• Non ci sono risultati.

• You can also insert your corrections in the proof PDF and email the annotated PDF.

N/A
N/A
Protected

Academic year: 2021

Condividi "• You can also insert your corrections in the proof PDF and email the annotated PDF."

Copied!
25
0
0

Testo completo

(1)

Dear Author,

Here are the proofs of your article.

• You can submit your corrections online, via e-mail or by fax.

• For online submission please insert your corrections in the online correction form. Always indicate the line number to which the correction refers.

• You can also insert your corrections in the proof PDF and email the annotated PDF.

• For fax submission, please ensure that your corrections are clearly legible. Use a fine black pen and write the correction in the margin, not too close to the edge of the page.

• Remember to note the journal title, article number, and your name when sending your response via e-mail or fax.

• Check the metadata sheet to make sure that the header information, especially author names and the corresponding affiliations are correctly shown.

• Check the questions that may have arisen during copy editing and insert your answers/

corrections.

• Check that the text is complete and that all figures, tables and their legends are included. Also check the accuracy of special characters, equations, and electronic supplementary material if applicable. If necessary refer to the Edited manuscript .

• The publication of inaccurate data such as dosages and units can have serious consequences.

Please take particular care that all such details are correct.

• Please do not make changes that involve only matters of style. We have generally introduced forms that follow the journal’s style.

Substantial changes in content, e.g., new results, corrected values, title and authorship are not allowed without the approval of the responsible editor. In such a case, please contact the Editorial Office and return his/her consent together with the proof.

• If we do not receive your corrections within 48 hours, we will send you a reminder.

• Your article will be published Online First approximately one week after receipt of your corrected proofs. This is the official first publication citable with the DOI. Further changes are, therefore, not possible.

• The printed version will follow in a forthcoming issue.

Please note

After online publication, subscribers (personal/institutional) to this journal will have access to the complete article via the DOI using the URL: http://dx.doi.org/[DOI].

If you would like to know when your article has been published online, take advantage of our free alert service. For registration and further information go to: http://www.springerlink.com.

Due to the electronic nature of the procedure, the manuscript and the original figures will only be

returned to you on special request. When you return your corrections, please inform us if you would

like to have these documents returned.

(2)

Metadata of the article that will be visualized in OnlineFirst

Please note: Images will appear in color online but will be printed in black and white.

ArticleTitle The “phase function” method to solve second-order asymptotically polynomial differential equations Article Sub-Title

Article CopyRight Springer-Verlag

(This will be the copyright line in the final PDF) Journal Name Numerische Mathematik

Corresponding Author Family Name Spigler

Particle

Given Name Renato

Suffix

Division Dipartimento di Matematica

Organization Universit à “Roma Tre”

Address 1, Largo S.L. Murialdo, 00146, Rome, Italy

Email marcov@math.unipd.it

Author Family Name Vianello

Particle

Given Name Marco

Suffix

Division Dipartimento di Matematica Pura e Applicata Organization Universit à di Padova

Address 63, Via Trieste, 35121, Padua, Italy Email

Schedule

Received 24 September 2010

Revised 29 November 2011

Accepted

Abstract The Liouville-Green (WKB) asymptotic theory is used along with the Borůvka’s transformation theory, to obtain asymptotic approximations of “phase functions” for second-order linear differential equations, whose coefficients are asymptotically polynomial. An efficient numerical method to compute zeros of solutions or even the solutions themselves in such highly oscillatory problems is then derived. Numerical examples, where symbolic manipulations are also used, are provided to illustrate the performance of the method.

Mathematics Subject Classification (2000) (separated by '-')

Primary: 65L99 - 34E20 - Secondary: 65D20

Footnote Information Work supported, in part, by the Italian GNFM-INdAM and GNIM-INdAM.

(3)

Author Query Form

Please ensure you fill out your response to the queries raised below and return this form along with your corrections

Dear Author

During the process of typesetting your article, the following queries have arisen. Please check your typeset proof carefully against the queries listed below and mark the

necessary changes either directly on the proof/online grid or in the ‘Author’s response’

area provided below

Query Details required Author’s response

1 Please confirm the corresponding author is correctly identified and amend if necessary.

Also check his e-mail (seems to be M. Vianello;

we have followed jobsheet).

Journal: 211

Article: 441

(4)

uncorrected

proof

Numer. Math.

DOI 10.1007/s00211-011-0441-9 Mathematik

The “phase function” method to solve second-order asymptotically polynomial differential equations

Renato Spigler · Marco Vianello

Received: 24 September 2010 / Revised: 29 November 2011

© Springer-Verlag 2011

Abstract The Liouville-Green (WKB) asymptotic theory is used along with the

1

Bor˚uvka’s transformation theory, to obtain asymptotic approximations of “phase func-

2

tions” for second-order linear differential equations, whose coefficients are asymp-

3

totically polynomial. An efficient numerical method to compute zeros of solutions

4

or even the solutions themselves in such highly oscillatory problems is then derived.

5

Numerical examples, where symbolic manipulations are also used, are provided to

6

illustrate the performance of the method.

7

Mathematics Subject Classification (2000) Primary: 65L99 · 34E20; Secondary:

8

65D20

9

1 Introduction

10

An efficient asymptotic-numerical method, to approximate zeros of solutions of

11

second-order linear differential equations,

12

y ′′ + q(x)y = 0, (1)

13

on a half-line, was developed in [28,29 ] for the oscillatory case with q(x) = a +b/x +

14

O(x − p ), a > 0, b ∈ R, p > 1. Such result has been obtained connecting Bor˚uvka’s

15

Work supported, in part, by the Italian GNFM-INdAM and GNIM-INdAM.

R. Spigler ( B )

Dipartimento di Matematica, Università “Roma Tre”, 1, Largo S.L. Murialdo, 00146 Rome, Italy e-mail: marcov@math.unipd.it

M. Vianello

Dipartimento di Matematica Pura e Applicata, Università di Padova, 63, Via Trieste, 35121 Padua, Italy

Author Proof

(5)

uncorrected

proof

transformation theory ([3,4], [21, §§5.1, 9.1], [30]) to an Olver’s idea developed to

16

compute zeros of cylinder (Bessel) functions [22,23]. This approach actually allows

17

to compute also the solutions to (1). In this paper, we extend our method to the case

18

of an asymptotically polynomial coefficient,

19

q(x) ∼ cx m , x → +∞, c, m ∈ R + . (2)

20

Equivalently, q(x) = cx m [1 + o(1)] as x → +∞. More precisely, q(x) is required to

21

be the restriction to the real half-line x > ρ of a function, q(z), analytic in the annular

22

sector

23

S ρ,γ := {z ∈ C : |z| > ρ, | arg z| < γ }, (3)

24

with ρ ≥ 0, 0 < γ ≤ π/2, and

25

q(z) ∼ cz m , z → ∞ in S ρ,γ . (4)

26

The exponent m does not need to be an integer, in which case it is understood that

27

the principal branch of z m has to be considered. Without any loss of generality, we

28

stipulate that q(z) does not vanish in S ρ,γ , and thus q(x) > 0 for x > ρ.

29

We shall confine ourselves to the reduced form (1), to which every smooth second-

30

order linear differential equation, y ′′ +a(x)y +b(x)y = 0, can be taken (see e.g. [ 4]),

31

with q(x) = b(x) − a 2 (x)/ 4 − a (x)/ 2. In particular, when a(z) ∼ αz n , b(z) ∼ βz m ,

32

as z → ∞ in S ρ

0

0

, with m, n ∈ N, m > 2n, and β > 0, assuming that a (z) ∼

33

αnz n −1 in an annular subsector S ρ,γ ⊆ S ρ

0

0

, we are led to an equation like (1)

34

with q(z) ∼ βz m in S ρ,γ , so that q(x) ∼ βx m → +∞ as x → +∞. Note that

35

the assumption concerning asymptotic differentiation of a(z) is satisfied, e.g., when

36

a(z) = αz n 1 + O(z − p ) , p > 0, see [23,31].

37

All the previous assumptions are satisfied, in particular, when a(x) := P a (x)/Q a

38

(x) and b(x) := P b (x)/Q b (x) , are rational functions, with b(x) → +∞ as x → +∞,

39

and deg (P b ) − deg (Q b ) > 2 (deg (P a ) − deg (Q a )). In this case, S ρ

0

0

= S ρ,γ is

40

an annular sector which leaves out the poles of a(x) and b(x), taking ρ sufficiently

41

large and/or γ sufficiently small. Note that, in fact, when a(x) and b(x) are rational,

42

so is q(x), and, in particular, when a(x) and b(x) are polynomials, so is q(x), with

43

deg (q) ≤ max{deg (b), 2 deg (a)}.

44

It is an immediate consequence of the Sturm comparison theorem (see e.g. [13,

45

Ch. 11]), that all solutions of (1)–(2) exhibit an extremely rapid oscillatory behavior,

46

with frequency increasing unboundedly. The numerical treatment of such problems

47

is known to represent a difficult task, which calls for special discretization methods

48

(see e.g. [14,24,27], and references therein). Since time-stepping methods, however,

49

require exceedingly small time-steps, and the global errors become very large due

50

to the large values taken by the higher derivatives of solutions [16,17], recently, a

51

number of alternatives have been proposed [6,18]. Some of these are, e.g., Magnus,

52

Cayley, and Neumann methods, which are related to Lie group techniques [18].

53

Others, like Filon-type methods, rest on suitable interpolations [6,7]. They are all

54

based on the reformulation of the differential equations in terms of highly oscillatory

55

Author Proof

(6)

uncorrected

proof

integrals. All the algorithms of the first group come from the Magnus series expansion

56

of solutions or variants of it, those of the second group involve suitable polynomial

57

(Hermite) interpolation of part of the integrand, namely the slowly varying amplitude

58

factor. In a recent application to forced oscillator equations arising from electronic

59

circuit simulation, Filon-type quadratures have been used very successfully [6,7]. An

60

extension of such problems was treated efficiently by still another method, based on

61

modulated Fourier series [8]. Throughout, extensive exploitation of asymptotics and,

62

when possible, of symbolic manipulations, have proved to be quite useful.

63

In this paper, we consider still another method to solve oscillator equations as (1),

64

whose coefficient q(x) is asymptotically polynomial. In some cases treated in [16,17],

65

q(x) was required to diverge to +∞ as x → +∞, while its derivatives should remain

66

small, in the same limit. The algorithm proposed in the present paper is not affected

67

by such limitation, even though here we do not consider such cases, e.g. when q(x)

68

diverges exponentially. Moreover, also the present algorithm avoids time-stepping and

69

seems to perform even better when oscillations are faster, as in the aforementioned

70

methods.

71

The inherent difficulties in handling rapidly oscillatory solutions strongly affects,

72

in particular, the numerical approximation of zeros of solutions, whenever one

73

computes them via the approximation of the solutions themselves. As in [29], instead

74

of evaluating in advance the solutions to (1)–(2), our approach for computing zeros of

75

any given solution consists in obtaining an asymptotic-numerical approximation of a

76

related function, that is one of the so-called “phases” or “phase functions”. A phase

77

function, α(x), is any C 3 -solution of tan α(x) = u(x)/v(x), where (u, v) denotes a

78

basis for equation (1). By differentiation, we obtain

79

α (x) = − W

u 2 (x) + v 2 (x) , (5)

80

where W := W [u, v] is the (constant) Wronskian of u, v. By further differentiations

81

and using (1), the following third-order nonlinear differential equation,

82

α ′2 (x) = q(x) − 1

2 {α, x}, {α, x} := α ′′′ (x) α (x) − 3

2

 α ′′ (x) α (x)

 2

, (6)

83

is obtained. In (6 ), {α, x} represents the so-called “Schwarzian derivative” of α,

84

and the close-form equation satisfied by α is a special instance of the “Kummer

85

differential equation”, which plays a key role in the transformation theory of linear

86

ordinary second-order differential equations [3,4,21,30]. Indeed, every solution of (6)

87

is the “kernel” of a transformation which takes equation (1) into Y ′′ + Y = 0. As a

88

consequence of this fact, it is easily shown that every solution to (6) is a phase function

89

related to the basis

90

u(x) = |α (x) | −1/2 sin α(x), v(x) = |α (x) | −1/2 cos α(x), (7)

91

[3,4]. Therefore, knowing a single phase function, it is possible to retrieve a basis

92

(and thus all solutions) to equation (1). In addition, zeros of solutions can be directly

93

Author Proof

(7)

uncorrected

proof

obtained by means of α(x), since it can be proved that

94

|α(x k ) − α(x k−1 ) | = π (8)

95

for any two consecutive zeros, x k −1 and x k , of any given solution of (1): From (8), in

96

fact, given α(x) and one of the two zeros, it is possible to evaluate the other, solving

97

a nonlinear equation like α(x) = const. (note that, by ( 5), all phase functions are

98

strictly monotone).

99

Even though, in principle, obtaining a phase function α(x) requires solving a third-

100

order nonlinear differential equation, we consider the iterative scheme

101

0 ) 2 (x) = q(x),n +1 ) 2 (x) = q(x) − 1

2 {α n , x }, n = 0, 1, 2, . . . , (9)

102

see [29], and show that, under the asymptotic assumptions in (4), α n (x) converges (in

103

a suitable sense) to a special phase function, α(x) ≡ α L G (x), the so-called Liouville–

104

Green phase. Such a convergence is usually very fast, and thus the method becomes

105

very competitive when compared to methods based on the preliminary numerical

106

solution of the original differential equation, (1).

107

Here is the plan of the paper. In Sect. 2, we develop the basic asymptotic anal-

108

ysis, obtaining the representation α ′2 L G (z) = q(z) + O(z −2 ) in a suitable subsector

109

of S ρ,γ . This makes it reasonable the initial approximation in the iterative scheme in

110

(9), whose convergence is proved in Sect. 3. In Sect. 4, we show that our method can

111

be implemented efficiently, in particular when q(x) is a polynomial, and numerical

112

examples are given for the purpose of illustration.

113

2 Preliminary asymptotic analysis

114

Setting

115

φ := (α ) 2 , φ n := (α n ) 2 , (10)

116

equation (6) transforms into

117

φ (x) = q(x) + [φ, x], [φ, x] := − 1 4

φ ′′ (x) φ (x) + 5

16

 φ (x) φ (x)

 2

, (11)

118

and the iterative scheme in (9) becomes

119

φ 0 (x) = q(x), φ n+1 (x) = q(x) + [φ n , x ], n = 0, 1, 2, . . . (12)

120

Note that the third-order problem in (6) has been reduced to a second-order problem

121

(in (11)), and α can be recovered by a quadrature. Moreover, the scheme in (12) has

122

the advantage, with respect to that in (9), of avoiding the evaluation of square roots.

123

This circumvents the difficulties due to the complexification of the problem. In fact,

124

Author Proof

(8)

uncorrected

proof

both, asymptotic estimates and proof of convergence, will be performed in suitable

125

sectors of C, since we need to control derivatives.

126

In the next section, we shall prove the convergence (in a suitable sense) of φ n to a

127

particular solution of equation (11 ), φ ≡ φ L G , which is the square of the derivative

128

of any “Liouville-Green phase”, that is any phase function related to a real-valued

129

Liouville-Green (WKB) basis of equation (1).

130

Recall that a complex-valued Liouville-Green (L G) basis is, according to Olver’s

131

theorem ([23, Ch. 6, Thm. 11.1], see also [32]),

132

U ± (z) = q −1/4 (z) e ±iξ(z) [1 + ε ± (z) ], (13)

133

where

134

ξ(z) =

z



q 1/2 (t ) dt, (14)

135

and

136

|ε ± (z) | ≤ exp V z, ± (F )  − 1, |ε ± (z) | ≤ |q 1/2 (z) | exp V z, ± (F )  − 1 , (15)

137

the z-function V z,∞ ± (F ) denoting the variation of

138

F ≡ F(z) :=

z



q −1/4 D 2 q −1/4

dt (16)

139

along the path ℓ ± , see [23]. The paths ℓ ± ≡ ℓ ± (z) connect z to ∞ remaining inside a

140

suitable annular subsector of S ρ,γ , and are required to be ξ -progressive, i.e., Im ξ(z)

141

must be nonincreasing along ℓ and nondecreasing along ℓ + . In (14), (15), (16), the

142

branches of the fractional powers of q(z) must be continuous, that of q 1/2 (z) being

143

the square of q 1/4 (z). Note that in (13) a family of bases is actually defined, corre-

144

spondingly to the arbitrary choice of the primitive, ξ(z).

145

When q(z) is as in (4), we can show that the estimates

146

ε ± (z) = O(z −m/2−1 ), q −1/2 (z)ε ± (z) = O(z −m/2−1 ), z ∈ S ρ

⊂ S ρ,γ ,

147

(17)

148

hold, for ρ > ρ sufficiently large, and γ < min {γ, π/(m + 2)}. Such estimates have

149

been obtained using as path ℓ + the circular arc η ≡ η z (θ ) = |z|e i θ , arg z ≤ θ ≤ γ ,

150

followed by the ray η ≡ η z (r ) = re i γ

, r ≥ |z|, whereas ℓ − is the circular arc η ≡

151

η z (θ ) = |z|e i θ , arg z ≥ θ ≥ −γ , followed by the ray η ≡ η z (r ) = re −iγ

, r ≥ |z|.

152

Let check that such paths are indeed ξ -progressive. In fact, on the circular arc of

153

+ , we have

154

d

Im ξ(η(θ )) = Im  dξ



= Im

q 1/2 (η(θ ))η (θ )

155

= √ c Im

η m/2 (θ )η (θ ) ( 1 + ω)

, (18)

156

Author Proof

(9)

uncorrected

proof

where ω → 0 as z → ∞ in S ρ,γ . Now,

157

η m/2 η = |z| m/ 2+1 e i [(m/2+1)θ+π/2] , (19)

158

hence

159

Im

η m/2 η ( 1 + ω)

= |z| m/ 2+1

cos m 2 + 1

θ 

( 1 + Re ω)

160

− sin m 2 + 1

θ  Im ω

∼ |z| m/ 2+1 cos m 2 + 1

θ 

(20)

161

as z → ∞ in S ρ,γ

, where 0 < γ < π/(m + 2). Therefore, the left-hand side is

162

strictly positive for |z| ≥ ρ 1 > ρ, with ρ 1 sufficiently large (and | arg z| ≤ γ ). This

163

shows that Im ξ(η(θ )) increases on the circular arc in such subsector.

164

Similarly, on the ray following the circular arc in ℓ + ,

165

d

dr Im ξ(η(r )) =c r m/2

sin m 2 +1

γ 

( 1+Re ω)+cos m 2 +1

γ  Im ω

166

∼ √

c r m/2 sin m 2 + 1

γ 

(21)

167

as z → ∞ in S ρ,γ

, ω being an infinitesimal. Again, the left-hand side is positive for

168

|z| ≥ ρ 2 > ρ, for ρ 2 sufficiently large.

169

An analogous procedure is adopted to show that the path ℓ is also ξ -progressive

170

in S ρ,γ

, with ρ sufficiently large. Below, we denote with ρ the largest among the

171

previous lower bounds for ρ (i.e., ρ 1 , ρ 2 , and the corresponding radii concerning ℓ ).

172

In order to establish the asymptotic estimate in (17), we compute first (by Ritt’s

173

theorem [23, Thm.4.2, p. 9], [31])

174

q −1/4 D 2 q −1/4 = O

z −m/2−2

, (22)

175

valid in any fixed subsector of S ρ,γ , for instance in S ρ

, and thus, after a little algebra,

176

we obtain

177

V z, ± (F ) :=



1±





 q −1/4 D 2 q −1/4





 | d z | +



2±





 q −1/4 D 2 q −1/4





 | d z |

178

≤ 2K

 2

M + 1 + γ



|z| −m/2−1 , z ∈ S ρ

, (23)

179

where the superscript appended to ℓ ± refers to a circular arc (index 1) or to a ray

180

(index 2), and K is the constant implied by the O-symbol in (22). From this, (17)

181

follows, since e V − 1 = O(V ) as V → 0.

182

Author Proof

(10)

uncorrected

proof

From any complex L G basis in (13), we obtain a real L G basis,

183

u(x) := Re U + (x) = U + (x) + U − (x)

2 , v(x) := Im U + (x) = U + (x) − U − (x)

2i ,

184

(24)

185

on the real half-line x > ρ , since U (x) = U + (x).

186

We stipulate that a primitive ξ(z), real for z real, has been chosen, which is possible

187

since q(z) has such property, q(x) > 0, and the appropriate branch of the square root

188

has been selected. Then, we define

189

φ L G (x) := W 2 [u, v]

[u 2 (x) + v 2 (x) ] 2 ≡ 1

[u 2 (x) + v 2 (x) ] 2 , (25)

190

which is the square of the derivative of every phase defined by the family of bases in

191

(24), see (10). Recall, once for all, that φ L G is invariant in the family of real L G bases

192

(obtained varying the real constant of integration in (14)). In fact, all such bases are

193

related to each other by a unimodal orthogonal transformation, and the corresponding

194

phases are related to each other by an additive constant. The Wronskian in (25), which

195

is clearly a constant (see (1)), turns out to be equal to 1 for any L G basis. In fact,

196

simple direct calculations, using (24) and (13), yield

197

W [u, v] = 1 2i

2i (1 + ε + )( 1 + ε − ) + q −1/2 ( 1 + ε − )ε + − q −1/2 ( 1 + ε + )ε  ,

198

(26)

199

and thus, by (17 ), we obtain W [u, v] → 1 as x → +∞, and hence W [u, v] ≡ 1.

200

Extending analytically φ L G to the complex sector S ρ

, by using (24) and (25),

201

we obtain the asymptotic representation for φ L G (z)

202

φ L G (z) = 1

q −1/2 (z) [1 + ε + (z) ][1 + ε − (z) ]  2 = q(z) 1 + O

z −m/2−1  ,

203

(27)

204

valid in S ρ

′′

, where |ε ± (z) | < 1 for |z| > ρ ′′ . Therefore, by (4), φ L G (z) =

205

O (z m ) + O z m/ 2−1  = O (z m ) in S ρ

, while, by Ritt’s theorem ([23, Thm.4.2,

206

p. 9]; see also [31]),

207

φ L G (z) = O z m −1

, φ ′′ L G (z) = O z m −2

, (28)

208

in any fixed annular subsector of S ρ

′′

, say S ρ

. Inserting these in the right-hand

209

side of the complexified differential equation (11) (recalling that φ L G (z) = 0 in

210

Author Proof

(11)

uncorrected

proof

S ρ

), we obtain

211

L G , z ] = − 1 4

O z m −2 

q(z) 1 + O z −m/2−1  + 5 16

 O z m −1  q(z) 1 + O z −m/2−1 

 2

212

= O z −2

, in S ρ

, (29)

213

and hence

214

φ L G (z) = q(z) + O z −2

, z ∈ S ρ

. (30)

215

Note that this represents a refinement of the asymptotic estimate in (27).

216

3 An iterative method for approximating L G phases

217

In order to prove the convergence of the iterative scheme in (12), we embed the

218

problem in the complex domain, that is we consider the analytic extension of both,

219

equation (11 ) for φ ≡ φ L G , and equation (12) for φ n +1 ,

220

φ L G (z) = q(z) + [φ L G (z), z ], (31)

221

φ 0 (z) = q(z), φ n+1 (z) = q(z) + [φ n (z), z ], n = 0, 1, 2, . . . , (32)

222

where z ∈ S ρ

(the annular sector in (30)). Such a procedure is suggested by the

223

need of “controlling” the derivatives in the iterative scheme (32). We start with a

224

technical lemma.

225

Lemma 3.1 Consider the complex-valued function

226

G(w 1 , w 2 , w 3 ) := q(z) − 1 4

w 3 w 1 + 5

16

 w 2 w 1

 2

(33)

227

of the three complex variables w 1 , w 2 , w 3 , defined in the closed polydisc of C 3

228

P = D 1 × D 2 × D 3 , where D k := B

φ (k L G −1) (z) ; ε k (z)

, k = 1, 2, 3, z ∈ S ρ

.

229

Here B(a ; r) ⊂ C denotes the open disc centered at a with radius r, and choose

230

ε 1 (z) = |φ L G (z) |/2 and ε k (z) = |φ (k L G −1) (z) |, k = 2, 3. Then, the following Lipschitz

231

estimate holds,

232

|G(w) − G(v)| = χ(z)||w − v|| ∞ , ∀w, v ∈ P, (34)

233

for every fixed z ∈ S ρ

, where we set w = (w 1 , w 2 , w 3 ), v = (v 1 , v 2 , v 3 ), and the

234

estimate

235

0 ≤ χ(z) ≤ C|z| −m (35)

236

holds, C being a suitable constant.

237

Author Proof

(12)

uncorrected

proof

Proof By Hartog’s theorem [26, Ch. 1], G is analytic in P, being separately analytic

238

in each of its arguments, w k ’s. In fact, w 1 ∈ B(φ L G (z) ; ε 1 (z)) implies that

239

|w 1 | ≥ | |φ L G (z) | − |φ L G (z) − w 1 | | ≥ |φ L G (z) | − ε 1 (z) = |φ L G (z) |

2 > 0. (36)

240

By the (generalized) mean value theorem,

241

|G(w) − G(v)| ≤ max

s ∈[w,v] {|grad G(s)|} ||w − v|| ∞ . (37)

242

This property holds in general for C 1 -mappings, G : (U ⊂ B 1 ) → B 2 , where U is

243

a convex subset of B 1 , B 1 and B 2 are Banach spaces, see e.g. [19], and the triple bar

244

denotes the norm of the Frêchét derivative of G. In (37 ), [w, v] denotes the segment

245

in C 3 , joining w and v, which is a subset of P, for any w, v ∈ P, P being a convex

246

set in C 3 . To estimate grad G, using (36) and observing that, by (27)–(28 ), |w 1 | −1 =

247

O z −m  , |w 2 | ≤ 2 

L G (z) 

 = O z m −1  , |w 3 | ≤ 2 

L G ′′ (z) 

 = O z m −2  in P, we

248

first evaluate

249









∂G

∂w 1







 ≤









 w 3 4w 2 1









 + 5

8









 w 2 2 w 3 1











= O z −2m

+ O z −m−2

,

250









∂G

∂w 2







 = 5 8









 w 2 w 1 2











= O z −2m

+ O z −m−1

, (38)

251









∂G

∂w 3







 = 1

4|w 1 | = O z −m  ,

252

valid for every m > 0. Then

253

sup

s ∈P |grad G(s)| ≤ sup

s ∈P 3



k =1









G

∂w k







 =: χ(z) = O z −m  , (39)

254

for every z ∈ S ρ

. Therefore, the estimate in (34)–(35) is proved, in view of (37),

255

C being the constant implied by the O-symbol in (39). ⊓ ⊔

256

We are now ready to state the main convergence theorem.

257

Theorem 3.2 Assume that equation (1) be given, with q(x) asymptotically polyno-

258

mial as in (2), (3), (4). Then, there exists x 0 ∈ (ρ, +∞), depending on all parameters

259

of the problem, such that φ n (x) in the scheme (12) converges to the function φ L G (x)

260

defined in (25), in the sense that

261

L G (x) − φ n (x) | ≤ C h n (x) x − n

2 −2

, x > x n , (40)

262

Author Proof

(13)

uncorrected

proof

where

263

h 0 (x) ≡ 1, h n (x) := C n

n −1



j =0

x − j

2 −m

, n = 1, 2, 3, . . . , (41)

264

being x n := x 0 + n

2/ sin γ , x 0 ≥ ρ and n = 0, 1, 2, . . .; γ and ρ are, respec-

265

tively, the semi-angle and the radius of the annular sector in (30), and C the constant

266

in (35).

267

Remark 3.3 While the convergence of φ n to φ L G in the case of an asymptotically

268

constant coefficient is linear (see [28,29 ], with m = 0), the inequality in ( 40) as well

269

as the numerical results in the examples below suggest that now the convergence might

270

be superlinear,

271

L G (x) − φ n (x) | ≤ C n +1 x 0 −2

n−1



j =0

x 0 + (n − j) √ 2  −m

272

= C x 0 2

 C 2 m/2

 n

⎣ Ŵ

x

0

√ 2 + 1 Ŵ

x

0

√ 2 + 1 + n

m

, uniformly for x > x 0 + n √ 2 sin γ , (42)

273

where Ŵ(·) is the Euler gamma function. Note that the convergence rate increases

274

with m, which is an essential feature of the present method. Large values of m imply

275

extremely rapid oscillations of solutions to equation (1).

276

Remark 3.4 As it will be clear in the proof below, the convergence results in (40)–(41)

277

are also valid in the complex domain and, more precisely, for z replacing x, z ∈ S n

278

S(0, γ ) + x 0 + n

2/ sin γ . Note that the geometrical structure of such a decreasing

279

sequence of sectors parallels the similar one introduced in [29].

280

Proof Theorem 3.2 has been stated in R, but we shall prove it in C, for technical

281

reasons. The proof will be inductive on n. By (30),

282

L G (z) − φ 0 (z) | = |φ L G (z) − q(z)| ≤ C |z| −2 , z ∈ S 0 := S(0, γ ) + x 0 , (43)

283

and assume

284

L G (z) − φ n (z) | ≤ C h n ( |z|)(|z| − n

2) −2 , z ∈ S n := S 0 + n √ 2

sin γ , (44)

285

for a fixed n, n = 1, 2, 3, . . ., where h n ( |z|) is defined in ( 41), and x 0 ≥ ρ will be

286

given below. Then,

287

L G (z) − φ n +1 (z) | = | [φ L G (z), z ] − [φ n (z), z ] |

288

≡ | G(φ L G (z), φ L G (z), φ ′′ L G (z)) − G(φ n (z), φ n (z), φ n ′′ (z)) |

289

≤ sup

s ∈[w,w

n

] {|grad G(s)|} ||w − w n || ∞ , (45)

290

Author Proof

(14)

uncorrected

proof

where w := (φ L G (z), φ L G (z), φ L G ′′ (z)), w n := (φ n (z), φ n (z), φ ′′ n (z)), and, by

291

Lemma 3.1,

292

L G (z) − φ n +1 (z) | ≤ C |z| −m ||w − w n || ∞

293

≤ C |z| −m max |φ L G (z) − φ n (z) |, |φ L G (z) − φ n (z) |, |φ ′′ L G (z) − φ n ′′ (z) | , (46)

294

provided that w n ∈ P (the polydisc in Lemma 3.1). By Cauchy integral formula,

295





 φ L G (k) (z) − φ n (k) (z) 



 =











 k ! 2π i



Ŵ

φ L G (u) − φ n (u) (u − z) k +1 du













296

k ! ( √

2) k sup

u∈ŴL G (u) − φ n (u) |, z ∈ S n +1 , k = 0, 1, 2, (47)

297

where Ŵ denotes the circle centered at z, with radius

2, and thus u = z +2e i θ

298

on Ŵ. Note that S n+1 ⊂ S n , with dist (S n+1 , S n ) = √

2, see (44). Using the inductive

299

assumption in (44), we are able to estimate the sup term in (47), obtaining

300





 φ (k) L G (z) − φ n (k) (z)





 ≤ k ! ( √

2) k C sup

u ∈Ŵ



h n ( |u|)

|u| − n √ 2 −2 

301

k ! ( √

2) k C n +1

n −1



j =0

|z| − ( j + 1) √ 2 −m

|z| − (n + 1) √ 2 −2

302

≤ C n +1

n



j =1

|z| − j √ 2 −m

|z| − (n + 1) √ 2 −2

, z ∈ S n +1 , k = 0, 1, 2.

303

(48)

304

Now, we should show that w n ∈ P, for every z ∈ S n +1 , which is true by (48) whenever

305

C n +1

n



j =1

|z| − j √ 2 −m

|z| − (n + 1) √ 2 −2

< min

k ε k (z) = O(z m −2 ),

306

(49)

307

see Lemma 3.1. This can be seen to be true, for x 0 sufficiently large, uniformly in n,

308

estimating the left-hand side of (49) by the right-hand side of (42) (with n replaced by

309

n + 1), observing that the right-hand side of ( 49) is of order O x 0 m−2

, uniformly in

310

z ∈ S n+1 . Finally, from (46),

311

Author Proof

(15)

uncorrected

proof

L G (z) − φ n +1 (z) | ≤ C n +2 |z| −m

n



j=1

|z| − j √ 2 −m

|z| − (n + 1) √ 2 −2

312

= C n+2

n



j =0

|z| − j √ 2 −m

|z| − (n + 1) √ 2 −2

, z ∈ S n +1 , (50)

313

that is inequality (44 ) with n + 1 replacing n. ⊓ ⊔

314

4 Numerical implementation and examples

315

In order to compute a given zero (x ≡ x k ) of any given solution to equation (1), know-

316

ing (an approximation of) the successive zero, x k +1 , we solve numerically the equation

317

α L G (x k +1 ) − α L G (x) =

x

k+1



x

φ 1/2 L G (t ) dt = π, x < x k +1 , (51)

318

see equation (8). In fact, any phase, α(x), is by (5) strictly monotone, and we can

319

choose it to be monotonic increasing selecting W ≡ const. < 0 in ( 5) (the two basis

320

functions, u and v, can be interchanged). Successive approximations to x k , say x k (n) ,

321

can be obtained from (51), replacing φ L G with φ n , see Theorem 3.2, and solving the

322

ensuing equation,

323

f k,n (x) :=

x

k+1



x

φ n 1/2 (t ) dt − π = 0. (52)

324

This can be done by Newton’s method, provided that f k,n (x) ≡ −φ n 1/2 (x) is

325

evaluated efficiently, and a suitable quadrature rule is adopted to compute f k,n (x).

326

It is worth noting that the familiar hypotheses ensuring global convergence of

327

Newton’s method hold, choosing x k+1 as an initial guess ( f k,n (x k+1 ) = −π),

328

f k,n ′′ (x) ∼ − 1 2 q −1/2 (x)q (x) < 0 as x → +∞, in view of the asymptotic results

329

obtained in the complex domain within the proof of Theorem 3.2 (see Remark 3.4).

330

In the following examples, Simpson’s rule with the usual a posteriori error estimate

331

has been used successfully throughout to compute f k,n (x).

332

The key point is, however, evaluating φ n (x) in practice, by means of the itera-

333

tions in the algorithm (12). As already observed in [29], numerical differentiation to

334

compute [φ n , x ] is impractical because of its inherent instability, which propagates

335

destructively as the iterations proceed. Alternatively, one may consider using sym-

336

bolic differentiation tools, but the simple example of the Airy equation, i.e. equation

337

(1 ) with q(x) = x, and thus, by ( 12),

338

φ 0 (x) = x, φ 1 (x) = x + 5 16 x 2 ,

339

φ 2 (x) = −25 − 780 x 3 + 960 x 6 + 1024 x 9 4 x 2 ( 5 + 16 x 3 ) 2 ,

340

Author Proof

(16)

uncorrected

proof

φ 3 (x) = N 3 (x)

D 3 (x) , where

341

N 3 (x) := −15625 − 4500000 x 3 + 55192500 x 6 + 833304000 x 9

342

+ 1367424000 x 12 + 6930432000 x 15 + 8945664000 x 18 (53)

343

+ 377487360 x 21 + 3019898880 x 24 + 1073741824 x 27 ,

344

D 3 (x) := 4 x 2 ( 5 + 16 x 3 ) 2 ( −25 − 780 x 3 + 960 x 6 + 1024 x 9 ) 2 ,

345

suggests a blow-up of the numerical coefficients of the (here rational) functions φ n (x),

346

as well as an (exponentially) increasing computational complexity. Note that in (54)

347

φ n (x) ∼ x as x → +∞, for n = 1, 2, 3, as it should be expected; for instance, the

348

coefficients of the leading terms in N 3 (x) and D 3 (x) both coincide with 1024 3 .

349

The method which turned out to be the most efficient one, exploits the analyticity

350

of the coefficient, q, and the ensuing analyticity of all functions φ n , reducing the com-

351

putational complexity to O(n 3 ) (see [29]). Indeed, knowing the Taylor expansion of

352

q(x) about a given point up to the order 2n suffices to obtain φ n at such a point, by

353

simple rational manipulations. From the iterative scheme in (12), it is in fact easily

354

seen that, having a number of coefficients, say k, for the Taylor expansion of φ n , results

355

in obtaining k − 2 coefficients for φ n +1 , the key-step being the trivial algorithm used

356

to compute ratios of (formal) power series (see for instance [23, Ch. 11, p. 20]). The

357

problem of computing the 2n Taylor coefficients needed to start such a local evaluation

358

algorithm, is straightforward when q(x) is a polynomial. Otherwise, when q(x) is a

359

rational function or, more generally, is an analytic function possessing the asymptotic

360

structure in (2), one can take advantage from the highly efficient methods for repeated

361

numerical differentiation of analytic functions developed, e.g., in [11,20] and even

362

better in the recent work in [2].

363

In the following subsections, we present first some examples of numerical

364

computation of zeros of solutions to asymptotically polynomial equations of the type

365

of equation (1), to illustrate the performance of our method. We then show how to solve

366

numerically, by the same “phase function” method, a given boundary-value problem

367

associated to equation (1). In fact, we stress that the present approach yields both a

368

way to compute zeros of solutions, without computing first the solutions themselves,

369

and also a procedure to approximate a certain basis of solutions, and thus the solution

370

to any given initial- or even boundary-value problem.

371

4.1 Approximating zeros

372

In this subsection we are concerned with the numerical approximation of zeros of

373

a given solution to equation (1) in four different cases. All examples have been

374

worked out within the Mathematica environment [34], with an extended precision of 18

375

significant figures. Part of these results were anticipated in [5,25,33].

376

Example 1 (Airy equation) Consider equation (1 ) with q(x) = x (Airy equation

377

[1,23]). The first 20 zeros of the solution J 1/3 (x) (Bessel function of the first kind

378

of order 1/3) have been computed starting from the 21th zero, which was obtained

379

independently with Mathematica.

380

Author Proof

(17)

uncorrected

proof

Fig. 1 Approximations of φ(x) by φ

n

(x) with n = 0, . . . , 4, when q(x) = x

Fig. 2 log( ˜x −1

k

−x

k

) for k = 1, 2, . . . , 20, using φ

n

(x) with n = 0, . . . , 4, when q(x) = x

In practice, equation (52) has been solved as described at the beginning of Sect. 4,

381

replacing x k +1 with the approximated value x k (n) +1 , in turn approximated by the Newton

382

method with an absolute error smaller than 10 −16 .

383

In Fig. 1, few functions φ n (x) approximating φ(x) are plotted for this case. In

384

Fig. 2, the corresponding errors made are shown as functions of n and k. Here one

385

can see clearly the fast approach of the functions φ n (x) to φ(x) as n increases, and

386

the decay of the error on the k-th zero as k grows, for fixed n.

387

Example 2 (generalized Airy equation) In Table 1, we show only the (absolute) errors

388

on the first 20 positive zeros (computed as above) of the solution x 1/2 J 1/4 (x 2 /2) to

389

equation (1 ) with q(x) = x 2 (see [1]).

390

Example 3 (a perturbed generalized Airy equation) Here, the case of a perturbed

391

generalized Airy equation is considered, taking q(x) = x 2 − 3/(4 x 2 ) in equation (1).

392

Author Proof

(18)

uncorrected

proof

Fig. 3 Approximations of φ(x) = x 2 by φ

n

(x) with n = 0, . . . , 3, when q(x) = x 24x 3

2

Fig. 4log( ˜x 1

k

−x

k

) for k = 1, . . . , 20 using φ

n

(x) with n = 0, . . . , 4, when q(x) = x 24x 3

2

This coefficient is obtained by equation (6 ) choosing the phase α(x) = x 2 /2, and thus

393

the zeros of the solution y(x) = |α (x) | −1/2 sin α(x) = x −1/2 sin (x 2 /2) are known

394

exactly, x k = √

2kπ . We get from this the value of the 21th zero, and then compute

395

the first 20 zeros by our algorithm, for n = 0, 1, . . . , 4. In Fig. 3, few functions φ n (x),

396

approximating the function φ(x) relevant to this case are plotted. In in Fig. 4, the

397

corresponding errors are shown, as in the previous example.

398

Example 4 (a cubic polynomial coefficient) Finally, we consider, as coefficient in (1),

399

the cubic polynomial q(x) = x 3 − x + 1. In this case, we show the errors made in

400

computing the 10 consecutive zeros of the solution vanishing at x = 6, located on the

401

left of such a zero. The same is done for the solution vanishing at x = 20. In Fig. 5,

402

we plotted again few approximating functions φ n (x).

403

Recall that the choice of a given zero defines a solution to equation (1) up to a

404

multiplicative constant.

405

Author Proof

(19)

uncorrected

proof

Fig. 5 Approximations of φ(x) by φ

n

(x) with n = 0, . . . , 4, when q(x) = x 3 − x + 1

4.2 Numerical solution of a boundary-value problem

406

The “phase function” method described in the previous sections, besides providing the

407

direct computation of zeros of solutions (see Sect. 4.1), makes it possible to approxi-

408

mate a (Liouville-Green) basis, and thus any solution, to a given initial or boundary-

409

value (BV) problem for equation (1). In fact, knowing any phase, α(x), a basis can be

410

constructed as in (7). In practice, we obtain an approximate basis through the iterative

411

scheme in (32),

412

u n (x) := φ −1/4 n (x) sin

x



φ n 1/2 (t ) dt,

⎠ , v n (x) := φ −1/4 n cos

x



φ n 1/2 (t ) dt

⎠ ,

413

(54)

414

recalling that φ n (x) approximates φ L G (x) = (α L G ) 2 , see Theorem 3.2. We also recall

415

that the choice of the primitive in (54) is irrelevant to the purpose of approximating a

416

basis, see [3,4,30].

417

Consider the following BV problem for equation (1): y ′′ + q(x)y = 0 on a < x <

418

b, y(a) = A, y(b) = B, with a coefficient q(x) belonging to the class described in

419

Sect. 1 . Then, its solution has the form y(x) = c 1 u(x) + c 2 v(x), (u(x), v(x)) being,

420

e.g., the L G basis in (7), and the constants c 1 , c 2 can be determined as functions of

421

a, b, A, B. We shall use, instead, the approximate solution, y n (x) := c 1 (n)u n (x) +

422

c 2 (n)v n (x), where u n (x), v n (x) are given in (54). The “constants” c 1 (n), c 2 (n) can

423

also be determined approximately as functions of the known data. It is lengthy but

424

elementary to establish that the error made in evaluating y(x) on the interval [a, b] by

425

means of y(x) can be estimated as

426

|y(x) − y n (x) | ≤ const. x n −m/2 sup

x>x

n

n (x) − φ L G (x) |, for b > x > a > x n ,

427

(55)

428

Author Proof

(20)

uncorrected

proof

where the sup term can be estimated as in Remark 3.3, and x n is defined in Theorem 3.2.

429

We give now an example of application of such a procedure.

430

Example 5 Consider the two-point BV problem for the perturbed generalized Airy

431

equation of Example 3,

432

y ′′ +

 x 2 − 3

4x 2



y = 0, a < x < b, y(a) = A, y(b) = B. (56)

433

Tables 4 and5 refer to two cases for the exact solution y(x) = |α (x) | −1/2 [sin α(x) −

434

cos α(x)] = (2/x) 1/2 sin(x 2 / 2 − 1/4), see Example 3. The two cases are obtained

435

choosing the intervals 6 < x < 8, and 9.5 < x < 11.5, respectively. Clearly, A

436

and B are evaluated from such a test solution. In Tables 4 and5, the effectiveness of

437

the method is shown, since the error turns out to be of order of 10 −11 and 10 −14 ,

438

respectively, already with n = 4.

439

5 Comparing the “phase function” method with a standard high-order

440

algorithm

441

The problems considered in the previous examples still concern solutions with an

442

oscillatory behavior not so hard, given the modest powers there, but they may become

443

more challenging on larger intervals. To test the effectiveness of the “phase func-

444

tion” method to solve highly oscillatory problems, we compared its performance with

445

that of a standard but highly accurate method. Instead of resorting to Runge–Kutta

446

methods combined with dense output, whose order seems to be not higher than 8 in

447

the existing literature, we prefer to use the MATLAB RKN12(10) code (also denoted

448

by RKN1210), which is a 12th/10th order Runge–Kutta–Nyström integrator with

449

automatic control of the step size. This is widely used in problems where extremely

450

stringent error tolerances are required [9,10,15].

451

Example 6 We considered two examples concerning a perturbed generalized Airy

452

equation, whose phase function α(x), and hence the coefficient q(x), can be

453

estimated, and the exact solutions are known.Thus, the accuracy of both, our method

454

and the RKN12(10) can be compared computing the discrepancy between the approx-

455

imate and the exact solution. In case (a) we chose the same equation of Example 5,

456

i.e., q(x) = (x 2 − 3/(4x 2 )) (for which α(x) = x 2 /2), in case (b) we chose

457

q(x) = (x 6 − 15/(4x 2 ) (corresponding to α(x) = x 4 /4); see Figs. 6 and 7. The

458

numerical solution based on RKN12(10) of course is faster, in this examples, since

459

our method exploited symbolic manipulations. On the other hand, our method works

460

for solutions with arbitrarily rapid oscillations, and has an infinite accuracy in those

461

parts of the code where symbolic manipulations are used.

462

Example 7 Moreover, in several cases the RKN12(10) method fails completely, for

463

instance when the step-size falls below the smallest acceptable value (t f i n −t i n )/10 12 ,

464

see [10]. An example of such occurrence is encountered solving the initial-value prob-

465

lem for the harmonic oscillator, y ′′2 y = 0, x > 0, y(0) = 0, y ( 0) = ω. Choosing

466

Author Proof

(21)

uncorrected

proof

Fig. 6 Absolute error, |y(x) − y

n

(x) |, made solving the Cauchy problem of case (a) in Example 6, with RKN12(10), on the interval [9, 11]

Fig. 7 Absolute error, |y(x) − y

n

(x) |, made solving the Cauchy problem of case (b) in Example 6, with RKN12(10), on the interval [9, 11]

increasingly larger values of ω, we found that when ω = 10 5 the algorithm RK12(10)

467

starts failing, while our method works, yielding a solution with an error about equal

468

to zero. This performance of our method was observed even for larger values of ω.

469

In closing, we remark that the method proposed by Glaser et al. [12] presents some

470

similarities with ours: Both methods need one starting zero accurately computed; both

471

try to compute zeros solving directly a suitable, related ODE, circumventing the need

472

Author Proof

Riferimenti

Documenti correlati

Nell’arte totale converge l’interesse di D’Annunzio per molti campi della cultura e delle arti.. Le arti figurative, sulla scorta di Berenson e Pater, (p. 90) sono viste

Anna Hospital Catanzaro, IT, 13Division of Cardiac Surgery, University of Parma Parma, IT, 14Department of Cardiac Surgery, Centro Cardiologico - Fondazione Monzino IRCCS, University

(A) Animal treated with the murine anti IL-1β antibody showed a significant reduction in the level of NETosis plasma marker MPO-DNA complexes (n = 7) (B) Levels of NET-associated

Per quanto riguarda il livello di sedazione raggiunto, riportato in tabella 3.9 sono emerse differenze tra il gruppo A che conteneva tutti e quattro i farmaci e i gruppi B (dove non

27 Sambon è coinvolto anche nell‘incanto degli avori del Comune di Volterra a Firenze il 2 dicembre 1880, una vendita di oggetti provenienti dal locale museo etrusco, orga-

L’ambiguo umore di Pechino ha certamente creato un problema di attendibilità statistica in merito alle stime numeriche, altalenanti tra l,5% e il 7,5% della popolazione,

CA_H: camel hair, CA_K: soluble camel hair keratin, CA_R: camel hair residue, CS_H: cashmere, CS_K: soluble cashmere keratin, CS_R: cashmere residue..

[r]