University of Torontozhou/papers/dcf_Nov7_4.pdf · University of Toronto September 15, 2014...

34
MEASURING NONLINEAR DEPENDENCE IN TIME SERIES, A DISTANCE CORRELATION APPROACH By Zhou Zhou 1 University of Toronto September 15, 2014 Abstract We extend the concept of distance correlation of Szekely, Rizzo and Bakirov (The Annals of Statistics, 2007) and propose the auto distance correlation function (ADCF) to measure the temporal dependence structure of time series. Unlike the classic measures of correlations such as the autocorrelation function, the proposed measure is zero if and only if the measured time series components are independent. In this paper, we propose and theoretically verify a subsampling methodology for the inference of sample ADCF for dependent data. Our methodology provides a useful tool for exploring nonlinear dependence structures in time series. 1 Introduction Measuring the temporal dependence structure is of fundamental importance in time series analysis. The autocorrelation function (ACF) (Box and Jenkins, 1970), which measures the strength of linear dependencies in time series, has been one of the primary tools for exploring and testing time series dependence for many decades. For a univariate time series, the ACF measures the Pearson correlation between observations of the series. In particular, the ACF equals 0 when the measured observations are not at all linearly related and it achieves its maximum absolute value 1 if measured observations are perfectly linearly related. 1 Corresponding author. Department of Statistics, 100 St. George St., Toronto, Ontario, M5S 3G3 Canada E-mail: [email protected] Key words and phrases. Auto distance correlation function, nonlinear time series, nonlinear dependence, subsampling. 1

Transcript of University of Torontozhou/papers/dcf_Nov7_4.pdf · University of Toronto September 15, 2014...

Page 1: University of Torontozhou/papers/dcf_Nov7_4.pdf · University of Toronto September 15, 2014 Abstract We extend the concept of distance correlation of Szekely, Rizzo and Bakirov (The

MEASURING NONLINEAR DEPENDENCE IN TIME SERIES,

A DISTANCE CORRELATION APPROACH

By Zhou Zhou1

University of Toronto

September 15, 2014

Abstract

We extend the concept of distance correlation of Szekely, Rizzo and Bakirov

(The Annals of Statistics, 2007) and propose the auto distance correlation function

(ADCF) to measure the temporal dependence structure of time series. Unlike the

classic measures of correlations such as the autocorrelation function, the proposed

measure is zero if and only if the measured time series components are independent.

In this paper, we propose and theoretically verify a subsampling methodology for the

inference of sample ADCF for dependent data. Our methodology provides a useful

tool for exploring nonlinear dependence structures in time series.

1 Introduction

Measuring the temporal dependence structure is of fundamental importance in time series

analysis. The autocorrelation function (ACF) (Box and Jenkins, 1970), which measures

the strength of linear dependencies in time series, has been one of the primary tools for

exploring and testing time series dependence for many decades. For a univariate time

series, the ACF measures the Pearson correlation between observations of the series. In

particular, the ACF equals 0 when the measured observations are not at all linearly related

and it achieves its maximum absolute value 1 if measured observations are perfectly linearly

related.

1Corresponding author. Department of Statistics, 100 St. George St., Toronto, Ontario, M5S 3G3

Canada

E-mail: [email protected]

Key words and phrases. Auto distance correlation function, nonlinear time series, nonlinear dependence,

subsampling.

1

Page 2: University of Torontozhou/papers/dcf_Nov7_4.pdf · University of Toronto September 15, 2014 Abstract We extend the concept of distance correlation of Szekely, Rizzo and Bakirov (The

Nevertheless, the traditional linear time series models are inadequate in explaining

many phenomenons observed in contemporary time series, such as volatility clustering,

asymmetric cycles, extreme value dependence, time irreversibility, bimodality and mean

reverting, among others. This leads to a recent surge in nonlinear time series research.

Prominent examples include Brillinger (1977), Engle (1982), Ashley et. al (1986), Priestley

(1988), Tong (1990), Chen and Tsay (1993), Franses and van Dijk (2000), Fan and Yao

(2003), Kantz and Schreiber (2000), Tjøstheim and Auestad (1994a, b) and Brockwell

(2007), among others.

Despite the fast advancement in nonlinear time series model building, on the other

hand, there have been few works on how to explore and measure the complex dependence

structures in nonlinear time series. Exceptions include, among others, Bagnato et. al

(2011) who considered a graphical device of displaying general dependence structure in time

series using the χ2 tests of independence in contingency tables. It seems that nowadays the

Pearson correlation related quantities, which measure the amount of linear dependence,

are frequently used to explore the temporal dependence structures in nonlinear time series.

Recently Szekely, Rizzo and Bakirov (2007) (SRB hereafter) proposed the distance

correlation to measure and test linear and nonlinear dependence between two samples

each composed of iid observations; see also Szekely and Rizzo (2009). The purpose of this

paper is to extend the concept of distance correlation into temporally dependent data and

analyze a corresponding measure of time series dependence, the auto distance correlation

function (ADCF), to explore and test nonlinear dependence structures in time series.

The distinct feature of ADCF is that the theoretical ADCF equals 0 if and only if

the measured time series components are independent. The latter feature implies that

theoretical ADCF is capable of measuring all forms of departures from independence. In

particular, the ADCF is capable of digging out complex nonlinear dependence structures

which are buried under the Pearson correlation related measures. For instance, it is well

known that many financial return time series show no signs of Pearson correlation but the

squared returns exhibits strong dependence. Hence the latter financial returns are strongly

nonlinearly related. The ACF plot of the original return series would show no interesting

signal at all. However, the ADCF of the original series usually shows a strong signal of

dependence, which correctly reflects the underlying nonlinear dynamics of the return series.

See also Section 6.1 for a detailed discussion of the SP500 monthly excess return data.

2

Page 3: University of Torontozhou/papers/dcf_Nov7_4.pdf · University of Toronto September 15, 2014 Abstract We extend the concept of distance correlation of Szekely, Rizzo and Bakirov (The

Time series with heavy tailed marginal distributions are frequently observed in areas

such as finance, hydrology and telecommunications. For instance, Rachev and Mittnik

(2000) did an extensive empirical study and found that the tail indices of many high-

frequency financial return data are between 1 and 2. It is difficult and sometimes misleading

to apply Pearson correlation related measures to such time series; see for instance Figure

1 in Section 3 and Davis and Mikosch (1998). On the other hand, however, the ADCF is

defined for multivariate time series with finite first moment. In Section 4, we prove that

the sample ADCF is consistent as long as the time series has (1 + r)th moment for some

r > 0. Therefore one extra nice feature of the ADCF is that it can be used to measure the

strength of dependence for many heavy tailed multivariate time series.

Just as the sample ACF lets the data ‘speak for themselves’ and provides a first step

in the analysis of linear time series, the sample ADCF is a nonparametric measure which

allows one to visually explore the pattern of the strength of nonlinear dependence in sta-

tionary series and provides guidance for subsequent parametric or semi-parametric analysis.

Specifically, three major uses of the ADCF are as follows. First, the sample ADCF displays

the existence/nonexistence of temporal dependence of any kind which can be used to select

the time lag of nonlinear time series models and check the adequacy of certain parametric

nonlinear models by investigating the independence of its residuals. Second, the pattern

of sample ADCF, such as speed of decay or periodicity, can be used to do a preliminary

selection of nonlinear time series models: a model whose ADCF does not follow the latter

pattern should be removed from consideration. Third, cut-off of ADCF at a small lag

directly suggests a nonlinear m-dependent model; see Section 3.2 for more details.

One of the most important tasks in the study of ADCF is to perform finite sample

inference. The presence of temporal dependence invalidates the permutation test in SRB

for the latter purpose. In this paper, we propose a subsampling methodology for the

inference of the ADCF under dependence. Utilizing the contemporary martingale theory

and empirical process theory, we are able to control the uniform oscillation rate of the

empirical characteristic function of dependent data and consequently theoretically verify

the latter subsampling procedure for weakly dependent data. In fact, our theoretical

results can be generalized to facilitate the asymptotic theory of a class of V-statistics

under dependence. Simulation studies in Section 5.2 shows satisfactory performance of the

subsampling procedure for low dimensional time series.

3

Page 4: University of Torontozhou/papers/dcf_Nov7_4.pdf · University of Toronto September 15, 2014 Abstract We extend the concept of distance correlation of Szekely, Rizzo and Bakirov (The

The rest of the paper is structured as follows. Section 2 defines the ADCF function

and its sample version. Section 3 discusses the properties of the ADCF for multivariate

linear time series models, multivariate nonlinear moving average processes and multivariate

nonlinear auto regressive processes. Comparisons between the ADCF and the ACF for

the latter processes will also be performed in Section 3. In Section 4, we will establish

the consistency of the sample ADCF and obtain the asymptotic null distribution of the

statistic for testing ADCF= 0. In Section 5, a subsampling method is proposed for finite

sample inference of the ADCF. A small simulation study on the accuracy of the finite

sample ADCF test is conducted in Section 5.2. The SP500 data and the Canadian lynx

data are analyzed in Section 6. Some previously unfound dependence signals are uncovered

from our data analysis. Finally the asymptotic results are proved in Section 7.

2 Definitions of ADCF

Throughout this paper the Euclidean norm of x ∈ Rp is denoted by |x|. The inner product

on Rp is denoted 〈·, ·〉. For a complex valued function f(·), the complex conjugate of f is

denoted by f and |f |2 = ff . Denote by fX the characteristic function of a random vector

X and the joint characteristic function of random vectors X and Y is denoted fX,Y . The

symbol i denotes the imaginary unit with i2 = −1.

Motivated by the notion of distance correlation in SRB, for a strictly stationary time

series {Xj}, the auto distance correlation function (ADCF) at order k, k ≥ 0 is defined

as the distance correlation between Xj and Xj+k. For the completeness of the paper, we

briefly introduce the definitions and notation in this section.

Let {Xj}∞j=−∞ be a strictly stationary multivariate time series of dimension p and

we observe {Xj}nj=1. Throughout this paper we assume that p � n and therefore it is

reasonable to assume that p is fixed and do not vary with the series length n. It is well

known that Xj and Xj+k are independent if and only if fXj ,Xj+k(t, s) = fXj(t)fXj+k(s) for

almost all t and s in Rp. The latter property naturally leads to the following definition of

auto distance covariance between Xj and Xj+k.

Definition 1 (Auto Distance Covariance ). For k ≥ 0, the auto distance covariance and

4

Page 5: University of Torontozhou/papers/dcf_Nov7_4.pdf · University of Toronto September 15, 2014 Abstract We extend the concept of distance correlation of Szekely, Rizzo and Bakirov (The

sample auto distance covariance between Xj and Xj+k are defined as

VX(k) =1

c2p

∫Rp

∫Rp

|fXj ,Xj+k(t, s)− fXj(t)fXj+k(s)|2

|t|p+1|s|p+1dtds and (1)

VnX(k) =1

c2p

∫Rp

∫Rp

|fnk (t, s)− fn(t)fn,k(s)|2

|t|p+1|s|p+1dtds (2)

respectively, where cp = π(1+p)/2/Γ((1 + p)/2),

fnk (t, s) =1

n− k

n−k∑j=1

exp{i〈t,Xj〉+ i〈s,Xj+k〉}

is the empirical characteristic function of {(Xi, Xi+k)} and

fn(t) =n−k∑j=1

exp{i〈t,Xj〉}/(n− k), fn,k(t) =n−k∑j=1

exp{i〈t,Xj+k〉}/(n− k)

are the marginal empirical characteristic functions of {Xj}n−kj=1 and {Xj}nj=k+1, respectively.

We call VX(0) and VnX(0) the distance variance and sample distance variance of (Xj),

respectively.

It is clear from Definition 1 that VX(k) equals 0 if and only if Xj and Xj+k are in-

dependent. The quantity VX(k) equals the distance covariance between Xj and Xj+k in

SRB.

Definition 2 (Auto Distance Correlation). For k ≥ 1,define the auto distance correlation

between Xj and Xj+k

RX(k) =

√VX(k)

VX(0), if VX(0) 6= 0; (3)

Otherwise let RX(k) = 0. The sample auto distance correlation between Xj and Xj+k is

the nonnegative number RnX(k) defined by

[RnX(k)]2 =

VnX(k)√VnY (0)VnX∗(0)

, if VnY (0)VnX∗(0) 6= 0, (4)

where VnY (0) is the sample distance variance of {Yj}n−kj=1 , Yj = Xj+k and VnX∗(0) is the

sample distance variance of {Xj}n−kj=1 . Otherwise let RnX(k) = 0.

5

Page 6: University of Torontozhou/papers/dcf_Nov7_4.pdf · University of Toronto September 15, 2014 Abstract We extend the concept of distance correlation of Szekely, Rizzo and Bakirov (The

It can be shown that the auto distance covariance, the auto distance correlation are well

defined as long as Xj has finite first moment. Additionally, the theoretical ADCF achieves

its minimum 0 if and only if Xj and Xj+k are independent. And the theoretical ADCF

achieves its maximum 1 if Xj and Xj+k are perfectly linearly related by an orthogonal

matrix. We refer to SRB and Szekely and Rizzo (2009) for more details on the properties

of the distance correlation.

Recall that Yj = Xj+k for j = 1, 2, · · · , n− k. Let arl = |Xr −Xl| and brl = |Yr − Yl|.Define

ar· =

∑n−kl=1 arln− k

, a·l =

∑n−kr=1 arln− k

, a·· =

∑n−kr,l=1 arl

(n− k)2, Arl = arl − ar· − a·l + a··. (5)

Define br·, b·l, b·· and Brl similarly.

Proposition 1.

VnX(k) =1

(n− k)2

n−k∑r,l=1

ArlBrl. (6)

Proposition 1 is Theorem 1 of SRB with the second sample therein replaced by the

lagged observation of the time series. We restate it here because of its fundamental im-

portance in the theory of ADCF. Proposition 1 shows a very interesting algebraic equality

which greatly reduces the computation time of VnX(k). Note that the original definition

of VnX(k) in (2) involves evaluating an integration on R2p. For a moderately large p such

as p = 5, direct numerical computation of such integration is formidable. On the other

hand, however, Proposition 1 claims that VnX(k) equals∑n−k

r,l=1ArlBrl/(n− k)2, which can

be calculated easily with an O(n2) time complexity.

2.1 Dependence measures

Throughout this paper, we shall assume that the the stationary process {Xj} admits the

following representation

Xj = G(· · · , εj−1, εj), (7)

where εj, j ∈ Z, are independent and identically distributed (iid) random variables and

G : R∞ → Rp is a function such that G(· · · , εj−1, εj) converges to an appropriate random

6

Page 7: University of Torontozhou/papers/dcf_Nov7_4.pdf · University of Toronto September 15, 2014 Abstract We extend the concept of distance correlation of Szekely, Rizzo and Bakirov (The

vector. The representation (7) can be viewed as a physical system with (εj) being the inputs

or shocks, (Xj) being the outputs and G being the filter that represents the underlying data

generating mechanism. Model (7) covers a very wide range of linear and nonlinear time

series models encountered in practice; See Wu (2005) for more discussions and examples.

Define the shift process Fj = (· · · , εj−1, εj). Let {(ε′j)}j∈Z be an iid copy of {(εj)}j∈Z.

For j ≥ 0, let the coupled process F∗j = (F−1, ε′0, ε1, · · · , εj). Define physical dependence

measures for the stationary time series {Xj} as follows

Definition 3 (Physical dependence measures). Assume that ‖Xj‖q <∞ with q > 0, where

‖ · ‖q := [E(| · |q)]1/q. Define physical dependence measures

δ(k, q) = ‖G(Fk)−G(F∗k )‖q, k ≥ 0. (8)

Let δ(k, q) = 0 if k < 0. Additionally, write ‖ · ‖ := ‖ · ‖2.

Calibrating the idea of coupling, δ(k, q) measures the dependence of Xk on the input ε0.

The above dependence measures are closely related to the data generating mechanism and

therefore are easy to work with theoretically. Wu (2005) contains detailed calculations of

δ(k, q) for a very general class of linear and nonlinear time series models. All the asymptotic

results of this paper will be expressed in terms of the physical dependence measures.

Another popular class of time series dependence measures is the mixing coefficients.

See for instance Rosenblatt (1956). Generally speaking, the mixing coefficient at lag k

measures the strength of dependence between the σ-fields generated by time series obser-

vations at least k steps away and a fast decay of the mixing coefficients suggests short range

dependence of the series. The mixing coefficients and the physical dependence measures

are defined in a purely mathematical fashion and are targeted at theoretical or asymp-

totic investigations of time series. However, the latter dependence measures generally do

not have corresponding sample versions and therefore are not suitable for the purpose of

exploring and testing dependence structures of time series.

Proposition 2. Assume ‖Xi‖ <∞ and δ(j, 2)→ 0 as j →∞. Then for any r such that

0 < r < 1, we have

VX(k) ≤ C0

∞∑j=0

{δ(j, 2)δ(k + j, 2)}1−r, (9)

7

Page 8: University of Torontozhou/papers/dcf_Nov7_4.pdf · University of Toronto September 15, 2014 Abstract We extend the concept of distance correlation of Szekely, Rizzo and Bakirov (The

where C0 is a finite constant that only depends on p and r. In particular, if δ(j, 2) = O(j−β)

for some β > 1, then VX(k) = O(k−β(1−r)) for any r ∈ (0, 1− 1/β). And if δ(j, 2) = O(χj)

for some 0 < χ < 1, then VX(k) = O(χk1) for some 0 < χ1 < 1.

Proposition 2 is proved in Section 7. Proposition 2 shows that the auto distance covari-

ances can be bounded by the physical dependence measures. In particular, if δ(j, 2) decays

algebraically (geometrically) to 0, then VX(k) also decays algebraically (geometrically) to

0. Generally, however, the physical dependence measures cannot be bounded by the auto

distance covariances. As an illustrative example, let us consider the ARMA(1, 1) model

(1 − B/2)Xj = (1 − 2B)aj, where aj’s are iid standard normal and B denotes the back

shift operator. Elementary calculations show that VX(k) = 0 for all k > 0. However, the

physical dependence measures are positive at all lags.

3 Examples

3.1 Vector linear models

Consider the vector linear model

Xt = A∗ +∞∑j=0

Ajεt−j, (10)

where εj’s are iid length-p random vectors with Eε0 = 0 and E|ε0| < ∞, and Aj’s are

p× p coefficient matrices with∑∞

j=0 |Aj| <∞. It is easy to see that δ(j, q) = O(|Aj|) for

all j ≥ 0 and q ≥ 1. By Proposition 2, if |Aj| decays algebraically (geometrically) to 0

and ‖ε0‖ < ∞, then VX(k) and RX(k) also decay algebraically (geometrically) to 0. In

particular, consider the following vector ARMA(h, q) model:

(I − φ(B))Xt = A∗ + (I − θ(B))εt, (11)

where φ(B) =∑h

j=1 φjBj, θ(B) =

∑qj=1 θjB

j and φj and θj are p× p coefficient matrices.

It is easy to see that (11) can be represented in the form of (10) with Aj decaying expo-

nentially if the roots of |I − φ(B)| are all outside of the unit circle. Therefore for model

(11), VX(k) and RX(k) decay exponentially as well.

8

Page 9: University of Torontozhou/papers/dcf_Nov7_4.pdf · University of Toronto September 15, 2014 Abstract We extend the concept of distance correlation of Szekely, Rizzo and Bakirov (The

We shall first compare the performance of ADCF and ACF for heavy tailed univariate

processes. Figure 1 below shows the sample ADCF and sample ACF for model (11) with

p = 1, φ(B) = 0.5B, θ(B) = 2B and εj’s iid t(2). The length of the series is 300. In the

upper panel of Figure 1, the dotted (solid) vertical lines are the true (sample) ADCF of

the model and the dotted horizontal curve are the critical values of testing RX(k) = 0

at 95% level. The critical values are obtained by subsampling which will be discussed in

detail in Section 5.1. Upper panel of Figure 1 clearly shows the exponential decay of the

sample ADCF. Note that, if E(ε2t ) < ∞, then Xt is a white noise. Even though the t(2)

distribution has infinite variance, the sample ACF behaves as if the series is white noise

and fails to capture the dependence structure of the process.

Let H(p) denote the p×p matrix with Hj,j(p) = 0.6 and Hj,j−1(p) = 0.2, where Hj,k(p)

denotes the (j, k)th entry of H(p). Lower panel of Figure 2 shows the sample ADCF of

model (11) with p = 5, φ(B) = H(5)B, θ(B) = 0 and εj’s iid standard 5-dimensional

Gaussian. Upper panel of Figure 2 shows the sample ADCF of model (11) with p = 5,

φ(B) = 0, θ(B) = H(5)B and εj’s iid standard 5-dimensional Gaussian. The series lengths

are both 300. Figure 2 shows that the ADCF is a very visual-friendly and neat way of

summarizing the dependence structure of multivariate time series: The ADCF for the

vector AR process decays exponentially and the ADCF for the vector MA process cuts off

at lag 1.

3.2 Vector nonlinear moving average processes

We call a multivariate time series (Xt) a nonlinear moving average process of order m

(NMA(m)) if it admits the following representation:

Xj = W (εj−m, εj−m+1, · · · , εj), (12)

where W is a (possibly) nonlinear function. Clearly a NMA(m) process is m-dependent.

Therefore the ADCF of Xt cuts off at lag m. In fact, if the ADCF of a process (Xt) cuts

off at a small lag m, then the NMA(m) model is a good candidate for the process. This

is analogous to the linear process case where if the ACF cuts off at lag m, then a MA(m)

process should be considered.

9

Page 10: University of Torontozhou/papers/dcf_Nov7_4.pdf · University of Toronto September 15, 2014 Abstract We extend the concept of distance correlation of Szekely, Rizzo and Bakirov (The

5 10 15 20 25

0.10

0.15

0.20

0.25

Series X

Lag

ADCF

0 5 10 15 20

0.00.4

0.8

Lag

ACF

Series X

Figure 1: Comparison of the sample ADCF and sample ACF for univariate ARMA(1, 1)

model : (1 − 0.5B)Xj = (1 − 2B)εj with εj’s iid t(2) and n = 300. The dotted (solid)

vertical lines are the true (sample) ADCF of the model and the dotted horizontal curve

are the critical values at 95% level.

5 10 15 20 25

0.20.3

0.40.5

0.60.7

Series X

Lag

ADCF

5 10 15 20 25

0.20.3

0.40.5

0.60.7

Series X

Lag

ADCF

Figure 2: The sample ADCF for vector linear models (1 −H(5)B)Xj = εj (lower panel)

and Xj = (1−H(5)B)εj (upper panel) with εj’s iid 5-dimensional standard Gaussian and

n = 300. The dotted (solid) vertical lines are the true (sample) ADCF of the model and

the dotted horizontal curve are the critical values at 95% level.

10

Page 11: University of Torontozhou/papers/dcf_Nov7_4.pdf · University of Toronto September 15, 2014 Abstract We extend the concept of distance correlation of Szekely, Rizzo and Bakirov (The

As an illustrative example, let us consider the following NMA(m) process

Xj =m∏k=0

εj−k, (13)

where εk’s are iid p vectors with mean 0 and the multiplication in (13) is coordinate-wise.

If ‖εj‖ < ∞, then (13) is a multivariate white noise process and thus the ACF cannot

reveal the underlying dependence structure. Figure 3 compares the sample ADCF and

sample ACF of model (13) with p = 1, m = 2 and εj’s iid standard Gaussian. It is clear

that the sample ADCF cuts off at lag 2 in this case.

5 10 15 20 25

0.10

0.20

0.30

Series X

Lag

ADCF

0 5 10 15 20

0.00.4

0.8

Lag

ACF

Series X

Figure 3: Comparison of the sample ADCF and sample ACF of model (13) with p = 1,

m = 2, εj’s iid standard Gaussian and n = 300. The dotted (solid) vertical lines are the

true (sample) ADCF of the model and the dotted horizontal curve are the critical values

at 95% level.

3.3 Vector nonlinear auto regressive processes

Many nonlinear time series models used in practice, such as bilinear models (Granger and

Andersen, 1978), threshold models (Tong 1990) and (G)ARCH models (Engle (1982) and

Bollerslev (1986)), have the following Markovian representation:

Xj = M(Xj−1, εj). (14)

11

Page 12: University of Torontozhou/papers/dcf_Nov7_4.pdf · University of Toronto September 15, 2014 Abstract We extend the concept of distance correlation of Szekely, Rizzo and Bakirov (The

If Xj = M ′(Xj−1, · · · , Xj−m, εj), then simply let Yj = (XTj , · · · , XT

j−m)T and (Yj) admits

representation (14). We shall call (14) a vector nonlinear auto regressive (NAR) model.

As shown in Wu (2005), if for some x0, ‖M(x0, εj)‖ <∞, and

L < 1, where L = supx 6=y

‖M(x, ε0)−M(y, ε0)‖|x− y|

. (15)

then (14) admits a unique stationary solution, and iterations of (14) lead to Xi = G(Fi).Furthermore, we have δ(j, 2) = O(Lj). Therefore Proposition 2 implies that

VX(k) = O(Lk1) for some L1 ∈ [0, 1). (16)

In other words, the ADCF decays exponentially to 0 for model (14). As an illustrative

example, let us consider the following univariate ARCH(2) model:

Xj = σjεj, where σ2j = 0.5 + 0.8X2

j−1 + 0.1X2j−2 (17)

and εj’s are iid standard Gaussian. It is well known that many financial returns show no

Pearson correlation but the correlation in square returns is strong. The later fact implies

that financial returns are nonlinearly correlated. Figure 4 compares sample ACF and

sample ADCF of model (17) with n = 300. It can be seen that the ACF exhibits no signal

for (Xt). However, the ACF shows moderate dependence in (X2t ). On the other hand, the

ADCF shows nice exponential decays for both (Xt) and (X2t ), which correctly reflects the

underlying nonlinear dependence structure.

4 Asymptotic results

Theorem 1. Suppose that ‖Xj‖1+r0 < ∞ for some r0 > 0 and∑∞

k=0[δ(k, 1 + r0)] < ∞.

Then for all k ≥ 0

VnX(k)→ VX(k) in probability.

Corollary 1. Under the conditions of Theorem 1, we have that RnX(k) is a weakly consis-

tent estimator of RX(k) for each k ≥ 1.

Theorem 1 and Corollary 1 establish the weak consistency of VnX(k) and RnX(k). The

consistency only requires that Xj has (1 + r0)th moment for some r0 > 0. The condition

12

Page 13: University of Torontozhou/papers/dcf_Nov7_4.pdf · University of Toronto September 15, 2014 Abstract We extend the concept of distance correlation of Szekely, Rizzo and Bakirov (The

5 10 15 20 25

0.10.2

0.30.4

0.5

Series X

Lag

ADCF

0 5 10 15 20

−0.2

0.00.2

0.40.6

0.81.0

Lag

ACF

Series X

5 10 15 20 25

0.10.2

0.30.4

0.50.6

Series X^2

Lag

ADCF

0 5 10 15 20

0.00.2

0.40.6

0.81.0

Lag

ACF

Series X^2

Figure 4: Comparison of the sample ADCF and sample ACF of processes (Xt) and (X2t )

in model (17) with εj’s iid standard Gaussian and n = 300. The dotted (solid) vertical

lines are the true (sample) ADCF of the model and the dotted horizontal curve are the

critical values at 95% level.

∑∞k=0[δ(k, 1 + r0)] < ∞ is a very classic short range dependence condition and it means

that the cumulative effect of ε0 in predicting future values of the time series is finite. For

the vector linear process (10), the latter condition is equivalent to∑∞

j=0 |Aj| <∞.

Theorem 2. Let ξ(t, s) denote a complex valued zero-mean Gaussian process with covari-

ance function and relation function

Γ∗(g, g0) =∑j∈Z

E[ψj,k(t, s)ψT0,k(t0, s0)] and R∗(g, g0) =

∑j∈Z

E[ψj,k(t, s)ψT0,k(t0, s0)]

respectively, where ψj,k(t, s) = [exp(i〈t,Xj〉) − fX(t)] × [exp(i〈s,Xj+k〉) − fX(s)], g =

(t, s) and g0 = (t0, s0). Assume that Γ(·, ·) is positive definite, that ‖Xj‖4p+2 < ∞, that∑∞j=0 δ(j, 4p + 2) < ∞ and that δ(j, 4) = O((j + 1)−β) for some β > 3. Then under the

assumption that Xj and Xj+k are independent, we have

nVnX(k)⇒ 1

c2p

∫Rp

∫Rp

|ξ(t, s)|2

|t|p+1|s|p+1dtds, (18)

where ⇒ denotes weak convergence.

13

Page 14: University of Torontozhou/papers/dcf_Nov7_4.pdf · University of Toronto September 15, 2014 Abstract We extend the concept of distance correlation of Szekely, Rizzo and Bakirov (The

Theorem 2 unveils the limiting distribution of the sample auto distance covariance

under the null hypothesis that Xj and Xj+k are independent. By Chapter 1 of Kuo (1975)

and the proof of Lemma 2 in Section 7, it is easy to show that

1

c2p

∫Rp

∫Rp

|ξ(t, s)|2

|t|p+1|s|p+1dtds =

∞∑j=1

λjZ2j ,

where Zj’s are iid standard normal, λj’s are nonnegative constants with

0 <∞∑j=1

λj =1

c2p

∫Rp

∫Rp

Γ(g, g)

|t|p+1|s|p+1dtds <∞.

It is straightforward to obtain that nVnX(k) → ∞ if Xj and Xj+k are dependent. We call

Γ(·, ·) and R(·, ·) long-run covariance and long-run relation functions which are due to the

dependence of the series. If (Xj) is an iid sequence, then simple calculations show that

Γ(g, g0) = [fX(t− t0)− fX(t)fX(t0)][fX(s− s0)− fX(s)fX(s0)] and R(g, g0) = [fX(t+ t0)−fX(t)fX(t0)][fX(s+s0)−fX(s)fX(s0)]. In this case the limiting distribution (18) coincides

with that in Theorem 5 of SRB. In the limiting distribution (18), the dependence of the

series is reflected in the long-run covariance and relation functions. It is worth mentioning

that in a short note, Remillard (2009) suggested using auto distance correlation to measure

nonlinear dependence in time series. However, the latter note claimed (without proof) that

the asymptotic distribution of the auto distance correlation is the same as that in SRB

when the series is a white noise. Based on Theorem 2, we shall point out here that

the latter claim of the limiting distribution was in fact incorrect because there are many

possible forms of nonlinear dependence in a white noise process which will result in very

different asymptotic distributions than that in SRB.

Due to the temporal dependence of the series, the permutation test of SRB is invalid

for making inference in the time series case. In Section 5.1 below, a subsampling based

test procedure will be introduced.

5 Finite sample performance

5.1 The subsampling

We see from Theorem 2 that the asymptotic null distribution of testing VX(k) = 0 involves

the long-run covariance of the series {exp(i〈t,Xj〉) exp(i〈t,Xj+k〉)}j∈Z and a possibly high

14

Page 15: University of Torontozhou/papers/dcf_Nov7_4.pdf · University of Toronto September 15, 2014 Abstract We extend the concept of distance correlation of Szekely, Rizzo and Bakirov (The

dimensional integration. Hence in practice it is difficult to directly use the asymptotic null

distribution to perform hypothesis testing, especially when p is large. Here we suggest

using the subsampling (Politis et. al, 1999) to avoid directly estimating the asymptotic

distributions. Specifically, the following steps can be adopted:

(a). Select a block size l and define subsamples Xl,j = (Xj, Xj+1, · · · , Xj+l−1), j =

1, 2, · · · , n− l + 1.

(b). Calculate V lXl,j(k) for j = 1, 2, · · · , n− l + 1.

(c). Let V l(k, 1) ≤ · · · ≤ V l(k, n− l+ 1) be the ordered sample auto distance covariances

obtained from the subsamples.

(d). For α ∈ (0, 1), the 100(1− α)% critical value of testing VX(k) = 0 can be estimated

by (l − k)V l(k, lα)/(n− k), where lα = b(1− α)(n− l + 1)c. Namely reject the null

hypothesis if VnX(k) > (l − k)V l(k, lα)/(n− k).

The above subsampling procedure is intuitively plausible since the subsamples preserve the

dependence structure of the series. Therefore it is expected that the empirical distribution

of the test statistic obtained from the subsamples coincides with the target distribution

asymptotically. The following theorem establishes the asymptotic validity of the above

subsampling procedure under minimal sufficient conditions for the block size l:

Theorem 3. Assume that l → ∞ with l/n → 0 and that Xj and Xj+k are independent.

Then under the assumptions of Theorem 2, we have

P[(n− k)VnX(k) ≤ (l − k)V l(k, lα)]→ 1− α as n→∞. (19)

We now discuss the choice of the block size l for moderate sample sizes. Here we suggest

use the minimum volatility (MV) method advocated in Chapter 9.4 of Politis et. al (1999).

The idea behind the MV method is that, if a block size is in a reasonable range, then the

estimated critical values for the independence test should be stable when considered as

a function of block size. Hence one could first propose a grid of possible block sizes and

then choose the block size which minimizes the volatility of the critical values near this

size. More specifically, let the grid of possible block sizes be {l1 < . . . < lM} and let the

15

Page 16: University of Torontozhou/papers/dcf_Nov7_4.pdf · University of Toronto September 15, 2014 Abstract We extend the concept of distance correlation of Szekely, Rizzo and Bakirov (The

estimated critical values be {Tlj}, j = 1, 2, . . . ,M . For each lj, calculate se(⋃3r=−3{Tlj+r}),

where se denotes standard error, namely

se({Tj}kj=1) =[ 1

k − 1

k∑j=1

|Tj − T |2]1/2

with T =∑k

j=1 Tj/k. Then one chooses the l∗j which minimizes the above standard errors.

In our simulation studies, the MV method performs reasonably well. We shall refer to

Chapter 9.4 of Politis et. al (1999) for a more detailed description and discussion of the

latter minimum volatility method.

5.2 A simulation study

In this section we shall conduct a small simulation to study the finite sample accuracy of the

auto distance covariance test of independence for multivariate time series. In particular, we

are interested in investigating how dimension p and strength of the dependence influence

the accuracy of the test. For this purpose, we shall consider the vector ARMA model (11)

with φ(B) = aIpB2 and θ(B) = 0, 0 ≤ a < 1, where Ip is the p × p identity matrix.

The errors εj are iid p-dimensional standard Gaussian. Then Xj is independent of Xj+1

and the dependence of the series gets stronger as a gets closer to 1. We are interested in

testing H0 : VX(1) = 0. In our studies we choose n = 100,200 and 300. We then perform

the subsampling test with the minimum volatility block size selection method to test H0.

Based on 5000 repetitions, the simulated type I error rates with respect to various choice

of p and a are reported in Table 1 below.

16

Page 17: University of Torontozhou/papers/dcf_Nov7_4.pdf · University of Toronto September 15, 2014 Abstract We extend the concept of distance correlation of Szekely, Rizzo and Bakirov (The

n = 100 n = 200 n = 300

a = 0.2 a = 0.5 a = 0.8 a = 0.2 a = 0.5 a = 0.8 a = 0.2 a = 0.5 a = 0.8

p = 1 11.7% 14.3% 18.1% 11.7% 11.6% 18.2% 9.7% 10.6% 13.3%

p = 2 9.7% 14.4% 23.3% 9.1% 12.9% 20.0% 9.7% 10.5% 14.2%

p = 3 9.4% 15.0% 24.5% 8.5% 12.1% 21.7% 9.5% 9.6% 14.8%

p = 4 10.6% 15.9% 26.7% 9.2% 11.2% 21.9% 9.3% 10.6% 16.5%

p = 5 12.5% 18.3% 27.8% 9.3% 11.9% 23.0% 9.0% 11.6% 18.3%

p = 6 15.2% 21.2% 30.4% 9.2% 13.2% 24.0% 9.7% 11.6% 18.3%

p = 7 17.2% 22.1% 30.9% 9.7% 14.0% 24.5% 9.5% 11.0% 18.4%

p = 8 19.9% 25.6% 33.6% 11.9% 15.1% 26.7% 9.0% 10.9% 20.4%

p = 9 20.4% 25.9% 37.3% 12.3% 16.1% 26.7% 9.8% 11.1% 21.4%

p = 10 22.4% 24.1% 35.6% 12.5% 17.5% 27.9% 10.0% 12.4% 22.4%

Table 1. Simulated type I error rates for model (11) with φ(B) = aIpB2 and θ(B) = 0 and

nominal level 10%. Series lengths n = 100, 200 and 300 with 5000 replicates.

Table 1 shows that the subsampling performs reasonably well in the simulations. We

also find that the type I error rate improves as the sample size enlarges. On the other

hand, for a fixed sample size, the performance of the subsampling drops as the dependence

strengthens. When making inference of the mean and variance functions, it is well known

that the accuracy of the subsampling decreases as the dependence of the series gets stronger

for a moderate sample size. See for instance Chapter 9.5 of Politis et. al (1999). In our

case the same phenomenon is observed. For strongly dependent data, generally we need

a relatively large sample size to guarantee the accuracy of the subsampling test. There-

fore one should be cautious when exploring temporal dependence structures of strongly

dependent time series.

Table 1 shows that the accuracy of the test decreases when the dimension p is large

compared to the sample size n. As we mentioned in Section 2, throughout the paper

we assume p � n and all the asymptotic theory of the paper are established under the

latter assumption. In our simulations, we find that, for series of moderate dependence,

the performance of the subsampling tests are similar and reasonably accurate for different

choices of p as long as p�√n. Additionally, for many real world applications such as the

ones which will be analyzed in Section 6 of this paper, the dimension p is typically much

17

Page 18: University of Torontozhou/papers/dcf_Nov7_4.pdf · University of Toronto September 15, 2014 Abstract We extend the concept of distance correlation of Szekely, Rizzo and Bakirov (The

smaller than the series length n and hence the theory and methodology of the paper apply.

0 200 400 600 800

−0.2

0.0

0.2

0.4

SP500 Return

Month

Retu

rn

0 5 10 15 20 25

0.0

0.2

0.4

0.6

0.8

1.0

Lag

ACF

Series SP500

0 5 10 15 20 25

0.0

0.2

0.4

0.6

0.8

1.0

Lag

ACF

Series SP500^2

5 10 15 20 250.

080.

100.

120.

14

Series SP500

Lag

ADCF

Figure 5: Time series plot of the SP500 monthly excess return data (upper left panel),

the ACF of the series and the squared series (upper right and lower left panels), and the

ADCF of the series (lower right panel).

6 Real data examples

6.1 The SP500 return data

The (G)ARCH models have been very successful in explaining the volatility clustering

phenomenon observed in financial returns. In this section we are interested in testing the

adequacy of the (G)ARCH models. As an illustrative example, we shall analyze the SP500

monthly excess return data starting from 1926 for 792 observations. The SP500 index

is widely used in the derivative markets. And hence investigating its volatility is of great

interest. Figure 5 shows the time series plot of the data as well as the ACF and ADCF plots.

The ACF of the original series shows no signal while the ACF plot of the squared returns

exhibit moderate level of dependence. The latter phenomenon is classic for time series of

financial returns. On the other hand, the ADCF of the original series shows strong signals

18

Page 19: University of Torontozhou/papers/dcf_Nov7_4.pdf · University of Toronto September 15, 2014 Abstract We extend the concept of distance correlation of Szekely, Rizzo and Bakirov (The

of dependence compared with the critical values of the independence test. The data set

was analyzed in Tsay (2005) and he found an AR(3)-GARCH(1,1) model fits the data well.

The upper part of Figure 6 summaries the ACF of the residuals and the squared residuals

of the AR(3)-GARCH(1,1) model. From the ACF angle, the residuals show no signs of

correlation. However, if we plot the ADCF of the residuals and the squared residuals as

shown in the lower panel of Figure 6 (with block size 10), interesting and surprising signals

at lags 1 and 3 float to the surface. The p-values of testing RX(k) = 0 and RX2(k) = 0 at

lags 1 and 3 are all less than 0.01. Ignoring the latter dependencies will lead to incorrect

inference of the GARCH parameters and the forecast limits. On the other hand, however,

the existence of strong dependencies at lags 1 and 3 of the residuals means that we could

possibly utilize the latter dependencies and further improve our forecasts of the volatility.

For instance, one could adopt an NMA(3) model for the squared residuals and forecast the

volatility accordingly. The detailed modeling and analysis of the residuals is beyond the

scope of this paper. In general, we suggest fitting the SP500 monthly excess return data

with a GARCH model with dependent innovations.

0 5 10 15 20 25

0.0

0.2

0.4

0.6

0.8

1.0

Lag

ACF

Series SP500_residuals

0 5 10 15 20 25

0.0

0.2

0.4

0.6

0.8

1.0

Lag

ACF

Series SP500_residuals^2

5 10 15 20

0.06

0.08

0.10

0.12

Series SP500_residuals

Lag

ADCF

5 10 15 20

0.05

0.06

0.07

0.08

0.09

Series SP500_residuals^2

Lag

ADCF

Figure 6: The ACF and ADCF of the residuals and the squared residuals of the SP500

monthly return data.

19

Page 20: University of Torontozhou/papers/dcf_Nov7_4.pdf · University of Toronto September 15, 2014 Abstract We extend the concept of distance correlation of Szekely, Rizzo and Bakirov (The

6.2 The Canadian lynx data

The time series is composed of the annual record of the number of the Canadian lynx

“trapped” in the Mackenzie River district of the North-West Canada for the period 1821-

1934. Upper panel of Figure 7 plots the data (at log10 scale). The series has become a

classic series for threshold auto regressive (TAR) time series models (Tong ,1990) since the

latter model successfully mimics the asymmetric cycle of the series with a nice predator

(lynx) and prey (hare) interaction interpretation in ecology. Chapter 4.1.4 of Fan and Yao

(2003) did a detailed analysis of the series and found that the following TAR model fits

the data well:

Xj = 0.546 + 1.032Xj−1 − 0.173Xj−2 + 0.171Xj−3 − 0.431Xj−4 + 0.332Xj−5

−0.284Xj−6 + 0.21Xj−7 + ε(1)j , if Xj−2 ≤ 3.116

Xj = 2.632 + 1.492Xj−1 − 1.324Xj−2 + ε(2)j , if Xj−2 > 3.116. (20)

Fan and Yao (2003) mentioned that the residuals of the above TAR model pass most

residual-based tests of autocorrelation comfortably. We shall use the ADCF of the residuals

to test if any nonlinear dependence structure is missed from model (20). Lower panel of

Figure 7 shows the ADCF plot of the residuals. The block size is chosen as 10 in our

analysis. It is clear that the residuals still show strong dependence. In particular, the

TAR model (20) did not fully explain the immediate dependence of the adjacent lynx

counts and the cyclic dependence. Therefore, if the purpose of the study is to forecast

future captures of lynx, then there is plenty of space for model improvement over (20).

7 Proofs

To prove consistency and asymptotic null distribution of the auto distance covariances in

Theorems 1 and 2, we first deal with the two singular points 0 and ∞ in the definition (1)

of the ADCF. Second half of the proof of Theorem 1 and Lemmas 2 and 3 below achieve

the latter task. Then we prove the ergodicity and weak convergence of the empirical char-

acteristic functions using martingale approximation and empirical process techniques. To

prove consistency of the subsampling in Theorem 3, The main techniques used are approx-

imating the empirical characteristic functions by m-dependent processes and controlling

20

Page 21: University of Torontozhou/papers/dcf_Nov7_4.pdf · University of Toronto September 15, 2014 Abstract We extend the concept of distance correlation of Szekely, Rizzo and Bakirov (The

0 20 40 60 80 100

2.0

2.5

3.0

3.5

The Canadian lynx data

Year

Log_

10(L

ynx C

ount

)

5 10 15 20

0.15

0.25

0.35

Residuals of the Canadian lynx data

Lag

ADCF

Figure 7: The time series plot of the Canadian lynx data (upper panel) and the ADCF of

the residuals of model (20) (Lower panel).

the variation of the empirical process of the corresponding m-dependent processes. See

Lemmas 6 and 7.

In this Section the symbol C denotes a finite constant which may vary from place to

place. Let

α(t, j) := exp{i〈t,Xj〉} and β(t, j) := exp{i〈t,Xj+k〉}, j = 1, 2, · · · , n− k. (21)

Write dw = 1c2p|t|p+1|s|p+1dtds. Further write

ζn(t, s) = fnk (t, s)− fn(t)fn,k(s) and ζ(t, s) = fk(t, s)− f(t)f(s), (22)

where fk(t, s) = fXi,Xi+k(t, s). For j ∈ Z define the projection operator

Pj· = E(·|Fj)− E(·|Fj−1). (23)

Lemma 1. For all j, r ∈ Z, η ∈ (0, 1] and q ≥ 1, we have

‖Prα(t, j)‖q ≤ 22−η|t|η[δ(j − r, ηq)]η,

‖Prα(t, j)β(s, j)‖q ≤ 22−η{|t|η[δ(j − r, ηq)]η + |s|η[δ(j + k − r, ηq)]η}. (24)

21

Page 22: University of Torontozhou/papers/dcf_Nov7_4.pdf · University of Toronto September 15, 2014 Abstract We extend the concept of distance correlation of Szekely, Rizzo and Bakirov (The

Proof. We will only prove the first inequality in (24) since the other inequality can be

proved using similar arguments. Let α∗(t, j) = exp{i〈t,X∗j,r〉}, where X∗j,r = G(F∗j,r) with

F∗j,r = (Fr−1, ε′r, εr+1, · · · , εj). By Theorem 1 in Wu (2005), we have

‖Prα(t, j)‖q ≤ ‖α(t, j)− α∗(t, j)‖q ≤ ‖ cos〈t,Xj〉 − cos〈t,X∗j,r〉‖q + ‖ sin〈t,Xj〉 − sin〈t,X∗j,r〉‖q.

Note that

‖ cos〈t,Xj〉 − cos〈t,X∗j,r〉‖q = 2‖ sin〈t,Xj +X∗j,r〉

2sin〈t,Xj −X∗j,r〉

2‖q

≤ 2‖ sin〈t,Xj −X∗j,r〉

2‖q ≤ 2‖min{1,

∣∣∣〈t,Xj −X∗j,r〉2

∣∣∣}‖q≤ 2‖

∣∣∣〈t,Xj −X∗j,r〉2

∣∣∣η‖q ≤ 2‖∣∣∣ |t||Xj −X∗j,r|

2

∣∣∣η‖q= 21−η|t|η[δ(j − r, ηq)]η.

The same inequality holds for ‖ sin〈t,Xj〉 − sin〈t,X∗j,r〉‖q. Therefore the lemma follows. ♦

Proof of Theorem 1. For γ > 0, define

D(γ) = {(t, s) ∈ Rp × Rp : γ ≤ |t| ≤ 1/γ, γ ≤ |s| ≤ 1/γ}. (25)

Let Vn,γ(k) =∫D(γ)|ζn(t, s)|2 dw, Vγ(k) =

∫D(γ)|ζ(t, s)|2 dw. For a fixed (t, s) ∈ D(γ),

||ζn(t, s)|2 − |ζ(t, s)|2| = (|ζn(t, s)|+ |ζ(t, s)|)(||ζn(t, s)| − |ζ(t, s)||)≤ 4|ζn(t, s)− ζ(t, s)|≤ 4

{|fnk (t, s)− fk(t, s)|+ |[fn(t)− f(t)]fn,k(s)|+ |[fn,k(s)− f(s)]f(t)|

}:= 4{I + II + III}.

Note that we used the fact that |ζn(t, s)| ≤ 2 and |ζ(t, s)| ≤ 2 for all (t, s) and n. Write

Ψn,l(t, s) =n−k∑j=1

Pj+k−l[α(t, j)β(s, j)].

Then the summands of Ψn,l(t, s) form a martingale difference series. Without loss of

generality, assume r0 < 1. Write q0 = 1 + r0. By Lemma 1 and Burkholder’s inequality,

we have

(‖Ψn,l(t, s)‖q0/Bq0)q0 ≤ E

( n−k∑j=1

|Pj+k−l[α(t, j)β(s, j)]|2)q0/2

≤n−k∑j=1

E|Pj+k−l[α(t, j)β(s, j)]|q0

22

Page 23: University of Torontozhou/papers/dcf_Nov7_4.pdf · University of Toronto September 15, 2014 Abstract We extend the concept of distance correlation of Szekely, Rizzo and Bakirov (The

≤ C(γ)(n− k){δ(l − k, q0) + δ(l, q0)}q0 ,

where Bq = 18q3/2(q − 1)−1/2, C(γ) = (2/γ)q0 . Note that I = |∑∞

l=0 Ψn,l(t, s)|/(n − k).

Therefore

(n− k)‖I‖q0 ≤∞∑l=0

‖Ψn,l(t, s)‖q0 ≤ Bq0 [C(γ)(n− k)]1/q0∞∑l=0

{δ(l − k, q0) + δ(l, q0)} ≤ C(n− k)1/q0 .

Similarly, it can be shown that ‖II‖q0 + ‖III‖q0 ≤ C(n− k)1/q0−1. Hence

‖Vn,γ(k)− Vγ(k)‖q0 ≤∫D(γ)

‖|ζn(t, s)|2 − |ζ(t, s)|2‖q0 dw

≤ 4

∫D(γ)

‖I‖q0 + ‖II‖q0 + ‖III‖q0 dw ≤ C(n− k)1/q0−1. (26)

In particular, |Vn,γ(k)−Vγ(k)| → 0 in probability for each fixed γ. Clearly Vγ(k) converges

to VX(k) as γ tends to 0. Therefore, to prove Theorem 1, it suffices to show that in

probability

lim supγ→0

lim supn→∞

|Vn,γ(k)− VnX(k)| = 0. (27)

Note that, for each γ > 0,

|Vn,γ(k)− VnX(k)| ≤∫|t|<γ|ζn(t, s)|2 dw +

∫|t|>1/γ

|ζn(t, s)|2 dw

+

∫|s|<γ|ζn(t, s)|2 dw +

∫|s|>1/γ

|ζn(t, s)|2 dw. (28)

Now for z = (z1, · · · , zp) ∈ Rp, define H(y) =∫|z|<y{1− cos z1}/|z|1+p dz.

Then it is clear that H(y) ≤ limy→∞H(y) ≤ cp. On the other hand,

H(y) =

∫|z|<y

2 sin2[z1/2]

|z|1+pdz ≤

∫|z|<y

z212|z|1+p

dz = C(p)y,

where C(p) is a constant that only depends on p. A careful check of the proof of Theorem

2 of Szekely et. al (2007) shows that∫|t|<γ|ζn(t, s)|2 dw ≤ 4

2

n− k

n−k∑j=1

(|Xj+k|+ E|X1|)2

n− k

n−k∑j=1

EX [|Xj −X|H(|Xj −X|γ)],

23

Page 24: University of Torontozhou/papers/dcf_Nov7_4.pdf · University of Toronto September 15, 2014 Abstract We extend the concept of distance correlation of Szekely, Rizzo and Bakirov (The

where X is identically distributed and independent of Xj and EX is taken with respect to

X. Now let q1 = 1 + r0/2. Then

|Xj −X|H(|Xj −X|γ) ≤ |Xj −X|min{cp, C(p)|Xj −X|γ}≤ cp|Xj −X|[C(p)|Xj −X|γ/cp]r0/2 = C∗(p)|Xj −X|q1γr0/2

≤ C∗(p)2q1−1{|Xj|q1 + |X|q1}γr0/2,

where C∗(p) = c1−r0/2p C(p)r0/2. Therefore,∫

|t|<γ|ζn(t, s)|2 dw ≤ Cγr0/2

2

n− k

n−k∑j=1

(|Xj+k|+ E|X1|)2

n− k

n−k∑j=1

(|Xj|q1 + E|X1|q1). (29)

Similar to the proof of (26), it can be shown that, in probability

1

n− k

n−k∑j=1

|Xj+k| → E|X1|,1

n− k

n−k∑j=1

|Xj|q1 → E|X1|q1 .

Hence (29) implies that

lim supn→∞

∫|t|<γ|ζn(t, s)|2 dw ≤ Cγr0/2E|X1|E|X1|q1 .

Therefore lim supγ→0 lim supn→∞∫|t|<γ |ζn(t, s)|2 dw = 0 in probability. Again, a careful

check of the proof of Theorem 2 of Szekely et. al (2007) shows that∫|t|>1/γ

|ζn(t, s)|2 dw ≤ 16γ2

n− k

n−k∑j=1

(|Xj+k|+ E|X1|).

Hence, in probability lim supγ→0 lim supn→∞∫|t|>1/γ

|ζn(t, s)|2 dw = 0. The other two terms

in (28) can be dealt with similarly. Therefore Theorem 1 follows. ♦

Lemma 2. Assume that ‖Xj‖4 < ∞ and δ(j, 4) = O((j + 1)−β) for some β > 3. Let

ξn(t, s) =√n− kζn(t, s). If Xj is independent of Xj+k, then there exists a finite constant

C which does not depend on n, such that

E∫|t|<γ|ξn(t, s)|2 dw + E

∫|s|<γ|ξn(t, s)|2 dw < Cγ. (30)

24

Page 25: University of Torontozhou/papers/dcf_Nov7_4.pdf · University of Toronto September 15, 2014 Abstract We extend the concept of distance correlation of Szekely, Rizzo and Bakirov (The

Proof. We will only deal with the first summand in (30) since the second summand can

be dealt with similarly. Let

uj = α(t, j)− Eα(t, j) and vj = β(s, j)− Eβ(s, j), j = 1, 2, · · · , n− k. (31)

Then ξn(t, s) =√n− k{

∑j ujvj/(n− k) +

∑j uj/(n− k)

∑j vj/(n− k)}. Therefore

E∫|t|<γ|ξn(t, s)|2 dw ≤ 2

n− kE∫|t|<γ|∑j

ujvj|2 dw +2

(n− k)3E∫|t|<γ|∑j

uj∑j

vj|2 dw

:= I0 + II0.

Write

I0 =2

n− kE∫|t|<γ,|s|≤1

|∑j

ujvj|2 dw +2

n− kE∫|t|<γ,|s|>1

|∑j

ujvj|2 dw := I ′0 + I ′′0 .

Denote by x the complex conjugate of x and <(x) the real part of x. Note that

I ′0 =2

n− k

∫|t|<γ,|s|≤1

E(∑j

ujvj∑r

urvr) dw =2

n− k

∫|t|<γ,|s|≤1

E<{∑j,r

ujvjurvr} dw

≤ 2

n− k∑j,r

∫|t|<γ,|s|≤1

|Eujvjurvr| dw. (32)

Since Eujvj = EujEvj = 0 for all j and Pj’s are orthogonal, we have

|Eujvjurvr| = |E∑h∈Z

Phujvj∑l∈Z

Plurvr| = |E∑h∈Z

PhujvjPhurvr|

≤∑h∈Z

E|PhujvjPhurvr| ≤∑h∈Z

‖Phujvj‖‖Phurvr‖. (33)

Similar to the definition of X∗j,r in Lemma 1, define u∗j,r and v∗j,r by replacing Fj with F∗j,r.Then Lemma 1 and Theorem 1 in Wu (2005) imply that

‖Phujvj‖ ≤ ‖ujvj − u∗j,hv∗j,h‖ ≤ ‖(uj − u∗j,h)vj‖+ ‖(vj − v∗j,h)u∗j,h‖≤ ‖uj − u∗j,h‖4‖vj‖4 + ‖vj − v∗j,h‖4‖u∗j,h‖4≤ 2|t|δ(j − h, 4)‖vj‖4 + 2|s|δ(j + k − h, 4)‖u∗j,h‖4. (34)

Since vj =∑

h∈ZPhvj, by Burkholder’s inequality and Lemma 1, we have

(‖vj‖4/B4)2 ≤ ‖

∑h∈Z

|Phvj|2‖ ≤∑h∈Z

‖Phvj‖24 ≤∑h∈Z

[2|s|δ(j + k − h, 4)]2 = C|s|2.

25

Page 26: University of Torontozhou/papers/dcf_Nov7_4.pdf · University of Toronto September 15, 2014 Abstract We extend the concept of distance correlation of Szekely, Rizzo and Bakirov (The

Recall that Bq is defined in the proof of Theorem 1. Therefore ‖vj‖4 ≤ C|s|. Similarly,

‖u∗j,h‖4 ≤ C|t|. Plugging the latter inequalities into (34), we have

‖Phujvj‖ ≤ C|t||s|(δ(j − h, 4) + δ(j + k − h, 4)). (35)

Similarly, ‖Phurvr‖ ≤ C|t||s|(δ(r − h, 4) + δ(r + k − h, 4)). Therefore (33) implies that

|Eujvjurvr| ≤ C|t|2|s|2∑h∈Z

(δ(j − h, 4) + δ(j + k − h, 4))(δ(r − h, 4) + δ(r + k − h, 4))

≤ C|t|2|s|2(|j − r|+ 1)−β.

Therefore (32) implies that

I ′0 ≤C

n− k∑j,r

∫|t|<γ,|s|≤1

|t|2|s|2(|j − r|+ 1)−β dw = γC

n− k∑j,r

(|j − r|+ 1)−β ≤ Cγ.(36)

Now assume |s| > 1. Note that ‖vj‖4 ≤ 2. By Lemma 1 and (34), we have

‖Phujvj‖ ≤ ‖uj − u∗j,h‖4‖vj‖4 + ‖vj − v∗j,h‖4‖u∗j,h‖4≤ 4|t|δ(j − h, 4) + 25/3|s|1/3δ1/3(j + k − h, 4/3)C|t|≤ C|t||s|1/3[δ(j − h, 4) + δ1/3(j + k − h, 4)].

Similarly, ‖Phurvr‖ ≤ C|t||s|1/3[δ(r − h, 4) + δ1/3(r + k − h, 4)]. Hence

I ′′0 ≤ C

n− k∑j,r

∫|t|<γ,|s|>1

|t|2|s|2/3(|j − r|+ 1)−β/3 dw

= γC

n− k∑j,r

(|j − r|+ 1)−β/3 ≤ Cγ. (37)

By (36) and (37), Lemma 2 follows. ♦

Lemma 3. Assume that ‖Xj‖4 < ∞ and δ(j, 4) = O((j + 1)−β) for some β > 3. If Xj

is independent of Xj+k, then there exists a finite constant C which does not depend on n,

such that

E∫|t|>1/γ

|ξn(t, s)|2 dw + E∫|s|>1/γ

|ξn(t, s)|2 dw < Cγ1/3. (38)

Lemma 3 can be proved using similar arguments as those in the proof of Lemma 2.

Details are omitted.

26

Page 27: University of Torontozhou/papers/dcf_Nov7_4.pdf · University of Toronto September 15, 2014 Abstract We extend the concept of distance correlation of Szekely, Rizzo and Bakirov (The

Lemma 4. Under the assumptions of Theorem 2 and for any fixed γ > 0, we have that

ξn(t, s) converges weakly to ξ(t, s) on D(γ).

Proof. We will first prove finite dimensional convergence. Namely for g1, g2, · · · , gl ∈ D(γ),

l ≥ 1,

(ξn(g1), ξn(g2), · · · , ξn(gl))T ⇒ (ξ(g1), ξ(g2), · · · , ξ(gl))T , (39)

where ⇒ denotes weak convergence. We will only prove the case when l = 1, since other

cases follow similarly by the Cramer-Wold device. Let g = (t, s) ∈ D(γ). Recall the

definition of uj and vj in (31). By (35), we have∑h∈Z

‖Phujvj‖ ≤ C|t||s|∑h∈Z

(δ(j − h, 4) + δ(j + k − h, 4)) <∞.

Therefore by Theorem 3 in Wu (2005), we have(<(∑j

ujvj/√n− k),=(

∑j

ujvj/√n− k)

)T⇒(<(ξ(t, s)),=(ξ(t, s))

)T,

where =(x) denotes the imaginary part of x. Now Similar arguments imply that

|∑j

uj| = OP(√n− k) and |

∑j

vj| = OP(√n− k).

Therefore |∑

j uj∑

j vj/(n− k)3/2| = OP(1/√n− k) = oP(1). Hence ξn(g)⇒ ξ(g).

We will now prove the tightness of ξn(t, s) on D(γ). For a fixed g0 ∈ D(γ), similar

arguments as those of (35) imply that ‖ξn(g0)‖ ≤ C. Hence for each positive η, there

exists an a, such that

P(|ξn(g0)| ≥ a) ≤ η, n ≥ 1. (40)

For (t, s) ∈ Rp × Rp, let t = (t1, · · · , tp)T and s = (s1, · · · , sp)T . Define

D∗(γ) =

p⋃j=1

{(t, s) : γ/√p ≤ |tj| ≤ 1/γ or γ/

√p ≤ |sj| ≤ 1/γ}.

Obviously D∗(γ) covers D(γ). For each ε > 0, partition each interval γ/√p ≤ |tj| ≤ 1/γ

and γ/√p ≤ |sj| ≤ 1/γ into v blocks, namely {tj : tj,k ≤ tj ≤ tj,k+1}, {sj : sj,k ≤ sj ≤

27

Page 28: University of Torontozhou/papers/dcf_Nov7_4.pdf · University of Toronto September 15, 2014 Abstract We extend the concept of distance correlation of Szekely, Rizzo and Bakirov (The

sj,k+1}, k = 1, 2, · · · , v, such that ε ≤ mink(|tj,k−tj,k+1|) ≤ 2ε, ε ≤ mink(|sj,k−sj,k+1|) ≤ 2ε

for all j. Note that v ≤ C/ε. Those grid points on each axis further partition D∗(γ) into

v2p := v(p) cubes. Denote by c1, · · · , cv(p) those cubes. Find points o1, o2, · · · , ov(p), such

that oj ∈ cj, j = 1, 2, · · · , v(p). Define the quantities

w(ε) = sup|g′−g′′|≤ε,g′,g′′∈D∗(γ)

|ξn(g′)− ξn(g′′)|, m∗(ε) = max1≤j≤v(p)

supg′,g′′∈cj

|ξn(g′)− ξn(g′′)|.

Note that if |g′ − g′′| ≤ ε, then g′ and g′′ must lie in the same cube or in adjacent ones.

Therefore it is easy to see that

w(ε) ≤ 2m∗(ε) ≤ 4 max1≤j≤v(p)

sup|g−oj |≤2

√2pε

|ξn(g)− ξn(oj)| := 4m(ε).

Let I = {α1, α2, · · · , αq} ⊂ {1, 2 · · · , 2p} be a nonempty set and 1 ≤ α1 < · · · < αq.

For g = (g1, · · · , g2p)T ∈ R2p, let gI = (g111∈I , · · · , g2p12p∈I) and Xj,I = Xj,α1×· · ·×Xj,αq .

For θ, y ∈ R2p, define∫ θI

0

∂qξn(y + gI)

∂gIdgI =

∫ θα1

0

· · ·∫ θαq

0

∂qξn(y + gI)

∂gα1 · · · ∂gαqdgα1 · · · dgαq .

For the nonempty set I defined above, let I1 = I ∩ {1, 2, · · · , p} and I2 = I ∩ {p + 1, p +

2, · · · , 2p} − p. Let χj(I) = Xj,I1α(t, j) − E[Xj,I1α(t, j)] and λj(I) = Xj+k,I2β(s, j) −E[Xj+k,I2β(s, j)]. If I1 is empty, then let Xj,I1 = 1. The same rule applies to Xj+k,I2 . For

g = (tT , sT )T ∈ R2p, we have

∂qξn(g)

∂gα1 · · · ∂gαq= i|I|

{ n−k∑j=1

χj(I)λj(I)√n− k

−∑n−k

j=1 χj(I)∑n−k

j=1 λj(I)√(n− k)3

}, (41)

where |I| denotes the number of elements in I. Let r(q) = (2p+ 1)/(q + 1). Then similar

to the proof of (35), we have∥∥∥ ∂qξn(g)

∂gα1 · · · ∂gαq

∥∥∥r(q)≤ C uniformly over D∗(γ). (42)

Since ‖max1≤j≤v(p) |∂qξn(oj+gI)

∂gI|‖r(q)r(q) ≤

∑v(p)j=1 ‖

∂qξn(oj+gI)

∂gI‖r(q)r(q) and v(p) ≤ C/ε2p, we have by

(42) ∥∥∥ max1≤j≤v(p)

∣∣∣∂qξn(oj + gI)

∂gI

∣∣∣∥∥∥r(q)≤ Cε−2p/r(q) uniformly over D∗(γ). (43)

28

Page 29: University of Torontozhou/papers/dcf_Nov7_4.pdf · University of Toronto September 15, 2014 Abstract We extend the concept of distance correlation of Szekely, Rizzo and Bakirov (The

Consequently,∫ εp

−εp· · ·∫ εp

−εp

∥∥∥ max1≤j≤v(p)

∣∣∣∂qξn(oj + gI)

∂gI

∣∣∣∥∥∥r(q)

dgI ≤ Cεqε−2p/r(q) ≤ Cεπ,

where π = 1/(2p+ 1) and εp = 2√

2pε. Note that

ξn(g)− ξn(oj) =∑

I⊆{1,2,··· ,2p}

∫ (g−oj)I

0

∂|I|ξn(oj + gI)

∂gIdgI .

Therefore

‖m(ε)‖ ≤∑

I⊆{1,2,··· ,2p}

∫ εp

−εp· · ·∫ εp

−εp

∥∥∥ max1≤j≤v(p)

∣∣∣∂qξn(oj + gI)

∂gI

∣∣∣∥∥∥ dgI≤

∑I⊆{1,2,··· ,2p}

∫ εp

−εp· · ·∫ εp

−εp

∥∥∥ max1≤j≤v(p)

∣∣∣∂qξn(oj + gI)

∂gI

∣∣∣∥∥∥r(q)

dgI

≤ C∑

I⊆{1,2,··· ,2p}

επ ≤ Cεπ.

By Markov’s inequality, we have

P(m(ε) ≥ επ/2) ≤ ‖m(ε)‖2

επ≤ Cεπ → 0. (44)

Hence (40), (44) and Theorem 7.3 in Billingsley (1999) imply that ξn(t, s) is tight on D(γ).

Proof of Theorem 2. By Lemma 4 and the continuous mapping theorem, we have for each

fixed γ > 0,

nVn,γ(k)⇒ 1

c2p

∫D(γ)

|ξ(t, s)|2

|t|p+1|s|p+1dtds.

Note that

n|Vn,γ(k)− VnX(k)| ≤∫|t|<γ|ξn(t, s)|2 dw +

∫|t|>1/γ

|ξn(t, s)|2 dw

+

∫|s|<γ|ξn(t, s)|2 dw +

∫|s|>1/γ

|ξn(t, s)|2 dw.

Therefore Lemmas 2 and 3 imply that Theorem 2 holds. ♦

29

Page 30: University of Torontozhou/papers/dcf_Nov7_4.pdf · University of Toronto September 15, 2014 Abstract We extend the concept of distance correlation of Szekely, Rizzo and Bakirov (The

Proof of Proposition 2. Note that VX(k) =∫|Cov(α(t, 0), β(s, 0))|2 dw. Furthermore,

|Cov(α(t, 0), β(s, 0))| = |E{∑j∈Z

Pjα(t, 0)∑j′∈Z

Pj′β(s, 0)}| = |E{∑j∈Z

Pjα(t, 0)Pjβ(s, 0)}|

≤∑j∈Z

‖Pjα(t, 0)‖‖Pjβ(s, 0)‖ =0∑

j=−∞

‖Pjα(t, 0)‖‖Pjβ(s, 0)‖. (45)

Now by the similar arguments as those in the proof of Lemmas 2 and 3, Proposition 2

follows. Details are omitted. ♦

Proof of Theorem 3. Theorem 3 follows from Proposition 3 below and very similar argu-

ments as those in the proof of Theorem 1 of Beran (1984). See also Theorem 3.2.1 of

Politis et. al (1999). Details are omitted. ♦

Proposition 3. Let U be a random variable following the distribution on the right hand

side of (18). Then under the assumptions of Theorem 3, we have for any x > 0

Ln(x)→ P(U ≤ x), as n→∞,

where Ln(x) = 1n−l+1

∑n−l+1j=1 I{(l − k)V lXl,j(k) ≤ x} and I{·} is the indicator function.

Proof. Note that for any x > 0, x is a continuous point for the distribution of U . Propo-

sition 3 follows by Lemmas 5, 6, 7 and Theorem 2 when letting γ → 0, m → ∞,

mγ2(p+1) →∞ and m/n→ 0. ♦

Lemma 5. For u = 1, 2, · · · , n− l + 1 and γ > 0, define

Vl,k(u, γ) =

∫D(γ)

∣∣∣ 1

l − k

u+l−k−1∑j=u

α(t, j)β(s, j)− 1

l − k

u+l−k−1∑r=u

α(t, r)1

l − k

u+l−k−1∑j=u

β(s, j)∣∣∣2 dw.

Let Ln(x, γ) = 1n−l+1

∑n−l+1u=1 I{(l − k)Vl,k(u, γ) ≤ x}. Then under the assumptions of

Theorem 3, we have

Ln(x) ≤ Ln(x, γ) and P[Ln(x)− Ln(x− γ1/6, γ) ≤ −γ1/12] ≤ C1γ1/12,

where C1 is a positive constant which does not depend on n and γ.

30

Page 31: University of Torontozhou/papers/dcf_Nov7_4.pdf · University of Toronto September 15, 2014 Abstract We extend the concept of distance correlation of Szekely, Rizzo and Bakirov (The

Proof. Obviously V lXl,u(k) ≥ Vl,k(u, γ), u = 1, 2, · · · , n− l+1. Therefore Ln(x) ≤ Ln(x, γ).

On the other hand, note that, for random variables X and Y and any x ∈ R and δ > 0,

we have

I{X ≤ x} − I{Y ≤ x− δ} ≥ −I{X − Y ≥ δ}. (46)

Therefore

Ln(x)− Ln(x− γ1/6, γ) ≥ − 1

n− l + 1

n−l+1∑u=1

I{(l − k)[V lXl,u(k)− Vl,k(u, γ)] ≥ γ1/6} := −Un(x, γ).

Hence by Markov’s inequality, it follows that

P[Ln(x)− Ln(x− γ1/6, γ) ≤ −γ1/12] ≤ P{−Un(x, γ) ≤ −γ1/12} ≤ E[Un(x, γ)]

γ1/12. (47)

For each u, by Lemmas 2 and 3 and the Markov’s inequality, we have

E[I{(l − k)[V lXl,u(k)− Vl,k(u, γ)] ≥ γ1/6}] = P{(l − k)[V lXl,u(k)− Vl,k(u, γ)] ≥ γ1/6}≤ E{(l − k)[V lXl,u(k)− Vl,k(u, γ)]}/γ1/6

≤ C1γ1/3/γ1/6 = C1γ

1/6, (48)

where C1 is a positive constant which does not depend on n and γ. Therefore E[Un(x, γ)] ≤C1γ

1/6. Now by (47), the Lemma follows. ♦

For an integer m > 0 and j = 1, 2, · · · , n− k, write

αm(t, j) := E[α(t, j)|εj,m], βm(t, j) := E[β(t, j)|εj,m], ηm(t, s, j) := E[η(t, s, j)|εj,m],(49)

where η(t, s, j) = α(t, j)β(s, j) and εj,m = (εj, εj−1, · · · , εj−m). For u = 1, 2, · · · , n− l+ 1,

let

Vl,k(u, γ,m) =

∫D(γ)

∣∣∣ 1

l − k

u+l−k−1∑j=u

ηm(t, s, j)− 1

l − k

u+l−k−1∑r=u

αm(t, r)1

l − k

u+l−k−1∑j=u

βm(s, j)∣∣∣2 dw.

Write Ln(x, γ,m) = 1n−l+1

∑n−l+1u=1 I{(l − k)Vl,k(u, γ,m) ≤ x}.

Lemma 6. Under the assumptions of Theorem 3, we have

P[Ln(x, γ)− Ln(x−m−1/γ2(p+1), γ,m) ≤ −m−1/2/γp+1] ≤ C2m−1/2/γp+1,

P[Ln(x+m−1/γ2(p+1), γ,m)− Ln(x, γ) ≤ −m−1/2/γp+1] ≤ C2m−1/2/γp+1. (50)

where C2 is a positive constant which does not depend on n and γ.

31

Page 32: University of Torontozhou/papers/dcf_Nov7_4.pdf · University of Toronto September 15, 2014 Abstract We extend the concept of distance correlation of Szekely, Rizzo and Bakirov (The

Proof. For u = 1, 2, · · · , n− l+ 1, write S∗(u, l) = S1(u, l)−S2(u, l)S3(u, l) and S∗m(u, l) =

S1,m(u, l)− S2,m(u, l)S3,m(u, l), where

S1(u, l) =u+l−k−1∑j=u

η(t, s, j)/(l − k), S1,m(u, l) =u+l−k−1∑j=u

ηm(t, s, j)/(l − k),

S2(u, l) =u+l−k−1∑j=u

α(t, j)/(l − k), S2,m(u, l) =u+l−k−1∑j=u

αm(t, j)/(l − k),

S3(u, l) =u+l−k−1∑j=u

β(s, j)/(l − k), S3,m(u, l) =u+l−k−1∑j=u

βm(s, j)/(l − k).

Then for each u, by Cauchy’s inequality,

‖Vl,k(u, γ)− Vl,k(u, γ,m)‖ = ‖∫D(γ)

|S∗(u, l)|2 dw −∫D(γ)

|S∗m(u, l)|2 dw‖

≤∫D(γ)

‖|S∗(u, l)|2 − |S∗m(u, l)|2‖ dw

≤∫D(γ)

‖S∗(u, l)− S∗m(u, l)‖4[‖S∗(u, l)‖4 + ‖S∗m(u, l)‖4] dw.(51)

By Lemma A.1.(i) of Liu and Lin (2009) with q = 4 and Lemma 1 with q = 4 and η = 1,

we have

√l − k‖S1(u, l)− S1,m(u, l)‖4 ≤ C(p)

∞∑j=m

{|t|δ(j, 4) + |s|δ(j + k, 4)} ≤ C(p)m−2/γ (52)

for any (t, s) ∈ D(γ), where C(p) is a constant which only depends on p and can vary

from place to place. Analogous results hold for Sj(u, l)− Sj,m(u, l), j = 2, 3. On the other

hand, by the similar Martingale decomposition method used in the proof of Theorem 1, it

follows that

maxj=1,2,3

√l − k{‖Sj(u, l)‖4 + ‖Sj,m(u, l)‖4} ≤ C(p)/γ. (53)

Therefore plugging (52) and (53) into (51), we have

(l − k)‖Vl,k(u, γ)− Vl,k(u, γ,m)‖ ≤∫D(γ)

C(p)m−2/γ2 dw ≤ C(p)m−2/γ4(p+1).

Now similar arguments as those in the proof of Lemma 5 leads to (50). Details are omitted.

32

Page 33: University of Torontozhou/papers/dcf_Nov7_4.pdf · University of Toronto September 15, 2014 Abstract We extend the concept of distance correlation of Szekely, Rizzo and Bakirov (The

Lemma 7. Assume that m/n→ 0. Then under the assumptions of Theorem 3, we have

Ln(x, γ,m)→ P{(l − k)Vl,k(u, γ,m) ≤ x} in probability.

Proof. It suffices to show that ‖Ln(x, γ,m)−ELn(x, γ,m)‖ → 0 as n→∞. Let ρ(u, x) :=

I{Vl,k(u, γ,m) ≤ x}. Since {ρ(u, x)}n−l+1u=1 is m+ l-dependent and 0 ≤ ρ(u, x) ≤ 1, we have

(n− l + 1)2Var(Ln(x, γ,m)) =∑

|j−r|≤m+l

Cov(ρ(j, x), ρ(r, x)) +∑

|j−r|>m+l

Cov(ρ(j, x), ρ(r, x))

=∑

|j−r|≤m+l

Cov(ρ(j, x), ρ(r, x)) ≤ 2(m+ l)2. (54)

Since m/n→ 0 and l/n→ 0, the lemma follows. ♦

REFERENCES

Ashley, R. A., Patterson, D. M. and Hinich, M. J. (1986). A diagnostic test for nonlinear serial

dependence in time series fitting errors. Journal of Time Series Analysis. 7 165–178.

Bagnato, L., Punzo, A. and Nicolis, O. (2011). The autodependogram: a graphical device

to investigate serial dependences. Journal of Time Series Analysis. To appear. DOI:

10.1111/j.1467-9892.2011.00754.x

Beran, R. (1984). Bootstrap methods in statistics. Jahresberichte des Deutschen Mathematischen

Vereins 86 14-30.

Billingsley, P. (1999). Convergence of Probability Measures. second edition. Wiley.

Bollerslev, T. (1986). Generalized autoregressive conditional heteroskedasticity. Journal of

Econometrics 31 307-327.

Box, G. E. P. and Jenkins, G. M. (1970). Time series analysis: Forecasting and control. San

Francisco: Holden-Day.

Brillinger, D. R. (1977). The identification of a particular nonlinear time series system. Biometrika

64 509–515.

Brockwell, P. (2007). Beyond linear time series. Statistica Sinica 17 3-7.

Chen, R. and Tsay, R. (1993). Functional-coefficient autoregressive models. Journal of the

American Statistical Association. 88 298–308.

Davis, R.A. and Mikosch, T. (1998). The sample ACF of heavy-tailed stationary processes with

applications to ARCH. Annals of Statistics 26 2049–2080.

Engle, R. F. (1982). Autoregressive conditional heteroscedasticity with estimates of the variance

of United Kingdom inflations. Econometrica 50 987-1007.

33

Page 34: University of Torontozhou/papers/dcf_Nov7_4.pdf · University of Toronto September 15, 2014 Abstract We extend the concept of distance correlation of Szekely, Rizzo and Bakirov (The

Fan, J. and Yao, Q. (2003). Nonlinear time series, nonparametric and parametric methods.

Springer, New York.

Franses, P. H. and van Dijk, D. (2000). Nonlinear time series models in empirical fiance. Cam-

bridge University Press.

Granger, C. W. J. and Andersen, A. (1978). Non-linear time series modelling. In Applied time

series analysis, ed. D. F. Findley, Academic Press, New York, 25-38.

Liu, W. and Lin, Z. (2009). Strong approximation for a class of stationary processes. Stochastic

Processes and their Applications 119 249–280.

Kantz, H. and Schreiber, T. (2000). Nonlinear time series analysis. Cambridge University Press.

Kuo, H. H. (1975). Gaussian measures in Banach spaces. Lecture notes in Math. 463 Springer,

Berlin.

Politis, D. N., Romano, J. P. and Wolf, M. (1999). Subsampling. Springer, New York.

Priestley, M. B. (1988). Non-linear and non-stationary time series. Academic Press.

Rachev, S. and Mittnik, S. (2000). Stable Paretian models in finance. Wiley, Chichester.

Remillard, B. (2009). Discussion of: Brownian distance covariance. Annals of Applied Statistics,

3 1295–1298.

Rosenblatt, M. (1956). A central limit theorem and a strong mixing condition. Proceedings of

the National Academy of Sciences, USA 42 43-47.

Szekely, G. J. and Rizzo, M. L. (2009). Brownian distance covariance (with discussions). Annals

of Applied Statistics 3 1236-1308.

Szekely, G. J., Rizzo, M. L. and Bakirov, N. K. (2007). Measuring and testing independence by

correlation of distances. Ann. Statist. 35 2769-2794.

Tjøstheim, D. and Auestad, B. H. (1994a). Nonparametric identification of nonlinear time series:

projections. Journal of the American Statistical Association. 89 1398–1409.

Tjøstheim, D. and Auestad, B. H. (1994b). Nonparametric identification of nonlinear time series:

selecting significant lags. Journal of the American Statistical Association. 89 1410–1419.

Tong, H. (1990). Non-linear time series: a dynamical system approach. Oxford University Press,

Oxford, UK.

Tsay, R. S. (2005). Analysis of financial time series, second edition. Wiley-Interscience.

Wu, W. B. (2005). Nonlinear system theory: another look at dependence. Proceedings of the

National Academy of Sciences USA, 102, 14150–14154.

34