1. Introduction
Suppose we want to measure the perimeter of a plane curve and we only have at our disposal a ruler, but not a wire that would perfectly match the boundary [
1]. How are we going to evaluate the perimeter? One can indeed look for polygons whose sides are measurable with the help of the ruler, and one can “approximate” the shape of the curve by means of such polygons, asking ourselves the question by how much such polygons are close to the figure whose perimeter we want to evaluate. Within this framework, the very fruitful idea of Caccioppoli [
2,
3] was to “measure” the distance between the polygon and the figure by means of the area of the difference set. In order to gain an idea of the perimeter, we have to take a sequence of polygons, such that the area of the difference set becomes smaller, and they will provide a sequence of approximate values of the perimeter. However, the area-type approximation of the figure can be obtained also by means of polygons having a perimeter unnecessarily large (for example, if we curl a polygon, we can achieve a very large perimeter). Thus, among all sequences of polygons which approximate a geometric figure, we must take, among the limits of their perimeters, the smallest value, i.e., the minimal limit of perimeters of the approximating polygons.
The development of geometric measure theory [
2,
3,
4,
5,
6,
7,
8,
9,
10,
11,
12] therefore led to the discovery of many important concepts including, in particular, finite-perimeter sets with their reduced boundary (see definitions in
Appendix A), two concepts that play an important role in modern mathematics. With hindsight, one can say that the divergence theorem does not hold on the topological boundary of a finite-perimeter set, but only on the reduced boundary, which is therefore the truly important concept of the boundary. Let us here review some key aspects of this framework, which provide a motivation of our research.
If
is a finite-perimeter set, the De Giorgi structure theorem ensures that its reduced boundary
can be written in the form
where the
are disjoint compact sets, and
is a set having vanishing
-dimensional Hausdorff measure. The sets
are contained in
-dimensional manifolds
of class
. If the point
x belongs to
, one finds that the generalized normal to
E, denoted by
, can be obtained by the equation
This means that the normal at x is given by the normal existing at x on the manifold (for this purpose, it is of crucial importance that the manifold should be of class ). Thus, the generalized normal is evaluated from the usual normal to a countable infinity of compact portions of smooth manifolds .
However, despite this profound result, the actual evaluation of the generalized normal may turn out to be impossible in some cases. In order to understand this feature, one can consider a dense sequence
of points of
. If the points
are vectors with
n components which have rational coordinates
then within each ball
centred at the point
and having radius
one can find infinitely many points, and the same is true outside such a ball. The desired set
E is in this case
and its volume can be majorized as follows:
where
is the ball of unit radius. We have therefore defined a finite-volume set. Although no picture can be drawn, one can think of
E as consisting of infinitely many balls about the points
of
, which are dense in
.
The topological boundary
of
E coincides with the whole of
minus
E. In fact, given a point
x of
, one can find a sequence of points with rational coordinates, extracted from the sequence
considered previously. Such an extracted sequence tends to
because the points with rational coordinates are dense. This means that
is an accumulation point for
E. Thus
x belongs to the closure
of
E. In turn, since
E is an open set (being formed by a countable union of open balls), the closure of
E is given by
and hence
Let us now prove that
E is a finite-perimeter set, despite the fact that its topological boundary has infinite volume. Indeed, by definition, the perimeter
is given by
where the integral over
E is a volume integral, and the defining conditions mean that
T is a vector field of
norm never bigger than 1, of class
on
and having compact support on
. By virtue of the definition (
3), one can write the volume integral of the divergence of
T in the form
The union of a finite number of balls
gives rise to an irregular set, because the balls do not have smooth intersections. Nevertheless, on such a set the divergence theorem holds, since such a set is piecewise smooth. One can therefore write, by virtue of the divergence theorem:
where
is the normal to the topological boundary of the union of open balls, and
is a surface measure. By virtue of (
8) and (
9), one finds the majorization (
being the measure of the surface given by the topological boundary of the union of open balls
)
where, on the first line, we have exploited the majorization of the
norm of
T and the unit norm of
, while on the second line we have exploited the property
This latter condition means that the topological boundary of the union of balls
is in general smaller than the union of the various topological boundaries of such balls. Furthermore, we can re-express the majorization (
10) as follows:
Indeed, if
, one can find a constant
M for which
By virtue of (
12), a real number
M exists such that
i.e., the set
E has a finite perimeter.
To sum up, the set
E defined in Equation (
3) is a finite-perimeter set of finite volume, while its topological boundary has infinite volume. Moreover, another consequence of De Giorgi’s structure theorem is that, if
E is a finite-perimeter set and if
T is a vector field of class
, then
Thus, the divergence theorem holds for a finite-perimeter set, but it does not involve the whole of the topological boundary (because might contain parts that are too irregular to see the divergence of T). The divergence theorem involves therefore only the reduced boundary , which represents the truly useful part of the topological boundary . With hindsight, this is the ultimate meaning of De Giorgi’s structure theorem and of the concept of reduced boundary. The finite-perimeter sets are therefore the most general objects for which the divergence theorem still holds.
In the example we have investigated it is impossible to explicitly evaluate the generalized normal. Nevertheless, by virtue of De Giorgi’s structure theorem, we know some qualitative properties of great value: the reduced boundary is given by the countable union of pieces of manifolds of class
, and the generalized normal is given by the normal to such manifolds. In such an example the reduced boundary is given by the countable union of pieces
of smooth manifolds, and such pieces are represented by awkward-looking pieces of spherical surfaces. One can also prove, for this example, the inclusion property
i.e., the reduced boundary is a subset of the countable union of the topological boundaries of all open balls centred at
and having radius
.
It is therefore clear that every explicit construction of generalized normal is a challenging task, which is nevertheless necessary if one wants to go beyond the purely qualitative properties. The following sections are devoted to two original examples, motivated by the desire to study geometric measure theory in non-Euclidean spaces. In particular, in
Section 2 we deal with hyperbolic geometry, whereas the case of elliptic geometry is investigated in
Section 3. The physical relevance of our framework is discussed in
Section 4, where we consider possible applications to the action principle for two-dimensional Euclidean quantum gravity. In particular, we deal with the Euclidean two-dimensional Callan–Giddings–Harvey–Strominger (hereafter referred to as CGHS) dilaton gravity model [
13,
14,
15] (see also Refs. [
16,
17,
18,
19]). As we will show, a two-dimensional model makes it possible to overcome the issue regarding the definition of the normal vector in higher-dimensional theories such as general relativity (for recent reviews of general relativity and beyond, see, e.g., Refs. [
20,
21,
22] and references therein). Last, the concluding remarks are made in
Section 5.
2. An Example of Generalized Normal in the Hyperbolic Plane
The hyperbolic (or Lobachevsky) plane [
23,
24,
25] is defined as the upper half-plane
in the complex plane
endowed with the metric
Geodesics (or hyperbolic lines) in
are defined in terms of Euclidean objects in
, being represented either by (the intersection of
with) Euclidean segments in
perpendicular to the real axis
or by (the intersection of
with) Euclidean circles (strictly speaking, we should distinguish between a circle and the associated circumference) in
having Euclidean centre on
and Euclidean radius
r. Any two points in
can be joined by a unique geodesic. Let
be the piecewise differentiable path
defining a geodesic in
. Bearing in mind (
17), one finds that the geodesic (
18) satisfies the following system of ordinary differential equations [
26]:
where
and
. When
, the solution of (19) is given by
representing a vertical line in
. On the other hand, if
is nonvanishing, the system (19) leads to
which describes a Euclidean positive semicircle with an Euclidean centre at
and Euclidean radius
r, since
satisfies
It is not difficult to show that the Euclidean center
and the Euclidean radius
r of the Euclidean circle through
can be written as [
24]
The hyperbolic length
of a piecewise differentiable path
is given by
The hyperbolic distance
between two points
is defined by the formula
where the infimum is taken over all paths
joining
. Equation (
26) defines a distance function on
, since it is non-negative, symmetric, and satisfies the triangle inequality [
23]. Moreover, it can be shown that for any
[
23] (when the points
are such that
, their hyperbolic distance reads as [
23,
24]
The hyperbolic distance
can also be written in terms of the Euclidean radius
r and the Euclidean center
of the geodesic connecting
as [
24]
The group of all isometries of
is isomorphic to the space
, which is defined as follows:
This is equivalent to expressing
as the quotient space
because, if we change sign to all matrix entries, both
and the determinant condition are preserved. Besides being a group,
is also a topological space in which the fractional linear transformation
can be identified with the point
of
. More precisely, as a topological space,
can be identified with the following subset of
:
If one defines
the map
is therefore a homeomorphism, and we can write that
which is a more precise expression of the quotient space formula involving
. Therefore
is a topological group, and the fractional linear maps
T have a norm induced from
given by
A hyperbolic
n-sided polygon
is a closed set of
(i.e., the Euclidean closure of
) bounded by
n hyperbolic geodesic segments. The vertices of
, defined as the points of intersection of two line segments, can lie in
, although no segment of
can belong to
[
23].
An example that makes it possible for us to display an explicit expression for the dual normal can be constructed as follows. Let
denote the
vertices of a hyperbolic
-sided polygon
. Such a polygon can be constructed by considering the unique geodesic connecting the following pairs of points:
and
,
and
,
and
, …,
and
, and eventually
and
. Since the points occurring in Equation (29) have different real parts, all the aforementioned geodesics will be Euclidean positive semicircles with a centre on the real axis
. The generic geodesic
joining
and
will be defined by
where, from Equation (19), we have
Therefore, we find that
where
The unit normal vector
to the geodesic (
30) will be
with
subjected to the conditions
and
.
We note that
which define an Euclidean degenerate circle. From the above equation jointly with formula (
28), we find that
We note that the above result cannot be obtained by employing Equation (
27), since
meaning that
does not belong to
when
(see
Appendix B for further details). In this limit, Equation (
27) would lead to a meaningless result.
The geodesic
connecting
and
will have a Euclidean centre at
and Euclidean radius
, whereas for the geodesic
through
and
the Euclidean centre lies at
and the Euclidean radius is
. From Equation (23), we find
and
In the limit of large
n, we obtain
which agree with Equation (39) (cf. also Equation (
37)).
The unit normal vector to the geodesic (
40) will be
satisfying the conditions
and
, whereas the unit normal vector to (
41) is
where
are such that the relations
and
are fulfilled.
The polygon
having vertices (29) has thus a generalized normal represented by Equations (
36), (
45), and (
46), i.e.,
It follows from the above equation that the
n-th segment of the polygon
admits a normal vector which differs from the one of the
-th segment. In other words, the generalized normal (
47) defines a discontinuous vector field and, in the limit of large
n, assumes an infinite number of values. Furthermore, we note that in our example the reduced boundary
of the polygon
has a form that agrees with the De Giorgi structure theorem (cf. Equation (
1)).
3. An Example of Generalized Normal in Elliptic Geometry
In Euclidean geometry, given a line and a point that does not lie on this line, there exists one and only one line that passes through the given point and is parallel to the given line. On the other hand, in hyperbolic geometry, which we have considered in the previous section, infinitely many distinct parallel lines can be found. A third option, where no parallel lines exist, gives rise to elliptic geometry [
27] and will be considered here. The most common model of elliptic geometry is represented by the surface of a sphere (however, it should be noted that in elliptic geometry two lines are usually assumed to intersect at a single point, while in spherical geometry two great circles intersect at two points, which is why spherical geometry is said to be a doubly elliptic geometry). Another relevant example of (two-dimensional) elliptic geometry is Klein’s conformal model of the elliptic plane [
27].
In this section, we consider an example of two-dimensional elliptic geometry having a squared line element [
28]
The resulting connection coefficients read as
and hence the geodesic equations take the form
, where the prime denotes differentiation with respect to the affine parameter
. If we eliminate the factor
from the second equation and substitute it in the first, we get
which leads to
Then, from Equations (50b) and (
52) we obtain the geodesic solutions
where
and
are integration constants.
We note that the system (50) is not affected if we interchange the role of
x and
y, i.e., it is invariant under the transformation
This means that
is a solution of Equation (50), where
are real-valued integration constants and
We can now set up a geodesic triangle
by means of Equations (
52)–(
54). Let
be the vertices of
(see
Figure 1).
Then, the side
of
connecting
A and
B is parametrized by
where
and the functions
and
can be obtained from Equations (
53) and (
54), respectively, by setting
The side
, which links the vertices
B and
C, is defined by
The functions
and
can be read off from Equations (
53) and (
54), respectively, with
Last,
can be obtained from
with
the functions
and
stemming from Equations (
53) and (
54), respectively, when
It should be noted that the solution having
can be regarded as a “limiting” case of Equations (
52)–(
54).
At this stage, we are ready to evaluate the normal vector field to the geodesic triangle
. Given the curve
, it is known that the equation of its normal to a generic point with the coordinates
is [
29]
the prime denoting differentiation with respect to the
x variable. In our example, the geodesics can be equivalently described by means of (cf. Equation (
52))
which means that, recalling Equation (62a), the equation of the side
reads also as (it is clear that Equation (
70) can also be obtained from Equation (
60))
and hence by virtue of Equation (
68) the normal vector field
to
is defined by
where
are the coordinates of a generic point belonging to
. As a consequence of (64a) (or, equivalently, Equation (
63)), we have for the side
from which we derive the form of its normal vector field
being the coordinates of a point lying on
. As pointed out previously, the side
cannot be obtained directly from (
69). Indeed, its equation is
which means that
is
Likewise the hyperbolic example discussed in
Section 2, Equations (
71), (
73), and (
75) give rise to a discontinuous normal vector field.
4. Application to Two-Dimensional Euclidean Quantum Gravity
As far as we know, the kind of thinking used so far in our paper has been lacking in the literature on fundamental interactions in physics. In recent times, much work in the mathematical literature has been devoted to the investigation of geometric measure theory in non-Euclidean spaces [
30]. Within such a framework, the occurrence of discontinuous normal vector fields has to be considered [
10], and our original explicit examples in
Section 2 and
Section 3 can be of interest.
On the other hand, the examples of
Section 2 and
Section 3 have a clear mathematical motivation, but they do not have an impact on theoretical physics. For this purpose, we are currently considering the case of Euclidean quantum gravity [
31,
32,
33,
34,
35,
36,
37,
38,
39,
40]. Within this framework, a prescription for functional integration is necessary, and we propose to consider only finite-perimeter sets that match the assigned data on their reduced boundary. We arrive at this prescription upon bearing in mind that measurable sets belong to two families: either they have finite perimeter, or they do not. The restriction to finite-perimeter sets might be severe, but it leads to mathematical properties which are under control and hence merits a careful assessment.
A second choice is also in order, and it has to do with the dimension of such finite-perimeter sets. In higher dimensions a problem arises, i.e., how to define a vector that plays the role of normal. Even just for a straight line in three-dimensional Euclidean space, what is defined is the plane orthogonal to such a line, but there is no coordinate-independent way of selecting two linearly independent normal vector fields to the line. Thus, a four-dimensional analogue of our
Section 2 and
Section 3 is not conceivable, as far as we can see. We have therefore resorted to two-dimensional Euclidean quantum gravity, by focusing on the Euclidean two-dimensional CGHS dilaton gravity model [
13,
14,
15] (see also e.g., Ref. [
41] for a general framework pertaining to Euclidean dilaton gravity in two dimensions). This example will clarify the role of the discontinuous normal
occurring in the divergence theorem (
14).
The Euclidean CGHS action reads as
g denoting the metric determinant,
R the Ricci scalar,
the dilaton field, and
a cosmological constant. Bearing in mind the recipes of Ref. [
42] (see also Refs. [
43,
44]), we can write the variations of the Lagrangian
occurring in Equation (
76) as
where
are the metric and the dilaton field equations, respectively, and
is referred to as the symplectic potential current density. By virtue of Equation (
76), we find
We aim at evaluating the integral
where
is a two-dimensional region having finite perimeter. For this purpose, we find it more convenient to consider an off-shell calculation, i.e., for metric and dilaton field which do not obey the Euclidean field equations. In such a way, we may choose an example as close as possible to the ones developed in the theory of finite-perimeter sets in
n-dimensional Euclidean space [
11], while avoiding the difficult task of solving the coupled partial differential equations for metric and dilaton.
Preparing the Ground for Evaluating the Integral
In order to prepare the ground for investigating the integral (
80), let us consider the following example of two-dimensional finite-perimeter set [
11], which provides a slight modification of the case discussed in
Section 1 (which we recall is valid in
with
, see Equation (
12)). Consider in
the set
E defined by
where, as before,
denotes the Euclidean open ball centred at the point
and having small positive radius
, which is supposed to satisfy
such that
with
being the Euler
-function. The perimeter of
is
denoting the one-dimensional Hausdorff measure. Starting from Equation (
85), it can be shown [
11] that for every
, the set
has finite perimeter since, by virtue of Equation (
83),
Therefore, as
(
being the volume of
E), we have
as
and hence
which means that
E has finite perimeter. Furthermore, it is possible to prove that
, which implies as a consequence
[
11].
We can now extend the previous example to a two-dimensional analytic Riemannian manifold
. For this reason, let us consider the two-dimensional geodesic ball
having small positive radius
r and centred at the point
. The volume
of
can be written by means of the power series [
45]
whereas the volume
of
reads as [
45]
If the trace
R of the Ricci tensor is positive, then it follows from Equations (
89) and (
90) that, for sufficiently small
r,
and
respectively.
At this stage, we have the necessary ingredients to generalize the previous example to the case when the manifold
admits a positive trace
R of Ricci. Indeed, along the same lines as before, let
be a surface of
constructed in terms of the geodesic balls
whose radius
is subject to the condition (
83). Bearing in mind Equation (
92), we have
which makes it possible for us to prove that the set
has finite perimeter, since, similarly to Equation (
87),
Therefore, the set
, which can be obtained from
in the limit
, has finite perimeter with
although
.