(VI.E) Jordan Normal Form

Total Page:16

File Type:pdf, Size:1020Kb

(VI.E) Jordan Normal Form (VI.E) Jordan Normal Form Set V = Cn and let T : V ! V be any linear transformation, with distinct eigenvalues s1,..., sm. In the last lecture we showed that V decomposes into stable eigenspaces for T : s s V = W1 ⊕ · · · ⊕ Wm = ker (T − s1I) ⊕ · · · ⊕ ker (T − smI). Let B = fB1,..., Bmg be a basis for V subordinate to this direct sum and set B = [T j ] , so that k Wk Bk [T]B = diagfB1,..., Bmg. Each Bk has only sk as eigenvalue. In the event that A = [T]eˆ is s diagonalizable, or equivalently ker (T − skI) = ker(T − skI) for all k , B is an eigenbasis and [T]B is a diagonal matrix diagf s1,..., s1 ;...; sm,..., sm g. | {z } | {z } d1=dim W1 dm=dim Wm Otherwise we must perform further surgery on the Bk ’s separately, in order to transform the blocks Bk (and so the entire matrix for T ) into the “simplest possible” form. The attentive reader will have noticed above that I have written T − skI in place of skI − T . This is a strategic move: when deal- ing with characteristic polynomials it is far more convenient to write det(lI − A) to produce a monic polynomial. On the other hand, as you’ll see now, it is better to work on the individual Wk with the nilpotent transformation T j − s I =: N . Wk k k Decomposition of the Stable Eigenspaces (Take 1). Let’s briefly omit subscripts and consider T : W ! W with one eigenvalue s , dim W = d , B a basis for W and [T]B = B. The operator N = T − sI 1 2 (VI.E) JORDAN NORMAL FORM d (with matrix [N]B = B − sId ) is nilpotent, and so fN(l) = l (since its only eigenvalue is 0 ). Since fN(l) must be the product of the invariant factors (of lI − N), the normal form of lI − N is quite limited: (r) (d) n f (lI − N) = diagf1, . , 1, lq ,..., lq g where1 q(r) + ... + q(d) = d. To express the corresponding rational canonical form for N we introduce the following notation: 0 s 1 B .. C q B 1 . C N := B C [q × q matrix]. s B .. .. C @ . A 1 s Then q(r) q(d) (1) R(N) = diag N0 ,..., N0 ; for instance, if 0 1 1 B C B 1 C B C n f (lI − N) = B 1 C B C B 2 C @ l A l3 then 0 1 0 B C B 1 0 C B C B C R(N) = B C . B 0 C B C B 1 0 C @ A 1 0 (Notice the apparent “gap” between the first and second 1’s.) 1once and for all: the numbers (r) ,..., (d) are not exponents, just superscripts. (VI.E) JORDAN NORMAL FORM 3 So for some change-of-basis matrix S we have −1 B − sId = [N]B = S R(N) S or −1 −1 B = S R(N) S + sId = S (R(N) + sId) S . Using (1) this means that if, say, S = SC!B then (r) (d) −1 q q [T]C = S BS = diag N0 ,..., N0 + sId q(r) q(d) = diag Ns ,..., Ns ; for example, 0 1 s B C B 1 s C B C B C n o B C = diag N 2, N 3 . B s C s s B C B 1 s C @ A 1 s q This is called a big Jordan block, and the “boxes” Ns are little Jordan blocks. They have very easy characteristic polynomials, namely (l − s)q (after all, T : W ! W has only eigenvalue s ). The little blocks correspond to a decomposition of W into N -cyclic subspaces W = W(r) ⊕ · · · ⊕ W(d) as in the rational canonical structure theorem (§VI.C). C is a basis subordinate to this decomposition, consisting of a choice of “cyclic vector” for each W(i) and its successive images under N ; more on this later. Going back to a transformation T : V ! V with multiple eigen- values, what we are aiming at is a further decomposition of each of s the Wk = ker (T − skI) into Nk -cyclic subspaces: (r1) (d1) (rm) (dm) V = W1 ⊕ · · · ⊕ W1 ⊕ · · · ⊕ Wm ⊕ · · · ⊕ Wm . 4 (VI.E) JORDAN NORMAL FORM With respect to some special basis C subordinate to this entire direct sum decomposition2 the matrix of T will then be (r ) (d ) ( ) ( ) q 1 q 1 rm dm J ( ) = N 1 N 1 N qm N qm T diag s1 ,..., s1 ;...; sm ,..., sm , where each semicolon separates two big Jordan blocks3 (and each comma two little Jordan blocks). As it stands this looks formidable, and at least the basis C looks very hard to compute. Decomposition of Wk (Take 2). So in order to facilitate compu- tation of the entire Jordan decomposition A = SJ S−1 , it is useful to have an approach to the “nilpotent building blocks” Nk : Wk ! Wk that does not appeal to rational canonical form. DEFINITION 1. The height h of a nilpotent transformation N : W ! W is defined by Nh = 0, Nh−1 6= 0. (The minimal polynomial of such a transformation is simply lh .) One may also define the (N-)height h(w~ ) of w~ 2 W by Nh(w~ )w~ = 0, Nh(w~ )−1w~ 6= 0 . Clearly the height of N is just the supremum of the (N-)heights of vectors 2 W . (The height of a subspace of W, similarly, just means the supremum of heights of vectors in that subspace.) h0 Notation: We shall write ZN (w~ ) for the N -cyclic subspace of W generated by w~ (w~ of height h0 ); that is, spanfw~ , Nw~ ,..., Nh0−1w~ g. CLAIM 2. Given N : W ! W nilpotent of height h , with h−1 h−1 fN w~ 1,..., N w~ tg a basis for im(Nh−1). Then we have a decomposition h h (t+1) (s) W = Z (w~ 1) ⊕ · · · ⊕ Z (w~ t) ⊕ W ⊕ · · · ⊕ W . N N | {z } cyclic of height ≤h−1 2but this is not enough; “special” means the basis also has “cyclic” properties (to be described below) 3there is one for each eigenvalue (VI.E) JORDAN NORMAL FORM 5 PROOF (BY INDUCTION ON h). h = 1 : W = kerN . Take any basis for W ; the spans of the individual basis vectors give the desired decomposition. inductive step : The following picture of W is useful, where N acts by sending you up one level, the top row is kerN , and the height of a vector is its “distance” from the top. h−1 im(N ) (top=0) h−1 h−1 N w N w ker(N) N 1 t N N w w 1 t In order to apply the inductive hypothesis (= the Claim for height h−1 h − 1 ), we consider U = kerN ⊆ W . Now N j U has height h − 1: h−2 im(N| U ) h−1 h−1 h−2 h−2 N w........ N w N u N u 1 t t+1 r Nw 1 ..........Nwt ut+1 ...........u r Set f~u1,...,~utg = fNw~ 1,..., Nw~ tg ⊆ U , and complete h−2 h−2 h−1 h−1 fN ~u1,..., N ~utg = fN w~ 1,..., N w~ tg 6 (VI.E) JORDAN NORMAL FORM h−2 h−2 h−2 to a basis for imf(N j U) g , by adding some fN ~ut+1,..., N ~urg as shown. By the inductive hypothesis, h−1 h−1 h−1 h−1 U = ZN (~u1) ⊕ · · · ⊕ ZN (~ut) ⊕ ZN (~ut+1) ⊕ · · · ⊕ ZN (~ur) ⊕ U(r+1) ⊕ · · · ⊕ U(s) . | {z } N-cyclic of height≤h−2 The picture is now h−1 h−1 h−2 h−2 N w........ N w N u t+1 N ur 2 1 2 t NwNw Nu Nu 1 t t+1 r Nw ..........Nw u ...........u 1 t t+1 r guaranteed by inductive hypothesis — U is the direct sum of the columns (= cyclic subspaces under the action of N ). So why not tack w~ 1,..., w~ t on to the bottom of the first t columns and obtain a decomposition . Nw Nw 1 t w w 1 t as promised? h−1 h−1 Here’s how to do this rigorously: since N w~ 1,..., N w~ t are h−1 a basis for im(N ) , no nontrivial linear combination of w~ 1,..., w~ t can be in ker(Nh−1) = U . Therefore (2) spanfw~ 1,..., w~ tg \ U = f0g, which implies that dim(spanfw~ 1,..., w~ tg) + dim U = dim(spanfw~ 1 ... w~ tg + U). (VI.E) JORDAN NORMAL FORM 7 Rank+Nullity applied to Nh−1 =) dim W = dim(spanfw~ 1,..., w~ tg + U) =) W = spanfw~ 1,..., w~ tg + U. Together with (2) this =) W = spanfw~ 1,..., w~ tg ⊕ U, which by our decomposition of U h−1 h−1 = spanfw~ 1,..., w~ tg ⊕ ZN (~u1) ⊕ · · · ⊕ ZN (~ut) h−1 h−1 (r+1) (s) ⊕ZN (~ut+1) ⊕ ... ⊕ ZN (~ur) ⊕ U ⊕ ... ⊕ U . Since f~u1,...,~utg was just fNw~ 1,..., Nw~ tg , the direct sum in the parentheses is just h h ZN(w~ 1) ⊕ · · · ⊕ ZN(w~ t) and we are done. The upshot of the Claim just proved is this (weaker) PROPOSITION 3. N : W ! W nilpotent =) (1) (s) h1 hs W = W ⊕ · · · ⊕ W = ZN (w~ 1) ⊕ · · · ⊕ ZN (w~ s); that is, W decomposes into a direct sum of N -cyclic subspaces of various heights.4 So if we take h1−1 hs−1 C = fw~ 1, Nw~ 1,..., N w~ 1;...; w~ s, Nw~ s,..., N w~ sg, then n o h1 hs [N]C = diag N0 ,..., N0 .
Recommended publications
  • Regular Linear Systems on Cp1 and Their Monodromy Groups V.P
    Astérisque V. P. KOSTOV Regular linear systems onCP1 and their monodromy groups Astérisque, tome 222 (1994), p. 259-283 <http://www.numdam.org/item?id=AST_1994__222__259_0> © Société mathématique de France, 1994, tous droits réservés. L’accès aux archives de la collection « Astérisque » (http://smf4.emath.fr/ Publications/Asterisque/) implique l’accord avec les conditions générales d’uti- lisation (http://www.numdam.org/conditions). Toute utilisation commerciale ou impression systématique est constitutive d’une infraction pénale. Toute copie ou impression de ce fichier doit contenir la présente mention de copyright. Article numérisé dans le cadre du programme Numérisation de documents anciens mathématiques http://www.numdam.org/ REGULAR LINEAR SYSTEMS ON CP1 AND THEIR MONODROMY GROUPS V.P. KOSTOV 1. INTRODUCTION 1.1 A meromorphic linear system of differential equations on CP1 can be pre­ sented in the form X = A(t)X (1) where A{t) is a meromorphic on CP1 n x n matrix function, " • " = d/dt. Denote its poles ai,... ,ap+i, p > 1. We consider the dependent variable X to be also n x n-matrix. Definition. System (1) is called fuchsian if all the poles of the matrix- function A(t) axe of first order. Definition. System (1) is called regular at the pole a,j if in its neighbour­ hood the solutions of the system are of moderate growth rate, i.e. IW-a^l^Odt-a^), Ni E R, j = 1,...., p + 1 Here || • || denotes an arbitrary norm in gl(n, C) and we consider a restriction of the solution to a sector with vertex at ctj and of a sufficiently small radius, i.e.
    [Show full text]
  • The Rational and Jordan Forms Linear Algebra Notes
    The Rational and Jordan Forms Linear Algebra Notes Satya Mandal November 5, 2005 1 Cyclic Subspaces In a given context, a "cyclic thing" is an one generated "thing". For example, a cyclic groups is a one generated group. Likewise, a module M over a ring R is said to be a cyclic module if M is one generated or M = Rm for some m 2 M: We do not use the expression "cyclic vector spaces" because one generated vector spaces are zero or one dimensional vector spaces. 1.1 (De¯nition and Facts) Suppose V is a vector space over a ¯eld F; with ¯nite dim V = n: Fix a linear operator T 2 L(V; V ): 1. Write R = F[T ] = ff(T ) : f(X) 2 F[X]g L(V; V )g: Then R = F[T ]g is a commutative ring. (We did considered this ring in last chapter in the proof of Caley-Hamilton Theorem.) 2. Now V acquires R¡module structure with scalar multiplication as fol- lows: Define f(T )v = f(T )(v) 2 V 8 f(T ) 2 F[T ]; v 2 V: 3. For an element v 2 V de¯ne Z(v; T ) = F[T ]v = ff(T )v : f(T ) 2 Rg: 1 Note that Z(v; T ) is the cyclic R¡submodule generated by v: (I like the notation F[T ]v, the textbook uses the notation Z(v; T ).) We say, Z(v; T ) is the T ¡cyclic subspace generated by v: 4. If V = Z(v; T ) = F[T ]v; we say that that V is a T ¡cyclic space, and v is called the T ¡cyclic generator of V: (Here, I di®er a little from the textbook.) 5.
    [Show full text]
  • MAT247 Algebra II Assignment 5 Solutions
    MAT247 Algebra II Assignment 5 Solutions 1. Find the Jordan canonical form and a Jordan basis for the map or matrix given in each part below. (a) Let V be the real vector space spanned by the polynomials xiyj (in two variables) with i + j ≤ 3. Let T : V V be the map Dx + Dy, where Dx and Dy respectively denote 2 differentiation with respect to x and y. (Thus, Dx(xy) = y and Dy(xy ) = 2xy.) 0 1 1 1 1 1 ! B0 1 2 1 C (b) A = B C over Q @0 0 1 -1A 0 0 0 2 Solution: (a) The operator T is nilpotent so its characterictic polynomial splits and its only eigenvalue is zero and K0 = V. We have Im(T) = spanf1; x; y; x2; xy; y2g; Im(T 2) = spanf1; x; yg Im(T 3) = spanf1g T 4 = 0: Thus the longest cycle for eigenvalue zero has length 4. Moreover, since the number of cycles of length at least r is given by dim(Im(T r-1)) - dim(Im(T r)), the number of cycles of lengths at least 4, 3, 2, 1 is respectively 1, 2, 3, and 4. Thus the number of cycles of lengths 4,3,2,1 is respectively 1,1,1,1. Denoting the r × r Jordan block with λ on the diagonal by Jλ,r, the Jordan canonical form is thus 0 1 J0;4 B J0;3 C J = B C : @ J0;2 A J0;1 We now proceed to find a corresponding Jordan basis, i.e.
    [Show full text]
  • Jordan Decomposition for Differential Operators Samuel Weatherhog Bsc Hons I, BE Hons IIA
    Jordan Decomposition for Differential Operators Samuel Weatherhog BSc Hons I, BE Hons IIA A thesis submitted for the degree of Master of Philosophy at The University of Queensland in 2017 School of Mathematics and Physics i Abstract One of the most well-known theorems of linear algebra states that every linear operator on a complex vector space has a Jordan decomposition. There are now numerous ways to prove this theorem, how- ever a standard method of proof relies on the existence of an eigenvector. Given a finite-dimensional, complex vector space V , every linear operator T : V ! V has an eigenvector (i.e. a v 2 V such that (T − λI)v = 0 for some λ 2 C). If we are lucky, V may have a basis consisting of eigenvectors of T , in which case, T is diagonalisable. Unfortunately this is not always the case. However, by relaxing the condition (T − λI)v = 0 to the weaker condition (T − λI)nv = 0 for some n 2 N, we can always obtain a basis of generalised eigenvectors. In fact, there is a canonical decomposition of V into generalised eigenspaces and this is essentially the Jordan decomposition. The topic of this thesis is an analogous theorem for differential operators. The existence of a Jordan decomposition in this setting was first proved by Turrittin following work of Hukuhara in the one- dimensional case. Subsequently, Levelt proved uniqueness and provided a more conceptual proof of the original result. As a corollary, Levelt showed that every differential operator has an eigenvector. He also noted that this was a strange chain of logic: in the linear setting, the existence of an eigen- vector is a much easier result and is in fact used to obtain the Jordan decomposition.
    [Show full text]
  • 18.700 JORDAN NORMAL FORM NOTES These Are Some Supplementary Notes on How to Find the Jordan Normal Form of a Small Matrix. Firs
    18.700 JORDAN NORMAL FORM NOTES These are some supplementary notes on how to find the Jordan normal form of a small matrix. First we recall some of the facts from lecture, next we give the general algorithm for finding the Jordan normal form of a linear operator, and then we will see how this works for small matrices. 1. Facts Throughout we will work over the field C of complex numbers, but if you like you may replace this with any other algebraically closed field. Suppose that V is a C-vector space of dimension n and suppose that T : V → V is a C-linear operator. Then the characteristic polynomial of T factors into a product of linear terms, and the irreducible factorization has the form m1 m2 mr cT (X) = (X − λ1) (X − λ2) ... (X − λr) , (1) for some distinct numbers λ1, . , λr ∈ C and with each mi an integer m1 ≥ 1 such that m1 + ··· + mr = n. Recall that for each eigenvalue λi, the eigenspace Eλi is the kernel of T − λiIV . We generalized this by defining for each integer k = 1, 2,... the vector subspace k k E(X−λi) = ker(T − λiIV ) . (2) It is clear that we have inclusions 2 e Eλi = EX−λi ⊂ E(X−λi) ⊂ · · · ⊂ E(X−λi) ⊂ .... (3) k k+1 Since dim(V ) = n, it cannot happen that each dim(E(X−λi) ) < dim(E(X−λi) ), for each e e +1 k = 1, . , n. Therefore there is some least integer ei ≤ n such that E(X−λi) i = E(X−λi) i .
    [Show full text]
  • MA251 Algebra I: Advanced Linear Algebra Revision Guide
    MA251 Algebra I: Advanced Linear Algebra Revision Guide Written by David McCormick ii MA251 Algebra I: Advanced Linear Algebra Contents 1 Change of Basis 1 2 The Jordan Canonical Form 1 2.1 Eigenvalues and Eigenvectors . 1 2.2 Minimal Polynomials . 2 2.3 Jordan Chains, Jordan Blocks and Jordan Bases . 3 2.4 Computing the Jordan Canonical Form . 4 2.5 Exponentiation of a Matrix . 6 2.6 Powers of a Matrix . 7 3 Bilinear Maps and Quadratic Forms 8 3.1 Definitions . 8 3.2 Change of Variable under the General Linear Group . 9 3.3 Change of Variable under the Orthogonal Group . 10 3.4 Unitary, Hermitian and Normal Matrices . 11 4 Finitely Generated Abelian Groups 12 4.1 Generators and Cyclic Groups . 13 4.2 Subgroups and Cosets . 13 4.3 Quotient Groups and the First Isomorphism Theorem . 14 4.4 Abelian Groups and Matrices Over Z .............................. 15 Introduction This revision guide for MA251 Algebra I: Advanced Linear Algebra has been designed as an aid to revision, not a substitute for it. While it may seem that the module is impenetrably hard, there's nothing in Algebra I to be scared of. The underlying theme is normal forms for matrices, and so while there is some theory you have to learn, most of the module is about doing computations. (After all, this is mathematics, not philosophy.) Finding books for this module is hard. My personal favourite book on linear algebra is sadly out-of- print and bizarrely not in the library, but if you can find a copy of Evar Nering's \Linear Algebra and Matrix Theory" then it's well worth it (though it doesn't cover the abelian groups section of the course).
    [Show full text]
  • Your PRINTED Name Is: Please Circle Your Recitation
    18.06 Professor Edelman Quiz 3 December 3, 2012 Grading 1 Your PRINTED name is: 2 3 4 Please circle your recitation: 1 T 9 2-132 Andrey Grinshpun 2-349 3-7578 agrinshp 2 T 10 2-132 Rosalie Belanger-Rioux 2-331 3-5029 robr 3 T 10 2-146 Andrey Grinshpun 2-349 3-7578 agrinshp 4 T 11 2-132 Rosalie Belanger-Rioux 2-331 3-5029 robr 5 T 12 2-132 Georoy Horel 2-490 3-4094 ghorel 6 T 1 2-132 Tiankai Liu 2-491 3-4091 tiankai 7 T 2 2-132 Tiankai Liu 2-491 3-4091 tiankai 1 (16 pts.) a) (4 pts.) Suppose C is n × n and positive denite. If A is n × m and M = AT CA is not positive denite, nd the smallest eigenvalue of M: (Explain briey.) Solution. The smallest eigenvalue of M is 0. The problem only asks for brief explanations, but to help students understand the material better, I will give lengthy ones. First of all, note that M T = AT CT A = AT CA = M, so M is symmetric. That implies that all the eigenvalues of M are real. (Otherwise, the question wouldn't even make sense; what would the smallest of a set of complex numbers mean?) Since we are assuming that M is not positive denite, at least one of its eigenvalues must be nonpositive. So, to solve the problem, we just have to explain why M cannot have any negative eigenvalues. The explanation is that M is positive semidenite.
    [Show full text]
  • Advanced Linear Algebra (MA251) Lecture Notes Contents
    Algebra I – Advanced Linear Algebra (MA251) Lecture Notes Derek Holt and Dmitriy Rumynin year 2009 (revised at the end) Contents 1 Review of Some Linear Algebra 3 1.1 The matrix of a linear map with respect to a fixed basis . ........ 3 1.2 Changeofbasis................................... 4 2 The Jordan Canonical Form 4 2.1 Introduction.................................... 4 2.2 TheCayley-Hamiltontheorem . ... 6 2.3 Theminimalpolynomial. 7 2.4 JordanchainsandJordanblocks . .... 9 2.5 Jordan bases and the Jordan canonical form . ....... 10 2.6 The JCF when n =2and3 ............................ 11 2.7 Thegeneralcase .................................. 14 2.8 Examples ...................................... 15 2.9 Proof of Theorem 2.9 (non-examinable) . ...... 16 2.10 Applications to difference equations . ........ 17 2.11 Functions of matrices and applications to differential equations . 19 3 Bilinear Maps and Quadratic Forms 21 3.1 Bilinearmaps:definitions . 21 3.2 Bilinearmaps:changeofbasis . 22 3.3 Quadraticforms: introduction. ...... 22 3.4 Quadraticforms: definitions. ..... 25 3.5 Change of variable under the general linear group . .......... 26 3.6 Change of variable under the orthogonal group . ........ 29 3.7 Applications of quadratic forms to geometry . ......... 33 3.7.1 Reduction of the general second degree equation . ....... 33 3.7.2 The case n =2............................... 34 3.7.3 The case n =3............................... 34 1 3.8 Unitary, hermitian and normal matrices . ....... 35 3.9 Applications to quantum mechanics . ..... 41 4 Finitely Generated Abelian Groups 44 4.1 Definitions...................................... 44 4.2 Subgroups,cosetsandquotientgroups . ....... 45 4.3 Homomorphisms and the first isomorphism theorem . ....... 48 4.4 Freeabeliangroups............................... 50 4.5 Unimodular elementary row and column operations and the Smith normal formforintegralmatrices .
    [Show full text]
  • On the M-Th Roots of a Complex Matrix
    The Electronic Journal of Linear Algebra ISSN 1081-3810 A publication of the International Linear Algebra Society Volume 9, pp. 32-41, April 2002 ELA http://math.technion.ac.il/iic/ela ON THE mth ROOTS OF A COMPLEX MATRIX∗ PANAYIOTIS J. PSARRAKOS† Abstract. If an n × n complex matrix A is nonsingular, then for everyinteger m>1,Ahas an mth root B, i.e., Bm = A. In this paper, we present a new simple proof for the Jordan canonical form of the root B. Moreover, a necessaryand sufficient condition for the existence of mth roots of a singular complex matrix A is obtained. This condition is in terms of the dimensions of the null spaces of the powers Ak (k =0, 1, 2,...). Key words. Ascent sequence, eigenvalue, eigenvector, Jordan matrix, matrix root. AMS subject classifications. 15A18, 15A21, 15A22, 47A56 1. Introduction and preliminaries. Let Mn be the algebra of all n × n complex matrices and let A ∈Mn. For an integer m>1, amatrix B ∈Mn is called an mth root of A if Bm = A.IfthematrixA is nonsingular, then it always has an mth root B. This root is not unique and its Jordan structure is related to k the Jordan structure of A [2, pp. 231-234]. In particular, (λ − µ0) is an elementary m k divisor of B if and only if (λ − µ0 ) is an elementary divisor of A.IfA is a singular complex matrix, then it may have no mth roots. For example, there is no matrix 01 B such that B2 = . As a consequence, the problem of characterizing the 00 singular matrices, which have mth roots, is of interest [1], [2].
    [Show full text]
  • A SHORT PROOF of the EXISTENCE of JORDAN NORMAL FORM Let V Be a Finite-Dimensional Complex Vector Space and Let T : V → V Be A
    A SHORT PROOF OF THE EXISTENCE OF JORDAN NORMAL FORM MARK WILDON Let V be a finite-dimensional complex vector space and let T : V → V be a linear map. A fundamental theorem in linear algebra asserts that there is a basis of V in which T is represented by a matrix in Jordan normal form J1 0 ... 0 0 J2 ... 0 . . .. . 0 0 ...Jk where each Ji is a matrix of the form λ 1 ... 0 0 0 λ . 0 0 . . .. 0 0 . λ 1 0 0 ... 0 λ for some λ ∈ C. We shall assume that the usual reduction to the case where some power of T is the zero map has been made. (See [1, §58] for a characteristically clear account of this step.) After this reduction, it is sufficient to prove the following theorem. Theorem 1. If T : V → V is a linear transformation of a finite-dimensional vector space such that T m = 0 for some m ≥ 1, then there is a basis of V of the form a1−1 ak−1 u1, T u1,...,T u1, . , uk, T uk,...,T uk a where T i ui = 0 for 1 ≤ i ≤ k. At this point all the proofs the author has seen (even Halmos’ in [1, §57]) become unnecessarily long-winded. In this note we present a simple proof which leads to a straightforward algorithm for finding the required basis. Date: December 2007. 1 2 MARK WILDON Proof. We work by induction on dim V . For the inductive step we may assume that dim V ≥ 1.
    [Show full text]
  • Facts from Linear Algebra
    Appendix A Facts from Linear Algebra Abstract We introduce the notation of vector and matrices (cf. Section A.1), and recall the solvability of linear systems (cf. Section A.2). Section A.3 introduces the spectrum σ(A), matrix polynomials P (A) and their spectra, the spectral radius ρ(A), and its properties. Block structures are introduced in Section A.4. Subjects of Section A.5 are orthogonal and orthonormal vectors, orthogonalisation, the QR method, and orthogonal projections. Section A.6 is devoted to the Schur normal form (§A.6.1) and the Jordan normal form (§A.6.2). Diagonalisability is discussed in §A.6.3. Finally, in §A.6.4, the singular value decomposition is explained. A.1 Notation for Vectors and Matrices We recall that the field K denotes either R or C. Given a finite index set I, the linear I space of all vectors x =(xi)i∈I with xi ∈ K is denoted by K . The corresponding square matrices form the space KI×I . KI×J with another index set J describes rectangular matrices mapping KJ into KI . The linear subspace of a vector space V spanned by the vectors {xα ∈V : α ∈ I} is denoted and defined by α α span{x : α ∈ I} := aαx : aα ∈ K . α∈I I×I T Let A =(aαβ)α,β∈I ∈ K . Then A =(aβα)α,β∈I denotes the transposed H matrix, while A =(aβα)α,β∈I is the adjoint (or Hermitian transposed) matrix. T H Note that A = A holds if K = R . Since (x1,x2,...) indicates a row vector, T (x1,x2,...) is used for a column vector.
    [Show full text]
  • Preliminary Exam 2018 Solutions to Afternoon Exam Part I. Solve Four of the Following Five Problems. Problem 1. Find the Inverse
    Preliminary Exam 2018 Solutions to Afternoon Exam Part I. Solve four of the following five problems. Problem 1. Find the inverse of the matrix 1 0 1 A = 00 2− 0 1 : @5 0 3 A Solution: Any sequence of row operations which puts the matrix (A I) into row- reduced upper-echelon form terminates in the matrix (I B), where j j 3=8 0 1=8 1 0 1 B = A− = 0 1=2 0 : @ 5=8 0 1=8A − 1 Alternatively, one can use the formula A− = (bij), where i+j 1 bij = ( 1) (det A)− (det Aji); − where Aji is the matrix obtained from j by removing the jth row and ith column. Problem 2. The 2 2 matrix A has trace 1 and determinant 2. Find the trace of A100, indicating× your reasoning. − Solution: The characteristic polynomial of A is x2 x 2 = (x+1)(x 2). Thus the eigenvalues of A are 1 and 2, and consequently− the− eigenvalues of −A100 are 1 and 2100. So tr (A) = 1 +− 2100. Problem 3. Find a basis for the space of solutions to the simultaneous equations x + 2x + 3x + 5x = 0 ( 1 3 4 5 x2 + 5x3 + 4x5 = 0: Solution: The associated matrix is already in row-reduced upper-echelon form, so one can solve for the pivotal variables x1 and x2 in terms of the nonpivotal variables x3, x4, and x5. Thus the general solution to the system is ( 2x3 3x4 5x5; 5x3 4x5; x3; x4; x5) = x3v1 + x4v2 + x5v3; − − − − − where v1 = ( 2; 5; 1; 0; 0), v2 = ( 3; 0; 0; 1; 0), v3 = ( 5; 4; 0; 0; 1), and x3, x4, − − − − − and x5 are arbitrary scalars.
    [Show full text]