Você está na página 1de 12

Metric space

In mathematics, a metric space is a set where a notion of distance (called a metric)


between elements of the set is defined. The metric space which most closely
corresponds to our intuitive understanding of space is the 3-dimensional Euclidean
space. The Euclidean metric of this space defines the distance between two points as
the length of the straight line connecting them. The geometry of the space depends on
the metric chosen, and by using a different metric we can construct interesting non-
Euclidean geometries such as those used in the theory of general relativity.

A metric space induces topological properties like open and closed sets which leads to
the study of even more abstract topological spaces.


Definition

A metric space is a tuple (M,d) where M is a set and d is a metric on M, that is, a
function

such that

1. d(x, y) 0 (non-negativity)
2. d(x, y) = 0 if and only if x = y (identity of indiscernibles)
3. d(x, y) = d(y, x) (symmetry)
4. d(x, z) d(x, y) + d(y, z) (triangle inequality).

The function d is also called distance function or simply distance. Often d is omitted and
one just writes M for a metric space if it is clear from the context what metric is used.
Removing one or more of these requirements leads to the concepts of a pseudometric
space, a quasimetric space, a hemimetric space, a semimetric space or most generally
a prametric space.

The first of these four conditions actually follows from the other three, since:

2d(x, y) = d(x, y) + d(y, x) d(x,x) = 0.


It is more correctly a property of a metric space, but one that many texts include in the
definition.

Some authors require the set M to be non-empty.

Metric spaces as topological spaces

The treatment of a metric space as a topological space is so consistent that it is almost a


part of the definition.

About any point x in a metric space M we define the open ball of radius r (>0) about x
as the set

B(x; r) = {y in M : d(x,y) < r}.

These open balls generate a topology on M, making it a topological space. Explicitly, a


subset of M is called open if it is a union of (finitely or infinitely many) open balls. The
complement of an open set is called closed. A topological space which can arise in this
way from a metric space is called a metrizable space; see the article on metrization
theorems for further details.

Since metric spaces are topological spaces, one has a notion of continuous function
between metric spaces. This definition is equivalent to the usual epsilon-delta definition
of continuity (which does not refer to the topology), and can also be directly defined
using limits of sequences.

Examples of metric spaces

The real numbers with the distance function d(x, y) = |y x| given by the absolute
value, and more generally Euclidean n-space with the Euclidean distance, are
complete metric spaces.
The rational numbers with the same distance function are also a metric space,
but not a complete one.
Hyperbolic space.
Any normed vector space is a metric space by defining d(x, y) = ||y x||, see also
relation of norms and metricshttp://en.wikipedia.org/wiki/Metric_%28mathematics
%29#Relation_of_norms_and_metrics. (If such a space is complete, we call it a
Banach space). Example:
o the Manhattan norm gives rise to the Manhattan distance, where the
distance between any two points, or vectors, is the sum of the distances
between corresponding coordinates.
o The maximum norm gives rise to the Chebyshev distance or chessboard
distance, the minimal number of moves a chess king would take to travel
from x to y.
The discrete metric, where d(x,y)=1 for all x not equal to y and d(x,y)=0
otherwise, is a simple but important example, and can be applied to all non-
empty sets. This, in particular, shows that for any non-empty set, there is always
a metric space associated to it.
The British Rail metric (also called the Post Office metric or the SNCF metric) on
a normed vector space, given by d(x, y) = ||x|| + ||y|| for distinct points x and y,
and d(x, x) = 0. More generally ||.|| can be replaced with a function f taking an
arbitrary set S to non-negative reals and taking the value 0 at most once: then
the metric is defined on S by d(x, y)=f(x)+f(y) for distinct points x and y, and d(x,
x) = 0. The name alludes to the tendency of railway journeys (or letters) to
proceed via London (or Paris) irrespective of their final destination.
If X is some set and M is a metric space, then the set of all bounded functions f :
X M (i.e. those functions whose image is a bounded subset of M) can be
turned into a metric space by defining d(f, g) = supx in X d(f(x), g(x)) for any
bounded functions f and g. If M is complete, then this space is complete as well.
The Levenshtein distance, also called character edit distance, is a measure of
the dissimilarity between two strings u and v. The distance is the minimal number
of character deletions, insertions, or substitutions required to transform u into v.
If X is a topological (or metric) space and M is a metric space, then the set of all
bounded continuous functions from X to M forms a metric space if we define the
metric as above: d(f, g) = supx in X d(f(x), g(x)) for any bounded continuous
functions f and g. If M is complete, then this space is complete as well.
If M is a connected Riemannian manifold, then we can turn M into a metric space
by defining the distance of two points as the infimum of the lengths of the paths
(continuously differentiable curves) connecting them.
If G is an undirected connected graph, then the set V of vertices of G can be
turned into a metric space by defining d(x, y) to be the length of the shortest path
connecting the vertices x and y.
Similarly (apart from mathematical details):
o For any system of roads and terrains the distance between two locations
can be defined as the length of the shortest route. To be a metric there
should not be one-way roads. Examples include some mentioned above:
the Manhattan norm, the British Rail metric, and the Chessboard
distance.
o More generally, for any system of roads and terrains, with given maximum
possible speed at any location, the "distance" between two locations can
be defined as the time the fastest route takes. To be a metric there should
not be one-way roads, and the maximum speed should not depend on
direction. The direction at A to B can be defined, not necessarily uniquely,
as the direction of the "shortest" route, i.e., in which the "distance"
reduces 1 second per second when travelling at the maximum speed.
Similarly, in 3D, the metrics on the surface of a polyhedron include the ordinary
metric, and the distance over the surface; a third metric on the edges of a
polyhedron is one where the "paths" are the edges. For example, the distance
between opposite vertices of a unit cube is 3, 5, and 3, respectively.
If M is a metric space, we can turn the set K(M) of all compact subsets of M into
a metric space by defining the Hausdorff distance d(X, Y) = inf{r : for every x in X
there exists a y in Y with d(x, y) < r and for every y in Y there exists an x in X
such that d(x, y) < r)}. In this metric, two elements are close to each other if every
element of one set is close to some element of the other set. One can show that
K(M) is complete if M is complete.
The set of all (isometry classes of) compact metric spaces form a metric space
with respect to Gromov-Hausdorff distance.
Given a metric space (X,d) and an increasing concave function f:[0,)[0,)
such that f(x)=0 if and only if x=0, then f o d is also a metric on X.
Given a injective function f from any set A to a metric space (X,d), d(f(x), f(y))
defines a metric on A.
Using T-theory, the tight span of a metric space is also a metric space. The tight
span is useful in several types of analysis.
The set of all n by m matrices over a finite field is a metric space with respect to
the rank distance d(X,Y) = rank(Y-X).

Notions of metric space equivalence

Comparing two metric spaces one can distinguish various degrees of equivalence. To
preserve at least the topological structure induced by the metric, these require at least
the existence of a continuous function between them (morphism preserving the topology
of the metric spaces).

Given two metric spaces (M1, d1) and (M2, d2):

They are called homeomorphic (topologically isomorphic) if there exists a


homeomorphism between them (i.e., a bijection continuous in both directions).

They are called uniformic (uniformly isomorphic) if there exists a uniform


isomorphism between them (i.e., a bijection uniformly continuous in both
directions)

They are called similar if there exists a positive constant k > 0 and a bijective
function f, called similarity such that f : M1 M2 and d2(f(x), f(y)) = k d1(x, y) for
all x, y in M1.

They are called isometric if there exists a bijective isometry between them. In
this case, the two spaces are essentially identical. An isometry is a function f :
M1 M2 which preserves distances: d2(f(x), f(y)) = d1(x, y) for all x, y in M1.
Isometries are necessarily injective.

They are called similar (of the second type) if there exists a bijective function f,
called similarity such that f : M1 M2 and d2(f(x), f(y)) = d2(f(u), f(v)) if and only if
d1(x, y) = d1(u, v) for all x, y,u, v in M1.

In case of Euclidean space with usual metric the two notions of similarity are equivalent.
Boundedness and compactness

A metric space M is called bounded if there exists some number r, such that d(x,y) r
for all x and y in M. The smallest possible such r is called the diameter of M. The space
M is called precompact or totally bounded if for every r > 0 there exist finitely many
open balls of radius r whose union covers M. Since the set of the centres of these balls
is finite, it has finite diameter, from which it follows (using the triangle inequality) that
every totally bounded space is bounded. The converse does not hold, since any infinite
set can be given the discrete metric (the first example above) under which it is bounded
and yet not totally bounded. A useful characterisation of compactness for metric spaces
is that a metric space is compact if and only if it is complete and totally bounded. Note
that compactness depends only on the topology, while boundedness depends on the
metric.

Note that in the context of intervals in the space of real numbers and occasionally
regions in a Euclidean space Rn a bounded set is referred to as "a finite interval" or
"finite region". However boundedness should not in general be confused with "finite",
which refers to the number of elements, not to how far the set extends; finiteness implies
boundedness, but not conversely.

By restricting the metric, any subset of a metric space is a metric space itself (a
subspace). We call such a subset complete, bounded, totally bounded or compact if it,
considered as a metric space, has the corresponding property.

Separation properties and extension of continuous functions

Metric spaces are paracompact Hausdorff spaces[1] and hence normal (indeed they are
perfectly normal). An important consequence is that every metric space admits partitions
of unity and that every continuous real-valued function defined on a closed subset of a
metric space can be extended to a continuous map on the whole space (Tietze
extension theorem). It is also true that every real-valued Lipschitz-continuous map
defined on a subset of a metric space can be extended to a Lipschitz-continuous map on
the whole space.
Distance between points and sets

A simple way to construct a function separating a point from a closed set (as required for
a completely regular space) is to consider the distance between the point and the set. If
(M,d) is a metric space, S is a subset of M and x is a point of M, we define the distance
from x to S as

d(x,S) = inf {d(x,s) : s S}

Then d(x, S) = 0 if and only if x belongs to the closure of S. Furthermore, we have the
following generalization of the triangle inequality:

d(x,S) d(x,y) + d(y,S)

which in particular shows that the map is continuous.

A metric space is a set with a global distance function (the metric ) that, for every
two points in , gives the distance between them as a nonnegative real number
. A metric space must also satisfy

1. if,

2. ,

3. The triangle inequality .

Here are some examples of compact spaces:

The unit interval [0,1] is compact. This follows from the Heine-Borel Theorem.
Proving that theorem is about as hard as proving directly that [0,1] is compact.

The half-open interval (0,1] is not compact: the open cover for

does not have a finite subcover.


Again from the Heine-Borel Theorem, we see that the closed unit ball of any
finite-dimensional normed vector space is compact. This is not true for infinite
dimensions; in fact, a normed vector space is finite-dimensional if and only if its
closed unit ball is compact.
Any finite topological space is compact.

Consider the set of all infinite sequences with entries in . We can turn

it into a metric space by defining , where is the

smallest index such that (if there is no such index, then the two

sequences are the same, and we define their distance to be zero). Then is a

compact space, a consequence of Tychonoff's theorem. In fact, is


homeomorphic to the Cantor set (which is compact by Heine-Borel). This
construction can be performed for any finite set, not just {0,1}.

Consider the set of all functions and defined a topology on

so that a sequence in converges towards if and only if

converges towards for all . (There is only one such


topology; it is called the topology of pointwise convergence). Then is a
compact topological space, again a consequence of Tychonoff's theorem.
Take any set , and define the cofinite topology on by declaring a subset of

to be open if and only if it is empty or its complement is finite. Then is a


compact topological space.
The prime spectrum of any commutative ring with the Zariski topology is a
compact space important in algebraic geometry. These prime spectra are almost
never Hausdorff spaces.
If is a Hilbert space and is a continuous linear operator, then

the spectrum of is a compact subset of . If is infinite-dimensional, then

any compact subset of arises in this manner from some continuous linear
operator on .
If is a complex C*-algebra which is commutative and contains a one, then the

set of all non-zero algebra homomorphisms carries a natural

topology (the weak-* topology) which turns it into a compact Hausdorff space.
is isomorphic to the C*-algebra of continuous complex-valued functions on
with the supremum norm.
Any profinite group is compact Hausdorff: finite discrete spaces are compact
Hausdorff, therefore their product is compact Hausdorff, and a profinite group is a
closed subset of such a product.
Any locally compact Hausdorff space can be turned into a compact space by
adding a single point to it (Alexandroff one-point compactification). The one-point

compactification of is homeomorphic to the circle ; the one-point

compactification of is homeomorphic to the sphere . Using the one-point


compactification, one can also easily construct compact spaces which are not
Hausdorff, by starting with a non-Hausdorff space.
Other non-Hausdorff compact spaces are given by the left order topology (or
right order topology) on bounded totally ordered sets.

A metric space is a set together with a real valued function

(called a metric, or sometimes a distance function) such that, for every ,

, with equality if and only if


For and with , the open ball around of radius is the set

. An open set in is a set which equals an


arbitrary (possibly empty) union of open balls in , and together with these open
sets forms a Hausdorff topological space. The topology on formed by these open
sets is called the metric topology, and in fact the open sets form a basis for this topology
(proof).

Similarly, the set is called a closed ball around

of radius . Every closed ball is a closed subset of in the metric topology.

The prototype example of a metric space is itself, with the metric defined by

. More generally, any normed vector space has an underlying


metric space structure; when the vector space is finite dimensional, the resulting metric
space is isomorphic to Euclidean space.

Let be a Hausdorff space, and be a compact subspace of . We prove that

is open, by finding for every point a neighborhood disjoint from


.

Let . , so by the definition of a Hausdorff space, there exist open

neighborhoods of and of such that . Clearly

but since is compact, we can select from these a finite subcover of


Now for every there exists such that . Since and

are disjoint, , therefore neither is it in the intersection

A finite intersection of open sets is open, hence is a neighborhood of disjoint from


.
A topological space, also called an abstract topological space, is a set together with a
collection of open subsets that satisfies the four conditions:

1. The empty set is in .

2. is in .

3. The intersection of a finite number of sets in is also in .

4. The union of an arbitrary number of sets in is also in .

Alternatively, may be defined to be the closed sets rather than the open sets, in which
case conditions 3 and 4 become:

3. The intersection of an arbitrary number of sets in is also in .

4. The union of a finite number of sets in is also in .

These axioms are designed so that the traditional definitions of open and closed
intervals of the real line continue to be true. For example, the restriction in (3) can be
seen to be necessary by considering , where an infinite
intersection of open intervals is a closed set.
In the chapter "Point Sets in General Spaces" Hausdorff (1914) defined his concept of a
topological space based on the four Hausdorff axioms (which in modern times are not
considered necessary in the definition of a topological space).

Você também pode gostar