14
IEEE TRANSACTIONS ON SIGNAL PROCESSING, VOL. 66, NO. 22, NOVEMBER 15, 2018 5803 Statistical Tests for Detecting Granger Causality Ribhu Chopra , Member, IEEE, Chandra Ramabhadra Murthy , Senior Member, IEEE, and Govindan Rangarajan Abstract—Detection of a causal relationship between two or more sets of data is an important problem across various scien- tific disciplines. The Granger causality index and its derivatives are important metrics developed and used for this purpose. How- ever, the test statistics based on these metrics ignore the effect of practical measurement impairments such as subsampling, addi- tive noise, and finite sample effects. In this paper, we model the problem of detecting a causal relationship between two time series as a binary hypothesis test with the null and alternate hypotheses corresponding to the absence and presence of a causal relationship, respectively. We derive the distribution of the test statistic under the two hypotheses and show that measurement impairments can lead to suppression of a causal relationship between the signals, as well as false detection of a causal relationship, where there is none. We also use the derived results to propose two alternative test statistics for causality detection. These detectors are analytically tractable, which allows us to design the detection threshold and determine the number of samples required to achieve a given missed detection and false alarm rate. Finally, we validate the derived results using extensive Monte Carlo simulations as well as experiments based on real-world data, and illustrate the dependence of detection per- formance of the conventional and proposed causality detectors on parameters such as the additive noise variance and the strength of the causal relationship. Index Terms—Causality detection, granger causality, perfor- mance analysis, hypothesis testing. I. INTRODUCTION A. Motivation C ONFIRMING the presence or absence of a causative re- lationship between different observed phenomena is an important problem across various disciplines of physical, bio- logical, and social sciences. It has been argued that if the phe- nomena of interest can be represented as a time series, then the corresponding relationship between them can be quantified by directed mutual information [1]. It is also known that for Manuscript received December 7, 2017; revised May 21, 2018, July 4, 2018, and August 12, 2018; accepted September 14, 2018. Date of publication Septem- ber 26, 2018; date of current version October 8, 2018. The associate editor coordinating the review of this manuscript and approving it for publication was Dr. Paolo Braca. The work of C. R. Murthy was supported in part by a re- search grant from the Aerospace Network Research Consortium and in part by the Young Faculty Research Fellowship from the Ministry of Electronics and Information Technology, Government of India. The work of G. Rangarajan was supported in part by research grants from J. C. Bose National Fellowship, DST Centre for Mathematical Biology Phase II, and in part by UGC Centre for Advanced Study. (Corresponding author: Chandra Ramabhadra Murthy). R. Chopra is with the Indian Institute of Technology Guwahati, Assam 781039, India (e-mail:, [email protected]). C. R. Murthy and G. Rangarajan are with the Indian Institute of Science, Bangalore 560012, India (e-mail:, [email protected]; [email protected]). Color versions of one or more of the figures in this paper are available online at http://ieeexplore.ieee.org. Digital Object Identifier 10.1109/TSP.2018.2872004 Gaussian distributed time series, the directed mutual informa- tion reduces to the logarithm of ratio of the prediction error variances of the time series of interest with and without in- cluding the purported causing time series into the prediction model [2]. This quantity, known as the Granger causality in- dex (GCI) [2], [3], has been applied extensively for the detection of causal relationships across numerous research areas such as neuroscience [4]–[11], physics [12], [13], climate science [14], econometrics [15], etc. Most of the present studies on Granger causality assume that the sampling process is noiseless and the sampling frequency is above the Nyquist rate. As a consequence, the tests are not designed to achieve a given false alarm/detection rate. It is also typically assumed that the second-order statis- tics of the signals of interest are perfectly known. However, in practice, the causality detector has to estimate the second- order statistics from a finite number of possibly undersampled and noisy measurements of the signals of interest. In this pa- per, our main focus is on the quantification of the effects of these measurement inaccuracies on the performance of Granger causality detectors. This, in turn, allows us to design the detector to achieve a desired false alarm/detection probability. The phenomenon of undersampling arises mainly due to the inability of the signal acquisition unit to sample the signal of interest above the Nyquist rate. As shown later in the paper, under-sampling can lead to a weakening or even a complete suppression of the causal relationship between two signals. In addition to this, additive noise can lead to suppression of an existing causal relationship, as well as may cause a false de- tection [16], [17]. The effect of these inaccuracies is further compounded by the fact that the second-order statistics of the signal of interest, required to compute the GCI, are not known, and have to be estimated using the acquired samples. Therefore, finite sample effects, in addition to undersampling and addi- tive noise, also contribute to errors in the detection of a causal relationship between the signals of interest. In this paper, we model the problem of detection of Granger causality as a binary hypothesis test, and evaluate the effects of the measurement im- pairments listed above on the probabilities of detection and false alarm. B. Related Work In the original formulation of Granger causality in [2], a time series y[n] is said to Granger cause another time series x[n], when the past samples of the y[n] can be used to improve the estimate of the x[n] over ‘what can be predicted using all other information in the universe’, conventionally considered to be the past samples of x[n]. Therefore, a causal relation- ship between x[n] and y[n] can be inferred by comparing the 1053-587X © 2018 IEEE. Personal use is permitted, but republication/redistribution requires IEEE permission. See http://www.ieee.org/publications standards/publications/rights/index.html for more information.

Statistical Tests for Detecting Granger Causality · reversed Granger causality is the most noise resilient. The problem of sub-sampling in Granger causality detection hasbeenstudiedintheliterature[27]–[34].In[27],[28],theau-thors

  • Upload
    others

  • View
    7

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Statistical Tests for Detecting Granger Causality · reversed Granger causality is the most noise resilient. The problem of sub-sampling in Granger causality detection hasbeenstudiedintheliterature[27]–[34].In[27],[28],theau-thors

IEEE TRANSACTIONS ON SIGNAL PROCESSING, VOL. 66, NO. 22, NOVEMBER 15, 2018 5803

Statistical Tests for Detecting Granger CausalityRibhu Chopra , Member, IEEE, Chandra Ramabhadra Murthy , Senior Member, IEEE, and Govindan Rangarajan

Abstract—Detection of a causal relationship between two ormore sets of data is an important problem across various scien-tific disciplines. The Granger causality index and its derivativesare important metrics developed and used for this purpose. How-ever, the test statistics based on these metrics ignore the effect ofpractical measurement impairments such as subsampling, addi-tive noise, and finite sample effects. In this paper, we model theproblem of detecting a causal relationship between two time seriesas a binary hypothesis test with the null and alternate hypothesescorresponding to the absence and presence of a causal relationship,respectively. We derive the distribution of the test statistic under thetwo hypotheses and show that measurement impairments can leadto suppression of a causal relationship between the signals, as wellas false detection of a causal relationship, where there is none. Wealso use the derived results to propose two alternative test statisticsfor causality detection. These detectors are analytically tractable,which allows us to design the detection threshold and determinethe number of samples required to achieve a given missed detectionand false alarm rate. Finally, we validate the derived results usingextensive Monte Carlo simulations as well as experiments basedon real-world data, and illustrate the dependence of detection per-formance of the conventional and proposed causality detectors onparameters such as the additive noise variance and the strength ofthe causal relationship.

Index Terms—Causality detection, granger causality, perfor-mance analysis, hypothesis testing.

I. INTRODUCTION

A. Motivation

CONFIRMING the presence or absence of a causative re-lationship between different observed phenomena is an

important problem across various disciplines of physical, bio-logical, and social sciences. It has been argued that if the phe-nomena of interest can be represented as a time series, thenthe corresponding relationship between them can be quantifiedby directed mutual information [1]. It is also known that for

Manuscript received December 7, 2017; revised May 21, 2018, July 4, 2018,and August 12, 2018; accepted September 14, 2018. Date of publication Septem-ber 26, 2018; date of current version October 8, 2018. The associate editorcoordinating the review of this manuscript and approving it for publication wasDr. Paolo Braca. The work of C. R. Murthy was supported in part by a re-search grant from the Aerospace Network Research Consortium and in partby the Young Faculty Research Fellowship from the Ministry of Electronicsand Information Technology, Government of India. The work of G. Rangarajanwas supported in part by research grants from J. C. Bose National Fellowship,DST Centre for Mathematical Biology Phase II, and in part by UGC Centre forAdvanced Study. (Corresponding author: Chandra Ramabhadra Murthy).

R. Chopra is with the Indian Institute of Technology Guwahati, Assam781039, India (e-mail:,[email protected]).

C. R. Murthy and G. Rangarajan are with the Indian Institute of Science,Bangalore 560012, India (e-mail:,[email protected]; [email protected]).

Color versions of one or more of the figures in this paper are available onlineat http://ieeexplore.ieee.org.

Digital Object Identifier 10.1109/TSP.2018.2872004

Gaussian distributed time series, the directed mutual informa-tion reduces to the logarithm of ratio of the prediction errorvariances of the time series of interest with and without in-cluding the purported causing time series into the predictionmodel [2]. This quantity, known as the Granger causality in-dex (GCI) [2], [3], has been applied extensively for the detectionof causal relationships across numerous research areas such asneuroscience [4]–[11], physics [12], [13], climate science [14],econometrics [15], etc. Most of the present studies on Grangercausality assume that the sampling process is noiseless and thesampling frequency is above the Nyquist rate. As a consequence,the tests are not designed to achieve a given false alarm/detectionrate. It is also typically assumed that the second-order statis-tics of the signals of interest are perfectly known. However,in practice, the causality detector has to estimate the second-order statistics from a finite number of possibly undersampledand noisy measurements of the signals of interest. In this pa-per, our main focus is on the quantification of the effects ofthese measurement inaccuracies on the performance of Grangercausality detectors. This, in turn, allows us to design the detectorto achieve a desired false alarm/detection probability.

The phenomenon of undersampling arises mainly due to theinability of the signal acquisition unit to sample the signal ofinterest above the Nyquist rate. As shown later in the paper,under-sampling can lead to a weakening or even a completesuppression of the causal relationship between two signals. Inaddition to this, additive noise can lead to suppression of anexisting causal relationship, as well as may cause a false de-tection [16], [17]. The effect of these inaccuracies is furthercompounded by the fact that the second-order statistics of thesignal of interest, required to compute the GCI, are not known,and have to be estimated using the acquired samples. Therefore,finite sample effects, in addition to undersampling and addi-tive noise, also contribute to errors in the detection of a causalrelationship between the signals of interest. In this paper, wemodel the problem of detection of Granger causality as a binaryhypothesis test, and evaluate the effects of the measurement im-pairments listed above on the probabilities of detection and falsealarm.

B. Related Work

In the original formulation of Granger causality in [2], atime series y[n] is said to Granger cause another time seriesx[n], when the past samples of the y[n] can be used to improvethe estimate of the x[n] over ‘what can be predicted using allother information in the universe’, conventionally consideredto be the past samples of x[n]. Therefore, a causal relation-ship between x[n] and y[n] can be inferred by comparing the

1053-587X © 2018 IEEE. Personal use is permitted, but republication/redistribution requires IEEE permission.See http://www.ieee.org/publications standards/publications/rights/index.html for more information.

Page 2: Statistical Tests for Detecting Granger Causality · reversed Granger causality is the most noise resilient. The problem of sub-sampling in Granger causality detection hasbeenstudiedintheliterature[27]–[34].In[27],[28],theau-thors

5804 IEEE TRANSACTIONS ON SIGNAL PROCESSING, VOL. 66, NO. 22, NOVEMBER 15, 2018

mean squared prediction error for x[n] with and without in-cluding y[n] in the prediction model [18]. The prediction modelconsidering only the past samples of x[n] is known as the re-strictive model (R-model), whereas the one including the ef-fects of y[n] is known as the unrestricted model (U-model). Inaddition to the GCI, several metrics for quantifying Grangercausality have been studied [19], [20]. The system model withtwo time series [2] has been also generalized to study causalrelationships between more than two time series [21]. However,all these studies assume the samples are noiseless, and ignorethe finite sample effects on the second-order statistics beingestimated.

The effect of noise on Granger causality was first discussedin [22], and was mathematically analyzed in [16] for a bivariateVAR-1 process. These results were extended to a generalizedVAR(p) (pth order autoregressive process) in [17], where aKalman filtering based technique to alleviate the detrimentaleffect of noise was also developed. An alternative test statisticfor causality detection, known as the phase slope index, wasintroduced in [23] and was shown to be more resilient to noise,as compared to the GCI. Another metric for causality, known asthe time reversed Granger causality, has been discussed in [18],[24], [25]. The authors in [26] used simulation techniques tocompare the robustness of GCI, time reversed Granger causality,and phase slope index to correlated noise, and found that timereversed Granger causality is the most noise resilient.

The problem of sub-sampling in Granger causality detectionhas been studied in the literature [27]–[34]. In [27], [28], the au-thors study the effects of the choice of the sampling frequencyon the detectability of a causal relationship between neurolog-ical signals. In [28], these derived results are used to identifyfavorable sampling rates for detecting Granger causality. In [29],two methods, based on the expectation maximization (EM) al-gorithm and the variational inference framework, are proposedto detect causal relationships from low resolution data, and theirapplicability to simulated and practical data is tested. The prob-lem of causal inference from an under-sampled time series byusing a general purpose Boolean constraint solver has been dis-cussed in [32]. It is shown that the method considered in [32]is independent of the modeling parameters, and can be conve-niently scaled to any number of variables. In [33], the authorspropose three sampling rate agnostic algorithms for the recov-ery of a causal relationship between observed time series. Theauthors in [34] incorporate the effect of both down-samplingand additive noise in their data model. However, none of thesestudies consider the effect of finite samples on the detectionperformance, or model the problem of detection of Grangercausality as a binary hypothesis testing problem. They also donot analyze the effects of these physical impairments on theprobabilities of detection and false alarm.

The authors in [1] consider the related problem of estimationof directed mutual information between two time series, using afinite number of samples. Here, the test for a causal relationshipbetween two time series is modeled as a binary hypothesis test,and it is shown that the probabilities of missed detection andfalse alarm asymptotically converge to zero as the number ofsamples is increased. In this paper, we propose to extend the

binary hypothesis testing framework for causality detection toGCI.

C. Contributions

In the present work, we consider the problem of detectinga causal relationship between two possibly sub-sampled timeseries, in the presence of additive white Gaussian noise. Wemodel the time series as two independent pth order autoregres-sive (AR(p)) processes in the absence of a causal relation, and asa bivariate vector autoregressive (VAR(p)) process in the pres-ence of a causal relationship. We model the test for causalityas a binary hypothesis testing problem with the null hypothe-sis corresponding to the absence of a causal relationship, andthe alternate hypothesis to its presence. The system model isdescribed in detail in Section II. Following this, the main ob-jective of this paper is to evaluate the detection performance ofGCI for finite samples of the two time series of interest. Ourcontributions in this direction are as follows:

1) We derive the effects of down-sampling on the causalrelationship between two signals. (See Section III.)

2) We derive the GCI for the down-sampled time series inthe presence of additive noise assuming knowledge ofthe second-order statistics of the signals of interest. (SeeSection IV.)

3) We derive the statistics of the GCI under the two hypothe-ses for finite number of samples, and use these to derivethe probabilities of detection and false alarm. Our analysisaccounts for the estimation of the second-order statisticsof the two signals using a finite number of samples whenthe generative model is unknown. (See Section V.)

4) Based on the statistics of GCI, we propose two alter-native test statistics to detect Granger causality and de-rive the corresponding probabilities of detection and falsealarm. (See Section VI.)

5) Using detailed simulations and numerical experimentsbased on real-world data, we validate the derived the-ory and compare the performance of the two proposedtest statistics against the GCI. (See Section VII.)

The results derived in this paper can be used to characterizethe behavior of GCI and related detectors under different phys-ical conditions. In addition to this, the two alternative causalitydetectors discussed in this work are easy-to-use replacementsfor the GCI. We next describe the system model considered inthis work.

Notation: Boldface lowercase and uppercase letters repre-sent vectors and matrices, respectively. The kth column of Ais denoted by ak . (.)H represents the Hermitian operation on avector or a matrix, and (.)† represents the Moore-Penrose pseu-doinverse of a matrix. IK ,0K and OK represent the identitymatrix, the all-zero vector and the all-zero matrix of dimensionK, respectively. The �2 norm of a vector and the Frobeniusnorm of a matrix are denoted by ‖ · ‖2 and ‖ · ‖F , respectively.CN (μ, σ2) represents a circularly symmetric complex Gaussianrandom variable with mean μ and variance σ2 . E[·] and var(·)represent the mean and variance of a random variable. In gen-eral, x and x denote the MMSE estimate and the corresponding

Page 3: Statistical Tests for Detecting Granger Causality · reversed Granger causality is the most noise resilient. The problem of sub-sampling in Granger causality detection hasbeenstudiedintheliterature[27]–[34].In[27],[28],theau-thors

CHOPRA et al.: STATISTICAL TESTS FOR DETECTING GRANGER CAUSALITY 5805

TABLE IA LIST OF NOTATION FOLLOWED IN THE PAPER

estimation error of a random variable x. Table I introduces thedifferent symbols used in this paper.

II. SYSTEM MODEL

We consider two complex valued discrete time zero meanGaussian random processes c[n] and d[n] having variances σ2

c

and σ2d , respectively, and expressed in the form of a bivariate

VAR(q) model as

c[n] =q∑

k=1

a∗1,k c[n − k] +

q∑

k=1

a∗2,k d[n − k] + ηc [n],

d[n] =q∑

k=1

a∗3,k c[n − k] +

q∑

k=1

a∗4,k d[n − k] + ηd [n], (1)

with ai,k , i ∈ {1, 2, 3, 4} being the regression coefficients, andηc [n] and ηd [n] the innovation components in c[n] and d[n],respectively. We assume that c[n] and d[n] are sampled at theNyquist rate, and the temporally white innovation process vectorη[n] = [ηc [n] , ηd [n]]T is distributed as

η[n] ∼ CN(

0,

[σ2

ηc0

0 σ2ηd

]).

For simplicity, we assume that only a unidirectional couplingfrom d[n] to c[n] can exist, i.e., a3,k = 0 , for 1 ≤ k ≤ q.

Defining cq [n] � [c[n], c[n −1], . . . , c[n −q +1]]T , dq [n]�[d[n], d[n −1], . . . , d[n −q +1]]T , ai � [ai,1 , ai,2 , . . . , ai,q ]T ,i ∈ {1, 2, 3, 4}, we can write (1) as,

[c[n]d[n]

]=[aH

1 aH2

0Hq aH

4

] [cq [n − 1]dq [n − 1]

]+[

ηc [n]ηd [n]

], (2)

with 0q being the q dimensional all zero vector. Definingrcc [τ ] � E[c[n]c∗[n − τ ]], rcd [τ ] � E[c[n]d∗[n − τ ]], and sim-ilarly rdc [τ ] and rdd [τ ], we can write

rcc [τ ] =q∑

k=1

a∗1,k rcc [τ − k] +

q∑

k=1

a∗2,k rdc [τ − k] + σ2

ηcδ[τ ].

(3)Letting rcd,q [τ ] = [rcd [τ ], . . . , rcd [τ − q + 1]]T , and simi-

larly rcc,q [τ ], etc., the above becomes

rcc [τ ] = aH1 rcc,q [τ − 1] + aH

2 rdc,q [τ − 1] + σ2ηc

δ[τ ]. (4)

We can similarly write

rcd [τ ] = aH1 rcd,q [τ − 1] + aH

2 rdd,q [τ − 1], (5)

and

rdd [τ ] = aH4 rdd,q [τ − 1] + σ2

ηdδ[τ ]. (6)

The above equations, along with the information that rcc [0] =σ2

c , and rdd [0] = σ2d can be used to compute rcc [τ ], rcd [τ ], and

rdd [τ ] for different values of τ .Defining

Rcc,q [τ ] � E[cq [n]cH

q [n − τ ]], (7)

Rcd,q [τ ] � E[cq [n]dH

q [n − τ ]], (8)

and similarly defining Rdc,p [τ ] and Rdd,q [τ ], it can be shownusing the Weiner-Hopf equations [35] that

[a1

a2

]=

[Rcc,q [0] Rcd,q [0]Rdc,q [0] Rdd,q [0]

]−1 [rcc,q [1]rcd,q [1]

]. (9)

We can therefore calculate the regression coefficients from thecorrelation coefficients when the latter are not known. The inter-ested reader is referred to [35, Chapter 2] for a detailed derivationof the Wiener-Hopf equations. We can then use these to computethe estimation error variance as

σ2ηc

= σ2c −[rcc,q [1]rcd,q [1]

]H [Rcc,q [0] Rcd,q [0]Rdc,q [0] Rdd,q [0]

]−1 [rcc,q [1]rcd,q [1]

].

(10)Also, it can be observed from (2) that the causal dependence ofc[n] on d[n] is determined by the coefficient vector a2 , that is,there exists a causal relationship only if a2 �= 0.

Now, in the absence of a causal relationship, c[n] can alter-natively be expressed using a univariate AR(q) model,

c[n] = gH cq [n − 1] + ζ[n], (11)

Page 4: Statistical Tests for Detecting Granger Causality · reversed Granger causality is the most noise resilient. The problem of sub-sampling in Granger causality detection hasbeenstudiedintheliterature[27]–[34].In[27],[28],theau-thors

5806 IEEE TRANSACTIONS ON SIGNAL PROCESSING, VOL. 66, NO. 22, NOVEMBER 15, 2018

with ζ[n] being the zero mean temporally white complex Gaus-sian innovation component having a variance σ2

ζ . Consequently,

rcc [τ ] = gH rcc,q [τ − 1] + σ2ζ δ[τ ], (12)

g = R−1cc,q [0]rcc,q [1], (13)

and

σ2ζ = σ2

c − rHcc,q [1]R−1

cc,q [0]rcc,q [1]. (14)

It is to be noted that in the presence of a causal relationshipbetween c[n] and d[n] the mean squared prediction error for c[n]using the bivariate model, σ2

ηc, will be smaller than the mean

squared prediction error using the univariate signal model, σ2ζ ,

whereas, in the absence of such a causal relationship betweenthem, σ2

ζ = σ2ηc

.The GCI uses this property to quantify the causal relationship,

and is defined as [2]

TG � log

(σ2

ζ

σ2ηc

). (15)

In Granger causality, we say that a causal relationship existsbetween c[n] and d[n] if TG > 0, and no causal relationshipexists if TG = 0.

However, both c[n] and d[n] are assumed to be noiseless andsampled at the Nyquist rate. Moreover, it also assumes that theexact second-order statistics for both c[n] and d[n] are known.However, these assumptions do not hold in a practical data ac-quisition system. Therefore, in the subsequent sections, we in-troduce these imperfections, viz. down-sampling, measurementnoise, and finite sample effects, into a GCI based detector, andanalyze their impact on the detection performance. In the nextsection, we discuss the effect of down-sampling on the GCI.

III. THE EFFECT OF DOWN-SAMPLING ON GCI

Let u[n] and v[n] respectively be versions of c[n] and d[n]down-sampled by a factor M , such that,

u[n] = c[nM − l], v[n] = d[nM − l]. (16)

Since we use the second-order statistics of the observed widesense stationary random processes for detection, we can set theoffset parameter l to zero without loss of generality. We can nowexpress u[n] and v[n] in the form of a bivariate linear regressionprocess of order p as

[u[n]v[n]

]=

[bH

1 bH2

0Hp bH

4

][up [n − 1]vp [n − 1]

]+

[εu [n]εv [n]

], (17)

with bi , i ∈ {1, 2, 4} representing new regression coef-ficients, and E [u[n − k]ε∗u [n]] = 0, E [v[n − k]ε∗v [n]] = 0,E [u[n − k]ε∗v [n]] = 0, and E [v[n − k]ε∗u [n]] = 0 for k > 0. Itis important to note that the model order p in this case maybe different from the model order q considered in the previoussection due to down-sampling.

Since the GCI is a function of the second-order statistics ofc[n] and d[n], we need to derive the second-order statistics ofu[n] and v[n] in terms of c[n] and d[n] in order to quantify theeffect of down-sampling on the GCI.

Let ruv [τ ] = E[u[n]v∗[n − τ ]], ruv ,p [τ ] = E[up [n]v∗[n −τ ]], and Ruv ,p [τ ] = E[up [n]vH

p [n − τ ]], with the absenceof the index [τ ] indicating τ = 0. We can write, ruu [τ ] =E[c[nM ]c∗[nM − τM ]] = rcc [τM ]. Similarly, ruu,p [τ ] =[rcc [τM ], rcc [(τ + 1)M ], . . . , rcc [(τ + p)M ]]T , and

ruu [τ ] = bH1 ruu,p [τ − 1] + bH

2 rvu,p [τ − 1] + σ2ηc

δ[τ ], (18)

ruv [τ ] = bH1 ruv ,p [τ − 1] + bH

2 rvv ,p [τ − 1], (19)

rvv [τ ] = bH4 rvv ,p [τ − 1] + σ2

ηdδ[τ ], (20)

and

Ruu,p

=

⎢⎢⎢⎣

rcc [0] rcc [M ] . . . rcc [(p − 1)M ]rcc [M ] rcc [0] . . . rcc [(p − 2)M ]

......

. . ....

rcc [(p − 1)M ] rcc [(p − 2)M ] . . . rcc [0]

⎥⎥⎥⎦ .

(21)

The weight vectors b1 , b2 , and b4 can then be determined viathe Wiener-Hopf equations [35] as

[b1

b2

]=

[Ruu,p Ruv ,p

Rvu,p Rvv ,p

]−1 [ruu,p [1]ruv ,p [1]

], (22)

and b4 = R−1vv ,prvv [1]. (23)

We can then write,

εu [n] = u[n] − bH1 up [n − 1] − bH

2 vp [n − 1], (24)

εv [n] = y[n] − bH4 vp [n − 1]. (25)

Consequently,

rεu εu[0] = σ2

c −[ruu,p [1]ruv ,p [1]

]H [Ruu,p Ruv ,p

Rvu,p Rvv ,p

]−1 [ruu,p [1]ruv ,p [1]

],

(26)

rεv εv[0] = σ2

c − rHvv,p [1]R−1

vv ,prvv ,p [1]. (27)

rεu εv[0] = rcd [0]

− [rHuu,p [1] rH

uv,p [1]][Ruu,p Ruv ,p

Rvu,p Rvv ,p

]−1[ruv ,p [1]rvv ,p [1]

]

− rHuv,p [1]R−1

vv ,prvv ,p [1] +

[ruu,p [1]rvv ,p [1]

]H

×[Ruu,p Ruv ,p

Rvu,p Rvv ,p

]−1 [Ruv ,p

Rvv ,p

]R−1

vv ,prvv ,p [1].

(28)

Hence, the covariance matrix of the innovation process need notbe diagonal after down-sampling.

Similarly, the single variable linear regression model for thedown-sampled signal u[n] can be written as

u[n] = fH up [n − 1] + ξ[n], (29)

Page 5: Statistical Tests for Detecting Granger Causality · reversed Granger causality is the most noise resilient. The problem of sub-sampling in Granger causality detection hasbeenstudiedintheliterature[27]–[34].In[27],[28],theau-thors

CHOPRA et al.: STATISTICAL TESTS FOR DETECTING GRANGER CAUSALITY 5807

with ξ[n] being the innovation component. Therefore, we get

f = R−1uu,pruu,p [1], (30)

and rξξ [0] = σ2u − rH

uu,p [1]R−1uu,pruu,p [1]. (31)

A causal relationship between u[n] and v[n] can now be deter-mined by considering the ratio TG = log2(

rξ ξ [0]rε u ε u [0] ).

To illustrate the effect of down-sampling on the GCI, considerthe simple special case: c[n] = ad[n − 1] + ηc [n], and d[n] =ηd [n]. Letting M = 2, we can write

u[n] = c[2n] = aηd [2n − 1] + ηc [2n], (32)

v[n] = d[2n] = ηd [2n]. (33)

Consequently, ruu [1] = rcc [2] = 0, ruv [1] = rcd [2] = 0, ruv [0]= rcd [0] = 0, and rξξ [0] = rεx εx

[0] = σ2c . We see that there ex-

ists a causal relationship between c[n] and d[n] for any a �= 0,but no causal relationship exists between u[n] and v[n]. Thus,down-sampling can suppress an existing causal relationship be-tween two signals. However, the degree of suppression of thecausal relationship depends on the structure of the VAR modelfollowed by the signals. We next analyze the effect of additivenoise on the performance of the GCI.

IV. GCI WITH DOWNSAMPLING AND ADDITIVE NOISE

Let us define x[n] and y[n] as the down-sampled signalsu[n] and v[n] corrupted by AWGN, such that,

x[n] = u[n] + νx [n],

y[n] = v[n] + νy [n], (34)

where νx [n] ∼ CN (0, σ2νx

) and νy [n] ∼ CN (0, σ2νy

).Using (17), we can write,[x[n]y[n]

]=

[bH

1 bH2

0Hp bH

4

][up [n − 1]vp [n − 1]

]+

[εu [n] + νx [n]εv [n] + νy [n]

]. (35)

If a causal relationship exists between c[n] and d[n], and is pre-served in u[n] and v[n] after down-sampling, it should also existbetween x[n] and y[n], and therefore, the past samples of x[n]and y[n] can be used to predict x[n] better than its prediction byusing the past samples of x[n] alone. Letting z[n] = [x[n] y[n]]T

and zp [n] = [xTp [n] yT

p [n]]T , we can express x[n] as

x[n] = wH zp [n − 1] + ϕ2 [n], (36)

with ϕ2 [n] being the prediction error for the bivariate VARmodel, and w being the weight vector minimizing the meansquared value of ϕ2 [n].

Defining σ2ϕ2

� E[|ϕ2 [n]|2 ] = E[|x[n] − wH zp [n − 1]|2 ], itcan be shown that [35],

σ2ϕ2

= σ2x − wH rzx,p [1] − rH

zx,p [1]w + wH Rzz ,pw, (37)

where

σ2x = E[x[n]x∗[n]] = σ2

c + σ2νx

, (38)

rzx,p [τ ] = E[zp [n]x∗[n − τ ]] = [rHxx [τ ] rH

yx [τ ]]H , (39)

and

Rzz ,p = E[zp [n]zH

p [n]]. (40)

Since the additive noise is independent of both u[n] and v[n],it can be shown that

rxx [τ ] = ruu [τ ] + σ2νx

δ[τ ], (41)

rxy [τ ] = ruv [τ ], (42)

ryy [τ ] = rvv [τ ] + σ2νy

δ[τ ]. (43)

Substituting equations (18)–(20) into (41)–(43) we obtain

rxx [τ ]=bH1 ruu,p [τ −1] + bH

2 rvu,p [τ −1] + (σ2εu

+σ2νx

)δ[τ ],(44)

rxy [τ ] = bH1 ruv ,p [τ − 1] + bH

2 rvv ,p [τ − 1], (45)

ryy [τ ] = bH4 rvv ,p [τ − 1] + (σ2

εv+ σ2

νy)δ[τ ], (46)

and similarly,

Rzz ,p =[Rxx,p Rxy ,p

Ryx,p Ryy ,p

]

=[Ruu,p Ruv ,p

Rvu,p Rvv ,p

]+[

σ2νx

Ip 0p

0p σ2νy

Ip

]. (47)

The optimal weight vector minimizing σ2ϕ2

can be obtainedvia the Wiener-Hopf equations as

w = R−1zz ,przx,p [1]. (48)

The minimized value of σ2ϕ2

is

σ2ϕ2

= σ2x − rH

zx,p [1]R−1zz ,przx,p [1]. (49)

However, in case no causal relationship exists between c[n]and d[n], and consequently between x[n] and y[n], x[n] can beexpressed using the single variable AR model as

x[n] = hH xp [n − 1] + ϕ1 [n], (50)

with ϕ1 [n] being the single variable prediction error, and h beingthe optimal weight vector, computed as

h = R−1xx,prxx,p [1]. (51)

Similar to the two variable case, the minimized value of theprediction error takes the form

σ2ϕ1

= σ2x − rH

xx,p [1]R−1xx,prxx,p [1]. (52)

The GCI therefore becomes

TG = log2

(σ2

ϕ1

σ2ϕ2

)= log2

(σ2

x − rHxx,p [1]R−1

xx,prxx,p [1]σ2

x − rHzx,p [1]R−1

zz ,przx,p [1]

).

(53)To demonstrate the effects of noise on the GCI, we again

consider our running example, d[n] = εv [n], c[n] = u[n] =ad[n − 1] + bd[n − 2] + εc [n]. Downsampling these by a fac-tor M = 2, we obtain, v[n] = εv [n], u[n] = bv[n − 1] + εu [n],with, σ2

v = σ2εv

, and σ2u = |b|2σ2

εv+ σ2

εu, and consequently,

σ2x = |b|2σ2

εv+ σ2

εu+ σ2

νx, σ2

y = σ2εv

+ σ2νy

, (54)

Page 6: Statistical Tests for Detecting Granger Causality · reversed Granger causality is the most noise resilient. The problem of sub-sampling in Granger causality detection hasbeenstudiedintheliterature[27]–[34].In[27],[28],theau-thors

5808 IEEE TRANSACTIONS ON SIGNAL PROCESSING, VOL. 66, NO. 22, NOVEMBER 15, 2018

rxy [0] = 0, rxy [1] = bσ2εv

. (55)

Therefore,

Rzz ,1 =[ |b|2σ2

εv+ σ2

εu+ σ2

νx0

0 σ2εv

+ σ2νy

], (56)

σ2ϕ2

= σ2εu

+ σ2νx

+|b|2σ2

εvσ2

νy

σ2εv

+ σ2νy

, (57)

and

σ2ϕ1

= σ2x = |b|2σ2

εv+ σ2

εu+ σ2

νx. (58)

We can simplify the GCI as

TG = log2

⎜⎝|b|2σ2

εv+ σ2

εu+ σ2

νx

|b|2σ2εv

(σ 2

ν y

σ 2ε v +σ 2

ν y

)+ σ2

εu+ σ2

νx

⎟⎠ . (59)

In the absence of additive noise in both x[n] and y[n], TG

reduces to TG = log2(|b|2 σ 2

ε v +σ 2ε u

σ 2ε u

), as in the previous section.

However, as either σ2νx

, or σ2νy

get large, the second term in theexpression for TG converges to zero. Thus, additive noise canalso lead to the suppression of a causal relationship between twosignals.

The discussion till now assumed that the second-order statis-tics of both the signals of interest are available at the detec-tor. However, in practice, we have to estimate the second-orderstatistics using a finite number of samples of x[n] and y[n].The effects of the use of estimated second-order statistics onthe detection of Granger causality in an under-sampled noisyenvironment are discussed in the next section.

V. FINITE SAMPLE EFFECTS ON THE GCI

In case only N samples of each of x[n] and y[n] are observed,and the generative model is unknown, we need to estimate theoptimal predictors for the single and two variable cases, thecorresponding prediction error variances, and the GCI or anequivalent test statistic using these samples. In this section, wefirst discuss the properties of the optimal predictor weights forthe one variable and two variable prediction models. Then, weuse these weights to calculate the statistics of the finite sampleestimates of the corresponding mean square prediction errors,and finally derive expressions for the performance of GCI withfinite number of observations.

A. Properties of Optimal Predictor Weights

It is known that given a finite number of observations, theleast squares estimate is the best linear unbiased estimator forx[n] [35]. Letting x[1], . . . , x[N ] be the N available samplesof x[n], y[1], . . . , y[N ] be the N available samples of y[n], andx[n] = wH zp [n − 1] be the least squares (LS) estimate of x[n]generated using the two variable model, we can write

xN−p [N ] = xN−p [N ] + xN−p [N ], (60)

where xN−p [N ] is the error term orthogonal to the measurementspace. Also, w is the least squares estimate of the weight vector,

given as [35]

w = Z†p [N ]xN−p [N ]. (61)

Here, w is a finite sample estimate of the true weight vectorw with Zp [N ] = [zp [N − 1], zp [N − 2], . . . , zp [p]]H , and ()†

represents the Moore-Penrose inverse of a matrix. Since the in-novation process and the measurement noise are white and zeromean, it can be shown that w has the following properties [35]:

1) w is an unbiased estimator of w.

2) The covariance matrix of w is cov(w) =σ 2

ϕ 2N−p R

−1zz ,p .

Therefore, w = w + w, with w ∼ CN(0,

σ 2ϕ 2

N−p R−1zz ,p

).

Similarly, considering the single variable prediction model,we can writexN −p [N ] in terms of its projection, xN −p [N ] on thedata matrix, Xp [N ] = [xp [N − 1],xp [N − 2], . . . ,xp [p]]H ,and the error component xN −p [N ] = xN −p [N ] − xN −p [N ].Again, xN −p [N ] = Xp [N ]h with h = X†

p [N ]xN −p [N ], and

h ∼ CN(

h,σ2

ϕ1

N − pR−1

xx,p

). (62)

We next use the above properties of the estimated regressionweights to derive the statistics of the finite sample estimate ofthe mean squared error for the two prediction models.

B. Mean Squared Prediction Errors

First, considering the prediction error for the two variablemodel, we can express x[n] as

x[n] = x[n] − x[n]

= x[n] − wH zK [n − 1] − wH zK [n − 1]. (63)

Since the innovation process of u[n] and the additive noise arewhite, x[n] is zero mean i.i.d. Gaussian with variance

E[|x[n]|2 ] = σ2ϕ2

+ E[wH zp [n − 1]zHp [n − 1]w]

− 2�{E[wH zp [n − 1]x∗[n]]}− 2�{E[wH zp [n − 1]zH

p [n − 1]w]}. (64)

It is common in the adaptive filtering literature to assume theweight error vector to be independent of the regression vectorsas well as the desired output [35]. Hence,

E[wH zp [n − 1]x∗[n]] = E[wH zp [n − 1]zHp [n − 1]w] = 0,

(65)and

E[wH zp [n − 1]zHp [n − 1]w]

= Tr(E[zp [n − 1]zHp [n − 1]E[wwH ])

= Tr

(Rzz ,p

σ2ϕ2

N − pR−1

zz ,p

)= σ2

ϕ2

2p

N − p. (66)

Substituting these in (64), we can write

E[|x[n]|2 ] = σ2ϕ2

(1 +

2p

N − p

). (67)

Page 7: Statistical Tests for Detecting Granger Causality · reversed Granger causality is the most noise resilient. The problem of sub-sampling in Granger causality detection hasbeenstudiedintheliterature[27]–[34].In[27],[28],theau-thors

CHOPRA et al.: STATISTICAL TESTS FOR DETECTING GRANGER CAUSALITY 5809

We can similarly argue, for the one variable model, that x[n]is also zero mean i.i.d. Gaussian distributed, with

E[|x[n]|2 ] = σ2ϕ1

(1 +

p

N − p

). (68)

Thus, the estimates of the prediction error variances from theone variable and two variable regression models of the systemcan be expressed as

t1 =1

N − p

N∑

n=p+1

|x[n]|2 , (69)

t2 =1

N − p

N∑

n=p+1

|x[n]|2 . (70)

These can now be used to calculate the GCI as

TG = log2

(t1t2

), (71)

However, since TG is a logarithm of ratios of random variables,its statistics become hard to determine. Instead, we define TE �2−TG = t2

t1as a test statistic, for simplicity of analysis.

Since x[n] and x[n] are the measurement errors for the sameprocess, t1 and t2 cannot be considered independent. However,since both t1 and t2 are sums of a large number of i.i.d. randomvariables, we can use the central limit theorem to approximatethese as Gaussian r.v.s [36]. Note that this approximation is validwhen the number of samples N is much larger than the modelorder p. Now

E[t2 ] = σ2ϕ2

(1 +

2p

N − p

), (72)

and

var(t2) =1

N − pσ4

ϕ2

(1 +

2p

N − p

)2

. (73)

Similarly,

E[t1 ] = σ2ϕ1

(1 +

p

N − p

). (74)

However, in the presence of a causal relationship betweeny[n] and x[n], wH zp [n − 1] will be a better estimate of x[n]than hH xp [n − 1], therefore, wH zp [n − 1] can be written aswH zp [n − 1] = hH xp [n − 1] + x[n], such that, x[n] is orthog-onal to hH xp [n − 1]. Consequently,

ϕ1 [n] = ϕ2 [n] + x[n], (75)

with E[x[n]ϕ∗2 [n]] = 0, and

σ2ϕ1

= σ2ϕ2

+ σ2x , (76)

where σ2x � var(x2).

Now, since x[n] = wH zp [n] − hH xp [n], it can be shown thatE[x[n]] = 0, and,

E[|x[n]|2 ] = wH Rzz,pw + hH Rxx,ph − 2�{wH Rzxh}= rH

zz,pR−1zz ,przz ,p + rH

xx,pR−1xx,prxx,p

− 2�{rHzz,pR

−1zz ,pRzx,pR−1

xx,prxx,p}. (77)

Therefore, x[n] = x[n] + x[n] − hH xp [n − 1] and

E[t1 ] =(σ2

ϕ2+ σ2

x

)(1 +

p

N − p

)

= E[t2 ] + σ2x2

(1 +

p

N − p

)− σ2

ϕ2

N − p. (78)

Similarly,

E[t21 ] = E2 [t1 ] +1

N − p(σ2

ϕ2+ σ2

x)2(

1 +p

N − p

)2

, (79)

and,

var(t1) =1

N − p

(σ4

ϕ2+ σ4

x + 2σ2ϕ2

σ2x

)

×(

1 +p2

(N − p)2 +2p

N − p

)

= var(t2) +1

N − p

(σ4

x + 2σ2ϕ2

σ2x

− σ4ϕ2

(3p2

(N − p)2 +2p

N − p

)). (80)

Consequently, it can be shown that

cov(t1 , t2) =σ4

ϕ2

N − p. (81)

The GCI can therefore be approximated as the ratio of twocorrelated Gaussian r.v.s t1 and t2 whose statistics are derivedabove. In the next subsection, we use the above statistics tocalculate the probabilities of detection and false alarm.

C. Performance of GCI

Now, we declare a causal relationship to exist between thetwo signals if the ratio TE = t2

t1is below a threshold λ. Since

t1 is the sum of non-negative terms, it is almost surely positive,and therefore

Pr{TE < λ} = Pr{λt1 − t2 > 0}. (82)

Since t1 and t2 are approximated as correlated Gaussian r.v.s,therefore, from (78), TD � λt1 − t2 can also be approximatedas a Gaussian r.v. with mean and variance

E[TD ] = (λ − 1)E[t2 ] + λσ2x

(1 +

p

N − p

)− λσ2

ϕ2

p

N − p,

(83)

and

var(TD ) = λ2 var(t1) + var(t2) − 2λ cov(t1 , t2), (84)

respectively.Now, under the null hypothesis, H0 , σ2

ϕ1= σ2

ϕ2and σ2

x|H0=

0. Hence, under H0 , the above simplify to

μTD ,0 � E[TD |H0 ]

= (λ − 1)σ2ϕ2

(1 +

p

N − p

)− σ2

ϕ2

(p

N − p

), (85)

Page 8: Statistical Tests for Detecting Granger Causality · reversed Granger causality is the most noise resilient. The problem of sub-sampling in Granger causality detection hasbeenstudiedintheliterature[27]–[34].In[27],[28],theau-thors

5810 IEEE TRANSACTIONS ON SIGNAL PROCESSING, VOL. 66, NO. 22, NOVEMBER 15, 2018

σ2TD ,0 � var(TD |H0)

= (λ2 + 1)σ4

ϕ2

N − p

(1 +

2p

N − p

)2

− λ2 σ4ϕ2

N − p

(3p2

(N − p)2 +2p

N − p

)− 2λ

σ4ϕ2

N − p.

(86)

Similarly, under the alternate hypothesis,H1 , σ2ϕ1

> σ2ϕ2

, andtherefore,

μTD ,1 � E[TD |H1 ] = λσ2x

(1 +

p

N − p

)

+ (λ − 1)σ2ϕ2

(1 +

p

N − p

)− σ2

ϕ2

p

N − p, (87)

σ2TD ,1 � var(TD |H1) = (λ2 + 1)

σ4ϕ2

N − p

(1 +

2p

N − p

)2

+ λ2 σ4x

N − p

(1 +

p

N − p

)2

+ 2λ2 σ2ϕ2

σ2x2

N − p

(1 +

p

N − p

)2

− λ2 σ4ϕ2

N − p

(3p2

(N − p)2 +2p

N − p

)− 2λ

σ4ϕ2

N − p.

(88)

Based on these, the probability of false alarm can be calcu-lated as

PFA = Pr{TD > 0|H0} = Q

(μTD,0

σTD,0

). (89)

Similarly, the probability of correct detection of a causal rela-tionship, PD , can then be calculated as

PD = Pr{TD > 0|H1} = Q

(μTD,1

σTD,1

). (90)

Using (89), a detector providing a given false alarm rate of P ∗FA

can be designed by selecting a threshold λ that satisfies

[σ2

ϕ2(λ − 1)

(1 +

p

N − p

)− λσ2

ϕ2

p

N − p

]2

= Q−1(P ∗F A )

[(λ2 + 1)

σ4ϕ2

N − p

(1 +

p

N − p

)2

− λ2 σ4ϕ2

N − p

(3p2

(N − p)2 +2p

N − p

)− 2λ

σ4ϕ2

N − p

]. (91)

The above reduces to a quadratic equation in λ that can be solvedto obtain the detection threshold for the detector.

As an example, consider N samples of signals u[n] and v[n]corrupted by AWGN and observed as x[n] and y[n], respectively,such that v[n] = εv [n] and

u[n] =

{εu,0 [n] H0

aεv [n] + εu,1 [n] H1, (92)

with var(εv ) = σ2εv

, var(εu,1) = σ2εu

, and var(εu,0) = |a|2σ2εv

+σ2

εu. Then, K = 1, σ2

ϕ1is given by (58), and

σ2ϕ2

=

⎧⎨

σ2ϕ1

H0

σ2εu

+ σ2νx

+|a |2 σ 2

ε v σ 2ν y

σ 2ε u +σ 2

ν y= σ2

ϕ1− Δ H1

, (93)

with Δ = |a|2σ2εv

(1 − σ 2

ν y

σ 2ε v +σ 2

ν y

).

Therefore,

μTD ,0 = σ2ϕ1

(λ − 1 +

λ − 2N − 1

), (94)

and

σ2TD ,0 =

1N − 1

[(λ2 + 1)σ4

ϕ1

(1 +

2N − 1

)2

− λ2σ4ϕ1

(3

(N − 1)2 +2

N − 1

)− 2λ

σ2ϕ1

N − 1

].

(95)

Consequently, the probability of false alarm is given as (96) atthe bottom of this page.

Under the alternate hypothesis,

E[TD |H1 ] = (λ − 1)σ2ϕ1

(1 +

1N − 1

)

− Δ(

1 +1

N − 1

)− (σ2

ϕ1− Δ)

1N − 1

,

(97)

and

var(TD |H1) =1

N − 1

[(λ2 + 1)σ4

ϕ1

(1 +

1N − 1

)2

+ Δ2(1+

2N −1

)2

− 2σ2ϕ2

Δ(1 +

2N −1

)2

+ σ4ϕ1

(3

(N − 1)2 +2

N − 1

)]. (98)

The probability of detection can thus be expressed as (99) at thebottom of the next page.

We observe from (99) that the probability of detection de-pends on the ratio Δ

σ 2ϕ 1

, which can in turn be viewed as the

PFA = Q

⎜⎜⎝

√(N − 1)

(λ − 1 + λ−2

N −1

)

√(λ2 + 1)σ4

ϕ2

(1 + 2

N −1

)2 − λ2σ4ϕ2

(3

(N −1)2 + 2N −1

)− 2λ

σ 4ϕ 1

N −1

⎟⎟⎠ . (96)

Page 9: Statistical Tests for Detecting Granger Causality · reversed Granger causality is the most noise resilient. The problem of sub-sampling in Granger causality detection hasbeenstudiedintheliterature[27]–[34].In[27],[28],theau-thors

CHOPRA et al.: STATISTICAL TESTS FOR DETECTING GRANGER CAUSALITY 5811

relative difference between the variances of the prediction er-ror under the one variable and the two variable models. Conse-quently, the difference between the mean square prediction errorfor the single and two variable prediction models, i.e., t1 − t2can be used as a test statistic to detect the presence of a causalrelationship. We inspect the use of t1 − t2 as a test statistic inthe next section.

VI. GRANGER CAUSALITY DETECTION BASED ON THE

DIFFERENCE OF ERROR TERMS

Let TC � t1 − t2 denote the difference in the finite sampleestimates of the mean squared errors of the one variable andthe two variable prediction models. It can be argued that TC

is the difference of two correlated Chi-squared r.v.s., each with2(N − p) degrees of freedom. Therefore, the pdf of TC can bewritten as [37]

fTC(y) =

|y|N −p−1

(m − 1)![2var(t1)var(t2) − cov2(t1 , t2)γ](N −p)

× exp

(14α−y

)N −p−1∑

i=0

(N − p + i − 1)!(N − p − i − 1)!

(2

γ|y|)i

, y < 0;

fTC(y) =

|y|N −p−1

(m − 1)![2var(t1)var(t2) − cov2(t1 , t2)γ](N −p)

× exp

(−1

4α+y

)N −p−1∑

i=0

(N − p + i − 1)!(N − p − i − 1)!

(2

γ|y|)i

, y ≥ 0,

(100)

with

γ =[(var(t1) + var(t2))2 − 4cov2(t1 , t2)]

12

var(t1)var(t2) − cov2(t1 , t2)), (101)

and

α± = γ ± [(var(t1) − var(t2))]var(t1)var(t2) − cov2(t1 , t2))

. (102)

The p.d.f. of TC is in the form of a series, which makes itdifficult to obtain the probabilities of detection and false alarmin closed form. However, for large N , TC can be approximatedas a Gaussian r.v., such that,

E[TC |H0 ] = − σ2ϕ2

p

N − p, (103)

E[TC |H1 ] = σ2x

(1 +

p

N − p

)− σ2

ϕ2

p

N − p, (104)

and

var(TC |H0) =1

N − p

[2σ4

ϕ2

(1 +

2p

N − p

)2

− σ4ϕ2

(3p2

(N − p)2 +2p

N − p

)− 2σ4

ϕ2

].

(105)

var(TD |H1) =1

N − p

[2σ4

ϕ2

(1 +

2p

N − p

)2

+ σ4x

(1+

p

N − p

)2

+ 2σ2ϕ2

σ2x2

(1+

p

N − p

)2

− σ4ϕ2

(3p2

(N − p)2 +2p

N − p

)− 2σ4

ϕ2

].

(106)

Using the above statistics, it is now straightforward to designthe detector to achieve a given probability of false alarm, andobtain closed-form expressions for the probability of false alarmand detection, similar to (89) and (90), respectively. In particular,when N p, i.e., the number of samples is much larger than themodel order, it can be easily shown that the detection thresholdλ is well approximated by

λ =

√2Q−1(PF A )σ2

ϕ2√N − 1

, (107)

and the probability of detection as a function of the target falsealarm probability P ∗

F A can be approximated as

PD ≈ Q

⎝Q−1 (P ∗F A )

√2σ2

ϕ2−√

N − pσ2x√

σ4x + 2σ2

ϕ2σ2

x2

⎠ . (108)

Based on the above, the number of samples required to attain agiven pair of probabilities of detection and false alarm can beapproximated as

N ≈ p +

(Q−1(PF A )

√2σ2

ϕ2

σ2x

+ Q−1(PD )

1 + 2σ2

ϕ2

σ2x

)2

.

(109)Revisiting the example considered in the previous sections,

we observe that, σ2ϕ1

= σ2x , σ2

ϕ2= σ2

x − Δ, using which, it iseasy to obtain the expressions for the probabilities of false alarmand detection. We omit writing out the expressions for the sakeof brevity.

We can also use the mean squared value of the difference oferror terms as a test statistic to detect the presence of a causalrelationship between x[n] and y[n]. Letting x[n] � x[n] − x[n],

PD = Q

⎜⎜⎝

√N − 1

((λ − 1)σ2

ϕ1

(1 + 1

N −1

)− Δ(1 + 1

N −1

)− (σ2ϕ1

− Δ) 1N −1

)√

(λ2 + 1)σ4ϕ1

(1 + 1

N −1

)2 + Δ2(1 + 2

N −1

)2 − 2σ2ϕ2

Δ(1 + 2

N −1

)2 + σ4ϕ1

(3

(N −1)2 + 2N −1

)

⎟⎟⎠ . (99)

Page 10: Statistical Tests for Detecting Granger Causality · reversed Granger causality is the most noise resilient. The problem of sub-sampling in Granger causality detection hasbeenstudiedintheliterature[27]–[34].In[27],[28],theau-thors

5812 IEEE TRANSACTIONS ON SIGNAL PROCESSING, VOL. 66, NO. 22, NOVEMBER 15, 2018

Fig. 1. Value of TE at different SNRs, for a = 0.5, with known second-orderstatistics.

and

TA � 1N − p

N∑

n=p+1

|x[n]|2 , (110)

it can be shown that

E[TA |H0 ] = σ2ϕ1

(3p

N − p

), (111)

and

E[TA |H1 ] = rHzz,pR

−1zz ,przz ,p + rH

xx,pR−1xx,prxx,p

− 2�{rH R−1zz ,pRzxR−1

xx,prxx,p} +2σ2

ϕ2p

N − p+

σ2ϕ1

p

N − p. (112)

However, the higher order statistics and the exact distributionof TA cannot be found in closed form. Therefore, we evaluateits performance via simulations. To the best of our knowledge,the test statistics TC and TA have not been used in the literatureto detect Granger causality.

VII. SIMULATION RESULTS

In this section, we use Monte Carlo simulations to cor-roborate our analytical expressions and illustrate the relativeperformance of different test statistics for detecting Grangercausality. We also illustrate the causality detection performanceof TC using real-world data with known ground truth, availablein [38]

For the simulation-based study, we consider a system modelsimilar to the example discussed in the previous sections, fordifferent values of the variance of the additive noise and vary-ing number of samples. The probabilities of detection and falsealarm are obtained by averaging over 10,000 independent real-izations of the signals of interest.

A. Effect of Additive Noise

In Fig. 1, we plot the mean value of the test statistic TG forthe example considered in Section IV, with b = 0.5. Here we

Fig. 2. Probabilities of (a) detection and (b) false alarm as a function of thedetection threshold λ for the test statistic TE .

inspect the effects of noise on the GCI. We define γ1 � σ 2u

σ 2ν 1

,

and γ2 � σ 2u

σ 2ν 2

as the observation SNRs for x[n] and y[n]. We

see that the mean value of TG converges to a value close to zeroas the noise variance increases. The reduction in the SNR willmake it harder to detect the presence of a causal relationshipbetween the signals of interest.

B. Finite Sample Effects

We now consider the effects of using a finite number of sam-ples on the GCI, as discussed in Section V. In Figs. 2(a) and 2(b),we compare the simulated detection and false alarm probabil-ity obtained by using the test statistic TE as a function of thedetection threshold λ and for different number of samples, witha = 0.25 and γ1 = γ2 = 0 dB. The simulated performance ofTE is found to be in close agreement with the values computedusing our analytical expressions in (96) and (99). It is also ob-served that the behavior of both the probability of detection andfalse alarm as a function of the detection threshold becomessharper with an increase in the number of samples, and asymp-totically approximates a step function. The transition point of thestep function occurs at the mean of the test statistic under the re-spective hypotheses, which corresponds to the 50% probability

Page 11: Statistical Tests for Detecting Granger Causality · reversed Granger causality is the most noise resilient. The problem of sub-sampling in Granger causality detection hasbeenstudiedintheliterature[27]–[34].In[27],[28],theau-thors

CHOPRA et al.: STATISTICAL TESTS FOR DETECTING GRANGER CAUSALITY 5813

Fig. 3. ROC for detection based on TE , at an SNR of 0 dB, for differentnumber of samples.

of detection/ false alarm point on Figs. 2(a) and 2(b). Therefore,the threshold λ = 1 gives a zero error rate when infinite samplesof the two signals of interest are available. However, in case of afinite number of samples, λ has to be chosen carefully to achievereliable performance.

In Fig. 3, we compare the theoretical and simulated receiveroperating characteristics (ROCs) for the example considered inSection V for a = 0.25 with different values of N , at an SNRof 0 dB. It is observed that for a small number of samples,the ROC is close to the PF A = PD line, whereas, for a largenumber of samples, the probability of detection becomes almostindependent of PF A . This is in line with the behavior predictedin (96) and (99).

Having established that the use of finite samples indeed causesa deterioration in the performance of the GCI, we now discussthe performance of the other two test statistics, viz. TC andTA , as discussed in Section VI. In Fig. 4(a), we compare thetheoretical and simulated ROCs for the example considered inSection VI, with a = 0.25, for different number of samples, andat an SNR of 0 dB. We observe that the ROC for TC shows abehavior similar to that for TE with an increase in the numberof samples. Therefore, an appropriate choice of the detectionthreshold λ becomes important for TC as well.

In practice, one is typically provided a given number of sam-ples of a pair of time-series data to perform detection. In thiscase, there is a trade-off between the window size N and thequality of the ROC obtained. Using a larger N improves thereliability of detection, but the resulting ROC has more “jumps”because it is computed by averaging over fewer instantiations.We illustrate this in Fig. 4(b), where we plot the ROC for thesetup of Fig. 4(a) with the total number of available samplesfixed at 10,000. We see that the ROC with N = 100 (averagedover 100 instantiations) matches well with the correspondingcurve in Fig. 4(a), where 10,000 instantiations were used. Forhigher N , the performance improves, but the quality of theROC deteriorates, because the probability of detection is av-eraged over a correspondingly fewer number of instantiations.

Fig. 4. ROC for detection based on TC , at an SNR of 0 dB, for differentnumber of samples (N ) (a) averaged over 10,000 realizations in all cases,(b) obtained for a fixed data record length (10,000) divided into windows oflength N .

Thus, given a data record of a given size, a good choice for Nis the largest one such that the ROC is averaged over about 100instantiations.

In Fig. 5, we repeat the experiment considered in Fig. 4(a)by fixing the number of samples of x[n] and y[n] at 100 andvarying the SNRs. Again, the theoretical and simulated plots arein close agreement, with the concavity of the ROC increasingwith the SNR.

In Figs. 6 and 7, we evaluate the performance of TA as a teststatistic for detection of a causal relation between x[n] and y[n]for a setup similar to the example considered in Section V, witha = 0.25. Since theoretical expressions for the behavior of TA

are not derived, these are not considered here. In Fig. 6, we plotthe ROC of TA for different SNRs for 100 samples of each ofx[n] and y[n] being used for detection. The performance of TA

under different SNRs is seen to mimic that of TC and TE , withthe concavity of the ROC increasing with the SNR. In Fig. 7,we plot the probabilities of detection for different numbers ofsamples as a function of the signal SNRs. In this case, we use theNeyman-Pearson criterion to fix the probability of false alarm at10%. As expected, an increase in the number of samples leads to

Page 12: Statistical Tests for Detecting Granger Causality · reversed Granger causality is the most noise resilient. The problem of sub-sampling in Granger causality detection hasbeenstudiedintheliterature[27]–[34].In[27],[28],theau-thors

5814 IEEE TRANSACTIONS ON SIGNAL PROCESSING, VOL. 66, NO. 22, NOVEMBER 15, 2018

Fig. 5. ROC for detection based on TC , with N = 200 samples, for variousvalues of the SNR.

Fig. 6. ROC for detection based on TA , with N = 100 samples, for variousvalues of the SNR.

Fig. 7. Probability of detection using TA at a fixed false alarm rate of 10% asa function of the SNR.

improved detection performances for the same SNR. Therefore,TA can also be used to detect a causal relationship between apair of signals in the finite sample regime.

In Fig. 8, we compare the simulated performance of TE ,TC and TA for the example considered previously, with b =0.25, at an SNR of 0 dB and for 100 samples. We observethat TA , despite being theoretically intractable, offers the best

Fig. 8. ROC of test statistics TE , TC and TA for N = 100 samples at anSNR of 0 dB.

Fig. 9. Performance of the test statistic TC on the real-world datasets 67–69in [38] for different numbers of samples.

performance among the three test statistics considered in thispaper, followed by TC and TE respectively. It is interesting tonote that if the exact knowledge about the second-order statisticsis available, then the performance of the three test statisticsshould be identical.

C. Detection Performance of TC With Real World Data

In Figs. 9 and 10, we plot the performance obtained fromthe test statistic TC on a real-world dataset with known groundtruth, available in [38]. Fig. 9 corresponds to data pairs 65–67 ofthis database. These data pairs consist of the returns from relatedstocks over a period of five years. Similarly, Fig. 10 correspondsto data pair 69 of the database, containing temperature measure-ments from inside and outside a room taken every five minutes.These four data sets were selected out of the 108 available datasets because:

1) They represent time series data, in accordance with oursystem model.

2) The number of samples in these time series were suffi-ciently large to estimate of the probabilities of detectionand false alarm with good accuracy.

Page 13: Statistical Tests for Detecting Granger Causality · reversed Granger causality is the most noise resilient. The problem of sub-sampling in Granger causality detection hasbeenstudiedintheliterature[27]–[34].In[27],[28],theau-thors

CHOPRA et al.: STATISTICAL TESTS FOR DETECTING GRANGER CAUSALITY 5815

Fig. 10. Performance of the test statistic TC on the real-world dataset 69in [38] for different numbers of samples.

Fig. 11. Performance of TC for the dataset considered in Fig. 10 for N = 200with different amounts of added noise.

For generating the experimental results, we assumed that thedata was uniformly sampled at Nyquist rate. However, neitherthe model parameters of the data nor the additive noise varianceare known. Hence, an order 1-VAR generative model is as-sumed, and each data set is preprocessed to make it zero mean.The detection threshold for a given false alarm rate (PF A ) iscalculated assuming the two sequences to be independent andwhite, and using the sample variance of the caused sequence asσ2

ϕ2in (107).

The data pairs are divided into windows of length N , andthe test statistic TC is calculated and compared to the thresholdin (107). The results of these comparisons are averaged over allthe sample windows to obtain the probability of detection for agiven probability of false alarm. The jitters in the plots are dueto the averaging over the relatively small number of samplesthat were available. The performance of data set 69 is poorerthan data sets 65–67, for the same value of N , indicating eithera weaker causal relationship, or larger amount of additive noise,or both. Nonetheless, the ROC follows a pattern similar to thatpredicted by (108).

In Fig. 11, we plot the performance of TC for real-worlddata used in Fig. 10 with different amounts of added noise fora window length of N = 200. Noisy samples are obtained byadding real valued white Gaussian noise to the caused sequence,suitably scaled to achieve a given SNR. These samples are usedto calculate the test statistic TC , which is then compared againstthe detection threshold in (107), with σ2

ϕ1as the sample vari-

ance of the corrupted caused signal. It is observed that for a fixednumber of samples per window, the addition of noise worsensthe detection performance of TC , and the variation in the detec-tion performance approximately matches with the Q-functionof the square root of the SNR, as expected.

VIII. CONCLUSIONS

In this work, we derived the effects of different data acquisi-tion impairments on the detection of a causal relation betweentwo signals. We showed that these effects, viz. down-sampling,additive noise, and finite sample effects, could lead to a signifi-cant deterioration in the performance of the GCI, both individ-ually as well as cumulatively. We derived the probabilities ofdetection and false alarm for the GCI under the impairments.Following this, we proposed two alternative test statistics, basedon the difference of mean squared error terms, and the meansquared value of the difference of error terms. Via extensivesimulations, we showed that the derived results corroborate thesimulated as well as real-world data, with the test statistic basedon the mean squared value of the difference of error terms out-performing the other two statistics. A more precise analysis ofthis test statistic is an interesting direction for future research.

ACKNOWLEDGMENT

The authors are grateful to the anonymous reviewers for theirfeedback, which significantly improved the overall presentation,and in particular the simulation results in the paper.

REFERENCES

[1] I. Kontoyiannis and M. Skoularidou, “Estimating the directed informa-tion and testing for causality,” IEEE Trans. Inf. Theory, vol. 62, no. 11,pp. 6053–6067, Nov. 2016.

[2] C. W. J. Granger, “Investigating causal relations by econometric modelsand cross-spectral methods,” Econometrica, vol. 37, no. 3, pp. 424–438,1969.

[3] D. A. Pierce, “R2 measures for time series,” J. Amer. Statist. Assoc.,vol. 74, no. 368, pp. 901–910, 1979.

[4] R. Goebel, A. Roebroeck, D.-S. Kim, and E. Formisano, “Investigatingdirected cortical interactions in time-resolved fMRI data using vectorautoregressive modeling and Granger causality mapping,” Magn. Reson.Imag., vol. 21, no. 10, pp. 1251261, 2003.

[5] W. Hesse, E. Mller, M. Arnold, and B. Schack, “The use of time-variantEEG Granger causality for inspecting directed interdependencies of neuralassemblies,” J. Neurosci. Methods, vol. 124, no. 1, pp. 27–44, 2003.

[6] Y. Chen, G. Rangarajan, J. Feng, and M. Ding, “Analyzing multiple non-linear time series with extended Granger causality,” Phys. Lett. A, vol. 324,no. 1, pp. 26–35, 2004.

[7] M. Dhamala, G. Rangarajan, and M. Ding, “Analyzing information flowin brain networks with nonparametric Granger causality,” NeuroImage,vol. 41, no. 2, pp. 354–362, 2008.

[8] S. Hu and H. Liang, “Causality analysis of neural connectivity: New tooland limitations of spectral Granger causality,” Neurocomputing, vol. 76,no. 1, pp. 44–47, 2012.

Page 14: Statistical Tests for Detecting Granger Causality · reversed Granger causality is the most noise resilient. The problem of sub-sampling in Granger causality detection hasbeenstudiedintheliterature[27]–[34].In[27],[28],theau-thors

5816 IEEE TRANSACTIONS ON SIGNAL PROCESSING, VOL. 66, NO. 22, NOVEMBER 15, 2018

[9] D. Sampath et al., “A study on fear memory retrieval and REM sleepin maternal separation and isolation stressed rats,” Behav. Brain Res.,vol. 273, pp. 144–154, 2014.

[10] M. Hu and H. Liang, “A copula approach to assessing Granger causality,”NeuroImage, vol. 100, pp. 125–134, 2014.

[11] A. K. Seth, A. B. Barrett, and L. Barnett, “Granger causality analysis inneuroscience and neuroimaging,” J. Neurosci., vol. 35, no. 8, pp. 3293–3297, 2015.

[12] R. Ganapathy, G. Rangarajan, and A. K. Sood, “Granger causality andcross recurrence plots in rheochaos,” Phys. Rev. E, vol. 75, Jan. 2007, Art.no. 016211.

[13] S. Stramaglia, J. M. Cortes, and D. Marinazzo, “Synergy and redundancyin the Granger causal analysis of dynamical networks,” New J. Phys.,vol. 16, no. 10, 2014, Art. no. 105003.

[14] E. Kodra, S. Chatterjee, and A. R. Ganguly, “Exploring Granger causal-ity between global average observed time series of carbon dioxide andtemperature,” Theor. Appl. Climatol., vol. 104, no. 3, pp. 325–335, 2011.

[15] H. White and D. Pettenuzzo, “Granger causality, exogeneity, cointegration,and economic policy analysis,” J. Econometrics, vol. 178, Part 2, pp. 316–330, 2014.

[16] H. Nalatore, M. Ding, and G. Rangarajan, “Mitigating the effects of mea-surement noise on Granger causality,” Phys. Rev. E, vol. 75, Mar. 2007,Art. no. 031123.

[17] H. Nalatore, S. N, and G. Rangarajan, “Effect of measurement noise onGranger causality,” Phys. Rev. E, vol. 90, Dec. 2014, Art. no. 062127.

[18] I. Winkler, D. Panknin, D. Bartz, K. R. Mller, and S. Haufe, “Validity oftime reversal for testing Granger causality,” IEEE Trans. Signal Process.,vol. 64, no. 11, pp. 2746–2760, Jun. 2016.

[19] T. Schreiber, “Measuring information transfer,” Phys. Rev. Lett., vol. 85,pp. 461–464, Jul. 2000.

[20] L. A. Baccala and K. Sameshima, “Partial directed coherence: A newconcept in neural structure determination,” Biol. Cybern., vol. 84, no. 6,pp. 463–474, 2001.

[21] E. Siggiridou and D. Kugiumtzis, “Granger causality in multivariate timeseries using a time-ordered restricted vector autoregressive model,” IEEETrans. Signal Process., vol. 64, no. 7, pp. 1759–1773, Apr. 2016.

[22] P. Newbold and N. Davies, “Error mis-specification and spurious regres-sions,” Int. Econ. Rev., vol. 19, no. 2, pp. 513–519, 1978.

[23] G. Nolte et al., “Robustly estimating the flow direction of information incomplex physical systems,” Phys. Rev. Lett., vol. 100, Jun. 2008, Art no.234101.

[24] S. Haufe, V. V. Nikulin, and G. Nolte, Alleviating the Influenceof Weak Data Asymmetries on Granger-Causal Analyses. Berlin,Heidelberg: Springer, 2012, pp. 25–33.

[25] S. Haufe, V. V. Nikulin, K.-R. Mller, and G. Nolte, “A critical assessmentof connectivity measures for EEG data: A simulation study,” NeuroImage,vol. 64, pp. 120–133, 2013.

[26] M. Vinck et al., “How to detect the Granger-causal flow direction in thepresence of additive noise?,” NeuroImage, vol. 108, pp. 301–318, 2015.

[27] A. K. Seth, P. Chorley, and L. C. Barnett, “Granger causality analysisof fMRI BOLD signals is invariant to hemodynamic convolution but notdownsampling,” NeuroImage, vol. 65, no. Supplement C, pp. 540–555,2013.

[28] L. Barnett and A. K. Seth, “Detectability of Granger causality for subsam-pled continuous-time neurophysiological processes,” J. Neurosci. Meth-ods, vol. 275, no. Supplement C, pp. 93–121, 2017.

[29] M. Gong, K. Zhang, B. Schoelkopf, D. Tao, and P. Geiger, “Discoveringtemporal causal relations from subsampled data,” in Proc. 32nd Int. Conf.Mach. Learn., vol. 37, Jul. 2015, 1898–1906.

[30] M. Gong, K. Zhang, B. Scholkopf, C. Glymour, and D. Tao, “Causaldiscovery from temporally aggregated time series,” in Proc. Conf. Uncer-tainty Artif. Intell., Aug. 2017, pp. 1–10.

[31] A. Tank, E. B. Fox, and A. Shojaie, “Identifiability of non-Gaussian struc-tural VAR models for subsampled and mixed frequency time series,” inProc. SIGKDD Workshop Causal Discovery, 2016, pp. 1–7.

[32] A. Hyttinen, S. Plis, M. Jrvisalo, F. Eberhardt, and D. Danks, “A constraintoptimization approach to causal discovery from subsampled time seriesdata,” Int. J. Approx. Reasoning, vol. 90, pp. 208–225, 2017.

[33] S. Plis, D. Danks, C. Freeman, and V. Calhoun, “Rate-agnostic (causal)structure learning,” Adv. Neural Inf. Process. Syst., vol. 28, pp. 3303–3311,2015.

[34] X. Wen, G. Rangarajan, and M. Ding, “Is Granger causality a viabletechnique for analyzing fMRI data?,” PLOS ONE, vol. 8, pp. 1–11, 2013.

[35] S. Haykin, Adaptive Filter Theory. London, U.K.: Pearson Education,2002.

[36] R. Tandra and A. Sahai, “SNR walls for signal detection,” IEEE J. Sel.Topics Signal Process., vol. 2, no. 1, pp. 4–17, Feb. 2008.

[37] M. K. Simon, Probability Distributions Involving Gaussian Random Vari-ables. New York, NY, USA: Springer, 2002.

[38] J. M. Mooij, J. Peters, D. Janzing, J. Zscheischler, and B. Scholkopf,“Distinguishing cause from effect using observational data: Methods andbenchmarks,” J. Mach. Learn. Res., vol. 17, no. 32, pp. 1–102, 2016.

Ribhu Chopra (S’11–M’17) received the B.E. de-gree in electronics and communication engineeringfrom Panjab University, Chandigarh, India, in 2009,and the M.Tech. and Ph.D. degrees in electronicsand communication engineering from the Indian In-stitute of Technology Roorkee, Roorkee, India, in2011 and 2016, respectively. He worked as a ProjectAssociate with the Department of Electrical Com-munication Engineering, Indian Institute of Science,Bangalore, from August 2015 to May 2016. FromMay 2016 to March 2017, he worked as an Institute

Research Associate with the Department of Electrical Communication Engi-neering, Indian Institute of Science, Bangalore, India. In April 2017, he joinedthe Department of Electronics and Electrical Engineering, Indian Institute ofTechnology Guwahati, Assam, India. His research interests include statisticaland adaptive signal processing, massive MIMO communications, and cognitivecommunications.

Chandra Ramabhadra Murthy (S’03–M’06–SM’11) received the B.Tech. degree in electricalengineering from the Indian Institute of Technol-ogy Madras, Chennai, India, in 1998, the M.S. andPh.D. degrees in electrical and computer engineer-ing from Purdue University and the University ofCalifornia, San Diego, USA, in 2000 and 2006, re-spectively. From 2000 to 2002, he worked as anEngineer for Qualcomm, Inc., where he worked onWCDMA baseband transceiver design and 802.11bbaseband receivers. From August 2006 to August

2007, he worked as a Staff Engineer at Beceem Communications, Inc., onadvanced receiver architectures for the 802.16e Mobile WiMAX standard. InSeptember 2007, he joined the Department of Electrical Communication En-gineering, Indian Institute of Science, Bangalore, India, where he is currentlyworking as a Professor.

His research interests include the areas of energy harvesting communica-tions, multiuser MIMO systems, and sparse signal recovery techniques appliedto wireless communications. His paper won the best paper award in the Com-munications Track at NCC 2014 and a paper co-authored with his student wonthe student best paper award at the IEEE ICASSP 2018. He has 50+ journalpapers and 80+ conference papers to his credit. He was an Associate Edi-tor for the IEEE SIGNAL PROCESSING LETTERS during 2012–2016. He is anelected member of the IEEE SPCOM Technical Committee for the years 2014–2016, and has been re-elected for the 2017–2019 term. He is the Past Chairof the IEEE Signal Processing Society, Bangalore Chapter. He is currentlyserving as an Associate Editor for the IEEE TRANSACTIONS ON SIGNAL PRO-CESSING, the Sadhana Journal, and an Editor for the IEEE TRANSACTIONS ON

COMMUNICATIONS.

Govindan Ranagarajan received the IntegratedM.Sc. degree from Birla Institute of Technology andScience, Pilani, India, and the Ph.D. degree from theUniversity of Maryland, College Park, MD, USA. Hethen was with the Lawrence Berkeley Lab, Univer-sity of California, Berkeley, before returning to Indiain 1992. He is currently a Professor of mathematicswith the Indian Institute of Science, Bangalore. He isalso the Chairman of the Division of InterdisciplinaryResearch.

His research interests include nonlinear dynamicsand chaos, time series analysis, and brain–machine interface. He is a J. C. BoseNational Fellow. He is also a Fellow of the Indian Academy of Sciences and theNational Academy of Sciences, India. He was knighted by the Government ofFrance: Knight (Chevalier) of the Order of Palms. He was also a Homi BhabhaFellow.