• Nie Znaleziono Wyników

Basic properties of Lipschitz functions Let (X, d) be a semimetric space, i.e., a metric space for which the condition d(x, y

N/A
N/A
Protected

Academic year: 2021

Share "Basic properties of Lipschitz functions Let (X, d) be a semimetric space, i.e., a metric space for which the condition d(x, y"

Copied!
20
0
0

Pełen tekst

(1)

doi:10.7151/dmdico.1170

SPACES OF LIPSCHITZ FUNCTIONS ON METRIC SPACES

Diethard Pallaschke Institute of Operations Research

University of Karlsruhe – KIT D–76128 Karlsruhe, Germany e-mail: diethard.pallaschke@kit.edu

and

Dieter Pumpl¨un

Faculty of Mathematics and Computer Science FernUniversitaet Hagen

D–58084 Hagen, Germany e-mail: dieter.pumpluen@fernuni-hagen.de

Dedicated to Zbigniew Semadeni, the founder of Categorical Functional Analysis

Abstract

In this paper the universal properties of spaces of Lipschitz functions, defined over metric spaces, are investigated.

Keywords: categories of Lipschitz spaces, Saks spaces, base normed spaces.

2010 Mathematics Subject Classification: 46M99, 26A16, 46B40.

1. Basic properties of Lipschitz functions

Let (X, d) be a semimetric space, i.e., a metric space for which the condition d(x, y) = 0 does not imply x = y. If there is no confusion, we will omit the semimetric and will only write X instead of (X, d). Moreover, to exclude the trivial case we will assume that d ≡ 0 implies card(X) = 1. If several semimetric spaces occur, we will write (X, dX), i.e., take the space as index.

(2)

Definition 1.1. If X, Y are (semi) metric spaces with (semi) metrics dX, dY, a mapping f : X −→ Y is called Lipschitz iff there exists an M ≥ 0 such that (L) dX(f (x), f (y)) ≤ M dY(x, y) for all x, y ∈ X.

One puts

(1) L(f ) := inf{M | M ≥ 0 and dY(f (x), f (y)) ≤ M dX(x, y) for all x, y ∈ X}.

L(f ) is called the Lipschitz constant of f. If L(f ) ≤ 1, then f is called a contrac- tion.

Special cases of Definition 1.1 are if Y is a normed or a seminormed linear space and dY the metric induced by the norm; in the following we will mostly study the case Y = R or Y = CI. If, for a semimetric space X, edX denotes the equiv- alence relation on X induced by the semimetric dX then X/

dX

, the set of equiv- alence classes, carries a canonical structure of a metric space. A Lipschitz map- ping f : X −→ Y between two semimetric spaces induces a Lipschitz mapping f : X/ˆ

dX

−→ Y/

dY

with L(f ) = L( ˆf ). Hence, the theory of Lipschitz mappings between semimetric spaces cannot yield more information than the theory of Lip- schitz mappings between metric spaces. This is the reason why, in the following, only Lipschitz mappings on metric spaces are investigated.

Lemma 1.2. Let A ⊂ X be a non-empty subset of a metric space (X, d) and let distA: X −→ R with distA(v) = inf

x∈Ad(v, x) be the distance function of A. Then:

• the distance function is a Lipschitz function with Lipschitz constant 0 ≤ L ≤ 1.

• if A = X then L = 0 and if A 6= X and there exists a y ∈ X \ A which has a closed point to A, i.e., there exists a z ∈ A with d(y, z) = distA(y), then L = 1 and distA is an isometry.

Proof. Let x, y ∈ E and ε > 0 be given. By the definition of the infimum there exists a point z ∈ A with distA(y) ≥ d(y, z) − ε. We get

distA(x) ≤ d(x, z) ≤ d(x, y) + d(y, z)

≤ d(x, y) + distA(y) + ε.

By interchanging x and y

(2) |distA(x) − distA(y)| ≤ d(x, y)

(3)

follows. Hence, the distance function is a Lipschitz function with Lipschitz con- stant L ≤ 1.

If A = X then distA= 0. Now assume that there exists an y ∈ X \ A which has a closed point to A, i.e., there exists a z ∈ A with d(y, z) = distA(y), then

|distA(z) − distA(y)| = |0 − distA(y)| = d(z, y), which implies L = 1 by (2).

Remark 1.3. The Lipschitz functions on X separate points:

More precisely, let (X, d) be a metric space and let {x1, . . . , xn} ⊂ X be a finite subset of pairwise disjoint points, i.e., xi 6= xj for i 6= j. Then there exist Lipschitz functions ϕi : X −→ R such that for i ∈ {1, . . . , n}

ϕi(xj) =

1 : i = j

0 : i 6= j .

Proof. For i ∈ {1, . . . , n} put Ai := {x1, . . . , xn} \ {xi} and because of Lemma 1.2 define ϕi: X −→ R by ϕi(x) = distdistAi(x)

Ai(xi). Then ϕiis a Lipschitz function with these properties.

2. Spaces of Lipschitz functions

In the following let Lip denote the category of metric spaces and Lipschitz maps and define

LIip(X) := Lip(X, R), X ∈ Lip.

Furthermore, we introduce the following subcategories of Lip called Lip1 which is the subcategory defined by all contractions, Lip the subcategory generated by all metric spaces (X, d) of finite diameter, i.e., diam(X) = sup{d(x, y) | x, y ∈ X} < ∞ and finally:

Lip1 = Lip1∩ Lip.

For a metric space X consider the space L(X) of all real valued bounded Borel- measurable functions endowed with the supremum norm k k. This is a Banach space for any metric space X. Put for a metric space X

Lip(X) := LIip(X) ∩ L(X) and take as norm

kf kL:= max{L(f ), kf k}.

(4)

Note that the same definition makes sense for f : X −→ E in Lip and E ∈ Vec, where Vec denotes the category of real normed linear spaces and continuous linear maps and Vec1 the subcategory defined by linear contractions (For more details see [5] and [9]).

Lip(X) carries two norms k k and k kL and we have obviously

(∗) k k≤ k kL.

L(Lip(X)) denotes the closed unit ball with respect to k kL and (Lip(X)) the closed unit ball with respect to k k. Obviously

L(Lip(X)) ⊂ (Lip(X)).

holds. One defines

C(Lip(X)) := {f | f ∈ Lip(X), f (x) ≥ 0 for all x ∈ X} .

C(Lip(X)) is a proper, generating cone in Lip(X), i.e., Lip(X) = C(Lip(X)) − C(Lip(X)). That C(Lip(X)) is proper is trivial. For every f ∈ Lip(X) and every x ∈ X one has

(∗∗) −kf k≤ f (x) ≤ kf k,

which implies that C(Lip(X)) is generating. Furthermore, C(Lip(X)) is closed with respect to k kbecause C(Lip(X)) =T

x∈X{f ∈ Lip(X) | f (x) ≥ 0} and, moreover, because of (∗) also with respect to k kL.

In order to avoid mixing up both normed spaces, let Lip(X) := Lip(X), k k

and LipL(X) := Lip(X), k kL . Moreover, let us point out that the product, as well as the pointwise maximum and minimum of two bounded Lipschitz functions is again a Lipschitz function and let us denote by 1I ∈ Lip(X) the constant function 1I(x) = 1, for all x ∈ X.

1I ∈ Lip(X) is an order unit with respect to k k (not k kL).

Furthermore, observe that Lip(X) is in general not a Banach space. To see this, take X := [0, 1] the unit interval with d(x, y) = |x − y|. Then the function f : [0, 1] −→ R with f (x) := √

x is not Lipschitz, but it is the uniform limit of the Lipschitz functions fn : [0, 1] −→ R with fn(x) := min{nx,√

x}, because supt∈[0,1]|fn(t) − f (t)| = 4n12.But LipL(X) is a Banach space ([12] 1.6.2 ).

Let us recall some notations: If C is an arbitrary cone of a vector space E, then a convex subset B ⊂ C is called a base of C if every z ∈ C \ {0} has a unique representation z = λb with λ > 0 and b ∈ B. Every cone C ⊂ E induces a

(5)

partial order ≤ by x ≤ y if and only if y − x ∈ C. A partial order ≤ is called archimedian if, for some y ≥ 0 and all λ > 0 x ≤ λy implies x ≤ 0. For a, b ∈ E let [a, b] := {z ∈ E | a ≤ z ≤ b}. Moreover, we use the notation a ∨ b = max{a, b}

if the maximum exists and correspondingly the minimum a ∧ b = min{a, b}, and write |a| for a ∨ (−a). A norm k k for E is called a Riesz norm if, for all a, b ∈ E, the inequality |a| ≤ |b| implies kak ≤ kbk. An element e ∈ C is called an order unit if for every z ∈ E there exists a λ > 0 such that λe ≤ z ≤ λe.

Remark 2.1. Let E be a vector space and C ⊂ E a generating cone.

• If C has an order unit e ∈ C then the function

z 7→ kzk = inf{λ > 0 | − λe ≤ z ≤ λe}

is a norm for E, which is called an order unit norm.

• If C has a base B ⊂ C such that the set S = conv (B ∪ −B) is order bounded, then the Minkowski functional

p(z) = inf{λ > 0 | z ∈ λS}

is a norm for E. We call this norm p a base norm and call E a base normed space. The base is denoted by B = Bs(E) and C = R+Bs(E) holds with R+ = [0, +∞).

For a real vector space E we denote by E the algebraic dual, that is the vector space of all linear forms from E to R. If E is endowed with a locally convex Haus- dorff linear topology τ then the pair (E, τ ) is called a locally convex topological vector space and we denote by E0 its topological dual, that is the vector space of all continuous linear forms from E to R.

A Saks space is a triple (E, k k, τ ) where k k is a norm on the real linear topological space E and τ is a locally convex Hausdorff linear topology τ on E such that the unit ball k k(E) is τ -closed and τ -bounded. For any normed vector space (E, k k) the triple (E0, k k0, σ(E0, E)) is a Saks space, where k k0 is the dual norm and σ(E0, E) the weak-* topology on E0.

Proposition 2.2. For a metric space X the space Lip(X) := Lip(X), k k endowed with the pointwise order of functions is a regular ordered order unit normed space with the closed and generating order cone C(Lip(X)) the order unit 1I ∈ Lip(X) and (Lip(X)) = [−1I, 1I]. Lip(X) is in general not complete.

Proof. The equation (Lip(X)) = [−1I, 1I] follows from (∗∗). The proof of the remaining assertions is straightforward.

(6)

Proposition 2.3. LipL(X) is a Banach space and the cone C(Lip(X)) is gener- ating and k kL-closed but k kL is not a Riesz-norm with respect to C(Lip(X)).

Proof. The completeness of LipL(X) is shown in [12], Proposition 1.6.2, and that C(Lip(X)) is a proper generating cone was shown above as was the closedness.

Example 2.4. Define the function

f : R −→ R given by f (x) :=

2x : x ∈ [−12,12] 1 : x ≥ 12

−1 : x ≤ −12 ,

then L(f ) = 2 and kf k = 1 which implies kf kL = 2, i.e., L(Lip(R)) ⊂6=

(Lip(R)).

Note furthermore, that if A ⊂ Lip(X) is k k-closed, then A is also k kL-closed, i.e., for the topologies,

τk k ⊂ τk kL, holds which follows from (∗).

Remark 2.5. For a metric space X, LipL(X), k kL, k k is a Saks space with the topology of k k, i.e., a 2−normed linear space in the sense of Semadeni [10], who investigated these spaces which led to the introduction of Saks spaces.

C(Lip(X)) is k kL-closed, proper and generating but does not make LipL(X) a regular ordered Saks space because k kL is not a Riesz norm with respect to C(Lip(X)) (see [7]).

As we have seen, k k ≤ k kL implies L(Lip(X)) ⊂ (Lip(X)), i.e., L(Lip(X)) is k k-bounded and (Lip(X)) is also k kL-closed.

Additionally it should be noted, for further use, that C(Lip(X)) is 1-normal (see [13], Prop. 9.2 (e), p. 86), i.e., g ≤ f ≤ h implies kf k≤ max {kgk, khk} . Moreover, Lip(X) is a Stonian vector lattice (see [2] p. 186).

As for any normed linear space (E, k k), (E, k k, σ(E, E0)) is a Saks space, we have the canonical Saks spaces Lip(X), k k, σ (Lip(X), Lip(X)0) and Lip(X), k kL, σ (LipL(X), LipL(X)0) . The first one is even a regularly ordered Saks space because Lip(X) ∈ Vec+1 (see [7], Example 3.2 iii) ).

3. The duals of Lipschitz functions spaces

Now the connection between Lip(X) and LipL(X) will be investigated as well as between their (topological) dual spaces Lip(X)0 and LipL(X)0 which are

(7)

both linear subspaces of Lip(X). Inequality (∗) implies that in the commutative diagram

LipL(X ) -

Id

#

Lip(X )

λ ∈ Lip(X )0

R Q

Q Q

Q Q

Q Q

QQs λ ◦ Id

the identity map is a contraction but not a quasi-isometry( see [12], pp. 3–4).

The dual norm of Lip(X)0 will be denoted by k k0 and the dual norm of LipL(X)0 by k k0L. Now Id induces an injective contraction κX : Lip(X)0 −→

LipL(X)0 with κX(λ) := λ ◦ Id, which may be considered as an inclusion, hence we often write κX(λ) := λ.

For λ ∈ Lip(X)0 one has:

X(λ)k0L = sup {|λ(f )| | kf kL≤ 1}

≤ sup {|λ(f )| | kf k≤ 1} (because of (∗))

= kλk0, i.e., taking κX as inclusion, then

(∗0) k k0L≤ k k0.

Hence, we have

0(Lip(X) ⊂ 0L(Lip(X)), i.e., one may regard Lip(X)0 as a subspace of LipL(X)0.

In the following we will show that in general Lip(X)0 is a proper subspace of LipL(X)0, Lip(X)06=LipL(X)0. For this we use the construction of point derivations from D.R. Sherbert [11] (see also [12] Chapter 7) which we briefly outline.

Consider the real Banach space

l:= {x := (xn)n∈N | (xn)n∈N bounded sequence } endowed with the supremum norm

kxk:= sup

n∈N

|xn|.

(8)

Let c ⊂ l denote the closed subspace of all convergent sequences and let lim : c −→ R be the continuous linear functional which assigns to every convergent sequence its limit. Consider a norm-preserving Hahn-Banach extension ”LIM” of the functional ”lim” to l:

c -

l

? LIM

R Q

Q Q

Q Q

Q Q

Q s lim

with the following additional properties:

(i) LIMn→∞xn= LIMn→∞xn+1, (ii) lim inf

n→∞ xn≤ LIMn→∞xn≤ lim sup

n→∞

xn.

We shall use the the notation LIMn→∞xn= LIM(x) for x := (xn)n∈N∈ l. These functionals ”LIM” are called translation invariant Banach limits (see [4], Chapter II.4, Exercise 22).

Now let (X, d) be a metric space and ∆ := {(x, x) ∈ X × X | x ∈ X} the diagonal of X × X. For a double sequence w ∈ ((X × X) \ ∆)N, w := (xn, yn)n∈N) one defines the sequence

Tw(f ) := f (yn) − f (xn) d(yn, xn)



| n ∈ N



, f ∈ Lip(X).

This yields a mapping

Tw : Lip(X) −→ l with Tw(f ) := f (yn) − f (xn) d(yn, xn)



n∈N

,

which satisfies kTw(f )k ≤ L(f ) ≤ kf kL, i.e., a continuous linear map Tw : LipL(X) −→ l. Hence for any Banach limit LIM the composition Dw= LIM◦Tw

is a continuous linear functional

Dw : LipL(X) −→ R with Dw(f ) = LIM(Tw(f )), i.e., Dw ∈ LipL(X)0.

(9)

As the definition of Dw resembles the classical definition of the derivation of functions, one is interested if and under which assumptions Dw is a derivation in the sense of Bourbaki [3] i.e., for f, g ∈ Lip(X) and x ∈ X,

(P R0) Dw(f g) = f Dw(g) + gDw(f )

holds, i.e., in our case, as the left side does depend (directly) on x ∈ X (P R) Dw(f g) = f (x)Dw(g) + g(x)Dw(f )

where the dependence of Dw(f g) on x ∈ X has to be specified. The answer to this is (see [11], Prop. 8.5)

Proposition 3.1. If (X, d) is a metric space and for w ∈ ((X × X) \ ∆)N, w := (xn, yn)n∈N) satisfies x0 = limn→∞xn = limn→∞yn, i.e., x0 ∈ X is non- isolated, then (P R) is satisfied and Dw is called a point derivation in x0.

Proof. If (an)n∈N, (bn)n∈N ∈ l, then, for convergent (an)n∈N, with a = limn→∞an one has for any Banach limit LIM

LIMn→∞(an· bn− αbn) = 0,

because |an· bn− αbn| = |an − α||bn| ≤ B|an− α| if k(bn)n∈Nk = B, which implies limn→∞(an· bn− αbn) = LIMn→∞(an· bn− αbn) = 0.

Hence

LIMn→∞(an· bn) = αLIMn→∞bn follows. This implies, for f, g ∈ Lip(X)

Dw(f g) = LIM(Tw(f g))

= LIMn→∞

 (f g)(yn) − (f g)(xn) d(yn, xn)



= LIMn→∞



f (yn)g(yn) − g(xn)

d(yn, xn) + g(xn)f (yn) − f (xn) d(yn, xn)



= f (x0)LIMn→∞

 g(yn) − g(xn) d(yn, xn)



+ g(x0)LIMn→∞

 f (yn) − f (xn) d(yn, xn)



= f (x0)Dw(g) + g(x0)Dw(f ) which completes the proof.

(10)

In [11], Proposition 8.5, it is shown that for x0 isolated any Dw ≡ 0 if w fulfills the condition of Proposition 3.1.

Proposition 3.2. In general, Lip(X)0 is a proper subspace of LipL(X)0, Lip(X)06=LipL(X)0.

Proof. Consider the metric space X := [0, 1] with the usual metric d(x, y) =

|x − y|. Let Dw ∈ LipL([0, 1])0 be any point derivation for x = 0 and define the sequence ϕk, k ∈ N, k > 1, in LipL(X) by

ϕk(x) := max (√

kx,

√k

1 − k(x − 1) )

.

Then

max ϕk(x) = ϕk(1 k) = 1

√k.

This implies that the sequence (ϕk)nk∈N converges to 0 with respect to k k. A trivial calculation shows Dwk) = √

k. Since the constant function 0 has Dw(0) = 0 and the sequence (Dwk))k∈N is unbounded, it follows that Dw 6∈

Lip(X)0 which completes the proof.

4. The space of point functionals

A central role in our investigations plays, for a metric space X, x ∈ X, the mapping

δX : X −→ Lip(X) with δX(x)(f ) := f (x), f ∈ Lip(X).

First note that for every x ∈ X the linear functional δX(x) ∈ Lip(X) is con- tinuous with respect to both norms k k and k kL. The restriction of δX to Lip(X)0 is denoted by δX and to LipL(X)0 by δLX. The upper indicees are omitted if misunderstandings are not possible.

Proposition 4.1. Let (X, d) be a metric space. Then (a) For all x ∈ X, kδX(x)k0= kδXL(x)k0L= 1 holds.

(b) δXL : X −→ LipL(X)0 is a contraction.

(c) δX is injective and δX(X) is a linearly independent set.

(11)

Proof. (a): For both norms k kand k kL one has kδLX(x)(f )k0L = sup {|δX(x)| | f ∈ L(Lip(X))}

= sup {|f (x)| | f ∈ L(Lip(X))} ≤ sup {|f (x)| | kf k≤ 1}

= kδX(x)k0≤ max {kf k, 1} ≤ max {kf kL, 1} ≤ 1.

Hence, as k1I(x)k = 1, kδX(x)k0L= kδX(x)k0= 1 follows.

(b): One has

LX(x) − δXL(y)k0L = sup|δXL(x)(f ) − δXL(y)(f )| | kf kL≤ 1

= sup {|f (x) − f (y)| | kf kL≤ 1}

≤ sup {kf kLd(x, y) | kf kL≤ 1} = d(x, y).

To prove (c) let α1, α2, . . . , αn∈ R and x1, x2, . . . , xn∈ X be given with xi 6= xk for i 6= k, and assume thatPn

j=1αjδX(xj) = 0 holds. By Remark 1.3 there exist ϕ1, ϕ2, . . . , ϕn∈ Lip(X) with

ϕi(xj) = δij =

1 : i = j

0 : i 6= j NowPn

j=1αjδX(xj)(ϕi) = 0 yields αi= 0.

Next we consider the linear subspace generated by the point functionals

D(X) := hδXi :=

( λ =

n

X

i=1

αiδX(xi) | x1, . . . , xn∈ X, α1, . . . , αn∈ R, n ∈ N )

of Lip(X).

Remark 4.2. By DL(X) we denote the space D(X) endowed with the dual norm k k0L and by D(X) we denote the space D(X) endowed with the dual norm k k0. The canonical injections are δXL and δX. At the beginning of Section 2 it was already pointed out, that a normed linear space E and a metric space X with metric d, the notions Lipschitz function for a mapping f : X −→ E and kf kas well kf kL are defined analogously to the case E = R. Hence, for δX and δLX one may try to compute these norms. As δX is not Lipschitz only δXL remains. For the sake of brevity we denote the ∞-norm of δXL : X −→ LipL(X)0 with kδXLk and get from Proposition 4.1

(12)

LXk= sup



X(x)k0L | x ∈ X



= 1, which implies kδLXkL= 1 for the L-norm, as L δLX ≤ 1.

Theorem 4.3. Let X be a metric space, E ∈ Vec1 with norm k k and ϕ : X −→ E, ϕ ∈ Lip with kϕkL≤ 1. Then there exists a unique ϕ0 : DL(X) −→ E in Vec1 with ϕ = (ϕ0XL and kϕ0k = kϕkL such that

X -

δLX

(DL(X))

? 0)

(E) Q

Q Q

Q Q

Q Q

Q s ϕ

commutes.

Proof. We may assume that ϕ 6= 0 because otherwise the statement is trivially true. The assumption kϕkL ≤ 1 implies ϕ(X) ⊂ (E), hence we may restrict ϕ to (E) in its image domain. As misunderstandings are not possible we will denote the restriction also by ϕ. As {δX(x) | x ∈ X} is a basis of DL(X) the linear mapping ϕ0: D(X) −→ E is well defined by

ϕ0

n

X

i=1

αiδX(xi)

! :=

n

X

i=1

αiϕ(xi)

and satisfies ϕ = ϕ0◦δX. A routine calculation shows that ϕ0is a linear mapping.

In order to prove kϕ0k = kϕkL, let λ ∈ E0 with λ 6= 0, then kλ ◦ ϕkL ≤ kλk0kϕkL and hence λϕ

kλk0kϕkL ∈ (LipL(X)) . For ξ =Pn

i=1αiδLX(xi) ∈ DL(X) there exists a λξ ∈ E0 withkλξk0 = 1 and λξ0(ξ)) = kϕ0(ξ)kE. Now:

0(ξ)kE = λξ ϕ0 n

X

i=1

αiδXL(xi)

!!

= λξ n

X

i=1

αiϕ(xi)

!

=

n

X

i=1

αiλξ◦ ϕ(xi)

≤ kϕkL

n

X

i=1

αi

λξ◦ ϕ kϕkL (xi)

= kϕkL

n

X

i=1

αiδXL(xi)

! λξ◦ ϕ kϕkL



(13)

= kϕkLsup



|ξ(f )| = |

n

X

i=1

αif (xi)|

f ∈ L(Lip(X))



≤ kϕkLkξkL

which implies kϕ0k ≤ kϕkL≤ 1.

On the other hand, kϕkL= kϕ0◦ δXLkL≤ kϕ0kkδXLkL= kϕ0k (because of Remark 4.2) and hence kϕkL= kϕ0k follows.

Remark 4.4. For a metric space X, Theorem 4.3 means that the mapping δXL −→ (DL(X)) is universal with respect to all Lipschitz mappings ϕ : X −→ E, kϕkL ≤ 1 where E is a norned real linear space. This is equivalent to the following statement:

The unit ball functor : Vec1 −→ Lip1 has DL : Lip1 −→ Vec1 as a left adjoint.

Proof. The proof is elementary because maps a contraction in Vec1 to a contraction in Lip1and Theorem 4.3 states that δXL −→ (DL(X)) is a canonical universal embedding for a metric space into a normed real linear space.

Proposition 4.5. Let (X, d) be a metric space. Then D(X) is a base normed linear space with base Bs (D(X)) = conv ({δX(x) | x ∈ X}) , the convex hull of δX(X), the order cone C (D(X)) = R+Bs (D(X)) and the base norm

kλkB =

k

X

i=1

αiδX(xi) B

=

k

X

i=1

i|

if λ =Pk

i=1αiδX(xi)is the representation of λ in the basis δX(X).

Proof. Lip(X) is an order unit normed linear space (see Proposition 2.2) with order unit 1I. The order in Lip(X) is pointwise. Hence, well-known results (cp, e.g. [13], Chapt. 9) imply that Lip(X)0 is a base ordered linear space with cone

C(Lip(X)0) :=λ | λ ∈ Lip(X)0, λ(f ) ≥ 0 for all f ∈ C(Lip(X)) and base

Bs Lip(X)0 = λ | f ∈ Lip(X)0, λ ≥ 0 and λ(1I) = 1

= C(Lip(X)0) ∩λ | λ ∈ Lip(X)0 and λ(1I) = 1 . We define

C (D(X)) := C Lip(X)0 ∩ D(X) and B := Bs Lip(X)0 ∩ D(X).

Of course, B is a base set in D(X) with B ⊂ C (D(X)) , hence, R+B ⊂ C (D(X)) .

(14)

If λ :=Pn

i=1αiδX(xi) ∈ C (D(X)) , then Remark 1.3 yields at once αi ≥ 0, 1 ≤ i ≤ n. The converse implication is obviously true. If λ > 0, then kαk :=

Pn

i=1αi> 0 follows and

(1) λ = kαk

n

X

i=1

αi

kαkδX(xi) holds. On the other hand, λ =Pn

i=1αiδX(xi) ∈ B is equivalent to λ(1I) = 1, i.e., λ(1I) =Pn

i=1αi = 1. This shows that B is the convex hull of {δX(x) | x ∈ X} , i.e., B = conv {δX(x) | x ∈ X} and because of (1), C (D(X)) = R+B, hence Bs (D(X)) := B is a base for C (D(X)) .

For a basis representation of λ =Pn

i=1αiδX(xi) ∈ D(X), λ 6= 0 put kα+k :=

n

X

i=1 αi≥0

αi , kαk :=

n

X

i=1 αi<0

αi.

Then

(2) λ = kα++− kα

with

λ+=

n

X

i=1 αi>0

αi

+X(xi) for kα+k 6= 0 and λ+ := 0 else, and analogously for λ.

As a subset of Bs (Lip(X)0) the set Bs (D(X)) is linearly bounded. Hence the base seminorm induced by Bs (D(X)) is a norm (cp. [13]). We denote this norm by k k0 for the moment. One of the possible representations for a base norm is:

kλk0 = inf {β + γ | β, γ ≥ 0, λ = βξ − γη, ξ, η ∈ Bs (D(X)) } . For λ :=Pn

i=1αiδX(xi) (2) implies

kλk0 ≤ kα+k + kαk =

n

X

i=1

i|.

Now, let λ = βφ − γψ be a second representation. We may take the union of δX(x), x ∈ X, which appear in the basis representation of λ, φ, and ψ and denote it by {δX(xi) | xi∈ X, 1 ≤ i ≤ n} . Let λ =Pn

i=1αiδX(xi), φ =Pn

i=1ϕiδX(xi), and ψ =Pn

i=1ψiδX(xi) where ϕi ≥ 0 and ψi ≥ 0 for 1 ≤ i ≤ n. Then

n

X

i=1

αiδX(xi) =

n

X

i=1

(βϕi− γψi) δX(xi)

(15)

hence αi = (βϕi− γψi) and |αi| = |βϕi− γψi| ≤ βϕi+ γψi. As φ, ψ ∈ Bs (D(X)) one hasPn

i=1ϕi =Pn

i=1ϕi= 1 andPn

i=1i| ≤ β + γ follows such that kλkB=Pn

i=1i| bas been proved.

In view of Theorem 4.3 it is natural to ask if a similar result can be proved for δX : X −→ Bs (D(X)) . This is indeed the case as the following Proposition shows.

Let us denote the category of base-normed linear spaces and linear base pre- serving mappings (which are, by the way contractions) by BN-Vec1. If, for E ∈ BN-Vec1, Bs (E) denotes the base of E, then

Proposition 4.6. Let X be a metric space, E ∈ BN-Vec1 and ϕ : X −→ Bs(E) a Lipschitz mapping. Then there exists a unique ϕ0 : D(X) −→ E in BN-Vec1 such that

X -

δX

Bs(D(X))

? Bs(ϕ0)

Bs(E) Q

Q Q

Q Q

Q Q

Q s ϕ

commutes, and 1 = kϕ0k = kϕk, where kϕk= sup {kϕkE | x ∈ X} .

Proof. As δX(X) is a basis of D(X) the linear mapping ϕ0 : D(X) −→ E is well defined by

ϕ0 n

X

i=1

αiδX(xi)

! :=

n

X

i=1

αiϕ(xi) and makes the above diagram commutative.

The proof of the equality of norms is trivial. As ϕ0 is a contraction kϕ0 ≤ 1 holds. Also kϕkk = 1 as ϕ(X) ⊂ Bs(E). Moreover

1 = kϕ(x)kE ≤ kϕ0kkδX(x)k0= kϕ0k ≤ 1, i.e., kϕ0k = 1.

Despite the fact that the Lipschitz constant or the norm k kL does not appear explicitly in Proposition 4.6 the result is nonetheless quite interesting for metric spaces.

(16)

Corollary 4.7. Let (X, dX) be a metric space. Then

(i) δX: X −→ Bs (D(X)) is a universal embedding into the base of the based normed linear space D(X), i.e. the canonical functor

Bs : BN-Vec1 −→ Lip1 has the functor D : Lip1 −→ BN-Vec1 as a left adjoint with the adjunction morphism δX: X −→ Bs (D(X)) . (ii) δX: X −→ Bs (D(X)) is a universal contractive embedding into a metric

convex module, i.e., if C is a convex module (see [6]) and f : X −→ C is in Lip1, then there is a unique affine mapping f0 : Bs (D(X)) −→ C with f = f0◦ δX.

If Met-Conv denotes the category of metric convex modules, or, what is the same, of metric convex subsets of real linear spaces [6] and affine mappings, Bs (D(X)) may be regarded as a functor BsD : Lip1 −→

Met-Conv which is left adjoint to the canonical forgetful functor U : Met-Conv −→ Lip1 assigning to every C ∈ Met-Conv its underlying metric space and δX induces the adjunction morphism.

Proof. (i) is just a reformulation of Proposition 4.6 (ii) results by straightforward arguments from Proposition 4.6 and the results in §2 of [6].

Corollary 4.7 (ii) shows an interesting fact, namely the canonical and close con- nection between metric and convex structures.

5. The predual

There is another interesting topology, which was first investigated by R.F. Arens and J. Eells [1], and which will be discussed in this section.

The following proof uses a method completely different from the one used in [1] and is considerably shorter.

Define

ג : LipL(X) −→ D0L(X) and k : D0L(X) −→ LipL(X) by

ג(f )(λ) := λ(f ), f ∈ LipL(X), λ ∈ D0L(X) and

k(λ)(x) := λ(δXL(x)), λ ∈ D0L(X), x ∈ X.

(17)

Theorem 5.1. ג : LipL(X) −→ D0L(X) and k : DL0 (X) −→ LipL(X) are in Vec1 and

ג ◦ k = idDL0(X) and k ◦ ג = idLipL(X), x ∈ X holds. Hence, ג and k are isometries.

Proof. Let us denote the norm dual to k k0L of DL(X) on D0L(X) by k k#L. For x, y ∈ X one has if λ ∈ DL0 (X) :

|k(λ)(x) − k(λ)(y)| = |λ δLX(x) − λ δXL(y) |

≤ kλk#Lk(δXL(x) − δLX(y)k0L≤ kλk#Ld(x, y) (because of Proposition 4.1).

Hence L(k(λ)) ≤ kλk#L follows. Moreover

|k(λ)(x)| = |λ δLX(x) | ≤ kλk#LXL(x)kL= kλk#L and we have

kk(λ)k= sup {|k(λ)(x)| | x ∈ X} = sup {|λ δXL(x) | | x ∈ X}

≤ kλk#Lsup {|δLX(x)| | x ∈ X} ≤ kλk#L. This yields

(∗) kk(λ)kL= max {kk(λ)k, L (k(λ)) } ≤ kλk#L i.e., kkk ≤ 1, k is a contraction.

For f ∈ LipL(X) one gets kג(f)k#L = supn

|ג(f)(λ)|

λ ∈ DL(X) and kλk0L≤ 1o

= sup n

|λ(f )|

λ ∈ DL(X) and kλk0L≤ 1o

≤ kλk0Lsupn kf kL

kλk0L≤ 1o

≤ kf kL, which gives

(∗∗) kגk = supn

kג(f)k#L

kf kL≤ 1o

≤ 1.

i.e., ג is also a contraction, kגk ≤ 1.

(18)

For λ ∈ DL0 (X) and x ∈ X ג(k(λ)) δLX(x)

= δLX(x) ((k(λ)) = k(λ)(x) = λ δXL(x) , holds for all x ∈ X, which yields ג◦k(λ) = λ, because δLX(x) | x ∈ X is a basis of DL(X) and hence

ג ◦ k = idD0L(X). Also for f ∈ LipL(X) and any x ∈ X

k(ג(f ))(x) = ג(f ) δLX(x) = δXL(x)(f ) = f (x) which results in k(ג)(f ) = f and hence

k ◦ ג = idLipL(X),

i.e., k and ג are inverse to each other. This together with (∗) and (∗∗) yields the assertion

Remark 5.2. To show the dependence of X, an index X will be added: גX

and kX, because both are natural transformations between interesting functors.

The interesting topology mentioned at the beginning of this section is the dual topology σ (D0L(X), DL(X)) transferred by k (and ג) to LipL(X).

References

[1] R.F. Arens and J. Eells, Jr., On embedding uniform and topological spaces, Pacific J. Math. 6 (1956) 397–403. doi:10.2140/pjm.1956.6.397.

[2] H. Bauer, Wahrscheinlichkeitstheorie (5te Auflage), de Gruyter Lehrbuch, Walter de Gruyter & Co. (Berlin, 2002).

[3] N. Bourbaki, ´El´ements de math´ematique. XI. Premi`ere partie: Les structures fon- damentales de l’ analyse, no. 1102. Hermann et Cie. (Paris, 1950).

[4] N. Dunford and J.T. Schwartz, Linear Operators: Part I, Interscience Publishers, Inc. (New York, 1957).

[5] D. Pumpl¨un, Elemente der Kategorientheorie, Hochschultaschenbuch (Spektrum Akademischer Verlag, Heidelberg, Berlin, 1999).

[6] D. Pumpl¨un, The metric completion of convex sets and modules, Result. Math. 41 (2002) 346–360. doi:10.1007/BF03322777.

[7] D. Pumpl¨un, A universal compactification of topological positively convex sets, J.

Convex Anal. 18 (4) (2011) 999–1012.

[8] S. Rolewicz, Metric Linear Spaces, PWN – Polish Scientific Publishers, Warszawa and D. Reidel Publishing Company (Dordrecht, 1972).

[9] Z. Semadeni, Banach Spaces of Continuous Functions, Vol. I PWN – Polish Scientific Publishers (Warszawa, 1971).

(19)

[10] Z. Semadeni, Some Saks-space dualities in harmonic analysis on commutative semi- groups, Special topics of applied mathematics (Proc. Sem., Ges. Math. Datenverarb., Bonn, 1979), 71–87 (North-Holland, Amsterdam-New York, 1980).

[11] D.R. Sherbert, The structure of ideals and point derivations in Banach Algebras of Lipschitz functions, Trans AMS 111 (1964) 240–272.

doi:10.1090/S0002-9947-1964-0161177-1.

[12] Nik Weaver, Lipschitz Algebras, World Scientific (Singapore, New Jersey, London, Hong Kong, 1999).

[13] Yau Chuen Wong and Kung Fu Ng, Partially Ordered Topological Vector Spaces, Oxford Mathematical Monographs. Clarendon Press, Oxford, 1973.

Received 31 May 2015

(20)

Cytaty

Powiązane dokumenty

Zhang [4] introduced cone metric space by generalized the concept of a metric space, replacing the set of real numbers, by an ordered Banach space and obtained some fixed point

Indeed, a rather useful theorem of Kechris, Louveau and Woodin [6] states that a coanalytic σ-ideal of compact sets is either a true coanalytic set or a G δ set.. For each

In Section 4, we develop a capacity theory based on this space; we study basic properties of capacity, including monotonicity and countable subadditivity, as well as several

It is well known that any complete metric space is isomet- ric with a subset of a Banach space, and any hyperconvex space is a non- expansive retract of any space in which it

Banach–Mazur distance, nearly (almost) isometric Banach spaces, ` 1 -preduals, hyperplanes in c, weak ∗ fixed point property, stable weak ∗ fixed point property, almost stable weak

this relationship, by showing that the density result is a straightforward consequence of the uniqueness condition via the above-mentioned “guided”.. selection theorem in L 1 and

Within this approach, in [2]–[7], [13], [18], [21] qualified decay was obtained for the first boundary value problem for the wave equation in an exterior domain for which the

In this section we shall develop some theorems involving equi­. valent