15
1 Statistical Tests for Detecting Granger Causality Ribhu Chopra, Chandra R. Murthy, and Govindan Rangarajan Abstract—Detection of a causal relationship between two or more sets of data is an important problem across various scientific disciplines. The Granger causality index and its derivatives are important metrics developed and used for this purpose. However, the test statistics based on these metrics ignore the effect of practical measurement impairments such as subsampling, additive 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 performance of the conventional and proposed causality detectors on parameters such as the additive noise variance and the strength of the causal relationship. I. I NTRODUCTION A. Motivation Confirming the presence or absence of a causative relation- ship between different observed phenomena is an important problem across various disciplines of physical, biological, and social sciences. It has been argued that if the phenomena 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 Gaussian distributed time series, the directed mutual infor- mation reduces to the logarithm of ratio of the prediction error variances of the time series of interest with and without including the purported causing time series into the prediction model [2]. This quantity, known as the Granger causality index (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 R. Chopra is with the Indian Institute of Technology Guwahati, Assam, India. (email: [email protected]). C. R. Murthy and G. Rangarajan are with the Indian Institute of Science, Bangalore, India. (emails: [email protected], [email protected]). The work of C. R. Murthy was supported in part by a research grant from the Aerospace Network Research Consortium and by the Young Faculty Research Fellowship from the Ministry of Electronics and Information Tech- nology, Govt. of India. The work of G. Rangarajan was supported by research grants from JC Bose National Fellowship, DST Centre for Mathematical Biology Phase II and UGC Centre for Advanced Study. 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 statistics 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 paper, 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 detection [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 additive 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 impairments 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 relationship between x[n] and y[n] can be inferred by comparing the mean squared prediction error for x[n] with and without including y[n] in the prediction model [18]. The prediction model considering only the past samples of x[n] is known as the restrictive model (R-model), whereas the one including the effects of y[n] is known as the unrestricted model (U- model). In addition to the GCI, several metrics for quantifying Granger causality have been studied [19], [20]. The system model with two time series [2] has been also generalized to study causal relationships between more than two time series [21]. However, all these studies assume the samples are noiseless, and ignore the finite sample effects on the second- order statistics being estimated.

Statistical Tests for Detecting Granger Causalitycmurthy/lib/exe/fetch.php?media=... · 2018. 9. 15. · The Granger causality index and its derivatives are important metrics developed

  • Upload
    others

  • View
    5

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Statistical Tests for Detecting Granger Causalitycmurthy/lib/exe/fetch.php?media=... · 2018. 9. 15. · The Granger causality index and its derivatives are important metrics developed

1

Statistical Tests for Detecting Granger CausalityRibhu Chopra, Chandra R. Murthy, and Govindan Rangarajan

Abstract—Detection of a causal relationship between two ormore sets of data is an important problem across various scientificdisciplines. The Granger causality index and its derivativesare important metrics developed and used for this purpose.However, the test statistics based on these metrics ignore the effectof practical measurement impairments such as subsampling,additive noise, and finite sample effects. In this paper, wemodel the problem of detecting a causal relationship betweentwo time series as a binary hypothesis test with the null andalternate hypotheses corresponding to the absence and presenceof a causal relationship, respectively. We derive the distributionof the test statistic under the two hypotheses and show thatmeasurement impairments can lead to suppression of a causalrelationship between the signals, as well as false detection of acausal relationship, where there is none. We also use the derivedresults to propose two alternative test statistics for causalitydetection. These detectors are analytically tractable, which allowsus to design the detection threshold and determine the numberof samples required to achieve a given missed detection andfalse 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 detectionperformance of the conventional and proposed causality detectorson parameters such as the additive noise variance and thestrength of the causal relationship.

I. I NTRODUCTION

A. Motivation

Confirming the presence or absence of a causative relation-ship between different observed phenomena is an importantproblem across various disciplines of physical, biological, andsocial sciences. It has been argued that if the phenomenaof interest can be represented as a time series, then thecorresponding relationship between them can be quantifiedby directed mutual information [1]. It is also known that forGaussian distributed time series, the directed mutual infor-mation reduces to the logarithm of ratio of the predictionerror variances of the time series of interest with and withoutincluding the purported causing time series into the predictionmodel [2]. This quantity, known as theGranger causalityindex (GCI) [2], [3], has been applied extensively for thedetection of causal relationships across numerous researchareas such as neuroscience [4]–[11], physics [12], [13], climatescience [14], econometrics [15], etc. Most of the presentstudies on Granger causality assume that the sampling process

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

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

The work of C. R. Murthy was supported in part by a research grantfrom the Aerospace Network Research Consortium and by the Young FacultyResearch Fellowship from the Ministry of Electronics and Information Tech-nology, Govt. of India. The work of G. Rangarajan was supported by researchgrants from JC Bose National Fellowship, DST Centre for MathematicalBiology Phase II and UGC Centre for Advanced Study.

is noiseless and the sampling frequency is above the Nyquistrate. As a consequence, the tests are not designed to achievea given false alarm/detection rate. It is also typically assumedthat the second-order statistics of the signals of interestareperfectly known. However, in practice, the causality detectorhas to estimate the second-order statistics from a finite numberof possibly undersampled and noisy measurements of thesignals of interest. In this paper, our main focus is on thequantification of the effects of these measurement inaccuracieson the performance of Granger causality detectors. This, inturn, allows us to design the detector to achieve a desiredfalse 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.In addition to this, additive noise can lead to suppression ofan existing causal relationship, as well as may cause a falsedetection [16], [17]. The effect of these inaccuracies is furthercompounded by the fact that the second-order statistics ofthe signal of interest, required to compute the GCI, are notknown, and have to be estimated using the acquired samples.Therefore, finite sample effects, in addition to undersamplingand additive noise, also contribute to errors in the detectionof a causal relationship between the signals of interest. Inthispaper, we model the problem of detection of Granger causalityas a binary hypothesis test, and evaluate the effects of themeasurement impairments listed above on the probabilitiesofdetection and false alarm.

B. Related Work

In the original formulation of Granger causality in [2], atime seriesy[n] is said toGranger causeanother time seriesx[n], when the past samples of they[n] can be used to improvethe estimate of thex[n] over ‘what can be predicted using allother information in the universe’, conventionally consideredto be the past samples ofx[n]. Therefore, a causal relationshipbetweenx[n] and y[n] can be inferred by comparing themean squared prediction error forx[n] with and withoutincluding y[n] in the prediction model [18]. The predictionmodel considering only the past samples ofx[n] is knownas the restrictive model (R-model), whereas the one includingthe effects ofy[n] is known as the unrestricted model (U-model). In addition to the GCI, several metrics for quantifyingGranger causality have been studied [19], [20]. The systemmodel with two time series [2] has been also generalizedto study causal relationships between more than two timeseries [21]. However, all these studies assume the samples arenoiseless, and ignore the finite sample effects on the second-order statistics being estimated.

Page 2: Statistical Tests for Detecting Granger Causalitycmurthy/lib/exe/fetch.php?media=... · 2018. 9. 15. · The Granger causality index and its derivatives are important metrics developed

2

The effect of noise on Granger causality was first discussedin [22], and was mathematically analyzed in [16] for abivariate VAR-1 process. These results were extended to ageneralized VAR(p) (pth order autoregressive process) in [17],where a Kalman filtering based technique to alleviate thedetrimental effect of noise was also developed. An alternativetest statistic for causality detection, known as the phase slopeindex, was introduced in [23] and was shown to be moreresilient to noise, as compared to the GCI. Another metricfor causality, known as the time reversed Granger causality,has been discussed in [18], [24], [25]. The authors in [26] usedsimulation techniques to compare the robustness of GCI, timereversed Granger causality, and phase slope index to correlatednoise, and found that time reversed Granger causality is themost noise resilient.

The problem of sub-sampling in Granger causality detectionhas been studied in the literature [27]–[34]. In [27], [28],the authors study the effects of the choice of the samplingfrequency on the detectability of a causal relationship betweenneurological signals. In [28], these derived results are usedto identify favorable sampling rates for detecting Grangercausality. In [29], two methods, based on the expectationmaximization (EM) algorithm and the variational inferenceframework, are proposed to detect causal relationships fromlow resolution data, and their applicability to simulated andpractical data is tested. The problem of causal inference froman under-sampled time series by using a general purposeBoolean constraint solver has been discussed in [32]. It isshown that the method considered in [32] is independentof the modeling parameters, and can be conveniently scaledto any number of variables. In [33], the authors proposethree sampling rate agnostic algorithms for the recovery ofacausal relationship between observed time series. The authorsin [34] incorporate the effect of both down-sampling andadditive 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 alsodonot analyze the effects of these physical impairments on theprobabilities of detection and false alarm.

The authors in [1] consider the related problem of estima-tion of directed mutual information between two time series,using a finite number of samples. Here, the test for a causalrelationship between two time series is modeled as a binaryhypothesis test, and it is shown that the probabilities of misseddetection and false alarm asymptotically converge to zero asthe number of samples is increased. In this paper, we proposeto extend the binary hypothesis testing framework for causalitydetection to GCI.

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.We model the time series as two independentpth orderautoregressive (AR(p)) processes in the absence of a causalrelation, and as a bivariate vector autoregressive (VAR(p))

process in the presence of a causal relationship. We model thetest for causality as a binary hypothesis testing problem withthe null hypothesis corresponding to the absence of a causalrelationship, and the alternate hypothesis to its presence. Thesystem model is described in detail in Section II. Followingthis, the main objective of this paper is to evaluate the detectionperformance of GCI for finite samples of the two time seriesof interest. Our contributions 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 twohypotheses for finite number of samples, and use theseto derive the probabilities of detection and false alarm.Our analysis accounts for the estimation of the second-order statistics of the two signals using a finite numberof samples when the generative model is unknown. (SeeSection V.)

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

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

The results derived in this paper can be used to characterizethebehavior of GCI and related detectors under different physicalconditions. 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 consideredin this work.

Notation: Boldface lowercase and uppercase letters repre-sent vectors and matrices, respectively. Thekth column ofAis denoted byak. (.)H represents the Hermitian operation ona vector or a matrix, and(.)† represents the Moore-Penrosepseudoinverse of a matrix.IK ,0K and OK represent theidentity matrix, the all-zero vector and the all-zero matrixof dimensionK, respectively. Theℓ2 norm of a vector andthe Frobenius norm of a matrix are denoted by‖·‖2 and‖·‖F , respectively.CN (µ, σ2) represents a circularly sym-metric complex Gaussian random variable with meanµ andvarianceσ2. E[·] and var(·) represent the mean and varianceof a random variable. In general,x and x denote the MMSEestimate and the corresponding estimation error of a randomvariablex. Table I introduces the different symbols used inthis paper.

II. SYSTEM MODEL

We consider two complex valued discrete time zero meanGaussian random processesc[n] andd[n] having variancesσ2

c

andσ2d, respectively, and expressed in the form of a bivariate

Page 3: Statistical Tests for Detecting Granger Causalitycmurthy/lib/exe/fetch.php?media=... · 2018. 9. 15. · The Granger causality index and its derivatives are important metrics developed

3

TABLE I: A list of notation followed in the paper.

Rxy,p[τ ] The p× p correlation matrix between tworandom vectorsx[n] andy[n− τ ] oflengthp each.

rxx,p[τ ] The p× 1 correlation vector between thescalarx[n] and itsp past values startingfrom x[n− τ ].

rxy[τ ] The correlation between the scalarsx[n] andy[n− τ ].

σ2x Variance of the WSS random processx[n].

d[n] A noiseless random process sampled at theNyquist rate.

c[n] A noiseless random process, potentiallycaused byd[n], sampled at the Nyquist rate.

q The model order for the VAR modelrelatingc[n] andd[n].

A The weight matrix relatingc[n] andd[n].u[n] A downsampled version ofc[n].v[n] A downsampled version ofd[n].p The model order for the VAR model

relatingu[n] andv[n].B The weight matrix relatingu[n] andv[n]x[n] A noisy observation ofu[n].y[n] A noisy observation ofv[n].zp[n] The concatenated observation vector of

length2p.w The weight vector relatingx[n] andzp[n].h The weight vector relatingx[n] andxp[n].

VAR(q) model as

c[n] =

q∑

k=1

a∗1,kc[n− k] +

q∑

k=1

a∗2,kd[n− k] + ηc[n],

d[n] =

q∑

k=1

a∗3,kc[n− k] +

q∑

k=1

a∗4,kd[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 inc[n] andd[n],respectively. We assume thatc[n] and d[n] are sampled atthe Nyquist rate, and the temporally white innovation processvectorη[n] = [ηc[n] , ηd[n]]

T is distributed as

η[n] ∼ CN(

0,

[

σ2ηc

00 σ2

ηd

])

.

For simplicity, we assume that only a unidirectional couplingfrom d[n] to c[n] can exist, i.e.,a3,k = 0 , for1 ≤ 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]

]

=

[

aH1 aH20Hq aH4

] [

cq[n− 1]dq[n− 1]

]

+

[

ηc[n]ηd[n]

]

, (2)

with 0q being theq dimensional all zero vector. Definingrcc[τ ] , E[c[n]c∗[n − τ ]], rcd[τ ] , E[c[n]d∗[n − τ ]], and

similarly rdc[τ ] andrdd[τ ], we can write

rcc[τ ] =

q∑

k=1

a∗1,krcc[τ−k]+

q∑

k=1

a∗2,krdc[τ−k]+σ2ηcδ[τ ]. (3)

Lettingrcd,q[τ ] = [rcd[τ ], . . . , rcd[τ−q+1]]T , and similarlyrcc,q[τ ], etc., the above becomes

rcc[τ ] = aH1 rcc,q[τ − 1] + aH2 rdc,q[τ − 1] + σ2ηcδ[τ ]. (4)

We can similarly write

rcd[τ ] = aH1 rcd,q[τ − 1] + aH2 rdd,q[τ − 1], (5)

andrdd[τ ] = aH4 rdd,q[τ − 1] + σ2

ηdδ[τ ]. (6)

The above equations, along with the information thatrcc[0] =σ2c , and rdd[0] = σ2

d can be used to computercc[τ ], rcd[τ ],andrdd[τ ] for different values ofτ .

Defining

Rcc,q[τ ] , E[

cq [n]cHq [n− τ ]

]

, (7)

Rcd,q[τ ] , E[

cq[n]dHq [n− τ ]

]

, (8)

and similarly definingRdc,p[τ ] andRdd,q[τ ], it can be shownusing the Weiner-Hopf equations [35] that

[

a1a2

]

=

[

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 fromthe correlation coefficients when the latter are not known. Theinterested reader is referred to [35, Chapter 2] for a detailedderivation of the Wiener-Hopf equations. We can then usethese to compute the 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 dependenceof c[n] on d[n] is determined by the coefficient vectora2, thatis, there exists a causal relationship only ifa2 6= 0.

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

c[n] = gHcq[n− 1] + ζ[n], (11)

with ζ[n] being the zero mean temporally white complexGaussian innovation component having a varianceσ2

ζ . Con-sequently,

rcc[τ ] = gHrcc,q[τ − 1] + σ2ζδ[τ ], (12)

g = R−1cc,q[0]rcc,q[1], (13)

andσ2ζ = σ2

c − rHcc,q[1]R−1cc,q[0]rcc,q[1]. (14)

It is to be noted that in the presence of a causal relationshipbe-tweenc[n] andd[n] the mean squared prediction error forc[n]using the bivariate model,σ2

ηc, will be smaller than the mean

Page 4: Statistical Tests for Detecting Granger Causalitycmurthy/lib/exe/fetch.php?media=... · 2018. 9. 15. · The Granger causality index and its derivatives are important metrics developed

4

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 relation-

ship, and is defined as [2]

TG , log

(

σ2ζ

σ2ηc

)

. (15)

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

However, bothc[n] and d[n] are assumed to be noiselessand sampled at the Nyquist rate. Moreover, it also assumesthat the exact second-order statistics for bothc[n] and d[n]are known. However, these assumptions do not hold in apractical data acquisition system. Therefore, in the subse-quent sections, we introduce these imperfections, viz. down-sampling, measurement noise, and finite sample effects, into aGCI based detector, and analyze their impact on the detectionperformance. In the next section, we discuss the effect ofdown-sampling on the GCI.

III. T HE EFFECT OFDOWN-SAMPLING ON GCI

Let u[n] andv[n] respectively be versions ofc[n] andd[n]down-sampled by a factorM , 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 setthe offset parameterl to zero without loss of generality. Wecan now expressu[n] andv[n] in the form of a bivariate linearregression process of orderp as[

u[n]v[n]

]

=

[

bH1 bH

2

0Hp bH

4

] [

up[n− 1]vp[n− 1]

]

+

[

ǫu[n]ǫv[n]

]

, (17)

with bi, i ∈ {1, 2, 4} representing new regression coeffi-cients, andE [u[n− k]ǫ∗u[n]] = 0, E [v[n− k]ǫ∗v[n]] = 0,E [u[n− k]ǫ∗v[n]] = 0, andE [v[n− k]ǫ∗u[n]] = 0 for k > 0.It is important to note that the model orderp in this case maybe different from the model orderq considered in the previoussection due to down-sampling.

Since the GCI is a function of the second-order statistics ofc[n] andd[n], we need to derive the second-order statistics ofu[n] and v[n] in terms ofc[n] and d[n] in order to quantifythe effect 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]vHp [n − τ ]], with the absence

of 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] + σ2ηcδ[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 vectorsb1, b2, andb4 can then be determinedvia the 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,

ǫu[n] = u[n]− bH1 up[n− 1]− bH

2 vp[n− 1], (24)

ǫv[n] = y[n]− bH4 vp[n− 1]. (25)

Consequently,

rǫuǫu [0] = σ2c

−[

ruu,p[1]ruv,p[1]

]H [Ruu,p Ruv,p

Rvu,p Rvv,p

]−1 [ruu,p[1]ruv,p[1]

]

, (26)

rǫvǫv [0] = σ2c − rHvv,p[1]R

−1vv,prvv,p[1]. (27)

rǫuǫv [0] = rcd[0]

−[

rHuu,p[1] rHuv,p[1]

]

[

Ruu,p Ruv,p

Rvu,p Rvv,p

]−1 [ruv,p[1]rvv,p[1]

]

− rHuv,p[1]R−1vv,prvv,p[1] +

[

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

]H

×[

Ruu,p Ruv,p

Rvu,p Rvv,p

]−1 [Ruv,p

Rvv,p

]

R−1vv,prvv,p[1]. (28)

Hence, the covariance matrix of the innovation process neednot be diagonal after down-sampling.

Similarly, the single variable linear regression model forthedown-sampled signalu[n] can be written as

u[n] = fHup[n− 1] + ξ[n], (29)

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

f = R−1uu,pruu,p[1], (30)

and rξξ[0] = σ2u − rHuu,p[1]R

−1uu,pruu,p[1]. (31)

A causal relationship betweenu[n] and v[n] can now be

determined by considering the ratioTG = log2

(

rξξ[0]rǫuǫu [0]

)

.

To illustrate the effect of down-sampling on the GCI,consider the simple special case:c[n] = ad[n − 1] + ηc[n],andd[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)

Page 5: Statistical Tests for Detecting Granger Causalitycmurthy/lib/exe/fetch.php?media=... · 2018. 9. 15. · The Granger causality index and its derivatives are important metrics developed

5

Consequently,ruu[1] = rcc[2] = 0, ruv[1] = rcd[2] = 0,ruv[0] = rcd[0] = 0, andrξξ[0] = rǫxǫx [0] = σ2

c . We see thatthere exists a causal relationship betweenc[n] and d[n] foranya 6= 0, but no causal relationship exists betweenu[n] andv[n]. Thus, down-sampling can suppress an existing causalrelationship between two signals. However, the degree ofsuppression of the causal relationship depends on the structureof the VAR model followed by the signals. We next analyzethe effect of additive noise on the performance of the GCI.

IV. GCI WITH DOWNSAMPLING AND ADDITIVE NOISE

Let us definex[n] and y[n] as the down-sampled signalsu[n] andv[n] corrupted by AWGN, such that,

x[n] = u[n] + νx[n],

y[n] = v[n] + νy[n], (34)

whereνx[n] ∼ CN (0, σ2νx) andνy[n] ∼ CN (0, σ2

νy).

Using (17), we can write,[

x[n]y[n]

]

=

[

bH1 bH

2

0Hp bH

4

] [

up[n− 1]vp[n− 1]

]

+

[

ǫu[n] + νx[n]ǫv[n] + νy[n]

]

.

(35)If a causal relationship exists betweenc[n] and d[n], and ispreserved inu[n] andv[n] after down-sampling, it should alsoexist betweenx[n] and y[n], and therefore, the past samplesof x[n] and y[n] can be used to predictx[n] better than itsprediction by using the past samples ofx[n] alone. Lettingz[n] = [x[n] y[n]]T and zp[n] = [xT

p [n] yTp [n]]

T , we canexpressx[n] as

x[n] = wHzp[n− 1] + ϕ2[n], (36)

with ϕ2[n] being the prediction error for the bivariate VARmodel, andw being the weight vector minimizing the meansquared value ofϕ2[n].

Defining σ2ϕ2

, E[|ϕ2[n]|2] = E[|x[n] −wHzp[n − 1]|2],it can be shown that [35],

σ2ϕ2

= σ2x −wHrzx,p[1]− rHzx,p[1]w+wHRzz,pw, (37)

whereσ2x = E[x[n]x∗[n]] = σ2

c + σ2νx, (38)

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

Hyx[τ ]]

H , (39)

andRzz,p = E

[

zp[n]zHp [n]

]

. (40)

Since the additive noise is independent of bothu[n] andv[n], it can be shown that

rxx[τ ] = ruu[τ ] + σ2νxδ[τ ], (41)

rxy[τ ] = ruv[τ ], (42)

ryy[τ ] = rvv[τ ] + σ2νyδ[τ ]. (43)

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

rxx[τ ]=bH1 ruu,p[τ − 1] + bH

2 rvu,p[τ − 1] + (σ2ǫu+σ2

νx)δ[τ ],

(44)rxy[τ ] = bH

1 ruv,p[τ − 1] + bH2 rvv,p[τ − 1], (45)

ryy[τ ] = bH4 rvv,p[τ − 1] + (σ2

ǫv+ 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

]

+

[

σ2νxIp 0p

0p σ2νyIp

]

. (47)

The optimal weight vector minimizingσ2ϕ2

can be obtainedvia the Wiener-Hopf equations as

w = R−1zz,przx,p[1]. (48)

The minimized value ofσ2ϕ2

is

σ2ϕ2

= σ2x − rHzx,p[1]R

−1zz,przx,p[1]. (49)

However, in case no causal relationship exists betweenc[n]andd[n], and consequently betweenx[n] and y[n], x[n] canbe expressed using the single variable AR model as

x[n] = hHxp[n− 1] + ϕ1[n], (50)

with ϕ1[n] being the single variable prediction error, andhbeing the 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 − rxx,p[1]

HR−1xx,prxx,p[1]. (52)

The GCI therefore becomes

TG = log2

(

σ2ϕ1

σ2ϕ2

)

= log2

(

σ2x − rHxx,p[1]R

−1xx,prxx,p[1]

σ2x − rHzx,p[1]R

−1zz,przx,p[1]

)

.

(53)

To demonstrate the effects of noise on the GCI, we againconsider our running example,d[n] = ǫv[n], c[n] = u[n] =ad[n− 1]+ bd[n− 2]+ ǫc[n]. Downsampling these by a factorM = 2, we obtain,v[n] = ǫv[n], u[n] = bv[n − 1] + ǫu[n],with, σ2

v = σ2ǫv

, andσ2u = |b|2σ2

ǫv+ σ2

ǫu, and consequently,

σ2x = |b|2σ2

ǫv+ σ2

ǫu+ σ2

νx, σ2

y = σ2ǫv

+ σ2νy, (54)

rxy[0] = 0, rxy[1] = bσ2ǫv. (55)

Therefore,

Rzz,1 =

[ |b|2σ2ǫv

+ σ2ǫu

+ σ2νx

00 σ2

ǫv+ σ2

νy

]

, (56)

σ2ϕ2

= σ2ǫu

+ σ2νx

+|b|2σ2

ǫvσ2νy

σ2ǫv

+ σ2νy

, (57)

andσ2ϕ1

= σ2x = |b|2σ2

ǫv+ σ2

ǫu+ σ2

νx. (58)

We can simplify the GCI as

TG = log2

|b|2σ2ǫv

+ σ2ǫu

+ σ2νx

|b|2σ2ǫv

(

σ2νy

σ2ǫv

+σ2νy

)

+ σ2ǫu

+ σ2νx

. (59)

In the absence of additive noise in bothx[n] and y[n],

Page 6: Statistical Tests for Detecting Granger Causalitycmurthy/lib/exe/fetch.php?media=... · 2018. 9. 15. · The Granger causality index and its derivatives are important metrics developed

6

TG reduces toTG = log2

(

|b|2σ2

ǫv+σ2

ǫu

σ2ǫu

)

, as in the previous

section. However, as eitherσ2νx

, or σ2νy

get large, the secondterm in the expression forTG converges to zero. Thus, additivenoise can also lead to the suppression of a causal relationshipbetween two signals.

The discussion till now assumed that the second-orderstatistics of both the signals of interest are available at thedetector. However, in practice, we have to estimate the second-order statistics using a finite number of samples ofx[n] andy[n]. The effects of the use of estimated second-order statisticson the detection of Granger causality in an under-samplednoisy environment are discussed in the next section.

V. FINITE SAMPLE EFFECTS ON THEGCI

In case onlyN samples of each ofx[n] and y[n] areobserved, and the generative model is unknown, we need toestimate the optimal predictors for the single and two variablecases, the corresponding prediction error variances, and theGCI or an equivalent test statistic using these samples. In thissection, we first discuss the properties of the optimal predictorweights for the one variable and two variable predictionmodels. Then, we use these weights to calculate the statisticsof the finite sample estimates of the corresponding meansquare prediction errors, and finally derive expressions for theperformance of GCI with finite 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 theN available samplesof x[n], y[1], . . . , y[N ] be theN available samples ofy[n],and x[n] = wHzp[n − 1] be the least squares (LS) estimateof x[n] generated using the two variable model, we can write

xN−p[N ] = xN−p[N ] + xN−p[N ], (60)

wherexN−p[N ] is the error term orthogonal to the measure-ment space. Also,w is the least squares estimate of the weightvector, 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. Sincethe innovation process and the measurement noise are whiteand zero mean, it can be shown thatw has the followingproperties [35]:

1) w is a an unbiased estimator ofw.

2) The covariance matrix ofw is cov(w) =σ2

ϕ2

N−pR−1

zz,p.

Therefore,w = w + w, with w ∼ CN(

0,σ2

ϕ2

N−pR−1

zz,p

)

.

Similarly, considering the single variable prediction model,we can writexN−p[N ] in terms of its projection,xN−p[N ] onthe data matrix, ,Xp[N ] = [xp[N−1],xp[N−2], . . . ,xp[p]]

H ,and the error componentxN−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,σ2ϕ1

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 expressx[n] as

x[n] = x[n]− x[n]

= x[n]−wHzK [n− 1]− wHzK [n− 1]. (63)

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

E[|x[n]|2] = σ2ϕ2

+ E[wHzp[n− 1]zHp [n− 1]w]

− 2ℜ{E[wHzp[n− 1]x∗[n]]}− 2ℜ{E[wHzp[n− 1]zHp [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[wHzp[n− 1]x∗[n]] = E[wHzp[n− 1]zHp [n− 1]w] = 0,(65)

and

E[wHzp[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

)

= σ2ϕ2

2p

N − p. (66)

Substituting these in (64), we can write

E[|x[n]|2] = σ2ϕ2

(

1 +2p

N − p

)

. (67)

We can similarly argue, for the one variable model, thatx[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 fromthe one variable and two variable regression models of thesystem can 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

(

t1

t2

)

, (71)

Page 7: Statistical Tests for Detecting Granger Causalitycmurthy/lib/exe/fetch.php?media=... · 2018. 9. 15. · The Granger causality index and its derivatives are important metrics developed

7

However, sinceTG is a logarithm of ratios of random variables,its statistics become hard to determine. Instead, we defineTE , 2−TG = t2

t1as a test statistic, for simplicity of analysis.

Sincex[n] andx[n] are the measurement errors for the sameprocess,t1 andt2 cannot be considered independent. However,since botht1 andt2 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 isvalid when the number of samplesN is much larger than themodel orderp. Now

E[t2] = σ2ϕ2

(

1 +2p

N − p

)

, (72)

and

var(t2) =1

N − pσ4ϕ2

(

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] andx[n], wHzp[n− 1] will be a better estimate ofx[n]than hHxp[n − 1], therefore,wHzp[n − 1] can be writtenas wHzp[n − 1] = hHxp[n − 1] + x[n], such that,x[n] isorthogonal tohHxp[n− 1]. Consequently,

ϕ1[n] = ϕ2[n] + x[n], (75)

with E[x[n]ϕ∗2[n]] = 0, and

σ2ϕ1

= σ2ϕ2

+ σ2x, (76)

whereσ2x , var(x2).

Now, sincex[n] = wHzp[n] − hHxp[n], it can be shownthatE[x[n]] = 0, and,

E[|x[n]|2] = wHRzz,pw+ hHRxx,ph− 2ℜ{wHRzxh}= rHzz,pR

−1zz,przz,p + rHxx,pR

−1xx,prxx,p

−2ℜ{rHzz,pR−1zz,pRzx,pR

−1xx,prxx,p}. (77)

Therefore,x[n] = x[n] + x[n]− hHxp[n− 1] and

E[t1] =(

σ2ϕ2

+ σ2x

)

(

1 +p

N − p

)

= E[t2] + σ2x2

(

1 +p

N − p

)

− σ2ϕ2

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

(

σ4ϕ2

+ σ4x + 2σ2

ϕ2σ2x

)

×(

1 +p2

(N − p)2+

2p

N − p

)

= var(t2) +1

N − p

(

σ4x + 2σ2

ϕ2σ2x

−σ4ϕ2

(

3p2

(N − p)2+

2p

N − p

))

. (80)

Consequently, it can be shown that

cov(t1, t2) =σ4ϕ2

N − p. (81)

The GCI can therefore be approximated as the ratio of twocorrelated Gaussian r.v.st1 andt2 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 ratioTE = t2

t1is below a thresholdλ. Since

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

Pr{TE < λ} = Pr{λt1 − t2 > 0}. (82)

Sincet1 andt2 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

)

−λσ2ϕ2

p

N − p,

(83)

and

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

respectively.

Now, under the null hypothesis,H0, σ2ϕ1

= σ2ϕ2

andσ2x|H0

= 0. Hence, underH0, the above simplify to

µTD ,0 , E[TD|H0] =

(λ− 1)σ2ϕ2

(

1 +p

N − p

)

− σ2ϕ2

(

p

N − p

)

, (85)

σ2TD ,0 , var(TD|H0)

= (λ2 + 1)σ4ϕ2

N − p

(

1 +2p

N − p

)2

− λ2σ4ϕ2

N − p

(

3p2

(N − p)2+

2p

N − p

)

− 2λσ4ϕ2

N − p. (86)

Similarly, under the alternate hypothesis,H1, σ2ϕ1

> σ2ϕ2

,

Page 8: Statistical Tests for Detecting Granger Causalitycmurthy/lib/exe/fetch.php?media=... · 2018. 9. 15. · The Granger causality index and its derivatives are important metrics developed

8

and therefore,

µTD ,1 , E[TD|H1] = λσ2x

(

1 +p

N − p

)

+ (λ− 1)σ2ϕ2

(

1 +p

N − p

)

− σ2ϕ2

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

+ 2λ2σ2ϕ2σ2x2

N − p

(

1 +p

N − p

)2

− λ2σ4ϕ2

N − p

(

3p2

(N − p)2+

2p

N − p

)

− 2λσ4ϕ2

N − p. (88)

Based on these, the probability of false alarm can becalculated as

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

(

µTD,0

σTD ,0

)

. (89)

Similarly, the probability of correct detection of a causalrelationship,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 ofP ∗

FA can be designed by selecting a thresholdλ that satisfies

[

σ2ϕ2(λ− 1)

(

1 +p

N − p

)

− λσ2ϕ2

p

N − p

]2

= Q−1(P ∗FA)

[

(λ2 + 1)σ4ϕ2

N − p

(

1 +p

N − p

)2

− λ2σ4ϕ2

N − p

(

3p2

(N − p)2+

2p

N − p

)

− 2λσ4ϕ2

N − p

]

. (91)

The above reduces to a quadratic equation inλ that can besolved to obtain the detection threshold for the detector.

As an example, considerN samples of signalsu[n] andv[n] corrupted by AWGN and observed asx[n] and y[n],respectively, such thatv[n] = ǫv[n] and

u[n] =

{

ǫu,0[n] H0

aǫv[n] + ǫu,1[n] H1, (92)

with var(ǫv) = σ2ǫv

, var(ǫu,1) = σ2ǫu

, and var(ǫu,0) =|a|2σ2

ǫv+ σ2

ǫu. Then,K = 1, σ2

ϕ1is given by (58), and

σ2ϕ2

=

{

σ2ϕ1

H0

σ2ǫu

+ σ2νx

+|a|2σ2

ǫvσ2

νy

σ2ǫu

+σ2νy

= σ2ϕ1

−∆ H1, (93)

with ∆ = |a|2σ2ǫv

(

1− σ2

νy

σ2ǫv

+σ2νy

)

.

Therefore,

µTD ,0 = σ2ϕ1

(

λ− 1 +λ− 2

N − 1

)

, (94)

and

σ2TD ,0 =

1

N − 1

[

(λ2 + 1)σ4ϕ1

(

1 +2

N − 1

)2

− λ2σ4ϕ1

(

3

(N − 1)2+

2

N − 1

)

− 2λσ2ϕ1

N − 1

]

. (95)

Consequently, the probability of false alarm is given as (96)at the top of the next page.

Under the alternate hypothesis,

E[TD|H1] = (λ− 1)σ2ϕ1

(

1 +1

N − 1

)

−∆

(

1 +1

N − 1

)

− (σ2ϕ1

−∆)1

N − 1, (97)

and

var(TD|H1) =1

N − 1

[

(λ2 + 1)σ4ϕ1

(

1 +1

N − 1

)2

+∆2

(

1 +2

N − 1

)2

− 2σ2ϕ2∆

(

1 +2

N − 1

)2

+ σ4ϕ1

(

3

(N − 1)2+

2

N − 1

)

]

. (98)

The probability of detection can thus be expressed as (99) inthe next page.

We observe from (99) that the probability of detectiondepends on the ratio∆

σ2ϕ1

, which can in turn be viewed asthe relative difference between the variances of the predictionerror under the one variable and the two variable models. Con-sequently, the difference between the mean square predictionerror for the single and two variable prediction models, i.e.,t1 − t2 can be used as a test statistic to detect the presence ofa causal relationship. We inspect the use oft1 − t2 as a teststatistic in the next section.

VI. GRANGER CAUSALITY DETECTION BASED ON THE

DIFFERENCE OFERROR 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 thatTC isthe difference of two correlated Chi-squared r.v.s., each with2(N − p) degrees of freedom. Therefore, the pdf ofTC canbe written as [37]

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;

Page 9: Statistical Tests for Detecting Granger Causalitycmurthy/lib/exe/fetch.php?media=... · 2018. 9. 15. · The Granger causality index and its derivatives are important metrics developed

9

PFA = Q

(N − 1)(

λ− 1 + λ−2N−1

)

(λ2 + 1)σ4ϕ2

(

1 + 2N−1

)2

− λ2σ4ϕ2

(

3(N−1)2 + 2

N−1

)

− 2λσ4ϕ1

N−1

. (96)

PD = Q

√N − 1

(

(λ− 1)σ2ϕ1

(

1 + 1N−1

)

−∆(

1 + 1N−1

)

− (σ2ϕ1

−∆) 1N−1

)

(λ2 + 1)σ4ϕ1

(

1 + 1N−1

)2

+∆2(

1 + 2N−1

)2

− 2σ2ϕ2∆(

1 + 2N−1

)2

+ σ4ϕ1

(

3(N−1)2 + 2

N−1

)

. (99)

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)]

1

2

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. ofTC is in the form of a series, which makes itdifficult to obtain the probabilities of detection and falsealarmin closed form. However, for largeN , 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

)

− σ2ϕ2

p

N − p, (104)

and

var(TC |H0) =1

N − p

[

2σ4ϕ2

(

1 +2p

N − p

)2

− σ4ϕ2

(

3p2

(N − p)2+

2p

N − p

)

− 2σ4ϕ2

]

. (105)

var(TD|H1) =1

N − p

[

2σ4ϕ2

(

1 +2p

N − p

)2

+ σ4x

(

1 +p

N − p

)2

+ 2σ2ϕ2σ2x2

(

1 +p

N − p

)2

− σ4ϕ2

(

3p2

(N − p)2+

2p

N − p

)

− 2σ4ϕ2

]

. (106)

Using the above statistics, it is now straightforward to designthe detector to achieve a given probability of false alarm,and obtain closed-form expressions for the probability of falsealarm and detection, similar to (89) and (90), respectively. Inparticular, whenN ≫ p, i.e., the number of samples is muchlarger than the model order, it can be easily shown that the

detection thresholdλ is well approximated by

λ =

√2Q−1(PFA)σ

2ϕ2√

N − 1, (107)

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

FA can be approximated as

PD ≈ Q

Q−1 (P ∗FA)

√2σ2

ϕ2−√

N − pσ2x

σ4x + 2σ2

ϕ2σ2x2

. (108)

Based on the above, the number of samples required to attaina given pair of probabilities of detection and false alarm canbe approximated as

N ≈ p+

(

Q−1(PFA)√2σ2ϕ2

σ2x

+Q−1(PD)

1 + 2σ2ϕ2

σ2x

)2

.

(109)

Revisiting the example considered in the previous sections,we observe that,σ2

ϕ1= σ2

x, σ2ϕ2

= σ2x − ∆, using which, it

is easy to obtain the expressions for the probabilities of falsealarm and detection. We omit writing out the expressions forthe sake of 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 betweenx[n] andy[n]. Letting x[n] , x[n]−x[n],and

TA ,1

N − 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 + rHxx,pR

−1xx,prxx,p

− 2ℜ{rHR−1zz,pRzxR

−1xx,prxx,p}+

2σ2ϕ2p

N − p+

σ2ϕ1p

N − p. (112)

However, the higher order statistics and the exact distribu-tion of TA cannot be found in closed form. Therefore, weevaluate its performance via simulations. To the best of ourknowledge, the test statisticsTC andTA have not been usedin the literature to detect Granger causality.

Page 10: Statistical Tests for Detecting Granger Causalitycmurthy/lib/exe/fetch.php?media=... · 2018. 9. 15. · The Granger causality index and its derivatives are important metrics developed

10

-20 -10 0 10 20

1(dB)

0

0.05

0.1

0.15

0.2

0.25

0.3

TG

2=20dB

2=15dB

2=10dB

2=5dB

2=0dB

2=-5dB

2=-10dB

2=-15dB

2=-20dB

Fig. 1: Value ofTE at different SNRs, fora = 0.5, with knownsecond-order statistics.

VII. S IMULATION 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 perfor-mance ofTC using real-world data with known ground truth,available in [38]

For the simulation-based study, we consider a system modelsimilar to the example discussed in the previous sections,for different values of the variance of the additive noise andvarying number of samples. The probabilities of detection andfalse alarm are obtained by averaging over10, 000 independentrealizations of the signals of interest.

A. Effect of Additive Noise

In Fig. 1, we plot the mean value of the test statisticTG forthe example considered in Section IV, withb = 0.5. Here weinspect the effects of noise on the GCI. We defineγ1 ,

σ2

u

σν1

,

and γ2 ,σ2

v

σν2

as the observation SNRs forx[n] and y[n].We see that the mean value ofTG converges to a value closeto zero as the noise variance increases. The reduction in theSNR will make it harder to detect the presence of a causalrelationship between the signals of interest.

B. Finite Sample Effects

We now consider the effects of using a finite number ofsamples on the GCI, as discussed in Section V. In Figs. 2aand 2b, we compare the simulated detection and false alarmprobability obtained by using the test statisticTE as a functionof the detection thresholdλ and for different number ofsamples, witha = 0.25 andγ1 = γ2 = 0 dB. The simulatedperformance ofTE is found to be in close agreement withthe values computed using our analytical expressions in (96)and (99). It is also observed that the behavior of both theprobability of detection and false alarm as a function of the

0.8 0.9 1 1.1 1.20

0.2

0.4

0.6

0.8

1

PD N=500

N=200N=100N=50

(a)

0.8 0.9 1 1.1 1.20

0.2

0.4

0.6

0.8

1

N=50N=100N=150N=200

Line: TheoryMarker: Simulation

(b)

Fig. 2: Probabilities of (a) detection and (b) false alarm asafunction of the detection thresholdλ for the test statisticTE.

detection threshold becomes sharper with an increase in thenumber of samples, and asymptotically approximates a stepfunction. The transition point of the step function occurs atthe mean of the test statistic under the respective hypotheses,which corresponds to the 50% probability of detection/ falsealarm point on Figs. 2a and 2b. Therefore, the thresholdλ = 1gives a zero error rate when infinite samples of the two signalsof interest are available. However, in case of a finite numberof samples,λ has to be chosen carefully to achieve reliableperformance.

In Fig. 3, we compare the theoretical and simulated receiveroperating characteristics (ROCs) for the example consideredin Section V for a = 0.25 with different values ofN , atan SNR of0 dB. It is observed that for a small number ofsamples, the ROC is close to thePFA = PD line, whereas,for a large number of samples, the probability of detectionbecomes almost independent ofPFA. This is in line with thebehavior predicted in (96) and (99).

Having established that the use of finite samples indeedcauses a deterioration in the performance of the CGI, we nowdiscuss the performance of the other two test statistics, viz.TC

andTA, as discussed in Section VI. In Fig. 4a, we compare thetheoretical and simulated ROCs for the example considered inSection VI, witha = 0.25, for different number of samples,

Page 11: Statistical Tests for Detecting Granger Causalitycmurthy/lib/exe/fetch.php?media=... · 2018. 9. 15. · The Granger causality index and its derivatives are important metrics developed

11

0 0.2 0.4 0.6 0.8 1P

FA

0

0.2

0.4

0.6

0.8

1P

D N=1000N=500N=200N=100N=50

Line: TheoryMarker: Simulation

Fig. 3: ROC for detection based onTE , at an SNR of0 dB,for different number of samples.

and at an SNR of0 dB. We observe that the ROC forTC

shows a behavior similar to that forTE with an increase inthe number of samples. Therefore, an appropriate choice ofthe detection thresholdλ becomes important forTC as well.

In practice, one is typically provided a given number ofsamples of a pair of time-series data to perform detection.In this case, there is a trade-off between the window sizeN and the quality of the ROC obtained. Using a largerN

improves the reliability of detection, but the resulting ROC hasmore “jumps” because it is computed by averaging over fewerinstantiations. We illustrate this in Fig. 4b, where we plottheROC for the setup of Fig. 4a with the total number of availablesamples fixed at10, 000. We see that the ROC withN = 100(averaged over100 instantiations) matches well with thecorresponding curve in Fig. 4a, where10, 000 instantiationswere used. For higherN , the performance improves, but thequality of the ROC deteriorates, because the probability ofdetection is averaged over a correspondingly fewer numberof instantiations. Thus, given a data record of a given size,agood choice forN is thus the largest one such that the ROCis averaged over about100 instantiations.

In Fig. 5, we repeat the experiment considered in Fig. 4aby fixing the number of samples ofx[n] and y[n] at 100and varying the SNRs. Again, the theoretical and simulatedplots are in close agreement, with the concavity of the ROCincreasing with the SNR.

In Figs. 6 and 7, we evaluate the performance ofTA as atest statistic for detection of a causal relation betweenx[n]and y[n] for a setup similar to the example considered inSection V, witha = 0.25. Since theoretical expressions forthe behavior ofTA are not derived, these are not consideredhere. In Fig. 6, we plot the ROC ofTA for different SNRs for100 samples of each ofx[n] andy[n] being used for detection.The performance ofTA under different SNRs is seen to mimicthat ofTC andTE, with the concavity of the ROC increasingwith the SNR. In Fig 7, we plot the probabilities of detectionfor different numbers of samples as a function of the signal

0 0.2 0.4 0.6 0.8 1P

FA

0

0.2

0.4

0.6

0.8

1

PD

N=1000N=500N=200N=100

Line: TheoryMarker: Simulation

(a)

0 0.2 0.4 0.6 0.8 1P

FA

0

0.2

0.4

0.6

0.8

1

PD

N=1000, 10 InstantiationsN=500, 20 InstantiationsN=200, 50 InstantiationsN=100, 100 Instantiations

(b)

Fig. 4: ROC for detection based onTC , at an SNR of0 dB,for different number of samples (N ) (a) averaged over 10,000realizations in all cases, (b)obtained for a fixed data recordlength (10,000) divided into windows of lengthN .

SNRs. In this case, we use the Neyman-Pearson criterion tofix the probability of false alarm at 10%. As expected, anincrease in the number of samples leads to improved detectionperformances for the same SNR. Therefore,TA can also beused to detect a causal relationship between a pair of signalsin the finite sample regime.

In Fig. 8, we compare the simulated performance ofTE,TC andTA for the example considered previously, withb =0.25, at an SNR of 0 dB and for 100 samples. We observethatTA, despite being theoretically intractable, offers the bestperformance among the three test statistics considered in thispaper, followed byTC and TE respectively. It is interestingto note that if the exact knowledge about the second-orderstatistics is available, then the performance of the three teststatistics should be identical.

Page 12: Statistical Tests for Detecting Granger Causalitycmurthy/lib/exe/fetch.php?media=... · 2018. 9. 15. · The Granger causality index and its derivatives are important metrics developed

12

0 0.2 0.4 0.6 0.8 1P

FA

0

0.2

0.4

0.6

0.8

1P

D

SNR=6dBSNR=3dBSNR=0dBSNR=-3dBSNR=-6dB

Line: TheoryMarker: Simulation

Fig. 5: ROC for detection based onTC , with N = 200samples, for various values of the SNR.

0 0.2 0.4 0.6 0.8 1P

FA

0

0.2

0.4

0.6

0.8

1

PD SNR= 6dB

SNR=3 dBSNR=0dBSNR=-3 dBSNR=-6 dB

Fig. 6: ROC for detection based onTA, with N = 100samples, for various values of the SNR.

C. Detection Performance ofTC with Real World Data

In Figs. 9 and 10, we plot the performance obtained fromthe test statisticTC on a real-world dataset with knownground truth, available in [38]. Fig 9 corresponds to datapairs 65 − 67 of this database. These data pairs consist ofthe returns from related stocks over a period of five years.Similarly, Fig 10 corresponds to data pair69 of the database,containing temperature measurements from inside and outsidea room taken every five minutes. These four data sets wereselected out of the108 available data sets because:

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

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

-20 -10 0 10 20SNR

0

0.2

0.4

0.6

0.8

1

PD

N=50N=100N=200N=500N=1000

Fig. 7: Probability of detection usingTA at a fixed false alarmrate of 10% as a function of the SNR.

0 0.2 0.4 0.6 0.8 1P

FA

0

0.2

0.4

0.6

0.8

1

PD

TA

TC

TE

Fig. 8: ROC of test statisticsTE , TC and TA for N = 100samples at an SNR of 0 dB.

For generating the experimental results, we assumed thatthe data was uniformly sampled at Nyquist rate. However,neither the model parameters of the data nor the additivenoise variance are known. Hence, an order 1-VAR generativemodel is assumed, and each data set is preprocessed to makeit zero mean. The detection threshold for a given false alarmrate (PFA) is calculated assuming the two sequences to beindependent and white, and using the sample variance of thecaused sequence asσ2

ϕ2in (107).

The data pairs are divided into windows of lengthN , andthe test statisticTC is calculated and compared to the thresholdin (107). The results of these comparisons are averaged overall the sample windows to obtain the probability of detectionfor a given probability of false alarm. The jitters in the plotsare due to the averaging over the relatively small numberof samples that were available. The performance of data set69 is poorer than data sets 65-67, for the same value ofN ,indicating either a weaker causal relationship, or larger amountof additive noise, or both. Nonetheless, the ROC follows apattern similar to that predicted by (108).

Page 13: Statistical Tests for Detecting Granger Causalitycmurthy/lib/exe/fetch.php?media=... · 2018. 9. 15. · The Granger causality index and its derivatives are important metrics developed

13

0 0.2 0.4 0.6 0.8 1P

FA

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1P

D

N=50N=20N=10

Fig. 9: Performance of the test statisticTC on the real-worlddatasets67− 69 in [38] for different numbers of samples.

0 0.2 0.4 0.6 0.8 1P

FA

0

0.2

0.4

0.6

0.8

1

PD

N=200N=150N=100

Fig. 10:Performance of the test statisticTC on the real-worlddataset69 in [38] for different numbers of samples.

In Fig. 11, we plot the performance ofTC for real-worlddata used in Fig. 10 with different amounts of added noise fora window length ofN = 200. Noisy samples are obtainedby adding real valued white Gaussian noise to the causedsequence, suitably scaled to achieve a given SNR. Thesesamples are used to calculate the test statisticTC , which isthen compared against the detection threshold in (107), withσ2ϕ1

as the sample variance of the corrupted caused signal. It isobserved that for a fixed number of samples per window, theaddition of noise worsens the detection performance ofTC ,and the variation in the detection performance approximatelymatches with theQ-function of the square root of the SNR,as expected.

ACKNOWLEDGMENTS

The authors are grateful to the anonymous reviewers fortheir feedback, which significantly improved the overall pre-

0 0.2 0.4 0.6 0.8 1P

FA

0

0.2

0.4

0.6

0.8

1

1.2

PD

No added noiseSNR=20 dBSNR=10 dBSNR=0 dB

Fig. 11: Performance ofTC for the dataset considered inFig. 10 forN = 200 with different amounts of added noise.

sentation, and in particular the simulation results in the paper.

VIII. C ONCLUSIONS

In this work, we derived the effects of different dataacquisition impairments on the detection of a causal relationbetween two signals. We showed that these effects, viz. down-sampling, additive noise, and finite sample effects, could leadto a significant deterioration in the performance of the GCI,both individually as well as cumulatively. We derived theprobabilities of detection and false alarm for the GCI undertheimpairments. Following this, we proposed two alternative teststatistics, based on the difference of mean squared error terms,and the mean squared value of the difference of error terms.Via extensive simulations, we showed that the derived resultscorroborate the simulated as well as real-world data, with thetest statistic based on the mean squared value of the differenceof error terms outperforming the other two statistics. A moreprecise analysis of this test statistic is an interesting directionfor future research.

REFERENCES

[1] I. Kontoyiannis and M. Skoularidou, “Estimating the directed infor-mation and testing for causality,”IEEE Trans. Inf. Theory, vol. 62,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,”Journal of the AmericanStatistical Association, 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 datausing vectorautoregressive modeling and Granger causality mapping,”MagneticResonance Imaging, vol. 21, no. 10, pp. 1251 – 1261, 2003.

[5] W. Hesse, E. Mller, M. Arnold, and B. Schack, “The use of time-variant EEG Granger causality for inspecting directed interdependenciesof neural assemblies,”Journal of Neuroscience Methods, vol. 124, no. 1,pp. 27 – 44, 2003.

[6] Y. Chen, G. Rangarajan, J. Feng, and M. Ding, “Analyzing multiplenonlinear time series with extended Granger causality,”Physics LettersA, vol. 324, no. 1, pp. 26 – 35, 2004.

Page 14: Statistical Tests for Detecting Granger Causalitycmurthy/lib/exe/fetch.php?media=... · 2018. 9. 15. · The Granger causality index and its derivatives are important metrics developed

14

[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. Seventh International Symposium on NeuralNetworks (ISNN 2010)Advances in Web Intelligence.

[9] D. Sampath, K. Sabitha, P. Hegde, H. Jayakrishnan, B. M. Kutty,S. Chattarji, G. Rangarajan, and T. Laxmi, “A study on fear memoryretrieval and{REM} sleep in maternal separation and isolation stressedrats,” Behavioural Brain Research, 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 analysisin neuroscience and neuroimaging,”Journal of Neuroscience, 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, p. 016211,Jan. 2007.

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

[14] E. Kodra, S. Chatterjee, and A. R. Ganguly, “Exploring Grangercausality between global average observed time series of carbon dioxideand temperature,”Theoretical and Applied Climatology, vol. 104, no. 3,pp. 325–335, 2011.

[15] H. White and D. Pettenuzzo, “Granger causality, exogeneity, coin-tegration, and economic policy analysis,”Journal of Econometrics,vol. 178, Part 2, pp. 316 – 330, 2014. Recent Advances in Time SeriesEconometrics.

[16] H. Nalatore, M. Ding, and G. Rangarajan, “Mitigating the effectsof measurement noise on Granger causality,”Phys. Rev. E, vol. 75,p. 031123, Mar 2007.

[17] H. Nalatore, S. N, and G. Rangarajan, “Effect of measurement noise onGranger causality,”Phys. Rev. E, vol. 90, p. 062127, Dec 2014.

[18] I. Winkler, D. Panknin, D. Bartz, K. R. Mller, and S. Haufe, “Validityof time reversal for testing Granger causality,”IEEE Transactions onSignal Processing, vol. 64, pp. 2746–2760, June 2016.

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

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

[21] E. Siggiridou and D. Kugiumtzis, “Granger causality inmultivariatetime series using a time-ordered restricted vector autoregressive model,”IEEE Transactions on Signal Processing, vol. 64, pp. 1759–1773, April2016.

[22] P. Newbold and N. Davies, “Error mis-specification and spurious re-gressions,”International Economic Review, vol. 19, no. 2, pp. 513–19,1978.

[23] G. Nolte, A. Ziehe, V. V. Nikulin, A. Schlogl, N. Kramer, T. Brismar,and K.-R. Muller, “Robustly estimating the flow direction of informationin complex physical systems,”Phys. Rev. Lett., vol. 100, p. 234101, Jun2008.

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

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

[26] M. Vinck, L. Huurdeman, C. A. Bosman, P. Fries, F. P. Battaglia, C. M.Pennartz, and P. H. Tiesinga, “How to detect the Granger-causal flowdirection in the presence of additive noise?,”NeuroImage, vol. 108,pp. 301 – 318, 2015.

[27] A. K. Seth, P. Chorley, and L. C. Barnett, “Granger causality analysis offMRI 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 forsubsampled continuous-time neurophysiological processes,” Journal ofNeuroscience Methods, 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,” inProceedings ofthe 32nd International Conference on Machine Learning(F. Bach andD. Blei, eds.), vol. 37 ofProceedings of Machine Learning Research,(Lille, France), pp. 1898–1906, PMLR, 07–09 Jul 2015.

[30] M. Gong, K. Zhang, B. Scholkopf, C. Glymour, and D. Tao,“Causaldiscovery from temporally aggregated time series,” inProceedingsConference on Uncertainty in Artificial Intelligence (UAI)2017, p. ID269, Association for Uncertainty in Artificial Intelligence (AUAI), Aug.2017.

[31] A. Tank, E. B. Fox, and A. Shojaie, “Identifiability of non-gaussianstructural var models for subsampled and mixed frequency time series,”in Proceedings SIGKDD Workshop on Causal Discovery 2016, 2016.

[32] A. Hyttinen, S. Plis, M. Jrvisalo, F. Eberhardt, and D. Danks, “Aconstraint optimization approach to causal discovery fromsubsampledtime series data,”International Journal of Approximate Reasoning,vol. 90, pp. 208 – 225, 2017.

[33] S. Plis, D. Danks, C. Freeman, and V. Calhoun, “Rate-agnostic (causal)structure learning,” inAdvances in Neural Information ProcessingSystems 28(C. Cortes, N. D. Lawrence, D. D. Lee, M. Sugiyama, andR. Garnett, eds.), pp. 3303–3311, Curran Associates, Inc.,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, 072013.

[35] S. Haykin,Adaptive Filter Theory. Pearson Education, 2002.[36] R. Tandra and A. Sahai, “SNR walls for signal detection,” IEEE Journal

of Selected Topics in Signal Processing, vol. 2, pp. 4–17, Feb 2008.[37] M. K. Simon, Probability Distributions Involving Gaussian Random

Variables. Springer US, 2002.[38] J. M. Mooij, J. Peters, D. Janzing, J. Zscheischler, andB. Scholkopf,

“Distinguishing cause from effect using observational data: Methods andbenchmarks,”Journal of Machine Learning Research, 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 IndianInstitute of Technology Roorkee, India in 2011 and2016 respectively. He worked as a project associateat Department of Electrical Communication Engi-neering, Indian Institute of Science, Bangalore fromAug. 2015, till May 2016. From May 2016 to March2017 he worked as an institute research associate at

the Department of Electrical Communication Engineering, Indian Institute ofScience, Bangalore, India. In April 2017, he joined the department of Elec-tronics and Electrical Engineering, Indian Institute of Technology Guwahati,Assam, India. His research interests include statistical and adaptive signalprocessing, massive MIMO communications, and cognitive communications.

Page 15: Statistical Tests for Detecting Granger Causalitycmurthy/lib/exe/fetch.php?media=... · 2018. 9. 15. · The Granger causality index and its derivatives are important metrics developed

15

Chandra R. Murthy (S’03–M’06–SM’11) receivedthe B. Tech. degree in Electrical Engineering fromthe Indian Institute of Technology Madras, India, in1998, the M. S. and Ph. D. degrees in Electricaland Computer Engineering from Purdue Universityand the University of California, San Diego, USA,in 2000 and 2006, respectively. From 2000 to 2002,he worked as an engineer for Qualcomm Inc., wherehe worked on WCDMA baseband transceiver designand 802.11b baseband receivers. From Aug. 2006to Aug. 2007, he worked as a staff engineer at

Beceem Communications Inc. on advanced receiver architectures for the802.16e Mobile WiMAX standard. In Sept. 2007, he joined the Departmentof Electrical Communication Engineering at the Indian Institute of Science,Bangalore, India, where he is currently working as a Professor.

His research interests are in the areas of energy harvestingcommunications,multiuser MIMO systems, and sparse signal recovery techniques appliedto wireless communications. His paper won the best paper award in theCommunications Track at NCC 2014 and a paper co-authored with his studentwon the student best paper award at the IEEE ICASSP 2018. He has 50+journal papers and 80+ conference papers to his credit. He was an associateeditor for the IEEE Signal Processing Letters during 2012-16. He is an electedmember of the IEEE SPCOM Technical Committee for the years 2014-16,and has been re-elected for the 2017-19 term. He is a past Chair of the IEEE

Signal Processing Society, Bangalore Chapter. He is currently serving as anassociate editor for the IEEE Transactions on Signal Processing, the SadhanaJournal and as an editor for the IEEE Transactions on Communications.

Govindan Ranagarajan obtained his IntegratedM.Sc. from Birla Institute of Technology and Sci-ence, Pilani, India and his PhD from Universityof Maryland, College Park, USA. He then workedat the Lawrence Berkeley Lab, University of Cali-fornia, Berkeley before returning to India in 1992.He is currently a Professor of Mathematics at theIndian Institute of Science, Bangalore. He is alsothe 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 andthe National Academy of Sciences, India. He was knighted by the Governmentof France: Knight (Chevalier) of the Order of Palms. He was also a HomiBhabha Fellow.