Viewing the Real Projective Plane in R3 ; the Cross-Cap and the Steiner

Total Page:16

File Type:pdf, Size:1020Kb

Viewing the Real Projective Plane in R3 ; the Cross-Cap and the Steiner Appendix A Viewing the Real Projective Plane in R3;The Cross-Cap and the Steiner Roman Surface It turns out that there are several ways of viewing the real projective plane in R3 as a surface with self-intersection. Recall that, as a topological space, the projective plane, RP2, is the quotient of the 2-sphere, S 2,(inR3) by the equivalence relation that identifies antipodal points. In Hilbert and Cohn–Vossen [2](andalsodoCarmo [1]) an interesting map, H , from R3 to R4 is defined by .x; y; z/ 7! .xy; yz;xz;x2 y2/: This map has the remarkable property that when restricted to S 2,wehave H .x; y; z/ D H .x0;y0; z0/ iff .x0;y0; z0/ D .x; y; z/ or .x0;y0; z0/ D .x;y; z/. In other words, the inverse image of every point in H .S 2/ consists of two antipodal points. Thus, the map H induces an injective map from the projective plane onto H .S 2/, which is obviously continuous, and since the projective plane is compact, it is a homeomorphism. Therefore, the map H allows us to realize concretely the projective plane in R4 by choosing any parametrization of the sphere, S 2,and applying the map H to it. Actually, it turns out to be more convenient to use the map A defined such that .x; y; z/ 7! .2xy; 2yz;2xz;x2 y2/; because it yields nicer parametrizations. For example, using the stereographic representation of S 2 where 2u x.u; v/ D ; u2 C v2 C 1 2v y.u; v/ D ; u2 C v2 C 1 u2 C v2 1 z.u; v/ D ; u2 C v2 C 1 J. Gallier and D. Xu, A Guide to the Classification Theorem for Compact Surfaces, 105 Geometry and Computing 9, DOI 10.1007/978-3-642-34364-3, © Springer-Verlag Berlin Heidelberg 2013 106 A Viewing the Real Projective Plane in R3; The Cross-Cap and the Steiner Roman . Fig. A.1 Jakob Steiner, 1796–1863 we obtain the following four fractions parametrizing the projective plane in R4: 8uv x.u; v/ D ; .u2 C v2 C 1/2 4v.u2 C v2 1/ y.u; v/ D ; .u2 C v2 C 1/2 4u.u2 C v2 1/ z.u; v/ D ; .u2 C v2 C 1/2 4.u2 v2/ t.u; v/ D : .u2 C v2 C 1/2 Of course, we don’t know how to visualize this surface in R4, but we can visualize its four projections in R3, using parallel projections with respect to the four axes, which amounts to dropping one of the four coordinates. The resulting surfaces turn out to be very interesting. Only two distinct surfaces are obtained, both very well known to topologists. Indeed, the surface obtained by dropping y or z is known the cross-cap surface, and the surface obtained by dropping x or t is known as the Steiner roman surface. These surfaces can be easily displayed. We begin with the Steiner surface named after Jakob Steiner (Fig. A.1). 1. The Steiner roman surface. This is the surface obtained by dropping t. Going back to the map A and renaming x;y;z as ˛; ˇ; ,if x D 2˛ˇ; y D 2ˇ; z D 2˛; t D ˛2 ˇ2; it is easily seen that A Viewing the Real Projective Plane in R3; The Cross-Cap and the Steiner Roman . 107 xy D 2zˇ2; yz D 2x 2; xz D 2y˛2: If .˛;ˇ;/is on the sphere, we have ˛2 Cˇ2 C 2 D 1, and thus we get the following implicit equation for the Steiner roman surface: x2y2 C y2z2 C x2z2 D 2xyz: It is easily verified that the following parametrization works: 2v x.u; v/ D ; u2 C v2 C 1 2u y.u; v/ D ; u2 C v2 C 1 2uv z.u; v/ D : u2 C v2 C 1 Thus, amazingly, the Steiner roman surface (Fig. A.2) can be specified by fractions of quadratic polynomials! It can be shown that every quadric surface can be defined as a rational surface of degree 2, but other surfaces can also be defined, as shown by the Steiner roman surface. It can be shown that the Steiner roman surface is contained inside the tetrahedron defined by the planes x C y C z D 1; x y C z D 1; x C y z D 1; x y z D 1; with 1 Ä x;y;z Ä 1. The surface touches these four planes along ellipses, and at the middle of the six edges of the tetrahedron, it has sharp edges. Furthermore, the surface is self-intersecting along the axes, and has four closed chambers. A more extensive discussion can be found in Hilbert and Cohn–Vossen [2], in particular, its relationship to the heptahedron. One view of the surface consisting of six patches is shown in Fig. A.2. Patches 1 and 2 are colored blue, patches 3 and 4 are colored red, and patches 5 and 6 are colored green. A closer look reveals that the three colored patches are identical under appropriate rigid motions, and fit perfectly. Another revealing view is obtained by cutting off a top portion of the surface (Fig. A.3). 108 A Viewing the Real Projective Plane in R3; The Cross-Cap and the Steiner Roman . Fig. A.2 The Steiner roman 1 y surface 0.5 0 -0.5-1 1 0.5 0 z -0.5 -1 1 0 0.5 -1 -0.5 x Fig. A.3 A cut of the Steiner 1 roman surface 0.5 y 0 -0.5 -1 0.5 0 z -0.5 -1 -1 -0.5 0 x 0.5 1 In this picture, it is clear that the surface has chambers. We now consider the cross-cap surface. 2. The cross-cap surface. This is the surface obtained by dropping either the y coordinate, or the z coordinate. Let us first consider the surface obtained by dropping y. Its implicit A Viewing the Real Projective Plane in R3; The Cross-Cap and the Steiner Roman . 109 equation is obtained by eliminating ˛; ˇ; in the equations x D 2˛ˇ; z D 2˛; t D ˛2 ˇ2; and ˛2 C ˇ2 C 2 D 1. We leave as an exercise to show that we get .2x2 C z2/2 D 4.x2 C t.x2 C z2//.1 t/: If we now consider the surface obtained by dropping z, the implicit equation is obtained by eliminating ˛; ˇ; in the equations x D 2˛ˇ; y D 2ˇ; t D ˛2 ˇ2; and ˛2 C ˇ2 C 2 D 1. We leave as an exercise to show that we get .2x2 C y2/2 D 4.x2 t.x2 C y2//.1 C t/: Note that the second implicit equation is obtained from the first by substituting y for z and t for t. This shows that the two implicit equations define the same surface. An explicit parametrization of the surface is obtained by dropping z: 8uv x.u; v/ D ; .u2 C v2 C 1/2 4v.u2 C v2 1/ y.u; v/ D ; .u2 C v2 C 1/2 4.u2 v2/ z.u; v/ D ; .u2 C v2 C 1/2 One view of the surface obtained by cutting off part of its top part to have a better view of the self intersection, is shown Fig. A.4. 3: The Steiner roman surface, again. The last projection of the projective plane is obtained by dropping the x coordinate. Its implicit equation is obtained by eliminating ˛; ˇ; in the equations y D 2ˇ; z D 2˛; t D ˛2 ˇ2; and ˛2 C ˇ2 C 2 D 1. We leave as an exercise to show that we get 4.y2 C z2/t 2 D .z2 y2/.y2 z2 C 4t/: 110 A Viewing the Real Projective Plane in R3; The Cross-Cap and the Steiner Roman . Fig. A.4 A cut of the 1 cross-cap surface y 0.5 0 -0.5 -1 0.5 0 z -0.5 -1 -1 -0.5 0 x 0.5 1 This time, it is not so obvious that it corresponds to the Steiner roman surface. However, if we perform the rotation of the y; z plane by =4,wehave p p 2 2 y D Y Z; 2 2 p p 2 2 z D Y C Z; 2 2 and we have y2 C z2 D Y 2 C Z2 and z2 y2 D 2YZ. Thus, the implicit equation becomes 4.Y 2 C Z2/t 2 D 2YZ.2YZ C 4t/; which simplifies to Y 2t 2 C Z2t 2 C Y 2Z2 D 2YZt; which is exactly the equation of the Steiner roman surface. Just for fun, we also get the parametrization 4v.u2 C v2 1/ x.u; v/ D ; .u2 C v2 C 1/2 4u.u2 C v2 1/ y.u; v/ D ; .u2 C v2 C 1/2 4.u2 v2/ z.u; v/ D : .u2 C v2 C 1/2 A Viewing the Real Projective Plane in R3; The Cross-Cap and the Steiner Roman . 111 Fig. A.5 Another view of the 1 Steiner roman surface y 0.5 0 -0.5 -1 1 0.5 z 0 -0.5 -1 -1 -0.5 0 x 0.5 1 Fig. A.6 A cut of the Steiner 1 roman surface y 0.5 0 -0.5 -1 0.5 0 z -0.5 -1 -1 -0.5 0 x 0.5 1 This Steiner roman surface displayed in Fig.
Recommended publications
  • Simple Infinite Presentations for the Mapping Class Group of a Compact
    SIMPLE INFINITE PRESENTATIONS FOR THE MAPPING CLASS GROUP OF A COMPACT NON-ORIENTABLE SURFACE RYOMA KOBAYASHI Abstract. Omori and the author [6] have given an infinite presentation for the mapping class group of a compact non-orientable surface. In this paper, we give more simple infinite presentations for this group. 1. Introduction For g ≥ 1 and n ≥ 0, we denote by Ng,n the closure of a surface obtained by removing disjoint n disks from a connected sum of g real projective planes, and call this surface a compact non-orientable surface of genus g with n boundary components. We can regard Ng,n as a surface obtained by attaching g M¨obius bands to g boundary components of a sphere with g + n boundary components, as shown in Figure 1. We call these attached M¨obius bands crosscaps. Figure 1. A model of a non-orientable surface Ng,n. The mapping class group M(Ng,n) of Ng,n is defined as the group consisting of isotopy classes of all diffeomorphisms of Ng,n which fix the boundary point- wise. M(N1,0) and M(N1,1) are trivial (see [2]). Finite presentations for M(N2,0), M(N2,1), M(N3,0) and M(N4,0) ware given by [9], [1], [14] and [16] respectively. Paris-Szepietowski [13] gave a finite presentation of M(Ng,n) with Dehn twists and arXiv:2009.02843v1 [math.GT] 7 Sep 2020 crosscap transpositions for g + n > 3 with n ≤ 1. Stukow [15] gave another finite presentation of M(Ng,n) with Dehn twists and one crosscap slide for g + n > 3 with n ≤ 1, applying Tietze transformations for the presentation of M(Ng,n) given in [13].
    [Show full text]
  • An Introduction to Topology the Classification Theorem for Surfaces by E
    An Introduction to Topology An Introduction to Topology The Classification theorem for Surfaces By E. C. Zeeman Introduction. The classification theorem is a beautiful example of geometric topology. Although it was discovered in the last century*, yet it manages to convey the spirit of present day research. The proof that we give here is elementary, and its is hoped more intuitive than that found in most textbooks, but in none the less rigorous. It is designed for readers who have never done any topology before. It is the sort of mathematics that could be taught in schools both to foster geometric intuition, and to counteract the present day alarming tendency to drop geometry. It is profound, and yet preserves a sense of fun. In Appendix 1 we explain how a deeper result can be proved if one has available the more sophisticated tools of analytic topology and algebraic topology. Examples. Before starting the theorem let us look at a few examples of surfaces. In any branch of mathematics it is always a good thing to start with examples, because they are the source of our intuition. All the following pictures are of surfaces in 3-dimensions. In example 1 by the word “sphere” we mean just the surface of the sphere, and not the inside. In fact in all the examples we mean just the surface and not the solid inside. 1. Sphere. 2. Torus (or inner tube). 3. Knotted torus. 4. Sphere with knotted torus bored through it. * Zeeman wrote this article in the mid-twentieth century. 1 An Introduction to Topology 5.
    [Show full text]
  • Equivelar Octahedron of Genus 3 in 3-Space
    Equivelar octahedron of genus 3 in 3-space Ruslan Mizhaev ([email protected]) Apr. 2020 ABSTRACT . Building up a toroidal polyhedron of genus 3, consisting of 8 nine-sided faces, is given. From the point of view of topology, a polyhedron can be considered as an embedding of a cubic graph with 24 vertices and 36 edges in a surface of genus 3. This polyhedron is a contender for the maximal genus among octahedrons in 3-space. 1. Introduction This solution can be attributed to the problem of determining the maximal genus of polyhedra with the number of faces - . As is known, at least 7 faces are required for a polyhedron of genus = 1 . For cases ≥ 8 , there are currently few examples. If all faces of the toroidal polyhedron are – gons and all vertices are q-valence ( ≥ 3), such polyhedral are called either locally regular or simply equivelar [1]. The characteristics of polyhedra are abbreviated as , ; [1]. 2. Polyhedron {9, 3; 3} V1. The paper considers building up a polyhedron 9, 3; 3 in 3-space. The faces of a polyhedron are non- convex flat 9-gons without self-intersections. The polyhedron is symmetric when rotated through 1804 around the axis (Fig. 3). One of the features of this polyhedron is that any face has two pairs with which it borders two edges. The polyhedron also has a ratio of angles and faces - = − 1. To describe polyhedra with similar characteristics ( = − 1) we use the Euler formula − + = = 2 − 2, where is the Euler characteristic. Since = 3, the equality 3 = 2 holds true.
    [Show full text]
  • INTRODUCTION to ALGEBRAIC GEOMETRY, CLASS 25 Contents 1
    INTRODUCTION TO ALGEBRAIC GEOMETRY, CLASS 25 RAVI VAKIL Contents 1. The genus of a nonsingular projective curve 1 2. The Riemann-Roch Theorem with Applications but No Proof 2 2.1. A criterion for closed immersions 3 3. Recap of course 6 PS10 back today; PS11 due today. PS12 due Monday December 13. 1. The genus of a nonsingular projective curve The definition I’m going to give you isn’t the one people would typically start with. I prefer to introduce this one here, because it is more easily computable. Definition. The tentative genus of a nonsingular projective curve C is given by 1 − deg ΩC =2g 2. Fact (from Riemann-Roch, later). g is always a nonnegative integer, i.e. deg K = −2, 0, 2,.... Complex picture: Riemann-surface with g “holes”. Examples. Hence P1 has genus 0, smooth plane cubics have genus 1, etc. Exercise: Hyperelliptic curves. Suppose f(x0,x1) is a polynomial of homo- geneous degree n where n is even. Let C0 be the affine plane curve given by 2 y = f(1,x1), with the generically 2-to-1 cover C0 → U0.LetC1be the affine 2 plane curve given by z = f(x0, 1), with the generically 2-to-1 cover C1 → U1. Check that C0 and C1 are nonsingular. Show that you can glue together C0 and C1 (and the double covers) so as to give a double cover C → P1. (For computational convenience, you may assume that neither [0; 1] nor [1; 0] are zeros of f.) What goes wrong if n is odd? Show that the tentative genus of C is n/2 − 1.(Thisisa special case of the Riemann-Hurwitz formula.) This provides examples of curves of any genus.
    [Show full text]
  • Homeomorphisms of the 3-Sphere That Preserve a Heegaard Splitting of Genus Two
    PROCEEDINGS OF THE AMERICAN MATHEMATICAL SOCIETY Volume 136, Number 3, March 2008, Pages 1113–1123 S 0002-9939(07)09188-5 Article electronically published on November 30, 2007 HOMEOMORPHISMS OF THE 3-SPHERE THAT PRESERVE A HEEGAARD SPLITTING OF GENUS TWO SANGBUM CHO (Communicated by Daniel Ruberman) Abstract. Let H2 be the group of isotopy classes of orientation-preserving homeomorphisms of S3 that preserve a Heegaard splitting of genus two. In this paper, we construct a tree in the barycentric subdivision of the disk complex of a handlebody of the splitting to obtain a finite presentation of H2. 1. Introduction Let Hg be the group of isotopy classes of orientation-preserving homeomorphisms of S3 that preserve a Heegaard splitting of genus g,forg ≥ 2. It was shown by Goeritz [3] in 1933 that H2 is finitely generated. Scharlemann [7] gave a modern proof of Goeritz’s result, and Akbas [1] refined this argument to give a finite pre- sentation of H2. In arbitrary genus, first Powell [6] and then Hirose [4] claimed to have found a finite generating set for the group Hg, though serious gaps in both ar- guments were found by Scharlemann. Establishing the existence of such generating sets appears to be an open problem. In this paper, we recover Akbas’s presentation of H2 by a new argument. First, we define the complex P (V ) of primitive disks, which is a subcomplex of the disk complex of a handlebody V in a Heegaard splitting of genus two. Then we find a suitable tree T ,onwhichH2 acts, in the barycentric subdivision of P (V )toget a finite presentation of H2.
    [Show full text]
  • Â-Genus and Collapsing
    The Journal of Geometric Analysis Volume 10, Number 3, 2000 A A-Genus and Collapsing By John Lott ABSTRACT. If M is a compact spin manifold, we give relationships between the vanishing of A( M) and the possibility that M can collapse with curvature bounded below. 1. Introduction The purpose of this paper is to extend the following simple lemma. Lemma 1.I. If M is a connectedA closed Riemannian spin manifold of nonnegative sectional curvature with dim(M) > 0, then A(M) = O. Proof Let K denote the sectional curvature of M and let R denote its scalar curvature. Suppose that A(M) (: O. Let D denote the Dirac operator on M. From the Atiyah-Singer index theorem, there is a nonzero spinor field 7t on M such that D~p = 0. From Lichnerowicz's theorem, 0 = IDOl 2 dvol = IV~Pl 2 dvol + ~- I•12 dvol. (1.1) From our assumptions, R > 0. Hence V~p = 0. This implies that I~k F2 is a nonzero constant function on M and so we must also have R = 0. Then as K > 0, we must have K = 0. This implies, from the integral formula for A'(M) [14, p. 231], that A"(M) = O. [] The spin condition is necessary in Lemma 1.l, as can be seen in the case of M = CP 2k. The Ricci-analog of Lemma 1.1 is false, as can be seen in the case of M = K3. Definition 1.2. A connected closed manifold M is almost-nonnegatively-curved if for every E > 0, there is a Riemannian metric g on M such that K(M, g) diam(M, g)2 > -E.
    [Show full text]
  • 1 University of California, Berkeley Physics 7A, Section 1, Spring 2009
    1 University of California, Berkeley Physics 7A, Section 1, Spring 2009 (Yury Kolomensky) SOLUTIONS TO THE SECOND MIDTERM Maximum score: 100 points 1. (10 points) Blitz (a) hoop, disk, sphere This is a set of simple questions to warm you up. The (b) disk, hoop, sphere problem consists of five questions, 2 points each. (c) sphere, hoop, disk 1. Circle correct answer. Imagine a basketball bouncing on a court. After a dozen or so (d) sphere, disk, hoop bounces, the ball stops. From this, you can con- (e) hoop, sphere, disk clude that Answer: (d). The object with highest moment of (a) Energy is not conserved. inertia has the largest kinetic energy at the bot- (b) Mechanical energy is conserved, but only tom, and would reach highest elevation. The ob- for a single bounce. ject with the smallest moment of inertia (sphere) (c) Total energy in conserved, but mechanical will therefore reach lowest altitude, and the ob- energy is transformed into other types of ject with the largest moment of inertia (hoop) energy (e.g. heat). will be highest. (d) Potential energy is transformed into kinetic 4. A particle attached to the end of a massless rod energy of length R is rotating counter-clockwise in the (e) None of the above. horizontal plane with a constant angular velocity ω as shown in the picture below. Answer: (c) 2. Circle correct answer. In an elastic collision ω (a) momentum is not conserved but kinetic en- ergy is conserved; (b) momentum is conserved but kinetic energy R is not conserved; (c) both momentum and kinetic energy are conserved; (d) neither kinetic energy nor momentum is Which direction does the vector ~ω point ? conserved; (e) none of the above.
    [Show full text]
  • The Geometry of the Disk Complex
    JOURNAL OF THE AMERICAN MATHEMATICAL SOCIETY Volume 26, Number 1, January 2013, Pages 1–62 S 0894-0347(2012)00742-5 Article electronically published on August 22, 2012 THE GEOMETRY OF THE DISK COMPLEX HOWARD MASUR AND SAUL SCHLEIMER Contents 1. Introduction 1 2. Background on complexes 3 3. Background on coarse geometry 6 4. Natural maps 8 5. Holes in general and the lower bound on distance 11 6. Holes for the non-orientable surface 13 7. Holes for the arc complex 14 8. Background on three-manifolds 14 9. Holes for the disk complex 18 10. Holes for the disk complex – annuli 19 11. Holes for the disk complex – compressible 22 12. Holes for the disk complex – incompressible 25 13. Axioms for combinatorial complexes 30 14. Partition and the upper bound on distance 32 15. Background on Teichm¨uller space 41 16. Paths for the non-orientable surface 44 17. Paths for the arc complex 47 18. Background on train tracks 48 19. Paths for the disk complex 50 20. Hyperbolicity 53 21. Coarsely computing Hempel distance 57 Acknowledgments 60 References 60 1. Introduction Suppose M is a compact, connected, orientable, irreducible three-manifold. Sup- pose S is a compact, connected subsurface of ∂M so that no component of ∂S bounds a disk in M. In this paper we study the intrinsic geometry of the disk complex D(M,S). The disk complex has a natural simplicial inclusion into the curve complex C(S). Surprisingly, this inclusion is generally not a quasi-isometric embedding; there are disks which are close in the curve complex yet far apart in the disk complex.
    [Show full text]
  • 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]
  • The Real Projective Spaces in Homotopy Type Theory
    The real projective spaces in homotopy type theory Ulrik Buchholtz Egbert Rijke Technische Universität Darmstadt Carnegie Mellon University Email: [email protected] Email: [email protected] Abstract—Homotopy type theory is a version of Martin- topology and homotopy theory developed in homotopy Löf type theory taking advantage of its homotopical models. type theory (homotopy groups, including the fundamen- In particular, we can use and construct objects of homotopy tal group of the circle, the Hopf fibration, the Freuden- theory and reason about them using higher inductive types. In this article, we construct the real projective spaces, key thal suspension theorem and the van Kampen theorem, players in homotopy theory, as certain higher inductive types for example). Here we give an elementary construction in homotopy type theory. The classical definition of RPn, in homotopy type theory of the real projective spaces as the quotient space identifying antipodal points of the RPn and we develop some of their basic properties. n-sphere, does not translate directly to homotopy type theory. R n In classical homotopy theory the real projective space Instead, we define P by induction on n simultaneously n with its tautological bundle of 2-element sets. As the base RP is either defined as the space of lines through the + case, we take RP−1 to be the empty type. In the inductive origin in Rn 1 or as the quotient by the antipodal action step, we take RPn+1 to be the mapping cone of the projection of the 2-element group on the sphere Sn [4].
    [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]
  • 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.
    [Show full text]