V (v i + W i ) (v i + W i ) is path-connected and hence is connected.

Similar documents
Math 396. Quotient spaces

Math 145. Codimension

Math 110, Spring 2015: Midterm Solutions

Math 676. A compactness theorem for the idele group. and by the product formula it lies in the kernel (A K )1 of the continuous idelic norm

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 ) =

Math 396. An application of Gram-Schmidt to prove connectedness

CHAPTER 7. Connectedness

Lemma 1.3. The element [X, X] is nonzero.

Chapter 2 Linear Transformations

(dim Z j dim Z j 1 ) 1 j i

Parameterizing orbits in flag varieties

Math 541 Fall 2008 Connectivity Transition from Math 453/503 to Math 541 Ross E. Staffeldt-August 2008

Math 396. Bijectivity vs. isomorphism

A PRIMER ON SESQUILINEAR FORMS

Part V. 17 Introduction: What are measures and why measurable sets. Lebesgue Integration Theory

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

THE MINIMAL POLYNOMIAL AND SOME APPLICATIONS

Math 203A, Solution Set 6.

Topological vectorspaces

Introduction to Real Analysis Alternative Chapter 1

The Cyclic Decomposition of a Nilpotent Operator

INVERSE LIMITS AND PROFINITE GROUPS

ALGEBRAIC GEOMETRY COURSE NOTES, LECTURE 2: HILBERT S NULLSTELLENSATZ.

Math 210C. A non-closed commutator subgroup

where Σ is a finite discrete Gal(K sep /K)-set unramified along U and F s is a finite Gal(k(s) sep /k(s))-subset

(2) For x π(p ) the fiber π 1 (x) is a closed C p subpremanifold in P, and for all ξ π 1 (x) the kernel of the surjection dπ(ξ) : T ξ (P ) T x (X) is

Footnotes to Linear Algebra (MA 540 fall 2013), T. Goodwillie, Bases

3 COUNTABILITY AND CONNECTEDNESS AXIOMS

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

Math 594. Solutions 5

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

1 The Local-to-Global Lemma

MAT2342 : Introduction to Applied Linear Algebra Mike Newman, fall Projections. introduction

MAT 445/ INTRODUCTION TO REPRESENTATION THEORY

Thus, X is connected by Problem 4. Case 3: X = (a, b]. This case is analogous to Case 2. Case 4: X = (a, b). Choose ε < b a

Characterizations of the finite quadric Veroneseans V 2n

Maths 212: Homework Solutions

Linear Algebra Notes. Lecture Notes, University of Toronto, Fall 2016

Math 210C. The representation ring

Connectedness. Proposition 2.2. The following are equivalent for a topological space (X, T ).

Dot Products, Transposes, and Orthogonal Projections

Part III. 10 Topological Space Basics. Topological Spaces

2: LINEAR TRANSFORMATIONS AND MATRICES

First we introduce the sets that are going to serve as the generalizations of the scalars.

Math 249B. Geometric Bruhat decomposition

k k would be reducible. But the zero locus of f in A n+1

LINEAR ALGEBRA BOOT CAMP WEEK 1: THE BASICS

Math General Topology Fall 2012 Homework 8 Solutions

10. Noether Normalization and Hilbert s Nullstellensatz

are Banach algebras. f(x)g(x) max Example 7.4. Similarly, A = L and A = l with the pointwise multiplication

Def. A topological space X is disconnected if it admits a non-trivial splitting: (We ll abbreviate disjoint union of two subsets A and B meaning A B =

2. Prime and Maximal Ideals

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

Math 210B. Artin Rees and completions

MATH 51H Section 4. October 16, Recall what it means for a function between metric spaces to be continuous:

ZORN S LEMMA AND SOME APPLICATIONS

2. Duality and tensor products. In class, we saw how to define a natural map V1 V2 (V 1 V 2 ) satisfying

Chapter 2: Linear Independence and Bases

Hilbert spaces. 1. Cauchy-Schwarz-Bunyakowsky inequality

Projective Schemes with Degenerate General Hyperplane Section II

MATH 54 - TOPOLOGY SUMMER 2015 FINAL EXAMINATION. Problem 1

Math 113 Midterm Exam Solutions

TOPOLOGY HW 2. x x ± y

Convergence in shape of Steiner symmetrized line segments. Arthur Korneychuk

Math 396. Subbundles and quotient bundles

DIVISORS ON NONSINGULAR CURVES

MATH Linear Algebra

Smith theory. Andrew Putman. Abstract

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

A NICE PROOF OF FARKAS LEMMA

Extension of continuous functions in digital spaces with the Khalimsky topology

arxiv: v1 [math.gr] 8 Nov 2008

Economics 204 Fall 2012 Problem Set 3 Suggested Solutions

HW 4 SOLUTIONS. , x + x x 1 ) 2

Honors Algebra 4, MATH 371 Winter 2010 Assignment 4 Due Wednesday, February 17 at 08:35

2. Intersection Multiplicities

Vector Space Basics. 1 Abstract Vector Spaces. 1. (commutativity of vector addition) u + v = v + u. 2. (associativity of vector addition)

Categories and Quantum Informatics: Hilbert spaces

Writing proofs for MATH 51H Section 2: Set theory, proofs of existential statements, proofs of uniqueness statements, proof by cases

ISOMETRIES OF R n KEITH CONRAD

Math 121 Homework 4: Notes on Selected Problems

INVARIANT PROBABILITIES ON PROJECTIVE SPACES. 1. Introduction

Theorem 5.3. Let E/F, E = F (u), be a simple field extension. Then u is algebraic if and only if E/F is finite. In this case, [E : F ] = deg f u.

Kernel and range. Definition: A homogeneous linear equation is an equation of the form A v = 0

COMPLEX VARIETIES AND THE ANALYTIC TOPOLOGY

Filters in Analysis and Topology

Reid 5.2. Describe the irreducible components of V (J) for J = (y 2 x 4, x 2 2x 3 x 2 y + 2xy + y 2 y) in k[x, y, z]. Here k is algebraically closed.

ρ Y (f(x 1 ), f(x 2 )) Mρ X (x 1, x 2 )

Algebraic Geometry. Andreas Gathmann. Class Notes TU Kaiserslautern 2014

BRILL-NOETHER THEORY, II

2.2. Show that U 0 is a vector space. For each α 0 in F, show by example that U α does not satisfy closure.

Thus, the integral closure A i of A in F i is a finitely generated (and torsion-free) A-module. It is not a priori clear if the A i s are locally

Math 210B. Profinite group cohomology

Where is matrix multiplication locally open?

Exercises on chapter 0

LECTURE 16: TENSOR PRODUCTS

Math 341: Convex Geometry. Xi Chen

0.2 Vector spaces. J.A.Beachy 1

7 Lecture 7: Rational domains, Tate rings and analytic points

Transcription:

Math 396. Connectedness of hyperplane complements Note that the complement of a point in R is disconnected and the complement of a (translated) line in R 2 is disconnected. Quite generally, we claim that the complement of a translated hyperplane in a finite-dimensional normed vector space over R is disconnected. In fact, this disconnectedness phenomenon is entirely an artifact of codimension 1. Higher codimensions never cause such problems. (Recall that if W is a subspace of a vector space V, the codimension of W in V is defined to be dim(v/w ); this may be infinite, and it may be finite even if dim V and dim W are infinite, but when dim V is finite it is equal to dim V dim W.) As an example of the situation in higher codimension, we can see that removing a translated line (or several such) from R 3 doesn t lead to disconnectedness: we can go around the line when thinking about paths linking up points. Likewise, removing the (codimension 2) origin from R 2 doesn t cause disconnectedness. Roughly speaking, there s enough elbow room complementary to codimensions > 1 to avoid disconnectedness. The aim of this handout is to explore this situation. 1. Main result We prove a connectivity theorem even when V is infinite-dimensional. Of course, to have a reasonable topology we suppose V is equipped with a norm, and we use the resulting metric topology (one can consider the possibility of putting a topology on V in other ways, but we will not discuss that here). Theorem 1.1. Let V be a normed vector space over R. Then for any finite set of (not necessarily distinct) closed linear subspaces W 1,..., W k of (not necessarily finite) codimensions > 1 and any v 1,..., v k V, the complement V (v i + W i ) is path-connected and hence is connected. For example, the complement of any finite configuration of lines in R 3 is path-connected; this is geometrically obvious. Note that the theorem does not require V to be finite-dimensional, nor the W i s to have finite codimension. However, we will construct the paths within well-chosen finitedimensional subspaces of V, so the finite-dimensional case is the essential one. Of course, in the finite-dimensional case the closedness hypothesis on the subspaces is automatically satisfied. Also, infinitude of codimension of W i is not a serious problem either: in fact, the higher the codimension of W i gets, the more room there is complementary to W i, and hence the easier life should be for finding paths! Before we prove the theorem, we record an interesting consequence. Corollary 1.2. Let V be a normed vector space over C, and W 1,..., W k a finite collection of (not necessarily distinct) proper closed linear subspaces. For any v 1,..., v k V, the complement is path-connected and hence is connected. V (v i + W i ) Proof. We can view everything as R-vector spaces at the expense of doubling dimensions and codimensions (when finite). In particular, V/W j is a non-zero C-vector space, whence as an R- vector space has dimension at least 2 (perhaps infinite, which is even better!). Hence, by the theorem, we re done. Now we give the proof of the theorem. 1

2 Proof. The case dim V 1 is trivial. Consider the special case dim V = 2. In this case the only linear subspace of codimension > 1 is the origin, so we re just looking at the complement of a finite set of points. The path-connectedness of this is left to the reader as a pleasant exercise in using definitions. Now consider the general case. Choose two points x, y V not in any of the v i + W i s. We want to find a path connecting them which avoids the complements. Translating everything in sight (i.e., x, y and the v i s) by x, we can assume x = 0 (so 0 is not in any v i + W i, so v i W i for all i). It is exactly the convenience of using such a translation (to reduce to studying paths joined to the origin) that forces us to formulate the original theorem in the context of translated subspaces. Since the linear subspaces W i (and hence translates of them) are closed, the complement V (v i + W i ) is open. We first reduce to the finite-dimensional case. Since V/W i has dimension at least 2, we get vectors v i,1, v i,2 V which induce linearly independent elements in V/W i, which is to say av i,1 + bv i,2 W i for all a, b R not both zero. Let Ṽ be the finite-dimensional subspace of V spanned by x = 0, y, v 1,..., v k, and the vectors v i,1, v i,2 for 1 i k, say given the induced norm from V. Let W i = W i Ṽ. Since av i,1 + bv i,2 W i for all i, clearly Ṽ / W i has dimension at least 2 for all i (this is why we forced the v i,j s to be in Ṽ ). Because of all of the vectors we ve forced into Ṽ, it is easy to see that Ṽ and the W i s with the v i s satisfy all of the original hypotheses (especially the codimension > 1 condition). Hence, if we could settle the finite-dimensional case then we could find a continuous path in Ṽ (v i + W i ) = Ṽ ( ) V (v i + W i ) which joins x = 0 to y. Since Ṽ V is an isometry, this path is also continuous when viewed inside of V, and hence does the job. Thus, we may now assume dim V <, and we will argue by induction on the dimension. Of course, in the finite-dimensional case all norms are equivalent and hence we can essentially suppress mention of the norm. As a preliminary step to help with the induction (basically to allow us to start the induction at dimension 2 rather than having to do an explicit argument in dimension 3 first), we reduce to the case where y is not in any of the W i s. We can find a small open ball B ε (y) around y which is inside of the complement of the closed (v i + W i ), and even avoids touching any of the (finitely many, closed) W i s which don t contain y. We claim there there is a y B ε (y) not contained in any W i s. Indeed, due to how we chose ε, we can use a translation by y to reduce to showing that inside of a given small ball around the origin we can always find a point which avoids any specified finite collection of hyperplanes. But any vector in V admits a non-zero scaling which is inside of B ε (0), so it is equivalent to show that V is not the union of finitely many hyperplanes. This follows from Lemma 2.3 below. Using such a choice of y B ε (y), if we can find a path from 0 to y in the complement of the (v i + W i ) s, then hooking this onto a radial path from y to y in the ball B ε (y) (which is likewise

disjoint from all (v i + W i ) s), we ll be done. Hence, replacing y with a well-chosen sufficiently close y lets us assume that y W i for all i. Now the idea is to take a suitably well-chosen hyperplane slice through y to drop the dimension of V without affecting codimensions of the W i s. This will reduce us to the case dim V = 2 which has already been treated. More specifically, we will find a 2-dimensional subspace V 0 in V which contains y but meets each W i in {0} (this would not be possible if y W i!). Now quite generally, if U, U V are linear subspaces and v, v V are points, then it is easy to see that { (v + U) (v + U, if v v U + U ) = (v + u ) + (U U ), if v v = u + u U + U Thus, back in our original situation, if V 0 is a 2-dimensional subspace of V which contains y and meets each W i in {0} then V 0 (v i + W i ) is either empty or a point. Thus, we have V 0 (V (v i + W i )) = V 0 (V 0 (v i + W i )) with each V 0 (v i + W i ) either empty or a point. Thus, slicing with the subspace V 0 which contains 0 and y brings us to a complement of a finite set in the 2-dimensional V 0, and such a complement is path-connected (and contains y and 0 = x). Our problem is now reduced to a statement in linear algebra which we can prove over an arbitrary infinite field, as in the theorem below (in which the auxiliary vector is y). The required result in linear algebra is treated in the next section. 2. A theorem in linear algebra To emphasize the essentially algebraic (as opposed to topological) aspect of what remains to be done, we now work over an essentially arbitrary field. Theorem 2.1. Let V be a finite-dimensional vector space over an infinite field F, with dim V 2, and let W 1,..., W k be (not necessarily distinct) linear subspaces of codimensions > 1. Choose an auxiliary vector v 0 V with v 0 W i for all i. Then there exists a 2-dimensional subspace V 0 such that v 0 V 0 and V 0 W i = {0} for all i. Remark 2.2. This lemma is false over finite fields. Indeed, over a finite field a finite-dimensional vector space V contains only finitely many vectors, so we can even find finitely many lines (e.g., the span of each non-zero element) whose union is the entire space. Taking {W i } to be the finite set of lines in V, any non-zero subspace certainly contains one of these lines and so no such V 0 as in the theorem can exist. It is a characteristic of infinite fields that a vector space of finite dimension > 1 over such a field cannot be expressed as a finite union of lower-dimensional subspaces, and we will reduce the proof of the theorem to exactly this fact, which will be proven afterwards as a separate lemma (that was already used in earlier arguments). Proof. We can drop any W i s which are equal to 0, so we may assume W i 0 for all i (and that there actually are some W i s). We induct on dim V 2, the case dim V = 2 being clear (as then the W i s are automatically {0}, so we may use V 0 = V ). When dim V > 2, we just need to find a codimension-1 subspace H such that v 0 H and W i H for all i. Indeed, in that case W i +H = V (as W i then surjects onto the 1-dimensional V/H), so dim(w i H) = dim(w i ) + dim(h) dim(w i + H) = dim(w i ) + dim(h) dim(v ) = dim(h) dim(v/w i ) 3

4 In other words, if we slice with a hyperplane H not containing W i, then the codimension dim H/(W i H) = dim(h) dim(w i H) = dim V/W i of W i H in H is equal to the codimension of W i in V, which we assumed to be > 1. Thus, once we find an H containing v 0 which does not contain any of the W i s, then we can replace V with H and each W i with W i H without destroying any of the hypotheses (and any W i H s which vanish can be dropped). Since dim H = dim V 1, by induction on the dimension of V we d be done. Our problem therefore is to find a codimension-1 subspace through v 0 which does not contain any of the non-zero codimension subspaces W 1,..., W k whose codimension in V is > 1. As a concrete example, for V = R 3 this says that, given any finite set of lines L 1,..., L k in R 3 and any point v 0 not on any of these lines, we can find a plane through v 0 which does not contain any of the lines. It is geometrically obvious in this case that a random choice of plane through v 0 will do the job (though a few bad planes may fail). The general argument goes as follows. We can view the problem of constructing a hyperplane H as the problem of finding a suitable non-zero linear functional l : V F (with H then taken to be the codimension-1 kernel of l). In such terms, we seek a point l in the dual space V with l(v 0 ) = 0 but l non-zero on each of the non-zero subspaces W i. This ensures that H = ker l is a hyperplane containing v 0 but not any of the W i s. Consider the annihilator Wi V, which is to say the subspace of functionals which vanish on W i. Since linear maps V V that kill a subspace W uniquely factor through the projection V V/W, by taking the case V = F we arrive at an evident linear isomorphism Wi (V/W i ), so this subspace of V has dimension dim V/W i < dim V = dim V, and hence it is a proper subspace of V (with codimension dim W i ). Let K = (F v 0 ), a codimension-1 subspace of V. Since v 0 W i, we have F v 0 W i, so Wi K (as otherwise applying ( ) to this via V V would yield the reverse inclusion F v 0 W i which we have assumed not to hold). For each i, we claim that the subspace K i = K Wi in K is a proper subspace. If not, so this intersection equals K, then we d get K Wi V, forcing Wi = K since K has codimension 1 in V (so there are no non-trivial intermediate subspaces between K and V ). But we ve just seen that Wi is not contained in K, so this would be a contradiction. Now the task is to show that the vector space K = (F v 0 ) inside of V contains an element which is not inside of any of the proper subspaces K i = Wi K. In other words, we want to show that the vector space K cannot be a union of finitely many proper subspaces (which would be false over a finite field). So far we have not used that F is an infinite field, but it is at this step that the infinitude of F plays the crucial role. We isolate the necessary fact in the form of a lemma below. Lemma 2.3. Let F be an infinite field, V a vector space, and V 1,..., V k finitely many proper subspaces. Then V is not the union of the V j s. Proof. The cases k = 0, 1 are clear. This also settles dim V 1. The idea now is to draw a random line in V and to find a point on this line which is not on any of the V i s. We may assume k > 1 and (by induction on k) the result is known for collections of < k proper subspaces. By induction we can choose a vector v not contained in V 1,..., V k 1. If also v V k, we re done. Otherwise, choose another vector v not contained in the proper subspace V k (so v v). Let L be the span of v v 0, so the translated line v + L = v + L passes through both v and

v. Note that L V i = {0} since this intersection is a proper subspace of the 1-dimensional L (as L contains both v and v, at least one of which is not in V i ). Consider the intersection (v + L) V i = (v + L) V i for 1 i k. Since L V i = {0}, this intersection (v + L) V i is either empty or a point (i.e., it cannot contain two points, as the difference would be a non-zero element in L V i = {0}). Thus, (v + L) ( V i ) = ((v + L) V i ) is a finite (perhaps empty) union of points. But v + L is infinite since L is 1-dimensional over an infinite field. Hence, we can find x v + L not contained in any V i. The union of the V i s is therefore not all of V. 5