MATH 131A: REAL ANALYSIS (BIG IDEAS)

Similar documents
Logical Connectives and Quantifiers

Analysis Finite and Infinite Sets The Real Numbers The Cantor Set

Definitions & Theorems

Math 117: Honours Calculus I Fall, 2002 List of Theorems. a n k b k. k. Theorem 2.1 (Convergent Bounded) A convergent sequence is bounded.

MATH 409 Advanced Calculus I Lecture 10: Continuity. Properties of continuous functions.

Undergraduate Notes in Mathematics. Arkansas Tech University Department of Mathematics

MAT 570 REAL ANALYSIS LECTURE NOTES. Contents. 1. Sets Functions Countability Axiom of choice Equivalence relations 9

Continuity. Chapter 4

Functions. Chapter Continuous Functions

Proof. We indicate by α, β (finite or not) the end-points of I and call

Lecture 10. Riemann-Stieltjes Integration

MATH 409 Advanced Calculus I Lecture 12: Uniform continuity. Exponential functions.

Solutions Final Exam May. 14, 2014

Copyright 2010 Pearson Education, Inc. Publishing as Prentice Hall.

5. Some theorems on continuous functions

Economics 204 Summer/Fall 2011 Lecture 5 Friday July 29, 2011

converges as well if x < 1. 1 x n x n 1 1 = 2 a nx n

Math 117: Honours Calculus I Fall, 2012 List of Theorems. a n k b k. k. Theorem 2.1 (Convergent Bounded): A convergent sequence is bounded.

M2PM1 Analysis II (2008) Dr M Ruzhansky List of definitions, statements and examples Preliminary version

Continuity. Chapter 4

M2P1 Analysis II (2005) Dr M Ruzhansky List of definitions, statements and examples. Chapter 1: Limits and continuity.

Math LM (24543) Lectures 02

REVIEW OF ESSENTIAL MATH 346 TOPICS

Solutions Final Exam May. 14, 2014

Principle of Mathematical Induction

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

MATH 202B - Problem Set 5

Copyright c 2007 Jason Underdown Some rights reserved. statement. sentential connectives. negation. conjunction. disjunction

1. Theorem. (Archimedean Property) Let x be any real number. There exists a positive integer n greater than x.

Maths 212: Homework Solutions

x y More precisely, this equation means that given any ε > 0, there exists some δ > 0 such that

2 Sequences, Continuity, and Limits

MATH 101, FALL 2018: SUPPLEMENTARY NOTES ON THE REAL LINE

Part 2 Continuous functions and their properties

1. Supremum and Infimum Remark: In this sections, all the subsets of R are assumed to be nonempty.

MATH NEW HOMEWORK AND SOLUTIONS TO PREVIOUS HOMEWORKS AND EXAMS

Mathematical Methods for Physics and Engineering

Real Analysis Math 131AH Rudin, Chapter #1. Dominique Abdi

A LITTLE REAL ANALYSIS AND TOPOLOGY

In N we can do addition, but in order to do subtraction we need to extend N to the integers

Math 140A - Fall Final Exam

Sequences. Chapter 3. n + 1 3n + 2 sin n n. 3. lim (ln(n + 1) ln n) 1. lim. 2. lim. 4. lim (1 + n)1/n. Answers: 1. 1/3; 2. 0; 3. 0; 4. 1.

MATH5011 Real Analysis I. Exercise 1 Suggested Solution

Contents Ordered Fields... 2 Ordered sets and fields... 2 Construction of the Reals 1: Dedekind Cuts... 2 Metric Spaces... 3

Carefully tear this page off the rest of your exam. UNIVERSITY OF TORONTO SCARBOROUGH MATA37H3 : Calculus for Mathematical Sciences II

MATH 409 Advanced Calculus I Lecture 7: Monotone sequences. The Bolzano-Weierstrass theorem.

Homework 4, 5, 6 Solutions. > 0, and so a n 0 = n + 1 n = ( n+1 n)( n+1+ n) 1 if n is odd 1/n if n is even diverges.

Advanced Calculus Math 127B, Winter 2005 Solutions: Final. nx2 1 + n 2 x, g n(x) = n2 x

1 Definition of the Riemann integral

MAS221 Analysis, Semester 1,

g 2 (x) (1/3)M 1 = (1/3)(2/3)M.

Quick Tour of the Topology of R. Steven Hurder, Dave Marker, & John Wood 1

Metric Spaces and Topology

Introductory Analysis 2 Spring 2010 Exam 1 February 11, 2015

Walker Ray Econ 204 Problem Set 3 Suggested Solutions August 6, 2015

In N we can do addition, but in order to do subtraction we need to extend N to the integers

1 Homework. Recommended Reading:

Review Notes for the Basic Qualifying Exam

Real Analysis - Notes and After Notes Fall 2008

Definition 2.1. A metric (or distance function) defined on a non-empty set X is a function d: X X R that satisfies: For all x, y, and z in X :

MATH 51H Section 4. October 16, Recall what it means for a function between metric spaces to be continuous:

Solution of the 8 th Homework

Math 117: Infinite Sequences

Real Analysis Chapter 4 Solutions Jonathan Conder

FUNCTIONS OF A REAL VARIABLE

d(x n, x) d(x n, x nk ) + d(x nk, x) where we chose any fixed k > N

Notes on Complex Analysis

Notes on uniform convergence

1 + lim. n n+1. f(x) = x + 1, x 1. and we check that f is increasing, instead. Using the quotient rule, we easily find that. 1 (x + 1) 1 x (x + 1) 2 =

2.1 Convergence of Sequences

Chapter 3 Continuous Functions

2.2 Some Consequences of the Completeness Axiom

THEOREMS, ETC., FOR MATH 515

A Short Journey Through the Riemann Integral

Supremum and Infimum

MATH 102 INTRODUCTION TO MATHEMATICAL ANALYSIS. 1. Some Fundamentals

Iowa State University. Instructor: Alex Roitershtein Summer Homework #5. Solutions

h(x) lim H(x) = lim Since h is nondecreasing then h(x) 0 for all x, and if h is discontinuous at a point x then H(x) > 0. Denote

Limit and Continuity

MATH 131A: REAL ANALYSIS

Induction, sequences, limits and continuity

Chapter 1 The Real Numbers

G1CMIN Measure and Integration

1/12/05: sec 3.1 and my article: How good is the Lebesgue measure?, Math. Intelligencer 11(2) (1989),

FINAL REVIEW FOR MATH The limit. a n. This definition is useful is when evaluating the limits; for instance, to show

Solutions to Homework 9

Week 2: Sequences and Series

Economics 204 Fall 2011 Problem Set 2 Suggested Solutions

REAL ANALYSIS II: PROBLEM SET 2

Lemma 15.1 (Sign preservation Lemma). Suppose that f : E R is continuous at some a R.

Introduction to Real Analysis

Econ Slides from Lecture 1

Set, functions and Euclidean space. Seungjin Han

P-adic Functions - Part 1

MASTERS EXAMINATION IN MATHEMATICS SOLUTIONS

Math 320: Real Analysis MWF 1pm, Campion Hall 302 Homework 8 Solutions Please write neatly, and in complete sentences when possible.

4. We accept without proofs that the following functions are differentiable: (e x ) = e x, sin x = cos x, cos x = sin x, log (x) = 1 sin x

We have been going places in the car of calculus for years, but this analysis course is about how the car actually works.

Hyperreal Calculus MAT2000 Project in Mathematics. Arne Tobias Malkenes Ødegaard Supervisor: Nikolai Bjørnestøl Hansen

CHAPTER 6. Limits of Functions. 1. Basic Definitions

Transcription:

MATH 131A: REAL ANALYSIS (BIG IDEAS) Theorem 1 (The Triangle Inequality). For all x, y R we have x + y x + y. Proposition 2 (The Archimedean property). For each x R there exists an n N such that n > x. Definition 3. A Cauchy sequence is a sequence of real numbers (a n ) such that for every ε > 0 there exists a positive integer N such that a m a n < ε whenever m, n N. Proposition 4. For each n 1 let H n := 1 + 1 + 1 + + 1. Then the sequence 2 3 n (H 1, H 2, H 3,...) is not a Cauchy sequence. Proof. For each n N, consider the difference H 2n H n = 1 n + 1 + 1 n + 2 + 1 n + 3 + + 1 2n. Notice that there are exactly n terms in the sum, and each term is larger than (or equal to) the smallest one 1. Thus 2n H 2n H n 1 2n + + 1 2n = 1 2. So for each N 1 let m = 2N and n = N; then H m H n 1. 2 Definition 5. A sequence (a n ) is bounded if there exists M such that a n M for all n 1. Proposition 6. Every Cauchy sequence is bounded. Proof. Suppose that (a n ) is a Cauchy sequence. Taking ε = 1, this means that there is some N N for which a m a n 1 for all m, n N. This splits our sequence into a finite piece (indices 1 through N) and an infinite piece (index N and after). By Lemma??, the finite piece is bounded, say by M. Using the Triangle Inequality we find that a n = a n a N + a N a n a N + a N < 1 + M. It follows that (a n ) is bounded by M + 1. Definition 7. Two sequences (a n ) and (b n ) are equivalent if for each ε > 0 there exists a natural number N (which can depend on ε) such that a n b n < ε whenever n N. Proposition 8 (Q is dense in R). For every pair of real numbers x, y with x < y there exists a rational number q such that x < q < y. Definition 9. Let (a n ) be a sequence of real numbers. We say that (a n ) converges to L R (and we write a n = L) if for every ε > 0 there exists a natural number N for which a n L < ε for all n N. 1

2 MATH 131A: REAL ANALYSIS (BIG IDEAS) Theorem 10 (R is complete). Every Cauchy sequence in R converges to a real number. Definition 11. Suppose that E R is a set of real numbers and let M, m R. We say that M is a least upper bound for E if (a) M is an upper bound for E and (b) if M is another upper bound for E, then M M. Similarly, we say that m is a greatest lower bound for E if (a) m is a lower bound for E and (b) if m is another lower bound for E then m m. Theorem 12 (Least upper bound property). Let E be a non-empty subset of R. If E has an upper bound, then it has a least upper bound. Proof. Since E is non-empty, there is some element a 1 E, and since E has an upper bound, we can choose some upper bound b 1. Clearly a 1 b 1. We will define two sequences (a i ) and (b i ) inductively such that a n b n for all n. For each n 1, suppose that we have constructed a n and b n such that a n b n. Let K = an+bn be the average of a 2 n and b n. Note that a n K b n. If K is an upper bound for E, let a n+1 = a n and let b n+1 = K = an+bn. Then b 2 n+1 a n+1 = an+bn a 2 n = bn an. 2 If K is not an upper bound for E, then there is some element x E such that x > K (otherwise K would be an upper bound). In that case, let a n+1 = x and let b n+1 = b n. Then b n+1 a n+1 = b n x < b n K = bn an. Note that in either case b 2 n+1 a n+1 bn an. 2 Also in either case a n is increasing and b n is decreasing. Since a n E for all n and b n is an upper bound for E for all n, we have a 1 a 2 a 3 b 3 b 2 b 1. Using the inequality b n+1 a n+1 bn an 2 one can show by induction that b n+1 a n+1 b 1 a 1. (1) 2 n We will show that the sequences (a n ) and (b n ) are Cauchy. Let ε > 0 be given. Choose N 1 such that 2 N > b 1 a 1. Let m, n N. Without loss of ε generality, assume that m n. Then a m a n, and we have a m a n = a m a n b n a n b 1 a 1 2 n < ε. Thus (a n ) is Cauchy. A very similar argument shows that (b n ) is Cauchy. Using (1) one can show that (a n ) and (b n ) are equivalent sequences. Therefore they have the same it, say L. I claim that L is the least upper bound of E. On the one hand, if x E then x b n for all n since every b n is an upper bound for E. Thus x L. Since x was an arbitrary element of E, we conclude that L is an upper bound for E. On the other hand, if L is any other upper bound, then L a n for all n since every a n is an element of E. Thus L L. It follows that L is the least upper bound of E. Proposition 13 (Bounded monotonic sequences are convergent). If a sequence is increasing and bounded above then it converges. Proof. Suppose that (a n ) is increasing and bounded above. Let E = {a n : n 1}. Then (a n ) has a least upper bound which we will call s = sup a n. We will show that a n = s. Let ε > 0 be given. Since s is the least upper bound, s ε is not an upper bound for E. So there exists some N 1 such that a N > s ε. Since (a n ) is increasing, it follows that a n a N > s ε for all n N. Thus, for all n N we have a n s = s a n s a N < s (s ε) = ε.

MATH 131A: REAL ANALYSIS (BIG IDEAS) 3 Therefore (a n ) converges to s. Definition 14. A subsequence of a sequence (a n ) is a sequence of the form (a nk ), where n k is a strictly increasing sequence of natural numbers. Theorem 15 (Bolzano-Weierstrass). Every bounded sequence has a convergent subsequence. Definition 16. Let (a n ) be a sequence. Then sup a n = N sup{a n : n > N}, inf a n = N inf{a n : n > N}. Definition 17. Let (a k ) be a sequence, and for each n 1 let s n = a k = a 1 + a 2 +... + a n be the nth partial sum. We say that a k converges if the sequence (s n ) is a convergent sequence. If a n is convergent, we say the series a n is absolutely convergent. Proposition 18 (Cauchy criterion for series). A series a n is convergent if and only if for each ε > 0 there exists a number N 1 such that a k < ε whenever n m N. k=m Corollary 19. If a n is convergent then a n = 0. Proof. Apply the Cauchy criterion with m = n to get: for every ε > 0 there exists a number N 1 such that a n < ε whenever n N. Proposition 20 (The comparison test). Let a n be a series of nonnegative terms. (1) If a n converges and b n a n for all n then b n converges. (2) If a n = + and b n a n for all n then b n diverges. Proof. (1) Suppose that a n converges. Then it satisfies the Cauchy criterion. Given ε > 0 there exists a number N 1 such that m m a k = a k < ε for all n m N (note a k 0 for all k). So for m n N we have m m m b k b k a k < ε, k=n k=n k=n k=n where we used the Triangle inequality in the first step. Thus b n satisfies the Cauchy criterion, and therefore it converges. (2) Let (s n ) be the sequence of partial sums for a n and let (t n ) be the sequence of partial sums for b n. Then t n s n for all n. Since s n = + we have t n = + too. k=n

4 MATH 131A: REAL ANALYSIS (BIG IDEAS) Proposition 21 (Alternating series test). If (a n ) is a decreasing sequence of nonnegative real numbers and a n = 0 then the alternating series ( 1) n+1 a n converges. Proof. Let (s n ) denote the sequence of partial sums. Let (e n ) be the subsequence of evenindexed terms and (o n ) the subseqeunce of odd-indexed terms. Then o n = s 2n 1 and e n = s 2n : The even-indexed terms are increasing since (s n ) = o 1, e 1, o 2, e 2, o 3, e 3,.... e n+1 e n = s 2n+2 s 2n = a 2n+2 + a 2n+1 0. Similarly, the odd-indexed terms are decreasing since o n+1 o n = s 2n+1 s 2n 1 = a 2n+1 a 2n 0. I claim that every even-indexed term is less than or equal to every odd-indexed term: e m o n for all m, n N. First, e n o n+1 o n because o n+1 e n = s 2n+1 s 2n = a 2n+1 0 and (o n ) is decreasing. Now, if m n, then because (e n ) is increasing, e m e n o n. On the other hand, if m n then o n o m e m. To summarize, e 1 e 2 e 3 o 3 o 2 o 1. Since (e n ) is increasing and bounded above, it converges, say to L e. Similarly, (o n ) converges, say to L o. But L o L e = o n+1 e n = (o n+1 e n ) = a 2n+1 = 0. Thus L o = L e. It follows that (s n ) converges. Definition 22. Let X R. A function f : X R is continuous at the point a X if, for every sequence (a n ) (with a n X) converging to a, the sequence (f(a n )) converges to f(a). Put another way, f is continuous at a if f(a n) = f for every sequence (a n ) that converges to a. ( a n Definition 23. Let X R. A function f : X R is continuous at the point a X if, for every ε > 0 there exists a δ > 0 such that f(x) f(a) < ε whenever x X and x a < δ. Theorem 24 (Extreme Value Theorem). Let f : [a, b] R be a continuous function. Then f is bounded (i.e. there exists a real number M such that f(x) M for all x [a, b]). Furthermore, f attains its maximum and minimum values on [a, b] (i.e. there exist x 0, y 0 [a, b] such that f(x 0 ) f(x) f(y 0 ) for all x [a, b]). Proof. We argue by contradiction. Suppose that f is not bounded on [a, b]. Then for each n N there exists a number x n [a, b] such that f(x n ) > n. So we get a sequence (x n ) which is bounded (it s always inside [a, b]). By the Bolzano-Weierstrass theorem, there exists a convergent subsequence (x nk ) of (x n ), say that x nk = L. Then L [a, b], so by assumption f is continuous at L. So we have a sequence (x nk ) which converges to L but the sequence (f(x nk )) doesn t converge to f(l) (since f(x nk ) + ). This contradicts the continuity of f, so we must have that f is bounded. )

MATH 131A: REAL ANALYSIS (BIG IDEAS) 5 Now let M := sup{f(x) : x [a, b]}. We just proved that M is a finite number. For each n N there exists a number y n [a, b] such that M 1 f(y n n) M (otherwise M wouldn t be the sup). By the squeeze theorem, f(y n ) = M. We aren t guaranteed that (y n ) is a convergent sequence, but it is bounded (always in [a, b]) so Bolzano-Weierstrass gives us a convergent subsequence (y nk ). Say y nk = y 0. Since f is continuous at y 0 we have f(y 0 ) = f( y nk ) = f(y nk ) = f(y n ) = M. (In the third equality we used that if a sequence is convergent, then every subsequence converges to the same it.) So f attains its maximum M. The assertion for the minimum is similar. Theorem 25 (Intermediate Value Theorem). Let I R be an interval and suppose that f : I R is continuous. Let a, b I such that a < b. Then for every y between f(a) and f(b) (i.e. either f(a) < y < f(b) or f(a) > y > f(b)) there exists at least one x (a, b) such that f(x) = y. Proof. Assume that f(a) < y < f(b) (the other case is similar). We will construct two sequences (a n ) and (b n ) which converge to x 0 such that f(a n ) y and f(b n ) y. Let S = {x [a, b] : f(x) < y}. Then S is not empty because a S. Since S is also bounded above (by b) it has a least upper bound, say x 0 = sup S. For each n N, the number x 0 1 is not an upper bound for S so there exists b n n S such that x 0 1 < b n n x 0. By the squeeze theorem, b n = x 0. Since f is continuous and since f(b n ) < y we have f(x 0 ) = f( b n ) = f(b n ) y. Now for each n N let a n = min{b, x 0 + 1 n }. Then for each n, a n / S so f(a n ) y. Again by the squeeze theorem, a n = x 0 so we have Thus f(x 0 ) = y. f(x 0 ) = f( a n ) = f(a n ) y. Definition 26. A function f : X R is uniformly continuous on X if for every ε > 0 there exists a δ > 0 such that f(x) f(y) < ε whenever x, y X and x y < δ. Proposition 27. If f : X R is uniformly continuous on X and if (a n ) is a Cauchy sequence then (f(a n )) is a Cauchy sequence. Proof. Let (a n ) be a Cauchy sequence and let ε > 0. Since f is uniformly continuous on X, there exists a δ > 0 such that f(x) f(y) < ε whenever x, y X and x y < δ. Since (a n ) is Cauchy, there exists and N 1 such that a m a n < δ for all m, n N. Thus, for such m, n we have so (f(a n )) is a Cauchy sequence. f(a m ) f(a n ) < ε Proposition 28. A function f : (a, b) R can be extended to a continuous function on [a, b] if and only if it is uniformly continuous on (a, b).

6 MATH 131A: REAL ANALYSIS (BIG IDEAS) Proof. ( ) Suppose that f can be extended to a continuous function f : [a, b] R. Then by Proposition?? f is uniformly continuous on [a, b]. It follows immediately that f is uniformly continuous on (a, b). ( ) Suppose that f is uniformly continuous on (a, b). We must decide how to define f(a) and f(b) to make f : [a, b] R continuous. It suffices to show how to do this for f(a) (since f(b) is similar). We define f(a) = f(a n ) for any sequence (a n ) in (a, b) converging to a. (2) Claim 1: if (a n ) is a sequence in (a, b) converging to a then the sequence (f(a n )) converges. To prove this, note that (a n ) is a Cauchy sequence, so by Proposition 27, (f(a n )) is Cauchy, so it converges. Claim 2: if (a n ) and (a n) both converge to a then f(a n ) = f(a n). To prove this, form the sequence (z n ) via (z n ) = (a 1, a 1, a 2, a 2, a 3, a 3,...). It should be clear that z n = a n = a n = a so by Claim 1, f(z n ) exists. But now (f(a n )) and (f(a n)) are subsequences of the convergent sequence (f(z n )) so they converge to the same it, i.e. f(a n ) = f(a n). Therefore (2) makes sense, and it is clear by construction that f is continuous at a. Definition 29. Suppose that X contains the interval (c, b) for some c < a < b. We say that f(x) = L if x a f(a n) = L for every sequence (a n ) in (c, a) (a, b) converging to a. Definition 30. Let f be a function defined on the set (c, a) (a, b) for some c < a < b. Then x a f(x) = L if for every ε > 0 there exists a δ > 0 such that f(x) L < ε whenever 0 < x a < δ. Definition 31. Let f : X R be a function and let a X such that there is an interval (c, b) X containing a. We say that f is differentiable at a if the it f(x) f(a) exists and is finite; we call this it the derivative of f at a, and write it f (a). Proposition 32 (Differentiability implies continuity). Suppose that f is defined on (c, b) for some c < a < b. If f is differentiable at a then f is continuous at a. Proof. Recall that when thinking about the it of a function at a we consider points x close to a but not equal to a. So we are justified in writing [ ] f(x) f(a) f(x) f(a) f(x) = (x a) + f(a) = f(a) + (x a). x a x a x a Using our assumption that f is differentiable at a, the it of the quotient on the right-hand side above exists and is finite. So we have f(a) (x a)f(x) = x a (x a) x a f(x) f(a) x a This shows that x a f(x) = f(a), as desired. = 0 [something finite] = 0.

MATH 131A: REAL ANALYSIS (BIG IDEAS) 7 Proposition 33 (Product rule). Let f and g be functions that are differentiable at a. Then fg is differentiable at a and (fg) (a) = f(a)g (a) + f (a)g(a). Proof. Adding zero in a clever way, we compute as desired. f(x)g(x) f(a)g(a) f(x)g(x) f(x)g(a) + f(x)g(a) f(a)g(a) = g(x) g(a) f(x) f(a) = f(x) + g(a) = f(a)g (a) + f (a)g(a), Theorem 34 (Rolle s Theorem). Suppose that f is continuous on [a, b] and differentiable on (a, b) and that f(a) = f(b). Then there exists a point c (a, b) such that f (c) = 0. Proof. By the Extreme Value Theorem, since f is continuous on the closed interval [a, b] it attains its maximum and minimum on [a, b]. So there exist y, z [a, b] such that f(y) f(x) f(z) for all x [a, b]. Suppose that z is an interior point, i.e. z (a, b). Let (z n ) be a sequence in (a, z) (z, b) converging to z and let (z + n ) and (z n ) be the subsequences of (z n ) such that z + n > z and z n < z. Then (since f(z) is a maximum) f(z n + ) f(z) z n + z 0 and f(z n ) f(z) z n z (Both its exist because f is differentiable at z.) sequence converge to the same it, we see that f(z n ) f(z) z n z f(z n ) f(z) 0 and z n z 0. Since subsequences of a convergent It follows that f (z) = 0. By the same reasoning, we have f (y) = 0 if y (a, b). The only case left is if y and z are both endpoints of [a, b]. But since f(a) = f(b) this forces f(y) = f(x) = f(z) for all x [a, b], so f (x) = 0 for all x (a, b). Theorem 35 (The Mean Value Theorem). Suppose that f is continuous on [a, b] and differentiable on (a, b). Then there exists a c (a, b) such that f f(b) f(a) (c) =. b a Proof. Define ( ) f(b) f(a) g(x) = f(x) (x b) + f(b). b a The function in parentheses is just the equation of the secant line connecting (a, f(a)) and (b, f(b)). We compute g(a) = 0 and g(b) = 0. Clearly g is continuous on [a, b] and differentiable on (a, b), so Rolle s theorem applies. So there is some point c (a, b) such that g (c) = 0. On the other hand, This proves the theorem. g (c) = f (c) f(b) f(a). b a 0.

8 MATH 131A: REAL ANALYSIS (BIG IDEAS) Definition 36. Suppose that f is bounded on [a, b]. A partition P of [a, b] is a finite ordered set P = {a = x 0 < x 1 < x 2 <... < x n = b}. The intervals [x k 1, x k ] are called subintervals of P. For each subinterval, let m k = inf{f(x) : x [x k 1, x k ]}, M k = sup{f(x) : x [x k 1, x k ]}. The lower Darboux sum L(f, P ) of f with respect to P is L(f, P ) = m k (x k x k 1 ) and the upper Darboux sum U(f, P ) of f with respect to P is U(f, P ) = M k (x k x k 1 ). We define L(f) = sup{l(f, P ) : P is a partition of [a, b]} and U(f) = inf{u(f, P ) : P is a partition of [a, b]}, and we say that f is integrable on [a, b] if L(f) = U(f). In that case we write b a f = L(f) = U(f). Proposition 37 (Cauchy criterion for integrals). Suppose that f is bounded on [a, b]. Then f is integrable if and only if for every ε > 0 there exists a partition P of [a, b] such that U(f, P ) L(f, P ) < ε. Proof. ( ) Suppose that such a partition P exists for each ε > 0. Then U(f) L(f) U(f, P ) L(f, P ) < ε. Since ε is arbitrary, this implies that U(f) = L(f). ( ) Now suppose that f is integrable. Let ε > 0 be given. Since U(f) is the greatest lower bound of the U(f, P ) over all partitions P, the number U(f) + ε is no longer a lower 2 bound, so there exists a partition P 1 such that U(f, P 1 ) < U(f) + ε 2. Similarly, there exists a partition P 2 such that L(f, P 2 ) > L(f) ε 2. Let P = P 1 P 2. Since integrability means that U(f) = L(f), we have This completes the proof. U(f, P ) L(f, P ) U(f, P 1 ) L(f, P 2 ) = U(f, P 1 ) U(f) + L(f) L(f, P 2 ) < ε 2 + ε 2 = ε.

MATH 131A: REAL ANALYSIS (BIG IDEAS) 9 Proposition 38. Every continuous function on [a, b] is integrable. Proof. First, we observe that because f is continuous on the closed interval [a, b] it is uniformly continuous on [a, b]. So, given ε > 0 there exists a δ > 0 such that f(x) f(y) < ε whenever x y < δ. b a Let P be a partition of [a, b] such that x k x k 1 < δ for every subinterval. By the extreme value theorem applied to [x k 1, x k ], the function f attains its maximum M k and its minimum m k for some points z k, y k in the interval [x k 1, x k ]. And whatever those points z k, y k are, they are separated by a distance of at most δ. So M k m k = f(z k ) f(y k ) < ε b a. It follows that U(f, P ) L(f, P ) = (M k m k )(x k x k 1 ) < ε (x b a k x k 1 ) = ε since the last sum is telescoping and equals b a. By the Cauchy criterion, f is integrable on [a, b]. Theorem 39 (The Fundamental Theorem of Calculus I). If g is continuous on [a, b] and differentiable on (a, b), and if g is integrable on (a, b) then b a g = g(b) g(a). Proof. Let P be a partition of [a, b]. For each k, apply the Mean Value Theorem to the interval [x k 1, x k ] to find a point c k (x k 1, x k ) such that (x k x k 1 )g (c k ) = g(x k ) g(x k 1 ). Then g(b) g(a) = (g(x k ) g(x k 1 )) = g (c k )(x k x k 1 ). For each k we clearly have It follows that inf{g (x) : x [x k 1, x k ]} g (c k ) sup{g (x) : x [x k 1, x k ]}. L(g, P ) g(b) g(a) U(g, P ). Since this inequality holds for every partition P (and the middle quantity does not depend on P ), we have L(g ) g(b) g(a) U(g ). But U(g ) = L(g ) since g is integrable, so b a g = U(g ) = g(b) g(a). Theorem 40 (The Fundamental Theorem of Calculus II). Suppose that f is integrable on [a, b] and define F (x) = a f(t) dt. Then F is continuous on [a, b]. If f is continuous at c (a, b) then F is differentiable at c and F (c) = f(c).

10 MATH 131A: REAL ANALYSIS (BIG IDEAS) Proof. Let M be such that g(x) M for all x [a, b]. For x, y [a, b] we have y y F (y) F (x) = f(t) dt f(t) dt = y f(t) dt f(t) dt M x y. a Therefore F is Lipschitz, so it is uniformly continuous on [a, b]. Suppose that f is continuous at c (a, b). Observe that and so a f(c) = 1 F (x) F (c) c = 1 x f(c) dt c f(t) dt, F (x) F (c) f(c) = 1 (f(t) f(c)) dt. c Let ε > 0 be given. Since f is continuous at c, there exists a δ > 0 such that f(t) f(c) < ε whenever t c < δ. For the integral above we only care about values of t between c and x, and if < δ then certainly t c < δ. For such x it follows that F (x) F (c) Therefore F (c) = f(c). f(c) 1 c f(t) f(c) dt < 1 x c ε dt = ε.