Skip to main content
Log in

On visual distances for spectrum-type functional data

  • Regular Article
  • Published:
Advances in Data Analysis and Classification Aims and scope Submit manuscript

Abstract

A functional distance \({\mathbb H}\), based on the Hausdorff metric between the function hypographs, is proposed for the space \({\mathcal E}\) of non-negative real upper semicontinuous functions on a compact interval. The main goal of the paper is to show that the space \(({\mathcal E},{\mathbb H})\) is particularly suitable in some statistical problems with functional data which involve functions with very wiggly graphs and narrow, sharp peaks. A typical example is given by spectrograms, either obtained by magnetic resonance or by mass spectrometry. On the theoretical side, we show that \(({\mathcal E},{\mathbb H})\) is a complete, separable locally compact space and that the \({\mathbb H}\)-convergence of a sequence of functions implies the convergence of the respective maximum values of these functions. The probabilistic and statistical implications of these results are discussed, in particular regarding the consistency of k-NN classifiers for supervised classification problems with functional data in \({\mathbb H}\). On the practical side, we provide the results of a small simulation study and check also the performance of our method in two real data problems of supervised classification involving mass spectra.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3

Similar content being viewed by others

References

  • Abraham C, Biau G, Cadre B (2006) On the kernel rule for function classification. Ann Inst Stat Math 58:619–633

    Article  MathSciNet  MATH  Google Scholar 

  • Alt H, Braß P, Godau M, Knauer C, Wenk C (2003) Computing the Hausdorff distance of geometric patterns and shapes. In: Aronov B, Basu S, Pach J, Sharir M (eds) Discrete and computational geometry. Springer, Berlin, pp 65–76

    Chapter  Google Scholar 

  • Baíllo A, Cuevas A, Cuesta-Albertos JA (2011) Supervised classification for a family of Gaussian functional models. Scand J Stat 38:480–498

    MathSciNet  MATH  Google Scholar 

  • Baíllo A, Cuevas A, Fraiman R (2011a) Classification methods for functional data. In: Ferraty F, Romain Y (eds) The Oxford handbook of functional data analysis. Oxford University Press, Oxford, pp 259–297

    Google Scholar 

  • Banks D, Petricoin E (2003) Finding cancer signals in mass spectrometry data. Chance 16:857

    MathSciNet  Google Scholar 

  • Billingsley P (1968) Convergence of probability measures. Wiley, New York

  • Bosq D (2000) Linear processes in function spaces. Theory and applications. Lecture Notes in Statistics, vol 149. Springer, Berlin

  • Bosq D, Blanke D (2007) Inference and prediction in large dimensions. Wiley, Chichester

    Book  MATH  Google Scholar 

  • Cérou F, Guyader A (2006) Nearest neighbor classification in infinite dimension. ESAIM Probab Stat 10:340–355

    Article  MathSciNet  MATH  Google Scholar 

  • Coombes KR, Baggerly KA, Morris JS (2007) Pre-processing mass spectrometry data. In: Dubitzky W, Granzow M, Berrar D (eds) Fundamentals of data mining in genomics and proteomics. Springer, New York, pp 79–102

    Chapter  Google Scholar 

  • Cuesta-Albertos JA, Fraiman R, Ransford T (2006) Random projections and goodness-of-fit tests in infinite-dimensional spaces. Bull Br Math Soc 37:1–25

    Article  MathSciNet  MATH  Google Scholar 

  • Cuevas A (2014) A partial overview of the theory of statistics with functional data. J Stat Plann Inference 147:1–23

    Article  MathSciNet  MATH  Google Scholar 

  • Cuevas A, Fraiman R (1998) On visual distances in density estimation: the Hausdorff choice. Stat Probab Lett 40:333–341

    Article  MathSciNet  MATH  Google Scholar 

  • Devroye L, Györfi L, Lugosi G (1996) A probabilistic theory of pattern recognition. Springer, New York

    Book  MATH  Google Scholar 

  • Dubuisson MP, Jain AK (1994) A modified Hausdorff distance for object matching. In: Proceedings of international conference on pattern recognition, Jerusalem, pp 566–568

  • Ferraty F, Vieu P (2006) Nonparametric functional data analysis: theory and practice. Springer, New York

    MATH  Google Scholar 

  • Folland GB (1999) Real analysis. Modern techniques and their applications. Wiley, New York

    MATH  Google Scholar 

  • Galeano P, Joseph E, Lillo R (2014) The Mahalanobis distance for functional data with applications to classification. Technometrics (2015, to appear)

  • Granas A, Dugundji J (2003) Fixed point theory. Springer, New York

    Book  MATH  Google Scholar 

  • Holá L (1992) Hausdorff metric on the space of upper semicontinuous multifunctions. Rocky Mt J Math 22:601–610

    Article  MathSciNet  MATH  Google Scholar 

  • Horváth L, Kokoszka P (2012) Inference for functional data with applications. Springer, New York

    Book  MATH  Google Scholar 

  • Huttenlocher DP, Klanderman GA, Rucklidge WJ (1993) Comparing images using the Hausdorff distance. IEEE Trans Pattern Anal Mach Intell 15:850–863

    Article  Google Scholar 

  • Ledoux M, Talagrand M (1991) Probability in Banach spaces. Isoperimetry and processes. Springer, New York

    Book  MATH  Google Scholar 

  • Natanson IP (1960) The theory of functions of a real variable, vol 2. Frederick Ungar Publishing Co., New York

    MATH  Google Scholar 

  • Nutanong S, Jacox E, Samet H (2011) An incremental Hausdorff distance calculation algorithm. Proc VLDB Endow 4:506–517

    Article  Google Scholar 

  • Ramsay JO, Silverman BW (2002) Applied functional data analysis. Methods and case studies. Springer, New York

    Book  MATH  Google Scholar 

  • Ramsay JO, Silverman BW (2005) Functional data analysis, 2nd edn. Springer, New York

    MATH  Google Scholar 

  • Rockafellar RT, Wets RJB (2009) Variational analysis. Springer, New York

    MATH  Google Scholar 

  • Sendov B (1990) Hausdorff approximations. Kluwer, Dordrecht

    Book  MATH  Google Scholar 

  • Sim DG, Kwon OK, Park RH (1999) Object matching algorithms using robust Hausdorff distance measures. IEEE Trans Image Process 8:425–429

    Article  Google Scholar 

  • Stone CJ (1977) Consistent nonparametric regression. Ann Stat 5:595–645

    Article  MATH  Google Scholar 

Download references

Acknowledgments

We are grateful to Prof. David Tudela (Department of Inorganic Chemistry, Universidad Autónoma de Madrid) for providing us with some useful information on spectra. A. Cuevas and R. Fraiman have been partially supported by Spanish Grant MTM2013-44045-P.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to A. Cholaquidis.

Appendix

Appendix

Proof of Proposition 2 To prove this Proposition we first must prove an auxiliary result:

Lemma 1

If \(f,g\in \mathcal {E}\), then there exist \(u\in \partial H_f\) and \(v\in \partial H_g\) such that

$$\begin{aligned} \mathbb {H}(f,g)=d(u,H_g)=d(v,H_f)=\Vert u-v \Vert . \end{aligned}$$
(4)

Proof

We have, by definition of \(\mathbb {H}\):

$$\begin{aligned} \mathbb {H}(f,g)=d_H(H_f,H_g)=\max \left\{ \sup _{a\in H_f} d(a,H_g), \sup _{b\in H_g}d(b,H_f)\right\} . \end{aligned}$$

Assume \({\mathbb H}(f,g)>0\). Otherwise the result is trivial. Let us suppose by contradiction that there is no pair \((u,v)\in \partial H_f\times \partial H_g\) such that (4) is fulfilled. In any case, the compactness of \(H_f\) and \(H_g\) guarantees the existence of \(x=(x_1,x_2)\) and \(y=(y_1,y_2)\) fulfilling (4) but, according to our contradiction argument, either x or y must be an interior point. For example, if \(x\in int(H_f)\), then \(0<x_1<1\). We will see that \(d((x_1,f(x_1)), H_g)\ge \mathbb {H}(f,g)\). For every \(t\in [0,1]\) such that \(|t-x_1|<\mathbb {H}(f,g)\) let us denote \(u^t=(t,u_2^t)\) and \(v^t=(t,v^t_2)\) the intersection points of \(\partial B(x,\mathbb {H}(f,g))\) and the line \(x_1=t\); with \(u^t_2<v^t_2\). From the assumption on x, \(d(x,H_g)={\mathbb H}(f,g)\). This entails that \(\mathring{B}(x,\mathbb {H}(f,g))\cap H_g=\emptyset \) and, since \(H_g\) is a hypograph (which implies that if \((a,b)\in H_g\) then the segment joining (ab) and (a, 0) is included in \(H_g\)) it is clear that \(g(t)\le u^t_2\), for all \(t\in [0,1]\) with \(|t-x_1|< \mathbb {H}(f,g)\). Therefore, if we move upwards the point \(x=(x_1,x_2)\) to \((x_1,f(x_1))\) (recall that from the USC assumption, \(x_2\le f(x_1)\)), we have \(\mathring{B}((x_1,f(x_1)),\mathbb {H}(f,g))\cap H_g=\emptyset \) and then \(d((x_1,f(x_1)), H_g)\ge \mathbb {H}(f,g)\). We cannot have \(d((x_1,f(x_1)), H_g)>\mathbb {H}(f,g)\) since \((x_1,f(x_1))\in H_f\) and \(\mathbb {H}(f,g)=d_H(H_f,H_g)\). So, we must have \(d((x_1,f(x_1)), H_g)= \mathbb {H}(f,g)\) with \(u:=(x_1,f(x_1))\in \partial H_f\). As a consequence, we must also have a point \(v\in \partial H_g\) such that \(\Vert u-v\Vert = \mathbb {H}(f,g)\). This contradicts the assumption we made about the non-existence of such a pair (uv). \(\square \)

We now prove Proposition 2.

Proof

Let us denote

$$\begin{aligned} d=\max \left\{ \sup _{\left\{ \begin{array}{c} x=(x_1,x_2)\in \partial H_g \\ g(x_1)\ge f(x_1) \end{array}\right\} }d\left( x,\partial H_f\right) ,\ \sup _{\left\{ \begin{array}{c} y=(y_1,y_2)\in \partial H_f \\ f(y_1)> g(y_1) \end{array}\right\} } d \left( y,\partial H_g\right) \right\} . \end{aligned}$$

The case \(\mathbb {H}(f,g)=0\) is trivial, so let us assume \(\mathbb {H}(f,g)>0\). We will first see that \(\mathbb {H}(f,g)\le d\). Since \(H_f\) and \(H_g\) are compact, there are two possibilities:

  1. (1)

    there exists \(x\in H_g\) such that \(\mathbb {H}(f,g)=d(x,\partial H_f)\), or

  2. (2)

    there exists \(y\in H_f\) such that \(\mathbb {H}(f,g)= d(y,\partial H_g)\).

Let us suppose that we are in the first case. By Lemma 1 we can assume that \(x=(x_1,x_2)\in \partial H_g\). Since \(\mathbb {H}(f,g)>0\) and \(H_f\) is a hypograph it must be \(g(x_1)> f(x_1)\), then \(x\in \{x=(x_1,x_2)\in \partial H_g: g(x_1)\ge f(x_1)\}\) from where it follows that \(\mathbb {H}(f,g)\le d\). If we are in case 2, again by Lemma 1, we can assume \(y=(y_1,y_2)\in \partial H_f\), as \(\mathbb {H}(f,g)>0\) and \(H_g\) is a hypograph it must be \(f(y_1)>g(y_1)\), then \(y \in \{y=(y_1,y_2)\in \partial H_f: f(y_1)> g(y_1) \}\). from where it follows that \(\mathbb {H}(f,g)\le d\). The inequality \({\mathbb H}(f,g)\ge d\) follows directly from the definition of \(\mathbb {H}\). \(\square \)

We also use the following proposition in the algorithm to calculate \({\mathbb H}(f,g)\) for f and g continuous.

Proposition 4

Let \(f,g\in \mathcal {E}\) be continuous functions, let u and v be the points of Lemma 1. Then, there exist \(t \in [0,1]\) and \(s\in [0,1]\) such that \(u=(t,f(t))\) and \(v=(s,g(s))\).

Proof

Again, assume \({\mathbb H}(f,g)>0\). By Lemma 1 \(u\in \partial H_f\), \(v\in \partial H_g\), and \(\mathbb {H}(f,g)=\Vert u-v\Vert \). So it is enough to prove that \(u=(t,f(t))\) and \(v= (s,g(s))\). Since f is continuous and \(u\in \partial H_f\), there are four possibilities: (and the same holds for \(v\in \partial H_g\)):

  1. 1.

    u is in the left border: \(u=(0,u_2)\) with \(u_2<f(0)\).

  2. 2.

    u is in the right border: \(u=(1,u_2)\) with \(u_2<f(1)\).

  3. 3.

    u is in the lower border: \(u=(u_1,0)\) with \(0\le u_1\le 1\).

  4. 4.

    u is in the upper border: \(0\le u_1\le 1\) y \(u_2=f(u_1)\).

We now prove that u can only be in Case 4. It is clear that Case 3 is not possible because both functions are non-negative. Cases 1 and 2 are also excluded following the ideas used in Lemma 1. For example, let us suppose that we are in Case 1 (see Fig. 4). First observe that \(\mathring{B}((0,f(0)),\mathbb {H}(f,g))\cap H_g=\emptyset \); otherwise there would exist \((t_1,t_2)\in \mathring{B} ((0,f(0)),\mathbb {H}(f,g))\cap H_g\), then, the segment joining the points \((t_1,0)\) and \((t_1,t_2)\) (which is included in \(H_g\)) intersects \(\mathring{B} (u,\mathbb {H}(f,g))\). But this contradicts \(\mathbb {H}(f,g)=d_H(H_f,H_g)\). So we conclude \(d ((0,f(0)),H_g)\ge \mathbb {H}(f,g)\). However \(d ((0,f(0)),H_g)> \mathbb {H}(f,g)\) leads to a contradiction with the definition of \(\mathbb {H}(f,g)\). Also, \(d(u,H_g)= d((0,f(0)),H_g)\) leads to another contradiction. Indeed, if this were the case, we would have two points (\((0,u_2)\) and (0, f(0))) on the vertical axis \(x_1=0\) which are equidistant to the hypograph \(H_g\). Then we have three possibilities:

  1. (a)

    \(u_2<g(0)<f(0)\). This contradicts \(d(u,H_g)= d ((0,f(0)),H_g)\), since all the points \((0,u_3)\) with \(u_2<u_3<g(0)\) belong to \(H_g\).

  2. (b)

    \(g(0)\le u_2\): this contradicts the continuity of g since \(H_g\) must have a point in the boundary of \(B((0,f(0)), d((0,f(0)),H_g))\) and no point in the open ball \(\mathring{B} ((0,u_2), d ((0,u_2),H_g))\).

  3. (c)

    \(g(0)\ge f(0)\): this is not compatible with \(d(u,H_g)= d ((0,f(0)),H_g)\).

Fig. 4
figure 4

We cannot have \(d(u,H_g)= d ((0,f(0)),H_g)\) with \(u=(0,u_2)\) and \(u_2<f(0)\)

Proof of Theorem 1

(a) To state the local compactness we will in fact prove a slightly stronger property: we will show that any closed and bounded set in \(({\mathcal E},{\mathbb H})\) is compact. Indeed, this would imply that the closed balls are compact. Since the family of balls with center at a given point is a local base, the local compactness will follow.

Since we are in a metric space compactness is equivalent to sequential compactness. Let us take \(\{f_n\}\subset \mathcal {E}\) a bounded sequence; we will prove that this sequence has necessarily a convergent subsequence. To see this, note that the corresponding sequence of compact sets \(H_{f_n}\) is bounded. So it has convergent subsequence, which we may denote again by \(H_{f_n}\), in the Hausdorff metric (since the closed and bounded sets are compact in the space of compact sets with the Hausdorff metric). Denote by C the limit of that subsequence. Therefore it is enough to prove that

$$\begin{aligned}&\text{ if } \left\{ H_{f_n}\right\} _n \text{ is } \text{ fulfils } H_{f_n}\rightarrow C \text{ for } \text{ some } \text{ compact } \text{ set } C, \text{ then } \text{ there } \nonumber \\&\text{ exists } \text{ a } \text{ USC } \text{ function } f:[0,1]\rightarrow [0,\infty )\text{, } \text{ such } \text{ that } C=H_{f}. \end{aligned}$$
(5)

Let us take \((x,y)\in C\) and \((x_n,y_n)\in H_{f_n}\) converging to (xy); note that there exists at least one such sequence because \(d_H(H_{f_n},C)\rightarrow 0\). Now, since the \(H_{f_n}\) are hypographs the vertical segment \([(x_n,0),(x_n,y_n)]\) joining the points \((x_n,0)\) and \((x_n,y_n)\) is included in \(H_{f_n}\). So \(H_{f_n}\rightarrow C\), implies

$$\begin{aligned} \left[ \left( x,0\right) ,\left( x,\limsup f_n(x_n)\right) \right] \subset C. \end{aligned}$$
(6)

Indeed, since \(H_{f_n}\) is a hypograph, \(f_n(x_n)\ge y_n\). Then if we take \(\limsup \) we obtain \(y\le \limsup f_n(x_n)\) and \(\{(x,z): 0\le z \le \limsup f(x_n)\}\subset C\).

Let us now define \(f:[0,1]\rightarrow [0,\infty )\) by

$$\begin{aligned} f(x)= \sup _{\{x_n\}:x_n\rightarrow x} \limsup f_n(x_n). \end{aligned}$$

Since \(\{f_n\}\) is bounded, f is well defined as a real-valued function. Let us prove that \(C=H_f\). Since C is closed, we have, by (6), \(H_{f}\subset C\). Moreover, if \((x,y)\in C\), taking \((x_n,y_n)\rightarrow (x,y)\) with \((x_n,y_n)\in H_{f_n}\), \(f(x)\ge \limsup f_n(x_n)\ge \limsup y_n=\, y\), we obtain \((x,y)\in H_f\).

It remains to prove that f is USC. Suppose by contradiction that there exists \(a \in [0,1]\) such that \(\limsup _{x\rightarrow a}f(x)>f(a)\). Then, we can take a constant \(\delta >0\) and a sequence \(x_n\rightarrow a\), \(x_n\ne a\) for all n such that \(f(x_n)>f(a)+\delta \) for n large enough, say \(n>n_0\). By the definition of f, for every \(x_n\) we can take a sequence \(z^k_n\rightarrow _k x_n\) (dependent on n), such that \(f(x_n)=\lim _{z^k_n\rightarrow _k x_n} \limsup f_k(z^k_n)\).

Given \(\varepsilon >0\), for every \(n>n_0\) let us take an increasing sequence \(k(n)>n_0\) with

$$\begin{aligned} \left| z_n^{k(n)}-x_n\right| <\frac{1}{n}\quad \text{ and }\quad \left| f_{k(n)}(z_n^{k(n)})-\limsup f_k(z^k_n)\right| <\varepsilon , \end{aligned}$$

that is, \(|f_{k(n)}(z_n^{k(n)})-f(x_n)|<\varepsilon \). But as \(z_n^{k(n)}\rightarrow _n a\) this contradicts \(f(x_n)>f(a)+\delta \) for \(n>n_0\).

Completeness follows directly from the fact that the space of compact sets endowed with the Hausdorff metric is complete, together with (5).

To prove separability, let \({\mathcal P}_n\) be the set of all partitions of [0, 1] defined by \(0=x_0<x_1<\dots <x_{n-1}<x_n=1\) where the \(x_i\) are rational numbers. Denote \(\mathcal {P}=\cup _n {\mathcal P}_n\). Note that \(\mathcal {P}\) in numerable.

Given a partition \(\mathcal {P}\in \mathcal {P}_n\) and a set \(q_0,\dots ,q_{n-1}\) of rational numbers, let us define

$$\begin{aligned} f_{\mathcal {P}}(x)={\left\{ \begin{array}{ll} q_0 &{} \text { if } x \in [0,x_1)\\ q_i &{}\text { if } x\in (x_i,x_{i+1})\quad 1 \le i \le n-3\\ q_{n-1} &{}\text { if } x\in (x_{n-1},1]\\ \max \{q_i,q_{i+1}\} &{}\text { if } x=x_i \quad 1 \le i \le n-2 \end{array}\right. } \end{aligned}$$
(7)

It is immediately seen that this function is USC and bounded. Let us see that the (numerable) set of all functions defined by (7), for all possible partitions \(\mathcal {P}\) and rational values \(q_i\) is dense in \(\mathcal {E}\) with respect to \(\mathbb {H}\). Let f be a non-negative USC function and take \(\varepsilon >0\). Consider \(\mathcal {P}\in \mathcal {P}_n\) a partition of the form \(0=x_0<x_1<\dots <x_{n-1}<x_n=1\) where \(x_i\) are rational numbers and such that \(\max _{i=0,\dots ,n-1} |x_{i+1}-x_i|<\varepsilon /2\). By Proposition 1 there exists \(f_i=\max _{x\in [x_i,x_{i+1}]} f(x)\). Let us take \(q_0,\dots ,q_{n-1}\) rational numbers such that \(q_i>f_i\) and \(q_i-f_i<\varepsilon /2\) for all i. For this partition and this set of rational numbers let us define \(f_{\mathcal {P}}\) as in (7). Now we claim that \(\mathbb {H}(f_\mathcal {P},f)\le \varepsilon \). Indeed, it is clear that \(f_\mathcal {P}(x)>f(x)\) for all \(x\in [0,1]\) so that \(H_f\subset H_{f_\mathcal {P}}\), and \(\mathbb {H}(f_\mathcal {P},f)= \sup _{z\in H_{f_\mathcal {P}}} d(z,H_f).\) Given \(z=(z_1,z_2)\in H_{f_\mathcal {P}}\), there exists \(0\le i_0 \le n-1 \) such that \(x_{i_0}\le z_1\le x_{i_0+1}\). Now, choose t such that \(x_{i_0}\le t\le x_{i_0+1}\) and \(f_{i_0}= f(t)\). We have \(z_2< f(t)+\varepsilon /2\) and then \(d(z,H_f)< \varepsilon \). Since z was an arbitrary point in \(H_{f_\mathcal {P}}\) we finally get \(\sup _{z\in H_{f_\mathcal {P}}} d(z,H_f)\le \epsilon \).

(b) By Proposition 1 (i) we know that there exists \(z\in [0,1]\) such that \(f(z)=\max _{x\in [0,1]} f(x)\). As \(\mathbb {H}(f_n,f)\rightarrow 0\) there exist \(x^n=(x_1^n,x_2^n)\in H_{f_n}\) such that \(x_n\rightarrow (z,f(z))\). Then, \(x^n_2\le f_n(x_1^n)\le \max _{x\in [0,1]} f_n(x)\) and, since \(x_2^n\rightarrow f(z)\), we obtain

$$\begin{aligned} \max _{x\in [0,1]} f(x) = f(z)\le \liminf _{n\rightarrow +\infty } \max _{x\in [0,1]} f_n(x) \le \limsup _{n\rightarrow +\infty } \max _{x\in [0,1]} f_n(x). \end{aligned}$$

Finally, let us prove that \(\limsup _n \max _{x\in [0,1]} f_n(x) \le \max _{x\in [0,1]} f(x)\). Denote \(z_0=\limsup _{n} \max _{x\in [0,1]} f_n(x)\). There exists \(x_n\in [0,1]\) such that \(f_n(x_n)\rightarrow z_0\) with \(f_n(x_n)= \max _{x\in [0,1]} f_n(x)\). Taking if necessary a subsequence, we can assume that \(x_n\rightarrow x_0\in [0,1]\). Since \((x_0,z_0)\in H_f\) we have \(f(x_0)\ge z_0\) then \(\max _{x\in [0,1]} f(x)\ge z_0\).

Proof of Proposition 3

Proof

This result is just a direct corollary from Th. 2 in Cérou and Guyader (2006) (recall that the continuity of \(\eta (x)\) is a sufficient condition for (2)), combined with the fact that the regression function \(\eta _Q(x)={\mathbb P}(Y=1|X=x)\) (i.e., the regression function under Q) can be approximated by a continuous compact supported function; we use here the local compactness of \({\mathcal E}\) (see Folland 1999, Proposition 7.9). Indeed, note that the joint distribution of (XY) is completely determined by \(\eta (x)={\mathbb P}(Y=1|X=x)\) and by the marginal distribution \(\mu \) of X. Then, given Q, one can construct P by just approximating \(\eta _Q(x)={\mathbb P}(Y=1|X=x)\) by a continuous compact-supported function \(\eta _P(x)\) which, without loss of generality, can be taken \(0\le \eta _P\le 1\). Then, the distribution P determined by \(\eta _P\) and the marginal distribution \(\mu \) of X is arbitrarily close to Q (just taking \(\eta _P\) close enough to \(\eta \)). Indeed, given any Borel set \(C \subset \mathcal E \times \{0,1\}\), consider the sets \(C_0 = \{x \in \mathcal E: (x,0) \in C\}\) and \(C_1 = \{x \in \mathcal E: (x,1) \in C\}\). Then,

$$\begin{aligned} Q(C) = \int _{C_0} (1-\eta _Q(x)) d\mu (x) + \int _{C_1} \eta _Q(x) d\mu (x), \end{aligned}$$

and

$$\begin{aligned} P(C) = \int _{C_0} (1-\eta _P(x)) d\mu (x) + \int _{C_1} \eta _P(x) d\mu (x), \end{aligned}$$

which can be made arbitrarily close. \(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Cholaquidis, A., Cuevas, A. & Fraiman, R. On visual distances for spectrum-type functional data. Adv Data Anal Classif 11, 5–24 (2017). https://doi.org/10.1007/s11634-015-0217-7

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11634-015-0217-7

Keywords

Mathematics Subject Classification

Navigation