Survey on Exterior Algebra and Differential Forms

Total Page:16

File Type:pdf, Size:1020Kb

Survey on Exterior Algebra and Differential Forms Survey on exterior algebra and differential forms Daniel Grieser 16. Mai 2013 Inhaltsverzeichnis 1 Exterior algebra for a vector space 1 1.1 Alternating forms, wedge and interior product . 1 1.2 A scalar product enters the stage . 3 1.3 Now add an orientation: Volume element, Hodge ∗ .................. 4 1.4 Formulas in an arbitrary basis . 5 1.5 Modifications for not positive definite inner product . 6 2 Differential forms 7 2.1 Pointwise ('tensorial') constructions . 7 2.2 Integration . 8 2.3 Derivative operations . 9 1 Exterior algebra for a vector space Let V be an n-dimensional real vector space. Whenever needed, we let e1; : : : ; en be a basis of V and e1; : : : ; en its dual basis. At first reading you may leave out the parts on Hodge ∗ and non-positive definite metrics. 1.1 Alternating forms, wedge and interior product k 1. Let k 2 N0.A k-multilinear form on V is a map ! : V ! R which is linear in each entry, i.e. 0 0 !(av1 + bv1; v2; : : : ; vk) = a !(v1; v2; : : : ; vk) + b !(v1; v2; : : : ; vk) 0 for all v1; v1; v2; : : : ; vk 2 V and a; b 2 R, and similarly for the other entries. The form is called alternating if it changes sign under interchange of any two vectors: !(v1; : : : ; vi; : : : ; vj; : : : ; vk) = −!(v1; : : : ; vj; : : : ; vi; : : : ; vk) Equivalent conditions (to alternating) are: !(v1; : : : ; vk) = 0 if any two of the vi are the same. Or: !(vσ(1); : : : ; vσ(k)) = sign(σ)!(v1; v2; : : : ; vk) for all permutations σ of f1; : : : ; kg. The space of alternating k-multilinear forms on V is denoted ΛkV ∗. This is a vector space with basis feI : jIj = kg, where I runs over subsets of f1; : : : ; ng with k elements and I i1 ik e := e ^ · · · ^ e for I = fi1 < i2 < ··· < ikg; with ^ defined below, or explicitly: I I e (ej1 ; : : : ; ejk ) = δJ for J = fj1 < ··· < jkg f1;2g 1 2 1 2 1 For example, e = e ^ e satisfies (e ^ e )(e1; e2) = 1, and this implies that (e ^ 2 1 2 P i P j e )(e2; e1) = −1 and all other (e ^ e )(ej1 ; ej2 ) are zero, hence for v = v ei; w = w ej we get (e1 ^ e2)(v; w) = v1w2 − v2w1. 1 k ∗ n n ∗ k ∗ It follows that Λ V has dimension k , in particular dim Λ V = 1, and Λ V = f0g if k > n. Also Λ1V ∗ = V ∗ and Λ0V ∗ = R.1 We also write k = deg ! if ! 2 ΛkV ∗ and call k the degree of the form !. 2. Wedge (or exterior) product: For ! 2 ΛkV ∗, ν 2 ΛlV ∗ define ! ^ ν 2 Λk+lV ∗ by X (! ^ ν)(v1; : : : ; vk+l) = sign(σ) !(vσ(1); : : : ; vσ(k))ν(vσ(k+1); : : : ; vσ(k+l)) σ where the sum runs over all permutations σ of f1; : : : ; k + lg preserving the order in the first and second 'block', i.e. satisfying σ(1) < ··· < σ(k), σ(k + 1) < ··· < σ(k + l). For example, for k = l = 1 (! ^ ν)(v; w) = !(v)ν(w) − !(w)ν(v) Rule: ! ^ ν = (−1)deg !·deg ν ν ^ ! For this property one says that ^ is 'graded commutative' (in the physics literature also 'super commutative'). Also, ^ is bilinear and associative. Remark (Relation to cross product in R3): The wedge product generalizes the cross product in the following sense. If V = R3 then dim Λ1R3 = dim Λ2R3 = 3. So we have identifications (isomorphisms) 1 3 3 1 2 3 Λ R ! R ;! = !1e + !2e + !3e 7! (!1;!2;!3) 2 3 3 2 3 3 1 1 2 Λ R ! R ; µ = µ1e ^ e + µ2e ^ e + µ3e ^ e 7! (µ1; µ2; µ3) 1 3 P i P j Now for !; ν 2 Λ R with ! = !ie , ν = νje we have 2 3 3 1 1 2 ! ^ ν = (!2ν3 − !3ν2)e ^ e + (!3ν1 − !1ν3)e ^ e + (!1ν2 − !2ν1)e ^ e so if !; ν are identified with vectors as in the first line, then ! ^ ν corresponds (as in the second line) to the cross product of these vectors. 3. Let v 2 V . The interior product with v is the linear operator k ∗ k−1 ∗ ιv :Λ V ! Λ V ;! 7! !(v; : : : ) that is, (ιv!)(v2; : : : ; vk) = !(v; v2; : : : ; vk) ('plug in v in the first slot'). Here k 2 N, but we 0 ∗ also define ιv = 0 on Λ V . Clearly, ιv depends linearly on v. With wedge products it behaves as follows (as a 'super derivation'): deg ! ιv(! ^ ν) = (ιv!) ^ ν + (−1) ! ^ (ιvν) (1) 3 P i For example, in R , if v = i v ei then 1 2 3 1 2 3 2 3 1 3 1 2 ιv(e ^ e ^ e ) = v e ^ e + v e ^ e + v e ^ e (2) (of course one could write −v2e1 ^ e3 for the middle term) 4. Behavior under maps: A linear map A : V ! W defines the pull-back map ∗ k ∗ k ∗ ∗ A :Λ W ! Λ V ; (A !)(v1; : : : ; vk) := !(Av1; : : : ; Avk) (3) 1 0 By definition, V = R, and a linear map R ! R is determined by its value at 1. 2 k ∗ where ! 2 Λ W , v1; : : : ; vk 2 V . For k = 1 this is also called the dual (or transpose) map A∗ : W ∗ ! V ∗. Pullback behaves naturally with wedge product A∗(! ^ ν) = A∗! ^ A∗ν ∗ ∗ and with interior product: ιv(A !) = A (ιA(v)!), as follows directly from the definitions. For V = W and k = n pullback relates to the determinant as follows: A∗ = (det A) Id on ΛnV ∗. Explicitly, this means !(Av1; : : : ; Avn) = (det A)!(v1; : : : ; vn) which follows directly from the facts that ! is multilinear and alternating, and the Leibniz formula for the determinant. 1.2 A scalar product enters the stage From now on assume that a scalar product is given on V , that is, a bilinear, symmetric, positive definite2 form g : V × V ! R. We also write hv; wi instead of g(v; w). This defines some more structures: 1. Basic geometry: The scalar product allows us to talk about lenghts of vectors and angles between non-zero vectors: g(v; w) jvj = pg(v; v); (v; w) = arccos \ jvj · jwj 2. Using the scalar product on V we get a map g# : V ! V ∗; v 7! g(v; ·) Since g is non-degenerate, this map is injective, hence bijective (since dim V = dim V ∗ < 1). The inverse of g# is called g[ : V ∗ ! V Therefore, we may identify vectors and linear forms (but we do this only when necessary).3 3. Using this identification, we get a scalar product on V ∗, which we also denote by h ; i: hα; βi := hg[(α); g[(β)i for α; β 2 V ∗. 4. More generally, we get a scalar product on ΛkV ∗ for each k. It is easiest to define it by the property: I k ∗ If e1; : : : ; en are orthonormal then the basis fe : jIj = kg of Λ V is orthonormal: P I P J P 1 k 1 k In other words, h I aI e ; J bJ e i := I aI bI . Then for arbitrary v ; : : : v ; w ; : : : ; w 2 V ∗ one has4 hv1 ^ · · · ^ vk; w1 ^ · · · ^ wki = det(hvi; wji) (4) This formula also shows that one obtains the same scalar product if one uses a different orthonormal basis in the definition. 2Everything can be done in the more general case that g is only non-degenerate, but one needs to be careful with the signs, see Section 1.5. 3The fact that g# is surjective, i.e. that every linear form on V can be represented by a vector using the scalar product, is sometimes called the Riesz lemma. It holds more generally when (V; g) is a Hilbert space, that is, if V is allowed to be infinite-dimensional but required to be complete with the norm defined by g. 4Proof: By definition this holds if all vi, wj are taken from the basis vectors e1; : : : ; en. Then it holds in general since both sides are multilinear in the 2k entries v1; : : : vk; w1; : : : ; wk. 3 1.3 Now add an orientation: Volume element, Hodge ∗ Now assume that on V a scalar product and an orientation is given. 1. The Hodge ∗ operator is the unique linear map (for each k) ∗ :ΛkV ∗ ! Λn−kV ∗ with the property that5 ∗ (e1 ^ · · · ^ ek) = ek+1 ^ · · · ^ en for any oONB, (5) 1 n that is, for any oriented orthonormal basis (oONB) e1; : : : ; en with dual basis e ; : : : ; e . Intuition: k-forms e1 ^ · · · ^ ek correspond to k-dimensional subspaces W = spanfe1; : : : ; ekg of V ∗. Then ∗(e1 ^ · · · ^ ek) corresponds to the orthogonal complement of W . Of course not every form can be written in this way, but using linearity ∗ is defined when it is defined on forms of this type. So one can say:6 • Alternating multilinear forms are a 'linear extension' of the notion of vector subspace. • Then ∗ corresponds to orthogonal complement. 2. Define the volume element (or volume form) of V as dvol = ∗1; dvol 2 ΛnV ∗: Why is this reasonable? Because for any oONB e1; : : : ; en we have, by definition of ∗, dvol = e1 ^ · · · ^ en and therefore dvol(e1; : : : ; en) = 1 for any oONB. (6) So the volume of a 'unit cube' is one, as it should be. 3. Properties of ∗: ! ^ ∗ν = h!; νi dvol for !; ν 2 ΛkV ∗ (7) Also, if ν is fixed then the validity of (7) for all ! defines ∗ν. (7) can easily be checked on basis elements, and then extends by linearity.7 ∗ (ei1 ^ · · · ^ eik ) = sign(σ)ej1 ^ · · · ^ ejn−k for oONB (8) where fj1; : : : ; jn−kg = f1; : : : ; ngnfi1; : : : ; ikg and σ is the permutation sending (1; : : : ; n) 7! 8 (i1; : : : ; ik; j1; : : : ; jn−k).
Recommended publications
  • Introduction to Linear Bialgebra
    View metadata, citation and similar papers at core.ac.uk brought to you by CORE provided by University of New Mexico University of New Mexico UNM Digital Repository Mathematics and Statistics Faculty and Staff Publications Academic Department Resources 2005 INTRODUCTION TO LINEAR BIALGEBRA Florentin Smarandache University of New Mexico, [email protected] W.B. Vasantha Kandasamy K. Ilanthenral Follow this and additional works at: https://digitalrepository.unm.edu/math_fsp Part of the Algebra Commons, Analysis Commons, Discrete Mathematics and Combinatorics Commons, and the Other Mathematics Commons Recommended Citation Smarandache, Florentin; W.B. Vasantha Kandasamy; and K. Ilanthenral. "INTRODUCTION TO LINEAR BIALGEBRA." (2005). https://digitalrepository.unm.edu/math_fsp/232 This Book is brought to you for free and open access by the Academic Department Resources at UNM Digital Repository. It has been accepted for inclusion in Mathematics and Statistics Faculty and Staff Publications by an authorized administrator of UNM Digital Repository. For more information, please contact [email protected], [email protected], [email protected]. INTRODUCTION TO LINEAR BIALGEBRA W. B. Vasantha Kandasamy Department of Mathematics Indian Institute of Technology, Madras Chennai – 600036, India e-mail: [email protected] web: http://mat.iitm.ac.in/~wbv Florentin Smarandache Department of Mathematics University of New Mexico Gallup, NM 87301, USA e-mail: [email protected] K. Ilanthenral Editor, Maths Tiger, Quarterly Journal Flat No.11, Mayura Park, 16, Kazhikundram Main Road, Tharamani, Chennai – 600 113, India e-mail: [email protected] HEXIS Phoenix, Arizona 2005 1 This book can be ordered in a paper bound reprint from: Books on Demand ProQuest Information & Learning (University of Microfilm International) 300 N.
    [Show full text]
  • Topology and Physics 2019 - Lecture 2
    Topology and Physics 2019 - lecture 2 Marcel Vonk February 12, 2019 2.1 Maxwell theory in differential form notation Maxwell's theory of electrodynamics is a great example of the usefulness of differential forms. A nice reference on this topic, though somewhat outdated when it comes to notation, is [1]. For notational simplicity, we will work in units where the speed of light, the vacuum permittivity and the vacuum permeability are all equal to 1: c = 0 = µ0 = 1. 2.1.1 The dual field strength In three dimensional space, Maxwell's electrodynamics describes the physics of the electric and magnetic fields E~ and B~ . These are three-dimensional vector fields, but the beauty of the theory becomes much more obvious if we (a) use a four-dimensional relativistic formulation, and (b) write it in terms of differential forms. For example, let us look at Maxwells two source-free, homogeneous equations: r · B = 0;@tB + r × E = 0: (2.1) That these equations have a relativistic flavor becomes clear if we write them out in com- ponents and organize the terms somewhat suggestively: x y z 0 + @xB + @yB + @zB = 0 x z y −@tB + 0 − @yE + @zE = 0 (2.2) y z x −@tB + @xE + 0 − @zE = 0 z y x −@tB − @xE + @yE + 0 = 0 Note that we also multiplied the last three equations by −1 to clarify the structure. All in all, we see that we have four equations (one for each space-time coordinate) which each contain terms in which the four coordinate derivatives act. Therefore, we may be tempted to write our set of equations in more \relativistic" notation as ^µν @µF = 0 (2.3) 1 with F^µν the coordinates of an antisymmetric two-tensor (i.
    [Show full text]
  • Introduction to Gauge Theory Arxiv:1910.10436V1 [Math.DG] 23
    Introduction to Gauge Theory Andriy Haydys 23rd October 2019 This is lecture notes for a course given at the PCMI Summer School “Quantum Field The- ory and Manifold Invariants” (July 1 – July 5, 2019). I describe basics of gauge-theoretic approach to construction of invariants of manifolds. The main example considered here is the Seiberg–Witten gauge theory. However, I tried to present the material in a form, which is suitable for other gauge-theoretic invariants too. Contents 1 Introduction2 2 Bundles and connections4 2.1 Vector bundles . .4 2.1.1 Basic notions . .4 2.1.2 Operations on vector bundles . .5 2.1.3 Sections . .6 2.1.4 Covariant derivatives . .6 2.1.5 The curvature . .8 2.1.6 The gauge group . 10 2.2 Principal bundles . 11 2.2.1 The frame bundle and the structure group . 11 2.2.2 The associated vector bundle . 14 2.2.3 Connections on principal bundles . 16 2.2.4 The curvature of a connection on a principal bundle . 19 arXiv:1910.10436v1 [math.DG] 23 Oct 2019 2.2.5 The gauge group . 21 2.3 The Levi–Civita connection . 22 2.4 Classification of U(1) and SU(2) bundles . 23 2.4.1 Complex line bundles . 24 2.4.2 Quaternionic line bundles . 25 3 The Chern–Weil theory 26 3.1 The Chern–Weil theory . 26 3.1.1 The Chern classes . 28 3.2 The Chern–Simons functional . 30 3.3 The modui space of flat connections . 32 3.3.1 Parallel transport and holonomy .
    [Show full text]
  • LP THEORY of DIFFERENTIAL FORMS on MANIFOLDS This
    TRANSACTIONSOF THE AMERICAN MATHEMATICALSOCIETY Volume 347, Number 6, June 1995 LP THEORY OF DIFFERENTIAL FORMS ON MANIFOLDS CHAD SCOTT Abstract. In this paper, we establish a Hodge-type decomposition for the LP space of differential forms on closed (i.e., compact, oriented, smooth) Rieman- nian manifolds. Critical to the proof of this result is establishing an LP es- timate which contains, as a special case, the L2 result referred to by Morrey as Gaffney's inequality. This inequality helps us show the equivalence of the usual definition of Sobolev space with a more geometric formulation which we provide in the case of differential forms on manifolds. We also prove the LP boundedness of Green's operator which we use in developing the LP theory of the Hodge decomposition. For the calculus of variations, we rigorously verify that the spaces of exact and coexact forms are closed in the LP norm. For nonlinear analysis, we demonstrate the existence and uniqueness of a solution to the /1-harmonic equation. 1. Introduction This paper contributes primarily to the development of the LP theory of dif- ferential forms on manifolds. The reader should be aware that for the duration of this paper, manifold will refer only to those which are Riemannian, compact, oriented, C°° smooth and without boundary. For p = 2, the LP theory is well understood and the L2-Hodge decomposition can be found in [M]. However, in the case p ^ 2, the LP theory has yet to be fully developed. Recent appli- cations of the LP theory of differential forms on W to both quasiconformal mappings and nonlinear elasticity continue to motivate interest in this subject.
    [Show full text]
  • 21. Orthonormal Bases
    21. Orthonormal Bases The canonical/standard basis 011 001 001 B C B C B C B0C B1C B0C e1 = B.C ; e2 = B.C ; : : : ; en = B.C B.C B.C B.C @.A @.A @.A 0 0 1 has many useful properties. • Each of the standard basis vectors has unit length: q p T jjeijj = ei ei = ei ei = 1: • The standard basis vectors are orthogonal (in other words, at right angles or perpendicular). T ei ej = ei ej = 0 when i 6= j This is summarized by ( 1 i = j eT e = δ = ; i j ij 0 i 6= j where δij is the Kronecker delta. Notice that the Kronecker delta gives the entries of the identity matrix. Given column vectors v and w, we have seen that the dot product v w is the same as the matrix multiplication vT w. This is the inner product on n T R . We can also form the outer product vw , which gives a square matrix. 1 The outer product on the standard basis vectors is interesting. Set T Π1 = e1e1 011 B C B0C = B.C 1 0 ::: 0 B.C @.A 0 01 0 ::: 01 B C B0 0 ::: 0C = B. .C B. .C @. .A 0 0 ::: 0 . T Πn = enen 001 B C B0C = B.C 0 0 ::: 1 B.C @.A 1 00 0 ::: 01 B C B0 0 ::: 0C = B. .C B. .C @. .A 0 0 ::: 1 In short, Πi is the diagonal square matrix with a 1 in the ith diagonal position and zeros everywhere else.
    [Show full text]
  • NOTES on DIFFERENTIAL FORMS. PART 3: TENSORS 1. What Is A
    NOTES ON DIFFERENTIAL FORMS. PART 3: TENSORS 1. What is a tensor? 1 n Let V be a finite-dimensional vector space. It could be R , it could be the tangent space to a manifold at a point, or it could just be an abstract vector space. A k-tensor is a map T : V × · · · × V ! R 2 (where there are k factors of V ) that is linear in each factor. That is, for fixed ~v2; : : : ;~vk, T (~v1;~v2; : : : ;~vk−1;~vk) is a linear function of ~v1, and for fixed ~v1;~v3; : : : ;~vk, T (~v1; : : : ;~vk) is a k ∗ linear function of ~v2, and so on. The space of k-tensors on V is denoted T (V ). Examples: n • If V = R , then the inner product P (~v; ~w) = ~v · ~w is a 2-tensor. For fixed ~v it's linear in ~w, and for fixed ~w it's linear in ~v. n • If V = R , D(~v1; : : : ;~vn) = det ~v1 ··· ~vn is an n-tensor. n • If V = R , T hree(~v) = \the 3rd entry of ~v" is a 1-tensor. • A 0-tensor is just a number. It requires no inputs at all to generate an output. Note that the definition of tensor says nothing about how things behave when you rotate vectors or permute their order. The inner product P stays the same when you swap the two vectors, but the determinant D changes sign when you swap two vectors. Both are tensors. For a 1-tensor like T hree, permuting the order of entries doesn't even make sense! ~ ~ Let fb1;:::; bng be a basis for V .
    [Show full text]
  • Multivector Differentiation and Linear Algebra 0.5Cm 17Th Santaló
    Multivector differentiation and Linear Algebra 17th Santalo´ Summer School 2016, Santander Joan Lasenby Signal Processing Group, Engineering Department, Cambridge, UK and Trinity College Cambridge [email protected], www-sigproc.eng.cam.ac.uk/ s jl 23 August 2016 1 / 78 Examples of differentiation wrt multivectors. Linear Algebra: matrices and tensors as linear functions mapping between elements of the algebra. Functional Differentiation: very briefly... Summary Overview The Multivector Derivative. 2 / 78 Linear Algebra: matrices and tensors as linear functions mapping between elements of the algebra. Functional Differentiation: very briefly... Summary Overview The Multivector Derivative. Examples of differentiation wrt multivectors. 3 / 78 Functional Differentiation: very briefly... Summary Overview The Multivector Derivative. Examples of differentiation wrt multivectors. Linear Algebra: matrices and tensors as linear functions mapping between elements of the algebra. 4 / 78 Summary Overview The Multivector Derivative. Examples of differentiation wrt multivectors. Linear Algebra: matrices and tensors as linear functions mapping between elements of the algebra. Functional Differentiation: very briefly... 5 / 78 Overview The Multivector Derivative. Examples of differentiation wrt multivectors. Linear Algebra: matrices and tensors as linear functions mapping between elements of the algebra. Functional Differentiation: very briefly... Summary 6 / 78 We now want to generalise this idea to enable us to find the derivative of F(X), in the A ‘direction’ – where X is a general mixed grade multivector (so F(X) is a general multivector valued function of X). Let us use ∗ to denote taking the scalar part, ie P ∗ Q ≡ hPQi. Then, provided A has same grades as X, it makes sense to define: F(X + tA) − F(X) A ∗ ¶XF(X) = lim t!0 t The Multivector Derivative Recall our definition of the directional derivative in the a direction F(x + ea) − F(x) a·r F(x) = lim e!0 e 7 / 78 Let us use ∗ to denote taking the scalar part, ie P ∗ Q ≡ hPQi.
    [Show full text]
  • Appendix a Spinors in Four Dimensions
    Appendix A Spinors in Four Dimensions In this appendix we collect the conventions used for spinors in both Minkowski and Euclidean spaces. In Minkowski space the flat metric has the 0 1 2 3 form ηµν = diag(−1, 1, 1, 1), and the coordinates are labelled (x ,x , x , x ). The analytic continuation into Euclidean space is madethrough the replace- ment x0 = ix4 (and in momentum space, p0 = −ip4) the coordinates in this case being labelled (x1,x2, x3, x4). The Lorentz group in four dimensions, SO(3, 1), is not simply connected and therefore, strictly speaking, has no spinorial representations. To deal with these types of representations one must consider its double covering, the spin group Spin(3, 1), which is isomorphic to SL(2, C). The group SL(2, C) pos- sesses a natural complex two-dimensional representation. Let us denote this representation by S andlet us consider an element ψ ∈ S with components ψα =(ψ1,ψ2) relative to some basis. The action of an element M ∈ SL(2, C) is β (Mψ)α = Mα ψβ. (A.1) This is not the only action of SL(2, C) which one could choose. Instead of M we could have used its complex conjugate M, its inverse transpose (M T)−1,or its inverse adjoint (M †)−1. All of them satisfy the same group multiplication law. These choices would correspond to the complex conjugate representation S, the dual representation S,and the dual complex conjugate representation S. We will use the following conventions for elements of these representations: α α˙ ψα ∈ S, ψα˙ ∈ S, ψ ∈ S, ψ ∈ S.
    [Show full text]
  • Multilinear Algebra and Applications July 15, 2014
    Multilinear Algebra and Applications July 15, 2014. Contents Chapter 1. Introduction 1 Chapter 2. Review of Linear Algebra 5 2.1. Vector Spaces and Subspaces 5 2.2. Bases 7 2.3. The Einstein convention 10 2.3.1. Change of bases, revisited 12 2.3.2. The Kronecker delta symbol 13 2.4. Linear Transformations 14 2.4.1. Similar matrices 18 2.5. Eigenbases 19 Chapter 3. Multilinear Forms 23 3.1. Linear Forms 23 3.1.1. Definition, Examples, Dual and Dual Basis 23 3.1.2. Transformation of Linear Forms under a Change of Basis 26 3.2. Bilinear Forms 30 3.2.1. Definition, Examples and Basis 30 3.2.2. Tensor product of two linear forms on V 32 3.2.3. Transformation of Bilinear Forms under a Change of Basis 33 3.3. Multilinear forms 34 3.4. Examples 35 3.4.1. A Bilinear Form 35 3.4.2. A Trilinear Form 36 3.5. Basic Operation on Multilinear Forms 37 Chapter 4. Inner Products 39 4.1. Definitions and First Properties 39 4.1.1. Correspondence Between Inner Products and Symmetric Positive Definite Matrices 40 4.1.1.1. From Inner Products to Symmetric Positive Definite Matrices 42 4.1.1.2. From Symmetric Positive Definite Matrices to Inner Products 42 4.1.2. Orthonormal Basis 42 4.2. Reciprocal Basis 46 4.2.1. Properties of Reciprocal Bases 48 4.2.2. Change of basis from a basis to its reciprocal basis g 50 B B III IV CONTENTS 4.2.3.
    [Show full text]
  • 28. Exterior Powers
    28. Exterior powers 28.1 Desiderata 28.2 Definitions, uniqueness, existence 28.3 Some elementary facts 28.4 Exterior powers Vif of maps 28.5 Exterior powers of free modules 28.6 Determinants revisited 28.7 Minors of matrices 28.8 Uniqueness in the structure theorem 28.9 Cartan's lemma 28.10 Cayley-Hamilton Theorem 28.11 Worked examples While many of the arguments here have analogues for tensor products, it is worthwhile to repeat these arguments with the relevant variations, both for practice, and to be sensitive to the differences. 1. Desiderata Again, we review missing items in our development of linear algebra. We are missing a development of determinants of matrices whose entries may be in commutative rings, rather than fields. We would like an intrinsic definition of determinants of endomorphisms, rather than one that depends upon a choice of coordinates, even if we eventually prove that the determinant is independent of the coordinates. We anticipate that Artin's axiomatization of determinants of matrices should be mirrored in much of what we do here. We want a direct and natural proof of the Cayley-Hamilton theorem. Linear algebra over fields is insufficient, since the introduction of the indeterminate x in the definition of the characteristic polynomial takes us outside the class of vector spaces over fields. We want to give a conceptual proof for the uniqueness part of the structure theorem for finitely-generated modules over principal ideal domains. Multi-linear algebra over fields is surely insufficient for this. 417 418 Exterior powers 2. Definitions, uniqueness, existence Let R be a commutative ring with 1.
    [Show full text]
  • Multilinear Algebra
    Appendix A Multilinear Algebra This chapter presents concepts from multilinear algebra based on the basic properties of finite dimensional vector spaces and linear maps. The primary aim of the chapter is to give a concise introduction to alternating tensors which are necessary to define differential forms on manifolds. Many of the stated definitions and propositions can be found in Lee [1], Chaps. 11, 12 and 14. Some definitions and propositions are complemented by short and simple examples. First, in Sect. A.1 dual and bidual vector spaces are discussed. Subsequently, in Sects. A.2–A.4, tensors and alternating tensors together with operations such as the tensor and wedge product are introduced. Lastly, in Sect. A.5, the concepts which are necessary to introduce the wedge product are summarized in eight steps. A.1 The Dual Space Let V be a real vector space of finite dimension dim V = n.Let(e1,...,en) be a basis of V . Then every v ∈ V can be uniquely represented as a linear combination i v = v ei , (A.1) where summation convention over repeated indices is applied. The coefficients vi ∈ R arereferredtoascomponents of the vector v. Throughout the whole chapter, only finite dimensional real vector spaces, typically denoted by V , are treated. When not stated differently, summation convention is applied. Definition A.1 (Dual Space)Thedual space of V is the set of real-valued linear functionals ∗ V := {ω : V → R : ω linear} . (A.2) The elements of the dual space V ∗ are called linear forms on V . © Springer International Publishing Switzerland 2015 123 S.R.
    [Show full text]
  • CLIFFORD ALGEBRAS Property, Then There Is a Unique Isomorphism (V ) (V ) Intertwining the Two Inclusions of V
    CHAPTER 2 Clifford algebras 1. Exterior algebras 1.1. Definition. For any vector space V over a field K, let T (V ) = k k k Z T (V ) be the tensor algebra, with T (V ) = V V the k-fold tensor∈ product. The quotient of T (V ) by the two-sided⊗···⊗ ideal (V ) generated byL all v w + w v is the exterior algebra, denoted (V ).I The product in (V ) is usually⊗ denoted⊗ α α , although we will frequently∧ omit the wedge ∧ 1 ∧ 2 sign and just write α1α2. Since (V ) is a graded ideal, the exterior algebra inherits a grading I (V )= k(V ) ∧ ∧ k Z M∈ where k(V ) is the image of T k(V ) under the quotient map. Clearly, 0(V )∧ = K and 1(V ) = V so that we can think of V as a subspace of ∧(V ). We may thus∧ think of (V ) as the associative algebra linearly gener- ated∧ by V , subject to the relations∧ vw + wv = 0. We will write φ = k if φ k(V ). The exterior algebra is commutative | | ∈∧ (in the graded sense). That is, for φ k1 (V ) and φ k2 (V ), 1 ∈∧ 2 ∈∧ [φ , φ ] := φ φ + ( 1)k1k2 φ φ = 0. 1 2 1 2 − 2 1 k If V has finite dimension, with basis e1,...,en, the space (V ) has basis ∧ e = e e I i1 · · · ik for all ordered subsets I = i1,...,ik of 1,...,n . (If k = 0, we put { } k { n } e = 1.) In particular, we see that dim (V )= k , and ∅ ∧ n n dim (V )= = 2n.
    [Show full text]