Back to Lecture Notes List

Real Line

The real line R consists of all the real numbers x Î (- , ) and is synonimous with the collection of all decimals x = ± tmtm-1…t1t0Χ d1d2d3 … dn …, with the usual arithmetic operations of addition and multiplication, subtraction, division etc. Here ‘± ’ represents the sign of the decimal, ‘tmtm-1…t1t0’ the integer part and ‘d1d2d3…dn…’ the decimal or the fractional part. The left most digit tm is the most significant digit, tm-1 is the next most significant digit, and so on till t0, and then the decimal digits d1, d2, … in that order are the most significant ones. Note that the decimal part continues ad infinitum and that a terminating decimal stands for the number with the rest of the decimal digits assumed as zero. If at least one of the digits at the unit (t0), tenth (t1), hundredth (t2), … place, or at any of the first (d1), second (d2), … , n-th (dn) decimal places is non-zero a decimal with a ‘+’ sign represents a positive real number, and one with a ‘-’ sign a negative real number. If all of ti’s and di’s are zero the decimal represents the number zero. Given two decimals we also know which is larger. Let a, b Î R = (- , ) and b > a. Sub-intervals of the real line R are of the form: open (- , b), semi-open (- , b], open (a, ), semi-closed [a, ), open (- , ), open (a, b), semi-open (a, b], semi-closed [a, b), closed [a, b]. Of these the last four are finite intervals and the first five infinite intervals. We would use Z to denote the set of integers, and N for the set of non-negative integers, or the natural numbers. The real line R could be picturized as an infinite straight line with 0, and the units ± 1, ± 2, … marked on it. Given a point x on the real line one could find its associated decimal representation ± tm…t1Χ d1…dn …, and conversely. The absolute value |x| of the decimal x = ± tm…t1Χ d1 …dn … is the number tm…t1Χ d1…dn… . The integral part of the positive number tm…t1Χ d1…dn… is the decimal tm…t1Χ 0, or the number t0 + 10 t1 + 100 t2 + 10mtm.

A simple analysis of the familiar procedure for dividing a decimal number by another shows that if m and n (non-zero) are integers the decimal representation of m/n is recurring (i.e., a certain block of digits continues to recur indefinitely). Conversely, one easily sees that a recurring decimal represents a rational number m/n. Thus a number is irrational (i.e., it cannot be represented as a quotient m/n of integers) iff its decimal representation is non-recurring.

The real numbers satisfy: a + (b + c) = (a + b) + c (associativity of addition), a(bc) = (ab)c (associativity of multiplication), a + b = b + a (commutativity of addition), a b = ba (commutativity of multiplication), a(b + c) = ab + ac (distributivity of multiplication over addition), and the triangle inequality: |a+b| £ |a| + |b|, a, b Î R .

Exercise: Prove that: (i) |a-b| ³ |a| - |b|, in which the equality holds iff ab ³ 0 and |a| ³ |b|; (ii) max {a, b} = [(a+b) + |a-b|]/2; (iii) min {a, b} = [(a+b) - |a-b|]/2. (The results (ii) and (iii) imply that picking up the largest and the smallest numbers in a list could be done through arithmetic operations!).

A subest S of R is called upper bounded, or, bounded above if for some M Î R , x £ M for all x Î S. Such an M is called an upper bound for S. The least upper bound (lub) of a set S is the smallest upper bound of S, i.e., an M such that x £ M for all x Î S and such that for any e > 0 there is an x Î S such that x > M-e . A constructive procedure to show the existence of lub of an upper bounded set is as follows: Without loss of generality (i.e., by adding to each element a fixed positive number and then dividing by a high power of 10, if necessary) we could assume that all elements of S are less than 1 and that at least one element is greater than zero. Consider the subset of S of numbers having the largest first decimal digit. Call this digit as d1. Next consider the subset of this set having the largest second decimal digit. Call this digit as d2. Continue this procedure to get d3, d4, … , dn, … ad infinitum. It then is clear that M = 0Χ d1d2d3 … dn … is the least upper bound of S.

S is called lower bounded, or, bounded below if for some m Î R , x ³ m for all x Î S. Such an m is a lower bound for S. The greatest lower bound (glb) of a set S is the largest lower bound of S, i.e., an m such that x ³ m for all x Î S and such that for any e > 0 there is an x Î S such that x < m+e . Similar to the construction of lub we can constructively show that each lower bounded set has a glb.

S is called bounded if there is an M Î R such that |x| £ M for all x Î S. Thus, a bounded set is both upper and lower bounded, and conversely.

If S is a subset of R , sup S stands for lub of S if it is upper bounded and , otherwise. Similarly, inf S denotes glb of S if it is lower bounded and - otherwise. Thus sup S and inf S are meaningful for all subsets of R , whereas lub S and glb S make sense only if S is, respectively, upper and lower bounded.

For a subset S of R , max S denotes the largest, and, min S denotes the smallest element of S. Such elements need not exist. If S is an upper bounded subset of R , sup S denotes the smallest M such that for all x Î S, x £ M. If S is not bounded above we define sup S to be equal to + . If S is a lower bounded subset of R , inf S denotes the largest m such that for all x Î S, x ³ m. If S is not bounded below we define inf S to be equal to - . Note that given any e > 0, howsoever small, there is an element x in S such that x > sup S - e . Similarly, there is an element x in S such that x < inf S + e .

The considerations above could be summarized as the following:

Theorem. If S is a subset of R , then: (i) S has a least upper bound iff S is bounded above; (ii) S has a greatest lower bound iff S is bounded below; and (iii) S has a least upper bound as well as a greatest lower bound iff S is bounded.

Convergence of Sequences

Let xn, n ³ 1 be real numbers. The notation {xn}Ί {xn}n³ 1 is an abbreviation for a sequence {x1, x2, x3, x4, … }. (Note that the xn’s in a sequence {xn} need not be distinct, whereas the xn’s in a countable set S = {x1, x2, x3, x4, …} are supposed to be distinct). If n1 < n2 < n3 < … are natural numbers, the sequence {xnk}k³ 1 is called a subsequence of the sequence {xn}n³ 1.

{xn} is called bounded if there exists a constant M such that |xn| £ M for all n. It is called upper bounded if for some M, xn £ M for all n, and lower bounded if for some M, xn ³ M for all n.

A sequence {xn}nÎ N is called convergent, or is said to converge if there is a real number x such that given an arbitrary e > 0, there is an integer N such that |xn – x| < e , for all n > N. The number x is said to be the limit of the sequence and we write: limn® xn = x, or that xn ® x.

It is immediate from the definiton of the convergence of a sequence that {xn}n³ 1 is convergent iff {xn}n³ m is convergent where m ³ 1 is any natural number. What it means is that a finite number of terms of a sequence can be dropped from it without affecting the convergence or divergence of the sequence. It is also clear that every convergent sequence is bounded.

We call x a limit point of a sequence {xn} if given an arbitrary d > 0, xn Î (x-d , x+d ) for an infinity of n’s. Similarly, a is called a limit point of a set S if for each d > 0, the set SÇ (x-d , x+d ) has an infinity of points in it.

Exercise. (a) Prove that a sequence cannot have two limits. (b) Give an example of a seqence with an infinity of limit points. (c) Show that the limit points of the set of irrationals constitute the entire real line. (d) Give an example of an infinite set having no limit point. (e) Prove that every uncountable subset of the real line has a limit point.

If {xn} is an upper bounded sequence lim supn® xn denotes the largest of the limit points of the sequence. If the sequence is not bounded above lim supn® xn is taken to be + . If {xn} is a lower bounded sequence lim infn® xn denotes the smallest of the limit points of the sequence. If the sequence is not bounded below lim supn® xn is taken to be - .

The Bolzano-Weierstrass Principle. Every bounded sequence {xn}n³ 1 of real numbers possesses a convergent subsequence {xnk}k³ 1.

Proof: After division by a power of 10 and a translation we could assume the sequence {xn} to be contained in [0, 1). Consider the ten semi-closed subintervals I10 = [0, 10-1), I11 = [10-1, 2΄ 10-1) …, I19 = [9΄ 10-1, 1), each of length 10-1. An infinity of xn’s must lie in one of these intervals, I1d1, say. Pick one member xn1 from I1d1 and then divide I1d1 into 10 semi-closed subintervals I20, … , I29, each of lengh 10-2. One of these, I2d2, say, contains an infinity of the points of the sequence. Pick one member xn2, (n2 > n1), from I2d2 and then divide I2d2 into 10 semi-closed subintervals I30, … , I39, each of lengh 10-3. Continuing like this we get the subsequence {xnk} of {xn} convergent to the decimal x = 0Χ d1d2d3d4d5 … dk… . #

{xn} is called Cauchy if for each e > 0 there exists an integer N such that:|xn–xm| < e , for all m, n > N.

The Cauchy Convergence Criterion. A real sequence {xn} is convergent iff it is Cauchy.

Proof: A sequnce satisfying the Cauchy criterion is bounded. Hence by the Bolzano-Weierstrass principle there exists a convegent subsequence {xnk}, say, with limit x. Hence, there is an M such that |xnk – x| < e /2, k > M. Also, given an e > 0, there is an N such that |xn – xm| < e /2, for all m, n > N. Choosing k > M large enough so that nk > N, we have |xn – x| < |xn – xnk| + | xnk - x| < e /2 + e /2 = e , n > N. Conversely, let {xn} be convergent, say, {xn} ® x. Then given an e > 0, there is an N such that |xn – x| < e /2, for all n > N. Hence |xn – xm| £ |xn – x| + |x – xm| < e /2 + e /2 = e , for all m, n > N, i.e., {xn} is Cauchy. #

The Cauchy convergence criterion could be alternately stated as follows: {xn} converges iff for arbitrary n > m, |xn - xm| ® 0, as m ® . The property that every Cauchy sequence in R converges to a point in R is termed as the completeness of R .

Exercise. Prove that: (i) limn® xn exists iff lim infn® xn = lim supn® xn.; (ii) if xn £ x, lim supn® xn £ x.; (ii) if xn ³ x, lim infn® xn ³ x; (iii) if xn £ yn, lim supn® xn £ lim infn® yn; (iii) if xn £ yn £ zn, and lim supn® xn = lim infn® zn = c, then limn® yn = c (sandwich theorem).

Exercise. Recall the construction of the lub of an upper bounded set S above to show that if the lub S does not belong to S then there exists a sequence of elements of S converging to the lub. Show similarly that if the glb S, for a lower bounded set S, does not belong to S there exists a sequence of elements of S converging to glb S.

A sequence {xn} is called: (i) increasing if xn+1 > xn for all n, (ii) decreasing if if xn+1 < xn for all n, (iii) non-increasing if xn+1 £ xn for all n, (iii) non-decreasing if xn+1 ³ xn for all n, (iv) monotone if {xn} is either non-increasing or non-decreasing, and (v) strictly monotone if it is increasing or decreasing. {xn} L means {xn} is non-increasing and lim xn = L. {xn}­ L means {xn} is non-decreasing and lim xn = L.

Theorem. Every bounded monotone sequence {xn} is convergent.

Proof: Using the Bolzano-Weierstrass principle, onsider a convergent subsequence {xnk} ® L. Then given an arbitrary e > 0, there is a K such that | xnk - L| < e , k > K. Taking N = nK+1, irrespectively of whether {xnk} is non-decreasing or non-increasing, |xn - L| < e , n > N. Hece {xn} is Cauchy and so convergent. #

Alternately, if {xn} is monotone but not Cauchy, for some e > 0, there exist pairs (mk, nk), k ³ 1, such that mk < nk < mk+1 < nk+1, k ³ 1 and |xmk - xnk| ³ e . Then |xmk – xmk+p| ³ pe ® , as p ® . Hence {xn} is un bounded. #

Exercise. (i) Give an example of a divergent sequence which has a convergent sub-sequence. (ii) If a monotone sequence has a convergent sub-sequence show that the sequence itself is convergent. (iii) If {n1, n2, n3, … } denotes a permutation of the ordered set {1, 2, 3, … } of natural numbers and {xn} is a given sequence, prove that {xn} is convergent iff {xnk} is convergent.

Properties of Limits

Let limn® an = a, limn® bn = b, and limn® cn = c Ή 0. Then, (i) limn® (an+bn) = a+b, (ii) limn® (an-bn) = a-b, (iii) limn® (anbn) = ab, (iv) limn® (1/cn) = 1/c, and (v) limn® (an/cn) = a/c.

Proof: For (i) and (ii), given an e > 0, choose N such that |an-a|, |an-a| < e /2, n > N. Then |an+bn –(a+b)| £ |an-a| + |bn-b| < e /2 + e /2 = e , and |an-bn –(a-b)| £ |an-a| + |bn-b| < e /2 + e /2 = e , n > N. # For (iii) given an 1 > h > 0 choose N such that |an-a| < h , |bn-b| < h , n > N. Then, for n > N, |anbn – ab| £ |an–a||bn| + |a||bn – b| £ h (|b|+h ) + |a|h < h (|b|+1) + |a|h . Taking h = e /(|a|+|b|+1), we have |anbn – ab| < e , n > N. # For (iv) given an |c|/2 > h > 0 choose N such that |cn-c| < h , n > N. Note that this implies |cn| ³ |c|-h > |c|/2. Then, for n > N, |1/cn – 1/c| £ |cn–c|/(|c||cn|) < h /(|c||c|/2) = 2h /c2. Taking h = e c2/2, we have |1/cn – 1/c| < e , n > N. # Finally (v) follows from (iii) and (iv). #

Series

Let an be real numbers. Just like a sequence {an} is an infinite extension of an n-tuple {a1, a2, …, an}, an infinite sum or a series S n³ 1 an Ί a1 + a2 + … + an + … , is an infinite extension of the notion of a finite sum a1+a2+…+an of n terms. The sum sn = a1+a2+…+an is called the n-th partial sum of the series S n³ 1 an. To associate a possible meaning to a series, we call S n³ 1 an convergent if the sequence {sn}n³ 1 of its partial sums is convergent. The limit s = limn® sn, if it exists, is called the sum of the series. If limn® sn, does not exist the series is said to be divergent, i.e., if a series is not convergent it is divergent and then the notion of its sum remains undefined. If {sn} ® , one says that S n³ 1 an diverges to + . Similarly, if {sn} ® - , one says that S n³ 1 an diverges to - . A series which is divergent, but neither diverges to + nor to - , is called oscillatory.

If the associated series S n³ 1 |an| is convergent, the series S n³ 1 an is called absolutely convergent. If {bn}n³ 1 is a permutation of the sequence {an}n³ 1, S n³ 1 bn is called a re-ordering or a re-arrangement of the series S n³ 1 an. Each subsequence {snk}k³ 1 of {sn}n³ 1 defines a re-grouping S k³ 1 (ank-1+1 + … + ank), (n0 = 1), of the series S n³ 1 an. In testing the convergence of a series, the corresponding result on the sequences implies that a finite number of terms of a series can be dropped without affecting its convergence or divergence.

The Cauchy Convergence Criterion. Let sn denote the n-th partial sum of a series S n³ 1 an. The series conveges iff for each e > 0, there exists a N = N(e ) such that |am+1+…+an| = |sn-sm| < e for all n > m > N.

Proof: The result is immediate from the definition of convergence of a series and the Cauchy criterion for the convergence of a sequence. #

The Cauchy convergence criterion for a series is equivalent to the following statement: The series S n³ 1 an conveges iff for arbitrary n > m > N, |sn - sm| ® 0, N ® . In particular: if S n³ 1 an then an ® 0, n ® .

The Weierstrass’ M-Test. A series S an of positive terms is convergent if there is a constant M such that sn = a1+…+an < M, for all n ³ 1.

Proof: The result follows from the theorem that every bounded monotone sequence {xn} is convergent. #

Comparison Test. If S n³ 1 bn is a convergent series of positive terms and |an| £ bn, n ³ 1, then S n³ 1 an is convergent.

Proof: The result follows as |am+1 + … + an| £ |bm+1 + … + bn|. #

Corollary. If an, bn > 0, and an+1/an £ bn+1/nn, n ³ 1, then S an converges if S bn converges.

Proof: We have an = (an/an-1) (an-1/an-2)… (a2/a1)a1 £ (bn/bn-1) (bn-1/bn-2)… (b2/b1)a1 = (a1/b1)bn, n ³ 1. #

Root Test. Let q = limn® |an|1/n exist. Then, the series S n³ 1 an converges if q < 1, and diverges is q > 1. (If q = 1, the test is inconclusive).

Proof: If q < 1, we can choose e > 0 such that q+e < 1. Now, we can find a N such that |an|1/n < q+e , n > N. Then, for n > m > N, |sn – sm| £ |am+1| + … + |an| < (q+e )m+1 + … + (q+e )n < (q+e )N/[1-(q+e )] ® 0, N ® . Consequently S n³ 1 an converges. Next, if q > 1, we can choose e > 0 such that q-e > 1. Next, we can find a N such that |an|1/n > q-e , n > N. Hence, for n > N, |an| ³ (q-e )n ® , as n ® . Consequently, as an does not tend to zero as n ® , the series S n³ 1 an diverges. (For both the series S [n(n+1)]-1 (sn = 1-1/(n+1) ® 1, n ® , convergent), and S 1 (sn = n ® , divergent), q = limn® |an|1/n = 1, illustrating the inconclusive cases!).

Ratio Test. If q = limn® |an+1/an| exists, the series S n³ 1 an converges if q < 1, and diverges if q > 1. (If q = 1, the test is inconclusive).

For the series S 1/n and S 1/n2, both ratio and the root tests give q = 1. Hence, even though q = 1 for both the series, one of them diverges while the other converges.

Proof: If q < 1, we can choose e > 0 such that q+e < 1. Now, we can find a N such that |an+1/an| < q+e , n > N. Then, |aN+n| < |aN+1|(q+e )n-1, n ³ 1. Hence the series S n³ 1 aN+n obtained by dropping the first N terms, is dominated by the convergent geometric series S n³ 1 |aN+1|(q+e )n-1 implying the convergence of S n³ 1 an. If q > 1, we can choose e > 0 such that q-e > 1. Now, we can find a N such that |an+1/an| > q-e , n > N. Then, |aN+n| > |aN+1|(q+e )n-1 ® , as n ® . Thus as |an| does not tend to zero as n ® , S n³ 1 an diverges. (The inconclusiveness of the test in the case q = 1 is illustrated by the series S [n(n+1)]-1 and S 1). #

Integral Test. Let f(x) > 0 be a function defined on [1, ) such that f(x) £ f(y) if x > y and let an = f(n), n ³ 1. Then the series S an converges iff ò (1, ) f(x)dx < .

Proof: Since an ³ ò (n,n+1) f(x)dx ³ an+1, n ³ 1, sn ³ ò (1,n+1) f(x)dx ³ sn+1 – a1. Hence if {sn} is bounded ò (1, ) f(x)dx < , and if the integral is bounded {sn+1} is bounded. The test follows from the Weierstrass’ M-test. #

Examples. Taking f(x) = 1/x, the associated series and the integral are S 1/n, and ò (1, ) (1/x)dx = , hence S 1/n diverges. Taking g(x) = 1/x2, the series and the integrals are S 1/n2, and ò (1, ) (1/x2)dx = [-1/x](1, ) = 1 < , hence S 1/n2 converges. More generally, defining ln[0](x) = x, ln[1](x) = ln (ln[0](x), ln[2](x) = ln (ln[1](x), ln[3](x) = ln (ln[2](x), and so on, so that ln[q+1](x) = ln (ln[q](x), q ³ 0, and taking h(x) = 1/(ln[0](x) ln[1](x) ln[2](x)… ln[q-1](x) [ln[q](x)]p), the associated series is S n³ m (n ln (n) ln (ln n)… ln[q-1](n) [ln[q](n)]p)-1, and the associated integral is I = ò (m, ) (x ln x ln(ln x)… ln[q-1](x) [ln[q](x)]p)-1 dx. Putting t = ln[q](x), dt = d(ln (ln ( … (ln x) … ))) = [1/ln[q-1](x)] [1/ln[q-2](x)] … [1/ln(x)] [1/x] dx. Hence, I = ò (ln[q](m), ) t-pdt < iff p > 1. Hence the series converges if p > 1 and diverges for p £ 1.

Cauchy’s Condensation Test. If an+1 ³ an ³ 0, n ³ 1, S an converges iff S 2na2n is converges.

Proof: As s2n+1 - s2n ³ 2na2n ³ (s2n+2 - s2n+1)/2 summing for 0 £ n £ m-1, s2m – a1 ³ a1 + S 1£ n£ m-1 2na2n ³ (s2m+1 – s2)/2, showing that the sequence of partial sums of S an is bounded iff the sequence of partial sums of S 2na2n is so. The result follows. #

Leibnitz Test (Alternating series). If an 0, n ® , the series a1 – a2 + a3 – a4 + … converges.

Proof: For n > m, |sn-sm| = |am+1 - am+2 + … +(-1)n-man| Î [am+1-am+2, am+1]. Hence |sn-sm| ® 0, m ® . #

Theorem (Regrouping). If a series S an is convergent, every regrouped series is convergent with the same limit.

Proof: Since the sequence of partial sums converges, every subsequence of it also converges. #

Theorem (Rearrangement). If S an is absolutely convergent, every rearranged series is convergent with the same limit.

Proof: Let S denote the sum of the series. By the absolute convergence, given any e > 0, there exists an N0 such that |sm – S| = |a1 + … + am – S| < e /2, and |am+1| + … + |an| < e /2, for all n > m > N0. Now consider a rearranged series S bn. Since each term of S an is represented precisely once in S bn, fixing m as before, we can find an N such that for l > N all of a1, … , am occur amongst b1, … , bl. Also we can find n > m such that all of b1, … , bl occur amongst a1, … , an. Now, |b1 + … + bl - S| £ |b1 + … + bl – (a1 + … + am)| +| a1 + … + am – S| < |am+1| + … + |an| + e /2 < e /2 + e /2 = e . It follows that liml® b1 + … + bl = S, i.e., S bn converges with sum S. #

A convergent series which is not absolutely convergent is called conditionally convergent.

Theorem. If S an is conditionally convergent and s is any given number, there exists a rearrangement S bn of the series S an such that the series S bn is convergent and its sum is s .

Proof: Let {p n} and {n n} denote the subsequences of {an} of respectively positive and negative terms. To begin with the construction of S bn, in the first step, if s is positive keep on adding terms from {p n} till the sum is greater than s , and if s is negative keep on adding terms from {n n} till the sum is less than s . Subsequently, in the second step, if the sum is greater than s keep adding terms from {n n} till the sum is less than s , and if the sum is less than s keep adding terms from {p n} till the sum becomes greater than s . Continue this process ad infinitum by performing third, fourth, etc. steps. Let the last term added at the n-th step be called b n and the sum after the addition of b n as s n. If b n is positive 0 < s n - s £ b n, and if b n is negative b n £ s n - s < 0. Hence |s n - s | £ |b n|. Since b n ® 0, it follows that s n ® s , as n ® . Also if t n is any other sum during the n-th step |t n - s | < |s n-1 - s |, so that also t n ® 0 as n ® . Hence the sequence of the partial sums of S bn converges to s . #

Convergence of a Power Series. The complex power series S n³ 0 an(z-a)n converges if |z-a| < R, and diverges if |z-a| > R, with the radius of convergence R given by R = liminfn® |an|-1/n.

Proof: If R < and |z-a| < R, for some e > 0, |z-a| < R-2e . Now for some N, n > N implies |an|-1/n > R-e , i.e., |an| < 1/(R-e )n. Hence, |an(z-a)n| < [(R-2e )/(R-e )]n, n > N so that, after a certain number of terms, being dominated by a geometric series with common ratio (R-2e )/(R-e ) < 1, the series S n³ 0 an(z-a)n converges. Next, if R = , given any complex z there exists an N such that |an|-1/n > 2|z-a|, i.e., |an(z-a)n| < 2-n for all n > N. Hence, after a certain number of terms, being dominated by a geometric series with common ratio ½, the series

On the other hand, if |z-a| > R, for some e > 0, |z-a| > R+2e . Also, for an infinity of n’s |an|-1/n < R+e , i.e., |an| > 1/(R+e )n, so that |an(z-a)n| > [(R+2e )/(R+e )]n ® , n ® . Hence the series S n³ 0 an(z-a)n can not converge. #

Examples on Convergence on the Circle of Convergence (a) The power series S n³ 1 zn/n2 converges uniformly on the circle of convergence |z| = 1. (b) The geometric series S n³ 0 zn converges nowhere on the circle of convergence |z| = 1. (c) The power series

 

Continuous Functions

A real valued function f(x) defined in a neighborhood of a point x = a is said to be continuous at x = a if given an arbitrary e > 0, howsoever small, there exists a d > 0 such that |f(x)–f(a)| < e , for all {x : |x–a| < d }. Note that |f(x)–f(a)| < e , and |x–a| < d are, resp., equivalent to f(a) - e < f(x) < f(a) + e , and x Î (a-d , a+d ). Also note that if f is defined only on the left of x = a the condition 0 < |x–a| < d is assumed to be applicable only to the points x that satisfy x < a; similarly, if f is defined only on the right of x = a, the condition 0 < |x –a| < d applies only to x > a. If f(x) is continuous at each point of an interval I, f(x) is said to be continuous on the interval I. In the sequel a function would mean a real valued function, unless otherwise stated.

The definition of continuity of f(x) at x = a implies that if f(a) > 0 there is a neighborhood (a-d , a+d ), (d > 0) of the point x = a, in which f(x) is positive. For, with e = f(a)/2, the d obtained satisfies f(x) > f(a) - f(a)/2 = f(a)/2, x Î (a-d , a+d ). Similarly if f(a) is negative, choosing e = -f(a)/2 we have f(x) < f(a) + (-f(a)/2) =f(a)/2 < 0, x Î (a-d , a+d ), i.e., f(x) is negative in the entire neighborhood (a-d , a+d ).

Properties of Continuous Functions

Proof: Let f(x) be continuous at x = a. Then given any e > 0 there is a d such that |f(x)-f(a)| < e , |x-a| < d . Also there is an N such that |xn-a| < e , n > N. Hence |f(xn)-f(x)| < e , n > N. Conversely, if f(x) is not continuous at x = a, for some e > 0 for whatever d > 0 there is an x Î (a-d , a+d ) such that |f(x)-f(a)| ³ e . Choosing d = d n ® 0, and calling the corresponding point x as xn we have xn ® a for which f(xn) does not converge to f(a). #

Proof: By the properties of limits, the results follow from (a) f(xn) + g(xn) ® f(a) + g(a), (b) f(xn) – g(xn) ® f(a) - g(a), (c) f(xn)g(xn) ® f(a)g(a), and, (d) f(xn)/g(xn) ® f(a)/g(a), provided g(a) Ή 0. #

Proof: Let xn ® a. As zn = g(xn) ® g(a), f(g(xn)) = f(zn) ® f(g(a)). Hence f(g(x)) is continuous at x = a. #

Some more properties of continuous functions are listed in the following:

Theorem. Let f(x) be a continuous real valued function defined on a closed interval [a, b]. Then: (i) f(x) is bounded, i.e., the set {f(x) : x Î [a, b]} is a bounded subset of R ; (ii) f(x) attains its maxima, i.e., there is a point y Î [a, b] such that f(y) ³ f(x) for all x Î [a, b]; (iii) f(x) attains its minima, i.e., there is a point z Î [a, b] such that f(z) £ f(x) for all x Î [a, b]; (iv) f(x) is uniformly continuous, i.e., given an arbitrary e > 0, there exists a d > 0, such that |f(y)–f(z)| < e , for all y, z Î [a, b] satisfying |y–z| < d .

Proof: (i) If f(x) is not bounded there is no M such that |f(x)| < M for all x Î [a, b]. Hence there is a sequence of points xn Î [a, b] such that |f(xn)| ® . By the Bolzano-Weierstrass principle there is a subsequnce {xnk} converging to a point c Î [a, b]. But then, since f(xnk) diverges to infinity in magnitude it can not converge to the finite value f(c), f(x) cannot be continuous at x = c. #

(ii) By (i) S = {f(x) : a £ x £ b} is bounded and so has a lub M. If the lub M does not belong to S there is a sequence {xn} of points in [a, b] such that f(xn) ® M. As {xn} is a bounded sequence, it has a convergent subsequence {zk} ® z Î [a, b], say. Then, by the continuity of f(x) at x = z, M ¬ f(zk) ® f(z), k ® , so that f(z) = M. Hence, after all, M is attained at the point x = z. #

(iii) The proof could be given analogously to (ii): Let f(xn) ® inf {f(x) : x Î [a, b]. Since {xn} is a bounded sequence, it has a convergent subsequence {zn} ® z Î [a, b]. Then f(zn) ® f(z) and so the miniima is attained at x = z. Alternatively, it follows from (ii) by considering the function –f(x) instead of f(x), as the negative of the maxima of –f(x) is the minima of f(x). #

(iv) If f(x) is not uniformly continuous, there is an e > 0 for which no d > 0 works. Hence, there exists a sequence of pairs (xn, yn) of points in [a, b] such that |xn-yn| < 1/n and |f(xn)-f(yn)| ³ e . Since {xn} is a bounded sequence it has a convergent subsequence {x k} ® c, say. The associated subsequence {h k}, say, of the corresponding yn’s is also convergent and moreover converges to the same limit c. Letting k ® in |f(x k)-f(h k)| ³ e , we have 0 = |f(c)- f(c)| ³ e , a contradiction. #

Intermediate Value Theorem. Let f(x) be continuous on [a, b]. If f(a) Ή f(b) and g is a value in between them, there exists a point c Î (a, b) such that f(c) = g .

Proof: Considering the function ± [f(x)-g ] instead of f(x), if necessary, we could assume that f(a) < 0 = g < f(b). Let S = {x Î [a, b] : f(x) < 0}. S is non-empty and bounded. Let c = lub S. Then f(c) ³ 0, else there would be a point on right to c belonging to S. However, for any d > 0, there exists a point zd such that c-d < zd Î S, i.e.,f(zd ) < 0. Letting d ® 0, by continuity of f(x) at x = c there follows f(c) = limd ® 0+ f(zd ) £ 0. It follows that f(c) = 0. #

Derivative of a Function

The first derivative or the derivative of a function f at a point x is defined by

df(x)/dx Ί (d/dx)f(x) Ί f (x) = limh® 0 [f(x+h) – f(x)]/h.

If the limit does not exist we say that the derivative does not exist at the point x, or that f is not differentiable at the point x. The second derivative f² (x) is the derivative of the first derivative at a point x, i.e., f² (x) = (f ) (x). The existence of f² (x) implies that f (t) exists in a neighborhood of the point x, and that f (x+h) = f (x) + hf² (x) + o(h), h ® 0. In general, if g(x) = f(n)(x), the n-th derivative of f at x, then f(n+1)(x) = g (x). The zeroth order derivative f(0)(x), by convention, is taken to be f(x) itself.

The right and the left derivatives f +(x) and f -(x) at a point x = a are defined by f +(x) = limh® 0+ [f(x+h)–f(x)]/h and f -(x) = limh® 0- [f(x+h)–f(x)]/h, respectively. It is clear that f (x) exists iff both f +(x) and f (x) exist and are equal.

By a quantity q(h) being o(h), h ® 0, means that q(h)/h ® 0, as h ® 0. We say Q(h) = O(h), h ® 0, if there exists a d > 0 and a constant M such that |Q(h)/h| < M, |h| < d .

By definition, if f is differentiable at x,

f(x+h) = f(x) + f (x)h + o(h), h ® 0.

Similarly, the existence of f +(x) is equivalent to the relation

f(x+h) = f(x) + f +(x)h + o(h), h ® 0+,

and the existence of f -(x) is equivalent to the relation

f(x+h) = f(x) + f -(x)h + o(h), h ® 0-.

Let the functions f and g be differentiable at a point x, and let c be a constant. It is clear from the definition that

(f(x) ± g(x)) = f (x) ± g (x),

and that

(cf(x)) = cf (x).

Since g, being differentiable, is continuous at the point x, [f(x+h)g(x+h)-f(x)g(x)]/h = [(f(x+h)-f(x))/h]g(x+h) + f(x)[g(x+h)-g(x)]/h ® f (x)g(x) + f(x)g (x). it follows that the derivative of a product of two functions could be obtained by

(f(x)g(x)) = f (x)g(x) + f(x)g (x).

As a consequence, by induction,

(f1(x)f2(x) … fn(x)) = f1 (x)f2(x) … fn(x) + f1(x)f2 (x) … fn(x) + … + f1(x)f2(x) … fn (x).

Proof: For, assuming the result for n-1, (f1(x)f2(x)…fn(x)) = (f1(x)f2(x)…fn-1(x)) fn(x) + (f1(x)f2(x)…fn-1(x))fn (x) = f1 (x)f2(x) … fn(x) + f1(x)f2 (x) … fn(x) + … + f1(x)f2(x) … fn (x). #

Choosing f1(x) = f2(x) = … = fn(x) = x, and using (x) = 1, we get

(d/dx)xn Ί dxn/dx Ί (xn) = nxn-1, n = 1, 2, … .

If g(x) Ή 0, using the previous result, f (x) = [(f(x)/g(x))g(x)] = (f(x)/g(x)) g(x) + (f(x)/g(x))g (x), which on simplification gives the formula for the derivative of a quotient

(f(x)/g(x)) = [f (x)g(x) – f(x)g (x)]/[g(x)]2, (g(x) Ή 0).

If g is differentiable at the point x and f differentiable at the point g(x),

f(g(x+h)) - f(g(x)) = f (g(x))(g(x+h)-g(x)) + o(g(x+h)-g(x)).

Dividing by h and passing to the limit as h ® 0, we get

(f(g(x)) = f (g(x))g (x), (chain rule).

Leibnitz Formula. Let n be a natural number and let f(n)(x) and g(n)(x) exist. Then [f(x)g(x)](n) exists and is given by:

(d/dx)n[f(x)g(x)] = S 0£ k£ n nCkf(n-k)(x)g(k)(x).

Proof: The formula is true if n = 1. Hence assuming it for n, if f(n+1)(x) and g(n+1)(x) exist, and differentiating it

[f(x)g(x)](n+1) = S 0£ k£ n nCk[f(n+1-k)(x)g(k)(x)+f(n-k)(x)g(k+1)(x)]

= f(n+1)(x)g(0)(x) + S 1£ k£ n [nCk+nCk-1]f(n+1-k)(x) g(k)(x) + f(0)(x)g(n+1)(x)

= S 0£ k£ n+1 n+1Ckf(n+1-k)(x)g(k)(x), as nCk+nCk-1 = [nCk+nCk-1], 1 £ k £ n. #

L’Hospital’s Rule (I). If f(a), g(a) = 0, f (a), g (a) exist, and g (a) Ή 0, then

limx® a f(x)/g(x) = f (a)/g (a).

Proof: f(x)/g(x) = {f(a) + [f (a)+o(1)](x-a)}/{g(a)+ [g (a)+o(1)](x-a)} = [f (a)+o(1)]/[g (a)+o(1)] ® f (a)/g (a). #

Exercise 1. If f (x) exists, prove that f is continuous at the point x.

Exercise 2. If f (x) exists, prove that f(x+h) = f(x) + O(h), h ® 0, but that the converse need not be true.

Rules of Differentiation

Let f, g be functions differentiable at a point x, and let c be a constant. Then

(f ± g) (x) = f (x) ± g (x),

(cf) (x) = cf (x),

(fg) (x) = f (x)g(x) + f(x)g (x),

(f/g) (x) = [f (x)g(x) – f(x)g (x)]/[g(x)]2, (if g(x) Ή 0).

Note that if f (x) and g (x) exist

f(x+h)g(x+h) = [f(x) + f (x)h + o(h)][ g(x) + g (x)h + o(h)] = f(x)g(x) + [f (x)g(x) + f(x)g (x)] + o(h),

and

f(x+h)/g(x+h) – f(x)/g(x) = [f(x) + f (x)h + o(h)]/[ g(x) + g (x)h + o(h)] - f(x)/g(x)

= [f (x)g(x) – f(x)g (x)]/[g (x)]2 + o(h).

The first three relations follow similarly. #

Rolle’s Theorem. If f(x) is continuous real valued on [a, b] and differentiable on (a, b) and f(a) = f(b) = 0, there exists a x on (a, b) such that f (x ) = 0.

Proof: If f Ί 0, any point on (a, b) could be taken as x . If not without loss of generality we can assume that f takes a positive value. Then f assumes its maximum value at some point x Î (a, b), say. For h > 0, sufficiently small, we have f(x +h) – f(x ) £ 0, implying f (x ) £ 0. But also we have f(x -h) – f(x ) £ 0, implying f (x ) ³ 0. Hence f (x ) = 0. #

Lagrange’s Mean Value Theorem of Differential Calculus. If f(x) is continuous on [a, b] and differentiable on (a, b), there exists a point x Î (a, b) such that f(b) - f(a) = (b - a)f (x ).

Proof: Consider the function g(x) = f(x) – [f(a) + (x – a){f(b) – f(a)}/(b-a)]. It is continuous on [a, b], differentiable on (a, b), and it vanishes at x = a, b. Hence by Rolle’s theorem, there exists x Î (a, b) such that g (x ) = 0, i.e., f (x ) – [f(b) – f(a)]/(b-a) = 0. #

Cauchy’s form of the Mean Value Theorem. If f(x) and g(x) are continuous on [a, b], differentiable on (a, b), g does not vanish on (a, b), and g(b) Ή g(a), then there exists a point x Î (a, b) such that: [f(b)-f(a)]/[g(b)–g(a)] = f(x)/ g(x).

Proof: Consider the function h(x) = [f(x)–f(a)] [g(b)–g(a)] - [f(b)–f(a)] [g(x)–g(a)]. We have h(a) = h(b), h(x) is continuous on [a, b] and differentiable on (a, b). By Rolle’s theorem for some x Î (a, b), h (x) = 0, i.e., f (x )[g(b) – g(a)] = g (x)[f(b) – f(a)]. #

L’Hospital’s Rule (II). Let f(x) and g(x) be continuous in a neighborhood of x = a, and differentiable in a deleated neighborhood U = (a-d , a+d )\{a}. If f(a), g(a) = 0, then limx® a f(x)/g(x) = limx® a f (x)/g (x), provided the latter limit exists.

Proof: By the Lagrange’s form of the mean value theorem, f(x)/g(x) = [f(x)-f(a)]/[g(x)–g(a)] = f (x )/g (x ), x Î U, where x lies in between a and x and therefore tends to a as x tends to a. The result follows. #

Taylor Series Approximation Theorem. Let n ³ 1. If f(n)(a) exists,

f(x) = f(a) + (x-a)f (a) + (x-a)2f² (a)/2! + … + (x-a)nf(n)(a)/n! + o((x-a)n), x ® a.

Proof: By a repeated use of the L’Hospital’s rule,

limx® a [f(x) – {f(a) + (x-a)f (a) + (x-a)2f² (a)/2! + … + (x-a)nf(n)(a)/n!}]/[(x-a)n/n!]

= limx® a [f (x) – {f (a) + (x-a)f² (a) + (x-a)2f(3)(a)/2! + … + (x-a)n-1f(n)(a)/(n-1)!}]/[(x-a)n-1/(n-1)!]

= limx® a [f² (x) – {f² (a) + (x-a)f(3)(a) + (x-a)2f(4)(a)/2! + … + (x-a)n-2f(n)(a)/(n-2)!}]/ [(x-a)n-2/(n-2)!]

= …………………………………………………………………

= limx® a [f(n-2)(x) – {f(n-2(a) + (x-a)f(n-1)(a) + (x-a)2f(n)(a)/2!}]/ [(x-a)2/2!]

= limx® a [f(n-1)(x) – {f(n-1)(a)+ (x-a)f(n)(a)}]/ [(x-a)]

= limx® a [{f(n-1)(x) – f(n-1)(a)}/(x-a) - f(n)(a)] = 0. #

L’Hospital’s Rule (III). Let f(x) and g(x) be differentiable on (a, b) and |f(x)|, |g(x)| ® as x ® a+. Then, if limx® a+ f (x)/g (x) exists, limx® a+ f(x)/g(x) = limx® a+ f (x)/g (x).

Proof: Let e > 0 and L = limx® a+ f (x)/g (x). There exists a d > 0 such that |f (x)/g (x)-L| < e /2, if x Î (a, a+d ). Let 0 < h < d . It is clear that limx® a+ f(x)/g(x) = limx® a+ [f(x)-f(a+h)]/[g(x)-g(a+h)]. Hence there exists an 0 < h < h such that for a < x < a+h , |f(x)/g(x) - [f(x)-f(a+h)]/[g(x)-g(a+h)]| <e /2. By the Lagrange’s form of the mean value theorem for a < x < a+h , there exists a x satisfying x < x < a+h such that [f(x)-f(a+h)]/[g(x)-g(a+h)] = f (x )/g (x ) Î (L-e /2, L+e /2). Hence, |f(x)/g(x)–L| £ |f(x)/g(x)-[f(x)-f(a+h)]/[g(x)-g(a+h)] + [f(x)-f(a+h)]/[g(x)-g(a+h)] – L| < e /2 + e /2 = e , for all a < x < h . It follows that limx® a+ f(x)/g(x) = L. #

Taylor’s Expansion with Remainder

Let I[x, x+h] denote the closed interval with endpoints x and x+h, and I(x, x+h) the corresponding open interval. The result of the mean value theorem of differential calculus may be re-written as: f(x+h) = f(x) + hf (x ), where f is continuous on I[x, x+h] and differentiable on I(x, x+h), and x Î I(x, x+h). A generalization of the same is:

Taylor’s Expansion with Mean Value Form of Remainder. If f is continuous on I[x, x+h] and (n+1)-times differentiable on I(x, x+h), then

f(x+h) = f(x) + hf (x) + h2f² (x)/2! + … + hnf(n)(x)/n! + hn+1f(n+1)(x )/(n+1)!, x Î I(x, x+h).

Proof: Consider the function g(t) = f(x+t) – [f(x) + tf (x) + t2f² (x)/2! + … + tnf(n) (x)/n! + tn+1K/(n+1)!], where K is the constant that makes g(h) = 0. We have g(0) = g (0) = … = g(n)(0) = 0. Since g(h) = 0, for some x 1 Î (0, h), g (x 1) = 0. Hence, for some x 2 Î (0, x 1), g² (x 2) = 0. Continuing like this, there exists a x n+1 Î (0, x n), g(n+1)(x n+1) = 0. Renaming x = x + x n+1, we have K = f(n+1)(x ), completing the proof. #

Generalized Mean Value Theorem. If f(t), g(t) are continuous on I[x, x+h], h Ή 0, and are (n+1)-times differentiable on I(x, x+h) and g(n+1)(t) Ή 0, t Î I(x, x+h), for some x Î I(x, x+h)

[f(x+h) - S 0£ k£ n f(k)(x)hk/k!]/[g(x+h) - S 0£ k£ n g(k)(x)hk/k!] = f(n+1)(x )/g(n+1)(x ).

Proof: Let F(t) = [f(x+t) - S 0£ k£ n f(k)(x)tk/k!], and G(t) = [g(x+t) - S 0£ k£ n g(k)(x)tk/k!]. Note that F(i)(t) = [f(i)(x+t) - S 0£ k£ n-i f(k+i)(x)tk/k!] and so F(i)(0) = 0, 0 £ i £ n.Consider the function H(t) = F(t)G(h) – F(h)G(t). We have H(0) = H(h) = 0. By the mean value theorem H (x 1) = 0 for some x 1 Î ((0, h), i.e., F (x 1)G(h) – F(h)G (x 1) = 0. Now, for the function H (t) = F (t)G(h) – F(h)G (t). We have H (0) = H (x 1) = 0. By the mean value theorem, for some x 2 Î ((0, x 1), H² (x 2) = 0, i.e., F² (x 2)G(h) – F(h)G² (x 2) = 0. Continuing like this, for the function H(i)(t) = F(i)(t)G(x i) – F(x i)G(i)(t), as H(i)(0) = H(i)(x i) = 0, by the mean value theorem for some x i+1 Î ((0, x i), H(i+1) (x i+1) = 0, i.e., F(i+1)(x i+1)G(h) – F(h)G(i+1)(x i+1) = 0, 1 £ i £ n. Putting x = x+x n+1, the result for i = n+1 implies f(n+1)(x )G(h) – F(h)g(n+1)(x ) = 0. Note that by the mean value theorem G(h) = hn+1g(n+1)(h ) Ή 0. The result follows after division by the non-zero quantity G(h)g(n+1)(x ). #

Choosing g(t) = (t-x)n+1, as g(n+1) = (n+1)!, the previous result on the remainder in the Taylor’s series follows.

Power Series Expansion of Certain Functions

The function ln x is defined by: ln x = ò (1,x) (1/t)dt, x > 0. Since for 0 < x < 1, ò (1,x) Ί - ò (x,1), the above definition implies that: ln x = - ò (x,1) (1/x)dx, 0 < x < 1. Note that the change of variable s = 1/t leads to

ln (1/x) = ò (1,1/x) (1/t)dt = - ò (1,x) (1/s)ds = - ln x, x > 0.

The additive formula: ln xy = ln x + ln y, (x, y > 0), could be established as follows: Putting s = t/y,

ln xy = ò (1,xy) (1/t)dt = ò (1/y,x) [1/(sy)]yds = [ò (1,x) -ò (1,1/y)](1/s)ds = ln x – ln (1/y) = ln x + ln y. #

The defintion of ln x and the fundamental theorem of integral calculus implies that: d(ln x)/dx = 1/x, x > 0.

Note that the function ln x has been defined on the domain (0, ), and has range (- , ). Since (ln x) is positive throughout the domain ln x is increasing and therefore has its inverse function defined on the domain (- , ), with range (0, ). This inverse function is called the exponential function ex, i.e., ex = ln-1x, -  < x < . Differentiating the identity ln (ex) Ί x, (1/ex)d(ex)/dx = 1, i.e., d(ex)/dx = ex, - < x < . Put x = ea, y = eb in ln xy = ln x + ln y, to get ln eaeb = a + b. Taking the exponetial of both sides, we have eaeb = ea+b, which is the same as the multiplicative formula for the exponetials: ex+y = exey, - < x, y < .

Power Series Expansion of the logrithmic function

log (1+x) = x – x2/2 + x3/3 – x4/4 + … , -1 < x £ 1.

Here, f(x) = log (1+x), f (x) = 1/(1+x), f² (x) = -1/(1+x)2, f (x) = 2!/(1+x)3, … , f(n)(x) = (-1)n-1(n-1)! /(1+x)n, … . Hence, log (1+x) = S 1£ k£ n [(-1)k-1(k-1)!/(1+x)k|x=0]xk/k! + Rn(x) = S 1£ k£ n (-1)k-1xk/k + Rn(x), where Rn(x) is the remainder term in the Taylor’s expansion. For 0 < x £ 1, Rn(x) = [(-1)n[x/(1+x )]n+1/(n+1). Since for 0 < x £ 1, 0 < x < x, and so x/(1+x ) < 1, the remainder term (-1)n [x/(1+x )]n/(n+1) ® 0, as n ® . Hence the expansion is valid on (0, 1]. For –1 < x £ 0, we use the relation

log (1+x) = ò (0,x) [1/(1+t)]dt = ò (0,x) [1 – t + t2 – t3 + … + (-1)n-1tn-1 + (-1)ntn/(1+t)]dt = S 1£ k£ n (-1)k-1xk/k + Rn(x),

from which |Rn(x)| = |ò (0,x) [(-1)ntn/(1+t)]dt| = ò (0,|x|) [tn/(1-t)]dt £ [1/(1-|x|)]|x|n+1/(n+1) ® 0, as n ® . #

Power Series Expansion of the exponential function

ex = 1 + x + x2/2! + x3/3! + … , - < x < .

Here, f(x) = ex, f (x) = ex, f² (x) = ex, f (x) = ex, … , f(n)(x) = ex, … . Hence, ex = S 1£ k£ n-1 [ex|x=0]xk/k! + [ex ]xn/n! = S 1£ k£ n-1 xk/k + ex xn/n!. As, n! = Ö (2p ) nn+1/2e-n, for each - < x < , the remainder goes to zero as n ® . Alternately, if 2|x| £ m, |x|n/n! £ (|x|m/m!)2-(n-m) ® 0, n ® . #

Since the power series for ex converges absolutely for all x, the muliplicative rule ex+y = exey could also be deduced by a simplification of the product of the two power series for the right hand side:

exey = (S n³ 0 xn/n!) (S m³ 0 ym/m!) = S l³ 0 S 0£ n£ l (xn/n!)[yl-n/(l-n)!] = S l³ 0 [S 0£ n£ l lCnxnyl-n]/l! = S l³ 0 (x+y)l/l! = ex+y. #

An analysis similar to that for ex is valid in establishing the expansions:

sin x = x – x3/3! + x5/5! – x7/7! + … , - < x < ;

cos x = 1 – x2/2! + x4/4! – x6/6! + … , - < x < .

The complex exponential function eiq :

The function eiq may be defined by: eiq = cos q + i sin q , - < q < ,

so that: e-iq = cos q - i sin q , cos q = (eiq + e-iq )/2, sin q = (eiq - e-iq )/(2i), - < q < . Using i2 = -1, and the expansions for sin q and cos q , we have

eiq = 1 + iq + (iq )2/2! + (iq )3/3! +(iq )4/4! + … + (iq )n/n! + … , - < q < + ,

a comparision of which with the expansion ex = 1 + x + x2/2! + x3/3! + … justifies the notation eiq .

The arctan function

arctan x Ί tan-1x = x - x3/3 + x5/5 – x7/7 + … , |x| £ 1.

Here, f(x) = tan-1x, f (x) = 1/(1+x2). In this case, instead of a direct approach, we integrate the geometric series of 1/(1+x2) with remainder: 1/(1+x2) =1 – x2 + x4 – x6 + … + (-1)nx2n + (-1)n+1 [x2n+2/(1+x2)], integrating which on (0, x) gives tan-1x = x – x3/3 + x5/5 – x7/7 + … + (-1)nx2n+1/(2n+1) + (-1)n+1 ò (0,x) [x2n+2/(1+x2)]dx. The magnitude of the remainder does not exceed ò (0,|x|) x2n+2dx = |x|2n+3/(2n+3) ® 0, n ® , if |x| £ 1. #

 

 

Maxima-minima of Differentiable Functions

A function f(x) is said to have a local maxima at x = a, if there is a d > 0 such that f(a) ³ f(x), |x-a| < d . The point x = a is said to be a strict local maxima if there is a d > 0 such that f(a) > f(x), |x-a| < d .

Now suppose f(x) has a local maxima at x = a. By definition then f +(x) £ 0 and f -(x) ³ 0 so that if f (x) exists at x = a, we must have f (a) = 0. Hence if f(x) is differentiable at x = a, the condition f (a) = 0 is necessary for f(x) to have a maxima at x = a.

One could also reason as follows: if f (x) < 0, by f(x) = f(a) + (x-a)f (a) + o(x-a), x ® a, one concludes that there exists a d > 0, such that for all x satisfying |x-a| < d , there holds f(x) < f(a) if x > a and f(x) > f(a) if x < a, contradicting that x = a is a local maxima. Similarly, if f (x) > 0, one concludes that there exists a d > 0, such that for all x satisfying |x-a| < d , there holds f(x) > f(a) if x > a and f(x) < f(a) if x < a, contradicting that x = a is a local maxima.

A function f(x) is said to have a local minima at x = a, if there is a d > 0 such that f(a) £ f(x), |x-a| < d . The point x = a is said to be a strict local minima if there is a d > 0 such that f(a) < f(x), |x-a| < d . As in the case of local maxima, if f (a) exists also a local minima at x = a implies that f (a) = 0.

Hence, if f(x) has a local maxima or a local minima at x = a and if f (a) exists then f (x) = 0, i.e., if f(x) is differentiable at x = a, f (a) = 0 is a necessary condition for f(x) to have a local maxima-minima at x = 0. Thus, the points of local maxima-minima of f(x) are to be found amongst those x’s that satisfy the equation f (x) = 0.

Note that for f(x) = x3, f (0) = 0 but x = 0 is not a local maxima-minima point of f(x).

Theorem. If for some m ³ 1, f(2m)(a) exists and is not equal to zero and f (a) = … = f(2m-1)(a) = 0, then x = a is a local maxima of f(x) if f(2m)(a) < 0, and, x = a is a local minima of f(x) if f(2m)(a) > 0.

Proof: As, f(x)-f(a) = (x-a)f (a)+(x-a)2f² (a)/2!+…+(x-a)2mf(2m)(a)/(2m)!+o((x-a)2m) = [f(2m)(a)+o(1)](x-a)2m/(2m)!, x ® a, so that there exists a d > 0 such that for 0 < |x-a| < d , f(x) – f(a) < 0 if f(2m)(a) < 0 and f(x) – f(a) > 0 if f(2m)(a) > 0, proving the result. #

Theorem. If for some m ³ 1, f (a) = … = f(2m-1)(a) = 0, and f(2m)(x) exists in a deleted neighborhood of x = a and is negative (positive), then x = a is a local maxima (minima) of f(x).

Proof: The result follows from

f(x) = f(a) + (x-a)f (a) + (x-a)2f² (a)/2! + … + (x-a)2m-1f(2m-1)(a)/(2m-1)! + (x-a)2mf(2m)(x )/(2m)!, x Î I(a, x). #

Theorem. If for some n ³ 1, f (a) = … = f(n-1)(a) = 0, and f(n)(x) exists in a deleted neighborhood of x = a and the quantity (x-a)n f(n)(x) is negative (positive) there, then x = a is a local maxima (minima) of f(x).

Proof: The result follows from

f(x) = f(a) + (x-a)f (a) + (x-a)2f² (a)/2! + … + (x-a)n-1f(n-1)(a)/(n-1)! + (x-a)nf(n)(x )/(n)!, x Î I(a, x),

since x -a and x-a have the same sign. #

A minima or a global minima of f(x) is the largest number m such that f(x) ³ m for all x in the domain of definition of f. If there exists a point a such that f(a) = m, x = a is called a minima point of f and we say that f(x) attains its minima. The maxima or global maxima, similarly, is the smallest number M such that f(x) £ M. A maxima point a satisfies f(a) = M, and we say that f attains its maxima. The maxima minima points are indeed also local maxima and minima points, without the converse being necessarily true.

Intermediate Value Theorem for a Derivative. Let f(x) be differentiable on [a, b]. If f +(a) Ή f -(b) and g is a value in between them, there exists a point c Î (a, b) such that f (c) = g .

Proof: Using the function ± [f(x)-g x] instead of f(x), if necessary, we could assume that f +(a) < 0 = g < f -(b). Then f(x) decreases at x = a and increases at x = b so that neither of these points can be a minima point of f(x). Let the minima point be c. Then, a < c < b and f (c) = 0. #

One could contrast the statement of intermediate value theorem for a function f(x) with that for a derivative f (x): whereas the former requires the continuity of the function f(x), in the latter the continuity of the derivative f (x) is not needed.

Back to Lecture Notes List