1. Introduction
Large deviations theory is one of the most classical subfields of probability theory with a wide range of applications in information theory, statistical physics, and rare-event simulation [8]. While the most refined results are available in the study of sequences of random variables and of time-dependent stochastic processes, the understanding of large deviations in spatial random systems is still in its infancy. Even the most basic statistics such as the edge counts in a random geometric graph can give rise to surprising and highly non-trivial condensation effects on the level of large deviations [6].
One of the earliest achievements in this context is [13], which proves a large deviation principle (LDP) for statistics of a homogeneous Poisson point process in under rather restrictive boundedness and locality assumptions. Later, these conditions could be relaxed substantially so as to allow for statistics with bounded exponential moments and stabilizing score functions [25, 26]. While the main focus there was on scalar and measure-valued LDPs in the thermodynamic regime, very recently also geometric statistics in the sparse and dense regimes of the Poisson point process were considered [16, 17].
All of the aforementioned results have in common that they deal with random geometric systems in Euclidean space. However, during the last years the hyperbolic space has received substantial attention in the context of complex networks [11]. Moreover, there has been vigorous activity to understand the asymptotic behavior of geometric functionals of random set systems in hyperbolic space as well. While there has been substantial progress for central and non-central limit theorems and Poisson approximation results [1, 2, 3, 10, 12, 14, 15, 18, 21, 22], large deviation principles have not been considered so far. By investigating the volume of -nearest neighbor (kNN) balls in hyperbolic space in the present work, we provide a first step in this direction and complement the recent findings [21] about their extremal behavior.
The general proof strategy is to extend and to develop further the ideas of [16] on a coarse graining scheme to introduce a blocked point process. Inside each block, a Poisson approximation theorem in the spirit of [21] is used to replace the original functional by a Poisson point process for which the LDP is given by Sanov’s theorem. However, the intricate geometry of the hyperbolic space makes it far more challenging to implement the blocking argument in comparison to the Euclidean setting considered in [16]. For instance, in the Euclidean setting the kissing number is finite so that in a box of side length of order only a uniformly bounded number of points with a nearest neighbor radius exceeding can be placed. In contrast, in hyperbolic space, this number grows exponentially in ; see [9]. Moreover, in the half-space model of hyperbolic space, the Poisson intensity is inhomogeneous in the vertical direction when considered in Euclidean terms, while in the -setting there is homogeneity in all directions. We deal with these problems by deriving more refined exponential moment bounds and analyzing the point configuration in vertical layers that individually can be considered approximately homogeneous.
We believe that the techniques developed in the present article open the door to the investigation of further LDPs in hyperbolic space. In particular, we think of extending the results for the Euclidean component counts from [17] in the sparse regime. Here, we note that while the study of component counts is restricted to trees in the dense regime [22] such constraints are no longer present in the sparse regime.
The rest of the manuscript is organized as follows. In Section 2, we properly introduce the considered model and state the LDP for the volumes of large kNN balls. Next, in Section 3, we give a proof outline, where we reduce the assertion to two key auxiliary results, namely Propositions 3.1 and 3.2. These results are proven separately in Sections 4 and 5, respectively.
2. Model and main result
By the hyperbolic space of dimension we mean the unique simply connected, -dimensional Riemannian manifold of constant sectional curvature , cf. [5, 24]. There are several models to represent in the -dimensional Euclidean space . To carry out our computations, we will work in this paper with the so-called half-space model, but we emphasize that all results we derive are actually model independent; for background material on half-space model of hyperbolic space we refer to [24, Chapter 4.6]. In the half-space model, we identify with the product space . The Riemannian metric is then determined by
|
|
|
and we denote by the hyperbolic distance of two points , which in terms of the Euclidean distance between and is given by
|
|
|
see [24, Theorem 4.6.1].
According to [24, Theorem 4.6.7], the hyperbolic volume measure on has density
(2.1) |
|
|
|
with respect to the Lebesgue measure on and we will use the notation
|
|
|
for the hyperbolic volume of a measurable set . In contrast, we shall write for the Lebesgue measure of the appropriate dimension, which will always be clear from the context. In addition, we denote for and by the hyperbolic ball of radius centered at . Its hyperbolic volume satisfies
(2.2) |
|
|
|
independently of , with being the surface content of the -dimensional Euclidean unit sphere; see [24, page 79]. In particular, grows like a constant multiple of , as .
Henceforth, we let be the random counting measure distributed as a Poisson point process with the hyperbolic volume measure as its intensity measure. We note that such is stationary in the sense that its distribution is invariant with respect to the full group of hyperbolic isometries. In particular, is the number of points of falling into a measurable set .
In this paper, we write or sometimes for the restriction of the measure to . In what follows, we shall restrict to the family of sampling windows
|
|
|
whose hyperbolic volume is given by .
For we study the asymptotics of large -nearest neighbor (kNN) balls centered in . To make this precise, we define the point process
|
|
|
where
|
|
|
and is a threshold sequence satisfying
|
|
|
In particular, the expected number of exceedances in the window is of order .
Our main result is a large deviation principle (LDP) for the volumes of large kNN balls. We recall that a family of random variables , defined on some probability space and taking values in a Polish space , satisfies an LDP with speed and rate function , provided that is lower semicontinuous and if for each measurable set one has that
|
|
|
where and stand for the interior and the closure of , respectively. To present our main result, we fix and introduce the space as well as the measure on which has Lebesgue density , . By we denote the space of finite measures on , which supplied with the weak topology becomes a Polish space [7, Proposition A2.5.III]. For we let
(2.3) |
|
|
|
be the relative entropy entropy of with respect to , where indicates that is absolutely continuous with respect to and denotes the corresponding Radon-Nikodym derivative; see [13, Equation (2.10)]. We are now prepared to present the main result of this paper.
Theorem 2.1 (LDP for kNN balls).
Let . Then, the family of random measures satisfies an LDP on with speed and rate function .
As highlighted in Section 1, Theorem 2.1 can be seen as the hyperbolic counterpart of [16, Theorem 2.1], which concerns large deviations of the empirical measure of recentered and rescaled kNN balls in Euclidean space. On a formal level, in the hyperbolic setting, we found it more convenient to consider a scaling where the expected number of Poisson points in the window grows exponentially in , whereas in [16] the corresponding scaling is linear in parameter . After this rescaling, our condition on the growth of corresponds precisely to the growth of in [16, Equation (2.1)]. Moreover, in the Euclidean setting the dense scaling in [16] can be equivalently transformed to the regime of growing windows considered here. However, we stress that the regimes of dense points and growing windows are no longer equivalent in hyperbolic setting, and only the growing-window asymptotics reflects the negative curvature effects from the hyperbolic space.
While the previous paragraph illustrates that there is a formal correspondence between Theorem 2.1 and the Euclidean analog [16, Theorem 2.1], the hyperbolic geometry creates substantial complications, which require us to develop novel methodological tools. First, while both Theorem 2.1 and [16, Theorem 2.1] rely on a Poisson approximation result for large kNN balls, such a result is substantially harder to establish in hyperbolic geometry. We therefore adapt the arguments of a very recent work [21]. Moreover, when establishing a central uniform integrability property, [16] relies crucially on a kissing-number argument. More precisely, while in Euclidean space, the number of non-intersecting balls of radius that can be put into a box of side length of order is uniformly bounded independently of , such a property does not hold in hyperbolic space. Therefore, we need to develop delicate exponential moment bound that ensure uniform integrability despite a potentially unbounded number of balls.
To make the presentation more accessible, we provide a brief summary of the notation used throughout this paper, most terms will formally be defined in the forthcoming sections:
- : cardinality of a set
- : indicator function
- : Landau symbols
- : underlying probability space
- : expectation(integration) wrt.
- : law of a random element
- : Dirac measure
- : space of locally finite measures
- : -dimensional hyperbolic space,
x identified with
- : relative entropy of a measure wrt.
- : restriction of a measure to a set
- :hyperbolic ball of radius centered at
- : hyperbolic volume measure
x with differential
- : Lebesgue measure
- : hyperbolic distance
- : Euclidean distance
- : total variation distance
- : Kantorovich-Rubinstein distance
In addition, simple point processes with will be identified with locally finite point configurations . Moreover, by abuse of notation, we write if is an atom of the counting measure . The intensity measure of is denoted by .
In this paper we will denote by a generic constant whose value might change from occurrence to occurrence. If several such constants are needed at the same time, we denote them by
Finally, since the parameters and are fixed in the rest of the manuscript, we omit them from the notation.
3. Proof outline
To prove Theorem 2.1, we partition into congruent blocks and set . We note that is not necessarily an integer. However, to keep notation simple we do not write rounding symbols.
The key step in the proof of Theorem 2.1 is to relate the empirical measure with the following separated point process:
|
|
|
Here, for a locally finite counting measure on we put with and let be an ‘internal region’ that we will define precisely in (3.1) below. The idea behind introducing these internal regions is that we want large kNN balls to occur independently for distinct regions. In other words, this step will remove the dependence of kNN radii exceeding the threshold , which is defined such that .
Having introduced the process , the proof of Theorem 2.1 can be split up into two steps regarding exponential equivalence (with respect to the total variation distance) in the sense of [8, Definition 4.2.10]:
-
(1)
show that the family of random measures is exponentially equivalent to , where for each , is a Poisson point process on whose intensity measure has density with respect to the Lebesgue measure;
-
(2)
show that the family of random measures is exponentially equivalent to .
Proposition 3.1 (Exponential equivalence of and ).
The families of random measures and are exponentially equivalent.
Proposition 3.2 (Exponential equivalence of and ).
The families of random measures and are exponentially equivalent.
The main task is to prove Propositions 3.1 and 3.2, which are the hyperbolic analogs of [16, Proposition 4.3] and [16, Propositions 4.4–4.6]. Before doing so in Sections 4 and 5, we briefly explain how to deduce Theorem 2.1 from these two results.
Proof of Theorem 2.1.
First, we note that by the Poisson variant of Sanov’s theorem, the family of random measures satisfies an LDP with speed and rate function ; see [8, Theorem 6.2.10]. Hence, taking into account Propositions 3.1 and 3.2, the result follows from the fundamental fact in large deviations theory that exponentially equivalent families of random elements satisfy the same LDP; see [8, Theorem 4.2.13].
∎
We conclude the present overview section, with a precise definition of the internal regions . To that end, we let be a diverging sequence with . Now, for each , we define
(3.1) |
|
|
|
where
with
, refers to the Euclidean distance in and stands for the boundary of the set .
The desired properties of the internal regions are described in the following proposition.
Proposition 3.3 (Containment and volume of ).
It holds that
-
(1)
for every .
-
(2)
.
To prove Proposition 3.3, we rely on the following distance formula taken from [10, Proposition 2.1]. We also recall that we identify points with pairs in the product space .
Lemma 3.4.
Let , and define and . Then,
|
|
|
where
In particular, by minimizing over ,
(3.2) |
|
|
|
holds for all sufficiently large .
Proof.
To carry out the minimization, we set the . This gives . Hence, is asymptotically equivalent to , which in turn is equivalent to for large .
∎
We now prove Proposition 3.3. For the proof, we note that
(3.3) |
|
|
|
for suitable constants ; see Equation (4.1) in [21].
Proof of Proposition 3.3.
We prove separately parts (i) and (ii).
Part (i)
Let and . Then, by (3.2),
|
|
|
In other words, .
Part (ii)
Note that is a cube of side length so that
|
|
|
Hence, for , by (2.1) and Fubini’s theorem, can be bounded as
|
|
|
|
|
|
|
|
Therefore,
|
|
|
as .
Next, consider . Here, the important observation is that if . That is, if . Now,
|
|
|
Moreover,
|
|
|
|
|
|
|
|
|
|
|
|
Noting that concludes the proof.
∎
4. Proof of Proposition 3.1
In this section, we prove Proposition 3.1, that is, the exponential equivalence of the empirical measures and a family of empirical measures of suitable Poisson point processes. We recall that is composed of its constituents in the individual boxes. The first step of the proof is therefore to proceed as in [16, Proposition 4.1] and establish a Poisson approximation result individually for each . We do this by showing the stronger assertion that for each the Kantorovich-Rubinstein distance between the law of and that of tends to zero, as . Here, we recall that for two point processes on ,
|
|
|
where the supremum runs over all measurable Lipschitz- functions on the space of point processes on the space with respect to the total variation distance . For two measures on the latter is given by
|
|
|
Proposition 4.1 (Poisson approximation for separated processes).
Let . Then,
|
|
|
where is a sequence of independent and identically distributed Poisson point processes on , each having intensity measure .
To prepare the proof, some elements of which are similar to the main computations in [21], we recall that for locally finite, we put . We also define and . Then, and are localized to in the sense of [4]. Namely, for every and all , we have that
Moreover, is a so-called stopping set, in the sense that if , then
for every compact set .
First, we set . Henceforth, we put
|
|
|
Proof of Proposition 4.1.
We shall check a set of sufficient conditions for convergence in Kantorovich-Rubinstein distance as provided in Theorem 6.4 in [4]. This requires the analysis of the total variance distance as well bounds on three error terms , and which will be defined below.
First, it is claimed that
|
|
|
Recall that if and only if . Hence, for , it follows from the Mecke equation for Poisson point processes [20, Theorem 4.1], that
|
|
|
|
|
|
|
|
This implies that has the Lebesgue density
|
|
|
and thus,
|
|
|
|
|
|
The second term above vanishes as because of the dominated convergence theorem. For the first term, we note that Proposition 3.3 gives
|
|
|
It is now concluded that , as .
To ease notation, we henceforth write instead of .
Subsequently, we demonstrate that
|
|
|
Notice that
|
|
|
where the second equality is due to the fact that for all . Now,
|
|
|
|
where we have used that .
Next, we turn our attention to the second error term
|
|
|
We here apply the following inequalities:
|
|
|
|
|
|
|
|
Hence,
(4.1) |
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
For the inequality at the last line, we have applied (3.3).
For the bound of we distinguish between the cases and . If we have that
|
|
|
and, hence,
|
|
|
|
|
|
|
|
|
|
|
|
where in the last step we used the volume estimate for and the fact that for any and ; see the discussion after (2.2).
For , we split into two terms:
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
where is so small that . The term is bounded by
|
|
|
|
|
|
|
|
Let be a fixed reference point (sometimes called the origin) and let denote the exponential map at in which is the tangent space at . Then, by the above
the bound,
|
|
|
It follows from the polar integration formula in hyperbolic geometry [5, pp. 123-125] that
|
|
|
|
|
|
|
|
|
|
|
|
where for some , the -dimensional unit sphere in . It follows from Lemma 5 in [21] that
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
Here, the last inequality follows from the constraint .
For , one can see that
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
where again for some .
Note that for every with ,
|
|
|
and further, by Lemma 4 in [21], .
Using this bound,
|
|
|
|
|
|
|
|
Applying the substitution , we find that the integral is bounded in and, hence, that , as .
We now put all bounds together and apply Theorem 6.4 in [4] to conclude that, for each ,
|
|
|
This completes the proof.
∎
By the maximal coupling lemma [19, Lemma 4.32], Proposition 4.1 gives for each couplings and satisfying
(4.2) |
|
|
|
Hence, we can define a coupling between and by setting
|
|
|
where we assume that the pairs are independent and identically distributed for different . Having constructed the couplings, we now proceed with the proof of the exponential equivalence.
Proof of Proposition 3.1.
First, we note that for every
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
where we used the exponential Markov inequality and the i.i.d. property of the pairs discussed above.
Moreover, by (4.2), we have that , as in probability. Hence, as in [16, Proposition 4.3], it suffices to verify the uniform integrability condition
(4.3) |
|
|
|
for every , where we let
|
|
|
denote the point process of exceedances. The key observation that we will use tacitly throughout the rest of the proof is that to determine the restriction of to a set , it suffices to know in an -neighborhood of . Therefore, we partition into vertical layers , where . In order to avoid dealing with an infinite number of such layers, we define a value such that . As in previous arguments, here, is assumed to be an integer; if this is not the case, we replace by which is a small perturbation of . We then set .
In particular, we deduce from Lemma 3.4 that is independent of whenever and are such that . Hence, it suffices to prove that
(4.4) |
|
|
|
The arguments for layers with or are similar.
We start with , where to ease notation, we set . Note that by the kNN property, we have that if is not among the nearest neighbors of , then
|
|
|
By the same reason, each point is covered by at most further balls of the form . Therefore,
|
|
|
where in the last step we used once again Lemma 4 in [21]. Now, note that and let for , be a Poisson variable with mean . Thus, by the multivariate Mecke equation for Poisson point processes [20, Theorem 4.4], for ,
|
|
|
|
(4.5) |
|
|
|
|
Now, we want to apply the concentration inequality for Poisson random variables [23, Lemma 1.2]. That means, we need to provide an upper bound for the expression
|
|
|
To that end, we note that the first exponential is bounded by
for a suitable . Moreover, since , we conclude that .
Plugging this into (4.5) gives that for and suitable ,
|
|
|
|
|
|
|
|
Therefore, by Markov’s inequality,
|
|
|
Hence, for large we have
|
|
|
which remains bounded since .
Next, to bound , we proceed similarly as in the proof of [16, Proposition 4.3] and decompose into independent horizontal blocks. More precisely, we consider the diluted family of boxes , where we set
|
|
|
Hence, applying Hölder’s inequality we reduce the verification of (4.4) to showing that
(4.6) |
|
|
|
where . Now, proceeding as in the case of shows that for large
|
|
|
Since the last expression belongs to , the proof is complete.
∎