Quantifying Statistical Interdependence PART III: N ... · Quantifying Statistical Interdependence...

48
Quantifying Statistical Interdependence PART III: N> 2 Point Processes J. Dauwels a,,1 T. Weber b F. Vialatte c,d T. Musha e A. Cichocki c a School of Electrical and Electronic Engineering (EEE), Nanyang Technological University (NTU), Singapore. b Operations Research Center, Massachusetts Institute of Technology, Cambridge, MA. c Laboratory for Advanced Brain Signal Processing, RIKEN Brain Science Institute, Saitama, Japan. d Laboratoire SIGMA, ESPCI ParisTech, Paris, France. e Brain Functions Laboratory, Inc., Yokohama, Japan. Abstract Stochastic event synchrony (SES) is a recently proposed family of similarity mea- sures. First, “events” are extracted from the given signals; next, one tries to align events across the different time series. The better the alignment, the more simi- lar the N time series are considered to be. The similarity measures quantify the reliability of the events (fraction of “non-aligned” events) and the timing precision. So far, SES has been developed for pairs of one-dimensional (Part I) and multi- dimensional (Part II) point processes. In this paper (Part III), SES is extended from pairs of signals to N> 2 signals. The alignment and SES parameters are again determined through statistical in- ference, more specifically, by alternating the following two steps: (i) one estimates the SES parameters from a given alignment; (ii) with the resulting estimates, one refines the alignment. The SES parameters are computed by maximum a posteriori (MAP) estimation (Step 1), in analogy to the pairwise case. The alignment (Step 2) is solved by linear integer programming. In order to test the robustness and reliability of the proposed N -variate SES method, it is first applied to synthetic data. We show that N -variate SES results in more reliable estimates than bivariate SES. Next N -variate SES is applied to two problems in neuroscience: it used to quantify the firing reliability of Morris-Lecar neurons, and to detect anomalies in EEG synchrony of Mild Cognitive Impairment (MCI) patients; those problems were also considered in Part I and II respectively. In both cases, the N -variate SES approach yields a more detailed analysis. Key words: timing precision, event reliability, stochastic event synchrony, maximum a posteriori estimation, Morris-Lecar neuron, firing reliability, Preprint submitted to Elsevier 20 August 2011

Transcript of Quantifying Statistical Interdependence PART III: N ... · Quantifying Statistical Interdependence...

Page 1: Quantifying Statistical Interdependence PART III: N ... · Quantifying Statistical Interdependence PART III: N > 2 Point Processes J. Dauwelsa,∗,1T. WeberbF. Vialattec,dT. MushaeA.

Quantifying Statistical Interdependence

PART III: N > 2 Point Processes

J. Dauwels a,∗,1T. Weber bF. Vialatte c,dT. Musha eA. Cichocki c

aSchool of Electrical and Electronic Engineering (EEE), Nanyang TechnologicalUniversity (NTU), Singapore.

bOperations Research Center, Massachusetts Institute of Technology, Cambridge,MA.

cLaboratory for Advanced Brain Signal Processing, RIKEN Brain ScienceInstitute, Saitama, Japan.

dLaboratoire SIGMA, ESPCI ParisTech, Paris, France.

eBrain Functions Laboratory, Inc., Yokohama, Japan.

Abstract

Stochastic event synchrony (SES) is a recently proposed family of similarity mea-sures. First, “events” are extracted from the given signals; next, one tries to alignevents across the different time series. The better the alignment, the more simi-lar the N time series are considered to be. The similarity measures quantify thereliability of the events (fraction of “non-aligned” events) and the timing precision.

So far, SES has been developed for pairs of one-dimensional (Part I) and multi-dimensional (Part II) point processes. In this paper (Part III), SES is extended frompairs of signals to N > 2 signals.

The alignment and SES parameters are again determined through statistical in-ference, more specifically, by alternating the following two steps: (i) one estimatesthe SES parameters from a given alignment; (ii) with the resulting estimates, onerefines the alignment. The SES parameters are computed by maximum a posteriori(MAP) estimation (Step 1), in analogy to the pairwise case. The alignment (Step2) is solved by linear integer programming.

In order to test the robustness and reliability of the proposed N -variate SESmethod, it is first applied to synthetic data. We show that N -variate SES results inmore reliable estimates than bivariate SES. Next N -variate SES is applied to twoproblems in neuroscience: it used to quantify the firing reliability of Morris-Lecarneurons, and to detect anomalies in EEG synchrony of Mild Cognitive Impairment(MCI) patients; those problems were also considered in Part I and II respectively.In both cases, the N -variate SES approach yields a more detailed analysis.

Key words: timing precision, event reliability, stochastic event synchrony,maximum a posteriori estimation, Morris-Lecar neuron, firing reliability,

Preprint submitted to Elsevier 20 August 2011

Page 2: Quantifying Statistical Interdependence PART III: N ... · Quantifying Statistical Interdependence PART III: N > 2 Point Processes J. Dauwelsa,∗,1T. WeberbF. Vialattec,dT. MushaeA.

time-frequency map, bump model, EEG, Alzheimer’s disease, mild cognitiveimpairment

1 Introduction

Neural synchrony may play an important role in information processing inthe brain. Although the details of this coding mechanism have not been fullyrevealed, it has been postulated that neural synchrony may be involved incognition (Varela et al., 2001) and even in consciousness (Singer, 2001). Thecorrelation between different brain signals has been studied intensively in re-cent years both by experimental neuroscientists (e.g., (Abeles et al., 1993))and computational neuroscientists (e.g., (Amari et al., 2003)), and also byneurologists; indeed, various medical studies have reported that neurologicaldiseases (such as Alzheimer’s disease and epilepsy) are related to perturbationsin neural synchrony (Matsuda et al., 2001; Jeong, 2004).

Motivated by the intensified interest in neural synchrony, numerous researchershave in the last years developed and refined methods to quantify the synchronybetween signals (see, e.g., (Stam, 2005; Quiroga et al., 2002; Pereda et al.,2005; Toups et al., 2011)). In recent work (Dauwels et al., 2007, 2009a,b), wehave proposed a new family of synchrony measures referred to as stochasticevent synchrony (SES); this class of synchrony measures is inspired by theVictor-Purpura distance metrics (Victor et al., 1997). The basic idea is thefollowing: First, we extract “events” from the given time series; next, we tryto align events from one time series with events from the other. The betterthe alignment, the more similar the time series are considered to be. We alsoquantify the timing jitter between matched (“coincident”) events; the smallerthe timing jitter, the larger the synchrony. SES thus considers two differentaspects of synchrony: reliability and timing precision. Those concepts werealso recently considered in (Toups et al., 2011), and they can be understoodfrom the following analogy: when you wait for a train in the station, the trainmay come at the station or it may not come at all; for example, it may be outof service due to some mechanical problem. If the train comes, it may or may

∗ Corresponding author.Email addresses: [email protected] (J. Dauwels), theo [email protected] (T.

Weber), [email protected] (F. Vialatte), [email protected] (T.Musha), [email protected] (A. Cichocki).1 Part of this work was carried out while J.D. was at the Laboratory for Informa-tion and Decision Systems (LIDS), Massachusetts Institute of Technology (MIT),Cambridge, MA.

2

Page 3: Quantifying Statistical Interdependence PART III: N ... · Quantifying Statistical Interdependence PART III: N > 2 Point Processes J. Dauwelsa,∗,1T. WeberbF. Vialattec,dT. MushaeA.

not be on time. The former uncertainty is related to reliability, whereas thelatter is related to precision.

So far, SES was restricted to pairs of signals. In this paper (Part III), weextend SES from pairs of signals to N > 2 signals. The underlying princi-ple is similar, but the inference algorithm to compute the SES parameters isfundamentally different. In bivariate SES (Dauwels et al., 2007, 2009a,b) weapplied the max-product algorithm to align pairs of sequences. We have im-plemented the max-product algorithm (and various refinements) for aligningN > 2 sequences as well; however, that approach unfortunately leads to inac-curate alignments, probably because N -wise alignment is substantially harderthan pairwise alignment. As an alternative, we solve the N -wise alignment byinteger linear programming, which yields optimal or near-optimal alignmentsat reasonable computational cost.

The extension from pairs of signals to N > 2 signals is non-trivial. In thefollowing, we briefly touch upon this issue. For N = 2, the problem of timeseries comparison is essentially that of finding an alignment between the timeseries. The “alignment” approach of Victor & Purpura (Victor et al., 1997)was the inspiration for bivariate SES; we have shown in (Dauwels et al., 2007,2009a,b) that the alignment between 2 time series (say x1 and x2) is the sameas finding an underlying time series v which can then be transformed to x1

and x2 by a sequence of steps involving jittering the events, and insertions anddeletions. Quantifying the distance between x1 and x2 is equivalent to findinga v that minimizes some combination of d(x1, v) and d(x2, v). By making thecombination rule for these two distances accelerating (such as a root-mean-square), one can ensure that v is in the “middle” of x1 and x2. Comparingx1 with x2 is equivalent to finding a hidden “consensus” process v that cangenerate both x1 with x2. Finding the v “interprets” the similarity of x1 and x2

in terms of a consensus that they both represent, but is not necessary to findv to quantify this similarity, since it can also be expressed as an alignment,without an explicit v.

For N = 3, one can attempt to jointly minimize some combination of d(x1, v),d(x2, v), and d(x3, v), without attempting to find an underlying v. Or, one canattempt to find a single v for which some combination of d(x1, v), d(x2, v),and d(x3, v) are smallest.

Finally, for N > 3, there is at least one other possibility: the point processmight be generated by two or more hidden processes. This can first happenat N = 4, with two hidden processes, say v and w, for which the distancesd(x1, v), d(x2, v), d(x3, w), d(x4, w) have a lower total than the distance fromany single v to all four of the observations x1, x2, x3, and x4. In other words,for N > 3, the problem of finding multivariate similarity is generally not thesame as finding the single underlying “consensus” process.

3

Page 4: Quantifying Statistical Interdependence PART III: N ... · Quantifying Statistical Interdependence PART III: N > 2 Point Processes J. Dauwelsa,∗,1T. WeberbF. Vialattec,dT. MushaeA.

In summary, there are at least 3 kinds of definitions of synchrony and similar-ity:

(1) a definition based on the N(N −1)/2 distances d(xi, xj) between pairs ofobserved processes,

(2) a definition based on finding a consensus process v, which minimizes theN distances d(xi, v),

(3) a definition based on finding multiple hidden processes.

For N = 2, these definitions are identical. For N = 3, the second and third areidentical, but differ from the first. For N > 3, all three definitions are distinct.In this paper, we consider the second definition: the N point processes aregenerated from a (hidden) consensus process v. In future work, we will extendN -variate SES to multiple hidden processes.

Stochastic event synchrony (SES) is applicable to any kind of time series(e.g., from finance, oceanography, and seismology). We will consider here neu-ral spike trains and electroencephalograms (EEG). More specifically, we willuse the SES method to quantify the reliability of Morris-Lecar neurons, andto predict mild cognitive impairment (MCI) from electroencephalograms; wealso considered those problems in Part I (Dauwels et al., 2009a) and PartII (Dauwels et al., 2009b) respectively.

This paper is organized as follows. In the next section, we explain how SEScan be extended from pairs of point processes to N > 2 point processes. Wedescribe the underlying statistical model (Section 3), and outline our inferencemethod 2 (Section 4). We consider various extensions of our statistical model(Section 5). We investigate the robustness and reliability of the SES inferencemethod by means of synthetic data (Section 6). We use SES to quantify thereliability of Morris-Lecar neurons (Section 7), and to detect abnormalities inthe EEG synchrony of MCI disease patients (Section 8). At last, we offer someconcluding remarks (Section 9).

Readers who are less interested in the technical details may wish to readSection 2, where the general idea is outlined, and Sections 7 and 8, where twoapplications are discussed; those sections can be read independently of themore technical Sections 3 and 4.

2 An implementation of N -variate SES for one-dimensionaland multi-dimensional point processes is available fromhttp://www.dauwels.com/SESToolbox/SES.html.

4

Page 5: Quantifying Statistical Interdependence PART III: N ... · Quantifying Statistical Interdependence PART III: N > 2 Point Processes J. Dauwelsa,∗,1T. WeberbF. Vialattec,dT. MushaeA.

2 Principle

Suppose that we are given N > 2 continuous-time signals (e.g., electroen-cephalogram (EEG) signals recorded from different channels), and we wish todetermine the similarity of those signals. In (Dauwels et al., 2009b) we con-sidered pairs of signals (N = 2); although the extension to N > 2 may seemstraightforward, it leads to a combinatorial problem that is much harder. Inthe following, we will closely follow the setting and notation of (Dauwels etal., 2009b).

As a first step, we extract point processes from those N > 2 signals, whichmay be achieved in various ways. As an example, we generate point processesin time-frequency domain: first the time-frequency (“wavelet”) transform ofeach signal is computed in a frequency band f ∈ [fmin, fmax]. Next those mapsare approximated as a sum of half-ellipsoid basis functions, referred to as“bumps” (see Fig. 1 and (Vialatte et al., 2007)). Each bump is described byfive parameters: time t, frequency f , width ∆t, height ∆f , and amplitude w.The resulting bump models represent the most prominent oscillatory activ-ity in the signals at hand. This activity may correspond to various physicalor biological phenomena. For example, oscillatory events in EEG and otherbrain signals are believed to occur when assemblies of neurons are spikingin synchrony (Buzsaki, 2006; Nunez et al., 2006). In the following, we willdevelop N -variate SES for bump models. In this setting, SES quantifies thesynchronous interplay between oscillatory patterns in N > 2 given signals,while it ignores the other components in those signals (“background activ-ity”). In contrast, classical synchrony measures such as amplitude or phasesynchrony are computed from the entire signal, they make no distinction be-tween oscillatory components and the background activity. As a consequence,SES captures alternative aspects of similarity, and hence, it provides comple-mentary information about synchrony.

Besides bump models, SES may be applied to other sparse representations ofsignals. Moreover, the point processes may be defined in other spaces than thetime-frequency plane, for example, they may occur in two-dimensional space(e.g., images), space-frequency (e.g., wavelet image coding) or space-time (e.g.,movies). Such extensions may straightforwardly be derived from the exampleof bump models; we refer to Section 5 and (Dauwels et al., 2009b) for moredetails.

It is also noteworthy that the events in the point processes may be labeledby discrete tags. For example, in multineuronal recordings, the labels wouldcorrespond to the neurons of origin. More generally, the labels may correspondto the sites or processes of origin. SES may then be applied to the point pro-cesses associated with each label, and it would quantify the coupling between

5

Page 6: Quantifying Statistical Interdependence PART III: N ... · Quantifying Statistical Interdependence PART III: N > 2 Point Processes J. Dauwelsa,∗,1T. WeberbF. Vialattec,dT. MushaeA.

t

tt

f

ff

(a) Time-frequency maps.

t

t

t

ff

f

(b) Bump models.

Fig. 1. Similarity of three EEG signals (N = 3); from their time-frequency trans-forms (top), one extracts two-dimensional point processes (“bump models”; bot-tom), which are then aligned.

the different sites (e.g., neurons).

We now consider the central question: how can we quantify the similarityof N > 2 point processes defined on a space S? Let us consider the exam-ple of bump models (see Fig. 1 and Fig. 2). Intuitively speaking, N bumpmodels (xi)i=1,2,...,N may be considered well-synchronized if bumps appear in(almost) all bump models simultaneously, apart from a constant offset in timeand frequency, and a “small” amount of jitter in time and frequency. If oneoverlays N well-synchronized bump models, and removes the potential av-erage offsets in time and frequency (denoted by δti and δfi respectively, fori = 1, 2, . . . , N), bumps naturally appear in clusters that contain precisely one

6

Page 7: Quantifying Statistical Interdependence PART III: N ... · Quantifying Statistical Interdependence PART III: N > 2 Point Processes J. Dauwelsa,∗,1T. WeberbF. Vialattec,dT. MushaeA.

t

f

1

4

2

5

3

6

7

Fig. 2. Five bump models overlayed on top of each other (N = 5); the dashed boxesindicate clusters. The average offset between the bumps is close to zero.

bump from all (or almost all) bump models, as illustrated in Fig. 2 for N = 5.In that example, cluster 1 and 6 contain bumps from all 5 bump models xi,cluster 2 and 7 contain bumps from 3 bump models, cluster 3 and 5 containbumps from 2 bump models, and cluster 4 contains bumps from 4 bump mod-els. In general, the average offset in time and frequency between the bumps isnot necessarily zero (as illustrated in Fig. 3), and it may be harder to recognizethe different clusters. The algorithm developed in this paper is able to extractsuch clusters, even in the general case of non-zero average offsets in time andfrequency (as in Fig. 3).

If the point processes are well-synchronized, almost all clusters contain (closeto) N bumps, specifically, one bump from each (or almost each) of the N bumpmodels. Therefore, an important similarity statistics is the average number ofevents per cluster, or more generally, the statistical distribution of the num-ber of events per cluster. Moreover, as in the pairwise case (Dauwels et al.,2009a,b), one can quantify how well the bumps are aligned within each clus-ter, by computing the jitter sti and sfi in time and frequency respectively, fori = 1, 2, . . . , N .

More generally, N point processes on a space S may be considered similar if

7

Page 8: Quantifying Statistical Interdependence PART III: N ... · Quantifying Statistical Interdependence PART III: N > 2 Point Processes J. Dauwelsa,∗,1T. WeberbF. Vialattec,dT. MushaeA.

t

f

14

2

5

3

6

7

Fig. 3. Five bump models overlayed on top of each other (N = 5); the dashed boxesindicate clusters. The average offset between the bumps is non-zero. For example,the red and brown bumps tend to be located at the bottom and top respectively ofthe clusters, whereas the green bumps tend to lie at the right side of the clusters.Such tendencies cannot be observed in Fig. 2.

events appear in clusters with (close to) N events and with “small” dispersion(computed by the distance measure on S); those clusters may only appearafter certain transformations have been applied, e.g., translation (to eliminateoffsets as in Fig. 3), rotation, and scaling. In other words, N point processesare considered similar, if they can be transformed into each other by a fewoperations, including deletions and insertions, “small” random perturbations,and transformations such as translation, rotation, and scaling.

Let us now return to bump models. We determine the SES parameters δti,δfi, sti and sfi (i = 1, 2, . . . , N), and the event clusters by statistical inference,along the lines of the pairwise case (Dauwels et al., 2009a,b). We start by con-structing a statistical model that captures the relation between the N bumpmodels; that statistical model contains the SES parameters, besides variablesrelated to the alignment of the different bumps. Next we perform inferencein that statistical model, resulting in estimates for the SES parameters andclusters. More concretely, we apply cyclic maximization, as in the pairwisecase.

8

Page 9: Quantifying Statistical Interdependence PART III: N ... · Quantifying Statistical Interdependence PART III: N > 2 Point Processes J. Dauwelsa,∗,1T. WeberbF. Vialattec,dT. MushaeA.

In the following section, we outline our statistical model; in Section 4 wedescribe how we conduct inference in that statistical model.

3 Statistical Model

3.1 Casual Description

The intuitive concept of similarity outlined in Section 2 may readily be trans-lated into a generative stochastic model. In that model, the N point processesxi are treated as independent noisy observations of a hidden process v. Theobserved sequences (xi)i=1,...,N are obtained from v by the following four-stepprocedure:

(1) COPY: generate N copies of the hidden point process v,(2) DELETION: delete some of the copied events,(3) PERTURBATION: shift the remaining copies over (δti, δfi)i=1,...,N , and

randomly perturb the positions, with variance (si)i=1,...,N = (sti, sfi)i=1,...,N ,amounting to the N point processes (xi)i=1,...,N .

(4) INSERTION: Additional events are inserted (“background events”), whichare unrelated to the hidden point process v, and are modeled as mutuallyindependent.

As a result, each sequence xi consists of “noisy” copies of hidden events (gen-erated by Step 1 to 3), besides background events (generated by Step 4). Thenoisy copies are related to each other through the hidden events v, whereasthe background events are all independent of each other. The point processesxi may be considered well-synchronized if there are only few deletions (cf. Step2) and insertions (cf. Step 4), and if the events of xi are “close” to thecorresponding hidden events (cf. Step 3), apart from offsets (δti, δfi)i=1,...,N .Fig. 4 illustrates a generative process that results in the bump models ofFig. 3. More generally, as we pointed out in the previous section, one may in-clude other transformations in the perturbation step besides translation over(δti, δfi)i=1,...,N , such as rotation and scaling.

Some readers may wonder why insertions need to be modeled explicitly. In-deed, inserting an event is equivalent to adding a hidden event with N “noisy”copies, followed by N − 1 deletions. In this way, the SES models for pairs ofpoint processes (Part I and II) are able to capture insertions, even thoughthey are not modeled explicitly (Dauwels et al., 2009a,b). However, for largeN (e.g., N > 10), the cost of an insertion becomes prohibitively large (due tothe N−1 deletions), and as a consequence, the statistical model does no longercapture insertions. The inserted event will be grouped with other events, lead-

9

Page 10: Quantifying Statistical Interdependence PART III: N ... · Quantifying Statistical Interdependence PART III: N > 2 Point Processes J. Dauwelsa,∗,1T. WeberbF. Vialattec,dT. MushaeA.

t

f

14

2

5

3

6

7

δ1

δ2

δ3δ4

δ5

Fig. 4. Generative model for the N = 5 bump models (xi)i=1,...,N of Fig. 3. One firstgenerates a hidden bump model v, indicated in dashed lines. Next one makes N = 5identical copies of v and shifts those over (δi)i=1,...,N = (δti, δfi)i=1,...,N , as indicatedby the colored arrows in cluster 1. The resulting events are then slightly shifted,with variance (si)i=1,...,N = (sti, sfi)i=1,...,N , as indicated by the black arrows incluster 1. Finally, some of those events are deleted (with probability pd), resultingin the bump models (xi)i=1,...,N . For example, 2 events are deleted in cluster 2.

ing to incorrect clusters. Therefore, it becomes necessary to model insertionsexplicitly. As an illustration, Fig. 11(a) shows several event clusters in additionto background events (indicated by red hexagons). More information on thisapplication can be found in Section 7. Note that insertions are also explicitlymodeled in the Victor-Purpura distance metrics (Victor et al., 1997), whichwere the source of inspiration for SES.

3.2 Formal Description

We now describe the underlying stochastic model in more detail. We referto Table 1 for a summary of all relevant variables and parameters. For conve-nience, we introduce the following notation. The length of the point process xi

is denoted by Li (with i = 1, 2, . . . , N). The individual events of point processxi are denoted by xij (with i = 1, 2, . . . , N and j = 1, . . . , Li). The occurrence

10

Page 11: Quantifying Statistical Interdependence PART III: N ... · Quantifying Statistical Interdependence PART III: N > 2 Point Processes J. Dauwelsa,∗,1T. WeberbF. Vialattec,dT. MushaeA.

Symbol Explanation

(xi)i=1,2,...,N N given bump models

Li length of xi

Ltot total number of bumps in the N models (xi)i=1,2,...,N

xij j-th bump in bump model xi

tij occurrence time of xij

fij frequency of xij

δti and δfi average offset in time and frequency for bump model xi

sti and sfi jitter in time and frequency for bump model xi

pd deletion probability

v hidden bump model from which the observed bump models (xi)i=1,2,...,N

are generated

ℓ length of v

cij cij = 0 if xij is a background event, otherwise index of event in v that generated xij

Ck set of nk copies of vk (cf. (13))

Ik index set of Ck (cf. (13))

K index set of bump clusters of size nk > 1 (cf. (12))

L number of bump clusters, i.e., number of hidden events vk

with at least one copy (nk > 0)

ρ fraction of missing bumps in the clusters

v background events in (xi)i=1,2,...,N

ℓ length of v

χ fraction of background events in (xi)i=1,2,...,N

d0 cost associated with a cluster/exemplar (cf. (25))

d0 cost associated with a background event (cf. (26))

d(tij , fij) cost associated with an event xij that belongs to a non-trivial cluster (cf. (27))

d(tij , fij, ti′j′, fi′j′) cost associated with an event xij that belongs to exemplar xi′j′ (cf. (48))

bk equal to one iff cluster k is non-empty (nk > 0 and k = 1, 2, . . . , Ltot), otherwise zero

bijk equal to one iff the xij belongs to cluster k, i.e., cij = k, otherwise zero

bij equal to one iff xij is an exemplar, otherwise zero

biji′j′ equal to one iff xij is associated with exemplar xi′j′ , otherwise zero

eij equal to one iff xij is a background event, otherwise zero

Table 1List of variables and parameters associated with models p(x, c, v, θ, ℓ) (16),p(x, c, θ) (23) (31), and p(x, b, e, θ) (43) (49).

time and frequency of those events are referred to as tij and fij respectively.Moreover, we will use the notation δi = (δti, δfi), si = (sti, sfi), θi = (δi, si)(with i = 1, 2, . . . , N), and θ = (θ1, θ2, . . . , θN).

11

Page 12: Quantifying Statistical Interdependence PART III: N ... · Quantifying Statistical Interdependence PART III: N > 2 Point Processes J. Dauwelsa,∗,1T. WeberbF. Vialattec,dT. MushaeA.

The hidden process v = {v1, . . . , vℓ}, which is the source of all events inx1, x2, . . . xN (besides the background events; cf. Step 4), is modeled as fol-lows. The number ℓ of points in v is geometrically distributed with parameterλ vol(S):

p(ℓ) = (1 − λ vol(S))(

λ vol(S))ℓ

, (1)

where vol(S) is the multi-dimensional volume of set S. (We motivate thischoice of prior in Part I (Dauwels et al., 2009a,b).) In the particular case ofbump models in the time-frequency domain, the space S is defined as:

S = {(t, f) : t ∈ [tmin, tmax] and f ∈ [fmin, fmax]}, (2)

and therefore

vol(S) = (tmax − tmin)(fmax − fmin). (3)

Each point vk for k = 1, . . . , ℓ is uniformly distributed in S:

p(t, f |ℓ) = vol(S)−ℓ, (4)

where t and f are the positions of the hidden events (vk)k=1,...,ℓ in time andfrequency respectively. The amplitudes, widths and heights of the bumps vk

are independently and identically distributed according to priors pw, p∆t andp∆f respectively. In the following, we will discard those priors since they areirrelevant. With those choices, the prior of the hidden process v equals:

p(v, ℓ) = p(ℓ)p(v|ℓ) ∝ (1 − λ vol(S))λℓ, (5)

where the priors for the amplitudes, widths and heights of the bumps vk havebeen discarded for convenience.

From the hidden process v, the point processes (xi)i=1,...,N are generated asfollows (see Fig. 4). We first generate N identical copies of v. Next the ampli-tudes, widths and heights of the bumps are replaced by random independentdraws from priors pw, p∆t and p∆f respectively. Again, those priors are irrele-vant for what follows, and we will omit them. Next, each event is removed withprobability pd (“deletion”), independently of the other events. The probabilitymass associated with the (remaining) nk copies of (vk)k=1,...,ℓ is given by:

p(nk) = pN−nk

d (1 − pd)nk . (6)

We will need the product of p(nk) for k = 1, . . . , ℓ:

ℓ∏

k=1

p(nk) = pNℓ−Ltot

d (1 − pd)Ltot

, (7)

12

Page 13: Quantifying Statistical Interdependence PART III: N ... · Quantifying Statistical Interdependence PART III: N > 2 Point Processes J. Dauwelsa,∗,1T. WeberbF. Vialattec,dT. MushaeA.

where Ltot is the total number of events in the N point processes (xi)i=1,...,N :

Ltot =N

i=1

Li. (8)

At last, the resulting N sequences are shifted over (δi)i=1,...,N = (δti, δfi)i=1,...,N ,and the occurrence times and frequencies are slightly perturbed, resulting inthe sequences (xi)i=1,...,N . As we pointed out earlier, there might be a non-trivial timing and frequency offset between the bump models (see Fig. 3and Fig. 4). The parameters (δti, δfi) are introduced in the model to accountfor such offsets. The offsets between v and (xi)i=1,...,N may be modeled as bi-variate Gaussian random variables with mean vectors (δti, δfi) and diagonalnon-isotropic covariance matrices Vi = diag(sti, sfi). It is reasonable to as-sume that the offsets in time are independent from the offsets in frequency,and vice versa. Therefore, we use diagonal matrices Vi. Statistical dependen-cies between the perturbations in time and frequency may be modeled bynon-diagonal covariance matrices Vi. Such extensions are straightforward, andwe do not consider them here.

We adopt the improper priors p(δti) = 1 = p(δfi) for (δti)i=1,...,N and (δfi)i=1,...,N

respectively, and conjugate priors for sti and sfi, i.e. scaled inverse chi-squaredistributions:

p(sti) =(st0νt/2)νt/2

Γ(νt/2)

e−νtst0/2sti

s1+νt/2ti

(9)

p(sfi) =(sf0νf/2)νf/2

Γ(νf/2)

e−νf sf0/2sfi

s1+νf/2fi

, (10)

where νt and νf are the degrees of freedom, and st0 and sf0 are the width ofthe scaled inverse chi-square distributions, and Γ(x) is the Gamma function.

In (Dauwels et al., 2009b) we normalized the parameters (δt, st) and (δf , sf) bythe width and height of the bumps respectively, in order to take the size of thebumps into account. For simplicity, we will discard such normalization factorsin the following. They can easily be incorporated in the statistical model, andwe will briefly address this issue in Section 5.

It is also noteworthy that in (Dauwels et al., 2009b) the variance of the timeand frequency perturbations in the generative process is defined as st/2 andsf/2 respectively (instead of st and sf ), so that the variance between the twoobserved sequences x1 and x2 is given by st and sf . Therefore, when comparingresults from bivariate and N -variate SES, a factor of two needs to be takeninto account.

For later convenience, we will introduce some more notation. We denote byvcij

the hidden event that generated xij (the jth event in point process xi).

13

Page 14: Quantifying Statistical Interdependence PART III: N ... · Quantifying Statistical Interdependence PART III: N > 2 Point Processes J. Dauwelsa,∗,1T. WeberbF. Vialattec,dT. MushaeA.

The function c is hence a clustering function, that groups the events xij intodifferent clusters. Since there is at most one event from point process i incluster k, the clustering function c fulfills the constraints:

Li∑

j=1

δ[

cij − k]

≤ 1, ∀i, k. (11)

Note that certain hidden events (vk)k=1,...,ℓ may not have any copies, since allN copies may have been deleted. Therefore, the function c does not necessarilytake ℓ different values. Without loss of generality, we will assume that c takesvalues in {1, 2, . . . , L}, where L is the number of clusters and 0 ≤ L ≤ ℓ.Note that the number L of clusters is at most Ltot, i.e., the total number ofevents; this maximum number occurs when each event is a cluster, and henceall clusters are of size 1. With this definition of L, the number of copies n1,. . . , nL are non-zero, whereas nL+1 = nL+2 = · · · = nℓ = 0. We introduce theindex set K of clusters with nk > 1:

K = {k ∈ {1, 2, . . . , L} : nk > 1}. (12)

We denote by Ck the set of nk copies of vk, and denote its index set by Ik:

Ck = {xij : cij = k} and Ik = {(i, j) : cij = k}. (13)

The fraction ρ of missing events in the clusters can be computed as:

ρ = 1 −∑L

k=1 nk

LN= 1 − n

N, (14)

where n is the average number of events per cluster. Another important statis-tics is the distribution (pj)

Nj=1 of the number of events per cluster:

pj =

∑Lk=1 δ[nk − j]

L, j = 1, 2, . . . , N (15)

In this notation, the overall probabilistic model may be written as:

p(x, c, v, θ, ℓ) ∝ p(st)p(sf)(1 − λ vol(S))(λ pNd )ℓp−Ltot

d (1 − pd)Ltot

·N∏

i=1

Li∏

j=1

N(

tij − tcij; δti, sti

)

N(

fij − fcij; δfi, sfi

)

. (16)

Note that the parameters N , (Li)i=1,...,N and Ltot are fixed for given pointprocesses. Likewise, for given clustering c, the parameters L and (nk)k=1,...,L

are fixed. The total number of deletions is given by Ldel,tot = Nℓ − Ltot. Thenumber of hidden events vk without copies is given by Ldel = ℓ − L.

As in Part I and II, we can marginalize the statistical model p(x, c, v, θ, ℓ)analytically w.r.t. v and ℓ (Dauwels et al., 2009a,b), resulting in p(x, c, θ) (cf.

14

Page 15: Quantifying Statistical Interdependence PART III: N ... · Quantifying Statistical Interdependence PART III: N > 2 Point Processes J. Dauwelsa,∗,1T. WeberbF. Vialattec,dT. MushaeA.

Appendix A):

p(x, c, θ) ∝ γ βNLp(st)p(sf)

·∏

k∈K

(i,j)∈Ik

N(

tij − tk; δti, sti

)

N(

fij − fk; δfi, sfi

)

, (17)

where

tk =

(i,j)∈Ik

wti(tij − δti)

(i,j)∈Ik

wti

(18)

fk =

(i,j)∈Ik

wfi(fij − δfi)

(i,j)∈Ik

wfi, (19)

with wti = s−1ti and wfi = s−1

fi . Interestingly, the parameters (tk, fk) may beinterpreted as the coordinates of the center of the nk copies associated with vk

(see Fig. 5(b)); in other words, those nk copies may be viewed as a cluster ofevents, whose center is located at (tk, fk). As easily can be shown, the latterparameters are also the maximum likelihood (ML) estimates of the time andfrequency of the hidden event vk, when the copies of vk and the parameters(δi, si) are given (i = 1, 2 . . . , N). In practice, the latter are usually not given,and therefore, the coordinates tk and fk depend on how the events xij areassigned to hidden events vk; in other words, those parameters depend on theclustering c.

The parameters β and γ in (17) are defined as:

β = pdN√

λ, (20)

and

γ =(

p−1d (1 − pd)

)Ltot

(1 − λvol(S))1

1 − λvol(S)pNd

. (21)

Note that we defined similar parameters γ and β in Part I and II (Dauwels etal., 2009a,b); the N -variate statistical model (17) is a natural extension of thepairwise statistical models of (Dauwels et al., 2009a,b). The constant γ doesnot depend on c or the SES parameters δi and si (with i = 1, . . . , N), andtherefore, it is irrelevant for estimating the latter parameters and the clusters;we will discard γ in the following.

So far, we have not yet taken background events into account. They can bemodeled as follows. Besides the hidden process v, we generate the backgroundevents as a point process v of length ℓ. We define the prior p(v, ℓ) similarly as

15

Page 16: Quantifying Statistical Interdependence PART III: N ... · Quantifying Statistical Interdependence PART III: N > 2 Point Processes J. Dauwelsa,∗,1T. WeberbF. Vialattec,dT. MushaeA.

p(v, ℓ) (5):

p(v, ℓ) = p(ℓ)p(v|ℓ) ∝ (1 − λ vol(S))λℓ. (22)

As a result, some of the events x are generated from v (according to Step 1 to3), and the other events (background events) are the process v. We will denoteby χ the fraction of background events in x. To account for the backgroundevents, we will now assume that c takes values in {0, 1, 2, . . . , L}, where cij = 0iff xij is a background event. From a given clustering function c, we can easilyinfer the background events (and hence also ℓ). An important statistics is thefraction χ of background events, which can also easily be computed form c.

We can include background events in statistical model (17) by multiplying itwith the prior p(v, ℓ) (22), resulting in:

p(x, c, θ) ∝βNLβ ℓp(st)p(sf)

·∏

k∈K

(i,j)∈Ik

N(

tij − tk; δti, sti

)

N(

fij − fk; δfi, sfi

)

, (23)

where the parameter β = λ (cf. (20)).

The exponent of β and β in (23) does clearly depend on c, and as a result,the parameters β and β affect the inference of c and the SES parameters. Asin Part I and II, we will interpret those parameters in terms of cost functions;the expressions log β and log β are part of the cost associated to each clusterand background event respectively. In all our experiments, we found that thesetting β = 10−20 yields satisfactory results.

3.3 Interpretation in Terms of Cost Functions

We can gain additional insight by considering the logarithm of statisticalmodel (23):

− log p(x, c, θ) = − log p(st) − log p(sf) − LN log β − ℓ log β

+∑

k∈K

(i,j)∈Ik

(

1

2log 2πsti +

1

2sti(tij − tk − δti)

2

+1

2log 2πsfi +

1

2sfi(fij − fk − δfi)

2)

+ ζ, (24)

where ζ is an irrelevant constant. The expression (24) may be considered asa cost function that associates certain costs with each event and cluster; weprovided a similar viewpoint in Part I and II (Dauwels et al., 2009a,b). The

16

Page 17: Quantifying Statistical Interdependence PART III: N ... · Quantifying Statistical Interdependence PART III: N > 2 Point Processes J. Dauwelsa,∗,1T. WeberbF. Vialattec,dT. MushaeA.

unit cost d0 associated with each of the L clusters is given by:

d0 = −N log β. (25)

Likewise, the unit cost d0 associated with each of the background events isgiven by:

d0 = − log β. (26)

The unit cost of each event xij associated with a cluster k of size nk > 1equals:

d(tij , fij; c, θ) =1

2log 2πsti +

1

2sti(tij − tk − δti)

2

+1

2log 2πsfi +

1

2sfi(fij − fk − δfi)

2. (27)

This cost depends on the choice c of clusters and on the parameters θ. Indeed,the parameters tk and fk are dependent on c and θ as follows:

tk =

∑Ni=1

∑Li

j=1 δ[cij − k] wti(tij − δti)∑N

i=1

∑Li

j=1 δ[cij − k] wti

(28)

fk =

∑Ni=1

∑Li

j=1 δ[cij − k] wfi(fij − δfi)∑N

i=1

∑Li

j=1 δ[cij − k] wfi

, (29)

with wti = s−1ti and wfi = s−1

fi . The distance d(tij, fij) (27) is illustratedin Fig. 5(b).

Note that the second and fourth term in the RHS of (27) are normalizedEuclidian distances. Since the point processes (xi)1,...,N are defined on the time-frequency plane (see Fig. 4), the (normalized) Euclidean distance is indeed anatural metric. In some applications, the point process may be defined onmore general spaces, in particular, curved spaces; in such situations, one mayadopt non-Euclidean distance measures. We refer to (Dauwels et al., 2008)for an example. To simplify the notation, we will define the unit cost of eachevent xij associated with a cluster k of size nk = 1 as d(tij , fij; c, θ) = 0.

We also define costs d(st) = − log p(st), d(sf) = − log p(sf), and d(θ) =d(st) + d(sf). With those definitions of unit costs, we can rewrite (24) as:

− log p(x, c, θ) = d(θ) + Ld0 + ℓd0 +L

k=1

(i,j)∈Ik

d(tij , fij; c, θ) + ζ. (30)

17

Page 18: Quantifying Statistical Interdependence PART III: N ... · Quantifying Statistical Interdependence PART III: N > 2 Point Processes J. Dauwelsa,∗,1T. WeberbF. Vialattec,dT. MushaeA.

t

f

δ1

δ2

δ3δ4

δ5

(a) Hidden event vk (dashed line) and its 5“noisy” copies. The arrows indicate the system-atic offsets (δi) and random offsets.

t

f

(b) For the sake of clarity, the off-sets δi (cf. (a)) have been eliminated.The dot corresponds to the centerof the 5 events (cf. (28) (29)), andthe dotted lines indicate the distancesd(tij , fij) (27).

t

f

(c) After eliminating the offsets δi,event 2 (dotted) lies the closest to thehidden event (dashed), and it serves asexemplar. The arrows indicate the dis-tances d(tij , fij, ti′j′, fi′j′) (48).

Fig. 5. Exemplar representation. A cluster with 5 events, generated from a hiddenevent (dashed lines). For the sake of clarity, we have eliminated the offsets δi in (b)and (c). The exemplar is marked in dotted lines in (c).

18

Page 19: Quantifying Statistical Interdependence PART III: N ... · Quantifying Statistical Interdependence PART III: N > 2 Point Processes J. Dauwelsa,∗,1T. WeberbF. Vialattec,dT. MushaeA.

This expression can be written as a function of c as follows:

− log p(x, c, θ) = d(θ) + d0 maxij

cij + d0

ij

δ[cij ]

+maxij cij

k=1

N∑

i=1

Li∑

j=1

δ[cij − k] d(tij, fij; c, θ) + ζ , (31)

where d(tij, fij; c, θ) is given by (27). Clearly, the RHS of (31) depends on cin a non-linear fashion (cf. (27), (28) and (29)).

4 Statistical Inference

A reasonable approach to infer (c, θ) is maximum a posteriori (MAP) estima-tion:

(c, θ) = argmax(c,θ)

log p(x, c, θ), (32)

subject to (11). There is no closed form expression for (32), therefore, weneed to resort to numerical methods. A simple technique to try to find (32) iscyclic maximization: We first choose initial values θ(0), and then perform thefollowing updates for κ ≥ 1 until convergence:

c(κ) = argmaxc

log p(x, c, θ(κ−1)) (33)

θ(κ) = argmaxθ

log p(x, c(κ), θ), (34)

where (33) is determined subject to (11). The update (34) of the parameters θis straightforward, and may be carried out by cyclic maximization; we referto Appendix B for more details. The update (33) is far less straightforward;it involves an intractable optimization problem. We circumvent this issue bysolving a related tractable optimization problem. In the following, we describethat problem.

The update (33) may be expanded as:

c(κ) = argminc

(

d0 maxij

cij + d0

ij

δ[cij ]

+maxij cij

k=1

N∑

i=1

Li∑

j=1

δ[cij − k] d(tij, fij; c, θ(κ−1))

)

, (35)

which is also determined subject to (11).

19

Page 20: Quantifying Statistical Interdependence PART III: N ... · Quantifying Statistical Interdependence PART III: N > 2 Point Processes J. Dauwelsa,∗,1T. WeberbF. Vialattec,dT. MushaeA.

4.1 Equivalent (Intractable) Optimization Problem

The optimization problem (35) is hard to solve directly, and therefore, we willsolve a related tractable optimization problem instead. In order to formulatethe latter, we introduce the following binary variables:

• bk is equal to one iff cluster k is non-empty (nk > 0 and k = 1, 2, . . . , Ltot),• bijk is equal to one iff the xij belongs to cluster k, i.e., cij = k,• eij is equal to one iff xij is a background event.

We will now rewrite (35) using that notation, which will allow us to simplifythe combinatorial problem in Section 4.2.

The binary variables b are related through the constraints:

bk = min

1,N

i=1

Li∑

j=1

bijk

, ∀k. (36)

These non-linear inequality constraints are equivalent to the linear constraints:

bijk ≤ bk, ∀i, j, k, andN

i=1

Li∑

j=1

bijk ≥ bk, ∀k. (37)

The constraints (11) correspond to:

Li∑

j=1

bijk ≤ 1, ∀i, k. (38)

Moreover, an event is either a background event, or belongs to a cluster, whichcan be encoded by the constraints:

L∑

k

bijk + eij = 1, ∀i, j, (39)

In this representation, we can rewrite (35) as:

(

b(κ), e(κ))

= argminb,e

(

d0

Ltot∑

k=1

bk + d0

N∑

i=1

Li∑

j=1

eij

+Ltot∑

k=1

N∑

i=1

Li∑

j=1

bijk d(tij, fij ; b, θ(κ−1))

)

, (40)

subject to (37)–(39). As we pointed out earlier, the number of clusters is atmost Ltot, i.e., the total number of events.

20

Page 21: Quantifying Statistical Interdependence PART III: N ... · Quantifying Statistical Interdependence PART III: N > 2 Point Processes J. Dauwelsa,∗,1T. WeberbF. Vialattec,dT. MushaeA.

If xij is associated with a cluster k of size nk > 1, and hence∑N

i=1

∑Li

j=1 bijk > 1,

the expression d(tij, fij ; b, θ(κ−1)) in (40) is given by (27) (see Fig. 5(b)), with

θ = θ(κ−1) and:

tk =

∑Ni=1

∑Li

j=1 bijk w(κ−1)ti (tij − δ

(κ−1)ti )

∑Ni=1

∑Li

j=1 bijk w(κ−1)ti

(41)

fk =

∑Ni=1

∑Li

j=1 bijk w(κ−1)fi (fij − δ

(κ−1)fi )

∑Ni=1

∑Li

j=1 bijk w(κ−1)fi

, (42)

where w(κ−1)ti =

(

s(κ−1)ti

)−1and w

(κ−1)fi =

(

s(κ−1)fi

)−1; otherwise d(tij , fij; b, θ

(κ−1)) =0.

By exponentiating the objective function in (40), and adding the priors in θ,we obtain the statistical model:

p(x, b, e, θ(κ−1)) ∝ p(s(κ−1)t )p(s

(κ−1)f )

Ltot∏

k=1

(

βN)bk

N∏

i=1

Li∏

j=1

β eij

·Ltot∏

k=1

N∏

i=1

Li∏

j=1

(

N(

tij − tk; δti, sti

)

N(

fij − fk; δfi, sfi

))bijk

, (43)

which is equivalent to (23). Likewise, the constrained combinatorial optimiza-tion problem (37)–(40) is equivalent to (35); it is a non-linear combinatorialoptimization problem, since the objective function (RHS of (40)) is non-linearin b. Since the problem is intractable, we will simplify it: we linearize theobjective function by using an exemplar representation. The resulting linearcombinatorial optimization problem (a.k.a. integer linear program or ILP) canthen be solved exactly by integer linear programming. In other words, we willapproximate the original non-linear and intractable integer program (37)–(40)(which is hard to solve directly) into a linear and tractable integer program(which is much easier to solve).

4.2 Related (Tractable) Optimization Problem

Ideally, we wish to find the cluster centers (tk, fk) (41) (42) that minimizethe combined total cost d(tij, fij) (27) of all events in the cluster. That isan intractable problem, and we simplify it as follows. Rather than searchingthrough all possible cluster centers, we only consider events xij as potentialcluster centers. In other words, we consider a restricted subset of potentialcenters. More specifically, we approximate the cluster center (tk, fk) (41) (42)by the event xij of the same cluster that lies the closest to the center, after

21

Page 22: Quantifying Statistical Interdependence PART III: N ... · Quantifying Statistical Interdependence PART III: N > 2 Point Processes J. Dauwelsa,∗,1T. WeberbF. Vialattec,dT. MushaeA.

eliminating the offset (δti, δfi) (see Fig. 5(c)). The events xij that serve as clus-ter centers are referred to as “exemplars”. This approach is inspired by otherexemplar-based clustering algorithms, including affinity propagation (Frey etal., 2007) and recent extensions (Lashkari et al., 2008; Givoni et al., 2009).

As a result, the non-linear cost d(tij, fij) (27) is approximated by a cost thatis independent of b; the non-linear objective function (RHS of (40)) becomeslinear in b, and hence the non-linear combinatorial optimization problem (40)is approximated by an integer linear program. We now derive this integer linearprogram. Similarly to the variables bk and bijk, we introduce the followingbinary variables:

• bij is equal to one iff xij is an exemplar,• biji′j′ is equal to one iff xij is associated with exemplar xi′j′.

In this formulation, we approximate the intractable optimization problem (37)–(40) by the following integer linear program in b:

(

b(κ), e(κ))

= argminb,e

(

d0

N∑

i=1

Li∑

j=1

bij + d0

N∑

i=1

Li∑

j=1

eij

+N

i,i′=1

Li∑

j=1

Li′∑

j′=1

biji′j′ d(tij, fij, ti′j′, fi′j′; θ(κ−1))

)

, (44)

subject to

i′j′biji′j′ + bij + eij = 1, ∀i, j, (45)

Li∑

j=1

biji′j′ ≤ bi′j′, ∀i, i′ 6= i, j′, (46)

bijij′ = 0, ∀i, j, j′, (47)

where

d(tij , fij, ti′j′, fi′j′; θ(κ−1)) =

1

2log 2πs

(κ−1)ti +

1

2s(κ−1)ti

(tij − ti′j′ − δ(κ−1)ti )2

+1

2log 2πs

(κ−1)fi +

1

2s(κ−1)fi

(fij − fi′j′ − δ(κ−1)fi )2. (48)

By exponentiating the objective function in (44), and adding the priors in θ,

22

Page 23: Quantifying Statistical Interdependence PART III: N ... · Quantifying Statistical Interdependence PART III: N > 2 Point Processes J. Dauwelsa,∗,1T. WeberbF. Vialattec,dT. MushaeA.

we obtain the statistical model:

p(x, b, e, θ(κ−1)) ∝ p(s(κ−1)t )p(s

(κ−1)f )

N∏

i=1

Li∏

j=1

(

βN)bij

β eij

·N∏

i,i′=1

Li∏

j=1

Li′∏

j′=1

(

N(

tij − ti′j′; δ(κ−1)ti , s

(κ−1)ti

)

N(

fij − fi′j′; δ(κ−1)fi , s

(κ−1)fi

)

)biji′j′

.

(49)

The objective function (44) is a linear approximation of the non-linear objec-tive function (35), and similarly, the statistical model (49) is an approximationof p(x, c, θ(κ−1)) (23). The resulting linear integer problem (44)–(47) is mucheasier to solve than the non-linear integer problem (37)–(40). Note that bothproblems lead to similar results, since exemplars are often close to the clustercenter (cf. Fig. 5(c)).

The sum∑

ij bij in (44) is equal to the number of exemplars; therefore, thefirst term in (44) assigns a cost d0 to each exemplar. Likewise, the secondterm in (44) assigns a cost d0 to each background event. The third term in (44)associates the cost (48) to each event xij , based on its associated exemplar xi′j′.This cost is independent of b, and consequently, the objective function (44) islinear in b.

The constraints (45) ensure that each event is either an exemplar, or is associ-ated with one exemplar (and not more than one exemplar), or is a backgroundevent (insertion). The constraints (46) encode the fact that an event xij canonly be associated to an exemplar xi′j′ (biji′j′ = 1) iff the latter is indeed anexemplar (bi′j′ = 1); they also ensure that at most one event xij from xi can beassociated with an exemplar xi′j′. Finally, the constraints (47) ensure that anevent xij cannot be associated with an exemplar from the same point processxi; without those constraints, multiple events from the same point process xi

may belong to the same cluster, which is not allowed.

The combinatorial optimization problem (44)–(47) is an integer linear programin b and e, since the objective function (44) and constraints (45)–(47) arelinear in the variables b and e. More specifically, it is a binary linear program,since all variables are binary. Instead of solving the intractable problem (35) orequivalently (37)–(40), we solve the tractable problem (44)–(47); in particular,we solve it by means of off-the-shelf integer programming software.

As an alternative, we have also implemented the max-product algorithm (andvarious refinements) to solve (44)–(47), similarly as for bivariate SES (Dauwelset al., 2009a,b). Unfortunately, that approach leads to poor results for N -variate SES. In particular, it does not always convergence, and sometimesit yields solutions that violate the constraints (45)–(47); those issues mightbe due to the fact that N -wise alignment is significantly more complex than

23

Page 24: Quantifying Statistical Interdependence PART III: N ... · Quantifying Statistical Interdependence PART III: N > 2 Point Processes J. Dauwelsa,∗,1T. WeberbF. Vialattec,dT. MushaeA.

pairwise alignment.

Note that it is straightforward to determine estimates ρ, (pj)Nj=1, and χ from b

and e. Also, an estimate c can easily be computed from b and e, as follows.We number all exemplars from 1 to L, in arbitrary order. We then set cij = kif xij is the kth exemplar or if it is associated with the kth exemplar. Likewisewe set cij = 0 if xij is a background event (eij = 1). The resulting estimate cis an approximation of (35). With this estimate of c, we can eventually refinethe estimate θ, following rule (34).

The resulting SES inference algorithm is summarized Table 2.

5 Extensions

So far, we have developed N -variate SES for the particular example of bumpmodels in time-frequency domain. The statistical model (23), and equivalentlythe cost function (31), may easily be generalized, and it may be applied todifferent kinds of point processes. One simply needs to define the cost func-tions d in (31) in a suitable manner. We will briefly outline several potentialextensions and alternative applications.

• As we pointed out earlier, in (Dauwels et al., 2009b) we normalized theparameters (δt, st) and (δf , sf) by the width and height of the bumps re-spectively, in order to take the size of the bumps into account. Such nor-malization factors can easily be incorporated in cost function (31); the unitcost d(tij, fij; c, θ) of an event xij is then defined as:

d(tij, fij ; c, θ) =1

2log 2πsti +

1

2sti(tij − tk − δti)

2

+1

2log 2πsfi +

1

2sfi(fij − fk − δfi)

2, (50)

with δti = δti ∆tk, δfi = δfi ∆fk, sti = sti ∆t2k, and sfi = sfi ∆f 2k , where ∆tk

and ∆fk are the average width and height respectively of the bumps xij incluster k.

• As outlined in (Dauwels et al., 2009b, Section 6), one can easily incorporatedifferences in amplitude, width and height between the bumps of the differ-ent point processes. Moreover, the bumps may be oblique, i.e., they are notnecessarily parallel to the time and frequency axes.

• Until now we have considered bump models in the time-frequency domain.However, the statistical model (23), and equivalently the cost function (31),is readily extendable to point processes in other Euclidean spaces, e.g., three-dimensional spatial point processes or one-dimensional point processes in

24

Page 25: Quantifying Statistical Interdependence PART III: N ... · Quantifying Statistical Interdependence PART III: N > 2 Point Processes J. Dauwelsa,∗,1T. WeberbF. Vialattec,dT. MushaeA.

INPUT:

One-dimensional or multi-dimensional point processes (xi)Ni=1 and parameters β, β,

νt, νf , st0, sf0, (δ(0)ti )Ni=1, (δ

(0)fi )Ni=1, (s

(0)ti )Ni=1, (s

(0)fi )Ni=1.

ALGORITHM:

Iterate the following two steps until convergence or the available time has elapsed:

(1) Update the clustering (b, e) (and equivalently c) by ILP:

(

b(κ), e(κ))

= argminb,e

(

d0

N∑

i=1

Li∑

j=1

bij + d0

N∑

i=1

Li∑

j=1

eij

+

N∑

i,i′=1

Li∑

j=1

Li′∑

j′=1

biji′j′ d(tij , fij, ti′j′ , fi′j′ ; θ(κ−1))

)

,

subject to (45)–(47).

(2) Update the SES parameters:

Solve the equations:

δ(κ)ti =

1

Li

Li∑

j=1

(

tij − tc(κ)ij

)

δ(κ)fi =

1

Li

Li∑

j=1

(

fij − fc(κ)ij

)

s(κ)ti =

νtst0 + Li s(κ)ti,sample

νt + Li + 2

s(κ)fi =

νfsf0 + Li s(κ)fi,sample

νf + Li + 2,

with

tk =

(i,j)∈I(κ)k

w(κ)ti (tij − δ

(κ)ti )

(i,j)∈I(κ)k

w(κ)ti

fk =

(i,j)∈I(κ)k

w(κ)fi (fij − δ

(κ)fi )

(i,j)∈I(κ)k

w(κ)fi

.

OUTPUT:

Clustering (b, e) (and equivalently c) and SES parameters ρ, χ, (pj)Nj=1, (δti)

Ni=1,

(δfi)Ni=1, (sti)

Ni=1, (sfi)

Ni=1.

Table 2Inference algorithm for N -variate SES.

25

Page 26: Quantifying Statistical Interdependence PART III: N ... · Quantifying Statistical Interdependence PART III: N > 2 Point Processes J. Dauwelsa,∗,1T. WeberbF. Vialattec,dT. MushaeA.

time domain. We will consider an example of the latter in Section 7; theunit cost d(tij; c, θ) of an event xij may then defined as:

d(tij; c, θ) =1

2log 2πsti +

1

2sti(tij − tk − δti)

2, (51)

where we did not take the widths of the events into account. If the latterneed to be taken into account, we may normalize the parameters (δti, sti) asin (50).

• In some applications, the point processes may be defined on curved mani-folds, and non-Euclidean distances are then more natural. For instance, thepoint processes may be defined on planar closed curves. We refer to (Dauwelset al., 2008) for an example. The unit cost d(tij; c, θ) of an event xij maythen be defined by:

d(tij ; c, θ) =1

2log 2πsti +

1

2sti

(

g(tij, tk) − δti

)2, (52)

where tk is the “center” of cluster k, defined as:

tk = argmint

(i,j)∈Ik

1

sti

(

g(tij, t) − δti

)2, (53)

where g is an arbitrary function, potentially non-linear; for g(x, y) = x− y,we recover (51).

6 Analysis of Synthetic Data

We investigate the robustness and reliability of N -variate SES by means ofsynthetic data. We consider one-dimensional and two-dimensional point pro-cesses, as in Part I (Dauwels et al., 2009a) and II (Dauwels et al., 2009b)respectively. We discuss the results for one-dimensional and two-dimensionalpoint processes in Section 6.1 and 6.2 respectively.

6.1 One-Dimensional Point Processes

We randomly generated 1’000 sets of N = 5 one-dimensional point processesaccording to the generative process outlined in Section 3.

We tested several values of the parameters pd, δti, and sti (σti), for i = 1, 2, . . . ,5. In particular, we tested the values pd = 0, 0.1, . . . , 0.4, and σti = 10ms, 30ms,and 50ms (for i = 1, 2, . . . , 5), tmin = 0ms, and tmax = ℓ0 · 100ms. The lengthℓ was chosen as ℓ = ℓ0/(1−pd), where we tested the values ℓ0 = 40, 100. With

26

Page 27: Quantifying Statistical Interdependence PART III: N ... · Quantifying Statistical Interdependence PART III: N > 2 Point Processes J. Dauwelsa,∗,1T. WeberbF. Vialattec,dT. MushaeA.

this choice, the expected length of the point processes is ℓ0, independently ofpd. In one set of experiments, we set δti = 0, for i = 1, 2, . . . , 5. In a secondset, the offsets δti are drawn uniformly within [−50ms, 50ms]. In each case, wedid not insert events (cf. Step 4).

We used the initial values δ(0)ti = 0ms, and s

(0)ti = (20ms)2, (30ms)2, for i = 1, 2,

. . . , 5. The parameter β was identical for all parameter settings, i.e., β = 0.04;it was optimized to yield the best overall results. We used an uninformativeprior for δti and sti,, i.e., p(δti) = p(sti) = 1, for i = 1, 2, . . . , 5. One could testvarious initial values of δti, for i = 1, 2, . . . , 5. However, the number of initialconditions grows exponentially with N , and therefore, it is not really practicalto test multiple values for each δ

(0)ti . For example, if we test 3 values for each

δ(0)ti , we need to test a total of 35 = 243 initial values δ

(0)ti . Alternatively, one

may use a small random subset of initial values; for the sake of conciseness,we do not consider that approach here.

We set β = 10−20, as in all our simulations in this paper; for the syntheticdata, no background events were inferred, i.e., χ = 0 for all parameter settings.

In order to assess the SES measures S = st, ρ, we compute for each abovementioned parameter setting the expectation E[S] and normalized standarddeviation σ[S] = σ[S]/E[S]. Those statistics are computed by averaging over1’000 sets of 5 point processes, randomly generated according to the generativeprocess outlined in Section 3.

The results are summarized in Fig. 6. From this figure we can make the fol-lowing observations:

• The estimates of st and pd are slightly biased, especially for small ℓ0, i.e., ℓ0

= 40, st ≥ (30ms)2 and pd > 0.2. However, the bias is significantly smallerthan for bivariate SES (see (Dauwels et al., 2009a)).

• The estimates of st do only weakly depend on pd, and vice versa.• The estimates of st and pd only weakly depend on δti (curved vs. solid lines);

they are robust to lags δt. Note that one could further reduce this depen-dency by testing various initial values δ

(0)ti . However, the number of initial

conditions grows exponentially with N , as we mentioned earlier; therefore,this approach is not really practical.

• The estimates of st and pd are less biased for larger ℓ0.

We have also observed from our experiments (not shown here):

• The estimates of δt are unbiased for all considered values of δt, st, and pd.• The normalized standard deviation of the estimates of δt, st and pd grows

with st and pd, but it remains below 30%. Those estimates are thereforereliable.

27

Page 28: Quantifying Statistical Interdependence PART III: N ... · Quantifying Statistical Interdependence PART III: N > 2 Point Processes J. Dauwelsa,∗,1T. WeberbF. Vialattec,dT. MushaeA.

0 0.2 0.40

10

20

30

40

50

σt = 10

σt = 30

σt = 50

E[σ

t][m

s]

pd

(a) ℓ0 = 40.

0 0.2 0.40

10

20

30

40

50

σt = 10

σt = 30

σt = 50

E[σ

t][m

s]

pd

(b) ℓ0 = 100.

0 0.2 0.40

0.1

0.2

0.3

0.4

σt = 10

σt = 30

σt = 50

E[ρ

]

pd

(c) ℓ0 = 40.

0 0.2 0.40

0.1

0.2

0.3

0.4

σt = 10

σt = 30

σt = 50

E[ρ

]

pd

(d) ℓ0 = 100.

Fig. 6. Results for N -variate stochastic event synchrony for one-dimensional pointprocesses: the figure shows the expected value E[σt] and E[ρ] for the parametersettings ℓ0 = 40, 100, σt = 10, 30, 50ms, and pd = 0, 0.1, . . . , 0.4. The solid linesare for zero delays δti, whereas the dotted lines are for offsets δti drawn uniformlywithin [−50ms, 50ms].

• The normalized standard deviation of the SES parameters decreases as thelength ℓ0 increases, as expected.

6.2 Two-Dimensional Point Processes

Similarly as in the one-dimensional case, we randomly generated 1’000 sets ofN = 5 two-dimensional point processes according to the generative processoutlined in Section 3.

We considered several values of the parameters pd, δti, sti (σti), δfi, and sfi

(σfi), for i = 1, 2, . . . , 5. In particular, we tested the values pd = 0, 0.1, . . . , 0.4,and σti = 10ms, 30ms, and 50ms, σfi = 1Hz, 2.5Hz, 5Hz (for i = 1, 2, . . . , 5),tmin = 0ms, and tmax = ℓ0 ·100ms, fmin = 0Hz, and fmax = ℓ0 ·1Hz. The lengthℓ was chosen as ℓ = ℓ0/(1−pd), where we tested the values ℓ0 = 40, 100. Withthis choice, the expected length of the point processes is ℓ0, independently of

28

Page 29: Quantifying Statistical Interdependence PART III: N ... · Quantifying Statistical Interdependence PART III: N > 2 Point Processes J. Dauwelsa,∗,1T. WeberbF. Vialattec,dT. MushaeA.

0 0.2 0.40

10

20

30

40

50

σt = 10

σt = 30

σt = 50

E[σ

t][m

s]

pd

(a) ℓ0 = 40.

0 0.2 0.40

10

20

30

40

50

σt = 10

σt = 30

σt = 50

E[σ

t][m

s]

pd

(b) ℓ0 = 100.

0 0.2 0.40

0.1

0.2

0.3

0.4

σt = 10

σt = 30

σt = 50

E[ρ

]

pd

(c) ℓ0 = 40.

0 0.2 0.40

0.1

0.2

0.3

0.4

σt = 10

σt = 30

σt = 50

E[ρ

]

pd

(d) ℓ0 = 100.

Fig. 7. Results for synthetic data: the figure shows the expected value E[σt] and E[ρ]and the normalized standard deviation σ[σt] and σ[ρ] for the parameter settingsℓ0 = 40, 100, σt = 10, 30, 50ms, pd = 0, 0.1, . . . , 0.4, and σf = 1Hz. The solid linesare for zero delays δti, whereas the dotted lines are for offsets δti and δfi drawnuniformly within [−50ms, 50ms] and [−5Hz, 5Hz] respectively. The curves for zeroand random delays are practically coinciding.

pd. In one set of experiments, we set δti = 0 = δfi, for i = 1, 2, . . . , 5. In asecond set, the offsets δti and δfi are drawn uniformly within [−50ms, 50ms]and [−5Hz, 5Hz] respectively. In each case, we did not insert events (cf. Step4).

We used the initial values δ(0)ti = 0ms, δ

(0)fi = 0Hz, s

(0)ti = (20ms)2, (30ms)2,

and s(0)fi = (2Hz)2, for i = 1, 2, . . . , 5. The parameter β was identical for all

parameter settings, i.e., β = 0.01; it was optimized to yield the best overallresults. We used an uninformative prior for δti, δfi, sti, and sfi, i.e., p(δti) =p(sti) = p(δfi) = p(sfi) = 1.

The results are summarized in Fig. 7 to 9. Overall, they are quite similarto the ones for one-dimensional point processes (cf. Fig. 6). We observe thefollowing:

• The estimates of st and pd are slightly biased. However, the bias is signif-icantly smaller than in the one-dimensional case, as for bivariate SES (cf.Fig.4 and Section 7 in (Dauwels et al., 2009b)).

• The bias increases with sf , which is in agreement with our expectations:the more frequency jitter, the more likely that some events are reversed in

29

Page 30: Quantifying Statistical Interdependence PART III: N ... · Quantifying Statistical Interdependence PART III: N > 2 Point Processes J. Dauwelsa,∗,1T. WeberbF. Vialattec,dT. MushaeA.

0 0.2 0.40

10

20

30

40

50

σt = 10

σt = 30

σt = 50

E[σ

t][m

s]

pd

(a) ℓ0 = 40.

0 0.2 0.40

10

20

30

40

50

σt = 10

σt = 30

σt = 50

E[σ

t][m

s]

pd

(b) ℓ0 = 100.

0 0.2 0.40

0.1

0.2

0.3

0.4

σt = 10

σt = 30

σt = 50

E[ρ

]

pd

(c) ℓ0 = 40.

0 0.2 0.40

0.1

0.2

0.3

0.4

σt = 10

σt = 30

σt = 50

E[ρ

]

pd

(d) ℓ0 = 100.

Fig. 8. Results for synthetic data: the figure shows the expected value E[σt] andE[ρ] and the normalized standard deviation σ[σt] and σ[ρ] for same the parametersettings as in Fig. 7, but now with σf = 2.5Hz. Again, the curves for zero andrandom delays are practically coinciding.

frequency, and hence are aligned incorrectly. The bias is about the same asfor bivariate SES (see (Dauwels et al., 2009b)).

• The estimates of st do only weakly depend on pd, and vice versa.• The estimates of st and pd are robust to lags δt and frequency offsets δf ,

since the latter can be estimated reliably.• The estimates of st and pd are less biased for larger ℓ0.

We have also observed from our experiments (not shown here):

• The estimates of δt and δf are unbiased for all considered values of δt, δf ,st, sf , and pd.

• The normalized standard deviation of the estimates of δt, st and pd growswith st and pd, but it remains below 30%. Those estimates are thereforereliable.

• The normalized standard deviation of the SES parameters decreases as thelength ℓ0 increases, as expected.

In summary, by means of the N -variate SES inference method, one may re-liably and robustly determine the timing dispersion st and event reliabilityρ of a set of N (one-dimensional or multi-dimensional) point processes. Aswe also observed in (Dauwels et al., 2009a,b) for bivariate SES, the timingdispersion and the number of event deletions are slightly underestimated due

30

Page 31: Quantifying Statistical Interdependence PART III: N ... · Quantifying Statistical Interdependence PART III: N > 2 Point Processes J. Dauwelsa,∗,1T. WeberbF. Vialattec,dT. MushaeA.

0 0.2 0.40

10

20

30

40

50

σt = 10

σt = 30

σt = 50

E[σ

t][m

s]

pd

(a) ℓ0 = 40.

0 0.2 0.40

10

20

30

40

50

σt = 10

σt = 30

σt = 50

E[σ

t][m

s]

pd

(b) ℓ0 = 100.

0 0.2 0.40

0.1

0.2

0.3

0.4

σt = 10

σt = 30

σt = 50

E[ρ

]

pd

(c) ℓ0 = 40.

0 0.2 0.40

0.1

0.2

0.3

0.4

σt = 10

σt = 30

σt = 50

E[ρ

]

pd

(d) ℓ0 = 100.

Fig. 9. Results for synthetic data: the figure shows the expected value E[σt] andE[ρ] and the normalized standard deviation σ[σt] and σ[ρ] for same the parametersettings as in Fig. 7, but now with σf = 5Hz. Again, the curves for zero and randomdelays are practically coinciding.

to the ambiguity inherent in event synchrony. However, this bias is smallerfor N -variate SES than for bivariate SES, especially for one-dimensional pointprocesses.

7 Application: Firing Reliablity of a Neuron

We consider here an application of SES that we also investigated in (Dauwelset al., 2009a): We use SES to quantify the firing reliability of neurons. We againconsider the Morris-Lecar neuron model (Morris et al., 1981), which exhibitsproperties of type I and II neurons (Gutkin et al., 1998; Tsumoto et al., 2007;Tateno et al., 2004). The spiking behavior differs in both neuron types, asillustrated in Fig. 10. In type II neurons, the timing jitter is small, but spikestend to drop out. In type I neurons, on the other hand, fewer spikes dropout, but the dispersion of spike times is larger. In other words, type II neuronsprefer to stay coherent or to be silent, on the other hand, type I neurons followthe middle course between those two extremes (Robinson, 2003).

In (Dauwels et al., 2009a) we applied bivariate SES to the data of Fig. 10. Herewe apply N -variate SES to the same data set. (We refer to (Dauwels et al.,

31

Page 32: Quantifying Statistical Interdependence PART III: N ... · Quantifying Statistical Interdependence PART III: N > 2 Point Processes J. Dauwelsa,∗,1T. WeberbF. Vialattec,dT. MushaeA.

0 500 1000 1500 2000 2500 3000 3500 40000

5

10

15

20

25

30

35

40

45

50

t [ms]

tria

l

(a) Spike trains from type I neuron.

0 500 1000 1500 2000 2500 3000 3500 40000

5

10

15

20

25

30

35

40

45

50

t [ms]

tria

l

(b) Spike trains from type II neuron.

Fig. 10. Raster plots of spike trains from type I (top) and type II (bottom) neurons;in each case 50 spike trains are shown.

32

Page 33: Quantifying Statistical Interdependence PART III: N ... · Quantifying Statistical Interdependence PART III: N > 2 Point Processes J. Dauwelsa,∗,1T. WeberbF. Vialattec,dT. MushaeA.

1000 1200 1400 1600 18000

5

10

15

20

25

30

35

40

45

50

t [ms]

tria

l

(a) Type I spike trains clustered by N -variate SES.

1000 1200 1400 1600 18000

5

10

15

20

25

30

35

40

45

50

t [ms]

tria

l

(b) Type II spike trains clustered by N -variate SES.

Fig. 11. Results of N -variate SES with β = 10−3; for the sake of clarity, we showthe range t ∈ [950, 1950]. Raster plots of spike trains from type I (top) and type II(bottom) neurons. Each cluster is indicated by a different combination of color andmarker type (e.g., star, circle); background events are marked by red hexagons.

33

Page 34: Quantifying Statistical Interdependence PART III: N ... · Quantifying Statistical Interdependence PART III: N > 2 Point Processes J. Dauwelsa,∗,1T. WeberbF. Vialattec,dT. MushaeA.

2009a) for more details on that data set.) In particular, we apply N -variateSES to the 50 trials simultaneously.

As an illustration, we show results of N -variate SES in Fig. 11. Each clusteris indicated by a different combination of color and marker type (e.g., star,circle); background events are marked by red hexagons.

We choose the parameters in the N -variate SES algorithm as follows. We setδ(0)t = 0, and s

(0)t = (3ms)2, (5ms)2, (7ms)2 and (9ms)2. Each initialization

of (δ(0)t , s

(0)t ) may lead to a different solution (c, δt, st); we choose the most

probable solution, i.e., the one that has the largest value p(x, c, δt, st) (23).We set β = 10−10. Larger values of β lead to a prohibitively large number ofbackground events, whereas smaller values yield no background events at all.

We computed the SES parameters for different values of β. Fig. 12 shows howst (σt), ρ, and χ (fraction of background events) depend on β for both neurontypes. From those figures it becomes immediately clear that the parametersst (σt) and ρ hardly depend on β. For values of β < 10−4, some of the eventsfrom the Type II neuron are considered as background events, which is ob-viously incorrect (cf. Fig. 11(b)). Therefore, only values β > 10−4 should beconsidered.

The parameter ρ is significantly smaller in type I than in type II neurons, incontrast, st is vastly larger. This agrees with our intuition: since in type IIneurons spikes tend to drop out, ρ should be larger. On the other hand, sincethe timing dispersion of the spikes in type I is larger, we expect st to be largerin those neurons. We have made the same observations in (Dauwels et al.,2009a).

Table 3 compares the numerical results for bivariate and N -variate SES. Aswe pointed out earlier, bivariate SES defines the variance of the perturbationsin the generative process as st/2 (instead of st), so that the variance betweenthe two observed sequences x1 and x2 is given by st. In Table 3 the standarddeviations (σt =

√st) in the generative process are reported, both for bivariate

and N -variate SES; the values in (Dauwels et al., 2009a) differ by a factor√

2,since there the standard deviation between the observed sequences is reported.

In (Dauwels et al., 2009a) we assessed the reliability of the bivariate SESestimates by means of bootstrapping (Efron et al., 1993); we follow a similarprocedure here for the N -variate SES estimates. In particular, for both types ofneurons we generated 1,000 sets of 50 spike trains; we followed the generativeprocess of Fig. 4, with the N -variate SES parameters of the actual spike trains,i.e., (ρ, σt, χ) = (10.6, 0.0025, 0.035) and (ρ, σt, χ) = (2.8, 0.18, 0.0) for typeI and II neurons respectively. Next we applied N -variate SES to the resulting1,000 sets of 50 spike trains. The expected value and normalized standard

34

Page 35: Quantifying Statistical Interdependence PART III: N ... · Quantifying Statistical Interdependence PART III: N > 2 Point Processes J. Dauwelsa,∗,1T. WeberbF. Vialattec,dT. MushaeA.

Type IType II

β

σt[m

s]

10−5 10−410−3 10−2

0

50

100

150

(a) The parameter σt as a function of β.

Type IType II

β10−5 10−4

10−3 10−20

0.05

0.1

0.15

0.2

0.25

0.3

ρ

(b) The parameter ρ as a function of β.

Type IType II

β10−5 10−4

10−3 10−20

0.1

0.04

0.06

0.08

0.02

χ

(c) The parameter χ as a function of β.

Fig. 12. The parameters σt, ρ, and χ estimated from spike trains of type I andtype II Morris-Lecar neurons (cf. Fig. 10): the top, middle, and bottom figure showhow σt, ρ, and χ respectively depend on β.

35

Page 36: Quantifying Statistical Interdependence PART III: N ... · Quantifying Statistical Interdependence PART III: N > 2 Point Processes J. Dauwelsa,∗,1T. WeberbF. Vialattec,dT. MushaeA.

deviation σ of those estimates is reported in Table 3. We can observe thatthe expected value corresponds well with the actual value, and the normalizedstandard deviations are small. Therefore, the N -variate SES estimates can beconsidered reliable.

We first discuss the results for the type I neuron: the estimate of σt frombivariate and N -variate SES is almost identical; however, the estimate of ρis much smaller for N -variate SES than for bivariate SES. Interestingly, N -variate SES inferred that about 3.5% of the events are background events,which accounts for the larger estimate ρ = 0.029 from the bivariate approach.In other words, type I neurons almost never fail to fire (firing reliability of99.75%), however, additional spikes may occur (3.5% of the spikes). It is note-worthy that this insight was obtained by N -variate SES, and could not berevealed through bivariate SES; this example thus illustrates that N -variateSES not only yields more accurate estimates of the SES parameters (cf. Sec-tion 6), but also can lead to a more refined and detailed analysis. Interestingly,the normalized standard deviation σ[ρ] is much larger for N -variate SES thanfor bivariate SES, since ρ is much smaller for N -variate SES. However, thestandard deviation σ[ρ] is small and about the same for both models.

We now elaborate on the results for the type II neuron. The N -variate ap-proach leads to larger and smaller estimates of σt and ρ respectively than thebivariate approach. None of the events are considered as background events(χ = 0). We have manually counted the number of deletions in Fig. 11(b),and obtained ρ = 0.184. The N -variate estimate of ρ is exact and clearly themost reliable, whereas the bivariate approach overestimates the number ofdeletions. Since the N -variate approach considers all point processes simulta-neously (N = 50), it infers the hidden process v more reliably than bivariateSES, and is able to associate events x more accurately with hidden events vk.

We have observed that the N -variate SES algorithm converges after at mostthree iterations, both for type I and type II neurons. In each of those iterations,one updates the decision variables b, c, and e, and the SES parameters θ.Since we allowed a maximum number of 30 iterations, we can conclude thatthe algorithm always has converged in our experiments.

8 Application: Diagnosis of MCI from EEG

Several clinical studies have shown that the EEG of Alzheimer’s disease (AD)patients is generally less coherent than of age-matched control subjects; this isalso the case for patients suffering from mild cognitive impairment (see (Jeong,2004; Dauwels et al., 2010b) for a review). In this section, we apply SES to de-tect subtle perturbations in EEG synchrony of MCI patients. We considered

36

Page 37: Quantifying Statistical Interdependence PART III: N ... · Quantifying Statistical Interdependence PART III: N > 2 Point Processes J. Dauwelsa,∗,1T. WeberbF. Vialattec,dT. MushaeA.

Bivariate SES N -variate SES

Statistics Type I Type II Type I Type II

σt 10.7 1.91 10.6 2.81

E[σt] 10.8 1.91 10.3 2.87

σ[σt] 1.8% 1.8% 3.7% 3.0%

ρ 0.0290 0.270 0.0025 0.184

E[ρ] 0.0283 0.273 0.0036 0.186

σ[ρ] 12% 3.1% 90% 15%

χ – – 0.035 0.0

E[χ] – – 0.037 0.0

σ[χ] – – 11% 0.0%

Table 3Estimates of bivariate SES estimates (ρ, σt) and N -variate SES parameters (ρ, σt,χ). Also shown are the results from the bootstrapping analysis of those estimates,in particular, the expected values and the normalized standard deviations σ. Theexpected values practically coincide with the actual estimates and the normalizedstandard deviations are small; therefore, the estimates may be considered reliable.

this application also in (Dauwels et al., 2009b), where we applied bivariateSES. We analyze here the same EEG data set, and use the same preprocess-ing and bump modeling procedures as in (Dauwels et al., 2009b). The onlydifference is that we here apply N -variate SES instead of bivariate SES.

We first conducted a similar statistical analysis as in (Dauwels et al., 2009b).The main results of that analysis are summarized in Fig. 13 and 14; theycontain p-values obtained by the Mann-Whitney test for the parameters ρand st respectively. This test indicates whether the parameters take differentvalues for the two subject populations. More precisely, low p-values indicatelarge difference in the medians of the two populations. The p-values are shownfor σ

(0)t =

√s0,t = 0.1, 0.15, . . . , 0.25, σ

(0)f =

√s0,f = 0.05, 0.1, . . . , 0.15, β =

0.01, 0.001, 0.0001, T = 0.21, 0.22, 0.23, 0.24, and the number of zones NR =3 and 5.

The results for the parameters ρ and st are quite similar to the results obtainedwith bivariate SES (see (Dauwels et al., 2009b)). The lowest p-values for ρ are

37

Page 38: Quantifying Statistical Interdependence PART III: N ... · Quantifying Statistical Interdependence PART III: N > 2 Point Processes J. Dauwelsa,∗,1T. WeberbF. Vialattec,dT. MushaeA.

0.05 0.1 0.1510

−4

10−3

10−2

10−1

100

T=0.21T=0.22T=0.23T=0.24

p-v

alue

σ(0)f(a) β = 0.01, 3 zones

0.05 0.1 0.1510

−4

10−3

10−2

10−1

100

T=0.21T=0.22T=0.23T=0.24

p-v

alue

σ(0)f(b) β = 0.001, 3 zones

0.05 0.1 0.1510

−4

10−3

10−2

10−1

100

T=0.21T=0.22T=0.23T=0.24

p-v

alue

σ(0)f(c) β = 0.0001, 3 zones

0.05 0.1 0.1510

−4

10−3

10−2

10−1

100

T=0.21T=0.22T=0.23T=0.24

p-v

alue

σ(0)f(d) β = 0.01, 5 zones

0.05 0.1 0.1510

−4

10−3

10−2

10−1

100

T=0.21T=0.22T=0.23T=0.24

p-v

alue

σ(0)f(e) β = 0.001, 5 zones

0.05 0.1 0.1510

−4

10−3

10−2

10−1

100

T=0.21T=0.22T=0.23T=0.24

p-v

alue

σ(0)f(f) β = 0.0001, 5 zones

Fig. 13. p-values obtained by the Mann-Whitney test for the parameter ρ for

σ(0)t =

√s0,t = 0.1, 0.15, . . . , 0.25, σ

(0)f =

√s0,f = 0.05, 0.01, 0.15, β = 0.01, 0.001,

0.0001, T = 0.21, 0.22, 0.23, 0.24 and the number of zones NR = 3 and 5. The

p-values seem to vary little with σ(0)t , σ

(0)f and β, but are more dependent on T and

the number of zones. The lowest p-values are obtained for T = 0.22 and NR = 5zones; the corresponding statistical differences are highly significant.

obtained for T = 0.22, β = 0.01, and NR = 5 (see Fig. 13(e)), i.e., the smallestvalue is p = 5.4 · 10−4; in bivariate SES, the smallest p-value (p = 2.1 · 10−4)was obtained for for T = 0.22, β = 0.001, and NR = 5 (see (Dauwels etal., 2009b)). Similarly as in bivariate SES, the results depend strongly on T(cf. (Dauwels et al., 2009b)); we provided an explanation for this dependency

in (Dauwels et al., 2009b). Interestingly, the results depend much less on σ(0)t ,

σ(0)f , and β than in bivariate SES.

From bivariate and N -variate SES analysis (see Fig. 13), we can conclude thatthe statistical differences in ρ are highly significant, especially for T = 0.22 andNR = 5: There is a significantly higher degree of non-correlated activity in MCIpatients, more specifically, a high number of non-coincident, non-synchronousoscillatory events. As in (Dauwels et al., 2009b), we did not observe a stronglysignificant effect on the timing jitter st of the coincident events (see Fig. 14):very few p-values for st are smaller than 0.001, which suggests there are nostrongly significant differences in st.

The N -variate SES model allows us to analyze the results for ρ in more de-tail; we have investigated how the statistics of bump clusters differ in controlssubjects and MCI patients. More specifically, we considered the relative fre-

38

Page 39: Quantifying Statistical Interdependence PART III: N ... · Quantifying Statistical Interdependence PART III: N > 2 Point Processes J. Dauwelsa,∗,1T. WeberbF. Vialattec,dT. MushaeA.

0.05 0.1 0.1510

−4

10−3

10−2

10−1

100

T=0.21T=0.22T=0.23T=0.24

p-v

alue

σ(0)f(a) β = 0.01, 3 zones

0.05 0.1 0.1510

−4

10−3

10−2

10−1

100

T=0.21T=0.22T=0.23T=0.24

p-v

alue

σ(0)f(b) β = 0.001, 3 zones

0.05 0.1 0.1510

−4

10−3

10−2

10−1

100

T=0.21T=0.22T=0.23T=0.24

p-v

alue

σ(0)f(c) β = 0.0001, 3 zones

0.05 0.1 0.1510

−4

10−3

10−2

10−1

100

T=0.21T=0.22T=0.23T=0.24

p-v

alue

σ(0)f(d) β = 0.01, 5 zones

0.05 0.1 0.1510

−4

10−3

10−2

10−1

100

T=0.21T=0.22T=0.23T=0.24

p-v

alue

σ(0)f(e) β = 0.001, 5 zones

0.05 0.1 0.1510

−4

10−3

10−2

10−1

100

T=0.21T=0.22T=0.23T=0.24

p-v

alue

σ(0)f(f) β = 0.0001, 5 zones

Fig. 14. p-values obtained by the Mann-Whitney test for the parameter st for

σ(0)t =

√s0,t = 0.1, 0.15, . . . , 0.25, σ

(0)f =

√s0,f = 0.05, 0.01, 0.15, β = 0.01, 0.001,

0.0001, T = 0.21, 0.22, 0.23, 0.24 and the number of zones NR = 3 and 5. Veryfew p-values are smaller than 0.001, which suggests there are no strongly significantdifferences in st.

quency pj = p(nk = j) of bump clusters of size nk = j, for j = 1, 2, . . . , NR.The results are summarized in Fig. 15, for the parameter settings that yieldedthe smallest p-values for ρ (NR = 5 and β = 0.01). From those figures, wecan observe strongly significant differences in clusters of size 1, 2 and 5 forT = 0.22; specifically, in MCI patients there are fewer clusters of size 5 andmore clusters of size 1 and 2. As a result, the fraction of missing events ρ islarger in MCI patients, as we mentioned earlier. The smallest p-values for theparameter p2 is p = 2 · 10−5, which is substantially smaller than the smallestp-value for ρ (p = 2.1 · 10−4 for bivariate SES and p = 5.4 · 10−4 for N -variateSES).

In (Dauwels et al., 2010a) we have applied a variety of classical synchronymeasures to the same EEG data set. The main results of that analysis aresummarized in Table 4. The most significant results were obtained with thethe full-frequency direct transfer function (ffDTF), which is a Granger mea-sure (Pereda et al., 2005), resulting in a p-value of about 10−3 (Mann-Whitneytest). In (Dauwels et al., 2009b) we have combined ρ with ffDTF as featuresto distinguish MCI from control subjects (see Fig. 16). About 84% of thesubjects are correctly classified. Here we combine ffDTF with the parameterp2, computed by N -variate SES; the classification rate slightly improves toabout 87%. This result is encouraging, however, it is too weak to allow us to

39

Page 40: Quantifying Statistical Interdependence PART III: N ... · Quantifying Statistical Interdependence PART III: N > 2 Point Processes J. Dauwelsa,∗,1T. WeberbF. Vialattec,dT. MushaeA.

0.05 0.1 0.15

10−4

10−3

10−2

10−1

100

T=0.21T=0.22T=0.23T=0.24

p-v

alue

σ(0)f(a) p1

0.05 0.1 0.15

10−4

10−3

10−2

10−1

100

T=0.21T=0.22T=0.23T=0.24

p-v

alue

σ(0)f(b) p2

0.05 0.1 0.15

10−4

10−3

10−2

10−1

100

T=0.21T=0.22T=0.23T=0.24

p-v

alue

σ(0)f(c) p3

0.05 0.1 0.15

10−4

10−3

10−2

10−1

100

T=0.21T=0.22T=0.23T=0.24

p-v

alue

σ(0)f(d) p4

0.05 0.1 0.15

10−4

10−3

10−2

10−1

100

T=0.21T=0.22T=0.23T=0.24

p-v

alue

σ(0)f(e) p5

Fig. 15. p-values obtained by the Mann-Whitney test for the parameters

pi = p(nk = i) (with i = 1, 2, . . . , 5) for σ(0)t =

√s0,t = 0.05, 0.1, 0.15,

σ(0)f =

√s0,f = 0.01, 0.15, . . . , 0.25, β = 0.01, T = 0.22, and the number of zones

NR = 5.

Measure Correlation Coherence Phase Coherence Corr-entropy Wave-entropy

p-value 0.025∗

0.029∗

0.041∗

0.032∗ 0.096

Measure MVAR coherence Partial Coherence PDC DTF ffDTF dDTF

p-value 0.15 0.16 0.60 0.29 0.0012∗∗†

0.029∗

Measure Kullback-Leibler Renyi Jensen-Shannon Jensen-Renyi IW I

p-value 0.065 0.067 0.069 0.074 0.052 0.060

Measure Nk Sk Hk S-estimator Omega complexity

p-value 0.029∗

0.045∗ 0.052 0.042

∗ 0.079

Measure Hilbert Phase Wavelet Phase Evolution Map Instantaneous Period GFS

p-value 0.96 0.082 0.64 0.73 0.031∗

Measure st (bivariate) ρ (bivariate) st (multivariate) ρ (multivariate) p2 (multivariate)

p-value 0.19 0.00021∗∗

0.018∗

0.00054∗∗

2 · 10−5 ∗∗

Table 4Sensitivity for prediction of MCI: uncorrected p-values for Mann-Whitney test; *and ** indicate p < 0.05 and p < 0.005 respectively; † indicates p-values thatremain significant after post-correction. The results for the standard measures andbivariate SES (st and ρ) are from (Dauwels et al., 2010a) and (Dauwels et al., 2009b)respectively, and we refer to those references for more details.

40

Page 41: Quantifying Statistical Interdependence PART III: N ... · Quantifying Statistical Interdependence PART III: N > 2 Point Processes J. Dauwelsa,∗,1T. WeberbF. Vialattec,dT. MushaeA.

0.045 0.05 0.055 0.060

0.05

0.1

0.15

0.2

MCICTR

p 2

F 2ij

Fig. 16. Combining p2 with ffDTF as features to distinguish MCI from age-matchedcontrol subjects. Note that ffDTF is a similarity measure whereas p2 is a dissimilaritymeasure. The (ffDTF, p2) pairs of the MCI and control subjects tend towards theleft top corner and bottom right corner respectively. The smooth curve (solid) yieldsa classification rate of 87%.

predict AD reliably. We would need to combine synchrony measures with com-plementary statistics, for example, spectral features. We refer to (Dauwels etal., 2010b) for more information on potential extensions. Moreover, the resultswould of course need to be verified on more datasets.

In summary, N -variate SES helped us to better understand the results fromthe bivariate SES analysis (Dauwels et al., 2009b): the fraction of missingevents ρ is larger in MCI patients, since in those patients there are fewerbump clusters of size 5 and more clusters of size 1 and 2. Moreover, N -variateSES allowed us to further improve the classification of MCI patients vs. controlsubjects from 84% to 87%.

9 Conclusion

We proposed an approach to determine the similarity of N > 2 (one- andmulti-dimensional) point processes; it is based on an exemplar-based statisti-cal model that describes how the point processes are related through a commonhidden process. The similarity of the point processes is determined by perform-ing inference in that model by means of integer programming techniques in

41

Page 42: Quantifying Statistical Interdependence PART III: N ... · Quantifying Statistical Interdependence PART III: N > 2 Point Processes J. Dauwelsa,∗,1T. WeberbF. Vialattec,dT. MushaeA.

conjunction with point estimation of the parameters. The proposed techniquemay be used for various applications in neuroscience (e.g., in brain-computerinterfaces, analysis of spike data), biomedical signal processing, and beyond.

In bivariate SES, we apply the max-product algorithm for aligning pairs ofsequences. As we have observed, the max-product algorithm performs poorlyfor aligning N > 2 sequences; we have also experimented with various refine-ments of the max-product algorithm, and none of them yielded satisfactoryresults. An interesting topic for future research is to develop extensions ofthe max-product algorithm that lead to optimal or close-to-optimal N -wisealignments. Such message passing algorithms are often simpler and faster thaninteger linear programming techniques.

Acknowledgements

The authors wish to thank the anonymous reviewers for their helpful feedbackthat has substantially improved the paper.

A Appendix: Derivation of the SES Model

In this appendix, we derive the N -variate SES model (17).

We first marginalize p(x, c, v, θ, ℓ) (16) over v; it is noteworthy that only theGaussian terms N (·) in (16) depend on v. In the following, we will focus onthose terms. For a given hidden event vk, three cases are possible:

• All copies of vk were deleted (nk = 0). There are no Gaussian terms as-sociated with vk in (16), therefore, the expression (16) may be consideredas constant w.r.t. vk. Integrating (16) over vk then leads to a term vol(S).There are Ldel such terms, since the number of hidden events vk withoutcopies is given by Ldel.

• There is one copy of vk (nk = 1), and therefore, there is only one Gaussianterm in (16) that corresponds to vk. Integrating that term over vk results inthe (trivial) term 1.

• There are more than one copies of vk (nk > 1), and as a consequence, thereare several Gaussian terms in (16) that correspond to vk. As easily can beshown, integrating the nk Gaussian terms over vk yields the terms:

(i,j)∈Ik

N(

tij − tk; δti, sti

)

N(

fij − fk; δfi, sfi

)

, (A.1)

42

Page 43: Quantifying Statistical Interdependence PART III: N ... · Quantifying Statistical Interdependence PART III: N > 2 Point Processes J. Dauwelsa,∗,1T. WeberbF. Vialattec,dT. MushaeA.

where

tk =

(i,j)∈Ik

wti(tij − δti)

(i,j)∈Ik

wti

(A.2)

fk =

(i,j)∈Ik

wfi(fij − δfi)

(i,j)∈Ik

wfi, (A.3)

with wti = s−1ti and wfi = s−1

fi .In summary, marginalizing over v results in Ldel terms vol(S), and also

in terms of the form (A.1), where there is one such term for each cluster ofsize nk > 1.

The result of marginalizing over v may then be written as:

p(x, c, θ, ℓ) ∝ p(st)p(sf)(1 − λ vol(S))(λvol(S)pNd )Ldel

· (λpNd )Lp−Ltot

d (1 − pd)Ltot

·∏

k∈K

(i,j)∈Ik

N(

tij − tk; δti, sti

)

N(

fij − fk; δfi, sfi

)

, (A.4)

where we used the decomposition:

ℓ = L + Ldel. (A.5)

Now we marginalize over the length ℓ. The first term in the decomposi-tion (A.5) is fixed for given clustering c. Therefore, marginalizing p(x, c, θ, ℓ) (A.4)over ℓ is equivalent to marginalizing over Ldel:

p(x, c, θ) =∞∑

ℓ=0

p(x, c, θ, ℓ) =∞∑

Ldel=0

p(x, c, θ, ℓ) (A.6)

∝ p(st)p(sf)(1 − λ vol(S))∞∑

Ldel=0

(λvol(S)pNd )Ldel

· (λpNd )Lp−Ltot

d (1 − pd)Ltot

·∏

k∈K

(i,j)∈Ik

N(

tij − tk; δti, sti

)

N(

fij − fk; δfi, sfi

)

(A.7)

∝ p(st)p(sf)(1 − λ vol(S))1

1 − λvol(S)pNd

· (λpNd )Lp−Ltot

d (1 − pd)Ltot

·∏

k∈K

(i,j)∈Ik

N(

tij − tk; δti, sti

)

N(

fij − fk; δfi, sfi

)

. (A.8)

In (A.7) occurs a sum of a geometric series; since |λvol(S)pNd | = λvol(S)pN

d <1, we can apply the well-known formula for the sum of a geometric series,

43

Page 44: Quantifying Statistical Interdependence PART III: N ... · Quantifying Statistical Interdependence PART III: N > 2 Point Processes J. Dauwelsa,∗,1T. WeberbF. Vialattec,dT. MushaeA.

resulting in (A.8). By defining β and γ as in (20) and (21) respectively, weobtain statistical model (17).

B Appendix: Derivation of the SES Inference Algorithm

In this appendix, we derive the inference update (34) for N -variate SES.

The point estimates δ(κ)ti and δ

(κ)fi are the (sample) mean of the timing and

frequency offset respectively, computed between all events in xij and theirassociated cluster centers:

δ(κ)ti =

1

Li

Li∑

j=1

(

tij − tc(κ)ij

)

(B.1)

δ(κ)fi =

1

Li

Li∑

j=1

(

fij − fc(κ)ij

)

, (B.2)

where Li is the number of events in xi, and tc(κ)ij

and fc(κ)ij

are the coordinates

of the (inferred) hidden event vc(κ)ij

associated with xij . The latter hidden event

is inferred as the center of the cluster associated with xij . The coordinates tkand fk are computed as:

tk =

(i,j)∈I(κ)k

w(κ)ti (tij − δ

(κ)ti )

(i,j)∈I(κ)k

w(κ)ti

(B.3)

fk =

(i,j)∈I(κ)k

w(κ)fi (fij − δ

(κ)fi )

(i,j)∈I(κ)k

w(κ)fi

, (B.4)

with w(κ)ti =

(

s(κ)ti

)−1, w

(κ)fi =

(

s(κ)fi

)−1, and I(κ)

k is the index set of the set C(κ)k

of events in cluster k (as specified by c(κ)):

C(κ)k =

{

xij : c(κ)ij = k

}

and I(κ)k =

{

(i, j) : c(κ)ij = k

}

. (B.5)

44

Page 45: Quantifying Statistical Interdependence PART III: N ... · Quantifying Statistical Interdependence PART III: N > 2 Point Processes J. Dauwelsa,∗,1T. WeberbF. Vialattec,dT. MushaeA.

The estimates s(κ)ti and s

(κ)fi are obtained as:

s(κ)ti =

νtst0 + Li s(κ)ti,sample

νt + Li + 2(B.6)

s(κ)fi =

νfsf0 + Li s(κ)fi,sample

νf + Li + 2, (B.7)

where s(κ)ti,sample and s

(κ)fi,sample are computed as:

s(κ)ti,sample =

1

Li

Li∑

j=1

(

tij − tc(κ)ij

− δ(κ)ti

)2(B.8)

s(κ)fi,sample =

1

Li

Li∑

j=1

(

fij − fc(κ)ij

− δ(κ)fi

)2, (B.9)

and where tk and fk are given by (B.3) and (B.4).

Interestingly, the RHS of (B.1) and (B.2) depends on δ(κ)ti and δ

(κ)fi respectively,

through (B.3) and (B.4). Therefore the equalities (B.1) and (B.2) need to besolved numerically; the same holds for (B.6) and (B.7). Those expressionsmay be evaluated numerically by alternating the updates (B.3) and (B.4)

with (B.1) (B.2) (B.6) and (B.7), with as initial estimates δ(κ−1)ti , δ

(κ−1)fi , s

(κ−1)ti ,

and s(κ−1)fi . It can easily be shown that this procedure is guaranteed to converge

to a local extremum; indeed, it is equivalent to cyclic maximization, where oneconditional maximization is in θ (resulting in (B.1) (B.2) (B.6) and (B.7)) andthe other is in the parameters {(tk, fk)}k=1,...,L (resulting in (B.3) and (B.4)).

45

Page 46: Quantifying Statistical Interdependence PART III: N ... · Quantifying Statistical Interdependence PART III: N > 2 Point Processes J. Dauwelsa,∗,1T. WeberbF. Vialattec,dT. MushaeA.

References

Abeles M., Bergman H., Margalit E., and Vaadia E., 1993. Spatiotemporalfiring patterns in the frontal cortex of behaving monkeys. J. Neurophysiol70(4), 1629–1638.

Amari S., Nakahara H., Wu S., and Sakai Y., 2003. Synchronous firing andhigher-order interactions in neuron pool. Neural Computation 15, 127–142.

Buzsaki G., 2006. Rhythms of the brain. Oxford University Press.Chapman R., Nowlis G., McCrary J., Chapman J., Sandoval T., Guillily

M., Gardner M., Reilly L., 2007. Brain event-related potentials: diagnos-ing early-stage Alzheimer’s disease. Neurobiol. Aging 28, 194–201.

Cichocki A., Shishkin S., Musha T., Leonowicz Z., Asada T., and KurachiT., 2005. EEG filtering based on blind source separation (BSS) for earlydiagnosis of Alzheimer’s disease. Clin. Neurophys 116, 729–37.

Dauwels J., Vialatte F., Rutkowski T., and Cichocki A., 2008. MeasuringNeural Synchrony by Message Passing. In Advances in Neural InformationProcessing Systems 20, Edited by Platt J., Koller D., Singer Y., Roweis S.,Cambridge, MA: MIT Press 361–368.

Dauwels J., Tsukada Y., Sakumura Y., Ishii S., Aoki K., Nakamura T., Mat-suda M., Vialatte F., and Cichocki A., 2008. On the synchrony of mor-phological and molecular signaling events in cell migration, InternationalConference on Neural Information Processing (ICONIP), Lecture Notes inComputer Science (LNCS) 5506: 469–477.

Dauwels J., Vialatte F., Weber T., and Cichocki A., 2009a. Quantifying Sta-tistical Interdependence by Message Passing on Graphs: I. One-DimensionalPoint Processes, Neural Computation 21(8): 2152–2202.

Dauwels J., Vialatte F., Weber T., and Cichocki A., 2009b. Quantifying Statis-tical Interdependence by Message Passing on Graphs: II. Multi-DimensionalPoint Processes, Neural Computation 21(8): 2203–2268.

Dauwels J., Vialatte F., and Cichocki A., 2010a. A comparative study of syn-chrony measures for the early diagnosis of Alzheimer’s disease based onEEG, Neuroimage 49:668–693.

Dauwels J., Vialatte F., and Cichocki A., 2010b. Diagnosis of Alzheimer’s Dis-ease from EEG Signals: Where Are We Standing?, Curr Alzheimer Research,Epub May 11 2010.

Delprat N., Escudie B., Guillemain P., Kronland-Martinet R., TchamitchianP., and Torresani B., 1992. Asymptotic wavelet and Gabor analysis: ex-traction of instantaneous frequencies. IEEE Trans. Information Theory 38,644–664.

Gutkin B. S. and Ermentrout G. B., 1998. Dynamics of membrane excitabil-ity determine interspike interval variability: a link between spike genera-tion mechanisms and cortical spike train statistics. Neural Computation 10,1047–1065.

Efron, B. and Tibshirani, R., 1993. An Introduction to the Bootstrap, Chapman& Hall/CRC.

46

Page 47: Quantifying Statistical Interdependence PART III: N ... · Quantifying Statistical Interdependence PART III: N > 2 Point Processes J. Dauwelsa,∗,1T. WeberbF. Vialattec,dT. MushaeA.

Frey B. and Dueck D., 2007. Clustering by passing messages between datapoints. Science 315:5814, 972–976.

Givoni I. and Frey B. J., 2009. A binary variable model for affinity propagation.Neural Computation 21:6, 1589–1600.

Goupillaud P., Grossman A., and Morlet J., 1984. Cycle-octave and relatedtransforms in seismic signal analysis. Geoexploration 23, 85–102.

Harris F.J., 2004. Multirate signal processing for communication systems. Up-per Saddle River, NJ: Prentice Hall PTR.

Herrmann C. S., Grigutsch M., and Busch N. A., 2005. EEG oscillations andwavelet analysis. In: Handy, T. (ed.), Event-Related Potentials: a MethodsHandbook, Cambridge, MIT Press, 229–259.

Hogan M. J., Swanwick G. R. J., Kaiser J., Rowan M., and Lawlor B., 2003.Memory-related EEG power and coherence reductions in mild Alzheimer’sdisease. Int. J. Psychophysiol. 49.

Jeong J., 2004. EEG dynamics in patients with Alzheimer’s disease. ClinicalNeurophysiology 115, 1490–1505.

Lashkari D. and Golland P., 2008. Convex clustering with exemplar-basedmodels. In Advances in Neural Information Processing Systems 20, Editedby Platt J., Koller D., Singer Y., Roweis S., Cambridge, MA: MIT Press361–368.

Matsuda H., 2001. Cerebral blood flow and metabolic abnormalities inAlzheimer’s disease. Ann. Nucl. Med. 15, 85–92.

Morris C. and Lecar H., 1981. Voltage oscillations in the barnacle giant musclefiber. Biophys. J. 35, 193–213.

Musha T., Asada T., Yamashita F., Kinoshita T., Chen Z., Matsuda H.,Uno M., Shankle W.R., 2002. A new EEG method for estimating corti-cal neuronal impairment that is sensitive to early stage Alzheimers disease.Clin. Neurophysiol 113, 1052–8.

Nunez P. and Srinivasan R., 2006. Electric Fields of the Brain: The Neuro-physics of EEG. Oxford University Press.

Pereda E., Quiroga R. Q., and Bhattacharya J., 2005. Nonlinear multivariateanalysis of neurophsyiological signals. Progress in Neurobiology 77, 1–37.

Quiroga R. Q., Kraskov A., Kreuz T., and Grassberger P., 2002. Performanceof different synchronization measures in real data: a case study on EEGsignals. Physical Review E 65.

Robinson H. P. C., 2003. The biophysical basis of firing variability in corticalneurons, Chapter 6 in Computational Neuroscience: A Comprehensive Ap-proach, Mathematical Biology & Medicine Series, Edited By Jianfeng Feng,Chapman & Hall/CRC.

Singer W., 2001. Consciousness and the binding problem, 2001. Annals of theNew York Academy of Sciences 929, 123–146.

Stam C. J., 2005. Nonlinear dynamical analysis of EEG and MEG: review ofan emerging field. Clinical Neurophysiology 116, 2266–2301.

Tallon-Baudry C., Bertrand O., Delpuech C., and Pernier J., 1996. Stimulusspecificity of phase-locked and non-phase-locked 40Hz visual responses in

47

Page 48: Quantifying Statistical Interdependence PART III: N ... · Quantifying Statistical Interdependence PART III: N > 2 Point Processes J. Dauwelsa,∗,1T. WeberbF. Vialattec,dT. MushaeA.

human. Journal of Neuroscience 16, 4240–4249.Tateno T. and Pakdaman K., 2004. Random dynamics of the Morris-Lecar

neural model. Chaos 14(3).Toups J., Fellous J.-M., Thomas P., Sejnowski T., and Tiesinga P., 2011.

Finding the event structure of neuronal spike trains, Neural Computation23(9), 2169–2208.

Tsumoto K., Kitajima H., Yoshinaga T., Aihara K., and Kawakami H., 2006.Bifurcations in Morris-Lecar neuron model. Neurocomputing 69, 293–316.

Uhlhaas P. and Singer W., “Neural synchrony in brain disorders: relevance forcognitive dysfunctions and pathophysiology,” Neuron, 52:155–168, 2006.

Varela F., Lachaux J. P., Rodriguez E., and Martinerie J., 2001. The brainweb:phase synchronization and large-scale integration. Nature Reviews Neuro-science 2(4), 229–39.

Vialatte F., Cichocki A., Dreyfus G., Musha T., Rutkowski T. M., and Ger-vais R., 2005. Blind source separation and sparse bump modelling of time-frequency representation of EEG signals: new tools for early detection ofAlzheimer’s disease. Proc. IEEE Workshop on Machine Learning for SignalProcessing, 27–32.

Vialatte F., 2005. Modelisation en bosses pour l’analyse des motifs oscilla-toires reproductibles dans l’activite de populations neuronales : applicationsa l’apprentissage olfactif chez l’animal et a la detection precoce de la maladied’Alzheimer, PhD Thesis, Paris VI University, Paris.

Vialatte F., Martin C., Dubois R., Haddad J., Quenet B., Gervais R., andDreyfus G., 2007. A machine learning approach to the analysis of time-frequency maps, and its application to neural dynamics. Neural Networks20, 194–209.

Victor J. D. and Purpura K. P., 1997. Metric-space analysis of spike trains: the-ory, algorithms, and application. Network: Comput. Neural Systems 8(17),127–164.

Womelsdorf T., Schoffelen J.M., Oostenveld R., Singer W., Desimone R., EngelA.K., and Fries P., “Modulation of neuronal interactions through neuronalsynchronization,” Science, 316:1609–1612.

48