Quick viewing(Text Mode)

L'hôpital's Rules and Taylor's Theorem for Product Calculus 1

L'hôpital's Rules and Taylor's Theorem for Product Calculus 1

L’Hˆopital’s Rules and Taylor’s Theorem for Product Alex B. Twist Michael Z. Spivey University of Puget Sound Tacoma, Washington 98416

1 Introduction

This paper is a continuation of the second author’s undergraduate-level survey [4] of product calculus, a variation of the standard calculus. In [4] the second author defines the product and , describes rules for finding product and , discusses applications, and proves product versions of some important theo- retical results in calculus, including a fundamental theorem. Other works on product calculus include Dollard and Friedman [2] and Grossman [3]; more can be found in the references in [2] and [4]. The major contributions of this article are product calculus versions of l’Hˆopital’s rules and Taylor’s theorem. In the process we prove an exponentiation rule, a , and a generalized for the product derivative. The former two imply rules for product and product integration by substitution, respectively. We need the definitions of the product derivative and integral and a theorem from [4]. The product derivative of a f at x is defined as

1   ∆x 1 ∗ f(x + ∆x) df dx = f (x) = lim . ∆x→0 f(x) We denote the nth product derivative of f by f [n]. The product integral of f > 0 over [a, b] is as follows: Let P = x0, x1, . . . , xn be a partition of [a, b] and let ck be any point on the interval [xk−1, xk]. Let ∆xk = xk − xk−1 and let kP k be the maximum value of ∆xk. Then the Riemann product integral of f over [a, b] is n b dx Y ∆xk Paf(x) = lim f(ck) , kP k→0 k=1 provided the exists and is independent of the choice of P and the ck’s. Finally, we have the following relating the product derivative and the usual Newto- nian derivative. Theorem 1. (Spivey 2006) If f(x) 6= 0, then f ∗(x) exists and is nonzero if and only if f 0(x) exists, in which case   d 0 f ∗(x) = exp ln |f(x)| = ef (x)/f(x). dx

1 In Section 2 we prove some additional product differentiation rules, which lead to additional product integration rules. Section 3 contains our results on l’Hˆopital’s rules using the product derivative. Taylor’s theorem with the product derivative is given in Section 4.

2 More on Product Calculus

Our results on l’Hˆopital’s rules and Taylor’s theorem require some results concerning the product derivative. Here we have the exponential rule, a parallel to the of Newtonian calculus. Theorem 2. (Exponential Rule) If f is a product differentiable function of x with f(x) > 0 and f ∗(x) > 0 and g is a differentiable function of x then

1 0 d f(x)g(x) dx = f ∗(x)g(x)f(x)g (x). Proof. 1 1 f(x + ∆x)g(x+∆x)  ∆x d f(x)g(x) dx = lim ∆x→0 f(x)g(x) 1 ! f(x + ∆x)g(x+∆x)  ∆x = lim exp ln ∆x→0 f(x)g(x)  g(x + ∆x) ln(f(x + ∆x)) − g(x) ln f(x) = exp lim ∆x→0 ∆x d (g(x) ln f(x)) = e dx 0 g(x) f (x) +g0(x) ln f(x) = e f(x)

 0 g(x) g0(x) = ef (x)/f(x) eln f(x) = f ∗(x)g(x)f(x)g0(x), via Theorem 1. Just as the product rule for Newtonian calculus yields the technique of integration by parts, the exponential rule for product calculus produces a product integration by parts. Corollary 1. (Product Integration by Parts) Let u be a product differentiable and prod- uct integrable function of x with u(x) > 0 and u∗(x) > 0. Let v be a differentiable function of x. Then v 1 vdx u P(du dx ) = .  dv dx P u dx

2 Proof. This follows directly from Theorem 2. We can retain a commutative-like property in Corollary 1 by letting v(x) = ln y(x). y0(x) ∗ ln y ln x Then, using the facts that y(x) = ln y (x) (Theorem 1) and x = y , Corollary 1 becomes

ln y ln u 1 vdx 1 ln y dx u y P(du dx ) = P(du dx ) = = . Puln(y∗(x))dx Py∗(x)ln u dx We also have a product chain rule. Theorem 3. (Product Chain Rule) Let f be a product differentiable function of u with f ∗(u) > 0, and let u be a differentiable function of x. Then

1 ∗ u0(x) df dx = f (u) .

Proof. First consider the composition of our functions, f(u(x)). Define the function

1  f(u(x)+t) t ( f(u(x)) ) 1 , t 6= 0; g(t) = df du 1, t = 0.

Note that as t → 0, the numerator of g(t), t 6= 0, is a product derivative. Thus, g(t) is continuous about t = 0. Now, letting t = ∆u 6= 0, we have

1  f(u(x)+∆u)  ∆u f(u(x)) g(∆u) = 1 df du

1 ∆u f(u(x) + ∆u) ⇔ (g(∆u)df du ) = . f(u(x)) Now, considering the product derivative of f(x), we can write

1   ∆x 1 f(u(x + ∆x)) df dx = lim ∆x→0 f(u(x)) 1 f(u(x) + ∆u) ∆x = lim ∆x→0 f(u(x))

1 where ∆u = u(x + ∆x) − u(x). Via substitution, we can express df dx as

1 1 ∆u 1 df dx = lim ((g(∆u)df du ) ) ∆x ∆x→0 1 ∆u = lim (g(∆u)df du ) ∆x ∆x→0

3 ∆u  1 lim∆x→0 ∆x = lim g(∆u) lim df du ∆x→0 ∆x→0 1 du = (df du ) dx = f ∗(u)u0(x).

The product chain rule leads directly to a product integration by substitution tech- nique, just as with the usual Newtonian calculus. We present the indefinite version; the proof of the definite version is similar.

Corollary 2. (Indefinite Integration by Substitution) Let g be a product integrable func- tion of u, and let u be a differentiable function of x. Let f(x) = g(u(x))u0(x). Then

dx u0(x) dx du Pf(x) = P(g(u(x)) ) = Pg(u) = C0G(u), where G(u) is a product of g(u).

Proof. If G is a product antiderivative of g, then G∗(u) = g(u), and by the product chain rule (Theorem 3) we have

u0(x) 1 du 1 f(x) = g(u(x)) = (d[G(u(x))] du ) dx = d[G(u(x))] dx .

Product integrating, we have

dx u0(x) dx 1 dx Pf(x) = P(g(u(x)) ) = P(d[G(u(x))] dx ) = C0G(u).

We now provide a generalized version of the mean value theorem for product deriva- tives presented by Spivey [4]. This theorem is pivotal to proving the product calculus versions of l’Hˆopital’s rules and Taylor’s theorem.

Theorem 4. (Generalized Mean Value Theorem for Product Derivatives) Let f be a function continuous on [a, b] and product differentiable on (a, b). Let g be continuous on [a, b] and differentiable on (a, b). Then, if g0(x) 6= 0 ∀x ∈ (a, b), f(a) 6= 0, and f(b) 6= 0, 1  f(b)  g(b)−g(a) 1 = f ∗(c) g0(c) f(a) for some c such that a < c < b.

4 g(x)−g(a)  f(x)   f(a)  g(b)−g(a) Proof. Let F (x) = f(a) f(b) . We have F (a) = F (b) = 1; thus we may appeal to Rolle’s theorem for product calculus from Spivey [4] to conclude that F ∗(c) = 1 for some c ∈ (a, b). So we have

g0(c) f(a) g(b)−g(a) 1 = F ∗(c) = f ∗(c) f(b)

1 1 1 f(a) g(b)−g(a) ⇒ 1 g0(c) = f ∗(c) g0(c) f(b) 1  f(b)  g(b)−g(a) 1 ⇒ = f ∗(c) g0(c) . f(a)

3 L’Hˆopital’s Rules

f(x) L’Hˆopital’s rules are useful for evaluating indeterminate limits in the form limx→a g(x) where, as x → a, either both f(x) → 0 and g(x) → 0 or f(x) → ∞ and g(x) → ∞. L’Hˆopital’s rules can also be used to evaluate limits that result in the indeterminate forms ∞0 and 1∞; however, doing so requires that the limiting expression first be put in 0 ∞ the form 0 or ∞ by appropriate algebraic manipulations. Using the product derivative, though, yields explicit l’Hˆopital-type rules for these indeterminate forms. First we need an alternate characterization of a finite limit.

Lemma 1. We have limx→c f(x) = L > 0 if and only if for every  > 1 there exists 1 f(x) δ > 0 such that if 0 < |x − c| < δ then  < L < . − f(x)  Proof. (⇐) Given  > 0, ∃ δ > 0 such that if 0 < |x − c| < δ, then e < L < e . Thus we have − < ln f(x) − ln L <  ⇒ lim ln f(x) = ln L x→c ⇒ lim eln f(x) = eln L x→c ⇒ lim f(x) = L. x→c Showing the converse entails a similar argument. We now present our two l’Hˆopital-type rules using the product derivative. As is the case with many results in [4] the proofs are not all that dissimilar from existing proofs for the usual versions of l’Hˆopital’s rules. (See, for example, Bartle and Sherbert [1, 178].)

5 Theorem 5. (L’Hˆopital’sRules with Product Derivatives, Part I) Let −∞ ≤ a < b ≤ ∞. Let f be a continuous and product differentiable function on (a, b). Let g be a differentiable function on (a, b) such that g0(x) 6= 0 ∀x ∈ (a, b). Suppose that limx→a+ f(x) = 1 and limx→a+ g(x) = 0.

1 1 ∗ g0(x) + g(x) a)If limx→a+ f (x) = L ∈ R , then limx→a+ f(x) = L. 1 1 ∗ g0(x) g(x) b)If limx→a+ f (x) = L ∈ {0, ∞}, then limx→a+ f(x) = L.

Proof. If a < α < β < b, then by Rolle’s theorem, g(α) 6= g(β). Further, by the generalized mean value theorem for product derivatives, ∃u ∈ (α, β) such that

1 1 f(β) g(β)−g(α) f ∗(u) g0(u) = . f(α)

Case (a): If L ∈ R+, given some  > 1, ∃ c ∈ (a, b) such that

L 1 < f ∗(u) g0(u) < L ∀ u ∈ (a, c),  from which we have

1 L f(β) g(β)−g(α) < < L ∀ α, β ∈ (a, c], α < β.  f(α)

Now taking the limit as α → a+, we have

L 1 < f(β) g(β) < L ∀ β ∈ (a, c].  Since  > 1 is arbitrary, we have our desired result. Case (b): If L = +∞ and given M > 0, then ∃ c ∈ (a, b) such that

1 f ∗(u) g0(u) > M ∀ u ∈ (a, c), from which we have

1 f(β) g(β)−g(α) > M ∀ α, β ∈ (a, c], α < β. f(α)

If we take the limit as α → a+, we have

1 f(β) g(β) > M ∀ β ∈ (a, c].

Since M > 0 is arbitrary, the assertion follows. If L = 0, the proof is similar.

6 Theorem 6. (L’Hˆopital’sRules with Product Derivatives, Part II) Let −∞ ≤ a < b ≤ ∞. Let f be a continuous and product differentiable function on (a, b). Let g be a differentiable function on (a, b) such that g0(x) 6= 0 ∀x ∈ (a, b). Suppose that limx→a+ g(x) = ±∞.

1 1 ∗ g0(x) + g(x) a)If limx→a+ f (x) = L ∈ R , then limx→a+ f(x) = L. 1 1 ∗ g0(x) g(x) b)If limx→a+ f (x) = L ∈ {0, ∞}, then limx→a+ f(x) = L.

Proof. Suppose that limx→a+ g(x) = ∞ (the case for −∞ is similar). If a < α < β < b, by Rolle’s theorem, g(α) 6= g(β). Further, by the generalized mean value theorem for product derivatives (Theorem 4), ∃ u ∈ (α, β) such that

1 f(β) g(β)−g(α) 1 = f ∗(u) g0(u) . f(α) Case (a): If L > 1, given some  > 1, then ∃ c ∈ (a, b) such that

L 1 < f ∗(u) g0(u) < L ∀ u ∈ (a, c).  Thus we have 1 L f(β) g(β)−g(α) < < L ∀α, β ∈ (a, c], α < β.  f(α) Since g(x) → +∞ as x → a+, we may assume that c is such that g(c) > 0. Letting β = c, we have 1 L  f(c)  g(c)−g(α) < < L ∀ α ∈ (a, c). (1)  f(α) g(c) + g(c) Since g(α) → 0 as α → a , we may assume 0 < g(α) < 1 ∀ α ∈ (a, c), from which it follows g(α) − g(c) g(c) = 1 − > 0 ∀ α ∈ (a, c). g(α) g(α) g(α)−g(c) If we exponentiate (1) by g(α) , we have

g(α)−g(c) 1 L g(α) f(α) g(α) g(α)−g(c) < < (L) g(α) .  f(c)

1 g(c) g(α) + Now, since g(α) → 0 and f(c) → 1 as α → a , then for any δ such that 0 < δ ≤ 1, 1 g(c) g(α) δ ∃ d ∈ (a, c) such that 0 < g(α) < δ and f(c) < e ∀α ∈ (a, d), giving

(1−δ) L 1 < f(α) g(α) < (L) eδ ∀α ∈ (a, d). 

7 1 δ 2 Now take δ = min{1, ln , ln  ln L }. Then we have (L)e ≤ L . We also have L(1−δ) L L L L > δ ≥ 1 = 1 = 2 .  L Lln  ln L ( ln L )ln L  1 L g(α) 2 Thus we have 2 < f(α) < L , and because  is arbitrary we have our desired result. The cases for L = 1 and 0 < L < 1 are similar. 1 Case (b): If L = ∞, let M > 1 be given and c ∈ (a, b) such that f ∗(u) g0(u) > M ∀ u ∈ (a, c). Then it follows that

1 f(β) g(β)−g(α) > M ∀ α, β ∈ (a, c], α < β. (2) f(α)

1 Since g(x) → ∞ as x → a+, we may select c, d, d < c such that g(c) > 0, f(c) g(α) > √ g(c) 1 1/ e, and 0 < g(α) < 2 ∀ α ∈ (a, d). If we take β = c in (2) and exponentiate by g(c) 1 − g(α) , we get 1   g(α) g(c) √ f(α) 1− > M g(α) > M. f(c) Thus, we have r 1 1 √ M f(α) g(α) > f(c) g(α) M > . e 1 g(α) Since M > 1 is arbitrary, it follows that limα→a+ f(α) = ∞. The argument is similar for L = 0.

4 Product Taylor’s Theorem

In the final section we prove a product version of Taylor’s theorem. Instead of repre- senting a function as the sum of a number of terms with an additive remainder, the product version shows how to represent a function as the product of a number of factors with a multiplicative remainder. Theorem 7. (Product Taylor Theorem) If f is an n + 1 times continuous, product differentiable function on an open interval I containing the point c such that f [i](c) 6= 0 for all i, 0 ≤ i ≤ n, then ∀x ∈ I such that x > c,

2 n ∗ x−c ∗∗ (x−c) [n] (x−c) f(x) = f(c)f (c) f (c) 2! . . . f (c) n! Rn(x), where (x−c)n+1 [n+1] Rn(x) = f (z) (n+1)! for some z such that c < z < x.

8 Proof. Define the functions F (t) and G(t) such that f(x) F (t) = n (x−t)k Q [k] k! k=0 f (t) (x − t)n+1 G(t) = . (n + 1)! Note here that, taking 00 = 1, we have F (x) = 1 and G(x) = 0. Thus, using the generalized mean value theorem for product calculus, we have for some z ∈ (c, x)

1 1 F (x) G(x)−G(c) 1 F ∗(z) G0(z) = = F (c) G(c) . F (c) It follows that 1 1 F (c) G(c) = F ∗(z) G0(z)  1  1 (x−t)n+1 ! d d(f(x)) dt (n+1)! dt = k t = z n (x−t) 1 Q [k] k! dt k=0 d(f (t) ) 1   −(x−z)n 1 ( n! ) =  (n−1)  −(x−z) (x−z)n f ∗(z)(f ∗(z)−1f ∗∗(z)(x−z))f ∗∗(z)−(x−z) . . . f [n](z) (n−1)! f [n+1](z) n! 1 ! −(x−z)n 1 ( n! ) = (x−z)n f [n+1](z) n! = f [n+1](z). Thus, we deduce that F (c) = f [n+1](z)G(c) f(x) (x−c)n+1 ⇔ = f n+1(z) (n+1)! n (x−c)k Q [k] k! k=0 f (c) 2 n (x−c)n+1 ∗ x−c ∗∗ (x−c) [n] (x−c) n+1 ⇔ f(x) = f(c)f (c) f (c) 2! . . . f (c) n! f (z) (n+1)! .

(x−c)n+1 n+1 From this, it follows that Rn(x) = f (z) (n+1)! . Corollary 3. If f [n](x) exists ∀ n and is continuous on the interval I containing the point c, and if limn→∞ Rn(x) = 1, then ∀ x > c such that x ∈ I, n k Y [k] (x−c) f(x) = lim f (c) k! . n→∞ k=0 Proof. This result follows directly from the previous theorem.

9 References

[1] Robert G. Bartle and Donald R. Sherbert. Introduction to Real . Wiley, New York, 3rd edition, 2000.

[2] John D. Dollard and Charles N. Friedman. Product Integration with Applications to Differential Equations. Addison-Wesley, Reading, MA, 1979.

[3] Michael Grossman. The First Nonlinear System of Differential and Integral Calcu- lus. MATHCO, Rockport, MA, 1979.

[4] Michael Z. Spivey. A product calculus. Submitted.

10