GEOMETRY FINAL 1 (A): If M Is Non-Orientable and P ∈ M, Is M

Total Page:16

File Type:pdf, Size:1020Kb

GEOMETRY FINAL 1 (A): If M Is Non-Orientable and P ∈ M, Is M GEOMETRY FINAL CLAY SHONKWILER 1 (a): If M is non-orientable and p ∈ M, is M − {p} orientable? Answer: No. Suppose M − {p} is orientable, and let (Uα, xα) be an atlas that gives an orientation on M − {p}. Now, let (V, y) be a coordinate chart on M (in some atlas) containing the point p. If (U, x) is a coordinate chart such that U ∩ V 6= ∅, then, since x and y are both smooth maps with smooth inverses, the composition x ◦ y−1 : y(U ∩ V ) → x(U ∩ V ) is a smooth map. Hence, the collection {(Uα, xα), (V, α)} gives an atlas on M. Now, since M − {p} is orientable, i ! ∂xα det j > 0 ∂xβ for all α, β. However, since M is non-orientable, there must exist coordinate charts where the Jacobian is negative on the overlap; clearly, one of each such pair must be (V, y). Hence, there exists β such that ! ∂yi det j ≤ 0. ∂xβ Suppose this is true for all β such that Uβ ∩V 6= ∅. Then, since y(q) = (y1(q), . , yn(q)), lety ˜ be such thaty ˜(p) = (y2(p), y1(p), y3(p), y4(p), . , yn(p)). Theny ˜ : V → y˜(V ) is a diffeomorphism, so (V, y˜) gives a coordinate chart containing p and ! ! ∂y˜i ∂yi det j = − det j ≥ 0 ∂xβ ∂xβ for all β such that Uβ ∩ V 6= ∅. Hence, unless equality holds in some case, {(Uα, xα), (V, y˜)} defines an orientation on M, contradicting the fact that M is non-orientable. Additionally, it’s clear that equality can never hold, since y : Uβ ∩ V → y(Uβ ∩ V ) is a diffeomorphism and, thus, has full rank. 1 2 CLAY SHONKWILER Therefore, the above contradiction implies that there exist β, γ such that ! ! ∂yi ∂yi det j > 0 det j < 0. ∂xβ ∂xγ Let c be a path in V from q1 ∈ Uβ ∩ V to q2 ∈ Uγ ∩ V such that c does not pass through p. Then the Uα cover c and c is compact, so a finite subcollection U1,...,Uk covers c, where U1 = Uβ and ∂yi ∂yi Uk = Uγ. Since det j > 0 and det j < 0, there must exist ∂x1 ∂xk l ∈ {1, . , k} such that ! ! ∂yi ∂yi det j > 0 det j < 0. ∂xl ∂xl+1 Now, W = Ul ∩ Ul+1 ∩ V 6= ∅, since Ul and Ul+1 overlap on some segment of c, which is contained in V . Note that i ! i ! i ! ∂y ∂y ∂xl j = j j ∂xl+1 ∂xl ∂xl+1 by the chain rule, so i ! i ! i ! ∂y ∂y ∂xl 0 > det j = det j · det j > 0. ∂xl+1 ∂xl ∂xl+1 From this contradiction, then, we conclude that M − {p} is not orientable. ♣ (b): Let f : M → N be a local diffeomorphism. If one of M or N is orientable, is the other orientable? Answer: If N is orientable, then so is M. If M is orientable, then N need not be. Suppose f : M → N is a local diffeomorphism and that N is orientable. Then, at each point p ∈ M, there exists a neighborhood Up of p such that f : Up → f(Up) is a diffeomorphism. Let (Vp, yp) be a coordinate chart containing the point f(p) and let 0 0 0 −1 0 Vp = f(Up) ∩ Vp. Then f(p) ∈ Vp ⊂ f(Up), so Up = f (Vp) is 0 0 0 a neighborhood of p such that f|Up : Up → Vp is a diffeomorphism. 0 0 n 0 0 0 Furthermore, yp|Vp ◦f|Up : Up → yp|Vp (Vp) ⊂ R is a diffeomorphism, since it is the composition of diffeomorphisms. 0 0 Now, suppose p, q ∈ M such that Up ∩ Uq 6= ∅. Then −1 −1 −1 −1 (yp|V 0 ◦ f|U 0 ) ◦ (yq|V 0 ◦ f|U 0 ) = yp|V 0 ◦ (f|U 0 ◦ f| 0 ) ◦ yq| 0 = yp|V 0 ◦ yq| 0 p p q q p p Uq Vq p Vq GEOMETRY FINAL 3 0 0 0 0 is a differentiable map from yq(Vp ∩Vq ) to yp(Vp ∩Vq ). Furthermore, since we can find such neighborhoods for each p ∈ M, [ 0 Up = M. p∈M 0 0 0 Therefore, (Up, yp|Vp ◦ f|Up ) defines a differentiable structure on M. 0 0 Finally, note that if Up ∩ Uq 6= ∅, then −1 −1 det(D((yp|V 0 ◦ f|U 0 ) ◦ (yq|V 0 ◦ f|U 0 ) )) = det(D(yp|V 0 ◦ yq| 0 )) > 0 p p q q p Vq since N is orientable. Therefore, M is orientable. On the other hand, if M is orientable and f : M → N is a bijective local diffeomorphism, then f −1 : N → M is a local diffeomorphism and we can apply the above result to see that N is orientable. However, if f is not surjective, then N is not necessarily orientable. To see why, let N be any non-orientable manifold. Let (U, x) be a coordinate chart containing some point p ∈ N. Then n x : U → x(U) ⊂ R n is a diffeomorphism; let M = x(U). Since M = x(U) ⊂ R is open, M is an orientable manifold and the map x−1 : M → N is a local diffeomorphism. However, despite M being orientable, N is not orientable. Furthermore, if f : M → N is not injective, then the orientabil- ity of M does not guarantee the orientability of N. For example, 2 2 2 consider the standard projection π : S → RP . Let M = S and 2 N = RP . Then M is orientable and N is not. Let (Uα, xα) be an oriented atlas on M, let p ∈ M and let (U, x) be a coordinate chart in this atlas containing p. Then, since N = S2/{±1}, we know that (π (U ), x ◦ π |−1) gives an atlas on N. Hence, both U and π(U) α α α α Uα are diffeomorphic to x(U) and so π|U : U → π(U) is a diffeomor- phism. Since our choice of p ∈ M was arbitrary, we see that such a neighborhood exists for all p ∈ M, so π is a local diffeomorphism. Therefore, we conclude that f : M → N being a local diffeomor- phism and M being orientable does not necessarily mean that N is orientable. ♣ 2 Let X be a vector field on a manifold M such that X(p) 6= 0 for all p ∈ M. Show that there exists a diffeomorphism f : M → M without any fixed points. 4 CLAY SHONKWILER 3 Let f : SL(2, R) → R where f(A) = tr(A). Find the regular values of f. What is the diffeomorphism type of f −1(q) for q regular? What does f −1(q) look like for q singular? a b Answer: Let ∈ SL(2, ). Then ad − bc = 1, so either a and d c d R are both non-zero or b and c are both non-zero. Let a b a b U = |a, d 6= 0 ,U = |b, c 6= 0 . 1 c d 2 c d 3 Define the map φ1 : U1 → R by a b 7→ (a, b, d). c d ad−1 3 Since b 6= 0, and ad−bc = 1, c = b , so we can define ψ1 : R −{(a, 0, d)} → SL(2, R) by a b (a, b, d) 7→ ad−1 . b d ad−1 Then det(ψ1(a, b, d)) = ad − b b = ad − (ad − 1) = 1 and a b φ1 ◦ ψ1(a, b, d) = φ1 ad−1 = (a, b, d) b d and a b a b ψ1 ◦ φ1 = ψ1(a, b, d) = ad−1 ; c d b d ad−1 since ad − bc = 1, we see that c = b . Therefore, φ1 and ψ1 are inverses; 4 since we can view SL(2, R) as a subset of R and φ1 and ψ1 are both clearly 4 3 smooth when viewed as maps between R and R − {(a, 0, d)}, we conclude 3 that, in fact, φ1 is a homeomorphism of U1 with R − {(a, 0, d)}. 3 Similarly, define the map φ2 : U2 → R by a b 7→ (a, b, c). c d 1+bc Then, since a 6= 0 and ad − bc = 1, d = a , so we can define ψ2 : 3 R − {(0, b, c)} → SL(2, R) by a b (a, b, c) 7→ 1+bc . c a Clearly, both φ2 and ψ2 are smooth. Furthermore, by a similar argument to 3 that given above, φ2 ◦ ψ2 = id and ψ2 ◦ φ2 = id, so φ2 : U2 → R − {(0, b, c)} is a homeomorphism. 3 Now, φ1(U1) ∩ φ2(U2) = R − ({(a, 0, d)} ∪ {(0, b, c)}), so we need to check −1 3 that φ1 ◦ φ2 is smooth on this overlap. To that end, let (a, b, c) ∈ R such GEOMETRY FINAL 5 that a, b 6= 0. Then −1 a b 1 + bc φ1 ◦ φ2 (a, b, c) = φ1 1+bc = a, b, . c a a Since a 6= 0, this is a smooth map. Similarly, −1 a b ad − 1 φ2 ◦ φ1 (a, b, c) = φ2 ad−1 = a, b, . b c b Since b 6= 0, this is also a smooth map. Therefore, we conclude that (Ui, φi) gives an atlas on SL(2, R) for i = 1, 2. Now, consider the trace map. If x1 6= 0, then x x tr ◦ φ−1(x , x , x ) = tr 1 2 = x + x , 1 1 2 3 x1x3−1 x 1 3 x2 3 so ∂(tr ◦ φ−1) ∂(tr ◦ φ−1) ∂(tr ◦ φ−1) D(tr) = 1 , 1 , 1 = (1, 0, 1), ∂x1 ∂x2 ∂x3 which has rank 1, so none of these points is critical.
Recommended publications
  • Recognizing Surfaces
    RECOGNIZING SURFACES Ivo Nikolov and Alexandru I. Suciu Mathematics Department College of Arts and Sciences Northeastern University Abstract The subject of this poster is the interplay between the topology and the combinatorics of surfaces. The main problem of Topology is to classify spaces up to continuous deformations, known as homeomorphisms. Under certain conditions, topological invariants that capture qualitative and quantitative properties of spaces lead to the enumeration of homeomorphism types. Surfaces are some of the simplest, yet most interesting topological objects. The poster focuses on the main topological invariants of two-dimensional manifolds—orientability, number of boundary components, genus, and Euler characteristic—and how these invariants solve the classification problem for compact surfaces. The poster introduces a Java applet that was written in Fall, 1998 as a class project for a Topology I course. It implements an algorithm that determines the homeomorphism type of a closed surface from a combinatorial description as a polygon with edges identified in pairs. The input for the applet is a string of integers, encoding the edge identifications. The output of the applet consists of three topological invariants that completely classify the resulting surface. Topology of Surfaces Topology is the abstraction of certain geometrical ideas, such as continuity and closeness. Roughly speaking, topol- ogy is the exploration of manifolds, and of the properties that remain invariant under continuous, invertible transforma- tions, known as homeomorphisms. The basic problem is to classify manifolds according to homeomorphism type. In higher dimensions, this is an impossible task, but, in low di- mensions, it can be done. Surfaces are some of the simplest, yet most interesting topological objects.
    [Show full text]
  • 2 Non-Orientable Surfaces §
    2 NON-ORIENTABLE SURFACES § 2 Non-orientable Surfaces § This section explores stranger surfaces made from gluing diagrams. Supplies: Glass Klein bottle • Scarf and hat • Transparency fish • Large pieces of posterboard to cut • Markers • Colored paper grid for making the room a gluing diagram • Plastic tubes • Mobius band templates • Cube templates from Exploring the Shape of Space • 24 Mobius Bands 2 NON-ORIENTABLE SURFACES § Mobius Bands 1. Cut a blank sheet of paper into four long strips. Make one strip into a cylinder by taping the ends with no twist, and make a second strip into a Mobius band by taping the ends together with a half twist (a twist through 180 degrees). 2. Mark an X somewhere on your cylinder. Starting at the X, draw a line down the center of the strip until you return to the starting point. Do the same for the Mobius band. What happens? 3. Make a gluing diagram for a cylinder by drawing a rectangle with arrows. Do the same for a Mobius band. 4. The gluing diagram you made defines a virtual Mobius band, which is a little di↵erent from a paper Mobius band. A paper Mobius band has a slight thickness and occupies a small volume; there is a small separation between its ”two sides”. The virtual Mobius band has zero thickness; it is truly 2-dimensional. Mark an X on your virtual Mobius band and trace down the centerline. You’ll get back to your starting point after only one trip around! 25 Multiple twists 2 NON-ORIENTABLE SURFACES § 5.
    [Show full text]
  • Equivariant Orientation Theory
    EQUIVARIANT ORIENTATION THEORY S.R. COSTENOBLE, J.P. MAY, AND S. WANER Abstract. We give a long overdue theory of orientations of G-vector bundles, topological G-bundles, and spherical G-fibrations, where G is a compact Lie group. The notion of equivariant orientability is clear and unambiguous, but it is surprisingly difficult to obtain a satisfactory notion of an equivariant orien- tation such that every orientable G-vector bundle admits an orientation. Our focus here is on the geometric and homotopical aspects, rather than the coho- mological aspects, of orientation theory. Orientations are described in terms of functors defined on equivariant fundamental groupoids of base G-spaces, and the essence of the theory is to construct an appropriate universal tar- get category of G-vector bundles over orbit spaces G=H. The theory requires new categorical concepts and constructions that should be of interest in other subjects, such as algebraic geometry. Contents Introduction 2 Part I. Fundamental groupoids and categories of bundles 5 1. The equivariant fundamental groupoid 5 2. Categories of G-vector bundles and orientability 6 3. The topologized fundamental groupoid 8 4. The topologized category of G-vector bundles over orbits 9 Part II. Categorical representation theory and orientations 11 5. Bundles of Groupoids 11 6. Skeletal, faithful, and discrete bundles of groupoids 13 7. Representations and orientations of bundles of groupoids 16 8. Saturated and supersaturated representations 18 9. Universal orientable representations 23 Part III. Examples of universal orientable representations 26 10. Cyclic groups of prime order 26 11. Orientations of V -dimensional G-bundles 30 12.
    [Show full text]
  • ON a THEOREM by FARB and MASUR Let Σg Be the Closed
    PROCEEDINGS OF THE AMERICAN MATHEMATICAL SOCIETY Volume 128, Number 12, Pages 3463{3464 S 0002-9939(00)05450-2 Article electronically published on June 7, 2000 ON A THEOREM BY FARB AND MASUR KOJI FUJIWARA (Communicated by Ronald A. Fintushel) Abstract. Farb and Masur showed that an irreducible lattice in a semisimple Lie group of rank at least two always has finite image by a homomorphism into the outer automorphism group of a closed, orientable surface group. We point out that their theorem extends to the outer automorphism groups of a certain class of torsion-free, freely indecomposable word-hyperbolic groups. Let Σg be the closed, orientable surface of genus g and let Mod(Σg)denotethe mapping class group of Σg. Farb and Masur proved the following in [FM]. Theorem 1 (Farb-Masur). Let Γ be an irreducible lattice in a semisimple Lie group of rank at least two. Let h :Γ! Mod(Σg) be a homomorphism. Then Im(h) is finite. Their proof uses a deep work on the Poisson boundary of mapping class groups by Kaimanovich-Masur [KM]. Let G be a torsion-free, freely indecomposable word-hyperbolic group [G]. By Sela, there is a graph of groups decomposition of G such that the edge groups are isomorphic to Z and the vertex groups are in two families; one is a collection of the fundamental groups of punctured surfaces fSigi, which are called CMQ subgroups, and the other is rigid in the sense that the outer automorphism groups are finite. This graph decomposition, which is called JSJ-decomposition of G, is unique in a certain sense; in particular the set fSigi is an invariant of G (Theorem 1.7 of [S]).
    [Show full text]
  • LECTURE 23: the STOKES FORMULA 1. Volume Forms Last
    LECTURE 23: THE STOKES FORMULA 1. Volume Forms Last time we introduced the conception of orientability via local charts. Here is a criteria for the orientability via top forms: Theorem 1.1. An n-dimensional smooth manifold M is orientable if and only if M admits a nowhere vanishing smooth n-form µ. Proof. First let µ be a nowhere vanishing smooth n-form on M. Then on each local chart (U; x1; ··· ; xn) (where U is always chosen to be connected), there is a smooth function f 6= 0 so that µ = fdx1 ^ · · · ^ dxn. It follows that µ(@1; ··· ;@n) = f 6= 0: We can always take such a chart near each point so that f > 0, otherwise we can replace x1 1 1 n 1 n by −x . Now suppose (Uα; xα; ··· ; xα) and (Uβ; xβ; ··· ; xβ) be two such charts, so that on the intersection Uα \ Uβ one has 1 n 1 n µ = fdxα ^ · · · ^ dxα = gdxβ ^ · · · ^ dxβ; where f; g > 0: Then on the intersection Uα \ Uβ, β β α n 0 < g = µ(@1 ; ··· ;@n ) = (det d'αβ)µ(@1 ; ··· ;@α) = (det d'αβ)f: It follows that det (d'αβ) > 0. So the atlas constructed by this way is an orientation. Conversely, suppose A is an orientation. For each local chart Uα in A, we let 1 n µα = dxα ^ · · · ^ dxα: Pick a partition of unity fραg subordinate to the open cover fUαg. We claim that X µ := ραµα α is a nowhere vanishing smooth n-form on M. In fact, for each p 2 M, there is a neighborhood U P Pk of p so that the sum α ραµα is a finite sum i=1 ρiµi.
    [Show full text]
  • Orientability of Manifolds
    Orientability of Manifolds Arlo Caine April 23, 2002 Abstract This brief discussion of orientability for differentiable manifolds fol- lows a manuscript of lecture notes by Prof. Peter Petersen, UCLA. The presentation uses the point of view of covering spaces. In this talk we will explore this viewpoint and establish its equivalence with the viewpoint we discussed last semester. Let us first review some linear algebra. Let B1 and B2 be two ordered bases of a finite dimensional vector space V . Definition: B1 and B2 represent the same orientation for V if the transition matrix M21 from B1 to B2 has positive determinant. This relation, call it ∼, on the set of all ordered bases of V is clearly reflexive −1 −1 since M11 = I, and symmetric since det M12 = det M21 = (det M21) > 0. If B3 is a third ordered basis for V and B1 ∼ B2 and B2 ∼ B3 then B1 ∼ B3 since det M31 = det(M32M21) = det M32 · det M21. So ∼ gives an equivalence relation; furthermore, it has exactly two equivalence classes. Definition: A choice of such an equivalence class is called an orientation for the vector space V . If M is a smooth manifold of dimension n then for each p ∈ M the tangent space TpM is a vector space of dimension n, and hence has two choices of ori- entation. We would like to use this scenario to construct a two-sheeted covering 1 n space OM called the orientation covering of M. If (x , ..., x ) are coordinates on a connected neighborhood U ⊂ M, then the set {∂1|p, ..., ∂n|p} form a basis for the tangent space at each p ∈ M, where ∂ ∂i|p = i .
    [Show full text]
  • Arxiv:1909.13149V1 [Math.DS] 28 Sep 2019 Ora Fdnmcladcnrlsses O.1,N.1 Jan 1, No
    Journal of Dynamical and Control Systems, Vol. 18, No. 1, January 2012, 21–33 ( c 2012) MORSE-SMALE SURFACED DIFFEOMORPHISMS WITH ORIENTABLE HETEROCLINIC A. MOROZOV and O. POCHINKA Abstract. In the present paper we consider preserving orientation Morse-Smale diffeomorphisms on surfaces. Using the methods of fac- torization and linearizing neighborhoods we prove that such diffeo- morphisms have a finite number of orientable heteroclinic orbits. 1. Introduction The Morse-Smale dynamics is a pattern of a regular structurally stable (rough) behaviour. The foundations of the theory of roughness of flows, laid down in the classic work by A. Andronov and L. Pontryagin [1], were developed by the associates of academician A. Andronov in the Gorky school of non-linear oscillations E. Leontovich and A. Mayer [12], [13]. Moreover, A. Meyer introduced the concept of roughness for a discrete dynamical system and obtained a complete topological classification of such systems (Morse-Smale systems) on the circle [14]. Since then, the theory of topological classification of Morse-Smale dy- namic systems has gained wide popularity and has undergone intensive de- velopment (see, for example, review [8]). So the topological classification of structurally stable flows on surfaces is exhaustively described in the work by M. Peixoto [16]. On three-dimensional manifolds, the necessary and suf- ficient conditions of the topological conjugacy of Morse-Smale flows follows from the works by Ya. Umanskiy [20] and A. Prishlyak [18]. There are arXiv:1909.13149v1 [math.DS] 28 Sep 2019 also multidimensional classification results, for example, the classification by S. Pilyugin [17] for Morse-Smale flows without heteroclinic intersections on the n-dimensional sphere.
    [Show full text]
  • The Topology of Fiber Bundles Lecture Notes Ralph L. Cohen
    The Topology of Fiber Bundles Lecture Notes Ralph L. Cohen Dept. of Mathematics Stanford University Contents Introduction v Chapter 1. Locally Trival Fibrations 1 1. Definitions and examples 1 1.1. Vector Bundles 3 1.2. Lie Groups and Principal Bundles 7 1.3. Clutching Functions and Structure Groups 15 2. Pull Backs and Bundle Algebra 21 2.1. Pull Backs 21 2.2. The tangent bundle of Projective Space 24 2.3. K - theory 25 2.4. Differential Forms 30 2.5. Connections and Curvature 33 2.6. The Levi - Civita Connection 39 Chapter 2. Classification of Bundles 45 1. The homotopy invariance of fiber bundles 45 2. Universal bundles and classifying spaces 50 3. Classifying Gauge Groups 60 4. Existence of universal bundles: the Milnor join construction and the simplicial classifying space 63 4.1. The join construction 63 4.2. Simplicial spaces and classifying spaces 66 5. Some Applications 72 5.1. Line bundles over projective spaces 73 5.2. Structures on bundles and homotopy liftings 74 5.3. Embedded bundles and K -theory 77 5.4. Representations and flat connections 78 Chapter 3. Characteristic Classes 81 1. Preliminaries 81 2. Chern Classes and Stiefel - Whitney Classes 85 iii iv CONTENTS 2.1. The Thom Isomorphism Theorem 88 2.2. The Gysin sequence 94 2.3. Proof of theorem 3.5 95 3. The product formula and the splitting principle 97 4. Applications 102 4.1. Characteristic classes of manifolds 102 4.2. Normal bundles and immersions 105 5. Pontrjagin Classes 108 5.1. Orientations and Complex Conjugates 109 5.2.
    [Show full text]
  • Math 396. Orientations on Bundles and Manifolds 1. Orientation Atlases and Orientation Forms Recall That an Oriented Trivializin
    Math 396. Orientations on bundles and manifolds 1. Orientation atlases and orientation forms Recall that an oriented trivializing atlas for E → M is a trivializing atlas ni {(Ui, φi : E|Ui ' Ui × R )} such that whenever Ui ∩ Uj is non-empty (so ni = nj) the ordered bases of E(m) induced by φi and φj for each m ∈ Ui ∩ Uj lie in the same orientation class (i.e., the change of basis matrix for the m-fiber has positive determinant). In other words, a trivializing atlas is oriented if it gives a well-defined orientation on each fiber E(m) (a non-trivial condition only for those m in a double overlap Ui ∩ Uj for i 6= j). Thus, when given an oriented trivializing atlas we get an orientation on every E(m). (Note that this would not make sense if we allowed E(m) = 0.) We say that two oriented trivializing atlases are equivalent if, for each m ∈ M, they put the same orientation on the fiber E(m). As we saw in class, this really is an equivalence relation, and in each equivalence class there is a unique maximal element (in the sense of containing all trivializing frames from all oriented atlases in the equivalence class). These maximal elements are called orientation atlases for E → M. Thus, two orientation atlases are equal if and only if they define the same orientation on each fiber E(m) (as equivalence is the same as equality for the maximal elements). We define an orientation form on E → M to be a nowhere-vanishing global section ω of det E, which is to say a trivializing section of the line bundle det E.
    [Show full text]
  • Anubhav Nanavaty
    AN INTRODUCTION TO DIFFERENTIAL FORMS, STOKES' THEOREM AND GAUSS-BONNET THEOREM ANUBHAV NANAVATY Abstract. This paper serves as a brief introduction to differential geome- try. It first discusses the language necessary for the proof and applications of a powerful generalization of the fundamental theorem of calculus, known as n 3 Stokes' Theorem in R . Further, geometry in R will be discussed to present Chern's proof of the Poincar´e-HopfIndex Theorem and Gauss-Bonnet The- 3 orem in R , both of which relate topological properties of a manifold to its geometric properties. Only a working knowledge of multivariable calculus is needed to understand this paper. All other concepts are introduced and dis- cussed. Contents 1. Introduction 1 2. Differential Forms and Manifolds 2 2.1. Differentiable Manifolds 2 2.2. Differentiable Forms 4 2.3. Exterior Derivatives 6 2.4. Integration of Forms 8 3. Stokes' Theorem 10 3.1. Applications 13 4. Riemannian Manifolds and Geometry in R3 14 4.1. Cartan's Structure Equations in Rn 15 4.2. Curvature in R3 18 5. The Gauss-Bonnet Theorem 20 6. Conclusion 26 Acknowledgments 26 References 26 1. Introduction We first introduce the concept of a manifold, which leads to a discussion of differential forms, the exterior derivative and pull-back map. We then discuss integration of forms in Rn in order to state and prove Stokes' Theorem in Rn. A few applications of Stokes' Theorem are also stated and proved, such as Brouwer's fixed point theorem. In order to discuss Chern's proof of the Gauss-Bonnet Theorem in R3, we slightly shift gears to discuss geometry in R3.
    [Show full text]
  • Arxiv:2105.07559V1 [Math.GT] 17 May 2021 Ooinal Ufcs W-Ie Curves
    RIGHT-ANGLED ARTIN GROUPS AND CURVE GRAPHS OF NONORIENTABLE SURFACES TAKUYA KATAYAMA AND ERIKA KUNO Abstract. Let N be a closed nonorientable surface with or without marked points. In this paper we prove that, for every finite full subgraph Γ of Ctwo(N), the right-angled Artin group on Γ can be embedded in the mapping class group of N. Here, Ctwo(N) is the subgraph, induced by essential two-sided simple closed curves in N, of the ordinal curve graph C(N). In addition, we show that there exists a finite graph Γ which is not a full subgraph of Ctwo(N) for some N, but the right-angled Artin group on Γ can be embedded in the mapping class group of N. 1. Introduction Let F be a closed orientable or nonorientable surface with or without marked points. When we are concerned with the orientability and topological type of a surface, by Sg,m and Ng,m we denote a closed orientable surface of genus g with m marked points and closed nonorientable surface of genus g with m marked points, respectively. We also write S and N simply for a closed orientable surface and closed non-orientable surface, respectively. Throughout this paper a curve on F means a simple closed curve on F . A curve c on F is essential if c bounds neither a disk nor a disk with one marked point, nor a M¨obius band. We denote by C(F ) the curve graph of F , which is the simplicial graph that the vertex set consists of the isotopy classes of all essential simple closed curves on F and the edge set consists of all non-ordered pair of essential simple closed curves which can be represented disjointly.
    [Show full text]
  • Problems (To Be Handed In)
    MATH 4030 Differential Geometry Homework 5 due 11/10/2015 (Tue) at 5PM Problems (to be handed in) n Unless otherwise stated, we use U; O to denote connected open subsets of R . The symbols S; S1;S2;S3 3 always denote surfaces in R . Recall that a critical point of a smooth function f : S ! R is a point p 2 S such that dfp = 0. 0 1. Let f; g : R ! R be smooth functions such that f(u) > 0 and g (u) > 0 for all u 2 R. Define a 3 smooth map X : R × (−π; π) ! R by X(u; v) := (f(u) cos v; f(u) sin v; g(u)): Show that the image S := X(R × (−π; π)) is a surface, and that all the normal lines of S pass through the z-axis. What does the surface S look like? 3 2. Let S ⊂ R be a surface. 3 (a) Fix a point p0 2 R such that p0 2= S, show that the distance function from p0 f(p) := jp − p0j defines a smooth function f : S ! R. Moreover, prove that p 2 S is a critical point of f if and only if the line joining p to p0 is normal to S at p. 3 (b) Fix a unit vector v 2 R , show that the height function along v h(p) := hp; vi defines a smooth function h : S ! R. Prove that p 2 S is a critical point of h if and only if v is normal to S at p.
    [Show full text]