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!
F be a linear homeomorphism. Prove that there exists k E N such that for every x £ E we have K. Now define h : J 2 —> J 2 by h(x,y) = ( H ( x , a ( y ) ) , y ) . By Theorem 1.8.2 it follows that h is a homeomorphism. If (x,/(x)) G F then (1) [ai,&i]. For every t G [ai,&i] let Ht be the unique homeomorphism of Ji taking [—!,£] linearly onto [—1,0] and [t, 1] linearly onto [0,1]. Then H: JJi x [01,61] -4 Ji defined by H(x,t) — Ht(x) is an isotopy (Exercise 1.8.1). Now put F(x,y) = ( H ( x , X ( y ) ) , y ) . Then F belongs to IK(Q) by Theorem 1.8.2. In addition, F has the property that for every ( x , y ) 6 /[AT], F(x,y) = ( H ( x , X ( y ) ) , y ) = ( H ( x , x ) , y ) = (0,y). Jn X Jm < « / 2 , (8) h^WftcW^, (9) h\K=lK. Proof. Observe that (7) follows as in the proof of Lemma 1.6.4. Since (8) and (9) are trivial, we are done. 0 By squeezing the (m — l)-cell J771"1 into its interior (see Figure 19) we can similarly construct a homeomorphism ip: JTm —> J m such that (10) (/) and let x G P. If x G A then x e Q and ?(/) (x) = f(x) and so there is nothing to prove. Suppose therefore that x g A. Let UQ be the collection of all elements of U which contain x. Observe that au 6 Q • CP(Y) be continuous and linear. Then supp(A) is bounded in X if A is bounded in Y. Proof. Striving for a contradiction, assume that supp(A) is not bounded. Then there exists an element / G CP(X) such that /[supp(A)j is not bounded in R. We may assume without loss of generality that N C /[supp(A)]. For every n G N, let Vn = f ~ l [ ( n - V 4 ,n + 1/4)]- Then V = {Vn : n e N} is a discrete family of open subsets of X. By induction on k we will construct a point y^ G A, an element n(k) G N, and an element fk G CP(X) such that (1) fk[X\VnW]C{0}, ( 2 ) n(l) 1 and suppose we completed the construction up to k — 1. Let Then P is clearly finite, so there exists n(k) > n(k — 1) so large that 1, C(k,m,n) = {y£Y : (3 At e(y,m,n)) ( A\ l/n- By continuity l/n- We can find a basic neighborhood V(B,p) of 0, where B C X is finite and p > 1, such that for all g G CP(X] with f-g€ V(B,p) we have \ l/n- Since the sets Ai \ ^4 are pairwise disjoint there exists an i such that (Ai \A)r\B = fy. Let g G CP(X) be such that Then g G F(yli,m) so that \(p(g}(y)\ fact that / - g G V(B,p). ). Observe that the subspace topology that A inherits from A((^>) is independent of 3~. Indeed, A \ {q} is discrete and a basic neighborhood of g has the form where for certain finite £ C D and n < u). We will now describe our example. As on Page 46, let E denote U n R by • v?(oo) = (0,1), which proves that ip is continuous. To prove that (p~l is continuous at (0, 1), let tn in [0, 1) and xn G X be arbitrary such that ((1 — tn)xn, tn) -> (0, 1). Then tn —> I and hence regardless of the behaviour of the sequence (xn)n, we have (xn,tn) -> oo in A(X). ^Exercise 1.2.9: By Theorem 1.2.2 there exist continuous functions /i : LI —> LI and /2 : 1/2 —> LI such that /i f Ai = /i and /2 t ^2 = /i"1. Define fl"i, Hi : L\ x L 2 —>• LI x L2 by That HI and /^2 are homeomorphisms of LI x £2 follows by elementary considerations (continuity is clear, and it is easy to display their inverses). Then H ~ H-zoHi is as required. ^•Exercise 1.4.2: Let tn —> t in I and xn —> x in L. We have to prove that {3(tn,xn) —>• /3(x,t) in L. [E] C s, B = • B by I be continuous. If / has no fixed-point, then I is the disjoint union of the open sets E = {x 6 I : x < /(#)} and F = {x E I : f(x) < x}. By connectivity of I, at most one can be non-empty, say E. But then 1 < /(I) contradicts the fact that the range of / is contained in I. The proof of (2) is trivial. For (3), let /: X -»• X be continuous. Put E = {(0, y) : -1 < y < 1}. If f[E] C E then we get a fixed-point in E by (1). So assume that /[F] is not contained in F. Then f[E]r\E = 0 since by Exercise A. 2. 11 it follows that /[F] is locally connected and no locally connected subcontinuum of X intersects both E and X\E. Observe that by compactness the function / is closed (Exercise A. 5. 5(1)). There is a closed neighborhood V of E such that f[V] O E = 0. We may assume that V = X \ {(x,sin %):e<x< 1/vr}
\\x\\/k< M*)|| <*N|. 1.2. Extending continuous functions
Suppose that X, Y and Z are topological spaces with Y a subspace of X and let /: Y —> Z be continuous. In topology it is often of interest to know whether / is the restriction to Y of a continuous function /: X —> Z. Easy examples show that in general this need not be the case. If / is the restriction to F of a continuous function /: X —> Z then we say that / is continuously extendable over X or that / is a continuous extension of /. In this section we shall present examples of spaces Z having the property that if Y is closed in an arbitrary space X, then every continuous function /: Y —>• Z is continuously extendable over X (respectively, over some neighborhood of F). Observe that Urysohn's Lemma can be looked at as a result on extending continuous functions. For let X be a space and let A and B be disjoint closed subsets ofX. If E = A\JF then by Lemma A.4.1 it follows that the continuous function /: E -» I denned by / \ A = 0 and f \ B = I can be extended to a continuous function /: X —> E. The function /: E —>• I is not very interesting and the question naturally arises whether it is also possible to extend more interesting functions. In order to study this question with success, let us first formulate and prove the following technical result.
22
1. BASIC TOPOLOGY
Lemma 1.2.1. Let X be a space, A a closed subset of X, and let A' C A be dense in A. Then there exist a locally finite open cover U of X \A and a sequence of points [au : U E U} in A' such that (1) for all U 6 U and x € U, g(x, au) < 2@(x, A), (2) if Un E U for every n and
lim g(Un,A) = 0
n —>-oo
then lim diam(t/ n ) = 0.
n—>oo
Proof. Let V={B(x,VtQ(x,A)):xeX\A}. Since A is closed, V is an open cover of X \ A. By paracompactness of X \ A (Corollary A. 7. 3) there exists a locally finite open cover U of X \ A that refines V. Since U < V, for each U E U there exists xu E X \ A with UCB(xu,l/4Q(xu,A)). In addition, since A' is dense in A, for each U E U there exists au E -4' with We claim that the C/'s and the Of/'s are as required. Claim 1. For every C7 G U and x G C/ the following inequalities hold:
Proof. The first inequality is easy since
g(x, au) < Q(X,
Also, Q(XU, A] < Q(XU,X) + g(x, A) < %g(xu, A) + g(x, A), from which it follows that %g(xu,A) < g(x,A), as required.
So it remains to verify (2). To this end, assume that Un G U for all n and that lim g(Un,A) = 0 . n—>ao
For each n pick pn G Un such that linin^oo g(pn, A) = 0. By the claim we obtain lim Q(xUn,A) < lim 4/3^Pn, A) = 0. n—too
n—>oo
1.2. EXTENDING CONTINUOUS FUNCTIONS
23
Since Un C B(xun, 1/4^(x[/n, A)) for all n, we get limn^.oo diam(t/n) = 0. So we are done. D An open cover It and a sequence of points {au '• U G U} such as in this lemma is called a Dugundji system for X and A. We now come to the main result in this section. The Dugundji Theorem (Part 1) 1.2.2. Let L be a locally convex linear space and let C C L be convex. Then for every space X with closed subspace A, every continuous function f : A -» C can be extended to a continuous function f : X — > C . Remark 1.2.3. For Part 2 of The Dugundji Theorem, see Page 394. Proof. Let the open cover U of X \ A and the sequence of points (au in A(= A) be a Dugundji system for X and A. In addition, let KU '• X\A for U G U be the Ac-functions with respect to U. Define /: X -» L by (±\
f f(~.\ _ J
f(T\ /W
' ~ i v^
,,-(~\
tt~^\
(rx pfc A\ l ^)i
(x £ X \ A)
We will first prove that / is well-defined and continuous at all points of X\A. To this end, fix an arbitrary x 6 X \ A. Since U is locally finite, there is a neighborhood W of x in X \ A meeting finitely many elements of U only, say C/i, . . . , Un. For U 6 U missing W we clearly have As a consequence, for every y £ W we have
n=l
So J(y] is a convex combination of points in C and therefore belongs to C. By continuity of the ^-functions, it also follows that / |~ W is continuous. It suffices to prove continuity of / at the points of A. Pick an arbitrary element a G A. An arbitrary neighborhood of /(a) is without loss of generality of the form /(a) + W, where W is a convex neighborhood of 0 (Exercise 1.1.4). So let /(a) + W be such a neighborhood. The continuity of / at a implies that there exists 6 > 0 such that B(a, 6) n A C f~l[f(a) + W]. Claim 1. /[B(a, %)} C /(a) + W. Proof. Pick an arbitrary x G -B(a, 5/s)- ~H x E A then there is nothing to prove. So assume without loss of generality that x 0 A. Then Q(x,A) < g(x,a) < %;
24
1. BASIC TOPOLOGY
as a consequence, if x E U E U then
g(a, au) < g(a, x) + g(x, < e(a,x) + 2g(x,A) <S by (1) of Lemma 1.2.1. From this we conclude that if x E U E U then we have au E #(a, 6) (~\ A and so /(at/) E /(a) + W. Consequently, / » - / » = ( J] ^ ( x ) - /(at/)) -/(a)
Since /(at/) - /(a) 6 W for every C7 6 U we see that f ( x ) — /(a) is a convex combination of elements of W. Hence by convexity of W we obviously get that f ( x ) — /(a) E W, as required. 0 We conclude that / is continuous at a.
D
Remark 1.2.4. It is a natural problem whether the local convexity assumption in Theorem 1.2.2 can be dropped. This was a fundamental open problem ever since DUGUNDJI'S paper [140] appeared in 1951. It was finally solved in 1994 by CAUTY [87] in the negative. His construction used in an essential way DRANISNIKOV'S result in [139] about the existence of an infinitedimensional compactum with finite cohomological dimension. As a corollary to the Dugundji Theorem we get (cf. Theorem A. 4. 6): The Tietze Theorem 1.2.5. For every space X with closed subspace A, every continuous function from A to E or I can be extended over X. A space X is called an Absolute Retract (abbreviated AR) provided it is a retract of every space Y containing it as a closed subspace. If X is an AR and / : X —> Y is a homeomorphism then Y is an A R as well. Consequently, X is an A R if and only if for every space Y containing a closed subspace Z which is homeomorphic to X, there exists a retraction r: Y —>• Z. Theorem 1.2.7 below implies that a retract of an AR is an AR. A space X is called an Absolute Neighborhood Retract (abbreviated ANR) provided it is a neighborhood retract of every space Y containing it as a closed subspace. The space X = {0, 1} is easily seen to be an ANR but is not an AR; simply observe that by continuity retractions preserve connectivity, and so there does not exist a retraction r:I-*X. Notice that every AR is an ANR. As above, Theorem 1.2.7 below easily implies that X is an A N R if and only if for every space Y containing a closed subspace Z which is homeomorphic to X, Z is a neighborhood retract of Y. Also, every neighborhood retract of an A N R is again an ANR (Proposition 1.2.10).
1.2. EXTENDING CONTINUOUS FUNCTIONS
25
We call a space X an Absolute (Neighborhood) Extensor (abbreviated A(N)E) provided that for every space Y and for every closed subspace A of y, every continuous function /: A —> X can be extended over Y (over a neighborhood (depending on /) of A'mY). We shall prove in Theorem 1.2.7 below that X is an A(N)R if and only if X is an A(N)E. This is of fundamental importance. In the sequel we shall not always conscientiously refer to Theorem 1.2.7 when dealing with A(N)R's. The reader should keep this in mind. We first prove an important fact. Lemma 1.2.6. Every space can be imbedded as a closed subspace of some AE. Proof. This is easy. First observe that every space is homeomorphic to a closed subspace of some normed linear space (Corollary 1.1.8). Now apply Theorem 1.2.2. D We shall now present the announced characterization of A(N)R's. Theorem 1.2.7. Let X be a space. The following statements are equivalent: (1) X is an A ( N ) R , (2) X is an A(N)E. Proof. The implication (2) =>• (1) is trivial. For (1) => (2), let us assume that X is an ANR. The proof for AR's is entirely similar, and shall therefore be omitted. Let Y be a space, A C Y be closed, and /: A —>• X be continuous. By Lemma 1.2.6, we may assume that X is a closed subspace of some AE, say Z. Let G: Y —> Z be a continuous extension of /. Since X is a closed subspace of Z, it follows that X is a neighborhood retract of Z. So let U be a neighborhood of X in Z for which there exists a retraction r: U —>• X. Put V = G~l[U]. Since G[A] C X, V is clearly a neighborhood of A in Y. Let h denote the restriction of r o G to V (observe that h is well-defined since G[V] C U). Then h is continuous and extends / since for every y G A we have h(y) = r(G(y}} = f ( y ) . D Corollary 1.2.8. Every space is homeomorphic to a closed subspace of an AR. Proof. This follows directly from Lemma 1.2.6 and Theorem 1.2.7.
D
Corollary 1.2.9. Let C be a convex subset of a locally convex linear space. Then C is an AR. Proof. This follows from Theorems 1.2.2 and 1.2.7.
D
26
1. BASIC TOPOLOGY
Notice that Cauty's linear space mentioned in Remark 1.2.4 shows that the local convexity assumption in this corollary cannot be dropped. Proposition 1.2.10. A neighborhood retract of an ANR is an ANR. As a consequence, an open subspace of an ANR is an ANR. Proof. Let X be an ANR, A C X closed, U an open neighborhood of A, and r: U —)• A a retraction. In addition, let B be a closed subspace of a space Y and let /: B —> A be continuous. Since X is an ANR there are a neighborhood V of B in Y and a continuous extension g: V —> X of /. Then clearly W = g~l\U] is a neighborhood of A. The function f = ro g: W —> A is continuous and extends / since for every y G B we have
Since an open subspace of a space is clearly a neighborhood retract, the second statement is indeed a consequence of the first. D Proving that a given space is an AR or an ANR is usually a difficult task. We will come back to this in Chapter 4. For the moment we shall prove a few elementary results about AR's and ANR's only. Proposition 1.2.11. A product of countably many nonempty spaces is an AR iff all factors are. Corollary 1.2.12. R n , ln, Q and s are AR's. Proof. Apply Proposition 1.2.11 and Corollary 1.2.9.
D
n
Corollary 1.2.13. For each n > 0, § is an ANR. Proof. Let U = Rn+1 \{(0, 0, . . . , 0)}. Then U is an ANR by Corollary 1.2.12 and Proposition 1.2.10. The function r: U —>• §n defined by x T(T\ — { ' \ \l lXx \l l\
is clearly a retraction. Consequently, S n is a neighborhood retract of an ANR and is therefore an A N R itself (Proposition 1.2.10). D In Chapter 2 we shall prove that no §n is an AR. Proposition 1.2.14. The product of finitely many ANR's is an ANR. Proof. The simple proof is left as an exercise to the reader.
D
A product of a countably infinite number of ANR's need not be an ANR (in contrast to Proposition 1.2.11 on AR's). The following result gives more information on this.
1.2. EXTENDING CONTINUOUS FUNCTIONS
27
Theorem 1.2.15. Let Xn be a nonempty space for every n £ N. For the product X = n^Li Xn, the following statements are equivalent: (1) X is an A N R , (2) each Xn is an ANR and there is an n E N such that Xm is an AR for every m > n.
Proof. Since for arbitrary n, the product X can be factorized as
n^x n x» n
oo
i=l
i=n+l
the implication (2) =$• (1) follows from Propositions 1.2.11 and 1.2.14. For (1) =>• (2), first observe that each Xn can be viewed as a retract of X (cf. the proof of Proposition 1.2.11), hence Xn is an ANR for every n. Now take points xn G Xn, n e N, and let C^ be an AR which contains the space Xn as a closed subset (Corollary 1.2.8). Since X is an ANR and is closed in C = fl^Li Cn (Exercise A. 1.13), there is a neighborhood U of X in C for which there exists a retraction r: C7 -> X. Put x = ( x i , X 2 , . . .)• Since C/ is a neighborhood of x, there are an n G N and neighborhoods V^ of Xi in Cj for every i < n — 1 such that x e W = Vi x • • • x Vn-i x Cn x Cn+i x • • • C [/. Observe that by Proposition 1.2.11, the product ~[\^=nCm is an AR. Now define a function
by An easy check shows that s is a retraction, from which it follows that the product rim=n^™ is an AR. Now apply Proposition 1.2.11. D The following result enables us to create many new A(N)R's from old ones in yet another way. Theorem 1.2.16. Let X = X\ 11X2, where Xi and X% are closed in X, and let X0 =Xi nX2. Then (1) I f X 0 , Xl and X2 are A(N)R's then X is an A ( N ) R . (2) IfX and X0 are A(N)R's then Xi and X2 are A(N)R's. Proof. For (1), assume that XQ, X\ and X% are ANR's. We shall prove that X is an ANR. The proof of (1) for AR's is similar.
28
1. BASIC TOPOLOGY
Assume that X is a closed subset of a space Z. Our task is to construct a neighborhood U of X in Z and a retraction r:U-*X. To this end, define
respectively (cf. the proof of Lemma A. 8.1). It is clear that Z; n X = Xi for i G (0, 1, 2}, that Zi HZ 2 = Z0 and finally that Zi UZ 2 = Z. Also observe that the Zi are closed in Z. Since XQ is closed in ZQ there are a closed neighborhood WQ of XQ in ZQ and a retraction r0 : W0 —>• XQ. Now for i E {1, 2} define Ti : WQ U Xi —>• Xj by rT
i
-
z
- z
Observe that r^ is a retraction. By Theorem 1.2.7 there exists a closed neighborhood Vi of W0 U Xi in Zi such that r; can be extended to a continuous function fi-.Vi^Xi (te{l,2». It is clear that there exists for i 6 {1, 2} a closed neighborhood Ui of Xi in Zi such that Ui C Vi and Ui n ZQ C W0. Then E/i n E/2 C W0
from which it follows that the function f : UiUU2 -^ X defined by fi(z)
(z€Ui),
r2(z)
(z G C7 2 ),
is a well-defined retraction. Since U\ U U2 is a neighborhood of X in the space Z, we are done. For (2), let XQ and X be ANR's. Again, the proof for AR's is similar. Since XQ is an A N R , there are a neighborhood UQ of XQ in X and a retraction r: UQ ->• X0. Define f : Xi U UQ -» Xi by (x 6 A"i),
An easy check shows that r is a retraction. Since X\ U f/o is a neighborhood of Xi and X is an ANR, it now follows that Xi is an A N R as well. The proof for X2 is the same. D Exercises for §1.2. 1. Prove that every ANR is locally contractible. 2. Prove that if A ( X ) is an ANR then so is X.
1.3. FUNCTION SPACES
29
3. Prove Proposition 1.2.11. 4. Let X - {0} U {l/n : n e N}. Prove that A(X) is an example of a contractible space which is not an ANR. 5. Let A C S™" 1 . Observe that Bn \ A is a convex subset of W1 and hence is an AR. Present a direct proof of this. 6. Prove that each AR is path-connected. Prove that every ANR is locally path-connected. Observe that, in particular, every ANR is locally connected. 7. Give an example of a subspace X of R, such that X is an A N R , but XU{p} is not an ANR for some p G R. 8. Let X be the sin (y x )-continuum in the plane. Prove that X is not an ANR. ^•9. Let .4i and A^ be closed subsets of the locally convex linear spaces LI and 1/2, respectively. Prove that for each homeomorphism h: A\ —> AI there exists a homeomorphism H: LI x 1/2 —> LI x Lz such that H(a,0) = (0,/i(a)) for every a G A. 10. Let L be a linear space. In addition, let X be a space, A C X be closed and /: A —> L be continuous. Finally, let It and {au • U £ It} be a Dugundji system for X and A. For each x G X \ A let 8.(x) = {U <E U : x 6 U}.
Define a function F =>• L by
Ffx) = ^ ' ' conv({/(a[/) : U e £ ( x ) } )
(x£X\A).
Prove that F is LSC. 11. Use Exercise 1.2.10 to prove that a normed linear space satisfies the conclusion of the Dugundji Theorem 1.2.2. 12. Let L be a linear space. Assume that for every space X, every LSC function F: X => L such that for every x G X the set F(x) is compact and convex, has a continuous selection. Prove that every linear subspace of L is an AR. 1.3. Function spaces
The space C*(X) from Example 1.1.5 is a special case of a more general construction. Let X and Y be spaces and let Q be an admissible metric on Y. Define Ce(X,Y) = {/ € C(X,Y) : diaine(/[X]) < oo}. It is sometimes convenient to refer to the elements of Ce(X, Y} as bounded functions. We shall endow CB(X, Y) with a useful topology.
30
1. BASIC TOPOLOGY
In §A.2 we observed that @ need not be a metric on C(X, Y). Fortunately, on Ce(X,Y) it is a metric. Lemma 1.3.1. Let X and Y be spaces and let Q be an admissible metric on Y. Then (1) for all /, g 6 Ce(X, Y) we have £(/, g) < oo, (2) the function g: Ce(X,Y) x Ce(X,Y) -> [0, oo) is a metric. Proof. For (1), take an arbitrary point z G X and observe that for all functions /, g G Ce(X, Y) the following holds: Q(f,9) < diam,(/pq) + Q ( f ( z ) , g ( z ) ) + diam, (g(X)) < oo. The proof of (2) is routine and is left as an exercise to the reader.
D
From now on we shall endow CQ(X, Y} with the topology induced by Q. Let us emphasize that the set Ce(X, Y} as well as its topology depend on the choice of the metric Q. It is a natural question to ask when the choice of Q is irrelevant. Lemma 1.3.2. Let X and Y be spaces with X compact. In addition, let Q\ and Q-2 be admissible metrics for Y. Then (1) Cei(X,Y)=Cea(X,Y)=C(X,Y), (2) the topologies on C(X,Y] induced by QI and £2 are the same. Proof. (1) is trivial. For each e > 0, y e Y and i G {1, 2} we put Bi(y,e) = {zeY : Qi(y,z) < e}. For (2), take / 6 Cei(X,Y) and e > 0, arbitrarily. Our aim is to prove that there exists 6 > 0 such that {g € CQ2(X,Y) : g2(f,g)
< 6} C {g e CB1(X,Y) : ^ ( f . g ] < e},
i.e., that the ^2-ball about / with radius 6 is contained in the ^i-ball about / with radius e. To this end, observe that since f[X] is compact, the open cover has a 02-Lebesgue number, say 6 (Lemma A. 5. 3). We claim that this 8 is as required. To see that this is indeed the case, take an arbitrary g 6 CQ2 (X, Y} such that Qi(f,g) < 5. For each x e X we have 02(f(x),g(x)') < 5, so there exists px e X such that { f ( x ) , g ( x } } C J3i(/(p z ), e/2). Consequently, for each element x G X we have Qi(f(x},g(x)) <e, from which it follows by Exercise A. 5. 4 that Qi(f,g) < £•
1.3. FUNCTION SPACES
31
So we conclude that the topology on C(X, Y) induced by £2 is finer than the topology on C(X,Y) induced by 0i. By interchanging the roles of g\ and £2 m the above argument we find that the induced topologies are indeed the same. D Let X and Y be spaces with X compact. From the above lemma we conclude that all the topologies we defined on the set C(X, Y) coincide. So for compact X and any (Y,g) we shall denote the space Ce(X,Y) simply by C(X, Y). The topology on C(X,Y) is called the topology of uniform convergence. Observe that the norm topology on C(X) for compact X defined in Example 1.1.5 coincides with the just defined topology on C(X, R). On Page 19 we defined the so-called compact-open topology on C(X}. This is again a special case of a more general construction. Indeed, let X and Y be spaces and for an arbitrary compact subset K in X and an arbitrary open subset U in Y define [K,U] =
{feC(X,Y):f[K]CU}.
Topologize C(X, Y) by taking the collection {[K, U] : K C X compact and U C Y open} as an open subbase. This topology is called the compact-open topology on the set C(X,Y). We will now show that for compact X and arbitrary Y the compact-open topology on C(X, Y) coincides with the topology of uniform convergence on C(X, Y). This allows us to prove quite easily that C(X, Y) is separable. Proposition 1.3.3. Let X and Y be spaces with X compact. The topology of uniform convergence on C(X, Y) coincides with the compact-open topology on C(X,Y). As a consequence, C(X,Y} is separable. Proof. Let g be an admissible metric on Y. In addition, let K C X be compact and U C Y open. If / G [K, U] then f[K] is a compact subset of U. By Corollary A.5.4 there exists 8 > 0 such that
B(f[K\,S)CU. This clearly implies that if g G C(X, Y} is such that £>(/, g) < 6 then g[K] is contained in U, i.e., {geC(X,Y):Q(f,g)<6}C[K,U}. From this we conclude that [K, U] is open in the topology of uniform convergence on C(X, Y) and hence that the topology of uniform convergence is finer than the compact-open topology.
32
1. BASIC TOPOLOGY
For the converse, let !B and £ be countable open bases for X and Y, respectively, which are both closed under finite unions. For B € *B and E € £ put By the above, each A(B,E) is open in the topology of uniform convergence on C(X,Y). Let A be the (countable) collection of all A(B:E)'s. We claim that the family A* of all finite intersections of elements of A is an open base for C(X, Y) endowed with the topology of uniform convergence. This proves on the one hand that compact-open topology is finer than the topology of uniform convergence and on the other hand that that C(X, Y) has a countable base. Let / G C(X,Y) and e > 0. We shall prove that there exists an element F € A* such that / G F C {g G C(X,Y) : g(f,g) < e}. By compactness of the set f[X], there are finitely many elements of £, say E\ , E-z, • • • •> En, such that (1) (2) for every i < n, diam(Ei) < e.
Let U = {f~l[Ei\ : i < n}. Since "B is a base, there clearly is a a cover V of X consisting of elements of 23 such that V < U. By compactness of X, we may assume that V is finite. For each i < n let Wi be the union of the elements of V the closures of which are contained in f~1[Ei], Since "B is closed under finite unions, W = {Wi : i < n} is a subcollection of !B, "W covers X, and W has the property that the closure of each Wi is contained in f~1[Ei\. (This is so since Wi is a finite union of sets the closures of which are contained in f-^Ei].) Now put
It is clear that / G F. In addition, F is open in the topology of uniform convergence. We claim that F C {g G C(X,Y) : £(/, 5) < e}. To this end, take g G F and x G X. There exists i < n with x G Wi. Since /, g G F, it follows that f[Wi] U g[Wi] C Ei. Consequently, both f(x) and g(x) belong to Ei from which we get by (2) that g(f(x},g(x)] < £. So we are done. D This result can be used to estimate the number of continuous functions. Corollary 1.3.4. Let X and Y be spaces with X compact. Then C(X,Y) has cardinality at most c. Proof. This follows from Proposition 1.3.3 and Exercise A. 2. 16. We now turn to completeness properties of C(X,Y) and CS(X,Y).
D
1.3. FUNCTION SPACES
33
Proposition 1.3.5. Let X and (Y, g) be spaces. Let ( f n } n be a g-Cauchy sequence in C(X,Y] such that for every x G X, lim^^oo fn(x) exists. Then the function f : X -> Y defined by f(x] — limn^oo / n (x) is continuous. In addition, if fn G Ce(X, Y) for every n then f G Cg(X, Y} and f = lim™-^ fn (inCe(X,Y)). Proof. For the first part of the proposition, see Lemma A. 3.1. For the remaining part, suppose that fn G CS(X, Y) for every n. We shall prove that diam(/[A"]) < oo. By Claim 1 in the proof of Lemma A. 3.1 there exists an M G N such that for every x G X, g ( f ( x ) , /M(^)) < 1- Take arbitrary x,z G X. Then g(f(x)J(z})
< g ( f ( x ) , f M ( x ) ) + Q(fM(x], f M ( z ) ) +
g(fM(z}J(z))
< 2 + diam(/ M [A]), so diam(/[A"]) < oo. It remains to prove that / = limTn.00 fn (in C8(X,Y)). again easily from Claim 1 in the proof of Lemma A. 3.1.
But this follows D
Corollary 1.3.6. Let X and (Y, g) be spaces. Then g is complete if and only if g is complete. Proof. Suppose that (Y, g) is complete and let (fn}n be a 0-Cauchy sequence in Ce(X,Y). Fix z G A" arbitrarily and let e > 0. There exists A7" G N such that g(fm fm) < £ for all n,ra > N. Since for all n and m,
we conclude that (fn(z)) is Cauchy in (Y,g). The completeness of (Y, g) and Proposition 1.3.5 now yield that the sequence (/ n )n converges. Now assume that g is complete. Let (yn)n be a ^-Cauchy sequence in Y. For each n let fn : X -> Y be the constant function with value yn. It is easy to see that ( f n ) n is a ^-Cauchy sequence in Ce(X,Y). So / = limn_^oo fn exists and belongs to Ce(X,Y). It is trivial to prove that / is constant and that the sequence (yn)n converges to the unique point in the range of /. D Corollary 1.3.7. Let X be compact and Y topologically complete. the space C(X,Y) is topologically complete.
Then
Groups of homeomorphisms. We will now restrict our attention to spaces of homeomorphisms and will derive some basic properties of them. Let X and Y be spaces and define §(A,F) = {/ G C(X,Y) : f is surjective}. There are spaces X and Y for which §(A", Y) is empty, for example this is the case if X has smaller cardinality than Y.
34
1. BASIC TOPOLOGY
Proposition 1.3.8. Let X and Y be spaces with X compact. Then §>(X, Y} is closed in C(X, Y). Proof. Take an arbitrary / ^ S(X, Y). There exists a point y G Y \ f[X]. So fe[X,Y\{y}]CC(X,Y)\S(X,Y). Since [X, Y \ {y}] is open in C(X,Y) (Proposition 1.3.3), this shows that
is a neighborhood of /. Hence C(X, Y) \ B(X, Y} is open in C(X, Y}.
D
Let X and Y be spaces with X compact, and let e > 0. A continuous function /: X —> Y is called an e-map if for every y G Y, s.
Let e > 0 and put
CE(X, Y) = {/ G C(X, Y) : f is an e-map} and
Se(X,Y)=Ce(X,Y)nS(X,Y), respectively. In addition, put
Lemma 1.3.9. Let X and Y be spaces with X compact. Then C£(X,Y) is open in C(X, Y) for every e > 0. Consequently, S£(X, Y} is a closed subset ofC(X,Y). Proof. Take / G C£(X,Y). Since X is compact, /: X ->• f[X] is a closed map (Exercise A. 5. 5). We claim that for every y G f[X] there exists an open neighborhood Uy (in /[-X"]) such that diam(f-l[Uy})
<e.
This will be achieved in two steps. Take an arbitrary y G f[X]. Then diam(f-l(y))
<e
since / is an £-map. There clearly exists an open neighborhood U of f ~ l ( y ) such that diam(C7) < e. Now by using that / is a closed map, Exercise A. 1.1 5 gives us the required neighborhood Uy. Let 6 > 0 be a Lebesgue number for the open covering
{Uy-.yz f[X}} off[X] (Lemma A. 5. 3). Let g G C(X, Y) be such that g(g, /) < %. We claim that g G C£(X, Y). This clearly suffices. To this end, take an arbitrary y G Y. Since we have 0(f,g) < 5/2 it follows easily that diam (fg~1(y)) < 5. There
1.3. FUNCTION SPACES
35 l
consequently exists a point z 6 f[X] such that fg~~l(y) C Uz. This implies that Since g~1(y) C f~1fg~1(y), we conclude that diam (g~1(y)} < £•, i.e., g is an e-map. That $ £ ( X , Y ) is closed now follows by Proposition 1.3.8. D Let X and Y be spaces. We introduce a few more interesting subsets of C(X, Y). Let 3(X, Y) denote the subset of C(X, Y) consisting of all imbeddings of X into Y , and let 9t(X, Y) denote the set of all homeomorphisms from X onto Y. If X = Y then for H ( X , X ) we shall simply write As usual, ^K(X) is called the autohomeomorphism group of X. Lemma 1.3.10. Let X and Y be spaces with X compact. Then
As a consequence, 3(X,Y) is a G^-subset o f C ( X , Y ) , and 3{,(X,Y) is a G$ subset of both C(X, Y) and S(X, Y). Proof. That 3(X,Y) C f}™=l Cif (X,Y] is a triviality. Pick an arbitrary
Then / is a 1/n-map for every n, hence / is one-to-one. So the compactness of X implies that / is an imbedding (Exercise A. 5. 9). The remaining statements are obvious. D Corollary 1.3.11. Let X and Y be spaces with X compact and Y topologically complete. Then both 3(X, Y} and 'H(X, Y) are topologically complete. Proof. Since C(X,Y) is completely metrizable by Corollary 1.3.6, this follows immediately from Lemma 1.3.10 and Theorem A. 6. 3. D So 'K(X) is complete if X is compact. It will be convenient to explicitly describe a complete metric for "H(X) that generates its topology. Proposition 1.3.12. Let X be a compact space. For /, g 6 'K(X)
define
Then a is a complete metric on "H(X) that generates its topology. Proof. That a is a metric is left as an exercise to the reader. We shall first prove that Q and a generate the same topology on J{,(X}.. Since for
36
1. BASIC TOPOLOGY
all /, g E 3t(X] we have £>(/,#) < cr(/, ), the only thing to verify is that for every e > 0 and every / E Oi(X) there exists S > 0 such that if g E K(X} and g(f,g) < 6 then a(f,g) < E. Choose arbitrary e > 0 and / E ^K(X). By compactness, /-1 is uniformly continuous (Exercise A.5.18) and consequently there exists 7 > 0 such that for all x,y E X with g ( x , y ) < 7 we have g ( f ~ l ( x ) , f ~ l ( y } } < £/2- Let Take g 6 3{(X) such that Q(f,g) < S. Pick an arbitrary x e X and put z = g~l(x). Since £>(/, g) < 6, it follows that Q(f(z),g(z))
= Q(fg-l(x},x}
< 6 < 7.
As a consequence, We conclude that 6(g~1,f~i) < Eli by Exercise A. 5.4. Therefore,
Now let (fn)n be a cr-Cauchy sequence in "H(X}. Then (fn)n is a QCauchy sequence in C(X, X) and therefore the limit / = limn_>oo fn exists and belongs to C(X, X] (Corollary 1.3.6). Similarly, by the definition of cr, the limit g = limn^.oo f~l exists and belongs to C(X,X). It is easily seen that fog = lx=gof from which it follows that / 6 'H(X). D Lemma 1.3.10 and Proposition 1.3.12 both imply by Theorem A. 6. 6 the following Corollary 1.3.13. If X is compact then 3i(X) is a Baire space. Exercises for §1.3. 1. Let N denote the discrete space of natural numbers. Prove that C(N, K; |-|) is not separable. 2. Let X, Y and Z be compact spaces. For / € C(Z, X) and g,h£ C(X, Y) prove that Q(g° f,ho f ) < g(g,h). In addition, show that if / is surjective then
3. Let X be compact, let /, g G C(X, X) such that g is a homeomorphism. Prove that
1.4. THE BORSUK HOMOTOPY EXTENSION THEOREM
37
4. Let X be a compact space. Prove that the function f : M(X) x :K(X)-» K(X) defined by ^(/,S) = /o 5 " 1 is continuous (i.e., 'K(X) is a topological group). 5. Prove that the function /: I —>• I defined by (0 < x < V 4 ),
x
belongs to the closure of IK(I) in C(I,I). (Hence IK (A") is even for compact A not necessarily a closed subspace of C(X, A).) 6. Prove that IK (I) has exactly two components. 7. Give an example of two nontrivial continua A and Y such that S(A", Y) is empty. 1.4. The Borsuk homotopy extension theorem The aim of the present section is to prove that a continuous function / from a closed subspace A of a space X into an A N R Z is extendable over X if and only if / is homotopic to an extendable function g: A —> Z. We shall need the following simple lemma: Lemma 1.4.1. Let A be a closed subset of a space X. Then for every neighborhood V of B — (X x {0}) U (A x I) in X x I there is a continuous function a: X x I —>• V which is the identity on B. I U
V 0
x
A
Figure 1. Proof. Let TT: X x I ->• X be the projection. Then by compactness of I, TT is closed by Exercise A.5.8. We may assume without loss of generality that V is open. So F = (X x I) \ U is a closed set which misses A x I. Hence
U = X\ 7t[F}
38
1. BASIC TOPOLOGY
is an open neighborhood of A in X such that U = U x I C V . Now let ft: X ^l be a Urysohn function (Lemma A. 4.1) such that 0\A=1,
0\X\U = Q.
Define a: X x I -»- X x l b y a(x,t) = (x,{3(x) • t). Then a. is clearly continuous. It is easily seen that it restricts to the identity on B. We shall prove that a[X x I] C V, thereby showing that a is as required. To this end, take (x, t) e X x I arbitrarily. If x £ U then j3(x) = 0 and consequently, a(x,t) = (x,0) 6 B C V. On the other hand, if x £ U then a(z, f) = (x, /3(:r) • t) € U x I C V . D We now come to the main result in this section. The Borsuk Homotopy Extension Theorem 1.4.2. Let A be a closed subspace of a space X, let Z be an A N R and let H : A x I —> Z be a homotopy such that HQ is extendable to a continuous function f : X —> Z. Then there is a homotopy F : X x I -> Z such that (1) ^o = /, (2) for every tel,Ft \A = Ht.
Proof. Using the notation as in Lemma 1.4.1, define a function £: B —>• Z
by
^,t) = lf(x) \H(x,t)
(»^,t = 0), (x e A,tei).
Since X x {0} and Axl are closed in B, it follows easily that £ is continuous. As Z is an ANR and as B is closed in X x I, we can find a neighborhood V of B in .X x I such that £ can be extended to a continuous function £' : F —> Z. Let a : X x I -> V be as in Lemma 1.4.1. Define F : X x I ->• Z by the formula
It is easily seen that F is as required.
D
Theorem 1.4.2 has some interesting corollaries. Corollary 1.4.3. Let A be a closed subset of a space X . Let Z be an A N R and let f : A —>• Z be continuous. The following statements are equivalent: (1) / can be extended over X, (2) / is homotopic to an extendable function g: A —>• Z.
1.4. THE BORSUK HOMOTOPY EXTENSION THEOREM
39
Proof. That (1) => (2) is clear since / ~ / by Lemma A. 12.1(1). For (2) =>• (1), let H : A x I -> Z be a homotopy with H0 = g and HI = f . Since g is extendable, / is extendable as well (Theorem 1.4.2). D Our next goal is to investigate contractibility within the class of ANR's. Theorem 1.4.4. Let X be an AR. Then X is contractible. Proof. By Corollary 1.1.8, X can be thought of as a closed subset of a normed linear space L. Fix an arbitrary point p £ L. Since the function F : L x I —>• L defined by the formula ( x , t ) \->t-p+(l-t)-x is a contraction, it follows that X is contractible, being a retract of L (Theorem A. 12.4). D These results yield the following interesting characterization of AR's. Corollary 1.4.5. Let X be a space. The following statements are equivalent:
(1) X is an AR, (2) X is a contractible A N R . Proof. The implication (1) =$• (2) is trivial by Theorem 1.4.4.
For (2) =» (1), let H: X xl->X be a homotopy such that HQ = 1 and HI is constant, say with constant value c. In addition, let Y be a space, A C Y be closed and /: A —)• X be continuous. Define F: A x I —>• X by Then F is clearly a homotopy with F0 — f and FI a constant function. Trivially, FI can be extended to a continuous function from Y to X. As X is an A N R , by Corollary 1.4.3 it follows that FQ = f can be extended over Y. We conclude that X is an AR. D For a nontrivial generalization of Corollary 1.4.5, see Theorem 4.2.20. Cones. Let X be a space and consider its cone A(JC). We observed in Exercise 1.2.2 that if A(X) is an ANR, then so is X. It is an interesting question whether the converse holds. Theorem 1.4.6. For a space X the following statements are equivalent:
(1) X is an ANR, (2) A(X) is an A N R ,
40
1. BASIC TOPOLOGY
(3) A (AT) is an AR.
Proof. Observe that (3) =>• (2) is trivial, and (2) =>• (1) follows from the above observation. So we only need to prove the implication (1) =>• (3). To this end, first observe that we may think of X as a closed and bounded subspace of a normed linear space L (Corollary 1.1.8). The cone A (A") is by Exercise 1.1.29 homeomorphic to the subspace
K = {(t,tx)
:tel,xeX}
of I x L (for notational simplicity, we switched the coordinates of L and II , and applied the homeomorphism t \-> 1 — £ of I). Since X is an A N R , there are an open neighborhood U of X in L and a retraction r : U —> X. Let V be an open neighborhood of L \ U such that V n X = 0 (Corollary A. 4. 3). By Lemma A. 4.1 there is a continuous function a: L —>• I such that a \ X = 1 and a \ V = 0. Define a function /3 : I x L —>• K by
t),ta(Vt)r(*/t))
(otherwise).
We claim that (3 is a retraction from I x L onto /f . This suffices for the proof, since clearly II x L is an AR (Corollary 1.2.9 and Proposition 1.2.11) and a retract of an AR is an AR. Notice that /2(0,Q) = (0,0). In addition, if x G X and t G II \ {0} are arbitrary then tx/t = x E U and so fi(t,tx) = (ta(x),ta(x)r(x))
= (t,tx).
This proves that j3 \ K is the identity. So all there remains to prove is that (3 is continuous, and this is left to the reader in Exercise 1.4.2. D Notice that the implication (2) =$> (3) also follows from Corollary 1.4.5. Simply observe that A(AT) is contractible (Exercise A. 12. 5).
Exercises for
§1.4.
1. Let A be a closed subspace of a space X and let T be the subspace (X x {0}) U (A x I)
of X x I. Prove that for every space Y, if a continuous function /: T —> Y can be extended over (X x {0}) U t/, where U is a neighborhood of A x I in X x I, then / can be extended over X x I. ^•2. Prove that the function /3 in the proof of Theorem 1.4.6 is continuous.
1.5. TOPOLOGICAL CHARACTERIZATION OF SOME FAMILIAR SPACES
41
1.5. Topological characterization of some familiar spaces The aim of this section is to present topological characterizations of two familiar spaces, namely, the Cantor middle-third set C and the closed unit interval I. (See §1.9 for topological characterizations of Q and P.) Topological characterization of the Cantor set. A space X is called zero- dimensional if it is nonempty and has a base consisting of clopen sets, i.e., if for every point x E X and for every neighborhood U of x there exists a clopen subset C C X such that x G C C U. It is clear that a nonempty subspace of a zero-dimensional space is again zero-dimensional and that products of zero-dimensional spaces are zero-dimensional. Observe that by Exercise A. 2. 12 it follows that a space is zero-dimensional if and only if it has a countable base consisting of clopen sets. It is clear that no nontrivial connected space is zero-dimensional. As a consequence, if a space contains a nontrivial connected subspace then it is not zero-dimensional. Proposition 1.5.1. Let X be zero- dimensional. Then for every open cover U of X there exists a refinement V of U consisting of pairwise disjoint clopen sets. Proof. Let U be an open cover of X. Since X is zero-dimensional, it can be refined by a clopen cover £. We may assume without loss of generality that £ is countable, say £ = {En : n 6 N} (Corollary A. 2. 3). Now put Vi = EI and put Vn = En\[j Ei i
for every n > 2. Then V = {Vn : n 6 N} is as required.
D
Corollary 1.5.2. Let X be a space. Then X is zero-dimensional if and only if for all disjoint closed sets A and B there exists a clopen subset U C X such that A C U and B n U = 0. Proof. First assume that X is zero-dimensional. For every x £ X let Ux be a neighborhood of x such that Ux D A = 0 or Ux fl B = 0 (here we use that A and B are disjoint closed sets). By Proposition 1.5.1 there is a refinement V of {Ux : x € X} consisting of pairwise disjoint clopen sets. Now put = \J{ V G V:
It is easy to see that U is as required. If for all disjoint closed subsets A and B of X there exists a clopen set C with A C C C X \ B, then the clopen subsets of X clearly form a base. So then X is zero-dimensional. D
42
1. BASIC TOPOLOGY
We shall now characterize the zero-dimensional subspaces of the real line R. Proposition 1.5.3. A nonempty subspace X of R is zero-dimensional if and only if it does not contain any (nondegenerate) interval. Proof. Assume that X C R is zero-dimensional. If X contains a nontrivial interval E then X contains a nontrivial connected subspace and is therefore not zero-dimensional. Now assume that X C R is such that it contains no nontrivial intervals. Then the set D = R \ X is dense in R. The collection {(di,d 2 ) r\X : di < d2 and c?i,d 2 € D} is easily seen to be a base for X consisting of clopen subsets of X. Consequently, X is zero-dimensional. D This proposition gives us a rich supply of zero-dimensional spaces. For example, the rational numbers Q, the irrational numbers P, the product QxP, etc., are all zero-dimensional. The following example is of particular interest: Example 1.5.4. The Cantor middle-third set C. From I = [0,1] remove the interval (Y 3 , 2/3), i.e., the 'middle-third' interval. From the remaining two intervals, again remove their 'middle-thirds', and continue in this way. At stage i of the construction we have a disjoint family 3^ of 2l closed subintervals of I, each of length 3~ l . The union of 3~; is denoted by Hi. Observe that Hi is a closed subset of I and hence is compact. The intersection of the H^s is called the Cantor middle-third set, C. Clearly, C is closed in I and hence compact. Also, C is zero-dimensional by Proposition 1.5.3. For if (a, b) is a nondegenerate subinterval of R then there exists i £ N such that 3~z < b — a which implies that (a, b) cannot be contained in C since C can be covered by a disjoint family of closed intervals each of length 3~ z .
Figure 2.
1.5. TOPOLOGICAL CHARACTERIZATION OF SOME FAMILIAR SPACES
43
Observe that if F 6 ^ then F n C / 0. For at stage i + 1 of the construction there exists an interval AI £ 3^+1 such that A\ C F. In addition, at stage i + 2 of the construction, there exists an interval AI e 3^+2 such that AI C AI. Proceeding in this way inductively, it is easy to construct a decreasing sequence (An}n of elements of 3*i+n such that An C F for every n. By compactness of I, 0 / fl^Li An C C, as desired. It is easy to see that C is the subspace of I consisting of all points that have a triadic expansion in which the digit 1 does not occur, i.e., the set oo
x=
\
51 ^T : Xi
G
{°' 2} for every *'
We claim that the set C has no isolated points. If x belongs to C and a < x < b then choose i so large that 3~z < min{;r — a, b — x}. Take F G 3^ with x & F. Then F C (a, 6). There are precisely two disjoint elements of 3~i+i that are contained in F. By the above, both these elements contain at least one point of C. We conclude that (a, b) D C contains at least two points. The Cantor set C is a very interesting space. There exists a simple topological characterization of it which we shall now present. Theorem 1.5.5. C is (topologically) the unique nonempty zero-dimensional compact space without isolated points. Proof. Since in Example 1.5.4 we showed that C is a zero-dimensional compact space without isolated points, it suffices to prove that if X and Y are such spaces then they are homeomorphic. To this end, let X and Y be compact zero-dimensional spaces without isolated points. By induction on n, we shall construct elements m(n) G N and finite disjoint clopen covers Un = {Un^ : i < m(n)},
V = {Vn^ : i < m(n}}
of X and Y, respectively, consisting of nonempty sets, such that and for n > 2,
(1) mesh(U n ) < Vn, meshCV n ) < Yn, (2) U n < U n _ i , V T O < V n _i, (3) for i < m(n — 1) and j < m(n): Unj C Un-\,i iff Ki,j C V^-i^. Suppose that we constructed the covers UTO and Vm for all m < n, n > 1. Pick an arbitrary i < m(ri). By Proposition 1.5.1 there exist covers £ and 3r of C/n)i and Vn>i, respectively, consisting of pairwise disjoint clopen sets, such that (4) £ <
(5) y-
44
1. BASIC TOPOLOGY
By compactness of Un,i and Vn,i, the covers £ and 3~ are finite. We assume that each element of £ as well as 3" is nonempty. We claim that without loss of generality we may assume that £ and 3" have the same cardinality. To see this, pick an arbitrary E £ £. Since X has no isolated points, there exist distinct x,y E E. In addition, since X is zero-dimensional, its clopen sets form a base and so there is a clopen subset C C E such that x E C and y $. C. So E can be split into two nonempty clopen sets, namely, C and E\C. So we can replace £ by a cover having |£| + 1 elements. By repeating the same procedure if necessary, we can therefore ensure that £ and 3~ have the same cardinality. We conclude that for i < m(n) there are covers £; and 3^ of Un^ and Vn,i, respectively, such that (6) |£t| = |^|,
(7) 8,i and 3^ are pairwise disjoint and consist of nonempty clopen sets, each of diameter less than l/n. Put Un+i — LJl^i ^"i and Vn_|_i = U™LI 3V By construction, we can enumerate lin+i and Vn+i in such a way that (3) is satisfied for n + l. This completes the inductive construction. Observe that IJ^Li ^ *s a base for X and, similarly, that U^Li "^n ^s a base for Y (Exercise A.2.15). We will now use these covers for the construction of the required homeomorphism. If x E X then for each n E N there is a unique i(n,x) < m(n) such that x E C/ n ,i(n,x)- By (3) it follows that the collection {Vn^n^}n is decreasing and since Y is compact, we get H^Li Ki,i(n,z) 7^ 0- So by (1), the intersection H^Li ^n,i(n,z) consists of precisely one point. By interchanging the roles of X and F, these remarks imply that the function /: X —> Y defined by f ( x ) = y <£=^ for every n e N and i < m(n)
(x 6 C7n)i iff y G V^^),
is a bijection. Also / is continuous since for n e N and i < m(n), f~l[Vn>i] is equal to C/^^. Similarly, f~l is continuous. So we conclude that / is a homeomorphism. D A space homeomorphic to C is called a Cantor set from now on. The characterization of C allows us to recognize many spaces as Cantor sets. Corollary 1.5.6. Let X be any compact nonempty zero-dimensional space. Then X x C is a Cantor set. Proof. Since both X and C are zero-dimensional it easily follows that X x C is compact and zero-dimensional. Since C has no isolated points it follows that .Y x C has no isolated points. An appeal to Theorem 1.5.5 now yields the desired result. D
1.5. TOPOLOGICAL CHARACTERIZATION OF SOME FAMILIAR SPACES
45
Since {0,1}°° is compact and zero-dimensional, and has clearly no isolated points, we conclude that, in particular, C « {0,1}°°. This consequence can also be verified directly: the function /: {0,1}°° —>• C defined by 00
£1 \
\~^
C)
X
i
/(*) = L 6 t=l -^ is a homeomorphism (see Page 43). Corollary 1.5.7. Every zero-dimensional space can be imbedded in C. Proof. Let X be zero-dimensional and let U be a countable base for X consisting entirely of clopen sets. Let {t/n : n 6 N} enumerate U and for every n define fn: X -> (0,1} by fn(x) = 0 <^ x e Un.
Now define an imbedding e: X —> {0,1}°° by putting e(x)n = fn(x]
(n G N)
(this is in fact the same imbedding as in the proof of Corollary A.4.4). Since by the above C « {0,1}°°, this completes the proof. D So every zero-dimensional space can be imbedded in R. This fact admits a nontrivial generalization that will be proved in §3.3. By Cantor's standard diagonal argument, it follows that {0,1}°° is uncountable. So C is uncountable as well. (To be precise, its cardinality is c.) Since by Corollary 1.5.6, C is homeomorphic to C x C, this remark implies that there is a family A consisting of uncountably many pairwise disjoint Cantor sets in R. (Let < ^ : C x C —> C b e a homeomorphism and put A — {^[{c} x C] : c € C}.) Since there are only countably many rational numbers in R, one of the elements of A misses Q, i.e., there exists a Cantor set K in R consisting entirely of irrational numbers. Let P denote the set of irrational numbers in R. Since as was observed in Corollary 1.5.7, every zero-dimensional space can be imbedded in C, we consequently get: Corollary 1.5.8. Every zero-dimensional space can be imbedded in P. We now aim at proving that for every compact space X there exists a continuous surjection from C onto X. Lemma 1.5.9. The function f : C x C —> J denned by /(x,y) = x — y is a continuous surjection. Proof. Since C C E, / is well-defined and clearly continuous. Consider the set FI — [0, 1/3] U [2/3,1], i.e., the first approximation to C. It is a triviality to verify that FI — FI, i.e., the set {x — y : x, y G FI}, is equal to J. By induction
46
1. BASIC TOPOLOGY
one can prove that all approximations Fn, n G N, to C that were used to define C, have the same property (the induction step can be performed by noting that Fn — (l/3-Fn-i)\j(l/3- Fn-i + 2/3), a union of two shrunken copies of Fn-i.}. Since the Fn decrease and C = H^Li Fn, an easy compactness-type argument shows that C-C-JJ, as required.
D
We now come to the announced Theorem 1.5.10. Every compact space is a continuous image of C. Proof. By Lemma 1.5.9 and Exercise A.1.13 it follows that there exists a continuous surjection /: C°° —>• Q. Now let X be any compact space. According to Corollary A.4.4, we may assume that X C Q. Put Y — f ~ l [ X } . The restriction g = f\Y:Y->X is a continuous surjection. Observe that Y is zero-dimensional. By Corollary 1.5.6, C x Y and C are homeomorphic. The required continuous surjection a: C —>• X can therefore be obtained as the composition
a = (3 o TT o g, where TT: C x Y —> Y is the projection and j3: C —>• C x Y is an arbitrary homeomorphism. D Remark 1.5.11. Theorem 1.5.10 can be used to construct so-called 'space filling curves', i.e., continuous surjections from I onto E2. For let g: C —> I 2 be a continuous surjection (Theorem 1.5.10). Then, by Corollary 1.2.12, g can be extended to a continuous surjection g: I —»• E 2 . The Cantor set can be imbedded in 'many' spaces. Since it is uncountable, it cannot be a subspace of any countable space. There are also examples of uncountable spaces of which it is not a subspace, but it is not so easy to construct such a space. That such spaces exist follows however from Corollary 1.5.14 below. Before we present the next theorem, we first fix some notation. By S we mean the set consisting of all finite sequences consisting of O's and 1's. (This includes the empty sequence.) If s G £ then \s\ denotes its length, i.e., if s = 0 then \s = 0 and if s = i\ii... in, where ij G {0,1} for each j < n, then \s\ — n. If s — i\i^ . • -in G S and k G (0,1} then s^k G S denotes the sequence i\i<± • • • ink. If s G S U {0,1}°° then t G S is an initial segment of s if there is an n G N such that t is equal to the sequence consisting of the first n elements of s. Finally, if s G Eu{0,l}°° then s extends t G S if t is an initial segment of s.
1.5. TOPOLOGICAL CHARACTERIZATION OF SOME FAMILIAR SPACES
47
Theorem 1.5.12. Let X be topologically complete and let f : X —>• Y be a continuous surjection. If Y is uncountable then X contains a Cantor set K such that f [ K is one-to-one (and hence an imbedding). Proof. Let Q be an admissible complete metric on X which is bounded by y2 (Exercise A. 1.6). Since Y is uncountable, X contains an uncountable subset on which / is one-to-one. This set contains an uncountable dense-in-itself subspace, say E (Exercise A.2.13). For s G £ we shall define a nonempty open subset Us of X having the following properties:
(1) t / s n £ ^ 0 ,
_
(2) if t 6 E and t extends s then U± C f/ s ,
(3) /[^ 0 ]n/[t7 s ^] = 0, (4) diamt/ s < 2 M . We put £70 = X. Now suppose that we constructed Us for all s 6 S such that |s| < n. Fix an arbitrary s G E such that |s| = n. By (1) and the fact that E is dense-in-itself, we may pick two distinct points x,y e Usr\E. There exists e > 0 such that D(x,e) U D(y,e) C Us. Since f(x] ^ f(y) and / is continuous, we may additionally assume that f[D(x,e)] r\f[D(y,e}] — 0 and that e < 2' s l. Define Us~0 — B(x,e) and Us~1 — B(y,e), respectively. It is easily seen that our choices are as required. Now for s G {0,1}°° put As = \\{Ut '• s extends t}. By (1), (2) and (4) and the fact that Q is complete it follows easily that As consists of a single point of X, say xs. Define a function £: {0,1}°° —> X by
f(s) =xs. We claim that £ is an imbedding. That £ is one-to-one is clear since (3) implies that / o £ is one-to-one. By Exercise A.5.9 it therefore suffices to prove that £ is continuous. Since by Exercise A.2.15 the collection [Us : s 6 S} is a local base at every point of the range of £, it suffices to prove that £-1[t/t] is open for every t G S. But this is clear since ^[Ut] = {s e {0,1}°° : s extends t} is a basic clopen subset of (0,1}°°.
D
Since the cardinality of the Cantor set is c, we immediately obtain from this that 'for topologically complete spaces the Continuum Hypothesis holds'. By this we mean that a topologically complete space cannot be of cardinality strictly between uj and c. (Observe that by Exercise A.2.16 it also cannot be of cardinality larger than c.)
48
1. BASIC TOPOLOGY
Corollary 1.5.13. Let X be an uncountable analytic space. Then X contains a Cantor set. As a consequence, if X is analytic then either X\ < uo or \X\ = c. Proof. Let Y be topologically complete, and let /: Y -» X be a continuous surjection. If \X\ < uj then there is nothing to prove. So assume that X is uncountable. Then Y is uncountable, and hence by Theorem 1.5.12 there exists a Cantor set K C Y such that / \K is one-to-one, and hence a topological imbedding (Exercise A.5.9). From this we get \X\ > c. Finally, \X\ < c by Exercise A.2.16. D This result enables us to prove the following interesting result. Corollary 1.5.14. Let X be a space. Then we can decompose X into two sets neither of which contains an uncountable topologically complete subspace (in particular, neither contains an uncountable compact subspace). Proof. If \X\ < c then by Corollary 1.5.13 there is nothing to prove. So assume that \X\ = c. First observe that the number of closed subsets of X is at most c (Exercise A.2.16). So there are also at most c closed subsets of X of cardinality c and we can list them as {Aa : a < A} (no repetitions permitted) for some cardinal A < c. By transfinite induction on a < A we will pick distinct points xa,ya E Aa such that {x<*,ya} n {xp,yp : /3 < a} = 0. This is easily done since each \Aa\ = c and the number of points that has to be avoided at stage a is smaller than c. Now put A = {xa : a < A} and B = X \ A, respectively. We claim that A and B are as required. To this end, let Y C A be a topologically complete subspace. If Y were uncountable, then it would contain a Cantor set by Theorem 1.5.12, and hence it would contain some Aa. But this would contradict ya G Aa \ A. The proof for B is similar. D We now present an interesting application of Corollary 1.3.11, which is a partial generalization of Theorem 1.5.12. Corollary 1.5.15. Let X be a topologically complete space, let K be compact, and assume that X contains an uncountable family ofpairwise disjoint homeomorphs of K. Then X contains a copy of the product C x K. Remark 1.5.16. Observe that the uncountably many pairwise disjoint homeomorphs of K may be irregularly imbedded. This result shows that they may be replaced by a Cantor set of 'regularly' imbedded copies of K.
1.5. TOPOLOGICAL CHARACTERIZATION OF SOME FAMILIAR SPACES
49
Proof. By Corollary 1.3.11 it follows that 3(K, X] is topologically complete, so we can pick an arbitrary complete admissible metric p for it and use the same technique as in the proof of Theorem 1.5.12 to create a Cantor set subspace of it. By assumption, 3(K, X] contains an uncountable subspace A consisting of functions having pairwise disjoint images. We perform our construction with that subspace. Since A is uncountable it contains an uncountable dense-in-itself subspace, say £ (Exercise A. 2. 13). We proceed as in the proof of Theorem 1.5.12, with one slight adaptation only. Recall that in each step of the construction we picked in some nonempty open subset U C 3(K, X) meeting £ two distinct elements /o,/i 6 £, together with small enough neighborhoods VQ,VI C U of /o and /i , respectively. Here we make the neighborhoods even smaller (see below). An inspection of the proof of Theorem 1.5.12 shows that this does not create problems. We do not only want to get a Cantor set C C 3(K, X) at the end of the process, but we additionally want the collection {h[K] : h 6 6}
(*)
to be pairwise disjoint. But this can easily be achieved. For recall that /o[/r]n/i[tf]=0. So by compactness of K there exists e > 0 such that
5e( (Corollary A. 5.4). Put
, Y) : g(fi, h) <e}
(t = 0, 1).
If hi G Jj for i = 0, 1 then hQ[K] n hi[K] = 0. So now we replace V; by the set Vi n 2F; for i = 0, 1. These neighborhoods are clearly as required. So at the end of our process we indeed get a Cantor set C C 3(K, X } having the property that the collection in (*) is pairwise disjoint. Now define the function £ : G x K —> X by
Then £ is clearly continuous. It is moreover one-to-one by the fact that the collection in (*) is pairwise disjoint. So it is an imbedding by an application of Exercise A. 5. 9. D Remark 1.5.17. An uncountable analytic space contains a Cantor set by Corollary 1.5.13. So if the compact space K contains one point only, then for K Corollary 1.5.15 not only holds for topologically complete spaces but also for the broader class of analytic spaces. The question therefore naturally arises whether Corollary 1.5.15 is also true for analytic spaces. This question was considered in BECKER, VAN ENGELEN and VAN MILL [45]; it was shown there that it is undecidable.
50
1. BASIC TOPOLOGY
We say that a space X is totally disconnected if for all distinct points x, y in X there exists a clopen set C in X such that x E C but y (£ C. It is clear that every zero- dimensional space is totally disconnected. Even more is true. Every space X that admits a one-to-one continuous function into a zero-dimensional space is totally disconnected. The question naturally arises whether every totally disconnected space is zero-dimensional. If this were true then checking whether a given space is zero-dimensional would be simpler. However, the answer to this question is in the negative, as the following example shows. Example 1.5.18. Erdos' space. Put E — {x G t2 : Xi G Q for every i}. We claim that E is totally disconnected but not zero-dimensional. On Page 9 it was observed that the topology on I2 is finer than the topology that i2 inherits from R°° . We conclude that E admits a one-to-one continuous function into the product Q°° , which is zero-dimensional (see Exercise 1.5.2). This implies that E is totally disconnected. We shall now prove that E is not zero-dimensional by showing that if U is open in i2 and 0 6 U C {x e i2 : \\x\\ < 1} then U D E is not clopen in E. To this end, let U be such a neighborhood of 0. We shall inductively define a sequence g^ e Q, i & N, such that for every i,
Let qi = 0 and assume that the QJ are defined for j < i — I for some i > 2. For 0 < ra < i put x(i,m) = ( g i , . . . , f t _ i , m / i? 0,0, . . . ) . Observe that x(i,0) — Xi-\ e U and that x ( i , i ] — (
g(xi,(? \U) < Q(x(i,m0),x(i,m0 + 1))
This completes the inductive construction. By (*) we have i Y^ n2 < 1
N
1.5. TOPOLOGICAL CHARACTERIZATION OF SOME FAMILIAR SPACES
51
and hence
We conclude that the point q = (q\,q-2, • • • ) belongs to £ 2 , and hence to E. Since clearly, lim Xi — q (in £ 2 ) i—>oo
(Lemma 1.1.12), it follows by (*) that on the one hand q belongs to the closure of U fl E in E, while on the other hand Q(q,P\U)=]im
i—too
Q(Xi,P\U)=Q,
so q $ U. We conclude that U fl E is not closed in £", hence not clopen. Topological characterization of the closed unit interval. We shall now present a topological characterization of the closed unit interval I . The characterization below in Theorem 1.5.19 is not very appealing and its proof is rather involved. But it turns out to be very useful as will be demonstrated in the proof of the Mazurkiewicz Theorem 1.5.22 about the existence of topological copies of I in certain spaces. Theorem 1.5.19. Let Y be a continuum containing two distinct points x and y. Assume that for every z G Y\{x,y} there exist closed sets Lz and Uz ofY such that (1) LZ(JUZ=Y, (2) xeLz,y£ Uz, (3) Lz n Uz = {z}. Then Y is homeomorphic to I. Remark 1.5.20. In Exercise A. 10. 6 we proved that every continuum has at least two noncut points. From the conditions in Theorem 1.5.19 it obviously follows that Y \ {x,y} consists entirely of cut points. So the points x and y are the only noncut points of Y. Proof. We first establish the following Claim 1. For every z 6 Y \ {x,y}, Lz and Uz are continua. Proof. Suppose that for certain z £ Y \ {x,y}, Lz is not a continuum. Then Lz can be written as the union of two disjoint nonempty relatively open sets, say E and F. Without loss of generality, z £ E. Then F is contained in A = Y \ Uz and since A is open in F, this easily implies that F is open in Y. However, since F is closed in Lz and Lz is closed in Y, it also follows that F is closed in Y. Consequently, F is a nonempty clopen subset of F,
52
1. BASIC TOPOLOGY
which contradicts the fact that Y is connected. (See also Exercise A.10.5.) The proof for Uz is similar. 0 Define an order ^ on Y by putting:
Claim 2. Let d, e e Y \ {x,y}. The following statements are equivalent: (4) d * e,
(5) d e L e , (6) Ld C L e , (7) e E C/d,
(8) Ue C t/d-
Proof. The implication (4) =>» (5) is trivial. For (5) =J> (6) and (5) => (8), assume that d & Le. If d = e then there is nothing to prove, so assume that d ^ e. This implies d <£ Ue, for otherwise d G Le n Ue - {e}.
Observe that Y \ {d} is the disjoint union of the open sets Ld\{d},
Ud\{d}.
Since y e f/e n (t/d \ {d}), it therefore follows by connectivity of Ue (Claim 1) that Ue C (t/d \ {c?}) (hence (8) follows) which also implies that
Ld C X \ Ue C Le (hence (6) follows). Observe that (8) => (7) is trivial. The implication (7) =^ (6) can be proved in precisely the same way as (5) => (8). Since (6) =$• (4) follows by the definition of =^, we are done. 0 Claim 3. =<; is a linear order on Y. Proof. This follows easily from Claims 1 and 2. Let D be the set of all rational numbers between 0 and 1, i.e., D = Qn(o,l). Claim 4. There is a subset E C D and a one-to-one function f:E-+Y\{x,y} such that (9) for all d,e£E, f(d) =$ /(e) iff d < e, (10) E is dense in I, (11) f[E] is dense in Y.
<0
1.5. TOPOLOGICAL CHARACTERIZATION OF SOME FAMILIAR SPACES
53
Proof. Let 23 = {Bn : n £ N, n even} be a base for F consisting of nonempty open sets. In addition, let 7 — {Fn : n £ N, n odd} be a base for D, also consisting of nonempty open sets. Observe that by connectivity, every Bn is infinite. In addition, since Q has no isolated points, the same holds for every Fn. By induction on n we shall pick a point en £ D and a point f ( e n ) £ Y such that the following conditions are satisfied: (12) if n is odd then en £ Fn, (13) if n is even then /(e n ) £ f? n , (14) / r { e i , . . . , e n } is one-to-one and order preserving (in the sense of (9)). For n = 1 pick an arbitrary point e\ in F\ and an arbitrary point f ( e \ ] in y \ {x,?/}. Observe that by connectivity of Y such a choice is possible. Now suppose that we completed the construction of the points
There are two cases to consider of course. Suppose first that n + I is odd. Pick an arbitrary point e n +i £ Fn+\ \ {ei, . . . , en} and define G = {m < n : em < e n +i},
H — {m < n : em > e n +i},
respectively. Let a = max{/(em) : m £ G} and b = min{/(em) : ra £ H}. Since / is one-to-one, a ^ b and so La n t/& = 0. In addition, La and E/6 are closed so that by connectivity of Y there exists a point /(e n +i) in F strictly between a and 6 (with respect to the order =^). It is easy to see that en+i and /(e n +i) are as required. For n + 1 even we can proceed by a similar argument. 0 Now fix t £ I for a moment and put At = {Lf(d) : d £ E and t < d} U {£//(<*) : d £ E and d < t}. Claim 5. At has the finite intersection property and ^\At consists of precisely one point. Proof. That At has the finite intersection property follows easily from the definition of =<(. Consequently, the compactness of Y implies that f ] A t ^ 0. Now suppose that there exist two distinct points a and b £ f\At- Without loss of generality, a =<; b. Since both La and Ub are closed in Y and disjoint, by connectivity and by the fact that f[E] is dense, there exists d £ E \ {t} such that f(d) lies strictly between a and b. Now if t < d then £/(d) € At which implies that b £ f ] A t C I//(d), which is a contradiction. However, if d < t then t//(d) £ At which implies that a £ f\At C t^/(d), which is also a contradiction.
54
1. BASIC TOPOLOGY
Define a function g : I —> Y by
Claim 6. g is a homeomorphism. Proof. First observe that g is order-preserving. Therefore, since g \ E = /, / is one-to-one and £ C I is dense, it follows that g is injective. Also observe that the range of g is dense in Y since it contains f[E]. By compactness of I it therefore suffices to prove that g is continuous (Exercise A. 5. 9). To this end, let U C Y be open, and take an arbitrary point t e g ~ l [ U ] . By compactness of Y, Claim 5 implies that there exist dQ,di € E such that do < t < d\ and £/(«M n U f ( d o ) C U. A straightforward verification shows that ( d o , d i ) C ^~ 1 [C7], i.e., g~1[U] is a neighborhood of t. <> So we are done with the proof of the theorem.
D
As an application we shall now prove the Mazurkiewicz Theorem about the existence of topological copies of E in certain spaces. Let X be a space and let a, b 6 X. A simple chain connecting a and b is a collection U\ , . . . , Un of open subsets of X such that
(1) « e (2) & e (3) ^ n C 7 j ^ 0 iff H - j | < 1. Lemma 1.5.21. Let X be a connected space and let U be an open cover o f X . For any two points x and y in X there exists a simple chain connecting x and y consisting of elements of U. Proof. Let V be the set of all points in X which are connected to x by a simple chain of elements of U. Then V is clearly open and x E V. We shall prove that V is closed as well. The connectivity of X will then imply that V = X. Take an arbitrary v e V. There exists U 6 U containing v, pick a point t € UnV. Then x and t can be connected by a simple chain C/i, . . . , Un of elements of U. Observe that U n Un ^ 0 and define k = mm{i
1.5. TOPOLOGICAL CHARACTERIZATION OF SOME FAMILIAR SPACES
55
The Mazurkiewicz Theorem 1.5.22. Let X be topologically complete, connected and locally connected. If x and y are distinct points of X then there exists an imbedding $: I —>• X such that $(0) = x and <&(!) = y.
Figure 3.
Proof. Let Q be an admissible complete metric on X. For each n 6 N we shall construct a simple chain t/n,i> • • • •> UnjTn(n) from x to y having the following properties (1) for each i < m(n), Un>i is connected and diam(t/ ni ;) < l/n, (2) for all ii < i% < m(n + 1) there exist j\ < J2 < m(ri) such that CU and The existence of t/i,i, . . . , t/i, m (i) follows from Lemma 1.5.21. Assume that the sets t/n,i> . . . , £/ n ,m(n) have been constructed for certain n. For j smaller than m(n) pick a point PJ £ f/ n ,j n t/nj+i ; put po — ^ and Pm(n) — VBy another application of Lemma 1.5.21 for each 1 < j; < m(n) there exists a simple chain Vj from PJ~I to p^ in t/^^- consisting of connected sets such that (3) if V € Vj then V C UnJ, (4) for every V e Vj, diam(y) < V n+ i. Unfortunately, we cannot simply join these chains together, because of doubling back (see Figure 3). We can obtain the desired simple chain by the following procedure. For every 1 < j < m(n) let Vj = {V^i, . . . , Vj>n(j)}. Put = min{/c < n(l) : (3A <
n
0)}-
Let A = max{/i < n(2) : Vi ;7r n V^ 7^ 0}- Now replace the collection Vi by {Vi,i,...,Vi ) 7 r }, and, similarly, V2 by |y 2 ,A, • • • , ^2,n(2)}- Repeat this with the 'new' Vi U V2 and the 'old' V 3 , etc. At the end of the process, the union of the 'new' Vj's is clearly the required simple chain. This completes the inductive construction.
56
1. BASIC TOPOLOGY
For every n G N put m(n) /
*-- n —
I I
Unjii
i=l
and let
We claim that Y is homeomorphic to the closed unit interval I. Observe that y is a closed subset of X and that x,y G Y. Claim 1. y is a continuum. Proof. That Y is compact follows from the construction and Corollary A. 5. 2. Now apply Proposition A. 10. 7. <0> We claim that Y satisfies the conditions mentioned in Theorem 1.5.19. Let z G Y \ {x,y}. For each n, at least one and at most two of the £/ nj i's contain z. Let An be the union of all the t/n/s preceding these and Bn be the union of all the Un/s following these. Put
n=l
n— 1
Observe that both A and B are open subsets of Y. Now if d &Y \{z} then (1) easily implies that for certain n, d G An U Bn. So A U B = Y \ {z}. We claim that A n 5 = 0. To the contrary, assume that for some n,p € N there exists d 6 yln fi Sp. Without loss of generality, n < p. Let i\ < m(p) be such that z G f/p,^. Since d £ Bp there exists i such that i\ < i < m(p) and d G t/p,;. By (2) there exist ji < j < m(n) such that Up,ii
^r ^n,j\t
Pji —
n,j •
Suppose that z G t/nj. Then by construction, Un>j D An = 0, which is impossible since d G An n [7niJ- . Therefore z 0 C7n> j . Now since z G f/p,H C Unjl ,
ji < j,
we obtain J7n)J- C Bn. But this is clearly a contradiction since d G Unj n A n and Anr\Bn = 0. Put Lz = A U {z} and l^ = B U {2;}, respectively. By Theorem 1.5.19 we conclude that Y w I.
D
Corollary 1.5.23. Let X be topologically complete and locally connected. IfUCX is open and connected and x and y are distinct points of U then there exists an imbedding $: I -> U such that $(0) = x and $(1) —y. Proof. By Theorem A. 6. 3, U is topologically complete. So it satisfies the conditions in Theorem 1.5.22. D
1.5. TOPOLOGICAL CHARACTERIZATION OF SOME FAMILIAR SPACES
57
Corollary 1.5.24. Let X be topologically complete, connected and locally connected. Then X is path-connected and locally path-connected. Remark 1.5.25. A disconnected space cannot be path-connected. So connectivity is essential in the Mazurkiewicz Theorem. The same holds for both completeness and local connectivity. Examples: • A connected and locally connected space which is not path-connected. For every i £ N let i—l
oo
Wi = U [-1 + Vi, 1 - Viln * {l}i X JJ n=l
[-1 + l/i} 1 - l/,]n.
n=i+l
It will be shown in Exercise 1.5.17 that the subspace oo
W=\JWi i—l
of Q is as required. • A connected, topologically complete space which is not locally connected. The sin(1/x)-continuum in the plane is clearly as required. Exercises for §1.5. A spaxe X is called nowhere locally compact if no point has a neighborhood with compact closure. A metric g on a set X is called non-Archimedean if e(x, y) < max{£(;r, z), Q(Z, y ) } for all x, y, z G X.
1. Prove that every space of cardinality less than the cardinality of R is zerodimensional. 2. Let Xn be zero-dimensional for every n G N. Prove that H^Li Xn is zero-dimensional. 3. Let X be a space. Suppose that for every n G N there is a discrete collection Jn of closed subsets of X such that (i) \jyn = x, (2) mesh(Jn) < y n . Prove that X is zero-dimensional. 4. Let X be a compact zero-dimensional space. Prove that the collection of clopen subsets of X is countable. (Observe that N has uncountably many clopen subsets, so the compactness assumption on X is essential.) 5. Let X be a zero-dimensional space and let /: X -> Y be an open and closed continuous surjection. Prove that Y is zero-dimensional. 6. Let X be zero-dimensional. Prove that X admits a zero-dimensional compactification.
58
1. BASIC TOPOLOGY
7. Let X be a zero-dimensional cr-compact space. Prove that there is a nullsequence (An)n of compact subsets of X such that An n Am = 0 if n ^ m and Ur=i An = X. 8. Let X be a locally connected and locally compact space. In addition, let U C X be connected and open. Prove that if K C U is compact then there is a continuum C C [7 which contains .K". 9. Let X be a Peano continuum and let U and V be connected open subsets of X such that [7 U V = X. Prove that for all compact subsets A and B of X with A C. U and B C V there are subcontinua /C and L m X such that A C K C [7 and B C L CV while moreover K\J L = X. 10. Prove that for every space X there exist a zero-dimensional space Y and a continuous surjection /: Y —*• X. 11. Let X be a space. Prove that the following statements are equivalent: (1) X is zero-dimensional, (2) X has an admissible metric which is non-Archimedean. 12. Let X be locally compact and totally disconnected. Prove that X is zerodimensional. 13. Let X be totally disconnected. Prove that if x € X and K C X \ {x} is compact then there is a clopen C C X such that x £ C C X \ K. 14. Let £7 be Erdos' space. Prove that E is homeomorphic to E x E. 15. Prove that both Q are P are nowhere locally compact. 16. Prove that if infinitely many factors in a product H^L} Xn are not compact then it is nowhere locally compact. 17. Prove that the space W described on Page 57 is connected and locally connected, but not path-connected.
1.6. The inductive convergence criterion and applications In this section we present a method enabling us to construct new homeomorphisms from old ones, namely, the Inductive Convergence Criterion. The method is illustrated with applications: we deduce the topological homogeneity of Q and we present results on countable dense homogeneity. Lemma 1.6.1. Let (X,g) be a complete metric space and let (An}n be a sequence of subsets of X. Suppose that (xn)n is a Cauchy sequence in X such that for every n, (*)
e(zn+i, xn) < 3~n • min{g(xl, A*) : 1 < i < n}.
Then lin^^oo xn does not belong to (J^Li -^-n-
1.6. THE INDUCTIVE CONVERGENCE CRITERION AND APPLICATIONS
59
Proof. Take an arbitrary n G N and let x = linin-^oo xn. We shall prove that x $_ An. To this end, observe that condition (*) implies that for arbitrary m we have Q(xm+n,x(m_i]+n)
< 3-« m ~ 1 )+ n ) •e(xn, An) < 3~m • g(xn,An).
From this it follows that TO
' ^\~^ni
Q(Xm+ni Xn) ^ Z—~i i=l
n
>'>
and as
we obtain oo £}{ Jb j Jt>ri)
Hill
77T.—$-00
tr ^^TTl-f-Tl 5 JL/TI )
_
/
* ^
^
y\t^rn ; -*~*~7l)
/2 c: V Tl 5 -**-7lJ ;
so x $. An.
D
Observe that in the above lemma the distance between x n +i and xn 'depends' only on the points £1,... ,xn. Hence if one wishes to choose the points (xn)n inductively so that (*) is satisfied, at stage n + I the choice of xn+i is subject to n - hence finitely many - conditions. Let X be a compact space and let (hn)n be a sequence in ^K(X). It is clear that for each n e N the function fn = hn o • • • o hi
belongs to 'H(X). If / = lim.^-^ fn exists then it will be denoted by lim hn o • • • o hi
n—>oo
and is called the infinite left product of the sequence (hn}n. We want to find conditions on the sequence (hn)n which ensure that linin^oo hno- • -ohi exists and belongs to "H(X). The Inductive Convergence Criterion 1.6.2. Let (hn)n be a sequence in "H(X) such that X is compact. In addition, assume that for all n G N, (1) Q(hn+ (2) e(hn+l,lx} Then h — limn_^00 hn o • • • o hi exists and belongs to 'H(X). Proof. For every n, put fn — hn o • • • o hi. Condition (1) and Exercise 1.3.2 imply that for all n, (3)
£(/n+i, /n) = e(hn+i o • • • o hi, hn o • • - o hi) = e(hn+i,lx) < 2~ n .
60
1. BASIC TOPOLOGY
We conclude that the sequence (fn)n
ls
£-Cauchy and consequently,
/ = lim fn n—>-oo
exists (Corollary 1.3.6). Since fn E S ( X , X ) for every n, Proposition 1.3.8 implies that / 6 S ( X , X ) . Now by (2), (3), Lemmas 1.6.1 and 1.3.10 we obtain n=l
as required.
D
The above result tells us that if the sequence (@(hn, 1)) converges rapidly to 0, then the infinite left product limn^.00 hn o • • • o h\ is a homeomorphism. It turns out that we are never interested in the precise speed at which (@(hn,l)} converges. We are always in the pleasant situation that while inductively defining the sequence (hn)n, we are able to choose the next homeomorphism 'sufficiently close' to the identity. This simplifies life considerably. Application 1: Topological homogeneity of the Hilbert cube.
Every nontrivial topological group has fixed-point free homeomorphisms. Simply observe that if (G, °) is a topological group then every translation of the form x i->- g o x, where g is not the identity element of G, is a fixed-point free homeomorphism. If Q were a topological group then the proof of its homogeneity would be trivial by the above remarks. But Q is not a topological group since it has the fixed-point property, as we will see later in Corollary 2.4.6. It is geometrically obvious that no I n , n G N, is homogeneous. For n = I this is a triviality. Simply observe that E\{0} is connected while I\{ l/2} is not. For larger n the proof is more complicated. See the proof of Theorem 2.4.14 for more details. As we announced, we wrill prove that Q is homogeneous, thereby demonstrating a striking difference between the 'finite dimensional' and the 'infinitedimensional' situation. We put 5= n-1
respectively. We call s the pseudo-interior and B(Q) the pseudo-boundary ofQ.
1.6. THE INDUCTIVE CONVERGENCE CRITERION AND APPLICATIONS
61
This notation is slightly confusing since s was already our abbreviation for R°°. But M°° and 0^=1 (~~1» l)n are homeomorphic and so this abuse of notation is not as bad as it seems at first glance. Lemma 1.6.3. Suppose that x,y e s C Q. Then there is a homeomorphism h 6 3t(Q) with h(x) = y. Proof. Since for each i, Xi, yi 6 (— 1, l) t , there is an element hi G ^K(JJ;) such that hi(xi) = yl, we may take h — h\ x /i2 x • • • (Exercise A. 1.13). D We see that all points in the pseudo-interior are topologically equivalent. Consequently, if we show that every point in Q can be 'homeomorphed' into s, then we have shown that Q is homogeneous. We need a preliminary lemma. Lemma 1.6.4. Suppose that x G Q, that m e N, and that e > 0. Then there is an element h 6 'H(Q) such that (1) e(h,!Q) <e,
(2) /i(z) m e(-i,i) m , (3) h does not affect
the first m — 1 coordinates of any point, i.e.,
h(y)i = yi
for all i < m — 1 and y e Q.
«UI m
1 l
— z,2~( n ~ 1 )
Figure 4. Proof. If |x m | 7^ 1, let /i = IQ. Therefore, without loss of generality assume that xm = 1. Let n > m be such that 2~( n ~ 2 ) < e. It is geometrically obvious that there is a homeomorphism / : J I m x J I n ^ j r m x J I T l such that (4) / ( p , g ) - ( p , g ) i f p < l - 2 - ( - 1 ) , (5) /[{I} x JJn] is contained in JJm x {1} and its projection on J m is contained in the open interval (1 — 2~( r i ~ 1 ), 1).
62
1. BASIC TOPOLOGY
For a moment we think of Q as (JJm x JJ n ) x R. where R denotes the product of all remaining factors. We define h = f x IR. So h only affects the n-th and m-th coordinate of any point. To verify that h is as required, first observe that (2) and (3) are clearly satisfied. For (1), notice that for every x £ Q, Q(h(x),x) = 2~m • \h(x)m -xm\+2-n- \h(x)n-xn\ < i.2-("- 1 ) +2~n - 2
< £.
From this it follows that g(h, IQ) < e (Exercise A. 5. 4).
D
Lemma 1.6.5. Let x G Q. There is an h G M(Q) with h(x) e s. Proof. The homeomorphism h shall be of the form lim hn o • • • o hi ,
n—>-
where (hn)n is an inductively constructed sequence of homeomorphisms of Q. To ensure convergence in the limit, at each stage of the construction the next homeomorphism is constructed in accordance with the smallness conditions in the Inductive Convergence Criterion 1.6.2. Specifically, the sequence (hn}n shall satisfy the following conditions: (l)n hn does not affect the first n — I coordinates of any point, (2) n (hn o • • • o hi(x))i G (-1, 1) for every i < n, (3)n Q(hn, IQ) is so small that the conditions of the Inductive Convergence Criterion 1.6.2 are satisfied. Apply Lemma 1.6.4 to find a homeomorphism / 6 "H(Q} such that /Or)! 6 (-1,1) and define hi = f . Then hi is as required since (l)i and (3)i are empty conditions. Now suppose that hi has been defined for every i < n. The Inductive Convergence Criterion 1.6.2 gives us a magic e > 0 and tells us that we must choose the next homeomorphism e-close to the identity. We are not interested at all in the precise value of e. The only thing we need is that such e > 0 exists. By Lemma 1.6.4 there is an element /in+i G 3t(Q) such that g(hn+i, IQ) < £, hn+i does not affect the first n coordinates of any point, and (hn+i(hno.-.ohl(x)))n+1 It is evident that hn+i is as required.
G (-1,1).
1.6. THE INDUCTIVE CONVERGENCE CRITERION AND APPLICATIONS
63
Now put h = linin-^oo hn o • • • o hi. By construction, h G ^K(Q). continuity of the projections 7rm : Q —>• J m we find that for every ra,
By
h(x)m = ( lim /in o • • • o /ii(x)) m n—>oo
= lim (hn o • • • o hi(x)) n>m
m
i.e., /i(x) 6s.
D
We now come to the announced result: Theorem 1.6.6. The Hilbert cube Q is homogeneous. Proof. Take x,y G Q. By Lemma 1.6.5 we find /ii,/&2 € ^(Q) such that both hi(x) and /i2(y) belong to s. In addition, by Lemma 1.6.3, there is an element / G ^K(Q) with f ( h i ( x ) ) = h2(y). Then takes x onto ?/.
D
Application 2: Countable dense homogeneity. A space X is called countable dense homogeneous provided that for all countable and dense subsets D,E C X there exists an element / G W(X) with f[D] = E. The topological sum of S1 and §2 is an example of a countable dense homogeneous space which is not homogeneous. But for connected spaces it will turn out that countable dense homogeneity implies homogeneity. Let X be a space and fix an arbitrary point x G X. Let r(x) denote the type of x, i.e., Observe that r(x) is invariant under "H(X}, that is, for all y G T(X) and all h G 'H(X) the image h(y) also belong to T(X). Theorem 1.6.7. If X is countable dense homogeneous and x G X then T(X) is clopen in X . Proof. We will show that r(x) is clopen. Claim 1. r(x) is closed. Proof. To the contrary, assume that there exists p G T(X) \ T(X). Let E be a countable dense subset of r(x), and let F be a countable dense subset ofU = X\ T(X). Then both
A = E U F,
B = A U {p}
64
1. BASIC TOPOLOGY
are countable dense subsets of X and hence there exists by assumption an element h e 'K(X) with h[A] = B. Since p 0 h[E] because h[r(x}] C T(X), there exists b E -F with /i(6) = p. But then /&[£/] is a neighborhood of p and hence intersects r(x). But this implies that for some q 6 C/ we have that /i(g) G r(x), i.e., q 6 T(#). This is a contradiction. 0 Claim 2. T(X) is open. Proof. It is easy to see that r(x) is open if and only if Intr(ar) 7^ 0. Striving for a contradiction, suppose therefore that Intr(x) = 0. By Claim 1 it follows that r(x) is a closed subset of X with empty interior, hence is nowhere dense. Let E C U = X \ T(X] be countable and dense in X. There is an element h e 3t(X) with h[E] = EU{x}. Pick e £ E such that /i(e) = x. But then e G r(a;), which is a contradiction.
D
Corollary 1.6.8. A connected countable dense homogeneous space is homogeneous. So for connected spaces, countable dense homogeneity is a strong form of homogeneity. There is an example of a connected homogeneous space which is not countable dense homogeneous, see VAN MILL [294] for details. A space X is called strongly locally homogeneous provided that it has an open base U such that for all U e U and points x,y G U there exists an element h E 3i(X) such that h(x) ~ y and h(z) = z for z 0 U. Such a homeomorphism is said to be supported on U. It is geometrically obvious that the Euclidean spaces W1 , n G N, are strongly locally homogeneous. Before we come to the main result in this Application, we make a few remarks. Let X be a locally compact space. Observe that if / G 'H(X) then the function af : aX —>• aX defined by
af(x) = ^ J v ' \ oo
,
x
(a; = oo),
belongs to CK(aX). It will sometimes be convenient to think of a locally compact space as subspace of its Alexandroff compactification. Theorem 1.6.9. Let X be locally compact and strongly locally homogeneous. Then X is countable dense homogeneous. Proof. Let A — {01,02,...} and B = { & i , & 2 ) - - - } be faithfully indexed dense subsets of X. The hypothesis of strong local homogeneity implies that for each neighborhood U of a point x G X, and for any dense G C X, there exists a homeomorphism of aX which is supported on U and takes x into G
1.6. THE INDUCTIVE CONVERGENCE CRITERION AND APPLICATIONS
65
(use that G n U ^ 0). Using the Inductive Convergence Criterion 1.6.2, we construct a sequence (hn}n of homeomorphisms of aX such that its infinite left product h is a homeomorphism and such that the following conditions (which ensure h[A] = B and h \ X G IK(JQ) are satisfied: (1) hn o • • • o hi (di) = h-2i ° • • • o h\_ (di) G B for each i and n > 2i, (2) (hn° • • • °hi)~l(bi) = (hii+i o • • -ohi)~l(bi) G A for each i and each n > li + l, (3) for all n, hn(oo) = oo. Assume hi, . . . , h
66
1. BASIC TOPOLOGY
(1) (2) (3) (4) Prove
lln is a star-refinement of lt n -i, mesh(Un) < 2"71, (hn-i o • • • o hi)[Un] has mesh less that 2~n, hn is limited by Un. that h = limn^.oo hn ° • • • ° h\ is a homeomorphism of X.
2. Use Exercise 1.6.1 to show that a topologically complete space which is strongly locally homogeneous is countable dense homogeneous. 3. Prove that both s and B(Q) are dense in Q. Prove that B(Q) is cr-compact. Are s and B(Q) homeomorphic? 4. Prove that an open subspace of a strongly locally homogeneous space is strongly locally homogeneous. 5. Let X be a compact space with closed subspace A. Prove that there is a continuous function /: X —> Q such that (2) / \ X \ A is one-to-one. 6. Let X be a homogeneous space. Prove that the following statements are equivalent: (1) X is a Baire space. (2) X is not meager in itself. 7. Let X be a space and x £ X. Prove that for every h e !K(A") we have h[r(x}] = r(x). 1.7.
Bing's shrinking criterion
In this section we provide a second method for constructing new homeomorphisms from old ones, namely Bing's Shrinking Criterion, which is one of the most powerful tools in geometric topology. As an application we will prove that Q is homeomorphic to its own cone. Let X and Y be compact spaces. A continuous surjection /: X —>• Y is called a near homeomorphism provided that for every e > 0 there is a homeomorphism h: X —> Y such that g(h, f) < e. Clearly, / e C(X,Y) is a near homeomorphism if and only if / belongs to the closure of 3-C(X, Y) in C(X, Y}. Simple examples show that a near homeomorphism need not be a homeomorphism (Exercise 1.3.5). Near homeomorphisms are very important in geometric topology. Let X and Y be compact spaces. A continuous surjection /: X —> Y is called shrinkable provided that for every e > 0 there is a homeomorphism h: X —> X satisfying (1) diam (hf~1(y}}
< e for every y € Y (i.e., / o h~l is an e-map),
Shrinkability means that the point-inverses of / can be uniformly shrunk to small sets by a homeomorphism of X that, from the standpoint of the
1.7. BING'S SHRINKING CRITERION
67
space F, does not change / very much. The reader should pause to get used to this definition. Ding's Shrinking Criterion says that shrinkablity is the same as being a near homeomorphism. If the domain of the function under consideration has a rich supply of homeomorphisms, e.g., if it is a manifold, then verifying shrinkability is usually easier than verifying directly that the function in question is a near homeomorphism. Lemma 1.7.1. Let X and Y be compact spaces and let f : X —> Y be shrinkable. Then for each e > 0 and for each (p 6 'K(X) there exists an element ip £ ^K(X) such that (1) f oip is an e-map, (2) Q ( f o < f > , f o i j , ) < £ . Proof. Let 7 > 0 be such that for A C X, if diam(A) < 7 then (*)
diam (p"1 [A]) < e
(use the fact that (p~l is uniformly continuous (Exercise A. 5. 18)). Since / is shrinkable, there is a homeomorphism h G 3t(X) such that (3) g(f,foh)<£, (4) V y e r jdiam^-Hy)) < 7Since £(/, / ° h) < e, it follows that g(fh~l(p,fhh~1(p} so that
< £ (Exercise 1.3.2),
We claim that f h ~ l ( p is an e-map. If so, then ijj = h~l(p is as required. To this end, take an arbitrary y E Y and consider (fh~l(p)~l(y} = ( p ~ l h f ~ l ( y } . Since diam ( h f ~ l ( y ) } < 7 we see by (*) that diam ((p~1hf~1(y)} < e, and this is what we had to prove. D Bing's Shrinking Criterion 1.7.2. Let X and Y be compact spaces and let f : X —>• Y be a continuous surjection. Then f is a near homeomorphism if and only if f is shrinkable. Proof. First assume that / is a near homeomorphism. Let e > 0. There is a homeomorphism g: X —>• Y such that g(g, /) < £/2. Find 7 > 0 such that if A C Y has diameter less than 7 then 5-1[A] has diameter less than e (use the fact that g~l is uniformly continuous (Exercise A. 5. 18)). Let p: X —>• Y be a homeomorphism such that g(p, /) < min{Y2? e/2J and put h = g~l ° p. We claim that h is a shrinking homeomorphism. Since £>(/,p) < £/2 and
= Q(g,f)
68
1. BASIC TOPOLOGY
(we used Exercise 1.3.2 twice), it follows that g(fjh]
< g(f,p) + g(pjh) < */2 + */2 = e.
Now take an arbitrary y G Y. If z G p(f~l(y)) then since g(p, /) < 0/2 we get g(y,z) < T/2- So, diam (p(f~1(y})) < 7 from which it follows that diam (g~lp(f~l(y}}} < e. We conclude that diam (h(f~1(y))} < £. Now assume that / is shrinkable. Let e > 0 and let ho be the identity homeomorphism on X. By applying Lemma 1.7.1 inductively, we find that there exists a sequence (hn}n>\ in 'K(X) such that if we put pn = f ° hn for each n > 1, then the following conditions are satisfied: (1) pn is a (2) Q(pn+i,pn) <3~n-e and (3) g(pn+i,pn) < 3 ~ n
By (2), the sequence (pn)n is Cauchy. Since g is complete, the function p = lim pn n—too
is well-defined and is continuous by Lemma A. 3.1. Observe that by Proposition 1.3.8, p G S(X,y). Again by (2), it follows that
i-l
Now apply Lemma 1.6.1 to the sequence (pn)n to obtain
n=l
i.e., p G ^(X^Y) (Lemma 1.3.10). We conclude that / is a near homeomorphism. D Corollary 1.7.3. Let X and Y be compact spaces and let /: X —>• Y be shrinkable. Then X and Y are homeomorphic. Application: The cone over the Hilbert cube is the Hilbert cube. Recall that for a compact space the cone over X is the space we get from X x I by identifying X x {1} to a single point oo (see Page 515). It is geometrically clear that for each n the cone over In is homeomorphic to I n+1 and so naively one would expect, by taking the 'limit' as n goes to infinity, that A(Q) « Q. That this is indeed true will be demonstrated here as an application of Bing's Shrinking Criterion. Lemma 1.7.4. For every n G N, t G [— 1, 1) and e > 0 there exists a homeomorphism h: J J n x J I ^ - J I n x J I such that
1.7. BING'S SHRINKING CRITERION
69
(1) h \ J n x [—!,£] is the identity function, (2) the diameter of h[$n x {!}] is less than e. Proof. The lemma is geometrically obvious. Alternatively, observe that the Lemma is clear for n — 1 since then h can be found such as in the proof of Lemma 1.6.4, and proceed inductively. D Lemma 1.7.4 allows us to conclude that: Theorem 1.7.5. A(Q) is homeomorphic to Q. Proof. Let p: Q x E —>• A(Q) be the natural quotient map. We will show that p is shrinkable. From Corollary 1.7.3 it wall then follow that
n ~n v TT ~ /\(n\ \cS —t* /-^ \cJ —f xx Jl r*~* L i \V\os T/ /11 Let e > 0 and let U be a neighborhood of oo G A(Q) having diameter less than e. There exists t G (0,1) such that (Q x [t, 1)) U {00} C U. Find n G N such that oo
]T 2~ m < e/2 m=n
and applying Lemma 1.7.4 find a homeomorphism h: J n x I —>• J n x I such that (1) /i C JJn x [0,t] = 1, (2) diam(/i[JJn x {!}]) < e/2.
Define // G tt(g x E) by (we make a few obvious identifications here). By the special choice of n it follows easily that dia,m(H[Q x {!}]) < e. We claim that g(p,po H} < e. Take a point (x, s) G Q x I. If s < i then #(x, s) = (,x, s), so
Q(p(x,s},pH(x,s}) = 0. If s > t then (p(x,s),pH(x,s)} C U. Therefore, since diam(L r ) < e, we find that g(p(x, s),pH(x, s)) < £. We conclude that p is shrinkable. D Consider the point oo G A(Q). Basic closed neighborhoods of this point have the form Ut = ( Q x [ t , l ) ) u { c x D }
(te (0,1)).
Observe that the boundary of Ut is equal to Q x {t}. Consequently, Theorem 1.7.5 implies that oo has arbitrarily small closed neighborhoods U such that (1) U is homeomorphic to Q, (2) Fr([7) is homeomorphic to Q.
70
1. BASIC TOPOLOGY
By the homogeneity of Q (Theorem 1.6.6), each point of Q has arbitrarily small closed neighborhoods U satisfying (1) and (2). In particular, each point of Q has arbitrarily small closed neighborhoods with contractible boundaries. This demonstrates a striking difference with the finite dimensional situation.
Exercises for §1.7.
1. Let X be a compact space. Let Z be the space we obtain from A(X) x I by identifying A(X) x {1} to a single point, and let TT: A(X) x I —>• Z be the decomposition map. Prove that TT is shrinkable and conclude from this that A(ApQ) « A(X) x I. 2. Let X be a compact space and let Y = A(X) x Q. Prove that Y is homeomorphic to its own cone 3. Let X, Y and Z be compact spaces and let / : X —>• Y and g : Y —> Z be near homeomorphisms. Prove that gof : X —>• Z is a near homeomorphism. 4. Let p: Q —>• 0^2^* be the projection. Prove that p is a near homeomorphism.
1.8. Isotopies
We will now present the third and last method for constructing new homeomorphisms from old ones by means of so-called isotopies. In comparison to the Inductive Convergence Criterion and Bing's Shrinking Criterion, at first glance this method does not seem to be as important as the other two since direct appealing applications are hard to find. However, it turns out that this method is as fundamental as the preceding two. Let X and Y be spaces. Recall that a homotopy from X to Y is a continuous function H: X x I —>• Y. We call a continuous function
H: X x K ^Y, where K is any compact space, a K-homotopy. If t G K then the function Ht: X —> Y defined by Ht(x) — H(x,i) is called the t-th level of H. A K-isotopy from X to Y is a K-homotopy H: X x K —> Y each level of which is a homeomorphism from X onto Y. If K C M is a compact interval then a K-homotopy (resp. K-isotopy) is simply called a homotopy (resp. isotopy). The proof of the following proposition is a triviality and is left as an exercise to the reader.
1.8. ISOTOPIES
71
Proposition 1.8.1. For each n G N let Hn : Xn x Kn ->• Yn be a Kn-isotopy. Then the function OO
H: l[Xnx n=l
OO
OO
*[[Kn-> JJr n , n=l
n=l
defined by H(x,t)n = Hn(xn,tn) is a
(n e N),
n^Li Kn-isotopy.
We shall now present our last method for constructing new homeomorphisms from old ones. Theorem 1.8.2. Let K be a compact space. If H: X x K —>• X is a Khomotopy and a: Y —>• K is continuous then so is the function /: X xY -*X xY
defined by (1) f ( x , y ) = (H(x,a(y»,y)Moreover, if X and Y are compact and H is a K-isotopy then f is a homeomorphism. Remark 1.8.3. Observe that the function / defined in (1) is 'level preserving', i.e., / does not change the second coordinate of any point. Proof. That / is continuous is trivial. Assume next that X and Y are compact and that H is a K-isotopy. Since / is 'level preserving' and since each 'level' of / is a homeomorphism from X onto Jf, it follows that / is surjective. By compactness we therefore only need to show that / is one-toone (Exercise A.5.9). To this end, take distinct (a, &),(«', 6') eXxY. If b / b' then clearly /(a, b) / /(a',6')- If b = b' then a(b) = a(fc') and a / a'. Consequently, /f (a,a(6)) ^ H(a',a(b')) since Ha^ is one-to-one. D Application: Extending certain homeomorphisms. Isotopies are very useful in situations where one wants to extend a given homeomorphism. This will be demonstrated in a simple situation here. The same ideas will be used to derive much more complicated results in §5.2. Let E, F C K. be compact subsets and let /: E —> F be a homeomorphism. It is easy to see that it is not always possible to extend / to a homeomorphism /: 1R —>• E. For example, let E = F = {0,1,2} and define / by/(0) = l,/(l) = O a n d / ( 2 ) = 2. Identify R and the £-axis in R 2 . Although in general a map / between compact subsets of E cannot be extended over R, an easy application of
72
1. BASIC TOPOLOGY
Exercise 1.2.9 yields that if / is a homeomorphism then it can be extended to a homeomorphism /: R2 —> R 2 . We shall now present a different proof of this which much more than the solution of Exercise 1.2.9 illustrates the technique of extending homeomorphisms that we are going to use in §5.2. Theorem 1.8.4. Let E,F C R be compact and let f : E —> F be a homeomorphism. Then f can be extended to a homeomorphism /: R2 -> R 2 . Proof. We shall prove the theorem for the interval (—1,1) instead of R. Since ( — 1, 1) and R are homeomorphic, we are allowed to do this. So let E,F C (—1,1) be compact and let /: E —>• F be a homeomorphism. Let K C (—1,1) be a compact interval containing both E and F. For each t G K let Ht '• J —> J be the unique homeomorphism taking [—!,£] linearly onto [—1,0] and [t, 1] linearly onto [0,1]. It is easily seen that the function H : JJ x K —> J defined by H(x,t) — Ht(x) is an isotopy (Exercise 1.8.1). Let F C ( — 1, 1)2 be the graph of the function /, i.e.,
Our aim is to find a homeomorphism h: J 2 —> J 2 that takes F onto {0} x F. W7e achieve this by applying Theorem 1.8.2. Define the function (p: F —>• K as follows:
h ( x , f ( x ) ) = (H(x,a(f(x»)JW)
= (H(x,x)J(x))
= (0, /(*)).
We conclude that h is as required. By precisely the same argumentation we can find a homeomorphism g: JJ2 —>• JI2 such that for every (x, /(x)) G F, (2)
g(x,f(x))=(x,Q).
Now define ^ : J 2 -> JJ2 by £(x, y) = (y, x) and put / = £ o h o g~l . Then / is a homeomorphism of J 2 and it is easily seen that /[(—I, I) 2 ] = ( — 1, 1)2. We claim that / extends / and is therefore as required. Take an arbitrary x G -E1. Then
(apply (1) and (2)).
D
1.9. HOMOGENEOUS ZERO-DIMENSIONAL SPACES
73
Exercises for §1.8.
1. Let -1 < a < b < I and for each t e K = [a, b] let Ht: J -> J be the unique homeomorphism taking [—!,£] linearly onto [—1,0] and [£, 1] linearly onto [0,1]. Prove that the function H: Jx K-» J
denned by H ( x , t ) = Ht(x) is an isotopy. ^•2. Prove that Q is strongly locally homogeneous. 3. Let D C Q be countable and dense. Prove that Q \ D is homogeneous. 1.9. Homogeneous zero-dimensional spaces Homeomorphism theory in the class of zero-dimensional spaces is particularly elegant. It turns out that various classical zero-dimensional spaces such as Q and F can be topologically characterized as nicely as the Cantor set. Homogeneous zero-dimensional spaces. Our first aim is to characterize the class of all homogeneous zero-dimensional spaces. Lemma 1.9.1. Let X be a zero-dimensional space and let x,y 6 X. If for each e > 0 there are clopen neighborhoods U and V of x and y, respectively, such that (1) diam(C7) < e, diam(V) < e, (2) U is homeomorphic to V, then there is a homeomorphism h: X —> X with h(x] — y. Proof. For each n > 0 we will construct a clopen neighborhood Un of y and a clopen neighborhood Vn of x and for each n > 1 an element hn € "K(X} such that (3) (4) (5) (6)
diam(C/n) < 2~ n , diam(K) < 2~ n , UnUhn^o-.-oh^Vn] C C7n_! and h^1 o • • • o h'^ [Un] U Vn CVn-i, hn°...ohi[Vn] = Un, hn is supported on Un-\.
Put UQ = VQ — X and suppose that for some n > 0, Vi and U% have been constructed for all 0 < i < n — 1 and hi for all 1 < i < n — 1. If hn-i o ••• ohi(x) = y then let V be any clopen neighborhood of x of diameter less than 2~ n such that both V and U = hn-\ o • • • o hi[V] have diameter less than 2~n. Put
74
1. BASIC TOPOLOGY
Then our choices are obviously as required. Suppose therefore that
hn-i o ••• ohi(x) -£ y. Let V C Vn-i be a clopen neighborhood of x and U C Un-\ be a clopen neighborhood of y such that (7) diam(V) < 2~ n , diam(t7) < 2~ n , (8) V is homeomorphic to [7, (9) hn_i o • • • o hi[V] CC7 n _!,
(10) fcn_i o... 0 /i 1 [v r ]nt/ = 0.
Let /: /in-i ° • • • ° hi [V] —> C7 be a homeomorphism. Define Un = U,Vn = V and /i n : A" —>• X by
{
x f(x)
(x ^C/ n U/i n _! o - - - o h i [ V n ] ) , (x£hn-io...ohi[[Vn]),
f-l(x]
(xeun).
It is clear that Un, Vn and hn satisfy our inductive requirements. If p £ X and g(p, x) > 2~n then p $ Vn and so hn ° • • • ° hi (p) £ hn o • • • o hi [Vn] = Un.
By (6) this implies that (11)
hk°'--°hi(p) = hno---ohi(p)
for all k > n. Consequently, if we define h: X —>• X by h = lim hn ° • • • o hi n—»oo
then /i is well-defined. Observe that by (4) and (5), h(x) = y. We claim that h is a homeomorphism. If g(p,x} > 2~n then, as remarked above, hn-i o • • • o / i 1 ( p ) £ C7n which implies by (11) that /i(p) ^ C/n. Since /i(x) = y, this implies that h is one-to-one. In addition, if g(p,y) > 2~n then p £ Un and therefore by (4) if
q = h~l o - - - o h ~ l ( p ) then h(q) = p. We conclude that h is surjective. By (3) and (6), h is continuous (Lemma A.3.1). It is clear that h is open at all points but x. We therefore only check openness of h at x. This will give us that h is a homeomorphism. To this end, let V be any neighborhood of x. Let n > 0 be so large that Vn C V. Then clearly y = h(x) e Un = h[Vn] C h[V]. We therefore conclude that h[V] is a neighborhood of y, which implies that h[V] is open. D Lemma 1.9.1 will be used among other things in Exercises 1.9.1 and 1.9.2 for the construction of a partition of E into two homeomorphic homogeneous parts.
1.9. HOMOGENEOUS ZERO-DIMENSIONAL SPACES
75
Corollary 1.9.2. Let X be zero-dimensional. The following statements are equivalent: (1) X is homogeneous. (2) if x, y E X then x and y have arbitrarily small homeomorphic clopen neighborhoods. Strongly homogeneous zero-dimensional spaces. Let X be zerodimensional. We say that X is strongly homogeneous provided that all nonempty clopen subsets of X are homeomorphic (and hence are homeomorphic to X). By Corollary 1.9.2 it follows that a strongly homogeneous (zero-dimensional) space is homogeneous. Corollary 1.9.3. A strongly homogeneous (zero-dimensional) space is homogeneous. A homogeneous zero-dimensional space need not be strongly homogeneous (VAN DOUWEN [128]), but at least the following weaker statement holds. Lemma 1.9.4. A homogeneous zero-dimensional space is strongly locally homogeneous. Proof. Let x £ X and let U be an arbitrary clopen neighborhood of x. In addition, pick an arbitrary y E U \ {x}. We claim that there is a homeomorphism h: X —>• X sending x onto y and which moreover restricts to the identity on the complement of U. Since X is homogeneous, x and y have arbitrary small clopen neighborhoods which are homeomorphic by homeomorphisms sending x onto y. So we can find disjoint clopen neighborhoods E and F of x and y, respectively, and a homeomorphism f : E —> F such that (1) E(JF CU, (2) f ( x ) = y.
Now define h: X -> X by
{ Then h is clearly as required.
p
(p 0 E U F),
f(p)
(P € £),
/-'(P) (peF). D
Lemma 1.9.5. The Cantor set C is strongly homogeneous and strongly locally homogeneous. Proof. This is clear from the topological characterization of C (see Theorem 1.5.5) since every nonempty clopen subspace of C is compact, zerodimensional and has no isolated points. D
76
1. BASIC TOPOLOGY
Topological characterization of Q and F. We now apply the results obtained so far to get interesting topological characterizations of various classical zero-dimensional spaces. Theorem 1.9.6. The space of all rational numbers Q is topologically the unique nonempty countable space without isolated points. Proof. It is clear that Q is countable, nonempty, and has no isolated points. It therefore suffices to prove that all nonempty countable spaces without isolated points are homeomorphic. To this end, let X and Y be such spaces. By Exercise 1.5.1, both X and Y are zero-dimensional. They consequently have zero-dimensional compactifications aX and aY by Exercise 1.5.6. Both aX and aY clearly have no isolated points because X and Y both have no isolated points. So aX w aY & C by Theorem 1.5.5. Since C is strongly locally homogeneous (Lemma 1.9.5), it is countable dense homogeneous by Theorem 1.6.9. So there is a homeomorphism /: aX —>• aY with f[X] = Y. In particular, X w Y. D Remark 1.9.7. Our next result is a topological characterization of the space of irrational numbers P. There is a proof of this characterization along the same lines as the proof of Theorem 1.9.6. But it is simpler by using the completeness of P to follow the method in the proof of Theorem 1.5.5. Theorem 1.9.8. The space of all irrational numbers P is topologically the unique nonempty, topologically complete, nowhere locally compact and zerodimensional space. Proof. That P is topologically complete, zero-dimensional and nowhere locally compact, follows easily from Remark A.6.5, Proposition 1.5.3 and Exercise 1.5.15. Claim 1. Every nonempty open subset of a nowhere locally compact zerodimensional space can be decomposed into infinitely many pairwise disjoint nonempty clopen sets. Proof. Let X be nowhere locally compact and zero-dimensional. Let U C X be nonempty and open. The proof of Proposition 1.5.1 shows that we can decompose U into nonempty clopen subsets of X (possibly, a finite number). So without loss of generality we may assume that U is clopen. Since X is nowhere locally compact (Exercise 1.5.15), U is not compact. It consequently has an open cover U without finite subcover. The proof of Proposition 1.5.1 shows that this cover can be refined by a clopen partition (here clopen means clopen inX). This partition is necessarily infinite since if it were finite then U would have a finite subcover.
1.9. HOMOGENEOUS ZERO-DIMENSIONAL SPACES
77
isomorphic trees of clopen sets which form bases of X and Y, respectively. A path P through these trees has a unique point as its interesting because of compactness and the fact that the elements of P have smaller and smaller diameters. The trees are finitely branching, i.e., each node has finitely many successors only. With topologically complete, zero-dimensional nowhere locally compact spaces similar trees of clopen sets can be constructed (their construction is easier than in the case of Cantor sets). By applying Claim 1 we can ensure that all levels of the trees are infinite, that they have countable height and that they are infinitely branching, i.e., each node has infinitely many (hence ui many) successors. So the trees are obviously isomorphic. By using complete metrics, one can again ensure that for each path P the intersection of its elements is a single point. So the proof can then be completed such as in the proof of Theorem 1.5.5. D Corollary 1.9.9. P and N°° are homeomorphic. Proof. Use Exercise 1.5.16 to conclude that N°° is nowhere locally compact. Zero-dimensionality and completeness follow from Proposition 1.5.3, Exercise 1.5.2, Remark A.6.5 and Lemma A.6.2. So we are done by Theorem 1.9.8. D Corollary 1.9.10. If X is topologically complete then it is a continuous image of P. Proof. By Theorem A.6.3, X is homeomorphic to a closed subspace of E°°. Claim 1. IR°° is a continuous image of P. Proof. Since P°° w (N00)00 w N°° % P (Corollary 1.9.9), it suffices to prove that M is a continuous image of P. Consider the zero-dimensional product space Y — NX C. Since {n} x C maps onto [—n, n] by Theorem 1.5.10, Y maps onto E. But P x Y satisfies the conditions mentioned in Theorem 1.9.8 (cf. the proofs of Theorem 1.5.10 and Corollary 1.9.9) and hence is homeomorphic to P. So we conclude that P maps onto Y and so in turn maps onto R. 0" So we are done.
D
This yields the following characterization of analytic spaces. Corollary 1.9.11. Let X be a space. The following statements are equivalent: (1) X is analytic, (2) X is a continuous image of P.
78
1. BASIC TOPOLOGY
A Hurewicz-type theorem. The space Q is clearly not a Baire space. It is in fact a 'test space' for the Baire property. Theorem 1.9.12. Let X be a space. If X is not a Baire space then X contains a closed subspace homeomorphic to Q. Proof. Let 9 be a countable collection of dense open subsets of X such that ("") S is not dense. There exists a nonempty open subset V C X such that
y n f | S = 0. Then V contains no isolated points. For if x is an isolated point of V then clearly x E V and so, V being open in X, x is an isolated point of X. But then x belongs to every dense subset of X, i.e.,
So by Theorem 1.9.6 it suffices to prove the following: Claim 1. Let Y be a space having no isolated points. If 9 is a countable collection of dense open sets in Y then Y has a countable subspace K having no isolated points such that K \ K C f] 9. Proof. For y € Y, if A is a collection of subsets of Y we say that A converges to y if every neighborhood of y contains all but finitely many A 6 A. It is easy to see that if y e Y and U is a neighborhood of y then there is a pairwise disjoint infinite collection V of nonempty open subsets of U \ {y} such that V converges to y and U V C U (Exercise 1.9.4). Enumerate 9 as {Gn '• n E N}. We will construct a sequence
{Un : n e N} of pairwise disjoint collections of nonempty open sets in Y and a sequence {Kn : n e N} of countable subsets of F, as follows. Let Ui = {Y}. At stage n + 1, where n > 0, for each U e Un, pick an arbitrary element kn(U) e U Pi Gn and choose an infinite pairwise disjoint collection Vn(U) of nonempty open subsets of Y \ {kn(U}} that converges to kn(U), and satisfies (1) \JVn(U) CUnGn.
Let Un+l = \J{Vn(U) :UeUn}
and Kn = {kn(U) : U G Un}. Note that
1.9. HOMOGENEOUS ZERO-DIMENSIONAL SPACES
79
(2) Kn C U Un, U Un+i C |J Un \ Kn and |J U n+ i C Gn. This completes the construction of the desired sequences. Put K = U^°=i Kn- Then K is clearly countable. It remains to show that K has no isolated points and has the property that K \ K C P) 9For each n and U £ Un the subset {kn+i(V) : V G V n+ i([/)} of Kn+i converges to kn(U) since Vn(U) converges to kn(U), and it does not contain kn(U) since kn(U) $ Kn+i. So kn(U] is not an isolated point of K. We will next show that 1C \ K C f| 9. It follows from (2) that (3) For all j G N and U G Uf.
Ur\K C {kj(U)} \J\JV j ( U ) . To see this, consider any j, s G N and C7 € Uj. If s > j then
and so Ks n C7 C U(T/ G U J+i : F C U} = \J Vj(U). Moreover, if s = j then kj(U] eKsnU = {k0(V) :VnUJ:Q} = {kj(U)}. Finally, if s < j then Ks n U C Ks n |J Uj+i C Ks n |J U s+ i = 0. The reason why the proof works is that by (1) and (3) we get: (4) For every J E N and U 6 Uj we have K n U C U. Now consider any x G K \K. We claim that for every j there is an element Uj e Uj which contains x. Since \J Uj C Gj this implies that
which is as required. Let U\ = Y. Next, consider any j ' G N and assume that we already found Uj. Since x £ K but x 7^ kj(Uj) and Vj(Uj) converges to kj(Uj) we conclude by (3) that there is an element UJ+i G Vj(Uj) with
x G c/j+i n /c. Then x G t/j+i because of (4). This completes the construction of the sequence {Un : n G N}. <> So we are done.
D
Corollary 1.9.13. Let X be a space. Then every closed subspace of X is a Baire space if and only if X has no closed subspace homeomorphic to Q. Proof. Necessity is clear. To prove the sufficiency, it is enough to prove that if X is not a Baire space then it contains a closed homeomorph of Q. But this is the statement of Theorem 1.9.12. D
80
1. BASIC TOPOLOGY
Exercises for §1.9. 1. Let A be a subset of R such that A+Q = A. Prove that A is homogeneous. 2. Let
A = { A C R : Q C A , A + Q = A, An
(J
A + mr = 0}.
n£Z\{0}
(1) Prove that A / 0. (2) Let 3C be a chain (under inclusion) of elements of A. Prove that (3) Apply the Kuratowski-Zorn Lemma to conclude that A contains a maximal element, say AQ. (4) Prove that if x £ K. \ AO then there exist n G Z \ {0} and a G AO such that x = a + mr. (5) For each n <E Z \ {0} let An = {x G R \ AO : (3 a <E A 0 )(or = a + mr)}. Prove that the collection {An : n G Z} partitions R. (6) Prove that for every n € Z we have An = An + Q. (7) Prove that R can be decomposed into two homeomorphic homogeneous sets. 3.
(1) Prove that there is a Borel subset E C R such that EnaR\E. (2) Prove that if E is a homogeneous subset of R such that E ~ R \ E then E is not Borel.
4. Let Y be a space without isolated points. Prove that if y £ y and £7 is a neighborhood of y then there is a pairwise disjoint infinite collection V of nonempty open subsets of U\{y} such that V converges to y and (J V C [7. 1.10. Inverse limits Sometimes it is possible to 'approximate' a complicated space 'arbitrarily closely' by less complicated objects. In this section we shall formally define what we mean by this and we shall derive an interesting 'approximation theorem'. On Page 137 we will discuss Freudenthal's Approximation Theorem about approximations of spaces by polyhedra. An inverse sequence is a sequence of pairs (Xn,fn)n of spaces Xn and continuous functions fn: Xn+i —> Xn. The spaces Xn are called coordinate spaces and the mappings fn are called bonding maps. The inverse limit of the inverse sequence (Xn,fn)n, denoted by is defined to be the following subspace of the product of the Xn: oo
G JJ Xn : (Vn e N)(/ n (z n + i) = z n ) . n=l
1.10. INVERSE LIMITS
81
It will sometimes be convenient to let Xoo denote the inverse limit of the sequence (Xn, fn)n. For every n 6 N, the restriction to hjn(JC n ,/ n ) n of the projection onto the n-th factor of the product O^Li Xn shall be denoted by /£°. By definition it follows that for every n, fn o f^_l = f^°. The functions /£° are called the projections of the inverse sequence. In addition, for ra > n the composition fm ° • • • ° fn • Xm+i -> Xn +1
shall be denoted by /™ . Observe that /£+1 = /n. Finally, /™ is defined to be the identity function from Xm to Xm. Lemma 1.10.1. Let (Xn,fn)
be an inverse sequence. Then the collection
is a base for the topology of]nn(Xn,fn)nIn addition, this base is closed under finite unions and finite intersections. Proof. Let U be a relatively open subset of X^. In addition, let x £ U. By the definition of the product topology, there exist N € N and for every i < N an open oen subset Vi C Xi such that N
oo
n
i=l
i=N+l
Since x\ G Vi, x% G V-2 and f\(x-2] = a?i, it follows that Continuing in this way, we obtain that AT
An easy verification shows that which is as required. The second part of the lemma follows easily from the fact that m
foo _
foo
/ n Jm Jn for all m > n. Simply observe that a basic open subset of .Y^ that 'depends' on the n-ih element of the inverse sequence, i.e., is of the form
for certain open U C X, also 'depends' on the m-th element of the sequence for every m > n. This is true since for every m > n and U C Xn we have
82
1. BASIC TOPOLOGY
So for a finite collection of basic open sets U in Xoo it is possible to identify an index TV € N such that every U £ U 'depends' on XN. D We will now present a few illuminating examples of inverse sequences and their limits. Example 1.10.2. Consider the inverse sequence (A r n ,/ n ) n , where Xn = X and fn is the identity on X for every n. This is a trivial inverse sequence. A moments reflection shows that its inverse limit is (homeomorphic to) X. The next example shows that inverse limits are generalizations of products (see the exercises for details). Example 1.10.3. For every n let Xn be a space. The sequence n
I I -<*-rm
ra=l
where rn : Om^i Xm —> 0™=! Xm is the projection, is an inverse sequence with the infinite product
71=1
as its inverse limit. Example 1.10.4. Let X be a space and for every n let Xn be a subspace of X such that Xn+i C Xn. Form an inverse sequence (Xn,in)n by letting in be the inclusion Xn+i °->- Xn. Then \Ym.(Xn,in}n is homeomorphic tofT=l*n.
From this example we see that an inverse limit of nonempty spaces can be empty. Since the intersection of a decreasing sequence of nonempty compact spaces is nonempty, for compact spaces one cannot get such an example (see Lemma 1.10.10(2) below). Example 1.10.5. Let X be a compact space. By Corollary A. 4. 4 we may assume that X is a subspace of Q. For every n let Xn = rn[X], n
where rn : Q —> J is the projection. In addition, let fn : Xn+i —>• Xn be defined by fn(Xi,...,Xn+i]
A trivial verification shows that
= (xi,...,Xn). A Hm(A n , fn)n = X .
There are several simple but useful observations that we have to make about inverse limits. First, let (Xn, fn)n be an inverse system, and let A C N be infinite. Write A as {mi, . . . , m n , . . . } in such a way that
mi < m-2 < • • • < mn < • • • .
1.10. INVERSE LIMITS
83
The sequence (Xmi, fm*+l)i is also an inverse sequence. By abuse of notation it will be denoted by (A A n ,/ n ) n e ^. It is called an inverse subsequence Of
(Xn,fn)n.
Lemma 1.10.6. If (Xn, fn}n^A is an inverse subsequence o f ( X n , f n ) n the limits \im(Xn, fn}n^A and ]im(Xn, fn)n are homeomorphic. Proof. The function (p: \im(Xn,fn)n
then
—>• Hm(A"n, fn}n^A defined by
is easily seen to be a homeomorphism.
D
So passing to a subsequence makes no difference in the limit (as with ordinary limits). Lemma 1.10.7. Let (Xn,fn)n and (Yn,gn}n be inverse sequences. Suppose that for every n G N there is a continuous function rn : Xn —>• Yn such that the diagram Xn 4- X
V JLn
g " V *•'J-n+l
commutes. Then the function (p: lim(Xn, fn)n -> lim(Yn,gn)n defined by <{>((zn)n) = (rn(xn)}n is continuous. Proof. This is clear since (p is the restriction to hm(J^ n ,/ n ) n of a product map from f]£Li Xn ->> H^Li Yn (Exercise A. 1.13). D These simple results will sometimes be used in the following situation. Suppose that we are inductively defining an inverse sequence (Xn, fn)n and that we would like to construct it in such a way that Hm(X n ,/ n ) n admits a selfmap (p with certain special properties. This is sometimes achieved by constructing for every n a function (pn : Xn+i -> Xn such that the diagram
commutes. By Lemma 1.10.6 it follows that the inverse sequences (Xn,fn)n and (Xn,fn}n>2 have the same limit, and Lemma 1.10.7 shows that the functions (pn induce a canonical selfmap of that limit.
84
1. BASIC TOPOLOGY
Lemma 1.10.8. Let (Xn: fn)n be an inverse sequence with surjective bonding maps. Then the projections of (Xn, fn)n are also surjective. Proof. Fix n G N and take an arbitrary x 6 Xn. Since fn is surjective, there exists xn+i G Xn+i with fn(xn+i) = x. Similarly, there exists x n+2 6 Xn+2 with / n+1 (x n+2 ) = xn+i, etc. Define
etc. The sequence (xn)n 6 O^Li Xn belongs to hm(.Y n ,/ n ) n and since f°°((r } } — ^n r — ^i r Jn \\^n)n)
we are done.
D
Remark 1.10.9. Let (Xn,fn)n be an inverse sequence. For every i define the subspace X(i) of fl^Li Xn by oo
X(i) = {x 6 J| Xn : (Vn < i)(fn(xn+l)
= x n )}.
n=l
It is clear that X(i) is closed in H^Li Xn and nonempty and that X(i + l) CX(i)
for every i, and, finally, that
1=1
So ^oo is a closed subspace of H^Li Xn. This shows that if IP is a topological property which is 'closed hereditary and (countably) productive' then X^ has O5 if all Xn have O3. This simple observation allows us to prove many properties of inverse limits. Compactness and completeness properties of inverse limits. It is natural to ask when an inverse limit is compact or topologically complete. Lemma 1.10.10. (1) The inverse limit of an inverse sequence of topologically complete spaces is topologically complete. (2) The inverse limit of an inverse sequence of nonempty compact spaces is nonempty and compact. Proof. Observe that compactness and completeness are closed hereditary and productive (Lemma A.6.2). So we are done by Remark 1.10.9. D
1.10. INVERSE LIMITS
85
Connectivity properties of inverse limits. Detecting connectivity in topological spaces is important. For products this is easy: a product is connected if and only if all factors are. But inverse limits do not share this property with products, as the following example shows. Example 1.10.11. A disconnected inverse limit of connected spaces. Let {qn : n 6 N} be an enumeration of the rational numbers in I. For every n G N define 00
Xn = ( I x { 0 , l } ) U (J ({ m }xl). m=n
Then X\ D X% D • • • D Xn ! > • • • , every Xn is connected and
n-l
is disconnected. So we are done by Example 1.10.4. Observe that the inverse limit in this example is compact. So compactness of the inverse limit is not enough to guarantee connectivity. If all spaces involved are continua, then there are no problems. Lemma 1.10.12. The inverse limit of an inverse sequence of continua is a continuum. Proof. Let (Xn, fn)n be an inverse sequence of continua. Let the spaces X ( i ) for every i be defined as in Remark 1.10.9. We claim that every X ( i ) is a continuum. Indeed, fix i and define a function (p: f l l i ^-j ~^ -^"(0 by Then ip is evidently a continuous surjection. It therefore follows that X ( i ) is a continuum since Hyl; ^-j ^ s It now suffices to apply Proposition A. 10. 7. D The dyadic solenoid. As an example of the power of inverse limits, we construct a very peculiar space here. Consider the following inverse system gi < f gi , f ... , f gi , f where / : S1 —> S1 is the function f(z} = z2 (complex multiplication). The inverse limit of this system is denoted by £2 and is called the dyadic solenoid. We list a few properties of £2 that we can derive easily from previous results. First, it is a continuum by Lemma 1.10.12. In addition, since S1 is a topological group under complex multiplication, and / is a continuous homomorphism, it follows easily that £2 is a closed subgroup of the product group (S1)00. So £2 is a topological group and hence a homogeneous space.
86
1. BASIC TOPOLOGY
A continuum X is called indecomposable if it is not the union of two proper subcontinua. So I is not indecomposable (= decomposable) since it is the union of [0, 1/2] and [l/2,1]. Lemma 1.10.13. Let X be a continuum. The following statements are equivalent: (1) X is indecomposable, (2) every proper subcontinuum of X is nowhere dense. Proof. That (2) =£> (1) is clear since no space is the union of two nowhere dense subsets. For (1) => (2), suppose that X contains a proper subcontinuum C with nonempty interior U. Let £ be the collection of components ofY = X\U. Then every E 6 £ meets U \ U by Corollary A.10.5. So every E e £ meets C but misses C7. If |£| = 1 then clearly X is the union of two proper subcontinua. Suppose therefore that £ has more than one element. Then Y contains a proper nonempty relatively clopen subspace W by Proposition A.10.3(1). Every member from £ either is contained in W or is disjoint from it. So W is the union of a family of subcontinua of Y which all meet C. So C U W is a proper subcontinuum of X. Similarly, C\J(Y\ W) is a proper subcontinuum of X. Since they cover X, we arrive at a contradiction. D Corollary 1.10.14. A nondegenerate locally connected continuum is decomposable. Proof. Let x and y be distinct points in X and let U be a connected neighborhood of x whose closure misses y. Then L7" is a proper subcontinuum of X which is not nowhere dense. So X is decomposable by Lemma 1.10.13. D Theorem 1.10.15. The dyadic solenoid is indecomposable. Proof. Let C C S2 be a proper subcontinuum. For every n put Cn — f%°[C]. Since C is proper there exists n E N such that Cn is a proper subset of S1. Observe that Cn is a subinterval of S1. Now f ~ l [ C n ] is the union of two pairwise disjoint nonempty subintervals of S1 of arc-length exactly one half the arc-length of Cn. In addition, it contains C n +i. By connectivity, Cn+i is contained in one of them. We conclude that the arc-length of Cn+i is less than or equal to half of the arc-length of Cn. So for large N the arc-length of CN is arbitrarily small. Now assume that C has nonempty interior. Then there are i E N and a nonempty open interval U C S1 such that (fi°)~l[U] C C. But the arclength of f~l[U] is the same as the arc-length of U, etc. So the arc-length of CN does not tend to 0 for large N. This is a contradiction.
1.10. INVERSE LIMITS
87
So we conclude that C has empty interior. Since C was arbitrary, we conclude that £ 2 is indecomposable by Lemma 1.10.13. D Remark 1.10.16. It is easy to show that the dyadic solenoid contains an arc. For every n put In = {ei7Vt :0
Then gn = f \ I n +i —»• I n is a homeomorphisrn. The inverse limit of the inverse sequence (I n?
C=\J C(B) B£-B is equal to the composant E of x. It is clear that C is a subset of E. Now if y 6 E then there is a proper subcontinuum K of X such that both x and y belong to K. Since K is proper, there is an element B e 23 which misses K. But then K C C(B) and hence y G C(B] C C. So E is the union of countably many proper subcontinua of X. An application of Lemma 1.10.13 therefore yields the desired result. D A continuum Z with Z2 w Z but Z°° $ Z. As a second example of the power of inverse limits, we construct another very peculiar space here. Most of the known examples of compact spaces X such that X2 w X (which implies that Xn w X for every n] have the property that X00 w X. Think of the Cantor set and the Hilbert cube as examples of such spaces. We will use inverse limits to demonstrate that this is not always the case. Theorem 1.10.18. Let Y be a locally compact space homeomorphic to its own square. Then Y has a compactification homeomorphic to its own square.
88
1. BASIC TOPOLOGY
Moreover, if the one-point compactification of Y is a continuum then this compactification can be chosen to be a continuum as well. Proof. The compactification X of Y is the inverse limit of a suitable inverse sequence (anY, fn}n of compactifications anY of Y, where for each n, fn is a continuous function from an+\Y —> anY which restricts to the identity on Y. This implies that f[an+iY \Y] = anY \ Y by Exercise A.4.2. In order to ensure that X2 w X we will construct the inverse sequence in such a way that for each n E N there is a homeomorphism hn: an+iY —>• anY x anY so that the diagram h
a2Y
<-^—
a3Y
<
•••
a2Y x a2Y i •• • commutes. Let a\Y be the one-point compactification of Y and let h be any homeomorphism from Y onto Y2. There is a compactification a2Y of y such that h extends to a homeomorphism hi: a2Y —> aiY x a\Y. Since a\Y is the smallest compactification of Y, the identity mapping on Y extends to a mapping /i: a2Y -» oiF. Similarly, there is a compactification a^Y of y such that h extends to a homeomorphism h2: a3Y —> a2Y x a2Y. Define /2 : a3Y ->• a 2 y
by
f2 = hl1 o(fi x fi)oh2. It is easily seen that /2 f y is the identity on Y. In the same way define the compactification a^Y', and the functions /ts and /s, etc. Since the hn's are homeomorphisms, it follows from Lemma 1.10.7 applied twice and Lemma 1.10.10(2) that X = Lirr^a^y, / n ) n is a compactification of Y homeomorphic to its own square (cf. Exercise 1.10.9). Observe that if aiY is a continuuum then so are all the other a n y's. As a consequence, X is a continuum as well by Lemma 1.10.12. D Now consider the Cantor set C, and let B be the topological sum of NX C and a one-point space. Clearly, B2 = ((N x C) © {pt}) x ((N x C) 0 {pt}) w (N2 x C 2 ) © (N x C) 0 (N x C) © {pt} « (N x C) 0 {pt} w JB. Put Y = B x [0,1) x Q.
Lemma 1.10.19. Y2 w y.
1.10. INVERSE LIMITS
89
Proof. Simply observe that F2 = (B x [0,1) x Q) x (B x [0,1) x Q)
&(BxB)x ([0,1) x [ 0 , l ) ) x ( Q x Q ) « £ x ([0,1) x I ) x Q w 5 x [0,1) x Q
= y. So we are done.
D
Let Z be the compactification of Y we get from Lemma 1.10.19. Observe that the one-point compactification of Y is connected since it is the cone over B x Q. From this it follows that Z is a continuum as well. Lemma 1.10.20. Z ^ Z°°. Proof. Striving for a contradiction, suppose that h is a homeomorphism from Z onto Z°°. Since Y, being locally compact, is open in Z, h[Y] is open in Z°°. In addition, since B has an isolated point, Y contains an open copy of [0.1) x Q. Therefore, Z°° contains an open copy of [0,1) x Q, say F. Since F is open and nonempty, it contains a basic open subset. So there are an integer n and for each i < n a nonempty open set Ui C Z so that Ui x U-2 x • • • x Un x Z x Z x • • • C F.
Hence Z is (homeomorphic to) a retract of some open subset of F, or, equivalently, Z is homeomorphic to a retract of some open subset of [0,1) x Q. Since [0,1) x Q is locally connected, this implies that Z is locally connected (Exercise A.12.4). Its open subspace Y is therefore locally connected as well. But this implies that C is locally connected, which is a contradiction. D Remark 1.10.21. In view of the existence of a compact space homeomorphic to its own square but not to its countable infinite product, the question naturally arises whether there is a compact space X which can be mapped onto X2 but not onto X00. Interestingly, the answer to this question is in the negative (see Exercise 1.10.8). Brown's Approximation Theorem. Before we are in a position to formulate and prove an interesting 'approximation theorem' for inverse sequences, we have to fix some notation. Let (Xn,fn)n be an inverse sequence with inverse limit X^. It is sometimes useful to have an explicit admissible metric for XQQ . For every n £ N let dn be an admissible metric for Xn which is bounded by 1 (Exercise A. 1.6). The formula oo
g(x,y) = ^1"ndn(xn,yn) n=l
90
1. BASIC TOPOLOGY
defines an admissible metric for the product of the Xn (Exercise A. 1.10), and hence also for XOQ. Observe that g is also bounded by 1. Lemma 1.10.22. Let (Xn,fn)n be an inverse sequence. In addition, for every n let dn be an admissible metric for Xn which is bounded by I. Then with respect to the metric Q for XOQ — hjn^n,
fn)n
defined above, we have that every /£° : Xoo —> Xn is a 2~(n~^ -mapping. Proof. This is easy. Fix n £ N and let p, q e ^oo be such that
fn(p) = fn(o)Observe that pi — qi for every i < n. Consequently,
We conclude that /£° is a 2~( n ~ 1 )-mapping.
D
Brown's Approximation Theorem 1.10.23. Let(Xn,fn)n be an inverse sequence of compact spaces with inverse limit X^,. If each fn is a near homeomorphism, then so is each f£°. In particular, if each fn is a near homeomorphism then Xoo is homeomorphic to X\ (and hence to every Xn). Proof. We shall prove that /f0 : Xoo —>• X\ is a near homeomorphism. For each n, let dn be an admissible metric for Xn which is bounded by 1 and let Q be the metric for X^ introduced above. Let E > 0. Inductively, we shall construct for every n G N a homeomorphism hn : Xn+i —>• Xn such that the functions gn : X^ —> X\ defined by gn = hi o • • • o hn o f^_l have the following properties: (1) di(5i,/r)< e / 2 , (2) dl(gn,gn+l) <3~n-e, (3) di(gn,gn+1)<3-n-inm{di(gi,Si/.(X00,Xl))
:l
Since j\ is a near homeomorphism and /f° = j\ o /^°, it is clear that h\ can be constructed so that condition (1) is met. Suppose that the homeomorphisms hi are defined for 1 < i < n. Observe that for every 1 < i < n, the function gi 6 S(Xoo, Xi) \ 9 I A (^00,^1) (Lemmas 1.10.8 and 1.10.22). Let li
n
6 = 3- - mm {e, min{di (^,9^(^00,^1) and notice that 6 > 0. Since gn = hlo---ohno f™+1 =h1o---ohno fn+1
1.10. INVERSE LIMITS
91
and fn+i is a near homeomorphism, we can find a homeomorphism hn+i '• Xn+2 —> Xn+i such that g(hi o---ohno fn+l o f™+2,hi o---ohno hn+i o /£j.2) < 8 (use the fact that hi o • • • o hn is uniformly continuous (Exercise A. 5. 18)). It is clear that hn+i is as required, which completes the inductive construction of the hn. By (2) it follows that the sequence (gn)n is
exists. Since gn £ ^ ( X o c ^ X i ) for every n, Proposition 1.3.8 implies that Observe that (1) and (2) imply that di(/i°,0)<e. Also, (3), Lemmas 1.6.1 and 1.3.10 show that
n=l
which is as required.
D
Dynamical systems. A dynamical system is a pair (X, f) where X is a space and /: X —>• X is continuous. In topological dynamics one studies the behaviour of the iterates of /. Inverse limits play a role in certain considerations there. We will not go into this, but we remark that the following considerations that will play a role at one point in our chapter about dimension theory were motivated by well-known ideas in topological dynamics. Let (X, /) be a dynamical system, and consider the following inverse sequence
X < f X < f ••• < f X < f with inverse limit X^. So the same map is repeated infinitely often in the inverse system under consideration. Define the so-called shift a: X00 -» X00 as follows: cr(xi,x2,...) = (/ (zi), £1,2:2,... )• It is clear that a is well-defined. The compositions / o TTi 7Tn_i
(n — 1), ) > ^ 1), i\ (n
are all continuous from which it follows that <j is continuous. It is in general not true that a is a homeomorphism.
92
1. BASIC TOPOLOGY
Define another shift function r: X00 —> X00 by T(>i,Z2,...) = (ar2,X3,...)It is clear that r is continuous as well. A moments reflection shows that <7[Xoo] C Xoo,
r[Xoo] C Xo
Moreover, if x £ A^ then cr(r(x)) = (7(2:2, £ Similarly, r(cr(;r)) = We conclude that a \ X^ is a homeomorphism, with inverse r \ X^. By abuse of notation, we will denote a \ X^ also by a. Lemma 1.10.24. Let (AT, /) be a dynamical system, and let a: X^ —>• Xc denote the shift homeomorphism. Then for each n the diagram Y
A.
' n Y \- .A no
,00
Y <'" Y AOO A
commutes. As a consequence, if f is fixed-point free then so is a. Proof. This is clear since if x e X^ then / o f£°(x) = f ( x n ) and which is equal to f ( x \ ) if n = I and x n _i — f ( x n ) if n > 1. This implies that if x is a fixed point of a then f£°(x) is a fixed point of /, which establishes the second statement. D So in some vague sense, every continuous function /: X —>• X can be 'approximated' by a homeomorphism (cf. Lemma 1.10.22). Exercises for §1.10. A space X is called isotopically homogeneous provided that for every x, y £ X there exists an isotopy H : X x I —>• X such that HO = lx and HI(X) = y. 1. Let A" be the inverse limit of an inverse sequence of copies of I. Prove that X has the fixed-point property. 2. Prove that the circle S1 is not homeomorphic to the inverse limit of an inverse sequence of copies of I. 3. Prove that the statement in Example 1.10.3 is correct.
1.10. INVERSE LIMITS
93
4. Prove that the statement in Example 1.10.4 is correct. 5. Let (X n , fn) be an inverse system with open surjective (continuous) functions. Prove that for every n € N the function is open. 6. Let (Xn,fn)n be an inverse sequence such that every fn is a homeomorphism. Prove that hm(X n ,/ n ) n is homeomorphic to X\ (and hence to every Xn). 7. Let (Xn, fn)n be an inverse sequence. In addition, for every n let An C Xn be such that fn[An+l] C An; put gn — fn f A n +i : An+i —> An . Prove that the inverse limit and the subspace (x £ hm(Xn,/n)n : (Vn G N)(x n G A n )} of Um(Xn, fn)n are homeomorphic. ^•8. Assume that X is a compact space. Prove that X maps onto X2 if and only if it maps onto X00 . 9. Let (anX,fn)n be an inverse sequence consisting of compactifications of the space X. Assume moreover that for every n the function fn restricts to the identity on X. Hence aiX < a2X <•••
g(y) = (g^(y)}n is well-defined and continuous. 11. Let X be a space and let
Ji C ^2 C • • • C yn C • • • be an increasing sequence of Wallman bases of X. (1) Prove that for every n there is a continuous function which extends the identity on X. (2) Prove that 3 = (Jn°=1 3n is a Wallman base of X. (3) Prove that the inverse limit of the sequence (a; (X,3~n ),/n)n and u; (X", 9") are equivalent compactifications of X.
94
1. BASIC TOPOLOGY
12. Let (Xn, fn)n be an inverse sequence with compact spaces. Prove that if A and B are disjoint closed subsets of X^ then there exists an N E N such that for each n > N the sets /~[A] and f%°[B] are disjoint. 13. Let X and Y be compact spaces and let /: X —>• Y and g: Y —> X be continuous functions such that both g o f : X —> X and f o g : Y —>• Y are near homeomorphisms. Prove that X and Y are homeomorphic. 14. Prove that the statement in Example 1.10.5 is correct. 15. Prove that in the statement in Remark 1.10.5 the compactness assumption is essential. 16. Prove that an indecomposable continuum is not path-connected. 17. Give an example of a homogeneous continuum which is not isotopically homogeneous. 18. Let X be a compact space and let /: X —> X be continuous. Prove that
K=f\f[X\ i=l
is nonempty and also that f[K] = K. 19. Let X be an indecomposable continuum and let E be one of its composants. Prove that if E is not dense then E is closed. 20. Let X be an indecomposable continuum and let E and F be distinct composants of X. Prove that E and F are disjoint. 21. Prove that for a continuum X the following statements are equivalent: (1) X is hereditarily indecomposable. (2) If C, D are arbitrary subcontinua of X such that C n D ^ 0 then C C D or D C C. 22. Let X be a hereditarily indecomposable continuum, and let /: X —> Y be monotone. Prove that Y is a hereditarily indecomposable continuum. 23. Let X be an indecomposable continuum. Prove that the family of composants of X is uncountable. It can be shown that the dyadic solenoid £2 cannot be imbedded in the plane. It is an interesting question whether there is a planar indecomposable continuum. This question will be considered in the remaining part of this section. Let o, b and c be three distinct points in the plane. Inductively, construct finite simple chains Ci, 62, • • • of open balls having among other things the following properties: (1) mesh(e ri ) < 2 ~ n , (2) Cn < C n _ i . But these are not all conditions. We require the C n 's to follow a certain pattern. The collection Ci is a simple chain from a to c through b, 63 is a simple chain from a to b through c, 63 is a simple chain from b to c through a (see Figure 5). Then repeat this process at infinitum. For every n let Cn = |J Cn and let P = H^Li Cn-
1.11. HYPERSPACES
95
24. Prove that P is a continuum. 25. Prove that F is indecomposable.
Figure 5. 1.11. Hyperspaces In this section we shall introduce the hyperspace of a topological space. Hyperspaces are used later to construct various interesting examples. Let X be a space. The collection of all non-empty compact subsets of X shall be denoted by 1X . We will endow 1X with a suitable topology. For E,F e 2X define the Hausdorff distance Qn(E,F} between E and F by
QH(E,F] = infje > 0 : E C B(F,e) and F C B(E,e)}. The compactness of E and F implies that gn(E,F) e [0, oo) (Exercise 1.11.1). Lemma 1.11.1. g# : 2X x 2 X —> [0, oo) is a metric. Proof. We shall only present a proof of the triangle inequality, the other verifications being trivial. Take E, F, G e 2 X , let e > 0 and put 6 = £/2. By definition we have (*)
E C B(F,6 + QH(E,F}) and F C B(G, 6 + QH(F,G}}.
Take an arbitrary point x G E. By (*) there exists y E F such that
Analogously there exists z G G such that e(y,z
We conclude that Q(X,Z)
<e
96
1. BASIC TOPOLOGY
Therefore, since x was an arbitrary point of E, E C B(G,e + QH(E,F] + QH(F,G)). Similarly, G C B(E,e + QH(E,F] + QH( Since e was an arbitrary positive number, we obtain as required.
D
The metric QH is called the Hausdorff metric on 2X . We endow 2X with the topology derived from this metric and call it the hyperspace of X. Define e: X —»• 2X by e(x] — {x}. It is easily seen that e is an isometry, hence an imbedding, and that e[X] is a closed subset of 2X (Exercise 1.11.3). It will sometimes be convenient to identify X and e[X], Compactness and completeness properties of hyperspaces. We
will show that compactness and completeness are 'preserved' by the hyperspace operator. We need a simple lemma first. Lemma 1.11.2. Let (X, Q) be a space and (An)n a Cauchy sequence in 2X . LetA= (J~=i^n. Then (1) The metric p = Q \ (A x A) is totally bounded. (2) If AOO — linin-^oo An exists in 2X then it is equal to (f) {x eX : ( V n G N)(3x n e A n )(lim n _, 00 a; n = x)}, and also to (+)
\\n-l U m > n ^ T i -
(3) If AOO — linin^oo An exists in 2X then A is compact and oo
A = Avo U |J An. n=l
(4) If Q is complete then limn-^ An exists in 2X (and hence has to be equal to the sets (f) and (J) in (2)J. Proof. For (1), let e > 0. There exists A" e N such that QH(An,Am) < l/±e for all n,m > N. Consider AN and let -F C AN be a finite 1/4e-net (by Theorem A. 6.1, g\(Apf x AN) is totally bounded, whence the desired net exists by Exercise A. 6.2). Fix n > N. Since QH(AN,An) < l/±e, An C B(AN, 1/4£:)So for every x e An there exists y £ AN such that Q(X, y) < V4e. For such a y there exists z e F such that Q(z,y] < l/4e. We conclude that F is a y2e-net for An U AN- Since n was arbitrary, F is even a l/2e-net for \Jn>N An, and hence an e-net for \Jn>N An- Since \Jn=i An is compact, it has a finite e-net
1.11. HYPERSPACES
97
by the same reason AN has one. The union of these two finite e-nets is a finite e-net for A. So p is totally bounded by Exercise A.6.2. For (2), let B = [x G X : (Vn G N)(3xn G ^(lim^^ xn = x)} and C = n^Li Um>n Am, respectively. We will prove that AOO CB C C C AOO. Pick an arbitrary x G AQO. For each n G N there exists a point xn G An such that ^(x,x n ) < 2g(x,An). We claim that the sequence (xn)n converges to x. To this end, let e > 0. Since (An)n converges to A^, there exists TV G N such that for each m > N, A^ C B(Am, l/<2^}- From this it follows that for m > N we have ^(x, A m ) < l/2£, i.e., 0(:r,:rm) < e. Hence AOQ C 5. That B C C is trivial. Striving for a contradiction, assume that there exists a point x G C\Aoo. Let e = g(x,AOQ) and observe that e > 0. Since A n —>• AOO in 2 X , there exists N G N such that for every n > N we have An C 5(Aoo, V2£)- As a consequence, (^ 4 XV G (J A n
(
which is a contradiction. For (3), let U be a cover of A* = AOO U (J^li ^™ by open subsets of X. Since AOO is compact, there is a finite y C U such that AOO C y 3. By Corollary A.5.4 there exists e > 0 such that B(A00^e) C y J. Since (A n ) n converges to AQQ, there exists TV G N such that An C 5(Aoo,£) C y ^ for n > TV, and so n>N
Since 9" is finite and Un_T1 An is compact, we conclude that A* can be covered by finitely many elements of U. So A* is compact, and since ^4^ C A by (2), it follows that A* C A. But since (J^Li An is dense in A and is contained in A*, it also follows that A C A*. For (4), we first observe that by (1) the metric p — Q \ (A x A] is totally bounded. Since A is closed, p is also complete. So A is compact by Theorem A.6.1. Put K = H^Li U m > n ^ ™ - Then K is the intersection of a decreasing sequence of nonempty closed subsets of the compact space A, and hence is nonempty and compact itself. Hence K G 2 X . To prove that An ->• K in 2 X , let £ > 0. By the compactness of K, there exists mi G N such that (5)
AnCB(K,e).
98
1. BASIC TOPOLOGY
Since (An)n is Cauchy, there exists ra2 E N such that QH(An,Am) < £/2 for n, ra > 7712- As a consequence, for n > rri2,
(6)
^ C |J ATO TT1>TO2
From (5) and (6) it now follows that @H(K,An) < e for n > max(rai,m2), which is as desired. D Theorem 1.11.3. Let X be a space with admissible metric Q. Then (1) If Q is complete then so is QH(2) If Q is totally bounded then so is QHProof. Observe that (1) follows from Lemma 1.11.2(4). For (2), assume that Q is totally bounded. Let E > 0 and let y be a finite open cover of X with nonempty sets of diameter less than e. For every F e 3 pick XF G F. For A e 2X, put BA = {xF : F n A ^ 0 } . Since J covers X, A C \J{F e y : F n A 7^ 0}, from which it easily follows that QH(A,BA) < £• Since 3* is finite, this shows that 2X can be covered by finitely many er-balls. D Corollary 1.11.4. If X is topologically complete then so is 2X . If X is compact then so is 2X . Proof. Both statements follow from Theorems 1.11.3 and A. 6.1.
D
Hyperspace maps. Let X and Y be spaces and let /: X —> Y be continuous. By Exercise 1.11.7 we may regard X and Y to be (canonical) subspaces of 2X and 2 y , respectively. There exists a canonical continuous extension 2^ of / over 2X , which is called the hyperspace map of /. Indeed, define 2-^ : 2X —>• 2Y in the obvious way by We shall prove below that hyperspace maps are continuous. They are of crucial importance in the process of analyzing hyperspaces. Lemma 1.11.5. Let X and Y be spaces and let /: X —>• Y be continuous. Then the hyperspace map 2? : 2X —> 2Y is continuous. Proof. Suppose that (An)n is a sequence in 2X converging to A 6 2X . Then obviously (An)n is Cauchy in 2X , and hence A — (J^Li An is compact by Lemma 1.11.2(3). As a consequence / \ A is uniformly continuous (Exercise A. 5. 18). Now let e > 0 and pick S > 0 such that if x, y £ A are arbitrary points with g(x,y) < 6 then 0(/(x),/(y)) < e. If D,E C A are nonempty
1.11. HYPERSPACES
and compact and 0H(D,E) < 6 then gH(f[D],f[E]) in 2Y.
99
< e. So, f[An] ->• f[A] D
The union operator. We shall now consider a more complicated 'hyperspacemap'. Let X be a space. Consider the space 22 , i.e., the hyperspace of the hyperspace of X. Each point A G 22 is a compact subset of 2X and therefore can be regarded as a 'compact' family of compact subsets of X. We shall prove that the union of such a family is compact. From this it follows that the 'union-operator' defined by \J(A)=\jA, is well-defined. We shall also prove that this operator is continuous. Lemma 1.11.6. Let X be a space and let A belong to 22 . Then \JA is a compact subset of X. Proof. Take a sequence (xn)n in \JA. We shall prove that it has a convergent subsequence in \JA. For each n there exists An G A such that xn e An. By compactness of the subspace A of 2X , we may assume without loss of generality that the sequence (An)n converges to a point A 6 A. From Lemma 1.11.2(3) it follows that B — A U U^=1 An is compact. Hence (xn)n has a convergent subsequence in B (Theorem A. 5.1). Since B C (J.A, it follows that this convergent subsequence is convergent in \JA. Now apply Theorem A. 5.1. D So from Lemma 1.11.6 we conclude that for every Jf , the union-operator
is well-defined. We shall now prove that it is continuous. Proposition 1.11.7. Let X be a space. Then the union-operator |J:2 2 "^2 X is Lipschitz and hence is continuous. Proof. Let QHH denote the Hausdorff metric on 22 . Take arbitrary points £ and y in 22 and let QHH (£)3~) — £• Fix £ > 0 and pick an arbitrary element x G U £. There exists E G £ such that x G E. Since E G £ there exists F G 3 such that QH(E, F) < 6 + e. This implies that E C B(F, 5 + e) and so there exists y G F C \J y such that Q(X, y) < 6 + e. We conclude that
100
1. BASIC TOPOLOGY
By interchanging the roles of £ and 3 we therefore conclude that
Since e was chosen arbitrarily, we get £>#(U £, (J IF) < $, as required.
D
Corollary 1.11.8. Let X andY be spaces and let g : X —> 2Y be continuous. Then the function g:2x^2Y defined by
g(A) = |J g(x) x£A
is continuous. Proof. The hyperspace map 29 : 2X —> 22 is continuous by Lemma 1.11.5. In addition, by Proposition 1.11.7, the union-operator (J : 22 —> 2Y is also continuous. Since g — [J o 25, we are done. D Let X be a space and for each n € N define yn(X] = {A e 2X : \A\ < n}.
Observe that ^(X) C ?2(X) C • • • C 3n(X) C • • • . The following simple consequence of Exercise 1.11.6, is basic for hyperspaces. Theorem 1.11.9. Let X be a space containing more than one point and let n € N. Then the function fn : Xn -»• 3n(X) denned by
is continuous. Corollary 1.11.10. Let X be a space and let AI, . . . ,An be nonempty subsets of2x. Then the function f : HlLi ^ ~> 2^ defined by f(A1,...,An)=A1\J---UAn is continuous. Proof. Again, let \J: 22 —> 2X denote the union-operator. By Theorem 1.11.9 we get that the function g: HlLi ^ ~* ^ defined by
is continuous. Since / = \J og, we are done.
D
1.11. HYPERSPACES
101
A canonical base for 1X . We used the Hausdorff distance to endow 2X with a natural topology. There is a different way to do this in terms of a canonical base for 2X yielding the so-called Vietoris topology on 2X . We will show that both topologies on 2X are the same. The alternative description of the topology on 2X in terms of a canonical base is sometimes very convenient. Let X be a space. For every finite collection A of subsets of X put (A) = {B e2x :B C\jA and for every A G A, A n B ^ 0}. Lemma 1.11.11. Let X be a space. Then (1) if A is a finite family of open subsets of X then (A} is open in 2X , (2) the collection 'B(X) = {(3) : 3 is a finite family of open subsets of X} is an open base for 2X which is closed under finite intersections. Proof. For (1), let .A be a finite family of open subsets of X and let D G (A}. For each A G A pick an arbitrary point XA £ An D. Since A is finite and consists of open sets, and D is compact, there exists by Corollary A. 5. 4 an e > 0 such that (3) for every A e A, B(xA,e) C A, (4) Now take E G 2X such that QH(D,E] < e. By (4), E C \JA. Fix an arbitrary element A G A. Since D C B(E,e), there exists an x G E such that Q(XA,X) < e. By (3), x belongs to A, i.e., E intersects A. This proves that the ball (in 2X) about D of radius e is contained in (.A). For the first part of (2), take arbitrary D G 2X and e > 0. By compactness of D, there is a finite family A of open subsets of X such that (5) D C\JACB(D,e}, (6) for every A&A,AnD^$, (7) for every A G A, diam(A) < e. So D G (A). Observe that (A) is open by (1). We claim that (.A) is contained in the e-ball in 2X about D. To this end, take an arbitrary E G (.A). By (5), we need only show that D C B(E, e). But this is trivial. For take an arbitrary element x G D. By (5) there exists A G A containing x. Since E G (.A) we have EnA 7^ 0, which implies by (7) that there exists x' G E with g(x, x') < e. We conclude that x G B(E,e). Finally, let fo = {Ui, . . . , Un} and 3\ = {Vi, . . . , Vm} be arbitrary collections open subsets of X. Put U = (J^o and V — U ^ i * respectively. An easy verification shows that
W n (9-1) - ({u n vlt . . . , u n vm, Ui n v, . . . , un n V}),
102
1. BASIC TOPOLOGY
as required.
D
Corollary 1.11.12. Let X be a space. Then
= |J n—l X
is dense in 2 . Connectivity in hyperspaces. We now turn to connectivity properties of 2*. Proposition 1.11.13. Let X be a space. Then (1) X is connected iff2x is connected, (2) X is locally connected iff 2X is locally connected. Proof. We prove (1). First suppose that X is connected. Since y^X) is dense in 2X (Corollary 1.11.12), it suffices to prove that it is connected, and for this it suffices to prove that every Jn(X) is connected. However, this follows directly by Theorem 1.11.9 since for each n, yn(X) is the image of Xn under the map fn : Xn -»• 2X denned by Conversely, assume that 2X is connected. We shall prove that X is connected. Striving for a contradiction, assume that X is not connected. So it is possible to write X as UuV, where U and V are disjoint nonempty open sets. Clearly,
In addition, ({U}},({V}}, and ({U, V}) are pair wise disjoint, nonempty and open by Lemma 1.11.11(1). This contradicts the connectivity of 2X . We prove (2).
Now assume that X is locally connected. Take A e 2X and let U be a neighborhood of A in 2X . By Lemma 1.11.11(2) there exists a finite family V of open subsets of X such that A £ (V) C U. Since X is locally connected, every component of every V e V is open (Exercise A. 2. 8). By compactness of A we may therefore assume without loss of generality that V consists of connected open sets. We claim that (V) is connected. First observe that for every V G V we may identify 2V and the subspace {A e 2X : A C V} of 2X (Exercise 1.11.7). Let V = {Vi, . . . , Vn}. Now define the function
i-l
1.11. HYPERSPACES
by
103
n
g(A1:...,An)= \jAi. 1=1 Then g is continuous by Corollary 1.11.10. By the first part of the Theorem, 2Vl is connected for every i < n. So (V) is connected, being the range of g. Conversely, assume that 2X is locally connected. Let x 6 X and let U be an open neighborhood of x in X. Since ({U}) is an open neighborhood of {x} in 2X (Lemma 1.11.11(1)), by local connectivity of 2X there is a connected neighborhood C of {x} in 2X such that {x} e C C {{£/}). Then V = C n X is a neighborhood of x in X and clearly
V CC = |J C C [7. We claim that C is connected. Suppose that this is not true. Let E and F be disjoint nonempty relatively open nonempty subsets of C that cover C. There are disjoint open subsets E',F' C X such that E1 n C = E and F' n C = F (Lemma A.8.1). Now observe that ( { E ' } ) , ({Ff}) and ({E' , F'}} are disjoint nonempty open subsets of 2X (Lemma 1.11.11(1)), that Gn({E'}} and en({F'}) are nonempty and that G C ({E'}) U({F'}) U ({E',F'}). This contradicts the connectivity of C. D Hyperspaces of subcontinua. Let X be a space and put G(X) = {A € 2X : A is connected}. It is called the hyperspace of subcontinua of X and is of particular interest in the case that X is a continuum. Lemma 1.11.14. Let X be a space. Then G(X) is a closed subspace of2x. So if X is compact then so is Proof. Let A be an element of 2X which does not belong to Q(X). Write A as the union of two disjoint relatively clopen nonempty subsets E and F. By Corollary A.4.3 there are disjoint open subsets E' and F' in X with E C E' and F C F' . Then ({E',F'}) is an open neighborhood of A which clearly misses C(X). The second part of the lemma follows from the first part and Corollary 1.11.4. D Whitney maps. A fundamental tool in hyperspace theory is the concept of a Whitney map which we will now discuss. Let X be a compact space. A Whitney map for 2X is a continuous function g: 2X —> [0, oo) having the following properties: (1) if A, B G 2X and A is a proper subset of B then g(A) < g ( B ) ,
104
1. BASIC TOPOLOGY
(2) g ( { x } ) = 0 for every x e X. Observe that (1) implies that if A G 2X and A\ > 1 then g(A] > 0. There are several ways to construct such functions. We will discuss only one of them. Consider the function space C(X,I). It is separable by Proposition 1.3.3. So let us fix a sequence (fi)i of functions in C(X, I) which form a dense subset of C(X, I). Define oj: 1X ->• I by
i—l
It is clear that uj is well-defined. Proposition 1.11.15. Let X be a compact space. The function u;: 2X —>• E denned in (*) is a Whitney map for 2 X . Proof. Let us first prove that for a fixed i the function 2X -> I defined by A H-> diam/; [A] is continuous. This is clear since it is the composition of the functions
here 2^ is the hyperspace map of /; (which is continuous by Lemma 1.11.5) and diam: 21 -> I is the function that sends every element A G 21 onto its diameter (which is continuous by Exercise 1.11.12). So u; is continuous by Exercise A. 3.1. Let A, .B G 2X be such that A is a proper subset of B. Fix an arbitrary element x e £?\ A There is by Lemma A. 3.1 a continuous function /: X -> 1 such that /[A] = {0} and /(x) = 1. Let n > 2 be such that
Observe that 0, 1 € /[-B] since A 7^ 0 which implies that diam fn [B] > %. In addition, f[A] = {0} so that diam fn [A] < i/4. From this we conclude that diam fn [A] < diamfn[B]. Also, diam ^ [A] < diam fl [B] for every i since A is a subset of B.
1.11. HYPERSPACES
105
This shows that
71—1
i=l
i=n+l
n— 1
oo 2
^ ~* i=l
diam
2
/i[#] + ~
n diam
/n[£] + XI z=n+l
which is as required. Since clearly u({x}) = 0 for every x G X, this completes the proof.
D
Let uj\ 2X -> [0,oo) be a Whitney function. For each t G [0,o;(A")] the collections {A £ 2X : u(A) = t} and {A € G(X) : w(A) = *}
are called Whitney levels. Lemma 1.11.16. Let X be a continuum, and w: 2X —>• [0,oo) a Whitney function. Then for every t G [0, w(X)] the Whitney level £ = {A G C(X) : u(A) = t}
is closed in Q(X} and moreover covers X. Proof. The first statement is trivial since £ is equal to
which is an intersection of two closed subsets of 2X (use that cj is continuous and Lemma 1.11.14). To check that £ covers X, let us consider any x E X: and let
Observe that 3C ^ 0 since Ct>({x}) = 0 and hence {x} G 3C. Since by continuity of a; and Exercise 1.11.2 the collection OC is closed in C(X) (and hence compact), there is an E G DC with : AT G 3C}.
To see that ui(E) = t, let us fix any e > 0, and let £/ be an open neighborhood of E with u(C) < u(E) + e for any E C C C 17 (Exercise 1.11.22). The
106
1. BASIC TOPOLOGY
component H of U which contains x intersects U \ U by Corollary A. 10. 5. So E is a proper subcontimmm of the continuum C = E U H . Hence
By (*), C g X for otherwise u(C) < u(E). Hence u(C) > t and so
u(E) >t-e. Since e > 0 was arbitrary, this shows that, indeed, x 6 E 6 £.
D
Corollary 1.11.17. Let JC be a hereditarily indecomposable continuum, and oj: 2X —>• [0, oo) a Whitney function. Then for every t G the Whitney Jevel £= 4 is an upper semi-continuous decomposition X . Proof. Suppose C and D are intersecting continua in X with o;(C) = Since ^C is hereditarily indecomposable, C C D or D C C (Exercise 1.10.21). If e.g., C C D then C = D for otherwise uj(C] < u(D}. This shows by Lemma 1.11.16 that £ is a partition of X. To prove that £ is upper semi-continuous, let A C X be an arbitrary closed set. Then
({A, X}} n £ = {E e £ : E n A ^ 0} X
is closed in 2 since both ({A,X}) Lemma 1.11.16). As a consequence,
and £ are (use Exercise 1.11.2 and
is closed in X by Lemma 1.11.6.
D
Exercises for §1.11.
1. Let X be a space and let E and F be compact nonempty subsets of X. Prove that gH(E,F) e [0, oo). 2. Let A" be a space and let F C X be closed. Prove that the sets {A € 2X : A n F ^ 0},
{A £ 2 X : .4 C F}
are both closed in 2A . 3. Prove that yn(X) is closed in 2X for every n e N. 4. Prove that 2 X is separable. 5. Let (A n ) n be a decreasing sequence of nonempty compact subsets of X. Prove that An ->• f L ^^ in 2X .
1.11. HYPERSPACES
107
6. Let X be a space containing more than one point and n G N. Prove that the function fn : Xn —>• 3~n(X) denned by
is continuous, and a homeomorphism if and only if n = 1. 7. Let A^ be a subspace of Y. Prove that the natural inclusion 2X <—>• 2y is an imbedding. 8. Let A be a space. In addition, let (An)n and (Bn)n be sequences in 2A converging to A and B, respectively. Prove that if An C Bn for every n then A C B. 9. Let X be a space and let (An)n be a sequence in 2 X converging to an element A 6 2X . Prove that if an G An for every n then there is a subsequence of (a n ) n converging to an element a G A 10. Let A <E 2 X , and for every n let An e 2* be such that A C An C D(A, l/n}. Prove that (A n ) n converges to A in 2 X . 11. Prove that 6(1) is homeomorphic to I 2 . 12. Let X be a space. Prove that the function diam: 2 X —> [0, oo) sending each element of 2X onto its diameter, is continuous. 13. Let X be compact and let E,F G 1X . Prove that , F) = maxjsup g(x, F), sup ^(x, E)}. 14. Prove that I" can be imbedded in 21 for every n. 15. Let X be compact and let 33 = {Bn : n e N} be an open base for it such that B\ = X. Let (A,){ be a sequence in 2 X . (1) Prove that there exists a decreasing sequence (Xn)n of closed subsets of X with the following properties: 1. Xi=X, 2. every neighborhood of Xn contains infinitely many Aj's, and 3. if Xn r\Bn ^ 0 then Xn\Bn has a neighborhood F containing at most finitely many A;'s. (2) Put D = Pl^Li Xn- Prove that there is a subsequence of the sequence (Ai)i converging to D. (So this is an alternative proof of the second part of Corollary 1.11.4.) 16. Let X Prove (1) (2)
and Y be compact, and let /: X —> Y be a continuous surjection. that the following statements are equivalent: / is open. The function F: Y ->• 2X defined by
is continuous.
108
1. BASIC TOPOLOGY
17. Let (X, £>) be a compact space and let /: X —> Y be a surjection which is both continuous and open. Prove that Y is homeomorphic to the subspace
of 2X . Conclude that formula d(yi,yi} =
QH(f~l(yi},f~l(y2}}
defines an admissible metric on Y. 18. Prove that the functions s,i: 21 —> I defined by s ( A ) = m i n A and £(/!) = max A are continuous. ^•19. Let Q be an admissible convex metric on the compact space X. Prove that the function e:2Xx [0,oo) -)• 2X defined by e(A,t) = Dt(A) is continuous. ^•20. Let X be a space and let k £ N. In addition, let A = {Al : i £ N} be a sequence of subsets of X, each of cardinality k. Prove by induction on k that there are an infinite subset E C N and a (possibly empty) subset A° of X and for every i € E a partition A®, A} of Ai such that that one of the following statements is true: (A) Every x £ X has a neighborhood meeting finitely many terms of the sequence (Ai)i^E only. (B) A V 0 and (1) the sequence (A°); £j E converges to A° in the Vietoris topology. (2) Every x € X has a neighborhood meeting finitely many terms of the sequence (A\}i^E only. 21. Let X be a compact space with disjoint closed subsets A and B. Define C = {C G 2X : C is a continuum from A to B}. Prove that C is a closed subspace of 2X . 22. Let X be a space, and let a; be a Whitney function for 2X . Let e > 0 be arbitrary. Prove that for every E € 2X there is a neighborhood U of E in X such that for every C e 2X with E C C C Z7 we have w(C) < u;(£) + e. . Let X be a hereditarily indecomposable continuum, and u: 2X —>• [0, oo) a Whitney function. For t e [0, u;(X)] consider the Whitney level £ = {A €
Prove that the natural quotient map TT : X —>• JC/£ is open.
1.11. HYPERSPACES
109
24. Let X be a compact space and let uj: 2X —> [0, oo) be a Whitney map. Prove that if t G (0,o;(X)] is such that its corresponding Whitney level & = {A£2X :u(A) = t} is nonempty then inf{diam(£) : E € 8} > 0. 25. Let X be a compact space and let u: 2X —>• [0, oo) be a Whitney map. Prove that for every e > 0 there is an element t £ (0, oo) such that if A e 2X is such that u>(A) < £ then diam(A) < s. 26. Let T and X be spaces and let ip: T —>• 2 X be continuous. In addition, let TT : X x T —> T be the projection. Prove that if E = {(x,t) : x <E
CXxT
then -n \ E: E —> T is a closed surjection.
This page intentionally left blank
CHAPTER 2
Basic combinatorial topology In this chapter we present some elementary combinatorial results and apply these to get nontrivial information about the topology of the Euclidean spaces R n , n e N. The main result is the Brouwer Fixed-Point Theorem. 2.1. Affine notions
In this section we will introduce simplicial complexes and present some basic results on them. So-called subdivisions of simplicial complexes are important in the proof of Brouwer's Fixed-Point Theorem as well as in ANRtheory. For completeness sake, we begin by reviewing some elementary Linear Algebra. The proofs of these results are left to the reader. Let V denote a fixed vector space. Let F be a finite subset of V, say F = {vi,... ,vn}. An affine combination of v i , . . . , vn is a vector v that can be written in the form J^ILi tivi with ti,..., tn G E and X^=i ^ = •*•• Such a v is called geometrically dependent on F. A linear (affine) subspace of V is a subset of V closed under the formation of linear (affine) combinations. Theorem 2.1.1. Let A be a subset ofV and let a 6 A. Then A is an affine subspace if and only if A — a is a linear subspace. If 5 C V, then the intersection of all affine (linear) subspaces of V containing 5 is the smallest affine (linear) subspace of V containing 5. This subset is called affine hull (linear hull) of 5 and is denoted by aff(S) (lin(5)). Theorem 2.1.2. aff(S) is the set of all affine combinations of elements of S. Moreover, for every a 6 5 the following equality holds: aff(S) — a — lin(5 — a). We see that if 5 C V then aff(5) is a translated linear subspace of V, since for every a £ S the equality aff (S) = a + lin(5 - a) 111
112
2. BASIC COMBINATORIAL TOPOLOGY
holds. We say that S spans the affine subspace aff(S'). Let v\,..., vn € V. Then (1) the vectors vi,..., vn are said to be geometrically independent if for all elements t i , . . . , tn G M with ^"=1 ^ = 0 and SILi ^^ = 0 we have t\ = • • • = tn = 0, (2) a subset 5 C V is linearly (geometrically) independent if and only if every finite subset is. Theorem 2.1.3. Let 5 C V. The following statements are equivalent: (1) (2) (3) (4)
5 is geometrically independent, no x G 5 is geometrically dependent on a finite subset F C 5 \ {x}, for every s G 5, {x — s : x E 5, x ^ s} is linearly independent, for some s e 51, {x — s : x 6 5, x 7^ s} is linearly independent.
Corollary 2.1.4. Let 5 C Rn be geometrically independent. Then S contains at most n + 1 points. Consider the affine subspace aff(S) and fix an arbitrary a G S. We saw that aff (5) = a + lin(5 — a). So there exists a subset T C S such that lin(5 — a) = lin(T — a) while moreover T — a is linearly independent. By Theorem 2.1.3 it follows that T' = {a} U T is geometrically independent. Then aff(T') = o + lin(T' - a) = a + lin(T - a) = a + lin(5 - a) = aff(S). So we conclude that 5 and its geometrically independent subset T' span the same affine subspace. An affine subspace of a linear space is called m-dimensional if it is spanned by m + 1 geometrically independent vectors. Proposition 2.1.5. Let S C V be geometrically independent. If A,B C 5 then aff(A) n aff (B) = aff (A n B}. The following result is a nice test for proving that subsets of V are geometrically independent. Theorem 2.1.6. Let OQ, o > i , . . . , an be elements ofV such that di+i & aff({a 0 ,... ,a»}) for every i < n. Then {O,Q. . . . , an} is geometrically independent. A function between affine subspaces of linear spaces is called affine if it preserves affine combinations. Images and preimages of affine sets under affine functions are again affine.
2.1. AFFINE NOTIONS
113
Theorem 2.1.7. Let Vi and V2 be linear spaces, let A\ C Vi and A? C V? be affine subspaces and let f : A\ —* A2 be a function. Then the following statements are equivalent: (1) f is affine, (2) the composition Al - ai ^4 Al —> A-2 ^ AI - a?
(ai 6 AI, a2
i.e., the function £: AI — a\ -» A2 — a2 defined by
£(x) = f(x + ai) -a 2 , is linear. From Theorem 2.1.7 it follows that an affine function, the domain and range of which are both contained in a finite dimensional normed linear space, is continuous. This can be seen as follows. We first claim that for arbitrary n,m (E N, linear functions
are continuous. First observe that each linear function /: M —> M is a multiplication and hence is continuous. Assume that every linear function /: W —> R is continuous for i < n, and let F: Rn+l -> M be linear. If fl = F \ W1 x {0} and /2 = F f {0} x {0} x • • • x {0} xM. n-l
then they are continuous by assumption. But then F is continuous by linearity, being the sum of the continuous functions
and
( x i , . . . , x n + i ) ^ / 2 (0, . . . , 0 , x n + i ) . Consequently, each linear function /: Mn —>• RTO is continuous since Wn is endowed with the product topology. By Exercise 1.1.18 it follows that each finite dimensional normed linear space is topologically isomorphic to some Mn . So we conclude that a linear function between finite dimensional normed linear spaces is indeed continuous. This gives us what we want since by Theorem 2.1.7 each affine function /, the domain and range of which are both contained in a finite dimensional normed linear space, is of the form
where both £ and 77 are translations and hence homeomorphisms (Exercise 1.1.3), and F is linear.
114
2. BASIC COMBINATORIAL TOPOLOGY
Simplices. Let V be a linear space. An n-simplex in V is a geometrically independent subset of V having precisely n + 1 points. Simplices are denoted by Greek letters. If a and r are simplices and a C T then a is called a face of r- to indicate that a is a face of r we shall also use the notation: a ^ r. An n-simplex in V is sometimes also called an n-dimensional simplex. Theorem 2.1.8. Let a be a simplex in V and let A = aff(a). Then every element x G A can be written uniquely as an affine combination ^v£a. tv • v of a. In addition, the functions av : A —> R denned by av(x] = tv are affine. Corollary 2.1.9. Let a and T be simplices in V such that r ^ a and let A = aff(cr),
B =
aff(r).
If x G B then av(x) — 0 for every v G a \ r. The real numbers av(x) for v G a are called the affine coordinates of x with respect to a. We call the av the coordinate functions of aff(cr). This notation will remain in force throughout the book. Observe that by the above remarks and Theorem 2.1.8 it follows that the coordinate functions of aff(cr) are continuous on aff(cr). A geometric simplex is the convex hull of a simplex. We use |cr as an abbreviation for conv(
2.1. AFFINE NOTIONS
115
Proof. We prove that diam(|cr|) < diam(<7). Take arbitrary p, q £ cr|. Then since £veo. av(p) = 1 we get
from which it follows by using standard properties of the norm that
= max||i; — q\ . As a consequence we obtain for every w G a that \\q-i So
<
t> — w ; .
\\P ~ Q\\ < max \\w — q\\ < max \\v — D
as required.
Figure 6. For each 0 < i < k let di 6 R/c+1 be the vector all coordinates of which are 0, except the (i + l)-th coordinate which equals 1. The vectors d0,..., dk are clearly linearly independent (hence geometrically independent). For x in aff ({do,..., dk}) we have that the affine coordinates of x = (XQ, ..., Xk) with respect to do,... ,dk are equal to XQ, . . . , X k , respectively. Put rk =
116
2. BASIC COMBINATORIAL TOPOLOGY
Observe that \Tk = {x G I fc+1 : J^ILi ^i = 1} is a closed and bounded subset of R/c+1 and hence is compact (alternatively, use that \Tk\ is a geometric simplex). Theorem 2.1.12. Let (V, || • ||) be a normed linear space, and let a — {v0,...,vk} C V be a k-simplex. Then the function f : \a\ —»• |r^| denned by /(a) = (a w o (a),...,a! w f c (a)) is a homeomorphism. Proof. First observe that the functions aVi are affine (Theorem 2.1.8). Consequently, / is continuous by the remark following Theorem 2.1.7, and oneto-one by the definition of affine coordinates (Theorem 2.1.8). Moreover, / is clearly surjective since if y G \r^\ then f ( x ) = y, where
By compactness of \cr\ it therefore follows that / is a homeomorphism (Exercise A. 5. 9). D Triangulation. A simplicial complex is a countable collection § of nonempty, finite sets such that: (SC)abstact
if
The elements of the set
s = Us
are called the vertices of the simplicial complex S. Although S need not be a subset of a normed linear space, without loss of generality we may assume that this is the case. In fact, we can think of S as being a linearly independent subset of some normed linear space. This can be seen as follows. For every n G N let xn G I2 be the vector all whose coordinates are 0 except for the n-th coordinate which equals 1. Observe that the set D = {xn : n G N} is linearly independent. Let /: S —> D be an injection, and identify every simplex a G S with /[
(sc)
MOM - ^ n ^ .
In the sequel, when dealing with a simplicial complex S, we shall always require that S = (J § is contained in a normed linear space L, and that (SC) holds. The set |S = [JIM : cr ^ §} is called the geometric realization of S
2.1. AFFINE NOTIONS
117
and |S| is said to be triangulated by S (or S is a triangulation of |8|). Observe that the collection Sy = {|a| : a G §} consists of geometric simplices in L and has the following properties: (SC)geometric
(1) if cr| £ §|| and r is a face of a then |r| 6 Sy, (2) for all |<Ji|, (72 1 € §>\\ with <TI fl <72| 7^ 0, |<TI| H |a2 is a face of <TI as well as |<7 2 |-
Let cr\ be a geometric simplex in L. The collection of all faces of |cr| is clearly a simplicial complex which triangulates \a\. It is called the standard triangulation of |cr| and is denoted by 3(0). Topologizing a simplicial complex. From this moment on, we will no longer distinguish between an abstract simplicial complex and its geometric realization. So a simplicial complex in a normed linear space L is a collection S of geometric simplices (which by abuse of terminology we will call simplices for short in the sequel) in L such that for all a\ , 0*2 G S we have (1) if a G S and r is a face of a then r € S, (2) ai n 02 / 0 =>• 0"i n <72 is a face of a\ as well as 02A subcollection T of 8 which is also a simplicial complex is called a subcomplex ofS. Let T be a simplicial complex in L. For each m > 0 define <j(m) = |s £ 'J : <j is at most m-dimensional}. So T 0 is the set of vertices of all simplices in T, T^1) is the collection of all at most one-dimensional faces of all simplices in T, etc. The elements of T'°) are called the vertices of T. It is a triviality that 7^m^ is a subcomplex of T; we call it the m- skeleton of T. Let L be a normed linear space and let T be a simplicial complex in L. We are interested in useful topologies on the set X = (J T. There are several natural candidates for such a topology. First, X is a subset of L and therefore carries the subspace topology inherited from L. It turns out that this is in general an interesting but not a useful topology. We shall now describe a better topology. To this end, let us first agree that there cannot be ambiguity about the topology that each simplex of T should carry. By Theorem 2.1.12, each ^-dimensional simplex in L with its subspace topology is naturally homeomorphic to a standard fc-dimensional simplex in Rfe+1 ; each simplex in L from this moment on is endowed with its subspace topology. It would be very unnatural if the useful topology we want to define on X should induce a different topology on one of the simplices of T. So there is a natural candidate for a topology on X, namely, the largest topology that induces the 'right' topology on every simplex of T, i.e., U C X is open iff for every simplex T e T, U D T is open in T.
118
2. BASIC COMBINATORIAL TOPOLOGY
This is called the Whitehead topology on X and X with this topology is usually denoted by \7\. Lemma 2.1.13. Let T be a simplicial complex in L and let X = \J 7. Then (1) The Whitehead topology on X is a topology and is finer than the topology that X inherits from L, i.e., the inclusion \7\ <-» X C L is continuous, (2) On every simplex of 7 the Whitehead topology and the topology on L induce the same subspace topology. Proof. The proof of (1) is a triviality the verification of which we leave to the reader. For (2), let r be a simplex of T. It suffices to prove that for every relatively open subset U C T there exists a Whitehead-open V C |T| such that V fl T = U. We claim that V — U U (|T| \ T) is as required. It is clear that V fl r = U, and so it remains to prove that V is Whitehead-open. To this end, let a 6 7 and put V = V D (r U a}. Then V is open in r U a since its complement is equal to r \ U which is compact and hence closed in r U a. Consequently, V fl a = V n a is open in cr, as required. D Observe that a subset A of |T| is closed if and only if A n r is closed for every r G 7. This immediately implies that |T| is a T\-space, i.e., a space in which every singleton is closed (this follows also from Lemma 2.1.13(1)). Due to our self-chosen (sometimes unpleasant) restriction to deal with separable metrizable spaces exclusively for the time being, there is a small problem with the Whitehead topology since it need not be metrizable. However, there is a simple combinatorial condition on 7 that ensures that |T| is separable and metrizable; call a simplicial complex T locally finite if every vertex of T is contained in at most finitely many simplices of T. In turns out that for our considerations it always suffices to consider locally finite simplicial complexes and these are precisely the simplicial complexes that are separable metrizable when given the Whitehead topology, see Proposition 2.1.21 below. For the moment, we shall not worry about the metrizability of |T| but we shall prove a few easy but important lemmas from which metrizability in the locally finite case will follow rather easily later. Lemma 2.1.14. Let 7 be a simplicial complex with subcollection S. Then (1) U S is a closed subspace of |T|, (2) if S is a subcomplex then the topology that |J § inherits from Tj coincides with the Whitehead topology on |JS.
2.1. AFFINE NOTIONS
119
Proof. For (1), for convenience, put Y — (J S. Since each simplex of T has finitely many faces only, and T is a simplicial complex, it is clear that for every r G T, Y D T is closed in r. By definition of the Whitehead topology, we therefore conclude that Y is closed in |T|. For (2) observe that since S C T, the topology that (J S inherits from |T| is coarser than its Whitehead topology. Now let A C |S| be closed. We shall prove that A is a closed subset of |T|. To this end, let r be an arbitrary simplex in T. Since T has finitely many faces only, there is a finite subcollection y of 8 such that
\j3~r\r = |J S n r. Consequently, A n T = (A n U S) n r = (A n U 5") n r. Now since A C |8| is closed, for every a G 3", A n a is closed in cr, and consequently, A n
and
carx = ( ] { T 6 T : x 6 r}, respectively. We shall prove that carx G T and it is therefore the smallest simplex of T that contains x. Corollary 2.1.17. Let T be a simplicial complex. Then for every x G |T|, (1) Six is open in T|,
120
2. BASIC COMBINATORIAL TOPOLOGY
(2) carx belongs to T. In addition, the family {Six : x G T^} covers |T|. Proof. For (1), let § = {r G T : x €" T}. From Lemma 2.1.14 we obtain that y 8 is closed in |T|, hence Sta: is open. For (2), first take T G 0" such that x G T. Since for every a G T such that x G cr we have cr n r =
(*)
xn G (Knrn+l)\ y rn. i=l
Since K is compact, and metrizable, we may by Theorem A.5.1 assume without loss of generality that x = limn_>.00 xn exists (and belongs to K}. Now let r G T be an arbitrary simplex. Since rn n r is either empty or a face of r, there exists m G N such that oo
m T
U n n=l
n T
= IJ Tn H T, n=l
from which it follows that {xn : n G N} O r C {XI,...,X T O }. By the definition of the Whitehead topology, this implies that {xn : n G N} is closed in |T (recall that |T| is TI). Since the sequence (xn)n converges, it therefore has to be eventually constant, which contradicts (*). D One of the reasons that the Whitehead topology on a simplicial complex is relatively simple to deal with is that it is easy to check that certain functions are continuous.
2.1. AFFINE NOTIONS
121
Lemma 2.1.19. Let 7 be a simplicial complex and let X be a space. A function f : \7\ —>• X is continuous if and only if the restriction of f to every simplex r € 7 is continuous on T. Proof. Suppose that /: |T| —>• X is such that the restriction of / to every simplex r G T is continuous. Let U C X be open. For each r G T, is open in T. By the definition of the Whitehead topology, it therefore follows that f~l[U] is open in |T|. D Corollary 2.1.20. Let T be a simplicial complex. Then \7\ is a normal topological space, i.e., for every pair of disjoint closed subsets A and B of\7\ there exists a continuous function f : 7 —> I such that f \A = 0 and f\B = l. Proof. Since 7 is countable, we can enumerate it as {rn : n G N}. By Lemma A. 4.1 there exists a continuous function /i : r\ —> I such that By induction on n, we shall construct continuous functions fn : UlLi Ti ~^ I such that
(1) /nKAnu: = 1 r,) = o, (2) / n t ( B n U r = i r O = l, (3) if n > 1 then fn extends / n -iSince we already defined /i, assume fi to be constructed for 1 < i < n. We shall construct fn+l. Let A1 = A n US ^ and B' = B n \J^ Ti, respectively. Define g: A' U 5' U UJU r» ->• I by (a; e A'),
Then 5f is clearly continuous and extends fn. Since UF^i Ti '1S compact, it is metrizable by Lemma 2.1.18. By Corollary 1.2.5 there consequently exists a continuous extension /n+i : UlLi T{ —>• I of ^. It is clear that fn+i is as required. Now put / = Ui^i fn- Then / is a well-defined function from |T| into I having the property that / \ A = 0 and / \ B = 1. Moreover, / is continuous by Lemma 2.1.19. D These simple results enable us to prove the following Proposition 2.1.21. Let 7 be a locally finite simplicial complex in the linear space L. Then \7\ is a locally compact separable metrizable space.
122
2. BASIC COMBINATORIAL TOPOLOGY
Proof. We shall first prove that each point of |T| has an open neighborhood which is separable, metrizable, and has compact closure in |T|. To this end, take an arbitrary x G |T|. By Corollary 2.1.17(1), Stx is open. Since 7 is locally finite, St x is contained in the union of finitely many simplices of T. This can be seen as follows. First observe that Six is contained in the union of all simplices that contains x. But, as we will show, there are only finitely many of such simplices. Striving for a contradiction, assume that S is an infinite sub collection of T such that x G P| S. Fix
2.1. AFFINE NOTIONS
123
Now let L and E be norrned linear spaces and let 7 and 8 be simplicial complexes in L and E, respectively. We say that 7 and § are combinatorially equivalent if there exists a bijection / : 7^ —>• S^0) such that F C 7^ spans a simplex in 7 if and only if f[F] spans a simplex in 8. Let us also say that |T| and S| are simplicially homeomorphic if there exists a homeomorphism h: \7\ —> §| such that for all r E T and a E §, we have h[r] 6 § as well as h~l[a] e T. Proposition 2.1.23. Let L and E be normed linear spaces and let 7 and 8 be simplicial complexes in L and E, respectively. The following statements are equivalent: (1) T and 8 are combinatorially equivalent, (2) |T| and |S| are simplicially homeomorphic. Proof. For (1) => (2), let /: 7^ -»• S<°) be a bijection such that F C T<°) spans a simplex in 7 if and only if f[F] spans a simplex in 8. So for every simplex T = |{x0, . . . ,Xk}\ € T, the set {/(x 0 ), . - . ,/(£&)} spans a simplex cr of 8. Define a homeomorphism fr : T —> a by
i=0
i=0
here ao(^), • • • ,(*k(x) denote the affine coordinates with respect to T of an arbitrarily chosen point x E T of course. It is easily seen that jr is indeed a homeomorphism, cf. Theorem 2.1.12. Now define /: |T| —>• |S| by riET
By unicity of affine coordinates (Theorem 2.1.8), / is well-defined. In addition, / is continuous since the restriction of / to every simplex of 7 is continuous (Lemma 2.1.19). It is easily seen that / is one-to-one and surjective and that f~l is continuous for the same reason / is. We conclude that / is a homeomorphism. Clearly both / and f~l are 'simplex preserving'. Conversely, let /: |T| —>• |S| be a homeomorphism such as given by (2). The function is clearly as required.
D
From the above proposition we conclude that the topology of a simplicial complex depends only on the combinatorial properties of its vertex set. For that reason when discussing a simplicial complex, it is no longer necessary to mention the normed linear space it is a subset of. Let X be a space. We say that X is a polytope if there exists a locally finite simplicial complex T such that X and T| are homeomorphic. In case 7
124
2. BASIC COMBINATORIAL TOPOLOGY
is finite we say that X is a polyhedron. Observe that each polyhedron is compact. Polytopes are very important since they are the 'bridge' between abstract topological spaces and concrete ones. If X is a polytope, say X is homeomorphic to |T|, then we will find it sometimes convenient to not distinguish between X and |T|. This will never cause confusion because the triangulation under consideration will always be explicitly defined but the reader should keep in mind that there usually are many different triangulations of the same polytope. If |T| is a polytope then a subset Y C |T| is called a subpolytope if there is a subcomplex 8 C T such that Y = |S|. Similarly for subpolyhedron. Theorem 2.1.24. Each polyhedron is an ANR. Proof. Let T be a finite simplicial complex. By induction on the cardinality of T we shall prove that |T| is an ANR. If T consists of one simplex only then |T| is an AR by Theorems 2.1.12 and 1.2.9. Suppose that for every simplicial complex 7 of cardinality at most n — I > 1, |T| is an ANR, and let T be a simplicial complex of cardinality n. Since T is finite, there exists a simplex T 6 7 such that for every u E 7 with r ^ cr, a = T (let T be any simplex of maximal dimension). Put S = T \ {T}. Then § is a simplicial complex and |T| = r U |S|. In addition, clearly T n |8| = \{T n cr : a G S}\. So by our inductive assumptions and by Theorem 1.2.16(1) it follows that |T| is an ANR. D In Corollary 4.3.6 we shall prove that every polytope is an A N R , which generalizes Theorem 2.1.24. Exercises for §2.1. 1. Prove Theorem 2.1.1. 2. Prove Theorem 2.1.2. 3. Prove Theorem 2.1.3. 4. Prove Proposition 2.1.5. 5. Prove Theorem 2.1.6. 6. Prove Theorem 2.1.8. 7. Prove Theorem 2.1.10. 8. Let A be an affine subspace of a finite dimensional linear space L. Prove that A is closed in L. ^•9. Prove that if a is an n-simplex in W1 then Int(|cr|) is equal to \a\°. 10. Let a be a simplex and let p G a. Prove that a \ {p} is convex if and only if p is one of the vertices.
2.2. BARYCENTERS AND SUBDIVISIONS
125
11. Let a be a simplex and let C be a convex set which is contained in da. Prove that C is contained in a proper face of a. 12. Let X be a set triangulated by the simplicial complex S. Prove that the collection of all geometric interiors of elements of |S| is a partition of X . 13. Let 11 be a countable collection of sets. Prove that the collection (J : y C U is finite and P| 3 ^ 0} is a simplicial complex. 14. Let T be a simplicial complex. Prove that |T| is paracompact (we do not assume that T is locally finite). 15. Let T be a simplicial complex that is not locally finite. Prove that |T| is not metrizable (this is the converse to Proposition 2.1.21). ^•16. Let T be a simplicial complex and S a subcomplex of T. Prove that the topological interior and topological boundary of |S| in |T| are given by {x e |S| : x e a £ T => a G §} and
|JCr\s)n|S|, respectively.
2.2. Barycenters and subdivisions Througout this section, L is a fixed normed linear space. All simplices under consideration are subsets of L. If a is a simplex in L then a denotes the set of its vertices. Barycenters. Let a be a simplex. The barycenter ba of a is the point in a whose affine coordinates (with respect to ci) are all equal: if a - {V0,. . . , f n }
then
Observe that ba belongs to the geometric interior a° of a. Lemma 2.2.1. Let S be a simplicial complex. Then if cr, a' G § and ba G a' then a is a face of a' . Proof. By (SC), a Pi a' is a face of a. This face contains bff so it must be a itself. D Lemma 2.2.2. Let r\ C r2 C • • • C Tk be a strictly increasing collection of faces of a simplex a. Then the barycenters bTl , bT2 , . . . , 6Tfc are geometrically independent.
126
2. BASIC COMBINATORIAL TOPOLOGY
Proof. By Theorem 2.1.6 it suffices to show that for i < k, bTi+1 £aff({6 Tl A 2 ,...,& Ti }). To this end, take an arbitrary vertex v 6 f^i \ T;. Then
Since {6 T1 ,6 T2 ,...,6 T J C aff(f.) we get
aff({6 T l ,6 T 2 ,...,6 T J) C aff(fi). This and Corollary 2.1.9 imply that for every x 6 aff (|&TI , 6T2 , . . . , bTi }) we have av(x) = 0. This is clearly as required. D Let a be a simplex and let 3C = (TI, T2, . . . , r/t} be a chain of faces of a (by a chain we mean of course a chain with respect to inclusion) with corresponding chain of vertices 3C = {ri,f 2 , . . . , fjt}. If 3C is a maximal chain of faces then clearly <7 € 3C and consequently |J JC = o\ This rather trivial remark will be used without explicit reference a few times in the remaining part of this section. For every v 6 (J 3C define its height ht(v) to be the first i < k with v 6 f;. Define a quasi-order '= M(v) < ht(w). Lemma 2.2.3. Let a be an n-simplex and let 9C be a maximal chain of faces of a with corresponding quasi-order =<;. Then (1) OC has cardinality n + 1 and tie corresponding quasi-order =<; is a linear order on a = (J 3C. (2) If/? is the geometrically independent set of all barycenters of elements of 3C then it is geometrically independent and \^\ — {x £ a : (Vv,w E cr)(v ^ w =>• Q W (X) > aw(a:)}. (3) If v ^, w and for certain x G |/?| we have av(x] = aw(x] then for every T G OC with v 6 f and w 0 f we have that the affine coordinate of x in \f3 1 with respect to bT is equal to 0. Proof. We claim that if a is the direct successor of /? in 3C then a \ /? is a singleton: if it contains two distinct points v and w then
would be a chain extending 5C; likewise one sees that the first element of 3C is a singleton. Thus OC = {TO,TI, . . . ,rn} where f^ has i + 1 points. Now write a = {VQ,VI, . . . ,vn} such that for every 0 < i < n, TV = {^0,^1, • • • , ^}Then Vi =
2.2. BARYCENTERS AND SUBDIVISIONS
127
For (2), observe that /3 is geometrically independent by Lemma 2.2.2. An arbitrary element x 6 \(3\ can be written as n
n
x — 2_. ti • bTi where ti > 0 for every i and i=0
(Lemma 1.1.1). Observe that
=0
j=0
i=0
j=i
j +1
J
For every 0 < i < n define n
and observe that
i=0
i=Q
j=Q
i=0
From this we conclude that the Si are the affine coordinates of x (Theorem 2.1.8) and satisfy SQ > si > • • • > sn. Conversely, let p be a point in a whose affine coordinates SQ, ..., sn have the property that SQ > si > • • • > sn. Then the numbers /L
n
— \^ld („
l-\\e
T^ ±JOn
are non-negative and have the property that n
2_^ ^i ' brt = (SQ — i=0
Si
i=0
= p. In addition, J^ILo ^ ~ Sr=o si ~ ^- ^e ^ are consequently the affine coordinates of p in \j3\ (Theorem 2.1.8). This proves (2). For (3), simply observe that if for p above we have that if Si = Sj for certain i < j < n then all tk for i < k < j are equal to 0. D
128
2. BASIC COMBINATORIAL TOPOLOGY
Lemma 2.2.4. Let a be a simplex and let 3Ci and OC? be chains of faces of a. Let @i and fa denote the sets of barycenters of elements of 3Ci and 3C2, respectively. Then
Proof. Since |/?i| n \fa\ is convex and contains fli n/3 2 , it is clear that
\Pinfol = conv(/?i n &) c |ft| n |/?2|. For the reverse inclusion, take an arbitrary x G |/?i | fl \fa\- Since (3Ci\3C2)n(3C2\3Ci)=0, without loss of generality we may assume that <7 0 UCi \ 3^2- We assume without loss of generality that XL \ 3C2 ^ 0 for otherwise we get what we want from Lemma 2.2.2. It is clear that there exists a linear order ^ on the set a such that for every element of T G 3Ci we have that f is an initial =^isegment. This order corresponds to a maximal chain of faces of a. Pick an arbitrary element r G KI \ 3C2; observe that T ^ a. The collection 3C2 can easily be extended to a maximal chain -C of simplices such that r $ £ (Exercise 2.2.1). Consequently, r is not an initial segment with respect to the corresponding linear order =<(2 (Lemma 2.2.3(1)), i.e., there exist distinct elements v , w G cr such that v G r, w; ^ r and w; =^2 ^; consequently, o^Oc) > a w (a;) (Lemma 2.2.3(2)). Observe that v =^i w from which it follows similarly that
av(x) > aw(x). By Lemma 2.2.3(3) we therefore conclude that the affine coordinate of x in the simplex |/?i | with respect to bT is equal to 0. D Subdivisions. Let X be a subset of L which is triangulated by S. A triangulation T of X is called a subdivision of S if for every simplex cr G S we have that the collection
7 (a) = {r € T : T C cr} is finite and is a triangulation of cr. Lemma 2.2.5. Let S and T be triangulations of X. If 7 is a subdivision of § then for every r G T and cr G S sucJi tAat bT G a we Lave r C cr. Proof. Take an element T' G T(cr) such that bT G T'. Then by Lemma 2.2.1 it follows that r C r' C cr, as required. D Theorem 2.2.6. Let S, T and $ be simplicial complexes. If § is a subdivision of 7 and 7 is a subdivision of^R then §> is a subdivision ofR.
2.2. BARYCENTERS AND SUBDIVISIONS
129
Proof. We have to prove that S(p) is a finite triangulation of p G ft. That the collection S(p) is a simplicial complex is clear since condition (SC) holds for all simplices in 8. Let a G S(p). We claim that there exists an element ra G T(p) such that <j C Tff. This is easy. Indeed, since IJT(p) = p there exists Tff G T(p) such that ba G ra. So by Lemma 2.2.5 it follows that ra is as required. Since T(p) is finite and for every r G T(p), S(r) is finite, we obtain that the collection S(p) is finite. To conclude the proof, observe that \JS(p) — p is trivial since (J T(p) = p and \J §(r) = r for every r G T(p). D Let a be a geometric simplex in L. We shall define a special triangulation of cr, the so-called barycentric triangulation. Let -B(cr) denote the set of all barycenters effaces of a. We shall define a simplicial complex 23(cr) consisting of subsets of B(a] as follows: A nonempty subset /3 of B(a) belongs to 13(0") if and only if the faces of a the barycenters of which belong to /3 form a chain. It is clear that if j3 G 23 (cr) and 7 C J3 then 7 G 23 (cr). So 23 (cr) is an abstract simplicial complex. By Lemma 2.2.4 it follows that the collection
satisfies condition (SC) (see Page 116). So the geometric realization of 23 ( is |23(cr) and, in particular, is contained in a. In order to prove that |23(a is a subdivision of cr, it therefore remains to prove that
Let x G a. Without loss of generality we have a = {VQ, . . . , vn} and ttvn (X}
<
a
Vn - 1 (X)
< • • • < ^f 0 (X} •
This ordering corresponds to a (maximal) chain of faces in cr, which in turn corresponds to a (maximal) simplex /? G 23(0"). By Lemma 2.2.3(2), x G |/?|. Let X be a set in L which is triangulated by the simplicial complex §. As above, for every a G S let 23 (cr) be the barycentric triangulation of cr. The collection ®(§) = {|r| G 23(a)| : c r G § } is called the bary 'centric subdivision of §. That 23(8) is a collection geometric simplices with union X is clear. We claim that it is a simplicial complex and is a subdivision of 8. For that we only need to verify condition (SC) (see Page 116) because for every a G 8, { | r | e | ® ( S ) | : | r | C < r } = 23(
130
2. BASIC COMBINATORIAL TOPOLOGY
Let fa G S(cTj) for <Ji G 8, i = 1, 2. If <TI D 02 = 0 then there is nothing to prove. If cri n o"2 ^ 0 then a = <TI n cr2 is a face of <TI as well as cr2. Now put Then 71 and 72 belong to 25 (
M = IA|n<7 (. = 1,2). That |7i| C |/3; | n a is clear. Take an arbitrary x G |/3;| n a. There exists an element /3 G 23 (o7) such that a: G |/?|. Observe that
Consequently,
By Lemma 2.2.4 we therefore conclude that
\/3i\ n \/32\ = |&| n |/32| n a = |7i| n |72| = |7l n 72 | = [ft n /?2|, which is as required. Sometimes it will be convenient to denote S by sd^S and 23 (S) by We define the second barycentric subdivision sd^S of S by !B ($(§)). Similarly, one defines sd^S, the n-th barycentric subdivision of S. Notice that by Theorem 2.2.6, sd^S is a subdivision of § for all n. Let S be a triangulation of a subset of L. The mesh of S, denoted by mesh(S), is defined to be the number supdiam(er). ff£§
We allow mesh(S) to be equal to oo. Theorem 2.2.7. Let a be an n-simplex in the normed linear space L and let *B(cr) be the barycentric triangulation of a. Then mesh(|-B(o-)|) < —^— -diam(cr). Proof. First observe that for every /? G *B((7) we have diam(/?) = diam(|/?|) (Theorem 2.1.11). Fix (3 G B(
of (T such that c is the barycenter of {^0,^1, • • • , Vk} and d is the barycenter of {^QJ^IJ ••• ,vm}, respectively. We may assume without loss of generality
2.2. BARYCENTERS AND SUBDIVISIONS
131
that k < m. Then \\d-c\ i=0
TTL ~T~ J-
m
1 k +l
j=0 ,
/v
;+1 + ltl
V i i "
fc / I V 1( 1 k + l j =0 \ V m + 1
M
3=0
max
0<j
Moreover,
m
Vi
-
Vn
m +l
i=0
i=0
m+l m diam(
In the last inequality we used that one of the terms (if i = j) is equal to 0. We conclude that diam(/3) <
ra
m+l
diam(a).
Finally observe that m < n.
D
Corollary 2.2.8. Let § be a triangulation of a set X in a normed linear space L such that every a G § is at most n-dimensional. IfB is the barycentric subdivision of S then mesh(S) <
mesh(S).
Corollary 2.2.9. Let \T\ be a geometric simplex in a normed linear space L. Then for every e > 0 there is an m 6 N such that
132
2. BASIC COMBINATORIAL TOPOLOGY
Proof. Assume that r is an n-simplex. By Corollary 2.2.8 it follows that for every m G N we have i
\
/
mesh (sd (m) J(r)) < (
^ iv
Ti I
\m
)
• diam(r).
-L *
We conclude that for a sufficiently large m, mesh(sd^ TO ^5'(r)) <e.
D
We shall now formulate a very important property of polyhedra. Theorem 2.2.10. Let T be a finite simplicial complex. Then for every open cover U of |T| there exists m G N such that sd^T refines U, i.e., for every simplex u G sd^ m 'T there exists an element U G U with a C U. Proof. Fix r 6 T for a moment. By Corollary 2.2.9 and Lemma A.5.3 there exists n(r) G N such that every a G sd^T''y(r) is contained in an element of U. Let m = max{n(r) : r G T}. Now take an arbitrary K G sd^T. By the definition of sd (m) T it is clear that there exists r G T with K G sd (m) 3~(r). Since every element of sd'm'5F(r) is contained in an element of sd^ n ^ T ^9 r (r), the simplex K is contained in an element of U. D The above theorem can be generalized as follows: for every polytope |T| and for every open cover U of |T| there exists a subdivision § of T such that each simplex a G § is contained in an element of U. However, for noncompact polytopes, S generally cannot be chosen to be an iterated barycentric subdivision. For details, see WHITEHEAD [408].
Exercise for §2.2.
^•1. Let a be a simplex and let 3C be a chain of faces of a. Prove that if r is a proper face of a such that r £ 3C then there is a maximal chain of faces L such that 3C C L but r 0 £. 2.3. The nerve of an open covering
Let X be a space and let U be a countable open cover of X. We say that a simplicial complex T is a nerve of U if 7^ can be indexed as {x(U} : U G U} such that for every n > 0, n
(*)
x(U0),... , x(Un) spans a simplex in T iff Q Ui ^ 0. i=0
If T is a nerve of U then it is convenient to adopt the notation {x(U} : U G U} for T(°\ where it is implicitly assumed that the 'indexing' is such that (*) holds.
2.3. THE NERVE OF AN OPEN COVERING
133
Proposition 2.3.1. Let X be a space. Each countable open cover of X has a nerve.
ur\w uni
u n v nw
unv
cover
V
nerve Figure 7.
Proof. Let U = {Un : n G N} be a countable open cover of X. For each n let xn G I2 be the vector all coordinates of which are 0, except for the n-th coordinate which equals 1. The sequence (xn)n is linearly independent, hence geometrically independent. Define 3 = {F C N : F is finite and |°| Un / 0}n<EF
2
For F 6 y let r(F) be the simplex in I spanned by {xn : n G F}. Put It is easy to verify that T is a simplicial complex and that it moreover is a nerve for U. D Clearly, any two nerves of the same open cover are combinatorially equivalent (hence 'isomorphic'), so by Proposition 2.1.23 we can now speak of the nerve N(U) of the cover U. If L is a normed linear space containing a simplicial complex T which is combinatorially equivalent to the nerve N(U) then we say that N(U) can be realized in L. Let N(U) be the nerve of the locally finite open cover U (recall that U is countable, see Page 486). We shall use the Ac-functions with respect to U to
134
2. BASIC COMBINATORIAL TOPOLOGY
define a canonical continuous function K: X —>• N(U)\. This function is the 'bridge' between the 'abstract' space X and the 'concrete' space |JV(U)| and is called the n-function of the cover U. Indeed, define K: X -»• \N(U)\ by
K(X) We claim that K is well-defined. To check this, for x G A" put
g~(x) = {C/ e U : x e U}. Observe that by the local finiteness condition, 3*(x) is finite, that KU(X) = 0 if U $ y(x), and that ICt/eg-r^) KU(X) = 1. This implies that K
u(x) • x(U).
Hence the infinite sum in the definition of K reduces to a finite sum. In addition, the {Ku(x) : U G ^(x)} are the affine coordinates of K(X) in the simplex r(x) spanned by the vertices {x(U} G U : U e ?(x)}
of 7V(U). This shows that K(X) is well-defined, and also that K(X) G r(x). Lemma 2.3.2. Let X be a space and let U be a locally finite open cover of X. Then for each x G X, r(x] is the carrier of K(X) in N(U). Proof. As remarked above, r(x) is a simplex of N(li) that contains K(X). As remarked above, the Kjj(xys are the affine coordinates of K(X) with respect to the simplex in N(U) spanned by the vertices {x(U) : U G y(x)}. Now since for every U G 3(x) we have that KU(X] ^ 0, T(X) is the smallest simplex containing K(X), i.e., r(x) = carK(x). D Theorem 2.3.3. Let X be a space and let U. be a locally finite open cover of X. Then the K-function K: X —> 1^(11)1 -has the following properties: (1) AC is continuous; (2) for every U G U: K~I [Stx(U)] = U. Proof. We shall first prove (2). Observe that for every x G X and U Eli, x gU «=> U g 3(x) ^=> x(U] g T(X) - carAt(x) (Lemma 2.3.2). Now if x(U) $ r(x] then since K(X) G r(x), K(X) g Stx(C7). This proves that K~1[Stx(U}] C U. Conversely, if K(X) g Stx(U) then there is a simplex T G N(U) which contains K(X) but not x(U). Since car K (x} C r, this gives us that x(U) 0 car/c(x) = T(X) and hence that x 0 U. This proves that K[U] C Stx(U).
2.3. THE NERVE OF AN OPEN COVERING
135
For (1), let L be a normed linear space which realizes 7V(U) (e.g., £ 2 , see the proof of Proposition 2.3.1). Take an arbitrary x € X. There is a neighborhood W of x meeting only finitely many elements of U. Since the infinite sum in the definition of K for points of W clearly reduces to a finite sum, and the functions KU are all continuous, by the continuity of the algebraic operations on L it follows that K regarded as a mapping from X into L is continuous. Now l e t W = { t / e l t : t / n V ^ ^ 0 } and let Y be the union of all simplices of N(U) the vertices of which correspond to elements of the collection W. Observe that Y is a finite union of elements of N(11) and hence is compact. Since x e W it follows that 3~(x) C W and so
K(X) G T(X) C Y. Since Y is compact, by Lemma 2.1.13(1) it follows that the topology that Y inherits from L is the same as the topology that Y inherits from |Ar(U)|. We conclude that K: X —>• |-/V(U)| is continuous at all points of W, and hence at the point x. D Let X and Y be spaces and let U be an open cover of X. A continuous function /: X —>• Y is called a It-mapping if there is an open cover V of Y such that /^[V] < U. This concept is related to the concept of a e-mapping on Page 34. For let /: X —> Y be an gr-map, where X is compact, and let U be the open cover of X consisting of all open sets of diameter less than e. We claim that / is a It-mapping. To this end, first observe that if
yeV = Y\f[X] then V is an open neighborhood of y such that /-1|V] = 0 is contained in every U € U. Next, if y € f[X] then let U be an open neighborhood of f ~ l (y) such that diam(C/) < e. Since by compactness of X the function / is closed (Exercise A.5.5), there is a neighborhood V of y such that /~ 1 [V r ] C U (Exercise A. 1.15). So we conclude that / is a It-mapping. We see that the concept of an e-map is indeed the 'compact' version of the concept of a It-map. Corollary 2.3.4. Let X be a space and let It be a locally finite open cover o f X . Then K: X -> \N(U)\ is a It-mapping. Proof. This follows from Theorem 2.3.3(2) and Corollary 2.1.17.
D
It is unfortunately not the case that a cover U of a space X is locally finite if and only if the simplicial complex N(U) is locally finite. Since, as we pointed out above, our main interest is in locally finite simplicial complexes, it is natural to wonder when JV(U) is locally finite. First a definition. An open cover U of X is said to be star-finite if for every U 6 U the set
{v e it: v n u + 0}
136
2. BASIC COMBINATORIAL TOPOLOGY
is finite. Observe that a star-finite open cover is locally finite. Consequently, every star-finite cover is countable. This definition leads us to the following Theorem 2.3.5. Let X be a space and let U be an open cover of X. Then (1) there exists an open refinement V of U such that V is star-finite, (2) N(W) is locally finite if and only if U is star-finite. Proof. For (1), we may assume without loss of generality that A" is a subspace of the Hilbert cube Q (Corollary A.4.4). For every U 6 U let U C Q be open such that U n X = U and put U = {U : U G U}. Then V = (J U is an open subset of Q and hence is locally compact and <j-compact. There are compacta Fn C V for n 6 N such that F! C Int F2 C F2 C Int F3 C • • • C F n _i C Int Fn C Fn C • • •
while moreover V — U^Li Fn- Put F0 — 0. For every n let £ n be a finite subcollection of U with Fn C (J £ n . Now put W = {U n (Q \ F n _i) n Int Fn+1 : U G £ n , n 6 N}.
It is clear that W covers V', is star-finite and that W < U. As a consequence, the collection V = W \ X is as required. For (2), first assume that N(li) is locally finite. If there exists U G U such that the set {V G U : V n t / ^ 0 } is infinite then the vertex x(U) of N(IV) is contained in infinitely many onedimensional simplices of TV(U), which contradicts the local finiteness condition on TV (It). The verification of the reverse implication is a triviality which we leave to the reader. D We shall now prove that every space can be 'approximated arbitrarily closely' by a polytope. Corollary 2.3.6. Let X be a space. For every open cover U of X there exists a polytope P and a U-mapping f : X — > P . If X is compact then P can be chosen to be a polyhedron. Proof. Let U be an open cover of X. By Theorem 2.3.5(1), there exists a star-finite open cover V of X which refines U. Let
K: X ^ \N(V)\ be the ^-mapping of V. Since V is star-finite, |./V(V)| is locally finite by Theorem 2.3.5(2). By Corollary 2.3.4, K is a V-mapping. Since V < U, we are done. If X is compact then V is a finite subcover of U and proceed as above. D
2.3. THE NERVE OF AN OPEN COVERING
137
Freudenthal's Approximation Theorem. We now come to an interesting 'Approximation Theorem'. See §1.10 for results that are in the same spirit. Freudenthal's Approximation Theorem 2.3.7. Every compact space is homeomorphic to the inverse limit of an inverse sequence consisting of polyhedra. Proof. We start as in Example 1.10.5. By Corollary A.4.4 we may restrict ourselves to a compact subspace X of Q. For every n let Xn = 7Tn[X\, n
where rn: Q —> JJ is the projection. In addition, let fn: Xn+i —>• Xn be defined by fn(xi,...,xn+i) = ( x i , . . . , a ; n ) . Then ]^m(XnJn)n = X (Example 1.10.5). Unfortunately, the sets Xn need not be polyhedra. It will require a little extra work to take care of this problem. Claim 1. Let X be a compact subspace of JP for some n. Then for every neighborhood U of X there exists a polyhedron P such that X C P C U. Proof. By Corollary A.5.4 there exists 6 > 0 such that if A C J n , diam A < 6 and AnX / 0 then A C U. Since J n is a polyhedron, it has by Theorem 2.2.10 a triangulation § such that for every cr E S we have diamcr < 6. Let P be the union of all simplices in S that intersect X. An easy check show that P is as required. 0 Now for every n G N we shall define a polyhedron Pn in JP with Xn C Pn, as follows. Let PL = JT. By Claim 1 there exists a polyhedron P2 in ]J2 such that X2 C P2 C B(X2, V 2 ). By the Dugundji Theorem 1.2.2, we can extend the function /i: X? —>• X\ C P! to a continuous function £1: P2 —> PI . For PS we have to do a little extra work. Since P2 is an A N R (Theorem 2.1.24), there is a neighborhood U of X% such that the function /2 : X3 —> X? ^ P2 can be extended to a continuous function 77: U —>• P2. Pick a neighborhood F C C7 of Xs such that r)[v] c B(x2, y3), & o
138
2. BASIC COMBINATORIAL TOPOLOGY
We think of Lhn(P n , f n ) and lim(Xn, / n ) = JC as subspaces of claim that Um(P n , £ n ) = ]im(Xn, fn) = X. It is clear that )jm(Xn,fn)
C h'm(P n ,£ n ).
Now take an arbitrary x E Hm(P n ,£ n ) and fix TV e N. Then,
m=N
and hence XN € XN by (*). Since £n extends fn for every n, this easily implies that x £ Hm(JC n ,/ n ). D
Exercises for §2.3. 1. Let X be a space and let It and V be open covers of X such that V < U. For every V £ V pick J7(V) e U such that V C f/(V). This defines a function / 0 : A^(V) (0) -+ N(U)W. Extend this function 'linearly' over each simplex of N(V) and prove that the resulting function from |W(V)| -^ \N(U)\ is continuous. 2. Prove that the following statements for a space X are equivalent: (1) X is compact. (2) Every locally finite open cover of X is star-finite. 2.4. Simplices in En In this section we shall formulate and prove some fundamental properties of simplices in W1 . Among other things, we shall present a proof of the Brouwer Fixed-Point Theorem and some of its applications. Lemma 2 A.I. Let a\ and a-2 be two (geometric) n-simplices in W1 which intersect in a common (n — l)-face T. Then a\ U cr2 is a neighborhood of the barycenter of T. Proof. Observe that the lemma is trivial if n — 1; we may therefore assume that n > 1. Let d\ = {a 0 ,ai,. . . ,a n _i,a^} and a2 = {a 0 ,ai, . . . ,a n _i,a^}, where The points {ao5 &i, . . . , a n-i; a n} are geometrically independent, from which it follows by Theorem 2.1.3 that the points — ao, 02 —
2.4. SIMPLICES IN Rn
139
are linearly independent. Now consider the standard basis ei, 62, . . . , en in W1 and a linear isomorphism gl : W1 —> W1 with the properties: gl(ei) = di - a0 (i < n), g1(en) = a1n-a0. This isomorphism followed by the translation x H-» x + ao is an affine isomorphism (Theorem 2.1.7) and hence a homeomorphism (see Page 113) which we denote by f1. Observe that f l has the following properties: '/1(Q)=a0,
In the standard basis we now change en into — en. By a similar argumentation as above we obtain an affine isomorphism / 2 : Mn —> Mn such that /2(Q)=a0,
Observe that fl and /2 agree on the plane P = {x £ Mn : xn = 0}. Also observe that /* f F = / 2 \ P is an affine isomorphism between P and the hyperplane spanned by the elements of r = (a 0 , ai, . . . , a n -i}- Now define a homeomorphism / : ffin —>• Mn as follows
Let b be the barycenter of r and let c = (Y n , Yn, . . . , Yn, 0). Then /(c) = 6 and the set is a neighborhood of c of which of which it is geometrically obvious that it is mapped by / into a\ U a? (see Exercise 2.4.1). We conclude that a\ U a^ is a neighborhood of b. D The next result is the key in the proof of the Brouwer Fixed-Point Theorem. Theorem 2.4.2. Let r be an n-simplex in W1 , and let S be a finite triangulation of r which subdivides its standard triangulation. Finally, let [i £ S be an (n — 1) -simplex and let b be the barycenter of//. (1) Ifb^dr then p, C dr and fj, is a face of precisely one n-simplex in §. (2) Ifb^dr then // is a face of precisely two n-simplices in §. Proof. Assume that b E dr. We will show that p, C dr. Since S subdivides the standard triangulation of r there is a simplex a G § with b € a such
140
2. BASIC COMBINATORIAL TOPOLOGY
that a is contained in a proper face of r, i.e., a C dr. By Lemma 2.2.1 we obtain that ju is a face of a so that // C dr. Now we shall prove that depending on whether b £ da or not, w is a face of at most one or at most two n-simplices in S. To this end, first assume that b G dr. If there exist two n-simplices in § having // as a common face then the union of these two simplices is a neighborhood of b (Lemma 2.1.5). But this contradicts the fact that b G dr which is equal to the topological boundary of r (Exercise 2.1.9). Ifb^dr and there exist three distinct n-simplices, say <Ji, <72 and a% G S, having IJL as a common face then since S is a simplicial complex, (<TI U cr2) H (cr2 U cr3) D (<73 U <TI) = (<TI H cr2) U (<72 n cr3) U (cr3 n <TI)
For the last equality observe e.g., that IJL is an (n — l)-simplex contained in the simplex a\ H <72. If a\ Pi a2 is an n-simplex then <j\ = <J2 which is a contradiction. So a\ fl cr2 is an (n — l)-simplex since it contains n as a face, hence it must be equal to fi. By Lemma 2.4.1 the above intersection is also a neighborhood of b in W1 . Consequently, // is a neighborhood of b in Mn which is impossible because JJL is contained in an (n — l)-dimensional hyperplane. For the remaining part of the proof, put
Observe that U is a nonempty relatively open subset of r. We shall now prove that // is a face of at least one n-simplices of 8. Striving for a contradiction, assume that there does not exist an n-simplex in S having IJL as a face. According to Lemma 2.2.1 this implies that IJL is the only simplex in S containing b which implies that U C IJL. Now if b 6 dr then we arrive at the desired contradiction because then U C ^ C dr and dr has empty interior in r. If b 0 dr then U n Int(r) is a neighborhood of b € Kn which is contained in yu. But this contradicts the fact that // is contained in an (n — l)-dimensional hyperplane of W1 . All there remains to prove is that if b ^ dr then there are at least two n-simplices having IJL as a face. Striving again for a contradiction, assume that this is not the case. We already know from the above that there is an n-simplex a e S having p as a face. Pick an arbitrary p e U and let 7 G S be such that p G 7. Then & G 7 by the definition of U and so by the same argumentation as above, JJL is a face of 7. But then either 7 = p or 7 = a. In either case, p £ a. We conclude that U C a. So Int(r) n C/ is a neighborhood of & G Mn which is
2.4. SIMPLICES IN R"
141
contained in a. Since )U is a proper face of a we also have b £ 5cr. This is a contradiction. D Let T — \{XQ, . . . ,Xk}\ be an arbitrary fc-simplex in E n . An m-Sperner map for r is a function from the vertices of sd^ m ^(r) to {0, . . . , k} such that if v is such a vertex and v € \{xio,...,xit}\ then /i(u) € {i 0 ,...,«*}. We call a fc-dimensional simplex a in sd^ m ^(r) full if
There are always full simplices as the following result shows. As to be expected, its proof is of a combinatorial nature. Sperner's Lemma 2.4.3. Let T = \{XQ, . . . ,Xk}\ be a k-simplex in Rfc and let h be an m-Sperner map for T. The number of full simplices in sd'm^(r) is odd and hence non-zero. Proof. We shall prove the lemma by induction on k. If k = 0 then T — \ {XQ}\ , and there is one full simplex. Assume that the theorem is true for all (k — 1)simplices, k — I > 0. We shall prove the theorem for T. Put /3 = {x0,...,xk-i}, 3>i = {cr e sd(m}5(T) : a is (k - l)-dimensional, h[cr] = {0, . . . , k - 1}}, ?2 = {a e sd (m) 3 r (r) : a isfc-dimensional,{0, . . . , k - 1} C h[a}},
respectively. Observe that C
By the definition of an m-Sperner map it therefore follows that the restriction of h to the vertices of sd^ m ^9 r (^) is an m-Sperner map for ft. As a consequence, the collection J*i (0) is the set of full simplices in the m-th barycentric subdivision of |/?|, so |Ti(0)| is odd by our inductive assumptions. Also, ^(O) is the set of full simplices in the ?n-th barycentric subdivision of r; so we have to prove that ^(O)! is an odd number. Put R = {(«, ju) € 3*1 x y2 '• K is a face of p,}.
142
2. BASIC COMBINATORIAL TOPOLOGY
We compute the cardinality of ft twice. For each K G CPi put #[«;] = {/JG ?2 : ( K ,/x) G ft}. Claim 1. |ft| - |0>i ( 0 ) | + 2 -|Ti(l)|. Proof. Clearly, jft| = £ |ftM|. «£?!
We claim that if K E ^i(O) then |ft[ft]| = 1. Simply observe that such K is a face of precisely one fc-dimensional simplex in sd^ m ^9 r (r) by Theorem 2.4.2, and this simplex clearly belongs to J*2. We next claim that if K G ^1(1) then |ft|V|| = 2. For take an arbitrary K G ?i(l) and assume first that it is contained in a (k — l)-dimensional face of r, say K C |{#;0, . . . , X i k _ 1 } \ . Then Ji[«] C { z o , . . . , « & - i } which implies that {^o, . . . , i/t-i} = {0, . . . , k — 1} and so K C |/?|, contradiction. So such a AC is not contained in a (k — 1) -dimensional face of r, and therefore, again by Theorem 2.4.2, is a (k — 1) -dimensional face of precisely twofc-simplicesin sd^S^r), and these simplices clearly belong to 3V ^ For each /z G O^ put ft- 1 ^] = { K e O ) 1 : (K,H) eft}. Claim 2. |fl| = | y 2 ( 0 ) | + 2 - | ? 2 ( l ) | . Proof. Clearly,
Take /z G T 2 (0). We claim that l^"1^]! = 1. Indeed, h[fi] = {0, . . . , k}, i.e., the function h is one-to-one on fi. So p, has exactly one (k — 1) -dimensional face K such that { 0 , . . . , f c — !} = /&[«] and this face is clearly the only element of Ift"1^]!- If V e y 2 (l) then /i[/t] = (0, . . . , A; - 1}. Since /i has size fc + 1, there exist precisely two subsets E and F of fi of cardinality fc with the property that Consequently, there exist precisely two (fc — l)-dimensional faces K± and ^2 of IJL with /I[KI] = {0, . . . , k — 1} = h[kz\. These simplices are obviously the only elements of Ift" 1 ^] | . 0 We find that
is even. As was remarked at the beginning of the proof, |CPi(0)| is odd, hence sois|T 2 (0)|. D
2.4. SIMPLICES IN R™
143
This result has interesting consequences. Lemma 2.4.4. Let T be a k-simplex in Rk , say r = \{XQ, . . . , Xk}\- In addition, let {Fi : 0 < i < k} be a collection of closed subsets ofW1 such that for every {IQ, . . . , ii} C {0, . . . , k} we have
Proof. If not, then r — (T \ F0) U • • • U (r \ Fk). Let e > 0 be a Lebesgue number for this covering (use that T is compact and apply Lemma A. 5. 3). By Corollary 2.2.9 we can choose ra so large that mesh(sd^ m ^3 r (r)) < e. Let V be the set of vertices of elements of sd^ m ^(r). Let v € V and put Iv = (0 < i < k : a»(u) > 0} (here a0 (v) , • • • , o.k (v ) are the affine coordinates of v with respect to x0 , . . . , Xk of course).
Figure 8. Let Iv = { I Q , . . . , i i } . Then v G \{xltt,...,xit}\CFi0\J---\JFit. So for v G V we can pick h(v) £ {0,.... A;} such that a / l ( u ) ( v ) > 0 , veFh(v). Then h is an m-Sperner map for T. By Theorem 2.4.3 there exists a full simplex
144
2. BASIC COMBINATORIAL TOPOLOGY
Proof. Let T = \{XQ, . . . , xn}\ be an n-simplex in W1 . Then r is homeomorphic to I n (Exercise 1.1.24) and so it suffices to prove the Theorem for T. To this end, let / : T —> T be continuous. For i — 0, 1, . . . , n let Fl = {xer:ai(f(x))
(here (XQ(X), . . . , an(x) are the affine coordinates of x with respect to XQ, . . . , xn of course) . Let | {xi0 ,...,Xit}\ be a face of r and x € | {xi0 , . . . , Xit } \ . Then for an element i i0 . . . ^
so that "zo (&) + • • • + QV (x) = l> alo (f(x))
+ • • • + a», (/(a;))
and hence there must be 0 < j < i with alj(x)
>aii(f(x)).
Consequently, x € Fij . We conclude that \{xi0 , . . . , x^ }\ C F;0 U • • • U Fit . Each Fi is closed by continuity of / and the functions a; (Theorem 2.1.8). So by Lemma 2.4.4 there exists x 6 fllLi -^- Then for every 0 < i < n,
and n
n
i=l
i-l
This implies oti(x] — «;(/(#)) for every 0 < i < n and consequently, x is equal to /(x). D Corollary 2.4.6. The Hilbert cube Q has the fixed-point property. Proof. Let /: Q —>• Q be continuous. For every n G N define J\ — JT*£^I/*ITI-I nr \ — i T i T M i T T T* 1 \\ -i\n — \-*' ^ ^ • V"^l' • ' • i -^n) — \J V x ) i i • ' • •> J V 1 ^ ) n ) ) •
It is clear that for every n the set Kn is closed in Q and that Kn+i C Kn. Fix n 6 N, let pn: Q —> J n denote the projection and define a continuous function fn: J n -> J n by By Theorem 2.4.5, fn has a fixed-point, say ( x i , . . . , x n ), from which it follows that ( x i , . . . , a ; n , 0 , 0 , . . . ) G ATn.
2.4. SIMPLICES IN Rn
145
We conclude that the Kn's form a decreasing collection of nonempty closed subsets of Q so that by compactness of Q,
n=l
It is clear that every point in K is a fixed-point of /.
D
Corollary 2.4.7. Let X be a compact ANR. If/: X —> X is nullhomotopic then f has a fixed-point. In particular, every compact AR has the fixed-point property. Remark 2.4.8. In view of this result, the following question naturally arises. Let X be a compact ANR and let / : X —> X be continuous. Assume that / is homotopic to a continuous function g: X —>• X such that g has a fixed-point. Does / have a fixed-point? The answer to this question is in the negative, see Exercise 2.4.7. So in some sense, Corollary 2.4.7 is 'best possible'. Proof. We may assume that X is a subspace of Q (Theorem A. 4. 4). Since every constant function X —> X is extendable to a constant function Q —> X, by Corollary 1.4.3 it follows that / can be extended to a continuous function Since by Corollary 2.4.6, Q has the fixed-point property, / has a fixed-point, say x. Since f[Q] C X, x belongs to X, and since / extends /, we conclude that x is a fixed-point of /. Since by Theorem A. 12.4 every compact AR is contractible, the second part of the Corollary immediately follows from the first part and Proposition A. 12. 2. D Remark 2.4.9. In §3.14 we will show by elementary but remote methods that Brouwer's Fixed-Point Theorem for I3, implies the Brouwer Fixed-Point Theorem for all dimensions. This seems to show that the 'real' power of the Brouwer Fixed-Point Theorem is already attained at dimension three. This is a rather curious phenomenon and it is not clear whether there is a direct route from Brouwer in I 3 to Brouwer in all dimensions. This is of course only of interest for 'philosophical' reasons. The proof of Sperner's Lemma 2.4.3 is by induction, so it is equally complicated in all dimensions (except for the dimensions 1 and 2). Similar remarks can be made for the various other proofs that exist of Brouwrer's Theorem in the literature. This seems to indicate that even if one finds a more direct route from Brouwer in I 3 to Brouwer in all dimensions, this will not result in a simpler proof of Brouwer's Theorem. But this does not mean that clarifying the rather curious phenomenon that we will encounter in §3.14 would not be interesting. We finish this section by presenting three applications of the techniques developed in this section.
146
2. BASIC COMBINATORIAL TOPOLOGY
Application 1: The No-Retraction Theorem. It is clear that the function r: JJn —> Bn defined by r(x)
~
1
-
"|o;|| < 1).
n
is a retraction. Consequently, B has the fixed-point property by Theorem 2.4.5 and Exercise A.12.10(2) (observe that in fact I n and Bn are homeomorphic, cf. Exercise 1.1.24). Theorem 2.4.10 (No-Retraction Theorem). For every n E N, §n-1 is not a retract of Bn. Proof. To the contrary, suppose that Sn-1 is a retract of Bn. As was remarked above, Bn has the fixed-point property. Consequently, §™-1 has the fixed-point property by Exercise A.12.10(2). However, the antipodal mapping on S™^ 1 clearly demonstrates that §n-1 does not have the fixed-point property. Contradiction. D Corollary 2.4.11. No Sn is contractible. Proof. Suppose that §n is contractible. Then it is an AR by Corollaries 1.2.13 and 1.4.5. Consequently, there exists a retraction r: Bn —> S71"1, which contradicts Theorem 2.4.10. D Application 2: The Theorem on Partitions. The following result is the basis for dimension theory (see Chapter 3). The Theorem on Partitions 2.4.12. Consider J n and for i
2.4. SIMPLICES IN R rl
147
and for every i < n,
C 5i, 0[£i] C A,. Therefore, g has no fixed-point, which contradicts Theorem 2.4.5.
D
Corollary 2.4.13. Consider the Hilbert cube Q and its opposite faces W-1 ={xEQ:Xi = -1},
W} = {x 6 Q : Xl = 1}.
1
If Ci is a partition between W^ and W^ for every i then HiLi ^ ^ $• Proof. For every m, define /m : JIm —> Q by j m \-Eli---i-Em) — \'E'i i • • • i • E m i ' J i U ) . . . ) .
Then /TO is clearly an imbedding. It is easily seen that f^i^i] is a partition between A; and £; for every « < m. Consequently, by Theorem 2.4.12,
By compactness of Q we therefore obtain that oo
c 0
n **'
i=l
which is as desired.
D
Application 3: The Non-Homogeneity Theorem. Theorem 2.4.14. Let n 6 N. TLen (1) if A C S71"1 then Bn \ A is contractible. (2) if A C 5n \ $n~l is nonempty then Bn\A is not contractible. Proof. (1) Define H: (Bn \ A) x I ->• Bn \ A byfl"(a;,*)= (1 - ^)x. It is easily seen that H is a contraction. (Alternatively, apply Exercise 1.2.5 and Theorem 1.4.4.) (2) Without loss of generality, 0 6 A. Assume to the contrary that H:(Bn\A)xl^Bn\A is a contraction. Define F: S71"1 x I ->• §n-1 by
Then F contracts §n~1 to a point which contradicts Corollary 2.4.11. Since JJn and 5n are homeomorphic, cf. Exercise 1.1.24, this yields Corollary 2.4.15. Let n 6 N. Then ln is not homogeneous.
D
148
2. BASIC COMBINATORIAL TOPOLOGY
Exercises for §2.4. Let X be a space. We say that a closed set A in X separates X if X \ A is not connected. ^•1. Make the 'it is geometrically obvious' part in the proof of Lemma 2.4.1 precise. 2. Prove that the 'No-Retraction Theorem', the 'Brouwer Fixed-Point Theorem' and the fact that no S n is contractible are equivalent statements, in the sense that they are easily deduced from each other. ^•3. Let C be a compact convex subset of a norrned linear space L. Prove that for every e > 0 there exists a map fe: C —>• C such that (1) 0 ( / e , l ) < e , (2) fs\C] is contained in a finite dimensional linear space subspace of L. 4. Let X (1) (2) Prove
be a space and let /: X —>• X be a map with the following properties: the closure of f[X] is compact, for each e > 0 there exists x 6 X with Q(X, f(x}] < e. that / has a fixed-point.
5. Let C be a compact convex subset of a normed linear space L. Prove that C has the fixed-point property. (This is called the Schauder FixedPoint Theorem.) 6. Prove that the antipodal map S™ —>• Sn is not nullhomotopic. 7. Give an example of a compact ANR X having a homotopy H: X x I —>• X such that HQ is the identity and HI is fixed-point free. ^•8. Let A C Rn be compact, where n > 1. Prove that if A separates M71 then there is a continuous function /: A —>• S n-1 which is not nullhomotopic. 9. Let A C Sn be compact, where n > I . Prove that if A separates Sra then there is a continuous function /: A —> §n-1 which cannot be extended over S™. 2.5. The Lusternik-Schnirelman-Borsuk theorem The Brouwer Fixed-Point Theorem which was proved in the previous section, is one of the most central results in topology. We presented a detailed proof of it for two reasons. First of all, its proof uses subdivisions of triangulations of simplices. Getting used to this is important because that will be used extensively in ANR-theory later. Secondly, it is the basis for dimension theory. A more powerful result is the so-called Lusternik-Schnirelman-Borsuk Theorem which can be proved along the same lines using a somewhat more complicated combinatorial lemma than the one of Sperner (Lemma 2.4.3). We will not present a proof of the Lusternik-Schnirelman-Borsuk Theorem here since it is used only twice in this book and for our purposes its proof does not give more insight than the proof of Sperner's Lemma.
2.5. THE LUSTERNIK-SCHNIRELMAN-BORSUK THEOREM
149
The Lusternik-Schnirelman-Borsuk Theorem 2.5.1. If M is an arbitrary closed covering of S n such that |M| = n + 1 then at least one element M € M contains a pair of antipodal points. For a nice combinatorial proof, see DUGUNDJI and GRANAS [143, Theorem 4.4]. The Borsuk-Lusternik-Schnirelman Theorem implies many deep theorems about the topology of the Euclidean spaces E n . Exercises for §2.5. A map /: Sm —> S™ is called antipodal-preserving provided that f(—x) = —f(x) for every x G STO. 1. Prove that the Lusternik-Schnirelman-Borsuk Theorem is sharp by showing that there exists a closed covering M of Sra such that |M = n + 2 while no M G M contains a pair of antipodal points. ^•2. Prove that the following statements are equivalent in the sense that they are easily deduced from each other: (1) The Lusternik-Schnirelman-Borsuk Theorem for S n . (2) There is no antipodal-preserving map /: S^-^S"" 1 . (3) (Borsuk Antipodal Theorem) An antipodal-preserving map /: S^^S71'1 is not nullhomotopic. (4) (Borsuk-Ulam Theorem) Every continuous /: Sn ->• Rn
sends at least one pair of antipodal points to the same point. 3. Prove that the Lusternik-Schnirelman-Borsuk Theorem implies the Brouwer Fixed-Point Theorem.
This page intentionally left blank
CHAPTER 3
Basic dimension theory Dimension theory enables us to assign to every topological space X a number, dimX, in {-l,0,l,...}U{oo} having, among other things, the following properties: (1) if X and Y are homeomorphic spaces then dim X = dim Y, (2) dimR n = n for every n € N. So dimX is a topological invariant of X, and by (2) it distinguishes between the euclidean spaces M n , n G N. A space X for which dim X < oo is called finite dimensional, and a space is in finite- dimensional if it is not finite dimensional. The aim of this chapter is to present some basic results from dimension theory. Some of the presented results are classical: they mostly concern finite dimensional spaces. However, during the last decades significant contributions were made concerning the topology of infinite-dimensional spaces and the topology of hereditarily indecomposable continua. We shall also present some of these results in detail. 3.1. The covering dimension Let X be a space and let F be an index set. The Theorem on Partitions 2.4.12 motivates the following definition: a family of pairs of disjoint closed sets r = {(Ai,Bi) : i £ F} of X is called essential if for every family
{Liiier}, where L z is an arbitrary partition between Ai and Bi for every i, we have
if r is not essential then it is called inessential. So Theorem 2.4.12 and Corollary 2.4.13 show that I n has an essential family of n pairs of disjoint closed sets and Q has essential families of size n for every n < NO- It follows easily that every subfamily of an essential family is again essential (Exercise 3.1.1). 151
152
3. BASIC DIMENSION THEORY
Theorem 3.1.1. I n has an essential family consisting of n pairs of disjoint closed sets, but every family consisting of at least n + l pairs of disjoint closed sets is inessential. Proof. By the above remarks we need only to consider the second part of the theorem. Let A and B be disjoint closed subsets of I n and let E C I be dense. Claim 1. There is a partition D between A and B such that
D C {x e I n : (3i < n)(xi € E}}. Proof. This is easy. Every point x € A has a neighborhood of the form n IJfatj&i], i—l
with o,i,bi E E U {0, 1} for every i < n, which misses B. There is a finite family 3 of these neighborhoods whose union covers A, and the boundary D of this union is contained in the union of the boundaries of the elements of f. We conclude that D is the required partition between A and B. <0 Now let r = {(Ai,Bi) : i < n + 1} be an arbitrary family consisting of n+l pairs of disjoint closed subsets of ln. We will prove that T is inessential. Indeed, there exist n + l pairwise disjoint dense subsets of I. One can take for example
Q) nl, E2 = ( > 3 + Q ) n l , E3 = ( \ 5 + Q) nl, . . . , respectively. By the above there exist partitions Di between Ai and Bi such that DiC{x<Eln: ( 3 j < H}(XJ G Er)} (i < n + 1). Since the Ei are pairwise disjoint, clearly HlLi Di = 0.
D
Observe that in Theorem 3.1.1 we formulated a topological property of I n shared by no ITO for m ^ n. In particular we obtain: Corollary 3.1.2. Let n,m € N. Ifn^m
then In ^ lm.
These remarks suggest the following definition: for a space X define its covering dimension, dimX 6 { — 1,0, 1, . . .} U {oo}, by
dim X < n
<&
dim X = n dim X = oo
<£> <£>
every family of n + 1 pairs of disjoint closed subsets of X is inessential, dim X < n and dim X ^ n — I , dim X / n for every n > — 1 .
3.1. THE COVERING DIMENSION
153
A space X with dimJY" < oo is called finite dimensional] if it is not finite dimensional then it is called infinite-dimensional. Observe that dim X > n if and only if there is an essential family consisting of n pairs of disjoint closed subsets of X. So in a sense, there are n different 'directions' in X. Also observe that it is not clear at all why we call dimX the covering dimension of X: the term partition degree seems to be more appropriate. We will explain our terminology later. It is easy to see that if X and Y are homeomorphic spaces then X and Y have the same covering dimension. In our new terminology, Theorem 3.1.1 reads as follows: Theorem 3.1.3. diml n = n for every n. In §1.5 we called a space zero-dimensional if it is nonempty and has a base consisting of clopen sets. In Corollary 1.5.2 it was shown that a space X is zero-dimensional if and only if 0 is a partition between any pair of disjoint closed subsets of X. So, in our new terminology, a space X is zero-dimensional in the sense of §1.5 if and only if dimX = 0. It will turn out that zero-dimensional spaces are, in a sense, the 'building blocks' of all the other finite dimensional spaces. Without too much trouble, the proof of Theorem 3.1.1 can be adapted to show that dimR" = n for every n. This equality however will turn out to follow trivially from Theorem 3.1.1 and results to be derived in §3.2. For that reason we will not verify it here. For later use we shall now study some elementary properties of essential families of pairs of disjoint closed sets. The following simple result is fundamental for dimension theory since it allows to extend partitions in subspaces to partitions in the whole space. Lemma 3.1.4. Let Y be a subspace of a space X. In addition, let A and B be disjoint closed subsets of X. If U and V are open neighborhoods of A and B, respectively, having disjoint closures, and if S is a partition in Y between Yr\U and Y Pi V, then there is a partition T in X between A and B such that T n Y C S. Proof. Write Y \ 5 as the disjoint union of two open (in Y) sets E and F such that
YnucE, Ynv CF. Since E n V = 0 we obtain E n B = 0, and similarly F n A = 0. From this it follows easily that A U E and B U F are separated. By Corollary A.8.2 there exist disjoint open neighborhoods U' and V of A\JE and B(JF, respectively. Clearly, T = X \ (U1 U V) is as required. D For a closed subspace we can do a little better.
154
3. BASIC DIMENSION THEORY
Corollary 3.1.5. Let Y be a closed subspace of a space X. In addition, let A and B be disjoint closed subsets of X. If S is a partition in Y between the sets Y n A and Y n B, then there is a partition T in X between A and B such that T n Y C S. Proof. Write Y \ S as the disjoint union of two open (in Y) sets E and F such that YnACE,
YHB CF.
Observe that S U F U B is closed in X and that A n (S U F U B) = 0. By Corollary A. 4. 3 there is an open neighborhood U of A in X such that £7n(SuFu£) = 0. By a similar argument it is possible to find an open neighborhood V of B in X such that
Vn(SuEuu) =0. By construction, S is a partition between U C\Y and F n Y. Now apply Lemma 3.1.4. D Corollary 3.1.6. Let Y be a zero- dimensional subspace of X . Then for all disjoint closed subsets A and B of X there exists a partition S between A and B in X such that S n Y = 0. Proof. Let A and B be disjoint closed subsets of X and let U and V be disjoint neighborhoods of A and B, respectively, such that UnV = 0 (Corollary A. 4. 3). Since Y is zero-dimensional, 0 is in Y a partition between the sets U n Y and V n Y. Now apply Lemma 3.1.4. D Corollary 3.1.7. If a space X is the union of at mostn + l zero-dimensional subspaces then dim X < n. Remark 3.1.8. We will prove later in Corollary 3.3.9 that the converse of this result is also true; so a space is at most n-dimensional if and only if it is the union of a family of n + I (not necessarily pairwise distinct) zerodimensional subspaces. Proof. Let X = \J^ Xl with dimXi < 0 for i < n + I . Let be an arbitrary family of n + 1 pairs of disjoint closed subsets of X. By Corollary 3.1.6, for each i < n + I there exists a partition L% between Ai and Bi in X such that LiCiXi = 0. Then f}™=i Li = 9 so that T is inessential. We conclude that dim X
3.1. THE COVERING DIMENSION
155
Theorem 3.1.9. Let X be a space, let {(Al^Bi) : i € F} be essential in X, and let F0 be a proper subset of T. If for every i £ F 0 , Li is a partition between Ai and Bi and L= p| Lz ier 0 then is essential in L. Proof. For i (£ TO let £^ be a partition in L between LnAi and L n Bi. By Corollary 3.1.5, for i £ F0 we can find a partition F; in X between Ai and ^ such that Fl H L C E^ Then
,n
e L n Pi E,= f|
which is as required.
D
Corollary 3.1.10. Let X be a space and let n > 0. If dim X > n then there exist disjoint closed subsets A and B of X such that if L is a partition between A and B then dim L > n — 1. Proof. If n — 0 then there is nothing to prove. So assume that n > 1. Since dim X > n there exists an essential family in X. We claim that AI and BI are as required. Indeed, let L be a partition between AI and BI. Observe that L ^ 0 since T is essential. We consider two cases. If n = 1 then since L / 0, dim!/ > 0 and so we are done. If n > 1 then by Theorem 3.1.9, L has an essential family of cardinality n — 1, namely the collection {(L n A^ L n Bi) : i = 2, . . . , n}. And so dim L > n — 1 . D Exercises for §3.1. A space X is called countable dimensional if it can be written as the union of countably many zero-dimensional subspaces (cf. Corollary 3.1.7). In addition, a space X is called strongly infinite-dimensional if it has an infinite essential family. See e.g., §3.13 for more information on these notions. Let X be a space with disjoint closed subsets A and B. We say that a closed subset S of X is an irreducible partition between A and B if S is a partition between A and B while no proper closed subset of S shares this property. 1. Prove every subfamily of an essential family is again essential.
156
3. BASIC DIMENSION THEORY
2. Let r = {(Ai, Bi) : i < n} be a family of pairs of disjoint closed subsets in the space X. Prove that if there exist different indices i-,j
{(AinE,Bir\E) :ier} is inessential in E. ^•9. Let X be a compact space and let r = {(Ai,Bi) : i 6 F} be an essential family of pairs of disjoint closed subsets of X. Prove that there is a component C of X such that the family r \C = {(AlnC,BinC)
:i€r}
is essential in C. Conclude that every strongly infinite-dimensional compact space has a strongly infinite-dimensional component. 10. Prove that 21, the hyperspace of I, is infinite-dimensional. 11. Let /: X —>• Y be continuous, and let r = {(Ai, Bi} : i 6 /} be a collection of pairs of disjoint closed sets in X. Assume that the collection consists of pairs of disjoint sets and is inessential. Prove that r is inessential. 12. Let /: X —> Y be continuous, and let T = {(Ai,ft) : i € /} be an inessential collection of pairs of disjoint closed sets in Y. Prove that the collection f = {(/-1[Ai],/-1[ft]):t€/} is inessential in X.
3.2. TRANSLATION INTO OPEN COVERS
157
13. Let X be a space and let {(Ai, Bi) : i £ F} be essential in X. Suppose that the set y C X is such that y fl P|i£r L; 7^ 0 for any choice of partitions Li of Ai and Bi in X. For each i E F let f/i and Vi be disjoint closed neighborhoods of Ai and Bi, respectively. Prove that
{(t/iny,v;ny): i e r} is essential in y. 14. Let X be a zero-dimensional space containing the pairwise disjoint closed subsets A, B and C'. Prove that C is a partition between A and B. (So in a zero-dimensional space every irreducible partition is empty.) ^•15. Give an example of a space X containing two disjoint closed subsets A and B such that no partition between A and B is irreducible. ^•16. Prove that the Brouwer Fixed-Point Theorem and the fact that there is a space X with dim X = oo are equivalent statements, in the sense that they are easily deduced from each other. (This complements the list in Exercise 2.4.2 with an additional statement.) 3.2. Translation into open covers We defined the covering dimension dim X in terms of properties of the family of all closed subsets of X. In this section we see that dimX is also describable in terms of properties of the family of all open covers of X. As an application of the results derived here we present a simple proof that W1 is n-dimensional for every n. We conclude from this that W1 and RTO are homeomorphic if and only if n = m. Let X be a space and let A and 13 = {B(A) : A e A} be families of subsets of X. We say that 13 is a swelling of A if (1) for every A e -A, A C B(A), (2) for every finite J C A, f| 3~ = 0 iff [}A^B(A] = 0. Observe that if B(Ao),B(Ai) G 13 are distinct then so are A0 and AI, but the converse need not hold. Before we state the following proposition, let us recall that a locally finite collection is countable (see the remark on Page 486 following the definition of locally finite collection). Proposition 3.2.1. Let 3 be a locally finite collection of closed subsets of a space X. Then 3 has a swelling consisting of open subsets of X. Proof. We can enumerate 3 as {F; : i £ N}. Without loss of generality, assume that F\ = 0. By induction on i £ N we shall construct an open neighborhood U% of F; such that
158
3. BASIC DIMENSION THEORY
is a swelling of 3. Put U\ = 0, and assume that for some i the sets U\ , . . . , have been constructed. Observe that 3i is locally finite. Put finite subcollection of 3l and ( P) £) n Fl+l = 01. Observe that 23 is locally finite as well and put B = \J eS>. Then B is closed by Exercise A. 7.1 and clearly B D Fi+\ — 0. By Corollary A. 4. 3 there consequently exists an open neighborhood Ui+i of FJ+I the closure of which misses B. It is clear that the set Ui+\ is as required. We claim that {Ui : i G N} is a swelling of 3. To this end, take arbitrary «'(!)) . . . ,i(n) G N and assume that D?=i ^i(j) = ®- Let m = max{z(l), . . . , i(n)}. Then {Ui, . . . , Um} is a swelling of {Fi, . . . , Fm} by our construction. We conclude that f|™=1 C/i(j) = 0 . D Corollary 3.2.2. Let 3 be a locally finite family of closed subsets of a space X. Also, for every F G 3 let V(F) be a neighborhood of F. Then there exists a swelling {U(F) : F 6 3} of 3 consisting of open subsets of X such that for every F e 3 we have U(F) CV(F). Proof. By Proposition 3.2.1 there exists an 'open' swelling {W(F) : F 6 3} of 3. By Corollary A. 4. 3 there exists for every F G 3 an open neighborhood U(F) such that U(F) C V(F) n W(F). We claim that [U(F) : F e 3} is as required. To this end, let £5 C 3 be finite such that H S = 0- Then D clearly f) F€S W(F] = 0 from which it follows that fl^es U(F) = 0 " We now show how finite closed collections can be modified. Lemma 3.2.3. Let A,Ai,...,An be closed subsets of a space X. Moreover, let V\ , . . . , Vn be open subsets of X with A n Ai C Vi for every i < n. If
1=1 for i < n then
(i) A u U I U ^ c i u U I U ^
(2) innr=i^cnr =1 Fr\/ z .
Proof. Observe that (1) is trivial. For (2), simply observe that for i < n,
AnAz = ( A \ v l } n ( A l u v l ) = [(An ^ \Vi\\j [An (Vi \Viil So we are done.
D
3.2. TRANSLATION INTO OPEN COVERS
159
One should think of the previous triviality in the following way. The closed sets A are changed into the closed sets A. But this is not done in an arbitrary way because (1) tells us that what is covered by the ^4's is also covered by the A's, and (2) tells us where the intersection of the A's can be found. The results obtained so far have nice applications in dimension theory. Corollary 3.2.4. Let X be a space such that dimX < n < oo. Then for every countable open cover U of X and for every subcollection 3 C U of cardinality n + 2 there exists an open shrinking V — {V(U) : U G U} o f U having the following properties: (1) for every U G U, V(U] C V(U) C U, (2) r\U€
y= {c/i,..., un+2}. By Proposition A.7.1 there exists a closed shrinking {Bl : i 6 N} of U. For every i < n + 1 define Ai = X \ Ui. Then {(Bi,Ai) : i < n + 1} is a family of n + 1 pairs of disjoint closed subsets of X. Since dimX < n, there exist open sets Vi C X for i < n + 1 such that (3) B1CV1CV1CX\A1 = Ui, Now consider the closed sets B\, . . . , Bn+2- For i < n + I put Bi = Vi and let Bn+2 = Bn+2 \ ljr=i ^- Observe that the B's are the 'improved' sets we get from Lemma 3.2.3 by considering the collections of closed sets
and open sets
Vi,.. . , So we get by (4) that n+2
(5)
n+1
f l ^ C p| i=l
i-l
Observe that ^ — {B\ , .62 , . . . , Bn+2 , Bn+3 , -Bn+4 , • • • } covers X since \J™=i Bi C Ul^i ^- ^n addition, the first n + 2 members of 23 have empty intersection by (5). Since Bi C Ui for every i < n + 2, the desired result now follows from Corollary 3.2.2. D
160
3. BASIC DIMENSION THEORY
Let U be a cover of a space X and let n > 0 (we do not assume U to be open). We say that the order of U is at most n, ord(U) < n, if for every x 6 X, \{U £U:x£ U}\
3.2. TRANSLATION INTO OPEN COVERS
161
disjoint closed refinement, yet E is one-dimensional (as will be shown in Example 3.4.13). We do not know whether spaces that satisfy condition (8') are at most (n + l)-dimensional. Proof. We prove (1) =>• (2). Since a refinement of a refinement is a refinement, by Corollary A. 7. 3 we may assume that U is locally finite. Enumerate U as {£/o,i : i & N}. In addition, let {F(j) : j G N} enumerate the collection of all subsets of N of cardinality precisely n + 2. By Corollary 3.2.4, for j G N there exists an open cover Vj = {Uj^ : i G N} of X having the following properties: (8) for each i, Uj i C Uj i C t/j-i ;, (9) Now for each i 6 N define
Claim 1. The collection V = {Si : i G N} covers X. Proof. Take an arbitrary x G A". Since U is locally finite, x is contained in finitely many elements of U only. Pick ra G N such that x ^ Ui> m ^o,i- We conclude that for every j' G N there is an index k(j) < m such that x G f^j,fc(j) • So there exists k < m such that a; belongs to infinitely many of the C/j,fe. By (8) this implies that x G Sk, which is as required. <> We conclude that V is a closed shrinking of U and hence is locally finite. Moreover, ord(V) < n by (8) and (9). We prove (2) => (3). By (2) there exists a locally finite closed refinement § of U such that ord(S) < n. For each S G 8 pick U(S) G U containing 5. By Corollary 3.2.2 there exists an open swelling V = {V(S) : S G S} of 8 such that for every element 5 G 8 we have V(S] C U(S). Since ord(S) < n we get ord(V) < n, and so V is as required. We prove (3) =>• (5). By (3) there exists an open refinement W of U such that ord(W) < n. For each W G W pick U(W) G U be such that W C U(W). Now for each U G U define
V(U) = \J{W G W : U(W) = U}. Clearly, V = {V(U) : U G U} is an open shrinking of U and it therefore suffices to prove that ord(V) < n. Take pairwise distinct V(U1),...,V(Un+2)eV
162
3. BASIC DIMENSION THEORY
and assume that x G C\i=i V(Ui)- For each i < n+2 there exists Wi € W such that U(Wi) = Ui and x G Wi. Since the [7; are pair wise distinct, this implies that the Wi are pairwise distinct. But ord(W) < n, and so x G [^^ Wi = 0. This is a contradiction. We prove (5) =>• (4). Observe that if A is a cover of X such that ord(,A) < n and I> is a shrinking of .A then ord(!B) < n. So (4) follows directly from (5) and Corollary 3.2.4. We prove (4) =3- (6). This is a triviality. We prove (6) => (7). This follows immediately from Corollary 3.2.2. We prove (7) => (1). Let {(A;,.E?i) : i < n + 1} be a family of n + 1 pairs of disjoint closed subsets of X. Observe that n+l
n+l
p| Ai n U 5, = 0 i=l
i=l
so that the collection n+l
{x\A1,X\A2,...,X\An+l,X\ covers X. By (7) there consequently exists an open cover V = {Vi : i < n + 2} of X such that (10) ViCX\Ai
(i
(12) ord(V) < n. Let J = {Fx : i < n + 2} be a closed shrinking of V = {Vi : i < n + 2} (Proposition A. 7.1). Now for i < n + 1, define Ai and Bi by
Ai = Ai\J(Fn+2\Vi),
Claim 2. For every i < n + 1, Ai n B» = 0. Proof. Simply observe that A% n Bi — 0, that F; C V^ from which it follows that Ai n Fi = 0 and (Fn+2 \ Vi) n F» = 0 and that Fn+2 n 5» = 0. 0
Claim 3.
1
u
& = X.
3.2. TRANSLATION INTO OPEN COVERS
163
Proof. Since 3~ covers X, we are done if we show that (JF^i -^ - ^=i &i and Fn+2 C (J^/ AY The first inclusion is a triviality. For the second one observe that Fn+2 C Vn+2 and that by (12), Vn+2C\^\^ Vi = 0. We conclude that n+l
n+1
i—l
i=l
Fn+2 = Fn+2 \ p| Vi c y A, which proves the claim.
Q
By Corollary A. 4. 3 there exists for every i < n + I a partition Li between Al and Bi. Since A^ C A; and Bi C B; for every z, L; is also a partition between the sets A; and Z?;. By Claim 3 we obtain n+l
n+l
as required.
D
Naturally, the reader wonders about the relation between the dimension of a space X and the various dimensions of its subspaces. We finish this section by deriving two results about that relation. The Countable Closed Sum Theorem 3.2.8. If X can be covered by countably many closed and at most n-dimensional sets then X is at most n- dimensional. Proof. It is clear that without loss of generality we may assume that n < oo. Enumerate the closed cover "3 as {Fi : i e N}, and put FQ — 0. Let U be a finite open cover of X. By induction on i > 0 we shall construct an open cover U(i) of X such that the following conditions are satisfied: (1) U(0) = U, (2) if 0 < j < i then U(i) is a shrinking of U(j'), (3) ord(U(i) r Fi) < n.
Observe that conditions (2) and (3) are satisfied for i = 0. Now assume that we completed the construction for all j with 0 < j < i. Since dim .Ft < n, by assumption, the open cover U(i — 1) \ Fi of Fi has an open shrinking V = {V(U} : U € U(i - 1)}
of order at most n (Theorem 3.2.5). For each U 6 U(i — 1) put W(U) - (U\Fi)\JV(U). It is clear that W = {VT(C7) : t / e U ( i - l ) } is an open shrinking of U(i — 1) such that ord(W \ Fi) < n. An easy application of Proposition A.7.1 and Corollary 3.2.2 now shows that there exists
164
3. BASIC DIMENSION THEORY
an open cover U(i) of X satisfying (2) and (3). This completes the inductive construction. Let the cardinality of U be k < oo. It is clear that for i e N we can enumerate li(i) as {Umji : m < k} such that for 0 < i < j and ra < k we have Umj C Um^. Now for each x £ X there exists m(x] < k such that x belongs to infinitely many of the t/ m ( x ),;. So by construction we have
By (2) and (3) this implies that oo
oo
Um,i : m < A;} - { f| t7m,i : m < fc}
is a closed shrinking of U and has order at most n. Consequently, dim X < n by Theorem 3.2.5. D Since Mn is a countable union of homeomorphs of I n , Theorems 3.1.1 and 3.2.8 imply that dimR n < n. In the remaining part of this section we shall present, among other things, two additional proofs of this inequality which are interesting in their own rights. The Subspace Theorem 3.2.9. Let A be a subspace of a space X. Then dim A < dim.Y. Proof. If dim.Y = oo then there is nothing to prove. So without loss of generality assume that n = dim X < oo. First assume that A is closed. Observe that if r is an essential family in A then it is also an essential family in X. From this it follows immediately that dim A < n. Next, assume that A is open. Observe that A is an F^-subset of X by Exercise A. 2. 3. That dim A < n therefore follows from the above and the Countable Closed Sum Theorem 3.2.8. Finally, assume that A is an arbitrary subspace of X. In addition, let U be a cover of A by sets that are open in A. For each U 6 U pick an open subset V(U] of X with V(U] C\A = U. Put V = \JUeli V(U). Then V is an open subset of X and hence, by the above, dim V < n. Since V = {V(U) : U G U}
covers V, Theorem 3.2.5 yields the existence of an open refinement W of V with ord(W) < n. Then X = W \ A is an open refinement of U with order at most n. Another application of Theorem 3.2.5 now gives us that dim A < n, as required. D
3.2. TRANSLATION INTO OPEN COVERS
165
A different proof of Theorem 3.2.9 shall be given in Corollary 3.4.14(1)2. Observe that since W1 is homeomorphic to (0, l) n we have by Theorems 3.1.1 and 3.2.9 that dirnlP < n. The results in this section can also be used to prove that various spaces are zero-dimensional. We shall demonstrate this by the following example. For all n € N and 0 < m < n let y\.n>m — {x G W1 : exactly m coordinates of x are rational}. Proposition 3.2.10. Let n G N. Then W1 = U^ =0 ^Vm dimensional for all 0 < m < n.
and
^n,m is zero-
Proof. The first part of this proposition is trivial. For the second part, observe that if m = 0 then there is nothing to prove since 9^n,o is the product of n copies of IP. So let m > 1 and let A = |i(l), . . - , i(m)} be a set of m indices in {1, 2, . . . , n} and let q\ , . . . , qm £ Q- Put X — {x € W1 : x^) = qj for every j < m}. Observe that X is a closed subspace of R™ and that X n *Rn,m — [x 6 W1 : £;(j) = QJ for every j < m and x% G P for i £ A}. Consequently, Xr\9tnjTn is a closed subspace of 9^n,m which is homeomorphic to the product of n — m copies of F. We conclude that X n ytn,m is zerodimensional. Since lHn;TO is the union of countably many sets of the above form X n 9ln,m, we conclude that dimlH njm = 0 by Theorem 3.2.8. D Corollary 3.2.11. For all n £ N and 0 < m < n, dimja: G IRn : at most m coordinates of x are rational} < m. Proof. Observe that the set of all x E Mn for which at most m coordinates are rational coincides with So we are done by Proposition 3.2.10 and Corollary 3.1.7.
D
It therefore again follows that dimM n < n. We now come to what is sometimes called the 'Fundamental Theorem of Dimension Theory'. Theorem 3.2.12. For all n e N, dim IT = diml n = dim§ n = n. Proof. By Theorem 3.1.1, dim I™ = n. By Theorem 3.2.9 this implies that dimM n > n. Since dimM n < n, we obtain dimM n = n. Also, §n is the union of two homeomorphs of I n , so an appeal to Theorem 3.2.8 gives us that dimS n < n. It now follows easily that dim§ n = n as well. D
166
3. BASIC DIMENSION THEORY
Theorem 3.2.12 is of fundamental importance since it confirms our geometric intuition that M n , ln and S n are n-dimensional. Since the covering dimension is a topological notion, we also obtain the following Corollary 3.2.13. Ifn^m then W1 and Mm are not homeomorphic. We finish this section with three characterizations of dimension, two of which are in the spirit of Theorem 3.4.4 and one of which is 'combinatorial'. Let r be an n-dimensional simplex in a linear space L. Exercise 1.1.24 gives us that T is homeomorphic to E n , hence dim r — n. Now let X = \7\ be a polytope. Since T is countable, by the Countable Closed Sum Theorem 3.2.8 and the above remark we get dim X = supjdim r : r e T} = supjA; : r is afc-simplex,r e T}. We conclude that the topological dimension of X equals its 'combinatorial' dimension. Theorem 3.2.14. Let X be a nonempty space and let n > 0. The following statements are equivalent: (1) dimX < n, (2) for every open cover li of X there exists a locally finite open refinement V of U such that such that ord(V) < n, (3) for every open cover U of X there exists a star-finite open refinement VofU such that such that ord(V) < n, (4) for every open cover U of X there exist a polytope P such that dim P < n and a ll-mapping f : X — > P . Proof. We prove (1) =» (3). Without loss of generality, U is star-finite (Theorem 2.3.5(1)). Now by Theorem 3.2.5 there exists an open shrinking V of U such that ord(V) < n. Since each shrinking of a star-finite cover is clearly star-finite, we are done. We prove (3) => (2) => (I). That (3) => (2) is a triviality and that (2) =>> (1) follows from Theorem 3.2.5. We prove (3) => (4). Let U be an open cover of X. By (3) there exists a (countable) star-finite open refinement V of U such that ord(V)
is a V-mapping by Corollary 2.3.4, and hence a U-mapping. Also, P is a polytope by Theorem 2.3.5. Finally, since ord(V) < n, N(V) has no simplices
3.2. TRANSLATION INTO OPEN COVERS
167
of dimension greater than n. By the Countable Closed Sum Theorem 3.2.8 it follows that dim P < n. We prove (4) =>• (1). Let U be an open cover of X. By (4) there exists a poly tope P such that dimP < n and a U-mapping f : X -> P. For each U € U let E(U) be the set of all y G P with an open neighborhood Vy such that /~ 1 [T / y ] C U. Then the collection £ = {E(U) : U £ 11} is an open cover of P while moreover f-l[E(U)]CU for every U G U. Since dimF < n, Theorem 3.2.5 implies the existence of an open shrinking V = {V(U) : U G U} of £ = {E(U) : £7 e U} such that ord(V) < n. It is clear that the collection W = {f~l[V(U)] : U G U} is an open shrinking of U such that ord(W) < n. By Theorem 3.2.5 we therefore conclude that dim A7" < n. D Exercises for §3.2. A space X is called almost zero-dimensional if it has an open base 23 such that every B e "B has the property that X \ B is the union of clopen subsets of X. 1. Let X be a space containing more than one point and let p £ X. Prove that dim(A \ {p}) = dimX. (So the dimension cannot be raised by the adjunction of a single point.) 2. Let (Xn, f n ) n be an inverse sequence of compact spaces such that
dim An < ra < oo for all n. Prove that the inverse limit of the sequence is at most mdimensional. (The compactness condition is superfluous in this result. But we are not in the position yet to prove this. See Exercise 3.5.2 for more information.) ^•3. Let A' be a dense-in-itself space such that dim X < n < oo. Prove that for every cover {Ui,...,Um} consisting of nonempty open sets there is a collection {Fi,...,F m } consisting of closed sets having the following properties: (1) Q^FiCUt for every i < m, (2) U™! K = X, (3) if i\ < i-2 < • • • < in+2 < m are arbitrary then P|n=i2 Fij — $• ^•4. Prove that every n-dimensional compact space has a component which is also n-dimensional. 5. Let X be a space. Prove that X is almost zero-dimensional if and only if X has a base !B such that for each pair G, H of elements of 3 with disjoint closures there is a clopen set W in A" with G C W C X \ H. 6. Prove that every almost zero-dimensional space is totally disconnected. ^•7. Prove that Erdos' space E is almost zero-dimensional.
168
3. BASIC DIMENSION THEORY
8. Prove that if X is almost zero-dimensional then every open cover of X has a countable refinement by pairwise disjoint closed sets. 9. Prove that C(I 2 ) is infinite-dimensional. 10. Let X be a homogeneous space. Prove that every nonempty open subspace of X has the same dimension as X. 3.3. The imbedding theorem The aim of this section is to prove that every space A" with dim X < n can be imbedded in E2n+1 . This is a result of fundamental importance. For n > 0. define yin = {x G E2n+1 : at most n coordinates of x are rational}.
So fc=0
where the sets D^n+i.fc ^ M2n+1 are as in the previous section. Observe that 9to is the space of irrational numbers P. The space 9tn is called Nobeling's universal n-dimensional space. For later use, we do not aim at imbeddings in R2n+1 but at imbeddings in 9tn. This does not require extra work and explains our terminology. Lemma 3.3.1. For n > 0, dim 9^ = n and 9Tn is the union of n + I zerodimensional subspaces. Proof. By Corollary 3.2.11 we get dimWn < n. To see that dim9T n > n, we claim that 91n contains a homeomorph of I n . This is easy. Define an imbedding i : I" -» 9?n by i(xi,...,xn) = ( x i , . . . , x n , \ / 2 , \/2, . ..,-s/2). By Theorems 3.2.9 and 3.2.12 we now obtain dim01n > n. Since, as observed above, 9Tn = !H2n+i,o U ^n+i,! U • • • U ^n+i.n? an application of Proposition 3.2.10 gives us that 9tn is the union of n + I zerodimensional subspaces. D A finite subset F = {xi, . . . , xn} of Mm is said to be in general position if for every collection i(0)
is geometrically independent, cf. §2.1.
3.3. THE IMBEDDING THEOREM
169
Lemma 3.3.2. Let G = {xi,... ,xn} and F = {?/i,... ,y/t} be subsets of Em such that G is in general position, and let e > 0. Then for each i < k there exists a point xn+i G Km such that (1) ||j/i -2; n+ i|| < £ >
(2) {xi,...,xn,xn+i,...,xn+k} is in general position. Proof. It is clear that the lemma follows by induction once we have established it for k — I . So we assume without loss of generality that k — 1. For every nonempty subset F C { l , . . . , n } of cardinality at most m, let v(F)=aS(F), the affine hull of F. Each v(F) is a nowhere dense closed subset of Mm since \F\ < m. Since G = {xi,... ,x n } is finite, there consequently exists a point
xn+l € 5T \ \J{v(F) : F C {1,..., n}, \F\ < m} such that g(y\, xn+\) < e. We claim that {x\,..., x n +i} is in general position. Since G is, we need only prove that if F C {!,..., n} has cardinality at most m then {xi : i 6 F} U {xn+i} is geometrically independent. This is clear however since by construction, xn+\ 0 v(F] (Theorem 2.1.6). D Now for the remaining part of this section, let X be a fixed space such that 0 < dim^ < n. Proposition 3.3.3. Let f 6 C(X,R2n+l; Q), let e > 0, and let H be an at most n-dimensional affine subspace of M 2n+1 . For every finite open cover li ofX there exists g e C(X, E 2n+1 ; g) such that (1) g ( f , g ) < e,
(2) g[X] n H = 0, (3) g is a U-mapping.
Proof. Since / is bounded, f[X] is compact. There consequently exists a finite open cover V of it such that mesh(V) < £/4. Let W be the common refinement of U and /-1[V]. Observe that W is finite. Since dim X < n, there exists by Theorem 3.2.5 a finite open shrinking £ of W of order at most n. The essential properties of the cover £ are the following ones: (4) £ has order at most n, (5) £ < U, (6) for each E e £, diam(/[£]) < e/4We assume without loss of generality that every member from £ is nonempty. Let £ = {-Bi, • • • , Em}. Now for each i < m, pick an arbitrary point
170
3. BASIC DIMENSION THEORY
There is a geometrically independent set 5 = {hi, . . . , hk] C R 2n+1 of cardinality at most n + 1 such that S spans H. By Lemma 3.3.2, for each i < m there exists yi e M2n+1 such that (7) g ( y l , z l ) < e/4, (8) {hi , . . . , hk , yi , • - • , ym } is in general position. Let KI, . . . , Aem : X —)• I denote the ^-functions with respect to the cover £. Define 5: X -»• R2n+1 by
Then g is clearly continuous and we claim that it is as required. First observe that g is bounded because its range is contained in the simplex spanned by the vectors {yi,... , y ^ } which is compact by Lemma 1.1.1(2). For F C {1, 2 , . . . , m} we let F denote {j/» : z 6 F}. For every x G X let Fx — {i < m : x G Ei}. Observe that Ki(x) > 0 if and only if i G Fx so that Y^ierx Ki(x} = 1It follows by (4) and (8) that Fx is geometrically independent. Consequently, the Ki(x), i G Fx, are the affine coordinates of g(x) with respect to Fx. In particular, g(x) G aS(Fx). Claim 1. 0(f,g) < e. Proof. Fix an arbitrary x G X. It follows by (6) and (7) that \\f(x)—yi\ for every i G Fx. Consequently, \\f(x) — 9(x)\
=
\\ / . Ki(x) ' f ( x ) — /
J
< e/2
KI(X) • yi
We conclude that g(f,g) < £/2 < £, which is as required.
•(>
Claim 2. g[X] C g[X] C \JX&X ztt(Fx) C M2^1 \ H. Proof. Let x e X. By (4), |FZ| < n + 1 and so if we put E = Fx U 5 then \E\ < (n + 1) + (n + 1) = 2n + 2.
By (8) this implies that E1 is geometrically independent. Now observe that by Proposition 2.1.5 we have aff(Fx)
n H = afi(Fx) n aff(S) = aff(F x n 5) = aff(0) = 0.
3.3. THE IMBEDDING THEOREM
171
Since g(x) G aff(F x ) and x is arbitrary it therefore follows that g[X] C |J aS(Fx) C R2n+l \ H. But since aS(Fx) is closed by Exercise 2.1.8 and there are only finitely many F^'s, we even have that g[X] C g\x] C |J aff(F x ) C R2n+l \ H, x£X
as desired.
§
It remains to prove that g is a U-mapping. Let p G [-X"]- By Claim 2 there is a subset Gp C {1, 2, . . . ,m} of cardinality at most n + 1 such that p G aff(G p ). We may assume that Gp is minimal with respect to this property. Since Gp is geometrically independent, the affine coordinates of p with respect to Gp are all non-zero. In this way we associated to every element p G g[X] a subset of indices Gp C {1,2, ... ,ra}. Claim 3. Let p G g[X] and i0 G Gp. If A C {y1? . . . , y m } \ {yi() } has size at most n + 1 then p g' a Proof. Our reasoning is similar to the one in the previous claim. Striving for a contradiction, assume that p G aff (A). Observe that p
U A| < (n + 1) + (n + 1) = 2n + 2.
By (8) this implies that Gp U A is geometrically independent. So by Proposition 2.1.5 we get p G aff (A) n aff (Gp) = aff (A n Gp). Since i0 ^ A and the affine coordinate of p corresponding to y^ is non-zero, by unicity of affine coordinates (Theorem 2.1.8), this is a contradiction. <0> Now let B be the complement of
and for some i G Gp,yi $ A}. By Claim 3, B is an open neighborhood of p in M2n+1 . Claim 4. g-i[B]C^GpEl. Proof. Pick an arbitrary z G g~~l[B] and let IQ G G p . By the definition of g we have g(z) G aff(F z ). Since |FZ| < n + 1 by (4), if i0 ^ F2 it would follow that which is a contradiction. So io G F2 or, equivalently, z G £^i() .
172
3. BASIC DIMENSION THEORY
If p E U = W1 \ g[X] then U is a neighborhood of p and g~l[U] = 0. Since by (5) we have £ < U, this completes the proof.
D
Proposition 3.3.4. Let X be a space 0 < dim^Y" < n. In addition, let
{Ui-.ie N} be a sequence of finite open covers of X such that mesh(Ui) < V; for every i. Then the function space C(X, R 2n+1 ; £>) contains a dense GSsubset G having the following properties: (1) (2) (3) (4)
Every f 6 C is an imbedding. Every f 6 G is a Ui-map for every i. The range of every f £ C is contained in 9Tn. The range of every / E C has compact closure in 9tn.
Proof. Let A = [i(l),... ,i(n + l)} C {1, 2 , . . . , 2n + l} and g i , . . . ,g n+1 G Q be arbitrary. Put # = (re G M2n+1 : xi(j) = q3 for every j < n + 1}. Observe that H is an n-dimensional affine subspace of M2n+1 and that
H n 9?n = 0. 2n+1
In addition, each point of R \ 9In has at least n + 1 rational coordinates and is therefore contained in an affine subspace of the form H. We conclude that the complement of 9tn is the union of countably many n-dimensional affine subspaces, say {Li : i G N}. Now consider the function space C(X, R 2n+1 ; Q). For every i G N, put C, = {/ 6 C(X, M 2n+1 ; Q) : f is a U z -map and f[X] n Li = 0}. It follows from Proposition 3.3.3 that Gz is dense in C(X, R 2n+1 ; Q) for every i. Claim 1. C, is open in C(X,R2n+1; Q) for every i. Proof. Take an arbitrary / 6 C^. Since / is bounded, f[X] is compact. Therefore, since f[X] fl Li = 0, there exists £ > 0 such that (5)
{yeE 2 ^ 1 :g(yJ[X])<e}nLi
=
(Corollary A.5.4). Now since / is a U^-map, every y in the compact set f[X] has a neighborhood Vy in R2n+1 such that f~l\Vy\ is contained in an element of Ui. Put V = {Vy : y e /[A"]}. By Lemma A.5.3 there exists 8 > 0 such that every A C M2n+1 with diam(A) < 6 and which moreover intersects f[X] is contained in an element of V. Let 7 — min{ 5/3, e}. We claim that the open
3.3. THE IMBEDDING THEOREM
173
ball about / with radius 7 is contained in C;. To this end, take an arbitrary function g G C(X, R 2n+1 ; g) such that 0(f,g) < 7. We shall first prove that
g[X] n L, = 0. This is easy. Take an arbitrary x G X. Then g(f(x),g(x)} that g(x) is contained in the compact set
< £, which implies
:Q(y,\X}}<e}. So by (5), g[X]r\Li = 0. We next prove that g is a Ut-map. Take an arbitrary element p G g[X}. There exists x G X such that g(x) G B(p, %). Consider the open neighborhood of x. We claim that B is contained in an element of U^. If z G B then clearly g(g(x),g(z))
< %.
Since g(f,g} < %, we also get
Q(g(x),f(x))<%, From these inequalities it follows that g(f(z)J(x))
<
g(g(z)J(z})<%.
e(f(z),g(z))+g(g(z),g(x))+e(g(x)J(x))
< % + % +% =8 and so
(6)
g-l[B(g(x),%}}Cf^[B(f(x),S})].
Since by the special choice of 6 the set B(f(x),6) is contained in an element of V, f~l[B(f(x),6)] is contained in an element of IL1. By (6) it therefore follows that g ~ l [ B ( g ( x ) , %)] is contained in an element of Hi. But this is as desired since B(p, %) C B(g(x), %). 0 By Corollary 1.3.6, C(X, M2n+1 ; g} is a topologically complete space, and is therefore a Baire space (Theorem A. 6. 6). Consequently, Claim 1 implies that
is a dense G^-subset of C(X, R2n+1 ; g). We shall prove that C is as desired. Observe that (2) follows by construction. Take any function / G G. Then f[X] n (JSi ^ = 0, i.e., f[X] C Wn. This proves (3) and (4) for /. So it suffices to prove that / is an imbedding. We shall prove that for every x G X and every neighborhood W of x there exists a neighborhood V of } ( x ) such that / -1 [V^] C W. From this it follows that / is one-to-one and that /: X —>• f[X] is open, i.e., / is an
174
3. BASIC DIMENSION THEORY
imbedding. Take an arbitrary x e X and a neighborhood W of x. Let e > 0 be such that B(x,e) C W. There exists i G N with l/i < e. Since / is a Uimap, there exists a neighborhood V of /(x) such that /~ 1 [T / ] is contained in an element U of Hi. So x G C/ and since diam(t/) < y^ < e, we conclude that f~l[V] CUC B(x,e) C W, as required. D We now come to the main result in this section. The Imbedding Theorem 3.3.5. Let X be a space with 0 < dim AT < n. Then there exists an imbedding i: X —>• 9ln such that i[X] has compact closure in 9ln. Proof. By Exercise A.6.5, X has an admissible metric Q for which there exists a sequence of finite open covers {Hi : i G N} such that for every z, mesh(Ui) < l/i (i G N). So we are in a position to apply Proposition 3.3.4 to get what we want. D Remark 3.3.6. Notice that we proved in fact the stronger statement that for a space X with 0 < dim^C < n, any bounded map /: X -> R2n+1 can be approximated arbitrarily closely by an imbedding in yin the range of which has compact closure in 9tn. Remark 3.3.7. Theorem 3.3.5 is 'best possible' for there exist n-dimensional spaces that cannot be imbedded in M 2n . An example of such a space is the union X of all n-faces of a (2n + 2)-dimensional simplex. This X is obviously n-dimensional and that it cannot be imbedded in R2n is due to FLORES [164]. The proof of this fact for n — I is simple and is based on the Jordan Curve Theorem for R 2 . For n > 1, the proof is more complicated. Its main ingredient is the Borsuk-Ulam Antipodal Theorem in Exercise 2.5.2(4), which is equivalent to the Lusternik-Schnirelman-Borsuk Theorem 2.5.1. See ENGELKING [155, p. 106] for details and additional information. Some applications. We now present some applications of the results in this section. The proofs are based on Theorem 3.3.5 which explains our interest in imbeddings in 9tn instead of imbeddings in R 2n+1 . Corollary 3.3.8. If dim X < n, n > 0, then X has a compactification *yX such that dim 7^ < n. Proof. Let i: X -» 9Tn be an imbedding such that jX = i[X] is a compact subset of yin (Theorem 3.3.5). Then dim7^Y < n by Lemma 3.3.1 and the Subspace Theorem 3.2.9. D Corollary 3.3.9. Let X be a nonempty space and let n > 0. The following statements are equivalent: (1) dim X
3.3. THE IMBEDDING THEOREM
175
Proof. Suppose that dim X < n. By Theorem 3.3.5 we may assume that A is contained in 9ln. By Lemma 3.3.1, 9tn is the union of n+1 zero-dimensional subspaces. Now apply the Subspace Theorem 3.2.9. Conversely, assume that X is the union of at most n + 1 zero-dimensional subspaces. Then apply Corollary 3.1.7 to conclude that dim A" < n. D We now prove a generalization of Corollary 3.1.6 to n-dimensional spaces. Corollary 3.3.10. Let 0 < n < oo and let X be an n-dimensional subspace of the space Y. Then for all disjoint closed subsets A and B ofY there exists a partition S between A and B in Y such that dim(S n A") < n — 1. Proof. Write A = (J^i ^> where dimA^ < 0 for every i (Corollary 3.3.9). Let A and B be disjoint closed subsets of A". By Corollary 3.1.6 there is a partition S between A and B such that 5 n X \ = 0 . But then SnX is contained in (J™J2 Xi and hence is at most (n — l)-dimensional by Corollary 3.3.9. D By a repeated application of the previous corollary, we get the following result. Corollary 3.3.11. Let A be a space with subspace Y such that dimY < n, where 0 < n < oo. For each collection T — {(F^Gi) : i < n + 1} of pairs of disjoint closed subsets of X there exists for every i < n + 1 a partition Si between Fi and Gi such that n+1
Y n n S, = 0. The following result is known as the G$ -Enlargement Theorem. Corollary 3.3.12. Let M be a subspace of a space X. If dim M < n then there exists a G$-subset S C X with M C S and dim S < n. Proof. We first prove the corollary for the special case dim M = 0. Let £ be a countable open base for M. For every pair (Eo,Ei) of elements of £ such that EX) C EI there exists by Corollary 3.1.6 an open set B C M such that E0 C B C B C EI
while moreover M n Frf? = 0. The collection I> of Z?'s obtained in this way is clearly a countable base for M. In addition,
F= \J FrB is an Fcr-subset of M which misses M. Let S = M\F. Then 5 is a G^-subset of the G^-subset M and hence is a G^-subset of A. It is clear that "B \ S consists of (relative) clopen sets, and so S is zero-dimensional.
176
3. BASIC DIMENSION THEORY
We will now use the just proved special case to prove the general case. To this end, write M = M0U---UMn, where Mi is zero-dimensional for every 0 < i < n (Corollary 3.3.9). By the above there exists for every 0 < i < n a zero-dimensional G$-subset M* of X with Mi C M*. The union of the M*'s is clearly a G^-subset of X and is at most n-dimensional by Corollary 3.3.9. D Exercises for §3.3. 1. Give an example of a space X, an open cover U of X, and an imbedding /: X —t X such that / is not a U-mapping. 2. Let X be a space. Prove that every continuous function /: X —>• Q can be approximated arbitrarily closely by an imbedding in Q. 3. Prove that for every n > 0 there exists a compact space Xn such that (1) dimXn = n, (2) every space Y with dimy < n can be imbedded in Xn. >4. Let X be a compact space without isolated points such that 0 < dimX < n < oo. Prove that there exists a continuous surjection /: C —>• X such that each fiber of / has cardinality at most n + 1 5. Let X be a compact space and let /: C —» X be a continuous surjection such that each fiber of / has cardinality at most n + 1. Prove that X is at most n-dimensional. 6. Let X be a compact space with 0 < n = dimX < oo. Prove that every open cover U of X has a closed refinement 7 such that for every F £ 7 the collection {F1 € 5 : F' n F ^ 0} has cardinality at most 32"+1. ^•7. Prove that every (n+l)-dimensional compactum contains an n-dimensional compactum (and hence an n-dimensional continuum by Exercise 3.2.4). 3.4.
The inductive dimension functions ind and Ind
There are two additional dimension functions that are important in dimension theory, namely, the small and the large inductive dimension function, abbreviated ind and Ind, respectively. It turns out that for a given space these functions and the dimension function dim take the same value. The functions ind and Ind are important because in certain situations it is easier to deal with them than with dim. For example, it is a triviality to verify that for all X and y, ind(X xY) < i
3.4. THE INDUCTIVE DIMENSION FUNCTIONS ind AND Ind
177
by equality of ind and dim it therefore follows that dim(X x Y) < dim A + dimF. However, to verify this straight from the definition of dim is unpleasant. The aim of this section is to study basic properties of the dimension functions ind and Ind and to prove equality of all three dimension functions. We shall now give the definition of ind, which differs from the definition of dim in the sense that it is an inductive definition. Indeed, for a space X define its small inductive dimension ind X £ { — 1 , 0 , 1 , . . . } U {00} as follows: indA = -l <£> A = 0, ind X < n <4> for every x G X and every closed subset A of X not containing x there exists a partition L between {x} and A such that indL < n — 1, ind X = n <& ind X < n and ind X <£. n — 1, ind X = oo <{=> ind X ^ n for every n > — I. It is clear that if X and Y are homeomorphic spaces then ind A" = indl'. The property of having small inductive dimension at most n can be expressed in terms of a special countable base. Lemma 3.4.1. A space X satisfies 0 < ind A" < n if and only if X has a countable base *B such that ind Fr B < n — 1 for all B e !B. Proof. If £ X < n then A" has clearly a base U such that ind Fr U < n — 1 for every U £ U. This base has a countable subcollection !B which is still a base (Exercise A.2.12). So !B is as required. The converse of the lemma is trivial. D A consequence of this lemma is that for a space X we have ind A" = 0 if and only if X is nonempty and has a base consisting of clopen sets. This means that X is zero-dimensional if and only if ind X — 0. We shall now derive a few additional properties of the dimension function ind. The following triviality shall be used several times in the forthcoming: if X is a space, Y C X. y 6 Y and A is a (relatively) closed subset of Y not containing y. then y does not belong to the closure of A in X. Proposition 3.4.2. Let X be a space. Then dim A" = 0 iff ind A = 0. Proof. This is a triviality since, as observed above, ind A = 0 iff the clopen subsets of X form a base iff dim A = 0 (see Page 41). D Lemma 3.4.3. Let X be a space, and let A be a subspace of X. ind A < ind X.
Then
Proof. There is nothing to prove if ind A = oo, so we assume that ind A is finite. Again, there is nothing to prove if ind A = — 1. So assume the lemma
178
3. BASIC DIMENSION THEORY
to be true for all spaces Y with 0 < ind Y < n — 1, and assume that ind X < n. Take x E A and let C be a closed subset of A not containing x. Since x $ C there is a partition L in X between x and C such that indL < n — 1. Then L n A is a partition between x and C in A so that, by our inductive assumption, ind(L n A)
The Addition Theorem 3.4.4. If A and B are subspaces of a space X then 'md(AuB) < ind A + ind B + I . Proof. If ind A — oo or ind B = oo then there is nothing to prove. So we assume that ind A and indL? are both finite. We induct on ind A + indB. If A and B are both empty, i.e., if ind A = indB = — 1, then there is again nothing to prove. So assume that the theorem is true for all subspaces E and F of X with
ind E + ind F < n - 1, where n > — 1. Let a = ind A and 0 — ind B and assume that a + fi = n. We shall prove that ind(Auf?)
ind (5 n (A U B))
and Y = F then indX = 1 and
This shows that Theorem 3.4.4 is 'best possible'.
Corollary 3.4.6. If X can be written as the union ofn + 1 zero-dimensional subspaces then indX < n. Proof. Since for every space F, dimF = 0 iff indl" = 0 (Proposition 3.4.2) the result follows immediately from Theorem 3.4.4. D
3.4. THE INDUCTIVE DIMENSION FUNCTIONS ind AND Ind
179
We want to prove the equality of dim and ind. Before being able to do that, we need to derive a preliminary lemma. Lemma 3.4.7. Let X be space and let n > 0. The following two statements are equivalent: (1) For every x G X and for every open neighborhood U ofx there exists a partition L between {x} and X \U such that dim L < n — 1. (2) For every pair A and B of disjoint closed subsets of X there exists a partition L between A and B such that dimL < n — 1. Proof. First observe that (2) => (1) is a triviality. For (1) => (2), let A and B be disjoint closed subsets of X. By assumption, for each x G X there exist open sets U(x] and V(x] such that (3) x G t/0), U(x)r\V(x) = 0, (4) if L(x) = X\ (U(x) U V(x)) then dim L(x) < n - 1, (5) U(x) n A = 0 or U(x) n B = 0. The cover {U(x) : x G X} has a countable subcover, say U (Corollary A.2.3). Put L = \J{L(x) : U(x) G U}. Observe that by (4) and the Countable Closed Sum Theorem 3.2.8 we have dim L < n - 1. Let U0 = {U G U : U n A ^ 0} and Ui = U \ U 0 . Enumerate U0 as {U0ii : i € N} and Ui as {E7M : i G N}, respectively. For each i G N put (6) Ei = U0ti \ \Jj
180
3. BASIC DIMENSION THEORY
hypothesis, all these partitions have covering dimension at most n — 1. We conclude that dim A < n by Lemma 3.4.7 and Corollary 3.1.10. We shall now prove that ind X < dimX. If dim A = — 1 or dim A^ = oo then this is a triviality. So assume that 0 < dim X — n. By Corollary 3.3.9, X is the union of at most n+1 zero-dimensional subspaces. So by Corollary 3.4.6 we get ind A
3.4. THE INDUCTIVE DIMENSION FUNCTIONS ind AND Ind
181
subsets A and B of X there exists a partition L between A and B such that dimL < n — 1. By our inductive hypothesis it follows that these partitions all have large inductive dimension at most n — I. From this we conclude that indeed Ind X < n = dim X. D Remark 3.4.11. From now on we will no longer formally distinguish between dim, ind and Ind. Theorems proved for dim will be used for ind and vice versa, etc. Let X and Y be spaces. It is natural to wonder about the relation between dim A, dimF and dim(A x Y}. Since dimIR71 = n for every n, one would expect the relation dim(A r x Y} — dim AT + dimF to hold for all nonempty X and Y. Unfortunately, this is not true, see Example 3.4.13. We shall now prove 'half of the expected equality.
Theorem 3.4.12. Let X and Y be nonempty spaces. Then dim(A x Y) < dimX + dimY. Proof. By the Coincidence Theorem 3.4.10, it suffices to prove that ind(A x y) < ind A" + indY. First observe that without loss of generality, 0 < ind X, ind Y < oo. We shall prove the theorem by induction on ind X + ind Y. If ind X + ind Y = 0 then both X and Y are zero-dimensional, and so is X x Y. Assume that the theorem is true for all spaces X and Y with ind A + ind 1" < n- 1,
where n > 1, and let X and Y be spaces such that ind A = a, indF = /3 and a + (3 = n. By Lemma 3.4.1 there is a countable base 23 for A such that dimFrS < a - I for every B 6 B. Put B' = \JBe^ FrE. Then B' is an FO--subset of X and hence by the Countable Closed Sum Theorem 3.2.8 it follows that dim B' < a - 1. Since Fr B n (A \ B') = 0 for every B e £ it also follows that *B \(X\B') consists of relatively clopen sets, i.e., dim(A\.B) < 0. Similarly, construct an F^-subset E C Y such that dimE < (3 — 1 and dim(y \ E) < 0. Now observe that by our inductive hypothesis and the Countable Closed Sum Theorem 3.2.8, B' x Y is an at most (a — 1) + /?dimensional Fff-subset of A" x y. It follows similarly that X x E is an at most a + (/3 ~ 1)-dimensional F^-subset of A x Y. Again by the Countable Closed Sum Theorem 3.2.8 we get dim ((B' x y) U (X x E))
Since the complement of (B' x Y} U (X x E} is zero-dimensional, being the product (A \ B') x (Y \ E), we get by Theorem 3.4.4 that dim(A xY)<(a + 0-I) + Q + l = a + /3, as required.
D
182
3. BASIC DIMENSION THEORY
The following example shows that the other 'half of the expected equality dim(X xY] = dimX + dimY need not hold. See Theorem 3.9.5 for a (strong) generalization of this result to all dimensions. Example 3.4.13. There exists a one-dimensional space E such that E x E is one-dimensional as well. We will show that the space E in Example 1.5.18, i.e., Erdos' space, is the required example. From Exercise 1.5.14 we know that E « E x E. So it suffices to prove that dim.E = 1. In Example 1.5.18 we showed that dim E > 1. We shall prove here that ind E < 1. First observe that E is a subgroup of I2, and hence is a topological group. By homogeneity, it therefore suffices to prove that indE < 1 at the zeroelement of E. We claim that for every e > 0, the sphere S£ consisting of all points in E with norm e, is zero-dimensional. Indeed, since all elements of S£ have the same norm, Lemma 1.1.12 shows that S£ is homeomorphic to a subspace of the countably infinite product of rational numbers, which is zerodimensional by Proposition 1.5.3 and the trivial observation that products of zero-dimensional spaces are again zero-dimensional. So S£ is indeed zerodimensional, being a subspace of a zero-dimensional space. We now summarize the results obtained so far in the following Corollary 3.4.14. (1) Let X be a nonempty space and let n > 0. The following statements are equivalent: 1. dimX < n, 2. for every subspace AofX, dim A < n, 3. X has a compactification 'yX such that dim^yX < n, 4. for every pair of disjoint closed subsets A and B of X there exists a partition L between A and B such that dim L < n — I , 5. X is the union of at most n + 1 zero-dimensional subspaces. (2) If X is a space and if A and B are subspaces of X then dim(AuB} < dimA + dimB + 1, dim(A x B) < dim A + dimB. Exercises for §3.4. Let X be a space. Define the compactness degree cmpX of X, as follows: cmpX = — 1 <=>• X is compact, cmp X < n <&• for every x G X and every closed subset A of X not containing x there exists a partition L between {x} and A such that cmp L < n — 1, cmpX = n «=>• cmpX < n and cmpX •£ n — 1, cmp X = oo <$ cmp X ^ n for every n > — 1.
3.5. DIMENSIONAL PROPERTIES OF COMPACTIFICATIONS
183
So the compactness degree of a space is intuitively its small inductive dimension modulo the class of all compact spaces. It is clear that if X and Y are homeomorphic spaces then they have the same compactness degree. Let *yX be a compactification of X. We call the space ^X \ X the remainder of the compactification ^X of X. Let the compactness deficiency be the least integer n G [ — 1, oo] such that X has a compactification with remainder of dimension n. Observe that homeomorphic spaces have the same compactness deficiency. Given a space X and n £ N, we shall denote by X(n) the set of all points in X that have arbitrarily small neighborhoods with at most (n — l)-dimensional boundaries. 1. Prove that Erdos' space E is not topologically complete. Give an example of a one-dimensional totally disconnected topologically complete space F such that F and F x F are one-dimensional 2. Let X be a space. Prove that cmpX < def X < 3. Let X be a totally disconnected space which is not compact and let ^X be a compactification of X. Prove that dim^X \ X > dimX, i.e., X cannot be compactified by adding a set of smaller dimension than the dimension of X. ^•4. Let X be a compact space such that 0 < dimX = n < oo, and let Y be the product Q x X. (1) Prove that if S is a topologically complete space containing Y then S\Y contains a homeomorph of X . As a consequence, the dimension of S \ Y is at least n. (2) Prove that def Y = n. 5. Let X be a space and n E N. Prove that there is a countable family open sets U such that (1) U is a local base at every point of X( n ), (2) dim(I7 \ C7) < n - 1 for every U e U. Prove that dimX( Tl ) < n. 3.5. Dimensional properties of compactifications In this section we collect some results on compactifications preserving certain dimensional properties. We saw in Corollary 3.3.8 that every m-dimensional space X has an mdimensional compactification. Such a compactification is called dimension preserving. The following result generalizes this and is central in this section. Theorem 3.5.1. Let X be a space, and let {(Fn,Gn) : n € N} be a countable collection of pairs of disjoint closed subsets of X. Then there is a dimension preserving compactification aX of X such that for every n the sets Fn and Gn have disjoint closures in aX .
184
3. BASIC DIMENSION THEORY
Proof. We will prove this in the case of finite dimensional X only, the infinite-dimensional case being simpler. Let Q be an admissible metric on X for which there exists a sequence {Un : n G N} of finite open covers of X such that mesh(Un) < l/n for every n (Exercise A.6.5). Since An = {X \ Fn,X \ Gn} is an open cover of X, we can replace lin by the common refinement of Vtn and An. So we may assume without loss of generality that for every n and U E Un we have that U n Fn — 0 or U n Gn = 0. Let ra = dimX. By Proposition 3.3.4 there is an imbedding i: X —> Otm such that (1) i[X] has compact closure in ytm. (2) i is a U n -map for every n. Put aX = i[X]. Then aX is a dimension preserving compactification of X (cf. Corollary 3.3.8). Fix n E N. We claim that i[Fn] and i[Gn] have disjoint closures in aX. Since i is a Un-map, there is an open cover 9 of M2m+1 such that
{rMGJrGeS} refines Un. Striving for a contradiction, assume that
pei[Fn]ni[Gn]. There is an element G E S such that p E G. Pick an element £/ E Un such that i~1[G] C U. We may assume without loss of generality that UC\Fn = 0. But then i~l[G] n ,Fn = 0 and so G n z[Fn] = 0. Since M2m+1 \ G is closed, this implies that p £ i[Fn], which is a contradiction. D Corollary 3.5.2. Let X be a space, and let bX be a compactincation of X. Then there is a compactincation aX ofX which is dimension preserving while moreover aX > bX. Proof. By Lemma A.5.6 there is a collection {(A'n,B'n) : n E N} of pairs of disjoint closed subsets of bX such that for any pair (E, F} of disjoint closed subsets in bX there is an ra E N such that E C A'm and F C B'm. For every n, let An = A'n C\X and Bn = B'n n JC, respectively. By Theorem 3.5.1 there is a compactification aX of X such that for every n the closures in aX of An and Bn are disjoint, while moreover dimaX = dimX. But this implies by Lemma A.8.3 that the identity function X —> X can be extended to a continuous function aX —> bX, i.e., aX > bX. D We now show that compactifications can also preserve the dimensions of countably many closures of closed sets.
3.5. DIMENSIONAL PROPERTIES OF COMPACTIFICATIONS
185
Theorem 3.5.3. Let X be a space, and let "H be a countable family closed subsets of X. Then there is a Wallman base 3 of X such that (1) :KC_J, (2) dimH = dimH for every H 6 "K (here closure means closure in u(X,f)). Proof. List 3~C as {Hi : i e N} such that every H 6 JC is listed infinitely often. For every H 6 !K put HH = dimH, and if H = Hl then m = HHLet 3~i be an arbitrary Wallman base for X (Lemma A. 9.1) and let ai X be the compactification cj(X, 3"i) of X. Consider the set H\. The closure of HI in a\X is a compactification of HI, say PiHi. By Corollary 3.5.2 there is a compactification /32Hi > ftiHi such that dim ,$2 HI = n\. By an application of Exercise A. 11. 8, there is a compactification £2-^ of X for which there is a continuous function /i : t2X ->• aiX
which extends the identity on X while moreover the closure of HI in t2X is ^2 HI . Hence the closure of HI in t2X has dimension n\ . By Corollary A. 9. 8 there is a Wallman base TI of X such that t\X and w(X, TI) are equivalent compactifications. Let S"2 be a Wallman base of X containing the collection 9~iUTiU{Hi} (Lemma A. 9.1), and let a2X be u(X,y2}. By Exercise A. 9. 4 there is a continuous function gi : a2X —>• iiX which extends the identity on X. Observe that at this stage we have the following compactifications and continuous functions: The Wallman bases fi and 32 have the property that 3"i C y2, and the compactification t\X has the property that the closure of HI in it has the 'right' dimension, namely n\. This is the first step in an inductive process. We now proceed recursively as above, dealing with the set Hi in the i-ih step of our process. We thus obtain an inverse sequence 1
9l
2
9
2
3
„ V j f + V j 4 V , V , f a\X
(3) 3"i_i c y^ (4) Hitfi, (5) u(X,?i) = aiX,
186
3. BASIC DIMENSION THEORY
(6) fa: tiX —> aiX extends the identity on X, (7) gi : di+iX -» tiX extends the identity on X, (8) the closure of Hi in tiX has dimension U{. Put y = Ui^i ^i- We claim that 3~ is the desired Wallman base of X (that it is a Wallman base follows from Exercise 1.10.11). Let a^X be the inverse limit of our inverse sequence. We already know by Exercise 1.10.11 that So docX is a compactification of X and every projection restricts to the identity on X. Now let H G "K be arbitrary. The closure of H in tiX is mapped by fa onto the closure of H in aiX. Similarly, the closure of H in di+iX is mapped by gi onto the closure of H in tiX. Also, the closure of H in a^X is mapped by the various projections onto the closures of H in the corresponding compactifications. This is all true because of the compactness of the spaces involved and the fact that every function in our inverse sequence restricts to the identity on X. So all the closures of H in the various compactifications form an inverse system of which the limit is the closure of H in a^X (Exercise 1.10.7). Since H is listed infinitely often, by (8) this sequence has a subsequence all elements of which have dimension UH- But then, since every subsequence of an inverse system has the same limit (Lemma 1.10.6), by Exercise 3.2.2 it consequently follows that the closure of H in a^X has dimension at most riH- But since H is a subspace of its own closure in a^X , by the Subspace Theorem 3.2.9 we get that this closure is at least HH -dimensional. So we are done. D Corollary 3.5.4. Let X be a space, and let 3 be a countable family closed subsets of X. Then there is a compactification aX of X such that dim F = dim F
for every F € J. Remark 3.5.5. An inspection of the proof of Theorem 3.5.3 shows that for this corollary it is not necessary to deal with Wallman bases. The proof of it can be completed entirely within the framework of inverse limits. The following corollary to Theorem 3.5.3, which will be important in §3.11, does however not follow from Corollary 3.5.4. An inspection of its proof will show that we need to control the dimensions of various partitions between points and closed sets, and for that control over closures alone does not suffice. It turns out that via the use of Wallman bases, one gets the desired control automatically. One simply has to put the right collection of closed sets in a suitable Wallman base, and the Wallman compactification theory does the rest.
3.5. DIMENSIONAL PROPERTIES OF COMPACTIFICATIONS
187
Corollary 3.5.6. Let X be a space with countable family closed subsets 9Then X has a compactification aX having the following properties: (1) dimG = dimG for all G G g, (2) X(n) C (aX)(n} for every n > 0. Proof. For n > 0 there is by Exercise 3.4.5 a countable collection of open sets Un in X such that (3) Un is a local base at every point of X(n), (4) dim(Z7 \ U) < n - I for every U G Un. Let IK be the collection {U : U G U n , n > 0} U {X \ U : U G U n ,n > 0} U
U{!7\[/:£/GUn,n>0}ug. Then IK is a countable collection closed subsets of X. By Theorem 3.5.3 there is a Wallman base f of X such that
(5) IKCJF, (6) dimH = dimH for every H G IK (here closure means closure in
u(x,y}}. Put aX = a;(X, IF). Observe that (1) follows from (6) since S C IK. So we need only worry about (2). To this end, let x G -X"(n) for certain n > 0 and let V be an arbitrary open neighborhood of x in aX. Let W" be a closed neighborhood of x in aX such that W C V. There is an element U G Un such that x 6 C7 C U C WnX (closure means closure in X here). Observe that
{u,u\u,x\u} cy. This means that
u*u(x\uy = u(x,?)
and
C/*n(.Y\C/)* = (C7\C7)*. (Lemma A. 9.4). Observe that U* is the closure of U in u}(X,3) (Exercise A.9.1). Similarly, (U\U)* is the closure of U \ U in u(X, J). Since W is a closed neighborhood of x, it consequently contains U* . Hence (U \ U)* is a partition between {x} and aX\W which is at most (n — l)-dimensional by (4) and (6). Since V was arbitrary, this proves that x G (aX)^, as required. D Let X be a space, and let / : X -> X be continuous. A set A C X such that f[A] n A — 0 or, equivalently, A n /-1[A] = 0, is called a color of /. A coloring of / is a finite collection of colors of / which covers X. We say
188
3. BASIC DIMENSION THEORY
that / is colorable if there is a finite cover of X consisting of closed colors of/Observe that every colorable map is fixed-point free. There is an easy to describe example of a fixed-point free map which is not colorable. The verification of this is difficult. See §3.12 for details. Let /: X —>• X be fixed-point free. Then every x E X has a closed neighborhood Vx such that Vx Pi f[Vx] = 0. By Corollary A. 2. 3, the cover {Vx:xe X} has a countable subcover. There consequently exists for every fixed-point free map / : X —>• X a countable and potentially infinite closed cover V of X such that every V G V misses its own image. So a fixed-point free map on X is colorable if such a cover can be chosen to be finite. Corollary 3.5.7. Let X be a space and let 3 be a countable collection of continuous functions from X to X. Then there is a compactification aX of X such that
(1) (2) Every f 6 y can be extended to a continuous function f : aX -> aX. If f is a homeomorphism then so is f. In addition, if f £ 3~ is colorable then f is fixed-point free. Proof. We assume without loss of generality that the identity on X belongs to 3~ and that for each homeomorphism / £ J we also have that f~l G 3~. Assume that / € 3 is colorable. Then there is a finite closed cover 9/ ofX such that Gr\f~l[G] = 0 for every G e S/. Put Kf = Sf U f~l[9f]Let
U r^n]. Observe that since the identity belongs to 3~ we have An C An+\ . In addition, since X G An we have dimo;(X, A n ) = dim JC. As in the proof of Theorem 3.5.1, oo
A = |^J .An
n=l
is a Wallman base such that aX = u)(X,A) is the inverse limit of an inverse sequence of compact spaces having the same dimension as X, hence aX is dimension preserving.
3.5. DIMENSIONAL PROPERTIES OF COMPACTIFICATIONS
Since
189
r^cyt
for every / G y we have that / can be extended to a continuous function /: aX —>• aX by Exercise A. 9. 4. If / is a homeomorphism then so is / because f~l G 5F can be extended as well and this extension is the inverse of /. Now assume that there exists an / G 9" which is colorable. The collection
is a sub collection of A. We claim that /: aX —>• aX is fixed-point free. To see this, observe that the finite collection 9 covers X and so does [f~l[G] : G G 5}- For a given p € aX there consequently exists G G 9/ such that p G f ~ l [ G ] . The continuity of /: aX —> aX therefore implies that
Since G and f ~ l [ G ] are disjoint and both belong to A, they have disjoint closures in aX (Lemma A. 9.4(2)). So f(p) ^ p. D The dimension of almost zero-dimensional spaces. We will use compactifications to prove that every almost zero- dimensional space is at most one-dimensional. A subset X of a compactum K is L-imbedded in K if for every open cover U of K there is a neighborhood V of X in K such that every subcontinuum of K which is a subset of V is contained in an element of U. Let A and B be disjoint closed sets in a space X, and let e > 0. A triple
is an L£ -enlargement for the pair (A, B} if (1) G and H are open neighborhoods of A and .B, respectively,
(2) G n # = 0, (3) y is a discrete collection of closed sets of diameter at most e such that X \ (G U H) = U ?• Theorem 3.5.8. Let X be an L-imbedded subspace of a compact space (K,g). If A and B are disjoint closed subsets of X having disjoint closures in K then for each e > 0 there is in X an L£ -enlargement (G, H, J) for the pair A, B such that for every F G J the sets
G D F,
~H n F
have disjoint closures in K (here closure means closure in K).
190
3. BASIC DIMENSION THEORY
Proof. Let 8 = Q(A,B). Our assumptions give us that 6 > 0. Pick an arbitrary e > 0. We will construct in X an Z/e-enlargement for the pair (A, B}. There is by assumption an open neighborhood U of X in K such that all continua in U have diameter at most y 2 min{c, 6}. Let (p: K —>• I be continuous such that U = (p~l [(0, 1]] (Lemma A. 4.1). For each n, let ^» = y" 1 [[ 1 /(n+l), 1 /n]].
Observe that every Kn is compact. Fix n for a moment. If C is a component of Kn then C is not a continuum from A n Kn to 5 n K"n . This is so because diam(C) < l/28 since C C U and 0(.A,.B) = 6. Hence Kn can be split into two disjoint closed sets Pn and Qn containing Ar\Kn and Br\Kn, respectively (Exercise A. 11. 5). Observe that the sets oo
oo
s= |J A n U p n t / ) , r = U Q 2 n U ( £ n £ / ) 71=1
n=l
are disjoint closed subsets of U. Let a : U —> I be a Urysohn function such that a[S] C {0} and a[T] C {!}. Put F = a-1 [[0,i/0],
W = a'1 [(3/4,1]].
Clearly, Vflt/ and WC\U are disjoint. Observe that U\(V(JW) is contained in the union of the discrete collection % = {K2n_! : n G N} in [/. Each K2n-i is a compactum whose components have all diameter at most i/2£. So by Corollary A. 10. 2 we can split K2n-i into a finite pairwise disjoint collection compacta y2n-i of diameter at most e. Observe that the traces of V and W on any of these compact pieces are disjoint. We conclude that the triple n—l
is the required //^-enlargement for (A, B).
D
Corollary 3.5.9. Let X be an L-imbedded subspace of a compact space K. Then dim X < I . Proof. In this proof A stands for the closure of A in K.
3.5. DIMENSIONAL PROPERTIES OF COMPACTIFICATIONS
191
Let F be a closed subset of A r , and let p G X \ F. We will prove that there are disjoint open subsets V, W in X with p E V, F C W and dim ((A" \(V\J W)) < 0. By Theorem 3.4.10, this suffices. We shall construct two increasing sequences of open sets
p G GI C G2 C • • • ,
F C HI C H2 C • • •
and for every n G N a collection Jn of closed subsets of X such that (1) G n n # n n A = 0, (2) the triple (Gn+i,Hn+i,yn) is an L^-enlargement for the pair
(Gnnx,Hnnx), (3) if F € 3~n is arbitrary then
F n Gn n F n #n = 0. Let GI , HI be any pair of open sets containing p and F, respectively, having disjoint closures in K. Suppose Gn,Hn have been defined. For n = 1, we get the triple (G2,#2j3~i) directly from Theorem 3.5.8. Assume n > 2. For each F 6 3>i-i consider the disjoint closed sets A(F] = Gn n F,
£(F) =Hn^F
in A. These sets have disjoint closures in K by (3). By Theorem 3.5.8 there is an L^-enlargement of the pair (A(F), -B(F)) , such that for every R 6 ft^ the sets (4)
G(F) n fl, fT(F) n R
have disjoint closures in K. For each F 6 J n _i let C/(F) D F be an open set in A" such that the collection (5)
is discrete in A (Exercise A.7.3). Put
Gn+l =GnU \J{G(F) n U(F) : F e ^.J, Hn+1 =HnU \J{H(F) n t/(F) : F 6 ^_i}. It is clear that Gn+i and Hn+i are open, and by (5) and (1) it follows that
G n + 1 n H n + i n A = 0. Put
Jn = {FnR:F e ^.RG^F}. Then 3"n is a discrete collection of closed sets with diameter at most l/n. It is easy to see that the triple (Gn+i,Hn+i, Jn) is an L^-enlargement for the pair
192
3. BASIC DIMENSION THEORY
This completes the inductive construction. Now put oo
oo
V = U Gn,
W=\jHn.
n=l
n=l
Observe that for each e > 0 the complement X \ (V U W) is the union of a discrete collection of closed sets of diameter at most e. As a consequence, the partition X \ (V U W) is at most zero-dimensional (Exercise 1.5.3). D We are now in a position to present the proof of the following interesting result.
Theorem 3.5.10. If X is almost zero-dimensional then dimX < 1. Proof. Let X be nonempty and almost zero-dimensional. In addition, let !B be a base for X with the properties stated in Exercise 3.2.5. We may assume that 1$ is countable (Exercise A.2.12). For every pair G, H G 23 such that G n # = 0,
pick a clopen set C in X with
G C C C H. Since 23 is countable, the collection C of clopen sets picked in this way is countable as well. There is a compactification ^X of X such that for every C E C we have that C and X\C have disjoint closures in jX (Exercise A.9.3). That is, the closure of C in ^X is clopen in jX. For every B E 23 let B* be an open subset of ^X such that B* n X = B. We claim that X is L-imbedded in ~/X. To this end, let e > 0, and put J{ = {B E 23 : diamtf* < Y3e}, and
U = \jK, respectively. If x G X then there is an element B 6 23 such that
B CB(x, y7e) n X. Since 5 is dense in B*, the set 5* is contained in the closed ball D(x, VT^), hence B* C U. We conclude that U is an open neighborhood of X in ~fX. To conclude the proof, let K C [7 be an arbitrary continuum. Pick two elements BQ,BI G ^K such that
(i)
B*nK^®^KnB*.
If .Bo and #1 have disjoint closures in X then there is an element C G C such that B0CC CBl.
3.6. MAPPINGS INTO SPHERES
193
The closure of C in jX is clopen. By (1), K intersects this closure, as well as its complement. But this contradicts the connectivity of K. So the closures of BQ and BI intersect. Since there is a subfamily £ of IK such that K C \J{B* :B e £}
this proves that the diameter of K is at most E. So an appeal to Theorem 3.5.9 finishes the job.
D
Exercises for §3.5.
1. Let X be a space and let Y be a compact space for which there is a continuous function /: X —> Y. Prove that there is a dimension preserving compactification ^X of X such that the function / can be extended to a continuous function g: ^X —>• Y. ^•2. Let ( X n , fn)n be an inverse sequence consisting of at most m-dimensional spaces. Prove that its inverse limit is at most m-dimensional. 3. Let X be a space which is not compact. Prove that X has a compactification *yX such that dim'jX = oo. >4. Give an example of a compact space X containing two disjoint closed subsets A and B such that no partition between A and B is irreducible. 5. Let X be a subspace of the compact space Y. Prove that the following statements are equivalent: (1) X is L-imbedded in Y. (2) For every e > 0 there is a neighborhood U of Y in X such that every continuum in U has diameter less than e. >-6. Let A and B be L-imbedded subspaces of the compacta X and Y, respectively. Prove that A x B is L-imbedded in X x Y. Conclude that A x B is at most one-dimensional. 3.6. Mappings into spheres In the previous sections we presented several characterizations of dimension, some 'internal' (e.g., Theorem 3.2.14(2) and (3)) and some 'external' (e.g., Theorem 3.2.14(4)). In this section we shall prove that a space X is at most n-dimensional if and only if every continuous function /: A —>• §n, where A C X is closed, can be extended over X, thereby deducing another important 'external' characterization of dimension. As an application of this result we shall present a proof of the Brouwer Invariance of Domain Theorem. We first formulate and prove two lemmas that are needed in the proof of the main result in this section. Lemma 3.6.1. Let X be a space and let A\ and A? be disjoint closed subsets of X such that 0 < dim (X \ (A\ U ^2))
194
3. BASIC DIMENSION THEORY
Proof. Put A = A\ U A 2 . By Corollary A. 4. 3 there exist open subsets U and V_ of X such that AI C U, A2 C V and UnV = 0. Since dim (A" \ A) < n and U H (X \ A) and V C\ (X \ A) are disjoint closed subsets of X \ A, by Corollary 3.4.14 there exist disjoint open sets E and F in X \ A such that
(1) u_n(X\A)cE, (2) VH(X\A}CF, (3) if L = (X \ A) \ (E U F) then dim L < n - 1. Since AiUE = UUE, and J5 is open in X being open in the subspace X\A, we conclude that AI U E is open in X. Similarly, A2 U F is open in X. Consequently, L is closed in X and is a partition between AI and A2 in X with dimL < n — 1 by (3). So L is as required. D Corollary 3.6.2. Let X be a space and let AI and A2 be closed subspaces of X such that 0 < dim (X \ (Ai U A 2 )) < n. Then there exist closed subspaces Xi and X2 of X such that (1) (2) (3) (4)
Ai=Xin(AiUA2), A2 = X 2 n ( A i U A 2 ) , dim ((Xi n X2) \ (Al U A 2 )) < n - 1, ^!UX 2 = X.
Proof. Put F = JC \ (Ai n A 2 ). By Lemma 3.6.1 there exists a closed set L in Y with dim L < n — 1, such that F \ L can be written as the disjoint union of two open sets E and F such that AI D Y C E and A2 n Y C F. It is easy to see that Xi = AI U E U L and JC2 = A2 U F U I/ are as required. D These simple results enable us to derive the following: Theorem 3.6.3. Let X be a space and let A be a closed subspace of X such that 0 < dim(X \ A) < n. Then every continuous function g: A -> §>n can be extended to a continuous function g: X —> §n. Proof. We shall prove the theorem by induction on n > 0. Suppose first that n = 0 and recall that S° = { — 1, 1}. By Corollary 1.5.2 there exists a clopen set C C X such that g'1^!) C C and g~l(l) C X \ C. Define the function g : X -t §° by
-i
xeC
An easy check shows that g is as required. Now assume that the theorem is true for n — 1, n > 1, and let g: A -» §n. Define §? = {x € §n : xi > 0} and §£ = {x G §n : Xi < 0}, respectively. Observe that §^ and §2 are both homeomorphic to I n and that S^ n Sg is
3.6. MAPPINGS INTO SPHERES
195
homeomorphic to ^n~1. Now put AI — g~v^\ and A2 = g~l[$2], respectively. Let Xi and X2 be such as in Corollary 3.6.2 for AI and A2. By our inductive assumption, we can extend g \ A\ n A2 to a continuous function
Now for i = 1,2 define &: AlU(X1nX2) -> S? by gl = (g\Ax)Uh. Since S? is homeomorphic to I n it is an AR (Corollary 1.2.12) and so we can extend gi for i = 1, 2 to a continuous function #; : Xi —> §". Clearly,
is a continuous extension of g.
D
Corollary 3.6.4. Let X be a space and ]et /, g : X —> §n be continuous. If
is at most (n — 1) -dimensional then f and g are homotopic. Proof. Consider the product X x I and put
Then A is closed in X xl (Exercise A. 1.9) and the function £: A -> §" defined by f /(*) (* = 0), ^(x,*) = «{ f ( x ) = g ( x ) (xeX\D(f,g)), { 9(x] (t = 1), is clearly continuous. In addition, (X x I) \ A is contained in
and hence is at most ((n — 1) + l) -dimensional by Theorem 3.4.12. So Theorem 3.6.3 implies that £ can be extended to a continuous function £: X x l - > § n , which means that / and g are homotopic.
D
Theorem 3.6.5. Let X be a space and let n > 0. TLe following statements are equivalent: (1) dimX
196
3. BASIC DIMENSION THEORY
every i < n + I let oti : X —»• I be a Urysohn function such that cti \ Ai = 0 and al \ Bl = 1 (Corollary A.4.1). Define a: X -> I n+1 by a(x) = ( a i ( x ) , . . . ,a n + i(a;)). Put A = \J*=i(Ai U Bi) and (3 = a \ A. Then j3[A] is contained in the boundary B of I n+1 which is homeomorphic to S n by Exercise 1.1.24. By assumption, there exists a continuous extension 7 : X —> B of J3. Now for every i < n + 1 put E^ = (^i ° ^}~l(lli}- Then clearly E^ is a partition between Ai and Bi for every « such that HlLi Ei = ®. Consequently, r is inessential. D Application 1: The Brouwer Invariance of Domain Theorem. To begin with, we shall first present an 'internal' characterization of the boundary points of an arbitrary closed subset of a fixed Rn . Proposition 3.6.6. Let n 6 N and let X be a closed subspace ofW1. Then a point x in X belongs to the boundary Fr(X) of X in W1 if and only if x has arbitrarily small neighborhoods U in X such that every continuous function g : X \ U -> Sn~1 can be extended continuously over X. Proof. First assume that x € Fi(X). Without loss of generality, x = 0. If V is a neighborhood of 0 in X then there is e > 0 such that B(x,e)r\X C V. Consequently, it suffices to consider for some e > 0 a 'spherical' neighborhood .0(0, e) of 0 and a continuous function g : X\B(Q,e") -» S71"1. Without loss of generality assume that e = 1 which means that the boundary of 5(0, e] is $n-1. Since 0 belongs to the boundary of X, there is a point q e B(Q, 1) \ X. For each p e Bn \ {q} let r(p) denote the 'projection' of p on S71"1 from q. Put X0 = X n Bn and Xi =X\B(Q,1), respectively. Then X0 and X1 are closed in X, cover X, and have the following additional property:
X0 n S71-1 = Xi n S71"1. Since dimS n-1 = n - 1 (Theorem 3.2.12), the function g \ Y can be extended to a continuous function g \ Y : S71"1 —>• S71"1 by Theorem 3.6.5. In addition, define h: XQ —>• S71"1 by h — g \ Y o T. Observe that r(y) = y for y 6 y so that This means that the functions 5 and /i agree on Y from which it follows that the function g = h Ug is the desired continuous extension of g.
3.6. MAPPINGS INTO SPHERES
197
Conversely, let x e X and assume that x is an interior point of X. There is e > 0 such that B = D ( x , e ] C X. Let U be an open neighborhood of x in X such that U C B(x,e). We identify Bn~l and the boundary of B. For each p G X \ U let r(p] denote the 'projection' of p on Sn-1 from x. Then r cannot be extended over X since this would yield a retraction from J5 onto its boundary, which is impossible by Theorem 2.4.10. D Corollary 3.6.7. Let n € N and Jet X and Y be closed subspaces ofW1. If f : X —> Y is a homeomorphism then f[Fr(X)] — Fr(F). We are now in a position to present a proof of the following interesting: Brouwer Invariance of Domain Theorem 3.6.8. Let n G N and let U be an open subset of M n . If/: U —> W1 is injective and continuous then the following statements hold: (1) f[U] is open in W1, (2) / : £ / — ) • f[U] is a homeomorphism. Proof. We shall first prove (1). Take an arbitrary x G U. We shall prove that f ( x ) belongs to the interior of f[U]. There exists e > 0 such that D(x,e) = {y€Wl
:\\x-y\\<e}CU.
Since B is compact, by Exercise A. 5. 9, / \ B is a homeomorphism. By Corollary 3.6.7 we conclude that f(x] belongs to the interior of f[B] and hence to the interior of f[U]. We shall now prove (2). This is easy. If V is an open subset of U then (1) applied to / f V shows that f[V] is open in W1 and hence in f[U]. We conclude that / : £ / — » f[U] is an open mapping and therefore by injectivity is a homeomorphism. D The question naturally arises whether something like Theorem 3.6.8 can also be derived for the class consisting of all infinite- dimensional linear spaces. As to be expected, this is not possible. For the first part of the theorem, this can be demonstrated quite easily by considering Hilbert space I2. The function / : I2 —> I2 defined by
is an imbedding with nowhere dense range. We now turn to the second part of Theorem 3.6.8. Theorem 3.6.9. For an infinite-dimensional normed linear space L there exists a bijective continuous function f : L —> L such that f is not a homeomorphism.
198
3. BASIC DIMENSION THEORY
Proof. Let S be the unit sphere in L. Since L is infinite-dimensional, S is not compact by Exercise 1.1.22. Consequently, by Exercise A. 5. 13 there exists a continuous function A: S —>• (0,1] such that infA(S) = 0. Now define f:L-^Lby the formula
/(O) = Q. Claim 1. If x £ L and a E [0, oo) then f ( a x ] = a f ( x ) . Proof. This is a triviality. If x = 0 or a = 0 then there is nothing to prove. In addition, if x ^ 0 and a ^ 0 then OLX \ —-J -ax = a f ( x ) , \ax\. as required. 0
(
We shall now prove that / is continuous. It is clear that this need only be checked at the origin. To this end, let (xn}n be a sequence in L \ {0} such that limn^.00 xn = 0. Then II £( Nil \ l *^n 1 I I l l ^ - r l l I I n—too n ||/(a:n)|| = A f Tj-^ J ' \\Xn\\ < I • \\Xn\\ -> 0. H^nll '
We conclude that limn_j.oo f ( x n ) = 0. We shall next prove that / is injective. To this end, take x,y € L and assume that f ( x ) = f ( y ) . Without loss of generality, x ^ 0. Then f ( x ) ^ 0 so that f ( y ) ^ 0 from which it follows that y ^ 0. So f ( x ) = f(y) gives us that y
MM
Ml'
-y,
and since A(5) C (0,1] this implies that there exists a > 0 such that x = ay. By the claim, this yields f ( x ) = a f ( y ] , i.e., a = I. We conclude that x = y. Now take y € L \ {0}. Put
x=
y
An easy application of the claim gives us that f ( x ) = y. We shall now prove that / is not a homeomorphism by showing that /-1 is not continuous. Since inf A(5) = 0, we can choose a sequence (xn)n in S such that lira A(x n ) — 0. n—¥oo
It follows that \\f(xn)\\=X(xn}-
\\Xn\\ ^ 0 - 1 = 0 ,
so f(xn} —> 0. However, \\xn\\ — 1 for every n, so xn -ft 0.
D
3.6. MAPPINGS INTO SPHERES
199
Application 2: Dimension and mappings. We shall now discuss the question how continuous mappings can lower or raise dimension. The results obtained earlier in this section are important in the proof of Theorem 3.6.10 below. They are not used in the proof of the subsequent Theorem 3.6.14. However, Theorem 3.6.14 is motivated by Theorem 3.6.10 and so it is natural to discuss it here. We already know that every compact space is a continuous image of C (Theorem 1.5.10). So continuous closed surjections, can arbitrarily raise dimension. In general it is therefore not possible to conclude anything. Certain conditions have to be imposed. It is natural to look at the fibres of the map under consideration. If something can be said about their dimensions, then our intuition tells us that there should be a relation between the dimensions of the domain and the range of the map. This is what the next result is about. Theorem 3.6.10. Let X and Y be spaces and let f : X —>> Y be a continuous closed surjection. If there is an integer k > 0 such that dim/~ 1 (y) < k for every y 6 Y then dimX Proof. Wre may assume without loss of generality that dimy < oo. We shall prove the theorem by induction on dim.y. If dimF = — 1 then X and Y are empty and so the theorem trivially holds. Now assume the theorem holds for all spaces X and Y and all continuous closed surjections f : X — t Y such that dimy < n, where n > 0, and assume that we are in the situation that Y is a space such that dim Y = n. Claim 1. If y € F and U is an open neighborhood of f ~ l (y) in X then there is an open neighborhood V of f ~ l ( y ) such that V C U while moreover dim Fr V < n - 1 + k.
Proof. Put W = Y \ f[X \ U]. Since the map / is closed, W is an open neighborhood of y. Let E be an open neighborhood of y in y such that EC E C W
while moreover dimFrE < n — I (Theorem 3.4.10). Then V = f~1[E] is as required since clearly V CU and FrV = Fr/-![£] = f~l[E] \ f~l[E} C rl[E] \ /^[E] = f~l[F,E] and so dim FrV < n — 1 + A; by the Subspace Theorem 3.2.9 and our inductive hypothesis applied to the closed map / \ f'1 [Fr E]: f'^Fr E] -> Fr £. 0 Put m = n + k. Now let A be an arbitrary closed subset of A", and let h: A —>• S m be continuous. We will prove that h can be extended to a
200
3. BASIC DIMENSION THEORY
continuous function h : X —> S m . It will then follow that dim X < m = n + k by Theorem 3.6.5. Let y 6 Y be arbitrary. Then dimf~l(y) < k < n + k and so we can extend h continuously over A U f ~ 1 ( y ) (Theorem 3.6.3). Since STO is an ANR, this function can be extended over a neighborhood Uy of A U f ~ l ( y ) (Corollary 1.2.13). By Claim 1 there is an open neighborhood Vy of f~~l(y) such that Vy C Uy while moreover dim Fr Vy < n — 1 + k. So h can evidently be extended over A\JVy. The open cover V = {Vy : y £ Y} has a countable subcover by Corollary A. 2. 3, say {Vyi : i G N}. We shall inductively define a sequence of continuous functions hi : A U Uj-=i Vy, -^ Sm such that z-l
(*)
hi\A\J \JVyj =hi-!, 3=1
for every i > 1. Let hi be an arbitrary continuous extension of h over AuVyi . Assume that the functions hi, . . . , hi-\ have been defined satisfying (*). Put t-l
i-l
3=1
3=1
respectively. Observe that K U L = A\J Uj=i ^y> • By the above we can extend the function / to a continuous function g: L —»• S m . Observe that i—l
i—l
i—l V
D = {xeLnK: h^(x) ± g ( x ) } C P = (J ^w \ U y, = U Frl ^' j=i j=i j=i Since dimP < n — l + k — m — 1 by the Countable Closed Sum Theorem 3.2.8, we have that dimD < m — 1 by the Subspace Theorem 3.2.9. We conclude that the functions hl-i \Kr\L and g \Kr\L are homotopic by Corollary 3.6.4. The Borsuk Homotopy Theorem 1.4.2 implies that the function h^-i \ Kr\L can be extended to a continuous function g: L -> Sm. Since hk-i and g agree on the closed set K n L it follows that their union is the desired function hk • This completes the inductive construction. Now define h: X ->• §m by
/i(x) = /ij(x)
(x e vyi}.
Then /i is clearly a continuous function extending h.
D
As we saw above, closed continuous surjections can arbitrarily raise dimension. It is natural to ask whether the same can happen for open continuous surjections. But things are slightly different here. For observe that if X is zero-dimensional, and / : X —> Y is an open and closed continuous surjection, then Y is zero-dimensional (Exercise 1.5.5). So open maps defined on
3.6. MAPPINGS INTO SPHERES
201
zero-dimensional compacta do not raise dimension. In contrast, open maps denned on one-dimensional compacta can arbitrarily raise dimension, as we will prove after the completion of some preparatory work. Proposition 3.6.11. For each n > 1 there exists a continuous function fn : Bn+l ->• 2s" such that for every x G S n , fn(x) — {x}. Proof. Define fn : Bn+l ->• 2 s ™ by t (r\ Jn\^J
{ {§-} - \) Lr f J \ I
f _ I _ on . Jn(X) —
Z —
x
i
-
^9 9 || II I 5- Z — ^ • ||X||
(z = 0), f i n\ (^T ^ UJ.
It is easy to visualize this function e.g., for n — 1. It is clear that fn is continuous and that / n (z) = {x} for every x 6 §n. D Lemma 3.6.12. Let X be a compact space with closed sets A,B,E and F such that A n B = 0 and E n F — 0. Then there exist a compact space Y and an open surjection f:Y-$X such that the collection of pairs is inessential in Y. Proof. Put Z = (A U B) U (E U F). In addition, pick Urysohn functions £,77: Z ->• I such that £[A] = Q,£[B] = l,rj[E] = 0 and n[F] = I (Lemma A. 4.1). The function h(x) = (£(x),ri(x)) 2
(xeZ) 2
maps Z into the boundary of I . Let h: X —>• I be a continuous extension of h (Corollary 1.2.12). By the fact that I 2 w B2 under a homeomorphism which (necessarily) maps its boundary W = dl2 onto S1 (Exercise 1.1.24), there exists by Proposition 3.6.11 a continuous function g : I 2 -> 2W such that g(x) — {x} for every x G W. Now define Y= U ( W x 0 ( & ( x ) ) ) C X x W x£X
This is the space Y we are looking for. We first claim that Y is closed in X x W. To this end, suppose that (xn,yn)n is a sequence in Y converging to an element ( x , y ) & X x W. Then xn —>• x and so 51 o h(xn] —> g o /i(x). Since yn£ g° h(xn} for every n, and y n -> y, it therefore follows that y G g o /i(x)
(Lemma 1.11.2(2)). This proves that Y is closed, whence compact.
202
3. BASIC DIMENSION THEORY
Now let / = ?TI \ Y —> X , where TTI : X x W —>• A^ is the projection. This is the map / we are looking for. The function F: X -> 2XxW defined by
F(x) = {x} x [0 o h(x)] is continuous because g and /i are. As a consequence, / is open by Exercise 1.11.16. Let 7T2 : X x W —> W denote the projection. Now take arbitrary elements (x,y] G /-1[A] and (p,q) 6 f~l[B\. Then obviously x € A and y G g ( h ( x ) ^ . But since x G Z, it follows that h(z) = h(x) E W and so
g(h(x))={h(x)}. So this implies that ( x , y ) = (x, /&(#)). Similarly it follows that p € B and that (p, g) = (p, h(p)} . Since h(x) ^ h(p) and (x, y) and (p, g) were arbitrarily chosen, this implies 7T2 [/^[A]] n 7T2 [/"^.B]] = 0. It follows similarly that
This implies that the collection of pairs is by Exercise 3.1.11 inessential since dimVF = 1.
D
Corollary 3.6.13. Let X be a compact space. Then there exist a compact space Y and an open surjective map /: Y —» X such that if A,B,E and F are arbitrary closed subsets of X such that A n B = 0 and E n F — 0 then the pairs are inessential. Proof. Let £ = {(En: Fn) : n e N} be a collection of pairs of disjoint closed sets for X such as in Lemma A. 5. 6. That is, if (A, B) is any pair of disjoint closed subsets of X then there exists n e N such that A C En and F C Bn. Let {(rii , mi) : i G N} be an enumeration of all pairs of natural numbers. We aim at applying Lemma 3.6.12 repeatedly. In the first step of the construction we find a compact space Y\ and an open continuous surjection /o : Y\ —> X such that the collection of pairs (**)
is inessential. Then we again apply Lemma 3.6.12 but now with Y\ and the collection of pairs
3.6.
MAPPINGS INTO SPHERES
203
obtaining an open surjection /i : Y-2 -» YI such that the collection of pairs
{ (/r1 1/o"1 [En2]} , /r1 [/o-1 [Fn2]] ) , (/r1 [.fir1 (Em2]\ , /r1 [/o-1 [^2]] ) } ,
is inessential. Observe that by (**) and Exercise 3.1.12 it follows that the collection of pairs is inessential as well. So now it is clear how to proceed. In an infinite process we can deal with all pairs in the collection £, thus obtaining an inverse sequence y
/O
y
fl
Y
_
_
V"
/"
y
with continuous, surjective, open bonding mappings. If YOO denotes its inverse limit, then the projection /oo : Y^ -> X is an open continuous surjection by Exercise 1.10.5. In addition, if A,B,E and F are closed subsets of X such that A n B — 0 and E n F — 0 then there exist n, m e N such that Since by construction the pairs { (&&„], &[Fn]),
(/^[
are inessential, the same holds for the pairs So the space 1^, and the map /^ are as required.
D
This leads us to the following interesting result. Theorem 3.6.14. Let X be a compaction. Then there exists an at most one-dimensional compactum Y which can be mapped onto X by an open map. Proof. By applying Corollary 3.6.13 we obtain an inverse sequence y
_ ~y*
/(I
y
fl
y
_
_
y
/».
of compact spaces, with continuous, open and surjective bonding mappings, having the following additional property: if for some n > 0, A, B,E and F are closed sets in Yn such that A n B = 0 and E n F = 0 then the collection of pairs is inessential in Yn+\. Let Y^ denote the inverse limit of the sequence. By similar arguments as the ones in the proof of Corollary 3.6.13 it follows that for every n > 0 the canonical projection /^ from 1^ onto Yn is an open surjection. So we are done if we can prove that dim YOO < 1. But this follows by a compactness argument. For let A:B,E and F be closed subsets of Y^ such that A n B = 0 and E n F = 0. By Exercise 1.10.12 it follows that
204
3. BASIC DIMENSION THEORY
there exists n € N such that f^[A] C\ f£[B] = 0 and /£[£] n /£[F] = 0. By construction we have that the pairs are inessential. So now we are done by using Exercise 3.1.11 and by observing that /"+1 [4 C/- 1 ^ [4], etc. D Exercises for §3.6. ^•1. Let n > 0 and let X be a compact space which can be written as the union of closed sets Fi, i (E N, such that for all distinct i,j £ N, dim(F; n F j ) < n. Prove that every continuous function from F\ into Sn is continuously extendable over X. 2. Observe that the previous exercise is a generalization of Theorem A. 10. 6. 3. Let X be a compact space and let H be the union of all components of X of dimension at most n. Prove that dim H < n. ^•4. Let A be a closed subspace of a space X such that dim(X \ A) < n, let 0 < k < n, and let /: A —> S fc be continuous. Prove that there exists a closed subspace B of X such that A D B — 0 and dim B < n — k — 1, while moreover the function / can be extended over X \ B. 5. Let n > 2 and consider the n-dimensional cube I n . Define En~l = {(xi,...,xn-i,l) : 0 < xl < l,i = l , . . . , n - 1}. Prove that if X = ln \ En~l then def X = n - 1. 6. Let X be an at most (n — l)-dimensional space. Prove that every continuous function /: X —> Sra is nullhomotopic. 7. Prove that S" is unicoherent for every n > 2. 3.7.
Dimension of subsets of W1 and certain generalizations
In this section we prove some interesting dimensional properties of subsets of the Euclidean spaces Rn , n G N. Theorem 3.7.1. Let X C Rn . Then
dimX = n
Proof. First, if Int^T ^ 0 then X contains a homeomorph of In and is therefore n-dimensional by Theorems 3.2.12 and Corollary 3.4.14. Next, assume that Int X = 0. There consequently exists a countable set D C W1 \ X which is dense in W1 .
3.7. DIMENSION OF SUBSETS OF R™ AND CERTAIN GENERALIZATIONS
205
By Proposition 3.2.10, Rn = 9tnj0 U 9tn,i U • • • U
Notice that 9ln,n consists of all points of IRn having rational coordinates only, and hence is a countable dense set. By Proposition 3.2.10 and Corollary 3.1.7 we get that dim Y < n — I . By the countable dense homogeneity of W1 (Corollary 1.6.10), there is an element h G CK (W1} such that h[D] = ftn,n. This implies that h[X] C Y, and hence by Theorem 3.2.9 that dim^C = dim/i[.Y] < n — 1, as required. D Corollary 3.7.2. Let U C W1 be nonempty and open. If U is not dense then dim Fr U = n — I . Proof. Put F = FrU. Then by Theorem 3.7.1 it follows that dim F< Pick an arbitrary point m £ Mn \ U.
n-l.
Striving for a contradiction, assume that dim F < n — 2. We will first prove that we may assume without loss of generality that U is compact. For this we will use a trick on 'exchanging' points at infinity. The one-point compactification X — W1 U {00} is homeomorphic to Sn (Exercise A. 4. 8). The boundary of U in X is F if it is compact and F U {00} otherwise. Since dimF = dim(F U {oo}) (Exercise 3.2.1) we arrive at the conclusion that the boundary of U in X is at most (n — 2)-dimensional. Now observe that Y — X \ {m} is homeomorphic to R n , that the closure of U in Y is compact, and that its boundary is at most (n — 2)-dimensional. So from now on assume that U is compact in Rn and that FrC/ is at most (n — 2)-dimensional. We may assume without loss of generality that the origin 0 of Mn belongs to U. By compactness of U, the collection {r-U:r G (0,1]} is a local base at 0 consisting of compact neighborhoods all of whose boundaries are at most (n — 2)-dimensional. But since W1 is homogeneous, the same is true for all points of E n . By Theorem 3.4.10 (applied twice) this implies that dim W1 < n - 1, which contradicts Theorem 3.2.12. D Corollary 3.7.3. Let L be a closed subset ofW1 with dim L < n-2. Then L does not separate W1 . Proof. Suppose that Mn \L = UUV, where U and V are disjoint nonempty open sets. Then Fr£7 C L and so dimFrL < n — 2 which contradicts Corollary 3.7.2. D
206
3. BASIC DIMENSION THEORY
We now aim at proving in Theorem 3.7.6 a much stronger version of this corollary. First we need to do derive some elementary results about essential families and continua. Proposition 3.7.4. Let X be a compact space, let {(Ai,Bi) : i G F} be an essential family of pairs of disjoint closed subsets of X and let F(0) C F. If for each i E F(0), Li is a partition between Ai and Bi in X and n G F \ F(0) then L = H w ^ contains a continuum from An to Bn. Proof. Suppose that L does not contain any continuum from An to Bn. Let H = An n L,
G = BnnL,
and put
£ = {C : C is a component of L and C n H / 0},
E = (J £,
3" = {C : C is a component of L and C n G ^ 0},
F = |J J,
and
respectively. From Proposition A. 10. 3 it follows that both E and F are closed. Also, by assumption, E D -F = 0. Take an arbitrary C E £. Since L \ F is an open neighborhood of C in L, by Lemma A. 10.1 there exists a clopen (in L) set Uc which contains C but misses F. By compactness, finitely many t/c's cover .E. We conclude that 0 is a partition in L between H and G. By Corollary 3.1.5 there is a partition T in X between An and Bn such that T n L = 0. But this contradicts the fact that is essential (Exercise 3.1.1).
D
Corollary 3.7.5. Let X be a compact space, let be an essential family of pairs of disjoint closed subsets of X and let n G F. Suppose that Y C X is such that Y meets every continuum from An to Bn. For each i G F \ {n} let Ui and Vi be disjoint closed neighborhoods of Ai and Bi, respectively. Then is essential in Y. Proof. By Proposition 3.7.4, if for every i G F \ {n} the set Li is a partition between Ai and Bi then Y^r\i^r\{n} ^i ^ ®- Now apply Exercise 3.1.13. D A space X is called continuum- connected provided that for any pair of points x,y G X there is a subcontinuum C of X containing both x and y. A continuum-connected space is obviously connected, but the converse need not be true (Exercise 3.7.1).
3.7. DIMENSION OF SUBSETS OF R™ AND CERTAIN GENERALIZATIONS
207
Theorem 3.7.6. Let M C G, where G C W1 is open and connected. If M is at most (n — 2)-dimensional then G\M is continuum-connected. Proof. We first prove the special case that G = W1. Take two arbitrary distinct points x, y G Mn \ M and let K denote the closed ball in Rn with center 1/2(x + y] and radius l/2\\x — y\\. Let /: ln —>• K be a continuous surjection which maps a pair of opposite faces A and B of I n to x and y, respectively, and has the property that the restriction g — / \In\ (A\JB] is a homeomorphism from In\ (A\JB) onto K\ {x, y}. Such a function can easily be constructed with the technique used in the solution of Exercise 1.1.24. The set f ~ l [ M r \ K ] is at most (n — 2)-dimensional. Hence by Corollary 3.7.5 there is a continuum C from A to B which misses f~l[Mr\K]. But then f[C] is a continuum containing x and y but missing M. Now consider an arbitrary open and connected set G C W1 and arbitrary points x,y £ G\ M. By Lemma 1.5.21 there exist open balls BI,B<21 • • •,B^ in G such that x E B\, y £ Bk and BiC\Bi+\ ^ 0 for alH < fc — 1 (here we use that G is open and connected). Since M has empty interior by Theorem 3.7.1 there exists for i < k— 1 a point Zi € (BiC\Bi+i)\M. Put ^o = x and 2^ = y. By the above for i
the complement X \ L is connected. The examples of n-dimensional totally disconnected spaces for all n that will be constructed in §3.9 show that a generalization of this concept to noncompact spaces makes no sense; simply observe that the singletons are the only connected subsets of a totally disconnected space. Observe that each Cantor-manifold is a continuum and that the one-dimensional Cantor-manifolds are precisely the one-dimensional continua. Proposition 3.7.7. Every n-dimensional compactification of W1 is an ndimensional Cantor-manifold. In particular, §n and In are n-dimensional Cantor-manifolds. Proof. Let 7Mn be an n-dimensional compactification of R n , and let A in jX be at most (n — 2)-dimensional. Then B = W1 \ A is connected by Corollary 3.7.3. In addition, A n W1 has empty interior in W1 by Theorem 3.7.1
208
3. BASIC DIMENSION THEORY
which implies that B is dense. We conclude that ^W1 \ A contains the dense connected set B and is therefore connected itself. D We will now show that there are 'many' n-dimensional Cantor-manifolds. Theorem 3.7.8. Let n > 1. Every compact n-dimensional space X contains an n-dimensional Cantor-manifold. Proof. By Theorem 3.6.5 there exists a closed subset A C X and a continuous function /: A —> Sn-1 such that / admits no continuous extension over X. Let "B = {Bi : i 6 N} be a countable open base for X. We define a decreasing sequence (Xi)i>o of closed subsets of X, as follows. First, let XQ = X. If Xi has been defined then we consider two cases. If / does not admit a continuous extension over A U (Xi \ Bi+i) then put
otherwise put We claim that M = Di^o -^ *s an ^--dimensional Cantor-manifold. Observe that by construction, for no i the function / admits a continuous extension over A U Xi . Claim 1. / does not admits a continuous extension over A U M. Proof. Striving for a contradiction, assume that f : A\J M —>• S71"1 is a continuous extension of /. By Corollary 1.2.13 it follows that there is an open neighborhood U of M such that / can be extended to a continuous function f:A\JU-> S™"1. Pick N so large that M C XN C U (here we use that X is compact, see Exercise A. 5. 17). Then / \ (Au XN) extends / which is impossible by construction. 0 So by Theorem 3.6.3 it follows that dimM > n. Since dimM < n being a subspace of X we conclude that dimM = n. To complete the proof it clearly suffices to prove that if MI and M2 are proper closed subsets of M with M = MI U M2 then dim (Mi n M2) > n — 1. Striving for a contradiction, assume that there exist proper closed subsets MI and M2 of M with M = M!UM 2 ,
dim(Mi n M2) < n - 2.
Put AI - A U MI and A2 = A U M2, respectively. Claim 2. / admits continuous extensions over AI and A 2 , say to /i and /2, respectively.
3.7. DIMENSION OF SUBSETS OF R n AND CERTAIN GENERALIZATIONS
209
Proof. It suffices to prove this for A\. To this end, pick x G M2 \ MI (here we use the fact that MI and M2 are proper subsets of M). Pick i G N so large that x E -B;+i C X \ M\. Consider the set Xi+i. It contains x and hence it intersects Bi+i. By construction this can only happen when / can be extended continuously over A U (Xl \ Bl+i). Since A\ C A U (Ar; \ -B;+i), we are done. 0 Put 5 = A U (Mi n M 2 ). The functions /i f # and /2 \ B only differ in a subset of MI n M2, and hence on a set of dimension n — 2 at most. We conclude that f i \ B and /2 f Z? are homotopic by Corollary 3.6.4. Since /2 \ B admits a continuous extension over A U M2 , the Borsuk Homotopy Extension Theorem 1.4.2 implies that the same is true for /i \ B, say to the function f[ : A U M2 ->• §n~l. Since B= ( A u M ! ) n ( A u M 2 ) , it follows that the formula
defines a continuous extension of / over AU M. This is a contradiction. D This yields another solution to Exercise 3.2.4. Corollary 3.7.9. If X is compact and dim X = n then one of its components is n- dimensional. Exercises for §3.7. 1. Give an example of a connected space which is not continuum-connected. 2. Give an example of a space X having the property that for every component C C X we have dimC < 3. Let X be compact and let A and B be disjoint closed subsets of X. In addition, let L be a closed subset of X such that no continuum in L intersects both A and B. Prove that there is an open neighborhood U of L such that no continuum in U intersects both A and B. >4. Let X be a compact space, let {(Ai,Bi) : i G F} be an essential family of pairs of disjoint closed subsets of X and let F(0) C F. Suppose that for each i G F(0), Li is a partition between Ai and Bi in X and let
L= ier(o) n L,.
Finally, let m,n 2 G F \ F(0) be distinct. Call a subset T C X small if no continuum in T meets both Ani and Bm .
210
3. BASIC DIMENSION THEORY
Prove that if N is a finite pairwise disjoint collection of small closed subsets of X then there is a continuum in from An2 to Bn2 .
Let X be a compact space such that dimX > 2. Prove that there are real numbers 5 > 0 and e > 0 such that if N is any finite collection of pairwise disjoint closed sets with mesh(!N) < 5 then there is a continuum C Q X \ (J N with diameter at least e. Let X be a compact space with dimX < I . Prove that for every real number e > 0 there is a family K of pairwise disjoint closed subsets of X with mesh(N) < e such that each continuum C C X \ (J [NT has diameter less than s. 3.8. Higher-dimensional hereditarily indecomposable continua Indecomposable continua are strange objects. We proved their existence in §1.10. In this section we go one step further, and prove the existence of 'many' hereditarily indecomposable continua of arbitrary large dimension. Higher-dimensional hereditarily indecomposable continua were first constructed in 1951. But only the last decade it was realized that they are fundamental objects that can be used for the solution of various open problems in dimension theory. See the notes for more information. The following result is the heart of the construction of hereditarily indecomposable continua. Theorem 3.8.1. Let X be a compaction and let F0,Fi be disjoint closed sets in X. Then there are disjoint open neighborhoods WQ and W\ of FQ and FI, respectively, such that any two intersecting continua K0, KI in are comparable, i.e., K0 C KI or KI C K0. Remark 3.8.2. A compact space X is called Bing provided that for all subcontinua A and B of X we have Ar\B = ® or A C B or B C A. So a zerodimensional compact space is Bing. Theorem 3.8.1 shows that disjoint closed sets in a compact space can always be partitioned by a Bing compactum. Proof. Let g be an admissible metric on X. The Euclidean metric on I will be denoted by d. Let P be the union of the five open rectangles in I 2 , described in Figure 9. The set I 2 \ P is the union of the disjoint open sets MQ and MI in the picture. Clearly, M0 H MI = 0 . Observe that (1)
(OjxlCMo,
{l}xICMi.
3.8. HIGHER-DIMENSIONAL HEREDITARILY INDECOMPOSABLE CONTINUA 211
Hence P is a partition between {0} x I and {1} x I in I2. Let WQ,O and W1)0 be open neighborhoods of FQ and F\, respectively, such that /O\
Tf7
yZ.)
o TT/
(ft
K^O o I I ''1 0 — W-
5
/6
2
/3
P
Vs
Figure 9. In addition, let F = {/; : z e N} be a dense subset of C(-X", E) (Proposition 1.3.3). Recursively, determine for every i > I open sets WQ^ and W\^ in X and a continuous function M; : X —>• I 2 , by the following formulas: (3)
^(x,Wo,i-i)
f(
Q(X,
(4)
Since Wo,onWo,i = 0 5 it follows that u\ is well-defined and continuous. Notice that by (1) and the fact that the first component of u\ is an appropriate Urysohn function, we have Ui[W0i0] C {0} x I C Mo,
WitVTi.o] C {1} x I C Mi.
By (4) and the fact that MQ fi MI = 0, we consequently obtain that By induction on i this reasoning can easily be extended. So oo
W0 = U W0ti,
oo
= U wi>*> i=0
212
3. BASIC DIMENSION THEORY
are disjoint open neighborhoods of FQ and F\ , respectively. We claim that these neighborhoods are as desired. Let KQ and K\ be arbitrary intersecting continua of X such that KoUKi
CX\(W0UWi).
Striving for a contradiction, assume that there exist points ao£K0\Ki,
ai£Ki\K0.
There is by Lemma A. 4. 2 a Urysohn function u: X —> I such that
u(o 0 ) = 0,
u(ai) = 1,
u[K0] C [0, 1/2],
u[#i] C [i/2, 1].
Since F is dense in C(J\T,I), there exists i such that d(u, fi) < l/6. So we have (5)
/,(a0) <E [0, i/6), £(01) e (%, 1], /i[^o] C [0, %], /^i] C [1/3, 1],
hence
(6)
K 0 n/- 1 (V 3 ,i]-0, K 1 n/- 1 [o,y 3 )=0.
This shows that the continua To = Ui[Ko],
T! =ui[Kl],
have the property that (7)
To C I x [0,2/3],
T0 C l x [1/3,1].
Since KQ and K\ are disjoint from Wo,i U Wi^, the continua TO and T\ are contained in P, cf. (4). By (5), TO intersects the strip E x [0, Ye) and, similarly, T\ intersects I x ( 5 / 6 , 1]. Since TO and T\ are continua, the geometry of P and (7) imply that they are disjoint. But this is a contradiction since K0 and KI intersect, hence T0 fl TI / 0. D Corollary 3.8.3. Every (n + 1) -dimensional compactum X contains an ndimensional hereditarily indecomposable subcontinuum. Proof. Let X be an (n + l)-dimensional compact space with essential family {(Ai,Bi) : i < n + I}. An arbitrary partition T between A\ and B\ must be at least n-dimensional by Theorem 3.1.9. Theorem 3.8.1 shows that for T we may take a closed set having the property that each of its components is hereditarily indecomposable (Exercise 1.10.21). One of those components, say C, must have the same dimension as T by Exercise 3.2.4 or Corollary 3.7.9. So C is a hereditarily indecomposable continuum of dimension n or n + 1. But every (n + l)-dimensional compactum contains an n-dimensional compactum by Exercise 3.3.7, and hence an n-dimensional continuum by Exercise 3.2.4. So X contains an n-dimensional hereditarily indecomposable continuum. D
3.8. HIGHER-DIMENSIONAL HEREDITARILY INDECOMPOSABLE CONTINUA 213
Corollary 3.8.4. Every strongly infinite-dimensional compact space contains a strongly infinite-dimensional hereditarily indecomposable continuum. Proof. If X is a strongly infinite-dimensional compactum with essential family {(Ai, BI) : i E N} then an arbitrary partition between AI and BI must be strongly infinite-dimensional by Theorem 3.1.9. By using Exercise 3.1.9 one of the components of such a partition T must be strongly infinite-dimensional as well. We can therefore use the same method as in the proof of Corollary 3.8.3. D Remark 3.8.5. In view of Corollary 3.8.3 and its proof, the question naturally arises whether there exists for n > 2 a hereditarily indecomposable continuum separating the top from the bottom in ln+1. The answer to this question is in the affirmative. Simply observe that I n+1 is locally connected and by contractibility is unicoherent (Corollary A. 12.10). As a consequence, the Bing partition between the top and the bottom of I 3 considered in the proof of Corollary 3.8.3 contains a connected partition by Exercises 3.10.2 and 3.10.3. This partition is clearly a hereditarily indecomposable continuum. Henderson compacta. A compactum X is called Henderson if it is infinite-dimensional but every finite dimensional closed subspace is zerodimensional. It is not clear at all that such a compactum exists. By using Corollary 3.8.4 we will present a surprisingly simple construction of such a space (for a stronger result, see Theorem 3.13.10). Let X be a compact space. For t > 0 and n > 0 put Xn,t — \\{C C X : C is a continuum with dimC < n and diamC > t}. Lemma 3.8.6. Let X be a Bing compactum. Then for every t > 0 and n > 0 we have dim Xnj < n. Proof. Let F be an arbitrary closed subset of X with diamF < t. Claim 1. The components of F of dimension at most n cover Xn,t n F. As a consequence, dim(Xntt n F) < n. Proof. Indeed, let x G Xn^ H F. There is a continuum A contained in Xnj with dim A < n and diamA > t such that x € A. Let B be the component of x in F. Then diam B < t and B n A 7^ 0. So A g. B and hence since X is Bing it follows that B C A. By the Subspace Theorem 3.2.9 we conclude that dimB < n. Hence x is contained in an at most n-dimensional component of F.
214
3. BASIC DIMENSION THEORY
Let H be the union of all components of F with dimension at most n. Then dimH < n by Exercise 3.6.3. We just observed that H D Xnj n F. So we are done by the Subspace Theorem 3.2.9. <£> Since X is compact, there is a finite closed cover 7 of X with mesh less than t. By the claim, Xn,t n F is at most n-dimensional for every F. Hence dimXn:t < n by the Countable Closed Sum Theorem 3.2.8. D This leads us to the following unexpected result. Theorem 3.8.7. Let X be a Bing compaction. The union of all nondegenerate finite dimensional subcontinua of X is countable dimensional. Proof. Every nondegenerate subcontinuum of X has positive diameter. So the set we are interested in is a subset of
71=1
This means that we are done by Lemma 3.8.6, Corollary 3.3.9 and Theorem 3.2.9. D This is all we need for the construction of a Henderson compactum. Corollary 3.8.8. Every strongly infinite-dimensional compactum contains a strongly infinite-dimensional Henderson compactum. Proof. By Corollary 3.8.4 we may assume without loss of generality that X is a hereditarily indecomposable continuum. Let be essential in X. By Theorem 3.8.7 there is a countable family
{N2i :i e N} of zero-dimensional subspace of X covering all finite dimensional nondegenerate subcontinua of X. For every i e N let Ti be a partition between the sets A-2i and B2i such that Tz n N2i = 0. By Theorem 3.1.9 it follows that the compactum
is strongly infinite-dimensional. In addition, T contains no nondegenerate finite dimensional subcontinua. For such a subcontinuum is contained in \J°^=1 NH and hence misses T. We claim that T is Henderson. For if A is a finite-dimensional closed subspace of T which is not zero-dimensional, then one of its components is not zero-dimensional (Exercise 3.2.4) and hence by construction misses T. So this is impossible. D
3.8. HIGHER-DIMENSIONAL HEREDITARILY INDECOMPOSABLE CONTINUA 215
Homogeneity. As we observed in §1.10, the dyadic solenoid £2 is a topological group and hence is homogeneous. So there is a homogeneous indecomposable continuum. Is there an example of a homogeneous hereditarily indecomposable continuum? Surprisingly, there is such a space which is even planar. It is the so-called pseudo-arc which can be constructed by a slightly more complicated procedure than the one used for the indecomposable continuum P in Figure 5 on Page 95. So the construction of the pseudo-arc is not difficult. The verification that it is hereditarily indecomposable is not difficult as well. But the proof of its homogeneity is not simple. This was established by BING in [55]. The pseudo-arc is clearly one-dimensional. So in view of the existence of higher-dimensional hereditarily indecomposable continua (Corollary 3.8.3), the question arises whether such a space can be homogeneous. The answer to this question is in the negative. Proposition 3.8.9. If X is a Bing compaction such that 2 < dim AT < oo then X is not homogeneous. Proof. Let dim AT = n. Let U be a finite open cover of X of mesh less than 1. Since every element of U is an F^-subset of X (Exercise A.2.3), the Countable closed Sum Theorem 3.2.8 shows that there are an element U G U and a compact subset F C U such that dimF = n. Corollary 3.7.9 shows that we may assume without loss of generality that F is a continuum. So X contains an n-dimensional continuum of diameter less than 1. Observe that besides compactness, we did not use any of the properties of X. So by repeating this reasoning, we can construct by induction a decreasing sequence (F;); of n-dimensional subcontinua of X such that diam Fi < l/i for every i. By compactness, the intersection Hi^i -^ ^ ®- In fact, it is a single point since diam Fi < l/i for every i. Let x be the unique point in that intersection. Claim 1. If C is a nondegenerate subcontinuum of X containing x then dimC = n. Proof. For every i, x G Fi n C. Since X is hereditarily indecomposable, this implies that Fi C C or C C Fi (Exercise 1.10.21). So if there exists i such that Fi C C then we are done since by the Subspace Theorem 3.2.9 (applied twice) we then have n = dim Fi < dim C < dim X = n. If there does not exist such i then C C Fi for every i and hence C f) # = i—l
But this contradicts the fact that C is nondegenerate.
216
3. BASIC DIMENSION THEORY
By equality of dim and ind (Theorem 3.4.8), there is a closed subspace G of X such that dimG = n — 1. Corollary 3.7.9 shows that we may assume that G is a continuum. Since n — 1 > 1, the set G is a nondegenerate subcontinuum. Let y be an arbitrary element in G. Then by Claim 1 there does not exist a homeomorphism of X which maps x onto y. Hence X is not homogeneous. D Remark 3.8.10. The proof of Proposition 3.8.9 clearly breaks down if X is infinite-dimensional. It would work if every infinite-dimensional compact space would contain a nondegenerate finite dimensional continuum. But this is not true, as was shown in Corollary 3.8.8 (see also §3.13). By a different technique, ROGERS [354] proved that there is no infinite-dimensional hereditarily indecomposable homogeneous continuum. It is unknown however whether there is an infinite-dimensional homogeneous indecomposable continuum (this is a question of ROGERS). Exercise for §3.8.
1. Let X be a space and let FQ, FI be disjoint closed sets in X. Prove that there are disjoint open neighborhoods WQ and W\ of .Fo and F\ > respectively, such that any two intersecting continua Ko, K\ in X \ (Wo U W\) are comparable, i.e., KQ C K\ or K\ C KQ. 3.9. Totally disconnected spaces In this section we shall prove that totally disconnected spaces of every dimension exist, cf. Examples 1.5.18 and 3.4.13 where we constructed an example of a one-dimensional totally disconnected space. Our main tool is the concept of a so-called 'G^-selection'. After some preparatory work, the construction of the examples is surprisingly simple. The lexicographic order. Suppose that X is compact and that
is a continuous surjection. The set-valued function
F: Y ^ X l
defined by F(y) = f ~ ( y ] in general unfortunately does not admit a continuous selection, see Page 13. We shall show however that there always exists a selection s for F such that the range of s is a G^-subset of X. Let n e N U {00} and consider the product JJn. It can be ordered by the so-called lexicographic order ^, as follows:
x Xy
^=>
(3 n € N) (xn < yn and Xi = yi if i < n).
A straightforward verification shows that -< is a linear order on J n .
3.9. TOTALLY DISCONNECTED SPACES
217
Lemma 3.9.1. Let n € N U {00} and let X C JJn be compact. Then X has a smallest element with respect to ^. Proof. Since X is compact and TTI is continuous, TTI[^] is compact as well and hence x\ — minTTipC] exists. Observe that X n Trf l ( x i ] is a nonempty compact subset of X. So x 2 = min^fX n Trf l ( x \ } ] exists as well. Observe that the set
x n^l(xi) n7T2"1(x2)
is again a nonempty compact subset of X. So the minimum of its projection onto the third factor exists. Etc. It is clear that the point x = (xi}i defined in this way is the smallest element of X with respect to ^ (observe that by compactness of X it belongs to X, even if n = oo). D Theorem 3.9.2. Let X be a compact space and let f : X ->• Y be a continuous surjection. There exists a G^-subset S of X which intersects each fiber of f in precisely one point. Proof. By Corollary A. 4. 4 we may assume that X is a subspace of Q. Since / is a surjection, for every y G Y the fiber f ~ l ( y ) is a nonempty compact subset of X. Hence it has a smallest element with respect to the lexicographic ordering ^ on Q (Lemma 3.9.1), say p(y). We claim that the set S = {p(y}
:yeY}
is a GVsubset of X and is therefore as required. Define for n, m G N the following sets: Un^m — {x G X : for all z G X such that Xi — Zi for i < n — 1 and zn < xn — 1/m we have f(x] ^ f ( z ) } . We first prove that the t/n,m's are open by establishing that their complements are closed. To this end, for each j G N take x(j) G X \ Un>m and assume that the sequence (x(j)) . converges to a point x G X. For each j there exists an element z(j) in X such that x(j)i = z(j)i for every i < n — I while moreover z ( j ) n < x(j}n - Vm and f(x(j)) = f ( z ( j ) ) . By compactness, without loss of generality we assume that (z(j)) . converges to a point z in X. By continuity of the projections and / we get Xi = Zi for every i
218
3. BASIC DIMENSION THEORY
We finally claim that the intersection U of all the t/n,m's is equal to S. Since S intersects every fiber of / in precisely one point, and as we just observed, S C £/, it suffices to prove that U intersects every fiber of / in at most one point. To this end, take an arbitrary y 6 Y and assume that there are distinct points x ( l ) , x ( 2 ) 6 U C \ f ~ 1 ( y ) . Let n 6 N be the first index such that x(l)n ^ x(2)n. We may assume without loss of generality that
x(l}n <x(2)n. There exists i G N such that x(l)n < x(2)n — l/i. Since x(2) 6 Un>i and x(l)m ~ x(2)m for all m < n — 1, we get f ( x ( l } ] ^ f ( x ( 2 ) } . This is a contradiction. D The existence of n-dimensional totally disconnected spaces. We are now in a position to prove the main result of this section. Theorem 3.9.3. For each n > 0 there exists an n-dimensional topologically complete totally disconnected space. There also exists a strongly infinitedimensional topologically complete totally disconnected space. Remark 3.9.4. The existence of a strongly infinite-dimensional topologically complete totally disconnected space turns out to have a surprising application: it is the main ingredient in the solution of a famous problem in dimension theory that remained open for more than 50 years. See Theorem 3.13.8 for details. Proof. We shall construct the examples simultaneously. Consider Y — — JJH X v JJlTn , -A.
where n 6 NU {oo}. Let TT: X —>• JJ be the projection onto its first factor and put A = TT~ I (— 1) and B = 7r~ 1 (l), respectively. Define C = {C € 2X : C is a continuum from A to B}. By Exercise 1.11.21 it follows that C is a closed subspace of 2X . Since 2X is compact by Proposition 1.11.4, this proves that C is compact. Let C be a Cantor set in J and since C / 0 there is a continuous surjection a: C -»> C (Theorem 1.5.10). Now put
Claim 1. Y is compact and 7r[Y] = C. Proof. We shall first prove that ir[Y] = C. Clearly, 7r[F] C C. Take an arbitrary t G C. Then a(t) is a continuum from A to B. Consequently, 7r[o;(i)] is a continuum in J from —1 to 1 and therefore is equal to J. So there exists an element y £ a(t) with TT(T/) = t. Since this y belongs to Y, this proves that C C 7r[y].
3.9. TOTALLY DISCONNECTED SPACES
219
We shall next show that Y is compact. To this end, let (yi)i be a sequence in Y converging to a point y G X. Then t; = 7t(yi) belongs to C for every i and (ti)i converges to t = TT(?/). Consequently, t belongs to C. Since a is continuous, the sequence of continua (a(t;)). converges to a(t) in C. We claim that y e a(t). If this is not the case then we can find e > 0 such that y£D(a(t},e}. Since (cu(£;)). converges to a(t), for all but finitely many i, yi£a(ti) which implies that y G D(a(t),e). that yeY.
CD(a(t),e) This is a contradiction. We conclude 0
Since TT is continuous and Y is compact (Claim 1), by Theorem 3.9.2 there exists a G^-subset 5 of Y which intersects each fiber of TT \Y in precisely one point. The compactness of Y implies that S is topologically complete (Theorem A.6.3). Claim 2. S intersects each continuum from A to B. Proof. Take an arbitrary C € C. Since a is surjective, there exists t G C such that a(t) = C. Observe that (TT r IT1^) = 7r ~ 1 W n r = ^H*) n «(*) = T"1^) n (7 C C. Since 5 intersects each fiber of TT \ Y we consequently get S fi C 7^ 0.
<(>
Observe that TT f S is a one-to-one continuous function from 5 onto the zero-dimensional space C. This implies that S is totally disconnected, cf. Example 1.5.18. Now assume that n is finite. Then dim S > n by Theorem 2.4.12 and Corollary 3.7.5. Also dim 5 < n since 5 is contained in C x JJ n , which is n-dimensional by Theorems 3.2.12 and 3.4.12. We conclude that dim S = n. Now if n — oo then S is strongly infinite-dimensional by Corollaries 2.4.13 and 3.7.5. D Application: The dimension of infinite products. As an unexpected application of the results obtained here we will show that there exists for every n > 0 a topologically complete n-dimensional space whose infinite power is also n-dimensional. This shows that there exists an example such as Example 3.4.13 in every dimension. We will in fact show that the spaces constructed in the proof of Theorem 3.9.3 are as required. We have to be a little careful though. In the proof of Theorem 3.9.3 the precise nature of the 'G^-selection' 5 is irrelevant. We used Theorem 3.9.2 to prove its existence, but how it was created in the proof of Theorem 3.9.2 did play no role. But this is important here. So we assume
220
3. BASIC DIMENSION THEORY
in the proof of Theorem 3.9.5 that the 'G^-selection' S was constructed via the method used in the proof of Theorem 3.9.2, i.e., via the lexicographic order on JIn. Theorem 3.9.5. For every n > 0 there exists an n-dimensional topologically complete space whose infinite power is also n-dimensional. Proof. Let S be the space constructed in the proof of Theorem 3.9.3. We claim that S is as required. Observe that S C JJ x JP and that the projection
TT: J J x J n -> JJ maps 5 in a one-to-one way onto the Cantor set C. Define a function
/: S x 5 -> C 2 x Rn as follows: f((x,z),(x',z')) 2
= ((x,x'),z + z').
n
Since dim(C x R ) < n (Theorems 3.2.12 and 3.4.12) it will follow that dim S2 < n if we can prove that / is an imbedding. It is clear that / is oneto-one and continuous. The potential problem is the continuity of the inverse. To check this, let ( f ( ( x i , Z i ) , (X,^)))^ be a sequence in /[52] converging to an element /((x, z), (x', z')) in f[S2]. By the definition of / this implies that Xi —> X,
x'i —> x',
Zi + z'i —¥ Z + z'.
By compactness we may assume without loss of generality that Zi —> a,
z'i —> &,
n
where a,6 e JJ . So a + b = z + z'. Since (xi,Zi) £ a(xj) for every i and Gi(xi) —>• a(x) by continuity of a it follows easily that (re, a) 6 a(x] (Lemma 1.11.2(b)). But this implies by construction that z -< a. It follows similarly that z' -< b. Since z + z' = a + b we get z = a and z' = b. This implies that ((zi,Zi),(z;,z;)) -> ( ( x , z ) , ( x ' , z ' ) ) , as required. So dim S2 < n from which it follows that dim S2 = n since S2 contains a (closed) copy of the n-dimensional space 5 (Theorem 3.2.9). The same method can be be used to show that the dimension of each finite power of S is equal to n. Since the countably infinite power of 5 is the inverse limit of the inverse sequence consisting of all of its finite powers (Example 1.10.3), we are done by an appeal to Exercise 3.5.2. D Remark 3.9.6. It was shown by POL [345] that if X is a compact space which is (n + l)-dimensional then it contains an n-dimensional G^-set 5 such that dim 5°° = n. The methods used in this section show that this is true for the special case X = J n+1 .
3.10. THE ORIGINS OF DIMENSION THEORY
221
Exercises for §3.9.
1. Let X be a compact space such that dimX > n + l. Prove that X contains a totally disconnected G^-subset S such that dim S > n. 2. Let X and Y be totally disconnected. Prove that X x Y is totally disconnected. 3. Prove that there exists a countable dimensional infinite-dimensional topologically complete totally disconnected space. 3.10. The origins of dimension theory The first definition of a dimension function is due to BROUWER [77]. He seems to have been put on the right track by PoiNCARE [337, 338] who had proposed to call a continuum n-dimensional, if it can be dissected in separate pieces by one or more (n — l)-dimensional continua. Brouwer refined this idea and defined a dimensional invariant intended for topologically complete spaces without isolated points (in the terminology of his days, normal sets in the sense of Frechet) which he called Dimensions grad. There is no reason for the restriction to topologically complete spaces without isolated points, so we will state his definition for arbitrary spaces instead. Indeed, define the Dimensionsgrad Dg X of a space X, as follows: 0 < Dg.Y < n
<£>
Dg X = n Dg X = oo
<£> <&•
X is nonempty and for every pair A, B of disjoint closed subsets of X there exists a closed set T C X which is a cut between A and B such that Dg T is at most n — 1, DgX < n and Dg^Y is not smaller than n, Dg X ^ n for every n > — 1 .
So this invariant is very similar to the large inductive dimension, which was defined much later. It is easy to see that if X and Y are homeomorphic spaces then Dg X is equal to DgY. Hence Dg is a topological invariant, as are all dimension functions considered in this chapter. It is clear that DgX = 0 if and only if X ^ 0 and X contains no continuum of cardinality larger than 1 . Such spaces are called punctiform. There is something very peculiar about the definition of Dimensionsgrad, namely, a cut between two closed subsets of an arbitrary space is not necessarily a partition, even in compact spaces. In the sin (llx} -continuum in the plane {(x,sin(V,)) :0<x< i/J U {(0,y) : -1 < y < 1} the point (0,1) cuts between the points (Y^O) and (0, — 1). But the complement of (0, 1) is connected! (See also Exercise A. 10. 7.) This was pointed
222
3. BASIC DIMENSION THEORY
out to Brouwer by Urysohn. As a result of this, Brouwer issued a flood of notes and papers correcting his 'mistake'. For a description of the troubles this caused in the early days of dimension theory, see e.g., JOHNSON [213]. Lemma 3.10.1. For every space X we have Dg X < dim X. Proof. This is easy. If dim X — — 1 then X = 0 and so Dg X = — 1. Suppose that the inequality is true for all spaces X with dimX < n — 1, where n > 0. By the Coincidence Theorem 3.4.10 it follows that for every pair of disjoint closed sets A, B C X there exists a partition T between A and B such that dimT < n — 1. By our inductive assumption, DgT < dimT < n ~ 1. Since a partition is a cut, it follows that DgJ\T < n, as required. D The proof of Lemma 3.10.1 and Urysohn's Example suggest the question whether every cut is a partition in some well-behaved class of spaces. Lemma 3.10.2. Let X be a space which is both tope-logically complete and locally connected. If A and B are closed subsets of X and T cuts between A and B then T is a partition between A and B. Proof. Let U be the family of all components of X \ T. Since X is locally connected, every member of U is open (Exercise A.2.8). Assume that some U G U intersects both A and B. Then U is a connected, locally connected and topologically complete space (Theorem A.6.3). But then by the Mazurkiewicz Theorem 1.5.22 it follows that there is an imbedding $: I —> U such that $[1] meets both A and B. But this contradicts the fact that T cuts between A and B. Now let U be the union of all elements of U that meet A and let V be the union of all the other elements of U. Then U is an open neighborhood of A, V is an open neighborhood of B and U U V = X \T. Hence T is a partition between A and B. D So from Lemma 3.10.2 one would hope to be able to prove that Dg and dim take the same values on locally connected, topologically complete spaces. This was indeed widely believed to be true for a long time. In the book by HUREWICZ and WALLMAN [208], one can read on Page 4 that Dimensionsgrad and dimension agree even in the class of all locally connected spaces. A proof of this assertion was not presented. Similar statements for various classes of locally connected spaces were subsequently made in various other books on dimension theory. None of these claims was supported by a proof. That there are troubles, even for compact locally connected spaces, is easily demonstrated. Let X be a compact space of Dimensionsgrad 0. Then it contains no nontrivial continuum, and hence is zero-dimensional (Corollary 3.7.9). This triviality will be used in the proof of Theorem 3.10.3 below. If moreover X is locally connected and DgX < 1 then by Lemma 3.10.2 every two disjoint closed sets can be partitioned by a zero-dimensional set,
3.10. THE ORIGINS OF DIMENSION THEORY
223
i.e., the large inductive dimension of X is at most 1. But if X is compact and locally connected and has Dimensionsgrad at most 2 then all there can be concluded from the above is that every pair of disjoint closed subsets can be partitioned by a closed subset with Dimensionsgrad at most 1. But that closed set does not need to be locally connected, so we are at a dead end. Brouwer always claimed that his Dimensionsgrad could serve as the basis of an equally important theory of dimension as ordinary dimension. But one would guess from the above that for a sufficiently 'bad' compact space dim and Dg can differ. The following surprising result shows that this is not true and that Brouwer, as usual, was right after all with his claim at least for compact spaces. Theorem 3.10.3. Let X be a compact space. Then dimX = DgX. Proof. By Lemma 3.10.1 we only need to prove that dim^C < DgX for all compact spaces X. If DgX = 0 then by the above remarks we have that dim X = 0. So now suppose that for every compact space X such that 0
we have dim X < Dg X.
Assume that X is a compact space such that Dg X = n. Let 13 be a countable base for X. For every pair (So, BI) of elements of 23 such that BQ C B\ there is a cut between BQ and X \ BI with Dimensionsgrad at most n — I . So by our inductive assumption, every such cut has dimension at most n — 1. By the Countable Closed Sum Theorem 3.2.8, the union of these cuts is at most (n — l)-dimensional. By Corollary 3.3.12 this union is contained in an at most (n — 1) -dimensional G$ -subset 5 of X. Then T = X \ S is an Fa -subset of X. We claim that T contains no nontrivial continuum. Striving for a contradiction, assume that C is a nontrivial continuum contained in T, and pick arbitrary distinct points a, b £ C. There exist elements BQ,BI G "B such that a e 5o C £0 C #! \ {b}. By construction, 5 contains a cut between BQ and X\Bi. But C is a continuum meeting both B0 and X \ BI but missing 5. This is a contradiction. So we conclude that T contains no nontrivial continua. But then every compact subset of T is zero-dimensional by the same reasons as before. So T is a countable union of compact zero-dimensional subspaces, and hence is zerodimensional itself (again by the Countable Closed Sum Theorem 3.2.8). So by Theorem 3.4.4 we conclude that dim X = dim(S U T) < dim S + dim T + l = ( n - l ) + 0 + l = n, as required.
D
224
3. BASIC DIMENSION THEORY
It is clear that there are spaces X such that DgX ^ dimX. For let A" be a totally disconnected n-dimensional space (Theorem 3.9.3). Then X is punctiform and so DgX = 0. The gap between Dimensionsgrad and dimension can therefore be arbitrarily large. In view of the claim in the book by HUREWICZ and WALLMAN [208], the question naturally arises whether there are similar examples within the class of locally connected spaces. Observe that Lemma 3.10.2 draws specific attention to the class of all topologically complete locally connected spaces, for there the notions of a cut and a partition coincide. C
00,01,10,11
0,1
Figure 10. Theorem 3.10.4. For each n — 2 , 3 , . . . , oo there exists a locally connected topologically complete space Xn such that DgXn = 1 < dimXn = n. Proof. Represent the set S, the set consisting of all finite sequences consisting of O's and 1's (including the empty sequence), by dots in the plane as in Figure 10. These dots are compactified in a natural way by a copy C of the Cantor set. We now connect every dot with its two 'successor' dots by an interval (see Figure 11). In this way we obtain a connected space which we call T. By U we denote one of the basic neighborhoods of a point of C. The set U is clearly connected, so T is locally connected. But it has a stronger property which we will now describe. Claim 1. For an arbitrary subset CQ C C the space TQ U CQ is locally connected. Proof. In fact, a basic neighborhood of an arbitrary point c G CQ C TO U CQ has the form
(C/nTo)u([/nC 0 ). This set is connected, since it contains the dense connected set U n TQ.
<0>
3.10. THE ORIGINS OF DIMENSION THEORY
225
Now let / : C —> Q be a continuous surjection from C onto the Hilbert cube Q (Theorem 1.5.10). Let Z = T(Jf Q and let g: T ->• Z be the quotient map. So Z is the disjoint union of TO and the Hilbert cube Q. For every n let Yn C Q be an n-dimensional totally disconnected Gs-set (Theorem 3.9.3) and let Xn = T0UYn. Claim 2. Xn is a topologically complete space. Proof. This is clear since the complement of Xn in Q is an TV -subset of Q by Theorem A. 6. 3. So the complement of Xn in the compact space Z is an TV -subset of Z as well.
C
Figure 11.
Claim 4. dimXn = n. Proof. It is clear that Yn is a closed n-dimensional subset of Xn. So Xn is at least n-dimensional by the Subspace Theorem 3.2.9. On the other hand, Xn \ Yn is a countable union of one-dimensional compacta. Hence dimXn < n by the Countable Closed Sum Theorem 3.2.8. 0
Claim 5. DgXn = 1. Proof. Let A and B be disjoint closed subsets of Xn. Take an arbitrary partition So in Yn between A n Yn and B n Yn. Corollary 3.3.10 gives us a partition S in Xn between A and B such that SC\Yn C S0 and dim(S\S 0 ) < 0.
226
3. BASIC DIMENSION THEORY
We claim that S is totally disconnected. So we must show that 0 is a partition between any pair of distinct points a and 6 of 5. Since S \ Yn is an open zero-dimensional subspace of S, a moments reflection shows that we may assume without loss of generality that both a and b belong to the totally disconnected closed space Yn of Xn. So 0 is a partition between a and b in 5 n Yn. By another appeal to Corollary 3.3.10 it follows that 0 can be extended to a partition between a and b in S such that the dimension of its intersection with S \ Yn is — 1. We conclude that 0 is a partition in S between a and b. Since a totally disconnected space is punctiform, it has Dimensionsgrad 0. We conclude that DgXn < I . That is clear since Xn contains copies of the unit interval.
0
This concludes the proof of the theorem.
D
Let us remark that there also is an infinite-dimensional locally connected topologically complete space with Dimensionsgrad 1. It suffices to take the topological sum of the spaces Xn in Theorem 3.10.4. This space is countable dimensional. There is also a strongly infinite-dimensional example since there is a strongly infinite-dimensional totally disconnected G<$-subset in Q (Theorem 3.9.3).
Exercises for §3.10.
1. Let X be a cr-compact space. Prove that Dg X = dimA^. ^•2. Let A be a space which is both topologically complete and locally connected. Suppose that A and B are disjoint closed subsets of A. Prove that every partition S between A and B contains an irreducible partition T between A and B. ^•3. Let A be a unicoherent Peano continuum containing two disjoint continua A and B. Prove that if S is an irreducible partition between A and B then S is connected. 4. Show that the fact that A and B are continua in the previous exercise is essential. ^•5. Prove that Sn for n > 2 cannot be imbedded in a product of the form ^\ 1 X
' * *
X
Y\ -ft ,
such that every Xi is at most one-dimensional. 6. Prove that the local connectivity assumption is essential in Exercise 3.10.2, even for compact spaces.
3.11. THE DIMENSIONAL KERNEL OF A SPACE
227
3.11. The dimensional kernel of a space For a space X and a point x 6 X, indx X denotes the dimension of X at the point x, and is defined as follows: mdxX = -l 0 < mdx X < n
<£> <£>
mdx X = n indj; X = oo
<£> <£>
X = 0, for every closed subset A of X not containing x there exists a partition L between {x} and A such that dimL < n — 1, indx X < n and ind z X ^ n — 1 , indx X ^ n for every n > — 1 .
Observe that dim X = ind X < n if and only if indx X < n for every x 6 X. If X is n-dimensional then its dimensional kernel A.(X) is the set {x £ X : indxX = n}. Lemma 3.11.1. Let X be an n-dimensional space, n>l. (1) The dimensional kernel of X is an Fff -subset of X. (2) The dimensional kernel of X has dimension > n — 1. Proof. Let E denote the complement of the dimensional kernel of X. For every i e N and for every x 6 E there is an open neighborhood Ulx of x such that indFr(C7*) < n - 2, diam(L^) < i/iThe cover {[/.£ : x G E} of E has a countable subcover, say Uz (Corollary A. 2. 3); put Ui = \JUi. Then Ui is open, and oo
E C U = p| t/i. i=l
We claim that L7" C E as well. Pick arbitrary x € C/ and e > 0. Let z be so large that l/i < e. Let V G Ui be such that x G F. Then diam(V ) = diam V < l/i < e which implies that V C B(x,e). Since indFr(y) < n — 2 and e > 0 was arbitrarily chosen, this implies that ind^ X < n — 1, i.e., x & E. We conclude that U C E and hence that E = U. As a consequence, £" is a G^-subset of JC, which proves (1).
For (2), let CO
F=U U Then F is the union of countably many closed subsets of X with dimension at most n — 2. By the above, the collection U = U^i ^i '1S a local base at every point of E (cf. Exercise A.2.15). So it follows that the collection U\(E\F]
228
3. BASIC DIMENSION THEORY
consists of open and closed subsets of E \ F and is in E \ F a local base at any point. From this we see that ind(£' \ F) < 0. Now let K = X \E be the dimensional kernel of A", and suppose that ind K < n — 2. We will derive a contradiction. By (1) and the above, K U F is an F0.-subset of X and by Theorems 3.4.10 and 3.2.8 it follows that md(K\JF) < n - 2 . But then by Theorem 3.4.4, ind X = ind ((K U F) U (E \ F ) ) < n - 2 + 0 + 1 = n - 1,
which is a contradiction.
D
For compact spaces, we can do a little better. Theorem 3.11.2. Let X be an n-dimensional compact space. Then the dimensional kernel of M is n-dimensional and a-compact. Proof. The dimensional kernel is cr-compact by Lemma 3.11.1(1). To prove that it is n-dimensional, first observe that if n = 0 then there is nothing to prove. So let n > 1. By Theorem 3.7.8, X contains an n-dimensional Cantor-manifold Y. We claim that Y is contained in the dimensional kernel of X. If so then we are done. To this end, let y G Y be arbitrary. Since dimF > 1 there exists a point z G Y \ {y}. If indy X < n then there is a partition F between {y} and {z} such that dimF < n — 2. But then F D Y separates Y and is at most (n — 2)-dimensional, which is a contradiction since Y is an n-dimensional Cantor-manifold. D Let X be zero-dimensional, and let x E X. It is clear that indx X — 0. So the dimensional kernel of X is X itself. This explains why in the following definition we require X to be at least one-dimensional. A space is called weakly n-dimensional, where n > 1, if it is n-dimensional, but its dimensional kernel is of dimension n — l. By Theorem 3.11.2, no compact subspace of a weakly n-dimensional space is n-dimensional. It is not clear at all whether weakly n-dimensional spaces exist. But in Theorem 3.11.8 we will construct for every n > 1 such a space. We have to so some preparatory work first. Lemma 3.11.3. Let X be weakly n-dimensional with subspace F. Then F is at most (n — I)-dimensional or weakly n-dimensional. Proof. Assume that dim F = n. We claim that if x belongs to the dimensional kernel of F then x belongs to the dimensional kernel of X. For assume that x £ F does not belong to the dimensional kernel of X. Then x has arbitrarily small neighborhoods U in X such that indFrt/ < n — 2. By Lemma 3.4.3 and Exercise A.1.16 it follows that the boundary of U n F in F
3.11. THE DIMENSIONAL KERNEL OF A SPACE
229
has dimension at most n — 2. So this shows that x does not belong to the dimensional kernel of F. So the dimensional kernel of F is a subspace of the dimensional kernel of X and hence is at most (n — l)-dimensional by the Subspace Theorem 3.2.9 (or by Lemma 3.4.3). D Lemma 3.11.4. Let n > 1. In addition, let X be compact such that there exists a continuous function f : X —t I such that for every q € Q D (0,1) we have dim f ~ l ( q ] < n. Then there is a G5-subset S C (0,1) with Qn(0,1) C S such that dim/" 1 (z) < n for every z € S. Proof. By the Countable Closed Sum Theorem 3.2.8 it follows that
U r'fo)
g€Q
is at most n-dimensional and hence can be enlarged to a G^-subset S of X with dim S < n (Corollary 3.3.12). Then T = f[X \ S] is by compactness of X an FCT-subset of I which misses Q. It is clear that for every z in the GS~ subset I \ T of I we have that f ~ l ( z } C 5. Hence f ~ l ( z ] is at most ndimensional. D Lemma 3.11.5. Let K be an (n + I)-dimensional compact space with disjoint closed subsets A and B. There exist a continuous function f : K —> I and a Cantor set A in [i/3, 2/3] such that f[A] C {0}, f[B] C {1} and dim ( f ~ l ( t ) ) < n for each t e A. Proof. Let a: K —>• I be a Urysohn function such that
a[A] C {0},
a[B] C {1}
(Lemma A.4.1). Identify I and the set I x {0} x ••• x {0} ^m^Mmv^B^H*. jtm^—^^^^^mi/
2n+2 2n
3
in R + . By Remark 3.3.6 there is an imbedding g : K -> Wn+i such that g ( a , g ) < Ys- It follows that we may think of K as subspace of yin+i having the additional property that Ki[A] C (—00, Ys) and ^2 [-6] C ( 2 / 3 , oo). Now consider an arbitrary element q G Q n [V3, 2/3]. The set 5 = {x € M 2n+3 : xl = q} can be identified with M 2n+2 , and the set K n S with a subspace of {x G M 2n+2 : at most n coordinates of x are rational}. But this implies that dim(K D 5) < n by Corollary 3.2.11. So the function £ = 7T! \K: K ->R
has the property that for every q £ Qn [Y3, 2/3] we have dim^" 1 ^) < n. By Lemma 3.11.4 it follows that we may enlarge Qn[Y3, 2/3] to a G^-subset S having the same property. Observe that by the Baire Category Theorem A.6.6
230
3. BASIC DIMENSION THEORY
and the fact that Q H [Y3, 2/3] is dense in [1/3, 2/3], it follows that S is uncountable. So 5 contains a Cantor set A by Theorem 1.5.12, and all we need to do is to adjust £ slightly so as to satisfy the additional property that it sends A to 0 and B to 1. This adjustment is left as Exercise 3.11.1 to the reader. D Lemma 3.11.6. If X is an n- dimensional space, where n > 0, then there exists a zero- dimensional F^-subset N C X such that dim(A" \ N) < n — 1. Proof. If dim X = 0 then N = X is as required. Suppose therefore that the lemma is true for all spaces Z with dim Z < n — 1, where n > I. Assume that dim A = n. Let $ be a countable base for X such that dimFr(-B) < n — 1 for every B € £ (Lemma 3.4.1). Put
Y = |J Then Y is an Fa-subset of X since every Fr(B) is closed in A, and so Y is at most (n — l)-dimensional by the Countable Closed Sum Theorem 3.2.8. In addition, dim(X \ Y} < 0 since T> \ X \ Y is a clopen base for A \ Y . By our inductive assumption there exists a zero-dimensional F^-subset N C Y such that dim(y \ N) < n — 2. Observe that N is an F^-subset of A as well. Since dim(A \ F) < 0, it follows by Theorem 3.4.4 that dim(A \ N) < dim(Y \N} + dim(A \ Y) + 1 < (n - 2) + 0 + 1
= n — 1, as required.
D
Lemma 3.11.7. Let X and Y be compact spaces, and let f : X —>• Y be a continuous surjection. IfY is zero-dimensional at y £ Y and if f ~ l ( y ) is zero- dimensional then X is zero- dimensional at every point of f ~ l ( y ) . Proof. Pick an arbitrary x £ f ~ l ( y ) , and let U be a neighborhood of x in X. Since dimf^1(y) = 0 there is a clopen subset C of f ~ l ( y ] such that x G C C U . Pick disjoint open subsets E and F in A such that
It is clear that without loss of generality we may assume that ECU. Observe that E U F is a neighborhood of f ~ l ( y ) in A. Consequently, since / is a closed map, there is a neighborhood V of y in Y such that f ~ l [ V ] C E U F (Exercise A. 1.1 5). Since Y is zero-dimensional at y, we may assume without loss of generality that V is clopen. Then f~l[V]r\E is a clopen neighborhood of x in A which is contained in U. D We now come to the main result in this section.
3.11. THE DIMENSIONAL KERNEL OF A SPACE
231
Theorem 3.11.8. Let n > 1 and Jet K be an (n + l)-dimensionaJ compactum. Then K contains a weakly n-dimensional G§ -subset X (hence X is topologically complete). Proof. Let r = {(A0, Z?0), . . . , (A n ,5 n )} be an essential family in K. By Lemma 3.11.5 there exist a continuous function /: K —>• I and a Cantor set A C (0, 1) such that f[A0] = {0}, f[B0] = {1} and dim ( f ~ l ( t ) ) < n for every t G A. We now closely follow the construction in Theorem 3.9.3. Put C = {C 6 2K : C is a continuum from A0 to B0}. Then C is closed in 2K and hence is a compact space (see the proof of Claim 1 of Theorem 3.9.3). There consequently is a continuous surjection 0: A —>• C (Theorem 1.5.10). Put
Then Z is compact, and f[Z] = A (see the proof of Claim 2 of Theorem 3.9.3). Now dim Z > n because Z intersects every continuum from AQ to BQ (Corollary 3.7.5). In addition, the fibers of g = f \ Z are at most n-dimensional and A is zero- dimensional. We conclude that dimZ < n by Theorem 3.6.10. Now let A denote the dimensional kernel of Z. Then because Z is compact, A is an n-dimensional (T-compact subset of Z (Theorem 3.11.2). There consequently exists a a-compact zero-dimensional subset N of A such that dim(A\AT) < n- 1 (Lemma 3.11.6). Now put X = Z\N. Then X is clearly a G<$-subset of K and we first claim that it is n-dimensional. Let C e 6. We will prove that C meets X. Pick t € A such that >(£) = C. Observe that f ~ l ( t ) nC C Z. If dimf~l(t} n C > 0 then f ~ l ( t } n c intersects X because the complement of X in Z is zero-dimensional. We may therefore assume that dimf~l(t) n C = 0. But since A is zero-dimensional, we get that Z is zero-dimensional at all points of f ~ l ( t ] nC (Lemma 3.11.7). Consequently, /-1(*) n C n A = 0, i.e., 0 ^ f ~ l ( t ) n C C X. So we conclude from Corollary 3.7.5, that dimX > n. However, because X is a subspace of Z, we also have dimX < n. Consequently, dimX — n, as required. We finally claim that X is weakly n-dimensional. This is however a triviality (cf. the proof of Lemma 3.11.3). Simply observe that if x is a point of X at which X is n-dimensional, then Z is n-dimensional at x, which implies that x E X n A. Since by construction, dim(X n A) < n — 1, we are done. D Remark 3.11.9. If one takes K = In+l in the above theorem then there is no need to use Lemma 3.11.5 in the proof because we can use for / the projection onto the first coordinate. It seems that this gives the easiest known examples of weakly n-dimensional spaces.
232
3. BASIC DIMENSION THEORY
A universal weakly n-dimensional space. In §3.3 we showed that for every n there is a universal space for n-dimensional spaces. That is, there is an n-dimensional space X which contains a homeomorphic copy of any n-dimensional space. We will prove now that a similar result holds for weakly n-dimensional spaces. Theorem 3.11.10. For each n > 1 there exists a weakly n-dimensional space E such that any weakly n-dimensional space imbeds in E. Proof. Let p: (0, 1}°° x Q ->• {0, 1}°° be the projection. It will be convenient in this proof to denote the hyperspace of Q, i.e., the space 2^, by 3C. For a fixed natural number n, let JVC be the subspace of the product X x OC°° x (DC x K)00, consisting of all elements of the form (1)
(K^Lfr^dtDfr)
satisfying the following conditions: (2)
LiCL2C---CK,
dimLi
(i e N),
(3)
K = Cl U Di:
dim(Cl n Dx] < n - 2
(z e N).
There are a subspace T of {0, 1}°° and a continuous mapping (p from T onto the space M (use Exercise 1.5.10 and Corollary 1.5.7). It will be convenient to denote (p as follows:
(4)
t ^ ( K ( t ) , (1^)^(0^}, n^t}):).
For any t € T, let
(5)
H(t) = {z € K(t) \ U~! Li(t) : for any neighborhood U of z in Q there is an 2 e N with 2 e C;£ C £7 and 2 £
For * e r, put
(6) i—l
Finally, let E C {0, 1}°° x Q be the subspace of all points of the form (t, z), where t e T and z £ E(t). We shall show that (7)
E is weakly n-dimensional,
and
(8)
any weakly n-dimensional space imbeds in E.
3.11. THE DIMENSIONAL KERNEL OF A SPACE
233
To this end let, us consider the sets K*={(t,z):t£T,ze #(*)},
LI = {(t,z}:t(ET,ze
Claim 1. For every i 6 N we have (9)
dim LI < n - 1, EC C* U £>*,
dim(Cf n £>*) < n - 2.
Proof. It follows from Exercise 1.11.26 that the projection TT = p \L* is closed. Observe that the fibers of TT are of the form
Hence by (2) the fibers of TT are at most (n — l)-dimensional. Since dim T = 0, the desired result follows from Theorem 3.6.10. By (2) and (3) the other statements follow by similar considerations. •(> By (9) and Theorem 3.2.8, dim(|J2i Ll )
(10)
E\\jL*CE(n_1).
This implies among other things that the dimension of the set E \ -E( n _i) is at most n — I . So if dim.E > n then we contradict Lemma 3.11.1. This implies that dimE < n and also by the same argument that E is weakly n-dimensional provided that E is n-dimensional. The latter fact will follow once we established (8). Simply observe that there are weakly n-dimensional spaces (Theorem 3.11.8) which imbed into E by (8). Hence E is at least n-dimensional since it contains n-dimensional subsets (Theorem 3.2.9). Proof. Let us consider any point a ~ (i, z] from the set on the left hand side of the inclusion (10), i.e., z G H(t) by (6). Let W be a neighborhood of a in {0,1}°° x Q, and let
U = {xeQ: ( t , x ) £W}. Then U is a neighborhood of z\ let i 6 N be an index given by (5). The projection p[C* \ W] is closed in T (cf. the proof of Claim 1). It also misses t since Ci(t) C U. Let V be a clopen neighborhood of t in T, disjoint from p[C* \ W]. Then C? np~l[V]CW and therefore
is an (n — 2)-dimensional partition between the point a and the set E \ W , cf. (9). 0
234
3. BASIC DIMENSION THEORY
Let us consider now an arbitrary weakly n-dimensional space X. We shall find an element t E T such that X embeds in E(t). The dimensional kernel of X is an Jf^-set in X (Lemma 3.11.1). Hence there exist closed sets Zi C Z2 C • • •
in X such that dimZi < n — 1 for every i and X \ U^i ^ f= -^"(n-i)Corollary 3.5.6 and the fact that every compact space imbeds in Q (Corollary A.4.4), provide an imbedding h: X —>• Q such that dim/?,[Zj] < n — 1, for i = 1, 2 , . . . , and h(x) G (h[X])(n_ij for any x G X \ (J^i ^- Let
K = h[X],
(11)
Ll =
Then oo
(12)
h[X]\\jLlCK(n_1}. i=l
By (12) there exist pairs (Ci,Di), i G N, of compact sets in K, with K = d\JDi,
dim(ClHDi)
such that for any z G /&LX"] \ Ui^i ^ an<^ any neighborhood C/ of 2 in Q, there is an i € N with 2 G C; C C7 and z $ Di. The sets K and I/; in (11), together with the pairs (Ci,Di) determine a point in the set M, cf. (1), (2) and (3). So there exists t G T such that (13)
K
Let us consider the section 2£(£) of the set .E, cf. (6). According to formula (5),
and hence by (13), h[X] C E(t). This completes the proof of (8) and ends the proof of the theorem.
D
Dimension of products of weakly one-dimensional spaces. In his paper [284], MENGER asked whether there exists a weakly one-dimensional space X such that dimXn — n for every n. We will prove here among other things that the product of an arbitrary family weakly one-dimensional spaces is at most one-dimensional, thereby giving a strong negative answer to MENGER'S question. Theorem 3.11.11. If X is weakly one-dimensional then X has a compactification 'jX in which it is L-imbedded. Proof. We put K = A.(X). In addition, let C be a countable collection clopen subsets of X having the property that it is a local base at every point ofX\K (Exercise A.2. 12).
3.11. THE DIMENSIONAL KERNEL OF A SPACE
235
There is a compactification ^X of X such that for every C G C we have that C and X\C have disjoint closures in ^X (Exercise A. 9. 3). That is, the closure of C in ^X is clopen in jX. We claim that X is L-imbedded in ~fX. To this end, let U be a finite open cover of X. For every point x G X\ K, pick elements Cx G C and Ux G U such that (1)
X 6 C X CCX CUX.
Put A = X \ \JX£x\i
= U Cx u U WA-
E
Then E is an open neighborhood of X in *yX. Let Z C E be an arbitrary continuum. If Z fl Cx ^ 0 for some x E X \ K then Z C Cz C Ux by connectivity of Z and (1). So we may assume without loss of generality that Z n Cx — 0 for every x E X \ K. But then
zc\JwA so that by (3) and the connectivity of Z there is an A G A such that Z C WA • So this shows that Z C UAd Corollary 3.11.12. The product of a finite family of weakly one-dimensional spaces is one-dimensional. Proof. Such a product if L-imbedded in one of its compactifications by Exercise 3.5.6 and Theorem 3.11.11. So the desired result follows by an application of Corollary 3.5.9. D In the light of this result, the question arises whether the dimension of products of weakly n-dimensional spaces can be estimated more accurately
236
3. BASIC DIMENSION THEORY
than the bounds we get from Theorem 3.4.12. It is unclear whether this is possible. Exercises for §3.11. 1. Let X be a space and let A and B be disjoint closed subsets of X. In addition, let /: X —> I be continuous such that /[A] C [0, V 3 ), g [ B ] C ( % , ! } . Prove that there is a continuous function /: X —>• I having the following properties: (1) /(x) = / ( x ) i f x € / - 1 [ [ 1 / 3 , 2 / 3 ] ] ,
(2) f[A] C {0} and f[B] C {!}. 2. For every n € N let Xn be weakly one-dimensional. Prove that O^Li Xn is one-dimensional.
We now present a different proof of the fact that the product of two weakly one-dimensional spaces is one-dimensional. Let X and Y be weakly one-dimensional. Pick an arbitrary point (x, y) G (X x Y) \ (A(AT) x A.(Y)). We claim that X x Y is at most one-dimensional at ( x , y ) . We first show that this suffices. Striving for a contradiction, assume that X x Y is two-dimensional. Then by our claim, A(AT x Y) C A(X) x A(y) hence dim (A.(X) x A(F)) < 0. This contradicts the fact that A.(X x F) is at least one-dimensional (Lemma 3.11.1(2)). We assume without loss of generality that y £ A(F). If in addition x £ A.(X) then X x Y is zero-dimensional at (x,y). So we assume further that x G A(AT). Let U and V be arbitrary open subsets of X and Y, respectively, such that x & U and y G F. We will construct an open subset E1 C X x F such that (x,y)£ ECU xV 1
while moreover indFrE < 0. Since y 0 A(Y) we may assume without loss of generality that V is clopen. Let U' be an open neighborhood of x such that C^CC/,
FrU' C X\\(X).
It is possible to pick U' since ind (A" \ A(A)) < 0 (Corollary 3.1.6). If dim V < 0 then Fr(U' xV}= Fr(U") x V is at most zero-dimensional and so we are done. Assume therefore that dim V = 1. Then V is weakly one-dimensional (Lemma 3.11.3), and so we may assume without loss of generality that V = Y. Put A = FrU' and let U" be an open subset of X such that A C U" C U" C U. 3. Prove that for every n € N there exist pairwise disjoint clopen subsets Uin, Uzn,... of X such that (1) Uin n A ^ 0 for every i,
3.12. COLORINGS OF MAPS
237
(2) diam Uin < l/n f°r every z, (3) A C U^! Uin C (J* i tt« C tf", (4) Fr(U~ i tfi«) C A(X). Now for every n e N let £ n denote the collection of all clopen subsets of Y of diameter at most l/n and put
Observe that if b e Y is such that for every n there exists a clopen subset Cn of Y containing 6 and of diameter at most l/n then Y is zero-dimensional at b. So if 6 € A(Y) then there exists n e N such that b £ Bn. 4. Prove that for every n there is a decreasing sequence Vin , i G N, of clopen subsets of Y such that H^i ^« = ^«5. Let 00
w
00
= U LK^"*^")n=l i=l
Prove that A x A(Y) C W. 6. Prove that FrW C (A(X) x A(Y)) U (A x Y). Put
E = W\J(U' x Y). Then E is an open neighborhood of ( x , y ) and by Exercise 3.11.2(3) it follows that 'ECU xY.
We claim that Fr E is zero-dimensional. First observe that (*)
FrE C FrWU(FrU' x Y) = FrW(J (A x Y).
Put Bo = Fr £ n (A x Y) and BI = Fr£ \ S0, respectively. Since A x A(Y) C W (Exercise 3.12.4) and Fr E n W = 0, it follows that B0CAx ( Y \ A ( Y ) ) , which is zero-dimensional. We conclude that BQ is a zero-dimensional closed subspace of Fr E. In addition, Exercise 3.12.5 and (*) imply that Bi = Fr E \ Bo C Fr E \ (A x Y) C Fr W \ (A x Y) C A(X) x A(Y),
which is also zero-dimensional. So B\ is zero-dimensional as well. Since BQ is closed, the Countable Closed Sum Theorem 3.2.8 now easily gives us that dimE < 0, as desired. 3.12. Colorings of maps Let X be a space and let /: X —>• X be a fixed-point free continuous function. On Page 188 we defined / to be colorable if there is a finite closed cover A of X such that A n f[A] — 0 for every A € A. Since finite open covers can be shrunk to closed covers, and finite closed covers can be swelled to open covers, finite open covers do equally well (see §3.2).
238
3. BASIC DIMENSION THEORY
There are two natural questions we consider here. First of all the question whether every fixed-point free continuous function can be colored. And secondly, whether something can be said about the number of colors needed for a colorable function. Negative results.
Theorem 3.12.1. There is a space X with a fixed-point free involution which admits no coloring. Proof. If an: S n —>• Bn is the antipodal map, then every closed cover 3~ of S n such that F n ct[F] = 0 for every F G 3~ has at least n + 2 elements by Theorem 2.5.1. Let X be the topological sum of the S™ (n e N), and let h be the involution on X which is defined by the requirement that h f S n = an for every n. Now assume that there is a finite closed cover 3 such that F n h[F] = 0 for every F <E 3~. Pick m so large that |3~| - 2 < m. Then 9 = 5 \ §m is a closed cover of STO no element of which contains an antipodal pair. This implies that m + 2 < |S| < |3*|, which is a contradiction. D Observe that the space X in this theorem is infinite-dimensional. It is therefore natural to ask whether there is also a finite dimensional example. The answer to this question is quite interesting. For continuous functions there is such an example, but for homeomorphisms this is not true. Before we can answer it, we need to make a very interesting detour which will lead us via topological Ramsey theory to the desired result. We consider the Cantor set {0,1}°° with its usual Tychonoff product topology. As was remarked on Page 458, the function /: J'(N) -> {0,1}°° defined by f ( A ) = XA, the characteristic function of A, is a bijection. The product topology on {0,1}°° therefore induces a topology on T(N). With this topology, !P(N) is homeomorphic to {0,1}00 and hence is a Cantor set. The collection of all sets of the form ( N ) : (a C A] A (A n T = 0)}, where a and T are arbitrary finite subsets of N, is a clopen base for J"(N). For this simply observe that if a and r are disjoint finite subsets of N then the basic open subset {x G {0,1}°° :(ie0=> Xi = l ) and (i e T => xx = 0)} of {0,1}°° corresponds under this function to (cr, T). So the finite subsets of N form a countable dense subset of T(N). The complement, i.e., the collection of all infinite subsets of N, is a G^-subset which is also dense by the Baire Category Theorem A.6.6.
3.12. COLORINGS OF MAPS
239
If H C N is infinite then [H]u denotes the collection of all infinite subsets ofH. The usual topology on [N]w is the subspace topology that [N]w inherits from 5)(N) (with the just defined topology). If A, B C N then we write A < B provided that max(A) < min(B}. So this makes sense only if A is finite. For a < A, where a C N is (necessarily) finite and A C N is infinite, let (*)
[a,A] = {Se [Nf
:aCSCa(JA}.
u
Observe that [0, A] is simply [A] . The Ellentuck topology on [N]w has as basic open sets all sets of the form (*). It is easy to verify that they satisfy the axioms for a base for a topology. There is a problem though with the Ellentuck topology since it does not have a countable base. So it violates our assumption that all topological spaces are separable and metrizable. This causes no problem, since the Ellentuck topology is used merely as a tool here to prove interesting theorems on [N]w endowed with its usual topology, and the language of topology which we are used to is convenient for that. Lemma 3.12.2. If H C N is infinite then the subspace [H]w of [N|w with the Ellentuck topology is homeomorphic to [N\u. Proof. Let /: H —> N be any bijection. It is easy to see that the function F: y(H) ->> T(N) defined by F(A) = f[A]
is a homeomorphism.
D w
Lemma 3.12.3. The Ellentuck topology on [N] contains the usual topology on mu. Proof. It suffices to prove that if a and T are arbitrary finite subsets of N then
<<7,r)n[Nr is open in the Ellentuck topology. To this end, let A C N be infinite such that A e (cr, T}. Put B = A \
or
[0,A] C [N]w \X.
It is called completely Ramsey if for every o, A C N with a finite and A infinite and a < A there is an infinite set B C A with [a, B] C X
or
[a, B] C [N]w \ X.
240
3. BASIC DIMENSION THEORY
We introduce some additional notation. Let U C [N]w be open in the Ellentuck topology. Call [a, A] good if for some B C A we have [a, B] C U. If [a, A] is not good then it is called bad. We shall call [a, A] very bad if it is bad and for every n G A, [a U {n}, A/n] is bad, where A/n = {m 6 A : m > n}. Notice that if [a, A] is (very) bad and B C A is infinite then [a, B] is (very) bad. Lemma 3.12.4. If [a, A] is bad then there is an infinite subset B C A such that [a, B] is very bad. Proof. Indeed, suppose that this is not true. Then [a, A] is bad but not very bad and so there exists no 6 A such that [a U {no}, A/HQ] is good. There consequently exists an infinite B0 C A/HO such that [a U {no},#o] ^ ~U. Since [a, BQ] is not very bad, there exists ni G B such that [a U {HI}, BO/HI] is good. Observe that no < n\. So there exists an infinite B\ C BO/HI with [a U {ni},Bi] C £7, etc. Let B = {no,ni,...}. Pick an arbitrary element C e [a, 5]. Then a C (7 C a U B. Let n^ be the first integer in (C n B) \ a. Since [a U {r^}, £?j] C U and C G [a U {n^}, #;] we conclude that C 6 U. Since C was arbitrary, it therefore follows that [a,B] C [/, so [a, A] is good, which is a contradiction. D Proposition 3.12.5. Let U C [N]w be open in the Ellentuck topology. Then U is completely Ramsey. Proof. Suppose that [a, A] is given. We may assume without loss of generality that it is bad. Claim 1. There is a decreasing sequence A D B0 D BI D • • • ,
with Hi — min(.B;) strictly increasing, such that for any b C {no,... , n^_i} we have that [a U b, Bi] is very bad. Proof. By Lemma 3.12.4 there exists an infinite BQ C A such that [a,-Bo] is very bad. Let n0 = min(E). Since [a U {no},-B/{no}] is bad, again by Lemma 3.12.4 there exists an infinite BI C B/HQ such that [a U {no}, .Si] is very bad. Then [a, Si] is very bad since BI C B and [a U {no}, BI] is very bad by construction. Now assume that for i > 1
A ^B0 D BI D ••• D Bi, are as we want. So n;_i < ni = min(.B;) and if b C { n o , . . . , n;_i} is arbitrary then [a U 6, Bi] is very bad. So [a U b U {n;}, B/rii] is bad and we can use Lemma 3.12.4 repeatedly (finitely often) to get an infinite set Bi+\ C Bi such that for every b C {no,..., n^_i} we have that
[a\Jb(J{ni},Bi+i]
3.12. COLORINGS OF MAPS
is very bad. Then Bi+i is clearly as required.
241
<0
Let B = {n0, rai,... }. We claim that [a, B] C [N]w \ U. If not then [a, B] H £7 / 0
and so since C7 is open there is [a', B'] C [a,B] such that [a',B'] C U. By Exercise 3.12.1 for some i, a' — aUb with b C { n o , . . . , nx} and B'/rii C Bi/rii. So since [a U 6, B'/Hi] C £/, we have that [a U 6, -Bz/n] is good. This is a contradiction. D Corollary 3.12.6. Let [N]w = P0 U • • • U Pk, where each Px is open in the natural topology. Then there are an infinite H C N and an i < k such that every infinite subset of H is contained in Pi. Proof. It is clear that we may restrict ourselves to the case that k = I (Lemma 3.12.2). Since PQ is open in the Ellentuck topology by Lemma 3.12.3 it is completely Ramsey by Proposition 3.12.5. So we are done. D We are now ready to present our second example of a map which cannot be colored. Theorem 3.12.7. There are a zero-dimensional space X and a fixed-point free continuous function f : X -+ X such that (1) f is closed, (2) / is finite-to-one, (3) / cannot be colored. Proof. Put A" = [N]w endowed with its natural topology. Define f : X — > X
by f ( A ) = A\{mmA} (AeX). It is clear that / is fixed-point free. We claim that it is continuous. To this end, let (An)n be a sequence in X converging to an element A 6 X. Let p = mmA,
a = {p},
/3 = {1, 2 , . . . ,p - 1},
respectively. Then A 6 (a, j3] and since An ->• A we may therefore assume without loss of generality that min An = minyl for all n. Now assume that A \ {min A} 6 (a, T)
for certain finite cr, T C N. Then if a' = a U {min A} and r' = T \ {min A}, it follows that A € U, where U — {cr', r'}. There exists N € N such that An € U for all n > N. But since minAn = min .4 it follows that a C An \ {mmAn} and T n (An \ {min An}} = 0. We conclude that f(An) € (a, T} for all n > N. To prove that / is closed, let A C X be closed, and assume that B g f[A]. Let p = min B. Observe that
242
3. BASIC DIMENSION THEORY
Since A is closed, for every i < p — I there exist finite sets a^Ti C N such that (**)
Bu{i}e(vl,Tl)C[N\"\A. {p}U(UfJ11
Put a = &i\{i}) and r = Uf=i Tii respectively, and let V — (a, r}. Then B 6 V . We claim that V n f[A] = 0. Striving for a contradiction, assume that there exists A G A such that A \ {min A} G F. Then since
a C A\ {min A} it follows that p € A \ {min A} and so min A < p. Put i — min A. Observe that Ti n (A\ {i}) = 0 and that i $ n. So A e {CT^T;} which contradicts (**). We already implicitly proved that / is finite-to-one. To show that it cannot be colored, let U be an arbitrary finite open cover of X. By Corollary 3.12.6 there are an element U 6 U and an infinite subset H C N such that [H]u C U. But then f ( H ] 6 U so that f[U] H t/ / 0, i.e., U does not color /. D Positive results. We will now present the main results on the existence of colorings. If /: X —> X is a homeomorphism then A C X is called /-invariant Theorem 3.12.8. Let X be finite dimensional, and let f be a fixed-point free homeomorphism. Then f can be colored. Proof. We will first prove that if dimJY" = 0 then / can be colored with at most three pairwise disjoint colors. By the remarks on Page 188 there exists an open cover U of X such that f [ U ] r \ U = 0 for all U e It. By Proposition 1.5.1 we may assume without loss of generality that U is a clopen partition. So U is countable and we may list it as {Un : n e N}, where U\ — 0. By induction on n we shall now construct a clopen partition {Vj71, V^1, V3n} of Un such that
a) i—i
i=l
for j < 3. Put Vj- — 0 for j < 3, and assume that the V^'s have been constructed for i < n and j; < 3. For j < 3 put
Claim 1.
3.12. COLORINGS OF MAPS
243
Proof. First observe that H 3 =i Ej ~ 0 from which it follows that both the sets /[f>)3:=1 Ej] and f~1[T\3-i Ej] are empty since / is one-to-one. So if the claim is not true then we may assume without loss of generality that e.g., But this is impossible since E\ n E<2 = 0. So the claim implies that is an open cover of Un , which consequently can be refined by a clopen partition (Proposition 1.5.1). So there exists a clopen partition n
T/ +l T/ 0/n+l 1 ' 1 i 3 V
V
n
+1\ /
of t/n+i such that
(2)
v/ +1 n(/- 1 [s j .]u/[£; j ]) = 0
for j < 3. Claim 2. The V^+^s are as required. Proof. If (1) fails for n + 1 then since it holds for n for some j < 3 either
which is impossible by (2), or
/[T/7+1] n V/+1 / 0 which is impossible since f[Un+i] n Un+i = 0, or
which is also impossible by (2).
<0
This completes the inductive construction. So Vj = Ui^i ^7 ^or J ^ 3 is the desired coloring of / with three colors. Now assume that dim X = n + 1 and that the theorem has been proved for all spaces of dimension at most n, where n > 0. Let D be a countable dense subset of X and let !B be a countable base for X such that dim Fr B < n for all B <E !B. Put 5 = £> U Uses Fr5 > and T=
\Jfn[S], nGZ
respectively. It is clear that T is a dense /-invariant _FCT-subset of X which by the Countable Closed Sum Theorem 3.2.8 has dimension at most n. In
244
3. BASIC DIMENSION THEORY
addition, Z = X \ T is also /-invariant and is of dimension at most 0 since Fr B n Z = 0 for all B € 33 (so 23 \ Z is a countable base for Z consisting of relatively clopen sets). There exists by our inductive assumption a finite cover U of T consisting of relatively open sets such that f[U] n U — 0 for all U 6 U. For every U € U let U' C X be open such that U' n T = U. Then /[£/'] H [/' = 0 since / is a homeomorphism and T is dense in X. We conclude that U' = [Ur : U £ U} is a finite collection of open subsets of X which 'colors' T. Now consider the zero-dimensional /-invariant set Z. We would like to apply the same technique as the one we just used. But Z need not be dense in X. But we know that a strong form of the theorem is true for zero-dimensional spaces, so we will proceed from there. There exist relatively open pairwise disjoint sets Vi, V% and V3 in Z that 'color' Z. Observe that if e.g., x 6 V\ then f(x) 6 Vz U V3. By Lemma A. 8.1 there are pairwise disjoint open sets V/, V2' and V3' in X such that VI n Z = Vi for z < 3. Now put V/' = V/ r\f~l[V^\J V3']. Then V/' is an open subset of X and clearly f[V"] D V/' = 0. We conclude by proving that V\ C V/'. Indeed, take an arbitrary element x E Vi. Then /(x) G V"2 U V3 from which it follows that x e f ~ l [ V 2 U V3] C f~l\yi U V3']. We conclude that x e V/'. In other words, V/' 'colors' all the points of Vi. Define the sets V2" and V3" similarly. So U U {V0", Vi", V3"} is the desired coloring of X. D Remark 3.12.9. Observe that the proof of Theorem 3.12.8 shows that if X is n-dimensional and / is a fixed-point free homeomorphism of X then / can be colored with at most 3n + 3 colors. The number of colors. It is a natural to try to reduce the number of colors needed to color a given function. This is sometimes possible and we will come to that after some preparatory work. Our main tool is the following theorem, which is interesting in its own right. Theorem 3.12.10. Let X be a space with dimX < n and let U be a finite open cover of X with |Uj > n + 3. In addition, let f : X —> X be a, homeomorphism. Then there is an open shrinking V = {Vu : U G U} ofU such that for any 3 C U with |J| = n + 3 we have C\u^(vu u f~^\Yu\) = 0It is tempting to think of the shrinking V in this result as a cover for which Jr[Vu} :UeU}
3.12. COLORINGS OF MAPS
245
In addition, let V = {{0,4}, {1,4}, {2,4}, {3,4}}. Then the collection W = [VUf-^V]
:V e V}
has precisely two elements, namely, the sets {0,1,4} and {2,3,4}. So the order of W is 1. But since P) W ^ 0, the cover V does not have the property stated in Theorem 3.12.10. Before presenting the proof of Theorem 3.12.10, we will first show how we can use it to reduce the number of colors of a homeomorphism. Theorem 3.12.12. Let X be a space with dim.Y < n, and let /: X —> X be a fixed-point free homeomorphism. If f is colorable, then it can be colored with n + 3 colors. Remark 3.12.13. The space X — {0, 1, 2} and the homeomorphism / = {(0,1), (1,2), (2,0)} of X show that Theorem 3.12.12 is sharp for n — 0. By results of STEINLEIN [380, 379] it follows that Theorem 3.12.12 is sharp for every n. Proof. Let k be the minimum cardinality of a coloring of /, and let be a coloring. Assume that k > n + 3. We will derive a contradiction. Observe that Ui / Uj if i ^ j. By Theorem 3.12.10 we may without loss of generality assume that for any 3~ C li of cardinality n + 3 we have
Since / is a homeomorphism, this means that for any 7 C U of cardinality n + 3 we have
(i) Let y = {FI, . . . , Ffc} be a closed shrinking of U (Proposition A. 7.1). For every i < k — 1, define
B, = F, u (Fk n f[x \ U,} n rl[x \ u,}) . We claim that £ = {£?i, . . . , Bk-i} is a coloring of /. To see that f[Bi] n B, - 0
observe that
/[Bi] - f[F,} u (f[Fk] n f2[x \ u,} n (X \ [/,)) . This gives us what we want since
Fx n /[F,] = 0,
F, n (X \ U,} = 0,
Fk n f[Fk] = 0.
246
3. BASIC DIMENSION THEORY
To see that !B covers, first note that \j*ll Fi C (J*!/ Bi. Hence it suffices to show that Fk is covered. To this end, pick an arbitrary x G Fk and consider the collection Since k — 1 > n + 2, (1) implies that there exists i < k — I such that But then x 6 Bi. This contradiction completes the proof of the theorem.
D
Before turning to the proof of Theorem 3.12.10, we make the following remarks. Recall that we wish to shrink an open cover U to an open cover V={Vu:U€ U} such that for any subfamily W C U of cardinality n + 3 we have
PI (uu A moments reflection shows that our task is to construct V in such a way that for any subfamily W of V of cardinality n + 3 and any partition 3" U S of W we have
Lemma 3.12.14. Let X be a space with dim .AT < n. Let Fi C Ui fori < m, with Fl closed and Ui open. In addition, Let Gj C U'- for j < k, with Gj closed and £/• open. I f m + k>n + 3 then there exist closed subsets
FI, . • . ,Fm,
GI, . . . , Gk
such that
(1) Fl C Ui for i < m, and G3 C C/j for j < k, (2) (J™! Ft C U^ x F,, and \Jkj=1 G3 C \Jk.=1 Gjt
(3) n^i^nn* =1 G,- = 0.
Proof. We assume without loss of generality that m + k = n + 3. Suppose first that m > n + 2. Since dim^ < n, there exist open sets Vi, . . . , Vn+i such that
(4) F, C Vl C V, C Ui for i < n + I , So we get what we want by a direct application of Lemma 3.2.3 (let Ai = Fi for i < n + 1 and A = Fn+z}. We may therefore assume without loss of generality that m < n + 1 and, similarly, that k
3.12. COLORINGS OF MAPS
247
Observe that for i < m — I and j < k — 1 we have Fi n Fm C Ui,
GJ n Gk C U'j.
Since dim X < n and (m — 1) + (k — 1) = n + 1, there exist open sets Vi,...,Vm_i,
Wi,...,Wfc_i
such that (6) F, n Fm C y, C Vi_^ Uz for i < m - 1, (7) Gj n G fc C Wj C Wj C tfj for j < k - 1, (8) nS'F^nn^FrW, =0.
So we again get what we want by a direct application of Lemma 3.2.3.
D
Proof of Theorem 3.12.10. Let 5F be a closed shrinking of li (Proposition A. 7.1). Fix n + 3 different elements of U, say,
g = {c/ 1 ,...,[/ m ,c/{,...,t/a, where m + k = n + 3. Let
be the elements of y corresponding to the elements in 9. So Fi C Ui for i < m and f - l [ G j ] C f - l [ U f i for j < k. Since m + k = n + 3, by Lemma 3.12.14, there exist closed sets A\ , . . . , ATO, BI , . . . , Bk such that (1) Ai C Ui for i < m, and Bj C f
[U',] for j < k,
Put Aj = f[Bj] for j < fc. In J replace Fl by A; for i < m and GJ by A'- for j < k. The other elements of 3r are not being replaced. We claim that the collection 3*' obtained in this way covers X, and hence is a closed shrinking of U. To see that it covers, pick an arbitrary x E X. If x E UI^i Fi then we are done by the first part of (2). If x E Uj=i GJ then f ~ l ( x ] E Uj=i f~1[Gj] Q Uj=i Bj by the second part of (2), As a consequence, for some j < fc, x E A'-. Finally, if x does not belong to any element of 'K then the element of 3" that contains x belongs to 3"'. Observe next that f|™ i A» n f|*U /"M^] = 0Since / is a homeomorphism, there exists an open swelling V of 3"' which is simultaneously a shrinking of U (Corollary 3.2.2) and which is as required. So we arrive at the conclusion that U has an open shrinking V such that the Vs and /-1[V]'s corresponding to 9 have empty intersection. The same procedure can be repeated with any subfamily of V of cardinality n + 3. So after finitely many steps, we arrive at the required shrinking of U. D
248
3. BASIC DIMENSION THEORY
Corollary 3.12.15. Let X be a space with dimX < n and let f be a fixedpoint free homeomorphism of X. Then f can be colored with at most n + 3 colors. Proof. By Theorem 3.12.8, / is colorable, and by Theorem 3.12.12 the number of colors can be reduced to n + 3. D Coloring continuous functions. An inspection of the proof of Theorem 3.12.12 shows that it very strongly depends on the fact that we are dealing with homeomorphisms. In fact, as Theorem 3.12.7 shows, homeomorphisms are essential for the basic coloring result Theorem 3.12.8 and they are also essential for the process of reducing the number of colors in Theorem 3.12.12. We will investigate now the natural question whether the results obtained so far can be generalized. In the light of Theorem 3.12.7, the only reasonable question is whether colorable maps can be colored by a color of small size. So we are aiming at a generalization of Theorem 3.12.12. The only place in the proof of Theorem 3.12.12 where we used that the map under consideration is a homeomorphism, is in the proof of Theorem 3.12.10. So we will first think about the question whether Theorem 3.12.10 is also true for continuous maps instead of homeomorphisms. For n = 0 there are no problems. To see this, assume that X is zerodimensional, U is a finite open cover of X, and /: X —> X is continuous. Since dim^f = 0, there is an open shrinking V = {Vy : U £ U} of U with ord(V) < 0, i.e., V is pairwise disjoint (Theorem 3.4.4). But then V is as required. For let x e X, and observe that there is precisely one element of V that contains x. Similarly, there is precisely one element of /^[V] that contains x. But for larger n, Theorem 3.12.10 does not hold for continuous functions, as the following simple example shows. Example 3.12.16. Let X be the topological sum of the spaces I n , n e N, where each I n is a copy of I. In addition, let {di : i > 2} be a countable dense subset of Ii. Now define /: X —>• X as follows. / f l i is a homeomorphism from Ii onto I 2 . In addition, / f Ii is the function with constant value di for every i > 2. Then / is clearly continuous (and has no fixed-point). For every i e N let Ei and Fi be open subsets of Ii such that
Ei\JFi= I, and Define
3.12. COLORINGS OF MAPS
249
Then U — {C/i, . . . , C/4} is an open cover of X. Let V = {Vi,...,^} be an arbitrary open shrinking of U. Then V\_ and V?, cover Ii, and are clearly proper subsets of Ii. So by connectivity of Ii, V\ D V? ^ 0. Pick i > 2 with di 6 Vi n V2- Since Vs and V4 cover \Ji>2 Ii, the same argument shows that V3 O 1/4 n Ii 7^ 0. We conclude that as required. This example shows that if one wishes to color continuous functions, the method used so far does not work. Fortunately, there is another way to obtain what we want. Theorem 3.12.17. Let X be a space and let f : X -> X be fixed-point free and continuous. If f is colorable and dimX < n, then f can be colored with at most n + 3 colors. Proof. We claim that by Corollary 3.5.7 we may assume without loss of generality that X is compact. For let aX be an at most n-dimensional compactification of X such that / can be extended to a fixed-point free continuous function /: aX —> aX. By compactness of aX the function / can be colored. So if the theorem for compact spaces is true then / can be colored with at most n + 3 colors. It suffices therefore to observe that if y colors / then 2F \ X colors /. So assume that X is compact. We also assume for the time being that / is surjective. The general case will be derived from this special case and will be dealt with later. Consider the dynamical system (X, /) and the inverse system \r
j
/_
~\r
.
J_
j
J_
~\r
.
J
_
It was shown on Page 92 that the shift function a : X^ -> X^ defined by a(xi,X2,...) = (/(zi), £1,2:2,...). is a homeomorphism. Moreover, for each n the diagram
commutes and a is fixed-point free (Lemma 1.10.24). Now observe that dim ^00 < n by Exercise 3.2.2. So by Theorem 3.12.10 there is a closed cover y of X^ such that a[F] fi F — 0 for every F e J
250
3. BASIC DIMENSION THEORY
while moreover |3"| < n + 3. Now for each F 6 3~ by compactness there exists an NF e N such that f ^ [ F ] n /^[cr[F]] = 0 for every m > NF (Exercise 1.10.12). So if n - max{7VF : F e 3~} then /~[F] n /£° [
K = n /TO oo
i—l
Then f[K] = K by Exercise 1.10.18. By the result just proved there is a cover C = {Ci : i < n + 3} of K consisting of relatively open sets such that f[Ci] n Ci — 0 for every i. Let {Di : i < n + 3} be a closed shrinking of C (Proposition A. 7.1). Then f[Di] n Di = 0 for every i. Let Ul C X be an open neighborhood of Di for i < n + 3 such that f \Uj\r\Ui = 0. The union U of the t/i's is an open neighborhood of K. By compactness, for a sufficiently large k we have fk[X] C [7. So
is an open cover of X . We claim that it is a coloring of /. To prove this, take an arbitrary i < n and assume that there exists Then f k ( x ) e C/; and there exists an element z 6 f~k[Ui] such that /(z) = x. Then on the one hand fk (x) 6 Ui while on the other hand This shows that f k ( x ) G C/i fl /[E/;], which is a contradiction.
D
Exercises for §3.12. Let /: X —> X be a fixed-point free and continuous. If / is colorable then its color number C(f) is the minimal cardinality of a coloring of /. If / cannot be colored then we put C(f) = oo. Suppose that / : X —> X is a fixed-point free and continuous. We say that a subset A of X is invariant if f[A] C A. Observe that a color of / \ A is a subset of A 1. Let a,b,A,B C N be such that a and b are finite, while A and B are infinite. Prove that [a, A] C [b, B] iff a D 6, a \ 6 C B and A C B. 2. Prove that the collection in (*) on Page 239 is a base for a topology on [N]'" .
3.13. VARIOUS KINDS OF INFINITE-DIMENSIONALITY
251
3. Let A be a countable collection infinite subsets of N. Prove that there is an infinite subset X C N such that for every A G A we have XnA^®^
An(N\X).
4. Prove that the Ellentuck topology on [N]" does not have a countable base. We now present an interesting and unexpected addition theorem for the color number of a fixed-point free map. Theorem 3.12.18. Let X be a space and let /: X —> X be a continuous fixedpoint free function such that X — A U B for invariant subsets A and B of X . Then
The following two simple exercises will be useful. 5. Let f : X —>• X be a fixed-point free map with invariant subspace A. Suppose that the subset C C A is a closed color of / \ A. Prove that there exists an open subset W of X such that (/ \ A)~l[C] C W and W is a color of /. 6. Suppose that / : X —> X is a fixed-point free map and that C is a collection of n colors of / each of which is open or closed. Prove that there exists an open coloring of / of size at most n. From Exercise 3.13.6 it follows that the color number of a map /: X —> X is the minimal cardinality of a cover of X consisting of colors that are open or closed. The following exercise relates an open coloring of an invariant subspace to a collection of open colors of the ambient space. 7. Suppose that / : X —>• X is a fixed-point free map and that the subset A of X is invariant. Let II be an open coloring of / \ A with |1I| = in < oo. Prove that there exists a collection V consisting of at most ra open colors of / such that A C \J V. We are now in a position to finish the proof of Theorem 3.12.18. 8. Prove Theorem 3.12.18. 3.13. Various kinds of infinite-dimensionality Recall that a space X is strongly infinite- dimensional if it has an infinite essential family of pairs of disjoint closed sets. By Corollary 2.4.13, the Hilbert cube Q is an example of a strongly infinite-dimensional space. A space X is called weakly infinite- dimensional if it is not strongly infinitedimensional, i.e., if for every family
of pairs of disjoint closed subsets of X there exist partitions Di between and Bi such that A = 0-
252
3. BASIC DIMENSION THEORY
Proposition 3.13.1. Let X C Q be weakly infinite-dimensional. Then for every sequence {(A;, Bi) : i G N} of pairs of disjoint closed subsets ofQ there exist partitions Di between Ai and Bi in Q such that
Proof. For every i 6 N let Ui and Vi be disjoint closed neighborhoods of Ai and Bi, respectively (Corollary A. 4. 3). Since X is weakly infinitedimensional, the sequence {(Ui n X, Vi D X),i 6 N} is inessential. Consequently, there exist partitions Si between UiC\X and Vi n X in X such that the intersection of the 5;'s is empty. By Lemma 3.1.4 there exists for every i a partition Di between Ai and Bi in Q such that Di n X C Si. Then clearly oo
oo
fl Di n x c p| s, = 0, i=l
i=l
as required.
D
Finally, recall that a space X is called countable dimensional if it can be written as the union of countably many zero-dimensional subspaces. So every finite dimensional space is countable dimensional by Corollary 3.3.9, but not conversely. An important example of a countable dimensional space which is infinite-dimensional is given in the following example. Example 3.13.2. As in §1.1, let a = {x e s : (3n e N)(Vm > n)(xm = 0)}.
For each n G N put
an = {x E s : (Vm > n)(xm = 0)}. It is clear that an is homeomorphic to R n , n e N, and hence is the union of n + 1 zero-dimensional subspaces (Corollary 3.3.9). Since a is equal to the union of the crn's, it follows that a is countable dimensional. Observe that a is infinite-dimensional by Theorem 3.2.12 and Corollary 3.4.14(1). Observe that every subspace of a countable dimensional space is again countable dimensional. Example 3.13.3. There exists an infinite-dimensional compact space which is countable dimensional. This is a triviality. Let X be the one-point compactification of the topological sum of the spaces I, I2, I3, • • • . Then X is infinite-dimensional for the same reason a is. In addition, X is countable dimensional by Corollary 3.4.14(1). Proposition 3.13.4. Every countable dimensional space is weakly infinitedimensional.
3.13. VARIOUS KINDS OF INFINITE-DIMENSIONALITY
253
Proof. Let X be countable dimensional. Then X is the union of countably many zero-dimensional subspaces, say Xi, i G N. Let {(Ai,Bi) : i 6 N} be a family of pairs of disjoint closed subsets of X. By Corollary 3.1.6 for every i there exists a partition Di between Ai and £?; such that Di D Xi = 0. Since the union of the X^s equals X, the intersection of the IV s is empty, which is as required. D Corollary 3.13.5. No strongly infinite- dimensional space is countable dimensional. Corollary 3.13.6. The Hilbert cube cannot be written as the union of countably many zero- dimensional subspaces. In view of the above results, the question naturally arises whether the converse to Proposition 3.13.4 holds. This question was known as Alexandroff's problem for decades, [8, §4, Hypothesis]. We shall answer this problem in the negative. However, we need to derive the following result first. Theorem 3.13.7. Every topologically complete space X has a compactification 'jX such that the remainder ^X \ X is countable dimensional. Proof. We may assume that X is a subspace of the Hilbert cube Q (Corollary A. 4. 4). Since X is topologically complete, there is a decreasing sequence (Un)n of open sets in Q such that
n—l
(Theorem A. 6. 3). Since Q is compact, for each e > 0 there is a finite open cover of Q with mesh less than e. Consequently, for each n there exist finitely many open sets £/n,i, . . . , Un,mn C Q such that (1) £7n,i U • • • U Unjmn = Un, (2) mesh({t/ n ) 1 ,...,*7 n > m n })< i/n. For each n G N and m < ran, define / n>m : Q —> I by fn,m(x)
= 1/2&(X, Q \ Un,m)
and define / : Q —> Q coordinate wise as follows: f ( x ) = (fi,i(x), • • • , f i , m i ( x ) , f 2 , i ( x ) , . . . , /2,m 2 (a:),/3,i(a;), . . . , - . . ) • By (2) it follows easily that / \ X is an imbedding. We claim that f[Q]\f[X]Ca. Take an arbitrary point f(q) E f[Q] \ f[X]. Since q $. X, by (1) there exists an N € N such that q £ UN. Consequently, since the C/n's decrease, for every n > N and m < mn we have fn,m(q] = 0,
254
3. BASIC DIMENSION THEORY
as required. Now let jX be the closure of f[X] in f[Q]. Then jX is a compactification of X the remainder 'jX \ X of which is contained in the countable dimensional space a (Example 3.13.2). So 7 A7" \ X is countable dimensional as well. D The topological completeness of the spaces under discussion in the above theorem is essential. It can be shown that a is an example of a countable dimensional space the remainder of each compactification of which is strongly infinite-dimensional, see Exercise 3.13.4. We shall now present the solution to Alexandroff's problem. Theorem 3.13.8. There exists a weakly in finite- dimensional compact space which is not countable dimensional. Proof. By Theorem 3.9.3 there exists a strongly infinite-dimensional totally disconnected and topologically complete space, say S. Let X — jS be a compactification of S with countable dimensional remainder (Theorem 3.13.7). We claim that X is the required example. To this end, first observe that X is not countable dimensional since by Corollary 3.13.5 its subspace 5 is not countable dimensional. Now let T = {(A;, BI) : i > 0} be a sequence of pairs of disjoint closed subsets of X . Write X \ S as the union of countably many zero-dimensional subspaces, say Xi, i € N. By Corollary 3.1.6 for each i e N there exists a partition Di between Ai and Bi such that Di Pi Xi = 0. We conclude that the compact set D — Hi^i Di is contained in 5, and S being totally disconnected, therefore has to be zero-dimensional (Exercise 1.5.12). By another appeal to Corollary 3.13.5 there exists a partition DO between AQ D and BQ such that D0 n D = 0. We conclude that f£0 D* = 0An infinite-dimensional space X is called hereditarily infinite- dimensional if for every nonempty subspace A of X either dim A = 0 or dim A = oc. A hereditarily infinite-dimensional compact space is clearly Henderson, but the converse need not be true. See the notes for more information. We finish this section by presenting an example of a hereditarily infinitedimensional space. Lemma 3.13.9. Let K C JJ be a Cantor set, let £ 6 N and Jet K C N\{£} be an infinite set the complement of which is also infinite. For each j; € K there exists a partition Sj between the opposite faces W~l and Wj ofQ such that JS each subset X C S w^n K infinite-dimensional. Proof. Without loss of generality we may assume that £ = I and K is the set of even natural numbers. For each i e N let Ci = vr-^-l, - 1/2] and A = ^[1/2, 1],
3.13. VARIOUS KINDS OF INFINITE-DIMENSIONALITY
255
respectively. Now put A = {/ e C(Q,Q) : (Vi 6 M)(/-1[I^-1] = C2t and f-^Wft
= D2l)}.
Then A is a separable metric space since by Proposition 1.3.3 C(Q, Q) is. By Exercise 1.5.10 there exists a subspace T of K such that T can be mapped onto A, say by the map T. Let £=TxJ2xJ3x---cg and define 4> : E —> Q by $0) =T(XI)(X). Then 4> is clearly continuous. Claim 1. For each i, (7TiO$)~ 1 (0) is a partition in £ between the sets C and D2i n E. Proof. Fix i 6 N. Then
and similarly, ^>~1[VFi1] = £^ n D 2 i- Now since 7rl~1(0) is a partition between W~l and VF/, the claim follows. 0> Since W^1 and Wg^ are in the interior of C^i and D 2 i, respectively, there exists by Lemma 3.1.4 a partition 6*2 i between W^1 and l/t^i m Q such that (*)
^ a . H ^ C (TTioS)- 1 ^).
We claim that
s
is as required. Claim 2. SnEC ^(O). Proof. This immediately follows from (*).
<0
Now let M be a subspace of S such that K C 7Ti[M]. We shall prove that M is strongly infinite-dimensional. Striving for a contradiction, suppose that this is not the case, i.e., that M is weakly infinite-dimensional. By Proposition 3.13.1 for i 6 N there exist partitions Ml in Q between C 2j and D-zi such that P l ^ 1 M i n M = 0. For i € N there exists a continuous function fi : Q —>• JJ; such that
256
3. BASIC DIMENSION THEORY
(Exercise A.4.9). Define /: Q -> Q by /(x) = (/!(x),/ 2 (x),...). Claim 3. / 6 A and M n /-1(0) = 0Proof. That / is continuous is trivial. Now take i G N. Then by the definition of /;. Similarly one proves that /^[W/] = D-U. We conclude that / G A. Finally, oo
f - l ( o ) n M = Pi Mi n M = 0, i=l
as required.
<)
Now since r is onto, there exists p G T such that r(p) = /. Since T C 7Ti[M], there exists x G M such that #1 = p. Then x £ S n E from which it follows by Claim 2 that f ( x ) = r(p)(x) = T(XI)(X)
= $(x) = 0.
But this contradicts Claim 3.
D
This result enables us to construct our final example. Theorem 3.13.10. There exists a hereditarily infinite-dimensional compact space. Proof. By Exercise 3.13.6 there exists a collection {Ki : i G N} of pairwise disjoint Cantor sets in JJ such that each nondegenerate subinterval of JJ contains one of the Ki. So for each i choose pairwise disjoint such Cantor sets Kn,Ki2, . . . in Ji = [— 1, l]i. In addition, let Kik (i, k G N) be a collection of pairwise disjoint infinite subsets of N \ {1} such that
i 0 Kik for all i and k. Fix i, k E M. By Lemma 3.13.9, for every j G Kik we can choose a partition Sj in Q between W~l and Wj such that each subset in the intersection
whose projection onto the i-fh axis contains Kik, is infinite-dimensional. Now put
We claim that 5 is hereditarily infinite-dimensional. Clearly 5 is a compact subspace of Q. Since 1 & \JikKik, by Corollary 2.4.13 S intersects every partition between W±l and W±. By Exercise 3.1.13 it therefore follows
3.14. THE BROUWER FIXED-POINT THEOREM REVISITED
257
that S is not zero-dimensional (for otherwise there would exist a partition P between W^1 and W^ such that P fi S — 0, contradicting the fact that the family of opposite faces of Q is essential). Now let M be an arbitrary nonempty subspace of S. There are two cases to consider. First assume that for each i e N, 7Tj[M] is zero-dimensional. Then M is contained in the set oo
Jj7T z [M],
which is clearly zero-dimensional. So then dim M = 0 as well. Next assume that there exists i 6 N such that 7T;[M] is not zero-dimensional. Then 7T;[M] contains a nondegenerate subinterval of JJ^ by Proposition 1.5.3. Consequently, by construction, TT^fM] contains some Kij which implies that M is (strongly) infinite-dimensional because M is contained in Sij. D Exercises for §3.13. A real-valued function / on a space X is countably continuous provided that X can be partitioned into countably many sets EI, #2, • • • such that for every i the restriction / \ Ei is continuous. 1. Prove that if a space X is the union of countably many weakly infinitedimensional subspaces then X is weakly infinite-dimensional. Conclude that the Hilbert cube is not the union of countably many weakly infinitedimensional subspaces. 2. Let X be a strongly infinite-dimensional compact space. Prove that X contains a hereditarily infinite-dimensional closed subspace. ^•3. Prove that if G is a GVsubset of s containing a then G\o contains a copy of the Hilbert cube. 4. Prove that if 70- is a compactification of a then the remainder 70- \ a contains a copy of the Hilbert cube and is therefore strongly infinitedimensional. ^•5. Give an 'explicit' example of an upper semi-continuous function that is not countably continuous. 6. Prove that there exists a collection {Ki : i e N} of pairwise disjoint Cantor sets in J such that each nondegenerate subinterval of JT contains one of the Ki. 3.14. The Brouwer fixed-point theorem revisited In this section we show that the existence of a two-dimensional hereditarily indecomposable continuum, which uses Brouwer's Fixed-Point Theorem for I3 only, implies the Brouwer Fixed-Point Theorem for all dimensions. This seems to show that the 'real' power of the Brouwer Fixed-Point Theorem is already attained at dimension three.
258
3. BASIC DIMENSION THEORY
Kelley's construction. Let X be a hereditarily indecomposable continuum with dimX > 2. The existence from such a space follows from Corollary 3.8.3. An inspection of the proof shows that for the construction of X one needs Brouwer's Fixed-Point Theorem for I 3 only. We will show that from X we can build an infinite-dimensional space. Our strategy will be to 'blow up' the points of X and to topologize them in such a way that we get an infinite-dimensional hereditarily indecomposable continuum. Since dimX > 2, by Exercise 3.7.5 the following holds for X: (1) there are real numbers 6 > 0 and e > 0 such that if X is any finite collection of pairwise disjoint closed sets with mesh(N) < 26 then there is a continuum C C X \ [J N with diameter at least e. Let us: 2X —>• I be a Whitney map (Proposition 1.11.15). By Exercise 1.11.25 there is an element t 6 (0, oo) such that if A E 2X is such that uj(A) < t then diam(A) < e. Consider the Whitney level
£ = [A e G(X) : u(A) = t}. Then £ is an upper semi-continuous decomposition of X by Corollary 1.11.17. Here we use for the first time that X is hereditarily indecomposable. By Exercise 1.11.24 we find that (2)
7 = inf{diam(E) : E £ £} > 0.
This property of the collection £ is crucial. So for the diameter of an arbitrary element E G £ we have (3)
0 < 7 < diam£<£.
Let Y be the space X/£, and let TT : X —>• Y be the natural quotient map. Since TT is monotone, we already know from general considerations that Y is a hereditarily indecomposable continuum (Exercise 1.10.22). Observe that TT is open by Exercise 1.11.23. So Y is homeomorphic to the subspace £ of 2X, and the formula
d(tfi,y2) = ^(r 1 (yi) J /~ 1 (!/2)) defines an admissible metric on Y (Exercise 1.11.17). This is the metric on Y that we will use in the sequel. Observe that since £ consists of continua, the space Y is in fact homeomorphic to a subspace of G(X). This observation will be used in the solution to Exercise 3.14.1. Having found the collection £ of "blown up" points, we fix any n > 2 and set (4)
6n
3.14. THE BROUWER FIXED-POINT THEOREM REVISITED
259
Observe that by (1) and (2), 6n > 0. Proposition 3.14.1. For any finite closed cover J ofY of mesh less than 6n there is an element A E y intersecting at least n elements of 9r Proof. It will be convenient to introduce the following notation. For A C Y we put Striving for a contradiction, let 9" be a finite closed cover of Y with mesh less than 6n such that each element of A 6 y meets at most n — I elements of y \ {A}. We shall associate to each element A G J a compact set yMCTr-1^]
(5) such that
(6) y(A] H E ^ 0 for any E G £(A), (7) diamip(A) < 26n, (8) if A, B e y are distinct then (p(A) n (p(B] = 0.
Assume that (p is already defined on 9 C 2F, and let A e 9" \ 9 be arbitrary. The set JC of all elements of 9 hitting A has by assumption cardinality less than or equal to n — 1. For each D e 9, let B(D) be the closed 5 about tp(D}. Observe that by (7) we have (9)
diamB(D) < 4Jn.
Let us fix an arbitrary element E(A) e £(^4.). We shall show that (10)
E(A) \ {]{B(D} :D € K] / 0.
Otherwise, ^(A) is covered by the collection {B(D} -.D^'K} and hence, -E(A) being a continuum, by Exercise A. 10. 3 and (9) we get < \3<\ • max diamB(D] < (n - 1) • 45n, which contradicts (4). Let a be any element given by (10) and let B be the closed 5n-ball about a. We set
(11)
(p(A) =BnK-1[A\.
We claim that
260
3. BASIC DIMENSION THEORY
If D E !K then a does not belong to the closed Jn-ball about (p(D}. So the closed Jn-ball about a misses tp(D], i.e., (p(A) r\(p(D] = 0. Moreover, if D belongs to $ \ 'H then A n D = 0 and hence tp(A) n ^>(D) = 0 since (p(A) Cn-^A],
(p(D) CTT-^D],
TT1^] n Tr'1^] = 0 -
So (8) is also satisfied (for g U {A}}. We conclude that we can extend (p over 9 U {A} without creating problems. By the fact that 3~ is finite, we may therefore assume that (p is defined on all of y. The collection (12)
N = {v?(A) : A e ?}
is disjoint and has by (7) mesh less than or equal to 26n < 26. So it can serve as input for (1). To reach a contradiction, let C be an arbitrary continuum in X \ (J N. Let E e £ intersect C and pick A 6 J such that E G £(A). By (6),
Since C does not intersect \J N, this shows that E1 is not a subset of C. But then since X is hereditarily indecomposable, it must follows that C is a subset of E. This implies by (3) that diam C < diam E < e. By (1) there is such a continuum C with diameter at least e. This contradiction establishes the proof. D By Exercise 3.3.6, we now come to the following surprising conclusion. Corollary 3.14.2. The existence of an at least two-dimensional hereditarily indecomposable continuum implies the existence of an hereditarily indecomposable in finite- dimensional continuum. Dimension theory. By Exercise 3.1.16 we know that the Brouwer FixedPoint Theorem and the fact that there is a space X with dim^Y = oo are equivalent statements, in the sense that they are easily deduced from each other. So from Brouwer's Fixed-Point Theorem for I3, implying the existence of a two-dimensional hereditarily indecomposable continuum, we derived the Brouwer Fixed-Point Theorem for all dimensions. This is a rather curious phenomenon and it is not clear whether there is a more direct route from Brouwer in I3 to Brouwer in all dimensions. This is of course only of interest for 'philosophical' reasons. The proof of Sperner's Lemma 2.4.3 is by induction, so it is equally complicated in all dimensions (except for the dimensions one and two). Similar remarks can be made for the various other known proofs of Brouwer's Theorem. This seems to indicate that even if one finds a
3.14. THE BROUWER FIXED-POINT THEOREM REVISITED
261
more direct route from Brouwer in I 3 to Brouwer in all dimensions, this will not result in a simpler proof of Brouwer's Theorem. But this does not mean that clarifying the rather curious phenomenon that we encountered in this section would not be interesting. Exercises for §3.14.
1. Let Z be a compact space with dimZ > 3. Prove that dimC(-Z) = oo. 2. Give an example of a continuum X such that dimC(A') = 2.
This page intentionally left blank
CHAPTER 4
Basic ANR theory In this chapter we shall present several basic results from A N R theory. We introduce and characterize ANR-pairs. They enable us to prove that the hyperspace of a Peano continuum is an AR, which is an important step in the proof of Curtis-Schori-West Hyperspace Theorem that 2X Kt Q if and only if X is a Peano continuum, and that there are deformations of hyperspaces through various interesting subspaces. A curious important result that we will prove is that for a Peano continuum X there are arbitrarily close to the identity maps from 2X into the finite subsets of X. Our concept of an ANRpair, which was motivated by TORUNCZYK's concept of a locally homotopy negligible set from [390], turns out to be a convenient framework for obtaining such results.
4.1. Some properties of AN R's
In this section we shall derive a few useful properties of AN R's. We have to introduce some terminology first. Let Y be a space and let U be an open cover of Y. Two continuous functions f,g:X-*Y are called U-close if for every x £ X there exists an element U G U such that {f(x),g(x)} CU. A homotopy H: X x I ->• Y is said to be limited by U provided that for any x G X there exists U G U such that H[{x} x I] C U.
Two continuous functions f , g : X -> Y are U-homotopic, if there is a homotopy H: X x I —>• Y which is limited by U and connects / and g, i.e., HQ = f and HI = g. Small functions and small homotopies. Here is our first property of ANR's.
Theorem 4.1.1. Let X be an A N R . Then for every open cover U of X there exists an open refinement V of U such that for every space Y, any two V-close maps f,g:Y-+X are U-homotopic. 263
264
4. BASIC ANR THEORY
Proof. By Corollary 1.1.8 we may assume that X is a closed subspace of a normed linear space L. Observe that L is locally convex (Exercise 1.1.6). Since X is an A N R , there are a neighborhood V of X in L and a retraction r: V -> X. Put £ = {r~l[U} : U E U}. Then £ is an open cover of V, and since V is open in L, there is an open cover If of V having the following properties:
(i) if < £ , (2) If consists of convex sets. Let V = y \ X. We claim that V is the required open cover of X. To this end, let Y be a space, and assume that f,g:Y—*X are continuous and V-close. Define a homotopy G : Y x I —> L by
Then G is continuous and clearly connects / and g. Since by (2) the elements of If are convex, by the special definition of G it follows that for every y E Y there exists F G If such that (3)
G({y} xl]CF.
We conclude that G can be considered to be a mapping from Y x I into V and that moreover G is an If-homotopy. Now define H : Y x I —>• X by
H = roG. Then H connects / and g since r is a retraction. Moreover, by (3) and (1) it follows easily that H is limited by U. D Remark 4.1.2. It is a natural problem whether the property of ANR's stated in Theorem 4.1.1 in fact characterizes the class of all ANR's. This was a difficult and fundamental problem which remained unanswered for decades. It was finally solved in 1994 by CAUTY [87] in the negative by his example of linear space which is not an ANR (Remark 1.2.4). To see that Cauty's Example really solves the problem, let L be a linear space. In addition, let U be an open cover of L. The function A : I/ x L x E —> L defined by X ( x , y , t ) = (l-t)-x + t-y, is defined in terms of the algebraic operations on L and is therefore continuous. For every x e L pick an element Ux 6 U containing x. Since A is continuous and X[{x} x {x} x I] = {x}, by compactness of I there exists for every x E X a neighborhood Vx of x such that X[VX x Vx x I] C Ux (Exercise A. 5. 7). Put V = {Vx : x £ L}. We claim that V is as required. To this end, let X be a space and let /, g : X —>• L be continuous V-close functions. Define a homotopy H : X x I —>• L in the obvious way by the formula
4.1. SOME PROPERTIES OF ANR'S
265
Then clearly H0 = f and HI = g. Fix an arbitrary x 6 X. Since / and g are V-close, there exists an element p 6 L such that f ( x ) , g ( x ) G Vp. But then [f(x)} x {g(x}} x I C V p X V p X l from which it follows that H(x,t) = \ ( f ( x ) , g ( x ) , t ) 6 Up for every t 6 I. So this indeed proves that / and g are U-homotopic. Homotopies with control. In many applications one needs homotopies 'with control'. We will therefore present a 'controlled' version of the Borsuk Homotopy Extension Theorem, cf. Theorem 1.4.2. Theorem 4.1.3. Let X be an ANR and let U be an open cover of X. Then for every closed subset A of a space Y and for every homotopy H: A x E —>• X such that (1) H is limited by U, (2) H0 can be extended to a continuous function h0 : Y ->• X there exists a homotopy H: Y x E —>• X such that (3) H is limited by U, (4) H0 = ho and H\(Axl} = H. Proof. Let H, A and Y be as in the formulation of the theorem. By Theorem 1.4.2 there exists a homotopy F: Y x I —> X such that FQ — ho and For each a G A there exists Ua 6 U containing H[{a} x E]. There consequently exists by compactness of I and continuity of H a neighborhood Ea of a in Y such that
F[Ea xl]CUa (Exercise A. 5. 7). Put E = (JaeA Ea . Then E is a neighborhood of A in Y. By Corollary A. 4.1 there is a Urysohn function X:Y-tI such that A f A = 1 and A \ Y \ E = 0. Now define H : Y x I -> X as follows: Then H is clearly continuous and H \ (A x I) = H. In addition, H(y,0)=F(y,Q)=ho(y) for every y 6 Y so ^o — ^o- It remains to prove that H is limited by U. This is a triviality. Take an arbitrary y G Y. If y £ E then there exists a G A such that y e Ea. Consequently, H({y} x I) C F[{y} x E] C t/a.
Now if y $ E then A(y) = 0 from which it follows that
266
4. BASIC ANR THEORY
consists of precisely one point and is therefore also contained in an element of U. D ANR-pairs. We shall now derive another important property of ANR's. For later use, we will state it in terms of ANR-pairs. Let X be a space with subspace Y. We say that the pair (X,Y) is an ANR-paz'r provided that for every space Z and every closed set A C Z and every continuous function / : A —> X there exist a neighborhood U of A in Z and a continuous extension /: U —> X of / having the additional property that f[U\A] C Y. Observe that if ( X , Y ) is an ANR-pair then so is (E,Y), where E is any subspace of X containing Y . Define the notion of an (\R-pair similarly. Observe that if (X, Y) is an A(N)R-pair then both X and Y are A(N)R's (see Exercise 4.1.7). Moreover, if X is an ANR then ( X , X ) is an ANR-pair. Also, Y is dense in X. For if Y is not dense in X then there exists a point x e X such that for some e > 0 we have that B(x,e) D Y = 0. Now let Z — I, A — {0} and /: A —>• X the constant function with value x. Then obviously / cannot be extended over a neighborhood of A with 'new' values iny. Let X be a space, and let Y C X. A deformation of X through Y is a homotopy H : X x I —>• X such that HQ is the identity on X, and Ht[X] C Y for every t 6 (0, 1}. Lemma 4.1.4. Let X be a space and let U be a neighborhood of X x {0}
in X xl. Then there is a continuous function a: X ->• (0, 1] such that {(x,t) :t (0, 1] defined by a(z) =
i/2d((x,Q),(Xxl)\U)
is continuous (and clearly well-defined since Q is bounded by 1). We claim that it is as required. To this end, consider for some x G X a point of the form (x,t), where t < a(x). If ( x , t ) € (X x I) \ U then a(x) =
So we conclude that a(x) — 0, which is a contradiction.
D
4.1. SOME PROPERTIES OF ANR'S
267
Corollary 4.1.5. Let X be a space with subspace Y. If X can be deformed through Y then for every open cover U of X there exists a homotopy H: X x I -> X
such that (1) H deforms X through Y, (2) H is limited by U. Proof. Let F: X x I —> X be a deformation through Y. For every x 6 X pick Ux G U containing x. Since F~1[[/x] is a neighborhood of (x, 0) in X x I, there are an open neighborhood Vx of x and 0 < tx < 1 such that F[Vxx[0,tx]]
CUX.
Put
V = |J Vxx[0,tx). x£X
By Lemma 4.1.4 there is a continuous function a: X —> (0,1] such that {(x,t) :x e X,t< a(x}} C V. Now define H: X x I ->• X by H(x,t) = F(x,t-a(x)). Since a(x) > 0 for every x, we clearly have that H deforms X through Y. Now if x € JC then (x,a(x)) G V. So there exists x' G X such that (z,a(z)) 6 Vx, x [O,^)But then { x } x [ 0 , a ( x } ] CVX, x [0,^) and so
H[{x] x I] C F[{x} x [0,a(x)]] C C/^, as required.
D
These results enable us to characterize A(N)R-pairs, as follows: Theorem 4.1.6. Let X be a space, with subspace Y. The following statements are equivalent: (1) ( X , Y ) is an A(N)R-pair. (2) X is an A ( N ) R and there exists a deformation from X through Y. Proof. We will show this for ANR-pairs only, the proof for AR-pairs being entirely similar. For (1) => (2), consider .Y x 3, its subspace X x {0} and the function /: X x {0} —>• X defined by /(x,0) = x. There exist a neighborhood t / o f ^ L x { 0 } i n ^ C x I and a continuous extension /: U —> X of / such
268
4. BASIC ANR THEORY
that f[U\(X x {0})] C Y. Let a: X -> (0, 1] be the function in Lemma 4.1.4 for U . Now define H : X x I —> X by the formula
Then H is a deformation through Y. Since, as observed earlier, X is an ANR, we are done. For (2) => (1), let X be an ANR and H : X x I ->• X be a deformation through Y. In addition, let Z be a space, A C Z be closed and f : A -> X continuous. Since X is an A N R , there are a neighborhood U of A in Z and a continuous extension f:U-*X. Let Q be an admissible metric for Z which is bounded by 1 (Exercise A. 1.6), and define g: Z -> X by the formula
Then g is clearly as required.
D
As we saw, it is sometimes possible to replace a homotopy by one with better properties. Another example of such a phenomenon is the following result, which shall be used frequently in the sequel. Proposition 4.1.7. Let (X, g) be a space with subspace Y and let H be a deformation of X through Y. Then H can be replaced by a deformation F of X through Y such that
g(F(x,t},x) < t for all x E X and t el. Proof. For every x 6 X and t G I put G(x,t) = {s e I : d i a m ( H [ { x } x [0,s]]) > *}. Observe that possibly G(x,t) = 0. Claim 1. G(x,t) is closed in I for every x and t. Proof. This follows easily from the continuity of the functions H and diam (Exercise 1.11.12). <> Now define f : X x I ->• I by J ^'
J
[
By Claim 1, £ is well-defined. Claim 2.
is Isc.
4.1. SOME PROPERTIES OF AIMR'S
269
Proof. Suppose that for some v G [0, 1), x G X and t G I we have £(x, i) > v. Suppose that there are sequences (x n ) n in X and (£ n ) n in I such that
while moreover £(xn,tn) < v for all n. We may assume without loss of generality that £(x n , t n ) —> w < t>. Observe that since v < 1, diam (H({xn} x [0, £(x n ,£ n )]) > tn for every n so that by continuity of H and diam (Exercise 1.11.12) we get
This shows that £(x,t) < w < v, which is a contradiction.
Q
Let j] : X xl —>• I be the constant function with values 0. Then r\ is continuous, and hence use. By Corollary A. 7. 6 there is a continuous function S : X xl —>• I such that for every ( x , t ) G X x I we have 0 < £(z,t) <£(£,*) and if 0 < ^(x, i) then 0 < 6 ( x , t ) <£(x,t). Now define the homotopy F : X x I —>• X by the formula
Then F is clearly continuous. We claim that it is as required. To this end, pick arbitrary x 6 X and t G I. Then g(x,F(x,t))
=g(x,H(x,6(x,t)))
is less than t if £(x, t} > 0 and 0 < £ otherwise. It therefore suffices to prove that F deforms X through Y' . It is clear that 0(z, F(x, 0)) = 6(x, H(x, e(ar, 0))) = 0(ar, ff (x, 0)) = 0,
i.e., F(x,0) = x. Finally, take arbitrary x e X and t € (0, 1]. If f(a:,£) = ° then 0 = di&m(H[{x} x {0}]) > t > 0 which is impossible. So £ ( x , t ) > 0 from which it follows that 6(x,t] > 0 and consequently F(x,t) = H(x,6(x,t)) as required.
EY, D
Corollary 4.1.8. Let Z C Y C X. If (Y, Z) and ( X , Y ) are A(N)R-pairs then so is (X, Z) .
270
4. BASIC ANR THEORY
Proof. We prove this for ANR-pairs only, the proof for AR-pairs being entirely similar. Let F: X x I —> X be a, deformation through Y (Theorem 4.1.6). We may assume by Proposition 4.1.7 that g ( H t ( x ) , t ) < t for all x 6 X and t £ I. Similarly, let G: Y x I —>• Y be a deformation through Z having the property that g ( G t ( y ) , t ) < t for all y e Y and t e I. Define the homotopy H: X x I —>• .Y by V
'
\G(F(x,t),t)
(*>0).
We only need to check that H is continuous at points of the form (x, 0). So let (xn)n be a sequence in X converging to x and let (tn)n be a sequence in (0,1] such that tn \ 0. Then
< 0(x,xn) + g ( x n , F t n ( x n ) ) + g ( F t n ( x n ) , G t n ( F t n ( x n ) ) _ Q\3-"> -En) i ^n
So we are done by another application of Theorem 4.1.6.
D
Partial realization. The following interesting property of ANR's is not very appealing at first glance. It turns out however that it is of crucial importance in the process of finding a usable topological characterization of ANR's. We need to introduce some terminology first. Let X be a space and let U be an open cover of X. In addition, let T be a locally finite simplicial complex and let § be a subcomplex of T containing all the vertices of T. A partial realization of T in X relative to (S, U) is a continuous function such that for every a £ T there exists U 6 U such that In case S = T we say that / is a full realization of T in X relative to U. Theorem 4.1.9. Let X be an ANR. Then for every open cover U of X there exists an open refinement V of U such that for every locally finite simplicial complex T and every subcomplex S of T containing all the vertices of 7, every partial realization of T in X relative to (S, V) can be extended to a full realization of 7 in X relative to U. Proof. Recall that by Corollary 1.1.8 we may think of X as a closed subspace of a normed linear space L. Since X is an ANR, there are a neighborhood V of X in L and a retraction r: V ->• X. Put £ = {r~l[U] : U G U}. Then £ is
4.1. SOME PROPERTIES OF ANR'S
271
an open cover of V, and since V is open in L, there is an open cover 3 of V having the following properties: (1) ? < £ , (2) J consists of convex sets.
As in the proof of Theorem 4.1.1, put V = 3 \ X. We claim that V is the required open cover of X. To this end, let T be a locally finite simplicial complex, let S be a subcomplex of T containing all the vertices of T, and let / : |S| -> X be a partial realization of T in X relative to (S, V). For each a G T let a* denote the convex hull of f[cr D |S|] in L; observe that by (2) a* is contained in an element of y. For each n > 0, let
Kn = 7(n) U § (here T^ n ^ denotes the n-skeleton of T of course, see §2.1). Then OCn is a subcomplex of T. By induction on n > 0 we shall construct a continuous function fn: \%n\ —>• V such that the following conditions are satisfied: (3) /o = /, (4) if n > 1 then fn extends / n _i, (5) for every a G T, fn[a n \3Cn\] C a*.
Observe that /o is well-defined since T^0) C §. Now assume that for certain n > 0 we constructed the function fn-i- We will define fn on every simplex a G OCn separately and will conclude by applying Lemma 2.1.19 that the union of all the constructed functions is continuous. We have to be careful of course since we want the union of all the functions to be well-defined. So take an arbitrary simplex a G OCn — T^ U §. We want to define fn on a. There are two cases to consider. First assume that a G 3Cn-i- Then put
\ a. Suppose therefore that a £ 3C n _i, i.e., a G 7^ \ 3C n _i. Then
(x G T' n >).
272
4. BASIC ANR THEORY
Observe that (p is clearly well-defined by (4) , that the range of (p is contained in V by (5) and that (p extends / by (3). In addition, by another appeal to Lemma 2.1.19 it follows that
g = ro(f. Since r is a retraction and (p extends /, g extends / as well. Now take an arbitrary a E T and an element W E V such that There exists n > 0 such that a E 7^n\ By (5) and the definition of (p we obtain (p[u\ C a*. There exists F E 3~ such that F D X = VF; observe that a* C F. By (1) there exists t/ E U such that F C r"1 [£/"]. We conclude that sM=r[
is finite. By Exercise A. 7.1 it therefore follows that Wx =
p| Un \ ( |J
n£Ax
Fn)
n£Ax
is open. Put W = {Wx : x G X}. Then W is an open cover of X. We claim that ,W) :x G X} < U .
To this end, take x G X and fix an arbitrary n G Ax. Let Wy e W be such that x G Wy. Then clearly Ax C Ay from which it follows that Wy C Un. We conclude that St({x}, W) C Un, as required.
4.1. SOME PROPERTIES OF ANR'S
273
Repeating this argument for W we get an open cover Wi of X such that for every x € X there is W € W with St({x},Wi) C VF. Observe that this implies that Wi refines W. Now let W\ G Wi be arbitrary. We are interested in computing St(Wi, Wi). To this end, pick an arbitrary x 6 Wi and let W{ e W be such that W{ C\Wi ^ 0, say y e W[ n Wi . There exists by construction an element W G W such that St({y},Wi) C W. We conclude that x e Wi C Wi U W[ C St({y}, Wi) C W and so ^ C St({x},W). This shows that St(^i,Wi) C St({x},W) from which it follows that St(VKi, Wi) C C7 for some Z7 e U. By Theorem 2.3.5(1) there is a star-finite refinement V of Wi. Then V is clearly as required. Now assume that U is finite. Then there is no need to refine U by a locally finite open cover. Observe that the cover W is finite in that case and that the same is true for Wi . So Wi is a star-refinement of U which is obviously star-finite, being finite. D Corollary 4.1.11. Let (X, Y) be an ANR-pair. Then for every open cover U of X there exists an open refinement VofU such that for every locally finite simplicial complex T and every subcomplex § of T, containing all the vertices of 7, and every partial realization f : |S| —> Y of 7 in Y relative to (S, V \Y) can be extended to a full realization g: |T| —> Y of 7 in Y relative to U \Y. Proof. Let U' be a star-refinement of U (Lemma 4.1.10). Since X is an ANR, we let V be the refinement of U' we get from Theorem 4.1.9. We claim that this V also works here. By Corollary 4.1.5 there exists a deformation H from X through Y which is limited by U'. Now let a locally finite simplicial complex T be given, let § be a subcomplex of T, containing all the vertices of T, and let /: |S| —> Y be a partial realization of T in Y relative to (§, V \Y). Since / is also a partial realization in X relative to (S, V), / can be extended to a full realization g of T in X relative to U'. We now use H to push g into Y, as follows. First observe that |T| is a metrizable space, and that |S| is a closed subspace of |T| (Lemma 2.1.14 and Proposition 2.1.21). Let Q be an admissible metric for |T| which is bounded by 1 (Exercise A. 1.6), and define /: |T| —> X by the formula
Then / is clearly continuous. If x € |S| then g(x, |S|) = 0, and so
274
4. BASIC ANR THEORY
Moreover, if x $ |8| then g(x, |S|) > 0 and so J ( x ) G Y. We conclude that the range of / is contained in Y. Moreover, since H is a U'-homotopy, the functions / and g are U'-close. Pick an arbitrary a G T. There is an element UQ G U' such that g[a] is contained in UQ. In addition, g and / are U'-close. For every x € a there consequently exists an element U'x G U' containing g(x] as well as J(x). We conclude that f[a] is contained in St([/o,U'). So we are done since U' is a star-refinement of U. D Dominating polytopes. Let X and Y be spaces. We say that X dominates Y if there are two maps g:Y-*X,
f-.X^Y,
such that the composition / o g \ Y —>• Y is homotopic to the identity on Y. The space X is called a dominating space for Y. Lemma 4.1.12. A space X is contractible if and only a one-point space dominates X . Proof. Suppose first that X is contractible and let H: X x I —> X be a homotopy contracting X to a point, say p. Now let g : X —>• {p} be the obvious function and let / : {p} —>• X be the inclusion. Then H connects the identity on X with f o g , i.e., {p} dominates X. Conversely, assume that there exists a point p and functions g : X —> {p} and /: {p} —> X such that f o g is homotopic to the identity on X. Since f o g is clearly a constant function we conclude that X is contractible. D As remarked above, one sometimes needs homotopies 'with control'. For that reason we define a 'controlled version' of the concept of domination. Again let X and Y be spaces and let U be an open cover of Y. We say that X U-dominates Y if there are two maps such that the composition f o g : Y —> Y is U-homotopic to the identity on Y. The space X is called a U- dominating space for Y. Theorem 4.1.13. Let X be a (compact) ANR. Then for every open cover U of X there exists a (polyhedron) polytope P such that P U-dominates X. Proof. Let V be an open refinement of U with the properties stated in Theorem 4.1.1. Let V be a star-refinement of V (Lemma 4.1.10). By Theorem 4.1.9 there exists an open refinement W of V such that for every locally finite simplicial complex T, for every subcomplex S containing all the vertices of T, and for every partial realization / : |S| —> X of T in X relative to (S, W) there exists a full realization g: |T| ->• X of T in X relative to V such that g extends /.
4.1. SOME PROPERTIES OF ANR'S
275
Claim 1. Such a full realization g has the following property: For every a e T and for every W € W with /[a n S|] C W there exists V € V such that #[0-] U W C V. Proof. Let cr e T and W e W be such that /[cr n |S|] C W. There exists an element V 6 V such that 0[
A0 n AI n • • • n An ^ 0. Since A is a star-refinement of W there exists W 6 W such that AQ U AI U • • • U An C
W.
Consequently, f[a n N(A)W] = f [ { x ( A o ) , x(Al], . . . , x(An)}] C AQ U AI U • • • U An
cw. So we are done.
<$•
Now by the special choice of W, the function / can be extended to a full realization g: \N(A)\ —> X relative to V. Claim 3. The identity function lx on X and the function go K are V-close. Proof. Take an arbitrary x G X. Since A is star-finite, there are only finitely many elements of A that contain x, say AQ, AI , . . . , An. Then a = \{x(A0),...,x(An)}\ is a simplex in N(A), and by Lemma 2.3.2, K(X) € a. There exists W G W such that AQ U AI U • • • U An C
W.
276
4. BASIC ANR THEORY
Then x G W and since f(x(Ai)}
G A{ for every 0 < i < n, we have
/[<jn|AT(.A)( 0 >|] C VK By Claim 1 there exists V G V such that g[a] U PF C I7. Since K(X) G cr we have 0 («(#)) € V and since W C V we have x 6 V. So {x,g(K,(x))} CV.Q By the special choice of V we now conclude that the indetity function lx and g o K are U-homotopic, i.e., P U-dominates X. D
Exercises for §4.1. Let X be a space. An equiconnecting function on X is a continuous function A : X x X x I —>• X such that A(x, x, i) = x for every £ £ I. The spaces X and Y have the same homotopy type if there are continuous functions /: X —> Y and g: Y —> X such that / o g is homotopic to the identity function on Y and g o f is homotopic to the identity function on X. 1. Present a different proof of Lemma 4.1.4 than the one in the text by using Corollary A.7.6. 2. Let r: X —>• Y be a retraction. Prove that X dominates Y. 3. Let X be an A N R with dimX < n. Prove that for every open cover II of X there exists a polytope P such that P U-dominates X and dim P < n. 4. Let X be a compact space with divaX — n < oo. Prove that there is a finite open cover U of X such that for every space Y that U-dominates X we have dimY > n. ^•5. Let X be a compact ANR. Prove that for every e > 0 there exists <5 > 0 such that for every space Y and for every closed subspace A of Y and for all maps /, g: A —> X with £(/, g) < (5, if / has a continuous extension J:Y ->X then g has a continuous extension
g:Y ^X such that g(f,g)
< e.
6. Let X be an AR. Prove that X has an equiconnecting function. 7. Show that if (X, Y) is an A(N)R-pair then both X and Y are A(N)R's. In addition, prove that X and Y have the same homotopy type. 8. Let (X, Y) be an ANR-pair. Assume that for some closed subspace A C X there is a retraction r: X —>• A having the additional property that
r[Y] C Y n A. Prove that (A, Y n A) is an ANR-pair. 9. Let X be a space with subspace Y. Suppose that there is a deformation of X through Y. Let Z be a space with closed subspace K and let e: Z —> I be continuous such that £ -1 (0) = X. Prove that for every continuous function /: Z —> X there is a continuous function g: Z —> X such that
4.2. A CHARACTERIZATION OF ANR'S AND AR'S
277
(1) g ( f ( z ) , g ( z ) ) < e ( z ) for every z £ Z, (2) g\K = f\K, (3) g[Z \ K] C Y.
4.2. A characterization of ANR's and AR's The aim of this section is to present purely topological characterizations of the classes of all ANR's and AR's. We will demonstrate the strength of our characterization by showing that the hyperspace of each Peano continuum, i.e., a locally connected continuum, is an AR. Here is the announced characterization, stated in terms of ANR-pairs. Theorem 4.2.1. Let X be a space and let Y C X be dense. The following statements are equivalent: (1) (X,Y) is an ANR-pair, (2) for every open cover U of X there exists an open refinement VoflL such that for every locally finite simplicial complex 7 and every subcomplex S of T containing all the vertices of T, every partial realization of T in Y relative to (§,V \ Y) can be extended to a full realization of 7 in Y relative toll \Y. Observe that by Corollary 4.1.11 we only need to verify the implication (2) => (1). This will be done by proving a series of lemmas. So let (X, Y) be a pair of spaces having the realization property stated in Theorem 4.2.1(2). By Corollary 1.1.8 we may assume that X is a closed subspace of a normed linear space L. We shall prove that there are a neighborhood U of X in L and a retraction r: U —> X such that r[U \ X] C Y. We first show that this suffices. Since we will not use anything specific about L, we can replace L by the normed space L' = L x M. We identify X and X x { 0 } i n L x I . By our claim, there are a neighborhood U of X in L' and a retraction r: U —> X such that r[U \ X] C Y. This can be used to show that (X, Y) is an ANR-pair, as follows. Let E be a space wdth closed subspace F, and let
be continuous. Since X is an ANR, being a neighborhood retract of L (Theorem 1.2.9), there are a neighborhood V of F in E such that / can be extended to a continuous function g: V —>• X. Since U is a neighborhood of X in I/, the set Ur\ (X x I) is a neighborhood of X x {0} in X x I. There consequently is a continuous function a : X -» (0, 1] such that {(x,t) :0
278
4. BASIC ANR THEORY
(Lemma 4.1.4). Let g be an admissible metric on E which is bounded by 1 (Exercise A. 1.6), and define £: V ->• X x I by t(x) =
(g(x),Q(x,F)-a(g(x))).
Then £[V] CU,£\F = fand£[V\F}CU\(Xx the function /: V -> X by
{0}). Now finally define
f ( x ) = (r°$(x). Then / extends / and by the special property of r, f[V \F] CY. So we are done. For a space X we will let rX denote the family of all open subsets of X. Lemma 4.2.2. Let X be a space with subspace Y. The function a: rY -+rX
defined by a(A) = {x <= X : g(x,A) < g(x,Y\A)} (where, by convention, g(x, 0) = ooj has the following properties: (1) CT(0) = 0, a(Y) = X, (2) a (A] n Y = A for every A 6 rY, (3) if A,B G rY then ACS iff a (A) C a(B], (4) if A, B e r r theno-(AnS) = (7(^)0(7(5).
Proof. Let K: pi" —> pX be as in Lemma A.8.1. Then for every U G rl", cr([7) = X \ / c ( y \ C 7 ) . An easy check shows that a is as required.
D
It will be convenient to fix some notation. Let a: rX —> rL be the function of Lemma 4.2.2. In addition, by applying our assumptions on X and Lemma 4.1.10, by induction on n > 0 it is a triviality to construct two sequences of open (= open in X) covers An and !Bn of X with the following properties: (i) Ao = {X}, (ii) *Bn < An and for every locally finite simplicial complex T and every subcomplex S of T containing all the vertices of T, every partial realization of T in Y relative to (S, Sn \ Y) can be extended to a full realization of T in Y relative to An \ Y, (ni) for n > 1, An is a star-refinement of *B n _i, (iv) mesh(.An) < l/n.
Finally, by Lemma 1.2.1 and the fact that Y is dense in X there exists a countable open cover U of L \ X and a sequence of points (au)ueu m Y with the following properties:
4.2. A CHARACTERIZATION OF ANR'S AND AR'S
279
(v) if p G U € U then g(p,au) < 2g(p,X), (vi) if Un G V for every n and lim
n—>-oo
then lim diam(C/n) = 0.
n—>-oo
By an appeal to Theorem 2.3.5(1) we find that there exists a countable open cover V of L \ X such that (vii) V is star-finite and refines U. So without loss of generality we may assume that U is star-finite. We also assume that every U e U is nonempty. Lemma 4.2.3. Let x 6 X and let E be a neighborhood of x in L. There exists a neighborhood F of x in L such that if U Eli meets F then U C E. Proof. Let E > 0 be such that B(x,e) C E. If the lemma is not true then for every n there exists Un E U such that
(1) un n B(x, */„) ^ 0, (2) Un\E^$. By (1), linin-^oo g(Un,X) — 0 so by (vi), lim.^-^ diam(t/ n ) = 0. This easily contradicts (1) and (2). D This simple result is used in the proof of the following lemma. Lemma 4.2.4. For each n > 0 there exists an open neighborhood Hn of X in L such that H0 = L and for n > I ,
(!)„ E_n C (2) n Hn C Hn-i, _ (3)n if U Eli meets Hn then there exists A 6 An such that
Proof. We shall construct the Hn's by induction on n > 0. Let n > 1 and suppose that Hn-i has been constructed. By Corollary A. 4. 3 there exists a neighborhood V of X in L such that For each x G X there exists Ax € .An such that x E Ax. Then cr(A x ) is a neighborhood of x in L and consequently by Lemma 4.2.3 there exists a neighborhood Bx of x in L such that if U € U intersects Bx then U is contained in a(Ax) D F. Put #n = |J B x .
280
4. BASIC ANR THEORY
We claim that Hn is as required. Observe that (l) n and (2) n are trivially satisfied. For the verification of (3) ra , assume that U G U meets Hn. Since U is open, U D #n 7^ 0 and so there exists x such that UC\BX / 0. From this it follows that U C a(Ax}r\V . Since V C Hn-i, we are done. D Fix U G U for a moment and put
Since #o 7^ 0, E 7^ 0. We claim that .E is finite. Striving for a contradiction, assume that E is infinite. Then by Lemma 4.2.4(1), g(U,X) = 0 from which it follows by (vi) that diam(f/) = 0. Since U is nonempty, we find that U consists of a single point which belongs to X since g(U, X) = 0; contradiction. Now for every U G U define n(U) = max{n > 0 : U D Hn ^ 0}. Let K: C \ X ->• |AT(U)| be the ^-function of the cover U, cf. §1.1. For every U G U select a point z\j G C/. By Lemma 4.2.4(3) n (j/) we may pick an element At/ 6 -An(u) such that C7 C cr(Au). Lemma 4.2.5. For each U € U there exists a point j/[/ G A[/ n Y such that
Proof. Let U G U. If aj/ G A{/ then we put ?/[/ = aj/ and are done by (v) and the fact that a\j G Y. So assume that au 0 -4t/- Since zv G Lr C cr(Afy) it follows by the definition of a that
Since au ^ AU, (v) consequently implies that Q(ZU,AU) < g(zu,au) < 2g(zu,X). So any point in AU fl F will do.
D
Recall that by convention the vertex of N(li) corresponding to U G U is denoted by x(U), cf. §2.3. Now we define $: |A^(U)^| -> F as follows:
Since 7V(U)(°)| is a discrete topological space (Corollary 2.1.15) it follows that $ is continuous. We now aim at extending $ over a large part of |JV(U)|. For each m > 0 put Bm — Hm \ HTO-I-I,
4.2. A CHARACTERIZATION OF ANR'S AND AR'S
281
and let
Clearly Tm is a subcomplex of N(IL). Lemma 4.2.6. If\m-n\>2 then \3>n n \fm\ = 0. Proof. This follows directly from Lemma 4.2.4(3).
D
Lemma 4.2.7. For each m > 0, <1> \ |OP m )^| is a partial realization of Tm in Y relative to (CP m )(°>,S m _i \Y). Proof. Let \{x(U^0)), . . . ,x(U^k))}\ be a simplex in 3>m. Then by the definition of lPm, for each 0 < j < k, U^ n Hm ^ 0, so that m < n(Ui^). Consequently, (ii) and (iii) imply that for each 0 < j < k,
so that by Lemma 4.2.4(3) and the fact that a is monotone there exists an element A, 6 Am such that Observe that by the definition of $ this implies that for every 0 < j < k, *(*(Ui(j)) eA^., CAj. Since |{x(t/ i(0 )), . . . ,x(C/ i(fc) )}| is a simplex in N(U), Ui(0) H • - • n C7i(fc) / 0, so that from which it follows by Lemma 4.2.2(1), (4) that A0 n • • • n Ak ^ 0.
Since by (iii) Am is a star-refinement of S m _i, there exists an element of "Bm-i which contains all Aj and hence all ^(x(Ui^. D By (ii) we may extend for every m > 1 the partial realization $ of CP2m in Y relative to ((J^m)^, ^m-i f F) to a full realization *2m: |T 2 m| ~ ^ y ,
relative to A^m-i \ Y. Now for m > 0 consider the subcomplexes O^m+i of A^(U). Observe that for every m the complex l^m+i meets its neighbors |J"2m| and |T2 m +2| only (Lemma 4.2.6). Define ^2ml =
^2ml
( 0 )
U T 2 m 1 H T2m
U ?2ml H T
2 m 2
.
282
4. BASIC ANR THEORY
Clearly, T2m_|_i is a subcomplex of O^m+i? containing every vertex of Define a function $2m+i : |^2m+i| —> Y by (X G |0> 2m
Since the functions *2m and * 2m+2 extend $ f|(T 2 m) ( 0 ) and respectively, and since by Lemma 2.1.14(1) the sets
are closed in |JV(U)|, the function $ 2m +i is well-defined and continuous. Lemma 4.2.8. For each m > 1, $ 2m+ i is a partial realization of 92m+i in relative to (T 2m+ i,!B 2m _ 2 Proof. Let m > 1 and let T G 0 5 2m _ ( _i. By Lemma 4.2.7 there exists an element B 6 !B2m such that
Since by (ii)
Let r' be a face of r such that T' 6 O^m+i n ? 2m . By construction there exists A,-/ G ^L2m-i such that
Observe that A n A T / ^ 0. Now let r" be a face of r such that T" G CP2m+i n J ) 2m+2 . By construction there also exists AT" G ^I2m+i such that
As above, A H A T // ^ 0. Now since .A2m+i < ^,2m, «^2m < ^- 2 m-i, and by (iii), yi 2m _i is a star-refinement of !B2m_2, we are done. D So by Lemma 4.2.8 and (ii), we may extend for each m > 1 the function 3>2m+i to a full realization
relative to the covering ^I 2m _ 2 \ Y. Lemma 4.2.9. For each n > 0, K,[Bn] C |?n .
4.2. A CHARACTERIZATION
283
OF ANR'S AND AR'S
Proof. Take an arbitrary x £ Bn = Hn \ Hn+i. By (vii) there exist finitely many elements of U that contain x only, say t/i(o), • • • , U^)- The simplex T =
\{x(Ui(0)),...,x(Ui(k))}\
is an element of CPn. By Lemma 2.3.2, K(X) belongs to T, and since r is a D subset of \yn\ we are done. For each n > 0 let Pn =
|J
Define * : P2 -> Y by *(ar) = * m (ar)
(x € 3 > m | ; m > 2 ) .
Then * is continuous by Lemma 2.1.19. Notice that for every m > 3, ^ \ |Pm| is a full realization of relative to Am-3 \ Y.
n
Finally, define r : H<2 —>• X by
Then r is well-defined by Lemma 4.2.9. We claim that r is a retraction. It is clear that r[H-2 \ X] C Y. Since by Theorem 2.3.3, K is continuous and ^Y is closed in L, it remains to check the continuity of r at the points of the boundary Fr.Y of X.
Lemma 4.2.10. r is continuous at every point ofFrX. Proof. Let x G FT X and let (tn)n be a sequence in H2 \ X such that limn_>oo tn — x. We shall prove that limn^00 r(tn) = x. To this end, for every n pick Un 6 U containing tn. In two steps we shall prove that lim r(tn) = lim
Claim 1. limn_j.oo Proof. Since limn-^
g(Un,X) = 0, by (vi) it follows that lim diam(f/ n ) = 0.
n—>-oo
From this we conclude that lim z\jn = x.
n—too
By Lemma 4.2.5 we therefore obtain lim g(zun,yun) < 2 lim g(zUn,X) — 0,
284
4. BASIC ANR THEORY
so that by the definition of \I>, lim fy(x(Un)} = lim yu = lim zu = x,
n—too
n—*oo
' ra— >oo
as desired.
As in §2.3 for every y e ~H2 \ X let F(y) = {U € U : y G U} and let r(y) be the simplex of -/V(U) spanned by the (finitely many) vertices {*(*/) : U € F(y)}. By Lemma 2.3.2, r(y) is the carrier of K,(y) in JV(U), so in particular,
«(2/) € r(y). Claim 2. l i m n ^ 0 0 ^ * « t n , * x C / n
=0.
Proof. For every n E N let fc(n) e N be such that t n e Bk(n} (without loss of generality, for every n, k(n) > 3). By Lemma 4.2.4(2) and the fact that the sequence (tn)n converges to x e X, it follows easily that lim^-^ k(n) — oo. Observe that for every n, r(tn) € fk(n)- By construction, for every n there exists An G -A,k(n)-3 such that *[r(t n )] C An, and since {x(f/ n ),«(f n )} C r(tn) we therefore obtain £7
*«*
C*rt
CA .
From limn^.oo fc(n) = oo we find by (iv) that lim diam(A n ) = 0,
n—>oo
so we are done.
<0>
This completes the proof.
D
Remark 4.2.11. Notice that if U in the above proof has order at most n + l, where n > 0, then the dimension of N(W) is at most < n+l. This follows from the Countable Closed Sum Theorem 3.2.8 and the obvious fact that N(U.) has no simplexes of dimension greater than n + 1 . This will be used in the proof of Theorem 4.2.30. Corollary 4.2.12. Let X be a space. The following statements are equivalent: (1) X is an A N R , (2) for every open cover U of X there exists an open refinement V of U such that for every locally finite simplicial complex T and every subcomplex § of T containing all the vertices of 7, every partial realization of 7 in X relative to (§, V) can be extended to a full realization of 7 in X relative to U.
285
4.2. A CHARACTERIZATION OF ANR'S AND AR'S
We shall now derive some applications. Call a space X homotopically trivial if for every n > 0, every continuous function /: §n —>• X can be continuously extended over Bn+l. So every AR is homopically trivial for obvious reasons. Notice that a space X is path-connected iff every function /: §° —>• X can be continuously extended over Bl. So a homotopically trivial space is one which is 'path-connected in every dimension'. Lemma 4.2.13. Every contractible space is homotopically trivial. Proof. Let X be a contractible space and let H: X x I —> X be a homotopy contracting X to a single point, say p. In addition, let n > 0 and let g be a continuous function from Sn to X. Define g: Bn+1 —> X by g(x)=H(g(*/M),l-\\x\
("A Uj,
An easy check shows that g is the required continuous extension of g.
D
There exist elementary examples of homotopically trivial spaces that are not contractible. A well-known example of such a space is the so-called Warsaw circle in the figure below. See Exercise 4.2.1.
Figure 12. We now present the following consequence of Theorem 4.2.1. Theorem 4.2.14. Let X be a locally connected space with dense subspace Y If X has an (open) base !B such that for every finite subfamily 3 ofB: if p| y is nonempty then every component L o property that L n Y is homotopically trivial. Then (X,Y) is an ANR-pair.
has the
286
4. BASIC ANR THEORY
Proof. Let it be an open cover of X. Since 33 is an open base for the topology of X, there exists a refinement V of U consisting entirely of elements of f£>. By Lemma 4.1.10 there exists a star-refinement W of V. Since a refinement of W is also a star-refinement of V, and since by local connectivity of X components of open subsets are open (Exercise A. 2. 8), we may assume without loss of generality that W consists of components of elements of "B. We aim at applying Theorem 4.2.1. To this end, let T be a locally finite simplicial complex, let 8 be a subcomplex of T containing all the vertices of T, and let /: |S| ->• Y be a partial realization of T in Y relative to (S, W \ Y). For each a e T pick Wff € W such that f[cr n |S|] C Wff. We first want to extend / to a continuous function g: \7^ U S| —>• Y such that g is a partial realization of 1 in Y relative to (T^1) U 8, V \ Y). As to be expected, we shall construct the function g 'simplex- wise'. To this end, take an arbitrary We distinguish between two subcases. If a € 8 then / is already defined on all of a: put gff — j \ a. If a £ 8 then a e T^1) \ T^°). Since § contains all the vertices of T, crn|8| = da. So by the fact that Wo-flY is path-connected, there exists an extension ga : a —>• Wff H Y of / \ da. Now define g : \7^ U 8| —> Y as follows: g(x)=ga(x) (xea 6T ( 1 ) US). Since T^ 1 ) U 8 is a simplicial complex, g is well-defined. Also, g is continuous by Lemma 2.1.19. We claim that g is a partial realization of T in Y relative to (T^1) U S, V \ Y). To this end, take an arbitrary simplex a 6 T. First observe that By construction it therefore follows that g[a n |T(1) U S|] C ( i ^ r : r is a one-dimensional face of a} U Wff. Again since S contains all the vertices of T, each of the W T 's intersects Wa. Since W is a star- refinement of V, there consequently exists Va e V such that which is as required. Now let a 6 T be an arbitrary simplex and put £(0-) = {T e T : ( j ^ r } . Since by the local finiteness of T each vertex of a is contained in at most finitely many simplexes in T, it is clear that £(cr) is finite. Define
F(a)=
H VT. r(E£(o-)
4.2. A CHARACTERIZATION OF ANR'S AND AR'S
287
Claim 1. For each a 6 T, every component of F(a] n Y is homotopically trivial and if a ^ T then F(a) C F(r). Moreover, if a € T(1) U S then g[o-]CF(o-). Proof. That every component of F(a) is homotopically trivial follows by the assumptions on !B and the fact that the £(cr)'s are finite. Also, if a ^ r then
£(r) C £( CT ) from which it follows that F(cr) C F(r). Now take an arbitrary a E T^1) U § and T e £(cr). Then
We conclude that #[
0
Now by induction on n > I we shall construct continuous functions / n : |o-< n >us|->r such that the following conditions are satisfied: (1) f i = 9 and /n+1 extends /n, (2) for every a e 7^ U S, /n[a] C F(a}. Assume that for certain n > 2, / n _i has been defined. We shall construct fn 'simplex- wise'. To this end, take an arbitrary a G 7^ U S . l f c r e T^"1) U S then / n _i is already defined on all of a: put fa = fn-i \ a. If a £ T(n-1) U 8 then a € 7^\7^n-^. Observe that
Since n > 2 and cr 0 T^" ), a is at least two-dimensional. Consequently, da is connected from which it follows that fn_i[da] is connected. By Claim 1 it follows that fn-i[da] is contained in a homotopically trivial subset of the intersection F(a] n Y. The function / n _i f 5cr can therefore be extended to a continuous function fff:
288
4. BASIC ANR THEORY
As to be expected, define h: |T| —>• Y by h(x)=fn(x)
(xe|T
It is clear that by (1), h is well-defined. By another appeal to Lemma 2.1.19 it follows that h is continuous. By construction, h extends / and by (2), it is a full realization of T in Y relative toV\Y and therefore to U \ Y. D Corollary 4.2.15. Let C be a convex subspace of a locally convex linear space. IfCCECC then (E, C) is an AR-pair. Proof. Since convex sets in a linear space are contractible and hence homotopically trivial, and the linear space under consideration is locally convex, and the intersection of an arbitrary family of convex sets is again convex, this follows immediately from Theorem 4.2.14. D Remark 4.2.16. Since the intersection of an arbitrary family of convex sets in a normed linear space is again convex, it seems that Corollary 4.2.15 gives a new proof of Theorem 1.2.9. This is not true however, since Theorem 1.2.9 was used in the proof of Theorem 4.2.1. Corollary 4.2.17. Let 0 < n < oo and let A be an arbitrary subset of§>n. Then Bn+l \ A is an AR. Corollary 4.2.18. Let X be a locally connected space. If X has an (open) base 3 such that for every finite subfamily J of "B: if p| y is nonempty then every component C of P) IF is homotopically trivial. Then X is an ANR. Naturally, Theorem 4.2.1 suggests the question whether there exists a characterization of the class of AR's in the same spirit. In Corollary 1.4.5 we showed that a space X is an AR if and only if X is a contractible ANR. So it seems that our question has already been answered. However, deciding whether a given space is contractible is sometimes quite a complicated task, cf. Corollary 2.4.11. Within the class of ANR's it turns out that contractibility and homotopy triviality are equivalent notions. Since for a given space X, verifying that it is homotopically trivial is usually much easier than verifying that it is contractible, we shall present a proof of our assertion in full detail. We first derive the following lemma. Lemma 4.2.19. Let X be a space. The following statements are equivalent: (1) X is homotopically trivial, (2) for every locally finite simplicial complex T and every subcomplex 8 of T, every continuous function g: |S| —> X can be extended to a continuous function g: |T| —>• X.
4.2. A CHARACTERIZATION OF ANR'S AND AR'S
289
Proof. The implication (2) =>• (1) is a triviality. Simply observe that there is a homeomorphism from Bn onto an n-dimensional simplex a taking §n-1 onto the geometric boundary da of a (Exercise 1.1.24). The implication (1) =>• (2) is also very simple to prove. Let us first show that without loss of generality we may assume that § contains all the vertices of T. Since |T^°^| is a discrete topological space by Corollary 2.1.15, and since both |S| and |T(°)| are closed in |T| by Lemma 2.1.14(1), we can extend g over |S U T(°)| in an arbitrary way. Now, by induction on n > 0 by a standard procedure we shall construct a continuous function gn: |S U 7^\ —>• X such that go = g,
gn+i extends gn.
Since S contains all the vertices of T, we may indeed put go = g. Assume that the function gn-i has been constructed for certain n > 1. As previously in this section, we shall construct gn 'simplex-wise'. To this end, take an arbitrary a€SUT(n). If a 6 S U T^71"1) put ga = gn-i \ a — 9 \ &• So assume that tre7(n)\(S\J7(n-V). Now we use our assumption on X to extend the function gn-i \da: da —> X to a continuous function ga: a —>• X. Define gn : S U 7^\ —I X by 9n(x}=gff(x}
OrEseSuT^).
As in the previous proofs in this section it follows that gn is well-defined and continuous. This completes the inductive construction of the functions gn for n 6 N. Now define g: \7\ —> X as follows: g(x}=gn(x]
(xe
Then g is clearly as required.
D
We shall now present a proof of the following Theorem 4.2.20. Let X be a space. The following statements are equivalent: (1) X is an AR, (2) X is a contractible A N R , (3) X is a homotopically trivial ANR.
Proof. Observe that the implications (1) =>• (2) => (3) are trivialities, see Corollary 1.4.5 and Lemma 4.2.13. So let X be a homotopically trivial ANR. We shall prove that X is an AR. Observe that X has the realization property stated in Corollary 4.2.1(2).
290
4. BASIC ANR THEORY
It will be convenient to think of X as the space X in the proof of Theorem 4.2.1. We adopt all notation and terminology introduced in that proof. We shall prove that the retraction r: H^, —> X can be extended over L so that the desired result follows from Corollary 1.2.9 and the obvious fact that a retract of an A R is an AR. Recall that = |J m=2
Define
Q = |iPo|u|a> 1 |u|iP2|. Observe that by Lemma 4.2.6,
P 2 ng = |3>2|, and also that by definition, P2 U Q — PQ. Now since T = y0 U CPi U CP2 is a simplicial complex, and 7 2 is a subcomplex of T, by Lemma 4.2.19 we can extend ^ f l^l to a continuous function
e-.Q->x. By Lemma 4.2.9 it follows that K\C \ H?\ C Q. Consequently, the function f: C ^ X
defined by ~( } rx
=
0OK)(x)
(cec\ff2),
is well-defined if for every x e D = H2\H2 we have r(x) = (^OK)(X). To verify this, take an arbitrary x £ D. Then x E B%. Consequently, Lemma 4.2.9 gives us that K(X) 6 [O^l- By the definition of 9 we therefore have
r(x) = (# o K)(X] = (0o K)(X], which is as required. Finally, f is continuous since the sets H? and C\H<2 are closed in C and the restrictions f \ H^ and f \ C \ H-2 are both continuous. D Corollary 4.2.21. Let (X, F) be an ANR-pair. If Y is homotopically trivial then (X,Y) is an AR-pair. Proof. By the previous theorem it suffices to prove that X is homotopically trivial. By Theorem 4.1.6 there is a deformation H : X x I —>• X through Y. Let /: Sn —>• ^f be continuous, where n > 0. For sufficiently small t > 0, the functions / and Ht ° / are as close as we please, so they are homotopic by Theorem 4.1.1 since X is an ANR. Since Y is homotopically trivial, the function Ht o / can be extended over Bn+l . So the functions Ht o / and / are homotopic functions from S n into X, one of which can be extended over Bn+l . But then the other one can also be extended by the Borsuk Homotopy Extension Theorem 1.4.2. D
4.2. A CHARACTERIZATION OF ANR'S AND AR'S
291
Application to hyperspaces. As in §1.11, for every compact space X we let 2X denote its hyperspace. As announced above, we shall apply the results in this section to prove that if X is a Peano continuum then 2X is an AR. This demonstrates the power of the techniques derived so far. We shall conclude from this by using TORUNCZYK'S Theorem from [391] that the hyperspace of each non-degenerate Peano continuum is homeomorphic to the Hilbert cube; this is the Curtis-Schori-West Hyperspace Theorem. (See also [298, Chapter 8] for details.) Since the Hilbert cube is an AR, at first glance there seems no reason for proving the AR-property for hyperspaces. However, that the hyperspace of each Peano continuum is an A R is a fundamental step in the proof of the Curtis-Schori-West Hyperspace Theorem. Lemma 4.2.22. For each n G N, there exists a map r : Bn+1 -> 5s (§n) such that for every x G S n , r(x] — {x}. Proof. First consider the case n = 1. Parametrize S1 as [0, 2?r], with 0 and 2?r identified. For 9 G [0, 27r], set to — K — \9 — TT|. Define a homotopy H: § x x [0,7r] ->> ^(S1)
by (te < t < TT).
Observe that Ho(9) — {0} for every 9 G S1. In addition, define a homotopy K: § x x [0,7r]-> Js^1) as follows: K(9,t) = {t,2K-t}\J ({0-te,0 + to}r\[t,2n-t]). It is easy to see that the homotopy H followed by K provides a homotopy from S1 into 3s (S1) connecting the inclusion map 9 -> {9} and the constant map 9 ->• {TT}. Now we proceed by induction on n. Assume that for certain n > 2 there exists a map /: Bn -»• ^(S71"1) such that f ( x ) = {x} for every x G S71"1. We identify Bn+l and the two-point compactification of Bn x (— 1, 1). Define a map g: Bn+1 -> J3(§n) as follows: ' g ( x , t ) = {(p,t) :pe f ( x } } ^(+00) = {+00}, -oo) = {-oo},
( ( x , t ) eBnx (-1,1)),
An easy check shows that g is as required.
D
A subcollection £ of3'00(X) is called an expansion hyperspace of X if for every E G £ and F G ^^(X) such that E C F we have F G £. Observe that ^oo(X) is an expansion hyperspace.
292
4. BASIC ANR THEORY
Dense expansion hyperspaces are AR's, as will be shown now. Theorem 4.2.23. Let X be a compact space. The following statements are equivalent: (1) (2) (3) (4)
X is a Peano continuum, 2X is a Peano continuum, 2X is an AR. If £ C 2X is a dense expansion hyperspace then (2 X , £) is an AR-pair.
Proof. We already know that the equivalence (1) <3> (2) is true. See Propositions 1.11.4 and 1.11.13. Since (4) =$• (3) => (2) are trivial (Exercise 1.2.6), it remains to verify the implication (1) & (2) =^ (4). Naturally, we aim at applying Theorem 4.2.14 to prove that (2X , £) is an AR-pair. Claim 1. £ is path-connected and locally path-connected. Proof. Let F 6 £, say F = {xi,...,x n }, where x% ^ Xj if i / j. If £7 is a neighborhood of F in 2X then there is a smaller neighborhood of the form {{Vi, . . . ,yn}}, where Vi is an open neighborhood of Xi for every i (Lemma 1.11.11(2)). We may assume without loss of generality that
Vi n Vj = 0 if i ^ j, and that each Vl is connected since X is locally connected. We claim that A = {{Vi, . . . , Vn}) n £ is path-connected, which will prove that £ is locally path-connected. This will implicitly also show that £ is pathconnected since £ = ({X}} n £ and X is connected. To this end, let G 6 A be arbitrarily chosen. For every i < n let Gi = G n Vi. Then every Gi ^ 0 while moreover G = \J^=1 Gi. Let us concentrate on GI = {ai, . . . , aTO} for a moment. By Theorem 1.5.22, Vi is path-connected. So there exists for every j < m a path Aj : I —> V\ such that Aj(0) = Q.J and Aj(l) = x\. The function fii : I -> {{Vi}) n £ defined by Hi(t) = {xl}\J{Xj(t)
:j<m}
is a path connecting {xi} U GI and {xi} (Corollary 1.11.10). Define the functions [ij for 2 < j < n similarly. Then the function // : I —> A defined by
is a path connecting F and F U G. Using the same technique there exists a path in A connecting FUG and G. Indeed, for j < m define Vi : I -> ({V\ }}n£ by
4.2. A CHARACTERIZATION OF ANR'S AND AR'S
293
Define the functions Vj for 2 < j < n similarly. Then the function v : I —>• A defined by
is a path connecting .F U G and G. The join of these paths /i and v is therefore the required path.
<0>
So we see in particular that £ is locally connected. Hence components of open sets of £ are open in £ (Exercise A. 2. 8). We aim at applying Theorem 4.2.14 to prove that (2 X , £) is an AR-pair. Consider the base 'B(X) for 2X identified in Lemma 1.11.11(2). Since T>(X) is closed under finite intersections (Lemma 1.11.11(2)), it suffices to prove that every component of an element of 'B(X) \ £ is homotopically trivial. To this end, let £/i, . . . , Un C X be nonempty and open, and consider a component C of B = {{t/i, . • . , Un}) n £. Observe that B ^ 0 since £ is dense, hence C ^ 0. Claim 2. C is homotopically trivial. Proof. We already know by Claim 1 that C is open in £. By connectivity of C and the fact that £ is locally path-connected, it therefore follows by Lemma A. 10. 8 that C is path-connected. So for the verification that C is homotopically trivial it suffices to consider a continuous function g : §m —>• C, where m > 1. By Lemma 4.2.22 there exists a map r: Bm+1 -> jF3(STO) such that for every x G S m , r(#) = {x}. By Corollary 1.11.8, the function g: Bm+l ~^2X defined by
g(p) = \Jg[r(p)] is continuous. It is clear that the range of g is contained in £ and that
g\$m = gWe claim that g[Bm+l] C B. To this end, take an arbitrary;? € Bm+l . Notice that g[r(p}} consists of at most three elements of B. So each of these elements meets every Ui and is contained in their union. But then clearly |J^[r(p)] has the same property. By connectivity of Bm+l it now follows that g[Bm+l] C C, which is as required.
<0>
By Theorem 4.2.14 it therefore follows that (2X , £) is an ANR-pair. However, implicitly, we also proved that £ is homotopically trivial itself since clearly £ = {{A"}} fi £ and £ is connected by Claim 1. So we conclude that (2X , £) is an AR-pair by Corollary 4.2.21. D
294
4. BASIC ANR THEORY
Corollary 4.2.24. Let X be a Peano continuum, and £ C 1X a dense expansion hyperspace. Then there is a deformation of2x through £. Proof. Apply the previous theorem and Theorem 4.1.6.
D
Observe that since 3~oo(X) is an expansion hyperspace, this result shows that we can approximate compacta in X by finite sets in a continuous way. This curious result will turn out to be of crucial importance later. Toruiiczyk's Theorem. Let X be a space. We say that X has the disjoint- cells property provided that for every n G N, every continuous function /: I n x {0,1} —> X is approximable (arbitrarily closely) by maps sending I n x {0} and I n x {1} to disjoint sets. Formally, for every n, every continuous function / : E n x {0, 1} —> X and every £ > 0 there is a continuous function g: In x {0, 1} such that (1) Q ( f , g ) < e , (2) 0prx{0}]n0pL n x{l}] = 0. It is clear that Q has this property. To see this, let e > 0 and a continuous function /: I n x {0, 1} —>• Q be given. There exists m e N such that 2-< m - 1 ) < e. Define g: I n x {0, 1} -»• Q as follows: 9(x,i) = (/fo * ) i > " - , / ( z , * ) m - i , M , - - . ) > (« £ {0,1})An easy check shows that g satisfies (1) and (2). Toruiiczyk's Theorem states (among other things) that Q is topologically the only AR with the disjointcells property. Toruiiczyk's Theorem 4.2.25. Let X be a space. The following statements are equivalent: (1) X is homeomorphic to the Hilbert cube Q. (2) X is a compact AR and has the disjoint-cells property. For details, see TORUNCZYK [391] or VAN MILL [298, Chapter 8]. So all there remains to prove for the Curtis-Schori-West Hyperspace Theorem is that 1X has the disjoint-cells property. Lemma 4.2.26. Let X be a Peano continuum and let "X, be a compact subset of TOO (X). Then for every e > 0 there is a continuous function
such that QH(/, 1) < £• Proof. Define £ = {F e 3"oo(X) : F is not contained in an element of OC}.
4.2. A CHARACTERIZATION OF ANR'S AND AR'S
295
Observe that £ is an expansion hyperspace of X and that £ Pi 3C = 0. We claim that £ is dense in 2X . Since yoo(X) is dense in 2X (Lemma 1.11.11(2)), it suffices to prove that for arbitrary F G J'00(X) and e > 0 there exists an element E e £ with gH(F,E] < e. So let F e Joo(^) and e > 0 be given. Pick a point x E F and a sequence (xn)n in X \ {x} converging to x, with Q(xn,x} < £ for each n. Define Observe that for each n, Fn € ^^(X) and gn(F,Fn) < e. We claim that for some n, Fn € £. If not, then there exists a sequence (Kn}n in 3C such that Fn C Kn for each n. By compactness of 3C we assume without loss of generality that Kn —»• K , n —> oo. Claim 1. (xi, £2,0:3,. . . } C ^. Proof. Suppose that for some n, xn £ K. Since Kn —> K, n —> oo, there is an ra > n such that Km C K \ {#„}. But this is impossible since xntFmCKmCK\ {xn}. So we are done.
•<)
From the claim we conclude that K is infinite, which contradicts the fact that K is finite. We see that £ is dense in 2X . By Corollary 4.2.24 there are arbitrarily small maps 2X —> £. Since £ n IK = 0, we are done. D The Curtis-Schori-West Hyperspace Theorem 4.2.27. Let X be compact space. The following statements are equivalent: (1) X is a Peano continuum, (2) 2X is homeomorphic to the Hilbert cube Q. Proof. As observed above, it suffices to prove that for a Peano continuum X, 2X has the disjoint-cells property. Since 2X admits arbitrarily small maps into ^^(X) (Corollary 4.2.24) and since for every compact subset % of jFoopsT) there are arbitrarily small maps from 2X into 2X \ % (Corollary 4.2.26), this is obvious. D Characterization of finite dimensional ANR's and AR's. We shall now characterize the class of all finite dimensional ANR's and AR's in simple topological terms. Let X be a space and let 0 < n < oo. We say that X is connected in dimension n, abbreviated C n , provided that for every 0 < ra < n, every continuous function /: STO —> X can be extended to a continuous function /: Bm+1 ->• X. So the homotopically trivial spaces are precisely those spaces that are connected in every dimension. In addition, we say that X
296
4. BASIC ANR THEORY
is locally connected in dimension n, abbreviated LC n , provided that for every x E X and for every neighborhood U of x and for every 0 < m < n there exists a neighborhood V of x such that every continuous function /: Sm —> V extends to a continuous function /: Bm+1 —>• C7. We shall prove that if dim A" = n < oo then X is an ANR if and only if X is LCn and also that A7" is an AR if and only if X is both LC n and C n . The proof of the following lemma is precisely the same as the proof of Lemma 4.2.13 and is therefore left as an exercise to the reader. Lemma 4.2.28. Let X be a space. IfX is locally contractible then X is LC n for every n. Proposition 4.2.29. Let X be a space and let 0 < n < oo. The following statements are equivalent: (1) X isLC n . (2) for every open cover U of X there exists an open refinement V of U such that for every locally finite simplicial complex T with
dim |T| < n + 1 and every subcomplex 8 of T containing all the vertices of T, every partial realization of T in X relative to (§, V) can be extended to a full realization of 7 in X relative to U. Proof. We prove (2) => (1). Take an arbitrary x G X and a neighborhood U of x. There exists an open neighborhood W of x such that W C U. Consider the open cover U = {U,X\W} of X. By assumption, for the open cover U there exists an open refinement V such as in (2). Pick V G V such that x G V. Fix an integer m < n, and let a be an arbitrary (m + l)-dimensional simplex in M m+1 . It will be convenient to identify Sm and da (Exercise 1.1.24). Now let 3~(cr) be the simplicial complex consisting of all the faces of a and let 8 be the subcomplex consisting of all proper faces. Consider a continuous function /: da —> WnV and observe that / is a partial realization of ^(cr) in X relative to (S, V). By assumption, / can be extended to a full realization /: a = |3~(0")| —> X relative to U. Consequently, there exists U' G U with f[cr] C U'. Since / extends /, U' n W + 0, i.e., U' = U. We prove (1) => (2). Let U be an open cover of X. By downward induction, we shall construct open covers U n , V n , U n _ i , V n _ i , . . . , U Q , V0 of X having the following properties:
(3) Un = U,
4.2. A CHARACTERIZATION OF ANR'S AND AR'S
297
(4) if 0 < i < n then for every V 6 Vi and for every continuous function /: Sl —>• V there exist U € Hi with V C U and a continuous extension /: Bi+l ->• C7 of /, (5) if 0 < i < n — 1 then Ui is a star-refinement of Vi+i (Observe that (4) implies Vi < Hi.) The construction of these covers is a triviality. By Lemma 4.1.10 there is never trouble with the construction of the U n 's. Let us construct the cover Vn. Take an arbitrary x € X and pick Ux G lin containing x. Since X is LCn there is an open neighborhood Vx of x such that Fx C Ux while moreover every continuous function / : S n —>• Vx can be extended to a continuous function /: Bn+l —> Ux. The cover Vn = {Vx : x G X ]
is clearly as required. The construction of the other V^'s is similar. Now put V = VQ, let T be a locally finite simplicial complex with dim |T < n + 1, let § be a subcomplex of T containing all the vertices of T, and let /: |S| —> X be a partial realization of T in X relative to (S, V). By induction, we shall construct for every 0 < i < n a continuous function
having the following properties: (6) /0 = / and for 1 < i < n, fi extends /;_i, (7) for 0 < i < n, fi is a partial realization of T in X relative to the pair Since § contains all the vertices of T, our choice /o = / is as required. Now assume that for certain 1 < i < n the function /;_i has been constructed. As in the previous proofs, we shall construct fl 'simplex- wise'. Let a be an element of T^ \ (S U1J^~1)). By (7) there exists an element V e V»_i such that fi[da] C F. So by (4) there exist U £ U z _i with V C. U and a continuous extension /CT : a -)• f/ of the function /;_i f 9cr. Now define £ : |§ U T^ | ->• A" as follows:
By Lemma 2.1.19 fi is continuous. We now claim that fi is a partial realization of T in X relative to (S U T^, V^). To this end, take an arbitrary simplex o 6 T. By (7) there exists V e V,_i such that f^i[(7r\\B\J7^-^\] C V. Let T be a face of a such that r e T^. By construction there exists U E Ui_i such that fi[r] C U. Observe that V intersects U. Consequently,
Now since V;_i < Ui-i and by (5) Ui_i is a star-refinement of Vi, we are done. This completes the inductive construction.
298
4. BASIC ANR THEORY
With the same technique it is clear that we can extend fn to a continuous function / n + 1 :|SUT< n + 1 > -+X such that for every a € T(n+1) there exists U £ Un — U such that fn+i[a]CU. n+1
However, by assumption j( tion of T in X relative to U.
) = T so that fn+i is the required full realizaD
We now come to the announced characterization. Theorem 4.2.30. Let X be a space and let 0 < n < oo. The following statements are equivalent: (1) X isLCn. (2) for every space Y and for every closed subspace AofY with dim(Y\A}
-
_ f z
f ( x )}
(zeY\ (zeY\A),
- \ /(*) (z e A).
is continuous. Proof. Let a: rA ->• rY be a function such as in Lemma 4.2.2 and let 'B be a countable open base for X. In addition, let 5" be a countable open base for Y \ A. For every B e 13 and n € N define B(ri) CC by B(n) = {y£(Y\A)n a ( f - l [ B ] ) : g(y, A) < i/n} U B.
4.2. A CHARACTERIZATION OF ANR'S AND AR'S
299
An easy check shows that the collection 3~ U {B(n) : B e 23, n G N} is as a base for the required topology on C. 0 We shall now prove that X is a neighborhood retract of C. Once this has been established, by part (5) of the claim we are done. The proof of our assertion is precisely the same as the proof of Theorem 4.2.1. Only two minor changes should be made. We shall adopt the notation and the terminology introduced there, we shall indicate the required changes and we shall leave the precise verification to the reader. By Proposition 4.2.29 we have a realization property available up to dimension n + 1. So in the construction of the covers An and 23n, one should replace 'locally finite simplicial complex T' by 'locally finite simplicial complex T with dim T| < n + 1' everywhere. The special open cover U of C \ X used in the proof of Theorem 4.2.1 should have the additional property that its order does not exceed n + 1. However, by the fact that dim(C \ X) < n + 1, this can be achieved by Theorem 3.2.5(5). This implies that the nerve N(U) has dimension at most n + 1 (Remark 4.2.11). With these two small changes, the proof of Theorem 4.2.1 can now be followed verbatim. We prove (2) => (3). Suppose that (3) is not true at the point p. Then there exists e > 0 such that for every z € N there exist a space Yi with dim Yi < n and a continuous function /;: Yi —> B(p, e/^) that is not homotopic within B(p, e) to a constant function. Define A and Y by
(j(Y, x n n i i u
^v and
r= i=l
x
Here oo ^ Z^i(^ ^)- Topologize r by requiring that X^i(-^ x ^) ^s an open subspace of Y and a basic neighborhood of oo has the form X^/c(^ x ^) for fc G N. It is easy to see that r is a separable metrizable space. Now define the function /: A —> X as follows
< f(x,V=P
(ser')i
( /(oo) = p Then / is clearly continuous. Observe that by Theorems 3.4.12 and 3.2.9 we have dim(l A \A) < n + l so that by assumption there exists a neighborhood W of A in Y such that / can be extended to a continuous function f:W—>X. Then V = f ~ l [ B ( p , e}] is a neighborhood of oo in Y and therefore contains all but finitely many of the Yi x I, i 6 N. But this implies that for all but
300
4. BASIC AMR THEORY
finitely many i £ N the function /j is homotopic within B(p,e) to a constant function. This is a contradiction. We prove (3) => (1). Since dim§ m = m for all m (Theorem 3.2.12), this is clear. D Theorem 4.2.31. Let X be a space and let 0 < n < oo. The following statements are equivalent: (1) X isLC n andCn. (2) For every space Y and for every closed subspace AofY with dim(Y\A) < n + 1, every continuous function f : A —> X can be continuously extended over Y, (3) X is LCn and for every space Y with dim Y < n, every continuous function f : Y —>• X is nullhomotopic. Proof. The proof of (1) =^ (2) is precisely the same as the proof of the implication (3) => (1) in the proof of Theorem 4.2.20 and is therefore left as an exercise to the reader. We shall now prove (2) =» (3). That X is LCn follows from Theorem 4.2.30. Let Y be a space with dimF < n and assume that /: Y —> X is continuous. Since by Theorems 3.4.12 and 3.2.8 we have dim A(Y~) < n + 1 (here A(F) denotes the cone over Y of course), by (2) we conclude that / can be extended to a continuous function /: A(F) —>• X. The contractibility of A(y) (Exercise A.12.5) implies that / is nullhomotopic. From this it follows easily that / is nullhomotopic as well. The proof of (3) => (1) is a triviality of course since dim§ m = m for every m (Theorem 3.2.12). D These results have the following corollaries. Corollary 4.2.32. Let X be a space. The following statements are equivalent: (1) X is a Peano continuum, (2) there is a continuous surjection f : I —> X. Proof. Assume first that X is a Peano continuum. By Theorem 1.5.10, there exists a continuous surjection g: C —>• X, In addition, by Theorem 1.5.22, X is LC° and C°. Consequently, an easy application of Theorem 4.2.31 yields that g can be extended to a continuous function /: I —> X. The implication (2) =>• (1) follows from Exercises A.5.5 and A.2.11. D
4.3. OPEN SUBSPACES OF ANR'S
301
We can now state our characterization theorem for finite dimensional ANR's and AR's. Theorem 4.2.33. Let X be a space, let 0 < n < oo and let dimX < n. Then (1) X is an ANR iff X is locally contractible iff X is LCn, (2) X is an AR iff X is LCn and Cn. Proof. We shall first present a proof of (1). To this end, first observe that if X is an ANR then X is locally contractible (Exercise 1.2.1) and consequently LCn by Lemma 4.2.28. Conversely, assume that X is LC n . Then X is locally contractible by Theorem 4.2.30(3). Consequently, X is LCm for every m (Lemma 4.2.28). Now let m = 2n + 1. By Theorem 3.3.5, we may assume that X is a subspace of §m. Put Z = {x 6 Bm+1 : \\x\\ < 1} U X. Then Z is an AR by Corollary 4.2.17 and X is clearly closed in Z. Since dimja: 6 Bm+l : \\x\\ < 1} < m + 1, by Theorem 4.2.30 it follows that X is a neighborhood retract of Z. Since Z is an AR, Proposition 1.2.10 implies that X is an ANR. We shall now present a proof of (2). To this end, assume that X is LC n and Cn. By Theorem 4.2.31 we conclude that X is contractible and hence Cm for every m (Lemma 4.2.13). Now define Z as in the proof for (1). By Theorem 4.2.31(2) there is a retraction r: Z —> X from which we conclude that X is an AR, as required. D Remark 4.2.34. Unfortunately, a characterization of infinite-dimensional ANR's and AR's as simple as Theorem 4.2.33 does not seem to be possible. BORSUK [70, p. 124] constructed an example of a contractible and locally contractible compactum which is not an ANR. Exercise for §4.2. 1. Give an example of a homotopically trivial compact space which is not contractible 4.3. Open subspaces of ANR's
In this section we shall prove that every open subspace of an A N R is again an ANR and that every space that admits an open cover by ANR's is an ANR as well. As an application we conclude that every polytope is an ANR. We also show that every A(N)R has an A(N)R completion. Theorem 4.3.1. Let (X, Y) be an ANR-pair and let U be an open subspace o f X . Then (U, U n Y) is an ANR-pair.
302
4. BASIC ANR THEORY
Proof. Let Z be a space, A C Z be closed, and let g : A —> U be continuous. Since (X,Y] is an ANR-pair, there exists an open neighborhood V of A in Z such that g can be extended to a continuous function / : V ->• X while moreover /(V \ A] C F. Since / is continuous and extends g, and since U is open in X, we conclude that W — f~l[U] is an open neighborhood of A. Then g — f \ W : W —>• C7 is a continuous extension of g. In addition, if z belongs to W \ A then g(z) — f(z] G U n F. So we are done. D Corollary 4.3.2. Let X be an ANR and ]et U be an open subspace of X. Then U is an ANR. Observe that [—1,0) U (0,1] is not an AR but is an open subspace of the AR JJ. So an open subspace of an AR need not be an AR. We now aim at proving that a 'local' ANR is an ANR. First we need to derive a few elementary results. In our proofs in this section we make use of cones. As we saw in Exercise A. 12. 6, if A is a (closed) subspace of a space X then A (A) can be thought of as a (closed) subspace of A(X). It will be convenient to do that. We will also use the fact that the cone of an an ANR is an AR (Theorem 1.4.6). Proposition 4.3.3. Let X be a space and assume that X can be written as U U V, where both U and V are open in X . I f U and V are ANR's then so is X. Proof. We shall prove that the cone over X is an AR so that X is an ANR by Theorem 1.4.6. By Proposition A. 7.1 there exist closed sets E and F in X such that ECU,FCV andE\jF = X. Observe that C A(C7),
A(F) C A(V),
A(E) U A(F) =
Now let Y be a space, let A C Y be closed, and let /: A —» A(X) be continuous. Put E' = f~l[A(E)] and F' = f~l[&(F)], respectively. By Lemma A. 8.1 there exist closed sets E" and F" in Y such that Since U n V is an open subspace of U, it is an ANR by Corollary 4.3.2. Consequently, Theorem 1.4.6 implies that the function / \ E' n F': E' n F' -)• A(E n F)
can be extended to a continuous function g: E" n F" —> A(t/ D V). Now define fE:E'U (E" n F") -+ A(t7) by
Then /# is obviously continuous. By another application of Theorem 1.4.6, there exists a continuous extension /i: E" —> A(£7) of JE- Similarly define fp
4.3. OPEN SUBSPACES OF ANR'S
303
and a continuous extension /2 : F" ->• A(V) of /F. Now define f:Y-> in the obvious way, namely,
Then / is the required continuous extension of /.
D
Lemma 4.3.4. Let X be a space having an open cover U by pairwise disjoint ANR's. Then X is an ANR. Proof. This is a triviality. Assume that X is a closed subspace of a space Y. By Lemma 4.2.2, for every U e It there exists an open subset V(U) of Y such that (2) the collection {V(U) : U e U} is pairwise disjoint. Note that U is closed in V(U) for every U e U. So by our assumptions on U, for every U € U there exist an open subset W(U) of V(U) which contains U and a retraction r\j\ W(U) —>• C7. Observe that the W(U)'s are open in Y since the V(C/)'s are. Now put VF = Uc/eu ^(^0- Then VF is a neighborhood of X in F and the function r: W —> X defined by r(x) = rv(x}
(x 6 W(U), U € U)
is a retraction.
D
We now come to the following Theorem 4.3.5. Let X be a space and suppose that X admits an open cover U consisting of ANR's. Then X is an ANR. Proof. Without loss of generality assume that U is countable. Enumerate U as {Un : n 6 N}. Since by Proposition 4.3.3 the union of finitely many elements of U is an ANR, we may assume without loss of generality that for every n, Un C Un+i. For every n, put Vn = {xeX:g(x,X\Un)>
l
/n}.
It is clear that each Vn is open, that Vn C V n +i, and that (J^Li Vn — X. Finally define Vn \ Vn-2 (n>3). Then X — U^Li Rn d for \m — n\ > 2, Rn D -Rm = 0. Also, ,Rn is an open subset of the ANR Un. So by Corollary 4.3.2, we conclude that Rn is an ANR. Consequently, Lemma 4.3.4 implies that an
~ LJ ^2n-!'
n=l
^= LJ
304
4. BASIC ANR THEORY
are open ANR subspaces of X. Since their union equals X, we infer by Proposition 4.3.3 that X is an ANR. D As announced in §2.1, we now get Corollary 4.3.6. Every poly tope is an A N R . Proof. Let 7 be a locally finite simplicial complex and consider P = |T|. Take an arbitrary x € |T(°)|. Since T is locally finite, the collection § = {r 6 T : x € r}
is finite. Consequently, J* = {a e 7 : (3 r 6 S)(<J ^ r)} is a finite subcomplex of T. From Theorem 2.1.24 we conclude that |?| is an ANR. Since Stx, the star of x, is contained in |T| and is open in P by Corollary 2.1.17, we conclude from Corollary 4.3.2 that Star is an ANR. Since by Corollary 2.1.17, is an open cover of P, Theorem 4.3.5 implies that P is an ANR.
D
Enlarging A(N)R's. We now show that every A ( N ) R has a completion which is also an A(N)R. Theorem 4.3.7. Let Y be a space with A ( N ) R subspace X. Then there is a subspace SofY such that X C S C X while moreover (1) 5 is a Gs-subset ofY, (2) ( S , X ) is an A(N)R-pair, Proof. It is clear that we may assume without loss of generality that X is dense in Y . We first claim that the theorem for ANR's implies the theorem for AR's. For let X be an AR. By the ANR-case of the theorem, we may assume that for some 5 with X C S C Y we have that (5, X] is an ANR-pair. Since X is homotopically trivial, being an AR, we conclude by Corollary 4.2.21 that ( S , X ) is an AR-pair. So now assume that X is an ANR. We begin the proof as the proof of Corollary 1.1.8. Let oY be a compactification of Y (Corollary A. 4. 8). By Lemma 1.1.6 we may think of aY as a linearly independent subspace of a Banach space of the form C(A) (here A is a compact space). By linear independence, if EQ is the linear hull of X then EQ D aY — X. Hence AT is a closed subspace of EQ. Since X is an A N R , there are an open neighborhood UQ of X in EQ and a retraction r: UQ —» X. Select an open subset U in C(A) such that U D EQ — UQ. Consider the closure EQ of EQ in C(A). Since X is dense in aY it follows that aY C EQ. Also,
U = UnE0 c E / n E -
4.3. OPEN SUBSPACES OF ANR'S
305
Since aY is topologically complete, being compact, there is a GVsubset T of U C\ EQ which contains UQ such that r can be extended to a continuous function f: T ->• aY (Corollary A.8.4). Observe that T is a GVsubset ofC(A) since Ur\E0 obviously is and T is a GVsubset of UC\EQ. In addition, X C T since X C UQ. Put S — T n Y. Observe that S is obviously a G^-subset of Y. In addition, f(y) — y for every y e S. This is so since X is dense in S (Exercise A. 1.9). Now put S^f-^S}. Then f f Si: Si —> 5 is a retraction. Observe that
C7n£ 0 = c/o c Si c unE0. Since (Eo,Eo) is an ANR-pair by Corollary 4.2.15 it follows that
(Ur\E0,ur\E0) is an ANR-pair as well (Theorem 4.3.1). We conclude from this that (S^UHEo) is an ANR-pair. Since f is a retraction from Si onto S such that
f[u n EQ] = r[U n EQ] = X it follows by Exercise 4.1.8 that that (S,X) is an ANR-pair.
D
Corollary 4.3.8. Let X be an A ( N ) R . Then there exists an A ( N ) R S and an imbedding i: X <-) S such that (1) i[X] is dense in S, (2) (S,i[X]) is an A(N)R-pair, (3) S is topologically complete. Proof. Let aX be a compactification of X (Corollary A.4.8). It follows from Theorem 4.3.7 that there is a G^-subset 5 of aX which contains X while moreover (S,X) is an A(N)R-pair. But then S is an A ( N ) R and moreover is topologically complete since aX is (Theorem A.6.3). D Exercises for §4.3. 1. Give an example of a space X which can be written as EUF, where both E and F are closed ANR subspaces of X, while X is not an ANR. 2. Let X be the union of two open subspaces U and V such that U, V and U n V are AR's. Prove that A' is an AR. ^•3. Let X be a space having a point x such that (1) x has arbitrarily small homotopically trivial neighborhoods, (2) A" \ {x} is an ANR. Prove that X is an ANR. (This generalizes Theorem 1.4.6.)
This page intentionally left blank
CHAPTER 5
Basic infinite-dimensional topology In this chapter we shall prove some elementary results from infinitedimensional topology. The results are elementary in the sense that no powerful apparatus is needed, but the proofs are not always easy. In infinite-dimensional topology one studies the topology of objects such as the Hilbert cube Q, the Hilbert space I2 and its infinite-dimensional linear subspaces, the countably infinite product of lines M°° and its infinitedimensional linear subspaces, and manifolds modeled on them. See the notes for more information. In this chapter we develop basic homeomorphism theory in Q. Much emphasis is on so-called absorbing systems. They are the basis for the topological classification of function spaces of low Borel complexity in the next chapter. 5.1. Z-sets Homeomorphism extension results are useful in infinite-dimensional topology. In this section we shall present a class of subsets of Q, the so-called class of Z-sets, for which an important homeomorphism extension result shall be derived in §5.3. The concept of a Z-set does not seem to be very appealing, yet it is the most central concept in infinite-dimensional topology. It turns out, although this is not immediately clear from the definition, that Z-sets are 'small'. Here smallness has a different meaning than being 'small' with respect to category or measure. It is 'small' in the sense of homotopy. This will be made precise in Exercise 5.1.7 below. Let X be a space. A closed subset A C X is called a Z-set in X provided that for every open cover U of X and every function / e C(Q,X) there is a function g € C(Q,X) such that (1) / and g are U-close, (2) g[Q] n A = 0.
A a Z-set is a countable union of Z-sets. The collection of Z-sets and aZ-sets in X are denoted by Z(X) and Z a ( X ) , respectively. 307
308
5. BASIC INFINITE-DIMENSIONAL TOPOLOGY
It will sometimes be convenient to have a 'metric' translation of the concept of a Z-set. For compact X the following lemma is a triviality since then every open cover has a Lebesgue number. For general X, the proof is only slightly more complicated. Lemma 5.1.1. Let X he a space and let A C X be closed. Then the following statements are equivalent: ( 2 ) V e > O V / e C ( g , X ) 39£C(Q,X\A): Q ( f , g ) < e . Proof. Since (1) =>• (2) is trivial, it suffices to establish the reverse implication. Let U be an open cover of X, and let /: Q —> X be continuous. By compactness of /[Q], there exists 6 > 0 with the property that every A C X with diam(A) < 6 and which moreover intersects f[Q], is contained in an element U E U (Lemma A. 5. 3). By (2) there is a function g G C(Q,X) with £>(/, g} < 6 and g[Q] n A = 0. We claim that / and g are U-close. This is a triviality. Pick an arbitrary x G Q. Then @(f(x),g(x)) < 6 and since f(x) £ /[Q], {f(x),g(x)} is contained in an element of U. D It is also possible to use finite dimensional cubes to characterize Z-sets, see Exercise 5.1.3. We shall now derive some elementary properties of Z-sets. Lemma 5.1.2. Let X be a space. Then (1) If A e Z(X) and B C A is closed in X then B e Z(X}. (2) If A G Z(X) then A has empty interior in X. (3) If (X, g) is complete, A 6 Zo-(X) and A is closed then A e Z(X); in particular, finite unions of Z-sets are again Z-sets. (4) If (X, Q) is complete and A 6 Za(X] then for every e > 0 and for every f E C(Q,X) there exists g 6 C(Q,X) such that g(f,g) < e and g[Q] n A = 0. (5) If A e Z(X) and Y is any space then A x Y e Z(X x Y}. (6) If A e Z(X) and h e ttpf) then h[A] e Z(X). (7) If A C X is closed and for every e > 0 there exists / e C(X, X \ A) such that g(f, !*•) < £ then A e Z(X). Proof. Statement (1) is trivial. For (2), (5) and (6), see Exercise 5.1.1. For (4), write A = \J^=lAn with An 6 Z(X) for every n G N. Using Lemma 5.1.1 and Exercise A. 5. 3, it is clearly possible to construct a sequence of maps fn e C(Q, X \ An), n 6 N, such that (8) e ( / i , / ) < * / 2 ,
5.1. Z-SETS
309
(9) £(/„+!, / n ) < 3 - n - * / 2 ,
(10) 0(/ n +i, /n) < 3~n •min{e(/ i [g],A i ) : 1 < i < n}. By (9) and Proposition 1.3.5, F = lim^,-^ fn exists and is an element of C(Q, X). Take x G Q arbitrarily. Then by (8) and (9) we find that e(F(x),f(x))
=KmoQ(fn(x)tf(x)) n-l
< lim
>
~~ n-»oo ^—' m=0
3~ m • e/2
- 3/2 ' £/2
From this we conclude that g(F, f) < e. It now suffices to prove that F[Q] misses A. This however is an immediate consequence of (10) and Lemma 1.6.1. Observe that (3) is a direct consequence of (4). For (7), let £ > 0 and g G C(Q,X). Find / G C(X,X \ A) such that g(f, lx) < £• Put h = f °g. Then clearly h[Q] r\A = 0, while moreover,
g(h, g) = g(f o g, g) < £(/, lx) <£ (Exercise 1.3.2). We conclude that A G Z(Q) (Lemma 5.1.1).
D
We are interested particularly in Z-sets in the Hilbert cube Q. The following lemma provides us with 'many' Z-subsets of Q. Lemma 5.1.3. Let A C Q be a closed set. Then: (1) A is a Z-set if and only if for every e > 0 there exists f G C(Q, Q) such that g(f, 1Q) < e and f[Q] n A = 0. (2) Ifirn[A] ^ Sn for infinitely many n G N, then A G Z(Q). (3) If 7rn[A] C {-1,1} for certain n G N, then A G Z(Q). Proof. For (1), let A G Z(Q). Since 1Q G C(Q,Q), the definition of a Z-set immediately gives us that for every e > 0 there exists / G C(Q,Q) with g(f, IQ) < e and f[Q] fl A = 0. The converse is a direct consequence of Lemma 5.1.2(7). For (2), choose £ > 0 and find n G N so large that 2~( n-1 ) < e while moreover ifn[A] ^ JJ^. Take an arbitrary t G JJ n \7r n [.4], and define /: Q —> Q by Then clearly f[Q] n A = 0 and g(f, 1Q) < e. So by (1), A G Z(Q). For (3), first observe that { — 1,1} G %(J). We consequently get what we want by first applying (5) and then (1) of Lemma 5.1.2. D
310
5. BASIC INFINITE-DIMENSIONAL TOPOLOGY
Remark 5.1.4. Observe that this lemma implies that each singleton subset of Q is a Z-set. Similarly, every endface of Q is a Z-set. So Q has 'many' Zsubsets, There are also spaces with 'few' Z-sets. It is easy to see for example that the collection £(J) is equal to
At first glance it seems a little surprising that all points in Q are Z-sets. However, we already know that Q is homogeneous (Theorem 1.6.6), so once one singleton Z-subset has been found it follows automatically that all singleton subsets share this property. Corollary 5.1.5. (1) B(Q) 6 Zff(Q), (2) ifKCs is compact then K e Z(Q).
Exercises for §5.1. A closed subset A of Q is called an A-set if for every open U C Q such that U is nonempty and homotopically trivial, the set U \ A is again nonempty and homotopically trivial. 1. Let X (1) (2) (3)
be a space. If A £ Z(X) If A £ Z(X) If A £ Z(X)
Prove the following statements: then A has empty interior in X. and Y is any space then A x Y € Z(X x Y). and h £ ft (A") then /i[A] £
2. Give an example of a space A" which can be covered by a countable union of Z-subsets. (This shows that in Exercise 5.1.1(2) it is not possible to replace Z-set by er Z-set.) 3. Let X be a space and let A C X be closed. Prove that the following statements are equivalent: (1) A £ Z(X), (2) V n £ N V e > 0 V / £ (7(1™, A) 3g E C(I n , X \ A) : g(f.g) < s. 4. Let n > 1 and n
A C Bn = {x € Rn : ^ Xi < l} i=l
n
be closed. Prove that A e 2,(B ) iff ^ C S""1. 5. Prove that an A-set in Q is nowhere dense. •6. Let A C Q be an A-set. Prove that for each polyhedron P, for each s > 0 and for each / £ C(P, Q) there exists 5 £ c7(P, Q) such that £(/, 5) < £ and 0 P n A = 0. Let A C
be closed. Prove that A is an A-set iff A is a Z-set.
5.2. EXTENDING HOMEOMORPHISMS IN s
311
8. For every n let Xn be a space. Assume that a closed set 00
ACX = Yl*n
has the property that 7rn[A] ^ Xn for infinitely many n. Prove that A is a Z-subset of X.
5.2. Extending homeomorphisms in s The aim of this section is to prove that if £", F C s are compact and / is a homeomorphism from E onto F which moves the points less than e then it can be extended to a homeomeomorphism of Q that satisfies the same smallness condition. An element h 6 ^K(Q) is boundary preserving if h[B(Q}] = B(Q), or, equivalently, h[s] — s. Lemma 5.2.1. For every compact subset K C s such that K ^ 0 there is a boundary preserving h 6 IK(Q) such that Tt\h\K] = {0}.
Figure 13.
Proof. Without loss of generality we may assume that
n=l
where — 1 < an < bn < 1 for n 6 N. For n > 1, let hn : Si x J n ->• Ji x J n be a homeomorphism of Ji x JJn having the following properties (see Figure 13): (1) hn does not change the Ji-coordinate of any point, (2) every horizontal line intersects hn [[ai, bi] x [an, bn]] in a set of diameter at most l/n.
312
5. BASIC INFINITE-DIMENSIONAL TOPOLOGY
It is geometrically obvious how to define hn. Let U be the region in Figure 14 below the dotted line. Similarly, V is the region above the dotted line. We will describe hn \ U only, hn \ V being entirely similar. Let t 6 JJ be fixed and let t be the vertical line through the point (t, — 1) in Ji x J n . In Figure 14 on the line i four bullets are drawn, the first one of which is x\ — (t, — 1) and the second one is x-2 = (t,an). The third and fourth bullet will be denoted by x% = (t, 6) and x± = (t,c), respectively. Here an < b < c. Now define hn \ {t} x $n by requiring that the interval from x\ to x2 is mapped linearly onto the interval from x\ to x 3 , and the interval from x 2 to x4 linearly onto the interval from xj, to x 4 . Etc. Then hn is as required. Define / : Q -»> Q by where (xi:yn) = hn(x\,xn] for every n > 1. The composition -K\ o f : Q is equal to vri and is therefore continuous. For n > 2, the composition is equal to the composition pohnO 7T{1>n},
where p: Ji x JJ n ->• J n is the projection.
Figure 14. Hence 7rn o / is continuous, being the composition of three continuous functions. By the definition of the product topology on Q it therefore follows that / is continuous. It is easy to display the continuous inverse of /. Alternatively, one can show that / is a bijection. It then suffices to apply Exercise A.5.9 to conclude that / is a homeomorphism. We shall prove here that / is surjective only. To this end, let y € Q be arbitrary. Put Xi = y\. Since hi: JJi x JJ2 -* Ji x J 2
5.2. EXTENDING HOMEOMORPHISMS IN s
313
is a homeomorphism, there exists a point (xi,a) 6 JJi x ^2 such that
Put X2 = a. Etc. Then clearly f ( x ) = y. That / is boundary preserving follows easily from the fact that all hn are boundary preserving.
Figure 15. Now think of Q as Ji x Qi and take two points in f[K] having the same second coordinates, i.e., points of the form (x,z) and ( y , z ) . If x ^ y then \x - y\ > l/n for certain n > 2. By the definition of / we have that the points (x,zn) and (y,zn) both belong to /i n [[ai,&i] x [a ra ,6 n ]]. But this contradicts (2). So we conclude that the function g: f[K] -> Qi defined by g(x,y) = y is one-to-one and by compactness is therefore an imbedding (Exercise A. 5.9). See Figure 15. Let B = gf[K] and let
be the inverse of g. In addition, let £: B —> [ai,&i] be
314
5. BASIC INFINITE-DIMENSIONAL TOPOLOGY
Since F is clearly boundary preserving, we find that the function h = F o f is as required. D Corollary 5.2.2. For every compact subset K C s and e > 0 there are an infinite TV C N and a boundary preserving homeomorphism h: Q ->• Q such that (1) the complement of N is infinite and ^n&N 2~ n < e, (2) Q(h,lQ)<e, (3) nnh[K] = {0} for every n e TV. Proof. Choose n e N so large that
Write N\ {1, 2, . . . , n — 1} as the disjoint union of infinitely many infinite sets, say C\ , C<2 , . . . . Recall that for every i
denotes the projection. Let
hi". be a homeomorphism with the following properties: (1) hi is boundary preserving (this has its obvious meaning), (2) hi^d [K] projects onto 0 in the first factor of the product (Lemma 5.2.1). Define h: Q ->• Q by
i
(J
It is clear that h is as required (observe that h is nothing but the product of the hi$ and the identity on the first n — I factors). Finally, observe that the set TV = {mm(Ci) : i € N} has infinite complement in M and TV C {n, n + 1, . . . } so E meJV 2~ m < e. D The following lemma is the key in deriving our main result. The strategy of the proof is similar to the one in Theorem 1.8.4, but is more complicated. It will be convenient to introduce some notation that shall be fixed throughout the remaining part of this section. Let A and B be complementary infinite subsets of N. Put QA = {x € Q : xn = 0 if n g A},
respectively.
QB = [x € Q : xn = 0 if n g B},
5.2. EXTENDING HOMEOMORPHISMS IN s
315
It will be convenient to think of Q as QA x QB- We will specify A and B later. Let 6 > 0 be such that
The 'origins' of Q, QA and Qs, i.e., the points having all coordinates 0, shall be denoted by 0, or, as we think of Q as QA x QB, by (0,0). Let X,Y and Z be compact subsets of s such that X U Y C Q^ and Z C Qs and let p: A^ —>• Z and q: Y —> Z be homeomorphisms such that
Q(Q IP,
(*)
7
for certain 7 > 0. Again we will specify p,q,*y,X,Y and Z later (see Figure 16). We would like to construct a homeomorphism h\ which takes X onto the 'graph' G(p) of p. Similarly, a homeomorphism hi which takes Y onto the 'graph' G(q) of q. And finally, a homeomorphism ^3 mapping G(p) onto G(). Then h^1 ° h% o hi will a homeomorphism of Q extending the homeomorphism q~l op: X —>• F. Before we can make this precise, we define two technical concepts. A homeomorphism /: Q —>• Q is called an yl-homeomorphism (respectively, .B-homeomorphism) if for any x E Q we have
for all n G A (respectively, n e £?). So an A-homeomorphism is a 'vertical action'. Similarly, a S-homeomorphism acts 'horizontally'.
Figure 16. The desired homeomorphisms are constructed in the following lemma. Lemma 5.2.3. (1) There is a boundary preserving A-homeomorphism hi : Q —> Q such that hi(x,0) = ( x , p ( x ) ) for every x 6 X.
316
5. BASIC INFINITE-DIMENSIONAL TOPOLOGY
(2) There is a boundary preserving A-homeomorphism h-2: Q —> Q such that hi(y, 0) = (y, q(y}) for all y € Y. (3) There is a boundary preserving B-homeomorphism h% : Q —> Q such that hs(x,p(x)] = (q~1p(x),p(x)} for every x € X while moreover Q(h3,lQ) < 7. The reader should pause to see what is going on. The homeomorphism hi indeed takes X onto the 'graph' G(p) of p. Similarly, h2 takes Y onto the 'graph' G(q) of q. Finally, /i3 is a 'small' homeomorphism mapping G(p] onto G(q). So h%1 ° /i3 o hi is a 'small' homeomorphism of Q extending the homeomorphism q~l op: X ^ Y (see Figure 17) if QB has 'small' diameter. Proof. Since X U Y U Z is compact we can find for every n an interval such that X U Y U Z C 0^=1 h r n> r n]- Put oo
X^ = {x £ JJ [-rn, r n ] : a;n = 0 if n n=l
and 71=1
respectively. Observe that both KA and KS are products of symmetric intervals (the symmetry of the intervals will be important later), that X U Y C KA ^ QA a n d that Z C K B B .
Figure 17. By applying the Tietze Extension Theorem 1.2.5 to each factor of KB separately, we can extend p: X —>• Z C KB to a continuous function p: QA ^ KB
5.2. EXTENDING HOMEOMORPHISMS IN s
317
(alternatively, apply Corollary 1.2.12). For every t € (—1,1) let (pt : J ->• J be the unique homeomorphism taking the interval [—1,0] linearly onto [— 1, t] and [0, 1] linearly onto [t, 1]. For each n 6 B let the isotopy # n : JJ x [-r n ,r n ] ->• J be defined by Hn(x,i] = (pt(x). The H^'s define a .ftTs-isotopy H:QB xKB^QB as follows:
H(y, x)n = Hn(yn, xn)
(n e B)
(Proposition 1.8.1) (we make an obvious identification here). Now define
hi : QA x QB -* QA * <2s
by hi(x,y) =
(x,H(y,p(x))).
Then /ii is an A-homeomorphism (Theorem 1.8.2). In addition, if x G X then (for the last equality, use the definition of the Hn's for n G -B). We conclude that hi is as required. The construction of h? is precisely the same. The construction of ^3 is similar but slightly more complicated because of the smallness condition involved. Consider p~l : Z -> X C KA and q~l : Z ->• Y C ^. By Exercise 1.3.8 and (*) on Page 315,
e(p~l,g~1} = Q(p~lp,q~lp) = o(^x,q~lp) < 7Claim 1. There exist continuous extensions £, rj: QB —> KA of p~l and g"1, respectively, such that Q(£, 77) < 7. Proof. By applying the Tietze Extension Theorem 1.2.5 to each factor of KA separately (alternatively, apply Corollary 1.2.12), we can extend p~l and q~l to continuous functions f , g : QB —> KA- Put
It is clear that U is an open neighborhood of Z in QB- Let a: QB —>• I be a Urysohn function such that a\Z = l,
a\(QB\U)=0
(Corollary A. 4.1). Now define £,r?: QB ->• KA by £(x)=a(x)-f(x),
r](x}=a(x}-g(x},
318
5. BASIC INFINITE-DIMENSIONAL TOPOLOGY
respectively. Observe that by the special choice of KA these functions are well-defined (and obviously continuous). Now if x 0 U then a(x) — 0, and so £(x) = r)(x). On the other hand, if x £ U then 1~nn/(r\ f(r\ — n(r\ z o^xj • I\J\X)n y\^)n\I n=l
= a(x] < 1 -7
-g(f(x),g(x))
From this we conclude that £(£,17) < 7 (Exercise A. 5. 4). If x G Z then clearly a(x) — 1, so £(x) = 1 • f(x] = f ( x ] . We conclude that £ extends /, and similarly, that T] extends g. Consequently, £ and TJ are as required. <0For all ( x , y ) € (-1,1)2 let
be the unique homeomorphism taking [— l,x] linearly onto [—I,?/] and [x, 1] linearly onto [y, 1]. Observe the following easy: Claim 2. If ( x , y ) e (-1,1)2 then £(<£>( x > y ) ,lj) = \x - y\. We now define a (KA x /d)-isotopy FrQ^x^x/iCO-X^ coordinate wise as follows: F(q, x, y)n = V ( X n , y n ) (qn)
(n £ A),
cf. Proposition 1.8.1. Now define h3: QAxQB ^ QAx QB
by h3(x,y) =
(F(t(y),r,(y))(z),y).
Then h% is a 5-homeomorphism (Theorem 1.8.2). We shall prove that ^3 is as required. Take an arbitrary x E X. Then h3(x,p(x)) = = =
(F(£p(x)^p(x)}(x},p(x)) (F(x,q-ip(x))(x),p(x)) ((1~1P(X):P(X))
(for the last equality, use the definition of F).
5.2. EXTENDING HOMEOMORPHISMS IN s
319
Finally, take an arbitrary (x,y) G QA x QB- Then
<7 (use Claims 1 and 2). We conclude that g(hs, IQ) < 7 (Exercise A. 5. 4).
D
We now come to the main result in this section. Theorem 5.2.4. Let .E, F C s be compact and let f : -E -> F be a homeomorphism such that £(/, Is) < e. Then / can be extended to a boundary preserving homeomorphism f:Q-*Q such that £>(/, IQ) < e. Proof. Let e\ = g(f, 1^) and put 5 = l/6(e — e^) (this specifies £). By Corollary 5.2.2 we find a boundary preserving homeomorphism g: Q -> Q and an infinite subset B C N such that (1) 0(g, 1Q) <S, (2) Kng[E U F] = {0} for every n € 5, (3) A = N \ 5 is infinite and £ n£B 2"n < %
(this specifies A and 5). Put X — g[E], Y ~ g[F] and h = g <=> f o g~l. Observe that Notice that ^C U F C Q^. Since Q^ PS Q, we can find a topological copy Z of X in s such that Z C QB (Corollary A. 4. 4) (Qs is not a subset of s but a smaller infinite-dimensional 'subcube' is a subset of s ) . Let p: X —>• Z be any homeomorphism and let g — p o h~l. Then clearly Put 7 = ei + 2(5. This specifies p, q and 7. Now let hi,h<2 and /is be as in Lemma 5.2.3. By (3) we have e(hi,!Q)<6, Q(h2,lQ)<6. Consequently, if we put 1 T = /i" o /i3 o hi, then Q(T, IQ} < 7 + 26 = ei + 46.
320
5. BASIC INFINITE-DIMENSIONAL TOPOLOGY
Since r clearly extends h, we find that f = g'1 or og
extends / and that £(/, IQ) < e\ + 46 + 26 — e. Since it is clear that / is boundary-preserving, this finishes the proof.
D
Exercise for §5.2. Let e > 0. An isotopy H: Q x I —> Q is called an e-isotopy provided that for every x G Q the diameter of the set H [{x} x I] is less than E. ^•1. Let E, F C s be compact and let /: E —>• F be a homeomorphism such that Q(/,IE) < £• Prove that there is an e-isotopy H: Q x I —>• Q such that Ho = IQ and Hi \ E — f (i.e., HI extends /). (This improves Theorem 5.2.4.) 5.3. The estimated homeomorphism extension theorem The aim of this section is among other things to prove that if E, F
(1) fc[W«]nU{Wf:t<m,Aie{-l,l}} (2) h[W°n}CW^ (3) g(h,lQ) <e, (4) h\K = l. Proof. Let e > 0 and choose m > n such that
= 0,
5.3. THE ESTIMATED HOMEOMORPHISM EXTENSION THEOREM
321
We first push W% into W^ and then away from the endfaces in the lower coordinate directions. Without loss of generality we assume that 0=1. It is geometrically obvious that there is a homeomorphism
having the following properties: (5) ^ [ { l } x J T O ] C J n x { l } , (6) if x < I — 2~( m ~ 1 ) then
Figure 18. This is precisely the same homeomorphism as the one in Figure 4 on Page 61. Let hi 6 J(.(Q) be the homeomorphism of Q that is defined by crossing
i x {l}m] C njLi (-1,1), x {l}m,
(11) (12)
Now let hi G N(Q) be the homeomorphism that is defined by crossing with the identity on the other factors of Q. It is clear that
5. BASIC INFINITE-DIMENSIONAL TOPOLOGY
322
(13)
h-t\K = l K ,
(14)
0(/l2,lQ)
From (7), (8), (9), (10), (13) and (14) it follows easily that h = h2 o hi is as required. D
rm —1
Figure 19.
Corollary 5.3.2. For each n G N, 9 G { —1,1} and e > 0 there is a homeomorphism h: Q —> Q with
(1) h[W»] C s, (2) e(h,iQ) <e, (3) h\K = \. Proof. Put HI = n. By Lemma 5.3.1 we can push W% into some other endface W^2 with n? > ni, by a homeomorphism hi G ^K(Q) which is as close to to the identity as we wish. Repeated applications of this yields a sequence (hi)i G CK(Q), each element of which is close enough to the identity in order to have h = lim hi o hi-i o • • • o hi G ^K(Q), i-*oo
g(h, IQ) < e
(Theorem 1.6.2). Moreover, each hi can be taken to be the identity on K, so that h \K = IKFinally, each hi pushes the face W^. off the faces W? for j < nl+i and p, in { — 1,1}. For hi close enough to the identity (as prescribed in Lemma 1.6.1), we can keep the limit homeomorphism h 'away from all faces of Q', that is, h[W%] Cs. D
5.3. THE ESTIMATED HOMEOMORPHISM EXTENSION THEOREM
323
We now know how to push an endface into s. We are going to use this result to push an arbitrary Z-set A away from a fixed endface, and then away from all endfaces, i.e., to push A into s. We need the following result. Proposition 5.3.3. Let X be a compact space, let X0 C X be closed and let f : X ->• s be continuous such that f \ XQ is an imbedding. Then for every e > 0 there is an imbedding g : X —>• s such that (1) g ( f , g ) < £ , (2) g\X0 = f\X0. Proof. Let 53 be a countable collection of compacta in X \X0 such that their interiors form a base for X \ XQ (observe that X \ XQ is locally compact). Assume that 0 £ !B. In addition, let {(Fj,G;) : i € N} enumerate the set
Choose ra 6 N such that
n=m+l
Since 7rm+i/[Xo] is a compact subset of (—1, l) m +;, for every i € N we can apply the Tietze Extension Theorem 1.2.5 to construct a continuous function gl : X —> (— 1, l)m+i such that (3)
9l
\ XQ = (7Tm+l O /)
\X0,
(4) #i[Xo], 9i[Fi] and gi[Gi] are pairwise disjoint. (Pick two distinct points p, q G (— 1, l)m+i \ Tfm+if[Xo\ and extend the function that is constant p on F{, constant q on G;, and is equal to 7rm+; o / on XQ.) Define g: X —> s by g(x) = (/(z)i,/(o;)2,...,/(:r)m,5i(z),02(20,....). It is clear that g is continuous, that g \ XQ = f \ XQ and that @(f,g) < £• Take x,y £ X such that x ^ y. If x,y £ X0 then g(x) ^ g(y) since f \ XQ is an imbedding. If x € X0 and y ^ XQ then we can find an index i 6 N such that y £ F{. By (4) and the definition of g it now follows that
Finally, if x,y ^ XQ then we can find an i G N such that x belongs to Fi and y belongs to Gi. As above it follows that g(x)m+i ^ g(y)m+i. We conclude that g is one-to-one and the compactness of X now yields that g is an imbedding (Exercise A. 5. 9). D Lemma 5.3.4. Let A G Z(Q). Then for each n G N, 9 € {-1, 1} and £ > 0 there is a homeomorphism h: Q —> Q with
324
5. BASIC INFINITE-DIMENSIONAL TOPOLOGY
(1) h[A] n wl = 0, (2) g(h, IQ) < e, (3) h\K=l.
Proof. By Corollary 5.3.2 we can find a homeomorphism hi £ !K(<5) such that g(hi,lQ) < £/3, hi \K = 1K and hi[W^\ C s. Since A\JK\JB(Q)
eZa(Q)
(Corollary 5.1.5), by Lemma 5.1.2(4) there is a continuous function such that g(a, IQ) < e/3- Consequently, a[Q] C s and hence by compactness, g(a[Q],A U /f) > 0. By Proposition 5.3.3 choosing XQ = 0 there is an imbedding [3 from Q into s with
(4) p[Q]n(A\JK) = Q,
(5) g(p,a)<*/3
(observe that we can achieve (4) since ^(a[<5],^4 U K) > 0). Notice that
Put
Since the function 7 is a homeomorphism from hi [W%] U K onto f3hi [W%] U A' such that £(7,1) < 2e/s- By Theorem 5.2.4, we can extend 7 to a homeomorphism F: Q -> <5 such that ^(P, IQ) < 2e/3. Observe that
(6) T\K = 9I, (7) r[/n[w n ]]nA = 0. Now put h = (P o /ii)" 1 . Then since
the function /i is as required.
D
We can now prove the following: Theorem 5.3.5. Let K C s be compact and let A e Z(Q). Then for every e > 0 there is a homeomorphism h : Q —> Q with (1) Q(h,\Q} <£,
(2) h\K = lK, (3) 7t[A] C s.
5.3. THE ESTIMATED HOMEOMORPHISM EXTENSION THEOREM
325
Proof. All we have to do is to apply Lemma 5.3.4 inductively to free A from all endfaces W% while keeping K pointwise fixed. This has to be done with a little care so that once A is free from an endface, the limit homeomorphism does not carry it back to that endface. But this can be done again writh the help of Lemma 1.6.1 (cf. the proof of Corollary 5.3.2). D This leads us to the following basic result. Corollary 5.3.6. Let A € Za(Q). Then for every e > 0 there exists a homeomorphism h: Q —> Q such that h[A] C s and g(h, IQ) < e. Proof. Write A = \J™=1 An, where An e Z(Q) for every n. By Theorem 5.3.5 there is a 'small' homeomorphism j\: Q —> Q such that /i[Ai] C s. Again by Theorem 5.3.5 there is a 'small' homeomorphism / 2 : Q —>• Q such that /2 [/i [^2]] C s while moreover /2 restricts to the identity on /ifylij. Etc. The infinite left product h = limn_j.00 fn o • • • o f1 is a homeomorphism of Q such that h[A] C s and g(h, IQ) < e. D We can now come to the main result in this section. Homeomorphism Extension Theorem 5.3.7. LetE,F € Z(Q) and let f
be a homeomorphism from E onto F such that g(f, IE) < £• Then f can be extended to a homeomorphism f'Q-^Q such that g(f, IQ) < s. Proof. Just apply Theorems 5.3.5 and 5.2.4 to E U F.
D
Perhaps the reader feels that it is not worth the trouble in Theorem 5.3.7 to prove that it is possible to extend 'small' homeomorphisms to 'small' homeomorphisms. At first glance it seems that the possibility of extending homeomorphisms is the most important fact and that the extra smallness condition is a technical curiosity. This is not true however. The possibility to extend 'small' homeomorphisms to 'small' homeomorphisms makes it possible to apply the inductive convergence criterion once again to create new homeomorphisms from old ones. This procedure turns out to be extremely powerful. Observe that the metric Q in Theorem 5.3.7 is the standard 'convex' metric on Q. It can be shown that the theorem is false if one replaces Q by an equivalent 'nonconvex' metric. For details see e.g., ANDERSON, CURTIS and VAN MILL [20]. Of course, homeomorphisms between Z-sets can be extended, but problems arise with the smallness condition. The following result should be compared with Corollary 5.3.6 A Hilbert cube is a space homeomorphic to Q. We would like to have an estimated homeomorphism extension theorem for Hilbert cubes but run into 'smallness' problems here as well. Fortunately, a weaker version of the
326
5. BASIC INFINITE-DIMENSIONAL TOPOLOGY
estimated homeomorphism extension theorem is true for Hilbert cubes which is powerful enough for the applications that we have in mind. It will be convenient to let M® denote an arbitrary Hilbert cube. Corollary 5.3.8. Let X = M® be a Hilbert cube with admissible metric p. Then for each e > 0 there is a 6 > 0 with the following property: if A, B G Z ( X ) are arbitrary and f : A —> B is a homeomorphism with p(/, I A) < 5 then there is a homeomorphism /: X ->• X such that f \ A = f and p(f, lx) < e. Proof. Let /: Q —>• X be any homeomorphism. Fix e > 0 and let 7 > 0 be such that if A C Q has diameter less than 7 then f[A] has diameter less than e. (Here we use that / is uniformly continuous (Exercise A. 5. 18).) Next, let 8 > 0 be a Lebesgue number for the cover (*)
{/[£/] : U C Q is open and diamt/ < 7}
(Lemma A. 5. 3). We claim that 6 is as required. To this end, let
A,B and let g: A ->• B be a homeomorphisms such that p(g, lx) < <5. It is clear that /~1[A] and f ~ l [ B ] are Z-sets in Q (cf. Lemma 5.1.2(6)) and that the function £: /~ 1 [A] —>• f ~ l [ B ] defined by £() = f ~ 1 g f ( q ) is a homeomorphism. We claim that £(£, IQ) < 7. To see this, fix an arbitrary q G /-1[A]. Then f ( q ) G A and so p ( f ( q ) , g f ( q ) } < 5. There is an open subset U C Q such that f[U] contains both f ( q ) and gf(q). This is so because 6 is a Lebesgue number for (*). But then q and £(q) — f ~ l g f ( q ) both belong to [7, i.e., g ( q , £ ( q ) ) < 7 which is as claimed (cf. Exercise A. 5. 4). By the Homeomorphism Extension Theorem 5.3.7, we can extend £ to a homeomorphism f : Q -> Q
such that £>(£, IQ) < 7. Now put g — f o £ o /-1. It is clear that ^ is a homeomorphism and extends g. The only thing that needs to be checked is the smallness condition. To this end, take an arbitrary x in X, and let q be f ~ l ( x ) . Then g ( q , £ ( q ) ) < 7 from which it follows that p(x,g(x))=p(f(q),g(f(q))=p(f(q),ft(q)) as required.
< e, D
Corollaries. We will now derive some interesting corollaries of the Homeomorphism Extension Theorem. Corollary 5.3.9. Let S G Z(7(Q} and let A be an infinite collection endfaces of Q. Then for every e > 0 there is a homeomorphism h: Q —>• Q such that g(h, IQ) < e while moreover h[S] C
5.3. THE ESTIMATED HOMEOMORPHISM EXTENSION THEOREM
327
Proof. For every A E A there are n(A) E N and 0(A) E {-1,1} such that A = Wn(A) • Without loss of generality we assume that there exists an infinite subcollection A' of A such that for every A E A', 6(A) = 1. By Corollary 5.3.6 we may assume without loss of generality that S C s. Write S as IJ^ S;, where Si is compact for every i. Let e > 0 and pick a natural number n so large that oo
E 2~m < vWrite N \ { l , 2 , . . . , n — l}as the disjoint union of infinitely many infinite sets, say {Ei : i E N}, such that for every i E N, min(^) = n(A) for certain A E A'. Let J.J.
H.
k^E,
k£Er
be a homeomorphism such that
Simply observe that {x G JSi : xmin(Ei) = 1} is a Z-set copy of Q by Lemma 5.1.2(5) and hence contains a topological copy of 7TEi[Si] by Corollary A. 4. 4. So hi is given by Theorem 5.3.7 since the compact set TT^ [Si] is contained in the pseudo-interior of the Hilbert cube n/ces Jfc and hence is a Z-set there (Corollary 5.1.5). Define h: Q —> Q by (J
D
Corollary 5.3.10. If 5 e Za(Q] then Q can be deformed through Q\S. Proof. By Corollary 5.3.9 we may assume without loss of generality that S is a subset of B(Q). So the homotopy H : Q x I —>• Q defined by H(x,t] = (l-t)-x for i 6 I and x € Q is as required.
D
Mapping Replacement Theorem. Our final application of the Homeomorphism Extension Theorem is a generalization of Proposition 5.3.3 which will be used several times in the forthcoming. Let X be a space. A continuous function /: X —>• Q is called Z-map provided that f[X] 6 Z(Q). If / is an imbedding as well as a Z-map then it is called Z-imbedding.
328
5. BASIC INFINITE-DIMENSIONAL TOPOLOGY
Mapping Replacement Theorem 5.3.11. Let X be compact with closed subset A. Let f : X —> Q be a continuous function such that f \ A is a Zimbedding. Then for every e > 0 there is a Z-imbedding g : X —>• Q such that g(f, g) < o and g \ A = / \ A. Proof. Since B(Q) 6 2>
D
Exercises for §5.3. ^•1. Prove the following generalization of Theorem 5.3.7: If E1, F £ £(Q) and /: E1 —>• F is a homeomorphism such that g(f, IE) < £ then there is an eisotopy H: Q x I —> Q with /fo = IQ and HI f E = /, i.e., //i extends /. 2. Prove that Q is isotopically homogeneous.
5.4. THE COMPACT ABSORPTION PROPERTY
329
3. Let X be a space. Suppose that A' = Qo U Qi, where
Qo w Qi K Qo n Qi K, Q, Qo n Qi € Z(Q 0 ) and Q0 n Qi e Z(Qi). Prove that X « Q. 5.4.
The compact absorption property
The aim of this section is to characterize all elements A e Za(Q) for which there is a homeomorphism h: Q —> Q such that h[A] = B(Q). As an application, we prove that an infinite product of intervals is homeomorphic to s iff infinitely many of them are not compact and that s \ E w s for every cr-compact subset E C s. We also prove the curious result that the subspace {A 6 21 : dim A = 0} of 21 is homeomorphic to s. This result will be important later. Capsets. Let MQ be a Hilbert cube. An element A £ Za(MQ] is called a capset1 provided that A can be written as the union of an increasing sequence AI C A-2 C • • • C An C • • • of Z-sets in M® such that the following absorption property holds: Ve > 0 Vn € N V Z € Z(MQ) 3m>n3he ^K(M^): (1) g(h,l) <e, (2) h\An = l, (3) h[Z]CAm. We will prove later that a subset A C Q is a capset if and only if there exists a homeomorphism h: Q -» Q such that h[A] = B(Q). So the property of being a capset characterizes the way B(Q) is topologically placed in Q. This implicitly also characterizes the way s is placed in Q, For proving that a subspace A of Q is homeomorphic to s it is sufficient (but not necessary) to prove that the complement B = Q \ A is a capset. Theorem 5.4.1. Let A be a capset in the Hilbert cube M®. Then for all K, L e Z(MQ] and e > 0 there is a homeomorphism h: MQ -> MQ such that (1)
Q(h,l)<£,
(2) h\K = lK, (3) h[L \K]CA.
Roughly speaking, A absorbs L \ K by a small motion keeping K pointwise fixed. 1
CAP abbreviates Compact Absorption Property.
330
5. BASIC INFINITE-DIMENSIONAL TOPOLOGY
Proof. Let the Z-sets A n , n G N, witness the fact that A is a capset. In addition, let K and L be Z-sets and let e > 0. Finally, let L0 C LI C . . . be a sequence of closed subsets of L\K such that LQ = 0 and L\K = \Jf*L1 Li. We shall inductively construct a sequence /o, / ! , - • • in ^K(M^) and a sequence of natural numbers n(0) < n(l) < • • • such that fo = l and n(0) = 1. while moreover the following statements hold: (1) g ( f i , l ) is so small that the conditions mentioned in the Inductive Convergence Criterion 1.6.2 are satisfied, (2) £(/,,!)< 3-* - e , (3) / i O / ^ o . - . o / o t L i l C A ^ ) , (4) fi\K\JAn(i,l) = l. This will establish the theorem since by (1) and (2), / = lim^oo fi° • • • ° f i exists, belongs to M(MQ), and satisfies £>(/,!) < e, while moreover by (3) and (4), /[L \ K] C A and / \ K = 1 K . It remains to perform the induction. Since /o and n(0) are already defined, assume that /; and n(i) have been chosen up to i > 0. Find 6 > 0 such that if we choose fi+i 5-close to 1 then (1) and (2) are automatically satisfied for i + 1. Let B - fc o j^ o ••• o / 0 [Li+i] and observe that B G Z(MQ) (Lemma 5.1.2(6)) and that B n K = 0. Let 7 = e(B,K). Let ^ > 0 be such that every homeomorphism between Z-sets of M® that moves the points less than £ can be extended to a homeomorphism of M® moving the points less than 6 (Corollary 5.3.8). Since A is a capset, there are ra > n(i) and a homeomorphism g : M® —>• M® such that (5) 0(0,1) (6) g\An(l) =1, (7) g[B]CAm. Observe that g[B] n K = 0. It is clear that g is not quite yet as required since g might move K. However, clearly (g\B)\JlK\JlAn(i) is a homeomorphism of B U K U An^ onto g[B]uK(J An^ . This homeomorphism can be extended to a homeomorphism h: M® —> MQ with g(h, 1) < 6. Now if we put n(i + 1) = m and fi+i = h then these choices are easily seen to be as required. D Corollary 5.4.2. Let A be a capset in the Hilbert cube MQ . Then for all K, L e Z(MQ] such that K C An L and e > 0 there is a homeomorphism h: MQ ->• MQ such that (1) g(h,l)<e,
(2) h\K = lK, (3) h[L] C A.
5.4. THE COMPACT ABSORPTION PROPERTY
331
We now present some fundamental properties of capsets. Theorem 5.4.3. Let MQ be a Hilbert cube. (1) If A C M® is a capset and h: M® —>• M® is a homeomorphism then h[A] is a capset. (2) If AC MQ is a capset and B e Za(MQ) then AUB is a capset. (3) If A and B are capsets in M® and Z C M® is a Z-set (possibly empty) then for each e > 0 there is a homeomorphism h : M® —> M® such that h[A\Z] = B\ Z, g(h, 1) < e and h \ Z = \z. Remark 5.4.4. It follows in particular from Theorem 5.4.3(3) that capsets are topologically unique, that is, if A and B are capsets in Q then there is an element h e 'K(Q) such that h[A] = B. Proof. For (1), observe that h is uniformly continuous (see e.g., the proof of Corollary 5.3.8). For (3), write OO
00
A\Z=\jAt,
B\Z=\jBz,
i=l
i=l
Q
where Al,Bl e Z(M ] for every i G N and AI - 0 = BI (observe that A is acompact, hence so is A \ Z, apply Lemma 5.1.2, etc.). We shall inductively construct a sequence /i, /2, . . . in Oi(M^) such that /i = 1 and (4) g(fi, 1) is small enough in order to apply Theorem 1.6.2, (5) g ( f i , l ) < 3-* - e , (6) BiCfiog^A], (7) fiog^AJCB, (8) / i r^uU;= 1 i(5i-i[A J -]Ufi J -) = l, where gl-l = /,_i o ••• o / j . Assume that /i, . . . , fi have been constructed. Find 6 > 0 such that if we choose fi+1 J-close to 1 then (4) and (5) are satisfied for i + 1. Observe that g;[/ii+i] E Z(M®) (Lemma 5.1.2(6)). Also, by our induction hypothesis,
Notice that K 6 Z(MQ] by Lemma 5. 1.2(3), (6), that Z n K = 0 and that Ai+i n Z = 0. We aim at applying Corollary 5.4.2 with the Z-sets K and K U g^Ai+i}. Indeed, since B is a capset which contains K, there is by Corollary 5.4.2 a homeomorphism Q: M® -» M® such that (9) g(a,l)< %, (10) a[^[yli + i]] C B,
(i.e., Al+i is absorbed by B),
332
5. BASIC INFINITE-DIMENSIONAL TOPOLOGY (11)
a\K=l.
There is a problem though, since a might move Z and we would like a to have the property that its restriction to Z is the identity. But this can easily be fixed by using the same method as in the proof of Theorem 5.4.1 since by (8), Zr\(K\Jgi[Ai+i]} = 0 (so this again depends on the Homeomorphism Extension Theorem). We may therefore additionally assume that (11)' a\(K\JZ) = l. Since a o 9i belongs to CK(M^), by (1) it follows that a o gi[A] is a capset. Notice that by (6) and (11)' we have, K' = K U a o 9l[Al+l} Cao 9i[A] and that Z U K' U Bi+\ is a Z-set. By a similar argumentation as the one above, there is a homeomorphism (3 : M® —> M® such that (12) g(P,l)<%, (13) j3[Bi+1] Cao 9i[A] (14) p\Z\JK' = 1.
(i.e., Bi+l is absorbed by a o gi[A}),
Now define /;+i = fl~l o a. Then clearly g(fi+i,l) and (14),
< 6. Moreover, by (10)
fi+i o gl[Al+i] = 0~l o a o # z [A i+ i] = a o ^[^+1] C B, and by (13),
Bi+i C/3'1 oao 9i[A] = fi+i o 9i[A]. Since clearly /i+i \ Z U K = 1, we see that /;+i is as required. Now put / = lim^co fi o • • • o /! = lim^oo ^. Clearly g(f, 1) < e. By (8) and (7) we have, 00
(15)
OO
f[A \ Z] = |J f[At\ = U 9^[A^} CB\Z. i=l
i=l
In addition, by (8) it follows that for every i,
= lim(/n o . . . o fi+l)[Bi] n>z
= Si,
so that by (6) and (8), We conclude that OO
(16)
B\Z=\jBiCf[A\Z]. i=l
By (15) and (16), f[A \Z} = B\Z and so we are done.
5.4. THE COMPACT ABSORPTION PROPERTY
333
Observe that the only thing that was used in the proof of (3) is that every capset satisfies the conclusion of Corollary 5.4.2. It therefore follows that we actually proved a (potentially) stronger result. This will be used in the proof of (2). For (2), let A C MQ be a capset, and let B 6 Zff(MQ}. Then A U B satisfies the conclusion of Corollary 5.4.2. So by the proof of (2), there exists a homeomorphism h: M® —>• MQ such that h[A] = A U B. So (1) implies that A U B is a capset. D A combination of (2) and (3) of the above theorem yields the following: Corollary 5.4.5. Let MQ beaHilbert cube, and let A, B e Zff(MQ) with A a capset. Then there is a homeomorphism h: MQ ->• MQ with h[A(JB] = A. B(Q] is a capset. The results on capsets proved so far are not very interesting if there were no capsets. Fortunately, B(Q] is a capset which we now aim at proving. Let n 6 N and define
i.e., En = {x e Q : (Vi e ®)(\Xi\ < 1 - 2~ n )}. Put oo
•=• _ I I= - - (J -n n=l
and observe that H is a
There obviously exists k > n such that for every i < m. For each j; > m let /ij : JJj —> Jj be a homeomorphism having the following properties: (1) ^[minTr^/^^maxTr^/^]] C [-1 + 2~ fc , 1 - 2~k] (2) /i,,- r[-l + 2 - n , l - 2 - n ] = l
334
5. BASIC INFINITE-DIMENSIONAL TOPOLOGY
(It is a triviality that these homeomorphisms exist.) Define h: Q —>• Q by •H^l i
i -Em — 1 1 %mi ' ' ' ) — V-^l j • • • ) -Em— 1 5 /^m V^ra/5 i^m+1 \%m+l J 5 • • • J
(i.e., /i is the product of the /im's and the identity on the first ra — 1 factors of Q). Then /i is a homeomorphism, £(/i, 1) < £/2 < e ? ^ T >^n — 1 and / i / C C E/C. So we are done. D We are now in a position to prove the converse of Corollary 5.4.2. Corollary 5.4.7. If an element A E Za(Q) satisfies the conclusion of Corollary 5.4.2 then it is a capset. Proof. By Proposition 5.4.6, E is a capset. The proof of Theorem 5.4.3(3) therefore gives us an h E ^K(Q) with h[A] = E. So A is a capset by Theorem 5.4.3(1). D We will now proceed to prove that B(Q] is a capset. Corollary 5.4.8. Let A be an infinite collection ofendfaces ofQ. Then (JA is a capset. Proof. By Corollary 5.3.9, the capset E can be pushed into U A by a homeomorphism of Q. Now since [JA E 2-0- (Q) (Lemma 5.1.3(3)), the desired result is a direct application of Theorem 5.4.3(2). D This easily yields the main result in this section. Theorem 5.4.9. An element A E 2-0- (Q) is a capset if and only if there is a homeomorphism h: Q —> Q such that h[A] — B(Q). Proof. Apply Corollary 5.4.8 and Theorem 5.4.3(3).
D
A useful characterization of capsets. We now derive derive an interesting characterization of capsets which can easily be applied in concrete situations. Let B = (Bi)i be a tower of subsets of Q. That is, Bi C Bl+l C Q for every i. We say that B has the deformation property if there is a homotopy H : Q x E —>• Q such that H0 = IQ while moreover for each t E (0, 1] there exists an i E N such that H[Qx[t,l]]
CBX.
A tower (Bi)i of compacta in Q is expansive if for each i there exists for some j an imbedding (fi : Bi x Q -> Bj with ?;(&, 0) = b for all b E Bi. Theorem 5.4.10. Suppose a crZ-set B C Q contains an expansive tower with the deformation property. Then B is a capset.
5.4. THE COMPACT ABSORPTION PROPERTY
335
Proof. By Corollary 5.4.7 and the Homeomorphism Extension Theorem 5.3.7 it suffices to show that for all compact a L C Q and K C L D B, and every £ > 0, there exists an imbedding e : L —>• B with g(e, 1) < e and e\K = 1. Assume for convenience that diam(Q) < Y2 (simply replace Q by the equivalent metric e/4). Also assume e < 1, and set 6 = 1/3e. Define the function A: L —> [0,6) by X(y) = d • g(y,K). By hypothesis, there exists an expansive tower (Bi)i in B with the deformation property. Let H : Q x I —> Q be a deformation such that for each t > 0, H[Q x [t, 1]] C Bi for some z, and assume that g(h(q,t),q) < t for all g and t (Exercise 5.4.4). Define a map /: L ->• 5 by the formula /(?/) = H(y,\(y)). Claim 1. / is well-defined, £(/, 1) < 6 and / f # = 1. Proof. If y e L \ A" then g(y,K) > 0 and so \(y) > 0, i.e., /(y) e B. Moreover, if y 6 K then A(y) = 0 and f(y) = H(y, 0) = y G -B. So the range of / is contained in B, and / \ K = 1. Pick an arbitrary y & L. Then
e(y, /(y)) = 0(i/, ^(y, A( y ))) < A(y) = <J • efo, ^) < y2(J, hence ^(/, 1) < i/26 < 6.
0
For each i > 1, let Ei =
{yeL:i/i+2
Thus L\K = (Ji^i -^i- For every i there exists a map »?i:
->
1
such that r)~ (Q) = Fr Ei (boundary with respect to L) and 77; f ^ \ Fr Ei is one-to-one (Exercise 1.6.5). Re-indexing the tower (Bi)i, we may assume for each i that
(a) and that there exists an imbedding (Pi : B, x Q ->• 5t+i with (^i(6, 0) = & for every b € Bi and (&)
diam^z[{6} x Q] < yi+2
for all 6. (The last fact follows easily from the compactness of Bi.) consider the map from EI to B% defined by
Now
Notice that for & e EI we have by (a) that /(&) G 5i and by hence that
vi(/(&)^i(&))efl 2 . So the map is well-defined, and clearly continuous.
336
5. BASIC INFINITE-DIMENSIONAL TOPOLOGY
Since rji[Fr EI] C {0}, for b € FrEi we get So the above map may be extended by / \ L \ E\ to a map f\ : L —>• B. Note that by (a) and the fact that B\ C B-2 it follows that fi[E-2\ C _B 2 - Next, the map from E2 to B3 defined by may be extended by /i f L \ E2 to a map /2 : I/ —)• -B with /2^s] C I?3. Continuing in this fashion, we obtain a sequence of maps fi : K —>• j5. Claim 2. The pointwise limit e = lim^oo fi exists and is continuous. It moreover has the following properties: (1) e\K = f\K = l, (2) Q(e,l) <e,
(3) e[L\K]r\K = 0. Proof. If y & K then g(y,K] = 0 and /(y) = y. From the construction it therefore follows that the sequence (fi(y)) • is the constant sequence with value f(y) — y. This shows that e is well-defined at points of K. Fixy E L\K, with l
/i+i
Then /t(y) = fi+i(y) = ••• = e(y). This shows that for this point e is also well-defined. Moreover, observe that Vi(fi-i(y),°)
= /i-i(y)
and so
By the definition of /; it therefore follows from (b) that Q(fMJi-M)
<5/r+2-
Now consider the points f i - \ ( y ] and /(?/). By construction we get
/(y) = /i- 2 (y) and so by a similar argumentation as the one above,
e(/i-i(y and so
5.4. THE COMPACT ABSORPTION PROPERTY
337
Since 6 = 1/3£ and g(y, K) < 1, we have g ( e ( y ) , y ) < e. By compactness of L this will imply that g(e, 1) < £ provided we will be able to prove that e is continuous (Exercise A. 5. 4). Observe that Vt+i <0(2/>*0,
<*
£<1,
so that (c)
e(e(y),y) < %+2 + %+1 + 5 - g(y, #) < 36 • g(y,K) <
Q(y,K}.
Continuity of e at points of L \ K is clear since each point of L \ K has a neighborhood on which / agrees with some /{, which is continuous. So it suffices to prove continuity at the points of K . A moments reflection shows that it is enough to show that if a sequence of points (yn)n in L\K converges to a point y € K then e(yn) —> e(y). From (c) we get e(yn) -» y. Since e \ K — I we have e(y) = y and so we are done. Observe that (c) also implies that e[L\K]r\K = ®.
<0>
We show that e is an imbedding by showing that e \ L \ K is one-to-one. This clearly suffices since by Claim 2, e[L \ K] n K — 0. Consider distinct points y,y' 6 L \ K. Suppose first that g(y,K] and g(y',K) lie in the same subinterval [yi+1, y). Then e(y) = fi(y) = <£»(/»-! (y),»7i(y)), and likewise for e(y'). Since r}i(y) ^ f]i(y'} it follows that e(y] / e(y'). On the other hand, if Q(y,K) e [yi+1, yz) and g(y',L) e [y,+1, y,), with z < j, then
e(y) = f^(y) e Bl+i c BJ, and
Since 77j(y') 7^ 0, e(y') ^ 5,. Thus e(y) + e(y').
D
Corollary 5.4.11. An element B E 2.CT(Q) is a capset if any only if B can be written as the union of a tower (Bi)i of Z -sets in Q such that: (1) each Bi is homeomorphic to Q, (2) each Bt is a Z-set in Bl+i , (3) (Bi)i has the deformation property. Proof. First consider the tower E = (En}^=l defined above. It is clear that Ei « Q and Ei is a Z-set in Ei+i for every i. For the latter statement, observe that Ei projects onto a proper subset of the projection of Ei+i in every coordinate direction (Lemma 5.1.3(2)). In addition, the function H: Q xl^Q defined by is a deformation of Q through E. If t 6 (0, 1] then for n such that 2~n < t we have H[[Q x [t, 1]] C En. So the tower (En)n has the deformation property.
338
5. BASIC INFINITE-DIMENSIONAL TOPOLOGY
Since capsets are topologically unique by Theorem 5.4.3, every capset can be written in the above form. Conversely, let (J3j); be a tower such as the one above. By the Z-set Homeomorphism Extension Theorem 5.3.7, the pair (Bi,Bi+i) is homeomorphic to the pair (Bi x {0}, Bi x Q). Thus (Bi)i is an expansive tower. So the desired result follows from Theorem 5.4.10. D Some other properties of capsets. We already know that the union of a capset and a countable union of Z-sets is again a capset. We now present a property of capsets in the same spirit. Theorem 5.4.12. Let M® be a Hilbert cube, let A C M® be a capset and let B C MQ be a Z-set. Then A\B is a capset. Proof. We may assume that MQ = Q and that B C B(Q) (Corollary 5.3.9). So 5 n B = 0. Now apply Theorem 5.4.3(3) to get a homeomorphism a: Q-+Q
such that a \ B is the identity, while moreover We conclude that A \ B is a capset by Theorem 5.4.3(1).
D
The following corollary, which at first glance seems rather technical, will turn out to be the basis for our considerations in the next section. Corollary 5.4.13. Let X be compact with closed subspace K. In addition, let A C X be a-compact. Then for every map f : X —>• Q that restricts to a Z -imbedding on K, there exists for every e > 0 a Z -imbedding g: X —» Q such that (1) g(g,f)<E, (2) g\K = f\K,
(3) g-1[B(Q)]\K = A\K. Remark 5.4.14. This result should be interpreted as follows. The (possibly empty) set K is already in place. So the function g should not interfere with this. The imbedding g pushes A \ K in place. Proof. By Theorem 5.3.11 we may assume without loss of generality that / is a Z-imbedding. Observe that f[A] £ ZCT(Q) and that f[X] e £(Q) so that by Theorems 5.4.3(3) and 5.4.12 there exists a homeomorphism a: Q —> Q such that a is arbitrarily close to the identity, a \ f[K] is the identity and a[B(Q) \ f[K}} = ((B(Q] \ f[X}) U f[A}) \ f[K}. Now put g = a~l o /. Observe that 9\K = f\K.
5.4. THE COMPACT ABSORPTION PROPERTY
339
Also, / and g are arbitrarily close, and
x e g-l[B(Q)} \ K & g(x) € B(Q) \ g[K]
\ f[K] <£> x € A \ K,
which is as required.
D
Application 1. We shall first present some other simple corollaries to the results obtained so far. Corollary 5.4.15. Let A be a a-compact subset of s. Then s\A and s are homeomorphic. Remark 5.4.16. For a generalization of this result, see Exercise 5.4.6. Proof. By Corollary 5.4.8, B(Q) is a capset. In addition, A e Zff(Q) by Corollary 5.1.5. Consequently, B(Q}\JA is a capset by Theorem 5.4.3(2). Theorem 5.4.9 therefore implies the existence of an h € Jt(Q) such that
i.e., h[s] = s\A.
D
Corollary 5.4.17. For every n, let An be a nondegenerate interval in H Then oo
n
4 ~ TO00 /l n r*j IK.
n=l
iff infinitely many of the An are not compact. Proof. If only finitely many of the An are not compact then the product of the A n 's is locally compact and hence is certainly not homeomorphic to M°° (Exercise 1.5.16). So assume that infinitely many of the An's are not compact. Observe that every An is homeomorphic to (—1,1), to (—1,1], or to [—1,1]. Also, infinitely often, An is not homeomorphic to [—1,1]. Consequently, the product of the A n 's is homeomorphic to the complement in Q of a set consisting of infinitely many endfaces. Now apply Corollary 5.4.8 and Theorem 5.4.9. D
340
5. BASIC INFINITE-DIMENSIONAL TOPOLOGY
Application 2. Here we will apply our results to the hyperspace of the unit interval I. The result presented here shall be used in the next section. If X is a compact space then for k G {0, 1 , 2 , . . . , 00} we let dim>k(X) denote the subspace of 2X consisting of all >k- dimensional elements. We define dimk(X) and dim
n=l
Proof. Let A G dim
ord(U) < k
(Theorem 3.2.5). Let K: rA —>• rX be the function in Lemma 4.2.2. For a cover U as the one above, the collection {K(U} :U£U}
has order at most k by (4) in Lemma 4.2.2. For a sufficiently small open neighborhood V of A we have that diam(/t(t/) n V) < e for every U G U. The collection {K(U) n V : U G U} therefore shows that A G $n,k(X) for every n such that e < l/n. Conversely, if A G D^Li 3n,k(X) then since for every collection of subsets U of X we have that ord(U \A)
D
Corollary 5.4.19. Let X be a Peano continuum. Then dim>i(X) is a crZset in 2X . Proof. By Exercise 5.4.9 and Lemma 5.4.18,
n—l
X
is an F^-subset of 1 . In addition, there is a deformation H : 1X x II —> 2X through yoo(X) by Corollary 4.2.24. Since every element of J^X) is zerodimensional, it follows that dim>i(X) is a crZ-set in 2X . D
5.4. THE COMPACT ABSORPTION PROPERTY
341
By the Curtis-Schorl- West Hyperspace Theorem 4.2.27 we know that 2 n and Q are homeomorphic. So it is natural to ask for 'natural' capset subspaces of 21. Here is one. Theorem 5.4.20. dinii(I) is a capset in 21. Proof. For convenience, let E* denote the subset {{t} : t e 1} of 21. It is a Z-set in 21 by Lemma 4.2.26. We will show that dinii (I) UE* is a capset of 2 n . But then dimi(I)= ( d i m i ( I ) u r ) \ I * is a capset as well by Theorem 5.4.12. For K C I, let UK = inf(-K) and b^ = sup(K), respectively. Notice that
aK + (1 - t) • (bK - aK) < bK 1
for all K 6 2 and t € I. For t € I define Mt = {K e 21 : [aK + ( ! - * ) • (bK - aK},bK] C K}. Claim 1. For t > 0, Mt is a Z-set in 21. Proof. We first claim that Mt is closed in 21. Let (Kn)n be a sequence in Mt converging to an element K 6 21. The points as and 65 depend continuously on S e 21 by Exercise 1.11.18. This implies that a,Kn —* &K and bxn —>• and hence [aXre + (1 - t) • (bKn - aKn), bKn] ->• [aK + (I - t) • (bK - OK), bK] in 2 n . So
K + ( l - t ) - ( ^ - a K ) , ^ ] CK by Exercise 1.11.8 and hence K 6 Mj. Every element of Mt is either a singleton or is one-dimensional. Since I* is a Z-set, and dinii(I) is a <jZ-set by Corollary 5.4.19, the fact that Mt is closed implies that it is a Z-set (Lemma 5.1.2(3)). <0>
For t 6 I and K 6 21 put
#t = (aK + (1 - t) • (A; - a*r) :fc€ ^}Observe that ^ is the image of K under a linear map which sends O-K onto UK and bK onto UK + (1 - t) • (bK — OK}Define H: 21 x I -» 21 by } = Kt(J [aK + (l-t)- (bK - aK), bK}. Claim 2. H is continuous, H0 is the identity, and Ht maps 21 homeomorphically onto Mt for every t > 0. In addition, gn(Ht, 1) < t for t € I.
342
5. BASIC INFINITE-DIMENSIONAL TOPOLOGY
Proof. That H is continuous follows easily from the fact that ax and bx depend continuously on K (Exercise 1.11.18). That H0 is the identity is trivial. Fix t e I and K e 2 H . Observe that mmHt(K) = ax,
maxHt(K) = bx-
This implies that Ht(K) £ Mf. We next claim that Ht is one-to-one. To this end, let K,K' 6 21. If aK < aK> then obviously Ht(K) ^ Ht(K'}. So we may assume without loss of generality that OK = ax> and, similarly, that bx = bx>- The linear maps that send K onto Kt and K' onto K[ are consequently identical. Since Kt and K't are both subsets of
To show that Ht maps 21 onto M^, take an arbitrary K € Mt. Let L = [ak,aK + (1 - t) • (bK - aK)} n K and let S be the image of L under the linear map that sends UK onto ax and the point ax + (I — t) • (bx ~ ax) onto bx- Then clearly, Ht(S) = K. That the distance between Ht and the identity is less than or equal t follows easily since bk ~ (ax + (1 - t) • (bK - aK)) = t(bk - ak) < t.
So we are done.
<>
Because for 0 < s < t < 1 it follows that
H~l[Mt]=M(t_s)/(l_s} is a Z-set in 21, this shows that Mt is a Z-set in M s . So now the desired result follows by Corollary 5.4.11 by using the tower (M^);. D Corollary 5.4.21. The spaces {At I1 :dimA = 0} and s are homeomorphic. Exercises for §5.4. For each i (E N let — 1 < cu < 6, < 1. The set {x G Q : Xi G [ai, bi\ for all but finitely many i G N} is called the basic core set structured on the core Hi^i [°M&i]^1. Prove that B(Q) x Q, B(Q) x B(Q) and (Q x Q) \ (s x s) are capsets for QxQ.
5.5. ABSORBING SYSTEMS
343
2. Let A e Z
5. Let A be a (relatively) closed subset of s and let B be its closure in Q. Prove that A G Z(s) iff £ G Z(<9). 6. Let A G Z CT (s). Prove that R°° \ A « s. 7. Let .4 C Q be closed and let / : Q —>• Q be continuous such that / \ A is a Zimbedding such that f ~ l [ s ] r \ A = sCiA. Prove that / can be approximated arbitrarily closely by a function g : Q —> Q having the following properties: (1) g is a Z-imbedding, (2) g\A = f\A, (3) 9-^8] = 8.
8. Prove that C = {x e Q : (3 k e N)(V j e N)(|o;j | < 2 fc ~ J ')} is a capset. 9. Let X be a compact space. Prove that for all n and k the set Sn,fc(^) is open in 2X . 0. Let /: Q —> Q be continuous, let A C Q be cr-compact and let K C Q be closed. In addition, let //: Q —>• [0, oo) be a continuous function such that //~ 1 (0) = K. Prove that there is a continuous function g: Q —» Q such that
(i) flr# = /rtf,
(2) 5 f Q \ K is one-to-one and g[Q \ K] e 2, CT (Q), (3) if x 6 Q \ K then (0(:c), /(x)) < ^(x), (4) 9 11. Let M and A/" be Hilbert cubes containing the capsets X and Y, respectively. Prove that if x G X and y G Y are arbitrary then there is a homeomorphism (p : M —> JV such that (1) v,[X] = Y, (2) (^(x) = y. 5.5. Absorbing systems
In this section we generalize the ideas in the previous section for obtaining more complex structures with 'absorption' properties. As applications we shall prove that the space of all infinite-dimensional compacta and the
344
5. BASIC INFINITE-DIMENSIONAL TOPOLOGY
space CQ endowed with the topology of coordinatewise convergence are homeomorphic to the countably infinite product of copies of B(Q). Notice that in Theorem 5.4.9 we characterized up to homeomorphism how the pseudo-boundary B(Q) of Q is placed in Q. We are interested in the question how the countably infinite product B(Q)°° is placed in the Hilbert cube Q°°. Notice that B(Q) is cr-compact, but not compact. So B(Q)°° is not <7-compact (Exercise A.5.19) and hence B(Q] and B(Q)°° are not homeomorphic. That this is not so surprising also follows from the following reasoning. Observe that B(Q) is the union of countably many endfaces which are easily seen to be nowhere dense in B(Q}. Hence B(Q) does not satisfy the Baire Category Theorem and by Theorem 1.9.12, it consequently contains a closed copy of Q. Hence B(Q)00 contains a closed copy of Q°°. So by Corollary A.13.4, B(Q}°° is an absolute Fa$ but not an absolute Gsa- (Theorem A.13.5). The conclusion is that we are dealing with something more complex than B(Q) here. Let M be a class of spaces. We say that M is topological if for every space B for which there exists an element A G M such that B K, A it follows that B e M. Similarly, M is closed hereditary if for every A e M and closed subset B C A we have B 6 M. Let F be an ordered set and let M7 be a collection of spaces for each 7 6 F. Each M7 is assumed to be topological and closed hereditary. Let Mr stand for the whole system (M 7 ) 7€ r- Let X = (JC7)7er be an order preserving indexed collection of subsets of a topological copy M® of Q. By this we mean that JC7 C X^< if and only if 7 < 7'. The system X is called Mr -universal if for every order preserving system (Ay)7 in Q such that A7 6 M7 for every 7 6 F, there is an imbedding /: Q -> M® with f'1^] = A7 for every 7 € F. So a system is universal if an arbitrary system of appropriate subsets of Q can be pushed in place by means of an imbedding of Q in M®. The system X is called strongly Mr -universal if for every order preserving system (A 7 ) 7 in Q such that A7 E M7 for every 7 6 F, and for every map /: Q —>• MQ that restricts to a Z-imbedding on some compact set K, there exists a Z-imbedding g: Q —> M® that can be chosen arbitrarily close to / with the properties: g \ K = f \ K and ^"^Xy] \ K = A7 \ K for every 7 6 F. So roughly speaking, a system is strongly Mr-universal if an arbitrary system of appropriate subsets of Q can be pushed in place off a Z-set by means of imbeddings approximating a given continuous function.
5.5. ABSORBING SYSTEMS
345
The system X is called reflexively universal, if for every continuous function / : M® —>• M® that restricts to a Z-imbedding on some compact set K, there exists a Z-imbedding g: M® -> M® that can be chosen arbitrarily close to / with the properties: g \ K = f \ K and g~l [X^\ \ K = X^ \ K for every 7 G F. So roughly speaking, a system is reflexively universal if it can be pushed in place off a Z-set by means of imbeddings approximating a given continuous function. Lemma 5.5.1. Let F be an ordered set and for every 7 G F let M7 be a topological and closed hereditary class of spaces. Let Mr stand for (M 7 ) 7e rIn addition, let M® be a topological copy of Q and let X = (Xy) 7£ r be an order preserving indexed collection of subsets of MQ . Then (1) If X is both Mr-universal and reflexively universal then it is strongly "Mr-universal. (2) If X is strongly Mr -universal and X^ € M7 for every 7 G F then X is both Mr -universal and reflexively universal. Proof. For (2) use that each M7 is topological. For (1), let (Ay) 7 be an order preserving system of subsets of Q such that Aj G M7 for every 7 G F, and let / : Q —» M® be a map that restricts to a Z-imbedding on some compact set K C Q. Since X is 3Vtr -universal there exists an imbedding a: Q —t M® such that o^fXy] = A^ for every 7 G F. Let e > 0.
Since Q is an AR (Corollary 1.2.12), we can extend the function
a~l : a[Q] -> Q to a continuous function T: M® —>• Q. Then obviously T o a(q) = q for every q G Q. Consider the composition
£ = f o r : MQ ->M Q . Observe that £ \ a[K] is a Z-imbedding from a[K] onto f[K]. Since X is reflexively universal, there consequently is a Z-imbedding g: M® — such that (3)
Q(g,foT)<e,
(4) V 7 G F : p-1^] \ <*[K] = X7 \ a[^], (5) g\a[K] = foT\a[K]. Now put h = g o a: Q ->• M^. Observe that by Exercise 1.3.2, £0, f) = e(g°a,f°Toa) < g(g, f OT) < e.
346
5. BASIC INFINITE-DIMENSIONAL TOPOLOGY
It is clear that h is a Z-imbedding since h[Q] C g[M®], and that h \K — f \K. Finally, for every 7 £ F we have h-1^] \K = a~i [g-l[X,]\ \ cr1
So we are done.
D
The system X is called Mr -absorbing if (1) X~f e M7 for every 7 e F, (2) U 7 er --^"7 ls contained in a crZ-set of M<2, (3) X is strongly Mr-universal. As to be expected, cf. Theorem 5.4.3(3), there is a uniqueness theorem for absorbing systems. Theorem 5.5.2. Let F be an ordered set and for every 7 E F Jet M7 be a topological and closed hereditary class of spaces. Let Mr stand for (M 7 ) 7e rFinally, let M® and M® be topological copies of Q. If X = (X 7 ) 7<Er ,
^ = (Ky) 7e r
are Mr -absorbing systems in M® respectively M® then there is a homeomorphism h: M® -» M® such that h[Xy] = Y^ for all 7 G F. If moreover M® = M® then h can be found arbitrarily close to the identity. Proof. We use the same back and forth argument as in the proof of Theorem 5.4.3. Obviously, we may assume that MQ - MQ' = Q. Let U7 ^7 Q USo A* and let |J7 F7 C [J^o B^ wnere 0 = A0 C Al C A2 C • • •
and 0 = Bo C Bl C B2 C • • • are sequences of Z-sets in Q. Let e > 0. By induction on i we shall construct a sequence of elements fi € ^K(Q) such that if gi = fi o • • • o fQ then (1) /o - 1, (2) for i > 1, g(fijl) is small enough in order to apply Theorem 1.6.2 later, (3) f o r i > l , £(/»,!) < 3 - * - e , (4) Al n X7 = Ai n g^[Y^] for alH > 0 and 7 e F,
5.5. ABSORBING SYSTEMS
347
(5) Bl n gi[X^\ = Bl n 1"7 for alH > 0 and 7 e F, (6) for i > 1, /< t te»-i[4i-i] U Bi_i) = 1, where 1 denotes as usual the identity function. Assume that fi has been constructed. Since X7 G M7 and M7 is topological and closed hereditary we have gi[X^]n(gi[Ai+l]UBi')
e M7
for every 7 e F. Put K = gi[Ai] U £?; and observe that gi[X^\ n K = F7 n K. Since ^ is strongly universal we can find an arbitrarily close to the identity ^-imbedding a: gi[Ai+\\ U Bi —>• Q that fixes K point wise and that has the additional property that for every 7 6 F,
Let a be an extension of a to a homeomorphism of Q that is arbitrarily close to IQ (Theorem 5.3.7). Since d o^[X] is just as X strongly universal, we can find a Z-imbedding /3 : aogi[Ai+i]\JBi+i —> Q that fixes K' = pointwise and that has the property that for every 7 e F,
j3~l [a o 9l[x^}] n Bl+l = r7 n B l+l . Let f3 be an extension of J3 to a homeomorphism of Q. Just as in the case of d, we may assume that /3 is arbitrarily close to the identity. If we put
then one can easily verify the induction hypothesis for i + I . So h = lim^oo Qi is a homeomorphism of Q such that g(h, 1) < e (this follows from (3)). It is easily verified, cf. the proof of Theorem 5.4.3, that h maps each X^ onto 17. D If the M7 for 7 G F are all equal to a fixed class M then we use the term ^-absorbing system. For such systems there are two important special cases. First, F = N with an inverted ordering. This means that X is a decreasing sequence of subsets of M®. In this case we use the term Mabsorbing sequence. The second special case is the case where F is a singleton. We then use the term Jti-absorber. An UV-absorber. Let yff denote the collection of cr-compact spaces. It follows from Corollary 5.4.13 that every capset in Q is an ^-absorber. Since by Theorem 5.5.2 3Vabsorbers are topologically unique, it follows that the property mentioned in Corollary 5.4.13 characterizes capsets. Theorem 5.5.3. B(Q] is an 3v-absorber.
348
5. BASIC INFINITE-DIMENSIONAL TOPOLOGY
Absorbing sequences in Q°° . We shall now consider the special case of absorbing sequences. So our system is a decreasing sequence Q 2 Xl D X2 3 • • • •
We put Xoc = n^Li Xn- If 3Vt is closed under finite intersections then M^ stands for the collection of countable intersections of elements of M. Lemma 5.5.4. Let X be an M-absorbing sequence in Q such that M is closed under finite intersections. Then X^ is an MS -absorber. Proof. Let X be the sequence Q D Xi D X2 D • • • . Let M = f|~=i Mn^Q be such that Mn e M for all n 6 N. In addition, let / : Q —>• Q be a map that restricts to a Z-imbedding on some closed subset K C Q. There exists by assumption a Z-imbedding a: Q —>• Q which approximates /, agrees with / on IT, and has the property that
i=i
for every n 6 N. Then clearly a"1^] \K required.
= f|£i Mt \ K, which is as D
Let X be a subset of Q. We define three decreasing sequences of subsets ofQ 0 0 : Sn(X) =X x ••• x X x Q x Q x ••• , n times
= {x € Q°° : at least n of the Xi's are in X}, S"(X) = {x £Q°° :xxeX for some i > n}. Note that Sn(X) C 5;(X) C 5;'(X), that
n=l
and that = S'^ = {x 6 Q°° : infinitely many xt belong to A^}. Theorem 5.5.5. Let X C Q be strongly M-imi versal. Then the sequences (Sn(X))n,
(S'n(X})n,
(SZ(X))n
are strongly M-universal in Q°° . If, in addition, M is closed under finite intersections then X00 and S'00(X) are strongly Ms-universal in Q°°. Proof. Let (gn}n be a sequence of admissible metrics on Q such that &(x,y) = max{gn(xn,yn) : n 6 N}
5.5. ABSORBING SYSTEMS
349
is an admissible metric on Q°° (Exercise A. 5. 16). Consider a map
f-Q-^Q00 that restricts to a Z-imbedding on some compactum K and a sequence of elements of M. We may assume that / is a Z-imbedding (Theorem 5.3.11). Write Q\K as a union of a sequence (^)°^o of compacta with Fi C Int(Fi+1) for every i and F0 = 0. Let e > 0 and put for every i. Consider now the n-th component fn: Q —> Q of /. We shall construct a sequence ao,Q;i,... of continuous functions from Q to Q such that for every and i: (1) gn(ai,cti-i) < Ei, ax \ Fi-i = a,i-i f-Fi-i, (2) QJ I" Q \ Fi+i = fn \ Q \ Fi+i and a; \ Fi is a ^-imbedding, (3) a^[X]nFi = Anr\Fi. Put ao = fn and assume that oti has been constructed. Using the strong M-universality of X we find a Z-imbedding f3: Fi+i —> Q, close to a.i f-Fi+i, with (3 \ Fi = ax \ Fx and 0~l[X] = An n FJ+I. Using the fact that Q is an AR (Corollary 1.2.12), extend /3 to a map a i+ i : Q ^ Q that restricts to / on Q \ Fi+2 and which is sufficiently close to o^ (Exercise 4.1.5). The a^s obviously form a Cauchy sequence and so the function gn = lim a% i—too
is continuous (Lemma A. 3.1). It is easy to verify that gn has the following properties: (4) Qn(9nJn) < £,
(5) if x e Fl+1 \Fi then en(gn(x),fn(x)) < (6) gn\K = fnlK, (6) gn \ Ft is a Z-imbedding for every i,
g(f[K]J[Fl+l]),
Define ^ = (gn)n' Q -^ Q°°- Note that g is one-to-one and hence an imbedding. The set g[Q] is contained in the crZ-set
i=0
and is therefore a Z-set (Lemma 5.1.2(3)). The maps / and g are e-close and / \K = g \K. Let x G Q \ K. If a; is an element of An then x e n?=i ^jConsequently, we have gj(x) e X for j = 1, 2, . . . , n. This means that
cs X c
350
5. BASIC INFINITE-DIMENSIONAL
TOPOLOGY
On the other hand, if g(x) is an element of S'^(X) then QJ(X] G X for some element j > n and hence x £ Aj C An. So we are done. D An 3>5-absorber. Consider the pseudo-boundary B(Q] of Q. It is an ^-absorber in Q by Theorem 5.5.3. Since ya is trivially closed under finite intersections we have by Theorems 5.5.2 and 5.5.5: Corollary 5.5.6. The sequences
(Sn(B(Q))) n ,
(Si(B(Q))) n ,
(s;'(B(Q)))n
are 3ff-absorbingand hence they are homeomorphic in Q°°. Moreover, B(Q)°° and S'00(B(Q}^ are yff$-absorbers. So just like B(Q) is the standard model of an 3~a-absorber, B(Q}°° is the standard model of an S^-absorber. Application 1: The space of all infinite-dimensional compacta. We are interested here in the spaces dim>k(Q), dim
(2) There exists a homeomorphism (3: 2® —> Q°° such that for every
fce{o,i,2,...}, fl[dim
Since dimoo(Q) = Pl^Lo dim>fc(Q), this shows that the space of all infinitedimensional compacta is homeomorphic to B(Q)°°. Corollary 5.5.8. There is a homeomorphism a: 2Q —> Q°° such that a[dim 00 (Q)] = S(Q)°°. We now present the proofs of these results. Proposition 5.5.9. The sequence (dim>fc(Q)) f c in 2«.
is refiexively universal
Proof. Let F: 2® —> 2^ be continuous and let OC be a closed subset of 2^ such that F \ "X. is a Z-imbedding. Since 2^ w Q (Theorem 4.2.27), we may assume that F is a Z-imbedding (Theorem 5.3.11) on all of 2®. Define the function e: 2Q -> I by =
5.5. ABSORBING SYSTEMS
351
There is by Corollary 4.2.24 a deformation H : 2Q x I ->• 2Q such that H0 is the identity and Ht(A) is finite for t > 0 and A £ 2^. By Proposition 4.1.7 we may assume without loss of generality that QH(Ht,l)
0
Define the homotopy a : 2Q x I -» 2^ by
where A is the subset of Hi^2 ^^ ^^a^ ^s obtained from A by a coordinate shift. Note that at(A) C [-t,t]°° and that a 0 (^) = {0}- The map G: 1Q ->• 2Q that approximates F is defined by G(A)=He(F(A})(F(A))+ae(F(A))(A). This function is continuous by Corollary 1.11.8 and the continuity of the homotopies H and a. Observe that QH(G(A),F(A)) <3e(F(A)) Q
for every A € 2 . If A 6 K then e(-F(A)) = 0 and hence G restricts to F on X Let A be an element of 2Q \ OC. Then t = e(F(A)) > 0 and hence Ht (F(A)] is finite. So G(A) is a finite union of translates of at(A) and consequently is the union of a finite set and a countable collection of copies
352
5. BASIC INFINITE-DIMENSIONAL TOPOLOGY
of A. This means that G preserves dimension by the Countable Closed Sum Theorem 3.2.8 and G-l[dim>k(Q}} \ X = dim> fc (Q) \ DC. We shall show that G is one-to-one. The restriction of G to OC is obviously one-to-one. If A <E 2Q \ 3C then QH(G(A),F(A})
< te(F(A)) <
6H(F[X\1F(A))
and hence G(A) is not in G[OC] = f[OC]. For the remaining case let the elements A and B in 2Q \OC be such that G(A) = G(B). Let ?r : Q ->• J be the projection onto the first coordinate. Define r — e(F(A)} and t = e(F(£?)). Select a point y = (a, x) E G(A) — G(B] such that a = min (?r(G(A))) = min (x(G(B))). Note that y is an element of both Hr (F(A)) and Ht (F(B}} . Since the latter sets are finite we can define A > 0 as one half of the distance of y towards the other points in Hr(F(A)) U Ht(F(B)). Let m and n be the first natural numbers that satisfy r/m < A and t/n < A. We now have: oo
({y} + [-A, A]00) n G(A] = {y} U \J {a + ^} x (x +
This implies
[a + r/m} x (x + r/mA) = {a+ Vn} x (x + */nB) and so r/m = ijn and T/mA = t/nB and hence that A ~ B. So G is one-to-one and therefore by compactness an imbedding (Exercise A. 5. 9). Observe that ?r[G(A)] is countable if A G 2Q \ % so G(A) is nowhere dense in Q. Since (A, t) ^Dt(A) = {xeQ: g(x, A) < t} is a deformation of Q through the complement of g[2Q \OC] (Exercises A. 5. 20 and 1.11.19), we have that G[2^ \ DC] is a crZ-set. Consequently, G[2Q] C F[OC] U
is a Z-set (Lemma 5.1.2(3)) and so G is a Z-imbedding. This completes the proof. D Proposition 5.5.10. The sequence (dim>fc(Q))^ 1 is 3a-absorbing in 2®.
5.5. ABSORBING SYSTEMS
353
Proof. By Corollary 5.4.19, dim>i(Q) is a aZ-set in 2®. In view of the previous proposition it suffices to show that the system (dim>fc(Q)) is ^-universal (Lemma 5.5.1). In Theorem 5.4.20 we showed that dinii(I) is a capset for the Hilbert cube 2 n . So the pairs (^dinnCI)),
(Q,B(Q))
are homeomorphic (Theorem 5.4.9). So by Corollary 5.5.6, S"(dimi(E)) is an 3Vabsorbing sequence in (21)00. Define the imbedding a: (21)00 —> 2Q by
Claim 1. Oi^i Pi 'IS ^-dimensional if and only if precisely k of the P;'s are in dimi(I). Proof. Consider H^i Pi- The dimension of each Pi is either 0 or 1. In the latter case, Pi contains an interval. So the dimension of O^i Pi equals
(Theorem 3.2.12) which proves the claim. So we have for every k. The sequence (dim>fc(Q)) f c is therefore ^-universal because the sequence (5^.(dimi(I))) is. D Corollary 5.5.11. dim^C,)) is an 3^5-absorber. We find Theorem 5.5.7 by by combining Theorem 5.5.2, Corollary 5.5.6 and Corollary 5.5.11. The fact that (2^, (dim>fc(Q)) ,_ 1 ) is homeomorphic to (Q°°,S(B(Q))] means that there is a homeomorphism a: 2® —> Q°° such that a[dim> fc (Q)j = B(Q) x • • • x B(Q) xQxQx-->
*«^BB^— ^^•^•KMh. ^ttm^^^^^^^^^^
K times
for every k. This implies that c^dinio^Q)] = B(Q)°°. By comparing the pairs (2«,dim> f c (Q)), (Q°°, we find part (2) of Theorem 5.5.7.
354
5. BASIC INFINITE-DIMENSIONAL TOPOLOGY
Application 2: CQ in the topology of coordinatewise convergence. The Banach space c0 was introduced in Example 1.1.9. It is the subset of s consisting of all elements the coordinates of which converge to 0 endowed with a natural norm topology. We will endow CQ here with the subspace topology it inherits from s, i.e., the topology of coordinatewise convergence (which incidentally coincides with the weak topology on c0, but we will not go into this). We will prove that CQ with this topology is homeomorphic to B(Q)°°. It will be convenient here to represent the Hilbert cube Q by E°° , where M stands for the compactification [—00,00] of IR. Consequently, R°° corresponds to the pseudo-interior of Q. In addition, for n 6 N let 0n = [x G M°° : \Xi\ < 2~ n for all but finitely many i}. Lemma 5.5.12. 0 = (Qn)n is a decreasing sequence of capsets in Q with intersection CQ. Proof. That the sequence (0 n )n is decreasing and has intersection CQ are trivialities. It is clear that 0 n is a basic core set for every n. So we are done by Exercise 5.4.3. D Proposition 5.5.13. The system 0 is ya-universal in Q. Proof. Let AI D A 2 3 • • • be a sequence of cr-compacta in Q. Let a: N x N - > N
be a bijection and put N; = {a(i,j) : j € N} for every i. For every i let Qi denote the Hilbert cube Claim 1. The set Cl = {xeQl:(3ke N)(Vj e is an Sv-absorber in Qi. Proof. This is Exercise 5.4.8.
0
Observe that for every x G Ci we have lim^oo xa^^ = 0. Define in Qi the aZ-set Di = {x G Qi : £ a (i,j) | < 2~ z for all but finitely many i}. Let fi : Q —> Qi be an imbedding such that f^l[Ci] = Ai and (Exercise 5.5.4). Consider the imbedding / = (fi)ii Q ->• O^i Qi ^ QLet x € An. If i > n then we have fi(x) € Qi and hence all components of f i ( x ) are in [— 2~ n ,2~ n ]. If i < n then we have x G Ai and hence f i ( x ) € Cx. Note that only finitely many components of f i ( x ) are outside [— 2~ n , 2~ n ] and
5.5. ABSORBING SYSTEMS
355
hence only finitely many components of f ( x ) are outside this interval. This means that f ( x ) is an element of 0 n . If x £ An then we have fn(x] <£ Dn. This means that infinitely many components of fn(x) have absolute value greater than 2~n and hence f ( x ) ^ 0 n . So we conclude that / -1 [0n] is equal to An. D For a space X and element * G X we let W(X, *) = {x e ^C00 : x; = * for all but finitely many i}. This set is called the weak Cartesian product of X. We will now prove that the system 0 is reflexively universal. To this end, let a : N2 ->• N be any bijection, and define $ : R°° ->• (E00)00 by the formula So $ simply rearranges coordinates. By letting * = 0 it is easily seen that with this map the system 0 satisfies the conditions of Lemma 5.5.14, except maybe for the fact that there is a deformation of IR00 through c0. To prove this, first observe that there is a deformation of R°° through M°° by Corollary 5.3.10. Next, c0 is a dense linear subspace of M°°. So there is a deformation of E°° through CQ by Corollary 4.2.15. This is clearly all we need (Corollary 4.1.8). Again let F be an ordered set. Lemma 5.5.14. Let X = (X 7 ) 7e r be an order preserving system in Q such that there is a deformation ofQ through X = fl 7 er -^7 and ^ * ^ (Tyer ^V Assume that there is a homeomorphism $ : Q —>• Q°° satisfying for all 7 G F. Then X is refiexively universal. Proof. Let /: Q —>• Q be a map that restricts to a Z-imbedding on some compact set K and let e > 0. The point * plays a central role in our considerations. If * E K then there is nothing to worry about. Suppose therefore that * <£ K. Since X is dense in Q (otherwise there would not be a deformation of Q through X) and f[K] is nowhere dense being a Z-set, there is point p £ X \ f[K] such that the distance between p and /(*) is arbitrarily small. Consider the function f : K ( J {*} —>• Q defined by
Then / is a Z-imbedding and is very close to / \ K U {*}. We may extend / to a function / : Q ->• Q which agrees with / on K\J {*} and which moreover is very close to / (Exercise 4.1.5).
356
5. BASIC INFINITE-DIMENSIONAL TOPOLOGY
The conclusion of these considerations is that we may assume without loss of generality that (1)
*?K
=*
f(*)eX\f[K].
We put K = K(J{*}. Let d be an admissible metric on Q°° so that (2)
if Xi = x'i for alH < k then d(x, x'} < 2~k~2
(cf. Exercise A. 1.10). Let U be the collection of all open subsets of Q of diameter less than l/2£, and let V = {$[U} : U 6 U} Then V is an open cover of the compact space Q°° and consequently has a Lebesgue number 7 > 0 (Lemma A. 5. 3). By Exercise 5.5.1 we may assume without loss of generality that (3)
f(Q\K]CX\f[K}.
Define 6: Q°° ->- [0,1) by 6(x) = min{7, d(x, $ o /[.K"])}. Claim 1. If a; € Q then <5($ o /(z)) = 0 iff x e K. Proof. This is easy. If x 6 K then clearly <5(4>o/(x)) = 0. Moreover, if x $ K then /(x) £ f[K] by (3) and so
(0 < * < i
l),2t-l)
(1/2 < * <
An easy check shows that # is as required.
5.5. ABSORBING SYSTEMS
357
Let Ci be the ^-th component of the map $ o /. For every k define a homotopy hk '• Q x I —>• Q by v
' '
\H(x,2-2t)
if l/2
It is easy to see that hk is well-defined and continuous. Observe that (/ijt)o is (k and (/ijt)i is the identity function. For k <E N, put
Observe that by Claim 1, U£U Pk=Q\K. For x £ Pk define /'(x) G <5°° by
hk+i(x,l x i > k + 4,
and extend /' on K by /' f ^ = $ o / f K. Claim 3. /' is well-defined, continuous and one-to-one. (Hence hence an imbedding by Exercise A. 5. 9.) Proof. Suppose e.g., that 6($ o /(x)) = 2~ fc for certain x G Q and k G N. Then on the one hand fofc+i (#, 1 + k - k] = hk+i (x, 1) = x H(x,l + k-k) =H(x,l) = *
i i i i i
< A;, = k + 1, G {k + 2 = k + 4, > A; + 4,
and on the other hand i — &,
x H(x, l + k-l-k) = H(x, 0) = x *
i e {fc + l : i = A; + 3, i > A;+ 3,
2},
So both definitions agree. It is clear that /' \ Pk is continuous for every k. Hence f'\Q\Kis continuous. If (xn}n is a sequence in Q \ K converging to an element x 6 K
358
5. BASIC INFINITE-DIMENSIONAL
TOPOLOGY
then 6($of(xn)} -> 8 ( $ o f ( x ) } = 0 (Claim 1) since all functions involved are continuous. We assume without loss of generality that xn G Pn for every n. Then lim f'(xn)=
lim (Ci(x n ),...,Cn(x n ), ? , ? , ? , . . . )
This proves that /' is continuous. We next prove that /' is one-to-one. To this end, let x, x' G Q be distinct points. We distinguish between several subcases. Case 1. There exists k G N such that x,x' G PkThen /'(aO fc+ 2 = x £ x' = f'(x')k+2. Case 2. There exist k G N such that x £ Pk and x' G -P&+1Then f'(x)k+3 = x and f'(x')k+3 = /'(z')(fc+i)+2 = z'- So we are again done in this case. Case 3. There exist k,k' G N such that k + I < k' while moreover x G Pk and x' G Pk' • Then fc + 4 < Jfc' + 3 and so f(x)k>+3 = * and /'(x') fc ' +3 = x'. But clearly x' $ K and since * G K this proves that f ' ( x ) ^ f ' ( x ' ) . Case 4. There exists k G N such that x G Pk and x' G /f . Striving for a contradiction, assume that f ' ( x ) = f ' ( x ' } . By the definition of /' this implies that $of(x) and $o/(x') agree in their first k coordinates. This implies by (2) that and hence So x ^ Pk, which is a contradiction. Since f'\K = 3>of\Kis one-to-one, we now conclude that /' is indeed one-to-one. <0> We next claim that /' is even a Z-imbedding. Claim 4. /' is a Z-imbedding. Proof. First observe that f'[K] G Z(Q°°} since f[K] G Z(Q) and $ is a homeomorphism (Lemma 5.1.2). Fix A; G N and consider f'[Pk\- Since the coordinates of every point in f'[Pk] are * beyond fc + 4, it follows that f'[Pk]
5.5. ABSORBING SYSTEMS
359
projects onto a proper subset of Q in infinitely many coordinate directions. So f [ P k ] G ^(Q00) by Exercise 5.1.8. This implies that f'[Q] G Za(Q°°) and hence by compactness of f'[Q] it is even a Z-set by Lemma 5.1.3(3). 0 Claim 5. g(3>~1 o / ' , / ) < e. Proof. If x G Q \ K is arbitrary, say x G Pfc, then /'(x) and $ o /(x) agree in their first fc coordinates, so d ( f ( x ) , $ o /(x)) < 2- fc - 2 < 5($ o /(x)) < 7. So there exists t/ G U such that
which implies that $-1o/'(x),/(x)6C7, i.e., Q($~I o /'(#), f ( x } } < l/2s. Since the functions / and $ o /' agree on ^, we are done. 0 Claim 6. If 7 6 T then f ' [ X ^ \K]C W(X~^ *). Proof, lix £ X~i\K then f ( x ) e X CX^ by (3) and hence (4)
$o/(x)
Pick A; such that x £ Pk. From the definition of /' we conclude by (4) that the first k coordinates of f ' ( x ) are in X^. Now observe that H is a deformation through X, and hence through Xy, and x,* G Xy. So /'(x) E W(-X" 7 ,*), as claimed. •(> Claim 7. For every 7 G F we have (f)'1^™]
\K = X^\K.
Proof. Pick an arbitrary point x G X-, \ K. By Claim 6,
Hence x G (/ / )~ 1 [^]\^Conversely, if x G (f')~l[X^°] \ K then x £ K and so x G Pfc for some element A; G N. But then f ' ( x ) k + 2 — x and so x G X7. <0 Now put g = $-1 o /'.
Claim 8. g is a Z-imbedding which agrees with / on K, is e-close to / and moreover satisfies (^[Xy] \ K = X^ \ K for every 7 G F.
360
5. BASIC INFINITE-DIMENSIONAL TOPOLOGY
Proof. That g is a Z-imbedding follows from Claim 4. In addition, g and / are e-close by Claim 5. If 7 E F is arbitrary then by Claim 7, g-i[X,} \ K = (/r1 [*[*,]] \ K C (/')- ^ \ = X7\£.
Conversely, if x E Xy \ K then by Claim 6,
From this we conclude that #(x) 6 X7 \ K.
0
So if * E .K" then K = K and we are done. Assume therefore that we are in the case that * ^ K . We ensured that /(*) E X \/[_K"] and that (*) = /(*)Observe that on the one hand for 7 E F, g~l[XJ \ (K U {*}) = (g-l[X,} \K)n(Q\ {*}), while on the other hand
This shows that But g(*) = /(*) E X C X7 and so * E g~l[X^]\K. Also, * E X\K The upshot is that g-l[X^]\K = X^\K, which is as required.
C X 7 \AT.
D
So we have: Theorem 5.5.15. The system 0 is 3a- absorbing and CQ is an 5FCT<5 -absorber inQ. Corollary 5.5.16. CQ is homeomorphic to B(Q)°° . Exercises for §5.5. ^•1. Suppose that there is a deformation of Q through A. In addition, let X be compact, let K C X be closed, and assume that /: X —> Q is continuous while moreover / \ K is a Z-imbedding. Finally, let e: X \ K —> (0, oo) be continuous. Prove that there is a continuous function g : X —>• Q such that (1) g\K = f\K, (2) g[X\K\CiA\g[K\, (3) for every x e X \ K we have Q(Q(X}, /(x)) < e(x).
5.5. ABSORBING SYSTEMS
361
^•2. Let A be strongly M-universal in Q, let X C M be such that there is a deformation of M through X. In addition, assume that Q x M a Q. Prove that A x X is strongly M-universal in Q x M. ^•3. Let B C Q be an 3~CT<5-absorber. Prove that there is an ^-absorber B1 C Q which is contained in B. To prepare ourselves for the applications in Chapter 6, we will identify a few other strongly 3~CT<5-universal sets in Exercise 5.5.5 below. Some preparatory work has to be done first before we can prove the results we are after. ^•4. Let A be an 3>-absorber and B a aZ-set with A C B C Q. Let C be a u-compactum in a compact space X. Prove that for each continuous function /: X —>• Q such that for some compact K C X, / \ K is a Zimbedding and for each e > 0 there is a Z-imbedding g : X —>• Q such that (1) g ( f , g ) < e , (2) g \ K = f \ K, (3) g-1[A]\K = C\K, (4) g[X\(C\JK)]r}B = ®. ^•5. For every n let En ~ Q. Prove that if Xn is an 3~CT-absorber which is contained in a aZ-set Cn C En for every n, then every set X with 00
00
Yl xn c x c Yl cn n=l
n=l
is strongly Sv^-universal in E — H^Li &"•• We will now prove some 'universality properties' of dense linear subspaces of R00. Proposition 5.5.17 below at first glance seems to be very technical. Its strength however is that the metric g under consideration beyond admissibility does not have any relation with the algebraic operations on L. For that reason a rather careful analysis is required. Observe that the function 6 in Proposition 5.5.17 is in some vague sense a simultaneous 'parametrization' of many straight-line segments. Proposition 5.5.17. Let L be a linear space with admissible metric Q. Then there is a continuous function 8: L x L x I —>• I such that for all t € I and x,p 6 L, (1) S(x,p,t) = 0 if and only ift = 0, (2) 6(x,p,t)
Proof. For all x,p € L and t e I put G(x, p, t) = {s 6 I : diam(p + [0, s ] - x ) >t}.
Observe that possibly G(x,p,t) = 0 and also that G(x,p, 0) = I. Claim 1. G(x,p,t) is closed in I for all x,p G L and t £ I. Proof. This follows easily from the continuity of the algebraic operations on L and the function diam (Exercise 1.11.12). 0
362
5. BASIC INFINITE-DIMENSIONAL TOPOLOGY
Now define £ : L x L x I —>• I by mmG(x,p,t) l
(G(x,p,t)?fy, (C(*,p,t) = 0).
By Claim 1, £ is well-defined. In addition, £(x,p, 0) = 0. Claim 2. £ is Isc. Proof. Suppose that for v £ [0,1), x,p £ L and £ £ I we have that £(x,p, £) > t>. Assume that there are sequences (x n ) n , (pn)n in L and (i n ) n in I such that
x n —>• x,
pn —>• p,
tn -* t
while moreover Sn = £ ( £ n , P n , £ n ) <
V
for all n. We will derive a contradiction. To this end, without loss of generality assume that sn —>• w < v. Observe that since v < 1, diam(pn + [0, sn] • xn) > tn for every n so that by continuity of the algebraic operations on L and the function diam (Exercise 1.11.12) we get diam(p + [0, s] • x) > t This shows that £(x,p, £) <w
•(>
iLxLxI-^Ib
Then £' is Isc by Exercise A. 7. 6. Let rj: L x L x I - ^ I b e the constant function with values 0. Then r\ is continuous, and hence use. By Corollary A. 7. 6 there is a continuous function 5 : L x L x I —> • ! such that for every (x,p,t) £ L x L x I with 0 < £'(x,p,t) we have 0 < S(x,p,t)
<£'(x,p,t).
Observe that if t > 0 and £(z,p, t) = 0 then 0 = diam({p» = diam(p + [0, 0] • x) > t > 0 which is impossible. So we conclude that if t > 0 then so are £(z,p, t} and £'(z,p, i) which implies that <5(x,p, i) > 0. Clearly, 5(z,p, 0) = 0. Observe that £'(x,p, t) < i for all a;, p £ L. As a consequence, 8(x,p,t) < t for all x,p G L. To check (3), pick arbitrary x,p G L and £ G I. If t = 0 then 5(x,p, t) = 0 and so there is nothing to prove. So assume that t > 0. By the above, £(z,p, £) > 0 so that by construction, 0 < 5(x,p,t) < £(x,p,t). This means that <5(x,p, £) 0 G(x,p, £), i.e., diam(p + [0, 5(x, p, i)] • x) < £. This means that g(p,p + 5(x,p, t)x) < t. We conclude that 8 is as required.
D
5.5. ABSORBING SYSTEMS
363
Corollary 5.5.18. Let Y be a dense linear subspace of R°°. Then for any admissible metric Q on R°° there is a homotopy H: R°° x R°° x I —>• R°° such that for aUx,y GR°° and t G I, (1) H0(x,y) = x, (2) if£ > 0 then Ht(x,y) G Y if and only if y G Y, (3) g(Ht(x,y),x)<2t. Proof. Since Y is a dense linear subspace of R°° it follows that (R°°, Y) is an AR-pair (Corollary 4.2.15). So by Proposition 4.1.7 there is a deformation
S: R°° x I ->• R°° such that
(4) S0 = 1, (5) St[R°°] C Y for every £ > 0, (6) g(St, 1) < t for every t G I. Let 5: R°° x R°° x I —>• I be continuous such that S(x,p, t) = 0 if and only if £ = 0 while moreover £>(p,p + 5(x,p, £)#) < £ for all £ G I and x,p € R°° (Proposition 5.5.17). Now define H: R°° x R°° x I ->• R°° by
#(*, y, t) = S(ar, t) + 5(y, S(x, t), t)y. We claim that H is as required. If t = 0 then H(x,y,0) = S(x,0) = x by (4). If y £ Y then clearly H(x,y,t) G 1" by (5) and the fact that Y is a linear subspace of R°°. In addition, if H(x,y,t) € Y and £ > 0 then 8(y,S(x,t),t) > 0 from which it follows that y G 1 since
<%, S(x, t), t}y = H(x, y, t) - S(x, t) G Y. It therefore suffices to verify (3). But this is trivial, since if x, y G M°° and t G I are arbitrary then Q(X, Ht(x, y ) ) < g(x, S(x, t)} + g(S(x, t), Ht(x, y ) ) < t + g(S(x, t), S(x, t) + 6 ( y , S(x, t ) , t ) y )
= It. So we are done.
D
Corollary 5.5.19. Let Y be a dense linear subspace of R°°. Then for any admissible metric Q on R°° there is a homotopy F: R°° x I —>• R°° such that
(1) F0 = l, (2) i f x G R°° and t > 0 then Ft(x) G Y if and only if x G Y, (3) g(Ft,l) < 2t for all t G I. Proof. Let H: R°° x R°° x I -» R°° be the homotopy from Corollary 5.5.18. It suffices to put F(x, t) = H(x, x, t) for all x G R°° and t G I. D
364
5. BASIC INFINITE-DIMENSIONAL TOPOLOGY We again represent the Hilbert cube Q by R00 . These results enable us to prove the following interesting result:
Theorem 5.5.20. Let (L 7 ) 7e r be a system of linear subspaces of R°°. Assume that the intersection H 7 er ^T JS dense in R°° and that there is a continuous function if) : Q ->• R°° such that T/T^L^nR 0 0 = L7 for each 7 G F. Then the system (L7 x R°°)7€r is refiexively universal in Q x Q. To start the proof of the theorem, let / = ( / i > / 2 ) : QxQ-^QxQbe continuous and let K C Q x Q be closed such that / \ K is a Z-imbedding. We let Q be an arbitrary admissible metric on Q which is bounded by 1 (Exercise A. 1.6). In addition, let 9 > 0. We may assume that / is a Z-imbedding by Theorem 5.3.11. Let d be the max metric on Q x Q that corresponds with g. Define e : Q x Q —> I by e(x) = min{d(f(x),f[K])/4:,9/4} and observe that £ -1 (0) = K since / is an imbedding. Since L = (~] L7 is a dense linear subspace of R°° , the pair (R°° , L) is an AR-pair by Corollary 4.2.15. In addition, (Q,R°°) is an AR-pair for the same reason. So by Corollary 4.1.8 it follows that (Q, L) is an AR-pair. There consequently is a deformation of Q through L. This implies that we can find a continuous function / i : Q x Q —> Q x Q such that 6(f1(x)Jl(x))<e(x)
for every x e Q x Q and with the property fi[(Q x Q) \ K] C L (Exercise 4.1.9). ^•6. Prove that there is a continuous £: Q x Q —>• I such that ^ -1 (0) = K and e ( f i (x, y), /i (a:, y) + f (ar, y}^(x}} < E(X, y ) for each (x, y) £ (Q x Q) \ K. Define gi : Q x Q -> Q by by ( I '«')
( ( * , » ) € if),
It is clear that g\ is well-defined. ^•7. Prove that g\ is continuous, that p(/i(aj),pi(a;)) < 2e(x) for every x in Q x Q and that (pf 1 ^] n (R°° x Q)) \ K = (L7 x Q) \ K for every 7. Let 5 denote the a-compact subset (Q x Q) \ (R°° x R°°) of Q x Q (it is even a capset of Q x Q by Exercise 5.4.1, but this is not needed in the proof). By Exercise 5.4.10 there is a continuous function Q2'- Q x Q —> Q such that 0(32(2;), / 2 (a;)) < £C*0 for every x e Q x Q,
5.5. ABSORBING SYSTEMS
365
while moreover g? \ (Q x Q] \ K is one-to-one, gi [(Q x Q) \ K] e £ CT (Q), and 52-
1
[5(g)]\X = J B \ K
Put # = (91,92). ^•8. Prove that ^ is a Z-imbedding, that # f /C = / |" K and that rf(/(x),^(x))
< 2e(x)
for every x £ Q x Q. The map satisfies by Exercise 5.5.7 and the properties of the map 92 for every 7 £ F the following: ^[L-r x M°°] \ K = (g^[L,] n 52"1[M00]) \ K
So this completes the proof of Theorem 5.5.20 since clearly d(f,g) < 9.
This page intentionally left blank
CHAPTER 6
Function spaces In this chapter we will discuss the topology of pointwise convergence on function spaces. During the last decades these function spaces have played a prominent role in both general- and infinite-dimensional topology. The systematic study of them was initiated by ARHANGEL'SKII [24]. Techniques from general topology, infinite-dimensional topology, functional analysis and descriptive set theory are primarily used for the study of function spaces. The mix of methods from several disciplines makes the subject particularly interesting. Several monographs, surveys and research papers were written containing many interesting and diverse results, see e.g., ARHANGEL'SKII [25, 27, 28], LUTZER and McCov [264], McCov and NTANTU [280], VAN MILL [297], BAARS and DE GROOT [35], DOBROWOLSKI, MARCISZEWSKI and MOGILSKI [123], CAUTY [88], CAUTY, DOBROWOLSKI and MARCISZEWSKI [91] and MARCISZEWSKI [268]. For a Tychonoff space X let CP(X) be C(X) endowed with the topology of pointwise convergence. The main result in this chapter is Theorem 6.12.15 which implies that if X is nondiscrete and countable and CP(X] is an Fassubset of Rx then CP(X] is homeomorphic to B(Q}°°. Many results of the previous chapters are applied in its proof. To demonstrate its power, observe that it implies that Q and co + I have homeomorphic function spaces. We will also discuss linear homeomorphisms. It is possible to prove that certain topological properties are shared by spaces having linearly homeomorphic function spaces. We will give three such examples: compactness, dimension and topological completeness. This shows for example that the function spaces of Q and uj+l are not linearly homeomorphic. An important recent result that we will prove is the statement that for metrizable spaces X and Y, if X is countable dimensional and CP(X) « CP(Y) then Y is countable dimensional. This proves that if X is any finite dimensional metrizable space then the spaces CP(X) and CP(Q) are not homeomorphic. All spaces under discussion here are Tychonoff, i.e., completely regular and TI .
367
368
6. FUNCTION SPACES
6.1. Notation For a Tychonoff space X let CP(X) be C(X) topologized as a subspace of the full product Rx endowed with the Tychonoff product topology. So CP(X) is endowed with the topology of pointwise convergence. The spaces CP(X) are of interest to topologists and functional analysts for various reasons. It is possible to restricts one's attention to bounded functions only. For a space X, C*(X) endowed with the subspace topology of Rx is denoted by C*(X). Note that C*(X) is a subspace of CP(X). If X is compact then C(X) — C*(X) so then there is no difference. In this chapter we are primarily interested in the spaces CP(X). The theory of bounded functions is more complicated and is less well understood. We will demonstrate this with a few examples only, cf. §6.10. Observe that basic neighborhoods of / 6 Rx have the form N ( f , F , e ) = {g € R* : (Vx e F ) ( \ f ( x ) - g(x)\ < e)}, where F is an arbitrary finite subset of X, and e is an arbitrary positive real number. N ( f , F , e ) is called the basic neighborhood of f determined by F and e. If X is a countable space which is not necessarily metrizable, then CP(X) is a linear subspace of the product Rx . So it is a linear subspace of a countable product of real lines, and is therefore a separable metrizable space. The space X can be rather 'bad' from a general topological perspective, its function space is always 'good'. As we will see, certain non-metrizable countable spaces X have very interesting function spaces and this prompts us to study non-metrizable spaces as well. This means that we sometimes have to be more careful than before. The classical Tietze-Urysohn Extension Theorem (ENGELKING [153, Theorem 2.1.8]) is no longer valid for example since a Tychonoff space need not be normal. But at least a portion of it is true since if X is Tychonoff and A C X is compact then every continuous function /: A —>• R can be extended to a (bounded) continuous function /: X —> R. This can be seen as follows. First think of X as subspace of its Cech-Stone compactification J3X. Then A, being compact, is closed in /3X. Since /3X is compact and Hausdorff, it is normal, and so by the Tietze-Urysohn Extension Theorem, / can be extended to a (necessarily bounded) continuous function /: (3X —> R. Hence / = f \ X is the required (bounded) extension over X. We will use this in this chapter in the case that A is a finite set often without explicit reference. Observe that /3X is characterized by the fact that it is the unique compactification having the property that every bounded and continuous function /: X —>• R can be extended to a continuous function /: /3X —>• R (cf. ENGELKING [153, Corollary 3.6.3]).
6.2. THE SPACES C P ( X ) : INTRODUCTORY REMARKS
369
If X is a space then X* denotes j3X \ X. We will also frequently use without explicit reference the well-known fact that metrizable spaces are paracompact. This is due to STONE [383]. A simpler proof was found by RUDIN [359]; she in fact proved by a direct construction avoiding theory on paracompact spaces that every open cover of a metrizable space has an open refinement which is both locally finite and a-discrete, i.e., the union of countably many discrete subfamilies. See ENGELKING [153, Chapter 4] for additional information. 6.2. The spaces CP(X}: Introductory remarks
Since CP(X) is a dense linear subspace of M.x, it consequently inherits many properties from M.x. For instance, its local convexity, its weight (see Corollary 6.2.2 below) and cellularity, etc. Lemma 6.2.1. Let X be a Tychonoff space. Then CP(X} is dense in M.x. Moreover, C*(X] is dense in CP(X}. Proof. Let / e Rx, F C X be finite and £ > 0. The function g = f \ F can be extended to a bounded continuous function g: X —> M since F is finite. This function is in the basic neighborhood of / determined by F and E. So this proves both statements simultaneously. D If X is a space then its weight is the smallest cardinality of a base for X. Observe that if X is infinite then the weight of Rx is equal to the cardinality of X. In addition, if Y C Rx is dense, then Y and M.x have the same weight. Corollary 6.2.2. If X is infinite then the weight of CP(X) is equal to the cardinality of X. As a consequence, if X and Y are infinite such that CP(X] and CP(Y) are homeomorphic then \X\ = \Y . Corollary 6.2.3. Let X be a space. The following statements are equivalent: (1) (2) (3) (4)
CP(X) is separable and metrizable. CP(X] is metrizable. CP(X) is first countable. X is countable.
Proof. If IRX contains a dense first countable subspace then it is first countable at some point. This implies that X is countable by the definition of the product topology on RA . So we are done by Corollary 6.2.2. D If L and M are linear spaces then L ~ M means that L and M are linearly homeomorphic. This means that there is a bijection (p: L —>• M which is simultaneously a homeomorphism and a linear isomorphism.
370
6. FUNCTION SPACES
i- and t-equivalence. Topologists are interested in the classification of topological spaces. Classifying topological spaces up to homeomorphism or homotopy type etc., is the ultimate goal for a topologist. In this chapter we classify spaces X up to homeomorphism or linear homeomorphism type of their function spaces CP(X). In its full generality this program would be much too complicated and the complete picture is presently beyond our reach. But for function spaces of low Borel complexity some definitive results are known, and it is our aim to present those in full detail. Many results of the previous chapters will be applied in this chapter. We say that spaces X and Y are l-equivalent provided that CP(X) and CP(Y) are linearly homeomorphic. Notation: X~iY. We define the corresponding notion of I* -equivalence similarly. Notation: X~i*Y. Homeomorphic spaces are obviously I- and i* -equivalent. But the converse need not be true. Example 6.2.4. There are non-homeomorphic ^-equivalent spaces. Proof. Let X = [0, 1] U [2, 3] and Y = [0, 2] U {3}. Then evidently, X and Y are not homeomorphic. We will prove that they are ^-equivalent. Indeed, define $: CP(X) -» CP(Y) by
(o < y < i),
f(y}
It is left as an exercise to the reader to prove that $ is a linear homeomorphism. D If X is compact then C(X) can also be endowed with its natural Banach space topology (see Page 4). It is clear that the identity C(X) —> CP(X) is continuous. This has by Corollary 1.1.16 the following interesting consequence. Theorem 6.2.5. Let X and Y be compact spaces, and let ip: CP(X] -+ CP(Y) be a linear homeomorphism. Then V\ C(X) -> C(Y)
is also a linear homeomorphism. The converse of this result is not true. Let us note that MILJUTIN [290] proved that all Banach spaces C(X) with X uncountable and compact metrizable are linearly homeomorphic. Hence C(I) and C(I2) are linearly homeomorphic, but CP(I) and CP(I2) are not, as we will see in §6.9 that ^-equivalent spaces have the same dimension.
6.2. THE SPACES CP(X): INTRODUCTORY REMARKS
371
Corollary 6.2.6. There are countably infinite compact (necessarily metrizable) spaces X and Y which are not l-equivalent. Proof. By BESSAGA and PELCZYNSKI [48] there are such spaces X and Y for which C(X) and C(Y) are not linearly homeomorphic (see also Corollary 6.10.9). So now an appeal to Theorem 6.2.5 finishes the proof. D We say that X and Y are t-equivalent provided that CP(X] are CP(Y) are homeomorphic as topological spaces. Notation: X~tY. Remark 6.2.7. Even for simple spaces it is in general difficult to decide whether they are £- or ^-equivalent. We will show in §6.12 that all countable non-discrete spaces X such that CP(X] is an F^-subset of Rx are tequivalent. This result implies that all countable non-discrete metrizable space are ^-equivalent (Theorem 6.3.10). But they are not all ^-equivalent, as was shown in Corollary 6.2.6. Products. We will prove that the spaces CP(X) behave well with respect to products. If {Xi : i G /} are spaces then @ i67 ^Q denotes their topological sum. Theorem 6.2.8. Let {Xx : i G /} be a family of spaces. Then
If I is finite, then C;(0l£/ Xx] ~ fU/ C;(Xi). Proof. It is left as an exercise to the reader to show that the function
defined by F(f)i = f \ Xi, i e /, is as required. This function also works for bounded functions if 7 is finite. D Observe that the same proof does not work for bounded functions if I is infinite. The situation for such functions is in fact much more complicated, see the notes for more information. Additional structures on CP(X). With the operations of pointwise addition and pointwise multiplication, CP(X] is a commutative topological ring with unit, the unit being the constant function with value 1. It is a famous theorem of GEL'FAND and KOLMOGOROFF [173] that the ring structure by itself determines the topological structure on X provided X is compact. They proved that if X and Y are compact and C(X) and C(Y] are isomorphic as rings then X and Y are homeomorphic. For details, see also
372
6. FUNCTION SPACES
DUGUNDJI [142, Theorem XIII.6.5]. (The proof in [142] makes use of the topology of pointwise convergence.) For noncompact spaces X, the algebraic structure of C(X) is, in general, not strong enough to determine the topology of X. For consider the spaces X = uii and Y = wi + 1. Then clearly C(X) and C(Y) are isomorphic as rings, but X and Y are not homeomorphic. For arbitrary spaces there is a result in the same spirit though. NAGATA [322] proved that CP(X) and CP(Y) are topologically isomorphic as topological rings if and only if X and Y are homeomorphic. That we deal with real-valued functions is essential in this result. It is known that the ring of all continuous functions X —> E°° , endowed with the topology of pointwise convergence, does not always determine the topological type of X (ARHANGEL'SKII [27, Page 12]). In this book we will not discuss these structures and other similar ones. We will study the topological and linear topological properties of CP(X) exclusively. 6.3. The Borel complexity of function spaces
Obviously, X is discrete if and only if CP(X) = RA . The question naturally arises how CP(X) is placed in E.x if X is not discrete. We show here that it cannot be a Gsa- or F^-subset of M.x but it can be an FCT ,5 -subset CP(X) is not G§. We will first show that CP(X) cannot be a G^-subset of Ex unless X is discrete. This result will be used in the proof of our main results. Lemma 6.3.1. Suppose that CP(X) contains a nonempty G$-subset of M.x . Then X is the topological sum of a countable space and a discrete space. Proof. Let S be a nonempty G^-subset of M.x which is contained in CP(X). Pick an arbitrary element / G 5. Let {Un : n G N} be a sequence of open subsets of M.x such that fl^Li Un = S. For every n there exists a finite subset Fn C X such that every g £ Rx with g \ Fn — f \ Fn belongs to Un. Put F = USJLi Fn. Then F is countable, and every g G E.x with g\F = f\F belongs to H^Li Un — S C CP(X) and hence is continuous. Let A be any subset ofX\F and let XA '• X —>• {0, 1} be the characteristic function. Since XA \ F = 0 it follows that So by the above, / -I- XA is continuous from which it follows that
6.3. THE BOREL COMPLEXITY OF FUNCTION SPACES
373
is continuous as well. Consequently, every A C X \ F is clopen in X .
D
Lemma 6.3.2. Let X be countable. If CP(X) contains a dense Gs-subset of Rx then X is discrete. Proof. Assume that X is not discrete. There is a function g e M.x \ CP(X). Define $ : Rx ->• Rx by Then $ is a homeomorphism of Rx onto itself and $[CP(X)] subspace of Rx \CP(X). Thus
is a dense
Rx \CP(X} also contains a dense G^-subspace of M.x , which violates the Baire Category Theorem. Simply observe that since X is countable, Rx is topologically complete (Lemma A. 6. 2 and Exercise A. 6. 10). D Corollary 6.3.3. If X is countable and CP(X) is both a Baire space and a Borel subset of M.x then X is discrete. Proof. By Exercise A. 13. 4, if CP(X) is both Borel and a Baire space then it contains a dense topologically complete subspace. Since a topologically complete subspace of M.x is a G^-subset of it (Theorem A. 6. 3), the previous lemma shows that X is discrete. D These results lead us to our first result on the descriptive complexity of CP(X) for general X. Theorem 6.3.4. IfCp(X)
is a G$-subset ofE.x then X is discrete.
Proof. By Lemma 6.3.1 it follows that X = F 0 Z), where F is countable and D is discrete. Since CP(X) is canonically equivalent to the subspace CP(F) x RD
of RF x RD = Rx (Theorem 6.2.8), it follows that CP(F) is a G5-subset of RF . So F is discrete as well by Lemma 6.3.2. D Remark 6.3.5. The basic ingredient in the proof of Lemma 6.3.2 is that M.x is a Baire space if X is countable. But M.x is a Baire space for arbitrary X as well. This follows easily from the fact that Rx satisfies the countable chain condition from which one gets that every nowhere dense set in Rx is contained in a nowhere dense G^-set. This allows a reduction to the known case of countable products. (Alternatively, one can use the observation in Remark 6.4.4 below or e.g., DE GROOT [178].) So the proof in Lemma 6.3.2 in fact shows that for arbitrary X, if CP(X) contains a dense GVsubset o then X is discrete.
374
6. FUNCTION SPACES
CP(X) is not Fff. In view of Theorem 6.3.4 the question naturally arises whether CP(X) can be an F^-subset of Rx . Theorem 6.3.6. If CP(X) is an Fa-subset of Rx then X is discrete. Proof. Assume that X is not discrete, say that XQ is a non-isolated point of X, and that U^Lo ^n — CpPO? where each Fn is closed in Rx and FQ = 0. By induction on n we will construct a function fn : X —> I and an open neighborhood Un of XQ such that the following conditions are satisfied: (1) /o < A < / 2 <••_!> (2) U0DUiDUi2U2DU2D---, (3) fn(x0)
= 1,
(4) fn\Un\{x0}
= l-2-n,
(5) fn \ X \ {XQ} is continuous, (6) fn+l
\X\Un = fn,
(7) if / e Rx is such that f\(X\Un)\J then / ^ Fn.
{XQ} = /„ \ (X \ Un) u {x0}
Define /o by /Q(XQ) — 1 and /o(#) = 0 if x ^ XQ, and put UQ = X. Suppose that we completed the construction for 0 < i < n. Let g : X —> [0, 2~n] be continuous such that g(xo) = 2~n and g[X \ Un] C {0}. Define fn+i by -(n+l)jp(a.)}
(x
^
XQ).
The set V = g~l [(2-( n+1 ),2~( n - 1 )]] is a neighborhood of x0 such that fn+l \ V \ {x0} = l-2'n + 2-( n+1 ) = 1 - 2-(n+i\ Since x0 is not isolated, fn+i is not continuous at x 0 , hence /n+i ^ F n+ i. Since Fn+i is closed in Rx there exists a finite subset A of X such that for each / e ffix, if / \ A = fn+i \ A then / does not belong to Fn+i. As a consequence, Un+i = (V\A)\J {XQ} is the required neighborhood of XQ. This completes the induction. Put / = linin^oo fn. each n,
Observe that / exists because of (1). Then for
f\(X\Un)\J
{X0} = fn\(X\
Un] U {Zo}.
By (7) we therefore obtain that / £ U^°=0 Fn. But / is continuous, as the following argument shows. If x is a point of X which does not belong to U = fl^Lo ^n then by (2), x 0 Un for some n. So by (6) and (5), / is continuous at x. But if x G U then Un is a neighborhood of x such that f[Un} C {1 — 2~ n , 1} by (3) and (4). Hence / is continuous at x by (1), (5), (6) and (3). So we conclude that / G CP(X) \ (J£L0 F ™-
D
6.3. THE BOREL COMPLEXITY OF FUNCTION SPACES
375
CP(X) is not Gsa- We will now use Lemma 6.3.2 to generalize Theorem 6.3.4. Let X be a space, and let Y C E. By CP(X, Y) we denote the set of continuous functions /: X -> Y, topologized as a subspace of the full product Yx . Since Yx is a subspace of Rx , it follows that CP(X,Y) is simply the subspace {/ 6 CP(X) : f[X] C Y} of CP(X). Let X be a space and e > 0. We put ) = {9£ CP(X] : (Vz e X)(|/(ar) - g(x}\ < e}}. Lemma 6.3.7. Let X be a countable space, let f 6 C'p(X) and Jet £ > 0. Tnen (1) C/(/,e) is a G5-subset of CP(X). (2) C/(/, e) and CP(X) are homeomorphic. (3) Uo< 5<£ ^(/, *) ^ ^nse in Proof. That U(f,e) is a G^-subset of C'p(X) is clear since for every x G X the set {9eCp(X): g(x)-f(x)\<e} is open in CP(X) and X is countable. To show that CP(X) and U(f,e) are homeomorphic, first consider the translation f : Ex ->• Ex defined by This function maps C7(/, e) onto C/(0, e:), where 0 is the constant function with value 0. So it suffices to show that this set and CP(X) are homeomorphic. But by observing that (— e,e) and E are homeomorphic, this follows by the remark preceding this lemma. It remains to prove (3). It suffices to prove that Uo<5<e ^(Q,^) is dense in V = U(0, e). So let g <E V, F C X be finite, and 7 > 0. Pick 0 < A < 1 such that e(l — A) < 7 and put h — \g. Then ,~f)n
\J
[7(0,5).
This is clearly as required.
D x
Theorem 6.3.8. If CP(X) is a Gsff-subset ofR
then X is discrete.
Proof. The proof is inspired by the proof of Theorem A. 13. 5. As above we may assume without loss of generality that X is countable. So let X be countable and non-discrete such that oo
CP(X) = |J G,, i=0
376
6. FUNCTION SPACES
where G; is a G^-subset of M.x for every i and GO = 0. We will derive a contradiction. We shall by induction on i define a sequence {/; : i > 0} in CP(X] and a decreasing sequence of positive real numbers {<£; : i > 0} such that (1) (2) (3) (4)
/o is the constant function with value 0, and £Q = 2°, U(fo, y^o) 5 U(fi,£i) D U(fi, y^i) ^ U(f2,£2) ^ ••• 2:^+1 < 3"1"1 • £j for every i, U(fi,£i) n Gi = 0 for every i.
5
Assume that we have proved the existence of such sequences. Observe that by (1) and (3), £i < 2~l for every i. So for every i we have by (2) and (3) that We conclude that the sequence (fi)i converges uniformly to a continuous function /: X -> R (cf. Lemma A.3.1). It follows moreover from (3) that for x G X, i > 0 and n > 2 we have L I f.(r\ _ Jn+i\-^)\ f , .(r\\ _^ < \Ji\' I f - ( Lr )} — Ji+l\i• , i fr")! \Ji\^J )\ 4i iI Ji-i +, l!Vf ^r V /
n-l < V4£i + /
3
£i.
This implies that sup|/i(x) -/(a;)| < for every i, i.e., / € flSo W*^) ^ C'P(^) \ USi G < = 0' which is a contradiction. It remains to perform the induction. Assume that fa and £i have been determined. By Lemma 6.3.7 it follows that C/(/j, l/^£i) is a G^-subset of CP(X) which in addition is homeomorphic to CP(X). The latter implies by Lemma 6.3.2 and Theorem A. 6. 3 that no dense subset of £/(/;, l/±£i) is a (7,5-subset of Mx . The former implies that Gi+\ fi f/(/i, V4£) is a G^-subset of Gi+i and hence is a G^-subset of Rx . The conclusion is that is not dense in £/(/;, V4 £ )- S° there are by Lemma 6.3.7(3) an element for some 0 < 8 < l/±£i, a finite set F C X and a 7 > 0 such that i,F,7) n (Gt+1 n U(ft, y4eO) = 0-
6.4. THE BAIRE PROPERTY IN FUNCTION SPACES
377
Now let £i+i be a positive real number smaller than min{7, y^ - <J, 3"'"1 • el}. It is clear that fi+i and £i+i satisfy the inductive hypotheses.
D
Remark 6.3.9. Observe that for countable spaces, Theorem 6.3.6 is a consequence of Theorem 6.3.8. CP(X) can be Fa&. We saw that unless X is discrete, CP(X) cannot be Fa or G&ff in Rx . But it can be Fffg in Rx .
Theorem 6.3.10. Let X be a countable metrizable space. Then CP(X) is an Fas-subset ofRx . Proof. Let Q be an admissible metric for X, and let x £ X. Then the e-S definition of continuity shows that CP(X) is the set oo
oo
PI n U {9 e ^X : g[B(x, y m )] C \g(x) - ^/n,g(x) + i/n]}. x£X n=l ra=l
Since X is countable and each set {g 6 R* : 0[B(z, i/m)] C \g(x) - i/n,g(x) + i/n]} x
is closed in R , C P (X) is indeed an .F^-subset of R x , as claimed.
D
So we conclude that all familiar countable metrizable spaces, such as the rational numbers or a convergent sequence with its limit, have function spaces of the same Borel complexity. In fact, as we will see later by using the machinery on jF^-absorbers developed in Chapter 5, they are all homeomorphic. The metrizability is not essential in this result. There are examples of countable, regular, nonmetrizable spaces X for which CP(X) is an F^-subset of E.x . We will come back to this later (see Example Ell in §6.13). 6.4. The Baire property in function spaces Being a Baire space is an important topological property for a space and it is therefore natural to ask when function spaces are Baire. We already saw in Corollary 6.3.3 that if X is countable and nondiscrete then if CP(X) is Baire it is not a Borel subset of Rx . This indicates that Baire CP(X) are rather rare. A family of subsets A of a, space X is called strongly discrete provided that every A G A has a neighborhood U(A) such that the family
{U(A} :AeA} is discrete.
378
6. FUNCTION SPACES
Lemma 6.4.1. Let X be a space and suppose that CP(X] is Baire. Then every pairwise disjoint sequence of finite subsets of X has an infinite strongly discrete subsequence. Proof. Let A be an infinite sequence of pairwise disjoint finite subsets of X. If A is finite then there is nothing to prove, so assume it is infinite. For each n 6 N let Gn = {/ € CP(X) :(3Ae A)(f[A]
C (n,n + 1/2))}
Then Gn is clearly an open subset of CP(X] since every A G A is finite. We claim that it is also dense in CP(X). To this end, let / e CP(X), F C X be finite and e > 0 be arbitrary. Since A is infinite and pairwise disjoint and F is finite there is an element A e A such that Ar\F = 0. Define /: AL)F ->• E by
fix} = and let g : X —>• M be an arbitrary continuous extension of /. Then clearly and so Gn meets all nonempty basic open sets in CP(X). By assumption, CP(X) is a Baire space and so there is an element oo
/ e f| Gn. n-1
For each n pick An E A such that f[An] C (n,n + 1/2). Since the family {(n,n+ 1/2) : n e N } is discrete in R and / is continuous, the same is true for the family irl[(n,n+%)}:neN}. We conclude that the subsequence {An : n 6 N} of A is strongly discrete.
D
Interestingly, the converse to Lemma 6.4.1 is also true. So there is a purely topological characterization of the spaces X for which CP(X) is a Baire space. Before we can prove this, we first formulate and prove the following result. Lemma 6.4.2. Let J be a closed and bounded interval in E and let be continuous. In addition, let A C X be closed, PCX finite, E > 0 and T : F —>• J be such that T(X] — f ( x ) < e for all x e F n A. Then there is a continuous function g-.X-^J such that g \ F = T and \g(x) — f ( x } \ < e for all x £ A.
6.4. THE BAIRE PROPERTY IN FUNCTION SPACES
379
Proof. The function /: X —> J can be extended to a continuous function /: /3X ->• J. We let A denote the closure of A in J3X. Extend the function r-f \FC\A: FftA-> (-£,£) to a continuous function h: A -> (— e,e) (use that F is finite). Since fiX is compact and Hausdorff it is normal from which it follows that the function (f+h\^)\J(T\F\A) can be extended to a continuous function g : f3X -> J. Then g = g \ X is clearly as required. D It will be convenient to introduce the following notation. If F ^ 0 and is the basic neighborhood of / G M.x determined by F C X and e > 0 then we put S(U) = F, m(U) = snp{\g(a}\ :g^U,a&F}. Observe that 0 < m(U] < oc, and also that N ( f , F , e ) = N(g,G,6) if and only if F = G, e = 6 and / \ F = g \ G. Theorem 6.4.3. Let X be a space. The following statements are equivalent: (1) CP(X) is a Baire space. (2) Every pairwise disjoint sequence of finite subsets of X has an infinite strongly discrete subsequence. Remark 6.4.4. Since a discrete space obviously has the property that every pairwise disjoint sequence of finite sets has a strongly discrete subsequence, a corollary to this this result is the nontrivial fact that M.x is a Baire space for arbitrary X. Proof. By Lemma 6.4.1, it suffices to prove the implication (2) =>• (1). For that it suffices by Exercise 1.6.6 to prove that CP(X) is not meager in itself. So let X be a space having the property stated in (2), and let [Fn : n G N}
be a sequence of nowhere dense subsets of C P ( X } . Since CP(X) is dense in Ex , each Fn is nowhere dense in R^ as well. We may assume without loss of generality that Fl C F2 C . . - C Fn C . . . . We define three sequences by induction as follows: an increasing sequence
{Sn : n G N} of finite subsets of X , an increasing sequence {mn : n G N} of positive numbers, and a collection {Un : n G N} of finite families of basic open sets in Rx such that, for each n, the following conditions are satisfied:
380
6. FUNCTION SPACES
(i) for every U G Un there exists an element U' E lin+i such that U' CE7, (ii) U n Fn = 0 for all U E U n , (iii) 5(C7) C Sn for all [7 E U n , (iv) ra(C7) < m n for all C7 E U n . (v) for all / E [—m n ,m n ] x there exists an element U E U n+ i such that Since F\ is nowhere dense, there is a nonempty basic open set U E Rx such that UnFi = 0. Put 5i = 5(C7), mi = m(U) and Ui = {f/}. Suppose that Sn: mn and lin have been defined satisfying (ii) through (v). Let Z be the compact cube [— mn,mn]x . There consequently exists a finite G C Z such that Consider the finite collection For every H E ^K let UH be a nonempty basic open subset of Rx such that UHCH\ Fn+l (here we use that Fn+i is nowhere dense). Now put 5n+1=5nU
S(UH),
m n+ i = mn Un+1 ={UH:He W}, respectively. Claim 1. These choices satisfy the inductive requirements. Proof. It is clear that (i) through (iv) are satisfied. To check (v) for n + 1, let / € [— mn,mn]x . There exists g 6 G such that Pick an arbitrary h 6 UH and let x € Sn. Then
\f(x) - h(x)\ < |/(:r) - g(x}\ + \g(x) - h(x)\
We conclude that UH C 7V(/, 5n, Yn) and since UH G U n +i, we are done. 0 The pairwise disjoint sequence (5n+i \ Sn : n € N}
6,4. THE BAIRE PROPERTY IN FUNCTION SPACES
381
has by our assumptions on X a strongly discrete subsequence, say, {Snk+1 \Snk:k£ N}. We may assume without loss of generality that rik+i > max{nfc + 2, 2k + 1} for every k. For each k G N let T%k-i — SVifc, -i = m = m n f c + i,
respectively. We claim that the following statements hold: (1) (2) (3) (4)
V n Fn = 0 for all V € V n , S(V) C Tn for all V e V n , /[5(F)j C [-Mn, Mn] for all / € V € V n , if / 6 [-Mn,Mn]x then there is an element V 6 V n+ i such that
Statements (1) through (3) are easily verified. Simply use that all sequences considered so far are increasing. For (4), let / € [— Mn, Mn]x and assume that n = 2k (the case for odd n follows by a similar argumentation). By (v) there exists an element U e U nfc+ 2 such that C/C7V(/,S n f c + 1 ,Vn f c + i). Since
> n/s + 2 it follows that Tn = S-2k ^ Snk+i- Moreover, Vn fc + l < l/2k = Vn,
so that UCN(f,Tn,
i/n),
as required. Let {H^fc : A; 6 N} be a discrete family of open subsets of X such that for each k:
(5)
\ r2jb _i c
(6)
_! n
For each k let (7) £>2* =
= 0.
382
6. FUNCTION SPACES
We will now define an increasing sequence {jn : n G N} of positive even integers, a sequence {Vn : n G N} of basic open subsets in EA , a sequence {fn '• n € N} of elements of CP(X) and, finally, a sequence {en : n G N} of positive real numbers, having for each n the following properties:
(10) fn£[-Mjn,M x (11) \fn+i(x) - fn(x}\ < en for all x G £>,„. To begin the induction, let jl = 2, let Vi G V2 and / G Vi. By (3), By the Tietze-Urysohn Extension Theorem there is an /i G C'p(X) such that /i G [-M 2 ,M 2 ] X and /i f S(Vi) = f \ 5(Vi). Then /i 6 Vi and since S(Vi) C T2 there exists e\ > 0 such that JV(/i,T 2 ,3ei)CVi. Assume that jn, V n , /n and en have been defined satisfying (8) through (11). Then let jn+i be an even integer greater than max{jn, 1 + l/e n }- By (10), (4), (i) and the fact that jn+i — jn > 2 there exists a Fn+i £ ^jri.+i sucri that Pick an arbitrary element / 6 Vn+\. By (5) and (7), (12)
Tjn+1 \T j n + 1 _! C W jn+1 C X \
By (3),
and by (10), /n 6 [-M3n,Mjn}x C [-Mjn+1, MJn+1]x. We want to apply Lemma 6.4.2 with the closed set A = Djn and the functions fn and Observe that if x 6 Djn n T Jn+1 then by (12), x G T Jn+1 _i so that since
/ we get
jn+i - I)So by Lemma 6.4.2 there is an element fn+i £ C*p(^) such that while moreover |/ n+1 (x) -
fn(x)
for all x G jDJn and r- T\
_ f i- T\
I -ijn + l
— J
I
J
Jn+ l -
6.4. THE BAIRE PROPERTY IN FUNCTION SPACES
383
Then fn+i € Vn+i and so there exists an 0 < e n +i < £n/2 such that N(fn+i,Tjn+1,3en+i)
C Vn+i.
This completes the inductive construction. Condition (11) can be modified to say that (13) \fm(x) — fn(x)\
< %£k whenever m,n > k and x G Djk.
This follows from (11) and (8) since \fm(x)
- fn(x)\
= fm(x)
- fm-i(x]
< \fm(x)
~ fm-l(x)\
+ ---- h fn+l (x) ~ H ----- h \fn+l(x)
~
fn(x)\ fn(x)\
In particular, (13) says that for each k, {fn : n e N} is uniformly convergent on Djk . Therefore, {fn : n £ M} converges pointwise to an / 6 Rx such that for every k the function / \ Djk is continuous (cf. Lemma A. 3.1). Since the collection {W2k • k e N} is discrete, / e CP(X). Finally, we show that f £ Vn for each n. To this end, let n be fixed and let x 6 5(K). By (2), (6) and (7), x 6 Tjn C Djn. If m > n then by (13), I\Jm\f (LT)\ — Jn\f (Lr)\ \\
This means that \f(x}- fn(x)\
<2en<3en.
So by (9), f£N(fn,Tjn,3en)CVn. So by (1) we get that
n=l
as required.
D
Corollary 6.4.5. Let X be a space for which CP(X] is Baire. If Y is a subspace of X then CP(Y) is Baire. Proof. This follows directly from Theorem 6.4.3 since the property that every sequence of pairwise disjoint finite sets has a strongly discrete subsequence is hereditary. D
384
6. FUNCTION SPACES
So a Baire CP(X) has many Baire subspaces. The question arises whether there is a nondiscrete space X whose function space CP(X) is hereditary Baire. For countable spaces X this is determined by the question whether Q admits a closed imbedding in CP(X] (Corollary 1.9.13). There is a consistent example of such a space, see Page 393 below. We now turn our attention to the Baire property in products of function spaces. Corollary 6.4.6. Let X and Y be such that both CP(X) and CP(Y] are Baire. Then CP(X] x CP(Y) is a Baire space. Proof. By Theorem 6.2.8, CP(X] x CP(Y) is homeomorphic to CP(X © Y}. Since X and Y both have the property in Theorem 6.4.3(2), it is clear that X © Y has this property as well. So we are done by the characterization result Theorem 6.4.3. D Observe that virtually the same proof shows that if {Xi : i G 1} is an arbitrary family of spaces such that Cp(Xi) is a Baire space for every i then the space Y\ieICp(Xi) is a Baire space as well. The precise verification of this is left as an exercise to the reader. In Theorem A.6.10 we proved that the product of two separable metrizable Baire spaces is again a Baire space. It is not possible to generalize the obvious proof of that result to metrizable spaces. The question whether the product of two metrizable Baire spaces is again a Baire space was raised in CHOQUET [99], OXTOBY [331] and SIKORSKI [374]. Using forcing techniques it was answered in the negative by COHEN [101]. It was known earlier from the work of OXTOBY [331] and KROM [232] that the Continuum Hypothesis implies the existence of two metrizable Baire spaces whose product is not Baire. In FLEISSNER and KUNEN [163] such spaces were constructed in ZFC alone. They even constructed in the same paper an example of a single metrizable Baire space whose square is not Baire. In view of Corollary 6.4.6 the question arises whether the product of linear Baire spaces is Baire. If so, then the corollary would be a consequence of the fact that CP(X) is a linear space. But there are even normed Baire spaces whose product is not Baire. This is due independently to VAN MILL and POL [302] and VALDIVIA [92]. The conclusion is that the spaces CP(X) are linear spaces with a particularly rich structure. (This conclusion is not very surprising.) We now present some examples of spaces X for which CP(X) is Baire, or not.
A countable dense-in-itself space X such that CP(X) is Baire. There is according to VAN DOUWEN [130, 134] a countable space X without
6.4. THE BAIRE PROPERTY IN FUNCTION SPACES
385
isolated points having the following curious property: no point of X is simultaneously an accumulation point of two disjoint subsets of X (hence X is perfectly disconnected, see the notes of §A.l and Exercise A. 1.2 for more information). We will show that this space X has the property that its function space CP(X) is a Baire space. So CP(X) can be Baire without X being Baire. We first claim that X is a so-called nodec space, that is, a space in which every nowhere dense set is closed. To prove this, let A C X be nowhere dense. If A is not closed then there exists an x G A \ A. Then clearly x E A\{x} which implies that there is an open neighborhood U of x which is contained in A U {x} (Exercise A. 1.2). Since X has no isolated points this implies that U\{x} is a nonempty open subset of X which is contained in A, which contradicts the fact that A is nowhere dense. Observe that in a nodec space every nowhere dense set is not only closed but also discrete. Let y be an arbitrary countably infinite collection pairwise disjoint finite subsets of X, and let F0 € 2F be arbitrary. Suppose first that every neighborhood U of FQ intersects all but finitely many F Claim 1. There are a point x € FQ and an infinite subcollection J' of 3*\{F0} such that every neighborhood of x intersects all but finitely many members from y1. Proof. We prove this by induction on the cardinality of FQ. If |F0| = 1 then y = y \ {FQ} is as required. Now suppose that we have what we want if |.Fo I —ni and suppose that |Fo| = n + I . Fix an arbitrary x e FQ. If every neighborhood of FQ \ {x} intersects all but finitely many members of J\ {FQ} then we get what we want by our inductive assumption. If not, then there is a neighborhood U of FQ \ {x} such that the collection is infinite. But then every neighborhood of x intersects infinitely many members from y\ {FQ}. 0 So let x and 3' be as in the claim. Since x £ \J3*', it follows that 3"' is infinite. Split 3"' into two disjoint infinite subcollections, say 3'0 and y{. Then by Claim 1 , x is an accumulation point of \J 3'0 as well as \J 3^ . This is a contradiction since |J 3~g and [J 3^ are disjoint and X is perfectly disconnected. So our assumption that every neighborhood U of FQ intersects all but finitely many F e 3~ \ {F0} leads to a contradiction. Let CQ be a clopen neighborhood of FQ for which there exists an infinite subcollection $Q of the
386
6. FUNCTION SPACES
collection J \ {F0} such that \J30 C X\C0. Let FI € 3~0 be arbitrary. By the same argument, there is a clopen neighborhood C\ C X \ CQ of FI and an infinite subcollection 3^ of 3^ \ {-^1} such that \Jffi C X\C]_. Observe that in fact U 3~i C X \ (Co U C\). Continuing in this way inductively, one easily constructs a subsequence {Fn : n G a;} of 3* having pairwise disjoint clopen neighborhoods. But then 5 = U n e^ ^n *s discrete, hence nowhere dense and hence closed since X is nodec. Since X is Lindelof, being countable, this easily implies that {Fn : n G w} is strongly discrete. For an easier example of a countable nondiscrete space whose function space is Baire, see Corollary 6.5.7. A pseudocompact space. It is clear that a pseudocompact space X has the property that its function space CP(X) is not Baire. This is so, for example, since in a pseudocompact space any discrete family of open sets is finite. It can also be shown directly, as follows. Simply observe that the subspaces {feCp(X):f[X]C[-n,n}} are closed and nowhere dense in CP(X) and cover CP(X) in case X is pseudocompact. Our aim now is to construct an example of a particular pseudocompact space which demonstrates that for a space X the property of having a Baire CP(X) is not determined by its countable subspaces. A point p of a topological space X is called a weak P-point if for every countable A C X \ {x} we have x £ A. KUNEN [236] proved that N* contains many weak P-points. Here N has the discrete topology. In fact, if U is a countably infinite collection of pairwise disjoint nonempty clopen subsets of N* then
contains a weak P-point. Let K denote the set of all weak P-points in N* and observe that K is dense in N*. Lemma 6.4.7. K is pseudocompact. Proof. Let A be a countably infinite discrete collection of nonempty clopen subsets of K. We will derive a contradiction. Since K is zero-dimensional, being a subspace of N*, this guarantees pseudocompactness of K. For every A e A let CA be an open subset of N* such that CA n K = A. Observe that the collection
{CA : A 6 A} is pairwise disjoint since K is dense. For every A £ A let BA be a nonempty clopen subspace of N* such that BA C CA • Since the collection
6.5. FILTERS AND THE BAIRE PROPERTY IN Cp(Ny)
387
is pairwise disjoint, by the above there is a weak P-point in |J BA\ |J BA, A£A
AeA
say x. So x £ K. But this easily contradicts the fact that A is discrete.
D
We conclude that CP(K] is not Baire since K is pseudocompact. However, every countable subspace of K is discrete. So every countable subspace A C K has the property that CP(A) = RA is Baire (even topologically complete). This proves indeed that for a space X the property of having a Baire CP(X) is not determined by its countable subspaces. 6.5. Filters and the Baire property in {7P(]%) For each filter 3 on N we can topologize the set N = N U {00} as follows: each element of N is isolated and the collection {U U {00} : U G jF} is a neighborhood base at oo. This topological space will be denoted by !%•. It is an easy exercise to verify that !%• is Tychonoff if and only if 3" is a free filter on N. The spaces CP(N?) are of crucial importance in our analysis of the topological type of function spaces of low Borel complexity in §6.12. Here we are interested in characterizing when such a space has the property that its function space is a Baire space. The Cantor set. We consider the Cantor set {0,1}°° with its usual Tychonoff product topology. If a and r are finite disjoint (possibly empty) subsets of N then the set
(cr, T) = {x £ {0, 1}°° : (n £ a =>> xn = 0) A (n e r => xn = 1)} is a basic clopen subset of {0, 1}°°. Notice that if a\ , r\ , a -2 and TI are pairwise disjoint finite subsets of N then
(cri,n) n (02, r2) = {<TI ucr 2 ,ri ur 2 ). Lemma 6.5.1. Let F C {0,1}°° be nowhere dense. Then for every finite subset A C N there are disjoint finite subsets a and (3 of N \ A such that if 7 and 6 are arbitrary complementary subsets of A then
Proof. The proof is basically a triviality. Let 71,^1, .. .7™, $n list all the complementary pairs of subsets of A. Consider the nonempty clopen subset {71,^1} of {0,1}°°. Since F is nowhere dense, the set (71,^1) \ F has nonempty interior. There are consequently disjoint finite subsets ai and fli of N \ A such that
388
6. FUNCTION SPACES
Now consider the nonempty clopen subset (72, £2} n <«!,/?!} = ( 7 2 U a i , ( $ 2 U / 3 i } of {0, 1}°°. By a similar argument there are disjoint finite subsets ot-z and ftz of N \ (A U QI U fti) such that (72,^2) n (ai,£i) n (a 2 ,#2> n F = 0. If we repeat this process n times, we dealt with all possibilities. It is clear that the finite disjoint sets
i=\
i—l
are as required.
D
Set theoretic characterization of meager filters. Each subset A of N can be identified with a point in the Cantor set {0, 1}°°, namely with its characteristic function XA- So 9(N) can be topologized by using the identification A ^ XA, as follows: U C ?(N) is open if and only if
{XA : A 6 U} is open in {0,1}°°. The set T(N) topologized in this way is a Cantor set because {0, 1}°° is. For A, S G T(N) put
V(A, S) = {B e ?(N) : B n 5 = A n 5}
(cf. Page 462). Observe that the topology of y(N) is generated by all sets of the form V r (A,5), where 5 is finite. Also notice that the finite subsets of N form a (countable) dense subset of 7(N). Lemma 6.5.2. Let 3 C 3>(N) be nowhere dense. Then for every finite set A C N there are a finite set C C N \ A and a subset D C C such that for every B C N we have V (B, A) n V (D,C)n?=®. Proof. This is exactly the statement in Lemma 6.5.1 translated in terms of the set theoretic language of ?(N). Let F = {XH '• H G 9"}. Then F is nowhere dense in {0, 1}°°. Let a and J3 be the sets given by Lemma 6.5.1 and put D — a and C = a U ft. Now let B C N be arbitrary, and put 6 = B n A and 7 = A \ B, respectively. Then (7, 6} n (a, ft) n F = 0 implies Simply observe that
6.5. FILTERS AND THE BAIRE PROPERTY IN Cp(N-j)
So we are done.
389
D
A family 3~ of subsets of a set X is said to be closed under supersets provided that for all F £ 3 and G e ^(X) such that F C G it follows that G e y. Note that any filter on X is closed under supersets. If J is a collection of subsets of N then by the phrase ' J is meager' we mean that J is a meager subset of the space 9(N). Similarly for the phrase '3~ is not meager'. We now come to the following interesting characterization of meager niters on N. Theorem 6.5.3. Let 3~ C 9(N) be closed under supersets. statements are equivalent:
The following
(1) 3 is meager. (2) There is a sequence (An}n ofpairwise disjoint finite subsets of N such that for every F 6 3~ the set ( n G N : F n A n = 0} is finite.
Proof. We first prove that (2) implies (1). This is simple. Indeed, let the sequence (An)n be given. For every n E N put D n - {B E ?(N) : (Vm > n)(B n A m ^ 0)}.
We will prove that every T>n is a closed and nowhere dense subset of O-'(N), which suffices since by assumption 3~ is covered by the D n 's. So fix n, and assume that B £ T>n. Then there exists m > n such that J5 Pi A m = 0. So V^(5,yl TO ) is a clopen neighborhood of B which misses T>n. This proves that T>n is closed. To prove it is nowhere dense, let B, A G J*(N) be arbitrary with A finite. We will show that the basic neighborhood Vr(B,A) of B is not contained in T)n. This clearly suffices. Indeed, since .4 is finite there is an m > n such that A n Am = 0. This implies (B\Am] HA = Bn A, and hence B\AmeV(B,A). But clearly, B \ Am g T>n. To prove (1) => (2), let oo
^ c |J Jn, n=l
where for each n the set !J"n C J*(N) is nowhere dense. WTe may assume without loss of generality that
*
9i c ?2 c • • • c yn c • • - .
390
6. FUNCTION SPACES
By Lemma 6.5.2 we may pick by induction on n a finite set An with subset B such that (1) AnnAm = 0 if n ^ra, (2) for every B G ?(N),
(B,
n v £ n A n n yn =
We claim that the sequence (An}n is as required. If not, then for some element X G 3~ we have that E = {i G N : X n Ai — 0} is infinite. Since 7 is closed under supersets, it follows that Y — X U \Ji<EE Bi G 7. By (*), there is i G E such that F 6 3^. Since X n A; = 0 and the sequence (An}n is pair wise disjoint, we get F n Ai = Bi. This shows that Y G V(Bi,Ai), and consequently that ' U A<
nV B
( " A^
nJ
-
This contradicts (2).
D
This result allows us to 'decompose' meager filters in meager filters. Corollary 6.5.4. Let J be a filter on N. In addition, let {N; : i G N} be a partition of N into infinite pairwise disjoint sets. Put 7l — J \ Nj for every i. TJien (1) if Fi G Ji for i < n then |J?=1 F, U U J > n N, G 7. (2) ^ imbeds as a closed subspace of 3" for every i, (3) if 3" is a free filter on N and a meager subset of CP(N) then there exists a partition {Nt : i G N} of N such that each 3^ is a free iilter on N; and a meager subset o Proof. For (1), for every i < n let F; G 3~ be such that F» n N» = F». So
and so the desired result follows by observing that 3" is closed under supersets. For (2), simply observe that the function which assigns to each F G 3~; the set F U (N \ Ni ) is a closed imbedding of 3~; into 3". For (3), let (An)n be a pairwise disjoint sequence of finite subsets of N such as in Theorem 6.5.3. There clearly is a partition (N^ : i G N} of N such that every Nj contains infinitely many terms of the sequence (An)n. Since 3" is free, so is every 3V In addition, by construction and Theorem 6.5.3, every 3^ is a meager subset of O'(Nj). D
6.5. FILTERS AND THE BAIRE PROPERTY IN C p (Nj-)
391
Corollary 6.5.5. Let 3 be an ultrafilter. Then 3 is not meager. Proof. We naturally use the characterization in Theorem 6.5.3. To this end, let A be a countably infinite pairwise disjoint collection of finite subsets of N. Split A into two countably infinite subcollections T> and C. Then
are disjoint subsets of N, so at most one of them belongs to 3. We may therefore assume without loss of generality that B $ 3. Then N \ B £ 7 by Exercise A. 1.1 and misses infinitely many elements from A. D The Baire property. We now come to the announced characterization result. Theorem 6.5.6. Let 3 be a free filter on N. The following statements are equivalent: (1) Cp(Ng-) is a Baire space. (2) 3 is not meager. Proof. We first prove that (1) => (2). Assume that 3 is meager. We will prove that C p (r%) is meager in itself and so is not a Baire space. By Theorem 6.5.3, there is a pairwise disjoint sequence (An}n of finite subsets of N such that for every F £ 3 the set {n £ N : F n An = 0}
is finite. If /: Ny -» R is continuous then /" 1 [(/(oo) — l,/(oo) + 1)] is a neighborhood of oo. So there exists an N £ N such that for every n > N there is an element xn £ An such that For every n put Bn = {/ £ M* : (Vro > n)(3xm e A m )(|/(oo) - f(xm)\ < 1)}. As we just observed, oo
£„(%) C |J Bn. n=l N
Since Cp(Ng-) is dense in R (Lemma 6.2.1), we are done if we can prove that each Bn is meager. Claim 1. Every Bn is closed and nowhere dense in M N . Proof. Fix n and assume that / G MN \ Bn. Then there exists m > n such that /[A m ]n[/(oo)-l,/(oo) + l] = 0.
392
6. FUNCTION SPACES
Since Am is finite there clearly is an open neighborhood U of / such that for every g 6 U we have
So U misses Bn and since / was arbitrary this shows that Bn is closed. We next prove that Bn is nowhere dense. If not then there are an element / € MN , a finite subset F C N and an e > 0 such that the set
is contained in Bn. Without loss of generality, oo G F. Since F is finite, there is an element m > n such that Am fl F — 0. Define h : F U ATO —> R by /(oo) + 2
(xeAm).
The function /i can be extended to a continuous function h: N —>• R by the Tietze-Urysohn Extension Theorem. By (*) it follows that h £ Bn, but by the definition of Bn it follows that h £ Bn. This contradiction establishes the proof of the claim. 0 We next prove that (2) =>• (1). Let (An)n be a pairwise disjoint collection finite subsets of N. Our aim is to apply Theorem 6.4.3, i.e., to prove that this sequence has an infinite strongly discrete subsequence, We may assume without loss of generality that oo 0 U^Li An, i.e., that An C N for every n. Since 3* is not meager, there exists by Theorem 6.5.3 an element F e 3~ such that the set E= {ne N : F n A n = 0} is infinite. A moments reflection shows that the sequence (An}n^E is strongly discrete in 1%. D Corollary 6.5.7. Let 5F be an ultrafilter on N. Then C p (f%) is a Baire space. Proof. This follows directly from Corollary 6.5.5 and Theorem 6.5.6.
D
Observe that for an ultrafilter J the space Cp(r%-) is not a Borel subset of R* by Corollary 6.3.3. Remark 6.5.8. For ultrafilters 5" there is an alternative way of looking at the spaces N^, which we will now explain. Consider the space /?N (here N is endowed with the discrete topology). If p € N* then the collection
5P = {A C N : p E A}
6.6. EXTENDERS
393
is easily seen to be an ultrafilter on N, and the subspace N U {p} of /3N is canonically homeomorphic to the space f% . A point p in a topological space X is called a P -point if the intersection of countably many neighborhoods of p is again a neighborhood of p. It is a wellknown result of RUDIN [360] that the space N* contains a P-point under the Continuum Hypothesis. SHELAH constructed a model of set theory in which there are no P-points in N* (see e.g., WlMMERS [410]). Notice that every Ppoint is obviously a weak P-point and that N* contains many weak P-points in ZFC alone (see Page 386). In [184], GUL'KO and SOKOLOV proved that if .Y = Nu{p}, where p 6 N* , then CP(X) is hereditary Baire if and only if p is a P-point (for a generalization, see MARCISZEWSKI [270]). So if p is not a P-point then CP(X) is Baire, but not hereditary Baire. Observe that there are many non-P-points in N* so the existence of a Baire but not hereditary Baire CP(X) does not require any additional set theoretic assumption. 6.6. Extenders Let A C X be a closed subspace. An extender is a function that simultaneously extends real-valued continuous functions on A to real-valued continuous functions on X. That is, a function p: C(A] ~
such that for every / e C(A). Theorem 6.6.1. Let X be a normal space, and let A C X be closed. Then there is an extender p: C(A) —> C(X) which is linear. Proof. This is easy. Simply observe that C(A) is a vector space. So let 3~ be a Hamel basis for C(A). By the Tietze-Urysohn Theorem, every / e J can be extended to a continuous function /: X —>• R. Now define p: C(A] -> C(X] by the formula
where f\ , . . . , fn 6 3~ and AI , . . . , Xn G M are arbitrary. It is clear that p is as required. D So linear extenders always exist and hence are not very interesting. For metrizable spaces, there always exists an extender which is both linear and continuous. This is much more interesting.
394
6. FUNCTION SPACES
The Dugundji Theorem (Part 2) 6.6.2. Let X be a metrizable space with closed subspace A. Then there is a linear extender y: C(A)
having the following properties: (1) (p: Cp(A) —> CP(X) is continuous. (2) If / € C * ( A ) then I I / H Remark 6.6.3. Observe that by (2),
A = UJ\
(See also ENGELKING [153, Exercise 4.5.20(c)j.) In Example E14 of §6.13 we will show that there is a countable space for which the Dugundji Extension Theorem fails. This is of interest since such a space X has the additional property that CP(X] is separable and metrizable. Proof. We return to the proof of Theorem 1.2.2 for the case L = R. First we note that nowhere in that proof we made an essential use of separability since all metrizable spaces are paracompact (see the introduction to this chapter). Consider the formula (*) from Page 23: (X G A)
. f ( a u ) (xex\' A}. It is evident that the assignment ?(/) = / is linear. We claim that it is also continuous. To prove this, it suffices to prove continuity at 0, the constant function with value 0. To this end, let P C X be finite and e > 0. Put
Then Q is finite since U is locally finite. Claim 1.
6.6. EXTENDERS
395
for every U E UQ so that \f(au}\ < £• Finally, observe that if U G U \ UQ then KU(X) — 0. These facts imply that
^ KU(X) ' f ( a u } ueii KU(X) • f ( a u ) KU(X) • f ( a u ) \
which is as required.
0
So the claim proves the continuity of (p. With a similar calculation we will prove that (p is norm preserving. Indeed, if / e C*(A), xeX\Aandt = \\f\\ then
ueu ueu
KU(X) - t = t.
This means that ?(/) and / have the same norm.
D
Using this result in a clever way, it is possible to prove the existence of extenders in various situations. Corollary 6.6.5. Let X be a space with compact metrizable subspace K. Then there is a continuous linear extender (p: CP(K) —>• CP(X). Proof. By Corollary A. 4.4 there is an imbedding i: K ->• Q. Since K is compact, the Tietze-Urysohn Theorem shows that the function i : K —>• Q can be extended to a continuous function i: X ->• Q. In addition, Theorem 6.6.2 gives us a continuous linear extender e: Cp(i[K]) —> CP(Q). Now define the function (p: CP(K) ->• CP(X) by the formula =e(foi~1)oi. It is easy to see that (p is as required.
D
396
6. FUNCTION SPACES
Various other results in the same spirit can be proved along the same lines. For example, an extender for a pair of spaces ( X , Y ) , where Y is a closed subspace of X, exists if X is normal and Y is completely metrizable and separable, or if X is paracompact and Y is completely metrizable. For details, see MICHAEL [285] and ARENS [22]. Factorizing CP(X] and C*(X}. Theorem 6.6.2 can be used to prove several interesting factorization properties of function spaces. Let X be a space and A C X. We say that an element / e CP(X) vanishes on A provided that / \ A is the constant function with value 0. For a topological space X and subset A C X we let CP,A(X) denote the subspace of CP(X) consisting of all functions vanishing on A. Similarly for C*iA(X). If A is a singleton, say {a}, then we denote CP,A(X] and C^A(X] simply by Cp,a(X] and C*^a(X), respectively. Proposition 6.6.6. Let X be a space and let A C X be closed. If there is a continuous linear extender e: CP(A) —> CP(X} then CP(X) ~ CP,A(X) x CP(A) In addition., if there is a continuous linear extender £: C*(A) —>• C*(X) then
Proof. We will prove this for C*(X] only. The proof for CP(X) is similar. Define p: C*(X) -»• C*(A) by p(f) = $ \ A. Then p is continuous (it is the restriction to C*(X) of a projection) and linear. Define
by We will first prove that (p is well-defined. Take an arbitrary / € C*(X). It is obvious that p(f) e C*(A) and that / - (f op)(/) e C*(X). Furthermore,
= 0,
hence / - ( e o p ) ( / ) e C ; ! A ( X ) . That (p is continuous and linear is trivial. We show that ip is a homeomorphism. Define
6.6. EXTENDERS
397
by It is clear that T/> is well-defined, continuous and linear. Furthermore, ?/> o (p is the identity on C*(X). We show that (poijj is the identity on CP,A (X)xC* (A) . Take arbitrary / e C*fA(X) and p e Cp(4). Note that p(f) = f \ A - Q, hence by linearity of £ it follows that (£ o /?)(/) = £(0) = 0. So
to)) ,
= (/,), as required.
D
This result will sometimes be used in a very trivial situation, namely that A is a singleton subset of X. For such a subset the conditions of Proposition 6.6.6 are trivially satisfied, and so we get: Corollary 6.6.7. Let X be a space and let x G X. Then CP(X] ~ {/ G CP(X) : f(x] = 0} x R.
Simi/arJy,
Proof. It suffices to observe that Cp({point}) = C* ({point}) ~ R.
D
Corollary 6.6.8. Let X be a metrizable space, and let A C X be closed. Then
cp(x) ~ cp,A(x) x CP(A\ c;(x) ~ c;tA(x) x c;(A). Proof. Again, we prove this for C*(X) only. Since X is metrizable, and A in A" is closed, there is by Theorem 6.6.2 a continuous linear and norm preserving extender £: C*(A) —> C*(X). So we are done by an application of Proposition 6.6.6. D Corollary 6.6.9. Let X be a space containing a nontrivial convergent sequence. Then
cp(x] ~ cp(x) x M, c;(x] ~ c;(x) x R Proof. Let A « o;+l be a nontrivial convergent sequence in X. First observe that (cj + 1) 0 {point} « w + 1 so that C p (w + 1) ~ Cp((u + 1) © {point}) - CP(UJ + 1) x M
398
6. FUNCTION SPACES
by Theorem 6.2.8. It now follows from Corollary 6.6.5 and Proposition 6.6.6 that CP(X) ~ CP,A(X) x CP(A) ~CP;A(X) x (CP(A) x R ) ~ (CP,A(X) x CP(A)) x R ~ CP(X) x R. The proof for bounded functions is similar.
D
Corollary 6.6.10. Let X be an infinite first countable space. Then
cp(x) ~ cp(x) x R, c;(x) ~ c;(x) x E. Proof. If X is discrete then there is nothing to prove. If X is not discrete then it contains a nontrivial convergent sequence. So then we are done by Corollary 6.6.9. D Remark 6.6.11. This result is not true for all spaces. MARCISZEWSKI [268] proved that there exists a compact space X such that CP(X) is not linearly homeomorphic to CP(X) x R. Corollary 6.6.9 indicates that Marciszewski's Example must be rather complex since it does not contain any nontrivial convergent sequence. If X is a space and A C X is closed, then TT : X ->• X/A denotes the natural decomposition map. The unique point of X/A contained in 7r[A] will be denoted by oo for convenience. Since X/A is endowed with the quotient topology, it is clear that for every (bounded) continuous function /: X —>• R vanishing on A there exists a unique (bounded) continuous function /: X/A —> R vanishing on Tr[A] such that / = / ° TT. So CP,A(X) and Cpj00(X/A) can be identified as sets. Fortunately, this identification is a linear homeomorphism. Similarly for bounded functions. Proposition 6.6.12. Let X be a space, and A C X be closed. Then
cp,A(x) ~ cp>00(x/A), c;tA(x) ~ c;t00(x/A). Proof. This will be proved for CP>A(X) only since the proof for bounded functions is entirely similar. Define ?: CPJA(X] -> CptOQ(X/A} by (/?(/) = / (see the remarks above). Then (p is a bijection as we just observed, and is clearly linear. It suffices to prove that (pis a homeomorphism. Take arbitrary 2 / 1 , . . . , yn G X/A and let x i , . . . , xn 6 X be such that TT(XJ) = yi for every i. Then for all e > 0 and / G Cp^^O it is easily seen that V>[N(f, {xi,...,xn},£)]= N ( f , { y i , . . . , 2/n}, e)]So ? is a homeomorphism.
D
6.7. THE TOPOLOGICAL DUAL OF CP(X)
399
Again, let X be a space, let A C X be closed. By Corollary 6.6.7 we get Cp(X/A)~Cpt00(X/A)xR and
These trivialities will be used in the proof of the following result. Corollary 6.6.13. Let X be an infinite metrizable space, and let A C X be closed. Then
cp(x] ~ CP(X/A) x cp(A), c;(x] ~ c;(x/A) x c;(A). Proof. We complete the proof in two steps. First, Corollaries 6.6.10 and 6.6.8 give us that CP(X) ~Cp(X)xR~
CP,A(X) x CP(A) x R.
But
Cp,A(X)~Cpt00(X/A) by Proposition 6.6.12. Again by Corollary 6.6.8, CP(X/A) ~ Cpt00(X/A) x Cp({point}) ~ CP>00(X/A) x R. It therefore follows by the above remarks that CP(X) ~ CP,A(X) x CP(A) x R ~Cpt00(X/A)xCp(A) x R ~ (CP>00(X/A) x R) x CP(A) - CP(X/A) x CP(A). The proof for bounded functions is again similar.
D
Remark 6.6.14. These simple results are basically the 'only' known general tools for proving that spaces are ^-equivalent. They will be used frequently in the forthcoming sections. 6.7. The topological dual of CP(X) For a space X let L(X] be the dual of CP(X), i.e., the set consisting of all continuous functionals on CP(X). For x € X we define the function ^ : CP(X) -» R by £c(/) = f ( x ) . As usual, £x is called the evaluation at x. Lemma 6.7.1. Let X be a space and x 6 X. Then £x is a continuous functional.
400
6. FUNCTION SPACES
Proof. It is evident that £x is linear. Observe that for e > 0 and / G CP(X) we have tx[N(f,{x},e)]C(f(x)-e,f(x)+e). From this continuity of ^x is obvious.
D
If x, y G X and x / y then there is / 6 C(X) such that f ( x ) / /(y). As a consequence, U/) = / ( * ) / / ( < / ) = £,(/)• So we conclude that the assignment x >->• £z is one-to-one. This makes it possible to identify X with a subset of L(X) . We will show that X is a linearly independent subset of L(X) (hence X ^ L(X)] and that it generates L(X}. We will topologize L(X} later and prove that X is in fact a subspace of L(X). Proposition 6.7.2. Let X be a space. Then the collection {£x : x G X} is a Hamel basis for L(X). Proof. We will first prove linear independence. Take pairwise distinct elements x\ , . . . , xn in X, and elements AI , . . . , An Gffi.such that X)ILi ^£E» = 0For every i < n let fi : X —>• M be continuous such that f i ( x i ) - 1,
fi(xj) = 0 (i^ j ) .
Then fc=l
k—l
for every i < n. Now let F: CP(X) ->• M be a continuous functional. Then F(0) = 0, where 0 is the constant function with value 0. By continuity of F there is a neighborhood of 0 which is mapped by F into the interval (—1,1). For this neighborhood we may take a basic neighborhood of 0, say A^(0, P, 5), where P C X is finite and 6 > 0. Let P — {xi, . . . , xn}. Claim 1. If f , g 6 CP(X) and / \ P = g \ P then F(f) = F(g). Proof. Let f,ge CP(X) be such that f \ P = g \ P. Then / - g \ P = 0 and hence k • (f - g) = 0 for every k e N. So fc • (/ - g} G AT(Q, P, 5) for all fe so that k \ F ( f ) - F ( g ) \ = \F(k-(f-g))\
0
6.7. THE TOPOLOGICAL DUAL OF CP(X)
401
Observe that by the Tietze-Urysohn Theorem it follows that TT is surjective. Claim 1 consequently shows that there is a unique function F' : M.p —> R such that F = F' o TT . Then F' is linear since F and TT are. Since P is finite, by elementary Linear Algebra there are AI, . . . , \n G K such that i—l
for every g G Rp . From this we conclude that F(f) = F ' o i r ( f ) = i-l
i=l
for every / G CP(X). But this is clearly as required.
D
We now topologize L(X) in a natural way. For / G C(X) let be defined by Then L(f) is clearly a functional on L(X). Let £(X) be the collection of functionals on L ( X ) obtained in this way, and consider the function
e: L(X] defined by (e(F})L(f)=L(f}(F}=F(f}. This function is clearly linear, and we claim that it is one-to-one. Since e is linear, it suffices to prove that Kere = {0}. Take an arbitrary F G L(X] such that F ^ 0. By Proposition 6.7.2 there are elements Xi,...,xn G X and ai , . . . , an G M such that
We may assume without loss of generality that the z's are pairwise distinct. Since F ^ 0, wre may also assume without loss of generality that a\ ^ 0. Let / G CP(X) be such that f ( x i ) = I and /[{ar2, . . . ,xn}] C {0}. Then
from which it follows that e(F) ^ 0. So we showed that £(JO separates the points of L(X] which implies that the function e is one-to-one (there are general theorems that allow us
402
6. FUNCTION SPACES
to conclude this without any work, but it is a good exercise to perform the above calculations). The topology on L(X) we are interested in is the weakest topology which makes all functionals L(f), f G C(X), continuous. That is, we identify L(X) and the linear subspace e[L(JQ] of RL(X) . It is clear that L(X) is a locally convex linear space. With this topology L(X) is called the topological dual of CP(X). It will be convenient to identify x G X and £x G L(X). Lemma 6.7.3. Let X be a space and f G C(X). Then L(f) : L(X) -»• E is the unique continuous functional that extends f . Proof. That L(f) is a continuous functional is clear. For x G X we clearly have L(/)(x) = f ( x ) , so !/(/) extends /. Since X is a Hamel basis for L(X) (Proposition 6.7.2) it follows that !/(/) is unique. D Proposition 6.7.4. The inclusion X <-^- L(X] is a closed imbedding. Proof. We show first that (£x : x G X} as subspace of L(X) is homeomorphic to X. The homeomorphism that does the job is of course the function which we denote for convenience by £. Let U C X be open, and let x G U. We claim that £[U] is a neighborhood of £(x) = £x in £[X]. Indeed, let / 6 C(X) be such that f(x) = 1 and f[X \U] C {0}. Put Then V is open in L(X) and £x G V D £[X] C ^[£7]. This shows that is open. To prove continuity, let V C ^[X] be open, and let £(x) = ^x G V be arbitrary. By the definition of the topology on L(X), there are functions /i , . . . , fn G C(X) and open subsets f/i , . . . , C/n of R with
Then a; G nT=i /i"1^] ^ ^~1[^/]- This proves continuity, and hence that ^ is a homeomorphism. This justifies the identification of X and £[X], We show next that X as subspace of L(X) is closed. To this end, take an arbitrary F 6 L(X) \ X. There are xi , . . . , xn G X and AI , . . . , Xn G IR such that F = ]>^™_1 AiX{ (Proposition 6.7.2). We may assume without loss of generality that xi ^ Xj \ii ^ j and also that A; ^ 0 for every i.
6.7. THE TOPOLOGICAL DUAL OF CP(X)
403
We assume first that n > 2. There are pairwise disjoint open sets Vi C X with Xi G Vi for every i. Let /; 6 C(X) be such that fl(xi) = 1 and for every i. We claim that F belongs to the open set HlLi ^(/i)"1^ \ {0}] and that this set is contained in L(X] \ X. Indeed, for each i < n, n
n
fl(xJ) = A, e R \ {0}. Moreover, for an arbitrary x £ X there exists i < n with x £Vi (since n > 2) so that L(fi)(x} = /i(ar) = 0 0 R \ {0} and so
We conclude that F has a neighborhood which misses X. Assume next that n = 1, i.e., F = XiX\. Since F £ X, clearly AI ^ 1. Put U = R \ { 1 } and let / be the constant function with value 1 . We claim that F 6 L(/)-1[C7] C L(X) \ X. Indeed, L(f)(F)
= AiL(/)(xi) - AI/(XI) = A! G U
and for x 6 X we have L(f)(x)
— f(x) — I ^ U. So we are done.
D
We now come to an interesting 'duality' result. Theorem 6.7.5. Let X and Y be spaces. Then CP(X] and CP(Y) are linearly homeomorphic if and only ifL(X) and L(Y) are linearly homeomorphic. Proof. First suppose that (p\ CP(X) —> CP(Y) is a linear homeomorphism. Define *: L(X) -> L(Y) by *(F) = Fop-1. Then * is obviously a welldefined linear function. To see that it is continuous, notice that for / £ C(Y) and open U C M we have which is open in Define $: L(F) ->• L(A") by $(<2) = G<xp. Then by symmetry it follows that $ is continuous. Since trivially $ = *-1, this shows that L(X] and L(y) are linearly homeomorphic. Conversely, suppose that -0: L(X) —>• L(F) is a linear homeomorphism. Define v?: CP(.Y) ->• CP(Y) by <^(/) = (L(/)o^- 1 ) \Y. Then y? is obviously a well-defined linear function. To prove continuity, take arbitrary finite P C Y" and let e > 0. For every y £Y there are *?,...,<€*,
Af,...,AJ[ 6M\{0}
404
6. FUNCTION SPACES
such that
1=1 (Proposition 6.7.2). Let N = max-E^ l A f l : 2/ e P}, 5 = S/N and Q = {x\ : y £ P and i < ny}. We claim that Indeed, if / € JV(Q, Q, £) then for y G P we have
So 9? is continuous at the origin, and hence is continuous everywhere by linearity. Define tf : CP(Y) ->• CP(.Y) by 1%) = (L(p) o ^) t X. By symmetry it follows that $ is continuous and linear. Since clearly i) = (p~1, we are done. D Remark 6.7.6. It is unknown whether a similar result can be derived for homeomorphisms. Are CP(X) and CP(Y) homeomorphic if and only if L(X) and L(Y) are homeomorphic? (Probably not.) 6.8. The support function If X and Y have linearly homeomorphic function spaces CP(X) and CP(Y) then X and Y have much in common. A certain multi-valued lower semicontinuous function plays a crucial role in investigating those similarities. In this section we define this function, and derive some of its basic properties. Applications of the results derived here are given in the next sections. Let X and Y every y € Y, the is defined to be neighborhood U
be spaces and let (p : C(X) —> C(Y ) be a linear function. For support in X of y with respect to (p, abbreviated supped/), set of all x 6 X satisfying the condition that for every of x, there exists an / G C(X) such that f[X \ U] C {0}
At first glance this seems to be a very technical concept. But it will turn out to be an important tool in understanding the function spaces CP(X) a little better. If ? is a bijection then (^-1 : CP(Y) —> CP(X) is also linear, so we can then consider the support of a point in Y with respect to (p and the support of a point in X with respect to y"1. When it is clear from the context which support we mean, we will suppress the index.
6.8. THE SUPPORT FUNCTION
405
For A C Y we define supp(A) = U( SU PP(2/) : ?/ ^ A}. There is an alternative description of supports in the special case of a continuous linear function that is very useful. Let X and Y be spaces and let (p: CP(X) —>• CP(Y) be continuous and linear. Fix y E Y. The function ipy: CP(X) -> 1R defined by 4>y = £y ° V 'ls continuous and linear. So i})y E L ( X ) . If i/}y ^ 0 there consequently exist by Proposition 6.7.2 pairwise distinct x\,..., xn E X and \\,..., An € R \ {0} such that
(Notice that if (p is a bijection then t/>y 7^ 0 for every y 6 Y.) This means that for every / E Cp(-X") we have
We claim that supp(^) = {xi,...,x n }. To see this, let x E supp(y) and suppose that x ^ {xi, . . . , xn|. Since X\{xi,...,xn} is an open neighborhood of x, there exists / E CP(X] such that f ( x i ) — 0 for every i < n and tp(f}(y] ¥" 0- But by (*),
which is a contradiction. Conversely, let i < n and let C/ C X \ {xj : j < n and j ^ i} be an arbitrary open neighborhood of Xi. Let / E C(X) be such that f[X\U] C {0} and f ( x i ) = 1. Then again by (*),
which proves that x^ E supp(y). So we conclude, in particular, that the support of a point is finite. Let us summarize what we just concluded for continuous linear functions. To this end, let X and Y be spaces and let (p: CP(X) —>• CP(Y) be continuous and linear. We proved that for every y E Y there exists a (possibly empty) finite subset supp(y) C X and for every z E supp(y) a non-zero real number A(y, z) depending on y and 2 such that for every / E CP(X),
(**)
406
6. FUNCTION SPACES
Observe that (**) is in fact identical to (*). Lemma 6.8.1. Let X and Y be spaces, y G Y and (p: CP(X) —>• CP( continuous and linear. If f,g G CP(X) coincide on supp(y) then
Proof. This follows easily from (*). Simply observe that if
then supp(y) = {x\ , . . . , xn} and
which is as required.
D
We will now present a few other useful properties of the support function. Lemma 6.8.2. Let X and Y be spaces, and let (f>: CP(X] continuous and linear. Then (1) (2) (3) (4) (5)
->• CP(Y] be
If ip is injective then supp(y) is dense in X. IfACY then supp(A) C supp(A). If (p is surjective then supp(y) ^ 0 for every y G Y. If supp(y) ^ 0 for every y 6 Y then supp: Y =$• X is LSC. If (p is a homeomorphism then for every x 6 X we have that
Proof. For (1), suppose there is an x € X which is not in the closure of supp(y). Let / G C(X] be such that f ( x ) = 1 and /[supp(Y)] C {0}. Since the function / agrees with the constant function 0 with value 0 on every support set, by Lemma 6.8.1 it follows that ip(f)(y) = ?(Q)(y) = 0 for every y e Y. So
supp(A).
Since x G supp(y), there exists an element / G C ( X ) with f[X \ U] C {0} and
But then
= 0, which is a contradiction.
For (3), take an arbitrary y G Y such that supp(y) = 0. Let / G C(Y) be such that f(y) ^ 0, and let g G C(X) be such that
6.8. THE SUPPORT FUNCTION
on 0 which is a neighborhood of supp(y), and so ip(g)(y) Lemma 6.8.1. This is a contradiction. For (4), let U C X be open, and put
407
— f(y] = 0 by
A = {y G Y : supp(y) n U ^ 0}. Let y G A, and take x G supp(y) n U. Let V C X be open such that
x C V C V C U. There exists / G C(A") such that /[X \ V] C {0} and (p(f)(y)
^ 0. Let
Then W is an open neighborhood of y. We claim that W C A. Suppose that there is an element z G W \ A. Then (p(f)(z) ^ 0 and supp(2;) n U = 0. So X \ V is a neighborhood of supp(z) and f [ X \ V ] C {0}, which implies by Lemma 6.8.1 that ?(/) (2) — 0, contradiction. For (5), let x G X and assume that x ^ A — supp^, (supp^-^x)). Since A is finite, there exists / G C(X) such that /(x) = 1 and f[A] C {0}. By Lemma 6.8.1 it follows that ?(/) = 0 on supp^-i^), and by the same argument that f ( x ) = (p~l (
408
6. FUNCTION SPACES
For every n, let xn G supp(yl) be such that f ( x n ) = n. There exists y\ G A such that x\ G supp(^i). Put n(l) = 1. Since Ki(i) is a neighborhood of x\ and x\ G supp(^i), there exists h G CP(X) such that Let
and put /i = A/i. Then /i is continuous, /[X \ V^i)] C {0}, and by linearity of
v% n p = 0 for every z > n(k). Since x n (fc) G supp(.4), there is an element y^ G A such that x n (fc) G suppd/fc). The set V n (fc) is a neighborhood of £ n (fc) and since x n (fc) G supp(?/A;), there consequently exists an element h G CP(X] such that h[X \ y n(fc) ] C {0}, v>(/i)(y*)^0. Let
_ k- hk v(h}(ykY and fk=Xh. Then /fc[A" \ F n(fc) ] C {0}, and by linearity of
v(/fc)(j/fc) = k-hk. Now let / = X^fcLi A- To see that / is well-defined and continuous, observe that every x & X has a neighborhood Ux meeting at most one member of the collection {Vn^} '• k G N}. So by (1), / \ Ux is a finite sum and hence / f Ux is well-defined and continuous. For every k let gk = ]Ci=i fi- Observe that for j > k we have
nv
=
and fj \ X \ Ui^fc+i ^n(i) = 0 - So / and gk coincide on supp(yfc), which implies that v ? (/)(?//c) —(p(9k)(yk) (Lemma 6.8.1). But
= hk = hk = k.
6.8. THE SUPPORT FUNCTION
409
So we conclude that tp(f)(yk) — k for every k. But this means that A is riot bounded in Y", which is a contradiction. D Corollary 6.8.4. Let X and Y be spaces, and let
Observe that if U covers X and supp(y) 7^ 0 for every y E Y then T\x covers Y. Corollary 6.8.5. Let X and Y be metrizable spaces and let
be a continuous linear surjection. IfU is a locally finite open cover of X then every y £ Y has a neighborhood Vy such that the collection
is finite. As a consequence, TU is a locally finite open cover ofY. Proof. By Lemma 6.8.2(3), supp(y) / 0 for every y € Y. So TU covers Y. Observe that TU is open in Y for every U E U since supp is LSC (Lemma 6.8.2(4)). If TU does not have the property stated in the corollary then there are a point y E Y and a sequence (yn)n in Y converging to y and distinct f/ n 's in U such that for every n, yn E Tun (here we use that Y is metrizable; in fact, for this argument it suffices to assume that Y is first countable). For every n let xn E supp(y n ) fl Un be arbitrary. Since the sequence (yn)n is bounded because it converges, supp{yn : n 6 N} is bounded as well by Theorem 6.8.3. Hence {xn : n € N} is bounded which implies that
Z = {xn : n G N} is compact since X is metrizable (cf. Exercise A.5.14). Since U is locally finite, this implies that Z intersects finitely many elements of U only. This is a contradiction. D
410
6. FUNCTION SPACES
Let X be a space. We say that a subset A C X is C'-imbedded in X if every continuous function /: A —> M can be extended to a continuous function /: X -> M. Observe that by the Tietze-Urysohn Theorem, every closed subspace of a normal space X is C-imbedded in X. The situation for nonnormal spaces is more complicated. Typical examples of closed and discrete C-imbedded subsets of an arbitrary space X can be obtained in the following way. Let f:X —> E be continuous and unbounded. It is easy to construct a sequence (xn)n in X such that \f(xn+l)\>\f(xn)\ +l for every n. We claim that D = {xn : n G N} is C-embedded in X. To prove this, first observe that f[D] is a closed discrete subspace of E. Hence f[D] is (7-imbedded in R by the Tietze-Urysohn Theorem. Observe that is a bijection. Let g : D —>• M be any function. Define the function
h: f[D] ^ R in the obvious way by h(f(d))=g(d). Then h is continuous, f[D] being discrete. By the above we can therefore extend h to a continuous function h : E —> R. Now let g = hof:X ->R. Then g is obviously continuous. In addition, if d G D then g(d) =
hof(d)=hof(d)=g(d).
We conclude that g is the required continuous extension of g. Proposition 6.8.6. Let X and Y be spaces, and let be a continuous linear surjection. Then for each closed and bounded subset K C X, the set L = {y£Y: supp(y) C K} is closed and bounded in Y. Proof. By Lemma 6. 8. 2(3), (4), supp is LSC and hence L is closed. If L is not bounded, it contains by the above remarks a discrete (faithfully indexed) Cimbedded subset D — {yn : n G N}. For each n, let z£supp(yn)
6.9. NONEXISTENCE OF LINEAR HOMEOMORPHISMS
411
(see (**) on Page 406). Then tn > 0. For if tn = 0 then \(yn,z) = 0 for every z G supp(yn) so that (p(f}(yn} = 0 for every / 6 CP(X). Since ? is surjective, this means that g(yn) = 0 for every g 6 CP(Y). This is absurd. Let g e C*(y) be such that g(yn) — tn. (That g exists follows from the fact that D is discrete and C-imbedded.) Since (f> is a surjection, there exists an element / E C(A^) such that ?(/) = g. Since X is bounded, there is c G M such that /[-K"] C [— c, c]. Let n e N be such that n > c. Then
This contradiction proves the proposition.
D
6.9. Nonexistence of linear homeomorphisms
It is possible to prove that certain topological properties are shared by spaces having linearly homeomorphic function spaces. We will give three such examples: compactness, dimension and topological completeness. See also the next section for a related result. We interpret these results as obstacles for obtaining linear homeomorphisms. Since it will follow that if X and Y are ^-equivalent then X is compact if and only Y is compact, C P (H) is not linearly homeomorphic to C*p(ffi). Interestingly, these space are homeomorphic by a result of GUL'KO and KHMYLEVA [183]. So the results that we prove here do not imply the existence of linear homeomorphisms, but do imply their nonexistence. There are many other examples and better theorems than the ones presented here. See e.g., the surveys of ARHANGEL'SKII [25, 27, 28] and BAARS and DE GROOT [35] for more information. Compactness.
Theorem 6.9.1. Let X and Y be spaces, and let (p: CP(X] -» CP(Y) be a linear homeomorphism. Then X is cr-compact if and only ifY is a-compact.
412
6. FUNCTION SPACES
Proof. Assume that X is cr-compact. We claim that L(X) is
is continuous, and has therefore cr-compact image. But the images of the a n 's cover L(X) since X generates L(X) (Proposition 6.7.2). As a consequence, the space L ( X ) is cr-compact. Now Y C L(Y) is closed by Proposition 6.7.4, and since L ( X ) and L(Y) are linearly homeomorphic by Theorem 6.7.5, it follows that L ( X ) contains a closed homeomorphic copy of Y. As a consequence, Y is u-compact D Corollary 6.9.2. Let X and Y be spaces, and let (p: CP(X) ->• CP(Y) be a linear homeomorphism. Then X is compact if and only if Y is compact. Proof. Let X be compact. Then Y is cr-compact by Theorem 6.9.1. But by an application of Corollary 6.8.4 it also follows that Y is pseudocompact. As a consequence, Y is compact (ENGELKING [153, Theorem 3.11.1]). D We already know by the result of GUL'KO and KHMYLEVA cited above that CP(I) and CP(R) are homeomorphic. So a generalization of Corollary 6.9.2 for (topological) homeomorphisms is not possible. We present other examples of this phenomenon which are based on the results obtained in Chapter 5. Example 6.9.3. A compact space X and a noncompact space Y such that their function spaces CP(X) and CP(Y) are homeomorphic. Proof. Let X = w+1. We claim that CP(X] w B(Q)°°. Indeed, let A = {oj}. By Corollary 6.6.8 we have CP(X) ~ CP,A(X) x CP(A) = CP,A(X) x E. But CP,A(X) is clearly linearly homeomorphic to c0 = {x e R°° : lim xn = 0}. Tl—>-00
Since c0 x E ~ CQ it follows by Corollary 5.5.16 that CP(X) - CP,A(X] x E ~ c 0 x M ~ c 0
6.9. NONEXISTENCE OF LINEAR HOMEOMORPHISMS
413
Let Xn = X for every n. By Theorem 6.2.8 we consequently obtain that
So the space Y = uj x (uj + 1) is a noncompact space having the property that CP(Y) is homeomorphic to CP(JQ. D In §6.12 we will obtain a far reaching generalization of this result. Dimension. Theorem 6.9.4. Let X and Y be separable metrizable spaces with CP(X] ~ CP(Y). Then X and Y have the same dimension. Proof. Let (p: CP(X) —»• CP(Y) be a linear homeomorphism. For each n 6 N, put Yn = {y 6 Y : \ supply)] = n}. Claim 1. The set Zn = Y \ \J™=1 Yi is open in Y for every n. Proof. Let y G Zn be arbitrary. Since supply) has cardinality greater than n, there is a family V consisting of n + I pairwise disjoint open subsets of X such that supplp(y) n V ^ 0 for every V 6 V. Since supp^ is LSC by Lemma 6.8.2(4), there is a neigborhood W of y such that for every y' 6 W and y E V we have supply') n V ^ 0. Clearly, W C Zn. 0 By Claim 1, YI is closed and hence an Fa -subset of Y. Similarly, Y% U YI is closed in Y and so Y?, being open in Y-z U Yi , is an ^-subset of 12 U Yi and hence of Y . By the same reasoning it follows that Yn is an Fff -subset of Y for every n. By symmetry, Xn = (x E X : \ supp _i(x)| = n} is an F^-subset of X for every n. Fix n £ N, and write Fn = U^-Ai, where each Ai is closed in Y". Fix * e N, and pick an arbitrary y e Ai. Let supply) = {xi, . . . ,xn}. For every j < n let Vj be an open neighborhood of Xj such that V^- r\Vj> = 0 if J 7^ j' • By the same reasoning as in the proof of Claim 1, there is a neighborhood U o f y such that for every z 6 U we have supptf(z) n V} ^ 0 for every j < n. As a consequence, £/' = UnAi is a neighborhood of y in A^ such that for every z 6 C/' we have supp^zjnV} ^ 0 for every j < n. We conclude that supp (z) CiVj is a single point for every j
414
6. FUNCTION SPACES
Now we do the same thing for every Xj. Fix j; < n and let HJ G N be such that Xj G Xnj, and let s u p p ^ - i ( x j ) = { 7 / 1 , . . . , y n j } - There are a neighborhood AJ of Xj in Xnj, and pairwise disjoint neighborhoods -Bi,... ,Bn. of ? / i , . . . , |/nj, respectively, such that for every k < rij the function pk that sends each t G A3 onto the unique element in supp^-i^) n Bk is continuous. Now by Lemma 6.8.2(5) we have y G supp^-i (supp (y)). Assume that j is the index for which y G s u p p ^ - i ( x j ) , say y = y^. Then p/t ° Kj(y) — Pk(%j) = Vk = y- So the continuous function Pfc OKJ: U' -+Y
has y as fixed-point. The set of all fixed-points F of pk° KJ is a closed subset of U' (Exercise A. 1.9) and is mapped by TT, homeomorphically onto a subset of X. Observe that F is an F^-subset of Y. Since each neighborhood involved in this process can be picked from a countable base, it follows that we in fact defined countably many functions only, Each of those is defined on an Fa-subset of F, and is the composition of a map into X followed by a map into Y. Their fixed-points sets are Fasubsets of Y, and, as we proved, cover Y. In addition, they are mapped homeomorphically onto subsets of X. By the Countable Closed Sum Theorem 3.2.8, one of those fixed-point sets must have the same dimension as Y. The Subspace Theorem 3.2.9 therefore implies that dim AT > dimY. By symmetry, dimF > dimA^. So we conclude that X and Y indeed have the same dimension. D Remark 6.9.5. It is an interesting open problem whether ^-equivalent spaces have the same dimension. CAUTY [90] recently proved that if X and Y are metrizable compacta such that there is an open surjection from CP(X) onto CP(Y) then Xk strongly infinite-dimensional for some integer k implies Yp strongly infinite-dimensional for some integer p. As a consequence, the Hilbert cube Q is not t-equivalent to any finite dimensional metrizable compactum. MARCISZEWSKI [271] proved, modifying Cauty's technique and using an idea from GUL'KO [182], that if X and Y are metrizable spaces such that CP(X) ft* CP(Y) then X is countable dimensional if and only if Y is. See §6.11 for details. Completeness. We will now present our last result in the same spirit. First, let us prove the following preliminary result. Lemma 6.9.6. Let (X, g) be a metric space which is not completely metrizable, and let {Un : n G N} be a collection of open covers of X such that mesh(Un) < l/n f°r every n. There are a strictly increasing sequence (in}n of natural numbers and for each n an element Un G Uit such that Un+i C Un
6.9. NONEXISTENCE OF LINEAR HOMEOMORPHISMS
415
Proof. Let (X,g) denote the completion of X (ENGELKING [153, p. 273]). For each U 6 Un there is an open set V\j C X such that V\j n X = U . Since U £ Un is dense in V\j and diam (£7) < yn it follows that diam Vu < l/n as well. Let Vn = IJi^f/ : U G Un}- Then 14 is open in X, and Jf C Vn. So
n—l
is a GVsubset of X, and hence is completely metrizable (cf. Theorem A. 6. 3). So V \ X / 0 and we may pick an arbitrary element x in that set. Claim 1. There are a strictly increasing sequence (in}n of natural numbers and for every n a Un G Uin such that x G Vj/n and Vun+1 C Vf/ n . Proof. Let zi = 1 and let C/i E Ui be such that x G V^ . Let m > 1 and suppose that ii, . . . , im-i and t/i, . . . , C/TO-i have been constructed. Let
6 = g(x,X\VUm_l). Pick im > im-i such that l/i m < 8. There is an element Um G Uim such that x G V[/m- Since diamVjy^ < l/im < 8 and x G Vf/ m , V^m C Vum_l. So we are done. <(> Since diam V[/m -»• 0, it follows that f|m=i ^- n^T = {x}nX = 0. Also, for all m we have
c/m+i = VUm+l nx cvUmnx = um. (Here the first bar means closure in X and the second bar closure in X.)
D
Theorem 6.9.7. Let X and Y be metrizable spaces, and let
be a continuous linear surjection. If X is completely metrizable then Y is completely metrizable. Proof. Suppose that Y is not completely metrizable. Let g be an admissible complete metric on X. There are by Corollary 6.8.5 and the fact that every metrizable space is paracompact, locally finite open covers Un of X and Vn of Y for n G N such that (1) mesh(U n ) < i/n and mesh(V n ) < Vn, (2) Un+i refines Un for every n, and similarly for the V's, (3) for each V G Vn the collection
is finite.
416
6. FUNCTION SPACES
By Lemma 6.9.6 we may assume without loss of generality that for each n there exists Vn 6 Vn such that 0 ^ Vn+\ C Vn and D^Li Vn = 0Let for every n, (4)
U'n = {U£Un:VnnTu^®}.
By (3), U'n is finite. Claim 1. \JU'n+lC(JU'n. Proof. Pick an arbitrary element E 6 U'n+1. Since Un_(-i refines Un there exists F 6 Un such that E C F. Since E E U^+1 there is ?/ 6 Vn+i such that supp(y) n E ^ 0. But then supp(?/) n F / 0 as well. Since y e Vn+i C Vn, this implies that F £ U^ and so E C F C |J U'n, as required. 0 Notice that by (4), supp(yn) C (JU'n for every n. For each n pick an arbitrary element yn G Vn. So supp(y n ) C Claim 2. JC is compact. Proof. Since 7^ is a closed subspace of the topologically complete space X, it suffices to prove that Q\K xK is totally bounded (Theorem A. 6.1). Let e > 0, and let j 6 N be so large that lfj < e/2. For U 6 U^-, pick an arbitrary element z(U) e L7". Since diamU < 1/j, clearly U C S(2;(L / "),e). This implies by Claim 1 that oo
U supp(y n ) C \J{B(z(U),e) : C/ 6 Uj}. n=j
su
But |Jn=i PP(?/n) is finite and so we are done.
<(>
By Lemma 6.8.6 it follows that L = {y E Y : supp(y) C K} is a closed and bounded subset of Y. Since Y is metrizable, this means that L is compact (cf. Exercise A. 5. 14). Since H^Li ^ — 05 (yn)n C L is a sequence without convergent subsequence, which is a contradiction (Theorem A. 5.1). D Remark 6.9.8. In §6.12 we will show that CP(Q) and Cp(co + 1) are homeomorphic. So Theorem 6.9.7 is not true for homeomorphisms since Q is not topologically complete (Corollary A. 6. 7). 6.10. Bounded functions In this section we derive a result on spaces of bounded functions that is in the same spirit as the results obtained in the previous section. Our results demonstrate a striking difference between CP(X) and C*(X), and provide for compact spaces new nonexistence results for linear homeomorphisms.
6.10. BOUNDED FUNCTIONS
417
Let X be a space, and let A C X. The derived set Ad of A in X is the set of all accumulation points of A in X. That is, the set of all points x G X such that every neighborhood U of x intersects A \ {x}. Observe that Ad is a closed subset of X . It is not so that Ad is always a subset of A. For let A = { 1/n : n G N} and X = E. Then 0 e Ad. For a space X and ordinal number a we define X^ a \ the a-th derivative of X, by transfinite induction as follows: (1) *° = X(2) if a is a successor, say a = ,3 + 1, then A'( a ) is the derived set (X^)d ofX^ inX; (3) if a is a limit ordinal then X^ = f| j g
Lemma 6.10.1. Let X be a space, and let a, (3 be ordinal numbers. Then (1) X^ is a closed subset of X. (2) Ifp
Proof. All three statements follow easily by transfinite induction noting that derived sets are closed. D Let A be a subspace of X. Then clearly A is dense-in-itself iff A C Ad. We say that A is scattered if A contains no dense-in-itself subsets, i.e., every subset of A contains a (relative) isolated point. Theorem 6.10.2. If X is scattered then there exists an ordinal number a such that X^ = 0. Proof. Since X is scattered, for every ordinal number a such that X^ ^ 0 we have that X^a+l^ is a proper subset of X^ . This is so since X^ has an isolated point. This means that if X^ is nonempty for every a then
for every a. But this is impossible since Lemma 6.10.1(2) shows that the family of subsets («+i) \ X(<*)
: a an
ordinal number}
of X is pair wise disjoint. Hence for an ordinal a > \ X\ we must have X^ = 0
and this is as required.
D
418
6. FUNCTION SPACES
The scattered height n(X) of a scattered space X is defined to be the smallest ordinal a such that X^ = 0. It will be convenient to denote C*(X) with the topology of uniform convergence (see Page 7) by C*(JC). Similarly, its subspace consisting of all (bounded) functions vanishing on a closed subset A of X is denoted A family 3~ C C(X) is equicontinuous if for every x G X and e > 0 there is a neighborhood U of x in X such that for each / £ 3~ and y € f/, Proposition 6.10.3. If ^ C C*(JC) is compact then it is equicontinuous. Proof. Let x € X and e > 0. The family {B(f, e/s) : / 6 9"} is an open cover of 5F. Since 9" is compact, there are /i,...,/ n e J such that the family {B(fi,£/3) : i < n} covers 9r. Since each fi is continuous, there is a neighborhood U of x such that for all y 6 U and for every i < n, Now let / £ ? and y 6 [/" be arbitrary. There is i < n such that / E -£?(/;, So Since y 6 C7, we consequently have as required.
D
We now come to our first main result. Theorem 6.10.4. Let X and Y be first countable spaces and i* -equivalent. Then X is discrete if and only ifY is discrete, or, in terms of K, K(X) < 2 if and only if K,(Y) < 2. Proof. Striving for a contradiction, suppose that K,(X) < 2 and K(Y) > 2. Since X is not empty we have n(X) = 1 so that X is discrete. Since n(Y) > 2, some y € Y is not isolated. Let {Un : n < u)} be a decreasing open base at y. For every n let f n : Y -> I be a Urysohn function with fn(y) — 1 and fn[Y \ Un] C {0}. Then fn —>• X{y} pointwise in RY , where X{y} is the characteristic function of {y}. Since X{y} & C*(Y), {fn : n < uj} is closed and discrete in C*(Y). Now let ip: C*(X) —> C*(Y) be a linear homeomorphism. By Corollary 1.1.16, (p: C^(X) —> C7*(y) is also a linear homeomorphism. Since C^(X) and C*(y) are Banach spaces, by Exercise 1.1.31 there is a k e N such that for every / € C*(X) we have
6.10. BOUNDED FUNCTIONS
419
Let gn = (p~l(fn) for every n. Then \\gn\\ < k\ f n \ = k. Hence (gn}n is a sequence contained in the compact space [ — k , k ] x . So the sequence (gn)n has an accumulation point g G [— k, k]x . Since X is discrete, g is continuous (and bounded) on X. However, since {fn : n < cu} is closed and discrete in C*(Y), the same is true for {gn : n < co} in C P ( X ) . This is a contradiction. D The question naturally arises whether for all scattered metrizable t*equivalent spaces X and Y and for all n 6 N we have K,(X) < n iff K,(Y) < n. For n — 1 this is trivially true for then both X and Y are empty, and forn = 2 it is true by Theorem 6.10.4. But for larger n it is false. Example 6.10.5. For each n > 2 there are scattered compact t* -equivalent spaces X and Y such that n(X) — 2 and K(Y) = n + 1. Proof. Indeed, put X — uj + 1 (this X works for all n simultaneously) and Yn = (a; + l) n for n > 2. Then K,(X] = 2 and K,(Y) = n + 1. We claim that Cp(Yn) ~ CppC) for all n > 2. We will only prove this for n = 2. The remaining part of the proof is left as an exercise to the reader. Indeed, put Y = Y% and let A = { ( x , y ) £Y : (x = u)V (y = u)}. Observe that A is homeomorphic to uj + 1 and that Y/A « a; + 1. The last statement follows since Y/A is the one point compactification of a countable discrete space, and one point compactifications are topologically unique. By Corollary 6.6.13 and Theorem 6.2.8 we therefore find that CP(Y) ~ CP(Y/A) x CP(A)
For the last homeomorphism, observe that if Z — (uj + 1) © (w + 1) and 5 denotes the two non-isolated points in Z then Z/B K LJ + 1, for which it follows by the same reasoning and Corollary 6.6.10 that CP(Z) ~ Cp(u + 1) x M2 - C p (w + 1),
as required.
D
Theorem 6.10.4 can be generalized though. We will show below that if X and Y are i* -equivalent scattered metrizable spaces then n(X] < uj if and only if K,(Y) < uj. Before proving this result, we need to derive two preliminary results first. Lemma 6.10.6. Let X be a metrizable space with n(X) < uj. Then there is a metrizable space Y such that K,(X) = K,(Y) and C*(X] ~ C^A(Y), where A = Y^ .
420
6. FUNCTION SPACES
Proof. We prove the lemma by induction on K(X). If K(X) = I then X is discrete, so Y = X will do. So suppose the lemma to be true for all metrizable spaces X for which n(X} < n, where n > 1. Let X be a metrizable space with K.(X) - n, and let B = X^. Then by Corollary 6.6.8 we have that Since n(B) = n — 1, there is by the induction hypothesis a metrizable space Z such that K(Z) = K(B) and C*(B) ~ C*>C(Z), where C = Z^ . Then
c;tC(z) x c;tB(x) ~ (cf. Theorem 6.2.8). Let Y = Z © X. Then yW = B U C and Ac(F) = So we are done.
K(X).
D
If -B is a set then we order its finite nonempty subsets by inclusion. Clearly (J^B) \ {0}, C) is a directed set. Let X be a space and B an infinite set. For every b E B let f\, 6 E.x be such that for every x £ X the set {6 G -B : fb(x) ^ 0} is finite. Define the function ^bB fb '• X —> R by the following formula:
Then obviously Y^b&B fb *s well-defined. Lemma 6.10.7. Let X be a space and B an infinite set. For every b e B let fb 6 Ex be such that for every x E X, {b£B: fb(x) ± 0}
is finite. For S e S = Joo(#) \ {0} let fs = Eb € s fb- Then { f S : S e §} is a net in Ex and limse§ /s = Ebes faProof. It is clear that {fs : S G §} is a net in Rx . Now let e > 0 and PCX be finite. For every p £ P let Sp = {b £ B : fb(p) ^ 0}. Put 50 = U pe p SPThen 5o E S. Let 5 6 § be such that 5 D 5o, p G P, and / = E 66 £ /&• Then fees
EeSp /*(p) - 6es E /* p
=0 < e.
Hence, Hm S6S fs = f .
D
6.10. BOUNDED FUNCTIONS
421
Theorem 6.10.8. Let X and Y be i* -equivalent metrizable spaces. Then
Proof. Striving for a contradiction, assume that K(X) < u and K(¥) > u and that X and Y are £* -equivalent metrizable spaces. By Lemma 6.10.6 we may replace C*(X) by C*^A(X) with A = X^ . Let y. C^A(X] -> C*(Y) be a linear homeomorphism. Then by Corollary 1.1.16, (p: C*u A(X) —>• C*(F) is also a linear homeomorphism. So there is by Exercise 1.1.31 a k E N such that for every / E C * A(X] we have \\f\\/k<Mf}\\
belongs to C* A(X). Notice that every / E C * * A ( X ) can be written as
where for x € B, ctx = f ( x ) . Put gx = (p(fx}. For every y E Y we put Cy = {x E B : gx(y] ^ 0}Claim 1. Cy is finite for every y € Y. Proof. Suppose that Cy is infinite for some y E Y. Let (xn)n be a sequence in Cy such that xn ^ xm if n ^ m. Define hn '• X —>• E by , n
_
fxn
—
„ (y)
/
N '
Then hn E ^^(A") for every n and the sequence (hn}n converges pointwise to 0, the constant function with value 0, in C* A(X}. Now pxri (y)
gXn (y)
.
Hence (p(hn) yV 0 in C*(Y). This contradiction proves the claim.
Now define b : Y ->• R by
Notice that for every y E F we have b(y) — X^xec 1^(^)1' hence b is welldefined. Claim 2. H6II < 2k.
6. FUNCTION SPACES
422
Proof. For y e Y, let =
{x€B:gx(y)>0},
C~ =
respectively. Notice that ^ 9.
Similarly we can prove that || ^x^c- gx\\ < k. So for every y £ Y we have
E
9x(y] - Y^
E
which proves the claim.
0
Now for P C B finite, let
Mp = I 2_, &xfx '• \&x\ < k for x G P >. Notice that xeP
X<EX\P
is a compact subset of M^, hence of C* A(X). It is evident that Mp is also a compact subset of C** A(X). Claim 3. For every y E Y, P C B finite, and e > 0, there is a neighborhood U(y, P, e) of y in F such that for each 2 £ C7(y, P, e) and / 6 y?[Mp] we have |/(y) - /(z)| < e. Proof. Since tp[Mp] is compact in C*(y), Proposition 6.10.3 implies that it is equicontinuous. From this, the claim follows immediately. 0 Now let A7' £ N be so large that > 2k.
Claim 4. There are yo, • • • , y./v £ y, PO, • • • , PN C S finite, and C / o , . . . , neighborhoods of yo ; • • • 5 yjv respectively, such that (1) (2) (3) (4) (5)
for every 0 < z < TV: Cyt C P,, PoCPi C - . - C P ^ , U0DUiD---DUN, for every 0 < i < N: U, C U(yl:Pl: i/4), for every 0 < i < N: yx e Y^N~^.
6.10. BOUNDED FUNCTIONS
423
Proof. We will prove this claim by induction. Since K(¥) > w, Y^ =£ 0 and hence we may pick Let P0 = Cya and C70 = U(y0,P0, y 4 ). Suppose
are found for 0 < n < TV. Since yn £ y(^~ n ) and N — n > 1, we can find
j/n+i € (!/„ \ {yt :i < n}) n y^-^1)). Let -fn+l
=
-fn U Cyn + l
and
c/n+i = c/ n nc/(y n+1 ,p n+1 , i/4). It is easy to see that these choices satisfy our inductive hypotheses. Now let g : Y —> JT be a continuous function such that g(yi) = (-lY
(0
Then \\g\\ = 1, so ll^"1!^)!! < ^- Hence t^"1(^) can be written as
where \ax < k for every x £ B. Notice that
for every 0 < i < N. Claim 5. g = ^x&Baxgx. Proof. Let § = ^(B) \ {0}, and for every S e § let /5 = Ex 6 5 a ^/^ Lemma 6.10.7, ^~l(g) = Iim56§ fs and E X J B a ^^ = Iim5es a x f/ x . So
which proves the claim.
<(>
6. FUNCTION SPACES
424
Let 0 < i < N. Since Cyi C P{ (Claim 4(1)), we have by Claim 5, (-1)1 = g(yi) =
a
x9x(yi) =
axgx(yi).
By Claim 4(3) and (4), yN G U(y^P^ Y 4 ). Furthermore,
So by Claim 3,
E
IN,
-
If i > 0 then by the same argument,
We conclude that in case i > 0, we have by the above and Claim 4(2),
axgx(yN) -
E a g (y*r\IN. - V^ x x
axgx(yl-i) -
V" axgx(yN) ± 2
> 2-
-E >3/4-
If i = 0 then with P_i = 0 we have 3
So by Claim 4(2),
x£PN
A-
6.10. BOUNDED FUNCTIONS
425
From this we conclude that
~(N + i)>2k, which contradicts Claim 2.
D
Corollary 6.10.9. Let X and Y be l-equivalent compact metrizable spaces. Then n(X) < uj if and only if K(¥) < u. So we see for example that if X = uj + 1 and Y is the one point compactification of the topological sum of the spaces (a; + l) n , n e N, then X 7^ Y because K(X) — 2 and n(Y) = u + 1. We are now in a position to present the following: Example 6.10.10. There are ^-equivalent countable, locally compact (hence metrizable) spaces which are not £* -equivalent. Proof. Let X be the topological sum of countably many copies of oj + 1. Then
n=l
by Theorem 6.2.8. In addition, let Y be the topological sum of the spaces
Then oo
P ((
W
oo
+ !)") ~ n
again by Theorem 6.2.8 and the computations in Example 6.10.5. Observe that n(X} = 2 and K,(Y) = ui. So X and Y are not £*-equivalent by Theorem 6.10.8. D Remark 6.10.11. It is unknown whether there are £* -equivalent (countable) spaces which are not ^-equivalent. BAARS [31] proved the following interesting result: Theorem 6.10.12. Let X and Y be locally compact zero- dimensional separable metrizable spaces. Then the following statements are equivalent: (1) X and Y are I* -equivalent. (2) There is a linear homeomorphism ?: CP(X)
—»• CP(Y) such that
426
6. FUNCTION SPACES
So if such X and Y are £*-equivalent then they are ^-equivalent. Examples of £*-equivalent spaces which are not ^-equivalent must therefore be rather complex. Observe that a linear homeomorphism (p\ CP(X) —> CP(Y) does not necessarily map C*(X) onto C*(Y) by Example 6.10.10. Remark 6.10.13. Similar questions about the relation between the function spaces CP(X) and C*(X] for topological homeomorphisms are also quite interesting. For countable spaces, some definitive answers are known. BANAKH and CAUTY [40] proved the following nice result: Theorem 6.10.14. If X is countable and nondiscrete then C*(X) is homeomorphic to CP(X] x a. This implies that if X and Y are countable, nondiscrete and ^-equivalent then they are t*-equivalent. The converse of this statement is false. By a result of MARCISZEWSKI and VAN MILL [272] there are countable nondiscrete t* -equivalent spaces which are not t-equivalent. 6.11. Nonexistence of homeomorphisms In view of the results obtained so far the question naturally arises what can be said about homeomorphisms between function spaces which are not necessarily linear. We will approach this as in §6.9, i.e., we try to find topological properties shared by spaces having homeomorphic function spaces. We interpret the results in this section as obstacles for obtaining homeomorphisms. In the next section we will present positive results, i.e., we will prove using the infinite-dimensional apparatus developed in the Chapter 6 that many function spaces are homeomorphic. Homeomorphisms are much more flexible than linear homeomorphisms. So they are easier to construct but, on the other hand, their nonexistence is harder to prove. An interesting result is due to OKUNEV: he proved (among other things) that if X is cr-compact and Cp(X) is homeomorphic to CP(Y) then Y is acompact (see OKUNEV [328] or ARHANGEL'SKII [27, Corollary III.2.12]). For metrizable spaces X the situation is even better: a metrizable space X is acompact if and only if CP(X) is analytic (CHRISTENSEN [100]). Another interesting result in the same spirit was obtained recently by CAUTY [90]. He proved that if X and Y are metrizable compacta such that there is an open surjection from CP(X) onto CP(Y] then Xk strongly infinitedimensional for some integer k implies Yp strongly infinite-dimensional for some integer p. As a consequence, the Hilbert cube Q is not t-equivalent to any finite dimensional metrizable compactum. We will present a refinement of the method CAUTY developed in [90] here which is due to MARCISZEWSKI [271] and which will lead us to the following result: if X and Y
6.11. NONEXISTENCE OF HOMEOMORPHISMS
427
are metrizable, X is countable dimensional and CP(X) & CP(Y), then Y is countable dimensional. We begin by deriving a few auxiliary results. It will be convenient to introduce the following notation. For an integer k let [X}k be the subspace of 3'00(X) consisting of all sets of size precisely k. Lemma 6.11.1. Let X and Y be metrizable spaces and let a: X —> [Y]n be continuous. Then there are a countable open cover U of X and for every element U G U and k < n a continuous function ffr: U -» Y such that for every x G U G U we have (*} \*)
rv(r} O^x; --
Proof. Let x G X be arbitrary and let a(x) = {y^ ..., y™}- There are pairwise disjoint open sets V£,. • • , V™ in Y such that yk G Vk for k < n. The set Wx = {x1 G X : (Vfc < n)(a(x'} n Vxk ^ 0)} is by continuity of a an open neighborhood of x in .Y and, for every k < n, the map x' ^>a(x')nV£ is single-valued and continuous on Wx. Since metrizable spaces are paracompact, the cover W = {Wx : x G X} has a a-discrete open refinement, say V. Let V = Ui^i "Vi, where each Vi is discrete. For every i, put Ul = \J V.L. Then U = {Ul: i G N}
is a countable open cover of X and we will show that it is as required. To this end, fix i and observe that for every V G V; there exists x(V) G X such that y C Wx(yy For A; < n we let £y be the restriction of the function
x' H- a(x') n v;fc(F) to y. Then ^y is evidently continuous. Since Vz is discrete, the function fk -- IUI f=^kv JUi vev z is continuous on Ul. It is clear that with these functions (*) is satisfied.
D
Lemma 6.11.2. Let A and B be Gs-subsets of the metrizable spaces X and Y, respectively. In addition, let f : A—^Y and g : B —>• X be continuous functions such that f o g is the identity on B. Then g : B —> g[B] is a homeomorphism and g[B] is a Gs-subset of X. Proof. The map g: B —>• g[B] is a homeomorphism since by assumption,
f \ g[B]: g[B] ^ B is its continuous inverse. It therefore remains to show that g[B] is a Gsof X. Put S = f ~ l [ B } . Then 5 is clearly a G^-subset of X since / is
428
6. FUNCTION SPACES
continuous. In addition, g[B] is contained in 5 and / \ g[B] : g[B] —> B is a homeomorphism. So by Exercise A. 8. 3 we conclude that g[B] is a G^-subset of 5 and hence of X. D We now formulate the main result in this section. Theorem 6.11.3. Let X and Y be metrizable spaces. If CP(X) and CP(Y) are homeomorphic then Y is a countable union of G§ -subsets which are homeomorphic to G$-subsets of X and vice versa. Corollary 6.11.4. Let X and Y be t-equivalent metrizable spaces. Then X is countable dimensional if and only if Y is countable dimensional. Proof. Suppose that X is countable dimensional. Since countable dimensionality is hereditary, it follows that the GVsubsets of Y we get from Theorem 6.11.3 are all countable dimensional. Hence Y is a countable union of countable dimensional subspaces, hence is countable dimensional itself. This completes the proof since the roles of X and Y can be interchanged. D Corollary 6.11.5. Let X be a finite dimensional metrizable space. Then the function spaces CP(Q) and CP(X) are not homeomorphic. Proof. Simply observe that Q is strongly infinite-dimensional and hence not countable dimensional (Corollary 2.4.13 and Exercise 3.1.6) and that every finite dimensional metrizable space is countable dimensional (cf. Corollary 3.3.9 and ENGELKING [155, Theorem 4.1.17]). D To begin the proof of Theorem 6.11.3, let us fix some notation. For A and B finite subsets of X and Y, respectively, we put for every n > 1, V(A,n) = {/ € CP(X) : |/(z)| < Vn for x G A}, W(B,n) = {ge CP(Y) : \g(y) < i/n for y E B}. Observe that V(A,n) is a basic open neighborhood of the zero 0 of CP(X), i.e., the constant function with value 0. Similarly for W(B, n). For x G X and y 6 Y we also put for every n > 1, V(x,n) = { f £ C p ( X ) : \ f ( x ) \ <
i/n},
Let (p: CP(X) —> CP(Y) be a homeomorphism. Since CP(Y) is homogeneous, being a linear space, we may assume without loss of generality that c/?(0) = 0. For n, m > 1 and y 6 Y we put e(j/, m, n) = {A e S^TO \ {0} :
6.11. NONEXISTENCE OF HOMEOMORPHISMS
429
Lemma 6.11.6. Y = \J^=1 \J™=1 C(k,m,n) for every n > 1. Proof. Fix n > I and y £.Y. Then W(y, n) is a neighborhood of 0 in CP(Y). Hence by continuity of (p and the fact that
D
Lemma 6.11.7. The set C(k, ra, n) is closed in Y for every k,m,n > I. Proof. Let (yi)i be a sequence in C(fc,ra,n) converging to a point y 6 Y. For each % let ^ C X be a set of size at most k such that Claim 1. There are an infinite subset E C M and a (possibly empty) subset A° of X and for every i E E a partition y4° , A\ of ^4Z such that that one of the following statements is true: (A) Every x E X has a neighborhood meeting finitely many terms of the sequence (Ai)i£E only. (B) A° ^ 0 and (1) the sequence (A^)i^E converges to A° in the Vietoris topology. (Observe that \A°\ < k because (A^)ieE -> A° and \A%\ < k for every i e £^ (Exercise 1.11.3).) (2) Every a; e X has a neighborhood meeting finitely many terms of the sequence (A\}i^E onlyProof. This is precisely the statement of Exercise 1.11.20.
0
If (A) holds then we redefine ^4° to be the empty set. We also put A® = 0 and A] — Al for every i € E. So every x G X has a neighborhood Ux meeting finitely many terms of the sequence (A\}i^E only. We may assume without loss of generality that E — N. We claim that which will show that y G C(k,m,n). Incidentally, this also shows that A° ^ 0, and hence that (A) does not hold, for if A° = 0 then and so (p[V(A°,m)] = CP(Y) <£. W(y,ri). (This observation will be used in the proof of Lemma 6.11.8.) Striving for a contradiction, suppose there is a function / e V(A°,m) such that
430
6. FUNCTION SPACES
of (f> there are p > 1 and a finite set B C X such that for every g G CP(X) with / — g G V(B,p) we have \<+>(g}(y}\ > l/n. We may assume that A° C B. Since / is continuous there is an open set U C X containing A° with
Since every point in X has a neighborhood meeting finitely many Aj's only, there is a neighborhood V of the finite set B and an integer ZQ such that
V n A] = 0 for every i > z'0. Put U' = U n V. If (A) holds then A° = 0 for every i, and so .4° C U' for every i. If (B) holds, A° —>• ,4° which implies that for all but finitely many i, A° C C7'. So there is an integer ii > IQ such that
Aj n (B u Z77) = 0 and
A°iCU' for every i> ii. Since B U U' and |Ji;>i ^1 are disjoint closed sets in X, there is by the Tietze-Urysohn Extension Theorem a function g G CP(X} such that 0(z) = /(x) for all x G 5 U U7 and 0 f U I>tl Al =Q.Sog\A\=Q for i > n so that ^ G V(Ai,m). But then |?(5 f )(2/i)| < Vn ^or * > H while on the other hand / — g G F(-B,p) and so |y?(<7)(y)| > Vn by the above. This contradicts the continuity of (p(g). CH For natural numbers m, n > 1 and A; > 2 we define _E(/c, m, n) = C(fc, m, n) \ C(k — 1, m, n). So E(k,m,n) consists of all points y G C(k,m,n) such that every element of C(y,m,n) has size at least k. We also put £7(1, m,n) = C(l,m,n) for all m,n > 1. From Lemma 6.11.7 it follows that each E(k,m,n) is a G$subset of Y. In addition, by Lemma 6.11.6, 00
Y=
00
UU k=l m=l
for every n > 1. For y G E(k, m, n) we put
Lemma 6.11.8. Let t/i,y G £7(A;,ra,n) for i G N and certain fc,m,n > 1. If yi —>• y and Ai G £(y,m,n) for every i then (A^)i contains a subsequence which converges to an element o
6.11. NONEXISTENCE OF HOMEOMORPHISMS
431
Proof. By the proof of Lemma 6.11.7 we may assume without loss of generality that there are nonempty finite A° C X and A® C Ai such that A° —>• A° in the Vietoris topology, such that A° witnesses the fact that y G <7(A;,ra,n). Since y G E(k,m,n) this means that \A°\ > k and so for all but finitely many i we must have \A®\ > kby the definition of the Vietoris topology. But then for those infinitely many i we have \A®\ = k since \Ai = k for every i. This implies that \A°\ < k. We obtain \A°\ = k and so A° G £(y,m,n). D Lemma 6.11.9. If k,m,n > 1 and y G E(k,m,n] then 8.(y,m,n) is finite. Proof. Striving for a contradiction, assume that £(?/,m,n) is infinite. Claim 1. There is a set A C X with \A\ < k and a sequence (A;); of sets from £(?/, m, n) such that Ai C\ Aj = A for all distinct i, j. Proof. This follows from Exercise A. 1.4.
0
We will prove that A G G(y,m,n) which will contradict y G E(k,m,n). To this end, take an arbitrary / G V(A,m}. Our task is to prove that Striving for a contradiction, assume otherwise, i.e., \
< l/n. This is a contradiction with the D
If y G E(k:m,n) then a min (?/) denotes the union of the family So o;m,n(y) is a finite set by Lemma 6.11.9. Let £ be an arbitrary admissible metric on X and let p > k and q > I be natural numbers. Define G(k,m,n,p,q) = {y £E(k,m,n) : \amtn(y)\ = p and g ( x , x ' } > l/q for all distinct x,x' G am>n(y)}. Obviously, OO
CXD
E(k,m,n) = [J \J G(k,m,n,p,q). p=k q=l
Lemma 6.11.10. For all natural numbers A;, m,n,p and q: (1) G(&,ra,n,p, q) isaGs-subsetofY. (2) am,n \ G(/c,m,n,p, q) : G(fc,m,n,p, q) —>• [^C]p is continuous.
432
6. FUNCTION SPACES
Proof. Fix k,m,n > 1. For p > k and q > I let K(p,q] be the set of all points y E E(k,m,n) such that am,n(y) has a subset C of cardinality p such that g(x,x') > l/q for all distinct x,x' E C. Claim 1. K(p,q) is closed in E(k,m,n}. Proof. Indeed, let (yi)i be a sequence of points in K(p,q) converging to an element y E E(k,m,n). For every i let Ci = {a;J,...,a^} C am,n(yi) witness the fact that yi E K(p,q). Consider the sequence (x]}i. For every i there is an element Ai E £(yi,ra,n) containing x\. By Lemma 6.11.8 we may assume without loss of generality that Ai —> A, where A E £(y,m,n). By the definition of the Vietoris topology this implies that the sequence (x]}i has a convergent subsequence (Lemma 1.11.2(3)), say converging to an element xl E OLn^m(y}. Let E\ be the infinite set of indices that correspond to the members of that convergent subsequence. Consider the sequence (x\ )^E • By the same reasoning, it contains a subsequence converging to an element x2 E amjn(y). Etc. Since p is finite, we can therefore assume without loss of generality that xl ->• x3 E a m , n (y) for every j < p. Put C — {xl,...,xp}. Then \C\ < p by construction. But the elements of every Ci are at least l/q apart. This shows that C also has p elements which are all at least l/q apart, i.e., C is a witness of the fact that ye K(p,q). 0 Since from Lemma 6.11.7 it follows that each E(k,m,n) is a G^-subset of Y, (1) follows from Claim 1 and the equality oo
G(k,m,n,p,q) = K(p:q}\
\jK(p+l,r). r=l
For (2), in the above argument, if we take ^ and y from G(k, m, n,p, q) then Ci = Oim,n(yi},
C - am,n(y}.
So by Lemma 6.11.8 we conclude that a.m,n(yi} —> Oim,n(y}.
D
Consider the inverse ip~l : CP(Y] -> CP(X}. We define for if"1 the sets and the functions j3n^m as the sets
,n,p, q) and the functions a n ,m were defined for (p. In fact, only the value n = 1 will be important in the remaining part of this section.
6.11. NONEXISTENCE OF HOMEOMORPHISMS
433
Lemma 6.11.11. Let y G E(k,m,I) for certain k,m > 1. For x G ot.m,i(y} let kx,lx G N be such that x G F(kx,lx,m+ 1). There exists an x G a m ,i(y) such that y 6 ^ 3 . iTn +i(x). Proof. Suppose that
By the definition of ctm^ there is a subset A C am,i(y) such that
This means that ^>~ (g] G F(A, TTI) and consequently which is a contradiction.
D
We are now finally in the position to present the proof of the main result in this section. Proof of Theorem 6.11.3. Since Y is the union of the countably many G$sets G(k, m, l,p, q) it suffices to prove that each of them itself is a countable union of G^-sets homeomorphic to G^-subsets of X. Fix k,m,p and q. By Lemmas 6.11.10 and 6.11.1 it follows that , l,p, q) =
G>,
r=l
where for each r G N, Gr is open in G(k, m, 1, p, q) (hence is a G^-subset of Y) while moreover there exist continuous functions /; : G> —> X for i = 1 , . . . , p such that for every y G G>, Fix r G N. Consider the family of GVsets {//(fc',m / ,m + l,p',g') : k',m',q' > l,p > k'} which covers X, and the corresponding continuous functions Pm',m+i- By an application of Lemma 6.11.1 for these families of sets and functions, we can represent X as a countable union of G^-sets Hs, s G N, such that for every s there are continuous functions gs- : Hs —> Y for j = I , . . . , p's having the following property: for every y G Gr and si, . . . , sp G cj with fi(y) G HSi ,
434
6. FUNCTION SPACES
there exist i < p and j < p's. such that g^ (fi(y)} = y. For this last property use Lemma 6.11.11. The set Gr is the union of the G^-sets
where s = (si , . . . , sp) £ W . For a fixed s we have
These sets are closed in Gg and by Lemma 6.11.2 are homeomorphic to G$subsets of X. D Remark 6.11.12. The converse of Theorem 6.11.3 does not hold. To see this, let X and Y be countable spaces. Then X is the union of the countably many G<$-sets {{x} : x G X}. Similarly, Y is the union of the countably many Gj-sets {{y} : y G Y}. Since points are topologically unique, X and Y satisfy the conclusion of Theorem 6.11.3. But it is not so that CP(X) and CP(Y) are necessarily homeomorphic. Simply observe that if X — N then CP(X) is topologically complete and if Y = Q then CP(Y) is not a Baire space. 6.12. Topological equivalence of certain function spaces In this section we prove that if X is nondiscrete and countable and CP(X) is an F^-subset of Rx then CP(X) is homeomorphic to B(Q}°°. This implies among other things that all function spaces of countable nondiscrete metrizable spaces (e.g., Q, w+1, etc.) are homeomorphic (Theorem 6.3.10). This interesting result is due to DOBROWOLSKI, MARCISZEWSKI and MOGILSKI [123] who proved it via the BESTVlNA-MociLSKi [53] technique. Our method of proof is via the absorber method developed in Chapter 5. This yields a slightly stronger result which is due independently to BAARS, GLADDINES and VAN MILL [32] and DIJKSTRA and MOGILSKI [118]. We have to make several detours in order to arrive at the desired conclusion. Topological characterization of meager niters. We characterized meager niters in purely set theoretic terms in §6.5. We will now do the same in purely topological terms. Theorem 6.12.1. Let J be a free filter on N. Then the following conditions are equivalent:
(1) J 6 (2) y is a meager subset o
6.12. TOPOLOGICAL EQUIVALENCE OF CERTAIN FUNCTION SPACES
435
Proof. It suffices to prove the implication (1) => (2). Let U, 3C,V be subsets of 1P(N) such that It is open, X and y are meager, and J = (It \ X) U ^. Striving for a contradiction, assume that 3 is not a meager subset of J*(N). Then U ^ 0, and hence there exist an element B € U and an element t £ N such that V(£,S) C It, where 5 = {1,2, . . .,fj>. Write N0 = N \ 5 and let 3~0 = y \ N 0 . Observe that the function TT: 0>(N) ->• ^(No) defined by
A H> A n N0 corresponds to the projection from {0, 1}N onto {0, 1}N°. Hence it is an open continuous surjection (Exercise A. 5. 5(3)). In addition, every fiber of TT has evidently cardinality 2l. We claim that Indeed, if A C N0 is arbitrary then A U (B n 5) <E V(B,S) since A n 5 = 0 and 7r(^4 U (B n 5)) = A. Since 3" is a free filter, so is 3~o and hence it is dense in 9(No). Moreover, since 3~ contains a dense G^-subset of V(B,S) and 7r[V(B,S)] — J'(No) and TT is a continuous open surjection with finite fibers, 3~o contains a dense G^-subset of CP(No) by Exercise A. 1.1 7. (This can also be proved directly.) Let £: T(No) —> J'(No) be the homeomorphism assigning to each B G J'(No) the set N0 \ B. Since 3~0 is a filter, Consequently, the compact space T(No) contains two disjoint dense GSsubsets which by Exercise A. 6. 10 is impossible. D Sequence spaces from filters. Let y be a free filter on a set E C N. Define cy = {x € RE : (Ve > 0)({z € E : \xz\ < e} G J}. Observe that cy is a dense linear subspace of M^ (use that 9" is free). It is clear that if J is the so-called Frechet filter ^={^CN:N\^lis finite} then cy is c0. Finally, observe that if J is a free filter on N then cy is a closed linear subspace of the space Cp(Ng-) considered in §6.5. Indeed, cy is canonically homeomorphic to the subspace {/ e C P (N T ) : /(oo) = 0} of Cp(Ngr). This implies by Corollary 6.6.7 that
Lemma 6.12.2. Let ^ be a free filter on N. If c^ is analytic then so is (topologized as subspace of J ) (N)j.
436
6. FUNCTION SPACES
Proof. Define £: 3 —»• cy as follows:
It is easy to see that £ is an imbedding of 3 into cy . To prove that it is a closed imbedding, simply observe that £[3] — cy n {0,1 }N. So 3~ is homeomorphic to a closed subspace of an analytic space, and hence is analytic itself (Lemma A. 13. 11). D Corollary 6.12.3. Let 3~ be a free filter on N. If cy is analytic then 5 is a meager subset o Proof. If cy is analytic then so is 3" by Lemma 6.12.2. This implies by Theorem A. 13. 13 that 3 e C(?(N)) . So 3" is meager by Theorem 6.12.1. D Note that for any free filter 5" we have that CQ C cy. We now aim at proving 'universality' properties for spaces of the form cy. The following simple corollary to Lemma 6.5.2 will be useful. Corollary 6.12.4. Let 3~ be a closed and nowhere dense subset of Then for every n G N, there is a set X C {n + 1, n + 2, . . . } such that the collection {E(JX :E C {l,2,...,n}} misses 3. Proof. Let A = {1, 2 . . . , n}. By Lemma 6.5.2 there are a finite set C C N\ A and a subset D CC such that for every S C N we have
Let X — D. We claim that X is as required. To this end, let E C {1, 2, . . . , n} be arbitrary, and put Y = E U X. Since 1" n C = B we have Ye
V(Y,A)r\V(B,C).
As a consequence, Y ^ J by (*).
D
As in §5.5 we let R be the two-point compactification [—00, oo] of M. We represent Q by R N . We use the following arithmetic on Q: y0 = esc,
±00 + a = ±00
(a finite), and oo — oo = 0, etc.
The following result is needed in several forthcoming results. Proposition 6.12.5. Let 3 be a free filter on N which is a meager subset of
6.12. TOPOLOGICAL EQUIVALENCE OF CERTAIN FUNCTION SPACES
437
Remark 6.12.6. We already know from Theorem 6.5.6 and the fact that Cp(Ny) ~ c? x R that cj is not a Baire space. For if it would be a Baire space then Cp(Ny) would be a Baire space as well by Theorem A. 6. 10. So cy is a meager subset of Q (Exercise 1.6.6) and hence is meager in itself. That it is contained in a aZ-set of Q therefore comes as no surprise. One could guess from these observations that no extra work is required since it is tempting to think that a linear space which is meager in itself is in fact a countable union of ^-subsets. If this would be true then some proofs in this section would have been simpler. However, it is not true, as was shown by BANAKH [39]. His example is the linear span in £2 of Erdos' space (Example 1.5.18). It is even absolutely Borel. See also BANAKH, RADUL and ZARICHNYI [41, Theorem 5.5.19] for details. Proof. Put f = R \ ((-3,-l)u(l,3)). Define g: f ->• R by
{
x (xeJJ), x-2 (x 6 [3,oo], x + 2 (x 6 [-00, -3]).
Let g°°: f°° -> R°° be denned by g°°(x)l = g(xl] for i e N. Then g°° is obviously continuous. There is for every n a nowhere dense closed subset An of 3?(N) such that J C \J£=lAn. p ut Bn = {x 6 f °° : {i € N : \x% < 2} 6 An}
for n 6 N.
Claim 1. Bn is closed in f°°. Proof. Let x € f °° \ Bn. Then K = {i e N : Xi\ < 2} £ An
and so there is an element m G N such that if S — {1,2,. . . , m} then For i < m such that \Xi\ < 1 let Vi — J. In addition, for i < m with \Xi\ > 3 let Vi = [3, oo] if xl > 0 and V, = [-00, -3] if x, < 0. Then
is an open neighborhood of x in TN which misses Bn. For if y G Vr and L = {ieN: yl\<2}
438
6. FUNCTION SPACES
then L n 5 = K n S. This implies that L G V(K, S) and hence L <£ An.
0
Claim 2. g°°[Bn] G Proof. For each m put Put Sm = {1, 2, . . . , m}. By Corollary 6.12.4 there exists Am C {m + 1, m + 2, . . . } such that the collection {TuAm:TCSm} misses An. Define fm : Q ->• Q by
{
Xj 0 1081
(j < m), (j 6 A m ), (jtAm).
Then /m is clearly continuous and the sequence (/ m ) m approximates the identity on Q. We claim that fm[Q] H <7°°[jBn] = 0 for every m, which is sufficient by Lemma 5.1.3(1). Striving for a contradiction, suppose that there are y G Q and x G .Bn such that
fm(y) = gac(x). Put
C = { / c G N : | z f c | <2}. Then C G .An and so C£{TuAm:TCSm}. There consequently exists j > m such that j 6 C£±Am. If j e C \ Am then fm(y}j = 1081. In addition, |xj| < 2 since j 6 C and so |^°°(a:)j| < 1. This is a contradiction. If j e Am\C then fm(y)j = 0. Moreover, since j £ C we have |xj| > 3 and so g°°(x)j > 1. This also leads to a contradiction. <0> ClaimS. cyC\J™=ig°°[Bn]. Proof. Pick an arbitrary xec^CW30. Define the point y G TN by
{
Xi (zi 6 J), Xi + 2 (ari G (l,oo), Xi - 2 (xi e (-oo,-l).
Then g°°(y) = x. Since x G eg-, /T = {i G N : |xi| < 1} G y.
Since /C is clearly equal to the set
{ z G N : \Vi\ < 2 } we conclude that y G U^=i ^-
"v"
This completes the proof of the lemma.
D
6.12. TOPOLOGICAL EQUIVALENCE OF CERTAIN FUNCTION SPACES
439
Lemma 6.12.7. Let 3* be a free filter on N such that cy is analytic. Then cy is both reflexively universal and J '&$ -universal in Q. Proof. We first prove that cy is reflexively universal. To this end, first observe that if 3 — $ then cy = CQ and so we are done by Theorem 5.5.15. We may therefore assume that y ^ #, i.e., that there is an element F G ff such that N \ F is infinite. Let $ = 3 \ F. Observe that 9 C y and that cy is canonically homeomorphic to eg x s. We aim at applying Theorem 5.5.20. For that we need to show that there is a function tjj : MF —> M.F such that
^-l[cs]nRF =cs. But this is easily done. If a: R —>• [—1081, 1081] is a homeomorphism then the function T/>: RF -> [-1081, 1081]F C RF defined by has the desired property. We next prove that cy is S^-universal. First observe that y is a meager subset of J"(N) by Lemma 6.12.2 and Corollary 6.12.3. Find a partition {N, : i G N}
of N such as in Corollary 6.5.4(3). As on Page 354, for every n G N, put Qn = {x G lNt : N < 2~ n for all but finitely many i}. Let yx = y \ N; for every i. Then every 3^ is meager. It is easily seen that oo
Co C c^ C JJ c?i . 1=1
For each i G N, let Q{ = [-2- i+1 ,2- i+1 ] N '. Then n~i Qi ^ Q. As in the proof of Proposition 6.12.5 it follows that cyi C\Qi is contained in a crZ-set Di ofQi. Let A = H^Li An where each An is a a-compact subset of Q while moreover An+i C An for every n G N. Define for every i, C, = {x G Qt : (Vj G Ni)(|xj| < 2 f c ~ j for some jfc)}. As in the proof of Lemma 5.5.12, C^ is an 3~CT-absorber in Qi. By Exercise 5.5.4, there is an imbedding fa : Q —> Qi such that f ~ l [ C i ] = Ai and Define / = (/;); : Q -> Hi^i Qi ^ Q- Then / is an imbedding and we claim that f ~ l [ c y ] = A. First, let x G A and n G N. If i > n, then fi(x] G Qi, so all components of fi(x) are in [— 2~ n ,2~ n ]. If i < n, then since x £ Ai we have /i(x) G C^. In this case only finitely many components of fi(x) are outside the interval [— 2~ n ,2~ n ]. We conclude that only finitely many components of f ( x ) are outside [— 2~ n ,2~ n ]. This implies that f ( x ) G Qn-
440
6. FUNCTION SPACES
Since n was arbitrary we get from Lemma 5.5.12 that f ( x ) G CQ C cy. Second, let f ( x ) G cy and fix n 6 N. Then fn(x) G cyn fl Qn C £)n, hence x € A n . We conclude that x £ A. D Function spaces of low Borel complexity: the metrizable case. We will now make our first steps towards proving that all F^-function spaces of countable spaces are homeomorphic. Some of the methods used here were used before, for example in §6.6. Since we are dealing with Q instead of s, things are slightly more complicated. For completeness sake we will include the routine verifications. We define the continuous function ^ : 1R —> RN by \I/(r)(n) = sign(r) • min{ r|,n}. Note that ^(r)(n) is finite for all n ^ oo and (*)
r - lim *(r)(n) = *(r)(oo). n—>-oo
Hence *[R] C M and *[R] C Cp(Ny}. For / 6 Q = MN let / be the extension of / over N that assigns 0 to oo. Define a function (p : tN x R -» R* by
Lemma 6.12.8. (p is a homeomorphism and (p[cy x E] = Cp(Ng-]. Proof. It is clear that (p is well-defined. To prove that it is surjective, take an arbitrary g G M N . Put r = g(oo) and f = g — \J>(r). Observe that by (*),
This means that if h = f \ N then h = f and so ¥>(/», r ) = / , as required. It is clear that ? is continuous. The above considerations easily allow to display the inverse of <£>, as follows: (**)
tN,/i(oo)).
This means that ? is a homeomorphism. Take arbitrary x 6 cy and r G K. Then ip(x,r) = x + *(r). By the above, both x and ^(r) are continuous functions on F%. So the same is true for their sum. Conversely, let / G Cp(]%) and let r = /(oo).
6.12. TOPOLOGICAL EQUIVALENCE OF CERTAIN FUNCTION SPACES
441
Then r € M and by (**), ^"1(/) = ( [ / - * ( r ) ] f N , r ) . Since /-*(r)(oo)=0, we have by (*) that / — \&(r) G cy, which is as required.
D
We will now investigate which CP(X) are contained in a crZ-set of Q. Before we can say more about this, we need to derive a simple result first. If y is closed under supersets then a subcollection $ of 3 is said to be a base for 3 provided that for every F G J there exists G G 9 such that G C F. Lemma 6.12.9. Let J be a free filter on N with base 15. If 15 is an analytic subset of?(N) then so is J. Proof. Define a function ip: T(M) x T(N) -> T(N) by (p(A,B] = AUB. We first claim that ( is continuous. Assume that
for certain E, F G 3>(N) such that F is finite. Put
respectively. We claim that OCxLCip-1[V(E,F)]J which is obviously as required. To this end, pick arbitrary A' G 3C and B' G £. Then and so
(A' u B') n F = (A u B) n F = E n F. This implies that ip((A',B'))
G V(.E,F), as claimed.
It is clear that CB x P(N) is analytic, being the product of two analytic spaces. Since (p[B x 05(N)] = 3 the continuity of ? shows that 5" is analytic as well.
D
This yields the following interesting result. Proposition 6.12.10. Let X be a countable nondiscrete space such that its function space CP(X) is analytic. Then CP(X) is contained in a aZ-set ofQ.
442
6. FUNCTION SPACES
Proof. Let x e X be a nonisolated point of X. Identify X \ {x} with N and x with oo; put 3 ~ = { C / C N : t / U {x} is a neighborhood of x}. Then 9" is a free filter on N and CP(X) C CP(N?). Let £ = {B C N : B U {x} is a clopen neighborhood of x}. Clearly, *¥> is a base for 3 since ^ is zero-dimensional being countable (cf. Exercise 1.5.1). The subspace !B of 05(N) and the subspace £ = {/ € CP(X) : (/[X] C {0,1}) A (f(x) = 1)} of CP(X] are canonically homeomorphic (identify each B G *B with its characteristic function x#), and moreover, £ is closed in C P (X) (cf. the proof of Lemma 6.12.2). So !B is analytic since £ is. But then 3 is analytic as well by Lemma 6.12.9. By Proposition 6.12.5, cy is contained in a crZ-set of Q. So by Lemma 6.12.8 the same is true for CP(I%) and hence in turn for its subspace CP(X] (Lemma 5.1.2). D Lemma 6.12.11. Let X be a countable space. If X contains a clopen subset Y containing exactly one accumulation point and CP(X] is an Fff$-subset ofRx then CP(X] is an 3ff&-absorber in Rx . Proof. By Proposition 6.12.10, CP(X) is contained in a crZ-set of Q. There is a free filter 3 on N such that Ngr is homeomorphic to Y. It easily follows that cy is an F^s-subset of Q, hence by Lemmas 6.12.7 and 5.5.1, cy is strongly ^F^-universal. Since there clearly is a deformation of E through M, by Exercise 5.5.2 and Lemma 6.12.8 we obtain that Cp(Ny) is strongly Ja5universal. Since Y is clopen in X, CP(X) is canonically homeomorphic to the product Cp(Y)xCp(X\Y) (Theorem 6.2.8). Since CP(X \ Y} is a dense linear subspace of E.X\Y, there is a deformation of M x \ y through CP(X \ Y} by Corollaries 4.2.15 and 4.1.8. By an application of Exercise 5.5.2 we consequently obtain that CP(X) is strongly 3>5-universal in Q. D Proposition 6.12.12. Let X be a countable nondiscrete metrizable space. Then CP(X] is an 3v<5-absorber in Rx . Proof. By Theorem 6.3.10 it follows that CP(X) is an 3^-subset of Rx. By Corollary A.13.12 and Proposition 6.12.10, CP(X) is contained in a crZ-set of Rx (there is a simple direct proof of this fact, but Proposition 6.12.10 is needed in full generality later so we might as well use it here). It remains to show that CP(X] is strongly ^o-
6.12. TOPOLOGICAL EQUIVALENCE OF CERTAIN FUNCTION SPACES
443
Consider the space Ng. Topologically this is a simple convergent sequence with its limit and hence it is a countable nondiscrete metrizable space. So Cp(Ng) is an S^-absorber by Lemma 6.12.11. We now reduce the problem for CP(X) to C p (f%), as follows. Let g be an admissible metric on X and let A be a nontrivial convergent sequence in X. Observe that such a sequence exists since X is nondiscrete. By the above it follows that CP(A) is strongly SVa-universal in M^ 4 . Claim 1. There is a retraction r: X —> A. Proof. Let a be the unique nonisolated point of A. Since X is metrizable (hence first countable) and zero-dimensional (cf. Exercise 1.5.1) there is a decreasing clopen neighborhood base (Un)n of a in X such that Un \ Un+i contains exactly one point from A \ {a} for every n, say a(n). Now define the retraction r : X -> A by the following formula: fa(n) r(x) = < la
(x e Un \ t/n+i), (x = a).
An easy check shows that r is as required.
0
The formula V(g)(x) = sign (g(r(x))) min||g(r(a;))|,
rrx\\}
defines a continuous extender that extends every g £ M"4 to an element of Rx . The map \I> has the following properties: *(#) \ A = g, ^(g} \ X \ A has its values in R and ij>[Cp(A)] C CP(X). If / e M X \ A then / is the extension of / over X with zeros. As in the proof of Lemma 6.12.8 it is easily seen that is a homeomorphism from RX\A x MA onto Rx . Let Cp(X,A) = {f\X\A:fe
CP(X) and / \ A = 0}
and note that Since CP(X, A) is a dense linear subspace of R^^ , there is as in the proof of Lemma 6.12.11 a deformation of R^"4 through CP(X,A). So Exercise 5.5.2 implies that CP(X) is strongly J^-universal in Rx . D Function spaces of low Borel complexity: the general case. We now come to the main result in this section. Proposition 6.12.13. Let X be a countable nondiscrete space. Then X has a clopen subspace Y with exactly one accumulation point or there is a partition of X into infinitely many pairwise disjoint nondiscrete clopen sets.
444
6. FUNCTION SPACES
Proof. Suppose that X does not have a clopen subspace with exactly one accumulation point. Enumerate X as {xn : n e N}. Put XQ — 0. Assume that we constructed pairwise disjoint nondiscrete clopen sets X\,..., Xn of X such that Yn = X \ \J™=0 Xi is also nondiscrete and n
{xi,...,xn} c (j Xi i=0
if n > 1. By assumption, Yn contains at least two accumulation points. By the zero-dimensionality of X (cf. Exercise 1.5.1), we can split Yn into two nondiscrete clopen sets, say A and B. If n
xn+l e A u |J x, i=0
then put Xn+i = A. If not then xn+i e B and we put Xn+i = B.
D
Lemma 6.12.14. Let X be a countable nondiscrete space. Then CP(X) contains an yab-absorber for M.x . Proof. Since X is zero-dimensional (cf. Exercise 1.5.1), for every pair of distinct points x,y £ X there is a continuous function /: X —> {0,1} such that f(x] = 0 and f(y) — 1. Picking such a function for each pair of distinct points of X yields a continuous injective function /: X —> {0,1}°°. Let Y be the image of X under /; clearly, Y is a countable metrizable space. In addition, Y is not discrete since / is injective. By Proposition 6.12.12 we have that CP(Y] is an J^-absorber in W. Let T be the metrizable topology on X that it inherits from Y. So a set U in X is r-open if and only if f[U] is open in Y. The set X endowed with the topology T will be denoted by XT. Then clearly XT « Y and hence CP(XT] is an Sv^-absorber in Rx. Observe that T is weaker than the original topology on X since / is continuous. This implies that every function g € Cp(Xr) is also continuous with respect to the original topology on X. From this it follows that CP(XT) C CP(X). Since Cp(Xr) is an ^-absorber in E x , we are done. D We now come to the main result in this section. Theorem 6.12.15. Let X be a countable nondiscrete space such that CP(X} is an Fa$-subset of Q. Then CP(X) is an yas-absorber in Q. Proof. By Proposition 6.12.13 and Lemma 6.12.11 we may assume without loss of generality that X admits a clopen partition {Xn : n G N} such that every Xn is not discrete. By Lemma 6.12.14 there is for every n an Jo-5-absorber Bn in Ex™ such that Bn C Cp(Xn). So by Exercise 5.5.3
6.13. EXAMPLES
445
we may pick an ^-absorber B'n in RXn such that B'n C Bn. Finally, Proposition 6.12.10 gives us that Cp(Xn} is contained in a aZ-set Cn C RXn . We now have by Theorem 6.2.8 the following inclusions: oo
oo
oo
[J B'n C JJ Cp(Xn) = CP(X) C H Cn. n=\
n=l
n=l
By Exercise 5.5.5 we consequently get that CP(X) is strongly .F^-universal, which finishes the proof. D Remark 6.12.16. In view of Theorem 6.12.15 the question naturally arises whether for a countable space X the Borel type of CP(X) determines the topological type of CP(X}. But this is not true, as was shown by CAUTY [88]. As we saw in §6.3, nontrivial function spaces are at least of 'Fff§ Borel complexity'. Theorem 6.12.15 proves that at the first nontrivial stage all function spaces are homeomorphic. What happens at more complex situations is a mystery, although Cauty's result just quoted shows that a result as elegant as Theorem 6.12.15 is not possible. 6.13.
Examples
In this section we collect several examples dealing with function spaces, some with and some without details. Several of the examples are already dealt with earlier in this chapter. El. Non-homeomorphic ^.-equivalent spaces. See Example 6.2.4. This shows that ^-equivalence is not the same as topological equivalence. E2. Countable compact (necessarily metrizable) spaces X and Y which are not i-equivalent. See Corollary 6.2.6. E3. Countable nondiscrete spaces Xn for every n such that
This shows that Theorem 6.2.8 is in general not true for spaces of bounded functions, even if one considers homeomorphisms instead of linear homeomorphisms. See MARCISZEWSKI and VAN MILL [272] for details. There are some special cases for which the version of Theorem 6.2.8 for homeomorphisms does hold for spaces of bounded functions. See e.g., BANAKH and CAUTY [40]. E4. C(uji) and C(o;i -I- 1) are isomorphic as rings, yet coi and u)\ + I are not homeomorphic.
446
6. FUNCTION SPACES
See Page 371 to see why this is of interest. E5. Countable spaces Xa for which Cp(Xa) is of arbitrary multiplicative Borel complexity. See CALBRIX [83, 84] and LUTZER, VAN MILL and POL [265] for details. This is of interest since on Page 375 it was shown that CP(X) cannot be a G8ffsubset of E.x, unless X is discrete. In addition, CAUTY, DOBROWOLSKI and MARCISZEWSKI [91] used this to prove the interesting result that if X is countable and CP(X) is a Borel subset of E.x then it is of multiplicative Borel class. So the results of Calbrix and Lutzer et all show that any possible Borel complexity for CP(X) can be attained. E6. A countable regular, nonmetrizable space X for which CP(X] is an FffS-subset ofRx. Let Xk = {(V n , i/nfc) : 1 < n < 00} U {(0,0)} for k > I and let X = \J{Xk :l
6.13. EXAMPLES
Ell.
447
An infinite compact space X such that CP(X] is not linearly homeomorphic to CP(X) x M.
Such an example was constructed by MARCISZEWSKI [268]. It is known that if X contains a nontrivial convergent sequence then CP(X) ~ CP(X) x R (Corollary 6.6.9). So Marciszewski's example does not contain nontrivial convergent sequences. The first example that comes to mind as a possible candidate is /3uj. But this space is homeomorphic to the topological sum of itself and a one-point space and so Cp((3u}) ~ Cp(j3uj} x E. The second space that comes to mind as a possible example is a;* = j3uj \ u. But a;* contains a copy of J3uj which is a retract. So CP(UJ*) ~ Cp(u*} x EL E12. A nonmetrizable compact space K such that CP(K] is not homeomorphic to CP(K) x CP(K).
See MARCISZEWSKI [266] and GUL'KO [181]. In view of Corollary 6.6.10 and Examples Ell and E13, the question naturally arises whether there is an infinite metrizable space X for which the spaces CP(X) and CP(X) x CP(X] are not (linearly) homeomorphic. Observe that this question would be trivial if every infinite-dimensional linear space L would have the property that L is (linearly) homeomorphic to its own square L x L. But this is not true as is clear from Example E13. It is not even true if L is metrizable, as was first shown by POL [342]. By a refinement of Pol's method, VAN MILL [296] proved that there is an infinitedimensional normed linear space L such that L is not homeomorphic to L x EL The question was answered by POL [344] who presented three examples, one of which is COOK'S continuum from [103] and another one is a very rigid subspace of M, basically due to KuRATOWSKi [237]. We will present the second example below. E13. An infinite subspace X o f ~ K such that CP(X] is not linearly homeomorphic to CP(X] x CP(X}. We will first show that there is a 'very rigid' subspace of EL Example 6.13.1. There exists a subset B C M having the following properties: (1) No nonempty open subset of B is meager in EL (2) If V C B is open and nonempty then each continuous function /: V^B
is either the identity or constant on a nonempty open subset W C V. Proof. Let "X. be the family of Cantor subsets of EL Put C = {(K, /) : K e OC, / G C(K)J[K]
is uncountable, f[K] n K = 0}.
448
6. FUNCTION SPACES
Since |3C| = c (Exercise A.2.16) and \C(K}\ < c for every K 6 X (Corollary 1.3.4), we can enumerate C as {(Ka,fa)
:a< c}.
By transfinite induction on a < c, we will pick points xa 6 Ka such that Let XQ be an arbitrary element of KQ, and assume that for some a < c we picked x/3 for every f3 < a. Now observe that fa[Ka] is uncountable, and consequently has cardinality c (Corollary 1.5.13). So there is a set B C fa[Ka] such that \B\ = c while moreover There exists a set A C Ka such that fa \ A is one-to-one and fa[A] Some A' C A has \A'\ = c while moreover
= B.
Let xa be any point of A'. It is easily seen that xa is as required. Put B = {xa : a < c}. We claim that B is the 'very rigid' subspace of E we are after. Observe that B has the property that for every a < c the point fa(xa] does not belong to B. Claim 1. B and R \ B intersect every Cantor set in E. Proof. For a Cantor set K C E, let /: K —> E be an imbedding such that (For this, all one needs to observe is that K is bounded.) The pair (K, /) belongs to C and hence there exists a < c such that By construction, xa G Ka. The pair (f[K], /~ x ) also belongs to C and so the same argument shows that K n (R \ B) / 0. 0 Claim 2. If U C 5 is nonempty and open (in B} then it is not meager in E. Proof. Let [/' C E be open such that U' C\B = U. Suppose that U is meager in E. Then there is a countable collection £ of nowhere dense closed subsets of E such that U C (J £. By the Baire Category Theorem, G = U'\\j£ is nonempty. In fact, the Baire Category Theorem implies that G is uncountable. So G is an uncountable G$ -subset of E and consequently contains a Cantor set K by Theorems A. 6. 3 and 1.5.12. So G intersects B by Claim 1. But this is obviously a contradiction since G n B — 0. 0
6.13. EXAMPLES
449
Claim 3. If U C B is nonempty and open (in B) and /: U -> B is continuous, then either / is the identity on C7, or is constant on some nonempty open subset V C U. Proof. Suppose that / is not the identity. Then there exists a nonempty open subset W C U such that the closure of W in B is contained in U while moreover W D f[W] = 0 (the closure is taken in R). We claim that f[W] is countable. Striving for a contradiction, suppose that it is uncountable. By Corollary A. 8. 4 there is a G<$-subset 5 C W such that W C S while moreover / \W can be extended to a continuous function /: 5 -> f[W]. Since /[S] contains /[W7], it is uncountable. By Theorem 1.5.12 there consequently exists a Cantor set K C 5 such that / \ K is one-to-one. There exists a < c such that As a consequence, xa € K n J3, which is contained in the closure of Ty in which is contained in U. But by construction, fa(xa} $ B, and so, since f\U = f, it also follows that which is a contradiction. Now for every q in the countable set f[W] let Aq — {x G W : f ( x ) = q}. Then A = {Aq : q 6 /[W7]} is a countable cover of W by (relative) closed sets. By Claim 2, one of those sets must have nonempty interior in B, so we are done. <0> This completes the construction of B.
D
We now proceed with the proof that CP(B] is not linearly homeomorphic to its own square CP(B) x CP(B). For this, we need to do some preparatory work first. Lemma 6.13.2. Let K:L be separable metrizable spaces. Assume that no nonempty open subset ofK is meager in K. Let S : K ->• 3oo(£) \ {0} be LSC. Then there exist a nonempty open set G C K and finitely many pairwise disjoint open sets in L, say Hi,...,Hn, and for every i < n continuous functions Si : G —> Hi such that for each t 6 G. Proof. For every n e N, the set Fn = {t 6 G : \S(t)\ = n} is an FCT-subset of G. So one of them has nonempty interior, say Fn. So we can find a nonempty open set G C Fn, small enough that there are n pairwise disjoint open sets Hi C L such that each S(t) with t & G intersects every Hi in a
450
6. FUNCTION SPACES
singleton. The functions s; are defined by letting Si(t) be the unique point in Hi H S(t) for teG. D Theorem 6.13.3. Let B be the space constructed in Example 6.13.1. Then its function space CP(B) is not linearly homeomorphic to CP(B] x CP(B}. Proof. First observe that
CP(B®B] ~CP(B]
xCp(B]
(Theorem 6.2.8). So it suffices to show that CP(B) is not linearly homeomorphic to CP(B ®B). Let B ® B = B' U B", where B' and B" are disjoint open copies of B. For each x G B and A C B, we denote by x' , x", A' and A" the corresponding copies of the point x and the set A in B' and 5", respectively. Striving for a contradiction, assume that there is a linear homeomorphism We will apply Lemma 6.13.2 to K — B' and the LSC function F:B'^?00(B)\{
supp^(y') = {y,a 2 ,...,o n } for y G V.
We now perform a similar construction with K — V" - the copy of the set V in B" . We end up with a nonempty open set W" C V" and points bi £ B for 2
supply") = { y , & 2 , . . . , & T O } for y G W C V.
6.13. EXAMPLES
Let J = {02, . . . , a n , &2; • • • 5 km}, having the property that (3)
an
451
d take an arbitrary element c 6 W \ J
c',c" ^ (supp^-i^) : x G J}.
(Again we use here that supports of points are finite.) We define a linear function T: E —> R2 in the following way. Given r 6 M, choose any / 6 CP(B] such that / f J = 0 and /(c) = r. Let T(r] = ((^(/)(c'),^(/)(c")). We claim that T(r] does not depend on the choice of /, and that the correspondence
r ^ T(r] is linear. To see that this is true, let g G C(X) also be such that g \ J = 0 and g(c] = r. Then
(see formula (**) on Page 406) and, similarly, By making a similar calculation for c", we conclude that T is well-defined. Now that we know this, linearity is clear. We will now show that T is surjective, which by linearity obviously yields a contradiction. Pick arbitrary r',r" 6 M. There exists by (3) a function g£Cp(B'\JB") such that 9 \ (J{supp^-i (x) : x e J} = 0, while moreover <7(c')=r', <7(c")=r". Let / = (p~1 (g) and r = /(c). Then again by using formula (**) on Page 406, we get that if x G J then =
E
A(z, *)(*) = 0.
zGsupp^-x (x)
From this we conclude that T(r) - (^(/)( C '),^(/)( C ")) = (9(<S),9(
D
E14. A countable space for which the Dugundji Extension Theorem fails. Countable spaces have separable metrizable function spaces and so in view of the result of GEBA and SEMADENI [172] mentioned in Example 6.6.4, the question naturally arises (see ARHANGEL'SKII [26, Problem 58]) whether some form of the Dugundji Extension Theorem holds for that class of spaces. Observe that there are always linear extenders by Lemma 6.6.1. But, as
452
6. FUNCTION SPACES
we will show, there are not always extenders which are continuous, even for countable spaces. One can interpret this from the point of view of selection theory, as follows. If X is countable then CP(X) is a separable metrizable locally convex vector space. In addition, if A C X is closed then the restriction operator p:
CP(X)-+CP(A)
defined by p(f) — f \ A is bounded, linear, onto (by the Tietze Extension Theorem) and open. So by our claim there does not always exist a continuous selection for the continuous set- valued function Interestingly, a result of MICHAEL [288] implies that for every countable subspace C C CP(A) there is a continuous selection for the set- valued function We will now present the construction of the desired example. Let D be an almost disjoint family of infinite subsets of u, i.e., \D\ = uj for all D 6 D and \D D D'\ < u for all distinct D, D' 6 T>. There are 'large' almost disjoint families of subsets of u. For example, if for every £ G IP, S(t) is a sequence of rational numbers converging to £, then the family {S(t} : t € F} is an almost disjoint family of c subsets of the countable set Q. Let 9" be a collection of functions from uj to uj and let (p: T) —>• y be a surjection. Finally, let q be any point not in uo x (uj + 1). Topologize as follows. The set uj x (oj + 1) is an open subspace of A (?), carrying the usual product topology. For finite £ C D and n G uj define y ( £ ) = { g } u [ ( o ; \ U £ ) x ( u ; + l ) ] u \J
t/(£,n) = F(£) \ ({0, 1, . . . ,n - 1} x (u + 1)), respectively. (We identify here the functions ip(E) \ E with their respective graphs.) The C/(£,n)'s form a neighborhood base for q. It is easy to see that A(cp) is regular and T\. Lemma 6.13.4. Let Del). Then ip(D) \D C uj x w C uj x (w + 1) converges to q in A (?). Proof. Let f/(£,n) be a basic open neighborhood of q. If D G £ then
6.13. EXAMPLES
453
and so all but finitely many elements of (p(D] \ D are contained in t/(£,n). Suppose therefore that D $. £. Then D n |J£ is finite since D is pairwise disjoint and £ is finite. So all but finitely elements of
454
6. FUNCTION SPACES
is pair wise disjoint. Proof. Enumerate A as {Aa : a < c} such that each A £ A is listed c times. By transfinite induction on a < c we will pick xa G Aa \ {xp : 13 < a}.
This is a triviality. Let XQ be any point of AQ. If x@ has been denned for every /3 < a, where a < c, then proceed as follows. The set A'a = Aa \ {Xf3 : J3 < a}
is nonempty since \Aa\ — c and \a\ < c. So we can let xa be any point of A'a. For every A 6 A let B(A) = {x7 : A7 = A}. Then the collection {B(A) : A G A} is clearly as required. D Observe that the cardinality of T>* is equal to c, that the family of all Cantor subsets of D* has size c (Exercise A.2.16) and that every Cantor set has size c. So by Lemma 6.13.5 there is a disjoint family "K of subsets of D* such that every Cantor subset of D* contains an element of ^K; moreover, each H G !H has size c. For every H e 'K let (/?#: H —> uj^ be a surjection. Now define
by the rule (p(x) —
is a basic clopen subset of D* and consequently intersects 9*- For convenience, put
9» = -D»n9*
(
We will now construct a function /: S -> a; having some special properties. To this end, take an arbitrary a G S. The value /(cr) is determined as follows. Take an arbitrary G^ 6 9*(o")- Since e(G^)((<7,a;)) = 1, there exists n 6 a; such that e(G*)(0,n)) > 1/2- put /(cr) = n. For cr G S put (*)
U((7) - { ^ E C P (A) : ii((t7,o;)) > ya,
u((cr,/(a))) > 1/2}-
6.13. EXAMPLES
455
Then U(
e[9»]nl%)^0, or, equivalently, g*(a)ne-1[U((r)]^0. For every k E uj define (**)
Ufc - \J{U(a) : length a > k}.
By the above and the continuity of e on 9*, for every A; we have that
e-^Ufcjng* is dense and open in 9*- We conclude that
is a dense G^-subset of 9*, and hence of D*, and it therefore contains a Cantor set (Theorem 1.5.12). So by construction, there exists
It follows that e(D^) G U^ for all k, and hence, by (*) and (**), e(D-}((a^}} > 1/2, e(D:)((aJ(a)))
>%
for infinitely many a. However, e(£>:)(((7,a;))= J D:((<7,a;))> 1/2 implies that cr E D, so in fact, (t) «ra ((*,/("))) > V 2 for infinitely many a £ D. But e(D^)(q) = 0 and / = vK^) and so f \ D converges to g. By continuity of e(D^) we must consequently have that (t) *(£:) ((*,/(*)))< i/2 for all but finitely many
456
6. FUNCTION SPACES
a first category set (in £). As a consequence, e is continuous on a dense Ggsubset of £. Let us now return to the space A and its closed subset A. Since D* is a Cantor set, by the above remarks and Theorem 6.13.6 it follows that C P (A) is not Borel. Also, there does not even exist a measurable extender from CP(A) to <7 P (A). Interestingly, CP(A) is Borel since the space A is similar to one of the examples considered in LUTZER, VAN MILL and POL [265]. E15. A compact space X and a noncompact space Y such that CP(X) and CP(Y) are homeomorphic. Let X = I and Y = E. Then CP(X) w CP(Y), as was shown by GUL'KO and KHMYLEVA [183]. See also Example 6.9.3. E16. An infinite-dimensional compact space X such that CP(X) is not homeomorphic to CP(Y) for any finite dimensional compact metrizable space Y. By CAUTY [90], X = Q is as required. See also Corollary 6.11.4. E17. Scattered compact I*-equivalent spaces X and Y such that K,(X) — 2 and K,(Y] = n + 1. See Example 6.10.5. E18. l-equivalent countable, locally compact (hence metrizable) spaces which are not I*-equivalent. See Example 6.10.10. E19. Countable t*-equivalent spaces which are not t-equivalent. See MARCISZEWSKI and VAN MILL [272]. E20. Countable spaces X and Y such that CP(X] and CP(Y) are nonhomeomorphic absolutely Borel sets of the same Borel type. See CAUTY [88].
APPENDIX A
Preliminaries In this appendix we will collect for easy reference and for completeness sake some facts that are important throughout this book. It is certainly not our intention that the results presented here will be used as a text for an introductory course in topology. For the more experienced reader there is no need to read this chapter in full detail. She or he should read §A.l only to get our notation straight. All spaces under discussion in Chapters 1 through 5 of the present book are separable and metrizable exclusively. A.I. Prerequisites and notation
We assume that the reader understands the material of basic courses in general topology, calculus and linear algebra. Concepts such as topological space, metric space, metrizable space, complete metric space, completely metrizable space, compact space, product space, continuous function, etc., should be familiar. For all undefined notions see any textbook on General Topology (e.g., DUGUNDJI [142], ENGELKING [153] and NAGATA [324]). Notation. We let I, JJ, R, N and Z denote the interval [0, 1], the interval [—1,1] and the sets of real and natural numbers and integers, respectively. In addition, Q denotes the subset of E consisting of all rational numbers, and IP denotes E \ Q. That is, IP is the subset of E consisting of all irrational numbers. Set theoretic notions. If A and B are sets then AAB denotes their symmetric difference, i.e., =
(A\B)(J(B\A).
If {Xl : i e /} is a family of sets then
is their (Cartesian) product. For every j e /, the projection Hie/ ^ ~^ Xj will be denoted by -KJ . If E is an arbitrary subset of / then KE denotes the projection from Hie/^i onto its subproduct }\^EXi. 457
458
A. PRELIMINARIES
Let {xi : i e /} be an indexing of a set X. We say that X is faithfully indexed by / provided that the function / —> X defined by
i H> Xi is a bijection. If U is a family of sets and X is a set then U \ X denotes the collection
{UHX :U ell}. If /: X —>• Y is a function and A C X then / \ A denotes the restriction of / to A. For t 6 R, the symbol / = t means that / is the constant function with value t. The identity function from X to X shall be denoted by 1 or by \xA function /: X —>• Y is sometimes identified with its graph T = {(xj(x))
: x e X} C A" x Y.
By this identification we can use the standard set theoretic terminology for functions. For example, if A,B are subsets of X and f : A -* Y, g: B ^ Y are functions such that f ( x ) = g(x) for every x £ AC\ B then fU9 is the function from A U Z? to Y defined as follows:
(zeA), (x € B).
Etc. If X and Y are sets and /: X —> Y then a fiber of / is a set of the form f ~ l [ { y } ] , where y e /[X]. If j/ G /[X] then its fiber f ~ l [ { y } ] will for simplicity be denoted by f ~ l ( y ) . If X is a set then y(X) denotes the power set of X. Each subset A of a set X can be identified with its so-called characteristic function \A '• X —> {0,1} defined by XA(X) — 1 if and only if x G A. This means that we can identify A C X with XA G {0,1}X- The function A (->• ^A is clearly a bijection from J>(^C) onto {0,1} X . It will be convenient to introduce the following notation. If U and V are covers of a set X then U < V means that U is a refinement of V (i.e., for every U e U there exists V G V such that U C V). A nonempty subcollection A of
(1) 003-,
A.I. PREREQUISITES AND NOTATION
459
(2) if F, G e 5 then F n G e 9", (3) if F E 3- and F C G C X then G € 3". A filter 3* is called /ree if f°| 3" = 0. If a filter is not free then it is called fixed. An ultrafilter is a filter not properly contained in another filter. The Kuratowski-Zorn Lemma easily implies that every filter is contained in some ultrafilter. It is easy to see that the intersection of a fixed ultrafilter is a singleton. So a fixed ultrafilter on a set X has the form {A C X : x e A}
for some x G X. Notice that each filter has the finite intersection property. If 3" C 7(X) has the finite intersection property then the collection
is easily seen to be a filter. It is called the filter generated by 3". The first infinite cardinal is uj and c denotes 2 W , i.e., the cardinality of the continuum. We usually identify uj and the set of positive integers. By uj + I we mean the sequence of positive integers converging to the point uj. So uj + 1 is homeomorphic to the subspace {Vn:neN}U{0} of M. That is, uo + 1 is nothing but a prototype of a convergent sequence with its limit. If X is a set then \X\ denotes its cardinality. Metric spaces. Let (X, g) be a metric space. If A C X then diam(/i) denotes the diameter of A. If U is a collection of subsets of X then the mesh of U is the number mesh(U) = sup{diam(t7) : U e U}. We allow mesh(U) to be equal to oo. In addition, g is called bounded if the diameter of X is finite For A C X and e > 0 we let
B(A,e) = {ytX : g(A,y) < e} and
denote the open and closed balls about A with radius e, respectively. If A is a singleton, say {x}, then we write B(x,e) instead of B({x},e) and D(x,e) instead of D({x},e). If it is unclear to which metric g we are referring, we occasionally use the notation Be(A,e} and BQ(x,e}. Similarly for D. We sometimes denote B(A,e) by BE(A) and D(A,e) by De(A).
460
A. PRELIMINARIES
Let (X, g) and (Y, d] be metric spaces. A surjective function /: X —>• Y is an isometry provided that for all x,y e X.
Metrizable spaces. Let X be a metrizable space. We say that a metric Q on X is admissible if it generates the topology on X. Each admissible metric can be replaced by one which in addition is bounded, see Exercise A. 1.6. Topological notions. Let X be a topological space with subset A. Then A and Int(A) denote its closure and interior, respectively. In addition, Fr(A) is the boundary of A. A subset of a space X is called clopen if it is both closed and open. If U is a collection of subsets of X then U denotes the collection {U : U € U}.
Let A C X and let U be an open cover of X. The star of A with respect to U is the set St(A U) = \J{U G U : U n A ^ 0}. The cover (St(C7,U) : U e U} is denoted by St(U). We say that an open cover V of X is a star-refinement of U if St(V) < U, i.e., if for every V G V there exists U G U such that St(V,V) C [7. An important characterization of compactness is the well-known fact that a space X is compact if and only if every collection of closed subsets of X with the finite intersection property has nonempty intersection. This will be used without explicit reference in the remaining part of this book. Let f:X-*Y. We say that / is closed provided that f[A] is closed for every closed set A C X. Similarly for open. If X and Y are spaces then / : X —> Y is an imbedding if the function
is a homeomorphism. If /: X —t Y is an imbedding and f[X] is a closed subset of Y then / is called a closed imbedding. Similarly for open. Moreover, if / : X -> Y is an imbedding and f[X] is dense in Y then / is called a dense imbedding. The symbol 1X rz F' means that X and Y are homeomorphic spaces. If X and Y are spaces then C(X, Y) denotes the collection of all continuous functions from X to Y.
A.I. PREREQUISITES AND NOTATION
461
We assume that all topological spaces under discussion are nonempty. Topological products. Let X = Hi^i %-i, where Xi is a nonempty topological space for every i, and X is endowed with the Tychonoff product topology. It is possible to identify each Xl in a natural way with a subspace of X, as follows. Pick an arbitrary element Xi £ Xi for every i. The subspace {Xi} X • • • X {Xi-i} X Xl X {Xi+l} X • • •
is clearly homeomorphic to Xi. Observe that the composition X, 4 X ^ X^
where i denotes the 'inclusion', is the identity on Xi. If [A, B} is a partition of / then X can be naturally be identified with the product U Xi x JJ Xi. i£A
i£B
We freely will make such identifications without warning. So we will permute or reorganize the factors of a product in any way that is convenient to us. A countably infinite product of copies of the same space X shall sometimes for convenience be denoted by X00. The i-th coordinate of a point x in a product space shall be denoted by Xi. The Euclidean spaces W1. The standard metric on W1 is the Euclidean metric which is defined by
Let Bn and S n 1 abbreviate the unit ball and unit sphere in W1, respectively. Two points x, y € §n will be called antipodal provided that x = —y. The function Sn —> S n defined by x t-> — x is called the antipodal mapping on S n . The Hilbert cube. The Hilbert cube, abbreviated Q, is the product JJ°° endowed with the Tychonoff product topology. Observe that Q is compact, being a product of compact spaces. A natural metric for Q that generates its topology is oo
Q(x,y) = ^2~l\Xi -yx\ i=l
(Exercise A. 1.10). Unless otherwise stated, Q will always be endowed with this metric. The diameter of Q is 2. The point in Q all coordinates of which are 0 will be denoted by 0.
462
A. PRELIMINARIES
Geometrically one should think of Q as an infinite-dimensional brick the sides of which get shorter and shorter. This can be demonstrated in the following way. Let x(n) £ Q be the point having all coordinates 0 except for the n-th coordinate which equals 1. So x(n) is the 'endpoint' of the n-th axis in Q. Intuitively, each x(n) has distance 1 from 0 and hence x(n) and 0 are far apart. However, the appearance of the factor 2~ z in the definition of Q implies that 0(z(n),0)=2- n , whence the sequence (x(n)) converges to 0 in Q. If n G N then by the above remarks, Q is naturally homeomorphic to oo
u n x TT n
«U
A
I I
aU m.
ra=n+l
It will be convenient to let Qn denote the product rim=n+i $™.For n € N and 9 e { — 1,1} by W% we denote the endface of the Hilbert cube Q. Groups. A group G which is also a topological space, is called a topological group if the function G x G —> G defined by (x,y) I-)- x -y~l
is continuous. This easily implies that a translation on G, i.e., a function from G into G of the form x !->• x • a or of the form x !->• a • x for some fixed a £ G, is a homeomorphism. Exercises for §A.l. Let X be a space with subset A. We say that p (E X is an accumulation point of A provided that p G A \ {p}. A topological space X is called homogeneous provided that for all x,y € A there exists a homeomorphism h: X —>• X such that /i(x) = y. So, loosely speaking, a space X is homogeneous if from the topological standpoint all points in X behave in the same way. For A,S e 0 3 (N), put V(A, S] = {B e 3>(N) : B n S = A n S}. 1. Let X be a set and let J C T(X). Prove that the following statements are equivalent: (1) 3 is an ultrafilter. (2) y has the finite intersection property, and if A C X then either
A.I. PREREQUISITES AND NOTATION
463
•2. Let X be a space without isolated points. Prove that the following statements are equivalent: (1) No point of X is simultaneously an accumulation point of two disjoint subsets of X. (2) If p is an accumulation point of A C X then p €E Int(A U {p}). (3) If x € X then the collection {A C X : x is an accumulation point of A} is an ultrafilter on X. 3. Let X be a space and let V be a countable (finite) collection of clopen subsets X . Prove that there is a countable (finite) collection clopen subsets W of X such that (1) every W G W is contained in an element V € V, (2) W is pairwise disjoint, (3) 4. Let X be a set, let k E N and let A be an infinite collection of subsets of X each of cardinality k. Prove that there are a set A C X with \A\ < k and a sequence (Ai)i of elements of A such that Ai C\ Aj = A for all distinct i and j. 5. Let A, B, E, F £ 7(N). Prove that the following statements are equivalent: (1) V(A,E)CV(B,F), (2) B n F C A n . E a n d F \ 5 C E \ A (hence F C E). 6. Let X be a space and let g be an admissible metric for X and let a 6 (0, oo). Prove that the function p: X —>• X x [0, oo) defined by p(x,y) = mm{a,g(x,y)} is an admissible bounded metric for X . Prove that if Q is complete, then so is p. 1 . Let A and B be subsets of a space X. Prove that for every x G X we have Q(x,A\JB)£{Q(x,A),e(x,B)}. 8. Let X and Y be spaces with Y Hausdorff. Prove that if / : X —>• Y is continuous then its graph
is a closed subspace of the product X x Y . (For a related result, see Theorem 1.1.15.) 9. Let X and Y be spaces with Y Hausdorff and let /, g : X —>• Y be continuous. Prove that the set
{x € X : f ( x ] = g(x)} is closed in X .
464
A. PRELIMINARIES
MO. For every n let Qn be an admissible metric for Xn. Assume that for some constant c G (0, oo) all the metrics gn are bounded by c. Define
by the formula
Prove n that g is an admissible metric for the Tychonoff product topology
° rinii^-
11. Let X be a metrizable space and let (xn)n be a Cauchy sequence in X. Assume that (xn)n has a convergent subsequence, say with limit x. Prove that linirj-^oo xn — x. 12. Let G be a topological group. Prove that G is homogeneous. 13. For every n let Xn and Yn be spaces and let fn : Xn —>• Yn be continuous. Prove that the function
defined by is continuous. Also prove that if every fn is an imbedding (homeomorphism) then / is an imbedding (homeomorphism). Conclude that if An is contained in Xn for every n then O^Li An ^s a subspace of H^Li Xn14. Let AI, . . . , An be a collection of subsets of a metric space X such that i-l
A, n U Ai ± 0 for every i < n with i > 2. Prove that 71
7T,
diam ( ^J An < ^ diam(Ai). 1=1 i=i 15. Let /: X —> Y be a closed map, let A C Y and let U be an open neighborhood of /"•'•[A] in X. Prove that there is an open neighborhood V of A in Y such that f~l[V] C U. 16. Let X be a space with subspace Y. Prove that if U C X is open then the boundary of U n Y in Y is a subset of the boundary of U in X. 17. Let /: X —> Y be a continuous surjection. Assume moreover that / is open and that f ~ l ( y } is finite for every y. Prove that if S is a dense GVsubset of X then f[S] is a dense GVsubset of Y. (See Exercise A.6.12 for an example showing that the finiteness of the fibers of the map / is essential.)
A.2. SEPARABLE METRIZABLE TOPOLOGICAL SPACES
465
A.2. Separable metrizable topological spaces
Our primary interest in this book is in infinite-dimensional and related topology and hence in separable metrizable spaces. This is why we do not bother to state our theorems in the first five chapters in their most general forms for this would only distract the reader from the essential parts of our considerations. For simplicity we therefore call a separable metrizable space simply a space. Non-separable metrizable spaces and non-metrizable spaces are considered in Chapter 6, where we study the topology of pointwise convergence on function spaces. If X is a countable space which is not necessarily metrizable, then its function space of real-valued continuous functions endowed with the topology of pointwise convergence is a linear subspace of the product Rx . So it is a linear subspace of a countable product of real lines, and is therefore a separable metrizable space. The space X can be rather 'bad' from a general topological perspective, its function space is always 'good'. As we will see, certain non-metrizable countable spaces X have very interesting function spaces and this prompts us to study non-metrizable spaces as well. It is not so that we consider the theory of separable metrizable spaces to be more important than the theory of general topological spaces. Topological spaces with applications in computer science have very bad separation properties and consequently are very far from being separable metrizable, yet they are important. Similar remarks can be made for spaces endowed with a weak topology of some sort in functional analysis, or CW-complexes in algebraic topology. However, we believe that for a good understanding of topology it is essential to make a thorough study of separable metrizable spaces first. In our belief generalizations should be postponed until having mastered the separable metrizable spaces. The separable metrizable spaces can be characterized in general topological terms, as follows. A topological space X is separable and metrizable if and only if X is a regular 7\-space with a countable base. This is called Urysohris Metrization Theorem. If X is separable and metrizable then X is regular and T\ for obvious reasons and has a countable base by Theorem A.2.1 below. Conversely, if X is regular and T\ and has a countable base then it is normal, being Lindelof, so by Urysohn's Lemma for normal spaces there are enough real-valued continuous functions to separate closed sets. The proof of Corollary A.4.4 then shows that X can be imbedded in Q and hence is separable and metrizable. (See ENGELKING [153, Theorem 4.2.9] for more details.) For any space X we let Q denote an admissible metric on X. Theorem A.2.1. Let X be a space. There is a countable collection of open subsets of X which is a base for its topology.
466
A. PRELIMINARIES
Remark A.2.2. This result will be used without explicit reference throughout the remaining part of this book. Proof. Let g be an admissible metric on X. There is by assumption a countable dense subset of X, say D. Put Then CB is a countable collection open subsets of X and we claim that it is a base. To this end, let x E X and U C X be open such that x E U. There is e > 0 such that B(x,e) C U. Let d E D be such that Q(x,d] < l/2£ and let q be rational such that 0(x, d) < q < l/2e. If 2 E -S(d, g) then
0(x, z) < 0(x, c?) + g(d, z) — e.
So we conclude that B(d,q) CB(x,e) C U. In addition, x E B(d,q) since Q(x,d] < q.
D
The following corollary to this result will be used quite frequently. Corollary A.2.3. Let X be a space and let U be a collection of open subsets of X. Then there is a countable subcollection VofU such that \JV — \JU. Proof. Let S be a countable base for X and put £' = {BeV: (3U eU}(B CU)}.
Then !B' is countable since !B is. For every B E !B' pick C/(5) E U such that 5 C U(B). We claim that the countable subcollection V ={£/(£) : B e ® ' > of U is as required. To this end, let x G U U be arbitrary. Pick U G U such that x G C/. Since !B is a base, there exists B G !B such that x £ B C. U. So 5 E 3' from which it follows that x E 5 C E/(.8). We conclude that x is contained in an element of V.
D
Exercises for §A.2. A subset D of a space X is called discrete if every x G D has a neighborhood Ux such that Uxr\ D = {x}. If X is a space then A C X is called a Gs -subset of X if there are open subsets Un Q X, n G N, such that A = H^Li ^- The complement of a (7,5-set is called an F&-subset and is characterized by being the union of a countable collection of closed subsets of X.
A.2. SEPARABLE METRIZABLE TOPOLOGICAL SPACES
467
Let X be a space and let x £ X. A collection open subsets U of X is called a local base at x if for open set V in X with x €E V there is an element U G II such that x eU CV. 1. Let X be a space, and suppose that A C A' is not closed. Prove that there is a sequence (an)n of points in A such that x = lhnn->-oo o,n exists and does not belong to A. 2. For each i G N let A; be a space. Show that a sequence (x(n)) in the product IlfcLi Xk converges to a point x 6 HfeLi Xk if and only if the sequence (x(n)i)n converges to x» for every i 6 N. (We say that flfcLi Xk is endowed with the topology of coordinatewise convergence.) 3. Let X be a space. Prove that every closed subspace of A is a G^-subset of X. Conclude that every open subspace of X is an FCT-subset of X. 4. Let X be a space. Prove that the intersection of a countable family of G<$-subsets of A is again a G<$-subset of X. Conclude that the union of a countable family of FCT-subsets of X is an Fa-subset of X. 5. Let X be a space with G^-subset A. Prove that if B is a GVsubset of A then B is a G<$-subset of X. Conclude that a GVsubset of a closed subset of A is a GS -subset of X. 6. Let A be a space. Prove that if D C X is discrete then D is a GVsubset of X. 7. Let X be a space and let DC be a chain of closed subsets of X. Prove that there is a countable subfamily 3C' of OC such that \J 3C' = (J 3C. 8. Prove that a space X is locally connected if and only if for every non-empty open subset U C X every component C of U is open. 9. Let A be a closed subset of a locally connected space X. Prove that if U is a component of A \ A then U U A is closed. 10. Let A be locally compact and locally connected. Prove that X has a base consisting of connected open sets with compact closures. 11. Let /: X —> Y be a continuous closed surjection. Prove that if X is locally connected then so is Y. 12. Let IX be a collection of open subsets of a space X which is a local base at every point of a subset A C X. Prove that there is a countable subcollection of U having the same property. Conclude that if U is a base for a space X then there is a countable subcollection of U which is also a base for X. ^13. Prove that each uncountable space can be written as the union of a countable subspace and a dense-in-itself subspace (this is the so-called CantorBendixson Theorem). 14. Let X be a space, and let A C A" be uncountable. Prove that if < is any linear order on A then there is an element x £ A such that
x € {a € A : a < x} D {a € A : x < a}.
468
A. PRELIMINARIES
15. Let X be a space and let E C X. For every n let lCn be a collection of open subsets of X such that E C \JUn and meshCU n ) < l/n. Prove that the collection \J^=l VLn is a local base at every point of E. 16. Let X be a space. Prove that X has at most c closed subsets. (Hence A^ has cardinality at most c and X has at most c open subsets.) 17. Let X and Y be spaces. Prove that the collection of all continuous functions X —> Y has cardinality at most c. 18. Let X be a space with closed subspace K. Let e: X \ K —>• (0, oo) be continuous. Prove that there is a continuous function S: X —>• [0, oo) such that (1) 6\K = 0, (2) ]fxeX\K then 0 < 5(x) < e(x). A.3. Limits of continuous functions Let X and (I7, g) be spaces. For all /, g G C(X, Y) put Q(f,9) = sup{g(/(x),5(a;)) : x € X}. Observe that £>(/,) E [0,oc]. If X — Y = N with the Euclidean metric g, f is the identity and g is the function n !->• 2n, then clearly g(f,g) = oo. As a consequence, £ need not be a metric on C(X, Y}, it is an 'extended' metric. Yet it is convenient to adopt some of the terminology of metrics and to treat Q as some sort of generalized metric. For example, we call a sequence (fn}n m C(X^Y) Cauchy if for each e > 0 there exists an N G N such that g ( f m f m ) < £ for all n,m > N. This is of course nothing but the ordinary definition of a Cauchy sequence in a metric space. Lemma A.3.1. Let X and (Y, g) be spaces. Let ( f n ) n be a g-Cauchy sequence in C(X,Y) such that for every x G X, limn_>00 f n ( x ) exists. Then the function f : X —> Y denned by f ( x ) = limn-^oo f n ( x ) is continuous. Proof. Let x 6 X. We shall prove that / is continuous at x. To this end, let e > 0 be arbitrary. There exists by assumption an TV E N such that £>(/ n ,/m) < e/s f°r ^1 n,m> N. Claim 1. For every y e X and m > TV, g ( f ( y ) J m ( y ) )
< %•
Proof. Let y 6 X. Since g ( f n ( y } , fm(y)) < £/z f°r all n,m > N and since f(y] is equal to lim^oo f n ( y ) , we obtain that indeed e ( f ( y ) , f m ( y } } < £/s for every m > TV. <0 So the sequence (fn)n converges uniformly to / on X. Fix an admissible metric d on X. Since /TV is continuous at x, there exists 6 > 0 such that
A.4. NORMALITY TYPE PROPERTIES
if d ( x , z } < S then Q(/N(X), /N(Z)} with d(x, z) < 6 we have g(f(x)J(z))
< e/3- Now for an arbitrary z 6 X
< (/(*), fN(x)) + g(fN(x), fN(z)) + g(fN(z), e
£
469
f(z))
£
< /3 + /3 + /3 = £.
Since e was arbitrary, we conclude that / is continuous at x.
D
Exercises for §A.3.
1. Let fn'. X —>• R. be continuous for every n. Let (Mn}n be a sequence of real numbers such that for every x £ X and n e N, \fn(x}\ < Mn- Prove that if Xl^Li Mn is convergent then J^^Li fn IS uniformly convergent (and hence continuous by Lemma A.3.1). 2. Give an example of a sequence ( f n ) n £ (7(1,1) such that for every x G I the limit f ( x ) = limra_>.oo fn(x) exists while the function / is not continuous. (This shows that the condition that the sequence ( f n ) n in Lemma A.3.1 is Cauchy cannot be omitted.) A.4. Normality type properties
Real-valued continuous functions are very important in topology. The following lemma shows that disjoint closed subsets can always be separated by such a function. Lemma A.4.1. Let X be a space. If A and B are pairwise disjoint nonempty closed subsets of X, then there exists a continuous function u: X —> I such that u~l(0) = A and u~l(l) = B. Proof. Let Q be an admissible metric for X. It is easily seen that the function a: X —> I defined by
_
e(x,A)
U(X)
~ g(x,A) + g(x,B)
is as required.
D
Let X be a topological space, and let A and B be disjoint closed subsets of X. A continuous function /: X —> R for which there are distinct elements a, b e R such that f[A] C {a} and f[B] C {6}, is called a Urysohn function. So the function a in Lemma A.4.1 is a 'special' Urysohn function since it has the extra property that it 'separates' A from B in a precise way. We will need a second type of Urysohn functions in the following situation later. Suppose that A and B are closed subsets of a space X. In addition,
470
A. PRELIMINARIES
let A' C A \ B and B' C B \ A be closed sets which are nonempty. Let g be an admissible metric for X. Define the sets A" and B" in X by A" = {x G X : 0(z,A) < £( Observe that these sets are closed in X , and that A" n (A U 5) = A,
B" n (A U B) = B,
A" U B" = X.
By the above there are Urysohn functions a, f3 : X —>• I such that Q- 1 (0) = A',
a- 1 (l) = JB",
p~1(ty=B',
/?- 1 (l) = A".
Define u : X ->• I by the following formula: u(x)
/V 2 «(x)
O r e A"),
~\i-Y 2 /?(z) (xeB»).
l
If x e A" n 5" then /2a(x) = l/2 and 1 - V2/3(x) = i/2. Since A" and B" are closed in _Y, this easily implies that w is well defined and continuous. So it is a Urysohn function. It is special since u-\0] = A',
u~l(l) =B',
and
«[A\fi]C[0,y2), u[5\A]C(i/2,l]. This can be verified quite easily. So we completed the proof of the following result: Lemma A. 4. 2. Let X be a space containing the closed sets A and B. If
A' CA\B,
B' C B \ A
are nonempty closed sets then there is a Urysohn function u : X —> I such that (1) u~l(0) = A' and u~l(l] = B' , (2) u[A \ B] C [0, i/2) and u[B \ A] C (i/ 2 , 1]. We will now present some simple consequences. Corollary A. 4. 3. Let X be a space and let A and B be disjoint closed subsets of X. Then there are disjoint open subsets U and V in X such that
ACU,
B C V.
Proof. By Lemma A. 4.1 there exists a continuous function u: X —>• I such that u\A = Q,andu\B = l.PutU = u~l [[0, i/2)] and V = u~l [( 1/2, 1]] , respectively. Then U and V are clearly as required. D Corollary A. 4. 4. Let X be a space. Then X can be imbedded in Q and in E°° .
A.4. NORMALITY TYPE PROPERTIES
471
Proof. Let !B be a countable open base for X. For every pair (E,F) of elements of 33 with E C F pick a Urysohn function u: X —>• I such that u[E]C{0},
u[X\F}C{l}
(Lemma A.4.1). Let U denote the countable collection of functions obtained in this way. The function e: X —>• I u defined by e(x) = (u(x))u€U is easily seen to be an imbedding (Exercise A.4.1). Since Q clearly imbeds in E°°, we are done.
D
We will now generalize Lemma A.4.1 in an interesting way. Lemma A.4.5. Let X be a space with closed subspace A, and let g: A ->• E be continuous such that \g(o)\ < K < oo for every a G A. Then there is a continuous function h: X —> E such that \h(x)\ < l/3K
for all x G A",
|0(o) - h(a)\ < 2/3K
for all a e A.
Proof. Let Bl = g~1[-K,-l/3K},B2 = g-l[i/3K,K]. Then Bl and B2 are closed and disjoint in A and so in X. By Lemma A.4.1 there exists a continuous function h: X —> I such that /i[-Bi] C {0} and h[B2] C {!}. Now let h — 2/3K(h — !/2). We claim that h is as required. Indeed, if x G X is arbitrary then 2 \h(x)\ 3K \ / l = / /O l
J h(x) 1/2 < 2/3^ /2 V / /^ / O / £ = Vs^IO
Moreover, if a G A then there are three cases to consider. First assume that a G -Bi. Then /?,(a) = 0 from which it follows that h(a) = —l/3K. Since g(a) G [—K, —l/3K] this shows that \g(d) — h(a}\ < 2/3K. A similar calculation can be made if a G B2. Finally, if a 0 B\ (JB2 then \g(d}\ < 1/3K, and since by the above \h(a)\ < 1/3K^ we get \g(a) — h(a)\ < 2/3K as well. D This result enables us to prove that certain continuous functions can be extended. Theorem A.4.6. Let X be a space with closed subspace A. If f : A —>• JJ is continuous then there is a continuous function F: X —> J such that F\A — f . Remark A.4.7. For nontrivial generalizations of this result, see §1.2. Proof. By induction on n > 0, we define functions hn: X —>• E such that \hn(x}\ < 1/3 • ( 2 /s) n
for
all x G X,
n
/(a) - ]T hi(a) < (2/3)n+1 i=0
for
all a G A.
472
A. PRELIMINARIES
We get the function ho by applying Lemma A. 4. 5 with g = f and K = I. If hi has been properly defined for i < n, then /in+i is obtained by applying Lemma A.4.5 with g = f - £ILo hi and K = (2/3)n+1. We claim that
is as required. Indeed, by Exercise A. 3.1, the series X^o hi is uniformly convergent, and so its sum F: X —>• R is continuous. Also, observe that
71 = 0
for every x G X. Finally, by construction it follows that F \ A = f .
D
Let X be a space. A compactification of X is a pair ( K , i ) , where K is a compact space and i : X —>• K is a dense imbedding. It will be convenient to identify X and i[X]. Compactifications are important because they allow us to study spaces as dense subspaces of compact spaces. This sometimes simplifies life considerably. If aX and bX are compactifications of X then we say that aX < bX provided that there is a continuous function / : bX —>• aX which extends the identity on X. If aX < bX and bX < aX then we call aX and bX equivalent compactifications. This terminology is justified by the following argumentation. Let /: bX —>• aX witness the fact that aX < bX. Similarly for g : aX -> bX . Then f o g : aX —>• aX extends the identity on X and hence has to be the identity on all of aX since by Exercise A. 1.9 continuous functions agree on a closed set (here we use that X is dense in aX}. It follows similarly that g o /: bX —>• bX is the identity. We conclude that / is a homeomorphism that leaves X pointwise fixed; its inverse is g. Corollary A.4.8. Every space has a compactification. Proof. Let X be a space. By Corollary A. 4. 4 we may assume that X is a subspace of the compact space Q. So the closure of X in Q is a compactification of X. D If X is a noncompact locally compact space then it can be compactified through the addition of a single point. This is the so-called Alexandroff (onepoint) compactification aX of X. The unique point in aX \ X is usually denoted by oo. Observe that X is necessarily open in aX (Exercise A. 4. 3) so that a basic neighborhood of a point x in aX that belongs to X has the form U with U an arbitrary open neighborhood of x in X. In addition, a
A.5. COMPACTNESS TYPE PROPERTIES
473
basic neighborhood of oo in aX has the form aX \ K, where K C X is an arbitrary compact set. Exercises for §A.4. 1. Prove that the function e: X —>• Iu in the proof of Corollary A.4.4 is an imbedding. 2. Let aX and bX be compactifications of X. Assume that there is a continuous function /: aX —>• bX such that / restricts to the identity on X. Prove that f [ a X \ X] = bX \ X. 3. Let X be locally compact, and let aX be a compactification of X. Prove that X is open in aX. ^•4. Let X be a space with subset t/, and /: X —> I continuous. Prove that the set
B = [s e (o, i ) : rl[(o,6)] n u / /^[[o,*)] nt/} is countable. ^•5. Let X be a space and let U be a countable collection subsets of X. Prove that for every closed set A C X and neighborhood V of A there is an open neighborhood W of A such that
(1) Wcv, _
_
(2) Py n U = U n W for every E7 € U. 6. Prove that every (separable metrizable) connected space contains one point only or has cardinality precisely c. 7. Observe that any endface W of Q is homeomorphic to Q. As a consequence, every space can also be imbedded in any endface of Q. 8. Prove that for every n the one-point compactification Rn U {00} of W1 is homeomorphic to S n . 9. Let X be a space containing the closed sets A and B such that A(JB = X. Prove that if A' C A \ B and B' C B \ A are nonempty closed sets then there is a Urysohn function u: X —> I such that
A.5. Compactness type properties A space X is called countably compact if every countable open cover of X has a finite subcover. In addition, X is called sequentially compact if every sequence in X has a convergent subsequence.
Theorem A.5.1. Let X be a space. The following statements are equivalent: (1) X is countably compact.
474
A. PRELIMINARIES
(2) X is sequentially compact. (3) X is compact. Proof. We prove (3) => (2). Let (xn)n be a sequence in X having no convergent subsequence. Then by Exercise A. 2.1, every subset of A = {xn : n E N} is closed. As a consequence, for every n there exists en > 0 such that B(xn,En)n(A\{xn})
= Q.
Then U = {(X \ A) U B(xn,£n} : n G N} is a countable open cover of X having no finite sub cover. We prove (2) => (1). Let U = {Un : n G N} be a countable open cover of X. If U has no finite subcover, then for every n there exists n
xn G X \ \J U,. 1=1
(*)
By assumption, there exists a convergent subsequence (xnk]k of (xn)nj say XTJ.J.
f
X.
There exists m with x G Um and so there exists K G N with xnk e Um for k > K . Now pick k > K such that rik > m. Then on the one hand ^rifc t Urni
but on the other hand 2-rifc ^ t7?7i
by (*). So we arrive at a contradiction. Finally (1) =>• (3) is trivial since J^C is Lindelof.
D
So in order to prove that a given space is compact, all one needs to do is to show that every sequence has a convergent subsequence. Corollary A. 5. 2. Let X be a space with complete admissible metric g. Suppose that for every n, lin is a finite collection of open subsets of X such that (1) mesh(U n ) < i/n, _ (2) for every V G Un+i there exists U G Un such that V C U. Then n—l
is a compact subset of X.
A. 5. COMPACTNESS TYPE PROPERTIES
475
Proof. By (2) and the finiteness of the collections Un it follows that K is closed in A, whence g \ K x K is complete. Let (xn)n be any sequence in K. We shall prove that it has a convergent subsequence. Since Ui is finite, there are an infinite set E\ C N and an element t/i G U such that {xn :n G £1} C Ul. In addition, since E\ is infinite, and U-2 is finite, there are an infinite subset E-2 C EI and an element U-2 G ^2 such that {xn : n G E2} C t/2.
Proceeding in this way recursively, we get a sequence #1 D £2 5 • • • 5 #n 5 • • •
of infinite subsets of N and elements Un G lin such that for every n, (*)
{xl : i G #4 C Z7n.
Pick n\ < n-2 < • • • such that n; G 2?i for every z. By (1) and (*) it clearly follows that (xni }i is Cauchy, and hence converges to a point p of A^. Since K is closed, p belongs to K. So we are done by Theorem A. 5.1. D The following simple result is quite important. Lemma A. 5. 3. Let X be a compact suhspace of a space Y and let U be a collection of open subsets ofY which covers X. Then there exists 6 > 0 with the property that every A C Y with diam(A) < 6 and which moreover intersects X is contained in an element U G U. Proof. Suppose, to the contrary, that such 6 does not exist. Then for every n G N we can find a subset An of Y such that (1) di (2) An intersects X, say xn G An n X, (3) An is not contained in any element of U.
Since X is compact, every sequence in A has a convergent subsequence (Theorem A. 5.1), so without loss of generality we may assume that
x = lim xn n—>oo
exists and belongs to A. There exists U G U such that x G U. Since U is open, there exists e > 0 such that J5(x,e) C U. In addition, there exists N G N such that xm G B(x, £/2) for every m > N. Now choose ra > N so large that l/m < £/2. Since the diameter of Am is less than l/m < e/2, it follows easily that Am C B(x,e] C f/, which is a contradiction. D A number 5 > 0 such as the one in the above lemma is called a Lebesgue number for U.
476
A. PRELIMINARIES
Corollary A.5.4. Let X be a compact subspace of a space Y and let U be an open neighborhood ofX in Y. Then there exists 6 > 0 such that B(X, 6) C U. Proof. Let 6 > 0 be a Lebesgue number for {U} (Lemma A.5.3). We claim that 6 is as required. To this end, take an arbitrary y £ B(X,6). There exists x £ X such that Q(X, y} < 6. As a consequence, {x: y} C U since diam({x,y}) < 6 and {x,y} n X ^ 0. D Corollary A.5.5. Let JVC be a finite closed cover of a compact space X. Then there exists 6 > 0 such that ifNCM and A C X satisfy (1) diam(A) < 6, (2) V7V 6 X: N n A ^ 0, then Proof. Fix a subcollection A C M with (~]A — 0 for a moment. Then { X \ A : A 6 .A}
covers X. Hence there there exists by Lemma A.5.3 a Lebesgue number
e(A) > 0 for that cover. This implies that if B C X has diameter less than e(A) then it is contained in an element of the cover and hence misses the complement of that element. We conclude that if B C X has diameter less than e(A] then for some A G A we have B n A = 0. The same argument can be repeated for every subcollection of M with empty intersection. Put 6 = min{e(.A) : A C M and ^\A = 0}. Since M is finite, 6 > 0. It is clear that 6 is as required.
D
We now state a result that at first glance seems rather curious. It will be used several times in the remaining part of this book. Lemma A.5.6. Let X be a compact space. There is a (countable) collection {(An,Bn] : r c e N } of pairs of disjoint closed subsets of X such that for any pair (E: F] of disjoint closed subsets in X there is n G N such that E C An and F C Bn. Proof. Let U be a countable open base for X. The collection V of all finite unions of the collection U = {U : U 6 U}
A. 5. COMPACTNESS TYPE PROPERTIES
477
is clearly countable. Let E and F be arbitrary disjoint closed subsets of X. Since U is a base and X is compact, E is the intersection of all elements of V which contain E. Similarly for F. Now if the collection
W = {V e V : (E C V) V (F C V}} has the finite intersection property then by compactness 0 ^ P | W = # n F = 0, which is a contradiction. Since V is closed under finite intersections it therefore follows that E and F are contained in disjoint members from V. D Exercises for §A.5. Let (X, g] and (Y, p) be spaces. A function /: X -» Y is called uniformly continuous provided that for every e > 0 there exists 6 > 0 such that for all a, b G X such that g(a, b) < 5 we have p(/(x), /(y)) < £• A metric g for a space X is called convex provided that Ds(De(A))
=DS+E(A)
for all closed AC X and S,£ > 0. Let /: X —> F be a continuous surjection. We say that / is irreducible if
f[A] ^ Y for all proper closed subsets AC X. If X is a space then A C X is called bounded in JC provided that f[A] is bounded in R for every / G 1. Let X be a compact space with admissible metric £>. complete.
Prove that g is
2. Let X be a space and let K be a compact subset of X. Prove that there exist x,y G K such that g(x,y) = diam(/f). 3. Let X be a space and let A and £? be disjoint closed subsets of X such that A is compact. Prove that g(A, B) > 0. 4. Let A' and (Y, Q) be spaces, where X is compact. Suppose that for / and g in C(X, Y) and £ e R we have e(/(oj),s(a;))
e(f,g)
(1) Let X and y be spaces where X is compact and let /: X —>• y be continuous. Prove that / is closed. (2) Give an example of a continuous surjection between compact spaces which is not open. (3) Let X and Y be spaces. Prove that the projection TT : X x Y —> X is open.
478
A. PRELIMINARIES
6. Let X, Y and Z be compact spaces. Assume that /: X —>• Y and g: Y —> Z are surjections such that / and g o f are continuous. Prove that g is continuous as well. 7. Let X and y be spaces and let A C X and £? C Y be compact. Prove that for every neighborhood E of A x B m X x Y there are open sets U C X and V C y such that Ax B C.U xV CE. 8. Let X and K be spaces with K compact. Prove that the projection
TT: X x K -+ X is closed. 9. Let X be compact and let /: X —»• Y be one-to-one. Prove that if f[X] is dense in Y then / is a homeomorphism from X onto y. 10. Prove that the Alexandroff compactification is topologically unique, i.e., prove that if aX and bX are compactifications of the locally compact space X such that \aX\X\ = l = \bX\X\ then there is a homeomorphism /: aX —>• bX which restricts to the identity on X. 11. Give an example of a locally compact space X and compactifications aX and bX of X, respectively, such that | a X \ X | = 2 = \bX\X\ while aX 96 bX. 12. Let X be a space and assume that X is not compact. Prove that X contains an infinite closed discrete subspace. 13. Let X be a non-compact space. Prove that there exists a continuous function A: X ->• (0,1] such that inf X[X] = 0. 14. Let X be a space. Prove that a closed subset of X is bounded in X iff it is compact. Let X be a compact space, and let Q be a metric on X. Assume that Q\ X x X —>• [0, oo) is continuous. Prove that g is an admissible metric. For every n, let Xn be a space. Prove that there are admissible metrics gn for Xn for all n such that the formula g ( x , y ) = mzy.{Qn(xn,yn} • n € N} defines an admissible metric for n^Lj. ^«17. Let (Cn)n be a decreasing sequence of compacta in a space X. Prove that for every neighborhood U of C — fl^Li Cn there exists an n 6 N such that CnCU. 18. Let X and Y be spaces, where X is compact, and let /: X —> Y be continuous. Prove that / is uniformly continuous, no matter which admissible metrics on X and Y are chosen.
A.6. COMPLETENESS TYPE PROPERTIES
479
19. For every n let Xn be a space which is not compact. Prove that fl^Li Xn is not cr-compact. 20. Prove that the standard metric Q of Q is convex. 21. Let X be a compact space and let /: X —> Y be a continuous surjection. Prove that there is a closed subspace A C X such that / \ A —>• Y is irreducible. 22. Let X be compact and let /: X —> y be irreducible. Prove that for every nonempty open subset U C. X there is a nonempty open subset f7* C y such that f~l[U^] is a dense subset of U. A. 6. Completeness type properties Let X be a space with admissible metric Q. As usual, we say that Q is totally bounded if for every e > 0 the open cover {B(x,e)
:xeX}
of X has a finite subcover. The following result improves Exercise A. 5.1. Theorem A. 6.1. Let (X, Q) be a space. The following statements are equivalent: (1) Q is totally bounded and complete. (2) X is compact. Proof. We prove (1) => (2). Let (xn)n be a sequence in X. We shall prove that it has a convergent subsequence. An appeal to Theorem A. 5.1 then does the job for us. Since Q is complete, it suffices to prove that (xn)n has a Cauchy subsequence. Since Q is totally bounded, the open cover {B(x,2~2} : x 6 X} has a finite subcover. The Pigeon Hole Principle gives us an infinite subset AI of N such that for all n, ra G AI we have
Continuing in this way recursively, we can construct a sequence of infinite subsets AI D A-2 3 • • • 3 Ak D • • • of N such that for all n, m G Afc we have
The subsequence ( x n k ] k of (x n ) n , where n^ = min A&, is clearly Cauchy. Tl^e prove (2) =» (1). This is Exercise A. 5.1.
D
480
A. PRELIMINARIES
So if one wants to verify that a given space is compact, all one needs to do is to show that it has an admissible metric Q which is both totally bounded and complete. Sometimes this is more convenient than proving that every open cover has a finite sub cover, or showing that every sequence has a convergent subsequence (Theorem A.5.1). A space X is called topologically complete if it admits an admissible complete metric. Observe that a closed subspace of a topologically complete space is topologically complete as well. But even more is true, we shall prove below that every GVsubset of a topologically complete space is topologically complete. Basic examples of topologically complete spaces are the compact spaces (Exercise A.5.1). They even have the curious property that no matter which admissible metric is chosen, it is always complete. We will generalize this in Corollary A.6.4 below by showing that locally compact spaces are topologically complete as well. Lemma A.6.2. If Xn is topologically complete for every n E N, then so is the product O^Li Xn. Proof. For every n 6 N let Qn be an abmissible complete metric on Xn. By Exercise A. 1.6 we may assume that every gn is bounded by 1. Now define
as in Exercise A. 1.10 by
n=l
We already know by that exercise that g is admissible on O^Li Xn- That it is complete is left to the reader in Exercise A. 6. 8. D Theorem A. 6. 3. Let X be a space. The following statements are equivalent: (1) (2) (3) (4)
X X X X
is topologically complete. can be imbedded in E°° as a closed subspace. is a G§ -subset of any space Y containing X . is homeomorphic to a Gs-subset of a topologically complete space Y
Proof. To begin with, let us establish the following Claim 1. Let 5 be a G^-subset of a space X. Then 5 can be imbedded in X x R°° as a closed subset.
A. 6. COMPLETENESS TYPE PROPERTIES
481
Proof. There exist closed sets Gi, i E N, in X such that
1=1 For i E N, let /; : S -> R be defined by
In addition, define f : S ^ X x R°° by Since the inclusion 5 <—>• Ar is an imbedding, so is /. We shall prove that is closed in X x R°°. To this end, let (xn)n be a sequence in 5 such that
Then (xn}n converges to a. Suppose that a £ 5, say a £ Gi. Since (xn)n converges to a we have linin^oo g(xn,Gi) = 0 from which it follows that lira fi(xn) — oo.
n—>oo
So by continuity of /, the i-ih coordinate of b must be equal to oo. This is a contradiction and we conclude that a E S. By continuity of / we have
/(a) = lim f ( x n ] = (a, 6), n —>-oo
so(o,6)€/[5].
0
We prove (1) =» (3). Let Y be a space containing X. It suffices to prove that X is a G$-subset of its closure (Exercise A. 2. 5). So assume, for convenience, that X is dense in Y. Let g be an admissible complete metric on X. For each n € N let 33n be a cover of X by open subsets of X each of ^-diameter less than l/n. There exists for each n E N a collection £ n of open subsets of Y such that £n I* X
= 23 n .
Let .En be the union of the family £ n . We claim that X is equal to the intersection of the En's and hence is indeed a G^-subset of Y. To see that this is true, take an arbitrary point y E fl^Li En- Since X is dense in Y, there exists a sequence (xi)i in X whose limit is y. This sequence is £>-Cauchy. For take e > 0 and let n e N be so large that l/n < e. There exists E e £ n such that y G E. Since (£{)i converges to y, all but finitely many of the Xi belong to E n X, which has ^-diameter less than l/n < e by construction. So the sequence (xi)i is 0-Cauchy which implies that y £ X because Q is complete. We prove (3) =>> (2).
482
A. PRELIMINARIES
By Corollary A.4.4 we may assume that X is a subspace of E°°. So (3) implies that X is a G^-subset of M°°. It consequently follows from the claim that X can be imbedded as a closed subspace of E°° x E°° which is obviously homeomorphic to R°°. We prove (2) =>• (4). This is trivial. We prove (4) ^ (1). By the claim, X is homeomorphic to a closed subspace of Y x R°°. Since the product of two topologically complete spaces is topologically complete (Lemma A.6.2), this establishes (1). D Corollary A.6.4. Every locally compact space is topologically complete. Proof. Let X be locally compact. Its Alexandroff compactification aX is topologically complete by Exercise A.5.1. Since aX \ X consists of one point only, X is open is aX. So X is topologically complete by Theorem A.6.3. D Remark A.6.5. Since R is topologically complete, and IP is a G^-subset of E, being the complement of the countable set Q, it follows that P is topologically complete by Theorem A.6.3(2). So it follows, in particular, that P x P is topologically complete by Lemma A.6.2. This will be used in Exercise A.6.71. A space X is a Baire space if the intersection of countably many dense open subsets of X is again dense. Or, equivalently, if the union of any countable family of nowhere dense subsets of X has empty interior. Theorem A.6.6. Every topologically complete space is a Baire space. Proof. Let g be an admissible complete metric on X, and let [Un : n e H} be a family of dense open subsets of X, where U\ = X. Let x £ X and e > 0. We will prove that B(x, e)nf]^ =1 Un ^ 0. By induction on n we will construct a point yn e X and a real number Sn > 0 such that (1) yi — x and 61 = e, (2) B(yn+i,6n+i) C B(yn, V2<$n) H Un+l for all n, (3) 6n+i < min{2-n, l/26n} for all n.
(We will not loose generality by assuming that e < I. If we do so then we do not need to require in (3) that 6n+i < 2~ n .) Suppose that yi,...,yn and 6i,...,6n have been constructed satisfying (1) through (3). Notice that B(yn, l/^n} is a non-empty open set and consequently intersects the 1 We will show later that P is homeomorphic to N°° (Corollary 1.9.9) from which these statements are trivial corollaries.
A.6. COMPLETENESS TYPE PROPERTIES
483
dense set Un+\. Let yn+i in Un+i fl B(yn, l/^n) be arbitrary. Since the intersection Un+\ fl B(yn, l/<2&n) is open, there exists 6 > 0 such that B(yn+i,S) C Un+l n B(yn, V 2 (J n ). Put 6n+i = min{5, 2~ n , V^n}- Then 8n+i is as required. The sequence (yn)n is clearly Cauchy since g(yn+i, yn) < 2~ n for every n. Hence it converges, say with limit y. If n E N then ym € B(yn, l/2^n) for every m > n and so
Hence 0(y,x) < e and y E H^Li Un- This is clearly as required.
D
Corollary A.6.7. Q is not topologically complete and so it is not a G$subset of M. Proof. It is clear that Q is not a Baire space. The collection {®\{q}:qe®}
consists of countably many dense open subsets of Q and has empty intersection. So we are done by Theorem A.6.6. D Remark A.6.8. So in some vague sense Q and P are of different 'complexity'. The field of mathematics dealing with such phenomena is called Descriptive Set Theory. See §A.13 for more information. It is not true that every Baire space is topologically compete. See Exercise A.6.7 for a counterexample. Since the product of topologically complete spaces is topologically complete (Lemma A.6.2), the question therefore naturally arises whether the product of Baire spaces is a Baire space. We will answer this question in Corollary A.6.10 below. A subset A of a space X is called meager if it is a countable union of nowhere dense sets. Clearly, a countable union of meager sets is meager and so is a subset of a meager set. Also observe that a set A C X is meager if and only if it is a subset of a countable union of closed and nowhere dense subsets of X. Lemma A.6.9. Let X and Y be spaces and let U be a countable collection dense open subsets of X x Y. Then there is a meager subset E C X such that for every x 6 X \ E and U e U we have that U n ({x} x Y) is dense in {x} x Y. Proof. Let J be a countable base for Y consisting of nonempty sets. For every F e J and U & U let
A(F, U) = {x € X : U n ({x} x F) = 0}.
484
A. PRELIMINARIES
Observe that U D (X x F) is dense in X x F. Since the projection IT: X xF->X is open and continuous (Exercise A. 5. 5), it follows that ir[U n (X x F)] is a dense and open subset of X. Since A(F, U)=X\7t[Un(X
x F)]
it follows that A(F, U} is a closed and nowhere dense subset of X. Pick an arbitrary xeX\ \J{A(F, U) : F € J, U 6 U}. We will show that C7 Pi ({x} x Y) is dense in {x} x F for every U E U, which is as required. Indeed, if this is not true for certain U E U then there is an F E J such that ( { x } x F ) n t / = 0. But then £ 6 A(F, [/), which is a contradiction. D This lemma easily implies the following interesting result. Theorem A.6.10. Let X and Y be Baire spaces. Then X x Y is a Baire space as well. Exercises for §A.6. Let (X, g) be a space and let e > 0. A subset F C X is called an e-net if for every x e X there exists y G F such that g(x, y) < e. 1. Let X be a space with subset A. Prove that if for some e > 0 the subset F of A is an e-net for A then F is an 2e-net for A. 2. Let (X, Q) be a space. Prove that g is totally bounded if and only if for every e > 0, X admits a finite e-net. ^•3. Let X be a space with admissible metric Q and subspace A. Prove that if Q is totally bounded then so is its restriction to A. 4. Prove that every space has an admissible totally bounded metric. 5. Let X be a space. Prove that X has an admissible metric Q for which there is a sequence {Un : n € N} of finite open covers of X such that mesh(II™) < Vn for every n. 6. For every n let Xn be topologically complete. Prove that the topological sum ^^L! Xn is topologically complete as well. 7. Prove that a space X is a Baire space if it contains a dense topologically complete subspace. Prove that the subspace Z = {(x,y)eR2 : ( x , j / e Q ) or (x,j/€P)} 2
of R is an example of a Baire space which is not topologically complete.
A.7. A COVERING TYPE PROPERTY
485
8. Prove that the admissible metric Q defined on O^Li Xn in the proof of Lemma A.6.2 is complete. ^•9. Let (X, g) and Y be spaces with Q complete and let /: X —>• Y be continuous. Prove that the following statements are equivalent: (1) / i s not both one-to-one and closed. (2) For some e > 0 there exist sequences (xk}k and (yk)k in X such that g(xk,yk) > £ for each k and liirifc-Kjo f(xk) = lim^oo f(yk)10. Let X be a Baire space. Prove that if S is a countable family of dense Cr<s-subsets of X then p) S is dense in X. (So in a Baire space any two dense Gj-subsets intersect.) 11. Let Xn be a Baire space for every n. Prove that fl^Li Xn is a Baire space. Give an example of a compact space X and an open continuous surjection /: X —> Y and a dense G^-subset S C X such that f[S] is not a G<$-subset of Y. A.7. A covering type property Let X be a space and let A and 13 = {B(A} : A e A} be covers of X (not necessarily by open or closed sets). We say that *B is a shrinking of A if for every A £ A, B(A] C A. Observe that if B(Ao),B(Ai) G £ are distinct then so are AQ and A\ and that the converse need not hold. We call *B an open shrinking if 2J consists of open subsets of X. A closed shrinking is a shrinking consisting of closed sets. Proposition A.7.1. Let X be a space and let U be an open cover of X. Then U admits a closed shrinking. Proof. First assume that U is countable. Enumerate it as {Un : n 6 N}. By Lemma A.4.1, for each n there exists a continuous function f n : X —> I such that /-1 [(0,1]] = Un. Define /: X -> I by
n=l
Since n=l
we get by Exercise A.3.1 that / is continuous. As U covers X, f(x] > 0 for every x £ X. Put
An = [x e X : fn(x] > ^}
(n e N).
By continuity of the functions fn and /, it follows easily that each An is closed. Also, since f ( x ) > 0 for all x, An C Un.
486
A. PRELIMINARIES
We claim that the collection {An : n G N} covers X. To this end, assume that there exists x € X such that for every n, x $ An. Then
for every n, so that /Or) =
2-"/n(x) < V2 ' n=l
^/^ ^
>
n=l
which is a contradiction since /(x) > 0. Now let U be an arbitrary open cover of X. Let V be a countable subcover of U (Corollary A. 2. 3). By the above, there exists a closed shrinking W = {W(V) : V e V}
of V. For each U 6 U define the subset E(U) C X by
_ r w(u) (u € v), Since W covers X, the collection £ = {£(£/) : U G U} is clearly a closed shrinking of U. D Let A be a collection of subsets of a space X. We say that A is locally finite provided that for every x G X there is a neighborhood Ux of x such that the collection {A 6 .A : A H [4 / 0} is finite. Observe that .A is countable. This can be seen as follows: every point in X has a neighborhood meeting only finitely many elements of A and countably many of these neighborhoods cover X by Corollary A.2.3. Proposition A.7.2. Let X be a space. Then for every open cover U of X there is an open shrinking V = {V(U) : U £ U} ofU which is locally finite. Proof. Let X be a space and let U be an open cover of X. If U has a finite subcover, say U' then there is nothing to prove. Simply put V(U) = U if U 6 U' and V(U] = 0 if U g U'. So assume without loss of generality that U does not have a finite subcover. By Proposition A.7.1 there is a closed shrinking {W(U} : U e U} of U. For each U G U let E(U) be an open subset of X such that
W(U] C E(U] C E(U) C U (Corollary A.4.3). The cover {E(U) : U 6 U} of X has a countable subcover by Corollary A.2.3, say £ = {E(U) : U e It'}, where U' C U is countable.
A.7. A COVERING TYPE PROPERTY
487
Observe that U' covers X as well. Let {Un : n G N} be a faithful indexing of U' (observe that U' is infinite) . For every n put Vn = Un\ \J
E^nj.
m
We claim that V = {Vn : n G N} is a locally finite open cover of X. Clearly V consists of open sets. For each x G X let n(x] be the smallest integer with x G Un(x). For this number n(x) we clearly have x G Vn(x). Consequently, V = {Vn : n G N} is a cover of X. We shall prove that V is locally finite as well. Take an arbitrary x G X. Since £ covers X, there is an n with x G E(Un}. Clearly E(Un] fl Vm = 0 for all m > 7i. Consequently, E(Un] is a neighborhood of x which intersects at most n members from V. Now for each U G U define V(U)CX by
Since V = {Vn : n G N} is a shrinking of U' — {Un : n G N}, this assignment is clearly as required. D A space X is called paracompact if for every open cover U of X there exists an open refinement V of U such that V is locally finite. Corollary A. 7. 3. Every space is paracompact. Let X be a space. A family 5F of continuous functions from X to I is called a partition of unity on X if for each x G X there exist a neighborhood Ux of x and a finite subset 3(x) of 3" such that (!) E/ 6 ?(x) /(?/) = * for each ytUx, (2) if / G J \ 3(x) and y G f/x then f(y) = 0. Each partition of unity on X is countable. This is so because countably many C/x's cover X (Corollary A. 2. 3) and the corresponding 2F(z)'s cover 3" with the possible exception of the constant function with value 0. If 7 is a partition of unity on X then we define Lemma A. 7. 4. Let X be a space and let "3 be a partition of unity on X. Then U(J) is a locally finite open cover of X. Proof. That U(3r) consists of open sets is clear. Take an arbitrary x G X and let Ux and 9~(z) be as in the above definition. By (1) there exists an / G y(x) such that f ( x ) ^ 0. We conclude that U(^) covers X. That 11(3) is locally finite follows immediately from (2). D
488
A. PRELIMINARIES
Let U be a locally finite open cover of a space (X, Q). We shall associate with U a certain family of continuous functions which will be useful here as well as in Chapter 3, as follows. For every U E U define KU : X —> E by
e(x,x\u) "
These functions are called the K-functions with respect to the cover U. Observe that the sum in the denominator of (*) contains at least one but at most finitely many non-zero terms, so that KU is well-defined. Also observe that KU(X) > 0. We next claim that each KU is continuous. This is easy. Take an arbitrary x G X. There is an open neighborhood W of x such that the set y={U eU:Ur\W ^0} is finite. Let pu denote the restriction KU \ W. Then for every y G W we have g(y,X\U)
pu(y) - ^ --f —„ . m Ev^o(y,x\v]
Since y is finite, pu is a continuous function on W. Since W is a neighborhood of x this implies that KU is continuous at x. Finally, observe that 5^E/eu Ku(x) = 1 for every x G X and that for every U G U, ^[(0,1]] CC7. We say that a partition of unity y on a space X is subordinated to a cover V of X if the cover 11(9") is a shrinking of V, in the following sense. We demand that y can be listed as {fv : V 6 V} such that
for every V G V. Theorem A. 7. 5. Let X be a space and let U be an open cover of X. Then there exists a partition of unity on X which is subordinated to U. Proof. By Proposition A. 7. 2, we may assume without loss of generality that U is locally finite. Let "K = {KU '• U £ U} be the set of K-functions with respect to U. Since U is locally finite, it is clear that % is a partition of unity on X which is subordinated to U. D Partitions of unity are very useful. We will demonstrate this in §1.1 by using them for obtaining a classical result in selection theory. Here we will present another useful application. Let X be a space. A function (not necessarily continuous) / : X -> R is called lower semi- continuous (abbreviated Isc) if f"1 [(£, oo)] is open in X
A. 7. A COVERING TYPE PROPERTY
489
for every t 6 EL Similarly, / is called upper semi- continuous (abbreviated use) if f~l [(— oo, £)] is open in X for every t 6 EL Corollary A. 7. 6. Let X be a space and let /, g: X —> ffi. be functions such that g < f while moreover f is lower semi-continuous and g is upper semicontinuous. Then there exists a continuous h : X —> M such that 9
Then (p is clearly continuous. Let x G U be given and let 3(x) = {KQI , . . . , Kqn } be the subset of consisting of those functions Kq for which Kq(x) / 0. Then x G Uqi fl • • -r\Uq so that g(x) < qi < f ( x ) for each i = 1, . . . , n. Hence iKqi (X) i-l
= (p(x)
< f(x)
i—1
•
KQi (x) =
f(x).
i-l
We conclude that g < (p < f on U. Define the function h : X —>• M by
(x e u}
>
We shall prove that h is continuous. Observe that h \U and h\X\U are continuous. Let (xn}n be an arbitrary sequence in U converging to an arbitrary
490
A. PRELIMINARIES
element x $. U. It suffices to prove that h(xn) —> h(x). There are two cases to consider. Suppose first that for infinitely many n we have h(xn) < h(x). Then g(xn) < g(x) for those n. Since g is use, g(xn) —> g(x) and so h(xn) -» g(x) = h(x).
The case that for infinitely many n we have h(x) < h(xn] can be completed similarly by using the fact that / is Isc. So h has all desired properties. D
Exercises for §A.7. Let (X, g) and (F, d) be metric spaces. Then /: X —» Y is called Lipschitz provided that d(f(x)J(y))
:F£j}
is discrete as well. 4. Let X be a space. Prove that the following statements are equivalent: (1) A^ is compact. (2) Every locally finite open cover of X is finite. •5. Let (X, g) be a space and 1C a collection of open subsets of X. Prove that there is a Lipschitz function e: X —>• I such that e~l [(0, 1]] = (JU while moreover for every x G X the open ball about x with radius e(x) is contained in an element of U. 6. Let X be a space and let /, g : X —>• R be Isc. For every x G X put h(x) =
mm{f(x),g(x)}.
Prove that h: X ->• R is Isc. A.8. Extension type properties
Let X be a space. It will be convenient to let pX denote the family of all closed subsets of X. In addition, for every x E X, we put g(x, 0) = oo.
A. 8. EXTENSION TYPE PROPERTIES
491
Lemma A. 8.1. Let X be a space with subspace Y. The function K : pY —>• pX
denned by K(A) = {x 6 X : g(x, A) < g(x, Y \A)} has the following properties:
(1) (2) (3) (4)
«(0) = 0, K(Y) = X, K(A) n Y = A for every A e pY, ifA,B epYandACB then «(A) C K ( B ) , if A, B e pY then K(A U B) = n(A) U «(#).
Proof. The straightforward verification that K is well-defined is left to the reader. Clearly, (1), (2), and (3) hold. For (4), take A,B e pY. Observe that by (3), K.(A) U K(B) C K(A U B ) . Let x e «(A U B). Since g(x,AUB)
€{g(x,A),g(x:B)}
(Exercise A. 1.7), without loss of generality g ( x , A ( J B ) = g(x,A). We shall prove that x E n(A). Since x G n(A U B}, (5)
Q(X, A) = Q(X,
A\JB)<e(x,Y\(A\JB)),
and trivially, g(x, A) = Q(X, A U B) < g(x, B}.
(6)
Since Y \ A C (Y\(A\JB))\JB, (5) and (6) and Exercise A.I. 7 imply that g(x, A) < g ( x , (Y \ (A U B)) U B) < g(x, Y\A),
so x e K(A).
n
Let X be a space. Subsets A and S of X are called separated if It is clear that if A and 5 are disjoint and are both closed, or both open, then A and B are separated. More interesting examples of separated sets are obtained in the following way. Let Y be a subspace of X and let U and V be disjoint subsets of Y that are open in Y. Then U and V are separated in X. For let U' be an open subset of X such that U' n Y = U. Then U1 n V = 0, i.e., F n U = 0. It follows similarly that U n V - 0. Corollary A. 8. 2. Let A and 5 be separated subsets of a space X. Then A and B can be separated by disjoint open subsets of X. Proof. It is clear that A and B are closed in their union A U B. Lemma A. 8.1, there exist closed subsets A' and B' of X such that AC A',
BCB',
BnA' = ®,
B'nA = ®,
A'uB'=X.
By
492
A. PRELIMINARIES
So U = X \ B' and V = X \ A' are disjoint open neighborhoods of A and £?, respectively. D We now turn to extendable continuous functions. Lemma A.8.3. Let Y be a dense subspace of X, let Z be compact, and let f : Y —> Z be continuous. The following statements are equivalent: (1) / can be extended to a continuous function J'.X—tZ. (2) For all closed sets A,B C Z we have
(Here closure means closure in X.) Proof. For (1) => (2), pick disjoint closed subsets A,B C Z. Since / extends / we have (*)
rl[A] C /-^A] and f-^B] C f^[B}.
Observe that by continuity of /, /-1[A] and f ~ l [ B ] are disjoint closed sets of X. So we get what we want from (*). The proof of (2) =£> (1) is more complicated. Pick an arbitrary x G X and let 'B(x) be the collection of all its open neighborhoods in X. Put
(Here closure means closure in Z.) Since for a finite subcollection U C we have P) U € *B(x) and C fl
f[UHY],
the family 3(x) has the finite intersection property (here we use that Y is dense in X). By compactness of Z, y(x) consequently has non-empty intersection. We shall prove that P)3~(x) consists of a single point. Let 7/1 and y-2 be distinct elements of Z. There exist closed neighborhoods V\ and 1/2 of y\ and y-2 respectively, such that V\ D V? = 0. From (2) it follows that
We may assume without loss of generality that x £ f
1
[l/i]. Thus
Since f[Y \ /"1[Vi]] misses Vi, its closure misses the interior of V\. This means that y\ $_ {~]3~(x). Observe that if x G Y then f ( x ) £ Pl^x) so that, by what we just proved, the unique point in p)3 r (x) is f ( x ) .
A.8. EXTENSION TYPE PROPERTIES
493
Assigning to x e X the unique point in (~) *F(x) , we define a function /: X —>• Z which extends /. It remains to check that / is continuous. Let V be an open neighborhood of f(x] in the space Z. Then
and so by compactness of Z there is a finite subfamily U C 'B(x) such that
| f[Ynu]cv. U£U
It is clear that f] U is a neighborhood of x for which /[H U] C I/.
D
This result has some very interesting consequences. Corollary A. 8. 4. Let Y be topologically complete. If X is a space, A C X and f : A —> Y is continuous, then there exists a Gs-subset S of X containing A so that f can be extended to a continuous function /: S -» Y. Proof. We may assume without loss of generality that A is dense in X (Exercise A. 2. 5). By Corollary A. 4. 8 we may assume that F is a subspace of a compact space Z. Let {(An,Bn) : n G N} be a collection of pairs of disjoint closed subsets of Z such as in Lemma A. 5. 6. For every n £ N put
Then Dn is closed in X and is disjoint from A. Now put T = X \ U^Li Dn. Then T is a GVsubset of X which contains A as a dense subset and / can be extended to a continuous function g: T —>• Z by Lemma A.8.3. Since Y is complete it is a 6?,$-subset of Z by Theorem A.6.3. So 5 = T n g~l\Y] is a G?5-subset of X containing A and / = g \ S is as required. D These elementary results imply an important result on extending homeomorphisms. The Lavrentieff Theorem A.8.5. Let X and Y be topologically complete spaces, let A C X and B C Y. For every homeomorphism f : A —> B there exist GS-subsets S C X and T C Y, respectively, such that ACS and B C T while moreover f can be extended to a homeomorphism f : S -> T. Proof. We may assume without loss of generality that A = X and B = Y (Exercise A.2.5). Let g: B —> A be the inverse of /. By Lemma A.8.4 there exist G^-subsets 50 of X and T0 of Y such that A C So and B C TO while moreover / can be extended to a continuous function F: SQ —> Y and g can be extended to a continuous function G: TO —> X. Put
s = So n F-^TO], T = TO n G-^SO],
494
A. PRELIMINARIES
respectively. Then S and T are Gj-subsets of X and F, respectively, such that A C 5 and B C T. Observe that GoF \S: S ->• A" is well-defined and is the identity on A. Since A is dense in 5, Exercise A. 1.9 gives us that GoF \ S is the identity on 5. Likewise, F o G \ T is the identity on T. D Exercises for §A.8. ^•1. Construct a continuous function /: P —»• I such that for no q e Q there exists a continuous extension /: PU {q} —>• I. 2. Let /: Q —>• I be continuous. Prove that there exists an element p (E P such that / can be extended to a continuous function /: Q U {p} —> I. ^•3. Let X and Y be spaces and let /: X —» Y be a continuous surjection. In addition, let S be a subset of X such that / \ S: S —>• Y is a homeomorphism. Prove that S is a G^-subset of X. A. 9. Wallman compactifications
Compactifications are very important in topology. They allow to study spaces as subspaces of compact spaces in a convenient way. There is a simple method of compactifying an arbitrary space. This is the so-called Wallman compactification method, which we will describe here. Let X be a space. A collection 3 of closed subsets of X is a base for the closed subsets of X (abbreviated: closed base) if the collection {X\F : F 6 3} is a base for the open subsets of X . It is clear that 5" is a closed base if and only if for every x £ X and closed subset A C Xsuch that x £ A there is an element F € 3 such that A C F and x £ F. A collection 7 of subsets of X is called a Wallman base of X if (1) y is countable and a closed base of X, (2) y is closed under finite intersections and finite unions, (3) if F0, FI 6 y are disjoint then there exist F^F[ e 3~ such that and
(4) if F e y and x 6 X \ F then there is an element F' e 3~ with x 6 F' and F' n F = 0. Lemma A. 9.1. Let S be an arbitrary countable family closed subsets of X. Then there is a Wallman base y of X such that S C J. Proof. Let "B be a countable open base of X, and put §! = {X \ B : B 6 £} U § U {B : B G S}.
A. 9. WALLMAN COMPACTIFICATIONS
495
Then §1 is a closed base of X and contains S. Now let §2 be the collection of all finite intersections of finite unions of elements of Si. Observe that 82 is closed under finite intersections and finite unions. Let So, Si G S be disjoint, and let / : X —> I be a Urysohn function such that
/[S 0 ]c{o}, (Lemma A.4.1). Put A = f ~ l [ [ 0 , l/2]] and B = f~l [[l/2, 1]], respectively. Then A and B are obviously closed, while moreover
and
So if we replace S2 by S2 U{^4, B} then (3) in the definition of a Wallman base is satisfied for the pair So, Si E 82. So by adding countably many closed sets to 82, condition (3) in the definition of a Wallman base is satisfied for all pairs of disjoint elements of 82- Let 83 be the collection of all finite intersections of finite unions of elements of the new collection. Now continue as above with 83, add new sets, and take finite unions of intersections, to create 84. Etc. At the end of this process, the collection
5= n=l
is the desired Wallman base of X. (Observe that (4) is trivially satisfied since J contains the collection {B : B E £}.) D Let X be a space with Wallman base 3~. A subcollection S C 5F is called centered if it has the finite intersection property. In addition, 8 C 3 is called maximally centered provided that S is centered but not properly contained in another centered subcollection of y. A maximally centered subcollection of y is also sometimes referred to as an J-ultrafilter. It is easy to prove the existence of 3~-ultrafilters. Indeed, if x E X then put
This is an ^-ultrafilter. To show this, first observe that i(x] is clearly centered. To check maximality, let F E $ be such that i(x] U {F} is centered. Our task is to prove that F £ i(x). This is equivalent to proving that x £ F. But this is trivial. For if x <£ F then there is by (4) in the definition of a Wallman base an element S £ 3~ with x £ S C X \ F. But then S £ i(x), which violates the fact that i(x] U {F} is centered. The following lemma shows that there are in fact 'many' 3~-ultrafilters. Lemma A.9.2. Let "3 be a Wallman base of a space X. Then every centered subfamily S C 3 is contained in an y-mtrafilter.
496
A. PRELIMINARIES
Proof. Enumerate 3 as {Fn : n £ N}, and, recursively, define subcollections Sn of y, as follows. Put Si = 8. Now assume that Sn has been defined. If §n U {Fn} is centered then S n +i = Sn U {-^n}, otherwise Sn+i = S n . So we try to 'catch' as many elements of 3 as possible. It is clear that
n=l
is a maximally centered subcollection of y which contains S.
D
We now study the J'-ultrafilters a little closer. Lemma A. 9. 3. Let 3 be a Wallman base of X. ultrafilter then
If S is an arbitrary 3-
(1) 00?, (2) (3) (4) (5)
3 is closed under finite unions and finite intersections, if S € S and F 6 3 contains S then F € S, if F € 3 \ S then there is an element S e S with F n 51 = 0, if A, 5 e y and A U B = X then A e 3F of B e J.
Moreover, if S' is an J-ultrafilter such that S 7^ 8' then there exist A 6 S and B e §' such that A n 5 = 0. Proof. Since S is centered, (1) is trivial. If S ^ S is an arbitrary finite subcollection, then
Su{p|S} is centered since S is. Since f) S £ ^ this shows that Q S 6 8. So 8 is closed under finite intersections. That S is also closed under finite unions, follows by a similar argument. This proves (2). Statement (3) is trivial. For (4), observe that
8U{F} is not centered since 8 is maximal. So there is a finite subcollection S Q 8 such that
p|SnF = 0. But S = P| 9 G 8 by (2). Hence S is as required. For (5), assume that A g S. By (4) for some 5 e S we have S n A = 0. So 5 C B which implies by (3) that B G 8. Since S ^ 8' we may without loss of generality assume that there is an element A e S\8'. There exists by (5) an element B e 8' such that Ar\B = 0. So we are done. D
A.9. WALLMAN COMPACTIFICATIONS
497
Let X be a space with Wallman base y. Put u(X, 3} = {§ C y : S is an J-ultrafilter}. To every x G X we associated an element i(x) G u(X, 3") above. If a:, y G -X" are distinct then there is an element F G 3* which contains y but not x. So F witnesses the fact that i(y] ^ i(x}. We therefore conclude that the function i: X -*u(X,y] is injective. It will be convenient to identify x G X with i(x] G w(A", 3"). So we regard X to be a subset of w(X, J). For every F G 3" put F* = {§ e u(X, J) : F G § } . Lemma A.9.4. Let X be a space with Wallman base 3". If F, G G 3" then
(1) F*HA: = F, (2) if F n G = 0 then F* n G* = 0,
(3) F*nG* = (FnG)*, (4) F * U G * = ( F U G ) * , (5) ifFuG = X then F* U G* = w(X, 3r). Proof. This is basically a reformulation of Lemma A.9.3. The proof of (1) is trivial. Similarly for (2). In addition, (3) and (4) follow from Lemma A.9.3(2) and (3). Finally, (5) follows from (5) of that same lemma. D So the collection (*)
J* = {F* : F G 3-}
is closed under finite unions and finite intersections. So the collection of its complements is closed under finite intersections and finite unions as well. As is thought in every first year topology course, such a collection It of subsets of a set Y can be used to define a topology on F, as follows: a set O C Y is open if and only if for every y G O there is an element U G U such that y G U C O.
We can therefore topologize u(X, 3") by using the collection H^,J)\F*:FG3-} as a base for its topology. So the collection 3~* is a base for the closed subsets of u(X, 3"). By Lemma A.9.4(1) this implies that the injective function i: X
-*u(X,f)
is an imbedding. We conclude that X is a subspace of u(X, 3r).
498
A. PRELIMINARIES
Lemma A. 9. 5. Let X be a space with Wallman base 3. Then (1) X is dense in u(X, J), (2) w(X, 9r) is a separable metrizable space, (3) Uj(X,y) is compact. Proof. For (1), let F G 5 be such that Our task is to prove that U n X / 0. This is trivial. For take an arbitrary element S G U. Then S 0 F* or, equivalently, F ^ §. So there is an element S G S with Sr\F = 0 by Lemma A. 9. 3(4). Then S is not empty and is contained in U n X. We claim that X is Hausdorff. To this end, let So, Si be distinct elements in u}(X,3). There are elements So G So and Si G §1 such that So fl Si = 0. By (3) in the definition of a Wallman base there are elements A, B G 3" such that S0 C A, SiCB, AUB = X, and
B n S0 = 0 = Si n A. But then by Lemma A. 9. 4, uj(X,?}\B*,
u(X,f)\A*
are disjoint open neighborhoods of §0 and Si , respectively. We next claim that u(X,y) is compact. This is simple. If U is an open cover of oj(X, 3") by basic open sets having no finite subcover then the collection of its complements is centered but has empty intersection. So there is a subfamily "K of 3 such that the collection {H* : H G 'K] is centered, but has empty intersection. This leads to a contradiction, as follows. Notice that the collection CK is centered by Lemma A. 9. 4. So by Lemma A. 9. 2 it is contained in an 5F-ultrafilter S. But then
which is a contradiction. So c<j(X, J) is a compact Hausdorff space with a countable base. Since a compact Hausdorff space is regular, it must be separable and metrizable by Urysohn's Metrization Theorem (see Page 465). D So we conclude that u(X, 7) is a compactification of X. It is called the Wallman compactification of X with respect to X. Perhaps the reader feels that Wallman compactifications are esoteric objects. But this is not true: we will show that every compactification of an
A. 9. WALLMAN COMPACTIFICATIONS
499
arbitrary space is a Wallman compactification. Once one is used to the language of Wallman compactifications, they are a very convenient tool. Proposition A. 9. 6. Let aX be a compactification of X. Suppose that aX has Wallman base J such that for every nonempty F E !f we have FC\X ^ 0. Then the compactifications aX and u(X, 3} of X are equivalent. Proof. Define y. aX ->• u(X, 3) by
^(p) = {F A moments reflection shows that (f> is a homeomorphism which extends the identity on X. D So now that we know how Wallman compactifications are characterized, we aim at proving that all compactifications are Wallman. Theorem A. 9. 7. Every space X has a Wallman base every nonempty element of which has nonempty interior in X. Proof. This will follow by using the same technique as in the proof of Lemma A. 9.1 and by applying Exercise A. 4. 5. Indeed, let 25 be a countable base for X. Let {(Bfi, B? ) : n € N} list all pairs of elements B0, #1 € $ such that BO C BI. By Exercise A.4.5 it is easy to construct for every n an open set Un C X such that
while moreover for any finite subset F C N the equality H^F ^ = HieF Ui holds. So if V is the collection of all finite unions of finite intersections of elements of {Un : n G N} then V has the following property: if £ C V is an arbitrary finite subcollection then f) £ = H^ge E. So if a finite intersection of closures of elements of V is nonempty then this intersection has nonempty interior. Precisely such as in the proof of Lemma A.9.1 we can continue this process, using Exercise A.4.5 at each stage of the construction to ensure that we do not create sets with empty interior. D Corollary A.9.8. Every compactification is a Wallman compactification. Proof. Let aX be a compactification of a space X. By Theorem A.9.7 the space aX has a Walman base 3~ each nonempty element has nonempty interior in aX. Hence each nonempty element of ^ intersects the dense set X. So Proposition A.9.6 now gives us that aX and u(X, 3 \ X} are equivalent compactifications. D
500
A. PRELIMINARIES
Exercises for §A.9. 1. Let X be a space with Wallman base y. Prove that for every F € 3~ the set F* coincides with the closure of F in u(X, 3}. 2. Let X be a space with Wallman base y. Prove that for all disjoint nonempty closed sets A and B in ui(X, 3") there are disjoint Fo,Fi € 3 such that A C F0* and B C F? . 3. Let X be a space and let S be a countable collection of closed subsets of X. Prove that X has a compactification jX such that if So, Si £ § are disjoint then their closures in ~fX are disjoint as well. 4. Let y and S be Wallman bases of X. Prove that the following statements are equivalent: (1)
u(X,S)
(2) for all disjoint FO, FI £ 3" there exist disjoint Go, G\ G S such that Fo C Go and FI C Gi. A. 10. Connectivity If x 6 X then it is easy to see that the component of x is equal to C X : C connected and x 6 C}. r
Since if A C A is connected then so is ,4, it follows that the component of x is a closed subset of X. Lemma A. 10.1. Let C be a component of the compact space X. Then for every neighborhood U of C there exists a clopen subset E of X such that C C E C U. Proof. Let C(x) denote the component of the point x 6 X. In addition, let 9 be the family of all clopen neighborhoods of x, and put
Claim 1. For every neighborhood U of Q(x] there exists a clopen subset E of X such that Q(x) C E C U. Proof. Observe that the intersection of finitely many elements of 9 again belongs to 9- Without loss of generality we assume that U is open. Since
it follows by compactness that U contains an element of S-
0
It is clear that C(x) is a subset of Q(x), since if A is any clopen subset of X then either C(x) n A = 0 or C(x) C A. We will prove that Q(x) is connected. If this is true then Q(x) C C(x), i.e., C(x) = Q(x). So then we get what we want by Claim 1.
A.10. CONNECTIVITY
501
Striving for a contradiction, assume that Q(x) is not connected. We shall derive a contradiction. We can write Q(x) as E U F, where both E and F are nonempty closed subsets of X and EnF = 0. Without loss of generality, we may assume that x € E. There exist disjoint open neighborhoods U and V of E and F, respectively (Corollary A.4.3). Since U U V is open and contains Q(x), by Claim 1 there exists a clopen subset G of X such that <2(V) C G C £7 U F. Observe that
Gnu=G\v is both open and closed. But this implies that Q(x) C G C\U and so F = 0, which is a contradiction. D Corollary A.10.2. Let X be a compact space every component of which has diemater less than E. Then X can be split into finitely many clopen sets of diameter less than E. Proof. Let C be an arbitrary component of X. By Lemma A.10.1 there is a clopen neighborhood Uc of C with diameter less than E. Finitely many t/c's cover X by compactness of X. So we are done by Exercise A.1.3. D Let y be a decomposition of X into pairwise disjoint nonempty closed subsets. We say that CP is upper semi-continuous provided that for every closed subset A of X, \J{P G ? : P n A ^ 0} is closed in X. Proposition A.10.3. Let X be compact. Then the collection of all components of X is an upper semi-continuous decomposition of X. Remark A. 10.4. The compactness of X is an essential in this result. See Exercises A.10.9 and A.10.11. Proof. Let C denote the collection of all components of X and let A be closed. We shall prove that if ? = {€ C : C n A / 0 } then (J f is closed in X. Take x $. \J J". Then the component C(x) of X which contains x misses A, hence by Lemma A. 10.1, for some clopen E in X, C(x] CECX\A. But then E n (J P = 0 and so x g \J 7.
D
A continuum is a space which is both compact and connected. Corollary A.10.5. Let A be a closed subspace of a continuum X such that 0 ^ A ^ X. Then every component C of A meets Fr A. Proof. First observe that if Fr A = 0 then A — X by connectivity of X. So Fr A 7^ 0. Striving for a contradiction, suppose that C is a component of A such that C n Fr A - 0. Then X \ Fr A is a neighborhood of C. By compactness of A there exists by Proposition A.10.3 and Lemma A.10.1 a
502
A. PRELIMINARIES
relatively clopen subset K of A such that C C K C X \ Fr A. Then K is closed in X since it is closed in A and A is closed in X. On the other hand, K misses Fr A and hence is contained in the interior of A in X which is open in X. So K is a relatively open subset of the interior of A in X and hence is open in X as well. We conclude that K is a proper nonempty clopen subspace of X since Fr A ^ 0, which contradicts connectivity. D This result enables us to formulate and prove the following interesting result. To put it into perspective, first observe that no connected space admits a partition into finitely many but at least two nonempty closed sets. With the extra condition of compactness added to connectivity we can do a little better. The Sierpiriski Theorem A.10.6. No continuum can be partitioned into countably many pairwise disjoint closed and nonempty sets. Proof. To begin the proof, we first establish the following. Claim 1. Let X be a continuum which admits a partition £ consisting of countably many closed sets of which at least two are nonempty. Then for every E E £ there is a continuum C C X \ E meeting at least two distinct elements of £. Proof. If E — 0 then we let C — X. So we may assume that E ^ 0. Pick a nonempty element E' € £ \ {E}. There are by Lemma A.4.1 disjoint open neighborhoods U and V of E and E', respectively. Let x be an arbitrary point in E' and let C be the component of x in V. Then clearly C n E = 0 and E'nC ^ 0. Also, CnFr V + 0 by Corollary A. 10.5. Since E' is contained in the interior of V it follows that Fr V D E' = 0. As a consequence, there exists y 6 C \E'. There consequently has to be an element E" £ £ \ {E'} such that C n E" ^ 0. <> The proof can now easily be completed by recursion, as follows. Assume that X is a continuum admitting a partition {Xi : i E N} consisting of closed sets of which at least two are nonempty. By the Claim there exists a decreasing sequence of continua Ci D C% D • • • in X such that Ci fi Xi = 0. To keep the recursion going, we assume that each Ci meets at least two distinct Xj's. Then by compactness, 0 / f)Si C, C X \ (J^ Xl = 0. This is a contradiction. D The following simple result is basic. Proposition A.10.7. Let X be a space and let (Cn}n be a decreasing sequence of subcontinua of X. Then C — f]^Li Cn is a continuum.
A.10. CONNECTIVITY
503
Proof. To the contrary, assume that C is not connected. Let E and F be disjoint relatively clopen nonempty subsets of C such that E U F — C. By Lemma A.8.1 there exist disjoint open subsets E' and F1 of X such that
E'nC = E, F' n C = F. Then U = E' UF' is a neighborhood of C and hence by Exercise A.5.17 there exists n G N such that Cn C E' U F'. But this contradicts the connectivity of Cn since C C Cn and C meets both E' and F'. D Let ^L be a space, and let x and y be elements of J\T. We say that x and y can be joined by a path provided that there is a continuous function a:I—tX such that a(0) = x and a(l) = y. A function such as a is called a pat/i from a; to y. Observe that if x and y can be joined by a path then so are y and x. Exercise A.10.1 below implies that for a space X the relation 'can be joined by a path' is an equivalence relation. An equivalence class of this relation is called a path-component of X. Observe that path-components are connected. A space X is called path-connected if it has one path-component only. A path-connected space is clearly connected. But the converse need not be true. Consider e.g., the well-known sin( l / x }-continuum {(ar,sin(i/ x )) . 0 < x < i/n} u {(0,y) : -1 < y < 1} in the plane (Exercise A. 10.10). A space X is called locally path-connected if every point has arbitrarily small path-connected neighborhoods. Lemma A.10.8. Let X be a connected space. Suppose that every x £ X has a path-connected neighborhood. Then X is path-connected. Proof. Pick an arbitrary x G X and put U = {y G X : x and y can be joined by a path}. Choose an arbitrary y 6 U and let V be a path-connected neighborhood of y. We claim that V C. U. To prove this, let w G V be arbitrary. Since y e U and V is a neighborhood of y, there is a point v 6 Vr\U. Since u, w 6 V and V is path-connected, w and v can be joined by a path. Moreover, since v E U it follows that v and x can be joined by a path. By Exercise A. 10.1 we get that w and x can be joined by a path, i.e., w E U. We conclude that indeed, W C U. Since y E U was chosen arbitrarily, it follows that U is both open and closed. Since clearly x E C7, the connectivity of X implies that U — X. Now choose arbitrary a, b E X. Then a and x can be joined by a path. Similarly, b and x can be joined by a path. By Exercise A. 10.1 we consequently get that a and b can be joined by a path. D
504
A. PRELIMINARIES
Corollary A.10.9. Every connected, locally path-connected space is pathconnected. Let X be a space and let A and B be two disjoint closed subsets of X. A partition between A and B is a closed subset S C X such that X \S can be written as the disjoint union of open sets U and V with A C [/ and S C I / . A typical partition between A and B has the form /~ 1 ( 1 /2) 5 where / is a Urysohn function with f[A] C {0} and f[B] C {!}. So by Lemma A.4.1 any pair of disjoint closed sets has a partition. Let X be a space and let A and B be disjoint closed subsets of X. By a continuum from A to B we mean a continuum C in X meeting A as well as B. Observe that if 5 is a partition between the closed sets A and B then every continuum from A to B meets S. Let A and B be disjoint subsets of a space X. A closed set K C X is called a cut in X between A and 5 if ^ n (A U S) = 0 and K n F ^ 0 for every continuum F C X from A to 5. In particular, if there is no continuum in X which meets both A and B then 0 is a cut in X between A and B. So a partition is a cut; the converse need not be true as we will see in Exercise A.10.7. A set T in a space X separates X if X \ T is disconnected. A point x in a space X is said to be a cut point of X if {x} separates X. In Exercise A. 10.4 we will show that if x is a cut point of X then X \ {x} can be written as U U V, where U and V are disjoint, open and nonempty subsets of X.
Exercises for §A.10. A continuous surjection /: X —> Y is called monotone provided that for every y G Y the fiber f~l(y) is connected. 1. Let X be a space. If x, y € X and y, z £ X are both joined by a path then so are a; and z.
2. Let X be connected and let 7 be a finite cover of X consisting of nonempty closed sets. Let n — \J\. Prove that we can list 7 as such that for any i < n with i > 2 we have i-l
Fl n (J F3- ^ 0.
A.11. THE QUOTIENT TOPOLOGY
505
3. Let X be a space with connected subset A. Prove that if {F . . . , Fn} is a collection of closed subsets of X which covers A then n
diam(A) < 1=1
4. Prove that if x is a cut point of X then X \ {x} can be written as U U V, where U and V are disjoint, open and nonempty subsets of X. 5. Let X be a continuum and let U and V be disjoint nonempty open subsets of X such that X\(U\JV) is a single point, say p. Prove that both U\J{p} and V U {p} are continua. ^•6. Let X be a nontrivial continuum. Prove that X contains at least two noncut points. 7. Give an example of a continuum X containing three distinct points p, q and r such that r is a cut between p and q but not a partition. 8. Let X be a continuum and let 3 be a countable family pairwise disjoint closed subsets of X. Prove that X \ [J 3~ is uncountable. 9. Construct an example of a space X having a point x for which the component C(x) of x in X does not satisfy the conclusion of Lemma A.10.1. 10. Prove that the sin(Y x )-continuum in the plane is not path-connected. 11. Construct an example of a space X the components of which do not form an upper semi-continuous decomposition of X. 12. Give an example of a path-connected compact space which is not locally connected. 13. Let /: X —>• Y be a continuous surjection which is both closed and monotone. Prove that for any connected E C Y we have that f~1[E] is connected. 14. Let n > 1 and let K C W1 be compact. Prove that W1 \ K has a component U such that W1 \ U is bounded. Conclude that U is the only unbounded component of W1 \ K. A . l l . The quotient topology
Let f be a decomposition of a space X into pairwise disjoint nonempty closed sets. In addition, let q: X —> 7* be the function sending x € X to the unique element of 3> containing x. Call a subset U of $ open if and only if g"1^] is open in X. In this way we endow J* with the quotient topology derived from X and q. With this topology we denote J" by X/1? and call it X modulo 7*. A function /: X —> Y is called quotient provided that U C. X is open if and only if f~l[U] is open. Observe that a quotient map is continuous. It
506
A. PRELIMINARIES
is clear that if / : X —>• Y is quotient and surjective then Y is canonically homeomorphic to the space XfP, where 5> is the decomposition l
{f- (y}--ytY}
ofX. If X is a space, and A C X is closed, then 3> = {A} U { {x} : x G X \ A} is a partition. It will be convenient to denote X/CP by X/A in this case. So X/A is the quotient space obtained from X by collapsing A to a single point. Lemma A. 11.1. Let 9 be a decomposition of a space X into pairwise disjoint closed sets. The following statements are equivalent. (1) The natural quotient mapping q: X —> XfP is closed. (2) CP is upper semi-continuous. Proof. For every closed set A C X we have
From this the implication (1) =>• (2) is obvious. On the other hand, if (2) holds then for every closed set A C X we have that q~l \q[A^ is closed from which it follows that q[A] is closed since we are dealing with the quotient topology. D Due to our self-chosen convention to deal with separable metrizable spaces exclusively, we are in a very unpleasant situation now. It is rather trivial to prove that the quotient topology just defined is indeed a topology, but it is not necessarily separable and metrizable (see Exercises A. 11.1 and A. 11. 3). We are primarily interested in decompositions of which the individual elements are compact. Fortunately, in that particular situation the decomposition spaces involved are separable and metrizable, as the following result shows. Theorem A. 11. 2. Let X be a space and let 9 be an upper semi-continuous decomposition of X consisting of compact sets. Then X/7 is separable and metrizable. Proof. First observe that Y = XfP is T\ since the elements of CP are closed sets. We next claim that Y is regular. Consider the natural quotient map q: X ^Y.
Then q is closed by Lemma A. 11.1. Let y E Y and let U be an open neighborhood of it. Since q ~ l ( y ) is closed, and q~l[U] is open, there is a closed neighborhood V of q ~ l ( y ) such that V C q~l[U] (Corollary A. 4. 3). Since q is closed, by Exercise A. 1.15 there is a neighborhood W of y such that q~l[W] C V. But then y 6 W C W C U since W C q[V] C U and q[V] is closed.
A. 11. THE QUOTIENT TOPOLOGY
507
So Y is regular and T\ and it consequently suffices to prove by Urysohn's Metrization Theorem mentioned on Page 465 that it has a countable base. To this end, let !B be a countable open base for X which is closed under finite unions. For every B € *B put (*)
U(B)=Y\q[X\B].
Since q is closed, U(B) is open. Now if U is an open neighborhood of y in Y then q~l [U] is an open neighborhood of the compact set q~l(y] in X. So there is a finite subcollection £ of 'B such that q~1(y} C \J £ C q~l[U}. Since !B is closed under finite unions, clearly (J £ 6 13. So there exists an element B £ !B such that Since g"1^) C B it follows that ?/ 6 U(B) C U. We conclude that the collection {U (B) : B e 13} is a countable open base for Y. D We shall now present a few important examples of upper semi-continuous decompositions. Our first example is the collection of all components of a compact space. Corollary A. 11. 3. Let X be a compact space and let G be its collection of components. Then X/G is a compact metrizable space with a base consisting of open and closed sets2 . Proof. That X/G is compact and metrizable follows from Proposition A. 10. 3 and Theorem A. 11. 2. Let q: X —$• Y = X/G be the natural quotient map. If y £ Y and U is a neighborhood of y in Y then g"1^] is a neighborhood of the component q~1(y] of X. By Proposition A. 10. 3 and Lemma A. 10.1 there is a clopen set W C X such that q~l(y] C W C q~1[U]. Since G consists of connected sets, it easily follows that q~l [q[W]\ = W. This shows that q[W] is a clopen neighborhood of y which is contained in U. D Our second example are the so-called adjunction spaces. Let X be a space with compact subspace A, and let / be a continuous surjection from A to a space B. We shall prove below that the decomposition
? = {f~l(b) : b € B} U {{x} : x G X \ A} is upper semi-continuous. The space XfP is denoted by X U/ B and is called LX attached to B by /', or the adjunction space obtained from X by attaching B to X by /. Observe that X \ A is naturally imbedded in X U/ B. So in a sense X/7* can be seen as X where A is replaced by B via the map /. 2
Spaces with such a base are called zero-dimensional.
508
A. PRELIMINARIES
Corollary A. 11. 4. Let X be a space, A C X be compact and f : A —» B be a continuous surjection. The decomposition 3> = {f~l(b) : b e B} U {{x} :xeX\A] ofX is upper semi- continuous. Hence X\J/B is a separable metrizable space. In addition, ifir-.X-^-XU/B is the natural quotient map then TT \ X \ A is a homeomorphism and ir[A] is homeomorphic to B. Remark A. 11. 5. This result shows that we may identify X\ A and 7r[X\A] as well as B and vr[A]. It will sometimes be convenient to do that. Proof. Let E C X be closed. Then
y {P e ' P : P n E ^ ® } = Eur1 [f[E n A}] is closed since f[E n A] is closed in B by Exercise A. 5. 5(1). Hence X U/ B is separable and metrizable by Theorem A. 11. 2. The second part of the corollary is a triviality. Observe that TT \X\A: X\A^
is a bijective quotient map, and hence a homeomorphism (Exercise A. 11. 2). A moments reflection shows that there is a canonical continuous bijection from B onto n[B]. So B and K[B] are homeomorphic by compactness of B (Exercise A. 5. 9). D We shall now discuss our last important example of an upper semicontinuous decomposition. A null-sequence in a space X is a sequence (An)n of closed subsets of A" such that limn_>.00 diam(An) = 0. Corollary A. 11. 6. Let X be a space, and let (An)n be a null-sequence consisting of pairwise disjoint nonempty compact subsets of X. Then the decomposition oo
= {An : n 6 N} U {{2:} : x # \J An n=l
is upper semi- continuous (and so X/7 is separable and metrizable). Proof. Let B C X be closed. We will prove that
C = B U |J {An : An n B ^ 0, n € N} is closed. If not, then there is a sequence (xn)n in C whose limit x does not belong to C (Exercise A.2.1). If infinitely many terms of the sequence belong to B then x E B since B is closed. If some An that meets B contains
A.11. THE QUOTIENT TOPOLOGY
509
infinitely many terms of the sequence then x G An since An is closed. So we may assume without loss of generality that there is a sequence of integers
m(l) < m ( 2 ) < • • • such that xn 6 Am(n) for every n. But since lim diam(A TO ( n )) = 0
n—>oo
^ '
and Am(n) n -B 7^ 0 for every n, this implies that x £ J3.
D
Exercises for §A.ll. 1. Let IP be a partition of a space into nonempty closed sets. Prove that the quotient topology on X/71 is indeed a topology. 2. Let /: X —)• Y be a bijective quotient map. Prove that / is a homeomorphism. ^•3. Prove that the quotient topology is not necessarily metrizable, even if X is a separable metrizable topological space. Demonstrate this by proving that R/N is not first countable at the point {N}. 4. Let X be a space, and let (An)n be a null-sequence consisting of pairwise disjoint nonempty compact subsets of X. In addition, for every n let /„.: An —» Bn be a continuous surjection. Prove that the decomposition
y = ifnl(y) • y e An,n e N} u {{x} •. x $ Q An] n=l
is upper semi-continuous (and so X/f is separable and metrizable). ^•5. Let X be a compact space with disjoint closed subsets A and B. In addition, let S be a closed subset of X such that every component of S misses A or B. Prove that S can be covered by a finite disjoint collection N of closed sets such that every N € N misses A or B. 6. Let X and y be compact spaces, and let /: X —> Y be a continuous surjection. Prove that / is quotient. 7. Let X and Y be compact spaces, and let /: X —>• Y be a continuous surjection. In addition, let A C X and C C F be closed sets such that f[A] = C. Finally, let I? be a compact space for which there exist continuous surjections f : A^ B,
77: B^C
such that
ftA = r)ot. Prove that if TT : JC —>• X U^ C is the natural quotient map, then there is a unique continuous function •y: X D£ B —> Y such that / = 7 O 7T.
Observe that 7 f S = 77.
510
A. PRELIMINARIES
Let X be a space with closed subspace F, and let ^X be a compactification of X. Prove that if aF is a compactification of F such that aF > F (here F is the closure of F in ^X] then there is a compactification bX of X with bX > X while moreover the closure of F in bX is aF. A. 12. Homotopies Let X and Y be spaces. A homotopy from X to F is a continuous function H: X xl-^-Y. If £ € I then the function Ht: X -*Y denned by Ht(x)=H(x,t) is called the t-level of H. One should think of a homotopy H as & continuous family of functions connecting HQ and HI . A homotopy from X to X is also called a deformation of JC. Two continuous functions f,g:X->Y are called homotopic, in symbols
if there is a homotopy # : X x I —>• F such that H0 — f and HI — g. We also say that / is homotopic to g. A continuous function that is homotopic to a constant function is said to be nullhomotopic Lemma A. 12.1. Let X and Y be spaces and let /, g,h: X —>• Y be continuous. Then (1) f^f, (2) iff ~g then g~f, (3) if / ~ g and g — h then f — h.
Proof. For (1), define H: X xl^Xby H(x,t) = f ( x ) . For (2), let H : X x I —>• Y be a homotopy such that H o = / and #1 = g. Define F: X x I -+ F by F(z,£) =#(a;,l - f ) . Then F is a homotopy with F0 = g and FI = f. For (3), let Jf, F: X x I -> y be homotopies with
Define 5: X x I -> y by c / r ^ _ / H(x,2t) ( 0 < t < 1/2), ^ ^ , t j - | ^ ^ t - l ) (1/2 < * < ! ) • Then 5 is a homotopy with SQ = f and Si = h.
D
A. 12. HOMOTOPIES
511
From the above lemma we conclude that the homotopy relation '~' is an equivalence relation in C(X,Y). If X and Y are spaces and if /: X —>• Y is continuous then [/] denotes its ^-equivalence class and [X, Y] denotes We say that [X, Y] is trivial if it contains one point only (i.e., all continuous functions from X to Y are homotopic and hence nullhomotopic) . As we will see, homotopic functions have much in common. A space X is called contractible provided that there exists a homotopy H: X x I ->• X such that H0 is the identity and HI is a constant function; the homotopy H is called a contraction of X . So a space X is contractible if and only if the identity function on X is nullhomotopic. It is clear that every contractible space is path-connected from which it follows that S° = { — 1,1} is not contractible. Path-connected spaces need not be contractible since no S n is contractible (Corollary 2.4.11). Observe that I is contractible. The homotopy ( x , t ) *-* t • x
contracts I to 0. Since products of contractible spaces are contractible (Exercise A. 12. 2), it follows that I n is contractible for every n < oo. We shall now present a simple but useful characterization of contractible spaces. Proposition A. 12. 2. Let X be a path-connected space. The following statements are equivalent: (1) X is contractible, (2) if Y is any path-connected space then [X, Y] is trivial, (3) if Y is any space then [Y, X] is trivial Proof. For (1) =^ (2), let F: X x I -» X be a contraction. If /: X ->> Y is continuous then / o F is a homotopy from X to Y that connects / with a constant function. It therefore suffices to prove that any two constant functions from X to Y are homotopic. Let cp,cq: X —>• Y be the constant functions with values p and g, respectively. Since Y is path-connected, there is a path a: I —> Y with a(0) = p and a(l) = q. Now define the required homotopy H : X x I —> Y by H(x,t) = a(t). For (2) => (3), first observe that X is contractible since [X, X] is trivial. Let F: X x I —>• X be a contraction such that F\ is the constant function with value c. If Y is any space and f : Y ^ X is continuous then H =
Fo(fxll):Yxl-^X
is a homotopy that connects / with the constant function with value c. Since (3) => (1) is a triviality, we are done.
D
512
A. PRELIMINARIES
Let X be a space with subspace A. We say that A is a retract of X provided that there is a continuous function r: X —> A such that r restricted to A is the identity on A. Such a function r is called a retraction. Lemma A.12.3. Let X be a space with subspace A. If A is a retract of X then A is a closed subset of X. Proof. Let r: X —>• A be a retraction. Consider the composition
x A A A x. This function is continuous and agrees with the identity function on X precisely on the set A. That A is closed is therefore a consequence of Exercise A.1.9. D We say that A is a neighborhood retract of X provided that there exists a neighborhood U of A in X such that A is a retract of U. Retractions are very interesting functions since they preserve many topological properties. For example, if r: X —>• A is a retraction and X has the fixed-point property, then A has the fixed-point property as well, etc. See Exercise A.12.10(2). Theorem A.12.4. Every retract of a contractible space is contractible. Proof. Let Y be a contractible space, let X be a subspace of Y, and let r: Y ->X
be a retraction. By definition, there exists a contraction H: Y x l - > y . Now define F: X x I -»• X by the formula F(x,t)=r(H(x,t)). Obviously, F is a contraction of X.
D
We shall sometimes identify R2 and the complex plane C. We will make use of the exponential function z^ez
(zeC).
As usual, if z = a + bi is a complex number then | z | = \/a2 + b2,
z = a — bi
denote its absolute value and its complex conjugate, respectively. Let X be a space. A continuous function /: X —>• S1 is called inessential if / is nullhomotopic. A continuous function that is not inessential is called essential. This divides the class of continuous functions from X to S1 into
A. 12. HOMOTOPIES
513
two classes. That for some X there are essential maps X —>• S1 will be shown in Exercise A. 12.8. A second, more technical concept is needed. We say that a function has a continuous logarithm provided that there is a continuous function
ltf x
such that / = e , i.e., f ( x ) = e ( "> for every x e X. We shall prove that for compact X a function / e C(X, S1) is inessential if and only if it has a continuous logarithm. If X is a space and /, g: X -> C are continuous then \f — g\ denotes sup{|/(x) — g(x)\ : x G X} whenever this supremum exists. Lemma A. 12. 5. Let X be a space and let f : X —)• S1 be continuous such that f[X] 7^ S1. Then f has a continuous logarithm. Proof. Choose q e M such that ei<7 £ f[X]. The function f i->- eil is a homeomorphism from the interval (q, q + 2?r) onto S1 \ {eiq}. Let be the inverse of this mapping. Define ?: X ->• IR by y?(x) = L(f(x)} . Then is clearly continuous and for every x e X. For this simply observe that if z E S1\{e*9} then z coincides lL ^). D Lemma A.12.6. Let X be a space and let f \ , f a ' . X —> S1 be continuous functions such that Then f i has a continuous logarithm if and only if fa has a continuous logarithm. Proof. Define h: X ->• S1 by h — fi/fax G X then \h(x)-l\ =
Observe that h is well-defined. If
_ |(A -fa}(x}\ _ .,,
, w ,, ^ ,
\h(x}\ MX] by assumption. It follows that the image of h is contained in the half plane
{z = a + bi : a > 0}, and consequently h is not surjective. According to Lemma A. 12.5 there is a continuous function tp: X —> M such that h — el(p. Since j\ — fa • h, the proof is complete. D
514
A. PRELIMINARIES
Lemma A.12.7. Let X be a compact space and let H: X x I -> S1 be a homotopy. Then H0 has a continuous logarithm if and only if HI has one. Proof. Since H is uniformly continuous by Exercise A.5.18, there is an n E N such that if s: t G I and |s — t\ < l/n then \H(x,t)-H(x,s)\
<1
for every x € X (in fact, much more is true). Put jj — Hj/n for 0 < j < n. Then |/j+i(z) - fj(x)\ = \H(j+1}/n(x) - H3/n(x}\ < 1, for every 0 < j < n — 1 and x G X. By applying Lemma A. 12.6 successively to the level functions fj for 0 < j < n, we arrive at the desired result. D We now come to the following result. Theorem A.12.8. Let X be compact and let f : X —> S1 be continuous. The following statements are equivalent: (1) f is inessential, (2) / has a continuous logarithm. Proof. For (1) =>• (2), let H: X x I ->• S1 be a homotopy such that H0 - f and HI is constant. Since HI has clearly a continuous logarithm we find that / = HQ has one by Lemma A. 12.7. For (2) =$• (1), suppose that / = eltf for a certain continuous function (f>: X —t R. It is easy to see that the function H: X x I —>• S1 defined
by
H(x,t) = eti(f>W
is a homotopy such that HQ is constant with value 1 and H\ — f .
D
A continuum X is unicoherent provided that whenever A and B are subcontinua of X such that A U B — X then A n B is connected. Theorem A. 12.9. Let X be a compact space. Assume that every continuous function f : S1 —>• X is nullhomotopic. Then X is unicoherent. Proof. Put
S^ = {(ar,y) G S1 : y > 0},
S1. - {(x,y) € S1 : y < 0}.
Observe that §+ and SL are arcs and that
s^ u §L = s1, §i n §L = {(1,0), (-1,0)}« s°. Assume that J£ is a continuum that is not unicoherent. Let A and B be subcontinua of X such that A\JB = X while An B is not connected. Observe that since X is a continuum we have A n B =£ 0. Write Ar\ B as the disjoint
A. 12. HOMOTOPIES
515
union of two nonempty compacta, say E and F. Define / : A n B —>• §>^_ D §L by Since S^. and §L are arcs we can extend / to continuous functions /i:A->Si_,
/2:j5^gi_
(Theorem A. 4. 6). Let g: X ->• S1 be the function /i U /2. Then g is clearly continuous and we claim that it is essential. Observe that (*)
g(A n B] = {(1, 0), (-1, 0)} = g[A] n g[B].
Striving for a contradiction, assume that g is inessential. By Theorem A. 12. 8 there consequently is a continuous function ?: X —> R such that 5 = el(p . Since A is connected, so is tp[A]. Similarly for (p[B\. Since A n B ^ 0, <£>[A] and <£>[.B] are intersecting intervals in R. As a consequence, S1 = ¥>[A] n y?[.B] is connected. By continuity of the exponential function, this implies that the set elS is connected. We will show that this leads to a contradiction. Indeed, by (*) and the fact that g = el{p we obtain c ei(
_
C ei
This shows that the connected set elS is also disconnected, being equal to the doubleton {(1,0), (-1,0)}. This is absurd . D Corollary A. 12. 10. Every contractible continuum is unicoherent. Proof. This is a trivial consequence of Theorem A. 12. 9 for if X is a contractible continuum then [X, S1] is trivial by Proposition A. 12. 2. D Remark A. 12.11. We will prove later that §n is unicoherent for every n > 2 (Exercise 3.6.7) and that no S n is contractible (Corollary 2.4.11). This proves that unicoherent continua need not be contractible. Cones. An important construction in topology is that of a cone over a topological space, which we will now describe. Let X be a space and let oo be a point not in X x [0, 1). Topologize ApO = (X x [0,l))u{oo| as follows: points of the form (x,t) have their usual product neighborhoods and a basic neighborhood of oo has the form (X x (a, 1)) U {oo},
where 0 < s < 1. We call A(X) the cone over X.
516
A. PRELIMINARIES
Observe that X x {0} is a closed subset of A(JQ, hence X can be thought of as a closed subspace of its own cone. It is easy to see that A(X) is contractible, see Exercise A. 12. 5 below. If X is compact then A(X) is the one-point compactification of X x [0, 1) and A(X) is homeomorphic to the quotient space obtained from X x I by identifying the subset X x {1} to a single point. Observe that A(X) has a countable base and therefore is a separable metrizable space (for compact spaces, this also follows from Theorem A. 11. 2.) Exercises for §A.12. A space X is called locally contractible (abbreviated LC) at x 6 X if for every neighborhood U of x in X there is a neighborhood V of x and a homotopy H: V x I —> U such that HI is the identity and HQ is constant. In addition, X is called locally contractible if AT is locally contractible at every point. A space X has the fixed-point property if every continuous function / : X —¥ X has a fixed-point, i.e., a point x € X with f(x) — x. 1. Prove Lemma A. 12.1. 2. For each n G N, let An be contractible. Prove that the product O^Li Xn is contractible. 3. For each n £ N, let Xra be a space. Prove that the following statements are equivalent: (!) nr=i X" is LC(2) Each Xn is LC and there is N G N such that Xm is contractible for every m > N. 4. Let X be locally connected. Prove that if Y is a subspace of X which is a retract of X then Y is locally connected as well. 5. Let X be a space. Prove that A(X) is contractible. 6. Let X be a space with subspace A. Prove that subspace V(A) = ( A x [0,1)) U{oo} of A(X) is homeomorphic to the cone A(A) over A. Moreover, if A is closed in X then V(A) is closed in A(X). 7. Let A be a contractible space, and let x € X be an arbitrary point. Prove that there is a homotopy H : X x I —t X such that HO is the identity and H\ is the constant function with value x. ^•8. Let n £ Z \ {0}. Prove that the mapping ipn : S1 -> S1 defined by V'n ( * ) = * "
is essential. 9. Prove that S1 is not contractible. ^•10.
(1) Prove that I has the fixed-point property.
A.13. BOREL AND SIMILAR SETS
517
(2) Let Y be a retract of X. Prove that if X has the fixed-point property then y has the fixed-point property. (3) Prove that the sin(y.j,)-continuum in the plane has the fixed-point property. . Let D = {z € C : \z\ < I}. Prove for every continuous function /: D —>• C with /[S1] C D there is a point x (E D with f(x) = x. 12. Prove that I 2 has the fixed-point property. A. 13. Borel and similar sets We will now briefly discuss Borel sets, analytic sets and sets with the property of Baire. Borel sets. Let X be a space. A Borel set in X is an element of the aalgebra of subsets of X generated by the open subsets of X. The family of all Borel subsets of X is denoted by ^>(X). It is clear that every closed subset of X is Borel, being the complement of an open set. So the smallest cr-algebra of subsets of X generated by the closed subsets of X is contained in 'B(X). It moreover contains all open sets, so it must be equal to ^(X). The conclusion is that we could also have started with the closed sets instead of the open sets in our definition of Borel set. Countable unions of closed sets are Borel, as well as countable intersections of open sets. So all ^-subsets and all 6^-subsets of X are Borel. Consider the subspace Q of E. It is an F^-set and is therefore Borel. But it is not a Gj-subset of M by Corollary A.6.7. The field in mathematics that studies among other things the fine distinctions between various Borel sets is called Descriptive Set Theory. See e.g., KuRATOWSKi and MOSTOWSKI [244], MOSCHOVAKIS [315], KECHRIS [218] and MILLER [310] for more information. We will only discuss some very basic material here. A family ft of sets is cr-additive (6-multiplicative) if for every countable subcollection S C ft we have that \J S e ft (respectively, f| § e ft). It is clear that the union and the intersection of an arbitrary family of cr-additive collections of subsets of a set is again cr-additive. Similarly for ^-multiplicative. Since y ( X ) is both <j-additive and 5-multiplicative, it follows that every collection of sets § C 'P(X) is contained in a smallest (with respect to inclusion) collection T C (X) which is cr-additive as well as ^-multiplicative. Theorem A. 13.1. Let X be a space. Then *B(X) is the smallest a-additive and 6-multiplicative collection of subsets ofX which contains all open subsets of X (respectively, all closed subsets of X). Proof. Let £ (respectively, 3~) be the smallest cr-additive and ^-multiplicative collection of subsets of X which contains all open subsets of X (respectively,
518
A. PRELIMINARIES
all closed subsets of X). Observe that 9 = [X \ F : F G J}
is also (T-additive and 6-multiplicative. Since 9 contains all open sets, this implies that £ C g. Since every open set is a countable union of closed sets (Exercise A.2.3), the collection of all open subsets of X is contained in 3~. So £ C 3". Since every closed set is a countable intersection of open sets, it follows similarly that y C £. We conclude that £ = 3". So if E 6 £ then E e 9 and hence there exists F e 3 such that E = X\F. Since F G £ this shows that X \ E 6 £. From this we conclude that £ is a er-algebra containing all the open subsets of X, hence 'B(^). On the other hand, 'B(X) is er-additive and Smultiplicative and contains all open subsets of X, hence £ C 'B(X). Since we already showed that £ = 5", this finishes the proof. D A countable intersection of F^-subsets of X is called an F^s-set. Observe that a countable intersection of FCT(5-sets is again an F^-set. The complement of an Fo-5-set is called a G^-set. It is clear that a G^-set is the union of countably many G^-subsets of X which implies that the union of countably many G^-sets is again a G^-set. Since closed subsets of X are G^-subsets (Exercise A.2.3), it follows that every F^-subset of X is a Gjcr-set. It follows similarly that every G^-subset is an F^-set. Finite intersections of G^-sets are again G§a, and similarly for finite unions of F^-sets. A space X is called an absolute G$ if it is a G^-subset of every space it is imbedded in. We define the notions absolute Fa, absolute F^ and absolute GSCT similarly. Theorem A.13.2. Let X be a space. (1) X is an absolute G$ if and only if X is topologically complete. (2) X is an absolute Fa if and only if X is a-compact. (3) X is an absolute G^ if and only if X is the union of countably many topologically complete subspaces. Proof. Observe that (1) follows directly from Theorem A.6.3. For (2), first observe that if X is a-compact then it is an absolute FCT. Conversely, use that X can be imbedded in Q (Corollary A.4.4), and observe that Q is compact and hence every Fo-subset of it is a-compact. Statement (3) can be proved by similar considerations. D
A.13. BOREL AND SIMILAR SETS
519
It is also possible to characterize the class of all absolute F^'s 'internally', but this would lead us to far. For our purposes the following result suffices. Theorem A.13.3. Let X be a space. The following statements are equivalent: (1) X is an absolute Fffs. (2) X can be imbedded in some topologically complete space as an Fffssubset. Proof. Since X can be imbedded in the topologically complete space Q (Corollary A.4.4), the implication (1) =$• (2) is obvious. For (2) =>• (1), assume that X is contained as an FCT($-subset of some topologically complete space F, as well as in a space Z. We may assume that Z is a subspace of Q. Since being an Fas-subset is inherited by subspaces, it clearly suffices to prove that X is an F^-subset of Q. By Theorem A.8.5, there are G^-subsets S of Y and T of Q, respectively, such that X C S and X C T while moreover the identity on X can be extended to a homeomorphism /: S —> T. It is clear that S \ X is a G^-subset of 5, hence f[S \ X] is a G^-subset of T. Since a G^-subset of a G^-subset is a G^-subset, it follows that f[S \ X] is a Ggo- subset of Q. Since Q \ T is an F^-subset of Q, it is also a Gsa subset of Q. So X is indeed an F^-subset of Q. D Corollary A. 13.4. If Xn is a-compact for every n then O^Li Xn is an absolute F^S. Proof. Assume that Xn C Q for every n (Corollary A.4.4). Then
is a subspace of Q00 by Exercise A. 1.13. Observe that the complement of FJ in Q°° is equal to 00
(J Qx -..xQx(Q\Xn) xQxQx--and hence is Gsa by Lemma A.6.2. So FJ is an F^-subset of Q°° and hence is an absolute Fff$ by Theorem A. 13.3. D So Q°° is an absolute Fffs. The question naturally arises whether it is an absolute Gsa as well. Theorem A.13.5. Q°° is not an absolute G§ff. Proof. Let {Ai : i e N} be a collection of topologically complete subspaces of Q°°. We claim that they do not cover Q°°, which suffices by Theorem A.13.2(3). Indeed, since Q^0 is not Baire, AI is not dense in it by
520
A. PRELIMINARIES
Exercise A. 6. 7. So there is a non-empty basic open subset U of
Q is absolute Fa but not G$, P is absolute G& but not Fff, Q30 is absolute F^s but not GS^, R°° \ Q°° is absolute GSff but not Fff5.
The property of Baire. If X is a space and A, B C X then we say that
A=* B if AAjF? is meager. Observe that if A =* 5 and B =* C then A =* C (Exercise A.13.11). This will be used without explicit reference in the remaining part of this section. Lemma A. 13. 7. Let X be a space with subsets A, B and M. (1) If M = AAB then A = MAS. (2) IfMCX is meager then A =* A U M =* A \ M. Proof. For (1), observe that
MAS = (AAS)AS = AA(SAS) = AA0 = A. For (2), simply observe that (A U M)AA C M and (A \ M)AA C M, and that M is meager. D Let C(X) be the collection of all subsets V of X for which there exist an open set U C X such that V =* U . We say that the elements of C(X) have the property of Baire3 . 3
This has nothing to do with Baire spaces. Being a Baire space is a topological notion, while having the property of Baire says something about how a subset of a topological space is placed therein.
A.13. BOREL AND SIMILAR SETS
521
Proposition A.13.8. If X is a space then C ( X ) is a cr-algebra which contains all open subsets as well as all closed subsets of X. Proof. That C(X) contains all open sets is clear. So if we prove that it is a cr-algebra then it also contains all closed subsets of X. For any A C X,FrAis nowhere dense, so IntA=*A. This implies that if U C X is open and F C X is closed then U=*U,
F=*IntF.
Claim 1. If A € C(X) then X \ A e C(X). Proof. Let A =* U for some open U C X. Then M — AAU is meager. In addition, AAU = (X\A)A(X\U)
and X\U=*Int(X\U). So we conclude from Lemma A. 13.7 that lnt(X \ U) =* X \ U = MA(X \A)=* X\A, as required.
^
Claim 2. If An 6 C ( X ) for every n then ]J™=1 An 6 C(X). Proof. For every n let An = f/ n AM n for C/n, Mn C X with Un open and Mn meager (Lemma A.13.7(1)). We will prove that OO
CX3
U An =* |J Un. n=l
n=l
First observe that OO
OO
\jAn=\J(Un\
OO
Mn) U\J(Mn\
Un).
The set \J^=1(Mn \ Un) is meager so that by Lemma A. 13.7(2), OO
OO
\jAn=*
\J(Un\Mn).
In addition, by the same lemma, OO
OO
OO
[\U nn=*\\U n n\[\ \_J \_J \ ^J n=l
M n n.
522
A. PRELIMINARIES
Moreover, oo
oo
oo
oo
( |J (Un \ M n )) A( U Un \ \J Mn) C (J Mn 71=1
n=l
n=l
n=l
is meager, which gives us what we want.
<0
This completes the proof.
D
Corollary A.13.9. If X is a space then (X} is the cr-algebra of subsets of X generated by its open subsets. D Proposition A. 13. 10. Let X be a space and A C X. The following statements are equivalent:
(1) A (2) A can be written as G U M where G is a G^ -subset of X and M is meager. (3) A can be written as F \ M, where F is an Fa-subset of X and M is meager. Proof. By Proposition A. 13.8, (2) => (1) and (3) =» (1). For (1) => (2), let U be open such that AAU is meager. Hence AAU is contained in a meager Fasubset, say F. Then G = U\F isG5 and G C A while moreover M = A \ G is meager. For (1) =>> (3), observe that X \ A 6 C(X) (Proposition A. 13. 8). Hence we can use (2) for X \ A to get what we want. D Analytic sets. Let X be a space. We say that it is analytic provided that it is a continuous image of a topologically complete space. We will show later that every topologically complete space is a continuous image of P. So an analytic space can also be defined as one which is a continuous image of IP (Corollary 1.9.11). Lemma A. 13.11. Let X be topologically complete. Then the collection of all analytic subspaces of X is a-additive and 6 -multiplicative and moreover contains all closed subsets of X. Proof. Every closed subspace of a topologically complete space is topologically complete, hence analytic for trivial reasons. Suppose that A is a countable collection of analytic subspaces of X. For every A there exists a topologically complete space YA and a continuous surjection JA '• YA —>• A. Then [J.A is a continuous image of the topological sum 0^^ YA which is topologically complete by Exercise A. 6. 6. So (J.A is analytic.
A. 13. BOREL AND SIMILAR SETS
523
The proof that f\A is analytic as well, is slightly more complicated. Consider the subspace
YA : (MA, A1 € A)(fA(yA}
= fA'(y
°f II = lI,4e.A YA- By continuity of the functions JA-, A G A, it easily follows that Z is a closed subspace of F] (cf. Exercise A. 1.9). So we conclude that Z is topologically complete since Yl is (Lemma A. 6. 2). Fix B e A The function tp: Z ^ X defined by
is well-defined and continuous since it is equal to the composition of the projection Yl ~^ YB and /#. We claim that (p maps Z onto [\A. If we can prove this, then we are done since we already observed that Z is topologically complete. So let x £ P) A be arbitrary. For every A € .A pick a point 3/^4 € YA such that /A^A] = x. Then y = (yyi)^^^ G -^ and
then for every A 6 .A we have
D
Corollary A. 13. 12. Let ^C be a topologically complete space. Then every Bore] subset of X is analytic. Proof. This is follows from Theorem A. 13.1 and Lemma A.I 3.11.
D
We finish this section with an important result.
Theorem A. 13. 13. Let X be a space with analytic subspace A. Then A belongs toC(X). Proof. We first claim that we may assume without loss of generality that X is compact. For let aX be a compactification of X (Corollary A. 4. 8), and suppose that we are able to prove that A e C(aX). Then there are subsets U and M of aX with U open and M meager such that A = [/AM. Since X is dense in aX, it follows that M n X is meager in X (Exercise A. 13. 7) and since A C X we obtain A=(UC\ X)A(M n X] G as required (Lemma A. 13. 7). We next claim that we may assume without loss of generality that A is dense in X. For if we can show that A G C(A) then there are [/, M C A such that U is relatively open and M is relatively meager such that M =
524
A. PRELIMINARIES
But this clearly implies that M is meager in X . Let U1 be an open subset of X such that U' n ~A = U. Then A= ( £ 7 ' A M ) n A G C(X) by Proposition A. 13. 8. So from now on we assume that X is compact and that A is dense in X. Let Z be Si topologically complete space which admits a continuous surjection / : Z —>• A. We will discuss one more reduction. Let It be the collection of all open subsets U of Z such that f[U] is meager in X. Then U has a countable subcollection U' such that \JUf = \JU (Corollary A. 2.3). Hence S = f[^U] is a countable union of meager sets, and hence is meager itself. Since A\SCA' = f[Z\\JU]CA and S is meager, it follows that A =* A'. In addition, Z \ \JU is closed in Z and hence is topologically complete. If V C Z' = Z \ |J U is relatively open and nonempty then /[V] is not meager. For assume otherwise, and let V C Z be open such that V n Z' = V. Then clearly f[V] is meager, being the union of the countably many meager sets
{f(unv']:Ueu'}(j{f[v}}. So V 6 U which implies that V H Z' - 0. But this contradicts the fact that V is nonempty. These considerations show that we may assume without loss of generality that for every nonempty open subset U C Z we have that f[U] is not a meager subset of X . So in particular, if U C Z is nonempty and open then the interior of f[U] in X is nonempty. Let Q be an admissible complete metric on Z which is bounded by 1 and let d be an admissible metric on X which is also bounded by 1 (Exercise A. 1.6). We claim that there is a meager Fa- subset F C X such that the dense (7<$-subset T — X \ F of X is contained in A. (Observe that T is dense by the Baire Category Theorem A. 6. 6.) To prove this, let !B be a countable open base for Z such that Z € 1$, and for every n let *Bn be an open cover of Z having the following properties: (1) £ n C - B , (2) mesh(®n) < 2~ n , (3) mesh(/[Sn]) < 2-". If B € "Bn then put
A.13. BOREL AND SIMILAR SETS
525
Observe that B C \J£(B). Claim 1. For every n and B G 25n the intersection of the open set
with f[B] is dense in f[B]. Proof. Let V C X be open such that V n /[B] 7^ 0. In addition, let W in X be open such that W C F and W n /[£] ^ 0. So /^[W] n 5 is a nonempty open subset of Z. There is an element B' G £ n +i such that So 5' G £(B) and since 0 ^ Int f[E] C Int /[B'] n Int f[B] nV C
we are done.
For every n and B G 23n , put Then D(B] is a nowhere dense closed subspace of AT by Claim 1. So oo
f=U U n=l
is a meager Fff-subset of A". Claim 2. T = X \ F C A. Proof. Pick an arbitrary element p G T. Put Ui = Z. Then p e /[t/i] since A is dense in A. By induction on n > 2 we will construct an element Un G 3n such that Un-i n f/n 7^ 0 and p G /[J/n]- Indeed, suppose that for some n > I we constructed Un. Since p $. D(Un), it follows that p G E(Un). So there exists t/n+i G £(t/ n ) such that p G f[Un+i]. But then t/n+i n Un ^ 0, which shows that t/n+i is as required. For every n > 2 pick a point 2n G UnnUn-i. By (2), the sequence (zn)n>2 is Cauchy and hence converges to an element z G Z. We claim that f ( z ) = p. Let V be an arbitrary neighborhood of p. Since linin^oo diam/[f/ n ] = 0 by (3), there is an TV G N such that f[Un] C V for all n > N. But this implies that all but finitely many terms of the sequence (f(zn))n>2 belong to V. So and since zn —>• z, by continuity of / we have f ( z ) = p.
0
We conclude that A contains a dense G^-subset of X, from which it follows that A =* X G C(X). D
526
A. PRELIMINARIES
Exercises for §A.13. A space X is called absolutely Borel if it is a Borel subset of every space it is imbedded in. Let A C X. Then A is said to be meager at the point p £ X if there is a neighborhood U of p such that U fl A is meager. The set of points of X at which A is not meager is denoted by D(A). 1. Let X be a space with subspace Y. Prove that
\Y. 2. Let Z be a space with subspaces X and y such that X C y. (1) Prove that if X e 'B(Z) then X € 3(Y). (2) Prove that if Y e B(Z) and X G B(Y) then X € ®(Z). 3. Let X be a space. Prove that the following statements are equivalent: (1) X is an absolute Borel set. (2) X can be imbedded as a Borel set in a topologically complete space y. ^4. Let X be an absolute Borel set. Prove that if X is a Baire space if and only if it contains a dense topologically complete subspace. 5.
(1) Prove that the number of G«5-subsets of Q is c. (2) Prove that the number of analytic subsets of Q is c. (3) Let X be a space. Prove that |33(X)| = c if and only if X is infinite.
6. Prove that there is a subset of R which is not Borel. 7. Let X be a dense subspace of Y . Prove that if M is meager in Y then Mr\X is meager in X. 8. Let X be a space with subspace A. Then (1) D(A) = 0 & A is meager. (2) D(A\D(A)) =0. (Hence A \ D(A) is meager.) (3) D(A) C A (4) D(A) is closed. (5) If E C X is meager then D(A) = D(A \E) = D(A U E). 9. Let X be compact and let / : X —> Y be irreducible. Prove that for A C X the following statements are equivalent: (1) A is meager, (2) f[A] is meager. Conclude that for B C Y the following statements are equivalent: (1) B is meager, (2) /^[B] is meager. 10. Prove that if X is absolutely Borel then it is analytic. 11. Prove that if A, B and C are subsets of a space X such that A=*JB, then A =* C.
B=*C
APPENDIX B
Answers to selected exercises It is of course best to try to figure out each exercise before looking for the answer here. But these answers are intended to be read, since some of them are used in the text and some of them occasionally provide additional information not covered by the text. ^•Exercise 1.1.10:
Define i: X —> C(X] by i(x)(y] = g(x,y). We will show that i is an isometry. For take arbitrary xi,x2 £ X. Then In addition, for every y G X we have \i(xi)(y) -i(x2)(y)\ = \e(xi,y) - g(x2,y)\ < e(xi,xz), from which it follows that (j(i(xi),i(x2))
< g ( x i , X 2 ) . Consequently,
which is as required. Observe that if x G X then i(x) is a function having nonnegative values only. As a consequence, a limit of functions of the form i(x) has nonnegative values only. To prove (1), let Y C X be an arbitrary subset. Assume that / 6 conv(z[F]) and that for some sequence (yn)n in Y , f = Hmn_».cxD i(yn}- We shall prove that / £ i[Y]. Since / e conv(i[K]) there exist distinct ai, . . . , am G Y and elements t\, . . . , tm € I such that
(Lemma 1.1.1). Without loss of generality we may assume that t\ ^ 0. Then for every n G N,
= /(»
Since the sequence (i(yn)}
converges to /, it therefore follows that lim yn = 01. 527
528
B. ANSWERS TO SELECTED EXERCISES
This obviously implies that / = i(ai}. (2) is a triviality. ^Exercise 1.1.13: We first prove that ip is a vector subspace of s. To this end, simply observe that since 0 < p < 1 for all x, y G R we have
x + y\p <\x\p + \y\p. This implies that if x,y G tp then 00
00
00
OO
n=l
n=l
n=l
u=l
E i*»+y-\ p < E(MP + MP) = E wp + E wp < °°.
and so x + y G F. Similarly, te G F for all £ G P and £ <E R. That £> is a metric and that the topology derived from this metric is compatible with the linear structure on lp follows from straightforward calculations which are left to the reader. We proceed to prove that ip is not locally convex. To this end, let U = B(0,1), the open ball of radius 1 about the origin of tp. We claim that 0 does not have a convex neighborhood which is contained in U. Striving for a contradiction, assume that V is such a neighborhood. There exists e > 0 such that D(0, e) C V'. Let n G N be so large that nl~pe> 1, and consider for every i < n the point x; G s having all coordinates 0 except for the z-th coordinate which equals e 1/fp . Observe that g(xi,Q) = (e 1 ' p ) p = e and so Xi G D(0, e) C V. Consider the point
Then a; is a convex combination of x\,...,xn and by convexity of V it follows that x E V (Lemma 1.1.1). But
1=1 Since V C 5(0,1), this is a contradiction. ^-Exercise 1.1.18: We prove by induction on n that ||-|| is continuous on R". WTe first claim that it suffices to prove continuity at the origin of R71. To see this, let (xn)n be a sequence in R™ converging to an element p G R n . Then p — xn -> 0, from which it follows by assumption that \\p — xn\\ —> ||0|| = 0. Since |||p|| — \\xn\\\ < \\p — xn\\ for every n, this shows that \\xn\\ —I ||p||, which is as required. Now let n = 1, and f ( x ) — ||x|| for every x G R. Put p = \\l\\. Then for every x G R we have Since p is fixed, this shows that / is continuous. Now assume the statement is true for n and consider a norm ||-|| on R n+1 . We think of Rn+1 as R" x R. Let (xi)i be a sequence in Rn+l converging to the origin of R n+1 . Write every Xi as (al,bi), where al € Rn and bi € R. Clearly a, -)• 0
B. ANSWERS TO SELECTED EXERCISES
529
and bi ->• 0. Since both ||-|| f Rn x {0} and ||-|| \ {0} x R are continuous by our inductive assumption,
INI < UK o)|| + ||(o, 6011 -» IIQH + IIQH = 0 + 0 = 0, as required. This completes the inductive proof. We claim that the metric d: R™ x R71 —>• [0, oo) defined by d ( x , y ) = \\x — y\\ is continuous. This follows easily because d is the composition of the function
defined by ip(x,y) = x — y and the norm ||-||. So by Exercise A. 5. 15 the topology generated by d on every compact ball of Rn is the Euclidean topology. But this implies that the topology on Rn generated by d is the Euclidean topology since Rn is locally compact. ^•Exercise 1.1.20: By an argument similar to the one in the first part of the solution of Exercise 1.1.18 it follows that / is continuous. In addition, -D(0, 1) is compact by the same exercise, hence so is f [ D ( 0 , 1)]. Let t be the maximum of /[-D(0, 1)]. Since / is not identically 0, there is an element y € Rn such that f ( y ) ^ 0. We may clearly assume that f ( y ) > 0 (replace if necessary y by —y). By linearity, y ^ 0 and so
We conclude that t > 0 and claim that it is as required. For this it suffices to prove that f ~ l ( t ) fl B(0, 1) = 0. Striving for a contradiction, assume that there is a vector x 6 f ~ l ( t ) D B(0, 1). Since t > 0, x ^ 0. Observe that
Since ||x|| < 1 we get which contradicts the fact that t is the maximum of /[£>((), 1)]. ^Exercise 1.1.21: We construct the en by induction on n. Let e.\ be any unit vector in V. Suppose that the vectors e i , . . . , e n - i have been constructed for some n > 2. The linear subspace W generated by ei, . . . , en-i is not equal to V by infinite-dimensionality. Take an arbitrary vector p £ V\W, and let R be the linear subspace of V generated by W and p. Then R is both topologically and algebraically isomorphic to Rn (Exercise 1.1.18). Let /: R —> R be a linear function with kernel W. By Exercise 1.1.20 there is t > 0 such that f ~ l ( t ] n B(0, 1) = 0 and /"*(£) H D(0, 1) ^ 0. Let x be an arbitrary vector in f ~ 1 ( t ) fl D(0, 1). So \\x\\ = 1 and x 0 W', i.e., {ei, . . . , e n -i, #} is linearly independent. Suppose that for some i < n — 1 we have \\x — a\\ < 1. Since f(x-el) = f ( x ) - f ( e i ) = t - 0 = t we get x - el £ /^(t) nB(0, 1) which is a contradiction. So by putting en = x we satisfy our inductive demands. ^-Exercise 1.1.24: We may assume without loss of generality that the closed ball D(0, 2) is contained in A. Suppose that XQ € R71 is such that for some b G [0, oo) we have bxo £ A.
530
B. ANSWERS TO SELECTED EXERCISES
Then by convexity of A and the fact that 0 € A it clearly follows that axo £ A for all a £ [0, 6]. By convexity of A we also have that \J{I(x,bx0) : z eS""1} C A. By Exercise 1.1.23 it consequently follows that every point of the form axo such that a £ [0, 6), belongs to the interior of A. Hence if bxo £ Fr A then (1) 0 < a < b implies that axo belongs to the interior of A, (2) b < a implies that axo does not belong to A. Now for an arbitrary XQ £ W1 put g(xo) = sup{o £ [0,oo) : axo £ A}. Observe that OXQ £ A and that A is bounded, so g(xo) exists. Also,
g(xo) • XQ £ A = A which implies that g(xo) • XQ £ A. Similarly, g(xo) • XQ £ W1 \ A. We conclude that g(xo) • XQ £ Fr A. So for every x £ S71"1 the element g ( x ) £ [0, oo) is the only element r € [0, oo) such that rx £ Fr A. If g(x) > a > 0 then ax € Int A and so there exists £ > 0 such that B(ax,s) C Int A. The set {y £ Rn : y £ B(ax,s) and ||y|| = ||ax|| = a} is open in {y£R" : \\y\\ = \\ax\\ = a}
so O = { y a y : y £ B(ax, e) and ||y|| = a} n-1
is open in S . And if z £ O then az £ B(ax,e) C Int A so g(^) > a. We conclude that p~ 1 [(a,oo)] is open. It follows similarly that g~l[(—oo,a)] is open. We conclude that g: S™"1 -> R is continuous. Define /: D(0,1) —> A by
and h: A ^ £>(0,1) by
respectively. There exists M > 0 such that 1 < g ( x ) < M for every x G S" i. So
||/(X)||<M.||X|| for every x £ D(0,1) and 11^)11 < Ikll
for every x £ A. Hence / and h are continuous at 0. Notice that if x ^ 0 then
1
x
/ x \
f
x
X i-)- T:—r? "->• T1—TT •-*• 91 71—TT M^ 91v 71—T7 x
\\\T\\ \x\\
\\T\\ ||x||
\^ll^lr T /
\
'x
\ ll\\r\\ lr )
is a composition of continuous functions, whence / is continuous at all points. By the same argument it follows that h is continuous. Finally notice that /i(/(a;)) = x
B. ANSWERS TO SELECTED EXERCISES
531
and f ( h ( y ) ] = y for every x E D(0,1) and y € A. We conclude that / and h are homeomorphisms, and /[Sn-1] = Fr A since g(xo) • XQ € Fr A for every XQ G W1. ^•Exercise 1.1.26:
We prove (1) =>• (2). The triangle inequality for ||-|| directly implies that Vx,y
< \\x-y\\-
Therefore, for every n G N,
0<
a; - \\xin]
x — xn —
Since limn_>.oo g(x,x(n)) = 0 this implies that limn_>.oo ||x(n)|| = ||o;||. That lim n ^oo x(n)i = Xi for every i G N is a triviality. We prove (2) => (1). Let e > 0. As x £ f.2 there is a p G N such that (3)
Now ||x(n)|| —> ||x|| as n —> oo and hence 3mo > 0 : n > mo => ||a;(n)||2 — ||a:||2 < e.
(4)
Finally, x(n)i —>• Xi for i = 1,... ,p — 1 as n —)• oo, and hence (5) and
:r(n)2 - x\
(6)
Then for n > max(?no, mi, m2) we have by (4) and (6), 00
00
p—1
p—1
i=l
i=l
p-l < Hn)|| a - Ikll
+
< e+e
from which it follows by (3) that 3e.
(7)
532
B. ANSWERS TO SELECTED EXERCISES
Since (a - 6)2 < 2a2 + 262 for all a, b <E R, oo
p— 1
oo
^(x(n)i - X i f — ^(x(n)i - X i f + ^(x(n)i - X i f p— 1
oo 2
< 5^(z(n)i -
Xi)
oo
+ 2 ]T x(ntf + 2 J^ z 2 . i=p
i=l
i=p
Consequently, for n > max(mo, mi, 7712) we obtain by (5), (7) and (3) that
We conclude that linin-^oo x(n) — x (in £ 2 ). ^Exercise 1.1.29: Recall that A(X) = (X x [0,1)) U{oc>, where oo £ X x [0, 1), topologized as explained on Page 515. The function cp: A(JO - + L X I defined by
(*e[o,i)),
99(00) = (0,1), is clearly a bijection. To prove that it is a homeomorphism, we only need to worry about the point oc. Let ((xn,tn)} be a sequence in X x [0, 1) converging to the point oo. It is clear that the sequence (tn)n converges to 1. Since X is bounded, there is an element e G [0, oo) such that \\x\\ < £ for every x G X. As a consequence, 0 < \\(l-tn}Xn\\
so that the sequence ((1 — tn)xn}
< (l-tn)e-»0,
converges to 0 in L. This shows that
B. ANSWERS TO SELECTED EXERCISES
533
There are several cases to consider. Suppose first that t — 0. We may assume without loss of generality that tn ^ 0 for every n. Observe that
since a[L] is a bounded subset of R. But then since X is a bounded subset of L, it follows by Exercise 1.1.12 that So from now on assume that t > 0. If x/t G X \ U C Vr then since X fi
X
77 ~^ T and V is open, there is an NI G N such that
for all n> NI. This show that
and hence Now assume that x/t 6 U. There is an ^2 € N such that
for all n > Ar2. But now the continuity of all functions involved clearly give us what we want. *> Exercise 1.6.1: Conditions (2) and (4) and Lemma A.3.1 show that h is continuous with dense image. We show that h is one-to-one and closed. Suppose that, for some e > 0 there exist sequences (xk)k and (yk)k in A" such that g ( x k , y k ) > £ for each A; and lim h(xk] = z = lim h(yk)
k—too
k—>oo
n
£
(Exercise A.6.9). Choose n such that 2~ < /5, and pick U € Un such that z £ U. Then for some k > n, {hk o • • • o /n(xfc), hk ° • • • ° h i ( y k ) } C U. By conditions (1) and (4) there exist V, W £ Un such that \hn o • • • o /ii(xfc), hk o • • • o h i ( x k ) } ^ V and
{hk o • • • o h i ( y k ) , h k o • • • o h i ( y k ) } C VF. There also exist I/', V^' € U n such that {hn-l
O • • • O / i l ( x f e ) , hn O • • • O / l l ( x f c ) } C V^'
and
Thus {V^', V7, t/, 1/F, VF'} is a chain in Ura with /i n _i o • • • o /ii(xfc) e V' and /i n _i o • • • o h\(yk) € W'.
534
B. ANSWERS TO SELECTED EXERCISES
Applying (hn-i ° • • • ° hi)~l, we obtain a 5-chain in (fcn-io.-.o/n)-1^] between the points Xk and yk- By condition (3), g ( x k , y k ) < 5 • 2~n < e, a contradiction. It follows that h is one-to-one and closed, and therefore a homeomorphism of X. ^•Exercise 1.8.2: Let q £ Q and let V be an open neighborhood of q in Q. There exists by Theorem 1.6.6 an element ip £ ^H(Q) such that V(?) = 0- So tj)[V] is a neighborhood of 0 and so there consequently exist 0 < 6 < 1 and JV E N such that
U = {x £ Q : a;, £ (-<S, 5) for i = 1, . . . , N} C V. It is clear that
U = {x £ Q : x, £ [-5, 5] for i = 1, . . . , N} is a Hilbert cube. The pseudo-interior of U shall be denoted by s(U) and the boundary of U by 9C7, i.e.,
Claim 1. For arbitrary x , y £ s(C/) there is an element / £ 3i(U) such that
• f ( x ) = y, • / \ 8U = 1. Observe such that such / can be extended to an element / £ ^K(Q) such that
f\Q\u = iQ. Proof. Let i £ {!,..., A^} and factorize U as [—5, 5]j x C/i, where Ui is the product of the remaining factors of U. Let pi be the projection from Ui onto the first N — 1 factors, and let pi be the projection from U onto f/j. For every z £ C/i put
and define the continuous function a : Ui —>• I by oa(z) — min{Ai(2), 1}. Define the isotopy gi : [— <5, 5] x I —> [— <5, 5] by requiring that for every i £ I the function gn maps the intervals [— 5, Xi] and [xi,8] linearly onto the intervals [-5, (1 - t)Xi + tyi],
[(1 - t)xi + tyi,S\.
Define the function /{ : U —> U by
Notice that fi £ J<(Z7) by Theorem 1.8.2, that fi\dU = l and that f i ( x l ) = yl. Identify U and HiLiC"" ^' $}* x Ili>Ar ^*- By a similar construction as the one above, it is clearly possible to construct an element F £ CK(f7) such that F \ B = 1 and F(x)i = yi for all i > N. Now put / = F O fN
O /jv_l O • • • O /! .
B. ANSWERS TO SELECTED EXERCISES Then / is as required.
535 0
So just as in the case of the homogeneity of Q we are done if we can prove that for every x G U there exists g G !H(U) such that g ( x } G s(U) and g \ dU = 1. (Again we extend such homeomorphism to a homeomorphism of Q by letting it restrict to the identity on Q \ U.) Take an arbitrary x G U and let uji : [—5, 5] —>• [—1,1] be the homeomorphism taking the intervals [— 5, Xi] and [xi,S] linearly onto the intervals [—1,0] and [0, 1]. Define the homeomorphism i7 : U to Q by
(i > N). Let :TO = Q(x) and
For every 0 < t < 1 put tft = {x <E Q : Xi € [-1 + t, 1 - t] for i = 1, . . . , TV}. Notice that x0 € K\. The function /3: Q -> I defined by /3(^) = sup{^ e I : z £ Kt} is obviously continuous. It therefore suffices to prove the following: Claim 2. Let z £ KI. Then there exists / e IK(Q) such that
Proof. The proof is similar to the proof of Lemma 1.6.5. Lemma 1.6.4 is not directly applicable, but instead we use the following analogous statement: Let m that (1) (2) (3) (4)
> N, z € KI and e > 0. Then there exists h e ^K(Q) such £(M)<e, fc(*) m e(-l,l), /i(tt)i = m for all u £ Q and z = l , . . . , m — 1, /i f W =
An inspection of the proof of Lemma 1.6.4 will show that this can be proved along the same lines. <$• The completes the solution to the first part of the exercise. The second part clearly follows from the first part and from Theorem 1.6.9. *• Exercise 1.10.8: Let f i : X —> X (i E {1, 2}) be continuous functions such that the function
536
B. ANSWERS TO SELECTED EXERCISES
maps X onto X2. Define the functions gn: X —> X (n G N) as follows:
9i = /i 02 = /i ° h 9s = /i ° /2 ° /2 Sn = /i ° /2 o • • • o /2,
for n G N.
n — 1 times
We claim that the function /: X —>• X00 defined by a"-* ( 0 i O ) , S 2 0 ) , . . . , 0 n ( a O , . . . ) 00
maps X onto X . Since X is compact, it suffices to show that the image of X under this map is dense in X00 (Exercise A.5.5(1)), and it turn it suffices to show that for every sequence y G X°° and every n G N there is an x G X such that /(z)» = yi for all i < n. Equivalently, we need to prove that for every n-tuple ( y i , . . . , yn) G Xn there is an x G X such that Qi(x) = yi for i < n. We prove this by induction on n. For n = 1 this is just the fact that /: is onto. Now we assume the claim is true for n and prove it for n + 1. By the induction hypothesis, find re' G X such that gi(x') = j/i+i, for 1 < i < n. Now pick £ such that /i(x) = t/i and fi(x} = x'. Then for 2 < i < n + I we have 0i (z) = S»-i(/2(oO) = 5i-i(ar') = j/i, and this completes the proof. ^•Exercise 1.11.19: Let U C X be open. We claim that e~1[{{[/})] and e~l[({U,X})] This establishes continuity of e by Lemma 1.11.11. Let (A,t) G e-1[{{t/,X})]. Pick a point
are both open.
x G e(A, t) n U = D(A, t) n U. There is e > 0 such that D(x,e) C [7. Pick a G A such that ^(x,a) < t. We claim that e[{{D(a, «/ 2 ),X}> x (t- « / 2 , < + e/2)] C <{I7,X}>. Observe that this is as required since ({D(a, £/2),X}) is a neighborhood of A and (t-%,t+%) is a neighborhood of t. Indeed, pick an arbitrary (A',t') G ({D(a, %),X}} x (t-
(1)
and £>(a, x) < g(a , a) + £>(a', x) and therefore g(a',x) > g(a,x) — g(a,a] > g(a,x) — £/2. We conclude that g(a',x) G [ g ( a , x ) - %,g(a,x) + %}. £
Case 1. g(a,x) - /2 < 0.
(2)
B. ANSWERS TO SELECTED EXERCISES
537
Then by (1), g(x,a) < Q(X,O) + £/2 = e. Hence since D(x, e) C U we have
a e A'nuc D(A',t')nu, and so e(A' , t') n U ^ 0. Case 2. £>(a,:r) - £/2 > 0. Observe that by (2) we have Q(O-',X) < (g(a,x] - %) + £ and hence, by convexity of £>, a'
£D(D(x,e),g(a,x)-%)).
So there is an element x' G D(x, e) such that g(a',x') < g(a,x] -
E
/2.
Observe that g(a,x)-%
which completes the solution to this exercise. *• Exercise 1.11.20: If k — 1 then there is nothing to prove. Simply observe that in that case we are dealing with a sequence with a convergent subsequence, in which case we are done, or with a sequence having no convergent subsequence. In the latter case, our sequence is discrete by Exercise A. 2.1, in which case we are done as well. Assume that the statement of the exercise is true for k — 1, where k > 2. For every i pick an arbitrary element Xi G Ai and let Bi = Ai\ {xi}. Let the infinite set E C N and the (possibly empty) subset B° C X be given by our induction hypothesis for the sequence (Bi]i. Assume first that every point in X has a neighborhood meeting finitely many terms of the sequence (-£?;),££ only. Wre distinguish between two subcases. If there is an infinite subset EQ C E such that the sequence ( x i ) i ^ E ( ) is convergent, say to x E X, then EQ, {x} and for every i € EQ the partition Bi, {xi} of Ai satisfy our inductive requirements. If there is no such EQ then the sequence (xi)i^E is discrete (Exercise A. 2.1) from which it follows that the sequence (Ai)i£E has the property that every x G X has a neighborhood meeting finitely terms of the sequence (Ai)i£E only. Simply observe that by assumption every x G X has a neighborhood Ux meeting finitely many terms of the sequence (Bi)i€E only- In addition, each x £ X has a neighborhood Vx meeting at most one element of the set {xi : i 6 E}. So Ux n Vx is a neighborhood of x meeting finitely many terms of
538
B. ANSWERS TO SELECTED EXERCISES
the sequence (Ai)ieE only. Suppose next that there are a nonempty subset B° of X and for every i G E a partition B®, B} of Bi such that that (1) (B^}i£E converges to B° in the Vietoris topology. (2) Every x e X has a neighborhood meeting finitely many terms of the sequence (Bi)i^E only. We argue as before. Case 3. There is an infinite subset EQ C E such that (xi)i^.E0 converges to a point x E X. We then put A° = {x} U B° and A° = {Xi} U 5°, A] = B] for every i £ E0. An easy check shows that our choices are as required (cf. Proposition 1.11.7). Case 4. (xi)i^E does not have a convergent subsequence. So the set {xi : i € E} is closed and discrete in X by Exercise A. 2.1. Now put A° = B°, A° = B° and A\ — B\ U {xi} for every i & E. As in the case k = l it follows that the sequence (Al)i^E has the property that every x £ X has a neighborhood meeting finitely many of its terms only. ^•Exercise 1.11.23:
Observe that TT is upper semi-continuous by Corollary 1.11.17. Let Y = X/8, and define/: Y ->• 2X by We claim that / is continuous. An application of Exercise 1.11.16 then does the job. To this end, let (yn)n be a sequence in Y converging to y £ Y. Put En ~ TT~l(yn) for every n, and E = 7r~1(y), respectively. Let S be an arbitrary limit point of the sequence (En)n in 2X . That is, S is the limit of a subsequence of (En)n- Observe that by compactness of 2X (Corollary 1.11.4) such limits exist. Claim 1. S C E. Proof. Striving for a contradiction, assume that there is an element x E S \ E. Then z = TT(X) ^ y, hence there is a closed neighborhood V of y which does not contain z. Since (yn)n converges to y, all but finitely many yn belong to V. So 7r~ 1 [y] is a closed subset of X which contains all but finitely many En (here we use that £ is upper semi-continuous, or, equivalently, that TT is continuous). But then each limit point of the sequence (En}n must be contained in Tr^jV] by Exercise 1.11.2. Since x is a witness of the fact that S is not contained in vr^ 1 [V], we reached the desired contradiction. •v' Since Q(X) is closed in 2X (Lemma 1.11.14), it follows that S is a subcontinuum of E. There is a subsequence of (En)n which converges to S. Since u> is continuous, this shows that u>(S) = t. Hence S cannot be a proper subcontinuum of E for otherwise t =ujS
B. ANSWERS TO SELECTED EXERCISES
539
We conclude that S = E. This proves that the sequence (En}n has only one limit point in 2A . Hence (En)n converges to E (in 2A ) and this is what we had to prove. ^•Exercise 2.1.9:
Let a — {fo, . . . , vn}, where the u's are geometrically independent. By Theorem 2.1.3 it follows that the vectors vi — t>o, • • • , vn — VQ are linearly independent in W1 . By a change of base we may assume that Vi — VQ is equal to the ^-th unit vector
of R n . Observe that a change of base is done via a linear isomorphism which consequently is a homeomorphism since linear maps are continuous (see Page 113). So the composition of the translation x !->• x — VQ and this linear isomorphism is a homeomorphism which maps \a\ onto the standard simplex : Vi < n,Xi > 0 and ^Xi < 1 j.
In addition, this homeomorphism sends \a\° onto \r ° . But for \T it is clear that its geometric interior coincides with its topological interior. So the same is true for |cr|. ^•Exercise 2.1.16:
Let U = {x G |S| : x G a G T => a G §}. We claim that [/ is open in |T|. To prove this, let r be an arbitrary element of T. If U D r = 0 then [7 D r is certainly open in r. Suppose therefore that t/Plr 7^ 0. Then r G § by the definition of [/. Suppose that y G r \ U. Then there has to be a simplex a G T such that y G a and cr 0 S. If r C
This problem is equivalent to the following set-theoretical statement: Let V be a finite set and let 9C be a chain of subsets of V . In addition, let T be an arbitrary nonempty subset ofV not belonging to 3CL){V}. Then there exists a linear order < on V such that every element 0/3C is an initial segment, while T is not an initial segment. We will solve this problem instead of the one in the text. There is a linear order < on V such that every element of OC is an initial <-segment.
540
B. ANSWERS TO SELECTED EXERCISES
If T is not an <-initial segment then we are done. If not then let t be the last element of T. Observe that t is not the last element of V for otherwise T = V. So let s be the successor of t, interchange s and £, and observe that the linear order we obtain in this way does the job for us. ^•Exercise 2.4.1: Our notation is the same as in the proof of Lemma 2.4.1. Take an arbitrary x G U and observe that
Assume without loss of generality that xn < 0 (the proof for xn > 0 is simpler). Then the equality n-l
n-l
X = (1 + £„ — E £i) ' Q + E Xi ' £i + ( — ^ n ) ' ( ~ g n ) i= l
i=l
shows that x is an affine combination of the {0, ei, . . . , e n _ i , — e n } with non-negative coefficients by (*). As a consequence,
belongs to
Then fe is continuous and we claim that it is as required. It is clear that the range of fe is contained in C and is contained in a finite dimensional linear subspace of L because each f e ( x ) is a convex combination of the points 7/1, . . . , ym. So it suffices to verify that /e does not move the points too far. This will be done in Claim 1 below. For every x € C, let Fx — {i < m : x £ Ei}. Observe that Ki(x) > 0
<^=»
i e Fx.
This shows that ^i€Fx K i ( x ) = 1. Claim 1. £(/ £ ,l) < e . Proof. Fix an arbitrary x € C. Since mesh(£) < e/2 it follows that that
B. ANSWERS TO SELECTED EXERCISES
541
for every i £ Fx. Consequently,
|| _ Je^Jll f (T}\\ — l l rx — < \ —
K -
/ ., Ki
We conclude that g ( f s , 1) < e/2 < £>
as
required.
<0>
>• Exercise 2.4.8: By Exercise A. 10. 14, W1 \ A has a unique unbounded component. Since A separates Kra , there consequently is a bounded component of W1 \ A, say [/. Pick an arbitrary element p€U. Define /3: Rn \ {p} -> S n-1 by
It is clear that (3 is well-defined and continuous. We claim that /3 \ A it cannot be extended over U U A, and hence not over Rn . Striving for a contradiction, assume that /: U U A —*• §ra-1 is a continuous extension of (3 \ A. Since both A and [/ are bounded, we can find r > 0 such that
B = B(p, r) D A U U. Let and define 5: Z?(p, r) —> S by the following formula: q(x) = , '
r-/3(x)
(x£
D(p,r)\U),
r-f(x)
Then g is well-defined and continuous since the intersection of the two closed sets D(p, r) \ U and U is contained in U \ U which is in turn contained in A by Exercise A.2.9. Let x G S be arbitrary. Then x g (/ x )\ = p +, r • ~P = x. r So g is a retraction from Z?(p, r) onto S, which contradicts the No-Retraction Theorem 2.4.10. We conclude that indeed (3 \ A: A —>• S71"1 cannot be extended over R n . So it cannot be nullhomotopic by Corollaries 1.2.13 and 1.4.3.
^Exercise 2.5.2: For (1) => (2), assume that there exists an antipodal-preserving map /: Sn —> S n-1 . By Exercise 2.5.1 we can cover S""1 by a family M of n + 2 closed sets such that no M G M contains a pair of antipodal points. Since / is antipodal preserving, the same is true for the family /~1[M]. But this contradicts (1). For (2) => (3), assume that there is an antipodal-preserving map /: S71"1 —»• S""1 which is nullhomotopic.
542
B. ANSWERS TO SELECTED EXERCISES
Identify S n-1 and the subspace S"""1 x {0} of Sn. By Theorem 1.4.2 we can extend / over the northern hemisphere of Sn to a continuous function /. The function / also gives us a function on the southern hemisphere of Sn via the formula x \-> — f(—x). Since / is antipodal preserving, these two functions agree on S71"1 and so their union is continuous and is obviously antipodal-preserving. But this contradicts (2). For (3) =>• (4), assume that /: S n -»• Rn is such that f(x) ^ f(-x) for every x <E S n . Define/: S71 ->• S71'1 by
This function is antipodal-preserving and hence so is its restriction to Sn-1. (We make the same identification as the one above.) But this restriction is nullhomotopic as it can be extended over the northern hemisphere of §". This contradicts (3). Finally, for (4) => (1), let JVC be a closed covering of Sn such that |M| = n + I and no M £ M contains an antipodal pair. Fix MO 6 M and let /M : §n —>• I for M 6 3Vt \ {Mo} be a Urysohn function such that /M \ M = 0 and /M \ — M = I . Define /: Sn —>• W1 by f(x) — (/M(X)) ^ . It suffices to observe that this function does not identify any antipodal pair of points. ^•Exercise 3.1.9:
Suppose that such component does not exist. Fix a component C of X for the time being. There exist partitions Di in C between Ai fl C and Bi n C for every i such that flier Di = 0. There consequently are by Corollary 3.1.5 for all i partitions Ei in X between Ai and Bi such that f\er Ei n C — 0. By Lemma A. 10.1, there is a clopen neighborhood Uc of C such that HiEr ^ ^ Uc — 0- This shows that the collection is inessential in UcBy compactness, finitely many of the clopen sets Uc cover X. So there is a finite clopen cover V of X such that is inessential in V for every V G V. By Exercises A. 1.3 and 3.1.8 we may assume without loss of generality that V is pairwise disjoint. For every i G F and V G V let Di(V) be a partition in V between Ai n VA and BiCiV such that for every V G V,
Since V is a finite disjoint clopen cover of X, it is easy to show that for every i € F the set %
= (J
vev is a partition in X between Ai and Bi. Since obviously flier ^ = ®i we arrived at the desired contradiction. The conclusion that every strongly infinite-dimensional compact space has a strongly infinite-dimensional component is trivial. *• Exercise 3.1.15:
Let E be Erdos' space. Then E is totally disconnected and not zero-dimensional
B. ANSWERS TO SELECTED EXERCISES
543
since if U C E is open such that
0 £ U C {x e E : \\x\\ < 1} then U is not clopen (see Example 1.5.18). This means that if S is any partition between the closed sets {0} and A = {x £ E : \\x\\ > 1} of E then 5 ^ 0 . Let S be any partition between {0} and A. We claim that S is reducible. As we observed, S ^ 0, so let x € S be arbitrary. Since E is totally disconnected, there is a clopen set C C E such that x E C but ID $• C1. We claim that T = S1 \ (7 is a partition between {0} and A. Since T is by construction a proper subset of S, this will do. Write E \ S as the disjoint union of open sets U and V such that 0 e U and A C T / . Put [/' = U\C and V" = V\JC, respectively. Then OeU' since Q g C and A C V'. Also, clearly, [/' n V = 0. So we done since E \ (U1 U V) = T. ^Exercise 3.1.16: By Theorem 3.1.3, diml n = n for every n. So the topological sum
contains for every n an essential family of n pairs of disjoint closed subsets. Hence dim A" = oo. Conversely, if X is infinite-dimensional then X contains for every n an essential family rn = {(Ai,Bi) : i < n} of pairs of disjoint closed subsets. Fix n. There clearly is a continuous function a: X —>• I n sending for every i < n the pair (Ai, Bi] into the z'-th pair of opposite faces of In. So the collection of pairs of opposite faces of ln must be essential, for otherwise rn is inessential (Exercise 3.1.11). But this clearly implies that dln is not a retract of In and so the Fixed-Point Theorem in dimension n (Exercise 2.4.2). *> Exercise 3.2.3: The assumptions imply that we may pick a point Xi (E Ui for every i such that i ^3
=>
x
i ^ XJ •
Let Vi be an open neighborhood of Xi such that Vi C Ui while moreover the collection is pairwise disjoint. There are two cases. First assume that m < n + 1. Then condition (3) is empty and so we need only to worry about (1) and (2). But it is clear how to do it. By Proposition A. 7.1 there exist closed sets d C Ui for every i such that \Ji™-i Gi = X. The problem of course is that the proposition does not guarantee that the GVs are nonempty. But that is easily fixed by putting Fi = d U Vi for every i. Assume next that m > n + 2. We will deal with the indices 1, . . . , n + 2 only. From the construction it will be clear how to proceed and to deal with all collections of n + 2 indices. By Proposition A. 7.1 there exist closed sets Gi C Ui for every i such that
544
B. ANSWERS TO SELECTED EXERCISES
For every i put
Then Bi is closed and nonempty because Vi C Bi, and is contained in Ui. In addition, the collection {Bi : i < m} covers X. For take an arbitrary x E X, and let i < m be such that x E Gi U Vi. If x 0 Uj^i ^j tnen x £ Bi. If there exists j 7^ i such that x E Vj then x £ Bj. For every i < m let VFi be an open neighborhood of Xi such that Wi C Vi and put Ai = (X\Ui)
\J\jWj. i+i
Then {(Ai,Bi) : z < n + l } i s a family of n + 1 pairs of disjoint closed subsets of X. Since dim .X < n, there exist open sets Si C X for i < n + 1 such that (1) Bi C £ C5i C X \ A i , Now consider the closed sets BI, . . . , Bn+z. For z < n + 1 put Bi = Si and let n +l
B n +2 = B n +2 \ ^J Si .
Observe that the B's are the 'improved' sets we get from Lemma 3.2.3 by considering the collections of closed sets Bra+2, BI, . . . , Bn+i and open sets Si, . . . , Sn+i. So we get by (2) that n+2
n+l
f | B , C p| FT Si = 0. i=l
(3)
i =i
Observe that ^B = {Bi , B-2 , . . . , Bn+2 , Bn+3 , Bn+4 , • • • , Bm }
2
(4)
2
covers X since IJ".^ Bi C y^ Bi. In addition, the first n + 2 members of 3 have empty intersection by (3). It is clear that Bi ^ 0 for every i < n+l. But Bn+2 is nonempty as well since Bn+i contains Wn+2 and Wn+2 n y?^1 S» = 0. It is clear how to proceed. We first use Corollary 3.2.2 to swell the collection (4) to appropriate open sets. The first n + 2 members of the new open cover have empty intersection. Then we pick another collection of n + 2 indices and proceed as above. After finitely many steps we are done. ^•Exercise 3.2.4:
Let r = {(Ai, Bi) : i < n} be an essential family of pairs of disjoint closed subsets of X. By Exercise 3.1.9 there is a component C of X such that the collection { ( A t n C , B ; n C ) : i < n} is essential in C. This prove that dim C > n. That dim C < n is a consequence of the Subspace Theorem 3.2.9. ^•Exercise 3.2.7:
The identity from E C £2 to E C Q°° C R°° is continuous by Lemma 1.1.12. Let
B. ANSWERS TO SELECTED EXERCISES
545
us denote this function by /. For every e > 0, consider the open ball 5(Q,e) = {zG E: \\x\\ < e}. Its closure in E is the closed ball D(Q,e) = {x G E : \\x\\ < e}. We claim that f[D] is a nowhere dense closed subset of Q°° , where D = D(0, e). To this end, let q e Q°° \ f[D}. Then \\q\\ > e, which implies that there exists N e N such that N V^ «2
There clearly exists 6 > 0 such that if \%/i — qi\ < 6 for every i < N then
\s So
is a neighborhood of q in R°° which misses f[D]. So we conclude that f[D] is closed in Q°° . It is clearly nowhere dense since every basic open subset of Q°° depends on finitely many coordinates only and hence contains elements of arbitrarily large norm. Since °° is zero-dimensional, the set Q°° \ f[D] can be covered by (countably many) clopen sets. So by continuity of / we get that in E C f? the complement of the closed ball D(Q, e) is the union of clopen subsets of E. So 0 has a neighborhood base of the required type. But since E is a topological group it is homogeneous, so the same phenomenon is true at any point. *• Exercise 3.3.4:
Observe that C(C,X) is topologically complete by Corollary 1.3.7. Since S(C,X) is closed in C(C,X) by Proposition 1.3.8 and nonempty by Theorem 1.5.10, it is a Baire space by Theorem A. 6. 6. For every k G N put such that i ^ j implies g(xi,Xj) > l/k)}We first claim that Gk is closed. Suppose that (fm)m is a sequence in Cfc with limit / G S(C, X). For every m there exists xm G X and a subset Fm C fml(xm) of cardinality n + 2, each two distinct points of which have distance at least l/k- By compactness of X we may assume without loss of generality that there is an x G X such that xm —> x (Theorem A. 5.1), and by compactness of 2X (Corollary 1.11.4) we may assume for the same reason that without loss of generality there exists an F such that Fm —> F. Since yn+2(X) is closed in 2 X (Exercise 1.11.3), it follows that \F\ < n + 2. Let e > 0 be so small that the collection 23 = {B(x,e) : x G F} is pairwise disjoint. We may assume without loss of generality that £ < l/^k- By the definition of the topology of 2X there exists m G N such that Fm C (J 23
546
B. ANSWERS TO SELECTED EXERCISES
and Fm 0 5 ^ 0 for every B £ 23. But every B € 23 has diameter less than l/k and hence contains at most one point of Fm • So 13 must have cardinality at least n + 2 which shows that |F| = n + 2. This argument also gives us that distinct points of F are at least Yfc apart. Since xm —> x, fm —>• /, Fm —>• F and Fm C /m 1 (x m ) for every m, we get F C f~l(x), i.e., / G Gfc. We next claim that Cfc is nowhere dense in S(C,X). To this end, let / €E S(C, X) be arbitrary and let e > 0. Let £ = {£Ji, . . . , £?m} be a clopen partition of C of mesh less than l/k such that diam(/[£Jj]) < £ for every i < m. For every i < m let V; be an open neighborhood of /[#i] of diameter less than e. By Exercise 3.2.3 there exist for i < m nonempty closed sets Bi C Vi such that UI*Li Bi = X and for any choice of indices i\ < • • • < in+2 < m we have H™=i -^b ~ ®- For every i < m let gi\ Ei -+ Bi be a continuous surjection (Theorem 1.5.10). The union g of the functions gi is a continuous surjection from C onto X. If x £ C, say x e Ei, then both f ( x ) and g(x) belong to Vi. So g ( f ( x ) , g ( x ) J < e from which it follows that g ( f , g ) < £ (cf. Exercise A. 5. 4). It remains to verify that g 0 Cfe. Striving for a contradiction, assume to the contrary that g E C^. Then there exist x G X and a subset F of g~l(x) of size n + 2 that witness this. Since the mesh of £ is less than Y/t, every £?; can contain at most one element of F. So F is covered by precisely n + 2 elements of £, say Eil , . . . , Ein+2 . But then n+2
which is a contradiction since B's with n + 2 different indices have empty intersection. By the Baire Category Theorem there consequently exists an element
If some fiber £ l(x) contains the pairwise distinct points pi, . . . ,pn+2 then for k G N so large that we have £ e Cfc. So this is impossible and we conclude that £ is at most (n + l)-toone. ^•Exercise 3.3.7: Let X be an (n + l)-dimensional compactum with essential family {(Ai,Bi) :i
B. ANSWERS TO SELECTED EXERCISES
547
Lavrentieff Theorem A.8.5 implies that the identity i: YC-S-^-YCT can be extended to a homeomorphism over G'.s-subsets T1 of T and 5" of 5, respectively. This shows that we are done, once we establish the following result. Claim 1. If T' C T be a GVsubset containing Y. Then there exists t £ P such that {y} x X C T'. Proof. For every n let Un C T be open such that T' = f£Li ^«. Fix n e N and let q £ Q be arbitrary. By Exercise A.5.8 the projection TT: T —> M is closed. There consequently exists by Exercise A.5.5 an open neighborhood Vq of q such that Tr" 1 ^] = Vq x X C Un- The union l/n of the collection {Vq : q £ Q} is a open neighborhood of Q such that Vn x X C Un. The intersection V = CC=i v^ is a G<$-subset of R such that V x X C T'. There exists £ E V \ Q by (Corollary A.6.7). Such a t is clearly as required. <0> For (2), observe that dim Y < n (Theorem 3.4.12). As a consequence, Y has a compactification 7!" such that dim7y < n (Corollary 3.3.8). So dim^y \ Y) < n by the Subspace Theorem 3.2.9. This implies that defy < n. However, defy > n as well since if aY is an arbitrary compactification of Y then aY \ Y contains a copy of X by (1) and hence has to be at least n-dimensional again by the Subspace Theorem 3.2.9. *> Exercise 3.5.2: Let .
v
AI
•<
fl
—
v
.
A2 ^
f-2
v
A3
be an inverse system consisting of at most m-dimensional spaces. By Corollary 3.3.8 the space X\ has an at most m-dimensional compactification a\X\. By Exercise 3.5.1 there exists an ra-dimensional compactification 02^2 such that /i can be extended to a continuous function /i: 0,2^2 —>• a\X\. Etc. This yields an inverse sequence v
CiAi
. i
fl
v . 02^2 i
/2
. v G3A_3 <
/3
••• ,
of at most m-dimensional compact spaces the inverse limit of which is at most mdimensional by Exercise 3.2.2. But the inverse limit of the original sequence is a subspace of that inverse limit. So we are done by the Subspace Theorem 3.2.9. ^•Exercise 3.5.4: As in the solution to Exercise 3.1.15, we start with Erdos' space E. Since E admits an injective function into the zero-dimensional space Q°° (see Example 1.5.18), there is a countable family Q of clopen subsets of E such that if x and y are any two distinct points of E then for some C G C we have x G C C E \ {y}. If U C E is open such that
0 C U C {x e E : \\x\\ < 1} then U is not clopen (see Example 1.5.18). This means that no partition between the closed sets {0} and A = {x G E : \\x\\ > 1} o f j E i s empty. Now let X = ^E be a compactification of E such that for every C E G the sets C and E \ C have disjoint closures in ^E (Theorem 3.5.1). This means that C is clopen in ^E for every C G C. We claim that X is as required. To this end, let let S be any partition between {0} and A. Then S is reducible.
548
B. ANSWERS TO SELECTED EXERCISES
Since S n E is a partition between {0} and A in E we have S H E ^ 0. Pick an arbitrary x £ S C\ E. There is an element C G C such that a; € C \ {0}. Now as in the solution to Exercise 3.1.15 it follows that T = 5 \ C is a partition between {0} and A. ^•Exercise 3.5.6:
This is basically a triviality. We use Exercise 3.5.5. Let Q and p be admissible metrics for X and Y, respectively. We endow A" x Y with the admissible metric d((x, y ) , (a, b)} = max{^>(a;, a), p ( y , b)}
(cf. Exercise A. 5. 16). Let e > 0. Let U be an open neighborhood of A in X such that every continuum in U has ^-diameter less than e. Pick V similarly for B. Now if C is a continuum in U x V then vrifC 1 ], where TTI : X x Y —>• X is the projection, is a continuum in [7, hence has ^-diameter at most e. It follows similarly that the projection of C into Y has p-diameter less than e. So C has obviously d-diameter less than e. *• Exercise 3.6.1:
By induction on n > 0 we shall prove that every continuous function from F\ into S" x M, where M is any AR, is continuously extendable over X. By taking
M = {pt} we get the statement in the exercise. Consider the case n = 0. So the collection {F; : i G N} is pairwise disjoint. Let
/ : Fi -» S° x M be continuous, where M is an AR. Assume that the closed set A = rl(-l) x M CFi
is nonempty. Let X be the space we obtain from X by identifying A to a single point a, i.e., X is the space X/A, let q: X —> X be the decomposition map and let C be the component of a in X. Then C is a continuum which admits the pairwise disjoint, closed covering {{a}, q[B] n C} U {q[Fi\ H C : i > 2}, where S = /^[{l} x M]. By Theorem A. 10.6 we have C = {a}. So there is by Proposition A. 10. 3 a clopen subset O of X with a E O and O fl S = 0. Because M is an AR we can find continuous functions
and
such that 0i f A = / \ A and p2 \ B = f \ B. Then gi U y2 : X ->• S° x M is as required. Assume now that we have what we want for n. So let X be a compact space, let {Fi : i G N}
B. ANSWERS TO SELECTED EXERCISES
549
be a closed cover of X such that dim(Fi D Fj) < n for all distinct indices i and j, let M be an AR and let f : F\ -* §"+1 x M be continuous. By the Countable Closed Sum Theorem 3.2.8 we have that B = [_){Fi n Fj : i, j G N and i ^ j} has dimension < n. Let x\ and xi be distinct points in S™ +1 and note that S r a + 1 \{xi,X2} « S " x R . By Corollary 3.3.10 there is a closed covering {Hi,Hz} of X such that
fiJ-n/-1[{xJ-}xM]
=0
for j = 1, 2 and dim(//i n #2 H 5) < n. Consider the compact space X' = #i n Hz and its closed covering {F; n X' : i £ N}. Then dim(Fi n Fj n X') < dim(B n A") < n for i 7^ j. Observe that / |~ FI n X' is a continuous function into (S n+1 \{zi,:r 2 }) x Af, n which is homeomorphic to S x R x M. Since R x M is an AR, being a product of AR's, we may apply our induction hypothesis to find a continuous function g: X ' - + ( S n + 1 \ {xi,x2}) x M w i t h ^ f F i n A " = / f F i n X ' . Observing that S n+1 \{xj} is homeomorphic to R n+1 , we may select continuous functions hj:Hj^(Sn+1\{Xj})xM,
j = l,2,
with /ij f A' = 0 and fy fF,- nFi = / \Hj nFi (Corollary 1.2.12). Then /» = /n U/i 2 is a continuous mapping from X into S ra+1 x M which extends /. So we are done. ^Exercise 3.6.4:
We prove this by induction on k + n. Assume first that k + n = 0, i.e., k = n = 0. Then / can be extended to a continuous function /: X —> S° by Theorem 3.6.3. So we put B = 0 in this case. Assume that we are done for all 0 < k < n, where k + n is less than or equal to m > 0. Take 0 < k < n such that k + n < ra + 1, and let A, X and / be as in the formulation of the exercise. Suppose first that k = 0. Put E = f ~ l ( — l) and F = /~ 1 (1), respectively. By Lemma 3.6.1 there is a partition B between E and F such that dim B < n — 1. Since X \ B is the disjoint union of two open sets one of which contains E and the other one contains F, it is clear that / can be extended to a continuous function /: X \ B —> S°. This completes the case that k = 0 Suppose next that k > 0. We argue as in the proof of Theorem 3.6.3. Define §i = {x£§k:xi >0},
§2 = {x € S fc : Xi < 0},
respectively. Observe that Si and S 2 are both homeomorphic to I fc and that SiPlS^ is homeomorphic to Sfc-1. Now put AI = ^[Si] and A2 — ^[S^j. Let X\ and X-2 be such as in Corollary 3.6.2. By our inductive assumption, there is a closed subset B of X\ n X-2 of dimension at most
550
B. ANSWERS TO SELECTED EXERCISES
and which moreover misses A\ fl AI so that we can extend g \ A\ r\A<2 to a continuous function h: (Xi n X2) \ B -> S" n SJ. Now for i = l,2 define 5j :
A,U[(X1nX2)\JB]-^Sr
by Qi = (g \ Ai) U h. Since S* is homeomorphic to I fe , we can extend Qi, i = 1, 2, to a continuous function gi: Xi \ B —> Sf (Corollary 1.2.12). Clearly, g = gl U g2 : X \ S ->• S fc
is a continuous extension of g. Since B is closed and has the correct dimension, we are done. ^Exercise 3.7.4: Striving for a contradiction, suppose that every continuum in L \ (J N misses one of the sets An2 and B n2 . Since N is finite and disjoint, there is a finite disjoint collection N' of closed subsets of X of such that (J N is contained in the interior W of U N' while moreover every continuum in (J N' misses A ni or Sni (Exercise 3.7.3). Each of the sets in the collection N' can be split into finitely many disjoint closed subsets each of which misses misses An2 or Bn2 (Exercise A. 11.5). All these sets together form the collection Ni. Since by assumption every continuum in L \ (J N misses misses Anz or B n2 , the same is true for L\W. Hence by the same argument as the one above, L\W can be covered by a finite disjoint collection !N2 of closed sets such that every N € K2 misses An2 or BU2 (Exercise A.11.5). For i = 1,2 let Si be a closed set in X separating the disjoint closed sets
d = Ani U \J{N e Xt : TV n Bni = 0}, D, = Bnt U \J{N Z^-.N^Bn^ 0}. Then clearly Si n 82 H L = 0. But this is a contradiction since the collection {(A,, 50 : i e r ( 0 ) U { m , n 2 } } is essential by Exercise 3.1.1. ^Exercise 3.7.6: Since dim X < 1 there is a finite closed cover y of X such that (1) mesh(T) < e. (2) ord(J) < 1. Let X be the collection of all intersections of precisely two distinct elements from 9". By (2), N is pairwise disjoint, and by (1), mesh(N) < e. For every F G 2F let
U(F) = F \ \J{F' G J : F' ^ F}. Then U(F) is open since if is finite and the collection It = {U(F} : F € J}
is pairwise disjoint. Let x G X \ IJN be arbitrary. Then there exists F G 3" containing F. But then for no F' e 7\ {F} we have x e F' for otherwise x £ IJNWe conclude that x e U(F). Now let C C X \ U N be any continuum. Then C is covered by the disjoint collection U and therefore has to be contained in an element of U and hence in an element of 9". We conclude that diam U < e, which is as required.
B. ANSWERS TO SELECTED EXERCISES
551
*• Exercise 3.10.2: If the empty set is a partition between A and B then we are clearly done. So assume that this is not the case. Let {Bn : n e N} be a countable base for X. Recursively, we will construct closed subsets Tn C S, as follows. Put To = S. If Tn has been defined, we perform the following test. If Tn \ Bn cuts between A and B then we put Tn+i --= Tn \ Bn, otherwise we put Tn+i = Tn. Then (Tn)n is a decreasing intersection of nonempty cuts between A and B. Put T = n^°=i ^™- If C* is a continuum in X meeting both A and B but missing T then the sequence (Tn H C)n is a decreasing sequence of closed subsets of the compact space C with empty intersection. There consequently is an TV £ N such that TN n C = 0 (Exercise A.5.17). But this contradicts the fact that TN cuts between A and B. By Lemma 3.10.2 it follows that T is a partition between A and B. If T' C. T is a proper closed set which is also a partition between A and B then there exists an n € N such that Bn D T 7^ 0 and Bn C\T' = 0. At stage n of our construction we looked at the set E = Tn \ Bn. If E is a, cut between A and B then T n +i misses Bn and hence T misses Bn which is a contradiction. On the other hand, if E is not a cut between A and B then there is a continuum C between A and 5 such that C n E = 0. But then
C n T' c c n Tn c c n sra c x \ T'. So C n T' = 0 which contradicts the fact that T1' is a cut between ^4 and B. ^-Exercise 3.10.3: Let us begin by making a few general remarks that will be used without explicit reference throughout the remaining part of the solution to this exercise. First, X is locally connected and topologically complete (Exercise A.5.1), so the notions of partition and cut agree in X (Lemma 3.10.2). Second, the local connectivity of X implies that components of open sets are open (Exercise A.2.8). This implies that if E C X is closed and C is an arbitrary component of X \ E then C U E is closed (Exercise A.2.9). Striving for a contradiction, assume that S is not connected. Write S as the disjoint union of the nonempty closed sets E and F. Since F is a proper closed subset of S, it is not a partition between A and B. As a consequence, there is a continuum C in X \ F intersecting both A and B. Let U be the component of X \ F which contains C. Then U is open and contains A U B since A U C U B is a continuum contained in X\ F. Observe that the sets UUF and E are closed. So if E' = UC\E then E' C (U U F) n E = U n E = E' C E'. We conclude that E' is a clopen subset of E and hence of S. So E' is a proper closed subset of S. By irreducibility of S, it is not a partition between A and B and hence not a cut. Let V be the collection of all components of X \ E'. As above, we can find an element V £ V containing A U B. Let V' be a component of X \ E' different from V. Then V' U E1 is closed and hence Fr(V') n £' ^ 0 since X is connected. Hence U (JV' is connected since £"' C U. We conclude that K = U\J\J{V
£V:V'^V}
is connected. In addition, K (JV = X since E' C U C K. Claim 1. tf n F n S = 0.
552
B. ANSWERS TO SELECTED EXERCISES
Proof. If x € K C\V n S then since components are disjoint we have x € U D V. Since E' C (7 and V n £' = 0 it follows that z <E F U (£ \ £'). But C7n (FU(E\E'))
= 0.
This is a contradiction.
<0
Since A U B C K and A (J B C V there are by connectivity of K and I/ and Exercise 1.5.9 subcontinua A' and B' of X such that
A U B C A' C K,
A U B C B' C I/,
A' U B' = X.
By Claim 1, A' n B' n S = 0. Since AUB C A' C\B' and S1 is a partition between A and B this proves that the set A fl B' is disconnected, contradicting the fact that X is unicoherent. ^•Exercise 3.10.5: We may assume without loss of generality that the spaces Xi are compact (Corollary 3.3.8). Assume that /i=(/ii,fc2):Sn->X x Y is an imbedding, where X and Y are compact and dimX < 1. We shall show that dim Y > n. This clearly suffices since the product of at most n — 1 at most one-dimensional spaces is at most (n — l)-dimensional (Theorem 3.4.12). Consider the compactum /ii[Sn]. If it is a singleton then S" clearly imbeds in Y. So then dim Y > n by the Subspace Theorem 3.2.9 since dimS n = n (Theorem 3.2.12). So we assume without loss of generality that /ii[Sn] is not a singleton. Since X\ is at most one-dimensional, there is an at most zero-dimensional closed subset A C X\ which separates hi [Sn]. Then h^1 [A] separates S". Since by Exercise 3.6.7 Sra is unicoherent, there is a continuum C C /ij^fA] which also separates S™ (Exercises 3.10.2 and 3.10.3). By Exercise 2.4.9 there is a continuous function /: C —> S"-1 which cannot be extended over S n . Since hi[C] is a connected subset of the at most zero-dimensional set A, it is a singleton and hence h-2 \ C: C —> Y is an imbedding. Put D = /i2[C]. Let u: D —> C be the inverse of the imbedding h? \ C. Because the function f = (fou)o(h2 t C ) : C ^§n~l cannot be extended over Sn, the map / o u: D —> S™" 1 cannot be extended over Y. But this implies that dimY > n (Theorem 3.6.5). ^•Exercise 3.13.3: Write G = Di^i Ui, where Ui is open for every i. For every n let Kn = [—l/n, l/n]. Since
n K~={°>' 00
by compactness of the set K]° it follows that for some index m G N we have that K^ C Ui. Observe that
Kni x {0} C U2. So again by compactness, there is an integer n^ > n\ such that Kni
X Kn2 X Kn2 X • • • C t/2-
B. ANSWERS TO SELECTED EXERCISES
553
Observe that
Kni x Kn2 x {0} C U3. So by the same argumentation there is an integer 713 > n-2 such that . . . , etc. At the end of our process we obtain that 00
OO
{=1
t=l
which proves our claim. ^Exercise 3.13.5: Let /: C —> Q be a continuous surjection (Theorem 1.5.10) and define
u: Q-> C C R by u(q) = max/~ (q). Claim 1. it is upper semi-continuous. Proof. Let r 6 R and q G it~ 1 [(—oo,r)]. Then u(q) < r and so f ~ l ( q ] C ( —oo,r). By compactness of C the function / is closed (Exercise A.5.5). There consequently exists a neighborhood V of q in Q such that /~ X [V] C (—oo, r) (Exercise A.1.15). Now for every p £ V we have u(p) < r. We conclude that u -1 [( —oo,r)] is open. O Claim 2. it is not countably continuous. Proof. Let Q = E\(J E%\J • • • . Since Q is not the union of countably many zerodimensional subspaces (Corollary 3.13.6), for some i we have dimEi > 0. We claim that u \ Ei is not continuous. Observe that the composition
is the identity on Ei and that / is continuous. But then if u I Ei were continuous this would imply that u \ Ei: Ei —>• u[Ei\ C C is an imbedding. But this is a contradiction since dim Ei > 0 but dim C = 0. •(> So we are done. ^•Exercise 4.1.5: Let 11 be the collection of all open subsets of X of diameter less than e. By Theorem 4.1.1 there is an open refinement V of U such that for every space V, any two V-close mappings f , g : Y —>• X are It-homotopic. Let 8 > 0 be a Lebesgue number for V (Lemma A.5.3). We claim that this 6 works. To this end, let Y be a space, A C Y be closed and f , g : A -> X continuous functions with £ ( / , g ) < 5. Assume moreover that / can be extended to a continuous function /: Y —> X. First observe that / and g are V-close, so they are U-homotopic, say by the 11homotopy H. By the 'controlled' Borsuk Homotopy Extension Theorem 4.1.3 there is a homotopy H: Y x I —>• X which is limited by U while moreover HO = f and H I (A x I) = H. Clearly HQ and H\ are U-close, so g(H0,Hi) <e. Since HO = f and HI \ A = g, we conclude that g = H\ is as required.
554
B. ANSWERS TO SELECTED EXERCISES
^Exercise 4.3.3:
Assume that X \ {x} is an ANR. We aim at applying the characterization Theorem 4.2.1. To that end, let U be an open cover of X. Fix U G U such that x G U. There exists a homotopically trivial neighborhood A of x such that A C U. Now let B be an open neighborhood of x with B C A. There exists an open refinement Ho of IX such that for every Uo € Ho with UQ n B ^ 0, we have Uo C A. Since X \ {x} is an ANR, there exists an open cover V of X \ {x} such that for every locally finite simplicial complex 7 and every subcomplex S of 7 containing all the vertices of 7, every partial realization of 7 in X \ {x} relative to (S, V) can be extended to a full realization of 7 in A" \ {x} relative to Uo n (X \ {x}). Put "W = V U {B}. Then "W is an open cover of A^. Now let T be a locally finite simplicial complex, let S be a subcomplex of 7 containing all the vertices of T, and let /: |S| —>• X be a partial realization of T in X relative to (S,W). Define To = {a e 7 : (3 V € V)(f[a n |S|] C V)},
TI = 7 \ T0.
In addition, put So = S fl TO. It is clear that To is a subcomplex of T and that So contains all the vertices of TO. Also, g = f \ |So| is a partial realization of TO relative to (So, V). Consequently, g can be extended to a full realization g of To in X \ {x} relative to U0 f (X x [0, 1)). Now put Si = S U To and define /i : |Si | ->• X by
It is clear that fi is well-defined and continuous. By a standard argument, by induction on n > 0, we shall construct a continuous function
having the following properties: (1) fi = f\ and for n > 1, fn+i extends / n , (2) for every a G T(re) n Ti, / n +i[a] C A. Observe that since S contains all the vertices of T, our choice f i = fi is possible. Now assume that for certain n > 0, the function / n +i has been constructed. Naturally, we shall construct fn+2 'simplex- wise'. To this end, take an arbitrary simplex
Proof. Let r be a proper face of a. There are two cases to consider. First assume that r G TO. Then there exists an element
u e Uo n (x x [o, i)) such that / n +i[r] = fi[r] C U. Since, as was just observed, fi maps every vertex of r in B, we conclude that U n B ^ 0. Consequently, by the special choice of Uo
B. ANSWERS TO SELECTED EXERCISES
555
we obtain, Jn+\[r] C A. Second, assume that r 0 To. Then by our inductive hypothesis (2) we get which is again as required.
<^
Now since A is homotopically trivial, there exists a continuous extension fa- : o —>• A of / n +i \ da (Lemma 4.2.13). Define / n+2 : §1 U T (n+1) | -> A" by (x 6 a e Sa U T (n+1) ).
/ n+2 (aO = f a ( x )
An application of Lemma 2.1.19 establishes the required continuity of /n+2- This completes the inductive construction. Now define /: |T| ->• A" by
Applying Lemma 2.1.19 again gives us that / is a full realization of T in A (A") relative to U. Since / clearly extends /, we are done. ^•Exercise 5.1.6: Let P be a polyhedron, e > 0 and / G C(P, Q) be arbitrary. Let U be a finite open cover of Q consisting of convex sets of diameter less than e. Let eo > 0 be a Lebesgue number for U (Lemma A. 5. 3). In addition, let 5 > 0 be such that if A C P and diam.4 < 6 then diam/[A] < £Q. Here we use that / is uniformly continuous (Exercise A. 5. 18). Let T be a triangulation of P such that every a <E T has diameter less than 8 (Theorem 2.2.10). We will now proceed by induction on the dimension of P. If dim P — 0 then P is finite and we obviously get what we want since A is nowhere dense (Exercise 5.1.5). Now assume that we have what we want for all polyhedra which are n-dimensional, where n > 0, and assume that dim P = n + 1. Let R be the union of all at most ndimensional simplices of T. Then R is an at most n-dimensional subpolyhedron of P. For every a € T there exists by construction an element Ua G 11 such that /[a] C Ua. Since Ua is open and f[a] is compact, (Exercise A. 5. 3). Since T is finite, 7 = min{7CT : cr G T} > 0.
We can apply our inductive hypothesis to R and the function / \ R to get a continuous function a : R —> Q \ A such that g(a,f\R)<^.
(1)
Now pick an arbitrary (n + l)-dimensional simplex in T. Since f[a] C Ua it follows from (1) that a[da] CU
o G T.
556
B. ANSWERS TO SELECTED EXERCISES
Then g is continuous by Lemma 2.1.19. It clearly suffices to prove that £(/, g) < e. To this end, let x G P be arbitrary and let a G T be a simplex with maximal dimension which contains x. Then both f[a] and g [a] are contained in Ua and hence @ ( f ( x ) , g ( x ) } < e since diamt/cr < e. ^Exercise 5.1.7: Let A C Q be an A-set. In addition, let e > 0. For every n we identify J n and the subspace Jnx{(0,0,...)} of Q. For a sufficiently large n the projection TT: Q —» J n moves the points less than 1/2£- Since J n is a polyhedron, there exists by Exercise 5.1.6 a continuous function /3:JT->Q\A which moves the points less than y2£. So the function /3 o TT : Q —>• Q \ A moves the points less than e. Now apply Lemma 5.1.2(7). Conversely, let A G £(X), and let J7 C Q be nonempty and homotopically trivial. Fix n > I and let /: §™-1 —>• C7 \ A be continuous. As in the solution to Exercise 5.1.6 it follows that U is an AR. So we can extend / to a continuous function /: Bn ->• U. Let 5 = g ( f [ B n ] , Q \ U). Since A is a Z-set, we may by Exercise 5.1.3 approximate / by a continuous function g: Bn —» Q \ A such that £(, /) < 5. Observe that this implies that g[Bn] C U \ A. There are two function from S71"1 to U \ A, namely, / and g \ S™" 1 . Since U \ A is an A N R , being an open subspace of an AR (Corollary 4.3.2), it follows that we can choose the approximation g so close to / that the functions / and g \ S™"1 are homotopic (Theorem 4.1.1). By the Borsuk Homotopy Extension Theorem 1.4.2, it therefore follows that we may extend / to a continuous function F : Bn —> U \ A. ^•Exercise 5.2.1: Let ei be such that £(/, 1^;) < e\ < e and let n G N be so large that
_ £i+5 Let K = E U F. Then K is a compact subset of s. Define an imbedding 0: K —>• s by the formula 0(oj) = (zi,. . . , # „ , 0,xi, x 2 , . . •)• It is clear that g(^lK}<
f; 2 - i - 2 = 2- n+1 . i=n+l
So by Theorem 5.3.3 there is an element 0 e 3t(Q) such that 4> = 4> and
For 0 < t < 1 let Ot '• J —> J be the unique homeomorphism that takes the intervals [— 1, 0] and [0, 1] linearly onto the intervals [— 1, Y2t] and [^t, 1], respectively. Define the isotopy G : Q x I ->• Q by
1).
B. ANSWERS TO SELECTED EXERCISES
557
It is clear that for all 0 < t < 1. Finally, let G" : Q x I ->• Q be the isotopy G't = G~l for 0 < t < I. Put A — 4>[E] and B = d o <j)[F]. Then A and B are compact disjoint subsets of s. This is clear since 7rn+i[yl] C {0} and 7rn+i[-B] C {Y2}. The functions a: E —>• Q and ft: F ^ Q defined by a (a:) = (V 2 aJi + l/2f(x)i, . . . , %xn + V 2 /(:r) n , 0, zi, z 2 , • • • ),
are well-defined imbeddings. Observe that a[E] and /3[F] are disjoint compact subsets of s. Now define a function h: A U B —»• s by
It is easy to see that h \ A and /t \ B are imbeddings. Moreover, h[A] = a[E],
h[B] = /3[B]
from which it follows that h is an imbedding as well. We claim that since A = {(xi,. . . ,xn,Q,xi,X2,. • •) '• x £ E}, B = {(xi,. ..,xn, l/2,xi,X2,.. . ) : x £ F}, it follows from the definitions of a and (3 and
^(/,I£) = ^ ( / ~ I , I F ) < £ I that £(/> \ A, 1A) < y 2ei , £(fe f 5, 1 B ) < y 2 £i + 2-"- 1 £i, so ^(/i, IAUB) < %ei + 2~n~lei = e2. For let y — (xi, . . . , x n , 0,a:i, X2, - . .) 6 -4 for certain x E E. Then
<
%g(x,f(x})
< V 2 £i. Moreover, if Z = (Xi, ... ,Xn, l/2,Xl,X2, . . .)
for certain x £ F then
558
B. ANSWERS TO SELECTED EXERCISES
and 0~ 1 (G^ 1 (2)) = x so that for s = f~l(x) we have a S
( ) = (l/2Sl + l/2Xl,.. . , l/2Sn + y 2 Z n , 0 , S l , S 2
and so
h ( z ) = (%8i This shows that
So by Theorem 5.2.4 there exists an element h 6 IK(Q) such that fo f A U B = and g(h, lg) < 62Define the function /f : Q x I —>• Q by
for all 0 < t < 1. We claim that H is as required. Since
0,
<^~ 1 ,
G,
G',
/i,
/i"1,
are uniformly continuous, we see that H is continuous, i.e., /f is an isotopy. It remains to verify the following: (1) Ho = lQ,
(2) H1\E = f , (3) g(Hu, Hv) < e for all u,v£l. Observe that (1) is trivial since Go = lg. For (2), let x e E. Then (j>(x) = <j>(x) G A, hence h(<j)(x}} = a(x). Also,
and so h(Gi o ^(/(x))) = P ( f ( x ) ) = Gi(a(x)). So Gi(a(x)) € /i[B] from which it follows that
Hence
which is as required. For (3), take arbitrary u, v € I. For every 0 < t < 1 we have t, lg) < ff(0, lg) + 2£(G t , lg) + p(ft, lg)
B. ANSWERS TO SELECTED EXERCISES
559
which implies that
Gu h
Gu,(f)
Gv h
Gv)
So H is an e-isotopy. *• Exercise 5.3.1: Pick ei > 0 such that g(f, IE] < £1 < e and put S — l/4(e — £i). Let K — E U F . Then K E Z(Q) by Lemma 5.1.2(3). By Theorem 5.3.5 there is an element
such that
Notice that
£(0, U) < 2e(p, IQ) + £(/, IE) < 26 + ei = s2. By Exercise 5.2.1 there is an £2-isotopy G: Q x.I—> Q such that
Go = IQ,
Gi \ A = g.
Now define the isotopy H: Q x I —> Q in the obvious way by
Then /fo = IQ, H\(x) =
= f ( x ) for x £ E, and for all w, w G I we
< 28 — 6.
So we are done. ^•Exercise 5.4.1: We will make use of the capset H defined on Page 333. Define the homeomorphism
An easy observation shows that
560
B. ANSWERS TO SELECTED EXERCISES
let {En : n G N} be a partition of N such that En is infinite and min(£'n) + 1 G En for every n. Factorize Q as
J El x J^2 x • • • x J E ™ x • • • . Define -A = {W^n(En} n W^n(En}+l : n G N} and notice that \jAC
and are homeornorphic to Q and are elements of Z,(JE?l). There consequently exists by the Homeomorphism Extension Theorem 5.3.7 a homeomorphism h: J E ™ —> J£" such that {z G J E " : x m i n ( S n ) = xmin(En)+l = I}. It follows that the homeomorphism
h = hi x hz x • • • x. hn x. • • • : Q -*• Q has the property that h[E] C \JA C ^[B(<9) x B(Q)]. So B(Q) x B(Q) is a capset by Proposition 5.4.6 and Theorem 5.4.3(2). [Remark: This exercise can also be solved by using the characterization of capsets in Corollary 5.4.11.] *> Exercise 5.4.10: If K = Q then there is nothing to prove. So assume without loss of generality that K ^ Q. Write Q\K as Ui*lo ^ wnere each d is compact, Co = 0 and each d is contained in the interior of Cj+i. For every i let
Observe that <5i > 62 > • • • > Si > • • • > 0, and that lim;_>.oo & = 0. Put go = f . By induction on i > 1 we will construct continuous functions
having the following properties: (1) §i \ K = f \ K and gi \ d is a Z-imbedding, (2) 9i \d-i=9i-i f C i - i , (3) to,Pi-i)<2-1, (4) if x G Q \ K then ^(^(x),5i-i(a;)) < T%n(x
(5) ( a n g
B. ANSWERS TO SELECTED EXERCISES
561
Suppose that for some i > 0 we constructed gi satisfying our inductive requirements. We will construct gi+i. Let e = min(2- (z+1) , 2- (l+1) ^ +2 ). Let 5 > 0 be such that for every space Y and every closed subspace A C Y and all continuous functions £, 77: A —> Q with g(£, rj) < 5, if £ has a continuous extension £: Y —> Q then 77 has a continuous extension fj: Y —> Q with £(£, fj] < e (Exercise 4.1.5). The function gi: Ci+i —>• Q can be approximated by a ^-imbedding h: d+i —> Q such that (6) M ^ = ^ f a , (7) (d+ir\h-1[B(Q)])\K
= (Ar\d+i)\K
(Corollary 5.4.13). So we may assume that g(h,gi I Ci+i) < S. Let V be the interior of d+2. By construction, C;+i C V. Put A = (Q \ V) U Ci+i. Define a function £: yl —>• Q by
The distance between £ and gi \ A is obviously smaller than S. Since gi \ A can be extended to gi : Q —>• Q we find that { : A -» Q can be extended to a continuous function £ : Q —>• Q such that ^(^, pi) < e. We claim that by putting gi+i = £ we satisfy our inductive demands. Observe that (l);+i, (2)i+i, (3)i+i and (5)i+i are trivially satisfied. We will check (4)i+i. Since gt+i and ^ agree on Q\ V, it suffices to consider a point x G Ci+%. Observe that for such x,
Since <5;+2 < f^(x) for every x G d+z, we are done. Put g = limi->.oo gi- Then g is continuous by (3) and Lemma A. 3.1. Clearly,
9 f K = f \ K. By (1) and (2) it follows that g f Q \ K is one-to-one. In addition, for x € d we get by (4) and (2) that
It therefore suffices to prove that g~1[B(Q)] \ K = A\K. easily from (5).
That this is true follows
^Exercise 5.5.1: By Exercise A. 2. 18 we may assume that e is defined on all of X and that e \ K = 0. By Corollary 5.3.10 there is a deformation S : Q x I -> Q through Q \ f [ K ] . We may assume without loss of generality that g(St(x),x) < t
562
B. ANSWERS TO SELECTED EXERCISES
for every x G Q and £ G I (Proposition 4.1.7). Define /' : X —> Q by the formula /'(») =
Se(x)/4(f(x)).
Then /' is continuous, /' \ K = f and (/(z), /'(*))< V 4 e(x)
(1)
for every x £ X. In addition, if x G X \ K then e(x) > 0 and so f' (x) £ f[K]. Let H : Q x. I —> Q be a deformation through A. Again by Proposition 4.1.7 we may assume that Q(Ht(x),x)< %t for every x G Q and t G I. Define £: X —>• [0, oo) by £(*) - min{ 1/46(3;), y 2 e(/'(a:), /'[#])}• Then £ is clearly continuous. Finally, define g : X —>• Q by
tf««)(/'OiO).
Then g is continuous, g \ K = f \ K = /. Also, 0(/'(*),
(2)
for every a; G X. So (1) and (2) show that for arbitrary x we have , /'(*)) + g ( f ' ( x ) , g ( x ) )
< %e + %e < e.
In addition, if x G X \ K then e(x) > 0 and so /'(x) 0 /'[•&"]• As a consequence, by (2) we get g(x) 0 /'[-K"] = s[-K]- So we are done. ^•Exercise 5.5.2: The solution to this exercise is similar to the proof of Theorem 5.5.5. Let
be a Z- imbedding. In addition, let K and C be subsets of Q such that K is closed and C € M. Just as in the proof of Theorem 5.5.5 we can find a continuous function g\ : Q —>• Q closely approximating /i such that
and 0i f Q \ K is a one-to-one map whose range is a aZ-set. Since there is a deformation of M through X, we can find a continuous function g-z : Q —>• M which closely approximates /2 such that 52 f K = ji \ K while moreover g^[Q \ K] C. X (Exercise 4.1.9). The map g — (#1,02) is a Z-imbedding of Q into Q x M which closely approximates / while moreover g \ K = / \ K and g~ 1 [A x X] \ K — C\K. ^Exercise 5.5.3: We adopt the notation we introduced on Page 333. We proved in Proposition 5.4.6 that the set 00
2= US., n=l
where
B. ANSWERS TO SELECTED EXERCISES
563
is a capset for Q. By the same argument as in the proof of that proposition, it follows that
s(oo) = y s~ oo
is a capset for Q°° . But S(oo) is contained in S°°, which is an 3~CT(5-absorber in Q°° by Theorem 5.5.5. So the J^-absorber 5°° contains a capset. By the topological uniqueness of Svj-absorbers (Theorem 5.5.2), this consequently holds for all 3a5absorbers. ^•Exercise 5.5.4: Let AI D A2 be an ( f a , 3v) -absorbing pair in Q (Corollary 5.5.6). By the proof of Theorem 5.5.2 there is a homeomorphism h: Q —> Q such that h[A] = A2,
h[B]CAi,
g(h,l)<%.
Since AI D AI is an (3>, ^a) -absorbing pair in Q and C D C is a decreasing pair of FCT -subsets of X, there is a Z-imbedding p: X —>• <5 such that (5) p- 1 [A 1 ]\K = C \ K = p- 1 [A 2 ]\^,
(6) p r tf = (ft o /) r K,
(7)
e(P,hof)<%.
Let g = h~l o p. Then 5 f ff = / f K . Furthermore
Finally,
cT'fA] \ K = p-1 [h(A]] \ K = p~l[A,} \K = C\K and g[X \ (C U K)] n B C /i"1 [p[X \ (C U K)]] n A'^Ai]
= h-1[P[x\c]r\p[x\K]r\Ai] = 0. So we are done. ^•Exercise 5.5.5: Let Qn be an admissible metric on En such that g ( x , y ) — mdLx{gn(xn,yn) : n G N} is an admissible metric on E (Exercise A. 5. 16). Pick an arbitrary subset X C E such that
Y[xn ex c We will modify the proof of Theorem 5.5.5 to obtain the desired result. Consider a map /: Q —>• E that restricts to a Z-imbedding on some compactum K and a sequence Q D AI D A-z D • • • of FCT-subsets. Put /loo = D^=i ^™- ^e may assume that / is a Z-imbedding (Theorem 5.3.11). Write Q \ K as a union of a
564
B. ANSWERS TO SELECTED EXERCISES
sequence (F;)°^0 of compacta with Fi C. Int(F;+i) for every i and Fo = 0. Let e > 0 and put for every i. Consider now the n-the component fn : Q —*• Q of f. We can construct just as in the proof of Theorem 5.5.5 a sequence 0:0,0:1,... of continuous functions from Q to En such that for every and i: (1) £„(<**, Qi-i) < £j, OLi \Fi-i = Q i - i f F i _ i ,
(2) Qi t Q \ ^i+i = /n I" Q \ F;+I (3) a t r 1 [A'n]nF i = A n n F i .
and
«i t # is a Z-imbedding,
By Exercise 5.5.4, we may assume that the function on : Q —>• EVi additionally satisfies the following property: (4) a i [ F i \ A n ] n C n = 0 . This has the effect that for the function gn = limi_»oo o.i we have
(5) We claim that the function g = (gn)n has the property that 0~ 1 [X] \ K = Aoo \ K. Let x G S"1^] \ #• Then 00*0 G X C []^i ^- So for every n e N, p n (x) e Cn, hence ^ n (x) i gn[Q\ (An U K)]. This gives x g Q \ (A n U K) for all n. Since x <£ K we consequently obtain x € A n for all n, hence x G Aoo- Conversely, if or £ Aoc \ K then for each n, re G An \ ff. Consequently, gn(x) G Sn for every n which implies that
*• Exercise 5.5.6: By Proposition 5.5.17 there is a continuous function 5: s x s x I —>• I such that (1) 6(x,p,t) = 0 if and only if t = 0, (2) 6(x,p,t) < t, (3) g(p,p + <5(x,p, i)x) < t for every t € I and x,p G s. Define £: Q x Q -> I by /
0
It is clear that ^ is well-defined. In addition, if (x,y) #. K then
by the definition of £ and property (3) of 5. If (x, y) £ K then £(rc, y) = 0. Moreover, if (x, y ) $. K then e(x, y) > 0 and so
by property (1) of 6. It suffices to prove that £ is continuous and for this it suffices to prove its continuity at points of K. To this end, let ((or n , y n )) be a sequence in (QxQ)\K converging to
B. ANSWERS TO SELECTED EXERCISES
565
an element (x, y) G K. Then e(xn, Vn) -> 0 from which it follows that £(x n , y-n) —>• 0 by property (2) of 6. ^•Exercise 5.5.7: The continuity of g\ need only be checked at points of K. To this end, let ((xn, yn)) be a sequence in (Q x Q) \ K converging to an element (x, y) G K. Then by Exercise 5.5.6, e(fi(xn,yn),gi(xn,yn)) < e(xn,yn) -» 0 from which it follows by continuity of /i that gi(xn,yn) —)• fi(x,y) which is as required. It is trivial that £>(/i(x), #i(x)) < 2e(x) for every x G Q x Q.
Fix 7 G r. if
(x,y)e(gil[Ljn(R°°
= gi(x,y),
xg))\AT
then x G R°°, (x,y) £ X, hence £(x,y) > 0, and gi(x,y) G L 7 . This implies that L~, 3 9i(x,y) = fi(x,y) + £(x,y)i(>(x). Since /i (x, ?/) G L C LT and L7 is a linear subspace of M°° we conclude that £(x,y)ij>(x)
G I7
and since ^(x,?/) > 0, we get ift(x) G L 7 . So x G ^~ 1 [L 7 ] n R°° = I/7, i.e., x G I/7. Consequently, (x,y) G (I/7 x Q) \ K. Since (x,y) was arbitrary, this shows that (g^(L^} n (R°° x Q)) \ K C (L7 x Q) \ K Conversely, if (x,y) G (L7 x Q) \ K then x G L7 = V'"1^] n R°° So x £ R°° and V)(a;) G ^7> hence
and
(^,2/) ^ K.
We conclude that (x,y)G(5r1[^]n(K00 xQ))\A-, which is as required. ^•Exercise 5.5.8: Since /» f A' = ^ \ K for i = 0, 1, it clearly follows that / \ K = g f K. It is also clear that d(/(x), #(x)) < 2e(x) for every x G Q x Q. We shall prove that # is one-to-one. To this end, take distinct points There are several cases to consider. If both (o;,y) and (x .y'} are in K then there is nothing to prove since / \ K = g \ K is an imbedding. Next, assume that e.g., (x,y)£K,
(x',y')£K.
Then e(x' ,y'} = min{d(f(x', y'), f [ K ] ) / 4 , 9/4} > 0 and so
d(g(x', y'),g[K}) > d(f(x', y'}, f[K])/2 > 0. Since g(x,y) G g[K] this proves that g(x'.y') ^ g(x,y). Assume finally that ( x , y ) , ( x ' , 2 / ) 0 K. But then gz(x,y) ^ 92 (x' , y') by construction and so g(x,y) ^ g(x',y') as well. So g is an imbedding by Exercise A. 5. 9. It remains to prove that g[Q x Q] is a Z-set in Q x Q. Since f[K] = g[K] is a Z-set, it
566
B. ANSWERS TO SELECTED EXERCISES
suffices to prove that L = g[(QxQ)]\K is a aZ-set (Lemma 5.1.2(3)). Since L is acompact, it consequently suffices to prove by Lemma 5.1.2(1),(3) that it is contained in a crZ-set. But this is clear from Lemma 5.1.2(5) since gi[(Q x Q) \ K] E Za(Q). ^•Exercise A.1.2:
For (1) =4> (2), observe that A misses X \ A. Since p is an accumulation point of A, by (1) it follows that p is not an accumulation point of X \ A. So there is a neighborhood U of p which misses
and hence is contained in A U {p}. For (2) =* (3), for some x £ X let 5= [ACX :x£ A\{x}}.
We will first show that 5 has the finite intersection property. To this end, let 9 be an arbitrary finite subcollection of 3. Then by (2), for every G £ 9 there is a neighborhood UG of x such that UG C G U {x}.
Put U = DGGS UG- Then U is a neighborhood of x and since X has no isolated points, there consequently is a point p E U \ {x}. So clearly p| 9 7^ 0 since this intersection contains p. We next prove that f is an ultrafilter by showing that if A and B are arbitrary complementary subsets of X then either A E 9" or B E 9" (Exercise A. 1.1). Assume the contrary, that is, there are complementary sets A and B in X such that p is neither an accumulation point of A nor B. Then there is a neighborhood U of p which misses
hence is contained in {p}. But this contradicts the fact that X has no isolated points. Since (3) =>• (1) is trivial, this completes the solution. ^•Exercise A. 1.4:
We prove this by induction on k. If k = 1 then 71 contains infinitely many singletons and so there is nothing to prove (let A = 0). Suppose that we are done for k — 1, where k > 2. For z E N let Ai E A be such that A* ^ A, if i / j. If (A; : i E N} contains an infinite pairwise disjoint subfamily, then we are clearly done. So assume that this is not true. Then there has to be an i E N such that the set E = {j€N\{i}:
A, H Ai ^ 0}
is infinite. We may assume without loss of generality that i = I . Since A\ is finite there consequently has to an infinite subset F C E and an element x € AI such that x E Aj for every j € F. For every i put Bi = Ai\ {x}. By our induction hypothesis, there are an infinite subset E C F and a set B' C X of size at most k — 2 such that if i,j £ E are distinct then
B. ANSWERS TO SELECTED EXERCISES
567
Now put B — B' U {x}. Then if i, j G E are distinct, we have
Ai n Aj = (Bl U {x}) n (Bj U {x})
= (Bi n B j ) U {x} = B' U {x} = 5. So we are done since clearly \B\ < k — 1. ^•Exercise A. 1.10: That Q is a metric is easily seen. To show it is admissible, consider a basic open subset U of Y[ = n^Li Xn first. It has the form
where N € N and 0 ^ Uj C Xi is open for every i < N. Pick an arbitrary x £ £7, and let e > 0 be such that Bgi(xi,e) C [7; for every i < N. Pick an arbitrary element y G H such that g(x,y) < 2~Ne. Then for i < N we have 2~iQi(xi,yi)
<2~Ne
and so Qi(Xi,yi)
<2~N+l£<£. -7V
We conclude that y €.U, i.e., 5 e (x, 2 £) C U. Conversely, take an arbitrary x £ H> and ^ £ > 0. There exists A^ € N such that
For i < N consider the ball Ui = BBi(xi, l/2N^)neighborhood
t/=n^ n A"' TV
n=l
x
We claim that the basic open
oo
n=N + l
of x is contained in Be(x,e). Indeed, take an arbitrary y G £7. Then
So we are done. *>Exercise A.2.13: Let X be uncountable, and put U = {U C X : U is open and countable}. The collection U of has a countable subcollection with the same union by Corollary A.2.3. So (J U is a countable union of countable sets and is therefore countable
568
B. ANSWERS TO SELECTED EXERCISES
itself. We claim that Y = X \ (J 11 is dense-in-itself. In fact we will show that it is locally uncountable, that is: every non-empty (relatively) open set is uncountable. This is clearly as required. Striving for a contradiction, assume that V C Y is a non-empty relatively open set which is countable. There is an open set V' C X such that V' Pi Y = V. But then V' , being the union of V and the countable set
is countable. This shows that V' e U and hence that 0 ^ V = V n Y — 0, which is a contradiction. ^•Exercise A. 4. 4: Observe that if E and F are subsets of X then EHF C En F. So if Er\F / ~E n F then
(F n F) \ E n F ^ 0. Striving for a contradiction, assume that B is uncountable. By the above, for every 6 G B there exists an element
a(S) € (r1 [[o, 5)] n u) \ r1 [[o, <j)] n t/.
(i)
If/(a((J)) < 6 then
a(<$) e r1 [[o, <$)] n C7 c /-1 [[o, 8)] n t/ which contradicts (1). So /(a(<5)) = 5 for every S G -B which means that we can order the set {a(6) :6£B} by the following rule:
a(60) < a ( S i )
o-
50 <6\.
By Exercise A.2.14 there is an element Si € B such that
Let V be an arbitrary open neighborhood of a(5i). There exists <5o G B such that SQ < 51 while moreover a (So) 6 V. But then a(60)£Vnf-l[[Q,6i)] This proves that
HU.
yn/- 1 [[o,<5i)]nt/^0.
Since V was an arbitrary neighborhood of a(5i), we get
But this contradicts (1). ^•Exercise A.4.5: We may assume without loss of generality that V is open. Let /: X ->l
by a Urysohn function with f(A) C {0} and f[X\V] C {1} (Lemma A.4.1). Since U is countable and (0,1) is uncountable, there is an element S G (0,1) such that
B. ANSWERS TO SELECTED EXERCISES
569
for every U G U (Exercise A.4.4). Hence
w = r1[[o,6)]
is as required. ^•Exercise A.5.15: Let Y be X with the topology generated by the metric Q. We claim that the identity X —> Y is continuous. This suffices since by compactness of X it will then follow that the identity is a homeomorphism (Exercise A.5.9). It is clearly enough to prove that if x G X and 6 > 0 then B(x,6) is open in X. Take an arbitrary y G B(x,5). Then g(x: y) < 6. Hence by continuity of g there are neighborhoods U and V of x and y, respectively, such that Q[UxV]C[Q,6). So g[{x} x V] C [0,5) showing that V^ C £(z,5). We conclude that B(x,S) is a neighborhood of y. ^•Exercise A.5.16: By Corollary A.4.8 we may assume that for every ra, Xn is a subspace of a compact space Zn. (This also follows from Corollary A.4.4.) By Exercise A.1.6 there is an admissible metric gn for Zn which are bounded by 2~n. We claim that g ( x , y ) = ma,x{gn(xn,yn) : n G N} is as required. That it defines a metric on Z = O^Li Zn is left to the reader. To show that it is admissible on Z, simply observe that Q : Z x Z -» [0, oo)
is continuous since Z is endowed with the topology of coordinatewise convergence (Exercise A.2.2) and the functions gn are continuous on Xn *Xn. So g is admissible on Z by Exercise A.5.15. Hence the restriction of g to X = H^Li Xn is admissible on X since X is a subspace of Z (Exercise A.1.13). ^Exercise A.6.3: Let e > 0. Since g is totally bounded, we may choose by Exercise A.6.2 a finite y2enet B for X. Put B' = {x£B:g(x,A)< %£}. For every x G B1 pick a point ax G A such that ^>(x,a x ) < y2e. We claim that the finite set {ax : x G B'} is an e-net for A. If so then we are done by another application of Exercise A.6.2. Let a G A be arbitrary. There exists x G B such that ^(x,a) < y^e. Hence x £ B' and £>(a, a x ) < £>(a, x) + g(x, ax} < l/2e + y2e = e. So we are done. ^Exercise A.6.9: For (1) => (2), assume first that / is not one-to-one. There are distinct x,y G X such that f(x] = f(y). Let Xk ~ x and yk = y for every k. The sequences
570
B. ANSWERS TO SELECTED EXERCISES
and the number £ = 2g(x,y) demonstrate that (2) holds. Assume therefore that / is not closed. There is a closed set A C X for which f[A] is not closed. By Exercise A. 2.1 there is a sequence (a n ) n in A such that lim an = z If (a-™)™ converges say to p, then p G A since A is closed and so f[A] E3 f ( p ) = z which is a contradiction. The completeness of Q implies that (a n ) n is not Cauchy. There consequently exists e > 0 such that for every N G N there exist n,m > N with 0(a n ,a m ) > e. A simple inductive argument proves the existence of subsequences (flnjfc,
(Omjfe
of (a n )n such that for every k we have g(ank,amk) > £• The existence of the sequences (ank)k and (amk}k and the point z proves (2). Conversely, assume that (2) holds. We shall prove (1). We assume that / is closed as well as one-to-one and will derive a contradiction. Claim 1. The sequence (xk)k has a convergent subsequence. Proof. If not then for every E C N we have that the set then the set {xn : n E E} is closed (Exercise A. 2.1). Consider the set {f(xn} '• n G N}. Since / is closed, it follows that for every E C N the set { f ( x n ) '• n G E} is closed as well. But then the sequence (/(a;™)) is eventually constant since it is convergent. But this contradicts / being one-to-one. •(> So we may assume without loss of generality that (xk}k converges, say to x. We may assume similarly that (yk)k converges, say to y. But then g(x,y] > e > 0 and f ( x ) = f ( y ) . This is a contradiction. ^-Exercise A. 6. 12: Consider the projection TT: I 2 -> I. We first claim that the set Qo = Q 0 1 is the TTimage of a Gg -subset of I2. This is easy. Enumerate Qo as {qn '• n G N}. For every n let zn = (qn,l- Vn) € I 2 The set Z = {zn '• n G N} is discrete and is therefore a G<$-subset of I 2 (Exercise A. 2. 6). Now let II: I3 —» I 2 be the projection onto the first two coordinates, i.e., n(z,3/,2) = (x,y). Then II is open by Exercise A. 5. 5(3). Put
S = ((0, 1] x I 2 ) U {(0, g n , 1 - Y n ) : n G N}. Then S is a dense GVsubset of I3 (since its complement is evidently a countable union of compact nowhere dense subsets of I 2 ) and
T = U[S] = {(x, y) G I 2 : (x = 0 => y G Qo)}. But T is not a Gg -subset of I 2 . Observe that {0} x I is a G5-subset of I 2 . So if T would be a G^-subset of I 2 then Qo would be a G^-subset of I. This leads to a contradiction in the same way as in the proof of Corollary A. 6. 7.
B. ANSWERS TO SELECTED EXERCISES
571
^•Exercise A.7.5: We may assume without loss of generality that It is locally finite in (J It (Corollary A.7.3). For every U € It define fv: X ->• I by
The function e: X —>• I defined by
s(x) = ma.x{fu(x) : U G It} is well-defined since It is locally finite. Clearly, e"1 [(0,1]] = [Jit. In addition, £ has the property that for every x G X,
is contained in an element of It. To prove this, first observe that if x £ A = X\(JU then e(x] = 0. So if x G A then the set in (*) is empty, and is therefore contained in every U G It. Moreover, if x G |J It then let U G It be such that e(a) = fu(x). Then clearly {y € X : o(y,x) < e(z) = min{l, g(x,X\ U)}} C U. It remains to verify that £ is Lipschitz. To this end, take arbitrary x, y G X. We have to distinguish several subcases. If x, y G A then e(x) — e(y} = 0 and so there is nothing to prove. Assume next that x 0 A, say t/ G It is such that
If y 0 A then y (£U and so
as desired. So we can assume that ?/ 0 A as well, say U' G It is such that
We may assume without loss of generality that e(x) < e ( y ) . If x £ U' then
So assume that x G U'. Claim 1. e(y] < Q(X, y) + e(x). Proof. If e(x] = I then there is nothing to prove. So we may assume that
Since min{l, g(x,X \ U'}} < e(x} < 1
we get e(x,X\U')<e(x-).
(1)
Observe that Q(V, X\U')< g(x, y) + g(x, X \ U')
(2)
and hence e(y) = min{l, g ( y , X \ U')} < min{l, g(x, y) + g(x, X \ £/')}•
(3)
572
B. ANSWERS TO SELECTED EXERCISES
If e(y) = 1 then by (3) and (1),
e(y) = i < e(x,y) + e(x,x\u'}< e ( x , y ) + e(x). On the other hand, if s(y) < 1 then e(y} = £>(y, X \ U'} and we are done as well by (2) and (1). 0 So by the Claim and e(x) < e(y), we get \E(X) - e(y)\ = e(y) - e(x) < g(x, y),
as required. ^Exercise A.8.1: Let {qn '• n 6 N} be a faithful indexing of Q. For every n G N let (pfj. be a sequence of pairwise distinct irrational numbers converging to qn. By Lemma A.4.1 there is for every n e N a continuous function such that / n ( p ? ) = 0 (ieven), /n(p?) = 1 (i odd). Define /: P -»• I by
Then / is continuous by Exercise A.3.1. We claim that / is as required. To this end, fix m G N. By continuity of the functions (/ n )n<m we can find a neighborhood U of x = qm in R such that
for every y £ U\{qi,... ,qm-i}. We pick arbitrary yo 6 C/ n {pr : « even},
y\ G C/ D (p^ : i odd}.
It follows that n<m
n>m
/(i/i) - E 5 ~"/»(»i) +5"m + E 5 ~ n
Since (*) implies
it follows that
B. ANSWERS TO SELECTED EXERCISES
573
Since y\ and yo were chosen arbitrarily, this shows that / cannot be extended to a continuous function /: PU {qm} —> I. ^ Exercise A. 8. 3: First Solution: We may assume without loss of generality that S is dense in X (Exercise A. 2. 5). By Corollary A. 4. 8 we may assume that X is a subspace of some compact space X, and Y is a subspace of some compact subspace Y (the proof in fact already 'works' if X and Y are just topologically complete). By the Lavrentieff Theorem A. 8. 5, there are G«5 -subsets So of X and YO of Y containing S and y, respectively, such that g = f\S:S—tY can be extended to a homeomorphism g: So —> YQ. The continuous functions
/ \ S0 n x -. S0 n x -» y and
g t So n X : S0 n X -)• y agree on a closed subspace of So HX (Exercise A. 1.9). Since they agree on S and S is dense in X , it follows that they agree on all of So D X. Assume that there exists an element x £ (So fl X) \ S. Then f(x) G Y and hence there is an element s G S such that /(s) = /(#) (here we use that / is surjective). But then s ^ x and yields a contradiction since g is injective. So So n X — S, proving that S is a G&subset of X. Second Solution: We may assume without loss of generality that S is dense in X (Exercise A. 2. 5). Put g = f \ S and let 23 be a countable open basis for y. For every pair E, F of elements of 23 such that E C F, consider the disjoint closed sets of S. Observe that the closed set V(E,F)=g-1[E\ng-i[Y\F\
misses S. Since 23 is countable, the collection S of all V(E,FYs obtained in this way is countable as well. It suffices to prove that X \ S = (J S- To this end, assume that there is an element x£X\(S\J\j9). Since S is dense in X there is a sequence (xn)n in S with limit x. The set {xn : n G N} is an infinite, closed and discrete subset of S, hence B = { f ( x n ) : n € N}
is an infinite, closed and discrete subset of Y. Simply observe that / \ Y is a homeomorphism. There is an element F G 23 such that f(x) G F while moreover F misses a neighborhood of infinitely many elements of B (in fact, all but one). Let E 6 23 be such that f(x) £ E C~E C F. Then f~l[E] is an open subset of X which contains x. Since S is dense in X, this implies that
574
B. ANSWERS TO SELECTED EXERCISES
Since F misses a neighborhood of infinitely many elements of B, by similar considerations we obtain that for infinitely many n we have
This gives us that x € g l[Y \F], hence x G V(E, F) C \J g. This is a contradiction. ^•Exercise A.10.6: Let x £ X be arbitrary. We claim that there is a noncut point y e X \ {x}. This clearly suffices. Striving for a contradiction, assume that every point y (E X \ {x} is a cut point. Let 3 = {Bn : n € N} be a countable base for X such that Bn 7^ 0 for every n. Put CQ = {x}. Recursively, we will construct for every n a subcontinuum Cn of X having the following properties: (1) C n -i is contained in the interior of C n , (2) C n n B n ^ 0 , (3) Cn ^ X. Suppose Cn has been constructed. Pick a point p € A^ \ Cn in the following way. If Bn+i \ Cn ^ 0 then pick p from that set. If B n+ i C Cn then let p <= X \ Cn be arbitrarily chosen (here we use (3)). By assumption, p is a cut point of X. So we can write X \ {p} as the union of two disjoint nonempty open sets, say U and V. We assume without loss of generality that x 6 U. Since p £ Cn and Cn is a continuum, it consequently follows that Cn C U. Put C n +i = U(J{p}. Then C n +i is a continuum by Exercise A.10.5. Moreover, C n +i 7^ A since it misses V. The interior of Cn+i contains U and hence Cn. If Bn+i C Cn then (2) is evidently satisfied for n + 1, and if not, then (2) is also satisfied since then p E Bn+i n C n +i. We conclude that Cra+i is as required. If Un°=1 Cn = X then the interiors of the Cn's cover A" by (1). The compactness of A" then implies that one of the C n 's is equal to X. But this contradicts statement (3). Pick an arbitrary point y E X \ (Jn°=i Cn- By assumption, y is a cut point. There consequently exist nonempty, disjoint and open sets E and F such that A" \ {y} is equal to E(JF. We may assume without loss of generality that x G E. Since y £ Cn we have that Cn C E for every n. Hence U^Li Cn is contained in E and so misses the nonempty open set F. But it is dense by (2). This gives us the contradiction we are after. ^Exercise A.11.3: Let {Un '• n E N} be a countable collection of neighborhoods of {N} in R/N and let TT: R —> R/N be the natural quotient map. For every n we have that the set 7T~ 1 [t/ n ] is an open neighborhood of N in R. For every n e N let an < n < bn be such that (a n ,6 n ) is a proper subset of (n — l/4,n-\- %} nf|"=1 ir~l[Ui]. Observe that the collection of intervals {(a n ,&n) : n e N} is pairwise disjoint. The set
oo V= \J(an,bn) n= l is a neighborhood of N such that -K~I [TT[V]] = V, hence Tr[V] is an open neighborhood of the point {N} in R/N. Assume that there exists N e N such that UN is
B. ANSWERS TO SELECTED EXERCISES
575
contained in ir[V]. Then 7r~ 1 [f/jv] Q V. But this is impossible since (a,/v,&/v) is a proper subset of 7r~l[UN] D (AT - %,N + y4) and (N-y4,N+ y4) n (n - y 4 , n + y 4 ) = 0 for every n ^ N. ^Exercise A. 11. 5: Let 6 be the collection of all components of S. By Corollary A. 11. 3, S/G has a base consisting of clopen sets. Let q: S —» S/G be the natural quotient map. Observe that S/G is compact since S is, and that by our assumptions, q[A n S] n g[B n 5] = 0. Since g is closed (Exercise A. 5. 5), the sets q[A n S] and g[B D S] are closed in S/G. Since they are also disjoint, for every x G S/G we may pick a clopen neighborhood Ux such that ux n q[A n s] = 0 or c/x n q[B n 5] = 0.
The cover has by compactness of 5/6 a finite subcover, say {Uxl,...,UXn}. For every z < n now put ^ = t4AU^ 3
Bn= B'nn F,
respectively. All the A n 's and £?n's are closed subsets of X. By the same reasoning, there is a (countable) collection {(H'n, G'n} : n G N} of pairs of disjoint closed sets in ^X such that for any pair (H, G) of disjoint closed subsets in ^X there is an m G N such that H C H'm and G C G'm. For every n put Hn = Hn n X,
Gn = Gn n X,
respectively. By Exercise A. 9. 3 there is a compactification j3X of X such that for every n the closures of An and f?n as well as the closures of Hn and Gn are disjoint. This implies by Lemma A. 8. 3 that the identity function X —>• X can be extended to a
576
B. ANSWERS TO SELECTED EXERCISES
continuous function /: f3X —> ^X. In addition, the identity function F —>• F can be extended to a continuous function £ : F* —> aF. Here F* means the closure of F in /3X. _ _ Since aF > F there is a continuous function r\ : aF —> F which extends the identity on F. Observe that the continuous functions r\ o £ and / [* F*, F*
_A> aF JL* F, F* ^ F,
both restrict to the identity on F. Since F is dense in F*, both functions are equal by Exercise A. 1.9. It is now clear that the adjunction space bX = j3X Ug F* is the required compactification of X. ^Exercise A. 12. 8: To the contrary, assume that i/>n is inessential. By Theorem A. 12. 8 there exists a continuous function (p: S1 —>• R such that for all z £ S1 (the constant '2?r' is added for technical reasons). Fix 9 E R. Then for z — e27"61 we have
i.e., From this we conclude that the continuous function / : R —>• R denned by takes its values in Z only. By connectivity of R it therefore follows that / is constant. So there exists N G Z such that (V0GR). From this we conclude that n-l + N = ^(e27"'1) =
B. ANSWERS TO SELECTED EXERCISES
577
for certain e > 0. Let p = (e,sin %). Then f(p) 0 E and so by the same argument as above, the image of the arc {(x,sinl/x) : e < x < 1/vr} misses E. So f[X] misses E and hence is contained in a subarc of X. Then / maps this arc into itself, and hence we find a fixed-point there by again applying part (1). *• Exercise A.12.11: To the contrary, assume that f(x) ^ x for every x £ D. Define r: C\ {0} —>• S1 and g: S1 —>• S1 by r(z) = z/\z\ and g(u) = r(u — f ( u ) ) , respectively. For every t € I and u G S1 we have H(u,t) = u-tf(u) ^0. For t — 1 this is clear and for t < 1 observe that \ t f ( u ) \ = t \ f ( u ) \
This page intentionally left blank
APPENDIX C
Notes and comments
We will be brief regarding historical notes. The books of ENGELKING [153, 155] are excellent sources for the reader interested in the history of topology. See also AULL and LOWEN [30], and JAMES [212]. >§!.!: Exercise 1.1.10 is due to Kuratowski [239] and Wojdyslawski [411]. Its generalization Corollary 1.1.8 was proved by Arens and Eels [23] and Blumenthal and Klee [63]. Their proofs yield isometric imbeddings of metric spaces in normed linear spaces of the form £00(A), for some set A. The elegant proof of Lemma 1.1.6 is basically due to Michael [286]. It was brought to my attention by Dobrowolski that Michael's arguments can be improved to get imbeddings in function spaces of the form C(A), where A is a, compact space. With his permission, his observation is published here. The Anderson Theorem from 1966 about the homeomorphy of £2 and s that was mentioned on Page 9 is due to Anderson [15] and gave the answer to a question posed by Frechet [167] in 1928 and also by Banach [38] in 1932. Theorem 1.1.13 is due to Banach. The proof presented here was taken from Brown and Page [79, pp. 316-317]. As in the proof of Theorem 1.1.13, separability is not used in the proof of Theorem 1.1.15. So the result is true for arbitrary Banach spaces. This will be used in Chapter 6. For many more classical results on Banach spaces, see Lindenstrauss and Tzafriri [261]. Theorem 1.1.22 is due to Michael [287]. Continuous selections are important tools in topology and functional analysis. For the first book containing a systematic and comprehensive study of the field of continuous selections of multivalued mappings, see Repovs and Semenov [351]. A more general result than Corollary 1.1.23 is due to Bartle and Graves [44]. The homeomorphism in that result can in general not be chosen to be linear. For let E be a separable Banach space not isomorphic to il. There is a continuous linear surjection T: tl —» E. Then KerT x E is not isomorphic to tl since all complemented linear subspaces of il are isomorphic to f1 by Pelczyriski [335]. In view of our remarks at the end of this section, the question naturally arises whether the linear space C(I) admits an equivalent norm ||-|| such that ((7(1), ||-||) is an inner product space. The answer to this question is in the negative. The proof of this fact requires techniques that are outside the scope of this book. The interested reader may consult for example Taylor [386, p. 195]. Exercise 1.1.13 shows that the formula ||a;|| = (Er=i W P ) 1/P does not define an admissible norm on tp for 0 < p < 1; it is known that if for p > I the set lp is defined similarly then this formula does define an admissible norm on it.
579
580
C. NOTES AND COMMENTS
Lemma 1.2.1 and Theorem 1.2.2 are due to Dugundji [140]. AR's and ANR's were introduced by Borsuk [64, 65]. The material on them presented in this section is standard. For more information, see Borsuk [70], Hu [197] and Chapter 4 of this monograph. In Exercise 1.2.4 we presented a very simple example of a contractible space which is not an ANR. There is also an example of a contractible LC space which is not an ANR, but this is much more complicated. See Borsuk [70, p. 124] for details. This is of particular interest, since in §4.2 we will show that contractibility and local contractibility implies the AR-property in finite dimensional spaces. Exercises 1.2.10 and 1.2.11 are due to van de Vel. See van Mill [298, Theorem 1.4.13] for more details. It is tempting to believe that the property in Exercise 1.2.12 characterizes the AR's among the linear spaces. But this is not true. Cauty [87] proved that there is a linear space L which is an A R having a closed linear subspace which is not an AR. It is not clear whether the AR-property in linear spaces can be characterized in terms of selections. All results in this section are standard. Hl-4: Theorem 1.4.2 and its applications are due to Borsuk [68]. That the cone of an ANR is an AR is well-known. The nice proof of this fact, which was presented here in Theorem 1.4.6, was communicated to me by Dobrowolski who learned it from Toruficzyk. The Cantor middle-third set C was introduced by Cantor [85]. The characterization Theorem 1.5.5 was proved by Brouwer [74]. Theorem 1.5.10 is due to Alexandroff and Urysohn [11]. The first space rilling curve was constructed by Peano in 1890. Theorem 1.5.12 and Corollary 1.5.13 are due to Souslin [375]. Sets A and B such as in Corollary 1.5.14 are called Bernstein sets. Corollary 1.5.15 is due to van Douwen [133]. The proof presented here is taken from Becker, van Engelen and van Mill [45]. It is also due independently to Pol. Example 1.5.18 was established in Erdos [157]. Theorem 1.5.19 was proved in Sierpiriski [367]. Theorem 1.5.22 is due to Mazurkiewicz [275]. Finally, Exercise 1.5.11 is due to Hausdorff [192] and de Groot [177]. The Inductive Convergence Criterion 1.6.2 is due to Fort [165] and was later rediscovered by Anderson [16]. Theorem 1.6.6 is due to Keller [219]. The proof presented here is due to Fort [165]. Theorems 1.6.7 and 1.6.9 are due to Bennett [46]. Exercises 1.6.1 and 1.6.2 can be found in Anderson, Curtis and van Mill [20]. Observe that Exercise 1.6.2 generalizes Theorem 1.6.9. The question naturally arises whether every strongly locally homogeneous Baire space is countable dense homogeneous. This question was answered in the negative by van Mill [294]. §1.7: Bing's Shrinking Criterion 1.7.2 is of course due to Bing [58]. It is one of the most
C. NOTES AND COMMENTS
581
powerful tools in geometric topology. It is easy to realize the cone of Q as a convex subset of £2. So Theorem 1.7.5 is a direct consequence of Keller's Theorem from [219] that all compact convex infinite-dimensional subspaces of I2 are homeomorphic (to Q). Exercise 1.7.1 is due to Schori [362]. Exercise 1.7.2 is folklore. It was observed in the late sixties while thinking about the following well-known problem of de Groot: does there exist a homogeneous continuum homeomorphic to its own cone but which is not the Hilbert cube? This problem is still unsolved. A related open problem is the one due to Bing and Borsuk [61]: is every homogeneous finite dimensional compact AMR a manifold?. ^§1.8: The 'graph method' of extending homeomorphisms used in the proof of Theorem 1.8.2 is due to Klee [221]. ^§1.9: Lemma 1.9.1 is due to van Mill [293]. Sierpiriski [369] proved the characterization of Q in Theorem 1.9.6. Theorem 1.9.8 is due to Alexandroff and Urysohn [10] who in the same paper also characterized the product Qx C as the topologically unique a-compact zero-dimensional space which is nowhere countable. The product QxP was topologically characterized in the same spirit by van Mill [292]: it is the topologically unique zero-dimensional space which is a countable union of closed topologically complete subspaces and which in addition is nowhere topologically complete and nowhere cr-compact. It was shown subsequently by van Engelen [148] that there are 'only' u>i many topologically distinct homogeneous zero-dimensional absolute Borel sets, and that they can all be topologically characterized. His proofs make use of complicated results from descriptive set theory. Corollary 1.9.13 is due to Hurewicz [202] and its proof is to van Douwen [129]. For results in the same spirit, see e.g., van Engelen and van Mill [150]. Exercises 1.9.1 and 1.9.2 are due to van Mill [293]. Call a space rigid if the identity is its only homeomorphism. The real line can also be decomposed into two homeomorphic rigid sets, as was shown independently by van Engelen [149] and Shelah [365]. This is much more complicated than the construction in Exercises 1.9.1 and 1.9.2. It is a natural question (due to van Douwen) whether there exist rigid zero-dimensional absolute Borel sets. The zero-dimensionality in this question is essential for there are simple examples of rigid continua (de Groot and Wille [180]). It was shown in van Engelen, Miller and Steel [151] that rigid zero-dimensional absolute Borel sets do not exist. *>§!.10: Inverse sequences and their limits are used extensively in topology. Especially for the construction of various counterexamples. They are in fact nothing but generalizations of products. Most of the general results on inverse limits that are mentioned here are standard. The dyadic solenoid is one of a family of peculiar spaces. Everything we said here about £2 also works if instead of the function /(z) = z2 in the inverse sequence one considers the function fp(z) = zp'. Here p = 2, 3 , . . . . One then gets the so-called p-adic solenoid Ep as inverse limit. The solenoids are
582
C. NOTES AND COMMENTS
due to van Dantzig [108]. They can be characterized topologically as the unique homogeneous continua having only arcs as proper subcontinua. This interesting result is due to Hagopian [185] (see also Bing [59]). The results we presented on indecomposable continua are all standard. See Kuratowski [243, pp. 208-214] for more details. The hereditarily indecomposable continua are in may respects generic objects, playing an essential role in dimension theory, cf. Mazurkiewicz [278], Krasinkiewicz [230], Sternfeld [382], Levin [253], Levin and Sternfeld [258. 257, 259], Lewis [260] and Nadler [320]. Theorem 1.10.18 is due to van Douwen [126]. The space in Lemma 1.10.20 is due to van Mill [291]. Theorem 1.10.23 is due to Brown [80] and was widely used in the early days of infinite-dimensional topology. For information on dyanmical systems, see e.g., Devaney [110] and de Vries [401]. Exercise 1.10.8 is due to Farah [158] and Sierpiriski [372]. Exercise 1.10.13 is also due to Brown [80]. Exercise 1.10.23 can be stated more precisely: the number of composants is in fact c. >§1.11: Hyperspaces were first considered in the early 1900's in the work of Hausdorff and Vietoris. For general information on hyperspaces see Nadler [320] and Illanes and Nadler [209]. Corollary 1.11.4 is due to Wazewski [404]. For a different proof, see van Mill [298, Proposition 4.7.2] and Exercise 1.11.15. Engelking [154] informed the author that the idea in the proof of [298, Proposition 4.7.2] is close to Kuratowski's proof of the Generalized Bolzano- Weierstrass Theorem (see [242, p. 340]). The Vietoris topology on 2X was introduced in Vietoris [400]. Whitney maps were in a context different from hyperspaces first constructed by Whitney [409]. The construction in the proof of Proposition 1.11.15 is due to Krasinkiewicz [227]. For other constructions and results on Whitney maps, see Nadler [320] and Illanes and Nadler [209]. The function e in Exercise 1.11.19 is called an expansion homotopy on 2X . The results in this section are standard. Simplicial complexes are very important in ANR-theory and algebraic topology. See e.g., Spanier [376] and Munkres [319].
§2.2: The results in this section are standard as well. Nerves of covers were introduced by Alexandroff [8], and their ^-mappings by Kuratowski [238]. Freudenthal's Approximation Theorem is due to Freudenthal [168]. The proof presented here was taken from Nagata [324]. In fact, a much more general statement is true (and proved by Freudenthal) than the one presented here. See Engelking [155, Theorem 1.3.12] for details.
§2.4: Theorem 2.4.3 is due to Sperner [377]. Theorem 2.4.5 is due to Brouwer [75]. The approach in this section for proving the Brouwer Fixed-Point Theorem is due to Knaster, Kuratowski and Mazurkiewicz [223]. A stronger result than Exercise 2.4.5 is due to Schauder [361]. In fact, Schauder claimed to have proved that every compact convex set in a linear space has the fixed-point property. But an inspection of
C. NOTES AND COMMENTS
583
his proof showed that it is valid for locally convex linear spaces only. For decades it was a formidable open problem whether the local convexity condition in Schauder's Theorem could be dropped. It was shown only recently by Cauty [89] that this is indeed the case. For more details, see also Dobrowolski's very informative notes in [119]. Exercises 2.4.8 and 2.4.9 are due to Borsuk [66]. The converse is also true. That is, if a compact subspace A C W1 does not separate Rra then every continuous function /: A —> S ra-1 is nullhomotopic. Similarly for S n . ^§2.5: The Borsuk-Lusternik-Schnirelman Theorem is due to Borsuk [67] and Lusternik and Schnirelman [263]. For more information, see the very informative notes in Dugundji and Granas [143, p. 171]. *§3.1: Motivated by work of Poincare [337, 338], Brouwer [77] presented the first definition of a topological dimension function. See §3.10 for details. Corollary 3.1.2 is due to Brouwer [77]. The covering dimension dim in terms of finite open covers was formally introduced in Cech [94]. Our definition of dim is usually called 'The Theorem on Partitions' and was proved in Eilenberg and Otto [145]. The results in §3.1 are well-known. ^§3.2: The proof of Theorem 3.2.5 consists mostly of small modifications of arguments that can be found in Engelking [155, Chapter 1]. Theorem 3.2.8 for the case of the small inductive dimension function ind (which takes the same value as dim, cf. Theorem 3.4.10) was proved for compact spaces by Menger [282] and by Urysohn [398]; finally, it was established in full generality by Tumarkin [393] and Hurewicz [199]. The Subspace Theorem 3.2.9 is a triviality for the small inductive dimension function ind. Theorem 3.2.12 and Corollary 3.2.13 are due to Brouwer [77]. Corollary 3.2.13 is also due to Brouwer [77]. Theorem 3.2.14 is due to Dowker [137] and Kuratowski [238]. Exercise 3.2.1 shows that within the class of separable metrizable spaces, the dimension cannot be raised by the adjunction of a single point. If one drops the separability condition, then this is no longer true, as was shown by van Douwen [125], and independently by Przymusinski [348]. Almost zero-dimensional spaces were introduced by Oversteegen and Tymchatyn [330]. They proved that almost zero-dimensional spaces are at most one-dimensional, and used this result to conclude that the homeomorphism groups of various interesting spaces such as Sierpiriski's Carpet and Menger's Universal Curve, are one-dimensional. For a simpler proof that almost zero-dimensional spaces are at most one-dimensional, see Levin and Pol [254] and Theorem 3.5.10. ^§3.3: That every n-dimensional space can be imbedded in R 2n+1 is due independently to Nobeling [327], Pontrjagin and Tolstowa [347] and Lefschetz [247]; that 9tn is universal for n-dimensional spaces is due to Nobeling [327]. Corollary 3.3.8 is due to Hurewicz [200]. Corollary 3.3.12 is due to Tumarkin [393].
M3.4: That dim, ind and Ind take the same values, is a consequence of the fact that we
584
C. NOTES AND COMMENTS
deal here with separable metrizable spaces exclusively. Even for the well-behaved class of metrizable spaces this does not hold in general. The famous result of Roy [355, 356] shows that for completely metrizable spaces, ind and dim need not agree (for recent developments, see Mrowka [317, 318] and Kulesza [235]). On the other hand, dim and Ind take the same values on metrizable spaces, as was shown independently by Katetov [216] and Morita [314]. The small inductive dimension function ind was first defined by Urysohn [396] and Menger [281]. The Addition Theorem 3.4.4 is due, for compact spaces, to Urysohn [398] and for general spaces to Tumarkin [393] and Hurewicz [199]. The large inductive dimension function Ind is related to Brouwer [77] but was formally first defined by Cech [93]. The Coincidence Theorem 3.4.10 is partly due to Hurewicz [200, 198], Brouwer [78], Menger [282], Urysohn [398] and Tumarkin [393]. Theorem 3.4.12 is due to Menger [283]. An example with similar properties as Erdos' space in Example 3.4.13 cannot be compact. Hurewicz [204] proved that if X is compact and Y is an arbitrary space and dimX < 1 then dim(X x y) — dimX + dim Y (see also [155, Problem 1.9(E)]). This result is 'best possible'. There exist compact spaces X and Y such that dimX = dimY = 2, while dim(X x Y) = 3, Pontrjagin [346]; see also Kodama [224]. Examples as the one in Example 3.4.13 exist in every dimension, as was shown by Anderson and Keisler [21]. Their spaces are not complete however. Topologically complete spaces with the same properties were constructed later by Kulesza [234]. For more information on dimension modulo a class of spaces, see Aarts and Nishiura [4]. The compactness degree and compactness deficiency are notions due to de Groot [176]. Exercises 3.4.2 and 3.4.3 are due to de Groot [176]. (See also de Groot and Nishiura [179].) He raised the problem whether cmpX = def X is true for all spaces X. This problem, posed in 1942, was finally solved in the negative by Pol [341] in 1982 (a simpler example was recently constructed by Levin and Segal [256]). See Aarts and Nishiura [4] for a comprehensive study of de Groot's compactification problem. It can be shown that the space Y in Exercise 3.4.4 also has the property that cmpY = n. See [4, Example 5.10(d)] for details. ^§3.5: The results on compactifications in this section are well-known. They follow for example from the work of Kuratowski [241]. The part of Corollary 3.5.7 not dealing with colorability is due to Engelking [152]. Theorem 3.5.8 and Corollary 3.5.9 are due to Pol and Levin [254]. Theorem 3.5.10 is due to Oversteegen and Tymchatyn [330]. Their proof, based on the notion of .R-trees, is rather complicated. The proof presented here is due to Levin and Pol [254]. It is not the most economical proof of Theorem 3.5.10. The proof in [254, §2] is shorter and more elegant and avoids the use of Corollary 3.5.9. However, Corollary 3.5.9 is of independent interest; its full strength will be used in §3.11 as well. Exercise 3.5.2 is due to Nagami [321]. Exercise 3.5.6 is due to van Mill and Pol [308] ^§3.6: Theorem 3.6.5 is basically due to Alexandroff [9]; contributions were also made by Hurewicz [205] and Hurewicz and Wallman [208]. Theorem 3.6.8 is due to Brouwer [76]. The proof presented here was taken from Hurewicz and Wallman [208]. Theorem 3.6.9 is due to Curtis and was published in [296]. Theorem 3.6.10 is for
C. NOTES AND COMMENTS
585
compact spaces due to Hurewicz [201] and for general spaces to Hurewicz and Wallman [208]. The fact that open maps on compact spaces can raise dimension is due to Kolmogoroff [225]. Anderson [13] proved that every Peano continuum is an open continuous image of the universal Menger curve. Theorem 3.6.14 is a corollary of this result. The proof presented here was brought to my attention by Uspenskii. It is due to Dranisnikov and is close to a construction of Kozlovskii [226], see also Fedorchuk [159, p. 186]. Exercise 3.6.1 is due to Dijkstra [113]. Exercise 3.6.4 is due to Borsuk [69]. Theorem 3.7.1 and Corollaries 3.7.2 and 3.7.3 are due to Menger [282] and Urysohn [397]. Proposition 3.7.4 can be found in Rubin, Schori and Walsh [357]. Theorem 3.7.6 is due to Mazurkiewicz [277]. The proof presented here is due to Rubin, Schori and Walsh [357]. The notion of a Cantor-manifold was introduced by Urysohn [397]. Theorem 3.7.8 was proved independently by Hurewicz and Menger [207] and by Tumarkin [394]. The proof presented here is due to Hurewicz [206]. A similar proof was given by Freudenthal [168]. Exercise 3.7.4 is due to Hart, van Mill and Pol [188]. The interesting Theorem 3.8.1 is due to Bing [57]. It is difficult to follow Bing's arguments. The proof presented here can be found in Hart, van Mill and Pol [188]. It is in fact close to Bing's original reasoning but is much easier to follow; variations of it circulated among various specialists in continua theory. The first Henderson compactum was constructed by Henderson [194]. Simpler examples can be found in Henderson [195], Bing [60], Zarelua [412], Rubin, Schori and Walsh [357] and Levin [252]. See §3.13 for information on stronger results. The proof of Corollary 3.8.8 is due to Levin [252]. Proposition 3.8.9 is due to Bing [57]. A nonmetrizable infinite-dimensional homogeneous indecomposable continuum was constructed by van Mill [299]. Theorem 3.9.2 is due to Bourbaki [72, p. 144, Exercise 9a]. The proof presented here is independently due to Kulesza [234] and van Mill [298, Theorem 4.7.3] and is inspired by Parthasarathy [332, Theorem 4.1]. The first topologically complete one-dimensional totally disconnected space was constructed by Sierpifiski [370]. Theorem 3.9.3 is due to Mazurkiewicz [276]. The proof presented here is due to Rubin, Schori and Walsh [357] and Pol [340]. The technique in the proof goes back to Mazurkiewicz [276] and Knaster [222] and has been used by several authors: see Lelek [250], Zarelua [412], Rubin, Schori and Walsh [357], Kulesza [234], Ivanov [210], Pol [343], van Mill and Pol [303] and Krasinkiewicz [228, 229]. These papers contain further references. The interesting Theorem 3.9.5 is due to Kulesza [234]. Without the completeness condition, it was proved earlier by Anderson and Keisler [21]. Theorem 3.10.3 is due to Fedorchuk, Levin and Shchepin [161]. Theorem 3.10.4 is due to Fedorchuk and van Mill [162]. Every n-dimensional space X can be imbedded in R 2 ™ +1 by Theorem 3.3.5 and hence in a product of 2n+l one-dimensional spaces.
586
C. NOTES AND COMMENTS
The number 2n + 1 is not optimal. It was proved by Lipscomb [262] that every ndimensional space imbeds in a product of a family of at most n+l one-dimensional spaces. See also Sternfeld [381], Bowers [73] and Olszewski [329]. Nagata [323, p. 163] asked whether every n-dimensional space can be imbedded in a product of at most n one-dimensional spaces. This question was answered in the negative by Borsuk [71]. He proved (essentially) the result stated in Exercise 3.10.5. The solution to this exercise presented here which is simpler than Borsuk's original arguments, is due to van Mill and Pol [306]. H3.ll: The proof of Lemma 3.11.5 presented here seems to be new. It follows immediately from Hurewicz's Theorem (Kuratowski [243, §45, statement IX]) that the zero-dimensional maps g: K —> ln+l are dense in the function space C(K, I n+1 ) (if p : I n+1 —>• I is the projection and g: K —> I n+1 is zero-dimensional then dim[(p o g)'1 (t)] < n for all t G I). Alternatively, one can also use Nagata's metric on K (Nagata [325, Theorem V.4]). The first examples of weakly n-dimensional spaces were given by Sierpiriski [370] and Mazurkiewicz [277]. A simpler construction can be found in Tomaszewski [389]. Theorem 3.11.8 is due to van Mill and Pol [303]. Theorem 3.11.10 is due to van Mill and Pol [307]. That the product of two weakly one-dimensional spaces is one-dimensional, is due to Tomaszewski [389]. Corollary 3.11.12 is due to van Mill and Pol [308]. The proof presented in the exercises of the fact that the product of two weakly one-dimensional spaces is one-dimensional, is basically due to to Tomaszewski [389]. (See van Mill and Pol [307].) ^§3.12: The notion of a coloring of a map is motivated on the one hand by the LusternikSchnirelman-Borsuk Theorem (see §2.5) and on the other hand by results in combinatorial set theory due to de Bruijn and Erdos [81], Katetov [217], and others. See also Frolik [170]. Theorem 3.12.1 is due to van Douwen [131]. The Ellentuck topology on [N]w was defined by Ellentuck in [147]. Corollary 3.12.6 of which the proof was taken from Kechris [218, §19], is a special case of the so-called Galvin-Prikry Theorem. The same conclusion can be derived if the Pi's are merely Borel subsets of [N]" . See Galvin and Prikry [171] and [218, §19] for details. Theorem 3.12.7 was first stated in [231, Theorem 3.4] and is due to Mazur. Theorem 3.12.8 is a special case of a result of van Douwen [131]. He showed that if X is finite dimensional and if /: X —>• X is closed, continuous and fixed-point free such that sup-d/" 1 ^)! : x e X} < oo then / can be colored. (Theorem 3.12.7 shows that this result is sharp.) Theorem 3.12.10 is due to van Mill [301]. The proof of Theorem 3.12.12 makes use of a technique independently due to Krawczyk and Steprans [231, Lemma 2.1] and Blaszczyk and Kim [62]. Corollary 3.12.15 is due to Aarts, Fokkink and Vermeer [3]. The proof presented here is due to van Mill [301]. As was already remarked in Remark 3.12.13, it is sharp for all n by results of Steinlein [380, 379]. It is of interest to note that it can be shown that if / is a fixed-point free involution on an at most n-dimensional space then / can be colored with at most n + 2 colors. For this the central thing one needs to verify for involutions is that in Theorem 3.12.10
C. NOTES AND COMMENTS
587
the n + 3 can be reduced to n + 2. For a generalization of Theorem 3.12.15 to nonmetrizable spaces, see van Hartskamp and Vermeer [190]. See also van Hartskamp [189]. The inverse limit trick used in the proof of Theorem 3.12.17 is due to Pol. Theorem 3.12.18 is due to Aarts and Fokkink [2]. ^§3.13: Theorem 3.13.7 is due to Lelek [249]. The proof presented here is due to Engelking and Pol [156]. The solution of Alexandroff's Problem presented in Theorem 3.13.8 is due to Pol [340]. Theorem 3.13.10 is due to Walsh [403]. The proof presented here was taken from Pol [343]. The first example of a Henderson compactum is due to Henderson [194], cf. §3.8. It is an interesting question whether there is a Henderson compactum containing a subspace with positive but finite dimension. The existence of such an example was established by Chatyrko and Pol [97]. There are even hereditarily indecomposable such spaces, as was recently proved by Pol [339]. The subspace with positive but finite dimension is one-dimensional in both examples. The problem whether similar spaces can be constructed containing subspaces with arbitrarily high finite dimension seems to be open. A stronger result than Exercise 3.13.2 is due to Rubin [358]. In Theorem 3.13.7 we saw that every topologically complete space can be compactified through the addition of a countable dimensional remainder. Exercise 3.13.4 shows that the completeness assumption is essential. In view of Corollary 3.5.7 the question naturally arises whether every homeomorphism of a topologically complete space can be 'compactified' through the addition of a countable dimensional remainder. The answer to this question is in the negative. Let us represent s by Rz and let a: s —>• s be the 'left shift' on s, i.e., a(x)i = x;+i for i G Z. It was shown by Dijkstra and van Mill [115] that if 75 is a compactification of s such that a can be extended to a continuous function a: 7.5 —>• 75 then 75 \ s contains strongly infinite-dimensional continua. Adjan and Novikov [5] constructed (answering a question of Lusin) the first upper semicontinuous function on I that is not countably continuous. A similar construction was performed also by Sierpiriski [373]. Jackson and Mauldin [211] proved, using notions from recursion theory, that Lebesgue measure A on the hyperspace 21 is not countably continuous (it is easy to see that it is upper semi-continuous). For a direct proof of this interesting fact avoiding recursion theory, see van Mill and Pol [305]. Exercise 3.13.5 is also due to them. ^§3.14: The result in §3.14 that the existence of a two-dimensional hereditarily indecomposable continuum implies the existence of an infinite-dimensional hereditarily indecomposable continuum is due to Kelley [220]. (See Levin and Rogers [255] for an improvement.) His method was used by Levin and Sternfeld [257, 258] to prove the interesting result that the hyperspace of nonempty subcontinua of an at least two-dimensional compactum is infinite-dimensional. The observation in §3.14 that Brouwer in I3 implies Brouwer in all dimensions is due to Hart, van Mill and Pol [188]. Exercise 3.14.1 is due to Rogers [353]. *§4.1: Most of the results in this chapter are well-known. Absolute (Neighborhood) Retracts were first defined by Borsuk [64, 65]. For a much more comprehensive study
588
C. NOTES AND COMMENTS
of the subject, see Borsuk [70], Hu [197] and van Mill [298]. Theorems 4.1.1 and 4.1.13 are due to Manner [187] and Theorem 4.1.9 is due to Dugundji [141] arid Lefschetz [248]. Our notion of ANR-pair seems to be new. There are similar notions in the literature but they are different from ours, see e.g., Mardesic and Segal [274], Moszynska [316] and van der Bijl [54]. Our notion was motivated by Toruhczyk's concept of a locally homotopy negligible set from [390]. It was a famous open problem due to Borsuk whether every compact ANR has the homotopy type of a (compact) polyhedron. The answer to this problem is in the affirmative, as was shown by West [406]. Theorem 4.2.1 is essentially due to Dugundji [141] and Lefschetz [247]. Our proof of Theorem 4.2.1 follows the exposition of Hu [197, Chapter IV §4]. Theorem 4.2.14 is essentially due to Toruriczyk [390]. Lemma 4.2.22 is due to Curtis and Nhu [106]. Wojdyslawski [411] proved that the hyperspace of every Peano continuum is an AR and conjectured that in fact 2X K Q if and only if X is a Peano continuum. In [364], Schori and West verified this conjecture in the case X = I. Their result was used in [363] to prove that 2° w Q for every connected graph. Based on these results of Schori and West, in [107] Curtis and Schori verified Wojdyslawski's conjecture. The proof that 2X has the disjoint-cells property presented in this section is implicit in Curtis and Nhu [106]. Using the existence of convex metrics on Peano continua it is possible to give different proofs of this fact (Toruriczyk [391] and Curtis (private communication)). Corollary 4.2.24 is due to Curtis [105]. The proof presented here, based on our notion of an ANR-pair, seems to be new. Theorem 4.2.25 is due to Toruriczyk [391]. He not only topologically characterized Q, but also manifolds modeled on it. In addition, he proved a corresponding result for t manifolds in [392] (see also Bestvina, Bowers, Mogilski and Walsh [52]). There are similar results for finite dimensional manifolds, but they are not as elegant as their infinite-dimensional counterparts since delicate algebraic obstructions enter the picture. See Daverman [109] for details. Finally, there are also similar results for manifolds defined on Menger compacta by Bestvina [51]. Theorem 4.2.30 is due to Kuratowski [240]. Theorems 4.3.1 and 4.3.5 are due to Hanner [187]. Our use of cones simplifies the proof of Theorem 4.3.5 slightly. Theorem 4.3.7 and its corollary 4.3.8 are due to Toruriczyk [390]. The proof presented here is due to Dobrowolski and Marciszewski [122]. Infinite-dimensional topology is the creation of Anderson. Several books were written on the subject, or deal with aspects of infinite-dimensional topology. See e.g., Chapman [96], Bessaga and Pelczyriski [50], van Mill [298], Chigogidze and Fedorchuk [160], Dijkstra [112], Gladdines [175], Chigogidze [98] and Banakh, Radul and Zarichnyi [41]. The highlights of infinite-dimensional topology are the theorems of Anderson [15] on the homeomorphy of I2 and s, of Chapman [95] on the invariance of Whitehead torsion, of West [406] on the finiteness of homototopy types of compact ANR's and of Toruriczyk [391, 392] on the topological characterization
C. NOTES AND COMMENTS
589
of manifolds modeled on various infinite-dimensional spaces. A large collection of open problems is West's paper [407]. The subjects that are being touched upon there range from absorbing sets and function spaces to ANR-theory. The central concept of a Z-set in infinite-dimensional topology is due to Anderson [16]. Intuitively, Z-subsets of X are 'small' in X. Here smallness has a different meaning than being 'small' with respect to category or measure. It is 'small' in the sense of homotopy (this will be made precise in Exercise 5.1.7). The notion of an A-set (the terminology is due to the author) is due to Anderson [16]. In the early days of infinite-dimensional topology, this was the central concept. But it soon became clear that it is equivalent to the notion of a Z-set defined in this section which turned out to be simpler to deal with in 'practical' situations. ^§5.2: Theorem 5.2.4 is due to Anderson [16] and Barit [43]. The technique used in the proof is motivated by work of Klee [221]. The solution to Exercise 5.2.1 by elementary means is not simple. It is a consequence of a known and stronger homeomorphism extension result than the one we derived in this section. However, the proof of this result requires the development of some nontrivial facts that play no further role in this book. So we decided to skip that. See Anderson and Chapman [19], Chapman [96] and van Mill [298, §7.4] for details. ^§5.3: Theorem 5.3.7 is due to Anderson [16] and Barit [43]. The other results in this section are standard applications of their result. Exercise 5.3.3 is due to Anderson [16]. It was asked in Anderson and Bing [18, Question 17] whether the condition QoCiQi is a Z-set both in Qo and Qi in this result can be dropped. For some classes of Hilbert cubes, the answer was subsequently shown to be affirmative. For example, if both Qo and Qi are Keller cubes in I2 (Quinn and Wong [350] and Mogilski [311]). It was shown by Handel [186] that the condition can be relaxed to Qo^Qi is a Z-set in Qo, or in Q\. This is of interest since it immediately implies the following old result of Anderson [14]: the product of the letter T and Q is a Hilbert cube. That is, T is a so-called Hilbert cube factor, i.e., a space whose product with the Hilbert cube is the Hilbert cube. It was finally shown by Sher [366] that the answer to the Anderson-Bing question is in the negative. *§5.4: The notion of a capset is independently due to Anderson [12] and Bessaga and Pelczyriski [49]. Many results in this section are standard. For references, see e.g., Bessaga and Pelczyriski [50]. A subset B G Za(Q) is called a boundary set provided that Q \ B w s. This notion is due to Curtis [104] who characterized the way boundary sets are topologically placed in Q. The standard example of a boundary set is B(Q). For 'pathological' examples of boundary sets, see e.g. Henderson and Walsh [196] and van Mill [295]. The statement in Corollary 5.4.2 comes close to saying that a capset is a so-called absorber or absorbing set in Q for the class of all compact spaces. Absorbers are related to capsets and were first defined by West [405]. ( See §5.5 for more information.) The influential paper by Bestvina and Mogilski [53] revived the interest in absorbers. Since then they are used quite effectively to prove the homeomorphy of various incomplete linear spaces. They are now
590
C. NOTES AND COMMENTS
a major tool in infinite-dimensional topology. See the monographs by Dijkstra [112] and Banakh, Radul and Zarichnyi [41] for more information. Theorem 5.4.9 is due to Anderson [12]. The interesting Theorem 5.4.10 is due to Curtis [104]. The Capset Characterization Theorem proved in Corollary 5.4.11 is also due to Curtis [104]. It was claimed earlier in Kroonenberg [233] that it is enough to replace condition (iv) by the weaker statement that for every t > 0 there exists for some n a map r\: Q —> Bn with 0(?y, 1) < e. But this is not true, as was shown by Anderson, Curtis and van Mill [20]. The argument given in Kroonenberg [233] breaks down at the attempted application of the Homeomorphism Extension Theorem 5.3.7 for the copies Bn of Q, using the restriction of the metric Q on Q, and these restrictions may be highly nonconvex. Theorem 5.4.20 and Corollary 5.4.21 are due to Kroonenberg [233]. They were the first results of this type. Subsequently, many similar results were obtained by a variety of techniques, see e.g., Curtis [105] and Curtis and Nhu [106] for details. Exercise 5.4.6 is due to Anderson [17]. ^§5.5: The results in this section are due to Dijkstra, van Mill and Mogilski [117]. For related results, see Bestvina and Mogilski [53]. A generalization of Theorem 5.5.7 can be found in Dobrowolski and Rubin [124]. Lemma 5.5.14 is an adaptation to our needs of Proposition 3.2 in Dobrowolski, Gul'ko and Mogilski [120]. A stronger result than Corollary 5.5.16 was first proved independently by Dobrowolski, Gul'ko and Mogilski [120] and Cauty [86] using the Bestvina-Mogilski approach from [53]. We will come back to this in Chapter 6. The first result in the spirit of Corollary 5.5.16 seems to be van Mill [297]. ^§6.1: The largest compactification of a space X is its so-called Cech-Stone compactification J3X. For more information on (3X, see e.g., Gillman and Jerison [174], Comfort and Negrepontis [102]. ^§6.2: Theorem 6.2.5 is due to Arhangel'skii [24]. Banakh and Cauty [40] proved that if / is countable and Cp(Xi) is analytic for every i G / then for X — ® i6/ Xi we have
Moreover, examples in Marciszewski and van Mill [272] show that if / is countably infinite then (*) in general is not true. ^§6.3: It was shown in Lutzer and McCoy [264] that if CP(X) is a G^-subset of Rx then X is discrete. The stronger result mentioned in Remark 6.3.5 and its simple proof are due to Dijkstra, Grilliot, Lutzer and van Mill [114]. Theorems 6.3.6 and 6.3.8 are also due to these authors. The simple proof of Theorem 6.3.8 presented here is due to van Mill [300]. Theorem 6.3.8 was used in Cauty, Dobrowolski and Marciszewski [91] to prove the interesting result that if X is countable and CP(X] is a Borel subset of RA then it is of multiplicative Borel class. For more results on the Borel complexity of function spaces, see e.g., Calbrix [83, 84], Lutzer, van Mill and Pol [265].
C. NOTES AND COMMENTS
591
A partial characterization of CP(X) being a Baire space is due to Lutzer and McCoy [264]. Theorem 6.4.3 is due independently to van Douwen [135], Pytkeev [349] and Tkachuk [388]. Our exposition of the proof of Theorem 6.4.3 closely followed McCoy and Ntantu [280, Theorem 5.3.8]. That there is a countable dense-in-itself space X such that CP(X) is Baire is due to Michalewski [289, Page 22]. The interesting Theorem 6.5.3 is due to Talagrand [384]. With the aid of topological game theory, results in the same spirit were obtained by Lutzer and McCoy [264]. Their proofs were simplified by Dobrowolski, Marciszewski and Mogilski [123]. Theorem 6.5.3 in the context of function spaces was used for the first time in Dobrowolski and Marciszewski [121]. Theorem 6.12.1 and Corollary 6.5.4 and their consequences are due to Dobrowolski, Marciszewski and Mogilski [123]. That C*P(I%) is a Baire space for an ultrafilter p is due to Lutzer and McCoy [264]. ^§6.6: Theorem 6.6.2 is essentially due to Dugundji [140]. Proposition 6.6.6 and Corollaries 6.6.8 and 6.6.10 are folklore. They were probably first proved by Arhangel'skii. The results in this section are folklore. They were used quite frequently in the Russian literature without explicit references and proofs, see e.g., Pavlovskii [333] and Pestov [336]. For another approach, see Baars and de Groot [35, §1.3]. See also Arhangel'skii [27]. The concept of support of a point is due to Arhangel'skii [24]. The alternative description of supports for continuous linear functions is due independently to Arhangel'skii and Pelant. It was of crucial importance to Baars and de Groot [35] in their analysis of function spaces of countable spaces. Most of Lemma 6.8.2 is due to Baars and de Groot [35, Chapter 1]. Theorem 6.8.3 and its Corollary 6.8.4 were proved by Arhangel'skii [24]. Corollary 6.8.5 is due to Baars and de Groot [35, Lemma 1.2.11]. Proposition 6.8.6 for normal spaces is Lemma 1.4.5 in Baars and de Groot [35]. The suggestion that it is also true for arbitrary spaces was made to the author by Marciszewski. ^§6.9: Theorem 6.9.1 are Corollary 6.9.2 are due to Arhangel'skii [24]. The interesting Theorem 6.9.4, which is due to Pestov [336], is in fact true for the class of all spaces, and not only for separable metrizable spaces. It has motivated the natural question whether dimension is preserved by ^-equivalence. This question is unsolved. See Remark 6.9.5 for more details. Theorem 6.9.7 is due to Baars, de Groot and Pelant [37]. For generalizations, see Marciszewski and Pelant [273] and Marciszewski [271]. It is also true, see Baars, de Groot and Pelant [37], that if X and Y are metrizable and p: C*(X) —>• CP(Y) is a continuous linear surjection then X completely metrizable implies Y completely metrizable. The proof of this is much more complicated than the proof of its corresponding result Theorem 6.9.7. *§6.10: The results in this section are taken from Baars, de Groot, van Mill and Pelant [36].
592
C. NOTES AND COMMENTS
The proof of Theorem 6.10.8 is a generalization of a method due to Pelant [334]. For other results in the same spirit, see e.g., Baars and de Groot [33, 34, 35]. *§6.11: It is an intriguing open problem whether ^-equivalent spaces have the same dimension. Marciszewski [271] proved the interesting Theorem 6.11.3, modifying Cauty's technique from [90] and using an idea from Gul'ko [182]. *§6.12: It is an interesting question which topological properties of the filter 3~ (as subspace of y(N)) translate into 'nice' topological properties of the function space cy. There are many of such results. We will mention only the result of Calbrix [83, 84] that cy is an absolute Borel set if and only if J is. The proof of Theorem 6.12.15 presented here contains elements that can be found in Dobrowolski, Marciszewski and Mogilski [123], Baars, Gladdines and van Mill [32] and Dijkstra and Mogilski [118]. However, the proof in the latter two papers has a gap. This gap was bridged in the paper Dijkstra and van Mill [116]. As we remarked at the beginning §6.12, the consequence of Theorem 6.12.15 that all Fa$ function spaces are homeomorphic was first shown by Dobrowolski, Marciszewski and Mogilski [123]. ^§6.13: Example Ell is due to Dijkstra, Grilliot, Lutzer and van Mill [114]. That there is an almost disjoint family of c subsets of u; is due to Sierpinski [371]. The Disjoint refinement Lemma 6.13.5 is a well-known result from set theory. It is in fact true in a much more general setting. The example A is due to van Mill and Pol [304]. It is based on a space constructed earlier in van Douwen and Pol [136]. For an example with similar properties, see Marciszewski [267]. *§A.l: Our notation is standard and all the results in this chapter are classical. For information on ultrafilters, see Comfort and Negrepontis [102]. Exercise A. 1.2 is due to van Douwen [130]. A space without isolated points and satisfying one of the equivalent statements in Exercise A. 1.2 is called perfectly disconnected. It is not clear at all that such spaces exist. They were first defined and constructed by van Douwen [134, 130]. It will be shown in §6.4 that a countable dense-in-itself perfectly disconnected space X has the property that its function space CP(X) is Baire. So X is not Baire but its function space is Baire, which is rather curious. Exercise A. 1.4 is usually called the A-System Lemma. See Juhasz [214] for this and many other related results. Exercise A. 1.10 and its solution are well-known. *§A.2: Exercise A. 2. 13 is the so-called Cantor-Bendixson Theorem which was proved by Cantor and Bendixson independently in 1883 for subsets of R. It is a natural question to ask when the limit of a sequence of continuous functions exists and is continuous. This basic question is studied in this section. Lemma A. 4.1 can be viewed to be some sort of an extension result. For let Y be
C. NOTES AND COMMENTS
593
the subspace A U B of X. It is clear that both A and B are clopen in Y. So the function / : Y —» I defined by / \ A = 0 and / \ B = I is continuous. Hence Lemma A. 4.1 tells us among other things that / can be extended to a continuous function a: X —> I. The possibility of extending continuous functions is very important in topology. We will come back to this in §1.2. Theorem A. 4. 6 can be found in Tietze [387] and Urysohn [399]. See also HausdorfT [191] for a related approach. In Corollary A. 4. 4 we proved in fact that every regular Ti-space with a countable base can be imbedded in Q and hence is metrizable, being homeomorphic to a subspace of a metrizable space. This is usually called Urysohn's Metrization Theorem. The most famous and important metrization theorem is the one due to Nagata, Smirnov and Bing. See e.g., Engelking [153, Chapter 4] for more details. The Alexandroff one-point compactification was first considered in Alexandroff [7]. Exercise A. 4. 4 is due to Berney [47].
§A.5: The notion of compactness is one of the most important in topology. As proved in this section, in the realm of metrizable spaces compactness can be characterized in terms of sequences. This simplifies life considerably (see for example the proof of Corollary 1.11.4). Lemma A. 5. 3 is essentially due to Lebesgue [246]. It was shown independently by Bing [56] and Moise [312] that every Peano continuum admits an admissible convex metric. This is a highly nontrivial result. So Exercise A. 5. 20 follows from their result. Theorem A. 6.1 is due to Frechet [166]. Theorem A. 6. 3 is due to Alexandroff [6] and Kuratowski [242]. The important class of Baire spaces was introduced by Bourbaki bourbaki. Theorem A. 6. 6 is due to Hausdorff [193]. Theorem A. 6. 10 and its generalization Exercise A. 6. 11 are due to Oxtoby [331].
§A.7: The results in this section are standard. Corollary A. 7. 6 is a classical result due to Dowker [138], Katetov [215] and Dieudonne [111]. §A.8: Lemma A. 8.1 is due to Kuratowski [243, §15, XIII, p. 122]. Lemma A. 8. 3 was proved independently by Taimanov [385] and by Eilenberg and Steenrod [146]. Theorem A. 8. 5 is due to Lavrentieff [245]. See van Douwen [132] for a partial converse to his theorem. In Engelking [153, Theorem 4.3.21] the reader can find more information on Lavrentieff 's Theorem. The solution to Exercise A. 8.1 is due to van Douwen [132]. Exercise A. 8. 3 is implicit in Marciszewski [271]. §A.9: Wallman compactifications were first defined by Frink [169], who was inspired by the work of Wallman [402]. In this section we discussed separable metrizable Wallman compactifications only, but from the construction it is clear that there also is a theory for general (Tychonoff) spaces. Frink [169] asked whether every compactification is a Wallman compactification. This was a central problem for some time, until it was solved in the negative by Ul'janov [395]. That every metric compactification is a Wallman compactification (Corollary A. 9. 8), is due independently to
594
C. NOTES AND COMMENTS
Aarts [1] and Steiner and Steiner [378]. Different proofs of the same result were given by Berney [47], van Douwen [127], Bandt [42], and others. We closely followed Berney's proof. >§A.10: Theorem A. 10. 6 is due to Sierpiriski [368]. It has a nontrivial generalization due to Dijkstra [113]. See Exercise 3.6.1 for more details. The compactness of the the Cn's in Proposition A. 10. 7 is essential, see Example 1.10.11. Exercise A. 10. 6 is due to Moore [313]. Exercise A. 10. 7 is due to Urysohn. For a description of the troubles this example caused in the early days of dimension theory, see e.g., Johnson [213]. It can even be shown that the set X \ \J 3" in Exercise A. 10. 8 has cardinality c. For the set X \ [J y is topologically complete, and an uncountable topologically complete space has size c (Corollary 1.5.13). The notion of adjunction space discussed in this section is very important in topology. We did not discuss it in its most general form. For more details, see Dugundji [142]. Null-sequences are used extensively in topology for creating decomposition spaces with various interesting properties. See e.g., Daverman [109] for more information. The material in this section is folklore. The approach to some of the results here is due to Burckel [82]. Theorem A. 12. 8 remains true for any space X. This is due to Eilenberg [144]. It can be shown that every cube I n has the fixed-point property. This is the Brouwer Fixed-Point Theorem, see §2.4. Borel subsets of R were introduced by Borel. The results in this section are all classical. The proof of Theorem A. 13. 5 is due to van Engelen [148, Lemma A. 2. 4]. In view of Corollary A. 13. 12, it is a natural question whether every analytic subset of a topologically complete space is Borel. This is nontrivial and was answered in the negative by Souslin. See Hurewicz [203] and Mazurkiewicz [279] for 'explicit' examples of such spaces. The classical proof of Theorem A. 13. 13 goes as follows. First one proves that the Baire property is invariant under the so-called A- operation (this is due to Nikodym [326]), and then one proceeds to prove that the analytic subsets of a topologically complete space X coincide with the sets that one obtains by performing the yi-operation on all closed subsets of X. For details, see e.g., Kuratowski and Mostowski [244]. The direct proof of Theorem A. 13. 13 presented here is motivated by these classical constructions. Call a subset of a space X coanalytic if its complement is analytic. Since C(X) is a a-algebra, it contains every coanalytic set by Theorem A. 13. 13. Since there are coanalytic spaces which are not analytic (see e.g., Kuratowski and Mostowski [244] or Kechris [218]), this shows that the converse to Theorem A. 13. 13 does not hold, i.e., that in general C(X) does not coincide with the collection of all analytic subsets of X. Exercise A. 13. 4 is a special case of a result due to Levi [251]. In Exercise A. 13. 6 we proved that there are subsets of R which are not Borel. If X = N with the discrete topology, then clearly all subsets of X are Borel. A more interesting example is X = Q: simply observe that every subset of Q is countable, hence Fa, and hence Borel. In view of
C. NOTES AND COMMENTS
595
this, it is tempting to believe that every uncountable space has a non-Borel subset. A so-called Q-set shows that this need not be true, see e.g., Miller [309].
This page intentionally left blank
Bibliography [1] J. M. Aarts, Every metic compactification is a Wallman-type compactification, Proc. Int. Symp. on Top. and its Appl., Herceg-Novi (Yugoslavia), 1968, Beograd, 1969, pp. 29-34. [2] J. M. Aarts and R. J. Fokkink, An addition theorem for the color number, 1999, to appear in Proc. Amer. Math. Soc. [3] J. M. Aarts, R. J. Fokkink, and J. Vermeer, Variations on a theorem of Lusternik and Schnirelmann, Topology 35 (1996), 1051-1056. [4] J. M. Aarts and T. Nishiura, Dimension and Extensions, North-Holland Mathematical Library, vol. 48, North-Holland Publishing Co., Amsterdam, 1993. [5] S. I. Adjan and P. S. Novikov, On a semicontinuous function, Moskov. Gos. Ped. Inst. Uchen. Zap. 138 (1958), 3-10, In Russian. [6] P. Alexandroff, Sur les ensembles de la premiere classe et les ensembles abstraits, C. R. Acad. Paris 178 (1924), 185-187. [7] P. Alexandroff, Uber die Metrisation der im Kleinen kompakten topologischen Rdume, Math. Ann. 92 (1924), 294-301. [8] P. Alexandroff, Une definition des nombres de Betti pour un ensemble ferme quelconque, C.R. Acad. Paris 184 (1927), 317-319. [9] P. Alexandroff, Dimensionstheorie. Eein Beitrag zur Geometrie der abgeschlossenen Mengen, Math. Ann. 106 (1932), 161-238. [10] P. Alexandroff and P. Urysohn, Uber nuldimensionale Punktmengen, Math. Ann. 98 (1928), 89-106. [11] P. Alexandroff and P. Urysohn, Memoire sur les espaces topologiques compacts, Verh. Akad. Wetensch. Amsterdam 14 (1929). [12] R. D. Anderson, On sigma-compact subsets of infinite-dimensional manifolds, unpublished manuscript. [13] R. D. Anderson, A characterization of the universal curve and a proof of its homogeneity, Annals of Math. 67 (1958), 313-324. [14] R. D. Anderson, The Hilbert cube as a product of dendrons, Notices Amer. Math. Soc. 11 (1964), 572. [15] R. D. Anderson, Hilbert space is homeomorphic to the countable infinite product of lines, Bull. Amer. Math. Soc. 72 (1966), 515-519. [16] R. D. Anderson, On topological infinite deficiency, Mich. Math. J. 14 (1967), 365— 383. [17] R. D. Anderson, Strongly negligible sets in Frechet manifolds, Bull. Amer. Math. Soc. 75 (1969), 64-67. [18] R. D. Anderson and R. H. Bing, A complete elementary proof that Hilbert space is homeomorphic to the countable infinite product of lines, Bull. Amer. Math. Soc. 74 (1968), 771-792. [19] R. D. Anderson and T. A. Chapman, Extending homeomorphisms to Hilbert cube manifolds, Pac. J. Math. 38 (1971), 281-293. 597
598
BIBLIOGRAPHY
[20] R. D. Anderson, D. W. Curtis, and J. van Mill, A fake topological Hilbert space, Trans. Amer. Math. Soc. 272 (1982), 311-321. [21] R. D. Anderson and J. E. Keisler, An example in dimension theory, Proc. Amer. Math. Soc. 18 (1967), 709-713. [22] R. Arens, Extension of functions on fully normal spaces, Pac. J. Math. 1 (1951), 353-367. [23] R. Arens and J. Eels, On embedding uniform and topological spaces, Pac. J. Math. 6 (1956), 397-403. [24] A. V. Arhangel'skii, On linear homeomorphisms of function spaces, Soviet Math. Doklady 25 (1982), 852-855. [25] A. V. Arhangel'skii, A survey of Cp-theory, Questions and Answers in Gen. Top. 5 (1987), 1-109. [26] A. V. Arhangel'skii, Cp-theory, Recent Progress in General Topology (M. Husek and J. van Mill, eds.), North-Holland Publishing Co., Amsterdam, 1992, pp. 1-56. [27] A. V. Arhangel'skii, Topological function spaces, Math. Appl., vol. 78, Kluwer Academic Publishers, Dordrecht, 1992. [28] A. V. Arhangel'skii, Some observations on Cp-theory and bibliography, Top. Appl. 89 (1998), 203-221. [29] W. Arveson, An invitation to C* -algebras, Springer-Verlag, Berlin, 1976. [30] C. E. Aull and R. Lowen, eds., Handbook of the History of General Topology, Kluwer Academic Publishers, Dordrecht, 1997. [31] J. Baars, On the £*-equivalence of certain locally compact spaces, Top. Appl. 52 (1993), 43-57. [32] J. Baars, H. Gladdines, and J. van Mill, Absorbing systems in infinite-dimensional manifolds, Top. Appl. 50 (1993), 147-182. [33] J. Baars and J. de Groot, An isomorphical classification of function spaces of zero dimensional locally compact separable metric spaces, Comrn. Math. Univ. Carolinae 29 (1988), 577-595. [34] J. Baars and J. de Groot, On the (.-equivalence of metric spaces, Fund. Math. 137 (1991), 25-43. [35] J. Baars and J. de Groot, On topological and linear equivalence of certain function spaces, CWI Tract, vol. 86, Centre for Mathematics and Computer Science, Amsterdam, 1992. [36] J. Baars, J. de Groot, J. van Mill, and J. Pelant, An example of tp-equivalent spaces which are not I*-equivalent, Proc. Amer. Math. Soc. 119 (1993), 963-969. [37] J. Baars, J. de Groot, and J. Pelant, Function spaces of completely metrizable spaces, Trans. Amer. Math. Soc. 340 (1993), 871-879. [38] S. Banach, Theorie des operations lineaires, PWN, Warszawa, 1932. [39] T. Banakh. Some properties of the linear hull of the Erdos set in I2, Bull. Polon. Acad. Sci. Ser. Math. Astronom. Phys. 47 (1999), 385-392. [40] T. Banakh and R. Cauty, Universalite forte pour les sous-ensembles totalement homes. Applications aux espaces CP(X), Colloq. Math. 73 (1997), 25-33. [41] T. Banakh, T. Radul, and M. Zarichnyi, Absorbing sets in infinite-dimensional manifolds, Mathematical Studies, vol. 1, VNTL Publishers, Lviv, 1996. [42] C. Bandt, On Wallman-Shanin-compactifications, Math. Nachr. 77(1977), 333-351. [43] B. Barit, Small extensions of small homeomorphisms, Notices Amer. Math. Soc. 16 (1969), 295. [44] R. G. Bartle and L. M. Graves, Mappings between function spaces, Trans. Arner. Math. Soc. 72 (1952), 400-413. [45] H. Becker, F. van Engelen, and J. van Mill, Disjoint embeddings of compacta, Mathematika 41 (1994), 221-232. [46] R. Bennett, Countable dense homogeneous spaces, Fund. Math. 74 (1972), 189-194.
BIBLIOGRAPHY
599
[47] E. S. Berney, On Wallman compactifications, Notices Amer. Math. Soc. 17 (1970), 215. [48] C. Bessaga and A. Pelczyriski, Spaces of continuous functions IV (on isomorphical classification of spaces of continuous functions), Studia Math. 19 (1960), 53-62. [49] C. Bessaga and A. Pelczyriski, The estimated extension theorem homogeneous collections and skeletons, and their application to the topological classification of linear metric spaces and convex sets, Fund. Math. 69 (1970), 153-190. [50] C. Bessaga and A. Pelczyriski, Selected topics in infinite-dimensional topology, PWN—Polish Scientific Publishers, Warsaw, 1975, Monografie Matematyczne, Tom 58. [51] M. Bestvina, Characterizing k-dimensional universal Menger compacta, Mem. Amer. Math. Soc. 71 (1988), no. 380, vi+110. [52] M. Bestvina, P. Bowers, J. Mogilski, and J. J. Walsh, Characterizing Hilbert space manifolds revisited, Top. Appl. 24 (1986), 53-69. [53] M. Bestvina and J. Mogilski, Characterizing certain incomplete infinite-dimensional absolute retracts, Michigan Math. J. 33 (1986), 291-313. [54] J. van der Bijl, Extension of continuous functions with compact domain, Ph.D. thesis, Vrije Universiteit, Amsterdam, 1991. [55] R. H. Bing, A homogeneous indecomposable plane continuum, Duke Math. J. 15 (1948), 729-742. [56] R. H. Bing, Partitioning a set, Bull. Amer. Math. Soc. 55 (1949), 1101-1110. [57] R. H. Bing, Higher-dimensional hereditarily indecomposable continua, Trans. Amer. Math. Soc. 71 (1951), 267-273. [58] R. H. Bing, A homeomorphism between the 3-sphere and the sum of two solid horned spheres, Annals of Math. 56 (1952), 354-362. [59] R. H. Bing, A simple closed curve is the only homogeneous bounded plane continuum that contains an arc, Pac. J. Math. 12 (1960), 209-230. [60] R. H. Bing, A hereditarily infinite dimensional space, Proceedings of the Second Prague Topological Symposium, Academia, Praha, pp. 56-62. [61] R. H. Bing and K. Borsuk, Some remarks concerning topologically homogeneous spaces, Annals of Math. 81 (1965), 100-111. [62] A. Blaszczyk and D. Y. Kim, A topological version of a combinatorial theorem of Katetov, Comm. Math. Univ. Carolinae 29 (1988), 657-663. [63] L. M. Blumenthal and V. L. Klee, On metric independence and linear independence, Proc. Amer. Math. Soc. 6 (1955), 732-734. [64] K. Borsuk, Sur les retractes, Fund. Math. 17 (1931), 152-170. [65] K. Borsuk, Uber eine Klasse von lokal zusarnmenhdngende Rdumen, Fund. Math. 19 (1932), 220-242. [66] K. Borsuk, Uber Schnitte der n-dimensionalen Euklidischen Rdume, Math. Ann. 106 (1932), 230-248. [67] K. Borsuk, Drei Sdtze iiber die n-dimensionale Euclidische Sphdre, Fund. Math. 20 (1933), 177-190. [68] K. Borsuk, Sur les prolongements des transformations continus, Fund. Math. 28 (1936), 99-110. [69] K. Borsuk, Un theoreme sur les prolongements des transformations, Fund. Math. 29 (1937), 161-166. [70] K. Borsuk, Theory of retracts, Paristwowe Wydawnictwo Naukowe, Warsaw, 1967, Monografie Matematyczne, Tom 44. [71] K. Borsuk, Remark on the Cartesian product of two 1-dimensional spaces, Bull. Polon. Acad. Sci. Ser. Math. Astronom. Phys. 23 (1975), 971-973. [72] N. Bourbaki, Topologie Ge.ne.rale. f2e ed.), Hermann, Paris, 1958.
600
BIBLIOGRAPHY
[73] P. Bowers, General position properties satisfied by finite products of dendrites, Trans. Amer. Math. Soc. 228 (1985), 739-753. [74] L. E. J. Brouwer, On the structure of perfect sets of points, Proc. Akad. Amsterdam 12 (1910), 785-794. [75] L. E. J. Brouwer, Uber Abbildung von Mannigfaltigkeiten, Math. Ann. 71 (1912), 97-115. [76] L. E. J. Brouwer, Invariantz des n-dimensionalen Gebiets, Math. Ann. 71—72 (1912— 1913), 305-313; 55-56. [77] L. E. J. Brouwer, Uber den natiirlichen Dimensionsbegriff, J. Reine Angew. Math. 142 (1913), 146-152. [78] L. E. J. Brouwer, Bemerkungen zum natiirlichen Dimensionsbegriff, Proc. Akad. Amsterdam 27 (1924), 635-638. [79] A. L. Brown and A. Page, Elements of functional analysis, Van Nostrand Reinhold, London,1970. [80] M. G. Brown, Some applications of an approximation theorem for inverse limits, Proc. Amer. Math. Soc. 11 (1960), 478-483. [81] N. G. de Bruijn and P. Erdos, A colour problem for infinite graphs and a problem in the theory of relations, Indag. Math. 13 (1951), 371-373. [82] R. B. Burckel, Inessential maps and classical Euclidean topology, Yearbook: Surveys of mathematics 1981, Bibliographisches Inst., Mannheim, 1981, pp. 119-137. [83] J. Calbrix, Classes de Baire er espaces d'applications continues, C. R. Acad. Sci. Paris 301 (1985), 759-762. [84] J. Calbrix, Filtres boreliens sur I'ensemble des entiers et espaces des applications continues, Rev. Roumaine Math. Pure Appl. 33 (1988), 655-661. [85] G. Cantor, Uber unendliche, lineare Punktmannichfaltigkeiten, Math. Ann. 21 (1883), 545-591. [86] R. Cauty, L'espace des fonctions continues d'un espace metrique denombrable, Proc. Amer. Math. Soc. 113 (1991), 493-501. [87] R. Cauty, Un espace metrique lineaire qui n'est pas un retracte absolu, Fund. Math. 146 (1994), 85-99. [88] R. Cauty, La classe Borelienne ne determine pas le type topologique de CP(X), Serdica Math. J. 24 (1998), 307-318. [89] R. Cauty, Solution du probleme de point fixe de Schauder, 1999, Preprint. [90] R. Cauty, Sur I'invariance de la dimension infinie forte par t-equivalence, Fund. Math. 160 (1999), 95-100. [91] R. Cauty, T. Dobrowolski, and W. Marciszewski, A contribution to the topological classification of the spaces CP(X}, Fund. Math. 142 (1993), 269-301. [92] M. Valdivia, Products of Baire topological vector spaces, Fund. Math. 125 (1985), 71-80. [93] E. Cech, Sur la theorie de la dimension, C.R. Acad. Paris 193 (1931), 976-977. [94] E. Cech, Prispevek k theorii dimense, Casopis Pest. Mat. Fys. 62 (1933), 277-291. [95] T. A. Chapman, Topological invariance of Whitehead torsion, Amer. J, Math. 96 (1974), 488-497. [96] T. A. Chapman, Lectures on Hilbert cube manifolds, American Mathematical Society, Providence, R. L, 1976, Expository lectures from the CBMS Regional Conference held at Guilford College, October 11-15, 1975, Regional Conference Series in Mathematics, No. 28. [97] V. A. Chatyrko and E. Pol, Continuum many Frechet types of hereditarily strongly infinite-dimensional Cantor manifolds, Proc. Amer. Math. Soc. 128 (2000), 12071213. [98] A. Chigogidze, Inverse spectra, North-Holland Mathematical Library, vol. 53, NorthHolland Publishing Co., Amsterdam, 1996.
BIBLIOGRAPHY
601
[99] G. Choquet, Lectures on analysis. Vol. I: Integration and topological vector spaces, W. A. Benjamin, Inc., New York-Amsterdam, 1969, Edited by J. Marsden, T. Lance and S. Gelbart. [100] J. P. R. Christensen, Topology and Borel structure, North-Holland Publishing Co., Amsterdam, 1974. [101] P. E. Cohen, Products of Baire spaces, Proc. Amer. Math. Soc. 55 (1976), 119-124. [102] W. W. Comfort and S. Negrepontis, The theory of ultrafilters, Grundlehren der mathematischen Wissenschaften, vol. 211, Springer-Verlag, Berlin, 1974. [103] H. Cook, Continua which admit only the identity mapping onto nondegenerate subcontinua, Fund. Math. 60 (1967), 241-249. [104] D. W. Curtis, Boundary sets in the Hilbert cube, Top. Appl. 20 (1985), 201-221. [105] D. W. Curtis, Hyperspaces of finite subsets as boundary sets, Top. Appl. 22 (1986), 97-107. [106] D. W. Curtis and N. T. Nhu, Hyperspaces of finite subsets which are homeomorphic to ^-dimensional linear metric space, Top. Appl. 19 (1985), 251-260. [107] D. W. Curtis and R. M. Schori, Hyperspaces of Peano continua are Hilbert cubes, Fund. Math. 101 (1978), 19-38. [108] D. van Dantzig, Uber topologisch homogene Kontinua, Fund. Math. 15 (1930), 102125. [109] R. J. Daverman, Decompositions of manifolds, Academic Press, New York, 1986. [110] R. L. Devaney, An introduction to chaotic dynamical systems, Addison-Wesley Publishing Co., Reading, 1989. [Ill] J. Dieudonne, Une generalisation des espaces compacts, J. de Math. Pures et Appl. 23 (1944), 65-76. [112] J. J. Dijkstra, Fake topological Hilbert spaces and characterizations of dimension in terms of negligibility, CWI Tract, vol. 2, Centre for Mathematics and Computer Science, Amsterdam. 1984. [113] J. J. Dijkstra, A generalization of the Sierpinski theorem, Proc. Amer. Math. Soc. 91 (1984), 143-146. [114] J. J. Dijkstra, T. Grilliot, J. van Mill, and D. J. Lutzer, Function spaces of low Borel complexity, Proc. Amer. Math. Soc. 94 (1985), 703-710. [115] J. J. Dijkstra and J. van Mill, On the dimension of Hilbert space remainders, Proc. Amer. Math. Soc. 124 (1996), 3261-3263. [116] J. J. Dijkstra and J. van Mill, The ambient homeomorphy of certain function and sequence spaces - addendum, 2001. [117] J. J. Dijkstra, J. van Mill, and J. Mogilski, The space of infinite-dimensional compact spaces and other topological copies of (I2,)", Pac. J. Math. 152 (1992), 255-273. [118] J. J. Dijkstra and J. Mogilski, The ambient homeomorphy of certain function and sequence spaces, Comm. Math. Univ. Carolinae 37 (1996), 595-611. [119] T. Dobrowolski, Cauty's proof of Schauder's fixed point theorem, 1999, Unpublished notes. [120] T. Dobrowolski, S. P. Gul'ko, and J. Mogilski, Function spaces homeomorphic to the countable product ofl^, Top. Appl. 34 (1990), 153-160. [121] T. Dobrowolski and W. Marciszewski, Classification of function spaces with the pointwise topology determined by a countable dense set, Fund. Math. 148 (1995), 35-62. [122] T. Dobrowolski and W. Marciszewski, Rays and the fixed point property in noncompact spaces, Tsukuba J. Math. 21 (1997), 97-112. [123] T. Dobrowolski, W. Marciszewski, and J. Mogilski, On topological classification of function spaces CP(X) of low Borel complexity, Trans. Amer. Math. Soc. 678 (1991), 307-324.
602
BIBLIOGRAPHY
[124] T. Dobrowolski and L. R. Rubin, The hyperspaces if infinite-dimensional compacta for covering dimension and cohomological dimension are homeomorphic, Pac. J. Math. 164 (1994), 15-39. [125] E. K. van Douwen, The small inductive dimension can be raised by the adjunction of a single point, Indag. Math. 35 (1973), 434-442. [126] E. K. van Douwen, Prime numbers, number of factors and binary operations, Dissertationes Math. (Rozprawy Mat.) 199 (1981), 1-35. [127] E. K. van Douwen, Special bases for compact metrizable spaces, Fund. Math. Ill (1981), 201-209. [128] E. K. van Douwen, A compact space with a measure that knows which sets are homeomorphic, Adv. Math. 52 (1984), 1-33. [129] E. K. van Douwen, Closed copies of the rationals, Comm. Math. Univ. Carolinae 28 (1987), 137-139. [130] E. K. van Douwen, Applications of maximal topologies, Top. Appl. 51 (1993), 125— 139. [131] E. K. van Douwen, /3X and fixed-point free maps, Top. Appl. 51 (1993), 191-195. [132] E. K. van Douwen, Nonextendible functions characterize dense G$ 's in metrizable spaces, Top. Appl. 51 (1993), 187-190. [133] E. K. van Douwen, Uncountably many pairwise disjoint copies of one metrizable compactum in another, Top. Appl. 51 (1993), 87-91. [134] E. K. van Douwen, Simultaneous extension of continuous functions, Ph.D. thesis, Collected papers I, II (J. van Mill, ed.), North-Holland Publishing Co., Amsterdam, 1994, p. 140. [135] E. K. van Douwen, Unpublished results, Collected papers I, II (J. van Mill, ed.), North-Holland Publishing Co., Amsterdam, 1994, p. 34. [136] E. K. van Douwen and R. Pol, Countable spaces without extension properties, Bull. Polon. Acad. Sci. Ser. Math. Astronom. Phys. 25 (1977), 987-991. [137] C. H. Dowker, Mapping theorems for non-compact spaces, Amer. Journ. of Math. 69 (1947), 200-242. [138] C. H. Dowker, On countably paracompact spaces, Can. J. Math. 3 (1951), 219-224. [139] A. N. Dranisnikov, On a problem of P.S. Alexandrov, Matem. Sbornik 135 (1988), 551-557. [140] J. Dugundji, An extension of Tietze's theorem, Pac. J. Math. 1 (1951), 353-367. [141] J. Dugundji, Absolute neighborhoods retracts and local connectedness in arbitrary metric spaces, Compositio Math. 13 (1958), 229-246. [142] J. Dugundji, Topology, Allyn and Bacon, Boston, 1966. [143] J. Dugundji and A. Granas, Fixed point theory. I, Pahstwowe Wydawnictwo Naukowe (PWN), Warsaw, 1982. [144] S. Eilenberg, Transformations continues en circonferences et la topologie du plan, Fund. Math. 26 (1936), 61-112. [145] S. Eilenberg and E. Otto, Quelques proprietes caracteristiques de la dimension, Fund. Math. 31 (1938), 149-153. [146] S. Eilenberg and N. Steenrod, Foundations of algebraic topology, Princeton University Press, Princeton, New Jersey, 1952. [147] E. Ellentuck, A new proof that analytic sets are Ramsey, J. Symb. Logic 39 (1974), 163-165. [148] A. J. M. van Engelen, Homogeneous zero-dimensional absolute Borel sets, CWI Tract, vol. 27, Centre for Mathematics and Computer Science, Amsterdam, 1986. [149] F. van Engelen, A decomposition o/R into two homeomorphic rigid parts, Top. Appl. 17 (1984), 275-285. [150] F. van Engelen and J. van Mill, Borel sets in compact spaces: some Hurewicz-type theorems, Fund. Math. 124 (1984), 271-286.
BIBLIOGRAPHY
603
[151] F. van Engelen, A. W. Miller, and J. Steel, Rigid Borel sets and better quasi-order theory, Logic and combinatorics (Arcata, Calif., 1985), Amer. Math. Soc., Providence, R.I., 1987, pp. 199-222. [152] R. Engelking, Sur la compactification des espaces metriques, Fund. Math. 48 (1960), 321-324. [153] R. Engelking, General topology, Heldermann Verlag, Berlin, second ed., 1989. [154] R. Engelking, Letter to the author (2-10-92), 1992. [155] R. Engelking, Theory of dimensions finite and infinite, Heldermann Verlag, Lemgo, 1995. [156] R. Engelking and R. Pol, Compactifications of countable-dimensional and strongly countable-dimensional spaces, Proc. Amer. Math. Soc. 104 (1988), 985-987. [157] P. Erdos, The dimension of the rational points in Hilbert space, Annals of Math. 41 (1940), 734-736. [158] I. Farah, Powers of N* , 2000, to appear in Proc. Amer. Math. Soc. [159] V. V. Fedorchuk, The Fundamentals of Dimension Theory, Encyclopedia of Mathematical Sciences (A. V. Arhangel'skii and L. S. Pontrjagin, eds.), vol. 17, SpringerVerlag, Berlin, 1990, pp. 91-202. [160] V. V. Fedorchuk and A. Chigogidze, Absolyutnye retrakty i beskonechnomernye mnogoobraziya, "Nauka", Moscow, 1992. [161] V. V. Fedorchuk, M. Levin, and E. V. Shchepin, On the Brouwer definition of dimension, Uspekhi Mat. Nauk 54 (1999), 193-194. [162] V. V. Fedorchuk and J. van Mill, Dimensionsgrad for locally connected Polish spaces, Fund. Math. 163 (2000), 77-82. [163] W. G. Fleissner and K. Kunen, Barely Baire spaces, Fund. Math. 101 (1978), 229240. [164] A. Flores, Uber n-dimensionale Komplexe die im R2n+i absolut selbstverschlungen sind, Ergebnisse eines math. Koll. 6 (1935), 4-6. [165] M. Fort, Homogeneity of infinite products of manifolds with boundary, Pac. J. Math. 12 (1962), 879-884. [166] M. Frechet, Les ensembles abstraits et le calcul fonctionnel, Rend, del Circ. Mat. di Palermo 30 (1910), 1-26. [167] M. Frechet, Les espaces abstraits, Hermann, Paris, 1928. [168] H. Freudenthal, Entwicklungen von Rdumen und ihren Gruppen, Compositio Math. 4 (1937), 145-234. [169] O. Frink, Compactifications and semi-normal spaces, Amer. J. Math. 86 (1964), 602-607. [170] Z. Frolik, Fixed points of maps of extremally disconnected spaces and complete Boolean Algebras, Bull. Polon. Acad. Sci. Ser. Math. Astronom. Phys. 16 (1968), 269-275. [171] F. Galvin and K. Prikry, Borel sets and Ramsey's Theorem, J. Symb. Logic 38 (1973), 193-198. [172] K. Geba and Z. Semadeni, Spaces of continuous functions V, Studia Mathematica 19 (1960), 303-320. [173] I. M. Gel'fand and A. N. Kolmogoroff, On rings of continuous functions on topological spaces, Comptes Rendus (Doklady) de 1'Academie des Sciences de 1'URSS 22 (1939), 11-15. [174] L. Gillman and M. Jerison, Rings of continuous functions, Van Nostrand, Princeton, 1960. [175] H. Gladdines, Absorbing systems in infinite-dimensional manifolds and applications, Ph.D. thesis, Vrije Universiteit, Amsterdam, the Netherlands, 1994. [176] J. de Groot, Topologische Studien, Ph.D. thesis, Groningen, the Netherlands, 1942.
604
BIBLIOGRAPHY
[177] J. de Groot, Non-archimedean metrics in topology, Proc. Amer. Math. Soc. 7 (1956), 948-953. [178] J. de Groot, Subcompactness and the Baire category theorem, Indag. Math. 25 (1963), 761-767. [179] J. de Groot and T. Nishiura, Inductive compactness as a generalization of semicompactness, Fund. Math. 58 (1966), 201-218. [180] J. de Groot and R. J. Wille, Rigid continua and topological group-pictures, Arch. Math. (Basel) 9 (1958), 441-446. [181] S. P. Gul'ko, Spaces of continuous functions on ordinals and ultrafilters, Mat. Zametki 47 (1990), 26-34. [182] S. P. Gul'ko, On uniform homeomorphisms of spaces of continuous functions, Proc. Steklov Inst. Math. 3 (1992), 87-93. [183] S. P. Gul'ko and T. E. Khmyleva, Compactness is not preserved by t-equivalence, Mat. Zametki 39 (1986), 895-903, in Russian. [184] S. P. Gul'ko and G. A. Sokolov, P-points in N* and the spaces of continuous functions, Top. Appl. 85 (1998), 137-142. [185] C. F. Hagopian, A characterization of solenoids, Pac. J. Math. 68 (1977), 425-435. [186] M. Handel, On certain sums of Hilbert cubes, Gen. Top. Appl. 9 (1978), 19-28. [187] O. Hanner, Some theorems on absolute neighborhood retracts, Arkiv Mat., Svenska Vetens. Akad. 1 (1951), 389-408. [188] K. P. Hart, J. van Mill, and R. Pol, Remarks on hereditarily infinite-dimensional continua, 2000, manuscript submitted for publication. [189] M. A. van Hartskamp, Colorings of Fixed-Point Free Maps, Ph.D. thesis, Vrije Universiteit, Amsterdam, the Netherlands, 1999. [190] M. A. van Hartskamp and J. Vermeer, On colorings of maps, Top. Appl. 73 (1996), 181-190. [191] F. Hausdorff, Uber halbstetige Fuktionen und deren Verallgemeinerung, Math. Zeitschr. 5 (1919), 292-309. [192] F. Hausdorff, Uber innere Abbildungen, Fund. Math. 23 (1934), 279-291. [193] F. Hausdorff, Grundzuge der Mengenlehre, Chelsea, New York, 1949. [194] D. W. Henderson, Finite dimensional subsets of infinite dimensional spaces, Topology Seminar Wisconsin 1965, Ann. of Math. Studies (60), 1965, pp. 141-146. [195] D. W. Henderson, An infinite-dimensional compactum with no positive-dimensional compact subsets-a simpler construction, Amer. J. Math. 89 (1967), 105-121. [196] J. P. Henderson and J. J. Walsh, Examples of cell-like decompositions of the infinitedimensional manifolds a and S, Top. Appl. 16 (1983), 143—154. [197] S. T. Hu, Theory of retracts, Wayne State University Press, Detroit, 1965. [198] W. Hurewicz, Uber stetige Bilder von Punktmengen, Proc. Akad. Amsterdam 29 (1926), 1014-1017. [199] W. Hurewicz, Normalbereiche und Dimensionstheorie, Math. Ann. 96 (1927), 736764. [200] W. Hurewicz, Uber das Verhdltnis separabler Rdume zu kompakten Rdumen, Proc. Akad. Amsterdam 30 (1927), 425-430. [201] W. Hurewicz, Uber stetige Bilder von Punktmengen (Zweite Mitteilung), Proc. Akad. Amsterdam 30 (1927), 159-165. [202] W. Hurewicz, Relativ perfekte Teile von Punktmengen und Mengen (A), Fund. Math. 12 (1928), 78-109. [203] W. Hurewicz, Zur Theorie der analytischen Mengen, Fund. Math. 15 (1930), 4-17. [204] W. Hurewicz, Sur la dimension des produits Cartesiens, Annals of Math. 36 (1935), 194-197. [205] W. Hurewicz, Uber Abbildungen topologischer Rdume auf die n-dimensionale Sphdre, Fund. Math. 24 (1935), 144-150.
BIBLIOGRAPHY
605
[206] W. Hurewicz, Bin einfacher Beweis des Hauptsatzes iiber Cantorsche Mannigfaltigkeiten, Prace Mat. Fiz. 34 (1937), 289-292. [207] W. Hurewicz and K. Menger, Dimension und Zusammenhangsstuffe, Math. Ann. 100 (1928), 618-633. [208] W. Hurewicz and H. Wallman, Dimension theory, Van Nostrand, Princeton, N.J., 1948. [209] A. Illanes and S. B Nadler, Hyper spaces: Fundamentals and Recent Advances, Marcel Dekker, New York, 1999. [210] A. V. Ivanov, An example concerning a theorem of Mazurkiewicz, Seminar in General Topology, Moskov. Gos. Univ., Moskow, 1981, pp. 49-51. [211] S. Jackson and R. D. Mauldin, Some complexity results in topology and analysis, Fund. Math. 141 (1992), 75-83. [212] I. M. James, ed., History of Topology, North-Holland Publishing Co., Amsterdam, 1999. [213] D. M. Johnson, The problem of the invariance of dimension in the growth of modern topology, Part II, Arch. Hist. Exact Sci. 25 (1981), 85-267. [214] I. Juhasz, Cardinal functions in topology, Mathematical Centre Tract, vol. 34, Mathematical Centre, Amsterdam, 1975. [215] M. Katetov, On real-valued functions in topological spaces, Fund. Math. 38 (1951), 85-91. [216] M. Katetov, On the dimension of non-separable spaces, Czech. Math. J. 2 (1952), 333-368. [217] M. Katetov, A theorem on mappings, Comm. Math. Univ. Carolinae 8 (1967), 431433. [218] A. S. Kechris, Classical descriptive set theory, Grad. Texts Math., vol. 156, SpringerVerlag, Berlin, 1994. [219] O. H. Keller, Die Homoiomorphie der kompakten konvexen Mengen in Hilbertschen Raum, Math. Ann. 105 (1931), 748-758. [220] J. L. Kelley, Hyperspaces of a continuum, Trans. Amer. Math. Soc. 52 (1942), 22-36. [221] V. L. Klee, Some topological properties of convex sets, Trans. Amer. Math. Soc. 78 (1955), 30-45. [222] B. Knaster, Sur les coupures biconnexes des espace euklidens de dimension n > 1 arbitrare, Mat. Sbornik 19 (1946), 9-18. [223] B. Knaster, K. Kuratowski, and S. Mazurkiewicz, Bin Beweis des fixpunktsatzes fur n-dimensionale Simplexe, Fund. Math. 14 (1929), 132 -137. [224] Y. Kodama, Cohomological dimension theory, Appendix to: K. Nagami, Dimension theory (New York) (1970). [225] A. Kolmogoroff, Uber offene Abbildungen, Annals of Math. 38 (1937), 36-38. [226] I. M. KozlovskiY, Dimension-raising open mappings of compacta onto polyhedra as spectral limits of inessential mappings of polyhedra, Soviet Math. Doklady 33 (1986), 118-121. [227] J. Krasinkiewicz, Shape properties of hyperspaces, Fund. Math. 101 (1978), 79-91. [228] J. Krasinkiewicz, Essential mappings onto products of manifolds, Geometric and Algebraic Topology (H. Torunczyk, S. Jackowski, and S. Spiez, eds.), (Banach Center Publications volume 18) PWN, Warszawa, 1986, pp. 377-406. [229] J. Krasinkiewicz, Homotopy separators and mapping into cubes, Fund. Math. 131 (1988), 149-154. [230] J. Krasinkiewicz, On mappings with hereditarily indecomposable fibers, Bull. Polon. Acad. Sci. Ser. Math. Astronom. Phys. 44 (1996), 147-156. [231] A. Krawczyk and J. Steprans, Continuous colourings of closed graphs, Top. Appl. 51 (1993), 13-26.
606
BIBLIOGRAPHY
[232] M. R. Krom, Cartesian products of metric Baire spaces, Proc. Amer. Math. Soc. 42 (1974), 588-594. [233] N. Kroonenberg, Pseudo-interiors of hyperspaces, Compositio Math. 32 (1976), 113131. [234] J. Kulesza, The dimension of products of complete separable metric spaces, Fund. Math. 135 (1990), 49-54. [235] J. Kulesza, New properties of Mrowka's space v^o, 1999, Preprint. [236] K. Kunen, Weak P-points in N* , Topology, Vol. II (Proc. Fourth Colloq., Budapest, 1978), North-Holland Publishing Co., Amsterdam, 1980, pp. 741-749. [237] K. Kuratowski, Sur la puissance de I'ensemble des "nombres de dimension" de M. Frechet, Fund. Math. 8 (1925), 201-208. [238] K. Kuratowski, Sur un theoreme fondamental concernant le nerf d'un systeme d'ensembles, Fund. Math. 20 (1933), 191-196. [239] K. Kuratowski, Quelques problemes concernant les espaces metriques non-separables, Fund. Math. 25 (1935), 534-545. [240] K. Kuratowski, Sur les espaces localement connexes et peaniens en dimension n, Fund. Math. 24 (1935), 269-287. [241] K. Kuratowski, Quelques theoremes sur le plongement topologique des espaces, Fund. Math. 30 (1938), 8-13. [242] K. Kuratowski, Topology I, Academic Press, New York, 1966. [243] K. Kuratowski, Topology II, Academic Press, New York, 1968. [244] K. Kuratowski and A. Mostowski, Set theory, North-Holland Publishing Co., Amsterdam, revised ed., 1976, With an introduction to descriptive set theory, Translated from the 1966 Polish original, Studies in Logic and the Foundations of Mathematics, Vol. 86. [245] M. Lavrentieff, Contribution a la theorie des ensembles homeomorphes, Fund. Math. 6 (1924), 149-160. [246] H. Lebesgue, Sur les correspondances entre les points de deux espaces, Fund. Math. 2 (1921), 256-285. [247] S. Lefschetz, On compact spaces, Annals of Math. 32 (1931), 521-538. [248] S. Lefschetz, Topics in Topology, Annals of Mathematical Studies 10, Princeton University Press, Princeton, 1942. [249] A. Lelek, On the dimension of remainders in compact extensions, Dokl. Akad. Nauk SSSR 160 (1965), 534-537. [250] A. Lelek, Dimension inequalities for unions and mappings of separable metric spaces, Coll. Math. 23 (1971), 69-91. [251] S. Levi, On Baire cosmic spaces, Proceedings of the Fifth Prague Topological Symposium, Heldermann Verlag, Berlin, pp. 450-451. [252] M. Levin, A short construction of hereditarily infinite dimensional compacta, Top. Appl. 65 (1995), 97-99. [253] M. Levin, Bing maps and finite-dimensional maps, Fund. Math. 151 (1996), 47-52. [254] M. Levin and R. Pol, A metric condition which implies dimension <1, Proc. Amer. Math. Soc. 125 (1997), 269-273. [255] M. Levin and J. T. Rogers, Jr., A generalization of Kelley's Theorem for C-spaces, Proc. Amer. Math. Soc. 128 (1999), 1537-1541. [256] M. Levin and J. Segal, A subspace of R3 with Cmp ^ def, Top. Appl. 95 (1999), 165-168. [257] M. Levin and Y. Sternfeld, Hyperspaces of two dimensional continua, Fund. Math. 150 (1996), 17-24. [258] M. Levin and Y. Sternfeld, The space of subcontinua of a 2-dimensional continuum is infinite dimensional, Proc. Amer. Math. Soc. 125 (1997), 2771-2775.
BIBLIOGRAPHY
607
[259] M. Levin and Y. Sternfeld, Atomic maps and the Chogoshvili-Pontrjagin claim, Trans. Amer. Math. Soc. 350 (1998), 4623-4632. [260] W. Lewis, The pseudo-arc, Boll. Soc. Math. Mexicana 5 (1999), 25-77. [261] J. Lindenstrauss and L. Tzafriri, Classical Banach spaces, I and II, Springer-Verlag, Berlin, 1996. [262] S. L. Lipscomb, On imbedding finite-dimensional metric spaces, Trans. Amer. Math. Soc. 211 (1975), 143-160. [263] L. Lusternik and L. Schnirelman, Topological methods in variational calculus (Russian), Moscow, 1930. [264] D. J. Lutzer and R. McCoy, Category in function spaces. I, Pac. J. Math. 90 (1980), 145-168. [265] D. J. Lutzer, J. van Mill, and R. Pol, Descriptive complexity of function spaces, Trans. Amer. Math. Soc. 291 (1985), 121-128. [266] W. Marciszewski, A function space C(K) not weakly homeomorphic to C(K)xC(K), Studia Mathematica 88 (1988), 129-137. [267] W. Marciszewski, On analytic and coanalytic function spaces CP(X), Top. Appl. 50 (1993), 241-248. [268] W. Marciszewski, A function space CP(X) not linearly homeomorphic to CP(X] x R, Fund. Math. 153 (1997), 125-140. [269] W. Marciszewski, On topological embeddings of linear metric spaces, Math. Ann. 308 (1997), 21-30. [270] W. Marciszewski, P-filters and hereditary Baire function spaces, Top. Appl. 89 (1998), 241-247. [271] W. Marciszewski, On properties of metrizable spaces X preserved by t-equivalence, 1999, to appear in Mathematika. [272] W. Marciszewski and J. van Mill, An example of t^-equivalent spaces which are not tp-equivalent, Top. Appl. 85 (1998), 281-285. [273] W. Marciszewski and J. Pelant, Absolute Borel sets and function spaces, Trans. Amer. Math. Soc. 349 (1997), 3585-3596. [274] S. Mardesic and J. Segal, Shape theory, North-Holland Publishing Co., Amsterdam, 1982. [275] S. Mazurkiewicz, O arytmetyzacji continuow, C.R. Varsovie 6 (1913), 305-311. [276] S. Mazurkiewicz, Sur les probleme K et X de Urysohn, Fund. Math. 10 (1927), 311319. [277] S. Mazurkiewicz, Sur les ensembles de dimension faible, Fund. Math. 13 (1929), 210-217. [278] S. Mazurkiewicz, Sur I'hyperespace d'un continu, Fund. Math. 18 (1932), 171-177. [279] S. Mazurkiewicz, Uber die Menge der differenzierbaren Funktionen, Fund. Math. 27 (1936), 244-249. [280] R. A. McCoy and I. Ntantu, Topological properties of spaces of continuous functions, Lectute Notes in Mathematics, vol. 1315, Springer-Verlag, 1988. [281] K. Menger, Uber die Dimension von Punktmengen I, Monatsh. fur Math, und Phys. 33 (1923), 148-160. [282] K. Menger, Uber die Dimension von Punktmengen II, Monatsh. fur Math, und Phys. 34 (1926), 137-161. [283] K. Menger, Dimensionstheorie, Leipzig und Berlin, 1928. [284] K. Menger, Bemerkungen uber dimensionelle Feinstruktur und Produktsatz, Prace Mat.-Fiz. 38 (1930), 77-90. [285] E. A. Michael, Some extension theorems for continuous functions, Pac. J. Math. 3 (1953), 789-806. [286] E. A. Michael, Local properties of topological spaces, Pac. J. Math. 21 (1954), 163171.
608
BIBLIOGRAPHY
[287] E. A. Michael, Continuous selections I, Annals of Math. 63 (1956), 361-382. [288] E. A. Michael, Continuous selections and countable sets, Fund. Math. Ill (1981), 1-10. [289] H. Michalewski, Przestrzenie funkcij ciaglych i przestrzenie dziedzicznie Baire'a (Master's Thesis), 1998. [290] A. A. Miljutin, Isomorphisms of the spaces of continuous functions over compact sets of the cardinality of the continuum (Russian), Teor. Funkcii Funkcional Anal i Prilozen. (Kharkov) 2 (1966), 150-156. [291] J. van Mill, A Peano continuum homeomorphic to its own square but not to its countable infinite product, Proc. Amer. Math. Soc. 80 (1980), 703—705. [292] J. van Mill, Characterization of some zero-dimensional separable metric spaces, Trans. Amer. Math. Soc. 264 (1981), 205-215. [293] J. van Mill, Homogeneous subsets of the real line, Compositio Math. 45 (1982), 3-13. [294] J. van Mill, Strong local homogeneity does not imply countable dense homogeneity, Proc. Amer. Math. Soc. 84 (1982), 143-148. [295] J. van Mill, A boundary set for the Hilbert cube containing no arcs, Fund. Math. 118 (1983), 93-102. [296] J. van Mill, Domain invariance in infinite-dimensional linear spaces, Proc. Amer. Math. Soc. 101 (1987), 173-180. [297] J. van Mill, Topological equivalence of certain function spaces, Compositio Math. 63 (1987), 159-188. [298] J. van Mill, Infinite-dimensional topology: prerequisites and introduction, NorthHolland Publishing Co., Amsterdam, 1989. [299] J. van Mill, An infinite-dimensional homogeneous indecomposable continuum, Houston J. Math. 10 (1990), 195-201. [300] J. van Mill, CP(X] is not GS&: a simple proof, Bull. Polon. Acad. Sci. Ser. Math. Astronom. Phys. 47 (1999), 319-323. [301] J. van Mill, Easier proofs of coloring theorems, Top. Appl. 97 (1999), 155-163. [302] J. van Mill and R. Pol, The Baire Category Theorem in products of linear spaces and topological groups, Top. Appl. 22 (1986), 267-282. [303] J. van Mill and R. Pol, On the existence of weakly n-dimensional spaces, Proc. Amer. Math. Soc. 113 (1991), 581-585. [304] J. van Mill and R. Pol, A countable space with a closed subspace without measurable extender, Bull. Polon. Acad. Sci. Ser. Math. Astronom. Phys. 41 (1993), 279-283. [305] J. van Mill and R. Pol, Baire 1 functions which are not countable unions of continuous functions, Acta Math. Hungar. 66 (1995), 289-300. [306] J. van Mill and R. Pol, Remark on products of 1-dimensional compacta, Q&A in General Topology 13 (1995), 97-98. [307] J. van Mill and R. Pol, Note on weakly n-dimensional spaces, 2000, (to appear in Monatsh. Math.). [308] J. van Mill and R. Pol, In preparation, 2001. [309] A. W. Miller, Special Subsets of the Real Line, Handbook of Set-Theoretic Topology (K. Kunen and J.E. Vaughan, eds.), North-Holland Publishing Co., Amsterdam, 1984, pp. 201-233. [310] A. W. Miller, Descriptive Set Theory and Forcing, Lecture Notes in Logic, vol. 4, Springer-Verlag, Berlin, 1995. [311] J. Mogilski, Unions of infinite-dimensional compact convex subsets of the Frechet space, Bull. Polon. Acad. Sci. Ser. Math. Astronom. Phys. 24 (1976), 1097-1102. [312] E. E. Moise, Grille decompositions and convexification theorems for compact locally connected continua, Bull. Amer. Math. Soc. 55 (1949), 1111-1121. [313] R. L. Moore, Concerning simple continuous curves, Trans. Amer. Math. Soc. 21 (1920), 333-347.
BIBLIOGRAPHY
609
[314] K. Morita, Normal families and dimension theory for metric spaces, Math. Ann. 128 (1954), 350-362. [315] Y. N. Moschovakis, Descriptive Set Theory, North-Holland Publishing Co., Amsterdam, 1980. [316] M. Moszyriska, Generalization of the theory of retracts on the pairs of spaces, Bull. Polon. Acad. Sci. Ser. Math. Astronom. Phys. 13 (1965), 13-19. [317] S. Mrowka, Small inductive dimension of completions of metric spaces, Proc. Amer. Math. Soc. 125 (1997), 1545-1554. [318] S. Mrowka, Small inductive dimension of completions of metric spaces. II, Proc. Amer. Math. Soc. 128 (2000), 1247-1256. [319] J. R. Munkres, Elements of Algebraic Topology, Addison-Wesley Publishing Co., Menlo Park, 1984. [320] S. B. Nadler, Hyperspaces of sets, Marcel Dekker, New York and Basel, 1978. [321] K. Nagami, Finite-to-one closed mappings and dimension, II, Proc. Japan Acad. 35 (1959), 437-439. [322] J. Nagata, On lattices of functions on topological spaces and functions on uniform spaces, Osaka Math. J. 1 (1949), 166-181. [323] J. Nagata, Modern dimension theory, Interscience Publishers John Wiley &: Sons, Inc., New York, 1965, Bibliotheca Mathematica, Vol. VI. Edited with the cooperation of the "Mathematisch Centrum" and the "Wiskundig Genootschap" at Amsterdam. [324] J. Nagata, Modern general topology, North-Holland Mathematical Library, vol. 33, North-Holland Publishing Co., Amsterdam, 1985. [325] J. I. Nagata, Modern dimension theory, Heldermann Verlag, Berlin, 1983, (revised and extended edition). [326] O. Nikodym, Sur une propriete de I'operation A, Fund. Math. 7 (1925), 149-154. [327] G. Nobeling, Uber eine n-dimensionale Universalmenge in R2""^1, Math. Ann. 104 (1931), 71-80. [328] O. G. Okunev, Weak topology of a dual space and a t-equivalence relation, Math. Notes 46 (1989), 534-538. [329] W. Olszewski, Embeddings of finite-dimensional spaces into finite products of 1dimensional spaces, Top. Appl. 40 (1991), 93-99. [330] L. G. Oversteegen and E. D. Tymchatyn, On the dimension of certain totally disconnected spaces, Proc. Amer. Math. Soc. 122 (1994), 885-891. [331] J. Oxtoby, Cartesian products of Baire spaces, Fund. Math. 49 (1961), 157-166. [332] K. R. Parthasarathy, Probability measures on metric spaces, Academic Press, New York, 1967. [333] D. Pavlovskii, On spaces of continuous functions, Soviet Math. Doklady 22 (1980), 34-37. [334] J. Pelant, A remark on spaces of bounded continuous functions, Indag. Math. 91 (1988), 335-338. [335] A. Pelczyriski, Projections in certain Banach spaces, Studia Math. 19 (1960), 209— 228. [336] V. G. Pestov, The coincidence of the dimension dim of (.-equivalent topological spaces, Soviet Math. Doklady 26 (1982), 380-383. [337] H. Poincare, L'espace et ses trois dimensions, Revue de Metaph. et de Morale 11 (1903), 281-301 and 407-429. [338] H. Poincare, Pourquoi I'espace a trois dimensions, Revue de Metaph. et de Morale 20 (1912), 483-504. [339] E. Pol, On hereditarily indecomposable continua, Henderson compacta and a question of Yohe, 2000, (to appear). [340] R. Pol, A weakly infinite-dimensional compactum which is not countabledimensional, Proc. Amer. Math. Soc. 82 (1981), 634-636.
610
BIBLIOGRAPHY
[341] R. Pol, A counterexample to J. de Croat's problem cmp = def, Bull. Polon. Acad. Sci. Ser. Math. Astronom. Phys. 30 (1982), 461-464. [342] R. Pol, An infinite-dimensional pre-Hilbert space not homeomorphic to its own square, Proc. Amer. Math. Soc. 90 (1984), 450-454. [343] R. Pol, Countable dimensional universal sets, Trans. Amer. Math. Soc. 297 (1986), 255-268. [344] R. Pol, On metrizable E with CP(E) % CP(E) X CP(E), Mathematika 42 (1995), 49-55. [345] R. Pol, Private communication (20-09-99), 1999. [346] L. S. Pontrjagin, Sur une hypothese fondamentale de la theorie de la dimension, C.R. Acad. Paris 190 (1930), 1105-1107. [347] L. S. Pontrjagin and G. Tolstowa, Beweis des Mengerschen Einbettungssatzes, Math. Ann. 105 (1931), 734-747. [348] T. C. Przymusiriski, A note on dimension theory of metric spaces, Fund. Math. 85 (1974), 277-284. [349] E. G. Pytkeev, The Baire property of spaces of continuous functions, Math. Zametki 38 (1985), 726-740. [350] J. Quinn and R. Y. T. Wong, Unions of convex Hilbert cubes, Proc. Amer. Math. Soc. 65 (1977), 171-176. [351] D. Repovs and P. V. Semenov, Continuous selections of multivalued mappings, Mathematics and its Applications, vol. 455, Kluwer Academic Publishers, Dordrecht, 1998. [352] C. A. Rogers, J. E. Jayne, C. Dellacherie, F. Tops0e, J. Hoffman-J0rgensen, D. A. Martin, A. S. Kechris, and A. H. Stone, Analytic sets, Academic Press, New York, 1980. [353] J. T. Rogers, Jr., Dimension of hyperspaces, Bull. Polon. Acad. Sci. Ser. Math. Astronom. Phys. 20 (1972), 177-179. [354] J. T. Rogers, Jr., Orbits of higher-dimensional hereditarily indecomposable continua, Proc. Amer. Math. Soc. 95 (1985), 483-486. [355] P. Roy, Failure of equivalence of dimension concepts for metric spaces, Bull. Amer. Math. Soc. 68 (1962), 609-613. [356] P. Roy, Nonequality of dimensions for metric spaces, Trans. Amer. Math. Soc. 134 (1968), 117-132. [357] L. Rubin, R. M. Schori, and J. J. Walsh, New dimension-theory techniques for constructing infinite-dimensional examples, Gen. Top. Appl. 10 (1979), 93-102. [358] L. R. Rubin, Noncompact hereditarily strongly infinite dimensional spaces, Proc. Amer. Math. Soc. 79 (1980), 153-154. [359] M. E. Rudin, A new proof that metric spaces are paracompact, Proc. Amer. Math. Soc. 20 (1969), 603. [360] W. Rudin, Homogeneity problems in the theory of Cech compactifications, Duke Math. J. 23 (1956), 409-419. [361] J. Schauder, Der Fixpunktsatz in Funktionalrdumen, Studia Math. 2 (1930), 171180. [362] R. M. Schori, Hyperspaces and symmetric products of topological spaces, Fund. Math. 63 (1968), 77-88. [363] R. M. Schori and J. E. West, Hyperspaces of graphs are Hilbert cubes, Pac. J. Math. 53 (1974), 239-251. [364] R. M. Schori and J. E. West, The hyperspace of the closed interval is a Hilbert cube, Trans. Amer. Math. Soc. 213 (1975), 217-235. [365] S. Shelah, Decomposing topological spaces into two rigid homeomorphic subspaces, Israel J. Math. 63 (1988), 183-211. [366] R. B. Sher, The union of two Hilbert cubes meeting in a Hilbert cube need not be a Hilbert cube, Proc. Amer. Math. Soc. 63 (1977), 150-152.
BIBLIOGRAPHY
611
[367] W. Sierpiriski, L'arc simple comme un ensemble de points dans I'espace a m dimensions, Ann. Mat. Pur. Appl. 26 (1917), 131-150. [368] W. Sierpiriski, Un theoreme sur les continus, Tohoku Math. 13 (1918), 300-303. [369] W. Sierpiriski, Sur une propriete topologique des ensembles denombrables denses en soi, Fund. Math. 1 (1920), 11-16. [370] W. Sierpiriski, Sur les ensembles connexes et non connexes, Fund. Math. 2 (1921), 81-95. [371] W. Sierpiriski, Sur une decomposition d'ensembles, Monatsh. fiir Math, und Phys. 35 (1928), 239-242. [372] W. Sierpiriski, Remarque sur la courbe peanienne, Wiadom. Mat. 42 (1937), 1-3. [373] W. Sierpiriski, Sur un probleme concernant les fonctions semi-continues, Fund. Math. 28 (1937), 1-6. [374] R. Sikorski, On the Cartesian product of metric spaces, Fund. Math. 34 (1977), 288-292. [375] M. Souslin, Sur une defenition des ensembles mesurables B sans nombres transfinis, Compt. Rend. Acad. Sci. 164 (1917), 89. [376] E. Spanier, Algebraic topology, Springer-Verlag, Berlin, 1982. [377] E. Sperner, Neuer Beweis fiir die Invarianz des Dimensionszahl und des Gebietes, Abh. Math. Semin. Hamburg. Univ. 6 (1928), 265-272. [378] A. K. Steiner and E. F. Steiner, Proucts of compact metric spaces are regular Wallman, Indag. Math. 30 (1968), 428-430. [379] H. Steinlein, Borsuk-Ulam Sdtze und Abbildungen mil kompakten Iterierten, Dissertationes Math. (Rozprawy Mat.) 177 (1980), 113. [380] H. Steinlein, On the theorems of Borsuk-Ulam and Ljusternik-Schnirelmann-Borsuk, Canad. Math. Bull. 27 (1984), 192-204. [381] Y. Sternfeld, Mappings in dendrites and dimension, Houston J. Math. 19 (1993), 483-497. [382] Y. Sternfeld, Stability and dimension—a counterexample to a conjecture of Chogoshvili, Trans. Amer. Math. Soc. 340 (1993), 243-251. [383] A. H. Stone, Paracompactness and product spaces, Bull. Amer. Math. Soc. 54 (1948), 977-982. [384] M. Talagrand, Compacts de fonctions mesurables et filtres non mesurables, Studia Math. 67 (1980), 13-43. [385] A. D. Taimanov, On extension of continuous mappings of topological spaces, Mat. Sb. (N.S.) 31 (1952), 459-463. [386] A. E. Taylor, Introduction to functional analysis, John Wiley and Sons, London, 1964. [387] H. Tietze, Uber Funktionen die auf einer abgeschlossenen Menge stetig sind, J. Reine Angew. Math. 145 (1915), 9-14. [388] V. V. Tkachuk, Characterization of Baire property in CP(X) by the properties of a space X, The mappings and the extensions of topological spaces (Ustinov, ed.), 1985, in Russian, pp. 21-27. [389] B. Tomaszewski, On weakly n-dimensional spaces, Fund. Math. 103 (1979), 1-8. [390] H. Toruriczyk, Concerning locally homotopy negligible sets and characterizations of l2-manifolds, Fund. Math. 101 (1978), 93-110. [391] H. Toruriczyk, On CE-images of the Hilbert cube and characterizations of Q manifolds, Fund. Math. 106 (1980), 31-40. [392] H. Toruriczyk, Characterizing Hilbert space topology, Fund. Math. Ill (1981), 247262. [393] L. A. Tumarkin, Beitrag zur allgemeinen Dimensionstheorie, Mam. Cb. 33 (1926), 57-86.
612
BIBLIOGRAPHY
[394] L. A. Tumarkin, Sur la structure dimensionelle des ensembles fermes, C.R. Acad. Paris 186 (1928), 420-422. [395] V. M. Ul'janov, Solution of a basic problem on compactifications of Wallman type, Soviet Math. Doklady 18 (1977), 567-571. [396] P. Urysohn, Les multiplicites Cantoriennes, C.R. Acad. Paris 175 (1922), 440-442. [397] P. Urysohn, Memoire sur les multiplicites Cantoriennes, Fund. Math. 7 (1925), 30 137. [398] P. Urysohn, Memoire sur les multiplicites Cantoriennes (suite), Fund. Math. 8 (1926), 225-359. [399] P. Urysohn, Uber die Mdchtigkeit der zusammenhdngenden Mengen, Math. Annalen 94 (25), 262-295. [400] L. Vietoris, Bereiche zweiter Ordnung, Monatsh. fur Math, und Phys. 32 (1922), 258-280. [401] J. de Vries, Elements of topological dynamics, Kluwer Academic Publishers, Dordrecht, 1992. [402] H. Wallman, Lattices and topological spaces, Annals of Math. 39 (1938), 112-126. [403] J. J. Walsh, Infinite-dimensional compacta containing no n-dimensional (n > 1) subsets, Topology 18 (1979), 91-95. [404] T. Wazewski, Sur un continu singulier, Fund. Math. 4 (1923), 214-245. [405] J. E. West, The ambient homeomorphy of an incomplete subspace of an infinitedimensional Hilbert space, Pac. J. Math. 34 (1970), 257-267. [406] J. E. West, Mapping Hilbert cube manifolds to ANR's: a solution to a conjecture of Borsuk, Annals of Math. 106 (1977), 1-18. [407] J. E. West, Problems in Infinite-dimensional Topology, Open Problems in Topology (J. van Mill and G. M. Reed, eds.), North-Holland Publishing Co., Amsterdam, 1990, pp. 523-597. [408] J. H. C. Whitehead, Simplicial spaces nuclei, and m-groups, Proc. London Mathematical Soc. 45 (1939), 243-327. [409] H. Whitney, Regular families of curves, I, Proc. Nat. Acad. Sci. 18 (1932), 275-278. [410] E. Wimmers, The Shelah P-point independence theorem, Israel J. Math. 43 (1982), 28-48. [411] M. Wojdyslawski, Retractes absolus et hyperespaces des continus, Fund. Math. 32 (1939), 184-192. [412] A. V. Zarelua, Hereditarily infinite-dimensional spaces, Theory of sets and topology (in honour of Felix Hausdorff, 1868-1942), VEB Deutsch. Verlag Wissensch., Berlin, 1972, pp. 509-525.
Special Symbols
Ad, 417 [a, A], 239 A, 460 AE, 25 aff(S), 111 av(x), 114 a A", 472 A/n, 240 ANE, 25 ANR, 24 AR, 24
), 418
459 D(x,e), 459 £>e(x), 459 De(A,e), 459 def A, 183 9|o- , 114 diam(A), 459 dim A, 152 dim/, (X), 340 dim>fc(A"), 340 ), 340
B(A,e), 459 £(x,e), 459 5n, 461 Se(z), 459 £e(A,e), 459 £A, 368 B(Q), 60 517
e-rnap, 34
f = t, 458 F: X =»y, 12 /~ 3,510
C, 42 carx, 119 CS(X,Y), 34 C(f), 250 eg-, 435 XA, 458 cmpX, 182 C, 459 conv(A), 1 conv00(J4), 2 convn(A), 1 C(X), 4 C(A", Y), 460 C*(X), 7 CP(X), 4, 368 Cp^PO, 396 C;>A(A'), 396 Ce(X,Y), 29 C^A"), 418
3"n(A), 100 Pr(A), 460 Fa, 466 3"^, 347 FaS, 518 / U 5, 458 , 100 Gs, 466 G^, 518 3e(X,Y), 34 Sn.fcW, 340
,457 613
614 lx, 458 IndX, 180 indX, 177 Int(A), 460 /(x,y), 19 J(X,y),35
J,457 «(*>> 418 ^'19 A(X), 227 LC, 516 /„)„, 80 ), HI £2 8 ' Isc, 488 L(X), 399 M
5, 348 mesh(U), 459 M«, 326
N, 457 N, 387 [N]" , 238 w, 459 w(X,J), 497
P, 42, 457 TTE, 457 P-Point, 393 IIi/*i,457 Q, 461 Q, 42, 457 M 457
>
g(f,g), 468 £>H, 96 s, 3, 60 (SC)abstact, 116 (SC)ge0metric, 117 (SC), 116 >, 130 5, 130 sd(n)S, 130 §e(X,Y), 34 S, 46
SPECIAL SYMBOLS
0 n ,354 A^X), 515 TV, 409 T
409
U, 460 It < V, 458 USC; 489
V(A,S), 462
VF^, 462
X ?s y, 460 X~/y, 370 X~^*y, 370 Xoo,348 [x]fc], 427 x/V, 505 X , 183 L[X,Y], J
511
Z,457 0, 2 5, 461 < Z>a(X), 307
Author Index
Cantor, 580, 592, 600 Cauty, 24, 26, 264, 367, 414, 426, 445, 446, 456, 580, 583, 590, 592, 598, 600 Cech, 583, 584, 600 Chapman, 588, 589, 597, 600 Chatyrko, 587, 600 Chigogidze, 588, 600, 603 Choquet, 384, 601 Christensen, 426, 601 Cohen, 384, 601 Comfort, 590, 592, 601 Cook, 447, 601 Curtis, 291, 325, 580, 584, 588-590, 598, 601
Aarts, 584, 586, 587, 594, 597 Adjan, 587, 597 Alexandroff,253, 254, 580-582, 584, 587, 593, 597 Anderson, 9, 325, 579, 580, 584, 585, 588-590, 597, 598 Arens, 396, 579, 598 Arhangel'skii, 367, 372, 411, 426, 451, 590, 591, 598, 603 Arveson, 455, 598 Aull, 579, 598 Baars, x, 367, 411, 425, 434, 591, 592, 598 Banach, 579, 598 Banakh, 426, 437, 445, 588, 590, 598 Bandt, 594, 598 Barit, 589, 598 Barov, x Bartle, 579, 598 Becker, 49, 580, 598 Bendixson, 592 Bennett, 580, 598 Berney, 593, 594, 599 Bessaga, 371, 588, 589, 599 Bestvina, 434, 588-590, 599 van der Bijl, 588, 599 Bing, 215, 580-582, 585, 589, 593, 597, 599 Blaszczyk, 586, 599 Blumenthal, 579, 599 Borel, 594 Borsuk, 301, 580, 581, 583, 585-588, 599 Bourbaki, 585, 593, 599 Bowers, 586, 588, 599, 600 Brouwer, 221-223, 580, 582-584, 600 Brown, 579, 582, 600 de Bruijn, 586, 600 Burckel, 594, 600
van Dantzig, 582, 601 Daverman, 588, 594, 601 Dellacherie, 610 Devaney, 582, 601 Dieudonne, 593, 601 Dijkstra, x, 434, 585, 587, 588, 590, 592, 594, 601 Dobrowolski, ix, x, 367, 434, 446, 579, 580, 583, 588, 590-592, 600-602 van Douwen, 75, 384, 580-583, 586, 591594, 602 Dowker, 583, 593, 602 Dranisnikov, 24, 585, 602 Dugundji, 23, 24, 29, 149, 372, 457, 580, 583, 588, 591, 594, 602 Eels, 579, 598 Eilenberg, 583, 593, 594, 602 Ellentuck, 239, 586, 602 van Engelen, 49, 580, 581, 594, 598, 602, 603 Engelking, 174, 368, 369, 394, 412, 415, 428, 457, 465, 579, 582-584, 587, 593, 603 Erdos, 580, 586, 600, 603
Calbrix, 446, 590, 592, 600 615
616
AUTHOR INDEX
Farah, 582, 603 Fedorchuk, 585, 588, 603 Fleissner, 384, 603 Flores, 174, 603 Fokkink, 586, 587, 597 Fort, 580, 603 Frechet, 221, 579, 593, 603 Freudenthal, 582, 585, 603 Frink, 593, 603 Frolfk, 586, 603 Galvin, 586, 603 Geba, 394, 451, 603 Gel'fand, 371, 603 Gillman, 590, 603 Gladdines, 434, 588, 592, 598, 603 Granas, 149, 583, 602 Graves, 579, 598 Grilliot, 590, 592. 601 de Groot, J., 373, 580, 581, 584, 603, 604 de Groot, J.A.M., 367, 411, 591, 592, 598 Gul'ko, 393, 411, 412, 414, 447, 456, 590, 592, 601, 604 Hagopian, 582, 604 Handel, 589, 604 Manner, 588, 604 Hart, x, 585, 587, 604 van Hartskamp, x, 587, 604 Hausdorff, 580, 582, 593, 604 Henderson, D.W., 585, 587, 604 Henderson, J.P., 589, 604 Hoffman-J0rgensen, 610 Hu, 580, 588, 604 Hurewicz, 222, 224, 581, 583-586, 594, 604, 605 Husek, 598 Illanes, 582, 605 Ivanov, 585, 605 Jackowski, 605 Jackson, 587, 605 James, 579, 605 Jayne, 610 Jerison, 590, 603 Johnson, 222, 594, 605 Juhasz, 592, 605 Katetov, 584, 586, 593, 605 Kechris, 517, 586, 594, 605, 610 Keisler, 584, 585, 598 Keller, 580, 581, 605 Kelley, 587, 605
Khmyleva, 411, 412, 456, 604 Kim, 586, 599 Klee, 579, 581, 589, 599, 605 Knaster, 582, 585, 605 Kodama, 584, 605 Kolmogoroff, 371, 585, 603, 605 Kozlovskii, 585, 605 Krasinkiewicz, 582, 585, 605 Krawczyk, 586, 605 Krom, 384, 606 Kroonenberg, 590, 606 Kulesza, 584, 585, 606 Kunen, 384, 386, 603, 606, 608 Kuratowski, 80, 447, 459, 517, 579, 582584, 586, 588, 593, 594, 605, 606 Lavrentieff, 593, 606 Lebesgue, 593, 606 Lefschetz, 583, 588, 606 Lelek, 585, 587, 606 Levi, 594, 606 Levin, 582-585, 587, 603, 606, 607 Lewis, 582, 607 Lindenstrauss, 579, 607 Lipscomb, 586, 607 Lowen, 579, 598 Lusin, 587 Lusternik, 583, 607 Lutzer, 367, 446, 456, 590-592, 601, 607 Marciszewski, ix, x, 18, 367, 393, 398, 414, 426, 434, 445-447, 456, 588, 590-593, 600, 601, 607 Mardesic, 588, 607 Martin, 610 Mauldin, 587, 605 Mazur, 586 Mazurkiewicz, 580, 582, 585, 586, 594, 605, 607 McCoy, 367, 590, 591, 607 Menger, 234, 583-585, 605, 607 Michael, 396, 452, 579, 607, 608 Michalewski, x, 591, 608 Miljutin, 370, 608 van Mill, 9, 49, 64, 294, 325, 367, 384, 426,434,445-447, 456, 580-582, 584592, 598, 601-604, 607, 608, 612 Miller, 517, 581, 595, 603, 608 Mogilski, ix, 367, 434, 588-592, 599, 601, 608 Moise, 593, 608 Moore, 594, 608 Morita, 584, 609
AUTHOR INDEX Moschovakis, 517, 609 Mostowski, 517, 594, 606 Moszyriska, 588, 609 Mrowka, 584, 609 Munkres, 582, 609 Nadler, 582, 605, 609 Nagami, 584, 609 Nagata, 372, 457, 582, 586, 593, 609 Negrepontis, 590, 592, 601 von Neumann, 455 Nhu, 588, 590, 601 Nikodym, 594, 609 Nishiura, 584, 597, 604 Nobeling, 583, 609 Novikov, 587, 597 Ntantu, 367, 591, 607 Okunev, 426, 609 Olszewski, 586, 609 Otto, 583, 602 Oversteegen, 583, 584, 609 Oxtoby, 384, 593, 609 Page, 579, 600 Parthasarathy, 585, 609 Pavlovskii, 591, 609 Peano, 580 Pelant, 591, 592, 598, 607, 609 Pelczyriski, 371, 579, 588, 589, 599, 609 Pestov, 591, 609 Poincare, 221, 583, 609 Pol, E., 587, 600, 609 Pol, R., x, 220, 384, 446, 447, 456, 580, 583-587, 590,592,602-604, 606-610 Pontrjagin, 583, 584, 603, 610 Prikry, 586, 603 Przymusiriski, 583, 610 Pytkeev, 591, 610
617
Schnirelman, 583, 607 Schori, 291, 581, 585, 588, 601, 610 Segal, 584, 588, 606, 607 Semadeni, 394, 451, 603 Semenov, 579, 610 Shchepin, 585, 603 Shelah, 393, 581, 610 Sher, 589, 610 Sierpiriski, 580-582, 585-587, 592, 594, 611 Sikorski, 384, 611 Smirnov, 593 Sokolov, 393, 604 Souslin, 580, 594, 611 Spanier, 582, 611 Sperner, 582, 611 Spiez, 605 Steel, 581, 603 Steenrod, 593, 602 Steiner, A.K., 594, 611 Steiner, E.F., 594, 611 Steinlein, 245, 586, 611 Sternfeld, 582, 586, 587, 606, 607, 611 Stone, 369, 610, 611 Taimanov, 593, 611 Talagrand, 591, 611 Taylor, 579, 611 Tietze, 593, 611 Tkachuk, 591, 611 Tolstowa, 583, 610 Tomaszewski, 586, 611 Tops0e, 610 Toruriczyk, 263, 291, 294, 580, 588, 605, 611 Tumarkin, 583-585, 611, 612 Tymchatyn, 583, 584, 609 Tzafriri, 579, 607
Quinn, 589, 610 Radul, 437, 588, 590, 598 Reed, 612 Repovs, 579, 610 Rogers, 455, 610 Rogers, Jr., 216, 587, 606, 610 Roy, 584, 610 Rubin, 585, 587, 590, 602, 610 Rudin, M.E., 369 Rudin, W., 393, 610 Salomon, x Schauder, 582, 610
Ul'janov, 593, 612 Urysohn, 222, 580, 581, 583-585, 593, 594, 597, 612 Uspenskii, 585 Ustinov, 611 Valdivia, 384, 600 Vaughan, 608 van de Vel, 580 Vermeer, 586, 587, 597, 604 Vietoris, 582, 612 de Vries, x, 582, 612
618
AUTHOR INDEX
Wallman, 222, 224, 584, 585, 593, 605, 612 Walsh, 585, 587-589, 599, 604, 610, 612 Wazewski, 582, 612 West, 291, 588, 589, 610, 612 Whitehead, 132, 612 Whitney, 582, 612 Wille, 581, 604 Wimmers, 393, 612 Wojdyslawski, 579, 588, 612 Wong, 589, 610 Yankov, 455 Zarelua, 585, 612 Zarichnyi, 437, 588, 590, 598 Zorn, 80, 459
Subject Index
preserving, 149 .A-operation, 594 Approximation Theorem Brown, 90 Freudenthal, 137
Absolute (Neighborhood) Extensor, 25 Absolute Neighborhood Retract, 24-29, 38, 39, 124, 145, 148, 263-265, 270, 274, 276, 284, 288, 289, 301-305, 580, 581, 588 pair, 266, 267, 273, 276, 277, 285, 290, 301, 304, 305, 588 Absolute Retract, 24-27, 29, 39, 145, 276, 288, 289, 294, 301, 305, 580 hyperspace, 292, 588 pair, 266, 288, 290 hyperspace, 292 absolute value, 512 absorber, 347, 348, 589 Bestvina-Mogilski type, 434, 589, 590 Dijkstra-van Mill-Mogilski type, 347 absorbing sequence, 347, 348 system, 346, 347 accumulation point, 462 Addition Theorem, 178 adjunction space, 507, 508 affine combination, 111, 114 coordinates, 114, 125, 126, 134 function, 112-114 continuous, 113 hull, 111 subspace, 111-113, 124 Alexandroff compactification, ^12 problem, 253, 254, 587 almost zero-dimensional, 167, 168, 189, 192, 583 analytic set, see set, analytic Anderson Theorem, 9, 579, 588 antipodal map, 148, 238, 461 points, 149, 461
Baire Category Theorem, ^82 Baire space, 3, 36, 66, 78, 79, 373, 378, 379, 383-385, 387, 391, 392, 446, 482-485, 526, 580, 591-593 hereditary, 393, 446 not topologically complete, 4&4 ball closed, 4^9 open, 459 Banach space, 3, 4, 9, 10, 12, 17, 21, 370, 394, 579 barycenter, 125, 126, 128, 129, 138, 139 barycentric subdivision, 129-132 triangulation, 129, 130 base, 41, 64, 122, 153, 167, 177, 238, 239, 285, 288, 369, 465, 467, 497, 507 closed sets, 494, 497 collection of sets, 441 countable, 593 Ellentuck topology, 250, 251 hyperspace, 101 inverse limit, 81 local, 19, 183, 368, 467, 468 Wallman, 93, 185, 494-500 basic core set, 342, 343 Bernstein set, 580 Bing compactum, 210, 212-215 Shrinking Criterion, 67, 580 bonding map, 80, 84, 93 Borel complexity, 445, 590 619
620
SUBJECT INDEX
set, 80, 373, 455, 456, 517-519, 523, 526, 586, 590, 595 absolute, 18, 437, 456, 526, 592 homogeneous, 581 rigid, 581 Borsuk Antipodal Theorem, 149 Example, 301, 580 Homotopy Extension Theorem, 38 controlled, 265 Problem, 588 Borsuk- Ulam Theorem, 149, 174 boundary, 20, 125, 196, 460, 464 geometric, 114 preserving homeomorphism, 311 set, 589 bounded component, 505 function, 29, 174, 368, 445 metric, 459, 463 set in a normed linear space, 3, 7, 20, 21 subspace, 407, 410, 477, 478 Brouwer Dimensionsgrad, 221 Fixed-Point Theorem, 143, 145, 148, 149, 260, 587, 594 Invariance of Domain Theorem, 197 Rn 56 R m , 166 Cantor set, 42-48, 75, 176, 229, 238, 254, 257, 387, 388, 434, 436, 581 Cantor-Bendixson Theorem, ^67 Cantor-manifold, 207, 208 capset, 329-331, 333, 334, 337, 338, 341343, 347, 354, 589 carrier, 119, 134, 284 Cauchy sequence, 4, 33, 58, 96, 4^4i 468, 469 Cauty Examples, see Example(s), Cauty Cech-Stone compactification, 368, 590 centered, 495 maximal, 495 characterization absolute F^g, 519 analytic set, 77 ANR, 284 A(N)R-pair, 267, 277 AR, 39, 289 boundary point, 196 Cantor set, 43
capset, 334, 337, 590 compact, 460 CP(X) Baire, 379 Cp(X) ^B(Q) 0 0 , 444 Cp(X} metrizable, 369 dimension, 160, 166, 174, 182, 195 finite dimensional manifold, 588 homogeneous Borel set, 77 indecomposable continuum, 86 hereditarily, 94 inessential map, 514 1,51 LC n , 296, 298 LCn and C", 300 ^-manifold, 588 meager filter, 389, 434 Menger manifold, 588 P, 76, 77 Q, 294 Q-manifold, 588, 589 Q, 76 topologically complete, 480 zero-dimensional, 58, 581 C-imbedded, 410 class of spaces closed hereditary, 344 topological, 344 clopen, 41, 50, 57, 58, 63, 443, 460, 500 closed ball, 459 base, 494 map, see map, closed shrinking, 485 closure, 460 coanalytic set, 594 color, 187 number, 250 open, 251 open or closed, 251 colorable, 188 coloring, 187, 188, 238, 241, 242, 244, 245, 248, 249, 251, 586 combinatorially equivalent, 123 compact-open topology, 19, 20, 31 compactification, 57, 64, 87, 93, 174, 182184, 188, 193, 207, 234, 253, 257, 472, 473, 478, 500, 510, 516 Alexandroff, 472 Cech-Stone, 368, 590 dimension preserving, 183 equivalent, 47% one-point, ^72
SUBJECT INDEX remainder, 183 Wallman, 93, 185, 494, 498, 499, 593 compactness deficiency, 183, 204 degree, 182, 183 compactum Bing, see Bing compactum Henderson, see Henderson compactum complete metric, see metric, complete topologically, see topologically complete with respect to a norm, 16 completely Ramsey set, 239 complex conjugate, 512 component, 37, 156, 167, 204, 209, 285, 467, 500, 501, 505, 507 path, 503 composant, 87, 94 cone, 21, 39, 70, 89, 302, 515, 516, 581 connected, 54, 55, 57, 58, 64, 85, 102, 156, 206, 207, 209, 473, 503-505 in dimension n, 295 locally, see locally, connected path, see path-connected continuous image C, 46 complete space, 522 Menger curve, 585 one-dimensional space, 203 P, 77 zero-dimensional space, 57, 58 continuous logarithm, 513 continuum, 37, 51, 58, 85-89, 94, 95, 105, 176, 209, 210, 216, 261, 501, 502, 505, 515, 581, 593 sin(V x ), 29, 57, 503, 505, 517 Cook, 447 decomposable, 86 from A to B, 206, 504 hereditarily indecomposable, 87, 94, 106, 108, 212, 213, 258, 260, 587 homogeneous, 215, 216 indecomposable, 86, 87, 94 homogeneous, 215 Peano, 58, 226, 292, 294, 295, 300, 340, 585, 588 unicoherent, 226, 514 Continuum Hypothesis, 47, 384, 393 continuum-connected, 206, 207, 209 contractible, 29, 39, 70, 146-148, 274, 285, 289, 301, 511, 512, 515, 516, 580
621
locally, see locally, contractible contraction, 511 convex, 1, 2, 15, 16, 19, 20, 23, 25, 29, 124, 125, 148, 288, 581, 582 combination, 1, 2 hull, 1 metric, 477, 479, 588 coordinate functions of a simplex, 114 space of inverse sequence, 80 countable closed sum theorem, 163 dense homogeneous, 63-66, 580 not homogeneous, 63 dimensional, 155, 156, 221, 252, 253, 428 count ably compact, 413 continuous, 257 cover, 458 locally finite, 4 $6 refinement, 458 star-finite, 135, 136 star-refinement, 4®0 covering dimension, 152, 160, 583 CP(X), 4, 368-375, 377-379, 383-387, 393-399, 402-404, 406, 407, 409-415, 425-428, 441, 442, 444-447, 452, 455, 456, 590-592 C*(X), 368, 371, 394, 396-399, 418, 421, 425, 426, 445, 590, 591 Criterion Bing Shrinking, 67 Inductive Convergence, 59, 65, 325 Curtis-Schori-West Hyperspace Theorem, 291, 295 cut, 222, 504 need not be a partition, 221, 505 point, 51, 504, 505 C(X), 4, 5, 19, 20, 31, 368, 370-372, 393, 394, 418 c 0 , 8, 354, 360, 435, 436 decomposable continuum, 86 deformation, 510 property, 334, 337, 343 through a subset, 266-268, 276, 294, 327, 355, 360, 361 degree compactness, 182, 183 partition, 153
622
SUBJECT INDEX
6-multiplicative, 517, 522 derivative, 417 derived set, 417 descriptive complexity, 373-375, 377, 446, 590 set theory, 483, 517, 581 diameter, 459 dimension at a point, 227, 230 coloring, 242 compactification, 183, 184, 186, 187 component, 167, 204 countable dimensional, 155, 214, 252, 254 covering dimension, 152, 160, 583 CP(X), 413, 414, 428, 456 Dimensionsgrad, 221-224, 226 hyperspace, 261 I", 165 infinite-dimensional, 156, 157, 168, 221, 252, 260, 456 hereditarily, 254, 256 inverse limit, 167, 193 large inductive dimension, 180 modulo a class, 183
counterexample, 224 discrete, 466 disjoint-cells property, 294, 588 dominating space, 274, 276 controlled, 274 Dranisnikov Example, 24 dual, 399, 400, 402-404 Dugundji system, 22, 23, 29, 278 theorems, 23, 29, 394, 451, 588 dyadic solenoid, 85, 86, 94, 215 dynamical system, 91, 92, 249 Ellentuck topology, 239, 240, 250, 251 endface, 310, 320, 322, 324, 326, 334, 462, 473 enlargement A(N)R, 304, 305 theorem, 175 e-map, 34, 66, 67, 90, 135 equiconnecting function, 276 equicontinuous, 418 Erdos space, see Example(s), Erdos essential continuous function, 512, 516 family, see family, essential euclidean space, 3, 64, 65, 149, 151, 461 topology, 19, 20 evaluation, 399 Example(s) Bing, 210, 212 Borsuk, 580 Cantor, 42 Cauty, 24, 26, 264, 445, 456, 580, 583 Cook, 447 Dranisnikov, 24 Erdos, 542, 547 Erdos, 50, 58, 160, 167, 182, 183, 437, 584 Henderson, 214 Kulesza, 220 Kuratowski, 447 Marciszewski, 447 Mazur, 241 Pol, 254, 447, 450, 587 Roy, 584 Sher, 589 Urysohn, 221, 594 expansion homotopy, 582
SUBJECT INDEX hyperspace, 291, 292, 294 exponential function, 512 extender, 393 continuous, 454 linear, 394-396 linear, 393 measurable. 455 extension, 21, 23, 24, 38, 40, 46, 72, 98, 148, 188, 193-196, 204, 265, 270, 273, 276, 277, 284, 288, 296-298, 300, 319, 325, 394, 451, 491-494 face of a simplex, 114, 125, 126, 128, 132, 138, 139, 174 opposite, 146, 147 proper, 114 factorization oiC^(X), 396-399 faithful indexing, 458 family Borel sets, 517 Cantor sets, 45 closed under supersets, 389 composants, 94 ^-multiplicative, 517 equicontinuous, 418 essential, 151-153, 155-157, 206, 209, 251 inessential, 151, 152, 156, 201, 202 open sets, 278 pairwise disjoint homeomorphs, 48 cr-additive, 517 solenoids, 581 strongly discrete, 377-379 fiber, 458 filter, 387, 389, 390, 434-436, 439, 441, 458 fixed, 459 Frechet, 435 free, 459 generated by, 459 finite dimensional, 153 fixed-point, 145, 148 free, 60, 92, 148, 188, 241, 249-251 homeomorphism, 242, 245, 248 involution, 238, 586 property, 92, 144-146, 148, 516, 517 theorem Brouwer, see Brouwer FPT Schauder, see Schauder FPT
623
Freudenthal's Approximation Theorem, 137 Fa, 87, 227, 230, 374, 466, 467, 517, 518, 522, 595 absolute, 518 3>-absorber, 347, 350, 361 Fa&, 377. 442, 445, 446, 518, 592 absolute, 344, 518-520 3^-absorber, 350, 353, 360, 361, 442, 444 full realization, 270, 273, 277, 284, 296 simplex, 141 ^-ultrafilter, 495, 496 function vanishing on a set, 396, 418 functional, 399, 400, 402 Fundamental Theorem of Dimension Theory, 165 >G: Galvin-Prikry Theorem, 586 Gs, 35, 172, 175, 217, 220, 221, 229, 231, 238, 257, 372, 373, 375, 427, 428, 434, 454, 464, 466, 467, 480, 483, 485, 493, 494, 517, 526, 590 absolute, 518 enlargement, 175, 304 selection, 217 G5a, 518 absolute, 344, 518-520 general position, 168, 169 geometric boundary, 114 dependence, 111, 112 independence, 112, 125, 126 interior, 114, 125 realization of a simplicial complex, 116 simplex, 114, 131, 138 graph method of Klee, 581 of a function, 12, 458, 463 group homeomorphisms, 35 topological, see topological group
Hamel basis, 18, 19 for L(X), 400 Hausdorff distance, 95 metric, 95, 96, 106, 107 space, 463 height, 418
624
SUBJECT INDEX
Henderson compactum, 213, 214, 254, 587 hereditarily indecomposable continuum, see continuum, hereditarily indecomposable Hilbert cube, 20, 26, 57, 60, 63, 66, 69, 73, 144, 147, 151,176,251-253, 257, 291, 294, 295, 309, 310, 325-331, 333, 334, 337, 338, 343, 345, 346, 348, 354, 355, 360, 361, 414, 426, 428, 436, 439, 441, 444, 461, 462, 470, 473, 479, 526, 581, 589, 590 manifold, 588 Hilbert space, 8, 9, 20, 21, 197, 437, 579, 588, 589 manifold, 588 Homeomorphism Extension Theorem, 320, 325 homogeneous, 63, 64, 66, 73, 75, 80, 94, 147, 168, 215, 216, 462, 4&4, 581, 582 countable dense, 63-66, 580 isotopically, 92, 94, 328 strong, 75 strong local, 64, 66, 73, 75, 580 homotopic, 38, 195, 263, 265, 510 homotopically trivial, 285, 288-290, 301, 305, 310 homotopy, 70, 71, 148, 510 expansion, 582 level, 70, 510 limited by a cover, 263 type, 276, 588 hyperspace, 95, 96, 101, 102, 107, 108, 156, 292, 587 Absolute Retract, 292 continua, 103, 261, 587 diam is continuous, 107 expansion, 291, 292, 294 finite sets, 100, 291, 294 Hilbert cube, 295 map, 98, 100 topologically complete, 98
dense, ^60 Wn, 174 open, 460 IP, 45 product, 586 Q, 470 R°°, 480 R 2n + 1 , 583 Z, see Z-imbedding indecomposable continuum, 86, 87, 94 homogeneous, 215, 216 Inductive Convergence Criterion, 59, 65 inessential continuous function, 512 family, 151, 152, 156, 201, 202 infinite left product, 59 infinite-dimensional, see dimension, infinitedimensional strongly, see dimension, strongly infinitedimensional Infinite-dimensional topology, 588 initial segment, 46 inner product space, 18 integers, 457 interior, 124, 460 geometric, 114 interval 1,^57 5,457 invariance domain, 197 Whitehead torsion, 588 invariant, 151, 242, 250, 251 inverse limit, 80, 82-85, 90-94, 137, 167, 193, 220, 249 sequence, 80-85, 90-94, 137, 167, 220, 249 subsequence, 83 involution, 238, 586 irrational numbers, 42, 45, 58, 76, 77, 168, 457, 482, 483, 494, 581 irreducible function, ^77, 479, 526 partition, 155, 157, 193, 226 isometry, 19, 96, 460 isotopically homogeneous, 92, 94, 328 isotopy, 70, 71, 73, 320, 328 joined by a path, 503, 504 Jordan Curve Theorem, 174
SUBJECT INDEX K-function, 134, 135, 275, 488, 582 Kuratowski-Wojdyslawski Isometric Imbedding Theorem, 19 Kuratowski-Zorn Lemma, 18, 4.59 large inductive dimension, 180 Lavrentieff Theorem, 493 Lebesgue measure, 587 number, ^75 Lemma A-system, 592 Disjoint Refinement, 453 Kuratowski-Zorn, 18, 4^9 length of a sequence, 46 L £ -enlargement, 189 ^-equivalent, 370, 371, 399, 413, 425, 445, 456 level homotopy, 70, 510 preserving, 71 Whitney, 105, 106, 108, 109, 258 lexicographic order, 216 L-imbedded, 189, 193, 234 limited by a cover, 65, 263, 265, 267 linear function, 12, 17, 20, 21 homeomorphism, 12, 20, 21, 369, 370, 398, 403, 411-413, 447, 450 hull, 111 order, 126, 216 space, 1-4, 8, 10, 12, 15-21, 26, 29, 111, 113, 114, 116, 123, 130, 131, 148, 197, 252, 288, 361, 363, 364, 369, 402, 435, 452 incomplete, 589 not ANR, 264, 580 not locally convex, 20 unit sphere, 20 subspace, 20 linear space, 29 linearly homeomorphic, 369, 370, 398, 403, 404, 447, 450 independent, 5, 7, 18, 20, 112 Lipschitz function, 99, 490 local base, 19, 183, 368, 467, 468 locally compact, 58, 64, 87, 121, 425, 456, 473, 478, 482
625
nowhere, 57, 58, 76 connected, 55-58, 86, 102, 222, 224, 226, 285, 288, 467, 505, 516 in dimension n, 296 contractible, 28, 296, 301, 516 at a point, 516 convex, 2, 18, 19, 23-25, 29, 288, 402, 452, 583 IP is not, 20 finite, 22, 108, 136, 138, 157, 158, 160, 409, 486, 487, 490 simplicial complex, 118, 121, 125, 273, 277, 284 homotopy negligible, 263, 588 path-connected, 29, 503, 504 uncountable, 568 lower semi-continuous, 404, 406, 4^8 set-valued function, 13 I* -equivalent, 370, 418, 419, 421, 425, 456 Lusternik-Schnirelman-Borsuk Theorem, 148, 149, 174, 586 L(X), 399, 400, 402-404 map closed, 109, 241, 460, 464, 467 dimension raising, 199 e, 34, 66, 67, 90, 135 essential, 512 inessential, 512 irreducible, 477 linear, see linear, function monotone, 504 open, 107, 108, 464, 477 quotient, 108, 505, 506, 508, 509 Sperner, 141 U, 169 Whitney, 103-106, 108, 109, 258 Mazurkiewicz Theorem, 55 meager, 66, 87, 447, 449, 483, 520, 522, 526 at a point, 526 collection of sets, 389 filter, 389-391, 434, 436 Menger compacta, 588 curve, 583, 585 mesh collection, 459 simplicial complex, 130 subdivision, 130, 131
626
SUBJECT INDEX
metric, 30, 90, 361, 363, 463, 464, 478, 484 admissible, 460 bounded, 459, 463 complete, 3, 21, 33, 35, 58, 96, 98, 463, 474, 477, 479, 480, 485 convex, 108, 477, 479, 588, 593 derived from a norm, 2, 3 Euclidean, 3 Hausdorff, 95, 96, 106, 107 Nagata, 586 non-Archimedean, 57, 58 nonconvex, 325 s, 19 totally bounded, 96, 98, 479, 484 Metrization Theorem Nagata-Smirnov-Bing, 593 Urysohn, 465, 593 Miljutin Theorem, 370 monotone map, 504 multi-valued function, 404 Nagata-Smirnov-Bing Theorem, 593 natural numbers, 457 near homeomorphism, 66, 67, 70, 90, 94 neighborhood retract, 512 nerve, 132, 133 net, 484 Nobeling's universal space, 168, 172, 174 nodec space, 385 norm, 2, 3, 579 Euclidean, 3 normal set in the sense of Frechet, 221 nowhere countable, 581 dense, 86, 197, 387, 388, 436 locally compact, 57, 58, 76 cr-compact, 581 topologically complete, 581 null-sequence, 508 nullhomotopic, 145, 148, 149, 204, 300, 510, 511, 514 open ball, 459 Ellentuck topology, 240 map, 10, 57, 107, 108, 460, 464, 477 problem, see problem, open shrinking, 4-85 swelling, 157, 158 order
cover, 160 lexicographic, 216 linear, 126, 216 preserving indexed collection, 344 origin of Q, 461
paracompact, 125, 369, 396, 487, 490 partial realization, 270, 273, 277, 281, 282, 284, 296 partition, 146, 153-155, 157, 175, 179, 182, 193, 206, 209, 221, 252, 504, 505 Bing, 210, 213 continuum, 502 cut is, 222 degree, 153 irreducible, 155, 157, 193, 226 R, 80 space, 125, 257, 443, 509 unity, 487, 488 path, 503 path-component, 503 path-connected, 29, 55, 57, 58, 94, 285, 503-505, 511 locally, 29, 57, 503, 504 Pea.no continuum, see continuum, Peano map, 580 perfectly disconnected, 385, 592 point accumulation, 462 cut, 51 P, 393, 446 weak P, 386 pointwise convergence topology, see topology, pointwise convergence Pol Examples, 447, 450, 587 polyhedron, 124, 136, 274, 310, 588 polytope, 123, 132, 136, 166, 270, 274, 276, 304 power set, 458 P-point, 393, 446 problem Alexandroff, 253, 254, 587 Anderson-Bing, 589 Banach and Frechet, 579 de Groot, 584 open, 145, 161, 216, 236, 260, 404, 414, 425, 446, 447, 580, 581, 587, 592 Schauder, 582
SUBJECT INDEX product, 26, 41, 58, 82, 235, 339, 368, 371, 450, 461, 467, 480, 516, 519 Cartesian, 457 weak, 355 projection, 70, 81, 84, 398, 457, 477, 478 property closed hereditary, 84 countably productive, 84 fixed-point, 92, 145, 146, 148, 516 property of Baire, 434, 520-523, 594 pseudo arc, 215 boundary, 60 interior, 60 punctiform, 221 Q-set, 595 quotient map, 108, 505, 506, 508, 509 quotient topology, 398, 505, 506, 509 not metrizable, 509 Ramsey set, 239 completely, 239, 240 theory, 238, 240 rational numbers, 42, 58, 76, 78-80, 182, 384, 416, 457, 483, 494, 517, 519, 520, 581, 594 not topologically complete, 483 real numbers, 457 realization full, 270, 273, 277, 284, 296 geometric, 116 partial, see partial realization refinement, 160, 166, 176, 263, 273, 277, 284, 296, 458 disjoint, 41, 168, 453, 463 star, 272, 460 star-finite, 136, 166, 272 reflexively universal, 345 retract, 146, 148, 276, 512, 516, 517 neighborhood, 26, 512 retraction, 512 rigid space, 447, 581 s, 3, 9, 19, 20, 26, 60, 66, 257, 310, 311, 314, 319, 320, 322-324, 339, 342, 343, 364, 579, 587, 588 scattered, 417, 419, 456 height, 418
627
Schauder Fixed-Point Theorem, 148, 583 segment initial, 46 straight-line, 19 selection, 13, 29 Gs, 217 continuous, 13, 16 semi-continuous lower, 13, 406, 488 upper, 257, 489 separated, 491 separating set, 148 separator, 504 sequence absorbing, 350 compacta, 4.18 ^-absorbing, 352 functions, 469 homeomorphisms, 59, 65 inverse, see inverse, sequence null, 58, 508, 509 sequentially compact, ^75, 4^4 set analytic, 48, 49, 77, 435, 436, 439, 441, 455, 522, 523, 526, 590, 594 Borel, see Borel set Cantor, see Cantor set coanalytic, 594 derived, 417 meager, 4^3 power, 458 property of Baire, 434, 520-523, 594 Ramsey, see Ramsey, set scattered, 417 aZ, see aZ-set Z, see .Z-set set-valued function, 12 shift inverse limit, 91, 92, 249 M z , 587 shrinkable, 66-68, 70 shrinking, 159, 160, 167, 244, 485, 486 closed, 4$5 open, 485 Sierpiriski Carpet, 583 Theorem, 502 a, 257 cr-additive, 517, 522 cr-algebra, 455, 458, 521
628
SUBJECT INDEX
a-compact, 58, 66, 226, 228, 339, 347, 361, 411, 412, 426, 479, 518, 519, 581 nowhere, 581 cr-discrete, 369 aZ-set, 307, 310, 334, 340, 346, 361, 436, 441 simple chain, 54 simplex, 114, 124 n-dimensional, 114 full, 141 geometric, 114 simplicial complex, 116, 117 locally finite, 118 simplicially homeomorphic, 123 sin(y x )-continuum, 29, 57, 503, 505, 517 skeleton, 117, 122, 271 solenoid dyadic, 85 p-adic, 581 space, 465 adjunction, 507, 508 Baire, see Baire space Banach, see Banach space euclidean, 3, 64, 65, 149, 151, 461 infinite-dimensional compacta, 350 linear, see linear, space normal, 121 perfectly disconnected, 385, 592 rigid, 447, 581 scattered, 417 topological, 465 Tychonoff, 367 space filling curve, 46 Sperner Lemma, 141 map, 141 sphere, 3, 9, 20, 26, 29, 63, 85, 92, 146149, 194-196, 201, 204, 207, 226, 288, 291, 297, 310, 461, 473, 511, 516 square, 87, 89, 93, 384, 450 standard simplex, 539 triangulation, 117, 139 star, 119, 460 finite, 135, 138, 166 refinement, 272, 460 straight-line segment, 19 parametrization, 361 strong local homogeneity, 64, 66, 73, 75, 580
strongly discrete family, 377-379 homogeneous, 75 infinite-dimensional, see dimension, stron infinite-dimensional subbase, 19, 31 subcomplex, 117 subpolyhedron, 124 subpolytope, 124 Subspace Theorem, 164 sum, 371 sup-norm, 4 support, 404-407 homeomorphism, 64 swelling, 157, 158 symmetric difference, 457 t-equivalent, 371, 414, 426, 428, 445, 456, 591, 592 t* -equivalent, 426, 456 Theorem Addition, 178 Anderson, 9, 579, 588 Borsuk, 149 Borsuk-Ulam, 149, 174 Brouwer, 143, 145, 148, 149, 157, 260, 587, 594 Brown, 90 Cantor-Bendixson, 467 Cauty, 24, 583 Chapman, 588 Closed Graph, 12 Coincidence of Dimension Functions, 180 Countable Closed Sum, 163 Curtis- Schori- West, 291 Dobrowolski-Marciszewski-Mogilski , 444 van Douwen, 48 Dugundji, 23, 29, 394, 451, 588 van Engelen, 581 Freudenthal, 137 Galvin-Prikry, 586 Jordan, 174 Keller, 581 Kuratowski-Wojdyslawski, 19 Lavrentieff, 4^3 Lusternik-Schnirelman-Borsuk, 149, 174 Marciszewski, 398, 447 Mazurkiewicz, 55 Michael, 16 Miljutin, 370
SUBJECT INDEX Nagata-Smirnov-Bing, 593 Pol, 254 Schauder, 148, 583 Sierpiriski, 502 Souslin, 47 Sperner, 141 Subspace, 164 Tietze, 24 Toruriczyk, 291, 294, 588, 589 Urysohn, ^65 West, 588 topological group, 1, 37, 60, 182, 215, 462 homeomorphisms, 35-37, 583 homogeneous, 4^4 Q is not, 60 S1, 85 £ 2 , 85 topological property closed hereditary, 84 countably productive, 84 topological space, 4^5 topologically complete, 47-49, 55, 57, 65, 66, 76, 77, 183, 222, 226, 480, 482, 484, 493, 518, 519, 522, 523, 526, 581, 594 A(N)R enlargement, 305 Baire, 482 compactification, 253 Continuum Hypothesis, 47 c;(x), 591 CP(X), 415 C(X,Y], 33 dense subspace, 4$4, 526 dimension, 220, 224, 584 I2, 9 H(X,Y), 35 hyperspace, 98 inverse limit, 84 nowhere, 581 P, 482 Q is not, 483, 517 topological sum, 4&4 totally disconnected, 183, 218, 221 weakly n-dimensional, 231 topology compact-open, 19, 20, 31 coordinatewise convergence, 9, 467 euclidean, 19, 20 pointwise convergence, 4, 368, 372 quotient, see quotient topology Tychonoff product, 368
629
uniform convergence, 7, 31 Vietoris, 101, 108, 582 weak, 354 Whitehead, 118, 119, 132 Toruriczyk Theorems, 291, 294, 588 totally disconnected, 50, 58, 167, 183, 218, 221 tower, 337 deformation property, 334, 343 expansive, 334 translation, 462 triangulation, 117 standard, 117 trivial homotopy class, 511 Tychonoff product topology, 368 space, 367 type homotopy, 276 of a point, 63 It-close, 263 ultrafilter, 391-393, 459, 462, 463, 495, 496 U-map, 135, 136, 166, 169, 176 unbounded component, 505 unicoherent, 204, 226, 514, 515 uniformly continuous, ^77, ^7# union operator, 99 unit ball, 3 Rn, 3, 461 interval, 451 sphere, 3 M n , 3, 461 universal curve, 583, 585 ya, 354 ?«S, 439 M r , 344, 345 reflexively, 345, 350, 355, 364, 439 space compact, 176 Nobeling, 168, 172, 174 weakly n-dimensional, 232 strongly yaS, 361 M, 348, 361 MS, 348 M r , 344, 346 upper semi-continuous. 257, 4&9
630
SUBJECT INDEX
decomposition, 106, 501, 505-509 Urysohn function, J^69 Metrization Theorem, 4 65, 593 usual topology on [N]w , 239 vertex, 116-118, 280 Vietoris topology, 101, 108, 582 Wallman base, 93, 185, 494-500 compactification, 93, 185, 4^4i 4^8, 499, 593 Warsaw circle, 301 homotopically trivial, 285 weak Cartesian product, 355 topology, 354, 465 weakly infinite-dimensional, 251, 252, 254, 257 not countable dimensional, 254 n-dimensional, 228, 231, 586 universal, 232 weakly one-dimensional, 235 weak P-point, 386 weight, 369 Whitehead topology, 118, 119, 132 torsion, 588 Whitney level, 105, 106, 108, 109, 258 map, 103-106, 108, 109, 258 >Z: zero-dimensional, 41-45, 57, 58, 73, 75, 76, 153, 154, 157, 165, 174, 177, 178, 182, 210, 230, 241, 253, 425, 581 almost, 167, 168, 189, 192, 583 Z-imbedding, 327, 328, 338, 343, 360, 361 Z-map, 327 Z-set, 307-311, 323-331, 333, 334, 337, 338, 343, 589