Lecture 2: Mutually Orthogonal Latin Squares and Finite Fields

Similar documents
Lecture 3: Latin Squares and Groups

Lecture 6: Finite Fields

Lecture 5: Latin Squares and Magic

Lecture 8: A Crash Course in Linear Algebra

Lecture 4: Constructing the Integers, Rationals and Reals

Lecture 2: Vector Spaces, Metric Spaces

Lecture 1: Latin Squares!

A Latin Square of order n is an n n array of n symbols where each symbol occurs once in each row and column. For example,

AN ALGEBRA PRIMER WITH A VIEW TOWARD CURVES OVER FINITE FIELDS

Rings If R is a commutative ring, a zero divisor is a nonzero element x such that xy = 0 for some nonzero element y R.

Outline. MSRI-UP 2009 Coding Theory Seminar, Week 2. The definition. Link to polynomials

1. Introduction to commutative rings and fields

LECTURE 10, MONDAY MARCH 15, 2004

WORKSHEET ON NUMBERS, MATH 215 FALL. We start our study of numbers with the integers: N = {1, 2, 3,...}

MATH CSE20 Homework 5 Due Monday November 4

Construction of latin squares of prime order

NOTES ON FINITE FIELDS

Factoring. there exists some 1 i < j l such that x i x j (mod p). (1) p gcd(x i x j, n).

Lecture 9: Taylor Series

1 What is the area model for multiplication?

Section 0.6: Factoring from Precalculus Prerequisites a.k.a. Chapter 0 by Carl Stitz, PhD, and Jeff Zeager, PhD, is available under a Creative

Recall, R is an integral domain provided: R is a commutative ring If ab = 0 in R, then either a = 0 or b = 0.

Lecture 14: The Rank-Nullity Theorem

Lecture 1: Period Three Implies Chaos

1. Introduction to commutative rings and fields

Q 2.0.2: If it s 5:30pm now, what time will it be in 4753 hours? Q 2.0.3: Today is Wednesday. What day of the week will it be in one year from today?

Figure 9.1: A Latin square of order 4, used to construct four types of design

MATH 115, SUMMER 2012 LECTURE 4 THURSDAY, JUNE 21ST

Math 4310 Solutions to homework 1 Due 9/1/16

Math Lecture 18 Notes

Finite arithmetic and Latin squares

Lecture 19: The Determinant

Proof by Contradiction

P-adic numbers. Rich Schwartz. October 24, 2014

Lecture 2: Period Three Implies Chaos

One-to-one functions and onto functions

Lecture 2: Eigenvalues and their Uses

Lecture 3: Sizes of Infinity

Homework 8 Solutions to Selected Problems

Minilecture 2: Sorting Algorithms

MA554 Assessment 1 Cosets and Lagrange s theorem

Math Lecture 3 Notes

8 Primes and Modular Arithmetic

Lecture 5. 1 Review (Pairwise Independence and Derandomization)

Commutative Rings and Fields

The Chinese Remainder Theorem

0 Sets and Induction. Sets

ALGEBRA II: RINGS AND MODULES OVER LITTLE RINGS.

Math 31 Lesson Plan. Day 5: Intro to Groups. Elizabeth Gillaspy. September 28, 2011

6.080 / Great Ideas in Theoretical Computer Science Spring 2008

Cosets and Lagrange s theorem

48 CHAPTER 2. COMPUTATIONAL METHODS

Math /Foundations of Algebra/Fall 2017 Foundations of the Foundations: Proofs

WORKSHEET MATH 215, FALL 15, WHYTE. We begin our course with the natural numbers:

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

In case (1) 1 = 0. Then using and from the previous lecture,

A field F is a set of numbers that includes the two numbers 0 and 1 and satisfies the properties:

Computations/Applications

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.

MATH 25 CLASS 21 NOTES, NOV Contents. 2. Subgroups 2 3. Isomorphisms 4

Solutions to odd-numbered exercises Peter J. Cameron, Introduction to Algebra, Chapter 2

Basic counting techniques. Periklis A. Papakonstantinou Rutgers Business School

The group (Z/nZ) February 17, In these notes we figure out the structure of the unit group (Z/nZ) where n > 1 is an integer.

Reading 11 : Relations and Functions

MATH 115, SUMMER 2012 LECTURE 12

Linear Algebra, Summer 2011, pt. 3

LECTURE 5, FRIDAY

Boolean circuits. Lecture Definitions

CHAPTER 3: THE INTEGERS Z

A Harvard Sampler. Evan Chen. February 23, I crashed a few math classes at Harvard on February 21, Here are notes from the classes.

MA257: INTRODUCTION TO NUMBER THEORY LECTURE NOTES

ORTHOGONAL SETS OF LATIN SQUARES AND. CLASS-r HYPERCUBES GENERATED BY FINITE ALGEBRAIC SYSTEMS. A Dissertation in. Mathematics. Daniel R.

Introduction to Algebra

Solutions Parabola Volume 49, Issue 2 (2013)

Math 581 Problem Set 9

x + 2y + 3z = 8 x + 3y = 7 x + 2z = 3

Rings. EE 387, Notes 7, Handout #10

CHMC: Finite Fields 9/23/17

Diary of Mathematical Musings. Patrick Stein

Chapter 4 Mathematics of Cryptography

Mutually Orthogonal Latin Squares: Covering and Packing Analogues

Recitation 9: Probability Matrices and Real Symmetric Matrices. 3 Probability Matrices: Definitions and Examples

Answers in blue. If you have questions or spot an error, let me know. 1. Find all matrices that commute with A =. 4 3

Introduction to finite fields

Error Correcting Codes Prof. Dr. P Vijay Kumar Department of Electrical Communication Engineering Indian Institute of Science, Bangalore

Math 3361-Modern Algebra Lecture 08 9/26/ Cardinality

Lesson 1 WHAT S A GROUP?

Elliptic Curves and Public Key Cryptography

Number Axioms. P. Danziger. A Group is a set S together with a binary operation (*) on S, denoted a b such that for all a, b. a b S.

Introduction to Ring Theory

USA Mathematical Talent Search Round 1 Solutions Year 27 Academic Year

Commutative Algebra MAS439 Lecture 3: Subrings

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

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

Partial Fractions. June 27, In this section, we will learn to integrate another class of functions: the rational functions.

Math1a Set 1 Solutions

Proving languages to be nonregular

Abstract Algebra. Joseph R. Mileti. January 22, 2012

NOTES ON OCT 4, 2012

* 8 Groups, with Appendix containing Rings and Fields.

Transcription:

Latin Squares Instructor: Padraic Bartlett Lecture 2: Mutually Orthogonal Latin Squares and Finite Fields Week 2 Mathcamp 2012 Before we start this lecture, try solving the following problem: Question Take a deck of playing cards, and remove the 16 aces, kings, queens, and jacks from the deck Can you arrange these cards into a 4 4 array, so that in each column and row, no two cards share the same suit or same face value? This question should feel similar to the problem of constructing a Latin square: we have an array, and we want to fill it with symbols that are not repeated in any row or column However, we have the additional constraint that we re actually putting two symbols in every cell: one corresponding to a suit, and another corresponding to a face value So: if we just look at the face values, we have a 4 4 Latin square Similarly, if we ignore the face values and look only at the suits, we should have a different 4 4 Latin square; as well, these two Latin squares have the property that when we superimpose them (ie place one on top of the other), each of the resulting possible 16 pairs of symbols occurs exactly once (because we started with 16 distinct cards) The generalization of this idea gives us an idea of orthogonality for Latin squares, which we define here: Definition A pair of n n Latin squares are called orthogonal if when we superimpose them (ie place one on top of the other), each of the possible n 2 ordered pairs of symbols occur exactly once A collection of k n n Latin squares is called mutually orthogonal if every pair of Latin squares in our collection is orthogonal Example The grid of playing cards you constructed earlier if you answered our first question is a pair of 4 4 squares, for the reasons we discussed earlier To further illustrate the idea, we present a pair of orthogonal 3 3 Latin squares: 1 2 3 1 2 3 (1, 1) (2, 2) (3, 3) 2 3 1, 3 1 2 (2, 3) (3, 1) (1, 2) 3 1 2 2 3 1 (3, 2) (1, 3) (2, 1) Like always, whenever we introduce a mathematical concept in combinatorics, our first instinct should be to attempt to count it! In other words: given an order n, what is the largest collection of mutually orthogonal Latin squares we can find? An upper bound is not too hard to find: Proposition For any n, the maximum size of a set of n n mutually orthogonal Latin squares is n 1 1

Proof Take any collection T 1, T k of mutually orthogonal Latin squares Then notice the following property: if we take any of our Latin squares and permute its symbols (ie switch all the 1 and 2 s), the new square is still mutually orthogonal to all of the other squares (Think about this for a bit if you are unpersuaded) Using the above observation, notice that we can without any loss of generality assume that the first row of each of our Latin squares is (1, 2, 3 n) Now, take any pair of mutually orthogonal Latin squares from our collection, and look at the symbol in the cell in the first column/second row (ie the symbol at (2, 1)): 1 2 n 1 2 n x y, We know that neither x nor y can be 1, because both of these squares are Latin squares As well, we know that they cannot agree, as the first row of the superimposition of these two squares contains the pairs (k, k), for every 1 k n This means that there are at most n 1 squares in our collection T 1, T k, because there are n 1 distinct choices for the cell (2, 1) that are not 1 We already know that sometimes n 1 is attainable: in our example above, we found 2 orthogonal Latin squares of order 3 When can we attain this bound? Perhaps surprisingly, the answer to this question is intimately related to the concept of finite fields! We define these here: Definition Roughly speaking, a field is a set F along with a pair of operations +, that act on our field, that satisfy the same properties that R does with respect to + and Formally, these properties are the following: Closure(+): a, b F, a + b F Identity(+): 0 F such that a F, 0 + a = a Commutativity(+): a, b F, a + b = b + a Associativity(+): a, b, c F, (a + b) + c = a + (b + c) Inverses(+): a F,!( a) F such that a + ( a) = 0 Closure( ): a, b F, a b F Identity( ): 1 F such that a F, 1 a = a Commutativity( ): a, b F, a b = b a Associativity( ): a, b, c F, (a b) c = a (b c) Inverses( ): a 0 F,!(a 1 ) F such that a (a 1 ) = 1 Distributivity(+, ) : a, b, c F, (a + b) c = (a c) + (b c) A finite field is simply a field that contains finitely many elements 2

Example You are invited to prove / check / simply believe that Z/pZ, the integers modulo a prime p, is always a field with respect to the operations +, (If you haven t seen this before: Z/pZ is simply the set {0, 1, 2 p 1} We define addition mod p, as follows: we say that a + b c mod p if the two integers a + b and c differ by a multiple of p Similarly, we define multiplication mod p by saying that a b c mod p if the two integers a b and c differ by a multiple of p These definitions give us a way to define addition and multiplication on our set: you can check that it satisfies all of the axioms listed earlier, if you wish) The reason we mention finite fields is that (like so very many things we will see in this class) we can turn them into Latin squares! Specifically, we have the following theorem Proposition Let F be a finite field that contains n elements Then there is a collection of n 1 mutually orthogonal Latin squares Proof For simplicity s sake, enumerate the elements F as {f 0, f 1 f n 1 }, such that f 0 = 0 and f 1 = 1 Now, notice the following fact: if a F is nonzero, then the grid af 0 + f 0 af 1 + f 0 af n 1 + f 0 af 0 + f 1 af 1 + f 1 af n 1 + f 1, af 0 + f n 1 af 1 + f n 1 af n 1 + f n 1 where we fill the cell (i, j) with af i + f j, is in fact a Latin square! To see why, suppose that there is some row i along which two cells (i, j) and (i, k) of this grid are the same: ie that af i + f j = af i + f k a(f i f i ) = (f k f j ) 0 = (f k f j ) f j = f k, and therefore that j = k and that these two cells are the same Similarly, if we pick any column j along which two cells (i, j) and (k, j) of this grid are the same, we get af i + f j = af k + f j a(f i f k ) = (f j f j ) a(f i f k ) = 0 f i f k = 0 f j = f k, and can again conclude that these two cells were the same This generates n 1 distinct Latin squares: label them T a, for every element a F We claim that this is fact a set of mutually orthogonal Latin squares! To see why, take any two 3

squares T a, T b, and suppose that there are two cells (i, j), (k, l) at which superimposing our two Latin squares yields the same ordered pair of symbols: ie that af i + f j = af k + f l and bf i + f j = bf k + f l Taking the difference of these two equations yields (a b)f i = (a b)f k f i = f k ; plugging this into our earlier equations yields f j = f l, and therefore that these two cells are the same Therefore, this is a set of n 1 mutually orthogonal Latin squares! For those of you taking Ruthi s class on finite fields, something you ll see there is the following theorem: Theorem 1 There is a finite field of order n if and only if n can be expressed as a prime power: ie that there is some prime p and natural number k such that n = p k We are not going to go into this theorem, as it s pretty far afield from the subject of our class However, it s worth taking a second to talk about how you can actually find these finite fields! We present a construction here: Construction Given a finite field F, we can form the ring of polynomials over F, F [x], by simply taking all of the polynomials of the form a 0 + a 1 x + a 2 x 2 + x n, where the elements a i are all elements in our field F (We multiply and add these polynomials as we would normally: ie (a + bx)(c + dx) = ac + (bc + ad) x + bd x 2, where we use our field to figure out how the multiplication and addition of these elements actually works A polynomial in F [x] is called irreducible if there is no way to write it as the product of two polynomials with smaller degrees For example, if F = F 2 = Z/2Z, the element x 2 + 1 of F [x] is not irreducible, because (x + 1) (x + 1) = x 2 + 2x + 1 x 2 + 1 mod 2 However, the polynomial x 2 + x + 1 is irreducible in F [x], which you can check by taking all of the polynomials of smaller degrees, taking their products, and checking that you never get x 2 + x + 1 Suppose that you take a finite field F, create the ring of polynomials F [x], and then find an irreducible polynomial g(x) in F [x] that s irreducible of degree n By multiplying by an appropriate constant, make it so that the coefficient of x n in g(x) is 1 Now, take F [x], and regard two polynomials as being the same if they differ by a multiple of g(x) (This is similar to the way we defined Z/nZ by saying that two things are the same if they differ by a multiple of n) Call this object F [x]/ g(x) This, rather surprisingly, is a finite field with F n -many elements 4

Example To illustrate this, take F = Z/2Z, and g(x) = x 2 + x + 1 The elements of F [x] look like 0, 1, x, x + 1, x 2, x 2 + 1, x 2 + x + 1, x 3, Suppose that we regard two elements in F [x] to be the same up to a multiple of g(x) Then, if we take any element h(x) of F [x] and repeatedly subtract appropriately chosen multiples of x k g(x) fromh(x), we can eventually insure that h(x) is a polynomial of degree at most 1 You can check that none of the remaining four elements of F [x] can be turned into each other via adding/subtracting multiples of x 2 + x + 1; therefore, we have F [x]/ g(x) = {0, 1, x, x + 1} By examining the addition and multiplication tables below, we can easily see that this forms a field: + 0 1 x x + 1 0 0 1 x x + 1 1 1 0 x + 1 x x x x + 1 0 1 x + 1 x + 1 x 1 0, 0 1 x x + 1 0 0 0 0 0 1 0 1 x x + 1 x 0 x x + 1 1 x + 1 0 x + 1 1 x, (To check the multiplication table above, in particular, we used the observation that x 2 x 2 + (x 2 + x + 1) = 2x 2 + x + 1 x + 1 mod 2, x 2 + x + 1, and similarly that x(x + 1) 1, (x + 1) 2 x mod 2, x 2 + x + 1 ) A beautiful consequence of this example is that it gives us a way to create three mutually orthogonal Latin squares of order 4, using our construction from earlier: T 1 = 0 1 x x + 1 1 0 x + 1 x x x + 1 0 1 x + 1 x 1 0, T x = 0 1 x x + 1 x x + 1 0 1 x + 1 x 1 0 1 0 x + 1 x 0 1 x x + 1 T x+1 = x + 1 x 1 0 1 0 x + 1 x x x + 1 0 1 These are all orthogonal! So: with this lecture, we ve shown (up to a nontrivial result on finite fields!) that whenever n is a prime power, we have sets of mutually orthogonal Latin squares that are as large as we could hope for (ie n 1) Given this, you might hope that we can always find sets of n 1 mutually orthogonal Latin squares, for any order n This turns out to be tragically false: 5

Theorem There are no pair of mutually orthogonal Latin squares of order 6 As far as I know, there are no known proofs of this result that boil down to anything much more elegant to just brute-force-checking all of the pairs (which, given that there are 6! 5! 9, 408 distinct Latin squares, is not pleasant Relatedly, this makes the fact that this was proved in 1899-1900 by Tarry, long before the advent of computers, even more impressive) In general, we don t know very much about this question: the maximum number of mutually orthogonal latin squares of order 10, I believe, is still not known, as is the value for almost all non-prime-powers 6