We next characterise the long-term behaviour of Riemannian Brownian motion on configurations of exactly two landmarks, endowed with a Riemannian structure induced by a radial kernel as described in the preceding section. The key observation is that for a radial kernel, the distance between the two landmarks is a diffusion process, whose dynamics is characterised by a scalar stochastic differential equation. It then remains to study the singularity at zero of this one-dimensional diffusion process. For this, we follow the classification of singular points by Cherny and Engelbert in [6].
To derive the Itô stochastic differential equation for the distance process, we significantly reduce the complexity of the required computations by working in a well chosen coordinate system. Specifically, unlike a brute-force application of Itô's formula to the stochastic differential equation (4), our approach circumvents the need to determine all Christoffel symbols. Instead, it requires only the computation of one divergence. The reduction in complexity is best illustrated in the case of two landmarks in , which is why we first discuss this case, even though the result is included in the subsequent more general analysis for two landmarks in .
3.1 Distance process between two landmarks
Let be a radial kernel (6) described in terms of a functions which is continuous on and continuously differentiable on . We further set .
3.1.1 Configurations with two landmarks in
For landmark configurations with two landmarks in
, we have
Since the radial kernel
takes the form (
6), we have, for
,
Due to (
3) describing the cometric induced by the Green's kernel
, it follows that the metric
on
induced by
is given as
We now change to a system of coordinates in which the expression for the metric
diagonalises. For
, we set
The constraint
then amounts to the condition
. Without loss of generality, we may work in the half plane
, which corresponds to the assumption that the Riemannian Brownian motion is started from a landmark configuration where the first landmark is bigger than the second one. The Brownian motion on
only leaves the half plane defined by
if the two landmarks collide or a landmark escapes to infinity in finite time, that is, the Riemannian Brownian motion explodes.
From
and
, we obtain
Since
in the half plane
, it follows that, for
,
(9) as well as
Therefore, the vector fields
and
on
defined by
(10) form an orthonormal frame
for the tangent bundle
with respect to the metric
. In particular, we can write
(11) where the divergence is taken with respect to the induced Riemannian volume measure. From the expression (
11), we can read off as for (
5) that the Brownian motion on
, that is, the diffusion process
on
with generator
is the unique strong solution to the Stratonovich stochastic differential equation
where
and
are independent one-dimensional standard Brownian motions.
Due to the form (
10) of the vector field
, it follows that, for
, the distance process
between the two landmarks induced by the Brownian motion on
is the unique strong solution to the Stratonovich stochastic differential equation
(12) It remains to compute the divergence of the vector field
explicitly and to express the Stratonovich stochastic differential equation as an Itô stochastic differential equation.
As a consequence of (
9), the Riemannian volume form
on
induced by the Riemannian metric
can be expressed in the coordinates
as
It follows that
which yields
We further note that, for
,
from which we can read off the drift term contribution arising from
. Since
we deduce that the Stratonovich stochastic differential equation (
12) can be rewritten as the Itô stochastic differential equation
(13) One can check that this is consistent with the expression obtained by starting directly from (
4).
3.1.2 Configurations with two landmarks in
The restriction to radial kernels of the form (6) allows us to describe the distance process between two landmarks, provided that no additional landmarks are present, by a one-dimensional stochastic differential equation. The argument, which was developed for two landmarks in in the previous subsection, carries over to two landmarks in , as shown next.
As before, we simplify the computations significantly by working in suitable coordinates, for which the metric tensor diagonalises. When considering configurations consisting of two landmarks in
, for
, we have
Moreover, the radial kernel
of the form (
6) is given by, for
,
where
denotes the
identity matrix. As in the analysis for two landmarks in
, it further follows from (
3) and the above expression for the radial kernel
that the induced metric
on
is determined by
We proceed by changing coordinates for the landmark space
from
to
, where
are given by
and by setting
Note that the constraint
is equivalent to the condition
. As two landmarks collide if and only if their distance process hits zero, it suffices to study the stochastic dynamics of the distance process, for landmarks evolving according to a Riemannian Brownian motion. For
, we obtain
as well as
Moreover, the radial vector field
satisfies at
that
In particular, the vector field
on
defined by
is of unit length. As the vector field
depends on the distance component
only, and as we can extend
locally to an orthonormal frame
for
, the distance process between the two landmarks is the one-dimensional diffusion process with generator
To determine the associated Itô stochastic differential equation, we first use
to determine
We further observe that we still have, for
,
and we compute
Thus, the distance process
between the two landmarks solves the Itô stochastic differential equation
(14) where
is a one-dimensional standard Brownian motion, and with the diffusivity
as well as the drift
given by
(15) Note that this is consistent with the Itô stochastic differential equation (
13) derived for
.
3.2 Collision analysis for two landmarks
In the previous section, we derived the Itô stochastic differential equation which governs the dynamics of the distance process between two landmarks in induced by the Riemannian Brownian motion on landmark configurations consisting of exactly two landmarks. Studying if the two landmarks collide and aiming for a classification which depends on the choice of the kernel as well as the dimension then amounts to analysing the one-dimensional diffusion process near the singularity at zero. Indeed, the landmarks collide if and only if the distance process hits zero. For our analysis, we follow the classification for singular points of one-dimensional diffusion processes by Cherny and Engelbert [6].
This classification depends on the asymptotic behaviour of the coefficients given in (
15) of the stochastic differential equation (
14) near zero, and hence on the small-distance asymptotics of the kernel. Motivated by our main examples, Gaussian and Sobolev kernels, we assume that there exist real-valued constants
such that, as
,
(16) To reduce the notational overhead, we write
as
if there exists a non-zero constant
such that, as
,
To start off analysing the singularity at zero for the one-dimensional diffusion process
, which is the unique strong solution to the Itô stochastic differential equation (
14), we remark that, as
,
For
, the function above is locally integrable near zero, and according to [
6, Definition 2.3], zero is then a regular point of (
14). In this case, as a consequence of [
6, Theorem 2.11], the diffusion process
hits zero in finite time with positive probability, meaning that landmarks collide with positive probability.
The remainder of this subsection is devoted to the case
, where
According to [
6, Definition 2.3], zero is then a singular point of (
14). For sufficiently small
, we have
(17) Hence, we can proceed with the classification of singularities by Cherny and Engelbert [
6], which is very well summarised on [
6, p. 39]. Throughout, we fix
such that (
17) is satisfied.
In the first step of the classification process, we need to consider the function
defined by
We compute that
By employing the change of variables
and subject to
being sufficiently small, we further obtain
which yields
From (
16), it then follows directly that, as
,
(18) which shows that
We further deduce that, as
,
implying that
To complete the classification for
and
, we observe that in this case, as
,
from which we conclude
Thus, according to the result [
6, Theorem 2.12], for
and
, the singularity of (
14) at zero is of type 2. In particular, the associated distance process
hits zero with positive probability, that is, the two landmarks collide with positive probability in finite time.
The remaining classification steps use the function
defined by
From (
18), we obtain that, as
,
Therefore, for
and
, we have, as
,
whilst in all other cases, we have, as
,
It follows that, irrespective of the dimension
,
At this stage, we deduce that if
and
, then by [
6, Theorem 2.13] the singularity of (
14) at zero is of type 1, which particularly implies that in this case, the two landmarks collide with positive probability in finite time. For
and
, it is a consequence of [
6, Theorem 2.17] that the singularity of (
14) at zero is then of type 5. Hence, in this case, any solution to (
14) started at a non-zero distance is strictly positive. Together with the argument presented in the following section, this implies that the associated landmark Brownian motion exists for all times.
To conclude the classification of the singularity at zero for the distance process
, it remains to observe that, for
and
, as
,
which yields that, for
and
,
Hence, by [
6, Theorem 2.15], in the case
and
, the singularity of (
14) at zero is of type 4. Whilst this together with Section
3.3 still establishes long-time existence of the associated landmark Brownian motion, the type of singularity detected implies that with positive probability the two landmarks draw arbitrarily close together, if measured with respect to the Euclidean metric.
A summary of the results derived is provided in Section 3.4 below.
3.3 Ruling out escape to infinity
Besides collision of landmarks, the only other source of Brownian incompleteness is escape to infinity, which we investigate next. We show that escape to infinity cannot occur before collision and therefore plays no role in the analysis of Brownian completeness or incompleteness in our setting. The presented proof requires that the kernel
and its derivative
are Lipschitz continuous away from zero and that
vanishes at infinity, as shall be assumed herewith. As before,
denotes Brownian motion on the space of two landmarks in
. Moreover, we write
for the inter-landmark distance of
. Then, the escape-to-infinity time
and the collision time
are defined as
We claim that
, which we establish by proving the equivalent result that
for all
. To this end, we truncate the stochastic differential equation (
4) for
using a Lipschitz function
such that
for
and
for
to obtain
(19) This truncated stochastic differential equation is well posed because its coefficients are Lipschitz continuous, as we show below. Hence,
does not escape to infinity in finite time. Moreover,
coincides with the Brownian motion
on the stochastic interval
. Consequently,
does not escape to infinity before
, that is,
for all
, as claimed.
We now show that the diffusivity coefficient in (
19) is Lipschitz continuous on the set
of all landmark configurations
with inter-landmark distance
. As the scalar kernel
is assumed to be Lipschitz on
, the cometric
is Lipschitz continuous on
. Moreover, as the scalar kernel
by assumption extends continuously to the compact set
, the set
has compact closure, namely, the set
itself together with the matrix
Thus, all of these matrices are positive definite. Taking the square root of a symmetric positive definite matrix is smooth by the implicit function theorem or, more generally, because the functional calculus is real analytic, see [
3]. In particular, the matrix square root is Lipschitz continuous on compacts. Consequently,
is Lipschitz continuous on
.
We next show that the drift coefficient of (19) is Lipschitz continuous on . The derivative of the scalar kernel is assumed to be Lipschitz continuous on . Therefore, the cometric has Lipschitz continuous coordinate derivatives on . Matrix inversion is real analytic and hence Lipschitz continuous on compacts. Thus, the metric is Lipschitz continuous on . The Christoffel symbol can be written as a contraction of the metric with coordinate derivatives of the cometric . Therefore, the Christoffel symbol is Lipschitz continuous on . Taken together, this implies the Lipschitz continuity of the drift in (19) on .
This concludes the proof that , that is, escape to infinity cannot occur before collision. An important consequence is that Brownian completeness follows as soon as collisions are ruled out.
3.4 Summary
The long-term behaviour of Riemannian Brownian motion on the configuration space of two landmarks in
depends on whether the ambient space has dimension
or
, and on whether the near-zero asymptotics (
16) of the kernel are given by
or
. Our characterisation follows [
6] and is well described in terms of the two hitting times
and
.
- If , then the two landmarks collide with positive probability. More specifically, for , the origin is a regular point of the distance process , which implies that and . For , it is a singular point of type 2 if and of type 1 if . This means that, subject to , in the case , there exists a unique non-negative solution to (14) up to such that and , whereas if , there exists a unique solution to (14) which is defined up to and has as well as .
- If and , then the landmark Brownian motion exists for all times. The singular point at zero is of type 5, meaning that for , any solution to (14) is strictly positive, and subject to there exists a unique solution defined up to and -a.s.
- If and , then the landmark Brownian motion exists for all times. The singularity at zero is of type 4, which implies that as long as any solution to (14) is strictly positive, and for , there exists a unique solution defined up to where as well as -a.s. on . In particular, the two landmarks almost surely do not collide, but with positive probability their Euclidean distance becomes arbitrarily small as time tends to infinity.
We next discuss implications for the Bessel potentials defined by (7), which are also known as Sobolev kernels. The Bessel potential of order in dimensions has asymptotics (16) with , as can be seen from (8). A minor modification is needed in the case to accommodate the logarithmic term in the asymptotics (8), but careful inspection of the arguments in Section 3.2 shows that the conclusion remains the same as for without the logarithmic term.
To summarise, the Sobolev metric of order gives rise to a Brownian complete Riemannian manifold if and to a Riemannian manifold which is Brownian incomplete otherwise. Note that this is also the threshold for the reproducing kernel Hilbert space to embed into , that is, for admissibility of this space of vector fields in the terminology of [20]. Interestingly, we have Brownian completeness not only above this threshold but also in the critical case . For the Gaussian kernel, one has both Brownian completeness and admissibility, in line with the interpretation of the Gaussian kernel as a Sobolev kernel of infinite order.