Writing proofs for MATH 61CM, 61DM Week 1: basic logic, proof by contradiction, proof by induction

Similar documents
Mathematical Induction

Proofs. Chapter 2 P P Q Q

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

Proofs. Chapter 2 P P Q Q

Basics of Proofs. 1 The Basics. 2 Proof Strategies. 2.1 Understand What s Going On

Some Review Problems for Exam 1: Solutions

Lecture 5 : Proofs DRAFT

Contradiction MATH Contradiction. Benjamin V.C. Collins, James A. Swenson MATH 2730

MATH10040: Chapter 0 Mathematics, Logic and Reasoning

MATH CSE20 Homework 5 Due Monday November 4

Techniques for Proof Writing

Section 3.1: Direct Proof and Counterexample 1

3 The language of proof

Basic Logic and Proof Techniques

Cool Results on Primes

Proofs: A General How To II. Rules of Inference. Rules of Inference Modus Ponens. Rules of Inference Addition. Rules of Inference Conjunction

Math 38: Graph Theory Spring 2004 Dartmouth College. On Writing Proofs. 1 Introduction. 2 Finding A Solution

means is a subset of. So we say A B for sets A and B if x A we have x B holds. BY CONTRAST, a S means that a is a member of S.

Sec$on Summary. Mathematical Proofs Forms of Theorems Trivial & Vacuous Proofs Direct Proofs Indirect Proofs

Introduction to Basic Proof Techniques Mathew A. Johnson

Proof by Contradiction

Section 2.1: Introduction to the Logic of Quantified Statements

Math 300 Introduction to Mathematical Reasoning Autumn 2017 Proof Templates 1

Topics in Logic and Proofs

Proof Techniques (Review of Math 271)

Basic Proof Techniques

Notes: Pythagorean Triples

1 Direct Proofs Technique Outlines Example Implication Proofs Technique Outlines Examples...

Mathematical Induction Again

Chapter 2. Mathematical Reasoning. 2.1 Mathematical Models

Mathematical Induction Again

Inference and Proofs (1.6 & 1.7)

What can you prove by induction?

LECTURE 1. Logic and Proofs

CISC-102 Fall 2017 Week 3. Principle of Mathematical Induction

Introducing Proof 1. hsn.uk.net. Contents

Notes on arithmetic. 1. Representation in base B

MATH10040: Numbers and Functions Homework 1: Solutions

Chapter 1: The Logic of Compound Statements. January 7, 2008

Basic Proof Examples

MA103 STATEMENTS, PROOF, LOGIC

MATH 2200 Final LC Review

PROBLEM SET 3: PROOF TECHNIQUES

Discrete Mathematics and Probability Theory Fall 2016 Seshia and Walrand Note 2

Proof by contrapositive, contradiction

Foundations of Advanced Mathematics, version 0.8. Jonathan J. White

HOW TO WRITE PROOFS. Dr. Min Ru, University of Houston

Chapter 1A -- Real Numbers. iff. Math Symbols: Sets of Numbers

AN INTRODUCTION TO MATHEMATICAL PROOFS NOTES FOR MATH Jimmy T. Arnold

Logic and Proofs 1. 1 Overview. 2 Sentential Connectives. John Nachbar Washington University December 26, 2014

Proofs. Introduction II. Notes. Notes. Notes. Slides by Christopher M. Bourke Instructor: Berthe Y. Choueiry. Fall 2007

Lecture 4: Constructing the Integers, Rationals and Reals

Mathematical Writing and Methods of Proof

MCS-236: Graph Theory Handout #A4 San Skulrattanakulchai Gustavus Adolphus College Sep 15, Methods of Proof

Direct Proof and Counterexample I:Introduction

Writing Mathematical Proofs

Direct Proof and Counterexample I:Introduction. Copyright Cengage Learning. All rights reserved.

MATH 271 Summer 2016 Practice problem solutions Week 1

Mathematical Induction

CITS2211 Discrete Structures Proofs

What is a proof? Proofing as a social process, a communication art.

FORMAL PROOFS DONU ARAPURA

THE DIVISION THEOREM IN Z AND R[T ]

Writing proofs. Tim Hsu, San José State University. May 31, Definitions and theorems 3. 2 What is a proof? 3. 3 A word about definitions 4

CS1800: Mathematical Induction. Professor Kevin Gold

Math Real Analysis

Theorem. For every positive integer n, the sum of the positive integers from 1 to n is n(n+1)

Your quiz in recitation on Tuesday will cover 3.1: Arguments and inference. Your also have an online quiz, covering 3.1, due by 11:59 p.m., Tuesday.

Lecture Notes on DISCRETE MATHEMATICS. Eusebius Doedel

8. Given a rational number r, prove that there exist coprime integers p and q, with q 0, so that r = p q. . For all n N, f n = an b n 2

Practice Test III, Math 314, Spring 2016

Practice Midterm Exam Solutions

Midterm Exam Solution

Mathematics 220 Midterm Practice problems from old exams Page 1 of 8

An Introduction to Mathematical Reasoning

The following techniques for methods of proofs are discussed in our text: - Vacuous proof - Trivial proof

Discrete Mathematics and Probability Theory Fall 2012 Vazirani Note 1

Foundations of Mathematics MATH 220 FALL 2017 Lecture Notes

Discrete Mathematics and Probability Theory Fall 2018 Alistair Sinclair and Yun Song Note 2

CMPSCI 250: Introduction to Computation. Lecture 11: Proof Techniques David Mix Barrington 5 March 2013

LECTURE NOTES DISCRETE MATHEMATICS. Eusebius Doedel

MATH 341, Section 001 FALL 2014 Introduction to the Language and Practice of Mathematics

Solutions to Homework Set 1

How to write maths (well)

CHAPTER 4: EXPLORING Z

Chapter 17. Proof by Contradiction The method

Chapter 1 Elementary Logic

MAT 300 RECITATIONS WEEK 7 SOLUTIONS. Exercise #1. Use induction to prove that for every natural number n 4, n! > 2 n. 4! = 24 > 16 = 2 4 = 2 n

Reading and Writing. Mathematical Proofs. Slides by Arthur van Goetham

At the start of the term, we saw the following formula for computing the sum of the first n integers:

Summer HSSP Week 1 Homework. Lane Gunderman, Victor Lopez, James Rowan

MATH 135 Fall 2006 Proofs, Part IV

Preparing for the CS 173 (A) Fall 2018 Midterm 1

Mathematical Reasoning Rules of Inference & Mathematical Induction. 1. Assign propositional variables to the component propositional argument.

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

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

Contribution of Problems

Foundations of Discrete Mathematics

We are going to discuss what it means for a sequence to converge in three stages: First, we define what it means for a sequence to converge to zero

Seunghee Ye Ma 8: Week 1 Notes September 29, 2016

Transcription:

Writing proofs for MATH 61CM, 61DM Week 1: basic logic, proof by contradiction, proof by induction written by Sarah Peluse, revised by Evangelie Zachos and Lisa Sauermann September 27, 2016 1 Introduction A proof is a mathematical argument for why a statement is true. In contrast to more subjective disciplines, concepts in math have rigorous descriptions and deductions must be 100% logically sound. The purpose of writing a proof is to communicate your argument to others. Therefore, it is important for you to learn how to write clear proofs that other people can understand. Math involves a lot of writing, and the standards for grammar and clarity are the same as in any other field. In particular, always write in complete sentences. Writing proofs is a separate skill from problem solving. When you do not have much experience writing proofs, many homework-type problems can appear to be much harder than they actually are. In these proof writing sections, we ll cover the proof strategies you will need to know for this class so that this artificial difficulty disappears. The formalisms you see in proofs are actually there to make your job easier, and with a little practice you will become fluent in them. Today we will cover basic logical terms and two general proof techniques: proof by contradiction and proof by induction. For each topic we ll present a proof of a simple result and break down the purpose of each line, except in the section on induction, for which instead there is an extended explanation and proof template. There will also be a more proof examples without the breakdown, and a collection of optional exercises. These optional exercises are meant to be easier than the homework assignment, so that they can serve as a warm-up if you are getting stuck. 2 Basic logic and notation This section will introduce some terms from classical logic and the corresponding notation. The basic mathematical proposition format is if P, then Q. For example, If n is an integer, then 2n is an integer. If n is an integer, then n + 1 is an integer. If n is a positive integer, then n + 1 is a positive integer. 1

* If n is an integer, then n + 1 is a positive integer. [This is a false statement! Propositions can sometimes be false.] Sometimes these statements are written slightly differently, as Q if P, or even with the the symbol = instead. For example: n is an integer = 2n is an integer. Even if a proposition if P, then Q is true, that does not mean that the converse if Q, then P is true. For example, using the three true statements above and taking the converse: *If 2n is an integer, then n is an integer. [FALSE] If n + 1 is an integer, then n is an integer. [TRUE] *If n + 1 is a positive integer, then n is a positive integer. [FALSE] If both the proposition and the converse are true, then the statement P if and only if Q (equivalently Q if and only if P ) is true. This statement can also be written as P iff Q or P Q. For example: n is an integer if and only if n + 1 is an integer. n is an integer iff n + 1 is an integer. n is an integer n + 1 is an integer. If you are proving an if and only if statement, or an equivalence, your proof will usually need to have two parts: one for the if part and one for the only if part. Although P = Q and Q = P are not equivalent statements, there is a statement equivalent to P = Q that is often useful when proving things. This statement is called the contrapositive and goes as follows: Examples: if not Q, then not P. If 2n is not an integer, then n is not an integer. If n + 1 is not an integer, then n is not an integer. If n + 1 is not a positive integer, then n is not a positive integer. Finally, we want to introduce two quantifiers, which you will see more of next week. These are for all and there exists. Often choosing the right quantifier can make a statement true or false, interesting or boring. Here are some examples: For all integers n, 2n is an integer. [TRUE] There exists an integer n so that 2n is an integer. [TRUE BUT DUMB] 2

*For all integers n, n/2 is an integer. [FALSE] There exists an integer n so that n/2 is an integer. [TRUE] There are also symbols for these two qualifiers. means for all and means there exists. integers n, 2n is an integer. an integer n so that n/2 is an integer. Finally, there are two abbreviations you might see: (1) i.e. stands for id est and means that is or in other words, (2) e.g. stands for exempli gratia and means for the sake of example or for example. Here are examples of how to use them: 7 is a prime number, i.e. a number divisible only by itself and one. We take a prime number, e.g. 5, and.... Note: when in doubt, write things in terms of words instead of symbols or abbreviations. 3 Proving universal statements using definitions Most mathematical statements that you ll encounter in your life are universal statements. These statements say that for all ( ) elements in a particular set, some result holds. For example: The cube of an odd integer is odd. If n, m Z and m is even, then nm is even. For n N, n j=1 j3 = n2 (n+1) 2 4 (HW 1, 61CM/DM) For any vectors x, y R n, x y 2 + x + y 2 = 2( x 2 + y 2 ) (HW 1, 61CM/DM) are all examples of universal statements. If the set in question is finite, you can prove the statement by checking each individual element (although you probably do not want to). However, if the set in question is infinite, then you cannot prove the statement by checking any number of individual elements. In this case, there are a number of approaches. One is to prove universal statements using definitions. The best way to start such an approach is to write down all definitions involved and noodle around until you have a logical proof. Here are some very simple examples so you can see how this works. (Note: some of the statements above need techniques in later sections.) 3

3.1 Examples Proposition. The square of an even integer is a multiple of 4. Short Proof. Any even integer a can be written as a = 2b for some integer b, and then a 2 = 4b 2. Because b 2 is an integer, a 2 is a multiple of 4. Longer Proof. Let a be an even integer. Then a = 2 b for some integer b. We have a 2 = (2 b) (2 b) = 4 b 2. Because the product of two integers is an integer and b is an integer, b 2 is an integer. This means that a 2 is the product of 4 and an integer, i.e. a 2 is a multiple of 4. Since a was an arbitrary even integer, we conclude that every even integer is a multiple of four. We will now break down the longer proof. Which type of proof should you be aiming for? In general, when writing a proof, you should write every fact that is not completely obvious to you. Thus your beginning proofs will be a bit longer than proofs that will appear in class or in books. This tendency is completely fine and is in fact to be encouraged! Short and clear proofs are a culmination of developed skill, not a starting point. Let a be an even integer. Computing 2 2, 4 2, 6 2, and 8 2 and seeing that they are all multiples of 4 is not a proof, even though by computing these examples you may see why the proposition should be true. Our goal is to show that the square of every even integer is a multiple of four, and to do this we need a general even number. Any specific even number will not suffice, even if we check a million of them. We give our general even number the name a so that we can refer to it in our argument. Then a = 2 b for some integer b. This is just a statement of the definition of an even number. Since a is a general even number, the above is really the only information available for us to use. Because a is a general even number, it is the product of 2 and a general integer, which we name b so that we can refer to it later. We have a 2 = (2 b) (2 b) = 4 b 2. We are interested in what the square of a general even number is, so we square our general even number a. The square turns out to be a nice expression, 4 b 2. Because the product of two integers is an integer and b is an integer, b 2 is an integer. 4

Recall that our goal is to show that a 2 is a multiple of 4, since a is our general even integer. We have shown that a 2 = 4 b 2, so we re done if b 2 is an integer, because then a 2 has the form of a general multiple of 4. You are free to use the fact that the product of two integers is again an integer without explanation. Because b 2 = b b, b 2 is thus an integer. This means that a 2 is the product of 4 and an integer, i.e. a 2 is a multiple of 4. Putting together the two previous sentences, we see that a 2 has the form of a general multiple of 4. Compare and contrast this sentence with the second sentence of the proof. There we unravel the definition of an even number, and here we reravel the definition of being a multiple of 4. Since a was an arbitrary even integer, we conclude that every even integer is a multiple of four. This line explains and justifies the inclusion of the first sentence of the proof, for we really have proved that the result holds for any even number. It is also good practice in to finish a proof by stating the result that you have just proved. It is nice to include a symbol to indicate the end of a proof. Many books use this box. Some people instead put Q.E.D., which stands for the Latin quod erat demonstrandum, meaning which was to be demonstrated/shown. This is a translation of the Ancient Greek phrase, óπɛρ ɛδɛι δɛίξαι, originally used by Euclid at the end of his proofs. Here is a second example of a universal statement and its proof, with a twist. Proposition. Let n be an integer. If n 2 is even, then n is even. To prove this proposition, we will use a trick from the previous section and prove the contrapositive. Note that this is still an example of a proof of a universal statement using definitions. Proof. We will prove the contrapositive: if n is odd, then n 2 is odd. If n is odd, then it can be written as n = 2a+1 where a is an integer. Then n 2 = (2a+1) 2 = 4a 2 +4a+1 = 2(2a 2 +2a)+1. Because 2a 2 + 2a is an integer, n 2 is an odd integer. Thus any odd number n has an odd square, and so we conclude by the contrapositive that any even square n 2 has an even root n. Proposition. The product of two rational numbers is rational. 5

Proof. Let a, c Q. We have b d a b c d = ac bd. By definition, a, b, c, d Z and b, d 0. Since the product of two integers is an integer, ac, bd Z. Since neither b nor d is zero, bd 0 as well. This implies that a c = p, where b d q p = ac and q = bd are integers and q 0. That is, a c is a rational number. Thus, since b d a, c Q were arbitrary rational numbers, we conclude that the product of two rational b d numbers is rational. 3.2 Exercises Exercise 1. Prove that the sum of two odd integers is even. Exercise 2. Prove that the product of two integers which are perfect squares is a perfect square. Exercise 3. Prove that the sum of two rational numbers is a rational number. Exercise 4. If a, b Z, we say that a divides b if there exists a c Z such that b = ac. We use the notation a b to mean that a divides b. For example, 2 4, 3 6, and 13 143. Show that 1 divides any integer. Show that any integer divides 0. Show that if n is an odd integer, then 4 n 2 1. Show that if a b and a c, then a (b + c). Show that if a b and b c, then a c. That is, the divisibility relation is transitive. Show that if d n and n is odd, then d is odd. 4 Proof by contradiction Suppose we want to prove some mathematical statement. One strategy is to show that if the statement weren t true, then we d be led to a conclusion that we know to be false. Thus, statement must be true, for assuming that it is false leads to a contradiction. 4.1 Examples We start with a simple example to understand this technique. Proposition. If z R is irrational, then so is z. Proof. Let z R be irrational. Assume for the sake of contradiction that z were rational. Then there exist p, q Z with q 0 such that z = p. But then z = p = p. This implies q q q that z is rational, which gives a contradiction since we assumed that z was irrational. 6

We will now break down the proof. Let z R be irrational. As in any proof of a universal statement, we take z to be an arbitrary irrational number. Assume for the sake of contradiction that z were rational. You should signal some way at the beginning of a proof by contradiction that this is the technique that you are planning to use. After we inform the reader of our strategy, we state the negation of the result that we are trying to prove. If z were not irrational, then it would be rational by definition. Then there exist p, q Z with q 0 such that z = p q. This is the definition of what it means for z to be rational. Since we re trying to derive a contradiction from z being rational, we should have to invoke the definition at some point. But then z = p q = p q. There is nothing stopping z from being rational when z R is arbitrary, but the proposition we re trying to prove has us assume that z is irrational. In fact, this is the only thing we re assuming, so we must use it some time in the proof. We want to connect what we know about z to what we know about z somehow. Since z is just z times ( 1), z must be equal to p q = p q. This implies that z is rational, which gives a contradiction since we assumed that z was irrational. The previous line told us that z can be written as the quotient of two integers p and q, q 0. So by definition, z is rational. But by hypothesis z is irrational. Since a number cannot be both rational and irrational, this is a contradiction. Here are some more substantial examples. Proposition. There are no integer solutions to the equation x 3 + x + 1 = 0. 7

Proof. Assume for the sake of contradiction that x = n is an integer solution of x 3 +x+1 = 0. There are two possibilities. Either n is even or n is odd. If n is even, then n = 2a for some integer a and then n 3 +n+1 = 8a 3 +2a+1 = 2(4a 3 +a)+1 is odd. If n is odd, then n = 2b+1 for some integer b and then n 3 +n+1 = 8b 3 +12b 2 +6b+1+2b+1+1 = 2(4b 3 +6b 2 +3b+b+1)+1 is odd. In either case, we have proven that 0 is odd, which gives us a contradiction. Thus no integer solution to x 3 + x + 1 = 0 exists. Proposition. For every real number x [0, π/2], sin x + cos x 1. Proof. Assume for the sake of contradiction that there exists such an x [0, π/2] so that sin x + cos x < 1. By the choice of domain, sin x, cos x 0 so that sin x + cos x 0. Then sin 2 x + cos 2 x + 2 sin x cos x < 1, and so 2 sin x cos x < 0. This means that either (sin x) or (cos x) must be less than 0, but that is a contradiction of our choice of x, because 0 x π/2. Thus for any x [0, π/2], sin x + cos x 1. 4.2 Proofs using the contrapositive vs. proofs by contradiction See if you can rewrite the proof that for any integer n, n 2 is even = n is even into a proof by contradiction. In fact any direct proof of the contrapositive can be worked into a proof by contradiction. Which one you use is a matter of taste. (I prefer proving the contrapositive generally. It is often shorter.) However, proofs by contradiction are more general. Let us say you want to prove if P, then Q. When proving the contrapositive, you assume (not Q) and conclude (not P). To rewrite this into a proof by contradiction, you would assume (P) and (not Q) and derive (not P), which contradicts (P). However, there are proofs by contradiction which assume (P) and (not Q) but derive a different contradiction. For example, when we proved that there were no integer solutions to x 3 + x + 1 = 0, the contradiction we derived was not that n was not an integer. Instead, we derived that 0 is an odd integer. This is an example of a proof that you can not simply rewrite to be a proof of the contrapositive. 4.3 Exercises Exercise 5. Prove that there are no rational solutions to the equation x 3 + x + 1 = 0. Exercise 6. Prove that there are no common prime factors of n and n + 1. Exercise 7. Prove that 5 is irrational. Exercise 8. Let z > 0 be irrational. Show that z is irrational. Show that n z is irrational for all integers n 3 as well. 5 Proof by induction The second strategy that we ll cover today is that of proof by induction. If you are interested in computer science, then this proof strategy will be very important to you in the future. 8

It is easiest to explain the method by presenting an example first. Let s prove the equality 1 + 2 + + 2 n = 2 n+1 1. Checking it for the case n = 1 is easy: the left hand side is 1 + 2 = 3, and the right hand side is 2 1+1 1 = 4 1 = 3 as well. Suppose that we have checked it by hand for n up to 10, and have gotten bored of all of the arithmetic. To prove it for the case n = 11, we just have to note that the left hand side is (1 + 2 + + 2 10 ) + 2 11, and we have already computed 1 + 2 + + 2 10 : 2 10+1 1 = 2 11 1. Hence, 1 + 2 + + 2 11 = 2 11 1 + 2 11 = 2 2 11 1 = 2 12 1, and we have proved the result for n = 11. We can generalize this argument. Suppose that we have proved the claim for all integers n k. That is, we have shown that 1 + 2 + + 2 n = 2 n+1 1 for all integers n k. Then we have all we need to prove the claim for n = k + 1, for 1 + 2 + + 2 k+1 = (1 + 2 + + 2 k ) + 2 k+1, and we know the result for n = k: 1 + 2 + + 2 k = 2 k+1 1. Thus, we can plug it in the equation above to get 1 + 2 + + 2 k+1 = (1 + 2 + + 2 k ) + 2 k+1 = 2 k+1 1 + 2 k+1 = 2 2 k+1 1 = 2 (k+1)+1 1, which gives the claim when n = k + 1. So we have shown that if the claim holds for n = k, then it also holds for n = k + 1. We have also shown that it holds for n = 1. Combining these together, we see that it holds for n = 2. Since it holds for n = 2, it must hold for n = 3. Similarly, since it holds for n = 3, it must hold for n = 4 as well, and so on. Thus, we have proved the claim for all n N. The outline above gives the basic structure of any inductive proof, where you are proving a statement P(n) with a parameter n. Such a proof needs both a base case, which here is n = 1, and an inductive step, which proves the desired statement P(k + 1) by assuming P(k). Often the statement for P(k) is called the inductive hypothesis. Using the inductive hypothesis to prove P(k + 1) is usually the difficult part. Note that the base case does not need to be at n = 1. Sometimes it is at n = 0, or n = 2, or n = 10. Whatever base case you pick, P (n) is true for all n beyond the base case. What are some examples of statements P(n) that can be proven by induction? 1 + 2 + + 2 n = 2 n+1 1 n i = n(n+1) 2. 9

Any positive integer n can be written as a product of prime numbers An under-determined system of n linear equations always has a nontrivial solution. The above explanation is actually a little oversimplified. What do you notice about the following proof by induction? Proposition. Every integer n 2 can be written as a product of prime numbers. Proof. We prove this statement by induction. The number is 2, so the statement is true for n = 2. If we assume that the statement is true for all n k, i.e. that all integers 2 n k have a prime factor, we will now prove the same for k + 1. If k + 1 is prime, then it itself provides a factorization into primes. If k +1 is not prime, then by definition it equals k +1 = a b where both a and b are integers smaller than k +1. By the inductive hypothesis, both a and b have a factorization into prime numbers. Multiplying these factorizations gives a factorization of k + 1 into prime numbers, as we wished to show. Thus every integer n 2 can be written as a product of prime numbers. First, note that here the base case is n = 2. The statement is false for n = 1, depending on how you view it. But, more importantly, note that our inductive hypothesis was not P(k). Instead, it was P(2) and P(3) and P(4) and... and P(k). Sometimes this is called strong induction. It is logically equivalent to the induction presented above, but sometimes it is much simpler to write! To help you stay organized, here is a template for writing a proof by induction: Proof. We proceed by induction on n. Base case: First we check that P( base case ) is true. proof of P( base case ) Thus, P( base case ) is true. Induction step: Suppose that P(n) is true for a general n show that P(n + 1) is true as well. proof of P(n + 1) Thus, P(n + 1) is true. This completes the induction base case. We will Can you revise this template to a template for strong induction? 5.1 Examples Proposition. For x R and n Z >0, d dx xn = nx n 1. 10

Proof. We proceed by induction on n. For the base case we take n = 1. Because x 1 = x is a linear function, its derivative is the constant slope 1 = 1 x 0. Now, assume the inductive hypothesis that d dx xn = nx n 1. We evaluate d dx xn+1 using the product rule: d dx xn+1 = x n d dx x + x d dx xn = x n + x(nx n 1 ) = (n + 1)x n Thus the inductive step is completed, and the statement is proven. Note that in the next example, the base case is n = 0, instead of n = 1 as above. Proposition. Let r R with r 1 and n Z, n 0. Then n r i = 1 + r + r 2 + r 3 + + r n = rn+1 1 r 1 Proof. We proceed by induction on n. Base case: First we check the claim when n = 0. We have 0 r 1 1 = 1. So indeed r 1 n r i = rn+1 1 r 1 ri = 1 and r0+1 1 r 1 = holds when n = 0. Induction step: Suppose that the result holds for a general n 0. We will now prove that the result holds for n + 1. We have ( n+1 n ) r i = r i + r n+1. By the induction hypothesis, n ri = rn+1 1 r 1, so ( n+1 n ) r i = r i + r n+1 = rn+1 1 r 1 + r n+1 = rn+1 1 + (r 1)r n+1 r 1 = rn+1 1 + r n+2 r n+1 r 1 = rn+2 1 r 1. Thus, the result holds for n + 1. This completes the induction. 11

5.2 Exercises Exercise 9. Prove by induction that n i = n(n + 1). 2 Exercise 10. Prove by induction that n (2i 1) = 1 + 3 + + (2n 1) = n 2. i=1 That is, the sum of the first n odd numbers equals n 2. Exercise 11. Prove by induction that for all n N, 6 (n 3 n). 6 General tips These tips will hold in general when you are writing proofs. We ll remind you of them next week because they are important! The the purpose of a proof is to communicate an argument. A common mistake with beginning proof writers is to write down a proof that is less complete than one they would give verbally. But your writing alone must convey the argument without you being there to explain what you meant to say. Here are some general guidelines to follow which will make your proofs clearer and more readable: Write in complete, grammatically correct sentences. Signal to your reader what you are doing in your proof. That means: If you are proving a universal statement about elements in some set (for example, a vector space), say that you are arguing with a general element of that set. Don t just start doing an argument with a vector v without saying what v is. If you are doing a proof by contradiction, say at the beginning of the proof that you are doing a proof by contradiction. If you are doing a proof by induction, say at the beginning of the proof that you are doing a proof by induction and clearly label the base case and induction step somehow. If at some point in the proof you want to break the argument up into cases, say that you are breaking the proof up into cases and label the cases. If your proof has multiple parts (for example, an proof or showing that two sets are equal), then before each of these parts state what you are proving. If you are handwriting your homework, try to figure out most or all of the proof before starting to write out your final copy. This technique will make it much easier to signal effectively. 12