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!
y. Assume \x\p > \y\p > 0. This means \x\p > p\y\p. We then see oo
oo
h 2k
^
k=0
(p-i)\y\PY,P~2hfc=0
Then we have >(*) - <j>{y) > \x\p - ^ \ y \
p
> p\y\p (l - -
^
=*
> 0.
On the other hand, if\x\p = \y\p, but a® = bq.ai = b\,... , a^-i = bj-x,a,j < bj, then j
(f>{x) >
\x\pJ2akP~2k fc=0
while 0(y)
< \y\pEi=0ykP"2k + \y\P^p-2U+1) = \x\P Ei=o akP~2k + \X\PVJP~2J + \AvP~2j Tpp-
This means
(XJ - y3 - —-jj
> 0.
So regardless,
for all x,y £ Qp. This means we are dealing with what is called a Lipschitz function. This also means (f> is continuous. Now we wish to look at the image of the whole domain, Qp. It turns out that K = (p(Qp) is related to a Cantor Set. We state the following without proof: In Qp, we can define the sphere with center a and radius pr to be Sr(a) = {x £ Qp : | a — x\p = pr}. Because the only distances in Q are {0} U pr where r £ Z there are countable many radii with which to be concerned. Then we decompose Qp into a collection of spheres all with center a = 0. Since the centers are the same, we will just call these spheres Sr. Now let Kr = 4>(Sr), and K = \Jr^j,Kr. lfr\ ^ r(S0) = Hi* £ Q P : k - 0| p = p0}).
P-ADIC RATIONAL NUMBERS
87
For any x € So, 1 <
= [0,p)
and K0\I leaves a Cantor Set. The details are in [76].
88
P-ADIC NUMBERS AND VALUATIONS
EXERCISES 5.1
Supply the missing proofs for Theorem 5.1
5.2 Prove that ifG is an abelian group, then for any positive integer n, (a*b)n an * bn. Here xn = (x * x * ■ ■ ■ * x).
=
n times Let G be a set with operation # such that
5.3
• G is closed under # . ■ # is associative. • There exists an element e G G such that e # a — a for all a G G. • For every a G G, there is ab e G such that b#a = e. Prove that G is a group. 5.4
Determine whether or not 7L with a*b = a — b is a group.
5.5 A non-empty subset, H, of a group G is called a subgroup of G if, using the product for G, H itself is a group. Let G be the integers 1 under +. Let H be the set of even integers. Prove that H is a subgroup. 5.6 Let G be a group and a G G. If n is a positive integer an is defined as in Problem 5.2; a0 = e, an ■ am = an+m, a~n = ( a " ) ' 1 . Prove that A = {an : n G Z} is a subgroup ofG. This is called the cyclic subgroup ofG generated by a and is denoted by < a >. 5.7
Prove that a cyclic group is abelian.
5.8
Show that any field is an integral domain.
5.9
Prove Theorem 5.2.
5.10 Show that the Euclidean metric on M. is also a norm. Prove that for all x, y G R, \xy\ = \x\ ■ \y\. 5.11
Convert each integer below into thep-adic version for the value ofp specified. a) 23, given p = 2 b) 23, given p — 3 c) 117, given p — 5 d) —6, given p — 2
5.12
Find the p-adic expansion of'1/3 ifp = 2.
5.13
Find the p-adic representation for 1/p.
5.14
Provide a (short) proof that (Z p , +) is, in fact, abelian.
EXERCISES
89
5.15 Our example of a sequence in Q that is Cauchy, but does not converge in the rationals, is X\ = 1 and xn+i = -^ -\ for n>2. L
Xn
What is the real number that is the limit of this sequence? (Note: This is not asking for a proof of convergence, which is assumed.) 5.16 Determine the p-adic distance between a = 252 and b = 140 for p = 2,3,5,7 5.17
Find the representation of a) - 1 / 5 in
5.18
Find the values ofp — 2,3, 5, 7 for which —1 has a square root in Q p .
5.19
Find the value of |90| p ifp — 2,3,5,7.
5.20
Prove that in Q 2 , if\x\2 = \y\2, then \x + y\2 = (l/2)|x| 2 .
5.21
Fill in the details for the diagonal argument to prove Theorem 5.5.
5.22 In many sources, the absolute value of a real number is written using the notation \X\OQ. Using this notation, explain why it is true that n2
Show that for any prime p, lp + Ylk>o(P ~ ^)Pk
=
*V
5.24 Show that in contrast to Theorem 5.7 if'ao = 0, then a = ... 0302^100 cannot have a multiplicative inverse. 5.25 Fill in the missing piece in Theorem 5.9. Suppose we have a non-Archimedian norm. So\\a + b\\ < max{||a||, ||6||}. Show that if \\a - b\\ < \\a\\, then \\a\\ = ||6||.
CHAPTER 6
SELF-SIMILAR OBJECTS
If we look at the construction of the Cantor Set, we see that for any k, a scaled copy of the picture kth level appears in subsequent levels. For instance, when k = 1, the set K\, which is the union of the intervals [0,1/3] and [2/3,1], is represented by the picture
For k = 3, the image of K3 looks like
which can be interpreted as four shrunken copies of K\. This leads us to the ideas of sets being self-similar which, in turn, takes us into the world of fractals and later, in Chapter 7, to various notions of dimension. 6.1
The Meaning of Self-Similar
We will be more formal later, but before we start we should say a little more about what makes a set self-similar. A set that is self-similar has the property that at any The Elements of Cantor Sets -With Applications, First Edition. By Robert W. Vallin Copyright © 2013 John Wiley & Sons, Inc.
91
92
SELF-SIMILAR OBJECTS
magnification there are pieces of the set that have the same shape as the whole. A line segment is a trivial example of such a set, but for a nontrivial example let us look at a fern branch
These are typically generated by repeatedly plugging an image into a system of func tions called similarities and letting the sequence approach a shape called an invari ant. The study of these sets really took off in the last half of the 20th century as re searchers were able to use computers to approximate pictures of these sets and the study of fractals rose in prominence.
6.2
Metric Spaces
The Cantor Set, of course, is a set in the real line, so we will begin this section with K. The usual idea for the distance between two points on the line is the absolute value of their difference. Symbolically we say that ifx,y e K the distance between x and y is d{x,y) = \y-x\. This is referred to as the Euclidean distance on M. and we will sometimes write this as | • |. A metric is the official name for a distance function, and we will begin by looking at the properties that make something a distance function. This was quickly remarked on in Chapter 2. We go into more detail here. Definition 6.1 A metric on a set S is a function d : S x S —»■ [0, oo) which satisfies the following three properties: ■ For all x,y £ S, d(x, y) > 0 (called nonnegativity) and d{x, y) = 0 if and only ifx = y. ■ For all x,y 6 S, d(x, y) = d(y, x) (symmetry) • For all x, y, z 6 S, d(x, y) < d(x, z) + d(z, y) (The Triangle Inequality) Some examples of spaces and accompanying metrics are as follows:
METRIC SPACES
93
1. LetS — R", where x = (xi,x2, ■ ■ ■ ,xn) and y = (yi, y2, ■ ■ ■ ,yn), and d(x, y) = V22/!=i(^fc — Uk)2y the Euclidean metric on R™. 2. Let S be any set and let
[0 ifx = y The function d is referred to as the trivial metric. 3. Let S be the plane R 2 with points written as ordered pairs; so x = (x\, x2) and y = (yi, y2). We may define distance as ~d(x,y) = |x/i - x i | + \y2 - x2\. This is referred to as the taxicab metric. 4. Let S = (—7r/2,7r/2) and define distance as D{x, y) = \ tan(y) — tan(x) |. We should give an example of a proof showing a function on a space is a metric, so we will show one way that we can define distance on the space of continuous functions. fl
EXAMPLE 6.1 Let C[0,1] = {/ : [0,1] —>■ K : / is a continuous function} and define p{f,g)=8w{\f{x)-g{x)\:x£[0,l]}. Then p is a metric (called the sup metric) on the space C[0,1].
■
Proof: First, since absolute value is nonnegative, we know \f(x) — g(x)\ > 0 for all x and hence so is its supremum. Furthermore, the supremum is equal to zero if and only if \f(x) — g(x)\ — Ofor every x £ [0,1] which, thanks to the properties of absolute value, means f(x) — g(x) = 0 or f(x) = g(x) for all x. The second criterion is straightforward. Since \f (x) — g(x)\ = \g(x) — f(x)\ these two quantities have the same supremum and p(f, g) = p(g, / ) . Finally, for any fixed value x and continuous functions / , g, and h, \f(x) - g(x)\ < |/(x) - h(x)\ + \h(x) - g(x)\. Thus sup{|/(x) - g(x)\} < sup{|/(:r) - h(x)\ + \h(x) - g{x)\} < sup{|/(x) h(x)\} + sup{\h(x)-g(x)\}.Sop(f,g) = sup{|/(x) -g{x)\ where x S [0,l}}fits all the requirements to be a metric. 4* We can now take some of the familiar properties ofM. with the Euclidean metric | • | and rewrite them in terms of a metric space, (S,d).
94
SELF-SIMILAR OBJECTS
Definition 6.2 For a point x £ S and a fixed r > 0, the open ball about x with radius r is given by B(x,r) = {s £ S : d(s,x) < r}. Note: What is contained in an open ball depends on the metric (see homework for some interesting examples), but our notation B(x, r) does not state what metric is being used. This should be clear from the context. S
EXAMPLE 6.2 Let S be any set and the metric the trivial metric, d. Then for any point s £ S there are two possible open balls around that point. If 0 < r < 1, then B(s, r) = {x}, while ifr > 1, then B{s, r) = S. m
Just as open intervals in the line generalize to open sets, we can generalize simi larly in a metric space. Definition 6.3 A set A C S in (S, d) is called open iffor every a in A there exists a value r so that B(a, r) C A. Definition 6.4 A set B is called closed if its complement Bc — S \ B is open. S
EXAMPLE 6.3 1. An non-empty open interval in (M, | • |) is an open set. (To generalize this, in any metric space (S, d) pick a point x and positive e. Then B(x, e) is an open set.) 2. The empty set 0 is both an open and closed set regardless of the metric space. 3. The set (—1, 0) U (2,3) in (R, | • j) is an open set. 4. Let S be any set and the metric the trivial metric, d. Then any set A C S is an open set since (for example) for all a £ A, B(a, 1/2) = {a} C A. 5. The set (1, 2] is neither open nor closed in (K, | • J). 6. The set of continuous functions f with /(0) = 0 is a closed set in (C[0,1], sup).
■
A couple of things are worth noting in these examples: First, let us take a closer look at the last example. If A = {/ £ C[0,1] : /(0) = 0}, then we will show A is closed by proving A = C[0,1] \ A is open. Pick a function g £ Ac. So we know (0) ^ 0, say g(0) = c. Let r\ = ^\c — 0|. Then for any function h £ B(g, rf) it must be true that h(0) ^ 0. Thus Ac is an open set and A must be closed. Second we may extend the ideas in examples above to a much larger union as we see in the theorem below.
METRIC SPACES
95
Theorem 6.1 For any metric space (S, d), the countable union of non-empty open sets is an open set. Proof: Note that a finite collection of objects is countable. If there are only finitely many A\, A2,.. ■, Am we can extend this to countably infinite by saying An = 0 for n > m, and so our proof has a union of infinitely many sets. Let A = W^L-^An where each An is an open set To show A is open, let x £ A. Then x G Aj for some j . Now Aj being open means there exists an £ > 0 such that x G B(x,e) C Aj C uAn. This shows A is an open set. 4 Corollary 6.1.1 For any metric space (S, d), the countable intersection of closed sets is a closed set. Proof: This is really the proof above and an application ofDeMorgan 's Law
(uAnf = n(A%). A Definition 6.5 A set A C S in (S, d) is called bounded if there is a point s € S and a number r > 0 such that d(s, a) < r for all a e A. Related to this is the idea of a totally bounded set. This will be useful in our study of fractals. Definition 6.6 Let A C S in (S, d). We say A is totally bounded iffor each e > 0 there is a finite set ofpoints {eti, «25 • • ■, a m } in A so that for any a & A there is an aj with d(a,a,j) < e. fl
EXAMPLE 6.4 In the real line, with the Euclidean metric, the closed interval [0,1] is totally bounded. Given any distance e > 0, let N € N such that l/N < e/2. We can use the set of points given by Xi = ei, where i = 0,1,2,..., 2N. For any x € [0,1] there is an IQ such that ei0 < x < e(i0 + 1) so \x -xio\ < \xio+1\ = e/2 < e. The metric is important. In the trivial metric [0,1] is not totally bounded as any two points must be a distance 1 from each other. ■
96
SELF-SIMILAR OBJECTS
Finally, we extend the notion of continuity to functions whose domain and range are metric spaces. Definition 6.7 Let f be a function with domain (S, d) and co-domain (T, p) (usually written f : (5, d) —> (T, p)). Then f is continuous at XQ £ S iffor every e > 0, there exists a 5 > 0 such that ifx £ S with d(xo, x) < 6, then p{f{x0),f(x))
<£.
Furthermore, f is a continuous function on S if it is continuous at each XQ £ S. Recall the definition of continuity we used when we dealt with only topological spaces, without a definition of distance. These spaces used the idea of open sets and their inverse images. The theorem below shows that idea still holds with metric spaces. Theorem 6.2 Let f : (5, d) —>• (T, p). Then f is a continuous function if and only if for any open setV £ T its inverse image f~l{V) is an open set in S. Proof: Since this is an "if and only if" theorem, we have two directions to prove. Note that we will deal with open balls in both the metric d and metric p. We are not differentiating between them in our notation. First, assume that f : (S, d) —> (T, p) is a continuous function and that V is an open set in T. If f~l{V) = 0, then we are done since 0 is an open set. Let us assume there is an XQ € S such that /(.To) = v £ V. Since V is open, there is an e > 0 such that v £ B(v,e) C V. That f is continuous means that for this e there is a 5 > 0 such that for all x £ B(XQ, S) we have p(v, f{x)) < e. This makes B(XQ, 5) a subset of f~l(V) hence f~l(V) is an open set. In the other direction, we start with the fact that the inverse image of every open set is an open set. Pick XQ £ S and let e > 0. Then V = B(f(xo),e) (1 T is an open set in T. Our hypothesis given us that f~y{V) is an open set in S. The definition of open set means there is a positive 5 such that B(XQ,6) C f~l(V). Thenify £ B(x0,5) we have p(f(xo),f{y)) < e. This proves f is continuous at xo and since this point was arbitrary, f is a continuous function. 4k Let us now provide proof of Theorem 3.13 from the last chapter. Recall the theo rem tells us, "Any compact metric space is the continuous image of the Cantor Set." Before we can proceed we need the following definition and fact. A space Y is called a Hausdorff space iffor any two points x,y £ Y there are open sets U and V in Y with x £ U, y £V and
t / n y = «. Any metric space is a Hausdorff space. However, this is a topological property, so a metric is not required to be on the space. There are topological spaces that are not Hausdorff. Fact: Let / : X —> Y where X is a compact metric space and Y is a Hausdorff space. If / is 1 — 1 and continuous, then / is a homeomorphism between X and the co-domain, f(X). The proof can be found in most topology books.
SEQUENCES IN (S, D)
97
Proof: (for Theorem 3.13) This proof is due to Rosenholtz (see [65]) and consists of a collection of small results put together. First, let I" be the product of countably infinitely many unit intervals; Iu — [0,1] x [0,1] x [0,1] x ■ • •. Then any compact metric space (X, d) is homeomorphic to a subset of IM. Without loss of generality we can assume the metric is bounded by 1. After all, d is equivalent to the metric d/(l + d). Since X is compact, it contains a countable dense set, call this xi, X2, ■ ■ ■■ Define H : X —>■ Iu by H(x) = (d(x,Xi),d(x, X2),d(x,X3),...). This function is continuous because each d(-, Xj) is a continuous function. H is one-to-one because if x 7^ y then there is some point Xj that is closer to one of x or y than it is to the other. Hence H(x) ^ H(y). Since X is compact and Iu is Hausdorff (being the product of Hausdorff spaces) our fact above shows that H is a homeomorphism onto H(X). We have already seen that the unit interval is the continuous image of the Cantor Set. This is the function g{^2a>i/3l) — J2ai/2l+l where a* — 0 or 2. The Cantor Set itself is homeomorphic to the countable product of sets of the form {0, 2}. The function h : C -* {0, 2} x {0, 2} x {0, 2} x • • • is the function where h(^2 di/31) = (0,1,0,2,0,3,...). This is a rather straightforward choice of function. Because the countable product of countable sets is itself countable (Theorem 3.6) we can extend our result to this: The Cantor Set is homeomorphic to the countable product of Cantor Sets. The set Iu is the continuous image of the Cantor Set. From the result above, any point in C can be thought of as (xi,X2,x-s,...), where each X{ € C. Thenwe create the function F : C —> I" where F(x1,x2,x3,...)
=
{g(xi),g(x2),g(x3),...).
Now let K C C be closed. Then K is the continuous image of C. The proof of this uses the "obvious" fact that C is homeomorphic to C, where C" = {x = ^2 bz/6l where bi = 0 or 5}. This new C has the additional property of being nonaveraging; if x,y E C, then (x + y)/2 £ C. Assume K' is a closed subset of C Then if x € C there is a unique point k £ K' where d(x,k) = d(x,K) = inf {d(x, y) : y £ K}. The function T : C —>• K' given by T{x) = k is continuous and onto, and is the function which we were working to find. There are other ways to prove this result, using heavier topology concepts. See [43] and [69]. 4(t 6.3
Sequences in (S, d)
A sequence in a metric space (S, d) can be defined as a function f with domain N and co-domain S. However, rather than writing f : N —» S we usually rely on the typical notation {i„}™=1 (or {xn} if the index is understood). Colloquially, we say {xn} converges to a point x, if as n grows, the distance between xn and x becomes arbitrarily small. To talk about distance precisely we need to pay attention to the metric used. So officially: Definition 6.8 Let (S,d) be a metric space. We say the sequence {xn} C S con verges to the point x € S iffor every e > 0 there exists a natural number N such
98
SELF-SIMILAR OBJECTS
that ifn > N, we have d{xn,x) S
< e.
EXAMPLE 6.5 In (R, j • |) the sequence given by xn = ^ - j - converges to 1. To prove this, let e > 0. For this e, set N £ N such that N < l/e. Then ifn > N, we have
ra+1
1
-1 ra + 1
1 n+1
1 1 < n- < — N <e
and we /jave our proof. S
■
EXAMPLE 6.6 /n 5 vw'tfi ?/ze trivial metric, d, every sequence is a convergent sequence. The proof of this will be left as a homework exercise. ■ In keeping with our example of the sup metric we have the following example:
S
EXAMPLE 6.7 The sequence fn(x) = cos(ir/n) converges in (C[0,1], p) to the function f(x) = I. M
Before we go too far, let us take a step back to work on sequences of functions and seeing from where their limits come. Let {fn} be a sequence of real-valued functions all with the same domain D. Being real-valued means for any x £ D, the image f(x) £ K. Thus if we fix a value x G D, the sequence {fn{x)} is a sequence of real numbers and, if lira f(xn) exists, we can call that the image of x under the the limit function f. Rephrasing that we have the next definition. fl
EXAMPLE 6.8 Let {fn} be a sequence of real-valued functions defined on the set D. We say {fn} converges pointwise to f iffor each x G D, the sequence {fn(%)} con verges in K to f{x). Notationally, we write this as fn —» / . ■
S
EXAMPLE 6.9 The sequence gn[x) = xn with domain [0,1] converges pointwise to
1,
X = 1.
SEQUENCES IN ( 5 , D)
99
Let us go as far as to look at the proof of this pointwise convergence. We wish to show that if we fix XQ, then for every e > 0 there is a N £ N such that if n > N, then \gn(x0) - g(x0)\ < e. Proof of Example 6.9: Fix XQ e [0,1] and let e > 0 be given. If x0 = 1, then gn(xo) = 1 = g(xo) and so \gn(xo) ~ 9{xo)\ < e for all n and the conclusion is trivially true. Assume XQ ^ 1. For this XQ and e there exists an N e N such that ln(xo) Then for n > N, \§n{xo)
~ g(,X0)\
= XQ < XQ < 6.
Thus gn converges pointwise to g. 4k Two things worth pointing out here. First, the definition ofN in the proof involves both e and XQ. This dependence on XQ is a clue that the convergence is pointwise. Secondly is how important the space is for Example 6.9. If the space is the set of all functions with bounded range, we will write this space BD[0,1], and we are fine, g fits that bill. However, g is not an element o/C[0,1], since g is not a continuous function despite all the gn being continuous. If we have put the sup distance on the set BD[0,1], gn still does not converge to g G BD[0,1] as that supremum of \g{%) - 9n{x)\ = 1 for all n. In order to create the fractal-type objects that are our goal, we will need to take sequences and transform them via a set offunctions. The following theorem shows that if we know the function is continuous, then the property of being a convergent sequence is preserved. Theorem 6.3 The continuous image of a convergent sequence is convergent. Proof: Suppose that f : (S, d) —► (T, p) is a continuous function and that {xn} is a sequence in S which converges to x 6 S. We claim {f(xn)} converges to the point f{x). Pick an e > 0 and consider V = B(f(x),e). Since V is an open set in T, f~l{V) is an open set in S that contains x. Because {xn} converges to x, there is an N e N such that xn £ / _ 1 ( ^ ) ifn > N. This, in turn, means f(xn) € V for n > N, which proves {f(xn)} is a convergent sequence. 4k The difficulty with convergence (or lack thereof) is that to prove convergence one needs to know the limit ahead of time. Under the right circumstances we can actually get around having to know the limit. First, we have to have a particular type of sequence and a certain type of space. Definition 6.9 Let (S, d) be a metric space. A sequence {xn} in S is called a Cauchy sequence iffor every e > 0, there exists an N £ N such that n,m> N implies &\Xnj
Xm)
100
fl
SELF-SIMILAR OBJECTS
EXAMPLE 6.10 Let xn = 1/n. Then {xn} is a Cauchy sequence in R with the Euclidean met ric, m
Proof: Let e > 0 be given. For this e, let N £ N such that N > 1/e. Then for n, m > N n—m 1 1 1 1 < —r-r T < — < £• n m nm mm{n, m\ N Since e was arbitrary, this holds for all e and thus Cauchy sequence. 4t fl
EXAMPLE6.il A sequence x = {xn} of real numbers is bounded if there exists a finite real number M such that \xn\ < M for all n. Let B denote the collection of all bounded sequences. On this set we put the distance d between sequences as
[2 *
ifx^y
where K represents the index of the first entry where x and y disagree. So if x = (1,1,1,2,1,1,...), y= (1,1,1,1,2,1,1,...) a « ^ = (1,1/2,1/3,...), then d*(x7 y) = 2~ 4 , d*(x, z) = 2^ 2 . If nn is the sequence which is zero in all entries except the nth where that entry is 1, then {K„} is a Cauchy sequence in (B,d*). * As we can see by these examples, the most obvious examples of Cauchy sequences all look like they should converge. The following theorem shows that this is trueconvergent sequences will be Cauchy sequences. This is not an "if and only if" statement though. Theorem 6.4 If {xn} is a convergent sequence in (S, d), then it is a Cauchy se quence. Proof: Suppose {xn} is a convergent sequence. Thus there exists a point x € S such that for any fixed e > 0 we can find a N g N such that if n > N, we have d(xn, x) < e/2. Then ifn, m > N, we have, via the Triangle Inequality, d{xn,xm)
< d(xn,x)
+ d(x,xm)
< e/2 + e/2 = e.
Since e is arbitrary this proves that the sequence is Cauchy. 4b As we said, the converse is not true. Below are examples ofnonconvergent Cauchy sequences. Thus having the property that all Cauchy sequences converge is a very special property for a metric space. So it gets a particular name. Definition 6.10 A metric space (S, d) is called complete if every Cauchy sequence {xn} converges to a point in S.
SEQUENCES IN (S, D)
S
101
EXAMPLE 6.12 Look at the sequence {1/n}. In the space S\ = K with the absolute value metric, this sequence converges to 0. In the set 5*2 = (0,1) with the same metric, the sequence does not converge. ■
S
EXAMPLE 6.13 In the metric space (Q, \-\) the sequence x\ = 2, xn+\ = ^f- + ■£- /or n > 2 is a nonconvergent Cauchy sequence. Thus Q is «of a complete metric space. ■
77je rea/ /ine WJY/J Euclidean metric is a complete metric space. In fact it is "the completion of the rational numbers." This phrase comes from one way of developing M. from Q. The reals can be shown to be the set of all possible limits of Cauchy sequences made up of points in Q. The space C[0,1] with the sup metric is also complete. Think about how this changes things: If we have a sequence of functions and we can prove they are Cauchy in this space, then we can conclude the sequence converges to a function f even if we don not have an explicit formula or graph for the limit function on hand. In order to start our discussion of self-similar objects and fractals, we need our sets to be compact. Without compactness we would not be able to define a metric that gives us the necessary convergence of our sequence of sets. In order to discuss compact sets we need the following: We are taking aspects of the real line (and R n ) and re-writing them in metric space terms. For these next definitions, let (5, d) be a metric space and A a subset ofS. However, the examples are in M. with the Euclidean metric. Definition 6.11 A point a is called a limit point of A if there is a sequence {xn} of points in A \ {a} such that xn —> a. fl
EXAMPLE 6.14 In (M, | ■ |) the point a = 0 is a limit point of the sets A\ = [0,1], A^ = (0,1), and A3 = {1/n}. The point 0 is the only limit point for A3. There are lots of other limit points for A\ and A2. Any point a with 0 < a < 1 has a sequence in A\ (or A2), not using a, that converges to a. ■
Limit points do not have to be members of the set. If we adjoin the limit points to the set, we get what we refer to as the closure. Definition 6.12 The closure of the set A, written as A, is the collection of the points of A and the limit points of A; that is, A = A U {x : x is a limit point of A}.
102
5
SELF-SIMILAR OBJECTS
EXAMPLE 6.15 In (R, | • |) the three sets above have A\ = [0,1], A2 = [0,1], and A3 =
{l/n}U{0}.
■
We have already defined closed sets, saying that closed is the complement of open. We can also define that in terms of limit points. A set A is closed if A contains all of its limit points. Definition 6.13 The set A is closed if A = A and is perfect if A is equal to the set of limit points (that is, every point in the set is a limit point of the set). fl
EXAMPLE 6.16 In (R, | ■ |) for the three sets above, A\ is perfect (hence closed) and neither A2 nor A3 is closed or perfect. The set A3 is closed, but still not perfect, no nonzero point is a limit point. There is no sequence in, for example, A3 \ {1/2} which converges to {1/2}. ■
6
EXAMPLE 6.17 The Cantor Set is a perfect set.
■
This is actually the historical origin of the Cantor Set. The set was a solution to the problem of finding a non-empty, perfect, nowhere dense set. Cantor was actually not the first person to solve this problem. Henry J. S. Smith, an English mathemati cian, first constructed such a set in 1875. However, his work was not well-known. Vito Volterra published his version of this work in 1881, but was still a graduate stu dent at the time, publishing in a somewhat obscure Italian journal. Finally Cantor's set was published in 1883. So sometimes this set is referred to as the Smith--VolterraCantor Set. Now we move on to sequences and compact sets. A beforehand, a sequence in a set S is a function with domain N and co-domain S. We will write a sequence as an ordered, countably infinite list {xn}'^L1 or just {x„}. Definition 6.14 Given a sequence {xn} in a space (S, d), a subsequence is an in finite sequence of elements taken from {xn}, in order. For the sequence {xn} the notation for a subsequence is {xnk} and the understanding that the elements are taken in order translates to the fact that n^ > k. S
EXAMPLE 6.18 The sequence {1/2,1/4,1/6,...} is a subsequence of{l/n}.
■
It is worth pointing out that in R where there is a notion of order, there is the notion of monotone subsequences (see Chapter 2 for the definition of monotone). The
SEQUENCES IN (S, D)
103
proof of the famous Bolzano-Weierstrass Theorem, which states that every bounded sequence in M. has a convergent subsequence, is based on the fact that every bounded sequence must have a monotone subsequence; this is where every sequence has a convergent subsequence. 8
EXAMPLE 6.19 The sequence {1, - 1 / 2 , 1 / 3 , - 1 / 4 , . . . , ( - l ) n + 1 / n , . . . } has {1,1/3,1/5,...} as a decreasing, hence monotone, subsequence. Yes, there is an "obvious" in creasing one, too. m
Definition 6.15 The set A is compact if for every sequence {xn} in A, there is a subsequence {xnk} which converges to a point in A. The notion of compactness is a topological concept. In a general topological space, the topology defines what makes an open set. We say that the collection {Ua} is an open cover of a set X if each Ua is an open set and X C \JUa. A finite subcover ofX is a finite collection {C/,}™=1 from the set ofUa where n
Xc{JUi. Now the topological definition of a compact set is that X is compact if and only if every open cover of X has a finite subcover. Since we are only dealing with metric spaces, we are using the equivalent idea of sequential compactness. Theorem 6.5 In a complete metric space (S, d) a set A is compact if and only if it is both closed and totally bounded. Proof: Suppose that A is a closed and totally bounded set in S. Let {a^} be a sequence of points in A. Since A is totally bounded, we can find a finite collection of closed balls of radius 1 such that A is contained in the union of those balls. The Pigeon-Hole Principle l implies that since there are more points in the sequence than balls in the collection, one of the closed balls must contain infinitely many points out of the sequence. Say that is the ball £?i and select a point XNX in B\. Since B\ n A is a subset of A, it too is totally bounded; thus we can cover it with a finite collection of closed balls of radius 1/2. Repeating our Pigeon-Hole argument, there exists one of these balls (call it Bi) containing infinitely many {xi} where i > N\. Choose a out £JV 2 of B2- Continuing in this fashion we construct a nested sequence B1 D B2 D B 3 D ■ ■ • , 'The Pigeon-Hole Principle states, in essence, that if you have N holes in which the pigeons may roost and more than N pigeons, there is at least one pigeon-hole that has two or more pigeons in it.
104
SELF-SIMILAR OBJECTS
where each Bn has radius 1/n. It is easy to see that the subsequence {XN .} is a Cauchy sequence in (S, d). Since A is closed, the sequence converges to a point in A. In fact, it converges to the point
X= H
B
n-
So we have that A is compact. In the other direction, assume A is compact. Let e > 0 be given. Suppose we cannot cover A using finitely many open balls of radius e. Then there is a sequence of points {xn} such that d(xn,xm) > e for i ^ j . Since A is compact, there must be a convergent subsequence {xn,}. But every convergent sequence is Cauchy, hence there is a pair of integers N\ and N2 such that d(xN1,XN2) < e, a contradiction.
Thus A is totally bounded.
4
One nice property of compactness is that under the right type of transformation, this property is preserved. Here, the right type of transformation is a continuous one. Theorem 6.6 The continuous image of a compact set is compact. Proof: Let f : (S, d) —> (T, p) be a continuous function and suppose A is a compact set in S. To show f(A) is compact, let {£/„} be an open cover of f(A). Then is an open cover of A. Since we know A is compact, there is a finite {fl{Un)} subcover. Without loss of generality, call it
{f-1(Ul)J-1(U2),f-1(Us),...J-1(Un)}. Our claim is {U\. U2, ■ ■ ■, Un} is a finite subcover for f(A). This is because y G f (A) means there exists x G f^1(A) with f(x) = y and x 6 f~1(Uj) for some j . Hence y £ U3 so f(A) C U]=1Uj. 4 Finally we come to what Barnsley in [5] calls "The Space Where Fractals Live." We will start with the space S where S is a non-empty closed subset o/R™. By H(S) we mean the collection of non-empty, compact subsets of S. Now that we have the space, we need to define the distance. We will adopt the ideas from [30]. Let Abe a set in our collection H(S). We define a (5-body of A as the set of points at most 8 far away from A; that is, As = {x e K" : \x - a\ < 6 for some a € A}. Given two sets A and B in rl{S), it is fairly obvious that for a large enough 5, B C A§. Our distance between two non-empty compact subsets will be the least distance so that we do get containment in both S-bodies.
SEQUENCES IN ( 5 , D)
105
Definition 6.16 Let H(S) denote the collection of non-empty, compact subsets of S C K n , where S is a non-empty closed set. We define the Hausdorff metric on H(S) by d(A, B) = inf{<5 : A C Bs and B C As}. That d(A, B) > 0 is obvious and, since there are two containments to take into consideration, d(A, B) = 0 if and only if A = B. The symmetry is obvious. We will leave proof of the Triangle Inequality as an exercise. Another way to frame this idea of distance is the following: Given a point b and a compact set A in the metric space (S, d) define D(b, A) as the minimum of all the dis tances between b and a where a £ A. Notationally, that is D(b, A) = min{D(&, a) : a G A}. Then if we have two compact sets A and B we say the distance from B to A is the largest of all the D(b, A); i.e. D(B,A)
=
max{iam{D(b,a)}}.
This idea, however, does not define a metric. It is another homework problem to show that it may not be true that D(B, A) = D(A, B). That is why we emphasize the words from and to in the language. Tofinallyget to the Hausdorff metric we have d{A, B) = max{D(^, B), D(B, A)}. As a consequence of this definition and A and B being compact sets, we have that there are points a £ A and b G B such that d{A,B)
=d(a,b),
where on the left the metric is the Hausdorff metric and on the right the metric is the original metric on S. As with the canonical construction of the Cantor Set, when we create these selfsimilar objects we have a process which leads to the final result, but the object in question is the result of infinitely many applications of the routine. This means that we have a sequence of iterations and must know that this sequence converges in the correct sense. In other words, we need ('H(S), d) to be a complete metric space. Theorem 6.7 The metric space (T-l(S), d) is a complete metric space. So if{An} a Cauchy sequence in (7i(S),d) there exists a set A £ "H(5) such that
is
A = lim An, n—»oo
where the limit is taken under the Hausdorff metric. Proof: To prove this, we must start with a metric space (S, p) yet work in the space (H(S), d). Suppose we have a sequence An which is Cauchy in (7i(S),d). Let A = {x : there is a sequence {XJ } with Xj G A, and Xj —> x}.
106
SELF-SIMILAR OBJECTS
(Note: Xj —> x in the metric p, not d.) We will show that the limit of the An is in fact A. Let e > 0 be arbitrary and fixed. Since Am is Cauchy in (%(S), d) there exists an N £ N such that for n,m > N we have d(Am, An) < e/2. Let n be a value greater than N. If x £ A, then there is a sequence {XJ} such that Xj £ Aj and for large k p(xk,x) < e/2. Now because d(Am, An) < e/2 when k > N there is a y £ A.m such that p(y, Xk) < e/2. By the Triangle Inequality, this means p(x, y) < p(x, xk) + p(xk, y) < e/2 + e/2 = e. Thus A C (Am)e. Now take y £ Am. Since {Am} is a Cauchy sequence we can choose integers j \ < J2 < J3 < ■ ■ ■ with ji = m and d(Ajk,Am) < £ for all m > j k . Now we define a sequence yj with each y.j £ Aj. For j < n, pick any y} £ Aj and let yn = y. Forjk < j < j k + : , choose y0 £ Aj where p(yjk,yJ) < ~. Then {yk} is Cauchy in (S, p) and converges to some x £ A. Then p(y, x) = limp(y, ijj) < e. So y £ Af and since y was arbitrary, Am C A£. Thus d(A7 Am) < e and {Am} converges to A. 4 6.4
Affine Transformations
While many of our notions will work in any metric space (S, d), in order to have illustrations of what is going on, we will mostly restrict ourselves to S — R or M2 with their respective Euclidean metrics along with TL(S) with the Hausdorff metric where S is a compact subset of X; however the proofs will tend to be for general metric spaces. We wish to apply a certain type of function to the compact subsets of S. This function is capable of doing three things (separately or in tandem) to a closed and bounded set. It can (I) shrink the object, (2) rotate the object, (3) translate the object. Such a function is called an affine transformation and is defined in full generality below. Definition 6.17 A function T : X —>■ Y is called an affine transformation iffor any s,t £ X and any c & [0,1] it is true that T(cx + (1 - c)y) = cT(x) + (1 -
c)T(y).
Theorem 6.8 An affine transformation of an interval [a, b] into M. must be of the form T(x) = mx + b, where m, b £ R. Proof: Suppose a < b (it is possible that a = b, but then the interval is trivial as is the result). For any x £ [a, b] we can rewrite it as x =
b— x x —a a+ b. b— a b— a
AFFINE TRANSFORMATIONS
107
Since x is between b and a, this is of the form ca + (1 — c)b where c G [0,1]. Using the fact that T is affine, we get h X x a T( T{x)A= ~ Tt T(a)\^+~ T(h\ T(b) = T(b)-T(a) x -\ b— a b—a b- a This shows that T is indeed linear. 4t
fl
bT(a)~aT(b) . b- a
EXAMPLE 6.20 The function T(x) = \x + 2 is an affine transformation on the set A = [0,1]. If we let T(A) — {T(x) : x e A}, we see T(A) = [2, 2.5], which is a shrunken version of the original interval that has been then shifted to the right 2 units. ■
Theorem 6.9 Every affine transformation from M.2 into M.2 has the form T(x1,x2)
= (ax i + bx2 +e,cxi,dx2
+ /),
where a, b, c, d, e, / are real numbers. For those who know matrix multiplication, this comes from the formula
Although transformations move points around in the space, there are points that do not move under T. These points are very important in the study of dynamical systems and will help us in creating fractals. Definition 6.18 Let T : S —>■ S be a transformation on a metric space. A point XQ E X is called a fixed point (or attractor or invariant) ifT(xo) = xo. fl
EXAMPLE 6.21 /. IfT\(x) = 3x + 1, then the fixed point is the solution to T(x) = 3x + 1 = x which is x — —1/2. 2. IfT2(xi,X2) = ( | x i —X2 + I, |^2)» then the fixed point is (3/2,0) (check this to see that it works). u
In order to create the self-similar objects we are heading for, we need a particular type of transformation, called a contraction, which ensures us that as we apply the transformation points are getting closer together. Definition 6.19 Let (S, d) be a metric space. A transformation T : S —)• S is called a contraction mapping if there is a constant r G [0,1) such that d(f(x)J(y))
108
fl
SELF-SIMILAR OBJECTS
EXAMPLE 6.22 The transformation T on (M, J • j) given by T{x) = x / 3 + 1 is a contraction. We see that \T(x) - T(y)\ = \\x - y\. ■
S
EXAMPLE 6.23 0 « M2 r/ie transformation T(x, y) = ( x / 4 + 2 / 3 , y/2) shrinks things by 1/4 in the x-direction and 1/2 in the y-direction, and then it translates the shrunken im age by 2/3 of a unit to the right. This is a contraction with d(T(x\, yi),T(x2,2/2)) < ■ ±d{{x-l,y1),(x2,y2))-
Theorem 6.10 IfT is a contraction on (S, d), then T is a continuous
function.
Proof: Given e > 0, we must show the corresponding S > 0 such that d(x, y) < 6 implies d(f(x), f{y)) < e. So let us start with e > 0 and suppose T has contraction factor r > 0. If we let 6 = ^rh-, then for d(x. y) < S d(T(x),T(y))
< rd(x.y)
< rS = r - ^ - < e r + 1 and we are done. (This is really a proof that T is uniformly continuous, but we shall not go into that distinction here.) 4 Contractions give a special quality to their invariants. While we cannot guarantee the existence of a fixed point (it depends on the space), if there is an invariant, we know there cannot be more than one. Theorem 6.11 Let T be a contraction on (S, d) with contraction factor r. Then the fixed point for T (if one exists) must be unique. Proof: Suppose not. Say T has two distinct fixed points x and y. Then we have the following: d(x,y) = d(T{x),T(y))
rW(s) = (ToTo..-or)(s). v
v
n copies ofT
'
Theorem 6.12 The Contraction Mapping Theorem Let (S, d) be a complete met ric space and let T : S —> S be a contraction with factor r. Then for any point s £ S, the sequence {T^ (,s)} converges to a point s and s is the invariant for T.
AFFINE TRANSFORMATIONS
109
Proof: Let s be a point in S. Then the triangle inequality gives us d(s, T^is))
< d{s, T(s)) + d(T{s), r ( J ' ( s ) ) < d(s, T(s)) + rd(s, T(s)).
In fact, for any positive integer n d{s,T^{s))
<
d(s,T(s))+d(T(s),r( 2 )(s)) + --- + d(T("- 1 )(s),T(")(s))
<
d(s, T(s)) + rd{s, T(s)) + ■■■ + rn~ld{s, 2
=
(1 + r + r + --- +
r ~ )d(s,T(s))
=
l=rld(s,T(s))
jh-A^ns))-
<
T(s))
n 1
In a similar vein, for positive integers m and n d{T^\s),T^m\s))
<
r^{m,n}d(s,T{s))
This shows that {T<~n\s)} is a Cauchy sequence. Since (S, d) is complete, it has to converge to some point which we will call s. Due to the fact that T is a continuous function, we can now see T(s) = T( lim T^n\s)) n—►oo
= lim T^n+1\s)
= s.
n—>oo
Thus s is the invariant. A Now we are ready to move on to the space (H(S), p) and apply our ideas. Theorem 6.13 Let T : S —> S be a contraction on the metric space (S, d) with contraction factor s. Then T is also a contraction on % with metric p (where T(A) — {T{a) : a £ A}) with contraction factor s. Proof: From Theorem 6.10 we have that T is continuous on S and Theorem 6.6 tells us the image of a compact set is compact, hence T maps % onto itself. The metric p looks for the largest distance between any two points in the compact sets U and V. Since the sets are compact, these are actual members of the set; that is, there exists u G U and v E V such that if 8 = d(u, v) we have U C VsandV C Us. Take A, B € H(S). Then there exist a e A and b € B so that d(T(A),T(B)) = d(T(a),T(b)) < sd(a,b) = sd(A,B). This means via the definition of the Hausdorff metric T(A) C T(B)S* where 5* = s ■ d{A, B). Similarly, T(B) C T(A)S. where S* = s ■ d(B, A). Thus p(T(A),T(B))<s-p(A,B), which is what we needed to prove. ♦
110
SELF-SIMILAR OBJECTS
By an Iterated Function System (IFS), we mean a finite collection {Ti}™=1 of contractions, each with corresponding contraction ratio Ti. With this collection we create the transformation
r(A) = T1(A)uT2(A)u-urn(4 where A is a non-empty, compact set. This is itself a contraction with contraction ra tio r = maxjri, r<2, ■ ■ ■, rn}. The invariant for T is guaranteed to exist by Theorem 6.13 and will be self-similar.
Let us give some examples of these systems and their invariants. 9
EXAMPLE 6.24 On the interval [0,1], ifJ\(x) — x/2> and T2{x) = x/3 + 2/3, then the IFS {Tj, T2} has the Cantor Set as its invariant. Note: We use [0,1] here and not the whole line because in order to get uniqueness of the invariant, we really need to work in the Hausdorff Space of non-empty, compact sets with the Hausdorff metric. If we allowed noncompact sets, then R = 7\(R) U T2(M) and we lose uniqueness of the fixed point. ■
9
EXAMPLE 6.25 In the unit square, let our three contractions be
Tx(x,y) = (x/2,y/2), T2{x,y) = {x/2 + l/2,y/2), T3(x,y) = (x/2+l/2,y/2 + l/2).
So a compact set (such as a triangle inside the unit square) is shrunk by a factor of 1/2 in both directions, one time shrunk and then translated a 1/2-unit to the right, and another time shrunk and then translated a 1/2-unit to the right and a 1/2-unit up. The resulting invariant is called the Sierpinksi Triangle.
AFFINE TRANSFORMATIONS
H
*
*
■
»
*
•
■
*
•
»
■
■
*
'
,m-
*
"
N »
»
* " " "* -
C\. tx Cx ; .
tx N
\
'
X
■
*
H
»
l
*
i
l
"
111
*
IF
i
Several Iterations of T = T\ U T2 U T3 /f is wort/i noting here that the end result being a triangle has nothing to do with the initial image being a triangle. The idea of invariant is that the end result will be the Sierpinski Triangle regardless of what shape we begin with. The only difference would be the speed with which the iterates converge.
S
EXAMPLE 6.26 Again we start in the unit square. Our starting image will be the unit inter val which we will turn into a curve. The Koch Curve (named for the Swedish mathematician Helge von Koch) first appeared in a 1904 paper entitled "On a Continuous Curve without Tangents, Constructed from Elementary Geometry." This example, unlike the others, uses rotation of the object. There are four transformations here. The first is shrinking by 1/3 in both directions. The second also shrinks by 1/3 in the x- and y-directions, but then translates over by 2/3 horizontally. The next shrinks by 1/3, then rotates by 60°, then translates 1/3 to the right. The final transformation is shrink by 1/3, rotate —60°, and then translate 2/3 to the right and up by ^ . For those who know matrix multiplication, a rotation by the angle 9 is given by
112
SELF-SIMILAR OBJECTS
Our four transformations are then
1/3
Ti{x,y)
'l/3
T'2(x,y)
0\
x\
/2/3N
0 \ fx\
, 0 1/3 i I J =
J
/(l/3)cos(60°) ^(l/3)sin(60°)
I 0, (-l/3)sin(60°)\M (l/3)cos(60°)J ^
/l/3\ ^ 0 ) '
(l/3)sin(60°)\ AA (l/3)cos(60°W \y
/ 1/2 N 1^3/6,
'Ll
/ (l/3)cos(60°) l-(l/3)sin(60°)
T^y)
=
T2(x,y)
=
(a:/3 + 2/3,y/3),
T 3 (x,y) T4(xV£/)
= =
( i / 6 - N / 3 ! / / 2 + l/3 ) a;v/3/6 + j//6), (x/6 + >/3!//2 + 1 / 2 , x V 3 / 6 - y / Q + >/3/6).
r
=
or
(x/3,y/3),
One interesting fact about this curve is that for at each step, k, the length of the curve is (4/3) k . Thus the final product has infinite length; but, since it fits inside the unit square, it covers finite area.
AFFINE TRANSFORMATIONS
113
„.r^.f\.
ft*
i]£
D.S
9.7
19
ftt
The first two iterations and a representation of the final curve. u When we defined the Cantor Ternary Set, we discussed one way to tweak it. This brought us to a generalization we wrote as Cr, 0 < r < 1. These sets are also the attractorfor an IFS, using the contractions T\{x) = rx andT2(x) — rx + (1 — r). This brings up the question, "Is every Cantor Set the attractor for an IFS?" The answer is, "No," but requires a different idea of metric. This comes from [15] where a Borel Probability Measure is used and the following theorem applied. Theorem 6.14 There exists a Cantor Set X% and a Borel probability measure p supported on X\ such that for every contractive map f : X\ —> X\, we have M/(A-i)) = 0. The set X\ has the property that for any iterated function system {fi, f2, ■ ■ ■, fn} the set Ufi (X\) has measure zero while X\ has full measure. Then X\ cannot be the attractor of an IFS. More ways to make Cantor Sets are covered in Chapter 10.
114
6.5
SELF-SIMILAR OBJECTS
An Application for an IFS
As an application of iterated function systems, let us consider the following. A theo rem from Calculus I states that if a function is differentiable at XQ, then it is contin uous at XQ, too. After the proof (an application of the limit definition of derivative), the next thing to show is an example to show the converse is not true. Continuity does not imply differentiability and the function f(x) = |a;| at XQ = 0 illustrates this. The question we wish to answer here is, "How much can we disconnect continuity and differentiability? " Since we can do this for one point, can it be done for two? Five? Countably infinitely many points? The answer is that this can be done for the entire line. There are functions which are continuous on all of R, yet differentiable at no point. The first example of a function with this property is credited to Weierstrass. His result, published in 1872, is defined by oo
/(X) = J > " C 0 S ( 6 " 7 T X ) , n=0
where 0 < a < 1, b is a positive integer, and ab > 1 + =y. Pictures of the graph are available nowadays thanks to computers. It is difficult, though, to see how the graph develops the properties we need. We will produce an example due to Katsuura [42] in which we can see the function developing because of the self-similar nature of the graph and the iterative process from which it is derived. We will work in the space of the closed unit square ([0, 1] x [0,1]). Our three contraction mappings are T^y)
=
T2(x,y)
=
T3{x,y)
=
(x/3,2y/3), (2-x/3,l+y/3), (2 + x/3,l
+
2y/3).
The first transformation takes an object and shrinks it by 1/3 in the x-direction and 2/3 in the y-direction. The second shrinks the image by 1/2 in each direction, trans lates it 2/3 to the right and up 1/2, then rotates it about the line y = 1/2. The last shrinks the image with the same ratios as the first, but then translates the image 2/3 to the right and 2/3 up. If we use the graph of y = x as the initial image (fo) then each successive application of T(A)=T1(A)UT2(A)UT3(A) will also yield the graph of a function. first three images are below.
We name them fn(x)
= Tn(fo(x)),
and the
AN APPLICATION FOR AN IFS
115
The initial graph and first two iterations Due to the least contraction being by 2/3, we see that for any k and any value x e [0,1], \fk{x) - /fe+i(x)| < (2/3)fe. This means for n < m sup{|/ m (x) - fn(x)\
:xe[0,1]}
< (2/3)11,
which shows that {fn} is a Cauchy sequence in (C[0, l],p). This is a complete metric space, hence the sequence must converge to some continuous function f. We still need to show that this function f does not have a derivative at any point. Rather than repeat all the technical details, we will say that the interested reader can find them in Katsuura's paper and we present the heuristic proof. The interval [0,1] can be split into two types of points: The first are points like x = 1/3, or x — 4/9; rational numbers in base 3. These points are where the fn have cusps (sharp corners). Notice that this never changes. Once a cusp appears at x, fn(x) will continue to have a cusp there. Thus f is not differentiable at that point. The other type of point, where there is not a cusp, are points like x = 1/2. At these locations the graph alternates direction. For / Q ( 1 / 2 ) the slope is positive 1. For / i ( l / 2 ) the slope is — 1. This alternating keeps going forever, meaning that there cannot be a limit to the sequence of slope values, which is the same as saying /'(1/2) does not exist. Thus f is a continuous function that is nowhere differentiable.
116
SELF-SIMILAR OBJECTS
EXERCISES 6.1 For the sequences xn = ( — 1)™ and yn — \jn give an example of monotone subsequences. 6.2 Prove that the function d*(x,y) = \x\ — x%\ + \y\ — 3/21 is a metric on the plane ]R2. 6.3
Show that in the trivial metric, d, every sequence is a convergent sequence.
6.4 A unit circle centered at the origin is the set of all points whose distance from (0,0) is equal to I. What this circle looks like depend on the metric. Draw two unit circles with center (0, 0) in the plane: one using the usual Euclidean metric and the other using d* from the exercise above. 6.5
Let S = C[0,1]. Given two functions f and g, define d(f,g)=
I Jo
\f(x)-g(x)\dx.
Show that d is a metric on S. Find an example to show this is not a metric if S is the set of Riemann integrable functions on [0,1]. 6.6 Show f(x) = rax + b is a similarity on the real line. If \m\ < 1, find the location of the fixed point. 6.7 Show that if'T is not a contraction, then the set of fixed points can have more than one element. 6.8 Although we could work with any non-empty, compact interval, we will stick with [0,1]. a) Let fi(x) = x/2 and f2{x) = 3a;/4 + 1/4. Show that both of these are similarities and
[o,i] = / 1 ([o,i])u/ 2 ([o,i]). b) Now come up with two different similarities,
Prove the Triangle Inequality for the Hausdorff metric. Show that for a finite set A, HS(A) = 0.
6.11 The Hausdorff metric needs the sets in the space to be compact and non empty. a) Find an example of two sets A and B so that if we allowed non-closed sets, we could have D(A, B) = 0 with A / B. (Hint: Keep things simple, stay in R.) b) Explain why d({0}, 0) = 00.
EXERCISES
117
6.12 Let p be a fixed prime number. Consider the following on the set of integers, Z. For i £ Z the value oford(x) is the largest natural number k such thatpk divides x. Define a function p : Z x Z —s- [0, oo) by
S P(X,V)
s _ J0' ~ \p-or^-y),
x = y, x^y.
Show that p is a metric on Z. This is an example of a metric that takes a countably infinite number of values. 6.13 Let A, B s M. A function f : A —> B is called uniformly continuous if for every e > 0 ?/?ere existe a (5 > 0 .SMC/Z that for all x,y £ A // |x — y| < 5, f/zen | / ( x ) — / ( y ) | < £■ 77HS, of course, does generalize to metric spaces A and B. a) Prove that f(x) — y/x is uniformly continuous when the domain is [1, oo). b) Prove that g{x) = x2 is not uniformly continuous on (—oo, oo). c) Prove that any linear function is uniformly continuous.
CHAPTER 7
VARIOUS NOTIONS OF DIMENSION
In Chapter 3 we looked at indicators of size (measure, category, cardinality). How ever, these did not show the differences in Cantor Sets of various ratios r. That is, for any fixed r, Cr is measure zero, nowhere dense, and has cardinality 2N°. So how is C1/3 different from C1/4 or C2/5? For that we turn to the notion of dimension.
7.1
Limit Supremum and Limit Infimum
Before we go into dimension, we must extend what we know about limits. Sometimes, such as for the sequence {(—1)"}, there is not a limit; however, we can look for a limit of a part of the sequence. What we mean by that is explained below. Definition 7.1 Given a set A of real numbers, the number u is an upper bound of A ifa< ufor all a £ A. Similarly, the number v is a lower bound of A if v < a for all a e A. S
EXAMPLE 7.1
The Elements of Cantor Sets -With Applications, First Edition. By Robert W. Vallin Copyright © 2013 John Wiley & Sons, Inc.
119
120
VARIOUS NOTIONS OF DIMENSION
LetA
= { 1 , 2 , 3 , 4 , 5 } , B = [0,1), and C = {x : x eQandx2
< 2}. Then
1. The upper bounds for A include 5, 7, and \/30, while some lower bounds are 0, —3, and —n. 2. For the set B, 0 is a lower bound and a member of the set and 1 is an upper bound, but not in B. 3. C has upper and lower bounds, but no such bound is a member of C.
u As we can see, upper/lower bounds are not unique nor do they have to exist. If a set has a lower bound, the set is called bounded below; and if it has an upper bound, the set is bounded above. Sets that are both bounded above and bounded below are just referred to as bounded. Sets without an upper bound or lower bound (or both) are unbounded. Definition 7.2 Let A c R. The number U is the supremum (also known as least upper bound) of A ifU is an upper bound of A, and for any other upper bound x we have U <x. If no such number exists, then we say sup(yl) = oo. Similarly, V is an infimum (or greatest lower bound) of A ifV is a lower bound of A, and for any other lower bound x we have
x
fl
EXAMPLE 7.2 For the sets A, B, and C in Example 7.I, sup( A) = 5, inf (A) = 1, sup(i?) = 1, inf (B) = 0, sup(C) = \ / 2 , and C has no infimum. ■
The reader may be familiar with maximums and minimums. Suprema and infima are different in that the max/min of a set must be a member of the set while sup/inf need not. The set B above does not have a maximum, but does have a supremum. Now we wish to combine this idea of sup/inf with the well-known concept of limit. Here we will stick with limits of real numbers. In spite of having already generalized limits to metric spaces (S, d), sup and inf need the space to have a total ordering to it. What we mean by that is for any two elements x and y, there needs to be a way to determine if' x < y, x > y, or x = y. A space having a metric attached to it does not say anything about order existing. Even in such a usual space as IR2 we usually do not refer to order, there is no typical rule to determine how the points (1, -\/3) and (0, 2) compare.
LIMIT SUPREMUM AND LIMIT INFIMUM
121
Suppose we have a sequence of real numbers, {xn}^=1. Associated with this se quence is the set of values of the sequence. We can then look at the sup and infofthe set. To avoid extra notation we will not distinguish between the sequence and the set of values of the sequence and so we will write this as sup n > 1 {x n } and'm£n>i{xn}. For example, the sequence tn = 1/n has sup n > 1 {£ n } = 1 and mfn>i{tn} = 0. = 1 and mln>i{yn~} = - 1 . Note With the sequence yn = (-1)™, supn>1{yn} that in these examples tn has an inf which is not a member of the set of sequence values and yn is an infinite sequence with a finite set of values. Lastly, we point out that for tn having a different starting index affects the value of the supremum; that is sup n > 1 {£„} = 1, sup„> 2 {£„} = 1/2, sup n > 3 {t n } = 1/3, etc. This brings us to the notion of limit supremum and limit infimum. Definition 7.3 Let {xn} be a sequence of real numbers. We definite the limit supre mum and limit infimum of the sequence as follows: lim sup x n = lim (supx^) and lim inf xn = lim (inf £&).
fl
EXAMPLE 7.3 1. Let xn = (—1)™. Then l i m s u p , ^ ^ xn = 1, while lim.infn_j.oo xn = —1.
{
n,
n is even ,
sin(l/n),
n is odd .
Then lim sup y = oo while lim inf y = 0. n
n
3. Let rn be an ordering of the rational numbers in [0, 2]. Then lim s u p ^ ^ ^ rn — 2, while lim inf^^oo rn = 0.
■
As the next two theorems show, our intuition is correct. The first shows that a lim sup is at least as big as a lim inf, and the second that if a sequence converges, then its lim sup and lim inf will converge to the same value. Theorem 7.1 Let {xn} be a sequence of real numbers. Then limsupa; n > lim inf xn. n->oo
n—>oo
Proof: If either lim sup xn = oo or lim inf xn = — oo, there is nothing to prove, so let us assume that there are real numbers a and b such that lim inf xn = a and
122
VARIOUS NOTIONS OF DIMENSION
lim sup xn = b. Consider a < a and j3 > b. By the definition of supremum and infimum, there exists an N G N so that for n > N we have a < xn < j5. But a < xn for infinitely many n means a < lim sup xn = b. Since a is arbitrary, letting a approach a gives us a < b which is our desired result, jt. Theorem 7.2 Let {xn}
be a sequence of real numbers and let c £ t . Then
lim xn = c if and only if lim sup xn = lim inf xn = c. Proof: Let e > 0. If lim sup xn = c, then there exists an N\ G N such that n > N\ implies xn < c+e. Similarly, there is an N2 G N s o thatn > N2 implies c — e < xn. Putting these together gives us, for n > max{7Vj. A^2 }, —e < xn — c < e, which is the same as \xn - c\ < e. Thus linin-^oo xn = c. In the other direction, assume lim„_>.00 xn = c. Then given e > 0 there exists N G N such that n > N implies \xn
-c\<e.
Thus for n > N we obtain c — e < lim inf xn < lim sup xn < c + e. But e is arbitrary. Letting it approach zero gives lim inf xn = lim sup xn = c. 6 This notion of'lim sup and lim inf extends to functions, the same way the ordinary idea of limit does. We will assume that our function has domain M. although we could be more general and have domain E C U . For more information, see [11]. Definition 7.4 Let f : M. —> R. and suppose XQ is fixed. Then lim sup f{x)
= inf { s u p { / ( x ) : x G {XQ — 5, x0 + S),x ^
xo}}
lim inf / ( x ) = supjinf{/(re) : x G (XQ — S, XQ + 5), x ^
xQ}}.
and
Continuing in a parallel fashion to what we had with sequences, we start with an example and two theorems (without proof this time) showing the same intuitive results.
TOPOLOGICAL DIMENSION
S
123
EXAMPLE 7.4 Let f(x) -1.
= sin(l/x). Then limsupa._>o f(x)
=
1> while liminf^^o
f{x)
Theorem 7.3 Let f : R —>• K and /ef £0 € M be fixed. Then lim sup/(x) > liminf f{x). Theorem 7.4 Let f : R —> R, /ef XQ € K be fixed, and let c 6 R. 77ie« lim /(#) = c if and only if lim sup /(x) = liminf/(x) = c.
Wfe wi// 500« utilize these to come up with various, non-integer notions of dimen sion. Let us begin with an easier idea, topological dimension, whose possible values are —1, 0,1, 2,....
7.2 Topological Dimension As we have seen earlier, topology concerns itself with the study of open and closed sets. We repeat the definition here. Given a space S, a topology is T is a collection of sets which satisfy certain rules:
1. %Se T. 2. IfUa(a&
A, an index set) is in T for all a £ A, then
\JuaeT.
aeA
3. IfU,V eT.thenUHV
e T.
The sets in T are referred to as the open sets. If A is an open set, then its complement S \ A is closed. A topology is generated by something called a basis. These are the sets in T which are apart of every open set. More specifically, we have the definition: Definition 7.5 Let T be a collection of open sets and let B be a family of subsets from T. We say that B is a basis for the topology iffor every open set A G T and every x G A, there is a set U G B such that x G U C A.
124
fl
VARIOUS NOTIONS OF DIMENSION
EXAMPLE 7.5 For M. with the Euclidean metric, the family B = {(a,b) : a,b G K, a < b} forms a basis. This is not the only one. The collection B* = {(a, b) : a, b G , a < b} is also a basis. ■
It is this idea (unstated) that we used when describing the real line and Sorgenfrey line in Example 3.7. Topological dimension is so named because it is based on the topology of the space and does not need the notion of distance (metric). There are three types of topological dimension: small inductive dimension, large inductive dimension, and Lebesgue covering dimension. In a normal space with a countable base (and we are staying in K™ with the Euclidean metric, which meets these criteria) the three notions have the same value. We first define small inductive dimension . As the name implies, it is defined inductively. For these sets we use the idea of boundary. Recall for a set A, the boundary of A is the set of points in the closure of A, but not in the interior of A. We write that as DA = A\A°. Definition 7.6 Begin with 0. Its inductive dimension is set to be — 1 (the only set with negative dimension). For a set A, the inductive dimension of A is the smallest value n such that for every x G A and every open set U with x G U we can find an open set V where x G V and V CU and the inductive dimension ofdV is at most n — 1. We will write this as
dimtop(A).
In our K", this reduces to V being an n-dimensional ball centered at x and the boundary ofV will have dimension n — 1. Thus a countable set inductive dimension 0, a line has inductive dimension 1, a plane 2, and so on. For a Cantor Set, given any point x and open U containing x there is an open interval I inside ofU so that x G / and I C U. The boundary of the interval has dimension 0 (just being singletons), thus dim top (C) = 0. Another way to define this is with separating components. To begin with, we say a topological space S is disconnected if there exists two non-empty, disjoint, open sets U and V so that S = UUV.Ifa space is not disconnected, then it is connected. A disconnected space consists of pieces called components. By the definition each component is a closed set. This seems a little mysterious, so before we go on to some examples, let us talk about inheriting a topology. Let S be a topological space with topology T and ACS. Then A inherits its own topology by TA = {A n T : T G T } . So as a subset ofM. with the Euclidean
TOPOLOGICAL DIMENSION
125
topology, the interval [0,1] is a closed set. However, in the space Y = [0,1] U [2, 3] with the inherited topology, [0,1] being equal to [0,1] (~l (—1, 2) is an open set in Y(!!). This tells us how each component can be open. The complement of the compo nent must (by being contained in one of U or V) must be an open set. The idea of connected and disconnected can also be inherited. For the set A to be disconnected, look atl/nA and V n A and see how A = (UnA)U(Un V).. fl
EXAMPLE 7.6 1. The interval [0,1] is a connected set in K with the Euclidean metric. 2. The interval [0,1] is a disconnected set in M with the discrete topology. Since each singleton is an open set we can just partition the interval into the open sets {0} and (0,1]. We can do even more and divvy this into uncountably many components. 3. The set A = [0,1] U [2, 3] is a disconnected set in K with the Euclidean metric that consists of two components. 4. The Cantor set in K and Euclidean is a disconnected set also with uncountably many components. In fact, it is what we call totally disconnected because the components are just single points. u
So the topological dimension of a set A can also be described this way: The dimension of the empty set is — 1. The dimension of any other set is one more that the dimension of the object with the smallest inductive dimension that can be used to separate the set into components. The condition of minimality is important. A finite set can also be separated by a single point that is not in the set. Let us redo some examples with the new point of view: If we have a finite set all the points are separate components and it takes nothing to separate the set. Hence the inductive dimension is — 1 + 1 = 0. Similarly, this argument shows a countable set has topological dimension zero. If we have an interval [a, b] we need a single point to separate it. Thus the interval has dimension 0 4- 1 = 1. A line is divided using a single point, which has dimension 0, so the line has topological dimension 1. A rectangle in the plane needs a line to separate it, so the topological dimension of a rectangle is 2. The Cantor Set is totally disconnected. Hence it need no points to separate the set into components. This means dim t o p (C) = 0. Topological dimension does have some nice properties. As the proofs bring us a little too far into topology, we will omit them. They can be found in [26]. Theorem 7.5 Let AC. B. Then dim top (A) < dimtop(_B). Theorem 7.6 If An are closed and d\mtop(An) d\mtop(UnAn)
= Ofor all n € N, then = 0.
126
VARIOUS NOTIONS OF DIMENSION
Recall that a homeomorphism between two sets is a bijection f where both f and f~l are continuous functions. The topic of topology concerns itself with properties that are conserved when a homeomorphism is applied. Theorem 7.7 Topological dimension is preserved by homeomorphism. are homeomorphic, then dim t o p (yl) = dim t o p (i?).
If A and B
Thus all types of Cantor Sets Cr have the same topological dimension. This is not helpful as we are looking for ways to distinguish Crfrom Cs. Now that we have the concept of homeomorphism and see how objects can be topologically equivalent, let us close this section with one of the more fascinating properties of Cantor Sets, the equivalence of compact, perfect, totally disconnected sets. The next theorem shows that all of these are the same as the Cantor Set. Theorem 7.8 Let Abe a compact, perfect, totally disconnected set. Then A is home omorphic to the Cantor Set C. Proof: In this proof, we do not create an explicit homeomorphism, F, but instead describe the construction of such a function. The set A is compact, hence bounded above and below. Let m be the greatest lower bound of A while M is the least upper bound of the set. Our construction, F, will be a function with domain [ra, M] and range [0,1] with the additional property that F(A) = C. Since A is totally disconnected, it cannot contain an interval, hence it is a nowhere dense set. This means the set [0,1] \ A is the union of countably infinitely many disjoint open intervals with the additional property that for any two intervals their closures are disjoint. Why? The intervals must be countable since each one contains a rational number. There cannot be finitely many since A is nowhere dense. Because A is a perfect set it cannot have any isolated points, hence the closure property. This collection of intervals will by denoted by I. The collection X is a set of bounded intervals. Take an interval I\ of maximal length (do you see why at least one must exist?). We match this up with the Cantor interval [1/3, 2/3] as an increasing line segment. Now let In and I\i represent the longest intervals to the left and right of 1\, respectively. Map these to the Cantor intervals [1/9, 2/9] and [7/9, 8/9] in an increasing linear manner. This process can continue with { / m , /112, I121 > ^122} being paired with {[1/27, 2/27], [7/27,8/27], [19/27, 20/27], [25/27, 26/27]} via increasing lines, et cetera. So now we have a strictly monotone function F : [m, M] \ A —> [0,1] \C which is clearly one-to-one and onto. To finish this, we must define F at each a € A. Define the set La as {x £ [m, M] \ A : x < a and x is an endpoint of an interval in I}. These points are to the left of a at which F has already been defined. We can then determine the value for F(a) by F(a) = sup{F(x)
:x e
LA}.
SIMILARITY DIMENSION
127
Thus we have a monotone bijection between [TO, M] and [0,1] and F(A) = C. Now we must show the continuity of F and F~l. First, note that since F is monotone, so is F~l. Second, F_1 is a bijection since F is one. So we will show that F must be continuous, the proof for F^1 being the same. Suppose (a, b) is an open interval in [0,1]. Then F" 1 ((a, b))
= = =
{x e [m, M] : F(x) £ (a, b)} {x e [TO, M] : a < F{x) < b} {xe[m.M}:F-l(a)<x
=
(F-\a),F-\b)),
which is open. This means the inverse image of open sets are open, hence F is continuous. 6 The last part of the proof above is that our function F and its inverse F^1 are continuous. This is a specific example of a much more powerful result: Iff : A —> B is a monotone bijection, then both f and f~l are continuous functions.
7.3
Similarity Dimension
We mention similarity dimension as it is easy to calculate and the equation involved will reappear later, but we shall see that similarity dimension does have a problem with it. As we noted in Chapter 6, the Cantor Set is an example of a self-similar object. It is the limit ofthe iteratedfunction systemTi(x) = ^xandT2(x) = j ^ + f Given an iterated function system with similarity ratios {r\, T-I, ..., rn} the simi larity dimension for the invariant K is the number s such that r{ + rs2 + ■ ■ ■ + rsn = I. So for our Cantor Set we have
,+
G) G)'which leads us to dim S j m (C) — |^-|. The sets to which we can apply this idea are somewhat limited. For a set A to be the invariant for an iterated function system, there must be a set of similarities {/$} such that A = Utf(A). This is not possible for some reasonably simple geometric shapes. This cannot be done for a unit circle. A second issue has to do with overlap. The unit interval I — [0,1] is self-similar and the invariant for transformations U\(x) = \x and U2 = \x + \ leads us to
128
VARIOUS NOTIONS OF DIMENSION
d i m s i m ( J ) = 1, which seems correct. However, it is also true that I = where V\(x) = ^x and V ^ x ) = | x + | . Since
Vi(I)UV2(I)
does not have s = lfor a solution, which counts as the "real" similarity dimension ? {UUU2} With the IFS {VUV2} we can see that VX{I) n V2(I) = [|, §] while has Ui(I) (~l {72 (-0 — {5}- So perhaps what we need is a condition that says the following for the invariant F: Ti(F) Pi Tj{F) must be small in some way. This will be formalized later with the Open Set Condition (Section 7.5).
7.4
Box-Counting Dimension
This type of dimension is popular since it is relatively easy to calculate. It has gone by different names over the years including capacity dimension, logarithmic dimen sion, and Kolmogorov dimension. At each stage n in the construction of the Cantor Set there are 2™ closed subintervals making up Kn. We can cover Kn using that many closed subintervals of length (1/3)™. Or we could use 4™ closed subintervals of length (1/6)™, or many oth ers. Box counting dimension looks at the relationship between the number of boxes used to cover a set and the diameter of the boxes involved (notice, unlike topological dimension, this requires us to have a metric in place). Let us look at this idea in generality. Definition 7.7 Let A be a non-empty, compact subset ofM.n. Let S be a positive number and let N$(A) be the smallest number of sets of diameter 5 needed to cover A. We then define the upper and lower box-counting dimension of A by dimbox(A)
=limsup S^o+
-ln(d)
and dimbox(A)=limmf <5-+o+
- \n(S)
When these two values agree, we have the box-counting dimension of A and write it as dimbox(A)
fl
= hm d~>o+
HNS(A)) -ln(<J)
.
EXAMPLE 7.7 We will not prove this; but as we would expect, diraf,ox({a}) = 0. In the line, dira(,ox([a,b]) = 1 where a < b; and in W', if B is the open ball of radius r > 0, d\m.box{B) = n (in fact, if A is an open set in W1, then dimb oa; (A) = n ■ since A must contain an open ball).
BOX-COUNTING DIMENSION
fl
129
EXAMPLE 7.8 For our Cantor Set, using the fact that at each stage there are 2n boxes (closed intervals since we are in the line) of diameter (1/3)™, it would seem we are dealing with ln(2") ,. nln2 In 2 lim ——r^-^ = -— n = lim n->oo — ln((l/3) ) n^oo n In 3 m3 (and this is the case); however, there is more work to be done before we can switch from the limits with functions to limits of sequences. ■
It is easy for us to see that it makes no difference whether the boxes we use on covering the Cantor Set are open or closed. This actually works in the definition, too. Open or closed does not change the dimensions. Are there other concessions to make ? Yes, many. To keep our arguments simple, let us stay in M.2 although all of this holds in M.n. Given 6 > 0 we can make a grid of side length 6 by considering all the squares of the form [niS, (ni +1)6} x [ri2<5, (n^ + 1)6} where n\ andn-i are integers. These squares have diameter 6\/2 and sides parallel to the axes. Any square in the grid with diameter 6 has side length 6/\/2. This means that an arbitrary set with diameter 6 needs at most 9 elements the grid to contain it (consider three rows of three boxes with diameter 6/y/2). So if N 1(A) is the number of grid boxes needed to cover A, we have N$(A) < 9Ng(A). On the other hand, a square of diameter 6 gives 5NS(A) < N%(A). Together H5NS(A)) -ln((J)
"
ln(JV|(A)) -ln(<5) "
\n(9N6(A)) -ln(6) '
and using ln(xy) = \n(x) + ln(y) and \n(6) —> —oo as 6 —> 0 + we have dinibox(A) A\mhox(A) are the same using open sets or grids. There are other equivalences (which we shall not prove) summarized below [30]. Definition 7.8 The upper and lower box-counting dimension for a set A C M.n are given by -PHNs(A)) dimbox(A) = limsup and dJmtoai(>l) = h m i n f — j ^ - , where Ng(A) is any of the following: • the smallest number of sets of diameter 6 that cover A; ■ the smallest number of closed balls of diameter 5 that cover A; ■ the smallest number of grid cubes of side length 6 that cover A; ■ the largest number of disjoint balls of radius 6 with centers in A.
130
VARIOUS NOTIONS OF DIMENSION
Lastly, we note here that our 5 approaching zero can be done through a specific sequence rather than all possible ways to go to zero. Specifically, for some c £ (0,1) we can let 5n = cn and use "linin^oo ." This brings us back to our Cantor Set example and the fact that dinifcox(C) = |^|. Let us show that there is a need to have both upper and lower Box-Counting dimensions. Sometimes the limsup and liminf do not agree. To find this example, we go back to the Cantor Set construction. The same as before, at each stage we have 2k closed intervals. However, this time we do not always have the same ratio at each stage. fl
EXAMPLE 7.9 There exists a set E in the line such that dim;,ox(.E) ^ dm\box(E).
■ fc
Proof: We begin with the unit interval [0,1] and the sequence n^ = 10 , k = 0,1, 2, 3 , . . . . For stages 0 , 1 , . . . , n\ we do our Cantor construction where the mid dle interval removed from each A is 1/3 of the length of A. Then from stage ri\ + 1 , . . . , ri2 the middle interval removed has length 9/10 the length of A. We continue to alternate, taking away 1/3 on stage 712 + 1 to rij, and 9/10 from 77.3 + 1 to 714, etc. Let E be the resulting set. This brings us to dui\box
( £ ) > 1^5 >&mbox(E) from using the fact that 1/3 > 1/5 > 1/10. 6 Now let us look at some of the properties (good and bad) for the Box-Counting Dimension. If we use closed sets to cover up A, it does not make a difference in the value ofNg(A). Nor will it matter if A is open or closed. The count will be the same. This brings us to the following theorem: Theorem 7.9 For a set A, d\vabox{A) = d.\m.box{A) and dimkcx(A)
=
dimbox(A).
Applying Thereom 7.9 leads us to the following example. ffl EXAMPLE 7.10 Let A be the set of rational numbers in [0,1]. This set is dense in the interval so its closure is [0,1]. This means dimfcocc(A) = dim{,o:r([0,1]) = 1. ■ This example serves to show that Box-Counting dimension does not have the prop erty of countable stability; that is, for a countable collection of sets An it need not be true that dimi,ox(\jAn) = sup n {dim(, oa; (j4 n )}. The case of a finite collection of sets has a little more subtlety. There are examples to show that we can have d\mbox{A U B) > max{d\mbox(A), dimbox(B)}. On the other hand, for upper dimension it is true that d\mbox{A UB) =
m&x{dimbox(A),dimbox(B)}.
HAUSDORFF MEASURE AND DIMENSION
131
The following theorem is stated without proof, but the idea is very straightforward. It states that Box-Counting dimension is monotonic. Theorem 7.10 If A C B, then d\mbox(A)
< dm\box(B)
anddimbcx(A)
<
dimbcx(B).
We end this section with one more example. This example is a little counterin tuitive. It shows that a countably infinite set (so small in measure, category, and cardinality) can have nonzero dimension. W
EXAMPLE7.il Let A = {l/n}£° =1 U {0}. This set has dimbox{A) = 1/2.
■
Proof: Take an open set U of diameter 5 < 1/2. Then there exists an integer k such that 1 1 g (k-l)k> ~ k{k + l)' Thus U can cover at most one of the points in { 1 , 1 / 2 , 1 / 3 , . . . , 1/fc}. So at least k sets of diameter S are required to cover A, giving us MN5{A)) -In 5
-]n(k{k
lnfc + l))'
This leads us to dimbox(A) > 1/2. In the other direction, if 1/2 > S > 0, again take k so that l/(k(k — 1)) > S > \/k(k + 1). Then it takes k + 1 intervals of length S to cover [0,1/fc], leaving k — 1 points to be covered by another k — 1 intervals. Thus \n(Ns(A)) -In 5 ordh^box(A) 7.5
ln(2fc) -ln(fc(fc-l))'
< 1/2.*
Hausdorff Measure and Dimension
We have already seen measure in terms of Lebesgue measure. We now go on to explore a related idea (Hausdorff measure) and show how this measure gives us another idea of dimension. Let Abe a set in R n ; and by a 5-cover of A we mean a collection of sets £/» (finite or countably infinite), each of whose diameter is less than S for which A C UC/,. For Lebesgue measure on the real line, we looked at
inf (fScE/O J ,
132
VARIOUS NOTIONS OF DIMENSION
where £(I) means the length of the interval I and the infimum was over all possible covers. It is reasonable to see that if we restricted ourselves, we would have gotten the same value with 6—covers. Moving into M.n we will change notation a bit. Since we no longer talk about intervals and length, given a set U we define its diameter as \U\ =swp{d(a, b)
:a,b&U}
with distance using the Euclidean metric. Definition 7.9 Let 4 c l " and s > 0. We define the Hausdorff s-outer measure of A with respect to 5 by
Hss(A) = mf{j2\Ut\s} where the infimum is taken over all 5-covers of A. This value depends on both s and 6 and is not yet what we want. As we shrink the value of5 the collection of acceptable covers gets smaller so the value of the infimum gets larger. Taking the limit we get Definition 7.10 Let A C R" and s > 0. Then we define the Hausdorff s-measure of A by US{A) = lim HI(A). We see right away that W (0) = 0; and if A C B, then any cover ofB also covers A so HS(A) < HS(B). When s = 1, this is the definition of Lebesgue measure. For R™, this is a constant multiple of Lebesgue n-dimensional measure , the constant being the volume of the unit ball in W1. Let us fix the set A and let S be small (usually, S < 1 is required). Then for s < t,
n\{A)
<
inflEI^I'l^mfiEl^r^8} IS^-^nf
{j:\U.n
<nss(A).
Taking the limit as 5 approaches zero gives us the following theorem. Theorem 7.11 Let A C M.n be fixed. For t > s, U\A) decreasing function of s.
< HS(A); that is, Hs is a
There is, in fact, more here. Suppose A has HS(A) < oo. Then for t > s, ft (A) < lim,5_>.o 5t~s'Hs(A). The right side here is zero, though. In an "opposite" way, for t < s the value ofTLt(A) = oo. At some point, the value of the Haus dorff measure jumps from infinity to zero and that is where we find the Hausdorff dimension. 1
Definition 7.11 Let A C R™. We define dim# atis (^4), the Hausdorff dimension of A, by Aim.Haus{A) = sup{s : HS{A) = 0} = sup{s : lis{A) = oo}.
HAUSDORFF MEASURE AND DIMENSION
Ifs = dim.Haus(A), it may be true that lis(A) 0 < %s (A) < oo, then we say A is an s-set.
133
is 0, infinite, or positive and finite. If
Let us move on to the following example, which shows that this dimension, much like measure, behaves in an expected way with finite sets having dimension zero. fl
EXAMPLE 7.12 Let A be a finite set. Then dim/f ous (A) = 0.
■
s
Proof: This is an immediate consequence ofH being a measure as for any value of s > 0, HS(A) = 0. Let us note, though, H°{A) ^ 0 if A ^ 0. * Two other quick properties that fall out ofHs being a measure are: ■ For the countable collection {Ai}, dimHausiViAi) -
swp{d\mHaus{Ai)},
which is called countable stability. ■ If A C B, then dimi^ aus (A) < dim# ous (.B), called monotonicity. This shows a big difference between Hausdorff and Box-Counting dimensions since we have dim# a u s(Q fl [0,1]) = 0 and dimBo:r(Q H [0,1]) = 1. One aid in computing the dimension is what is referred to as the Open Set Con dition (OSC). It is a characteristic that shows the set, in a way, consists of separate pieces that do not overlap too much. We use this with an Iterated Function System {/i, / 2 , . . . , / „ } . The IFS satisfies the OSC if its invariant set contains a set O that is open relative to the metric space on the attractor A, such that
■ ■ Uf=1f(0)
fi(0)nfj{0)=9ifi?j,and C O.
If the IFS does not satisfy both criteria, then we call it overlapping. fl
EXAMPLE 7.13 1. For the IFS fi(x) = l/2(x) and f2(x) = 3x/4 + 1/4 on E, the attractor is the unit interval. This is an overlapping IFS. 2. On R, / i (x) = x/Z and f2(x) = x/3 + 2/3, whose attractor is the Cantor Set, and the IFS satisfies the OSC. Just use the open set (0,1). 3. In the plane, the IFS 7i(x, y) = (x/2, y/2), T2{x, y) = (x/2 + 1/2, y/2), and Ta(x, y) = (x/2 + 1/4, y/2 + 1/2) from Example 6.25 satisfies the OSC using the open unit square.
134
VARIOUS NOTIONS OF DIMENSION
If a set satisfies the OSC, then computing dimensions becomes easy. The proof of this is a little lengthy for us, though. Theorem 7.12 Let A be the invariant for the IFS {/i, / 2 , • • •, fn} with ratios {c\, C2, c%,..., cn}. Suppose the system satisfies the Open Set Then = dimBox{A) = s, dimHaus(A)
contraction Condition.
where s is the solution to the equation c\ + c | + c | + • • • + csn = 1. Also, for that value ofs, 0 < rls(A) < oo. Let us return to our continuing example of the Cantor Set and find its Hausdorff dimension. S
EXAMPLE 7.14 Using Theorem 7.12, we find that the Hausdorff dimension of the Cantor Set is also In 2 / In 3, just like its Box-Counting dimension. ■
Because Hausdorff dimension is based on measures, various techniques have been devised for determining the measure and, from that, the dimension. In computing Hausdorff measure for a set, A, there is usually a two-pronged ap proach. We look at open covers of the set to find an upper bound for the measure and then find a lower bound. When they agree, we get HS(A). Lower bounds can be difficult to find. One technique used to get this involves the idea of a mass distribu tion. We define this for a measure /J, on the space M". The support of a measure /i is the smallest closed set X such that /i(R™ \ X) = 0. This set, supp(p), is closed. It is also true that x G supp{p) if and only if for every positive r, p,(B(x, r)) > 0, where B(x, r) denotes the open ball with center x and radius r. Now we define a mass distribution. Definition 7.12 A mass distribution is a measure on a bounded subset ofW1 that 0 < fi(Rn) < oo.
such
This is often thought of as taking a finite mass and spreading it out on a set X. S
EXAMPLE 7.15 Let a be a point in the set W1. Let Abe a subset ofW1.
^)
=
We define fi by
{ 0 a^ A
This is a mass distribution where all of the mass has been concentrated on a single point. m
HAUSDORFF MEASURE AND DIMENSION
fl
135
EXAMPLE 7.16 Let L be a line segment in R. Then a mass distribution with support L is given by fi(A) = m(L n A), where m is Lebesgue measure. m
A typical method for constructing a mass distribution follows closely with the canonical Cantor Set construction. Recall from Chapter 2 that a Borel set is formed by the a—algebra (collection of countable unions and complements) of open sets. We start with a Borel set A and an initial collection, EQ which consists solely of A. Inductively, Ek is a collection of disjoint Borel subsets of A such that for U € Ek is a subset of one of the sets in Ek-i and contains finitely many of the sets in Ek+i, call them {U\, U2, ■ ■ ■, Uj}. It is also required that for the sets in Ek, the maximum diameter of the sets decreases to zero as k —► 00. Finally, the mass in set U is divided among the {Ui, U2, ■ ■ ■, Uj} so that n(U1) + n(U2) + --- + n(Uj) = fi{U). Define £k to be the union of all the sets in Ek- We see that p(Rn \£k) = 0. We now get the result we are seeking, without proof. Theorem 7.13 Let p be defined on a collection of sets £ as above. Then p can be extended to all subsets in the space R™ by p(A) = p(A n £) and this is a measure on R™. If A is a Borel set, then this number is unique. The support of p is contained in C\£kNow we move on to applying this result to Hausdorff measure and dimension. This is the so-called Mass Distribution Principle found in [30]. Mass Distribution Principle Let \ibe a mass distribution on a set A C R n and suppose that for some s there are positive values c and 5 such that p(U)
< dimboa;(A) <
dimbox(A).
Proof: The result is an application of the following. If {Ui} is any cover of A, then
0 < p(A) = piUiU,) < 5>(£^) < cJ2 l^lsi
i
Apply the infimum process to each part and we see ~ris(A) > p(A)/c and for small 8 we have Ha (A) >
p(A)/c.^
136
VARIOUS NOTIONS OF DIMENSION
Let us apply this principle to the Cantor Ternary Set. For our mass distribution on C, let us suppose the unit interval has mass 1. At Step k, we have 2k subintervals of length 3~k. On each of those subintervals, let the mass be 2~k. Essentially, the mass then belongs to subintervals I is split equally among subintervals Ii and Ir at the next step. Let U be a set with \U\ < 1. Let k be the integer such that 3-k+1
< \U\
<3-k.
Then U can intersect at most one interval at Step k so H(U) < 2~k = (3"fc) l n 2 1 "3 <
(3\U\)ln2/ln3
and so %la2l l n 3 ( C ) > 0 by the Mass Distribution Principle. Thus dim Haus{C) < In 2 / In 3. In Chapter 10, we will apply mass distributions to a generalization of the Cantor Set construction to get a fix on the dimension of this new type of set. Some other results on Hausdorff dimension are the following. Theorem 7.14 If S is a complete metric space, dimin(i(S) [26]).
< dimHaus(S)
(see
Theorem 7.15 Let f : X —> Y be a similarity with ratio r > 0. If A C X is a Borel set, then "H s (/(^4)) = rsT-L(A) for any positive s and dirri# a M S (/(A)) = d.\mHaus(A). Proof: For any set U, since f is a similarity, \f(U)\ = r ■ \U\. Thus the sets which make a 5-cover of A form an rd-cover off(A). The first result follows from this and the second from the first. 4* Some of this will be especially useful when looking at products of sets.
7.6
Miscellaneous Notions of Dimension
Packing dimension is similar to Hausdorff dimension, but involves requiring the covering set to be disjoint. Let A be a set and give S > 0, let B = {Bi} be a collection of disjoint balls which cover A where each Bi has radius at most 6 and is centered at a point in A. Define the the S-packing number as
PZ(A) = sup I. Y,\Bi\a\, where the supremum is taken over all possible covers. As 8 approaches zero, the value of VI {A) must decrease and therefore has a limit. Let VZ(A) =
llmVZ(A).
MISCELLANEOUS NOTIONS OF DIMENSION
137
This, unlike for Hausdorff, is not a measure. If we have a countable, dense set each individual element has Vo({xi}) = 0, but as a whole Vo{U{xi}) > 0. We modify this by looking not just at the set A, but at sets containing A. So we come to the s-dimensional packing measure of the set A:
P°(A) = inf J5> 0 S (^) : A C U ^ ^ I , where the infimum is over all covers {Ai}. As with Hausdorff, as s increases, the value of Vs goes from positive infinity to zero, and that is where we find the packing dimension: dimp(A) = sup{s : VS{A) = 00} = inf{s : VS{A) = 0}. As the covers are more restrictive than we use for Hausdorff, yet do count as what we use for box-counting, we find that dimHaus(A)
< dimp(A) <
dimbox(A).
In a similar vein, we can change our definition of box-counting by using countable covers of the set. Our result will give us countable sets have dimension zero (just as with topological dimension), even when they are dense. The upper and lower modified box counting dimension of A are given by dhnMB(A)
= inf I sup{dim Bo:c (A i )} : A C U ^ ^
and
dirnMB(A) = inf Lup{dmBox(Ai)} ■ A Q ^ZiAi I. Now for any set A C M™ we have 0 < dimHaus{A) < d i m M B ^ ) < dim M B(-4) < d\mBox{A)
< n.
Now we turn to Besicovitch-Taylor dimension which first appeared in [7]. This is based on the gaps in the set and is defined for closed sets in the real line. There are some attempts to expand this to M.n ([64]), but we shall forego them here. Let E be a bounded and closed set in R and, without loss of generality, let us assume E C [0,1]. Then [0,1] \ E is a countable union of open intervals, {On}. This countable union can be ordered in such a way that the sequence {an} where an = (-{On) forms a nonincreasing sequence of positive numbers called the gap sequence. For the Cantor Set, this sequence is
138
VARIOUS NOTIONS OF DIMENSION
1/3,1/9,1/9,1/27,1/27,1/27,1/27, If E has Lebesgue measure zero, then
E°" In the other direction, ifty = {an} satisfies ^an = 1, then there exists closed sets B c K such that [0.1] \E consists of intervals with lengths in ^. Then we say E belongs to ty. If we let rn = 5Z7-_„ aj, then rn is a decreasing sequence with rn —> 0. This leads us to the following results about Hausdorff measure and dimension. Theorem 7.16 Suppose ty is a nonincreasing sequence of positive numbers satisfy ing Y^ o-n = 1 and E is a closed set in R belonging to *&. Then for 0 < s < 1 HS{E)
Proof: Define X(s, <3>) by A(s,*) = liminf n
. r, n
There is nothing to prove if\{s, *) — oo, so we will assume this is not the case. So lim infn^oo [n ( ^ ) ' ] < oo and let E be a fixed set belonging to Vf and [0,l}\E
= UnQn,
where the Qn are disjoint open intervals with £(Qn) = an. Define En recursively as Ei = [0,1] and En = Ei \ U™=1 Qi, n = 2, 3,4,.... Then En consists ofn closed intervals of total length rn and E is a subset of Enfor each n. Given e > 0 and r\ > 0, there exists an integer m such that rm < V and
mC-^-X
Vm ) Then each Em contains m intervals of total length rm, and we can see that the sum of the lengths of the intervals in Em is bounded above by X(s, \&) + e; and since r\ and e are arbitrary, we get our result. 4b To convert this to a result about Hausdorff dimension, let an be defined by (rn\an
i
MISCELLANEOUS NOTIONS OF DIMENSION
and let
139
a($>) = liminf an.
Corollary 7.16.1 Suppose E C R is closed and belongs to ^ with a(^) as above. Then 0 < dimHaus(E) < a ( # ) . Besicovitch and Taylor also show the following result about Cantor Sets Ca: If t < dimHaus{Ca), then for any 7 > 0 there is an arrangement of E, the gaps in [0,1] \ Ca, such that ri^E) = 7. This work on gap sequences gives us another type of dimension. This is called the Besicovitch-Taylor dimension. Given £ c l with gap sequence <3/ = {an} we let <jp = XI a n where 0 < /3 < 1 since we are on the real line. Then the BesicovitchTaylor dimension of the set E, written dim.B-T(E), is the smallest (3 for which that sum is finite; that is, &\YO.B-T(E)
= inf{/3 : 073 < 00}.
For the Cantor Set, the Besicovitch-Taylor dimension agrees with Hausdorff and others, diuiB-T(C) = In 2/ In 3. We will revisit this topic in Chapter 10. The following gives us another criterion for Hausdorff s-measure zero sets in the line. Corollary 7.16.2 Let E be closed in M. where E belongs to * and assume ^ a* < 00 for some a s G (0,1], Then Hs{E)=0. The gap sequence also relates to upper Box dimension for a set. If a set has Lebesgue measure 0 and is compact in R, then the gap sequence determines the value ofdimBox(E). The proof of this can be found in [71]. Theorem 7.17 Let E be a compact subset ofM. with Lebesgue measure 0 and let {an} be the gap sequence of E. Then -— Inn d\m.Box{E) = limsup — . n—>oo
For the Cantor Set, - 4 p ~ = — In otn
lnO!ra
' n 2 " „ = In 2/ In 3. — In i
n
'
As with many of the topics we cover, there are different paths down which to travel, yet we cannot look at much. Doing so would bring us too far from the topic of Cantor Sets. However, the interested reader is encouraged to learn more about the different notions of dimension and how to compare and contrast them.
140
VARIOUS NOTIONS OF DIMENSION
EXERCISES 7.1 For each set below, determine if it is bounded above. If so, give three examples of upper bounds. If not, declare the set not bounded above. Then go over them again, this time for lower bounds. a) [-1,1] b) {1/n : n e N } c) j r e Q : r < 2 ] d) {xeQ:x2<2} e) UneN(n,n+ 1/n] f) n~ = 1 [0,l/n] 7.2 For the sequences xn = ( — 1)", yn = 1/n, lim sup and lim inf. 7.3
an
d zn = sin(n7r/2) find the
Let an and bn be the sequences defined below: {a n } = {0,l, 2,0,1,2,0,1, 2 , 0 . . . } {6„} = {-l, 0 , 1 , 1 , - 1 , 0 , 1 , 1 , - 1 , 0 , . . . }
Find a) b) c) d) e) f) 7.4
lim sup an + lim sup bn lim sup{a„ + bn} lim inf an + lim inf bn lim inf {an + bn} lim sup a„ + lim inf bn limsup{anfe„}
The following statement is incorrect:
Let {xn} be a bounded sequence and k £ K. Then limsup{fcx„} = fclimsup.T„. Find a counterexample that shows this is wrong. Determine changes in the hypothe sis that will make the conclusion true. 7.5 What do the solutions to the two problems in Exercise 6.8 say about similarity dimension s ? 7.6
Show that if A is a finite set, then &uni,ox(A) = 0.
7.7 Suppose E is a bounded set and f : E —» K has the property that for all x and y, \f(x) — f(y) | < k\x — y\ where k is constant (such a function is called a Lipschitz function). Show that (MLBox(f(E)) 7.8
< dmBox(E)
anddimBox(f(E))
<
dimBox(E).
Let E = {x\,X2, X3,... ,xn}. Show that'H0(E) is the number ofpoints in E.
EXERCISES
7.9
141
Prove that if {E^} is a collection of sets
us{uEk)
Suppose f is a Lipschitz function. Show that dimHaus(f(E))
<
dimHaus(E).
7.11 Let f : R -» R be the function f(x) = x2. Show that for any set E C R, dim H a u s (.E) = dim f f o u s (/(£)). 7.12
Find the Hausdorjf dimension of the set {0,1,1/4,1/9,1/16,...}.
7.13
Show that % 0 (C) — oo and V} (C) — 0, where C is the Cantor Set.
7.14 Find the Box-Counting dimension of a general Cantor Set, Cr. What happens asr -> 0+ ? 7.15 Compute the Box-Counting dimension for the Sierpinski Triangle (see Exam ple 6.25).
CHAPTER 8
POROSITY AND THICKNESS-LOOKING AT THE GAPS
The main purpose of this chapter is to introduce readers to some reasonably new ideas and results which involve the Cantor Set. The sources for much of this are some papers which are very technical in their definitions, proofs, and constructions. To keep things at the correct level, there will be many statements here that are not matched with proofs. The list of references in the back should help the interested reader find out more.
8.1
The Porosity of a Set
Porosity was first studied by Denjoy in his work on trigonometric series in 1920 [20]. The subject arose again, this time with the modern nomenclature, in the 1960's with Dolzenko's work in cluster sets [25]. Similar to the definition of continuity, we begin with a pointwise idea and then expand it. Porosity can be defined for any metric space (S, d), but we shall begin with sets in the real line, moving later on into metric spaces. By starting in R, we will have an additional restriction beyond just looking pointwise. Wefirstlook only to one side ofxQ, finding the right porosity. The Elements of Cantor Sets -With Applications, First Edition. By Robert W. Vallin Copyright © 2013 John Wiley & Sons, Inc.
143
144
POROSITY AND THICKNESS-LOOKING AT THE GAPS
Definition 8.1 Let £ c l and let XQ £ £ ' . For a fixed r > 0 let X(E, XQ, r) denote the supremum of the lengths of the open intervals in Ec fl (XQ, XQ + r) (this can be any number in [0, r]). Then the right porosity of E at XQ is given by + /TP
\
v
\{E,x0,r) r->o
r
Similarly, there is the left porosity of E at x$, written p (E,XQ) r,xo) n Ec. Finally, the porosity of E at x0 is given by p(E,x0)
m&x{p+(E,xo),p~{E,x0)}.
=
\(E,x0,r)
x 0_
fl
and using (XQ —
EXAMPLE 8.1 Let E = {0} U { l / 2 n } ~ = 1 . Then p+(E,0)=
lim
^ " 2 ^
1 ~ 2
Notice that although the definition is written as a l i m s u p of a function (ofr), we could find our answer as a limit of a sequence. This is often the case. The following are properties and definitions that emerge from the concept of right porosity. 1. Porosity at a point is a value between zero and one. Ifp(E, XQ) = 0, we say E is nonporous at XQ. Ifp(E, Xo) — 1 we say E is strongly porous at XQ. 2. A set E is a called a porous set ifp(E, x) > Ofor each x 6 E. 3. If a set A = U'^L1En and each En is a porous set, then we say E is a-porous. The following theorem and its corollary are presented without proof, but details on this and other properties can be found in [80]. Theorem 8.1 If E is a porous set, then E is both measure zero and nowhere dense. Corollary 8.1.1 If E is a a-porous set, then E is both measure zero and first cate gory. Some definitions do not require XQ to be a point in E, but we are only interested in points in the set.
THE POROSITY OF A SET
145
These results say that porosity has both an analytical (measure) and topological (category) nature to them. The advantage to using porosity arguments is that we get two results via only one proof. As stated before, these ideas generalize to metric spaces. In such a setting a porous set will be nowhere dense, but measure zero depends on whether or not the underlying space has a measure on it. These results do not reverse themselves. There are sets that are measure zero and nowhere dense that are not porous. In fact, they are not even o -porous. Now let us look at porosity and the Cantor Set. As we can see in the picture A(C,0,2/3) = l / 3
0
" -v-
r=2/3
and via the self-similar nature ofC, we have P+(C,0)>\. Of course, p~ (C, 0) = 1 which leads to p(C, 0) = 1. This is actually the typical value. The following theorem is due to Denjoy and shows that most of the points in a set are strongly porous from both sides (bilaterally strongly porous). Theorem 8.2 If E C R is closed and nowhere dense, then the set of points in E at which E is strongly porous from both the left and the right sides is a set which is residual in E. The literature also contains what is now called porosity index . This was what Denjoy was working with in [20]. Definition 8.2 Let £ c l and let XQ S E. For a fixed r > 0 let h, k be positive real numbers such that (XQ + h, XQ + h + k) D E = 0 and h + k < r. Then the right porosity index of E at XQ is given by i+(E,xo)
k = lim sup —. r->0
h
Similarly, there is the left porosity index of E at XQ, written i~(E, XQ). Finally, the porosity index of E at XQ is given by i{E,x0)
=
max{i+(E,x0),i~(E,x0)}.
The index can take any value in [0, oo). In a similar vein as before, E is small ifi(E,xo) > 0 for all xo £ E. As we now see, porosity and porosity index are equivalent notions. Relating the two concepts is the following lemma whose proof is obvious.
146
POROSITY AND THICKNESS-LOOKING AT THE GAPS
Lemma 8.1 Let p = P+(E,XQ)
andi = i+(E,xo). p—
1+i
and i
Then
l-p
We should at some point look at an application of this idea. One of the most com mon uses of porosity is in the description of an exceptional set in the real line. The quick definition of an exceptional set would be when something happens everywhere except on some set E. What we then need is a good way to classify E. The usual ways are measure, category, and cardinality. Since there are measure zero, nowhere dense sets that are not porous, describing E as porous is a stronger result that even both of the other two. Recall a quick definition before our example theorem, we all know that for f : R —► K to be differentiable at XQ we must have that H
f{xo + h) - f(x0)
fr->o
h
exists and is finite. In a similar vein, we can use h —» 0 + (h —> 0~) to define the right (left) derivative of f at XQ. If f is differentiable at XQ then f has equal left and right derivatives at XQ, while f(x) = \x\ and XQ — 0 shows that having finite left/right derivatives does not imply differentiability. So our application of porosity is this theorem from [6]. Theorem 8.3 For an arbitrary function f : R —> R, the set of points of which f has a finite right derivative, but not a two-sided derivative, is a a-porous set.
8.2
Symmetric Sets and Symmetric Porosity
One aspect of Cantor Sets that we have not looked at yet is the symmetry of the set. If XQ is a point in C, then so is its reflection about the point 1/2, call this yo To find that reflected point, it is the same distance from 1/2 as XQ, that means either y0 = 1/2 + (1/2 - x0) (ifx0 < 1/2; or y 0 = 1/2 - (x0 - 1/2) (when x0 > 1/2). Either way, that brings us to yo ~ 2(1/2) — XQ. This leads us to a new definition. Definition 8.3 A set S is called symmetric if there exists a point t (not necessarily in S) such that for all x 6 S the point 2x — s is also in S.
&
EXAMPLE 8.2 An open (or closed) and bounded interval is symmetric about the midpoint of the interval. ■
fl
EXAMPLE 8.3 The Cantor Set is a symmetric set.
■
SYMMETRIC SETS AND SYMMETRIC POROSITY
147
In fact, all Cantor Sets Cr are symmetric about t = 1/2 by virtue of the con tiguous interval being taken from the center of each subinterval. We move from symmetric sets to a symmetric version of porosity for the real line. A word on nota tion here, for a point x and a set A, the symmetric copy of A about the point x is 2x - A = {2x - a:a e A}. Definition 8.4 Let £ c l and let XQ G E. For a fixed r > 0 let j(E, XQ, r) denote the supremum of the lengths of the open intervals A such that A n E = 0 and (2xo — A) fl E = 0 (this can be any number in [0, r]j. Then the symmetric porosity of E at xo is given by sfjp
N
,-
p (is, XQ) = limsup -
l(E,x0,r)
Just as before, this porosity can be any value between 0 and 1, inclusive, if ps(E,xo) > 0 we say E is symmetrically porous at XQ- The set E is symmetri cally porous if it is symmetrically porous at each of its points. In addition, there is the notion of a-symmetrically porous. If a set is symmetrically (asymmetrically) porous, then it obviously is porous (aporous). This idea does not flow in the other direction. The following result is from Evans [28]. S
EXAMPLE 8.4 There exist sets which are porous, yet not even asymmetrically porous.
■
We change our focus a little in this section. When we refer to the number r, it will reference the relative since of the interval removed; that means for each r, the Cantor Set in question is C(l-r)/2-
By the way that the contiguous intervals are removed, one would expect the Cantor Set to be a symmetrically porous set.This is true for the ternary set. In fact, for r > 0 fixed, any C'(i- r )/2 will be a symmetrically porous. Suppose though we let the size of the interval removed change at each iteration. Definition 8.5 Let rn be a sequence of real numbers with 0 < rn < 1. We define the Cantor Set C(i_ rn )/2 by the following construction. At Stage I, the interval 1% = [0,1] is divided into three pieces IQ, O, and I\ such that I = J0 U O U h and
£(I0) £{i)
=
1(h) e(i)
=
1 2y
l)
'
At stage k, there are 2k subintervals of size YljZi(~ip~)- ^a^ tnese {la}, where a is a string ofk — 10's and 1 's and we write the length of the string by \a\. Take each
1 48
POROSITY AND THICKNESS-LOOKING AT THE GAPS
Ia and divide it into three pieces Ia,\, Oa, and 7^2 such that I =
IafiUOaUlatl
and
KLfi) 1{I„)
=
*(/„,!) _ i nU l{h) 2
,
Tkh
The Cantor Set C( 1 _,.)/ 2 is then defined to be
c
\\
j
Now we have the following result relating Cantor Sets and asymmetric porosity. The proof is very detailed and will be left out. Theorem 8.4 For the Cantor Set C(i_ r )/2 the following are equivalent: 1. C(i_ r )/2 is porous. 2. C(i_ r )/2 is a-porous. 3. C( j _ r ) /2 is asymmetrically porous. 4. limsup n _ >(X) r n > 0. In the list above, symmetrically porous is left off, This is due to the following example by Evans, Humke, and Saxe [27]. S
EXAMPLE 8.5 There is a asymmetrically porous Cantor Set which is not symmetrically porous. I We will not prove this example, but the proof begins with the statement Let S = {k2/2 + 9k/2} and for each n set
(_ 0
otherwise
This means we are removing intervals, but not at every step and with quickly decreas ing frequency. We will still be able to break this apart into symmetrically porous pieces, but have so many O's in the sequence that we destroy the possibility of being symmetrically porous as a whole. As a comparison with another indicator of a set's size, in [74] a strongly symmet rically porous set with Hausdorff dimension 1 was constructed. This was done in a fashion similar to that of the above example.
A NEW AND DIFFERENT DEFINITION OF CANTOR SET
8.3
149
A New and Different Definition of Cantor Set
Previously, we used a fixed ratio r to create the set Cr. This point of view looked at the relative size of the open interval removed versus the original interval. Here we shall just collect together the countably many intervals removed (0\, 02,03,...) and focus on the order in which these Oi are listed. Let us define a Cantor Set C by
c = i\\Jot, i>l
where I is a finite closed interval and {Oi\i > 1} is a (finite or infinite) collection of disjoint open intervals contained in I. Notice here that this is a broader defini tion than we have had before; that is, there are sets which are Cantor Sets by this definition which are not Cantor Sets by the previous one. Some terminology and notation: We are going to restart our notation, beginning at ground zero in order to match the notation the interested reader would find in the literature on this topic. By a tree T> we mean a collection of strings of O's and 1 's where for our purposes, 0 denotes the left and 1 the right. Each finite string of O's and 1 's, UJ, is called a word and \UJ\ will denote the length of the word. Let I be the root of the tree. We'll refer to {1} as the zeroth level of the tree. Now suppose the tree has been defined to the nth level. We define the (n + l)st level as follows: Let Iu be an nth level vertex (a word consisting ofn O's and 1 's) of our tree. Assume first that
^n IU °A ±0Let Oju be the interval in the set {Ot\i > 1} of least index which is contained in 1^, and let I^0 and I"1 be the closed intervals with
I" = f ° u O , - U f 1 . We then let I"0 and 7""1 be subvertices ofIu in V. If
l n
" ( U °\ = 0'
then we set I"0 = Iu and let I"0 be a subvertex of I" in V. We repeat this pro cess for every vertex in 1^ in the nth level ofV. The (n + l) s * level of the tree is the set of vertices in Iu in V with \u\ = n + 1, where \u\ denotes the length of the word v. We continue this process inductively, creating the infinite tree V. The intervals 1,1°,I1,!00,101,... are called the bridges while the open intervals Oj, Ojo, Oji,... are called the gaps. Note that {O/c \I" is a bridge ofV} = {0{\i > 1},
150
POROSITY AND THICKNESS-LOOKING AT THE GAPS
hence oo
/
Any tree with this property is called a derivation of the Cantor Set C from I. While the end result will always be the same, the set one has at the nth step is determined by the order in which the Oi are labeled. In the picture below, which is the construction of the ordinary Cantor Set, on the left side the interval (1/9, 2/9) is the first interval removed, while on the right side the interval (1/3, 2/3) is removed first.
Thus different orderings of {Oi}, called derivations , give one different bridges and gaps. Comparing the relative sizes of bridges and gaps and looking across all order ings it the idea behind the thickness of a Cantor Set.
8.4
Thickness of a Cantor Set
This idea of constructing a Cantor Set in more than one fashion by using different ordering of the intervals removed leads us to the idea of thickness. Thickness is yet another way of measuring the size of a set. Using C to represent a Cantor Set, we use T(C) to denote the thickness of the set. If {Oi : i > 1} = 0, then we set T(C) = oo. If{Oi : i > 1} ^ 0, then r 6 ) = sup inf mm < ——F, ——r } , ■D A&V \\0A\ \OA\) where the supremum is over all derivations V ofC, the infimum is over all bridges A ofV, and \A\ means the length of the interval A. This supremum is always attained. Determining the thickness of Cantor Sets is not as intimidating as it sounds, thanks to the following Definition and Lemma. Definition 8.6 We say a derivation is ordered iffor intervals A and B in the deriva tion with A = A0 U OA U A1, B = B° U 0B U B\ and B C A, then \OA\
> \OB\.
Lemma 8.2 Let T> be any derivation ofC from I. We shall denote the thickness ofC with respect to a fixed derivation V as r-p. Then there exists an ordered derivation T>o of C from I with MC)
APPLYING THICKNESS
151
Furthermore, ifT>\ and T>2 are two ordered derivations of C from I, then TVl(C)
8.5 Applying Thickness Now let us apply some of this. For two sets A and B, their sum A + B is defined as A + B = {a + b: a<= Aandb e B}. This idea generalizes something already known, translation of a set. If we have a set P and add the number k to each point in P, the result is just P translated over k units. Typically this is written as k + P. So if P = [0,1], then 2 + P = [2,3] and if S = {-1,2}, then S + P= [-1,0] U [2,3]. Another interesting fact about the Cantor Set C has to do with the sum of C and itself. As we saw in the previous chapter, C is nowhere dense and measure zero and thus very small in two different ways. It turns out, however, that in another way this set is big. Theorem 8.5 For the Cantor Ternary Set C, C + C =[0,2]. Proof: Recall that C = { i £ [0,1] : the tertiary expansion of x consists of only 0's and 2 's}. Then 1/2(C) = {\x where x E C) are the numbers whose base 3 expansion con sists of just 0's and 1's. It is easy to see that if x € [0,1], we can decompose its expansion into the sum y + z where y and z consist of only 0's and 1 's in base 3. This shows
\c + \c = M and our desired result comes from multiplying both sides by 2. 4 This is the typical proof one finds in the literature. There is also a visual proof, the so-called "proof without words." In this, we see that any line of the form x + y = k
152
POROSITY AND THICKNESS-LOOKING AT THE GAPS
where k £ [0, 2] will, by use of the self-similar nature ofC, pass through a point in C x C, giving us our result. J L
\
J U
3 ^
P
:: d
n c.
I B I r
J L_ "1 C
-a r:
aL
U L
i
■ I'
\
\
n r a Ei
a ■
\ fl ■
\
\ \
J L 9 p
\ "i r \
n r nn
V«wa/ Proof that C + C ={0,2} Now we look at the sum of two Cantor Sets through the lens of thickness and porosity. This leads in the obvious fashion to the sum of a Cantor Set and itself. Theorem 8.6 For j = 1,2 let C0 be a Cantor Set derived from Ij with 0} a gap of maximal size in Cj. Assume that |Oi| < \I2\ and \02\ < \h\. IfT(C\) ■ T(C2) > 1, then C\ + C2 = h + hUsing the relationship between thickness and porosity, [73] has the following, which is a corollary of the corresponding thickness result in Theorem 8.6. For this, the Cantor Sets, Cj, have a constant ratio, Tj. Corollary 8.6.1 For j = 1,2 let Cj be a Cantor Set derived from I3 with Oj a gap of maximal size in Cj. Assume that |Oi| < |/ 2 | and \02\ < \h\. Ifp(Ct) + p(C2) > 1, then C1+C2
= Il+
I 2.
This result is not an "if and only if" type. Falconer, in [29] has an example of Cantor Sets C\ and C2 that are strongly porous (p(Cj) = I) yet C\ + C2 is an interval. In [56] and [61] the following question was asked and answered: For any k £ [0, 2], what is the cardinality of the set ofpoints (x,y) e C xC such that x + y = k. The short answer is, depending on k, the set is either finite or uncountable. The detailed answer, given by Nymann, depends on the base 3 expansion of h = k/2. For this we also need to define a value c(k). Let c(k) be the cardinality of the set
A BIT MORE ON THICKNESS
153
{n € N : an = 1}. Suppose h has a unique ternary expansion. Then the expansion cannot end in a string of zeroes. Define h as
where an = 0,1, or 2, an ^ Ofor infinitely many n, and an ^ 2 for infinitely many n. Now ifx, y £ \C, then x = Y,eJTandy
=
Y,Zn/3n
with £n,£n = 0,1. Given x + y = h, then if an = 0, it must be en = e'n — 0, and if an = 2, it must be en — e'n = 1. However, when an = 1 we can have either en — \ and e'n = 0 or en = 0 and e'n = 1. Thus there are 2c(k' choices for (x, y) (this is uncountable ifc(k) is infinite). Ifh has two different ternary expansions, then they are of the form h = Q,a\a2 ■ ■ ■ ar222 ... = Q.a\CL2 ■ ■ ■ ar~ibr, where a-y, a,2, ■ ■ ■, ar-i are 0, 1, or 2, and ar = 0 or 1, with br = ar + 1. Then we have either h(2c{k)) if ar = 0, [ ] * \3(2 C W- 1 ) if ar = l. For more on sums of Cantor Sets, see Section 10.7. 8.6
A Bit More on Thickness
The sum of two sets (our big application of thickness) can also be looked at as an intersection of sets issue. For a value x, S + x = {s + x : s G S} and S + T = {s + t: s e S andteT}. Then S +T =
{x:Sn(x-T)^®}.
In Theorem 8.6 we assume that \0\\ < \I2\and\O2\ < |/i|. The idea behind this is that neither of the Cantor Sets can lie in an open interval from the construction of the other. When this condition is met, the two Cantor Sets are said to be inter leaved. Williams, in [78], shows the counterintuitive result that intersections can be small, even when the thickness of the two Cantor Sets greater than 1 and the sets are interleaved. In [39], the authors answer the question, When is the intersection of two Cantor Sets empty, non-empty but possibly a singleton, and non-empty, but large in terms of thickness? Let C\ and C2 be Cantor Sets with thicknesses T\ and T-I, respectively. We will call this a thickness pair (T\ , T2). First we need the following two functions ofr.
154
POROSITY AND THICKNESS-LOOKING AT THE GAPS
/ T) =
r 2 + 3r + 1
andg(j) =
(2r + l) 2 7
.
Afext, we define three regions in the first quadrant (since r is non-negative) as fol lows: one region (called Region A) is in the quadrant, but below T\ ■ T2 — 1; the second is above Region A and below the area defined by n > T2, TI > / ( r 2 ) , andr2
>
g(n).
T"2 > n , T2 > f{n),
>
g{r2).
or andri
This is Region B. Finally, Region C is the area above Region B. This leads us to Theorem 8.7 Let C\ and C2 be Cantor Sets with associated thickness pair (TI, T2). If the pair lies in Region A, then C\ fl C2 can be empty. If the pair lies in Region B, then C\ fl C2 is non-empty, but can be just a singleton. If the pair lies in Region C, then the C\ fl C2 must be non-empty and can be either an interval or a Cantor Set, but in either case the intersection must contain a set of positive thickness. One last quick note using thickness. In [55] there is a lower bound result on Hausdorff dimension. This is important as upper bounds can usually be quickly obtained using covers of the set, but lower bounds are harder to come by. This paper shows that if the thickness of a Cantor Set C is r, then dunHaus(C)
>
^Jf^.
The bound is sharp as this is exact for the Cantor Ternary Set.
8.7
Porosity in a Metric Space
Like measure, category, and cardinality, porosity lends itself to the idea of a small set from Chapter 3. For porous sets, the small sets are the sets which are a-porous and when written as UE^, p(Ek) = 0. The fact that this is small becomes obvious when we recall that p(E) = 0 implies the set is measure zero. Just as with those other indicators of size, porosity can be taken to other spaces. For this we do need the idea of distance, so we cannot have just a topological space, but need a metric space. Definition 8.7 Let (X, d) be a metric space. Furthermore, let At C X and x £ M. For R > 0, we define A(x, R, M) as the supremum of all r > 0 so that there is a z e X such that B(z,r) C B(x,R) and B(z,r) fl M = 0, where B(z,r) is the open ball with center z and radius r. We then define the porosity of M at x by , w s ,. 2A(x, R.M) p(M,x) = Inn sup — .
POROSITY IN A METRIC SPACE
155
As before, we have the following: • p(M, x) is a value between 0 and 1. ■ The set M is porous if it is porous at each of its points. • The set M is a-porous if it can be written as the countable union ofporous sets. The appearance of the "2 " in the numerator of the definition seems a bit out of place, but recall from the definition in K the ratio is a diameter versus a radius. The following lemma from [80] is useful. Lemma 8.3 Let M c K and let N = M x Rn c R n + 1 . Then N is a porous subset (or a-porous subset) ofM.n+1 if and only if M is also as a subset ofM.. We leave this section pointing out one other type of porosity that appears in lit erature. In this version, there is a real-valued function associated with the distance A. Definition 8.8 Let g : K —> M. be a function which is increasing on the interval [0, h) for some positive value h with g(0) — 0. Let (X, d) be a metric space and M C X and x £ M. We say x is (g)-porous at x if limsup H->O+
—
> 0, R
where A(x, R, M) is defined as before. There are numerous other definitions that generalize the notion of porous. These include shell porous sets and superporous sets. Superporous sets also lend them selves to creating a topology. As with dimension and fractals, there are many roads upon which to travel, and the interested are encouraged to find out more.
156
POROSITY AND THICKNESS-LOOKING AT THE GAPS
EXERCISES 8.1 In the definition of Cantor Sets that we use for thickness, the set {Oj} can be finite. Explain how this opens us up to Cantor Sets that are not allowed in the canonical definition of a Cantor Set. 8.2 Note in the definition it does not say {0{\ has to be non-empty. What happens if it is empty? 8.3 One more difference. For an interval I, the closure of I, written I, is the in terval union its limit points2. So [0,1] is the closure 0/(0, 1) (and also the closure of (0,1], [0,1), and [0,1}). Anyway, other people, in their definition of this construction, insist that when i ^ j . Find an example of a "Cantor Set" that meets Astels definition [1] but where for some i.j. 8.4
Determine the value of
*HM}) for the Cantor Ternary Set using the usual ordering of the intervals removed (the typical derivation). 8.5
Show that the usual derivation for the Cantor Ternary Set is ordered.
8.6
Prove
T(C1/3)=1.
8.7 Find r(Cr) where 0 < r < 1. 8.8 Use the thickness, r(Cr), to determine the values of r so that Cr + Cr is an interval. 8.9 What about Cr and Cs ? Ifr is fixed, find the values of s which create a Cs so that Cr + Cs has an interval. 8.10
Give an example of a set in K2 that is porous, yet connected.
8.11
A set E is called very porous from the right at x if p^(E,xo)
= nm mi r—>0
. T
Similarly there is very porous from the left, very porous at a point, and very porous sets. Give an example of a set E and a point x such that E is porous at x, but not very porous there. 2
A limit point for a set S is a point x such that there is a sequence sn E S with limn-joo s„ = x.
CHAPTER 9
CREATING PATHOLOGICAL FUNCTIONS VIA C
Let / be a function. We say / has the Intermediate Value Property if for any two values a and b in the domain of / , and any y between / ( a ) and f(b), there is some c between a and b with /(c) = y. For a time, having the Intermediate Value Property was thought to be equivalent to being continuous. We will present counterexamples to this thought and use the Cantor Set to show a function with this property can have lots of discontinuities.
9.1
Sequences of Functions
Functions that illustrate a property, or more often where a nice property goes wrong, are sometimes referred to as pathological functions. For instance, continuity is a nice property. The pathological function then would be a nowhere continuous function. An example of this is the characteristic function of the rationals; the function is
11, ifx e Q, [0, ifx £ Q. The Elements of Cantor Sets -With Applications, First Edition. By Robert W. Vallin Copyright © 2013 John Wiley & Sons, Inc.
157
158
CREATING PATHOLOGICAL FUNCTIONS VIA C
Differentiability is also a nice property to have. A function which is not differentiable at any point (nowhere differentiable) is easy, just use the f above. This comes from the theorem "If f is differentiable at XQ, then f is continuous at XQ." The contrapositive then says, "If f is not continuous at XQ, then f cannot be differentiable there." The absolute value function shows that functions can be continuous at a point, but not differentiable there. So the real pathological example in this case is a function that is continuous everywhere, but nowhere differentiable. In Chapter 6 we used iterated function systems to create such a function. However, we actually did more. It showed an important technique in creating pathological functions. We could not explicitly draw the graph of such an f so instead we created a sequence offunctions which converged to our goal function. That is very typical when creating pathological functions. We create a sequence and show that the limit gives us what we want. There are, however, many meanings to the idea of "limit" and that is where we begin. We will start with a repeat of some earlier ideas on sequences of functions, the pointwise limit of a sequence offunctions. This is probably the easiest type of con vergence since it takes a problem involving functions and changes it to a problem involving real numbers. Definition 9.1 Let D be a set in the real line and fn : D —> Hfor n = 1, 2, 3 , . . . and f : D —> R. We say fn converges pointwise to f iffor every point x G D, we have liirin-Kjo fn(x) = f(x). We write this as fn —» / . fl
EXAMPLE 9.1 If D = [0,1] and fn — xn, then fn converges pointwise to the function
/(*)
For this type of convergence, we choose a point, XQ, then look at fn(xo) and find the limit as n approaches infinity. We do this for each XQ € D and we have the limit function, f. In Example 9.1,if0<x
EXAMPLE 9.2 If we let D = R and gn{x) = x ^ n + 1 , then gn —> g where (0,
g{x) = I 1/2, ll,
(fare (-1,1),
ifx = ±1, otherwise.
SEQUENCES OF FUNCTIONS
159
What can happen, and does happen in these examples, is that even though the sequence offunctions converge, they do so at different rates for different points. For example, with /io, (1/2) 10 is much closer to zero than (99/100) 10 . If the values do converge at more or less similar rates, we say that fn converges uniformly to f and define that rigorously now. Definition 9.2 Let D be a set in the real line and fn : D —> Hfor n = 1, 2, 3 , . . . and f : D —> R. Then fn converges uniformly to f (fn —>• f) iffor every e > 0 there exists an N G N such that n > N implies that for all x G D we obtain \fn{x)fl
f{x)\
<e.
EXAMPLE 9.3 Let D = [0,1] and fn(x) — x/n. We claim fn —» 0. This is seen as follows: Given any e > 0 choose N G N so that N > \je. So for any XQ G D and any n > N we have
1/nM - 0|
1
1
1
So for all the x in the [0,1], the fn(x) and f{x) values are e-close at the same time (time here means after the same index N). Note how important the domain is in this example. If D = R, the sequence still converges pointwise to the zero function, but the convergence is not uniform. Uniform convergence brings with it some important properties. For instance, it preserves the notion of continuity. If we know each fn is a continuous function, then the uniform limit must also be continuous. Theorem 9.1 Let fn : D —>ffibe a sequence of continuous functions. If {fn} converges uniformly to some function f : D —> K, then f must be a continuous function. Proof: Pick an XQ G D. We will show that f must be continuous there, and since XQ was arbitrary, this must hold for all XQ. We will use the typical e/3 argument. Let e > 0. Since fn —> f, for this e there is an N G N such that if n > N, then \fn(x) — f{x)\ < e/3 for every x £ D. We will then concentrate on f^. We know /JV is continuous at XQ, hence there exists a 5 > 0 so that for all x G D with \x — XQ\ < 5 we get \!N{X)
- fN(x0)\
< e/3.
160
CREATING PATHOLOGICAL FUNCTIONS VIA C
This implies that when \x — XQ\ < S we obtain \f{x)-f(x0)\
=
\f(x)-fN{x) x
+fN(x)-fN(x0)
<
\f( )
<
e/3 + e/3 + e/3 = e.
- /N{X)\
+ \!N{X)
+
- /WOo)
fN(x0)-f(xQ)\ + |/JV(ZO) -
/Oo)j
So we have shown that given e > 0 there exists a 5 > 0 such that for all x e D if \x — XQ\ < 5, then \f(x) — f{xo)\ < e. Hence f is continuous at XQ. 4 This fits right in with the continuous, nowhere differentiable function from Chap ter 6. Then we were using the fact that C[0,1] with the sup metric was a complete metric space. This metric is also called the uniform convergence metric/norm. The convergence of a sequence of functions using this metric is equivalent to the conver gence being uniform. Are there other properties which are held together by uniform convergence? How about differentiation? Integration? It turns out that the uniform limit of differen tiable functions does not have to be differentiable, but the uniform limit of integrable functions must be integrable. fl
EXAMPLE 9.4 Let fn be the function on [0,1] defined by l>|,
\x\>l/n,
We can see that fn converges to the absolute value function and the convergence is uniform since \fn(x) — f(x)\ < l/n. Each fn is differentiable at every point in R while f is not differentiable at zero. ■ Now we turn our attention to integrable functions and a positive result on conver gence. Theorem 9.2 If fn : [a, b] —> R is a sequence of Riemann integrable functions and suppose {fn} converges uniformly to f, then f is a Riemann integrable function and
Sa f = lim »^oo C /„. Proof: We shall use the equivalent idea of the Darboux Integral (Definition 3.24) found in Chapter 3 to show that f must be integrable. Again we have an e/3 argument. Let e > 0. Since {fn} converges uniformly to f, there is an N £ N such that \fn{x) — f(x)\ < e/(3(& — a)) for all x <E [a, b] and for all n > N. Thus I
SUP{/W(.T)}
- sup{/(a;)}| < e/(3(6 - a))
and \mf{fN(x)}-m{{f(x)}\<e/(3(b-a))
THE CANTOR FUNCTION
161
where the supremum and infimum are taken over all x 6 [a, b]. Now since JN is an integrable function, let P = {a = XQ,X\,X2, ■ ■ ■ ,Xk-i,Xk b} be a partition of [a, b] with the property that U(/N, P) — L(fN,P) < e/3. Using this partition on f, we obtain \U(f, P) - U(fN, P)\
< | sup{fN(x)} - sup{/(x)}| E K + i - *k)
Similar reasoning gives us \L(f, P) — £(/yv, P)\ < e/3. Thus \U(f,P)-L(f,P)\ <\U(f,P)-U(fN,P)
+
< \U(f,P) - U(fN,P)\ + \U(fN,P) < e/3 + £/3 + e/3 = e.
U(fN,P)-L(fN,P)+L(fN,P)-L(f,P)\ - L(fN,P)\
+
|L(/JV,P)
-
L(f,P)\
Taking the limit as e approaches zero gives us that f is an integrable function. Due to | J g\ < J \g\ and the uniform convergence of fn, we have that {J / „ } is a Cauchy sequence in K. Call the limit of this sequence L. Another application of the Triangle Inequality will show that \ J f — L\ < efor any positive e. This means
la f = l i m n^°° !l fn- * These results answer some questions about interchanging operations (limits, dif ferentiation, and integration). They say that for limits and integration, we can switch the order if we know the convergence is uniform; while for limits and differentiation, even uniform convergence is not enough to let us move things around. We will not venture into more here, but we can modify things so that differentiabil ity is preserved. We just need more conditions than just uniform convergence. This theorem is stated without proof and comes from [33]. Theorem 9.3 Let {/«} be a sequence of dijferentiable functions on the interval I with the property that f is a continuous function (such functions are called contin uously differentiable). Suppose fn converges pointwise to f on I and f'n converges uniformly to g on I. Then f is differentiable on I and f = g. 9.2
The Cantor Function
The first real construction that we will do is the Cantor Function , also known as the Devil's Staircase. It is a function F : [0,1] —» [0,1] which is increasing and onto (so by necessity, F(0) = 0 and F(l) = 1), but essentially flat. There is an explicit definition for F which we will see later. It is more instructive to look at this as an iterative process. So we will define Fn : [0,1] —> [0,1] and show that these do what they need to do. At each step Fn, will be defined on the closure
=
162
CREATING PATHOLOGICAL FUNCTIONS VIA C
of the intervals removed at the nth step of the construction of the Cantor Set along with Fn(0) = 0 and Fn(l) = 1. The rest of the function is made by joining points linearly. Stage 1: Fx{x) = 1 /2 ^ £ [ 1 / 3 , 2 / 3 ] .
'1/4, Stage2:F 2 (x) = < 1/2, ^3/4,
ifxe ifxe ifxe
[1/9,2/9] [1/3,2/3] [7/9,8/9]
7 1 - 1 jf _ J Stage n: Fn (x) = 2 72k l J ^ -Lm,k) where k represents the stage at which the contiguousinterval is removed and m refers to the position (first, second, third, etc.) of the interval removed at stage n ordered left to right on the number line.
The Cantor Function It is (somewhat) obvious that when Fn is defined on the closure of the contiguous intervals from Stage n, then the values on those intervals are determined for Fk where k > n and that each Fn is a continuous function. Less obvious is that the limit function F is a continuous function. This can be accomplished in a few ways. One is to find a way to explicitly describe^ the function F and then show that Fn —>• F. We then use the result that the uniform limit of continuous functions is a continuous function. However, the explicit formula is difficult to work with. Instead we would like to take a page from sequences of real numbers and use the idea of Cauchy sequences and the fact that (C[0,1], sup) is a complete space. Definition 9.3 Let {fn : [0,1] —► R} be a sequence offunctions. This sequence is uniformly Cauchy iffor every e > 0, there is an N £ N such that for all x £ [0,1] ;/ 'For example, let x g [0,1] have base 3 expansion x = 0.12101203 . . . and let A^ be 00 if x = 0 and otherwise N is the smallest number such that ajv = 1. Then define b„ = an/2 for n < N and fejv = 1. Finally, N ,
F(x) = E
THE CANTOR FUNCTION
163
n,m > N, we have \fn(x) - fm(x)\
< e.
Then basically the same theorem as before holds true: If {fn : [0,1] —> R.} is a uniformly Cauchy sequence of continuous functions, then fn converges to a continuous function f in C[0,1] with the supremum distance. So showing that our {Fn} is a uniformly Cauchy sequence is enough to prove that the Cantor Function F is continuous. This is usually done by showing \Fk(x)-Fk+1(x)\<2-k. The graph of F, which we shall denote by Q{F) has some additional interesting properties. Due to the symmetric nature of the contiguous intervals and of the choice of y-values on those intervals the graph is symmetric about the point (1/2,1/2). In addition, there is a self-similar nature to the graph. Despite this look to it, the graph does not have a fractional (Hausdorjf) dimension. We will not prove this, but dim.Haus{G{F)) = 1. The next result we will prove. This is the value of the length
ofG(F).
Theorem 9.4 The graph of the Cantor Function, as a subset of the plane, has length 2. Proof: At the kth step of the construction of C, there are 2h+l endpoints for the contiguous intervals, write them as {0 = X\, x2, x^,..., x2k+i = 1 } - The length of Fk, the approximation at step k, is 2k+i
Lk = Yl [^ - **-i)2 + (^fc) - ^fc-i)) 2 ] 1/2 • i=2
Now on contiguous intervals x2i+\ - x2i = (l/3) fc while Fk(x2i+1) - Fk(x2i) = 0. On the complement of those intervals, F{x2i) — Fk(x2i-i) = (l/2) fc . This means Q{Fk) consists of2k horizontal intervals of total length 1 — (2/3)fc and 2k intervals where Fk increases that have total length [(2/3) 2 + l] 1 / 2 . Then the length ofQ(F) is at least lim 1 - (2/3)fc + [(2/3)2fc + l] fc-4-OO
=2.
L
When a and b are real numbers, [a2 + b2}1/2 < \a\ + \b\. Thus the summation above has the property
E t r [te - Zi-i)2 + (Fk(Xi) - Fk{x^))2}112 2k+1
< Ei=2 Kxi - xi~l) + (Fk(Xi) ~ = {b-a) + {f{b)-f{a)) = 2.
Fk(xi-i))}
Thus the total length of the Cantor Function, which is 2 fc+i
sup ] T [{Xi - x^)2 i=2
+ (Fk(Xi) - F f c ^ . i ) ) 2 ] 1 / 2
164
CREATING PATHOLOGICAL FUNCTIONS VIA C
where the supremum is over all partitions cannot exceed 2 and our proof is done. 4ft We can also look at this function with the ideas of differentiation and integration. Firstly, the derivative of F is nearly always zero. From Calculus I, we know that if F' = 0 everywhere, then the function must be constant. Here the function is not constant, in fact it is onto [0,1], yet the derivative is zero where the graph is horizontal and that happens on all of [0,1] \ C. In measure terms, we say that F'(x) = 0 almost everywhere. At the points of C, F' does not exist. Secondly, since F is continuous on [0,1], it must be integrable there. Given the symmetry of the function, we see it divides the unit square into two equal-sized pieces and that /o F{x) dx = 1/2. This function is also used when the topic of absolutely continuous functions is brought up. This is one of many generalizations of continuity for functions. Definition 9.4 Let f : [a, b] —> R.. Then f is absolutely continuous if for each e > 0 there exists a 5 > 0 such that if {[an,bn]} is any countable collection of non-overlapping closed intervals in [a, b] with Yln°=i(bn — (in) < 8, then oo
£I/(M - / M I
<e
-
On a closed interval, most of our usual "nice " functions are absolutely contin uous; e.g. x2, sin(a;), ex. Functions that have a discontinuity are not absolutely continuous. It is not too difficult to see a function from [0,1] into R like
<1A /(*)' = {°' * \l, x> 1/2 V
is not absolutely continuous. If £ = 1/2, then for any 8 > 0 the interval [1/2 — <5/4,1/2 + 5/4} has length less than 5 yet | / ( l / 2 + 6/4) - / ( 1 / 2 - 5/4)\ > 1/2. This is a jump discontinuity. The function f above has a jump discontinuity. The function
fsin(l/x), 5(x) =
1n 10,
i/O, n x = 0
has an essential discontinuity at the origin (an essential discontinuity at XQ means lim x ^ X o f(x) does not exist). This time for any value M > 0 and any 5 > 0 there is a finite collection of closed, non-overlapping {[an, bn}} in [0, 5} with
J2\f(bn)-f(an)\>M. This brings up the topic of bounded variation.
THE CANTOR FUNCTION
165
Definition 9.5 Let f : [a, b] —> K and let P = {a = XQ, X \ , a?2, ■ ■ ■, xn-i,xn be a partition of [a, b]. Then the variation of f with respect to P is n
V(f,P) =
Y,\f(*k)-f(xk-i)\. fc=i
Finally, the variation of f on [a, b] is defined by V(/, [a, b]) = sup{V(/, P) : P is a partition of [a, 6]}. IfV(f, P) < oo, then we say f is o/bounded variation. S
EXAMPLE 9.5 The Cantor Function, F, is an example of a continuous function that is not an absolutely continuous function. ■
Intuitively, we can see this from the fact that we know F(\) — F(0) = 1, but F is constant on the contiguous intervals. Thus all the increasing goes on in the Cantor Set and that has measure zero. So continuous functions do not have to be absolutely continuous. The follow ing result tells when a continuous function will also have the absolute continuity property. A function is absolutely continuous if and only if it is continuous and has bounded variation. Now let us look at some further applications of this function. The first is in probability theory. Probability is the measure of the likelihood of the occurrence of an event. For example, if we roll a single die there are six (simple) outcomes, {1, 2, 3,4, 5, 6}. Each of these has probability 1/6, assuming the die is fair. We can summarize this in a chart. Outcome, 0
Probability, P(0)
1 2 3 4
1/6 1/6 1/6 1/6
5 6
1/6 1/6
This is an example of a probability distribution for a discrete random variable. The rules are: (1) For each outcome Oi, P{0{) > 0. (2) ^ P(Oi) = 1, where the sum is over all possible outcomes of the experiment. A random variable can also be continuous. In this instance an event is an interval of values the random variable can take, and the probability of the event is determined
= b}
166
CREATING PATHOLOGICAL FUNCTIONS VIA C
via integration. So f(x) is the continuous density function and for a < b in the domain of f the probability that a randomly chosen X falls in the interval [a, b] is given by
P(a < X < b) = [ f{x) dx. Ja
The Cantor Distribution is a probability density function f(x) with the property that the cumulative distribution function is the Cantor Function F; that is,
F(x)= f
J — oo
f(t)dt.
This is an example of a singular distribution, where all the probability is concen trated on a set of measure zero. To see another application of the Cantor Function, we go back for a moment to metrics. Many texts (including this one) have a problem in the homework to show that if d, is a metric on a set S, then so is d 1+ef The purpose of the exercise is to show how any metric space can be turned into a bounded metric space, as the distance between any two points is between 0 and 1. One can look at this in terms of composition of functions. The bounded metric is the composition f o d where f(x) = x/(l + x). This brings us to metric preserving functions. Definition 9.6 Let f be a function from [0, oo) into [0, oo). We say f is metric pre serving if for every metric space (S, d), the composition f o d is still a metric on S. fl
EXAMPLE 9.6
x is a metric preserving function. ■ 1+ x There are lots of properties and interesting examples of metric preserving func tions, including continuous, nowhere differentiable examples and examples with (ex tended) derivative infinity at No many points, but we will just stick with the following: a function is metric preserving iff(0) = 0 and f is subadditive (for all a, b G [O.oo), f(a + b) < / ( a ) + f(b)). In [24], it was shown that the Cantor Function F is subadditive, therefore it is an example of a metric preserving function. The function f(x) —
The concept of a Cantor Function can be generalized using the so-called Fat Cantor Sets. Recall that a Cantor Set can have positive measure if at each stage the proportion removed is tending toward zero. Consider 0 < A < 1. At the kth step, rather than removing an interval of length fc l/3 we now remove an interval of length A/3fc. This is a bit of notational abuse,
THE CANTOR FUNCTION
167
but easier than writing A/3fc. To clarify, when we use the Greek letter X we mean A/3fc is the proportion removed (£(Ij)/£(I), where j = 0,1) at step k, where using the English letter a means at every stage the proportion removed is a. Then the total lengths of the intervals removed in the construction of this Cx is A/3 + A/3 2 + • ■ • + 2 fc - 1 A/3 fc + • • • = A, resulting in a Cantor Set with positive measure. If we fix a A < 1, then we can create a Cantor Function, F\. Again, this function is increasing with F\(0) = 0 and Fx(l) = 1, and F\ is horizontal on the contiguous intervals. Arguments similar to before lead us to the length of F\, A + v / ( l - A ) 2 + l. Note when A = 1 we have the usual Cantor Function and length 2. On intervals of the form [0, x) we conform a measure, px, given by fix([0,x)) = Fx(x). The details of the proof of the length ofF\, which are found in [18], include the fact that Fx{y)-Fx{x)< 1 ~ 1 — A' y—x this means that (ix is an absolutely continuous function. Thus for every measurable (/jx-measurable) set E we have
px(E)=
f F x. JE
Since F^(x) = 0 on the complement of the set Cx, measurable subset of I,
/JLX(I
\CX)
= 0, and, if E is a
where m(E) is Lebesgue measure. Now if E is a measurable subset ofCx, then \i(C — A) = 1 — A and 1
and so m(E)/{\
= = < = =
Lix(I) = v\(CX) + pX(I\Cx) [i\(Cx) = fix(E) + px(Cx\E) JEF'x + m(Cx\E)/(l-X) JEFx + (m(Cx)-m(E))/(l-X) JEFx + l-m(E)/(l-X)
— A) < JE Fx. This proves that Fi(x) =
for almost every point in Cx.
l/(1-X)
168
9.3
CREATING PATHOLOGICAL FUNCTIONS VIA C
Space-Filling Curves
In 1878, Cantor showed that the cardinality of the unit interval, I = [0,1], and the unit square, I x / = [0,1] x [0,1], were the same. In modern notation, we show that these two sets are the same as follows: Obviously, any horizontal line segment in the unit square can be put into 1 — 1 and onto correspondence with [0,1] so n(I x I) > n(I). In the other direction, any point in I can be written as 0.010203 . . . where (since there is a choice) we say any representation that can be written in terminating fashion is instead written with a tail of9's. Then the function P(0.010203 . . . ) = (0.010305 . . . . 0.020406 . . . ) is an injection from I to I x I. Thus n(I) > n(I x / ) and all together n{I) =n(I
x I).
Cantor used continued fractions for actually constructing a bijection (see [34]) and was not his first try at proving this result. Cantor's original attempt was based on the decimal expansion of numbers, but Dedekind showed him that his function was, in fact, not onto. The trouble he ran into was due to some numbers having more than one expansion. The gist of his successful argument was thus: Every real number x in (0,1) can be expressed as a continued fraction 1
x= a\ H
j
, i
where the a,i are all positive integers. The two keys to Cantor's success here is that the continued fraction representation is infinite if and only if x is irrational and, when the representation is infinite, the representation is also unique. He then used a method similar to the function P defined above to show the unit square has the same cardinality as the irrational numbers in (0,1). Finally he showed there is a bijection between [0,1] \ Q and [0,1]. Almost immediately the question of continuity arose. Is it possible to construct a continuous bijection between I and I x I? The answer is no. This was proven by Eugen Netto in 1879. So, as mathematicians do, the question was changed to something similar, but with slightly different hypotheses. The requirement of being one-to-one was dropped. So now the question is, "Is there a continuous function from I onto / x I?" The answer to this is yes, with different constructions coming from mathematical luminaries like Hilbert, Peano, Sierpinski, and Lebesgue. All of these appeared in the early 1900 's. The first construction we will show is due to Hilbert. This is the most well-known of the ways to find a space-filling curve. As with many of these constructions, we create things in steps and the actual object we need is the limit of a sequence of iterations. We start with the unit square. At Stage 1, we bisect each side of the unit square to create four subsquares. Since the square is in four pieces, we divide the
SPACE-FILLING CURVES
169
unit interval into four parts. Each of these subintervals is mapped continuously to one of the subsquares in such a manner that adjacent intervals map to squares that abut (first interval to the lower left corner, second to the upper left, third to the upper right, last to the lower right) still preserving the continuity. This is illustrated via the U-shaped notation in the figure below. This mapping can be done because we know that bijections from an interval to a square exists.
Step 1 We repeat at the next step. Each subsquare is divided into four pieces, the unit interval is divided into 16 subintervals, and we piece things together as illustrated. The next two figures show steps 2 and 3. Our pattern ensures that if subinterval In is mapped to subsquare Qn, then the subsquares that Qn are divided into at the next step are the images of the subsets that In is divided into. This continues inductively.
Step 2
Step 3
So we have a sequence offunctions hn from the unit interval onto the unit square. Let h : I —> I x I be the pointwise limit of hn. Any point (2/1,2/2) G [0,1] x [0,1] is the limit of a nested sequence of subsquares Qn. The inverse images of the Qnform
170
CREATING PATHOLOGICAL FUNCTIONS VIA C
a nested sequence of subintervals in I. These subintervals will converge to a point i £ l and so h(x) = (2/1,2/2)As for continuity ofh, each hn is a continuous function onto I x I and suppose we pick x\ < xi such that \x\ — x^\ < 1/2 2 ". Then the interval [x\, X2} is contained in at most two subintervals from step n. This means \hn(x\) — hn(x2)\ < v 5 / 2 2 n , the diameter of two adjoined subsquares. Thus h is continuous. In [50] and [51], McClure took Hubert's space-filling curve and wrote it as h(x) = (X(x),Y(x)), a pair of coordinate functions. The self-similar nature of the coordinate functions were analyzed, and it was shown that for X and Y the graphs of X(x) and Y(x) d i m / / a u s ( X ) = dimBox(X)
= dimHaus(Y)
= dimBox(Y)
= 3/2.
As mentioned before, Henri Lebesgue (of Lebesgue Measure fame) also produced a space-filling curve. His construction is peripherally related to the Cantor Function (Section 9.2). Lebesgue was interested in the differentiability properties of such a curve. As is known, just because a function is continuous at a point does not ensure that it is differentiable at that point. In this instance, the 2 2 n subintervals of I that are mapped onto I x I to create gn are contiguous intervals from the 2nth step in creating C, the Cantor Set. Also, the four subsquares are not connected by a Ushaped pattern, but by an N-shaped pattern. The end result, g, is a space-filling curve that is continuous and differentiable on [0,1] \ C. Lance and Thomas, in [48], construct a new type of space-filling curve. One feature is that this does not involve any rotations of step k in the construction of step k + 1. Another is that one can see in the construction the making of a Cantor Dust in the plane, thus giving insight to the Lebesgue measure of the arcs in the construction. At each step in the recursion there is a subsquare which will be decomposed into four subsquares. Given that the original square has side length L, the subsquare will have side length s/2 ■ L, where s G (0,1]. The four subsquares are then linked together by line segments as shown:
For the recursion, each subsquare will have a copy of the graph from the previous step where the graph begins in the upper left corner and ends in the lower right, thus
BAIRE CLASS ONE FUNCTIONS
171
giving us a connected arc. We write the recursion operator as Rs (), where g refers to the graph. A pictorial example of this, where at step 0 the graph was the diagonal connecting (1,0) and (0,1) and s is set at 1/2, is
The function with domain [0,1] that describes this arc is found as follows: On the intervals [0,1/7], [2/7, 3/7], [4/7, 5/7], and [6/7,1] the function maps to the upper left, upper right, lower left, and lower right squares, respectively. The three intervals left in [0,1] are mapped to the connecting line segments to give us a continuous function. The picture above is (of course) from step 1. One continues to iterate, and these functions converge uniformly to a limit function. Because the convergence is uniform, the limit is continuous; and because the construction starts with choosing a value s, we will denote this as fs. Officially the function we have depends on both the step (n) and the width used in R at this step (s). With / s o the diagonal between (1,0) and (0,1), we then have fs,n
= R(l +
s)/2(f2s/(l+s),n-l)
and fs^n converges uniformly to fi- Due to the Cantor Dust nature of this construc tion, it can be seen that the Lebesgue measure of the graph of fs is s2. Our ultimate result here is f\, which is a space-filling curve.
9.4
Baire Class One Functions
Since the Cantor Set is in [0,1], in this section we will restrict ourselves to func tions with domain the unit interval. These can, through translations, be extended to functions with domain K. The Baire Functions are defined as a hierarchy offunctions starting with Class 0. The Baire Class 0 functions are the continuous functions. A function f is called Baire Class One (or of the first Baire Class) if it is the pointwise limit of Baire Class 0 functions, but not Baire Class 0. This means f is not continuous, but there is a sequence of continuous functions, fn, with fn —t f.
172
S
CREATING PATHOLOGICAL FUNCTIONS VIA C
EXAMPLE 9.7 The function 1,
J/X
= 1
is B\. This is the pointwise limit of fn{x) = xn.
m
We can, of course, generalize this to any finite ordinal number k. We say f is in Baire Class k if it is not in Baire Class j for j < k and there exists /„ in Baire Class k — 1 such that fn —> f. This idea actually goes even further, to the transfinite ordinals, but we will not proceed in that direction. Our notation for Baire Class k functions will be BkS
EXAMPLE 9.8 The function
JO,
ifxt®,
\l,
ifxeQ
is B2- It is the pointwise limit of the function Jl, = <
9n(X)
10,
ifx =
qi,q2,q3,...,qn,
otherwise,
th
where qi refers to the i rational number which is a B\ function for each n. The fact that XQ is not in B\ will be left as a homework exercise. ■ If you recall, in Chapter 2 we looked at the Borel Sets. These are the collection of sets generated from the open sets via union, intersection, and complement. We start with the open sets and then, via complement, the closed sets. Next, a set is called Qg if it is the countable intersection of open sets while a set is Ta if it is the countable union of closed sets. Continuing onward, we call a set which is the countable union of Gs sets will be considered to be in the collection (G$)a, which for convenience sake, we will write at G$a. Following this pattern we have Tas sets being the countable intersection of Fa sets. The classes then make up two sequences yS ? y5a i y5aS i • • • ; J~a 7 J~ad : ^aSa
, ••■ •
For a function f, one way to show that the function is continuous is to look at A\ = {x : f(x) > a} and A-i = {x : f(x) < a}. Then f is continuous iffor all a £ M., both A\ and A2 are open sets. This extends into the Baire Classification of functions on the real line. It can be shown that f £ B\ implies that both A\ and A2 are J>. At the next level, f £ B2 means that A\ and A2 are both Q$a. The rest of the hierarchy (at least for finite n) can be extrapolated by the reader. For ease in determining what functions are or are not in Baire Class One, we shall rely on the theorem below.
DARBOUX FUNCTIONS
173
Theorem 9.5 The following are equivalent. • f is in Baire Class One. • For any closed set, D, the restriction of f to D, f\r), has a point of continuity. • For any closed set P and real numbers a < b, at most one of the sets {xeP:
f{x) >b} {xeP:
f{x) < a]
is dense in P. S
EXAMPLE 9.9 Let C denote the Cantor Set. The function
(l,
ifxec,
Xc(x) = < lu,
otherwise
is in Baire Class one, while the function H(x)Ji,
ifxeC\E, lO,
otherwise
where E = {x : x is a left endpoint of a contiguous interval ofC}, is not. This can be seen by using the second part of Theorem 9.5. Given any closed set D, if D\C 7^ 0, then\c is continuous at a point in that set difference. IfD C C, then the function is constant when restricted to D. However, for H, the restriction H\D has no point of continuity since E is dense in C and C is dense in E (this is saying the sets are co-dense). ■ 9.5
Darboux Functions
In Calculus I, we have the Intermediate Value Theorem which states, "If f(x) is continuous on a closed interval [a, b] and f(a) ^ f(b), then for any C between /(a) and f(b), there exists at least one c E (a, b) such that /(c) = C." A function need not be continuous for this to be true. Let us start then, by turning the Intermediate Value Theorem into the Intermediate Value Property. Intermediate Value Property A function f has the Intermediate Value Property (IVP) iffor any two points a < b with f{a) ^ f(b) and for any value C between f(a) and f(b), there exists a c £ (a, b) such that /(c) = C. A function with the Intermediate Value Property is called a Darboux ("Darboo") function. Typically the set of Darboux functions is denoted by V.
174
CREATING PATHOLOGICAL FUNCTIONS VIA C
For awhile, it was thought that being Darboux was equivalent to being continu ous. This is not true as the following simple example shows. S
EXAMPLE 9.10 I sin 1/x, i / O , The function f (x) = < is not a continuous function, yet has I 0, x = 0 the Intermediate Value Property. ■
As we can see, this function has a special type of discontinuity. This is not a jump discontinuity, but an essential one, for every y E (liminf^^o f(x), n m s u P : c ^ o f(x)) there is a sequence xn converging to 0 such that f(xn) = y. While this is a good example, we can to better. It is straightforward to have a function with two points where it is not continuous, yet is Darboux. Just use y = f(x) + f(x — 1). This idea can be extended to any finite set of points. We can even extend this to an uncountable set of points, the Cantor Set. This is a good example because we can write it as the pointwise limit of fn and show pictures of some of the steps. fl
EXAMPLE9.il Let / i be the function which is zero everywhere except the first interval removed in the Cantor construction. On the interval [1/3, 2/3] the function is linear, join ing the points (1/3, —1) and (2/3,1). The next function fa looks like /\ except on the two intervals removed at Step 2 (1/9, 2/9) and (7/9, 8/9)) there are line segments connecting the left endpoint and y = — 1 with the right endpoint and y = 1. This process continues indefinitely.
The third iteration, fs We can actually write the limit function explicitly.
DARBOUX FUNCTIONS
175
otherwise, where (a, b) is an interval removed during the Cantor Set construction. Using Theorem 9.5 we can see this is not a Baire Class One function. The Cantor Set is a closed set, but f\c has no point of continuity. We could convert this to a Baire Class One function by making the linear parts be on the open intervals (a, b) and therefore f\c = 0, however, this function is no longer Darboux. ■ Still this function is continuous almost everywhere. Is it possible to keep the Inter mediate Value Property yet the function not be continuous? Yes, to do so just requires switching our numbers to base 2. S
EXAMPLE 9.12 Let x € [0,1] and let x be written in base 2 as x = 0.010203 ..., where a^ € {0,1} and if x has two representations we use the one that ends in a string of 0's. Now define f(x) by f(x) = limsup
~y^Qfc fc=l
This function has the property that on every nontrivial interval (a, b), f takes on every value between 0 and 1. We'll give a heuristic argument for this. For any y € [0,1] the tail of the decimal expansion is what influences the f(x) value and gives us the ability to have f(x) = y. The beginning of the decimal expansion then is what can be fixed in order to place x in the interval (a, b). ■ When we define a new collection offunctions, it is natural to ask if there is some old collection offunctions which is either equivalent or contained in this new idea. Here we show that a very familiar set offunctions is a subset ofTJ, the collection of derivatives. Be careful here, we are not saying that the differentiable functions (those f for which an / ' exists) are Darboux, but that the derivates (the / ' themselves) are inV. Theorem 9.6 Iff is a derivative, then f is the pointwise limit of continuous functions and / has the Intermediate Value Property. Proof: If f is a derivative, that means there exists a function F such that for each x in the domain of f, we have
lim n^h)-F(x) /i-s-o
= K
h
'
Since F is differentiable, then it must also be continuous. So let fn(x)=n{F(x
+
l/n)-F(x)).
176
CREATING PATHOLOGICAL FUNCTIONS VIA C
Each fn is continuous and fn —» / . In some books, this is considered proof that derivatives are in Baire Class One, but since we are saying that continuous functions are in Baire Class 0 and Baire Class 1 functions do not contain Baire Class 0, we need to be a bit careful. As for IVP, suppose that we have two values a and b and that without loss of generality we have f(a) < f(b) (if f is constant, there is nothing interesting going on as far as intermediate values go). Take a C between f(a) and f(b). The function G(x) = F(x) — Cx is continuous on [a, b] and differentiable on (a, b). Furthermore, G'(a) < 0 < G'(b). Since G is not linear (why?), it must reach an extreme value in the interior of [a, b}. So there must be a c, with a < c < b, such that G'(c) = 0. This gives us f(c) = C as needed. 4 This gives us a quick way to discount some functions as derivatives. For example,
X{0}{X)
=
fl, \0,
x = 0, x^O
cannot be the derivative of any function as it does not satisfy the IVP. So derivatives are a subset of both Baire Class One and Darboux functions. There is a lot of classical real analysis devoted to the functions in both categories. These are the Baire One, Darboux (B\D) functions. When put together, these two proper ties lead to some interesting results. First, some equivalences. Theorem 9.7 Let f : [a,b] —» ~R be a Baire One, Darboux function. This is equiva lent to ■ For each x, there exist sequences xn and yn where xn < x < yn and lim f(xn)
n—>oo
= f{x) = lim n—>oo
f(yn).
• The graph of f is connected as a subset of [a, b] x R ■ The sets A\ = {x : f(x) dense in themselves2.
> a} and A2 = {x : f(x)
< a} are (bilaterally)
Now we can proceed with an example to show that there are more than just deriva tives in B\ V. fl
EXAMPLE 9.13 Begin with the Cantor Set and {(an.b„)} the set of intervals removed in the construction. Define cn = (an + bn)/2, the midpoint of the intervals. We define
2
A set E C K is bilaterally dense in itself if for any e > 0 and any x € E, both (x,x + e)C\ E ^ % and (x — e, x) n E 7^ 0; that is, for any x in E, there are points in E that are arbitrarily close to x on both the left and right sides of x.
LINEARLY CONTINUOUS FUNCTIONS
177
f(x) to be 0 on C, arbitrary on cn, and linearly on [an, cn] and [c„, bn\. Then f £ B\D (any of the equivalencies in Theorem 9.7 will do). However, f is not continuous at x = 0 unless liniu^oo f(cn) = 0. Thus we can make sure that f is not a derivative. I Another pathological example can be found in [10]. There a Baire One, Darboux function is constructed which is zero on all but a measure zero set. A property which is true except on a measure zero set is said to be true almost everywhere, abbreviated a.e.
9.6
Linearly Continuous Functions
For this section, we move off of the real line and look at functions whose domain is in the plane with the Euclidean metric. For f : R2 —> R to be continuous at the point (xQ,yo), it means that for any e > 0 there exists a 6 > 0 such that for all (x,y) 6 M2 with d((x, y), (xo, yo)) < 6 we have \f(x, y) — f(xo, 2/o)| < £• Thus there is an open disk around (xo, yo) with radius S so that everything in the disk is mapped e—close to /(xo, yo). There is an alternate way to look at this involving sequences. Definition 9.7 Let f : R2 —> R. We say that f is continuous at (xo, yo) iffor every sequence {(xn, yn)} C R 2 that converges to (XQ, yo) we have f{xn,yn)
-> f{X0,V0)-
Since it is difficult to look at all possible sequences, this is not often used to prove continuity. Instead, this definition is mostly used to show a function is not continuous at a point. For example, the function
f(xv) = {^i" [0,
(*'»>* <0'°>' (x,y) = {0,0)
is discontinuous at (0,0) because the sequence (1/n, 1/n) converges to (0,0), but
/ ( l / n , l / n ) = l.
The function f above is the canonical example of a special type offunction, the separately continuous function. Definition 9.8 Let f : R 2 —>■ R. For a fixed value ofx0, the function fXo : R —> R defined by fXo (y) = f(x, y) is called the x-section of f at XQ. Similarly there is the y-section at yo, fyo (x). We say that f is separately continuous if each fXo and fyo is a continuous function for all Xo,yo £ ROur example f is separately continuous. The only possible trouble spot being at the origin. However, fo(x) = 02_? J2 if x ^ 0 and 0 at the origin, so trivially continuous. The same is true for fo{y).
178
CREATING PATHOLOGICAL FUNCTIONS VIA C
The generalization of this idea would be to have f\e, the restriction of f to a fixed line £, be a continuous function. This is what is known as a linearly continuous function. It is here that we will bring in our Cantor Set. That said, as with many constructions, we will start small and build our way up. We begin by defining the function g : M2 —> R by
g{x, y)
0, y/x2, x2/y, 1,
x 0 0 y
= 0 or y = 0, < x2 < y, < y < x2, = x2.
This function is continuous everywhere except the origin. At the origin is it linearly continuous (intuitively, for any line through the origin, if we look close enough the line is not involved with the parabola y = x2 and so the parabola does not interfere with the continuity at the origin). Let (a, b) be any point in the plane. We can then create a function with the dis continuity at (a, b) by using translation and reflection. Let go(x, y) be zero on the lines x = a and y = b. If a < x and y < b, go(x, y) = g(x — a,y — b). To the left ofx = a and above y — b, find go by reflecting go about the vertical line. Finally, on the rest of the square, go is the reflection of what we have so far about the horizontal line y = b. Now let {(an, 6„)}^Lj be any countable collection of points in the plane. Then each of the functions gn(x, y) = go(x — an, y — bn) is continuous everywhere except (a„, bn), where it is linearly continuous. We can then define the function G by oo
G{x,y) =
1
J2^9n{x,y). 71=1
Since each gn is bounded by 1, the Weierstrass M-Test 3 gives the convergence is uniform. We know that G(x, y) — gn(x, y) is continuous at (x, y), and this means that G is linearly continuous at (an,bn). Thus we can have a linearly continuous function with countably infinitely (even dense if we want) points of discontinuity. Can we have uncountably many such points? Yes, but for that we will need a Cantor Set. First, for a, b £ K with a < b, let (a, b) be an interval in the real line and let M denote the midpoint of that interval. Let Fo(x, y) be the function which is 0 outside of {a, b) x M., 1 on the parabola y = (x — b)2, and, on the strip bounded by [M, b), FQ(X, y) = while Fo(x, y) —
3
(x - b)2
if x is above the parabola,
y
=— if x is below the parabola and above the line y = 0. (x — b)2
Let {Mfc} be a sequence of nonnegative real numbers such that "^2 Mk < oo. If \gk(x)\ < M^ for all x 6 S, then Y2 9k converges uniformly on S.
LINEARLY CONTINUOUS FUNCTIONS
179
The values for FQ on the other three parts of the strip are found via reflection about the x-axis and x = M. This FQ is continuous everywhere in the place except at (a, 0) and (b, 0)) where it is linearly continuous. Letting (a\, b\), (02, 62), («3, bs),... be the intervals removed during the con struction of the Cantor Set, C. Let Fn be defined as above with respect to the interval (an,bn). The function „,
,
\Fn{x,y), I 0,
x £ (an,bn), otherwise
is continuous everywhere except C x 0, an uncountable set. This can be extended to a set dense in the plane, via C x q where q G Q. See [79] for the details.
180
CREATING PATHOLOGICAL FUNCTIONS VIA C
EXERCISES 9.1 Draw the graph of Fi{x), i = 1, 2,3, where F^ refers to the functions converge to the Cantor Function. 9.2
Show why fn = xn does not converge uniformly on the domain [0,1].
9.3 Prove that the sequence fn(x) converges uniformly if D = [0,1].
— x/n convergespointwise
9.4 Explain how if we know \Fk(x) — Fk+i(x)\ {Fk} is a uniformly Cauchy sequence. 9.5
which
if D = [0, oo), yet
< 2~k, then we can claim that
Provide the proof for Example 9.8 showing that \Q is in Baire Class 2.
9.6 Let D he a finite set in [0,1]. Show that XD(X) is Baire One. (Proof by picture is not rigorously legitimate, but it is okay here.) 9.7 Prove that \c is a Baire One function (thus showing that Baire One functions can have uncountably many points of discontinuity). 9.8
Divide the Cantor Set into two pieces: A = {x : x is an endpoint of a contiguous
interval}
and B =
C\A.
Show that for any a £ A and any £ > 0 there is ab £ B such that \a — b\ < e. Then show this is true if the roles of A and B are reversed. This means that A is dense in B and B is dense in A (that is, they are co-dense in each other). 9.9
Divide the Cantor Set into two pieces: A = {x : x is an endpoint of a contiguous
interval]
and B =
C\A.
Prove that XA is not a Baire One function. 9.10 Draw a graph of the Cantor Function F on the contiguous intervals ( 1 / 3 , 2/3), ( 1 / 9 , 2 / 9 ) , and (7/9, 8/9). 9.11
Explain why F has the Intermediate Value Property.
9.12
Prove that F is not B\.
CHAPTER 10
GENERALIZATIONS AND APPLICATIONS
There are many ways to tweak a Cantor Set. The literature is rife with papers on these ways and their consequences. Many of these results are intriguing, but highly technical. Hence in this chapter we present several short, and mostly "proof-less" sections. These are ideas which may serve to whet the appetite for further study by the reader. The author strongly encourages such studies, whether in class or out.
10.1 Generalizing Cantor Sets Our first generalization of the Cantor Ternary Set was seen in Chapter 3, where we dropped the "ternary" requirement. This construction begins with a fixed number r, 0 < r < 1 and at each step in this process we have an interval I and from that interval we remove a middle subinterval J so that £(J)/£(I) = r. This left two subintervals behind. We wrote this Cantor Set by Cr. This next generalization appears in [14]. We still take an interval and divide it into three pieces, keeping the outer two, but this time the ratios of the kept pieces do not have to be the same. Specifically, each closed interval I is partitioned into 1$, O, The Elements of Cantor Sets -With Applications, First Edition. By Robert W. Vallin Copyright © 2013 John Wiley & Sons, Inc.
181
182
GENERALIZATIONS AND APPLICATIONS
and 11, where I and IQ share a left endpoint, I and I\ share a right endpoint, and
with r\ + r2 < 1. Then K0
=
[0,1],
Kx
=
[0,ri]U[l-r2,l],
K2
=
[0,rf]u[r1(l-r2),r1]U[l-r2,l-r2(l-r2)]U[l-r1r2,l].
As before, each Krl is the union of2n disjoint closed intervals. We then set Cri t,.2 = n^L0-K„. This is again a perfect, uncountable, nowhere dense set in the unit interval. The most obvious question, as we had in the beginning, is, "How much is taken away in this method?" Well, informing K\ we remove from KQ an open interval of some length, call it r. Making E2 we remove two open intervals, one of length r\T, the other of length r2r. Continuing, at Step n+l we remove from the set En exactly 2" open intervals: one has length T\T, then n open intervals of length r™~lr2r, next n(n — l)/2 of length r\r2x^2T, etc., ending with one interval of length r2r. The total of these sum is ^J'
j=0
This is at Stage n. Thus the total length removed becomes oo
^
l-(ri+r2)
as a convergent geometric series. This set satisfies the Open Set Condition, and therefore the Hausdorff dimension can be found easily as the solution to r« + r«
= i.
We can use a similar idea to find the Besicovitch—Taylor dimension ofCri
f
\
/ \ ■ i 3=0 \J<
T
r
i
r2
=
rrj
where ri = r\ + r^. Thus sum of all the terms is r' 3 (1 + 77 + rj2 + r/3 + • • •), which simplifies to ?7< 1, ■q > 1.
GENERALIZING CANTOR SETS
183
This idea was further generalized by Baekfor what he referred to as Perturbed Cantor Sets . Let {an} and {bn} be real sequences where 0 < an,bn < 1/2. As always, let I® = [0,1]. At step 1 we take away an open subinterval, leaving behind the closed subintervals Too and loi where
KIo)=a^ndKLj-h and define E\ — loo U IQ\. At stage k we have a set Ek that consists of the union of sets Ia, where a is a string ofk 0's and 1 's (we will also say this as a s {0, l} fc and the length of a, \a\, is k). For stage k + 1, each Ia is divided into Iafl U O U I„tl where O is an open interval and
—
=ak+1and1(I-y=bk+l.
Then Ek+1 = U\a\=k+\I
For our results here, we make the assumption that an,K,dn
= 1 - (an + bn) > a > 0.
In [3] and [4], these sets were investigated in terms of (Hausdorff) measure and dimensions. In order to present these results, we introduce two new functions: n s
fc (C0„,fcn) = l i m i n f T T K + &£) = liminf V fc=l
*\I°),
\a\=n
and n
Qs{Can,bn) = limsup Y[(a% + bsk) = limsup V n
^°°
fc=l
t(Ia).
\a\=n
With these we can then define the upper and lower Cantor dimension for C 0 i i j n by dim
cSS5F( C '«»,6n) = S U P { S > ° : QS(Can,bn) = oo}
and dimcaritor(Can,bn)
= SUp{s > 0 : hs (Can,bJ
= Oo}.
Now we can present our first result. Lemma 10.1 Let {sn} be the sequence defined by solutions to a^1 + 6^" = 1. Then 0 < liminf sn < d\ra.cantor(Can,bn)
< ^mcantor(Ca„,bn)
< lim sup s„ < 1.
184
GENERALIZATIONS AND APPLICATIONS
Proof: Since {an} and {&„} are bounded away from 0 we know that inf sn < sups„ < 1. Now choose s and s' such that 0 < s < s' < lim inf sn. Then there exists an N such that for k > N, Sk > s' > s. We can choose a number j5 where 1 > j3 > a^, bkfor all k and this gives us n
n
n
k=N
k=N
k= N
This gives us liminf'n^oo rifc=/v(afc + ^fc) = °°- Thus hs(Caii_t,„) = oo ifO < s < liininfs n . This means lim inf ,sn < dinic a „ t o r (C a ? i _{,,„). In a similar manner, we can prove qs{Canbn) = 0 for t > lim sup sn and arrive at dim^r^-^:(C an :f,?i) < lim sup sn . 4 Recall mass distributions from Chapter 7. We can apply this concept to our con struction for Can ,6„ and get our results in measure and dimension. For the dimension result we see that Cantor dimensions are just a new way of looking at Hausdorff and packing dimensions. These are the following two theorems. Theorem 10.1 dinica„tor(Car,,b,:)
(Ca„,6 n ).
and Theorem 10.2 dim Contor ((?„„, b J = dimp(C a , lj6n ). In terms of measure, [4] shows the equivalence of hs(Canib„) andT-l{C'a„.b„) via two results which we combing here into one theorem. Theorem 10.3 Ifhs{Can^rl) = oo, then'H{Can,bn) ~ oo and if 0 < hs(Ca„,b„) < oo, then 0 < "H(Ca?Mbn) < oo. Now we move on to more than one subinterval, all with the same size. Let m, be an integer greater than 2 and fix r* so that 0 < r* < 1/m. The change to the steps in the construction is that at each stage k, the closed interval I is replaced by m equally spaced subintervals of length r*\I\, the ends of the far left and far right subintervals coinciding with the left and right endpoints of I. It is an exercise in arithmetic and geometric series to show that with fixed m and r* the set will be measure zero. As far as dimension goes, in [30] it is shown that the resulting set F has I n 777,
dim W a u s (F) = dmiBox(F) = — — lnr* and
0
Below is the picture when m = 3.
FAT CANTOR SETS
10.2
185
Fat Cantor Sets
In the homework in Chapter 3 we saw one example of a Cantor Set that had positive measure. We looked at another type offat construction in Chapter 9. This construc tion was used in both [8] and [18]. Below is an interesting result from the former paper, but first we repeat a construction we have for this type of set. Pick A G (0,1]. At stage k of the Cantor Set construction, rather than removing 2k subintervals of length 3~k from the middle, we remove middle subintervals of length A/3fe. Then the total length removed is A/3 + 2A/3 2 + • ■ • + 2 fc - 1 A/3 fe + • • • = A. We will denote such a set as C\ (A bit ofnotational abuse, but easier than writing C\/3k. When we use the Greek letter A we mean A/3fc is the proportion removed at step k, where using the English letter a means that at every stage the proportion removed is a.). When A < 1 the set C\ must have positive measure. The question asked and answered in [8] was about the type of numbers that are captured by these Cantor Sets. Define the set [x] as [x] = {A e [0,1] : x e Cx}. It turns out that these sets are small in terms of category. Theorem 10.4 For each x, the set [x] is a closed, nowhere dense subset of[0,1]. Proof: For any point x G C\, we can think ofx as the limit of a nested sequence of closed intervals. The intervals come from keeping either a left or right subinterval at each step. For X fixed there is a one-to-one correspondence between the point x in C\ and the elements S G <S, where S is the set of subsets of the natural numbers. This is done by saying for x G C\ the positive integers n is in Sx if x was in a right subinterval at Step n. When A G (0,1] we use <j)\ to be the mapping taking Sx to x. When A = 0 (then Co = [0,1]) there can be more than one subset ofN which corresponds to an x. For instance, the point x = 1/2 corresponds to {1} and {2, 3 , 4 , . . . } . Still we can define
186
GENERALIZATIONS AND APPLICATIONS
SoifliS^%, then r](S) < /i(S) while i f l € S / N , then p(S) < r]{S). After Step 1 of the construction, we have two intervals of length I\ = 2 _ 1 (1 — A/3); after Step 2, we have four intervals of length I<± = 2 _ 1 (I\ — A/3 2 ). In general, Step n has 2n intervals of length In
=
2-1(/n-i-A/3")
= =
2~ n (l - A/3)[l + (2/3) + ■ • • + (2/3)"- 1 ]) 2 - n ( l - A ) + 3-"A.
Now ifnG S € S and ifx&C\ xn+1 -xn=In
is the point that corresponds to S, then + A/3 = 2-™(l - A) + 2 • 3 " n • A,
where xn denotes the left endpoint of the interval that contains x at the nth stage. So x = 4>\(S), where
sup{|0 Q (S)-<MS)|} = |a-/9|/6 the set [x] is a closed subset of [0,1]. Define a linear ordering < on S as follows. We say E < F if there exists a positive integer n such that En^\ = Fn-\ and En C Fn, where Hk = H D {0,1, 2 , . . . , k}, H G S, and k > 0. We proceed to use this to show that [x] is nowhere dense in [0,1]. Suppose a, (3 € (0,1) and 4>a(E) = x =
10.3
Sums of Cantor Sets
As we saw earlier, in Chapter 8, a very interesting result occurs when adding together two Cantor Ternary Sets; namely C + C = {x + y : x,y € C} = [0,2]. One immediate consequence is that a set which is small in terms of measure and category can sum with itself to be large in both of those terms. In this section we
SUMS OF CANTOR SETS
187
look at some applications and generalizations that come from this addition. These results come from sources such as [38], [45], [54], [56] and [57]. For our first investigation we switch up our point of view. Normally we use Cr to denote a Cantor Set where the proportion removed from the middle of an interval is r. Associated with this r is a value X, the relative length of each subinterval retained. So A = (1 — r)/2. Then for 0 < A < 1 we have a Cantor Set C\ defined by Cx
e[0,l]:x
{l-\)Yiai\i\,
=
where aj £ {0,1}. We want to consider sums of these Cantor Sets, C 7 + C\, in terms of Hausdorff dimension and Lebesgue measure. As we have seen before, dim HausV-^X (CA and so if
In 2 ln(l/7)
In 2
RVA): In 2
+ ln(l/A)
<1,
then d\mHaus(C1
+ C\) < dim/f a u s (C 7 ) + dimHaus(C\)
< 1.
This implies that p(Cj + C\) < 1, and thus the sum of these Cantor Sets cannot have interior. In the other direction, we use the Gap Lemma from [60]. This tells us that if 1
I-27
A >1, 1 - 2A
then C1 + C\ must contain an interval. However, these two inequalities do not cover all of the square [0,1/2] x [0,1/2] (the region between the two graphs below). The sum of Cantor Sets with subscripts from this region is unknown as far as interior is concerned. For most of the points in the region, things do work out nicely where most is in terms of measure and nicely in nonzero measure. This result is from [70].
188
GENERALIZATIONS AND APPLICATIONS
Theorem 10.5 For any 7 with 0 < 7 < 1, there exists a set E7 C (0,1/2) with /x(£' 7 ) = 0 such that ifX^E^ and In 2
M1J1)
+
In 2
RVA)
>
*'
//zen p,(C1 + C\) > 0. In [13], the authors are answering questions concerning nonmeasurable sets. As intuitively stated in their paper, The Cantor middle-third set, when added to itself, gives an entire interval, [0,2]. So certainly there exists a measure zero set that when added to itself gives a nonmeasurable set. These results use the concepts of Bernstein Sets and Transfinite Induction . So there is some background information we must go over before we can proceed to results. A total order on a set S is a binary relation (usually written <) on S with the properties of totality (for all a,b G S either a < b or b < a), antiy symmetry (a < b and b < a implies a = bfor all a.b G S), and transitivity (for all a,b,c G S, if a < b and b < c, then a < c). The set S is well-ordered if it has a total ordering on it and every non-empty subset of S has a least element in this ordering. The set Z + with < is a well-ordered set while IR+ with the same relation is not; (0,1) is a non-empty subset ofW+ with no least element. Ordinal numbers are the sizes one can have for a well-ordered set. These include: the nonnegative integers { 0 , 1 , 2, 3 , 4 , . . . } ; the value to, the first transfinite ordinal and the size of the whole numbers; w + 1 (we will not go into the details here, but U> + 1>1+UJ
fl
= u>), LU + 2, ..., bj\, u>2, ■■■ y OJ^; and more.
EXAMPLE 10.1 The natural numbers N in their usual ordering is a well-ordered set of type to with the property that every number except 1 has a predecessor. Another well-ordering of the natural numbers is 1,3,5, 7 , 9 . . . 2 , 4 , 6 , 8 , . . . . Here, the order type is LO+UJ and neither the numbers 1 nor 2 have a predecessor. Thus order types are not unique. ■
A
EXAMPLE 10.2 In the plane, M2, we usually do not think of ordering elements, comparing ( — 1,2) with ( 2 , - 1 ) . However, working in full generality, there is a way to take ordered pairs (a, b) from A x B and put a total order on things, assuming
SUMS OF CANTOR SETS
189
A and B are themselves totally ordered sets. This is the lexicographic order. We say that (a, b) < (a*,b*) if a < a* or (a = a* and b < b*).
Transfinite Induction is an extension of mathematical induction to sets that can be well-ordered. These sets can be beyond countably infinite, in which case a proof by Transfinite Induction usually has three parts: • The Zero Case: P(0) is true. • The Successor Case: Let a + 1 be the successor for a. We must show that if P(a) is true (or possibly P{0) is true for all ft < a), then P(j3) is true. • The Limit Case: If X is a limit ordinal (a limit ordinal has no predecessor; that is, there is no a such that a + 1 = A), then P(X) follows from P{0), /3 < X. fl
EXAMPLE 10.3 For an example of applying Transfinite Induction we present a result due to Sierpinski: There exists a set X in M? such that for any line £, £ n X contains exactly two points. Start by well-ordering the set of all lines in the plane (there are c, the size of the continuum, of these) in such a way that each line has fewer than c prede cessors. Let £a be the line associated with the ordinal number a. To create X, assume that for each (3 < awe have a set Xp with the property that iff < j3 the line £j intersects the set Xp in exactly two points and no line £ intersects Xp in more than two points. Additionally, we insist that (1) iff < /3, then X1 C Xp, and (2) each Xp has less than c many points. Define Ya by Ya =
Up
Then Ya contains less than continuum many points and ya^£ contains at most 2 elements for any line and is exactly two for every j3 < a. If the cardinality ofYa n £a = 2, then let Xa = Ya. If not, since there are less than c points in Ya n £p where /3 < a and continuum many points on a line, we can find a point on £a which is not a point on £p, (3 < a. Adding one or two (however many is necessary) points to Ya will then create our Xa. Either way stage a is complete. We now have a set Xa defined and let X = \JaXa. ■ As stated, we also need the concept of a Bernstein Set. A Bernstein Set is a set B CM. with the property that for any uncountable closed set, S, in the real line both BnS and Bc n S are non-empty. It is not obvious that Bernstein Sets must exist, so we shall go through the construction. In order to do so, we need two lemmas, which we present without proof.
190
GENERALIZATIONS AND APPLICATIONS
Lemma 10.2 Any uncountable closed set in K has cardinality continuum. Lemma 10.3 There are continuum many uncountable closed subsets in K. Now on to the construction. Begin by well-ordering the real line. Call this order ing -<. If we use c denote the cardinality of the continuum, via Lemma 10.3 we can order all the uncountable closed subsets ofM. via ordinals smaller than continuum, {Fa : a < c}. By Lemma 10.2 each Fa has cardinality c. Using -< choose the first two elements in FQ, call them XQ and yo- Let X\ and yi be the first two elements of F\ (again using -<) that are different from XQ and yo. Now suppose a < c and that for each /3
-Up
is non-empty since Fa has cardinality c while the cardinality of\Jp
G D x D x Pa x Pa : a < c}
such that for each a < c ca=aa
+ ba and {dp : (3 < a} n {Aa+1 + Aa+i)
=0
SUMS OF CANTOR SETS
191
where Aa = Up . This is the set o/subsums of the series. Also, let oc
rn = 5Z
afc
'
k=n+l
the nth tail of the series. The following facts have been produced several times. The first time seems to be in 1914 [41]. • E is a perfect set. ■ E is a finite union of closed intervals if and only if an < rnfor n sufficiently large. E is an interval if and only ifan < rnfor all n. ,
Ifan>
rnfor n sufficiently large, then E is homeomorphic to the Cantor Set.
Now let us look at the sum of copies ofE. Using the notation from [57], we let ®mS = {si + s2 + s 3 H
h sm : Si e S}.
This is the algebraic sum of S with itself m times. Applying this type of sum to the set E, we have the following theorem. Theorem 10.7 There is a positive integer rnfor which (BmE is a finite union of intervals if and only if limsup — < oo. rn Moreover, the smallest positive integer rnfor which ®mE is a finite union of intervals is the smallest integer m such that an/rn < rnfor all but a finite number of integers n.
192
GENERALIZATIONS AND APPLICATIONS
Proof: Let m be a fixed positive integer. We construct a new sequence {c„} by c
= c(q-l)m+2
(q-l)m+l
= ■ • • = Cqm — aq
for q = 1,2,3,.... We can easily see that the ^ cn is a convergent series with 0 < c n + i < cnfor all n, and ®mE is also the set of sub sums of"^2 cn. Let oo k=n+l th
c
be the n tail of Y2 n- Then by the second bullet item above, (BmE is a finite union of intervals if and only if cn < rnfor all but finitely many index values. This inequality is clearly true if rn ^ nq, for some q. If n = mq, then cn = aq and fn — mrq. Thus ®mE is an interval if and only if aq < mrq for all but a finite number of integer values of q. The conclusion of the theorem now follows, tfk With only small changes to the proof we can state this next corollary. Corollary 10.7.1 The smallest integer m such that an/rn < m for all n is the small est integer for which ®mE is an interval. The Cantor Ternary set is the set of subsums ofY2 &n, where an = 2/3 n . Then an
rn
2/3"
ZZn+i
2/3fc
•
Thus, again, C + C = (B2C is an interval. Again, let E be the set of subsums of^an. This time we wish to look at sets of the form aE + bE. Letting s = Y2 an> the following are easy to see.
• E=
s-E,
• aE + bE = aE + (—b)E + bsfor any real numbers a and b, • aE + bE = a(E + (b/a)E) for any real numbers a, b with a / 0. The result of these facts is that we can transform aE + bE in such a way that we only really need to study the set E + xE, where 0 < x < 1. Now we come to a new theorem. Theorem 10.8 Assume an+i/an
< x < I for all n. Then
1. E + xE is an interval if and only if (a„ - r n ) / r n _ i < x < rn/(an for all nfor which an > rn.
- rn)
DIFFERENCES OF CANTOR SETS
193
2. E + xE is homeomorphic to the Cantor Set if {an ~ r „ ) / r „ _ i < x < rn/(an
- rn)
for n sufficiently large. Proof: Construct a new sequence {cn} for which C2n-i = an and C2n = xanfor n = 1,2,3,... and note that cn+\ < cn. Then E + xE is the set of sub sums of the series ^2 cn- Again let DO
k=n+l th
the n tail of ^2 cn. As before, E + xE is an interval if and only if cn < fn for all n. Now r^n = (1 + x)rn and r
Differences of Cantor Sets
The difference between two Cantor Sets concerns itself with sets of the form Cr, the so-called middle Cantor Set where in the construction each closed interval, I, is divided into three subintervals and the middle open interval, whose proportional length isr-£(I), is removed. Of use here is the value s, defined by s = (1 —r)/2. This is the proportional length of the intervals kept and allows us to write the construction fo Cr as the invariant of the two contractions T\(x) = sx andTi{x) = sx + (1 — s). Note here that TQ(0) = 0 and T\(1) = 1, showing that the interval removed is, in fact, in the middle. Now given two values a and b, the Cantor Sets Ca and C\, are both measure zero, both uncountable, and both nowhere dense. To separate these, we look at their set differences with themselves Ca-Ca
= {x-y
:x,y e Ca).
A second way to think of this is via translation. For a fixed t, Ca + t = {x + t : x € Co]. Then we can write the difference as Ca-Ca
=
{t:Can(Ca+t)^(D}.
1 94
GENERALIZATIONS AND APPLICATIONS
The first result, straightforward and not needing proof, isthatCa — Ca C [—1,1]. The rest of the results in this section follow the ideas that the "more" that is in a set, the more often that set should intersect itself; and ifa>b, then Ca has less to it than Cb. If the idea of less/more is measure, then we get this theorem which does a small bit of segregating for r values. Theorem 10.9 If s < 1/3, then Cr — Cr is a Cantor Set with measure zero. If s > 1/3, then Cr - Cr = [-1,1]. The next result shows that under the correct circumstances we know exactly how many points of intersection there are. This result holds only for some values of s. Theorem 10.10 For any s £ (0, y/2 - 1) and any natural number n, let
<—"(£)■ Then the intersection ofCr and Cr + tn will contain exactly 2™ points. There are intersections that are singletons. To describe those, we define the set A r = 1Cr — 1, the set difference of Cr and 1 — Cr. We then arrive at a lemma on intersections. Lemma 10.4 Choose s 6 (0,1/2). IfCr fl (Cr + t) is a single point, then t G A r . Continuing, let us define a set A^ as the union of all the images of A r under the transformation n-l
T(x) = snx + ^vls\l
- s),
i=0
where n is any positive integer and V{ is either — 1, 0, or 1. Some things to point out:
■ Afl c [-1,1], ■ Aft has measure zero, and • AR contains all of the points ofCr — Cr for which Cr n (Cr + t) is a finite set of points. Now that we know something about the intersection being finite, we turn to the question of how often that happens. When s is large enough, we would think that the Cantor Sets are big enough for the intersection to not be so sparse. That turns out to be correct. Theorem 10.11 If s > 1/3, then CT D (Cr + t) is a Cantor Set for almost all t 6
[-1.1].
PRODUCTS OF CANTOR SETS
195
We close this section with a result that is not for almost all t, but for all t. It looks at when the difference contains a Cantor Set. Theorem 10.12 Ifs e (V% - 1,1/2). then Ca n (C 0 +1) contains a Cantor Set for allt€ (-1,1). There are more results of this type to be found, using thickness from Chapter 8. See [46]. 10.5
Products of Cantor Sets
Let A and B be sets in Rn and Rm, respectively. The Cartesian product of A and B is the set of ordered pairs for which the first coordinate is an element of A and the second coordinate an element of B. Formally we have, Definition 10.1 Let A C K" and B C M.m. The Cartesian product of A and B is the set given by Ax B = {(a,b) :aeA,beB}. The set Ax B is a subset ofM.n+m. All of these ideas below can be further generalized from R 2 to higher dimensional spaces, but we will stay in the plane. S
EXAMPLE 10.4 Let A = [0,1] C R, the unit interval. Then A x 1/2 is a horizontal line and ■ A x A = {(x,y) : x,y G [0,1]} which is the unit square in the plane.
\M EXAMPLE 10.5 Let C denote the Cantor Set in the line. Then C x C is called the Cantor Dust. We will denote this set in M2 by D. A picture of the construction of D is below. ■
1 96
GENERALIZATIONS AND APPLICATIONS
Let us look at some properties of products and, by doing so, properties of the Cantor Dust. Theorem 10.13 Let A J c l a measure zero set in M2.
both with measure zero. Then their product Ax
B is
Proof: First we note that if [a, b] and [c, d] are intervals in the real line, then [a. b] x [c, d] is a rectangle in K 2 and the area of the rectangle is the value of the (outer) measure of the rectangle; m*([a, b] x [c, d}) = (b — a) ■ (d — c). Now since A and B are measure zero sets, given e > 0 there exists finite collections of open sets {Ui} and < \fe. {Vj} in the line such that A C UUi, B C UVj and J2m*(Ui),Jl'm*(Vj) Then Ax
B c \J(Ui
x Vj)
and
m*{A x B) < J2m*(Ui x Vj) < J2rn*(ui) -J2m*(Vj) Since e is arbitrary, this shows the outer measure of Ax measurable with (Lebesgue) measure zero. 4k
< £-
B is zero and such sets are
Corollary 10.13.1 The Cantor Dust has measure zero. The result above does generalize to measurable sets with positive measure: The Cartesian product of two measurable sets is a measurable set and the measure of the product is equal to the product of the individual measures. We now turn our attention to the various notions of dimension. Many of these results can be found in [30] and are presented without proof. For the first example above, when the product of the unit intervals is the unit square, it is obvious that d i m H o n a ( [ 0 , 1 ] x [0,1]) = dim/, o ,„ s ([0,1]) + dim// a ,, i s ([0, l]); however, this is not the usual result. The best we tend to get is a lower bound. Theorem 10.14 Let A c t " and B C Wn. d\mHaus(A
Then
x B) > dimHaus(A)
+
dimHaus(B).
In his book, Falconer shows a construction of two sets E and F, both with Hausdorff dimension zero, while d\mHaus{E
x F) > 1.
In order to get an upper-bound result, we must replace one of the Hausdorff di mensions in the right side with Box-Counting Dimension.
CANTOR TARGET
197
Theorem 10.15 For any sets A C l " and B C WLm dimHaus(A
x B) < d\mHaus(A)
+ dimfcox(jB).
This leads to the following corollary: C R m , ifdimHaus{B) Corollary 10.15.1 For any sets AcRnandB then dmvHaus{A x B) = dmiHaus(A) +dimHaus(B). S
=
d\mbox{B),
EXAMPLE 10.6 For any two Cantor Sets, Cr and Cs, we have dimHaus(Cr
x Cs) = dimHaus(Cr)
+ dim f f a u s (C s ).
Specifically, for the Cantor Dust, D, we obtain dim H<XUS{D) = 2 dim# a u s (C) = £ | = 1.261.... . If we use just (Upper) Box-Counting Dimension, then the inequality we get is a reverse of the one we have for Hausdorff dimension. Theorem 10.16 For any two sets AcRn dimbox(A
andB C Mm
x B ) < dimbox(A)
+
dimbox(B).
Proof: It is straightforward to see that if A can be covered by Ns (A) sets of diameter d and B can be covered by N$(B) sets of diameter S, then Ax B can be covered by the NS(A)Ns(B) "squares" formed by the product of these sets. Using the laws for logarithms and some simple algebra we arrive at \n(NS(AxB)) , ]n(NS(A)) ,. hmsup —^ ^-^r—JJ- < hmsup , ; " + limsup ,5^0+ -ln(5) ^0+ -ln(S) s^0+
\n(N5(B)) , -ln(5)
.\/;
which is what we need. 4 10.6
Cantor Target
A slightly different way to take the Cantor Set from M. to R 2 is via the Cantor Target. This is created by rotating the Cantor Set about the origin. Specifically, we can write the target, T, in polar coordinates via T = {(r,0)
:reC,0<e<2Tr}.
This set is used in the investigation of electromagnetic waves through a self-similar aperture. The one result we shall show is determining its Hausdorff dimension.
198
GENERALIZATIONS AND APPLICATIONS
Theorem 10.17 The Cantor Target, T, has Hausdorff dimension given by d\mHaus{T)
= 1 + ——. In 3
Proof: Let f : R 2 —>■ R 2 be the function f{x, y) = (x cos j/, y cos x). This mapping has the property \f(x,y)-f(x',y')\<\(x,y)-(x',y')\ where distance in R 2 is measured by the Euclidean metric. This property d{g(a),g(b))
= =
dim f f a u s / ( C x [0, 2n}) < dimHaus(C x [0, 2TT]) dimHaus C+ dimHaus[0, 27r] = j ^ | + 1.
In the other direction, if f is restricted to the domain [|, 1] x [0, n], then f~x is also a Lipschitz function. NowT is a subset o / / ( C n [ | , l ] x [0, 7r]) and so dimHaus(T)
> = =
10.7
d i m / / a u s / ( ( C n [§, 1]) x [0, TT}) dimifous((Cn[§)l])x[0,7r]) dimHaus(C
n [|, 1]) + dim ffaus ([0, TT]) = £ f + 1. 4
Ana Sets
Recall from Chapter 8 the set difference definition of a Cantor Set:
C = I\\JOi, i>i
where I is a finite closed interval and {Od\i > 1} is a (finite or infinite) collection of disjoint open intervals contained in I. In this section, we introduce a sequence
ANA SETS
199
and show how these sequences can be used to create sets that fit this definition of a Cantor Set. We begin with what we call an Ana Set and then proceed to the Golden Ana Set. The Ana Set is based on a verbal sequence. We begin with the letter "a" in stage 0. In stage 1, we replace the "a " with "ana ". In the subsequent stages every "a" is replaces by "ana" and every "n" is replaced by "ann". This is really just pronouncing each letter as "An a" or "An n". Thefirstfour stages are a ana ana ann ana ana ann ana ana ann ann ana ann ana Pictorially, this can be illustrated by starting with the closed unit interval. At stage k the unit interval is divided into 3fe equal-sized subintervals (corresponding to the number of letters at that stage). The intervals corresponding to an a are kept, and the open intervals corresponding to an n are removed. Ai= A2 = ^ 3 =
A4= Ana Sets (and then Golden Ana Sets) were introduced in [63]. In the article by Pe ([62]), he answers the question, "At stage k, how many a's and n's are there?" There are (3 fc " 1 + l ) / 2 a's and (S^1 - l ) / 2 n's. The Golden Ana Set is created by bringing up the following misconception: Since n is a consonant, then perhaps it should be said as "A n". Thus the Golden Ana sequence starts with "a" and there after every "a" is replaced with "ana" while every "n" is replaced by "a n". Thefirstfour stages here are then a ana ana an ana ana an ana ana an ana an ana He then shows that at stage k the number of a's is a(k) = F(2k ~ 1) and the number of n's is n(k) = F(2k — 2) where F(k) is the kth Fibonacci number. We then have a(k) _ F ( 2 f c - 1 ) , _ l + \/5 njkj ~ F(2k-2) ~*^~ 2 ' the Golden Ratio, hence the name of the set. For the Golden Ana set thefirstfour iterates look like
200
GENERALIZATIONS AND APPLICATIONS
G3=. Cx4 =
Corresponding to each of these constructions, there are three sets which we can then define. These are • The intersection of the Ak,
A = nf=lAk • The limsup of the sets: A = limsupk^,^ Ak = n j ^ (U^ = f c A m ), which is the set of points in infinitely many of the Ak• The liminf of the sets: A = lim inffe^oo Ak = U^L1 (C\™=kAm), the set of points in all but finitely many of the Ak• And the corresponding G = C\Gk, G = lira sup Gk, and G_ = liminf GkThe sets A and G are of the form [0,1] \ (uOj), hence Cantor Sets. The set A is a familiar set, it is the Cantor Ternary set, C. The limsup and liminf are different. Looking at the pictures we can see that later stages reintroduce points that were take out previously; e.g. there will be points in (1/3,2/3) despite this being the first interval removed. In [40], Jones was able to characterize some of the more interesting numbers that are in G. His main result was Theorem 10.18 Let Fj be the j t h Fibonacci number, Lj the j t h Lucas numbers, and 4> represents the Golden Ratio. Then the following numbers are all in the Golden Ana Set: L ■ ' i
-L-, + 1
3+1
£i>
^j+1
1
^
£23 A f p M 2j + l '
Since the Ana Set and the Golden Ana Set are both Cantor Sets, it is natural to want to characterize them in terms of some of the "usual" ideas about Cantor Sets. This was done in [72]. We start with the Golden Ana Set, since the Ana Set coincides with the Cantor Set. Theorem 10.19 The Golden Ana Set is a porous set in R. As with most applications of porosity, the result gives an immediate corollary. Corollary 10.19.1 The Golden Ana Set is both measure zero and nowhere dense. Proof: This is a consequence of the following: For any k 2 <
_ ^
< 3
•f'2fc-2
AVERAGE DISTANCE
201
so each interval I is divided into three pieces and the gap created is at least the same size as the sub-intervals on either side which came from I. That is, given e > 0 there is an m such that 3/F2 m < e < l/-p2m-2 and this means that in the interval (x,x + 3/F 2 m ) there is a gap of length \jFim which is in the complement of G. Thus we get our result. 4b Moving on to limsup's and liminf's we obtain some differing results. Theorem 10.20 The set A, A, G, and G_ are each dense in [0,1]. Let us sketch the proof of this. At each stage in the construction the unit interval is divided into 3n (or F(2k)) pieces. At least one of every three pieces relates to the letter "a" and so contains points that will be in A, A, G, or G. As the number of pieces increase without bounds this means the points in these sets will be dense in the unit interval. The liminf's turn out to be small sets. Somewhat expectedly. Theorem 10.21 The sets A and G are each a-porous sets. Hence they are measure zero and first category. However, when we turn our attention to the sets that come from limsup, these turn Q
out to have some largeness to them. The complements A being a-porous. Hence
Q
and G are small, both
Theorem 10.22 The sets A and G have full measure and are residual in [0,1]. One counterintuitive result on Cantor Sets is how such a sparse set can be added to itself to form a large set (an interval). We close this section with some summation results for Ana and Golden Ana Sets and an open question. Now A C A C A and A + A = [0, 2] so any sum of two of those sets gives an interval. The results are different for Golden Ana Sets. The sum G + G is not an interval. For example, G + G misses all of the points between 11/24 and 1/2. However, using what is called normalized thickness, 7(G) = T ( G ) / ( 1 + T(G)), it has been shown that
G + G + G + G= [0,4]. Nothing is known about either G + G + G or the sum ofG_ and itself. However, since G has positive measure, the famous theorem of Steinhaus proves that G+ G contains an interval.
10.8 Average Distance In [37], the authors introduce a new way to distinguish self-similar sets from one another and use Cantor Sets in many of their examples. Unlike the last section, this method can work for sets where an interval is divided into more than three sections and more than two of the subintervals are retained. For example, they introduce two
202
GENERALIZATIONS AND APPLICATIONS
generalized Cantor Sets. In the first, called M\, is created using an iterative process where one starts with the closed interval [0,1], divides it into 16 equal subintervals, and then keeps only subintervals numbered 1, 5, 11, and 16. The second, Mi, uses the same division of the unit interval, but discards all but pieces 7, 2, 75, and 16. Thus Mi contains larger gaps than M\. However, in terms of measure (measure zero), category (nowhere dense), cardinality (uncountably many points), and dimen — dim.Box(Mi) = 1/2), they both agree. Thus the need for a sion (dimtfaus(Mi) new indicator of size. We begin with the usual Cantor construction, Cr, where the unit interval is divided into three pieces and the middle one removed, and r refers to the proportion removed. Now fix an r. At each stage, Kn, in the construction of Cr there are 2™ closed subintervals and we can compute the distance between points, T>(Kn). The average distance for Cr is then the limit of this calculation. Definition 10.2 Let CT be a Cantor Set. The average distance between points for C, is [CK- W \X ~ v\dx dy r -u i V(Cr) = lim V{Kn) = lim ■ " * " * * " ' / ' = L±i n->oo TWOO JJK axay r +6 K Thus the usual Cantor Set has average distance between points given by V{C) = \ . With r fixed, the Cantor Set Cr is guaranteed to be a measure zero set. There are Cantor Sets with positive measure, the so-called Fat Cantor Sets. What we need for a fat set is a variable ratio r^ at stage k and r^ —>• 0. In this instance, we will fix r and at stage k remove a middle interval of proportion r . We will write this as C\.k. This type of set has an average distance formula. Theorem 10.23 The average distance between points of the set Crk is 3(2 - r) As we have seen with Cantor Dust, it is possible to take these ideas into higher dimensions. Our notation here will be that C\ = C r C l and C " + 1 = CT x C™ C M™+1. Before we can continue we must define distance as these higher dimensions admit different generalizations of the Euclidean distance. We are using the taxicab metric; recall, in R 2 x = {x\, x^) and y = (y\, 1/2) and d(x,y)
= \x2 - xi\ + Ij/2 ~V\\-
Then it turns out that Theorem 10.24 IfC™ is the n-dimensional
v{cr;
Cantor Dust, then
'r + V r+3
NON-AVERAGING SETS
203
As for the motivating examples Mi and M2, the average distance for these can be computed using results on self-similar sets. This is Corollary 9 in [37], using the fact that the similitudes are linear function on M with a fixed contraction ratio. Theorem 10.25 If {Ti, T 2 , . . . , T/v} is a set of contractions on [0,1] satisfying the Open Set Condition with the added property that Ti = ax + bi, then for the invariant K we obtain T,(K\ 1
;
=
2
^l
2
~
b
i)
N -Nr
This leads us to T>(M\) = 25/63 andT>(Mi) = 29/63. This gives us quantitative evidence that the points in M% are more spread out than the points in M\. 10.9
Non-Averaging Sets
Let S be a set of real numbers. We say that S is non-averaging iffor any two points a,b £ S, their arithmetic mean, (a + b)/2, is not an element of S. For a fixed natural number n, it is easy to come up with a non-averaging set containing n elements. An active problem in number theory is as follows: Estimate the maximum number v(n) of elements in a non-averaging subset of {1,2,3,..., n}. It is currently known that cnl/1° < u(n) < cn2/s, where c is some constant, but the exact fractional exponent is not known. The Cantor Set is almost a non-averaging set. Keeping in mind the base three expansion of points in the set, for x,y E C we can write x — Y2ai/3l and y = ^2 bi/y where a,i, bi is either a 0 or a 2. If k represents the first natural number where
204
GENERALIZATIONS AND APPLICATIONS
To prove the dimension is large he using the theorem below. Theorem 10.26 Let {nk} be a subsequence of the natural numbers and let n(m) be the number of j < m which are not in the subsequence. Let f be a function which takes each finite sequence of digits into a digit. Suppose E is the set of x such that if x = Yl fli/101, then for each k, ank = f(a,i, a-2,..., ank-\). Suppose there is a number a and a natural number N such that for each m > N, n(m) > ma — N. Then dimHaus(E) > a andUa(E) > 0. Notice that not only does this say the Hausdorff dimension is 1, but it also says the Hausdorjf measure is positive. This is a special circumstance as there are many sets with dimension 1, but measure 0. Now we give Foran's construction. It is Cantor-like in that blocks of decimals are made 0 representing intervals being removed. Let S = {Np} — {10p} be our subsequence of natural numbers and note that Np+i > Np + 2. Let E(S) be the set of all x G [0,1] such that if
where a* = 0 , 1 , . . . , 9 and s(j) = j(j + l ) / 2 and bj = 0 , 1 , , . . . , s(j) — 1 and the a,j and bj satisfy for each x G E{S), i. bn + bn+l ■ ios
n
= Nv>
2. an+i = 0 when n ^ Np, 3. bn+2 = 0 when n = Np. then E(S) is a non-averaging set. By item 2 the elements ofE(S) can be averaged by doing so to the blocks bn of their decimals. By item 1, ifx, y,z G E(S) and z = (x + y)/2, then x = y = z follows from the fact that each of their block are equal. By item 3, E(S) is a porous set. For ifx = Y.h3l®~s(i)> n = Np, a = £™ + 1 bjlO'8^, then x G [a, a + 10"S("+2)] and (a + KT S ("+ 2 ), a + Kr^™ + 1 )) n E = 0. Hence the porosity of E at x is at least lim n->oo
jQ-s(n+l) _ lQ-s(n+2) -,——r = 1. 10_s(n+1)
Now fix m and suppose n, K are such that n(n + l)/2 > m > (n — l)n/2 and s{10K) <m< s(10K+1). Then n(M > m — n — 3(10 i f + 1 — l ) / 9 — 2>K since at most n of the a,j are fixed by item 2 and at most 3(10^ + 1 — 1) + 2>K ai are fixed by items] and3 Thusn(m) > m - v / 2 m - 3 ( 1 0 \ / 2 m / 9 + l/21n(2m)). Leta G (0,1); then n(m) — am > m(l — a) — \/2m — 3(10\/2m/9 + l/21n(2m)). Since the right side of the inequality approaches infinity, there is an N such that for m > N, n(m) — am > —N. Thus the set E(S) satisfies the hypotheses of the theorem for each a G (0,1) and hence d\ra.Haus{E{S)) = 1.
CANTOR SERIES AND CANTOR SETS
205
10.10 Cantor Series and Cantor Sets As stated Chapter 1, Cantor worked in several areas of mathematics including in cluding set theory, number theory, and trigonometric sequences. Let us introduce a new topic here, the Cantor Series. The idea behind series representations is to take an x £ (0,1) and un(x) so that oo n=l
where each un(x), n = 1, 2 , . . . is a function of x which is one-to one. An example such a series (not a Cantor Series) is as follows: Define un(x) as un(x) = en(x) ■ 10 _ n with en(x) a non-negative integer such that 0 < en(x) < 9 for each n and there are infinitely many n such that en(x) ^ 0. Definition 10.3 Let B — {bj} be a sequence of integers where bj > 2 for all j and let ej = ej(x) be integers satisfying 0 < ej < bj - 1, Then if ej ^ Ofor infinitely many j , then
for j = 1,2,
oo
.
. &1&2 * • -&j
3 — i-
is the Cantor Series of x. We note here that these x will be irrational numbers in (0,1). From this series idea we can then define a Cantor Set Ea by I
oo
Ea = I x — y
—-;—-—— : 0 < a,i < 6j_i, ai + 0 unless a,_i = 0
with the understanding that oo ^ 0. This can then bring us back over to continued fractions. For each Cantor Set E there is a corresponding a that has a regular continued fraction representation given by 1 a & i - l &2-1+T6 3 - 1 + " .
In the paper [81] several properties of these sets and continued fractions are established. Also worth pointing out, this paper links these ideas to applications in the study of the quasi-crystal spectrum, high energy physics, and E-infinity theory. For example, they state that the fractal dimension of spacetime can be expressed as 4+ . ,
l
l
=[4:4,4,4,...].
206
GENERALIZATIONS AND APPLICATIONS
First, let us look at a new, recursive way to describe the continued fraction sion of an x 6 (0,1). Let T : (0,1) —> (0,1) be defined by T(x)
= x
where |_VXJ is the integer part ofl/x.
Then we can write 1
Tk~la where T°a = a and Tka
expan
6fc - 1 +
Tka'
T(Tk-1a).
=
As we recall from Chapter 4, a simple continued fraction [oo : a\,a,2,Q>3, ■ ■ •] has a,i}, which in rational number form is written —, convergents Ci ~ [CLQ : ai,a^,..., Pi is an integer and qi a positive integer. Then the values of the numerators and denominators satisfy the equations Pi
=
qi
=
diPi-i+Pi-2,
algi_i+ft_2
for i — 2, 3, 4, 5 , . . . with initial values Po = a0,
p\ = aia.0 + 1,
<7o = 1,
qi
=ai.
We can now write a in terms of the transformation T as a
pk + qk +
Tapk_i Tkaqk^
for any k. The main result in [81] is the following theorem on the Hausdorff dimension of these Cantor Sets. Theorem 10.27 For any integer sequence {bn} with bi > 2, let a and T^a be de fined as above. Then the Hausdorff dimension of the Cantor Set Ea is An(a-Ta-T2a---Tk-1a) a i m f f a u s Ea = hm mf ■ fe-voo — ln(6i • 62 • • • bk) Although a little off topic, something we have covered already is the construction of continuous, nowhere differentiable functions. In [77], Wen gives another such construction that is accessible to first year calculus students and uses a Cantor Se ries.
LIOUVILLE NUMBERS AND IRRATIONALITY EXPONENTS
10.11
207
Liouville Numbers and Irrationality Exponents
A number is algebraic if it is the root of a polynomial with integer coefficients. Thus 3, 5/2, and \f2 are all algebraic numbers. If a number is not algebraic, then it is called transcendental. As easy as examples of algebraic numbers were, examples of transcendental numbers were difficult to find. In 1844, Joseph Liouville established Liouville numbers and became the first to prove the existence of transcendental num bers. Definition 10.4 A real number x is called a Liouville number if x is irrational and for each positive integer n there exists integers p and q such that <
1
where q > 1. Some definitions call for x to be a real number, but no rational number can be a Liouville number. This is easy to see. Suppose there exists integers c, d with d > 0 so that x = c/d. Choose a positive integer n so that 2n~l > d. Then for any integers p, q with q > 1 and p/q ^ c/d we have cq — pd 1 ~ ~ dq dq
1 1 > r 2 ~q " q n l
Thus a rational x cannot be a Liouville number. In [58] we see a proof that every Liouville number is a transcendental number. Since we have looked at various notions of size in Chapter 3, we will take a moment to look at the set L of Liouville numbers in terms of category and measure. Theorem 10.28 The set of Liouville numbers is second category set, yet of measure zero. Proof: To begin with, since every point in L is irrational we will write it as
L = Qcn (n„G„) where the Gn are open sets made by the following union of open intervals: Gn = U~ 2 U~-co (p/q - l/qn,p/q
+
l/qn).
Obviously all rational numbers are in each Gn hence each Gn is dense in the line. Each Gn is also an open set. Thus G% is nowhere dense. The complement of L is then Q c U (UG£), a first category set. To look at measure, we will show that for any positive integers m and n we can contain L n (—m,m) is a sequence of intervals whose length adds to less than (Am + l)/(n — 2). Thus the total length can be as small as we desire.
208
GENERALIZATIONS AND APPLICATIONS
ForfixedTO,n L n (-TO, m)GGnn
(-TO, m) C U~ 2 U™J_m, (p/q - l/qn,p/q
+
l/qn).
This means L n (—TO,TO)/ias an open cover using intervals whose lengths total to
E ^ £r=-m, 2 /«" = E ~ 2(2^9 + l)(2/g») < E ~ 2(4mg + g)(l/g") = (4TO + 1) £ ~
2
l / 9 n _ 1 < (4™ + 1) /i°° ^1_™ dx = ^ ± 1 .
Tn« gives us L fl (—m, m) liaj measure zero, hence L is measure zero. 4 77?e irrationality measure of a number is a measurement of how the value can be approximated by a rational number. Definition 10.5 For any real number x the irrationality exponent of x, /x(x) is the supremum of all real numbers p such that 1 P < — q qn has infinitely many rational number solutions, p/q. For a rational number, x, the irrationality measure is p(x) = 1. The ThueSiegel-Roth Theorem states that if x is irrational, but algebraic, then p(x) — 2. Louiville numbers are those numbers whose irrationality exponent is infinite. In [12], Bugeaud applied this idea to the Cantor Set to answer the question, "How close can irrational elements of Cantor Sets be approximated by rational number in Cantor's Set?" We present his answer, without proof, in the next theorem. Theorem 10.29 Let T be a real number with r > 2. Then the Cantor Set, C, contains uncountably many elements whose irrationality exponent is equal to r. In fact, for X > Oifwesetno(T,X) = 1 + max{0, |_(ln(2/A))/lnrJ}, then the real number
n>n ( )fr,A)
is an element ofC with p{rjT^\) = r ([-\ is the greatest integer function). Although we are not presenting the proof of this, we will note that it makes use of the so-called Folding Lemma ofMendes France [32] which deals with continued fractions using only the numbers 0 and 1. These relate to the peaks and valleys one gets when folding a sheet of paper, hence the name. Regardless, this furthers the connections between the topics in this book.
SETS OF SUMS OF CONVERGENT ALTERNATING SERIES
10.12
209
Sets of Sums of Convergent Alternating Series
Let an be a sequence of numbers where an = 0 or 1, and let {bn} be a fixed sequence of real numbers. We wish to look at all the possible series we can create via the product of(—l)a" and bn; that is OO
71=1
and the collections
5=JE(-l)a"^:K}G{0,l}N|. It is known that ifbn < J^'kLn+i ^k, then S is just a closed interval. The results in [22] show that under some circumstances the set S will be a generalization of a Cantor Set. There are several assumptions we will place on the situations. To begin with, S depends on the choice of bn, but not the sign of bn- Therefore we may assume bn > Ofor all n. If ^ bn = oo and lim^^oo bn = 0, then the Calculus II result on rearranging series shows that S — R. Hence we just need to investigate absolutely convergent series so assume oo
Y2 bn < oo. 71 =
1
One property of absolutely convergent series is that the sum is unchanged if the terms of the series are reordered. Thus our last assumption is h > b2 > b3 > ■ ■ ■ > bn > ■ ■ ■ > 0,
the series of terms is nonincreasing. Let us now relate this to a Cantor-like Set. We begin with the usual construction: At each step in this process we have an interval I and from that interval we remove a middle subinterval J so that £(J)/£(I) = 1/3, where £(■) refers to the length of an interval. Our first generalization, creating a new Cantor Set, was done by insisting the middle interval removed have ratio £(J)/£(I) = r, where 0 < r < 1. We will denote this Cantor Set by Cr. Now, rather than keep r constant, we use the sequence {r„}. For explanation's sense, let us label the two subintervals left after J is removed. If I = [a, b] and J — [c, d], then the left/right intervals remaining will be [a, c] and [d, b], respectively. This, again, then gives us Kn where each Kn consists of2n disjoint, closed intervals and CTn = C\Kn. The Ternary Cantor Set has rn constant at 1/3 and we have seen before that if rn —> Owe have a Fat Cantor Set, one with positive measure. In this generalization, we loosen the restriction on rn to let it be negative. So we will assume rn £ ( — 1,1) for all n G N. We now need an additional requirement: There is an increasing sequence of integers n^ such that rnk > 0.
210
GENERALIZATIONS AND APPLICATIONS
Our new construction is as before with the exceptions that we now say £(J)/t(I)
= \rn\
and, ifrn < 0 then the remaining left/right intervals will be (again, respectively) [a, d] and [c, b\. Still we have Kn, although this means Kn and Kn+\ Crn = C\Kn. We now have the following result:
can be the same set, and
Theorem 10.30 Let {rn} be a sequence of numbers in (—1,1) such that there is a subsequence {rHk} which is positive. Then one of the following is true: 1. C is a Cantor-like Set. 2. C can be written as a closure of a countable union of disjoint closed intervals. The proof will be omitted. It is straightforward, based on what we already know about a Cantor Set when rn > 0. There is a question of uniqueness here. Is it possible for two sequences rn and rn such that CTn = Cfn ? The answer is yes, for example with
r„ = {1/7,1/3,1/7,1/3,...} and fn = {-3/7, -1/5,1/3, -3/7, -1/5,1/3,...}. Our attention turns back to the set S. The statements of the results and their proofs are quite long and detailed, but can be found in [22]. Theorem 10.31 Let {bn} be a nonincreasing and positive sequence satisfying oc
b
^2
n>
bk
'
k=n + l
Then the corresponding set of sums S where S= lreR:r
= Y2{-l)a"bnforsome{an}
€ {0,1} N I
is Cantor-like. More precisely, the starting interval I of the construction is I
J2bk'Y.bk fc=l fc=l
and the positive sequence {r„} is defined by bn - 2^k=n+l b
T,T=n k
"•
n = 1,2,3,.
THE MONTY HALL PROBLEM
211
Conversely, if a > 0, I = [—a, a], and {rn}, rn e (0,1) then there is a correspond ing Cantor-like set C. For this C there exists a unique decreasing positive sequence {bn} satisfying bn > E f c l n + i bk Siven h
2 whose set of sums is C. If we reverse the requirement on the tails of the sum ofbn, then we end up with an interval rather than a Cantor-like set. Theorem 10.32 Let {bn} be a nonincreasing and positive sequence
satisfying
oo
bn < ^2
bk
k=n+l
-
Then the corresponding set of sums S where S = I r G E : r = ] T ( - l ) a " & n for some {an}
e {0,1}N I
is the interval oo
oo
■ Hfefc>J2
h
fe=l fc=l
More results can be found in various papers by Dindos and by Menon. T. Saldt has also published many papers on this topic, but in Slovak.
10.13
The Monty Hall Problem
As we have stated many times before, one of the more intriguing properties of the Cantor Set is its pervasiveness. It often shows up where it is least expected. In this section, we see how a variation of the Cantor Set appears in a generalization of a famous recreational math problem. First, the original problem. The Monty Hall Problem You are a contestant on the game show Let's Make a Deal. The host, Monty Hall, shows you three giant boxes. Behind one of the boxes is a brand new car, the other two contain "zonks," gag prizes worth nothing. You do not know which box hides the car. You choose one of the boxes to be yours. Monty then reveals one of the boxes you did not choose and shows you the zonk inside that box. Now you are given a choice. Keep the box you originally chose, or switch to the other box whose contents are unknown. The question is, is it better to keep your box or change prizes?
212
GENERALIZATIONS AND APPLICATIONS
The answer, which surprises many including mathematicians and statisticians, is you should always switch boxes. The probability of getting the car is doubled by switching. We shall not go into explicit detail, although the interested reader may like [66]. The heuristic argument is easy enough to follow. When the contestant makes his/her first choice, Box 1, Box 2, or Box 3, there is no extra information and we are looking at a random guess. Thus the probability of choosing the box with the car is 1/3. The fact that Monty reveals a non-winning box does not change this. So when there are two boxes left, the probability that the other box contains the car is 1 — 1/3 = 2/3, hence the probability of winning is doubled when you switch. There are many articles written on this topic and variations on the game. A par ticularly good one for undergraduates is [67]. We will look at a problem from [59]. In a TV contest, the host randomly hides a single prize in one of several boxes. The contestant chooses a box and then the host - who knows where the prize is hidden - picks a box that is not the one the contestant chose, opens it, and shows the contestant and the audience that the box does not contain the prize. The empty box is removed and the contestant is given the choice of remaining with the chosen box or picking a new box. If the contestant chooses a new box, the remaining boxes, including the one previously held by the contestant, are jumbled randomly. This process continues until there are only two boxes left: the one the contestant holds and one other. The contestant is then offered one last chance to change. The question, as always, is what are the probabilities of winning associated with each box? If the contestant always stays with the box originally chosen, then the probability of having the winning box is 1/n. On the other end of the spectrum, if the contestant holds onto the original box until the last possible moment, and then switches, the probability of having the winning box is 1 — 1/n = — - . Of course, when n = 3 this is the Monty Hall Problem and we have the previous probabilities ofl/3and2/3. Let us now look at the probabilities associated with other strategies. During the course of the game, there are n — 2 offers to switch boxes. This means 2™~2 differ ent strategies which can be represented by a finite sequence of decreasing positive integers {n, a,k, flfc-i, • • •, a,2, ai} where n is the number of boxes at the start of the game, k changes were made, and a,i denotes that a change of boxes was made when there were ai boxes with which one is left to switch. For example, {5, 3,1} means the game started with 5 boxes and changes were made twice, once when there were 3 boxes left, and once when there was 1 box left (note: this 1 box, and the previ ous 3 boxes, are the number of choices of boxes with which to switch; this means if a i = 1 there are two boxes, one the contestant is holding, the other with which the contestant can swap) giving 02 — 3 and a\ — 1. Two items worth pointing out in our notation • cik 7^ n — 1. One box is held by the contestant, one box has already been revealed to be empty, thus the number of boxes from which the contestant can choose is at most n — 2.
THE MONTY HALL PROBLEM
213
■ n > cik > fflfe-i > ■ ■ ■ > ci2 > di > I. The sequence is actually strictly decreasing. Now we turn our attention to computing probabilities. If the first switch is made when there are ak boxes left for choosing, then let pk denote the probability of choos ing correctly among the ak boxes. The value ofpk is given by \
nj
ak
this product meaning that the wrong box was chosen at first and the right box chosen at the switch. In the same vein, pk-i
= (1 -ak)
. flfc-i
This last expression can be rewritten by substituting in the formula for pk to arrive at Pfe-i = 1 1 1 1 • Ofe-i flfe-iafc ak-\akn Continuing this process we arrive at the probability of obtaining the winning box by switching when there is one box left from which to choose. This value is p\ and is given by 1 1 (-l)fc_1 (-l) f c pi = + • ■ • + -i '+ —i 1 . a^ai ■ ■ ■ akn a\ a\a-2 aia'2---ak So, in our example {5,3,1} we have the contestant's probability of winning is 1 1 1 11 1 1-3 1-3-5 15 Surprisingly, this can be worked in the backwards direction. Given a rational number pi, we can determine the values ofn and the a». This is due to the following theorem. Pi
Theorem 10.33 Any rational number p/q € (0,1] has a unique representation q
a\
a^a-i
a\ai---a,k
where a* are positive integers such that 1 < ai < a2 < • • • < a.fc-i < afc — 1. Proof: The interval (0,1] can be written at the union of intervals of the form {{n 1) _ 1 , n - 1 ] . Thus any value a £ (0,1] belongs to exactly one of the intervals ((n l ) - 1 , n - 1 ] . And so 1 / 1 a = n Ai \n
1 ^ * 1 n+lj n
Ai n{n+\)
214
GENERALIZATIONS AND APPLICATIONS
where Ai G [0,1). If we say ai = \\/{n + 1), we get that a = (1 — a\)/n and cti £ (0, (n + 1) _ 1 ). Doing as we did before, this time with (0, (n + l ) " 1 ) , we find ot\ = (1 — a2)/mfor some m > n, and this leads us to 1 n
1 nm
-OL2-
We summarize the recursive process as 1
with ao = a,
ati-i
and OL\
1 -
aj-iftj,
where \x\ is the greatest integer function. Now, when a is an irrational number, then ai must be irrational for all i and the process never terminates. If a is a rational number, then the process is a modified Euclidean algorithm with a strictly decreasing sequence of remainders, hence the process must terminate. In applying the Euclidean algorithm to our rational number we create a sequence of dividends a\. a,2, ■ ■ ■, a^ with the property that ak-i < a*; — 1. This, plus the fact that 1 1 1 1 < < gives us the uniqueness of expansion that we needed. 4* As a consequence of this proof we now have the fact that for any number a G (0,1] we have 1 1 (-l)*-1 a= + •■• + + ••• , ai aia,2 aia 2 ---a/ c where 1 < ai < a 2 < 03 < • • •. This the referred to as the Pierce expansion of the number. We will denote this expansion by < a\, 02, ^3, • • • >• For a quick example where the sequence is infinite, let us look at < 1,2,3,4,... >. This represents the infinite series 1 1 1 1 lT~2!+3!~4!+'"' which is the Taylor Series expansion of f(x) = 1 — ex evaluated at x = —1. Thus 1 1 l!~2!
+
1 1 3!~4!
+
1 '"~ ~e' i
Incidentally, this proves that e is an irrational number. For this application, we go back to the historical origins of the Cantor Set. The set answered the question over the existence of a non-empty, perfect, nowhere dense set. Similar to the Cantor Ternary Set consisting ofpoints in [0,1] that have a ternary expansion that has no 1 's, we leave something out of the Pierce expansion.
THE MONTY HALL PROBLEM
fl
215
EXAMPLE 10.8 Let C be the set of real numbers in (0,1] whose Pierce expansion contains no odd integers. In [59] we see that this is a perfect and nowhere dense set. We also note that it is measure zero, due to the fact that
n—1
N
'
Additionally, this set is uncountable as there is a one-to-one correspondence between the element ofC and (0,1] given by /ai
a2 a 3
\
■
CHAPTER 11
EPILOGUE
We have how seen what began as a search for a pathological example; an example of a perfect, nowhere dense set is now so much more. The Cantor Set is useful in multiple areas, including topology, analysis, abstract algebra, and fractal geometry. It lends itself to pathological examples in studies such as probability theory, excep tional (e.g., measure, category, dimension) sets, and generalized continuity. Still, there are many different directions for new research by professors, graduate stu dents, and undergraduates. This text serves as a window into some of the areas one can search for new results. Of course this book contains just a small amount of the work done on this topic. As of May 2013, MathSciNet® lists over 2600 papers and books containing the keyword "Cantor Set."
The Elements of Cantor Sets -With Applications, First Edition. By Robert W. Vallin Copyright © 2013 John Wiley & Sons, Inc.
217
REFERENCES
1. Steven Astels, "Thickness measures for Cantor sets," Electron. Res. Announc. Am. Math. Soc. 5(1999), 108-111. 2. George Bachman, "Introduction to p-adic Numbers and Valuations Theory," Academic Press, 1964. 3. In Soo Baek, "Dimensions of the perturbed Cantor set," Real Anal. Exchange, 19(1), 1993/94, 269-273. 4. In Soo Baek, "Hausdorff measure on perturbed Cantor sets," Real Anal. Exchange, 20(2), 617-621. 5. Michael Barnsley, "Fractals Everywhere," Academic Press, 1988. 6. C.L. Belna, G.T Cargo, M.J. Evans, andP.D. Humke, "Analogues of the Denjoy-YoungSaks Theorem," Trans. AMS, 271 (1982), 253-260. 7. A.S. Besicovitch and S.J. Taylor, "On the complementary intervals of a linear closed set of zero Lebesgue measure," J. London Math. Soc, 29(1954), 449-459. 8. Duane Boes, Richard Darst, and Paul Erdos, "Fat, symmetric, irrational Cantor sets," Am. Math. Monthly 88 (1981), no. 5, 340-341. 9. Z.I. Borevich and I.R. Shafarevich, "Number Theory," Academic Press, 1966. 10. Andrew M. Bruckner, "Differentiation of Real Functions," American Mathematical So ciety, 1994. 11. Andrew M. Bruckner, Judith B. Bruckner, and Brian S. Thomson, "Elementary Real Analysis," Prentice-Hall, 2001. The Elements of Cantor Sets -With Applications, First Edition. By Robert W. Vallin Copyright © 2013 John Wiley & Sons, Inc.
219
220
REFERENCES
12. Yann Bugeaud, "Diophantine approximation and Cantor sets," Math. Ann., 341 (2008), 677-684. 13. Krzystof Ciesielski, Hajrudin Fejzic, and Chris Freiling, "Measure zero sets with nonmeasurable sums," Real Anal. Exchange, 27 (2001/02), 783-793. 14. W. A. Coppel, "An interesting Cantor set," Am. Math. Monthly, 90 (1983), 456-460. 15. Sylvain Crovisier and Michal Rams, "IFS attractors and Cantor sets," Topology Apps., 153 (2006), 1849-1859. 16. T.W. Cusick, "Sums and products of continued fractions," Proc. AMS, 27 (1971), 35-38. 17. T.W. Cusick, "On M. Hall's continued fraction theorem," Proc. AMS, 38 (1973), 253254. 18. Richard B. Darst, "Some Cantor sets and Cantor functions," Math. Mag., 45 (1972), 2-7. 19. Joseph Warren Dauben, "Georg Cantor-His Mathematics and Philosophy of the Infi nite," Princeton University Press, 1990. 20. A. Denjoy, "Sur Une Propriete des Series Trigonometrigues," Verlag v.d.G.V. der Wis-en Natuur. Afd., 1920. 21. Keith Devlin, "The Joy of Sets," Springer-Verlag, 1993. 22. Martin Dindos, "Generalized Cantor sets and sets of sums of convergent alternating series," J. Appl. Anal, 7 (2001), 131-150. 23. Bohuslav Divis, "On the sums of continued fractions," Ada Arith., 22 (1973), 157-173. 24. Josef Dobos, "The standard Cantor function is subadditive," Proc. AMS, 124 (1996), no. 11, 3425-3426. 25. E.P. Dolzenko, "Boundary properties of arbitrary functions (in Russian)," Izv. Ada. Nauk SSSR Ser. Mat., 31 (1967), 3-14. 26. Gerald Edgar, "Measure, Topology, and Fractal Geometry," Springer-Verlag, 1990. 27. M.J. Evans, PD. Humke, and K. Saxe, "A characterization of asymmetrically porous symmetric Cantor sets" Proc. AMS, 112 (1994), 805-810. 28. M.J. Evans, "Some theorems whose a-porous exceptional sets are not a -symmetrically porous," Real Anal. Exchange, 17(1991/92), 809-814. 29. Kenneth Falconer, "The Geometry of Fractal Sets," Cambridge University Press, 1985. 30. Kenneth Falconer, "Fractal Geometry, Mathematical Foundations and Applications," J. Wiley & Sons, 1990. 31. James Foran, "Non-averaging sets, dimension, and porosity," Canad. Math. Bull., 29 (1986), 60-63. 32. Michel Mendes France, "Sur les fractions continues limitees ," Ada Arith., 23 (1973), 207-215. 33. Russell A. Gordon, "The Integrals of Lebesgue, Denjoy, Perron, and Henstock," AMS, 1994. 34. Fernando Q. Gouvea, "Was Cantor surprised?" Am. Math. Monthly 118 (2011), no. 3, 198-209. 35. Marshall Hall, Jr., "On the sum and product of continued fractions," Ann. Math., 48 (1947), 966-993.
REFERENCES
221
36. Paul R. Halmos, "Naive Set Theory" Springer-Verlag, 1974. 37. Gregg Hartvigsen, Christopher C. Leary, and Dennis A. Ruppe, "Fractals, average dis tance, and the Cantor set," Fractals, 18 (2010) 327-341. 38. Kathryn E. Hare, Franklin Mendivil, and Leandro Zuberman, "The sizes of rearrange ments of Cantor sets," Canadian Math. Bull., 56 (2013), no. 2, 354-365. 39. Brian R. Hunt, Ittai Kan, and James A. Yorke, "When Cantor sets intersect thickly," Trans. Am. Math. Soc, 339 (1993), no. 2, 869-888. 40. Corey Jones, (private communication) 41. K. Kakeya, "On the partial sums of an infinite series," Tohoky Sci, Rep., 3 (1914), 159164. 42. Hidefumi Katsuura, "Continuous nowhere-differentiable Functions-an application of contraction mappings," Am. Math. Monthly, 98. no. 5 (1990), 411-416. 43. John L. Kelley, "General Topology," Van Nostrand, 1975. 44. Neal Koblitz, "p-adic Numbers, p-adic Analysis, and Zeta-Functions (2nd ed.)," Springer-Verlag, 1984. 45. Roger Kraft, "What's the difference between Cantor sets?," Am. Math. Monthly, 101 (1994), 640-650. 46. Roger Kraft, "Intersections of Thick Cantor Sets," Mem. Am. Math. Soc, 97 (1992), no. 468. 47. Steven G. Krantz, "An Episodic History ofMathematics: Mathematical Culture Through Problem Solving" MAA, 2010. 48. Timothy Lance and Edward Thomas, "Arcs with positive measure and a space-filling curve," Am. Math. Monthly 98 (1991), no. 2, 124-127. 49. R. Daniel Mauldin, "The Scottish Book: Mathematics from the Scottish Cafe," Birkhauser, 1982. 50. Mark McClure, "Self-similar structure in Hilbert's space filing curve," Math. Mag., 76 (2003), no. 1, 40-47. 51. Mark McClure, "The Hausdorffdimension of Hilbert's coordinate functions." Real Anal. Exchange 24 (1998/99), no. 2, 875-883. 52. P.K. Menon, "On a class of a perfect set," Bull. Am. Math. Soc, 54(1948), 706-711. 53. Ioana Mihaila, "The rationals of the Cantor set," College Math J., 35 (2004), 251-255. 54. Pedro Mendes, "Sum of Cantor sets: self-similarity and measure," Proc AMS, 127 (1999), 3305-3308. 55. Sheldon Newhouse, "The abundance of wild hyperbolic sets and nonsmooth stable sets for diffeomorphisms," Inst. Hautes Etudes Sci. Publ. Math., (1979), no. 50, 101-151. 56. J.E. Nymann, "The Sum of the Cantor Set with Itself," I'Enseign. Math., 39(1993), 177178. 57. J.E. Nymann, "Linear combinations of Cantor sets," Colloq. Math., LXVIII(1995) 259264. 58. John C. Oxtoby, "Measure and Category," Springer-Verlag, 1987. 59. Jaume Paradis, Pelegri Viader, and Liuis Bibiloni, "A mathematical excursion: From the three-door problem to a Cantor-type set," Am. Math. Monthly, 106 (1999), 241-251.
222
REFERENCES
60. J. Palis and F. Takens, "Hyperbolicity and Sensitive Chaotic Dynamics at Homoclinic Bifurcations," Cambridge University Press, 1993. 61. M. Pavone, "The Cantor Set and a Geometric Construction," I'Enseign. Math., 35 (1989), 41-49. 62. Joseph L. Pe, "Ana's golden fractal," Fractals 11 (2003), no. 4, 309-313. 63. Clifford Pickover, "Wonders of Numbers," Oxford Univ. Press, 2001. 64. Hui Rao, Huo-Jun Ruan, and Ya-Min Yang, "Gap sequence, Lipschitz equivalence, and box dimension offractal sets," Nonlinearity, 21 (2008), 1339-1347. 65. Ira Rosenholtz, "Another proof that any compact metric space is the continuous image of the Cantor set," Am. Math. Monthly, 83 (1976) 646-647. 66. Jason Rosenhouse, "The Monty Hall Problem: The Remarkable Story of Math's Most Contentious Brain Teaser," Oxford University Press, 2009. 67. Jeffrey Rosenthal, "Monty Hall, Monty fall, Monty crawl," Math Horizons, 16 (2008), 5 -7. 68. H.L. Royden, "Real Analysis, 2nd ed.," Macmillan Publishing, 1968. 69. Alan H. Schoenfeld, "Continuous surjections from Cantor sets to compact metric spaces," Proc. AMS, 46 (1974), 141-142. 70. Boris Solomyak, "On the measure of arithmetic sums of Cantor sets," Indag. Math., N.S. 8(1997), 133-141. 71. Claude Tricot, "Douze definitions de la densite logartihmique," C.R. Acad. Sci. Paris, 342 (1981), 549-552. 72. Robert W. Vallin, "Sums and sizes of Ana sets," (to appear) 73. Robert W. Vallin, "Newhouse thickness and porosity of Cantor sets," Real Anal. Ex change, 27(2001/02), 349-358. 74. Robert W. Vallin, "Symmetric porosity, dimension, and derivates," Real Anal. Exchange, 17 (1991/92), 314-321. 75. A.C.M. Van Rooij and W.M. Schikhoff, "A Second Course on Real Functions," Cam bridge University Press, 1982. 76. VS. Vladimirov, I.V Volovich, and E.I. Zelenov, "p-adic Analysis and Mathematical Physics," World Scientific, 1994. 77. Lu Wen, "A nowhere differentiable continuous function," Am. Math. Monthly, 107 (2000), 450-453. 78. R.F. Williams, "How big is the intersection of two thick Cantor sets," Contemp. Math., 117, Am. Math. Soc, (1991) 163 - 175, 79. Grace C. Young and W. H. Young, "Discontinuous functions continuous with respect to every straight line," The Q. J. of Pure Appl. Math., 41 (1910), 87-95. 80. Ludek Zaji'cek, "Porosity and a-porosity," Real Anal. Exchange, 13 (1987-88), 314-350. 81. Ting Zhong, Jing-Jing Zhang, and Liang Tang, "A class of Cantor sets associated with the regular continued fractions," Comput. Math. Appl, 61 (2011) 2251-2255.
Index
Ta set, 8, 172 Qs set, 8, 172 affine transformation, 106 algebraic numbers, 27 Ana set definition, 198 Golden, 199 Archimedean Property, 54 Baire Category Theorem, 33 Bernstein set, 188 Besicovitch-Taylor dimension, 137 Bolzano-Weierstrass Theorem, 103 Borel sets, 172 boundary, 8 box counting dimension definition, 128 lower, 128 upper, 128 Cantor Cantor Cantor Cantor Cantor
dimension, 183 dust, 195, 202 Function, 161 Series, 205 Set
Ternary, 17 construction, 17, 59, 149 derivation, 150 differences, 193 fat, 38, 49, 185 generalzied, 181 in base 3, 20 p-adic, 80, 86 perturbed, 183 products, 195 sums, 152, 186 thickness, 150 Cantor target, 197 Cantor, Georg, 2 cardinality, 22 Cartesian product, 26 Cauchy sequence, 99 closed set, 8, 30 continued fraction definition, 52 simple, 52 Continuum Hypothesis, 29 Contraction Mapping Theorem, 108 convergence in a topological space, 30 convergence of a function
The Elements of Cantor Sets -With Applications, First Edition. By Robert W. Vallin Copyright © 2013 John Wiley & Sons, Inc.
224
INDEX
pointwise, 158 uniform, 159 convergent?, 56 Darboux integral, 42 Darboux sum, 42 dimension Besicovitch-Tayior, 137 box counting, 127, 128 Cantor, 183 Hausdorff', 132 inductive, 124 modified box-counting, 137 packing, 136 diophantine equation definition, 60 linear, 60 division ring, 71 equivalence class, 23 equivalence relation, 23 Euclidean Algorithm, 61 fat Cantor Set, 38, 49, 185 Fibonacci number, 199 field, 71 first category; 32 fixed point, 107 function Baire class one, 171 bijection, 9 characteristic, 9, 40 continuous, 11, 31 Darboux, 173 decreasing, 29 definition, 9 homeomorphism, 35 increasing, 29 inverse, 9 Lebesgue integral of, 45 linearly continuous, 177 Lipschitz, 10, 14, 15, 141 measurable, 40 restriction, 9 Riemann integral of, 42 separately continuous, 177 simple, 43 Golden Ana set, 199 golden ratio, 56, 199 group abelian, 69 definition, 68 Hausdorff s—measure, 132
Hausdorff s—outer measure, 132 Hausdorff dimension, 132, 154, 182 Hausdorff metric, 105 homeomorphism, 35 inductive dimension, 124 infimum, 120 Integral Lebesgue, 45 Riemann, 42 integral domain, 70 interior, 8 Intermediate Value Property, 173 Iterated Function System, 110 Koch Curve, 111 Lebesgue integral, 44, 45 Lebesgue measure, 132 limit infimum function, 122 of a sequence, 121 limit point, 101 limit supremum function, 122 of a sequence, 121 linearly continuous function, 177 Liouville numbers, 207 Lipschitz function, 10, 14, 15, 141 lower bound, 119 Lucas number, 199 Maclaurin series, 64 mass distribution, 134 Mass Distribution Principle, 135 measurable function, 40 measurable set, 37 metric p—adic, 73 complete metric space, 100 definition, 12 Hausdorff, 105 space, 13, 92 modified box-counting dimension, 137 Monty Hall Problem, 211 non-averaging sets, 203 norm, 73 open ball, 94 open cover, 36 open set, 8 Open Set Condition, 128, 133, 182 ordered derivation, 150 outer measure, 36
p-adic Cantor Set, 80, 86 integer, 77 numbers, 72 rational, 82 packing dimension, 136 partition, 41 perturbed Cantor Set, 183 Pigeon-Hole Principle, 103 porosity definition, 144 generalized, 154 right, 144 porosity index, 145 porous set, 144 quadratic irrational, 53 relation definition, 22 equivalence, 23 Riemann integral, 42 Riemann sum, 41 ring, 70 second category, 32 self-similar, 91 sequence bounded, 100 Cauchy, 99 definition, 10 in a metric space, 97 set Ta, 8, 172 Qs, 8, 172 bounded, 8, 95 closed, 8, 30, 94, 102 closure, 101 compact, 103 countable, 24 definition, 5 dense, 32 equivalent sets, 22 finite, 24 infinite, 24 intervals, 8 measurable, 37 nowhere dense, 32 open, 8, 29, 94 perfect, 102 porous, 144 small, 21 symmetric, 146 totally bounded, 95 uncountable, 24
very porous, 156 Sierpinski Triangle, 110, 141 similarity dimension, 127 simple continued fraction, 52 simple function, 43 space-filling curve, 168 subsequence, 102 supremum, 120 symmetric porosity, 147 thickness, 150 Thomae's Function, 34 topological space, 29 topology basis for, 123 definition, 29, 123 transcendental numbers, 27 transfinite induction, 188 ultra-metric, 74 uniformly Cauchy, 162 upper bound, 119 valuation, 73