Lectures on Lie Groups

Total Page:16

File Type:pdf, Size:1020Kb

Lectures on Lie Groups Lectures on Lie Groups Dragan Miliˇci´c Contents Chapter1. Basicdifferentialgeometry 1 1. Differentiable manifolds 1 2. Quotients 4 3. Foliations 11 4. Integration on manifolds 19 Chapter2.Liegroups 23 1.Liegroups 23 2. Lie algebra of a Lie group 49 3.HaarmeasuresonLiegroups 79 Chapter3.CompactLiegroups 83 1.CompactLiegroups 83 Chapter4. BasicLiealgebratheory 103 1. Solvable, nilpotent and semisimple Lie algebras 103 2. Lie algebras and field extensions 110 3. Cartan’s criterion 115 4. Semisimple Lie algebras 120 5.Cartansubalgebras 131 Chapter 5. Structure of semisimple Lie algebras 143 1.Rootsystems 143 2. Root system of a semisimple Lie algebra 152 3. BasesofrootsystemsandWeylgroups 158 4. Relations defining semisimple Lie algebra 165 5. Semisimple Lie algebras and their root systems 177 Chapter6. ClassificationofcompactLiegroups 181 1. Compact semisimple Lie groups 181 iii CHAPTER 1 Basic differential geometry 1. Differentiable manifolds 1.1. Differentiable manifolds and differentiable maps. Let M be a topo- logical space. A chart on M is a triple c = (U,ϕ,p) consisting of an open subset U M, an integer p Z+ and a homeomorphism ϕ of U onto an open set in Rp.⊂ The open set U is∈ called the domain of the chart c, and the integer p is the dimension of the chart c. The charts c = (U,ϕ,p) and c′ = (U ′, ϕ′,p′) on M are compatible if either U U ′ = or U U ′ = and ϕ′ ϕ−1 : ϕ(U U ′) ϕ′(U U ′) is a C∞- diffeomorphism.∩ ∅ ∩ 6 ∅ ◦ ∩ −→ ∩ A family of charts on M is an atlas of M if the domains of charts form a covering of MAand any two charts in are compatible. Atlases and of M are compatibleA if their union is an atlas on M. This is obviously anA equivalenceB relation on the set of all atlases on M. Each equivalence class of atlases contains the largest element which is equal to the union of all atlases in this class. Such atlas is called saturated. A differentiable manifold M is a hausdorff topological space with a saturated atlas. Clearly, a differentiable manifold is a locally compact space. It is also locally connected. Therefore, its connected components are open and closed subsets. Let M be a differentiable manifold. A chart c = (U,ϕ,p) is a chart around m M if m U. We say that it is centered at m if ϕ(m) = 0. ∈If c = (U,ϕ,p∈ ) and c′ = (U ′, ϕ′,p′) are two charts around m, then p = p′. There- fore, all charts around m have the same dimension. Therefore, we call p the dimen- sion of M at the point m and denote it by dimm M. The function m dimm M is locally constant on M. Therefore, it is constant on connected components7−→ of M. If dimm M = p for all m M, we say that M is an p-dimensional manifold. Let M and N be two differentiable∈ manifolds. A continuous map F : M N is a differentiable map if for any two pairs of charts c = (U,ϕ,p) on M and−→d = (V,ψ,q) on N such that F (U) V , the mapping ⊂ ψ F ϕ−1 : ϕ(U) ϕ(V ) ◦ ◦ −→ is a C∞-differentiable map. We denote by Mor(M,N) the set of all differentiable maps from M into N. If N is the real line R with obvious manifold structure, we call a differentiable map f : M R a differentiable function on M. The set of all differentiable func- tions on M −→forms an algebra C∞(M) over R with respect to pointwise operations. Clearly, differentiable manifolds as objects and differentiable maps as mor- phisms form a category. Isomorphisms in this category are called diffeomorphisms. 1 2 1.DIFFERENTIALGEOMETRY 1.2. Tangent spaces. Let M be a differentiable manifold and m a point in M. A linear form ξ on C∞(M) is called a tangent vector at m if it satisfies ξ(fg)= ξ(f)g(m)+ f(m)ξ(g) for any f,g C∞(M). Clearly, all tangent vectors at m form a linear space which ∈ we denote by Tm(M) and call the tangent space to M at m. Let m M and c = (U,ϕ,p) a chart centered at m. Then, for any 1 i p, we can define∈ the linear form ≤ ≤ ∂(f ϕ−1) ∂i(f)= ◦ (0). ∂xi Clearly, ∂i are tangent vectors in Tm(M). 1.2.1. Lemma. The vectors ∂1, ∂2, . , ∂p for a basis of the linear space Tm(M). In particular, dim Tm(M) = dimm M. Let F : M N be a morphism of differentiable manifolds. Let m M. Then, for any ξ −→T (M), the linear form T (F )ξ : g ξ(g F ) for g C∞∈(N), ∈ m m 7−→ ◦ ∈ is a tangent vector in TF (m)(N). Clearly, Tm(F ): Tm(M) TF (m)(N) is a linear map. It is called the differential of F at m. −→ The rank rank F of a morphism F : M N at m is the rank of the linear m −→ map Tm(F ). 1.2.2. Lemma. The function m rank F is lower semicontinuous on M. 7−→ m 1.3. Local diffeomorphisms, immersions, submersions and subimmer- sions. Let F : M N be a morphism of differentiable manifolds. The map F is a local diffeomorphism−→ at m if there is an open neighborhood U of m such that F (U) is an open set in N and F : U F (U) is a diffeomorphism. −→ 1.3.1. Theorem. Let F : M N be a morphism of differentiable manifolds. Let m M. Then the following conditions−→ are equivalent: ∈ (i) F is a local diffeomorphism at m; (ii) T (F ): T (M) T (N) is an isomorphism. m m −→ F (m) A morphism F : M N is an immersion at m if T (F ): T (M) −→ m m −→ TF (m)(N) is injective. A morphism F : M N is an submersion at m if Tm(F ): T (M) T (N) is surjective. −→ m −→ F (m) If F is an immersion at m, rankm F = dimm M, and by 1.2.2, this condition holds in an open neighborhood of m. Therefore, F is an immersion in a neighbor- hood of m. Analogously, if F is an submersion at m, rankm F = dimF (m) N, and by 1.2.2, this condition holds in an open neighborhood of m. Therefore, F is an submersion in a neighborhood of m. A morphism F : M N is an subimmerson at m if there exists a neighbor- hood U of m such that the−→ rank of F is constant on U. By the above discussion, immersions and submersions at m are subimmersions at p. A differentiable map F : M N is an local diffeomorphism if it is a local diffeomorphism at each point of −→M. A differentiable map F : M N is an immersion if it is an immersion at each point of M. A differentiable map−→F : M N is an submersion if it is an submersion ant each point of M. A differentiable−→ map F : M N is an subimmersion if it is an subimmersion at each point of M. The rank of−→ a subimmersion is constant on connected components of M. 1.DIFFERENTIABLEMANIFOLDS 3 1.3.2. Theorem. Let F : M N be a subimmersion at p M. Assume that −→ ∈ rankm F = r. Then there exists charts c = (U,ϕ,m) and d = (V,ψ,n) centered at p and F (p) respectively, such that F (U) V and ⊂ (ψ F ϕ−1)(x ,...,x ) = (x ,...,x , 0,..., 0) ◦ ◦ 1 n 1 r for any (x ,...,x ) ϕ(U). 1 n ∈ 1.3.3. Corollary. Let i : M N be an immersion. Let F : P M be a continuous map. Then the following−→ conditions are equivalent: −→ (i) F is differentiable; (ii) i F is differentiable. ◦ 1.3.4. Corollary. Let p : M N be a surjective submersion. Let F : N P be a map. Then the following conditions−→ are equivalent: −→ (i) F is differentiable; (ii) F p is differentiable. ◦ 1.3.5. Corollary. A submersion F : M N is an open map. −→ 1.4. Submanifolds. Let N be a subset of a differentiable manifold M. As- sume that any point n N has an open neighborhood U in M and a chart (U,ϕ,p) centered at n such that∈ ϕ(N U) = ϕ(U) Rq 0 . If we equip N with the induced topology and define its∩ atlas consisting∩ of×{ charts} on open sets N U given by the maps ϕ : N U Rq, N becomes a differentiable manifold. With∩ this differentiable structure,∩ the−→ natural inclusion i : N M is an immersion. The manifold N is called a submanifold of M. −→ 1.4.1. Lemma. A submanifold N of a manifold M is locally closed. 1.4.2. Lemma. Let f : M N be an injective immersion. If f is a homeo- morphism of M onto f(M) N−→, f(M) is a submanifold in N and f : M f(M) is a diffeomorphism. ⊂ −→ Let f : M N is a differentiable map. Denote by Γf the graph of f, i.e., the subset (m,f(m−→)) M N m M . Then, α : m (m,f(m)) is a continuous { ∈ × | ∈ } 7−→ bijection of M onto Γf . The inverse of α is the restriction of the canonical projection p : M N M to the graph Γ . Therefore, α : M Γ is a homeomorphism. × −→ f −→ f On the other hand, the differential of α is given by Tm(α)(ξ) = (ξ,Tm(f)(ξ)) for any ξ T (M), hence α is an immersion.
Recommended publications
  • A Note on Presentation of General Linear Groups Over a Finite Field
    Southeast Asian Bulletin of Mathematics (2019) 43: 217–224 Southeast Asian Bulletin of Mathematics c SEAMS. 2019 A Note on Presentation of General Linear Groups over a Finite Field Swati Maheshwari and R. K. Sharma Department of Mathematics, Indian Institute of Technology Delhi, New Delhi, India Email: [email protected]; [email protected] Received 22 September 2016 Accepted 20 June 2018 Communicated by J.M.P. Balmaceda AMS Mathematics Subject Classification(2000): 20F05, 16U60, 20H25 Abstract. In this article we have given Lie regular generators of linear group GL(2, Fq), n where Fq is a finite field with q = p elements. Using these generators we have obtained presentations of the linear groups GL(2, F2n ) and GL(2, Fpn ) for each positive integer n. Keywords: Lie regular units; General linear group; Presentation of a group; Finite field. 1. Introduction Suppose F is a finite field and GL(n, F) is the general linear the group of n × n invertible matrices and SL(n, F) is special linear group of n × n matrices with determinant 1. We know that GL(n, F) can be written as a semidirect product, GL(n, F)= SL(n, F) oF∗, where F∗ denotes the multiplicative group of F. Let H and K be two groups having presentations H = hX | Ri and K = hY | Si, then a presentation of semidirect product of H and K is given by, −1 H oη K = hX, Y | R,S,xyx = η(y)(x) ∀x ∈ X,y ∈ Y i, where η : K → Aut(H) is a group homomorphism. Now we summarize some literature survey related to the presentation of groups.
    [Show full text]
  • Abelian Varieties with Complex Multiplication and Modular Functions, by Goro Shimura, Princeton Univ
    BULLETIN (New Series) OF THE AMERICAN MATHEMATICAL SOCIETY Volume 36, Number 3, Pages 405{408 S 0273-0979(99)00784-3 Article electronically published on April 27, 1999 Abelian varieties with complex multiplication and modular functions, by Goro Shimura, Princeton Univ. Press, Princeton, NJ, 1998, xiv + 217 pp., $55.00, ISBN 0-691-01656-9 The subject that might be called “explicit class field theory” begins with Kro- necker’s Theorem: every abelian extension of the field of rational numbers Q is a subfield of a cyclotomic field Q(ζn), where ζn is a primitive nth root of 1. In other words, we get all abelian extensions of Q by adjoining all “special values” of e(x)=exp(2πix), i.e., with x Q. Hilbert’s twelfth problem, also called Kronecker’s Jugendtraum, is to do something2 similar for any number field K, i.e., to generate all abelian extensions of K by adjoining special values of suitable special functions. Nowadays we would add that the reciprocity law describing the Galois group of an abelian extension L/K in terms of ideals of K should also be given explicitly. After K = Q, the next case is that of an imaginary quadratic number field K, with the real torus R/Z replaced by an elliptic curve E with complex multiplication. (Kronecker knew what the result should be, although complete proofs were given only later, by Weber and Takagi.) For simplicity, let be the ring of integers in O K, and let A be an -ideal. Regarding A as a lattice in C, we get an elliptic curve O E = C/A with End(E)= ;Ehas complex multiplication, or CM,by .If j=j(A)isthej-invariant ofOE,thenK(j) is the Hilbert class field of K, i.e.,O the maximal abelian unramified extension of K.
    [Show full text]
  • Material on Algebraic and Lie Groups
    2 Lie groups and algebraic groups. 2.1 Basic Definitions. In this subsection we will introduce the class of groups to be studied. We first recall that a Lie group is a group that is also a differentiable manifold 1 and multiplication (x, y xy) and inverse (x x ) are C1 maps. An algebraic group is a group7! that is also an algebraic7! variety such that multi- plication and inverse are morphisms. Before we can introduce our main characters we first consider GL(n, C) as an affi ne algebraic group. Here Mn(C) denotes the space of n n matrices and GL(n, C) = g Mn(C) det(g) =) . Now Mn(C) is given the structure nf2 2 j 6 g of affi ne space C with the coordinates xij for X = [xij] . This implies that GL(n, C) is Z-open and as a variety is isomorphic with the affi ne variety 1 Mn(C) det . This implies that (GL(n, C)) = C[xij, det ]. f g O Lemma 1 If G is an algebraic group over an algebraically closed field, F , then every point in G is smooth. Proof. Let Lg : G G be given by Lgx = gx. Then Lg is an isomorphism ! 1 1 of G as an algebraic variety (Lg = Lg ). Since isomorphisms preserve the set of smooth points we see that if x G is smooth so is every element of Gx = G. 2 Proposition 2 If G is an algebraic group over an algebraically closed field F then the Z-connected components Proof.
    [Show full text]
  • Matroid Automorphisms of the F4 Root System
    ∗ Matroid automorphisms of the F4 root system Stephanie Fried Aydin Gerek Gary Gordon Dept. of Mathematics Dept. of Mathematics Dept. of Mathematics Grinnell College Lehigh University Lafayette College Grinnell, IA 50112 Bethlehem, PA 18015 Easton, PA 18042 [email protected] [email protected] Andrija Peruniˇci´c Dept. of Mathematics Bard College Annandale-on-Hudson, NY 12504 [email protected] Submitted: Feb 11, 2007; Accepted: Oct 20, 2007; Published: Nov 12, 2007 Mathematics Subject Classification: 05B35 Dedicated to Thomas Brylawski. Abstract Let F4 be the root system associated with the 24-cell, and let M(F4) be the simple linear dependence matroid corresponding to this root system. We determine the automorphism group of this matroid and compare it to the Coxeter group W for the root system. We find non-geometric automorphisms that preserve the matroid but not the root system. 1 Introduction Given a finite set of vectors in Euclidean space, we consider the linear dependence matroid of the set, where dependences are taken over the reals. When the original set of vectors has `nice' symmetry, it makes sense to compare the geometric symmetry (the Coxeter or Weyl group) with the group that preserves sets of linearly independent vectors. This latter group is precisely the automorphism group of the linear matroid determined by the given vectors. ∗Research supported by NSF grant DMS-055282. the electronic journal of combinatorics 14 (2007), #R78 1 For the root systems An and Bn, matroid automorphisms do not give any additional symmetry [4]. One can interpret these results in the following way: combinatorial sym- metry (preserving dependence) and geometric symmetry (via isometries of the ambient Euclidean space) are essentially the same.
    [Show full text]
  • Modules and Lie Semialgebras Over Semirings with a Negation Map 3
    MODULES AND LIE SEMIALGEBRAS OVER SEMIRINGS WITH A NEGATION MAP GUY BLACHAR Abstract. In this article, we present the basic definitions of modules and Lie semialgebras over semirings with a negation map. Our main example of a semiring with a negation map is ELT algebras, and some of the results in this article are formulated and proved only in the ELT theory. When dealing with modules, we focus on linearly independent sets and spanning sets. We define a notion of lifting a module with a negation map, similarly to the tropicalization process, and use it to prove several theorems about semirings with a negation map which possess a lift. In the context of Lie semialgebras over semirings with a negation map, we first give basic definitions, and provide parallel constructions to the classical Lie algebras. We prove an ELT version of Cartan’s criterion for semisimplicity, and provide a counterexample for the naive version of the PBW Theorem. Contents Page 0. Introduction 2 0.1. Semirings with a Negation Map 2 0.2. Modules Over Semirings with a Negation Map 3 0.3. Supertropical Algebras 4 0.4. Exploded Layered Tropical Algebras 4 1. Modules over Semirings with a Negation Map 5 1.1. The Surpassing Relation for Modules 6 1.2. Basic Definitions for Modules 7 1.3. -morphisms 9 1.4. Lifting a Module Over a Semiring with a Negation Map 10 1.5. Linearly Independent Sets 13 1.6. d-bases and s-bases 14 1.7. Free Modules over Semirings with a Negation Map 18 2.
    [Show full text]
  • Platonic Solids Generate Their Four-Dimensional Analogues
    1 Platonic solids generate their four-dimensional analogues PIERRE-PHILIPPE DECHANT a;b;c* aInstitute for Particle Physics Phenomenology, Ogden Centre for Fundamental Physics, Department of Physics, University of Durham, South Road, Durham, DH1 3LE, United Kingdom, bPhysics Department, Arizona State University, Tempe, AZ 85287-1604, United States, and cMathematics Department, University of York, Heslington, York, YO10 5GG, United Kingdom. E-mail: [email protected] Polytopes; Platonic Solids; 4-dimensional geometry; Clifford algebras; Spinors; Coxeter groups; Root systems; Quaternions; Representations; Symmetries; Trinities; McKay correspondence Abstract In this paper, we show how regular convex 4-polytopes – the analogues of the Platonic solids in four dimensions – can be constructed from three-dimensional considerations concerning the Platonic solids alone. Via the Cartan-Dieudonne´ theorem, the reflective symmetries of the arXiv:1307.6768v1 [math-ph] 25 Jul 2013 Platonic solids generate rotations. In a Clifford algebra framework, the space of spinors gen- erating such three-dimensional rotations has a natural four-dimensional Euclidean structure. The spinors arising from the Platonic Solids can thus in turn be interpreted as vertices in four- dimensional space, giving a simple construction of the 4D polytopes 16-cell, 24-cell, the F4 root system and the 600-cell. In particular, these polytopes have ‘mysterious’ symmetries, that are almost trivial when seen from the three-dimensional spinorial point of view. In fact, all these induced polytopes are also known to be root systems and thus generate rank-4 Coxeter PREPRINT: Acta Crystallographica Section A A Journal of the International Union of Crystallography 2 groups, which can be shown to be a general property of the spinor construction.
    [Show full text]
  • Note on the Decomposition of Semisimple Lie Algebras
    Note on the Decomposition of Semisimple Lie Algebras Thomas B. Mieling (Dated: January 5, 2018) Presupposing two criteria by Cartan, it is shown that every semisimple Lie algebra of finite dimension over C is a direct sum of simple Lie algebras. Definition 1 (Simple Lie Algebra). A Lie algebra g is Proposition 2. Let g be a finite dimensional semisimple called simple if g is not abelian and if g itself and f0g are Lie algebra over C and a ⊆ g an ideal. Then g = a ⊕ a? the only ideals in g. where the orthogonal complement is defined with respect to the Killing form K. Furthermore, the restriction of Definition 2 (Semisimple Lie Algebra). A Lie algebra g the Killing form to either a or a? is non-degenerate. is called semisimple if the only abelian ideal in g is f0g. Proof. a? is an ideal since for x 2 a?; y 2 a and z 2 g Definition 3 (Derived Series). Let g be a Lie Algebra. it holds that K([x; z; ]; y) = K(x; [z; y]) = 0 since a is an The derived series is the sequence of subalgebras defined ideal. Thus [a; g] ⊆ a. recursively by D0g := g and Dn+1g := [Dng;Dng]. Since both a and a? are ideals, so is their intersection Definition 4 (Solvable Lie Algebra). A Lie algebra g i = a \ a?. We show that i = f0g. Let x; yi. Then is called solvable if its derived series terminates in the clearly K(x; y) = 0 for x 2 i and y = D1i, so i is solv- trivial subalgebra, i.e.
    [Show full text]
  • Matrix Lie Groups
    Maths Seminar 2007 MATRIX LIE GROUPS Claudiu C Remsing Dept of Mathematics (Pure and Applied) Rhodes University Grahamstown 6140 26 September 2007 RhodesUniv CCR 0 Maths Seminar 2007 TALK OUTLINE 1. What is a matrix Lie group ? 2. Matrices revisited. 3. Examples of matrix Lie groups. 4. Matrix Lie algebras. 5. A glimpse at elementary Lie theory. 6. Life beyond elementary Lie theory. RhodesUniv CCR 1 Maths Seminar 2007 1. What is a matrix Lie group ? Matrix Lie groups are groups of invertible • matrices that have desirable geometric features. So matrix Lie groups are simultaneously algebraic and geometric objects. Matrix Lie groups naturally arise in • – geometry (classical, algebraic, differential) – complex analyis – differential equations – Fourier analysis – algebra (group theory, ring theory) – number theory – combinatorics. RhodesUniv CCR 2 Maths Seminar 2007 Matrix Lie groups are encountered in many • applications in – physics (geometric mechanics, quantum con- trol) – engineering (motion control, robotics) – computational chemistry (molecular mo- tion) – computer science (computer animation, computer vision, quantum computation). “It turns out that matrix [Lie] groups • pop up in virtually any investigation of objects with symmetries, such as molecules in chemistry, particles in physics, and projective spaces in geometry”. (K. Tapp, 2005) RhodesUniv CCR 3 Maths Seminar 2007 EXAMPLE 1 : The Euclidean group E (2). • E (2) = F : R2 R2 F is an isometry . → | n o The vector space R2 is equipped with the standard Euclidean structure (the “dot product”) x y = x y + x y (x, y R2), • 1 1 2 2 ∈ hence with the Euclidean distance d (x, y) = (y x) (y x) (x, y R2).
    [Show full text]
  • Lecture 5: Semisimple Lie Algebras Over C
    LECTURE 5: SEMISIMPLE LIE ALGEBRAS OVER C IVAN LOSEV Introduction In this lecture I will explain the classification of finite dimensional semisimple Lie alge- bras over C. Semisimple Lie algebras are defined similarly to semisimple finite dimensional associative algebras but are far more interesting and rich. The classification reduces to that of simple Lie algebras (i.e., Lie algebras with non-zero bracket and no proper ideals). The classification (initially due to Cartan and Killing) is basically in three steps. 1) Using the structure theory of simple Lie algebras, produce a combinatorial datum, the root system. 2) Study root systems combinatorially arriving at equivalent data (Cartan matrix/ Dynkin diagram). 3) Given a Cartan matrix, produce a simple Lie algebra by generators and relations. In this lecture, we will cover the first two steps. The third step will be carried in Lecture 6. 1. Semisimple Lie algebras Our base field is C (we could use an arbitrary algebraically closed field of characteristic 0). 1.1. Criteria for semisimplicity. We are going to define the notion of a semisimple Lie algebra and give some criteria for semisimplicity. This turns out to be very similar to the case of semisimple associative algebras (although the proofs are much harder). Let g be a finite dimensional Lie algebra. Definition 1.1. We say that g is simple, if g has no proper ideals and dim g > 1 (so we exclude the one-dimensional abelian Lie algebra). We say that g is semisimple if it is the direct sum of simple algebras. Any semisimple algebra g is the Lie algebra of an algebraic group, we can take the au- tomorphism group Aut(g).
    [Show full text]
  • LECTURE 12: LIE GROUPS and THEIR LIE ALGEBRAS 1. Lie
    LECTURE 12: LIE GROUPS AND THEIR LIE ALGEBRAS 1. Lie groups Definition 1.1. A Lie group G is a smooth manifold equipped with a group structure so that the group multiplication µ : G × G ! G; (g1; g2) 7! g1 · g2 is a smooth map. Example. Here are some basic examples: • Rn, considered as a group under addition. • R∗ = R − f0g, considered as a group under multiplication. • S1, Considered as a group under multiplication. • Linear Lie groups GL(n; R), SL(n; R), O(n) etc. • If M and N are Lie groups, so is their product M × N. Remarks. (1) (Hilbert's 5th problem, [Gleason and Montgomery-Zippin, 1950's]) Any topological group whose underlying space is a topological manifold is a Lie group. (2) Not every smooth manifold admits a Lie group structure. For example, the only spheres that admit a Lie group structure are S0, S1 and S3; among all the compact 2 dimensional surfaces the only one that admits a Lie group structure is T 2 = S1 × S1. (3) Here are two simple topological constraints for a manifold to be a Lie group: • If G is a Lie group, then TG is a trivial bundle. n { Proof: We identify TeG = R . The vector bundle isomorphism is given by φ : G × TeG ! T G; φ(x; ξ) = (x; dLx(ξ)) • If G is a Lie group, then π1(G) is an abelian group. { Proof: Suppose α1, α2 2 π1(G). Define α : [0; 1] × [0; 1] ! G by α(t1; t2) = α1(t1) · α2(t2). Then along the bottom edge followed by the right edge we have the composition α1 ◦ α2, where ◦ is the product of loops in the fundamental group, while along the left edge followed by the top edge we get α2 ◦ α1.
    [Show full text]
  • Tight Immersions of Simplicial Surfaces in Three Space
    View metadata, citation and similar papers at core.ac.uk brought to you by CORE provided by Elsevier - Publisher Connector Pergamon Topology Vol. 35, No. 4, pp. 863-873, 19% Copyright Q 1996 Elsevier Sctence Ltd Printed in Great Britain. All tights reserved CO4&9383/96 $15.00 + 0.00 0040-9383(95)00047-X TIGHT IMMERSIONS OF SIMPLICIAL SURFACES IN THREE SPACE DAVIDE P. CERVONE (Received for publication 13 July 1995) 1. INTRODUCTION IN 1960,Kuiper [12,13] initiated the study of tight immersions of surfaces in three space, and showed that every compact surface admits a tight immersion except for the Klein bottle, the real projective plane, and possibly the real projective plane with one handle. The fate of the latter surface went unresolved for 30 years until Haab recently proved [9] that there is no smooth tight immersion of the projective plane with one handle. In light of this, a recently described simplicial tight immersion of this surface [6] came as quite a surprise, and its discovery completes Kuiper’s initial survey. Pinkall [17] broadened the question in 1985 by looking for tight immersions in each equivalence class of immersed surfaces under image homotopy. He first produced a complete description of these classes, and proceeded to generate tight examples in all but finitely many of them. In this paper, we improve Pinkall’s result by giving tight simplicial examples in all but two of the immersion classes under image homotopy for which tight immersions are possible, and in particular, we point out and resolve an error in one of his examples that would have left an infinite number of immersion classes with no tight examples.
    [Show full text]
  • Automorphism Groups of Locally Compact Reductive Groups
    PROCEEDINGS OF THE AMERICAN MATHEMATICALSOCIETY Volume 106, Number 2, June 1989 AUTOMORPHISM GROUPS OF LOCALLY COMPACT REDUCTIVE GROUPS T. S. WU (Communicated by David G Ebin) Dedicated to Mr. Chu. Ming-Lun on his seventieth birthday Abstract. A topological group G is reductive if every continuous finite di- mensional G-module is semi-simple. We study the structure of those locally compact reductive groups which are the extension of their identity components by compact groups. We then study the automorphism groups of such groups in connection with the groups of inner automorphisms. Proposition. Let G be a locally compact reductive group such that G/Gq is compact. Then /(Go) is dense in Aq(G) . Let G be a locally compact topological group. Let A(G) be the group of all bi-continuous automorphisms of G. Then A(G) has the natural topology (the so-called Birkhoff topology or ^-topology [1, 3, 4]), so that it becomes a topo- logical group. We shall always adopt such topology in the following discussion. When G is compact, it is well known that the identity component A0(G) of A(G) is the group of all inner automorphisms induced by elements from the identity component G0 of G, i.e., A0(G) = I(G0). This fact is very useful in the study of the structure of locally compact groups. On the other hand, it is also well known that when G is a semi-simple Lie group with finitely many connected components A0(G) - I(G0). The latter fact had been generalized to more general groups ([3]).
    [Show full text]