Sie sind auf Seite 1von 12

http://www.amaderforum.

com

CHAPTER 8. TOPOLOGY

Topology
Basic Concepts in Topology

Product Topology

Connectedness in Topology

Compact Space

Urysohn's lemma

http://www.amaderforum.com
http://www.amaderforum.com

What is topology?
Topologists are mathematicians who study qualitative questions about geometrical
structures. We do not ask: how big is it? but rather: does it have any holes in it? is it all
connected together, or can it be separated into parts? A commonly cited example is the
London Underground map.

This will not reliably tell you how far it is from Kings Cross to Picadilly, or even the
compass direction from one to the other; but it will tell you how the lines connect up
between them. In other words, it gives topological rather than geometric information.
Again, consider a doughnut and a teacup, both made of BluTack. We can take one of
these and transform it into the other by stretching and squeezing, without tearing the
BluTack or sticking together bits which were previously separate. It follows that there is
no topological difference between the two objects. Consider the problem of building a
fusion reactor which confines a plasma by a magnetic field.

http://www.amaderforum.com
http://www.amaderforum.com

Imagine a closed surface surrounding the plasma. At each point on the surface, the
component of the magnetic field parallel to the surface must be nonzero, or the plasma
will leak out. We are thus led to the following question: given a surface S, is it possible
for there to be a field of vectors tangent at each point of S which is nowhere zero? It turns
out that this depends solely on the topological nature of S. If S is the surface of a sphere,
it is not possible. Magnetic confinement of a ball of plasma just does not work. The only
type of surface for which this approach is possible is the inner tube shape, which is of
course the solution universally used for such reactors. (I do not claim that engineers
needed topologists to point this out; on the contrary, this is a nice example precisely
because many people can see for themselves that the claim is true.) Another amusing
consequence of the same argument is that at any given time, some point on the Earth's
surface is windless.

While these examples are interesting, they are somewhat limited. Topologically speaking
there are few possibilities for solid objects; essentially the only variable is the number of
holes. Interesting things happen, however, when we work with abstract objects in more
than three dimensions. This is not the dry academic game that it might appear at first
sight. Consider the ubiquitous use of graphs in science. One often sees graphs of (say)
pressure against temperature. Neither variable is naturally a geometric quantity.

http://www.amaderforum.com
http://www.amaderforum.com

Nonetheless, they can be represented by distances on a sheet of paper. Moreover,


geometric concepts such as slopes of lines and areas of regions can usefully be
interpreted in terms of thermodynamics. In principle, none of these geometric ideas are
neccessary; all the arguments can be carried through algebraically without any diagrams.
Nonetheless, most of us find diagrams helpful and illuminating, because we have a better
intuitive feel for geometry than for algebra. Now suppose that instead of two variables
(pressure and temperature) we are faced with a problem involving five or six. We might
like to draw a six dimensional graph, but we do not have enough space. We can try to
imagine what six dimensional geometry might be like, but it is difficult to have
confidence in our intuition about such a theoretical construct so far removed from
experience. On the other hand, we can try to do the problem by pure algebra, but in doing
so we forgo the visual insight which is one of the most powerful faculties of the human
mind. In fact, mathematicians have developed a rather strange yet highly fruitful
approach to such problems. The questions addressed are officially formulated in terms of
logic and algebra rather than geometry, and solutions are required to be expressed in such
terms. Where geometrical concepts are used, they must be defined in terms of algebraic
ones. All use of such concepts is to be justified purely in terms of such definitions, rather
than in terms of geometric intuition. The visual imagination guides the course of the
argument, and suggests what one should try to prove, but the argument must stand in its
own right. There is an interesting, if limited, analogy with physical science, with rigorous
proof playing the role of experiment and geometrical speculation (amongst other things)
that of the more mysterious processes by which new theories emerge. Let us consider an
example with more than three dimensions. Imagine an industrial robot.

The position and orientation of the hand is determined by deflection of the various joints
of the robot arm. Suppose that there are n joints, each of which bends in only one plane
(this assumption is not essential). We then have to give n numbers in order to specify the
state of the joints. In the case n=2, we could draw a diagram on a sheet of paper with one
axis for each joint, so that a point on the diagram would correspond to a possible state of
the joints. If n>3, then such a diagram would have to be n-dimensional, so we could not

http://www.amaderforum.com
http://www.amaderforum.com

draw it. Nevertheless, we can still reason about a theoretical n-dimensional space (call it
X) in which a single point represents a state of the arm. Now consider the position and
orientation of the hand. In the usual way, we need three coordinates to specify the
position. We next need to specify the direction in which the central axis of the hand
points. This needs two numbers: the angle of elevation above the horizontal and the
horizontal compass bearing. Note that these two numbers are not entirely independent or
well defined. If the elevation is vertical then the compass bearing is meaningless. Also,
the physical meaning of the compass bearing is unchanged if we add 360 degrees. We
can suppress this by insisting that all angles must be taken between 0 and 360, but then
we have a different problem in that the numerical value of the angle could change
suddenly (as the arm moves from 1 degree to 359 degrees) without any genuine
discontinuity in the behaviour of the robot. There is a topological reason for these
problems. If the space of possible orientations of the hand could be simply parametrised
by two variables, then it would be topologically a plane. However, it is not a plane. It can
instead be thought of as the surface of a sphere, which one imagines centered at a point
on the axis of the hand. If the hand held a torch, it would mark a point on the sphere
which would determine the orientation. There is one further variable to consider, viz. the
twist of the hand about its central axis. This gives a third angle, which again interacts in a
complicated way with the other two. All told, we need six parameters to describe the
position and orientation of the hand. We can thus consider a six dimensional space Y,
each point in which corresponds to a possible state of the hand. We can analyse the
structure of Y using the methods of algebraic topology, and learn a number of interesting
and nontrivial things about it. Each point in our space X corresponds to a state of the
joints of the robot. The state of the joints determines the position and orientation of the
hand. Thus, to each point x in the space X we associate a point of Y, which we shall call
f(x). However, if we want to achieve a specified position and orientation of the hand,
there will in general be many ways to do this. In other words, many different points of X
will be associated to the same point of Y. It would be convenient in the design and use of
such robots to have a so-called ``inverse mapping''. For each position and orientation of
the hand (corresponding to a point y in Y) one would like to choose a state of the joints
(corresponding to a point g(y) of X) which puts the hand in that situation (so that
f(g(y))=y). Moreover, one would like g(y) to vary continuously with y, so that similar
positions of the hand are achieved by similar states of the joints. It turns out, however,
that this is topologically impossible. It is thus necessary to devise a more complex control
strategy, in which the same hand position may be achieved in different ways at different
times. This is typical of the kind of information which can be supplied by topological
theory.

http://www.amaderforum.com
http://www.amaderforum.com

Product Topology
The topology on the Cartesian product of two topological spaces whose open sets
are the unions of subsets , where and are open subsets of and , respectively.

This definition extends in a natural way to the Cartesian product of any finite number of
topological spaces. The product topology of

where is the real line with the Euclidean topology, coincides with the Euclidean
topology of the Euclidean space .

In the definition of product topology of , where is any set, the open sets are
the unions of subsets , where is an open subset of with the additional condition
that for all but finitely many indices (this is automatically fulfilled if is a finite
set). The reason for this choice of open sets is that these are the least needed to make the
projection onto the th factor continuous for all indices . Admitting all products
of open sets would give rise to a larger topology (strictly larger if is infinite), called the
box topology.

The product topology is also called Tychonoff topology, but this should not cause any
confusion with the notion of Tychonoff space, which has a completely different meaning.

http://www.amaderforum.com
http://www.amaderforum.com

Connectedness in topology
A topological space is said to be connected if it cannot be contained in two disjoint
nonempty open sets. A set is open if it contains no point lying on its boundary; thus, in an
informal, intuitive sense, the fact that a space can be partitioned into disjoint open sets
suggests that the boundary between the two sets is not part of the space, and thus splits it
into two separate pieces.

Other notions of connectedness

Fields of mathematics are typically concerned with special kinds of objects. Often such
an object is said to be connected if, when it is considered as a topological space, it is a
connected space. Thus, manifolds, Lie groups, and graphs are all called connected if they
are connected as topological spaces, and their components are the topological
components. Sometimes it is convenient to restate the definition of connectedness in such
fields. For example, a graph is said to be connected if each pair of vertices in the graph is
joined by a path. This definition is equivalent to the topological one, as applied to graphs,
but it is easier to deal with in the context of theory. Graph theory also offers a context-
free measure of connectedness, called the clustering coefficient.

Other fields of mathematics are concerned with objects that are rarely considered as
topological spaces. Nonetheless, definitions of connectedness often reflect the topological
meaning in some way. For example, in category theory, a category is said to be
connected if each pair of objects in it is joined by a morphism. Thus, a category is
connected if it is, intuitively, all one piece.

There may be different notions of connectedness that are intuitively similar, but different
as formally defined concepts. We might wish to call a topological space connected if
each pair of points in it is joined by a path. However this concept turns out to be different
from standard topological connectedness; in particular, there are connected topological
spaces for which this property does not hold. Because of this, different terminology is
used; spaces with this property are said to be path connected.

Terms involving connected are also used for properties that are related to, but clearly
different from, connectedness. For example, a path-connected topological space is simply
connected if each loop (path from a point to itself) in it is contractible; that is, intuitively,
if there is essentially only one way to get from any point to any other point. Thus, a
sphere and a disk are each simply connected, while a torus is not. As another example, a
directed graph is strongly connected if each rdered pair of vertices is joined by a directed
path (that is, one that "follows the arrows").

Other concepts express the way in which an object is not connected. For example, a
topological space is totally disconnected if each of its components is a single point.

http://www.amaderforum.com
http://www.amaderforum.com

Connectivity

Properties and parameters based on the idea of connectedness often involve the word
connectivity. For example, in graph theory, a connected graph is one from which we must
remove at least one vertex to create a disconnected graph. In recognition of this, such
graphs are also said to be 1-connected. Similarly, a graph is 2-connected if we must
remove at least two vertices from it, to create a disconnected graph. A 3-connected graph
requires the removal of at least three vertices, and so on. The connectivity of a graph is
the minimum number of vertices that must be removed, to disconnect it. Equivalently, the
connectivity of a graph is the greatest integer k for which the graph is k-connected.

While terminology varies, noun forms of connectedness-related properties often include


the term connectivity. Thus, when discussing simply connected topological spaces, it is
far more common to speak of simple connectivity than simple connectedness. On the
other hand, in fields without a formally defined notion of connectivity, the word may be
used as a synonym for connectedness.

Another example of connectivity can be found in regular tilings. Here, the connectivity
describes the number of neighbors accessible from a single tile:

3-connectivity in a triangular
4-connectivity in a square tiling,
tiling,

6-connectivity in a hexagonal 8-connectivity in a square tiling (note that distance


tiling, equity is not kept)

http://www.amaderforum.com
http://www.amaderforum.com

Compact space
From Wikipedia, the free encyclopedia
Jump to: navigation, search
"Compact" redirects here. For other uses, see Compact (disambiguation).

In mathematics, a subset of Euclidean space Rn is called compact if it is closed and


bounded. For example, in R, the closed unit interval [0, 1] is compact, but the set of
integers Z is not (it is not bounded) and neither is the half-open interval [0, 1) (it is not
closed).

A more modern approach is to call a topological space compact if each of its open covers
has a finite subcover. The HeineBorel theorem shows that this definition is equivalent to
"closed and bounded" for subsets of Euclidean space. Note: Some authors such as
Bourbaki use the term "quasi-compact" instead, and reserve the term "compact" for
topological spaces that are Hausdorff and "quasi-compact".

A single compact set is sometimes referred to as a compactum; following the Latin


second declension (neuter), the corresponding plural form is compacta.

Definitions

Compactness of subsets of Rn

For any subset of Euclidean space Rn, the following four conditions are equivalent:

Every open cover has a finite subcover. This is the most commonly used
definition.
Every sequence in the set has a convergent subsequence, the limit point of which
belongs to the set.
Every infinite subset of the set has an accumulation point in the set.
The set is closed and bounded. This is the condition that is easiest to verify, for
example a closed interval or closed n-ball.

In other spaces, these conditions may or may not be equivalent, depending on the
properties of the space.

Note that while compactness is a property of the set itself (with its topology), closedness
is relative to a space it is in; above "closed" is used in the sense of closed in Rn. A set
which is closed in e.g. Qn is typically not closed in Rn, hence not compact.

http://www.amaderforum.com
http://www.amaderforum.com

Compactness of topological spaces

The "finite subcover" property from the previous paragraph is more abstract than the
"closed and bounded" one, but it has the distinct advantage that it can be given using the
subspace topology on a subset of Rn, eliminating the need of using a metric or an ambient
space. Thus, compactness is a topological property. In a sense, the closed unit interval
[0,1] is intrinsically compact, regardless of how it is embedded in R or Rn.

Other forms of compactness

There are a number of topological properties which are equivalent to compactness in


metric spaces, but are inequivalent in general topological spaces. These include the
following.

Sequentially compact: Every sequence has a convergent subsequence.


Countably compact: Every countable open cover has a finite subcover. (Or,
equivalently, every infinite subset has an -accumulation point.)
Pseudocompact : Every real-valued continuous function on the space is bounded.
Weakly countably compact (or limit point compact): Every infinite subset has an
accumulation point.

While all these conditions are equivalent for metric spaces, in general we have the
following implications:

Compact spaces are countably compact.


Sequentially compact spaces are countably compact.
Countably compact spaces are pseudocompact and weakly countably compact.

Not every countably compact space is compact; an example is given by the first
uncountable ordinal with the order topology. Not every compact space is sequentially
compact; an example is given by 2[0,1], with the product topology.

A metric space is called pre-compact or totally bounded if any sequence has a Cauchy
subsequence; this can be generalised to uniform spaces. For complete metric spaces this
is equivalent to compactness. See relatively compact for the topological version.

Another related notion which (by most definitions) is strictly weaker than compactness is
compactness.

http://www.amaderforum.com
http://www.amaderforum.com

Urysohn's lemma
From Wikipedia, the free encyclopedia
Jump to: navigation, search

In topology, Urysohn's lemma, sometimes called "the first non-trivial fact of point set
topology", is commonly used to construct continuous functions of various properties on
normal spaces. It is widely applicable since all metric spaces and all compact Hausdorff
spaces are normal. The lemma is generalized by (and usually used in the proof of) the
Tietze extension theorem.

The lemma is named after Pavel Samuilovich Urysohn. See also Cutoff function.

Formal statement

Urysohn's lemma states that X is a normal topological space if and only if, whenever A
and B are disjoint closed subsets of X, then there exists a continuous function from X into
the unit interval [0, 1],

f : X [0, 1],

such that f(a) = 0 for all a in A and f(b) = 1 for all b in B.

Any such function f is known as an Urysohn function.

Note that in the statement above, we do not, and in general cannot, require that f(x) 0
and 1 for x outside of A and B. This is only possible in perfectly normal spaces.

A corollary of the lemma is that normal T1 spaces are Tychonoff.

Sketch of proof

Illustration of Urysohn's "onion" function.

http://www.amaderforum.com
http://www.amaderforum.com

For every dyadic fraction r (0,1), we are going to construct an open subset U(r) of X
such that:

1. U(r) contains A and is disjoint from B for all r


2. for r < s, the closure of U(r) is contained in U(s)

Once we have these sets, we define f(x) = inf { r : x U(r) } for every x X. Using the
fact that the dyadic rationals are dense, it is then not too hard to show that f is continuous
and has the property f(A) {0} and f(B) {1}.

In order to construct the sets U(r), we actually do a little bit more: we construct sets U(r)
and V(r) such that

A U(r) and B V(r) for all r


U(r) and V(r) are open and disjoint for all r
for r < s, V(s) is contained in the complement of U(r) and the complement of V(r)
is contained in U(s)

Since the complement of V(r) is closed and contains U(r), the latter condition then
implies condition (2) from above.

This construction proceeds by mathematical induction. Since X is normal, we can find


two disjoint open sets U(1/2) and V(1/2) which contain A and B, respectively. Now
assume that n1 and the sets U(a/2n) and V(a/2n) have already been constructed for a =
1,...,2n-1. Since X is normal, we can find two disjoint open sets which contain the
complement of V(a/2n) and the complement of U((a+1)/2n), respectively. Call these two
open sets U((2a+1)/2n+1) and V((2a+1)/2n+1), and verify the above three conditions.

http://www.amaderforum.com

Das könnte Ihnen auch gefallen