vectors v, w ∈ V , it still holds if thesevectors are restricted to vary in a subspace V2 ⊆ V , i.e.,

〈SRv, w〉 = 〈v, SRw〉 ∀v, w ∈ V2. (2.5.1)

Therefore we can find an eigenvector v2 ∈ V2 with eigenvalue λ2 ∈R. Arguing inductively, we obtain an orthonormal basis consisting ofeigenvectors v1, . . . , vn ∈ V with eigenvalues λ1, . . . , λn ∈ R. Let

P = [v1 . . . vn]

be the orthogonal n × n matrix whose columns are the vectors vi, sothat we have

P−1 = P t. (2.5.2)

Consider the diagonal matrix Λ = diag(λ1, . . . , λn). By construction,we have

S = PΛP t

from (2.5.2), or equivalently,

SP = PΛ.

Indeed, to verify the relation SP = PΛ, note that both sides are equalto the square matrix [λ1v1 λ2v2 . . . λnvn]. �

2.6. Classification of conic sections: diagonalisation

Definition 2.6.1. A conic section2 (or conic for short) in the planeis by definition a curve defined by the following equation in the (x, y)-plane:

ax2 + 2bxy + cy2 + dx+ ey + f = 0, a, b, c, d, e, f ∈ R. (2.6.1)

Here we chose the coefficient of the xy term to be 2b rather than b

so as to simplify formulas like (2.6.2) below. Let X =

(xy

)and let

S =

(a bb c

). (2.6.2)

ThenX tSX = ax2 + 2bxy + cy2.

2chatach charut

Page 25: Analytic and Differential geometry Mikhail G. Katz∗

2.7. CLASSIFICATION OF CONICS: TRICHOTOMY, NONDEGENERACY 25

Theorem 2.6.2. Up to an orthogonal transformation, every conicsection can be written in a “diagonal” form

λ1x′2 + λ2y

′2 + d′x′ + e′y′ + f = 0, (2.6.3)

where the coefficients λ1 and λ2 are the eigenvalues of the matrix S.

Proof. Consider the row vector

T =(d e

).

Then TX = dx+ ey. Thus equation (2.6.1) becomes

X tSX + TX + f = 0. (2.6.4)

We now apply Theorem 2.5.2 to orthogonally diagonalize S to ob-tain S = PΛP t. Substituting this into (2.6.4) yields

X tPΛP tX + TX + f = 0.

We setX ′ = P tX . Then X = PX ′ since P is orthogonal. Furthermore,we have

(X ′)t = (P tX)t = (X t)(P t)t = (X t)P.

Hence we obtain

(X ′)tΛX ′ + TPX ′ + f = 0.

Letting T ′ = TP , we obtain

(X ′)tΛX ′ + T ′X ′ + f = 0,

where Λ is the diagonal matrix

Λ =

(λ1 00 λ2

).

Letting x′ and y′ be the components of X ′, i.e. X ′ =

(x′

y′

), we ob-

tain (2.6.3), as required. �

To obtain more precise information about the conic, we need tospecify certain nondegeneracy conditions, as discussed in Section 2.7.

2.7. Classification of conics: trichotomy, nondegeneracy

We apply the diagonalisation result of Section 2.6 to classify conicsections into three types (under suitable nondegeneracy conditions):ellipse, parabola, hyperbola.

Page 26: Analytic and Differential geometry Mikhail G. Katz∗

26 2. EIGENVALUES OF SYMMETRIC MATRICES, CONIC SECTIONS

Theorem 2.7.1. Suppose det(S) 6= 0, i.e., S is invertible. Then,up to an orthogonal transformation and a translation, the conic sectioncan be written in the form

λ1(x′′)2 + λ2(y

′′)2 + f ′′ = 0 (2.7.1)

(note that the constant term is changed).

Proof. If the determinant is nonzero then both eigenvalues λi arenonzero. The term d′x′ in (2.6.3) can be absorbed into the quadraticterm λ1x

′2 by completing the square as follows:

λ1x′2 + d′x′ = λ1

(x′2 + 2

d′

2λ1x′)

= λ1

(x′2 + 2

d′

2λ1x′ +

(d′

2λ1

)2)

− λ1

(d′

2λ1

)2

= λ1

(x′ +

d′

2λ1

)2

− d′2

4λ1,

and we set

x′′ = x′ +d′

2λ1.

Similarly e′y′ can be absorbed into λ2y′2. Geometrically this corre-

sponds to a translation along the axes x′ and y′, proving the theo-rem. �

Definition 2.7.2. A conic section of type (2.7.1) is called a hy-perbola if λ1λ2 < 0, provided the following nondegeneracy condition issatisfied: the constant f ′′ in equation (2.7.1) is nonzero.

Corollary 2.7.3 (Degenerate case). If det(S) < 0 and the con-stant f ′′ is zero, then instead of a hyperbola, the solution set is a de-generate conic given by a pair of transverse lines.

Definition 2.7.4. A conic section is called an ellipse if λ1λ2 > 0,provided the following nondegeneracy condition is satisfied: the con-stant f ′′ in equation (2.7.1) is nonzero and has the opposite sign ascompared to the sign of λ1.

See Example 11.6.6 for an application.

Corollary 2.7.5 (Degenerate case). If det(S) > 0 and the con-stant f ′′ is zero, then the ellipse degenerates to a single point x′′ = y′′ =0. If the constant f ′′ 6= 0 has the same sign as λ1 then the solution setis empty.

Page 27: Analytic and Differential geometry Mikhail G. Katz∗

2.8. CHARACTERISATION OF PARABOLAS 27

To summarize our conclusions for the ellipse and the hyperbola,note that since the determinant of the matrix S is the product of itseigenvalues:

ad− bc = λ1λ2,

we obtain the following corollary.

Corollary 2.7.6. If the conic ax2 +2bxy+ cy2+ dx+ ey+ f = 0is an ellipse then ac − b2 > 0. If ac− b2 > 0 and the solution locus isneither empty nor a single point, then it is an ellipse.

Corollary 2.7.7. If the conic ax2 +2bxy+ cy2+ dx+ ey+ f = 0is a hyperbola then ac− b2 < 0. If ac− b2 < 0 and the solution locus isnot a pair of transverse lines, then the conic is a hyperbola.

2.8. Characterisation of parabolas

If det(S) = 0 then one cannot eliminate the linear terms by com-pleting the square as in the previous cases. Therefore we continueworking with the equation

λ1x′2 + λ2y

′2 + d′x′ + e′y′ + f = 0 (2.8.1)

from (2.6.3) (note that diagonalizing S does not depend on S beinginvertible).

Definition 2.8.1. The conic is a parabola if the following two con-ditions are satisfied by the coefficients in (2.8.1):

(1) the matrix S is of rank 1 (equivalently, λ1λ2 = 0 and λ1 6= λ2);(2) if λ1 = 0 then d′ 6= 0; if λ2 = 0 then e′ 6= 0.

Corollary 2.8.2 (Degenerate case). Suppose S is of rank 1 andafter absorbing the linear term in (2.8.1) the equation becomes λ1x

′′2+f ′′ = 0 or λ2y

′′2 + f ′′ = 0 then the solution set is either empty, a line,or a pair of parallel lines.

Example 2.8.3. The equation x2 + 1 = 0 has an empty solutionset in the (x, y)-plane. The equation x2 = 0 has as solution set a singleline, namely the y-axis. The equation x2 − 1 = 0 has as solution set apair of parallel lines.

Page 28: Analytic and Differential geometry Mikhail G. Katz∗
Page 29: Analytic and Differential geometry Mikhail G. Katz∗

CHAPTER 3

Quadratic surfaces

3.1. Exercise on index notation

Theorem 3.1.1. Every 2×2 matrix A = (aij) satisfies the identity

aikakj + q δij = akka

ij , (3.1.1)

where q = det(A).

Note that indices i and j and free indices, whereas k is a summationindex (see Section 1.11).

Proof. We will use the Cayley-Hamilton theorem which assertsthat pA(A) = 0 where pA(λ) is the characteristic polynomial of A. Inthe rank 2 case, we have pA(λ) = λ2 − (trA)λ + det(A), and thereforewe obtain

A2 − (trA)A+ det(A)I = 0,

which in index notation gives aikakj−akkaij+q δij = 0. This is equivalent

to (3.1.1). �

Example 3.1.2 (Exercise on index notation). For a matrix A ofsize 3 × 3, the characteristic polynomial pA(λ) has the form pA(λ) =λ3 − Tr(A)λ2 + s(A)λ − q(A)λ0. Here q(A) = det(A) and s(A) =λ1λ2 + λ1λ3 + λ2λ3, where λi are the eigenvalues of A. By Cayley-Hamilton theorem, we have

pA(A) = 0. (3.1.2)

Express the equation (3.1.2) in index notation.

3.2. Summary: classification of conics

The analysis of the cases presented in Sections 2.7 and 2.8 resultsin the following classification.

Theorem 3.2.1. Assume not all of a, b, c are 0. A real conic ax2+2bxy + cz2 + dx + ey + f = 0 is represented, up to congruence andtranslation, by one of the following possible sets:

(1) the empty set ∅;(2) a single point;

29

Page 30: Analytic and Differential geometry Mikhail G. Katz∗

30 3. QUADRATIC SURFACES

(3) union of a pair of transverse lines;(4) a single line or a pair of parallel lines;(5) ellipse;(6) parabola;(7) hyperbola.

The first four cases are known as degenerate cases. A parabola canoccur only if ac− b2 = 0. An ellipse can occur only if ac− b2 > 0. Ahyperbola can occur only if ac− b2 < 0.

3.3. Quadratic surfaces

Definition 3.3.1. A quadratic surface in R3 is the locus of pointssatisfying the equation

ax2 + 2bxy + cy2 + 2dxz + fz2 + 2gyz+ hx+ iy+ jz + k = 0, (3.3.1)

where a, b, c, d, f, g, h, i, j, k ∈ R.

For quadratic surfaces, a complete classification is too detailed to betreated here, but we will consider a few central features and examples.

To bring this to standard form, we apply an orthogonal diagonali-sation procedure similar to that employed in Section (2.7). Thus, wedefine matrices S, X , and D by setting

S =

a b db c gd g f

, X =

xyz

, T =

(h i j

),

so that the quadratic part of (3.3.1) becomes X tSX , and the linearpart becomes DX . Then equation (3.3.1) takes the form

X tSX +DX + k = 0.

Theorem 3.3.2. Orthogonally diagonalizing S, we conclude thatthe general equation of a quadratic surface can be simplified to

ax2 + by2 + cz2 + dx+ fy + gz + h = 0, (3.3.2)

with new variables x, y, z and new coefficients a, b, c, d, f, g, h ∈ R.

The classification of quadratic surfaces is more involved than thecase of curves, and will not be pursued here. However, we point outsome important special cases.

Definition 3.3.3. A quadratic surface is called an ellipsoid if thecoefficients a, b, c in (3.3.2) are nonzero and have the same sign, andmoreover the solution locus is neither a single point nor the empty set.

Page 31: Analytic and Differential geometry Mikhail G. Katz∗

3.4. EXAMPLES OF QUADRATIC SURFACES 31

Remark 3.3.4. The nondegeneracy condition can be ensured byassuming that the linear terms in (3.3.2) all vanish, and the constantterm g has the opposite sign as compared to the sign of, say, the coef-ficient a.

3.4. Examples of quadratic surfaces

Additional special cases of quadratic surfaces are the following. It isimportant to think through each of these examples as they will provideimportant illustrations of the behavior of the Gaussian curvature (tobe introduced in Section 9.6) of surfaces.

Example 3.4.1. The (elliptic) paraboloid z = ax2 + by2, ab > 0.The Gaussian curvature is positive at each point.

Example 3.4.2. The hyperbolic paraboloid z = ax2 − by2, ab > 0.The Gaussian curvature is negative at each point.

Theorem 3.4.3. Consider the surface M = {(x, y, z) ∈ R3 : z =ax2 + by2}. If a and b have the same sign then the surface is a pa-raboloid. If a and b have opposite sign then the surface is a hyperbolicparaboloid.

This is immediate from the discussion above.

Example 3.4.4. The hyperboloid of one sheet1 with equation z2 =x2 + y2 − 1. The Gaussian curvature is negative at each point.

Example 3.4.5. The hyperboloid of two sheets2 with equation z2 =x2 + y2 + 1. The Gaussian curvature is positive at each point.

3.4.1. Jacobi’s criterion. This section is optional. The type of qua-dratic surface one obtains depends critically on the signs of the eigenvaluesof the matrix S. The signs of the eigenvalues can be determined withoutdiagonalisation by means of Jacobi’s criterion. Given a matrix A over afield F , let ∆k denote the k × k upper-left block, called a principal minor.Matrices A and B are equivalent if they are congruent (rather than similar),meaning that B = P tAP (rather than by conjugation P−1AP ). Equivalentmatrices don’t have the same eigenvalues (unlike similar matrices). Here Bis of course not assumed to be orthogonal.

Theorem 3.4.6 (Jacobi). Let A ∈ Mn(F ) be a symmetric matrix, andassume det(∆k) 6= 0 for k = 1, . . . , n. Then A is equivalent to the matrix

diag(

1∆1, ∆1

det∆2, . . . , det∆n−1

det∆n

).

1chad-yeriati2du-yeriati

Page 32: Analytic and Differential geometry Mikhail G. Katz∗

32 3. QUADRATIC SURFACES

For example, for a 2 × 2 symmetric matrix

(a bb d

)Jacobi’s criterion

affirms the equivalence to

(1a 00 a

ad−b2

).

Proof. Take the vector bk = ∆−1k ek ∈ F k of length k, and pad it with

zeros up to length n. Consider the matrix B = (bij) whose column vectorsare b1, . . . , bn. By Cramer’s formula, the diagonal coefficients of B sat-

isfy bkk = det

(∆k−1 00 1/det∆k

)= det∆k−1/∆k, so det(B) =

∏nk=1 bkk =

1/det(A) 6= 0. Compute that BtAB is lower triangular with diagonalb11, . . . , bkk. Being symmetric, it is diagonal. �

If some minor ∆k is not invertible, then A cannot be definite. Applyingthis result in the case of a real symmetric matrix, we obtain the followingcorollary. Let A be symmetric. Then A is positive definite if and only if alldet(∆k) > 0.

Define minors in general (choose rows and columns i1, . . . , it). Permutingrows and columns, we obtain the following corollary.

Corollary 3.4.7. Let A be symmetric positive definite matrix. Thenall “diagonal” minors are positive definite (and in particular have positivedeterminants).

An application of this is determining whether or not a quadratic surfaceis an ellipsoid, without having to orthogonally diagonalize the matrix ofcoefficients, as we will see below (care has to be taken to show that thequadratic surface is nondegenerate). For example, let us determine whetheror not the quadratic surface

x2 + xy + y2 + xz + z2 + yz + x+ y + z − 2 = 0 (3.4.1)

is an ellipsoid. To solve the problem, we first construct the corresponding

matrix S =

1 1/2 1/21/2 1 1/21/2 1/2 1

and then calculate the principal minors ∆1 =

1, ∆2 = 1 · 1− 12 · 1

2 = 3/4, and

∆3 = 1 +1

8+

1

8− 1

4− 1

4− 1

4=

1

2.

Thus all principal minors are positive and therefore the surface is an ellip-

soid, provided we can show it is nondegenerate. To check nondegeneracy,

notice that (3.4.1) has at least two distinct solutions: (x, y, z) = (1, 0, 0)

and (x, y, z) = (0, 1, 0). Therefore it is a nondegenerate ellipsoid.

3.5. Hessian, minima, maxima, saddle points

Consider a C2-smooth function f(u1, . . . , un) of n variables.

Page 33: Analytic and Differential geometry Mikhail G. Katz∗

3.5. HESSIAN, MINIMA, MAXIMA, SADDLE POINTS 33

Definition 3.5.1. The gradient of f at a point p = (u1, . . . , un) isthe vector

∇f(p) =

∂f∂u1∂f∂u2...∂f∂un

Definition 3.5.2. A critical point p of f is a point satisfying

∇f(p) = 0,

i.e. ∂f∂ui

(p) = 0 for all i = 1, . . . , n.

We now consider the second partial derivatives of f , denoted fij =∂2f

∂ui∂uj.

Definition 3.5.3. the Hessian matrix of f

Hf = (fij)i=1,...,n; j=1,...,n

is the n× n matrix of second partials.

The antisymmetrisation notation was defined in (1.9.2). We canthen formulate what is known as Schwarz’s theorem or Clairaut’s the-orem.

Theorem 3.5.4 (equality of mixed partials). The following threestatements are equivalent and true:

(1) the Hessian Hf is a symmetric matrix;(2) in terms of the antisymmetrisation notation, f[ij] = 0;(3) we have fij = fji for all i, j.

Example 3.5.5 (maxima, minima, saddle points). Let n = 2. Thenthe sign of the determinant

det(Hf) =∂2f

∂u1∂u1∂2f

∂u2∂u2−(

∂2f

∂u1∂u2

)2

at a critical point has geometric significance. Namely, if

det(Hf(p)) > 0,

then p is a local maximum or a local minimum. If det(Hf(p)) < 0,then p is a saddle point.3

Example 3.5.6 (Quadratic surfaces). Quadratic surfaces are a richsource of examples.

3Nekudat ukaf

Page 34: Analytic and Differential geometry Mikhail G. Katz∗

34 3. QUADRATIC SURFACES

(1) The origin is a critical point for the function whose graph isthe paraboloid z = x2 + y2. In the case of the paraboloid thecritical point is a minimum.

(2) Similar remarks apply to the top sheet of the hyperboloid of

two sheets, namely z =√x2 + y2 + 1, where we also get a

minimum.(3) The origin is a critical point for the function whose graph is

the hyperbolic paraboloid z = x2 − y2. In the case of thehyperbolic paraboloid the critical point is a saddle point.

In addition to the sign, the value of Hf(p) also has geometric sig-nificance, expressed by the following theorem.

Theorem 3.5.7. Let p ∈ R2 be a critical point of f . Then the valueof det(Hf(p)) is precisely the Gaussian curvature at (p, f(p)) ∈ R3 ofthe surface given by the graph of f in R3.

See Definition 9.6.1 for more details. We will return to surfaces inSection 5.2.

3.6. Parametric representation of a curve

There are two main ways of representing a curve in the plane: para-metric and implicit.

A curve in the plane can be represented by a pair of coordinatesevolving as a function of time t, called the parameter:

α(t) = (α1(t), α2(t)), t ∈ [a, b],

so that x(t) = α1(t) and y(t) = α2(t). Thus a curve can be viewed asa map

α : [a, b] → R2. (3.6.1)

Let C be the image of the map (3.6.1). Then C is the geometric curveindependent of parametrisation. Thus, changing the parametrisationby setting t = t(s) and replacing α by a new curve β(s) = α(t(s))preserves the geometric curve.

Definition 3.6.1. A parametrisation α(t) is called regular if forall t one has α′(t) 6= 0.

3.7. Implicit representation of a curve

A curve in the (x, y)-plane can also be represented implicitly as thesolution set of an equation

F (x, y) = 0,

where F is a function always assumed sufficiently smooth.

Page 35: Analytic and Differential geometry Mikhail G. Katz∗

3.8. IMPLICIT FUNCTION THEOREM 35

Definition 3.7.1. The curve CF ⊆ R2 is the locus of the equation

CF = {(x, y) : F (x, y) = 0}.Example 3.7.2. A circle of radius r corresponds to the choice of

the function F (x, y) = x2 + y2 − r2.

Further examples are given below.

(1) The function F (x, y) = y − x2 defines a parabola.(2) The function F (x, y) = xy − 1 defines a hyperbola.(3) The function F (x, y) = x2 − y2 − 1 also defines a hyperbola.

Remark 3.7.3. In each of these cases, it is easy to find a parametri-sation (at least of a part of the curve), by solving the equation for oneof the variables. Thus, in the case of the circle, we choose the positivesquare root to obtain y =

√r2 − x2, giving a parametrisation of the

upperhalf circle by means of the pair of formulas

α1(t) = t, α2(t) =√r2 − t2.

Note this is not all of the curve CF .

Unlike the above examples, typically it is difficult to find an explicitparametrisation for a curve defined by an implicit equation F (x, y) = 0.Locally one can always find one in theory under a suitable nondegen-eracy condition, expressed by the implicit function theorem.

3.8. Implicit function theorem

Theorem 3.8.1 (implicit function theorem). Let F (x, y) be a smoothfunction. Suppose the gradient of F does not vanish at a specificpoint p = (x, y) ∈ CF , in other words

∇F (p) 6= 0.

Then there exists a regular parametrisation (α1(t), α2(t)) of the curve CFin a suitable neighborhood of p.

A useful special case is the following result.

Theorem 3.8.2 (implicit function theorem: special case). Let F (x, y)be a smooth function, and suppose that the partial derivative with re-spect to y satisfies

∂F

∂y(p) 6= 0.

Then there exists a parametrisation y = α2(x), in other words α(t) =(t, α2(t)) of the curve CF in a suitable neighborhood of p.

Page 36: Analytic and Differential geometry Mikhail G. Katz∗

36 3. QUADRATIC SURFACES

Example 3.8.3. In the case of the circle x2+y2 = r2, the point (r, 0)on the x-axis fails to satisfy the hypothesis of Theorem 3.8.2. Thecurve cannot be represented by a differentiable function y = y(x) in aneighborhood of this point due to the problem of a vertical tangent.

3.9. Unit speed parametrisation

We review the standard calculus topic of the curvature of a curve.This is indispensable to understanding the principal curvatures of asurface (cf. Theorem 9.8.1 and Theorem 10.1.4).

Consider a parametrized curve α(t) = (α1(t), α2(t)) in the plane.Denote by C the underlying geometric curve, i.e., the image of α:

C ={(x, y) ∈ R2 : (∃t)

(x = α1(t), y = α2(t)

)}.

Definition 3.9.1. We say α = α(t) is a unit speed curve if∣∣dαdt

∣∣ = 1,

i.e.(dα1

dt

)2+(dα2

dt

)2= 1 for all t ∈ [a, b].

Definition 3.9.2. The parameter of a unit speed parametrisation,usually denoted s, is called arclength.

Example 3.9.3. Let r > 0. Then the curve

α(s) =(r cos

s

r, r sin

s

r

)

is a unit speed parametrisation of the circle of radius r. Indeed, we

have∣∣dαds

∣∣ =√(

r 1r

(− sin s

r

))2+(r 1rcos s

r

)2=√sin2 s

r+ cos2 s

r= 1.

3.10. Geodesic curvature

Our main interest will be in space curves, when one cannot in gen-eral assign a sign to the curvature. Therefore in the definition below wedo not concern ourselves with the sign of the curvature of plane curves,either. In this section, only local properties of curvature of curves willbe studied. A global result on the curvature of curves may be found inSection 11.6.

Definition 3.10.1. The (geodesic) curvature function kα(s) ≥ 0of a unit speed curve α(s) is defined by setting

kα(s) =

∣∣∣∣d2α

ds2

∣∣∣∣. (3.10.1)

Theorem 3.10.2. The curvature of the circle of radius r is 1rat

every point of the circle.

Page 37: Analytic and Differential geometry Mikhail G. Katz∗

3.11. SHAPE OPERATOR FOR CURVES 37

Proof. With the above parametrisation, we have

d2α

ds2=

(r1

r2

(− cos

s

r

), r

1

r2

(− sin

s

r

))

at s = 0, and so kα =∣∣1r

(− cos s

r,− sin s

r

)∣∣ = 1r. �

Note that in this case, the curvature is independent of the point,i.e. is a constant function of s.

In Section 11.1, we will give a formula for curvature with respectto an arbitrary parametrisation (not necessarily arclength).

In Section 11.3, the curvature will be expressed in terms of theangle θ formed by the tangent vector with the positive x-axis.

3.11. Shape operator for curves

Consider a regular curve α(s) parametrized by arclength s.

Definition 3.11.1. The unit tangent vector of α is the vectorv(s) = α′(s), written as v(s) = (x(s), y(s)). The unit normal vectorto the curve is the vector n(s) = (−y(s), x(s)) (alternatively n(s) =(y(s),−x(s))).

Lemma 3.11.2. With respect to the arclength parameter s, we havekα(s) = | 1

ydxds|.

Proof. Since x2 + y2 = 1 we obtain xdxds

+ y dyds

= 0. Therefore weobtain

α′′(s) = (dx

ds,dy

ds) = (

dx

ds,−dx

ds

x

y) =

dx

ds(1,−x

y) =

1

y

dx

ds(y,−x) = 1

y

dx

dsn

(up to sign), proving the lemma. �

Proposition 3.11.3. Differentiating the normal vector n along thecurve gives a vector proportional to v(s) (i.e., a tangent vector) withcoefficient of proportionality up to sign given by the geodesic curvaturekα(s):

d

dsn(s) = ±kα(s)v(s).

Page 38: Analytic and Differential geometry Mikhail G. Katz∗

38 3. QUADRATIC SURFACES

Proof. Indeed,

n′(s) = (−dyds,dx

ds)

= (x

y

dx

ds,dx

ds)

=dx

ds(x

y, 1)

=1

y

dx

ds(x, y)

=1

y

dx

dsv

and we apply Lemma 3.11.2 to prove the proposition. �

Remark 3.11.4. A similar phenomenon occurs for surfaces wherethe directional derivative of the normal vector is used to define theWeingarten map and ultimately the Gaussian curvature of the surface;see Section 9.1.

Page 39: Analytic and Differential geometry Mikhail G. Katz∗

CHAPTER 4

Curvature of curves

4.1. Osculating circle of a curve

To give a more geometric description of the curvature of a curve, wewill exploit its osculating circle at a point. We first recall the followingfact about the second derivative.

Theorem 4.1.1. Let f ∈ C2. The second derivative of a func-tion f(x) may be computed from a triple of points f(x), f(x+h), f(x−h)that are infinitely close to each other, as follows:

f ′′(x) = limh→0

f(x+ h) + f(x− h)− 2f(x)

h2.

The theorem remains valid for vector-valued functions; see Defini-tion 4.1.2. We see that the second derivative can be calculated fromthe value of the function at a triple of nearby points x, x+ h, x− h.

Definition 4.1.2. The osculating circle to the curve parametrizedby α with arclength parameter s at the point α(s) is obtained by choos-ing a circle passing through the three points

α(s), α(s− h), α(s+ h)

for infinitesimal h. The standard part1 of the resulting circle is theosculating circle; equivalently, we take limit as h tends to zero.

Remark 4.1.3. The osculating circle and the curve are “better thantangent” (they have second order tangency).

Since the second derivative is computed from the same triple ofpoints for α(s) and for the osculating circle (cf. Remark 4.1.1), wehave the following [We55, p. 13].

Theorem 4.1.4. The curvatures of the osculating circle and thecurve at the point of tangency are equal.2

1See Keisler [Ke74].2For example, let y = f(x). Compute the curvature of the graph of f

when f(x) = ax2. Let B = (x, x2). Let A be the midpoint of OB. Let C bethe intersection of the perpendicular bisector of OB with the y-axis. Let D =

39

Page 40: Analytic and Differential geometry Mikhail G. Katz∗

40 4. CURVATURE OF CURVES

4.2. Center of curvature, radius of curvature

Recall that we denote the geometric curve by C ⊆ R2 and itsparametrisation by α(s) (in the case of arclength).

To help geometric intuition, it is useful to recall Leibniz’s andCauchy’s definition of the radius of curvature of a curve [Cauchy 1826]as the distance from the curve to the intersection point of two infinitelyclose normals to the curve. In more detail, we have the following.

Definition 4.2.1. The center of curvature is the intersection pointof the normals to the curve at infinitely close points p and p′ of C.

Theorem 4.2.2 (Relation between center of curvature and oscu-lating circle). The intersection point of two infinitely close normals isthe center of the osculating circle at p.

Definition 4.2.3. The radius of curvature of a curve C at a point pis the distance from p to the center of curvature.

4.3. Flat Laplacian; Bateman–Reiss operator

Recall that we have two types of presentations of a curve, eitherparametric or implicit.

In Section 3.10 the curvature of a curve was calculated starting witha parametric representation of the curve. If a curve is given implicitlyas the locus (solution set) of an equation F (x, y) = 0, one can alsocalculate the geodesic curvature, by means of the theorems given inthe Section 4.4. We start with a definition.

Definition 4.3.1. The flat Laplacian ∆0 is the differential operatoris defined by

∆0 =∂2

∂x2+

∂2

∂y2.

Remark 4.3.2. The formula means that when we apply ∆ to asmooth function F = F (x, y), we obtain

∆0F =∂2F

∂x2+∂2F

∂y2.

(0, x). Triangle OAC yields sinψ = OAOC

=1

2

√x2+(ax2)2

r, triangle OBD yields

sinψ = BDOB

= ax2

√x2+a2x4

, and1

2

√x2+a2x4

r= ax2

√x2+a2x4

, so that 12 (x

2 + a2x4) = arx2,

and 12 (1 + a2x2) = ar, so that r = 1+a2x2

2a . Taking the limit as x→ 0, we ob-

tain r = 12a , hence k = 1

r= 2a = f ′′(0). Thus the curvature of the parabola at

its vertex equals the second derivative with respect to x (even though x is not thearclength parameter of the graph).

Page 41: Analytic and Differential geometry Mikhail G. Katz∗

4.4. GEODESIC CURVATURE FOR AN IMPLICIT CURVE 41

We also introduce the traditional shorter notation

Fxx =∂2F

∂x2, Fyy =

∂2F

∂y2, Fxy =

∂2F

∂x ∂y.

Then we obtain the equivalent formula

∆0(F ) = Fxx + Fyy.

Example 4.3.3. If F (x, y) = x2 + y2 − r2, then ∂2F∂x2

= 2 and

similarly ∂2F∂y2

= 2, hence ∆0F = 4.

We will be interested in the following operator that appears in aformula for curvature.

Definition 4.3.4. The Bateman-Reiss operator DB is defined by

DB(F ) = FxxF2y − 2FxyFxFy + FyyF

2x , (4.3.1)

which is a non-linear second order differential operator.

The subscript “B” stands for Bateman, as in the Bateman equa-tion FxxF

2y − 2FxyFxFy + FyyF

2x = 0. Alternatively, the operator can

be represented by the determinant

DB(F ) = −det

0 Fx FyFx Fxx FxyFy Fxy Fyy

This was treated in detail in [Goldman 2005, p. 637, formula (3.1)].The same operator occurs in the Reiss relation in algebraic geometry;see Griffiths and Harris [GriH78, p. 677].3 See also Valenti [2, p. 804]who refers to geodesic curvature as isophote curvature in the contextof a study of luminosity and eye center location. See also [3] and (two)references therein.

4.4. Geodesic curvature for an implicit curve

The operator defined in the previous section allows us to calculatethe curvature of a curve presented in implicit form, without having tospecify a parametrisation. Let CF ⊆ R2 be a curve defined implicitlyby F (x, y) = 0.

Theorem 4.4.1. Let p ∈ CF , and suppose ∇F (p) 6= 0. Then thegeodesic curvature kC of CF at the point p is given by

kC =|DB(F )||∇F |3 ,

where DB is the Bateman-Reiss operator defined in (4.3.1).

3Michel Reiss (1805-1869).

Page 42: Analytic and Differential geometry Mikhail G. Katz∗

42 4. CURVATURE OF CURVES

4.5. Curvature of conic sections

Example 4.5.1. In the case of the circle of radius r defined by theequation F (x, y) = 0, where F = x2 + y2 − r2, we obtain

Fx = 2x, Fy = 2y, ∇F = (2x, 2y)t,

and therefore |∇F | = 2√x2 + y2 = 2r. Meanwhile, Fxx = 2, Fyy =

2, Fxy = 0, hence

DB(F ) = 2(2y)2 + 2(2x)2 = 8r2,

and therefore the curvature is kC = 8r2

8r3= 1

r, as before (seee Theo-

rem 3.10.2).

Example 4.5.2. Consider the parabola C = CF given by the zerolocus of F (x, y) = y−x2. Then Fx = −2x, Fy = 1, Fxx = −2, Fyy = 0,Fxy = 0. Applying DB(F ) = FxxF

2y − 2FxyFxFy + FyyF

2x we obtain

DB(F ) = −2(1)2 = −2, while |∇F | =√1 + 4x2. Thus the curvature

satisfies kC = 2(1+x2)3/2

allowing us to identify the point of maximal

curvature in Theorem 4.5.3.

Theorem 4.5.3. The apex of the parabola y = x2 is its point ofmaximal curvature.

Proof. The calculations above produce kC = 2(1+x2)3/2

which at-

tains its maximal value k = 2 when x = 0. �

Example 4.5.4. Consider the hyperbola C = CF given by the zerolocus of F (x, y) = xy − 1. Then Fx = y, Fy = x, Fxx = 0, Fyy = 0,and Fxy = 1. Hence DB(F ) = −2xy. To simplify the expression forDB(F ) we exploit the defining equation of the curve xy = 1. Hencewe have DB(F ) = −2 at every point of the hyperbola. Meanwhile

|∇F | =√x2 + y2. Hence kC = 2

(x2+y2)3/2.

Theorem 4.5.5. The curvature of the hyperbola xy = 1 is maximalat the point (1, 1).

Proof. The calculations above produce kC = 2(x2+y2)3/2

. To maxi-

mize this expression along the curve, it suffices to minimize the expres-sion (x2+y2) in the denominator. Now exploiting the defining relationxy = 1 of the curve we obtain x2 + y2 = (x+ y)2 − 2xy = (x+ y)2 − 1.To minimize this expression, it suffices to minimize x+ y = x+ 1

x. The

sum x+ 1xis known to be minimal when x = 1. Namely, to show that

x + 1x≥ 2, rewrite the inequality as x2 + 1 ≥ 2x or x2 + 1 − 2x ≥ 0

or equivalently (x − 1)2 ≥ 0 which is a true inequality. Hence the

Page 43: Analytic and Differential geometry Mikhail G. Katz∗

4.6. CURVATURE OF GRAPH OF FUNCTION 43

maximum of the curvature is when x = 1 and so y = 1, proving thetheorem. �

Exercise 4.5.6. An interesting further example that should becalculated out is the curve y = ln x which of course is not a conic.

4.6. Curvature of graph of function

Theorem 4.6.1. Let x0 be a critical point of f(x), and consider thegraph of f at (x0, f(x0)). Then the curvature k of the graph equals

k = |f ′′(x0)|.Proof. We parametrize the graph by α(t) = (t, f(t)). Then we

have α′′(t) = (0, f ′′(t)) and |α′′(t)| = |f ′′(t)|, which is the expectedanswer at a critical point. However, the parametrisation (t, f(t)) is nota unit speed parametrisation of the graph.

Intuitively, the second order Taylor polynomial of f at x0 has thesame osculating circle as f , and therefore it suffices to check the resultfor a standard curve such as a circle or a parabola.

We check that applying the characterisation of curvature in termsof the Bateman operator DB(F ) = FxxF

2y − 2FxyFxFy + FyyF

2x gives

the same answer. Let F (x, y) = −f(x) + y. At the critical point x0,we have

∇F = (−f ′(x0), 1)t = (1, 0)t,

whileDB(F ) = −f ′′(x)(1)2 − 0 + 0 = −f ′′(x).

Hence the curvature at this point satisfies

k =|DBF ||∇F |3 =

|f ′′(x0)|1

= |f ′′(x0)|,

as required. �

Page 44: Analytic and Differential geometry Mikhail G. Katz∗
Page 45: Analytic and Differential geometry Mikhail G. Katz∗

CHAPTER 5

From arclength to surfaces

5.1. Existence of arclength

The unit speed parametrisation is sometimes called the arclengthparametrisation of the geometric curve C.

Theorem 5.1.1. For each regular curve α(t), there exists a unitspeed parametrition β(s) = α(t(s)) of the underlying geometric curve C,

where the parameter s defined by equation s(t) =

∫ t

a

∣∣∣dαdτ

∣∣∣dτ .

Proof. Recall the formula for the length of the graph of f(x)from a to b:

L =

∫ b

a

√1 + (f ′(x))2 dx .

The element of length (infinitesimal increment) ds along the graphdecomposes by Pythagoras’ theorem as follows:

ds2 = dx2 + dy2.

More generally, for a curve α(t) = (α1(t), α2(t)), we have the followingformula for the length:

L =

∫ b

a

ds

=

∫ b

a

√dx2 + dy2

=

∫ b

a

√(d(α1)

dt

)2+(dα2

dt

)2dt

=

∫ b

a

∣∣∣dαdt

∣∣∣dt.

We define the new parameter s = s(t) by setting

s(t) =

∫ t

a

∣∣∣dαdτ

∣∣∣dτ, (5.1.1)

45

Page 46: Analytic and Differential geometry Mikhail G. Katz∗

46 5. FROM ARCLENGTH TO SURFACES

where τ is a dummy variable (internal variable of integration). By the

Fundamental Theorem of Calculus, we haveds

dt=∣∣∣dαdt

∣∣∣. Let β(s) =

α(t(s)). Then by chain rule

ds=dα

dt

dt

ds=dα

dt

1

ds/dt=dα

dt

1

|dα/dt| .

Thus∣∣dβds

∣∣ = 1. �

Example 5.1.2 (Cusp). The plane curve α(t) = (t3, t2) is smoothbut not regular. Its graph exhibits a cusp. In this case it is impossibleto find an arclength parametrisation of the curve.

Remark 5.1.3 (Curves in R3). A space curve may be written incoordinates as

α(s) = (α1(s), α2(s), α3(s)).

Here s is the arc length if∣∣∣dαds∣∣∣ = 1 i.e.

∑3i=1

(dαi

ds

)2= 1.

Example 5.1.4. Helix α(t) = (a cosωt, a sinωt, bt).(i) make a drawing in case a = b = ω = 1.(ii) parametrize by arc length.(iii) compute the curvature.

5.2. Arnold’s observation on folding a page

The differential geometry of surfaces in Euclidean 3-space startswith the observation that they inherit a metric structure from the am-bient space (i.e. the Euclidean space).

Question 5.2.1. We would like to understand which geometricproperties of this structure are intrinsic?1

Part of the job is to clarify the sense of the term intrinsic.

Remark 5.2.2. Following Arnold [Ar74, Appendix 1, p. 301], notethat a piece of paper may be placed flat on a table, or it may be rolledinto a cylinder, or it may be rolled into a cone. However, it cannotbe transformed into the surface of a sphere, that is, without tearing orstretching. Understanding this phenomenon quantitatively is our goal,cf. Figure 12.2.1.

1This is atzmit rather than pnimit according to Vishne.

Page 47: Analytic and Differential geometry Mikhail G. Katz∗

5.3. REGULAR SURFACE; JACOBIAN 47

5.3. Regular surface; Jacobian

Our starting point is the first fundamental form, obtained by re-stricting the 3-dimensional inner product. What does the first funda-mental form measure? A helpful observation to keep in mind is that itenables one to measure the length of curves on the surface.

Consider a surface M ⊆ R3 parametrized by a map x(u1, u2) or

x : R2 → R3. (5.3.1)

We will always assume that x is differentiable. From now on we willfrequently omit the underline in x, and write simply “x”, to lighten thenotation.

Definition 5.3.1. The Jacobian matrix Jx is the 3× 2 matrix

Jx =

(∂xi

∂uj

), (5.3.2)

where xi are the three components of the vector valued function x.

Definition 5.3.2. The parametrisation x of the surfaceM is calledregular if one of the following equivalent conditions is satisfied:

(1) the vectors ∂x∂u1

and ∂x∂u2

are linearly independent;

(2) the vector product x1 × x2 is nonzero, where xi =∂x∂ui

;(3) the Jacobian matrix (5.3.2) is of rank 2.

Example 5.3.3. Let b > 0 be a fixed real number, and consider thefunction f(x, y) =

√b2 − x2 − y2. The graph of f can be parametrized

as follows:

x(u1, u2) = (u1, u2, f(u1, u2)).

This provides a parametrisation of the (open) northern hemisphere.

Then ∂x∂u1

= (1, 0, fx(u1, u2))t while ∂x

∂u2= (0, 1, fy(u

1, u2))t. These vec-tors are linearly independent and therefore we have a regular parametri-sation.

Note that we have the relations

fx =−xf

=−x√

b2 − x2 − y2, fy =

−yf. (5.3.3)

The southern hemisphere can be similarly parametrized by

x′ = (u1, u2,−f(u1, u2)).The formulas for the ∂x

∂uiare then ∂x

∂u1= (1, 0,−fx)t = (1, 0, x√

b2−x2−y2),

while ∂x∂u2

= (0, 1,−fy)t = (0, 1, y√b2−x2−y2

).

Page 48: Analytic and Differential geometry Mikhail G. Katz∗

48 5. FROM ARCLENGTH TO SURFACES

5.4. Coefficients of first fundamental form of a surface

Let 〈 , 〉 denote the inner product in R3. For i = 1, 2 and j = 1, 2,define functions gij = gij(u

1, u2) called “metric coefficients” by

gij(u1, u2) =

⟨ ∂x∂ui

,∂x

∂uj

⟩. (5.4.1)

Remark 5.4.1. We have gij = gji as the inner product is symmet-ric.

Example 5.4.2 (Metric coefficients for the graph of a function).The graph of a function f satisfies

⟨∂x

∂u1,∂x

∂u1

⟩= 1 + f 2

x ,

⟨∂x

∂u1,∂x

∂u2

⟩= fxfy,

⟨∂x

∂u2,∂x

∂u2

⟩= 1 + f 2

y .

Therefore in this case we have

(gij) =

(1 + f 2

x fxfyfxfy 1 + f 2

y

)

In the case of the hemisphere, the partial derivatives are as in for-mula (5.3.3).

5.5. Metric coefficients in spherical coordinates

Spherical coordinates2 (ρ, θ, ϕ) in 3-space will be reviewed in moredetail in Section 7.1. Briefly, ϕ is the angle formed by the positionvector of the point with the positive direction of the z-axis. Mean-while, θ is the polar coordinate angle θ for the projection of the vectorto the x, y plane.

Example 5.5.1 (Metric coefficients in spherical coordinates). Con-sider the parametrisation of the unit sphere S2 by means of sphericalcoordinates, so that x = x(θ, ϕ). We have

x = sinϕ cos θ, y = sinϕ sin θ, z = cosϕ .

Setting u1 = θ and u2 = ϕ, we obtain

∂x

∂u1= (− sinϕ sin θ, sinϕ cos θ, 0)

2Koordinatot kaduriot

Page 49: Analytic and Differential geometry Mikhail G. Katz∗

5.6. TANGENT PLANE 49

and∂x

∂u2= (cosϕ cos θ, cosϕ sin θ,− sinϕ),

Thus in this case we have

(gij) =

(sin2 ϕ 00 1

)

so that det(gij) = sin2 ϕ and√

det(gij) = sinϕ.

5.6. Tangent plane

Consider a parametrisation x(u1, u2) of a surface M ⊆ R3 andlet p = x(u1, u2) ∈ R3 be a point on the surface.

Definition 5.6.1. The vectors xi =∂x

∂ui, i = 1, 2, are called the

tangent vectors to the surface M at the point p = x(u1, u2).

Definition 5.6.2. Given a finite set S = {vi}i=1,...,n in Rb, wedefine its Gram matrix as the matrix of inner products

Gram(S) = (〈vi, vj〉)i=1,...,n; j=1,...,n. (5.6.1)

In Section 5.5 we defined the metric coefficients gij = gij(u1, u2).

We now state a relationship between the Gram matrix and the metriccoefficients.

Theorem 5.6.3. The 2× 2 matrix (gij) is the Gram matrix of thepair of tangent vectors:

(gij) = JTx Jx,

cf. formula (15.6.1).

The proof is immediate.

Definition 5.6.4. The plane spanned by the vectors x1(u1, u2)

and x2(u1, u2) is called the tangent plane to the surface M at the

point p = x(u1, u2), and denoted

TpM.

Page 50: Analytic and Differential geometry Mikhail G. Katz∗
Page 51: Analytic and Differential geometry Mikhail G. Katz∗

CHAPTER 6

First fundamental form, Gamma symbols

6.1. First fundamental form

The tangent plane TpM of a surface M ⊆ R3 was defined in Sec-tion 5.6.

Definition 6.1.1. The first fundamental form Ip of the surface Mat the point p is the bilinear form on the tangent plane Tp, namely

Ip : TpM × TpM → R,

defined by the restriction of the ambient Euclidean inner product:

Ip(v, w) = 〈v, w〉R3,

for all v, w ∈ TpM . With respect to the basis (frame) {x1, x2}, it isgiven by the matrix (gij), where

gij = 〈xi, xj〉.The coefficients gij are sometimes called ‘metric coefficients.’

Like curves, surfaces can be represented either implicitly or para-metrically (see Section 3.6 for representations of curves).

6.2. Plane and cylinder

The example of the sphere was discussed in Section 5.5. We nowconsider two more examples.

Example 6.2.1 (Plane). The x, y-plane in R3 is defined implic-itly by the equation z = 0. Consider the parametrisation x(u1, u2) =(u1, u2, 0) ∈ R3. This is a parametrisation of the xy-plane in R3.Then x1 = (1, 0, 0)t, x2 = (0, 1, 0)t, and

g11 = 〈x1, x1〉 =⟨100

,

100

= 1,

etcetera. Thus we have (gij) =

(1 00 1

)= I.

51

Page 52: Analytic and Differential geometry Mikhail G. Katz∗

52 6. FIRST FUNDAMENTAL FORM, GAMMA SYMBOLS

Example 6.2.2 (Cylinder). Let x(u1, u2) = (cosu1, sin u1, u2). Thisformula provides a parametrisation of the cylinder. We have

x1 = (− sin u1, cosu1, 0)t

and x2 = (0, 0, 1)t, while

g11 =

⟨− sin u1

cosu1

0

,

− sin u1

cosu1

0

= sin2 u1 + cos2 u1 = 1,

etc. Thus (gij) =

(1 00 1

)= I.

Remark 6.2.3. The two examples above illustrate that the firstfundamental form does not contain all the information (even up toEuclidean congruences) about the surface. Indeed, the plane and thecylinder have the same first fundamental form, but are geometricallydistinct imbedded surfaces.

6.3. Surfaces of revolution

Definition 6.3.1 (Surfaces of revolution). Here it is customary touse the notation u1 = θ and u2 = φ. The starting point is a generatingcurve C in the xz-plane, parametrized by a pair of functions

x = r(φ), z = z(φ).

The surface of revolution (around the z-axis) generated by C is para-metrized as follows:

x(θ, φ) = (r(φ) cos θ, r(φ) sin θ, z(φ)). (6.3.1)

Example 6.3.2. If we start with a generating curve which is thevertical line r(φ) = 1, z(φ) = φ, the resulting surface of revolution isthe cylinder.

Example 6.3.3. The generating curve r(φ) = sin φ, z(φ) = cos φyields the sphere S2 in spherical coordinates as discussed in Section 5.5;see Example 6.3.5 for more details.

Theorem 6.3.4. If φ is the arclength parameter of a parametrisa-tion (r(φ), z(φ)) of the generating curve C, then the first fundamentalform of the corresponding surface of revolution (6.3.1) is given by

(gij) =

(r2(φ) 00 1

).

Page 53: Analytic and Differential geometry Mikhail G. Katz∗

6.3. SURFACES OF REVOLUTION 53

Proof. We have

x1 =∂x

∂θ= (−r sin θ, r cos θ, 0)t,

x2 =∂x

∂φ=

(dr

dφcos θ,

dr

dφsin θ,

dz

)t

therefore

g11 =

∣∣∣∣∣∣∣

−drf sin θr cos θ

0

∣∣∣∣∣∣∣

2

= r2 sin2 θ + r2 cos2 θ = r2

and

g22 =

∣∣∣∣∣∣

drdφ

cos θdrdφ

sin θdzdφ

∣∣∣∣∣∣

2

=

(dr

)2

(cos2 θ + sin2 θ) +

(dz

)2

=

(dr

)2

+

(dz

)2

and

g12 =

⟨−r sin θr cos θ

0

,

drdφ

cos θdrdφ

sin θdzdφ

= −r drdφ

sin θ cos θ + rdr

dφcos θ sin θ = 0.

Thus

(gij) =

(r2 0

0(drdφ

)2+(dzdφ

)2).

In the case of an arclength parametrisation of the generating curve C,we obtain g22 = 1, proving the theorem. �

Example 6.3.5. Consider the curve (sinφ, cosφ) in the x, z-plane.The resulting surface of revolution is the sphere, where the φ parame-ter coincides with the angle ϕ of spherical coordinates. Thus, for thesphere S2 we obtain

(gij) =

(sin2 φ 00 1

).

Page 54: Analytic and Differential geometry Mikhail G. Katz∗

54 6. FIRST FUNDAMENTAL FORM, GAMMA SYMBOLS

6.4. From tractrix to pseudosphere

Example 6.4.1. The pseudosphere (so called because its Gauss-ian curvature equals −1) is the surface of revolution generated by acurve called the tractrix. This curve is parametrized by (r(φ), z(φ))where r(φ) = eφ and

z(φ) =

∫ φ

0

√1− e2ψdψ = −

∫ 0

φ

√1− e2ψdψ,

where −∞ < φ ≤ 0. The tractrix generates a surface of revolutionwith g11 = e2φ, while

g22 = (eφ)2 + (√

1− e2φ)2

= e2φ + 1− e2φ = 1.

Thus (gij) =

(e2φ 00 1

).

6.5. Chain rule in two variables

How does one measure the length of curves on a surface in terms ofthe metric coefficients of the surface x(u1, u2)? Let

α : [a, b] → R2, α(t) = (α1(t), α2(t))

be a plane curve. We will exploit the Einstein summation convention.We will also use the following version of chain rule in several variables.

Theorem 6.5.1 (Chain rule in two variables). Consider the curveon the surface defined by the composition β(t) = x ◦ α(t) where x =

x(u1, u2). Then dβdt

=∑2

i=1∂x∂ui

dαi

dt, or in Einstein summation conven-

tiondβ

dt=

∂x

∂uidαi

dt= Jx

(dαdt

)t

(the superscript t is the transpose).

6.6. Measuring length of curves on surfaces

Theorem 6.6.1. Let β(t) = x ◦ α(t) be a curve on the surface x.Then the length L of β is given by the formula

L =

∫ b

a

√gijdαi

dt

dαj

dtdt.

Page 55: Analytic and Differential geometry Mikhail G. Katz∗

6.7. NORMAL VECTOR, SYMBOLS Γkij OF A SURFACE 55

Proof. The length L of β is calculated as follows using chain rule:

L =

∫ b

a

∣∣∣∣dβ

dt

∣∣∣∣ dt =∫ b

a

∣∣∣∣∣

2∑

i=1

∂x

∂uidαi

dt

∣∣∣∣∣ dt

=

∫ b

a

⟨xidαi

dt, xj

dαj

dt

⟩1/2

dt

=

∫ b

a

(〈xi, xj〉

dαi

dt

dαj

dt

)1/2

dt =

=

∫ b

a

√gij(α(t))

dαi

dt

dαj

dtdt.

We therefore obtain L =∫ ba

√gij(α(t))

dαi

dtdαj

dtdt as required. �

Corollary 6.6.2. If gij = δij is the identity matrix, then the lengthof the curve β = x ◦ α on the surface equals the length of the curve αof R2.

Indeed, if the first fundamental form of the surface satisfies gij = δij,

then L =

∫ b

a

√(dα1

dt

)2

+

(dα2

dt

)2

dt = the length of α(t) in R2. The

parametrisation in this case is an isometry.To simplify notation, let x = x(t) = α1(t) and y = y(t) = α2(t).

Then the dt cancels out, and the infinitesimal element of arclength is

ds =√dx2 + dy2,

and the length of the curve is∫ds =

∫ √dx2 + dy2.

6.7. Normal vector, symbols Γkij of a surface

The symbols Γkij, roughly speaking,1 account for how the surfacetwists in space. They are, however, coordinate dependent.

We will see that the symbols Γkij also control the behavior of geodesicson the surface. Here geodesics on a surface can be thought of as curvesthat are to the surface what straight lines are to a plane, or what greatcircles are to a sphere, cf. Definition 7.11.2.

Let x : R2 → R3 be a regular parametrized surface, and let xi =∂x

∂ui.

1In the first approximation: kiruv rishon.

Page 56: Analytic and Differential geometry Mikhail G. Katz∗

56 6. FIRST FUNDAMENTAL FORM, GAMMA SYMBOLS

Definition 6.7.1. The normal vector n(u1, u2) to a regular surfaceat the point x(u1, u2) ∈ R3 is defined in terms of the vector product,cf. Definition 1.12.7, as follows:

n =x1 × x2|x1 × x2|

,

so that 〈n, xi〉 = 0 ∀i.

6.8. Gamma symbols

The vectors x1, x2, n form a basis (frame) for R3 (since x is regularby hypothesis). Recall that the gij were defined in terms of first partialderivatives of x. Meanwhile, the symbols Γkij are defined in terms ofthe second partial derivatives (since they are not even tensors, one doesnot stagger the indices). Namely, they are defined as the coefficients ofthe decomposition of the second partial derivative vector xij , definedby

xij =∂2x

∂ui∂uj

with respect to the basis, or frame,2 (x1, x2, n), as follows.

Definition 6.8.1. The symbols Γkij are uniquely determined by theformula

xij = Γ1ijx1 + Γ2

ijx2 + Lijn

(the coefficients Lij will be discussed in Section 9.7).

6.9. Basic properties of the Gamma symbols

Proposition 6.9.1. We have the following formula for the Gammasymbols:

Γkij = 〈xij , xℓ〉gℓk,where (gij) is the inverse matrix of gij.

Proof. We have

〈xij, xℓ〉 = 〈Γkijxk + Lijn, xℓ〉 = 〈Γkijxk, xℓ〉+ 〈Lijn, xℓ〉 = Γkijgkℓ.

We now multiply by gℓm and sum:

〈xij , xℓ〉gℓm = Γkijgkℓgℓm = Γkijδ

mk = Γmij .

This is equivalent to the desired formula. �

Remark 6.9.2. We have the following relation: Γkij = Γkji, or Γk[ij] =

0 ∀ijk.2Maarechet yichus

Page 57: Analytic and Differential geometry Mikhail G. Katz∗

6.10. INTRINSIC NATURE OF THE GAMMA SYMBOLS 57

Example 6.9.3 (The plane). Calculation of the symbols Γkij forthe plane x(u1, u2) = (u1, u2, 0). We have x1 = (1, 0, 0), x2 = (0, 1, 0).Thus we have xij = 0 ∀i, j. Hence Γkij = 0 for all i, j, k.

Example 6.9.4 (The cylinder). Calculation of the symbols for thecylinder x(u1, u2) = (cosu1, sin u1, u2). The normal vector is n =(cosu1, sin u1, 0), while x1 = (− sin u1, cosu1, 0), x2 = (0, 0, 1). Thuswe have x22 = 0, x21 = 0 and so Γk22 = 0 and Γk12 = Γk21 = 0 ∀k.

Meanwhile, x11 = (− cosu1,− sin u1, 0). This vector is proportionalto n:

x11 = 0x1 + 0x2 + (−1)n.

Hence Γk11 = 0 ∀k.Definition 6.9.5. We will use the following notation for the partial

derivative of the function gij = gij(u1, u2):

gij;k =∂

∂uk(gij).

Lemma 6.9.6. In terms of the symmetrisation notation introducedin Section 1.9, we have

gij;k = 2gm{iΓmj}k.

Proof. Indeed, by Leibniz rule

gij;k =∂

∂uk〈xi, xj〉 = 〈xik, xj〉+ 〈xi, xjk〉 = 〈Γmikxm, xj〉+ 〈Γmjkxm, xi〉.

By definition of the metric coefficients, we have

gij;k = Γmikgmj + Γmjkgmi = gmjΓmik + gmiΓ

mjk = 2gm{jΓ

mi}k

or 2gm{iΓmj}k. �

An important role in the theory is played by the intrinsic3 natureof the coefficients Γ. Namely, we will prove that they are determinedby the metric coefficients alone, and are therefore independent of theambient (extrinsic) geometry of the surface, i.e., the way it “sits” in3-space.

6.10. Intrinsic nature of the Gamma symbols

Theorem 6.10.1. The symbols Γkij can be expressed in terms of thefirst fundamental form and its derivatives as follows :

Γkij =1

2(giℓ;j − gij;ℓ + gjℓ;i)g

ℓk,

where gij is the inverse matrix of gij.

3This is atzmit and not pnimit according to Vishne.

Page 58: Analytic and Differential geometry Mikhail G. Katz∗

58 6. FIRST FUNDAMENTAL FORM, GAMMA SYMBOLS

Γ1ij

j = 1 j = 2

i = 1 0 1fdrdφ

i = 2 1fdrdφ

0

Table 6.11.1. Symbols Γ1ij of a surface of revolution (6.11.1)

Proof. Applying Lemma 6.9.6 three times, we obtain

giℓ;j − gij;ℓ + gjℓ;i = 2gm{iΓmℓ}j − 2gm{iΓ

mj}ℓ + 2gm{jΓ

mℓ}i

= gmiΓmℓj + gmℓΓ

mij − gmiΓ

mjℓ − gmjΓ

miℓ + gmjΓ

mℓi + gmℓΓ

mji

= 2gmℓΓmji .

Thus1

2(giℓ;j−gij;ℓ+gjℓ;i)gℓk = Γmij gmℓg

ℓk = Γmij δkm = Γkij, as required. �

6.11. Gamma Symbols for a surface of revolution

Recall that a surface of revolution is obtained by starting with acurve (r(φ), z(φ)) in the (x, z) plane, and rotating it around the z-axis,obtaining the parametrisation x(θ, φ) = (r(φ) cos θ, r(φ) sin θ, z(φ)).Thus we adopt the notation

u1 = θ, u2 = φ.

Lemma 6.11.1. For a surface of revolution we have Γ111 = Γ1

22 = 0,

while Γ112 =

r drdφ

r2, cf. Table 6.11.1.

Proof. For the surface of revolution

x(θ, φ) = (r(φ) cos θ, r(φ) sin θ, z(φ)), (6.11.1)

the metric coefficients are given by the matrix

(gij) =

(r2 0

0(drdφ

)2+(dzdφ

)2).

We will use the formula Γkij = 12(giℓ;j − gij;ℓ + gjℓ;i)g

ℓk (see Theo-rem 6.10.1). Since the off-diagonal coefficient g12 = 0 vanishes, thediagonal coefficients of the inverse matrix satisfy

gii =1

gii. (6.11.2)

Page 59: Analytic and Differential geometry Mikhail G. Katz∗

6.11. GAMMA SYMBOLS FOR A SURFACE OF REVOLUTION 59

We have∂

∂θ(gii) = 0 since gii depend only on φ. Thus the terms

gii;1 = 0 (6.11.3)

vanish. Let us now compute the symbols Γ1ij for k = 1. Using formulas

(6.11.2) and (6.11.3), we obtain

Γ111 =

1

2g11(g11;1 − g11;1 + g11;1) by formula (6.11.2)

= 0 by formula (6.11.3).

Similarly,

Γ112 =

1

2g11(g11;2 − g12;1 + g12;1) =

g11;22g11

=

ddφ(r2)

2r2=r dr

r2,

while Γ122 =

12g11

(g12;2 − g22;1 + g12;2) =g12;2g11

=ddφ

(0)

g11= 0. �

Page 60: Analytic and Differential geometry Mikhail G. Katz∗
Page 61: Analytic and Differential geometry Mikhail G. Katz∗

CHAPTER 7

From Clairaut’s relation to geodesic equation

7.1. Spherical coordinates

Spherical coordinates are useful in understanding surfaces of revo-lution (see Section 6.3).

Definition 7.1.1. Spherical coordinates (ρ, θ, ϕ) in R3 are definedby the following formulas. We have

ρ =√x2 + y2 + z2 =

√r2 + z2

is the distance to the origin, r =√x2 + y2, while ϕ is the angle with

the z-axis, so that cosϕ = zρ. Here θ is the angle inherited from polar

coordinates in the x, y plane, so that tan θ = yx.

Remark 7.1.2. The interval of definition for the variable ϕ is ϕ ∈[0, π] since ϕ = arccos z

ρand the range of the arccos function is [0, π].

Meanwhile θ ∈ [0, 2π] as usual.

The unit sphere S2 ⊆ R3 is defined in spherical coordinates by thecondition

S2 = {(ρ, θ, ϕ) : ρ = 1}.Definition 7.1.3. A latitude1 on the unit sphere is a circle satis-

fying the equationϕ = constant.

Example 7.1.4. The equator, defined by the equation {ϕ = π2},

is the only latitude that is also a great circle (see Section 7.2) of thesphere.

Each latitude of the sphere is parallel to the equator (i.e., liesin a plane parallel to the plane of the equator). A latitude can beparametrized by setting θ(t) = t and ϕ(t) = constant. Note that for apoint on the unit sphere defined by equation {ρ = 1}, we have

r = sinϕ, (7.1.1)

where r is the distance from the point to the z-axis.

1kav rochav

61

Page 62: Analytic and Differential geometry Mikhail G. Katz∗

62 7. FROM CLAIRAUT’S RELATION TO GEODESIC EQUATION

7.2. Great circles, Clairaut’s relation

In Section 8.2, we will encounter the geodesic equation. This is asystem of nonlinear second order differential equations.

We seek first to provide a geometric intuition for this equation. Wewill establish a connection between solutions of this system, on the onehand, and spherical great circles, on the other. Such a connection isestablished via the intermediary of Clairaut’s relation for a variablepoint q on a great circle:

r(t) cos γ(t) = const

where r is the distance from q to the (vertical) axis of revolution and γis the angle at q between the direction of the great circle and thelatitudinal circle (cf. Theorem 7.4.1).

Remark 7.2.1. In this section, we will verify Clairaut’s relationsynthetically for great circles, and also show that the latter satisfy afirst order differential equation. In Section 8.2 we will complete theconnection by deriving Clairaut’s relation from the geodesic equation.

Let S2 ⊆ R3 be the unit sphere defined by the equation

x2 + y2 + z2 = 1.

A plane P through the origin is given by an equation

ax+ by + cz = 0, (7.2.1)

where the vector abc

6=

000

is nonzero.

Definition 7.2.2. A great circle G of S2 is given by an intersection

G = S2 ∩ P.Example 7.2.3 (A parametrisation of the equator of S2). The

equator is parametrized by

α(t) = (cos t)e1 + (sin t)e2.

In general, every great circle can be parametrized by

α(t) = (cos t)v + (sin t)w

where v, w ∈ S2 are orthonormal vectors in R3.

Page 63: Analytic and Differential geometry Mikhail G. Katz∗

7.3. SOME BASIC LEMMAS 63

Example 7.2.4 (an implicit (non-parametric) representation of agreat circle). Recall that we have

x = r cos θ = ρ sinϕ cos θ; y = ρ sinϕ sin θ; z = ρ cosϕ . (7.2.2)

If the circle lies in the plane ax+ by+ cz = 0 where a, b, c are fixed, thegreat circle in coordinates (θ, ϕ) is defined implicitly by the equation

a sinϕ cos θ + b sinϕ sin θ + c cos θ = 0,

as in (7.2.1).

7.3. Some basic lemmas

Lemma 7.3.1. Scalar product of vector valued functions satisfiesLeibniz’s rule:

〈f, g〉′ = 〈f ′, g〉+ 〈f, g′〉. (7.3.1)

Proof. See [Leib]. In more detail, let (f1, f2) be components of f ,and let (g1, g2) be components of g. Then

〈f, g〉′ = (f1g1 + f2g2)′ = f1g

′1 + f ′

1g1 + f2g′2 + f ′

2g2 = 〈f ′, g〉+ 〈f, g′〉.

The same proof goes through for arbitrary number of components, e.g.,for functions with values in R3. �

Lemma 7.3.2. Let α : R → S2 be a parametrized curve on the

sphere S2 ⊆ R3. Then the tangent vectordα

dtis perpendicular to the

position vector α(t), or in formulas:

⟨α(t),

dt

⟩= 0.

Proof. We have 〈α(t), α(t)〉 = 1 by definition of S2. We apply theoperator d

dtto obtain

d

dt〈α(t), α(t)〉 = 0.

Next, we apply Leibniz’s rule (7.3.1) to obtain

⟨α(t),

dt

⟩+⟨dαdt, α(t)

⟩= 2⟨α(t),

dt

⟩=

d

dt(1) = 0,

completing the proof. �

Page 64: Analytic and Differential geometry Mikhail G. Katz∗

64 7. FROM CLAIRAUT’S RELATION TO GEODESIC EQUATION

7.4. Clairaut’s relation

Our interest in Clairaut’s relation lies in the motivation it providesfor the general geodesic equation on a surface of revolution. We will useNewton’s dot notation for the derivative with respect to t as in α(t).

Theorem 7.4.1 (Clairaut’s relation). Let α(t) be a regular para-metrisation of a great circle G on S2 ⊆ R3. Let r(t) denote the dis-tance from the point α(t) to the z-axis, and let γ(t) denote the anglebetween the tangent vector α(t) to the curve and the vector tangent tothe latitude through the point α(t). Then

r(t) cos γ(t) = const.

Here the constant has value const = rmin, where rmin is the least Eu-clidean distance from a point of G to the z-axis.

7.5. Spherical sine law

Let S2 be the unit sphere in 3-space.

Definition 7.5.1. A spherical triangle is the following collectionof data: the vertices are points of S2, the sides are arcs of great circles,while the angles are defined to be the angles between tangent vectorsto the sides. Here we assume that

(1) all sides have length < π,(2) the three vertices do not lie on a common great circle.

Theorem 7.5.2 (Spherical sine law).

sin a

sinα=

sin b

sin β=

sin c

sin γ

Example 7.5.3. Let γ = π2, then sin a = sin c sin γ. Note that for

small values of the sides a, c we recapture the Euclidean sine law

a = c sinα

as the limiting case.

7.5.1. This subsection is optional.

First proof of spherical sine law. Let A,B,C be the angles. Wechoose a coordinate system so that the three vertices of the spherical triangleare located at (1, 0, 0), (cos a, sin a, 0) and (cos b, sin b cosC, sin b sinC). Thevolume of the tetrahedron formed from these three vertices and the originis 1

6 sin a sin b sinC. Since this volume is invariant under cyclic relabeling ofthe sides and angles, we have

sin a sin b sinC = sin b sin c sinA = sin c sin a sinB

Page 65: Analytic and Differential geometry Mikhail G. Katz∗

7.5. SPHERICAL SINE LAW 65

and thereforesinA

sin a=

sinB

sin b=

sinC

sin cproving the law. See http://math.stackexchange.com/questions/1735860.

Second proof. Denote byA,B,C the vertices opposite the sides a, b, c.The points A,B,C can be thought of as unit vectors in R3. Note that sin a =|B × C|, etc. Meanwhile sinα = |C ′ × B′| where C ′ is normalisation of theorthogonal component of C when the component parallel to A is eliminated(through the Gram-Schmidt process). Here the component of C parallelto A is (C · A)A of norm cos b. The orthogonal component is C − (C · A)Ais of norm sin b = |A× C| = |C − (C ·A)A|. Thus we have

C ′ =C − (C ·A)A

|A× C| , B′ =B − (B ·A)A

|A×B| .

Therefore

sinα

sin a=

∣∣∣C−(C·A)A|A×C| × B−(B·A)A

|A×B|

∣∣∣|B × C| =

|(C − (C ·A)A) × (B − (B ·A)A)||A×B| |A× C| |B × C|

In the denominator we get the product of the three vector products, namely|A×B| |B×C| |C×A|, which is symmetric in the three vectors. Meanwhilein the numerator we get the norm of the vector

(C − (C · A)A)× (B − (B ·A)A). (7.5.1)

The triple of vectors A,B,C is transformed into the triple

A, (C − (C ·A)A), (B − (B ·A)A)by a volume-preserving transformation. This combined with the fact that Ais a unit vector shows that the norm of the vector (7.5.1) equals the absolutevalue of the determinant of the 3× 3 matrix [AB C] (i.e., absolute value ofthe volume of the parallelopiped spanned by the three vectors), which is alsosymmetric in the three points. Thus we have

sinα

sin a=

|det[AB C] |sin a sin b sin c

,

proving the spherical sine law. �

Third proof. There is an alternative proof that relies on the iden-tity (a × b) × (a × c) = (a · (b × c))a for each triple of vectors a, b, c ∈ R3.Indeed,

sinα =|(A×B)× (A× C)|

|A×B| |A× C| =det(AB C)

sin b sin c

since A is a unit vector, and therefore

sinα

sin a=

det(AB C)

sin a sin b sin c

Page 66: Analytic and Differential geometry Mikhail G. Katz∗

66 7. FROM CLAIRAUT’S RELATION TO GEODESIC EQUATION

as required. The sources for the spherical sine law are Smart 1960, pp. 9-10;Gellert et al. 1989, p. 265; Zwillinger 1995, p. 469.

Gellert, W.; Gottwald, S.; Hellwich, M.; Kastner, H.; and Kunstner,H. (Eds.). ”Spherical Trigonometry.” 12 in VNR Concise Encyclopedia ofMathematics, 2nd ed. New York: Van Nostrand Reinhold, pp. 261-282,1989.

Smart, W. M. Text-Book on Spherical Astronomy, 6th ed. Cambridge,England: Cambridge University Press, 1960.

Zwillinger, D. (Ed.). ”Spherical Geometry and Trigonometry.” 6.4 inCRC Standard Mathematical Tables and Formulae. Boca Raton, FL: CRCPress, pp. 468-471, 1995. �

7.6. Longitudes

Definition 7.6.1. A longitude (meridian)2 on S2 is the arc of agreat circle connecting the North Pole e3 = (0, 0, 1) and the SouthPole (−1, 0, 0).

Lemma 7.6.2. Assume a great circle G ⊆ S2 does not pass throughthe north pole, i.e., does not contain a longitude (meridian). Let p ∈ Gbe the point with the maximal z-coordinate among the points of G.Let α(t) be a regular parametrisation of G with α(0) = p, and denoteby α(0) the tangent vector to G at p. Then the great circle G has thefollowing three equivalent properties:

(1) α(0) is proportional to the vector product e3 × p;(2) α(0) is perpendicular to the longitude (meridian) passing through p,

and(3) α(0) is tangent at p to the latitudinal circle.

Proof. We will prove (1). Note that 〈α(0), p〉 = 0 by Lemma 7.3.2.It remains to prove that α(0) is orthogonal to e3. By definition of thepoint p, the function 〈α(t), e3〉 achieves its maximum at t = 0. Hencewe have

d

dt

∣∣∣t=0

〈α(t), e3〉 = 0.

By Leibniz’s rule,⟨dαdt

∣∣∣t=0, e3

⟩= −

⟨α(t),

de3dt

⟩= 0,

since e3 is constant. Thus 〈α(0), e3〉 = 0. �

Remark 7.6.3 (Equation of great circle). This material is optional.Note that a great circle on the sphere satisfies the equation cot(ϕ) =tan(γ) cos(θ − θ0), where γ is the angle of inclination of the plane

2kav orech

Page 67: Analytic and Differential geometry Mikhail G. Katz∗

7.7. PROOF OF CLAIRAUT’S RELATION 67

(see below). Indeed, a great circle is the intersection of the spherewith a plane through the origin. Let a unit normal to that plane beu = [− sin(γ), 0, cos(γ)], where for convenience we choose our x and yaxes so that y = 0 (the plane contains the y-axis). Then the equationu · [sinϕ cos θ, sinϕ sin θ, cosϕ] = 0 becomes cot(ϕ) = tan(γ) cos(θ).Rotating around the z axis, this becomes cot(ϕ) = tan(γ) cos(θ − θ0)(this gamma has anything to do with the gamma from Clairaut’s rela-tion).

7.7. Proof of Clairaut’s relation

Definition 7.7.1. The spherical distance on S2 is the distanced(p, q) between p, q ∈ S2 measured along arcs of great circles.

Lemma 7.7.2. Let p be the north pole: p = e3. Then the sphericaldistance d(e3, q) is the ϕ-coordinate of the point q.

Proof. This follows from the fact that the arclength along a unitcircle equals the subtended angle. Note that the plane containing e3and q includes the meridian passing through the point q. �

First we note that the complementary angle of γ is the angle withthe longitude.

Lemma 7.7.3. Let α(t) be a regular parametrisation of a great circle.If γ(t) is the angle between α(t) and the latitude, then there is anangle of π

2− γ(t) between α(t) and the vector tangent to the longitude

(meridian) at the point α(t).

Proof. By vanishing of the metric coefficient g12 (see Section 6.1)for a surface x(u1, u2) of revolution (see Section 6.11), the tangentvectors x1 = ∂x

∂θand x2 = ∂x

∂ϕare orthogonal. These are respectively

the tangents to the latitude and the longitude. Therefore the two anglesadd up to π

2. �

Proof of Clairaut’s relation. Let pmax ∈ S2 be the pointof α(t) with maximal z-coordinate. Let ϕ(t) be the spherical coordi-nate ϕ at the point α(t), so that r(t) = sinϕ(t). Consider the sphericaltriangle with vertices at the three points α(t), pmax, and the northpole e3. By Lemma 7.6.2, the angle of the triangle at pmax is π

2.

Let the arc c = c(t) be the arc of longitude (meridian) joiningthe variable point α(t) to e3. Let the arc b be the arc of longitude(meridian) joining pmax to e3. By Lemma 7.7.2, the lengths b and c arerespectively the ϕ-coordinates of the points pmax and α(t). Then bythe law of sines,

sin c sin(π2− γ)= sin b.

Page 68: Analytic and Differential geometry Mikhail G. Katz∗

68 7. FROM CLAIRAUT’S RELATION TO GEODESIC EQUATION

Note that r(t) = sin c(t) by (7.1.1). Since b is independent of t, weobtain the relation r(t) cos γ(t) = sin b, proving Clairaut’s relation. �

7.8. Differential equation of great circle

Recall that the sphere is defined by the equation ρ = 1. Notethat the parametrisation in Clairaut’s formula need not be arclength.Assume G is not a longitude (meridian), so we can parametrize it bythe value of the spherical coordinate θ.

Remark 7.8.1. By the implicit function theorem, we can thinkof G as defined by a suitable function

ϕ = ϕ(θ).

Theorem 7.8.2. The great circle G satisfies the following differen-tial equation for ϕ = ϕ(θ):

1

r2+

1

r4

(d ϕ

)2

=1

const2, (7.8.1)

where r = sinϕ and const = sinϕmin from Theorem 7.4.1.

Proof. An element of length ds along the great circle G decom-poses into a longitudinal (along a longitude (meridian), north-south)displacement dϕ, and a latitudinal (east-west) displacement rdθ. Theseare related by

ds sin γ = dϕ

ds cos γ = rdθ

so that ds2 = (dϕ)2 + (rdθ)2. Hence

rdθ tan γ = ds sin γ = dϕ,

or tan γ =dϕ

rdθ. Expressing cosine in terms of tangent, we obtain

cos2 γ =1

1 +(dϕrdθ

)2 .

Substituting cos γ = constr

from Clairaut’s relation (Theorem 7.4.1), weobtain3

3Equivalently, 1 +(

dϕrdθ

)2=(

r

const

)2or(

dϕrdθ

)2= r2

const2− 1 or dϕ

rdθ=

√r2

const2 − 1 or 1r

dϕdθ

=√

r2

const2 − 1 or

1

sinϕ

dθ=

√sin2 ϕ

const2− 1 (7.8.2)

Page 69: Analytic and Differential geometry Mikhail G. Katz∗

7.9. METRICS CONFORMAL TO THE FLAT METRIC AND THEIR Γkij 69

(const

r

)2(1 +

(dϕ

rdθ

)2)

= 1. (7.8.3)

Equivalently,

1

r2+

1

r4

(dϕ

)2

=1

const2where r = sinϕ.

This equation is solved explicitly in terms of integrals in Example 8.2.3below. �

Remark 7.8.3. At a point where ϕ is not extremal as a functionof θ, theorem on the uniqueness of solution applies and gives a uniquegeodesic through the point.

However, at a point of maximal ϕ, the hypothesis of the uniquenesstheorem does not apply. Namely, the square root expression on theright hand side of (7.8.2) does not satisfy the Lipschitz condition asthe expression under the square root sign vanishes. In fact, uniquenessfails at this point, as a latitude (which is not a geodesic) satisfies thedifferential equation, as well. Here we have r = const, and at anextremal value of ϕ one can no longer solve the equation by separationof variables (as this would involve division by the radical expressionwhich vanishes at the extremal value of ϕ). At this point, there is adegeneracy and general results about uniqueness of solution cannot beapplied.

7.9. Metrics conformal to the flat metric and their Γkij

A particularly important class of metrics are those conformal to theflat metric, in the following sense. We will use the symbol

λ = λ(u1, u2)

for the conformal factor of the metric, as below.

Definition 7.9.1. We say that a metric is conformal to the stan-dard flat metric if there is a function λ(u1, u2) > 0 such that gij = λδijfor all i, j = 1, 2.

Let λi =∂λ∂ui

.

Lemma 7.9.2. Consider a metric conformal to the standard flatmetric, so that gij = λ(u1, u2)δij. Then we have Γ1

11 = λ12λ, Γ1

22 = −λ12λ

,

and Γ112 =

λ22λ.

The values of the coefficients are listed in Table 7.9.1.

Page 70: Analytic and Differential geometry Mikhail G. Katz∗

70 7. FROM CLAIRAUT’S RELATION TO GEODESIC EQUATION

Γ1ij j = 1 j = 2

i = 1 λ12λ

λ22λ

i = 2 λ22λ

−λ12λ

Γ2ij j = 1 j = 2

i = 1 −λ22λ

λ12λ

i = 2 λ12λ

λ22λ

Table 7.9.1. Symbols Γkij of a conformal metric λδij

Proof. The general formula, as in Section 6.10, is

Γkij =1

2(giℓ;j − gij;ℓ + gjℓ;i)g

ℓk.

By hypothesis, we have g11 = g22 = λ(u1, u2) while g12 = 0. We have

Γ1ij =

1

2λ(gi1;j − gij;1 + gj1;i),

and the lemma follows by examining the cases.For example, let i = j = 1. Then Γ1

11 =g11;12λ

= λ12λ.

Let i = j = 2. Then Γ122 =

0−g22;1+0

2λ= −λ1

2λ.

Similar calculations yield the formulas for the coefficients Γ2ij . �

Coordinates with respect to which the metric is expressed by gij =λ(u1, u2)δij are called isothermal (see Section 10.3 in more detail).

Corollary 7.9.3. A metric in isothermal coordinates (u1, u2) sat-isfies the identity Γ1

11 + Γ122 = 0 and similarly Γ2

11 + Γ222 = 0.

Proof. From Table 7.9.1 we have Γ111 + Γ1

22 =λ12λ

− λ12λ

= 0. Simi-

larly, Γ211 + Γ2

22 = −λ22λ

+ λ22λ

= 0. �

This corollary will be used in Section 10.3.

7.10. Gravitation, ants, and geodesics on a surface

What is a geodesic on a surface?A geodesic on a surface can be thought of as the path of an ant

crawling along the surface of an apple, according to Gravitation, aPhysics textbook [MiTW73, p. 3].

Remark 7.10.1. Imagine that we peel off a narrow strip of theapple’s skin along the ant’s trajectory, and then lay it out flat on atable. What we obtain is a straight line, revealing the ant’s ability totravel along the shortest path.

On the other hand, a geodesic is defined by a certain nonlinearsecond order ordinary differential equation, cf. (7.11.1). To make the

Page 71: Analytic and Differential geometry Mikhail G. Katz∗

7.11. GEODESIC EQUATION 71

geodesic equation more concrete, we will examine the case of the sur-faces of revolution. Here the geodesic equation transforms into a con-servation law (conservation of angular momentum)4 called Clairaut’srelation. The latter lends itself to a synthetic verification for sphericalgreat circles, as in Theorem 7.4.1.

We will derive the geodesic equation using the calculus of varia-tions. I once heard R. Bott point out a surprising aspect of M. Morse’sfoundational work in this area. Namely, Morse systematically used thelength functional on the space of curves. The simple idea of using theenergy functional instead of the length functional was not exploiteduntil later. The use of energy simplifies calculations considerably, aswe will see in Section 9.9.2.

7.11. Geodesic equation

Consider a plane curve Rs→α

R2

(u1,u2)where α = (α1(s), α2(s)). If a

map x : R2 → R3 is a parametrisation of a surface M , the composition

Rs→α

R2

(u1,u2)→xR3

yields a curve

β = x ◦ αon M .

Proposition 7.11.1. Every regular curve β(s) on M satisfies theidentity

β ′′ =(αi

′αj

′Γkij + αk

′′)xk +

(Lijα

j ′αi′)n

Proof. Write β = x ◦ α, then β ′ = xiαi′ by chain rule. Differenti-

atinga again, we obtain

β ′′ =d

ds(xi ◦ α)αi′ + xiα

i′′ = xijαj ′αi

′+ xkα

k′′.

Meanwhile xij = Γkijxk + Lijn. Thus

β ′′ =(Γkijxk + Lijn

)αj

′αi

′+ xkα

k′′.

Rearranging the terms proves the proposition. �

Definition 7.11.2. A curve β = x◦α is a geodesic on the surface xif the one of the following two equivalent conditions is satisfied:

4tena’ zaviti

Page 72: Analytic and Differential geometry Mikhail G. Katz∗

72 7. FROM CLAIRAUT’S RELATION TO GEODESIC EQUATION

(a) we have for each k = 1, 2,

(αk)′′

+ Γkij(αi)

(αj)′

= 0 where′

=d

ds, (7.11.1)

meaning that

(∀k) d2αk

ds2+ Γkij

dαi

ds

dαj

ds= 0;

(b) the vector β ′′ is perpendicular to the surface and one has

β ′′ = Lijαi′αj

′n. (7.11.2)

Remark 7.11.3. The equations (7.11.1) will be derived using thecalculus of variations, in Section 9.9.2. Furthermore, by Lemma 8.1.2,such a curve β must have constant speed.

Page 73: Analytic and Differential geometry Mikhail G. Katz∗

CHAPTER 8

Geodesic equation, Clairaut

8.1. Equivalence of definitions

In Section 7.11 we defined a geodesic curve β(s) = x ◦ α(s) ona surface with parametrisation x = x(u1, u2) in two ways that wereclaimed to be equivalent:

(a) we have for each k = 1, 2, (αk)′′

+ Γkij(αi)

(αj)′

= 0;

(b) the vector β ′′ is normal to the surface and β ′′ = Lijαi′αj

′n.

Proof of equivalence. If β is a geodesic, then applying Propo-

sition 7.11.1 to the effect that β ′′ =(αi

′αj

′Γkij + αk

′′)xk+

(Lijα

j ′αi′)n

we see that the tangent component vanishes and we obtain

β ′′ = Lijαi′αj

′n,

showing that the vector β ′′ is normal to the surface. On the otherhand, if β ′′ is proportional to the normal vector n, then the tangentialcomponent of β ′′ must vanish, proving the equation (7.11.2) to theeffect that β ′′ = Lijα

i′αj′n. �

Corollary 8.1.1. Every geodesic in the plane is a straight line.

Proof. In the plane, all coefficients Γkij = 0 vanish. Then the

equation becomes (αk)′′

= 0. Integrating twice, we obtain that α(s) isa linear function of s. Therefore the graph is a straight line, provingthe corollary. �

Lemma 8.1.2. A curve β satisfying the geodesic equation (7.11.1)is necessarily constant speed.

Proof. From formula (7.11.2), we have

d

ds

(‖β ′‖2

)= 2〈β ′′, β ′〉 = Lijα

i′αj′αk

′〈n, xk〉 = 0,

proving the lemma. �

73

Page 74: Analytic and Differential geometry Mikhail G. Katz∗

74 8. GEODESIC EQUATION, CLAIRAUT

8.2. Geodesics on a surface of revolution

We use the notation u1 = θ, u2 = φ for the parameters in the caseof a surface of revolution generated by a plane curve (r(φ), z(φ)), andwrite x(θ, φ). Here the function r(φ) is the distance from the point onthe surface to the z-axis.

Lemma 8.2.1. On a surface of revolution, the angle γ between aregular curve (not necessarily geodesic) β(s) (where s as the arclengthparameter) and the latitude satisfies the relation cos γ = r dθ

ds.

Proof. The tangent vector to the latitude is x1. Recall that wehave g11 = r2, in other words |x1| = r. We have

cos γ =

⟨x1|x1|

,dβ

ds

⟩where x1 =

∂x

∂θ(θ, φ).

Recall that by chain rule β ′ = x1θ′

+ x2φ′

. Thus

cos γ =1

|x1|〈x1, x1θ

+ x2φ′

︸ ︷︷ ︸chain rule

〉 = θ′

|x1|〈x1, x1〉︸ ︷︷ ︸

|x1|2

|x2|〈x1, x2〉︸ ︷︷ ︸g12=0

= θ′ |x1| = rθ

,

proving the lemma. �

Theorem 8.2.2. Let x = r(φ) cos θ, y = r(φ) sin θ, and z = z(φ)be the components of a surface of revolution. Then the geodesic equa-tion for k = 1 is equivalent to Clairaut’s relation r cos γ = const(cf. Theorem 7.4.1).

Proof. The symbols for k = 1 are given by Lemma 6.11.1. Namely,

we have Γ111 = Γ1

22 = 0, while Γ112 =

r drdφ

r2. We will use the shorthand

notation θ(s), φ(s) respectively for the components α1(s), α2(s) of thecurve. Let ′ = d

ds. The differential equation of geodesic β = x ◦ α

for k = 1 becomes

0 = θ′′ + 2Γ112θ

′φ′

= θ′′ +2r dr

r2θ′

φ′

.

Multiplying by r2 we obtain

0 = r2θ′′ + 2rdr

dφθ′

φ′

= (r2θ′

)′

Page 75: Analytic and Differential geometry Mikhail G. Katz∗

8.2. GEODESICS ON A SURFACE OF REVOLUTION 75

by the Leibniz rule. Integrating (r2θ′

)′

= 0 we obtain

r2θ′

= const. (8.2.1)

Since the geodesic β is constant speed by Lemma 8.1.2, we can as-sume the parameter s is arclength by scaling by a constant. Now, byLemma 8.2.1 we have r cos γ = r2θ

= const from (8.2.1). Thus we ob-tain r(s) cos γ(s) = const, proving Clairaut’s relation for an arbitrarysurface of revolution. �

8.2.1. This section is optional. This section uses the notation f(φ) =r(φ). In the case of a surface of revolution, the geodesic equation can beintegrated by means of primitives :

1 =

⟨dx

ds,dx

ds

⟩= 〈x1θ

+ x2φ′

, x1θ′

+ x2φ′〉

= g11(θ′

)2 + g22(φ′

)2 = f2(θ′

)2 +

(dr

)2

+

(dg

)2

︸ ︷︷ ︸g22

)2.

Thus, 1 = f2(dθ

ds

)2

+ g22

(dφ

ds

)2

. Now multiply by

(ds

)2

:

(ds

)2

= f2 + g22

(dφ

ds

ds

)2

= f2 + g22

(dφ

)2

.

Now from Clairaut’s relation we have

ds

dθ=f2

c,

hence we obtain the formula f4

c2= f2 + g22

(dφdθ

)2or

1

c2=

1

f2+g22f4

(dφ

)2

which is the equation (7.8.1) for a geodesic on the sphere. Furthermore, wehave

dθ=

√f4

c2− f2

g22=f

c

√√√√f2 − c2

(drdφ

)2+(dgdφ

)2 .

Thus

θ = c

∫1

f

√√√√(drdφ

)2+(dgdφ

)2

f2 − c2dφ+ const.

Page 76: Analytic and Differential geometry Mikhail G. Katz∗

76 8. GEODESIC EQUATION, CLAIRAUT

Example 8.2.3. On S2, we have f(φ) = sinφ, g(φ) = cosφ, thus weobtain

θ = c

∫dφ

sinφ√

sin2 φ− c2+ const,

which is the equation of a great circle in spherical coordinates. Integratingthis by a substitution u = cot φ we obtain

cotφ = a cos(θ − θ0)

where a = 1−c2c . See

http://www.damtp.cam.ac.uk/user/reh10/lectures/nst-mmii-handout2.pdf

where θ and φ are switched.

8.3. Integration in polar, cylindrical, spherical coordinates

Following some preliminaries on areas, directional derivatives, and Hes-sians, we will deal with a central object in the differential geometry ofsurfaces in Euclidean space, namely the Weingarten map, in Section 9.1.Spherical coordinates were reviewed in Section 7.1. In this section, we re-view material from calculus on polar, cylindrical, and spherical coordinatesas regards their role in integration. The polar coordinates (r, θ) in the planearise naturally in complex analysis (of one complex variable).

Definition 8.3.1. Polar coordinates (koordinatot koteviot) (r, θ) satisfyr2 = x2 + y2 and x = r cos θ, y = r sin θ.

It is shown in elementary calculus that the area of a region D ⊆ R2 inthe plane in polar coordinates is calculated using the area element

dA = r dr dθ.

Thus, an integral is of the form∫

DdA =

∫∫rdrdθ.

Cylindrical coordinates in Euclidean 3-space are studied in Vector Cal-culus.

Definition 8.3.2. 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 ⊆ R3 is calculated with respect tocylindrical coordinates using the volume element

dV = r dr dθ dz.

Namely, an integral is of the form∫

DdV =

∫∫∫rdr dθ dz.

Page 77: Analytic and Differential geometry Mikhail G. Katz∗

8.3. INTEGRATION IN POLAR, CYLINDRICAL, SPHERICAL COORDINATES77

Example 8.3.3. Find the volume of a right circular cone with height hand base a circle of radius b.

Spherical coordinates1 (ρ, θ, ϕ) in Euclidean 3-space are studied in VectorCalculus.

Definition 8.3.4. Spherical coordinates (ρ, θ, ϕ) were already definedin Section 7.1. 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 tothe (x, y)-plane, the polar coordinates of its image, (r, θ), satisfy x = r cos θand y = r sin θ. The last coordinate ϕ of the spherical coordinates is theangle between the position vector of the point and the third basis vector e3in 3-space (pointing upward along the z-axis). Thus z = ρ cosϕ while r =ρ sinϕ .

Here we have the bounds 0 ≤ ρ, 0 ≤ θ ≤ 2π, and 0 ≤ ϕ ≤ π (notethe different upper bounds for θ and ϕ). Recall that the area of a sphericalregion D ⊆ S2 on the unit sphere is calculated using an area element

sinϕ dθ dϕ .

Meanwhile the volume of a region in 3-space is calculated using the volumeelement of the form

dV = ρ2 sinϕ dρ dθ dϕ,

so that the volume of a region D is∫

DdV =

∫∫∫

Dρ2 sinϕ dρ dθ dϕ .

Example 8.3.5. Calculate the volume of the spherical shell betweenspheres of radius ρ0 > 0 and ρ1 ≥ ρ0.

The area of a spherical region on a sphere of radius ρ = ρ1 is calculatedusing the area element

dA = ρ21 sinϕ dθ dϕ .

Thus the area of a spherical region D on a sphere of radius ρ1 is given bythe integral

DdA =

∫∫ρ21 sinϕ dθ dϕ = ρ21

∫∫sinϕ dθ dϕ

Example 8.3.6. Calculate the area of the spherical region on a sphereof radius ρ1 contained in the first octant, (so that all three Cartesian coor-dinates are positive).

1koordinatot kaduriot

Page 78: Analytic and Differential geometry Mikhail G. Katz∗

78 8. GEODESIC EQUATION, CLAIRAUT

8.4. Measuring area on surfaces

The area of the parallelogram spanned by the tangent vectors x1and x2 can be calculated as the square root of their Gram matrix,namely the matrix of the first fundamental form. This motivates thefollowing definition.

Definition 8.4.1 (Computation of area). The area of the sur-face x : U → R3 is computed by integrating the area element

√g11g22 − g212 du

1du2 (8.4.1)

over the domain U of the map x. Thus area(D) =∫U

√det(gij) du

1du2

where D is the region parametrized by x(u1, u2).

The presence of the square root in the formula is explained in in-finitesimal calculus in terms of the Gram matrix, cf. (5.6.1).

Example 8.4.2. Consider the parametrisation the unit sphere pro-vided by spherical coordinates. Then the integrand is

sinϕ dθ dϕ,

and we recover the formula familiar from calculus for the area of aregion D on the unit sphere:

area(D) =

∫∫

D

sinϕ dθ dϕ .

8.5. Directional derivative

We will represent a vector v ∈ Rn as the velocity vector v = dαdt

of acurve α(t), at t = 0. Typically we will be interested in the case n = 3(or 2).

Definition 8.5.1. Given a function f of n variables, its directionalderivative2 ∇vf at a point p ∈ Rn, in the direction of a vector v isdefined by setting

∇vf =d (f ◦ α(t))

dt

∣∣∣∣t=0

where α(0) = p.

Lemma 8.5.2. The definition of directional derivative is indepen-dent of the choice of the curve α(t) representing the vector v.

Proof. The lemma is proved in Elementary Calculus [Ke74]. �

Let p = x(u1, u2) be a point of a surface in R3.

2nigzeret kivunit

Page 79: Analytic and Differential geometry Mikhail G. Katz∗

8.6. EXTENDING V.F. ALONG SURFACE TO AN OPEN SET IN R3 79

Definition 8.5.3 (Orthogonal decomposition). The tangent planeto the surface x = x(u1, u2) at the point p, denoted Tp, was definedin Section 5.6 and is the plane passing through p and spanned byvectors x1 and x2, or alternatively as the plane perpendicular to thenormal vector n at p, resulting in an orthogonal decomposition

R3 = Tp + Rn.

Example 8.5.4. Suppose x(u1, u2) is a parametrisation of the unitsphere S2 ⊆ R3. At a point (a, b, c) ∈ S2, the normal vector is the

position vector itself: n =

abc

. In other words, n(u1, u2) = x(u1, u2).

Note that the sphere is defined implicitly by F (x, y, z) = 0 whereF (x, y, z) = x2 + y2+ z2 − 1. We have ∇F = (2x, 2y, 2z) and thereforethe gradient ∇F at a point is proportional to the radius vector of thepoint on the sphere.

8.6. Extending v.f. along surface to an open set in R3

Remark 8.6.1. The notion of vector field is defined in infi 4 butonly the maslul of applied mathematics requires it. The maslul of puremathematics takes course on rings instead.

Now let us return to the set-up

Rt→α

R2

(u1,u2)→xR3.

We consider the curve β = x ◦ α. Let v ∈ Tp be a tangent vector at a

point p ∈ M , defined by v = dβdt

∣∣t=0

, where β(0) = p. By chain rule,we have

v =dαi

dtxi.

Now consider the normal vector

n ◦ α(t)along β(t). This vector varies from point to point on the surface, but isnot defined in any open neighborhood in R3. We would like to extendit to a vector field in an open neighborhood of p ∈ R3, so as to be ableto differentiate it.

Remark 8.6.2. The curve β represents the class of curves withinitial vector v.

Lemma 8.6.3. The gradient ∇F of a function F = F (x, y, z) at apoint where the gradient is nonzero, is perpendicular to the level sur-face F (x, y, z) = 0 of the function F .

Page 80: Analytic and Differential geometry Mikhail G. Katz∗

80 8. GEODESIC EQUATION, CLAIRAUT

This was shown in elementary calculus.

Theorem 8.6.4. Given a regular parametrisation x(u1, u2) of Mand its normal vector n = n(u1, u2), one can extend n to a differentiablevector field N(x, y, z) defined in an open neighborhood of p ∈ M ⊆ R3,in the sense that we have

n(u1, u2) = N(x(u1, u2)

). (8.6.1)

Proof. We apply a version of the implicit function theorem forsurfaces to represent the surface implicitly by an equation F (x, y, z) =0, where the function F is defined in an open neighborhood of p ∈ M ,and ∇F 6= 0 at p. By Lemma 8.6.3, the normalisation

N =1

|∇F |∇F

of the gradient ∇F of F gives the required extension in the senseof (8.6.1) of the normal n. �

8.7. Differentiating n and N

We will need the directional derivative to define the Weingartenmap in Section 9.1.

Proposition 8.7.1. Let p ∈M , and v ∈ TpM where v = β ′(0) andβ(t) = x(α(t)). Let N = N(x, y, z) be a vector field in R3 extendingthe normal vector field along the surface, namely n(u1, u2). Then thedirectional derivative ∇vN satisfies

∇vN =d (n ◦ α(t))

dt

∣∣∣∣t=0

.

Proof. By (8.6.1), the function n(t) satisfies the relation

n(α(t)) = N(β(t)),

where β = x ◦ α. The directional derivative ∇vN can be calculatedusing this particular curve β since by Lemma 8.5.2, the gradient isindependent of the choice of the curve. We therefore obtain

∇vN =d (N ◦ β(t))

dt=d (n ◦ α(t))

dt,

proving the proposition. �

Page 81: Analytic and Differential geometry Mikhail G. Katz∗

8.8. HESSIAN OF A FUNCTION AT A CRITICAL POINT 81

8.8. Hessian of a function at a critical point

This section is intended to motivate the definition of the Weingartenmap in Section 9.1. The key observation is Remark 8.8.3 below, relatingthe Hessian and the Weingarten map.

Consider the graph in R3 of a function f(x, y) of two variables inthe neighborhood of a critical point p, where ∇f = 0.

Remark 8.8.1. The tangent plane of the graph of f at a criticalpoint p of f is a horizontal plane.

The Hessian (matrix of second derivatives) of the function at thecritical point captures the main features of the local behavior of thefunction in a neighborhood of the critical point up to negligible higherorder terms. Thus, we have the following typical result concerning thesurface given by the graph of the function in R3.

Theorem 8.8.2. At a critical point p of f , assume that the eigen-values of the Hessian Hf(p) are nonzero.

(1) If the eigenvalues of the Hessian have opposite sign, then thegraph of f is a saddle point.

(2) If the eigenvalues have the same sign, the graph is a local min-imum or maximum.

Remark 8.8.3. If one thinks of the Hessian matrix as defining alinear transformation (an endomorphism) of the horizontal plane, givenby the matrix of second derivatives, then the Hessian of a function ata critical point becomes a special case of the Weingarten map, definedin Section 9.1.

Page 82: Analytic and Differential geometry Mikhail G. Katz∗
Page 83: Analytic and Differential geometry Mikhail G. Katz∗

CHAPTER 9

Weingarten map

9.1. Definition of the Weingarten map

The definition of the Weingarten operator can be viewed as ananalogue of a formula we saw in the context of curves.

Remark 9.1.1. For a plane curve α(s) with tangent vector v(s)and normal vector n(s), we had the equation

d

dsn(s) = ±kα(s)v(s); (9.1.1)

see Proposition 3.11.3. Thus, curvature is closely related to the rate ofchange of the normal vector.

The Weingarten mapW ofM can be thought of as a surface analogof formula (9.1.1). The point is that W carries the information aboutthe curvature of M .

In Section 8.8, we considered the special case of a surface given bythe graph of a function f of two variables near a critical point of f .Now consider the more general framework of a parametrized regularsurface M in R3 with a regular parametrisation x(u1, u2).

Remark 9.1.2. Instead of working with a matrix of second deriva-tives, we will give a definition of an endomorphism of the tangent plane.This is basically a coordinate-free way of talking about the Hessian ma-trix.

As in Theorem 8.6.4, we extend the normal vector field n along thesurface to a vector field N(x, y, z) defined in an open neighborhoodof p ∈ R3, so that we have n(u1, u2) = N (x(u1, u2)).

Definition 9.1.3. Let p ∈ M . Denote by Tp = TpM the tangentplane to the surface at p. The Weingarten map (also known as theshape operator)

Wp : Tp → Tp

83

Page 84: Analytic and Differential geometry Mikhail G. Katz∗

84 9. WEINGARTEN MAP

is the endomorphism of the tangent plane given by directional deriva-tive of the vector-valued function N (extension of n as above):

Wp(v) = ∇vN =d

dt

∣∣∣∣t = 0

n ◦ α(t), (9.1.2)

where the curve β = x ◦ α is chosen so that β(0) = p while β ′(0) = v.

9.2. Properties of Weingarten map

The Weingarten map Wp : Tp → Tp at a point p ∈ M was definedin Section 9.1.

Lemma 9.2.1. The map Wp is well defined in that the image isincluded in the tangent plane Tp at p.

Proof. We have to show that the right hand side of formula (9.1.2),which is a priori a vector in R3, indeed produces a vector in the tangentplane. By Leibniz’s rule,

〈Wp(v), n(α(t))〉 = 〈∇vN, n(α(t))〉 =1

2

d

dt〈n ◦ α(t), n ◦ α(t)〉 = 0

since n is a unit vector. Therefore the image vectorWp(v) is orthogonalto n. Hence it lies in Tp, proving the lemma. �

The connection with the matrix of second derivatives is given in thefollowing theorem.

Theorem 9.2.2. The operator Wp is a selfadjoint endomorphism

of the tangent plane Tp, and satisfies 〈Wp(xi), xj〉 = −⟨n, ∂2x

∂ui∂uj

⟩.

Proof. By definition of the Weingarten map, ∇vN is the deriva-tive of N along a curve with initial vector v ∈ Tp. To simplify notation,assume that p = x(0, 0). We can choose the particular “coordinate”curve of the first coordinate, γ(t) = x(t, 0). Then we have γ′(t) = x1.Therefore

∇x1N =d

dtN(γ(t)) =

∂n

∂u1

by definition of the directional derivative, and similarly ∇x2N = ∂n∂u2

.Thus the coordinate lines u1 7→ x(u1, 0) and u2 7→ x(0, u2) of thechart (u1, u2) allow us to write

∇xjN =∂

∂ujn for each j = 1, 2.

Page 85: Analytic and Differential geometry Mikhail G. Katz∗

9.3. RELATION TO THE HESSIAN OF A FUNCTION 85

Therefore we have by Leibniz rule

〈W (xi), xj〉 =⟨∂n

∂ui, xj

=∂

∂ui〈n, xj〉 − 〈n, ∂

∂uixj〉

= −⟨n,

∂2x

∂ui∂uj

⟩,

which is an expression symmetric in i and j. Thus,

〈W (xi), xj〉 = 〈xi,W (xj)〉.This proves the selfadjointness of W by verifying it for a set of basisvectors. �

Theorem 9.2.3. The eigenvalues of the Weingarten map are real.

By Corollary 2.4.2, this is immediate from the fact that the endo-morphism Wp of the vector space TpM is selfadjoint.

9.3. Relation to the Hessian of a function

Theorem 9.3.1 (Relation to the Hessian). Given a function f oftwo variables, consider its graph x(u1, u2) = (u1, u2, f(u1, u2)). Then ata critical point p of f , the inner products 〈Wp(xi), xj〉 form the Hessianmatrix of second derivatives Hf of f .

Proof. The second partial derivatives of the parametrisation arexij = (0, 0, fij)

t. and n = e3 at a critical point p. Therefore we obtain〈xij, n〉 = fij up to sign, as required. �

Definition 9.3.2. The principal curvatures at p ∈ M , denoted k1and k2, are the eigenvalues of the Weingarten map Wp.

These will be discussed in more detail in Section 10.1 as well asDefinition 9.9.10.

Example 9.3.3. Consider the plane x(u1, u2) = (u1, u2, 0). Wehave x1 = e1, x2 = e2, while the normal vector field n

n =x1 × x2(x1 × x2)

= e1 × e2 = e3

is constant. It can therefore be extended to a constant vector field Ndefined in an open neighborhood in R3. Thus ∇vN ≡ 0 andWp(v) ≡ 0,and the Weingarten map is identically zero.

In the next section we will present nonzero examples of the Wein-garten map.

Page 86: Analytic and Differential geometry Mikhail G. Katz∗

86 9. WEINGARTEN MAP

9.4. Weingarten map of sphere and cylinder

Theorem 9.4.1. The Weingarten map of the sphere of radius r > 0at every point p of the sphere is the scalar map Wp : Tp → Tp given by1rId where Id is the identity map of Tp.

Proof. Represent a tangent vector v ∈ Tp by v = β ′(0). On the

sphere of radius r, we have n ◦ α(t) =1

rβ(t) (normal vector is the

normalized position vector). Hence

Wp(v) = ∇vN(β(t))

=d

dt

∣∣∣∣∣t=0

n ◦ α(t)

=1

r

d

dt

∣∣∣∣∣t=0

β(t)

=1

rv.

Thus Wp(v) =1

rv for all v. In other words, Wp =

1

rId. �

Note that the Weingarten map has rank 2 in this case.

Example 9.4.2 (Cylinder). For the cylinder, we have the parametri-sation

x(u1, u2) = (cosu1, sin u1, u2),

with x1 = (− sin u1, cosu1, 0)t and n = (cosu1, sin u1, 0)t. As before,this can be extended to a vector field N defined in an open neighbor-hood in R3, with the usual relation ∇xiN = ∂

∂uin. Hence we have

∇x1N =∂

∂u1n = (− sin u1, cosu1, 0). (9.4.1)

Similarly,

∇x2N =∂

∂u2n = (0, 0, 0). (9.4.2)

Now let v = vixi be an arbitrary tangent vector. Then

∇vN = ∇vixiN = v1∇x1N + v2∇x2N

by linearity of directional derivatives. Hence (9.4.1) and (9.4.2) yield

∇vN = v1∇x1N = v1

− sin u1

cosu1

0

,

Page 87: Analytic and Differential geometry Mikhail G. Katz∗

9.6. GAUSSIAN CURVATURE 87

and therefore,

Wp(v) = v1

− sin u1

cos u1

0

= v1x1,

where v = vixi. Thus, Wp(x1) = x1, while Wp(x2) = 0. Note that theWeingarten map has rank 1 in this case.

9.5. Coefficients Lij of Weingarten map

Since the two vectors (x1, x2) from a basis of the tangent plane Tp,we can introduce the following definition.

Definition 9.5.1. The coefficients Lij of the Weingarten map aredefined by W (xj) = Lijxi = L1

j x1 + L2j x2.

Example 9.5.2. For the plane, we have (Lij) =

(0 00 0

).

Example 9.5.3. For the sphere, we have (Lij) =

(1r

00 1

r

)= 1

rδij .

Example 9.5.4. For the cylinder, we have L11 = 1. The remaining

coefficients vanish, so that (Lij) =

(1 00 0

).

9.6. Gaussian curvature

We continue with the terminology and notation of the previoussection. The skew-symmetrisation notation was defined in Section 1.9.The signed curvatures will be defined in Section 11.1.

Definition 9.6.1. TheGaussian curvature function, denotedK =K(u1, u2), of a surface M with a regular parametrisation x(u1, u2) canbe defined in the following two equivalent ways:

(1) K the determinant of the Weingarten map:

K = det(Lij) = L11L

22 − L1

2L21 = 2L1

[1L22];

(2) K is the product of signed curvatures k1 and k2 of the curvesM ∩ P1 and M ∩ P2, where the planes Pi are defined by Pi =Span(n, vi) and v1, v2 are the eigenvectors of the Weingartenmap.

We will mostly work with definition (1) of Gaussian curvature; weincluded definition (2) above for general culture as it was the originaldefinition by Gauss.

Page 88: Analytic and Differential geometry Mikhail G. Katz∗

88 9. WEINGARTEN MAP

Example 9.6.2. Cylinder, plane K = 0, cf. Figure 12.2.1. For a

sphere of radius r, we have K = det

(1r

00 1

r

)=

1

r2.

Remark 9.6.3 (Sign of Gaussian curvature). Of particular geomet-ric significance is the sign of the Gaussian curvature. The geometricmeaning of negative Gaussian curvature is a saddle point. The geo-metric meaning of positive Gaussian curvature is a point of convexity,such as local minimum or local maximum of the graph of a function oftwo variables.

9.7. Second fundamental form

Recall that we extended the normal vector field n(u1, u2) along thesurface M near a point p ∈ M , to a smooth vector field N(x, y, z)defined in an open neighborhood of p ∈ R3.

Definition 9.7.1. The second fundamental form IIp is the bilinearform on the tangent plane Tp defined for u, v ∈ Tp by

IIp(u, v) = −〈∇uN, v〉.It may be helpful to keep in mind the observation that the second

fundamental form measures the curvature of geodesics on x viewed ascurves of R3 (cf. 3.10.1) as illustrated by Theorem 9.8.1 below.

Definition 9.7.2. The coefficients Lij of the second fundamentalform are defined to be

Lij = IIp(xi, xj) = −⟨∂n

∂ui, xj

⟩.

Lemma 9.7.3. The coefficients Lij of the second fundamental formare symmetric in i and j, more precisely Lij = +〈xij , n〉.

Proof. We have 〈n, xi〉 = 0. Hence ∂∂uj

〈n, xi〉 = 0, i.e.⟨

∂ujn, xi

⟩+ 〈n, xij〉 = 0

or〈n, xij〉 = −〈∇xjn, xi〉 = +II(xj , xi)

and the proof is concluded by the equality of mixed partials. �

Lemma 9.7.4. The second fundamental form allows us to identifythe normal component of the second partial derivatives of the parametri-sation x; namely we have

xij = Γkijxk + Lijn.

Page 89: Analytic and Differential geometry Mikhail G. Katz∗

9.8. GEODESICS AND SECOND FUNDAMENTAL FORM 89

Proof. Let xij = Γkijxk+ cn and form the inner product with n toobtain Lij = 〈xij , n〉 = 0 + c〈n, n〉 = c. �

9.8. Geodesics and second fundamental form

Theorem 9.8.1. Let β(s) be a unit speed geodesic on a surfaceM ⊆R3, so that β ′(s) ∈ TpM at a point p = β(s). Then the absolute valueof the second fundamental form applied to β ′ is the curvature of β at p:

|IIp(β ′, β ′)| = kβ(s),

where kβ is the curvature of β as a curve in R3.

Proof. We apply formula (7.11.2). Since |n| = 1, we have

kβdef= |β ′′| =

∣∣∣Lijαi′αj ′∣∣∣ .

Recall that β = x ◦ α. We have

II(β ′, β ′) = II(xiαi′, xjα

j ′) = αi′αj

′II(xi, xj) = αi

′αj

′Lij ,

completing the proof. �

Proposition 9.8.2. We have the following relation between theWeingarten map and the second fundamental form:

Lij = −Lkjgki.

Proof. By definition,

Lij = 〈xij , n〉 = −⟨

∂ujn, xi

⟩= −〈Lkjxk, xi〉 = −Lkjgki

proving the lemma. �

Corollary 9.8.3. We have the following relation: Li j = −gikLkj.

Proof. We start with the relation Lij = −Lkjgki of Proposi-tion 9.8.2, multiply it on both sides by giℓ, and sum over i, obtaining

Lijgiℓ = −Lkjgkigiℓ

= −Lkjδℓk= −Lℓj

as required. �

Page 90: Analytic and Differential geometry Mikhail G. Katz∗

90 9. WEINGARTEN MAP

9.9. Three formulas for Gaussian curvature

Remark 9.9.1. We will continue making use of the Einstein sum-mation convention, i.e. suppress the summation symbol

∑when the

summation index occurs simultaneously as a subscript and a super-script in a formula, cf. (10.7.1).

Theorem 9.9.2. We have the following three equivalent formulasfor the Gaussian curvature:

(a) K = det(Lij) = 2L1[1L

22];

(b) K = det(Lij)/det(gij);(c) K = − 2

g11L1[1L

22].

Proof. Here formula (a) is our definition of K.To prove formula (b), we use the formula Lij = −Lkjgki of Propo-

sition 9.8.2. By the multiplicativity of determinant with respect tomatrix multiplication,

det(Lij) = (−1)2det(Lij)/det(gij).

Let us now prove formula (c). The proof is a calculation. Note that bydefinition, 2L1[1L

22] = L11L

22 − L12L

21. Hence

− 2

g11

(L1[1L

22]

)=

1

g11

((L1

1g11 + L21g21

)L2

2 −(L1

2g11 + L22g21

)L2

1

)

=1

g11

(g11L

11L

22 + g21L

21L

22 − g11L

12L

21 − g21L

22L

21

)

= det(Li j) = K

as required. �

Remark 9.9.3. We can either calculate using the formula Lij =−Lkjgki, or the formula Lij = −Lkigkj. Only the former one leads tothe appropriate cancellations as above.

9.9.1. Normal and geodesic curvatures. The material in this sec-tion is optional. Let x(u1, u2) be a surface in R3. Let α(s) = (α1(s), α2(s))a curve in R2, and consider the curve β = x ◦ α on the surface x. Con-sider also the normal vector to the surface, denoted n. The tangent unitvector β′ = dβ

ds is perpendicular to n, so β′, n, β′ × n are three unit vector

spanning R3. As β′ is a unit vector, it is perpendicular to β′′, and there-fore β′′ is a linear combination of n and β′×n. Thus β′′ = knn+ kg(β

′×n),where kn, kg ∈ R are called the normal and the geodesic curvature (resp.).Note that kn = β′′ · n, kg = β′′ · (β′ ×n). If k is the curvature of β, then wehave that k2 = ||β′′||2 = k2n + k2g .

Exercise 9.9.4. Let β(s) be a unit speed curve on a sphere of radius r.Then the normal curvature of β is ±1/r.

Page 91: Analytic and Differential geometry Mikhail G. Katz∗

9.9. THREE FORMULAS FOR GAUSSIAN CURVATURE 91

Remark 9.9.5. Let π be a plane passing through the center of the sphereS in Exercise 9.9.4, and Let C = π ∩ S. Then the curvature k of C is 1/r,and thus kg = 0. On the other hand, if π does not pass through the centerof S (and π ∩ S 6= ∅) then the geodesic curvature of the intersection 6= 0.

Proposition 9.9.6. If a unit speed curve β on a surface is geodesic thenits geodesic curvature is kg = 0.

Proof. If β is a geodesic, then β′′ is parallel to the normal vector n, soit is perpendicular to n× β′′, and therefore kg = β′′ · (n× β′′) = 0. �

Exercise 9.9.7. The inverse direction of Proposition 9.9.6 is also cor-rect. Prove it.

Example 9.9.8. Any line γ(t) = at + b on a surface is a geodesic, asγ′′ = 0 and therefore kg = 0.

Example 9.9.9. Take a cylinder x of radius 1 and intersect it with aplane π parallel to the xy plane. Let C = x ∩ π - it is a circle of radius 1.Thus k = 1. Show that kn = 1 and thus C is a geodesic.

Definition 9.9.10. The principal curvatures, denoted k1 and k2, arethe eigenvalues of the Weingarten map W .

Assume k1 > k2.

Theorem 9.9.11. The minimal and maximal values of the absolute valueof the normal curvature |kn| at a point p of all curves on a surface passingthrough p are |k2| and |k1|.

Proof. This is proven using the fact that kg = 0 for geodesic curvesand from Theorem 3.9.4?1 �

9.9.2. Calculus of variations and the geodesic equation. Thematerial in this subsection is optional. Calculus of variations is known astachsiv variatsiot. Let α(s) = (α1(s), α2(s)), and consider the curve β =x ◦ α on the surface given by [a, b]s→αR2 →xR3. Consider the energy

functional E(β) =∫ ba ‖β′‖2ds. Here ′ = d

ds . We have dβds = xi(α

i)′

by chainrule. Thus

E(β) =∫ b

a〈β′, β′〉ds =

∫ b

a

⟨xiα

i′, xjαj ′⟩ds =

∫ b

agijα

i′αj′ds. (9.9.1)

Consider a variation α(s) → α(s)+ tδ(s), t small, such that δ(a) = δ(b) = 0.If β = x ◦α is a critical point of E , then for any perturbation δ vanishing at

1See which theorem this is.

Page 92: Analytic and Differential geometry Mikhail G. Katz∗

92 9. WEINGARTEN MAP

the endpoints, the following derivative vanishes:

0 =d

dt

∣∣∣∣∣t = 0

E(x ◦ (α+ tδ))

=d

dt

∣∣∣∣∣t = 0

{∫ b

agij(α

i + tδi)′

(αj + tδj)′

ds

}(from equation (9.9.1))

=

∫ b

a

∂ (gij ◦ (α + tδ))

∂tαi

αj′

ds

︸ ︷︷ ︸A

+

∫ b

agij(αi

δj′

+ αj′

δi′)ds

︸ ︷︷ ︸B

,

so that we have

A+B = 0. (9.9.2)

We will need to compute both the t-derivative and the s-derivative of thefirst fundamental form. The formula is given in the lemma below.

Lemma 9.9.12. The partial derivatives of gij = gij ◦ (α(s) + tδ(s))along β = x ◦ α are given by the following formulas:

∂t(gij ◦ (α+ tδ)) = (〈xik, xj〉+ 〈xi, xjk〉)δk,

and∂gik∂s

= (〈xim, xk〉+ 〈xi, xkm〉)(αm)′

.

Proof. We have

∂t(gij ◦ (α+ tδ)) =

∂t〈xi ◦ (α+ tδ)), xj ◦ (α+ tδ)〉

=

⟨∂

∂t(xi ◦ (α+ tδ)), xj

⟩+

⟨xi,

∂t(xj ◦ (α+ tδ))

= 〈xikδk, xj〉+ 〈xi, xjkδk〉 = (〈xik, xj〉+ 〈xi, xjk〉)δk.

Furthermore, g′ik = 〈xi, xk〉′ = 〈xim(αm)′

, xk〉+ 〈xi, xkm(αm)′〉. �

Lemma 9.9.13. Let f ∈ C0[a, b]. Suppose that for all g ∈ C0[a, b] we

have∫ ba f(x)g(x)dx = 0. Then f(x) ≡ 0. This conclusion remains true if

we use only test functions g(x) such that g(a) = g(b) = 0.

Proof. We try the test function g(x) = f(x). Then∫ ba (f(x))

2ds = 0.

Since (f(x))2 ≥ 0 and f is continuous, it follows that f(x) ≡ 0. If wewant g(x) to be 0 at the endpoints, it suffices to choose g(x) = (x− a)(b −x)f(x). �

Theorem 9.9.14. Suppose β = x ◦ α is a critical point of the energyfunctional (endpoints fixed). Then β satisfies the differential equation

(∀k) (αk)′′+ Γkij(α

i)′(αj)

′= 0.

Page 93: Analytic and Differential geometry Mikhail G. Katz∗

9.9. THREE FORMULAS FOR GAUSSIAN CURVATURE 93

Proof. We use Lemma 9.9.12 to evaluate the termA from equation (9.9.2)as follows:

A =

∫ b

a(〈xik, xj〉+ 〈xi, xjk〉)(αi)

(αj)′

δkds

= 2

∫ b

a〈xik, xj〉αi

αj′

δkds,

since summation is over both i and j. Similarly, B = 2∫ ba gijα

i′

δj′

ds =

−2∫ ba

(gijα

i′)′

δjds by integration by parts, where the boundary term van-

ishes since δ(a) = δ(b) = 0. Hence

B = −2

∫ b

a

(gikα

i′)′δkds

by changing an index of summation. Thus

1

2

d

dt

∣∣∣∣∣t=0

(E) =∫ b

a

{〈xik, xj〉αi′αj ′ −

(gikα

i′)′}

δkds.

Since this is true for any variation (δk), by Lemma 9.9.13 we obtain theEuler-Lagrange equation

(∀k) 〈xik, xj〉αi′αj ′ −(gikα

i′)′

≡ 0,

or〈xik, xj〉αi

′αj

′ − gik′αi

′ − gikαi′′ = 0. (9.9.3)

Using the formula from Lemma 9.9.12 for the s-derivative of the firstfundamental form, we can rewrite the formula (9.9.3) as follows:

0 = 〈xik, xj〉αi′

αj′

− 〈xim, xk〉αm′

αi′

− 〈xi, xkm〉αi′

αm′ − gikα

i′′

= −〈Γnimxn, xk〉αm′

αi′

− gikαi′′

= −Γnimgnkαm′

αi′

− gikαi′′

,

where the cancellation of the first and the third term in the first line resultsfrom replacing index i by j and m by i in the third term. This is true ∀k.Now multiply by gjk :

gjkΓnimgnkαm′

αi′

+ gjkgikαi′

= δjnΓnimα

m′

αi′

+ δjiαi′′

= Γjimαm′

αi′

+ αj′′

= 0,

which is the desired geodesic equation. �

Page 94: Analytic and Differential geometry Mikhail G. Katz∗
Page 95: Analytic and Differential geometry Mikhail G. Katz∗

CHAPTER 10

Principal curvatures, minimal surfaces, thmegregium

10.1. Principal curvatures

In Section 9.1 we defined the Weingarten map Wp : Tp → Tp whichis an endomorphism of the tangent plane Tp at a point p ∈ M of asurface M .

Definition 10.1.1. The principal curvatures, denoted k1 and k2,are the eigenvalues of the Weingarten map W .

Remark 10.1.2. The curvatures k1 and k2 are real numbers. Thisfollows from the selfadjointness of W (Theorem 9.2.2) together withCorollary 2.4.2.

Theorem 10.1.3. The Gaussian curvature K(u1, u2) = det(Wp) ata point p = x(u1, u2) equals the product of the principal curvatures at p.

Proof. The determinant of a 2 by 2 matrix equals the product ofits eigenvalues: K = det(Lij) = k1k2. �

Theorem 10.1.4. Let v ∈ Tp be a unit eigenvector belonging toa principal curvature k at the point p. Let β(s) be a geodesic on Msatisfying β ′(0) = v. Then the curvature of β as a space curve is theabsolute value of k :

kβ(0) = |k|.

Corollary 10.1.5. The absolute value of the Gaussian curvatureat a point p of a regular surface in R3 is the product of curvatures oftwo perpendicular geodesics passing through p, whose tangent vectorsare eigenvectors of the Weingarten map at the point.

Remark 10.1.6. The shortcoming of this corollary is the fact thatat this point we are only able to express the absolute value |K| interms of curvatures of curves through p. This will be rectified when weintroduce a finer invariant called signed curvature of an oriented planecurve in Section 11.1.

95

Page 96: Analytic and Differential geometry Mikhail G. Katz∗

96 10. PRINCIPAL CURVATURES, MINIMAL SURFACES, THM EGREGIUM

Proof of Theorem 10.1.4. Let v = vixi be a unit eigenvectorof W represented by a curve β(s) = x ◦ α(s), so that vi = αi

′. Since β

is a unit speed geodesic, we have β ′′ = Lijαi′αj

′n. Hence

kβ(0) = ‖β ′′(0)‖=∣∣∣Lijαi′αj ′

∣∣∣= |IIp(β ′, β ′)|= |⟨Wp (β

′), β ′⟩ |= |〈kβ ′, β ′〉|= |k〈β ′, β ′〉|= |k|

proving the theorem. �

10.2. Mean curvature, minimal surfaces

Definition 10.2.1. The mean curvature H of the surface M at apoint p is half the trace of the Weingarten map: H = 1

2trace(Wp) =

1

2(k1+k2) =

1

2Lii. A surface is called minimal if H = 0 at every point,

i.e. k1 + k2 = 0.

See Table 12.2.1 on plane and cylinder.

Theorem 10.2.2 (Scherk surface). The Scherk surface, parametrizedby (x, y, ln( cos y

cos x)), is a minimal surface.

Proof. Let f(x) = ln cos x. Then the Scherk surface is by defini-tion x(x, y) = (x, y, f(y) − f(x)). Then x12 = 0 and hence L12 = 0.Thus the matrix (Lij) is diagonal. By Corollary 9.8.3, the mean cur-vature H satisfies 2H = traceWp = L11g

11 + L22g22 and therefore the

condition H = 0 is equivalent to

L11g22 + L22g11 = 0. (10.2.1)

Let h(x) = f ′(x). We have x1 = (1, 0,−h(x))t and x2 = (0, 1, h(y))t.Then g11 = 1 + h2(x) and g22 = 1 + h2(y).

The normal vector is the normalisation of (−h(x), h(y), 1)t. Let C =√1 + h2(x) + h2(y). Then CL11 = −f ′′(x) and CL22 = f ′′(y). Thus

the condition H = 0 expressed by (10.2.1) is equivalent to

f ′′(x)(1 + h2(y)) = f ′′(y)(1 + h2(x)).

Therefore it is sufficient to show that the function 1+h2(x)f ′′(x)

is a constant

function of x. Indeed, if f(x) = ln cos x then h(x) = f ′(x) = − tan x

Page 97: Analytic and Differential geometry Mikhail G. Katz∗

10.3. MINIMAL SURFACES IN ISOTHERMAL COORDINATES 97

Figure 10.2.1. Scherk’s minimal surface

and f ′′(x) = −1cos2 x

. Therefore 1+h2(x)f ′′(x)

= −(1 + tan2 x) cos2 x = −1 as

required. See Figure 10.2.1. �

Remark 10.2.3. Geometrically, a minimal surface is representedby a soap film.1 See Figure 10.2.2 for another example (helicoid).

10.3. Minimal surfaces in isothermal coordinates

In this section and the following, we study the relation between min-imal surfaces and harmonic functions familiar from complex functiontheory.

Definition 10.3.1. A parametrisation x(u1, u2) is called isothermalif there is a function f = f(u1, u2) such that gij = f 2δij , i.e. 〈x1, x1〉 =〈x2, x2〉 and 〈x1, x2〉 = 0.

Definition 10.3.2. The function f 2 is called conformal factor ofthe metric.

Proposition 10.3.3. Assume x(u1, u2) is isothermal. Then it sat-

isfies the partial differential equation ∂2x∂(u1)2

+ ∂2x∂(u2)2

= −2f 2Hn.

1krum sabon, as opposed to bu’at sabon. Dip a wire (tayil) into soapy water.

Page 98: Analytic and Differential geometry Mikhail G. Katz∗

98 10. PRINCIPAL CURVATURES, MINIMAL SURFACES, THM EGREGIUM

Figure 10.2.2. Helicoid: a minimal surface

Proof. We have ∆x = x11+x22 = Γ111x1+Γ2

11x2+L11n+Γ122x1+

Γ222x2 + L22n = (Γ1

11 + Γ122) x1 + (Γ2

11 + Γ222)x2 + (L11 + L22)n. But by

Corollary 7.9.3, with respect to isothermal coordinates we necessarilyhave Γ1

11 +Γ122 = 0 and Γ2

11 +Γ222 = 0. Therefore ∆x = (L11 + L22)n =

−(L11 + L2

2)f2n = −2Hf 2n as required. �

10.3.1. Alternative proof. This subsection is optional.

Proof. Note that Lij = −Lmjgmi = −Lmjf 2δmi = −f 2Lij. Thus

Lij = −Lijf 2

so that the mean curvature H satisfies

H =1

2Lii = −L11 + L22

2f 2. (10.3.1)

Differentiating 〈x1, x2〉 = 0 we obtain ∂∂u2

〈x1, x2〉 = 0. Therefore

〈x12, x2〉+ 〈x1, x22〉 = 0. (10.3.2)

So −〈x12, x2〉 = 〈x1, x22〉. By Definition 10.3.1 of isothermal coordi-nates, we have

〈x1, x1〉 − 〈x2, x2〉 = 0. (10.3.3)

Page 99: Analytic and Differential geometry Mikhail G. Katz∗

10.4. MINIMALITY AND HARMONIC FUNCTIONS 99

Differentiating (10.3.3) and applying formula (10.3.2), we obtain

0 =∂

∂u1〈x1, x1〉 −

∂u1〈x2, x2〉

= 2〈x11, x1〉 − 2〈x21, x2〉= 2〈x11, x1〉+ 2〈x22, x1〉= 2〈x11 + x22, x1〉.

Inspecting the u2-derivatives, we similarly obtain 〈x11 + x22, x2〉 = 0.Since x1, x2 and n form an orthogonal basis, the sum x11 + x22 isproportional to n. Write x11 + x22 = cn. Applying (10.3.1), we obtain

c = 〈x11 + x22, n〉= 〈x11, n〉+ 〈x22, n〉= L11 + L22

= −2f 2H,

as required. �

10.4. Minimality and harmonic functions

We first recall a definition from Section 4.3.

Definition 10.4.1. Let F : R2 → R, (u1, u2) 7→ F (u1, u2). TheLaplacian, denoted ∆(F ), is the second-order differential operator onfunctions, defined by

∆F =∂2F

∂(u1)2+

∂2F

∂(u2)2.

Functions in the kernel of the Laplacian are called harmonic:

Definition 10.4.2. We say that F is harmonic if ∆F = 0.

Harmonic functions are important in the study of heat flow, or heattransfer.2

Corollary 10.4.3. Let x(u1, u2) = (x(u1, u2), y(u1, u2), z(u1, u2))in R3 be a parametrisation of a surfaceM and assume that x is isother-mal. Then M is minimal if and only if the coordinate functions x, y, zare harmonic.

Proof. From Proposition 10.3.3 we have

‖∆(x)‖ =√

(∆x)2 + (∆y)2 + (∆z)2 = 2f 2|H|,proving the corollary. �

2ma’avar chom.

Page 100: Analytic and Differential geometry Mikhail G. Katz∗

100 10. PRINCIPAL CURVATURES, MINIMAL SURFACES, THM EGREGIUM

Example 10.4.4. The catenoid is parametrized as follows:

x(θ, φ) = (a coshφ cos θ, a coshφ sin θ, aφ).

Here f(φ) = a cosh φ, while g(φ) = aφ. Then g11 = a2 cosh2 φ =f 2, g22 = (a sinhφ)2 + a2 = a2 cosh2 φ = f 2, g12 = 0. Calculate:

x11 + x22 = (−a cosh φ cos θ,−a cosh φ sin θ, 0)+ (a coshφ cos θ, a coshφ sin θ, 0)

= (0, 0, 0).

Thus the catenoid is a minimal surface. In fact, it is the only surfaceof revolution which is minimal [We55, p. 179].

10.5. Introduction to theorema egregium; intrinsic vs

extrinsic

Some of the material in this chapter has already been covered inearlier chapters. We include such material here for the benefit of thereader primarily interested in understanding Gaussian curvature.

The present Chapter is an introduction to Gaussian curvature. Un-derstanding the intrinsic nature of Gaussian curvature, i.e. the theo-rema egregium of Gauss, clarifies the geometric classification of sur-faces. A beautiful historical account and an analysis of Gauss’s proofof the theorema egregium may be found in [Do79].3

We present a compact formula for Gaussian curvature and its proof,along the lines of the argument in M. do Carmo’s book [Ca76]. Theappeal of an old-fashioned, computational, coordinate notation proof isthat it obviates the need for higher order objects such as connections,tensors, exponental map, etc, and is, hopefully, directly accessible to astudent not yet familiar with the subject, cf. Remark 12.4.2.

Remark 10.5.1. We would like to distinguish two types of prop-erties of a surface in Euclidean space: intrinsic and extrinsic. Under-standing the extrinsic/intrinsic dichotomy is equivalent to understand-ing the theorema egregium of Gauss. The theorema egregium is the keyinsight lying at the foundation of differential geometry as conceived byB. Riemann in his essay [Ri1854] presented before the Royal ScientificSociety of Gottingen in 1854 (see Section 10.6).

3Our formula (10.9.2) for Gaussian curvature is similar to the traditional for-mula for the Riemann curvature tensor in terms of the Levi-Civita connection (notethe antisymmetrisation in both formulas, and the corresponding two summands),without the burden of the connection formalism.

Page 101: Analytic and Differential geometry Mikhail G. Katz∗

10.6. RIEMANN’S FORMULA 101

The theorema egregium asserts that an infinitesimal invariant of asurface in Euclidean space, called Gaussian curvature, is an “intrinsic”invariant of the surface M . In other words, Gaussian curvature of Mis independent of its isometric imbedding in Euclidean space. Thistheorem paves the way for an intrinsic definition of curvature in modernRiemannian geometry.

We will prove that Gaussian curvature K is an intrinsic invariantof a surface in Euclidean space, in the following precise sense: K canbe expressed in terms of the coefficients of the first fundamental form(see Section ) and their derivatives alone. A priori the possibility ofthus expressing K is not obvious, as the naive definition of K involvesthe coefficients of the second fundamental form (alternatively, of theWeingarten map).

10.6. Riemann’s formula

The intrinsic nature of Gaussian curvature paves the way for a tran-sition from classical differential geometry, to a more abstract approachof modern differential geometry (see also Subsection 12.11.2). The dis-tinction can be described roughly as follows. Classically, one studiessurfaces in Euclidean space. Here the first fundamental form (gij) ofthe surface is the restriction of the Euclidean inner product. Mean-while, abstractly, a surface comes equipped with a set of coefficients,which we deliberately denote by the same letters, (gij) in each coordi-nate patch, or equivalently, its element of length. One then proceedsto study its geometry without any reference to a Euclidean imbedding,cf. (17.4.2).

Such an approach was pioneered in higher dimensions in Riemann’sessay. The essay contains a single formula [Ri1854, p. 292]. See Spivak[Sp79, p. 149]), where the formula appears on page 159. This is theformula for the element of length of a surface of constant (Gaussian)curvature K ≡ α:

1

1 + α4

∑x2

√∑dx2 (10.6.1)

where today, of course, we would incorporate a summation index i aspart of the notation, as in

1

1 + α4

∑i(x

i)2

√∑

i

(dxi)2 (10.6.2)

cf. formulas (17.4.3), (17.17.1), and (18.3.1).

Page 102: Analytic and Differential geometry Mikhail G. Katz∗

102 10. PRINCIPAL CURVATURES, MINIMAL SURFACES, THM EGREGIUM

10.7. Preliminaries to the theorema egregium

The material in this section has mostly been covered already inearlier chapters.

Definition 10.7.1. Given a surface M in 3-space which is thegraph of a function of two variables, consider a critical point p ∈ M .The Gaussian curvature of the surface at the critical point p is thedeterminant of the Hessian of the function, i.e. the determinant of thetwo-by-two matrix of its second derivatives.

The implicit function theorem allows us to view any point of aregular surface, as such a critical point, after a suitable rotation. Wehave thus given the simplest possible definition of Gaussian curvatureat any point of a regular surface.

Remark 10.7.2. The appeal if this definition is that it allows oneimmediately to grasp the basic distinction between negative versuspositive curvature, in terms of the dichotomy “saddle point versuscup/cap”.

It will be more convenient to use an alternative definition in termsof the Weingarten map, which is readily shown to be equivalent toDefinition 10.7.1.

We denote the coordinates in R2 by (u1, u2). In R3, let 〈 , 〉 be theEuclidean inner product. Let x = x(u1, u2) : R2 → R3 be a regularparametrized surface. Here “regular” means that the vector valuedfunction x has differential dx of rank 2 at every point. Consider partialderivatives xi =

∂∂ui

(x), where i = 1, 2. Thus, vectors x1 and x2 form abasis for the tangent plane at every point. Similarly, let

xij =∂2x

∂ui∂uj∈ R3.

Let n = n(u1, u2) be a unit normal to the surface at the point x(u1, u2),so that 〈n, xi〉 = 0.

Definition 10.7.3 (symbols gij, Lij , Γkij). The first fundamental

form (gij) is given in coordinates by gij = 〈xi, xj〉. The second fun-damental form (Lij) is given in coordinates by Lij = 〈n, xij〉. Thesymbols Γkij are uniquely defined by the decomposition

xij = Γkijxk + Lijn

= Γ1ijx1 + Γ2

ijx2 + Lijn,(10.7.1)

where the repeated (upper and lower) index k implies summation, inaccordance with the Einstein summation convention, cf. Remark 9.9.1.

Page 103: Analytic and Differential geometry Mikhail G. Katz∗

10.8. AN IDENTITY INVOLVING THE Γkij AND THE Lij 103

Definition 10.7.4 (symbols Lij). The Weingarten map (Lij) is anendomorphism of the tangent plane Rx1 + Rx2. It is uniquely definedby the decomposition

nj = Lijxi

= L1jx1 + L2

jx2,

where nj =∂∂uj

(n).

Definition 10.7.5. We will denote by square brackets [ ] the anti-symmetrisation over the pair of indices found in between the brackets,e.g.

a[ij] =12(aij − aji).

Note that g[ij] = 0, L[ij] = 0, and Γk[ij] = 0. We will use the

notation Γkij;ℓ for the ℓ-th partial derivative of the symbol Γkij. Alsox[ij] = 0.

10.8. An identity involving the Γkij and the Lij

The following technical result will yield the theorema egregium asan easy consequence.

Proposition 10.8.1. We have the following relation

Γqi[j;k] + Γmi[j Γqk]m = −Li[jLqk]

for each set of indices i, j, k, q (with, as usual, an implied summationover the index m).

Proof. Consider the third partial derivative xijk =∂3x

∂ui∂uj∂uk. Let

us calculate its tangential component relative to the basis {x1, x2, n}.Recall that nk = Lpkxp and xjk = Γℓjkxℓ + Likn. Thus, we have

(xij)k =(Γmijxm + Lijn

)k

= Γmij;kxm + Γmijxmk + Lijnk + Lij;kn

= Γmij;kxm + Γmij

(Γpmkxp + Lmkn

)+ Lij (L

pkxp) + Lij;kn.

Grouping together the tangential terms, we obtain

(xij)k = Γmij;kxm + Γmij

(Γpmkxp

)+ Lij (L

pkxp) + (. . .)n

=(Γqij;k + ΓmijΓ

qmk + LijL

qk

)xq + (. . .)n

=(Γqij;k + ΓmijΓ

qkm + LijL

qk

)xq + (. . .)n,

since the symbols Γqkm are symmetric in the two subscripts.

Page 104: Analytic and Differential geometry Mikhail G. Katz∗

104 10. PRINCIPAL CURVATURES, MINIMAL SURFACES, THM EGREGIUM

Now the symmetry in j, k (equality of mixed partials) implies thefollowing identity: xi[jk] = 0. Therefore

0 = xi[jk]

= (xi[j)k]

=(Γqi[j;k] + Γmi[jΓ

qk]m + Li[jL

qk]

)xq + (. . .)n,

and therefore Γqi[j;k] + Γmi[jΓqk]m + Li[jL

qk] = 0 for each q = 1, 2. �

10.9. The theorema egregium of Gauss

Recall from Section 9.9 that we have K = − 2g11L1[1L

22], where by

definition

K = det(Lij) = 2L1[1L

22]. (10.9.1)

Theorem 10.9.1 (Theorema egregium). The Gaussian curvaturefunction K = K(u1, u2) of a surface can be expressed in terms of thecoefficients of the first fundamental form alone (and their first andsecond derivatives) as follows:

K =2

g11

(Γ21[1;2] + Γj1[1Γ

22]j

), (10.9.2)

where the symbols Γkij can be expressed in terms of the derivatives of gijbe the formula Γkij =

12(giℓ;j − gij;ℓ + gjℓ;i)g

ℓk, where (gij) is the inversematrix of (gij).

Proof of theorema egregium. We present a streamlined version ofdo Carmo’s proof [Ca76, p. 233]. The proof is in 3 steps.

(1) We express the third partial derivative xijk in terms of both theΓ’s (intrinsic information) and the L’s (extrinsic information).

(2) The equality of mixed partials yields an identification of a suit-able expression in terms of the Γ’s, with a certain combinationof the L’s.

(3) The combination of the L’s is expressed in terms of Gaussiancurvature.

The first two steps were carried out in Proposition 10.8.1. Wechoose the valus i = j = 1 and k = q = 2 for the indices. ApplyingTheorem 9.9.2(c), we obtain

Γ21[1;2] + Γm1[1Γ

22]m = −L1[1L

22]

= g1iLi[1L

22]

= g11L1[1L

22]

Page 105: Analytic and Differential geometry Mikhail G. Katz∗

10.9. THE THEOREMA EGREGIUM OF GAUSS 105

since the term L2[1L

22] = 0 vanishes. This yields the desired formula

for K and completes the proof of the theorema egregium. �

Page 106: Analytic and Differential geometry Mikhail G. Katz∗
Page 107: Analytic and Differential geometry Mikhail G. Katz∗

CHAPTER 11

Signed curvature of plane curves

11.1. Curvature with respect to an arbitrary parameter

The formula for the curvature of a curve α in Euclidean space is par-ticularly simple with respect to the arclength parameter s, viz. kα(s) =|α′′(s)|; see formula (3.10.1). For plane curves, the curvature can beexpressed in terms of an arbitrary parameter t of a regular parametri-sation, as follows.

Theorem 11.1.1. With respect to an arbirary regular parameter t,the formula for the curvature of a regular plane curve α(t) = (x(t), y(t))is

kα(t) =|xy − yx|

(x2 + y2)3/2=

|det [α α]|‖α‖3 . (11.1.1)

Here x = dx/dt, y = dy/dt, and α = dα/dt is the tangent vector tothe curve α(t) = (x(t), y(t)). Here we used the 2-by-2 matrix

[α α] ,

where α and α are written as column vectors, yielding the determinantdet [α α] = xy − yx. Thus, we can write

kα =|det [α α]|

‖α‖3 . (11.1.2)

For a proof, see formula (11.4.2).This approach motivates the introduction of a finer invariant called

signed curvature of parametrized plane curves, obtained by removingthe absolute value signs from formula (11.1.2).

Definition 11.1.2. For a parametrized regular plane curve α(t),

we define a signed curvature function kα(t) by setting

kα =det [α α]

‖α‖3 .

This is discussed in more detail in Section 11.7.

Remark 11.1.3. In Section 11.3, curvature will be expressed interms of the angle θ formed by the tangent vector of the curve withthe positive direction of the x-axis.

107

Page 108: Analytic and Differential geometry Mikhail G. Katz∗

108 11. SIGNED CURVATURE OF PLANE CURVES

Example 11.1.4. Calculate the curvature of the graph of y = f(x)at a critical point x0 using formula (11.1.1).

11.2. Jordan curves in the plane

Definition 11.2.1. A Jordan curve in the plane is a non-selfintersectingclosed curve.

A Jordan curve can be represented by a continuous one-to-one map

α : S1 → R2.

Theorem 11.2.2 (Jordan curve theorem). A Jordan curve sepa-rates the plane into two open regions: a bounded region and an un-bounded region.

The bounded region is called the interior region.We will only deal with smooth (i.e., infinitely differentiable) regular

(i.e, α 6= 0) maps α.

11.3. Gauss map

We identify R2 with C so that a vector in the plane can be writtenas a complex number. Consider a smooth Jordan curve, denoted C, inthe plane identified with C:

C ⊆ C.

By Theorem 5.1.1 there is an arclength parametrisation α(s) of C. Let

v(s) = α′(s)

be its velocity vector at the point α(s) ∈ C. The vector v(s) can bethought of as a point of the unit circle S1 ⊆ C (by translating theinitial point of the vector v(s) to the origin):

v(s) ∈ S1

Remark 11.3.1. The Gauss map is usually defined using the nor-mal vector n(s) (see Remark 11.6.5). For plane curves the tangentvector v(s) and the normal vector differ by a 90◦ rotation in the plane:n(s) = iv(s) for an arclength parameter s so it could be defined interms of v as well.

The normal vector n(s) = iv(s) defines a map to the unit circlecalled the Gauss map, as follows.

Definition 11.3.2. The Gauss map is the map

n : C → S1, α(s) 7→ n(s) = iv(s) (11.3.1)

where α(s) is an arbitrary point of the curve C and v(s) = α′(s).

Page 109: Analytic and Differential geometry Mikhail G. Katz∗

11.4. CURVATURE EXPRESSED IN TERMS OF θ(s) 109

11.4. Curvature expressed in terms of θ(s)

Let s be an arclength parameter along a curve α(s) with tangentvector v(s) = α′(s).

Definition 11.4.1 (function theta). The angle θ(s) is defined inone of the following two equivalent ways:

(1) We write v(s) = eiθ(s), where the angle θ(s) is measured coun-terclockwise, from the positive ray of the x-axis, to the vec-tor v(s).

(2) Using the complex logarithm, we can also express θ(s) as fol-lows: θ(s) = 1

ilog v(s) = −i log v(s).

Theorem 11.4.2. We have the relation ddθeiθ = ieiθ.

This was proved in complex functions.

Theorem 11.4.3. We have the following expression for the curva-ture kα of the curve α:

kα(s) =

∣∣∣∣dθ

ds

∣∣∣∣ . (11.4.1)

Proof. Recall that v(s) = eiθ(s). By chain rule, we have

dv

ds= ieiθ(s)

ds,

and therefore the curvature kα(s) satisfies

kα(s) = ‖α′′(s)‖ =

∥∥∥∥dv

ds

∥∥∥∥ =

∣∣∣∣dθ

ds

∣∣∣∣ ,

since |i| = 1 and∣∣eiθ(s)

∣∣ = 1. �

Corollary 11.4.4. Let α(t) = (x(t), y(t)) be an arbitrary regularparametrisation (not necessarily arclength) of a plane curve. We havethe following expression for the curvature kα:

kα =|xy − yx|‖α‖3 (11.4.2)

Proof. With respect to an arbitrary parameter t, the componentsof the tangent vector v(t) = α(t) are x(t) and y(t). Hence we have

tan θ =y

x(11.4.3)

so that, differentiating (11.4.3) with respect to t, we obtain

1

cos2 θ

dt=xy − xy

x2. (11.4.4)

Page 110: Analytic and Differential geometry Mikhail G. Katz∗

110 11. SIGNED CURVATURE OF PLANE CURVES

Meanwhile by chain rule,

x =dx

dt=dx

ds

ds

dt= cos θ

ds

dtand

dt=dθ

ds

ds

dt. (11.4.5)

Furthermore, dsdt

= |v(t)| = |α(t)| = |dαdt|. Solving (11.4.5) for dθ

ds, we

obtain from (11.4.4) that ∣∣∣∣dθ

ds

∣∣∣∣ =|xy − yx|

|α|3and by (11.4.1) we obtain (11.4.2). �

11.5. Convex Jordan curves

Recall that the set-theoretic complement of a subset B ⊆ A isdenoted A \B.

Recall that any line ℓ ⊆ R2 divides the plane into two open half-planes, namely the two connected components1 of the set comple-ment R2 \ ℓ.

Definition 11.5.1. A smooth Jordan curve C is called strictlyconvex 2 if one of the following two equivalent conditions is satisfied:

(1) the interior of each segment joining a pair of points of C iscontained in the interior region of C;

(2) consider the tangent line Tp to C at a point p ∈ C; then thecomplement C \ {p} lies entirely in one of the open halfplanesof R2 \ Tp defined by the tangent line, for all p ∈ C.

Theorem 11.5.2. Assume that a smooth regular Jordan curve Cis strictly convex, and parametrized counterclockwise. Then the Gaussmap n : C → S1 of (11.3.1) is one-to-one and onto, and the func-tion θ(s) is monotone increasing.

Proof of the “onto” part. We will work instead with the mapv : C → S1 since it only differs from n by a 90 degree rotation. Letα(s) be an arclength parametrisation of C. First we consider the caseof “horizontal” vectors v(s) in the (x, y)-plane. These occur at pointson C with maximal and minimal imaginary part, i.e., the y-coordinate.

We think of the y corodinate as defining a “height” function on thecurve. By applying Rolle’s theorem to the “height function” y(s), weobtain a point of minimum smin and a point of maximum smax.

1rechivei kshirut2kamur with kuf

Page 111: Analytic and Differential geometry Mikhail G. Katz∗

11.6. TOTAL CURVATURE OF A CONVEX JORDAN CURVE 111

The points α(smin) and α(smax) have “opposite” horizontal tangentvectors. Such points correspond to the values θ = 0 and θ = π.

To treat the general case, we will now show how one obtains thepair of “opposite” tangent vectors

v = eiθ and − v = ei(θ+π).

First consider the vector wθ = ei(θ+π2) orthogonal to v. We then seek

the extrema of the function f given by the scalar product

f(s) = 〈α(s), wθ〉where α(s) is a parametrisation of the curve C. This is thought of asthe replacement of the “height function” above. At an extremum s0 ofthe function f , we have

d

ds(f(s))|s=s0 =

⟨dα

ds, wθ

⟩= 0.

Hence the tangent vector v(s0) = α′(s0) at each extremum of f isparallel to v. As v ranges over S1, we thus obtain the points on thecurve where the tangent vector is parallel to v. �

Proof of the “one-to-one” part. We would like to show thatthe Gauss map (11.3.1) is one-to-one. We proceed by contradiction(relying on the law of excluded middle). Suppose on the contrarythat two distinct points p ∈ C and q ∈ C have identical tangent vec-tors v(s) = eiθ. Then the tangent lines Tp and Tq to C at p and q areparallel. By definition of convexity, the curve C lies on the same side ofeach of the tangent lines Tp and Tq. Hence the tangent lines must coin-cide: Tp = Tq. Thus, both p and q must lie on the common line Tp = Tq.This implies that the arc of the curve between p and q is contained in C.This contradicts the hypothesis that the curve is strictly convex. Thecontradiction proves that the map is one-to-one. �

Remark 11.5.3. A counterclockwise parametrisation correspondsto θ(s) increasing, while a clockwise parametrisation corresponds to θ(s)decreasing.

11.6. Total curvature of a convex Jordan curve

The result on the total curvature3 of a Jordan curve is of interest inits own right. Furthermore, the result serves to motivate an analogousstatement of the Gauss–Bonnet theorem for surfaces in Section 12.9.

3Akmumiyut kolelet

Page 112: Analytic and Differential geometry Mikhail G. Katz∗

112 11. SIGNED CURVATURE OF PLANE CURVES

Definition 11.6.1. The total curvature Tot(C) of a curve C witharclenth parametrisation α(s) : [a, b] → R2 is the integral

Tot(C) =

∫ b

a

kα(s)ds. (11.6.1)

Example 11.6.2. Let us show that the total curvature of a circle Cr(of radius r) is 2π and therefore indepedent of r. Indeed, consider aparametrisation α(s) : [0, 2πr] → R2 of the circle given by the usualtrigonometric functions. Then

Tot(Cr) =

∫ 2πr

0

kα(s)ds =

∫ 2πr

0

1

rds = 2π.

Definition 11.6.3. If C is a closed curve, i.e, α(a) = α(b), we willwrite the integral (11.6.1) using the notation of a contour integral

Tot(C) =

C

kα(s)ds. (11.6.2)

Theorem 11.6.4. The total curvature of each strictly convex Jor-dan curve C with arclength parametrisation α(s) equals 2π, i.e.,

Tot(C) =

C

kα(s)ds = 2π.

Proof. Let α(s) be a unit speed parametrisation so that α(0) bethe lowest point of the curve, and assume the curve is parametrizedcounterclockwise. Let v(s) = α′(s). Then v(0) = ei0 = 1 and θ(0) = 0.The function θ = θ(s) is monotone increasing from 0 to 2π by The-orem 11.5.2. Applying the change of variable formula for integration,we obtain

Tot(C) =

C

kα(s)ds =

C

∣∣∣∣dv

ds

∣∣∣∣ ds =∮

C

dsds =

∫ 2π

0

dθ = 2π,

proving the theorem. �

Remark 11.6.5. We can also consider the normal vector n(s) tothe curve. The normal vector satisfies n(s) = ei(θ+

π2) = iei(θ) and |dn

ds| =

|dvds| = dθ

ds. Therefore we can also calculate the total curvature as follows:

C

kα(s)ds =

C

∣∣∣∣dn

ds

∣∣∣∣ ds =∮

C

dsds =

∫ 2π

0

dθ = 2π,

with n(s) in place of v(s). For surfaces the Gauss map is defined bymeans of the normal vector.

Page 113: Analytic and Differential geometry Mikhail G. Katz∗

11.7. SIGNED CURVATURE kα OF JORDAN CURVES IN THE PLANE 113

Example 11.6.6. The ellipse being a closed convex curve, we ob-tain that 2π is the total curvature of an arbitrary ellipse defined by aquadratic equation ax2+2bxy+cy2+dx+ey = f , ac−b2 > 0, providedthe curve is nondegenerate; see Definition 2.7.4 for details.

Example 11.6.7. The hyperbola H defined by λ1x2 + λ2y

2 = k isnot a closed curve but in the case λ1 = −λ2 (see Definition 2.7.2) theasymptotes of H are orthogonal and the image of θ is precisely half thecircle. Therefore a similar integration argument shows that the totalcurvature is π in this case: Tot(H) = π.

Remark 11.6.8. The theorem in fact holds for an arbitrary regularJordan curve (even though in general θ(s) will not be a monotonefunction) provided we adjust the definition of curvature to allow fornegative sign. We have dealt only with the convex case in order tosimplify the topological considerations. See further in Section 11.7.

Remark 11.6.9. A similar calculation will yield the Gauss–Bonnettheorem for convex surfaces in Section 12.9.

11.7. Signed curvature kα of Jordan curves in the plane

Let α(s) be an arclength parametrisation of a Jordan curve in theplane. We assume that the curve is parametrized counterclockwise. Asin Section 11.3, a continuous branch of θ(s) can be chosen where θ(s)is the angle measured counterclockwise from the positive x-axis to thetangent vector v(s) = α′(s).

Note that if the Jordan curve is not convex, the function θ(s) willnot be monotone and at certain points its derivative may take negativevalues:

θ′(s) < 0.

Once we have such a continuous branch, we can define the signed cur-vature kα as follows.

Definition 11.7.1. The signed curvature kα(s) of a plane Jordancurve α(s) oriented counterclockwise is defined by setting

kα(s) =dθ

ds

where θ = 1ilog v in other words v(s) = α′(s) = eiθ(s).

We have the following generalisation of Theorem 11.6.4 on the totalcurvature.

Page 114: Analytic and Differential geometry Mikhail G. Katz∗

114 11. SIGNED CURVATURE OF PLANE CURVES

Definition 11.7.2. The total signed curvature of a smooth Jordancurve C with arclength parametrisation α(s), s ∈ [a, b] is defined to be

Tot(C) =

∫ b

a

kα(s)ds.

Theorem 11.7.3. The total signed curvature of a smooth Jordancurve C with counterclockwise arclength parametrisation α(s) is 2π:

Tot(C) =

∫ b

a

kα(s)ds = 2π.

Proof. We exploit the continuous branch θ(s) as in Theorem 11.6.4,but without the absolute value signs on the derivative of θ(s). Applyingthe change of variable formula for integration, we obtain

Tot(C) =

C

kα(s)ds =

C

dsds =

∫ 2π

0

dθ = 2π,

proving the theorem modulo Theorem 12.1.4 below. �

Page 115: Analytic and Differential geometry Mikhail G. Katz∗

CHAPTER 12

Laplacian formula K, Hyperbolic metric,Gauss–Bonnet

12.1. Gauss map for nonconvex curves

Remark 12.1.1. For nonconvex Jordan curves, the Gauss map tothe circle will not be one-to-one, but will still have an algebraic degreeone. This phenomenon has an analogue for imbedded surfaces wherethe algebraic degree is proportional to the Euler characteristic.

Definition 12.1.2. Let α(s) be a parametrisation of a smoothclosed curve C. Consider an arbitrary smooth map (not necessarilythe Gauss map) α(s) 7→ eiθ(s) from C to the circle. The algebraicdegree of the map at a point z ∈ S1 is defined to be the sum

θ−1(z)

sign

(dθ

ds

),

where the summation is over all points in the inverse image of z.

Here by Sard’s theorem z can be chosen in such a way that theinverse image is finite so that the sum is well-defined.

Example 12.1.3. The map defined by z 7→ zn restricted to the unitcircle gives a map eiθ 7→ einθ of algebraic degree n. The preimage of 1 =

ei0 consists precisely of the nth roots of unity e2πikn , k = 0, 1, 2, . . . , n−1.

Altogether there are n of them and therefore the algebraic degree is n.

Theorem 12.1.4. For Jordan curves (i.e., imbedded curves), the al-gebraic degree of the Gauss map is 1 if oriented counterclockwise and −1if orientated counterclockwise.

This is the topological ingredient behind the theorem on the totalcurvature of a Jordan curve.

12.1.1. Rotation index of a closed curve in the plane. This sec-tion is optional. Given a regular closed plane curve of length L and arclengthparametrisation α(s) with α′(s) = eiθ(s), a continuous branch of θ(s) can bechosen even if the curve is not simple, obtaining a map θ : [0, L] → R. Asin Section 11.7, we can assume that θ(0) = 0. Then θ(L) is necessarily aninteger multiple of 2π. See Millman & Parker [MP77, p. 55].

115

Page 116: Analytic and Differential geometry Mikhail G. Katz∗

11612. LAPLACIAN FORMULA K, HYPERBOLIC METRIC, GAUSS–BONNET

Figure 12.1.1. Three points on a curve with thesame tangent direction, but the middle one correspondsto θ′(s) < 0.

Definition 12.1.5. The rotation index ια of a closed unit speed plane

curve α(s) is the integer ια = θ(L)−θ(0)2π .

Example 12.1.6. The rotation index of a figure-8 curve is 0.

Theorem 12.1.7. Let α(s) be an arclength parametrisation of a geomet-ric curve C. Then the rotation index is related to the total signed curva-

ture Tot(C) as follows:

ια =Tot(C)

2π. (12.1.1)

The proof is the same as in Section 11.7.

Remark 12.1.8. An analogous relation between the Euler characteristicof a surface and its total Gaussian curvature appears in Section 12.11.

Example 12.1.9. The figure-8 has rotation index 0. Describe an im-mersed curve with rotation index 2.

12.1.2. Connected components of curves. This section is optional.Until now we have only considered connected curves. A curve may in generalhave several connected components.1

1Rechivei k’shirut.

Page 117: Analytic and Differential geometry Mikhail G. Katz∗

12.3. THE LAPLACE-BELTRAMI FORMULA FOR K 117

Definition 12.1.10. Two points p, q on a curve C ⊆ R2 are said to lie inthe same connected component if there exists a continuous map h : [0, 1] → Csuch that h(0) = p and h(1) = q.

This defines an equivalence relation on the curve C, and decomposes itas a disjoint union C =

⊔iCi.

Definition 12.1.11. The set of connected components of C is denoted π0(C).The number of connected components is denoted |π0(C)|.

Example 12.1.12. Let F (x, y) = (x2 + y2 − 1)((x− 10)2 + y2 − 1), andlet CF be the curve defined by F (x, y) = 0. Then CF is the union of a pair ofdisjoint circles. Therefore it has two connected components: |π0(CF )| = 2.

The total curvature can be similarly defined for a non-connected curve,by summing the integrals over each connected components.

Definition 12.1.13. The total curvature of a curve C =⊔iCi is Tot(C) =∑

iTot(Ci).

We have the following generalisation of the theorem on total curvatureof a curve.

Theorem 12.1.14. If each connected component of a curve C is a Jordancurve parametrized counterclockwise, then the signed total curvature of C is

Tot(C) = 2π |π0(C)|.

Proof. We apply the previous theorem to each connected component,and sum the resulting total curvatures. �

12.2. Mean versus Gaussian curvature

Theorem 12.2.1. Unlike Gaussian curvature K, the mean curva-ture H = 1

2Lii cannot be expressed in terms of the gij and their deriva-

tives.

Proof. The plane and the cylinder have parametrisations withidentical (gij), but with different mean curvature, cf. Table 12.2.1. Tosummarize, Gaussian curvature is an intrinsic invariant, while meancurvature an extrinsic invariant, of the surface. �

12.3. The Laplace-Beltrami formula for K

A (u1, u2) chart in which the metric becomes conformal (see Defi-nition 17.7.1) to the standard flat metric, is referred to as isothermalcoordinates . The existence of the latter is proved in [Bes87].

Page 118: Analytic and Differential geometry Mikhail G. Katz∗

11812. LAPLACIAN FORMULA K, HYPERBOLIC METRIC, GAUSS–BONNET

Weingartenmap (Lij)

Gaussian cur-vature K

mean cur-vature H

plane

(0 00 0

)0 0

cylinder

(1 00 0

)0 1

2

invarianceof curvature

yes no

Table 12.2.1. Plane and cylinder have the same intrinsicgeometry (K), but different extrinsic geometries (H)

Definition 12.3.1. The Laplace-Beltrami operator for a metric λδijin isothermal coordinates is

∆LB =1

λ

(∂2

∂(u1)2+

∂2

∂(u2)2

).

The notation means that when we apply the operator ∆LB to afunction h = h(u1, u2), we obtain

∆LB(h) =1

λ

(∂2h

∂(u1)2+

∂2h

∂(u2)2

).

In more readable form for h = h(x, y), we have

∆LB(h) =1

λ

(∂2h

∂x2+∂2h

∂y2

).

Theorem 12.3.2. Given a metric in isothermal coordinates withmetric coefficients gij = λ(u1, u2)δij, its Gaussian curvature is minushalf the Laplace-Beltrami operator applied to the ln of the conformalfactor λ:

K = −1

2∆LB lnλ. (12.3.1)

Here ln x is the natural logarithm so that (ln x)′ = 1x.

Proof. Let λ = e2µ. We have from Table 12.3.1:

2Γ21[1;2] = Γ2

11;2 − Γ212;1 = −µ22 − µ11.

Page 119: Analytic and Differential geometry Mikhail G. Katz∗

12.4. HYPERBOLIC METRIC IN UPPERHALF PLANE 119

Γ1ij j = 1 j = 2

i = 1 µ1 µ2

i = 2 µ2 −µ1

Γ2ij j = 1 j = 2

i = 1 −µ2 µ1

i = 2 µ1 µ2

Table 12.3.1. Symbols Γkij of a metric e2µ(u1,u2)δij

It remains to verify that the ΓΓ term in the expression (10.9.2) for theGaussian curvature vanishes:

2Γj1[1Γ22]j = 2Γ1

1[1Γ22]1 + 2Γ2

1[1Γ22]2

= Γ111Γ

221 − Γ1

12Γ211 + Γ2

11Γ222 − Γ2

12Γ212

= µ1µ1 − µ2(−µ2) + (−µ2)µ2 − µ1µ1

= 0.

Then from formula (10.9.2) we have

K =2

λΓ21[1;2] = −1

λ(µ11 + µ22) = −∆LBµ.

Meanwhile, ∆LB log λ = ∆LB(2µ) = 2∆LB(µ), proving the result. �

Setting λ = f 2, we restate the theorem as follows.

Corollary 12.3.3. Given a metric in isothermal coordinates withmetric coefficients gij = f 2(u1, u2)δij, its Gaussian curvature is minusthe Laplace-Beltrami operator of the log of the conformal factor f :

K = −∆LB log f. (12.3.2)

12.4. Hyperbolic metric in upperhalf plane

This example is also discussed in Section 17.17. Let x = u1 andy = u2.

Theorem 12.4.1. The metric with coefficients

gij =1

y2δij (12.4.1)

in the upperhalf plane

H2 = {(x, y) | y > 0} ,called the hyperbolic metric of the upper half plane, has constant Gauss-ian curvature K = −1.

Page 120: Analytic and Differential geometry Mikhail G. Katz∗

12012. LAPLACIAN FORMULA K, HYPERBOLIC METRIC, GAUSS–BONNET

Proof. By Theorem 12.3.3, we have

K = −∆LB log f = ∆LB log y = y2(− 1

y2

)= −1,

as required. �

Either one of the formulas (10.9.2), (12.3.1), or (12.3.2) can serveas the intrinsic definition of Gaussian curvature, replacing the extrinsicdefinition (10.9.1), cf. Remark 10.6.

Remark 12.4.2. For a reader familiar with elements of Riemanniangeometry, it is worth mentioning that the Jacobi equation y′′+Ky = 0of a Jacobi field y on M (expressing an infinitesimal variation bygeodesics) sheds light on the nature of curvature in a way that nomere formula for K could. Thus, in positive curvature, geodesics con-verge, while in negative curvature, they diverge. However, to provethe Jacobi equation, one would need to have already an intrinsicallywell-defined quantity on the left hand side, y′′ + Ky, of the Jacobiequation. In particular, one would need an already intrinsic notion ofcurvature K. Thus, a proof of the theorema egregium necessarily pre-cedes the deeper insights provided by the Jacobi equation. Similarly,the Gaussian curvature at p ∈ M is the first significant term in theasymptotic expansion of the length of a “small” circle of center p. Thisfact, too, sheds much light on the nature of Gaussian curvature. How-ever, to define what one means by a “small” circle, requires introducinghigher order notions such as the exponential map, which are usuallyunderstood at a later stage than the notion of Gaussian curvature,cf. [Ca76, Car92, Ch93, GaHL04].

12.5. Gaussian curvature as product of signed curvatures

We originally defined the Gaussian curvature K = K(u1, u2) as thedeterminant of the Weingarten map Wp at p = x(u1, u2) ∈ M . We alsosaw that the absolute value |K| can be represented as the product of thecurvatures kβ1 and kβ2 of geodesics in the direction of the eigenvectors viof Wp, namely |K| = kβ1kβ2. In this spirit, some textbooks actuallydefine the Gaussian curvature not as the determinant of the Weingartenmap but rather as the product of the signed curvatures kα of planecurves obtained by intersecting M with the planes spanned by n andeach of the eigenvectors vi, as already mentioned in Definition 9.6.1,item 2.

Theorem 12.5.1. The Gaussian curvature Kp of M ⊆ R3 at psatisfies

Kp = kβ1kβ2

Page 121: Analytic and Differential geometry Mikhail G. Katz∗

12.5. GAUSSIAN CURVATURE AS PRODUCT OF SIGNED CURVATURES 121

where kβi =dθids

is the signed curvature of the plane curve βi =M ∩ Piwhere the plane Pi = Span(vi, n) is spanned by eigenvector vi of Wp

and n, relative to the basis (vi, n) for Pi.

Remark 12.5.2. To explain this in terms of the Hessian, we can as-sume without loss of generality thatM is the graph of a function f(x, y)vanishing at the origin, with a critical point at the origin, and suchthat the eigenspaces of the Hessian of f are precisely the x-axis andthe y-axis. Then the Gaussian curvature at the origin is the product ofsigned curvatures of the curves obtained by restricting f respectivelyto the x-axis and the y-axis.

If we adjust the coordinates by an isometry in such a way that the xand y axes become the eigenspaces, while the z-axis is the direction ofthe normal, then Hf becomes a diagonal martix Hf = diag(λ1, λ2).The curve β1 in the (x, z) plane is then described by z = λ1x

2 + o(x2)while the curve β2 in the (y, z) plane is described implicitly by z =λ2y

2 + o(y2). With respect to the natural orientations in these planes,the curves have signed curvature respectively λ1 and λ2 at the origin,so that K = λ1λ2 as required.

12.5.1. Binet–Cauchy identity. This section is optional.

Theorem 12.5.3 (Binet–Cauchy identity). The 3-dimensional caseof the Binet–Cauchy identity is the identity (a · c)(b · d) = (a · d)(b · c)+(a× b) · (c× d), where a, b, c, and d are vectors in R3.

The formula can also be written as a formula giving the dot productof two wedge products, namely (a×b)·(c×d) = (a·c)(b·d)−(a·d)(b·c).

Corollary 12.5.4 (Special case of Binet–Cauchy). In the specialcase of vectors a = c and b = d, we obtain |a× b|2 = |a|2|b|2 − |a · b|2.

Example 12.5.5. When both vectors a, b are unit vectors, we ob-tain the usual relation 1 = cos2(φ) + sin2(φ) (here the vector productgives the sine and the scalar product gives the cosine) where φ is theangle between the vectors.

Corollary 12.5.6. Let a = x1 and b = x2 be the two tangentvectors to the surface with parametrisation x(u1, u2). Then |x1×x2|2 =g11g22 − g212 = det(gij).

Equivalently, we have |x1 × x2| =√

det(gij). This can be thoughtof as the area of the parallelogram spanned by the two vectors.

Page 122: Analytic and Differential geometry Mikhail G. Katz∗

12212. LAPLACIAN FORMULA K, HYPERBOLIC METRIC, GAUSS–BONNET

12.6. Area elements of the surface and of the sphere

Consider a surface M imbedded in 3-space. We have√

det(gij) =∣∣x1 × x2

∣∣ ,

where xi =∂x∂ui

(cf. the Binet–Cauchy identity 12.5.3).Generalizing the case of the area element rdrdθ in the plane in polar

coordinates as in Section 8.3, as well as the spherical element of areasinϕdθdϕ, we obtain the following definition.

Definition 12.6.1. The area element dAM of the surface M is

dAM =√

det(gij)du1du2 = |x1 × x2|du1du2 (12.6.1)

where the gij are the metric coefficients of the surface with respect tothe parametrisation x(u1, u2).

This notion of area is discussed in more detail in Section 17.6.We have used the subscript M so as to specify which surface we are

dealing with.

Definition 12.6.2. A surface M is orientable if it admits a con-tinuous unit normal vector field N = Np, defined at each point p ∈ M .

The vector field N is a globally defined field along an orientablesurface. Note that the image vector N can be thought of as an elementof the unit sphere: Np ∈ S2 ⊆ R3.

12.7. Gauss map for surfaces in R3

Definition 12.7.1. The Gauss map of an orientable surface M ,denoted

G : M → S2,

is the map sending each point p = x(u1, u2) to the normal vector G(p) =n(u1, u2).

Remark 12.7.2. Each point p ∈ M lies in a neighborhood witha suitable parametrisation x(u1, u2). At a point p = x(u1, u2), wehave a normal vector n(u1, u2), obtained by normalizing the vectorproduct x1 × x2, which coincides with the globally selected normal N ,i.e., n(u1, u2) = Nx(u1,u2).

Lemma 12.7.3. If K 6= 0 at p ∈ M then the tangent vectorsWp(x1),Wp(x2) ∈ Tp are linearly independent.

Page 123: Analytic and Differential geometry Mikhail G. Katz∗

12.8. SPHERE S2 PARAMETRIZED BY GAUSS MAP OF M 123

Proof. The coefficients of the matrix (Li j) the coordinates of∂n∂u1, ∂n∂u2

with respect to the basis (x1, x2). Hence independence of thevectors n1 and n2 is equivalent to the condition det(Li j) 6= 0. �

Theorem 12.7.4. Choose a specific point p = x(u10, u20) ∈M of the

surface. If the Gaussian curvature is nonzero at p, then the map n(u1, u2)from (a neighborhood of) M to S2 produces a regular parametrisationof a spherical neighborhood of the point

G(p) = n(u10, u20) ∈ S2

on the sphere.

Proof. The parametrisation is given by the map G to the sphere.We have ∂

∂ui(G(p)) = ∂

∂ui(n(u1, u2)) = ∂n

∂ui=Wp(xi). By Lemma 12.7.3,

the two partials of G are linearly independent. Hence the parametri-sation defined by G is regular. �

12.8. Sphere S2 parametrized by Gauss map of M

Remark 12.8.1. We are no longer thinking of n(u1, u2) as a normalto the original surface M , but rather as a parametrisation of an openneighborhood on the unit sphere. Now consider the area element ofthe unit sphere.

Definition 12.8.2. Assume K 6= 0 at a point of M . Let (gαβ) bethe metric coefficients of the regular parametrisation n(u1, u2) of the

sphere, where nα = ∂n(u1,u2)∂uα

for each α = 1, 2.

Theorem 12.8.3. Consider a parametrisation n(u1, u2) of a neigh-borhood on the sphere S2 as in Theorem 12.7.4. Then the area ele-ment dAS2 can be expressed as

dAS2 =√

det(gαβ)du1du2 = |n1 × n2|du1du2.

Proof. This follows from the usual formula for the element ofarea for surfaces, applied to the chosen parametrisation n(u1, u2) asdefined in Theorem 12.7.4, in place of the traditional parametrisationin spherical coordinates (cf. Binet–Cauchy in Section 12.5.1). �

Remark 12.8.4. We have used the subscript to distinguish thearea element dAS2 of the sphere S2 from the area element dAM of thesurfaceM as in (12.6.1). Note that we need to distinguish the two areaelements

dAS2 and dAMbecause both will occur in the proof of the Gauss–Bonnet theorem.

Page 124: Analytic and Differential geometry Mikhail G. Katz∗

12412. LAPLACIAN FORMULA K, HYPERBOLIC METRIC, GAUSS–BONNET

12.9. Comparison of two parametrisations

When the Gaussian curvature is nonzero, the normal vector of Mallows us to define local coordinates n(u1, u2) on the unit sphere S2 asin Section 12.7.

Lemma 12.9.1. Consider the metric coefficients (gij) of the paramet-risation x(u1, u2) of the surface M and the metric coefficients (gαβ) ofthe sphere S2 relative to the parametrisation n(u1, u2). Then we havethe identity

det(gαβ) =(K(u1, u2)

)2det(gij)

where K = K(u1, u2) is the Gaussian curvature of the surface M .

Proof. Let L = (Lij) be the matrix of the Weingarten map W =

Wp with respect to the basis (x1, x2). Here p = x(u1, u2). By definitionof curvature we have K = det(L). Recall that the coefficients Lij ofthe Weingarten map are defined by

nα = Liαxi = xiLiα. (12.9.1)

Consider the 3 × 2-matrices A = [x1 x2] and B = [n1 n2]. Thenformula (12.9.1) implies by definition of matrix multiplication that

B = AL.

Therefore the Gram matrices satisfy

Gram(n1, n2) = BtB = (AL)tAL = LtAtAL = Lt Gram(x1, x2) L.

Applying the determinant, we complete the proof of the lemma. �

Remark 12.9.2. By the Cauchy-Binet formula, the desired iden-tity is equivalent to the formula |nu1 × nu2 | = |det(Lij)| |xu1 × xu2 |,immediate from the observation that a linear map multiplies the areaof parallelograms by its determinant. Namely, the Weingarten mapsends each vector xi to ni: Wp(xi) = ni ∈ Tp. A stronger iden-tity is true that is sensitive to the sign of the Gaussian curvature:W (u)×W (v) = det(W )(u× v) where (u, v) is any basis of the tangentplane; e.g., the basis (x1, x2).

12.10. Gauss–Bonnet theorem

The Gauss–Bonnet theorem for surfaces is to a certain extent anal-ogous to the theorem on the total curvature of a plane curve (Theo-rem 11.6.4). In both cases, an integral of curvature turns out to havetopological significance.

Page 125: Analytic and Differential geometry Mikhail G. Katz∗

12.10. GAUSS–BONNET THEOREM 125

Theorem 12.10.1 (Special case of Gauss–Bonnet). Let M be anorientable closed surface in R3 of positive Gaussian curvature (a typicalexample is an ellipsoid). Then the curvature integral satisfies

M

Kp dAM = 4π = 2πχ(S2),

where in each coordinate patch we have Kp = K(u1, u2) is the Gaussiancurvature function defined at every point p = x(u1, u2) of the surface,and χ(S2) = 2.

We will use the existence of global coordinates (with singularitiesonly at the north and south poles) on the sphere to write down aconcrete version of the calculation.

We have the Gauss map G : M → S2 defined by the normal vector.Let (θ, ϕ) be the usual spherical coordinates on the unit sphere, thoughtof as functions on S2. Thus we have a usual parametrisation

σ = σ(θ, φ) : [0, 2π)× (0, π) → S2

(omitting the problematic poles) as in Section 5.5. Consider the inversemap

σ−1 : S2 → [0, 2π)× (0, π)

defined away from the poles, so that for a point s ∈ S2 we have

(θ(s), ϕ(s)) = σ−1(s).

Definition 12.10.2. We choose global coordinates (u1, u2) on Mby setting

u1(p) = θ(G(p)), u2 = ϕ(G(p))

onM by precomposing the spherical coordinates by the Gauss map G.

Here we can dispense with local coordinate charts and compute theGauss–Bonnet integral in these global coordinates. The element of areacan therefore be written globally as

dAM =√

det(gij)du1du2

where (gij) is the metric of M with respect to these coodinates.

Page 126: Analytic and Differential geometry Mikhail G. Katz∗

12612. LAPLACIAN FORMULA K, HYPERBOLIC METRIC, GAUSS–BONNET

Proof of Gauss–Bonnet Theorem. We will exploit the rela-tion K(u1, u2)

√det(gij) =

√det(gαβ) from Section 12.9. We obtain

M

K(u1, u2) dAM =

∫∫

M

K(u1, u2)√det(gij)du

1du2

=

∫∫

S2

√det(gαβ)dθdϕ

=

∫∫

S2

dAS2

= area(S2) = 4π,

as required. �

Note that each of the double integrals∫∫

can be replaced by theiterated integral

∫ π

0

(∫ 2π

0

dθ(. . .)

),

where the appropriate integrand is placed where the dots (. . .) are.

12.11. Euler characteristic

The Euler characteristic of a closed imbedded surface in Euclidean3-space can be defined via the integral of the Gaussian curvature. Itcan be thought of as a generalisation of the rotation index of a planeclosed curve.

Here we no longer assume that the curvature is positive. The fol-lowing is either a definition or a theorem, depending where you start.

The Euler characteristic χ(M) of a surface M is defined by therelation

2πχ(M) =

M

KpdAM , (12.11.1)

where in a coordinate patch Kp = K(u1, u2) is the Gaussian curvatureat the point p = x(u1, u2) ∈M .

Remark 12.11.1. The relation (12.11.1) is similar to the line inte-gral expression for the rotation index in formula (12.1.1).

To show that this definition of the Euler characteristic agrees withthe usual one, it is necessary to use the notion of algebraic degree formaps between surfaces, similar to the algebraic degree of a self-map ofa circle.

Page 127: Analytic and Differential geometry Mikhail G. Katz∗

12.11. EULER CHARACTERISTIC 127

12.11.1. Alternative proof. This material is to be skipped. Herewe present another proof of Gauss–Bonnet theorem.

Proof. The convexity of the surface guarantees that the map nis one-to-one (compare with the proof of Theorem 11.5.2 on closedcurves).

We examine the integrand K dAM in a coordinate chart (u1, u2)where it can be written as K(u1, u2) dAM . By Lemma 12.9.1, we have

K(u1, u2) dAM = K(u1, u2)√det(gij)du

1du2 =√

det(gαβ)du1du2 = dAS2.

Thus, the expression K dAM coincides with the area element dAS2 ofthe unit sphere S2 in every coordinate chart. Hence we can write∫MK dAM =

∫S2 dAS2 = 4π, proving the theorem. �

12.11.2. Transition from classical to modern diff geom. Thetheorema egregium of Gauss marks the transition from classical differ-ential geometry of curves and surfaces imbedded in 3-space, to moderndifferential geometry of surfaces (and manifolds) studied intrinsically.

More specifically, once we have a notion of Gaussian curvature (andmore generally sectional curvature) that only depends on the metricon a surface (or manifold), we can study the geometry of the surface(or manifold) intrinsically, i.e., without any reference to a Euclideanimbedding.

To formulate the intrinsic viewpoint, one needs the notion of dualityof vector and covector. This theme is treated in Chapter 13.

Page 128: Analytic and Differential geometry Mikhail G. Katz∗
Page 129: Analytic and Differential geometry Mikhail G. Katz∗

CHAPTER 13

Duality in linear algebra, calculus, diff. geometry

13.1. Duality in linear algebra; 1-forms

Let V be a finite-dimensional real vector space.

Example 13.1.1. Euclidean space Rn is an example of a real vectorspace of dimension n with basis (e1, . . . , en).

Example 13.1.2. The tangent plane TpM of a regular surface Mat a point p ∈ M is a real vector space of dimension 2.

Definition 13.1.3. A linear form, also called 1-form, φ on V is alinear functional

φ : V → R.

Definition 13.1.4 (dx, dy). In the usual Euclidean plane of vectors

v = v1e1 + v2e2

represented by arrows, we denote by dx the 1-form which extracts theabscissa of the vector, and by dy the 1-form which extracts the ordinateof the vector:

dx(v) = v1,

anddy(v) = v2.

Example 13.1.5. For a vector v = 3e1+4e2 with components (3, 4)we obtain

dx(v) = 3, dy(v) = 4.

Definition 13.1.6. The quadratic forms dx2 and dy2 are definedby squaring the value of the 1-form on v:

dx2(v) = (dx(v))2.

Such quadratic forms are called rank-1 quadratic forms.

Thus,dx2(v) = (v1)2, dy2(v) = (v2)2.

Example 13.1.7. In the case v = 3e1 + 4e2 we obtain dx2(v) = 9and dy2(v) = 16.

129

Page 130: Analytic and Differential geometry Mikhail G. Katz∗

130 13. DUALITY IN LINEAR ALGEBRA, CALCULUS, DIFF. GEOMETRY

13.2. Dual vector space

Definition 13.2.1. The dual space of V , denoted V ∗, is the spaceof all linear forms on V :

V ∗ = {φ : φ is a 1-form on V } .

Evaluating φ at an element x ∈ V produces a scalar φ(x) ∈ R.

Definition 13.2.2. The evaluation map1 is the natural pairing be-tween V and V ∗, namely a linear map denoted

〈 , 〉 : V × V ∗ → R,

defined by evaluating y at x, i.e., setting 〈x, y〉 = y(x), for all x ∈ Vand y ∈ V ∗.

Remark 13.2.3. Note we are using the same notation for the pair-ing as for the scalar product in Euclidean space. The notation is quitewidespread.

Definition 13.2.4. If V admits a basis of vectors

(xi) = (x1, x2, . . . , xn),

then the dual space V ∗ admits a unique basis, called the dual ba-sis (yj) = (y1, . . . , yn), satisfying

〈xi, yj〉 = δij, (13.2.1)

for all i, j = 1, . . . , n, where δij is the Kronecker delta function.

Example 13.2.5. Let V = R2. We have the standard basis e1, e2for V . The 1-forms dx, dy form a basis for the dual space V ∗. Thenthe basis (dx, dy) is the dual basis to the basis (e1, e2).

Example 13.2.6. Let V = R2 identified with C for convenience.We have the basis (1, e

iπ3 ) for V where 1 = e1 + 0e2 while

eiπ3 =

1

2e1 +

√3

2e2.

Find the dual basis in V ∗. The answer is y1 = dx− dy√3, y2 =

2√3dy.

1hatzava?

Page 131: Analytic and Differential geometry Mikhail G. Katz∗

13.3. DERIVATIONS 131

13.3. Derivations

Let E be a space of dimension n, and let p ∈ E be a fixed point.The following definition is independent of coordinates.

Definition 13.3.1. Let

Dp = {f : f ∈ C∞}be the ring of real C∞-functions f defined in an arbitrarily small openneighborhood of p ∈ E.

Remark 13.3.2. The ring operations are pointwise multiplicationand pointwise addition, where we choose the intersection of the twodomains as the domain of the new function.

Note that Dp is infinite-dimensional as it includes all polynomials.

Theorem 13.3.3. Choose coordinates (u1, . . . , un) in E. A partialderivative ∂

∂uiat p is a 1-form

∂ui: Dp → R

on the space Dp, satisfying Leibniz’s rule

∂(fg)

∂ui

∣∣∣∣p

=∂f

∂ui

∣∣∣∣p

g(p) + f(p)∂g

∂ui

∣∣∣∣p

(13.3.1)

for all f, g ∈ Dp.

Formula (13.3.1) can be written briefly as

∂∂ui

(fg) = ∂∂ui

(f)g + f ∂∂ui

(g),

with the understanding that both sides are evaluated at the point p.This was proved in calculus. Formula (13.3.1) motivates the followingmore general definition of a derivation.

Definition 13.3.4. A derivation X at the point p ∈ E is a 1-form

X : Dp → R

on the space Dp satisfying Leibniz’s rule:

X(fg) = X(f)g(p) + f(p)X(g) (13.3.2)

for all f, g ∈ Dp.

Remark 13.3.5. Linearity of a derivation is required only withregard to scalars in R, not with respect to functions.

Page 132: Analytic and Differential geometry Mikhail G. Katz∗

132 13. DUALITY IN LINEAR ALGEBRA, CALCULUS, DIFF. GEOMETRY

13.4. Characterisation of derivations

It turns out that the space of derivations is spanned by partialderivatives.

Proposition 13.4.1. Let E be an n-dimensional space, and p ∈E. Then the collection of all derivations at p is a vector space ofdimension n.

Proof in case n = 1. We will first prove the result in the case n =1 of a single variable u at the point p = 0. Let X : Dp → R be a deriva-tion. Then

X(1) = X(1 · 1) = 2X(1)

by Leibniz’s rule. Therefore X(1) = 0, and similarly for any constantby linearity of X .

Now consider the monic polynomial u = u1 of degree 1, viewed asa linear function

u ∈ Dp=0.

We evaluate the derivation X at the element u ∈ D and set c = X(u).By the Taylor remainder formula, any function f ∈ Dp=0 can be writtenas

f(u) = a+ bu+ g(u)u

where b = ∂f∂u(0) and g is smooth and g(0) = 0. Now we have by

linearityX(f) = X(a+ bu+ g(u)u)

= bX(u) +X(g)u(0) + g(0) ·X(u)

= bc+ 0 + 0

= c∂

∂u(f).

Thus the derivation X coincides with the derivation c ∂∂u

for all f ∈ Dp.Hence the tangent space is 1-dimensional and spanned by the element∂∂u, proving the theorem in this case. �

13.4.1. The material in this subsection is optional.

Proof in case n = 2. Let us prove the result in the case n = 2 of twovariables u, v at the origin p = (0, 0). Let X be a derivation. Then

X(1) = X(1 · 1) = 2X(1)

by Leibniz’s rule. Therefore X(1) = 0, and similarly for any constant bylinearity of X. Now consider the monic polynomial u = u1 of degree 1, i.e.,the linear function

u ∈ Dp=0.

Page 133: Analytic and Differential geometry Mikhail G. Katz∗

13.5. TANGENT SPACE AND COTANGENT SPACE 133

We evaluate the derivation X at u and set c = X(u). Similarly, consider themonic polynomial v = v1 of degree 1, i.e., the linear function

v ∈ Dp=0.

We evaluate the derivation X at v and set c = X(v). By the Taylor re-mainder formula, any function f ∈ Dp=0, where f = f(u, v), can be writtenas

f(u, v) = a+ bu+ bv + g(u, v)u2 + h(u, v)uv + k(u, v)v2,

where the functions g(u, v), h(u, v), and k(u, v) are smooth. Note that the

coefficients b and b are the first partial derivatives of f at the origin. Nowwe have

X(f) = X(a+ bu+ bv + g(u, v)u2 + h(u, v)uv + k(u, v)v2)

= bX(u) + bX(v) +(X(g)u2 + g(u, v)2uX(u) + · · ·

)

= bX(u) + bX(v)

= bc+ bc

= c∂

∂u(f) + c

∂v(f)

by evaluating at the point (0, 0). Thus the derivation X coincides with thederivation

c∂

∂u+ c

∂vfor all test functions f ∈ Dp. Hence the two partials span the tangent space.Therefore the tangent space is 2-dimensional, proving the theorem in thiscase. �

13.5. Tangent space and cotangent space

Definition 13.5.1. Let p ∈ E. The space of derivations at p iscalled the tangent space Tp = TpE at p.

The results of the previous section can be formulated as follows.

Corollary 13.5.2. Let (u1, . . . , un) be coordinates for E, and letp ∈ E. Then a basis for the tangent space Tp is given by the n partialderivatives (

∂ui

), i = 1, . . . , n.

Definition 13.5.3. The space dual to the tangent space Tp is calledthe cotangent space, and denoted T ∗

p .

Definition 13.5.4. An element of a tangent space is a vector, whilean element of a cotangent space is called a 1-form, or a covector.

Page 134: Analytic and Differential geometry Mikhail G. Katz∗

134 13. DUALITY IN LINEAR ALGEBRA, CALCULUS, DIFF. GEOMETRY

Definition 13.5.5. The basis dual to the basis(∂∂ui

)is denoted

(duj), j = 1, . . . , n.

Thus each duj is by definition a linear form, or cotangent vector(covector for short). We are therefore working with dual bases

(∂∂ui

)

for vectors, and (duj) for covectors. The evaluation map as in (13.2.1)gives ⟨

∂ui, duj

⟩= duj

(∂

∂ui

)= δji , (13.5.1)

where δji is the Kronecker delta.

13.6. Constructing bilinear forms out of 1-forms

Recall that the polarisation formula (see definition 1.7.3) allowsone to reconstruct a symmetric bilinear form B = B(v, w), from thequadratic form Q(v) = B(v, v), at least if the characteristic is not 2:

B(v, w) =1

4(Q(v + w)−Q(v − w)). (13.6.1)

Similarly, one can construct bilinear forms out of the 1-forms dui,as follows.

Example 13.6.1. Consider a quadratic form

ai(dui)2

defined by a linear combination of the rank-1 quadratic forms (dui)2,as in Definition 13.1.6.

Polarizing the quadratic form, one obtains a bilinear form on thetangent space Tp.

Example 13.6.2. Let v = v1e1+ v2e2 be an arbitrary vector in theplane. Let dx and dy be the standard covectors, extracting, respec-tively, the first and second coordinates of v. Consider the quadraticform Q given by

Q = Edx2 + Fdy2,

where E, F ∈ R. Here Q(v) is calculated as

Q(v) = E(dx(v))2 + F (dy(v))2 = E(v1)2 + F (v2)2.

Polarisation then produces the bilinear form B = B(v, w), where vand w are arbitrary vectors, given by the formula

B(v, w) = E dx(v) dx(w) + F dy(v) dy(w).

Page 135: Analytic and Differential geometry Mikhail G. Katz∗

13.8. DUAL BASES IN DIFFERENTIAL GEOMETRY 135

Example 13.6.3. Setting E = F = 1 in the previous example, weobtain the standard scalar product in the plane:

B(v, w) = v · w = dx(v) dx(w) + dy(v) dy(w) = v1w1 + v2w2.

13.7. First fundamental form

Definition 13.7.1. A metric (or first fundamental form) g is asymmetric bilinear form on the tangent space at p:

g : Tp × Tp → R,

defined for all p and varying continuously and smoothly in p.

Remark 13.7.2. In Riemannian geometry one requires the associ-ated quadratic form to be positive definite. In relativity theory oneuses a form of type (3, 1).

Recall that the basis for Tp in coordinates (ui) is given by the tan-gent vectors

∂ui.

These are given by certain derivations (see Section 13.3).The first fundamental form g is traditionally expressed by a matrix

of coefficients called metric coefficients gij , giving the inner product ofthe i-th and the j-th vector in the basis:

gij = g

(∂

∂ui,∂

∂uj

),

where g is the first fundamental form.

Definition 13.7.3. The square norm the i-th vector is given bythe coefficient

gii =

∥∥∥∥∂

∂ui

∥∥∥∥2

.

We will express the first fundamental form in more intrinsic notationof quadratic forms built from 1-forms (covectors).

13.8. Dual bases in differential geometry

Let us now restrict attention to the case of 2 dimensions, i.e., thecase of surfaces. At every point p = (u1, u2), we have the metric coef-ficients gij = gij(u

1, u2). Each metric coefficient is thus a function oftwo variables.

We will only consider the case when the matrix is diagonal. This canalways be achieved, in two dimensions, at a point by a suitable changeof coordinates, by the uniformisation theorem (see Theorem 13.8.2).

Page 136: Analytic and Differential geometry Mikhail G. Katz∗

136 13. DUALITY IN LINEAR ALGEBRA, CALCULUS, DIFF. GEOMETRY

We set x = u1 and y = u2 to simplify notation. In the notationdeveloped in Section 13.6, we can write the quadratic form associatedwith the first fundamental form as follows:

g = g11(x, y)(dx)2 + g22(x, y)(dy)

2. (13.8.1)

For example, if the metric coefficients form an identity matrix: gij =δij, we obtain the standard flat metric

g = (dx)2 + (dy)2 (13.8.2)

or simply

g = dx2 + dy2.

Example 13.8.1 (Hyperbolic metric). Let g11 = g22 = 1y2

at each

point (x, y) where y > 0. This means that∣∣ ∂∂u1

∣∣ = 1yand

∣∣ ∂∂u2

∣∣ = 1y.

The resulting hyperbolic metric in the upperhalf plane {y > 0} isexpressed by the quadratic form

1

y2(dx2 + dy2

).

Note that this expression is undefined whenever y = 0. See Sec-tion 12.4. The hyperbolic metric in the upper half plane is a completemetric.

Closely related results are the Riemann mapping theorem and theconformal representation theorem.

Theorem 13.8.2 (Riemann mapping/uniformisation). Every met-ric on a connected2 surface is conformally equivalent to a metric ofconstant Gaussian curvature.

From the complex analytic viewpoint, the uniformisation theoremstates that every Riemann surface is covered by either the sphere, theplane, or the upper halfplane. Thus no notion of curvature is needed forthe statement of the uniformisation theorem. However, from the dif-ferential geometric point of view, what is relevant is that every confor-mal class of metrics contains a metric of constant Gaussian curvature.See [Ab81] for a lively account of the history of the uniformisationtheorem.

2kashir

Page 137: Analytic and Differential geometry Mikhail G. Katz∗

13.9. SURFACES OF REVOLUTION IN ISOTHERMAL COORDINATES 137

13.9. Surfaces of revolution in isothermal coordinates

The uniformisation theorem as stated it in Theorem 13.8.2 is anexistence theorem. It does not not provide an explicit recipe as to howone would go about finding a representation of the surface in termsof a metric of constant curvature, together with a suitable conformalfactor. In this section we describe an explicit presentation of this kindin the case of tori of revolution. We will make use of the differentialnotation for metrics developed in earlier sections.

Below we will express the metric of a surface of revolution in isother-mal coordinates. Recall that if a surface is obtained by revolving acurve (r(φ), z(φ)), we obtain metric coefficients g11 = r2 and g22 =(drdφ

)2+(dzdφ

)2.

Thus we obtain the following theorem.

Theorem 13.9.1. We have the following expression of the metricof surface of revolution:

r2dθ2 +

((drdφ

)2+(dzdφ

)2)dφ2. (13.9.1)

Proposition 13.9.2. Let (r(φ), z(φ)), where r(φ) > 0, be an ar-clength parametrisation of the generating curve of a surface of revolu-tion. Then the change of variable

ψ =

∫1

r(φ)dφ,

produces an isothermal parametrisation in terms of variables (θ, ψ).With respect to the new coordinates, the first fundamental form is givenby a scalar matrix with metric coeficients gij = r(φ(ψ))2δij, i.e., themetric is

r(φ(ψ))2(dθ2 + dψ2).

Remark 13.9.3. We thus obtain an explicit conformal equivalencebetween the metric on the surface of revolution and the standard flatmetric on the quotient of the (θ, ψ) plane. Such coordinates are referredto as “isothermal coordinates” in the literature.

The existence of such a parametrisation is of course predicted bythe uniformisation theorem (Theorem 13.8.2) in the case of a generalsurface.

Proof. Consider an arbitrary change of parameter φ = φ(ψ). Bychain rule,

dr

dψ=dr

dψ.

Page 138: Analytic and Differential geometry Mikhail G. Katz∗

138 13. DUALITY IN LINEAR ALGEBRA, CALCULUS, DIFF. GEOMETRY

Now consider again the first fundamental form (13.9.1). We need toimpose the condition

g11 = g22to ensure the condition that the matrix of metric coefficients is a scalar

matrix. Thus, we need to solve the equation r2 =(drdψ

)2+(dzdψ

)2, or

r2 =

((dr

)2

+

(dz

)2)(

)2

.

In the case when the generating curve is parametrized by arclength, weare therefore reduced to the equation

r =dφ

dψ,

or ψ =∫

dφr(φ)

. Substituting the new variable φ(ψ) in place of φ in

the parametrisation of the surface, we obtain a new parametrisation ofthe surface of revolution in coordinates (θ, ψ), such that the matrix ofmetric coefficients is a scalar matrix by construction. �

13.10. More on dual bases

Recall that if (x1, . . . , xn) is a basis for a vector space V then thedual vector space V ∗ possesses a basis called the dual basis and de-noted (y1, . . . , yn) satisfying 〈xi, yj〉 = yj(xi) = δij.

Example 13.10.1. In R2 we have a basis (x1, x2) =(∂∂x, ∂∂y

)in the

tangent plane Tp at a point p. The dual basis of 1-forms (y1, y2) for T∗p

is denoted (dx, dy). Thus we have⟨∂

∂x, dx

⟩= dx

(∂

∂x

)= 1

and ⟨∂

∂y, dy

⟩= dy

(∂

∂y

)= 1,

while ⟨∂

∂x, dy

⟩= dy

(∂

∂x

)= 0,

etcetera.

Similarly, in polar coordinates at a point p 6= 0 we have a basis(∂∂r, ∂∂θ

)for Tp, and a dual basis (dr, dθ) for T ∗

p . Thus we have⟨∂

∂r, dr

⟩= dr

(∂

∂r

)= 1

Page 139: Analytic and Differential geometry Mikhail G. Katz∗

13.10. MORE ON DUAL BASES 139

and ⟨∂

∂θ, dθ

⟩= dθ

(∂

∂θ

)= 1,

while ⟨∂

∂r, dθ

⟩= dθ

(∂

∂r

)= 0,

etcetera.Now in polar coordinates we have a natural area element r dr dθ.

Area of a region D is calculated by Fubini’s theorem as∫∫

D

r dr dθ =

∫ (∫rdr

)dθ.

Thus we have a natural basis (y1, y2) = (rdr, dθ) in T ∗p when p 6= 0,

i.e., y1 = rdr while y2 = dθ. Its dual basis (x1, x2) in Tp can be easilyidentified. It is

(x1, x2) =

(1

r

∂r,∂

∂θ

).

Indeed, we have

〈x1, y1〉 =⟨1

r

∂r, rdr

⟩= rdr

(1

r

∂r

)= r

1

rdr

(∂

∂r

)= 1,

etcetera.

Page 140: Analytic and Differential geometry Mikhail G. Katz∗
Page 141: Analytic and Differential geometry Mikhail G. Katz∗

CHAPTER 14

A hyperreal view

14.1. Successive extensions N, Z, Q, R, ∗R

Our reference for true infinitesimal calculus is Keisler’s textbook[Ke74], downloadable athttp://www.math.wisc.edu/~keisler/calc.html

We start by motivating the familiar sequence of extensions of num-ber systems

N → Z → Q → Rin terms of their applications in arithmetic, algebra, and geometry.Each successive extension is introduced for the purpose of solving prob-lems, rather than enlarging the number system for its own sake. Thus,the extension Q ⊆ R enables one to express the length of the diagonalof the unit square and the area of the unit disc in our number system.

The familiar continuum R is an Archimedean continuum, in thesense that it satisfies the following Archimedean property.

Definition 14.1.1. An ordered field extending N is said to satisfythe Archimedean property if

(∀ǫ > 0)(∃n ∈ N) [nǫ > 1] .

We will provisionally denote the real continuum A where “A” standsfor Archimedean. Thus we obtain a chain of extensions

N → Z → Q → A,

as above. In each case one needs an enhanced ordered1 number systemto solve an ever broader range of problems from algebra or geometry.

The next stage is the extension

N → Z → Q → A → B,

where B is a Bernoullian continuum containing infinitesimals, definedas follows.

Definition 14.1.2. A Bernoullian extension of R is any properextension which is an ordered field.

1sadur

141

Page 142: Analytic and Differential geometry Mikhail G. Katz∗

142 14. A HYPERREAL VIEW

Any Bernoullian extension allows us to define infinitesimals and dointeresting things with those. But things become really interesting ifwe assume the Transfer Principle (Section 14.4), and work in a truehyperreal field, defined as in Definition 14.3.4 below. We will providesome motivating comments for the transfer principle in Section 14.3.

14.2. Motivating discussion for infinitesimals

Infinitesimals can be motivated from three different angles: geo-metric, algebraic, and arithmetic/analytic.

θ

Figure 14.2.1. Horn angle θ is smaller than every rec-tilinear angle

(1) Geometric (horn angles): Some students have expressedthe sentiment that they did not understand infinitesimals untilthey heard a geometric explanation of them in terms of whatwas classically known as horn angles. A horn angle is thecrivice between a circle and its tangent line at the point oftangency. If one thinks of this crevice as a quantity, it is easy toconvince oneself that it should be smaller than every rectilinearangle (see Figure 14.2.1). This is because a sufficiently smallarc of the circle will be contained in the convex region cutout by the rectilinear angle no matter how small. When onerenders this in terms of analysis and arithmetic, one gets apositive quantity smaller than every positive real number. Wecite this example merely as intuitive motivation (our actualconstruction is different).

Page 143: Analytic and Differential geometry Mikhail G. Katz∗

14.3. INTRODUCTION TO THE TRANSFER PRINCIPLE 143

(2) Algebraic (passage from ring to field): The idea is torepresent an infinitesimal by a sequence tending to zero. Onecan get something in this direction without reliance on anyform of the axiom of choice. Namely, take the ring S of allsequences of real numbers, with arithmetic operations definedterm-by-term. Now quotient the ring S by the equivalencerelation that declares two sequences to be equivalent if theydiffer only on a finite set of indices. The resulting object S/Kis a proper ring extension of R, where R is embedded by meansof the constant sequences. However, this object is not a field.For example, it has zero divisors. But quotienting it furtherin such a way as to get a field, by extending the kernel K to amaximal ideal K ′, produces a field S/K ′, namely a hyperrealfield.

(3) Analytic/arithmetic: One can mimick the construction ofthe reals out of the rationals as the set of equivalence classesof Cauchy sequences, and construct the hyperreals as equiva-lence classes of sequences of real numbers under an appropriateequivalence relation.

14.3. Introduction to the transfer principle

The transfer principle is a type of theorem that, depending on thecontext, asserts that rules, laws or procedures valid for a certain num-ber system, still apply (i.e., are “transfered”) to an extended numbersystem.

Example 14.3.1. The familiar extension Q ⊆ R preserves the prop-erty of being an ordered field.

Example 14.3.2. To give a negative example, the extension R ⊆R∪{±∞} of the real numbers to the so-called extended reals does notpreserve the property of being an ordered field.

The hyperreal extension R ⊆ ∗R (defined below) preserves all first-order properties (i.e., properties involving quantification over elementsbut not over sets.

Example 14.3.3. The formula sin2 x+cos2 x = 1, true over R for allreal x, remains valid over ∗R for all hyperreal x, including infinitesimaland infinite values of x ∈ ∗R.

Thus the transfer principle for the extension R ⊆ ∗R is a theo-rem asserting that any statement true over R is similarly true over ∗R,

Page 144: Analytic and Differential geometry Mikhail G. Katz∗

144 14. A HYPERREAL VIEW

and vice versa. Historically, the transfer principle has its roots in theprocedures involving Leibniz’s Law of continuity.2

We will explain the transfer principle in several stages of increasingdegree of abstraction. More details can be found in Section 14.4.

Definition 14.3.4. An ordered field B, properly including the fieldA = R of real numbers (so that A $ B) and satisfying the TransferPrinciple, is called a hyperreal field. If such an extended field B is fixedthen elements of B are called hyperreal numbers,3 while the extendedfield itself is usually denoted ∗R.

Theorem 14.3.5. Hyperreal fields exist.

For example, a hyperreal field can be constructed as the quotient ofthe ring RN of sequences of real numbers, by an appropriate maximalideal.

Definition 14.3.6. A positive infinitesimal is a positive hyperrealnumber ǫ such that

(∀n ∈ N) [nǫ < 1]

More generally, we have the following.

Definition 14.3.7. A hyperreal number ε is said to be infinitelysmall or infinitesimal if

−a < ε < a

for every positive real number a.

In particular, one has ε < 12, ε < 1

3, ε < 1

4, ε < 1

5, etc. If ε > 0

is infinitesimal then N = 1εis positive infinite, i.e., greater than every

real number.A hyperreal number that is not an infinite number are called finite.

Sometimes the term limited is used in place of finite.Keisler’s textbook exploits the technique of representing the hyper-

real line graphically by means of dots indicating the separation betweenthe finite realm and the infinite realm. One can view infinitesimals withmicroscopes as in Figure 14.6.1. One can also view infinite numberswith telescopes as in Figure 14.3.1. We have an important subset

{finite hyperreals} ⊆ ∗R

2Leibniz’s theoretical strategy in dealing with infinitesimals was analyzed ina number of detailed studies recently, which found Leibniz’ strategy to be morerobust than George Berkeley’s flawed critique thereof.

3Similar terminology is used with regard to integers and hyperintegers.

Page 145: Analytic and Differential geometry Mikhail G. Katz∗

14.3. INTRODUCTION TO THE TRANSFER PRINCIPLE 145

Finite

−2 −1 0 1 2

Positive

infinite

Infinite

telescope1ǫ− 1

1/ǫ

1ǫ+ 1

Figure 14.3.1. Keisler’s telescope

which is the domain of the function called the standard part function(also known as the shadow) which rounds off each finite hyperreal tothe nearest real number (for details see Section 14.6).

Example 14.3.8. Slope calculation of y = x2 at x0. Here we usestandard part function (also called the shadow)

st : {finite hyperreals} → R

(for details concerning the shadow see Section 14.6). If a curve is de-fined by y = x2 we wish to find the slope at the point x0. To this endwe use an infinitesimal x-increment ∆x and compute the correspond-ing y-increment

∆y = (x0+∆x)2−x20 = (x0+∆x+x0)(x+∆x−x0) = (2x0+∆x)∆x.

The corresponding “average” slope is therefore

∆y

∆x= 2x0 +∆x

which is infinitely close to 2x, and we are naturally led to the definitionof the slope at x0 as the shadow of ∆y

∆x, namely st

(∆y∆x

)= 2x0.

Page 146: Analytic and Differential geometry Mikhail G. Katz∗

146 14. A HYPERREAL VIEW

The extension principle expresses the idea that all real objects havenatural hyperreal counterparts. We will be mainly interested in sets,functions and relations. We then have the following extension principle.

Extension principle. The order relation on the hyperreals con-tains the order relation on the reals. There exists a hyperreal numbergreater than zero but smaller than every positive real. Every set D ⊆ Rhas a natural extension ∗D ⊆ ∗R. Every real function f with domain Dhas a natural hyperreal extension ∗f with domain ∗D.4

Here the naturality of the extension alludes to the fact that suchan extension is unique, and the coherence refers to the fact that thedomain of the natural extension of a function is the natural extensionof its domain.

A positive infinitesimal is a positive hyperreal smaller than everypositive real. A negative infinitesimal is a negative hyperreal greaterthan every negative real. An arbitrary infinitesimal is either a positiveinfinitesimal, a negative infinitesimal, or zero.

Ultimately it turns out counterproductive to employ asterisks forhyperreal functions (in fact we already dropped it in equation (14.4.1)).

14.4. Transfer principle

Definition 14.4.1. The Transfer Principle asserts that every first-order statement true over R is similarly true over ∗R, and vice versa.

Here the adjective first-order alludes to the limitation on quantifi-cation to elements as opposed to sets.

Listed below are a few examples of first-order statements.

Example 14.4.2. The commutativity rule for addition x+y = y+xis valid for all hyperreal x, y by the transfer principle.

Example 14.4.3. The formula

sin2 x+ cos2 x = 1 (14.4.1)

is valid for all hyperreal x by the transfer principle.

Example 14.4.4. The statement

0 < x < y =⇒ 0 <1

y<

1

x(14.4.2)

holds for all hyperreal x, y.

4Here the noun principle (in extension principle) means that we are going toassume that there is a function ∗f : B → B which satisfies certain properties.It is a separate problem to define B which admits a coherent definition of ∗f forall f : A → A, to be solved below.

Page 147: Analytic and Differential geometry Mikhail G. Katz∗

14.5. ORDERS OF MAGNITUDE 147

Example 14.4.5. The characteristic function χQ of the rationalnumbers equals 1 on rational inputs and 0 on irrational inputs. Bythe transfer principle, its natural extension ∗χQ = χ∗Q will be 1 onhyperrational numbers ∗Q and 0 on hyperirrational numbers (namely,numbers in the complement ∗R \ ∗Q).

To give additional examples of real statements to which transferapplies, note that all ordered field -statements are subject to Trans-fer. As we will see below, it is possible to extend Transfer to a muchbroader category of statements, such as those containing the functionsymbols exp or sin or those that involve infinite sequences of reals.

A hyperreal number x is finite if there exists a real number r suchthat |x| < r. A hyperreal number is called positive infinite if it isgreater than every real number, and negative infinite if it is smallerthan every real number.

14.5. Orders of magnitude

Hyperreal numbers come in three orders of magnitude: infinitesi-mal, appreciable, and infinite. A number is appreciable if it is finitebut not infinitesimal. In this section we will outline the rules for ma-nipulating hyperreal numbers.

To give a typical proof, consider the rule that if ǫ is positive in-finitesimal then 1

ǫis positive infinite. Indeed, for every positive real r

we have 0 < ǫ < r. It follows from (14.4.2) by transfer that that 1ǫis

greater than every positive real, i.e., that 1ǫis infinite.

Let ǫ, δ denote arbitrary infinitesimals. Let b, c denote arbitraryappreciable numbers. Let H,K denote arbitrary infinite numbers. Wehave the following theorem.

Theorem 14.5.1. We have the following rules for addition:

• ǫ+ δ is infinitesimal;• b+ ǫ is appreciable;• b+ c is finite (possibly infinitesimal);• H + ǫ and H + b are infinite.

We have the following rules for products.

• ǫδ and bǫ are infinitesimal;• bc is appreciable;• Hb and HK are infinite.

We have the following rules for quotients.

• ǫb, ǫH, bH

are infinitesimal;

• bcis appreciable;

Page 148: Analytic and Differential geometry Mikhail G. Katz∗

148 14. A HYPERREAL VIEW

• bǫ, Hǫ, Hbare infinite provided ǫ 6= 0.

We have the following rules for roots, where n is a standard naturalnumber.

• if ǫ > 0 then n√ǫ is infinitesimal;

• if b > 0 then n√b is appreciable;

• if H > 0 then n√H is infinite.

Note that the traditional topic of the so-called “indeterminate forms”can be treated without introducing any ad-hoc terminology by meansof the following remark.

Remark 14.5.2. We have no rules in certain cases, such as ǫδ, HK,Hǫ,

and H +K.

These cases correspond to what are known since Moigno5 as inde-terminate forms.

Theorem 14.5.3. Arithmetic operations on the hyperreal numbersare governed by the following rules.

(1) every hyperreal number between two infinitesimals is infinites-imal.

(2) Every hyperreal number which is between two finite hyperrealnumbers, is finite.

(3) Every hyperreal number which is greater than some positiveinfinite number, is positive infinite.

(4) Every hyperreal number which is less than some negative infi-nite number, is negative infinite.

Example 14.5.4. The difference√H + 1 −

√H − 1 (where H is

infinite) is infinitesimal. Namely,

√H + 1−

√H − 1 =

(√H + 1−

√H − 1

) (√H + 1 +

√H − 1

)(√

H + 1 +√H − 1

)

=H + 1− (H − 1)(√H + 1 +

√H − 1

)

=2√

H + 1 +√H − 1

is infinitesimal. Once we introduce limits, this example can be refor-mulated as follows: limn→∞

(√n + 1−

√n− 1

)= 0.

5Elaborate on this historical note.

Page 149: Analytic and Differential geometry Mikhail G. Katz∗

14.7. DIFFERENTIATION 149

Definition 14.5.5. Two hyperreal numbers a, b are said to be in-finitely close, written

a ≈ b,

if their difference a− b is infinitesimal.

It is convenient also to introduce the following terminology andnotation.

Definition 14.5.6. Two nonzero hyperreal numbers a, b are saidto be adequal, written

a pq b,

if either ab≈ 1 or a = b = 0.

Note that the relation sin x ≈ x for infinitesimal x is immediatefrom the continuity of sine at the origin (in fact both sides are infin-itely close to 0), whereas the relation sin x pq x is a subtler relationequivalent to the computation of the first order Taylor approximationof sine.

14.6. Standard part principle

Theorem 14.6.1 (Standard Part Principle). Every finite hyperrealnumber x is infinitely close to an appropriate real number.

Proof. The result is generally true for an arbitrary proper orderedfield extension E of R. Indeed, if x ∈ E is finite, then x induces aDedekind cut on the subfield Q ⊆ R ⊆ E via the total order of E. Thereal number corresponding to the Dedekind cut is then infinitely closeto x. �

The real number infinitely close to x is called the standard part, orshadow, denoted st(x), of x.

We will use the notation ∆x, ∆y for infinitesimals.

Remark 14.6.2. There are three consecutive stages in a typicalcalculation: (1) calculations with hyperreal numbers, (2) calculationwith standard part, (3) calculation with real numbers.

14.7. Differentiation

An infinitesimal increment ∆x can be visualized graphically bymeans of a microscope as in the Figure 14.7.1.

The slope s of a function f at a real point a is defined by setting

s = st

(f(a+∆x)− f(a)

∆x

)

Page 150: Analytic and Differential geometry Mikhail G. Katz∗

150 14. A HYPERREAL VIEW

r r + β r + γr + α

st

−1

−1

0

0

1

1

2

2

3

3

4

4r

Figure 14.6.1. The standard part function, st, “roundsoff” a finite hyperreal to the nearest real number. The func-tion st is here represented by a vertical projection. Keisler’s“infinitesimal microscope” is used to view an infinitesimalneighborhood of a standard real number r, where α, β,and γ represent typical infinitesimals. Courtesy ofWikipedia.

whenever the shadow exists (i.e., the ratio is finite) and is the samefor each nonzero infinitesimal ∆x. The construction is illustrated inFigure 14.7.2.

Definition 14.7.1. Let f be a real function of one real variable.The derivative of f is the new function f ′ whose value at a real x isthe slope of f at x. In symbols,

f ′(x) = st

(f(x+∆x)− f(x)

∆x

)

whenever the slope exists.

Equivalently, we can write f ′(x) ≈ f(x+∆x)−f(x)∆x

. When y = f(x)we define a new dependent variable ∆y by settting

∆y = f(x+∆x)− f(x)

called the y-increment, so we can write the derivative as st(∆y∆x

).

Page 151: Analytic and Differential geometry Mikhail G. Katz∗

14.7. DIFFERENTIATION 151

a

a a+∆x

Figure 14.7.1. Infinitesimal increment ∆x under the microscope

Example 14.7.2. If f(x) = x2 we obtain the derivative of y = f(x)by the following direct calculation:

f ′(x) ≈ ∆y

∆x

=(x+∆x)2 − x2

∆x

=(x+∆x− x)(x+∆x+ x)

∆x

=∆x(2x+∆x)

∆x= 2x+∆x

≈ 2x.

Page 152: Analytic and Differential geometry Mikhail G. Katz∗

152 14. A HYPERREAL VIEW

a

|f(a+∆x)− f(a)|

a a+∆x

Figure 14.7.2. Defining slope of f at a

Given a function y = f(x) one defines the dependent variable ∆y =f(x+∆x)−f(x) as above. One also defines a new dependent variable dyby setting dy = f ′(x)∆x at a point where f is differentiable, and setsfor symmetry dx = ∆x. Note that we have

dy pq ∆y

as in Definition 14.5.6 whenever dy 6= 0. We then have Leibniz’s nota-tion

f ′(x) =dy

dxand rules like the chain rule acquire an appealing form.

Page 153: Analytic and Differential geometry Mikhail G. Katz∗

CHAPTER 15

Lattices and tori

15.1. Circle via the exponential map

We will discuss lattices in Euclidean space Rb in Section 15.2. Herewe give an intuitive introduction in the simplest case b = 1. Everylattice (discrete1 subgroup; see definition below in Section 15.2) in R =R1 is of the form

Lα = αZ = {nα : n ∈ Z} ⊆ R,

for some real α > 0. It is spanned by the vector αe1 (or −αe1).The lattice Lα ⊆ R is in fact an additive subgroup. Therefore

we can form the quotient group R/Lα. This quotient is a circle (seeTheorem 15.1.1). The 1-volume, i.e. the length, of the quotient circleis precisely α. We will give a description in terms of the complexfunction ez.

Theorem 15.1.1. The quotient group R/Lα is isomorphic to thecircle S1 ⊆ C.

Proof. Consider the map φ : R → C defined by

φ(x) = ei2πxα .

By the usual addition rule for the exponential function, this map is ahomomorphism from the additive structure on R to the multiplicativestructure in the group C \ {0}. Namely, we have

(∀x ∈ R)(∀y ∈ R) φ(x+ y) = φ(x)φ(y).

Furthermore, we have φ(x + αm) = φ(x) for all m ∈ Z and therefore

ker φ = Lα. By the group-theoretic isomorphism theorem, the map φdescends to a map

φ : R/Lα → C,

which is injective. Its image is the unit circle S1 ⊆ C, which is a groupunder multiplication. �

1b’didah

153

Page 154: Analytic and Differential geometry Mikhail G. Katz∗

154 15. LATTICES AND TORI

15.2. Lattice, fundamental domain

Let b > 0 be an integer.

Definition 15.2.1. A lattice L ⊆ Rb is the integer span of a linearlyindependent set of b vectors.

Thus, if vectors v1, . . . , vb are linearly independent, then they spana lattice

L = {n1v1 + . . .+ nbvb : ni ∈ Z} = Zv1 + Zv2 · · ·+ Zvb

Note that the subgroup is isomorphic to Zb.

Definition 15.2.2. An orbit of a point x0 ∈ Rb under the actionof a lattice L is the subset of Rb given by the collection of elements

{x0 + g : g ∈ L} .These can also be viewed as the cosets of the lattice in Rb.

Definition 15.2.3. The quotient

Rb/L

is called a b-torus.

At this point tori are understood at the group-theoretic level as inthe case of the circle R/L.

15.3. Fundamental domain

Definition 15.3.1. A fundamental domain for the torus Rb/L is aclosed set F ⊆ Rb satisfying the following three conditions:

• every orbit meets F in at least one point;• every orbit meets the interior Int(F ) of F in at most one point;• the boundary ∂F is of zero b-dimensional volume (and can bethought of as a union of (b− 1)-dimensional hyperplanes).

In the literature, one often replaces “n-dimensional volume” by n-dimensional “Lebesgue measure”.

Example 15.3.2. The interval [0, α] is a fundamental domain forthe circle R/Lα.

Example 15.3.3. The parallelepiped spanned by a collection ofbasis vectors for L ⊆ Rb is a fundamental domain fot L.

More concretely, consider the following example.

Page 155: Analytic and Differential geometry Mikhail G. Katz∗

15.4. LATTICES IN THE PLANE 155

Example 15.3.4. The vectors e1 and e2 in R2 span the unit squarewhich is a fundamental domain for the lattice Z2 ⊆ R2 = C of Gaussianintegers.

Example 15.3.5. Consider the vectors v = (1, 0) and w = (12,√32)

in R2 = C. Their span is a parallelogram giving a fundamental domainfor the lattice of Eisenstein integers (see Example 15.4.2).

Definition 15.3.6. The total volume of the b-torus Rb/L is bydefinition the b-volume of a fundamental domain.

It is shown in advanced calculus that the total volume thus definedis independent of the choice of a fundamental domain.

15.4. Lattices in the plane

Let b = 2. Every lattice L ⊆ R2 is of the form

L = SpanZ(v, w) ⊆ R2,

where {v, w} is a linearly independent set. For example, let α and βbe nonzero reals. Set

Lα,β = SpanZ(αe1, βe2) ⊆ R2.

This lattice admits an orthogonal basis, namely {αe1, βe2}.

Example 15.4.1 (Gaussian integers). For the standard lattice Zb ⊆Rb, the torus Tb = Rb/Zb satisfies vol(Tb) = 1 as it has the unit cubeas a fundamental domain.

In dimension 2, the resulting lattice in C = R2 is called the Gaussianintegers LG. It contains 4 elements of least length. These are the fourthroots of unity. We have

LG = SpanZ(1, i) ⊆ C.

Example 15.4.2 (Eisenstein integers). Consider the lattice LE ⊆R2 = C spanned by 1 ∈ C and the sixth root of unity e

2πi6 ∈ C:

LE = SpanZ(eiπ/3, 1) = Z eiπ/3 + Z 1 ⊆ C. (15.4.1)

The resulting lattice is called the Eisenstein integers. The torus T2 =

R2/LE satisfies area(T2) =√32. The Eisenstein lattice contains 6 ele-

ments of least length, namely all the sixth roots of unity.

Page 156: Analytic and Differential geometry Mikhail G. Katz∗

156 15. LATTICES AND TORI

15.5. Successive minima of a lattice

Let B be Euclidean space, and let ‖ ‖ be the Euclidean norm.Let L ⊆ (B, ‖ ‖) be a lattice, i.e., span of a collection of b linearlyindependent vectors where b = dim(B).

Definition 15.5.1. The first successive minimum, λ1(L, ‖ ‖) is theleast length of a nonzero vector in L.

We can express the definition symbolically by means of the formula

λ1(L, ‖ ‖) = min{‖v1‖

∣∣ v1 ∈ L \ {0}}.

We illustrate the geometric meaning of λ1 in terms of the circle ofTheorem 15.1.1.

Theorem 15.5.2. Consider a lattice L ⊆ R. Then the circle R/Lsatisfies

length(R/L) = λ1(L).

Proof. This follows by choosing the fundamental domain F =[0, α] where α = λ1(L), so that L = αZ, cf. Example 15.2 above. �

Remark 15.5.3. When α = 1, we can choose a representative fromthe orbit of x to be the fractional part {x} of x.

Definition 15.5.4. For k = 2, define the second successive min-imum of the lattice L with rank(L) ≥ 2 as follows. Given a pair ofvectors S = {v, w} in L, define the size2 |S| of S by setting

|S| = max(‖v‖, ‖w‖).Then the second successive minimum, λ2(L, ‖ ‖) is the least size of apair of non-proportional vectors in L:

λ2(L) = infS|S|,

where S runs over all linearly independent (i.e. non-proportional) pairsof vectors {v, w} ⊆ L.

Example 15.5.5. For both the Gaussian and the Eisenstein integerswe have λ1 = λ2 = 1.

Example 15.5.6. For the lattice Lα,β we have λ1(Lα,β) = min(|α|, |β|)and λ2(Lα,β) = max(|α|, |β|).

2Quotation marks: merka’ot.

Page 157: Analytic and Differential geometry Mikhail G. Katz∗

15.7. SPHERE AND TORUS AS TOPOLOGICAL SURFACES 157

15.6. Gram matrix

The volume of the torus Rb/L (see Definition 15.2.3) is also calledthe covolume of the lattice L. It is by definition the volume of a funda-mental domain for L, e.g. a parallelepiped spanned by a Z-basis for L.

The Gram matrix was defined in Definition 5.6.2.

Theorem 15.6.1. Let L ⊂ Rb be a lattice spanned by linearly inde-pendent vectors (v1, . . . , vb). Then the volume of the torus Rb/L is thesquare root of the determinant of the Gram matrix Gram(v1, . . . , vb).

In geometric terms, the parallelepiped P spanned by the vectors {vi}satisfies

vol(P ) =√

det(Gram(S)). (15.6.1)

Proof. Let A be the square matrix whose columns are the columnvectors v1, v2 . . . , vn in Rn. It is shown in linear algebra that

vol(P ) = |det(A)|.Let B = AtA, and let B = (bij). Then

bij = vti vj = 〈vi, vj〉Hence B = Gram(S). Thus

det(Gram(S)) = det(At A) = det(A)2 = vol(P )2

proving the theorem. �

15.7. Sphere and torus as topological surfaces

The topology of surfaces will be discussed in more detail in Chap-ter 17.17. For now, we will recall that a compact surface can be eitherorientable or non-orientable. An orientable surface is characterizedtopologically by its genus, i.e. number of “handles”.

Recall that the unit sphere in R3 can be represented implicitly bythe equation

x2 + y2 + z2 = 1.

Parametric representations of surfaces are discussed in Section 5.3.

Example 15.7.1. The sphere has genus 0 (no handles).

Theorem 15.7.2. The 2-torus is characterized topologically in oneof the following four equivalent ways:

(1) the Cartesian product of a pair of circles: S1 × S1;(2) the surface of revolution in R3 obtained by starting with the

following circle in the (x, z)-plane: (x − 10)2 + z2 = 1 (forexample), and rotating it around the z-axis;

Page 158: Analytic and Differential geometry Mikhail G. Katz∗

158 15. LATTICES AND TORI

(3) a quotient R2/L of the plane by a lattice L;(4) a compact 2-dimensional manifold of genus 1.

The equivalence between items (2) and (3) can be seen by markinga pair of generators of L by different color, and using the same colors toindicate the corresponding circles on the imbedded torus of revolution,as follows:

Figure 15.7.1. Torus viewed by means of its lattice(left) and by means of a Euclidean imbedding (right)

Note by comparison that a circle can be represented either by itsfundamental domain which is [0, 2π] (with endpoints identified), or asthe unit circle imbedded in the plane.3

15.8. Standard fundamental domain

We will discuss the case b = 2 in detail. An important role is playedin this dimension by the standard fundamental domain.

3The Hermite constant γb is defined in one of the following two equivalent ways:

(1) γb is the square of the maximal first successive minimum λ1, among alllattices of unit covolume;

(2) γb is defined by the formula

√γb = sup

{λ1(L)

vol(Rb/L)1

b

∣∣∣∣L ⊆ (Rb, ‖ ‖)}, (15.7.1)

where the supremum is extended over all lattices L in Rb with a Euclideannorm ‖ ‖.

A lattice realizing the supremum may be thought of as the one realizing the densestpacking in Rb when we place the balls of radius 1

2λ1(L) at the points of L. Indimensions b ≥ 3, the Hermite constants are harder to compute, but explicit values(as well as the associated critical lattices) are known for small dimensions, e.g. γ3 =

21

3 = 1.2599..., while γ4 =√2 = 1.4142....

Page 159: Analytic and Differential geometry Mikhail G. Katz∗

15.9. CONFORMAL PARAMETER τ OF A LATTICE 159

Definition 15.8.1. The standard fundamental domain, denoted D,is the set

D =

{z ∈ C

∣∣∣∣ |z| ≥ 1, |Re(z)| ≤ 1

2, Im(z) > 0

}(15.8.1)

cf. [Ser73, p. 78].

The domain D a fundamental domain for the action of PSL(2,Z)in the upperhalf plane of C.

Lemma 15.8.2. Multiplying a lattice L ⊆ C by nonzero complexnumbers does not change the value of the quotient

λ1(L)2

area(C/L).

Proof. We write such a complex number as reiθ. Note that multi-plication by reiθ can be thought of as a composition of a scaling by thereal factor r, and rotation by angle θ. The rotation is an isometry (con-gruence) that preserves all lengths, and in particular the length λ1(L)and the area of the quotient torus.

Meanwhile, multiplication by r results in a cancellation

λ1(rL)2

area(C/rL)=

(rλ1(L))2

r2 area(C/L)=

r2λ1(L)2

r2 area(C/L)=

λ1(L)2

area(C/L),

proving the lemma. �

15.9. Conformal parameter τ of a lattice

Two lattices in C are said to be similar if one is obtained from theother by multiplication by a nonzero complex number.

Theorem 15.9.1. Every lattice in C is similar to a lattice spannedby {τ, 1} where τ is in the standard fundamental domain D of (15.8.1).The value τ = eiπ/3 corresponds to the Eisenstein integers (15.4.1).

Proof. Let L ⊆ C be a lattice. Choose a “shortest” vector z ∈L, i.e. we have |z| = λ1(L). By Lemma 15.8.2, we may replace thelattice L by the lattice z−1L.

Thus, we may assume without loss of generality that the com-plex number +1 ∈ C is a shortest element in the lattice L. Thuswe have λ1(L) = 1. Now complete the element +1 to a Z-basis

{τ ,+1}for L. Here we may assume, by replacing τ by−τ if necessary, that Im(τ) >0.

Page 160: Analytic and Differential geometry Mikhail G. Katz∗

160 15. LATTICES AND TORI

Now consider the real part Re(τ). We adjust the basis by adding asuitable integer k to τ :

τ = τ − k where k =[Re(τ) + 1

2

](15.9.1)

(the brackets denote the integer part), so it satisfies the condition

−1

2≤ Re(τ) ≤ 1

2.

Since τ ∈ L, we have |τ | ≥ λ1(L) = 1. Therefore the element τ lies inthe standard fundamental domain (15.8.1). �

Example 15.9.2. For the “rectangular” lattice Lα,β = SpanZ(α, βi),we obtain

τ(Lα,β) =

{ |β||α|i if |β| > |α||α||β| i if |α| > |β|.

Corollary 15.9.3. Let b = 2. Then we have the following valuefor the Hermite constant: γ2 = 2√

3= 1.1547.... The corresponding

optimal lattice is homothetic to the Z-span of cube roots of unity in C(i.e. the Eisenstein integers).

Proof. Choose τ as in (15.9.1) above. The pair

{τ,+1}

is a basis for the lattice. The imaginary part satisfies Im(τ) ≥√32, with

equality possible precisely for

τ = eiπ3 or τ = ei

2π3 .

Moreover, if τ = r exp(iθ), then

|τ | sin θ = Im(τ) ≥√3

2.

The proof is concluded by calculating the area of the parallelogramin C spanned by τ and +1;

λ1(L)2

area(C/L)=

1

|τ | sin θ ≤ 2√3,

proving the theorem. �

Definition 15.9.4. A τ ∈ D is said to be the conformal parameterof a flat torus T 2 if T 2 is similar to a torus C/L where L = Zτ + Z1.

Page 161: Analytic and Differential geometry Mikhail G. Katz∗

15.10. CONFORMAL PARAMETER τ OF TORI OF REVOLUTION 161

15.10. Conformal parameter τ of tori of revolution

The results of Section 13.9 have the following immediate conse-quence.

Corollary 15.10.1. Consider a torus of revolution in R3 formedby rotating a Jordan curve of length L > 0, with unit speed param-etisation (f(φ), g(φ)) where φ ∈ [0, L]. Then the torus is conformallyequivalent to a flat torus

R2/Lc,d.

Here R2 is the (θ, ψ)-plane, where ψ is the antiderivative of 1f(φ)

as in

Section 13.9; while the rectangular lattice Lc,d ⊂ R2 is spanned by theorthogonal vectors c ∂

∂θand d ∂

∂ψ, so that

Lc,d = Span

(c∂

∂θ, d

∂ψ

)= cZ⊕ dZ,

where c = 2π and d =∫ L0

dφf(φ)

.

Figure 15.10.1. Torus: lattice (left) and imbedding (right)

In Section 15.8 we showed that every flat torus C/L is similar tothe torus spanned by τ ∈ C and 1 ∈ C, where τ is in the standardfundamental domain

D = {z = x+ iy ∈ C : |x| ≤ 12, y > 0, |z| ≥ 1}.

Definition 15.10.2. The parameter τ is called the conformal pa-rameter of the torus.

Corollary 15.10.3. The conformal parameter τ of a torus of rev-olution is pure imaginary:

τ = iσ2

Page 162: Analytic and Differential geometry Mikhail G. Katz∗

162 15. LATTICES AND TORI

of absolute value

σ2 = max

{c

d,d

c

}≥ 1.

Proof. The proof is immediate from the fact that the lattice isrectangular. �

15.11. θ-loops and φ-loops on tori of revolution

Consider a torus of revolution (T 2, g) generated by a Jordan curve Cin the (x, z)-plane, i.e., by a simple loop C, parametrized by a pair offunctions f(φ), g(φ), so that x = f(φ) and z = g(φ).

Definition 15.11.1. A φ-loop on the torus is a simple loop ob-tained by fixing the coordinate θ (i.e., the variable φ is changing). A θ-loop on the torus is a simple loop obtained by fixing the coordinate φ(i.e., the variable θ is changing).

Proposition 15.11.2. All φ-loops on the torus of revolution havethe same length equal to the length L of the generating curve C (seeCorollary 15.10.1).

Proof. The surface is rotationally invariant. In other words, allrotations around the z-axis are isometries. Therefore all φ-loops havethe same length. �

Proposition 15.11.3. The θ-loops on the torus of revolution havevariable length, depending on the φ-coordinate of the loop. Namely, thelength is 2πx = 2πf(φ).

Proof. The proof is immediate from the fact that the function f(φ)gives the distance r to the z-axis. �

Definition 15.11.4. We denote by λφ the (common) length of allφ-loops on a torus of revolution.

Definition 15.11.5. We denote by λθminthe least length of a θ-

loop on a torus of revolution, and by λθmaxthe maximal length of such

a θ-loop.

15.12. Tori generated by round circles

Let a, b > 0. We assume a > b so as to obtain tori that are imbeddedin 3-space. We consider the 2-parameter family ga,b of tori of revolutionin 3-space with circular generating loop. The torus of revolution ga,bgenerated by a round circle is the locus of the equation

(r − a)2 + z2 = b2, (15.12.1)

Page 163: Analytic and Differential geometry Mikhail G. Katz∗

15.13. CONFORMAL PARAMETER OF TORI OF REVOLUTION, RESIDUES163

where r =√x2 + y2. Note that the angle θ of the cylindrical coordi-

nates (r, θ, z) does not appear in the equation (15.12.1). The torus isobtained by rotating the circle

(x− a)2 + z2 = b2 (15.12.2)

around the z-axis in R3. The torus admits a parametrisation in termsof the functions4 f(φ) = a + b cosφ and g(φ) = b sinφ. Namely, wehave

x(θ, φ) = ((a+ b cos φ) cos θ, (a+ b cosφ) sin θ, b sinφ). (15.12.3)

Here the θ-loop (see Section 15.11) has length 2π(a+ b cos φ). Theshortest θ-loop is therefore of length

λθmin= 2π(a− b),

and the longest one is

λθmax= 2π(a+ b).

Meanwhile, the φ-loop has length

λφ = 2πb.

15.13. Conformal parameter of tori of revolution, residues

This section is optional.We would like to compute the conformal parameter τ of the stan-

dard tori as in (15.12.3). We first modify the parametrisation so as toobtain a generating curve parametrized by arclength:

f(ϕ) = a+ b cosϕ

b, g(ϕ) = b sin

ϕ

b, (15.13.1)

where ϕ ∈ [0, L] with L = 2πb.

Theorem 15.13.1. The corresponding flat torus is given by the

lattice L in the (θ, ψ) plane of the form L = SpanZ

(c ∂∂θ, d ∂

∂φ

)

where c = 2π and d = 2π√(a/b)2−1

. Thus the conformal parameter τ

of the flat torus satisfies τ = imax(((a/b)2 − 1)−1/2, ((a/b)2 − 1)1/2

).

Proof. By Corollary 15.10.3, replacing ϕ by ϕ(ψ) produces isother-mal coordinates (θ, ψ) for the torus generated by (15.13.1), where ψ =∫

dϕf(ϕ)

=∫

dϕa+b cos ϕ

b, and therefore the flat metric is defined by a lattice

in the (θ, ψ) plane with c = 2π and d =∫ L=2πb

0dϕ

a+b cos ϕb. Changing

the variable to to φ = ϕbwe obtain d =

∫ 2π

0dφ

(a/b)+cos φ, where a/b > 1.

4We use φ here and ϕ for the modified arclength parameter in the next section

Page 164: Analytic and Differential geometry Mikhail G. Katz∗

164 15. LATTICES AND TORI

Let α = a/b. Now the integral is the real part Re of the complex

integral d =∫ 2π

0dφ

α+cosφ=∫

dφα+Re(eiφ)

. Thus

d =

∫2dφ

2α+ eiφ + e−iφ. (15.13.2)

The change of variables z = eiφ yields dφ = −idzz

and along the circle

we have d =∮ −2idz

z(2α+z+z−1)=∮ −2idz

z2+2αz+1=∮ −2idz

(z−λ1)(z−λ2) , where λ1 =

−α+√α2 − 1 and λ2 = −α−

√α2 − 1. The root λ2 is outside the unit

circle. Hence we need the residue at λ1 to apply the residue theorem.The residue at λ1 equals Resλ1 = −2i

λ1−λ2 = −2i2√α2−1

= −i√α2−1

. The inte-

gral is determined by the residue theorem in terms of the residue at thepole z = λ1. Therefore the lattice parameter d from Corollary 15.10.3can be computed from (15.13.2) as d =

(2πiResλ1

)= 2π√

(α)2−1. proving

the theorem. �

Page 165: Analytic and Differential geometry Mikhail G. Katz∗

CHAPTER 16

Systole

16.1. Definition of systole

The unit circle S1 ⊂ C bounds the unit disk D. A loop on asurface M is a continuous map S1 →M .

Definition 16.1.1. A loop S1 → M is called contractible if themap f extends from S1 to the disk D by means of a continuous map F :D →M .

Thus the restriction of F to S1 is f .The notions of contractible loops and simply connected spaces were

reviewed in more detail in Section 17.18.

Definition 16.1.2. A loop is called noncontractible if it is notcontractible.

Definition 16.1.3. Given a metric g onM , we will denote by sys1(g),the infimum of lengths, referred to as the “systole” of g, of a non-contractible loop β in a compact, non-simply-connected Riemannianmanifold (M, g):

sys1(g) = infβlength(β), (16.1.1)

where the infimum is over all noncontractible loops β in M . In graphtheory, a similar invariant is known as the girth [Tu47].1

It can be shown that for a compact Riemannian manifold, the in-fimum is always attained, cf. Theorem 17.10.1. A loop realizing theminimum is necessarily a simple closed geodesic.

In systolic questions about surfaces, integral-geometric identitiesplay a particularly important role. Roughly speaking, there is an inte-gral identity relating area on the one hand, and an average of energiesof a suitable family of loops, on the other. By the Cauchy-Schwarzinequality, there is an inequality relating energy and length squared,hence one obtains an inequality between area and the square of thesystole.

1The notion of systole expressed by (16.1.1) is unrelated to the systolic arraysof [Ku78].

165

Page 166: Analytic and Differential geometry Mikhail G. Katz∗

166 16. SYSTOLE

Such an approach works both for the Loewner inequality (16.2.1)and Pu’s inequality (16.1.6) (biographical notes on C. Loewner andP. Pu appear in respectively). One can prove an inequality for theMobius band this way, as well [Bl61b].

Here we prove the two classical results of systolic geometry, namelyLoewner’s torus inequality as well as Pu’s inequality for the real pro-jective plane.

16.1.1. Three systolic invariants. The material in this subsec-tion is optional.

LetM be a Riemannian manifold. We define the homology 1-systole

sys1(M) (16.1.2)

by minimizing vol(α) over all nonzero homology classes. Namely, sys1(M)is the least length of a loop C representing a nontrivial homologyclass [C] in H1(M ;Z).

We also define the stable homology systole

stsys1(M) = λ1 (H1(M)/T1, ‖ ‖) , (16.1.3)

namely by minimizing the stable norm ‖ ‖ of a class of infinite order(see Definition 17.30.1 for details).

Remark 16.1.4. For the real projective plane, these two systolic in-variants are not the same. Namely, the homology systole sys1 equals theleast length of a noncontractible loop (which is also nontrivial homolog-ically), while the stable systole is infinite being defined by a minimumover an empty set.

Recall the following example from the previous section:

Example 16.1.5. For an arbitrary metric on the 2-torus T2, the1-systole and the stable 1-systole coincide by Theorem 17.22.3:

sys1(T2) = stsys1(T

2),

for every metric on T2.

Using the notion of a noncontractible loop, we can define the ho-motopy 1-systole

sys1(M) (16.1.4)

as the least length of a non-contractible loop in M .In the case of the torus, the fundamental group Z2 is abelian and

torsionfree, and therefore sys1(T2) = sys1(T

2), so that all three invari-ants coincide in this case.

Page 167: Analytic and Differential geometry Mikhail G. Katz∗

16.1. DEFINITION OF SYSTOLE 167

16.1.2. Isoperimetric inequality and Pu’s inequality. Thematerial in this section is optional.

Pu’s inequality can be thought of as an “opposite” isoperimetricinequality, in the following precise sense.

The classical isoperimetric inequality in the plane is a relation be-tween two metric invariants: length L of a simple closed curve in theplane, and area A of the region bounded by the curve. Namely, everysimple closed curve in the plane satisfies the inequality

A

π≤(L

)2

.

This classical isoperimetric inequality is sharp, insofar as equality isattained only by a round circle.

In the 1950’s, Charles Loewner’s student P. M. Pu [Pu52] provedthe following theorem. Let RP2 be the real projective plane endowedwith an arbitrary metric, i.e. an imbedding in some Rn. Then

(L

π

)2

≤ A

2π, (16.1.5)

where A is its total area and L is the length of its shortest non-contractible loop. This isosystolic inequality, or simply systolic inequal-ity for short, is also sharp, to the extent that equality is attained onlyfor a metric of constant Gaussian curvature, namely antipodal quotientof a round sphere, cf. Section 17.11. In our systolic notation (16.1.1),Pu’s inequality takes the following form:

sys1(g)2 ≤ π

2area(g), (16.1.6)

for every metric g on RP2. See Theorem 17.13.2 for a discussion of theconstant. The inequality is proved in Section 19.5. Pu’s inequality canbe generalized as follows. We will say that a surface is aspherical if itis not a 2-sphere.

Theorem 16.1.6. Every aspherical surface (M, g) satisfies the op-timal bound (16.1.6), attained precisely when, on the one hand, thesurface M is a real projective plane, and on the other, the metric g isof constant Gaussian curvature.

The extension to aspherical surfaces follows from Gromov’s inequal-ity (16.1.7) below (by comparing the numerical values of the two con-stants). Namely, every aspherical compact surface (M, g) admits ametric ball

B = Bp

(12sys1(g)

)⊆M

Page 168: Analytic and Differential geometry Mikhail G. Katz∗

168 16. SYSTOLE

of radius 12sys1(g) which satisfies [Gro83, Corollary 5.2.B]

sys1(g)2 ≤ 4

3area(B). (16.1.7)

16.1.3. Hermite and Berge-Martinet constants. The mate-rial in this subsection is optional.

Most of the material in this section has already appeared in earlierchapters.

Let b ∈ N. The Hermite constant γb is defined in one of the followingtwo equivalent ways:

(1) γb is the square of the biggest first successive minimum, cf. Defi-nition 17.14.1, among all lattices of unit covolume;

(2) γb is defined by the formula

√γb = sup

{λ1(L)

vol(Rb/L)1/b

∣∣∣∣L ⊆ (Rb, ‖ ‖)}, (16.1.8)

where the supremum is extended over all lattices L in Rb witha Euclidean norm ‖ ‖.

A lattice realizing the supremum is called a critical lattice. A crit-ical lattice may be thought of as the one realizing the densest packingin Rb when we place balls of radius 1

2λ1(L) at the points of L.

The existence of the Hermite constant, as well as the existence ofcritical lattices, are both nontrivial results [Ca71].

Theorem 15.9.1 provides the value for γ2.

Example 16.1.7. In dimensions b ≥ 3, the Hermite constants areharder to compute, but explicit values (as well as the associated crit-

ical lattices) are known for small dimensions (≤ 8), e.g. γ3 = 213 =

1.2599 . . ., while γ4 =√2 = 1.4142 . . .. Note that γn is asymptotically

linear in n, cf. (16.1.11).

A related constant γ′b is defined as follows, cf. [BeM].

Definition 16.1.8. The Berge-Martinet constant γ′b is defined bysetting

γ′b = sup{λ1(L)λ1(L

∗)∣∣L ⊆ (Rb, ‖ ‖)

}, (16.1.9)

where the supremum is extended over all lattices L in Rb.

Here L∗ is the lattice dual to L. If L is the Z-span of vectors (xi),then L∗ is the Z-span of a dual basis (yj) satisfying 〈xi, yj〉 = δij,cf. relation (17.4.1).

Page 169: Analytic and Differential geometry Mikhail G. Katz∗

16.2. LOEWNER’S TORUS INEQUALITY 169

Thus, the constant γ′b is bounded above by the Hermite constant γbof (16.1.8). We have γ′1 = 1, while for b ≥ 2 we have the followinginequality:

γ′b ≤ γb ≤2

3b for all b ≥ 2. (16.1.10)

Moreover, one has the following asymptotic estimates:

b

2πe(1 + o(1)) ≤ γ′b ≤

b

πe(1 + o(1)) for b→ ∞, (16.1.11)

cf. [LaLS90, pp. 334, 337]. Note that the lower bound of (16.1.11)for the Hermite constant and the Berge-Martinet constant is noncon-structive, but see [RT90] and [ConS99].

Definition 16.1.9. A lattice L realizing the supremum in (16.1.9)or (16.1.9) is called dual-critical.

Remark 16.1.10. The constants γ′b and the dual-critical latticesin Rb are explicitly known for b ≤ 4, cf. [BeM, Proposition 2.13]. Inparticular, we have γ′1 = 1, γ′2 =

2√3.

Example 16.1.11. In dimension 3, the value of the Berge-Martinet

constant, γ′3 =√

32= 1.2247 . . ., is slightly below the Hermite con-

stant γ3 = 213 = 1.2599 . . .. It is attained by the face-centered cu-

bic lattice, which is not isodual [MilH73, p. 31], [BeM, Proposition2.13(iii)], [CoS94].

This is the end of the three subsections containing optional material.

16.2. Loewner’s torus inequality

Historically, the first lower bound for the volume of a Riemannianmanifold in terms of a systole is due to Charles Loewner. In 1949,Loewner proved the first systolic inequality, in a course on Riemanniangeometry at Syracuse University, cf. [Pu52]. Namely, he showed thefollowing result, whose proof appears in Section 19.2.

Theorem 16.2.1 (C. Loewner). Every Riemannian metric g on thetorus T2 satisfies the inequality

sys1(g)2 ≤ γ2 area(g), (16.2.1)

where γ2 = 2√3is the Hermite constant (16.1.8). A metric attaining

the optimal bound (16.2.1) is necessarily flat, and is homothetic to thequotient of C by the Eisenstein integers, i.e. lattice spanned by the cuberoots of unity, cf. Lemma 15.9.1.

Page 170: Analytic and Differential geometry Mikhail G. Katz∗

170 16. SYSTOLE

The result can be reformulated in a number of ways.2 Loewner’storus inequality relates the total area, to the systole, i.e. least lengthof a noncontractible loop on the torus (T2, g):

area(g)−√32sys1(g)

2 ≥ 0. (16.2.2)

The boundary case of equality is attained if and only if the metric ishomothetic to the flat metric obtained as the quotient of R2 by thelattice formed by the Eisenstein integers.

16.3. Loewner’s inequality with defect

Loewner’s torus inequality can be strengthened by introducing a“defect” (shegiya, she’erit) term, similar to Bonnesen’s strengtheningof the isoperimetic inequality. To write it down, we need to review theconformal representation theorem (uniformisation theorem).

By the conformal representation theorem, we can assume that themetric g on the torus T2 is of the form

f 2(dx2 + dy2),

with respect to a unit area flat metric dx2+dy2 on the torus T2 viewedas a quotient R2/L of the (x, y) plane by a lattice. The defect termin question is simply the variance (shonut) of the conformal factor fabove. The inequality with the defect term looks as follows.

Theorem 16.3.1. Every metric on the torus satisfies the followingstrengthened form of Loewner’s inequality:

area(g)−√32sys(g)2 ≥ Var(f). (16.3.1)

Here the error term, or isosystolic defect, is given by the variance3

Var(f) =

T2

(f −m)2 (16.3.2)

of the conformal factor f of the metric g = f 2(dx2 + dy2) on the torus,relative to the unit area flat metric g0 = dx2+dy2 in the same conformalclass. Here

m =

T2

fdxdy (16.3.3)

2Thus, in the case of the torus T2, the fundamental group is abelian. Hencethe systole can be expressed in this case as follows: sys1(T

2) = λ1(H1 (T

2;Z), ‖ ‖),

where ‖ ‖ is the stable norm.3shonut

Page 171: Analytic and Differential geometry Mikhail G. Katz∗

16.5. AN APPLICATION OF THE COMPUTATIONAL FORMULA 171

is the mean4 of f . More concretely, if (T2, g0) = R2/L where L is alattice of unit coarea, and D is a fundamental domain for the actionof L on R2 by translations, then the integral (16.3.3) can be written as

m =

D

f(x, y)dxdy

where dxdy is the standard Lebesgue measure of R2.

16.4. Computational formula for the variance

The proof of inequalities with isosystolic defect relies upon the fa-miliar computational formula for the variance5 of a random variable6

X in terms of expected values.Namely, we have the formula

Eµ(X2)− (Eµ(X))2 = Var(X), (16.4.1)

where µ is a probability measure. Here the variance is

Var(X) = Eµ((X −m)2

),

where m = Eµ(X) is the expected value (i.e. the mean) (tochelet).

16.5. An application of the computational formula

Keeping our differential geometric application in mind, we will de-note the random variable (mishtaneh akra’i) f .

Now consider a flat metric g0 of unit area on the 2-torus T2. Denotethe associated measure by µ. In other words, µ is given by the usualLebesgue measure dxdy. Since µ is a probability measure, we can applyformula (16.4.1) to it. Consider a metric g = f 2g0 conformal to the flatone, with conformal factor f > 0. Then we have

Eµ(f2) =

T2

f 2dxdy = area(g),

since f 2dx dy is precisely the area element of the metric g.Equation (16.4.1) therefore becomes

area(g)− (Eµ(f))2 = Var(f). (16.5.1)

Next, we will relate the expected value Eµ(f) to the systole ofthe metric g. Then we will then relate (16.4.1) to Loewner’s torusinequality.

4tochelet (with “het”)5shonut6mishtaneh mikri

Page 172: Analytic and Differential geometry Mikhail G. Katz∗

172 16. SYSTOLE

By the uniformisation theorem, every metric on the torus T2 isconformally equivalent to a flat metric of unit area. Denoting such aflat metric by g0, such equivalence can be written as

g = f 2g0.

16.6. Conformal invariant σ

Let us prove Loewner’s torus inequality for the metric g = f 2g0on T2, using the computational formula for the variance. We firstanalyze the expected value term

Eµ(f) =

T2

fdxdy

in (16.5.1).By the proof of Lemma 15.9.1, the lattice of deck transformations

of the flat torus g0 admits a Z-basis similar (domeh) to {τ, 1} ⊆ C,where τ belongs to the standard fundamental domain (15.8.1). In otherwords, the lattice is similar to

Zτ + Z1 ⊆ C.

Definition 16.6.1. We define the conformal invariant σ as follows:consider the imaginary part Im(τ) and set

σ2 := Im(τ) > 0.

Lemma 16.6.2. We have σ2 ≥√32, with equality if and only if τ is

the primitive cube or sixth root of unity.

Proof. This is immediate from the geometry of the fundamentaldomain. �

16.7. Fundamental domain and Loewner’s torus inequality

Thinking of a unit area flat metric g0 as a lattice quotient R2/L, wecan write down the flat metric as dx2+dy2. We think of the conformalfactor f(x, y) > 0 of the metric g on T2 as a doubly periodic (i.e., L-periodic) function on R2.

Theorem 16.7.1. Let σ be the conformal invariant as above. Thenthe metric f 2(dx2 + dy2) satisfies

area(g)− σ2sys(g)2 ≥ Var(f). (16.7.1)

Proof. Since g0 is assumed to be of unit area, the basis for itsgroup of deck tranformations can therefore be taken to be the pair{

τ

σ,1

σ

}.

Page 173: Analytic and Differential geometry Mikhail G. Katz∗

16.8. BOUNDARY CASE OF EQUALITY 173

Here Im(τσ

)= σ. Thus the torus C/ Span

(τσ, 1σ

)has unit area:

area(C/ Span

(τσ, 1σ

))= 1.

With these normalisations, we see that the flat torus is ruled by a pencilof horizontal closed geodesics, denoted

γy = γy(x),

each of length σ−1, where the “width” of the pencil equals σ, i.e. theparameter y ranges through the interval [0, σ], with γσ = γ0. Note thateach of these closed geodesics is noncontractible in T2, and thereforeby definition

length(γy) ≥ sys1(g).

By Fubini’s theorem, we pass to the iterated integral (nishneh) to ob-tain the following lower bound for the expected value:

Eµ(f) =

∫ σ

0

(∫

γy

f(x)dx

)dy

=

∫ σ

0

length(γy)dy

≥ σsys(g).

See also [Ka07, p. 41, 44]. Substituting into (16.5.1), we obtain therequired inequality

area(g)− σ2sys(g)2 ≥ Var(f). (16.7.2)

where f is the conformal factor of the metric g with respect to the unitarea flat metric g0. �

Since we have in general σ2 ≥√32, we obtain in particular Loewner’s

torus inequality with isosystolic defect,

area(g)−√32sys(g)2 ≥ Var(f). (16.7.3)

cf. [Pu52].

16.8. Boundary case of equality

Corollary 16.8.1. A metric satisfying the boundary case of equal-ity in Loewner’s torus inequality is necessarily flat and homothetic tothe quotient of R2 by the lattice of Eisenstein integers.

Proof. If a metric f 2(dx2 + dy2) satisfies the boundary case of

equality area(g) −√32sys(g)2 = 0, then the variance of the conformal

factor f must vanish by (16.7.3). Hence f is a constant function. The

Page 174: Analytic and Differential geometry Mikhail G. Katz∗

174 16. SYSTOLE

proof is completed by applying Lemma 15.9.1 on the Hermite constantin dimension 2. �

Now suppose τ is pure imaginary, i.e. the lattice L is a rectangularlattice of coarea 1.

Corollary 16.8.2. If τ is pure imaginary, then the metric g =f 2g0 satisfies the inequality

area(g)− sys(g)2 ≥ Var(f). (16.8.1)

Proof. If τ is pure imaginary then σ =√

Im(τ) ≥ 1, and theinequality follows from (16.7.1). �

In particular, every surface of revolution satisfies (16.8.1), since itslattice is rectangular, cf. Corollary 15.10.1.

Page 175: Analytic and Differential geometry Mikhail G. Katz∗

CHAPTER 17

Manifolds and global geometry

17.1. Global geometry of surfaces

Discussion of Local versus Global: The local behavior is by def-inition the behavior in an open neighborhood of a point. The localbehavior of a smooth curve is well understood by the implicit functiontheorem. Namely, a smooth curve in the plane or in 3-space can bethought of as the graph of a smooth function. A curve in the plane islocally the graph of a scalar function. A curve in 3-space is locally thegraph of a vector-valued function.

Example 17.1.1. The unit circle in the plane can be defined im-plicitly by

x2 + y2 = 1,

or parametrically by

t 7→ (cos t, sin t).

Alternatively, it can be given locally as the graph of the function

f(x) =√1− x2.

Note that this presentation works only for points on the upper halfcir-cle. For points on the lower halfcircle we use the function

−√1− x2.

Both of these representations fail at the points (1, 0) and (−1, 0).To overcome this difficulty, we must work with y as the independent

variable, instead of x. Thus, we can parametrize a neighborhood of(1,0) by using the function

x = g(y) =√1− y2.

Example 17.1.2. The helix given in parametric form by

(x, y, z) = (cos t, sin t, t).

It can also be defined as the graph of the vector-valued function f(z),with values in the (x, y)-plane, where

(x(z), y(z)) = f(z) = (cos z, sin z).

175

Page 176: Analytic and Differential geometry Mikhail G. Katz∗

176 17. MANIFOLDS AND GLOBAL GEOMETRY

In this case the graph representation in fact works even globally.

Example 17.1.3. The unit sphere in 3-space can be representedlocally as the graph of the function of two variables

f(x, y) =√1− x2 − y2.

As above, concerning the local nature of the presentation necessitatesadditional functions to represent neighborhoods of points not in theopen northern hemisphere (this example is discussed in more detail inSection 17.3).

17.2. Definition of manifold

Motivated by the examples given in the previous section, we give ageneral definition as follows.

A manifold is defined as a subset of Euclidean space which is lo-cally a graph of a function, possibly vector-valued. This is the originaldefinition of Poincare who invented the notion (see Arnold [1, p. 234]).

Definition 17.2.1. By a 2-dimensional closed Riemannian mani-fold we mean a compact subset

M ⊆ Rn

such that in an open neighborhood of every point p ∈ M in Rn, thecompact subsetM can be represented as the graph of a suitable smoothvector-valued function of two variables.

Here the function has values in (n− 2)-dimensional vectors.The usual parametrisation of the graph can then be used to calcu-

late the coefficients gij (see Definition 10.7.3) of the first fundamentalform, as, for example, in Theorem 17.13.2 and Example 6.4.1. Thecollection of all such data is then denoted by the pair (M, g), where

g = (gij)

is referred to as the metric.

Remark 17.2.2. Differential geometers like the (M, g) notation, be-cause it helps separate the topology M from the geometry g. Strictlyspeaking, the notation is redundant, since the object g already incorpo-rates all the information, including the topology. However, geometershave found it useful to use g when one wants to emphasize the geome-try, and M when one wants to emphasize the topology.

Note that, as far as the intrinsic geometry of a Riemannian mani-fold is concerned, the imbedding in Rn referred to in Definition 17.2.1is irrelevant to a certain extent, all the more so since certain basicexamples, such as flat tori, are difficult to imbed in a transparent way.

Page 177: Analytic and Differential geometry Mikhail G. Katz∗

17.4. DUAL BASES 177

17.3. Sphere as a manifold

The round 2-sphere S2 ⊆ R3 defined by the equation

x2 + y2 + z2 = 1

is a closed Riemannian manifold. Indeed, consider the function f(x, y)

defined in the unit disk x2 + y2 < 1 by setting f(x, y) =√

1− x2 − y2.Define a coordinate chart

x1(u1, u2) = (u1, u2, f(u1, u2)).

Thus, each point of the open northern hemisphere admits a neighbor-hood diffeomorphic to a ball (and hence to R2). To cover the southernhemisphere, use the chart

x2(u1, u2) = (u1, u2,−f(u1, u2)).

To cover the points on the equator, use in addition charts x3(u1, u2) =

(u1, f(u1, u2), u2), x4(u1, u2) = (u1,−f(u1, u2), u2), as well as the pair of

charts x5(u1, u2) = (f(u1, u2), u1, u2), x6(u

1, u2) = (−f(u1, u2), u1, u2).

17.4. Dual bases

Tangent space, cotangent space, and the notation for bases in thesespaces were discussed in Section 13.3.

We will work with dual bases(∂∂ui

)for vectors, and (dui) for cov-

ectors (i.e. elements of the dual space), such that

dui(

∂uj

)= δij , (17.4.1)

where δij is the Kronecker delta.Recall that the metric coefficients are defined by setting

gij = 〈xi, xj〉,where x is the parametrisation of the surface. We will only work withmetrics whose first fundamental form is diagonal. We can thus writethe first fundamental form as follows:

g = g11(u1, u2)(du1)2 + g22(u

1, u2)(du2)2. (17.4.2)

Remark 17.4.1. Such data can be computed from a Euclideanimbedding as usual, or it can be given apriori without an imbedding,as we did in the case of the hyperlobic metric.

We will work with such data independently of any Euclidean imbed-ding, as discussed in Section 10.6. For example, if the metric coefficientsform an identity matrix, we obtain

g =(du1)2

+(du2)2, (17.4.3)

Page 178: Analytic and Differential geometry Mikhail G. Katz∗

178 17. MANIFOLDS AND GLOBAL GEOMETRY

where the interior superscript denotes an index, while exterior super-script denotes the squaring operation.

17.5. Jacobian matrix

The Jacobian matrix of v = v(u) is the matrix

∂(v1, v2)

∂(u1, u2),

which is the matrix of partial derivatives. Denote by

Jacv(u) = det

(∂(v1, v2)

∂(u1, u2)

).

It is shown in advanced calculus that for any function f(v) in a do-main D, one has

D

f(v)dv1dv2 =

D

g(u)Jacv(u)du1du2, (17.5.1)

where g(u) = f(v(u)).

Example 17.5.1. Let u1 = r, u2 = θ. Let v1 = x, v2 = y. Wehave x = r cos θ and y = r sin θ. One easily shows that the Jacobianis Jacv(u) = r. The area elements are related by

dv1dv2 = Jacv(u)du1du2,

or

dxdy = rdrdθ.

A similar relation holds for integrals.

17.6. Area of a surface, independence of partition

Partition1 is what allows us to perform actual calculations witharea, but the result is independent of partition (see below).

Based on the local definition of area discussed in an earlier chapter,we will now deal with the corresponding global invariant.

Definition 17.6.1. The area element dA of the surface is the ele-ment

dA :=√det(gij)du

1du2,

where det(gij) = g11g22 − g212 as usual.

1ritzuf or chaluka?

Page 179: Analytic and Differential geometry Mikhail G. Katz∗

17.6. AREA OF A SURFACE, INDEPENDENCE OF PARTITION 179

Theorem 17.6.2. Define the area of (M, g) by means of the formula

area =

M

dA =∑

{U}

U

√det(gij)du

1du2, (17.6.1)

namely by choosing a partition {U} of M subordinate to a finite opencover as in Definition 17.2.1, performing a separate integration in eachopen set, and summing the resulting areas. Then the total area is in-dependent of the partition and choice of coordinates.

Proof. Consider a change from a coordinate chart (ui) to anothercoordinate chart, denoted (vα). In the overlap of the two domains, thecoordinates can be expressed in terms of each other, e.g. v = v(u), andwe have the 2 by 2 Jacobian matrix Jacv(u).

Denote by gαβ the metric coefficients with respect to the chart (vα).Thus, in the case of a metric induced by a Euclidean imbedding definedby x = x(u) = x(u1, u2), we obtain a new parametrisation

y(v) = x(u(v)).

Then we have

gαβ =

⟨∂y

∂vα,∂y

∂vβ

=

⟨∂x

∂ui∂ui

∂vα,∂x

∂uj∂uj

∂vβ

=∂ui

∂vα∂uj

∂vβ

⟨∂x

∂ui,∂x

∂uj

= gij∂ui

∂vα∂uj

∂vβ.

The right hand side is a product of three square matrices:

∂ui

∂vαgij

∂uj

∂vβ.

The matrices on the left and on the right are both Jacobian matrices.Since determinant is multiplicative, we obtain

det(gαβ) = det(gij)det

(∂(u1, u2)

∂(v1, v2)

)2

.

Page 180: Analytic and Differential geometry Mikhail G. Katz∗

180 17. MANIFOLDS AND GLOBAL GEOMETRY

Hence using equation (17.5.1), we can write the area element as

dA = det12 (gαβ)dv

1dv2

= det12 (gαβ) Jacv(u)du

1du2

= det12 (gij) det

(∂(u1, u2)

∂(v1, v2)

)Jacv(u)du

1du2

= det12 (gij) du

1du2

since inverse maps have reciprocal Jacobians by chain rule. Thus theintegrand is unchanged and the area element is well defined. �

17.7. Conformal equivalence

Definition 17.7.1. Two metrics, g = gijduiduj and h = hijdu

iduj,onM are called conformally equivalent , or conformal for short, if thereexists a function f = f(u1, u2) > 0 such that

g = f 2h,

in other words,gij = f 2hij ∀i, j. (17.7.1)

Definition 17.7.2. The function f above is called the conformalfactor (note that sometimes it is more convenient to refer, instead, tothe function λ = f 2 as the conformal factor).

Theorem 17.7.3. Note that the length of every vector at a givenpoint (u1, u2) is multiplied precisely by f(u1, u2).

Proof. More speficially, a vector v = vi ∂∂ui

which is a unit vectorfor the metric h, is “stretched” by a factor of f , i.e. its length withrespect to g equals f . Indeed, the new length of v is

√g(v, v) = g

(vi

∂ui, vj

∂uj

) 12

=

(g

(∂

∂ui,∂

∂uj

)vivj

) 12

=√gijvivj

=√f 2hijvivj

= f√hijvivj

= f,

proving the theorem. �

Page 181: Analytic and Differential geometry Mikhail G. Katz∗

17.8. GEODESIC EQUATION 181

Definition 17.7.4. An equivalence class of metrics on M confor-mal to each other is called a conformal structure on M (mivneh con-formi).

17.8. Geodesic equation

The material in this section has already been dealt with in an earlierchapter.

Perhaps the simplest possible definition of a geodesic β on a surfacein 3-space is in terms of the orthogonality of its second derivative β ′′ tothe surface. The nonlinear second order ordinary differential equationdefining a geodesic is, of course, the “true” if complicated definition.We will now prove the equivalence of the two definitions. Consider aplane curve

Rs

−→α

R2

(u1,u2)

where α = (α1(s), α2(s)). Let x : R2 → R3 be a regular parametrisationof a surface in 3-space. Then the composition

Rs−→α

R2

(u1,u2)−→x

R3

yields a curveβ = x ◦ α.

Definition 17.8.1. A curve β = x◦α is a geodesic on the surface xif one of the following two equivalent conditions is satisfied:

(a) we have for each k = 1, 2,

(αk)′′

+ Γkij(αi)

(αj)′

= 0 where′

=d

ds, (17.8.1)

meaning that

(∀k) d2αk

ds2+ Γkij

dαi

ds

dαj

ds= 0;

(b) the vector β ′′ is perpendicular to the surface and one has

β ′′ = Lijαi′αj

′n. (17.8.2)

To prove the equivalence, we write β = x ◦ α, then β ′ = xiαi′ by

chain rule. Furthermore,

β ′′ =d

ds(xi ◦ α)αi′ + xiα

i′′ = xijαj ′αi

′+ xkα

k′′.

Since xij = Γkijxk + Lijn holds, we have

β ′′ − Lijαi′αj

′n = xk

(αk

′′+ Γkijα

i′αj′).

Page 182: Analytic and Differential geometry Mikhail G. Katz∗

182 17. MANIFOLDS AND GLOBAL GEOMETRY

17.9. Closed geodesic

Definition 17.9.1. A closed geodesic in a Riemannian 2-manifoldMis defined equivalently as

(1) a periodic curve β : R → M satisfying the geodesic equa-

tion αk′′+ Γkijα

i′αj′= 0 in every chart x : R2 →M , where, as

usual, β = x◦α and α(s) = (α1(s), α2(s)) where s is arclength.Namely, there exists a period T > 0 such that β(s+T ) = β(s)for all s.

(2) A unit speed map from a circle R/LT → M satisfying thegeodesic equation at each point, where LT = TZ ⊆ R is therank one lattice generated by T > 0.

Definition 17.9.2. The length L(β) of a path β : [a, b] → M iscalculated using the formula

L(β) =

∫ b

a

‖β ′(t)‖dt,

where ‖v‖ =√gijvivj whenever v = vixi. The energy is defined

by E(β) =∫ ba‖β ′(t)‖2dt.

A closed geodesic as in Definition 17.9.1, item 2 has length T .

Remark 17.9.3. The geodesic equation (17.8.1) is the Euler-Lagrangeequation of the first variation of arc length. Therefore when a pathminimizes arc length among all neighboring paths connecting two fixedpoints, it must be a geodesic. A corresponding statement is valid forclosed loops, cf. proof of Theorem 17.10.1. See also Section 16.1. InSections 17.11 and 17.15 we will give a complete description of thegeodesics for the constant curvature sphere, as well as for flat tori.

17.10. Existence of closed geodesic

Theorem 17.10.1. Every free homotopy class of loops in a closedmanifold contains a closed geodesic.

Proof. We sketch a proof for the benefit of a curious reader, whocan also check that the construction is independent of the choices in-volved. The relevant topological notions are defined in Section 17.18and [Hat02]. A free homotopy class α of a manifold M correspondsto a conjugacy class gα ⊂ π1(M). Pick an element g ∈ gα. Thus g actson the universal cover M of M . Let fg : M → R be the displacementfunction of g, i.e.

fg(x) = d(x, g.x).

Page 183: Analytic and Differential geometry Mikhail G. Katz∗

17.12. REAL PROJECTIVE PLANE 183

Let x0 ∈M be a minimum of fg. A first variation argument shows thatany length-minimizing path between x0 and g.x0 descends to a closedgeodesic in M representing α, cf. [Car92, Ch93, GaHL04]. �

17.11. Surfaces of constant curvature

By the uniformisation theorem 13.8.2, all surfaces fall into threetypes, according to whether they are conformally equivalent to metricsthat are:

(1) flat (i.e. have zero Gaussian curvature K ≡ 0);(2) spherical (K ≡ +1);(3) hyperbolic (K ≡ −1).

For closed surfaces, the sign of the Gaussian curvature K is that ofits Euler characteristic, cf. formula (17.26.1).

Theorem 17.11.1 (Constant positive curvature). There are onlytwo compact surfaces, up to isometry, of constant Gaussian curva-ture K = +1. They are the round sphere S2 of Example 17.3; andthe real projective plane, denoted RP2.

17.12. Real projective plane

Intuitively, one thinks of the real projective plane as the quotientsurface obtained if one starts with the northern hemisphere of the 2-sphere, and “glues” together pairs of opposite points of the equatorialcircle (the boundary of the hemisphere).

More formally, the real projective plane can be defined as follows.Let

m : S2 → S2

denote the antipodal map of the sphere, i.e. the restriction of the map

v 7→ −vin R3. Then m is an involution. In other words, if we consider theaction of the group Z2 = {e,m} on the sphere, each orbit of the Z2

action on S2 consists of a pair of antipodal points

{±p} ⊆ S2. (17.12.1)

Definition 17.12.1. On the set-theoretic level, the real projectiveplane RP2 is the set of orbits of type (17.12.1), i.e. the quotient of S2

by the Z2 action.

Denote by Q : S2 → RP2 the quotient map. The smooth structureand metric on RP2 are induced from S2 in the following sense. Let x :

Page 184: Analytic and Differential geometry Mikhail G. Katz∗

184 17. MANIFOLDS AND GLOBAL GEOMETRY

R2 → S2 be a chart on S2 not containing any pair of antipodal points.Let gij be the metric coefficients with respect to this chart.

Let y = m◦x = −x denote the “opposite” chart, and denote by hijits metric coefficients. Then

hij = 〈∂y

∂ui,∂y

∂uj〉 = 〈− ∂x

∂ui,− ∂x

∂uj〉 = 〈 ∂x

∂ui,∂x

∂uj〉 = gij. (17.12.2)

Thus the opposite chart defines the identical metric coefficients. Thecomposition Q ◦ x is a chart on RP2, and the same functions gij formthe metric coefficients for RP2 relative to this chart.

We can summarize the preceeding discussion by means of the fol-lowing definition.

Definition 17.12.2. The real projective plane RP2 is defined inthe following two equivalent ways:

(1) the quotient of the round sphere S2 by (the restriction to S2

of) the antipodal map v 7→ −v in R3. In other words, a typicalpoint of RP2 can be thought of as a pair of opposite points ofthe round sphere.

(2) the northern hemisphere of S2, with opposite points of theequator identified.

The smooth structure of RP2 is induced from the round sphere.Since the antipodal map preserves the metric coefficients by the cal-culation (17.12.2), the metric structure of constant Gaussian curva-ture K = +1 descends to RP2, as well.

17.13. Simple loops for surfaces of positive curvature

Definition 17.13.1. A loop α : S1 → X of a space X is calledsimple if the map α is one-to-one, cf. Definition 17.18.1.

Theorem 17.13.2. The basic properties of the geodesics on surfacesof constant positive curvature as as follows:

(1) all geodesics are closed;(2) the simple closed geodesics on S2 have length 2π and are de-

fined by the great circles;(3) the simple closed geodesics on RP2 have length π;(4) the simple closed geodesics of RP2 are parametrized by half-

great circles on the sphere.

Proof. We calculate the length of the equator of S2. Here thesphere ρ = 1 is parametrized by

x(θ, ϕ) = (sinϕ cos θ, sinϕ sin θ, cosϕ)

Page 185: Analytic and Differential geometry Mikhail G. Katz∗

17.15. FLAT SURFACES 185

in spherical coordinates (θ, ϕ). The equator is the curve x◦α where α(s) =(s, π

2) with s ∈ [0, 2π]. Thus α1(s) = θ(s) = s. Recall that the metric

coefficients are given by g11(θ, ϕ) = sin2 ϕ, while g22 = 1 and g12 = 0.Thus

‖β ′(s)‖ =√gij(s, π

2

)αi′αj ′ =

√(sin π

2

) (dθds

)2= 1.

Thus the length of β is∫ 2π

0

‖β ′(s)‖ds =∫ 2π

0

1ds = 2π.

A geodesic on RP2 is twice as short as on S2, since the antipodal pointsare identified, and therefore the geodesic “closes up” sooner than (i.e.twice as fast as) on the sphere. For example, a longitude of S2 isnot a closed curve, but it descends to a closed curve on RP2, sinceits endpoints (north and south poles) are antipodal, and are thereforeidentified with each other. �

17.14. Successive minima

The material in this section has already been dealt with in an earlierchapter.

Let B be a finite-dimensional Banach space, i.e. a vector space to-gether with a norm ‖ ‖. Let L ⊆ (B, ‖ ‖) be a lattice of maximal rank,i.e. satisfying rank(L) = dim(B). We define the notion of successiveminima of L as follows, cf. [GruL87, p. 58].

Definition 17.14.1. For each k = 1, 2, . . . , rank(L), define the k-thsuccessive minimum λk of the lattice L by

λk(L, ‖ ‖) = inf

{λ ∈ R

∣∣∣∣∃ lin. indep. v1, . . . , vk ∈ Lwith ‖vi‖ ≤ λ for all i

}. (17.14.1)

Thus the first successive minimum, λ1(L, ‖ ‖) is the least length ofa nonzero vector in L.

17.15. Flat surfaces

A metric is called flat if its Gaussian curvature K vanishes at everypoint.

Theorem 17.15.1. A closed surface of constant Gaussian curva-ture K = 0 is topologically either a torus T2 or a Klein bottle.

Let us give a precise description in the former case.

Page 186: Analytic and Differential geometry Mikhail G. Katz∗

186 17. MANIFOLDS AND GLOBAL GEOMETRY

Example 17.15.2 (Flat tori). Every flat torus is isometric to aquotient T2 = R2/L where L is a lattice, cf. [Lo71, Theorem 38.2].In other words, a point of the torus is a coset of the additive ac-tion of the lattice in R2. The smooth structure is inherited from R2.Meanwhile, the additive action of the lattice is isometric. Indeed, wehave dist(p, q) = ‖q − p‖, while for any ℓ ∈ L, we have

dist(p+ ℓ, q + ℓ) = ‖q + ℓ− (p+ ℓ)‖ = ‖q − p‖ = dist(p, q).

Therefore the flat metric on R2 descends to T2.

Note that locally, all flat tori are indistinguishable from the flatplane itself. However, their global geometry depends on the metric in-variants of the lattice, e.g. its successive minima, cf. Definition 17.14.1.Thus, we have the following.

Theorem 17.15.3. The least length of a nontrivial closed geodesicon a flat torus T2 = R2/L equals the first successive minimum λ1(L).

Proof. The geodesics on the torus are the projections of straightlines in R2. In order for a straight line to close up, it must pass througha pair of points x and x+ℓ where ℓ ∈ L. The length of the correspondingclosed geodesic on T2 is precisely ‖ℓ‖, where ‖ ‖ is the Euclidean norm.By choosing a shortest element in the lattice, we obtain a shortestclosed geodesic on the corresponding torus. �

17.16. Hyperbolic surfaces

Most closed surfaces admit neither flat metrics nor metrics of pos-itive curvature, but rather hyperbolic metrics. A hyperbolic surface isa surface equipped with a metric of constant Gaussian curvature K =−1. This case is far richer than the other two.

Example 17.16.1. The pseudosphere (so called because its Gauss-ian curvature is constant, and equals −1) is the surface of revolution

(f(ϕ) cos θ, f(ϕ) sin θ, g(ϕ))

in R3 defined by the functions f(φ) = eφ and

g(φ) =

∫ φ

0

(1− e2ψ

)1/2dψ,

where φ ranges through the interval −∞ < φ ≤ 0. The usual formu-

las g11 = f 2 as well as g22 =(drdφ

)2+(dgdφ

)2yield in our case g11 = e2φ,

Page 187: Analytic and Differential geometry Mikhail G. Katz∗

17.17. HYPERBOLIC PLANE 187

while

g22 =(eφ)2 +

(√1− e2φ

)2

= e2φ + 1− e2φ

= 1.

Thus (gij) =

(e2φ 00 1

). The pseudosphere has constant Gaussian

curvature −1, but it is not a closed surface (as it is unbounded in R3).

17.17. Hyperbolic plane

The metric

gH2 =1

y2(dx2 + dy2) (17.17.1)

in the upperhalf plane

H2 = {(x, y) | y > 0}is called the hyperbolic metric of the upper half plane.

Theorem 17.17.1. The metric (17.17.1) has constant Gaussiancurvature K = −1.

Proof. By Theorem 12.3.3, we have

K = −∆LB log f = ∆LB log y = y2(− 1

y2

)= −1,

as required. �

In coordinates (u1, u2), we can write it, a bit awkwardly, as

gH2 =1

(u2)2

((du1)2

+(du2)2)

.

The Riemannian manifold (H2, gH2) is referred to as the Poincare up-perhalf plane. Its significance resides in the following theorem.

Theorem 17.17.2. Every closed hyperbolic surface M is isomet-ric to the quotient of the Poincare upperhalf plane by the action of asuitable group Γ:

M = H2/Γ.

Here the nonabelian group Γ is a discrete subgroup Γ ⊆ PSL(2,R),

where a matrix A =

(a bc d

)acts on H2 = {z ∈ C|ℑ(z) > 0} by

z 7→ az + b

cz + d,

Page 188: Analytic and Differential geometry Mikhail G. Katz∗

188 17. MANIFOLDS AND GLOBAL GEOMETRY

called fractional linear transformations, of Mobius transformations. Allsuch transformations are isometries of the hyperbolic metric. The fol-lowing theorem is proved, for example, in [Kato92].

Theorem 17.17.3. Every geodesic in the Poincare upperhalf planeis either a vertical ray, or a semicircle perpendicular to the x-axis.

The foundational significance of this model in the context of theparallel postulate of Euclid has been discussed by numerous authors.

Example 17.17.4. The length of a vertical interval joining i to cican be calculated as follows. Recall that the conformal factor is f(x, y) =1y. The length is therefore given by

∣∣∣∣∫ c

1

1

ydy

∣∣∣∣ = | log c|.

Here the substitution y = es gives an arclength parametrisation.

17.18. Loops, simply connected spaces

We would like to provide a self-contained explanation of the topo-logical ingredient which is necessary so as to understand Loewner’storus inequality, i.e. essentially the notion of a noncontractible loopand the fundamental group of a topological space X . See [Hat02,Chapter 1] for a more detailed account.

Definition 17.18.1. A loop in X can be defined in one of twoequivalent ways:

(1) a continuous map β : [a, b] → X satisfying β(a) = β(b);(2) a continuous map λ : S1 → X from the circle S1 to X .

Lemma 17.18.2. The two definitions of a loop are equivalent.

Proof. Consider the unique increasing linear function

f : [a, b] → [0, 2π]

which is one-to-one and onto. Thus, f(t) = 2π(t−a)b−a . Given a map

λ(eis) : S1 → X,

we associate to it a map β(t) = λ(eif(t)

), and vice versa. �

Definition 17.18.3. A loop S1 → X is said to be contractible ifthe map of the circle can be extended to a continuous map of the unitdisk D → X , where S1 = ∂D.

Definition 17.18.4. A space X is called simply connected if everyloop in X is contractible.

Page 189: Analytic and Differential geometry Mikhail G. Katz∗

17.20. CYCLES AND BOUNDARIES 189

Theorem 17.18.5. The sphere Sn ⊆ Rn+1 which is the solution setof x20+ . . .+x2n = 1 is simply connected for n ≥ 2. The circle S1 is notsimply connected.

17.19. Orientation on loops and surfaces

Let S1 ⊆ C be the unit circle. The choice of an orientation on thecircle is an arrow pointing clockwise or counterclockwise. The standardchoice is to consider S1 as an oriented manifold with orientation chosencounterclockwise.

If a surface is imbedded in 3-space, one can choose a continuousunit normal vector n at every point. Then an orientation is defined bythe right hand rule with respect to n thought of as the thumb (agudal).

17.20. Cycles and boundaries

The singular homology groups with integer coefficients, Hk(M ;Z)for k = 0, 1, . . . ofM are abelian groups which are homotopy invariantsof M . Developing the singular homology theory is time-consuming.The case that we will be primarily interested in as far as these notesare concerned, is that of the 1-dimensional homology group:

H1(M ;Z).

In this case, the homology groups can be characterized easily withoutthe general machinery of singular simplices.

Let S1 ⊆ C be the unit circle, which we think of as a 1-dimensionalmanifold with an orientation given by the counterclockwise direction.

Definition 17.20.1. A 1-cycle α on a manifold M is an integerlinear combination

α =∑

i

nifi

where ni ∈ Z is called the multiplicity (ribui), while each

fi : S1 →M

is a loop given by a smooth map from the circle to M , and each loopis endowed with the orientation coming from S1.

Definition 17.20.2. The space of 1-cycles on M is denoted

Z1(M ;Z).

Let (Σg, ∂Σg) be a surface with boundary ∂Σg , where the genus gis irrelevant for the moment and is only added so as to avoid confusionwith the summation symbol

∑.

Page 190: Analytic and Differential geometry Mikhail G. Katz∗

190 17. MANIFOLDS AND GLOBAL GEOMETRY

The boundary ∂Σg is a disjoint union of circles. Now assume thesurface Σg is oriented.

Proposition 17.20.3. The orientation of the surface induces anorientation on each boundary circle.

Thus we obtain an orientation-preserving identification of each bound-ary component with the standard unit circle S1 ⊆ C (with its counter-clockwise orientation).

Given a map Σg → M , its restriction to the boundary thereforeproduces a 1-cycle

∂Σg ∈ Z1(M ;Z).

Definition 17.20.4. The space

B1(M ;Z) ⊆ Z1(M ;Z)

of 1-boundaries in M is the space of all cycles∑

i

nifi ∈ Z1(M ;Z)

such that there exists a map of an oriented surface Σg →M (for some g)satisfying

∂Σg =∑

i

nifi.

Example 17.20.5. Consider the cylinder

x2 + y2 = 1, 0 ≤ z ≤ 1

of unit height. The two boundary components correspond to the twocircles: the “bottom” circle Cbottom defined by z = 0, and the “top”circle Ctop defined by and z = 1. Consider the orientation on thecylinder defined by the outward pointing normal vector. It induces thecounterclockwise orientation on Cbottom, and a clockwise orientationon Ctop.

Now let C0 and C1 be the same circles with the following choice oforientation: we choose a standard counterclockwise parametrisation onboth circles, i.e., parametrize them by means of (cos θ, sin θ). Then theboundary of the cylinder is the difference of the two circles: C0 − C1,or C1 − C0, depending on the choice of orientation.

Example 17.20.6. Cutting up a circle of genus 2 into two once-holed tori shows that the separating curve is a 1-boundary.

Theorem 17.20.7. On a closed orientable surface, a separatingcurve is a boundary, while a non-separating loop is never a boundary.

Page 191: Analytic and Differential geometry Mikhail G. Katz∗

17.22. STABLE NORM IN 1-DIMENSIONAL HOMOLOGY 191

17.21. First singular homology group

Definition 17.21.1. The 1-dimensional homology group ofM withinteger coefficients is the quotient group

H1(M ;Z) = Z1(M ;Z)/B1(M ;Z).

Definition 17.21.2. Given a cycle C ∈ Z1(M ;Z), its homologyclass will be denoted [C] ∈ H1(M ;Z).

Example 17.21.3. A non-separating loop on a closed surface rep-resents a non-trivial homology class of the surface.

Theorem 17.21.4. The 1-dimensional homology group H1(M ;Z)is the abelianisation of the fundamental group π1(M):

H1(M ;Z) = (π1M)ab .

Note that a significant difference between the fundamental groupand the first homology group is the following. While only based loopsparticipate in the definition of the fundamental group, the definitionof H1(M ;Z) involves free (not based) loops.

Example 17.21.5. The fundamental groups of the real projectiveplane RP2 and the 2-torus T2 are already abelian. Therefore one ob-tains

H1(RP2;Z) = Z/2Z,and

H1(T2;Z) = Z2.

Example 17.21.6. The fundamental group of an orientable closedsurface Σg of genus g is known to be a group on 2g generators with asingle relation which is a product of g commutators. Therefore one has

H1(Σg;Z) = Z2g.

17.22. Stable norm in 1-dimensional homology

Assume the manifoldM has a Riemannian metric. Given a smoothloop f : S1 → M , we can measure its volume (length) with respect tothe metric of M . We will denote this length by

vol(f)

with a view to higher-dimensional generalisation.

Definition 17.22.1. The volume (length) of a 1-cycle C =∑

i nifiis defined as

vol(C) =∑

i

|ni| vol(fi).

Page 192: Analytic and Differential geometry Mikhail G. Katz∗

192 17. MANIFOLDS AND GLOBAL GEOMETRY

Definition 17.22.2. Let α ∈ H1(M ;Z) be a 1-dimensional homol-ogy class. We define the volume of α as the infimum of volumes ofrepresentative 1-cycles:

vol(α) = inf {vol(C) | C ∈ α} ,where the infimum is over all cycles C =

∑i nifi representing the

class α ∈ H1(M ;Z).

The following phenomenon occurs for orientable surfaces.

Theorem 17.22.3. LetM be an orientable surface, i.e. 2-dimensionalmanifold. Let α ∈ H1(M ;Z). For all j ∈ N, we have

vol(jα) = j vol(α), (17.22.1)

where jα denotes the class α+ α + . . .+ α, with j summands.

Proof. To fix ideas, let j = 2. By Lemma 17.22.4 below, a mini-mizing loop C representing a multiple class 2α will necessarily intersectitself in a suitable point p. Then the 1-cycle represented by C can bedecomposed into the sum of two 1-cycles (where each can be thoughtof as a loop based at p). The shorter of the two will then give a min-imizing loop in the class α which proves the identity (17.22.1) in thiscase. The general case follows similarly. �

Lemma 17.22.4. A loop going around a cylinder twice necessarilyhas a point of self-intersection.

Proof. We think of the loop as the graph of a 4π-periodic func-tion f(t) (or alternatively a function on [0, 4π] with equal values at theendpoints). Consider the difference g(t) = f(t) − f(t + 2π). Then gtakes both positive and negative values. By the intermediate functiontheorem, the function g must have a zero t0. Then f(t0) = f(t0 + 2π)hence t0 is a point of self-intersection of the loop. �

17.23. The degree of a map

An example of a degree d map is most easily produced in the caseof a circle. A self-map of a circle given by

eiθ 7→ eidθ

has degree d.We will discuss the degree in the context of surfaces only.

Definition 17.23.1. The degree

df

Page 193: Analytic and Differential geometry Mikhail G. Katz∗

17.24. DEGREE OF NORMAL MAP OF AN IMBEDDED SURFACE 193

of a map f between closed surfaces is the algebraic number of pointsin the inverse image of a generic point of the target surface.

We can use the 2-dimensional homology groups defined elsewherein these notes, so as to calculate the degree as follows. Recall that

H2(M ;Z) = Z,

where the generator is represented by the identity self-map of the sur-face.

Theorem 17.23.2. A map

f :M1 →M2

induces a homomorphism

f∗ : H2(M1;Z) → H2(M2;Z),

corresponding to multiplication by the degree df once the groups areidentified with Z:

Z → Z, n 7→ df n.

17.24. Degree of normal map of an imbedded surface

Theorem 17.24.1. Let Σ ⊆ R3 be an imbedded surface. Let p beits genus. Consider the normal map

fn : Σ → S2

defined by sending each point x ∈ Σ to the normal vector n = nx at x.Then the degree of the normal map is precisely 1− p.

Example 17.24.2. For the unit sphere, the normal map is theidentity map. The genus is 0, while the degree of the normal mapis 1− p = 1.

Example 17.24.3. For the torus, the normal map is harder to vi-sualize. The genus is 1, while the degree of the normal map is 0.

Example 17.24.4. For a genus 2 surface imbedded in R3, the degreeof the normal map is 1 − 2 = −1. This means that if the surface isoriented by the outward-pointing normal vector, the normal map isorientation-reversing.

Page 194: Analytic and Differential geometry Mikhail G. Katz∗

194 17. MANIFOLDS AND GLOBAL GEOMETRY

17.25. Euler characteristic of an orientable surface

The Euler characteristic χ(M) is even for closed orientable surfaces,and the integer p = p(M) ≥ 0 defined by

χ(M) = 2− 2p

is called the genus of M .

Example 17.25.1. We have p(S2) = 0, while p(T2) = 1.

In general, the genus can be understood intuitively as the number of“holes”, i.e. “handles”, in a familiar 3-dimensional picture of a pretzel.We see from formula (17.26.1) that the only compact orientable surfaceadmitting flat metrics is the 2-torus. See [Ar83] for a friendly topo-logical introduction to surfaces, and [Hat02] for a general definition ofthe Euler characteristic.

17.26. Gauss-Bonnet theorem

Every imbedded closed surface in 3-space admits a continuous choiceof a unit normal vector n = nx at every point x. Note that no suchchoice is possible for an imbedding of the Mobius band, see [Ar83] formore details on orientability and imbeddings.

Closed imbedded surfaces in R3 are called orientable.

Remark 17.26.1. The integrals of type∫

M

will be understood in the sense of Theorem 17.6.2, namely using animplied partition subordinate to an atlas, and calculating the integralusing coordinates (u1, u2) in each chart, so that we can express themetric in terms of metric coefficients

gij = gij(u1, u2)

and similarly the Gaussian curvature

K = K(u1, u2).

Theorem 17.26.2 (Gauss-Bonnet theorem). Every closed surfaceMsatisfies the identity

M

K(u1, u2)√

det(gij)du1du2 = 2πχ(M), (17.26.1)

where K is the Gaussian curvature function on M , whereas χ(M) isits Euler characteristic.

Page 195: Analytic and Differential geometry Mikhail G. Katz∗

17.28. GAUSS MAP 195

17.27. Change of metric exploiting Gaussian curvature

We will use the term pseudometric for a quadratic form (or theassociated bilinear form), possibly degenerate.

We would like to give an indication of a proof of the Gauss-Bonnettheorem. We will have to avoid discussing some technical points. Con-sider the normal map

F :M → S2, x 7→ nx.

Consider a neigborhood in M where the map F is a homeomorphism(this is not always possible, and is one of the technical points we areavoiding).

Definition 17.27.1. Let gM the metric of M , and h the standardmetric of S2.

Given a point x ∈M in such a neighborhood, we can calculate thecurvature K(x). We can then consider a new metric in the conformalclass of the metric gM , defined as follows.

Definition 17.27.2. We define a new pseudometric, denoted gM ,on M by multiplying by the conformal factor K(x) at the point x.Namely, gM is the pseudometric which at the point x is given by thequadratic form

gx = K(x)gx.

If K ≥ 0 then the length of vectors is multiplied by√K.

The key to understanding the proof of the Gauss-Bonnet theoremin the case of imbedded surfaces is the following theorem.

Theorem 17.27.3. Consider the restriction of the normal map Fto a neighborhood as above. We modify the metric on the source M bythe conformal factor given by the Gaussian curvature, as above. Thenthe map

F : (M, gM) → (S2, h)

preserves areas: the area of the neighborhood in M (with respect to themodified metric) equals to the “area” of its image on the sphere.

17.28. Gauss map

Definition 17.28.1. The Gauss map is the map

F :M → S2, p 7→ Np

defined by sending a point p of M the unit normal vector N = Np

thought of as a point of S2.

Page 196: Analytic and Differential geometry Mikhail G. Katz∗

196 17. MANIFOLDS AND GLOBAL GEOMETRY

The map F sends an infinitesimal parallelogram on the surface, toan infinitesimal parallelogram on the sphere.

We may identify the tangent space to M at p and the tangentspace to S2 at F (p) ∈ S2. Then the differential of the map F is theWeingarten map

W : TpM → TF (p)S2.

The element of area KdA of the surface is mapped to the element ofarea of the sphere. In other words, we modify the element of areaby multiplying by the determinant (Jacobian) of the Weingarten map,namely the Gaussian curvature K(p). Hence the image of the areaelement KdA is precisely area 2-form h on the sphere, as discussed inthe previous section.

It remains to be checked that the map has topological degree givenby half the Euler characteristic of the surface M , proving the theoremin the case of imbedded surfaces. Since degree is invariant under contin-uous deformations, the result can be checked for a particular standardimbedding of a surface of arbitrary genus in R3.

17.29. An identity

Another way of writing identity (17.22.1) is as follows:

vol(α) =1

jvol(jα).

This phenomenon is no longer true for higher-dimensional mani-folds. Namely, the volume of a homology class is no longer multiplica-tive. However, the limit as j → ∞ exists and is called the stable norm.

Definition 17.29.1. Let M be a compact manifold of arbitrarydimension. The stable norm ‖ ‖ of a class α ∈ H1(M ;Z) is the limit

‖α‖ = limj→∞

1

jvol(jα). (17.29.1)

It is obvious from the definition that one has ‖α‖ ≤ vol(α). How-ever, the inequality may be strict in general. As noted above, for2-dimensional manifolds we have ‖α‖ = vol(α).

Proposition 17.29.2. The stable norm vanishes for a class of finiteorder.

Proof. If α ∈ H1(M,Z) is a class of finite order, one has finitelymany possibilities for vol(jα) as j varies. The factor of 1

jin (17.29.1)

shows that ‖α‖ = 0. �

Page 197: Analytic and Differential geometry Mikhail G. Katz∗

17.31. FREE LOOPS, BASED LOOPS, AND FUNDAMENTAL GROUP 197

Similarly, if two classes differ by a class of finite order, their stablenorms coincide. Thus the stable norm passes to the quotient latticedefined below.

Definition 17.29.3. The torsion subgroup of H1(M ;Z) will bedenoted T1(M) ⊂ H1(M ;Z). The quotient lattice L1(M) is the lattice

L1(M) = H1(M ;Z)/T1(M).

Proposition 17.29.4. The lattice L1(M) is isomorphic to Zb1(M),where b1 is called the first Betti number of M .

Proof. This is a general result in the theory of finitely generatedabelian groups. �

17.30. Stable systole

Definition 17.30.1. Let M be a manifold endowed with a Rie-mannian metric, and consider the associated stable norm ‖ ‖. Thestable 1-systole of M , denoted stsys1(M), is the least norm of a 1-homology class of infinite order:

stsys1(M) = inf{‖α‖ | α ∈ H1 (M,Z) \ T1(M)

}

= λ1(L1(M), ‖ ‖

).

Example 17.30.2. For an arbitrary metric on the 2-torus T2, the1-systole and the stable 1-systole coincide by Theorem 17.22.3:

sys1(T2) = stsys1(T

2),

for every metric on T2.

17.31. Free loops, based loops, and fundamental group

One can refine the notion of simple connectivity by introducing agroup, denoted

π1(X) = π1(X, x0),

and called the fundamental group of X relative to a fixed “base”point x0 ∈ X .

Definition 17.31.1. A based loop is a loop α : [0, 1] → X satisfyingthe condition α(0) = α(1) = x0.

In terms of the second item of Definition 17.18.1, we choose a fixedpoint s0 ∈ S1. For example, we can choose s0 = ei0 = 1 for theusual unit circle S1 ⊆ C, and require that α(s0) = x0. Then thegroup π1(X) is the quotient of the space of all based loops modulo

Page 198: Analytic and Differential geometry Mikhail G. Katz∗

198 17. MANIFOLDS AND GLOBAL GEOMETRY

the equivalence equivalence relation defined by homotopies fixing thebasepoint, cf. Definition 17.31.2.

The equivalence class of the identity element is precisely the setof contractible loops based at x0. The equivalence relation can bedescribed as follows for a pair of loops in terms of the second item ofDefinition 17.18.1.

Definition 17.31.2. Two based loops α, β : S1 → X are equiva-lent, or homotopic, if there is a continuous map of the cylinder S1 ×[c, d] → X whose restriction to S1×{c} is α, whose restriction to S1×{d} is β, while the map is constant on the segment {s0} × [c, d] of thecylinder, i.e. the basepoint does not move during the homotopy.

Definition 17.31.3. An equivalence class of based loops is calleda based homotopy class. Removing the basepoint restriction (as well asthe constancy condition of Definition 17.31.2), we obtain a larger classcalled a free homotopy class (of loops).

Definition 17.31.4. Composition of two loops is defined most con-veniently in terms of item 1 of Definition 17.18.1, by concatenating theirdomains.

In more detail, the product of a pair of loops, α : [−1, 0] → Xand β : [0,+1] → X , is a loop α.β : [−1, 1] → X , which coincideswith α and β in their domains of definition. The product loop α.βis continuous since α(0) = β(0) = x0. Then Theorem 17.18.5 can berefined as follows.

17.32. Fundamental groups of surfaces

Theorem 17.32.1. We have π1(S1) = Z, while π1(Sn) is the trivial

group for all n ≥ 2.

Definition 17.32.2. The 2-torus T2 is defined to be the followingCartesian product: T2 = S1 × S1, and can thus be realized as a subset

T2 = S1 × S1 ⊆ R2 × R2 = R4.

We have π1(T2) = Z2. The familiar doughnut picture realizes T2 asa subset in Euclidean space R3.

Definition 17.32.3. A 2-dimensional closed Riemannian manifold(i.e. surface) M is called orientable if it can be realized by a subsetof R3.

Theorem 17.32.4. The fundamental group of a surface of genus gis isomorphic to a group on 2g generators a1, b1, . . . , ag, bg satisfying

Page 199: Analytic and Differential geometry Mikhail G. Katz∗

17.32. FUNDAMENTAL GROUPS OF SURFACES 199

the unique relationg∏

i=1

aibia−1i b−1

i = 1.

Note that this is not the only possible presentation of the group interms of a single relation.

Page 200: Analytic and Differential geometry Mikhail G. Katz∗
Page 201: Analytic and Differential geometry Mikhail G. Katz∗

CHAPTER 18

Pu’s inequality

18.1. Hopf fibration h

To prove Pu’s inequality, we will need to study the Hopf fibrationmore closely.

The circle action in C2 restricts to the unit sphere S3 ⊂ C2, whichtherefore admits a fixed point free circle action. Namely, the circle

S1 = {eiθ : θ ∈ R} ⊆ C

acts on a point (z1, z2) ∈ C2 by

eiθ.(z1, z2) = (eiθz1, eiθz2). (18.1.1)

The quotient manifold (space of orbits) S3/S1 is the 2-sphere S2 (fromthe projective viewpoint, the quotient space is the complex projectiveline CP1).

Definition 18.1.1. The quotient map

h : S3 → S2, (18.1.2)

called the Hopf fibration.

18.2. Tangent map

Proposition 18.2.1. Consider a smooth map f :M → N betweendifferentiable manifolds. Then f defines a natural map

df : TM → TN

called the tangent map.

Proof. We represent a tangent vector v ∈ TpM by a curve

c(t) : I →M,

such that c′(o) = v. The composite map σ = f ◦ c : I → N is a curvein N . Then the vector σ′(0) is the image of v under df :

df(v) = σ′(0)

by chain rule, proving the proposition. �

201

Page 202: Analytic and Differential geometry Mikhail G. Katz∗

202 18. PU’S INEQUALITY

18.3. Riemannian submersion

Definition 18.3.1. A projection (R2, g) → R to the u1-axis iscalled a Riemannian submersion if the metric coefficients in the hori-zontal direction are independent of u2, or more precisely,

g = h(u1)(du1)2

+ k(u1, u2)(du2)2, (18.3.1)

for suitable functions h and k (note that the matrix of metric coeffi-cients is diagonal but not necessarily scalar).

Replacing the coordinates by (x, y), we obtain a somewhat morereadable formula

g = h(x)dx2 + k(x, y)dy2.

Being a Riemannian submersion is, of course, a local property, and westated it for Euclidean space for convenience. We will use it mostly inthe context of maps between compact manifolds.

Example 18.3.2. The metric (17.17.1) of the Poincare upperhalfplane admits a Riemannian submersion to (the positive ray of) the y-axis but not to the x-axis, as the conformal factor depends on y.

18.4. Riemannian submersions

Riemannian submersions in the 2-dimensional case were dealt within an Section 18.3.

Consider a fiber bundle map f :M → N between closed manifolds,i.e., a smooth map whose df is onto. Let F ⊆ M be the fiber overa point p ∈ N . The differential df : TM → TN vanishes on thesubspace TxF ⊆ TM at every point x ∈ F . More precisely, we havethe vertical space1

ker(dfx) = TxF.

Given a Riemannian metric on M , we can consider the orthogonalcomplement of TxF in TxM , denoted

Hx ⊆ TxM.

ThusTxM = TxF ⊕Hx

is an orthogonal decomposition.

Definition 18.4.1. A fiber bundle map f : M → N betweenRiemannian manifolds is called a Riemannian submersion if at everypoint x ∈ M , the restriction of df : Hx → Tf(x)N is an isometry.

1anchi

Page 203: Analytic and Differential geometry Mikhail G. Katz∗

18.5. HAMILTON QUATERNIONS 203

Proposition 18.4.2. Let S3 be the 3-sphere of constant sectionalcurvature +1, i.e., the unit sphere in R4. The Hopf fibration

h : S3 → S2

is a Riemannian submersion, for which the natural metric on the base S2

is a metric of constant Gaussian curvature +4.

The latter is explained by the fact that the maximal distance be-tween a pair of S1 orbits is π

2rather than π, in view of the fact that

a pair of antipodal points of S3 lies in a common orbit and thereforedescends to the same point of the quotient space.

A pair of points at maximal distance is defined by a pair v, w ∈ S3

such that w is orthogonal to Cv.

18.5. Hamilton quaternions

The quaternionic viewpoint will be applied in Section ?The algebraof the Hamilton quaternions is the real 4-dimensional vector space withreal basis {1, i, j, k}, so that

H = R1 + Ri+ Rj + Rk

equipped with an associative (chok kibutz), non-commutative productoperation. This operation has the following properties:

(1) the center of H is R1;(2) the operation satisfies the relations i2 = j2 = k2 = −1 and ij =

−ji = k, jk = −kj = i, ki = −ik = j.

Lemma 18.5.1. The algebra H is naturally isomorphic to the com-plex vector space C2 via the identification R1 + Ri ≃ C.

Proof. The subspace Rj + Rk can be thought of as

Rj + Rij = (R1 + Ri)j,

and we therefore obtain a natural isomorphism

H = C1 + Cj,

showing that {1, j} is a complex basis for H. �

Theorem 18.5.2. Nonzero quaternions form a group under quater-nionic multiplication.

Proof. Given q = a+bi+cj+dk ∈ H, let N(q) = a2+b2+c2+d2,and q = a − bi − cj − dk. Then qq = N(q). Therefore we have amultiplicative inverse

q−1 =1

N(q)q,

proving the theorem. �

Page 204: Analytic and Differential geometry Mikhail G. Katz∗

204 18. PU’S INEQUALITY

18.6. Complex structures on the algebra H

Page 205: Analytic and Differential geometry Mikhail G. Katz∗

CHAPTER 19

Approach using energy-area identity

19.1. An integral-geometric identity

Let T2 be a torus with an arbitrary metric. Let T0 = R2/L be theflat torus conformally equivalent to T2. Let ℓ0 = ℓ0(x) be a simple

closed geodesic of T0. Thus ℓ0 is the projection of a line ℓ0 ⊆ R2.Let ℓy ⊆ R2 be the line parallel to ℓ0 at distance y > 0 from ℓ0 (here

we must “choose sides”, e.g. by orienting ℓ0 and requiring ℓy to lie to

the left of ℓ0). Denote by ℓy ⊆ T0 the closed geodesic loop defined by

the image of ℓy. Let y0 > 0 be the smallest number such that ℓy0 = ℓ0,

i.e. the lines ℓy0 and ℓ0 both project to ℓ0.Note that the loops in the family {ℓy} ⊆ T2 are not necessarily

geodesics with respect to the metric of T2. On the other hand, thefamily satisfies the following identity.

Lemma 19.1.1 (An elementary integral-geometric identity). Themetric on T2 satisfies the following identity:

area(T2) =

∫ y0

0

E(ℓy)dy, (19.1.1)

where E is the energy of a loop with respect to the metric of T2, seeDefinition 17.9.2.

Proof. Denote by f 2 the conformal factor of T2 with respect tothe flat metric T0. Thus the metric on T2 can be written as

f 2(x, y)(dx2 + dy2).

By Fubini’s theorem applied to a rectangle with sides lengthT0(ℓ0)

and y0, combined with Theorem 17.7.3, we obtain

area(T2) =

T0

f 2dxdy

=

∫ y0

0

(∫

ℓy

f 2(x, y)dx

)dy

=

∫ y0

0

E(ℓy)dy,

205

Page 206: Analytic and Differential geometry Mikhail G. Katz∗

206 19. APPROACH USING ENERGY-AREA IDENTITY

proving the lemma. �

Remark 19.1.2. The identity (19.1.1) can be thought of as thesimplest integral-geometric identity, cf. equation (19.4.3).

19.2. Two proofs of the Loewner inequality

We give a slightly modified version of M. Gromov’s proof [Gro96],using conformal representation and the Cauchy-Schwarz inequality, ofthe Loewner inequality (16.2.1) for the 2-torus, see also [CK03]. Wepresent the following slight generalisation: there exists a pair of closedgeodesics on (T2, g), of respective lengths λ1 and λ2, such that

λ1λ2 ≤ γ2 area(g), (19.2.1)

and whose homotopy classes form a generating set for π1(T2) = Z×Z.

Proof. The proof relies on the conformal representation

φ : T0 → (T2, g),

where T0 is flat, cf. uniformisation theorem 13.8.2. Here the map φmay be chosen in such a way that (T2, g) and T0 have the same area.Let f be the conformal factor, so that

g = f 2((du1)2 + (du2)2

)

as in formula (17.7.1), where (du1)2+(du2)2 (locally) is the flat metric.Let ℓ0 be any closed geodesic in T0. Let {ℓs} be the family of

geodesics parallel to ℓ0. Parametrize the family {ℓs} by a circle S1 oflength σ, so that

σℓ0 = area(T0).

Thus T0 → S1 is a Riemannian submersion, cf. Definition 18.3.1. Then

area(T2) =

T0

f 2.

By Fubini’s theorem, we have area(T2) =∫S1 ds

∫ℓsf 2dt. Therefore by

the Cauchy-Schwarz inequality,

area(T2) ≥∫

S1

ds

(∫ℓsfdt)2

ℓ0=

1

ℓ0

S1

ds (lengthφ(ℓs))2 .

Hence there is an s0 such that area(T2) ≥ σℓ0lengthφ(ℓs0)

2, so that

lengthφ(ℓs0) ≤ ℓ0.

This reduces the proof to the flat case. Given a lattice in C, we choosea shortest lattice vector λ1, as well as a shortest one λ2 not proportionalto λ1. The inequality now follows from Lemma 15.9.1. In the boundary

Page 207: Analytic and Differential geometry Mikhail G. Katz∗

19.2. TWO PROOFS OF THE LOEWNER INEQUALITY 207

case of equality, one can exploit the equality in the Cauchy-Schwarzinequality to prove that the conformal factor must be constant. �

Alternative proof. Let ℓ0 be any simple closed geodesic in T0.Since the desired inequality (19.2.1) is scale-invariant, we can assumethat the loop has unit length:

lengthT0(ℓ0) = 1,

and, moreover, that the corresponding covering transformation of theuniversal cover C = T0 is translation by the element 1 ∈ C. Wecomplete 1 to a basis {τ, 1} for the lattice of covering transformationsof T0. Note that the rectangle defined by

{z = x+ iy ∈ C | 0 < x < 1, 0 < y < Im(τ)]}is a fundamental domain for T0, so that area(T0) = Im(τ). The maps

ℓy(x) = x+ iy, x ∈ [0, 1]

parametrize the family of geodesics parallel to ℓ0 on T0. Recall thatthe metric of the torus T is f 2(dx2 + dy2).

Lemma 19.2.1. We have the following relation between the lengthand energy of a loop:

length(ℓy)2 ≤ E(ℓy).

Proof. By the Cauchy-Schwarz inequality,

∫ 1

0

f 2(x, y)dx ≥(∫ 1

0

f(x, y)dx

)2

,

proving the lemma. �

Now by Lemma 19.1.1 and Lemma 19.2.1, we have

area(T2) ≥∫ Im(τ)

0

(length(ℓy)

)2dy.

Hence there is a y0 such that

area(T2) ≥ Im(τ) length(ℓy0)2,

so that length(ℓy0) ≤ 1. This reduces the proof to the flat case. Givena lattice in C, we choose a shortest lattice vector λ1, as well as ashortest one λ2 not proportional to λ1. The inequality now followsfrom Lemma 15.9.1. �

Page 208: Analytic and Differential geometry Mikhail G. Katz∗

208 19. APPROACH USING ENERGY-AREA IDENTITY

Remark 19.2.2. Define a conformal invariant called the capacityof an annulus as follows. Consider a right circular cylinder

ζκ = R/Z× [0, κ]

based on a unit circle R/Z. Its capacity C(ζκ) is defined to be itsheight, C(ζκ) = κ. Recall that every annular region in the plane is con-formally equivalent to such a cylinder, and therefore we have defined aconformal invariant of an arbitrary annular region. Every annular re-gion R satisfies the inequality area(R) ≥ C(R) sys1(R)

2. Meanwhile, ifwe cut a flat torus along a shortest loop, we obtain an annular region R

with capacity at least C(R) ≥ γ−12 =

√32. This provides an alternative

proof of the Loewner theorem. In fact, we have the following identity:

confsys1(g)2C(g) = 1, (19.2.2)

where confsys is the conformally invariant generalisation of the homol-ogy systole, while C(g) is the largest capacity of a cylinder obtainedby cutting open the underlying conformal structure on the torus.

Question 19.2.3. Is the Loewner inequality (16.2.1) satisfied byevery orientable nonsimply connected compact surface? Inspite of itselementary nature, and considerable research devoted to the area, thequestion is still open. Recently the case of genus 2 was settled as wellas genus g ≥ 20

19.3. Hopf fibration and the Hamilton quaternions

The circle action in C2 restricts to the unit sphere S3 ⊂ C2, whichtherefore admits a fixed point free circle action. Namely, the circle S1 ={eiθ : θ ∈ R} ⊆ C acts on (z1, z2) by

eiθ.(z1, z2) = (eiθz1, eiθz2).

It is possible to show that the quotient manifold (space of orbits) S3/S1

is the 2-sphere S2 (from the projective viewpoint, the quotient is thecomplex projective line CP1). Moreover, the quotient map

S3 → S2, (19.3.1)

called the Hopf fibration, is a Riemannian submersion, for which thenatural metric on the base is a metric of constant Gaussian curva-ture +4 rather than +1. The latter is explained by the fact that themaximal distance between a pair of S1 orbits is π

2rather than π, in

view of the fact that a pair of antipodal points of S3 lies in a commonorbit.

Furthermore, the sphere S3 can be thought of as the Lie groupformed of the unit (Hamilton) quaternions inH = C2. The Lie group SO(3)

Page 209: Analytic and Differential geometry Mikhail G. Katz∗

19.4. DOUBLE FIBRATION OF SO(3) AND INTEGRAL GEOMETRY ON S2209

can be identified with S3/{±1}. Therefore the fibration (19.3.1) de-scends to a fibration

SO(3) → S2, (19.3.2)

exploited in the next section.

Remark 19.3.1. The group G = SO(3) can be identified with theunit tangent bundle of S2, as follows. Choose a fixed unit tangentvector v to S2, and send an element g ∈ G, acting on S2, to the imageof v under dg : TS2 → TS2.

Remark 19.3.2. Let us elaborate a bit further on the quaternionicviewpoint, with an eye to aplying it in Section 19.4. A choice of a purelyimaginary Hamilton quaternion in H = R1+Ri+Rj+Rk, i.e. a linearcombination of i, j, and k, specifies a complex structure on H. Such acomplex structure leads to a fibration of the unit sphere, as in (19.3.1).Choosing a pair of purely imaginary quaternions which are orthogonal,results in a pair of circle fibrations of S3 such that a fiber of one fibrationis perpendicular to a fiber of the other fibration whenever two suchfibers meet. The construction of two “perpendicular” fibrations alsodescends to the quotient by the natural Z2 action on S3.

19.4. Double fibration of SO(3) and integral geometry on S2

In this section, we present a self-contained account of an integral-geometric identity, used in the proof of Pu’s inequality (16.1.6) in Sec-tion 19.5, as well as in our argument in Lemma The identity has itsorigin in results of P. Funk [Fu16] determining a symmetric functionon the two-sphere from its great circle integrals; see [Hel99, Proposi-tion 2.2, p. 59], as well as Preface therein.

The sphere S2 is the homogeneous space of the Lie group SO(3),cf. (19.3.2), so that S2 = SO(3)/SO(2), cf. [Ar83, p. 82]. Denoteby SO(2)σ the fiber over (stabilizer of) a typical point σ ∈ S2. Theprojection

p : SO(3) → S2 (19.4.1)

is a Riemannian submersion for the standard metric of constant sec-tional curvature 1

4on SO(3). The total space SO(3) admits another

Riemannian submersion, which we denote

q : SO(3) → RP2, (19.4.2)

whose typical fiber ν is an orbit of the geodesic flow on SO(3) when thelatter is viewed as the unit tangent bundle of S2, cf. Remark 19.3.1.

Here RP2 denotes the double cover of RP2, homeomorphic to the 2-sphere, cf. Definition 17.12.2. A fiber of fibration q is the collection of

Page 210: Analytic and Differential geometry Mikhail G. Katz∗

210 19. APPROACH USING ENERGY-AREA IDENTITY

SO(2)

##GGG

GGGG

GGG

(RP2, dν

)

SO(3)

q99ttttttttt

p

&&LLL

LLLL

LLL

ν

::uuuuuuuuuu

great circle// (S2, dσ)

Figure 19.4.1. Integral geometry on S2, cf. (19.4.1) and (19.4.2)

unit vectors tangent to a given directed closed geodesic (great circle)on S2. Thus, each orbit ν projects under p to a great circle on the

sphere. We think of the base space RP2 of q as the configuration spaceof oriented great circles on the sphere, with measure dν. While fibra-tion q may seem more “mysterious” than fibration p, the two are actu-ally equivalent from the quaternionic point of view, cf. Remark 19.3.2.

The diagram of Figure 19.4.1 illustrates the maps defined so far.Denote by Eg(ν) the energy, and by Lg(ν) the length, of a curve ν

with respect to a (possibly singular) metric g = f 2g0.

Theorem 19.4.1. We have the following integral-geometric iden-tity:

area(g) =1

˜RP2

Eg(ν)dν.

Proof. We apply Fubini’s theorem [Ru87, p. 164] twice, to both qand p, to obtain

˜RP2

Eg(ν)dν =

˜RP2

(∫

ν

f 2 ◦ p ◦ ν(t)dt)

=

SO(3)

f 2 ◦ p

=

S2

f 2 dσ

(∫

SO(2)σ

1

)

= 2π area(g),

(19.4.3)

completing the proof of the theorem. �

Proposition 19.4.2. Let f be a square integrable function on S2,which is positive and continuous except possibly for a finite number of

Page 211: Analytic and Differential geometry Mikhail G. Katz∗

19.6. A TABLE OF OPTIMAL SYSTOLIC RATIOS OF SURFACES 211

points where f either vanishes or has a singularity of type 1√r, Then

there is a great circle ν such that the following inequality is satisfied:(∫

ν

f(ν(t))dt

)2

≤ π

S2

f 2(σ)dσ, (19.4.4)

where t is the arclength parameter, and dσ is the Riemannian measureof the metric g0 of the standard unit sphere. In other words, there is agreat circle of length L in the metric f 2g0, so that L2 ≤ πA, where A isthe Riemannian surface area of the metric f 2g0. In the boundary caseof equality in (19.4.4), the function f must be constant.

Proof. The proof is an averaging argument and shows that theaverage length of great circles is short. Comparing length and energyyields the inequality(∫

˜RP2 Lg(ν)2)

2π≤∫

˜RP2

Eg(ν)dν. (19.4.5)

The formula now follows from Theorem 19.4.1, since area(RP 2) = 4π.In the boundary case of equality in (19.4.4), one must have equality

also in inequality (19.4.5) relating length and energy. It follows that theconformal factor f must be constant along every great circle. Hence fis constant everywhere on S2. �

19.5. Proof of Pu’s inequality

A metric g on RP2 lifts to a centrally symmetric metric g on S2.Applying Proposition 19.4.2 to g, we obtain a great circle of g-length Lsatisfying L2 ≤ πA, where A is the Riemannian surface area of g. Thus

we have(L2

)2 ≤ π2

(A2

), where sys1(g) ≤ L

2while area(g) = A

2, proving

inequality (16.1.6). See [Iv02] for an alternative proof.

19.6. A table of optimal systolic ratios of surfaces

Denote by SR(M) the supremum of the systolic ratios,

SR(M) = supg

sys1(g)2

area(g),

ranging over all metrics g on a surface M . The known values of theoptimal systolic ratio are tabulated in Figure 19.6.1. It is interesting tonote that the optimal ratio for the Klein bottle RP2#RP2 is achievedby a singular metric, described in the references listed in the table.

Page 212: Analytic and Differential geometry Mikhail G. Katz∗

212 19. APPROACH USING ENERGY-AREA IDENTITY

SR(M) numerical value where to find it

M = RP2 =π

2[Pu52] ≈ 1.5707 Section 19.5

infinite π1(M) <4

3[Gro83] < 1.3333 . . . (16.1.7)

M = T2 = 2√3(Loewner) ≈ 1.1547 (16.2.1)

M =RP2#RP2

23/2≈ 1.1107 [Bav86, Bav06,

Sak88]

M of genus 2 > 13

(√2 + 1

)> 0.8047 ?

M of genus 3 ≥ 87√3

> 0.6598 [Cal96]

Figure 19.6.1. Values for optimal systolic ratio SR of surface M

Page 213: Analytic and Differential geometry Mikhail G. Katz∗

CHAPTER 20

A primer on surfaces

In this Chapter, we collect some classical facts on Riemann surfaces.More specifically, we deal with hyperelliptic surfaces, real surfaces, andKatok’s optimal bound for the entropy of a surface.

20.1. Hyperelliptic involution

Let M be an orientable closed Riemann surface which is not asphere. By a Riemann surface, we mean a surface equipped with a fixedconformal structure, cf. Definition 17.7.4, while all maps are angle-preserving.

Furthermore, we will assume that the genus is at least 2.

Definition 20.1.1. A hyperelliptic involution of a Riemann sur-face M of genus p is a holomorphic (conformal) map, J : M → M ,satisfying J2 = 1, with 2p + 2 fixed points.

Definition 20.1.2. A surface M admitting a hyperelliptic involu-tion will be called a hyperelliptic surface.

Remark 20.1.3. The involution J can be identified with the non-trivial element in the center of the (finite) automorphism group of M(cf. [FK92, p. 108]) when it exists, and then such a J is unique,cf. [Mi95, p.204] (recall that the genus is at least 2).

It is known that the quotient of M by the involution J produces aconformal branched 2-fold covering

Q :M → S2 (20.1.1)

of the sphere S2.

Definition 20.1.4. The 2p + 2 fixed points of J are called Weier-strass points. Their images in S2 under the ramified double cover Q offormula (20.1.1) will be referred to as ramification points.

A notion of a Weierstrass point exists on any Riemann surface, butwill only be used in the present text in the hyperelliptic case.

213

Page 214: Analytic and Differential geometry Mikhail G. Katz∗

214 20. A PRIMER ON SURFACES

Example 20.1.5. In the case p = 2, topologically the situationcan be described as follows. A simple way of representing the figure 8contour in the (x, y) plane is by the reducible curve

(((x− 1)2 + y2)− 1)(((x+ 1)2 + y2)− 1) = 0 (20.1.2)

(or, alternatively, by the lemniscate r2 = cos 2θ in polar coordinates,i.e. the locus of the equation (x2 + y2)2 = x2 − y2).

Now think of the figure 8 curve of (20.1.2) as a subset of R3. Theboundary of its tubular neighborhood in R3 is a genus 2 surface. Rota-tion by π around the x-axis has six fixed points on the surface, namely,a pair of fixed points near each of the points −2, 0, and +2 on the x-axis. The quotient by the rotation can be seen to be homeomorphic tothe sphere.

A similar example can be repackaged in a metrically more preciseway as follows.

Example 20.1.6. We start with a round metric on RP2. Nowattach a small handle. The orientable double cover M of the resultingsurface can be thought of as the unit sphere in R3, with two littlehandles attached at north and south poles, i.e. at the two points wherethe sphere meets the z-axis. Then one can think of the hyperellipticinvolution J as the rotation of M by π around the z-axis. The sixfixed points are the six points of intersection of M with the z-axis.Furthermore, there is an orientation reversing involution τ onM , givenby the restriction toM of the antipodal map in R3. The composition τ◦J is the reflection fixing the xy-plane, in view of the following matrixidentity:

−1 0 00 −1 00 0 −1

−1 0 00 −1 00 0 1

=

1 0 00 1 00 0 −1

. (20.1.3)

Meanwhile, the induced orientation reversing involution τ0 on S2 canjust as well be thought of as the reflection in the xy-plane. This isbecause, at the level of the 2-sphere, it is “the same thing as” thecomposition τ ◦ J . Thus the fixed circle of τ0 is precisely the equator,cf. formula (20.3.3). Then one gets a quotient metric on S2 which isroughly that of the western hemisphere, with the boundary longitudefolded in two, The metric has little bulges along the z-axis at northand south poles, which are leftovers of the small handle.

Page 215: Analytic and Differential geometry Mikhail G. Katz∗

20.3. OVALLESS SURFACES 215

20.2. Hyperelliptic surfaces

For a treatment of hyperelliptic surfaces, see [Mi95, p. 60-61].By [Mi95, Proposition 4.11, p. 92], the affine part of a hyperellip-tic surface M is defined by a suitable equation of the form

w2 = f(z) (20.2.1)

in C2, where f is a polynomial. On such an affine part, the map J isgiven by J(z, w) = (z,−w), while the hyperelliptic quotient map Q :M → S2 is represented by the projection onto the z-coordinate in C2.

A slight technical problem here is that the map

M → CP2, (20.2.2)

whose image is the compactification of the solution set of (20.2.1), isnot an imbedding. Indeed, there is only one point at infinity, given inhomogeneous coordinates by [0 : w : 0]. This point is a singularity.A way of desingularizing it using an explicit change of coordinates atinfinity is presented in [Mi95, p. 60-61]. The resulting smooth surfaceis unique [DaS98, Theorem, p. 100].

Remark 20.2.1. To explain what happens “geometrically”, notethat there are two points on our affine surface “above infinity”. Thismeans that for a large circle |z| = r, there are two circles above it satis-fying equation (20.2.1) where f has even degree 2p+2 (for a Weierstrasspoint we would only have one circle). To see this, consider z = reia. Asthe argument a varies from 0 to 2π, the argument of f(z) will changeby (2p + 2)2π. Thus, if (reia, w(a)) represents a continuous curve onour surface, then the argument of w changes by (2p+2)π, and hence weend up where we started, and not at −w (as would be the case were thepolynomial of odd degree). Thus there are two circles on the surfaceover the circle |z| = r. We conclude that to obtain a smooth compactsurface, we will need to add two points at infinity, cf. discussion around[FK92, formula (7.4.1), p. 102].

Thus, the affine part of M , defined by equation (20.2.1), is a Rie-mann surface with a pair of punctures p1 and p2. A neighborhood ofeach pi is conformally equivalent to a punctured disk. By replacingeach punctured disk by a full one, we obtain the desired compact Rie-mann surface M . The point at infinity [0 : w : 0] ∈ CP2 is the imageof both pi under the map (20.2.2).

20.3. Ovalless surfaces

Denote by M ι the fixed point set of an involution ι of a Riemannsurface M . Let M be a hyperelliptic surface of even genus p. Let J :

Page 216: Analytic and Differential geometry Mikhail G. Katz∗

216 20. A PRIMER ON SURFACES

M → M be the hyperelliptic involution, cf. Definition 20.1.1. Let τ :M →M be a fixed point free, antiholomorphic involution.

Lemma 20.3.1. The involution τ commutes with J and descendsto S2. The induced involution τ0 : S

2 → S2 is an inversion in a circleC0 = Q(M τ◦J ). The set of ramification points is invariant under theaction of τ0 on S2.

Proof. By the uniqueness of J , cf. Remark 20.1.3, we have thecommutation relation

τ ◦ J = J ◦τ, (20.3.1)

cf. relation (20.1.3). Therefore τ descends to an involution τ0 on thesphere. There are two possibilities, namely, τ is conjugate, in the groupof fractional linear transformations, either to the map z 7→ z, or to themap z 7→ −1

z. In the latter case, τ is conjugate to the antipodal map

of S2.In the case of even genus, there is an odd number of Weierstrass

points in a hemisphere. Hence the inverse image of a great circle is aconnected loop. Thus we get an action of Z2 × Z2 on a loop, resultingin a contradiction.

In more detail, the set of the 2p + 2 ramification points on S2 iscentrally symmetric. Since there is an odd number, p + 1, of ramifi-cation points in a hemisphere, a generic great circle A ⊆ S2 has theproperty that its inverse image Q−1(A) ⊆ M is connected. Thus bothinvolutions τ and J , as well as τ ◦ J , act fixed point freely on theloop Q−1(A) ⊆M , which is impossible. Therefore τ0 must fix a point.It follows that τ0 is an inversion in a circle. �

Suppose a hyperelliptic Riemann surfaceM admits an antiholomor-phic involution τ . In the literature, the components of the fixed pointsetM τ of τ are sometimes referred to as “ovals”. When τ is fixed pointfree, we introduce the following terminology.

Definition 20.3.2. A hyperelliptic surface (M,J) of even positivegenus p > 0 is called ovalless real if one of the following equivalentconditions is satisfied:

(1) M admits an imaginary reflection, i.e. a fixed point free, an-tiholomorphic involution τ ;

(2) the affine part of M is the locus in C2 of the equation

w2 = −P (z), (20.3.2)

where P is a monic polynomial, of degree 2p + 2, with realcoefficients, no real roots, and with distinct roots.

Page 217: Analytic and Differential geometry Mikhail G. Katz∗

20.4. KATOK’S ENTROPY INEQUALITY 217

Lemma 20.3.3. The two ovalless reality conditions of Definition 20.3.2are equivalent.

Proof. A related result appears in [GroH81, p. 170, Proposi-tion 6.1(2)]. To prove the implication (2) =⇒ (1), note that complexconjugation leaves the equation invariant, and therefore it also leavesinvariant the locus of (20.3.2). A fixed point must be real, but P is pos-itive hence (20.3.2) has no real solutions. There is no real solution atinfinity, either, as there are two points at infinity which are not Weier-strass points, since P is of even degree, as discussed in Remark 20.2.1.The desired imaginary reflection τ switches the two points at infinity,and, on the affine part of the Riemann surface, coincides with complexconjugation (z, w) 7→ (z, w) in C2.

To prove the implication (1) =⇒ (2), note that by Lemma 20.3.1,the induced involution τ0 on S

2 =M/ J may be thought of as complexconjugation, by choosing the fixed circle of τ0 to be the circle

R ∪ {∞} ⊆ C ∪ {∞} = S2. (20.3.3)

Since the surface is hyperelliptic, it is the smooth completion of thelocus in C2 of some equation of the form (20.3.2), cf. (20.2.1). Here Pis of degree 2p+2 with distinct roots, but otherwise to be determined.The set of roots of P is the set of (the z-coordinates of) the Weierstrasspoints. Hence the set of roots must be invariant under τ0. Thus theroots of the polynomial either come in conjugate pairs, or else are real.Therefore P has real coefficients. Furthermore, the leading cofficientof P may be absorbed into the w-coordinate by extracting a squareroot. Here we may have to rotate w by i, but at any rate the coefficientsof P remain real, and thus P can be assumed monic.

If P had a real root, there would be a ramification point fixed by τ0.But then the corresponding Weierstrass point must be fixed by τ , aswell! This contradicts the fixed point freeness of τ . Thus all roots of Pmust come in conjugate pairs. �

20.4. Katok’s entropy inequality

Let (M, g) be a closed surface with a Riemannian metric. De-

note by (M, g) the universal Riemannian cover of (M, g). Choose apoint x0 ∈ M .

Definition 20.4.1. The volume entropy (or asymptotic volume) h(M, g)of a surface (M, g) is defined by setting

h(M, g) = limR→+∞

log (volg B(x0, R))

R, (20.4.1)

Page 218: Analytic and Differential geometry Mikhail G. Katz∗

218 20. A PRIMER ON SURFACES

where volg B(x0, R) is the volume (area) of the ball of radius R centered

at x0 ∈ M .

Since M is compact, the limit in (20.4.1) exists, and does not de-

pend on the point x0 ∈ M [Ma79]. This asymptotic invariant describesthe exponential growth rate of the volume in the universal cover.

Definition 20.4.2. The minimal volume entropy, MinEnt, ofM isthe infimum of the volume entropy of metrics of unit volume on M , orequivalently

MinEnt(M) = infgh(M, g) vol(M, g)

12 (20.4.2)

where g runs over the space of all metrics on M . For an n-dimensionalmanifold in place of M , one defines MinEnt similarly, by replacing theexponent of vol by 1

n.

The classical result of A. Katok [Kato83] states that every metric gon a closed surfaceM with negative Euler characteristic χ(M) satisfiesthe optimal inequality

h(g)2 ≥ 2π|χ(M)|area(g)

. (20.4.3)

Inequality (20.4.3) also holds for hom ent(g) [Kato83], as well as thetopological entropy, since the volume entropy bounds from below thetopological entropy (see [Ma79]). We recall the following well-knownfact, cf. [KatH95, Proposition 9.6.6, p. 374].

Lemma 20.4.3. Let (M, g) be a closed Riemannian manifold. Then,

h(M, g) = limT→+∞

log(P ′(T ))

T(20.4.4)

where P ′(T ) is the number of homotopy classes of loops based at somefixed point x0 which can be represented by loops of length at most T .

Proof. Let x0 ∈M and choose a lift x0 ∈ M . The group

Γ := π1(M,x0)

acts on M by isometries. The orbit of x0 under Γ is denoted Γ.x0.Consider a fundamental domain ∆ for the action of Γ, containing x0.Denote by D the diameter of ∆. The cardinal of Γ.x0 ∩ B(x0, R) isbounded from above by the number of translated fundamental do-mains γ.∆, where γ ∈ Γ, contained in B(x0, R+D). It is also bounded

Page 219: Analytic and Differential geometry Mikhail G. Katz∗

20.4. KATOK’S ENTROPY INEQUALITY 219

from below by the number of translated fundamental domains γ.∆contained in B(x0, R). Therefore, we have

vol(B(x0, R))

vol(M, g)≤ card (Γ.x0 ∩B(x0, R)) ≤

vol(B(x0, R +D))

vol(M, g).

(20.4.5)Take the log of these terms and divide by R. The lower bound becomes

1

Rlog

(vol(B(R))

vol(g)

)=

=1

Rlog (vol(B(R)))− 1

Rlog (vol(g)) ,

(20.4.6)

and the upper bound becomes

1

Rlog

(vol(B(R +D))

vol(g)

)=

=R +D

R

1

R +Dlog(vol(B(R +D)))− 1

Rlog(vol(g)).

(20.4.7)

Hence both bounds tend to h(g) when R goes to infinity. Therefore,

h(g) = limR→+∞

1

Rlog (card(Γ.x0 ∩B(x0, R))) . (20.4.8)

This yields the result since P ′(R) = card(Γ.x0 ∩ B(x0, R)). �

Page 220: Analytic and Differential geometry Mikhail G. Katz∗
Page 221: Analytic and Differential geometry Mikhail G. Katz∗

Bibliography

[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?.

[1] Arnol’d, V. I. On the teaching of mathematics.(Russian) Uspekhi Mat.Nauk 53 (1998), no. 1(319), 229–234; translation in Russian Math. Sur-veys 53 (1998), no. 1, 229–236.

[Ba93] Babenko, I.: Asymptotic invariants of smooth manifolds. Russian Acad.Sci. Izv. Math. 41 (1993), 1–38.

[Ba02a] Babenko, I. Forte souplesse intersystolique de varietes fermees et depolyedres. Annales de l’Institut Fourier 52, 4 (2002), 1259-1284.

[Ba02b] Babenko, I. Loewner’s conjecture, Besicovich’s example, and relativesystolic geometry. [Russian]. Mat. Sbornik 193 (2002), 3-16.

[Ba04] Babenko, I.: Geometrie systolique des varietes de groupe fondamen-tal Z2, Semin. Theor. Spectr. Geom. Grenoble 22 (2004), 25-52.

[BB05] Babenko, I.; Balacheff, F.: Geometrie systolique des sommes connexeset des revetements cycliques, Mathematische Annalen (to appear).

[BaK98] Babenko, I.; Katz, M.: Systolic freedom of orientable manifolds, AnnalesScientifiques 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 inoptimal Loewner-type inequalities, Trans. Amer. Math. Soc., to appear.See arXiv:math.DG/0406008

[BCIK05] Bangert, V; Croke, C.; Ivanov, S.; Katz, M.: Filling area conjecture andovalless real hyperelliptic surfaces, Geometric and Functional Analysis(GAFA) 15 (2005) no. 3. See arXiv:math.DG/0405583

221

Page 222: Analytic and Differential geometry Mikhail G. Katz∗

222 BIBLIOGRAPHY

[BK03] Bangert, V.; Katz, M.: Stable systolic inequalities and cohomologyproducts, Comm. Pure Appl. Math. 56 (2003), 979–997. Available atarXiv:math.DG/0204181

[BK04] Bangert, V; Katz, M.: An optimal Loewner-type systolic inequality andharmonic 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.: Inegalite isosystolique pour la bouteille de Klein.Math. Ann.274 (1986), no. 3, 439–441.

[Bav88] Bavard, C.: Inegalites isosystoliques conformes pour la bouteille deKlein. 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 geometrie systolique de la bouteille deKlein. Arch. Math. (Basel) 87 (2006), no. 1, 72–74.

[BeM] Berge, A.-M.; Martinet, J. Sur un probleme de dualite lie aux spheresen geometrie des nombres. J. Number Theory 32 (1989), 14–42.

[Ber72] Berger, M. A l’ombre de Loewner. Ann. Scient. Ec. Norm. Sup. (Ser. 4)5 (1972), 241–260.

[Ber80] Berger, M.: Une borne inferieure pour le volume d’une variete rieman-nienne en fonction du rayon d’injectivite. Ann. Inst. Fourier 30 (1980),259–265.

[Ber93] Berger, M. Systoles et applications selon Gromov. Seminaire N. Bour-baki, expose 771, Asterisque 216 (1993), 279–310.

[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 ihrerGrenzgebiete (3), 10. Springer-Verlag, Berlin, 1987.

[BesCG95] Besson, G.; Courtois, G.;, Gallot, S.: Volume et entropie minimale desespaces localement symetriques. Invent. Math. 103 (1991), 417–445.

[BesCG95] Besson, G.; Courtois, G.;, Gallot, S.: Hyperbolic manifolds, amalga-mated products and critical exponents. C. R. Math. Acad. Sci. Paris336 (2003), no. 3, 257–261.

[Bl61] Blatter, C. Uber Extremallangen auf geschlossenen Flachen. Comment.Math. Helv. 35 (1961), 153–168.

[Bl61b] Blatter, C.: Zur Riemannschen Geometrie im Grossen auf dem Mobius-band. Compositio Math. 15 (1961), 88–107.

[Borovik & Katz 2012] Borovik, A., Katz, M. “Who gave you the Cauchy–Weierstrass tale? The dual history of rigorous calculus.” Foundationsof Science 17, no. 3, 245-276.See http://dx.doi.org/10.1007/s10699-011-9235-x

[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 rationalhomotopy type. Mem. Amer. Math. Soc. 8, 179, Providence, R.I., 1976.

[BI94] Burago, D.; Ivanov, S.: Riemannian tori without conjugate points areflat. Geom. Funct. Anal. 4 (1994), no. 3, 259–269.

Page 223: Analytic and Differential geometry Mikhail G. Katz∗

BIBLIOGRAPHY 223

[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 derMathematischen 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 oflarge 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 dela 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. Birkhauser 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.

[Cauchy 1826] Cauchy, A. L. Leons sur les applications du calcul infinitesimal a lageometrie. Paris: Imprimerie Royale, 1826.

[Ch93] Chavel, I.: Riemannian geometry—a modern introduction. CambridgeTracts in Mathematics, 108. Cambridge University Press, Cambridge,1993.

[CoS94] Conway, J.; Sloane, N.: On lattices equivalent to their duals. J. NumberTheory 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.; Tanre, D.: Lusternik-Schnirelmanncategory. Mathematical Surveys and Monographs, 103. American Math-ematical Society, Providence, RI, 2003.

[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.[CK03] Croke, C.; Katz, M.: Universal volume bounds in Riemannian mani-

folds, Surveys in Differential Geometry VIII, Lectures on Geometry andTopology held in honor of Calabi, Lawson, Siu, and Uhlenbeck at Har-vard University, May 3- 5, 2002, edited by S.T. Yau (Somerville, MA:International Press, 2003.) pp. 109 - 137. See arXiv:math.DG/0302248

[DaS98] Danilov, V. I.; Shokurov, V. V.: Algebraic curves, algebraic manifoldsand schemes. Translated from the 1988 Russian original by D. Corayand V. N. Shokurov. Reprint of the original English edition from theseries Encyclopaedia of Mathematical Sciences [Algebraic geometry. I,Encyclopaedia Math. Sci., 23, Springer, Berlin, 1994]. Springer-Verlag,Berlin, 1998.

Page 224: Analytic and Differential geometry Mikhail G. Katz∗

224 BIBLIOGRAPHY

[Do79] Dombrowski, P.: 150 years after Gauss’ “Disquisitiones generales circasuperficies curvas”. With the original text of Gauss. Asterisque, 62.Societe Mathematique de France, Paris, 1979.

[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. GraduateTexts 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. IndianaUniv. Math. J. 24 (1974), 351–407.

[FHL98] Felix, Y.; Halperin, S.; Lemaire, J.-M.: The rational LS category ofproducts and of Poincare duality 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.

[Goldman 2005] Goldman, Ron. Curvature formulas for implicit curves and sur-faces. Comput. Aided Geom. Design 22 (2005), no. 7, 632–658. Seehttp://u.math.biu.ac.il/ katzmik/goldman05.pdf

[GG92] Gomez-Larranaga, J.; Gonzalez-Acuna, F.: Lusternik-Schnirel’manncategory 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, andcohomology. Vol. III: Cohomology of principal bundles and homogeneousspaces, 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 [1994].

[Gro81a] Gromov, M.: Structures metriques pour les varietes riemanniennes.(French) [Metric structures for Riemann manifolds] Edited by J. La-fontaine and P. Pansu. Textes Mathematiques, 1. CEDIC, Paris, 1981.

[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.

Page 225: Analytic and Differential geometry Mikhail G. Katz∗

BIBLIOGRAPHY 225

[Gro96] Gromov, M.: Systoles and intersystolic inequalities, Actes de la TableRonde de Geometrie Differentielle (Luminy, 1992), 291–362, Semin.Congr., 1, Soc. Math. France, Paris, 1996.www.emis.de/journals/SC/1996/1/ps/smf sem-cong 1 291-362.ps.gz

[Gro99] Gromov, M.: Metric structures for Riemannian and non-Riemannianspaces, Progr. in Mathematics, 152, Birkhauser, Boston, 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 PublishingCo., Amsterdam, 1987.

[Hat02] Hatcher, A.: Algebraic topology. Cambridge University Press, Cam-bridge, 2002.

[He86] Hebda, J. The collars of a Riemannian manifold and stable isosystolicinequalities. Pacific J. Math. 121 (1986), 339-356.

[Hel99] Helgason, S.: The Radon transform. 2nd edition. Progress in Mathemat-ics, 5. Birkhauser, 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 andGeometric Topology 4 (2004), 1103-1109.

[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.

[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.

[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

Page 226: Analytic and Differential geometry Mikhail G. Katz∗

226 BIBLIOGRAPHY

[Ka03] Katz, M.: Four-manifold systoles and surjectivity of period map, Com-ment. Math. Helv. 78 (2003), 772-786. See arXiv:math.DG/0302306

[Ka05] Katz, M.: Systolic inequalities and Massey products in simply-connectedmanifolds, Geometriae Dedicata, to appear.

[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.

[KL04] Katz, M.; Lescop, C.: Filling area conjecture, optimal systolic inequali-ties, and the fiber class in abelian covers. Proceedings of conference andworkshop 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 systoliccategory of low dimensional manifolds. Communications on Pure andApplied Mathematics, to appear. See arXiv:math.DG/0410456

[KR05] Katz, M.; Rudyak, Y.: Bounding volume by systoles of 3-manifolds. SeearXiv:math.DG/0504008

[KatzRS06] Katz, M.; Rudyak, Y.; Sabourau, S.: Systoles of 2-complexes, Reebgraph, 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 andasymptotic 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. SeearXiv:math.DG/0505007

[Ke74] Keisler, H. J.: Elementary calculus: An infinitesimal approach. Availableonline at http://www.math.wisc.edu/∼keisler/calc.html

[Kl82] Klein, F.: Uber Riemanns Theorie der algebraischen Funktionen undihrer Integrale. – Eine Erganzung der gewohnlichen Darstellungen.Leipzig: B.G. Teubner 1882.

[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 MatrixProceedings 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.

[La75] Lawson, H.B. The stable homology of a flat torus. Math. Scand. 36

(1975), 49–73.

Page 227: Analytic and Differential geometry Mikhail G. Katz∗

BIBLIOGRAPHY 227

[Leib] Leibniz, G. W.: La naissance du calcul differentiel. (French) [The birthof differential calculus] 26 articles des Acta Eruditorum. [26 articles ofthe Acta Eruditorum] Translated from the Latin and with an introduc-tion and notes by Marc Parmentier. With a preface by Michel Serres.Mathesis. Librairie Philosophique J. Vrin, Paris, 1989.

[Li69] Lichnerowicz, A.: Applications harmoniques dans un tore. C.R. Acad.Sci., Ser. A 269, 912–916 (1969).

[Lo71] Loewner, C.: Theory of continuous groups. Notes by H. Flanders andM. 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.

[Mi95] Miranda, R.: Algebraic curves and Riemann surfaces. Graduate Stud-ies in Mathematics, 5. American Mathematical Society, Providence, RI,1995.

[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.[MiTW73] Misner, Charles W.; Thorne, Kip S.; Wheeler, John Archibald: Gravi-

tation. W. H. Freeman and Co., San Francisco, Calif., 1973.[Mo86] Moser, J. Minimal solutions of variational problems on a torus. Ann.

Inst. H. Poincare-Analyse non lineaire 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 geodesicon a 2-dimansional sphere, Int. Math. Res. Not. 2002:23 (2002), 1211-1222.

[OR01] Oprea, J.; Rudyak, Y.: Detecting elements and Lusternik-Schnirelmanncategory of 3-manifolds. Lusternik-Schnirelmann category and relatedtopics (South Hadley, MA, 2001), 181–191, Contemp. Math., 316, Amer.Math. Soc., Providence, RI, 2002. arXiv:math.AT/0111181

[Pa04] Parlier, H.: Fixed-point free involutions on Riemann surfaces. SeearXiv: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, Hypotheses qui servent de fondement a la geometrie, inOeuvres mathematiques de Riemann. J. Gabay, 1990.

Page 228: Analytic and Differential geometry Mikhail G. Katz∗

228 BIBLIOGRAPHY

[RT90] Rosenbloom, M.; Tsfasman, M.: Multiplicative lattices in global fields.Invent. Math. 101 (1990), no. 3, 687–696.

[Ru87] Rudin, W.: Real and complex analysis. Third edition. McGraw-Hill BookCo., 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 singulieres 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 Konvexitat fur Variationsprobleme auf dem n-dimensionalen Torus. manuscripta math. 71 (1991), 45–65.

[Sen91b] Senn, W. Uber Mosers regularisiertes Variationsproblem fur mini-male Blatterungen 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 inducedby 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.

[Sp79] Spivak, M.: A comprehensive introduction to differential geometry.Vol. 2, chapter 4. Third edition. Publish or Perish, Inc., Wilmington,Del., 1999.

[St00] Streit, M.: Field of definition and Galois orbits for the Macbeath-Hurwitz curves. Arch. Math. (Basel) 74 (2000), no. 5, 342–349.

[Tr93] Tralle, A.: On compact homogeneous spaces with nonvanishing Masseyproducts, 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.

[2] Valenti, R.; Sebe, N.; Gevers, T. Combining head pose and eye locationinformation for gaze estimation. IEEE Trans. Image Process. 21 (2012),no. 2, 802–815.

[3] Valenti R.; Gevers T. Accurate Eye Center Location and Tracking UsingIsophote Curvature. 2014.

Page 229: Analytic and Differential geometry Mikhail G. Katz∗

BIBLIOGRAPHY 229

[Wa83] Warner, F.W.: Foundations of Differentiable Manifolds and LieGroups. 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. Bombieried.), 293-301. Annals of Mathematics Studies 103. Princeton UniversityPress, 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.


Recommended