Almost Injective Mappings of Totally Bounded Metric Spaces into Finite Dimensional Euclidean Spaces ()
1. Introduction
It is a classical question in topology, that what kind of topological spaces
can be embedded into finite dimensional Euclidean spaces (endowed with the usual Euclidean topology). To motivate the present paper, first we recall some well known results in related investigations.
A classical characterization theorem can be recalled as follows. The covering dimension of
is finite if and only if there exists
such that for every open covering C of
there exists an open covering
that refines C and each point of X belongs to at most
elements of
(the smallest n satisfying this property is defined to be the dimension of
; if there are no such n, then
is defined to be infinite dimensional). Clearly, if
can be embedded into a finite dimensional Euclidean space, then
is metrizable. By a classical theorem, for a compact, metrizable space
the followings are equivalent:
(1)
is finite dimensional (say, its dimension is n);
(2)
is homeomorphic to a subspace of a finite dimensional (more concretely,
-dimensional) Euclidean space
(for a proof we refer to e.g. Corollary 50.9 of [1] ). Recall, that for a positive real number
, a function
is defined to be an
-map if and only if for all
the diameter of
is at most
. Thus, if
is “small” then an
-map is “almost injective”. The usual proof for the harder direction of the above equivalence is based on showing that if the dimension of
is
, then for all
, the set of continuous
-mappings of
into
is dense and open; then the Baire category theorem implies the existence of a continuous embedding of
into
.
Further, as explained in Theorem 2.2 of [2] , for a compact, metrizable space
, the above (1) and (2) are also equivalent with
(3) for all positive
,
admits an
-map to an n-dimensional simplicial complex.
In fact, Theorem 2.2 of [2] provides several statements equivalent with the above (1), (2) and (3) and recalls further notions of dimensions of topological spaces. All of these dimensions are equivalent for compact metric spaces. For further details on related notions and results we refer to the rather comprehensive survey article [2] and to the references therein.
Recall e.g. from [3] , that the metric space
is defined to be totally bounded if and only if for all positive
there exists a finite collection of
-balls of
that covers X. Further,
is compact if and only if it is totally bounded and complete (that is, every Cauchy sequence of
is convergent).
The main result of the present paper is Theorem 10 where we show, that
(i) if
is totally bounded, then for all
, there exists a continuous
-mapping of
into a finite dimensional Euclidean space
, further
(ii) estimations for n are also provided in terms of
and structural properties of
.
We should make the following remarks: item (i) above is well known. Thus, “almost injective” functions still exist, even if
is infinite dimensional—of course, if
is infinite dimensional, then n depends on
. The known proofs are existential, they cannot provide any upper bound for n. However our proof for (i) is effective in the sense, that (as stated in (ii) above), based on it, one can establish estimations on n in terms of
and structural properties of
.
The structure of the paper is rather simple: we close this section by fixing our notation and Section 2 contains the proofs. As we mentioned, our main goal is to prove Theorem 10.
Notation
Our notation is mostly standard, but the following list may help.
Throughout
denotes the set of natural numbers. In addition,
and
denotes the set of real numbers, and the set of positive real numbers, respectively.
Let
be a metric space,
and let
be a non-negative real number. As usual, the open
-ball
at a is the set
2. Proofs
Definition 1 A family
of
-balls is defined to be a
-net if and only if it covers X, that is,
(1)
Thus,
is a completely bounded metric space if and only if for all positive
there exists a finite
-net in
. In addition, as it is well known,
is compact if and only if it’s metric is totally bounded and complete (i.e. every Cauchy sequence is convergent in
). For further details we refer to [3] , as well.
If
is a completely bounded metric space, then
denotes the smallest cardinality
for which there exists a
-sized
-net of
.
Definition 2 Let
be a metric space, let
and let
. Then the
-type of b over A in
is defined to be the function
, such that for all
we have
(2)
Thus, the
-type of b is just the function describing the distances of b from elements of A.
By a
-type over A we mean a function
which is of the form
for some
.
Keeping the notation introduced so far, we say, that
realizes the
-type p in
if and only if
.
Remark 3 We motivate and explain the above terminology as follows. Let
be a metric space. One can associate a relational structure to
in the following way. If d is a distance of
, that is,
then the binary relation
is defined to be
(3)
Thus, the relational structure
completely describes
and, at the same time, it can be treated as a model for an appropriate first order language. Then our definition of
-types is essentially the same, as atomic types of the first order relational structure
in the usual, model theoretic sense.
Lemma 1 Let
be a metric space, let
be finite and let
. Then for the function
,
we have
(4)
particularly, f is continuous.
Proof. Enumerate
and for each
let
,
. By a slight abuse of notation, we have
(5)
for all
.
Let
be fixed in this paragraph. Then for any
the triangle inequality yields
(6)
so we have
(7)
Similarly, interchanging x and y in the previous estimations, we obtain
(8)
and hence
(9)
Consequently, for all
we have
(10)
as desired.
The following notion is an approximate version of splitting introduced in [4] .
Definition 4 Let
be a metric space, let
, let p be a
-type over A in
and let
be non-negative real numbers. Then we say, that p is
-splitting over B if and only if there exist
such that for all
we have
(11)
but whenever a realizes p, we have
(12)
Keeping the notation introduced in the above definition, intuitively
is
-splitting over B if and only if there exist
such that
and
are “indiscernible from the viewpoint of B modulo
”, but a “distinguishes them modulo
”.
The following theorems will be essential in this paper. Some variants of them (in different contexts) had been utilized e.g. in [5] , [6] and in [7] .
Theorem 5 Let
be a totally bounded metric space, let
,
let
and let
be arbitrary. Suppose
(13)
is a strictly increasing sequence of subsets of
such that for all
the type
is
-splitting over
. Then
.
Proof. Let
be a
-net of
with smallest possible cardinality. By our assumption on splitting, for each
there exist
such that for all
we have
(14)
but
(15)
Assume, seeking a contradiction, that
. By the pigeonhole principle, there exists
such that
(the
-ball in our net) contains at least two
's; more precisely, there exist
with
. Since
is a
-ball, it follows, that
. Therefore, by the triangle inequality,
(16)
and by symmetry,
(17)
It follows, that
(18)
By construction,
(19)
particularly,
(20)
Therefore,
(21)
and similarly,
(22)
It follows, that
(23)
Combining these, we get
(24)
Since
, this contradicts to (15) above, and the proof is complete.
Definition 6 Let
be a metric space, let
and let
. Then
is defined to be an
-basis for a if and only if for any
with
, the type
does not
-split over A.
Remark 7
1) Clearly, if A is an
-basis for
and
, then A is a
-basis for a, as well.
2) In addition, if A is an
-basis for
and
then B is an
-basis for
, as well.
Theorem 8 Let
be a totally bounded metric space, let
and let
. Then there exist
and
such that
is an
-basis for a.
In fact, arbitrary
is suitable and
can be chosen so, that
is satisfied, as well.
Proof. Let
be an arbitrary positive real number.
Suppose, seeking a contradiction, that the consequence of the theorem is not true. By recursion, we define finite subsets
for every natural number n, such that the following stipulations are satisfied:
(i)
, in fact,
;
(ii)
is
-splitting over
.
Let
and suppose
has already been defined for all
such that stipulations (i) and (ii) are satisfied. Then, by our indirect assumption, there exists
with
such that
is
-splitting over
. This means, that there exist
such that for all
we have
(25)
but
(26)
Let
. Then stipulations (i), (ii) remain true. In this way, one can define
for all
; this contradicts to Theorem 5. Thus, the proof is complete:
can be chosen to be some
(note, that an inspection shows, that each
has cardinality at most 2n and
).
Lemma 2 Let
be a totally bounded metric space, let
and let
. If A is an
-basis for
and
, moreover
and
are pairwise distinct and
for all
, then A is an
-basis for b.
Proof. Assume, seeking a contradiction, that A is an
-basis for
and
, but not an
-basis for b. Then there exist
such that for all
we have
(27)
but
(28)
By assumption,
and
are pairwise distinct. It follows, that there exists
such that
, that is,
.
Observe, that
(29)
and hence
(30)
Similarly,
(31)
and hence
(32)
Combining these estimations, we get
(33)
Completely similarly, one also can conclude, that
(34)
Observe moreover, that
because A is an
-basis for
. But then,
(35)
contradicting to (28).
Theorem 9 Let
be a totally bounded metric space. Then for each
there exists a finite set
such that A is an
-basis for all
.
In fact, we have
.
Proof. Let
be an
-net of
with
. Let
(36)
and let
. Enumerate
as
For each
there exist pairwise different
. By Theorem 8, for each
and
there exists an
-basis
for
with
. Similarly, again by Theorem 8, for each
and for each
there exists an
-basis
for a with
. Finally, let
(37)
We claim, that A satisfies the conclusion of the theorem. By construction, clearly
(38)
as desired. In addition, let
be arbitrary. Then there exists
such that
.
If
then by construction, A contains an
-basis
for b, which is also an
-basis for b because of Remark 7 (1). Hence, by Remark 7 (2), A is an
-basis for b.
If
, then
for some
. By construction, for all
we
have
. Therefore, by the triangle inequality, for all
we have
(39)
Hence, by Lemma 2,
is an
-basis for b. Since
, it follows from Remark 7 (2), that A is an
-basis for
b, as well. Since b was arbitrary, the proof is complete.
Lemma 3 Let
be a metric space and let
be arbitrary.
Assume
is an
-basis for all
. If
is not isolated,
then for all
implies
.
Proof. Since x is not an isolated point of
, there exists
. By assumption, A is an
-basis for z, hence
(40)
But
, so it follows, that
. Hence, by the triangle inequality,
(41)
as desired.
Let
be a metric space, let
and let Y be any set. As we mentioned in the introduction, according to the terminology of e.g. [2] , a function
is defined to be an
-map if and only if for all
, the diameter of
is at most
, or equivalently,
Now we are able to state and prove the main result of the paper: we give a new proof for the fact, that each totally bounded metric space admits a continuous
-map into some finite dimensional Euclidean space
(endowed with the usual Euclidean metric). Further, we provide upper bounds for n.
Theorem 10 Let
be a totally bounded metric space, let
and let
.
1) if
does not contain isolated points, then there exist
and an
-map
such that, for all
we have
, particularly f is continuous.
2) if
has countably many isolated points, then there exist
and a continuous
-map
.
3) if
is compact, then there exist
and a continuous
-map
.
Proof. First we show (1). By Theorem 9, there exists
such that A is an
-basis for all
and
. Let
and let
,
(42)
for all
. Then, by Lemma 1, for all
we have
(43)
Further, by assumption,
does not contain isolated points. Hence, by Lemma 3 f is an
-map, as desired.
To show (2), enumerate all the isolated points of X as
. As in (1),
by Theorem 9 there exists
such that A is an
-basis for all
and
. Let
and let
,
(44)
for all
. Then, by Lemma 1, for all
we have
(45)
Let
, let
be the function
(46)
and let
,
First we show, that f is continuous. To do so, let
and let
be arbitrary. We shall show, that there exists
such that for all
we have
.
Case 1: x is isolated in
. Then there exists
such that
, hence, for all
we have
.
Case 2: x is not isolated in
. Then, by 45, there exists
such that for all
we have
. In addition, there exists
with
. Choose
such that if
then
and finally let
. Now let
.
If y is not isolated, then
(47)
If y is isolated, then
for some
and hence
(48)
thus, f is continuous, as desired.
Next we show, that f is an
-map. To do so, assume
are such, that
. Observe, that by construction, for any
, the
isolated point
is the unique element
such that the last coordinate of
is equal to
. Hence neither x nor y are isolated. Therefore
(49)
hence
. Combining this with the definition of
, we obtain
. But then, Lemma 3 implies
.
To show (3), we note, that (e.g. by Corollary 4.1.16 of [3] ) a compact metric space is second countable, hence it may contain countably many isolated points, only. Thus, (2) implies (3).
3. Concluding Remarks
In Theorem 10 we have given a new proof for the fact, that each totally bounded metric space
admits a continuous
-map into some finite dimensional Euclidean space
(endowed with the usual Euclidean metric). Further, we provided upper bounds for n in terms of
and structural properties of
. Our proof had been obtained as follows:
• As recalled in Remark 3, there is a well known method that associates a first order structure
to a metric space
;
• if
is totally bounded, then
has nice properties inspired by (model theoretic) stability theory, more concretely, as shown in Theorem 5, if
, then each increasing chain of
-splitting
-types in
should have finite length;
• as shown in Theorem 10, using the above observation one can construct an appropriate continuous
-map from
to a finite dimensional Euclidean space.
The second item of the above list motivates the following problem which may serve as a starting point of a further investigation.
Open problem 11 Is the converse of Theorem 5 true, as well? That is, if
is a metric space such that, for all
the length of any increasing chain of
-splitting
-types in
is finite, then does it follow, that
is
totally bounded?
Acknowledgements
This work has been supported by Hungarian National Foundation for Scientific Research grant K129211.