The Symmetric Space for SL n (R)

Similar documents
The Hodge Star Operator

CALCULUS ON MANIFOLDS. 1. Riemannian manifolds Recall that for any smooth manifold M, dim M = n, the union T M =

LECTURE 25-26: CARTAN S THEOREM OF MAXIMAL TORI. 1. Maximal Tori

Mostow Rigidity. W. Dison June 17, (a) semi-simple Lie groups with trivial centre and no compact factors and

Some notes on Coxeter groups

274 Curves on Surfaces, Lecture 4

Notes on nilpotent orbits Computational Theory of Real Reductive Groups Workshop. Eric Sommers

ISOMETRIES OF R n KEITH CONRAD

LECTURE 10: THE ATIYAH-GUILLEMIN-STERNBERG CONVEXITY THEOREM

BRUHAT-TITS BUILDING OF A p-adic REDUCTIVE GROUP

Quadratic forms. Here. Thus symmetric matrices are diagonalizable, and the diagonalization can be performed by means of an orthogonal matrix.

Hadamard s Theorem. Rich Schwartz. September 10, The purpose of these notes is to prove the following theorem.

Notation. For any Lie group G, we set G 0 to be the connected component of the identity.

Algebra I Fall 2007

A CONSTRUCTION OF TRANSVERSE SUBMANIFOLDS

1 v >, which will be G-invariant by construction.

(x, y) = d(x, y) = x y.

Clifford Algebras and Spin Groups

MATH 112 QUADRATIC AND BILINEAR FORMS NOVEMBER 24, Bilinear forms

MATH 223A NOTES 2011 LIE ALGEBRAS 35

Basic Properties of Metric and Normed Spaces

Groups up to quasi-isometry

1 Euclidean geometry. 1.1 The metric on R n

Hyperbolic Geometry on Geometric Surfaces

THE FUNDAMENTAL GROUP OF THE DOUBLE OF THE FIGURE-EIGHT KNOT EXTERIOR IS GFERF

fy (X(g)) Y (f)x(g) gy (X(f)) Y (g)x(f)) = fx(y (g)) + gx(y (f)) fy (X(g)) gy (X(f))

William P. Thurston. The Geometry and Topology of Three-Manifolds

4.2. ORTHOGONALITY 161

REGULAR TRIPLETS IN COMPACT SYMMETRIC SPACES

homogeneous 71 hyperplane 10 hyperplane 34 hyperplane 69 identity map 171 identity map 186 identity map 206 identity matrix 110 identity matrix 45

LIE ALGEBRAS AND LIE BRACKETS OF LIE GROUPS MATRIX GROUPS QIZHEN HE

The Lorentz group provides another interesting example. Moreover, the Lorentz group SO(3, 1) shows up in an interesting way in computer vision.

L(C G (x) 0 ) c g (x). Proof. Recall C G (x) = {g G xgx 1 = g} and c g (x) = {X g Ad xx = X}. In general, it is obvious that

The Spinor Representation

Lecture 22 - F 4. April 19, The Weyl dimension formula gives the following dimensions of the fundamental representations:

Math 350 Fall 2011 Notes about inner product spaces. In this notes we state and prove some important properties of inner product spaces.

B 1 = {B(x, r) x = (x 1, x 2 ) H, 0 < r < x 2 }. (a) Show that B = B 1 B 2 is a basis for a topology on X.

Lecture 7: Positive Semidefinite Matrices

HOMEWORK 2 - RIEMANNIAN GEOMETRY. 1. Problems In what follows (M, g) will always denote a Riemannian manifold with a Levi-Civita connection.

The Hurewicz Theorem

A Convexity Theorem For Isoparametric Submanifolds

Tangent spaces, normals and extrema

Since G is a compact Lie group, we can apply Schur orthogonality to see that G χ π (g) 2 dg =

2. Intersection Multiplicities

SYMMETRIC SUBGROUP ACTIONS ON ISOTROPIC GRASSMANNIANS

Geometry of the Special Unitary Group

The Geometrization Theorem

MATH 23a, FALL 2002 THEORETICAL LINEAR ALGEBRA AND MULTIVARIABLE CALCULUS Solutions to Final Exam (in-class portion) January 22, 2003

A strong Schottky Lemma for nonpositively curved singular spaces

Root systems and optimal block designs

DS-GA 1002 Lecture notes 0 Fall Linear Algebra. These notes provide a review of basic concepts in linear algebra.

Course Summary Math 211

The Gelfand-Tsetlin Basis (Too Many Direct Sums, and Also a Graph)

MATRIX LIE GROUPS AND LIE GROUPS

Revisiting Poincaré's Theorem on presentations of discontinuous groups via fundamental polyhedra

Mapping Class Groups MSRI, Fall 2007 Day 2, September 6

Lecture Notes 1: Vector spaces

implies that if we fix a basis v of V and let M and M be the associated invertible symmetric matrices computing, and, then M = (L L)M and the

REFLECTIONS IN A EUCLIDEAN SPACE

1 Dirac Notation for Vector Spaces

NOTES on LINEAR ALGEBRA 1

A geometric proof of the spectral theorem for real symmetric matrices

Math 249B. Geometric Bruhat decomposition

An introduction to arithmetic groups. Lizhen Ji CMS, Zhejiang University Hangzhou , China & Dept of Math, Univ of Michigan Ann Arbor, MI 48109

08a. Operators on Hilbert spaces. 1. Boundedness, continuity, operator norms

Irreducible subgroups of algebraic groups

Vector Spaces, Orthogonality, and Linear Least Squares

We simply compute: for v = x i e i, bilinearity of B implies that Q B (v) = B(v, v) is given by xi x j B(e i, e j ) =

Partial cubes: structures, characterizations, and constructions

Lecture Notes Introduction to Cluster Algebra

MATH 423 Linear Algebra II Lecture 33: Diagonalization of normal operators.

IRREDUCIBLE REPRESENTATIONS OF SEMISIMPLE LIE ALGEBRAS. Contents

PICARD S THEOREM STEFAN FRIEDL

4. Images of Varieties Given a morphism f : X Y of quasi-projective varieties, a basic question might be to ask what is the image of a closed subset

Fix(g). Orb(x) i=1. O i G. i=1. O i. i=1 x O i. = n G

MILNOR SEMINAR: DIFFERENTIAL FORMS AND CHERN CLASSES

MATH Linear Algebra

Pullbacks, Isometries & Conformal Maps

ALGEBRAIC GROUPS J. WARNER

Problems in Linear Algebra and Representation Theory

Functional Analysis. Franck Sueur Metric spaces Definitions Completeness Compactness Separability...

( files chap2 to chap

Topics in Representation Theory: SU(n), Weyl Chambers and the Diagram of a Group

GROUP THEORY PRIMER. D(g 1 g 2 ) = D(g 1 )D(g 2 ), g 1, g 2 G. and, as a consequence, (2) (3)

Smooth Dynamics 2. Problem Set Nr. 1. Instructor: Submitted by: Prof. Wilkinson Clark Butler. University of Chicago Winter 2013

WARPED PRODUCTS PETER PETERSEN

Elementary linear algebra

Black Holes and Thermodynamics I: Classical Black Holes

Random Walks on Hyperbolic Groups III

Introduction to Arithmetic Geometry Fall 2013 Lecture #18 11/07/2013

CHAPTER 8. Smoothing operators

x i e i ) + Q(x n e n ) + ( i<n c ij x i x j

Finite affine planes in projective spaces

How curvature shapes space

(1) is an invertible sheaf on X, which is generated by the global sections

Math 302 Outcome Statements Winter 2013

REPRESENTATIONS OF U(N) CLASSIFICATION BY HIGHEST WEIGHTS NOTES FOR MATH 261, FALL 2001

Representations of algebraic groups and their Lie algebras Jens Carsten Jantzen Lecture III

Survey on exterior algebra and differential forms

Review of Linear Algebra Definitions, Change of Basis, Trace, Spectral Theorem

Transcription:

The Symmetric Space for SL n (R) Rich Schwartz November 27, 2013 The purpose of these notes is to discuss the symmetric space X on which SL n (R) acts. Here, as usual, SL n (R) denotes the group of n n matrices of determinant 1. 1 The Group Action The space X denotes the sets of n n real matrices M such that M is symmetric: M t = M. det(m) = 1. M is positive definite: M(v) v > 0 for all v 0. The action of SL n (R) on X is given by g(m) = gmg t. (1) One can easily verify that g(m) X. Only the third step has anything to it: gmg t (v) v = M(g t (v)) g t (v) = M(w) w > 0. The action we have described really is a group action, because g(h(m)) = g(hmh t )g t = (gh)m(gh) t = gh(m). 1

2 The Tangent Space The identity matrix I is the natural origin for X. The tangent space T I (X) consists of derivatives of the form V = dm ds (0), (2) where M s is a path of matrices in X and M 0 = I. Clearly V is a symmetric matrix. Also, to first order, we have M s = I +sv, and det(i +sv) = 1+s trace(v)+... (3) Hence trace(v) = 0. Conversely, if trace(v) = 0 and V is symmetric, then I +sv belongs to X to first order. More precisely M s = exp(sv) = I +sv + s2 2! V 2 + s3 3! V 3 + (4) gives a path in X, for small s, having V as derivative. In short, T I (X) is the vector space of n n symmetric matrices of trace 0. The differential action of SL n (R) on T I (X) works very conveniently. If g SL n (R) we have In short dg(v) = d ds ( ) g(m s ) = d ds (gm sg t ) = gvg t. dg(v) = gvg t. (5) In other words, the action of g on the tangent space T I (X) is really just the same as the action of g on X. Lemma 2.1 The stabilizer of I is SO(n). Proof: The stabilizer of I consists of those matrices g SL n (R) such that g(i) = I. This happens if and only of gg t = I. These are precisely the orthogonal matrices. That is g SO(n). To see this, note that g(v) g(w) = g t g(v) w. So, g preserves the dot product if and only if g t g = I, which is the same as gg t = I. 2

3 Lemmas about Symmetric Matrices Lemma 3.1 Every M X has the form M = gdg 1 where g SO(n) and D is diagonal. Proof: Introduce the quadratic form Q(v,v) = M(v) v. The lemma is equivalent to the claim that there is an orthonormal basis which diagonalizes Q. To see this, choose a unit vector w 1 which maximizes Q(, ). Next, consider the restriction of M to the subspace W = (w 1 ). If w 2 w 1 = 0 then M(w 2 ) w 1 = 0 because 0 = d ds Q(w 1 +sw 2,w 1 +sw 2 ) = 2M(w 1 ) w 2. (6) But this means that W is an invariant subspace for M, and M W is the matrix defining the quadratic form Q W. By induction, Q W is diagonalizable, with basis {w 2,...,w n }. Then {w 1,...,w n } is the desired basis on R n. Corollary 3.2 Every V T I (X) has the form T = geg 1 where g SO(n) and E is diagonal. Proof: By the previous result, and by definition, we can find a smooth path q s SO(n) and a smooth path D s of diagonal matrices so that V = d ds (g sd s g t s). By the product rule, d ds (g sd s gs) t = gd g t + ( g g t +g(g ) t). 0 Here we have set g = g 0 and D = D (0), etc. The last two terms cancel, because 0 = di dt = d ds (g sg t s) = gd g t +g g t +g(g ) t. Hence V = geg t, g = g 0, E = D. This completes the proof 3

4 The Invariant Riemannian Metric Given A,B T I (X), we define A,B = trace(ab). (7) Lemma 4.1 The metric is invariant under the action of SO(n). Proof: Given g SO(n), we have g t = g 1. Hence g(a),g(b) = gag 1,gBg 1 = trace(gag 1 gbg 1 ) = trace(gabg 1 ) = trace(ab) = A,B. This completes the proof. Lemma 4.2 The metric is positive definite. Proof: We can write V = geg 1, where E is diagonal and g is orthogonal. For this reason V,V = trace(e 2 ) > 0. The point is that E 2 has non-negative entries, not all of which are 0. 5 The Diagonal Slice There is one subspace of X on which it is easy to understand the Riemannian metric. Let X denotethesubsetconsistingofthediagonalmatriceswith positive entries. We can identify with the subspace R n 0 = {(x 1,...,x n ) x i = 0}. (8) The identification carries a matrix to the sequence of logs of its diagonal entries. This makes sense because all the diagonal entries of elements of are positive. In the log coordinates, the diagonal subgroup of SL n (R) acts by translations. Moreover, our metric agrees with the standard dot product on R n 0 4

at the origin, by symmetry. Hence, our log coordinates give an isometry between and R n 0 equipped with its usual Euclidean metric. Itremainstoshowthat isgeodesicallyembedded. Beforewedothis, we mention a cautionary case: In hyperbolic space, if we restrict the hyperbolic metric to a horosphere, we get the Euclidean metric. However, horospheres are not geodesically embedded, and so the zero curvature is a consequence of the distorted embedding. We want to rule out this kind of thing for. Lemma 5.1 is geodesically embedded in X. Proof: We ll show that the shortest paths connecting two points in belong to. Choose some M 1 and let M s be a path connecting M 1 to I = M 0. We can choose continuous paths g s and D s such that g 0 = g 1 = I. g s and D s vary smoothly. Here g s is orthogonal and D s is diagonal. M s = g s D s g t s for all s [0,1]. The path D s also connects M 0 to M 1. We just have to show that D s is not a longer path than M s. We compute d ds M s = A s +B s, where A s = g s D sg t s, B s = g sd s g t s +g s D s (g s) t. (9) Here the primed terms are the derivatives. Note that A s,a s Ms = D s,d s Ds (10) because the action of g s is an isometry of our metric. To finish the proof, we just have to check that A s and B s are orthogonal. To so this, we check that dms 1 (A s ) and dms 1 (B s ) are orthogonal. These are two tangent vectors in T I (X), and we know how to compute their inner product. Dropping the subscript s and using g t = g 1, we compute dm 1 (A) = (gd 1 g 1 )(gd g 1 )(gd 1 g 1 ) = dg 1 D D 1 g 1 = gωg 1, (11) 5

where Ω is some diagonal matrix. Similarly dm 1 (B) = g(ψ 1 +Ψ 2 )g 1, Ψ 1 = D 1 g 1 g, Ψ 2 = (g 1 ) gd 1 = g 1 g D 1. (12) The last equality comes from Equation 6. We compute A,B M = trace ( g(ωψ 1 +ΩΨ 2 )g 1) = trace(ωψ 1 +ΩΨ 2 ) = trace(ωd 1 g 1 g ) trace(ωg 1 g D 1 ) = 1 trace(ωd 1 g 1 g ) trace(d 1 Ωg 1 g ) = 2 trace(ωd 1 g 1 g ) trace(ωd 1 g 1 g ) = 0. (13) Equality 1 is comes from the fact that XY and YX in have the same trace for any matrices X and Y. Equality 2 comes from the fact that D 1 and Ω, both diagonal matrices, commute. 6 Maximal Flats Now we know that X is a totally geodesic slice isometric to R n 1. By symmetry, any subset g( ) X, for g SL n (R), has the same properties. We call these objects the maximal flats. As the name suggests, these are the totally geodesic Euclidean slices of maximal dimension. (We will not prove this last assertion, but will stick with the traditional terminology just the same.) Lemma 6.1 Any two points in X are contained in a maximal flat. Proof: By symmetry, it suffices to consider the case when one of the points is I and the other point is some M. We have M = gdg t where g is orthogonal and D is diagonal. But then g( ) contains both points. It is not true that any two points lie in a unique maximal flat. For generic choices of points, this is true. However, sometimes a pair of points can lie 6

in infinitely many maximal flats. This happens, for instance, when the two points are I and D, and D has some repeated eigenvalues. The maximal flats are naturally in bijection with the subgroups conjugate in SL n (R) to the diagonal subgroup. There is a nice way to picture this correspondence geometrically. Let P denote the real projective space of dimension n 1. Say that a simplex is a collection of n general position points in P. The diagonal subgroup preserves the simplex whose vertices are [e 1 ],...,[e n ], the projectivizations of the vectors in the standard basis. In general, a conjugage group preserves some other simplex; the vertices of the simplex are fixed by all elements of the subgroup. Thus, there is a bijection between the maximal flats and the simplices in P. 7 The Weyl Group and Weyl Chambers The Weyl group is usually defined in terms of the Lie Algebra, but here is a rough and ready geometric description. The Weyl group associated to X is the subgroup W O(n) which acts isometrically on the diagonal slice. When n is odd, we can take W SO(n). In general, W is generated by the permutation matrices. W contains a number of reflections, for instance, the permutation matrix which swaps the first two coordinates. Each such reflection g W fixes some hyperplane H g. The complement H g is a finite union of convex cones. Each such cone is called a Weyl chamber. Here is the first nontrivial example. When n = 3 there are 3 reflections. In log coordinates, the corresponding lines in R 3 0 are the intersections with the coordinate planes. Geometrically, these lines branch out along the 6th roots of unity, and the Weyl chambers fit together like 6 slices of pizza. In general, the points in the interior of the Weyl chambers correspond to matrices having no repeated eigenvalues. Using the action of the diagonal group on, we define, for p, the union C p of convex cones to be the translation of the Weyl chambers to p. Thus, we think of the Weyl chambers as something akin to the way we think of a light cone in Minkowski space: The cone is something that really lives at every point, in the tangent space at that point. We say that a line L in a maximal flat F is regular if, for some p L, the line L points into the interior of some chamber of C p. This definition is independent of the choice of p. If L is not regular, we call L singular. For 7

n = 3, the singular lines in are parallel to the 6th roots of unity, when we use log coordinates. Lemma 7.1 If L F is regular, then L lies in a unique maximal flat. Proof: We will argue by contradiction. Let F 1 = F and let F 2 be some second maximal flat containing L. Let S 1 and S 2 be the corresponding simplices in P. We can assume by symmetry that F 1 = and L contains the origin. Then the matrices of L I have no repeating eigenvalues. Hence, the fixed points of these matrices determine S 1 and S 2. But then S 1 = S 2. Hence F 1 = F 2. Lemma 7.2 If L F is singular, L is contained in infinitely many flats. Proof: Again, we normalize so that L is contained in and goes through I. In this case, the matrices of L I have some repeated entries, and the same entries repeat for all the matrices. Moreover, all the elements of L I have common eigenspaces. Let Σ P denote the union of the projectivizations of the eigenspaces. There are infinitely many simplices S which are compatible with Σ in the sense that each point of S is contained in some projectivized eigenspace. Any such simplex corresponds to a flat containing L. Remark: The last proof is a little bit opaque, so consider the example when n = 3 and L consists of matrices having the first two entries equal. Then Σ is a union of the origin O 0 and the line Λ at infinity in the projective plane. If we choose any two distinct points O 1,O 2 Λ, then the triangle (O 0,O 1,O 2 ) is compatible with Σ. So, the maximal flats form a network of Euclidean slices in X. The maximal flats intersect along the boundaries of the Weyl chambers. This picture suggests a kind of higher dimensional graph, called a building, and indeed when one considers SL n (Q p ), the linear group over the p-adics, the corresponding space X p is indeed known as a building. 8

8 Hyperbolic Slices It is worth mentioning that X contains some totally geodesic copies of H 2, the hyperbolic plane. When n = 2, the corresponding space X 2 is isometric to the hyperbolic plane. We can fit X 2 inside X n by considering diagonal matrices. Precisely, we can make the upper 2 2 block an arbitrary element of X 2, and then we can put (1)s for the other diagonal entries. This embeds X 2 isometrically inside X n. Lemma 8.1 X 2 is a totally geodesic subspace of X n. Proof: Let G SL n (R) be the subgroup of block diagonal matrices having (1) s for the first two entries and then some (n 2) (n 2) diagonal matrix. The action of G fixes X 2 pointwise. Moreover, for any point p X n X 2, there is some g G such that g(p) p. Suppose that x,y X 2 are two points, connected by a geodesic segment γ. Choosing x and y close enough together, we can arrange that γ is the unique distance minimizing geodesic connecting x and y. But if γ X 2 then we can find some g G such that g(γ) γ. But then γ and g(γ) are two distinct geodesic segments, having the same length, connecting x to y. This is a contradiction. Hence γ X 2. Since geodesics locally stay in X 2, the subspace X 2 is geodesically embedded. Once we have one totally geodesic copy of H 2 in X n, we get others by usingtheactionofsl n (R). Thusg(X 2 ) X n isatotallygeodesicembedded copy of H 2 as well. We call these the hyperbolic slices. It turns out that X has non-positive sectional curvature, and the Euclidean slices are the slices of maximum curvature (zero) and the hyperbolic slices are the slices of minimum curvature. 9