Dimensions, Embeddings, and Attractors
Dimensions, Embeddings, and Attractors
General Editors
B. BOLLOB Á S, W . FU L T O N , A . K A T O K , F . K I R W A N ,
P. SARNAK, B. SIMON, B. TOTARO
JAMES C. ROBINSON
University of Warwick
cambridge university press
Cambridge, New York, Melbourne, Madrid, Cape Town, Singapore,
São Paulo, Delhi, Dubai, Tokyo, Mexico City
Cambridge University Press
The Edinburgh Building, Cambridge CB2 8RU, UK
Published in the United States of America by Cambridge University Press, New York
www.cambridge.org
Information on this title: www.cambridge.org/9780521898058
C J. C. Robinson 2011
A catalogue record for this publication is available from the British Library
Preface page xi
Introduction 1
vii
viii Contents
The main purpose of this book is to bring together a number of results concern-
ing the embedding of ‘finite-dimensional’ compact sets into Euclidean spaces,
where an ‘embedding’ of a metric space (X, ̺) into Rn is to be understood as a
homeomorphism from X onto its image. A secondary aim is to present, along-
side such ‘abstract’ embedding theorems, more concrete embedding results
for the finite-dimensional attractors that have been shown to exist in many
infinite-dimensional dynamical systems.
In addition to its summary of embedding results, the book also gives a unified
survey of four major definitions of dimension (Lebesgue covering dimension,
Hausdorff dimension, upper box-counting dimension, and Assouad dimension).
In particular, it provides a more sustained exposition of the properties of the box-
counting dimension than can be found elsewhere; indeed, the abstract results
for sets with finite box-counting dimension are those that are taken further in
the second part of the book, which treats finite-dimensional attractors.
While the various measures of dimension discussed here find a natural
application in the theory of fractals, this is not a book about fractals. An
example to which we will return continually is an orthogonal sequence in an
infinite-dimensional Hilbert space, which is very far from being a ‘fractal’. In
particular, this class of examples can be used to show the sharpness of three of
the embedding theorems that are proved here.
My models have been the classic text of Hurewicz & Wallman (1941) on
the topological dimension, and of course Falconer’s elegant 1985 tract which
concentrates on the Hausdorff dimension (and Hausdorff measure). It is a
pleasure to acknowledge formally my indebtedness to Hunt & Kaloshin’s 1999
paper ‘Regularity of embeddings of infinite-dimensional fractal sets into finite-
dimensional spaces’. It has had a major influence on my own research over the
last ten years, and one could view this book as an extended exploration of the
ramifications of the approach that they adopted there.
xi
xii Preface
Part I of this book treats four different definitions of dimension, and investigates
what being ‘finite dimensional’ implies in terms of embeddings into Euclidean
spaces for each of these definitions.
Whitney (1936) showed that any abstract n-dimensional C r manifold is C r -
homeomorphic to an analytic submanifold in R2n+1 . This book treats embed-
dings for much more general sets that need not have such a smooth structure;
one might say ‘fractals’, but we will not be concerned with the fractal nature
of these sets (whatever one takes that to mean).
We will consider four major definitions of dimension:
1
2 Introduction
(iv) The Assouad dimension dA (X), a ‘uniform localised’ version of the box-
counting dimension: if B(x, ρ) denotes the ball of radius ρ centred at
x ∈ X, then N(X ∩ B(x, ρ), r) ∼ (ρ/r)dA (X) for every x ∈ X and every
0 < r < ρ (Chapter 9). This definition appears unfamiliar outside the
area of metric spaces and most results are confined to research papers
(e.g. Assouad (1983), Luukkainen (1998), Olson (2002); but see also
Heinonen (2001, 2003)).
and there are examples showing that each of these inequalities can be strict. We
will check that each definition satisfies the natural properties of a dimension:
monotonicity (X ⊆ Y implies that d(X) ≤ d(Y )); stability under finite unions
(d(X ∪ Y ) = max(d(X), d(Y ))); and the dimension of Rn is n (a consistent
way to interpret this so that it makes sense for all the definitions above is that
d(K) = n if K is a compact subset of Rn that contains an open set). We will
also consider how each definition behaves for product sets.
Our main concern will be with the embedding results that are available
for each class of ‘finite-dimensional’ set. The embedding result for sets with
finite covering dimension, due to Menger (1926) and Nöbeling (1931) (given
as Theorem 1.12 here), is in a class of its own. The result guarantees that when
dim(X) ≤ d, a generic set of continuous maps from a compact metric space
(X, ̺) into R2d+1 are embeddings.
The results for sets with finite Hausdorff, upper box-counting, and Assouad
dimension are of a different cast. They are expressed in terms of ‘prevalence’
(a version of ‘almost every’ that is applicable to subsets of infinite-dimensional
spaces, introduced independently by Christensen (1973) and Hunt, Sauer, &
Yorke (1992), and the subject of Chapter 5), and treat compact subsets of
Hilbert and Banach spaces. Using techniques introduced by Hunt & Kaloshin
(1999), we show that a ‘prevalent’ set of continuous linear maps L : B → Rk
provide embeddings of X when d(X − X) < k, where
X − X = {x1 − x2 : x1 , x2 ∈ X}
and d is one of the above three dimensions (see Figure 1). Note that if one
wishes to show that a linear map provides an embedding, i.e. that Lx = Ly
implies that x = y, this is equivalent to showing that Lz = 0 implies that z = 0
for z ∈ X − X. This is why the natural condition for such results is one on the
‘difference’ set X − X; but while dB (X − X) ≤ 2dB (X), there are examples of
Introduction 3
sets for which dH (X) = 0 but dH (X − X) = ∞ (and similarly for the Assouad
dimension).
Where the embedding results for these three dimensions differ from one
another is in the smoothness of the parametrisation of X provided by L−1 . In the
Hausdorff case this inverse can only be guaranteed to be continuous (Chapter 6);
in the upper box-counting case it will be Hölder (Chapter 8); and in the Assouad
case it will be Lipschitz to within logarithmic corrections (Chapter 9). Simple
examples of orthogonal sequences in ℓ2 (or related examples in c0 , the space of
sequences that tend to zero) show that the results we give cannot be improved
when the embedding map L is linear.
Chapter 4 presents an embedding result for subsets X of RN with box-
counting dimension d < (N − 1)/2. The ideas here form the basis of the results
for subsets of Hilbert and Banach spaces that follow, and justify the development
of the theory of prevalence in Chapter 5 and the definition of various ‘thickness
exponents’ (the thickness exponent itself, the Lipschitz deviation, and the dual
thickness) in Chapter 7.
Part II discusses the attractors that arise in certain infinite-dimensional
dynamical systems, and the implications of the results of Part I for this class of
finite-dimensional sets. In particular, the embedding result for sets with finite
box-counting dimension is used toward a proof of an infinite-dimensional ver-
sion of the Takens time-delay embedding theorem (Chapter 14) and it is shown
that a finite-dimensional set of real analytic functions can be parametrised using
a finite number of point values (Chapter 15).
Chapter 10 gives a very cursory summary of some elements of the theory
of Sobolev spaces and fractional power spaces of linear operators, which are
4 Introduction
Finite-dimensional sets
1
Lebesgue covering dimension
There are a number of definitions of dimension that are invariant under home-
omorphisms, i.e. that are topological invariants – in particular, the large and
small inductive dimensions, and the Lebesgue covering dimension. Although
different a priori, the large inductive dimension and the Lebesgue covering
dimension are equal in any metric space (Katětov, 1952; Morita, 1954; Chapter
4 of Engelking, 1978), and all three definitions coincide for separable metric
spaces (Proposition III.5 A and Theorem V.8 in Hurewicz & Wallman (1941)).
A beautiful exposition of the theory of ‘topological dimension’ is given in the
classic text by Hurewicz & Wallman (1941), which treats separable spaces
throughout and makes much capital out of the equivalence of these definitions.
Chapter 1 of Engelking (1978) recapitulates these results, while the rest of his
book discusses dimension theory in more general spaces in some detail.
This chapter concentrates on one of these definitions, the Lebesgue covering
dimension, which we will denote by dim(X), and refer to simply as the covering
dimension. Among the three definitions mentioned above, it is the covering
dimension that is most suitable for proving an embedding result: we will show
in Theorem 1.12, the central result of this chapter, that if dim(X) ≤ n then a
generic set of continuous maps from X into R2n+1 are homeomorphisms, i.e.
provide an embedding of X into R2n+1 .
There is, unsurprisingly, a topological flavour to the arguments involved
here, and consequently they are very different from those in the rest of this
book. However, any survey of embedding results for finite-dimensional sets
would be incomplete without including the ‘fundamental’ embedding theorem
that is available for sets with finite covering dimension.
7
8 Lebesgue covering dimension
The order of a covering is the largest integer n such that there are n + 1
members of the covering that have a nonempty intersection. A covering β is a
refinement of a covering α if every member of β is contained in some member
of α.
β ′ := {U ∈ β : U ∩ B = ∅}
1 In the context of metric spaces it is somewhat artificial to make the definition in this form, since
(A, ̺) is a metric space in its own right. But our main focus in what follows will be on subsets
of Hilbert and Banach spaces, where the underlying linear structure of the ambient space will be
significant.
1.1 Covering dimension 9
Proof We will say that an open covering α of X has order at most n at points
of Y if every point in Y lies in no more than n + 1 elements of α.
First we show that any open covering α of X has a refinement that has order
at most n at points of X1 . Any such covering of X provides a covering of X1 ,
which has a refinement β ′ that has order at most n. For every V ∈ β ′ , there
exists an element UV ∈ α such that V ⊂ UV . Then
β = {UV : V ∈ β ′ } ∪ {U \ X1 : U ∈ α}
is the required refinement of α. We can repeat this argument starting with the
covering β of X, and obtain a covering γ that refines β and has order at most
n at points of X2 .
We now define a further covering of X, which will turn out to be a refinement
of α of order at most n. As a first step in our construction, define a map
f : γ → β by choosing, for each G ∈ γ , an f (G) ∈ β such that G ⊂ f (G)
(this is possible since γ refines β). Now for each B ∈ β, let
and let δ be the union of all the sets d(B) (over B ∈ β).
Now, δ is a refinement of α, since d(B) ⊂ B for every B ∈ β, and β is a
refinement of α. Also, δ still covers X since γ covers X and every G ∈ γ is
contained in some B ∈ β (as γ refines β). All that remains is to show that δ
has order at most n.
Suppose that x ∈ X with x ∈ d(B1 ) ∩ · · · ∩ d(Bk ), with all the d(Bk ) distinct
(thus B1 , . . . , Bk are distinct). It follows that for each j = 1, . . . , k, x ∈ Gj
where f (Gj ) = Bj ; since B1 , . . . , Bk are distinct, so are G1 , . . . , Gk . Thus
x ∈ G1 ∩ · · · ∩ Gk ⊂ d(B1 ) ∩ · · · ∩ d(Bk ) ⊂ B1 ∩ · · · ∩ Bk .
Theorem 1.4 Any continuous map f : In → In has a fixed point, i.e. there
exists an x0 ∈ In such that f (x0 ) = x0 .
(We only use the notation diam(U ) when |U | would be ambiguous.) The mesh
size of a covering of A is the largest of the diameters of the elements of the
covering.
For two sets A, B ⊂ X we write
for the Hausdorff semidistance between A and B. Note that if B is closed then
dist(A, B) = 0 implies that A ⊆ B.
Theorem 1.5 Let I2 = [− 12 , 12 ]2 ⊂ R2 . Then dim(I2 ) ≥ 2.
Proof We want to show that any covering α of I2 with sufficiently small mesh
size contains at least three sets with nonempty intersection. To this end, take
a covering α with mesh size < 1 so that no element of the covering contains
points of opposite faces.
The first step is to construct a refinement α̃ of α consisting of closed, rather
than open, sets. To do this, observe that every x ∈ I2 is contained in some
Ux ∈ α, and we can find an open set Vx such that x ∈ Vx ⊂ V̄x ⊂ Ux . Since I2
is compact and {Vx : x ∈ I2 } is an open cover of I2 , there is a finite subcover
{Vxj }. We take α̃ to be the collection of all the closed sets {V̄xj }. By construction
this is a refinement of α consisting of closed sets.
We now show that α̃ contains at least three sets with nonempty intersection,
from which it is immediate (since α̃ is a refinement of α) that α contains at
least three sets with nonempty intersection.
Let Ŵ1 denote the side of I2 with x = − 21 , Ŵ1′ the side with x = 12 , Ŵ2 the
side with y = − 21 , and Ŵ2′ the side with y = 12 . Let L1 denote the union of
those elements of α̃ that intersect Ŵ1 ; L2 the union of those elements of α̃ that
are not in L1 and intersect Ŵ2 ; and let L3 be the union of all the other elements
of α̃ (those that intersect neither Ŵ1 nor Ŵ2 ). See Figure 1.1(a).
If we define K1 = L1 ∩ L3 then K1 separates Ŵ1 and Ŵ1′ in I2 , i.e. there exist
open sets U1 and U1′ such
I2 \ K1 = U1 ∪ U1′ , U1 ∩ U1′ = ∅
and Ŵ1 ⊂ U1 , Ŵ1′ ⊂ U1′ . The set K2′ = L1 ∩ L2 ∩ L3 separates Ŵ2 ∩ K1 from
Ŵ2′ ∩ K1 in K1 . One can then find a new closed set K2 , with K2 ∩ K1 ⊆ K2′ ,
that separates Ŵ2 and Ŵ2′ in I2 , i.e. such that there exist open sets U2 and U2′
such that
I2 \ K2 = U2 ∪ U2′ , U2 ∩ U2′ = ∅
and Ŵ2 ⊂ U2 , Ŵ2′ ⊂ U2′ . These constructions are illustrated in Figure 1.1(b). (If
the ‘proof by diagram’ of this last step is unconvincing, see IV.3 A) in Hurewicz
& Wallman (1941), or Exercise 1.3.)
Now for each x ∈ I2 , let v(x) be the 2-vector with components
⎧
⎨dist(x, Ki ) x ∈ Ui ,
⎪
⎪
vi (x) = 0 x ∈ Ki ,
⎪
⎪
⎩−dist(x, K ) x ∈ U ′ ,
i i
12 Lebesgue covering dimension
Γ2
L1 K1
Γ1
K2
K2
Γ1
L2
(a) Γ2 (b)
Figure 1.1 (a) A covering of I2 , divided into sets L1 (lightly shaded), L2 (more
heavily shaded), and L3 (not shaded). (b) K1 (lightly shaded) separates Ŵ1 and Ŵ1′
in I2 ; K2′ (a subset of K1 , shaded more heavily) separates K1 ∩ Ŵ2 and K1 ∩ Ŵ2′ in
K1 ; K2 (the dark line) separates Ŵ2 and Ŵ2′ in I2 , with K2 ∩ K1 ⊆ K2′ .
and set f (x) = x + v(x); note that f (x) ∈ I2 , and that f is continuous. It
follows from the Brouwer Fixed Point Theorem (Theorem 1.4) that f has
a fixed point, i.e. there exists an x0 ∈ I2 such that f (x0 ) = x0 . In particu-
lar, this implies that dist(x0 , K1 ) = dist(x0 , K2 ) = 0, i.e. that K1 ∩ K2 ⊂ K2′ =
L1 ∩ L2 ∩ L3 is nonempty. Since each of the original elements of α̃ is contained
in only one of the Lj s, there are three elements of α̃ that contain a common
point.
is a dense subset of X.
Lemma 1.7 A compact set A ⊆ (X, ̺) has dim(A) ≤ n if and only if it has
coverings of arbitrarily small mesh size and order ≤ n.
We will show that for each n ∈ N, the set of all 1/n-mappings is open and
dense, and our embedding result will then follow using the Baire Category
Theorem (Theorem 1.6) and the following simple lemma.
It is easy to check that this is the same as the affine space through x1 spanned
by {xj − x1 }kj =2 , i.e. all points of the form
k
x1 + cj (xj − x1 )
j =2
for all cj ∈ R (one could also form the same space by taking any xi and
considering all points of the form xi + j =i cj (xj − xi )). We say that the
points {x1 , . . . , xn } in RN are in general position in RN if no xj lies in the affine
space generated by any subcollection of the {xi } consisting of ≤ N elements
that does not contain xj .
An equivalent and more elegant definition makes use of the following
concept. We say that a set {y1 , . . . , yk } of k points in RN are geometrically
1.3 Embedding sets with finite covering dimension 15
independent if
k
k
aj yj = 0 and aj = 0,
j =1 j =1
for each j .
Now use Lemma 1.10 to select points {pj }rj =1 in R2n+1 such that
wi (x) = dist(x, X \ Ui ).
We now set2
r
g(x) = ϕi (x) pi . (1.4)
i=1
Since the only nonzero terms in the sum are for values of i for which x ∈ Ui ,
it follows from (1.2) and (1.3) that for such values of i, |pi − f (x)| < η and
hence
r
r
|g(x) − f (x)| = ϕi (x) (pi − f (x)) ≤ ϕi (x)|pi − f (x)| < η
i=1 i=1
for all x ∈ X.
2 In fact g maps X into an n-dimensional polyhedron. Since no more than n + 1 of the ϕk s are
nonzero at any one time, for every x ∈ Ui the image g(x) is contained in some fixed simplex
Si of dimension ≤ n. Then g(X) ⊂ ∪ri=1 Si , where the right-hand side defines a polyhedron of
dimension ≤ n. This remark will prove useful later in the proof of Theorem 2.12.
1.4 Large and small inductive dimensions 17
Exercises
1.1 Suppose that for every open cover {U1 , . . . , Un+2 } of X, there exists a
cover of X by closed sets {F1 , . . . , Fn+2 }, with Fj ⊆ Uj and ∩n+2
j =1 Fj = ∅.
Show that dim(X) ≤ n. [Hint: first show that the assumption implies that
the same is true with the {Fj } open.] (Theorem 3.2.1 in Edgar (2008)
shows that in fact the assumption here and dim(X) ≤ n are equivalent.)
1.2 Find a covering of R2 of order 3 and mesh size no larger than 1. Deduce that
any compact subset X of R2 has dim(X) ≤ 2. [Hint: any open covering of
X has a Lebesgue number that is strictly positive, see the proof of Lemma
1.7.]
1.3 Let A1 , A2 , and B be mutually disjoint subsets of a space X. We say that
B separates A1 and A2 in X if there exist two disjoint sets U1 and U2 ,
open in X, such that
A1 ⊂ U1 , A2 ⊂ U2 , and X \ B ⊂ U1 ∪ U 2 .
Now let A be a closed subset of X, C and C ′ a pair of disjoint closed
subsets of X, and K a closed subset of A that separates A ∩ C and A ∩ C ′
in A. Show that there exists a closed set B that separates C and C ′ in X
and satisfies A ∩ B ⊂ K.
1.4 Show that a one-to-one continuous mapping of a compact set is a homeo-
morphism.
Exercises 19
1.5 Assume that if M is a subspace of (X, ̺) with ind(M) ≤ 0 then given any
two open sets U1 and U2 that cover M, there exist disjoint open sets V1
and V2 with V1 ⊂ U1 and V2 ⊂ U2 such that V1 and V2 still cover M. Use
induction to show that if {U1 , . . . , Ur } is an open cover of M then there
exists an open cover {V1 , . . . , Vr } of M such that
Vj ⊂ Uj and Vi ∩ Vj = ∅ for i = j;
i.e. that ind(M) ≤ 0 implies that dim(M) = 0.
1.6 As mentioned immediately before the statement of Proposition 1.3, a
fundamental result in the theory of the small inductive dimension is that
a set A ⊆ (X, ̺) has ind(A) ≤ n if and only if it is the union of n + 1
subspaces of dimension ≤ 0. Use this result along with that of the previous
exercise to show that dim(A) ≤ ind(A).
1.7 Deduce from the following three facts that dim(X) = ind(X) for any sep-
arable metric space (reference is given to the relevant results in Hurewicz
& Wallman (1941)):
(i) any separable metric space X with dim(X) ≤ n can be embedded
n
into M2n+1 ∩ I2n+1 , the set of points in I2n+1 at most n of whose
coordinates are rational (Theorem V.5);
(ii) ind(M2n+1 ) = n (Example IV.1); and
(iii) A ⊆ B implies that ind(A) ≤ ind(B) (Theorem III.1).
2
Hausdorff measure and Hausdorff dimension
20
2.1 Hausdorff measure and Lebesgue measure 21
Note that any sets {Ui } are allowable in this cover of X (they need not be open).
Lemma 2.1 Hδs in an outer measure on (X, ̺) for each δ > 0.
Proof Fix δ > 0. Clearly (i) Hδs (∅) = 0 and (ii) Hδ(s) (A) ≤ Hδs (B) when-
ever A ⊆ B. To prove (iii) let {Aj }∞j =1 be a collection of subsets of X. Given
(i) ∞
ǫ > 0 there exists a sequence {Bj }i=1 of subsets of X such that
∞
∞
Aj ⊂ Bj(i) , |Bj(i) | ≤ δ, and |Bj(i) |s ≤ Hδs (Aj ) + ǫ2−j .
i=1 i=1
It follows that
⎛ ⎞
∞
∞
∞
∞
∞
Aj ⊂ Bj(i) and Hδs ⎝ Aj ⎠ ≤ |Bj(i) |s ≤ǫ+ Hδs (Aj ).
j =1 i,j =1 j =1 i,j =1 j =1
∞ ∞
Since this is valid for any ǫ > 0, Hδs j =1 Aj ≤ j =1 Hδs (Aj ) as
required.
One obtains the s-dimensional Hausdorff measure by refining the cover
involved in the definition of Hδs , i.e. taking the limit as δ → 0:
H s (X) = lim Hδs (X).
δ→0
The limit exists (it may be infinity) since Hδs (X) increases as δ decreases. It
follows immediately from Lemma 2.1 that H s is an outer measure; in fact
more is true.
Theorem 2.2 H s is a metric outer measure, i.e.
H s (A ∪ B) = H s (A) + H s (B) (2.2)
22 Hausdorff measure and Hausdorff dimension
which implies that L n (A) ≤ 2−n n H n (A). The lower bound relies on the
Vitali Covering Theorem, and since we only make use of the upper bound in
what follows we refer to Falconer (1985) for both the covering theorem (his
Theorem 1.10) and the proof of the lower bound (his Theorem 1.12).
Proof Suppose that H s (A) = 0. Then for any δ > 0 there is a cover of
A by sets {Uj } with |Uj | < δ such that (2.3) holds; in particular one such
cover exists. Conversely, given any δ > 0, choose ǫ > 0 such that ǫ 1/s < δ,
24 Hausdorff measure and Hausdorff dimension
and find a covering that satisfies (2.3). Then this must be a covering by sets
with |Uj | < δ that satisfies (2.3), and hence H s (A) < ǫ for every ǫ > 0, i.e.
H s (A) = 0.
We now prove some basic properties of the Hausdorff dimension.
Proposition 2.8
(i) If A, B ⊆ (X, ̺) and A ⊆ B then dH (A) ≤ dH (B);
(ii) the Hausdorff dimension is stable under countable unions: if Xk ⊆ X
then
∞
dH Xk = sup dH (Xk ); (2.4)
k=1 k
Then {f (Uj )} is a cover of f (X) and |f (Uj )| ≤ C|Uj |θ , from which it follows
that
|f (Uj )|s/θ < C s ǫ.
j
2.3 The Hausdorff dimension of products 25
Lemma 2.7 guarantees that H s/θ (f (X)) = 0, and hence dH (f (X)) ≤ s/θ .
Note that it is immediate from (ii) that the Hausdorff dimension of any
countable set is zero. In subsequent chapters we will frequently have recourse
to the example of an orthogonal sequence in a Hilbert space,
{aj ej }∞
j =1 ∪ {0}
Theorem 2.9 Let X be a closed subset of Rn . Then H s (X) > 0 if and only
if there exists a probability measure μ supported on X such that
Proof We only prove the ‘if’ part. Take any cover {Bri (xi )} of X with ri ≤ δ;
then
1 = μ(X) = μ X ∩ Bri (xi ) ≤ μ X ∩ Bri (xi ) ≤ c ris .
i i
dH (X × Y ) ≥ dH (X) + dH (Y ).
Proof Given s < dH (X) and t < dH (Y ), H s (X) > 0 and H t (Y ) > 0. It fol-
lows from Theorem 2.9 that there exist probability measures μ and ν supported
26 Hausdorff measure and Hausdorff dimension
The reverse inequality does not hold in general, as the following example
shows (Theorem 5.11 in Falconer (1985)). Let m0 = 1, mk+1 = k kj =0 mj .
Let X consist of those numbers in [0, 1] that have a zero in the rth decimal
place for mk + 1 ≤ r ≤ mk+1 and k even, and let Y consist of the numbers in
[0, 1] with a zero in the rth decimal place for mk + 1 ≤ r ≤ mk+1 and k odd.
Each of these sets X and Y has Hausdorff dimension zero. For X, consider
the first mk+1 decimal places, with k even; X can be covered by 10nk intervals
{Ij } of length 10−mk+1 , where
Then
k k
|Ij |1/k ≤ 10nk × 10−mk+1 /k ≤ 10 j =0 mj
× 10− j =0 mj
= 1,
j
1 = dH (0, 1) ≤ dH (f (X × Y )) ≤ dH (X × Y )
Hurewicz & Wallman (1941), is due to Edgar (2008, Theorem 6.3.11). A sim-
ilar argument, valid in any separable metric space, is given by Charalambous
(1999).
Note that an immediate consequence of this result is that dim(X) ≤ n for
any compact subset of Rn , since Proposition 2.8(iii) shows that dH (Rn ) = n,
and dH is monotonic (part (i) of the same proposition), so that dH (X) ≤ n for
any subset of Rn .
Theorem 2.11 Let X be a compact metric space. Then dim(X) ≤ dH (X).
Proof We use the characterisation of covering dimension from Exercise 1.1.
Let n = dim(X), so that it is not true that dim(X) ≤ n − 1. Then there must
exist an open cover {Ui }n+1
i=1 of X such that for any closed sets {Fi } with Fi ⊂ Ui
that still form a cover of X, ∩n+1
i=1 Fi = ∅.
Now define
δi (x) = dist(x, X \ Ui ) i = 1, . . . , n + 1
and δ(x) = δ1 (x) + · · · + δn+1 (x). Then each δi is Lipschitz continuous, and
hence so is δ:
|δi (x) − δi (y)| ≤ ̺(x, y) and |δ(x) − δ(y)| ≤ (n + 1)̺(x, y).
Since the {Ui } form a cover of X, x ∈ Ui for some i, and so δi (x) > 0; it
follows that δ(x) > 0 for every x ∈ X, and so since X is compact, there exist
b > a > 0 such that a ≤ δ(x) ≤ b for every x ∈ X. Define h : X → Rn+1 by
δ1 (x) δ2 (x) δn+1 (x)
h(x) = , ,··· , .
δ(x) δ(x) δ(x)
The function h is again Lipschitz, since
δj (x) δj (y) |δ(y)δj (x) − δ(x)δj (y)|
− =
δ(x) δ(y) δ(x)δ(y)
≤ a −2 [δ(y)|δj (x) − δj (y)| + δj (y)|δ(y) − δ(x)|]
≤ a −2 b(n + 2) ̺(x, y),
and so
√
|h(x) − h(y)| ≤ a −2 b(n + 2) n ̺(x, y).
Now, since h is Lipschitz, dH (h(X)) ≤ dH (X) (Proposition 2.8(iv)). The
proof is concluded by showing that h(X) contains the simplex
n+1
T = {(t1 , . . . , tn+1 ) ∈ Rn+1 : ti > 0 and ti = 1},
i=1
28 Hausdorff measure and Hausdorff dimension
Since ∩n+1
i=1 Fi = ∅, there exists an x ∈ X with δi (x)/δ(x) ≥ ti for each i. But
[δ
i i (x)/δ(x)] = 1 and i ti = 1, whence it follows that δi (x)/δ(x) = ti , i.e.
that h(x) = t, and so h(X) ⊇ T .
Of course, this inequality can be strict, since dim(X) is an integer-valued
definition of dimension, and there exist sets for which dH (X) ∈
/ N. However,
we always have equality for some homeomorphic image of X:
The proof is taken from Hurewicz & Wallman (1941, Theorem VII.4).
Proof Take s > n, and consider the collection Ks of all those functions f ∈
C(X, R2n+1 ) for which H s (f (X)) = 0. The condition that H s (f (X)) = 0
means that for each i ∈ N there exists a finite cover {Xj }kj =1 of X such that
k
|f (Xj )|s < 1/ i. (2.5)
j =1
Let X denote a finite cover {Xj }kj =1 of X, and denote by GX i,s the set of all
functions f ∈ C(X, R2n+1 ) that satisfy (2.5) for this decomposition. Then
⎡ ⎤
∞
Ks = ⎣ GX i,s .
⎦
i=1 all possible X
dH (f (X)) ≤ dH (polyhedron) ≤ n.
Since all such maps must therefore lie in Ks , Ks contains a dense subset of
C(X, R2n+1 ), so is itself dense.
We have shown that Ks is a dense Gδ in C(X, R2n+1 ), and Theorem 1.12
guarantees that the set of maps EX that are embeddings of X is also a dense
Gδ . It follows from the Baire Category Theorem (Theorem 1.6) that
EX ∩ Kn+(1/j )
j ≥1
Mandelbrot (1982) defined a ‘fractal’ as a set for which dim(X) < dH (X).
Luukkainen (1998) makes the nice comment that the result of this corollary
implies that ‘there is no purely topological reason for X to be fractal’.
Exercises
2.1 Let X be a subset of Rn , and let f : X → Rm satisfy
with C > 0 and θ ∈ [0, 1]. Show that the Hausdorff dimension of the
graph
then [0, 1] \ ∪Jq has s-dimensional Hausdorff measure zero. (This result
has applications to bounding the set of singular times in weak solutions of
the three-dimensional Navier–Stokes equations, see Scheffer (1976).)
2.3 Define the ‘d-dimensional spherical Hausdorff measure’ of a set X by
S d (X) = limδ→0 Sδd (X), where
∞
d d
Sδ (X) = inf ri : X ⊆ B(xi , ri ) : ri ≤ δ .
i=1
d
Show that dH (X) = inf{d : S (X) = 0}.
2.4 Suppose that X is a bounded subset of Rn that is covered by a collec-
tion of balls {B(x, r(x))}x∈X . Show that there exists a finite or countably
infinite disjoint subcollection of this cover, {B(xj , r(xj ))}∞
j =1 , such that
{B(xj , 5r(xj ))}∞
j =1 still covers X. [Hint: set M = sup x∈X r(x). Define
Use the results of the previous two exercises to show that H d (S) = 0.
[Hint: first show that L n (S) = 0.] (This result forms the final piece in
the proof of the partial regularity of the three-dimensional Navier–Stokes
equations due to Caffarelli, Kohn, & Nirenberg (1982).)
3
Box-counting dimension
The study of the box-counting dimension forms the core of Part I of this
book. We concentrate on the upper box-counting dimension (Definition 3.1),
since this is the least restrictive definition of dimension that allows one to
obtain a parametrisation of a ‘finite-dimensional’ set, using a finite number
of parameters, that has a well-defined degree of continuity (Hölder). But it
is also of interest since the upper box-counting dimension of many attractors
arising in the infinite-dimensional dynamical systems is finite. We explore the
implications of this fact in Part II.
As with the topological and Hausdorff dimensions, we give general results
for subsets of metric spaces; but as we switch to particular examples (and then
later embedding results) we specialise to subsets of Hilbert and Banach spaces.
log N(X, ǫ)
dbox (X) = lim ; (3.1)
ǫ→0 − log ǫ
31
32 Box-counting dimension
Definition 3.1 Let (X, ̺) be a metric space, and A ⊆ X. Let N(A, ǫ) denote
the minimum number of closed balls of radius ǫ with centres in A required to
cover A. The upper box-counting dimension of A is
log N(A, ǫ)
dB (A) = lim sup , (3.2)
ǫ→0 − log ǫ
and the lower box-counting dimension of A is
log N(A, ǫ)
dLB (A) = lim inf .
ǫ→0 − log ǫ
The inequality dLB (A) ≤ dB (A) is clear, with the above Cantor-like set show-
ing that it can be strict. The ‘box-counting dimension’ (3.1) only exists when the
lower and upper box-counting dimensions coincide; but since we will (almost)
always be interested in the upper box-counting dimension in what follows, we
will usually refer to the quantity dB (X) defined in (3.2) as the box-counting
dimension.1 (We will see in Section 8.2 that sets with finite lower box-counting
dimension do not enjoy the same embedding properties as sets with finite upper
box-counting dimension.)
It is immediate from the definition that if d > dB (A) then there exists an ǫ0
such that
1 In much of the dynamical systems literature the upper box-counting dimension is referred to as
the ‘fractal dimension’; although a little inelegant, ‘box-counting dimension’ is to be preferred
for obvious reasons.
3.2 Basic properties of the box-counting dimension 33
vary depending on the space in which one views X as lying. In the final chapter
of this book, the following simple observation will prove useful: if B1 and B2
are two Banach spaces with X ⊂ B1 ⊆ B2 , then
uB2 ≤ cuB1 ⇒ dB (X; B2 ) ≤ dB (X; B1 ) (3.5)
(the proof is immediate since NB2 (X, cǫ) ≤ NB1 (X, ǫ)).
3 This is not in fact a basis for ℓ∞ , but only for c0 , the subspace of ℓ∞ consisting of sequences
that tend to zero.
3.4 Orthogonal sequences 37
Furthermore,
log n
dLB (A) = lim inf . (3.12)
n→∞ − log |an |
We will use (3.10) and (3.12) immediately to consider some simple exam-
ples. The alternative form of (3.10), (3.11), will be useful later.
Proof First we prove (3.10). Given any ǫ with 0 < ǫ < |a1 |, let n = n(ǫ) be
the integer such that
The set A can be covered by n + 1 ǫ-balls, one centred at the origin and the
other n centred at {α1 , . . . , αn }. Thus
log N(X, ǫ) log n(ǫ) + 1 log n
dB (A) = lim sup ≤ lim sup ≤ lim sup .
ǫ→0 − log ǫ ǫ→0 − log |an(ǫ) | n→∞ − log |an |
The upper bound in (3.12) follows similarly.
To prove the reverse inequality, for any n large enough that |an | < 1, let n′
denote the integer n′ ≥ n for which
and set ǫ(n) = 41 (|an′ | + |an′ +1 |). It follows that any two elements from
{α1 , . . . , αn′ } are at least |an′ | > 2ǫ(n) apart (in any ℓp norm), and hence
N(A, ǫ(n)) ≥ n′ .
Since n′ ≥ n, |an | = |an′ |, and |an′ | < 4ǫ(n), it follows that
log n log n′ log N(A, ǫ(n))
≤ ≤ ,
− log |an | − log |an′ | − log(4ǫ(n))
and hence that
log n log N(A, ǫ(n)) log N(A, ǫ)
lim sup ≤ lim sup ≤ lim sup = dB (A).
n→∞ − log |an | n→∞ − log(4ǫ(n)) ǫ→0 − log(4ǫ)
Again, the lower bound in (3.12) follows similarly.
We now show that the right-hand sides of (3.10) and (3.11), which we call
d1 and d2 respectively, are equal. Take d > d2 , so that ∞ d
n=1 |an | = M. Then
d
since |an | is nonincreasing, this implies that n|an | ≤ M for any n, from which
it is easy to see that d1 ≤ d, and hence d1 ≤ d2 . Conversely, if d > d1 then for
all n sufficiently large, log n/(− log |an |) ≤ d, and so |an | ≤ n−1/d . It follows
that for any d ′ > d, ∞ d′ ′ ′
n=1 |an | < ∞. Thus d2 ≤ d for all d > d1 , and so
d2 ≤ d1 . It follows that d1 = d2 .
38 Box-counting dimension
Hα = {0} ∪ {n−α en }∞
n=1 .
Then
log n 1
dB (Hα ) = lim sup = . (3.13)
n→∞ α log n α
More strikingly (since all these examples have zero Hausdorff dimension), the
set
has
log n
dB (Hlog ) = lim sup = +∞.
n→∞ log log n
By combining these two examples, one can obtain a set with dB (Ĥ ) = ∞ but
dLB (Ĥ ) < ∞: the idea is to choose the coefficients {an } such that |an+1 | ≤ |an |
and there exist sequences nj → ∞ such that anj = 1/nj and mj → ∞ such
that amj = 1/ log mj . In more detail, define
e(x) = ⌊ex ⌋,
a1 = a2 = 1,
a3 = 1/(log 3),
a4 = a5 = · · · = ae(4) = 1/4,
ae(4)+1 = 1/ log(e(4) + 1),
ae(4)+2 = · · · = ae(e(4)) = 1/(e(4) + 2),
ae(e(4))+1 = 1/ log(e(e(4)) + 1),
etc. Then
log n
dB (Ĥ ) = lim sup = ∞,
n→∞ − log |an |
but
log n
dLB (Ĥ ) = lim inf = 1.
n→∞ − log |an |
Exercises 39
Exercises
3.1 Show that
log M(A, ǫ)
dB (A) = lim sup ,
ǫ→0 − log ǫ
if M(A, ǫ) denotes:
(i) the minimum number of closed balls of radius ǫ with arbitrary centres
that are required to cover A;
(ii) the largest number of disjoint balls of radius ǫ with centres in A; or
(iii) for a subset of Rn , the number of boxes of the form
is open, where No (X, ǫ) is the number of open balls of radius ǫ that covers
X. Then use the characterisation
Note that dH (X) ≤ dMB (X) since dH (X) ≤ dB (X) and the Hausdorff
dimension is stable under countable unions (2.4). Let X be a compact
subset of a Hilbert space H , and suppose that
dB (X ∩ U ) = dB (X)
for all open subsets U of H that intersect X. Use the Baire Category
Theorem (Theorem 1.6) to show that dMB (X) = dB (X).
3.7 Set
Pδs (X) = sup{ |Bi |s : {Bi } are disjoint balls with centres in X}
i i
and define
P0s (X) = lim Pδs (X).
δ→0
X − X = {x − y : x, y ∈ X}.
Lx = (L1 x, L2 x, . . . , Lk x);
41
42 An embedding theorem for subsets of RN
We will consider a restricted set E of linear maps, namely those of the form
E = {(l1∗ , . . . , lN∗ ) : lj ∈ BN },
where√BN = BN (0, 1) is the unit ball in RN . Note that any L ∈ E has norm at
most N.
Identifying E with (BN )k , we define a probability measure μ on E to be that
induced by choosing each lj according to the uniform probability measure λ on
BN (λ is the Lebesgue measure L N normalised by N ), i.e. μ is the product
measure ⊗kj =1 λ on (BN )k .
The following estimate lies at the heart of the proof of the embedding
theorem of this chapter (Theorem 4.3).
Lemma 4.1 For any α ∈ Rk and x ∈ RN ,
k
ǫ
μ{L ∈ E : |α + Lx| ≤ ǫ} ≤ cN k/2 , (4.1)
|x|
where c is an absolute constant.
Proof Let α = (α1 , . . . , αk ). Then
k
#
μ{L ∈ E : |α + Lx| ≤ ǫ} ≤ μ{L ∈ E : |αj + Lj x| ≤ ǫ}
j =1
k
#
= λ{l ∈ BN : |αj + (l · x)| ≤ ǫ},
j =1
Proof Consider
Q = ∩∞ ∞
n=1 ∪j =n Qj .
Theorem 4.3 Let X be a compact subset of RN . If k > 2dB (X) then given
any α with
2d
0<α <1−
k
and any linear map L0 ∈ L (RN , Rk ), for μ-almost every linear map L ∈ E
there exists a C = CL such that L′ = L0 + L satisfies
With L0 = 0 the theorem says that μ-almost every L′ ∈ E satisfies (4.3); but
the slight strengthening here is the key idea in the notion of ‘prevalence’ which
is defined in the next chapter and allows for similar results when X is a subset
of an infinite-dimensional space (Theorem 8.1). We prove a generalised version
of this theorem for subsets of RN using a wider class of mappings from RN
into Rk (but without the Hölder continuity of the inverse) in Lemma 14.4. The
proof of the result in this form is due to Hunt & Kaloshin (1999), but earlier
results along these lines, using density instead of prevalence, can be found in
Ben-Artzi et al. (1993) and Eden et al. (1994).
Zn = {z ∈ X − X : |z| ≥ 2−n }
and set
This Qn is essentially the set of ‘bad’ linear maps for which (4.3) (with C = 1)
does not hold for some (x, y) with |x − y| ≥ 2−n .
We now use the fact that dB (X − X) ≤ 2dB (X). Choose and fix d > dB (X);
then Zn ⊂ X − X can be covered by a collection of no more than N := 22nd/α
balls of radius 2−n/α , {B(zj , 2−n/α )}, whose centres zj lie in Zn .
Let Yj = Zn ∩ B(zj , 2−n/α ). Now note that if
√
|(L0 + L)zj | > 2−n/α (1 + 2 k + 2L0 )
then
n/α
n/α
n/α
Figure 4.2 If |(L0 + L)zj | > 2−n/α M then |(L0 + L)z| > 2−n/α for every z ∈ Y .
Thus the total measure of Qn , i.e. those maps for which things fail for some
z ∈ Zn , is bounded by
′
μ(Qn ) ≤ 22nd/α · CN,k,L 0
′
2nk(1−(1/α)) = CN,k,L 0
· 2[k−(k−2d)/α]n .
The term ‘prevalence’ was coined by Hunt et al. (1992), for a generalisation of
the notion of ‘almost every’ that is appropriate for infinite-dimensional spaces.
Essentially the same definition was used earlier by Christensen (1973), although
for him a set was prevalent if its complement was a Haar null set; we adopt here
the more recent and more descriptive terminology. A nice review of the theory
of prevalence is given by Ott & Yorke (2005). We only develop the theory here
as far as we will need it in what follows; more details can be found in the above
papers and in Benyamini & Lindenstrauss (2000, Chapter 6).
Once we have introduced prevalence, we show how the idea can be adapted to
treat certain classes of linear maps from infinite-dimensional spaces into finite-
dimensional Euclidean spaces (Section 5.2), and then prove a generalisation of
the inequality (4.1) that is a key element of the subsequent embedding proofs.
5.1 Prevalence
Let V be a normed linear space. First we define what it means for a subset
of V to be ‘shy’, the equivalent in this setting of ‘having measure zero’; the
complement of a shy set is said to be ‘prevalent’.
Definition 5.1 A Borel set S ⊂ V is shy if there exists a compactly supported
probability measure1 μ on V such that
μ(S + v) = 0 for every v ∈ V. (5.1)
More generally, a set is shy if it is contained in a shy Borel set.
1 Hunt et al. (1992) in fact make what initially appears to be a weaker definition: there need only
exist some measure μ such that 0 < μ(U ) < ∞ for some compact set U , for which (5.1) holds.
They then, however, make the observation that given such a measure one can always take instead
an appropriately weighted restriction of μ to U to obtain a compactly supported probability
measure for which (5.1) still holds.
47
48 Prevalence, probe spaces, and a crucial inequality
It is easy to show that in Rn a set is shy if and only if it has measure zero.
Proof Since subsets of Borel sets with Lebesgue measure zero also have
Lebesgue measure zero, and the same is true of ‘shyness’, we need only consider
Borel sets. If a Borel set S has Lebesgue measure zero then one can take μ to
be Lebesgue measure on the unit ball in Rn (weighted by the inverse of the
volume of the ball), and clearly μ(S + v) = 0 for every v ∈ Rn .
Conversely, let S be a Borel set and suppose that there exists a compactly
supported probability measure μ such that (5.1) holds for every v ∈ Rn ; let ν
be Lebesgue measure. Then by the Tonelli Theorem
" "
0= μ(S − y) dν(y) = ν(S − x) dμ(x) = ν(S)μ(Rn ) = ν(S),
Rn Rn
Proof We show that the union of two shy sets is shy, and the result then
follows by induction. To this end, given two shy sets S ′ and T ′ , find shy Borel
sets S and T that contain them, with corresponding probability measures μ
and ν.
Let μ × ν be the product measure on V × V , and for a Borel set S ⊂ V
define
S = {(x, y) ∈ V × V : x + y ∈ S}.
μ ∗ ν(S) = (μ × ν)(S ).
5.2 Measures based on sequences of linear subspaces 49
Since
" "
μ ∗ ν(S) = μ(S − y) dν(y) = ν(S − x) dμ(x),
V V
it follows that
With a little more work one can show that the countable union of shy sets is
shy, and so the countable intersection of prevalent sets is prevalent. We will not
require this (potentially powerful) result in what follows; a proof can be found
in Hunt et al. (1992, Fact 3′′ ), Ott & Yorke (2005, Axiom 3), or Benyamini &
Lindenstrauss (2000, Proposition 6.3).
Proof We identify Vj with Rdj and Sj with Bdj in the obvious way. Noting
that for v ∈ Sj we have (v, x) = (v, Pj x), the estimate follows immediately
from (4.2).
Given the result of the previous lemma, the following key estimate is rela-
tively straightforward.
∞ −γ
Lemma 5.5 shows that for α = f0 (x) + i =j i (φi , x) fixed, the bound on
and the inequality (5.3) now follows from the estimate (5.2).
In the proof of Theorem 9.18 we will require a more refined result. We
specialise to the case in which the {Vj } are mutually orthogonal, and dim(Vj ) ≤
d for every j . Rather than using Sj , the unit ball in Vj , in our construction of
E, we instead use a ‘unit cube’ Cj , where
Cj = {u ∈ Vj : |(u, ej,i )| ≤ 21 , i = 1 . . . , j },
dj
with {ej,i }i=1 an orthonormal basis for Vj . The measure on Cj is now induced
by Lebesgue√measure on Idj := [− 21 , 12 ]dj . Since any element of Cj has norm
bounded by d, we can use the orthogonality of the {Vj } to allow any γ > 1/2
in the definition of E.
We will require the following result of Ball (1986) about the volume of
(d − 1)-dimensional slices through the unit cube in Rd . The key point is that
the upper bound (which is sharp) does not depend on the dimension d. (Hensley
(1979) proved a similar result but with the upper bound 5.)
Lemma 5.8 In the situation described above, given any x ∈ H and any
f ∈ L (H, Rk ), for any j
k
ǫ
μ{ L ∈ E : |(L + f )(x)| < ǫ } ≤ c j γ d 1/2 , (5.4)
j x
where c is a constant independent of j and f , and j is the orthogonal
projection onto V1 ⊕ V2 ⊕ · · · ⊕ Vj .
Proof Arguing as in the proof of Lemma 5.6, the left-hand side of (5.4) is
bounded by
⎡ ⎤
)n n
# n
⎣ λj ⎦ {(φ1 , . . . , φn ) ∈ Cj : j −γ φj∗ (j x) < ǫ}.
j =1 j =1 j =1
5.2 Measures based on sequences of linear subspaces 53
and let μ denote the uniform probability measure on ID (i.e. Lebesgue measure),
the problem is to bound, for any y ∈ R,
1
μ{x ∈ ID : |y + (x · a ′ )| ≤ ǫ} = μ{x ∈ ID : |y + (x · â)| ≤ ǫ}
|a ′ |
ns
≤ μ{x ∈ ID : |y + (x · â)| ≤ ǫ},
|a|
where â = a ′ /|a ′ |. The result is now a consequence of Theorem 5.7, since
√
μ{x ∈ ID : |y + (x · â)| ≤ ǫ} ≤ 2ǫ|(Sâ − y â) ∩ ID | ≤ 2ǫ 2,
where Sâ is the hyperplane through the origin with normal â.
Proof Write ρ for the left-hand side of (5.5). If g(x) = 0 then the inequality
is trivially true. So assume that g(x) = 0, and let P be the subspace of B ∗ that
annihilates x.
If h is any other element of Sj with h(x) = 0 then since
[g(x)h − h(x)g](x) = 0,
it follows that g(x)h = h(x)g + p for some p ∈ P . One can therefore write any
element of Sj in the form p + rg for some p ∈ P and r ∈ R. That this expansion
is unique can be seen easily by applying both sides of p1 + r1 g = p2 + r2 g
to x.
Now, ρ is bounded above by the probability that φ ∈ Sj lies between
r ǫ r ǫ
− − g+P and − + g + P.
g(x) |g(x)| g(x) |g(x)|
If P is represented by the hyperplane in Rdj , and g by the vector γ , then
by definition this is the fraction of the measure of Uj that lies between
Uj ∩ ( + sγ ) = Ks . (5.6)
Π γ
K+
θ
Uj
K0
K−
is a convex symmetric subset of Rn whose volume is n−1 /n, and hence the
ratio of the largest (n − 1)-dimensional ‘slice’ through the origin (n−1 ) to the
volume is precisely n.
We now follow the argument of Lemma 5.6, using the estimate (5.5), to
obtain the Banach-space version of (5.3).
56 Prevalence, probe spaces, and a crucial inequality
Exercises
11
5.1 Show that 0 f (x) dx = 0 for a prevalent set of functions in f ∈ L1 (0, 1).
5.2 Show that Eγ (V ) is a compact subset of L (H, Rk ).
5.3 The Brunn–Minkowski Inequality (see Gardner (2002), for example) says
that if L and M are two convex subsets of Rn then
L n ((1 − t)L + tM)1/n ≥ (1 − t)L n (L)1/n + tL n (M)1/n
for t ∈ [0, 1]. Use this to show that the map s → L dj −1 (Ks )1/(dj −1) is
concave, where Ks is defined in (5.6), and deduce that L dj −1 (Ks ) attains
its maximal value when s = 0.
6
Embedding sets with dH (X − X) finite
We now give the first application of the constructions of the previous chapter
to prove a ‘prevalent’ version of a result first due to Mañé (1981). He showed
that if X is a subset of a Banach space B and dH (X − X) < k, then a residual
subset of the space of projections onto any subspace of dimension at least k are
injective on X.
We show here that in general no linear embedding into any Rk is possible
if we only assume that dH (X) is finite (Section 6.1). If we want an embedding
theorem for such sets, we must fall back on Theorem 1.12 which guarantees
the existence of generic embeddings of sets with finite covering dimension (we
can apply this result since dim(X) ≤ dH (X) by Theorem 2.11).
While we prove in Theorem 6.2 the existence of a prevalent set of linear
embeddings into Rk when dH (X − X) < k, we will see that even with this
assumption one cannot guarantee any particular degree of continuity for the
inverse of the linear mapping that provides the embedding (Section 6.3).
In this chapter and those that follow, we will often wish to show that certain
embedding results are sharp, in the sense that the information we obtain on
the modulus of continuity for the inverse of the embedding map cannot be
improved. In this context, the following decomposition lemma, which allows
us to reduce the analysis of general linear maps to the analysis of orthogonal
projections, is extremely useful.
Lemma 6.1 Let H be a Hilbert space and suppose that L : H → Rk is a
linear map with L(H ) = Rk . Then U = (ker L)⊥ has dimension k, and L can
be decomposed uniquely as MP , where P is the orthogonal projection onto U
and M : U → Rk is an invertible linear map.
Proof Let U = (ker L)⊥ and suppose that there exist m > k linearly indepen-
dent elements {xj }m m
j =1 of U for which Lxj = 0. Then {Lxj }j =1 are elements of
k
R ; since m > k at least one of the {Lxj } can be written as a linear combination
57
58 Embedding sets with dH (X − X) finite
of the others:
Lxi = cj (Lxj ).
j =i
It follows that
% &
L xi − cj xj = 0,
j =i
and hence
xi − cj xj ∈ ker L ∩ (ker L)⊥ = {0}.
j =i
2
(e.g. Mk = 0 and Mk = 2k for k ≥ 1). The set C can be covered by 2rk intervals
of length 2−M2k+1 , where
k
rk = k + (M2j − M2j −1 ) = k + M2k − 1.
j =1
The set A is given as the union of m sets Aj , each lying on a face of the unit
m-cube. Aj consists of points a = (a1 , . . . , am ) with aj = 0 and ai for i = j
an element of the Cantor set C constructed above.
Since for i = j , the one-dimensional orthogonal projection of Aj onto the
ith coordinate axis is precisely C, it follows that Aj can be covered by 2(m−1)rk
cubes whose edges have length 2−M2k+1 . It is easy to see that for any s > 0,
2(m−1)rk [2−M2k+1 ]s → 0
0 = Pv = Pb − Pa ⇒ P b = P a,
Clearly a ∈ Aj , b ∈ Bj , and v = b − a.
Given an infinite-dimensional Hilbert space H , take a countable orthonormal
set {ej }∞ ′
j =1 , and let Km be the subset of H obtained from Km by identifying the
coordinate axes of R with {ej }m
m
j =1 . Set
∞
K = {0} ∪ 2−m Km′ .
m=1
using the decomposition lemma (Lemma 6.1) this yields a rank k projection in
Rm that is injective on Km , a contradiction.
In the light of the result of Theorem 6.2, the set K has dH (K) = 0 but
dH (K − K) = ∞.
Theorem 6.2 Let X be a compact subset of a real Banach space B such that
dH (X − X) < k, where k is a positive integer. Then a prevalent set of linear
maps L : B → Rk are one-to-one between X and its image.
Zn = {z ∈ X − X : z ≥ 2−n }
be the set of all linear maps in E for which f + L fails to be injective for some
pair x, y ∈ X with x − y ≥ 2−n .
We will now show that ∞ n=1 μ(Qn ) = 0.
Choose δ > 0, and for each n (which is taken to be fixed for this portion of
the argument) cover Zn with a collection of balls B(zj , ǫj ) such that
−1
ǫjk < 2−n δ dnk n2k 2nk , (6.1)
j
whence
Now,
∞
Zn = (X − X) \ {0},
n=1
and so
∞
Qn = {L ∈ E : (f + L)(z) = 0 for some nonzero z ∈ X − X}
n=1
62 Embedding sets with dH (X − X) finite
Proof Suppose that P has rank k. Then there exists an orthonormal basis
{u1 , . . . , uk } for P H , so that for any x ∈ H ,
k
Px = (x, uj )uj .
j =1
k
In particular, P ei = j =1 (ei , uj )uj , so that
k
k
P ei 2 = (P ei , P ei ) = (P ei , ei ) = (ei , uj )(uj , ei ) = |(ei , uj )|2 .
j =1 j =1
It follows that
∞
∞
k ∞
k k
P ei 2 = |(ei , uj )|2 = |(ei , uj )|2 ≤ uj 2 = k,
i=1 i=1 j =1 j =1 i=1 j =1
Now given any choice of f , set φn = nf (1/n), let Nn be the first integer
j
greater than or equal to 1/φn , and define Tj = n=1 Nn ; for Tj ≤ i ≤ Tj +1 set
αi = 1/j . This gives an orthogonal sequence X for which the right-hand side
of (6.4) is infinite, and hence no finite-rank orthogonal projection can satisfy
(6.3).
Since 0 ∈ X, X ⊂ X − X; so there can be no finite-dimensional projection
P for which
P (x1 − x2 ) ≥ ǫf (x1 − x2 ) for all x1 , x2 ∈ X,
for any value of ǫ > 0. It follows from the decomposition lemma (Lemma 6.1)
that if one can rule out such a modulus of continuity for orthogonal projections,
the same follows for more general finite-rank linear maps.
Note that this argument also shows that one cannot prove a better embedding
theorem than Theorem 6.2 if one strengthens the assumption to one on the
modified box-counting dimension introduced in Exercise 3.6: all the above
examples are countable sets, and so have modified box-counting dimension
zero.
7
Thickness exponents
dist(X, Vj ) ≤ 2−j /3
then, recalling that in the proof x was an element of the set of differences
X − X, i.e. x = x1 − x2 with x1 , x2 ∈ X, it follows that
While this gives a lower bound on Pj x of the required form, the dimension
of Vj occurs in the estimate (7.1). In order to carry the argument through
successfully, we will need some control on how the dimension of Vj grows
with j . This is provided by the thickness exponent, τ (X), introduced by Hunt
64
7.1 The thickness exponent 65
& Kaloshin (1999), and discussed in Section 7.1. This exponent can be shown
to be zero when the set X consists of C ∞ functions, see Lemma 13.1.
A related quantity which can be defined for subsets of Hilbert spaces is the
Lipschitz deviation dev(X), covered in Section 7.2. Introduced by Olson &
Robinson (2010) and refined further by Pinto de Moura & Robinson (2010b),
this can replace the thickness exponent in the generalised (infinite-dimensional)
version of Theorem 4.3, and can be shown to be zero for the attractors arising in
the infinite-dimensional dynamical systems generated by a number of canonical
partial differential equations (Section 13.2).
Finally, in Section 7.3 we define a version of the thickness, the ‘dual thick-
ness’ τ ∗ (X), appropriate for subsets of Banach spaces. In a Hilbert space
τ ∗ (X) ≤ dev(X) ≤ τ (X); it is not clear how the thickness and dual thickness
are related for subsets of Banach spaces, but one can prove the useful result
that τ (X) = 0 implies that τ ∗ (X) = 0 (Proposition 7.10).
distB (X, V ) ≤ ǫ,
We will usually drop the space B from the notation in what follows, pre-
ferring the simpler d(X, ǫ) and τ (X). But note that, as with the box-counting
dimension, the definition depends on the space in which we consider X. We note
here for use later that if B1 and B2 are two Banach spaces with X ⊂ B1 ⊆ B2 ,
then
Proof Given ǫ > 0, cover X with N(X, ǫ) balls of radius ǫ. Then every point
of X lies within ǫ of the linear subspace V that is spanned by the centres of
these balls. (This is essentially the way that the idea was used by Foias & Olson
(1996).) Since the dimension of V is no greater than N(X, ǫ), this implies that
d(X, ǫ) ≤ N(X, ǫ) and the lemma follows.
We now show that for the example of an orthogonal sequence in a Hilbert
space (as considered in Lemma 3.5), the thickness is in fact equal to the box-
counting dimension. To show this we will require the following lemma due to
M. Doré (personal communication).
d(X, ǫ) ≥ n(1 − ǫ 2 /M 2 ),
Proof If d(X, ǫ) = d then there exist vi′ ∈ H such that vi′ − vi < ǫ, and
such that the space spanned by {v1′ , . . . , vn′ } has dimension d. Let P be the
orthogonal projection onto U , the n-dimensional space spanned by {v1 , . . . , vn }
and let vi′′ = P vi′ . Since P vi = vi we still have the inequality vi′′ − vi < ǫ
and clearly the dimension of the linear span of {v1′ , . . . , vn′ } is at least that of
the linear span of {v1′′ , . . . , vn′′ }.
Suppose that the linear span of {v1′′ , . . . , vn′′ } has dimension n − r. We can
write any element of U in terms of the {vj′′ } and an orthonormal basis for their
r-dimensional orthogonal complement in U , {u1 , . . . , ur }. So
n
n
r
nǫ 2 ≥ vi′′ − vi 2 ≥ |(vi , uj )|2
i=1 i=1 j =1
r
n 2
vi
= vi 2 (uj , )
j =1 i=1
vi
r
n 2
vi
≥ M2 (uj , ) = M 2 r.
j =1 i=1
vi
since devm (X) is nonincreasing in m the limit clearly exists provided that
devm (X) is finite for some m > 0.
Note that dev(X) is bounded above by τ (X), since devm (X) ≤ τ (X) for
every m > 0 (one can always approximate by the graph of the zero function,
which is m-Lipschitz). We now show, following Pinto de Moura & Robinson
(2010b), that this inequality can be strict.
and so
e1 e1
|φ(e1 /n) − φ(e1 /m)| ≤ 3 − .
n m
The function φ can be extended to a 3-Lipschitz function defined on the whole
of E1 (see Wells & Williams (1975), for example). It follows that dev3 (X) = 0,
and so dev(X) = 0.
We now show that τ (X) ≥ 1, following the argument used above to prove
Lemma 7.4. For n ≥ 1 set
e1 en+1
an = + ;
n + 1 (n + 1)2
note that an ≥ an+1 and limn→∞ an = 0. Let X = {a1 , a2 , . . .}. Set ǫn2 =
(an 2 + an+1 2 )/4. Since aj 2 ≥ an 2 > 2ǫn2 for j = 1, . . . , n, it follows
from the above lemma that
ǫn2 n
d(X, ǫn ) ≥ d({a1 , . . . , an }, ǫn ) ≥ n 1 − 2
≥ .
an 2
Definition 7.5 Given θ > 0, let nθ (X, ǫ) denote the lowest dimension of any
linear subspace V of B ∗ such that for any x, y ∈ X with x − y ≥ ǫ there
exists an element ψ ∈ V such that ψ∗ = 1 and
Set
log nθ (X, ǫ)
τθ∗ (X) = lim sup ,
ǫ→0 − log ǫ
∗
and define the dual thickness τ (X) by
Definition 7.6 Given θ > 0, let mα (X, ǫ) denote the lowest dimension of any
linear subspace V of B ∗ such that for any x, y ∈ X with x − y ≥ ǫ there
exists an element ψ ∈ V such that ψ∗ = 1 and
Define
log mα (X, ǫ)
σα∗ (X) = lim sup .
ǫ→0 − log ǫ
It would now be natural to define σ ∗ (X) = limα→0 σα∗ (X); this gives another
possible definition of a ‘dual thickness’, but with the more tortuous definition of
τ ∗ (which is never larger than σ ∗ , see below) we can still prove an embedding
theorem, and more importantly we can show that zero thickness (in the sense of
Hunt & Kaloshin’s definition) implies zero dual thickness (Proposition 7.10);
this does not seem to be possible using σ ∗ .
The following lemma shows that σα∗ (X) provides an upper bound for τ ∗ (X)
for any α > 0.
for all ǫ small enough that ǫ θ < α, and so τθ∗ (X) ≤ σα∗ (X) for all θ > 0.
The following simple corollary shows that τ ∗ (X) ≤ dB (X), i.e. that the dual
thickness is well adapted for use with sets that have finite box-counting dimen-
sion. Corollary 8.4 shows that there are sets for which this upper bound is
attained, so τ ∗ (X) is not always zero (there are possible definitions of ‘thick-
ness exponents’ that one might expect to be useful but which turn out to be
7.3 Dual thickness 71
zero whenever X has finite box-counting dimension, see Exercise 7.2 for one
example).
Proof Take d > dB (X). Then there exists an ǫ0 > 0 such that for all ǫ < ǫ0 ,
X can be covered with a collection B(xj , ǫ/12) of N ≤ ǫ −d balls, where the
xj are chosen to be linearly independent. To see that this is possible, first cover
X with a collection of balls B(zj , ǫ/13). This is a finite collection; since B is
infinite-dimensional one can perturb each zj in turn to some xj such that the
resulting collection {x1 , . . . , xn } is linearly independent for each n ≤ N. One
can then enlarge slightly the radius of each ball.
Now use the Hahn–Banach Theorem to define a collection of linear func-
tionals ψj with the property
1 ≤ ψj − ψk ∗ ≤ 2,
Exercises
7.1 Let X be a subset of a Banach space B, and denote by ε(X, n) the minimum
distance between X and any n-dimensional linear subspace of B. Show
74 Thickness exponents
that
log n
τ (X) ≤ lim sup . (7.4)
n→∞ − log ε(X, n)
(One can in fact prove equality here, see Lemma 2 in Kukavica & Robinson
(2004).)
7.2 One could try to define another ‘thickness measure’ for a set X ⊂ H as
follows. For each ǫ > 0, let dLE (X, ǫ) be the smallest n such that there
exists a 2-Lipschitz map φ : Rn → H such that dist(X, φ(Rn )) < ǫ. Define
log dbL (X, ǫ)
τLE (X) = lim sup .
ǫ→0 − log ǫ
The Johnson–Lindenstrauss Lemma (Johnson & Lindenstrauss, 1984)
guarantees that given a set of m points in a Hilbert space H , and an
n > O(ln m), there is a function f : H → Rn such that
1
2
u − v ≤ |f (u) − f (v)| ≤ 2u − v.
Show that τLE (X) = 0 for any set X with dB (X) finite.
7.3 Show that if U is a finite-dimensional Banach space dim(U ) = n then
there exists an ‘Auerbach basis’ for U : a basis {e1 , . . . , en } for U and cor-
responding elements {f1 , . . . , fn } of U ∗ such that ej = fj ∗ = 1 for
j = 1, . . . , n and fi (ej ) = δij , i, j = 1 . . . , n. [Hint: by identifying U with
(Rn , · ) one can work in Rn . For x1 , . . . , xn ∈ Rn let det(x1 , . . . , xn )
denote the determinant of the n × n matrix with columns formed by the
vectors {xj }. Choose {e1 , . . . , en } with ej = 1 such that det(e1 , . . . , en )
is maximal. Define candidates for the {fj } and check that the fj satisfy
the required properties.]
7.4 Use the result of the previous exercise to show that if U is any n-
dimensional subspace of a Banach space B, there exists a projection
P onto U whose norm is no larger than n, P ≤ n.
8
Embedding sets of finite box-counting
dimension
τ (X) = 0 ⇒ τ ∗ (X) = 0
(Proposition 7.10).
75
76 Embedding sets of finite box-counting dimension
that
x − y ≤ CL |Lx − Ly|θ for all x, y ∈ X. (8.2)
In particular, L is injective on X.
Foias & Olson (1996) first proved a result along these lines. They showed
that in a Hilbert space there is a dense set of orthogonal projections that are
injective on X and have a Hölder inverse, but did not give any explicit bound
on the Hölder exponent. The proof of the theorem given here, almost identical
to that of Theorem 4.3, is due essentially to Hunt & Kaloshin (1999), but
incorporates the generalised estimate of Lemma 5.10 and the dual thickness,
following Robinson (2009). Separating the more geometric elements of the
proof (the estimates contained in Section 5.2) serves to clarify the argument.
Proof If θ satisfies (8.1), then there exist β > 0, σ > τβ∗ (X), and δ > d such
that
k − 2δ
0<θ < . (8.3)
k(1 + β + ασ )
Since σ > τβ∗ (X), there exists a subspace of B ∗ , Vj , of dimension dj ≤ C1 2j θσ
such that for any x, y ∈ X with x − y ≥ 2−j θ , one can find a ψ ∈ Vj with
ψ∗ = 1 and |ψ(x − y)| ≥ 2−j θ(1+β) . (8.4)
With V = {Vj }∞j =1 , choose γ > 1 and let E = Eγ (V ) and μ the corresponding
probability measure as defined in Section 5.2.
Now let
Zj = {z ∈ X − X : z ≥ 2−θj },
and for a fixed choice of f ∈ L (B, Rk ) let Qj be the set of all those linear
maps in E for which (8.2), with L replaced by f + L, fails for some z ∈ Zj ,
Qj = {L ∈ E : |(f + L)(z)| ≤ 2−j for some z ∈ Zj }.
Since dB (X − X) ≤ 2dB (X) < 2δ, X − X can be covered with no more than
C2 22j δ balls of radius 2−j . Let Y be the intersection of Qj with one of these
balls.
Let M be a Lipschitz constant that is valid for all f + L, L ∈ E. If z, z0 ∈ Y ,
then since z − z0 ≤ 2−(j −1) ,
|(f + L)(z0 )| > (2M + 1)2−j ⇒ |(f + L)(z)| > 2−j for all z ∈ Y.
Thus μ(Y ) is bounded by the μ-measure of those L ∈ E for which
|(f + L)(z0 )| ≤ (2M + 1)2−j .
8.2 Sharpness of the Hölder exponent 77
It follows from Lemma 5.10 (in the Banach space case) or Lemma 5.6 (in the
Hilbert space case) that
μ(Y ) ≤ [(2M + 1)2−j j 2 C1α 2j θσ α |g(z0 )|−1 ]k
for any g ∈ Sj (recall that dj ≤ C1 2j θσ ). Using the definition of the dual
thickness there exists a ψ ∈ Sj such that
|ψ(z0 )| ≥ z0 1+β ≥ 2−θ(1+β)j
(cf. (8.4)), and so
μ(Y ) ≤ [(2M + 1)2−j j 2 C1α 2j θσ α 2θ(1+β)j ]k .
Since Qj is covered by no more than C2 22j δ balls, we obtain
μ(Qj ) ≤ C2 22j δ [(2M + 1)2−j j 2 C1α 2j θσ α 2θ(1+β)j ]k
= C3 j 2k 2−j [k(1−θ(1+β+σ α))−2δ] .
The assumption (8.3) implies that
k(1 − θ (1 + β + σ α)) − 2δ > 0,
and so the sum ∞ j =1 μ(Qj ) is finite. It follows from the Borel–Cantelli Lemma
(Lemma 4.2) that μ-almost every L belongs to only finitely many of the Qj ,
which implies (8.2) for some appropriate constant CL (the argument is identical
to the concluding part of the proof of Theorem 4.3).
One could try to repeat this proof, replacing the upper box-counting dimen-
sion by the lower box-counting dimension, but the assumption that dLB (X) < ∞
is not enough to show that L−1 is Hölder, as we will now see.
where {ej }∞ p
j =1 is the canonical basis of ℓ (or c0 ), and q is the conjugate
exponent to p.
Note that the estimate of Lemma 6.3 for the Hilbert space case is better
(rank P ≥ ∞ 2
j =1 P ej ) since one can use orthogonality in the proof, rather
than just the triangle inequality. However, this does not affect the argument in
Corollary 8.4, where we only require the rank of P to be finite.
8.2 Sharpness of the Hölder exponent 79
Therefore,
∞
∞
n ∞
n
q q−1 q q−1 q
P ej ℓp ≤n |λij | = n |λij | ≤ nq .
j =1 j =1 i=1 i=1 j =1
As a corollary, one can show that for the orthogonal sets introduced in
Lemma 3.5, the box-counting dimension and the dual thickness coincide (for
a direct argument that gives this value for the thickness in ℓ2 and does not use
80 Embedding sets of finite box-counting dimension
Theorem 8.1 see Lemma 7.4); from this it follows that the Hölder exponent
in Theorem 8.1 is asymptotically sharp, using ℓ2 (p = q = 2) for the Hilbert
space case and c0 (‘p = ∞’, q = 1) for the Banach space case.
Corollary 8.4 Let A = {aj ej }∞ p
j =1 ∪ {0} be an ‘orthogonal’ subset of ℓ as
in Lemma 3.5. Then τ ∗ (A) = dB (A) and if there exists a finite-dimensional
projection P in ℓp and a θ ∈ (0, 1) such that
αℓp ≤ CP αθℓp , for each α ∈ A, (8.5)
then
1
θ≤ . (8.6)
1+ (τ ∗ (A)/q)
Proof Since P (aj ej ) = aj P ej , it follows from (8.5) applied to aj ej that
|aj | ≤ C|aj |θ P ej θℓp , i.e. P ej ℓp ≥ C −1/θ |aj |(1/θ)−1 .
Lemma 8.3 implies that for such a P ,
⎛ ⎞1/q ⎛ ⎞1/q
∞
∞
q
rank P ≥ ⎝ P ej ℓp ⎠ ≥ C −θ ⎝ |aj |q[(1/θ)−1] ⎠ .
j =1 j =1
it follows that
dB (A) ≤ q[(1/θ ) − 1],
which implies that
1
θ≤ . (8.8)
1 + (dB (A)/q)
We now deduce that in fact τ ∗ (A) = dB (A), which will give (8.6). We know
that in general τ ∗ (A) ≤ dB (A) (Lemma 7.8), so suppose that τ = τ ∗ (A) <
dB (A). Then the result of the embedding theorem (Theorem 8.1) coupled with
the Decomposition Lemma (Lemma 8.2) implies that for some k sufficiently
Exercises 81
large one can find a projection P of rank k such that (8.5) holds for some
θ > 1/(1 + (dB (A)/q)). But this contradicts (8.8), and so τ ∗ (A) = dB (A).
When p = 2, i.e. in the Hilbert space case, one can deduce that τ ∗ (A) =
τ (A) = dev(A) = dB (A) for this particular class of examples.
The same argument shows that one cannot replace the upper box-counting
dimension by the lower box-counting dimension and still obtain a Hölder
inverse. Indeed, the sequence Ĥ defined at the end of Chapter 3 has dLB (Ĥ ) =
1/α but dB (Ĥ ) = ∞. For p = 2 (the Hilbert space case) the condition (8.7)
becomes
∞
|aj |2[(1/θ)−1] < ∞. (8.9)
j =1
But the values of the {aj } used to define Ĥ are constant, of order 1/x, for ∼ex
values of j , for ever larger values of x. It follows that whatever the value of θ ,
the sum on the left-hand side of (8.9) diverges.
It is natural to ask how much the requirement of linearity restricts the
regularity of L−1 that can be attained. In particular, one can ask whether it is
possible to find, in general, an embedding (not necessarily linear) of X into some
Rk that is bi-Lipschitz (i.e. L and L−1 are Lipschitz). In the next chapter we
introduce the Assouad dimension, and show that a necessary condition for the
existence of such an embedding is that the Assouad dimension of a set is finite
(this condition is not sufficient, however, see Section 9.4); a simple example
(Lemma 9.9) shows that there are sets with finite box-counting dimension but
infinite Assouad dimension, so that a bi-Lipschitz embedding result for sets
with finite box-counting dimension is not possible. A more involved example
that illustrates the same thing was given by Movahedi-Lankarani (1992); again,
his set is one with infinite Assouad dimension.
Exercises
8.1 Foias & Olson (1996) prove that if P0 and P are orthogonal projections
on a real Hilbert space H of equal (finite) rank and P x = 0 implies that
P x ≤ ǫx for some ǫ ∈ (0, 1) then P − P0 ≤ ǫ. Use this result
along with Lemma 6.1 to deduce from Theorem 8.1 that if X ⊂ H , kǫN
with k > 2dB (X), and θ satisfies (8.1) with α = 1/2 then a dense set of
rank k orthogonal projections in H are injective on X and satisfy
for some C > 0. (This is essentially the result of Foias & Olson (1996),
although they do not give an explicit bound on θ .)
8.2 Suppose that {Xn }n∈Z is a family of subsets of B, such that dB (Xn ) ≤ d for
each n ∈ Z. Show that if k > 2d then there exists a linear map L : B → Rk
and a θ > 0 such that L is injective on j ∈Z Xj , and for each n ∈ N there
exists a Cn > 0 such that
x − y ≤ Cn |Lx − Ly|θ for all x, y ∈ Xj .
|j |≤n
[Hint: use the fact that a countable intersection of prevalent sets is preva-
lent.] (A version of this result is proved in Langa & Robinson (2001) for
the attractors of nonautonomous systems, in Langa & Robinson (2006) for
random dynamical systems, and in Robinson (2008) for general cocycle
dynamical systems.)
8.3 Use Lemma 6.3 (or Lemma 8.3) to show that if X ⊂ H contains a set of
the form {0} ∪ {αj }∞j =1 , where the αj are orthogonal, then no linear map
L : H → Rk can be bi-Lipschitz on X. (As discussed above, in the next
chapter we will see examples of sets for which no map, whether linear or
not, can provide a bi-Lipschitz embedding into a Euclidean space.)
9
Assouad dimension
83
84 Assouad dimension
Lemma 9.3 Suppose that (X, ̺X ) is (M, s)-homogeneous and that the map
f : (X, ̺X ) → (Y, ̺Y ) is bi-Lipschitz:
Proof Take y ∈ f (X) and consider BY (y, r) ∩ f (X). Then y = f (x) for some
x ∈ X. Since f −1 is Lipschitz,
dA (A) = dA (B) = dA (A × B) = 1.
4n
Proof Clearly dA (A) ≤ 1 and dA (B) ≤ 1. Let rn = 2−2 , and consider
∞
(−rn , rn ) ∩ A = {0} ∪ K4j .
j =n
4n
Then, since this contains K4n , it requires at least 22 − 1 intervals of length
4n+1
ρn = 2−2 to cover it. So
4n
NA (rn , ρn ) ≥ 22 − 1
and
4n
rn 2−2 4n
= −24n+1 = 22 .
ρn 2
So A cannot be (M, s)-homogeneous for any s < 1. It follows that dA (A) = 1,
and similarly dA (B) = 1.
Now consider A × B. Since A × B contains a copy of A, dA (A × B) ≥ 1.
2m
Take r = 2−2 ; this is the smallest value of r such that
⎡ ⎤ ⎡ ⎤
∞ ∞
B(0, r) ∩ [A × B] = ⎣{0} ∪ K2m+4j ⎦ × ⎣{0} ∪ K2(m+1)+4j ⎦ .
j =0 j =0
while
r 2n+1 2m
= 22 −2 .
ρ
It follows that N ≤ (r/ρ)s , where
Now cover B(0, rm ) ∩ X by balls of radius rm /2; each point in this set of norm
more than rm /2 will require its own ball, and since
K −1 α n ≤ an ≤ Kα n .
Proof Take 0 < ρ < r. Consider a ball of radius r centred at the origin; then
A cover of B(0, r) ∩ X by balls of radius ρ will require a separate ball for each
point of norm greater than ρ; since
K −1 α n > ρ ⇒ an > ρ,
it follows that an > ρ for n < (log ρ + log K)/ log α, so certainly the same is
true for n ≤ [(log ρ + log K)/ log α] + 1. Thus
log ρ + log K log r − log K
N(B(0, r) ∩ X, ρ) ≤ − +2
log α log α
1 r 2 log K
= log + + 2.
− log α ρ log α
So X is (M, s)-homogeneous for any s > 0, i.e. dA (X) = 0.
The lower bound in this result is necessary:
Lemma 9.11 There are sequences an converging arbitrarily fast to zero for
which
X = {an en }∞
n=1 ∪ {0}
has dA (X) = ∞.
Proof Let bj be a sequence that converges to zero. Let an = bj for 2j −1 ≤
n ≤ 2j − 1. Now let r = bj + ǫ and consider
B(0, r) ∩ X = {an en : n ≥ 2j −1 }.
The number of balls of radius r/2 required to cover B(0, r) ∩ X is larger than
2j − 2j −1 , which is unbounded as j → 0. So X is not doubling. Since bj can
converge arbitrarily fast to zero, so can an .
Assumptions on the set of differences X − X are the key to proving embed-
ding results that use linear maps. We have seen that while for the box-counting
dimension dB (X − X) ≤ 2dB (X), one can have sets with zero Hausdorff dimen-
sion for which dH (X − X) = ∞. Unfortunately the same is true of the Assouad
dimension.
Lemma 9.12 There exists a set X with dA (X) = 0 and dA (X − X) = ∞.
Proof Let {xj } be an orthogonal sequence of the type constructed in the previ-
ous lemma, with xj ≤ 4−j . Suppose that the complement of the linear span
of the {xj } is infinite-dimensional, and choose a second orthogonal sequence
{yj } in this complement with yj = 4−j .
Let X be the closure of the set {aj }, where
a2j = yj and a2j +1 = xj + yj .
90 Assouad dimension
while for k = 2j + 1,
and
Note that in a very roundabout way we have shown that the Assouad
dimension can increase under Lipschitz continuous transformations, since
dA (X × X) ≤ 2dA (X), and X − X is the image of X × X under the Lipschitz
mapping (x, y) → x − y.
The following result, again due to Olson (see Olson & Robinson (2010)) is
more positive, and will be useful below.
Cover each of BX (r, x), BX (r, −x), BX (r, y), and BX (r, −y) by M(2r/ρ)s balls
of radius ρ/2. An argument similar to that used before yields a cover of B by
1 + 4M 2 (2r/ρ)2s balls of radius r/2.
Since NX−X (r, ρ) ≤ 1 + 4M 2 (2r/ρ)2s it follows that dA (X − X) ≤ 2s.
X0
X1
X2
Figure 9.1 The first steps of the construction of the geodesic metric space (X, ̺).
At each stage the bold subset is isometric to X0 .
distance: the shortest distance that one needs to travel on the graph from x
to y. For every j > i, (Xj , ̺j ) contains an isometric copy of (Xi , ̺i ), and
dist(Xj , Xi ) < (1/4)i+1 (see Figure 9.1) and so {(Xi , ̺i )}∞
i=1 forms a Cauchy
sequence in the Gromov–Hausdorff metric1 (see Chapter 3 of Gromov (1999),
or Heinonen (2003)). It follows that this sequence converges to some limiting
compact metric space (X, ̺); there remain isometric copies of (Xi , ̺i ) in (X, ̺).
The exact details of this limiting argument are not necessary here, the key point
is that this process leads to such a limit set containing isometric copies of every
(Xi , ̺i ).
Lemma 9.14 The space (X, ̺) is doubling with doubling constant 6, and if
H is a Hilbert space and f : Xi → H satisfies f (x) − f (y) ≥ ̺(x, y) then
the Lipschitz constant of f is bounded below by (1 + (i/4))1/2 . In particular,
there is no bi-Lipschitz embedding of X into a Hilbert space.
Proof Take x ∈ X and r with 0 < r ≤ 21 . Choose i with
i
r 1
≤ < 2r,
2 4
1 First, given two subsets A, B of a metric space (X, ̺), define the symmetric Hausdorff distance
distH (A, B) = max(dist(A, B), dist(B, A)). The Gromov–Hausdorff distance between two met-
ric spaces A and B is defined as
dGH (A, B) = inf distH (A′ , B ′ ),
A′ ,B ′ ∈M
where the infimum is taken over all isometric images A′ and B ′ of A and B as subsets of ℓ∞ (at
least one such isometry always exists, see Exercise 9.2).
9.4 Homogeneity is not sufficient for a bi-Lipschitz embedding 93
this ‘quadrilateral inequality’ holding in any inner product space, see Exercise
9.1. Since
2 2 k−1
f (x0 ) −f (x2 ) + f (x1 ) − f (x3 ) ≥ 1 + ̺(x0 , x2 )2 + ̺(x1 , x3 )2
4
k
= 1+ ̺(x0 , x2 )2 ,
4
it follows that for some j ∈ {0, . . . , 3}
1 i k
f (xj ) − f (xj +1 )2 ≥ 1+ ̺(x0 , x2 )2 = 1 + ̺(xj , xj +1 )2 .
4 4 4
Now take x = xj and x ′ to be one of the midpoints between xj and xj +1 to
obtain (9.3) for i = k.
94 Assouad dimension
Proof Write
j = {z ∈ Z : 2−(j +1) ≤ z ≤ 2−j }.
Since j ⊂ B(0, 2−j ) it can be covered by Mj balls of radius 2−(j +3) , with the
(j ) Mj (j )
centres {ui }i=1 of these balls satisfying ui ≥ 2−(j +2) , where
Mj = NX (2−j , 2−(j +3) ) ≤ 8s M = M ′ .
(j )
For each of the points ui , use the Hahn–Banach Theorem to find a norm 1
(j )
element φi of B ∗ such that
(j ) (j ) (j )
φi (ui ) = ui .
(j ) (j ) (j )
For each n ≥ 0 let Vj be the subspace of B ∗ spanned by {φ1 , φ2 , . . . , φMj }.
By the above, dim Vj ≤ M ′ for all j ≥ 0.
For any z ∈ n there exists a u = u(n)
i such that
z − u < 2−(n+3) .
Writing φ for φi(n) ∈ Vn ,
|φ(x)| = |φ(u) − φ(u − z)| ≥ u − u − z
≥ 2−(n+2) − 2−(n+3) = 2−(n+3) ,
and the lemma follows.
In a Hilbert space it is more helpful to use the following result; note that
the spaces Vj are now mutually orthogonal, but that the space Vn alone is not
sufficiently ‘rich’ to obtain (9.5) (cf. (9.4), where ψ ∈ Vn is enough).
Lemma 9.16 Suppose that Z is a compact homogeneous subset of a Hilbert
space H . Then there exists an M ′ > 0 and a sequence {Vj }∞j =0 of mutually
′
orthogonal linear subspaces of H , with dim Vj ≤ M for every j , such that for
any z ∈ Z with 2−(j +1) ≤ z ≤ 2−j ,
j z ≥ 2−(j +2) , (9.5)
j
where j is the orthogonal projection onto ⊕i=1 Vi .
Proof Write
Zj = {z ∈ Z : 2−(j +1) ≤ z ≤ 2−j }.
Since Zj ⊂ B(0, 2−j ) it can be covered by Mj balls of radius 2−(j +2) , with
(j ) Mj
centres {ui }i=1 , where
Mj = N(2−j , 2−(j +2) ) ≤ 4s M = M ′ .
96 Assouad dimension
(j ) M
j
Let Uj be the space spanned by {ui }i=1 ; clearly dim(Uj ) ≤ M ′ , and if Pj
denotes the projection onto Uj ,
Pj z ≥ z − z − Pj z ≥ 2−(j +1) − 2−(j +2) = 2−(j +2) .
Finally, define mutually orthogonal subspaces Vj such that
n
3 n
3
Vj = Uj
j =1 j =1
Thus
(j ) (j )
|(f + L)(z)| ≥ |(f + L)(zi )| − |(f + L)(z − zi )|
(j )
≥ |(f + L)(zi )| − Kj −γ 2−j ,
for any ψ ∈ Vj .
(j )
Lemma 9.15 implies that there exits a ψ ∈ Vj such that |ψ(zi )| ≥ 2−(j +3) ,
and since Nj ≤ Mj γ s and dj ≤ M ′ ,
N
μ(Qj ) ≤ c j γ s (M ′ )α j ζ j −γ = c j γ s+N(ζ −γ ) .
while in the Banach space case (taking ‘p = ∞’, i.e. c0 ) we cannot improve
on γ > 1.
Exercises
9.1 Show that if x, y, z are elements of any inner product space then
2 22
2x − y + z 2 ≤ 1 x − y2 + 1 x − z2 − 1 y − z2 ,
2 2
(9.13)
2 2 2 2 2 4
Exercises 101
Finite-dimensional attractors
10
Partial differential equations and
nonlinear semigroups
The second part of this book concentrates on the implications of Theorem 8.1
(embedding into Rk for sets with finite upper box-counting dimension) for the
attractors of infinite-dimensional dynamical systems.
105
106 Partial differential equations and nonlinear semigroups
for every ϕ ∈ Cc∞ () (infinitely differentiable functions with compact support
in ).
We use the standard notation H s () for the Sobolev space of functions that,
together with their (weak) partial derivatives of order ≤ s, are square integrable
10.2 Sobolev spaces and fractional power spaces 107
(i) If β > α then D(Aβ ) is a subset of D(Aα ), and the embedding is com-
pact, i.e. a bounded subset of D(Aβ ) is a compact subset of D(Aα ) (see
Exercise 10.2).
(ii) If A is a second order linear elliptic operator with constant coefficients
then there exist constants Cs and Cs′ such that
(The first of these is straightforward, the second relies on the theory of ellip-
tic regularity; see Evans (1998), Gilbarg & Trudinger (1983), or Proposition
6.18 in Robinson (2001) for a proof when A = − with Dirichlet boundary
conditions.)
for some α ∈ [0, 1). Given any u0 ∈ D(Aα ), there exists a unique solution
u(t; u0 ) : [0, T ) → D(Aα ), where T depends on u0 α , and this solution is
given by the variation of constants formula
" t
u(t; u0 ) = e−At u0 + e−A(t−s) g(u(s)) ds, (10.6)
0
10.4 The two-dimensional Navier–Stokes equations 109
where
∞
e−At u = e−λj t (u, wj )wj
j =1
(see Henry (1981, Lemma 3.3.2 and Theorem 3.3.3)). Solutions are continu-
ous from [0, T ) into D(Aα ) and depend continuously on the initial condition
(Henry’s Theorem 3.4.1).
If we assume that unique solutions of (10.4) exist for all t ≥ 0 (this usually
requires an equation-by-equation approach tailored to the particular model
under consideration), then the solutions generate a semigroup on D(Aα ) via
the definition S(t)u0 = u(t; u0 ). Properties (i) and (iii) are immediate from the
above results, and property (ii), S(t + s) = S(t)(s) for all t, s ≥ 0, follows from
the uniqueness of solutions.
The following estimates for the action of e−At between different fractional
power spaces are extremely useful:
γ −γ −γ
γ e t 0 < t < γ /λ1 ,
e−At L (H,D(Aγ )) = Aγ e−At L (H ) ≤ γ (10.7)
λ1 e−λ1 t t ≥ γ /λ1 ,
and consequently
" ∞ −(1−γ )
e−γ λ1
Aγ e−At L (H ) dt ≤ Iγ := if γ ∈ [0, 1), (10.8)
0 1−γ
see Exercise 10.3. Sometimes it is convenient to rewrite (10.7) as
Classically, these are equations for the two-component velocity u(x, t) and
the scalar pressure p(t),
∂u
− u + (u · ∇)u + ∇p = f (x) ∇ · u = 0, (10.11)
∂t
where we have set the kinematic viscosity (the coefficient of the Laplacian term)
equal to 1. The right-hand side f represents a body forcing that maintains the
motion (with f = 0 every solution decays to zero and the attractor is trivial,
see Exercise 10.5).
For mathematical simplicity we will concentrate on the periodic case, when
x ∈ = [0, 2π ]2 and
u(x + 2π ei , t) = u(x, t), i = 1, 2, (10.12)
where e1 and e2 are orthonormal vectors in R2 . It is also convenient to assume
the zero-average conditions
" "
f (x) dx = 0 and u0 (x) dx = 0; (10.13)
the condition on f ensures that the zero average of u(x, t) is preserved under
the time evolution.
The natural phase space for the problem we will denote2 by H : it is the
completion in the [L2 ()]2 -norm of
"
∞
H = {u ∈ [Cper ()]2 : ∇ · u = 0 and u(x) dx = 0}, (10.14)
∞ ∞
where Cper ()
is the space of all C functions that are periodic as in (10.12);
we equip H with the L2 norm. (Roughly speaking H consists of all functions
in [L2 ()]2 with zero average and (generalised) divergence zero.)
Given f ∈ H and u0 ∈ H , for all t ≥ 0 there exists a unique solution which
we denote by u(t) = u(t; u0 ) (suppressing the x dependence) that is continuous
from [0, ∞) into H , and depends continuously (in the H -norm) on u0 (see
Constantin & Foias (1988), Robinson (2001), or Temam (1977)). As with the
abstract semilinear equation in the previous section, we can use the solution to
define a semigroup on H by setting S(t)u0 = u(t).
It is a standard approach (particularly in the literature that views the two-
dimensional equations as a dynamical system) to reformulate (10.11) in ‘func-
tional form’, essentially as an ordinary differential equation on an appropriate
space. This reformulation is one way to eliminate the pressure from the equa-
tions, capitalising on the observation that the pressure term ∇p is orthogonal
2 Whenever we are dealing with the Navier–Stokes equations our ‘primary’ Hilbert space will be
H , so this should not cause any confusion with more general abstract considerations.
10.4 The two-dimensional Navier–Stokes equations 111
where ·, · denotes the pairing between V ∗ and V , and define a bilinear form
B : V × V → V ∗ by
B(u, u), v = ((u · ∇)u, v) for all u, v ∈ V .
for some cs , Cs .
In the analysis that follows we will only require the following properties of
A and B:
3 2 2
The notation (Du, Dv) denotes i,j =1 (Di uj , Di vj ); in particular Du2 = i,j =1 Di uj 2 .
112 Partial differential equations and nonlinear semigroups
(this follows from an integration by parts, and remains true for other boundary
conditions and in the three-dimensional case), and
which is only true in the two-dimensional periodic case (the proof relies on
expanding the expression ((u · ∇)u, u), then using the divergence-free con-
dition repeatedly in many pairwise cancellations). We will also make use of the
Poincaré inequality,
which follows making use of the zero-average condition (10.13), and is easy to
see using the Fourier expansion of u (see Exercise 10.6).
Finally we note that the Navier–Stokes equations can be recast in the abstract
form (10.4), where A is the Stokes operator and
In this case g(u) is locally Lipschitz from D(Aα ) into H for any α > 1/2:
For the first term we have (since u∞ ≤ cuH 2α as α > 1/2)
For the first and final time we use the two-dimensional Sobolev embedding
result (see Evans (1998), for example)
(which in particular shows that any Lp norm is bounded by the H 1 norm) with
s = 2α − 1 and obtain
It follows that
g(u) − g(v)L2 ≤ c[uα + vα ]u − v1/2 , (10.20)
α
and in general g is locally Lipschitz from D(A ) into H .
One could therefore could treat the two-dimensional Navier–Stokes equa-
tions within the abstract framework of Section 10.3. However, this would
require us to take an initial condition in D(Aα ) with α > 1/2, and to consider
the dynamical system generated on this space. Instead, it is more useful to
obtain the existence of a solution via other methods (above we stated that a
unique solution exists for any u0 ∈ H ), and then use the more abstract setting
when it makes the analysis more convenient.
We will adopt this approach in Section 13.2, in which we investigate the
Lipschitz deviation of attractors. There, the following observation will be cen-
tral: it follows from (10.20) that if we restrict our attention to a set X that is
bounded in D(Aα ) with α > 12 then g is Lipschitz from D(A1/2 ) into H :
g(u) − g(v)L2 ≤ Cu − v1/2 for all u, v ∈ X.
Exercises
ikx
10.1 If f (x) = k∈Z ck e then f ∈ H 1 (0, 2π ) provided that
f 2H 1 = (1 + |k|2 )|ck |2 < ∞.
Show that f ∞ ≤ cf H 1 , and deduce that f ∈ C 0 ([0, 2π ]).
10.2 Show that D(Aβ ) is compactly embedded in if f ǫH 1 |0, 2π ) then D(Aα )
if β > α.
10.3 If u = ∞ j =1 cj wj then
∞
2γ
Aγ e−At u2 = λj e−2λj t |cj |2 .
j =1
dist(S(t)B, X) → 0 as t → ∞.
dist(S(t)A2 , A1 ) → 0 as t → ∞.
115
116 Attracting sets in infinite-dimensional systems
Proposition 11.2 Suppose that there exists a compact attracting set K. Then
for any bounded set B, the set
ω(B) = S(s)B (11.1)
t≥0 s≥t
Proof Since there is a compact attracting set, Lemma 11.1 combined with
(11.2) shows that ω(B) ⊆ K; that ω(B) is nonempty follows similarly, taking
any initial sequences {bn } ∈ B and tn → ∞. Using (11.1), ω(B) is a decreasing
sequence of closed sets, and so is a closed subset of the compact set K; thus
ω(B) is compact.
Now suppose that x ∈ ω(B). Then there exist sequences {tn } with tn → ∞
and {bn } with bn ∈ B such that x = limn→∞ S(tn )bn . Then, since S(t) is con-
tinuous,
% &
S(t)x = S(t) lim S(tn )bn = lim S(t + tn )bn ,
n→∞ n→∞
But (by Lemma 11.1) {S(tn )bn } has a convergent subsequence, whose limit
must lie in ω(B), a contradiction.
Theorem 11.3 There exists a global attractor A if and only if there exists a
compact attracting set K, in which case A = ω(K).
The set A is clearly compact (since each ω(B) is contained in the compact
set K), invariant, and attracts every bounded set B, so is the global attractor.
It only remains to show that A = ω(K). It is immediate from (11.3) that
A ⊇ ω(K), while since A is the minimal closed set that attracts bounded sets
(Exercise 11.1) we must have A ⊆ K, and hence A = ω(A ) ⊆ ω(K).
i.e. the orbits of all bounded sets eventually enter and do not leave X. Clearly
the existence of a compact absorbing set implies the existence of a compact
attracting set, which we know implies the existence of a global attractor. In Sec-
tion 11.3 we prove that the existence of a bounded absorbing set for an abstract
semilinear parabolic equation implies the existence of a compact absorbing set,
and hence of the global attractor; in Section 11.4 we will prove the existence of
a global attractor for the two-dimensional Navier–Stokes equations by showing
directly the existence of a compact absorbing set.
Finally we give an alternative, more analytical characterisation of attractors
in terms of complete bounded orbits. This shows that while these objects have
a definition in terms of dynamics, they are of interest independent of their
dynamical interpretation.
118 Attracting sets in infinite-dimensional systems
First, if one takes the inner product (in H ) of (11.6) with u, since (Au, u) =
Du2 (10.16) and (B(u, u), u) = 0 (10.17) one obtains
1 d
u2 + Du2 = (f, u) ≤ f u. (11.7)
2 dt
Using the Poincaré inequality u ≤ Du (10.19) on the left-hand side, and
Young’s inequality (2ab ≤ a 2 + b2 ) on the right-hand side, one obtains the
differential inequality
d
u2 + u2 ≤ f 2 .
dt
This can be readily integrated (using the integrating factor et ) to deduce that
and so u(t)2 ≤ 2f 2 for all t ≥ t0 (u0 ). This provides a bounded absorbing
set in H . To obtain a compact absorbing set, we will show that there is a bounded
absorbing set in H 1 , i.e. that Du(t)2 is asymptotically bounded, uniformly
in terms of the L2 norm of the initial condition.
We first require a subsidiary estimate. Dealing with (11.7) differently, one
can use the Poincaré inequality and Young’s inequality on the right-hand side
to obtain
d
u2 + Du2 ≤ f 2 .
dt
Integrating this differential inequality from t to t + 1 gives
" t+1
u(t + 1)2 + Du(s)2 ds ≤ f 2 + u(t)2 .
t
Now take the inner product of (11.6) with Au, and use the special two-
dimensional periodic orthogonality relation (10.18) to obtain
1 d 1 1
Du2 + Au2 = (f, Au) ≤ f Au ≤ f 2 + Au2 . (11.9)
2 dt 2 2
Absorbing the Au2 from the right-hand side into the same term on the left-
hand side, and dropping the resulting + 12 Au2 , we obtain
d
Du2 ≤ f 2 . (11.10)
dt
Exercises 121
using (11.8). Since this is valid for all t ≥ t0 (u0 ), it follows that
Du(t)2 ≤ 6f 2 for all t ≥ t0 (u0 ) + 1.
(This ‘double integration’ trick can be formalised as the ‘Uniform Gronwall
Lemma’, see Exercise 11.5.)
This implies the existence of a bounded absorbing set in H 1 , and since H 1
is compactly embedded in L2 , this gives a compact absorbing set in H and
guarantees the existence of a global attractor A for the semigroup on H . With
a little further work one can show that the global attractor is a bounded subset
of D(A) (and hence of H 2 ), see Exercise 11.6.
Exercises
11.1 Show that the global attractor is the maximal compact invariant set, and
the minimal closed set that attracts all bounded sets.
11.2 Show that (11.1) and (11.2) are equivalent.
11.3 Show that A is connected whenever B is connected. [Hint: argue by
contradiction.]
11.4 If X is an invariant set, the unstable set of X is defined by
U (X) := {u0 ∈ B : there exists a globally defined solution u(t) with
u(0) = u0 and dist(u(t), X) → 0 as t → −∞}.
Show that U (X) ⊂ A for any invariant set X.
11.5 Use the double integration method used in Section 11.4 to prove the
‘Uniform Gronwall Lemma’: if x, a, and b are positive functions such
that
dx/dt ≤ ax + b
with
" t+r " t+r " t+r
x(s) ds ≤ X, a(s), ds ≤ A, and b(s) ds ≤ B
t t t
122 Attracting sets in infinite-dimensional systems
123
124 Bounding the box-counting dimension of attractors
1 Alternatively, dB (K) ≤ γ whenever θ γ M < 1. This formulation will be useful in Exercise 12.5.
12.1 Coverings of T [B(0, 1)] via finite-dimensional approximations 125
Now fix η with 0 < η < 1 − α, and let r0 = r0 (η). Cover K with N(K, r0 )
balls of radius r0 , {B(xj , r0 )}N j =1 , with centres xj ∈ K. Apply f to every
element of this cover. Since f (K) = K, this provides a new cover of K,
{f (B(xj , r0 ))}N
j =1 . It follows from (12.3) that each of these images can be
covered by M balls of radius (α + η)r0 , ensuring that
which via Lemma 3.2 (on taking the lim sup through a geometric sequence)
yields
log M
dB (K) ≤ .
− log(α + η)
Since η > 0 was arbitrary we obtain (12.2).
The key to applying this approach is to be able to prove (12.1), i.e. to
find a way of estimating the number of balls of radius α required to cover
Df (x)B(0, 1). When Df (x) is the sum of a compact map and a contraction,
we reduce the problem of covering Df (x)[B(0, 1)] to the problem of covering
Df (x)[BZ (0, 1)], where Z is some finite-dimensional subspace of B. We then
prove a covering result for balls in finite-dimensional subspaces. If Df (x) is
the sum of a compact map and a contraction for every x ∈ K in some suitably
uniform way we can then obtain (12.1) with the same α and M for every
x ∈ K.
Lemma 12.2 Let B be a Banach space and T ∈ Lλ/2 (B). Then there exists
a finite-dimensional subspace Z of B such that
Suppose that this is not the case. Choose some x1 ∈ B with x1 = 1, and let
Z1 = span{x1 }. Then
Cx2 − Cx1 ≥ ǫ.
Continuing inductively one can construct in this way a sequence {xj } with
xj = 1 such that
Cxi − Cxj ≥ ǫ i = j,
Hence,
dist(T [B(0, 1)], T [BZ (0, 1)]) ≤ λ̃ + dist(C[B(0, 1)], C[BZ (0, 1)])
< λ,
We now need to be able to cover T [BZ (0, 1)] with B-balls of a smaller
radius. Since
which gives the first inequality in (12.5). On the other hand, if x ∈ U with
x = nj=1 zj xj and xU ≤ 1 then since zj = fj (x),
If x ∈ B(0, 1), it follows from (12.6) that there is a y ∈ T [BZ (0, 1)] such
that T x − y < λ. Since y ∈ T [BZ (0, 1)], it follows from (12.7) that
y − yi ≤ λ for some i ∈ {1, . . . , k}, and so
T x − yi ≤ T x − y + y − yi < 2λ,
1
12.2 A dimension bound when Df ∈ Lλ/2 (B), λ < 2
We now show, following Mañé (1981), that if Df (x) ∈ Lλ/2 for every x ∈ K for
some λ with 0 < λ < 21 then we have sufficient control to bound the dimension
of K.
Then n = supx∈K νλ (Df (x)) and D = supx∈K Df (x) are finite, and
log((n + 1)D/λ)
dB (K) ≤ n .
− log(2λ)
Proof First we show that n = supx∈K νλ (Df (x)) is finite. For each x ∈ K,
there exists a finite-dimensional linear subspace Zx such that
Since Df (·) is continuous, it follows that there exists a δx > 0 such that
for all y ∈ B(x, δx ), i.e. νλ (y) ≤ νλ (x) for all such y. The open cover of K
formed by the union of B(x, δx ) over x has a finite subcover, whence it follows
that n < ∞.
Now, since n = supx∈K νλ (Df (x)) < ∞, we can use Corollary 12.4 to
deduce that
' (
D n
N (Df (x)[B(0, 1)], 2λ) ≤ (n + 1) for all x ∈ K.
λ
The bound on the dimension now follows using Lemma 12.1.
130 Bounding the box-counting dimension of attractors
It follows that if Df (x) ∈ Lα (B) with α < 1 then [D(f p )](x) ∈ Lαp (B). Thus
for p large enough, D(f p )(x) ∈ Lλ for some λ < 1/4, for every x ∈ K. One
can now apply Theorem 12.5 to f p in place of f (noting that f p (K) = K) to
deduce that dB (K) < ∞.
(see Henry (1981, Theorem 3.4.4)). Using the bound on Aγ e−At L (H ) in
(10.9) we obtain
DS(t; u0 )L (D(Aα )) ≤ e−At L (D(Aα ))
" t
+ Aα e−A(t−s) L (H ) Dg(S(s(u0 )))L (D(Aα ),H ) DS(s; u0 )L (D(Aα ))
0
" t
≤ 1 + cM (t − s)−α DS(s; u0 )L (D(Aα )) ds,
0
where
M = sup{Dg(x)L (D(Aα ),H ) : x ∈ A }. (12.9)
It follows from this inequality, using the result of Exercise 12.2, that
1/(1−α)
DS(t; u0 )L (D(Aα )) ≤ K := 2e[2cMŴ(1−α)] t ∈ [0, 1].
for all
(12.10)
Now choose ǫ > 0 such that α + ǫ < 1. Taking advantage of (12.10) one
can use very similar estimates to those above to show that
" t
DS(t)L (D(Aα ),D(Aα+ǫ )) ≤ ct −ǫ + cMK (t − s)−(α+ǫ) ds.
0
α+ǫ α
Since D(A ) is compactly embedded in D(A ), this shows that DS(t) is
compact for any t > 0. That dB (A ) is finite now follows immediately from
Corollary 12.6 applied with B = D(Aα ).
This approach is also applicable to the two-dimensional Navier–Stokes equa-
tions. We saw in Section 10.4 that the Navier–Stokes equations can be cast in
the form (12.8) with g locally Lipschitz from D(Aα ) into H provided that
α > 1/2, and Exercise 11.6 guarantees that if f ∈ H then the attractor is
bounded in D(A), so A is certainly bounded in D(A3/4 ) (to choose some fixed
α with 1/2 < α < 1). Corollary 12.7 then implies that the dimension of A
measured in D(A3/4 ) is finite; so certainly the dimension of A measured in H
is finite (see (3.5)).
As remarked at the beginning of this chapter, obtaining good bounds on the
dimension of the Navier–Stokes attractor has been an active area of research.
Of course, such bounds use the Hilbert space theory rather than the Banach
space approach developed above. In the periodic case, Constantin, Foias, &
132 Bounding the box-counting dimension of attractors
there are example forcing functions f for which the dimension is bounded
below by c′ f 2/3 (Liu, 1993), so this bound is essentially sharp. For a simpli-
fied proof and further discussion, see Doering & Gibbon (1995).
Exercises
12.1 Let B be a Banach space and assume that f ∈ C 1 (X), that K is a compact
set such that f (K) = K, and that for every x ∈ K the derivative Df (x)
has finite rank ν(x) with supx∈K ν(x) := ν < ∞. Show that dB (K) ≤ ν.
12.2 Suppose that X(t) satisfies
" t
X(t) ≤ a + b (t − s)−α X(s) ds.
0
1/(1−α)
With K = (2bŴ(1 − α)) show that Y (t) = 2aeKt satisfies
" t
Ẏ ≥ a + b (t − s)−α Y (s) ds,
0
Kt
and hence that X(t) ≤ 12ae .
∞
12.3 Recalling that Ŵ(z) = 0 t z−1 e−t dt, take up the argument of Corollary
12.7 immediately after (12.10), and show that
cKMŴ(1 − α)
Qn DS(1; u0 )L (D(Aα )) ≤ e−λn+1 + , (12.11)
(λn+1 − 1)1−α
where Pn is the orthogonal projection onto the space spanned by the first
n eigenfunctions of A and Qn = I − Pn is its orthogonal complement.
(Choosing n large enough that the right-hand side is strictly less than
1/8, one can then apply Theorem 12.5 with B = D(Aα ) and λ = 1/4 to
deduce that
log[4K(n + 1)]
dB (A ) ≤ n .
log 2
Dropping the first term in (12.11) we require λ1−α n+1 cKM, which
using the estimate on K in (12.10) becomes λn+1 cMα ecMα with
1/γ
Mα = M 1/(1−α) . Assuming that λn ∼ nγ , this yields n Mα ecMα , and
1+(1/γ ) cMα
hence dB (A ) Mα e .)
Exercises 133
The remainder of the exercises in this chapter outline the more refined theory
that is available to estimate the dimension of attractors in the Hilbert space case.
Exercise 12.4 provides a good estimate on coverings of T [B(0, 1)] in terms
of the singular values of T . Exercise 12.5 converts this into a bound on the
dimension via Lemma 12.1, and the remaining exercises give an indication of
how to apply this method efficiently in applications (which usually arise from
continuous time systems).
The covering argument of Exercise 12.4 requires the following two results.
The proof of the first is essentially the same as that of Lemma 14.2, below,
and the proof of the second can be found in Chepyzhov & Vishik (2002,
Lemma 2.2).
α1 (T ) ≥ α2 (T ) ≥ α3 (T ) ≥ · · · ,
where ᾱ1 ≥ ᾱ2 ≥ · · · and ᾱnn ≤ ω̄n . Show that, for any choice of d ∈ N,
where
√ 1/d 4j ω̄j
θ= 2ω̄d and M = max j/d
.
1≤j ≤d ω̄d
134 Bounding the box-counting dimension of attractors
1/d
[Hint: for each u ∈ K, consider the two cases ω̄d < α1 (Df (u)) and
1/d
ω̄d ≥ α1 (Df (u)) separately.]
12.5 Under the same conditions as in the previous exercise, use Lemma 12.1
to show that if
γ ω̄jd
ω̄d < 1 and ω̄d max j
< 1, (12.12)
1≤j ≤d ω̄d
where
⎧ ⎫
⎨n ⎬
Trn (L) = sup (ψj , Lψj ) : {ψj }nj=1 are orthonormal in H .
⎩ ⎭
j =1
12.8 Consider a semigroup S(·) defined on H that arises from the semilinear
evolution equation
Ott, Hunt, & Kaloshin (2006) conjectured that ‘many of the attractors associated
with the evolution equations of mathematical physics have thickness exponent
zero’.
In this chapter we give two results in this direction. The first, due to Friz &
Robinson (1999), shows that in some sense the thickness exponent is ‘inversely
proportional to smoothness’: if U ⊂ Rm and A is a subset of L2 (U ) that is
bounded in the Sobolev space H s (U ) then τ (A ) ≤ m/s, where the thickness
of A is measured in L2 (U ). So if an attractor is ‘smooth’ (i.e. is bounded in
H s () for every s) then it has zero thickness exponent.
The second result, due to Pinto de Moura & Robinson (2010c), is closer in
spirit to the above conjecture. This shows that the attractors of equations that
can be written as semilinear parabolic equations
136
13.1 Zero thickness 137
We will use a similar argument to what follows for Lemma 15.5. The proof
is due to Friz & Robinson (1999), see also Robinson (2008).
Proof A proof for the case n = 1 is sufficient; if n > 1 the argument can
be applied to each component of the functions in X. Let U ′ , U ′′ be smooth
bounded domains such that
U ⊂ U′ and U ′ ⊂ U ′′
and the support of E[u] is contained in U ′ (e.g. Theorem 7.25 in Gilbarg &
Trudinger (1983)).
Let A denote the Laplacian operator on U ′′ , with Dirichlet boundary con-
ditions (u = 0 on ∂U ′′ ). The Laplacian on such a domain has a sequence
{wj } of eigenfunctions with corresponding eigenvalues λj (Awj = λj wj )
which, if ordered so that λj +1 ≥ λj , satisfy λj ∼ j 2/m (see Davies (1995), for
example).
Since, for any u ∈ X, the support of E[u] is contained in U ′ , E[u] ∈
D(As/2 ). It follows from (13.2) and the inequality
Theorem 13.3 Take α ∈ [0, 21 ], and suppose that (13.3) has an attractor A
that is bounded in D(Aα ). Pick M0 such that A ⊂ BH (0, M0 /8).
(i) There exists a constant C > 0 such that
A1/2 (u − v)2
<C for all u, v ∈ A .
u − v2 log(M02 /u − v2 )
(ii) For each n ∈ N there exists a 1-Lipschitz function n : Pn H → Qn H ,
Qn = I − Pn ,
such that
(iii) If in addition
λn
lim =∞ (13.6)
n→∞ log n
then dev(A ) = 0 (where the Lipschitz deviation is measured in H).
The proof of (i) is due to Kukavica (2007); Pinto de Moura & Robinson
(2010c) show that (ii) is a consequence of (i) (using an argument of Foias,
Manley, & Temam (1988)) and observe that (iii) follows from (ii) (see also
Pinto de Moura & Robinson (2010b).)
The condition (13.6) will be satisfied in most interesting examples, since for
an elliptic operator of order 2p defined in ⊂ Rm , λn ∼ n2p/m (see Davies
(1995), for example).
Proof (i) Let w(t) = u(t) − v(t), L(t) = log(M02 /w(t)2 ) (so in particular
L ≥ 1), and set
A1/2 w2 Q(t)
Q(t) = and Q̃(t) = .
w2 L(t)
140 Thickness exponents of attractors
Now we use the fact that g is locally Lipschitz (10.5) and that A is bounded
in D(Aα ) to deduce that
h(t) = g(u(t)) − g(v(t)) ≤ Ku(t) − v(t)α = Kw(t)α
for some K > 0; combine this with the interpolation inequality wα ≤
w1−2α w2α
1/2 (see Exercise 13.4) to bound the right-hand side:
Since λn+1 Qn w2 ≤ A1/2 w2 it follows from (13.12) that
and hence
from (13.5). Following an argument similar to that used in Lemma 3.2, take
ǫ > 0 with εn+1 ≤ ǫ < εn , and then
log δ1 (A , ǫ) log δ1 (A , εn+1 ) log(n + 1)
≤ ≤ .
− log ǫ − log εn (λn+1 /2C) − 2 log M0
Then dev(A ) ≤ dev1 (A ) = 0 provided that (13.6) holds.
The increased power of this result over that of Lemma 13.1 is clearly
demonstrated by the following consequence for the two-dimensional Navier–
Stokes equations (recall that H is essentially the space L2 of square integrable
functions).
Proof We saw in Section 10.4 that the Navier–Stokes equations can be written
in the form (13.3) where g satisfies
for any α > 1/2 (this was (10.20)). For f ∈ H the attractor is bounded in D(A)
(see Exercise 11.6) and so it follows that uβ is bounded for all u ∈ A , for
any 0 ≤ β ≤ 1. Hence
i.e. g satisfies (13.4) with α = 1/2. One can therefore apply Theorem 13.3,
since the eigenvalues of the Stokes operator on a two-dimensional periodic
domain satisfy λn ∼ cn (see Exercise 13.3).
A first version of the result of part (iii) of Theorem 13.3 was proved in Pinto
de Moura & Robinson (2010a), using the dynamical ‘squeezing property’ due
to Eden et al. (1994), see Exercise 13.5. One could also appeal more directly to
results on families of approximate inertial manifolds of exponential order due
to Debussche & Temam (1994) and Rosa (1995).
Exercises
13.1 Suppose that Aβ is Lipschitz on A for some β > 0:
for all t ≥ 0. [Hint: consider y(t) = z(t) + (δ/γ )1/2 for t ∈ [0, t0 ] with
t0 chosen so that z(t) ≥ 0.] A more general version of this result, due to
Ghidaglia, can be found as Lemma III.5.1 in Temam (1988).
13.3 The eigenvalues of the Stokes operator on a two-dimensional periodic
domain [0, 2π ]2 are the sums of two square integers. If {λn } are these
eigenvalues arranged in nondecreasing order, show that 12 n ≤ λn ≤ 2n.
13.4 Use Hölder’s inequality and the definition of the fractional powers of A
in (10.2) to show that if u ∈ D(Aβ ) then for any α < β
13.5 Let S(t) be the semigroup generated by (10.4), and assume the existence
of a global attractor A ⊂ BH (0, M). Eden et al. (1994) show that there
exists a time t ∗ and an n0 such that for all n ≥ n0 there exists an orthogonal
144 Thickness exponents of attractors
145
146 The Takens Time-Delay Embedding Theorem
This expresses Mx as j ξj êj where the {êj } are orthonormal vectors in the
directions of Mej ; clearly
ξj 2
≤ 1,
j
αj
We now use this bound to prove the following lemma, which is the main
component of the proof of the Takens time-delay theorem; it is a version of
Lemma 4.6 from Sauer et al. (1991), but valid for Hölder continuous maps
{Fj }. In some sense it is a generalised version of Theorem 4.3, but without
the Hölder continuity of the inverse; the embedding part of that theorem is an
immediate corollary if one takes F0 , . . . , Fm to be a basis for the linear maps
148 The Takens Time-Delay Embedding Theorem
from Rk into Rn (so that θ = 1), and notes that in this case M(x,y) always has
rank k (see Exercise 14.2).
has rank at least r, and suppose that S0 = ∅ and dB (Sr ) < rθ for 1 ≤ r ≤ k.
Then for almost every α ∈ Rm the map Fα : X → Rk given by
m
Fα = F0 + αj Fj
j =1
is one-to-one on A.
We follow Sauer et al. (1991) and say that ‘Gα has property Ŵ with prob-
ability p’ if the Lebesgue measure of the set of α ∈ Bm (R) for which Gα has
property Ŵ is p times the measure of Bm (R).
Since the rank of Mz is at least r for every z ∈ Sr , the rth largest singular value
is always positive, and so Sr = ∪∞ j =1 Sr,j .
Note that
Gα (z) = G0 (z) + Mz α,
and so Lemma 14.3 implies that for z ∈ Sr,j the probability that |Gα (z)| < δ is
no larger than Cm,n (j δ/R)r .
Now fix r and j , and choose d with dB (Sr ) < d < rθ . Since d > dB (Sr )
there exists an ǫ0 > 0 such for any 0 < ǫ < ǫ0 , S r,j can be covered by no more
14.1 The finite-dimensional case 149
(k ∈ N), and assume that the set Xp of p-periodic points of g (i.e. x ∈ X such
that g p (x) = x) satisfies dB (Xp ) < pθ 2 /2 for all p = 1, . . . , k.
Let h1 , . . . , hm be a basis for the polynomials in N variables of degree at
most 2k, and given any θ -Hölder function h0 : RN → R define
m
hα = h0 + αj hj .
j =1
is one-to-one on X.
In order to apply Lemma 14.4 we need to check, for each x = y, the rank of
the matrix
with all of the z1 , . . . , zq distinct (we have q ≤ 2k), and where J(x,y) is a
k × q matrix each of whose rows consists of zeros except for one 1 and one
−1. Given any ξ ∈ Rq , we can find1 a set of coefficients {αj }m j =1 such that
j αj hj (zl ) = ξl , i.e. such that H(x,y) α = ξ . This implies that the rank of H is
q; since J : Rq → Rk , we only need to check the rank of J .
We split the set {(x, y) : x, y ∈ X, x = y} into three disjoint sets of pairs
(x, y), and show that Fα (x) = Fα (y) for almost every α on each of these sets.
It then follows that Fα (x) = Fα (y) for almost every α, for any x, y ∈ X with
x = y.
Case 1: x and y are not both periodic of period ≤ k. In this case with-
out loss of generality {x, g(x), . . . , g k−1 (x)} consists of k discrete points and
{y, . . . , g k−1 (y)} consists of r ≥ 1 points distinct from the iterates of x. So
⎛ ⎞
1 0 ··· ··· 0 −1 0 · · · · · · 0
⎜0 1 ··· ··· 0 0 −1 · · · · · · 0 ⎟
⎜ ⎟
⎜ .. .. ⎟
J(x,y) = ⎜ 0 0
⎜ . ··· 0 0 0 . ··· 0⎟ ⎟,
⎜. . .. .. .. .. .. .. .. ⎟
⎝ .. .. . . 0 . . . . .⎠
0 0 ··· 0 1 0 · · · −1 · · · 0
where the left block is k × k and the right block is r × k (with the top r × r
entries being minus the identity).
It follows that the rank of J : Rk+r → Rk is k, and so is the rank of F = J H .
Since the set of pairs x = y has box-counting dimension at most 2d, and we
have just shown that rank M(x,y) = k > 2d/θ by assumption, the conditions of
Lemma 14.4 are met for this choice of (x, y), i.e. for all such (x, y), Fα (x) =
Fα (y) for almost every α.
Case 3: x and y lie on the same periodic orbit of period ≤ k. Suppose that
p and q are the minimal integers such that g p (x) = x and g q (x) = y, with
1 ≤ q < p ≤ k. As an illustrative example, if p = 7 and q = 4, then J is of
1 One can make a linear change of coordinates so that the first components of z1 , . . . , zq are
distinct, and then simply interpolate in one variable.
152 The Takens Time-Delay Embedding Theorem
the form
⎛ ⎞
1 0 0 0 −1 0 0
⎜ 0 1 0 0 0 −1 0 ⎟
⎜ ⎟
⎜ 0 0 1 0 0 0 −1 ⎟
⎜ ⎟
⎜ ⎟
J(x,y) = ⎜ −1 0 0 1 0 0 0 ⎟.
⎜ ⎟
⎜ 0 −1 0 0 1 0 0 ⎟
⎜ ⎟
⎝ 0 0 −1 0 0 1 0 ⎠
0 0 0 −1 0 0 1
The rank of such a matrix is at least p/2, and hence rank M(x,y) ≥ p/2 in
this case. The box-counting dimension of the set of pairs that lie on the same
periodic orbit is bounded by dB (Ap )/θ , since any such (x, y) is contained in
the image of Ap under one of the mappings x → (x, g j (x)) for some j =
1, . . . , k.
so that there must be some x ∗ ∈ Ŵ with h̃(x ∗ ) = 0. But this implies that h(x ∗ ) =
h(g(x ∗ )), so that x ∗ and g(x ∗ ) (which are distinct points) are mapped to the
same point in Rk whatever the value of k.
So it is interesting if we can find a T sufficiently small that there are no
periodic orbits of any period ≤ T . Yorke (1969) showed that if the ordinary
14.2 Periodic orbits and the Lipschitz constant for ODEs 153
differential equation
has a periodic orbit of period T , then one must have T ≥ 2π/L. (Yorke also
gives an example to show that the factor of 2π is sharp.) We give a version
of this result here, following ideas in Kukavica (1994); the result is no longer
sharp (we only obtain T ≥ 1/L), but this sacrifice seems worthwhile given the
simplicity of the proof, which also provides a model for the proof of a similar
infinite-dimensional result (Theorem 14.8).
Theorem 14.6 Any periodic orbit of the equation ẋ = f (x), where f has
Lipschitz constant L, has period T ≥ 1/L.
i.e.
" T " T
|x(t) − x(t − τ )| ≤ |v̇(s)| ds = |f (x(s)) − f (x(s − τ ))| ds
0 0
" T
≤L |x(s) − x(s − τ )| ds.
0
Therefore
" T " T
|x(t) − x(t − τ )| dt ≤ LT |x(s) − x(s − τ )| ds,
0 0
is one-to-one on X.
Now consider the k-fold observation map on A given by
T
Fk [hα , g](Lx) = hα (Lx), hα (L(x)), · · · , hα (Lk−1 (x)) .
Since L is one-to-one between A and X, and Fk [h, α, g] is one-to-one between
X and its image, it follows that Fk ◦ L is one-to-one between A and its image.
If we define fj (x) = hj (Lx), then each fj is a Lipschitz map from A into
Rk , and we can write
T
Fk [hα , g](Lx) = Dk [fα , ](x) = fα (x), fα ((x)), . . . , fα (k−1 (x)) ,
where
M
fα = f0 + αj fj .
j =1
Then if we take
⎧ ⎫
⎨M ⎬
E= αj fj : (α1 , . . . , αM ) ∈ BM (0, 1)
⎩ ⎭
j =1
156 The Takens Time-Delay Embedding Theorem
and equip E with the measure induced by the uniform measure on BM (0, 1), it
follows that a prevalent set of Lipschitz f : A → Rk make the map Dk [f, ]
one-to-one on A .
Note that the condition on the number of delay coordinates required increases
with the dual thickness of the set A . In the case when A has zero dual
thickness (so, for example, if A is ‘smooth’ so that Lemma 13.1 guarantees
that its thickness is zero, or if the equation is in the right form that Theorem
13.3 shows that the Lipschitz deviation of A is zero) then the condition on
k reduces to the k > 2d one would obtain in the finite-dimensional Lipschitz
case (Theorem 14.5 with θ = 1).
Theorem 14.8 For each α with 0 ≤ α ≤ 1/2 there exists a constant Kα such
that if
g(u) − g(v) ≤ LAα (u − v) for all u, v ∈ D(Aα )
then any periodic orbit of (14.2) must have period at least Kα L−1/(1−α) .
While we assume that g is uniformly Lipschitz, this uniformity need only
hold for u, v contained in the (necessarily bounded) periodic orbit.
Proof On a periodic orbit of period T we have
" T
u(t) = u(t + T ) = e−AT u(t) + e−A(T −s) g(u(s + t)) ds,
0
14.4 Periodic orbits in semilinear parabolic equations 157
and so
" T
(I − e−AT )u(t) = e−A(T −s) g(u(s + t)) ds.
0
It follows that
u(t) − u(t + τ )
" T
= (I − e−AT )−1 e−A(T −s) [g(u(t + s)) − g(u(t + τ + s))] ds.
0
Since u is T -periodic,
" T " T
g(u(s + t)) ds = g(u(s + t + τ )) ds,
0 0
Therefore
u(t) − u(t + τ )
" T
6 7
= (I − e−AT )−1 (e−A(T −s) − cI ) (g(u(s + t)) − g(u(s + t + τ ))) ds.
0
|(Aα D(t), wk )|
" T 1/2 " T 1/2
λαk −λk s 2 2
≤ (e − c) ds (G(t + s), wk ) ds .
1 − e−λk T 0 0
158 The Takens Time-Delay Embedding Theorem
where
' (1/2
μα 1 − e−2μ (1 − e−μ )2
(μ) := − .
1 − e−μ 2μ μ2
Now, (μ)√ is bounded on [0, ∞) by some constant Cα : it is clear that (μ) ∼
α−1/2
μ √/ 2 as μ → ∞, while a careful Taylor expansion shows that (μ) ∼
μα /2 3 as μ → 0 (see Exercise 14.4).
It follows that for each k ∈ N
" T
|(Aα D(t), wk )|2 ≤ Cα2 T 1−2α |(G(t + s), wk )|2 ds.
0
Now integrate the left- and right-hand sides of this expression with respect to
t between t = 0 and t = T to obtain
" T " T
|Aα D(t)|2 dt ≤ Cα2 T 2−2α L2 |Aα D(s)|2 ds.
0 0
1−α
Therefore if Cα T L < 1 we must have
" T
|Aα (u(t) − u(t + τ ))|2 dt = 0.
0
It follows that u(t) = u(t + τ ) for all t ∈ [0, T ], and since this holds for any
τ > 0, u(t) must be a constant orbit. Therefore any periodic orbit must have
period at least Kα L−1/(1−α) .
Exercises
14.1 Show that the nonzero singular values of M are also the square roots of
the eigenvalues of MM T .
Exercises 159
The aim of this final chapter is to show that if the attractor A consists of
real analytic functions defined on some domain then one can parametrise
the attractor using a sufficient number of point values. We will show that if k
is large enough (proportional to the box-counting dimension of A ) then for
almost every choice of k points {xj } in the mapping
u → (u(x1 ), . . . , u(xk ))
is an embedding of A into Rkd .
More precisely, we will give a proof of the following theorem, first proved
(in a slightly different form) by Friz & Robinson (2001). The proof makes use
of a number of the ideas that have already been discussed: both the Hausdorff
and box-counting dimensions, the thickness and dual thickness, and, of course,
the embedding result of Theorem 8.1. (For a related result in the context of
purely analytic systems see Sontag (2002).)
Theorem 15.1 Let A be a compact subset of L2per (, Rd ) with dB (A )
finite. Suppose also that A consists of real analytic functions. Then, for
k ≥ 16dB (A ) + 1 almost every set x = (x1 , . . . , xk ) of k points in makes
the map Ex , defined by
Ex [u] = (u(x1 ), . . . , u(xk ))
one-to-one between X and its image.
8
In the statement of the theorem, we set = m j =1 [0, Lj ], and denote by
L2per (, Rd ) those functions in L2loc (Rm , Rd ) that are periodic with period
Lj > 0 in the {ej } direction,
u(x + Lj ej ) = u(x) for all j = 1, . . . , m.
‘Almost every’ is with respect to Lebesgue measure on k .
160
15.1 Real analytic functions and the order of vanishing 161
Although we give the result for attractors that consist of periodic functions,
with a little additional work essentially the same techniques can be used to
prove a more general result, valid (for example) on bounded domains with
Dirichlet boundary conditions, see Kukavica & Robinson (2004).
Before starting the proof proper, we give an idea of why the analyticity
condition is required. Suppose that we have chosen (x1 , . . . , xk ), and that these
are in fact a ‘bad’ set of points, which means that there are u, v ∈ A with
u = v such that
and
|D β fk (x)| ≤ M|β|!τ −|β| for all k = 1, . . . , d. (15.3)
w : K × → Rd ,
has order of vanishing at most j < ∞, and is such that ∂ α w(x; p) depends on
p in a θ -Hölder way for all |α| ≤ j . Then the zero set of w(x; p), i.e.
clearly Aj +1 ⊇ Aj , and A = ∪∞
j =1 Aj .
Thus, using the simple results of (3.5) and (7.2) (on box-counting dimension
and thickness for sets considered as subsets of different spaces) it is sufficient to
prove the lemma with C r (, Rd ) replaced by D(Ar ). We note here that since
functions in Aj enjoy uniform bounds on their derivatives, Aj is uniformly
bounded in D(Ar ) for each r ∈ N,
We start with the box-counting dimension. If s > r, then for any u ∈ D(As )
we have the interpolation inequality
(see Exercise 13.4). It follows from (15.6) that the identity map from Aj onto
itself is Hölder continuous as a map from L2 (, Rd ) into D(Ar ) with Hölder
exponent as close to 1 as we wish. That the box-counting dimension of Aj in
the space D(Ar ) is bounded by dB (Aj ; L2 ) ≤ dB (A ; L2 ) is a consequence of
part (iv) of Lemma 3.3.
In order to show that the thickness exponent is zero, let Pn denote the
projection onto the space spanned by the first n eigenfunctions of A,
n
Pn u = (u, wj )wj
j =1
Therefore,
CRk+r
distD(Ar ) Aj , Pn D(Ar ) ≤ 2k/d ,
n
15.3 Proof of Theorem 15.1 165
Ex (u) = Ex (v)
Thus, using Proposition 7.10 (zero thickness implies zero dual thickness),
τ ∗ (Wj,r ; C r (, Rd )) = 0, and hence Theorem 8.1 (embedding with Hölder
continuous inverse) guarantees that for any
(x ′ , xj (x ′ ; ε); ε),
N + (m − 1)k + k(1 − θ ),
and using the fact that the Hausdorff dimension is stable under countable unions
(Proposition 2.8(iii)) the same goes for the whole countable collection.
The projection of this collection onto k enjoys the same bound on its
dimension (since Lipschitz maps do not increase the Hausdorff dimension,
Proposition 2.8(iv)), and so to make sure that these ‘bad choices’ do not cover
k ⊂ Rmk we need
15.4 Applications
15.4.1 Determining nodes
The theorem provides an instantaneous version of the ‘determining nodes’ intro-
duced by Foias & Temam (1984): they called a collection of points {x1 , . . . , xk }
in (asymptotically) ‘determining’ if for two solutions u(x, t) and v(x, t),
max |u(xj , t) − v(xj , t)| → 0 as t →∞ (15.8)
j =1,...,k
implies that
sup |u(x, t) − v(x, t)| → 0 as t → ∞.
x∈
Foias & Temam showed that for the two-dimensional Navier–Stokes equations
there exists a δ such that if for every x ∈
|x − xj | < δ for some j ∈ {1, . . . , k}
then the collection of nodes is determining.1 Under a mild additional condition
our ‘instantaneous determining nodes’ are also asymptotically determining.
Lemma 15.6 Suppose that the conditions of Theorem 15.1 hold, and that the
attractor A attracts solutions in the norm of L∞ (). Then almost every set of
k nodes {x1 , . . . , xk } in is asymptotically determining.
Proof Since A is a compact subset of L∞ , the map Ex−1 : Rkd → L∞ () is
continuous. Thus given any ǫ > 0 there exists a δ, 0 < δ < ǫ, such that for
u, v ∈ A ,
ǫ
max |u(xj ) − v(xj )| < δ ⇒ u − v∞ < .
j =1,...,k 3
Now let u(t) and v(t) be two solutions that agree asymptotically on the nodes
x1 , . . . , xk as in (15.8). Since A attracts in L∞ (), there exists a time T > 0
1 In the same paper they conjectured that for solutions on the attractor coincidence of the values
of u and v at k points (for k large enough) should imply coincidence of u and v; our theorem
proves this conjecture when the attractor consists of real analytic functions. A proof for systems
that possess an inertial manifold was given by Foias & Titi (1991).
168 Parametrisation of attractors via point values
It follows that u∗ (t) − v ∗ (t)L∞ < ǫ/3, which combined with (15.9) shows
that u(t) − v(t)L∞ ≤ ǫ for all t ≥ T .
l ∼ [dB (A )]−1/n .
Exercises 169
Exercises
15.1 Let α, β be n-component multi-indices, and x ∈ Rn with |xi | < 1 for all
i = 1, . . . , n. Show that
α! β!
x α−β = , (15.10)
α: α≥β
(α − β)! (1 − x)1+β
1
where 1 = (1, . . . , 1) (n times). [Hint: the left-hand side is D β (1−x)1 .]
n
15.2 Suppose that f : R → R is real analytic at x with
f (y) = cα (y − x)α (15.11)
α≥0
for all |y − x| ≤ ǫ. Fix q ∈ (0, 1). Show that for any multi-index β ≥ 0,
for |y − x| ≤ qǫ, the derivative D β f (y) can be obtained from term-by-
term differentiation of (15.11), and
|D β f (y)| ≤ M|β|!τ −|β| ,
where
μ
M= and τ = (1 − q)ǫ.
(1 − q)n
Deduce that cα = (1/α!)D α f (x).
15.3 Suppose that f ∈ C ω (). Show that for any compact subset K ⊂ there
exist positive constants M and τ such that
|D β f (x)| ≤ M|β|!τ −β
for every x ∈ K.
Solutions to exercises
F1 ⊆ V1 ⊆ V1 ⊆ U1 ∩ (X \ ∩n+2
i=2 Fi ).
9n+2
So V1 ⊂ U1 and V1 ∩ i=2 Fi = ∅. Now, there exists an open set V2 with
n+2
F2 ⊆ V2 ⊆ V2 ⊆ U2 ∩ X \ V1 ∩ Fi ,
i=3
9n+2
and so V2 ⊆ U2 and V1 ∩ V2 ∩ i=3 Fi = ∅. Continuing in this way shows
that one can take the {Fi } open in the original assumption.
Now let {U1 , . . . , Uk } be an open cover of X. If k ≤ n + 1 then this cover
already has order ≤ n, so we can assume that k ≥ n + 2. Set Vj = Uj for
j = 1, . . . , n + 1, and
k
Vj +2 = Ui .
i=n+2
These sets cover X, and so there exist open sets {Fi }n+2
i=1 such that Fi ⊆ Vi ,
X ⊆ ∪n+2 F
i=1 i , and ∩n+2
F
i=1 i = ∅. Let W i = F i for i ≤ n + 1 and Wi = Fn+2 ∩ Ui
for i ≥ n + 2. Then for every i, Wi ⊆ Ui ,
k
n+2
X⊆ Wi , and Wi = ∅.
i=1 i=1
One can perform the same construction for every subset of {1, . . . , k} consisting
of n + 2 elements, to deduce that every intersection of n + 2 of the {Wi } is
empty, and hence that dim(X) ≤ n.
170
Solutions to exercises 171
(n, n + 1) × (m, m + 1) n, m ∈ Z;
all the elements of A2 are disjoint. Let α1 consist of all the open edges of these
squares,
but expanded to open subsets of R2 , such that the resulting sets are pairwise
disjoint. Let α0 be the collection of all open balls of radius 21 about the points
(n, m) of the integer lattice. Then α = α0 ∪ α1 ∪ α2 is a cover of R2 of mesh
size 1 and of order 3, see the figure below.
α2 α1 α0
The three sets α0 , α1 , and α2 that provide a cover of R2 of order 3.
A \ K ⊂ U ∪ U ′, A ∩ C ⊂ U, and A ∩ C′ ⊂ U ′.
: ;
2.1 Let ǫ > 0. Since dH (X) ≤ n, we can cover X by a collection B(xi , ri ) i∈I
of balls with centres xi ∈ X such that
rin+ǫ < ∞.
i∈I
Since
mi
(2ri )n+(1−θ)m+ǫ = 2n+(1−θ)m+ǫ mi rin+(1−θ)m+ǫ
i∈I j =1 i∈I
≤ K2n+(1−θ)m+ǫ C m rin+ǫ + rin+(1−θ)m+ǫ < ∞,
i∈I i∈I
n := [0, ∞) \ (J0 ∪ · · · ∪ Jn ).
2.3 Clearly Hδd (X) ≤ 2d Sδd (X), since any cover by balls of radius δ provides
a 2δ-cover of X. Also, if {Ui } is a δ-cover of X then any Ui is contained in some
ball of radius δ, so that Sδd (X) ≤ Hδd (X). It follows that S d (X) ≤ H d (X) ≤
2d S d (X), and so the value of d at which S d (X) jumps from ∞ to 0 is the
same as that at which H d (X) makes the same jump.
while the right-hand side of (S.3) is non-empty. The balls {B(xi ), r(xi )} are
disjoint by definition, and lie in a compact subset of X, so there can be only a
finite number of them, say k1 . Thus
k1
A1 ⊆ B(xi , r(xi )).
i=1
Now let
A2 = {x ∈ X : ( 43 )2 M < r(x) ≤ 34 M}
and
k1
A′2 = {x ∈ A2 : B(x, r(x)) ∩ B(xi , r(xi )) = ∅}.
i=1
Thus
k1
A2 \ A′2 ⊆ B(xi , 3r(xi )). (S.4)
i=1
As above, there exists a k2 such that the balls B(xi , r(xi )), i = 1, . . . , k2 are
disjoint and
k2
A′2 ⊆ B(xi , 3r(xi )).
i=k1 +1
Now, using the result of the previous exercise, find a disjoint subcollection of
these balls {Bri (xi )} such that S is still covered by {B5ri (xi )}. Since these balls
are disjoint,
"" ""
|f | ≥ |f | ≥ δ rid .
V i Bri (xi ) i
3.1 In each case it suffices to show that there exist constants c1 , c2 > 0 and
α1 ≥ 1, α2 ≤ 1 such that
(i) It is clear that N(A, ǫ) ≤ M(A, ǫ). In order to prove the lower inequality
in (S.5) consider a cover of A by N (A, ǫ) balls of radius ǫ, B(xi , ǫ).
Discarding any unnecessary balls from this cover, each ball B(xi , ǫ) must
contain a point yi ∈ A. Since
3.2 Take d > dB (X). Then for ǫ sufficiently small X can be covered by ǫ −d
balls of radius ǫ centred in X. It follows that O(X, ǫ) can be covered by ǫ −d balls
of radius 2ǫ, and so L n (O(X, ǫ)) ≤ ǫ −d (2ǫ)n n , from whence c(X) ≤ n − d.
For the opposite inequality, if d < dB (X) then there is a sequence ǫj → 0 such
that there are at least ǫj−d disjoint balls of radius ǫj with centres in X: then
L n (O(X, ǫj )) ≥ ǫj−d n ǫjn , and c(X) ≥ n − d.
3.3 Let n = dim(X), and let Kn consist of all mappings f ∈ C(X, R2n+1 )
such that dB (f (X)) ≤ n. Now, by the characterisation of dLB given in the hint,
Solutions to exercises 177
dLB (K) ≤ n if and only if for every k ∈ N, there exists a ǫ > 0 such that
No (K, ǫ) ≤ ǫ −n /k. Let Kn,k be the class of all mappings f such that this
inequality holds for some ǫ > 0. Clearly Kn,k is open, and
∞
Kn = Kn,k .
k=1
Now, as noted during the proof of the embedding theorem for sets with
dim(X) finite, the embedding map g defined in (1.4) maps X into an n-
dimensional polyhedron. Thus the set of maps f ∈ C(X, R2n+1 ) that map X
into such a polyhedron is dense; for any such map,
3.5 Choose s > dH (X) and t > dB (Y ). Then there exists a δ0 > 0 such that
N(Y, δ) ≤ δ −t for all δ ≤ δ0 . Let {B(xi , ri )} be a cover of X such that
ris < 1,
i
which is possible since H s (X) = 0 for s > dH (X). Now for each i cover Y
with Ni := N(Y, ri ) balls of radius ri , {B(yi,j , ri )}N i
j =1 . Thus
X×Y ⊂ B(xi , ri ) × B(yj , ri ).
i j
178 Solutions to exercises
Thus
H2δs+t (X × Y ) ≤ (2ri )s+t
i j
≤ N(y, ri )2s+t ris+t
i
≤2 s+t
ri−t ris+t < 2s+t .
i
s+t
It follows that H (X × Y ) < ∞, and hence that dH (X × Y ) ≤ s + t. Since
s > dH (X) and t > dB (Y ) were arbitrary, this completes the proof.
3.6 Let X ⊆ ∪∞ i=1 Xi with each Xi closed. Then using the Baire Category
Theorem there is an index j and an open set U ⊂ Rn such that X ∩ U ⊂ Xj .
So dB (X) = dB (Xj ). It follows using the definition that dMB (X) ≥ dB (X), and
we already have the reverse inequality.
3.7 If X ⊆ ∪i Xi then
dP (X) ≤ sup dP (Xi ) ≤ sup dB (Xi ).
i i
It follows from the definition of dMB that dP (X) ≤ dMB (X). Conversely, suppose
that s > dimP (X). Then P s (X) = 0, and so we can find a collection of sets Xi
such that X ⊂ ∪i Xi with P0s (Xi ) < ∞ for each i. In particular, Nδ (Xi )δ s is
bounded as δ → 0 for each i, from which it follows that dB (Xi ) ≤ s for each
i, and hence dMB (X) ≤ s.
5.1 Take a probe space E of constant functions
E = {gc ∈ L1 (0, 1) : gc (x) = c for all x ∈ [0, 1], 0 ≤ c ≤ 1},
i.e. a set isometric to [0, 1], equipped with Lebesgue measure. Then
" "
f (x) + gc (x) dx = f (x) dx + c,
1
which is zero for at most one c ∈ [0, 1]. Thus f + g = 0 for almost every
gc ∈ E.
5.2 To show that E is compact it suffices to consider only one ‘component’
of L, i.e. to prove the compactness of E0 . Given a sequence l (n) ∈ E0 with
∞
l (n) = j −γ [φj(n) ]∗ , φj(n) ∈ Sj ,
j =1
since each Sj is compact one can extract successive subsequences and then use
a diagonal argument to find a subsequence (which we relabel) such that for
Solutions to exercises 179
where clearly l ∈ E0 .
5.3 Note that for each s one can view Ks as a subset of Rdj −1 , and that the
set (1 − t)Ka + tKb is precisely the intersection of the convex hull of Ka and
Kb with + ((1 − t)a + tb)γ , and so in particular is a subset of K(1−t)a+b . It
follows from the Brunn–Minkowski inequality that
L dj −1 (K(1−t)a+b )1/(dj −1) ≥ (1 − t)L dj −1 (Ka )1/(dj −1) + tL dj −1 (Kb )1/(dj −1) ,
i.e. that the map s → L dj −1 (Ks )1/(dj −1) is concave. Since Uj is symmetric this
map is also symmetric, and hence it attains its maximum value when s = 0.
7.1 Denote the right-hand side of (7.4) by τ̃ . Taking any σ ∈ (0, τ (X)), there
is a sequence ǫj ∈ (0, 1) converging to 0 such that
< =
d(X, ǫj ) > ǫj−σ ≥ ǫj−σ = nj ,
7.2 For any d > dB (X), there exists an ǫ0 such that for any ǫ < ǫ0 one can
cover X by no more than Nǫ = ǫ −d balls of radius ǫ, with centres {xj }N ǫ
j =1 .
n
Use the Johnson–Lindenstrauss Lemma to find a function f : H → R , where
n = O(ln Nǫ ), such that
1
xi − xj ≤ |f (xi ) − f (xj )| ≤ 2xi − xj for all i, j = 1, . . . , Nǫ .
2
The mapping f −1 |{f (x1 ),...,f (xN )} is 2-Lipschitz onto {x1 , . . . , xN }. In particular,
it can be extended to a 2-Lipschitz map from Rn into H . Since any x ∈ X lies
within ǫ of one of the {xj }, this shows that there exists a 2-Lipschitz mapping
ϕ : Rn → H such that
dist(X, ϕ(Rn )) ≤ ǫ.
7.4 Let {e1 , . . . , en } be an Auerbach basis for U , and {f1 , . . . , fn } the asso-
ciated elements of U ∗ such that fi U ∗ = 1 and fi (ej ) = δij . Extend each fi
to an element φi ∈ B ∗ with φi B∗ = 1, and define
n
Pu = φj (u)ej .
j =1
has dB (Xn ) ≤ d, and hence τ ∗ (Xn ) ≤ d (Lemma 7.9). It follows from Theorem
8.1 that if k > 2d and
k − 2d
0<θ <
k(1 + d)
there is a prevalent set of maps L : B → Rk that are injective on Xn and satisfy
and (9.14) follows using the triangle inequality. (In fact this inequality holds in
any CAT(0) space, see Sato (2009).)
9.2 First note that
|sj (x)| = |̺(x, xj ) − ̺(xj , x0 )| ≤ ̺(x, x0 ),
and so s(x) ∈ ℓ∞ . Then
|sj (x) − sj (y)| = |̺(x, xj ) − ̺(y, xj )|,
from which it follows immediately that s(x) − s(y)ℓ∞ ≤ ̺(x, y). The lower
bound s(x) − s(y)ℓ∞ ≥ ̺(x, y) follows using the fact that {xj } is dense:
in particular for any ǫ > 0 there exists a j ∈ N such that ̺(y, xj ) < ǫ, and
hence
|sj (x) − sj (y)| ≥ ̺(x, y) − 2̺(y, xj ) > ̺(x, y) − 2ǫ.
ikx
10.1 Since f (x) = k ck e it follows that
1 1/2
|f (x)| ≤ |ck | ≤ 2
(1 + |k|2 )1/2 |ck |
k
1 + |k|
1/2 1/2
1
≤ 2
(1 + |k|2 )|ck |2
k
1 + |k| k
≤ Cu2H 1 .
Since |k|≤n ck eikx is continuous for each n, f is the uniform limit of continuous
functions, so continuous.
10.2 Take a sequence {un }∞ β
n=1 with un bounded in D(A ). Let
∞
n
2β
un = cn,j wj , so that Aβ un 2 = λj |cn,j |2 ≤ M
j =1 j =1
∞ ∗ 2β
Let u = n=1 cn wn ; note that n λn |cn∗ |2 ≤ M. Now,
∞
Aα (un − u)2 = λ2α ∗ 2
j |cn,j − cn |
n=1
m ∞
−2(β−α) 2β
≤ λ2α ∗ 2
j |cn,j − cn | + λm+1 λj |cn,j − cn∗ |2
n=1 n=m+1
m
−2(β−α)
≤ λ2α ∗ 2
j |cn,j − cn | + 2Mλm+1 .
n=1
Given ǫ > 0, choose m sufficiently large that the second term is ≤ ǫ/2; one
can then choose j large enough to ensure that the first term (with involves only
a finite number of coefficients) is also ≤ ǫ/2.
10.3 Since
d γ −λt
(λ e ) = γ λγ −1 e−λt − tλγ e−λt = λγ e−λt [γ λ−1 − t],
dλ
and so
" t
Aβ u(t) = Aβ e−At u0 + Aβ e−A(t−s) g(u(s)) ds.
0
184 Solutions to exercises
and so u(t) → 0 as t → ∞.
Thus
2
Du2 = Dj u2 = |k|2 |ck |2 ≥ |ck |2 = u2 .
j =1 k∈Ż2 k∈Ż2
dist(S(t)X, A ) = dist(X, A ) → 0 as t → ∞,
dist(S(t)A , Y ) = dist(A , Y ) → 0 as t → ∞,
11.3 Suppose that A is not connected. Then there exist open sets O1 and O2
such that O1 ∩ A = ∅ and O2 ∩ A = ∅,
O1 ∪ O2 ⊃ A , and O1 ∩ O2 = ∅.
Since the compact attracting set K ⊂ B(0, R) for some R > 0, and a ball in
B is connected, it follows that S(t)B(0, R), the continuous image of B(0, R),
is connected. Since A attracts B(0, R), for t sufficiently large, S(t)B(0, R)
is contained either wholly in O1 or wholly in O2 . Since B(0, R) ⊃ K,
ω(B(0, R)) ⊃ ω(K) and hence A = ω(B(0, R)) is contained in either O1 or
O2 , contradicting our initial assumption.
as t → ∞, and hence u0 ∈ A .
11.5 1Rearrange the governing inequality and multiply by the integrating factor
t
exp(− s a(u) du):
d % − 1 t a(u) du & 1t
e s x(t) ≤ b(t)e− s a(u) du ≤ b(t).
dt
Integrate this inequality from s to t + r, so that
1 t+r
" t+r
e− s a(u) du x(t + r) ≤ x(s) + b(u) du ≤ x(s) + B,
s
186 Solutions to exercises
i.e.
1 t+r
a(u) du
x(t + r) ≤ e s [x(s) + B] ≤ eA [x(s) + B].
Now integrate again from t to t + r with respect to s to obtain
rx(t + r) ≤ eA [X + rB],
and hence x(t + r) ≤ eA [B + (X/r)].
11.6 (i) Returning to (11.9), we retain the term in Au2 to give
d
Du2 + Au2 ≤ f 2 .
dt
Integrating from 0 to 1 then yields
" 1
2
Du(1) + Au(s)2 ds ≤ f 2 + Du0 2 ≤ 7f 2 .
0
using the orthogonality property (10.17) and the inequality (11.11). Since
Du(t)2 ≤ 6f 2 it follows that
d
ut 2 + Dut 2 ≤ c22 ut 2 Du2 ≤ 6c22 f 2 ut 2 .
dt
Dropping the Dut 2 and integrating with respect to t from s to 1 yields
" 1
2 2 2 2
ut (1) ≤ ut (s) + 6c2 f ut (r)2 dr,
s
(iv) Once again we use the triangle inequality on (11.6), this time to obtain
12.1 Clearly, for each λ > 0 and x ∈ K, Df (x) ∈ Lλ/2 (X) for all λ > 0 and
νλ (Df (x)) = ν(x). Consequently, for each 0 < λ < 12 ,
log (ν + 1) Dλ
dB (K) ≤ ν .
log(1/2λ)
Taking the limit as λ → 0 we obtain dB (K) ≤ ν.
1t
12.2 For Y (t) to be a supersolution of ẏ = a + b 0 (t − s)α y(s) ds we require
" t
2aeKt ≥ a + 2ab (t − s)−α eKs ds,
0
i.e.
" t " t
2 ≥ e−Kt + 2b (t − s)−α e−K(t−s) ds = e−Kt + 2bK −(1−α) u−α e−u du.
0 0
188 Solutions to exercises
1∞
Since Ŵ(z) = 0 t z−1 e−t dt, this is certainly ensured if
2bŴ(1 − α)
2 ≥ e−Kt + .
K 1−α
So it suffices to choose K = (2bŴ(1 − α))1/(1−α) . (In the case considered in this
exercise, this argument, due to Robinson (1997), offers a significantly simpler
proof than that due to Henry (1981, Lemma 7.1.1); while Henry’s result is
sharper, the bound here is often sufficient in applications.)
12.3 Using the bounds on Aα e−At in (10.9) and (10.10), there exists a
constant c > 0 such that
Therefore
4j ωj (Df (u))
j/d
,
ω̄d
j/d 1/d
where j is the largest integer such that ω̄d ≤ αj . Since ω̄d ≥ ᾱd , it follows
that j ≤ d. So no more than
4j ωj (Df (uj )) ω̄j
max j/d
≤ max 4j j/d
=: M
1≤j ≤d ω̄d 1≤j ≤d ω̄d
√ 1/d
balls of radius 2ω̄d are required to cover Df (u)[B(0, 1)].
Solutions to exercises 189
1/d
Alternatively, if ω̄d ≥ α1 (Df (u)) then
1/d
Df (u)[B(0, 1)] ⊆ B(0, α1 (Df (u))) ⊆ B(0, ω̄d ),
√ 1/d
and it requires only one ball of radius 2ω̄d to cover Df (u)[B(0, 1)] in this
case.
1/d
Thus Df (u)[B(0, 1)] can always be covered by M balls of radius 2ω̄d .
12.5 We need to have θ < 1 in order to apply Lemma 12.1. Using the hint
repeatedly, note that for each u ∈ K
and
1/d ¯j k/d γ ω̄jk
¯ d )γ max 4j
(2 = (2ω̄d ) max 4j
1≤j ≤d ¯ j/d
d
1≤j ≤d ω̄d
kj/d
* +k/d
γ d γ ω̄jd
≤ 2 4 ω̄d max j < 1. (S.8)
1≤j ≤d ω̄
d
We now apply Lemma 12.1, making use of the observation in the footnote.
12.8 Given a set {φj }nj=1 that is orthonormal in H , for any u ∈ A we have
n
n
n
(φj , −Aφj + Dg(u)φj ) ≤ − A1/2 φj 2 + M Aα φj φj
j =1 j =1 j =1
n
n
≤− A1/2 φj 2 + M A1/2 φj 2α φj 2−2α
j =1 j =1
⎛ ⎞α
n
n
≤− A1/2 φj 21/2 +M ⎝ A1/2 φj 2 ⎠ n1−α
j =1 j =1
⎡ ⎤
n
≤ (1 − α) ⎣− A1/2 φj 2 + M 1/(1−α) n⎦
j =1
⎡ ⎤
n
1/(1−α)
≤ (1 − α) ⎣− λj + M n⎦ .
j =1
It follows from (12.13) that the final line provides an upper bound for
qn (DS(1; u0 )) which is uniform over all u0 ∈ A . Since this bound is a con-
cave function of n, it follows from Exercise 12.6 that dB (A ) ≤ n once the
right-hand side is negative, which gives (12.15).
and
Whence
2β
(λn+1 − L2 )Qn w2 ≤ L2 Pn w2 .
2β
If n is large enough that λn+1 > 2L2 then
13.2 If y(0) ≤ (δ/γ )1/2 then clearly y(t) ≤ (δ/γ )1/2 for all t ≥ 0. If y(0) ≥
(δ/γ )1/2 then there exists t0 ∈ (0, ∞) such that
and
and so
2 2 δ
ż + γ z ≤ ẏ + γ y − ≤ 0.
γ
Integrating ż + γ z2 ≤ 0 yields
1 1
z(t) ≤ ≤ .
z0−1 + γ t γt
This implies (13.15) for t ∈ [0, t0 ], and since y(t) ≤ (δ/γ )1/2 for all t ≥ t0 the
result follows.
13.3 The eigenvalues are the sums of two square integers (positive and
negative); so we will have reached 2k 2 once we have taken [2(k − 1)]2 + 1
combinations of integers with modulus ≤ k. So if 4(k − 1)2 < n ≤ 4k 2
then
since k − 1 ≤ (n/4)1/2 ≤ k,
1 1/2
2
n ≤ k < k + 1 < n1/2 ,
∞
13.4 Put u = j =1 cj wj . Then
∞
Aα u2 = λ2α
j |(u, wj )|
2
j =1
∞
= λ2α
j |cj |
2α/β
|cj |2(1−(α/β))
j =1
⎛ ⎞β/α ⎛ ⎞1−(α/β)
∞ ∞
2β
≤⎝ λj |cj |2 ⎠ ⎝ |cj |2 ⎠
j =1 j =1
β α/β 1−(α/β)
= A u u ,
using Hölder’s inequality with exponents (α/β, 1/(1 − α/β)).
13.5 Fix an n ≥ n0 , and let X be a subset of A that is maximal for the relation
Qn (u − v)α ≤ Pn (u − v)α for all u, v ∈ X.
As in the proof of Theorem 13.3(ii), it follows that there exists a 1-Lipschitz
function n : Pn H → Qn H such that X ⊂ GPn H [n ].
If u ∈ A but u ∈/ X then there is a v ∈ X such that
Qn (u − v)α ≥ Pn (u − v)α . (S.10)
Since S(t ∗ )A = A (because A is invariant) there exist ū, v̄ ∈ A such that
u = S(t ∗ )ū and v = S(t ∗ )v̄; since (S.10) implies that (13.16) cannot hold, it
follows from (13.17) that
u − v ≤ δn ū − v̄ ≤ 2Mδn .
Since v ∈ GPn H [n ] this implies that A lies within a 4Mδn neighbourhood of
GPn H [n ] as claimed.
13.6 Given a u∗ ∈ H 2 \ H 3 , set f = Au∗ + B(u∗ , u∗ ). Then f ∈ L2 and u∗
is a stationary solution of the equations. The attractor must contain u∗ , and
hence cannot be bounded in H 3 .
13.7 Assume that w(0) = 0; we will show that w(t) = 0 for any t ≥ 0.
Taking the inner product of (13.7) with w yields
1 d
w2 + A1/2 w2 = (w, h(t)).
2 dt
Dividing by L(t)w2 we obtain
1 d 2 (w, h(t)) 1 h2
w + Q̃(t) = ≥ − − ,
2L(t)w2 dt L(t)w2 L(t) L(t)w2
Solutions to exercises 193
using Young’s inequality (ab ≤ (a p /p) + (bq /q) when p−1 + q −1 = 1). Thus
1 d
w2 + cQ̃(t) ≥ −c′ .
2L(t)w2 dt
14.1 If M T Me = λe then
14.4 We have
* +1/2
μα 2μ − 2μ2 + 34 μ3 − · · · (μ − 21 μ2 + 16 μ3 − · · · )2
(μ) ∼ −
μ − ··· 2μ μ2
μα 6 71/2
= (1 − μ + 32 μ2 − · · · ) − (1 − 12 μ + 61 μ2 )2
μ − ···
μα 6 71/2
= (1 − μ + 32 μ2 − · · · ) − (1 − μ + 14 μ2 + 13 μ2 + . . .)
μ − ···
μα 1 2 μα
= [ 12 μ + · · · ]1/2 ∼ √ .
μ − ··· 2 3
using (15.10). It follows that for |y − x| ≤ qǫ, D β f (y) is continuous and given
by
α!
D β f (y) = cα (y − x)α−β .
α≥β
(α − β)!
Assouad, A. (1983) Plongements lipschitziens dans Rn . Bull. Soc. Math. France 111,
429–448.
Ball, K. (1986) Cube slicing in Rn . Proc. Amer. Math. Soc. 97, 465–473.
Ben-Artzi, A., Eden, A., Foias, C., & Nicolaenko, B. (1993) Hölder continuity for the
inverse of Mañé’s projection. J. Math. Anal. Appl. 178, 22–29.
Benyamini, Y., & Lindenstrauss, J. (2000) Geometric Nonlinear Functional Analysis,
Vol. 1., AMS Colloquium Publications 48 (Providence, RI: American Mathematical
Society).
Boichenko, V. A., Leonov, G. A., & Reitmann, V. (2005) Dimension Theory for Ordinary
Differential Equations (Wiesbaden: Teubner).
Bollobás, B. (1990) Linear Analysis (Cambridge: Cambridge University Press).
Bouligand, M. G. (1928) Ensembles impropres et nombre dimensionnel. Bull. Sci. Math.
52, 320–344 and 361–376.
Caffarelli, L., Kohn, R., & Nirenberg, L. (1982) Partial regularity of suitable weak
solutions of the Navier–Stokes equations. Comm. Pure Appl. Math. 35, 771–831.
Carvalho, A. N., Langa, J. A., & Robinson, J. C. (2010) Finite-dimensional global
attractors in Banach spaces. J. Diff. Eq. 249, 3099–3109.
Charalambous, M. G. (1999) A note on the relations between fractal and topological
dimensions. Q & A in General Topology. 17, 9–16.
Chepyzhov, V. V. & Ilyin, A. A. (2004) On the fractal dimension of invariant sets:
applications to Navier–Stokes equations. Discr. Cont. Dynam. Syst. 10, 117–135.
Chepyzhov, V. V., & Vishik, M. I. (2002) Attractors for Equations of Mathematical
Physics, AMS Colloquium Publications 49 (Providence, RI: American Mathemat-
ical Society).
Christensen, J. P. R. (1973) Measure theoretic zero sets in infinite dimensional spaces
and applications to differentiability of Lipschitz mappings. Publ. Dép. Math. (Lyon)
10, 29–39.
Chueshov, I. D. (2002) Introduction to the Theory of Infinite-dimensional Dissipative
Systems (Kharkiv, Ukraine: ACTA Scientific Publishing House).
Constantin, P. & Foias, C. (1985) Global Lyapunov exponents, Kaplan–Yorke formulas
and the dimension of the attractor for 2D Navier–Stokes equation. Commun. Pure.
Appl. Math. 38, 1–27.
196
References 197
Friz, P. K. & Robinson, J. C. (1999) Smooth attractors have zero “thickness”. J. Math.
Anal. Appl. 240, 37–46.
Friz, P. K. & Robinson, J. C. (2001) Parametrising the attractor of the two-dimensional
Navier–Stokes equations with a finite number of nodal values. Physica D 148,
201–220.
Friz, P. K., Kukavica, I., & Robinson, J. C. (2001) Nodal parametrisation of analytic
attractors. Disc. Cont. Dyn. Sys. 7, 643–657.
Gardner, R. J. (2002) The Brunn–Minkowski inequality. Bull. Amer. Math. Soc. 39,
355–405.
Gilbarg, D. & Trudinger, N. S. (1983) Elliptic Partial Differential Equations of Second
Order, Grundlehren der mathematischen Wissenschaften 224 (Berlin: Springer-
Verlag).
Gromov, M. (1999) Metric Structures for Riemannian and Non-Riemannian Spaces,
Progr. Math. vol. 152 (Boston, MA: Birkhäuser).
Guillopé, C. (1982) Comportement á l’infini des solutions des équations de Navier–
Stokes et propriété des ensembles fonctionnels invariants (ou attracteurs). C. R.
Acad. Sci. Paris Ser. I Math. 294, 221–224.
Hale, J. K. (1988) Asymptotic Behavior of Dissipative Systems, Mathematical Surveys
and Monographs 25 (Providence, RI: American Mathematical Society).
Hale, J. K., Magalhães, L. T., & Oliva, W. M. (2002) Dynamics in Infinite Dimen-
sions, second edition, Applied Mathematical Sciences 47 (New York, NY:
Springer).
Heinonen, J. (2001) Lectures on Analysis on Metric Spaces, Universitext (New York,
NY: Springer-Verlag).
Heinonen, J. (2003) Geometric Embeddings of Metric Spaces, Report University of
Jyväskylä Department of Mathematics and Statistics, 90 (Jyväskylä: University of
Jyväskylä).
Henry, D. (1981) Geometric Theory of Semilinear Parabolic Equations, Springer Lec-
ture Notes in Mathematics 840 (Berlin: Springer-Verlag).
Hensley, D. (1979) Slicing the cube in Rn and probability. Proc. Amer. Math. Soc. 73,
95–100.
Heywood, J. G. & Rannacher, R. (1982) Finite approximation of the nonstationary
Navier–Stokes problem. Part I: Regularity of solutions and second-order error
estimates for spatial discretization. SIAM J. Numer. Anal. 19, 275–311.
Howroyd, J. D. (1995) On dimension and on the existence of sets of finite positive
Hausdorff measure. Proc. London Math. Soc. 70, 581–604.
Howroyd, J. D. (1996) On Hausdorff and packing dimension of product spaces. Math.
Proc. Cambridge Philos. Soc. 119, 715–727.
Hunt, B. (1996) Maximal local Lyapunov dimension bounds the box dimension of
chaotic attractors. Nonlinearity 9, 845–852.
Hunt, B. R. & Kaloshin, V. Y. (1997) How projections affect the dimension spectrum
of fractal measures. Nonlinearity 10, 1031–1046.
Hunt, B. R. & Kaloshin, V. Y. (1999) Regularity of embeddings of infinite-dimensional
fractal sets into finite-dimensional spaces. Nonlinearity 12, 1263–1275.
Hunt, B. R., Sauer, T., & Yorke, J. A. (1992) Prevalence: a translation-invariant almost
every for infinite dimensional spaces. Bull. Amer. Math. Soc. 27, 217–238; (1993)
Prevalence: an addendum. Bull. Amer. Math. Soc. 28, 306–307.
References 199
Hurewicz, W & Wallman, H. (1941) Dimension Theory, Princeton Math. Ser. 4 (Prince-
ton, NJ: Princeton University Press).
John, F. (1948) Extremum problems with inequalities as subsidiary conditions, pages
187–204 in Studies and Essays Presented to R. Courant on his 60th Birthday,
January 8, 1948 (New York, NY: Interscience Publishers, Inc., New York).
John, F. (1982) Partial Differential Equations, fourth edition (New York, NY: Springer).
Johnson, W. & Lindenstrauss, J. (1984) Extensions of Lipschitz maps into a Hilbert
space. Contemp. Math. 26, 189–206.
Kadec, M. I. & Snobar, M. G. (1971) Certain functionals on the Minkowski compactum.
Math. Notes 10, 694–696.
Kakutani, S. (1939) Some characterizations of Euclidean space. Japanese J. Math. 16,
93–97.
Katêtov, M. (1952) On the dimension of non-separable spaces. I. Czech. Math. Journ.
2, 333–368.
Kraichnan, R. H. (1976) Inertial ranges in two-dimensional turbulence. Phys. Fluids 10,
1417–1423.
Kukavica, I. (1994) An absence of a certain class of periodic solutions in the Navier–
Stokes equations. J. Dynam. Differential Equations 6, 175–183.
Kukavica, I. (2007) Log–log convexity and backward uniqueness. Proc. Amer. Math.
Soc. 135, 2415–2421.
Kukavica, I. & Robinson, J. C. (2004) Distinguishing smooth functions by a finite
number of point values, and a version of the Takens embedding theorem. Physica
D 196, 45–66.
Laakso, T. J. (2002) Plane with A∞ -weighted metric not bi-Lipschitz embeddable to
RN . Bull. London Math. Soc. 34, 667–676.
Landau, L. D. & Lifshitz, E. M. (1959) Fluid Mechanics, Course of Theoretical Physics
6. (London: Pergamon, London).
Lang, U. & Plaut, C. (2001) Bilipschitz embeddings of metric spaces into space forms.
Geom. Dedicata 87, 285–307.
Langa, J. A. & Robinson, J. C. (2001) A finite number of point observations which
determine a non-autonomous fluid flow. Nonlinearity 14, 673–682.
Langa, J. A. & Robinson, J. C. (2006) Fractal dimension of a random invariant set. J.
Math. Pures Appl. 85, 269–294.
Larman, D. G. (1967) A new theory of dimension. Proc. London Math. Soc. 17, 178–192.
Liu, V. X. (1993) A sharp lower bound for the Hausdorff dimension of the global
attractors of the 2D Navier–Stokes equations. Comm. Math. Phys. 158, 327–
339.
Luukkainen, J. (1981) Approximating continuous maps of metric spaces into manifolds
by embeddings. Math. Scand. 49, 61–85.
Luukkainen, J. (1998) Assouad dimension: antifractal metrization, porous sets, and
homogeneous measures. J. Korean Math. Soc. 35, 23–76.
Mallet-Paret, J. (1976) Negatively invariant sets of compact maps and an extension of a
theorem of Cartwright. J. Differential Equations 22, 331–348.
Mandelbrot, B. B. (1982) The Fractal Geometry of Nature (W. H. Freeman & Co, San
Francisco, CA).
Mañé, R. (1981) On the dimension of the compact invariant sets of certain nonlinear
maps. In Rand, D. A., & Young, L. S. (eds) Dynamical Systems and Turbulence,
200 References
Warwick 1980, Springer Lecture Notes in Mathematics 898, pp. 230–242 (New
York, NY: Springer).
Mattila, P. (1975) Hausdorff dimension, orthogonal projections and intersections with
planes. Ann. Acad. Sci. Fennicae A 1, 227–244.
Mattila, P. (1995) Geometry of Sets and Measures in Euclidean Spaces, Cambridge
Studies in Advanced Mathematics (Cambridge: Cambridge University Press).
Meise, R. & Vogt, D. (1997) Introduction to Functional Analysis (Oxford: Oxford
University Press).
Menger, K. (1926) Über umfassendste n-dimensionale Mengen. Proc. Akad. Wetensch.
Amst. 29, 1125–128.
Menger, K. (1998) Ergebnisse eines Mathematischen Kolloquiums. Reissued and edited
by E. Dierker & K. Sigmund (Vienna: Springer).
Morita, K. (1954) Normal families an dimension theory for metric spaces. Math. Ann.
128, 350–362.
Movahedi-Lankarani, H. (1992) On the inverse of Mañé’s projection, Proc. Amer. Math.
Soc. 116, 555–560.
Munkres, J. R. (2000) Topology, second edition (Upper Saddle River, NJ: Prentice Hall).
Nöbeling, G. (1931) Über eine n-dimensionale Universalmenge im R 2n+1 . Math. Ann.
104, 71–80.
Olson, E. J. (2002) Bouligand dimension and almost Lipschitz embeddings. Pacific J.
Math. 2, 459–474.
Olson, E. J. & Robinson, J. C. (2010) Almost bi-Lipschitz embeddings and almost
homogeneous sets. Trans. Amer. Math. Soc. 362, 145–168.
Ott, W. & Yorke, J. A. (2005) Prevalence. Bull. Amer. Math. Soc. 42, 263–290.
Ott, W., Hunt, B. R., & Kaloshin, V. Y. (2006) The effect of projections on fractal sets
and measures in Banach spaces. Ergodic Theory Dynam. Systems 26, 869–891.
Pesin, Ya. B. (1997) Dimension Theory in Dynamical Systems: Contemporary Theory
and Applications, Chicago Lectures in Mathematics (Chicago, IL: University of
Chicago Press).
Pinto de Moura, E. & Robinson, J. C. (2010a) Orthogonal sequences and regularity of
embeddings into finite-dimensional spaces. J. Math. Anal. Appl. 368, 254–262.
Pinto de Moura, E. & Robinson, J. C. (2010b) Lipschitz deviation and embeddings of
global attractors. Nonlinearity 23, 1695–1708.
Pinto de Moura, E. & Robinson, J. C. (2010c) Log-Lipschitz continuity of the vector
field on the attractor of certain parabolic equations, arXiv preprint 1008.4949.
Pinto de Moura, E., Robinson, J. C., & Sánchez-Gabites, J. J. (2010) Embedding of
global attractors and their dynamics, Proc. Amer. Math. Soc., to appear.
Pontrjagin, L. & Schnirelmann, G. (1932) Sur une propriété métrique de la dimension.
Ann. Math. 33, 156–162.
Prosser, R. T. (1970) Note on metric dimension. Proc. Amer. Math. Soc. 25, 763–765.
Robinson, C. (1995) Dynamical Systems: Stability, Symbolic Dynamics, and Chaos
(London: CRC Press).
Robinson, J. C. (1997) Some closure results for inertial manifolds. J. Dyn. Diff. Eq. 9,
373–400.
Robinson, J. C. (2001) Infinite-dimensional Dynamical Systems, Cambridge Texts in
Applied Mathematics (Cambridge: Cambridge University Press).
References 201
202
Index 203
preserved under: countable unions, 49; finite thickness exponent, 46, 64, 65
unions, 49 alternative definition, 73
probability measure, 21 bounded by the upper box-counting
probe space, 48 dimension, 66
products of sets bounds the Lipschitz deviation, 68
Assouad dimension of, 86 in two spaces, 65
box-counting dimensions of, 35 of an orthogonal sequence, 67
covering dimension of, 9, 40 of set of real analytic functions, 163
Hausdorff dimension of, 25, 35, 40 related to bounds in Sobolev spaces, 137
projection, see also orthogonal projection zero for: ‘smooth’ sets, 67, 137; the
rank in ℓp , 78 Navier–Stokes attractor when f is
smooth, 138
quadrilateral inequality, 93, 100 zero implies zero dual thickness, 73
zero thickness implies zero dual thickness,
rank 70
of orthogonal projection, 62 turbulence, 169
of projection in ℓp , 78
real analytic function, 160, 161 Uniform Gronwall Lemma, 121
bounds on derivatives, 169 unit ball, volume, 22
has finite order of vanishing, 162 unit cube, hyperplane slices of, 52
zero set, 163 unstable set, 121
refinement of a covering, 8 contained in global attractor, 121
upper box-counting dimension, 32
semigroup, 105 alternative definitions, 39
generated by: a semilinear parabolic bounded by Assouad dimension, 85
equation, 109; the Navier–Stokes bounds: Hausdorff dimension, 34; the dual
equations, 110 thickness, 71; the thickness exponent, 66
semilinear parabolic equation, 108 can be calculated through a geometric
existence of global attractor, 118 sequence, 33
finite-dimensional attractor, 130 embedding into: Rk for subsets of RN , 43;
generates a semigroup, 109 Euclidean spaces, 75
minimal period of periodic orbits, 156 finite for attractors of semilinear parabolic
shyness, 47 equations, 130
for subsets of Rn , 48 impossibility of bi-Lipschitz embeddings in
preserved under finite unions, 48 general, 81
simplex, 15 in two spaces, 35
singular values of a matrix, 146, 158 modified, see modified upper box-counting
Sobolev spaces, 106 dimension
and continuous functions, 107, 113 monotonicity, 33
and fractional power spaces, 108 of an invariant set, 124
spherical Hausdorff measure, 30 of orthogonal sequences, 36
squeezing property, 143 of products, 35
Stokes operator, 111 of set of real analytic functions, 163
eigenvalues on [0, 2π ]2 , 143 of unions, 33
of unit cube in Rn , 34
Takens Time-Delay Embedding Theorem under Hölder continuous maps, 34
in RN , 149 variation of constants formula, 108, 113, 130,
in a Banach space, 154 156
obstruction from existence of periodic Vitali Covering Theorem, 23
orbits, 152
τ (X), see thickness exponent weak derivative, 106