This content was uploaded by our users and we assume good faith they have the permission to share this book. If you own the copyright to this book and it is wrongfully on our website, we offer a simple DMCA procedure to remove your content from our site. Start by pressing the button below!
(s-t)(
0.
(1.163)
48
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
Let X be a Banach space with dual space X*, and let T> c X. As before, we use the pairing (x, j) to denote j(x), x G X, j G X*. Definition 1.19 A mapping T : V t-» X* is said to be monotone if for each u,v eV, Re(u - v, T(u) - T(v)) > 0.
(1.164)
It is called strongly monotone if for some c > 0, Re(u - v, T{u) - T{v)) > c\\x - y\\2.
(1.165)
A natural analogue of the above for mappings taking values in X is the following definition. Definition 1.20 A mapping T : V H-> X is said to be accretive if for all u, v G V and some j G J(u — v), Re(T(u) - T(v),j) > 0.
(1.166)
It is called strongly accretive if for some c > 0, (1.167)
Re(T(u) - T(v),j) > c\\x - yf.
Here J denotes the normalized duality mapping introduced earlier: for xGX, J(x) = {j G X* : (x,j) = \\x\\2 = \\j\\2}.
(1.168)
We note first that if X is a Hilbert space then X — X*, the classes of monotone (respectively, strongly monotone) and accretive (respectively, strongly accretive) mappings defined in X coincide, and (1.164) denotes the usual inner product. Thus, if X is specialized further to X = R, (1.164) becomes {u - v)(T(u) - T(v)) > 0. One connection between accretive mappings and nonexpansive mappings is immediate. If F : V H-> X is nonexpansive, then for T — I -F, x,yeV, and j G J(x - y), (T(x) - T(y),j) = {x-y=
(F(x) - F(y)),j)
\\x-y\\2-(F(x)-F(y),j)
> \\x - yf - \\F(x) - F(y)\\\\x - y\\ > 0. (1.169) Thus T is accretive. In addition, if F : V i-> X is a strict contraction, ||F(a;)-F(y)||
x,yeV,
0 < k < 1,
(1.170)
Mappings in Metric and Normed Spaces
49
then T = I — F is strongly accretive. On the other hand, not all accretive mappings are of the form I — F with F nonexpansive. A complete characterization of accretive mappings in metric terms has been given by Kato [Kato (1967)] (see also [Deimling (1974)]. Proposition 1.26 Let X be a Banach space, V C X, and T : V i-> X. Then T is accretive if and only if for each x,y € P and A > 0,
II* - 2/11 <\\x-y
+ \{T(x) - T(y))\\.
(1.171)
Thus a mapping T : T> i-> X is accretive if and only if the mapping J\ = (I + AT)" 1 (called the resolvent ofT) is nonexpansive on its domain for each positive A. Remark 1.2 Using the above, it is possible to extend the definition of accretivity in a natural way to the multivalued case. For a given subset B :xeB}. ofX, let \B\ = ini{\\x\\ A mapping T : V K-> 2X{V C X) is said to be accretive (respectively, strongly accretive) if there exists e > 0 (respectively, e > 0) such that for each x,y£T>, z£ T(x), w G T(y), and A > 0, (1 + e)\\x - 2/|| < \\x - y + X(z - w)\\.
(1.172)
Again, it can be shown that the mapping J\ = (I + AT)" 1 is singlevalued and nonexpansive (respectively, strict contraction) on its domain. If it is the case that the domain of J\ is all of X for some (hence all) A > 0, then T is said to be m-accretive (sometimes called hyperaccretive). The theory of accretive operators is extensive. The solvability of many equations involving partial differential operators (e.g., Laplacians) can be formulated as questions concerning accretive operators. We will not discuss the details here but instead refer the reader to [Browder (1976); Deimling (1992); Barbu (1976)], and [Martin (1973)]. Our motivation for including this notion here is simply to note the usefulness of fixed point theory for nonexpansive mappings in another context and to illustrate the dependence of each theory upon the other (see also [Kirk and Sims (2001)] and references therein). It is also not difficult to see that the resolvent J\ of an accretive mapping is firmly nonexpansive. As a matter of fact, F is firmly nonexpansive if and only if it is the resolvent (/ + A)~1 for some accretive mapping A C X x X.
Chapter 2
Differentiable and Holomorphic Mappings in Banach Spaces 2.1
Differentiable Mappings. Frechet Derivatives
The fundamental notions of abstract differentials, polynomials and power series were introduced by Maurice Prechet around 1909. The crux of the theory of functions between normed spaces is the question of differentiability. In this general situation the differentials of Prechet appear to be the most appropriate concepts. In the present chapter we develop Prechet differentials. Let X, Y be normed spaces and let U be an open set in X. As above, we denote by the same symbol || • || the norms of both X and Y. Definition 2.1 A mapping / : U H-> Y is said to be differentiable at x £ U if there exists a linear map A (= Ax) £ L(X, Y) such that / and the continuous affine linear map h £ X —> f(x) + Ah £ Y are tangent at x, i.e.,
a, !/<*+*>-/(->-*»!_„.
fc-o
M
\\h\\
If there is such a linear map Ax satisfying (2.1), then it is unique. We call A(= Ax) the Prechet derivative of / at x, and A will be denoted by T>f(x) or F'(x). The element f'(x)h £ Y is called the Prechet differential or the differential of / at x in the direction of h £ X. We note that differentiability depends only on the topologies of X and Y, and not on the particular norms used to define these topologies. If / : U i-> Y is differentiable at each point of U, then / is said to be differentiable on T>. In this case the mapping Vf : z € £/ H-> f'(x) £ L{X, Y) 51
(2.2)
52
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
is called the derivative (or differential) of / on U. If V is continuous on U, the mapping / is said to be of class C 1 on U. Note that the differential T>f is always vector-valued and not scalar-valued even when we may be interested in the differentiation of scalar-valued functions / : U i-» K (i.e., when Y = K). In fact, in this case we have Vf : U t-> L(X,K) = X*. Let X, Y be normed spaces and V a non-empty open subset of X. Then the notion of differentiability is a local property in the following sense. Proposition 2.1 If / : M H-> Y is differentiable on M, then f is differentiable on any open subset U of M, and W\u) Proposition 2.2
= (Vf)\u.
(2.3)
If M = [j Ui where each Ui is open in M, then f is i€l
differentiable on M if and only if it is differentiable on each Ui. Theorem 2.1 (Lipschitzian property) Let X and Y be normed spaces and U a nonempty open subset of X. If f : U — i > Y is differentiable at x € U, then there exist C > 0 and 6 > 0 such that Il/Cx) - / ( y ) | |
(2.4)
for y £ U, \\x — 2/|| < S. In particular, it follows that f is continuous at x. Theorem 2.2 (Chain rule) Let X,Y,Z be normed spaces, U an open i > V and g : V >—» Z. subset of X, and V and open subset of Y. Let f : U — If f is differentiable at x e U and g is differentiable at f(x) G V, then g o f : U f-> Z is differentiable at x and {9of)'{x)=g'{f{x))of'{x).
(2.5)
Corollary 2.1 Suppose F :T> >->Y is differentiable at a point XQ € V and assume that there exists a mapping G : Y — i > X defined on a neighborhood U of the point j/o = F(xo), and satisfying the relations GoF = Iv,
FoG = Iu
(2.6)
(we write in this case G = F~1). IfF~x is differentiable at yo = F(xo) G U, then the linear operator F'(xo) has a continuous inverse and [F'(xo)}-1 = (F-'Yiyo).
(2.7)
Differentiable and Holomorphic Mappings in Banach Spaces
2.1.1
53
Examples
Example 2.1 (Constant mappings) Let X and Y be normed spaces. Then a constant mapping is a mapping of the form f : x e X >-> b £Y for a fixed point b in Y. If / : X i-> Y is a constant mapping, then Vf = 0, i.e., f'(x) = 0 for all x e X. Example 2.2 (Linear mappings) If A € Lpf, Y), then PA is the constant mapping satisfying A'(x) = A for all x e X. Let / : £> t-> Y be differentiable a t i G O and let A& L(Y,Z). Then (Ao/)'(*) = A o / ' ( z ) .
(2.8)
Example 2.3 (Urysohn operator) Let K(t, s, x) be a function denned on a < t, s < b, \x\ < r, such that K(t, s, x) and K'x(t, s, x) are everywhere continuous. Let X = C[a, b\ be the space of all real- (or complex-)valued continuous functions on [a,b\ with the norm ||x|| = sup \x(t)\. Then one a
can define a mapping
/ : P H I , P = {XGI:
||:r|| < r}, by the formula
f(x(t))= [ K(t,s,x(s))ds. (2.9) Ja This mapping is usually called the nonlinear Urysohn integral operator. The operator / takes values in C[a, b] and is Frechet differentiable on the open ball ||x|| < r. Moreover, for any XQ € C[a, b] with ||:ro|| < r, and for every h 6 C[a, 6], we have
(f'(xo)h)(t)= f K'x(t,s,x0(s))h(s)ds.
(2.10)
Ja Note that the Urysohn operator (2.9) can be also considered on the space C[a, b] of all complex-valued continuous functions; then / is Frechet differentiable on C[a, b] if the kernel K(t, s, x) has continuous partial derivatives with respect to the variable x in the disk { x e C : | a ; | < r } o f the complex plane. Example 2.4 (Hammerstein operator) A particular case of the operator (2.9) is the Hammerstein operator denned by the equality
f(x(t))= [ K(t,s)f(s,x(s))ds:
(2.11)
Ja where the functions K(t, s), f(s, x) and fx(s, x) are continuous with respect to all variables for a < t, s < b, \x\ < r. Example 2.3 implies that the
54
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
Frechet derivative of / is given by (f'(x)h)(t)= I K{t,s)f'x{s,x{s))h(s)ds.
(2.12)
Ja
Example 2.5 (Nemytski operator) Suppose that the functions (p(s,x) and
•-> X = C[a, b] denned by (2.13)
*(x)(8):=
1~; (c) V ( T ^ ) = V ( * ) + V ( * ) - = lim Fn'+1 = tp. j—too ), i.e., ip2 = ip. In other words, if we let V = ) is a holomorphic retraction of T>, i.e., T = Fix(ir) if and only if A is power convergent in the norm topology of L(X) to a projection P : X —> Ker(/ — ^4). Moreover, we will see below that if the domain V is bounded, then the set T consists of only quasi-regular (or regular) points. The converse, of course, is not true. To see this, consider, for example, a rotation F(x) = eiex, 0 < 9 < 2TT, of the open unit disk A in C. This mapping has a unique and regular fixed point (the origin) in A, but F is not power convergent. Remark 5.7 Let o-(A) denote the spectrum of the linear operator A : X —> X, and let A be the open unit disk in C. As above, assume that T = Fix(F) ^ 0 for some F € Hol(P) and that aeV. Setting A = F'(a) and using the Cauchy integral formula, the chain rule and the boundedness ofD, we see that the powers of A are uniformly bounded, i.e., \\An\\ < M < oo. This implies that o-(A) C A. ) with Fix(F) = Fhc( . For the finite dimensional case this was established by J.-P Vigue [Vigue (1986); Vigue (1991a)], and in the general case by P. Mazet and J.-P Vigue [Mazet and Vigue (1991)]. They used nonlinear analogues of mean ergodic constructions. More precisely, they considered the Cesaro averages 1 "~1 . But a deficiency of Theorem 5.22 is that we do not know how to choose a convergent subsequence of {Cn} in order to approximatefixedpoints. Theorem 5.23 (which also covers the finite-dimensional case) would improve this situation if we could determine the minimal number p for which the mapping Cp defined by (5.119) is power convergent. We give below, inter alia, the simplest possible answer to this question, namely, p = 2. A partial generalization of Theorem 5.23 to unbounded domains was obtained by Do [Do (1992)] who has proved that if P is a convex hyperbolic domain in X, and F e Hol(P) is such that F(V) is contained in some compact convex subset of X with T = Fix(F) ^ 0, then this set T is a holomorphic retract of T>. (We recall, in passing, that a domain T> in a complex Banach space X is said to be hyperbolic if the Kobayashi pseudometric Kf) generates the relative topology of V in X (see Chapter 3.)) This result is only a partial generalization of Theorem 5.23 because of its compactness hypothesis. Indeed, according to a theorem of Krasnoselskii [Krasnoselskii and Zabreiko (1984)], F'(x) is compact for all x £ V and hence each point in Fix(F) is quasi-regular. However, we will show below that this compactness hypothesis is unnecessary. Moreover, we present a simple method for constructing a retraction onto Fix(F). This will also provide a proof of Theorem 5.23 [Reich and Shoikhet (1998c)]. Theorem 5.24 Let T> be a hyperbolic convex domain in X, and let F € Hol(P) be bounded on each subset which is strictly inside T>. (i) If T — Fix(F) contains at least one quasi-regular point a e V, then for 0, —> T> which is a retraction onto the fixed point set of Tr. But this set coincides with W, and we are done. (ii) Now let a e V be a regular null point of / . Once again, by Lemmata 8.5 and 8.6, this means that for each r > 0, the spectral radius of the operator (Tr)'(a) is less than 1. Thus there is an equivalent norm || • || i on X such that 11(7^(0)1^ < 1. It follows by continuity that in this norm there is a ball BR(O) centered at a with radius R{= R(r)) such that BR(O) CC V and ||('2^.)'(a;)|j1 < qr < 1 for each x £ Br(a). Fix r > 0 and take any t > r. Using the resolvent identity and the equality Tr(a) = a, t > 0, we have for all x € Br(a), 0. In the first case the mapping / is not a generator. In the third case / is a strongly semi-complete vector field. In the second case / is a group generator. The third case applies to the often used function ip(x) = 1 + x\ + x\ (see, for example, [Jordan and Smith (1987]). Thus the solution to the system f ± i - x a + xi(l + x? + a^) = 0 o be the semigroup of holomorphic self-mappings on Q, generated by ip. Then it is clear that the family {G(t)}t>o defined by G(t) = r 1 o S(t) o /
Theorem 3.2
The function p(-, •) defined by (3.7) is a metric on A.
Proof. It is clear tht p(z,w) > 0 for all (z,w) £ A x A, and since |m_z(u;)| = |m_ w (z)|, we have p(z,w) = p(w,z) and p{z,w) = 0 if and only if z — w. Now we need to show the triangle inequality p(z,w) < p(z,u) + p{u,w)
(3.16)
for each triple of points z, w, u in A. Indeed, let 7 : [0,1] H-> A be a curve joining the points z and w such that j'(t) is piecewise continuous. We will call such a curve an admissible curve. Consider the function 6(z, w) = inf{L7 : 7 is an admissible curve joining 2 and w}, (3.17)
84
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
where L 7 is the "length" of 7:
''-jfi^F*
<3'18)
We claim that 5(z,w) = p(z,w). Indeed, it is sufficient to show that 6(-, •) satisfies the conditions of Theorem 3.1. In fact, if F e Hol(A) and 7 is an admissible curve joining z and w, then F o 7 is an admissible curve joining F(z) and F(w). It follows by the Schwarz Lemma that
f'lF'jjmij'jt)]
LFO" = h
ri
\7'(t)\
i-iiW))i2 * - L T^HW dt = L^
(3J9)
Hence, 6(F(z),F(w))<6(z,w).
(3.20)
If now F e Aut(A), then applying (3.20) to F " 1 we get the equality 6(F(z), F(w)) = 6(z,w),
(3.21)
which proves (i). To prove (ii) and (iii), it is sufficient to evaluate 6(0, s) for 0 < s < 1. If 7 joins 0 and 5, then so does 0 = Rej. Hence r1 \j{t)\dt 1
r1 (3>{t)dt
Jo 1-I7WP-./0 i - [ / W /
Jo
2 =tanh- 1 (s),
i- — r
(3.22)
i.e., 6(0, s) > tanh - 1 (s). On the other hand, for ^(t) = ts, L1 = tanh - 1 (s). Consequently, we have J(0,s)=tanh~ 1 (s).
(3.23)
Now it follows, by properties (b) and (c) (see Remark 3.1), that the function 6 satisfies conditions (ii) and (iii) of Theorem 3.1. Thus, 5(z,w) = p(z,w).
(3.24)
Now it can be seen directly that (3.16) is a consequence of the definition of *(-,•)• • As a matter of fact, we have established somewhat more in our proof.
85
Differentiable and Holomorphic Mappings in Banach Spaces
Theorem 3.3 Each mapping F G Hol(A) is nonexpansive with respect to the metric p, i.e., (3.25)
p(F(z)), F(w)) < p(z,w). Moreover,
if equality
in (3.25) holds for some pair of points
z,w £ A , then
F G Aut(A). Actually, the conclusion of this theorem is also a direct consequence of the Schwarz-Pick inequality and (3.4). This metric, the existence of which we have just established, is called the Poincare hyperbolic metric on A and will be denoted by p. Theorem 3.4 The pair (A, p) is a complete unbounded metric space, and p defines on A the same topology as the original Euclidean topology on A. Proof. We list several topological properties of the Poincare metric which prove our assertion. (a) lim p(0,zn) = oo if and only if \zn\ —> l~. Indeed, by (3.9) we have n—>oo
p(0, zn) = p(0, \zn\) = t a n l T 1 \zn\, and the assertion follows. (b) Each p-ball B{a,R) = {z G A : p(a, z) < R) is the disk Ar(a) by the formula B(a,R) = {ze A:\z-sa\
< t-t&nhR},
(3.26) defined (3.27)
where .
l-(tanhfl)2 l-(tanhfl)2|a|2'
1-H2 l-(tanhfl)2|a|2 '
(328) K
'
In particular, if a = 0, then B(0,R) = {z G A : \z\ < tanhil} is a disk centered at zero. (c) B(a,R) = ma(B(0,r)), i.e., each p-ball is the image of a disk centered at zero under the corresponding Mobius transformation. Now, it is clear that if the sequence {zn} C A converges to z G A in the sense of the /9-metric, i.e., p(zn, z) —> 0, then \zn - z\ < \zn — sz\ + (1 - s)\z\ —> 0 because s —> 1~ when R —> 0 (see (3.28)). Conversely, if \zn — z\ —> 0 for zn, z G A, then there is r G (0,1) such that zn G A r = {z G A : \z\ < r). Hence, p(zn, z) = tanh" 1 \m-z(zn)\
< tanh" 1 ' ^ ~ *} -> 0. 1 — rz
(3.29)
86
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
Finally, note that each p-ball is bounded away from the boundary of A; this fact implies the completeness of the metric space (A,/j).
•
3.2
The Infinitesimal Poincare Metric and Geodesies
Let z be any point of A and let u e C be such that z + u S A. If p is the Poincare metric on A, then p(z,z + u)=
'7
i - \z\
(l + e(ti)),
(3.30)
where e(u) —» 0 as it —> 0. Formula (3.30) shows that the linear differential element dpz in the Poincare metric is denned by the formula dp'
(3-31)
= T^M2-
The form a{z,u)=
\u\ ' ' ,
zeA,
uGC,
(3.32)
is called the infinitesimal (or differential) Poincare hyperbolic metric on A. The following properties are an immediate consequence of the definition: (a) a(z, u)>0, z £ A, u e C . (b) a(z,tu) = \t\a{z,u), t e C . A consequence of the Schwarz-Pick Lemma is that each F G Hol(A) is a contraction for the infinitesimal Poincare metric. If F e Hol(A), then dpF(z) < dpz, or equivalently, a(F(z), dF{z)) < a(z, dz).
(3.33)
If F G Aut(A), then equality in (3.33) holds for all z e A. Moreover, if equality in (3.33) holds for at least one z G A, then F € Aut(A). This notion allows us, using Riemann integration, to define the "length" of any admissible curve in A.
Differentiable and Holomorphic Mappings in Banach Spaces
87
Let 7 : [0,1] •—> A be an admissible curve in A joining two points z and win A. Then the quantity
L 7 (= Ly(z, w)) = J dPy{t) = £ ~r^2dt
(3-34)
is called the hyperbolic length of 7. We already used this notion in the previous section and saw that the hyperbolic length is greater than, or equal to, the hyperbolic distance between its end points, i.e., (3.35)
p(z,w) < Ly(z,w).
A curve 7 joining the points z, w in A is called a geodesic segment in A if its length is equal to the hyperbolic distance between its end points z and w, i.e., (3.36)
L7(z,w) = p(z,w).
Proposition 3.1 For each pair of points z and w in A, there is a unique geodesic segment joining z and w and it is either a linear segment (if z and w lie on a diameter of A) or a segment of the circle in C which passes through z and w and is orthogonal to dA, the boundary of A. Proof. Indeed, for the points 0 and s, 0 < s < 1, the curve 71 (t) = ts is the unique curve joining 0 and s such that
'
M
=/ , * " = / ^ W
=
to"h-'w-
(3'37)
i.e., 71 (t) is the geodesic segment joining 0 and s. If z and w are arbitrary points in A, then the automorphism g = rtp om_ 2 , where m-z is a Mobius transformation and r v is the rotation with tp = — argm- z (w), takes z into 0 and w into s = |ra_ z (iy)|. If we now define 71 (t) as before, that is, 7i(t) = f|m_z(u>)|
(3.38)
i{t)=g-1{n{t)),
(3.39)
and
we obtain 7(0) = 5 -1 (7i(0)) = g'1^) = z and 7(1) = 5 -1 (7i(l)) = p~1(|m_z(u;)|) = mz(ei'p\m-z(w)\) — (mzom-z)(w) = w. Thus j(t) is an
88
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
admissible curve joining z and w. In addition,
r = f1 W\dt 7
=
Jo 1-I7WI 2
f1 Kr'Yhim • |7j(t)| Jo
i-lff-H-nC*))! 2
(3.40) But p{z,w) = p(0,m-z(w)) = L 7l and hence we obtain relation (3.36). The second part of our assertion is a direct consequence of Proposition 3.1.
•
3.3
The Poincare Metric on the Hilbert Ball and its Powers
Let B denote the open unit ball of a complex Hilbert space H. Recall that the Mobius transformation ma on B is denned by
ma(x)=
]
1 -(- [x, a)
(^l-\\a\\Qa + Pa)(x + a),
(3.41)
where i s l , P a is the orthogonal projection of H onto the subspace {Xa : X e C}, and Qa = I - Pa- The Poincare metric on the Hilbert ball B is the function PB • B x B —> 1R+ given by pM(x,y) = tanh- 1 \\m^(y)\\ = \ log \+_ | | ^ ^ | | .
(3.42)
Note that 1 — jlm—^d/)]!2 = a(x,y), where
(3.43) Applying this property of Mobius transformations we get the following more explicit formula for p, namely, pM(x,y) = tznh-1(l-<7(x,y))i.
(3.44)
It is clear that each ma is a pi-isometry and has a norm continuous injective extension from 1 onto itself. In general, if Bi and B2 are the open unit balls in complex Hilbert spaces Hi and if2, respectively, p\ and P2 are the Poincare metrics assigned to Bi and B2, respectively, and f € Hol(Bi,B2), then P2(f(x),f(y))
(3.45)
Differentiable and Holomorphic Mappings in Banach Spaces
89
In particular, each holomorphic self-mapping of B is nonexpansive with respect to p&. It is not difficult to see that for the Cartesian product B n of n Hilbert balls, n > 2, the function PB* : B n x Bn i-> R+ given by P*»(x,y) = max {pTiiixj, yj)},
(3.46)
l<j
where x = ( n , . . . , xn) and y = (j/i,. • •, j/ n ) belong to B n , defines a metric on B n which we call the Poincare metric on B n . Again, each holomorphic self-mapping of B n is nonexpansive with respect to this metric. We will see below that the Poincare metric pB" (as well as, of course, pjg itself) is a particular case of the so-called Caratheodory and Kobayashi pseudometrics assigned to any domain in a complex Banach space. 3.4
The Caratheodory and Kobayashi Pseudometrics
We are now in a position to define a pseudometric on an arbitrary domain T> which generalizes the Poincare metric on the unit disk, the Hilbert ball and on the powers of the Hilbert ball. Prom now on we let D b e a domain in an arbitrary complex Banach space X. 3.4.1
The Caratheodory
pseudometric
For x £ V we set BT(x) := {y £ X : \\y - x\\ < r} C T> and for y £ Br(x), y ^ x, we let 0 : A —> V be denned by
(3.47)
\\y-x\\ where A is the open unit disk in the complex plane. If g £ Hol(T>, A), then g o cf> £ Hol(A) and consequently
PA (W(0)), 5 ^ ( J ^ J [ ) ) ) < PA(O, t ^ ) ,
(3.48)
where p& is the Poincare metric on A. Hence, PA(g(x), g(y)) < tanh-1 (^y^j
•
(3.49)
If x and y are arbitrary points in V then, by connectedness, we can find xn} of points of V such that x 0 = x,xn— y, 5 > 0 and a chain {xo, x\,...,
90
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
and xi+i G Bs(xi) C T> for i = 0 , . . . , n - 1. By (3.49), we have
PA((*), g(y)) < f^tanh- 1 ^ H ^ - ^ - i l h
( 3. 5 0 )
for all g G Hol(£>, A). Consider the nonnegative function C (= Cv) on V xT> defined as follows: Cv(x,y)
= sup{PA(f(x),
f(y)) : f G Hol(P,A)}.
(3.51)
By (3.50), Cv{x,y) < oo for all x,y G V. It is easily checked that Cv is a pseudometric on T>. Definition 3.1 The function C (= Cv) is called the Caratheodory pseudometric on an arbitrary domain T>. Theorem 3.5 IfC\ andC2 are the Caratheodory pseudometrics assigned to the domains T>\ and T>2, respectively, and f G Hol(Pi,Z>2), then (3.52)
C2(f(x)J(y))
In particular, each holomorphic self-mapping of a domain V is nonexpansive with respect to the Caratheodory pseudometric C-p • Proof. Let / G Hol(X>i,X>2) be given. If h G Hol(P 2 , A), then h o f G Hol(X>i, A) and so for any x, y G £>i, we have (3.53)
pA{h(f(x)), h(f(y))) < d{x,y) and C2(f(x), f(y)) = sup{pA(h(f(x)),
h(f(y)) : h G Hol(2>2, A)}
(3.5g
IfV = A, then CA = PA is the Poincare metric on A.
Proof.
Using the mapping f(z) = z in A, we see that C&(z,w) > for z,w G A. Since holomorphic mappings from the disk into itself are nonexpansive with respect to the Poincare metric, we have CA(Z, W) < P&(z, w), and so equality holds. • PA(Z,W)
We will see below that also for the Hilbert ball B the Caratheodory pseudometric CB assigned to B coincides with the Poincare metric ps on B. Even in the one-dimensional case, however, it may happen that C-p is not a true metric. If, for example, T> = C, then Liouville's theorem implies that
Differentiable and Holomorphic Mappings in Banach Spaces
91
every / in Hol(C, A) is a constant mapping and hence pA(f(z), f{w)) = 0 for all z, w G C, and consequently Cp = 0. Nevertheless, if V is a bounded domain in an arbitrary complex Banach space X, then C-p is, in fact, a metric on T>. Proposition 3.3 Let V be a bounded domain in a complex Banach space X. Then Cv is a metric on V which induces the original topology on V defined by the norm topology of X. Furthermore, the metric space (D,C-p) is complete. We will prove this assertion later in a more general setting. 3.4.2
The Kobayashi
pseudometric
As the Caratheodory pseudometric on a domain V is defined by means of mappings from T> into A it is natural to look at mappings from A into T> to see if they also define a pseudometric on T> which generalizes the Poincare metric for the case of the unit disk. We will present two equivalent definitions of the Kobayashi pseudometric Kx> on V. The first definition is, in fact, that of the Lempert function S ([Lempert (1982)]; see also [Dineen (1989)]). In this case we can define a pseudometric on V after a minor adjustment to obtain the triangle inequality. For points x and y in a domain V we let 5v(x,y) = ini{PA{z,w)
: f G Hol(A,D), f(z) = x, f(w) = y). (3.55)
This definition presupposes the existence of a mapping / G Hol(A, V) whose range contains both x and y. This can be shown by applying the vectorvalued Stone—Weierstrass theorem. If V — A is the open unit disk in the complex plane, then using the homogeneity of the unit disk we see that 6&(z, w) = 6A(W, Z) for all z, w G T>. In addition, for z,w € A, the identity mapping shows that S&(z,w) < PA(Z> W), where p& is the Poincare metric on A. If u, v G A and / G Hol(A) is such that f(u) = z and f(v) = w, then we have that PA(Z, W) < PA(U, V) and hence PA(Z, W) < 8A{Z, W). SO we get the following assertion. Lemma 3.1 IfT> = A is the open unit disk in the complex plane, then 5A = PA is the Poincare metric on A. Remark 3.2 Unfortunately, some examples show that 5t> does not always satisfy the triangle inequality. Therefore it cannot serve as a pseudometric generalizing the Poincare metric. Indeed, if V (= T>r) = {(21,22) G C2 :
92
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
\zi\ < 1, |z 2 | < 1, \z\Z2\
The Kobayashi pseudometric Kp on a domain V is deKv(x,y)
=
inflj25v(xi>xi+i):n<=N,{x
= x1,...,xn+1
=y}cT>\.
(3.56)
Suppose that x, y and z belong to V, and {x = x\, X2, •. •, xn+i = y} and {y = 2/i 12/2, • • •, J/m+i = z} are finite subsets of V. Then n
m
Kv(x,z) < ^5v{xi,Xi+i)
+ Y^st>(yj,yj+i)
i=l
(3.57)
j=l
and, upon taking the infimum of each sum on the right-hand side independently, we see that Kv(x, z) < K-D(X,
Z)
+ Kv(y, z).
(3.58)
It is clear that K-p < Sx>- It is also clear that if 6x> does satisfy the triangle inequality then, in fact, Sv — KT>- AS a matter of fact, we have a more general assertion. Lemma 3.2 If p is any pseudometric on T> such that p < 5x>, then p < Kx>. In particular, C-r>(x,y) < Kt>{x,y). Proof.
If {x = xi, X2,. • •, xn+i n
= y} is a subset of V, then n
p{x,y) < Y^P(xu Xi+\) < ^foixi, i=l
x i + i),
(3.59)
i=l
and upon taking the infimum of the right-hand side, we find that p(x, y) < Kv{x,y). • We will see below that for a bounded convex domain the Caratheodory and Kobayashi metrics actually coincide. Meanwhile, we just mention this fact for the one-dimensional case. Proposition 3.4
IfV = A, then K& = PA is the Poincare metric on A.
Differentiable and Holomorphic Mappings in Banach Spaces
93
Suppose now that T>\ and X>2 are arbitrary domains and h £ Hol(A, T>{) with h(z) = x and h(w) = y for some z and u; in A. Let / belong to H o l ( P i , P 2 ) . T h e n fohe
H(A,V2)
a n d f(h(z))
= f(x),
f(h(w))
= f(y).
Hence <5p2(/(x), f(y)) < p&{z,w); taking the infimum of the right-hand side of this inequality over the appropriate set of holomorphic mappings, we see that W/0»0>/(y))<*Pi(s.v)-
(3-60)
Thus we arrive at the following assertion. Theorem 3.6 / / K\ and K-z are the Kobayashi pseudometrics assigned to the domains V\ and T>i, respectively, and f e Hol(X>i,X>2), then K2(f(x),f(y))
(3.61)
In particular, each holomorphic self-mapping of a domain V is nonexpansive with respect to the Kobayashi pseudometric K-p. In order to present the second definition of the Kobayashi pseudometric, we need the Kobayashi infinitesimal pseudometric which is a particular case of the so-called Finsler infinitesimal pseudometric considered in the next section. 3.5
Infinitesimal Finsler Pseudometrics
In this section we describe an infinitesimal approach to the definition of a pseudometric on a domain. It is based on the idea mentioned in previous sections regarding the measurement of the lengths of smooth curves. Definition 3.3 An infinitesimal Finsler pseudometric on a domain 23 in a Banach space X is a nonnegative function a : T> x X —> 1R such that (a) a(x,tv) = \t\a(x,v) for all (x,v) eV x X and t € C; (b) a is upper semicontinuous. If, in addition, a(x, v) > 0 for v ^ 0, we call a an infinitesimal Finsler metric. For two points x and y in V, let 7 : [0,1] —> V denote a parametrized curve, with a piecewise continuous derivative, joining x and y. We call 7 an admissible curve joining x and y. If a is an infinitesimal Finsler pseudometric and 7 : [0,1] —> T> is an admissible curve, then the function
94
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
t € [0,1] —> a(7(i),7'(£)) is a bounded measurable function and the length La(i) of 7 is defined by L a ( 7 ) - / a( 7 (t), i(t))dt. Jo
(3.62)
a(a;, iu) = ta(x, v) for ( i , u ) e P x X and t € R +
(3.63)
a(x, v) = a(x, -v) for (x,v) eVx X
(3.64)
The properties
and
are clearly implied by condition (a) and are equivalent, respectively, to the fact that curve length is independent of the parametrization and of the orientation of the curve. The integrated form of a is the function da :T> xT> —> R + defined by da(x,y) = inf{L Q (7) : 7 is an admissible curve joining x and y}. (3.65) It is easily verified that da is a pseudometric on V. If oti and OL-i are infinitesimal Finsler pseudometrics on T>\ and 2?2, respectively, and / : T>\ —> 2?2 *s a C 1 function, then / is called nonexpansive (respectively, an isometry) with respect to a.\ and c*2 if a2(f(x),f'(x)(v))
(3.66)
a2(f(x)J'(x)(v))=a1(x,v),
(3.67)
respectively,
for all (x,v)
eVxX.
Lemma 3.3 Let «i and a2 be infinitesimal Finsler pseudometrics on V\ andT>2, respectively, and let di andd2 be the integrated forms ofct\ anda2, respectively. If f : V\ —> V2 is a smooth function such that f : (T>i,ai) —> — > {T>2,d2) is also nonexpansive, (T>2,Oi2) is nonexpansive, then f : (T>i,di) 1 i.e., d2{f{x),f{y))
(3.68)
Differentiable and Holomorphic Mappings in Banach Spaces
95
Proof. Let 7 be an admissible curve in T>\ joining x and y. Then / o 7 is an admissible curve in X>2 joining f(x) and f(y)- Hence
< W/°7) = I «j((/o7)W.(/»7)'(tP ./o
=
f1a2(f(l(t)),f'(j(t))h'(t)))dt Jo
< I ai( 7 (t), -/(*))* = £O1(7) Jo
(3-69)
and "2(/(a;)) /(j/)) < inf{^ Ql (7) : 7 is an admissible curve joining x and j/} (3.70)
= d1(x,y).
D 3.5.1
Examples
Example 3.1 (The infinitesimal Poincare metric) Let a(z, v) = t_yLa for (z,v) € A x C. The integrated form of a is the Poincare metric p on A since inf{LQ(7) : 7 is an admissible curve joining z and u>} = tanh^ 1 \m-z(w)\, where "»—•
(3-71)
This function a is called the infinitesimal Poincare metric on A. Example 3.2 (The infinitesimal Caratheodory pseudometric) The infinitesimal Caratheodory pseudometric on a domain T> at the point (x, v) in V x X is defined by CD(X,V) = sup{|/'(z)(t,)| : / G Hol(2>, A)}.
(3.72)
If V is a domain in a complex Banach space X and / € Hol(X>, A), then f'(x) G X* and hence cv{x,\v)
= \X\cv{x,v) for all A G C and (z,u) e 23 x X.
(3.73)
The Cauchy estimates show that c-p is finite and locally bounded. If V = A is the open unit disk in the complex plane C, then it is clear that CA is
96
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
the infinitesimal Poincare metric; hence the integrated form of CA is the Poincare metric on A. If T>\ and X>2 are domains in complex Banach spaces X\ and Xi, respectively, and / € Hol(£>i,r>2), then cv2 (f{x), f'(x)(v))
< oDl(x,v)
(3.74)
for all (x,v) £ V\ x X\. Thus holomorphic mappings are nonexpansive with respect to the integrated forms d\ and cfo of cvx and cp 2 , respectively. In finite dimensions the definition of the pseudometric dc(x, y) = inf {Z/Cl)(7) : 7 is an admissible curve joining x and y} (3.75) is due to Reiffen [Reiffen (1965)] and the pseudometric is sometimes called the Caratheodory-Reiffen-Finsler (or simply CRF) pseudometric on a domain T>. Observe also that the infinitesimal Caratheodory pseudometric cv is a continuous infinitesimal Finsler pseudometric on the domain V. Finally, it can be shown (see, for example, the next section) that for any domain P in a complex Banach space, Cv < dc,
(3.76)
where C© is the Caratheodory pseudometric on V. We remark in passing that a strict inequality holds for some domains. Example 3.3 (The infinitesimal Kobayashi pseudometric) The infinitesimal Kobayashi pseudometric on a domain T> at the point (x, v) in T> x X is defined by kv(x,v)
= inf{77 > 0 : 3 / <E Hol(A,D), /(0)=x, f'(0)(r))=v}.
(3.77)
In contrast with the Caratheodory pseudometric, the integrated form dk(%, y) — inf{£fcD(7) : 7 is an admissible curve joining x and y} (3.78) of the infinitesimal Kobayashi pseudometric coincides with the Kobayashi pseudometric K-p defined above for any domain V, i.e., Kv(x,y) = dk(x,y), x,y£V.
(3.79)
Differentiable and Holomorphic Mappings in Banach Spaces
97
Therefore the last formula can serve as the second definition of the Kobayashi pseudometric.
3.6
Schwarz-Pick Systems of Pseudometrics
We are now in a position to define the so-called Schwarz-Pick systems of pseudometrics introduced by L. A. Harris which include the Caratheodory and Kobayashi pseudometrics as the smallest and the largest pseudometrics, respectively. Definition 3.4 A system which assigns a pseudometric to each domain in each normed linear space is called a Schwarz-Pick system if the following conditions hold: (i) the pseudometric assigned to A is the Poincare metric; (ii) (the Schwarz-Pick inequality) if p\ and p2 are the pseudometrics assigned to T>\ and Z>2> respectively, and / £ Hol(T>i,X>2)) then P2(f(x),f(y))
(3.80)
for &llx,y e D i . Note that condition (ii) implies that all biholomorphic mappings are isometries. A further obvious but useful consequence of (ii) is the following fact: if T>\ C 2?2 and p\ and pi are the pseudometrics assigned to V\ and T>z, respectively, then p2(x,y) < p\{x,y) for all x,y £ T>\. Proposition 3.5 The Caratheodory pseudometrics form the smallest Schwarz-Pick system. Proof. Let pz> be assigned to the domain T> by any Schwarz-Pick system. Let x, y € V be arbitrary. If / G Hol(£>, A), then PA{f{x), f{y))
(3.81)
Hence Cv{x,y) = sup{pA(f(x),f(y))
: f e Hol(D, A)} (3.82)
•
98
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
We have already seen that the Kobayashi pseudometrics {K-D} form a Schwarz-Pick system. Another immediate consequence of Lemma 3.2 is the following one. Proposition 3.6 The Kobayashi pseudometrics form Schwarz-Pick system.
the largest
Proof. Suppose that p-r> is assigned to the domain V by any SchwarzPick system of pseudometrics and suppose that x,y € T>. If / € Hol(A,D) and z, w S A are points such that x = f(z) and y = f{w), then (3.83)
Pv(x,y) < PA(Z,W).
Taking the infimum of the right-hand side over all such / we get pz>(x, y) < 8t>{x,y). By Lemma 3.2, this proves our assertion. • In proving the finiteness of the Caratheodory pseudometric we used the mapping
A-. + rJ^A.
(3.84)
\\y~x\\ The same mapping shows that
Kv(x,y) < Sv(x,y) < pA U, MzA\ = tanh"1 (^^)
(3.85)
for any domain T> with Br(x) C T>. We summarize the previous results by listing the relations we have established between the various pseudometricss on a domain T>: (3.86)
Cv
Since the Kobayashi pseudometrics form the largest Schwarz-Pick system, (3.85) implies the following proposition. Proposition 3.7 then
If {Pv} is a Schwarz-Pick system of pseudometrics,
pv(x,y)
x,y€V.
(3.87)
Differentiable and Holomorphic Mappings in Banach Spaces
99
The question that arises at this point is: does p induce the initial topology of T>1 The following theorem shows that for any bounded domain T>, a pseudometric px> assigned to I? by a Schwarz-Pick system is locally equivalent to the norm || • || of X. We denote by r{x) = dist|,.,| {x, dV) = inf {\\x - z\\ : z £ dV}
(3.88)
the distance in X between the point x and the boundary &D of the domain V. We also set R{x) = s\xp{\\x - z\\, z £ V). Theorem 3.7 then
(3.89)
If V is a bounded domain in a complex Banach space X,
tanh-1 ( ^ j ^ p ) < Pv(x, y)
(3.90)
for all x,y £ T>, and pv(x,y)
rCta.nh-1^^
(3.91)
whenever \\x — y\\ < dist||.||(a;,92)). Proof. The second inequality is a direct consequence of the previous proposition. Regarding the first inequality, it is enough to prove it for the Caratheodory pseudometrics since they form the smallest Schwarz-Pick system. So let Cv(x,y) = snp{pA(f(x), f(y)) : / £ Hol(P,A)}
(3.92)
be the Caratheodory pseudometric on T>. For x and y in T>, let I £ X* be a bounded linear functional on X of norm 1 such that (x — y,£) = \\x — y\\. Consider the holomorphic mapping / : V —» A defined by
(3.93) Clearly, f(x) = 0. Now it follows from the definition of the Caratheodory pseudometric Cv(x,y)
that
Cv(x,y) > PA(f(x),f(y)) = pA(0,/(y)) = ^nh'1 (\f(y)\)
(3.94)
100
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
Definition 3.5 A domain V in a complex Banach space X is said to be /3-hyperbolic (or just hyperbolic) if there exists a metric p assigned to V by a Schwarz-Pick system of pseudometrics which is equivalent to the norm of X. Corollary 3.1
Each bounded domain is hyperbolic.
Definition 3.6 A domain P in a complex Banach space X is said to be homogeneous if for each pair x and y in T> there exists an automorphism F of V such that F{x) = y. A simple consequence of the above theorem is the following assertion which we will use in the sequel. Corollary 3.2 Let X be a complex Banach space such that its open unit ball B is a homogeneous domain. Then all Schwarz-Pick systems of pseudometrics (actually metrics) on B coincide. Moreover, if for x € B we denote by Fx an automorphism of B such that Fx{x) = 0, then
pB(x,y) = tuab.-1(\\Fx(y)\\).
(3.95)
Proof. Let p (= ps) be any metric assigned to B by a Schwarz-Pick system. Then for each z £ B we have, by the above theorem, that p(0,«)=tanh- 1 (||z||).
(3.96)
Consequently, if Fx is an automorphism of B such that Fx(x) = 0, then p(x,y) = p{0,Fx(y)) = tanh^dlF^y)!).
(3.97)
• In particular, it is well known that if X is a J*-algebra, i.e., X is a closed subspace of L(H), the space of bounded linear operators on a complex Hilbert space H, such that for each x € X, xx* x £ X,
(3.98)
and V is the open unit ball of X, then, given x £ V, one can define the generalized Mobius transformation Mx by Mx(y) = (I-xx*)-l{x-y)(I-
x*y)-\l - x*x)i,
(3.99)
101
Differentiable and Holomorphic Mappings in Banach Spaces
where / is the identity operator in L(H). Thus each metric of a SchwarzPick system on D can be represented by p(x, 2 /)=tanh- 1 (||M : c ( 2 /)||).
(3.100)
Remark 3.3 Note that if X is a complex Banach space such that its open unit ball B is a homogeneous domain, then B is a bounded symmetric domain in X. The converse is also true: each bounded symmetric domain can be realized as the open unit ball of a complex Banach space. Obviously, this ball is a homogeneous domain. Since any Schwarz-Pick system is preserved under a biholomorphic mapping it follows that all Schwarz-Pick systems of pseudometrics (actually metrics) on bounded symmetric domains coincide. Actually, it was shown by S. Dineen, R. M. Timoney and J.P. Vigue [Dineen et al. (1985)] (see also [Lempert (1982)] for the finite dimensional case) that this fact holds in a more general setting. Theorem 3.8 If D is a convex domain in a complex Banach space X, then Cx> = Kx>, i.e., all Schwarz-Pick systems of pseudometrics on T> coincide. If, in addition, T> is bounded, then, as we already know, this pseudometric is, in fact, a metric on T>. This unique metric is sometimes called the hyperbolic metric on T>. Finally, we observe that a convex domain T> in Cn (which is not necessarily bounded) is p-hyperbolic if and only if it does not contain a complex affine line (see, for example, [Dineen (1989)]).
3.7
Bounded Convex Domains and Metric Domains in Banach Spaces
In this section we consider some special properties of bounded convex domains with respect to the hyperbolic metric. We show that this metric is compatible with the convex structure of a domain. Lemma 3.4 Let V be a bounded convex domain in a complex Banach space X endowed with the hyperbolic metric p. Then each p~ball in the metric space (T>,p) is bounded away from the boundary ofT>. Proof. We may suppose without loss of generality that 0 S V. It suffices to show that Br := {x € V : p(0, x) < r} is bounded away from the
102
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
boundary. Let £ e X* be such that supRe^(x) < 1.
(3.101)
Let /(x) = 2 ^ j for x £ V. Then /(0) = 0 and |/(x)| < 1 for all x G V. Hence / e Hol(D, A). Now if x € Br, then tanh-^l/Cx)!) = /9A(/(0), /(*)) < p(0,x) < r.
(3.102)
Hence ||/||B r < tanh(r) < 1 and there exists a, 0 < a < 1, such that Re^(x) < a for all x £ Br. Consequently, we get that Re^(x) < 1 for all x £ Br + -piy (1 — a)Xo, where XQ is the open unit ball of X. Since x = 0 is an interior point in V, it follows that there is M > 0 such that \\£\\ < M for all t satisfying (3.101). Hence by the Hanh-Banach separation theorem, we have that Br + -Ar, (1 - a)X0 c T> for some £ £ X*. Hence BT is bounded • away from the boundary of V. Thus we see that any p-Cauchy sequence is bounded away from the boundary. This yields the following important assertion. Theorem 3.9 A bounded convex domain V in a complex Banach space X endowed with the hyperbolic metric p is complete and p-hyperbolic. Morebounded away from the boundary of V if and only over, a subset KofVis if it is p-bounded. If a subset K of T> is bounded away from the boundary of T> we will sometimes also say that K lies strictly inside T>. We need the following simple lemma, which will be useful in the sequel. Lemma 3.5 Let T> be a bounded convex domain in a complex Banach space X, and let z be a point in V. Then for a fixed s € [0,1), the set K = {sx + (1 - s)z : x € V)
(3.103)
is strictly inside T>. Moreover, if r = d\st{z, dD), then 5 — dist(K, &D) > (1 - s)r. Proof. Let u S X be such that ||u|| < r. Then the element z + u belongs to T>. Since T> is convex, it follows that for all x € T> the elements sx + (1 - s){z + u) = sx + (1 - s)z + (1 — s)u also belong to V. This • completes the proof. Proposition 3.8 Let V be a bounded convex domain in a Banach space X, and let p be the hyperbolic metric on T>.
Differentiable and Holomorphic Mappings in Banach Spaces (i) Ifx,yeT>
and s,t€ [0,1], then p(sx + (1 - s)y, tx + (l-
(ii) Ifx,y,ztV
103
t)y) < p(x, y).
(3.104)
and s € [0,1], then p(sx + (1 - s)z, sy + (l- s)z) * diamP + ^ d i a t ^ a P ) **>V)'
(Hi) Ifx,y,w,z
(3J°5)
€ V and s € [0,1], then
p(sx + (1 - s)y, sw + (1 - s)z) < max[/o(x, w), p(y, z)]. (3.106) Proof. Assertions (i) and (iii) follow directly from the definition of the Lempert function and the Kobayashi pseudometric on a domain. To prove assertion (ii), let usfixz 6 P and s G [0,1). Consider the set K — {sx+(l — s)z : x E T>} and let r = dist(z, dV). Since V is bounded, one can find e > 0 such that e\\x-y\\ < ( l - s ) r < S = dist(K,&D) (say, e = £ * ^ ) for all x, y in V. Hence all elements of the form sx + (1 - s)z + se(x — y) also belong to V. Fix any y 6 V and consider the affine (hence holomorphic) mapping / defined as follows: fix) = sx + (1 - s)z + se(x - y).
(3.107)
Let p be the (unique) hyperbolic metric on V. Then p can be represented as the integrated form of the infinitesimal Caratheodory pseudometric on the domain V defined at the point (x, v) in T> x X by cv(x,v)=sup{\f'(x)iv)\:f£Eol(V,A)},
(3.108)
i.e.,
r
r
1 p(x,y) = m£\ L cc ( 7 ) = / d,( 7 (t) f -/(t))dt: Jo I 7 is an admissible curve joining x and y >.
(3.109)
Since the mapping / is a holomorphic self-mapping of V, we have that cv(f(x), f'(x)(v)) < c-oix, v) for all x e V and veX. At the same time, noticing that f'(x) = (1 -f e)s/ (/ is the identity mapping on X) does not depend o n i e l and f(y) = sy + (1 - s)z, we
104
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
see that cv{sy + (1 - s)z, (1 + e)slv) < cv(y, v), or cv(sy+(l-s)z,
(3.110)
slv) < Y^cv(y,v).
Since y is arbitrary, we can replace it with any element in V. Let now 7 : [0,1] —> T> be any admissible curve joining x and y. Then 71 = sj+ (1 — s)z is an admissible curve joining sx + (1 — s)z and sy + (1 — s)z. We have co(7i(*).7i(*)) = cu(s 7 (i) + (1 - s)z,
sf'(t))
(3.111)
Consequently, p(sx + (1 - s)2, 51/ + (1 - s)2;) < j - ^ p(x, y). Now substituting e = - ( 1 - ~ 3 ^ ( p ' 9 P ) we get assertion (ii).
(3.112) •
Definition 3.7 We say that V is a metric domain in X if there exists a metric p on T> such that (i) for each x 6 V and for each 0 < 5 < dist(a;,3D), there are positive numbers L (= L(6)), r(= r(S)) and m(= m(S)) such that p{x,y) < £||^ — 2/|| whenever ||a; — y\\ < S,
(3.113)
and p(x,y) > m\\x — y\\ whenever p(x,y) < r;
(3.114)
(ii) each p-ball is strictly inside V. A bounded convex domain in a complex Banach space with a metric assigned to it by the Schwarz-Pick system is a metric domain. It is clear that if V is a metric domain, then the metric space (V, p) is complete. For a bounded convex domain in a real Banach space such a metric can be induced by the complexification of X and by using a Schwarz-Pick metric p on the direct product of T> by itself in the complex sense. Other constructions of such domains can be given by using Hilbert's projective metric or Thompson's metric on a cone associated with a convex bounded domain V in X (see [Nussbaum (1994)]).
105
Differentiable and Holomorphic Mappings in Banach Spaces
Definition 3.8 Let D be a metric convex domain in a Banach space X with a corresponding metric p. We say that the metric p is compatible with the convex structure of T> if the following conditions hold: (i) If x,y,w,zGT>
and s £ [0,1], then
p(sx + (1 - s)y, sw + (l-
s)z) < m&x[p(x, w), p(y, z)]; (3.115)
(ii) for each z £ V, there is a real function (p : [0,1) —> [0,1) such that 1—s limsup T-T- < oo, s^il-y(s)
(3.116)
and for each pair of points x,y S T> and s 6 [0,1], the following inequality holds: p(sx + (1 - s)z,
sy + (l-
s)z) <
(3.117)
The following assertion is a direct consequence of the above propositions. Proposition 3.9 For each bounded convex domain V in a Banach space X, there is a metric p onT> such that (T>,p) is a complete metric space, T> is a metric domain, and the metric p is compatible with the convex structure
ofV.
Proof. Choose p as the hyperbolic metric on V and use inequality (3.112) with e = dfamp'9P^ • Since e —> 0 + as s —» 1~, the last inequality proves our assertion. •
Chapter 4
Some Fixed Point Principles
4.1
The Banach Principle
Let V be a topological space and assume that F is a self-mapping of V, i.e., F(V) C V. Definition 4.1 if F(z) = z.
(4.1)
A point z € V is called a fixed point of the mapping F
Condition (4.1), which is referred to as the invariance of the set T> under the mapping F, enables us to define the iterations Fn of F by the recursive relations F1 = F, Fn+1 = Fn o F, where "o" denotes the composition operation of mappings. The subset of 23 consisting of all the fixed points of a mapping F :T> —> Z> will be denoted by ¥\X.T>(F) (or simply by Fix(F)). Fixed point theory is mostly concerned with the following basic issues: (a) (b) (c) (d)
Existence (i.e., is Fix(F) ^ 0 ?). Uniqueness (or whether the fixed points are isolated). Approximation (i.e., devising algorithms for locating fixed points). Properties (e.g., structure, connectedness, etc.) of Fix(F). The following properties of F and its iterates Fn are obvious:
(i) Fix(F) C Fix(F n ) for all n > 1; (ii) if F is continuous and Fn(x) —> y G V as n —» oo, then y € Fix(F); (iii) {Fn}n>i has a natural one-parameter semigroup structure: Fn+m = FnoFm. Recall that if (M,p) is a metric space and F : M —> M is a mapping, then we say that F is nonexpansive if p(F(x),F(y)) < p(x,y) for all 107
108
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
x,y € M. If, moreover, there exists k, 0 < k < 1, such that < kp(x,y) for all x,y € M , then we call F a sirict conp(F(x),F(y)) traction or a fc-contraction. It is clear that nonexpansive mappings are continuous. The following principle is among the more widely used fixed point theorems. Theorem 4.1 (Banach's Contraction Principle) Let (M,p) be a complete metric space and let F : M —> M be a strict contraction. Then F has a unique fixed point in M, and for each x0 € M, the sequence of iterates {Fn(xo)} converges to this fixed point. Proof. Select XQ G M and define the iterative sequence {xn} by xn+i = F(xn) (equivalently, xn = Fn(x0)), n = 0 , 1 , 2 , . . . . Observe that for any indices n, p £ N, P(xn,xn+P)
= p{Fn(x0),Fn+P(x0))
= p{Fn(x0),F"
oF*(xo))
+•••+
p{Fp-\x0), Fv{xo))}
+ k+--- + kp-1)P(x0,
F(x0))
(4.2) This shows that {xn} is a Cauchy sequence, and since M is complete, there exists x e M such that lim xn = x. To see that x is the unique fixed point n—too
of F, observe that
x = lim xn = lim xn+i = lim F(xn) = F(x) n—>oo
n—>oo
(4.3)
n—>oo
and, moreover, x = F(x) and y = F(y) imply .
p(x,y) = p(F(x), F(y)) < kp(x, y),
(4.4)
yielding p(x, y) = 0. Finally, letting p —> oo in (4.2), we obtain a rate of convergence: p(xn, x) = P(Fn(x0), x) < - ^ - p(xo,F(xo)). i-k
(4.5) n
109
Some Fixed Point Principles
The Banach fixed point theorem fails for nonexpansive mappings. A simple standard example is the shift of the real axis M. defined by F(x) = x + l. There is a natural class of mappings which falls properly between the class of strictly contractive mappings and nonexpansive mappings. A mapping F : M —> M is called contractive if p(F(x),F{y))
< p(x,y),
(4.6)
x,y € M, x^y.
Obviously, a mapping of this type can have at most one fixed point. The mapping F : R —> R defined by F(x) = 1 + l n ( l + ex) provides a simple example of a fixed point free contractive mapping. (In fact, \x — F(x)\ > 1 for all x € K.) However, in compact spaces such mappings always have fixed points. Theorem 4.2 Let (M, p) be a compact metric space and let F : M —* M be contractive. Then F has a unique fixed point in M, and for any xo € M, the sequence {F n (xo)} of iterates converges to this fixed point. Proof. The function
= p(Fn+i(x0),Fx)
< p(Fn(x0),x)
= an, (4.7)
{an} is a decreasing sequence of nonnegative real numbers and so has a limit, say a. Again by compactness, {Fn(x0)} has a convergent subsequence {Fn"(x0)}, say lim Fnk(x0) = z. Obviously, p{z,x) = a. If a > 0, then fe—*oo
we obtain the contradiction: a=limop(Fn"+1{xo),x)=p(F{z),x)
= p(F{z),F(x)) < p(z, x) = a.
Thus a = 0. Therefore any convergent subsequence of {Fn(x0)} converge to x, so by compactness, lim Fn(xo) = x.
(4.8) must n
n—>oo
The theorem is generally attributed to Edelstein [Edelstein (1962)] (see also [Edelstein (1964)]), who actually proved the following slightly more general result (cf. [Goebel and Reich (1984)]).
110
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
Theorem 4.3 Let M be a metric space and F : M —> M a contractive mapping. If there is a point z in M such that a subsequence of {Fn(z)} converges to y, then y is the unique fixed point of F, and lira Fn(z) = y. n—>oo
Finally, we give an example which shows that additional conditions for the existence of fixed points are necessary even if M is bounded. Example 4.1 Let X = c0 and let M be the closed unit ball of X. The mapping F defined by F(x) = ( a , x x , . . . , x n . . . ) ,
0 < a < l ,
(4.9)
is an affine isometry of M, but it is fixed point free.
4.2
The Theorems of Brouwer and Schauder
Brouwer's theorem is very useful in applications in Analysis and its discovery has had a tremendous influence in the development of several branches of mathematics, most notably algebraic topology. The simple formulation of Brouwer's theorem belies the fact that it seems to require a 'nonelementary' (nonmetric) proof. All known proofs require facts not commonly used in metric fixed point theory. On the other hand, attempts to find other versions of Brouwer's theorem have given rise to several interesting questions of metric type which will be discussed later.
Theorem 4.4 (Brouwer's Fixed Point Theorem) // F : V -> T> is a continuous self-mapping of a compact convex subset V of C n into itself, then F has a fixed point. One of the most important fixed point principles is the well-known principle due to J. Schauder, which is a generalization of the finite-dimensional fixed point principle of Brouwer. Theorem 4.5 (J. Schauder) Let F be a mapping which maps a closed convex subset V of a Banach space X into itself. If F(D) is contained in a compact subset of T>, then F has at least one fixed point in T>. All the proofs of this theorem we know about rely on topological and geometric considerations. As a rule, these proofs use Brouwer's theorem and finite-dimensional approximations (see, for instance, [Trenogin (1980)] and [Krasnoselskii and Zabreiko (1984)]).
Some Fixed Point Principles
111
Note also that the Schauder principle has no constructive features; it fails to indicate the number of fixed points, as well as methods for their approximation. In the following section we will consider another class of mappings, namely, those mappings which are either contractive or nonexpansive with respect to the hyperbolic metric.
4.3
Holomorphic Fixed Point Theorems
Let 23 be a domain in a complex Banach space X, and let Hol(X>), as above, denote the family of all holomorphic self-mappings of V. Among the fixed point principles for a mapping F £ Hol(P) in general Banach space there are two useful classical criteria which ensure the existence and uniqueness of a fixed point of F. Recall that a subset S is said to lie strictly inside T> if there exists e > 0 such that the open ball Be(x) centered at x of radius e is a subset of V whenever x G V. In other words, inf{||x-y|| :ye&D, x € S} >e.
(4.10)
The following theorem, originally due to Earle and Hamilton [Earle and Hamilton (1970)], may be viewed as a holomorphic version of the Banach contraction principle. We give a slightly more general formulation of it [Harris (2003)]. Theorem 4.6 (Earle—Hamilton) Let X> be a nonempty domain in a complex Banach space X and let F : V —> V be a bounded holomorphic mapping. If F{V) lies strictly inside V, then F has a unique fixed point in V. Moreover, the sequence of iterates {Fn(x)}^=1 converges to this point uniformly on each bounded subset of V. Proof. We use a pseudometric p, called the Caratheodory-ReiffenFinsler pseudometric (CRF-pseudometric), with respect to which F is a strict contraction. Let A be the open unit disk of the complex plane. Define a(x,v) = sup{\g'(x)v{ : g : V —» A holomorphic}
(4-11)
for x G T> and v e X, and set
L ( 7 ) = f a( 7 (t),7'(*))* Jo
(4.12)
112
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
for 7 in the set T of all curves in T> with piecewise continuous derivative. Clearly a specifies a seminorm at each point of T>. Define a pseudometric p : V x V -» R+ by p(x, y) = inf {L( 7 ) : 7 G T, 7 (0) = x, 7 (1) = y}
(4.13)
for x, y € T>.
Let x € 2? and » e l . By the chain rule, (g o F)'(x)v = 5'(F(a;))F'(a;)u
(4.14)
for any holomorphic function g : V —» A. Hence, a(F(x),F'(a;)i;)
(4.15) < L(j) (4.16)
holds for all x,y GT>. Now by hypothesis, there exists an e > 0 such that B£(F(x)) C P whenever x £ T>. We may assume that V is bounded by replacing V with the subset U{B£(F(x)) : i £ D } . Fix t with 0 < t < e/S, where 6 denotes the diameter of F(D). x € T>, define F(y) = F{y) + t[F(y) - F(x)}
(4.17) Given (4.18)
and note that F : V —> V is holomorphic. Given x 6 V and v € X, it follows from /"(*)<; = (1 + *)*"(*)«
(4.19)
and (4.15) with F replaced by F that a(F(x),F'(x)v)<-L-a(x,v). L -j~ t
(4.20)
Integrating this as before, we obtain p(F(x),F(y))<^-tp(x,y)
(4.21)
Some Fixed Point Principles
113
for all x, y £ V. Since V is bounded, (D, p) is a complete metric space. Now our assertion is seen to follow from by the Banach fixed point theorem. • The Earle-Hamilton theorem still applies in some cases where the holomorphic mapping does not necessarily map its domain strictly inside itself. In fact, Theorem 4.8 below is a generalization of the Earle-Hamilton theorem for convex domains. The following well known principle follows directly from a result of Krasnoselskii and Zabreiko [Krasnoselskii and Zabreiko (1984), Theorem 21.5] and a special property of holomorphic mappings [Krasnoselskii and Zabreiko (1984), Theorem 23.5]. Theorem 4.7 Let V be a convex domain in X, and let F € Hol(P) be continuous on T>. If F(T>) is contained in a compact subset ofT> and x £ F(x)
(4.22)
for all x S dV, then F has a unique fixed point in V. Theorem 4.8 (Khatskevich-Reich-Shoikhet [Khatskevich et al. (1995a)]) Let T> be a nonempty bounded convex domain in a complex Banach space and let F :T> —» T> be a holomorphic mapping having a uniformly continuous extension to T>. If there exists an e > 0 such that \\F(x) — x\\ > e whenever x 6 &D, then F has a unique fixed point in T>. Of course, the hypotheses of this theorem are satisfied on some homothety of T> under the assumptions of the Earle-Hamilton Theorem. The hypothesis ||F(a;) — a;|| > e for all x £ dV is also satisfied when F(T>) is contained in a compact subset of V and condition (4.22) holds. If F is not necessarily compact, this condition is also satisfied when T> contains the origin and S»P»<1. xedv INI
(4.23)
The proof we present here is due to L. A. Harris [Harris (2003)]. Proof. by
Given 0 < t < 1 and x£~D, define a holomorphic map / ( : V -> V
ft(y) = (l-t)x + tF(y)
(4.24)
114
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
and let 6 > 0 be such that Bs(x) C V. To show that ft(V) lies strictly inside T>, take e = (1 - t)6. Let y G V and let w G BE(ft(y)). Then
z=^m
(4.25)
is in 2? since z G -B*(x), so w = (l-t)x
+ tF{y) G P .
(4.26)
Hence B£(ft(y)) C £> for all j , e D . By the Earle-Hamilton theorem, ft has a unique fixed point gt(x) in £>. Since the hyperbolic CRF-metric is continuous, the proof of the Banach contraction principle shows that the iterates of ft at a chosen point yo € T> are holomorphic and locally uniformly Cauchy in x. Hence the limit mapping gt : T> —» T> is holomorphic. Now a point x S Z> is a fixed point of gt if and only if a; is a fixed point of F. Thus, by the Earle-Hamilton theorem, it suffices to show that gt(T>) lies strictly inside V for some t > 0. Since F has a uniformly continuous extension to T>, by hypothesis there exist e > 0 and 6 > 0 such that ||-F(a;) — x\\ > e whenever x €T> and dist(a:,dX>) = inf{||a; - y\\ : y € dV} < 6.
(4.27)
Since V is bounded, there is an M with ||:r|| < M for all x € V. If x G T>,
F(gt(x)) - gt(x) = (1 - t)[F( 5t (i)) - z],
(4.28)
so \\F(gt(x))-gt(x)\\< 2(1-t)M.
(4.29)
Choose t close enough to 1 so that 2(1 — t)M < e. If d(gt(x),dT>) < S for some x G T>, then e<\\F(gt(x))-gt(x)\\, a contradiction. Thus, B&(gt(x)) C P for all x G P , as required.
(4.30) •
Example 4.2 The hypotheses on the behavior of F on dV cannot be omitted in Theorem 4.8. This follows by considering a translate of the shift operator. Specifically, let X = CQ and define F(x)=(^,xux2...)
(4.31)
Some Fixed Point Principles
115
for x 6 X. Clearly, F is an afRne isometry on X and F maps the ball Br(0) into itself for each r > | . However, if F(a;) = x, then -=Xl=x2
= ...,
(4.32)
contradicting the requirement that x belongs to CQ. Thus F has no fixed point in X. 4.4
Fixed Points in the Hilbert Ball
Despite Example 4.2, it is still an open problem whether if B is the open unit ball of a separable reflexive complex Banach space and F : B —> B is a holomorphic mapping with a continuous extension to B, then F must have a fixed point in B. However, Hayden and Suffridge [Hayden and Suffridge (1976)] have proved that e%eF has a fixed point in B for almost every 9. Also, Goebel, Sekowski and Stachura [Goebel et al. (1980)] have solved the problem in the affirmative for the case where X is a Hilbert space and this has been extended by Kuczumow [Kuczumow (1984); Kuczumow (1985)] to the case where X is a finite product of Hilbert spaces (with the max norm). An example of Kakutani [Kakutani (1941)] shows that holomorphy is essential in the hypotheses since he exhibited a fixed point free homeomorphism of the closed unit ball of any infinite dimensional Hilbert space. A related problem is to weaken the hypotheses of the Earle-Hamilton theorem by showing that if F : B —> B is a holomorphic mapping such that the sequence of iterates {Fn(x)} lies strictly inside B for some x £ B, then F has a fixed point is B. This has been established when X is a Hilbert space (see also [Goebel and Reich (1984)]) and when X is a finite product of Hilbert spaces in [Kuczumow (1985)] (see also [Kryczka and Kuczumow (1997)]). These results have been extended to bounded convex domains in a more general class of reflexive Banach spaces by Budzyriska [Budzynska (2004)]. The example presented above gives a counterexample for the general case. Let B be the open unit ball of a Hilbert space H. Applying the asymptotic center method due to M. Edelstein [Edelstein (1972)] and the properties of the Poincare metric p on the Hilbert ball we get the following fixed point theorem. Definition 4.2
Let F : B —> B be a p-nonexpansive mapping. We
116
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
shall call a sequence {yn} =0. lim p(yn,Fyn)
C B an approximating sequence for F if
n—KX>
Theorem 4.9 (see [Goebel and Reich (1984)]) Let F : 1 -+ B be a pnonexpansive mapping. Then the following are equivalent: (a) F has a fixed point; (b) There exists a point x in B such that the sequence of iterates {Fnx} p-bounded; (c) The sequence of iterates {Fnx} is p-bounded for all x in M; (d) There exists a p-bounded approximating sequence for F.
is
The following proposition can also be found in [Goebel and Reich (1984)]. Proposition 4.1 B such that {Fnx} point.
Let F : B —» B be p-nonexpansive. If there exists x in converges weakly to a point in B, then F has a fixed
Theorem 4.10 ([Goebel (1982b)], see also [Goebel and Reich (1984)] and [Shafrir (1992b)]) / / a norm continuous mapping F : B —> B is pnonexpansive on M, then it has a fixed point in B.
4.5
Fixed Points in Finite Powers of the Hilbert Ball
It is not difficult to observe that the hyperbolic metric in the Cartesian product B n of n open unit balls B is given by Pi"(X,y) = max pB(zj,Vj)-
(4.33)
In this section N(Mn) will denote the class of all pBn-nonexpansive selfmapping on B™. The class of those mappings in N(En) which have a continuous (in norm) extension to B" will be denoted by CN(Mn). It will also be convenient to consider the slightly more general class of mappings N(Mn) which consists of all norm continuous mappings / : B n —+ B" such N(Mn) for all 0 < t < 1 [Kuczumow (1985)] and [Shafrir that tf\Bn£ (1992b)]. Note that when / e iV(B") it may happen that f(x) £ d(Mn) for x £ Bjf. But in this case, if f(x) = v with ll«jill = ••• = »«*. 11 = 1.
(4-34)
Some Fixed Point Principles
117
then f(y)h = vh. • • •. f(v)h = vik
(4-35)
for all y 6 1". The following simple generalization of Theorem 4.9 uses induction with respect to n and is based also on the asymptotic center method. Theorem 4.11 (c/. [Kuczumow et al. (2001a)]) Let F : B n -> B" be a holomorphic mapping or more generally, a p%,n-nonexpansive mapping. Then the following statements are equivalent: (i) F has afixedpoint; (ii) there exists i £ l " such that {Fk(x)} lies strictly inside B n (this means that {Fk(x)} is pun-bounded); (Hi) there exists a ball B(x,r) in (Bn,pBn) which is F-invariant; (iv) there exists a nonempty, pBn-bounded, pmn-closed and convex subset of B" which is F-invariant. Theorem 4.12 ([Kuczumow (1985)] and [Shafrir (1992b)]) If F € JV(I"), then Fix(F) ^ 0. Results on common fixed points of commuting families of mappings in JV(B~") can also be found in [Kuczumow (1984)] and [Shafrir (1992b)]. Remark 4.1 Note that Theorems 4.11 and 4.12 do not explain what happens when one tries to approximatefixedpoints by simply iterating F. converge to a fixed point of In other words, do the iterations {Fn(x)}^1 F (locally or globally) under the conditions of these theorems? This question arises, in particular, in the context of Theorem 4.12 when FixB"(F) = 0. For the one-dimensional case, when n = 1 and B = A is the open unit disk in the complex plane C, a complete answer to this question is given by the classical theorem of Denjoy and Wolff which asserts that if F € Hol(A) has no interior fixed point in A, then the sequence {Fn{z)}™=1 converges to a unique boundary point r G dA for all z G A. This result has given a powerful thrust to the study of the asymptotic behavior of iterates in different situations. In the next section we will describe some generalizations of the DenjoyWolff Theorem as well as point out its crucial connections with the classical Julia Lemma, the Schwarz-Wolff boundary theorem, and geometric function theory in complex spaces.
Chapter 5
The Denjoy-Wolff Fixed Point Theory
Theorem 5.1 (Denjoy-Wolff) Let A be the open unit disk in the complex plane C. If F € Hol(A) is not the identity and is not an automorphism of A with exactly one fixed point in A, then there is a unique point a in the closed unit disk A such that the iterates {Fn}'^L1 of F converge to a, uniformly on compact subsets of A. Let D be a domain in a complex Banach space X, and F S Hol(£>). The main goal of this chapter is to discuss the behavior of the iterates of F in the spirit of Theorem 5.1. Definition 5.1 We say that a mapping F S Hol(P) is power convergent to a mapping h € Hol(Z>, X) if the sequence of iterates {-F"}£Li converges to h uniformly on each ball strictly inside T>. If ft = a is a constant mapping, then the point a S V will be called a locally uniformly attractive fixed point of F. Recall that for V = A an automorphism F € Aut(A) which has exactly one fixed point is called elliptic. So, the Denjoy-Wolff theorem asserts that F € Hol(A) is power convergent if and only if F is not an elliptic automorphism. 5.1 5.1.1
The One-Dimensional Case Iterates of holomorphic self-mappings of A with an interior fixed point
In this section we prove the Denjoy-Wolff theorem for the case when F € Hol(A) has an interior fixed point. 119
120
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
The Schwarz-Pick inequality provides us with information regarding the invariant behavior of a holomorphic self-mapping of a domain V when F € Hol(2?) has an interior fixed point in T>. Namely, if p is a pseudometric on T> assigned to 2} by a Schwarz-Pick system, then F(Q(a)) C Q(F(a)),
(5.1)
where Cl(a) is any />-ball centered at a. Consequently, Fn(fl(a)) C n(Fn(a))
(5.2)
for all n = 0,1,2, In particular, for the one-dimensional case where T> = A is the open unit disk of the complex plane C, using a Mobius transformation, one can see that the p-ball Q of radius r (= fir(a)) can be written in the form n r (a) = {w G A : \m-a(w)\ < tanhr},
0 < r < oo.
(5.3)
Solving this inequality we get that, in fact, the p-ball ilr(a) is the disk Qr(a) = {w G A : \w - sa\ < dt},
(5.4)
where d = tanhr, S=
T^W
(5-5)
IHW
Thus we have that for each F € Hol(A), F(fi r (a))C[) r (F(«)),
r €(0,(50),
a e A,
(5.6)
i.e., F maps the disk Qr(a) centered at sa into the disk Qr(F(a)) centered at sF(a) with the same radius td. However, for a fixed d G (0,1) this radius tends to zero when a tends to the boundary. We consider this case later. The next question which naturally follows is: does the sequence {Fn(z)} converge to a fixed point a of F for each z G A? The answer is "no" if F is an elliptic automorphism. But for the other cases the answer is affirmative. Proposition 5.1 Suppose that F G Hol(A) is not an elliptic automorphism and has a fixed point a G A. Then the iterates Fn of the mapping F converge uniformly on compact subsets of A to a holomorphic mapping tp G Hol(A). Moreover, if F is not the identity, then
121
The Denjoy-Wolf Fixed Point Theory
Proof. If F is not an elliptic automorphism of A and it has a fixed point a in A, then the Schwarz-Pick Lemma (Proposition 1.1.3) implies that \F'(a)\ < 1.
(5.7)
Since F'(z) is continuous in A there is a disk A(a, e) C A centered at a with radius e > 0 such that \F'{z)\ < 1
(5.8)
for all z e A(a,e), the closure of A(a,e). In turn, (5.8) implies that F satisfies the Lipcshitz condition (5.9)
\F(z) - F(w)\ < q\z - w\,
where q = max{\F'(z)\, z 6 A(a, e)}. In addition, from (5.9) we have that F maps A(a, e) into itself. So, F is a g-contractive self-mapping of A(a, e), and it follows from the Banach Fixed Point Theorem that {Fn(z)} converges to a for all z 6 A(a,e). Using the Vitali property, we get our assertion. D 5.1.2
Iterates of holomorphic self-mappings interior fixed point
of A with no
By several technical transformations one can show that
nP(o) = jiu £ A : j ^ 0 - < K} ,
(5-10)
where d = tanhr and K = i = ^ " j , (see formulas (5.4) and (5.5)). Now suppose that \a\ = 1. For an arbitrary K > 0 define the domain V(a,K) by the same formula as (5.10):
V(a,K) = jz e A : ! i f ^ £
(5.11)
It is not difficult to see that T>(a,K) is also a disk in A centered at the point j — • a e A with radius j ^ < 1, i.e.,
%i q ={ 2 £ A:| Z - I La|<^ I }.
(5.12)
This disk is internally tangent to the boundary of A at the point a and it is called a horocycle.
h
The following assertion establishes a property of the horocycles T>(a, K), with respect to the family Hol(A), which is analogous to the property of the domains Clr(a). Theorem 5.2 (Julia's lemma) Let F e Hol(A) and let a e dA be a unimodular point. Suppose that there exists a sequence {an} c A converging to a such that the limits
a=limlH£M
(6.13)
6 = lim F{an)
(5.14)
an->a. 1 - \an\
and n—*oo
exist. Then the following inequality holds for each z G A ; |i-5F(*)| 3 1 - \F(z)\> Proof.
a
|i-a«|» i - |*|»
(5-15)
It is clear that |6| = 1. For z,w G A we define the function
(5.16) It is a simple exercise to show that the Schwarz-Pick Inequality is equivalent to (5.17)
a(z,w)
(5.18)
o-{z,an)
l
°-y; •
The following assertion obtained by J. Wolff [Wolff (1926c)] (see also [Wolff (1926a)] and [Wolff (1926b)]) is often called the Wolff-Schwarz Boundary Lemma.
The Denjoy-Wolf Fixed Point Theory
123
Theorem 5.3 (Wolff) Let F € Hol(A) have nofixedpoint in A. Then there is a unique unimodular point a e 9A such that for each K > 0, the horocycle
V(a, K) = {* e A : ^ - g £ < if} ,
(5.20)
internally tangent to dA at a, is F-invariant. Moreover, there is a number a 6 (0,1] such that Fn(V(a, K)) C V{a, anK) for alln =
(5.21)
0,1,2,....
Proof. It is sufficient to show that there is a sequence {dnY^Li C A converging to a, which satisfies the conditions of Julia's lemma, i.e., the limits a =lunilM
(5.22)
on->» 1 — \an\ and lim F(an) = a exist. Moreover, we will show that a < 1. Then n—*oo
our assertion will be seen to be a consequence of Julia's lemma. Take an arbitrary positive sequence {rn} increasing to 1 and consider the mappings Fn = rnF. It is clear that Fn is continuous on A rn = {z £ A : \z\ < rn} and that it maps this set into itself. Thus Fn has a fixed point an G A rn C A. Passing to a subsequence, we may assume that {an} converges to a point a € A. If a € A, then, by continuity, we have a = lim an = lim rnFn(a) = F(a), (5.23) n—»oo
n—>oo
contradicting the assumption that F has no fixed point in A. Consequently, \a\ = 1. In addition,
idSsf«_lziwsl 1 - \dn\
(5.24)
1 - \an\
for all n € N. Once again, perhaps by passing to a subsequence, we conclude that a in (5.22) exists and is less than or equal to 1. It is clear that F(an) — ^~an converges to a and by inequality (5.15) of Julia's Lemma, we obtain 1 - \F(z)\*
S
1 - |*|* '
(b-2b)
124
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
which implies (5.21).
•
Definition 5.2 Given a point a G dA and a real number k > 1, a nontangential approach region at a is the set T{a,k)
= {z£ A : \ z - a \ < jfc(l - | z | ) } .
(5.26)
Definition 5.3 Given a function h G Hol(A,C), we say that h has the angular limit L at a boundary point a G dA if h(z) —> L as z —> a, z G F(a,/c) for each fc > 1. We write in this case L = Z lim
ft(z).
z—>a
(5.27)
Remark 5.1 /£ is easy to see that h has the angular limit L at a point a G dA if and only if f(z) —> L as z —> a for each angle region S — Iz G A : | arg(l - az)| < /?, |z - a| < r, /3G(0,|),
rG(0,2cos^)},
(5.28)
which is a sector in A bounded between two straight lines in A that meet at a and are symmetric about the radius to a. This set is usually called a Stolz angle at a. Definition 5.4 A mapping F G Hol(A) is said to have an angular derivative at a G dA if the angular limit Z. lim
F(2)
~ b := ZF'(a)
z~*a z - a
(5.29)
exists for some b £ dA. The following essential contribution to the Denjoy-Wolff Theory, which is a complement of the Julia Lemma, was made by C. Caratheodory [Caratheodory (1929)] and [Caratheodory (1954)]. Theorem 5.4 (Julia-Caratheodory theorem) Let F G Hol(A) and a G dA. The following are equivalent: (i) S = S(F) = liminf ^Luf^ < °°> where the limit is taken unrestrictedly in A. (it) F has a finite angular derivative at a. (in) F has the angular limit b G dA and Z lim F'(z) = Sab. z—>a
125
The Denjoy-Wolf Fixed Point Theory
We will prove this theorem in the next section in a more general setting. An important consequence of this theorem and Theorems 5.2 and 5.3 is the following version of the so-called Julia-Wolff-Caratheodory theorem (c/. [Shapiro (1993)] and [Cowen and MacCluer (1995)]). Theorem 5.5
Let F £ Hol(A). The following are equivalent;
(i) F has no fixed point in A. (ii) There is a boundary point a £ dA such that 0 < lim F'(z) < 1. z—*a
Such a point a £ dA, which satisfies the conclusion of Theorem 5.3 of Wolff, will be called a sink point of F on dA (or Wolff's point). A mapping F £ Hol(A) with a sink point a 6 dA is said to be of hyperbolic type if the number a in (5.22) (Julia's number) is strictly less than 1. Otherwise (a — 1) the mapping F is said to be of parabolic type. Of course, if 0 < a < 1, then F is power convergent. The question is whether this point is also attractive when a = 1. The affirmative answer to this question is given in the next assertion, following Wolff and Denjoy [Wolff (1926a); Wolff (1926b); Wolff (1926c)], [Denjoy (1926)]. Theorem 5.6 If F £ Hol(A) has no fixed points in A, then there is a unique unimodular point a £ dA which is a sink point of F and the iterates {Fn} converge locally uniformly on A to a constant mapping ip(z) = a. Proof. If F £ Hol(A) is of parabolic type, and an automorphism of A, then the assertion is an exercise requiring just simple calculations. Thus we may assume that F £ Aut(A). Since {Fn} is a normal family, there is a subsequence {Fn->} which converges to a mapping
j—>oo
(5.30)
Since
126
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
fixed point in A. Hence h is the identity. At the same time, g o F = lim Fq* oF= j—>oo
lim FPi =1=
j —»oo
lim F o Fqi = F og.
j—»oo
(5.31)
Thus, i*1 is an automorphism. Having reached a contradiction, we conclude that (fi(z) = a is a constant after all. Now, if there is a subsequence {nj} C N such that {Fni(z)} converges for z e A to a constant 6 6 A different from a, then one can find K > 0 such that the closure of the horocycle T>(a, K) does not contain b. Since Fn^(V(a,K)) C V(a,K), it is impossible that for z € V(a,K), the sequence {Fn'{z)} converges to b. Hence {Fn(z)} converges to a for any z £ A. • Thus the classical Denjoy-Wolff theorem is, in fact, a summary of the following three assertions due to Denjoy, Wolff and Julia. Each one of them has been extended to different situations. • If F € Hol(A) is not an automorphism of A and has a fixed point c in A, then this point is unique in A, and the sequence {Fn}n°=l converges to c, uniformly on compact subsets of A. • (The Wolff-Schwarz Lemma) If F e Hol(A) has no fixed point in A, then there is a unique unimodular point a S dA such that every disk T>a in A, internally tangent to <9A at a, is F-invariant, i.e., F{Va) C Va.
(5.32)
• If F 6 Hol(A) has no fixed point in A, then there is a unique unimodular point b G dA such that the sequence {Fn}^=1 converges to b, uniformly on compact subsets of A. The limit point of the sequence {Fn(z)}'^L1, z £ A, will be called the Denjoy-Wolff point of F. The point a and the point b in the last two assertions are, of course, one and the same. In other words, the sink point of F is also the Denjoy-Wolff point of F. However, this is not always the case in higher dimensional situations. In fact, there are many situations in the higher dimensional case when a holomorphic fixed point free mapping has a sink point, but is not power convergent (see the next section). Returning to the one-dimensional case, we refer the reader to the paper by R. Burckel [Burckel (1981)] and to the books [Shapiro (1993)] and
127
The Denjoy-Wolf Fixed Point Theory
[Cowen and MacCluer (1995)] for a modern interpretation of the DenjoyWolff Theorem and its applications. Here we only mention the following observation concerning this case. Remark 5.2 By using the Schwarz Lemma, Proposition 5.1 can be rephrased in the following manner: • Let F € Hol(A) have afixedpoint a £ A. IfF is not the identity, then F is power convergent if and only if \F'(a)\ < 1. It turns out that by using the notion of the derivative and its spectral properties, one can also study power convergent mappings in higher dimensional spaces. See Sections 5.4 and 5.5 below. We are now in a position to formulate several generalizations of Theorems 5.2 and 5.3. 5.2
The Unit Hilbert Ball
Let if be a complex Hilbert space with the inner product (•, •), and let B be the open unit ball in H. The following generalization of the Wolff-Schwarz Boundary Lemma (Theorem 5.3) is due to K. Goebel [Goebel (1982b)]. For the finite dimensional case, H = C n , this result was independently obtained by B. MacCluer [MacCluer (1983)] and G. Chen [Chen (1984)]. Theorem 5.7 IfF € Hol(B) has nofixedpoint, then there exists a unique point a e dB such that for each 0 < R < oo, the set E(a, R) = \x 6 B : | 1 ~ ^ a '
I
! ~~ INI
< R)
(5.33)
J
is F-invariant. Geometrically, the set E(a, R) is an ellipsoid the closure of which intersects the unit sphere dM at the point a. It is a natural analogue of the horocycle P(a, R). As a matter of fact, Theorem 5.7 holds in a more general setting (cf. Theorem 25.2 in [Goebel and Reich (1984)]). Theorem 5.8 Let p be the Poincare hyperbolic metric on B, and let F : B —> B be a p-nonexpansive mapping. If F is fixed point free, then there is a unique boundary point a e 9 B such that all the ellipsoids E(a, k), k > 0,
128
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
are F-invariant.
In other words, (5.34)
<Pa(F{x)) < if>a(x), where _ |l-(a,a)|2 ^
a W
~
i _ ||_i|2
•
(i).6b)
To prove this theorem we use the following generalized version of Julia's Lemma. Lemma 5.1 Let F : B —> B be a p-nonexpansive self-mapping of M. Suppose that for some sequence {an}'^L1 c B which converges to a boundary point a s dB, the following conditions hold: 6{F)
= lim i f J I ^ H < oo
n-oo l - | | a n | | and b = lim^oo F(an) € 3 1 . Then for each xeM,
(5.36)
(5.37)
Since F is /o-nonexpansive, we have p(F(x),F(an))
(5.38)
a(F(x),F(an))>o-(x,an),
(5.39)
or, equivalently,
where
(5.40) x,yeM. Writing the latter inequality explicitly we get
(1 - \\F(x)f
)(1 - \\F(an)f)
\l-(F(x),F(an))\2
>
( 1 - N | 2 ) ( l - Hanil2) |l-(^,an)|2
(5.41)
or
ll-(F(*),F(a n ))l 2 l-|lf(on)|| a \l-(x,an)\2 i-\\F(x)\\* ~ l - K I I 2 " i-INI 2 •
(
^
129
The Denjoy-Wolf Fixed Point Theory
Proof of Theorem 5.8 Let F : B —>IB be a p-nonexpansive self-mapping of I . Consider a mapping Ft : B -> B denned by Ft = tF, 0 < t < 1. We have />(#(:«;),Ft(i/)) < tp(F(x),F(y)) < tp(x,y).
(5.43)
So, for each t G [0,1) the mapping Ft : B —> B is a strict contraction with respect to the metric p. Hence it has a unique fixed point xt = Ft(xt). In addition, xt - F{xt) = Ft(xt)
- F{xt) = (t- l)F(xt).
(5.44)
So ||xt — -F(a;t)|| —> 0 when t tends to 1. Thus, if F has no fixed point in B we must have that lim_ ||a;t|| = 1.
(5.45)
On the other hand, since B is weakly compact, there is a sequence tn —> 1~ such that {xtn} weakly converges to an element a e B. But if we assume that ||a|| < 1, then we have p(F(xtn),F(a))
(5.46)
(5.47)
or, equivalently,
where a(x v)
- (i-INI2)(i-llg|la)
|i-(z,y>|2 Explicitly the latter inequality implies a{x>y)-
(5 48)
•
l-4ll^H 2 .(l-HI 2 )ll-^K,F(a))l 2 ~ 1-IKII 2 - (l-||F(o)P)|l-|2 •
(5"48)
^
y ;
Letting n —> oo, we get a(a,F(a)) > 1. Hence, a = F(a). This contradiction shows that a must lie on the boundary of B, i.e., \\a\\ = 1. It follows that the convergence of Xtn to a is, in fact, strong, that is, \\xtn — a|| —»• 0 as n —> oo.
(5.50)
In addition, the strong limit lim F{xtn) = lim — xtn = a
n—>oo
n—+oo t n
(5.51)
130
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
and (by passing to a subsequence, if necessary)
(5.52) Setting an = xtn and a = b in the above version of Julia's Lemma, we obtain the required inequality: <Pa(F{x)) < <pa(x).
(5.53)
Finally, we claim that a point a G <9B satisfying <pa(F(x)) < (fa(x) must be unique. Indeed, assume that there is b G dM such that ipb(F(x)) <
o/B. G IB. point curve (5.54)
converges to a as t goes to 1. R e m a r k 5.3 As a matter of fact, it can be shown (see [Goebel and Reich (1982)], [Goebel and Reich (1984)] and [Shafrir (1992b)]) that for each y G B the approximating curve xt = {l-
t)y + tF(xt)
(5.55)
is well defined and converges to a fixed point of F. If F has no interior fixed point in B, then for each j / £ l , this curve converges to the same (Wolff's) sink boundary point a G dM. Now the question is whether the sink point a in Theorem 5.7 is also the Denjoy-Wolff point of F, i.e., is it attractive?
The Denjoy-Wolf Fixed Point Theory
131
For the finite dimensional case (B is then the open Euclidean ball in H = Cn) and when F £ Hol(B) the affirmative answer was given by B. MacCluer [MacCluer (1983)]. Theorem 5.9 Let H = Cn be a Euclidean complex space (finite dimensional Hilbert space) and let B be the open unit ball in H. If F e Hol(B) has no fixed point in B, then there is a point a of norm I (a € dM) such that the iterates Fn(z) converge to a for all z £ B, uniformly on compact subsets ofM. For infinite dimensional Hilbert balls, A. Stachura [Stachura (1985)] has constructed a counterexample which shows that this convergence result fails even for biholomorphic self-mappings. Nevertheless, some restrictions on a mapping from Hol(B) lead to a generalization of Theorem 5.6. The following result was obtained by C.-H. Chu and P. Mellon [Chu and Mellon (1997)]. Theorem 5.10 Let B be the open unit ball in a Hilbert space H, and let F € Hol(B) be a compact mapping with no fixed point in B. Then the sink point a in Theorem 5.7 is attractive, i.e., the sequence {Fn} of iterates of F converges locally uniformly on B to the constant mapping taking the value a. Although this theorem contains Theorem 5.6 (as well as the abovementioned finite-dimensional result of B. MacCluer), it does not apply to the automorphisms of B because of the compactness restriction. In this connection we have to mention a result by T. L. Hayden and T. J. Suffridge [Hayden and Suffridge (1971)] which complements our information (see also [Suffridge (1974)]). Theorem 5.11 Let B be as above, and let F E Hol(B) be an automorphism of B with exactly two fixed points on dM. Then one of them is an attractive sink point of F. Remark 5.4 It is known that each automorphism o / B can be extended to an automorphism of B and that it has a fixed point in B (see [Hayden and Suffridge (1971)]). Moreover, if it does not have a fixed point in M, then it has one or two fixed points on the boundary. For holomorphic self-mappings of the Hilbert ball the following versions of the Julia-Caratheodory and the Julia-Wolff-Caratheodory theorems are very important complements to the Denjoy-Wolff theorem.
132
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
We need some additional notions which are well-known in the finitedimensional case (see, for example, [Rudin (1980)]). Definition 5.5 A curve A : [0,1) —> B is said to be asymptotically normal at a point T € B if
(5.56) M ^ < M < o o ,
0<.
(5.57)
where X(s) is the ortogonal projection of A(s) onto the complex line through 0 and T: (5.58)
X(S) = (A(S),T)T.
Let h be a holomorphic function on B with values in the complex plane. We say that h has a restricted limit L at r € dB if h has the limit L along every curve which is asymptotically normal at r. Definition 5.6 Let / : B —> H be a holomorphic mapping on B, and let r € dB. We say that / has a finite angular derivative at r if for some element y C B, the function / i : B —» C defined by h{X}-
(5.59)
l-(x,r) (5>59)
has a finite restricted limit at r. We denote this limit
Zf'(r).
Theorem 5.12 Let B be the open unit ball in the finite dimensional Hilbert space H = Cn, let F be a holomorphic self-mapping of B, and let a be a boundary point. The following are equivalent: (i) S(F) = lim inf ~|}.AyH < oo, where the limit is taken as x approaches a unrestrictedly in B. (ii) F has finite angular derivative at a; (Hi) F has a restricted limit b e 5B and (F'(x)a,b) has a restricted limit K ata£ dB. Moreover, if one of these conditions holds, then 5(F) = K. We will call the number S(F) the Julia number of F at a. Theorem 5.13 Let B be the open unit ball in the finite dimensional Hilbert space H = Cn, and let F be a holomorphic self-mapping ofM with
133
The Denjoy-Wolf Fixed Point Theory
no fixed point in B. Then there is a unique boundary point a € 3B such that
0 < S(F) = flim inf 1 ~ l l ^ ( f l l > ) - Z F » < 1. Y
x—>a
1 — ||x|| /
(5.60)
Remark 5.5 As a matter of fact, a similar assertion also holds for the infinite dimensional case if we replace the angular derivative by the so-called radial derivative of F at its boundary sink point
T F'{a)=j^_
^ y ^ .
(5.6D
Moreover, rather surprisingly, this derivative turns out to exist even if F is not necessarily holomorphic, but only p-nonexpansive. Theorem 5.14 Let B be the open unit ball in a complex Hilbert space H and let F : B —> B be a p-nonexpansive self-mapping of B with no fixed point in B. Assume that a € dM is the sink point of F and 5 = 6(F) is the Julia number at a. Then the radial derivative T F>{a) l= l i m r—»1~
l-WaW
(5.62)
1— T
exists and equals 8. Proof.
First, by the generalized Julia's Lemma 5.1, we have |l-(F(ra),a)l2 h
(1-rf
1-r
(5.63)
This implies that lim_(F(ro),o) = l.
(5.64)
Now we calculate l-|(F(ra),a))|
(l-|(F(ra),q)|)2
1-r
{l-r)(\-\{F(Ta),a)\)
<
H W f
-1-|^(~).->|2
l + |
r
'
( ] (5.63)
134
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
Therefore it follows from (5.63) that l
l - | < F1(-r ra )1+r ,a)| ' '
+
|(F(m),q)|
Thus, by using (5.64), we obtain lim
r—i-
l-\(F{ra),a)\ ' \ v h / [ < S.
(5.67)
1 - r
v
'
On the other hand,
1-II^MII < l-KF(ra),a)| _ 1-r
(5.68)
1-r
Since ^hminf1-"^", x-»a 1 — ||x||
(5.69)
where the limit is taken when x approaches a unrestrictedly, we get that hm
r—i-
l-||F(nO|| 1-r
=
n m l-\{F{ra),a)\ r->i1-r
=&
(5.70)
Returning to inequality (5.63), we can rewrite it as follows:
|i-<W,,LJ£WE.
( , 7I)
Letting r tend to 1~ and using (5.70), we get |l-(F(ra),o)|
lim •!
i—»i-
But again, |l — (F(ra),a)\ by (5.70) and (5.72),
N V
;
1— r
'
/[
<<5.
> 1 — |{F(ra),a)| and therefore we have,
aJi-W,.)!^
r-»i-
(5.72)
1— r
(5.73)
Combining now (5.70) and (5.73), we get
ll-(F(ra), a )| r^i- l-|(F(ra),a)|
=
(5.74)
The Denjoy-Wolf Fixed Point Theory
135
It follows that arg(l — (F(ra),a)) tends to 0 as r tends to 1~. Thus we obtain I h n r-i-
- ^ ' ^ . 1— r
1
The theorem is proved.
(5.75) •
Now, using the Lindelof principle (see [Rudin (1980)]) and Theorem 5.14, we obtain Theorem 5.13 as well as Theorem 5.12 by applying an appropriate rotation of the unit ball. Combining now Theorem 5.14 with the proof of Theorem 5.8, we obtain a stronger version of the latter theorem; it is a generalization of the JuliaWolff-Caratheodory theorem. Theorem 5.15 Let F : B —» B be a p-nonexpansive self-mapping o / B . If F has no fixed point in B, then there is a unique point a s dM such that fa(Fn(x))
< 6n<pa(x),
(5.76)
where
0 « - r-»ito I 1 '^"'•"I-Umbf •-'??»
(5.77)
is the radial derivative of F at this point a.
Definition 5.7
A mapping F s 7VP(B) is said to be of hyperbolic type
if it is fixed point free and its radial derivative at its sink (boundary Wolff's) point is strictly less than 1. Corollary 5.2 Each p-nonexpansive mapping of hyperbolic type is power convergent to its boundary sink point.
5.3
Convex Domains in C n
In 1941 M. H. Heins [Heins (1941)] extended the Denjoy-Wolff Theorem to a finitely connected domain bounded by Jordan curves in C. His approach is specific to the one-dimensional case. Another look at the Denjoy-Wolff Theorem is provided by a useful result of P. Yang [Yang (1978)] concerning a characterization of the horocycle in
136
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
terms of the Poincare hyperbolic metric in A. More precisely, he established the following formula: Urn [p(X, /i) - p(0, /x)] = | log '* ~ ^
.
(5.78)
So, in these terms the horocycle T>a in A can be described by the formula
V(a,R) = {z € A : lim (p(z,u) - p(0,w)) < i logfl).
(5.79)
Since a hyperbolic metric can be defined in each bounded domain in C n , one can try to extend this formula and use it as the definition of the horosphere in a domain in C". Unfortunately, in general the limit in (5.79) does not exist. Therefore, the new idea of M. Abate [Abate (1988a)] was to study two kinds of horospheres. More precisely, he defined the small horosphere Ezo(x, R) of center x, pole z0 and radius R by the formula
EZ0(x,R) = \z € V : \imsup[Kv(z,w)
- KT>(ZO,W)} < - logR,}, (5.80)
and the big horosphere of center x, pole ZQ and radius R by the formula
FZ0(x,R) = {z e V : limmf [Kv(z,w) - Kv(z0,w)) < i logi?},(5.81) where D is a bounded domain in C" and Kx> is its Kobayashi metric. For the Euclidean ball in C n , EZo(x,R) = FZo(x,R) (see [Abate (1988a); Mercer (1992); Mercer (1993)]). Thus each assertion which states for a domain V in C n the existence of a point a e &D such that fn(Ez(a,R))cFz(a,R)
(5.82)
for all z G T>, R > 0, / e Hol(£>) and n = 1,2,... is a generalization of Wolff's Theorem 5.3. This is true, for example, for a bounded convex domain in C n . A more general result was established by M. Abate in another work [Abate (1989b)]. Theorem 5.16 Let V be a bounded complete hyperbolic domain with a simple boundary, and let z e V. Suppose that for some f € Hol(Z?) the sequence of iterates {/"} is compactly divergent. Then there is a £ &D such that for all z €T>,R> 0 and n = 1,2,... the inclusion (5.82) holds.
The Denjoy-Wolf Fixed Point Theory
137
In fact, the assumptions of Theorem 5.16 are not sufficient to ensure a convergence result. Indeed, in order to generalize the notion of a sink point we give the following definition. Let V be a domain in a Banach space X and let F G Hol(P). We will say that a point x G dT> is a boundary sink point for F if there exist two sets of neighborhoods {Ua} and {Va}, a £ A {& directed set), in V such that the following conditions hold: (i) (ii) (iii) (iv) (v)
Ua cVa C P; xeUa; HaeAUa = naeAVa = 0; F(Ua) C Va; For ax < a2, ax,a2 G A, Ua2 C Uai and Va2 C Vai.
The following example is due to C.-H. Chu and P. Mellon [Chu and Mellon (1997)]. Example 5.1 Let V be the open unit bidisk in C 2 , i.e., V = A x A, and let h € Hol(A) be a fixed point free mapping with the Denjoy-Wolff point C, € dA. Consider the mapping F G Hol(P) denned as follows: F(z, w) = (eilfiz, h(w)),
(5.83)
where 0 < (p < 2TT. It is clear that if F has a sink point a £ dV, then a — (0, £) G dT>. At the same time, the sequence of iterates {Fn} does not converge to any boundary point of V. Nevertheless, the convergence result does hold for bounded strongly convex C 2 domains, and for strongly pseudo-convex hyperbolic domains with a C 2 boundary [Abate (1988a); Abate (1988b)]. To generalize these facts, M. Abate denned the following notion [Abate (1989b)]: A domain D c c C is said to be F-convex at x G dV if for all z G V and R > 0, Fz(x,R)DdV={x}.
(5.84)
The domain V is said to be F-convex if (5.84) holds for each x G &D. His result is the following one: Theorem 5.17 Suppose that in addition to the assumptions of Theorem 5.16 the domain V is F-convex. Then there is a point a G dV which is attractive for f.
138
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
Certain types of domains are known to be F-convex. We mention, for example, strictly pseudo-convex domains with a C 2 boundary, and the so-called m-convex domains (see [Abate (1988a); Abate (1988b); Mellon (1996)]).
5.4
Domains in Banach Space.
Unfortunately, our knowledge does not include positive results regarding extensions of Theorems 5.3 and 5.6 to general Banach spaces. Indeed, a simple example shows that there is a situation where even a sink point does not exist. Example 5.2 Consider the Banach space CQ of all complex sequences x — (x\,X2, • • •,xn,...) such that xn —> 0 as n tends to infinity, with the max norm, and the affine (hence holomorphic) mapping F : Co —> Co defined by F(x) = (a,X\,X2-- •), where o ^ O , \a\ < 1. It is clear that F maps the open unit ball V of Co into itself and that it is continuous on V. If F had a sink point on dT>, then it would necessarily be a fixed point of F in T>. But F has no fixed point there. However, the question is still open for reflexive Banach spaces. Moreover, if V has a strictly convex boundary and F is compact, then the convergence result holds [Kapeluszny et al. (1999a) and (1999b)]. Proposition 5.2 Let X be a strictly convex Banach space, and let F be a holomorphic compact self-mapping of the open unit ball T> of X without a fixed point in V. Then F is power convergent on T>. The situation is more fully understood when F has a fixed point in the domain. From now on we will assume that T> is a bounded domain in a Banach space X and that F 6 Hol(2?) has an interior fixed point a £ T>. Of course, as simple examples show, one cannot expect that a is always an attractive point even if F is not an automorphism (consider Example 5.1 with h(w) — w2). Nevertheless, rephrasing the Denjoy-Wolff Theorem 5.6 in the form of Remark 5.2, one can point out several generalizations of this result. Proposition 5.3 ([Vesentini (1983); Vesentini (1985); Khatskevich and Shoikhet (1984)]). Let V be a bounded domain in X and let F € Hol(£>). Suppose that F has an interior fixed point a GT> and let F'(a) be its Frechet
The Denjoy-Wolf Fixed Point Theory
139
derivative at this point. Then a is a locally uniformly attractive point of F if and only if the spectral radius of F'(a) is strictly less than 1. Proof. The necessity of the assertion immediately follows from the Cauchy inequality if we note that (Fn)'(a) = [F'(aW
(5.85)
by the chain rule. To establish sufficiency, denote F'(a) by A. If p(A) < 1, then it follows from the Rutickii theorem that there is a norm || • ||i equivalent to the original norm of X such that ||A||i(= sup ||Aa;||i) < 1. Thus there is a V
IWIi=i
'
neighborhood (a ball) U CC V of the point a such that sup||F'(a;)||1=q
(5.86)
x€U
The mean-value theorem implies that for all x and y £U, \\F(x)-F(y)\\1
(5.87)
Now, by the Banach Fixed Point Principle, we get that the sequence { F " } ^ ! converges uniformly to a on U. Using the Vitali Theorem we obtain our assertion. • Remark 5.6 As a matter of fact, we will see below that this assertion is a consequence of a much more general assertion due to E. Vesentini (see Theorem 5.18). Combining Proposition 5.3 with the Earle-Hamilton [Earle and Hamilton (1970)] fixed point theorem, one can formulate a geometrical characterization of the attractive fixed point. Proposition 5.4 ([Khatskevich and Shoikhet (1984); Khatskevich and Shoikhet (1994a)]). Let V be a bounded domain in X and let F e Hol(X>). Then F has an attractive fixed point in V if and only if there exist a domain P C D and an integer n > 0 such that Fn maps V strictly inside itself. If a is an attractive fixed point of F, then by definition it is unique and F is power convergent. For P c C , the converse is also true: If F € Hol(I>) is not the identity and has an interior fixed point, and F is power convergent, then a is unique.
140
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
However, this is not always the case in higher dimensions. Indeed, the following example exhibits a function F, mapping the bidisk B 2 in C2 into itself, which is power convergent, but itsfixedpoint set consists of an infinite number of points (actually it is a submanifold of V of dimension 1). Example 5.3 Let V = B2 = A x A and F = (zi, § {z\ + z\)). Then F is a power convergent mapping, but all the points of the form (zi, 1 - ^1 — zi) are fixed points of F. A full description of such a situation was obtained by E. Vesentini [Vesentini (1983); Vesentini (1985)]. Theorem 5.18 (Vesentini) Let V be a bounded convex domain in a Banach space X, and let F belong to Hol(P). Suppose that F has a fixed point a e V, and denote the spectrum of the linear operator F'(a) by a(F'(a)). Then F is power convergent if and only if the following two conditions hold: (i) a(F'(a)) C AU{1}; and (ii) I is a pole of the resolvent of F'(a) of order at most 1. Comments.
Condition (ii) is actually equivalent to the condition Ker(J - F'{a)) 0 Im(/ - F'(a)) = X
(5.88)
(see, for example, [Gohberg and Markus (I960)] and [Lyubich and Zemanek (1994)]). It is also known that conditions (i) and (ii) are equivalent to F'(a) being power-convergent to a projection P onto Ker(/ — F'(a)). So, if R £ Hol(P) is the limit point of {Fn} under these conditions, then R = a is constant if and only if P = 0. In general, there is a neighborhood U of a such that R2 — Ron U, that is, R is an idempotent of the algebraic semigroup Hol(t/). In this case the mapping R is said to be a local holomorphic retraction, and its image R(U) is called a holomorphic retract of U. In our setting this means that FixF n U = R(U) is a submanifold of U tangent to Ker(/ - F'(a)) (see [Cartan (1986)]). This fact describes the general structure of the fixed point set. Actually, only condition (ii) or (5.88) is sufficient for FixF to be a local retract (see, for example, [Shoikhet (1986); Shoikhet (1993)], [Vigue (1986); Vigue (1991a)], [Mazet and Vigue (1991); Mazet and Vigue (1992)]). Moreover, if V is convex, FixF is a global retract of T>, and hence a connected complex submanifold of T>. To prove this assertion, as well as Theorem 5.18, we need the following lemma.
141
The Denjoy-Wolf Fixed Point Theory
Lemma 5.2 Let F be a holomorphic mapping defined on a neighborhood V of the origin and suppose F(0) = 0. Assume that all the iterates {Fn}^=0 are well defined on V and bounded, i.e., \\Fn{x)\\ < M < oo,
(5.89)
xeV.
Uf Ker(J - F'(0)) 0 Im(/ - F'(0)) = X,
(5.90)
then there is a neighborhood U C V such that F'ix(F) nU is a complex submanifold ofU tangent to Ker(7 — F'(0)). Moreover, there exists a holomorphic mapping * defined on U such that \I>2 = * and Fixtf (¥) = Fixu(F).
(5.91)
Proof. Let P be a linear projection onto Ker(7 — F'(0)) and Q its complement, i.e., P + Q = I. Let U\ be a neighborhood of the origin in the space PX, and consider any holomorphic mapping / defined on U\ with values in KerP (= Im(7 - F'(0)). We claim that if / satisfies the condition ||/(u)|| =o(||u||),
(5.92)
PF{u + /(«)) = u
(5.93)
then
for all u £ U\ such that u-\- f(u) S U. Indeed, fix such u G Ui and consider the vector-functions (5.94)
<^(A) = PFn(Xu + f(Xu))
with A e A = { A G C : |A| < l } . By our assumption the sequence {ipn} is uniformly bounded, i.e., |K(A)||<||P||.M,
„ = 0,1,2,...,
AeA.
(5.95)
Since ipn : A —> PX is holomorphic, one can represent it in the form fn(A) =
(5.96)
where hmrl £ X is the first nonvanishing coefficient in the power series expansion of
(5.97)
142
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
Denoting S = F — F'(0), we get by a direct computation,
(5.98)
Similarly, by using an induction argument we obtain Y?n(A) = Xu + nhmiXmi + . . . . Thus we get that mn = mi and hmn = n • hmi, n = 1,2, other hand, it follows from the Cauchy inequalities that \\hmj
< \\P\\ • M.
(5.99) But on the (5.100)
This inequality shows that hmi must be zero, and therefore
(5.101)
Setting A = 1, we obtain
(5.102)
Since u is arbitrary, our claim is proved. Now consider the set FixF n U defined by the equation x = F(x),
x€U.
(5.103)
Using the projections P and Q, one can rewrite this equation in the form f u = PF(u + v) i \v = QF{u + v)
(5.104)
where u = Px and v = Qx. Since the operator I—QF'(0) is invertible on Im(7—.F'(O)), it follows by the implicit function theorem that the second equation of the latter system has a unique solution v = f(u) for all u sufficiently small. It is also easy to see that ||/(u)|| = O(||M||)- By using our previous claim we have that the first equation is satisfied with the substitution v = f(u), i.e., u = PF(u + f{u)).
(5.105)
In other words, all the solutions of the equation x = F(x) can be represented in the form x = u + f(u).
(5.106)
The Denjoy-Wolf Fixed Point Theory
143
Thus, shrinking U if necessary, we obtain that Fix(F) f~l U is a complex submanifold of U. Finally, if we define the mapping *:£/—» X by V(x) = Px + f(Px),
(5.107)
we get * 2 = * and Fix(#) n U = Fix(F) n U. This concludes our proof.D Proof of Theorem 5.18 First we note that by Lemma 5.2 and the Vitali property of holomorphic mappings in the topology of local uniform convergence over V, we can assume that V = U is a convex domain in X. Then there is a retraction ip : D —> Fix(F) which satisfies the condition 7poF = ip.
(5.108)
In addition, we have shown (see also the H. Cartan Theorem [Cartan (1986)]) that in a neighborhood U of the fixed point a of F we can find a local chart g : U —> V such that g{a) = 0 and such that gotpog~1 = P
(5.109)
is a linear projection. Now consider the mapping G = goFog~1,
(5.110)
defined on some neighborhood W of zero, together with its iterates Gn = goFnog~1. (Indeed, by the boundedness of {Fn} this sequence is uniformly Lipshitzian in some neighborhood of a. Hence, since Fn(a) = a,we can find a neighborhood W such that Fn(g~1(W)) C U.) We now have PG = got/jog-lgoFog-1
= gotjjog-1 = P.
(5.111)
In addition, G'(0) = g'{a) o F'(a) o [p'(a)]- 1 and therefore
144
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
However, once again, a retraction onto the fixed point set can be obtained by this simple iteration method only if condition (i) holds, i.e., only if F'(a) has no spectrum points on the unit circle except, possibly, 1. This brings us to another interesting aspect of fixed point theory: the study of the holomorphic retracts of a domain. 5.5
Holomorphic Retracts and the Structure of the Fixed Point Sets.
Let, as above, X be a complex Banach space, and let T> be a domain (open connected subset) in X. Lemma 5.2 in the previous section provides us with information on the local structure of the fixed point set of a holomorphic self-mapping of T> under a certain condition of regularity. To elaborate on this result in more detail we give the following definition. Definition 5.8 Let F £ Hol(£>). A point a € Fix(F) is said to be quasi-regular if the following condition holds: (*)
K e r ( / - F ' ( G ) ) © I m ( / - F ' ( a ) ) = X.
(5.112)
If, in addition, Ker(7 — F'(a)) = {0}, i.e., the linear operator / — F'(a) is invertible, then we say that a is a regular fixed point of F. By the implicit function theorem, it is clear that a regular fixed point is an isolated point of the set Fix(F), and that in the case of a finitedimensional X each fixed point is quasi-regular (or, in particular, regular). Thus, in these terms Lemma 5.2 can be reformulated as follows: Theorem 5.19 Let V be a domain in a complex Banach space X, and let F € Hol(P) have a quasi-regularfixedpoint a e Fix(F) c V. If F(D) is bounded, then there is a neighborhood U C V with a £ U such that Fix(F) n U is a complex submanifold of U tangent to Ker(7 — F'(a)). If, in addition, V is a bounded convex domain, then it can be shown (see Theorems 5.22 and 5.23 below) that Fix(i?) is a connected subset of V. This result for finite dimensional Banach spaces was obtained by J.P. Vigue [Vigue (1986)] and for the general case by P. Mazet and J.P. Vigue [Mazet and Vigue (1991)] (see also [Shoikhet (1986)]) by using a retraction method.
The Denjoy-Wolf Fixed Point Theory
145
Earlier, W. Rudin [Rudin (1978)] (see also [Rudin (1980)]) established a surprising result regarding the global structure of holomorphic selfmappings of the Hilbert ball. See also [Goebel and Reich (1984), p. 121]. Theorem 5.20 Let B be the open unit ball in a complex Hilbert space H, and let F € Hol(B). Then Fix(F) is an affine submanifold ofM. Proof. First we note that by direct calculation one can easily verify that each automorphism of the unit ball translates an affine subset of B onto another affine subset of B. Therefore it is sufficient to prove the theorem under the assumption that Fix(F) contains the origin. In other words, we assume that F(0) = 0. We show that in this case Fix(F) = Fix(A) n B, where A is the linear operator defined by the Frechet derivative of F at the origin, i.e., A = F'(0). Indeed, it follows by the Schwatz Lemma that ||F(i)|| < ||x||,
xeB,
(5.113)
and \\A\\ < 1.
(5.114)
Further, one can write each x G B in the form x = ru, where u G dM and 0 < r < 1. Let A be the open unit disk in the complex plane C and consider the holomorphic function g on A defined by g(X) = (F(Xu),u),
(5.115)
where (•, •) denotes the inner product in H. Clearly, g(0) = 0 and g'(0) — (Aii.u). If now x (= ru) € Fix(F), then g(r) = r and, by the one-dimensional Schwarz Lemma, g(X) = A for all A G A. Consequently, g'(0) = (Au,u) = 1 which implies Au = u. Hence Ax = x. Conversely, assume that x (= ru) G Fix(A). Then again g'(0) = (Au, u) = 1 and, by the Schwarz Lemma , g(X) = A G A. In particular, g(r) (= (F(ru),u)) = r. On the other hand, we have that ||F(ru)|| < ||r||. Setting y = r - 1 F ( r u ) , we have ||y|| < 1 and (y,u) = 1. Therefore y = u. This means that F{ru) = ru, or, which is the same, F(x) = x. The theorem is proved. • The simple example from the previous section (Example 5.3) shows that this result does not hold even in C2 if we replace the Euclidian ball by the
146
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
polydisk (i.e., the unit ball of C2 in the Chebyshev norm). Nevertheless, as we have already mentioned, the global structure of the fixed point set can be described by using the retraction method and H. Cartan's theorem. Definition 5.9 A mapping
The Denjoy-Wolf Fixed Point Theory
147
Even if the retraction onto Fix(F) exists, it need not be the limit of the iterates of the operator, even in the case of a linear operator A such that {0} ± Ker(J -A)± Ker(J - A)2. We will now mention several results concerning both these questions. Recall that a net {FJ}J&A C Hol(P, X) is said to converge to a mapping F £ 1101(1?, X) in the topology of locally uniform convergence over T> (or, briefly, T-converge) if for every ball B c c D , lim supllFjfc) - F(a;)|| = 0. "
j€Ax€B"
(5.116)
In this case we also write F = T-lim
Fj.
(5.117)
It is obvious that if F € Hol(P) is power convergent, then T = Fix(i?) ^ 0 and tp = T — lim Fn is a retraction onto T. So, by using a result due n—»oo
to J. J. Koliha [Koliha (1974)] Vesentini's theorem (Theorem 5.18) can be reformulated as follows. Theorem 5.21 Let V be a bounded convex domain in X, and let F G Hol(P) with T = Fix(F) ^ 0. If a € Fix(F) is a quasi-regularfixedpoint and A = F'(a), then F is power convergent to a retraction
(5.118)
148
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
It is clear that if in the setting of the above theorem cr(A) c A, then a G J- is a regular point. In addition, it was shown that such a point is the unique fixed point of F in V. As a matter of fact, as we will see in the sequel, for a bounded convex domain, the regularity of a point a G Fix(F) is already sufficient for its uniqueness, i.e., if a is regular, then Fix(F) = {a}. R e m a r k 5.8 Vesentini's theorem is also true for bounded domains which are not necessarily convex, but satisfy the following maximum modulus principle: For each f G Hol(£>, T>) such that f(D) D &D ^ 0, it follows that f(V) c &D. However, we will mainly consider convex domains in X because this is all that is needed in order to describe Fix(F) locally in each bounded T>. Indeed, it has been shown by P. Mazet [Mazet (1992)] that if V is bounded, then for each F G Hol(D) and each a G F\x(F), there is a convex neighborhood U
In addition, we need the convexity to consider different types of mean ergodic procedures. As a matter of fact, if the conditions of Theorem 5.20 hold for at least one point of T, then they hold for all the points of T. Moreover, if 1 G cr(A), then T contains infinitely many points because it is a retract of T>, hence a connected submanifold of V tangent to Ker(7 — A). We will see below that the latter fact is true whenever condition (5.112) holds, even if the spectrum a(A) contains other points on the boundary dA of the open unit disk A different from 1. But in this case, of course, by Vesentini's theorem, F is not power convergent and therefore the question of approximating its fixed points is still open. Nevertheless, if V is a bounded convex domain in X, and F G Hoi(T>) has at least one quasi-regular fixed point in V, then there is another mapping
Cn = - V Fk
(5.119)
The Denjoy-Wolf Fixed Point Theory
149
and proved the following results. Theorem 5.22 Let V be a bounded convex domain in X = Cn, and let F e Hol(X>) with T = Fix(F) ^ 0. Then there is a subsequence {Cnk} of {Cn} which T-converges to a power-convergent holomorphic mapping ip:T> -> P with Fix(
150
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
each A S (0,1), the averaged mapping FX = XI + {I- X)F
(5.120)
is power convergent. (ii) If for at least one A € (0,1) the mapping F\ defined by (5.120) is power convergent, then Fix(i r ) consists of quasi-regular points. First we observe that it is sufficient to establish this theorem for bounded domains. Indeed, if a is a fixed point of F (hence of F\), the hyperbolicity of T> implies the existence of a ball B(a,R) (with respect to the Kv pseudometric) centered at a which is bounded and strictly inside T>. Since F\ is nonexpansive with respect to Kx>, this ball is invariant under F\. Suppose now that (i) is established for that ball and let B(c,r) be any ball (with respect to the norm) which is strictly inside V. Then the family {F£} is bounded on the convex hull of B(a, R) UB(c, r) (which is also strictly inside T>), and therefore FA is power convergent on B{c,r) by the Vitali property. Proof of Theorem 5.24 0) Let a e Fix(F) be a quasi-regular point, and let A\ = XI + (1 - X)F'(a) for A € [0,1). It is clear that a £ Fix(F A ), where F\ is denned by (5.120) and Ax = (Fx)'{a),
Ae[0,l),
(5.121)
with Ao = A = F'(a). We intend to show that for each A G (0,1) the mapping F\ satisfies the conditions of Vesentini's Theorem. First we note that condition (*) (see formula (5.112) in Definition 5.8) is obvious because I-AX
= (1-X)A.
(5.122)
Now we must show that the set a(A\)\{l} lies inside the open unit disk A for each A G (0,1). Once again, the Cauchy integral formula shows that the operator A\ is power bounded. Suppose now that there exists C £ <9A fl o-{A\) and £ ^ 1. Then we have for such C, CI-Ax = (l-X)(tI-A),
(5.123)
where t=YZje
a{A).
(5.124)
The Denjoy-Wolf Fixed Point Theory
151
It is clear that \t\ > 1. But on the other hand, \t\ < 1, since a(A) C A (see (5.118)). So, |i| = 1 and we have, by (5.124), £ = A + (l-A)te8A.
(5.125)
But this is possible only if £ — t = 1. This contradiction proves our assertion. (ii) Now, if for some A 6 (0,1) the mapping F\ is power convergent, then it follows from the Cauchy inequalities that for each o € Fix(F) the operator A\ = (F\)'(a) is power convergent too. This fact in turn implies condition (*) for A\. Hence, it follows from (5.122) that a is quasi-regular. The proof is complete. D So, setting A = ^ in Theorem 5.24, we have that the Cesaro average (see (5.119)) C2 = \{I + F)
(5.126)
is power convergent, i.e., it is sufficient to takep = 2 in Theorem 5.23. As a matter of fact, it turns out that the Cesaro averages defined by (5.126) are power convergent for all p = 2,3 Moreover, we are able to show that all proper convex combinations of the iterates of F are also power convergent. We will describe below a general scheme for constructing a retraction onto Fix(F). A simple but very important consequence of Theorem 5.24 is the socalled Cartan's uniqueness theorem, which is a generalization of the second part of the classical Schwarz Lemma. Corollary 5.3 Let V be a hyperbolic convex domain in X, and let F € Hol(P) be a bounded on each subset which is strictly inside T>. If a € T> is a fixed point of F such that F'{a) = I, the identity operator on X, then F(x) = x for all x € V. Now we return to a general approach to find a retraction onto Fix(F). It follows by Theorem 5.21 that to construct a retraction onto Fix(i r ) by using a power convergent mapping, we must find a mapping $ £ Hoi (D) such that T = Fix(F) = Fix($) and for some point a € T the operator <3>'(a) is power convergent. Lemma 5.3 be such that
Let A be a bounded linear operator in X, and let Ao £ &(A) Ker(A07 - A) © Im(A0/ - A) = X.
(5.127)
152
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
Suppose that there exist a domain fi C C and a holomorphic function f defined in a neighborhood ClofQ, with the following properties: (a) U D a(A) and Ao € dCl; (b) f(Q) C A; (c) Ao is a simple root of the equation /(Ao) = 1;
(5.128)
(d) |/(A)| ^ 1 for all A S 90, A ^ Ao. Then the linear operator B — f{A) : X —» X defined by the formula I /(A)(A/ - A)~ld\,
B=TT-.
(5.129)
where T C 0 is a closed path around cr(A), is power convergent. Proof.
It follows from Dunford's Spectral Mapping Theorem that a{B) = f(a(A)).
(5.130)
Hence, conditions (a), (b) and (d) imply that a(B)\{l] c A.
(5.131)
Consider now the function 3(A) = [ l - / ( A ) ] ( A 0 - A ) - 1 .
(5.132)
Condition (c) implies that <7(A) is holomorphic in a neighborhood of the point Ao e dfl, and /(A o ) ^ 0. In addition, g(X) ^ 0 for all A £ a(A), by condition (d), and therefore the operator C = g(A), defined by the formula
C=^~. [ 0(A)(AJ - Ar'dX,
(5.133)
is invertible in X. Furthermore, it follows from the multiplicative property of the calculus of L(X)-valued functions denned by (5.129) and (5.133) (see [Rudin (1973)]) and by formula (5.132) that I - B = C(\0I - A) = (\0I - A)C.
(5.134)
This implies Ker(J - B) = Ker(A0/ - A) and Im(J - B) = Im(A 0 / - A). Now (5.127) and (5.128) imply that B is power convergent. The lemma is proved.
•
153
The Denjoy-Wolf Fixed Point Theory
Theorem 5.25 Let V be a hyperbolic convex domain in X, and let F e Hol(P) with F = Fix(F) ^ 0. If' T contains a quasi-regular point a € V, p
P
then each mapping $ of the form $ = JT akF , where ^ a* = 1 and fc=O fc=O
0 < afc y£ 1 for all k, is power convergent.
Proof. As in the proof of Theorem 5.24, we may assume that V is bounded. Let $ G Hol(P) be defined by p
(5.135)
$ = y£iakFk, fc=o V
where J2 ak = 1 and 0 < at ^ 1 for all A; = 0,1,.. .p. k=o Consider the holomorphic function (polynomial) / : A —» A defined by the following formula: p
f(\) = j2akXk>
AGA-
( 5 - 136 )
fc=0
It is clear that / satisfies conditions (a)-(d) of Lemma 5.3, with Q. = A, and therefore the operator B defined by (5.129) is power convergent to a projection onto Ker(7— A), where A = F'(a). But it follows from the chain rule that $'(a) = B = f(A). Hence, $ is power convergent onto Fix($) (see Remark 5.10). Furthermore, (5.135) implies that Fix(F) C Fix($). At the same time, these sets are connected submanifolds in T> tangent to the same subspace Ker(J — A). Therefore they coincide in V. This completes • the proof. Remark 5.10 combination
IfT> 3 0, then it can be shown that each infinite convex oo
$ = 5> f c F f c ,
(5.137)
fc=o oo
where Yl sk = 1 and 0 < afc ^ 1 for all k, belongs to Hol(2>) and is also fc=o
power convergent to a retraction onto Fix(.F). We now describe an implicit method for approximating fixed points of holomorphic mappings which has been used many times in the theory of nonexpansive mappings (see, for example, [Goebel and Reich (1984)]).
154
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
Let V be a bounded convex domain in X, and let F belong to Hol(X>). For t G [0,1) and a fixed y £ V, consider the mapping $(x) = tF(x) + (1 — t)y, x G V. Since $ maps V strictly inside itself, the Earle-Hamilton theorem [Earle and Hamilton (1970)] implies that there exists a unique fixed point z = zt(y) G V of the mapping $. Moreover, z = T-
lim $ n .
(5.138)
n—>oo
The X-valued function zt{y), 0 < t < 1 is called an approximating curve. It was shown in [Kuczumow and Stachura (1990)] that if X = C™ and Fix(F) 7^ 0, then there is a sequence tn G (0,1), tn —> 1 such that for each y G V, the sequence ztn (y) converges to a fixed point of F in V. At the same time, changing our point of view, for each t G [0,1), zt = zt(y) holomorphically depends on y G T>, by (5.138). We denote this mapping by Tt. In other words, Tt is the unique solution of the nonlinear operator equation Tt=tFoTt + (l-t)I.
(5.139)
The mapping % belongs to Hol(2>) and it is easy to check that Fix(7i) = Fix(F)
t G (0,1).
(5.140)
For the Hilbert ball B (as we have already mentioned) the situation is as follows: The net {Tt} is convergent in B as t —» oo to a mapping ^ : B —> B. / / Fix(F) ^ 0, then \& is a holomorphic retraction onto Fix(F) (otherwise, $ is a constant a G SB which is the sink point of F). Also, since for each t > 0 the mapping {Tt} is a firmly holomorphic self-mapping ofM, the iterates {^n}^Li converge weakly to a holomorphic mapping ip : B —» B, which is a retraction onto Fix(F) if it is not empty. A similar situation occurs in general Banach spaces. Theorem 5.26 Let D be a bounded convex domain in a complex Banach space X, and let F G Hol(X>) with T = Fix(F) ^ 0. If F contains a quasi-regular point in V, then the mapping Tt defined by (5.139) is power convergent to a retraction onto J-'.
The Denjoy-Wolf Fixed Point Theory
Proof. Let aeT, operator equation
155
and let B = (Tt)'(a), t G (0,1). Then B satisfies the B = tAoB
+ (l-
t)I,
(5.141)
where A = F'(a). Setting r = ^ , we have B = [I + r(I - A)}'1.
(5.142)
Therefore, B can be defined by formula (5.129) with ^
l
+ ra-A)'
ASA.
(^43)
Since r > 0, this function maps Cl = A into itself and satisfies all the condition of Lemma 5.3. Thus, by this lemma, B is power convergent to a projection onto Ker(7 — A) and so Tt is power convergent to a retraction onto Fix(F). The theorem is proved. • Remark 5.11 Note that it follows from the Neumann series representation of the operator B in (5.129) that the mapping $ t defined by (5.137), oo
$t = J > f e F \
(5.144)
k=o
with ak = t(l-t)k, t e (0,1), has B = [ J + r ( / - A ) ] ~ 1 as its Frechet derivative at the point a. However, generally speaking, $ t and % are different in the nonlinear case. Finally, we observe that P. Mazet and J.-P. Vigue ([Mazet and Vigue (1991)]) have shown that under certain conditions on the Banach space X (for example, reflexivity), condition (*) of Definition 5.8 is not necessary. In particular, they proved the following result: Theorem 5.27 ([Mazet and Vigue (1991)]) Let V be a bounded convex domain in a reflexive complex Banach space X, and let F S Hol(X>) with T — Fix(F) ^ 0. Then T is a holomorphic retract ofT>. The scheme of the proof is as follows. Since T> is a bounded convex domain in X, the Cesaro averages 1
n"1
Cn = -YJFk n
fc=o
(5-145)
156
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
are well defined and bounded on V. Then the Montel property implies that {Cn} contains a subsequence {Cnk} which weakly converges to a holomorphic mapping * 6 Hol(P). It is clear that if a £ Fix(F) C T>, then Cn(a) = a and *(o) = a. At the same time, since X is reflexive, it follows n-l
by the mean ergodic theorem that (Cn)'{a) = £ £ (Fk)'(a) strongly conk=o
verges to P = ('t)'(a) ) which is a projection onto Ker(J — F'(a)). Hence the point a £ Fix(F) C Fix(^) is a quasi-regular fixed point of *&. Moreover, by Theorem 5.22 the mapping ^ is power convergent to a retraction \P G Hol(X>) onto Fix(4r). Without loss of generality, assume now that V is a neighborhood of the origin and a = 0. The crucial point of the proof is the following technical auxiliary lemma: Lemma 5.4 ([Mazet and Vigue (1991)]) Let Sn(h) denote the Taylor polynomial in the power series expansion of a holomorphic mapping h at a neighborhood of the origin. Then for the mappings F and $ defined above, the following equality holds:
Sn{F o *") = S n (* n ) = S n (tf n+1 ).
(5.146)
Now this lemma implies that F o $ = $, hence Fix(<3>) c Fix(F). On the other hand, it follows by the constructions of the mappings * and $ that Fix(F) C Fix(^) = Fix($). Thus we see that Fix(F) = Fix($) = Fix($) is a retract of V, hence a connected submanifold of V tangent to Ker(J - F'(a)) = Ker(J - P).
Chapter 6
Generation Theory for One-Parameter Semigroups 6.1
6.1.1
Continuous and Discrete One-Parameter Semigroups on Metric Spaces Discrete and continuous flows on a domain
Let D be a topological space. A family M(D) of self-mappings of D forms a semigroup with respect to the composition operation on D if FoGGM(D)
(6.1)
whenever F and G belong to M(D). If, in particular, D is a metric space with a metric p on D, then the set Np(D) of p-nonexpansive mappings on D forms a semigroup with respect to the composition operation. If D is a complex domain in a Banach space X, then the set Hol(D) of all holomorphic self-mappings of D also forms a semigroup with respect to the composition operation. The set Aut(D) of all the automorphisms of D forms a group with respect to the composition operation because of the inclusion F-1 £ Hol(£>)
(6.2)
which holds whenever F € Aut(D). Hence F~l G Aut(£>). Let now A be a topological (additive) semigroup with zero, and let there exist a natural ordering of A, i.e., r >t if and only if there i s s e i such that r — t + s. We will mostly consider two cases: (a) A = N U {0} = { 0 , 1 , 2 , . . . } (respectively, A = Z = { 3,-2,-1,0,1,2,3...}; (b) A=[0,T), 0 < T < oo (respectively, A = (-T,T), 0 < T < oo). 157
158
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
Definition 6.1 Let A be as above and let M(D) be the semigroup of selfmappings with respect to the composition operation of a topological space D. A family 5 = {Ft} C M(D) is called a one-parameter semigroup on D with respect to A if the following conditions hold: (i) Ft+s = FfFt, whenever s, t and s + t belong to A; (ii) Fo = / , the identity operator on D. In other words, a one-parameter semigroup S c M(D) is defined by a mapping A i-> M(D) which preserves the additive structure of A with respect to the composition operation on M{D). If A = N U {0} = { 0 , 1 , 2 , . . . } , then S = {F0,F1,F2,...,Fn,...}, Fn e M(£>), is called a one-parameter discrete semigroup. Actually, such a semigroup consists of the iterates of a self-mapping F = F\, because of conditions (i) and (ii), i.e., Fo = I, Fn = Fn, n = 1,2, If, in particular, (D,p) is a metric space and M(D) = NP{D), then for each F G NP(D) the set S = {Fn} U {/} forms a one-parameter semigroup of /9-nonexpansive mappings on D. If for each n > 0, Fn is invertible and F~n e NP(D), then S can be extended to a one-parameter group GT = {Fn} U {F~n}, n > 0. In this case condition (i) holds for all t and s in A = Z = {... — 2, —1,0,1,2...}. This set G is called a one-parameter discrete group of isometries on D because of the relation x,y € D.
p(F(x),F(y))=p(x,y),
(6.3)
We have already mentioned that for F G Hol(D) the family of iterates S = {Fn}%L0, Fo = I, can be considered a one-parameter discrete-time semigroup. In this case the vector field / = / — F is referred to as the generator of 5 (see [Kato (1966)]). It is clear that S can be extended to a one-parameter discrete group on D if and only if F € Aut(D). Definition 6.2 Let D be a topological space. A family 5 = {Ft : t £ (0,T)}, T > 0, of self-mappings Ft of D is called a (one-parameter) continuous semigroup if (i) F9+t = FtoFs,
0<s
+ t
(6.4)
Generation Theory for One-Parameter Semigroups
159
(ii) for each x € D, lim Ft(x) = x,
(6.5)
t->o+
where the limit is taken with respect to the topology of D. 6.1.2
Examples
Example 6.1 Let X be a Banach space and let L(X) be the space of all bounded linear operators on X with the norm \\A\\= sup ||Ac||, AGL(X),
(6.6)
xeX.
11*11=1 For a linear operator A e L(X) we can define the family {Bt}tzM. of bounded linear operators by using the following exponential expansion:
Bt = e~tA := J2 ^ p *kAk-
(6-7)
It is clear that the family {Bt} satisfies the semigroup property (i) and (ii), i.e., BtoBs=
e-tA o e~sA = e~^t+s)A - Bt+S,
(6.8)
and Bo = I, the identity operator on X. So if D = X this family forms a one-parameter group of self-mappings of D because etA = [e-*- 4 ]- 1 : X -» X
exists for each
* > 0.
(6.9)
If D is the unit ball of X and A is an accretive operator on X, then the family {Bt}t>o = {e~tA}t>o forms a one-parameter semigroup of so-called proper contractions of D by the inequality ||A||<1,
«>0.
(6.10)
Obviously, this family satisfies the limit property lim Bt = I
t-*o+
(6.11)
with respect to the operator topology on AT, i.e., ^lim \\Bt - I\\L{X) = 0.
(6.12)
160
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
Moreover, it is easy to see that, in fact, ||Bt-J||iW
(6.13)
Now it follows by the semigroup properties that (lim
||Bt+,-JB,|Uw = O
(6.14)
for all real t and s. Therefore such a semigroup is called a uniformly continuous semigroup of linear operators. Remark 6.1 As a matter of fact, we will see below that each uniformly continuous semigroup (group) of linear bounded operators on X can be represented by the exponential form (6.15)
Bt = e~tA for some A G L{X). In addition, it follows from the exponential series for e~tA that hm^(I-Bt) = A,
(6.16)
where the limit again is taken with respect to the norm topology of L(X). In view of the last property, the operator A is usually called the infinitesimal generator of the semigroup Bt (see a more general definition below). Example 6.2 Let X be the space of all bounded uniformly continuous functions defined on R = (—oo, oo) with the topology of uniform convergence. Define the family {Bt}t£R of linear operators on X by the formula [Bt(x)](y) = x(t + y), t G R, y G R,
(6.17)
where x e X is a bounded uniformly continuous function on R = (—oo, oo). It is clear that [Bo(x)](y) — x(y), i.e., Bo = I - the identity operator on X, and
[Bt+.(x)] (y) = x(t + s + y) = x(t + (s + y)) = [Bt(B3(x))} (y), (6.18) i.e., Bt+s = BtoBs,
t,s£R.
(6.19)
Generation Theory for One-Parameter Semigroups
161
So the family {Bt} forms a semigroup of bounded linear operators on X. Actually this semigroup is a group because of [5 t ]- 1 = B_ t
for all
t > 0.
(6.20)
However, regarding the continuity, one can assert only the pointwise convergence of this semigroup to the identity, i.e., lim (7 - Bt){x) = lim {sup|x(i + y) - x{y) I) = 0
t—»0+
*—*0+ t - y g n
J
(6.21)
for each x € X. Moreover, the limit (6.22)
\imUl-Bt)(x)
does not exist for all a; € X. Thus such a semigroup is not uniformly continuous on X. Other examples of linear semigroups can be found, for instance, in [Hille and Phillips (1957)], [Daletskii and Krein (1970)] and [Krein (1971)]. Example 6.3 Let X = C be the complex plane, and let D = A be the open unit disk in C. Let {Ft}t€m be the family of holomorphic self-mappings of A defined by _, . . z + tanh t . Ft(z) = , z&A. ztanh t + 1
, (6.23
It is easy to see that this family forms a one-parameter group of automorphisms of A which has the locally uniform continuity property (6.24)
]\m+Ft(z) = z,
uniformly on each compact subset of A. Moreover, there exists the locally uniform limit lim \{z- Ft(z)) = l - z \
(6.25)
which defines a holomorphic mapping on A. Example 6.4 Another example of a semigroup on A which does not consist of automorphisms may be defined as follows: ft(2) =
e'-^-l)'
**°-
{6"26)
162
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
Despite the fact that each Ft is a fractional linear transformation mapping A into itself, its inverse is not a self-mapping on A. So this semigroup cannot be extended to a group of self-mappings of A. At the same time we again have that
(6.27) uniformly on each compact subset of A, and (6.28)
}™=\(z-Ft(z)) = z-z2
exists for each z € A. Note that both the semigroups in Example 6.3 and 6.4 consist of mappings which are nonexpansive with respect to the Poincare hyperbolic metric of A, i.e., (6.29)
p(Ft(z),Ft(w))
for all pairs z and w in A and for each t > 0. Moreover, by the Schwarz-Pick Lemma, in Example 6.4 the strict inequality (6.30)
p(Ft(z), Ft(w))
holds for all z and w in A, z ^ UJ, and t > 0, while in Example 6.3 we have equality p(Ft(z),
Ft(w))
=p{z,w)
z,weA,
t>0.
(6.31)
In this case one says that the semigroup (group) {Ft}t>o consists of pisometries of the metric space (D,p). As a matter of fact, all one-parameter semigroups (groups) of automorphisms of A can be described as follows. Example 6.5
(see [Berkson et al. (1974)]).
(i) Each one-parameter group which consists of elliptic automorphisms of A is of the form F*W
-
(e^-H2)z + a(l-e^) _ l)z + x _ H 2 e ^ t - * > 0 ,
Q(e^t
(6-32)
where ip S R, a G A are given. (ii) Each one-parameter group of hyperbolic automorphisms of A is of the form Ft{z)
= (e*-l)z + /3(l-ae*)'
(6"33)
Generation Theory for One-Parameter Semigroups
163
where
(6.34)
where c G M, c 7^ 0, a £ dA are given. Note that in (i) the point a G A is the unique common interior fixed point of the semigroup, i.e., F t (a) = a,
t > 0.
(6.35)
However, this fixed point is not attractive, i.e., Ft(z) does not converge to a as t —> 00 for all z G A, z 7^ a. In cases (ii) and (iii), a G dA is a boundary locally uniformly attractive fixed point for {Ft}t>o, i.e., Ft(a)=a,
t>0,
and
lim Ft(z) = a,
(6.36)
t—»oo
uniformly on each subset strictly inside A. Observe, however, that in case (ii) there is another (boundary) fixed point (5 G dA of the semigroup {Ft}t>o, while in (iii), a is the unique fixed point in the whole plane. These phenomena will be considered later in the context of the asymptotic behavior of more general semigroups. It turns out that many important classes of linear semigroups are closely connected to corresponding classes of nonlinear semigroups. Moreover, if these connections are one-to-one, one may use different techniques to study both types. An example of such a semigroup is the linear semigroup of isometries with respect to the so-called indefinite metric on Krein spaces. For simplicity we consider this construction just for the Cartesian product of complex planes although this scheme works for any product of Hilbert spaces (and sometimes even Banach spaces) (see, for example, [Vesentini (1987a-b); (1991); (1992); (1994a-b)] and [Khatskevich et al. (1995a)]). Example 6.6 Let H be a complex Hilbert space with the inner product (-, •) and let X — H © C be the Hilbert space of the direct sum of H and C
164
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
with the inner product (•, •) denned by (6.37)
(x, y) = (u, v) + zw,
where u, v e H, z, w € C, and x = u + z, y = v + w are elements of X. Let a : X —> C be a continuous hermitian sesquilinear form on X defined by (6.38)
a(x, y) = {u, v) - zw.
This form is called the indefinite metric on X. The space X endowed with this metric is variously called either a Krein or a Pontryagin space. If J denotes a self-adjoint operator on X defined by the block matrix J =(o-°i)'
< 6 - 39 )
where / is the identity operator on H, then the indefinite metric a : X —> C can be written as (6.40)
a(x,y) = (Jx,y). A linear operator A : X —> X is said to be an a-isometry if
(6.41)
a(Ax, Ay) = a(x, y). In operator form this condition becomes
(6.42)
A'JA = J,
where A* is the adjoint operator of A. It turns out that a one-parameter semigroup of linear block operators on X, which are a-isometries, can be completely described by a (nonlinear) semigroup of fractional linear transformations on H, which preserve the infinitesimal Poincare metric on the open unit ball B of H (see [Vesentini (1987a); Vesentini (1991)]). For simplicity, we consider the case where H — C is the complex plane and B = A - the open unit disk in C. First observe that all groups of automorphisms of A of types (i)-(iii) described in Example 6.5 can be written in the general form Ft{z) = ei6<^^-, 1 - atz
H<1,
MR,
t€R,
(6.43)
Generation Theory for One-Parameter Semigroups
165
where the semigroup properties can be verified by the relations e i(fl, + .-»t-8.)( 1
at+°=l
+ a3ateif>°) = l +
ate~i6'+a3 + asate-«s>
a°
asate-i9°
= *° = °'
(6.44)
(See, for example, [Berkson et al. (1974)].) Let us define the families {a{t)}t>o and {b(t)}t>o by a{t) =
.
l
V 1 - l«t|2
• e'*
(6.45)
and b(t) = -ata(t).
(6.46)
M = ei9'
(6.47)
Then a(t) and
(6.48) Hence,
, _ a(t) ^ + S J
F^-W)^w-Z
a(t)z + b(t)
= W^WY
(6>49)
If we now define the family {At}t > 0 of linear operators on C © C (= C2) by the matrix At=\
fa(t) b(t)\ \b(t) a(t)J
,
6.50
then it is easy to verify that this family forms a one-parameter linear semigroup on C 2 . Moreover, clearly Al JAt = J,
(6.51)
i.e., the family {At}t>o is a semigroup of a-isometries on C 2 , where a : C2 —> C is defined by a(z,w) = ziiBi — z-iw-i, and z = (z\, z2) and w =
166
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
(wi,W2) are elements of C2. In addition, for each t > 0, we have deb At = \a(t)\2 - \b(t)\2 = ^—L-p - M 2 \a(t)\2 1 ~ l"tl 2 - I3^|2 = 1-
(6-52)
Now let A : C2 -» C2 be any matrix operator on I = C2, i4= (cd)«
(6-53)
preserving the a-metric on X and normalized by the condition det A = 1. It then follows by the condition A*JA = J
(6.54)
that c = b and d = a, i.e., A has the form
(i!)
^
with |a|2 - |6|2 = 1. If we define the fractional linear transformation fW then we have
= ]rr^ bz + a
™-it?h-
(6-56)
(6-57)
Since | | | = 1 and |£| < 1, we obtain that F is an automorphism of A. Another useful class of one-parameter semigroups of linear operators is the so-called one-parameter semigroup of composition operators defined on Hardy spaces. Example 6.7 Let F : A —> A be a holomorphic self-mapping on A. If 0 < p < oo, then the operator C : HP(A) -> HP(A) defined by CFf = foF
(6.58)
is a well defined linear operator on HP(A). It is called the composition operator on HP(A) (see, for example, [Shapiro (1993)] and [Cowen and
Generation Theory for One-Parameter Semigroups
167
MacCluer (1995)]). If S = {Ft}t>o is a semigroup of holomorphic selfmappings on A, then the family {At}t>o defined by Atf = foFt,
t>0,
(6.59)
defines a semigroup of linear operators on HP(A). In addition, if 5 = {Ft}t>o is a continuous semigroup, i.e., lim FAz) = z,
t-»o+
z G A,
(6.60)
then {At}t>o is a continuous semigroup on HP(A) in the sense of the strong topology in this space, i.e., lim Atf = /
t->o+
(6.61)
for all / G HP(A). Note that this convergence is not uniform on HP(A). The converse is also true: i.e., for each strongly continuous semigroup {At}t>o of composition operators the corresponding semigroup {Ft} is continuous on the unit disk A in the topology of compact convergence on A (see [Berkson and Porta (1978)]). Of course, a subject of interest is the question when At is an isometry with respect to the norm topology on HP(A). For p ^ 2 the answer was given by F. Forelli [Forelli (1964)]. Namely, each one-parameter semigroup, (actually, group) {At} of composition operators on HP(A), p ^ 2, p > 1, is given by
(Atf)(z) = eimt[Fl(z)]h(Ft(z)), teR, weR,
(6.62)
where Ft is an automorphism of A. On the other hand, for each 1 < p < oo, such an operator At is an isometry on HP(A).
6.2
Linear semigroups
Let D be a domain in a Banach space X. A semigroup S = {Ft}te[o,T) is linear if for each t G [0, T) the mapping Ft : D —> D is the restriction of a linear operator At acting on X. We will mostly be interested in the cases where D is either all of X or the open unit ball in X. So, let us first assume that D — X, and let S = {Bt}t>o be a continuous semigroup of linear bounded operators on X, i.e., (i) Bt+s = BtoBa,
M>0,
(ii) Bo = lim Bt = I ( lim Btx = x, x € X),
(6.63) (6.64)
168
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
where the limit in (ii) is taken with respect to the strong topology on X. Usually, such a semigroup is known as a Co-semigroup. It is easy to see by the semigroup property (i) that condition (ii) implies continuity at each point to > 0, i.e., (6.65)
lim Bt = Bt0.
t—•to
Again we understand the last equality as lim Btx = Btax
(6.66)
t—*to
for each x e X. For each t > 0, define At = (I - Bt)/t. Clearly, At belongs to the space of bounded linear operators on X, for each t > 0. Theorem 6.1 Bt)xj
The set fi =
[x I
G X
:
lim Atx (= V
t -.o+
L(X),
lim \ (I *
t _o+
exists \ in dense X.
Proof. Let X* denote (as usual) the dual of X, and let f(t) be the continuous scalar-valued function on E + = [0, oo) defined by (6.67)
f{t) = t(Btx),
where a; is a fixed element of X and t G X*. Now it follows by the mean value theorem that 1 ft+s f(t) = lim - / f(6)(W. s _0+ S Jt
(6.68)
Since I € X* is arbitrary, we see that for each t > 0 and each x G X, Btx=
1 ft+s lim - / B9xd0. s _o+ s Jt
(6.69)
Define now y=
[ B9xd6, Jo
xeX,
(6.70)
t>0.
Then 1
-[Is
1 fs
1
s Jo
s Jt
B3]y = - / Bexd9 - - /
ft+s
B0xd9.
(6.71)
Generation Theory for One-Parameter Semigroups
169
Letting here s —> 0 + we obtain from (6.69) that (6.72)
lim - (I - Bs)y = x - Btx
s—>0 •
S
exists. In other words, y € SI. Let now a; be an arbitrary element of X. Define xn=n
Jo
As we have seen, xn s fi for all n = 1,2,3 represent xn in the form xn=
(6.73)
Bgxd6.
On the other hand, we can
f Bitxde. Jo
(6.74)
Letting n —> oo, we see that xn —> x strongly. Thus the closure of Q equals X, as claimed. • Definition 6.3
The mapping A : il —> X defined by Ax = lim -(x-
(6.75)
Btx)
is called the infinitesimal generator of the (strongly) continuous semigroup S = {-Bt}t>o- Obviously, the mapping A : Q —> X is linear, but not necessarily bounded unless Q, is all of X. The main property of the infinitesimal generator (which actually is an explanation of the name) is that a given semigroup and its generator are related by the Cauchy problem with an arbitrary initial data in fi. Theorem 6.2 let
Let S = {Bt}t>0 be a strongly continuous semigroup and
il= {x€X
: Ax:=
lim+-(I-Bt)x
existsl.
(6.76)
Then for each i e ! l , Btx G Cl and — (Btx) + ABtx = 0.
(6.77)
Proof. First we note that by the semigroup property, for t and s > 0 the operators Bt and Bs commute. By definition, the right derivative of Bt at
170
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
t > 0 is given by d+ 1 1 — (Btx) = lim - (Bt+3x - Btx) = lim Bt- (B3x - x) at s-»o+ s s-»o+ s = lim - (Bs - I)Btx = -ABtx, s—+0~'~ S
xeQ.
(6.78)
Also d+ — Btx = Bt Ax,
x e Cl.
(6.79)
Thus ABtx = BtAx,
x£il,
t>0,
(6.80)
which means that fi is J5t-invariant for each t > 0. If t > 0, then to obtain the left derivative it suffices to verify that for 0 < s < t, lim - (Btx - Btsx) + BtAx\\ = 0. s—»0+
5
II
(6.81)
To do this, we first observe that it can be shown by using the uniform boundedness principle and the semigroup property that for each t > 0, there is a number 1 < M < oo and w e (-oo, oo)
(6.82)
||-Bt|| < Me"1.
(6.83)
such that
Now we have, for x G fi, \\- (Bt% - Bt-.sx) + BtAx\\ \\s II =
Bt-s(-(Bsx-x))+BtAx
= \\Bt-.([^ (Bsx -x) + (Bs - I)Ax])\\ < Meu(t-J» | | - {Bsx -x) + Ac| + ||(B,-/)i4i|||-»0,
s-^0 + .
(6.84)
171
Generation Theory for One-Parameter Semigroups
Hence, for t > 0 we obtain ^ (Btx) = -ABtx. dt Corollary 6.1
(6.85) n
Each semigroup generator generates a unique semigroup.
Examples of strongly continuous semigroups are given in examples 6.1, 6.2 and 6.6 of the previous section. Note, however, that the semigroup 5 = {Bt = e~tA}t>o defined in Example 6.1 satisfies the so-called property of uniform continuity. Definition 6.4 if
A semigroup S = {Bt}t>o is called uniformly continuous
(6.86) lim | | / - 5 t | | = 0 , t->o+ where the limit is taken with respect to the operator topology on X. As we have already seen, for each bounded linear operator A : X —> X the semigroup S = {-Bt}, defined by Bt = e~tA, is a uniformly continuous semigroup. In this case A=
lim+-(I-Bt)
(6.87)
is its infinitesimal generator. Moreover, we have
1 (e~tA) = -Ae~tA
(6.88)
at uniformly on X. As a matter of fact, each uniformly continuous semigroup S can be represented in the form S = {e~*A}t>o for some A £ L(X). To show this we just need to prove the following assertion. Theorem 6.3 Let S = {Bt}t>o be a uniformly continuous semigroup, and let A : £1 -* X be its infinitesimal generator. Then A £ L(X). Proof.
It follows by the continuity in the uniform operator topology that
lim \\l-\ t->o+ll
f BTdA =0. II
t Jo
(6.89) v
'
Hence there exists to > 0 such that
IIJ - j - I BTdr\\ < 1. II
to Jo
M
(6.90)
172
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
Now if we define AQ S L(X) by 1 Ao = -
fto
H> Jo
(6.91)
Brdr,
then AQ is invertible and AQ1 = (/ - (/ - Ao))~ . In addition, 1 1 f fto fto 1 - (I - Bt)B0 = — { BTdr - / Bt+Tdr \ t tot [Jo Jo J 1 (1 /"* 1 /' t+to 1 = -{BTdr-/ B T drL
(6.92)
Letting t —» 0 + , we see that the expression on the right-hand side converges to the limit ^ ( / - Bto) which belongs to L(X). Hence, .4£ 0 = ± (I- Bto) and 4 = i (/ - B^BQ1 E L(X), and we are done. • Despite the fact that the class of uniformly continuous semigroups is quite narrow, this class has a large variety of applications to various problems. In particular, each strongly continuous semigroup in a finite dimensional space is uniformly continuous and is generated by a matrix operator. Another useful example of a uniformly continuous semigroup can be given by the integro-differential equation % + f k(x,y)f(t,y)dy = O, (6.93) ot JD where D is a domain in R" and k(x, y) is a L<2.(D x D) kernel. If we define a linear operator A : L<2(D) —> L?,{D) by (Ag)(x) = f k(x,y)g(y)dy, JD
x G R",
(6.94)
we have that the solution of the Cauchy problem
| U + |^ ( I , vmm.o 1/(0, *) = «(*)•
(695)
is given by the exponent f(t,x) = e~tAx.
(6.96)
Hence we obtan a uniformly continuous semigroup on X = L2(D). An important question in the general semigroup theory is when a given semigroup leaves a given domain D invariant.
Generation Theory for One-Parameter Semigroups
173
For linear semigroups this question mostly arises when D is the open unit ball in X. In this case the semigroup is said to be a semigroup of contractions on X. In other words, the problem is to find the properties of the infinitesimal generator A of a semigroup {Bt}t>o with ||Bt|| < 1
(6.97)
for all t > 0. Basic results in this connection are given in the famous Hille-Yosida and Lumer-Phillips theorems. Although these theorems are known for strongly continuous semigroup in locally convex spaces, it is enough for our further purposes to formulate them for uniformly continuous semigroups in Banach spaces. Definition 6.5 Let X be a Banach space and let A G L{X). The set p(A) = {X e C : (XI - A)'1 G L(X)} is called the resolvent set of the operator A. If p(A) is nonempty, then the operator-valued function R = (XI — A)"1 is called the resolvent of the operator A. Now we formulate a version of the Hille-Yosida theorem (see [Yosida (1974)]) for a uniformly continuous semigroup of contractions on X. Theorem 6.4 Let A € L(X). Then A is an (infinitesimal) generator of a semigroup of proper contractions S = {Bt}, \\Bt\\ < 1, if and only if (i) p(A)D(-oo,0);
(ii) XR(X, —A) = (I + j Aj
is a contraction for all X > 0.
Moreover, the following exponential formula holds:
Bt = e~tA= lim | 7 / + - A V 1 1 .
(6.98)
Remark 6.2 It can be shown that, actually, the spectrum a(A) of the operator A lies in the closed right-half plane. This, of course, implies condition (i). Definition 6.6 A linear operator A c L(X) is said to be accretive (respectively, dissipative) if for each x € X, there is x* £ J(x) such that Re(Ax,x*)>0
(6.99)
Re«x,a:')<0).
(6.100)
(respectively,
174
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
A very useful characterization of accretive (respectively, dissipative) operators can be given as follows. Theorem 6.5
A linear operator A G L(X) is accretive if and only if ||(A/ + A)a:||>A||i||
(6.101)
\\{XI-A)x\\>X\\x\\)
(6.102)
(respectively,
for all x £ X and A > 0. This result and the Hille-Yoside theorem imply the following important assertion known as the Lumer-Phillips theorem (see [Yosida (1974)]). Theorem 6.6 Let A G L(X). If A is accretive and for some Ao G R j = (0,oo), (AoJ + A)'1 G L(X), then A is the generator of a one-parameter semigroup of proper contractions on X, i.e, \\e-tA\\ < 1,
t>0.
(6.103)
t > 0,
(6.104)
Conversely, if for some A G L(X) \\e-tA\\ < 1,
then A is accretive (or - which is the same - A is dissipative). Definition 6.7 hermitian) if
An operator A G L(X) is said to be conservative (or Re{Ax,x*) = 0
(6.105)
for all x G X and x* G X*. A conservative operator A is simultaneously accretive and dissipative. The following assertion holds. Theorem 6.7 A linear operator A G L(X) is conservative if and only if its spectrum lies on the imaginary axis of the complex plane and it is a generator of a group {Bt = e~tA : t G (—00,00)} of linear isometrics, i.e., \\e~tAx\\ = \\x\\ for alltGR
andx£X.
(6.106)
175
Generation Theory for One-Parameter Semigroups
6.3
Generated Semigroups of Nonexpansive and Holomorphic Mappings
Now let D be a domain (open, connected subset) in a Banach space X with the topology induced by the norm of X. Definition 6.8 A semigroup S = {Ft : t e (0,T)}, T > 0, on D is said to be generated if for each i € D , there exists the strong limit f(x)=\im+j{x-Ft(x)).
(6.107)
In this case the mapping / : D —> X is called the (infinitesimal) generator of S. One of the most important problems in the context of semigroup theory is to find out when a continuous semigroup of nonexpansive mappings in the norm or metric sense (in particular, />-nonexpansive or holomorphic mappings) has a generator. To trace an analogy with the classical linear case, we note that each semigroup of bounded linear operators which is continuous in the operator topology is differentiable at zero, and its generator is also a bounded linear operator. And conversely, if / is a linear holomorphic mapping, then it is bounded by definition, and we obtain the simplest case: the semigroup 5 generated by / is a uniformly continuous linear semigroup Ft = e~ff. In addition, we have the exponential formula e~tf = lim (/ + tf/n)~n. n—>oo
(6.108)
For the nonlinear case the analogous facts are not trivial. For nonlinear semigroups of norm nonexpansive mappings this problem has been considered by many mathematicians, mostly in the framework of the study of the asymptotic behavior of semigroups as well as the structure of their stationary point sets. For semigroups of holomorphic mappings it was also studied in connection with the theory of branching processes. In the onedimensional case, the differentiability with respect to the parameter of nonlinear semigroups of holomorphic mappings was proved by E. Berkson and H. Porta [Berkson and Porta (1978)] in their study of linear semigroups of composition operators on Hardy spaces. This nice result was extended by M. Abate [Abate (1992)] to the case of C n . However, it is no longer true in the infinite dimensional case. E. Vesentini has investigated semigroups of those fractional linear transformations which are isometries with respect to
176
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
the infinitesimal hyperbolic metric on the unit ball of a Banach space. He used this approach to study several important problems in the theory of linear operators on indefinite metric spaces. Observe that in general, such semigroups are not everywhere differentiable. As a matter of fact, it turns out that a continuous one-parameter semigroup of holomorphic self-mappings of a domain in a Banach space is differentiable with respect to the parameter if and only if it is right locally uniformly continuous with respect to the parameter at zero. Let us recall that a net {fj}jeA C Hol(D,X) is said to converge to a mapping f 6 Ho\(D,X) in the topology of locally uniform convergence over D (or, briefly, T-converge) if for every ball B CC D (B is strictly inside D), limsup \\fj(x) - f(x)\\ = 0.
(6.109)
jeAxGB
We write in this case / = T— lim /,•. For the finite dimensional case jeA
this topology coincides with the compact open topology on D. Definition 6.9 A family S = {Ft}t>o C HOI(JD) is said to be a locally uniformly continuous one-parameter semigroup (or, briefly, a T-continuous semigroup) if it satisfies the semigroup property s,t>0, (i) F.+t = FtoFt, and (ii) T - lim Ft = I\Dt-»o+
Theorem 6.8 ([Reich and Shoikhet (1998b)]) Let D be a bounded domain in X, and let S = {Ft}t>o C Hol(D) be a strongly continuous semigroup. The following conditions are equivalent: (a) S is a T-continuous semigroup; (b) The differences ft = -AI- Ft) c are uniformly bounded on each subset strictly inside D; (c) For each x G D, there exists the strong limit hm+-t(I-Ft)(x)
= f(x),
which is bounded on each subset strictly inside D.
(6.110)
(6.111)
Generation Theory for One-Parameter Semigroups
Remark 6.3
177
It is remarkable that actually f = T-hm\{I-Ft),
(6.112)
i.e., the convergence in (c) is actually locally uniform convergence over D (see [Reich and Shoikhet (1998a)]). It is clear that if (c) holds, then the infinitesimal generator f belongs to Ho\(D,X). Since for the finite dimensional case a continuous semigroup of holomorphic self-mappings is T-continuous, it follows that such a semigroup always has a holomorphic infinitesimal generator (see also [Abate (1992)]). We now prove the above assertions in a more general setting. Definition 6.10 Let X be an arbitrary Banach space and let D be a in X. We say that a family [G3 : s e (0, T)}, T > 0, of self-mappings of domain D satisfies the approximate semigroup property if for each subset D strictly inside D the following conditions hold: (i) for each e > 0, there is a positive S = S(D,s) < T such that sup\\Gs(x) - Gp3/p\\ < es
x€D
(6.113)
for all positive integers p and all s 6 (0,6); (ii) for each pair s,t £ (0,T), s + t
(6.114)
Theorem 6.9 Let D be a domain in a complex Banach space X, and let {G s : s £ (0, T)} be a family of holomorphic self-mapping of D which satisfies the approximate semigroup property. Suppose that Gs converges to the identity as s —> 0 + , uniformly on each subset D strictly inside D, i.e., limsup ||G,(a;) - x\\ = 0.
(6.115)
s->0+ x€D
Then the strong limit lim - (/ - Gs) = f
(6.116)
s—»0+ S
exists and is a holomorphic mapping from D into X, which is bounded on each subset strictly inside D.
178
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
To prove our theorem we need two lemmas. Lemma 6.1 Let Dfeea domain in a complex Banach space and let <j) € Hol(.D). Suppose that for some subset Di c D withdist(Di,dD) > 0 there are two numbers /i and d, 0 < n < d, an integer p > 1, and a domain D2, DxCD2cD with dist(Di,dD2) > d such that sup \\x-<j>k(x)\\ < / i
(6.117)
for all k = 0,1,2,..., p— 1. Then for x S D\ the following inequality holds: \\x - 4?{x) - p{x - 4>(x))\\ < -/— (p - l)\\x - # c ) | | .
(6.118)
Proof. Let x € D\ and z G D2 be such that \\z — x\\ < fi. Then the ball Bd-n(z) with its center at z and radius d — fi lies in D2. Hence it follows from (6.117) and the Cauchy inequality that ||(J-0*)'(*)||<-Ji-.
(6.119)
Qi — fJi
Therefore, for x G D\ and y G D such that ||a; — y|| < \i we have, by (6.119), ||z - 4>k{x) ~{y- $k(y))\\ < - / - \\x - 2/H-
(6.120)
CL iX
Now setting y = (f>(x) and using (6.117) and (6.120) we obtain by the triangle inequality NP-1
||z - r(x) - P{x - 0(a:))|| = Y, [ ^ t o -
<^2\\4>k{x)-x-[4>k(
<-^-(p-l)\\x-
(6.121) •
and we are done.
Lemma 6.2 Let D be a domain in a complex Banach space and let a family {Gs : 0 < s < T}, Ga € Hol(D), satisfy the approximate semigroup property. Suppose that G3 converges to the identity as s —> 0 + , uniformly on each subset strictly inside D. Then for s > 0 small enough the net fB = -AI-G.) 5
(6.122)
179
Generation Theory for One-Parameter Semigroups
is uniformly bounded on each subset strictly inside D. Proof. Let D\ be a subset strictly inside D and let 0 < d < dist(dD, Di). Take any domain D2 CC D such that D\ CC D2 and dist(Di,dD2) > d. Choose fi, 0 < (i < d, such that (i(d — /z)" 1 < ^, and choose a, 0 < a < T, such that for all T € (0,
(6.123)
l
In addition, it follows by the approximate semigroup property (i) that there exists 0 < 8 < | such that \\Gk3(x)-Gsk(x)\\<%
(6.124)
Now set n = [^]. For s £ (0, S), we for all x £ £>2 and each k = 1,2, have n > 2, ns > | , and ks < a for all k = 1,2,..., h. Hence it follows from (6.123) and (6.124) that sup \\x - Gks(x)\\ <»,
x€D2
se {0,6).
(6.125)
Now for all x € D\ we get, by Lemma 6.1, n\\x - G.(x)\\ - ||z - Gna(x)\\ < \\n(x - Gs(x)) - (x - G?(z))|| <^n||a:-G.(x)||,
(6.126)
or
\\x-G3(x)\\
(6.127)
Therefore, by (6.122)-(6.127), we obtain
||/x(*)|| < ^ (IN - G ns (x)|| + \\Gns(x) - GJ(x)||)
(6.128) whenever s G (0,6). The lemma is proved.
•
Proof of Theorem 6.9. Let D be a domain in a complex Banach space X, and let {Gs : s £ (0,T)} be a family of holomorphic self-mappings of D which satisfies the approximate semigroup property. Suppose that G3
180
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
converges to the identity uniformly on each subset strictly inside D. To show that the net / . = - ( / - G.)
(6.129)
is a Cauchy net, as s —> 0 + , on each subset D\ strictly inside D, assume that e > 0 has been given. Choose 0 < d < dist(Di,dD), 0 < // < d, such that -7^— < e,
(6.130)
d — ^jj
and choose 0 < w < T such that sup ||i-G T (a;)|| <%
(6.131)
for all r e (0,w). Let 0 < S = (5(£>!,e) < w b e such that condition (i) of the definition is satisfied, i.e., \\GT(x) ii
GPT(X)\\ ii
p
(6.132)
< ET
whenever x e D\ and r e (0,6), p= 1,2, Now choose an integer N > 0 such that AT"1 < 6 and e • AT"1 < 5 /i. Then, for all integers m,n> N and allfc= 0,l,2,...,p = max{m,n}, we have, by (6.131) and (6.132), sup \\x - Gk_L. (x) I < sup \\G\_ (x) - G_k_ (x)\\
X6-D1
x€Di
m n
m
"
mn
+ J S " X - G ^ ( i B ) l l < e ^ + 2<'Z-
(6133)
Therefore, by (i) and Lemma 6.1, setting in this lemma (j> = GJL. and nm
p = m we get for all x G D\,
\\x - Gi (x) - mix - G_i_ (x)) II < |x - G^_ (a:) - mix - G j _ (x)) II + ||G!!L(a;)-Gj.(x)|| < T ^ - m l | x - G ^ ( x ) | | + e - - . 11
nra
n
"
d — (1
'
n m
(6.134)
Tl
Multiplying this inequality by n and using (6.129) and (6.130) we obtain, for x £ D\, ||/i(x)-/s!_(a:)||<e(||/_j_(»)|| + l ) .
(6.135)
Generation Theory for One-Parameter Semigroups
181
Now it follows, by Lemma 6.2, that there is C — C(Di) such that ||/^(x)||<£
(6.136)
for all x € D\, whenever N (and therefore n • m) is big enough. by (6.131) we have \\fi(x)-f^.(x)\\<e(C + l).
So,
(6.137)
In a similar way we can get \\fi(x)-f^.(x)\\<e(C + l)
(6.138)
for all x in Dx and n,m> N, and hence \\f±(x)-fx(x)\\<2e(C+l)
(6.139)
for all x € D\ whenever n,m> N. This inequality means that the sequence {/A }n-jv c o n v e r ges as n —> oo uniformly on each subset D\ strictly inside D. In particular, it converges uniformly on each ball strictly inside D and is uniformly bounded on such a ball. Therefore, its limit /=
lim fx
n—»oo «
(6.140)
is a holomorphic mapping from D into X. Now we show that the net {/s}s€(o,T) converges to / uniformly on each subset D\ strictly inside D. This will conclude the proof of our theorem. For given e > 0 and x € D\, setting n = [^-], we can choose s so small that ||/jL(aO-/(z)||<e.
(6.141)
In addition, for such s and n we have fs-fi.
= -s(I-Gs)-n(l-Gx)
+ H( j - G l r 1 )- n *( j - G i)]-
<6-142)
Observe that in our setting n = [^r], so that we have ns —» oo and J2|l -> 1 as s -» 0. Thus we can find 8 > 0 such that 1 - &£ < e and Gjsni (x) e £>2 CC D whenever s G (0,6) and a; € Di. Using the n
182
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
approximate semigroup property (ii) (see the Definition) we get for such s and all x G £>i, - \\G^ (x) - G.(x)|| < - llGiad (x) - G on
S
+ p^oG
o G,_^ (x)|| n
i£n1{x)-G3(x)\\
n
n
<-\M\\x-G S [
H n
TI
"
S~
^11
n
M
(6.143) where M =
sup
xeZ?2,s6(0,i)
||(Gs)'(x)||. Once again, using Lemma 6.2, we have
\\x-G.-M\\
(6-144)
and therefore (6.143) implies that - ||
n
(6.145)
II
Now condition (i) of the definition implies i \\G[?n](x)\\<e[-^<e. (6.146) s II n II sn Finally, by Lemmas 6.1 and 6.2, we obtain for x G Di and s G (0,5),
- ||x - G[tn] (x) -ns(xS II
n
V
Gx (x)) II "
/ II
< - \\x - G[inl(a:) - [ns](x - Gi(x)) II + - \[ns] - ns\ \\x - Gi (x)||
< fej H + 1 | H - ns\) \\x - Gx (x)\\
< ^ H + H _ 1 \ c < 2£e.
(6.147)
Thus for \ xns6 £>i and ns s € (0, / J) we get from (6.141)-(6.147) ||/a(x) - /(x)|| <
||/A(X)
- /(x)|| + ||/a(x) - /i(x)||
<e{2 + MC + L + 2C),
(6.148)
Generation Theory for One-Parameter Semigroups
and we are done.
183
•
Proof of Theorem 6.8. It follows by Theorem 6.2 that condition (a) implies condition (c). By Lemma 6.2, conditions (a) and (b) are equivalent. The implication (c)=^(b) is obvious. • 6.4
The Cauchy Problem and the Product Formula
We know already that a T-continuous semigroup of holomorphic mappings is right-differentiable with respect to the parameter t at zero. As a matter of fact, this implies, in turn, that Ft is differentiable at each point t € M+ and that the function u(t, x) = Ft(x) is the solution of the Cauchy problem u't(t,x) = -f(u(t,x)), u(0,x) = x [Cartan (1967)]. In the context of the Hille-Yosida theory the following question is also of interest. If in the infinite dimensional case we have a family of holomorphic mappings which satisfies the approximate semigroup property and converges to the identity uniformly on each subset strictly inside D, is this family differentiable with respect to the parameter and does its derivative generate a semigroup which may be represented by the so-called product or exponential formula? To answer these questions we first formulate the following general assertion. Theorem 6.10 Let (D,p) be a complete metric space, and let {Gs : 0 < s < T} be a family of p-nonexpnsive mappings on (D,p) with the following properties: (i) For each p-ball B C (D, p) and each e > 0, there is a positive 6 = 6{B, e)
G|(:r)) < e - s
(6.149)
for all x € B and for all integers p whenever s G (0, <5); (ii) For each p-ball B C (D,p) there exist /i = /x(B) > 0 and C = C(B) such that p(Gs{x),x)
(6.150)
for all x e B, whenever s G (0, /i). Then for each pair s € (0, T) and t > 0 and each sequence of integers {£„}
184
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
such that - ^ - > 1 , as n-»oo,
(6.151)
Tit
there exists the limit lim G*£ = Ft
n—*oo
(6.152)
n
uniformly on each p-ball in (D,p). This limit does not depend on the sequence {tn} and it is a locally uniformly continuous one-parameter semigroup with respect to t > 0. Proof. First we establish two simple inequalities. For each T 6 (0,T) and each integer £, we have e-i
p(GeT(x),x) < J2p{Gl+1(x),
Gl(x)) < tp(GT(x),x).
(6.153)
3=0
Now if £i and £2 are two arbitrary integers, (6.153) implies p{&?{x), G^{x)) = pi^a^M{G^-t2\x)),
G™in^2)(z))
< p(G^-^{x),x) < \h-e2\p(GT(x),x) (6.154) for each r G (0,T). Take a p-ball B C (D, p) and choose /i > 0 so that condition (ii) holds. Then, for each r G (0,/x) we have , by (6.153) and (6.154), p(Gir(x),x)<£-CT
(6.155)
p(GeTHx), G?(x)) < \h-l2\CT
(6.156)
and
for all x G B and for all integers l,l\,ti. For a given s G (0, T) and t > 0, consider the sequence of mappings NOO Gj 1 > on 5 , where {i n } is a sequence of integers which satisfies (6.151).
{
n J 1
Taking an integer N so that s/N < n we get, by (6.155), p (G'jUx), X) < ^
L < 00
(6.157)
for all n > iV and all x & B. In addition, for each j = 1,2,... ,tn and m = 1,2,..., pfG'7(a;),xN) <mj-~
C < 00
(6.158)
Generation Theory for One-Parameter Semigroups
185
whenever n> N This means that there exists a p-ball B\ c (D, p) such that the sequences (G*r(z)| In
and \G"J!_(Z)\
J jV
I
nm
are in Bi for all x e B, j = J JV
1,2,..., tn, m = 1,2, Now for a given e > 0 we can choose, by (i), 6 = 8(e, Bi)
(6.159)
for all z £ Bi and all m = 1,2,..., whenever 0 < r < 8. Taking N so large that s/N < min{/i, 8} and setting z = GJ_7. (a:), x £ B, m = 1,2,..., j = 1,2,..., tn, n > N and T = ^ , we obtain, by the triangle inequality, the nonexpansiveness of Gs and inequality (6.159), that
p((ft{x), G^ix)) < ^(C'r^fef^)), G^-UG^^ix)))
GT^HG^^ix)))
< t n • e • - < a • e, (6.160) n where a = sup {^f2} < oo because of (6.151). In the same way and for the same e > 0 we obtain the inequality P ( G V {X), G V " {X)) <e-a \
(6.161)
/
nm.
m
for all x € B wheneverTO> AT, n = 1,2, by (6.156) that
In addition, it follows
p ( G V m (X), Gtjr (x)) < ^2- - ^=2- £ < e • a \
nm
nm
'/
n
nTO
(6.162)
and ,o fG*1?" (a;), G V (a)) < ^ 2 . - ^ ^ L < e • a
(6.163)
186
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
for n,m> N, where N is large enough. Thus, by the triangle inequality, (6.160)-(6.163) imply that for a given e there is N > 0 such that p (G*? (X), G*r (x)) < 4ae
(6.164)
whenever n,m > N. This means that < Glr i is a Cauchy sequence uniformly on each p-ball B C (D,p), and since {D,p) is complete, its limit exists and is a /^-nonexpansive mapping on (D,p). Once again it follows from (6.156) that it {r n } is another sequence of integers such that — -»t
(6.165)
then, for a given e > 0 and x 6 B, p (G*? (X), GrP (x)) < — - — £ < e \
"
n
/
n
Tl
(6.166)
whenever n is large enough. This means that Ft = lim G'l1
(6.167)
does not depend on the sequence {tn} satisfying (6.151). Now let s e (0,T), t > 0 and r > 0 be given numbers, let {tn}i° and {rn}T be two sequences of integers such that ^ p —» t and ^ ^ - » r a s n —> oo. Then, for a given e > 0 and x € B, p(Ft(*V(x)), F t + r (x)) < p(F t (F r (x)), Gi" +r "(x)) +p(Fi" +I -"(x), T t + r (x))
<^(Ft(Fr(x)), ^ - ( G t ^ J + e < p(Ft(Fr{x)),
G'HFrix)))
+p(Gt£(Fr(x)),
G*j(Gt(x)))+e
< p(F r (x), G r /(x)) + 2e < 3e
(6.168)
whenever n is big enough. Since e > 0 is arbitrary, we have F t + r = Ft • F r .
(6.169)
187
Generation Theory for One-Parameter Semigroups
Thus, Fr : R+ —> (D, p) is a one-parameter semigroup which is uniformly continuous on each p-ball in {D,p) with respect to t > 0. The theorem is proved. • A consequence of this theorem is the following important assertion. Theorem 6.11 Let D be a metric domain in X with a metric p G (SPS) and let {Gs : s G (0,T)} be a family of holomorphic self-mapping ofD which satisfies the approximate semigroup property. Suppose that Gs converges to the identity as s —> 0 + , uniformly on each subset D strictly inside D, i.e., limsup ||Gs(x) - z|| = 0.
(6.170)
s-»0+ x€D
Then, for each pair s and t, s G (0, T), t > 0, and each sequence of integers {£„} such that — ->1 as n->oo, nt there exists the strong limit
(6.171)
lim G'a" = Ft, n—»oo
(6.172)
n
uniformly on each subset strictly inside D. This limit does not depend on oo} is a one-parameter {tn} and s in (6.171), and the family {Ft :0
f=
lim I (J _,).
(6.173)
s-»0+ S
For x G D the mapping u(t,x) — Ft(x) defined by (6.172) is the solution of the Cauchy problem
I
9t
;
(6 . 174)
lim u(t, x) = x
K. t->o+
where f is defined by (6.173).
Proof. Let D be a metric domain in X with some metric p G (SPS). Then it follows that conditions (i) and (ii) of the above theorem are satisfied. Therefore, by this theorem, for each pair s and t, s G (0,T), t > 0, and each sequence of integers {£„} such that ^ —> 1, there exists the strong limit Ft — limn-Kx, G*? uniformly on each subset strictly inside D. This
188
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
limit, F t : R+ —> Hol(D), is a one-parameter semigroup of holomorphic self-mappings of D. Now we already know that the limit
(6.175) exists and is a holomorphic mapping on D. We want to show that this mapping / : D —» X generates the semigroup {Ft}, i.e., that it also satisfies the condition / = thm
T-^
,
(6.176)
where the convergence in (6.176) is uniform on each Di CC D. Indeed, for given e > 0 and D\ CC D, we can find a small enough 5 > 0 such that
|/(s)_£z£(E)|<e
(6.177)
for all x € D\ and all t £ (0,6). In addition, setting s = t and tn = n we can find <5i < <5 such that
j||GT(x)-Ft(x)||
(6.178)
(see (6.160)) for all t £ (0,5). Once again, using (6.175), we take S2 < S such that - \\Gt(x) - G\ (x)|| < s for 0 < t < S2.
(6.179)
Thus we get for t e (0,J 2 ),
|| / W _£^M|<| / ( I ) _£Z|*)| + i ||C, - 0 1 (x)|| + -t ||G1 (i) - F,(x)|| < 3e,
(6.180)
and we are done. Finally, it follows by the semigroup property and (6.176) that for each x e D the mapping u(t,x) = Ft(x) is a solution of the Cauchy prob• lem (6.174). This concludes the proof of the theorem. Corollary 6.2 Let D be a metric domain in X with a metric p G (SPS) and let {Ft : t € (0,T)}, T > 0, be a one-parameter semigroup of holomor-
Generation Theory for One-Parameter Semigroups
189
phic self-mappings of D such that lim Ft = / ,
(6.181)
t-»o+
uniformly on each subset strictly inside D. Then this semigroup can be continuously extended to a flow {Ft : 0 < t < oo} on all of M + . Remark 6.4 Formula (6.172) is called the product formula. The crucial point in establishing this formula is that the family {Gs : s € (0, T)} of holomorphic mappings in D satisfies the approximate semigroup property and has a right-hand derivative at s = 0 which is equal to f. The question is what happens when we have an arbitrary continuous family {G s } s >o C Hol(D) which is differentiate at s = 0 + . Actually, we will see below that for a bounded convex domain, the above theorem and the so-called resolvent method and range condition imply a somewhat more general assertion. Theorem 6.12 Let D be a bounded convex domain in a complex Banach be an arbitrary family of holomorphic selfspace X, and let {G3}3e^,T) mappings of D such that lim
s->0+
X~G^X)
(6.182)
= f{x)
S
exists uniformly on each subset strictly inside D and is bounded on such subsets. Then (1) the Cauchy problem (6.174) has a global solution u(-, •) defined on R+ x D; (2) this solution can be obtained by the following product formula: u(t,-)=
lim Gl,
n-»oo
n
(6.183)
where the limit is uniform on each subset strictly inside D. An immediate but very important consequence of this theorem is the following assertion. Corollary 6.3 Let D be as in Theorem 6.10, and let f and g be two holomorphic generators of one-parameter semigroups on D, i.e., {Ft}t>o and {Gt}t>o, respectively. Then the mapping h = / + g is also a generator and the semigroup Ht generated by it can be obtained by the formula Ht=
lim [Ft. -G±]n,
(6.184)
190
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
where the limit is uniform on each subset strictly inside D. Corollary 6.4 The set of holomorphic generators on a bounded convex domain is a real cone. 6.5
Nonlinear Resolvents, the Range Condition and Exponential Formulas
Definition 6.11 We will say that a mapping / : D —> X satisfies the range condition if there exists a positive T > 0 such that for each s € (0,T), (I + sf)(D) D D
(6.185)
and Ts = (I + s / ) " 1 is a well-defined self-mapping of D. This mapping Ts is called the (nonlinear) resolvent of / . It turns out that similarly as in the linear Hille-Yosida theory, the resolvent method plays an important role in the study of nonlinear generators of semigroups, as well as in the study of the structure of the fixed point sets of self-mappings and null point sets of vector fields. In this context it is natural to look for geometrical conditions which will ensure that any semigroup of holomorphic mappings can be represented by exponential formulas or, in other words, to find out when the range condition holds for each holomorphic generator. To answer this query we need the following formula, which is usually called the resolvent identity. Lemma 6.3 Let D be a convex domain in a Banach space X, and let f : D —» X be a mapping which satisfies the range condition. Then for 0 < s
(6.186)
Proof. For each x 6 D the element y = f z - f (l- j)Tt(x) belongs to D, by the convexity of D. It follows by the definition of the resolvent that I-Tt = tf(Tt) for t 6 [0,T). Thus,
y = Tt(x) + -t(x- Tt{x)) = Tt(x) + sf(Tt(x)) = (I + sf)Tt(x)l6.187) Hence Ts{y) = (/ + sf)~l{y) = Tt(x), and we are done.
•
191
Generation Theory for One-Parameter Semigroups
Lemma 6.4 Let D be a convex metric domain with a metric p G (SPS). If f G Hol(D, X) is bounded on each subset strictly inside D and satisfies the range condition, then the family {Ts = (I+sf)~l : s £ (0,T)} converges to the identity, uniformly on each subset strictly inside D and satisfies the approximate semigroup property. Proof. Indeed, let D be a metric domain with a metric p £ (SPS), and let D\ be a subset of D such that dist(£>i, dD) > 0. Denote Mi = sup{||/(j/)|| : y 6 £>i} and 6 = min { ^ , T } , where 0 < d < dist(D\,dD). Setting y = x + sf(x) for x £ D and s £ (0, J), we have Ts(y) = cc, ||a;-1/11
< LMxs. (6.188)
Since for each T G (0, T) and each integer £,
p{T?{x),x) < "£P(TJ+I(x),
(6.189)
Tj{x)) < £p(TT(x),x),
3=0
we have that for x € D, n = 1,2,..., andfc= 1,2,..., n, | | 7 | (a:) - x \ \ < ± p ( T t ( x ) , x ) < C - s < ± ,
(6.190)
whenever 0 < s < 5^ = min{<5, ^ } . Firstly, (6.188) and (6.190) mean that the subset D2 = Di (J
Bd(x),
x€Di
which lies strictly inside D, where Bd(x) is the closed ball with its center at x and radius d, contains the sets {Ts(x)}^(o,^) and {Tt}, s G {0,6), n = 1,2,..., and k — l,...,n, where x € D\. Secondly, it follows from (6.188) that the net {7^}s€(o,<$) converges to the identity uniformly on D\. Now denote Mi = sup {||/(x)||}. It follows from the Cauchy inequalities that for each x € D\ and y £ D such that \\x ~ v\\ < f ) ll/'(y)ll < 2M/d, and hence, for such i and y we have
ll/W-/(2/)ll<^alk-2/||-
(6.191)
Now, because of the identity x - T.(x) = s/(T s (x)), x G D,
(6.192)
192
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
we have that /(*) = lim X~T^X\ s—>0+
(6.193)
S
uniformly on D\. In addition, (6.190)-(6.192) imply that
\\sf(x)-x + T2(x)\\ < ElJ/W + r, (rt\x)) - (rl-\x))\\
<£.^-s2,i6D.
(6.194)
Thus we obtain, by (6.188), (6.192) and (6.194), \\%(x) - Tf || < \\x - Tf (i) - s/(x)|| + ||s/(i) -x + T.{x)\\ < C2-^ s> + s\\f(x) - f(Ts(x))\\
f o r a l l x e D i , s € (0,<Ji), n = 1,2,{6.195)
This inequality shows that condition (i) of the definition of the approximate semigroup property is satisfied. Now take positive s,t such that s + t < Si. Then it follows by the resolvent identity and (6.188) that
\\Ts+t(x)-Tt(x)\\ < ip(r. + t (x), Tt(x))
^hp(Tt(^-tx
+
7T-tTt+s{x))'Tt{x))
^ip{jhx+7rtTt+°{x)>x) <^\\X-Ts+t{x)\\^-t<^.C.s.
(6.196)
Generation Theory for One-Parameter Semigroups
193
Thus we have
\\Ts+t(x)-Z(Tt(x))\\
< ^p{Ts+t(x), Tit
T,Tt(x))
\C+-
~ m s+t L
m
c]
2
(6.197)
This proves condition (ii) of the definition of the approximate semigroup property, and we are done. • Thus we have proved the main theorem of this section. Theorem 6.13 Let D be a convex metric domain in a complex Banach space X with a metric p € (SPS), and let f £ Ho\(D,X) be bounded on each subset strictly inside D and satisfy the range condition. Then f is the infinitesimal generator of the one-parameter semigroup of holomorphic self-mappings of D, which can be defined by the following analogs of the exponential formula:
Ft= lim ( / + - / ) " " n—>oo \
n
(6.198)
/
or /
1
\ [-**»]
Ft = lim ( / + - / )
,
(6.199)
where the convergence in (6.198) and (6.199) is uniform on each subset strictly inside D. As a matter of fact, for bounded convex domains the converse also holds. We prove this assertion in a more general setting. Let D be a metric convex domain in a Banach space X with a corresponding metric p on D. We recall that the metric p is compatible with the convex structure of D if the following conditions hold:
(i)
p(sx + (1 - s)y, sw+ (1 - s)z) < max[p(x, to), p(y,z)];
(6.200)
194
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
(ii) there is a real function
(6.201)
and for each three elements x,y,z £ D and s G [0,1], the following inequality holds: p(sx + (1 - s)z, sy + (l-
s)z) <
(6.202)
Of course, each convex bounded domain in a complex Banach space is a compatible metric domain with the hyperbolic metric p on D. Theorem 6.14 Let D be a metric convex domain in a Banach space X with a corresponding metric p on D, and let the metric p be compatible with the convex structure of D. Suppose that {Ft : 0 < t < T} is a family of p-nonexpansive self-mappings of D, i.e., (6.203)
p(Ft(x),Ft(y))
L=A
=:
/
(6.204)
exists uniformly on each p-ball in D and / : £ ) — » X is a continuous mapping, then f satisfies the range condition. Moreover, the resolvent Ts — ( / + sf)~x : D —> D is defined for each s > 0 and is a p-nonexpansive selfmappings of D.
Proof. Suppose that the conditions of the theorem are satisfied. For a fixed z € D and s, t > 0, we consider the equation x = - ^ - Ft(x) + -^-z S ~y L
S "|
t
(6.205)
= G.,t(x).
Since p is compatible with the convex structure of D we have, by definition, that
p(G.,t{x), G.,t{y)) < K l T i ) ^ ^ '
-(p(7Ti)p(x'y)'
Ft(j/))
(6"206)
Generation Theory for One-Parameter Semigroups
where 0 < f{^+i)
195
< 1> a n d limsup
T
r
< oo.
(6.207)
So, by the Banach fixed point theorem, it follows that equation (6.205) has a unique solution in D, x = xS}t for each s, t > 0. Since the equation (6.208)
x + sft(x) = z, z€D,
where ft = \ {I-Ft) is equivalent to (6.205), this implies that the mapping TStt = (I + sft)*1 is a well-defined self-mapping of D and T3
(6.209) where n = 1,2,..., and x° t (z) = y is an arbitrary element of D. Thus it follows by induction that
(6.210)
because Ft is a p-nonexpansive mapping on D. Hence Ts,t : D —> D is also /)-nonexpansive on D. Now we want to show that the net {TStt)s>o converges to a mapping 7^ : £> —> D, as t —» 0 + . If this holds it is clear that Ts = (I + s/)" 1 is a /9-nonexpansive mapping of D, and this will conclude our proof. First we show that for each z G D and each s > 0, the net {TStt(z)}(= {xs,t}) is strictly inside D for £ small enough. Indeed,
p{T.A*),z)=p(7±rtFt{x.,t) +
-KlTi)^:**
7L-z,z)
i) + piFt(z), z)].
(6.211)
Thus it follows by the approximate semigroup property of the resolvent and
196
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
by (6.207) that limsupp(:rs t, z) < limsup
-.
r- —
tJ^J±
< C < oo. (6.212)
So, there is a p-ball in D which contains the set {xS]t(z)} whenever t is small enough. Now we want to show that for a fixed s > 0 and a given z G D this net is a Cauchy net as t —> 0 + . Indeed, if we denote, as before, xa,t = Ts,t(z) a n d zt,r = (I + sfr)(xSit), we have z = (I + s/ t )(x,,t), and {zt,r} converges to z as t,r -» 0 + , because {ft} is a Cauchy net. But Ts,r(zt,r) = xs,t and we get p(Z,t(z),
Zir(z))
< p(Tt,t(z), Ts,r(zt,r))+p{Ts,r(zt,r),
Z,r(z)) (6.213)
as t, r —> 0 + . The proof is complete.
•
We denote by RNp(D) the class of all mappings / : D —» X for which the resolvent (J + rf)~1 is a well-defined /9-nonexpansive self-mapping of D for each positive r. We say that a mapping / : D —> X is an infinitesimal p-generator if for some T > 0, there exists a one parameter semigroup S = {St}te(o,T) of self-mappings St • D ^ D (St+r(x) = St{Sr{x)), x £ D, 0 < t + r < T) which are /O-nonexpansive for each t € (0,T): p(St(x), St(y)) < p(x,y),
x,y G D,
(6.214)
and such that lim \ p(x - tf(x), St(x)) = 0
(6.215)
exists for all x e D uniformly on each p-ball in D. In this case we will also say that / is a p-generator on (0, T). if T = oo we will write / € GNP(D). So if, in particular, D is a bounded convex domain in a complex Banach space and p is its hyperbolic metric, then a continuous p-generator on some interval (0,T) belongs to the class RNP(D). Thus we have the following analogs of the Hille-Yosida theorem. Theorem 6.15 Let D be a bounded convex domain in a complex Banach space X, and let p be its hyperbolic metric. Suppose that f : D —+ X is
Generation Theory for One-Parameter Semigroups
197
bounded and uniformly continuous on each p-ball in D. Then f 6 GNP(D) if and only if f e RNP{D). Theorem 6.16 Let D be a bounded convex domain in a complex Banach space X, and let f : D —> X be a holomorphic mapping. Then f generates a one-parameter semigroup on M.+ of holomorphic self-mappings of D if and only if for some T > 0 and for each r £ (0, T], the mapping Jr = ( 7 + r / ) " 1 is a well-defined holomorphic self-mapping of D. Moreover, in this case, Jr = (7 + rf)~l is a well-defined holomorphic self-mapping of D for all r >0.
Chapter 7
Flow-Invariance Conditions
7.1
Boundary Flow Invariance Conditions
Let V be a convex subset of a Banach space X and let / : V —» X be a continuous mapping on V, the closure of V. Then the following tangency condition of flow invariance lim dist(z - hf(x), V)/h = 0,
i£»,
(7.1)
is a necessary condition for the solvability of the Cauchy problem
(t + '<">-
«,2,
{ u(o) = i e D
oni?+ = [0,oo). It was shown in [Reich (1975); Reich (1976)] that condition (7.1) can be rewritten in another form which sometimes is more convenient. Namely, let X be a Banach space, and let V be a convex domain in X with 0 G V. We denote by p (— p-o) the Minkowski functional of V, that is, p(x) := inf{A > 0 : x G XD},
x € X.
(7.3)
(Note that since V is open, it is an absorbing set, i.e., for each x G X there is A > 0 such that x & \iD for all /x with |/x| > A.) It is known that p : X —* [0, oo) is continuous on X and that V = {x € X : px>(ar) < l } , ^ = { i £ X : p D (x) < l } .
(7.4)
Note also that p is a gauge function: p(x + y)< p(x) +p(y), p(Xx) = Xp(x), A > 0, 199
(7.5)
200
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
whence each one of the sets Ve = {x e X : p(x) < £}
(7.6)
is an open convex absorbing subset of X, the closure of which is Vt = {x S X : p(x) <£}. For 0 < £ < 1 we will call {Ve} the level sets of p. Now it follows by the Hahn-Banach theorem that for each x G X, there is a linear functional x* £ X such that
{
Re(x,x*)
=pv{x)
and
(7.7)
Re(y,x*)
(7.8)
and this means that j x is a support functional of T> at x. If T> is the open unit ball in X, then x* = j x for x G dV is a selection of the duality map J : X —* 2X' defined by J{x) = {x* G X* : (x,x*) = ||x||2 = ||z*|| 2 }.
(7.9)
Note that we can assume that j \ x — j x for all X > 0. If now / : T> —> X generates a strongly continuous semigroup S = {Ft}t>o, Ff.V^V, t>0, i.e., thm
ft(x) = f(x),
x£V,
(7.10)
where ft(%) = a ~ t ' x ' and the limit is taken with respect to the norm of X, then we have Re{ft(x),jx) = \ Re(z - Ft(x),jx) = i (l - Re(F t (x), j x )) > 0, a; G dV. Thus, it is evident that / satisfies the following boundary condition: Re{f(x),jx)
> 0, xe&D,
j x is a support functional of V at x. (7.11)
As a matter of fact, it follows by a result in [Reich (1975)] that (7.1) is equivalent to condition (7.11) if we require it to hold for all support
201
Flow-Invariance Conditions
functional j x . To see this, we define the set /(£>,x) = {z £ X : z = x + a(y - x) for some y € V and a > 0}. (7.12) The closure I(V, x) of this subset is sometimes called the support cone to V at x. Lemma 7.1 Let V be a convex subset of X. For z in X and x in V, the following are equivalent: (1) z G I(V,x).
(7.13)
(2) lim dM,({l-h)x
+ hz,V)/h
=
ti.
(7.14)
h—»0+
(3) Ifx* € X* supports V at x, then Re(z,x*) < Re(x,x*). Proof. (1)=>(3) is immediate. If z does not belong to I(D,x), then there exists x* in X* such that Re(z,x*) > sup{Re(y,x*) : y £ I(D,x)}. The functional x* must support I{V,x) at x. Consequently, (3)=^(1). The implication (2)=>(1) is also immediate. To prove (l)^=s>(2), let e > 0 be given. There are a > 0 and y € T> such that \z — (x + a(y — x))\ < e. LetO
Re(f(x),x*) > 0,
a; € dV.
(7.15)
For the classes of monotone and accretive mappings, theflowinvariance condition (7.1) (or equivalently, (7.11)) was systematically used to study the null point set of / and the Cauchy problem (7.2) (see, for example, [Brezis (1973)], [Reich (1975); Reich (1976)], [Goebel and Kirk (1990)], [Brezis (1970)], [Martin (1973)] and [Webb (1996)]. A result of Martin [Martin (1973)] shows that if V is a convex subset of X and / : T> —* X is a continuous accretive mapping on T>, then (7.1)
202
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
(respectively, (7.11) or (7.15)) is also sufficient for the existence of solutions to the Cauchy problem (7.2). This yields a continuous semigroup of nonexpansive mappings on T>. Theorem 7.1 Let T> be a convex subset of a Banach space X and let f :T> —> X be a continuous mapping on V. Assume that f is an accretive mapping on T>. Then f is a generator of a continuous semigroup of norm nonexpansive mappings on T> if and only if it satisfies the boundary flow invariance (7.1): lim dist(x-hf(x),V)/h
= Q, x € V.
(7.16)
For holomorphic mappings an analog of Martin's theorem can be formulated as follows. Theorem 7.2 ([Aizenberg et al. (1996)]) Let V be a bounded convex domain in a complex Banach space X and let f : T> —> X be a uniformly continuous mapping on T>, which is holomorphic in T>. Then f is a generator of a continuous semigroup of holomorphic self-mappings of T> if and only if it satisfies the flow invariance (7.1). Note that in contrast with Martin's theorem, we require that / should be uniformly continuous on V. The first question one can now raise is: What happens when f is not necessarily uniformly continuous on the closure VofV? We will answer this question as well as prove the above theorem (in a somewhat more general setting) by using the notion of the so-called numerical range. 7.2
Numerical Range of Holomorphic Mappings
Let V be a convex domain in a complex Banach space X, and suppose that V contains the origin. Let h : V —> X be a holomorphic mapping. Define hs(x) = h(sx),
0 < s < 1,
(7.17)
and note that hs is holomorphic in the enveloping domain £ V. Clearly, this domain contains T>, the closure of T>. For x e dT>, let J(x) be the set of all continuous linear functionals on X which are tangent to (support) V at x, i.e., J(x) = {I G X* : l(x) = 1, ReZ(y) < 1 for all y€V}.
(7.18)
Flow-Invariance Conditions
203
Let Q(x) be a non-empty subset of J(x). Definition 7.1 If h has a continuous extension to V, the numerical range of h (taken with respect to Q) is the set W(h)={l(h{x)):leQ(x),
xGdV}.
(7.19)
We write V in place of W when the numerical range is taken with respect to J. Note that in the case where V is the open unit ball B = {x£X: \\x\\ < 1}.
(7.20)
This definition agrees with that given by L. A. Harris in [Harris (1971b)]. We now define the upper and lower bounds L+(h) and L~{h) for the numerical range of holomorphic mappings as follows: L+(h) := lim supReV(/is)>
(7-21)
L~(h) := lim inf ReV(hs).
(7.22)
8-*l~
respectively, s—>1~
Using the maximum principle one can show that the limits in the above definition exist when V is balance. Definition 7.2 A holomorphic mapping h : T> —> X is said to be holomorphically dissipative (respectively, holomorphically accretive) if L+{h) < 0,
(7.23)
L'{h)>0.
(7.24)
respectively,
It is clear that h is holomorphically dissipative if and only if the mapping —h is holomorphically accretive. Definition 7.3 A holomorphic mapping h : V —» X is said to be holomorphically conservative if both h and — h are holomorphically dissipative (respectively, holomorphically accretive) i.e., L-(h) = L+(h) = 0.
(7.25)
Let p ( = px>) be the Minkowski functional of V and h : V —> X be a holomorphic mapping. Put ||ft||2> = sup{||M*)||:a:€P},
(7.26)
204
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
when this is finite. Suppose h has a uniformly continuous extension to T>. Then h is holomorphicaly dissipative (respectively, holomorphically accretive) if and only if supReV(/i) < 0 (respectively, inf ReV(/i) > 0), since ||/i — /i s || —> 0 as s —» 1~. Moreover, exactly as it was shown in [Harris (1971b), Theorem 2], one can prove that lun+ l | J + ^ l | p ~ 1 = sup ReW(h),
(7.27)
lim I I ^ - ^ H P - 1
(7.28)
respectively, =
infRe^(fc).
Thus, even though the numerical range may differ when taken with respect to different choices of Q, the number supReW(/i) (respectively, inf ReW(h)) remains fixed. Hence we can replace it by supRef(/i) (respectively, inf ReV(h)). In fact, this remains true for mappings which do not necessarily have a continuous extension to T>. (See the end of this section.) For the case of the open unit ball, the only properties of J{x) our arguments use are that J{x) ^ 0 and (7.29)
XJiXx) C J(x)
whenever |A| = 1 and ||a;|| = 1. For the case of bounded linear operators the notion of dissipativeness coincides with the classical one (see, for example, [Yosida (1974)]). In the theory of linear operators and its applications (to evolution equations, probability and ergodic theory), the classical Lumer-Phillips theorem (see, for example, [Lumer and Philips (1961)] and Theorem 6.6 in Chapter 6) plays an important role. For the case of bounded linear operators the Lumer-Phillips theorem can be reformulated as follows: A bounded linear operator A : X —> X (where X is a Banach space) is dissipative if and only if its resolvent (I — A-A)"1 is well defined on X for all X > 0 and satisfies the condition
||(7-AA)- 1 1| < 1,
A>0.
In other words, for all t > 0, the operator (I — tA)"1 contraction on X.
(7.30) : X —» X is a
205
Flow-Jnvariance Conditions
It turns out that for holomorphic mappings the notion of dissipativeness is equivalent to the flow invariance condition. First, we formulate the following extension of the Lumer-Phillips theorem. Theorem 7.3 Let B be the open unit ball of a complex Banach space X and let h : B —> X be holomorphic. Then h is holomorphically dissipative (respectively, holomorphically accretive) if and only if (I — th)(B) D B and (I — th)-1 is a well-defined holomorphic mapping of B into itself for each t > 0 (respectively, (I + th)(B) D B and (I + th)~1 is a well-defined holomorphic mapping of B into itself for each t > 0). In other words, h £ Hol(B, X) is holomorphically accretive if and only if it satisfies the range condition on B. We will, in fact, prove (the necessity part of) this theorem in a more general setting. Theorem 7.4 Let V be a bounded convex domain in X and leth :T> —> X be a holomorphically dissipative mapping bounded on each subset strictly inside V. Then for each t > 0, we have (I - th){V) D T> and (I -th)'1 is a well-defined holomorphic mapping ofV into itself. Proof. Fix y £ V and t > 0 and define (suspending our subscript convention) (7.31)
g.{x) = -(y + th{8x)) s
for x e V and 0 < s < 1. Given x £ dV and / £ J(x), we have
where
Rel(gs(x)) < - (ReJ(y) + s as = supReV(/i s ).
tas),
(7.32)
(7.33)
Assume without loss of generality that V contains the origin and let p be the Minkowski functional of V. Since p(y) < 1, it follows from our assumption that there exist a constant k < 1 and a number 6 > 0 such that - {p{y) + taa) < k s whenever 1 - S < s < 1. Then f := I — rga satisfies P(f(x)) > Re/(x - rgs(x)) > 1 - rk
(7.34)
(7.35)
206
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
whenever r > 0. The space X is an F-space with respect to p since V is open and bounded. Moreover, by hypothesis and the mean value theorem, there exists a number Ms satisfying P(9s(x) - 9s{y)) < Msp(x -y),
x,yeV
(7.36)
for each s with 0 < s < 1. Fix s with 1 — 6 < s < 1. Then by the inverse function theorem, for all small r > 0 the function / maps T> homeomorphically onto a subset of X containing 0. Hence f{V) 2 ( 1 - rk)V,
(7.37)
by (7.35). Therefore the function F:=f-\(l-r)I)
(7.38)
is a holomorphic mapping of cD into V, where c=(l-rk)/(l-r).
(7.39)
Clearly, c > 1. Since V lies strictly inside cD, the Earle-Hamilton theorem implies that F has a unique fixed point x* in T>. This is also a fixed point for g3 and hence z = sx* satisfies (I~th)(z) = y.
(7.40)
Suppose now that there is another solution z\ of (7.40). Choose s\ < 1 so that s\ > s and si > p{z{). Clearly, x\ = zi/si is in T) and satisfies gSl{xi) = x\. One can show as above that x\ is the unique fixed point of gSl in T>. On the other hand, if we set xi = z/si, then X2 £ "D and 9si(x2) = %2 by (7.40). Hence #2 = xi, so z\ = z. It can be shown (as, for example, in [Harris (1977), Theorem5]) that the solution z of (7.40) depends holomorphically on y £ V. This completes the proof of our asser• tion. Thus we have proved the necessity part of our generalization of the Lumer-Phillips theorem. The sufficiency part is proved in [Harris et al. (2000)]. Another immediate consequence of this theorem and the holomorphic analog of the Hille-Yosida theorem is the following boundary flow invariance condition.
Flow-Invariance Conditions
207
Corollary 7.1 Let T> be a bounded convex domain in a complex Banach space X, and let f :V —> X be a uniformly continuous mapping on V which is holomorphic in T>. If f satisfies the boundary condition inf
x*GJ(x)
Re(a;),a;*)>0,
x £ dV,
(7.41)
then it generates a semigroup of holomorphic self-mappings of T>. This corollary proves, in fact, the last theorem of the previous section. On the other hand, regarding the sufficiency part of the generalized Lummer-Phillips theorem, it does not seem natural to consider a boundary condition to characterize generators. This is because a bounded convex domain itself is a complete metric space with respect to its hyperbolic metric. In addition, there are many examples of holomorphic generators which have no continuous extension to V (see the examples below). In particular, if F £ Hol('D), then / = I — F is a generator because it satisfies the range condition. Thus, the following question arises: 7s there an interior flow invariance condition which characterizes the class of generators ? 7.3
Interior Flow Invariance Conditions
It would be desirable to find such an interior flow invariance condition from which (7.41) could be derived in the case where / £ Hol(P, X) has a continuous extension to V. For the Euclidean ball T> in X = C n , a certain condition in this direction was established by M. Abate [Abate (1992)]. Namely, he proved that / G Hol(I', C n ) is a generator if and only if it satisfies the estimate 2[\\g(x)\\2
-\(g(x),x)\2]Re(g(x),x)
+ (1 - ||a;||2)2Re{/'(x)/(a;), g(x)) > 0,
(7.42)
where g{x) = (1 - ||z|| 2 )/(z) + {f{x), x)x.
(7.43)
For n = 1 this condition becomes Ref(z)z>-~Ref'(z)(l-\z\2),
(7.44)
where z £ A, the open unit disk in the complex plane C and / € Hol(A, C).
208
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
Despite the simplicity of condition (7.44) it is not clear how (7.41) can be derived from (7.44) when / has a continuous extension to A. On the other hand, this condition may be useful in studying the behavior of the derivative of a semi-complete vector field in A. Therefore it is natural to ask the following question: Can this condition be extended to a general Banach space in a form similar to (7.44) (instead of condition (7A2))? Note also that in the one-dimensional case it follows from the maximum principle for harmonic functions that if / E Hol(A,C) has a continuous extension on A, then the boundary flow invariance condition Ref(z)z > 0,
(7.45)
z e A
implies the following interior condition: Ref(z)z>Ref(0)z(l-\z\2),
z e A.
(7.46)
Conversely, it is clear that (7.45) does result from (7.46) if / has a continuous extension to all of A. Thus another question arises: Are (7.46) and its Banach space analog necessary and sufficient for f to be a generator? The following result provides affirmative answers, in any Banach space, to all the questions raised above. Theorem 7.5 Let V be the open unit ball in a complex Banach space X. Then f G Hol(P, X) is a generator on T> if and only if it is bounded on each subset strictly inside T> and one of the following conditions holds: (a) For each x € V there exists x* G J(x) such that (7.47)
Re(f(x)-{l-\\x\\2)f(0),x*)>0; (b) inf
x*& J(x)
R e ( 2 | | z | | 2 / ( z ) + (1 - \\xf)f'(x)x,
x*)>0
x e P ; (7.48)
(c) for each x g T> and for each x* £ J{x),
Re(^-^f'(0)x+(l-\\xe)f(0),x*)
I
I
(7.49)
Furthermore, equality in one of the conditions (a), (b) or (c) holds if and only if it holds in the other conditions and f is a generator of a group of automorphisms ofV.
209
Flow-Invariance Conditions
Note that in the one-dimensional case condition (a) reduces to (7.46) and condition (b) becomes Abate's condition (7.44). Proof. First we observe that condition (a) (as well as condition (c)) implies that / must be a holomorphically accretive mapping, hence satisfy the range condition on V. Thus, to prove our theorem we just have to show the equivalence of conditions (a), (b) and (c), and that the property of / to be a generator implies condition (a). So let / be a generator of the one-parameter semigroup {Ft}t>o of holomorphic self-mappings of V. We intend to prove that condition (a) of the theorem is satisfied. Indeed, fix any x e V and x* e J(x), and set u = a;/||a;||, u* = a;*/||a;||. Consider the holomorphic function / on the unit disk A C C defined as follows: /(A) = (/(Au),u'>,
AeA.
(7.50)
Similarly we define a family {Ft}t>o of holomorphic self-mappings of A: F t (A) = (Ft (Au), u*),
A e A , t > 0.
(7.51)
It is clear that Fo(A) = A and that there exists the limit tlim+
\ (A - Ft (A)) = /(A).
(7.52)
Now let Mb denote the Mobius transformation on A defined by Mb(X) = ^ % , 1 — Ao
(7.53)
be A.
Consider the family {Ht}t>o of holomorphic self-mappings of A denned by Ht(\)
= Mh{0)(Ft(X)),
(7.54)
\£A,t>0.
Note t h a t since W«(0) = 0 for all t, it follows by the Schwarz L e m m a t h a t |Wt(A)| < |A| for all A e A and t > 0.
(7.55)
Now, by simple calculations, one can conclude t h a t for each A e A t h e curve Wt(A) : M + —> A is right-differentiable at zero a n d t Um
A
~ ^ t ( A = / ( A ) - / ( 0 ) + ?(0)A 2 : = h(X),
AeA.
(7.56)
It follows now by (7.55) a n d (7.56) t h a t for all A e A ,
Reh(\)X > 0
(7.57)
210
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
and ft(0) = 0.
(7.58)
Re/(A)A > Re/(0)A(l - |A|2).
(7.59)
But (7.57) means that
Setting in this inequality A = ||a;|| we have by (7.50), Re{f(x),x*} > Re(0),z*)(l - ||x||2).
(7.60)
Since x and x* £ J(x) are arbitrary, this proves the following implication: If / is a generator, then / satisfies (a). Now we will show that (a) is equivalent to (b). To do this, we return again to the function /(A) = /(0)-7(0)A 2 + MA)
(7-61)
denned by (7.50), where h(X) satisfies (7.56) and (7.57). But these conditions are equivalent to the conditions h(X) = X-p(X),
A€ A
(7.62)
with Rep(A) > 0.
(7.63)
So, (7.59) is equivalent to (7.62) and (7.63) for / of the form (7.61). Now, in the same terms, we translate condition (b). If we define / as above by (7.50), then (b) implies Re[2/(A)A + /'(A)(l - |A|2)] > 0.
(7.64)
If we substitute here / in the form (7.61) with h(X) = Xp)X) we see that (7.64) is equivalent to the condition
Re [Ap'(A) + i ± | ^ p ( A ) ] > 0, A e A.
(7.65)
We intend to show that (7.65) is equivalent to (7.63). Then setting again A = ||z|| in (7.50) and (7.64) and noting that /'(A) = (f'(Xu)u,u') = (f'(x)x,x*)Tr^ for x ^ 0, we will get by continuity the equivalence of conditions (a) and (b) of the theorem.
211
Flow-Invariance Conditions
So, let p e Hol(A, C) satisfy (7.63). Define F = (p - l)(p + I)" 1 . Since the mapping w = |^y maps the right half-plane into A, F is a self-mapping of A. Applying the Schwarz-Pick Lemma to F, we obtain (p-l\ \p+l)
_
2\p'\ |p + l | 2 - | p - l | 2 | l + p P - | P+ 1|2(1_|A|2)
V™>
or
b ' W < ^ .
(7.67)
This implies
Re(-Ap'(A)) < |Ap'(A)| < ' ' ^ f f l < \ ^ Rep(A), (7.68) which is equivalent to (7.65). In the opposite direction ((7.65) implies (7.63)) we prove the following somewhat more general fact: Let p S Hol(A,C) and suppose that there is a positive function ip : [0,1) —> R+ such that the following condition holds: Re(Ap'(A) + V(|A|)p(A)) > 0,
A e A.
(7.69)
Then Rep(A) > 0 everywhere on A. Indeed, setting A = re%e we have V(A)=r^
(7.70)
and (7.69) becomes Re (r^j
+ V>(r)Rep(A) > 0,
A = reie e A.
(7.71)
Assume now that there exists Ao = roei9° in A such that Rep(Ao) < 0.
(7.72)
Since (7.69) implies that Rep(0) > 0, there exist 0 < J~I < ro such that Rep(7-iei9°) = 0 and Rep(roeie°) < 0. Thus one can find r 2 e (ri,r0) such that Rep(r2ei9°) < 0
(7.73)
Re^(rac<*")<0.
(7.74)
and
212
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
This implies Re (r-2^)
+ rp(r2)Rep(r2eieo) < 0,
(7.75)
a contradiction. Thus Rep(A) > 0 for all A e A and we are done. In other words, conditions (a) and (b) are equivalent. To obtain (c) we now represent /(A) = (f(Xu),u*) in the form
f(X) = a + ibX-aX2 + X f I±££dji(O. JdA 1 - C*
(7 . 76)
and we calculate: 1 + C A = 1-|A[2 1 - C"A |1 - CA|2
=
1 - 1 * 1 (1 + |A|) 2 1 + |A| |1 - CA|2
=
1 + 1A| ( l - j A | ) 2 1 - |A| ' |1 - CAP • (7.77)
Since |C| = 1, this equality shows t h a t
U * l < Rel±CA 1+JA[
(7.78)
Noting that p(0) = /'(0) (see (7.61) and (7.62)), we obtain from (7.76) and (7.78)
Re/(0)A(l - |A|2) + | A | 2 ^ j /'(0) > Re/(A)A > Re/(0)A(l - |A|2) + | A | 2 [ ^ J /'(0).
(7.79)
Once again, setting A = ||x|| we get from the last inequality condition (c). Since by (b), Re(f'(0)x,x*) > 0, it is clear that (c) is stronger than (a).
•
The following fact is a direct consequence of Theorem 7.5. Corollary 7.2 Let f 6 Hol(P, X) be a holomorphic generator on T>. Then the linear operator A = f'(0) is (totally) accretive, i.e., inf
y*€J(y)
Re(Ay,y*) > 0,
y G X.
(7.80)
Proof. Substitute x = ty, x* = ty* in (b), where \\y\\ = \\y*\\ = 1, y* £ J(y) and t 6 (0,1). Letting t tend to zero we get (7.80). In its turn, (7.80) implies that the left-hand inequality in (c) is sharper than (a). Moreover, it implies that (a) holds for all x* £ J(x). Thus we
Flow-Invariance Conditions
213
have that if / has a continuous extension to V, then (a) yields the flow • invariance boundary condition (7.41) (equivalently, (7.1)).
7.4
Semi-Complete and Complete Vector Fields
In connection with flow invariance conditions we give the following definition. Definition 7.4 A holomorphic vector field
(7.81) on a domain V is determined by a holomorphic mapping / £ Hol(£>, X) and can be regarded as a linear operator mapping Hol(X>, X) into itself, where Tfg £ Hol(£>, X) is defined by (Tfg)(x)=g'(x)f(x),
xGV.
(7.82)
The set of all holomorphic vector fields on V is a Lie algebra under the commutator bracket
[Tg,Th] = \g(x)^
, h ( x ) ^ ] := {g'(x)h(x) - h'(x)g{x))^.
(7 . 8 3)
(see, for example, [Dineen (1989)]). Furthermore, each vector field (7.81) is locally integrable in the following sense: for each x £ T> there exist a neighborhood ft C T> of x and 8 > 0 such that the Cauchy problem (7.2) has a unique solution {u(t, x)} C T> defined on the set {\t\ <S}xQ,cRxV. Definition 7.5 A holomorphic vector field Tf defined by (7.81) and (7.82) is said to be (right) semi-complete (respectively, complete) on T> if the solution of the Cauchy problem (7.2) is well-defined on all of R+ x V (respectively, R x P), where R + = [0, oo) (respectively, R = (-00,00)). Thus, if V is hyperbolic, then Tg is semi-complete (respectively, complete) if and only if g is the generator of a one-parameter continuous semigroup (respectively, group). On the other hand, if T> is bounded and Hol(P, X) is the subspace of Hol(P, X) consisting of all those g £ Hol(r>, X) which are bounded on each ball strictly inside V, then a semigroup (group) {Ft}, t G R +
214
Nonlinear Semigroups, Fixed Points, and Geometry of Domainsh
(respectively, t G R), induces ^Jinear semigroup (group) {L(t)} of linear mappings L(t) : Kol(D,X) i-> Hol(£>,X), denned by (L(t)g)(x) := g(Ft(x)),
(7.84)
where i € l + ( t 6 K) and x e V. This semigroup is called the semigroup of composition operators on Hol(P,X). If {F t }, t € R+(t € R), is T-continuous (that is, differentiate), then {L(t)}, t e R + (i £ R), is also differentiable and
(7.85)
I £(0)<7 = 5
for all g 6 Hoi(P, X), where f = ^ _ . In other words, a holomorphic vector field Tf, denned by (7.81) and (7.82) and considered a linear operator on Hol(£>, X), is the infinitesimal generator of the semigroup {L(t)}. It is sometimes called the Lie generator. Thus a holomorphic vector field Tf is semi-complete (respectively, complete) if and only if it is the Lie generator of a linear semigroup (respectively, group) of composition operators on 1101(1?, X). This follows from the observation that (7.86)
L(t)Iv = Ft and Tflv = /,
(7-87)
where I-D is the restriction of the identity operator to V. Moreover, using the exponential formula representation for the linear semigroup: L®9
= E ~ljr-
Tf9
= eM-tTf]g
(7.88)
= <*PHT/]J O .
(7.89)
fc=O
(see Chapter 6), we also have 5W
= E -TT~ fc=o
K-
Tffv
Flow-Invariance Conditions
215
So, a locally uniformly continuous semigroup of holomorphic self-mappings can be represented in exponential form by the holomorphic vector field induced by its generator. In particular, Theorem 7.5 (see the previous section) asserts that if / £ Hol(D, X) then T/ (= f{x)-§^) is a semi-complete vector field if and only if / satisfies the interior flow invariance condition (a) (respectively, (b) or (c)). Furthermore, condition (c) also leads to a characterization of a semicomplete vector field to be complete. Following S. G. Krein [Krein (1971)] (see also [Vesentini (1996a]), we say that a linear operator A : X —> X is conservative if for all x € X and x* £ J{x) there holds Re(Ax,x*) = 0.
(7.90)
In order to simplify our terminology we will sometimes identify a semicomplete (complete) vector field Tf with its determining holomorphic mapping / , which is, in fact, the infinitesimal generator of a semigroup (group) of holomorphic self-mappings on T>. The following corollaries are consequences of Theorem 7.5 in the previous section. Corollary 7.3 Let f e Hol(X>, X) be a semi-complete vector field. Then f is actually complete if and only if its derivative at zero, /'(0), is a conservative linear operator. It is well known that a complete vector field g on the open unit ball V in a Banach space X is a polynomial of degree at most 2 (see, for example, [Arazy (1987)], [Dineen (1989)] and [Upmeier (1986)]). More precisely, g has the form g{x) = a + Ax + Pa{x),
(7.91)
where a is an element of X, A is a conservative operator on X, and Pa is a homogeneous form of the second degree such that Pia = iPa. One of the consequences of this representation is an infinitesimal analog of Cartan's uniqueness theorem ([Arazy (1987)], [Dineen (1989)] and [isidro and Stacho (1984)]): If g 6 Hol(Z>,X) is a complete vector field such that 5(0) = 0 and g'(0) = 0, then g = 0. Applying Corollary 7.3, we obtain at once the following extension of this theorem.
216
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
Corollary 7.4 If f £ Hol(D, X) is a semi-complete vector field on V such that /(0) = 0 and /'(0) = 0, then f = 0. We know already that if / e HolCD, X) has the form (7.92)
f = I-F,
where F is a self-mapping of V, then / is semi-complete. Thus Corollary 7.4 is also a generalization of Cartan's uniqueness theorem (sometimes this theorem is also called the Generalized Schwarz Lemma (see Chapter 3)). Suppose now that a complex Banach space X is a so-called JB* triple system. This is equivalent to saying that its open unit ball I? is a homogeneous domain, i.e., for each pair x,y £ T> there exists a holomorphic automorphism of V such that F(x) = y (see, for example, [Upmeier (1986); Dineen (1989); Isidro and Stacho (1984)]). Then it is well-known that for each a £ X there exists a homogeneous polynomial Pa{x) such that Pia = iPa and the mapping g :T> —> X defined by g{x)=a-Pa{x)
(7.93)
is a complete vector field on T>, which is called a transvection of T>. Using this fact and Corollary 4 in [Aharonov et al. (1999b)] we get the following representation theorem. Corollary 7.5 Let X be a JB* triple system and let V be its open unit ball. Then the cone Q of semi-complete vector fields on V admits the decomposition (7.94)
g = go@g+,
where QQ is the real Banach subspace of HitOO(V,X) consisting of transvections and Q+ is the subcone of Q such that for each h £ Q+, XD£ Re(h(x),x*)>0
for all z £ P .
(7.95)
In other words, / G Q admits a unique representation f = 9+ h
(7.96)
where g = /(0) - Pf(0)(x) is complete, h € G+ and h(0) — 0. The natural examples of JB* triple systems are a complex Hilbert space H, the space of bounded linear operators L(H) on H, and its subspaces J such that A £ J if and only if AA* A £ J (such subspaces are called
217
Flow-Invariance Conditions
J*-algebras); see [Harris (1974a); Dineen (1989)]. In the latter case the general form of transvections on V is (7.97)
g(x) = a - xa*x,
where a £ J and a* is its conjugate. Thus each semi-complete vector field on the open unit ball of a J* -algebra has the form (7.98)
f(x) = f(0)-xf*(0)x + h(x),
where he 6+ and h(0) = 0. In particular, when X = C is the complex plane and V = A the open unit disk in C, (7.98) becomes /(*) = / ( 0 ) - 7 ( 0 ) * 2 + *P(*),
(7-99)
where p{z) e Hol(A, C) and Rep(z) > 0,
(7.100)
zGA.
That is, p(z) is a function in the class of Caratheodory. Using the RieszHerglotz integral characterization of this class (see, for example, [Duren (1983)] and [Aleksandrov (1994)]) we deduce the following conclusion: / € Hol(A, C) is a semi-complete vector field if and only if it admits the representation
f{z) = a + ibz-az2 + z [
1+iZ
JdA 1 - C*
d/x(C),
(7.101)
where a e C, b G M and fi is a positive measure on dA. As a matter of fact, this representation is the key to proving the results in the previous section because we have partially used a reduction to the one-dimensional case. We have also obtained (7.99) with (7.100) by an independent method.
Chapter 8
Stationary Points of Continuous Semigroups 8.1
Generalities
Let V be a topological space, and let a family S = {Ft € (0, T)}, T> 0, of self-mappings Ft of V form a (one-parameter) continuous semigroup, i.e., (i) F3+t = Ft-Fs,
0<s + t
(8.1)
(ii) hmFt(x) = x,
(8.2)
where the limit is taken with respect to the topology of V. Definition 8.1 A subset W of V is said to be the stationary point set of S if it consists of all the points a £ T> such that Ft(a) = a for all te (0,T).
(8.3)
In other words,
W=
P) FixF t .
(8.4)
0
The first problem in this context is the existence problem of a stationary point of S, in other words, to determine when W ^ 0. The second problem is the uniqueness problem, that is, to find out when W consists of a unique point a € T>. This problem is connected with a more general, rather important in applications, problem - to determine the structure of the set W in relation to the topological structure of T>. Another important problem is to find constructive methods for the approximation of W. In 219
220
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
the case when T = oo this problem, in particular, is closely related to the study of the asymptotic behavior of semigroups. Definition 8.2 A point a 6 P i s said to be a periodic point of a semigroup S = {Ft : t £ (0, T)} if for some t0 > 0 this point is a fixed point of Fto, i.e., a = Fto(a), hence a = -Ft"(a) for all n = 0,1,2,... . The simplest situation is, of course, when for some t0 > 0 the mapping Ff0 : T> —> T> has a unique fixed point a = Ft0 (a) in T>. Then directly from the semigroup property (i) it follows that W = {a} ^ 0. But even in this case the approximation problem is still open. Sometimes one can study the existence and the structure of W as the common fixed point set of the commuting family {Ft}. Indeed, if, for example, V = B™, where B is the open Hilbert ball, then the following assertion, due to I. Shafrir ([Shafrir (1992a)]) is useful for the study of the common fixed point sets of one-parameter semigroups, which consist of nonexpansive mappings with respect to the hyperbolic metric p on V (= B n ). Lemma 8.1 Let {Fa : a £ J} be a commuting family of pnonexpansive (holomorphic mappings) self-mappings o/B n and assume that there is a p-bounded subset M C B™ which is Fa-invariant for all a £ J. Then there exists a common fixed point a £ B n of this family. Remark 8.1 Observe that if in Lemma 8.1 we omit the assumption that there exists a p-bounded subset M c B n which is Fa-invariant for all a £ J, then one can construct a counterexample even for n = 1. Example 8.1 ([Kuczumow et al. (2001a)]) Let B be the open unit ball in a separable Hilbert space (H, (•, •)) with the orthonormal basis {en}n6pjWith each affine subset Y = (x + Y) nB of B, where Y is a closed subspace of X and x £ B is the unique point of least norm in Y, we associate the nearest point holomorphic retraction of © onto Y given by the formula R = Mxo Projy o M_ x ,
(8.5)
where Mx is a Mobius transformation. We construct the family {-Fn}n6N in the following way: we choose a sequence of positive real numbers {an}nen such that this sequence is strictly decreasing, 0 < an < h n and lim an = n—*oo
0. Let us set Xo = X, Co = B, a0 = 0, r0 = ||ci - ao||.
(8.6)
221
Stationary Points of Continuous Semigroups
For n = 1, the mapping R\ is the nearest point holomorphic retraction of B onto the affine set C\ in B, where
(8.7) gi = e 2 - a0 - j .
j72 (e 2 - a0, ei - o 0 )(ei - a 0 ) ,
ra*-
*i =
(8.8)
(8-9)
ai = a o + ro ( - + - cosai)/ x + ( - s i n a ^ M ,
(8.10)
ri = ||ei - a i | | ,
(8.11)
Xi = {xeX0:x
= a1+y
and (ai,y) = 0}
(8.12)
and Ci=Xinl.
(8.13)
£ X, fi, f2,..., / „ £ If we already have r 0 , n,..., rn £ M, ai, a2,...,an X, ffi,ff2, • • • i3n € X, hi,h,2,...,hne X, X0,Xi,... ,Xn, CQ,C\, . . . , C n and Ri,R2,...,Rn, where each Xi is an affine set in X, e i , a , 6 d c Xuek € ^ for /c > i + 2(i = 0 , . . . ,n),Xi+1 C Xi for i = 0 , . . . , n - 1, and each Rt is the nearest point holomorphic retraction of B onto d for i = 1 , . . . , n, then the next holomorphic retraction Rn+i is the nearest point holomorphic retraction onto Cn+1=Xn+1nB,
(8.14)
where fn+i = ij jT (ei - an), \\ei - an\\ 9n+l = e-n+2 - On — T, ftn+i = M
r,9n+\,
(8.15) TTo ( e "+2 "" an el ~ an){ei
||ei — an\\
~ an),
(8.16) (8-17)
an+i = a\\9n+l\\ n + rn U - + - c o s a n + i j / n + i + ( - sina n + ij/i n + i ,(8.18) r-n+i = llei-On+iH,
(8.19)
222
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
and Xn+1 = {x e Xn : x = an+i + y and ( a n + 1 , y ) = O } .
(8.20)
Now it is sufficient to take Fn = RnORn^
o.-.oRx
(8.21)
= Fmo Fn
(8.22)
for n = 1,2,... to get FnoFm for every m, n £ N, FixF n = Cn
(8.23)
lim diamCn = 0,
(8.24)
for each n s N, n—too
and oo
f ) FixF n = 0.
(8.25)
n=l
Nevertheless, for a one-parameter continuous semigroup of pnonexpansive mappings the following assertion holds. Theorem 8.1 Let B be the open unit ball in a complex Hilbert space H and let p be the hyperbolic metric on B n , n > 1. If S = {Ft : t > 0} is a continuous semigroup of p-nonexpansive self-mapping ofW1, then the following statements are equivalent: (i) S has a periodic point in W1, i.e., there is to > 0 such that Ft0 has a fixed point in B n ; (ii) Ft has a fixed point in B n for each t > 0; (in) there is a stationary point a S B n of the semigroup S. We prove this theorem in a somewhat more general setting. In this connection we give the following definition. Definition 8.3 Let V be a topological space. We say that a class G of self-mappings of V has the compact commuting property (CCP) if for each commuting family {Fa : a £ J} the set f] Fix.Fa is not empty a£j
whenever there is a compact subset M of V which is Fa-invariant for all aE J.
Stationary Points of Continuous Semigroups
223
Proposition 8.1 Let T> be a topological space, and let G be a class of self-mappings of V which has the compact commuting property (CCP). If S = {F t : t > 0} C G is a continuous semigroup of self-mappings of V which has a periodic point in V, i.e., there is SQ > 0 such that FSo has a fixed point in T>, then there is a stationary point of the semigroup S, i.e., (1 FixF t ^ 0. t>o
Proof. Let a £ V be a periodic point of S, that is, a is a fixed point of FSo for some s 0 > 0. Consider the compact subset M of V defined by M = {Ft(a) : 0 < t < s0}. We claim that FS{M) C M for all s > 0. Indeed, if x G M, then there is 0 < p < so such that x = Fp(a), and we have, by the semigroup property (i), that Fs(x) = Fs(Fp(a)) = Fs+P(a).
(8.26)
On the other hand, the number s+p can be represented as s + p = USQ +1 for some 0 < t < SQ. Once again, by the semigroup property (i), we see that Fs(x) = Fs+P(a) = Ft+n30(a) = Ft(Fnso(a)) = Ft(F?0(a)) = Ft(a)eM.
(8.27)
This proves our claim which, in turn, implies the existence of a common fixed point of S. O Now Theorem 8.1 is seen to be a consequence of Proposition 8.1 and Lemma 8.1. Corollary 8.1 Let S be a continuous semigroup of holomorphic mappings ofM such that for some to € (0,oo), Fto — I- Then S is a group of automorphisms ofM of elliptic type, i.e., Ft = M~l o etA o M, where A is a linear conservative operator (Re{Ax, x) — 0 for all x in H) and M is a Mobius transformation o/B. Proof. Since Fto = /, Theorem 8.1 shows that S has a common fixed point o £ l . Let M = M-a denote a Mobius transformation of B such that M(a) — 0, and consider the semigroup $(£) = M o Fto M" 1 defined on B. It is clear that $(t)(0) = 0 for all t > 0, and that $(t 0 ) = /. Since $(s) o $(t 0 - s) = $(*o) = /
(8.28)
224
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
for all 0 < s < t0, we see that for all 0 < s < t0, <J>~J(s) = $(£0 — s) and $(s) is an automorphism of B. Hence $(s) is the restriction to B of a linear isometry of H onto itself for each s > 0. This implies our assertion. • By using the notion of the duality mapping one can easily reformulate this result for B n too. Occasionally, the mappings Ft, t £ M.+ , will also be denoted by F(t). The above approach implies also the following observation. If V is a strongly convex bounded C 2 domain in C n and {F(t) : t > 0} is a semigroup of holomorphic self-mappings of V such that for each t > 0, F(t) has a fixed point in V, then it follows by a result in [Abate (1989a)] (see also [Abate and Vigue (1991)]) that this semigroup has a common fixed point (stationary point) in I? as a commutative family of holomorphic mappings. Moreover, as in the case of the Hilbert ball, it is enough to require the existence of an interior fixed point only for one to > 0 to provide the existence of such a point for the whole semigroup. Theorem 8.2 (cf. [Abate (1988b)]) Let V be a strongly convex bounded C2 domain in Cn and let {Ft : t > 0} be a semigroup of holomorphic self-mappings ofV. Then the following assertions are equivalent: (a) The semigroup {Ft} has a stationary point in T>. (b) The semigroup {Ft} has a periodic point in V, i.e., there exists to > 0 such that Ft0 has a fixed point in T>. (c) There exists x £ T> and a sequence tn —> oo such that {Ftn(x)} is strictly inside T>. (d) For each x e V there is a sequence tn —» oo such that {Ftn(x)} is strictly inside V. We will give a proof of this theorem which is different from the one in [Abate (1988b)] by using the infinitesimal generator of a semigroup in a somewhat more general situation. Unfortunately, we cannot assert the full analog of this theorem for the infinite dimensional case (even for the Hilbert ball), but for some classes of p-nonexpansive or holomorphic mappings one can formulate a similar statement. We need the following notion. Definition 8.4 Let V be a domain in a Banach space X. We will say that a class G of self-mappings of V has the Denjoy-Wolff iteration property (DWIP) if whenever F G G has no fixed point in V, the sequence of iterates {Fn} strongly converges to a point on the boundary of V, uniformly on each compact subset of V.
225
Stationary Points of Continuous Semigroups
Proposition 8.2 Let G be a class of p-nonexpansive which satisfies the Denjoy-Wolff iteration property. Let a one-parameter continuous semigroup of p-nonexpansive which has no common fixed point inM. If Fto € G for at then there is a point e € <9B such that S converges toeast uniformly on each compact subset o/B.
mappings on B S = {Ft}t>o be mappings on B least one to > 0, tends to infinity,
Proof. First we note that by Theorem 8.1, Fto has no fixed point in B. Therefore there is a point e £ dM such that Fnt0 converges to e uniformly on each compact subset of B. Let C be a compact subset of B. Since the semigroup S = {Ft}t>o is continuous, the set C : = {zs = Fs(z) :z&C,
0 < s < t0}
(8.29)
is also a compact subset of B". Therefore, for each e > 0 one can find n 0 € N = {1,2,... } such that s u p | | F t " W - e | | = sup sup | | F n t o ( z s ) - e | | z€C
i€c
0<s
= sup sup ||F nto+s (5) - e|| < e sec o<s
(8.30)
for all n> UQ. If follows now that ||Ft(z)-e||<e for each t > no^o a n d z € C.
(8.31) •
Remark 8.2 For the Hilbert ballM the following classes of self-mappings o/B are known (see [Kuczumow et al. (2001a)]) to satisfy the Denjoy-Wolff iteration property: (1) the class Q\ consisting of the condensing holomorphic mappings; (2) the class Qi consisting of the firmly p-nonexpansive mappings of the first kind; (3) the class Qz consisting of the firmly p-nonexpansive mappings of the second kind; (4) the class Gi consisting of the averaged mappings of the first kind, i.e., F = (1 — c)I © cT, where T is p-nonexpansive and c G (0,1); (5) the class Q5 consisting of the averaged mappings of the second kind, i.e., F = (1 — c)I + cT, where T is p-nonexpansive and c £ (0,1). Thus if a semigroup S = {Ft}t>o contains at least one element Ft0, to > 0, of the class Gi,l
226
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
In a somewhat more general situation one can say the following. Proposition 8.3 Let D be a domain in a Banach space X, and let G be a class of self-mappings of V which has both the (CCP) and the (DWIP) properties. Suppose that {Ft : t > 0} C G is a semigroup of self-mappings of T>. Then the following assertions are equivalentfa) The semigroup {Ft} has a stationary point in T>. (b) The semigroup {Ft} has a periodic point in V, i.e., there exists to > 0 such that Ft0 has a fixed point in V. (c) There exists x G T> and a sequence tn —•> oo such that {Ftn(x)} is strictly inside V. (d) For each x G V there is a sequence tn —* oo such that {Ftn{x)} is strictly inside T>.
8.2
Generated Semigroups
As a matter of fact, we already know that if P is a bounded domain in C n , then each continuous semigroup of holomorphic self-mappings of V is locally uniformly continuous and hence differentiable at t = 0. We consider now such a situation in general. That is, we assume that T> is a domain in a Banach space X and a semigroup S — {Ft : t > 0} is generated on T>, i.e., for each x € T> there exists the strong limit f(x)=lim+-t(x-Ft(x)).
(8.32)
If in this case the mapping / : V —> X (the infinitesimal generator of S) is locally Lipshitzian on V, then, by using the uniqueness of the solution to the Cauchy problem
lim u(t, x) = x, \ t-»o+
(8.33)
where u(t, x) = Ft(x) and / is defined by (8.32), it follows that the stationary point set W is the null point set (Nullp/) of / in V, i.e., W = p ) Fixx,Ft = Null©/.
(8.34)
l>0
Once again, we continue with a somewhat more general situation.
227
Stationary Points of Continuous Semigroups
Definition 8.5 We say that a class M of self-mappings of V has the common fixed point property (CFPP) if the following condition holds: If {F3}S£A i s a ne * of commuting mappings in Ai such that for each s € A, Fs has a fixed point in V, then f] Fixpi^ ^ 0. s€A
Proposition 8.4 Let X be a real or complex Banach space, and let T> be a domain in X. Suppose that M is a class of self-mappings ofT> with both the (CFPP) and the (DWIP). Let f : V -> X be the infinitesimal generator of a one-parameter continuous semigroup {F(t) : t > 0} C M such that {F(t)} is locally uniformly Lipschitzian on V. If f has no null point in V, then for each x 6 V the net {F(t)(x)} strongly converges to a point b on the boundary ofV. Proof. First we note that there is an interval (0, /i) such that for each t G (0, /J,), F(t) has no null point in V. Indeed, if we suppose that there is a sequence tn —> 0 such that for each n the mapping F(tn) has a fixed point in V, then by the (CFPP) there is a point x GV such that F(tn)(x) = x for all n and hence f[x) = 0. This is a contradiction. So, for each m large enough (m > l//x) all the mappings F ( ^ ) have no fixed point in V. This means that for each x e V the sequence F ( ^ ) (x) (= Fn (~^)) converges strongly as n —> oo to a point bm on the boundary of T>. But it follows from the semigroup property that Fnm ( ^ ) (x) = Fn(l)(x) -> bm, a s n - » o o , and hence bm = b does not depend on m. We are now able to show that the semigroup {F(t)} strongly converges to 6 as t tends to infinity. In fact, for any given e > 0 and x G V, we can choose S > 0 such that ||.F(t)(a:) - F(t)(y)\\ < e/2 for all t > 0 whenever y € V and ||y-a;|| < 5. For such 8 we take m e N so large that m" 1 € (0, /x) and \\F(h){x) - x\\ < 6 for all h € [0,1/m). Finally, for such m and t > 0, setting n = [tm], we have \\F(£)(X)
~b\\ = \\Fn(±)(x)-b\\ < e/2
(8.35)
for t > 0 big enough. Since h = t- % € [0, i ) , we get for such t > 0, \\F(t)(x)-b\\ = \\F(~)(F(h)(x))-b\\<
IK)w)w)-^)w||+HS)w-*i
(8.36)
228
iVon.Kn.ear Semigroups, Fixed Points, and Geometry of Domains
and we are done.
•
Corollary 8.2 Under the conditions of Proposition 8.4, assume in addition that f is also locally Lipschitzian. Then the semigroup {F(t)} has a stationary point in V if and only if for some t 0 > 0 the mapping F(to) has a fixed point in V. Note now that by results in [Abate (1988a)] (see also [Kuczumow and Stachura (1990)]) for each strongly convex bounded C 2 domain V in C n the class of holomorphic self-mappings of T> has the Denjoy-Wolff iteration and the common fixed point properties. In addition, it is known [Abate (1992)] that each one-parameter semigroup of holomorphic self-mappings on V is differentiable with respect to the parameter. In this way we are led to a proof of Theorem 8.2. Remark 8.3 The following example shows that formula (8.34) is no longer true for the closure of V even in the case when f is continuous onV. Example 8.2 Let T> be the open unit disk in the complex plane C, i.e., V = {x e C : |x| < 1}. Consider f(x) — x - 1 + y/l - x. It is clear that / € Ho^X*, C) and that it is continuous on T>. In addition, Nul%/ = {0,1}. Ho"wever, the Cauchy problem (8.33) has the solution Ft : V ^ V, t>0
(8.37)
Ft(x) = l - [ l - e - ^ + e - ^ v / T 1 ^ ] 2 ,
(8.38)
defined by the formula
and for all £ > 0 we have Ft(l) = l - [ l - e - 5 f ] 2 < l .
(8.39)
Thus % ^ Nul%/. We will study boundary null points of generators later.
8.3
The Resolvent Method
Since in most situations the generator of a one-parameter semigroup satisfies the range condition, one can study the common fixed point set of the
Stationary Points of Continuous Semigroups
229
semigroup as the fixed point set of a single self-mapping defined by the resolvent of the generator. Lemma 8.2 Let T> be a domain in X, and let f : T> —> X satisfy the range condition on T>, i.e., the resolvent JT = (I + rf)~1 is well-defined for allr>0. Then for each r > 0, Fix c Jr = Nullp/. Proof. identity
(8.40)
Indeed, by definition, for each x £V and each r > 0 we have the Jr(x)+rf(Jr(x)) = x.
(8.41)
If a £ FixpJr, then we have by (8.41) f(a) = f{Jr(a)) = 0, i.e., a £ NUIID/. Conversely, if /(a) = 0, then (/ + rf)(a) = a for all r > 0. Thus J r (a) = (/ + r / ) - 1 ( a ) = a. D We now describe the resolvent method for the study of the null point set of generators. This method can be used to determine its structure, to solve the existence problem, as well as to find methods for approximating null points of generators. First we will trace a parallel with an implicit method for the study of the fixed point set of a single self-mapping of a domain. Let P be a bounded convex metric domain in X endowed with a metric p compatible with the convex structure of V (in particular, p may be the hyperbolic metric on V). Suppose that F : V —> V is a /)-nonexpansive self-mapping of V. For t e [0,1) and a fixed y &T>, consider the mapping $(x) = tF(x) + (1 — t)y, x e V. Since $ is a strict contraction on (V, p) (in particular, $ maps T> strictly inside itself), the Banach principle implies that there exists a unique fixed point z = zt(y) £ V of the mapping $. Moreover, zt(y) = lim $ n (y), n—>oo
(8.42)
where the limit is uniform on each p-ball in (T>, p). (For holomorphic mappings the same conclusion follows from the Earle-Hamilton theorem.) The X-valued function zt{y), 0 < t < 1, is called an approximating
curve. At the same time, changing our point of view, formula (8.42) implies that for each t £ [0,1), zt = zt{y) can be considered a /9-nonexpansive selfmapping of V depending on y 6 V (in particular, if F is holomorphic, then
230
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
zt depends holoraorphically on y G P). We denote this mapping by %. In other words, % is the unique solution of the nonlinear operator equation Tt=tFo%
+ {\- t)I.
(8.43)
The mapping % belongs to Np(V) (in particular, to Hol(2?)), and as we already know Fix(Ti) = Fix(F),
t e (0,1).
(8.44)
On the other hand, we know that the mapping g = I — F is a generator of a continuous flow of p-nonexpansive self-mappings of D. In addition, g : V —» X satisfies the range condition, i.e., for each r > 0 the resolvent Gr = (I + rg)-1 is a well defined p-nonexpansive (holomorphic) self-mapping of V. Claim 8.1 Gr = % with t = j ^ . Indeed, since Gr = (I + r(I - F))~x is the solution of the equation (/ + r(I - F)) oGr = I, we have (1 + r)Gr - rF o Gr = I or
Gr = -?—FoGr + -±-I.
(8.45)
1+r 1+r Setting t = jq^;, we get our claim by the uniqueness of the solution of this equation. Returning to the general case, let / : T> —> X be the generator of a continuous flow of p-nonexpansive self-mappings of V, and let Js = (I + s/)" 1 , s > 0, be its resolvent. Setting F — Js in our previous considerations, we see that for each r > 0 the mapping Gr = (I + r(I — J s ) ) - 1 is a well-defined p-nonexpansive self-mapping of V. The following relation is the key to our approach in the sequel. Lemma 8.3 Let f : V —> X be the generator of a continuous flow of p-nonexpansive self-mappings ofV, and let Ja = (I + s/)" 1 , s > 0, be its resolvent. Define GT = (/ -I- r(I - Js))-1 : V -> V. Then J(r+l)s = Js(Gr) = J.{I + r(I - Js))-1.
(8.46)
Proof. Fix s > 0 and set F = J3. By the previous claim we know that for each r > 0, the mapping Gr = (I + r(I — Ja))~x is a well-defined pnonexpansive self-mapping of V and for each x G V this mapping can be defined as the unique solution of the equation °r(X)
= 7^1
J'(Gr(x))
+ ^l*-
(8-47)
Stationary Points of Continuous Semigroups
231
On the other hand, since [I + r(I - Ja)]Gr = / and I-Js
— sf(Js) we have
[/ + rsf(Js)]Gr(x) = Gr{x) + rsf(Js(Gr{x)))
= x,
xGV.
(8.48)
Hence it follows from (8.47) that Js(Gr(x)) + (1 + r)sf(Ja(Gr(x))) = x,
xeV.
(8.49)
Since the equation z + (1 + r)sf(z) = x,
x€T>,
(8.50)
has the unique solution z = J(r+i)s(x) this means that
J(r+1)s(x)
= J3(Gr(x)).
(8.51J
Definition 8.6 Let 2> be a bounded convex metric domain in X endowed with a metric p compatible with the convex structure of V. We say that V satisfies the implicit approximation property if for each x € V and for each /J-nonexpansive self-mapping F of V its approximating curve {zt(x) : t 6 [0,1)} denned by the equation zt{x) = tF{zt{x)) + {\-t)x
(8.52)
is convergent to a point of T>, the closure of V. Proposition 8.5 Let V be as above and let f :V —» X satisfy the range condition, i.e., for each r > 0 the resolvent Jr = (I+rf)^1 is a well-defined p-nonexpansive self-mapping ofD. Then (1) For each x 6 T>, there exists a (= a(x)) € T> such that Jr(x) —> a and f(Jr(x)) —> 0 in the topology of X as r —» oo. (2) The following hypotheses are equivalent: (a) Nullx,/ ^ 0; (b) for some x G T>, the net {«7r(x)}, where Jr = (I + r / ) - 1 is strictly inside V; (c) for each x 6 V, the net {Jr(x)} is strictly inside V. Proof. Fix s > 0 and x e V. For r > 0, consider Gr{x) = (/ + r ( / J s ))~ 1 (a;). By our assumptions and the above claim, GT{x) —> a for some a (= a{x)) e 2?. In turn, by (8.47), we have that J,(Gr{x)) - Gr{x) -» 0,
r -> oo.
(8.53)
232
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
Together with (8.46) we get that for a given x G T>, lim Jr(x) = a (= a(x)) G V.
(8.54)
On the other hand, it follows by the definition of the resolvent that X
~ Jrr{x)
= f(Jr(x)),
r > 0.
(8.55)
Thus f(Jr(x)) —* 0 as r —> oo, and assertion 1 is proved. To prove assertion 2, we just have to show that (b) implies (a) and (a) implies (c) since the implication (c) =>• (b) is trivial. Indeed, if we assume that for some x G V the net {Jr{x)} is strictly inside V, then a = lim Jr(x) G V, and it follows from assertion 1 that a is a null point of r—>oo /•
Conversely, if / has a null point a G V, then for each x G V we have by (8.40), p(Jr(x),a) = p(Jr(x),Jr(a)) < p(x,a). Hence for each x G V, the net {Jr(x)} is /3-bounded in (T>,p), or, which is the same, this net is strictly inside T>. • Remark 8.4 Two standard examples of metric domains which satisfy the implicit approximating property are once again the open Hilbert ball and bounded convex domains in Cn endowed with the hyperbolic metric (see Chapters 4 and 5). In addition, it follows that Nullp/ is a p-nonexpansive retract of T>, and therefore, by Theorem 5 in [Kuczumow and Stachura (1990)] it is a metrically convex subset of (T>,p). We consider these two cases independently in the next section in the situation where a generator f has no interior null points in T>. 8.4
Null Point Free Generators
Let V be a bounded convex metric domain in X endowed with a metric p compatible with the convex structure of V. Assume also that V satisfies the implicit approximating property. Let / be the generator of a one parameter semigroup S = {Ft : t G (0, oo)} of /9-nonexpansive self-mappings of V. Suppose that / is null point free, i.e., the stationary point set W=
f)
FixpFt = Null©/ = 0.
(8.56)
0
In this case, we already know by Proposition 8.5 that for each x & V the net of resolvents {JT(x) (= (/ + rf)~1(x)) : t > 0} strongly converges to a
233
Stationary Points of Continuous Semigroups
boundary point a (= a(x)). We will now show that for the two cases when V is the open Hilbert ball or a bounded strongly convex domain in C n this point a G dT> does not depend on x G T> and is the sink point (Wolff point) for the semigroup {Ft : t G (0,oo)} generated by / . Now let X = H be a complex Hilbert space with the inner product (•,•), and let V = M be the open unit ball in H with the hyperbolic metric p. We recall that a boundary point a G dM is called the sink (or Wolff) point for a self-mapping F G NP(M) if all the ellipsoids E ( a , K ) = {x<= B : \ l - ( x , a ) \ 2 / ( I - \ \ x \ \ 2 )
K > 0,
(8.57)
internally tangent to dM at the point a are invariant under F. Theorem 8.3 Let f be the generator of a one parameter semigroup S = {Ft : t G (0,oo)} of p-nonexpansive self-mappings ofD and let f have no null point in B. Then there is a unique boundary point a G dM such that (i) for each x G P the net of resolvents {Jr(x) (= (/ + rf)~1(x)) : r > 0} strongly converges to a; (ii) for each r > 0 the point a G dM is the sink point for the resolvent mapping Jr : B —> B; (in) for each t > 0 the point a G dM is the sink point for the mapping Ft : M —> B, an element of the semigroup S. Proof. If N U I I B / = 0, then we claim that there exists a point a G dM such that for each x G B, Jr(x) strongly converges to a as r —> oo. Indeed, for each i £ l , the net {Jr(x)} is />-unbounded, i.e., ||J r (a;)|| —> 1 as r —> oo. Now fix ro > 0 and consider the mapping Jro : B —> B. Since Jro has no fixed points in B, it is known (see Chapter 5) that there exists a point a G dM (the sink point) such that all the ellipsoids
E(a,K)={x€B:\l-{x,a}\2/(l-\\x\\2)
K > 0, (8.58)
are invariant under Jro. But it follows from the resolvent identity Jrx - Jro ( ^ a; + ( l - - ) J r (z)) ,
(8.59)
r > ro, that E(a, k) is also invariant under Jr for all r > ro. NOW if for a fixed i £ l w e choose K > 0 such that x G E(a, K), then we get \l-(Jr(x),a)\2
(8.60)
234
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
and consequently, \l - (Jr(x),a)\ -» 0 as r -> oo. Hence Jr(x) converges strongly to a as r —> oo. If we now assume that for some 0 < r < TQ there exists K > 0 such that the ellipsoid E(a, K) is not invariant under Jr we get that there is another boundary sink point b G dM of Jr for such r G (O,ro). That is, all the ellipsoids E(b,K)={x£M:\l-(x,b)\2/(l-\\x\\2)
K > 0, (8.61)
are invariant under Jr. Once again, using the resolvent identity Jro(x) = Jrf—x+(l— ) Jro(x)) , (8.62) \f0 \ To/ J we get that all the ellipsoids E(b,K),K > 0, are invariant under JTQ. Clearly, we can find K > 0 such that the set M = E(a, K) n £(6, if) is not empty and is contained in B. But this set is a p-bounded subset of B which is also invariant under J r o . Thus Jro must have an interior fixed point in M. This is a contradiction. So, for all r > 0, all the ellipsoids E(a,K) are invariant under J r . Finally, by using the exponential formula Ft = lim UtT we see that these ellipsoids E(a,K) t > 0. The theorem is proved.
(8.63)
are also invariant under Ft for each •
Corollary 8.3 Let B be the open unit ball in a complex Hilbert space H, and let f : B —> H satisfy the range condition on B. If f has a continuous extension to M, then it has a null point in B. Proof. If / has no null point in B, then as we saw above there exists a sink point a G dM such that for each x G B, Jr(x) strongly converges to a, as r —• oo. But since f(Jr(x)) converges to zero as r —> oo, it follows by continuity that f(a) = 0 and we are done. D Remark 8.5 We will see below that, in fact, for holomorphic generators the following assertion is true: if f has no null point in B, then there is a point a G dM such that lim f{ra) = 0.
(8.64)
i—>i-
Moreover, we will see that for the finite dimensional case (when M is the open unit ball in Cn), if a £ 9B is the sink point of the semigroup S = {Ft •
Stationary Points of Continuoxis Semigroups
235
t G (0, oo)}, then there exists the so-called angular derivative Zf'(a) = P,
(8.65)
which is a real nonnegative number. It is clear that (8.65) implies (8.64). In addition, in this case, the semigroup S = {Ft : t G (0,oo)} strongly converges to a as t —> oo. Unfortunately, we cannot claim the latter fact in general. More details on the asymptotic behavior of semigroups in general Hilbert spaces can be found in Chapter 9. Here we show that in the particular case when the generator / has the special form / = / — T, where T is a self-mapping of B, then the semigroup 5 = {Ft : t £ (0, oo)} strongly converges to the sink point a G dM of the mapping T as t —> oo. Theorem 8.4
Let T be a p-nonexpansive mapping ofM.
(1) The Cauchy problem
| « « M +„(,,„ -T W l l I )) = ,
(866)
[u{0,x) = x has a unique solution {u(t, x) : t € R + } c B for each i £ l . (2) The semigroup {F(t)}t>o defined by F(t)x := u(t, x)
(8.67)
consists of p-nonexpansive mappings F(t), t > 0. (3) If T has no fixed point in B, then F(t) is also fixed-point-free for all t>0, and F(t)x converges as t tends to infinity to a point a G dM, the boundary ofM, for all x G B. Proof. First we note that the curve Gt = (1 — i)I + tT converges to / as t -> 0+ and -t (I - Gt) = I - T.
(8.68)
p(Gt(x),Gt(y)) < max[p(x,y),p(T(x),T(y))],
(8.69)
Since
236
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
each Gt is /j-nonexpansive. Hence it follows again by the product formula that for each t > 0 the mapping F(t) given by F(t) = lim G l n—>oo
n
(8.70)
is the solution operator for the Cauchy problem (8.66) and is also a pnonexpansive mapping on B. If now T has no fixed point in B, then it follows that
p | FixF(t) = 0.
(8.71)
t>o
By Theorem 8.1 we now see that each Ft has no fixed point in B. At the same time direct calculations show that the following formula holds:
F(t)(x) = e-*x + [ e-<*-«T(F(0(a:))de. Jo
(8.72)
\\T(F(0(x))\\ < 1
(8.73)
Since
for all £ € [0, t] and x € B, we can write F(t) = e-*I + (1 - e-*)A,
(8.74)
A = r J - j j * e-(*-«(T o F(0)d£
(8.75)
where
is a p-nonexpansive mapping on B. Hence Ft is an averaged mapping of the second kind and the result follows by Proposition 8.2 and Remark 8.2. D In the setting of finite dimensional Banach spaces it is known that each convex domain in C™ satisfies the implicit approximation property. If, in particular, V is a strongly convex C2 domain, and / £ Hol(£>, X) has no null point in V, then there exists a unique point a € &D such that for each x £ V, the net {J r (x)} converges to a, as r —> oo. Moreover, in this case the class of holomorphic self-mappings of V satisfies the Denjoy-Wolff iteration property, so the semigroup generated by / strongly converges to a boundary point b £ &D. The question is, of course, whether these two points o and b coincide.
237
Stationary Points of Continuous Semigroups
Theorem 8.5 Let V be a strongly convex bounded C2 domain in C" and let {F(t) : t > 0} be a semigroup of holomorphic self-mappings of T>. If {F{t)} has no stationary point in T>, then there exists a unique point a € dT> such that (a) For all x €T> the net {F(t)(x)} strongly converges, as t —> oo, to a. (b) For all x 6 V the net Jr{x) = I — rF'(Q)~l{x) strongly converges as r —» oo to the same point a. Proof. To establish our assertion it is enough to identify in our situation the points a and b obtained in Propositions 8.4 and 8.5. Setting / = —F'(0), we consider again the resolvents Js = (I + s/)" 1 , s > 0, which are holomorphic self-mappings of T>. Since / has no null points in ~D, J3 has no fixed point in V and as we saw above, the net {G r (0)} defined by formula (8.47) converges to a € &D as r -> oo (Gr(0) = ^ j J,(G r (0))V It follows from the proof of Theorem 2.3 in [Abate (1988a)] that for each n € N and for every x € T> and R > 0, the following inclusion holds: (8.76)
J?{Ex(a,R))cFx{a,R),
where Ex(a, R) and Fx(a, R) are the small and the big horospheres of center a, pole x, and radius R defined by Ex{a,R) = \z GV :limsup[p(z,u;) - p(x,w)] < \
loS#).
(8-77)
Fx(a,R) = [z e V : limmf [p(z,w) - p(x,w)] < ^ logi?},
(8.78)
where p is the hyperbolic metric on V. Therefore it follows again by the exponential formula (8.63) that for all t > 0, F(t)(Ex(a,R))
C Fx(a,R).
(8.79)
Hence, if 6 = lim F(t)x, then b = a and we are done. t—>oo
8.5
n
The Structure of Null Point Sets of Holomorphic Generators. Retractions
By Nullp/ we denote the analytic set defined as the null point set of / € Hol(P, X). Even in the finite dimensional case it is a complicated problem
238
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
to recognize when an analytic set N consists only of irreducible components (see, for example [Aizenberg and Yuzhakov (1983)] and [Chirka (1985)]). It is known that this is the case when N is locally a complex analytic manifold. By using the above approach one can study the null point set of a generator / as the common fixed point set of the commuting family {-Ft}, the semigroup generated by / . Indeed, if D is a bounded convex domain in C n and {Ft} is a semigroup of holomorphic mappings, then it is well known ([Abate and Vigue (1991)]) that the commonfixedpoint set of this commutative family is a holomorphic retract of V. This approach becomes less transparent if X is an infinite dimensional space. Equality (8.40) and the Mazet-Vigue theorem [Mazet and Vigue (1991)] (for example) immediately imply that the null point set of a holomorphic generator on a convex bounded domain in a complex reflexive Banach space is also a holomorphic retract of T>. (This, in turn, implies that Nullp/ is a connected submanifold of V [Cartan (1986)].) Theorem 8.6 Let V be a convex bounded domain in a complex Banach space X and let f € QHol(T>). Suppose that a G Nullp/ and that one of hte following hypotheses holds: (1) X is reflexive, (2) KerA © Im A = X, where A = f'(a). Then Nullp/ is a connected complex analytic submanifold in T>, which is tangent to KerA Proof. It is sufficient to note the a G Null©/ = FixpJr(f) for all r > 0, where FixpJ r (/) is the fixed point set of the resolvent Jr{f) of / in V. In addition, it follows by the chain rule that / — Jr(A) — rJr(A), where Jr{A) = [Jr(f)]'(a) is the resolvent of the linear operator A for r > 0. Thus Ker(/ - [Jr(f)]'(a)) = KerA, and the theorem follows from the MazetVigue theorem. • Corollary 8.4 Let V be a bounded convex domain in a reflexive complex Banach space X, and let {Ft : t £ (0, oo)} be a locally uniformly continuous semigroup of holomorphic self-mappings of V. Then W=
p | F\xT,Ft 0
is a holomorphic retract ofT>.
(8.80)
Stationary Points of Continuous Semigroups
239
Thus W is a connected analytic submanifold o}T>. Moreover, ifD = B is the open unit ball in a Hilbert space H, then W is an affine submanifold ofB. In the context of the above theorem we will also need the following notion. Definition 8.7 Let / be a holomorphic mapping from V into X and let W = Null-p/ ^ 0. A point a € W is said to be quasi-regular if the condition Kerf (a) © Im /'(a) = X
(8.81)
holds. If, in addition, Ker/'(a) = {0}, then we say that a is a regular null point of / . Corollary 8.5 Let T>,X and f be as in Theorem 8.6. If a € V is an isolated point o/Nullx>/, then it is unique. In particular, if a £ Nullp/ is regular, i.e., f'(a) is invertible, then a is unique. Thus from Theorem 8.6 we can obtain the global description of the (interior) stationary point set of a semigroup {Ft}, t > 0, generated by a holomorphic mapping. But even in this case we only know that a retraction exists, but we have no constructive approximation process for finding the points in W. So, the question is how to construct a retraction onto this set. A possible way is to investigate the asymptotic behavior of the semigroup as t —> oo. It will become clear that this may be done only if we know a priori at least one point a EW and the spectrum of the linear operator f'(a) satisfies certain conditions (i.e., it does not intersect the imaginary axis with, perhaps, the exception of zero). Another way would be to apply equality (8.40), and for a fixed r > 0 to construct the sequence of the discrete Cesaro averages
G" = ^ £ ^ ' j=0
(8-82)
so that a subsequence {Gnic} weakly converges to a mapping G : T> —» W which is a holomorphic retraction of V onto W. As a matter of fact, this method turns out to be supefluous because as we will see below, the iterates of the resolvents Jr strongly converge to a holomorphic retraction of V onto W.
240
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
Theorem 8.7 Let V be a bounded convex domain in X', and let f e Hol(D, X) be the generator of a one-parameter semigroup of holomorphic self-mappings ofD. Suppose that W = Nullp/ ^ 0. Then (i) If W contains a quasi-regular point a G T>, then for each s > 0 the sequence {J™ = (I + sf)~n}f converges to a holomorphic retraction ofD onto W, as n —> oo, uniformly on each ball strictly inside T). (ii) IfW contains a regular point a € V, then W = {a} and the net {Js — (/ + s/)~1}s>o converges to a as s —> oo, uniformly on each ball strictly inside T>. To prove our theorem we need the following lemma. Lemma 8.4 (cf. [Lyubich and Zemanek (1994)]) Let A be a bounded linear operator on a Banach space X such that \\{I-A)n\\<M,
n=l,2,...
(8.83)
for some M < oo. Then the following conditions are equivalent: KerA©ImA = X
(8.84)
I m i = ImA
(8.85)
Proof. The implication (8.84) => (8.85) is obvious. Now let (8.85) hold, and let a functional x* £ X* vanish on the sum KerA © ImA Then x* £ KerA*. Furthermore, it follows from (8.85) and the Banach-Hausdorff theorem that the condition (u, x*) — 0 for all u € KerA implies that x* £ ImA*. Thus x* £ KerA* n ImA*. But because of (8.83), KerA* n ImA* = {0} by the Yosida mean ergodic theorem ([Yosida (1974)]). So x* = 0, and this implies (8.84). We recall that a linear operator A : X —> X is said to be m-accretive if for each r > 0, the operator Xr = (I + rA)^1 is well denned on X and • ||(7 + rA)~l\\ < 1. Lemma 8.5 Let A be a bounded linear operator on X which is maccretive with respect to some norm equivalent to the norm of X. If A satisfies condition (8.84), then for each r > 0, the linear operator Xr = (I + rA)~l satisfies the condition Ker(7 - I r ) © Im(7 - Ir) = A.
(8.86)
Stationary Points of Continuous Semigroups
Proof.
241
Returning to the original norm of X, we have by the definition ||2?||<M
71=1,2,...,
r>0.
(8.87)
By Lemma 8.4 and (8.87), it is sufficient to show that Im(/ — I r ) is closed in X. Indeed, if yn G Im(7 — lr) converge to y G X, we get a sequence {xn} C X such that (/ - lr)xn
= rAIrxn
-> y G ImA.
(8.88)
Note that it follows by the definition of J r that Ker(7 - lr) = KerA.
(8.89)
Therefore, if we represent xn £ X'm the form xn = un+vn, where un G ImA and vn G KerA (see (8.84)), we get from (8.88) and (8.89), (/ - lr)un = r(Alrun + AIrvn) = r(Alrun + Avn) = rAIrun -^>ye ImA Denote zn = lrun
(8.90)
G X. We have by (8.90), Zn — un - vAzn
(8.91)
and hence zn G ImA. Since ImA is closed and invariant under A, and Azn —» p y G ImA, the sequence {z n } converges to some element z € X and hence un = zn + rAzn —> z + y = x. Once again, it follows by (8.90) a that (/ - Ir)x = y. Lemma 8.6 Let the conditions of Lemma 8.5 hold. Then er(Xr), the spectrum of the operator IT, is contained in the open unit disk A, except perhaps, for 1, i.e.,
(8.92)
Proof. Fix r > 0. It follows by (8.87) that
(8.93)
242
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
It follows by the spectral mapping theorem (see, for example, [Dunford and Schwarz (1958)]) that a(f(A)) = f(a(A)). Thus, if we assume that ei
Proof of Theorem 8.7 (i) Let / € Hol(D,X) be a generator and let {-Fi}t>o be the semigroup of holomorphic self-mappings of V generated by / . Suppose that W = Nullp/ = f) FixpF t contains a quasir>0
regular point a e T> (see Definition 8.7). Observe that the linear operator A = f'(a) is the infinitesimal generator of the semigroup {Ut}t>o, where Ut = (Ft)'(a). Since V is bounded, it follows by the Cauchy inequalities that {Ut = e.~At}t>o is uniformly bounded and therefore A is an maccretive operator with respect to some norm equivalent to the norm of X. In addition, for each r > 0, (/ + rA)~l = [{I + r / ) - 1 ] ' ^ ) by the chain rule. Thus by applying the Vesentini theorem [Vesentini (1985)] and Lemmata 8.4-8.6, we see that for each r > 0, the sequence {T^jJLj, where TT = (I + rf)~l :T> —» D converges locally uniformly to some holomorphic mapping
= |T r (^+(l-^)T t (x))-T r (a)|| i
\
1
V '
(8.94)
Thus we obtain the inequality ||Tt(x) - al^ <
^—p-
L-qr{L-
j)
\\x - ah,
which implies that %{x) converges to a as t —> oo, uniformly on
(8.95) BR(Q).
243
Stationary Points of Continuous Semigroups
Now it follows by the Vitali property that {Tt}t>o converges to a, locally uniformly on all of V. The theorem is proved. • Another retraction method is based on the following observation. Let S = {Ft}t>o be a uniformly continuous semigroup with T = N u l b / = f| FixFt ^ 0. 0
Pick o 6 f and denote At = (Ft)'(a), t > 0. It follows from the Cauchy inequalities that {At}t>o is a uniformly continuous semigroup of bounded linear operators, i.e., At converges to the identity in the operator norm as t —> 0 + . Therefore it has an infinitesimal generator B : X —> X, i.e., I-—^. (8.96) lim t K ' t-o+ Moreover, since the infinitesimal generator / : V —» X of S = {Ft}t>o is defined by the formula
B=
/ =t
ilm+^'
(8-97)
we have f'(a) = B. Now we can state our assertion. Theorem 8.8 Let V and S = {Ft}t>o be as above. Suppose that for some s e f = Dt>oFix(Ft), the following condition holds: KerB ® ImB = X,
(8.98)
where B is defined by (8.96) with At = {Ft)'{a). continuous Cesdro average
Then, for each t>0,
the
1 f*
$t = T / F3ds * Jo is power convergent to a retraction onto T. manifold ofT>.
(8.99) Thus T is a connected sub-
Proof. It follows from (8.99) that <J>t(a) = a for each t > 0, and hence $t G Hol(D). In addition, ($t)'{a) = Ct, where Ct is denned by
Ct = \ ( Asds = \ f e~Bsds. t Jo t Jo
(8.100)
fl = { A 6 C : ReA > 0},
(8.101)
Therefore, if we set Ao = 0,
244
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
and 1 - e~tX ft{\) = — ^ — ,
t>0,
(8.102)
we get that Ct =
^df
/t(A)
~B)~ldX'
(8-103)
where T is a contour that surrounds o(B) in fte = {A € : ReA > - e } for some e > 0. Furthermore, it follows by the maximum modulus principle that /(ft\{0}) C A and hence A = 0 is the unique and simple root of the equation /t(A) = 1 in fi. So, Ct is power convergent to a projection onto KerB. But T = f| Fix(Ft) coincides with the null point set of / t>o in V, where / is the generator of S = {Ft}t>o and f'(a) = B. Since ($t)'(a) = Ct, $t is power convergent to a retraction onto Fix($ t ), which is tangent to KerB. Together with the inclusion T C Fix($f) this implies • the equality T = Fix($ t ), which proves our assertion. 8.6
A Stabilization Phenomenon
Here we establish another interesting feature of the null point set of a holomorphic generator and then use it to construct holomorphic retractions [Khatskevich et al. (1998)]. Let us consider a semigroup S = {Ft}, t € M+, generated by / S Hol(X>, X). Let ft be the difference approximations of / , i.e., ft = j(I-Ft),
t>0.
(8.104)
If Nullp/ is not empty, then Nullp/ C Nulb/ t ,
t > 0.
(8.105)
Moreover, it is natural to expect that for sufficiently small t, the set Nullp ft will approximate Nullp/ in some sense. As a matter of fact, in the linear case, as well as in the holomorphic case, there is a stabilization phenomenon of Nullp ft for sufficient small t. Theorem 8.9 Let S = {Ft}t>o be a uniformly continuous semigroup generated by f € Hol(P, X) and let ft = t~l(I-Ft). Suppose that Nullp/ ^ 0 and that one of the following conditions holds:
Stationary Points of Continuous Semigroups
245
(1) X is reflexive; (2) Ker/'(a) © Im/'(a) = X for some a e Nullp/. Then there exists S > 0 suc/i t/iat /or all t £ (0,6), FixpF t (= Nullp/t) = Nullp/.
(8.106)
Proof. Since both Nullp/ t and Nullp/ are connected complex submanifolds of V and Nullp/ C Nullp/t,
(8.107)
it suffices to show that their tangent spaces coincide. A simple calculation shows that for a e Nullp/, (ft)'{a) = £ (/ - e~ M ), where A = f'(a). Thus our claim is that there exists a positive 5 such that for all t £ (0,S) Fix(e~ M ) = KerA
(8.108)
In order to prove (8.108) when X is reflexive, we first note that the semigroup e~tA = (Ft)'(a) is uniformly bounded by the Cauchy inequalities. We then let P denote the projection of X onto KeiA obtained from the mean ergodic theorem. Now let gt = \ Jo e~sAds. There is a positive S such that gt is invertible for all 0 < t < 6. For such t, let Pt be the mean ergodic projection onto Fix(e~tA). A computation shows that for all natural numbers m
*** = ( - f y °-i)MV
(8-109)
Letting m —> oo, we see that P = Ptgt = gtPt- Hence Pt = g^1P and KerP C KerP t . Since X = FixP © KerP = FixPt © KerPt
(8.110)
and FixP C FixP t , it follows that KerA = FixP = FixPt = Fixe~ M . When hypothesis (2) holds, the following simple direct argument is due to V. Khatskevich. In this case there is a positive e such that ||Az||>e||*|| for all z e ImA
(8.111)
246
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
Let x = y + z, where y e Ker.A and z 6 ImA, belong to Fix(e~tA). Then e~tAz = z and , 42
f\z
(8.112)
If 0 < t < min{l,£/(ellAH - 1 - \\A\\)} and z ^ 0, it follows that ||Az|| < e||z||, which contradicts (8.111). Hence z = 0 and x = y belongs to KerA
•
Theorem 8.10 Let T> be a bounded convex domain in a complex Banach space X, and let f :T> —> X be a holomorphic generator of a flow {Ft}t>o onV. Suppose that T = f] FixFt is not empty and that a e T. Then {Ft} t>o converges, as t —> 00, in the topology of locally uniform convergence over D to a retraction p : T> —> T if and only if one of the following conditions holds: either (i) ReA > e > 0 for all X <E cr(/'(a)), or (ii) {A £ C : ReA = 0}n
ofa(f'(a)),
and
Ker(/'(a)) 0 Im/'(a) = X.
(8.113)
In particular, when X is finite-dimensional, {Ft} converges uniformly on compact subsets of V to a retraction p : V —> X if and only if the spectrum of f'(a) does not intersect the imaginary axis except perhaps at the origin. Note also that in the presence of (i), T = {a}, and a is an asymptotically stable stationary point of {Ft}. Proof. Let T> and / be as in the statement of the theorem, let S = {Ft}, t > 0, be the semigroup generated by / , and choose a G T = p | FixFt ¥= 0t>o
(8-114)
Denote At = (Ft)'x(a), t > 0, and B = f'{a). It is easy to see that At = exp(-tB), t>0. Suppose now that one of the hypotheses, either (i) or (ii), holds. Then using the Cauchy inequalities and direct computations,
Stationary Points of Continuous Semigroups
247
we see that At satisfies either condition (i) or (ii). In addition, for t > 0 small enough, we have Ker(7 - At) @ Im(J - At) = X.
(8.115)
Indeed, it follows from Theorem 8.9 that Ker(J-v4t) = Ker£(= f| FixAt) v t>o ' for each t > 0 which is small enough. Once again, the Cauchy inequalities imply that B = lim \ (I — At) in the uniform operator topology. Finally, each At commutes with B, i.e., AtB = BAf. Thus Im(7 — At) = lmB. Now for such t > 0 the mapping Ft satisfies all the conditions of Vesentini's theorem (see also Lemma 5.2 in [Mazet and Vigue (1991)]). Thus we can conclude that for a small enough fixed r > 0, say r € (0,6), the sequence {FnT} converges in the topology of locally uniform convergence over V to a mapping pT : V —> T>, which is a retraction onto the set FixFT = J7. Taking T = \ with q a big enough integer, we deduce that {Fn} = {FnqT} converges to pT as n —• oo in the topology of locally uniform convergence. Hence pi is independent of such q. We denote this retraction by q
P-
Let Bi CC 52 CC 2? be two arbitrary balls strictly inside V. Choose <7i large enough, so that Fs(x) £ Bi for all x £ Bi and 0 < s < — < 6, and choose M such that {Ft} is Lipschitzian on Bi with constant M. We claim that {Ft} converges to p uniformly on Bi. Indeed, given e > 0 we first choose q2 > q\ such that ||-Fs(x) - a;|| < e/2M for all 0 < s < ^ and x e JBI . Now for k — [tq^] and for 0 < s = t — — < — we have \\Ft{x) -p{x)\\ = \\F±(F.(x))
-p(x)\\
< \\F±(Fs(x)) - F i ( x ) | | + \\Fj,(x) - p(x)\\ <|
+
£=
e >
(8.116)
wherever t (hence A;) is large enough. Conversely, let the net {Ft} converge in the topology of locally uniform convergence to a mapping p : V —> V. Then, by Vesentini's theorem, we have that Ft satisfies the conditions analogous to either condition (i) or condition (ii) for each fixed t > 0. Repeating the above considerations and using the spectral mapping theorem we obtain either (i) or (ii) (with (8.113)), respectively. This concludes the proof of the theorem. • Theorem 8.11
Let V be a bounded convex domain in a complex Banach
248
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
space X, and let F : T> —> T> be a holomorphic self-mapping such that F i x F ^ 0 . Then (1) the mapping f = I — F is the generator of a flow on V; (2) if for some a £ FixF, Ker(J - F'{a)) © Im(I - F'{a)) = X,
(8.117)
then {Ft}, the semigroup generated by f, converges, as t —» oo, to a retraction p : V —» FixF in the topology of locally uniform convergence over T>. Proof. 1. Let V and F : V -> V satisfy the hypotheses. Then / = / — F is bounded and has a null point in T>. for arbitrary r > 0, setting t = r(r + I ) " 1 we have that the equation x + r(x — F(x)) = z is exactly the equation x = (1 — t)z + tF(x), which has a unique solution x = xr(z) for each z £ V and t £ (0,1) by the Earle-Hamilton theorem because T> is convex. Since this solution may be obtained by iteration, it is holomorphic in z £ V. Thus by [Reich and Shoikhet (1996)], we see that I - F is a generator of a flow {Ft}t>o2. We already know that cr(F'(a)) c A. Therefore a{f'(a)) = <j{I — F'{a)) satisfies either (i) or (ii). The result now follows from Theorem 8.10.
•
8.7 8.7.1
Local and Spectral Characteristics of Stationary Points Cartan's uniqueness theorem
The following simple consequences of the above results indicate that some local characteristics of a null point of a generator can influence the global structure of the whole null point set and the global behavior of the semigroup. Theorem 8.12 Let D be a convex bounded domain in X, and let f be the holomorphic generator of a uniformly continuous one-parameter semigroup and have a null point a £V. If f'(a) = 0, then / = 0. Proof. Indeed, it is clear that a £ Nullp/ is a fixed point of the resolvent Jr = (I + rf)-1 £ Hol(P), r > 0. In addition {Jr)'{a) = I is the identity mapping on X. Thus, by Cartan's theorem (see, for example, [Franzoni
Stationary Points of Continuous Semigroups
249
and Vesentini (1980)], [Khatskevich and Shoikhet (1994a)]), Jr = Iv is the identity mapping on V. This implies that / = 0 in V. Moreover, we can establish a continuous form of this assertion, it is a generalization of the Harris-Schwarz Lemma [Harris (1971a)]. • Theorem 8.13 Let T> be a convex bounded domain in X, and let {fn} C Hol(T>, X) be a sequence of holomorphic generators which is uniformly bounded on each subset strictly inside T>. Assume that for some a G V, the following conditions hold: (a) {/n(a)} strongly converges to zero; (b) {fn{a)} converges to 0 in the operator topology. Then {/„} T-converges to 0, i.e., T- lim /„ = 0. n—>oo
8.7.2
Harris' spectrum of a semi-complete vector field
Following L. A. Harris [Harris (1971b)] we give the following definition. Definition 8.8 Let V be an open subset of X, a G T>, and let h € Ho\(D,X). The spectrum of h with respect to a, denoted by o-a(h), is the set of all A G C such that it is not possible to find open sets U C D, with a £ U and V C X with the property that XI — h is a biholomorphism of U onto V. Proposition 8.6 ([Harris (1971b)]) aa{h) = a(h'(a)) is the spectrum of the linear operator h' (a). Theorem 8.14 Let f G Ho\(T>, X) be the holomorphic generator of a one-parameter semigroup on T> and let a G Nullp/. Then (1) cra(f) lies in the right half-plane; (2) ifO £ aa{f), then a is the unique null point of f in V; (3) o~a{f) lies strictly inside the right hall-plane iff a is a globally asymptotically stable (in the Lyapunov sense) stationary point of the semigroup Sf = {Ft}, t > 0, i.e., {Ft} converges to a locally uniformly in T> as t->oo. Proof. Set A = f'{a). It is easy to see that A is the infinitesimal generator of a uniformly continuous semigroup Ut = e~tA and that Ut = (Ft(x))'x=a. Thus it follows by the Cauchy inequalities that Ut is a uniformly bounded semigroup of linear operators, it is well known that the
250
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
resolvent R(X,A) = (XI — A)~l is defined on the open left half-plane, i.e., ReA > 0 for all A G cr(A). Thus assertion (1) follows from Proposition 8.6. (2). If 0 ^ cr a (/), then the operator A — f'(a) is invertible. Hence a is an isolated null point of / in V, and our assertion follows from Corollary 8.5. (3). Suppose now that aa(f) = cr(A) lies strictly inside the right halfplane of C. As it is well known, this fact implies the estimate ||e- M || < Ne~vt
(8.118)
for some N > 0 and v > 0 (see, for example, [Daletskii and Krein (1970)], [Yosida (1974)]). Rewrite now the Cauchy problem in the form of a perturbed equation: x'(t) = -Ax(t)+g(x(t)), x(0) = xeV,
(8.119)
where g = A — f. Since /(a) = 0, there is some ball BT(a) CC V centered at a with radius r, such that g admits the representation
9(s) = JT/PJk)(a)o(x-a),
(8.120)
fc=2
where Pf , k > 2, are homogeneous forms of order k. Setting M = sup ||g(x) ||, we have, by the generalized Schwarz Lemma, ||5(z)||<Mr-2||z-a||2 for all x £ Br(a). Choosing now p < ry/u(MN)~1, in (8.118), we obtain the inequality ||<7(i)||<^||a:-a||
(8.121) where v and N are as (8.122)
for all x £ Mp(a) = {x G V : \\x - a\\ < p}. Thus Theorem VII.2.1 from [Daletskii and Krein (1970), p. 403] implies that problem (8.119) has a uniformly asymptotically stable solution on Bp(a) x R + . In other words, the net Ft\B,,(a) = x(t) converges uniformly to the point a, uniformly on Bp(a). An appeal to Vitali's property concludes now the proof of our assertion in one direction. Conversely, let {Ft}t>o T-converge to a G Nullp/. Then it follows from the Cauchy inequalities that the linear semigroup Ut = e~tA = (Ft)'x=a
Stationary Points of Continuous Semigroups
251
uniformly converges to zero as t —> oo. This is equivalent to that fact that for all t > 0, the spectral radius rff{Ut) < 1. By Dunford's theorem on the spectrum, it follows that a(A) = cra(f) lies strictly inside the right half-plane and we are done. • Definition 8.9 Let V be a domain in X and let / £ Hol(X>, X). A point a € Nullp/ is said to be regular if 0 ^ cr(/'(a)), i.e., /'(a) is an invertible linear operator. It is said to be strictly regular if (A) does not intersect the imaginary axis of the complex plane C According to this definition we obtain the following direct consequence of Theorems 8.14 and 8.6. Corollary 8.6 Let V be a bounded convex domain in X, and let f be a bounded semi-complete vector field in V. Suppose that f is a Fredholm mapping and that a G Nullp/. Then NUIID/ = {a} if and only if the point a is regular. Remark 8.6 If f is not Fredholm, but X is reflexive, we have, in general, two singular situations. Namely, if a £ Null©/ and 0 G o~a(f), then either (1) a is the unique null point in V, or (2) there are infinitely many null points of f in T>, and they form a connected complex submanifold ofT>. The following example shows that situation (1) actually may exist in the case of an infinite dimensional space (even if it is reflexive). Despite its uniqueness, such a point has no "good" property such as regularity. Example 8.3 Let X be the complex Hilbert space I2 with basis {ei}^, and let 0 < Oj < 1 satisfy a, —> 1 as i —> oo. Let V be the unit ball in X and define the linear mapping A : V i-» X by Ae, = (1 — a,)ej. This mapping has a unique null point x = 0, but it is not regular (0 is a point in the continuous spectrum of A). It is clear that A is the generator of a semigroup of self-mappings of V. Now we turn to the same questions concerning the approximation of fixed points. Remark 8.7 Let V be as above, and let F : V *-> V be a holomorphic self-mapping of V. Its iterates Fn : V i-> V, Fn = F " " 1 o F, n = 1,2,..., F° = I, are well defined and holomorphic. However, even when X is finite dimensional and F has a unique fixed point, there are many
252
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
situations when the sequence of iterates {Fn(x)}%L0 does not converge to the fixed point a for x ^ a. For example, let V be a unit ball and F = ellfil, 0 < ip < 2n. More generally, such a situation arises when the spectrum
f-^=F(F t )-F t , I Fo = h,
(gi23)
T-converges to a when t tends to infinity. As a simple example, consider again the mapping F = il mentioned above, the iterates of which do not converge to zero for each x ^ 0. At the same time, the Cauchy problem (8.123) has the solution Ft(x) = elt • e~tx which evidently uniformly converges to zero as t tends to infinity.
Chapter 9
Asymptotic Behavior of Continuous Flows 9.1
Strongly Semi-Complete Vector Fields in Banach Spaces
Let V be a domain in a Banach space X and let / be a semi-complete vector field on a domain V in X with Nullp/ =fc 0, i.e., the Cauchy problem
{
du
...
¥ + «»>=°
(9.1)
u(0) = x has a solution u(-, x) : 1R+ —> V which is well-defined on all of R + for each initial datum x 6 T>. In other words, / is the generator of a one-parameter semigroup S = {F(t)}t>o of self-mappings F(t) = u(t, •) e Hol(P). Definition 9.1 A point a G Nullp/ is said to be locally uniformly attractive if the semigroup S = {F(t)}t>o generated by / converges to a in the topology of locally uniform convergence over V. Definition 9.2 Let V be a domain in a Banach space X and let G(V) be the family of all semi-complete vector fields on V. A mapping / e Q{V) is said to be a strongly semi-complete vector field if it has a unique null point in V which is a locally uniformly attractive fixed point for the semigroup generated by / . Let a{A) denote the spectrum of a bounded linear operator A : X —» X. It is known that if T> is a bounded domain, then / £ Q^P) with / ( T ) = 0, r € V, is strongly semi-complete if and only if there is e > 0 such that ReA > e > 0 for all A 6 cr(/'(r)). Such a point r is sometimes said to be strictly regular . 253
254
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
In this section we will give several sufficient conditions for / 6 Hol(2?, X) to be strongly semi-complete on the open unit ball T> of X and obtain rates of convergence for the semigroups generated by such mappings. In this context boundary flow invariance conditions are seen to be quite useful. However, it does not seem natural to consider only boundary conditions because there are many examples of semi-complete vector fields denned on a domain D which have no continuous extension to T>. Since we will mainly concentrate our discussions on domains which are biholomorphically equivalent to a ball, we will assume in the sequel that T> is the open unit ball of a complex Banach space X. If X = H is a Hilbert space with the inner product (•,•), then we will use the letter B to denote its open unit ball. This will enable us to point out special features of semicomplete vector fields in this case. Our approach to the search for different (but equivalent) characterizations of the class of (strongly) semi-complete vector fields on T> is based on the following lemma. Lemma 9.1 Let V be the open unit ball in a complex Banach space X and let f S Hol(Z>, X) satisfy the inequality (9.2)
Re(f(x),x*)>a(\\x\\)\\x\\
for all x &T> and some x* € J(x), where a is a real continuous function on [0,1) such that for all /J, £ [0,1) and for all r > 0 the equation s + Xa{s) = \x
(9.3)
has a unique solution s(fi) in [0,1). Then (i) f is a semi-complete vector field on T>; (ii) if /3(t, s) is the solution of the Cauchy problem
(9.4) [0(0,8) = 8[0,l) and u(t,x) is the solution of (9.1) then the following estimate holds:
||u(t,x)||< / 9(t 1 ||i||), Proof.
xeV.
(9.5)
Fix r 6 [0,1) and A > 0, and consider the equations x + Xf{x) = y
(9.6)
255
Asymptotic Behavior of Continuous Flows
and (9.7)
s + Xa(s) = \y\,
where y 6 T>r = {a; € X : ||a;|| < r < 1} and s € [0,1). It follows from our assumption that equation (9.7) has a unique solution SQ = so(y) £ [0,1). Setting -y(s) = s + Xa(s) - \y\ and 6 > 0, we can find e > 0 such that 7(s + 6) >e. Taking x 6 V such that ||a;|| = s = SQ + 6, we have, by (9.2), for such x and some x* € J(z), Re(x + \f[x) - y, x*) > s2 + Aa(s)s - ||y||s = sj{s) > s • e.
(9.8)
Now it follows from Theorem 3 in [Aizenberg et a/. (1996)] that equation (9.6) has a unique solution x = x(y) such that ||a:(y)|| < So + 6. Since S is arbitrary, we must have
(9.9)
Mv)\\ < so-
In terms of nonlinear resolvents the latter inequality can be rewritten as
WMyn^il
(9.10)
+ Xar'Wyl
Now we obtain our assertion by [Reich and Shoikhet (1996)] and the exponential formula given there. • As a consequence of Lemma 9.1 we also get the following assertion [Elin et al. (2004)]. Theorem 9.1 Let T> be the open unit ball in an arbitrary complex Banach space X, and let f € H.o\(V,X). Then (1) f is semi-complete on V if and only if (i) it is bounded on each subset strictly inside T>; (ii) A = /'(0) is an accretive linear operator on X, i.e., Re(Ax,x*) > 0 , x€X,
X* e J(x),
(9.11)
and (Hi) for each x € V and x* € J{x), the following inequality holds:
Re ( Y ^ § A0)* + (! - M2)/(°). *')
< Re ( f r f | f'®x
+
^~ IN 2 )/(°)' x *) • (9-12)
256
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
(2) If {F(t)}t>o is the semigroup of holomorphic self-mappings ofT> generated by f, then it satisfies the following estimate: \\F(t)(x)\\<0(t,\\x\\,a,m),
(9.13)
where m = inf Re(/(O),a;*), a = inf (/'(0)a, a*), and P(t,s,a,m) ll*ll=i ll*ll=i satisfies the following algebraic equations: _
&
r-at
(I-/?)2
m
S
_
(1-s)2
0
.
U'
(1-/3) 2
(1-s)2 a ^ -4m, m ^ 0.
(9.14)
Now if /(O) = 0 we have m = 0, and we get that {F(t)}t>o satisfies the following explicit estimate which gives a rate of convergence of {F(t)}t>o to the origin, its stationary point. Corollary 9.1 Let f £ Q{V) be such that /(0) = 0 and A = /'(0) is accretive with Re(Ax,x*) > fc||a;||2, k > 0. Suppose that {F(t)}t>o is the semigroup generated by f. Then the following estimates hold: (t)
||F(t)a;|| < I l i l l a " * ^ ! * ,
{n)
(1 - \\F{t)x\\Y
\\F(t)x\\
kt
x £ V, t>0;
(9.15)
\\x\\
(1 - llxll)2 •
(9J6)
Proof. Both estimates follow directly from Lemma 9.1 (or Lemma 9.2) if we set a(s) = ks —- . 1+ s
(9.17)
(3(t,s)<se-kT^t,
(9.18)
In this case
where {/3(t, -)}t>o is the real-valued semigroup generated by a. • Remark 9.1 The estimate (i) is due to Gurganus [Gurganus (1975)], while (ii) was obtained by Poreda [Poreda (1987)]. Note that the condition /(0) = 0 is essential in their considerations as well as in our approach above.
Asymptotic Behavior of Continuous Flows
257
Remark 9.2 For the case of Hilbert space we will show below how more general estimates can be obtained when f has an arbitrary null point which is strictly regular . Remark 9.3 If k in Corollary 9.1 is positive, then the operator A is strongly accretive, hence its spectrum lines strictly in the right-half plane. Thus in this case f is strongly semi-complete. The question is whether a one-sided estimate or a flow invariance condition could recognize f a priori to be strongly semi-complete. We now give some sufficient conditions for f to be strongly semicomplete on the open unit ball T> of X, and obtain rates of convergence for the semigroups generated by such mappings in terms of a metric on T> assigned to it by a Schwarz-Pick system. We recall in passing that for a bounded convex domain T> in X, all metrics assigned to it by a Schwarz-Pick system coincide. We call this unique metric the hyperbolic metric on T>. Lemma 9.2 Let V be the open unit ball in X and let f 6 Hol(I>, X) satisfy the condition Re(f(x),x*)>a(\\x\\)-\\x\\,
xeV,
x*€J(x),
(9.19)
where a is a real continuous function on [0,1] such that a(l) = w > 0.
(9.20)
Then (i) f is strongly semi-complete; (ii) if {Ft}t>o is the semigroup generated by f, then for each pair of points x and y in T>, the estimate p(Ft(x), Ft{y)) < e " * lp{x, y), where p is the hyperbolic metric on V, holds. In particular, the null point of f, then p(Ft(x),T)<e-^tp(x,r)
(9.21) ifrGVis (9.22)
for all x € V. Remark 9.4 Lemma 9.2 is different from Lemma 9.1 because we impose different conditions on the function a. However, the proof of Lemma 9.2 is a modification of the proof of Lemma 9.1.
258
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
Proof of Lemma 9.2 Consider for each n = 1,2,... the mappings fn € Hol(D, X) denned by fn(x) = x+-
f(x) -y,
i£D,
(9.23)
where t > 0 and y G V. Let Vr be the open ball centered at the origin of radius r G [0,1). For all x G dVT = {x G X : \\x\\ = r} and for all x* € J{x) we have, by (9.19),
Re(/ n (z), x') = \\xf + 1 Be(f(x), x*) - Re^, *•) > r 2 + 1 r o ( r ) _ r || y || =r(r+t-
a(r) - ||y||). (9.24)
Since w = a(l) > 0, it follows that for n big enough the equation
(9.25)
Tt
has a solution rn G [0,1). Indeed, >n(0) = £a(0) < 1 for n > t|a(0)| and c/jn(l) = 1 + £w > 1. The inequality (9.24) implies in turn that for such n and r n , and for all x with ||a;|| = rn and x* G J(x), the inequality Re(/ n (z),z*)>r n (l-|l2/ll)-
(9-26)
holds. Since fn is bounded on T>rn, it follows from [Aizenberg et al. (1996)] that the equation fn(x) = x + ^f(x)-y
=0
(9.27)
has a unique solution x = J± (y) : = ( / + £ / ) (y) G I?rn for each y € V. In other words, the resolvent mapping J± maps P into VTn. It now follows from the well-known Earle-Hamilton fixed point theorem [Earle and Hamilton (1970)] that J± has a unique fixed point r in T>. This point is also a null point of / . In addition, repeating the proof of the Earle-Hamilton theorem as presented in [Goebel and Reich (1984)], we obtain the estimate
P(JX(«), J4(»))
<J T ^ s y ^ y )
for each pair of points x and y in V.
0.28)
259
Asymptotic Behavior of Continuous Flows
Since a(r) is continuous on the interval [0,1], it follows from (9.25) that r n —» 1 and a(rn) —> w as n —» oo. Therefore, by using the exponential formula Ft(x)= lim Jl(x) n—>oo n
(9.29)
and (9.28), we get by induction the estimates (9.21) and (9.22). Lemma 9.2 is proved. Example 9.1 Let T> = A be the open unit disk in the complex plane C and let / e Hol(A, C) be defined by f(z) = a-az2 + b z \ ^ , 1 + cz where a G C, Re b > 0 and 0 < c < 1. If we take a(s) := -|a|(l - s2) + (Reb)s\^-,
(9.30)
(9.31)
then we get Ref(z)z>a{\z\)\z\
(9.32)
and a ( l ) = R e & j ^ > 0. Hence f(z) is a strongly semi-complete vector field on A. Example 9.2 In the theory of autonomous systems the following system is often considered: (xi-x2 \x2
+ xi(p(x1,x2) = 0
+ X! +x2
(9.33)
We assume that the function ip is holomorphic in the open unit ball |xi| 2 4\x212 < 1. It is clear that for any point x — {x\,x2) e B the support functional x* is denned by (y,x*) = yix1+y2x2.
(9.34)
Hence, for the mapping f(x) = (-x2 + x\(p(x), xi + x2
(9.35)
Thus we have to examine three cases: (1) There exists a point x° = (xj,i§) G B such that Re<^(a;0) < 0; if this does not hold, then either
260
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
(2)
(9.36)
\ ±2 + XX + X2(l + x\ + x\) = 0
is well defined for all t > 0 and for all initial values in B, and converges globally on B to the origin. To illustrate another application of Lemma 9.2 we consider a question regarding the solvability of an autonomous differential equation of order n. Example 9.3 Let X be the n-dimensional complex space C n = {(zi, z 2 , . . . , zn) : Zj G C} with the ^,-norm
11*11 = ( E i * * r ) , \fc=i
<9-37)
/
and let T> be the open unit ball in X. Suppose that g : T> — i » C is a holomorphic function on V which has a continuous extension to T>. Denning / = (/i, • • •, /n) : P ^ C n in (9.1) by the formulae fi(zi,z2,---,zn)
= -zt+x,
fn(zx,z2,...,zn)
= g(zx,...,zn),
l < i < n - l , (9.38)
and using the standard method of rewriting an n-th order differential equation as a first order system of n equations, we deduce that the boundary condition Re f(z)zn
> \zn\2~p Re £ zk+1 ^ fc=i
,
z G dV,
(9.39)
Zk
is fulfilled if and only if the equation x^+g(x,x',...,x(n-V)=0
(9.40)
261
Asymptotic Behavior of Continuous Flows
with the initial data i(0) = 2 ll x'{0) = z2,...,x^-1\0)
(9.41)
= zn
has a unique solution x = x(t, z\,... ,zn), defined for all t > 0 and (z\,..., zn) € T>, which satisfies the estimate,
iix(t)iip,T = 0 ^«,(K*)r+ia:'(*)r+-" + k ( n - 1 ) (t)r) i < i (9-42) for each T > 0. It is clear that if we set a(s) := inf {fcfosi,.. . , i n . O : \x\p + \Xl\" + • • • + |x n _i| p = sp},(9.43) where *,
N -c
\g(x,Xi,...,Xn-i)\xn-i\p
= Re — L
&(x,xi,x2,...,xn-i) XX\X\P
X
-
xn-l
X a |n|P
Sn-l|jn-2| P l
X\
Xn-2
^
u
)
J'
then inequality (9.19) holds. Consequently, if liminf{$(o;,a;i,...,a; n _i) : \x\p + • • • + |a; n _i| p = sp} > 0,
(9.45)
then, by Lemma 9.2, the solution to the Cauchy problem in this example converges to the point (a;o,O,... ,0) which is the unique null point of the generator / . Remark 9.5 Note that if f £ Ho\(T>,X) is known to be a semi-complete vector field on V, then condition (9.20) can be replaced by a slightly more general condition, namely, a(l) > 0 for some
I € (0,1],
(9.46)
which will still ensure assertion (i) of Lemma 9.2. This implies the following very simple and interesting sufficient condition. Recall that a bounded operator A : X H-> X is said to be strongly accretive if Re{Ax,x*)>k\\x\\2 for some k > 0 and all x € X, x* € J(x).
(9.47)
262
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
Corollary 9.2 Let f S G{V), and suppose that the linear operator A = /'(0) is strongly accretive, that is, it satisfies (9.47) for some k > 0. / / o>4||/(0)||,
(9.48)
then f is a strongly semi-complete vector field. Proof. Consider the function a{s) = - | | / ( 0 ) | | ( l - s 2 ) + ks±=j. Using (9.48) we see that a(l) = 0 and a'(l) < 0. Hence there is I G (0,1) such that a(l) > 0. Since Theorem 9.1 shows that Re(/(x),a;*)>o(||a:||)-||a;||,
(9.49)
the result follows by Remark 9.5. Note that if A = / ' ( 0 ) is strongly accretive and /(0) = 0, then condition (9.48) is fulfilled automatically. Hence the origin is an attractive fixed point of the semigroup generated by / . Actually, this fact also follows from Corollary 9.1 above and the exponential rate of convergence obtained there.
•
9.2
Asymptotic Behavior of Flows of p-Nonexpansive Mappings on the Hilbert Ball
Let B be the open unit ball of a complex Hilbert space 7i with inner product (•, •), and let p : B x B — i > R+ be the hyperbolic metric on B, i.e., p{x,y) = tanh"1
\Jl-a(x,y),
^ • ' > - ( 1 - | ^ W " ' ) - *•-»•
<9-M)
As above, we denote by Np the class of all self-mappings F : 1 H> B which are nonexpansive with respect to p (/9-nonexpansive), i.e., p(F(x),F(y))
(9.51)
This class Mp properly contains the class Hol(B) of all holomorphic selfmappings of B. Definition 9.3 A mapping / : B t-» % is said to be strongly pmonotone (/)-monotone) if for each pair i , y € l there is e = e(x,y) > 0 (e = 0) such that p(x + rf(x), y + rf(y)) > (1 + re(x, y))p{x, y)
(9.52)
263
Asymptotic Behavior of Continuous Flows
for all r > 0 such that x + rf(x) and y + rf(y) belong to 1. it was shown in [Reich and Shoikhet (1997a)] that / : B »-> H is pmonotone if and only if it satisfies the condition
[1-llzH2
l-||y||2J
L
l-(x,y)
J
(9.53)
It is also known [Reich and Shoikhet (1997a), (1998b)] that if <S C Mp is the flow generated by a generator / £ QAfp(B) and / is bounded and uniformly continuous on each /3-ball in B, then the following relations hold: for each r > 0, W = Null(/) = Fix(J r ),
(9.54)
where Jr = (I + rf)'1. In the study of the asymptotic behavior offlowsof ,0-nonexpansive (or holomorphic) self-mappings of B, the two cases W ^ 0 and W = 0 are usually considered separately. In particular, if / € GAfp(B), then one can look for conditions which would imply the strong p-monotonicity of / with s(x, y) = e = const, (see formula (9.52)). If this is the case, then the exponential formula implies that W contains a unique point r which is globally attractive with an exponential rate of convergence: p(F(i)x, r) < exp(-et)p{x,T). However, such an approach cannot work when W — 0. Therefore, our aim is to find some sufficient (and perhaps necessary) conditions for global convergence of the flow generated by / which do not depend on W being empty or not. We now recall the following definition. Definition 9.4 Let 5 = {F(t)}t>o be aflowon B which is generated by / . We will say that a point T £ B, the closure of B, is a globally attractive point for S if for each x € B the strong limit lim F(t)x = T,
t—>oo
(9.55)
uniformly on each p-ball in B. If T £ B, then r is the unique asymptotically stable stationary point of S. If T £ OB, the boundary of B, we will call it the attractive sink point of S. As we have seen in the previous section, in the case of holomorphic generators the attractivity of a stationary point can be completely described in terms of their derivatives. For holomorphic generators with no null point
264
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
the situation can be described by using the so-called angular derivative at the boundary sink point (see the next section). However, in both cases (W ^ 0 and W = 0), the characteristics of the derivatives are not relevant in general, since / ' does not exist for / £ ^ATp(B) in the complex sense if / is not holomorphic. In this section we study such a situation. For a fixed T 6 B, the closure of B, and an arbitrary x £ B, we define a non-Euclidean "distance" between a; to r by the formula
(9.56)
Mx)=llr:^f(l-a(x,T)),
llxll where a(x, T) is defined by formula (9.50). 1
We already know that geometrically, the sets E(T, S) = {X€M: dr(x) <s},
s> 0,
(9.57)
are ellipsoids. If T € B, then these sets are exactly the p-balls (9.58)
E ( T , S) = { X € E : p(x, T) < r)
centered at r £ B and of radius r = tanh" 1 ^/3+1_f||T|ia . If r € dM, the boundary of B, then these sets E(T, s) = ( i 6 l : dr(x) = I
~ ^ 2 I 1 \\x\\
| 1
< s], )
s>0,
(9.59)
are ellipsoids which are internally tangent to the unit sphere dM at r . Now for fixed T E B and x 6 8E(T,S), X ^ T, consider the non-zero vector x*
= -,
1 T ( \ 1. n a S - i
1-(T(X,T)
\ 1 - \\X\\2
7 TTV 1-(T,X)
(9"6°)
J
As in [Aharonov et al. (1999a)], it can be shown that x* is a support functional of the smooth convex set E(T, S) at x, normalized by the condition lim(a;-T,a;*) = l.
(9.61)
X—>T
Then for a mapping / : B —> H, the fiow-invariance condition Re(/(x),x*>>0
(9.62)
Asymptotic Behavior of Continuous Flows
265
is necessary for / to be the generator of a continuous flow for which the sets E(T, S) are invariant. In our situation, when / £ GAfP(R), this is exactly the case if r £ B is a null point of / , since p(F(t)x, T) = p(F(t)x, F(t)r) < p(x, r).
(9.63)
Note also that condition (9.62) can be obtained directly from the pmonotonicity of / if we substitute y = r and / ( r ) = 0 into (9.53). In fact, inequality (9.63) shows that if condition (9.62) holds for some T £ B and all x £ B, then T must be a stationary point of <S = {F(t)}t>o, hence a null point of / . If / has no null point, then there is a unique boundary point r £ dM such that (9.62) holds. This point r is the sink point for the flow generated by/In order to classify the asymptotic behavior of flows we will consider a finer condition than (9.62). More precisely, for a point r e B and / e QAfp{M) we consider the following two real nonnegative functions on (0, oo): Wb(s) :=
inf
2Re(a;),a;*)>
s > 0,
(9.64)
inf
2Re(/(x),a;*),
s > 0,
(9.65)
dT(x)<s
and J(s):=
dT(x) = s
where x* is defined by (9.60). It is clear that w*(s) >w b (s) > 0
(9.66)
and that u>\,(s) is decreasing on (0,oo). Let .M(0, oo) denote the class of all positive functions UJ on (0, oo) such that jj is Riemann integrable on each closed interval [a, b] C (0, oo) and f
ds
/ , . is divergent. Jo+ u>(s)s 6
(9.67) *• ;
Note that for each OJ € M(0, oo), the function fi defined by dT(x)
/
i\
dm
<9-68)
is a strictly decreasing positive function on (0, dT(x)] which maps this interval onto [0, oo). We denote its inverse function by V : [0, oo) H-> (0, dT(x)\.
266
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
Theorem 9.2 Let f e &A/"P(B) be continuous and let S = {F(t)}t>0 be the flow generated by f. Given a point r € B and a function u> £ M(0, oo), the following conditions are equivalent: (i) For all se (0,oo), (9.69)
J{s) > w(s)
where u}*(s) is defined by (9.65); (ii) for any differentiate function W on [0,oo) such that V(t) < W{t), V{0) = W{0) and V'{Q) = W(0), dT(F(t)x) < W(t),
i £ l , t>0,
(9.70)
where V = fi"1 and Q is defined by (9.68). In particular, dT(F(t)x) < V(t); hence r is a globally attractive point forS. Proof. Consider the function $ : R + x l n R + defined by (9.71)
V(t,x) = dT(F(t)x). By direct calculations we have
^
[ = o + = -2*(0,z)Re
(9.72)
Let us first assume that condition (ii) holds. Since $(0,:r) = dT(x) = W(0), we get by (9.72) and (ii) that 2*(0,x)Re{f(x),x*)
= - ^ | t = o + > ~Jt
[W(t)}t=Q+
(9.73) Varying x € 0E{T, s) = { i e B : dT(x) = s}, we see that this inequality immediately implies (i). Conversely, let condition (i) hold. It follows from (9.71) and the semigroup property that for all x € B and s, t > 0, *(« + t , i ) = *(s,F(t)x).
(9.74)
Hence by (9.72) and the continuity of / , ^ is differentiable at each t > 0 and we deduce from (i) and (9.72) that ^
^ at
< -tt(t,*)«»(*(*, x)) < - * ( t , x)u>(*(t,x)).
(9.75)
267
Asymptotic Behavior of Continuous Flows
Separating variables we get d r(z)
Jit,
——
/ T (F(t)i)
= n(dT(F(t)x)) > t,
(9.76)
W^WJW
which is equivalent to condition (ii). Theorem 9.2 is proved. • We will call a function w £ .M(0, oo) which satisfies condition (i), an appropriate lower bound for / e QJ\fp(M). Remark 9.6 Of course, if the function u>\, defined by (9.64) belongs to .M(0,oo), then one can use it as an appropriate lower bound. However, examples show that sometimes UJ\, may be identically zero, while u>* itself belongs to the class M(0, oo). Moreover, we will see below that for a semigroup of holomorphic mappings with a boundary sink point, LJ\, is always a constant which determines the best rate of uniform exponential convergence of the flow. To illustrate Theorem 9.2 and to motivate our next definition, we now present several one-dimensional examples [Elin et al. (2002)]. Example 9.4 Let A be the open unit desk of the complex plane C, let n be a positive integer and let / : A — i > C be defined by
(9.77)
f{z) = -(l-z?)±^. If we set T = 1 and
(9.78) then we get
Re/(2)r = T^W Re r^S = di(z)Re r^S
> °- (9-79)
Since / is holomorphic on A, this inequality implies that / generates a flow 5 = {F(t)}t>o of holomorphic self-mappings of A. In addition, it can be shown (see Theorem 9.4 below) that w(s) = wb(s) =
inf 2Ref(z)z*
= const. = - .
di(z)<s
(9.80)
n
Hence / satisfies the conditions of Theorem 9.2. In this case, _. ,
n /- rfl(x) dX
n{s)=2js
n ,
s
T = ~2lnaW)
, (9 - 81)
268
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
and
V(t) = fi"1^) = e x p j - ^ j di(z).
(9.82)
Thus we have an exponential rate of convergence of the flow <S to the boundary point r = 1:
(9.83) Note also that although / has n + 1 null points {a^ :fc= 1,2,..., n +1} on the unit circle, only a\ = 1 is an attractive point of S = {F(t)}t>o. The reason is that Re/'(ai) > 0, while Re/'(a*,) < 1, k = 2 , 3 , . . . ,n + 1 (see Theorem 9.5 below). Example 9.5 Let A be as above and let / : A —> C be defined by
/(z) =
_(1_2)2l±ig
(9.84)
with \c\ < 1. Once again, if we define z* as in Example 9.4, we have
Ref(Z)l* = i ^ j j Re i ± g > dl(z) i ^ M > 0.
(9.85)
In this case, w\, (s) = 0 for all s £ (0, oo) and we cannot use it as an appropriate lower bound. However, we can define w(s) = as, where a = j ^ | 4 , and we find 1/1 1 \ 1 rdl^ d\ Q(s) = ^ = i(i__L^). ajs A2 a\s di(z)J Thus we get, by Theorem 9.2, the following rate of convergence:
i^W<_J Example 9.6
iizif
(9.86)
(987)
1 - \F(t)z\2 ~ 1 + atdiiz) 1 - \z\2 ' v " ' Let A be as above and let z = x + iy £ A. Define / : A —>
Cby f(z) = x*+iyl.
(9.88)
Since Ref(z)z = x¥ +y¥ > 0,
(9.89)
269
Asymptotic Behavior of Continuous Flows
/ is />-monotone and the origin is the unique null point of / . Hence, if we set r = 0, then we have do(z) = Y
(9.90)
^
and J{s) = 2 inf r - ^ —Bef{z)S Mz)=s \z\2(l - |z|2) = 2
inf , **+y* a p 2 M ( l + ,)i. *2+y2=?h (z2 + y 2 )(l-:E 2 -y 2 )
(9.91)
Setting u>(s) = w"(s), we get ^ ^ = J_ / w(A)A 2 * 7 , Finally, we obtain the estimate
dA
Al(A + l)*
.
(9.92) V
7
do(F(t)z) < V(t) = — ^ s . (9.93) [^-tdo(z)l + (do(«) + l)5] a -do(z) The latter inequality is equivalent to the estimate J^J -y .
\F(t)z\ <
(9.94)
Note that one can calculate F(i) directly by solving the Cauchy problem and get \F(t)z\2 =
+
r
(Ixit + l)2
Vr . (|yft+l)'
(9.95)
Thus for x = y we obtain \F(t)z\ =
J^
r
.
(9.96)
So, the rate of nonexponential convergence we have obtained is sharp. Remark 9.7 We will see below that a similar phenomenon is impossible for holomorphic mappings, namely:
270
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
If a flow of holomorphic self-mappings converges locally uniformly to an interior stationary point, then the convergence must be of exponential type. Definition 9.5 Let 5 = {F(t)}t>o be a flow with a stationary (or sink) point T e B. We will say that the asymptotic behavior of S at T is of order not less than a > 0 if there is a function u> € A4(0, oo) such that l i m i n f / ^ H >0,
(9.97)
and MF{t)x)-(i
+ iu\u*)))adr{x)
(9-98)
for all x G B and t > 0. Definition 9.6 We will say that the asymptotic behavior of <S at r is of exponential type if there is a decreasing function w e -M(0, oo) such that dT(F(t)x) < exp(-tu){dT(x)))dT(x)
(9.99)
for all x e 1 and t > 0. In particular, if OJ can be chosen to be a positive constant a, then we will say that 5 has a global uniform rate of convergence: dT{F(t)x) < exp(-ta)dT(x).
(9.100)
The following assertion is a consequence of Theorem 9.2 [Elin et al. (2002)]. Theorem 9.3 Let S = {F(t)}t>o be a flow generated by f £ QAfp{B) with a null (or sink) point T € B. Then the asymptotic behavior of S at T is of order not less than a > 0 if and only if there exists an appropriate lower bound u> € M(0, oo) for f such that —^
is decreasing on (0,oo).
(9.101)
Proof. We first observe that condition (9.98) with some UJ e X(0,oo) satisfying (9.97) is equivalent to the same condition with a function u\ S .M(0, oo) which satisfies both (9.97) and (9.101). Indeed, for a given CJ € A4(0, oo), define a function \i: (0, oo) •-+ (0, oo) by /i(s) = inf | ^
: le (0,s]|, s > 0.
(9.102)
271
Asymptotic Behainor of Continuovs Flows
It is clear that /i(s) is decreasing. Setting now wi(s) = s« • fJ>(s), we clearly see that LO\ satisfies (9.97) and that LJI(S) < u(s). Hence
/
-rr
(9-103)
Jo+ wi(s)s is divergent and CJI G A-f(0,oo). In addition, we have the inequality
[i+^oor - [ i + ^ w ] t t
(9'104)
which proves our claim. Thus we can assume for the rest of the proof that w satisfies (9.101). It remains to show that UJ is an appropriate lower bound for / . Indeed, defining fi : (0,dT(a;)] -» [0,oo) by (9.68) and using (9.101) we have ,dT(x) U{S)-Ja
dx
^
rdT(x)
^AJA"7 3
- w(dT(x)) i s
x±dx
w(A)Ai+i
A^+1
= ^b) [ a "* ( d r < a : ) ) i - l ] -
(9'105)
Inverting this expression we get y ( t ) := n - 1 w
* (i + i ^ w ) ) ^ ( x )
:=
^(t)-
(9 - 106)
It is clear that the function W(t) satisfies all the conditions of Theorem 9.2. D This completes the proof of Theorem 9.3. Our next assertion is a direct consequence of Theorems 9.2 and 9.3. Corollary 9.3 Let S — {F(t)}t>o be the flow generated by a mapping f with a null (or sink) point r € M. Then (i) the asymptotic behavior of S at r is of exponential type if and only if inf{io l l (/):/e(0,s]} > 0 ,
s > 0;
(9.107)
(ii) the flow S has a global uniform rate of exponential convergence if and only if OJ*(S)
for some a > 0.
>a
(9.108)
272
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
Indeed, in both cases (i) and (ii), there is one function w £ .M(0, co) such that the asymptotic behavior of <S at T is of order not less than a for all positive a. In case (i), UJ can be chosen to be u(s):=M{J(l):l£(0,s}} > 0,
s > 0,
(9.109)
while in case (ii) u> can be chosen to be a constant a. Remark 9.8 However, we will see in the next section that, for holomorphic mappings, condition (9.108) holds, in fact, for some a > 0 whenever condition (9.107) holds. In other words, for holomorphic flows any convergence of exponential type implies global uniform exponential convergence. The following example shows that for a semigroup of p-nonexpansive (but not holomorphic!) mappings an asymptotic behavior of exponential type does not imply, in general, a global uniform exponential rate of convergence. Example 9.7 formula:
i > C by the following Define a continuous mapping / : A — f(x + iy) = x(l - xf + iy(l - y)2.
(9.110)
Since Ref(z)z > 0 for all z = x + iy e A, it follows that / is the generator of a semigroup S = {F(t)}t>o of p-nonexpansive mappings such that each disk A r = {z G C : \z\ < r < l } is F(t)-invariant. Setting r = 0 and z* = i^p^l^p), we have
w»(s)= inf Re/(z)r= «*,(*)=«
inf
f{} ~ f^+ ^
" ^ {9.111)
^+?/2=rfT (z 2 + y 2 ) ( i - z 2 - z / 2 )
It is easy to see that lim w"(s) = 1 while w'(s) —» 0 as s —> oo (take, s->0+
for example, y = 0 and x = */j^i —» !)• 9.3
Flows of Holomorphic Mappings on the Hilbert Ball
In this section we will study in more detail flows S = {F(t)}t>o of selfmappings generated by holomorphic mappings / £ C/Hol(B) with stationary (or sink) points r € B. We already know that the asymptotic behavior of a flow S at T is of exponential type (Definition 9.6) if and only if the function J(s)=
inf 2Re(/(z),z*),
dT(x)=s
s > 0,
(9.112)
Asymptotic Behavior of Continuous Flows
273
satisfies (9.107). It turns out (see Theorem 9.4) that in this case this function and even the function w b (s)=
inf
dT(x)<s
2Re(/(a;),a;*)
(9.113)
are bounded from below by a positive number. Moreover, for a boundary sink point the function u)\, is just a constant. In both cases (interior stationary point or boundary sink point) the asymptotic behavior of a flow is completely determined by the value o/"(0) := liminf u>^(s) which is related to the value of the derivative of / at its null point (for the interior case) or the so-called angular derivative (for the boundary case). We begin with the following general assertion.
Theorem 9.4 (Theorem on universal rates of convergence) Let f G (?Hol(B) and let {F(t)}t>o be the flow generated by f. If for some point T £ B there is a decreasing function LJ : (0, oo) t-> (0, oo) such that i 6 l , t > 0,
dT{F{t)x) < e-tul{dAx))dT{x),
(9.114)
then there exists a number fj, > 0 such that dT{F(t)x)
< e-^drix),
x G B, t > 0.
(9.115)
Moreover, (i) if T G B, then /u can be chosen as \i = " b j , but pi cannot be larger than wJO) (= lim uds)); s->0+
(ii) ifr G 9B, then the maximal \x, for which (9.115) holds, is exactly W|,(0), that is, 0 < / i < W[,(0). We will prove and discuss this theorem separately for the case where T G B is a null point of / and for the case where r G dB, that is, when / is null point free.
9.3.1
Interior stationary
point
Lemma 9.3 Let f G £Hol(B) with /(0) = 0 and let uA and UJ\, be defined by (9.112) and (9.113). Then (i) w«(0) = wb(0) = 2 inf
||x|| = l
(ii) ^
Re(f'(0)x,x);
274
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
Proof.
First we show that J(0) < 2v,
(9.116)
inf Re(/'(0)z,a;).
(9.117)
where v=
Since in our case r = 0, we have Re(/(x), x*> =
(9.118)
Re{f(x),x).
l
II-HI K1
\\x\\ )
Now fixing u G dE, we set x = ru, where r € (0,1). Then we get
Re(/(z),z*) = R e — ! ^ / - / ( ™ ) , A 1 — rz \ r
(9.119) /
Therefore, J(s) <2Re-J-5- (-f(ru),u\, 1 — rz \r
I
where r2 =
||:E||2
= —?—. (9.120) s +1
Letting s (hence, r) tend to zero we obtain w"(0)<2Re(/ / (0)«,u).
(9.121)
Since u is arbitrary, (9.116) follows. On the other hand, it follows from the generalized Harnack inequality (see, for example, [Aharonov et al. (1999b)]) that for all i g l ,
Re(z), x) > Re(/'(0)z, x) ^ M > Hl^ll2 [ ^ J •
(9-122)
This implies that 2Re{f(x),x') =
| | x | | 2 ( 1 2 _ | | x | | 2 ) Re(/(x),x)
2HNI2 - ||*P(l-||sP)
i - N i _ 2, l + ||x|| (1 + NI) 2 "
^ ^
Hence
=
,i.«^aTHP = (^^) 5 (^)-
(9'124)
275
Asymptotic Behavior of Continuous Flows
Letting s tend to 0 + in (9.124), we see that UJ\,(0) > 2v. Since obviously u>t(O) > u>b(0), comparing the latter inequality with (9.116) we obtain (i). On the other hand, substituting now in (9.124) v — ^p-, we get asser• tion (ii). This completes the proof. To proceed, we denote by MT the Mobius transformation of B defined by
"'M-l4^('-W-' / r r F F (*-^)> ( M J 5 ) Note that MT is an automorphism of B (see, for example, [Rudin (1980)] and [Goebel and Reich (1984)]) which has the following properties: (a) M~l — MT (involution property) with MT(0) = T and MT(T) = 0; (b) l-\\MT(x)\\2 = a(x,y);
(c) l-(M r (x),r) = if{gg.
These properties imply the equality dT(MT(x)) = (1 - ||r|| 2 ) do(x).
(9.126)
Now let us consider the flow {G(£)}t>o C Hol(B) denned by G(t) = MTo F(t) o MT,
(9.127)
and let g G
= [(MO'Wr1/^*)).
(9.128)
Then Gt(0) = 0 for all t > 0 and g(0) = 0. Lemma 9.4
The following equality holds:
(9.129) where y — MT(x). Thus the functions w"(s) andw^(s) are invariant under the transformations (9.127) and (9.128).
276
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
Proof.
We have already seen in (9.71) and (9.72) that
= - 2 T^7) **('«• T^-T^fe))-
<9130)
On the other hand, by (9.126) and (9.127) and property (a) of MT we have
| [dT(F(t)x)]t=0+
= | K(M r G(%)] t = = Q + = ^[(l-||T|| 2 )do(G(i) 2 /)] t = o + .
(9.131)
Since
'-(i-H*')i»ipiBt(aM'iif)' we obtain (9.129) from (9.130) and (9.131).
(9132)
•
Now we are able to complete the proof of Theorem 9.4 for the case of an interior stationary point r £ B of 5. To this end, let us assume that condition (9.114) holds. Then it follows from Corollary 9.3 (i) and Lemma 9.3 (i) that W|,(0) = wB(0) > 0. Let the flow {G(t)}t>0 C Hol(B) and its generator g G £Hol(B) be defined by (9.127) and (9.128). By Lemmata 9.4 and 9.3 we have inf
Re(g(y),y*) ^
.
(9.133)
for all y 6 B.
(9.134)
Then, by Corollary 9.3 (ii), we have do(G(t)y) < do{y)e-t
*
Finally, setting y = MTx and using (9.126) we conclude that w.
(0)
dT(F(t)x) < dT(a;)e-t ~T~ .
(9.135)
Thus, the proof of Theorem 9.4 for the case where r e B is an interior stationary point is complete.
277
Asymptotic Behavior of Continuous Flows
Corollary 9.4 Let f 6 £Hol(l) with /(r) = 0, t € 1, and let {F(t)}t>0 be the flow generated by f. Then {F(t)}t>o has a global uniform rate of exponential convergence if and only ifcj^(O) > 0. Corollary 9.5 Let {F(t)}t>0 be the flow generated by f £ £Hol(B) and let T £ B. Then the following estimates are equivalent: (i) dT(F{t)x)
< e-^dT(x),
xeM, t>0;
< \\MT(x)\\ • e - ^ " l | M 2 W I | 2 t ,
(ii) \\MT(F(t)x)\\
l-||MT(x)||
x€B,t>0;
t
(Hi) | | M r ( F ( t ) a ; ) | | < ||M T (a;)|| • e~"' *+ll*r^ WH *,
x e I , t > 0 ,
where the numbers fi in (i) and (ii) can be chosen to be one and the same such that 0 < ^^- < (j, < LJ^(0) and v in (Hi) is defined by v = iw b (0) = ^w»(0) = inf Ke{Bf'{T)B-lx,x). 2
||z||=l
I
(9.136)
Here B is the linear operator defined by B = PT + y/l — ||r|| 2 (J — PT) if T ^ 0 andB = IifT = 0. Proof. First we note that inequalities (ii) and (iii) are equivalent to the following ones: (ii*) \\G(t)y\\<\\y\\-e-'ll=^lt,
t > 0;
(iii*) \\G(t)y\\<\\y\\.e-'^Mt,
t>0,
where y = MTx e B and the flow {G(t)}t>o is denned by (9.127). First, let us suppose that estimate (i) holds. By using (9.126) for the flow G we have (9.137)
do(G(t)y) < e~^do(y). Rewriting the latter inequality in the form
i - \\G(tM2
i - bll2
'
(9.137)
we get by direct calculations, I|G(%"2
-
IM|2Ni2 +
which coincides with (ii*).
(i-IMl^
£
I W I 2 - e ^ ( M l ! / " 2 ) ' (9-139)
278
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
Now we will assume that inequality (ii) (and hence (ii*)) holds. Differentiating both sides of this inequality with respect to t at t = 0 + , we obtain
-±Re(g(y),y) < _|| y || M i^M .
(9.140)
This implies that w'(s) > /x. Thus the decreasing function w(s) = a is an appropriate lower bound and the implication (ii) =>• (i) follows from Theorem 9.2. The claimed estimate for the number /x is contained in Theorem 9.4 (i). Now let us suppose again that inequality (ii) (hence, (ii*) and (9.140)) holds with some number /j, > 0. Setting in (9.140) y = ru, u € dB, r € (0,1) and letting r tend to zero (cf. the proof of Lemma 9.3), we get Re(g'(0)u, u) > f > 0. A direct calculation shows that g'(0) = [(MTy(0)}-lf'(T)(MT)'(0) = Bf'{r)B-\
(9.141)
and so v > 0. Therefore, again by the Harnack inequality, we have
Re (g(y), y) > Re (g'(0)y, y) ^ M > v\\yf i ^ M .
(9.142)
On the other hand, dln\\G(t)y\\
at
1 dln\\G(t)y\\2
=2
at
=
1
Jew
vJdG{t)y
\
Re\~dT'G{t)y)
=-wkp**{g{G{t)y)'G{t)y)-
(9.143)
Also, it follows from the Schwarz Lemma that ||G(£)y|| < \\y\\. Thus we have
M , . ^ .
(,114)
Integrating this inequality, we obtain the following estimate:
ln\\G(t)y\\ - ln\\y\\ < - " [ ^ | j | [ *, which coincides with (iii*).
(9-145)
279
Asymptotic Behavior of Continuous Flows
Finally, if condition (iii*) holds, then differentiating it with respect to t at t = 0 + , we get
^^'^-(TTW-^ 0 -
(9J46)
Thus (^(O) > 0 and, in view of Theorem 9.4, the result follows.
•
Remark 9.9 The above Corollary asserts that an exponential rate of convergence in the sense of the "distance" dT(-) is equivalent to the same rate of convergence in the norm ofH. We remark in passing that when r = 0, estimate (iii) can be extended to an arbitrary Banach space (see [Poreda (1987)] and [Suffridge (1977)]). Recall also that the original topology and the topology generated by the hyperbolic metric on B are locally equivalent. Although global equivalence does not hold, we will prove that an exponential rate of convergence in the norm coincides with a global rate of exponential convergence with respect to the hyperbolic metric. Corollary 9.6 Let {F(t)}t>o be the flow generated by f e £Hol(!) with / ( r ) = 0 , r £ B. Then the flow {F(t)}t>o has an exponential rate of convergence if and only if the following estimate of convergence in the hyperbolic metric holds: There exists a number r] > 0 such that p(F(t)x,r)
< P(X,T)
• e-"A^,
where A(x) = e-M*,T)^
x
<E B and t > 0.
(9.147)
Moreover, the maximal value of number n for which this inequality holds is v = i n f M = 1 Re {Bf'{r)B~^x, x), where B = PT + ^1 - \\T\\2 (I - PT) if T ^ 0 and B = I if r = 0. Proof. Let us suppose that {F(t)}t>o has an exponential rate of convergence. By Theorem 9.4, Corollary 9.5, Lemma 9.3 and formula (9.135) we have
||MT(F(t)a:)|| < ||AfT(i)|| • e-'&ffitfffl*,
x 6 B,
(9.148)
where the number u is defined in (9.136). Therefore one can write the estimate
p(F(t)x,T) = p(0, \\MT(F(t)x)\\) < p(0, ]\y\\)e-
\ (9.149)
280
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
which coincides with (9.147) for 77 = v. Suppose now that inequality (9.147) holds with some number rj > 0. Again, by using the flow {G(*)}t>o defined by (9.127) we can rewrite (9.147) as p(G(t)y,0)
where A(y) = e-2p{-vfi).
(9.150)
Differentiating this inequality with respect to t at t = 0 + , we obtain
Ito(9(y),y->>i?^gj^.
(9.151)
Setting now y = ru, u £ dM, r € (9,1), and letting r tend to zero (cf. the proofs of Lemma 9.3 and Corollary 9.5), we get v > 77, as claimed. • 9.3.2
Boundary sink point. Continuous version of the Julia-Wolff-Caratheodory theorem
As a matter of fact, if S = {F(t)}t>o converges to a boundary sink point T G dM with a rate of convergence of exponential type, dT(F(t)x) < exp(-tuj(dT(x)))
• dT(x),
(9.152)
where us € A^(0, 00) is a decreasing function, then this estimate can be improved as follows: dT(F(t)x) <exp(-tcj(0))dT(x),
(9.153)
where w(0) := lim w(s). In other words, we claim that if the inequality s-»0+
w'(s) > w(«)
(9.154)
holds for a decreasing w, then the stronger inequality J(s)
> CJ(O)
(9.155)
also holds. In particular, this fact holds for the function u = w\,. This implies, in turn, that w\, is actually constant: W|,(s) = wi,(0) = /3 for all s € (0,00) and is equal to the so-called angular derivative of / (if it exists) at the point r S dM. Moreover, this number /? gives the best rate of exponential convergence of S — {G(t)}t>o. These facts can be proved by using a continuous analog of the classical Julia-Wolff-Caratheodory Theorem. We will need some additional notions, well-known in the finite-dimensional case (see, for example, [Cowen and MacCluer (1995)] and [Rudin (1980)]).
281
Asymptotic Behavior of Continuous Flows
Definition 9.7 A curve A : [0,1) i-> 1 is said to be asymptotically normal at a point T £ dE if
(ii)
fM<M<0O,
<><.
where X(s) is the orthogonal projection of A(s) onto the complex line through 0 and T: (9.156)
A(S) = (A(S),T)T.
Definition 9.8 Let h be a holomorphic function on B with values in the complex plane C. We say that h has a restricted limit L at T £ dM if h has the limit L along every curve which is asymptotically normal at r. i > Ti be a holomorphic mapping on B and let Definition 9.9 Let / : B — T £ <9B. We say that / has a finite angular derivative at r if for some element i / e H , the function h : B H-> W, denned by
(9.157) has a finite restricted limit at r. We denote this limit Zf'(T). Theorem 9.5 Let f £ Hol(B, H) be the generator of a semigroup S = {F(t)}t>o of holomorphic self-mappings o/B. Suppose that f has no nullpoint in B and that r £ dE is the boundary sink point for S. Then the following are equivalent: (1) the asymptotic behavior of S at T is of exponential type; (2) there is a positive number 7 such that dT(F(t)x) < d'^drix),
x e B and t > 0.
Moreover, if the angular derivative /3 = £f'(j)
(9.158)
of f at r exists, then
(a) (3 is positive real number with /3 = 2inf{Re(/(x),x*), x £ B}; (b) the maximal 7 which satisfies condition (ii) is exactly ft. To prove Theorem 9.5 we will need the following assertion [Elin et al. (2002)]. Lemma 9.5 Let F £ Hol(B) be a holomorphic self-mapping of B with no fixed point in B, and let r be its boundary sink point. Then the curve
282
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
A : [0,1) H-> B defined by A(s) = F(ST)
(9.159)
is asymptotically normal at T. Proof. Since r is a sink point of F, it follows by Julia's lemma that there is a number 0 < S(F) < 1 such that dT(F(x)) < S(F) • dT(x),
x € B.
(9.160)
For 0 < s < 1 we have, by (9.156), 1-||A(*)|| 1+ 5 l-||A(a)|| 1+5 1 + ||A(5)||1-5 1 + ||A(S)|| 1-5 ||r-Al| 2 1 - s 2 ||r-Aj|2 1 - S2 -1-||A(S)||2(1-S)2-1-||A(5)||2(1-5)2
(9.161) Hence, lim sup
i-HAMH < lim
sup
i - IIAWII < 6{F)
(9J62)
On the other hand, the Julia-Wolff-Caratheodory theorem asserts that 6{F) = lim inf
X
~ ^ ^ ^ = Z f (T).
(9.163)
Thus, by (9.161)-(9.163), we get
Um 2-JAWII = s-tl-
1— S
= Um s->l-
lim
1-llAMH
s-»l-
1 — S
ilizAWji
= J(
164)
1 — S
The last equality implies that
(9.165) which proves condition (ii) of Definition 9.7.
283
Asymptotic Behairior of Continuous Flows
To prove condition (i), we calculate as follows: ||A(s)-A(s)|| 2 _
.IT=
||A(s)||2 + ||A(s)|| 2 -2Re(A( S ),A(s))
l-||A(s)||2 ~ . i T l-||A(s)|| 2 llA(s)||2+l|A(5)H2-2Re{A(5),r)(A(5),r)
*
THIAW
_ ,. HA(s)||2-||A(s)U2 _ _ 1-|1A(S)H2 1-5 -.iT1-||A(S)||2 - 1 , T - l - | | A ( a ) P ' l - « = l - g | = 0, by(9.164). This proves condition (i), and we are done.
(9.166) D
Proof of Theorem 9.5 Let condition (i) of the theorem hold, i.e., for some decreasing function w € ,M(0, oo), dT(F(t)x) < e- taJ ( d ^ x »d T (x)
(9.167)
or, explicitly, 2f^^il-(^T)l2
\l-(F(t)xtT)\a 1 - ||F(t)a;||2 -
1-NI2'
(9.167)
This is equivalent to the inequality \l-(F(t)x,r)\2 l-(x,r)|2
2tu(dr(x))l-\\F(t)x\\>
-
C
1-|H|2
•
(9.167)
Once again it follows from the Julia-Wolff-Caratheodory theorem that for a fixed t > 0,
-ZIFWrM^lim'^fgy,
(9.170)
where that last limit is taken along an asymptotically normal curve at r. Let us denote w(0)= lim_ u(dr(sT)).
(9.171)
Thus, setting x = ST in (9.169) and letting s tend to 1~, we get 62{F(t)) < e-2t"(0)6(F(t)),
(9.172)
284
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
or (9.173)
5(F(t)) < e-*\
where we set 7 = 2u;(0). Now by using Julia's lemma, we obtain the implication (i)=»(ii). The converse implication can be established by differentiating the inequality in (ii) at t = 0 + . Namely, we get (9.174)
Re(f(x),x*)>l>0.
So, one can set u(s) = ^, and the asymptotic behavior of <S at r is seen to be of exponential type. To prove the second part of the theorem, we first observe that / is a generator if and only if the equation x + tf(x) = y
(9.175)
is solvable for all* > 0 and y G B [Reich and Shoikhet (1996)]. The solution x = Jt(y) = (I + tf)~1 is the (nonlinear) resolvent of / . It has the following properties: (1) for each t > 0, Jt : B i-> B is a holomorphic self-mapping of B; (2) for each x G B, the function J7i : M+ H B is continuous and the semigroup 5 = {F(t)}t>o generated by / can be represented by the exponential formula lim f.7iln(a;) = F(t)x,
x € B,
(9.176)
where the limit is taken with respect to the locally uniform topology of B; (3) if / has no null point in B and r G dE is the sink point of 5, then for each t > 0, r is also the sink point of J7i, and moreover, the following approximations hold: lim Jt(x) = T ,
t—>00
xeM,
(9.177)
and lim f(Jt(x)) =0,
t—»oo
xeM.
(9.178)
Now let us suppose that /? = Z / ' ( T ) exists (finitely). Then it follows by properties (1) and (3) above and Lemma 9.5 that for each t > 0, the curve
Asymptotic Behavior of Continuous Flows
285
At(s) := Jt(sr) : [0,1) — i » B is an asymptotically normal curve at T £ dE. In addition, by equation (9.175) we have the identity At(s) + t / ( A t ( s ) ) = s r
(9.179)
for all s € [0,1) and t > 0. Denote the angular derivative Z.[Jt]'{T) of Jt : B t-> B at the point T by ct. Again, by the Julia-Wolff-Caratheodory theorem, Ct =
1-(MST,T)
lim
s->l"
1— S
(9.180)
By (9.179) and (9.180) we get lim /(A t (s)) = 0 and s—>1~
P =
(/(At(a)),r) 1 ((gr - At(s))) (At(s),r) - 1 *"?- t " (At(s),r) - 1 i im I -*+(A«(*)),r) 1 /(A f (.)),r)-1 l-3 \ , " i - t " l-(Af(s),r) s-t?-t\l-(At(s),T)+l-(At(s),T)J S "T-
= i(-l
+
l).
(9.181)
Thus we obtain that /3 is a non-negative real number and C*
= TTW
(9-182)
Hence, dr(Mx))
(9-183)
by Julia's lemma. Applying now the exponential formula (see property (2) above) we get dT(F(t)x) < e-t0dT{x).
(9.184)
To conclude the proof of Theorem 9.5, it remains to be shown that if condition (ii), or, equivalently, inequality (9.174) holds for some 7 > 0, then Indeed, setting x = ST in (9.174) we get
Re (f(sr), (ST)*) = Re (f(sr), ^ ^ " ^ }
. 5 ^ . ^ ^ .
(,185)
286
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
Letting now s tend to 1~, we get
(9.186) i.e., /? > 7. This completes the proof of Theorem 9.5. • For the one-dimensional case, that is, when B = A, the open unit disk in the complex plane C, the situation can be described as follows. Theorem 9.6 ([Elin and Shoikhet (2001)]) A mapping f e gHol(A) has no null point in A if and only if for some r £ dA the angular derivative ^f'{r) = 0
(9.187)
exists (finitely) with Re/3 > 0. Moreover, ifS— {F(t)}t>Q is the flow generated by f, then \F(t)z - T\2
\Z - T\2
1 . \F{t)z\2 * exP (-«Ra« 1YZ^ .
(9-188)
i.e., the point r is unique and the (globally) attractive sink point ofS. This assertion is an infinitesimal version of the Julia-WolffCaratheodory theorem. We remark in passing that more information on the asymptotic behavior of holomorphic and p-nonexpansive mappings and semigroups in the Hilbert ball can be found, for instance, in [Reich (1985); (1991); (1992)], [Reich and Shafrir (1987); (1990)] and [Reich and Shoikhet (1997b)]. 9.4
Admissible Lower and Upper Bounds and Rates of Convergence
In this section we intend to examine the influence of certain estimates involving the generator on the asymptotic behavior of the semigroup it generates. As in the previous section, we will compare the solutions of the Cauchy problem (9.1) with the solutions of certain one-dimensional Cauchy problems, namely,
[ft + *<0=°
(9,89)
[/3(0,s) = s € [ 0 , l ) . We begin with the following auxiliary assertions [Elin et al. (2004)].
287
Asymptotic Behavior of Continuous Flows
Lemma 9.6 Let w be a continuous positive function on [0,1). Then for all s € [0,1), the solution (3{t,s) of the Cauchy problem (9.189) is defined for all t > 0 and converges to 0 as t —> +oo. In addition, if m(s) and M(s) are the minimum and maximum, respectively, of the function w on the interval [0,s], then (3(t,s) satisfies the following estimate: se-M(s)t
< ^
s)
< se-m(s)t
(g
190)
Proof. First, rewriting the differential equation of (9.189) in the form | j | = —/3o;(/3), we note that the solution (3{t,s) of (9.189) is decreasing with respect to t (when it is denned) and positive. Second, since
\
J^
=
_
(9 191)
the convergence of the solution (3{t, s) to a certain limit so > 0 as t —> oo is equivalent to the divergence of the integral (9.192)
-^-r.
xu(x) But this holds if and only if so = 0. Finally, since ln/3(t,s) - Ins = - / w(/3(r,s))dr, Jo the monotonicity of /3(-, 5) implies the last statement of the lemma.
(9.193) •
Remark 9.10 For a given 0 < t < 00, consider the two monotone sequences defined as follows: Mi := M(s) = max{a;(x) : x G [0, s]}, mi := m(s) = min{w(x) : a; € [0,s]},
(9.194)
and Mn+i := max {w{x), x € [se~Mnt, s]} , m n + i := min {u(x),
x e [se" M n t , s]} .
(9.195)
Note that the sequence {Af n }^ =1 is decreasing while the sequence {^n}^=i is increasing. Hence the limits A — lim Mn and B = lim m n exist. Iterating
288
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
the proof of the last statement of the preceding lemma, we obtain the validity of the estimates (9.196)
se~tA < P(t,s) < se~tB.
Furthermore, if we have more information about u>, then the latter inequalities can be rewritten in a more precise form, namely, ifui is increasing, then A = UJ(S) and so /3(t,s) > sexp(—tu>(s)); if LU is decreasing, then B = u>(s) and 0(t, s) < sexp(—tw(s)). Lemma 9.7 Let w\, andoj2 be two continuous positive functions on [0,1) such that u>\ < u2- Let f3\ and (32 be the solutions of the following Cauchy problems:
f^+/WA) = o
If +/WA) = o
md
l/9i(0,a) = a
l/%(0,a) = s, (9.197)
where s e [0,1). Let m : [0, oo) i-y [0,1) be a differentiable nonnegative function such that m(0) = s and -mw 2 (m) < -i- < -muAm). at Then 02(t,s) < m(t) < Pi(t,s).
(9.198)
Proof. The assertion is evidently true when s = 0. Let to > 0 be small enough (so that m{t) > 0 for all 0 < t < to)- We have
r(''^<_t
Js y"i(y) Js xu2{x) On the other hand, the definitions of (3\ and fa imply that
_ ,
=
/ * < " > * / * < " > *
Js
ywi(y)
Js
(9.199)
(,200)
XCJ2(X)
Prom this we deduce that / r r < 0 and / rr < 0 Jpi(t,s) yui(y) Jm(t) xu2(x) for t small enough. Since the integrands are positive, we conclude that Now we will show that this inequality holds for all t > 0. Assume that there is t > 0 such that m(t) < 02(t, s). Let to be the infimum of all such
289
Asymptotic Behavior of Continuous Flows
t's. It is clear that /?2(*o,s) = m(to) = «o- Repeating our arguments, we get fait + to,s) < m(t + to) for t small enough. But this contradicts the choice of to. In a similar way one can also show that (3\{t, s) > m(t)forall t>0. O By NT we denote the class of all semi-complete vector-fields on B vanishing at the point T eB, i.e., (9.201)
JVT = { / 6 S H O 1 ( B ) : / ( T ) = 0 } .
Lemma 9.8 and only if
A mapping f 6 Hol(B, H) belongs to NT for some
R
JlM^>ReMML±, 1 - \\x\\2
Proof. i.e.,
X£K
1 - {X, T)
TGBI/
(9.202)
First we note that / e £Hol(B) if and only if it is p-monotone,
R e [|M4 + mj|i> R e . Ll-||a;||2
l-||j/||2J
(9.203)
l~(x,y)
If now / ( r ) = 0, T G B, then setting y = r in (9.203) we get inequality (9.202). Conversely, let (9.202) hold for some r e B. Denote
(9.204) Then (9.202) can be rewritten as Re(f(x),x+) >0,
xeB.
(9.205)
Now it can be shown that for each k > 1 — ||r|| 2 , and for each pair of points x e dET(k) and y € ET(k), the following inequality holds: Re(a;,a;+) > Re(y,x+),
(9.206)
where gr(fc) = | 3 : g B : | 1 i ~ _ ( ^ ) 2 | 2 < f c | ,
fc>l-||r||2,
(9.207)
are p-balls in B. Thus x+ is a support functional of the convex set ET(k) at the point x G dET(k). Now condition (9.205) implies that / G gUo\(ET(k)). Letting k tend to infinity, we also see that / G £Hol(B), because (J ET{k) = M. This concludes the proof. fc>l-||rP U
290
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
Proposition 9.1 Let / G Hol(B,if) and let u (= u(t,x)) be a local solution of the Cauchy problem (9.1). Then for any two given continuous functions u>e and u>u on [0, q) such that u>t is decreasing and nonnegative, and u)u is increasing, the following assertions are equivalent: (a) for some r G B and for each x G B, there is a number T = T(x) > 0 such that ||M_ T (z)||exp(-t Wu (||M_ T (z)||)) < ||M_T(u(t,a:))|| < ||M_ T (x)||exp(-^(||M_ T (x)||))
(9.208)
whenever t G [0,T); (b) for some T G B and for each x G B\{r},
M\\M-T(x)\\) < p£^p R e (*)'*') <M\\M.T(x)\\)t (9.209)
wherex*= rwx~
T=fcx)T-
When these assertions hold, the point r in (a) and (b) is one and the same, and moreover, f is a strongly semi-complete vector field with f(r) = 0. In addition, u (= u(t, x)) is globally defined on R+ x B and condition (a) holds for all t € R + . Proof.
Denote je(t, s) = se- ta "W
ju(t, s) = se~tuM,
(9.210)
and (9.211)
mT(t,x) = \\M-T(u(t,x))\\. Then (a) can be written as 7 u (t,
\\M_r(u(t,x))\\) < mT(t,x) < 7i(t, \\M-T(u(t,x))\\),
t G [0,T). (9.212)
Since 7u(0, ||M_T(u(t,x))||) = mT(0,x) = it(t, \\M-T(u(t,x))\\), these inequalities imply that dlu(t,\\M.T(x)\\)
di
dmT(t,x)
-
m
dye{t,\\M-r{x)\\)
-
at
, {9-2U)
291
Asymptotic Behavior of Continuous Flows
at the point t = 0, or equivalently, - W u (||M_ T (x)||).||M_ T (x)|| <
~^{T^Re(f(x),x*)
< -^(||Af_ T (a;)||) • ||Af_T(a:)||.(9.214) This yields the implication (a)=»(b). In the other direction, let (b) hold. Then it follows by Lemma 9.8 that / is semi-complete with / ( r ) = 0. Hence the solution of the Cauchy problem < dt ( «(0) = x As above, denote
' i s globally denned.
m(t) = r>v (t, x) = ||M_T(u(t, i))||
(9.215)
and s = ||M_r(a;)||. Then (b) can be written in the form -mu)u{m) < 4 ^ < -mwe(m),
(9.216)
0X
which coincides with (9.198). As a consequence of Lemma 9.7, we have WM)<m(t)
(9-217)
where /3U and fa are the solutions of the Cauchy problems
f^+/WA.) = 0 (Pu(0, s) = s
and
(f+PeMPe) = 0 lft(0,a) = a. (9.218)
This implies the required inequality (a) for all t > 0 (see Remark 9.10). D Definition 9.10 Those functions we and CJU which satisfy the assumptions of Proposition 9.1 and condition (b) of this proposition will be called admissible lower and upper bounds for / e Hol(B, H) (with respect to the point T e B). Thus the existence of an admissible lower bound for / implies that / is a strongly semi-complete vector field and the flow generated by / is exponentially norm convergent to its stationary point T, uniformly on each subset strictly inside B. Example 9.8 The mapping f(z) = (z1( z2) is obviously a strongly semicomplete vector field on the unit ball in C 2 equipped with any norm.
292
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
Clearly, for small a £ C the mapping (9.219)
fa(z) = (zl-azlz2) is still a generator. For which a € C is the mapping fa a generator in the domain D p = {zeC2:\Zl\p
+ \z2\p < 1 } ,
p>l?
(9.220)
A related calculation is given in [Suffridge (1977)]. The answer is the following: (9.221)
\a\< —{p + lY+iip-l^-1.
In the Hilbert ball B = D2 we observe that fa satisfies the estimate ||*|| 2 (1 - A||z||) < Re (fa(z),
z) < \\z\\\l
+ X\\z\\),
(9.222)
2 where A = \a\ —-= G [0,1]. Those inequalities coincide exactly with con3v3 dition (b) of Proposition 9.1 for the admissible lower and upper bounds denned by we(s) = 1 - As
and
wu(s) = 1 + As.
(9.223)
Hence by this proposition we obtain the following estimate for the semigroup {u(t, •)} generated by fa: ||2||e-t(i+A||z||) < || u ( t i Z )|| < ||2||e-*(i->ll*H).
(9.224)
Remark 9.11 If p is the Poincare metric on B, then using the known (see [Goebel and Reich (1984)]) properties of p, (9.225)
p{x,y)=p{0,M-y(x)) and p(0, kx) < kp(0, x),
(9.226)
0
we see that condition (a) in Proposition 9.1 implies the inequality p(u(t, x), T) < e-^W-rWpfa
T).
(9.227)
In other words, the exponential norm convergence implies the same rate of convergence in the hyperbolic metric.
293
Asymptotic Behavior of Continuous Flows
Remark 9.12 Generally speaking, the converse is not clear, because we know only local estimates: For each x € B there are a neighborhood U of x and numbers M = M(x, U) > 1 and m = m(x, U) < 1 such that (9.228)
m\\x-y\\
for all y € U (see [Goebel and Reich (1984)])- Therefore the following question arises: Are there an admissible lower bound u>e and an admissible upper bound uu for f £ NT such that condition (b) of Proposition 9.1 is equivalent to the same or to a similar estimate of convergence in the metric p? We answer this question by establishing the following assertion. Proposition 9.2 Let f £ Hol(B, H) and let u = u(t,x) be the local solution of the Cauchy problem (9.1). Let u>u and u>e be continuous positive functions defined on [0,1) such that
{
uu{s)
is increasing ut{s)a
. ,
.
(9-229)
— ^p — is decreasing. Then the following(1 assertions are equivalent: — sz) arctanns (a) LJU is an admissible upper bound and uie is an admissible lower bound (with respect to some point r € B) for f. (b) For some r £ B and for each x £ B, there is T = T{x) such that
pfozjexp (-iu,u(||M_r(a0||) J ^ T ^ L ) ^ Pfo «(*,*))
< P( ^)exp (-M||Af_T(aO||) J ^ t l ' l ) ) '
(9-230)
whenever t € [0,T). The point T in (a) and (b) is one and the same. Moreover, for each x e B, the solution u(-,x) is well defined globally and (b) holds for allt>0. Proof. It is clear that condition (a) coincides with condition (b) of Proposition 9.1. As above, consider the function m(t) = ||M_ T (u(t,x))||.
(9.231)
If (a) holds, then it implies that -mwu(m) < —- < ~mwe(m).
(9.232)
294
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
By Lemma 9.7, we obtain /?„(*) < m(t) <
fa{t),
(9.233)
where /?„ and fa are the solutions of the following Cauchy problems:
and { A,(0) = s = ||M_T(z)||
{ &(0) = a = ||M_ T (i)||. (9.234)
This implies that
In /aretanhflA
V arctanhs 7
=
_ I* {3u(r)uu({3u(r))dr Jo (1 -/J2(r))arctanhA.(r)
(9.235) {
'
and the same equality holds for fa and ue. Now by using (9.233) and (9.227), we get condition (b). The implication (b)=S-(a) follows from the comparison of the derivatives in inequality (b) with respect to t at the point D t = 0+. Now we will show that there are "universal" admissible lower and upper bounds u>A and w A . Proposition 9.3 For f £ Hol(B, H) and any point r € B, the following assertions are equivalent: (i) f has an admissible lower bound wi and an admissible upper bound u)u with respect to the point r S l . (ii) f G NT and there are numbers a > 0 and b> a such that
||M_r(x)||exp ( - t t l ± M = | I ) < ||M_r(u((,x))|| < | | M _ , W l l e X p ( - t o ^ M ) , (9.236) where u(t,x) is the solution of the Cauchy problem (9.1); (Hi) f G NT and there are numbers a > 0 and b> a such that
*•x)exp ("T6 n - i S ' , . ) ) - "(T>""'x) (9.237)
Asymptotic Behavior of Continuous Flows
295
The numbers a and b in (ii) and (Hi) can be chosen to be the same. In particular, if (i) holds, then one can put a = uv(O) and b = wu(0). Proof.
We intend to show that (i) is equivalent to
(i1) The functions UJA(S)
=a— 1+s
and
wA(a) = b i ^ 1— s
(9.238)
(with some numbers a and b) are admissible lower and upper bounds. In fact, all we need to prove is that (i)=>(i'), since the reverse implication is evident. Set v(t,z) = M.T(u(t,MT(x))) tanSnp{x) = - dv%+t'x)• If / € Hol(B,JJ) has admissible lower and upper bounds u>t and uu, then by Proposition 9.1, v(t, x) is well defined for all t > 0 and
||a;|| expt-^dliH)) < \\v(t,x)\\ < \\x\\ exp(-MIMI))-
(9-239)
Again by Proposition 9.1 we see that M M ) < jj^jjj Re (
(9-240)
This implies that
^ Re (
(9.241)
\\x\\ Now it follows by Proposition 9.1 that we also have
•Tr^pp 8 -<«**>*'££[
(9-242)
when 0 < a < w^(0) and b > uiu(0). This means that the functions uA{s) = a\^-
and
wA(s) = &^-±£
(9.243)
are also admissible lower and upper bounds for ip (and consequently for / ) . It is easy to verify that the function uA is increasing and that the function WA(S)S/(1 — s 2 )arctanhs is decreasing. Finally, (i') is equivalent to condition (ii) by Proposition 9.1 and to (iii) by Proposition 9.2. •
296
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
Remark 9.13 Actually, conditions (i)-(iii) of this proposition are equivalent to the following one: f G NT and there are numbers 0 < a < b such that a\\Bxf < Re(Bf'(T)x,
Bx) < b\\Bxf
(9.244)
for all x G H, where B is the linear operator defined by B = PT + V^l - ||T|| 2 (7 - PT), T G B\{0}, PT = r | ^ and B = I when r = 0. Indeed, by direct calculations one can show that y'(0) = Bf'(r)B~1 and (9.242) becomes (9.244),). In addition, by Proposition 9.1 and substituting the admissible lower and upper bounds (9.243), we conclude that the above conditions of Proposition 9.3 including (9.244) are equivalent to the inequality
(9.245) (compare with ft')).
Chapter 10
Geometry of Domains in Banach Spaces 10.1
Biholomorphic Mappings in Banach Spaces and Generators on Biholomorphically Equivalent Domains
Let X and Y be two Banach spaces over the field of complex numbers C, and let T> c X and ft C Y be domains (open connected subsets) in X and Y, respectively. Definition 10.1 A mapping / G Hol(D, ft) is said to be univalent on V if for each pair of distinct points x\ and x% in V we have f(x\) ^ /(a^). In this case one can define the inverse mapping f~l : f(V) H-> T>. It is well known (see, for example, [Herve (1963a)]) that if X = C" and Y = C m are finite dimensional complex spaces, then / : T> C C n — t > Cm, / G Hol(Z>, C m ), is univalent if and only if n = m and f"1 is also holomorphic on fi = f(T>), i.e., f~x S Hol(f2,T>). However, this fact is no longer true in the infinite dimensional case (see counterexamples in [Abts (1980)] and [Suffridge (1972)]). Therefore, in the general case we give the following definition. Definition 10.2 A univalent mapping / £ Hol(£>, fi), 9, = f(V), is said i > 2? belongs to Hol(fi,D). to be biholomorphic if f~x : fi — It is also known (see, for example, [Franzoni and Vesentini (1980)] and [Khatskevich and Shoikhet (1994a)] that if / € Hol(I>, ft) is biholomorphic, then for each point x £ V, the Frechet derivative A = f'(x) is a linear isomorphism between X and Y In this situation we will say that V and ft are biholomorphically equivalent. As we have already mentioned, in this case X and Y must be linearly isomorphic Banach spaces. Generally speaking, the converse is not true. That is, even if X and Y are isomorphic Banach spaces and / G Hol(P, V) has at each point x in V a continuously invertible Frechet 297
298
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
derivative, the mapping / needs not to be univalent on V. Nevertheless, in this case the mapping / is biholomorphic on a neighborhood of each x £ T> by the inverse function theorem. In this situation we will say that / e Hol(X>, Q,) is locally biholomorphic. The set of all univalent mappings from a domain T> C X into X will be denoted by Univ(P). For the special case when V is the open unit ball of X, the subset of Univ(P) normalized by the conditions /(0) = 0 and /'(0) = /
(10.1)
will be denoted by 5(2?). This notation conforms to the one used in the classical one-dimensional case, when V = A = {z € C : \z\ < 1}. In this case we simply write 5 (= S(A)) = {/ e Univ(A) : /(0) = 0 and /'(0) = l } .
(10.2)
That is, S consists of all the mappings / S Univ(A) such that / has the following Taylor series at the origin: oo
f{z) = z + Y,akZk.
(10.3)
k=1
The following simple assertion is the key to our subsequent considerations. L e m m a 10.1 Let T> and fl be two domains in a complex Banach space X such that Q, = f(D) for some biholomorphic mapping f : V H-> fi. Then the classes G(Q) andQ(T>) of generators (semi-complete vector fields) on f2 and T>, respectively, are linearly isomorphic, i.e., there is a linear invertible operator T from the space Hol(J7, X) onto the space Hol(£>, X) which takes the set (fi) onto the set g(D) (i.e., Q(V) = T(Q(SY)). Moreover, such an i > Q(T>) can be given by the formulae isomorphism T : Q(iY) —
?(?)(•) = [/'(-)]"V(/(-))
(10-4)
r-1(s)(-) = [/'(r1(-))M/"1(-)),
(io-5)
and
where ip e £(fi) and g e Q{T>).
Proof. Let
(10.6)
Geometry of Domains in Banach Spaces
299
is a semigroup of holomorphic self-mappings on V. Since the X-valued i » Q is differentiable for all t E [0, oo), so is the function S(-)(x) : M + — function G(-){x) : R + •-» V. Thus, for each x 6 T>, the strong limit g(x) = tUm i (x - G(t)(x))
(10.7)
exists and g E Q(V) is a generator on V. In other words, for each tp E G(ty, the mapping g = T(ip) belongs to £(£>), where T : Hol(Q, X) H-> Hol(I>, X) is defined by (10.4). It is clear that T is an invertible linear operator and T " 1 is given by (10.5). Changing the roles of ip and g, we see, by repeating the above considerations, that T " 1 takes Q{V) onto 5(fi). Lemma 10.1 is proved. O R e m a r k 10.1 It is clear that if if € aut(fi), then g = Tip € aut(2?) and conversely. If, in particular, T> = Q and f S Aut(2?), then the sets G(P) of ail semi-complete vector fields and aut(Z?) of all complete vector fields are invariant under the operator T : Hol(£>, X) •-> Hol(2?,X) defined by (10.4). i » Cl Remark 10.2 Let T> and Cl be two domains in X and let f : T> — be a biholomorphism of V onto fi. Define the operator T : Hol(f2,X) H-> Hol(I>, X) by (10.4). IftpE Hol(ft, X) has a null point b in fi, then so does g = Tip in V and a = f~1{b) is a null point of g in V. In addition, by direct calculations, we get g'(a) = [f'(a)}-1oiP'(b)of'(a).
(10.8)
Thus, the linear operators g'(a) : X t—» X and tp'(b) : X >—> X are similar, hence they have the same spectrum. In particular, if ip £ G(ty has a null point b € J7, which is either quasi-regular, regular or strictly regular, then a = / - 1 ( 6 ) is also a quasi-regular, regular or strictly regular null point of g, respectively. Using Lemma 10.1 and Remarks 10.1 and 10.2, one can present different parametric representations of semi-complete and complete vector fields. These representations are useful, inter alia, in finding geometric characterizations of biholomorphic mappings in Hilbert and Banach spaces. In the next section we will turn to a general description of spirallike and starlike mappings defined on the unit ball in Banach space.
300
10.2
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
Starlike, Convex, and Spirallike Mappings
Definition 10.3 A set M in X is called starshaped with respect to the point w0 if given any w G M, the point wt = (1 — t)w0 + tw also belongs to M for every t with 0 < t < 1. That is, if M contains w, then it also contains the entire line segment joining w to WQ. Definition 10.4 If V is a domain in X, then a biholomorphic mapping / G Hol(2?, X) is said to be a starlike mapping on V with respect to the point wo G ft, the closure of the image ft = /(£>) of V, if ft is a starshaped set with respect to the point WQ. • If WQ G ft, then we say that / is starlike with respect to an interior point. • If WQ G <9ft, the boundary of ft, then we say that / is starlike with respect to a boundary point. Definition 10.5 If P is a domain in X, then a biholomorphic mapping / G Univ(X>) is said to be a convex mapping on V if its image f(D) = ft is a convex domain in X. It is clear that a biholomorphic mapping / G Univ(2?) is convex if and only if it is starlike with respect to any point ifo G ft, the closure of the image ft = f(V). • A starlike mapping on V with respect to the origin is simply said to be, as in the classical case, starlike. In other words, a biholomorphic mapping / G Hol(P, X) is said to be starlike on T> if the closure ft of the image ft = f(D) of V is a starshaped set with respect to the origin. In these definitions the origin is in ft. If, in particular, the origin belongs to ft, then the mapping / has a null point r in T>. The set of all biholomorphic mappings on T> which are starlike on T> will be denoted by Star(P). If there is indeed a point r £T> such that f(r) = 0,
(10.9)
then we will write / G S*(T>). Of course, in this case such a point T is unique because S*T(V) C Star(D) c Univ(X>).
(10.10)
Again, in the one-dimensional case, when X = C, we will simply write S* to denote the family of all biholomorphic (univalent) starlike functions
Geometry of Domains in Banach Spaces
301
/ on the unit disk A normalized by the conditions /(0) = 0 and /'(0) = 1. That is, S* = SnS£{A).
(10.11)
In the sequel the spectrum of a linear operator A will be denoted by v(A). Definition 10.6 A set M in X is said to be spiral-shaped (with respect to the origin) if there is a bounded linear operator A : X —> X and a positive e such that ReA > e > 0 for all A e a{A) and such that for each w 6 M and t > 0, the point e~tAw also belongs to M. Definition 10.7 If P is a domain in X, then a biholomorphic mapping / s Hol(X>, X) is said to be a spirallike mapping on T> if the closure Cl of its image Q = f(T>) is a spiral-shaped set. • Once again, if the origin belongs to fi, then, as in the classical case, we will say that / is a spirallike mapping on V with respect to an interior point. • Otherwise, if the origin belongs to dQ. the boundary of ft, then we say
that / is spirallike with respect to a boundary point. The set of all biholomorphic mappings on T> which are spirallike (with respect to the origin) on V will be denoted by Spiral(D). If, in addition, there is a point r GT> such that / ( r ) = 0,
(10.12)
then we will write / € SpT(T>). Note also that if in Definition 10.6 the operator A = I, then fl is actually starlike, i.e. Star(Z>) C Spiral(P). Consequently, S*(V) C SpT{V).
10.2.1
Starlike functions on the unit disk
The concept of univalent starlike functions was first introduced by Alexander [Alexander (1915)] in 1915. In 1921, Nevanlinna [Nevalinna (1921)] made a more detailed study of this class. In particular, the following characterization of the class S* = S n SQ(A) is due to him. Theorem 10.1 disk A such that
Let f be a univalent holomorphic mapping on the unit /(0) = 0.
(10.13)
302
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
Then f is a starlike function on A if and only if
Re[^]>0, :eA.
(10.14)
Intuitively, this result follows (as does most of the work on starlike functions on the unit disk) from the identity
l-'<*">-"{=7$?}
<*»>
which is valid whenever the function / is holomorphic on A and not equal to zero at z = reie, r > 0, 0 G [0,27r]. Note also that if / G Hol(A.C) is locally biholomorphic, i.e., f'(z) ^ 0 everywhere, and satisfies (10.14), then it is necessarily univalent. Furthermore, condition (10.14) leads to the study of other interesting subclasses of Univ(A). In particular, in 1936 Robertson [Robertson (1936)] introduced the class S* (A) of starlike functions of order A: 5*(A) = if € S* : Re \^JT^\
> A > 0, Z € A | .
(10.16)
In 1978 Wald [Wald (1978)] characterized starlike functions with respect to another center. Using our notions, his result can be reformulated in the following way. Theorem 10.2
Suppose that f € Hol(A, C) is either of the form
(10.17)
f(z) = z + f2akzk fc=2
or of the form f(z) = Y^T=i bkZk with /'(r) = 1 for some r € A. Then the function g(z) = f(z) — /(r) belongs to S*(A) if and only ifReq(z) > 0, where 9{)~
f(z)-f(r)
~ z^r,
g(z) q(r) = 1 - |r| 2 .
(10.18)
We will see below that condition (10.18) can easily be obtained by using another approach in more general settings. Different applications of (10.18) are presented in Wald's thesis [Wald (1978)] (see also [Goodman (1983)]).
303
Geometry of Domains in Banach Spaces
10.2.2
Convex and close-to-convex functions disk
on the unit
Historically, the notion of a convex function arose earlier than the notion of a starlike function. This seems quite natural, since convexity has played a crucial role in the development of analysis and geometry. In 1913 Study [Study (1913)] described univalent functions on a closed disk, the image of which is a convex set. However, his condition employs the second derivative of the function. On the other hand, it is clear that a domain ft is convex if and only if it is starlike with respect to each one of its points. Using this fact, Sufferidge ([Suffridge (1973)] and [Suffridge (1977)]) gave another characterization of convex functions which also holds for higher dimensions (see Section 10.5). Here we quote another classical result due to Alexander [Alexander (1915)] which provides an analytic connection between convex and starlike functions. Theorem 10.3 Suppose that f is a locally biholomorphic function on the disk Ar = {z : \z\ < r}. Then f is convex on Ar if and only if the function f(z) = zf'(z) is starlike on A r . As we mentioned above, in the study of the functions of the form oo
f(z) = Y,*kZk
(10.19)
that are holomorphic and univalent on the unit disk A, certain subclasses, the members of which share some simple geometric property, arise rather naturally. Further developments in the classical theory have often had analytic generalizations and extensions. In 1952 Kaplan [Kaplan (1952)] defined the class of close-to-convex functions as those functions of the form (10.19) with ai = 1 such that
Re \^A]
> 0, z G A,
(10.20)
for some univalent convex function
Re[^]>0,
(10.21)
304
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
where h is starlike on A. If, in particular, / is starlike itself, then we can take h = f, and (10.21) holds bacause
(io-22)
Re[j(£]>0
for z G A. Note also that condition (10.21) is often used as the definition of a close-to-convex function on A, and that it is employed as a basic condition in higher dimensional generalizations (see Section 10.5). 10.2.3
Spirallike functions on the unit disk
It seems that the first occurrence of the class Spiral(A) = Spo(A) arose when condition (10.14) was modified analytically by inserting the factor ei9:
ReU£fflj>0,
Z£A.
(10.23)
(See Montel [Montel (1933)] and [Spacek (1933)].) Actually, this definition is compatible with our Definition 10.7. Proposition 10.1
Let f G Hol(A,C) have the form oo
f(z) = z+j2a*ak fc=2
( 10 - 24 )
(i.e., /(0) = 0 and /'(0) = I). Then f G Spo(A) if and only if condition (10.23) holds for some 9 e (-TT/2,TT/2). Note also that for afixed0 G (—TT/2, 7r/2), a function / satisfying (10.23) is called ^-spirallike. 10.3
Higher-Dimensional Extensions and the Dynamical Approach
Until 1970 the literature on geometric properties of biholomorphic mappings in higher dimensional space (C n , Hilbert spaces and Banach spaces) is rather limited (see [Cartan (1933)], [Matsuno (1955)] and [Suffridge (1970)]). Cartan [Cartan (1933)] was the first mathematician who suggested the study of starlike and convex mappings in several complex variables despite the fact that many properties of univalent functions on the unit disk (e.g., the Riemann mapping theorem, Caratheodory's theorem
305
Geometry of Domains in Banach Spaces
on kernel convergence and the Bieberbach-de Branges theorem) fail in the higher dimensional case. In 1970 Suffridge [Suffridge (1970)] established, inter alia, a necessary and sufficient condition for starlikeness which generalizes Theorem 10.1 to higher dimensions. He used the principle of subordination and onedimensional ideas due to Robertson [Robertson (1961)]. Furthermore, he also used a similar approach to describe starlike and spirallike domains in Banach spaces (see [Suffridge (1973)], [Suffridge (1977)], [Heath and Suffridge (1979)], [Gong (1999)], [Pfaltzgraff and Suffridge (1975)]). Pfaltzgraff and Suffridge [Pfaltzgraff and Suffridge (1975)] gave a characterization of close-to-starlike holomorphic mappings which in the onedimensional case coincides with a characterization of close-to-convex functions. Roughly speaking, the idea in these considerations is the following one. If / S S (i.e., f is a univalent holomorphic mapping on the unit disk A with /(0) = 0 and /'(0) = 1), then the condition of starlikeness (10.14) can be rewritten in the form = /'(*)$(*),
M
(10-25)
where g g Hol(A, C) has the form (10.26)
g(z) = zp(z) with Rep(z)>0,
z G A.
(10.27)
Conditions (10.26) and (10.27) are equivalent to 5(0) = 0
(10.28)
and Reg(z)z>0,
z <E A, z =£ 0.
(10.29)
The latter condition can easily be generalized to the case of Hilbert and Banach spaces. Namely, let X be a complex Banach space and let X* be the dual of X. By {x, x*) we denote the action of a linear functional x* in X* on an element x of X. Recall that the mapping J : X —* 2X" denned by J(x) = {x* £X*:(x,x*)
= \\x\\2=\\x*\\2},
x£X,
(10.30)
306
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
is called the (normalized) duality mapping. Let V be the open unit ball in X. We now define two families of holomorphic mappings on V. Let N denote the subset of those mappings g G Hol(£>, X) which satisfy 5(0) = 0 and
Re(g{x),x*) > 0
(10.31)
for all 0 ± x e V and x* 6 J{x). Let M={g€N: g'(0) = I},
(10.32)
where / denotes the identity operator on X. The following three assertions are due to Suffridge [Suffridge (1973)] and [Suffridge (1977)]. Theorem 10.4 Let f be a locally biholomorphic mapping on T> (i.e., f'(x) is a bounded linear operator with a bounded inverse for each x G T>) with /(0) = 0. Then f is starlike if and only if the following condition holds: /(*) = /'(*)[(*)],
XGV,
(10.33)
for some g G M. Note that condition (10.33) and the inclusion g € N imply the inclusion geM. Theorem 10.5
Let f be a locally biholomorphic mapping on T> and set f(x)-f(y)
= f'(x)[w(x,y)}
(10.34)
for all x,y G T>. Then f is a convex mapping on T> if and only if Re(w(x,y),x*) > 0 whenever \\y\\ < \\x\\ and x* G J(x). As it turns out, the class N is also useful in the characterization of spirallike mappings. Theorem 10.6 Let A be a bounded linear operator which is strongly accretive, i.e., there is e > 0 such that Re{Ax,x*} >e\\x\\2
(10.35)
for all x £ X and x* G J{x). Suppose that f € Hol(P, X) is a locally biholomorphic mapping on V which satisfies the conditions /(0) = 0
(10.36)
Geometry of Domains in Banach Spaces
307
and /'(0) = /.
(10.37)
Then f is spirallike (relative to A) if and only if it satisfies the equation A f{x) = f'{x)[g{x)}
(10.38)
with some g € N. Remark 10.3 Actually, Theorem 10-4 follows from Theorem 10.6. Indeed, if f is a locally biholomorphic mapping, then /'(O) = B is invertible. Setting A = I, one can consider the starlike mapping f = B~1 f which satisfies conditions (10.36) and (10.37) and equation (10.33) with g € N. As we have mentioned above, g must belong to M. Hence f also satisfies (10.33) with the same g. Remark 10.4 In the above remark we used the auxiliary mapping f because of the normalizing condition (10.37) of Theorem 10.6. This condition is essential in the proof of Theorem 10.6 in [Suffridge (1973)] and [Suffridge (1977)]. The crucial point in that proof is that condition (10.37), f'(0) = I, implies (0) — A. In fact, these conditions are needed because of the assumption (10.35) that A is strongly accretive. This assumption was used by Sufiridge [Suffridge (1977)] in his definition of spirallike mappings. Our definition 10.7 is more general, since the spectrum of each strongly accretive operator lies strictly in the right half-plane. This enables us, in particular, to avoid the normalization (10.37). Note also that in contrast with (10.35), our requirement of A is independent of any equivalent norm on X. Since we are interested in the geometric properties of fi, our definition seems to be more natural. Remark 10.5 It seems that for Cn, Hilbert and Banach spaces the papers [Pfaltzgraff (1975)] and [Pfaltzgraff and Suffridge (1975)] (see also [Gurganus (1975)]) were the first ones where the ideas of applying dynamical systems appeared. (In the one-dimensional case such ideas were employed earlier by Lowner, Kufarev [Golusin (1969); Goodman (1983)], Robertson [Robertson (1961)] and Brickman [Brickman (1973)].) Pfaltzgraff, for example, applied generalized Lowner differential equations to characterize subordination chains in Cn and univalent mappings on the unit ball. He also suggested this approach for Banach spaces.
308
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
Now we will turn to a general description of spirallike and starlike mappings denned on the unit ball in Banach space (see [Aharonov et al. (2003); Elin et al. (2001)] for the one-dimensional case and compare [Elin et al. (2000)] for a more general situation). Theorem 10.7 Let T> be the open unit ball in a Banach space X and let f € Hol(P, X) be e biholomorphic mapping on T>. Then f is spirallike on T> if and only if there exist a linear operator A such that its spectrum lies strictly inside the right half-plane and a real number m, m < 0, such that the following condition holds:
Re {{fizy'Afiz), z*) > m{\ - \\zf)
(10.39)
for all z CD and z* £ J(z). If A = I in (10.39), then f is actually starlike. In addition, if f has a null point in V, then it is spirallike (starlike) with respect to an interior point. Proof. If / is spirallike, then there exists a linear operator A with its spectrum strictly in the right-half plane such that for each y € $7 = f(T>) and t > 0, the element e~tAy is in Q. This means that the vector field (p — A is semi-complete on Cl. Hence, by Lemma 10.1, the vector field g = T(A) = [f'{-)]-1Af(-)
(10.40)
is semi-complete on T>. Therefore (10.39) is a consequence of formula (3.5.19) in [Elin et al. (2004)]. Conversely, if (10.39) holds, then, again by the above-mentioned formula, the vector field g defined by (10.40) is semi-complete on T>. Therefore A = T~1{g) is semi-complete on fi = /(£>). That is, for each t / e f i , the curve {e~tAy}t>o C 9. is the solution of the Cauchy problem
{
du •dl +
Au
= 0
(10.41)
u(0) = y. Now it follows from the assumption on A and Definition 10.7 that $7 is spiral-shaped. Note also that if / has a null point a £ T>, then so does g, and f(a) = g(a) = 0 S fi is the limit point of e~tAy for each y £ fi. Hence in this case / is spirallike with respect to an interior point. Otherwise, that is, if 0 £ dil, the mapping / is spirallike with respect to a boundary point. In addition, it is clear that if A = / , then the curve {e^ty}t>o is a straight line. In other words, Cl is star-shaped, and / is starlike. •
Geometry of Domains in Banach Spaces
309
Remark 10.6 As we saw in the proof of Theorem 10.7, in order to determine when f € Hol(P, X) satisfying (10.39) is spirallike (starlike) with respect to an interior point, we need to know whether f has an interior null point. If it does not, then there is a sequence {xn} C T> such that xn —• y € &D and f(xn) -» 0,
n -> oo.
Indeed, it is sufficient to set xn — f(e~nAz),
f
(10.42)
n = 0,1,2,... for any z £
Generally speaking, the above question is equivalent to the existence of a null point of the semi-complete vector field g defined by (10.40). Moreover, one can understand the condition of spirallikeness (starlikeness) via the differential equation f(z)g{z) = Af(z),
(10.43)
where g is a strongly semi-complete vector field. In addition, it turns out that if / is a solution of (10.43) such that f'(z) is invertible for all z £T>, then / is actually univalent, hence biholomorphic. Indeed, since g is strongly semi-complete, then by definition, it has an interior null point r £ V, which is locally uniformly attractive for the semigroup {Sg(t)}t>o generated by g. Furthermore, there is a neighborhood U C V of the point r such that / is biholomorphic on U. Denote V = f(U). Suppose by way of contradiction that / is not univalent on T>, that is, there are two distinct points x\ and xi in V such that f{x{) = ffa) = y S f(D). But then {Sg(t)xi} and {Sg(t)x2} both converge to r as t tends to infinity, and therefore there is to > 0 such that both Sg(to)x\ = u\ and Sg(to)x2 = u^ belong to U. Since u(t) = e~tA solves the Cauchy problem (10.41), we have Sg(t0) = f-1 o e~tA o / .
(10.44)
Hence Vi = /(«.-) = e~toAy eV,
» = 1,2.
(10.45)
This implies that yi = y2 and consequently, u\ = Sg(to)x\ = u 2 = Sg(to)x2 which is impossible by the univalence of Sg(to). Finally, note that (10.43) and the chain rule also imply the equality '
A = f'(T)og'{T)o[f'{T)r1,
(10.46)
310
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
which means that the spectrum of A lies strictly in the right half-plane. So we have proved the following assertion. Proposition 10.2 Let T> be a domain in X and let f € Hol(P, X) be locally biholomorphic on T>. Then f is spirallike if and only if there are a strongly semi-complete vector field g onV and a linear operator A : X —> X such that f satisfies the differential equation (10.43). //, in addition, T> is a hyperbolic domain endowed with a metric p which induces the norm topology, then there is a point r €T> such that / ( r ) = 0 and for each p-ball Br centered at T, Br = {x S V : p(x,r) < r}, the image f(Br) = £lr is spiral-shaped (star-shaped, when A = I). Remark 10.7 The last assertion follows from the fact that each p-ball Br centered at T is invariant under the semigroup {S(t)t>o generated by g. As a matter of fact, we will see below that in the above situation there is an equivalent norm of X and a ball B in this norm centered at r such that f(B) is spiral-shaped (star-shaped). Corollary 10.1 Let V be the open unit ball in X and let f G Hol(X>,X) be a locally biholomorphic mapping on T> which satisfies the condition
Re<[/ / (z)]- 1 i4/(z),z*>>a(||z||)||z||,
z 6 V, z* € J(z)
(10.47)
for some linear operator A : X —> X and some real continuous function a : [0,1] —> R such that a ( l ) > 0. Then f is a univalent spirallike (starlike) mapping on V. Remark 10.8 Condition (10.47) is a sufficient condition for f to be spirallike (or starlike) but it is not necessary. It would be nice, of course, to find a condition which is both necessary and sufficient for f to be spirallike (starlike). However, as we have already mentioned, to tackle this problem we need to find a condition which recognizes whether f e Hol(P, X) (or equivalently, g = [/'(•)]~ 1 ^/(") € Q{V)) has an interior null point in V. Generally speaking, these problems seem to be quite complicated. Moreover, there are examples of semi-complete vector fields on the closed unit balls T> of Banach spaces which have no mull point in T> [Khatskevich et al. (1995a)]. Nevertheless, these problems can be solved completely in the case of the Hilbert ball [Elin et al. (2004)]. Recall that for a bounded linear operator A : X - t I w e denote by a(A) the spectrum of A. The number K+(A) = max{ReA : A € a(A)}
(10.48)
311
Geometry of Domains in Banach Spaces
is called the upper exponential index of A. it is well known (see [Daletskii and Krein (1970)]) that K+(A) = Urn
loelleMll
( 10 - 49 )
Sllt
and for each u; > K + there is a positive number N = N(LJ) such that ||e M || < Ne"\
t > 0.
(10.50)
The number K-(A)
= min{ReA : A e a(A)} = ^
]2%t-—!!
(10.51)
is called the lower exponential index of A. We also have that for each UJ < K-, there is a positive number m = m(ui) such that | | e - M | | < me~ut. If now V is a bounded domain in X and g € G(V) has a null point T £ ~D, then in a neighborhood of r the mapping g can be represented by the Taylor series 00
g(x) = A(x-T) + 52Pk(x-T),
(10.52)
fc=fc where A = g'(r), I > 2, and P/t are homogeneous forms of order fc. Thus g is a strongly semi-complete vector field if and only if K~(A) > 0. Proposition 10.3 Let V be a bounded domain in X and let g e £/(£>) admit the representation (10.52) in a neighborhood of T €"D. Suppose that A = (T) satisfies the condition 0
(10.53)
Then for a given invertible operator B e L(X) such that BA = AB, the differential equation Af(x) - f'(x)g(x)
(10.54)
has a unique solution f € Ho\(V, X) which satisfies the initial conditions
(10.55)
312
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
If {S(t)}t>o is the semigroup generated by g, then this solution can be represented by the formula f(x) = lim BeAt(S(t)x t—>oo
- r),
(10.56)
where the limit in (10.56) is taken with respect to the topology of locally uniform convergence over T>. Proof. We assume for simplicity that T = 0. Step 1. First we note that there is a norm \\-\\A equivalent to the original norm of X and a ball B in this norm centered at r = 0 such that g e Q(B), i.e., B is invariant under the semigroup {S(t)}t>o generated by g. Indeed, there is m > 0 such that ||e~ M || < me" 11 '. for a given /J, G (Q,K-(A)), Setting ||a;|U = s u p | | e ( ' i / - 4 ) t x | | t>o
(10.57)
we get ||a;|| < ||a:||.4 < m.||a;|| and ||e- y l 'a:|U<e-'"||a;|U,
s > 0.
(10.58)
The last inequality implies in turn that for each x* € X* such that Re(z,z*> = \\x\\2A = \\x*\\2A (i.c, x* G JA{x)), Re(Ax, x*) > fj,\\x\\A > 0,
x ^ 0.
(10.59)
Now choose any ball Br = {x G X : \\X\\A < r} C T> centered at the origin with radius r > 0 and represent g G Q{D) by using the Taylor series (10.52) for T = 0. We have oo
g(x)=Ax
+ Y,Pk{x),
*>2,
(10.60)
k=i
where Pk are homogeneous polynomials of order k > I. If M = sup \\g{x)-Ax\\A, ||x||
(10.61)
then it follows by the Schwarz Lemma that \\g{x) - Ax\\A < ™ \\x\\lA
(10.62)
Geometry
of
Domains
in
Banach
Spaces
313
for all x £ BT. Therefore we get Re<5(z),z*>>M||o€i-^Nl!4+1 = IWIA(/^-^INI!4~1)>0
(10.63)
for all x £ B := {x £ X : \\x\\A < m i n - f ^ ) 1 ^ } } , x ^ 0 and x* £ JA(x). Hence g is a semi-complete vector field on B. Step 2. Let now h : B —> X be any bounded holomorphic mapping on B such that h(0) = 0 and T>nh(0) = 0 for 1 < n < I - 1.
(10.64)
we claim that lim etAh(S(t)x)
=0
(10.65)
for all x G B. Indeed, without any loss of generality we may assume that the radius of B is 1. As above, (10.64) and the Schwarz Lemma imply that \\h(x)\\A<M(h)\\x\\lA,
zeB,
(10.66)
where M(h) = sup H / i ^ m . In addition, it follows by Proposition 3.6.4 in [Elin et al. (2004)] that for each x&B,
l | 5 M ^ e " M t (T J M^( 1 -ll^ll-) 2
(10-67)
*e~"ii-tiu)*'
Now using (10.66), (10.67) and assumption (10.53), we can choose \i £ (0,rc_)and uj > K+ such that w < l^i. Then we get \\etAh(S(t)x)\\A
<
N(uj)e^\\h(S(t)x)\\A
< N{UJ) • M(h)eut\\S{t)x\\lA
< N{w)M{h)e^-1^
(1i^||)21
when t tends to infinity. The claim is proved.
- 0
(10.68) •
314
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
Step 3. Now we show that the limit in (10.56) exists for all s eT>. First, let us restrict ourselves to B. Consider the mapping u : R + X B H ^ defined as follows: u{t, x) =: eAtS(t)x,
t > 0, x G B.
(10.69)
By calculations and the definition of S(t), we get ^
^
= AeMS(t)x - eAtg(S(t)x) = -eAth(S(t)x),
(10.70)
where h(x) = g(x) - Ax obviously satisfies (10.64). So we get from (10.70) and (10.68) that for each x G B and h, t2 G WL+,
\\u(tliX) - (t2,x)\\A < N{u)M{h) jf*' e^"'")* • ( 1 ^ | | ) 2 . (10.71) Thus the limit in (10.56) exists pointwise. Denote this limit by / . Let now V be anyj>all strictly inside V. It follows from (10.53) that S{t) -> 0 uniformly on V. Hence there is a positive p such that y = S(p)x G B for each x G V. Furthermore, it follows from the semigroup property that for x£T>, lira BeA(p+t>S(p + t)x = lim = (liin BeApeAtS(t)y
BeApeAtS(t)S(p)x
= eApf(y) = eA"f(S(p)x).
(10.72)
This concludes Step 3. Step 4. Next we will show the solvability of equation (10.54). In fact, we claim that the mapping / : T> i-> X defined by (10.56) is a solution of this equation with the initial data /(0) = 0 and /'(0) = B. Indeed, substituting / = B lim eAtS(t) in the right-hand side of (10.54), we have by equation (5.2) in [Reich and Shoikhet (1996)],
f'(x)g(x) = BU^e" [ ^ ^ ( * ) ] = Stone«^(S(*)s) = B ton [etAg'(0)S(t)x + h(S(t)x)],
(10.73)
where h = g — A. Again, by Step 2, we conclude that lim eMh{S{t)x) = 0
(10.74)
t—too
and f'(x)g(x) = BA \ lim eAtS(t)x] = Af(x). U—»oo
J
(10.75)
315
Geometry of Domains in Banach Spaces
Step 5. Finally, it remains to show that if equation (10.54) has another solution which satisfies the same initial conditions, then it must coincide with the mapping / defined by (10.56). In fact, if / : V •-> X satisfies (10.54), /(0) = 0 and /'(0) = B, then we obtain
= -f>(S(t)x)^^
(10.76)
for all x £ V and t > 0. If we denote G(t, x) = f(S(t)x), equality means that
t > 0, then this
Af(S(t)x) = f'(S(t)x)g(S(t)x)
(10.77)
AG{tlx) = -™&*).. In addition,
(10.78)
G(0,x) = f(x). Therefore, solving (10.77) with the initial condition (10.78), we get G(t,x)=e-tAf(x),
(10.79)
or f{x) = etAG(t, x) = etAf(S(t)s) = etA [BS(t)x + h(S(t)x)], i £ D , t > 0, (10.80) where h = f — B. Letting t tend to infinity and using Step 2, we get f(x) = f(x). Proposition 10.3 is proved. Corollary 10.2 Let V be a bounded domain in X and let g e Q(T>) satisfy the conditions 9{r) = 0
(10.81)
g'(r) = /
(10.82)
and
for some r e V. Then for a given invertible operator B : X — i > X, the differential equation /(:r) = f'(x)g(x)
(10.83)
has a unique solution which satisfies the initial conditions /(r) - 0 and /'(r) = B.
(10.84)
316
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
If {S(t)}t>o is the semigroup generated by g, then this solution can be represented as f = Blhnet(S(t)-T).
(10.85)
Moreover, f is biholomorphic on V and f(D) is a star-shaped domain. Corollary 10.3 Let V be a bounded domain in X and let f :V »-> X be a starlike mapping on V with respect to an interior point r inV. Then there is a ball B C V centered at T such that f(B) is star-shaped. Moreover, if p is a metric on V assigned to V by a Schwarz-Pick system, then the image of each p-ball B centered at r is a star-shaped domain. Proposition 10.4 A bounded domain V in X is biholomorphically equivalent to a star-shaped domain if and only if there exists g € G(D) such that for some r e V, g(r) = 0 and g'(r) = I. 10.4
Distortion Theorems for Starlike Mappings on the Unit Ball
A celebrated theorem of Kobe asserts that the image of any univalent funcoo
tion on the unit disk of the form z + X) akZk covers the disk of radius fc=2
j centered at the origin. This fact is no longer true in higher dimensions (see [Cartan (1933)]). Nevertheless, Barnard, Fitzgerald and Gong showed that a similar assertion can be established for starlike mappings on the unit ball of a Banach space normalized at the origin (see [Barnard et al. (1991)]). An improved result was obtained in [Chuaqui (1995)] for the so-called strongly starlike mappings. Actually, covering estimates (as well as growth estimates) turn out to depend on lower and upper admissible bounds for semi-complete vector fields related to starlike mappings. Let T> be (as above) a domain in a Banach space X and let / € Hol(2?, X). We recall that / is starlike on V (with respect to an interior point r), i.e., f £ S*(D), if and only if it satisfies the differential equation f(x) = f'(x)g(x)
(10.86)
for some strongly semi-complete vector field g on V such that g(r) = 0
(10.87)
Geometry of Domains in Banach Spaces
317
and 9'(T) = I.
(10.88)
We will call such a mapping g the characteristic mapping of / , and the evolution equation
i^+9(a(t,x)),0
(io89)
(u(o,z) = x ev will be called the characteristic equation for / defined by (10.86). Recall also that if T> is bounded, then / can be found by the formula f{x) = A lim e*[u(t, x) - T], t—»oo
(10.90)
where A = / ' ( T ) and there is a ball B C T> centered at r such that / is starlike on B. Set fi{x) = A^f(x), x€T>. Then fx also satisfies (10.86) and is starlike on both V and B. So, one can try to establish growth and covering estimates of both / and f\ on B. If, in addition, V is a symmetric domain, then it can be realized as the unit ball in a 7S*-triple system, and we are able to study the behavior of these mappings on the whole of T> by using similar tools to the ones we employed in a Hilbert space. Indeed, first we note that in this case the domain T> is homogeneous, i.e., for each x £ V there is an automorphism M_T € Aut(P) such that M^T(T) = 0. The mapping / = /i °M-T is also starlike, since f(D) = f\(V) = J 4~ 1 /(P). At the same time, Lemma 10.1 implies that the mapping S=([M_ T (.)]')" 1 [S(M T (-))]
(10.91)
belongs to Q(T>) and is normalized by 9(0) = 0
(10.92)
g'(0) = / .
(10.93)
and
Hence, by the chain rule and (10.86), we get f{x) = /(M_T(z)) =
f'{M-T(x)MM-T(x))]
= ~nx)\[M-T{x))')-l[9(M-r{x))] - f'(x)g(x).
(10.94)
318
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
Thus g is the characteristic mapping for / and v(t, x) = M_T (u(t, MT(x)))
(10.95)
is the solution of the characteristic equation for / . Let now LJ = (o>i,o>2) be a pair of positive real functions on [0,1) such that the Cauchy problems
imA+^s)UJi{ms))=Q
I A ( 0 , S ) = SG[0,1)>
(io9g) i =
l,2,
have unique solutions 0i(t, s) G [0,1) for all t > 0, s G [0,1). Assume also that o>i is an increasing function while u>2 is decreasing. We will say that / is w-starlike on V if u^ and o>2 are admissible lower and upper bounds for g, i.e., for each x CD and x* G J(x), W2(\\x\\)\\x\\2 > Re
> wi(||a:|i)||x|| 2 .
(10.97)
Since ^'(0) = / , the existence of u> = (011,012) is again provided by the universal lower and upper bounds.
(10.98) However, even for the one-dimensional case there are many examples when wi and ^2 can be improved and may be chosen more precisely (see examples below). Let now 7$ be functions denned on (0,1) such that 7-(s) 1 T~r = 7~x i * = 1> A (10.99) 7i(s) suii{s) i.e., lnji(s) is a primitive function for a"'.v°A , i — 1,2. Observe that, since the Cauchy problem (10.96) is globally solvable for t G [0,00), and its solution Pi(t, s) goes to zero as t tends to 00, it follows that the integrals
IE®
(lal00)
are divergent and tend to 00. Hence 7* can be continuously extended to [0,1) by 7,(0) = 0. In order that 7* be denned uniquely, we set, for example, 7^(1/2) = 1. Then it is an increasing function on (0,1) by (10.99).
319
Geometry of Domains in Banach Spaces
Now (10.96) and (10.99) imply that 7i(A(t, s)) = e-' 7 i (s),
i = 1,2.
(10.101)
Recalling (10.95) and arguing as in the Hilbert space case, we obtain the inequalities
&(*, Ml) < H t , i ) | | < 0i(i, Ml)
(10.102)
which, in turn, will be used in the proof of the following assertion, where w = (wi,w2) [Elin et al. (2004)]. Proposition 10.5 Let X be a complex Banach space such that its open unit ball V is homogeneous and let f G Hol(I>, X) be w-starlike on V, with / ( T ) = 0 , T € V. Let 7i, i = 1,2, be the solutions of (10.99) defined above. Then the following estimates hold:
||A-1||-1A;(lkll)72(IN||) < ||/(Af_T(x))|| < tf(||s||)7i(|M|)|M|, (10.103) where A = /'(r), sup - i r , «e[o,P]7i(s)
0
(10.104)
k{p) := sup - 4 - r ,
0 < p < 1,
(10.105)
K(p):=
»€[0,p] 72 (S)
and M_ r is an automorphism ofV which takes r into 0. /n particular, /(£>) D P r ,
(10.106)
where T>r is the ball centered at the origin of radius
r = | | A - 1 | [ - 1 lim 72(s)ft(s).
(10.107)
s—>1-
Proof. Since 71 and 72 are increasing functions, we have by (10.101) and (10.102) that 7i(K*,a:)ll) <7i(/3i(t,IMD) = e - t 7 l ( b | | )
(10.108)
72(||«(t,x)||) > 72(/32(t, Ml)) = e-Sdl*!!)-
(10.109)
and
320
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
These inequalities and definitions (10.104) and (10.105) imply that
e - S ( H ) ' MINI) < Mt,x)\\ < e-S(NI) • K(\\x\\) (10.110) because \\v(t, x)\\ < \\x\\ by the Schwarz Lemma. Using now the fact that f(x) = A'ifiM-rix)) = l i m ^ o c e * ^ , ! ) , we get (10.103). Since f(V) = A~1f(T>) we also obtain (10.106) from the left-hand inequality of (10.103JI Remark 10.9 Observe that it follows by (10.92), (10.93) and (10.97) that Wi(0) < 1 while W2(0) > 1. If the limits lim i'As), i — 1,2, exist s—»0+
and are positive, then K(p) in (10.104) is finite and k(p) in (10.105) is positive. In this case u>i(0), i = 1,2, must actually equal 1. Indeed, it follows by (10.99) that
lim J^1-
s->0+
S
lim w^s) = 7-(0+) - ^ ( 0 ) ,
s->0+
i = 1,2.
(10.111)
Hence Wj(0) = 1, i = 1,2. In other words, in this case we must find functions uii such that UJJ(O) = 1, i = 1,2. Example 10.1 For each starlike mapping / G Hol(£>, X), lower and upper admissible bounds Ui(s), i = 1,2, can be chosen as follows: Wl(s)
= i_JI£
and
w2(s) = f ^
(10.112)
for some c £ [0,1]. If 0 < c < 1, then / is said to be strongly starlike. We have
^=(rr^'
72(S) =
(TT^
(10J13)
and K(p) = k(p) = -c .
(10.114)
So, we get the growth and covering estimates
Tnw- l / ( "- ( I ) ) l s ii^*iF-
(10115)
In particular, f(T>) D V
i „.
(10.116)
Geometry of Domains in Banach Spaces
321
Corollary 10.4 Let X be a complex Banach space such that its open unit ball is a homogeneous domain. Then the image of every starlike mapping (with respect to an interior point) covers the ball of radius | centered at the origin. Proof. Set c = 1 in (10.116). Note that the characteristic functions of many classical examples of starlike mappings have lower and upper bounds of the form Wu(s)
= lTAs.
(10.117)
In this case, solving the equations
(10.118) we get 7i,2(s) = j ~ ^ and K(p) = k(p) = 1. Thus we obtain the improved estimates
TT^<\\m\\
(10-119>
Now let us consider some different situations. Example 10.2
Let V = A be the open unit disk in C, and let f(z) = - L _ . -1-
Z
(10.120)
It is easy to see that / is a convex, hence a starlike function, /(0) = 0, and that its characteristic function is g(z) = z - z2 (recall that g(z) = f{z)/f'{z)). In this case u;,(s), i = 1,2, can be chosen as u>i(s) = 1 — s and w2(s) = 1 + s. Consequently, A = 1 and
TW\-W]-T^\'
(iai21)
In particular, /(A) D A I . Example 10.3 Let T>2 be the open unit ball in C 2 , a e C, and let ga:T>2>-> C2 be defined by 9a(zi,z2) := (zi-az%,z2).
(10.122)
322
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
We already know that this mapping is a generator if ||a|| < =^p and that wi,2 are of the form (10.117) with A = |a|;r4s . If now fa is the starlike mapping corresponding to ga, then we see that it satisfies the estimates (10.119). Example 10.4 In a similar way as in Example 10.3 one can consider a more general case: 9a,b(zi,z2) := (zi - azl, z2 - bz\).
(10.123)
Now the question is, for which a and b the mapping ga^ is a semi-complete vector field on Vp = {(zi.za) G C2 : |zi| p + N " < 1}.
(10.124)
We will give an answer for the cases p = 1 and p = 3 (for p ^ 1,3 the calculations are not so simple). 1. p = 1. In this case a support functional z* at z £ C2 is given by
(w,z*) = fwi^ii + ™ 2 ^ ) ( N + N ) \
zl
Z2
)
(10.125)
and
{ga,b(z), z*) = (\Zl\ + H - azl^ - bz^j (\zx\ + \z2\). (10.126) After some computations we get MI 2 (1 - A||z||) < Re(ffo,6(z), z*) < \\z\\\l + A||z||), (10.127) where
IW^Inl + lzal,
x=^%-
(10-128)
Thus 5a,b is a semi-complete vector field if J^J- + ^ > 1. (In fact, this condition is also necessary.) Therefore the starlike mapping corresponding to ga,b satisfies the estimates (10.119). In particular,
T T H s iA"Wi £ r n a
<10-129)
and
5*L
do.130)
323
Geometry of Domains in Banach Spaces
2. p = 3. Now the support functional at z G C2 is given by
(W,Z')= fulfil! + W2 N ! ) /(|2l|3 + N 3 ) i .
(lom)
Consequently, | Zl |3 + U |3_ flz 2]xij!L_ 6z 2j«j!
<Sa,t(*), O = —
—
21
,
Z2
(M 3 + M 3 ) 8
•
(10-132)
It is easy to observe that
Nl2 (i - ^
Ml) < **{9M, O < P f (i + ^
INIl), (10.133)
where ||z|| = (Ja;ij3 + |z2| 3 ) 3 . Therefore the mapping ga>b is a semicomplete vector field if \a\ + \b\ < 25 (Again, this condition is also necessary.) The corresponding starlike mapping fa^ satisfies the estimate (10.119) with A = 121^1. 23
Example 10.5
Consider the mapping / : X>2 >-> C2 defined by Uz)
:=
—!LZi2—
/2(z) := — i l ^ ^
*, 22 e Z2 .
(10.134) (10.135)
Calculations show that / is univalent on T>2 and 9(Zl,Z2) := [f'(zUZ2)]-1f(z1,Z2)
= (Zl(l + 22),*2(l + «l)) (10.136)
(cf. [Chuaqui (1995)]). Since g : V2 *-* C2 is a semi-complete vector field, it follows that / is starlike. In addition, in this case lower and upper bounds u)\ and ui2 can be chosen as in (10.117) with A = A- . Hence
iJ't'L.. ^ H/WH * T "?|| Z || and /(P2) covers an open ball of radius j(? .
(10-137)
324
10.5
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
Differential Equations for Starlike and Spirallike Mappings in H = Cn
In what follows we consider the complex Euclidean space H = Cn. In this case we write n
(10.138)
(z,w) = J2zkWk for vectors z = ( z i , . . . , zn) and w = (wi,...,
wn) in H. Thus
M = { z £ H : \\z\\ = y / \ Z l \ 2 + ••• + \ z n \ 2 < 1 }
(10.139)
is the open unit ball in 7i. For a linear operator F £ L{H, H) we will write ReF > 0 if ReA > 0 for all eigenvalues A of F. Note that this is equivalent to Re(F:r, x) > e\\x\\2, e > 0 , x £ H. As above we denote by / the identity operator on H. Recall that a univalent mapping h € Hol(B, H) is said to be spirallike on B if there exists a bounded linear operator T : H i-> H with ReF > 0 such that for each z € B and t > 0, e-trh(z) e h(M).
(10.140)
If this inclusion holds with F = / , the mapping h is called starlike. Furthermore, if h(z) = 0 for some point z £ B, the mapping h is said to be spirallike (starlike) with respect to an interior point. Otherwise, 0 € dh(M) and the mapping h is said to be spirallike (starlike) with respect to a boundary point. In this section we study a system of partial differential equations in B c C " connected to the classes of spirallike and starlike mappings on B with respect to a boundary point. It was shown in [Elin et al. (2000)] that such mappings are solutions of the differential equation Th(z) - h'(z)g(z),
(10.141)
where F G L(H, Ti) with ReF > 0 and / is an (infinitesimal) generator on B without null point. By h'(z) we denote the linear operator on H = C
325
Geometry of Domains in Banach Spaces
defined by the Jacobi matrix:
(
3fti(z)\
dhi(z)
:
:
:
dzx
at the point z G B.
(10.142)
dzn
Here we concentrate on generators of one-dimensional type (o.d.t.). In dhn(z) dhn(z) this case the above equation can be written in the following form: dzi '" dzn I Vh{z) = (1 - (z, T))P (z)h'(z)(z - r),
(10.143)
where r € 5 1 , p € Hol(B,C) with Rep(z) > 0, and T € L(H) is a bounded linear operator with ReF > 0. In particular, we will discuss in the sequel the existence, uniqueness and univalence of solutions of (10.143). By C : B >-> H we denote the Cayley transform of the unit ball: C ( Z ) =
1-(*,T)
( Z
+
T)-
(10-144)
By II we denote Siegel's domain in H: n = {w E H : Re(io,T) +
\(W,T)\2
> \\w\\2} .
(10.145)
Sometimes it is convenient to use the orthogonal projections P and Q denned by PZ = {Z,T)T
and
Qz = z - Pz,
(10.146)
and "partial coordinates": Z! = (Z,T)
and
z2 = Qz.
(10.147)
So, we write z — (zi,z2) for z = z\T + z2 £ H. Note that zx is a scalar while z2 is a 'vector coordinate'. We have
(10.148) and II = {w e H : Rewi > \\w2\\2 where wi - (W,T), W2 = QW). We also note that for any mapping F holomorphic in a domain fl C 7i, ^=F'(z)r,
(10.149)
326
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
where F'(z) denotes the Prechet derivative of F at z. The following more or less known fact will be useful in the sequel (c/. [Rudin (1980)]). Lemma 10.2 The Cayley transform maps B biholomorphically onto II and its inverse mapping is defined by z = C~1(w) = — ^ — W-T. 1 + Wl
(10.150) '
Moreover, (1 - zi)C(z)(z - r) = - 2 T
(10.151)
and d-^
= —T-T^-2™.
(10-152)
Lemma 10.3 A mapping h G Hol(B, H) is a solution of equation (10.143) if and only if the mapping h G Hol(II, H) defined by h(w) = ho C~1(w) is a solution of the differential equation ^ ^ + p (w)rh(w) = 0 ou>i
(10.153)
for some p G Hol(II, C) with Rep (w) > 0. The functions p G Hol(B,C) in (10.143) andpG Hol(II,C) in (10.153) are connected by the formula p{C{z))-p{z) = \.
(10.154)
Proof. Verifying (10.143) (respectively, (10.153)) with h = hoC (respectively, h = h o C~l) and p = ^ c (respectively, p = 2 p o ^ _ t ) , we get our assertion.
•
Using now Lemma 10.2 one can study the solvability and properties of equation (10.153) instead of equation (10.143). To proceed, we need the following lemma. Lemma 10.4
327
Geometry of Domains in Banach Spaces
(a) (fde Fabritiis (1994), Proposition 3.3]) For every function Hol(II, C) there exists k G Hol(II, C) such that ^
= b(W).
b G
(10.155)
(b) For any two functions k\,k2 G Hol(II,C), both satisfying (10.155), the difference re(u>) = ki(w) — k2(w) does not depend on w\ and so is an entire function on H, i.e.., re G Ho\(H,C). (c) The difference k(w\,W2) — k(l, W2) does not depend on the choice of a function k € Hol(II,C) satisfying equation (10.155). Proof. Assertion (a) was proved in [de Fabritiis (1994)]. Assume now that fci and k2 are two solutions of (10.155), i.e., dkQ^ = b(w), i = 1,2. Then for the function re(iu) — ki(w) — k2(w) we have dg^ = 0. So, this function is holomorphic in II and does not depend on w\ = {W,T). Hence, n is holomorphic in 7i. Indeed, for each W2 € Qti, one can find wi £ C such that Rexui > ||w2|| 2 , i.e., w = (11/1,1112) belongs to II. Since K G Hol(II, C) and does not depend on wi, it follows that K belongs to Ro\(QH,C), hence re e Hol(W,C), too. To prove assertion (c), we can now write k\ = k2 + re with dg^ = 0. Therefore, ki(wi,w2) - ki(l,w2) = [k2(wi,w2) + K(WI,W2)] - [k2{l,w2) + = [k2(wi,w2) + k2(l,w2)] +
[K{W!,W2)
-
K(1,W2)] K(1,W2)]
(10.156) D
= [k2{w1,w2)-k2(l,w2)].
Further, for any function k e Hol(2?,C), V C H, and any bounded linear operator F S L(H), one can define a holomorphic operator-valued mapping erk^ £ Hol(£>, L{H)) as follows:
e r f c (-) a = g^M r n a j n=0
aen_
(10.157)
"-
Let now k G Hol(II,C) satisfy (10.155) with b(w) =p(w) and ReF > 0 as in equations (10.143) and (10.153). For h € Hol(n,W) we define g G Hol(n,ft) by g{w) = erk^h{w).
(10.158)
328
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
L e m m a 10.5 A mapping h 6 Hol(Et, H) is a solution of equation (10.153) if and only if the mapping g defined by (10.158) does not depend on w\ and, consequently, g E Rol(H,H). Proof. Differentiating the mapping h G Hol(II, TC) denned by h(w) = e~rk(w^g(w), we have h(w + a) - h(w) = e-rfc(™){e-r(fc(™+a)-fc(™))p(w
= e-rk^{
+ a)
_
[e~nHw+a)-k(W)) _ j]g(w + a) + [5(w + a) _
g(w)}
g(w)]y
(10.159) Hence, we conclude that h'(w)a = e~rk(w) {g'(w)a - (k'(w)a)Tg(w)}.
(10.160)
h'(w)T = e- rfe W | ^ ^ - p (w)Tg(w) \ .
(10.161)
Consequently,
Substituting this expression in (10.153), we obtain that ^
= 0.
(10.162)
As in the proof of Lemma 10.4 one can conclude also that g £ Ho\(H,H). Conversely, let g be defined by (10.158) and assume that g'(w)T = 0. Then g{w + a) - g(w)
= e r f c ^ {[er(fc(tH-a)-fc(i")) _ j ^
w + a) +
fi(w
+ a)
_~h(w^J (10.163)
and g'(w)a = erk^{~ti(w)a + (k'(w)a)Th(w)}.
(10.164)
By our assumption for a = T we get the equality erk(w)
(l>^T
which implies our assertion.
+
p(w)Th(w))
= 0,
(10.165) n
Geometry of Domains in Banach Spaces
329
To formulate our next results we introduce the two following classes:
*:={&eHol(n,C):^=pM, where p (w) = 2 p ( c ! 1 ( w ) ) , R*P H > 0 J
(10.166)
and $ := {<j> £ Hol(QB, W) : e ^ 1 - ^ e Hol(QW.W) for some k £ K}. Theorem 10.8 (a) Let h £ Hol(B.W) be a solution of (10.143): Th{z) = (1 - zi)p (z)h'(z)(z - T).
(10.167)
Thenh(0,-) £ $ . (b) Let <j) £ <3>. Then equation (10.143) has a unique solution h £ Hol(B, H) which satisfies the initial data (10.168)
h(0,O =
h(z) = e-rk^z»
[erfc^-«^(O] | _ ,
(10-169)
and does not depend on the choice of k £ K. (c) //>£$, then the mapping h, defined by (10.169), is locally univalent on B if and only if, det(I>«) 0'(O) ± 0.
(10.170)
Proof, (a) Let h be a solution of (10.143). Then by Lemma 10.3 the mapping h = h o C " 1 is a solution of (10.153). It is clear that (10.171)
h(O,C) = h(l,C).
Define g by (10.158): g{w) = erk<-w)h(w), where k £ K. By Lemma 10.5, the mapping g does not depend on w\, i.e., g'(w)T = 0. Hence, actually, g £ Hol(QH,H). Then h(0,0 = h(l, C) = e-W^giO
£ *.
(10.172)
(b) Let >€$, i.e., for some k £ K,
0(0 = e-r^'StO,
(10-173)
330
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
and g G Hol(Q7t,?t). By Lemma 10.5 the mapping h, where h(w) = e-rk^g{w2),
(10.174)
w2 = Qw,
is a solution of (10.153), and the mapping h = hoC solves (10.143). Moreover, h(0,0 = ho C(0,C) = Ml,C) = e-W'QgiQ
= 4>(Q.
(10.175)
Let h*(0, C) = 0(0- As w e n a v e seen, for somefc*G K the mapping g*(w) = e ^ - ^ f t ^ C - 1 ^ ) )
(10.176)
does not depend on w\ and *, and, in fact, belongs to Hol(QH,H). Furthermore, by Lemma 10.4 we already know that k*(wi,w2) - k+(l,W2) = k(wi,W2) -
fc(l,w2),
(10.177)
where k G K. Then
M«0 = ^ r t > w ) ^ W L = c ( 2 ) _ e-r(fc.(tu1,iu2)-fe.(l,u;2))+rfe.(l,ii;2)o ( W 2 ) ru=C(x) =
re-r(fc(Wl,-lU2)-fc(i,-u,2)) rerk.(i,Qg^yii
= e-rfc(C(*))^rfc(i.C)0(C)]|
I
.
(10.178)
Consequently, a solution of (10.143) with (10.168) is unique and does not depend on the choice of a function fc G K. (c) Note that representation (10.169) can be rewritten in the form h = hoC, where
h(wi,o = Jwa-^-aUtt),
(f 1,0 e n,
(10.179)
and C is as above, the Cayley transform of the unit ball B. Since the Cayley transform is a biholomorphic mapping (see Lemma 10.2), it is enough to show that the mapping h defined by (10.179) is locally univalent if and only if (10.170) holds. To do this, we just calculate the Jacobian J(h) of the mapping h:
J(S) =
to(§^) = ( « % ^ d(tui,o) y dwi
cA ac
(Io.i8O)
J
331
Geometry of Domains in Banach Spaces
where gfe(QWl'C)
= er^0-^'0){-p(Wu<))r<(>(0
(10.181)
and
(10.182) So, J(fc) = -p( U ; 1 ,C)det(e r W 1 -«-^>O))det ^ 0 ( 0 ^ ) (10.183) (here we understand the operator er^li^^~*^f"1'^ as a matrix). This for• mula implies our claim. Example 10.6
Consider the function p € Hol(B, C) defined as follows:
(10-184)
P(*) = \±4-
Then using (10.154), we find the function p(w) = jf^pr • So solutions of equation (10.143) can be found by formula (10.168): h{z) = e~ir]°S\1^)iKO\
i - ,
,
.
( 10 - 185 )
where <j> € $. Take the operator (Hi 0 . . . 0 \ 0 /x2 . . . 0 r= . . . , \0
(10.186)
0 . . . /in/
where fit € C, i = 1,... ,n, and suppose that the mapping > satisfies the following condition: <W(O = -&(OC,
where
fa = (4>,T)£Hol(Qm,C)-
(10.187)
In this case, the coordinates of the solution h are seen to be
M z ) = ( ( i V « ? y •0i
(^
Q z
) a -zi)'
(i °- i88)
332
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
and C1_
2,^4-1
/
1
\
ftiW = " 7 4^T--^i i <5Z P*' * = 2,...,n. (10.189) (1 + Zf)a V 1 -- 2 ! / It is easy to verify (by using Theorem 10.8 (c)) that this mapping is locally univalent if and only if the function <j>\ does not vanish on QM. This fact will be also useful in the sequel. Prom the point of view of Theorem 10.8, the question of the (global) univalence of solutions remains open. Below we will prove the existence of a globally univalent solution for the important case in which T = 01, where f3 =-\(fT)'(T) := lim ((/ T )'(rr)). We already know that /? is a real T—>1~
nonnegative number. In other words, in the remaining part of this section we will deal with the differential equation Ph(z) = h'{z)f(z),
(10.190)
where f(z) = (1 - (z, T))P (Z)(Z - T), r G 3B, Rep(z) > 0, z £ 1.(10.191) Observe also that a univalent solution h (if it exists) of this equation is a star like mapping on B with respect to the boundary point r, because h(r) = lim h(rr) = 0 (see, for example, [Elin et al. (2000)]). r—>1~
Recall that a curve A : [0,1) i-» B is said to be asymptotically normal at a boundary point r 6 dM if
(10.192) and
i^-yj<M
0<s
(10.193)
where X(s) = (A(s),r) (see, for example, [Cowen and MacCluer (1995)] or [Rudin (1980)]). A holomorphic function ip : 1 H-» C is said to be differentiable at a boundary point r G 9 1 if for some number c € C and for every curve A : [0,1) i-> 1 asymptotically normal at r, the limit lim
fW ~c
z—>r, z€A \Z — T, r )
=:
ZIP'(T)
(10.194)
Geometry of Domains in Banach Spaces
333
exists finitely and does not depend on A. This limit is called the angular derivative of the function •$ at the point r G dM. We say that an o.d.t. generator (10.191) with r G dM is of a strongly hyperbolic type, if the angular derivative Z(/ T )'(r) := lim
(/(z)'T)
(= Z lim(l - {z,r))p(z))
Z—*T ^ 2 , T I — 1
= (3 (10.195)
Z—*T
exists with /? > 0, and the semigroup 5 = {Ft}t>o generated by / converges to T nontangentially (see [Cowen and MacCluer (1995)]). Theorem 10.9 Let f be a generator of a strongly hyperbolic type with Z(/ r )'(r) = p. Then the equation (10.190): ph(z) = h'(z)f(z)
(10.196)
has a globally univalent solution h € Hol(B, C n ) which is a starlike mapping of one-dimensional type. To prove this theorem we need the following lemma. Lemma 10.6 Let f £ G[T], T £ dB, be a generator of a strongly hyperbolic type with the angular derivative Z ( / T ) ' ( T ) = (3. Assume that there exists a real function c(t) : R+ H-» R+, c(t) ^ 0, and a sequence tk G M + , k = 1,2,..., tk —> oo, such that the locally uniform limit h{z) = lim c(tk){Ftk(z) k—>oo
- T)
(10.197)
exists. Then h satisfies the equation (10.198)
f3h(z) = h'(z)f(z). Proof.
First we note that by the Weierstrafi Theorem h'(z)=
lim c(tk)^^-.
fc—>oo
(10.199)
OZ
On the other hand, it is well-known that Ft, t > 0, satisfies the differential equation
^W+^£)/(2)
=o
(10 . 200)
(see, for example, [Reich and Shoikhet (1996)]). In addition, it follows from the Cauchy problem that ^^-f(z)
= f(Ft(z)).
(10.201)
334
Nonlinear Semigroups, Fixed Points, and Geometry of Domains
Thus, for given c(t) and tk, we obtain c(tk)^^-f(z)
= c(tk)(l - {Ftk(z),r))p(Ftk(z))(Ftk(z)~r). (10.202)
Noting that Jim (1 - (Ftk(z),r))P(Ftk(z)) = Jim $ $ $ $
= A (10-203)
and letting k tend to infinity in (10.202), we get ph(z) = h'(z)f(z), and we are done.
(10.204) •
Proof of Theorem 10.9 Let S = {Ft}t>o be the semigroup generated by / . Define ht £ Hol(B,W), t > 0, by
W-im-T.T)''*0-
(10 - 205)
Consider {ht(z),r) = [fy^Zl'l] • F « r any « > 0 and for all z £ B, the numerator and the denominator of this fraction take values in the left halfplane C_ = {£ G C : Re£ < 0} only. Hence, arg(/i t (^),r) < n, i.e., (ht{z),r) is never a real negative number. Therefore, the family {/it}t>o is normal (see, for example, [Montel (1933)]). Therefore there exists a sequence tk —* oo such that {htk} locally uniformly converges either to infinity or to a holomorphic mapping h £ Hol(B, H). But it follows from Theorem 10.8 that QFt(0) = 0, hence, ht(0) = T for all t > 0. Thus, for some sequence {tk}, tk —> oo, there exists the locally uniform limit h{z)= lim htk(z)£Eol(M,H). k—too
(10.206)
Setting c{t) = (Ft(o)-T,T)> w e °btain by Lemma 10.6 that the mapping h denned by (10.206) satisfies the differential equation 0h{z) = h'(z)f(z).
(10.207)
335
Geometry of Domains in Banach Spaces
Next we prove that the mapping h is locally univalent. Indeed, by Theorem 10.8, ™,M
lhn
1-(*U*),T)
-(HZ),T)
(10.208) Hence the restriction (f> = h\QM (which, by Theorem 10.8 (a), belongs to $ ) must satisfy the same condition: Q
=
(l{-(((%
QC =
~(
(10>209)
Since in our assumptions T = (31, we have
det(r0(O 0'(O) = Pdet{