1 s2.0 S0022247X04003105 Main
1 s2.0 S0022247X04003105 Main
Abstract
We provide a local as well as a semilocal convergence analysis for two-point Newton-like methods
in a Banach space setting under very general Lipschitz type conditions. Our equation contains a
Fréchet differentiable operator F and another operator G whose differentiability is not assumed.
Using more precise majorizing sequences than before we provide sufficient convergence conditions
for Newton-like methods to a locally unique solution of equation F (x) + G(x) = 0. In the semilocal
case we show under weaker conditions that our error estimates on the distances involved are finer
and the information on the location of the solution at least as precise as in earlier results. In the
local case a larger radius of convergence is obtained. Several numerical examples are provided to
show that our results compare favorably with earlier ones. As a special case we show that the famous
Newton–Kantorovich hypothesis is weakened under the same hypotheses as the ones contained in
the Newton–Kantorovich theorem.
2004 Elsevier Inc. All rights reserved.
0022-247X/$ – see front matter 2004 Elsevier Inc. All rights reserved.
doi:10.1016/j.jmaa.2004.04.008
I.K. Argyros / J. Math. Anal. Appl. 298 (2004) 374–397 375
1. Introduction
In this study we are concerned with the problem of approximating a locally unique
solution x ∗ of the nonlinear equation
F (x) + G(x) = 0, (1)
where F, G are operators defined on a closed ball Ū (w, R) centered at point w and of
radius R 0, which is a subset of a Banach space X with values in a Banach space Y .
F is Fréchet-differentiable on Ū (w, R), while the differentiability of the operator G is not
assumed.
A large number of problems in applied mathematics and also in engineering are solved
by finding the solutions of certain equations [4,7,16]. For example, dynamic systems are
mathematically modeled by difference or differential equations, and their solutions usually
represent the states of the systems. For the sake of simplicity, assume that a time-invariant
system is driven by the equation ẋ = Q(x) (for some suitable operator Q), where x is
the state. Then the equilibrium states are determined by solving Eq. (1). Similar equa-
tions are used in the case of discrete systems. The unknowns of engineering equations can
be functions (difference, differential, and integral equations), vectors (systems of linear
or nonlinear algebraic equations), or real or complex numbers (single algebraic equations
with single unknowns). Except in special cases, the most commonly used solution methods
are iterative—when starting from one or several initial approximations a sequence is con-
structed that converges to a solution of the equation. Iteration methods are also applied for
solving optimization problems. In such cases, the iteration sequences converge to an op-
timal solution of the problem at hand. Since all of these methods have the same recursive
structure, they can be introduced and discussed in a general framework.
We use the two-point Newton method
(see (20)), whereas in the local case we can provide a larger convergence radius using the
same information (see (134) and (135)).
Part A. Motivation
Deuflhard and Heindl [13] have proved the following affine invariant form of the
Newton–Kantorovich theorem [16] which is the motivation for this study.
In order for us to show that these observations hold in a more general setting we first
need to introduce the following assumptions.
Let R 0 be given. Assume there exist v, w ∈ X such that A(v, w)−1 ∈ L(Y, X), and
for any x, y, z ∈ Ū (w, r) ⊆ Ū (w, R), t ∈ [0, 1], the following hold:
A(v, w)−1 A(x, y) − A(v, w) h0 x − v, y − w + a (25)
and
A(v, w)−1 F y + t (z − y) − A(x, y) (z − y) + G(z) − G(y)
h1 y − w + tz − y − h2 y − w + h3 z − x + b z − y, (26)
where h0 (r, s), h1 (r + r̄) − h2 (r) (r̄ 0), h2 (r), h3 (r) are monotonically increasing func-
tions for all r, s on [0, R]2 , [0, R]2 , [0, R], [0, R], respectively, with h0 (0, 0) = h1 (0) =
h2 (0) = h3 (0) = 0, and the constants a, b satisfy a 0, b 0. Given y−1 , y0 , v, w in X,
define parameters c−1 , c, c1 by
y−1 − v c−1 , y−1 − y0 c, v − w c1 . (27)
Remark 1. Conditions similar to (21)–(26) but less flexible were considered by Chen and
Yamamoto in [11] in the special case when A(x, y) = A(x) for all x, y ∈ Ū (w, R) (A(x) ∈
L(X, Y )) (see also Theorem 4). Operator A(x) is intended there to be an approximation to
the Fréchet-derivative F (x) of F . However we also want the choice of operator A to be
more flexible, and be related to the difference G(z) − G(y) for all y, z ∈ Ū (w, R). It has
already been shown in special cases in [3,4,10] that this way the ratio of convergence for
method (2) is improved (see also Application 1). Note also that if we choose
With the above choices, we show the following result on majorizing sequences for
method (2).
δ n+1 δ n+2
1− 1−
h0 2
η + c + c−1 , 2
η + r0 + a < 1, (31)
1− δ
2 1− δ
2
and
1 δ n+1 n+1
n+1
1− δ 1 − 2δ
2 h1 2
η+θ η + r0 dθ − 2h2 η + r0
1− δ
2
2 1 − 2δ
0
n
n+1 n+2
δ δ 1 − 2δ 1 − 2δ
+ 2h3 1+ η + δh0 η + c + c−1 , η + r0
2 2 1 − 2δ 1 − 2δ
δ (32)
for all n 0. Then, iteration {tn } (n −1) given by
t−1 = r0 , t0 = c + r0 , t1 = c + r0 + η,
1
{ 0 [h1 (tn −t0 +r0 +θ(tn+1 −tn ))−h2 (tn −t0 +r0 )+b] dθ+h3 (tn+1 −tn−1 )}(tn+1 −tn )
tn+2 = tn+1 + 1−a−h0 (tn −t−1 +c−1 ,tn+1 −t0 +r0 )
(33)
is monotonically increasing, bounded above by
2η
t ∗∗ = + r0 + c, (34)
2−δ
and converges to some t ∗ such that
0 t ∗ t ∗∗ R. (35)
Moreover the following error bounds hold for all n 0:
n+1
δ δ
0 tn+2 − tn+1 (tn+1 − tn ) η. (36)
2 2
0 tk+1 − tk , (38)
and
h0 (tk − t−1 + c−1 , tk+1 − t0 + r0 ) + a < 1 (39)
for all k 0.
380 I.K. Argyros / J. Math. Anal. Appl. 298 (2004) 374–397
1
2 h1 (r0 + θ η) dθ − h2 (r0 ) + b + h3 (c + η) + δ a + h0 (c + c−1 , η + r0 ) δ,
0
0 t1 − t0 ,
h0 (c + c−1 , η + r0 ) + a < 1,
1
2 h1 tk+1 − t0 + r0 + θ (tk+2 − tk+1 ) − h2 (tk+1 − t0 + r0 ) + b dθ
0
+ h3 (tk+2 − tk ) + δ a + h0 (tk+1 − t−1 + c−1 , tk+2 − t0 + r0 )
by (29) and (32). Hence we showed (37) holds for k = n + 2. Moreover, we must show
tk t ∗∗ , (40)
t−1 = r0 t ∗∗ , t0 = r0 + c t ∗∗ , t1 = c + r0 + η t ∗∗ ,
δ 2+δ
t2 c + r0 + η + η = η + r0 + c t ∗∗ .
2 2
Assume that (40) holds for all k n + 1. It follows from (33), (37)–(39) that
I.K. Argyros / J. Math. Anal. Appl. 298 (2004) 374–397 381
δ δ δ
tk+2 tk+1 + (tk+1 − tk ) tk + (tk − tk−1 ) + (tk+1 − tk )
2 2 2
2 k+1
δ δ δ
· · · c + r0 + η + η + η + ···+ η
2 2 2
k+2
1 − 2δ 2η
= η + r0 + c + r0 + c = t ∗∗ . (41)
1− 2δ 2−δ
Hence, sequence {tn } (n −1) is bounded above by t ∗∗ . Inequality (39) holds for k = n+2
by (30) and (31). Moreover (38) holds for k = n + 2 by (41) and since (37) and (39) also
hold for k = n + 2. Furthermore, sequence {tn } (n 0) is monotonically increasing by (38)
and as such it converges to some t ∗ satisfying (35).
That completes the proof of Lemma 1. 2
We provide the main result on the semilocal convergence of method (2) using majorizing
sequence (33).
Proof. We first show estimate (44), and yn ∈ Ū (w, t ∗ ) for all n −1. For n = −1, 0,
(44) follows from (27), (33) and (43). Suppose (44) holds for all n = 0, 1, . . . , k + 1; this
implies in particular (using (27), (42))
382 I.K. Argyros / J. Math. Anal. Appl. 298 (2004) 374–397
x−1 = v, x0 = w,
xn+1 = xn − A(xn−1 , xn )−1 F (xn ) + G(xn ) (n 0). (53)
Iteration {xn } (n −1) is a special case of {yn } (n −1). Hence, we have
xk+1 − xk t¯k+1 − t¯k , lim xn = x ∗ ,
n→∞
and
x ∗ − xk t ∗ − t¯k , lim t¯k = t ∗ , (54)
where {t¯n } is {tn } (n −1) for r0 = 0.
We shall show
y ∗ − xk t ∗ − t¯k . (55)
For k = 0, (54) holds since y∗ ∈ Ū (w, t ∗ ). Suppose (54) holds for all n k. Then as in
(51) we obtain the identity
y ∗ − xk+1 = y ∗ − xk + A(xk−1 , xk )−1 F (xk ) + G(xk )
− A(xk−1 , xk )−1 F (y ∗ ) + G(y ∗ )
= A(xk−1 , xk )−1 A(x−1 , x0 ) A(x−1 , x0 )−1 F (xk ) − F (y ∗ )
− A(xk−1 , xk )(y ∗ − xk ) + G(xk ) − F (y ∗ ) . (56)
Using (56) we obtain in turn
1
[ 0 h1 (xk −x0 +t y ∗ −xk ) dt −h2 (xk −x0 )+h3 (y ∗ −xk−1 )+b]y ∗ −xk
y ∗ − xk+1 1−a−h0 (xk−1 −x−1 ,xn −x0 )
1
[ h1 ((1 + t)t ) dt − h2 (t ∗ ) + h3 (t ∗ ) + b] ∗
∗
0 y − xk
1 − a − h0 (t ∗ + c1 , t ∗ )
< y ∗ − xk t ∗ − t¯k → 0 as k → ∞. (57)
That is, x∗ = y∗.
If y ∗ ∈ Ū (x0 , R0 ) then as in (57) we get
1
∗ [ 0 h1 (t ∗ + tR0 ) dt − h2 (t ∗ ) + h3 (R0 ) + b] ∗
y − xk+1 y − xk
1 − a − h0 (t ∗ + c1 , t ∗ )
< y ∗ − xk . (58)
x∗
Hence, again we get = y∗.
That completes the proof of Theorem 2. 2
Remark 2. Conditions (31), (32) can be replaced by the stronger but easier to check
2η 2η
h0 + c + c−1 , + r0 + a 1 (59)
2−δ 2−δ
384 I.K. Argyros / J. Math. Anal. Appl. 298 (2004) 374–397
and
1
2η δ 2η
2 h1 + θ + r0 dθ − 2h2 + r0
2−δ 2 2−δ
0
δ 2η 2η
+ 2h3 1 + η + δh0 + c + c−1 , + r0
2 2−δ 2−δ
δ, (60)
respectively. Note also that conditions (29)–(32), (59), (60) are of the Newton–Kantoro-
vich-type hypotheses (see also (5)) which are always present in the study of Newton-like
methods [4,8,11,12,24,26].
Application 1. Let us consider some special choices of operator A, functions hi , i = 0, 1,
2, 3, parameters a, b and points v, w.
Define
for all x, y, z, w ∈ Ū (y0 , r) are used there instead of (67)–(70), where [x, y, z; G] denotes
a second order divided difference of G at (x, y, z), and γi , i = 6, 7, 8, 9, are non-negative
parameters.
Using the method of chord (i.e., (66) with A(wn ) = [wn−1 , w1 ; G]) with w0 = (5, 5),
w−1 = (1, 0), we obtain
(1) (2)
n wn wn wn − wn−1
0 5 5
1 1 0 5.000E+00
2 0.989800874210782 0.012627489072365 1.262E−02
3 0.921814765493287 0.307939916152262 2.953E−01
4 0.900073765669214 0.325927010697792 2.174E−02
5 0.894939851625105 0.327725437396226 5.133E−03
6 0.894658420586013 0.327825363500783 2.814E−04
7 0.894655375077418 0.327826521051833 3.045E−04
8 0.894655373334698 0.327826521746293 1.742E−09
9 0.894655373334687 0.327826521746298 1.076E−14
10 0.894655373334687 0.327826521746298 5.421E−20
Using our method (64) with y0 = (5, 5), y−1 = (1, 0), we obtain
(1) (2)
n yn yn yn − yn−1
0 5 5
1 1 0 5
2 0.909090909090909 0.363636363636364 3.636E−01
3 0.894886945874111 0.329098638203090 3.453E−02
4 0.894655531991499 0.327827544745569 1.271E−03
5 0.894655373334793 0.327826521746906 1.022E−06
6 0.894655373334687 0.327826521746298 6.089E−13
7 0.894655373334687 0.327826421746298 2.710E−20
We did not verify the hypotheses of Theorem 3 for the above starting points. However, it is
clear that the hypotheses of Theorem 3 are satisfied for all three methods for starting points
closer to the solution
x ∗ = (0.894655373334687, 3.27826421746298)
Application 2. Returning back to Remark 1 and (28), iteration (2) reduces to the famous
Newton–Kantorovich method (8).
Condition (29) reduces to
1 k+1 k+1
2η δ δ
2 γ1 1− +θ η dθ
2−δ 2 2
0
k+1
k+1
2η δ 2η δ
− 2γ1 1− + δγ0 1−
2−δ 2 2−δ 2
1
2γ1 θ η dθ + δγ0 η
0
or
k+1
γ0 δ 2 δ
− γ1 1− 1,
2−δ 2
which is true for all k 0 by the choice of δ. Furthermore (31) gives
2γ0η
2γ0 η < 1.
2−δ
Hence in this case conditions (29), (31) and (32) reduce only to (72) provided δ ∈ [0, 1].
Condition (72) for say δ = 1 reduces to (20).
Case 2. It follows from Case 1 that (29), (31) and (32) reduce to (72),
2γ0η
1 (73)
2−δ
and
γ0 δ 2
γ1 , (74)
2−δ
respectively, provided δ ∈ [0, 2).
Case 3. It turns out that the range for δ can be extended (see also Example 3). Introduce
conditions
1
γ0 η 1 − δ for δ ∈ [δ0 , 2),
2
where
√
−b + b2 + 8b γ1
δ0 = , b = , and γ0
= 0.
2 γ0
Indeed the proof of Theorem 2 goes through if instead we show the weaker condition
k+1 k+2
δ 2γ0 δ δ
γ1 + 1− δ,
2 2−δ 2
or
k+1
bδ δ 2 δ
b− − 0,
2 2 2
388 I.K. Argyros / J. Math. Anal. Appl. 298 (2004) 374–397
or
δ δ0 ,
which is true by the choice of δ0 .
Example 3. Returning back to Example 1 but using Case 3 we can do better. Indeed, choose
√
5 − 13
p = p0 = 0.4505 < = 0.464816242 . . ..
3
Then we get
η = 0.183166 . . ., γ0 = 2.5495, γ1 = 3.099, and δ0 = 1.0656867.
Choose δ = δ0 . Then we get
δ0
γ0 η = 0.466983415 < 1 − = 0.46715665.
2
√
That is the interval [(5 − 13 )/2, 1/2) can be extended to at least [p0 , 1/2).
In the example that follows we show that γ1 /γ0 can be arbitrarily large. Indeed
In order to compare with earlier results, we consider the case when x = y and v = w
(single step methods). We can then prove along the same lines to Lemma 1 and Theorem 2,
respectively, the following results by assuming that there exists w ∈ X such that A(w)−1 ∈
L(Y, X), for any x, y ∈ Ū (w, r) ⊆ Ū (w, R), t ∈ [0, 1],
A(w)−1 A(x) − A(w) g0 x − w + α (76)
and
A(w)−1 F (x + t (y − x)) − A(x) (y − x) + G(y) − G(x)
g1 x − w + ty − x − g2 x − w + g3 (r) + β y − x, (77)
where g0 , g1 , g2 , g3 , α, β are as h0 (one variable), h1 , h2 , h3 , a, b, respectively.
Then we can show the following result on majorizing sequences.
Lemma 2. Assume that there exist η 0, α 0, β 0, δ ∈ [0, 2), r0 ∈ [0, R] such that
I.K. Argyros / J. Math. Anal. Appl. 298 (2004) 374–397 389
1
h̄δ = 2 g1 (r0 + θ η) dθ − g2 (r0 ) + g3 (r0 + η) + β + δ α + g0 (r0 + η) δ,
0
(78)
2η
+ r0 R, (79)
2−δ
n+1
2η δ
g0 1− + r0 + α < 1, (80)
2−δ 2
1 n+1 n+1
2η δ δ
2 g1 1− + r0 + θ η dθ
2−δ 2 2
0
n+1
n+1
2η δ 2η δ
−2g2 1− + r0 + 2g3 1− + r0
2−δ 2 2−δ 2
n+1
2η δ
+ δg0 1− + r0
2−δ 2
δ (81)
for all n 0. Then, iteration {sn } (n 0) given by
s0 = r0 , s1 = r0 + η,
1 s
{g1 (sn +θ(sn+1 −sn ))−g2 (sn )+β} dθ(sn+1 −sn )+ snn+1 g3 (θ) dθ
sn+2 = sn+1 + 0
1−α−g0 (sn+1 ) (82)
Theorem 3. Assume that hypotheses of Lemma 2 hold and there exists y0 ∈ Ū (w, r0 ) such
that
A(y0 )−1 F (y0 ) + G(y0 ) η. (86)
Then, sequence {wn } (n 0) generated by Newton-like method (66) is well defined, re-
mains in Ū (w, s ∗ ) for all n 0, and converges to a solution x ∗ of equation F (x) + G(x)
= 0. Moreover the following error bounds hold for all n 0:
wn+1 − wn sn+1 − sn (87)
390 I.K. Argyros / J. Math. Anal. Appl. 298 (2004) 374–397
and
wn − x ∗ s ∗ − sn . (88)
Furthermore the solution x∗ is unique in Ū (w, s ∗ ) if
1
g1 (s ∗ + θ s ∗ ) − g2 (s ∗ ) dθ + g3 (s ∗ ) + g0 (s ∗ ) + α + β < 1, (89)
0
We state the relevant results due to Chen and Yamamoto [11, p. 40]. We assume that
A(w)−1 exists, and for any x, y ∈ Ū (w, r) ⊆ Ū (z, R),
0 < A(w)−1 F (w) + G(w) η̄, (91)
A(w)−1 A(x) − A(w) ḡ0 x − w + ᾱ, (92)
A(w)−1 F x + t (y − x) − A(x)
ḡ1 x − w + t y − x − ḡ0 x − w + β̄, t ∈ [0, 1], (93)
A(w) G(x) − G(y) g3 (r)x − y,
−1
(94)
where ḡ0 , ḡ1 , ᾱ, β̄ are as g0 , g1 , α, β, respectively, but ḡ0 is also differentiable with ḡ0 (r)
> 0, r ∈ [0, R], and ᾱ + β̄ < 1.
As in [11], set
r r
ϕ(r) = η̄ − r + ḡ1 (t) dt, ψ(r) = g3 (t) dt, (95)
0 0
χ(r) = φ(r) + ψ(r) + (ᾱ + β̄)r. (96)
Denote the minimal value of χ(r) on [0, R] by χ ∗,
and the minimal point by r ∗.
If χ(R)
0, denote the unique zero of χ by r0∗ ∈ (0, r ∗ ]. Define scalar sequence {rn } (n 0) by
u(rn )
r0 ∈ [0, R], rn+1 = rn + (n 0), (97)
g(rn )
where
u(r) = χ(r) − x ∗ (98)
and
g(r) = 1 − ḡ0 (r) − ᾱ. (99)
With the above notation they showed
I.K. Argyros / J. Math. Anal. Appl. 298 (2004) 374–397 391
Theorem 4 [11, p. 40]. Suppose χ(R) 0. Then Eq. (1) has a solution x ∗ ∈ Ū (w, r0∗ ),
which is unique in
Ū (w, R) if χ(R) = 0 or ψ(R) = 0, and r0∗ < R,
Ũ = (100)
U (w, R) if χ(R) = 0 and r0∗ < R.
Let
u(r)
∗ −1
D = Ūr∈[0,r ∗) y ∈ Ū (w, r) A(y) F (y) + G(y) . (101)
g(r)
Then, for any y0 ∈ D, sequence {yn } (n 0) generated by Newton-like method (66) is well
defined, remains in Ū (w, r ∗ ) and satisfies
yn+1 − yn rn+1 − rn (102)
and
yn − x ∗ r ∗ − rn (103)
provided that r0 is chosen as in (97) so that r0 ∈ Ry0 , where for y ∈ D ∗ ,
∗
u(r)
−1
Ry = r ∈ [0, r ) A(y) F (y) + G(y) , y − z r . (104)
y(r)
Remark 4. The conclusions of Theorem 4 hold (i.e., the results in [11] were improved) if
the more general conditions (76), (77) replace (92)–(94), and
ḡ0 (r) g2 (r), r ∈ [0, R], (105)
is satisfied. Moreover if strict inequality holds in (105) we obtain more precise error
bounds. Indeed, define the sequence {r̄n } (n 0), using (77), g2 instead of (93), ḡ0 , re-
spectively (with ḡ1 = g1 , α = ᾱ, β = β̄), by
r̄0 = r0 , r̄1 = r1 ,
u(r̄n ) − u(r̄n−1 ) + (1 − g2 (r̄n−1 ) − ᾱ)(r̄n − r̄n−1 )
r̄n+1 − r̄n = (n 1). (106)
g(r̄n )
It can easily be seen using induction on n (see also the proof of Proposition 1 that follows)
that
r̄n+1 − r̄n < rn+1 − rn , (107)
r̄n < rn , (108)
∗ ∗ ∗
r̄ − r̄n r − rn , r̄ = lim r̄n , (109)
n→∞
and
r̄ ∗ r ∗ . (110)
Furthermore condition (77) allows us more flexibility in choosing functions and constants.
392 I.K. Argyros / J. Math. Anal. Appl. 298 (2004) 374–397
Remark 5. Returning back to Newton’s method (8) (see also (28)), the iterations cor-
responding to (97) and (106) are (13) and (17), respectively. Moreover condition (105)
reduces to (19), and in case γ0 < γ1 , estimates (22)–(24) hold.
Remark 6. Our error bounds (87), (88) are finer than the corresponding ones (102) and
(103), respectively in many interesting cases.
Let us choose
α = ᾱ, β = β̄, g0 (r) = ḡ0 (r), g1 (r) = g2 (r) = ḡ1 (r), and
ḡ3 (r) = g3 (r) for all r ∈ [0, R].
Then we can show
Proof. It suffices to show (112) and (113), since then (114) and (115), respectively, can
easily follow. Inequality (112) holds for n = 1 by (111). By (82) and (97) we get in turn
1
{g1 (s0 +θ(s1 −s0 )) dθ−g2 (s0 )+α}(s1 −s0 )+ ss1 g3 (θ) dθ
s2 − s1 =
0 0
1−β−g0 (s1 )
1 r
0 {ḡ1 (r0 +θ(r1 −r0 )) dθ−ḡ2 (r0 )+ᾱ}(r1 −r0 )+ r ḡ3 (θ) dθ
1
0
< 1−β̄−ḡ0 (r1 )
u(r1 ) − u(r0 ) + g(r0 )(r1 − r0 ) u(r1 )
= = = r2 − r1 . (116)
1 − β̄ − ḡ0 (r1 ) g(r1 )
Assume
sk+1 < rk+1 (117)
and
sk+1 − sk < rk+1 − rk (118)
hold for all k n.
Using (82), (88), and (118) we obtain
I.K. Argyros / J. Math. Anal. Appl. 298 (2004) 374–397 393
1 s
{g1 [sk +θ(sk+1 −sk )] dθ−g2 (sk )+α}(sk+1 −sk )+ skk+1 g3 (θ) dθ
sk+2 − sk+1 = 0
1−β−g0 (sk+1 )
1 rk+1
0 {ḡ1 [rk +θ(rk+1 −rk )] dθ−ḡ2 (rk )+ᾱ}(rk+1 −rk )+ rk ḡ3 (θ) dθ
< 1−β̄−ḡ0 (rk+1 )
u(rk+1 ) − u(rk ) + g(rk )(rk+1 − rk ) u(rk+1 )
= = = rk+2 − rk+1 .
g(rk+1 ) g(rk+1 )
That completes the proof of Proposition 1. 2
In order for us to include a case where operator G is nontrivial, we consider the follow-
ing example for Theorem 2 (or Theorem 3).
1
x(t) = k t, s, x(s) ds, (119)
0
where the kernel k(t, s, x(s)) with (t, s) ∈ [0, 1] × [0, 1] is a nondifferentiable operator on
Ū (x0 , R/2). Define operators F, G on Ū (x0 , R/2) by
Choose x0 = 0, and assume there exists a constant θ0 ∈ [0, 1), a real function θ1 (t, s) such
that
k(t, s, x) − k(t, s, y) θ1 (t, s)x − y (122)
and
1
sup θ1 (t, s) ds θ0 (123)
t ∈[0,1]
0
In order to cover the local case, let us assume x ∗ is a zero of Eq. (1), A(x ∗ , x ∗ )−1 exists
and for any x, y ∈ Ū (x ∗ , r) ⊆ Ū (x ∗ , R), t ∈ [0, 1],
A(x ∗ , x ∗ )−1 A(x, y) − A(x ∗ , x ∗ ) h̄0 x − x ∗ , y − x ∗ + ā (125)
and
A(x ∗ , x ∗ )−1 F x ∗ + t (y − x ∗ ) − A(x, y) (y − x ∗ ) + G(y) − G(x ∗ )
h̄1 y − x ∗ (1 + t) − h̄2 y − x ∗ + h̄3 x − x ∗ + b̄ y − x ∗ , (126)
where h̄0 , h̄1 , h̄2 , h̄3 , ā, b̄ are as h0 , h1 , h2 , h3 , a, b, respectively. Then exactly as in (56)
but using (125), (126), instead of (25), (26) we can show the following local result for
method (2).
Denote by λ0 the smallest of the solutions in [0, R]. Then, sequence {xn } (n −1) gen-
erated by Newton-like method (2) is well defined, remains in Ū (x ∗ , λ0 ) for all n 0 and
converges to x ∗ provided that x−1 , x0 ∈ Ū (x ∗ , λ0 ). Moreover the following error bounds
hold for all n 0:
x ∗ − xn+1 pn , (129)
where
{ 01 [h̄1 ((1+t )xn−x ∗ )−h̄2 (xn −x ∗ )] dt +ā+h̄3 (xn−1 −x ∗ )}
pn = 1−b̄−h̄0 (xn −x ∗ )
xn − x ∗ . (130)
Application 3. Let us again consider Newton’s method, i.e., F (x) = A(x, y), G(x) = 0,
and assume
∗ −1
F (x ) F (x) − F (x ∗ ) λ1 x − x ∗ (131)
and
∗ −1
F (x ) F (x) − F (y) λ2 x − y (132)
for all x, y ∈ Ū (x ∗ , r) ⊆ Ū (x ∗ , R). Then we can set
xn+1 − x ∗ en , (138)
∗
xn+1 − x en1 , (139)
where
λ2 xn − x ∗ 2
en = (140)
2[1 − λ1 xn − x ∗ ]
and
λ2 xn − x ∗ 2
en1 = . (141)
2[1 − λ2 xn − x ∗ ]
That is
en en1 (n 0). (142)
If strict inequality holds in (136) then (137) and (142) hold as strict inequalities also (see
also Example 4).
Remark 7. As noted in [1–9,25] the local results obtained here can be used for projection
methods such as Arnoldi’s, the generalized minimum residual method (GMRES), the gen-
eralized conjugate residual method (GCR), for combined Newton/finite projection methods
and in connection with the mesh independence principle to develop the cheapest and most
efficient mesh refinement strategies.
Remark 8. The local results can also be used to solve equations of the form F (x) = 0,
where F satisfies the autonomous differential equation [4,8,16]
F (x) = P F (x) , (143)
where P : Y → X is a known continuous operator. Since F (x ∗ ) = P (F (x ∗ )) = P (0), we
can apply our results without actually knowing the solution x ∗ of Eq. (1).
396 I.K. Argyros / J. Math. Anal. Appl. 298 (2004) 374–397
The results obtained here can be extended to m-point methods (m > 2 an integer) [4–8,
17], and can be used in the solution of variational inequalities [23].
References
[1] F.L. Allgower, K. Böhmer, F.A. Potra, W.C. Rheinboldt, A mesh independence principle for operator equa-
tions and their discretizations, SIAM J. Numer. Anal. 23 (1986) 160–169.
[2] J. Appel, E. DePascale, P.P. Zabrejko, On the application of the Newton–Kantorovich method to nonlinear
integral equations of Uryson type, Numer. Funct. Anal. Optim. 12 (1991) 271–283.
[3] I.K. Argyros, Improving the rate of convergence of Newton methods on Banach spaces with a convergence
structure and applications, Appl. Math. Lett. 10 (1977) 21–28.
[4] I.K. Argyros, Advances in the Efficiency of Computational Methods and Applications, World Scientific,
River Edge, NJ, 2000.
[5] I.K. Argyros, On the radius of convergence of Newton’s method, Internat. J. Comput. Math. 77 (2001)
389–400.
[6] I.K. Argyros, A Newton–Kantorovich theorem for equations involving m-Fréchet-differentiable operators
and applications in radiative transfer, J. Comput. Appl. Math. 131 (2001) 149–159.
[7] I.K. Argyros, F. Szidarovszky, Convergence of general iteration schemes, J. Math. Anal. Appl. 168 (1992)
42–52.
[8] I.K. Argyros, F. Szidarovszky, The Theory and Applications of Iteration Methods, CRC Press, Boca Raton,
FL, 1993.
I.K. Argyros / J. Math. Anal. Appl. 298 (2004) 374–397 397
[9] P.N. Brown, A local convergence theory for combined inexact-Newton/finite-difference projection methods,
SIAM J. Numer. Anal. 24 (1987) 407–434.
[10] E. Cǎtinas, On some iterative methods for solving nonlinear equations, Rev. Anal. Numér. Théor. Approx. 23
(1994) 47–53.
[11] X. Chen, T. Yamamoto, Convergence domains of certain iterative methods for solving nonlinear equations,
Numer. Funct. Anal. Optim. 10 (1989) 37–48.
[12] J.E. Dennis, Toward a unified convergence theory for Newton-like methods, in: L.B. Rall (Ed.), Nonlinear
Functional Analysis and Applications, Academic Press, New York, 1971, pp. 425–472.
[13] P. Deuflhard, G. Heindl, Affine invariant convergence theorems for Newton’s method and extensions to
related methods, SIAM J. Numer. Anal. 16 (1979) 1–10.
[14] P. Deuflhard, F.A. Potra, Asymptotic mesh independence of Newton–Galerkin methods via a refined
Mysovskii theorem, SIAM J. Numer. Anal. 29 (1992) 1395–1412.
[15] J.M. Gutiérrez, M.A. Hernandez, M.A. Salanova, Accessibility of solutions by Newton’s method, Internat.
J. Comput. Math. 57 (1995) 239–247.
[16] L.V. Kantorovich, G.P. Akilov, Functional Analysis, Pergamon, Oxford, 1982.
[17] F.A. Potra, On an iterative algorithm of order 1.839 . . . for solving nonlinear equations, Numer. Funct. Anal.
Optim. 7 (1984–85) 75–106.
[18] F.A. Potra, V. Ptǎk, Sharp error bounds for Newton’s process, Numer. Math. 34 (1980) 67–72.
[19] W.C. Rheinboldt, An adaptive continuation process for solving systems of nonlinear equations, Banach
Center Publ. 3 (1977) 129–142.
[20] J.W. Schmidt, H. Schwetlick, Ableitungsfreie Verfahren mit hohere Konvergenzgeschwindigkeit, Comput-
ing 3 (1968) 215–226.
[21] J.W. Schmidt, H. Leonhardt, Eingrenzung von Losungen mit der Regula Falsi, Computing 6 (1970) 318–329.
[22] J.W. Schmidt, Untere Fehlerschranken fur Regula-Falsi Verhafren, Period. Math. Hungar. 9 (1978) 241–247.
[23] L.U. Uko, J.O. Adeyeye, Generalized Newton-iterative methods for nonlinear operator equations, Nonlinear
Studies 8 (2001) 465–477.
[24] T. Yamamoto, A convergence theorem for Newton-like methods in Banach spaces, Numer. Math. 51 (1987)
545–557.
[25] T.J. Ypma, Local convergence of inexact Newton methods, SIAM J. Numer. Anal. 21 (1984) 583–590.
[26] P.P. Zabrejko, D.F. Nguen, The majorant method in the theory of Newton–Kantorovich approximations and
the Ptǎk error estimates, Numer. Funct. Anal. Optim. 9 (1987) 671–684.