1. Introduction
The stereographic projection [
1,
2] is an important tool in Algebraic Topology. Since
(with the Euclidean topology) is Hausdorff and locally compact, it is Hausdorff-compactifiable by one point, and its one-point compactification turns out to be
. The inverse problem to this is to remove one point from
(usually the north pole) to obtain again
. This is known as the stereographic projection, that is,
minus the north pole is homeomorphic to
.
Infinite-dimensional real or complex Banach spaces are Hausdorff but not locally compact, therefore it does not make sense to talk about their Hausdorff one-point compactification (there exists their one-point compactification but it will not be Hausdorff). However, the stereographic projection can still be constructed in infinite-dimensional real or complex Banach spaces, and it turns out that the unit sphere minus any one point is homeomorphic to a closed hyperplane. This is accomplished in [
3], where the stereographic projection is transported to infinite-dimensional real Banach spaces to prove that the unit sphere of any infinite-dimensional real Banach space minus any one point is homeomorphic to a closed hyperplane. Many geometric techniques from the Geometry of the real Banach spaces are employed to accomplish the results of [
3]. It is worth mentioning that neither the unit sphere nor the unit ball of an infinite dimensional real or complex Banach space is compact.
The purpose of this manuscript is to construct the stereographic projection in topological modules in order to find sufficient conditions for such stereographic projection to become a homeomorphism. This way, we will have nontrivial examples of bodies in a topological module such that, after removing one point from their boundary, this boundary minus the point becomes homeomorphic to a hyperplane (Theorem 4).
2. Methodology
The topological interior of a subset
A of a topological space will be denoted by
. The closure and the boundary of
A are respectively denoted by
and
. If
B is a subset of
A, then the interior of
B relative to
A is denoted as
, the closure of
B relative to
A is denoted as
, and the boundary of
B relative to
A is denoted as
. It is clear that
,
, and
. Recall that a body in a topological space is simply a closed subset with nonempty interior. We will be working with bodies in topological modules. Since translations are homeomorphisms in topological modules, which mostly preserve the geometrical structure, we will focus on bodies of topological modules which are neighborhoods of zero. Only nonzero left-modules over nonzero associative and unitary rings are considered throughout this manuscript. Refer to [
4,
5,
6,
7,
8,
9,
10] for a wide perspective on topological rings, modules, and algebras.
An affine manifold in a module M over a ring R is a translation of a submodule of M, that is, where and N is a submodule of M. Observe that an affine manifold contains 0 if and only if . Alongside this manuscript, we will be working with the following affine manifolds: for and for and in the range of . Notice that for all because for all , and it geometrically represents a straight line passing through . We will pay special attention to a notable subset of given by . Note that if and , then . Another trivial observation is the fact that for all in view of the fact that if for some , then . On the other hand, for all and it geometrically represents a hyperplane, which will be closed if R is Hausdorff.
A topological ring
R is said to be practical provided that
, that is, 0 belongs to the closure of the invertibles
of
R. Practical rings serve to extend the classical Operator Theory to the topological module setting. An extensive study on practical topological rings can be found in [
11]. Let
M be a topological module over a topological ring
R. Let
. We say that
A is bounded if for each 0-neighborhood
U in
M there is an invertible
such that
. On the other hand, a point
is said to be an internal point of
A provided for every
there exists a 0-neighborhood
satisfying that
. The set of internal points of
A is denoted by
. In this sense,
A is called absorbing provided that
.
An ordered ring is a ring endowed with a partial order that is compatible with the addition and the multiplication. In other words, a partial order ≤ in a ring R is a ring order provided that for all , and . The set of positive elements is denoted by . It is not hard to check that if are comparable, then and the ring has null characteristic, that is, , because for all . Whenever we talk about an ordered topological ring we mean a ring that is endowed with a ring topology and a ring order, but there does not have to be any relation between the ring order and the ring topology. However, if we talk about topological ordered rings, then we mean a ring endowed with a ring order in such a way that the order topology is well-defined and turns out to be a ring topology.
Finally, we will make use of the Axiom of Choice, which states that for every nonempty set X there exists a choice function , that is, for every .
3. Results
We will construct the stereographic projection in topological modules by relying on [
3]. One of the geometrical principles upon which this construction is based is the fact that, in Euclidean spaces, straight lines passing through zero must intersect the boundary of bounded bodies whose interior contains zero. This is itself based upon the following topological fact: if
X is a topological space and
L is a connected subset of
X such that
and
for a subset
, then
. Nevertheless, we will first study certain types of bodies in topological modules, fit bodies and Minkowski bodies, that will provide the necessary insight to properly construct the stereographic projection.
3.1. Fit Bodies
Let us begin by bearing in mind the trivial fact that if M is a module over a ring R, then for every .
Definition 1 (Fit body).
Let M be a topological module over a topological ring R. A subset is said to be a fit body if A is closed, , and every satisfies that .
Under the settings of the previous definition, it is immediate to realize that if a fit body A is additively symmetric, then for every . Before stating and proving the main result of this subsection, we will provide nontrivial examples of fit bodies.
Proposition 1. Let R be a topological totally ordered ring with no holes. Then is a fit body.
Proof. By assumption, the ring topology of R is the order topology, thus is open. Let us show next that . Indeed, fix an arbitrary open interval containing 1. Notice that , so we may assume that . By hypothesis, A has no holes, therefore is not empty, meaning that . As a consequence, . In a similar way, . On the other hand, is closed in the order topology, which assures that . Then A is an additively symmetric body whose interior contains 0. Finally, , meaning that A is a fit body. □
A nontrivial example of a ring satisfying the properties of Proposition 1 is provided next.
Example 1. , endowed with the inherited topology from , is a topological totally ordered ring with no holes.
We are now in the right position to state and prove the main theorem of this subsection.
Theorem 1. Let M be a torsionfree topological module over a Hausdorff topological-inversion integral domain ring R of such that is open. Let , a fit body, and such that but . Then Proof. Fix an arbitrary
. A 0-neighborhood
exists such that
. Another 0-neighborhood
can be found in such a way that
. Next, there are 0-neighborhoods
and
satisfying that
. Take
a 0-neighborhood in
R so that
. Let
be a 0-neighborhood in
R such that
. Since
and
s is invertible, we have that
, thus, since
R is Hausdorff, there exists a 0-neighborhood
such that
. Consider the continuous function (recall that
R is a topological-inversion ring):
where
is a 0-neighborhood in
M such that
. Note that (
1) maps 0 to 1. Thus, we can find
a 0-neighborhood verifying that
for all
. Finally, let
. We will show that
. Indeed, fix an arbitrary
with
. Denote
and
. Notice that
. Therefore,
. Additionally,
. As a consequence,
. Since
A is a fit body, either
or
. Next,
because
and
. Additionally,
M is torsionfree and
R is an integral domain, hence either
or
. In the latter case,
which is impossible by bearing in mind that
but
. As a consequence,
meaning that
, so
. □
3.2. Minkowski Bodies
Let
M be a topological module over a topological ring
R. Let
be an absorbing subset [
12,
13,
14]. Notice that if
, then there exists a 0-neighborhood
such that
. If
R is practical, then we can find an invertible
. This fact motivates the following definition, based also upon the classical Minkowski functional [
15] in real or complex topological vector spaces.
Definition 2 (Minkowski functional).
Let M be a topological module over a practical topological-inversion ring R. Let be an absorbing subset. A Minkowski functional is a functionsatisfying that and, for every , and . Under the settings of the previous definition, observe that if , then for every also defines a Minkowski functional.
Proposition 2. Let M be a topological module over a practical topological-inversion ring R. Let be an absorbing subset. There exists a Minkowski functional for A.
Proof. According to the Axiom of Choice, there exist choice functions and . They satisfy that for all and for all . Let us then construct the Minkowski functional for A by relying on the choice functions and . Indeed, fix an arbitrary . There exists a 0-neighborhood such that . Therefore, the set is not empty, hence . Observe that for all . On the other hand, since R is practical, for every and every , we have that , meaning that . Note that . Then it suffices to define for and . □
For the next definition, bear in mind that every body of a topological module which is also a neighborhood of zero is absorbing.
Definition 3 (Minkowski body).
Let M be a topological module over a practical topological-inversion ring R. A subset is said to be a Minkowski body if A is closed, , and there exists a continuous Minkowski functional satisfying that for every and for all .
The simplest example of a Minkowski body is the unit ball of a real or complex normed space. The following is a nontrivial example of a Minkowski body.
Example 2. Let be endowed with the inherited topology from . Notice that R is a division ring because is algebraic over . In fact, R is a non-discrete topological division ring, so in particular it is a practical topological-inversion ring. Take . Then A is a Minkowski body. Indeed, A is closed and , and is a continuous Minkowski functional satisfying that for every and for all .
It is time now to state and prove the main result of this subsection.
Theorem 2. Let M be a topological module over a practical topological-inversion ring R. Let , a Minkowski body, and such that but . Then Proof. Fix an arbitrary
. There exists a 0-neighborhood
such that
. Another 0-neighborhood
can be found in such a way that
. Next, there are 0-neighborhoods
and
satisfying that
. Take
a 0-neighborhood in
R so that
. Notice that the map
is continuous and maps
a to 1 (keep in mind that
because
and
). Let
be a 0-neighborhood in
M such that
. Finally, let
. We will show that
. Indeed, fix an arbitrary
with
. Observe that
because
, hence
is invertible. Denote
and
. Notice that
. Therefore,
. Additionally,
. As a consequence,
. Then we obtain that
. Since
s is not a right 0-divisor, we conclude that
meaning that
. □
3.3. Exposed Faces
This section is devoted to construct nontrivial examples of topological modules and bodies for which there exists a hyperplane intersecting the boundary of the body but not the interior. We will rely on ordered topological rings and on the notion of exposed face [
16], taken from the Geometry of (real) Banach Spaces.
Lemma 1. Let M be a topological module over a topological ring R. If is so that , then is an open map.
Proof. Let such that . We will show that is a 0-neighborhood in R for every 0-neighborhood . Indeed, there exists a 0-neighborhood such that . Then . Finally, is a 0-neighborhood in R because is invertible. As a consequence, is a 0-neighborhood in R. □
Definition 4 (Exposed face).
Let M be a topological module over an ordered topological ring R. Let and such that exists in R. The set is called an exposed face.
Under the settings of the previous definition, .
Theorem 3. Let M be a topological module over an ordered topological ring R. Let and such that exists in R. If is an open map, , and , then .
Proof. Suppose on the contrary that there exists . Since is open and , we can find such that . Since a is an internal point of A, there exists a 0-neighborhood such that . Take any . Notice that , hence reaching a contradiction with the fact that . □
Under the settings of Theorem 3, observe that because . Therefore, . As a corollary of Theorem 3, we obtain the following final result.
Corollary 1. Let R be a nondiscrete, integral domain, totally ordered topological ring R. Let M be a topological R-module. Let A be a subset of M. If , then
Proof. In the first place, since
R is an integral domain,
. Next, let us show that
. Indeed, let
V be any additively symmetric neighborhood of 0 in
R. Since
R is not discrete,
, so we can find
. Since
V is additively symmetric, by choosing
if necessary, we may assume that
. As a consequence,
. Nevertheless, with the hypotheses we have, we cannot guarantee that
will be an open map. Nonetheless, we can achieve the desired result. Indeed, fix an arbitrary
such that
. Since
, we can fix
such that
. By choosing
if necessary, we may assume that
. By hypothesis, there exists an additively symmetric neighborhood
of 0 such that
. We have just shown the existence of
. We have also seen that
. Next,
, thus
which is a contradiction. □
3.4. The Stereographic Projection
This final section is aimed at constructing the stereographic projection. For this, we will strongly rely on the affine manifolds given by the straight lines and the hyperplanes.
Definition 5 (Stereographic projection).
Let M be a topological module over a topological-inversion ring R. Let be an additively symmetric body with . Let and such that but . Let such that . The following map is known as a stereographic projection: Observe that the stereographic projection (
4) is well-defined and continuous (note that multiplicative inversion on
R is continuous). In fact, notice also that
for all
.
Remark 1. Let M be a topological module over a topological division ring R. Let be an additively symmetric body with . If and are such that but is a singleton, then the only element satisfies that . Indeed, if satisfies that , then , meaning the contradiction that . As a consequence, the stereographic projection (4) is well-defined and continuous. The point is to determine under what circumstances the stereographic projection (
4) is a homeomorphism. For this, the following definition, based upon the classical notion of strongly rotund point [
17,
18,
19,
20,
21,
22] from the Geometry of Banach Spaces, will be employed.
Definition 6 (Strongly rotund point).
Let M be a topological module over a topological-inversion ring R. Let be an additively symmetric body with . An element is said to be a strongly rotund point of A provided that there exist and , called the supporting functional and support value, respectively, satisfying the following conditions:
is relatively compact in R.
.
.
.
for all .
for all .
If is a net converging to , then is not convergent.
Notice that strongly rotund points are trivially preserved by isomorphisms of topological modules. The following remark gathers several considerations about strongly rotund points.
Remark 2. Let M be a topological module over a topological ring R. Let be an additively symmetric body with . Let be a strongly rotund point of A with supporting functional and support value . Observe the following:
because and . Additionally, if , then because otherwise we conclude that in view of the fact that . As a consequence, . For these reasons, the condition is well-imposed in Definition 6.
If , then or equivalently . Indeed, if , then , hence , so because s is invertible, contradicting that .
Recall that, given a real or complex Banach space
X with unit sphere
and unit ball
, a closed subspace
is said to be an
-summand subspace of
X if
M is
-complemented in
X, that is, there exists a closed subspace
such that
, in the sense that
for all
and all
. A point
is said to be an
-summand vector of
X provided that
is an
-summand subspace of
X, where
. By bearing in mind ([
3], Lemma 2.1), if
is an
-summand vector of a real Banach space
X, then
x is a strongly rotund point of the unit ball
of
X in the sense of Definition 6. A couple of technical lemmas will be needed before constructing the stereographic projection.
Lemma 2. Let be topological spaces. Let be bijective. Let . Suppose that for every net such that converges to , there exists a subnet of convergent to x. Then is continuous at .
Proof. Assume to the contrary that is not continuous at . Then there exists a net such that converges to but does not converge to x. There exists a neighborhood U of x and a subnet of such that for all . Note that converges to . Thus, by hypothesis, there exists a subnet of convergent to x, which is a contradiction in view of the fact that for all . □
Lemma 3. Let M be a topological module over a topological ring R. Let and be nets such that is convergent to some and is convergent to some . If is bounded, then converges to m.
Proof. It suffices to show that converges to 0. Fix an arbitrary 0-neighborhood . There are 0-neighborhoods and such that . By hypothesis, there exists an invertible such that for all . There exists such that for all . Then, for all , . □
At this stage, we are finally in the right position to construct the stereographic projection.
Theorem 4. Let M be a topological module over a topological-inversion ring R satisfying that . Let be an additively symmetric body with . If there exists a strongly rotund point with supporting functional and support value , then the stereographic projection (4) is a homeomorphism. Proof. First off, let us denote by
to the stereographic projection (
4). We already know that
is well-defined,
for all
, and
is trivially continuous due to the fact that
R is a topological-inversion ring. Let us check now that
is surjective. Fix an arbitrary
. If
, then
. Thus let us assume that
. Since
a is a strongly rotund point of
A, by definition we have that
. Let
such that
. We will show that
. Indeed,
Next step is to prove that is one-to-one. Indeed, take with . Then , meaning that . Let us finally prove that is continuous. We will rely on Lemma 2. Fix an arbitrary . Take a net such that converges to . We will show the existence of a subnet of convergent to b. Indeed, is relatively compact in R, therefore there exists a subnet of such that is convergent to some . Then converges to . This is equivalent to saying that converges to . Since is convergent to , we conclude that converges to , in other words, converges to , which is equivalent to stating that converges to . At this point, observe that since otherwise we obtain that converges to , reaching the contradiction that is not convergent by bearing in mind Definition 6. As a consequence, , meaning that . On the other hand, is closed, hence . Since a is a strongly rotund point of A, . As a consequence, either or . If , then , which is impossible. Thus, . By relying on Lemma 2, we conclude that is continuous at b. □
4. Discussion
We will discuss a nontrivial example of a stereographic projection in a topological module other than a real topological vector space. Observe that if M is a module over a ring R, is a subring of R, and , then is an S-submodule of M.
Theorem 5. Let M be a topological module over a topological division ring R. Let be an additively symmetric body with . Let and such that but . Let such that . Consider the stereographic projection (4). If is a dense subring of R such that and , then is an additively symmetric body of the S-module with and the following stereographic projection is well-defined and continuous: Proof. Note that
B is trivially closed in
since
A is closed in
M. In fact,
. As a consequence,
B is an additively symmetric body of the
S-module
with
. It only remains to show that
Indeed,
, hence
Finally, notice that, since S is dense in R and R is a topological division ring, then is dense in M which implies, together with the fact that A is closed in M, that and . □
Under the settings of Theorem 5, observe that if
S is a proper subring of
R, then
is a proper submodule of
M, hence the stereographic projection (
5) might not be necessarily surjective. Theorem 5 allows plenty of examples of nontrivial stereographic projections.
Example 3. Let X be a real Banach space of dimension strictly greater than 1. Let and such that . According to Remark 1, the stereographic projectionis well-defined and continuous. Here, , , , , and . Now, take and let N denote as S-module. Additionally, let . In view of Theorem 5, the stereograpahic projectionis well-defined and continuous.