Solution of the 8 th Homework

Similar documents
Solution of the 7 th Homework

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

Continuity. Chapter 4

Continuity. Chapter 4

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 131A: REAL ANALYSIS (BIG IDEAS)

Principles of Real Analysis I Fall VII. Sequences of Functions

MATH 1A, Complete Lecture Notes. Fedor Duzhin

Functional Analysis HW #3

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

,... We would like to compare this with the sequence y n = 1 n

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.

Functional Analysis HW #1

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

Week 2: Sequences and Series

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

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

Math 104: Homework 7 solutions

Assignment-10. (Due 11/21) Solution: Any continuous function on a compact set is uniformly continuous.

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.

HOMEWORK ASSIGNMENT 6

MATH 301 INTRO TO ANALYSIS FALL 2016

Analysis II - few selective results

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

MATH 117 LECTURE NOTES

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

M17 MAT25-21 HOMEWORK 6

Module 7 : Applications of Integration - I. Lecture 20 : Definition of the power function and logarithmic function with positive base [Section 20.

ALGEBRAIC GEOMETRY COURSE NOTES, LECTURE 2: HILBERT S NULLSTELLENSATZ.

A LITTLE REAL ANALYSIS AND TOPOLOGY

Undergraduate Notes in Mathematics. Arkansas Tech University Department of Mathematics

Thus f is continuous at x 0. Matthew Straughn Math 402 Homework 6

Limits and Continuity

Definitions & Theorems

MA103 Introduction to Abstract Mathematics Second part, Analysis and Algebra

We are now going to go back to the concept of sequences, and look at some properties of sequences in R

Math 312 Spring08 Day 13

2.1 Convergence of Sequences

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

6.2 Deeper Properties of Continuous Functions

Math LM (24543) Lectures 02

Lecture 4 Lebesgue spaces and inequalities

Functions. Chapter Continuous Functions

Econ Slides from Lecture 1

Notes for Math 290 using Introduction to Mathematical Proofs by Charles E. Roberts, Jr.

Solution. 1 Solution of Homework 7. Sangchul Lee. March 22, Problem 1.1

CONSEQUENCES OF POWER SERIES REPRESENTATION

Sets, Functions and Metric Spaces

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

Sequences of Real Numbers

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

APPLICATIONS OF DIFFERENTIABILITY IN R n.

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

Analysis Finite and Infinite Sets The Real Numbers The Cantor Set

A lower bound for X is an element z F such that

Math 421, Homework #6 Solutions. (1) Let E R n Show that = (E c ) o, i.e. the complement of the closure is the interior of the complement.

COT 2104 Homework Assignment 1 (Answers)

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

Static Problem Set 2 Solutions

Chapter 2. Metric Spaces. 2.1 Metric Spaces

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.

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

FUNCTIONS OF A REAL VARIABLE

5.5 Deeper Properties of Continuous Functions

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.

Exam 2 extra practice problems

Notes on Integrable Functions and the Riesz Representation Theorem Math 8445, Winter 06, Professor J. Segert. f(x) = f + (x) + f (x).

(a) For an accumulation point a of S, the number l is the limit of f(x) as x approaches a, or lim x a f(x) = l, iff

A lower bound for X is an element z F such that

Notes on uniform convergence

Most Continuous Functions are Nowhere Differentiable

1 Lecture 25: Extreme values

Scalar multiplication and addition of sequences 9

Spring 2014 Advanced Probability Overview. Lecture Notes Set 1: Course Overview, σ-fields, and Measures

From now on, we will represent a metric space with (X, d). Here are some examples: i=1 (x i y i ) p ) 1 p, p 1.

2.3 Some Properties of Continuous Functions

L p Spaces and Convexity

Calculus (Real Analysis I)

MA651 Topology. Lecture 9. Compactness 2.

EXTINGUISHING THE DISTINGUISHED LOGARITHM PROBLEMS. University of California Irvine, University of California Irvine, and Auburn University

Solutions Manual for Homework Sets Math 401. Dr Vignon S. Oussa

Optimization and Optimal Control in Banach Spaces

Course 212: Academic Year Section 1: Metric Spaces

Sequences. Limits of Sequences. Definition. A real-valued sequence s is any function s : N R.

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

Math 430/530: Advanced Calculus I Chapter 1: Sequences

Immerse Metric Space Homework

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.

A CONTINUOUS AND NOWHERE DIFFERENTIABLE FUNCTION, FOR MATH 320

Partial differential equation for temperature u(x, t) in a heat conducting insulated rod along the x-axis is given by the Heat equation:

Chapter 3 Continuous Functions

a n b n ) n N is convergent with b n is convergent.

Chapter 2 Metric Spaces

2. Prime and Maximal Ideals

CHAPTER 6. Limits of Functions. 1. Basic Definitions

We saw in the previous lectures that continuity and differentiability help to understand some aspects of a

Chapter 8. P-adic numbers. 8.1 Absolute values

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 Sequences, Continuity, and Limits

Transcription:

Solution of the 8 th Homework Sangchul Lee December 8, 2014 1 Preinary 1.1 A simple remark on continuity The following is a very simple and trivial observation. But still this saves a lot of words in actual proofs. Lemma 1.1. Let f : X R be a function, and let x 0 X. Then the followings are equivalent: (i) f is continuous at x 0. I.e, x x0 ;x X f (x) = f (x 0 ). (ii) x x0 ; f (x) = f (x 0 ). Proof. This follows by just a few seconds of meditation. To be slightly precise, this follows because f (x 0 ) f (x 0 ) < ε for any ε > 0. 1.2 (optional) Weierstrass-Carathéodory formulation of differentiability Proposition 1.2. Let X R, let x 0 X be a it point of X, and let f : X R be a function. Then the followings are equivalent: (i) f is differentiable at x 0. (ii) There exists a function φ f : X R which is continuous at x 0 and satisfies Moreover, if these hold, then φ f (x 0 ) = f (x 0 ). f (x) = f (x 0 ) + φ f (x)( ) (1.1) It is loosely read as differentiability at x 0 holds if and only if the graph of a function is well approximated by a line (with finite slope) near x 0. What is good for this formulation is that various differentiation rules follows almost for free. Proof. (i) = (ii) : Define φ f : X R by Now, by the differentiability of f at x 0, we have φ f (x) = f (x) f (x 0 ) x x 0, x x 0 f (x 0 ), x = x 0. φ f (x) = = f (x 0 ) = φ f (x 0 ). This proves that φ f is continuous at x 0. Moreover, (1.1) holds for all x X by definition of φ f. ((1.1) is clearly true when x x 0. When x = x 0, both sides of (1.1) are zero.) Therefore (ii) follows. 1

(ii) = (i) : For x X \ {x 0 }, solving (1.1) in terms of φ f (x) gives φ f (x) = f (x) f (x 0). Taking it as x x 0, by the continuity of φ f at x 0, Therefore f is differentiable at x 0 and f (x 0 ) = φ f (x 0 ). = φ f (x) = φ f (x 0 ). 2 Solutions Exercise 1. Let X be a subset of R and let f : X R be a function. Then the following two statements are equivalent. (i) f is uniformly continuous on X. (ii) For any two equivalent sequences (a n ) n=m, (b n ) n=m, the sequences ( f (a n )) n=m, ( f (b n )) n=m are also equivalent sequences. Solution. (i) = (ii) : Let (a n ) n=m, (b n ) n=m be sequences in X which are equivalent. Then for any ε > 0, By the uniform convergence of f, there exists δ > 0 such that whenever x, y X and x y < δ we have f (x) f (y) < ε. By the equivalence, there exists N N such that whenever n N we have a n b n < δ. Combining two statements, it follows that whenever n N we have f (a n ) f (b n ) < ε. This proves that ( f (a n )) n=m, ( f (b n )) n=m are also equivalent sequences. (ii) = (i) : We prove contrapositive. Let us assume that f is not uniformly continuous. By negating Definition 3.31, There exists ε > 0 such that for any δ > 0 there exists x, y X such that x y < δ but f (x) f (y) ε. Now for each choice δ = n 1 (with n {1,2, }) pick such two elements a n, b n X (that is, a n b n < δ = n 1 but f (a n ) f (b n ) ε). Then (a n ) n=1, (b n) n=1 are equivalent but ( f (a n)) n=m, ( f (b n )) n=m cannot be equivalent. This proves the contrapositive as desired. Exercise 2. Give an example of a continuous function f : R (0, ) such that, for any real number 0 < ε < 1, there exists x R such that f (x) = ε. Remark. In view of the Maximum Principle, a continuous function f : R (0, ) must attain positive minimum on any finite closed interval [a, b] R. Consequently, if ε > 0 is small, any solution of f (x) = ε have large size. This in particular suggests that any continuous function f : R (0, ) satisfying x f (x) = 0 serves an example. 1 st Solution. Let f (x) = 2 x. Then clearly f is a continuous function with range (0, ). Moreover, for any 0 < ε < 1 we have f (log 2 ε) = ε. Therefore f satisfies all the desired properties. If you raise an objection by claiming that we have never learned both exponential function and logarithm in this course, here is another solution: 2 nd Solution. Let f : R (0, ) by f (x) = 1 1 + x. 2

Since f is defined by applying arithmetic operations on continuous functions such that denominator never vanishes, f defines a continuous function on R. Also, whenever 0 < ε < 1 we have ε 1 1 > 0 and hence f (ε 1 1 1) = 1 + (ε 1 1) = ε. Therefore f satisfies all the desired property. Exercise 3. Let X be a subset of R and let f : X R be a uniformly continuous function. Let x 0 be an adherent point of X. Then x x0 f (x) exists (and so it is a real number.) Solution. We divide the proof into two steps. Each step contains some important ideas which deserve their own right. Step 1. We prove the following lemma: Lemma. Let X be a subset of R and let f : X R be a uniformly continuous function. Then for every Cauchy sequence (a n ) n=m in X, the sequence ( f (a n )) n=m is also a Cauchy sequence. Proof of Lemma. Let ε > 0. Then There exists δ > 0 such that whenever x, y X and x y < δ we have f (x) f (y) < ε. There exists N m such that whenever j, k N we have a j a k < δ. Combining two statements, we have f (a j ) f (a k ) < ε whenever j, k N. This proves that ( f (a n )) n=m as desired. Remark. Before moving to the next step, we remark that this Lemma does not characterize uniform continuity. Indeed, f (x) = x 2 is not uniformly continuous, yet still mapping Cauchy sequences to Cauchy sequences. In fact, it is equivalent to the condition that f : X R is uniformly continuous on any bounded subsets of X. Step 2. To prove the original problem, we utilize Proposition 2.16. Indeed, let (a n ) n=0 and (b n) n=0 be any sequences in X that converge to x 0. Since ( f (a n )) n=0 and ( f (b n)) n=0 are Cauchy sequences by the Lemma above, they are also convergent. Let L = f (b n ). n Note that (a n ) n=0 and (b n) n=0 are equivalent, since we have n (a n b n ) = x 0 x 0 = 0. Then by Proposition 3.37 (Exercise 8.1), ( f (a n )) n=0 and ( f (b n)) n=0 are also equivalent. Thus f (a n) = ( f (a n ) f (b n )) + f (b n ) = 0 + L = L. n n n Therefore by Proposition 2.16, f converges to L at x 0. Exercise 4. Let X be a subset of R, and let f : X R be a uniformly continuous function. Assume that E is a bounded subset of X. Then f (E) is also bounded. 1 st Solution. We first present a proof by contradiction. Assume that f (E) is unbounded. Then for each n N, there exists a n E such that f (a n ) n. But since E is bounded, (a n ) n=0 is also bounded. Thus by the Bolzano-Weierstrass Theorem, there exists a subsequence (a j(n)) n=0 of (a n ) n=0 which is convergent. Let x 0 = n a j(n) denotes the it. Then by Lemma 2.13, x 0 is an adherent point of X. Thus by Corollary 3.40 (Exercise 8.3), we know that n f (a j(n) ) exists. But since f (a j(n) ) j(n) n, by choosing a sufficiently large n we get a contradiction. (Just pick n to be larger than a bound of ( f (a j(n) )) n=0.) 3

2 nd Solution. We present a direct proof. To make proof simple, we introduce some notations: Pick an arbitrary ε > 0. Then by the uniform continuity of f, we can choose δ > 0 such that whenever x, y X and x y < δ we have f (x) f (y) < ε. Let K Z be defined by K = {k Z : E [kδ, (k + 1)δ) }. Since E is bounded, K is a finite set. (Indeed, choose a bound M > 0 of E, and choose N N such that N δ > M. Then K { N,, N}.) Now we are ready. For each k K, pick an element x k E [kδ, (k + 1)δ). Since K is finite, L := ε + max{ f (x k ) : k K} is a well-defined positive real number. Moreover, for any x E, there exists k K such that x E [kδ, (k + 1)δ). Then x x k < δ and hence f (x) f (x k ) < ε. This shows that This proves that f (E) is also bounded. f (x) f (x) f (x k ) + f (x k ) ε + max{ f (x k ) : k K} = L. Exercise 5. Let X be a subset of R, let x 0 be a it point of X, and let f : X R be a function. If f is differentiable at x 0, then f is also continuous at x 0. Solution. We have ( ) = ( = ( ) f (x) f (x0 ) ( ) ) ( Therefore, by adding f (x 0 ) to both sides, it follows that f is continuous at x 0. ) ( ) = f (x 0 ) 0 = 0. Exercise 6. Let X be a subset of R, let x 0 be a it point of X, let f : X R be a function, and let L be a real number. Then the following two statements are equivalent. (i) f is differentiable at x 0 on X with derivative L. (ii) For every ε > 0, there exists a δ = δ(ε) > 0 such that, if x X satisfies < δ, then f (x) [ f (x 0 ) + L( )] ε. Solution. The proof is almost a tautology. Nevertheless we spell out every detail. (i) = (ii) : By definition, for every ε > 0, there exists δ = δ(ε) > 0 such that, if x X \ {x 0 } satisfies < δ, then L < ε. Now multiply both sides by. Then we have Since this continues to hold when x = x 0, we get (ii). f (x) [ f (x 0 ) + L( )] ε. (ii) = (i) : For every ε > 0, pick δ = δ(ε/2) > 0 as in (ii). Then whenever x X \ {x 0 } and < δ, we have f (x) [ f (x 0 ) + L( )] 1 2 ε. 4

Now divide both sides by. (This is possible because is positive.) Then we get L 1 x x 2 ε < ε. 0 This yields x x0 ;( )/( ) = L, which is equivalent to (i). Exercise 7. Let X be a subset of R, let x 0 be a it point of X, and let f : X R and g : X R be functions. (i) If f is constant, so that there exists c R such that f (x) = c, then f is differentiable at x 0 and f (x 0 ) = 0. (ii) If f is the identity function, so that f (x) = x, then f is differentiable at x 0 and f (x 0 ) = 1. (iii) If f, g are differentiable at x 0, then f + g is differentiable at x 0, and ( f + g) (x 0 ) = f (x 0 ) + g (x 0 ). (Sum Rule) (iv) If f, g are differentiable at x 0, then f g is differentiable at x 0, and ( f g) (x 0 ) = f (x 0 )g(x 0 ) + g (x 0 ) f (x 0 ). (Product Rule) (v) If f is differentiable at x 0, and if c R, then c f is differentiable at x 0, and (c f ) (x 0 ) = c f (x 0 ). (vi) If f, g are differentiable at x 0, then f g is differentiable at x 0, and ( f g) (x 0 ) = f (x 0 ) g (x 0 ). (vii) If g is differentiable at x 0, and if g(x) 0 for all x X, then 1/g is differentiable at x 0, and (1/g) (x 0 ) = g (x 0 )/(g(x 0 )) 2. (viii) If f, g are differentiable at x 0, and if g(x) 0 for all x X, then f /g is differentiable at x 0, and ( f /g) (x 0 ) = g(x 0) f (x 0 ) f (x 0 )g (x 0 ) (g(x 0 )) 2. (Quotient Rule) 1 st Solution. (i) Notice that = 0, x X \ {x 0 }. Taking it as x x 0 for x X \ {x 0 }, it follows that f is differentiable at x 0 and f (x 0 ) = 0. (ii) Notice that = 1, x X \ {x 0 }. Taking it as x x 0 for x X \ {x 0 }, it follows that f is differentiable at x 0 and f (x 0 ) = 1. (iii) Notice that [ f (x) + g(x)] [ f (x 0 ) + g(x 0 )] = f (x) f (x 0) + g(x) g(x 0), x X \ {x 0 }. Taking it as x x 0 for x X \ {x 0 }, we know from both the assumption and the Limit Laws that the right-hand side converges to f (x 0 ) + g (x 0 ). This proves the Sum Rule. (iv) Notice that f (x)g(x) f (x 0 )g(x 0 ) = f (x) f (x 0) g(x 0 ) + g(x) g(x 0) f (x), x X \ {x 0 }. Taking it as x x 0 for x X \ {x 0 }, we know from both the assumption and the Limit Laws that the right-hand side converges to f (x 0 )g(x 0 ) + g (x 0 ) f (x 0 ). This proves the Product Rule. (v) This follows by (i) and (iv). (vi) This follows by (iii) and (v), with the choice c = 1. (vii) Notice that [1/g(x)] [1/g(x 0 )] = g(x) g(x 0) 1 g(x)g(x 0 ), x X \ {x 0}. 5

Taking it as x x 0 for x X \ {x 0 }, we know from both the assumption and the Limit Laws that the right-hand side converges to g (x 0 )/g(x 0 ) 2. This proves (vii). (viii) This follows from (iv) and (vii). 2 nd Solution. We utilize the Weierstrass-Carathéodory formulation, Proposition 1.2. (i) Notice that (1.1) is satisfied with φ f (x) = 0. Therefore f is differentiable at x 0 with f (x 0 ) = φ f (x 0 ) = 0. (ii) Notice that (1.1) is satisfied with φ f (x) = 1. Therefore f is differentiable at x 0 with f (x 0 ) = φ f (x 0 ) = 1. (iii) Let φ f and φ g be as in (1.1) for f and g, respectively. Then f (x) + g(x) = [ f (x 0 ) + φ f (x)( )] + [g(x 0 ) + φ g (x)( )] = [ f (x 0 ) + g(x 0 )] + [φ f (x) + φ g (x)]( ), and hence φ(x) = φ f (x) + φ g (x) satisfies Proposition 1.2. This shows that f + g is differentiable at x 0 and ( f + g) (x 0 ) = φ(x 0 ) = φ f (x 0 ) + φ g (x 0 ) = f (x 0 ) + g (x 0 ). (iv) Let φ f and φ g be as in (1.1) for f and g, respectively. Then and hence f (x)g(x) = [ f (x 0 ) + φ f (x)( )][g(x 0 ) + φ g (x)( )] = f (x 0 )g(x 0 ) + [φ f (x)g(x 0 ) + φ g (x) f (x 0 ) + φ f (x)φ g (x)( )]( ), φ(x) = φ f (x)g(x 0 ) + φ g (x) f (x 0 ) + φ f (x)φ g (x)( ) satisfies Proposition 1.2. This shows that f g is differentiable at x 0 and (v) This follows by (i) and (iv). ( f g) (x 0 ) = φ(x 0 ) = φ f (x 0 )g(x 0 ) + φ g (x 0 ) f (x 0 ) + 0 = f (x 0 )g(x 0 ) + g (x 0 ) f (x 0 ). (vi) This follows by (iii) and (v), with the choice c = 1. (vii) Let φ g be as in (1.1) for g. Then ( 1 g(x) = 1 g(x 0 ) 1 g(x 0 ) 1 ) 1 = g(x) g(x 0 ) φ g (x) g(x)g(x 0 ) (), and hence φ(x) = φ g (x)/g(x)g(x 0 ) satisfies Proposition 1.2. This shows that 1/g is differentiable at x 0 and (1/g) (x 0 ) = φ(x 0 ) = φ g (x 0 ) g(x 0 ) 2 = g (x 0 ) g(x 0 ) 2. (viii) This follows from (iv) and (vii). Exercise 8. Let X, Y be subsets of R, let x 0 X be a it point of X, and let y 0 Y be a it point of Y. Let f : X Y be a function such that f (x 0 ) = y 0 and such that f is differentiable at x 0. Let g : Y R be a function that is differentiable at y 0. Then the function g f : X R is differentiable at x 0, and (g f ) (x 0 ) = g (y 0 ) f (x 0 ). Remark. This would have followed easily it it were true that g( f (x)) g( f (x 0 )) = g( f (x)) g( f (x 0)). Unfortunately, this is not always true as may vanish infinitely many times near x 0. We need to circumvent this technical issue. 6

1 st Solution. Let (a n ) n=0 be any sequence in X \ {x 0} which converges to x 0. Then we deduce from Exercise 8.5 that ( f (a n )) n=0 converges to f (x 0). Now we divide into two cases: Case 1) Assume that f (x 0 ) 0. Then for the choice ε = 1 2 f (x 0 ) > 0, there exists δ > 0 such that, whenever x X and < δ we have Then it follows from the reverse triangle inequality that f (x 0 )( ) ε. f (x 0 )( ) f (x 0 )( ) f (x 0 ) ε = ε since f (x 0 ) = 2ε. In particular, since 0, this implies that > 0 whenever < δ. Next, we pick N sufficiently large so that a n x 0 < δ whenever n N. Since a n X \ {x 0 }, it follows that f (a n ) f (x 0 ) > 0. Then our intuitive idea works and we have ( ) g( f (a n )) g( f (x 0 )) g( f (an )) g( f (x 0 )) f (a n ) f (x 0 ) = n a n x 0 n f (a n ) f (x 0 ) a n x 0 ( = n = g (y 0 ) f (x 0 ). g( f (a n )) g( f (x 0 )) f (a n ) f (x 0 ) ) ( n f (a n ) f (x 0 ) a n x 0 Case 2) Now assume that f (x 0 ) = 0. For ε = 42, pick δ > 0 such that, if y Y and y y 0 < δ then Then by the triangle inequality, we have g(y) g(y 0 ) g (y 0 )(y y 0 ) 42 y y 0. g(y) g(y 0 ) g (y 0 ) y y 0 + g(y) g(y 0 ) g (y 0 )(y y 0 ) ( g (y 0 ) + 42) y y 0. Also, choose N sufficiently large so that whenever n N, we have f (a n ) f (x 0 ) < δ. Then by noting that g( f (a n )) g( f (x 0 )) ( g (y 0 ) + 42) f (a n ) f (x 0 ), a n x 0 a n x 0 it follows from the squeezing theorem, together with n [ f (a n ) f (x 0 )]/(a n x 0 ) = f (x 0 ) = 0, we have g( f (a n )) g( f (x 0 )) = 0 = g (y 0 ) f (x 0 ). n a n x 0 Therefore, in any cases the Chain Rule follows. 2 nd Solution. Let φ f : X R and φ g : Y R, respectively, be as in the Weierstrass-Carathéodory formulation (Proposition 1.2) for f and g, respectively. Then g( f (x)) = g( f (x 0 )) + φ g ( f (x))[ ] = g( f (x 0 )) + φ g ( f (x))φ f (x)( ). Thus the function φ(x) = φ g ( f (x))φ f (x) is continuous at x 0 and satisfies (1.1). Thus it satisfies Proposition 1.2. This shows that g f is differentiable at x 0 and (g f ) (x 0 ) = φ(x 0 ) = φ g ( f (x 0 ))φ f (x 0 ) = g (y 0 ) f (x 0 ) ) Exercise 9. Let a < b be real numbers, and let f : (a, b) R be a function. If x 0 (a, b), if f is differentiable at x 0, and if f attains a local maximum or minimum at x 0, then f (x 0 ) = 0. Solution. We first consider local maximum case. Since x 0 is a local maximum point of f, there exists a sufficiently small 7

δ > 0 such that f attains a minimum on (x 0 δ, x 0 + δ) (a, b). Thenwe have 0, x 0 δ < x < x 0 0, x 0 < x < x 0 + δ. Thus taking right-it and left-it, we find that f (x 0 ) = 0 and f (x 0 ) = 0. + x x0 This proves that f (x 0 ) = 0. Local minimum case can be tackled in the same way. Exercise 10. Let a < b be real numbers, and let f : [a, b] R be a continuous function which is differentiable on (a, b). Assume that f (a) = f (b). Then there exists x (a, b) such that f (x) = 0. Solution. Utilizing the Maximum Principle, pick two points x 0, x 1 [a, b] such that f attains a global maximum at x 0 and a global minimum at x 1. If f (x 0 ) = f (x 1 ), then f must reduce to a constant function and hence the claim follows from Exercise 8.7.(i). (In this case, you can pick any point x (a, b).) In view of the previous observation, we may ssume f (x 0 ) f (x 1 ). This implies that either f (x 0 ) f (a) or f (x 1 ) f (a). In either cases, there exists a global extremum x of f, which is neither a nor b. Thus x is also a global extremum of f (a,b). Then x is a local extremum of f (a,b) and by the previous exercise, we have f (x) = 0. Exercise 11. Let X be a subset of R, let x 0 be a it point of X, and let f : X R be a function. If f is monotone increasing and if f is differentiable at x 0, then f (x 0 ) 0. If f is monotone decreasing and if f is differentiable at x 0, then f (x 0 ) 0. Solution. We first assume that f is monotone. If x X \ {x 0 }, then by dividing the cases based on whether x > x 0 or x < x 0, we find that 0 always holds. Now assume further that f is differentiable at x 0. Taking x x 0 in X \ {x 0 }, the inequality is preserved and hence we have f (x 0 ) = 0. For the monotone-decreasing case, this argument works exactly in the same way (but with inequalities reversed) to yield the corresponding proof. Exercise 12. Let a < b be real numbers, and let f : [a, b] R be a differentiable function. If f (x) > 0 for all x [a, b], then f is strictly monotone increasing. If f (x) < 0 for all x [a, b], then f is strictly monotone decreasing. If f (x) = 0 for all x [a, b], then f is a constant function. Solution. All these statements follows from the Mean Value Theorem in a direct manner. For example, assume first that f (x) > 0 for all x [a, b]. Then for any x < y in [a, b], the Mean Value Theorem says that there exists c [x, y] such that f (y) f (x) = f (c) > 0, y x 8

where the second inequality follows form the assumption. Multiplying y x to both sides, we get f (y) f (x) > 0, which implies that f is strictly increasing. The other two assertions are also proved in a similar manner. 9