Lattices and Codes (a brief introduction)

Similar documents
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

Classification of root systems

QUIVERS AND LATTICES.

Lecture 6: Finite Fields

Root systems. S. Viswanath

Problems in Linear Algebra and Representation Theory

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

A Field Extension as a Vector Space

A PRIMER ON SESQUILINEAR FORMS

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

REPRESENTATION THEORY OF S n

THE SEMISIMPLE SUBALGEBRAS OF EXCEPTIONAL LIE ALGEBRAS

Some notes on Coxeter groups

Classification of semisimple Lie algebras

Root systems and optimal block designs

Final Review Sheet. B = (1, 1 + 3x, 1 + x 2 ) then 2 + 3x + 6x 2

The Leech lattice. 1. History.

Linear Algebra (part 1) : Vector Spaces (by Evan Dummit, 2017, v. 1.07) 1.1 The Formal Denition of a Vector Space

FINITE GROUP THEORY: SOLUTIONS FALL MORNING 5. Stab G (l) =.

Lecture 6: Lies, Inner Product Spaces, and Symmetric Matrices

Background on Chevalley Groups Constructed from a Root System

is an isomorphism, and V = U W. Proof. Let u 1,..., u m be a basis of U, and add linearly independent

Math Linear Algebra II. 1. Inner Products and Norms

MATH 223A NOTES 2011 LIE ALGEBRAS 35

(1) A frac = b : a, b A, b 0. We can define addition and multiplication of fractions as we normally would. a b + c d

LINEAR EQUATIONS WITH UNKNOWNS FROM A MULTIPLICATIVE GROUP IN A FUNCTION FIELD. To Professor Wolfgang Schmidt on his 75th birthday

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

5 Quiver Representations

Algebraic number theory Revision exercises

ELEMENTARY SUBALGEBRAS OF RESTRICTED LIE ALGEBRAS

Math 250A, Fall 2004 Problems due October 5, 2004 The problems this week were from Lang s Algebra, Chapter I.

Algebraic Methods in Combinatorics

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.

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

AN ALGEBRA PRIMER WITH A VIEW TOWARD CURVES OVER FINITE FIELDS

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 ) =

Topics in linear algebra

[Disclaimer: This is not a complete list of everything you need to know, just some of the topics that gave people difficulty.]

Hans Wenzl. 4f(x), 4x 3 + 4ax bx + 4c

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

arxiv: v1 [math.rt] 14 Nov 2007

A 2 G 2 A 1 A 1. (3) A double edge pointing from α i to α j if α i, α j are not perpendicular and α i 2 = 2 α j 2

GENERATING SETS KEITH CONRAD

Yale University Department of Mathematics Math 350 Introduction to Abstract Algebra Fall Midterm Exam Review Solutions

ALGEBRA EXERCISES, PhD EXAMINATION LEVEL

EXERCISES IN MODULAR FORMS I (MATH 726) (2) Prove that a lattice L is integral if and only if its Gram matrix has integer coefficients.

SOME EXAMPLES OF THE GALOIS CORRESPONDENCE

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

Linear Algebra I. Ronald van Luijk, 2015

Representation Theory. Ricky Roy Math 434 University of Puget Sound

1 Fields and vector spaces

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

GALOIS GROUPS OF CUBICS AND QUARTICS (NOT IN CHARACTERISTIC 2)

Isomorphisms between pattern classes

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

Spanning, linear dependence, dimension

Math 594. Solutions 5

0.2 Vector spaces. J.A.Beachy 1

ALGEBRA II: RINGS AND MODULES OVER LITTLE RINGS.

Presentation 1

Symplectic representation theory and the Weyl algebra in positive characteristic

Lecture 4.1: Homomorphisms and isomorphisms

Category O and its basic properties

Math 121 Homework 5: Notes on Selected Problems

Math 210C. The representation ring

Representation Theory

Linear Algebra. Min Yan

6 Cosets & Factor Groups

FOURIER COEFFICIENTS OF VECTOR-VALUED MODULAR FORMS OF DIMENSION 2

There are two things that are particularly nice about the first basis

Linear Algebra, Summer 2011, pt. 3

Representation theory and quantum mechanics tutorial Spin and the hydrogen atom

FFTs in Graphics and Vision. Groups and Representations

Linear Algebra, Summer 2011, pt. 2

Longest element of a finite Coxeter group

October 25, 2013 INNER PRODUCT SPACES

A linear algebra proof of the fundamental theorem of algebra

chapter 12 MORE MATRIX ALGEBRA 12.1 Systems of Linear Equations GOALS

A linear algebra proof of the fundamental theorem of algebra

Introduction to modules

SEPARABLE EXTENSIONS OF DEGREE p IN CHARACTERISTIC p; FAILURE OF HERMITE S THEOREM IN CHARACTERISTIC p

9. Finite fields. 1. Uniqueness

Binomial Exercises A = 1 1 and 1

The value of a problem is not so much coming up with the answer as in the ideas and attempted ideas it forces on the would be solver I.N.

Algebraic Methods in Combinatorics

ORDERS OF ELEMENTS IN A GROUP

NOTES ON POLYNOMIAL REPRESENTATIONS OF GENERAL LINEAR GROUPS

(τ) = q (1 q n ) 24. E 4 (τ) = q q q 3 + = (1 q) 240 (1 q 2 ) (1 q 3 ) (1.1)

ROOT SYSTEMS AND DYNKIN DIAGRAMS

INTRODUCTION TO REPRESENTATION THEORY AND CHARACTERS

The Major Problems in Group Representation Theory

Classification of Root Systems

Math 61CM - Solutions to homework 6

MAT 445/ INTRODUCTION TO REPRESENTATION THEORY

IRREDUCIBLE REPRESENTATIONS OF SEMISIMPLE LIE ALGEBRAS. Contents

THE EULER CHARACTERISTIC OF A LIE GROUP

Isomorphisms and Well-definedness

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

LECTURE 4: REPRESENTATION THEORY OF SL 2 (F) AND sl 2 (F)

The Affine Grassmannian

Transcription:

Lattices and Codes (a brief introduction) Ian Coley May 28, 2014 0 Preamble The goal here to is to introduce irreducible root systems, and through them try to classify certain lattices. As we ve seen, lattices are just about the best thing there is in number theory (maybe), since it s easy to pop out a modular form from a lattice under the right circumstances. Luckily, we ll be working in those very circumstances! To get there, we re going to have to define what a code is, somewhat unfortunately. I m going to avoid as much of the code talk as possible, but some of it is pretty useful, such as define the dual lattice and examining un-normalised lattices (i.e. volume not equal to 1). All the things covered in class I will attempt to summarise rather than detail. The source of this talk is Lattices and Codes by Wolfgang Ebeling. 1 Lattices and Codes 1.1 Lattices Definition 1.1. A lattice Γ R n is a subset such that there exists a basis {e 1,..., e n } so that Γ = Ze 1 Ze n. The fundamental parallelotope of Γ is The volume of a lattice is R n /Γ = P = {λ 1 e 1 + + λ n e n : λ i [0, 1]}. vol(r n /Γ) = vol(p ) = det((e 1,..., e n )). Under this definition, we get vol(p ) = 1 for the standard lattice Z n R n. Given two lattices Γ Γ R n, we view Γ as a subgroup of Γ of finite rank, and we see vol(r n /Γ ) = vol(r n /Γ) Γ : Γ. Definition 1.2. Recall that R n is self-dual as a vector space. Specifically, we can let x x be the explicit isomorphism. We define the dual lattice of Γ, denoted Γ, by Γ := Hom(Γ, Z) = {x R n : x y Z for all y Γ}. 1

We will use dot notation because it will be fairly clear when we re multiplying vectors and scalars, so no need for brackets. Note that we don t necessarily have Γ Γ, but we would like to. A lattice is integral in this case, i.e. x y Z for all x, y Γ. It turns out that integral lattices are something a priori worth studying. Proposition 1.3. The integral lattices in R n are precisely the symmetric symmetric bilinear forms over Z for some b : S S Z, where S is a free abelian group of rank n. If Γ is an integral lattice with basis {e i }, then define A := ((e i e j )). Then det A Z, as all its entries are integers. This number is actually independent of choice of basis (which should be pretty clear). In fact, we can derive from the previous discussion that det A = Γ : Γ. We call a lattice unimodular if Γ = Γ. This is also equivalent to, for an integral lattice, det A = 1 or vol(p ) = 1. Hence we have been looking at unimodular lattices for a while! We need one more definition: Definition 1.4. An integral lattice is called even if x 2 = x x 2Z for all x Γ. Equivalently, the matrix A has even elements along its diagonal. 1.2 Codes I will restrict the discussion as much as possible to what is already relevant. Definition 1.5. A code of length n, usually denoted C, is a nonempty proper subset of F n q, where q is a prime power. A code is called linear if it is a linear subspace. Given two codewords x, y C, we can define the (Hamming) distance between x and y as the number of different terms, entry-wise. We might formally use w(x y) to count the number of nonzero terms; hence d(x, y) = w(x y). Since we only really care about linear codes, we will immediately jump into that notation. We write C is an [n,k,d]-code if it is a dimension k subspace of F n q of minimum distance d, that is d = min{d(x, y) : x, y C}. In our case, we always have q = 2. We will now build up (though you don t know it yet) special types of codes analogous to lattices. We define the dual code to C, denoted C, to be C := {y F n q : x y = 0 for all x C}. (This isn t really the definition, but it ought to be.) A linear code C is self-dual if, you guessed it, C = C. Equivalently, C is self-dual if dim n is even, k = n/2, and C C. For a binary code, we say it is doubly even if w(x) 4Z for all x C, where w counts the number of nonzeroes. A doubly even code automatically satisfies C C, which is lovely for us. One code that will be an important example is the extended Hamming code, denoted H, which is a binary [8,4,4]-code which is doubly even and self-dual. More on that later. 2

1.3 Codes Lattices Henceforth all codes are linear, binary [n, k, d]-codes unless otherwise noted. Let ρ be the reduction of Z n modulo 2. Then we may view C ρ(z n ), and it has index 2 n k. Hence ρ 1 (C) Z n is a finite index subgroup, which makes it a lattice with vol(r n /ρ 1 (C)) = 2 n k. Definition 1.6. Γ C := 1 2 ρ 1 (C). Why the fraction? Let s see. It s certainly a lattice in R n. Now take x, y Γ C. Then we can write x = 1 (c + 2z), y = 1 c + 2z 2 2 for c, c C, z, z Z n. Now we see x 2 = 1 2 (c2 + 4cz + 4z 2 ) and x y = 1 2 (c c ) (mod Z). Then we get the following proposition. Proposition 1.7. Let C be a (binary) linear code. (a) C C if and only if Γ C is an integral lattice. (b) C is doubly even if and only if Γ C is an even lattice. (c) C is self-dual if and only if Γ C is unimodular. This was the entire point of mentioning codes to begin with. Now we have two dictionaries to work with in discussion lattices. We now want to talk more about lattices. 1.4 Root Lattices We are going to classify lattices by something called irreducible root systems. It certainly sounds like a good thing to do, even if the words don t make sense yet. Consider an even (integral) lattice Γ. Then define We call the elements roots. R := {x Γ : x 2 = 2}. Definition 1.8. An even lattice Γ R n is a root lattice if R generates Γ. Now let s let Γ be a root lattice. Let x, y R. Then (x y) 2 x 2 y 2 = 4 = x y {0, ±1, ±2}. Further, x y = ±2 if and only if x = ±y, and if x y = 1, then (x y) 2 = 2 and thus x y R and conversely. Now we begin to get to the point. 3

Theorem 1.9. Let Γ R n be a root lattice. Then Γ contains a basis {e i } so that e 2 i = 2 for all i {1,..., n} and e i e j {0, 1}, i j, i, j {1,..., n}. This is the goal of the section. We have the following definition which will help in the proof. Definition 1.10. A subset S R is called a fundamental system of roots if (a) S is a basis of Γ; (b) Each β R can be written as a linear combination β = α S k α α such that all k α are non-positive or all are non-negative. Proving that every root lattice contains a fundamental system of roots proves the theorem. We can construct these systems in general, and so let us begin to do so. For each α R, let us define the hyperplane H α := {x R n : x α = 0}. Since R is finite, we know α R H α R n, so there exists some t R n so that t α 0 for all α R. We can divide R into two groups; for such a (particular) t, let R + t := {α R : t α > 0} so that R = R t + R t +. We will call R t + the positive roots. A root α R t + is called decomposable if there exist β, γ R t + so that α = β + γ, and indecomposable otherwise. Let S t R t + be the set of indecomposable positive roots. Proposition 1.11. S t is a fundamental system of roots for Γ. Conversely, given a fundamental system of roots S, then there exists t R n so that S = R + t. Proof. We will proceed in steps. is a linear combination of elements of S t with non- Lemma 1.12. Each element of R t + negative integer coefficients. Suppose not. Then let α R t + without this property. Further, we may assume we have chosen this α so that t α > 0 is minimal (since we are working with a finite set). Then suppose α is decomposable. Let α = β + γ. Then since we are working with positive roots, t α = t β + t γ = t α > t β, t γ. Therefore β, γ must be a linear combination of elements of S t with non-negative integer coefficients. But the sum of two such combinations also has non-negative coefficients, a contradiction. We have thus reduced to α S t lacking this property, which is absurd. 4

Lemma 1.13. For α β S t, we have α β {0, 1}. As we showed above, α β = ±2 if and only if α = ±β and α β = 1 if and only if α β R. In the first case, we cannot have α, α S t, as exactly one of these may be positive. In the second case, let γ = α β. Then α = β + γ and β = α γ, which contradictions the indecomposability of α, β. Lemma 1.14. The elements of S t are linearly independent. Suppose we had a minimal relation α S t a α α = 0. Then splitting the k α into positive and negative terms, we may rewrite this bβ β = c γ γ, b β, c γ > 0, {β} {γ} =. Let λ be this lefthand sum. Then λ 2 = β,γ b β c γ (β γ) 0 since β γ 0 for each β γ S t. Therefore λ = 0. Further, all its coefficients are zero as 0 = t λ = b β (t β), as each t β > 0. This will also show c γ are all zero, so we are done. The above lemmas show that S t is a basis for R n, so the fundamental system of roots is the required basis in any root lattice. Lemma 1.15. Let {γ 1,..., γ n } be a basis of R n. Then there exists an element t R n with t γ i > 0 for all i {1,..., n}. Let δ i be the projection of γ i onto the subspace spanned by the other n 1 basis vectors, and let δ i := γ i δ i. Then { > 0 i = j δ i γ j = 0 i j. This is seen by appealing to the orthonormal basis associated to {γ i } by the Gram-Schmidt process. Therefore the element t = n i=1 δ i fits the bill. Now for the converse of the proposition, given a fundamental system of roots S for R, the existence of a t R n so that t α > 0 for all α S is guaranteed by the last lemma. Let R + R be those roots which are purely nonnegative linear combinations of elements of S. Then R + R t + and R + R t +, so this shows R + = R t +. Therefore the elements of S are the indecomposable elements of R t +, and S S t. Since S = S t = n, we are done. So we come to the end, and the good question here is so what? Given a fundamental system of roots S = {e 1,..., e n }, we can associate to it a graph, called a Coxeter-Dynkin diagram, constructed as follows: let V = S be the vertex set, and draw an edge between e i and e j if and only if e i e j = 1. Let G be this graph. What can we say about it from our proofs so far? Lemma 1.16. G has no cycles. 5

Proof. Let x 1 x r be a minimal cycle of G, r 3. Then we have (x 1 + + x r ) 2 = r x 2 i + i j i=1 x i x j = 2r 2r = 0. But this is impossible as all roots have x 2 = 2. Lemma 1.17. G does not contain a subgraph of the form (for r 1) r+1 r+4 1 2 r 1 r r+2 r+3 Proof. Given that graph (and throwing in x for notation), we have (2x 1 + + 2x r + x r+1 + + x r+r ) 2 = 4 This is also impossible. r x 2 i + 4 i=1 1 i j r x i x j + r+4 i=r+1 + 2(x r+4 x r + x r+3 x r + x r+2 x 1 + x r+1 x 1 ) = 8r + 8 2(r 1) 16 = 0. Therefore if G is a connected graph, then it must be of the form below, for q, p, r 1. x 2 i 6

y q 1 x p 1 y q 2 x p 2...... y 1 x 1 c z1. zr 2 We can now analyse G further. Define the point zr 1. w := c + 1 p 1 (p i)x i + 1 q 1 (q j)y j + 1 r 1 (r k)z k. p q r i=1 j=1 Then the collection {x i, y j, z k, w} are a Q-basis for Γ Q. Additionally, the x i, y j, z k are (between each other) pairwise orthogonal, and w is orthogonal to all other basis elements. Therefore w 2 = w c = 1 p + 1 q + 1 r 1. Since we must have w 2 > 0, we have 1 p + 1 q + 1 r > 1. This yields two general solutions and three particular ones: (p, q, r) = (1, q, r), k=1 q r arbitrary, (2, 2, r), r arbitrary, (2, 3, 3), (2, 3, 4), (2, 3, 5). 7

A lattice Γ is called reducible if it can be written Γ = Γ 1 Γ 2 for two lattices Γ 1 R n 1 and Γ 2 R n 2 with n 1, n 2 1 and n 1 + n 2 = n. A lattice is irreducible if it is not reducible, and in this case its Coxeter-Dynkin diagram is connected. In general, a root lattice Γ can be decomposed as Γ = Γ i, where Γ i are irreducible, and the Coxeter-Dynkin diagram of Γ has as connect components the Coxeter-Dynkin diagrams of the Γ i. Hence Theorem 1.18. Every root lattice is the orthogonal direct sum of the irreducible root lattices with the following Coxeter-Dynkin diagrams: (n vertices) A n (n vertices) D n (n 4) E 6 E 7 E 8 These are the only theoretically possible Coxeter-Dynkin diagrams. But are there actual lattices that correspond to these? This is the question of existence, and indeed, they do exist. To construct A n : take the standard basis {e i } of R n+1. Then consider the subspace orthogonal to e = e i, which is isomorphic to R n. We may take the sublattice Γ of the standard lattice Z n+1 R n+1 contained in e, which is of rank n. This lattice is generated by elements of the form e i e j for i j, which indeed are its roots. We have R = n(n + 1), and there is a fundamental system of roots {e 2 e 1, e 3 e 2,..., e n+1 e i }. The Coxeter-Dynkin diagram of this basis is seen to be A n, as (for i, j {1,..., n}) (e i+1 e i ) (e j+1 e j ) = e i+1 e j+1 e i e j+1 e i+1 e j + e i e j = 2δ i (j) δ i (j + 1) δ i+1 (j) 2 i = j = 1 i = j ± 1. 0 else 8

To construct D n : Note that D 2 makes no sense, because to form the requires at least 3 vertices. Further, D 3 = A3, so we ignore this case to make our classification unique. Therefore for n 4, consider Γ = {(x 1,..., x n ) Z n : x i 2Z}. This is indeed a lattice, and it is generated by (for the standard basis {e i } of R n ) the points ±e i ± e j and ±e i e j for i j, of which there are 2n(n 1). Then we have a fundamental system of roots {e 2 e 1,..., e n e n 1, e 1 + e 2 }. Then we have A n 1 embedded as a subgraph corresponding to the obvious basis elements. We need only check e 1 + e 2 against an arbitrary element: for i {1,..., n 1}, (e 1 + e 2 ) (e i+1 e i ) = e 1 e i+1 e 1 e i + e 2 e i+1 e 2 e i = δ 1 (i + 1) δ 1 (i) + δ 2 (i + 1) δ 2 (i) { 1 i = 2 =. 0 else Therefore we attach the vertex e 1 + e 2 to the second point in A n 1, which forms D n. It turns out that the extended Hamming code H described above generates E 8. It is now time to write it down: we have H := span{(0, 1, 1, 0, 1, 0, 0, 1), (0, 0, 1, 1, 0, 1, 0, 1), (0, 0, 0, 1, 1, 0, 1, 1), (1, 0, 0, 0, 1, 1, 0, 1) (0, 1, 0, 0, 0, 1, 1, 1), (1, 0, 1, 0, 0, 0, 1, 1), (1, 1, 0, 1, 0, 0, 0)} F 7 2. Recall that we define Γ H by 1 2 times the pullback of the code to R 8, which we have already shown is an even unimodular lattice (viz. the most interesting kind of lattice). Let f 1,..., f 7 be defined by the above vectors scaled by 1 2. Then we have f 2 i = 2 for each vector, and f i f j = 1 for i j. From here, we can (begin to) construct a fundamental root system for Γ H: let e 1 = f 1 and for i > 1, let e i = f i f i 1. Together, these elements form a subgraph isomorphic to A 7 of the diagram of Γ H. However, we need one more basis element to finish off the construction. We take e 8 := 1 2 ( 1, 1, 0, 0, 1, 0, 1, 0), which satisfies e 2 8 = 2, e 8 f i = { 0 i = 1, 2, 3, 4 1 i = 5, 6, 7 Hence e 8 e i = 1 if and only if i = 5 and e 8 e i = 0 otherwise. This yields E 8. A slight modification of this construction will also yield E 7 and E 6. At this point it is worth noting that we are missing a few letters in the above, obviously B and C, but also F and G, though you wouldn t know it. The above process is precisely how one goes about classifying simple Lie algebras, of which there are four infinite families and five exceptional ones (and maybe I will write these down at this point). In the general, case, the graphs considered are directed (as the inner product defined on the roots is not symmetric), and we have situations where α i α j {±2, ±1, 0} are all possible. The reader should go look this up if it s interesting. 9.

1.5 The Weyl Group Let Γ be a root lattice in R n, and let α R be a root. Consider the linear transformation s α : R n R n, x x 2(x α)α the reflection of R n about the hyperplane α. This is an invertible linear transformation, so lives in GL n (R). Let W (Γ) GL n (R) be the subgroup generated by s α for all α R. Lemma 1.19. Let Γ R n be an irreducible root lattice. Then R n is irreducible as a W (Γ)-module. Proof. Suppose U R n is a W (Γ)-submodule. Then since R[W (Γ)] is a semisimple algebra, U is a W (Γ)-submodule as well. Let α Γ be a root, and suppose that α / U. Let u U be any element. Since s α (U) = U, we have u (u α)α U, whence u α = 0 else we find α U. Therefore α U. Therefore every root lies in U or U. Since R spans Γ, Γ = (U Γ) (U Γ). But Γ was assumed to be irreducible, so U Γ = or U Γ = Γ, which implies that U = 0 or R n. Lemma 1.20. Let Γ R n be a root lattice. Then W (Γ) acts transitively on the set of roots R. Proof. Let α, β R be two roots. Then the set {w(α ) : w W (Γ)} spans R n, as it is a nonzero W (Γ)-submodule of R n. Therefore choose some α := w(α ) not orthogonal to β. If α = β, then we are done as we have found an element of the Weyl group transformation any root into another. Hence suppose α β, so α β = ±1. Without loss of generality, we may replace β by s β (β) so that α β = 1. Then This completes the proof. s α s β s α (β) = s α s β (β (β α)α) = s α (β α β) = α. This may be a terrible aside (as it is not in Ebeling s book), but the following discussion is very interesting, which I presented in the Fall for the Algebraic Groups seminar. The extra material comes from Humphreys Linear Algebraic Groups. Viewing a root lattice Γ R n with roots R, we can consider Aut R := {σ GL n (R) : σ(r) = R}, those automorphisms of R n that preserve the roots. It is clear that W (Γ) Aut R. Further, we can define the automorphism group of the Coxeter-Dynkin diagram G of Γ, denoted Aut G, by permutations of the vertices that preserve the edge relations. Further, we can view an automorphism of G as an automorphism of R by pulling back the permutation of the vertices indexed by a fundamental system of roots S to a permutation of S as a generating set of R. This gives us a sequence of groups 1 W (Γ) Aut R Aut G 1. Proposition 1.21. (Humphreys 12.2) The above sequence is exact. Proof. We need a few lemmas to prove this. 10

Lemma 1.22. (Humphreys 9.2) W (Γ) is a normal subgroup of Aut R. It suffices to show that the set of generators {s α } is fixed under conjugation by elements of Aut R. Let σ Aut R. Then note that, for a root β, since s α (β) R. But σs α σ 1 (σ(β)) R σs α σ 1 (σ(β)) = σ(β (β α)α) = σ(β) (β α)σ(α) = σ(β) (σ(β) σ(α))σ(α), where we use the fact that α β = σ(α) σ(β), which I believe is reasonably clear, though caused much arguing when first presented. Since σ(β) runs over R as β runs over R, we have shown that σs α s 1 fixes R. Moreover, it sends σ(α) to σ(α) and fixes the hyperplane orthogonal to σ(α), so it is equal to s σ(α). Recall that the choice S t of a fundamental system of roots depended on the choice of t R n, but we proved that any choice would do. Of course by a simple counting argument for a given system S there are many many t R n such that S t = S. Hence we can consider the set S of fundamental systems of roots of R (of a lattice Γ). There is an obvious action of W (Γ) on S where w : {e 1,..., e n } {w(e 1 ),..., w(e n )}, but it is not a priori clear that what we get is still a fundamental system of roots. Fix a system of roots S = {e i }. Let w W (Γ). Then since W (Γ) acts transitively on the set of roots, for each α R there exists β R so that w(β) = α. Thus if we have ( n n ) k i w(e i ) = α = w(β) = w k i e i = w(β) = k i 0 or k i 0, i=1 i=1 so this (along with the other discussion above) shows that this action makes sense at all. Lemma 1.23. (Humphreys 10.3) The above action of W (Γ) on S is transitive. Additionally, the action of W (Γ) on a particular fundamental system of roots S is simply transitive, that is for any e i, e j S, there is a unique w W (Γ) so that w(e i ) = e j. This is a fairly nontrivial proof, and since I am likely running over time already, I will omit it. To complete the proof, we will show that Aut R is a semidirect product of Aut G and W (Γ). For a fixed basis S, define the subgroup H := {σ Aut R : σ(s) = S} Aut R. This subgroup is isomorphic to Aut G, as it is precisely those automorphisms preserving the vertices and edge relations of G. To show that Aut R = W (Γ) H, it is necessary and sufficient to show that W (Γ) Aut R(which we have done), H W (Γ) = {1}, and W (Γ) H = Aut R. Suppose that σ H W (Γ). By the second lemma, we have σ = 1 by simple transitivity. Now for any τ Aut R, we have shown that τ(s) is still a fundamental system of roots for R. Since W (Γ) acts transitively on S, there exists w so that w(τ(s)) = S. Therefore wτ H, so τ W (Γ) H. This completes the proof. 11

This allows us to easily calculate the Weyl groups for A n, since these graphs have very simple automorphism groups. Indeed, we have Aut G = Z/2Z for each of these graphs. Recall that the roots for A n (in the construction above) are R = {e i e j : 1 i j n + 1}. We may permute these roots by S n+1, giving the embedding S n+1 Aut R. Further, there is the map e i e j e n+1 i e n+1 j, which is not given by any permutation, which corresponds to the flip of the Coxeter-Dynkin diagram. This gives Aut R = S n 1 Z/2Z, and W = S n+1. There are a few more interesting things to say about root systems in general. The index of Γ Γ and something called the highest root of a root lattice would be fun to discuss, but fearing lack of time, I m not going to bother writing anything at this time. Thus ends our discussion of root lattices per se. 2 Even Unimodular Lattices 2.1 Motivation As stated a couple times, these are the most interesting kinds of lattices. Recall that for a general lattice Γ R n, we associate the theta function ϑ Γ (τ) := x Γ q 1 2 x2 defined on the upper half plane H. If Γ is an even integral lattice, then 1 2 x2 N for all x Γ. Therefore we may write ϑ Γ (τ) = a r q r, where a r = #{x Γ : x 2 = 2r}. We ve further proved in class that for an even unimodular lattice Γ R n, we have n 0 (mod 8) and ϑ Γ is a modular form of weight n/2, in the sense that ( ) aτ + b ϑ Γ = (cτ + d) n/2 ϑ Γ (τ). cτ + d Using all the notation from class, let M k be the space of weight k modular forms (with trivial character and with respect to the full modular group). Then M := k=0 M k is a graded C-algebra. We proved the following proposition in class: r=0 Proposition 2.1. as a C-algebra, where M = C[E 4, E 6 ], E 4 (τ) = 1 + 240 σ 3 (r)q r, E 6 (τ) = 1 504 σ 5 (r)q r r=1 r=1 are the usual Eisenstein series. This gives the following corollary. 12

Corollary 2.2. Let Γ be an even unimodular lattice in R 8. Then Γ is isomorphic to E 8. Proof. Let ϑ Γ be the theta function of Γ. This is a modular form of weight 4, and we know that M 4 = C E 4 as a C-vector space. Since we know that ϑ Γ has constant term 1, we know that ϑ Γ = E 4 on the nose. Therefore a 1 = #{x Γ : x 2 = 2} = 240, so Γ has 240 roots. These roots generate a root sublattice of rank at most 8. However, we calculated the number of roots in A 8 and D 8 to be 72 and 102, respectively, and it can be shown E 6 and E 7 have 72 and 126 roots, respectively. Hence these roots must generate E 8, and thus Γ = E 8. 2.2 Theta Functions with Spherical Coefficients The goal now will be to classify even unimodular lattices of rank 24 via their fundamental systems of roots by studying a special class of theta functions. Definition 2.3. Let P C[x 1,..., x n ] be a homogeneous complex polynomial of degree r. We say that P is spherical of degree r if P = 0, where = n x 2 i=1 i is the Laplace operator. We have a convenient characterisation of these polynomials. Theorem 2.4. P C[x 1,..., x n ] is spherical of degree r if and only if P is a linear combination of functions of the form (ξ x) r, where ξ C n, ξ 2 = 0 if r 2. The proof of this theorem is not particularly enlightening for the present discussion. Now we can construct functions on the upper half plane based around spherical polynomials. Definition 2.5. Let Γ R n be a lattice, z R n a point, and P a spherical polynomial of degree r. Then we define ϑ z+γ,p (τ) := P (x + z)e πiτ(x+z)2. x z+γ P (x)e πiτx2 = x Γ These functions are holomorphic in H and satisfy ϑ z+γ,p = ϑ z +Γ,P if z z Γ. Ideally, these functions are modular forms, so we can apply the corpus of knowledge we have built up. Hence, we examine how they transform with respect to the modular group. Proposition 2.6. We have the identity ( ϑ z+γ,p 1 ) ( ) n+2r τ = i r 1 P (y)e 2πiy z e πiτy2. τ i vol(r n /Γ) y Γ 13

This follows via Poisson summation and the relative niceness of spherical polynomials. Now we make the assumption that Γ is an even integral lattice, n is even, and we define k := n/2+r. In this case, Γ Γ, and for y, y Γ, we have y y Γ = y x y x Z and y 2 y 2 2Z. Therefore we have the (much more useful) formulae, if we let ρ Γ : ϑ ρ+γ,p (τ + 1) = e πiρ2 ϑ ρ+γ,p (τ), ( ϑ ρ+γ,p 1 ) 1 ( τ ) k = i r τ vol(r n /Γ) i σ Γ /Γ e 2πiσ ρ ϑ σ+γ,p (τ). This is not quite what we would like. Therefore we find a subgroup of the modular group under which these theta functions transform more favourably. Something that I did not mention earlier is that, given a binary linear code C, we have 2Γ Γ. We want to do something similar in the general case. Definition 2.7. The level of Γ, denoted N, is the least N N so that Nµ 2 2Z for all µ Γ. Lemma 2.8. If N is the level of Γ, NΓ Γ. We define the subgroups {( ) } a b Γ 0 (N) := SL c d 2 (Z) : N c, {( ) ( ) a b a b Γ(N) := SL c d 2 (Z) : c d ( ) 1 0 0 1 } mod N, where N is the level of Γ. Since we alluded to this in class, I will skip the foreplay and get right to the point. Theorem 2.9. For Γ(N), ϑ ρ+γ,p usual notation, is a modular form of weight k. In particular, if k is the ϑ ρ+γ,p k A = ϑ ρ+γ,p ( ) ϑ Γ,P k A = ϑ Γ,P d A Γ(N), A Γ 0 (N), where ( ) is the Legendre symbol and = ( 1) n/2 Γ : Γ = ( 1) n/2 vol(r n /Γ) 2. 2.3 Root Systems in Even Unimodular Lattices That last section was perhaps somewhat rushed, but the point was to make it this far. Corollary 2.10. Let Γ R n be an even unimodular lattice, and let P be as spherical polynomial in n variables of degree r. Then ϑ Γ,P is a modular form of weight n/2 + r, and a cusp form if r > 0. 14

Proof. The first assertion follows because, for an even unimodular lattice, N = 1 and Γ = Γ, and (as we have proved) n 0 (mod 8). Now we know ϑ Γ,P (τ) = c s q s, s=0 where q = e 2πiτ, and c s := x Γ x 2 =2s P (x). If r > 0, then c 0 = P (0) = 0, so ϑ Γ,P is a cusp form. We want to use this to classify even unimodular matrices of dimension 24, since modular forms of weight 12 are pretty much the only thing we have a terrific handle on. Let Γ be such a lattice and let R be its set of roots. Let f(x) := (x y) 2 x2 y 2 for some fixed y R n. Then f is a spherical polynomial of degree 2, and hence ϑ Γ,f is a cusp form of weight n/2 + 2. So for n = 8, 16, 24, we have a cusp form of weight 6, 10, and 14 respectively. But recall that for forms of weight 6 and 10, M k = C E k, and this turns out to be true as well for k = 14. Thus if we write ϑ Γ,f = c r q r, r=0 we have c r = 0 for all r. Recalling from a few lines ago what these coefficients mean, we have Proposition 2.11. For an even unimodular lattice Γ R n for n = 8, 16, 24, we have for a fixed y R n (x y) 2 x 2 y2 n = 0. In particular, x Γ x 2 =2r x Γ x 2 =2r n (x y) 2 = y2 n 2 R. x R Corollary 2.12. Either R = or R spans R n. Proof. If R does not span R n, then let 0 y R n which is orthogonal to all the elements of R. Then the above formula proves R = 0. Corollary 2.13. Let Γ R n be an even unimodular lattice, and let n = 8, 16, 24. Then all irreducible components of the root lattice spanned by R have the same Coxeter number h, where h = 1 n R. 15

Proof. This proof comes from the above proposition and the following one: Proposition 2.14. Let Γ R n be an irreducible root lattice. Then for a fixed y R n, we have (x y) 2 = 2 h y 2, x R We will take this proposition on faith. Now what does this mean? We have already shown that for n = 8, any even unimodular matrix is isomorphic to E 8. Take n = 16, and let Γ R 16 be an even unimodular lattice. Then θ Γ is a modular form of weight 8, and we have shown that ϑ Γ = E 2 4. Therefore R = 480, as we can directly calculate the coefficient of q of E 2 4 given we know E 4 (τ) = 1 + 240q + From the above, we see that R spans R 16 and each irreducible component of the root lattice spanned by R, which we denote R Z have Coxeter number 30. Since we have exhausted the possibilities for the irreducible components, we can see that R Z = E 8 + E 8 or D 16. Proposition 2.15. Let Γ R 24 be an even unimodular lattice. Then R Z is one of the following 24 lattices: 0, 24A 1, 12A 2, 8A 3, 4A 6, 3A 8, 2A 12, A 24, 6D 4, 4D 6, 2D 12, D 24, 4E 6, 3E 8, 4A 5 + D 4, 2A 7 + 2D 5, 2A 9 + D 6, A 15 + D 9, E 8 + D 16, 2E 7 + D 10, E 7 + A 17, E 6 + D 7 + A 11. Proof. These are the only possibilities for root systems R that satisfy the following properties: (a) R spans R 24 ; (b) All of the irreducible components of R Z have the same Coxeter number. If we let then condition (a) means that 24 24 R Z = α i A i + δ j D j + i=1 j=1 24 24 24 = i α i + j δ j + i=1 j=1 8 η k E k, k=6 8 k η k. Condition (b) gives the following quantities, all equal to 24, each of which corresponds to a set of solutions: And that is all! k=6 i α i, j δ j, k η k, (2j 3)α 2j 3 + j δ j 11α 11 + 7δ 7 + 6η 6, 17α 17 + 10δ 10 + 7η 7, 16δ 16 + 8η 8 16

We will state the last two steps of the classification, as they involve some things about codes that s not worth talking about. Recall that for a code C, we associate a lattice Γ C as 1 2 ρ 1 (C). Theorem 2.16. The correspondence C Γ C induces a one-to-one correspondence between equivalence classes of doubly even codes C of F n 2 and isomorphism classes of even lattices in R n containing a root lattice of type na 1. Self-dual codes correspond to unimodular lattices. Theorem 2.17. Up to isomorphism, there exist precisely 24 even unimodular lattices in R 24. Each lattice is uniquely determined by its root sublattice. The possible root sublattices are listed above. 17