SUMS OF SQUARES. Lecture at Vanier College, Sept. 23, Eyal Goren, McGill University.

Similar documents
The Quaternions. The Quaternions. John Huerta. Department of Mathematics UC Riverside. Cal State Stanislaus

Hypercomplex numbers

The Quaternions & Octonions: A Basic introduction to Their Algebras. By: Kyle McAllister. Boise State University

Quaternions and Octonions

sin(α + θ) = sin α cos θ + cos α sin θ cos(α + θ) = cos α cos θ sin α sin θ

Sums of four squares and Waring s Problem Brandon Lukas

A Brief History of Ring Theory. by Kristen Pollock Abstract Algebra II, Math 442 Loyola College, Spring 2005

Group, Rings, and Fields Rahul Pandharipande. I. Sets Let S be a set. The Cartesian product S S is the set of ordered pairs of elements of S,

ALGEBRA. 1. Some elementary number theory 1.1. Primes and divisibility. We denote the collection of integers

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

REVIEW Chapter 1 The Real Number System

Contents. D. R. Wilkins. Copyright c David R. Wilkins

Quaternion Algebras. Edgar Elliott. May 1, 2016

5. Vector Algebra and Spherical Trigonometry (continued)

John Huerta is finishing his Ph.D. in mathematics at the University of California, Riverside. His work tackles the foundations of supersymmetry.

Quaternions and the four-square theorem

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

QUADRATIC RINGS PETE L. CLARK

Complex Numbers. Rich Schwartz. September 25, 2014

SOLUTIONS FOR IMO 2005 PROBLEMS

Mathematics E-15 Seminar on Limits Suggested Lesson Topics

Solutions Exam 1 Math 221

Quaternions An Algebraic View (Supplement)

Course 2BA1: Hilary Term 2007 Section 8: Quaternions and Rotations

Math Circle Beginners Group February 28, 2016 Euclid and Prime Numbers Solutions

1 First Theme: Sums of Squares

Quaternions and Arithmetic. Colloquium, UCSD, October 27, 2005

Lattices. SUMS lecture, October 19, 2009; Winter School lecture, January 8, Eyal Z. Goren, McGill University

Winter Camp 2009 Number Theory Tips and Tricks

= 5 2 and = 13 2 and = (1) = 10 2 and = 15 2 and = 25 2

ORDERS OF ELEMENTS IN A GROUP

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.

A LITTLE BIT OF NUMBER THEORY

An integer p is prime if p > 1 and p has exactly two positive divisors, 1 and p.

Perspectives of Mathematics

Lecture 6: Finite Fields

Course MA2C02, Hilary Term 2010 Section 4: Vectors and Quaternions

Metacommutation of Hurwitz primes

Chapter 12: Ruler and compass constructions

UAB MATH TALENT SEARCH

Introduction. Question 1

irst we need to know that there are many ways to indicate multiplication; for example the product of 5 and 7 can be written in a variety of ways:

DOES XY 2 = Z 2 IMPLY X IS A SQUARE?

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.

GRE Quantitative Reasoning Practice Questions

Octonions? A non-associative geometric algebra. Benjamin Prather. October 19, Florida State University Department of Mathematics

1 Implication and induction

Sums of Squares (FNS 195-S) Fall 2014

A little bit of number theory

Beautiful Mathematics

SOME AMAZING PROPERTIES OF THE FUNCTION f(x) = x 2 * David M. Goldschmidt University of California, Berkeley U.S.A.

3.4 Complex Zeros and the Fundamental Theorem of Algebra

CSE 206A: Lattice Algorithms and Applications Spring Minkowski s theorem. Instructor: Daniele Micciancio

Quadratics. Shawn Godin. Cairine Wilson S.S Orleans, ON October 14, 2017

4 Hilbert s Basis Theorem and Gröbner basis

Algebra Revision Guide

Algebra. Here are a couple of warnings to my students who may be here to get a copy of what happened on a day that you missed.

Follow links for Class Use and other Permissions. For more information send to:

Bounds for the Representations of Integers by Positive Quadratic Forms

A Four Integers Theorem and a Five Integers Theorem

CONSTRUCTION OF THE REAL NUMBERS.

Math Circle Beginners Group February 28, 2016 Euclid and Prime Numbers

ON QUATERNIONS. William Rowan Hamilton. (Proceedings of the Royal Irish Academy, 3 (1847), pp ) Edited by David R. Wilkins

August 27, Review of Algebra & Logic. Charles Delman. The Language and Logic of Mathematics. The Real Number System. Relations and Functions

Chapter VI. Relations. Assumptions are the termites of relationships. Henry Winkler

EXERCISES IN QUATERNIONS. William Rowan Hamilton. (The Cambridge and Dublin Mathematical Journal, 4 (1849), pp ) Edited by David R.

Bott Periodicity and Clifford Algebras

Subalgebras of the Split Octonions

Discrete Mathematics and Probability Theory Fall 2018 Alistair Sinclair and Yun Song Note 6

2 Arithmetic. 2.1 Greatest common divisors. This chapter is about properties of the integers Z = {..., 2, 1, 0, 1, 2,...}.

Ramsey Theory over Number Fields, Finite Fields and Quaternions

Midterm 1 Review. Distance = (x 1 x 0 ) 2 + (y 1 y 0 ) 2.

ALGEBRA II: RINGS AND MODULES OVER LITTLE RINGS.

6 Cosets & Factor Groups

The Octonions. John C. Baez and John Huerta. Department of Mathematics University of California Riverside, CA USA.

Real Division Algebras

Linear Algebra for Beginners Open Doors to Great Careers. Richard Han

MA232A Euclidean and Non-Euclidean Geometry School of Mathematics, Trinity College Michaelmas Term 2017 Vector Algebra and Spherical Trigonometry

UNM - PNM STATEWIDE MATHEMATICS CONTEST XLVI - SOLUTIONS. February 1, 2014 Second Round Three Hours

EXAMPLES OF MORDELL S EQUATION

Complex numbers. Learning objectives

Algebra Exam. Solutions and Grading Guide

1. Prove that the number cannot be represented as a 2 +3b 2 for any integers a and b. (Hint: Consider the remainder mod 3).

Notice how these numbers thin out pretty quickly. Yet we can find plenty of triples of numbers (a, b, c) such that a+b = c.

DIHEDRAL GROUPS II KEITH CONRAD

Exponents and Logarithms

Abstract Algebra I. Randall R. Holmes Auburn University. Copyright c 2012 by Randall R. Holmes Last revision: November 11, 2016

Handout #6 INTRODUCTION TO ALGEBRAIC STRUCTURES: Prof. Moseley AN ALGEBRAIC FIELD

3 The language of proof

Sydney University Mathematical Society Problems Competition Solutions.

8 Primes and Modular Arithmetic

The Two Faces of Infinity Dr. Bob Gardner Great Ideas in Science (BIOL 3018)

MODEL ANSWERS TO THE FIRST HOMEWORK

MITOCW ocw f99-lec05_300k

CHAPTER 6. Prime Numbers. Definition and Fundamental Results

Modern Algebra Prof. Manindra Agrawal Department of Computer Science and Engineering Indian Institute of Technology, Kanpur

On The Non-Real Nature of x 0 (x R 0 ) The Set of Null Imaginary Numbers I Part I: Fundamentals

Definition 6.1 (p.277) A positive integer n is prime when n > 1 and the only positive divisors are 1 and n. Alternatively

Lecture 7: More Arithmetic and Fun With Primes

Discrete Math, Second Problem Set (June 24)

Transcription:

SUMS OF SQUARES Lecture at Vanier College, Sept. 23, 2011. Eyal Goren, McGill University. email: eyal.goren@mcgill.ca Two Squares. Which numbers are a sum of two squares? More precisely, which positive integers are the sum of squares of two integers? Here is a table: 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 Y Y N Y Y N N Y Y Y N N Y N N Y Y Y N Y N N N N Y Y N Here is some statistics Figure 1. Statistics The data raises many questions. For one, the plot strongly suggest an asymptotic. But what exactly is this asymptotic? what is the type of convergence? 1

2 SUMS OF SQUARES A theorem of Edmund Landau (1908) tells us that the proportion of numbers between 0 and 2 y K that are sum of two squares is assymptotically, where K 0.764 is a constant, called log(2) y the Landau-Ramanujan constant. 1 A lot is also known about the type of convergence. Another question is whether there is a pattern to which numbers are a sum of 2 squares? Consider for example the case of n a prime number. Can you tell whether it is a sum of 2 squares or not? Recall that an integer A is said to be congruent to a modulo n if A leaves a residue of a when divided by n. We write A a (mod n). For example: 13 1 (mod 4), because 13 = 3 4 + 1. There are the following facts: if A a (mod n), B b (mod n), then A + B a + b (mod n), A B a b (mod n), AB ab (mod n). We can in fact make the residues modulo n into a new system of numbers: to add or multiply two residues, add or multiply them as usual integers and then pass to the residue modulo n. For example, to multiply 3 (mod 5) with 4 (mod 5) we first multiply 3 and 4 to get 12, which gives the residue 2 (mod 5). Therefore, 3 4 = 2 (mod 5). In that way we get a system of numbers that satisfies all the usual identities (such as a(b + c) = ab + ac, and so on). The implication: is not always true 2, but is true if n is a prime. xy = 0 = x = 0 or y = 0, Theorem.(Pierre de Fermat, 1640)An integer n is a sum of two squares if and only if for every prime q congruent to 3 modulo 4, q divides n to an even power. Example. 2 3 5 7 2 13 = 25480 and so, according to the theorem, 25480 is a sum of two squares, Indeed, 25480 = 42 2 + 154 2. One can get some insight as to the meaning and proof of the theorem by reasoning as follows: If the theorem is true, then it has the following consequences: Any power of 2 is a sum of two squares. This is easy to check: 2 2a = (2 a ) 2. 2 2a+1 = (2 a ) 2 + (2 a ) 2. A prime congruent to 1 modulo 4 is a sum of two squares. An integer congruent to 3 modulo 4 is not a sum of two squares. The product of two numbers that are each a sum of 2 squares is a sum of two squares. Some of these are easy to prove: Note that (2x) 2 = 4x 2 is congruent to zero modulo 4 and (2x + 1) 2 = 4x 2 + 4x + 1 is congruent to 1 modulo 4. If n is a sum of two squares then n is therefore congruent to 0, 1 or 2 modulo 4. 1 In fact, Landau s result is more general. 2 For example 2 3 = 6 = 0 (mod 6), but neither 2 nor 3 are zero, even when viewed as residues (mod 6).

SUMS OF SQUARES 3 To explain the consequence that the sum of numbers each a sum of two squares is itself a sum of two squares, we recall the complex numbers. It is not necessary, but it is enlightening and it offers a perspective that can be used in many similar situations. To construct the complex numbers, a symbol i is added to the real numbers and it is decreed that i 2 = 1 The complex numbers are then numbers of the form z = x+yi, where x and y are real numbers. These numbers can be added and multiplied much like usual numbers. For example: (1 + i)(3 + 2i) = 1 3 + i 3 + 1 2i + i 2i = 3 + 3i + 2i 2 = 1 + 5i. We can depict the complex numbers in the (x, y) plane, and then define the absolute value of z as z = x 2 + y 2. It is precisely the length of the line connecting the origin to the point z, i.e., to the point (x, y). Multiplication, as one can check by direct calculation, satisfies (1) z 1 z 2 = z 1 z 2. This tells us the following: since (c + di)(e + fi) = ce df + (cf + de)i, after taking the square of the absolute values, we have by equation (1) the identity (ce df) 2 + (cf + de) 2 = (c 2 + d 2 )(e 2 + f 2 ). To make sure you understand the trick, express 5707026 as a sum of two squares (note that 9 = 3 2 is associated with complex number 3. So, for example, to express 90 as a sum of squares, we write it as 2 5 9. 2 is associated with 1 + i, 5 with 1 + 2i, 9 with 3 and so 90 is associated with (1 + i) (1 + 2i) 3 = 3 + 9i giving us that 72 = ( 3) 2 + 9 2 = 3 2 + 9 2.

4 SUMS OF SQUARES It took quite a long time for mathematicians to accept the existence of a number whose square is 1. We tend to think about numbers as depicting a reality one plus one is two, etc. and, as such, we accept the real numbers, numbers of the form 13.2768, and even 2.333333... as, well..., real! From this perspective, it was difficult to accept that there is a new number, called i, that satisfies i 2 = 1. And, at first, mathematicians found it more comfortable to think of i as an operator; namely, a symbol that denotes a concrete operation. The operation here is that i is the transformation of the plane taking (x, y) to ( y, x); that is, rotation counterclockwise by 90. This interpretation follows directly from the equation i(x + yi) = y + xi. Then, i 2 sends (x, y) to ( x, y) and so is like multiplication by 1. We can therefore interpret i 2 as 1. Based on the identities above, to prove the if part of the theorem, it is therefore enough to prove that every prime congruent to 1 modulo 4 is a sum of two squares. To prove that such a prime is a sum of squares, we will need the following fact: Lemma. There is an integer 0 < u < p such that u 2 leaves residue 1 modulo p. For example, 2 2 = 4 = 1 (mod 5), 5 2 = 25 = 1 (mod 13), 12 2 = 144 = 1 (mod 29). We assume the lemma. We mention that it does not hold for primes congruent to 3 modulo 4. We now introduce a new idea: a lattice L in the plane is a collection of points such that the sum, or difference, of two points in L is again in L. For example the set of points (x, y) such that Figure 2. The lattice of integers and some fundamental parallelograms, and the hexagonal lattice x and y are integers is a lattice. A lattice has a fundamental parallelogram, a parallelogram based at zero, whose vertices are points of L, that contains no point of L in its interior.

SUMS OF SQUARES 5 Minkowski s theorem. If the fundamental parallelogram of L has area V then a disc of radius r such that πr 2 4V must contain a point of L besides (0, 0). Application. We consider the lattice L of points (x, y) such that y ux is divisible by p. The parallelogram with vertices (0, 0), (0, p), (1, u), (1, p + u) is a fundamental parallelogram. Given a point (x, y) of the lattice, we can write it as x(1, u) + y xu p (0, p). This shows that such a point cannot be inside the fundamental parallelogram. The area of this parallelogram is p. Let us take a disc of radius r = 4π 1 p, a radius that satisfies Minkowski s condition on the nose. There is a lattice point (x, y) in this disc, that is, a point such that x 2 + y 2 < 4π 1 p < 2p and y ux is divisible by p. Now, y 2 is congruent to u 2 x 2 = x 2 modulo p and so, Since x 2 + y 2 < 2p it follows that p (x 2 + y 2 ). p = x 2 + y 2. Proof of Minkowski s theorem. Call the disc D and consider d = 1 2D - the disc shrunk by a factor of 2. Suppose that d is disjoint from l + d for all l L. Then, we can translate the various parts of d lying in different parallelograms so that they all lie in the fundamental one, and they are all disjoint. Therefore, the area of d, which is 1 4 πr2 is less than that of L, which is V. That is, 1 4 πr2 < V. This is a contradiction, and so d is not disjoint from l + d for some l 0. That means, that for some s, t D we have 1 2 s = l + 1 2t. Thus, l = 1 (s t). 2 It follows that the distance of l from the origin is at most 1 4 (r + r) < r, and so l lies in D. Four Squares. Here we would like to discuss the following theorem. Theorem. Every positive integer is a sum of 4 squares. The strategy of showing that if x is a sum of 2 squares and y is a sum of 2 squares than so is xy was very useful. Can we prove such a statement for 4 squares? To prove that, we introduce a generalization of complex numbers called Quaternion numbers. These are numbers of the shape a + bi + cj + dk where a, b, c, d are real numbers and i, j, k are new symbols that satisfy i 2 = j 2 = k 2 = 1, ij = k = ji.

6 SUMS OF SQUARES Using these identities one can add and multiply quaternions in a formal way. identities hold, for example x(yz) = (xy)z, but some identities do not hold. xy yx in general. Indeed, ij = ji. Still, we can define Many familiar For example, and we have, by laborious calculation, z = a 2 + b 2 + c 2 + d 2, z 1 z 2 = z 1 z 2. It now follows that if x is a sum of 4 squares and y is a sum of 4 squares then so is xy. And, in fact, the presentation of xy as a sum of 4 squares can be calculated from the presentation of x and of y, similar to the way we had done it for sum of two squares. Therefore, to show that every integer is a sum of 4 squares, it is enough to show that for primes congruent to 3 modulo 4. The proof of this result is again an application of Minkowski s theorem, but now using lattices in 4-dimensional space. The concept of a lattice generalized easily and so is the concept of a ball and its volume. Minkowski considered lattices in a space of arbitrary dimension n and proved the following general result, whose proof is a generalization of the proof given for n = 2. Minkowski s theorem 1889. Let L be a lattice in n-dimensional space. If the fundamental parallelogram of L has volume V then a ball whose volume is at least 2 n V must contain a point of L besides (0, 0). The choice of lattice in this case is much more tricky and we don t discuss this further. Three Squares. Also this problem admits a precise solution. Legendre s theorem. 4 n (8m + 7). A positive integer is a sum of three squares, unless it is of the form Note that 3 = 1 2 + 1 2 + 1 2 is a sum of 3 squares, as is 5 = 0 2 + 1 2 + 2 2, but 3 5 = 15 = 8 1 + 7 is not a sum of 3 squares (as one can verify directly). Thus, the paradigm of our previous proofs fails. There is no 3-dimensional generalization of the integers, similar to the two dimensional generalization a + bi (a, b integers) - the Gaussian integers, or the four dimensional generalization a + bi + cj + dk (a, b, c, d integers) - the Hurwitz integers. In fact, for many years, Sir William Rowan Hamilton tried to find generalizations of the complex numbers to three dimensional situation, namely, how to add and multiply triples (a + bi + cj). Addition is not a problem, but multiplication yielded a situation where many identities we expect just fail. One novelty of Hamilton s approach was the treatment of the expressions a + bi + cj as

SUMS OF SQUARES 7 generalized numbers and not as operators. Indeed, from the point of view of operators it is hard to see why there shouldn t be such a multiplication rule. The problem was not trivial by any means, certainly not at the time. Hamilton s investigations came on the heels of breakthroughs achieved by the use of complex numbers (for example, Gauss s proof of the fundamental theorem of algebra in 1799) at a time when many mathematicians still had problems accepting the complex numbers as numbers, namely, as purely algebraic quantities. In fact, Hamilton was the first to give a satisfactory definition of complex numbers. What Hamilton was doing was cutting-edge research. Hamilton was a prodigy. Born in Dublin at 1805 he could translate Latin, Greek and Hebrew at the age of 5. At twevlth he knew at least 9 languages and engaged in contests with the American calculating boy Zerah Colburn (when Colburn was seven years old he took six seconds to give the number of hours in thirty-eight years, two months, and seven days). The competition was one of the events turning Hamilton s interest to mathematics. He entered Trinity college in Dublin at 1823, but was appointed as a professor while still an undergraduate, at 1827, at the age of 22. Thus, we can appreciate the difficulty of the problem of generalizing complex numbers in the light of the efforts of such an intellect; he had devoted 15 years to the problem. The story goes that Hamilton s little sons, aged eight and nine, had, during the climatic month of October 1843, greeted him at breakfast with Well, Papa, can you multiply triplets. Hamilton later wrote to his son: Every morning in the early part of the above-cited month, on my coming down to breakfast, your (then) little brother William Edwin, and yourself, used to ask me: Well, Papa, can you multiply triplets? Whereto I was always obliged to reply, with a sad shake of the head: No, I can only add and subtract them. The inspiration that one can multiply quadruples (a + bi + cj + dk) in a reasonable way came to Hamilton in a flash on October 16, 1843, while walking with his wife along the Royal Canal in Dublin, on their way to a concert. He describes it as an electric circuit seemed to close and a spark flashed forth. He documented the insight by engraving the identities i 2 = j 2 = k 2 = ijk = 1 on a stone of the Brougham Bridge - performing also the first and most famous act of mathematical graffiti. The bridge has a plaque commemorating the event (Figure 3). In fact, there are deep reasons as to why one cannot multiply triples. Suppose there was such a multiplication of triples (x, y, z) real numbers that preserved length and extended the multiplication of usual real numbers (x, 0, 0). There is thus a multiplication rule for the sphere x 2 + y 2 + z 2 = 1 in three dimensional space. Take a tangent vector at the point (1, 0, 0) and use the multiplication to transport it to a tangent vector at a point (x, y, z) ((1, 0, 0) stands for the number one and so we want (1, 0, 0) (x, y, z) = (x, y, z) for any (x, y, z), whatever the multiplication rule may be). This way, we get a field of tangent vectors consisting of unit vectors, and in particular, a field for which no vector is zero. We were able to comb the ball. But there is a theorem in topology

8 SUMS OF SQUARES Figure 3. A plate commemorating Hamilton s discovery of the quaternions. The hairy ball theorem. One cannot comb the ball. The proof of this theorem uses totally different ideas from those we were pursuing so far and we shall not discuss the proof. It has the following consequence: Application. At any moment, somewhere on earth, the air stands perfectly still. Indeed, we may think about the direction of the wind as a field of tangent vectors; they cannot be all non-zero due to the hairy ball theorem. Very Recent Developments. Other number systems. Much as the complex numbers are a system of numbers in which we can do arithmetic, there are many others system of numbers (they are called number fields) in which we can do arithmetic. For example, we may consider the system of numbers of the form a + b 2, a, b Z. The sum and product of two such (real) numbers is a (real) number of the same sort and so we may ask, which numbers are sums of two, three, four, etc., squares in this system?

SUMS OF SQUARES 9 Legendre s theorem doesn t generalize easily. about 10 years ago, Cogdell, Sarnak and Piatetski - Shapiro gave an analogue that holds for every large enough number. Conway-Schneeberger Theorem. The generalization of expressions of the form x 2 + y 2, x 2 + y 2 + z 2 are quadratic forms. They are expressions of the form ij a ijx i x j where the a ij are constants and the x i are variables. For instance, x 2 + xy + 3y 3, 2xz + y 2 5yz, etc. Let n be an integer. If there is an assignment of integers values for the variables x i such that ij a ijx i x j = n, we say that the quadratic form represents n. For example, x 2 + y 2 represents 5, but does not represent 3. A natural question is when does a quadratic form represent every integer? We have the following striking theorem: Theorem. (Bhargava, Conway-Schneeberger) If a quadratic form ij a ijx i x j, where a ij are integers and are even integers if i j, represents the integers 1, 2,..., 15 then it represents every integer.