nips nips2009 nips2009-238 nips2009-238-reference knowledge-graph by maker-knowledge-mining

238 nips-2009-Submanifold density estimation


Source: pdf

Author: Arkadas Ozakin, Alexander G. Gray

Abstract: Kernel density estimation is the most widely-used practical method for accurate nonparametric density estimation. However, long-standing worst-case theoretical results showing that its performance worsens exponentially with the dimension of the data have quashed its application to modern high-dimensional datasets for decades. In practice, it has been recognized that often such data have a much lower-dimensional intrinsic structure. We propose a small modification to kernel density estimation for estimating probability density functions on Riemannian submanifolds of Euclidean space. Using ideas from Riemannian geometry, we prove the consistency of this modified estimator and show that the convergence rate is determined by the intrinsic dimension of the submanifold. We conclude with empirical results demonstrating the behavior predicted by our theory. 1 Introduction: Density estimation and the curse of dimensionality Kernel density estimation (KDE) [8] is one of the most popular methods for estimating the underlying probability density function (PDF) of a dataset. Roughly speaking, KDE consists of having the data points “contribute” to the estimate at a given point according to their distances from the ˆ point. In the simplest multi-dimensional KDE [3], the estimate fm (y0 ) of the PDF f (y0 ) at a point N y0 ∈ R is given in terms of a sample {y1 , . . . , ym } as, 1 ˆ fm (y0 ) = m m i=1 1 K hN m yi − y0 hm , (1) where hm > 0, the bandwidth, is chosen to approach to zero at a suitable rate as the number m of data points increases, and K : [0.∞) → [0, ∞) is a kernel function that satisfies certain properties such as boundedness. Various theorems exist on the different types of convergence of the estimator to the correct result and the rates of convergence. The earliest result on the pointwise convergence rate in the multivariable case seems to be given in [3], where it is stated that under certain conditions for f and K, assuming hm → 0 and mhm → ∞ as m → ∞, the mean squared ˆ ˆ error in the estimate f (y0 ) of the density at a point goes to zero with the rate, MSE[fm (y0 )] = E ˆ fm (y0 ) − f (y0 ) 2 = O h4 + m 1 mhN m as m → ∞. If hm is chosen to be proportional to m−1/(N +4) , one gets, ˆ MSE[fm (p)] = O 1 m4/(N +4) , (2) as m → ∞. This is an example of a curse of dimensionality; the convergence rate slows as the dimensionality N of the data set increases. In Table 4.2 of [12], Silverman demonstrates how the sample size required for a given mean square error for the estimate of a multivariable normal distribution increases with the dimensionality. The numbers look as discouraging as the formula 2. 1 One source of optimism towards various curses of dimensionality is the fact that although the data for a given problem may have many features, in reality the intrinsic dimensionality of the “data subspace” of the full feature space may be low. This may result in there being no curse at all, if the performance of the method/algorithm under consideration can be shown to depend only on the intrinsic dimensionality of the data. Alternatively, one may be able to avoid the curse by devising ways to work with the low-dimensional data subspace by using dimensional reduction techniques on the data. One example of the former case is the results on nearest neighbor search [6, 2] which indicate that the performance of certain nearest-neighbor search algortihms is determined not by the full dimensionality of the feature space, but only on the intrinsic dimensionality of the data subspace. Riemannian manifolds. In this paper, we will assume that the data subspace is a Riemannian manifold. Riemannian manifolds provide a generalization of the notion of a smooth surface in R3 to higher dimensions. As first clarified by Gauss in the two-dimensional case (and by Riemann in the general case) it turns out that intrinsic features of the geometry of a surface such as lengths of its curves or intrinsic distances between its points, etc., can be given in terms of the so-called metric tensor1 g without referring to the particular way the the surface is embedded in R3 . A space whose geometry is defined in terms of a metric tensor is called a Riemannian manifold (for a rigorous definition, see, e.g., [5, 7, 1]). Previous work. In [9], Pelletier defines an estimator of a PDF on a Riemannian manifold M by using the distances measured on M via its metric tensor, and obtains the same convergence rate as in (2), with N being replaced by the dimensionality of the Riemannian manifold. Thus, if we know that the data lives on a Riemannian manifold M , the convergence rate of this estimator will be determined by the dimensionality of M , instead of the full dimensionality of the feature space on which the data may have been originally sampled. While an interesting generalization of the usual KDE, this approach assumes that the data manifold M is known in advance, and that we have access to certain geometric quantities related to this manifold such as intrinsic distances between its points and the so-called volume density function. Thus, this Riemannian KDE cannot be used directly in a case where the data lives on an unknown Riemannian submanifold of RN . Certain tools from existing nonlinear dimensionality reduction methods could perhaps be utilized to estimate the quantities needed in the estimator of [9], however, a more straightforward method that directly estimates the density of the data as measured in the subspace is desirable. Other related works include [13], where the authors propose a submanifold density estimation method that uses a kernel function with a variable covariance but do not present theorerical results, [4] where the author proposes a method for doing density estimation on a Riemannian manifold by using the eigenfunctions of the Laplace-Beltrami operator, which, as in [9], assumes that the manifold is known in advance, together with intricate geometric information pertaining to it, and [10, 11], which discuss various issues related to statistics on a Riemannian manifold. This paper. In this paper, we propose a direct way to estimate the density of Euclidean data that lives on a Riemannian submanifold of RN with known dimension n < N . We prove the pointwise consistency of the estimator, and prove bounds on its convergence rates given in terms of the intrinsic dimension of the submanifold the data lives in. This is an example of the avoidance of the curse of dimensionality in the manner mentioned above, by a method whose performance depends on the intrinsic dimensionality of the data instead of the full dimensionality of the feature space. Our method is practical in that it works with Euclidean distances on RN . In particular, we do not assume any knowledge of the quantities pertaining to the intrinsic geometry of the underlying submanifold such as its metric tensor, geodesic distances between its points, its volume form, etc. 2 The estimator and its convergence rate Motivation. In this paper, we are concerned with the estimation of a PDF that lives on an (unknown) n-dimensional Riemannian submanifold M of RN , where N > n. Usual, N -dimensional kernel density estimation would not work for this problem, since if interpreted as living on RN , the 1 The metric tensor can be thought of as giving the “infinitesimal distance” ds between two points whose P coordinates differ by the infinitesimal amounts (dy 1 , . . . , dy N ) as ds2 = ij gij dy i dy j . 2 underlying PDF would involve a “delta function” that vanishes when one moves away from M , and “becomes infinite” on M in order to have proper normalization. More formally, the N -dimensional probability measure for such an n-dimensional PDF on M will have support only on M , will not be absolutely continuous with respect to the Lebesgue measure on RN , and will not have a probability density function on RN . If one attempts to use the usual, N -dimensional KDE for data drawn from such a probability measure, the estimator will “try to converge” to a singular PDF, one that is infinite on M , zero outside. In order to estimate the probability density function on M by using data given in RN , we propose a simple modification of usual KDE on RN , namely, to use a kernel that is normalized for n-dimensions instead of N , while still using the Euclidean distances in RN . The intuition behind this approach is based on three facts: 1) For small distances, an n-dimensional Riemannian manifold “looks like” Rn , and densities in Rn should be estimated by an n-dimensional kernel, 2) For points of M that are close enough to each other, the intrinsic distances as measured on M are close to Euclidean distances as measured in RN , and, 3) For small bandwidths, the main contribution to the estimate at a point comes from data points that are nearby. Thus, as the number of data points increases and the bandwidth is taken to be smaller and smaller, estimating the density by using a kernel normalized for n-dimensions and distances as measured in RN should give a result closer and closer to the correct value. We will next give the formal definition of the estimator motivated by these considerations, and state our theorem on its asymptotics. As in the original work of Parzen [8], the proof that the estimator is asymptotically unbiased consists of proving that as the bandwidth converges to zero, the kernel function becomes a “delta function”. This result is also used in showing that with an appropriate choice of vanishing rate for the bandwidth, the variance also vanishes asymptotically, hence the estimator is pointwise consistent. Statement of the theorem Let M be an n-dimensional, embedded, complete Riemannian submanifold of RN (n < N ) with an induced metric g and injectivity radius rinj > 0.2 Let d(p, q) be the length of a length-minimizing geodesic in M between p, q ∈ M , and let u(p, q) be the geodesic (linear) distance between p and q as measured in RN . Note that u(p, q) ≤ d(p, q). We will use the notation up (q) = u(p, q) and dp (q) = d(p, q). We will denote the Riemannian volume measure on M by V , and the volume form by dV . Theorem 2.1. Let f : M → [0, ∞) be a probability density function defined on M (so that the related probability measure is f V ), and K : [0, ∞) → [0, ∞) be a continous function that satisfies vanishes outside [0, 1), is differentiable with a bounded derivative in [0, 1), and satisfies, n z ≤1 K( z )d z = 1. Assume f is differentiable to second order in a neighborhood of p ∈ M , ˆ and for a sample q1 , . . . , qm of size m drawn from the density f , define an estimator fm (p) of f (p) as, m 1 1 up (qj ) ˆ fm (p) = (3) K n m j=1 hm hm where hm > 0. If hm satisfies limm→∞ hm = 0 and limm→∞ mhn = ∞, then, there exists m non-negative numbers m∗ , Cb , and CV such that for all m > m∗ we have, ˆ MSE fm (p) = E ˆ fm (p) − f (p) 2 < Cb h4 + m CV . mhn m If hm is chosen to be proportional to m−1/(n+4) , this gives, E (fm (p) − f (p))2 = O as m → ∞. (4) 1 m4/(n+4) Thus, the convergence rate of the estimator is given as in [3, 9], with the dimensionality replaced by the intrinsic dimension n of M . The proof will follow from the two lemmas below on the convergence rates of the bias and the variance. 2 The injectivity radius rinj of a Riemannian manifold is a distance such that all geodesic pieces (i.e., curves with zero intrinsic acceleration) of length less than rinj minimize the length between their endpoints. On a complete Riemannian manifold, there exists a distance-minimizing geodesic between any given pair of points, however, an arbitrary geodesic need not be distance minimizing. For example, any two non-antipodal points on the sphere can be connected with two geodesics with different lengths, namely, the two pieces of the great circle passing throught the points. For a detailed discussion of these issues, see, e.g., [1]. 3 3 Preliminary results The following theorem, which is analogous to Theorem 1A in [8], tells that up to a constant, the kernel becomes a “delta function” as the bandwidth gets smaller. Theorem 3.1. Let K : [0, ∞) → [0, ∞) be a continuous function that vanishes outside [0, 1) and is differentiable with a bounded derivative in [0, 1), and let ξ : M → R be a function that is differentiable to second order in a neighborhood of p ∈ M . Let ξh (p) = 1 hn K M up (q) h ξ(q) dV (q) , (5) where h > 0 and dV (q) denotes the Riemannian volume form on M at point q. Then, as h → 0, K( z )dn z = O(h2 ) , ξh (p) − ξ(p) (6) Rn where z = (z 1 , . . . , z n ) denotes the Cartesian coordinates on Rn and dn z = dz 1 . . . dz n denotes the volume form on Rn . In particular, limh→0 ξh (p) = ξ(p) Rn K( z )dn z. Before proving this theorem, we prove some results on the relation between up (q) and dp (q). Lemma 3.1. There exist δup > 0 and Mup > 0 such that for all q with dp (q) ≤ δup , we have, 3 dp (q) ≥ up (q) ≥ dp (q) − Mup [dp (q)] . In particular, limq→p up (q) dp (q) (7) = 1. Proof. Let cv0 (s) be a geodesic in M parametrized by arclength s, with c(0) = p and initial vedcv locity ds0 s=0 = v0 . When s < rinj , s is equal to dp (cv0 (s)) [7, 1]. Now let xv0 (s) be the representation of cv0 (s) in RN in terms of Cartesian coordinates with the origin at p. We have up (cv0 (s)) = xv0 (s) and x′ 0 (s) = 1, which gives3 x′ 0 (s) · x′′0 (s) = 0. Using these v v v we get, dup (cv0 (s)) ds s=0 the absolute value of the third d3 up (cv0 (s)) ds3 d2 up (cv0 (s)) = ds2 s=0 derivative of up (cv0 (s)) = 1 , and 0. Let M3 ≥ 0 be an upper bound on for all s ≤ rinj and all unit length v0 : 3 ≤ M3 . Taylor’s theorem gives up (cv0 (s)) = s + Rv0 (s) where |Rv0 (s)| ≤ M3 s . 3! Thus, (7) holds with Mup = M3 , for all r < rinj . For later convenience, instead of δu = rinj , 3! we will pick δup as follows. The polynomial r − Mup r3 is monotonically increasing in the interval 0 ≤ r ≤ 1/ 3Mup . We let δup = min{rinj , 1/ Mup }, so that r − Mup r3 is ensured to be monotonic for 0 ≤ r ≤ δup . Definition 3.2. For 0 ≤ r1 < r2 , let, Hp (r1 , r2 ) = Hp (r) = inf{up (q) : r1 ≤ dp (q) < r2 } , Hp (r, ∞) = inf{up (q) : r1 ≤ dp (q)} , (8) (9) i.e., Hp (r1 , r2 ) is the smallest u-distance from p among all points that have a d-distance between r1 and r2 . Since M is assumed to be an embedded submanifold, we have Hp (r) > 0 for all r > 0. In the below, we will assume that all radii are smaller than rinj , in particular, a set of the form {q : r1 ≤ dp (q) < r2 } will be assumed to be non-empty and so, due to the completeness of M , to contain a point q ∈ M such that dp (q) = r1 . Note that, Hp (r1 ) = min{H(r1 , r2 ), H(r2 )} . (10) Lemma 3.2. Hp (r) is a non-decreasing, non-negative function, and there exist δHp > 0 and MHp ≥ H (r) 0 such that, r ≥ Hp (r) ≥ r − MHp r3 , for all r < δHp . In particular, limr→0 pr = 1. 3 Primes denote differentiation with respect to s. 4 Proof. Hp (r) is clearly non-decreasing and Hp (r) ≤ r follows from up (q) ≤ dp (q) and the fact that there exists at least one point q with dp (q) = r in the set {q : r ≤ dp (q)} Let δHp = Hp (δup ) where δup is as in the proof of Lemma 3.1 and let r < δHp . Since r < δHp = Hp (δup ) ≤ δup , by Lemma 3.1 we have, r ≥ up (r) ≥ r − Mup r3 , (11) for some Mup > 0. Now, since r and r − Mup r3 are both monotonic for 0 ≤ r ≤ δup , we have (see figure) (12) r ≥ Hp (r, δup ) ≥ r − Mup r3 . In particular, H(r, δup ) ≤ r < δHp = Hp (δup ), i.e, H(r, δup ) < Hp (δup ). Using (10) this gives, Hp (r) = Hp (r, δup ). Combining this with (12), we get r ≥ Hp (r) ≥ r − Mup r3 for all r < δHp . Next we show that for all small enough h, there exists some radius Rp (h) such that for all points q with a dp (q) ≥ Rp (h), we have up (q) ≥ h. Rp (h) will roughly be the inverse function of Hp (r). Lemma 3.3. For any h < Hp (rinj ), let Rp (h) = sup{r : Hp (r) ≤ h}. Then, up (q) ≥ h for all q with dp (q) ≥ Rp (h) and there exist δRp > 0 and MRp > 0 such that for all h ≤ δRp , Rp (h) satisfies, (13) h ≤ Rp (h) ≤ h + MRp h3 . In particular, limh→0 Rp (h) h = 1. Proof. That up (q) ≥ h when dq (q) ≥ Rp (h) follows from the definitions. In order to show (13), we will use Lemma 3.2. Let α(r) = r − MHp r3 , where MHp is as in Lemma 3.2. Then, α(r) is oneto-one and continuous in the interval 0 ≤ r ≤ δHp ≤ δup . Let β = α−1 be the inverse function of α in this interval. From the definition of Rp (h) and Lemma 3.2, it follows that h ≤ Rp (h) ≤ β(h) for all h ≤ α(δHp ). Now, β(0) = 0, β ′ (0) = 1, β ′′ (0) = 0, so by Taylor’s theorem and the fact that the third derivative of β is bounded in a neighborhood of 0, there exists δg and MRp such that β(h) ≤ h + MRp h3 for all h ≤ δg . Thus, h ≤ Rp (h) ≤ h + MRp h3 , (14) for all h ≤ δR where δR = min{α(δHp ), δg }. Proof of Theorem 3.1. We will begin by proving that for small enough h, there is no contribution to the integral in the definition of ξh (p) (see (5)) from outside the coordinate patch covered by normal coordinates.4 Let h0 > 0 be such that Rp (h0 ) < rinj (such an h0 exists since limh→ 0 Rp (h) = 0). For any h ≤ h0 , all points q with dp (q) > rinj will satisfy up (q) > h. This means if h is small enough, u (q) K( ph ) = 0 for all points outside the injectivity radius and we can perform the integral in (5) solely in the patch of normal coordinates at p. For normal coordinates y = (y 1 , . . . , y n ) around the point p with y(p) = 0, we have dp (q) = y(q) [7, 1]. With slight abuse of notation, we will write up (y(q)) = up (q), ξ(y(q)) = ξ(q) and g(q) = g(y(q)), where g is the metric tensor of M . Since K( up (q) h ) = 0 for all q with dp (q) > Rp (h), we have, ξh (p) = 1 hn K y ≤Rp (h) up (y) h ξ(y) g(y)dy 1 . . . dy n , (15) 4 Normal coordinates at a point p in a Riemannian manifold are a close approximation to Cartesian coordinates, in the sense that the components of the metric have vanishing first derivatives at p, and gij (p) = δij [1]. Normal coordinates can be defined in a “geodesic ball” of radius less than rinj . 5 where g denotes the determinant of g as calculated in normal coordinates. Changing the variable of integration to z = y/h, we get, K( z )dn z = ξh (p) − ξ(p) up (zh) h K z ≤Rp (h)/h up (zh) h K = z ≤1 ξ(zh) K z ≤1 z ≤1 g(zh) − 1 dn z + ξ(zh) up (zh) h K( z )dn z g(zh)dn z − ξ(0) ξ(zh) − K( z ) dn z + K( z ) (ξ(zh) − ξ(0)) dn z + z ≤1 K 1< z ≤Rp (h)/h up (zh) h ξ(zh) g(zh)dn z . Thus, K ( z ) dn z ≤ ξh (p) − ξ(p) (16) t∈R z ≤1 sup |ξ(zh)| . sup K( z ≤1 z ≤1 dn z + g(zh) − 1 . sup K(t) . sup |ξ(zh)| . sup z ≤1 (17) z ≤1 up (zh) ) − K( z ) . h dn z + (18) z ≤1 K( z )(ξ(zh) − ξ(0))dn z + (19) z ≤1 sup K(t) . t∈R sup g(zh) . 1< z ≤Rp (h)/h sup dn z . (20) |ξ(zh)| . 1< z ≤Rp (h)/h 1< z ≤Rp (h)/h Letting h → 0, the terms (17)-(20) approach zero at the following rates: (17): K(t) is bounded and ξ(y) is continuous at y = 0, so the first two terms can be bounded by constants as h → 0. In normal coordinates y, gij (y) = δij + O( y 2 ) as y → 0, so, sup z ≤1 g(zh) − 1 = O(h2 ) as h → 0. (18): Since K is assumed to be differentiable with a bounded derivative in [0, 1), we get K(b) − u (zh) K(a) = O(b − a) as b → a. By Lemma 3.1 we have p h − z = O(h2 ) as h → 0. Thus, K up (zh) h − K( z ) = O(h2 ) as h → 0. (19): Since ξ(y) is assumed to have partial derivatives up to second order in a neighborhood of y(p) = 0, for z ≤ 1, Taylor’s theorem gives, n zi ξ(zh) = ξ(0) + h i=1 as h → 0. Since h → 0. z ≤1 zK( z )dn z = 0, we get ∂ξ(y) ∂y i z ≤1 + O(h2 ) (21) y=0 K( z )(ξ(zh) − ξ(0))dn z = O(h2 ) as (20): The first three terms can be bounded by constants. By Lemma 3.3, Rp (h) = h + O(h3 ) as h → 0. A spherical shell 1 < z ≤ 1 + ǫ has volume O(ǫ) as ǫ → 0+ . Thus, the volume of 1 < z ≤ Rp (h)/h is O(Rp (h)/h − 1) = O(h2 ) as h → 0. Thus, the sum of the terms (17-20), is O(h2 ) as h → 0, as claimed in Theorem 3.1. 6 4 Bias, variance and mean squared error ˆ Let M , f , fm , K, p be as in Theorem 2.1 and assume hm → 0 as m → ∞. ˆ Lemma 4.1. Bias fm (p) = O(h2 ), as m → ∞. m u (q) Proof. We have Bias[fm (p)] = Bias h1 K ph m follows from Theorem 3.1 with ξ replaced with f . , so recalling Rn K( z )dn z = 1, the lemma Lemma 4.2. If in addition to hm → 0, we have mhn → ∞ as m → ∞, then, Var[fm (p)] = m O 1 mhn m , as m → ∞. Proof. Var[fm (p)] = 1 1 Var n K m hm up (q) hm (22) (23) Now, Var 1 K hn m up (q) hm =E 1 K2 h2n m up (q) hm 1 hn m f (q) − E 1 K hn m 2 up (q) hm , (24) and, E 1 K2 h2n m up (q) hm = M 1 2 K hn m up (q) hm dV (q) . (25) By Theorem 3.1, the integral in (25) converges to f (p) K 2 ( z )dn z, so, the right hand side of (25) is O 1 hn m ˆ Var[fm (p)] = O as m → ∞. By Lemma 4.1 we have, E 1 mhn m 1 hn K m up (q) hm 2 → f 2 (p). Thus, as m → ∞. ˆ ˆ ˆ Proof of Theorem 2.1 Finally, since MSE fm (p) = Bias2 [fm (p)] + Var[fm (p)], the theorem follows from Lemma 4.1 and 4.2. 5 Experiments and discussion We have empirically tested the estimator (3) on two datasets: A unit normal distribution mapped onto a piece of a spiral in the plane, so that n = 1 and N = 2, and a uniform distribution on the unit disc x2 + y 2 ≤ 1 mapped onto the unit hemisphere by (x, y) → (x, y, 1 − x2 + y 2 ), so that n = 2 and N = 3. We picked the bandwidths to be proportional to m−1/(n+4) where m is the number of data points. We performed live-one-out estimates of the density on the data points, and obtained the MSE for a range of ms. See Figure 5. 6 Conclusion and future work We have proposed a small modification of the usual KDE in order to estimate the density of data that lives on an n-dimensional submanifold of RN , and proved that the rate of convergence of the estimator is determined by the intrinsic dimension n. This shows that the curse of dimensionality in KDE can be overcome for data with low intrinsic dimension. Our method assumes that the intrinsic dimensionality n is given, so it has to be supplemented with an estimator of the dimension. We have assumed various smoothness properties for the submanifold M , the density f , and the kernel K. We find it likely that our estimator or slight modifications of it will be consistent under weaker requirements. Such a relaxation of requirements would have practical consequences, since it is unlikely that a generic data set lives on a smooth Riemannian manifold. 7 MSE MSE Mean squared error for the hemisphere data Mean squared error for the spiral data 0.000175 0.0008 0.00015 0.000125 0.0006 0.0001 0.000075 0.0004 0.00005 0.0002 0.000025 # of data points 50000 100000 150000 200000 # of data points 50000 100000 150000 200000 Figure 1: Mean squared error as a function of the number of data points for the spiral data (left) and the hemisphere data. In each case, we fit a curve of the form M SE(m) = amb , which gave b = −0.80 for the spiral and b = −0.69 for the hemisphere. Theorem 2.1 bounds the MSE by Cm−4/(n+4) , which gives the exponent as −0.80 for the spiral and −0.67 for the hemisphere. References [1] M. Berger and N. Hitchin. A panoramic view of Riemannian geometry. The Mathematical Intelligencer, 28(2):73–74, 2006. [2] A. Beygelzimer, S. Kakade, and J. Langford. Cover trees for nearest neighbor. In Proceedings of the 23rd international conference on Machine learning, pages 97–104. ACM New York, NY, USA, 2006. [3] T. Cacoullos. Estimation of a multivariate density. Annals of the Institute of Statistical Mathematics, 18(1):179–189, 1966. [4] H. Hendriks. Nonparametric estimation of a probability density on a Riemannian manifold using Fourier expansions. The Annals of Statistics, 18(2):832–849, 1990. [5] J. Jost. Riemannian geometry and geometric analysis. Springer, 2008. [6] F. Korn, B. Pagel, and C. Faloutsos. On dimensionality and self-similarity . IEEE Transactions on Knowledge and Data Engineering, 13(1):96–111, 2001. [7] J. Lee. Riemannian manifolds: an introduction to curvature. Springer Verlag, 1997. [8] E. Parzen. On estimation of a probability density function and mode. The Annals of Mathematical Statistics, pages 1065–1076, 1962. [9] B. Pelletier. Kernel density estimation on Riemannian manifolds. Statistics and Probability Letters, 73(3):297–304, 2005. [10] X. Pennec. Probabilities and statistics on Riemannian manifolds: Basic tools for geometric measurements. In IEEE Workshop on Nonlinear Signal and Image Processing, volume 4. Citeseer, 1999. [11] X. Pennec. Intrinsic statistics on Riemannian manifolds: Basic tools for geometric measurements. Journal of Mathematical Imaging and Vision, 25(1):127–154, 2006. [12] B. Silverman. Density estimation for statistics and data analysis. Chapman & Hall/CRC, 1986. [13] P. Vincent and Y. Bengio. Manifold Parzen Windows. Advances in Neural Information Processing Systems, pages 849–856, 2003. 8


reference text

[1] M. Berger and N. Hitchin. A panoramic view of Riemannian geometry. The Mathematical Intelligencer, 28(2):73–74, 2006.

[2] A. Beygelzimer, S. Kakade, and J. Langford. Cover trees for nearest neighbor. In Proceedings of the 23rd international conference on Machine learning, pages 97–104. ACM New York, NY, USA, 2006.

[3] T. Cacoullos. Estimation of a multivariate density. Annals of the Institute of Statistical Mathematics, 18(1):179–189, 1966.

[4] H. Hendriks. Nonparametric estimation of a probability density on a Riemannian manifold using Fourier expansions. The Annals of Statistics, 18(2):832–849, 1990.

[5] J. Jost. Riemannian geometry and geometric analysis. Springer, 2008.

[6] F. Korn, B. Pagel, and C. Faloutsos. On dimensionality and self-similarity . IEEE Transactions on Knowledge and Data Engineering, 13(1):96–111, 2001.

[7] J. Lee. Riemannian manifolds: an introduction to curvature. Springer Verlag, 1997.

[8] E. Parzen. On estimation of a probability density function and mode. The Annals of Mathematical Statistics, pages 1065–1076, 1962.

[9] B. Pelletier. Kernel density estimation on Riemannian manifolds. Statistics and Probability Letters, 73(3):297–304, 2005.

[10] X. Pennec. Probabilities and statistics on Riemannian manifolds: Basic tools for geometric measurements. In IEEE Workshop on Nonlinear Signal and Image Processing, volume 4. Citeseer, 1999.

[11] X. Pennec. Intrinsic statistics on Riemannian manifolds: Basic tools for geometric measurements. Journal of Mathematical Imaging and Vision, 25(1):127–154, 2006.

[12] B. Silverman. Density estimation for statistics and data analysis. Chapman & Hall/CRC, 1986.

[13] P. Vincent and Y. Bengio. Manifold Parzen Windows. Advances in Neural Information Processing Systems, pages 849–856, 2003. 8