<<

Euler-Maclaurin formula

Michael S. Floater May 2, 2019

In these notes we derive the Euler-Maclaurin formula and apply it to .

1 Bernoulli numbers

We start by defining the Bernoulli numbers Bn, n = 0, 1, 2,.... We define them as the coefficients in the expansion

∞ x X xj = B . ex − 1 j j! j=0

This means that x B0 = lim = 1. x→0 ex − 1 x We obtain Bn, n ≥ 1 recursively. By the Maclaurin expansion of e we have

∞ i ! ∞ j ! ∞ n−1 ! X x X x X X Bj x = B = xn, i! j j! (n − j)!j! i=1 j=0 n=1 j=0 which implies that n−1 X n B = 0, n ≥ 2. (1) j j j=0 Thus, 1 1 B = − B = − , 1 2 0 2

1 1 1 B = − (B + 3B ) = , 2 3 0 1 6 1 B = − (B + 4B + 6B ) = 0, 3 4 0 1 2 and so on, and we find

n 0 1 2 3 4 5 6 Bn 1 -1/2 1/6 0 -1/30 0 1/42

We will show that Bn = 0 for n ≥ 3 and n odd. This follows from the fact that x f(x) := − (B + B x) ex − 1 0 1 is symmetric, i.e., f(x) = f(−x). To show this observe that the constant B0 is trivially symmetric, and so it is sufficient to show that

x x x(ex + 1) g(x) := + = ex − 1 2 2(ex − 1) is symmetric. To show this we compute

−x(e−x + 1) −x(1 + ex) g(−x) = = = g(x). 2(e−x − 1) 2(1 − ex)

2 Bernoulli polynomials

We define the Bernoulli polynomial of degree n ≥ 0 as

n X n B (t) = B tn−j. n j j j=0

2 So

B0(t) = 1, 1 B (t) = t − , 1 2 1 B (t) = t2 − t + , 2 6 3 1 B (t) = t3 − t2 + t, 3 2 2 1 B (t) = t4 − 2t3 + t2 − , 4 30 5 5 1 B (t) = t5 − t4 + t3 − t, 5 2 3 6 5 1 1 B (t) = t6 − 3t5 + t4 − t2 + , 6 2 2 42 and so on. From the definition and using (1) we deduce the following prop- erties: • Endpoint property:

n X n B (1) = B = B = B (0), n ≥ 2. n j j n n j=0

• Differentiation:

n−1 X n B0 (t) = (n − j) B tn−j−1 = nB (t), n ≥ 1. n j j n−1 j=0

• Integration:

Z 1 Z 1 1 0 Bn(t) dt = Bn+1(t) dt = 0, n ≥ 1. 0 n + 1 0

We will also derive the • Upper bound: |B2r(t)| ≤ |B2r|, r ≥ 0.

3 To do this we use Fourier . We define

Bbn(t) = Bn(t − j), j ≤ t ≤ j + 1, j ∈ Z, the 1-periodic extension of Bn : [0, 1] → R. By repeated differentiation, n! B(j)(t) = B (t), j = 0, 1, . . . , n, n ≥ 1, n (n − j)! n−j which implies that

(j) (j) Bn (0) = Bn (1), j = 0, 1, . . . , n − 2,

n−2 and therefore Bbn ∈ C (R). Consider the of Bbn,

∞ X 2πikt Bbn(t) = cke . k=−∞

Because Bbn(t) is real, c−k = ck. So letting ck = ak + ibk, we have a−k = ak and b−k = −bk, and so

∞ ∞ X X Bbn(t) = a0 + 2 ak cos(2πikt) + 2 bk sin(2πikt). k=1 k=1

The coefficients of Bbn are Z 1 −2πikt ck = Bn(t)e dt. 0 We find Z 1 c0 = Bn(t) dt = 0, n ≥ 1, 0 and for k 6= 0,

 −2πikt 1 Z 1 e −2πikt ck = Bn(t) + n Bn−1(t)e dt = −2πik 0 0 Z 1 n! −2πikt −n! ··· = n−1 B1(t)e dt = n = ak + ibk. (2πik) 0 (2πik)

4 It follows that a0 = b0 = 0 and for k 6= 0, (2r)! a = (−1)r−1 , b = 0, if n = 2r, k (2πk)2r k

(2r − 1)! a = 0, b = (−1)r , if n = 2r − 1. k k (2πk)2r−1 We now deduce that

∞ r−1 X cos(2πkt) Bb2r(t) = (−1) 2(2r)! , (2πk)2r k=1 and ∞ r X sin(2πkt) Bb2r−1(t) = (−1) 2(2r − 1)! . (2πk)2r−1 k=1 From this we obtain the claimed upper bound since

∞ X 1 |Bb2r(t)| ≤ 2(2r)! = |Bb2r(0)| = |B2r|. (2πk)2r k=1 3 Local Euler-Maclaurin expansion

We next use the Bernoulli polynomials to expand the of a function F : [0, 1] → R in terms of the trapezoidal rule and endpoint of odd order. We use successively.

Lemma 1 For r ≥ 0 and F ∈ C2r+2[0, 1],

Z 1 r+1 1 X B2k F (t) dt = (F (0) + F (1)) − (F (2k−1)(1) − F (2k−1)(0)) + R , 2 (2k)! r 0 k=1 where Z 1 1 (2r+2) Rr := B2r+2(t)F (t) dt. (2r + 2)! 0

5 Proof. Using integration by parts twice,

Z 1 Z 1 1 Z 1 h i 0 F (t) dt = B0(t)F (t) dt = B1(t)F (t) − B1(t)F (t) dt 0 0 0 0 1 Z 1 1 1h 0 i 1 00 = (F (0) + F (1)) − B2(t)F (t) + B2(t)F (t) dt 2 2 0 2 0 1 B = (F (0) + F (1)) − 2 (F 0(1) − F 0(0)) + R , 2 2 0 which proves the lemma in the case r = 0. Otherwise, suppose r ≥ 1, and assume that the lemma holds with r replaced by r − 1. Then again using integration by parts twice,

Z 1 1 (2r) Rr−1 = B2r(t)F (t) dt (2r)! 0 1 Z 1 1 h (2r) i 1 (2r+1) = B2r+1(t)F (t) − B2r+1(t)F (t) dt (2r + 1)! 0 (2r + 1)! 0 1 Z 1 1 h (2r+1) i 1 (2r+2) = − B2r+2(t)F (t) + B2r+2(t)F (t) dt (2r + 2)! 0 (2r + 2)! 0 1 = − (F (2r+1)(1) − F (2r+1)(0)) + R , (2r + 2)! r which completes the proof. 2

The Bernoulli polynomial B2r+2(t) in the remainder term Rr is not of one sign, so we cannot apply the . However, we can fix this by combining Rr with the last term in the expansion. Lemma 2 For r ≥ 0 and F ∈ C2r+2[0, 1], there is some ξ ∈ (0, 1) such that

Z 1 r 1 X B2k F (t) dt = (F (0) + F (1)) − (F (2k−1)(1) − F (2k−1)(0)) − R, 2 (2k)! 0 k=1 where B R = 2r+2 F (2r+2)(ξ). (2r + 2)!

6 Proof. The last term in the expansion in Lemma 1 plus the remainder can be written as B − 2r+2 (F (2r+1)(1) − F (2r+1)(0)) + R (2r + 2)! r Z 1 B2r+2 (2r+2) = − F (t) dt + Rr = −R, (2r + 2)! 0 where Z 1 1 (2r+2) R := (B2r+2 − B2r+2(t))F (t) dt. (2r + 2)! 0

Using the upper bound on B2r(t) of the previous section, the difference B2r − B2r(t) is of one sign in [0, 1], because

sgn(B2r)(B2r − B2r(t)) = |B2r| − sgn(B2r)B2r(t) ≥ |B2r| − |B2r(t)| ≥ 0. So by the mean value theorem, there is some ξ ∈ (0, 1) such that

Z 1 1 (2r+2) B2r+2 (2r+2) R = (B2r+2 − B2r+2(t)) dt F (ξ) = F (ξ). (2r + 2)! 0 (2r + 2)! 2

4 Global Euler-Maclaurin expansion

We apply the local expansion to obtain the global one. Given an interval [a, b] choose n ≥ 1 and let h = (b − a)/n and xi = a + ih, i = 0, 1, . . . , n. Theorem 1 For r ≥ 0 and f ∈ C2r+2[a, b], there is some ξ ∈ (a, b) such that Z b r X B2k f(x) dx = T (h) − h2k(f (2k−1)(b) − f (2k−1)(a)) − R, (2k)! a k=1 where n−1 h X T (h) = (f(a) + f(b)) + h f(x ), 2 i i=1 and B R = 2r+2 (b − a)h2r+2f (2r+2)(ξ). (2r + 2)!

7 Proof. Let F (t) = f(xi−1 + ht), t ∈ [0, 1], i = 1, . . . , n. Then Lemma 2 gives

Z xi Z 1 f(x) dx = h F (t) dt = xi−1 0 r h X B2k (f(x ) + f(x )) − h2k(f (2k−1)(x ) − f (2k−1)(x )) − R , 2 i−1 i (2k)! i−1 i i k=1 where B R = 2r+2 h2r+3f (2r+2)(ξ ) i (2r + 2)! i for some ξi ∈ (xi−1, xi). Summing this equation over i = 1, . . . , n yields the desired expansion except that the remainder term is

n n X B2r+2 X R = R = h2r+3 f (2r+2)(ξ ). i (2r + 2)! i i=1 i=1 However, an application of the mean value theorem gives

n X (2r+2) (2r+2) f (ξi) = nf (ξ) i=1 for some ξ ∈ (a, b), and this leads to the remainder term as claimed. 2

5 Applications

5.1 Endpoint corrections We can view the endpoint derivatives in the Euler-Maclaurin expansion as correction terms that greatly improve the accuracy of the trapezoidal rule, if they are available. From Theorem 1 we have

Corollary 1 For r ≥ 0 and f ∈ C2r+2[a, b],

Z b r X B2k f(x) dx = T (h) − h2k(f (2k−1)(b) − f (2k−1)(a)) + O(h2r+2) (2k)! a k=1 as h → 0.

8 5.2 Superconvergence of the trapezoidal rule For some functions, the trapezoidal rule itself has higher accuracy than usual. Corollary 2 Suppose r ≥ 0 and f ∈ C2r+2[a, b]. If f (2k−1)(b) = f (2k−1)(a) for k = 1, . . . , r, then Z b f(x) dx = T (h) + O(h2r+2) as h → 0. a This we will be the case for any r ≥ 0 for functions f ∈ C∞(R) that are periodic with period b − a. In fact for some functions of this type, the trapezoidal rule is exact. Theorem 2 Let [a, b] = [0, 2π] and let

n−1 n−1 X X f(x) = ak cos(kx) + bk sin(kx), k=0 k=1 for any choice of ak and bk in R. Then T (h) is exact for f. Proof. It is sufficient to show that T (h) is exact for f(x) := eikx, k = 0, 1, . . . , n − 1. The integral of this f is ( Z 2π 2π, k = 0; f(x) dx = 0 0, k > 0. On the other hand, since f(0) = f(2π),

n−1 ! 1 X 2jπ  T (h) = h (f(0) + f(2π)) + f 2 n j=1 n−1 n−1 X 2jπ  X = h f = h e2jkiπ/n. n j=0 j=0 So if k = 0, T (h) = hn = 2π, and if 1 ≤ k ≤ n − 1,

n−1 X e2kiπ − 1 T (h) = h (e2kiπ/n)j = h = 0. e2kiπ/n − 1 j=0 2

9 5.3 Sums of p-th powers We can also use the expansion to obtain a formula for sums of p-th powers. Corollary 3 For p ≥ 1,

n−1 X 1 jp = (B (n) − B ). p + 1 p+1 p+1 j=1 The first examples are

n−1 X 1 1 j = n2 − n, 2 2 j=1 n−1 X 1 1 1 j2 = n3 − n2 + n, 3 2 6 j=1 n−1 X 1 1 1 j3 = n4 − n3 + n2, 4 2 4 j=1 n−1 X 1 1 1 1 j4 = n5 − n4 + n3 − n, 5 2 3 30 j=1 n−1 X 1 1 5 1 j5 = n6 − n5 + n4 − n2, 6 2 12 12 j=1

Proof. Let f(x) = xp, p ≥ 1, and [a, b] = [0, n]. Then Z b Z n np+1 f(x) = xp dx = . a 0 p + 1 With h = 1, the trapezoidal rule for f on [0, n] is

n−1 1 X T (h) = np + jp. 2 j=1 Let r be such that p = 2r or p = 2r + 1. Then Theorem 1 gives Z n r X B2k p! xp dx = T (h) − np−2k+1, (2k)! (p − 2k + 1)! 0 k=1

10 and therefore,

n−1 r X np+1 1 1 X p + 1 jp = − np + B np−2k+1 p + 1 2 p + 1 2k 2k j=1 k=1 p 1 X p + 1 = B np−k+1 p + 1 k k k=0 1 = (B (n) − B ). p + 1 p+1 p+1 2

11