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

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

8 Primes and Modular Arithmetic

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

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

MATH 215 Final. M4. For all a, b in Z, a b = b a.

Direct Proof MAT231. Fall Transition to Higher Mathematics. MAT231 (Transition to Higher Math) Direct Proof Fall / 24

Chapter 3 Basic Number Theory

Chapter 5. Number Theory. 5.1 Base b representations

a = qb + r where 0 r < b. Proof. We first prove this result under the additional assumption that b > 0 is a natural number. Let

2x 1 7. A linear congruence in modular arithmetic is an equation of the form. Why is the solution a set of integers rather than a unique integer?

#26: Number Theory, Part I: Divisibility

Senior Math Circles Cryptography and Number Theory Week 2

1. multiplication is commutative and associative;

Math 110 FOUNDATIONS OF THE REAL NUMBER SYSTEM FOR ELEMENTARY AND MIDDLE SCHOOL TEACHERS

1 Overview and revision

Ch 4.2 Divisibility Properties

Lecture 2. The Euclidean Algorithm and Numbers in Other Bases

MATH 2112/CSCI 2112, Discrete Structures I Winter 2007 Toby Kenney Homework Sheet 5 Hints & Model Solutions

PUTNAM TRAINING NUMBER THEORY. Exercises 1. Show that the sum of two consecutive primes is never twice a prime.

2.2 Some Consequences of the Completeness Axiom

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).

Chapter 2. Divisibility. 2.1 Common Divisors

MATH 4400 SOLUTIONS TO SOME EXERCISES. 1. Chapter 1

Math 109 HW 9 Solutions

COMP239: Mathematics for Computer Science II. Prof. Chadi Assi EV7.635

Homework #2 solutions Due: June 15, 2012

CISC-102 Fall 2017 Week 6

2x 1 7. A linear congruence in modular arithmetic is an equation of the form. Why is the solution a set of integers rather than a unique integer?

The theory of numbers

12x + 18y = 50. 2x + v = 12. (x, v) = (6 + k, 2k), k Z.

What is proof? Lesson 1

Wednesday, February 21. Today we will begin Course Notes Chapter 5 (Number Theory).

NUMBER SYSTEMS. Number theory is the study of the integers. We denote the set of integers by Z:

This is a recursive algorithm. The procedure is guaranteed to terminate, since the second argument decreases each time.

Mathematical Reasoning & Proofs

The following is an informal description of Euclid s algorithm for finding the greatest common divisor of a pair of numbers:

Chapter III. Basic Proof Techniques. Proving the obvious has never been easy. Marty Rubin

Chapter One. The Real Number System

All variables a, b, n, etc are integers unless otherwise stated. Each part of a problem is worth 5 points.

Number Theory Math 420 Silverman Exam #1 February 27, 2018

Introduction to Abstract Mathematics

Contribution of Problems

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

CHAPTER 3: THE INTEGERS Z

Beautiful Mathematics

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

11 Division Mod n, Linear Integer Equations, Random Numbers, The Fundamental Theorem of Arithmetic

INTEGERS. In this section we aim to show the following: Goal. Every natural number can be written uniquely as a product of primes.

Midterm Preparation Problems

Example: This theorem is the easiest way to test an ideal (or an element) is prime. Z[x] (x)

REVIEW Chapter 1 The Real Number System

10 Problem 1. The following assertions may be true or false, depending on the choice of the integers a, b 0. a "

Modular Arithmetic Instructor: Marizza Bailey Name:

Math 110 HW 3 solutions

Proof 1: Using only ch. 6 results. Since gcd(a, b) = 1, we have

Wilson s Theorem and Fermat s Little Theorem

Part V. Chapter 19. Congruence of integers

Commutative Rings and Fields

Chapter 5: The Integers

Solution Set 2. Problem 1. [a] + [b] = [a + b] = [b + a] = [b] + [a] ([a] + [b]) + [c] = [a + b] + [c] = [a + b + c] = [a] + [b + c] = [a] + ([b + c])

Elementary Properties of the Integers

The Integers. Math 3040: Spring Contents 1. The Basic Construction 1 2. Adding integers 4 3. Ordering integers Multiplying integers 12

Introductory Mathematics

Math 131 notes. Jason Riedy. 6 October, Linear Diophantine equations : Likely delayed 6

Fall 2017 Test II review problems

Proof Techniques (Review of Math 271)

2. Prime and Maximal Ideals

Module 1. Integers, Induction, and Recurrences

MATH 433 Applied Algebra Lecture 4: Modular arithmetic (continued). Linear congruences.

4.4 Solving Congruences using Inverses

CS 360, Winter Morphology of Proof: An introduction to rigorous proof techniques

MATH10040: Chapter 0 Mathematics, Logic and Reasoning

PRACTICE PROBLEMS: SET 1

NOTES ON SIMPLE NUMBER THEORY

EUCLID S ALGORITHM AND THE FUNDAMENTAL THEOREM OF ARITHMETIC after N. Vasiliev and V. Gutenmacher (Kvant, 1972)

Divisibility = 16, = 9, = 2, = 5. (Negative!)

Proof by Contradiction

CHAPTER 6. Prime Numbers. Definition and Fundamental Results

Try the assignment f(1) = 2; f(2) = 1; f(3) = 4; f(4) = 3;.

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

3 The language of proof

Cool Results on Primes

Chapter 1 A Survey of Divisibility 14

PGSS Discrete Math Solutions to Problem Set #4. Note: signifies the end of a problem, and signifies the end of a proof.

Notes on Systems of Linear Congruences

MATH 501 Discrete Mathematics. Lecture 6: Number theory. German University Cairo, Department of Media Engineering and Technology.

Winter Camp 2009 Number Theory Tips and Tricks

2 Elementary number theory

Mathematics-I Prof. S.K. Ray Department of Mathematics and Statistics Indian Institute of Technology, Kanpur. Lecture 1 Real Numbers

CHAPTER 4: EXPLORING Z

MI 4 Mathematical Induction Name. Mathematical Induction

The set of integers will be denoted by Z = {, -3, -2, -1, 0, 1, 2, 3, 4, }

2. Two binary operations (addition, denoted + and multiplication, denoted

Math 13, Spring 2013, Lecture B: Midterm

a + b = b + a and a b = b a. (a + b) + c = a + (b + c) and (a b) c = a (b c). a (b + c) = a b + a c and (a + b) c = a c + b c.

MATH 361: NUMBER THEORY FOURTH LECTURE

Properties of the Integers

Definition For a set F, a polynomial over F with variable x is of the form

MTH 310, Section 001 Abstract Algebra I and Number Theory. Sample Midterm 1

Chapter 7. Number Theory. 7.1 Prime Numbers

Transcription:

WORKSHEET ON NUMBERS, MATH 215 FALL 18(WHYTE) We start our study of numbers with the integers: Z = {..., 2, 1, 0, 1, 2, 3,... } and their subset of natural numbers: N = {1, 2, 3,...} For now we will not worry about defining or justifying the most basic properties : we will assume that we know what the integers and natural numbers are, that we know how to add and multiply them, and that rules like commutativity, associativity, and distributivity all hold. The first thing we wish to study is divisibility: Definition 0.1. Let a and b be two integers. We say that a divides b if there is an integer q so that b = aq. We will use the notation a b as shorthand for this property. Proposition 0.2. Let a, b, and c be integers. If a b and a c then a (b + c). Proposition 0.3. Let a, b, and c be integers. If a b then a bc. Question 0.4. Which integers divide zero? Which are divisible by zero? Proposition 0.5. Let a, b, and c be integers. If a b and a c then a (b c). Question 0.6. Which integers divide one? Which are divisible by one? Proposition 0.7. Let a, b, c, s and t be integers. a (sb + tc). If a b and a c then Proposition 0.8. Let a, b, and c be integers. If a b and b c then a c. Question 0.9. Does 2 4? Does 2 3? How can you prove your answers? 1

2 WORKSHEET ON NUMBERS, MATH 215 FALL 18(WHYTE) As the last question indicates, we are missing some of the basic structure of the integers. The next piece that we want to incorporate is that we can compare integers to see which is bigger. This is what mathematicians call an ordering. To be more precise: Definition 0.10. An integer n is positive if and only if n N Definition 0.11. For any two integers a and b, we say a < b if and only if b a is positive. We can now prove many of the basic facts we know about comparing integers Problem 0.12. Prove the following (you may use that the sum and product of natural numbers are natural numbers) (1) If a < b then for any c we have a + c < b + c (2) If a < b and b < c then a < c (3) For any a and b exactly one of the following holds: a < b, b < a, or a = b (4) If a is positive and b is negative then ab is negative (5) If a is negative and b is negative then ab is positive (6) Let a and b be integers with a < b, then for all c > 0 we have ac < bc and for all c < 0 we have bc < ac We can also now establish some facts that get used quite often: Proposition 0.13. Let a and b be integers with ab = 0 then a = 0 or b = 0 Proposition 0.14. Let a, b, and c be integers with c = 0 then ac = bc = a = b This almost resolves the question at the end of the last page: Proposition 0.15. If q is an integer with 2q = 3 then 1 < q < 2. To show that there aren t any such q we need one more fact about the integers: Axiom 0.16. There are no integers n with 0 < n < 1. Using this you can now show: Proposition 0.17. 2 does not divide 3

WORKSHEET ON NUMBERS, MATH 215 FALL 18(WHYTE) 3 We now have an almost complete list of axioms for the integers and so can prove most facts we want. What we are missing is an axiom to guarantee that if we start counting we eventually get to every natural number. This can be formulated a number of ways, but we will start with : Well-ordering Axiom: If S is any non-empty subset of N then there is a smallest element s 0 S (here smallest means that for any s 0 < s for all other s S.) This may not at first seem to be the same idea as claiming that we eventually reach every element of N by counting. Here is a more direct translation of that idea Proposition 0.18. Let A N and suppose that 1 A and for every a A we also have a + 1 A, then A = N. The idea here is that we are assuming 1 A, and applying the second assumption, 1 + 1 = 2 must also be in A. But then, applying it again, 2 + 1 = 3 is in A, etc. In other words, the assumptions here say roughly that A contains all the numbers you can reach by counting. The conclusion asserts that this must be all the natural numbers, which is a way of saying that every natural number can be reached by counting. We can prove this from the Well-ordering axiom as follows: Let S be the set N \ A. If S is empty then A = N as claimed, so we just need to rule out any other possibility. If S is not empty then by the well-ordering axiom there is a smallest s 0 ins. What can s 0 be? It can t be 1 since we assumed that 1 is in A, so we must have s 0 > 1. Then consider the number n = s 0 1. Since s 0 > 1 we have n > 0 so n N. We also know that n < s 0 so, since s 0 is the smallest element of S, n must not be in S. That means that n A. But then our assumption says that n + 1 is also in A, and n + 1 = s 0 so s 0 is in A. That is impossible since s 0 is also known to be in S. This proof outline is common : we what to show something is true for all n N, so we look at the set of numbers where it fails to hold and try to prove it is empty. If it isn t empty then it has a smallest element s (by well-ordering again). This is the smallest natural number for which our desired statement is false, so it must be true for s 1 (and all other smaller numbers), and we can often use this to contradict things. Here s a more concrete example: Proposition 0.19. Let n be a natural number. Either n = 2q for some integer q or n = 2q + 1 for some integer q. As before, let S be the set of natural numbers which are neither 2q for any q nor are 2q + 1 for any q. Our goal is to show that S is empty. If not,

4 WORKSHEET ON NUMBERS, MATH 215 FALL 18(WHYTE) then it has a smallest element, s 0. Since 1 = 2(0) + 1 we know that s 0 > 1, so s 0 1 is also a natural number. Since s 0 1 < s 0 and s 0 is the smallest element of S we must have s 0 1 not in S. This means that s 0 1 is either of the form 2q or 2q + 1. In the first case, s 0 1 = 2q we can add one to both sides to get s 0 = 2q, which is impossible since s 0 S. Similarly, if s 0 1 = 2q + 1 then s 0 = 2q + 2 = 2(q + 1) so s 0 is twice some integer, which is again impossible since s 0 S. The only possibility left is that S is empty, so that no such s 0 exists. Alternatively, you can prove such statements via Proposition 0.18. For example, let A be the set of natural numbers which can be written as 2q or 2q + 1 for some integer q. Then since 1 = 2(0) + 1 we have 1 A and, if a is in A then either: a = 2q for some q, so a + 1 = 2q + 1 and so a + 1 A or a = 2q + 1 for some q, so a + 1 = 2q + 2 = 2(q + 1) so a + 1 A] Since, in both cases, we have a + 1 A we can apply Proposition 0.18 to conclude A = N. Problem 0.20. Here are a few more things one can prove with these methods, try to write the proof both directly from the well-ordering axiom and via proposition 0.18: (1) Every natural number is 3q, 3q + 1, or 3q + 2 for some integer q. (2) For every natural number n, n < 2 n d (3) For all n, dx xn = nx n 1 (this assumes you know the product rule and that d dx x = 1

WORKSHEET ON NUMBERS, MATH 215 FALL 18(WHYTE) 5 We often want to use these arguments in more general settings: for subsets of the integers, with largest elements rather than smallest, and so on. To state these we start with some definitions: Definition 0.21. If S is a subset of Z then an integer n is a lower bound for S if n s for all s S. A set which has a lower bound is called bounded from below. Similarly, an integer m is an upper bound for S if s m for all s S, and a set with an upper bounded is called bounded from above. A set which is both bounded from above and from below is simply called bounded. Proposition 0.22. Here are some basic facts and questions: (1) If n is a lower bound for S and m n then m is also a lower bound for S. (2) If n is a lower bound for S and T S then n is also a lower bound for T. (3) Can you formulate versions of the above for upper bounds? (4) The set N is bounded from below. Is it bounded from above? (5) If S is the set of odd integers, is S bounded from above? from below? With these terms in mind we can state a more general form of the well ordering principle: Proposition 0.23. If S is a non-empty set of integers which is bounded from below then S contains a smallest element. (Hint: shift S into N by adding something, apply the basic well ordering principle, and then shift back) Similarly, we have: Proposition 0.24. If S is a non-empty set of integers which is bounded from above then S contains a largest element. (Hint #1: multiply by -1, apply the lower bound version, then multiply by -1 again. Hint #2: For a completely different proof, let T be the set of upper bounds for S and apply the well ordering principle to T ) These are all the basic rules (called axioms ) that we need to say what the properties of the integers are that distinguish them from all other kinds of numbers. These are listed on the next page. A good exercise is to look back over all the properties we assumed up until this point and make sure you know how they follow from the things listed (the most challenging to deduce is probably that there is no integer between 0 and 1).

6 WORKSHEET ON NUMBERS, MATH 215 FALL 18(WHYTE) AXIOMS for the Integers: (1) If a and b are integers then a + b and ab are integers. (2) For any a and b, a + b = b + a and ab = ba. (3) For any a, b, and c. a + (b + c) = (a + b) + c and a(bc) = (ab)c. (4) For any a,b, and c, a(b + c) = ab + ac. (5) There are integers 0 and 1 so that for any a, a = a + 0 = a1. (6) For any a there is a b with a + b = 0. (7) For any a and b exactly one of the following is true: a = b a < b b < a (8) If a < b and b < c then a < c. (9) If a < b then for any c, a + c < b + c. (10) If a > 0 and b > 0 then ab > 0. (11) (Well ordering) If S Z is non-empty and bounded from below then S contains a smallest element.

WORKSHEET ON NUMBERS, MATH 215 FALL 18(WHYTE) 7 Definition 0.25. Let a and b be integers and let n be a natural number. We say a is congruent to b modulo n if n (a b). We use the notation a b mod n for this property. Problem 0.26. Decide whether each of the following statements is true and justify your answers: 2 93 mod 13 27 4 mod 5 15 6 mod 7 3 8 mod 2 Proposition 0.27. If a is an integer and n a natural number then a a mod n. Proposition 0.28. Let a and b be integers and n a natural number. If a b mod n then b a mod n. Proposition 0.29. Let a, b, and c be integers and n a natural number. If a b mod n and b c mod n then a c mod n. Proposition 0.30. Let a, b, and c be integers and n a natural number. If a b mod n then a + c b + c mod n. Proposition 0.31. Let a, b, and c be integers and n a natural number. If a b mod n then ac bc mod n. Proposition 0.32. Let a, b, and c be integers and n a natural number. If a b mod n and b c mod n then a c mod n. Proposition 0.33. Let a and b be integers and n a natural number. If a b mod n then a 2 b 2 mod n. Proposition 0.34. Let a and b be integers and n a natural number. If a b mod n then a 3 b 3 mod n. The last two propositions suggest the following: Conjecture 0.35. Let a and b be integers and n and m natural numbers. If a b mod n then a m b m mod n. Do you believe this conjecture? Can you prove it?

8 WORKSHEET ON NUMBERS, MATH 215 FALL 18(WHYTE) Proposition 0.36. Let a be an integer and b a natural number. There is an integers q and r with 0 r < b such that: a = bq + r Here q is called the quotient and r is called the remainder. Here s a hint: think about all possible q and r 0 that make the equation hold (without the assumption r < b) and then use well ordering to find the smallest such r. This proposition is often called the division algorithm because it is tells you exactly what one gets from old fashioned long division of natural numbers - a quotient and a remainder. However there s a subtlety here - the proposition says that q and r exist, but not that they are unique - in other words the division problem might have more than one right answer. Obviously that s not what we expect. Here s how that is phrased precisely (make sure you understand why, then prove it): Proposition 0.37. Let b a natural number and q 1, q 2, r 1, r 2 integers with 0 r 1 < b and 0 r 2 < b such that q 1 b + r 1 = q 2 b + r 2 then q 1 = q 2 and r 1 = r 2. In particular, we get the following about congruences: Proposition 0.38. For any integer a and natural number n there is exactly one integer b with 0 b < n such that a b mod n. Definition 0.39. A natural number n > 1 is prime numbers m with m n are m = 1 and m = n if the only natural Proposition 0.40. If n > 1 is a natural number then there is a prime number p such that p n. Proposition 0.41. Every natural number n > 1 can be written as a product of primes: n = p k 1 1 pkm m where p 1, p 2,..., p m are prime numbers and k i are natural numbers. This is the prime factorization theorem. It usually also comes with a uniqueness statement. Can you figure out what this should say? It tuns out that the uniqueness is more subtle than it appears, so we will need to develop some other ideas before tackling it. We start with another definition: Definition 0.42. Given two integers a and b a common divisor is an integer d satisfying such that d a and d b. Proposition 0.43. Show that unless a and b are both zero they have a greatest common divisor (hint: fist prove that if x and y are natural numbers and x y then x y ) Problem 0.44. Let n be a natural number. divisor of n and 0. Find the greatest common

WORKSHEET ON NUMBERS, MATH 215 FALL 18(WHYTE) 9 Proposition 0.45. Let p be a prime number. Show that for any integer a the greatest common divisor of p and a is either p ( if p a) or is 1. Proposition 0.46. Let a and b be natural numbers. Write a = bq + r using the division algorithm. Show that the common divisors of a and b are the same as the common divisors of r and b. This gives a practical way to compute greatest common divisors: take the larger of a and b and replace it with its remainder when divided by the other, and repeat until one of the numbers is zero. Problem 0.47. What is the greatest common divisor of 120 and 168? of 59 and 1016? Question 0.48. Can you show that this process always works? There is another surprising way of characterizing the gcd. For two numbers a and b, we think about all the numbers you can get by adding multiples of a and b together. We can imaging this by thinking of a and b as dollar values of bills and then asking what prices can paid with them. For example, if your country only issues a 6 dollar bill and a 14 dollar bill, can you buy something that costs 10 dollars? Yes - you pay with two 14 dollar bills and get three 6 dollar bills back in change. Can you buy something that costs 15 dollars? No - all the bills are worth an even number of dollars so there is no way to get an odd net transaction. Formulated more abstractly: Let S(a, b) = {na + mb : n, m Z}. Proposition 0.49. If c is a common divisor of a and b then c s for all s S(a, b) Proposition 0.50. If s S(a, b) then gcd(a, b) s. Proposition 0.51. If s S(a, b) then sx S(a, b) for all x Z Proposition 0.52. If S(a, b) = Z if and only if 1 S Proposition 0.53. The set S(0, 0) is {0}. For any other a and b the set S(a, b) is infinite. Proposition 0.54. If a b then S(a, b) is precisely the set of multiples of a. The main fact we are aiming to prove is a more general version of the last statement: Proposition 0.55. For any a and b in Z, not both zero, the set S(a, b) is precisely the set of multiples of gcd(a, b). Problem 0.56. Show that proposition 0.55 is equivalent to the statement that gcd(a, b) S(a, b).

10 WORKSHEET ON NUMBERS, MATH 215 FALL 18(WHYTE) The key ingredient in the proof of proposition 0.55 is this: Proposition 0.57. Let s S(a, b) and write s = aq + r using the division algorithm, then r S(a, b) (note that the same works if we divide by b instead of a). Using that you can prove: Proposition 0.58. If c is the smallest positive element of S(a, b) then c is a common divisor of a and b. This is almost enough to finish the proof of proposition 0.55. The only missing piece is to show that this c must be the greatest common divisor, not a smaller one. But using the earlier propositions we know that c must be divisible by gcd(a, b). Problem 0.59. Put all of the above together to give a complete proof of Proposition 0.55.