Vaughan's Lecture Notes on Real Analysis

76
1. 01/18/11: Building Up to the Reals, Part One 1.1. Symbols. We will start with a list of mathematical symbols that we will use throughout the course. : for all : there exists or : implies ⇐⇒ : if and only if : belongs to {x | condition}: set notation : intersection : union q: disjoint union, or if A = B q C, then A = B C and B C = S \ A: set difference, or {x S | x/ A} An example of a formula written (almost) completely in symbols: (> 0)(δ> 0) s.t. (|x - y| ⇒|f (x) - f (y)| <) 1.2. The Natural Numbers, or N. The main idea for the first part of the course will be R, or the real numbers. There are three main ways to prove their existence: decimal expansions, Dedekind cuts, or sequences of rationals. We will eventually use a hybrid of the second and third methods, but that is for later. For now... N = {1, 2, 3, 4, 5, 6,... } * . This is a commutative semiring with an identity (built off 1 with operators + and ·). There is a notion of order, or x>y, contained in the algebraic structure of N. Axiom (Induction). If S N and: (1) 1 S (2) x S (x + 1) S hold, then S = N. In other words, if the two conditions above are true, then nothing can fail to be in S. Note that we will say 0 / N. 1.3. The Integers, or Z. Z = N ∪{0}∪ (-N). This is a commutative ring with multiplicative identity, meaning that it is an abelian group under (+, 0) and is associative, commutative, and distributive (i.e. x(y + z)= xy + xz) under multiplication. We also define 0x =0 x. Note that a group G is a set with an operation ? such that g?h G for g,h G, along with these properties: (1) G has an identity e, for which the following holds: e?g = g?e = g g. (2) (g?h) ?k = g? (h?k) (3) (g G)(h G) such that g?h = e = h?g. * Read the book for notes on their structure. 1

description

Lecture notes on real analysis.

Transcript of Vaughan's Lecture Notes on Real Analysis

  • 1. 01/18/11: Building Up to the Reals, Part One

    1.1. Symbols.

    We will start with a list of mathematical symbols that we will use throughout thecourse.

    : for all: there exists or : implies : if and only if: belongs to{x | condition}: set notation: intersection: unionq: disjoint union, or if A = B q C, then A = B C and B C = S \A: set difference, or {x S |x / A}

    An example of a formula written (almost) completely in symbols:

    ( > 0)( > 0) s.t. (|x y| < |f(x) f(y)| < )1.2. The Natural Numbers, or N.

    The main idea for the first part of the course will be R, or the real numbers. Thereare three main ways to prove their existence: decimal expansions, Dedekind cuts,or sequences of rationals. We will eventually use a hybrid of the second and thirdmethods, but that is for later. For now...N = {1, 2, 3, 4, 5, 6, . . . }. This is a commutative semiring with an identity (built

    off 1 with operators + and ). There is a notion of order, or x > y, contained inthe algebraic structure of N.

    Axiom (Induction). If S N and:(1) 1 S(2) x S (x+ 1) S

    hold, then S = N.

    In other words, if the two conditions above are true, then nothing can fail to be inS. Note that we will say 0 / N.

    1.3. The Integers, or Z.

    Z = N {0} (N). This is a commutative ring with multiplicative identity,meaning that it is an abelian group under (+, 0) and is associative, commutative,and distributive (i.e. x(y + z) = xy + xz) under multiplication. We also define0x = 0 x.

    Note that a group G is a set with an operation ? such that g ?h G for g, h G,along with these properties:

    (1) G has an identity e, for which the following holds: e ? g = g ? e = g g.(2) (g ? h) ? k = g ? (h ? k)(3) (g G)(h G) such that g ? h = e = h ? g.Read the book for notes on their structure.

    1

  • Math 104, Spring 2011 Vaughan Jones Lecture 1

    An abelian group is a group with the property that g?h = h?g. From the propertiesof a group, it follows that the h in (3) is unique, namely h = g1 or g, and that(a ? b = a ? c) b = c.

    There is an order on Z, or some notion such that x > y. Namely,

    x y x y N {0} .We then obtain three properties about this ordering:

    (1) x x(2) (x y and y x) x = y(3) (x y and y z) x z

    These are the axioms of a partially ordered set. Another example is the orderingon the subsets of a set with instead of .

    1.4. The Rationals, or Q, an Ordered Field.

    Q = {mn | (m,n) = 1 and m,n Z and n 6= 0}. Note that (x, y) is equivalent tothe expression gcd(x, y).

    We will define an order on Q: Let P = {mn |m N, n N}. Then,x y y x P or y = x .

    Definition. A field F is a set with two operations +, and two distinguishedelements 0, 1 such that, for 0 6= 1,

    (F,+0) is an abelian group (F \ {0}, , 1) is an abelian group x(y + z) = xy + xz

    As a precaution, we can extend to all of F by 0x = 0 x.Q is not only a field, however, but also an ordered field.

    Definition. An ordered field is a field F with a subset P F with the propertyF = P q {0} q P

    where P = {p | p P}, and axioms(1) P + P P , where P + P = {p1 + p2 | p1, p2 P}(2) P P P, where P P = {p1 p2 | p1, p2 P}

    Proposition. 1 P.

    Proof. We will proceed by contradiction. Note that if x P, then (1)x P.Then, if 1 P, then (1) P, which is a contradiction because (1) (1) = 1,but 1 / P.

    Proposition. x1 P.

    Proof. We will proceed by contradiction, and assume that x1 P. If that is thecase, then x (x1) = 1. Then, we obtain a contradiction because x P and(x1) P, but 1 / P.

    2

  • Math 104, Spring 2011 Vaughan Jones Lecture 1

    Thus, given an ordered field (F,+, , 0, 1,P), we can define:x y [x < y] y x P {0} [y x P] .

    This ordering gives us the following consequences, taken as axioms in other sourcessuch as the book:

    (1) a b b a(2) c 0 ac bc(3) a2 0(4) (0 a, 0 b) ab 0(5) a > 0 a1 > 0(6) 0 < a < b 0 < b1 < a1

    3

  • Math 104, Spring 2011 Vaughan Jones Lecture 2

    2. 01/20/11: Building Up to the Reals, Part Two

    2.1. More on Q.

    Last time, we discussed ordered fields (F,+, , 0, 1,P) where F = P q {0} q P,and defined an order x y y x {0} P. Note that if x, y F , theneither x y or y x. This property is known as the Total Order Property.

    Now, its time for our first proof.

    Theorem. The equation x2 = 2 has no solution in Q.

    Proof. By contradiction; suppose x Q satisfies x2 = 2. Since (x)2 = x2 = 2,we may suppose x is positive. Write x = mn with m,n N and (m,n) = 1. Sincex2 = 2, (mn )

    2 = 2, which means that:

    (*) m2 = 2n2 .

    Therefore, m2 is even. However, because the square of an odd number is odd, itfollows that m is even. Say m = 2k with k N. Substituting m = 2k in (*), wesee:

    2k2 = n2

    So n2 is even, and by a previous argument, n is even. However, this means that 2divides both m and n, contradicting (m,n) = 1.

    Note: When writing proofs, make sure that you are thinking about whether whatyoure writing makes sense and whether youre writing too much.

    2.2. Terminology.

    Let (X,) be a poset, and let S 6= be a subset of X.Definition. S is bounded above if there is an x X with x s for all s S.Such an x is called an upper bound for S.

    Definition. Same as above, but with below, x s, lower bound.Definition. x X is a least upper bound for S if

    (1) x is an upper bound.(2) If y is an upper bound for S, then y x.

    The least upper bound of S is called the supremum and is denoted supS or S. If|S|

  • Math 104, Spring 2011 Vaughan Jones Lecture 2

    2.3. The Reals, or R, the Complete Ordered Field.

    R is a complete ordered field. Note that all complete ordered fields are isomor-phisms of one another, so there is only one complete ordered field.

    Proposition (-characterization of the supremum). Let S R, r R. Thenr = supS if and only if

    (1) r is an upper bound of S, and(2) ( > 0)(s S) with s > r .

    Proof. We will go in both directions.

    () (1) is trivial. For (2), let > 0 be given. By contradiction, suppose nosuch s exists, i.e. s S, s r r is an upper bound. However,r < r, so r is not the least upper bound, and we have a contradiction.

    () By (1) we only have to show that r is the least upper bound. Suppose p isan upper bound for S (namely, p S), and p < r. Apply (2) with = rpto obtain an s S such that s > r (r p) = p, which contradicts that pis an upper bound.

    Theorem. R is Archimedean, i.e. a Rn N such that n > a. Equivalently,a R (a > 0) n N such that 1n < a.Proof. By contradiction, suppose there is an a N. Then, N is bounded above.By the completeness of R, we can let s = supN. By the -characterization of thesupremum with = 1, we obtain that n N with n > s1. However, this impliesthat n+ 1 > s, which is a contradiction. Theorem. Q is dense in R, i.e. a, b R where a 0 and b > a, q Q suchthat a < q < b.

    Proof. We know that b a > 0, so 1ba > 0. Therefore, by the Archimedeanproperty, we can choose an m such that m > 1ba , which is equivalent to sayingmbma > 1. Consider {n N |n > ma}. This set is nonempty by the Archimedeanproperty. Let p be its least element (as a subset of N). Then, p < mb, because ifp mb, then p 1 would be greater than ma. Therefore, ma < p < mb, whichimplies a < pm < b. Note: In this proof, we have implicitly assumed that a 0. However, this does notcause any problems for us.

    5

  • Math 104, Spring 2011 Vaughan Jones Lecture 3

    3. 01/25/11: Completeness and Sequences

    3.1. Completeness.

    Theorem. The equation x2 = 2 has a solution in R.

    Proof. Let S = {x R |x2 = 2, x > 0}. Then, S is nonempty, because 1 S. S isalso bounded above, because if x, y 0, then

    x2 < y2 x2 y2 < 0 (x y)(x+ y) < 0 x y < 0so x < y. Thus, if y = 2, then y2 = 4 > 2, so 2 is an upper bound for S. Therefore,by the completeness of R, supS exists.

    Let x = supS. We want to show x2 = 2.To show x2 cannot be less than 2, we will assume that x2 < 2 and try to show

    that we can add > 0 to x such that (x+ )2 < 2, so that x is not an upper boundfor S.

    Scratchwork: We have x2 + 2x + 2 < 2 2x + 2 < 2 x2 = . So, we willtry to make 2x and 2 both less than 2 . Choose 0.

    Then, we obtain:

    (x+ )2 = x2 + 2x+ 2 < x2 + = x2 + 2 x2 = 2 .Now, we must show that x2 cannot be greater than 2. We will use the -

    characterization of the supremum, and try to find an > 0 such that (x )2 > 2,so that x is not the supremum of S.

    Scratchwork: We have x22x+2 > 2 x22 > 2+2x. Let = x22 > 0.We want to make 2x less than 2 . We can choose x2 2x > x2 = 2 .We know that x < x, but this implies that x > s for all s S, which is acontradiction because x is supposed to be the least upper bound of S.

    Therefore, x2 must equal 2, and thus x = supS is a solution to the aboveequation in R.

    This leads to the fact that every positive element in R is a square, a fact whichagain uses the completeness of R.

    3.2. Infinity.

    is a very nice symbol, with the properties that > r and < r for allr R.

    supS = means that S is not bounded above.Recall interval notation: [a, b) = {x | a x < b}, (a, b] = {x | a < x b}, and so

    on. Likewise, we define:

    [a,) = {x | a x}(, y) = {x |x < y}

    and so on.

    6

  • Math 104, Spring 2011 Vaughan Jones Lecture 3

    3.3. Sequences.

    Definition. A sequence beginning at m is a function from {n Z |n m} to R,or some other set, and is denoted (sn).

    Note: We need for a sequence to be well-defined, such as (sn) = n2 or (sn) = (1)n,

    not just something like (0, 12 ,14 , 0, 2, . . . ).

    Definition. limx(sn) = s, or sn s, is equivalent to:( > 0)(N N) s.t. (n > N) |sn s| < .

    Note: There are two irrelevancies in this definition: n N and |sn s| lead tothe same definition. Also, think of |a b| as the distance between a and b. If youadd the condition that Q, it is still the same definition by the density of Q.Example: limx 1+n1+3n =

    13 .

    Scratchwork: Notice that for any n, it is the case that:

    sn s = 1 + n1 + 3n

    13

    =3 + 3n 1 3n

    (1 + 3n)3=

    2

    3 + 9n

    and this is the same regardless of the absolute value sign. Therefore, we want2

    3+9n < .

    2

    3 + 9n< 2

    < 3 + 9n 2

    3 < 9n n > 1

    9(2

    3)

    Proof. Let > 0 be given. Choose any N N such that N > 19 ( 2 3). Then, forn > N :

    |sn s| = 1 + n1 + 3n 13

    = 23 + 9n < 23 + 9N <

    7

  • Math 104, Spring 2011 Vaughan Jones Lecture 4

    4. 01/27/11: Convergence, Monotone Sequences

    4.1. Some Basic Notes.

    If A B R and B is bounded above, then supA supB.Proposition. If (sn) is a sequence where lim sn = s and lim sn = t, then s = t.

    Proof. Suppose s 6= t and put = |st|50 into the definition of the convergence ofa sequence. Then, choose N such that n N |sn s| < |st|50 . Also, sincelim sn = t, choose M such that n M |sn t| < |st|50 . Choose k such thatk M and k N .

    Then, |st| |ssk|+|skt| by the triangle inequality. However, then |ssk| 0 such that |sn| < B for all n. Now sincetn t, choose N1 such that (n N1) |tn t| < 2B where > 0 is given.Also, since sn s, choose N2 such that (n N2) |t||sn s| < 2 . Now, chooseN = max{N1, N2}. Then, if n N :

    |sntn st| |sntn snt|+ |snt st| = |sn||tn t|+ |t||sn s| .Therefore, since |sn||tn t| B 2B = 2 because n N1 and |t||sn s| 2because n N2, their sum is less than or equal to 2 + 2 = .

    8

  • Math 104, Spring 2011 Vaughan Jones Lecture 4

    Note: This theorem is very convenient! For instance:

    1 + n

    1 + 3n=

    1 + 1n3 + 1n

    13

    .

    So what about when a sequence doesnt converge? Well, first of all, an un-bounded sequence obviously doesnt converge. But what about sn = (1)n?

    If this sequence converges, then for all > 0, then for all sufficiently large n,|sn a| < . However, if n is odd, then |1 + a| < and if n is even, |1 a| < .This means:

    2 = |1 (1)| |a 1|+ |a+ 1| < 2Set = 12 , and we have a contradiction.

    4.3. Monotonic Sequences.

    Definition. A sequence (sn) is nondecreasing if sn+1 sn for all n. (sn) isnonincreasing if an an+1 for all n. (sn) is monotonic if it is either nonde-creasing or nonincreasing.

    Theorem. In an ordered field F , F is complete if and only if every bounded mono-tone sequence converges.

    Proof. In next lecture.

    9

  • Math 104, Spring 2011 Vaughan Jones Lecture 5

    5. 02/01/11: Monotone, limsup, Cauchy

    5.1. Monotone Sequences, continued.

    Theorem. In an ordered field F , F is complete if and only if every bounded mono-tone sequence converges.

    Proof. We will go in both directions.

    () Let (sn) be monotonic and bounded. Without loss of generality, suppose(sn) is nondecreasing. Then, {sn |n N} is bounded above and nonempty.By completeness, there exists an s = sup{sN |n N}.

    We will claim that sn s. Let > 0 be given. We want to find anN such that if n N , then |sn s| < . By the -characterization of thesupremum, there exists N N such that sN > s . Then, for all n N ,we have:

    sN sn s s < sN snwhich implies |sn s| < .

    () We will again assume, without loss of generality, that (sn) is nondecreasing.Let S 6= be a subset of F that is bounded above by B. We must showthe existence of supS. We will construct inductively two sequences, (xn)and (yn), with the following properties:(1) xn S and yn is an upper bound for S.(2) xn+1 xn and yn+1 yn.(3) |xn+1 yn+1| 12 |xn yn|.

    To do this, let x0 S and y0 be any upper bound for S. Supposex1, x2, . . . , xn and y1, y2, . . . , yn are already constructed, and satisfy thethree properties. Then, either xn+yn2 is an upper bound for S or it is not.

    If it is, then set xn+1 = xn and yn+1 =xn+yn

    2 . If it is not, then there is an

    s S such that s xn+yn2 . Set xn+1 = s and yn+1 = yn. Note then thatevery xn stays in S and every yn is an upper bound for S, so (1) is satisfied.By construction, (2) is true, and so is (3). Therefore, the sequence we haveconstructed satisfies all of the given properties.

    Note that (xn) is a nondecreasing sequence bounded above by y0 andthat (yn) is a nonincreasing sequence bounded below by x0. Therefore, byhypothesis, there are x, y F such that xn x and yn y. We will claimthat x = y. It is obvious from (3) that |xn yn| 12n |x0 y0|. Suppose forthe moment we can show that 12n 0. Then, by choosing sufficiently largen, it will be the case that for any given > 0, |x xn| < , |yn y| < ,and |xn yn| < . Then, we can obtain that:

    |x y| = |x xn + xn yn + yn y| |x xn|+ |xn yn|+ |yn y| .Therefore, |x y| < 3, so x = y.

    However, we still need to show that 12n 0. We know 12n is a nonin-creasing bounded sequence, so it does have a limit z F . Also, by theelementary properties of limits which are valid in any ordered field, thesequence (2 12n ) 2z. However, the limit of (2 12n ) is the same as that of12n . Therefore, 2z = z, which means that z = 0. Therefore, we now knowthat x = y.

    This condition is what we need the convergence of 12n

    for.

    10

  • Math 104, Spring 2011 Vaughan Jones Lecture 5

    Finally, we need to show that x = y = supS. y is an upper bound forS because yn s for all s S and for all n N, so y s for all s S aswell. Also, if there existed an upper bound t < y, then there would be anxn such that xn > t, because (xn) x. Therefore, t cannot be an upperbound, and thus y is the least upper bound.

    Note: The way we proved that x = y is equivalent to proving lim(xn yn) = 0,which, for convergent sequences, implies that limxn = lim yn.

    Definition.

    limn sn = + M R, (N N) s.t. (n N) (sn M).limn sn = M R, (N N) s.t. (n N) (sn M).

    Note: is not a number.Proposition. Any unbounded monotone sequence tends to or .Proof. The proof is supplied in the textbook (10.4).

    5.2. lim sup, or lim, and lim inf, or lim.

    Suppose (sn) is a sequence in R. Let Tn = {sk | k n}. It is obvious thatTn Tn+1 Tn+2 . . . , and if A B for any sets A and B, then it is the casethat supA supB and inf A inf B.

    Therefore, the sequence (supTn) is nonincreasing and the sequence (inf Tn) isnondecreasing. By the completeness of R as well as the suitable consideration of in the unbounded cases, limn supTn and limn inf Tn always exist!

    Examples: lim sup(1)n = 1 and lim inf(1)n = 1.Let sn be defined such that sn = 1 1n if n is even and sn = n if n is odd. Then,lim sup(sn) = 1 and lim inf(sn) = .Let sn be defined such that sn = 1 +

    1n if n is even and sn = n if n is odd. Then,

    lim sup(sn) is still 1!

    Remember: lim sup, lim inf always exist.

    Theorem. Let (sn) be a sequence. Then:

    (1) sn s lim inf sn = lim sup sn = s(2) If lim inf sn = lim sup sn, then lim sn = lim inf sn = lim sup sn.

    Proof. We will prove each part.

    (1) We know that ( > 0)(N) s.t. {TN} (s , s+). Therefore, supTN [s , s+ ], and supTn [s , s+ ] for n > N . Because this is true forall , lim sup sn = s. The proof of lim inf is similar.

    The T stands for tail.

    11

  • Math 104, Spring 2011 Vaughan Jones Lecture 5

    (2) Suppose lim sup sn = r = lim inf sn for some r R. Let > 0 be given.We want to show that sn r. We know that:

    N1 s.t. supTn < r + n N1 andN2 s.t. inf Tn > r n N2 .

    Let N = max{N1, N2}. Then, if n N , sn Tn, so sn supTn < r + and sn inf Tn > r , which implies |sn r| < .

    5.3. Cauchy Sequences.

    Definition. Let (sn) be a sequence of real numbers. We say (sn) is Cauchy if:

    ( > 0)(N N) s.t. (m,n N) (|sn sm| < ) .Note: This makes sense in spaces with an arbitrary notion of distance.

    Proposition. A convergent sequence is Cauchy.

    Proof. Assume sn s for a sequence (sn). Let > 0 be given and choose N sothat (n N) |sn s| < 2 . Then, if m,n N , we obtain:

    |sm sn| = |sm s+ s sn| |sm s|+ |s sm| < 2

    +

    2= .

    12

  • Math 104, Spring 2011 Vaughan Jones Lecture 6

    6. 02/03/11: Cauchy Sequences, Subsequences

    6.1. A Simple Proof.

    Say we want to show that for 0 < r < 1, rn 0.One way to show this is to assume rn x. Then, we know that krn kx. Let

    k = r. Then, rrn rx, but rn+1 has the same limit as rn. Therefore, rx = x andx = 0.

    Another way is to show that ( 1r )n , as that implies rn 0. Write 1r as

    1 + b > 0. Then (1 + b)n 1 + nb, which tends to .

    6.2. Cauchy Sequences, again.

    Last time, we learned about lim sup, lim inf, and Cauchy sequences. Recall that(sn) is Cauchy if and only if:

    ( > 0)(N N) s.t. (n,m N) |sn sm| < .Note: This definition makes sense in any ordered field.

    Proposition. A Cauchy sequence is bounded.

    Proof. Choose = 17 (arbitrarily, of course). Then, we know that

    N N s.t. n,m N, |sn sm| < 17 .This means that for n N , |sn sN | < 17, or sN 17 < sn < sN + 17. So for alln, it is the case that

    sn max{s1, s2, . . . , sN1, sN + 17}and likewise,

    sn min{s1, s2, . . . , sN1, sN 17} .

    Theorem. If (sn) is a sequence in R, then (sn) is Cauchy if and only if (sn)converges.

    Proof. We will proceed in both directions.

    () This was proved in the previous lecture.() Let (sn) be a Cauchy sequence in R. Note that because (sn) is bounded,

    lim sup sn and lim inf sn exist and are real numbers. Therefore, it sufficesto show that lim inf sn = lim sup sn. Because lim inf lim sup in all cases,we only need to show that lim inf sn lim sup sn. To do this, we will showthat lim inf sn + lim sup sn for any > 0.

    Let lim sup sn = x and lim inf sn = y. Let be given, and choose anN N such that |sn sm| < for n,m N . Note that this impliessn sm < . Then, because lim sup sn = x, there exists an n N suchthat sn > x , and similarly there exists an m N such that sm < y+ .Therefore:

    x y = x sn + sn sm + sm y < 3 .

    13

  • Math 104, Spring 2011 Vaughan Jones Lecture 6

    Proposition. Let (sn) be a sequence in an (Archimedean) ordered field. Then, ifthere exist k, r (where 0 < r < 1) such that |sn sn+1| < rnk for all n, then (sn)is Cauchy.

    Proof. Let > 0 be given. Choose N such that n N rn < 1rk . This ispossible because rn 0, which we may need Archimedean for. Without loss ofgenerality, we will assume that m > n. Then, for n,m N , we have:

    |sn sm| = |sn sn+1 + sn+1 sn+2 + + sm1 sm| |sn sn+1|+ |sn+1 sn+2|+ + |sm1 + sm| < krn(1 + r + + rmn) =

    krn(

    1 rmn+11 r

    ) 0 and limn sn = s be given. ChooseN such that n N |sn s| < . Then, if k N , then |snk s| < becausenk N . Theorem (Bolzano-Weierstrass). In R, any bounded sequence has a convergentsubsequence.

    Proof Sketch. Take the interval in which the sequence lies, and divide it in half.Then, choose a half with infinitely many terms (there must be at least one) andpick a term in that interval. Then, divide that half in half and choose a half withinfinitely many terms again. Pick another term from that half, such that that termappears later in the sequence than the first. Repeat this sequence indefinitely. Thiswill lead to a limit for the subsequence that you will have picked.

    14

  • Math 104, Spring 2011 Vaughan Jones Lecture 7

    7. 02/08/11: Defining the Reals, Part One

    7.1. Equivalence Relations.

    Definition. Given a set S, an equivalence relation on S is a relation x y with 3properties:

    (1) x x x S (Reflexivity)(2) x y y x (Symmetry)(3) x y and y z x z (Transitivity)

    Given x S, define [x] = {y | y x}.

    Here are some facts about [x]:

    (1) x y [x] = [y](2) [x] [y] = if x y and [x] [y] = [x] if x y.

    Therefore, S is the disjoint union of equivalence classes.However, if S =

    iI Si, then define by:x y i I s.t. {x, y} Si

    It is easy to check that this relation is an equivalence relation, meaning that apartition on a set gives you an equivalence relation, not just the other way around.

    7.2. Defining the Reals.

    The Equivalence Relation. First, we will define a relation on sequences (sn) ofrationals, such that (sn) (tn) if and only if:

    limn sn tn = 0, i.e. ( > 0, Q) N s.t. n N |sn tn| < .

    This is an equivalence relation. Note that in an Archimedean field, ( > 0) can bereplaced by ( > 0 | Q), by the denseness of the rationals.

    We will show that if (sn) is Cauchy and (sn) (tn), then (tn) is Cauchy. Weknow that |sn sm| < and want to show that this implies |tn tm| < .|tn tm| = |tn sn + sn sm + sm tm| |tn sn|+ |sn sm|+ |sm tm| < 3Therefore, (tn) is Cauchy.

    So, we will define F to be the set of equivalence classes of Cauchy sequences. Wewill give it the structure of an ordered field and show it is complete.

    Checking the Field Properties. First, we will define + such that [(sn)] + [(tn)] =[(sn + tn)]. We must check to make sure that this is well-defined. Suppose (sn) (sn) and (tn) (tn). Then, we know that lim sn sn = 0 and lim tn tn = 0.Therefore:

    lim(sn sn + tn tn) = lim((sn + tn) (sn + tn)) = 0 + 0 = 0and + is well-defined. Let [(0n)] be the additive identity and note [(sn)] = [(sn)],so inverses exist. + is also commutative, so this is an abelian group.

    Next, we will define such that [(sn)] [(tn)] = [(sntn)]. Again, we want this tobe well-defined. First, we will check that (sntn) is also Cauchy:

    |sntn smtm| = |sntn sntm + sntm smtm| |sn||tn tm|+ |tm||sn sm| .15

  • Math 104, Spring 2011 Vaughan Jones Lecture 7

    Obviously, the rightmost side goes to zero, so (sntn) is Cauchy. Now, supposelim sn sn = 0 and lim tn tn = 0. We want to show that lim(sntn sntn) = 0.

    |sntn sntn| = |sntn sntn + sntn sntn| |tn||sn sn|+ |sn||tn tn|Again, obviously, the rightmost side goes to zero, so multiplication is also well-defined. It is obviously commutative, associative, distributive over addition, andhas an identity, namely [(1n)]. However, the existence of inverses is not trivial. Forthis, we will make an important observation.

    Proposition. Let (sn) be a Cauchy sequence of rationals. Then, either:

    (1) sn 0,(2) N s.t. sn > n N , or(3) N s.t. sn < n N .

    where Q and > 0.Proof. Suppose sn does not converge to 0. Then, the negation of the definition ofconvergence gives us that:

    > 0 s.t. N N,r N s.t. |sr| .By the Cauchiness of (sn), there exists an N s.t. |sn sm| < 2 n,m N , where is the same as the in the above expression. Choose the r guaranteed for this such that r N and |sr| .

    Suppose sr . Then, for n N , we have |sr sn| < 2 , so sn > 2 , which iscase (2).

    Now, suppose sr 2 . Then, for n N , we have |sr sn| < 2 , so sn < 2 ,which is case (3). Note that it is obvious to see that (1), (2), and (3) depend only on the class of (sn).

    If [(sn)] 6= [(0n)], then sn does not converge to 0, and we are in case (2) or (3).Then, we can change finitely many terms, by the above proposition, in order tohave sn 6= 0 for all n. Then, [(sn)] [( 1sn )] = [(1n)], but we must also make sure( 1sn ) is Cauchy. We obtain: 1sn 1sm

    = sn smsnsm < B2

    where |sn| < B, because Cauchy sequences are bounded.Therefore, finally, we have established that this is an abelian group under mul-

    tiplication, and thus that F is a field which contains the rationals as constantsequences, i.e. q Q is defined as (qn).Ordering the Field. We will define P as such:P = {[(sn)] | N, > 0 with sn > n N for some, hence any (sn) [(sn)]} .

    Essentially, P is the set of equivalence classes of Cauchy sequences that fall intocategory (2) of the above proposition. This trivially satisfies all of the conditionsnecessary for an order.

    Observe immediately that F is Archimedean. For some [(sn)], sn < k for somelarge integer k, because Cauchy sequences are bounded. Thus, 2k sn > 0 for alln, and (2kn) therefore bounds (sn) for all n. A similar argument can be used toestablish the nonexistence of infinitesimals.

    16

  • Math 104, Spring 2011 Vaughan Jones Lecture 8

    8. 02/10/11: Defining the Reals, Part Two

    8.1. Recap.

    Recall that the intuition of this construction is that we are approximating realsby sequences of truncations of rational approximations of each real number; this iswhere the notion of equivalence classes comes from.

    So far, we have that the additive group with identity [(0n)] and multiplicativegroup with identity [(1n)] exist and are abelian. Therefore, the previously defined Fis a field. Also, we have an order with P = {[(sn)] | case two of the proposition fromlast time}, so F is an ordered field. Furthermore, F is Archimedean because givena Cauchy, and thus bounded, [(sn)], there exists a q sn + for all n where q N,so in the field, q > [(sn)]. Thus, we will use the fact that in an Archimedeanfield, if all Cauchy sequences converge then it is complete in order to prove thecompleteness of F .

    8.2. Defining a Concept of Less Than.

    Note that |[(sn)]| = [(|sn|)]. We could define a notion of less than as such:|[(sn)]| < [(|sn|)] > 0 > 0 s.t. for large n, |sn| > , or |sn| < In this definition, , , sn are all rationals. This is a perfectly serviceable definition,but in order to get away from , we will use an alternate definition.

    Definition. [(sn)] is less than is defined through these two implications:

    (1) [(sn)] < N N s.t. sn < n N , and(2) N N s.t. sn < n N [(sn)] .

    Now, we are finally ready.

    8.3. Convergence of Cauchy Sequences in F .

    Theorem. Every Cauchy sequence in F converges.

    Proof. What we have to do is given ([(sn)]m) Cauchy, come up with a [(tn)] wherelimm[(sn)]m = [(tn)]. Our notation will be:

    [(sn)]1 = s1,1, s2,1, s3,1, s4,1, . . .

    [(sn)]2 = s1,2, s2,2, s3,2, s4,2, . . .

    [(sn)]3 = s1,3, s2,3, s3,3, s4,3, . . .

    ...

    [(sn)]k = s1,k, s2,k, s3,k, s4,k, . . . , sk,k, . . .

    We are going to shoot for the diagonal, (sn,n). However, we are dealing withequivalence classes of Cauchy sequences, so we could have made a bad choice ofrepresentatives and (sn,n) may go to infinity. Therefore, we will change [(sn)]k sothat it only varies by 1k in each term. We will do this by changing all elementsbefore N (Note, a finite number) so they all live within of the next. or, moreformally...

    Proposition. Let (sn) be a Cauchy sequence. For all > 0, there exists a (tn)such that (sn) (tn) and |tn tm| < for all n,m.

    17

  • Math 104, Spring 2011 Vaughan Jones Lecture 8

    Proof. N s.t. |sn sm| < n,m N . Then, fix:t1 = sN , t2 = sN , . . . , tn1 = sN , and tn = sn n N .

    It is obvious that [(tn)] = [(sn)]. So, because [(sn)]m is Cauchy, we can suppose:

    (**) |sn,m sk,m| < 1mk,m

    Now, set the aforementioned tn = sn,n. Now, all that remains for us to do is toshow that [(sn)]m [(tn)] as m.

    First, we will deal with the fact that [(tn)] must be Cauchy. We want to compare|tn tm|, or |sn,n sm,m|. Let > 0 be given, where in this case Q. Choose anN > 1 s.t. |[(sn)]m [(sn)]k| < m, k N . Also, choose a p s.t. |sp,msp,k| < .Now, suppose m, k N .|tm tk| = |sm,m sk,k| |sm,m sp,m|+ |sp,m sp,k|+ |sp,k sk,k| < 3

    where the first term is bounded by (**), the second term is bounded by the choiceof N and the Cauchiness of ([(sn)]m), and the third term is also bounded by (**).Therefore, we are done, and [(tn)] F .

    Now, we only need to show that limm[(sn)]m = [(tn)], or in other words that > 0 N s.t. m N |[(sn)]m [(tn)]| . Note that it is less than or equal to because of the second implication in the definition of less than above. Let > 0be given. By the Cauchiness of (tn), we have N > 2 s.t. m,n N, |tntm| < 2 .Then, we obtain:

    |tn sn,m| = |sn,n sn,m| = |sn,n sm,m + sm,m sn,m| |tn tm|+ |sm,m sn,m|

    2+

    2=

    Where the first term is bounded by the Cauchiness of (tn) and the second is boundedby (**). Therefore, |(tn) [(sn)]m| .

    Thus, since every Cauchy sequence in F converges and F is Archimedean, F iscomplete! 8.4. Metric Spaces.

    Definition. A metric space is a set X together with a function d : X X R,known as a metric, where d(x, y) satisfies:

    (1) d(x, y) 0 and d(x, y) = 0 x = y,(2) d(x, y) = d(y, x), and(3) d(x, y) d(x, z) + d(z, y) x, y, z X.

    Basically, all we are given when talking about metric spaces is a notion of distance.

    Example: In R, a possible metric is d(x, y) = |x y|.Example: Take any X, and define d such that:

    d(x, y) =

    {0, if x = y;

    1, if x 6= y.This is known as the discrete metric.

    18

  • Math 104, Spring 2011 Vaughan Jones Lecture 9

    9. 02/15/11: Metric Spaces, Part One

    9.1. Metric Spaces on Rn.

    Definition. A metric space is a set X and a function d : X X R such that:(1) d(x, y) 0 and d(x, y) = 0 x = y,(2) d(x, y) = d(y, x), and(3) d(x, y) d(x, z) + d(z, y)

    Remember that some examples are |x y| on R and the discrete metric from lastlecture.

    Now consider Rn = {(x1, x2, . . . , xn) |xi R}. We will give this set severaldifferent metrics.

    9.1.1. The Euclidian Metric. This metric is defined by:

    d((x1, . . . , xn), (y1, . . . , yn)) =

    (x1 y1)2 + (x2 y2)2 + + (xn yn)2

    For this metric, (1) and (2) are obvious. We will come back to (3).

    We will now explore the concept of norms, with the intention of obtaining somenew metrics for Rn during this exploration.

    Definition. A Norm on a (real) vector space V is a function || || : V R suchthat:

    (i) ||v|| 0 and ||v|| = 0 v = 0,(ii) ||v|| = || ||v|| for R, and(iii) ||v + w|| ||v||+ ||w||.

    9.1.2. The Maximum Norm. We will define this norm such that for V = Rn:

    ||(x1, . . . , xn)|| = max{|x1|, |x2|, . . . , |xn|}(i) and (ii) are obvious. For (iii), use the triangle inequality on R.

    9.1.3. The Sum Norm. We will define this norm such that for V = Rn:

    ||(x1, . . . , xn)||1 =ni=1

    |xi|

    (i), (ii), and (iii) are all trivial for this norm.

    Norms can also come from inner products (sometimes known as scalar products),which we will further investigate.

    Definition. An inner product is a function , : V V R such that:(a) v, v 0 and v, v = 0 v = 0,(b) v, w = w, v, and(c) v + w, x = v, x+ w, x and x, v + w = x, v+ x,w.

    19

  • Math 104, Spring 2011 Vaughan Jones Lecture 9

    9.1.4. The Squaring Norm. We will define this norm in terms of an inner product,such that for V = Rn:

    ||v||2 = n

    i=1

    x2i

    This norm has been defined from an inner product, whereni=1 x

    2i = v, v for

    v = (x1, x2, . . . , xn). While properties (i) and (ii) are immediate from the propertiesof an inner product, we must still verify (iii) for this norm. In other words, we mustshow that:

    (*) ||v + w||2 ||v||2 + ||w||2 .However, because both terms are positive, showing that the inequality holds whenboth sides are squared is equivalent to showing that it holds as presented in (*).Therefore, we have, by (c) (also known as bilinearity),:

    ||v + w||22 ||v||22 + 2 ||v||2 ||w||2 + ||w||22 v + w, v + w ||v||22 + 2 ||v||2 ||w||2 + ||w||22

    v, v+ 2 v, w+ w,w ||v||22 + 2 ||v||2 ||w||2 + ||w||22 | v, w | ||v||2 ||w||2

    Note that the absolute value signs in the last inequality are valid because if it istrue without the absolute value signs, then it is true with them. This last inequalityis known as the Cauchy-Schwarz inequality, which is equivalent to:xiyi x2i y2iSo now, to show that || ||2 is a norm, we will prove Cauchy-Schwarz.Proof. Take v, w V . Consider v + w, v + w 0, R.

    v + w, v + w 0v, v+ 2 v, w+ 2 w,w ||v||22 + 2 v, w+ 2 ||w||22 0(

    ||w||2 +v, w||w||2

    )2+ ||v||22

    v, w2||w||22

    0

    Because this is true for all , we can choose so that the first term is 0. Then, weobtain:

    ||v||22 v, w2||w||22

    0 ||v||22 ||w||22 v, w2 | v, w | ||v||2 ||w||2 .

    Therefore, since we have established the Cauchy-Schwarz inequality, || ||2 is a norm.

    The point of all of this norm-taking was that given a norm, we can make it ametric on Rn by d(v, w) = ||v w||. For instance, || ||2 is exactly the same as theaforementioned Euclidian metric on Rn!

    20

  • Math 104, Spring 2011 Vaughan Jones Lecture 9

    9.2. Metric Space Concepts. Let (X, d) be a metric space.

    Definition. The open ball centered at a of radius r for r > 0, a X is:B(a, r) = {x X | d(a, x) < r}

    Example: In R, the open ball is denoted (a r, a + r). In the discrete metric, theballs are either the single point or the whole space. In R2, || ||2, B(a, r) is theopen circle of radius r around a. With || ||, B(0, 1) is the square length 2 aroundO. With || ||1, B(0, 1) is the interior of the diamond created by taking the linex+ y = 1 and reflecting it across all of the axes. B(0, 1) is known as the unit ball.

    Definition. Given a sequence (xn) and an x X, (xn) converges to x if:( > 0)(N N) s.t. d(xn, x) < n N

    Definition. Given a sequence (xn) and an x X, (xn) is Cauchy if:( > 0)(N N) s.t. n,m > N d(xn, xm) <

    Definition. (X, d) is complete if and only if every Cauchy sequence converges.

    Note: Convergence in Rn (with any of the norms defined above) is componentwiseconvergence. So, is Rn complete in its three metrics? Yes! The discrete metric isalso complete.

    Definition. A subset U X is open if:x U r > 0 s.t. B(x, r) U

    Proposition. B(a, r) is open.

    Proof. Let x B(a, r) be given. Then, r > d(a, x), so r d(a, x) > 0.We will claim that B(x, r d(a, x)) B(a, r). Suppose y B(x, r d(a, x)).

    Then, we know that d(x, y) < r d(a, x) d(a, x) + d(x, y) < r. So, by thetriangle inequality, d(a, y) < r. Therefore, for any y B(x, r d(a, x)), y is inB(a, r). Therefore, B(x, r d(a, x)) B(a, r), and B(a, r) is open. As another example, every subset of the discrete metric is open.

    Definition. A subset E X is closed if and only if X \ E is open.Note that closed does not mean not open!Example: In the discrete metric, every subset is both open and closed.

    Proposition. E X is closed if and only if (xn E)(xn x) x E.Proof. We will go in both directions.

    () Assume X \ E is open. Also, suppose (xn) E and xn x, andtoward contradiction, that x / E, or x X \ E. This implies thatr > 0 s.t. B(x, r) X \ E. But since xn x, d(xn, x) < r for somen. Then, xn B(x, r) X \ E, so xn (X \ E) (E), which is a contra-diction.

    () Toward contradiction, suppose X \ E is not open. Then, x X \ E suchthat B(x, 1n ) E 6= n. This means that for every n N, there existssome yn E such that d(yn, x) < 1n . Thus, consider the sequence of yns,which we will call (yn). (yn) then obviously converges to x, but then byassumption and hypothesis, x (X \ E) (E), which is a contradiction.

    21

  • Math 104, Spring 2011 Vaughan Jones Lecture 10

    10. 02/17/11: Metric Spaces, Part Two

    10.1. Review.

    Given (X, d), U X is open if x U r > 0 s.t. B(x, r) U. If Ui where i Iis open for all i, then

    iI Ui is open.

    E X is closed if X \E is open. If Ei is closed for i I, theniI Ei is closed.

    10.2. More Metric Space Concepts.

    Definition. For x X, a neighborhood of x is a set S X such that there existsan open set U X with x U S.Definition. The closure of E is:

    E =

    F closed,EFF

    Observe that E is closed. This is the smallest closed set that contains E. Notethat B(a, r) is not necessarily C(a, r) = {x X | d(a, x) r}; for instance, this isnot true in the discrete metric space.

    Note that subsets can be both open and closed! For example, define a metricspace where X = [1, 1] (5, 17). Then, each interval is an open and closed subsetof the space. Another example is if X = Q, then (2,2) is open and closed.

    Definition. The interior of S X is S = {s S | r > 0 s.t. B(s, r) S}, or{s S |S is a neighborhood of S}.

    Definition. The boundary of S is S = S \ S.10.3. The p-adic Metric.

    We will define a metric on Z based on any p for p prime. We will call this a p-adicmetric.

    Define a metric by a norm, |j|p. Write j = pr m such that (m, p) = 1. Let|0|p = 0 and |j|p = 2r where 2 can be any prime.

    Then it is the case that

    |j|p 0 and j = 0 |j|p = 0as well as that

    |i+ j|p max{|i|p, |j|p} |i|p + |j|p .Furthermore, dp(i, j) = |i j|p, and this metric makes Z into a metric space. Notethat, for instance, for p = 3, 3n 0.

    Note that the empty set is open.Note that the empty set is closed.

    22

  • Math 104, Spring 2011 Vaughan Jones Lecture 10

    Example:(1 3)(1 + 3 + 32 + + 3n) = 1 3n+1 1

    Therefore, 2 + 2 3 + 2 32 + + 2 3n + = 1.We can also define | |p on Q, where q = pr mn for (m, p) = 1, (n, p) = 1. Then,

    |q|p = 2r and dp(q1, q2) = |q1 q2|p.Note that we can write numbers in decimal form by writing them as such:

    x =

    i=1

    ai pi +j=1

    di pi := . . . a3a2a1.d1d2d3 . . .

    Then, we get a decimal where as you add terms to the left it converges and asyou add to the right it diverges!

    We can define a field with this metric by defining it like we did R, but with thisnew metric instead of the familiar one for R. This field is called the p-adic numbers.Note that this technique of completing a metric space (as we did when constructingR from Q) can be applied to any metric.

    10.4. Even More Metric Space Concepts.

    Definition. Let (X, d) a metric space and S X. Then, x X is called anaccumulation point (or limit point or cluster point) of S if every neighborhoodof x contains a point in S other than x.

    Note that E X is closed if and only if it contains all of its accumulation points.Definition. Let (X, d) be given, and consider T := {U |U open in X}. This col-lection of subsets is called a topology on X defined by d. A topology satisfies threeaxioms:

    (1) T , X T .(2) If Ui T , then i I

    iI Ui T .

    (3) For U1, U2, . . . , Un T , it is the case thatni=1 Ui T .

    Any property that can be expressed only using T , or open sets, is called a topo-logical property. For instance, we can say...

    Proposition. Convergence is a topological property, or in other words:

    limnxn = x U open in X and x U, N s.t. n N xn U .

    Proof. We will go in both directions.

    () Let U be open with x U given. Since U is open, r > 0 s.t. B(x, r) U .Therefore, N s.t. n N d(xn, x) < r xn B(x, r) U .

    () Let > 0 be given. We know that B(x, ) is open. Thus, N s.t. n N xn B(x, ) d(xn, x) < .

    Therefore, changing the metric does not change the notion of convergence on a giventopology. Note that Cauchiness and completeness are not topological properties,but closures and interiors are.

    Definition. Let U, V be open subsets of X. A topological space X is Hausdorffif given that x U , y V , and x 6= y, then U V = .

    23

  • Math 104, Spring 2011 Vaughan Jones Lecture 11

    11. 02/24/11: Continuity and Cardinality, Part One

    11.1. More About Metric Spaces.

    Proposition. x E ( > 0) B(x, ) E 6= .Proof. We will go in both directions.

    () We will proceed by contradiction. Suppose there exists an > 0 suchthat B(x, ) E = . Then, X \ B(x, ) is a closed set containing E. Bythe definition of a closure, this implies x E X \ B(x, ). This is acontradiction because x B(x, ).

    () Let F be closed and E F . Suppose x / F . Then there exists an > 0such that B(x, ) X \ F . Because E F , X \ F X \ E. This is acontradiction, so for x F and F where F is closed, E F x E.

    Definition. Given (X, d) and A,B X, we say that A is dense in B if B A.Note that AX B = AB B (where AX is the closure of A under X and AB is theclosure of A under B), by the -criteria for x E.

    11.2. Continuity.

    Definition (-). Given metric spaces (X, d) and (Y,D), a function f : X Y issaid to be continuous at x if

    ( > 0)( > 0) s.t. d(x, y) < D(f(x), f(y)) < .Note that this is continuity at a single point !Example: f : R R such that

    f(x) =

    {0, if x Q;x, otherwise.

    This is a function which is continuous at 0 and discontinuous anywhere else.

    Proposition. f is continuous at x X (xn x) (f(xn) f(x)).Note: This is known as the limit definition of continuity at a point.

    Proof. We will go in both directions.

    () Suppose xn x and let > 0 be given. Choose a such thatd(x, t) < d(f(x), f(t)) < .

    Using in the definition of xn x, get an N s.t. n N , d(xn, x) < .This implies that d(f(xn), f(x)) < .

    () Suppose f is not continuous at x X. This means that: > 0 s.t. > 0,t B(x, ) s.t. d(f(t), f(x)) .

    Choose a sequence xn B(x, 1n ) with d(f(xn), f(x)) . Then, we have acontradiction, as xn x but f(xn) 9 f(x).

    24

  • Math 104, Spring 2011 Vaughan Jones Lecture 11

    Note that this has become the commonly accepted definition of continuity. However,this is because of the shift in focus from topological spaces to metric spaces (areversal of a previous trend). In a general topological space, the limit definition ofcontinuity does not hold!

    Definition. Given metric spaces X,Y and f : X Y , f is continuous if f iscontinuous at every point of X.

    Proposition. f : X Y is continuous f1(U) is open for all open U Y .

    Proof. We will go in both directions.

    () Let U be open in Y . Take x f1(U). This implies that f(x) U . By thedefinition of continuity, B(x, ) s.t. f(B(x, )) U , so B(x, ) f1(U).

    () Take x X and let > 0 be given. We must show that > 0 s.t. f(B(x, )) B(f(x), ) .

    However, B(f(x), ) is open, so f1(B(f(x), )) is open by hypothesis. Also,x f1(B(f(x), )), so > 0 such that B(x, ) f1(B(f(x), )), whichimplies that f(B(x, )) B(f(x), ).

    Definition. (X, d) and (Y,D) are isometric if there exists a bijection f : X Ywith D(f(x), f(y)) = d(x, y). This implies that f1 is also an isometry.

    Definition. f : X Y is called a homeomorphism if f is a continuous bijectionwhose inverse is continuous.

    Note: The last condition is necessary! For instance, the identity function from(R, discrete) to (R, usual) is continuous but f1 is not.

    11.3. Countability and Cardinality.

    We will take a little detour from metric spaces in order to talk about Cardinality,so as to prepare for the guest lecture in two lectures from Professor Hugh Woodin.

    First, we will explore the Cantor Set, defined as such:

    S0 = [0, 1]

    S1 =

    [0,

    1

    3

    ][

    2

    3, 1

    ]S2 =

    [0,

    1

    9

    ][

    2

    9,

    1

    3

    ][

    2

    3,

    7

    9

    ][

    8

    9, 1

    ]S3 =

    [0,

    1

    27

    ][

    2

    27,

    1

    9

    ][

    2

    9,

    7

    27

    ][

    8

    27,

    1

    3

    ][

    2

    3,

    19

    27

    ][

    20

    27,

    7

    9

    ][

    8

    9,

    25

    27

    ][

    26

    27, 1

    ]The Cantor set is defined as C :=

    nN Sn. Essentially, we start with [0, 1] and

    remove the middle third each time, for every interval. The endpoints of the intervalsin the set are definitely countable, but the set itself as actually uncountable!

    The related concept of uniform continuity is defined in a footnote in Lecture 16.Reminder: Given sets A,B and f : A B, if:

    S A, define f(S) = {f(x) |x S}, andT B, define f1(T ) = {x S | f(x) T}.

    In this case, f1 is not the inverse of f ! f1 always exists when applied to a set.25

  • Math 104, Spring 2011 Vaughan Jones Lecture 12

    12. 02/24/11: Cardinality, Part Two

    12.1. The Cantor Set, Continued.

    Recall the Cantor set from last time. Here is an inductive way to define it.

    Take an interval I = [a, b]. We will define C(I) = {[a, a + ba3 ], [a + 2(ba)3 , b]}.So, given J0 = {[0, 1]}, we have Jn+1 =

    IJn C(I). Then,

    C =nN

    (IJn

    I) .

    Proposition. The interior of the Cantor set is empty, or C = .

    Proof Sketch. Choose n, with 3n < 10 . Then, for some n, x is in an interval ofwidth 3n, so there is no ball around x.

    12.2. Random Facts.

    There exists a continuous nondecreasing function f : [0, 1] [0, 1] with f(0) = 0and f(1) = 1 such that f is constant on [0, 1] \ C. The graph of f is known as theDevils Staircase.

    Russells Paradox. Given a set S, S S or S / S. Define A = {S |S / S}. DoesA belong to A? There is no way for A to belong to or not belong to itself, as bothresult in contradictions. Thus, Russells Paradox will act as a cautionary tale forthe discussions to come; we must be careful when were talking about sets of allsets.

    12.3. Cardinality.

    We will denote the cardinality of a set A as |A| (or #(A)).Definition. Given sets A and B, |A| = |B| if there exists a bijection f : A B.Definition. A is finite if there exists an n N such that |A| = |{1, 2, . . . , n}|. Inother words, for finite sets, |A| = |B| they have the same number of elements.A is infinite if it is not finite.

    Definition. A is countable if A is finite or |A| = |N|, or equivalently, you canenumerate the elements of A as a sequence.

    Proposition. S N is countable.

    Proof. Define (cn) as follows. Define c1 to be the smallest element of S, and defineS1 = S \ {c1}. Then, define cn+1 to be the smallest element of Sn, and defineSn+1 = Sn \ {cn}. Then, it is easy to check that {ck | k N} = S.

    Definition. |A| |B| (also written A - B) if there exists an injection f : A B.26

  • Math 104, Spring 2011 Vaughan Jones Lecture 12

    Example: If S T , then S - T . Therefore, this order on sets says that up to achange of name, A B.Proposition. A - B a surjection g : B A.Proof. We will go in both directions.

    () Given an injection f : A B, define f such that for a A, g(f(a)) = aand for everything else, g(x) = a for some a.

    () We want to find an injection f : A B. For each a A, choose an x Bwith g(x) = a, and define f(a) = x.

    Proposition. Z, Z Z, and Q are countable.Proof. We will prove them one at a time.

    (1) Z: We will arrange the elements of Z in a sequence, namely (cn) such thatc0 = 0, c2n1 = n, c2n = n for all n N.

    (2) Z Z: By the previous argument, |Z| = |N|, so it suffices to show thatN N is countable. We will arrange the elements of N N as such:

    (1, 1) (1, 2) (1, 3) (1, 4) . . .

    (2, 1) (2, 2) (2, 3) (2, 3) . . .

    (3, 1) (3, 2) (3, 3) (3, 4) . . ....

    ......

    .... . .

    Now, proceed down each diagonal, starting at (1, 1), then moving to thenext diagonal containing (1, 2) and (2, 1), then the next, containing (1, 3),(2, 2), (3, 1), and so on. Alternately, define a function f : N N R suchthat f((m,n)) = 2m3n. By the unique prime factorization of integers, thisis certainly an injection.

    (3) Q: Take the function f : Z N Q such that f((m,n)) = mn . This isclearly a surjection.

    12.4. Counting R.

    We will define 2N = {sequences (an) such that an {0, 1}}, such as 01101001 . . .Note that this is the same as the set of all subsets of N, and in general, 2A is thesame as subsets of A, also known as the power set of A.

    Theorem (Cantor). 2N is uncountable.

    Proof. Suppose it is countable. Then, we can enumerate them as such:

    (an)1 = a1,1 a2,1 a3,1 . . .

    (an)2 = a1,2 a2,2 a3,2 . . .

    (an)3 = a1,3 a2,3 a3,3 . . ....

    ......

    .... . .

    27

  • Math 104, Spring 2011 Vaughan Jones Lecture 12

    Construct bn = an,m + 1 (mod 2). Obviously, bn is in 2N, but it cannot be the nth

    term in the sequence of sequences because it differs from the nth term in the nthplace. Therefore, it is not in the list, and we have a contradiction. Corollary. R is uncountable.

    Proof. It suffices to define an injection f : 2N R. Let f((an)) =n=1 an10

    n.We will claim that f is an injection. The proof is in the lecture notes on Cardinality,page 2. There is another proof in the notes, page 3, which shows that R has cardinality noless than |2N|.Theorem (Schroeder-Berinstein). If |A| |B| and |B| |A|, then |A| = |B|.Proof. Next lecture. By the remark after the Corollary and this Theorem, we can see that |R| = |2N|.

    The Continuum Hypothesis asks whether there is a subset of R that is uncount-able but that does not have the same cardinality as 2N. Or in other words, if thereexists a set A such that |N| |A| |2N|. The answer is still not known.

    Note that series are not defined yet. However, partial sums are Cauchy sequences, so they arereals.

    28

  • Math 104, Spring 2011 Vaughan Jones Lecture 13

    13. 03/01/11: Set Theory: Guest Lecture by Hugh Woodin

    13.1. Schroeder-Bernstein.

    Two prominent set theorists were Cantor and Godel, both of whom went insane.Anyway...

    Definition. Two sets X,Y have the same cardinality if there exists a bijection: X Y . We denote this as |X| = |Y |.Caution: This definition is not as intuitive as it seems. If X is infinite and a X,then it can happen that X and X \ {a} have the same cardinality.Definition. Suppose X,Y are sets. Then, the cardinality of X is less than or equalto the cardinality of Y if there exists an injection : X Y . We denote this as|X| |Y |.

    Now, we will turn our attention to the Schroeder-Bernstein Theorem.

    Theorem (Schroeder-Bernstein). Let X,Y be sets and suppose |X| |Y | and|Y | |X|. Then, |X| = |Y |. In other words, suppose X,Y are sets and there existsinjections f : X Y and g : Y X. Then, there exists a bijection : X Y .

    Before proving the theorem, we will first develop some concepts and propertiesthat will help us along the way.

    Consider a set Z and an injection h : Z Z. Start with some point a Z, andkeep applying h to it. Then, one of three things will happen:

    (1) We will get into a loop, so hn(a) = a for some n N.(2) We will reach a point where hn(a) is not in the range of h, so we cannot

    back up any farther, and effectively, the chain starts at a certain point.(3) We will never cycle, but we will be able to back up endlessly, so for all

    n N, hn(a) is in the range of h.Notice that case 3 is just a degenerate case of case 1, as it is simply just an endlessloop backwards. Now, we will define a concept using this intuition.

    Definition. Suppose h : Z Z is injective and suppose a Z. Then the orbit ofa given by h is the subset of Z generated by a, h, h1 where h1 : {range of h} Z.More formally, define X0 = {a}. Then, define inductively:

    xn+1 = {h(b) | b Xn} {c Z |h(c) Xn} XnThen, the orbit of a given by h, denoted Za, is defined to be

    n=0Xn where n N.

    Example: X1 = {a, h(a)} {c} where h(c) = a if such a c exists. Furthermore,X2 = X1 {h(h(a))} {d} where h(d) = c if c existed and if d exists.Example: Let Z = {0, 1, 2, . . . }, and let h(n) = n2 for n Z. Then, Z0 = {0},Z1 = {1}, and Z2 = {2, 4, 16, 256, . . . }, or in other words the even powers of 2. Z3can also be seen as the even powers of 3.

    Note that we can also denote the orbit of a given by h as Zha in order to explicitlystate the injection that we are dealing with.

    Lemma 1. Suppose h : Z Z is injective and suppose a, b Z. Let Zha and Zhbbe the orbits of a and b relative to h, as defined above. Also, suppose Zha Zhb 6= .Then, Zha = Z

    hb .

    29

  • Math 104, Spring 2011 Vaughan Jones Lecture 13

    Proof. Exercise.

    Another intuitive definition follows from our examination of orbits above.

    Definition. Suppose h : Z Z is injective and suppose a Z. Let Zha be the orbitof a given by h. Zha has an eve-point if there exists a b in Z

    ha such that b is not

    in the range of h.

    Lemma 2. If an orbit has an eve-point, then the eve-point of that orbit is unique.

    Proof. Exercise.

    Lemma 3. Let h : Z Z be injective and a Z. Suppose Zha has no eve-point.Then, h|Zha : Zha Zha is a surjection, and hence a bijection.Proof. Exercise.

    Now, we are ready for the actual proof.

    Proof (Schroeder-Bernstein). We are given injections f : X Y and g : Y X.We need to construct a bijection : X Y . Consider the functions g f : X Xand f g : Y Y . These are both injections because the composition of twoinjections is also always injective. So now, we can consider the orbits of points in Xgiven by gf and the orbits of points in Y given by f g. We will define : X Yby defining |Xgfa for each orbit of X given by g f . There are two cases for eachorbit:

    (1) Xgfa has no eve-point. Then, for each b Xgfa , (b) = f(b).(2) Xgfa has an eve-point. Let a0 be the eve-point. Then, it follows that Y

    fgf(a)

    must also have an eve-point. Let b0 be the eve-point of Yfgf(a), and define

    (a0) = b0, (a1) = b1, and so on.

    We can see that every point of every orbit is hit exactly once by , and thus isa bijection from X to Y .

    This theorem proves the rather trivial-looking statement that if |X| |Y | and|Y | |X|, then |X| = |Y |. However, while it may look trivial, it really is not.Proposition. |R| = |P (N)|, where P (N) = {A |A N}.Proof. It suffices to show that |R| |P (N)| and |P (N)| |R|.

    (1) Clearly, there is a bijection : N Q, so we just need an injection f : RP (Q). Define f(r) := {q | q Q, q < r} for all r R. This is clearlyinjective, so |R| |P (Q)| = |P (N)|.

    (2) We need an injection g : P (N) R. Define g(a), where a P (N), as such:

    g(a) =

    {ia(

    110 )

    i if a 6= 0 if a =

    This is clearly injective, so |P (N)| |R|.

    Theorem (Generalized Cantor). Suppose X is a set. Then, |X| 6= |P (X)|, whereP (X) = {A |A X}.Note that this implies |X| < |P (X)| and |R| = |P (N)| 6= |N|.

    30

  • Math 104, Spring 2011 Vaughan Jones Lecture 13

    Proof. Suppose f : X P (X). We need to produce an A X such that A is notin the range of f . Define A = {a X | a / f(a)}. Then, A cannot be in the rangeof f . To see this, suppose A = f(b). Then, we have:

    b A b / f(b) b / Awhich is clearly a contradiction. This theorem implies that there are an infinite number of infinities, each one biggerthan some others.

    Therefore, we will say 0 = |N|. Where does |P (N)| = |R| fit in? Is it 1?The Continuum Hypothesis states that |R| = 1. We have gotten to the point

    where we know that given our standard axioms, we can build universes where thisis true and universes where this is false. For another illustration of a conjecture likethis, consider the Parallel Postulate. It is true in R2, but false in a universe thatis a circle without the edge, despite that the usual axioms of geometry hold in thisnew space.

    In recent times, weve gotten closer to an answer. Behind this question and a lotof other set theory lie many philosophical issues about the meaning of things thatdont exist...

    31

  • Math 104, Spring 2011 Vaughan Jones Lecture 14

    14. 03/03/11: Compactness, Part One

    14.1. Basic Concepts.

    Let A X, and let (X, d) be a metric space.Definition. An open cover of A is a family {U | I} of open subsets of Xsuch that A I U.Example: [0, 1] R. Then, {(1, 2)} is an open cover. {(x , x + ) |x [0, 1]}is also an open cover. Consider = 13 . Then, the cover is very redundant, as it isan uncountably large collection of intervals, and we could come up with somethingmore economical.

    Definition. If {U | I} is an open cover of A, a subcover is a subfamily{U | J}, where J I, that is also an open cover of A.Definition. A X is compact if for every open cover, there is a finite subcover.It is the case that A X is compact if and only if A A with the inheritedmetric is compact. Therefore, the notion of compactness is an inherited notion, likeclosedness. Note that this is a very difficult definition to work with, if we want toshow a set is compact. However, this does make it very easy to show that a set isnot compact.

    Example: Finite metric spaces are always compact for any A X.Example: Take R. {(x , x+ ) |x R} is an open cover of R. Furthermore, youcannot cover R with finitely many intervals, so R is not compact. Another exampleis, for Z, the open cover {(n 32 , n+ 32 )}.14.2. Basic Properties.

    Proposition. In any metric space, a compact subset is closed and bounded.

    Proof. We will prove the two points separately.

    (1) First, we will prove boundedness. Take x A. Consider {B(x, n) |n N}.This family of sets is trivially an open cover, even of all of X. Take a finitesubcover {B(x, n1, . . . , B(x, nmax)}. Then,

    maxi=1 B(x, ni) = B(x, nmax).

    A maxi=1 B(x, ni) because its an open subcover, so A B(x, nmax) andA is bounded.

    (2) Now, we will prove closedness. Let x X \A, and define:Un = X \ {y | d(x, y) 1n where n N} .

    We will claim that the Uns cover A, or that every a A is in the complementof some {y | d(x, y) 1n}. In other words, for some n, d(a, x) > 1n , orn > 1d(a,x) . This is clearly true by the Archimedean property of R, so ourclaim is true. Now, extract a finite subcover {U1, U2, . . . , UM}. Notice thatMi=1 Ui = UM , so A UM . Then, we know that:

    {y | d(x, y) 1M } X \A UMso it follows that B(x, 1M ) X \ A. Thus, because x was arbitrary, thereis an open ball around every point in X \A, and A is closed.

    A subset S X is bounded if there exists an a X and an R > 0 such that S B(a,R).32

  • Math 104, Spring 2011 Vaughan Jones Lecture 14

    Proposition. If (X, d) is a compact metric space, then any closed subset of X iscompact.

    Proof. Let A be closed, and letI U be an open cover of A. It is clear that

    {U | I} {X \A} is an open cover of X. By the compactness of X, we knowthat X must then also be covered by

    NnN UnX \A, for some N N. Then, it is

    clear thatNnN Un is a finite subcover of A, because X \A cannot cover anything

    in A. 14.3. Sequential Compactness.

    Definition. Given (X, d) and A X, A is sequentially compact if every se-quence in A has a convergent subsequence in A.

    Example: By the Bolzano-Weierstrass Theorem, every closed interval [a, b] is se-quentially compact. Furthermore, by extension, [a1, b1] [a2, b2] [an, bn] issequentially compact in Rn. To see this in R2, first get a convergent subsequencein the first coordinate, which we can do because if we consider just the sequence(an), it is a sequence in R. Then, fix that subsequence and get a convergent sub-subsequence in the second coordinate. Any subsequence of a convergent sequenceconverges, so both coordinates converge and we have a convergent subsequence inR2. This process can easily be generalized to any n.

    Recall that when we defined compactness, we said that the definition was veryhard to work with. This definition is much easier, so we would like for them to beequivalent. Lucky for us, they are!

    Theorem. Let (X, d) be a metric space. Then, X is compact if and only if X issequentially compact.

    Proof. In the next lecture.

    33

  • Math 104, Spring 2011 Vaughan Jones Lecture 15

    15. 03/08/11: Compactness, Part Two

    15.1. (Sequential) Compactness.Remember: E X is compact if every open cover of E has a finite subcover, andE X is sequentially compact if every sequence in E has a convergent subsequencein E.

    Theorem. Given a metric space (X, d), X is compact if and only if X is sequen-tially compact.

    Proof. We will go in both directions.

    () Let (xn) be a sequence in X. If {xn |n N} is finite, then some xn mustbe hit infinitely many times, so we would be done. Therefore, assume{xn} is infinite. Let x X. Either x is the limit point of a subsequenceof (xn), or there exists an r > 0 such that |B(x, r) {xn |n N}| 1.To see this, assume that the latter is not true for some x X. Then,for all r > 0, |B(x, r) {xn |n N}| > 1. In this case, we could takeever-decreasing values of r, and there would be at least one xn such thatd(x, xn) < r for every r > 0. Therefore, if you picked all of those terms,we would have a convergent subsequence of (xn). So, in other words, thereis a convergent subsequence of (xn) or

    xX B(x, r) forms an open cover

    for X. By compactness, we can obtain a finite subcover of this open cover,so only finitely many B(x, r) cover X, and hence {xn}. Then, because forevery x, |B(x, r) {xn |n N}| 1, the finiteness of this subcover wouldimply that {xn} was also finite, which is a contradiction.

    () Let {U} be an open cover of X.First, we will prove this statement:

    R > 0 s.t. x X,U s.t. B(x,R) U .Proceeding by contradiction, we will assume that no such R exists. Becauseno such R exists, in particular, for each n N, there exists an xn X suchthat B(xn,

    1n ) * U for any U. Now, because X is sequentially compact,

    we can get a convergent subsequence of this sequence which converges tosome x X, which we will denote (xnk) x. We know that this x Ufor some U, and because U is open, there exists an r > 0 such thatB(x, r) U. Then, by the convergence of (xnk), there is some N Nsuch that for all n N , the following are true:(1) 1n

    R2 . Then, pick another point x3

    outside of B(x1, R)B(x2, R) such that d(xi, x3) > R2 for i = 1, 2. Continuethis process, such that xn+1 is not in

    ni=1B(xi, R) and d(xi, xn+1) >

    R2

    for i = 1, 2, . . . , n, for all n N. Then, the latter statement is obviouslytrue for all subsequences of (xn). Therefore, this sequence cannot have anyconvergent subsequences, a contradiction of our initial assumption. Thus,it must be the case that finitely many balls of radius R cover X, so X mustbe compact.

    Theorem (Heine-Borel). A subset of Rn is compact if and only if it is closed andbounded.

    Proof. We will proceed in both directions.

    () We already proved this statement for an arbitrary metric in the last lecture.() Let A Rn be closed and bounded. By the boundedness of A, we know

    that A B(0, R + 1) for some R > 0. Then, it is clearly the case thatA [R,R][R,R] [R,R]. The right hand side of that expressionis sequentially compact (from last lecture) and hence compact (from thislecture), and we also proved last time that a closed subset of a compact setis also compact. Therefore, A is compact.

    From this theorem, we obtain some very useful results, such as that the Cantor setis compact, the unit sphere is compact, and so on.

    Now, we will build up to another major, but seemingly unrelated, result.

    15.2. The Equivalence of Norms on Rn.

    Proposition. Let X,Y be metric spaces such that X is compact. Then, given acontinuous function f : X Y , f(X) is compact.Proof. Let U be an open cover of f(X). Then, it is clear that f

    1(U) covers X.Because f is continuous, we know that for any open set U Y , f1(U) is also anopen set. Thus, f1(U) is an open cover of X, and by the compactness of X, forsome n N, we have that:

    X ni=1

    f1(Ui)

    This, in turn, implies that f(X) ni=1 Ui , so f(X) is compact. Proposition. Let X be a metric space. If X is compact, then any continuousfunction f : X R attains its minimum and maximum values.

    35

  • Math 104, Spring 2011 Vaughan Jones Lecture 15

    Proof. From the proposition above, we know that f(X) R must be compact, soit is therefore closed and bounded. Thus, sup f(X) and inf f(X) exist, and by theclosedness of f(X), must lie inside f(X) itself. Corollary. Let X be a compact metric space. Then, given a continuous functionf : X R such that f(x) > 0 for all x X, it is the case that there exists an > 0such that f(x) for all x X.Proof. This follows immediately from the above proposition. Theorem. Any two norms on Rn are equivalent, or in other words, given twonorms | | and || || on Rn, there exist A,B > 0 such that for all x Rn,

    A|x| ||x|| B|x| .Proof. In next lecture.

    36

  • Math 104, Spring 2011 Vaughan Jones Lecture 16

    16. 03/10/11: Norms, Separability, Connectedness

    16.1. Equivalence of Norms on Rn.

    Theorem. Any two norms on Rn are equivalent, or in other words, given twonorms || || and ||| ||| on Rn, there exist A,B > 0 such that for all x Rn,

    A||x|| |||x||| B||x|| .Proof. We will make two trivial observations to start. First, note that if we replace||| ||| with || ||2, we will obtain the same result. Second, the statement is obviouslytrue for any A,B if x = 0, so we will assume that x 6= 0.

    First, we will prove the first inequality. Let ei be a unit vector with a 1 in theith place. Then, x =

    ni=1 xiei. Then, we can obtain that

    ni=1

    xiei

    ni=1

    (|xi| ||ei||) n

    i=1

    |xi|2 n

    i=1

    ||ei||2

    by the triangle inequality, followed by the Cauchy-Schwarz inequality. Furthermore,ni=1 |xi|2 = ||x||2, so this implies

    1ni=1 ||ei||2

    ||x|| ||x||2

    which proves the first inequality. Observe that this result implies that || || isuniformly continuous for || ||2, or, in other words, that the identity function from(Rn, || ||) to (Rn, || ||2) is uniformly continuous (and vice-versa).

    Now, we will prove the second inequality. Let > 0 be given. Then, choose = A, where A is the term we found above. Then, if ||x y||2 < , then

    ||x y|| 1A||x y||2 < 1

    A =

    A

    A= .

    As noted above, this means that the identity function from (Rn, || ||2) to (Rn, || ||)is (uniformly) continuous. In particular, let Sn = {x Rn | ||x||2 = 1}. Then,clearly, the identity function from (Sn, || ||2) to (Rn, || ||) is also continuous. Weknow that for any x Sn, ||x||2 =

    ni=1 x

    2i = 1, so S

    n is obviously closed andbounded, and hence compact, by the Heine-Borel Theorem. Furthermore, || || is anorm, so ||x|| > 0 for all x Sn. By the theorem from last lecture, this means thatthere exists an > 0 such that ||x|| > for all x Sn. So, if we take an arbitraryv 6= 0 in Rn, then we have:

    =

    v||v||22

    0, there exists a > 0 such that for all x, y X,d(x, y) < D(f(x), f(y)) < .

    The difference between continuity and uniform continuity is that the in uniform continuity mustwork for all x, y.

    37

  • Math 104, Spring 2011 Vaughan Jones Lecture 16

    Now, we will discuss two final metric space topics before diving into the subjectof Calculus.

    16.2. Separability.

    Definition. Let (X, d) be a metric space. X is separable if there exists a count-able dense subset of X.

    Example: R is separable because R Q and |Q| = |N|.Proposition. Every subset of a separable metric space is separable with the inher-ited metric.

    Proof. Let (X, d) be a metric space, Q be the countable dense subset of X, andA X. Consider {B(q, 1n ) | q Q, n N}. This is a countable set, as Q N iscountable. For each q, n, if B(q, 1n ) A 6= , choose a point in it, which we willdenote xq,n. Then, let D = {xq,n | q Q, n N}. It is clear that D is countableand a subset of A. Therefore, we only need to show that it is dense, or in otherwords, given a A and > 0, that there exists an x D such that x B(a, ).Choose an n N such that 1n < 2 . Because Q is dense in X, there exists a q Qsuch that d(q, a) < 1n R, anxn does not converge.

    Proof. We will prove the two cases.

    (1) We want to show that |anxn| converges. We can write because |anxn| =

    |an||xn|, we must show that |an||xn| converges. We will use the nth root

    test.

    lim sup n|an||xn| = lim sup( n

    |an| |x|) = |x| lim sup n

    |an| = |x| .

    So, the power series converges if |x| < 1, or |x| < 1 = R.

    Note that can be infinite.R is known as the radius of convergence.

    43

  • Math 104, Spring 2011 Vaughan Jones Lecture 18

    (2) Assume x > 1. Then, working backwards from the chain of equalities

    above, we have that lim sup n|an||xn| > 1. If n|an||xn| > 1, then it is

    clearly the case that |an||x|n > 1. However, this would mean that the termsof the power series do not tend to zero, so the series as a whole does notconverge.

    Here are some very useful power series to know.

    ex = exp(x) =n=0

    xn

    n!

    sin(x) =n=0

    (1)n x2n+1

    (2n+ 1)!

    cos(x) =n=0

    (1)n x2n

    (2n)!

    11 x =

    n=0

    xn for |x| < 1.

    log(1 x) = n=1

    xn

    nfor |x| < 1.

    18.3. Differentiability.

    Let be a function f : (x, y) R where x, y R, and let a (x, y).Definition. f is differentiable at a if lim

    xaf(x)f(a)

    xa = L for some L R. Then,we will write L = f (a).

    However, we have not really defined what limxa f(x), or in other words, the limitof a function at a, means, so we will do that now, for general metric spaces.

    Definition. Let f : X Y where (X, d) and (Y,D) are metric spaces, and leta X and y Y . Then:

    limxa f(x) = y ( > 0)( > 0) s.t. 0 < d(x, a) < D(f(x), y) < .

    An alternate definition of this concept is that if limxa f(x) = y, then:

    F (x) =

    {f(x), if x 6= ay, if x = a

    is continuous.

    Theorem. If f is differentiable at a, then f is continuous at a.

    Note that this does not apply for general metric spaces! We are returning to the definitionsof f and a given in the very beginning of this section.

    44

  • Math 104, Spring 2011 Vaughan Jones Lecture 18

    Proof. We can prove this using the - definition of the limit of a function. However,we will try using the alternate F (x) method instead.

    Because we know that limxa

    f(x)f(a)xa = L for some L R, we can define

    F (x) =

    {f(x)f(a)

    xa , if x 6= aL, if x = a.

    where F (x) is continuous. We want to show that f(x) is continuous at a. Inparticular, we know that F (x) is continuous at x = a. Thus, for x 6= a, we have:

    F (x) =f(x) f(a)

    x a f(x) = f(a) + F (x)(x a) .Then, because F (x) is continuous at a, for any > 0, there exists a > 0 such that|a x| < |L F (x)| < . This means:

    |L F (x)| =L f(x)f(a)xa = |L(x a) + f(a) f(x)| (1)|L + f(a) f(x)| < .(2)

    In order to prove the continuity of f(x) at a, we must find some such that|a x| < |f(a) f(x)| < . We will proceed casewise, for when L is positiveand negative.

    Let L be positive and > 0 be given. Then, set = . Assuming |ax| < = ,we obtain:

    |f(a) f(x)| < |L + f(a) f(x)| < .Therefore, f(x) is continuous in this case.

    Now, let L be negative and > 0 be given. Using a simple change of variables,we will simply write L = L, and keep L > 0. Notice that, in the series ofinequalities (1) above, | L F (x)| = |F (x) + L|. Therefore, we can write (2) as|f(x) f(a) + L| < . Then, letting = , and assuming |a x| < , we obtain:

    |f(a) f(x)| = |f(x) f(a)| < |f(x) f(a) + L| < .Therefore, f(x) is also continuous in this case, completing our proof.

    Now, we will simply list some basic properties of derivatives. We will not provethese because the proofs are fairly elementary, and involve only simple limit compu-tations. However, the proofs are in the textbook and any other elementary calculustextbook. We will just assume that this theorem is true, and use its results liberally.

    Theorem. Let f, g be functions on an open interval in R and let a be in thatinterval. Assume that f, g are differentiable at a. Then, the following are true:

    (1) (cf)(a) = cf (a), where c is a constant.(2) (f + g)(a) = f (a) + g(a).(3) (fg)(a) = f(a)g(a) + f (a)g(a).

    (4) If g(a) 6= 0, then(fg

    )(a) =

    f (a)g(a) f(a)g(a)g(a)2

    .

    For the quotient rule (part 4 above), notice that because g(a) 6= 0 and g is contin-uous at a, there exists an open interval containing a such that g(x) 6= 0 for all x inthat interval, making the theorem in fact valid as stated.

    45

  • Math 104, Spring 2011 Vaughan Jones Lecture 19

    19. 03/29/11: Theorems About the Derivative

    19.1. Review.

    Suppose we have a function f : (a, b) R where (a, b) is an open interval. Iflimxa

    f(x)f(a)xa = L, then L = f

    (a), and f is differentiable on a. Equivalently,

    we can define a function F (x) such that F (x) = f(x)f(a)xa if x 6= a and F (x) = L ifx = a. Then, F is differentiable at a if and only if F is continuous at a.

    Example: f(x) = x2 where (x) is 1 if x Q and 0 if x / Q. f is differentiableat x = 0. Because all sorts of functions are differentiable at a single point, thisconcept isnt really useful. What we really want is differentiability in the wholeinterval.

    19.2. The Main Results.

    Theorem (Chain Rule). Let I and J be open intervals, and let f : I R andg : J R be functions such that if a I, then f(a) J . If f is differentiable at aand g is differentiable at f(a), then g f is differentiable at a, and

    (g f)(a) = g(f(a))f (a) .Note: This is just the familiar chain rule, or in other words, dydx =

    dydu dudx .

    Proof. Define a function h : I R such that:

    h(y) =

    {g(y)g(f(a))yf(a) , if y 6= f(a),

    g(f(a)), if y = f(a).

    Then, h(y) is clearly continuous because g is differentiable at f(a). Also, define afunction GF : I R such that:

    GF (x) =

    {g(f(x))g(f(a))

    xa , if x 6= a,g(f(a))f (a), if x = a.

    Now, it suffices to show that GF is continuous at x = a.

    Notice that (hf)(x) = g(f(x))g(f(a))f(x)f(a) . Then, we haveGF (x) = hf(x) f(x)f(a)xawhen x 6= a and GF (x) = g(f(a))f (a) when x = a. We know that (h f)(x) goesto g(f(a)) as x approaches a by the definition of h and f(x)f(a)xa approaches f

    (a)as x approaches a by the differentiability of f , so we are done.

    Theorem. Let I be an open interval in R, and let f : I R. Suppose x0 Iand that f(x0) is a maximum [or minimum] on I. If f is differentiable at x0, thenf (x0) = 0.

    Proof. Suppose without loss of generality that f (x0) > 0. Then, we will claim thatit is the case that for all > 0, there exists an x such that x0 < x < x + andf(x) > f(x0). Assume toward contradiction that this is not the case. Then, we

    have, for every , some x such that f(x) f(x0). This implies that f(x)f(x0)xx0 0, which furthermore implies that lim

    xx0f(x)f(x0)

    xx0 0, which is a contradictionbecause f (x0) > 0. Therefore, because there exists an x such that f(x) > f(x0),f(x0) cannot be a maximum. A similar argument can be used to see that f(x0)cannot be a minimum.

    46

  • Math 104, Spring 2011 Vaughan Jones Lecture 19

    Theorem (Rolle). Let f : [a, b] R be continuous and differentiable on (a, b), andsuppose f(a) = f(b). Then, there exists an x (a, b) such that f (x) = 0.

    Proof. By the compactness of [a, b], f attains its maximum (and minimum) on[a, b]. If the maximum is attained at an endpoint, then it must attain its minimuminside (a, b) or the function must be constant (as the minimum would then be equalto the maximum). In the former case, the previous theorem proves our claim, andin the latter, the derivative is 0 everywhere.

    Theorem (Mean Value). Let f : [a, b] R be continuous and differentiable on(a, b). Then, there exists an x (a, b) such that

    f (x) =f(b) f(a)

    b a

    Proof. Let t(x) = f(x) (f(a) + (x a) f(b)f(a)ba

    ). Notice that t(a) = 0 and

    t(b) = 0. Also, t(x) = f (x) f(b)f(a)ba . So, by Rolles Theorem, there exists an xsuch that 0 = f (x) f(b)f(a)ba , or f (x) = f(b)f(a)ba .

    Corollary. Let f : (a, b) R and suppose f (x) = 0 for all x (a, b). Then, f isa constant function.

    Proof. If p > q in (a, b), then 0 = f(p)f(q)pq , which implies that f(p) = f(q) by theMean Value Theorem.

    Corollary. If f : (a, b) R is differentiable and f (x) > 0 for all x (a, b), thenf is strictly increasing.

    Proof. Notice that f (x) = f(p)f(q)pq for some p > q in (a, b), so f(p) > f(q).

    Note that this corollary can be generalized to all of the variations of >, namely,

  • Math 104, Spring 2011 Vaughan Jones Lecture 19

    Definition. Given a function f C(a, b) such that a < 0 < b, the Taylor seriesof f is defined as

    n=0

    f (n)(0)xn

    n!

    This is clearly a power series, but we would like to know if it converges, what itconverges to, and if it actually has anything to do with the function. Luckily, wehave a Theorem that tells us just that.

    Theorem (Taylor). Suppose f has n derivatives on (a, b), where a < 0 < b. Then,for x (a, b), there exists a y between 0 and x such that

    f(x) = f(0) + f (0)x+f (0)x2

    2!+ + f

    (n1)(0)xn1

    (n 1)! +fn(y)xn

    n!

    Proof. Let M be the solution to the equation

    f(x) =

    n1k=0

    f (k)(0)xk

    k!+Mxn

    n!

    for some fixed nonzero x. Furthermore, define the function

    g(t) =

    n1k=0

    f (k)(0) tk

    k!+Mtn

    n! f(t)

    Notice that g(x) = 0. Furthermore, g(0) = 0 and g(k1)(0) = 0 for all k suchthat 2 k n. So, by Rolles Theorem, there exists an x1 between 0 and xsuch that g(x1) = 0. Then, again by Rolles Theorem, because g(0) = 0 andg(x1) = 0, there exists an x2 between 0 and x1 such that g(x2) = 0. Continuing,we get that there exists an xn1 between 0 and xn2 such that g(n1)(xn1) = 0.Finally, there exists a y between 0 and xn1 such that g(n)(y) = 0. However,g(n)(y) = M f (n)(y), so M = f (n)(y) for some y between 0 and x, proving thetheorem.

    Technically, the Maclaurin seriesThis means that if 0 < x, then 0 < y < x, and if x < 0, then x < y < 0.

    48

  • Math 104, Spring 2011 Vaughan Jones Lecture 20

    20. 03/31/11: Differentiation of Power Series

    20.1. Consequences of Taylors Theorem.

    Recall that Taylors Theorem states that for any n-times differentiable functionf(x), there exists a y between 0 and x such that:

    f(x) = f(0) +f (0)x

    1!+ + f

    (n1)(0)xn1

    (n 1)! +f (n)(y)xn

    n!

    A simple corollary of this theorem follows.

    Corollary. If there exists a bound on the set {f (n)(y) | y varies as n varies}, thenthe Taylor series of f converges to f .

    To see this, notice that if there is a bound on the set, as n increases, xn

    n! convergesto 0.

    Taylors Theorem is very powerful, as it is what allows us to compute exactvalues of differentiable functions!

    20.2. Term-by-Term Differentiation of Power Series.

    Theorem. Letanx

    n be a power series with radius of convergence R. Definef(x) =

    anx

    n on (R,R). Then, the following are true:(1)

    n=0 nanx

    n1 converges.(2) f is differentiable on (R,R).(3) f (x) =

    n=0 nanx

    n1 on (R,R).Before we get to the bulk of the Theorem, we will make a few easy observations.

    Recall that R1 = lim |an| 1n for any power series. So, for the series in part (1)above, we have R11 = lim(n|an|)

    1n . However, we can show that limn n

    1n = 1,

    so by the limit theorems, R = R1, so ifanx

    n converges, thennanx

    n1 alsoconverges.

    Also, the following identity is true for any a, b R such that a b 6= 0:(3)

    an bna b = a

    n1 + an2b+ + abn2 + bn1

    Notice that the right hand side is an expression with n terms.The following is the infinite analogue to the Triangle Inequality:

    (4)

    n=0

    cn

    n=0

    |cn|

    Finally, recall the very simple identity

    (5) 1 + 2 + + n = n(n 1)2

    Now, we are ready to start.

    Proof. We have already proved (1) in the first paragraph above, and proving (2)and (3) are basically equivalent.

    In order to prove (2) and (3), it suffices to show that

    (6) limyx

    n=0

    (an(y

    n xn)y x nanx

    n1) = 0

    49

  • Math 104, Spring 2011 Vaughan Jones Lecture 20

    If R is finite, then choose an r such that 0 < r < R, |x| < r < R, and |y| < r < R.If R is infinite, then this step is not necessary, and we can simply pick any r.

    Now, by (4), we have that (6) is equal to

    limyx

    n=0

    an(yn1 + yn2x+ yn3x2 + + yxn2 + xn1 nxn1)

    Because (3) is n terms long, we can split up the last nxn1 and distribute them toevery other term to obtain:

    limyx

    n=0

    an((yn1 xn1) + x(yn2 xn2)

    + x2(yn3 xn3) + + xn2(y x))

    Then, if we factor out a yx from each term and apply (3) to each one individually,we obtain

    limyx |y x|

    n=0

    an((yn2 + yn3x+ + xn2)

    + x(yn3 + yn4x+ + xn3) + + xn2)

    Notice that every term in the above expression is less than rn2 because |x| < r < Rand |y| < r < R and there are n 1 terms in the first parenthesized expression,n 2 in the second, and so on until there is only 1 term in the last expression.Therefore, by (5), the whole expression above is less than or equal to:

    limyx |y x|

    n=0

    |an|n(n 1)2

    rn2

    Now, remember that the radius of convergence ofn=0 bnx

    n is equal to B1, where

    B = n|bn|. It can be shown that nn(n1)2 converges to 1, so the coefficients of

    this power series converge simply to |an|. Therefore, because r < R, the wholesecond term is a convergent power series, so the entire term can converges as y goesto x, completing the proof.

    20.3. Elementary Functions.

    Now, we can define all of these nice functions weve previously seen, and provetheir familiar properties.

    exp(x) =n=0

    xn

    n!

    cos(x) =n=0

    (1)n x2n(2n)!

    sin(x) =n=0

    (1)n x2n+1(2n+ 1)!

    50

  • Math 104, Spring 2011 Vaughan Jones Lecture 20

    20.3.1. The exponential function is never negative. Suppose there exists an a suchthat exp(a) = 0. We can assume a < 0 because if a > 0, then we can reach acontradiction immediately through direct calculation. Consider the real number A,defined to be sup{a | exp(a) = 0}. Clearly, exp(A) = 0. Consider exp(A + x) forsome x R. It is easy to see by direct calculation that the derivative of exp(x) isexp(x). Therefore, by the Chain Rule, the derivative of exp(A+ x) = exp(A+ x).

    Now, consider the Taylor series expansion of exp(A+ x), or

    F (x) = exp(A+ x) = F (0) + F (0)x+ + F(n)(y)xn

    n!

    for some y between 0 and x. Notice that exp(A+x) is increasing by direct calcula-tion, and every derivative is exactly exp(A+ x). Therefore, f (n)(y) = exp(A+ x),so we have a bound on y, and the Taylor series converges to F . However, it isthe case that F (0) = exp(A + 0) = exp(A) = 0, so plugging that in to the Tay-lor series, we get that exp(A + x) = 0 for all x. This is clearly false, becauseexp(A+ (A)) = exp(0) = 1. Thus, we have a contradiction, and exp(x) 6= 0.20.3.2. Adding real numbers in the exp function. Consider the following equation.

    d

    dx

    (exp(x+ a)

    exp(x)

    )=

    exp(x+ a) exp(x) exp(x+ a) exp(x)(exp(x))2

    = 0

    By a corollary to the Mean Value Theorem, this implies that exp(x+a)exp(x) is constant.

    Therefore, we have exp(x + a) = K exp(x) for all x R and some K R. Letx = 0. Then, exp(a) = K, so exp(x+ a) = exp(a) exp(x), a familiar identity.

    20.3.3. The limit as exp approaches zero. exp(R) is connected because exp is acontinuous function, and the range does not contain 0, so the function is alwayspositive.

    Now, choose > 0 such that exp() < 1. Then, it is the case that, for anyn N,

    exp(n) = (exp())nwhich is arbitrarily small as n goes to . Therefore, limx exp(x) = 0.20.3.4. exp is a bijection from R to (0,). This follows immediately from theobservation that exp is strictly increasing.

    20.3.5. The log function. We can define log(x) as the inverse function of exp(x),where log has domain (0,). Therefore, log exp(x) = x, exp log(x) = x, andlog(xy) = log(x) + log(y). Also, it is the case that the derivative of log is 1x by theChain Rule on log exp(x).

    20.3.6. Arbitrary exponential functions and e. We will define ar = exp(r log(a))for a, r R such that a > 0. Also, we will define the real number e = exp(1).Therefore, we can see that ex = exp(x).

    20.3.7. The Pythagorean trigonometric identity. Consider the following equation.

    d

    dx(cos2(x) + sin2(x)) = 2 cos(x) sin(x) + 2 sin(x) cos(x) = 0

    Thus, as before, cos2(x) + sin2(x) is a constant function. Plugging in 0, we get thatcos2(x) + sin2(x) = 1.

    51

  • Math 104, Spring 2011 Vaughan Jones Lecture 20

    20.3.8. Pi. Because cos(1) > 0 and cos(2) < 0 by the error of the alternating seriesfunction, it is the case that there exists an x between 1 and 2 such that cos(x) = 0.We will define pi2 = inf{x |x > 0, cos(x) = 0}.

    There are many more properties that we can prove at this stage, but they areleft as exercises.

    20.4. Some Extra Tidbits.

    This section includes some counterexamples and odd functions related to differen-tiability, as well as another theorem.

    (1) limx0 sin 1x does not exist.(2) The function

    f(x) =

    {x sin 1x , if x 6= 00, if x = 0

    is continuous at 0. However, it is not differentiable at 0, as

    lim0

    f() f(0)

    = lim0

    sin 1

    = lim0

    sin1

    which does not exist.(3) The function

    g(x) =

    {x2 sin 1x2 , if x 6= 00, if x = 0

    is everywhere differentiable, as if x = 0, then lim02 sin 1

    = 0. However,

    consider its derivative. If x 6= 0, then dgdx = 2x sin 1x cos 1x , and if x = 0,then dgdx = 0, from the above calculation. Therefore, g

    (x) is not continuousat 0.

    Theorem. Let f be differentiable on (a, b), where a < x1 < x2 < b, and let c R.If f (x1) < c < f (x2), then there exists a y such that x1 < y < x2 and f (y) = c.

    Proof. Consider the function g(x) = f(x) cx. Clearly, g is differentiable on (a, b),and its derivative is f (x) c. [x1, x2] is compact, so the restriction of g to [x1, x2]attains its minimum and maximum. Therefore, for some y [x1, x2], g(y) = 0, orf (y) c = 0. If y was equal to either x1 or x2, then f (x1) = c or f (x2) = c,which contradict our assumptions. Thus, y must not equal either of those points,and we are done.

    This section was origi