Skip to main content
Log in

Nonparametric statistics of dynamic networks with distinguishable nodes

  • Original Paper
  • Published:
TEST Aims and scope Submit manuscript

Abstract

The study of random graphs and networks had an explosive development in the last couple of decades. Meanwhile, techniques for the statistical analysis of sequences of networks were less developed. In this paper, we focus on networks sequences with a fixed number of labeled nodes and study some statistical problems in a nonparametric framework. We introduce natural notions of center and a depth function for networks that evolve in time. We develop several statistical techniques including testing, supervised and unsupervised classification, and some notions of principal component sets in the space of networks. Some examples and asymptotic results are given, as well as two real data examples.

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
Fig. 4
Fig. 5
Fig. 6

Similar content being viewed by others

References

  • Adar E, Zhang L, Adamic LA, Lukose RM (2004) Implicit structure and the dynamics of blogspace. Implicit Struct Dyn Blogspace 13:16989–16995

    Google Scholar 

  • Ahmed S, Li D, Rosalsky A, Volodin A (2001) Almost sure lim sup behavior of bootstrapped means with applications to pairwise i.i.d. sequences and stationary ergodic sequences. J Stat Plan Infer 98:126–137

    Article  MathSciNet  MATH  Google Scholar 

  • Aparicio D, Fraiman D (2015) Banking networks and Leverage dependence in emerging countries. Adv Complex Syst 18:1550022

    Article  MathSciNet  Google Scholar 

  • Arcones MA, Cui H, Zuo Y (2006) Empirical depth processes. TEST 15:151–177

    Article  MathSciNet  MATH  Google Scholar 

  • Auer J (1995) An elementary proof of the invertibility of distance matrices. Linear Multilinear A 40:119–124

    Article  MathSciNet  MATH  Google Scholar 

  • Barabasi A, Albert R (1999) Emergence of scaling in random networks. Science 286(5439):509–512

    Article  MathSciNet  MATH  Google Scholar 

  • Bickel P, Chen A (2009) A nonparametric view of network models and Newman–Girvan and other modularities. Proc Natl Acad Sci USA 106:21068–21073

    Article  MATH  Google Scholar 

  • Bollobás B, Janson S, Riordan O (2007) The phase transition in inhomogeneous random networks. Random Struct Algor 31(1):3–122

    Article  MATH  Google Scholar 

  • Bonanno G, Caldarelli G, Lillo F, Micciche S, Vandewalle N, Mantegna R (2004) Networks of equities in financial markets. Eur Phys J B 34:363–371

    Article  Google Scholar 

  • Breiman L (1968) Probability. Classics in applied mathematics, SIAM

  • Brown B (1983) Statistical uses of the spatial median. J R Stat Soc B 45:25–30

    MathSciNet  MATH  Google Scholar 

  • Bullmore E, Sporns O (2009) Complex brain networks: network theoretical analysis of structural and functional systems. Nat Rev Neurosci 10:186–196

    Article  Google Scholar 

  • Chatterjee S, Diaconis P (2013) Estimating and understanding exponential random graph models. Ann Statist 41:2428–2461

    Article  MathSciNet  MATH  Google Scholar 

  • Cuesta-Albertos J, Nieto-Reyes A (2008) The Tukey and the random Tukey depths characterize discrete distributions. J Multivar Anal 10:2304–2311

    Article  MathSciNet  MATH  Google Scholar 

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

    Article  MathSciNet  MATH  Google Scholar 

  • Cuesta-Albertos J, Fraiman R, Ransford T (2007) A sharp form of the Cramer–Wold theorem. J Theor Probab 20:201–209

    Article  MathSciNet  MATH  Google Scholar 

  • Dehling H, Wendler M (2010) Central limit theorem and the bootstrap for u-statistics of strongly mixing data. J Multivar Anal 101:126–137

    Article  MathSciNet  MATH  Google Scholar 

  • Devroye L, Fraiman N (2014) Connectivity of inhomogeneous random graphs. Random Struct Algor 45(3):408–420

    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 

  • Donges J, Petrova I, Loew A, Marwan N, Kurths J (2015) How complex climate networks complement eigen techniques for the statistical analysis of climatological data. Clim Dyn 45(9):2407–2424

    Article  Google Scholar 

  • Doukhan P, Neumann MH (2008) The notion of \(\psi \)-weak dependence and its applications to bootstrapping time series. Probab Surv 5:146–168

    Article  MathSciNet  MATH  Google Scholar 

  • Fraiman D (2008) Growing directed networks: stationary in-degree probability for arbitrary out-degree one. Eur Phys J B 61:377–388

    Article  MATH  Google Scholar 

  • Fraiman D, Balenzuela P, Foss J, Chialvo D (2009) Ising-like dynamics in large-scale functional brain networks. Phys Rev E 79:61922

    Article  Google Scholar 

  • Fraiman D, Saunier G, Martins E, Vargas C (2014) Biological motion coding in the brain: analysis of visually driven EEG functional networks. PloS ONE 9:e84612

    Article  Google Scholar 

  • Gao X, Xiao B, Tao D, Li X (2010) A survey of graph edit distance. Pattern Anal Appl 13:113–129

    Article  MathSciNet  Google Scholar 

  • Gauzere B, Brun L, Villemin D (2012) Two new graphs kernels in chemoinformatics. Pattern Recogn Lett 33:2038–2047

    Article  Google Scholar 

  • Gozolchiani A, Yamasaki K, Gazit O, Havlin S (2008) Pattern of climate network blinking links follows El Niño events. Europhys Lett 83:28005

    Article  Google Scholar 

  • Greco L, Farcomeni A (2016) A plug-in approach to sparse and robust principal component analysis. TEST 25:449–481

    Article  MathSciNet  Google Scholar 

  • Guigoures R, Boulle M, Rossi F (2015) Advances in data analysis and classification. Springer, New York

    Google Scholar 

  • Holme P (2015) Modern temporal network theory: a colloquium. Eur Phys J B 88:1–30

    Article  Google Scholar 

  • Jiang X, Münger A, Bunke H (2001) On median graphs: properties, algorithms, and applications. IEEE T Pattern Anal 23:1144–1151

    Article  Google Scholar 

  • Jo HH, Karsai M, Kertsz J, Kaski K (2012) Circadian pattern and burstiness in mobile phone communication. New J Phys 14:013055

    Article  Google Scholar 

  • Karrer B, Newman M (2011) Spectral methods for network community detection and network partitioning. Phys Rev E 83:8016107

    Article  Google Scholar 

  • Kolar M, Song L, Ahmed A, Xing E (2010) Estimating time-varying networks. Ann Appl Stat 4:94–123

    Article  MathSciNet  MATH  Google Scholar 

  • Kumar G, Garland M (2006) Visual exploration of complex time-varying graphs. IEEE T Vis Comput Gr 12:805–812

    Article  Google Scholar 

  • Kumar R, Novak J, Raghavan P, Tomkins A (2005) On the bursty evolution of blogspace. World Wide Web 8:159–178

    Article  Google Scholar 

  • Liu R (1988) On a notion of simplicial depth. Proc Natl Acad Sci USA 97:1732–1734

    Article  MathSciNet  MATH  Google Scholar 

  • Livi L, Rizzi A (2013) The graph matching problem. Pattern Anal Appl 16:253–283

    Article  MathSciNet  MATH  Google Scholar 

  • Mahalanobis P (1936) On the generalized distance in statistics. Proc Natl Inst Sci 2:49–55

    MATH  Google Scholar 

  • Mastrandrea R, Fournet J, Barrat A (2015) Contact patterns in a high school: a comparison between data collected using wearable sensors, contact diaries and friendship surveys. PloS ONE 10:e0136497

    Article  Google Scholar 

  • Micchelli C (1986) Interpolation of scattered data: distance matrices and conditionally positive definite functions. Constr Approx 2:11–22

    Article  MathSciNet  MATH  Google Scholar 

  • Mikosch T, Sorensen M (2002) Empirical process techniques for dependent data. Springer, New York

    MATH  Google Scholar 

  • Newman M (2006) Modularity and community structure in networks. Proc Natl Acad Sci USA 103(23):8577–8582

    Article  Google Scholar 

  • Newman M, Girvan M (2004) Finding and evaluating community structure in networks. Phys Rev E 69:026113

    Article  Google Scholar 

  • Peixoto TP (2015) Inferring the mesoscale structure of layered, edge-valued, and time-varying networks. Phys Rev E 92:042807

    Article  Google Scholar 

  • Peligrad M (1986) Recent advances in the central limit theorem and its weak invariance principle for mixing sequences of random variables (a survey). Dependence in probability and statistics, pp 193–224

  • Pignolet Y, Roy M, Schmid S, Tredan G (2015) Exploring the graph of graphs: network evolution and centrality distances. arXiv:1506.01565

  • Pollard D (1981) Strong consistency of k-means clustering. Ann Stat 9:135–140

    Article  MathSciNet  MATH  Google Scholar 

  • Robins G, Pattison P, Kalish Y, Lusher D (2007) An introduction to exponential random graph (p*) models for social networks. Soc Netw 29:173–191

    Article  Google Scholar 

  • Rohe K, Chatterjee S, Yu B (2011) Spectral clustering and the high-dimensional stochastic blockmodel. Ann Stat 39(4):1878–1915

    Article  MathSciNet  MATH  Google Scholar 

  • Rosenblatt M (1956) A central limit theorem and a strong mixing condition. Proc Natl Acad Sci USA 42:43–47

    Article  MathSciNet  MATH  Google Scholar 

  • Small C (1996) A survey of multidimensional medians. Int Stat Rev 58:263–277

    Article  Google Scholar 

  • Tang J, Scellato S, Musolesi M, Mascolo C, Latora V (2010) Small-world behavior in time-varying graphs. Phys Rev E 81:055101

    Article  Google Scholar 

  • Tsonis A, Swanson K (2008) Topology and predictability of El Niño and La Niña Networks. Phys Rev Lett 100:228502

    Article  Google Scholar 

  • Tukey J (1975) Mathematics and the picturing of data. In: Proceedings of the international congress of mathematicians, Vancouver, pp 523–531

  • Vardi Y, Zhang C (2000) The multivariate \(l_1\)-median and associated data depth. Proc Natl Acad Sci USA 97:1423–1426

    Article  MathSciNet  MATH  Google Scholar 

  • Vishwanathan SVN, Schraudolph NN, Kondor R, Borgwardt KM (2010) Graph kernels. J Mach Learn Res 11:1201–1242

    MathSciNet  MATH  Google Scholar 

  • Watts D, Strogatz S (1998) Collective dynamics of ’small-world’ networks. Nature 393:440–442

    Article  Google Scholar 

  • Xing E, Fu W, Song L (2010) A state-space mixed membership blockmodel for dynamic network tomography. Ann App Stat 4:535–566

    Article  MathSciNet  MATH  Google Scholar 

  • Xu K, Hero O (2013) Dynamic stochastic block models: statistical models for time-evolving networks. In: International conference on social computing, behavioral-cultural modeling, and prediction, vol 1, pp 201–210

  • Yang T, Chi Y, Zhu S, Gong Y, Jin R (2011) Detecting communities and their evolutions in dynamic social networks—a Bayesian approach. Mach Learn 82:157–189

    Article  MathSciNet  MATH  Google Scholar 

  • Zhao Y, Levina E, Zhu J (2011) Community extraction for social networks. Proc Natl Acad Sci USA 108(18):7321–7326

    Article  Google Scholar 

  • Zhao Y, Levina E, Zhu J (2012) Consistency of community detection in networks under degree-corrected stochastic block models. Ann Stat 40(4):2266–2292

    Article  MathSciNet  MATH  Google Scholar 

  • Zuo Y, Serfling R (2000) General notions of statistical depth function. Ann Stat 28(2):461–482

    Article  MathSciNet  MATH  Google Scholar 

Download references

Acknowledgements

The authors would like to thank two anonymous reviewers for helpful comments and criticism on earlier versions of the paper.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Daniel Fraiman.

Additional information

This article was produced as part of the activities of FAPESP Center for Neuromathematics (Grant#2013/07699-0, S.Paulo Research Foundation).

Appendix: Proofs

Appendix: Proofs

1.1 Characterization of the central set

Proof of Proposition 1

We have that the expected distance from a network H to a random network \({\mathbf {G}}\) is

$$\begin{aligned} {{\mathbb {E}}}\left( d({{\mathbf {G}}}, H) \right) = \sum _{G \in {\mathscr {G}}} d(G,H)p_G. \end{aligned}$$
(8)

Let A(G) be the adjacency matrix of the network G and \({{\mathbf {A}}}\) the adjacency matrix of the random network \({{\mathbf {G}}}\). Then expression (8) can be written as

$$\begin{aligned} \sum _{G \in {{\mathscr {G}}}} \sum _{i>j} \vert A(G)_{ij} - A(H)_{ij} \vert p_G&= \sum _{i>j} \sum _{G \in {{\mathscr {G}}}} \vert A(G)_{ij} - A(H)_{ij} \vert p_G \\&= \sum _{i>j}{{\mathbb {P}}}\left( {{\mathbf {A}}}_{ij} \ne A(H)_{ij} \right) , \end{aligned}$$

which is minimized by any network H with adjacency matrix \(A(H)_{ij}=1\) if and only if \({{\mathbb {P}}}\left( {{\mathbf {A}}}_{ij}=1 \right) \ge 1/2\). Moreover, if for all ij

$$\begin{aligned} {{\mathbb {P}}}\left( {{\mathbf {A}}}_{ij}=1 \right) \ne 1/2, \end{aligned}$$
(9)

there is a unique network S that minimizes expression (8) and the corresponding adjacency matrix satisfies \(A(S)_{ij}=1\) if and only if \({{\mathbb {P}}}\left( {{\mathbf {A}}}_{ij} =1 \right) > 1/2\).

On the other hand, if condition (9) does not hold, there are many solutions. The maximal center L is the network whose adjacency matrix fulfills \(A(L)_{ij} =1\) when \({{\mathbb {P}}}\left( {{\mathbf {A}}}_{ij}=1 \right) \ge 1/2\) and the set \({\mathscr {C}}\) contains exactly all subnetworks of L for which S is a subnetwork.

The proof for the empirical version is completely analogous. \(\square \)

1.2 Depth determines measure

Proof of Proposition 2

Let \(\mu = (\mu _1, \ldots , \mu _K)\) and \(\nu = (\nu _1, \ldots , \nu _K)\) be two distributions on \({{\mathscr {G}}}\) where \(K=2^m\) is the cardinal of the space of networks. For any \(H \in {{\mathscr {G}}}\), let \(d(H) = (d(H_1,H), \ldots d(H_{K},H))\). Then, the population depth is

$$\begin{aligned} D_\mu (H) = m - d(H)^T \mu . \end{aligned}$$

Therefore, the depth determines the measure if and only if

$$\begin{aligned} d(H)^T \mu - d(H)^T \nu = 0 \text { for all} H\in {{\mathscr {G}}}\text { implies } \mu = \nu . \end{aligned}$$
(10)

Denote by F the matrix with rows given by \(d(H_1), \ldots d(H_{K})\). Then expression (10) is equivalent to \(F (\mu -\nu ) = 0\), having a unique solution. The result follows from the invertibility of distance matrices. This result was initially proved by Micchelli (1986) and later on by Auer (1995), who provided an elementary proof, that the only uses the triangle inequality. In particular, our metric is just the \(L^1\) distance between the adjacency matrices and the result holds. For the sake of completeness, we now state the result in Auer (1995) for distance matrices.

Theorem 4

Let \(P_1, \ldots , P_n\) be distinct points in \({\mathbb {R}}^k\), and \(d_{i,j} = \Vert P_i - P_j \Vert \). If \(F_n\) is the distance matrix with entries \(d_{i,j}\), then

  1. (a)

    The \(\det F_n\) is positive if n is odd and negative if n is even; in particular, \(F_n\) is invertible.

  2. (b)

    The matrix \(F_n\) has one positive and \(n-1\) negative eigenvalues.

Applying Theorem A to our setup, we get that the matrix F is invertible. \(\square \)

1.3 Convergence of empirical depth

Proof of Theorem 2

From the ergodic theorem, we have that \({\hat{D}}_\ell (H) \rightarrow D(H)\) almost surely as \(\ell \rightarrow \infty \) for each \(H \in {\mathscr {G}}\). Since \({\mathscr {G}}\) is finite, we get uniform convergence. \(\square \)

Proof of Theorem 3

Recall that a sequence of random elements \({\mathbf {X}} := ({\mathbf {X}}_t, t \ge 1)\) is a strong mixing sequence if it fulfills the following condition. For \( 1 \le j < \ell \le \infty \), let \({\mathscr {F}}_j^\ell \) denote the \( \sigma \)-field of events generated by the random elements \(X_k,\ j \le k \le \ell \ (k \in \mathbf{N})\) . For any two \( \sigma \)-fields \( {\mathscr {A}}\) and \( {\mathscr {B}}\), define

$$\begin{aligned} \alpha ({\mathscr {A}}, {\mathscr {B}}):= \sup _{A \in {\mathscr {A}}, B \in {\mathscr {B}}} \vert {{\mathbb {P}}}\left( A \cap B \right) - {{\mathbb {P}}}\left( A \right) {{\mathbb {P}}}\left( B \right) \vert . \end{aligned}$$

For the given random sequence \({\mathbf {X}}\), for any positive integer n, define the dependence coefficient

$$\begin{aligned} \alpha (n) = \alpha ({\mathbf {X}},n) := \sup _{j \ge 1}\; \alpha ({\mathscr {F}}_{1}^j, {\mathscr {F}}_{j + n}^{\infty }). \end{aligned}$$

The random sequence \({\mathbf {X}}\) is said to be “strongly mixing,” or “\( \alpha \)-mixing,” if \( \alpha (n) \rightarrow 0\) as \( n \rightarrow \infty \). This condition was introduced by Rosenblatt (1956). By assumption, we have that the sequence of random networks \(\{{{\mathbf {G}}}_t: t\ge 1 \}\) is a strongly mixing sequence. In order to prove the theorem, we use the following result (see, for instance, Peligrad 1986).

Theorem 5

Let \(\{{\mathbf {X}}_t: t \ge 1\}\) be a strictly stationary centered \(\alpha \)-mixing sequence, and let \({\mathbf {S}}_\ell = \sum _{t=1}^\ell {\mathbf {X}}_t\). Assume that for some \(C>0\)

$$\begin{aligned} \vert {\mathbf {X}}_1 \vert< C \ {\textit{almost surely, and}}\; \sum _{n=1}^\infty \alpha (n) < \infty . \end{aligned}$$

Then,

$$\begin{aligned} \sigma ^2 := {{\mathbb {E}}}\left( {\mathbf {X}}_1^2 \right) + 2 \sum _{k=2}^\infty {{\mathbb {E}}}\left( {\mathbf {X}}_1 {\mathbf {X}}_k \right) , \end{aligned}$$

is absolutely summable. If in addition \(\sigma ^2 >0\), then \({\mathbf {S}}_\ell /\sqrt{\ell }\sigma \) converges weakly to a standard normal distribution.

First observe that

$$\begin{aligned} \beta ^T {\mathbf {Z}}_{\ell } = \frac{1}{\ell } \sum _{k=1}^{\ell } {\mathbf {W}}_k, \quad \text {with}\quad {\mathbf {W}}_k= \sum _{j=1}^{2^m} \beta _j \left( d(H_j, {{\mathbf {G}}}_k) - {{\mathbb {E}}}\left( d(H_j,{{\mathbf {G}}}_1) \right) \right) , \end{aligned}$$

where \(\{{\mathbf {W}}_t: t \ge 1 \}\) is a strictly stationary, bounded, centered \(\alpha \)-mixing sequence, fulfilling \(\sum _{n=1}^\infty \alpha (n) < \infty \). On the other hand, we have that

$$\begin{aligned} {{\mathbb {E}}}\left( {\mathbf {W}}_1^2 \right) = \beta ^T {{\mathbb {E}}}\left( {\mathbf {Y}}_1^T {\mathbf {Y}}_1 \right) \beta \ \text{ and } \ {{\mathbb {E}}}\left( {\mathbf {W}}_1 {\mathbf {W}}_k \right) = \beta ^T {{\mathbb {E}}}\left( {\mathbf {Y}}_1^T {\mathbf {Y}}_k \right) \beta , \end{aligned}$$

and the result follows from Theorem B. \(\square \)

1.4 Characterization of principal components

Proof of Proposition 3

Note that

$$\begin{aligned}&{\text {var}}\left( \frac{|{{\mathbf {G}}}\wedge Q|}{|Q|} \right) \\&\quad = \frac{1}{\vert Q \vert ^2}\sum _{G \in {\mathscr {G}}} \left( \sum _{i>j} c_{ij} A(G)_{ij} A(Q)_{ij} - \sum _{H \in {\mathscr {G}}} \sum _{i>j} c_{ij} A(H)_{ij} A(Q)_{ij}p_H \right) ^2p_G. \end{aligned}$$

We first consider the case when the network Q has only one link \((k,\ell )\), and find within this family the one that maximizes the objective function. Next we prove that for any other network Q the objective function is bounded by the maximum restricted to the former family. Finally, we show that the principal component space is generated by the one link networks.

Let \(Q_1\) such that \(A(Q_1)_{k_1\ell _1}=1\), and 0 otherwise. Also, let \({\mathscr {G}}_{Q_1}^+ = \{G\in {\mathscr {G}}: A(G)_{k_1\ell _1} = 1\}\) and \({\mathscr {G}}_{Q_1}^- = \{G\in {\mathscr {G}}: A(G)_{k_1\ell _1} = 0 \}\). When we search within the one link networks, the objective function reduces to

$$\begin{aligned} {\text {var}}\left( \frac{|{{\mathbf {G}}}\wedge Q|}{|Q|} \right)&= \frac{1}{c_{k_1\ell _1}^2}\sum _{G \in {\mathscr {G}}} \left( c_{k_1\ell _1} A(G)_{k_1\ell _1} - \sum _{H \in {\mathscr {G}}} c_{k_1\ell _1} A(H)_{k_1\ell _1} p_H \right) ^2 p_G \\&=\sum _{G \in {\mathscr {G}}} \left( A(G)_{k_1\ell _1} -{{\mathbb {P}}}\left( {{\mathbf {A}}}_{k_1\ell _1}=1 \right) \right) ^2 p_G \\&=\sum _{G \in {\mathscr {G}}^+} \left( 1- {{\mathbb {P}}}\left( {{\mathbf {A}}}_{k_1\ell _1}=1 \right) \right) ^2 p_G + \sum _{G \in {\mathscr {G}}^-} {{\mathbb {P}}}\left( {{\mathbf {A}}}_{k_1\ell _1}=1 \right) ^2 p_G \\&= ( 1- {{\mathbb {P}}}\left( {{\mathbf {A}}}_{k_1\ell _1}=1 \right) )^2 {{\mathbb {P}}}\left( {{\mathbf {A}}}_{k_1\ell _1}=1 \right) + {{\mathbb {P}}}\left( {{\mathbf {A}}}_{k_1\ell _1}=1 \right) ^2 {{\mathbb {P}}}\left( {{\mathbf {A}}}_{k_1\ell _1}=0 \right) \\&= {{\mathbb {P}}}\left( {{\mathbf {A}}}_{k_1\ell _1}=1 \right) (1- {{\mathbb {P}}}\left( {{\mathbf {A}}}_{k_1\ell _1}=1 \right) ), \end{aligned}$$

and the solution is the one link graph for which \({{\mathbb {P}}}\left( {{\mathbf {A}}}_{k_1\ell _1}=1 \right) \) is closest to 1 / 2.

In the general case, we want to find Q that maximizes

$$\begin{aligned} {\text {var}}\left( \frac{|{{\mathbf {G}}}\wedge Q|}{|Q|} \right)&= \frac{1}{\vert Q \vert ^2} \sum _{G \in {\mathscr {G}}} \left( \sum _{(i,j)\in Q} c_{ij} A(G)_{ij} - \sum _{H \in {\mathscr {G}}} \sum _{(i,j)\in Q} c_{ij} A(H)_{ij} p_H \right) ^2 p_G \\&= \sum _{G \in {\mathscr {G}}} \ \sum _{(i,j)\in Q} w_{ij} \left( A(G)_{ij} - {{\mathbb {P}}}\left( {{\mathbf {A}}}_{ij}=1 \right) \right) ^2 p_G, \end{aligned}$$

where \(w_{ij}= c_{ij} / \sum _{(p,q)\in Q} c_{pq}\). Now, since the weights \(w_{ij}\) add to one, we have

$$\begin{aligned} {\text {var}}\left( \frac{|{{\mathbf {G}}}\wedge Q|}{|Q|} \right)&\le \sum _{(i,j)\in Q} w_{ij}\; \max _{(i,j)}\; {{\mathbb {P}}}\left( {{\mathbf {A}}}_{ij}=1 \right) \left( 1 - {{\mathbb {P}}}\left( {{\mathbf {A}}}_{ij}=1 \right) \right) \\&= \max _{(i,j)}\; {{\mathbb {P}}}\left( {{\mathbf {A}}}_{ij}=1 \right) \left( 1-{{\mathbb {P}}}\left( {{\mathbf {A}}}_{ij}=1 \right) \right) , \end{aligned}$$

which corresponds to the one link optimum.

If there exists a unique one link graph (\(Q_1\)) that verifies \({{\mathbb {P}}}\left( {{\mathbf {A}}}_{k_1\ell _1}=1 \right) \) is closest to 1 / 2, then the principal component space is generated just by \(Q_1\), i.e., \({{\mathscr {S}}}_1 ={\mathscr {G}}_{Q_1}^+\). If there exist multiple one link graphs, \(Q_1, Q_2, \ldots , Q_p\), that minimize \(|{{\mathbb {P}}}\left( {{\mathbf {A}}}_{k\ell }=1 \right) -1/2|\), then the principal component space is \({{\mathscr {S}}}_1 = \cup _{i=1}^p {\mathscr {G}}_{Q_i}^+\). The second principal component space verifies the same. In this case, the maximization of the variance is over \(\{G\in {\mathscr {G}}: G \notin {{\mathscr {S}}}_1 \}\). Analogous for the rest of the components, for example, for finding the \(k-esima\) principal component space just maximizes the variance over \(\{G\in {\mathscr {G}}: G \notin {{\mathscr {S}}}_1, G \notin {{\mathscr {S}}}_2, G \notin {{\mathscr {S}}}_{k-1} \}.\) \(\square \)

1.5 Consistency of principal components

Proof of Proposition 4

For each \(Q \in {\mathscr {G}}\) from the ergodic theorem, we have the following.

  1. (a)

    \(\displaystyle \varLambda _{\ell } (Q) = \frac{1}{\ell } \sum _{k=1}^{\ell } \vert G_k \wedge Q \vert \rightarrow {{\mathbb {E}}}\left( \vert {{\mathbf {G}}}\wedge Q\vert \right) \), almost surely as \(\ell \rightarrow \infty \).

  2. (b)

    which converges almost surely to

    $$\begin{aligned} {{\mathbb {E}}}\left( \frac{\vert {{\mathbf {G}}}\wedge Q \vert ^2}{\vert Q \vert ^2} \right) + {{\mathbb {E}}}\left( \frac{\vert {{\mathbf {G}}}\wedge Q \vert }{\vert Q \vert ^2} \right) ^2 - 2 {{\mathbb {E}}}\left( \frac{\vert {{\mathbf {G}}}\wedge Q \vert }{\vert Q \vert ^2} \right) ^2 = {\text {var}}\left( \frac{\vert {{\mathbf {G}}}\wedge Q \vert }{\vert Q \vert }\right) . \end{aligned}$$
    (11)
  3. (c)

    Since the space \({\mathscr {G}}\) is finite, expression (11) entails that \(\hat{{\mathscr {Q}}}_1 \rightarrow {\mathscr {Q}}_1\) almost surely, i.e., \( \hat{{\mathscr {Q}}}_1 = {\mathscr {Q}}_1\) for \(\ell \) large enough almost surely, which entails that the principal components converge because the geodesics coincide eventually.

For the next principal component, the proof is analogous. \(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Fraiman, D., Fraiman, N. & Fraiman, R. Nonparametric statistics of dynamic networks with distinguishable nodes. TEST 26, 546–573 (2017). https://doi.org/10.1007/s11749-017-0524-8

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11749-017-0524-8

Keywords

Mathematics Subject Classification

Navigation