Is the diurnal pattern sufficient to explain intraday variation in volatility? A nonparametric assessment∗ Kim Christensen†

Ulrich Hounyo‡,†

Mark Podolskij§,†

March, 2018 Abstract In this paper, we propose a nonparametric way to test the hypothesis that time-variation in intraday volatility is caused solely by a deterministic and recurrent diurnal pattern. We assume that noisy high-frequency data from a discretely sampled jump-diffusion process are available. The test is then based on asset returns, which are deflated by the seasonal component and therefore homoskedastic under the null. To construct our test statistic, we extend the concept of pre-averaged bipower variation to a general Itˆo semimartingale setting via a truncation device. We prove a central limit theorem for this statistic and construct a positive semidefinite estimator of the asymptotic covariance matrix. The t-statistic (after pre-averaging and jump-truncation) diverges in the presence of stochastic volatility and has a standard normal distribution otherwise. We show that replacing the true diurnal factor with a modelfree jump- and noise-robust estimator does not affect the asymptotic theory. A Monte Carlo simulation also shows this substitution has no discernable impact in finite samples. The test is, however, distorted by small infinite-activity price jumps. To improve inference, we propose a new bootstrap approach, which leads to almost correctly sized tests of the null hypothesis. We apply the developed framework to a large cross-section of equity high-frequency data and find that the diurnal pattern accounts for a rather significant fraction of intraday variation in volatility, but important sources of heteroskedasticity remain present in the data. JEL Classification: C10; C80. Keywords: Bipower variation; bootstrapping; diurnal variation; high-frequency data; microstructure noise; pre-averaging; time-varying volatility. ∗

This paper was previously entitled “Testing for heteroscedasticity in jumpy and noisy high-frequency data: A resampling approach.” We appreciate the thoughtful feedback received from two anonymous referees, the associate editor, and Yacine A¨ıt-Sahalia as part of the revision process, which helped to significantly improve the paper. We also thank for comments on an earlier version of this paper made by conference participants at the 10th Computational and Financial Econometrics (CFE) conference in Seville, Spain; the Vienna-Copenhagen (VieCo) 2017 conference in Vienna, Austria; the 10th Annual SoFiE meeting in New York, USA; the Canadian Econometric Study Group (CESG) 2017 meeting in Toronto, Canada; and at seminars in Aarhus, Aix-Marseille, Albany, California Polytechnic, Connecticut, Cornell, Erasmus (Rotterdam), Glasgow, Manchester, Nottingham and Queen’s University. The authors acknowledge research funds from the Danish Council for Independent Research (DFF – 4182-00050). In addition, Podolskij thanks the Villum Foundation for the grant “Ambit fields: Probabilistic properties and statistical inference.” This work was also supported by CREATES, which is funded by the Danish National Research Foundation (DNRF78). Please address correspondence to: [email protected]. † Aarhus University, Department of Economics and Business Economics, CREATES, Fuglesangs All´e 4, 8210 Aarhus V, Denmark. ‡ University at Albany – State University of New York, Department of Economics, 1400 Washington Avenue, Albany, New York 12222, United States. § Aarhus University, Department of Mathematics, Ny Munkegade 118, 8000 Aarhus C, Denmark.

1

Introduction

There is a widespread agreement in the literature that any dynamic model of volatility should—at a minimum—account for two distinct features in order to explain the formation of diffusive risk in financial markets. On the one hand, a mean-reverting but highly persistent stochastic component is needed at the interday horizon to capture volatility clustering (e.g., Fama, 1965; Mandelbrot, 1963). On the other, a pervasive diurnal effect is required as one of the most critical determinants to describe the recurrent behavior of intraday volatility. In stock markets, for example, there is a tendency for absolute (or squared) price changes during the course of a trading day to form a socalled “U”- or reverse “J”-shape with notably larger fluctuations near the opening and closing of the exchange than around lunch time (see, e.g., Harris, 1986; Wood, McInish, and Ord, 1985, for earlystage documentation of this attribute).1 In addition to these effects, volatility may exhibit large, sudden shifts around the release of important economic news, such as macroeconomic information (e.g., Andersen and Bollerslev, 1998). A recent strand of work, fueled by access to high-frequency data and complimentary theory for model-free measurement of volatility, has taken a more detailed close-up of these components and largely confirmed their presence.2 The diurnal U-shape, in particular, has emerged as a potent—if not predominant—source of within-day variation in volatility. It is therefore common to formulate parametric models of time-varying volatility targeted for high-frequency analysis (be it in continuous- or discrete-time) as a composition of a stochastic and deterministic process (with suitable restrictions imposed to ensure the parameters are separately identified). A standard approach is to assume that the stochastic process is constant within a day but is evolving randomly between them (thus enabling volatility clustering), while the deterministic part is a smooth periodic function that is allowed to change within the day but is otherwise time-invariant (thus capturing the diurnal effect), see, e.g., Andersen and Bollerslev (1997, 1998); Boudt, Croux, and Laurent (2011); Engle and Sokalska (2012) and references therein. Indeed, a major motivation behind the preferred use of realized measures of return variation that are temporally aggregated to the daily frequency is to avoid dealing with the diurnal effect, since it is widely believed to make them intrinsically robust against its presence. However, as stressed by Andersen, Dobrev, and Schaumburg (2012); Dette, Golosnoy, and Kellermann (2016) diurnal effects inject a strong Jensen’s inequality-type bias in some of these estimators; an effect that is reinforced and magnified with a high “volatility-of-stochastic volatility” (e.g., Christensen, Oomen, and Podolskij, 2014). This can, for instance, alter the finite sample properties of jump tests designed to operate at either the intraday or interday horizon (e.g., Andersen, Bollerslev, and Dobrev, 2007; 1

Moreover, diurnal effects are present in other financial variables, such as asset covariances and correlations, bid-ask spreads, trade durations, trading volume, and quote updates. 2 A comprehensive list of papers in this field, including several reviews of the literature, is available at the webpage of the Oxford-Man Institute of Quantitative Finance’s Realized Library: http://realized.oxfordman.ox.ac.uk/research/literature.

1

Barndorff-Nielsen and Shephard, 2006; Lee and Mykland, 2008) and make them significantly leaned toward the alternative and cause spurious jump detection as result.3 As such, further investigation of diurnal effects appears warranted. In this paper, we develop a nonparametric framework to assess if diurnal effects can, in fact, explain all of the intraday variation in volatility, as stipulated by such a setup. A casual inspection of high-frequency data does not offer conclusive evidence about the validity of this conjecture. In concrete applications, inference is obscured by microstructure noise at the tick-by-tick frequency (e.g., Hansen and Lunde, 2006) and the existence of price jumps that are potentially very small and highly active (e.g., A¨ıt-Sahalia and Jacod, 2012a).4 Moreover, even if stochastic volatility is truly present, in practice its components may be so persistent that it is acceptable (and convenient) to regard it as absent on small time scales.5 Consistent with the above, we model the asset log-price as a general arbitrage-free Itˆo semimartingale, which is contaminated by microstructure noise. In this framework, the asymptotic theory is infill, i.e. the process is assumed to be observed on a fixed time interval with mesh tending to zero. There are several existing tests of constant volatility available in the high-frequency volatility area (see, e.g., Dette, Podolskij, and Vetter, 2006; Dette and Podolskij, 2008; Vetter and Dette, 2012). To our knowledge, none allow for the joint disturbance of jumps and microstructure noise, nor do they directly study the extension to diurnal variation advocated here. We formulate a test on the back of log-returns that are only homoskedastic under the null, after they are filtered for diurnal effects. We then study a jump-robust version of the pre-averaged bipower variation, where we extend the bivariate central limit theorem of Podolskij and Vetter (2009a) to the jumpy setting (see Barndorff-Nielsen, Hansen, Lunde, and Shephard, 2008; Jacod, Li, Mykland, Podolskij, and Vetter, 2009; Zhang, Mykland, and A¨ıt-Sahalia, 2005; Zhang, 2006, for further work on noiserobust volatility estimation). The test is constructed via the asymptotic distribution implied by a transformation of such statistics and an application of Cauchy-Schwarz for a particular—but standard—choice of the parameters. As an aside, we add that a slightly different configuration of our t-statistic (based on the comparison of suitably non-truncated and truncated statistics) can serve as a basis for a jump test, which is robust to diurnal effects, but we do not pursue this idea in the present paper. As the diurnal pattern is unknown in practice, we follow the “two-stage” approach of Andersen and Bollerslev (1997). In the first stage, the diurnal factors are pre-estimated. We propose a 3 The effects are deeply intertwined, however, because jumps can also induce substantial biases in and distort estimates of both integrated variance (e.g., Barndorff-Nielsen and Shephard, 2004; Christensen, Oomen, and Podolskij, 2014) and the diurnal pattern (e.g., Andersen, Bollerslev, and Das, 2001; Boudt, Croux, and Laurent, 2011). 4 In view of this, a related topic is whether a diffusive component is needed in the first place to represent risk formation in financial asset prices. While the prevailing evidence is slightly mixed, it appears largely affirmative (see, e.g., A¨ıt-Sahalia and Jacod, 2012b; Kolokolov and Ren`o, 2017; Todorov and Tauchen, 2010). 5 Of course, the locally constant approximation of stochastic volatility is one of the most heavily exploited in the analysis of high-frequency data, see, e.g., Mykland and Zhang (2009).

2

nonparametric estimator, which is both inherently jump- and noise-robust and therefore applicable in practice. It extends previous work of, e.g., Andersen and Bollerslev (1997); Andersen, Dobrev, and Schaumburg (2012); Boudt, Croux, and Laurent (2011); Taylor and Xu (1997); Todorov and Tauchen (2012) to the noisy setting (see also Hecq, Laurent, and Palm, 2012). In the second stage, the estimator is inserted and therefore replaces the true value in the deflation step. The estimate contains a sampling error, however, which afflicts the calculation and may propagate through the system and invalidate the analysis. This problem has largely gone unnoticed (or at least been ignored) in previous work.6 In a double-asymptotic setup with mesh going to zero, as before, and time span now increasing to infinity, we show our feasible statistic has a sampling error of sufficiently small order to not affect the previous infill asymptotic theory. As such, the inference procedure only applies if the researcher has high-frequency data over a large number of days, but is interested in the behavior of volatility on smaller time scales, which is an inherent but unavoidable limitation of our approach. A simulation study reveals that this substitution also has no discernable impact in finite samples. The test is, however, severely distorted by the presence of small infinite-activity price jumps, which it understandably appears to confuse with stochastic volatility. To improve inference, we suggest a new bootstrap approach. It is of independent interest and can be viewed as an overlapping version of the wild blocks of blocks bootstrap by Hounyo, Gon¸calves, and Meddahi (2017). We prove the first-order validity of the bootstrap, while in simulations it helps to restore an almost correctly sized test. The paper is structured as follows. In Section 2, we introduce our theoretical framework and the main assumptions. We extend the asymptotic theory of the pre-averaged bipower variation and construct a jump- and noise-robust test of the hypothesis that all intraday variation in volatility is captured by the diurnal pattern. In Section 3, we propose an estimator of the latent diurnal pattern and show that the sampling error of the feasible statistic is sufficiently small to not affect the previous results. In Section 4, we introduce the bootstrap and show its consistency for testing the null hypothesis in a noisy jump-diffusion setting. In Section 5, we present the Monte Carlo results, while an empirical illustration is conducted in Section 6. We conclude in Section 7. The mathematical proofs and some auxiliary results are relegated to the Appendix.

2

Theoretical setup

We let X denote a latent efficient log-price defined on a filtered probability space (Ω, F, (Ft )t≥0 , P ) and recorded in the window [0, T ], where T is the number of days in the sample and the subinterval [t − 1, t] is the tth day, for t = 1, . . . T . Throughout, T is mainly fixed and the asymptotic theory is infill, so we often impose T = 1 as a normalization, but please note the important digression in 6

A notable exception is Todorov and Tauchen (2012). In the supplemental material to that paper, available at Econometrica’s website, the authors treat the impact of diurnal filtering on their realized Laplace transform estimator of volatility, but microstructure noise is supposed to be absent.

3

Section 3, where a long-span analysis (T → ∞) is required, and there the additional notation and interpretation of [0, T ] is helpful. As consistent with the no-arbitrage restriction (e.g., Delbaen and Schachermayer, 1994), we model X as an Itˆo semimartingale: Z t Z t σs dWs + Jt , t ∈ [0, T ], (1) as ds + Xt = X0 + 0

0

where (at )t≥0 is a predictable, locally bounded drift process, (σt )t≥0 is an adapted, c`adl`ag volatility process, while (Wt )t≥0 is a Brownian motion. Jt is a jump process defined by the equation:     Jt = δ1{|δ|≤1} ? µt − ν t + δ1{|δ|>1} ? µt ,

(2)

where µ is a Poisson random measure on R+ × R and ν is a predictable compensator of µ, such that ν(ds, dx) = ds ⊗ λ(dx) and λ is a σ-finite measure. As explained above, the main idea of the paper is to construct a test, which tells whether a diurnal component is adequate to describe the evolution of within-day volatility.7 To this end, we need to put some structure on the problem, starting with: Assumption (D1):

σt = σsv,t σu,t .

σsv,t and σu,t represent two distinct sources of time-varying volatility in many financial return series. The first term, σsv,t , denotes a stochastic process, which allows for randomness in the evolution of σt over time. The second term, σu,t , is a deterministic seasonal component that represents the diurnal pattern. The multiplicative structure means our test can be formed by deflating the log-return series with σu,t and checking if the outcome is homoskedastic, as we do in Section 2.2. Assumption (D2):

2 σsv,t



2 is stationary with E(σsv,t ) = σ 2 and t≥0

P∞

 2 2 cov σ , σ sv,t sv,t+k < ∞. k=0

2 Assumption (D3): (σu,t )t≥0 is a continuously differentiable 1-periodic function with bounded R1 2 derivative for t → 0 and t → 1 and normalized such that 0 σu,s ds = 1.

Assumption (D2) – (D3) are sufficient to ensure identification of both volatility components from the data.8 Apart from the stationarity of the stochastic volatility, the former also restricts its memory, which implies the process is ergodic. The latter says the diurnal component has to be recurrent, so that we can gradually infer it from gathering a larger sample.9 Taken together, the 7 We do not speak about the dynamics of volatility during market close. Thus, our framework is consistent with random changes in between-day volatility. 8 The normalization in Assumption (D3) is known as a “standardization condition,” which ensures that the decomposition in Assumption (D1) is unique under (locally) constant volatility, see, e.g., Andersen and Bollerslev (1997); Boudt, Croux, and Laurent (2011); Taylor and Xu (1997). 9 It can be extended to accommodate a day-of-the-week effect (e.g., Andersen and Bollerslev, 1998).

4

conditions imply that an average of (an estimate of) volatility sampled at a fixed time of the day P P p 2 2 2 2 σ 2 , as T → ∞, thereby delivering → σu,s T −1 Tt=1 σsv,t−1+s = σu,s s ∈ (0, 1), i.e. T −1 Tt=1 σt−1+s 2 after suitable normalization. σu,s In Section 3, we use this idea to recover σu from noisy high-frequency data, so that we can

compute the test in practice, but—for the moment—we treat it as observed. We further rule out jumps in σsv,t : Assumption (V):

σsv,t is of the form: Z t Z t Z t v˜s dBs , σ ˜s dWs + a ˜s ds + σsv,t = σ0 + 0

0

(3)

0

where (˜ at )t≥0 , (˜ σt )t≥0 and (˜ vt )t≥0 are adapted, c`adl`ag stochastic processes, while (Bt )t≥0 is a standard Brownian motion that is independent of W . Assumption (V) is common in the realized volatility literature (see, e.g., Barndorff-Nielsen, Hansen, Lunde, and Shephard, 2008; Gon¸calves and Meddahi, 2009; Mykland and Zhang, 2009; Christensen, Podolskij, and Vetter, 2013; Hounyo, 2017). It facilitates the control of some approximation errors in the proofs, but it can potentially be relaxed. In recent work, Christensen, Podolskij, Thamrongrat, and Veliyev (2016) operate with a power variation-based statistic and impose a weaker set of assumptions, which allows for rather unrestricted jump dynamics in volatility (see their equation (2.3) and Theorem (3.2), which is based on Assumption (H1) from Barndorff-Nielsen, Graversen, Jacod, Podolskij, and Shephard (2006)). It may be possible to extend our setting in that direction, but we leave a full exploration of it for future research. In some of our results, we also assume that the volatility is bounded away from zero. In particular, we sometimes adopt the following condition: Assumption (V’):

σsv,t > 0 and σu,t > 0, for all t ≥ 0.

At last, we impose that: Assumption (J): There exists a sequence of stopping times (e τn )∞ n=1 increasing to ∞ and a R deterministic nonnegative function γ˜n such that R γ˜n (x)β λ(dx) < ∞ and ||δ(ω, t, x)|| ∧ 1 ≤ γ˜n (x), for all (ω, t, x) with t ≤ τ˜n (ω), where β ∈ [0, 2]. β captures the activity of the jump process. As β approaches two, the jumps are smaller but more vibrant. As explained by Todorov and Bollerslev (2010), the harder they are to distinguish from the diffusive part of X. Below, we impose Assumption (J) to hold for any β ∈ [0, 1), thus restricting attention to jump processes with sample paths of finite length.

5

2.1

Microstructure noise

The presence of market frictions (such as price discreteness, rounding errors, bid-ask spreads, gradual response of prices to block trades and so forth) prevent us from observing the true, efficient log-price process Xt . Instead, we observe a noisy version Yt , which we assume is given by Yt = Xt + t ,

(4)

where t is a noise term that collects the market microstructure effects. We assume that t is independently distributed and independent of Xt , such that  2 ω2, E(t ) = 0 and E 2t = σu,t

(5)

for any t, where Yt is observed. As consistent with, e.g, Bandi and Russell (2006); Kalnina and Linton (2008), the second moment of the noise is allowed to be heteroskedastic and exhibit diurnal variation. We assume it is identical to the volatility diurnality, which conveniently makes the detrended noise asymptotically i.i.d (cf. (12)). We return to this later in Remark 3, where we highlight the impact of weakening it to a general form of heteroskedasticity. About the noise distribution, we follow Podolskij and Vetter (2009a): Assumption (A):

(i)  is distributed symmetrically around zero, and (ii) for any 0 > a > −1,

a

it holds that E(|t | ) < ∞. Assumption (A’):

Cramer’s condition is fulfilled, that is lim supt→∞ |χ(t)| < 1, where χ denotes

the characteristic function of .

2.2

Test of heteroskedasticity

To develop a test of the “no heteroskedasticity after diurnal correction” assumption, we partition the sample space Ω into the following two subsets: ΩH0 = {ω : σsv,t is constant for t ≥ 0},

(6)

and ΩHa = Ω{H0 . The null hypothesis can then formally be defined as H0 : ω ∈ ΩH0 , whereas the alternative is Ha : ω ∈ ΩHa . Our goal is to find a test with a prescribed asymptotic significance level and with power going to one to test the hypothesis that ω ∈ ΩH0 . The key challenge we address is how to construct such a test, when X—apart from being driven by a Brownian component—is subject to diurnal variation, potentially discontinuous and observed with measurement error. The solution is based on computing a set of estimators, which reveal information about the presence of time-variation in the stochastic volatility σsv,t robustly to the above features.

6

The differential form of (1) scaled by σu,t yields: dXt at σt dJt = dt + dWt + , σu,t σu,t σu,t σu,t

(7)

dXtd = adt dt + σsv,t dWt + dJtd ,

(8)

or

where a superscript d is used to represent a process that has been adjusted by the seasonal component of volatility.10 Then, we study the quadratic variation of X d : Z t X d 2 ds + |∆Xsd |2 , [X ]t = σsv,s 0

(9)

s≤t

Rt 2 P ds is the integrated variance of X d , while s≤t |∆Xsd |2 is the sum of the squared where 0 σsv,s d . deflated jumps, where ∆Xsd = Xsd − Xs− We note that if the stochastic volatility process is constant, say σsv,t = σ, (1) reduces to Z t      Xt = X0 + as ds + σσu,t Wt − W0 + δ1{|δ|≤1} ? µt − ν t + δ1{|δ|>1} ? µt ,

(10)

0

while [X d ]t = σ 2 t +

X

|∆Xsd |2 .

(11)

s≤t

The construction of the t-statistic now progresses in three steps. Firstly, we account for microstructure noise by doing local pre-averaging of Y d . Secondly, we tease out the continuous part of the quadratic variation by suitably removing the jump component in (11). Thirdly, we develop a fully feasible theory by proposing a statistic that can replace σu in the computations.

2.3

The pre-averaging approach

In this section, we confine the clock to t ∈ [0, 1], i.e. we set T = 1. In our simulations and empirical work, we implement the test “day-by-day,” so that here the unit interval is naturally interpreted as a trading day’s worth of data. The noisy log-price Yt is observed at regular time points ti = i/n, for i = 0, . . . , n. Then, the deflated intraday log-returns (at frequency n) can be computed as: d d ∆ni Y d ≡ Yi/n − Y(i−1)/n ,

i = 1, . . . , n.

(12)

As Ytd = Xtd + dt , we can split ∆ni Y d into ∆ni Y d = ∆ni X d + ∆ni d , 10

(13)

Note that many parts of this paper can be applied to both the raw and deflated log-returns series (e.g., the pre-averaging theory in the next subsection). We base the theoretical exposition on the seasonally adjusted version to minimize the notational load, while we present results for both series in the empirical application.

7

d d where ∆ni X d = Xi/n − X(i−1)/n denotes the n-frequency return of the efficient log-price, while

∆ni d = di/n − d(i−1)/n is the change in the microstructure component. To lessen the noise, we adopt the pre-averaging approach of Jacod, Li, Mykland, Podolskij, and Vetter (2009); Podolskij and Vetter (2009a,b). To describe it, we let kn be a sequence of positive integers and g a real-valued function. kn represents the length of a pre-averaging window, while g assigns a weight to those noisy log-returns that are inside it. g is defined on [0, 1], such that g(0) = R1 g(1) = 0 and 0 g(s)2 ds > 0. We assume g is continuous and piecewise continuously differentiable with a piecewise Lipschitz derivative g 0 . A canonical function that fulfills these restrictions is g(x) = min(x, 1 − x). We introduce the notation: Z Z 1 0 0 g (u)g (u − s)du and φ2 (s) = φ1 (s) =

1

g(u)g(u − s)du,

(14)

s

s

and for i = 1, 2, we let ψi = φi (0). For instance, if g(x) = min(x, 1 − x), it follows that ψ1 = 1 and ψ2 = 1/12. Also, we write: ψ1n

= kn

kn X j=1



j g kn





j−1 −g kn

!2 and

ψ2n

  kn −1 1 X j 2 = g . kn j=1 kn

(15)

In the appendix, after freezing the volatility locally, ψ1n and ψ2n appear in the conditional expectation of the squared pre-averaged return in (17). As n → ∞, ψ1n → ψ1

and ψ2n → ψ2 ,

(16)

 while ψin − ψi = O n−1/2 , for i = 1, 2, so we can work with ψi and not worry about the effect of this substitution in the asymptotic theory. In contrast, ψin can differ a lot from ψi , if kn is small, so as a practical guide it is better to work with (15). The pre-averaged return, say ∆ni Y¯ d , is then found by computing a weighted sum of consecutive n-frequency deflated log-returns over a block of size kn :  kX n −1  j n¯d g ∆i Y = ∆ni+j−1 Y d , i = 1, . . . , n − kn + 2. k n j=1

(17)

As readily seen, pre-averaging entails a slight “loss” of summands compared to n. Thus, while the n−kn +2 original sample size is n, there are only n − kn + 2 elements in (∆ni Y¯ d )i=1 . It follows from the n¯d n ¯d n d decomposition in (13) that ∆i Y = ∆i X + ∆i ¯ and, as shown by Vetter (2008), r !   kn 1 n ¯d n d ∆i X = Op and ∆i ¯ = Op √ . (18) n kn Thus, the noise is dampened, thereby reducing its influence on ∆ni Y¯ d . As an outcome, we retrieve a basically noise-free estimate, which can substitute the efficient log-return ∆ni X d in subsequent com-

8

n−kn +2 11 putations, taking proper account of the dependence introduced in (∆ni Y¯ d )i=1 . The reduction

increases with larger kn , but too much pre-averaging also impedes the accuracy of estimators of the quadratic variation, yielding a trade-off in selecting kn . To strike a balance and get an efficient n−1/4 rate of convergence, Jacod, Li, Mykland, Podolskij, and Vetter (2009) propose to set:  √ (19) kn = θ n + o n−1/4 , ¯ d and ∆n ¯d are balanced and equal to for some θ ∈ (0, ∞). With this choice, the orders of ∆ni X  i √  −1/4 . An example of (19) used throughout this paper is kn = θ n . Op n 2.3.1

The pre-averaged bipower variation

n +2 With the pre-averaged return series, (∆ni Y¯ d )n−k , available, Podolskij and Vetter (2009a) propose i=1

the bipower variation statistic: d

n

BV (Y , l, r) = n

l+r −1 4

Nn 1 X y(Y d , l, r)ni , µl µr i=1

(20)

where l, r ≥ 0, y(Y d , l, r)ni = |∆ni Y¯ d |l |∆ni+kn Y¯ d |r , Nn = n − 2kn + 2 and µp = E(|N (0, 1)|p ).12 In the following, if we write BV (l, r)n and y(l, r)ni , we assume that they are implicitly defined with respect to Y d . Podolskij and Vetter (2009a) show that under suitable regularity conditions, in particular that X is a continuous Itˆo semimartingale (i.e., X follows (A.7)), then as n → ∞  l+r Z 1 2 1 n p 2 2 BV (l, r) → BV (l, r) = θψ2 σsv,s + ψ1 ω ds, (21) θ 0 and n1/4

BV (l1 , r1 )n − BV (l1 , r1 )

!

BV (l2 , r2 )n − BV (l2 , r2 )

 d →s M N 0, Σ ,

(22)

 ds l1 ,r1 ,l2 ,r2 with l1 , r1 , l2 , r2 ≥ 0, where “→” is stable convergence, Σ = Σij the conditional 1≤i,j≤2  | n n | 1/4 covariance matrix of the limiting process n BV (l1 , r1 ) , BV (l2 , r2 ) , and the transpose.13 2.3.2

A truncated pre-averaged bipower variation

The estimator in (20) can also be made jump-robust in both the stochastic limit and its asymptotic distribution, but—as explained by Podolskij and Vetter (2009a)—this puts strong restrictions on l 11

If kn is even, it follows with the above definition of g(x) = min(x, 1 − x) that the pre-averaged returns in (17) can Pkn /2 d Pkn /2 d n +2 be rewritten as ∆ni Y¯ d = k1n j=1 Y i+kn /2+j − k1n j=1 Y i+j . Thus, the sequence (2∆ni Y¯ d )n−k can be interpreted i=1 n

n

as constituting a new set of increments from a price process that is constructed by simple averaging of the rescaled d n noisy log-price series, (Yi/n )i=0 , in a neighbourhood of i/n, thus making the use of the term pre-averaging and the associated notation transparent. 12 In order to avoid a finite sample bias in the construction of BV (l, r)n , we only divide it by Nn (the number of summands in the estimator) in our simulations and empirical work. We stick with n in the theoretical parts of the paper, as it involves less notation. 13 The formal definition of Σ is given in Appendix A.

9

and r. Firstly, the central limit theory in (22) is not valid for the popular choice l = r = 1. Indeed, Vetter (2010) shows that this estimator is not even mixed Gaussian, which severely constrains our ability to draw inference. Secondly, the version with l = r = 2 as implemented below, does not converge to the limit in (21), if X jumps, and while that is true for the pre-averaged (1,1)-bipower variation, asymptotically, it is well-known that the latter typically has a pronounced upward bias in finite samples (e.g., Christensen, Oomen, and Podolskij, 2014). Thus, to achieve a better jumprobustness and enlarge the feasible set of powers for which we can do hypothesis testing, we follow Corsi, Pirino, and Ren`o (2010) in the no-noise and finite-activity jump setting by combining the bipower idea with the truncation approach of Mancini (2009); Jacod and Protter (2012); Jing, Liu, and Kong (2014). To introduce our t-statistic for the homoskedasticity test, we therefore start by deriving a result as above for a truncated pre-averaged bipower variation, which verifies that the probability limit and asymptotic distribution of this new estimator are identical to those given by (21) and (22) in the general setting, where X follows the Itˆo semimartingale in (1). Thus, we propose to set: Nn X l+r −1 1 n ˇ 4 yˇ(l, r)ni , BV (l, r) = n µl µr i=1

where yˇ(l, r)ni = |∆ni Y¯ d |l 1{|∆n Y¯ d |<υn } |∆ni+kn Y¯ d |r 1{|∆n

¯ d |<υn }

i+kn Y

i

(23)

and 1{·} is the indicator function,

which discards pre-averaged log-returns that exceed a predetermined level υn = αu$ n , for α > 0 and $ ∈ (0, 1/2),

(24)

such that un = kn /n. Theorem 2.1 Let l1 , r1 , l2 and r2 be four positive real numbers and X that  be given by (1). Suppose  2 −1 1 −1 ≤$< ∨ 2(ll22+r Assumption (J) holds for some β ∈ [0, min{1, l1 , r1 , l2 , r2 }) and that 2(ll11+r +r1 −β) +r2 −β) 1/2. Furthermore, we assume (D1), (V), (A), and impose the moment condition E(|t |s ) < ∞, for some s > (3 ∨ 2(r1 + l1 ) ∨ 2(r2 + l2 )). If any li or ri is in (0, 1], we postulate (V0 ), otherwise either (V0 ) or (A0 ). In addition, suppose that kn → ∞ as n → ∞ such that (19) holds. Then, as n → ∞, ! ˇ (l1 , r1 )n − BV (l1 , r1 ) BV d n1/4 →s M N 0, Σ). (25) n ˇ (l2 , r2 ) − BV (l2 , r2 ) BV Theorem 2.1 shows that (23) is robust to the jump part in its limiting distribution. Note that Σ is identical to the matrix in (22). To our knowledge, the result is new with the main innovations being the statistic is (23) and the underlying process is a general Itˆo semimartingale given by (1). It extends Theorem 3 of Podolskij and Vetter (2009a) to discontinuous X by establishing a joint asymptotic distribution, as in (22), for the class of truncated pre-averaged bipower variation. In previous work, Jing, Liu, and Kong (2014) prove—under some regularity conditions—the consistency ˇ (2, 0)n , and CLT for the truncated pre-averaged realized variance, i.e. the statistic of the form BV when X follows (1). Our paper generalizes the latter article to the bipower setting with—subject 10

to the above constraint—arbitrary powers. The lower bound on $ is determined by an interplay between the bipower parameters and the activity of the jump process. The crude intuition is that small jumps tend to resemble Brownian motion, so if the threshold vanishes too slowly, it can impair the jump-robustness, and this effect is aggravated for larger bipowers. We therefore normally work with $ close to a half in practice. In our Monte Carlo, l1 = r2 = 2, l2 = r2 = 1 and β = 0.5, so that any $ ∈ [1/3, 1/2) is valid. Throughout, we always set $ = 0.49. The above enables extraction of an essentially noise-free and jump-robust estimate of the continuous piece of the quadratic variation in (9) and thus facilitates the construction of a test for the presence of time-variation in σsv,t . An implication of (25) is that for any l1 , r1 , l2 , r2 ≥ 0, which adhere to the conditions of Theorem 2.1 and such that l1 + r1 > l2 + r2 , as n → ∞, l +r

l +r

1 1 p ˇ (l1 , r1 )n − (BV ˇ (l2 , r2 )n ) l21 +r21 → BV BV (l1 , r1 ) − (BV (l2 , r2 )) l2 +r2

Z = 0

1

"Z  # ll1 +r 1  l1 +r  l2 +r  1 2 1 2 +r2 2 2 1 1 2 2 2 2 θψ2 σsv,s + ψ1 ω ds − θψ2 σsv,s + ψ1 ω ds ≥ 0, θ θ 0

(26)

with equality if and only if σsv,t is constant. We thus build a test of H0 via the infeasible t-statistic:  l1 +r1  n n l2 +r2 1/4 ˇ ˇ BV (l1 , r1 ) − (BV (l2 , r2 ) ) n d n √ Tinf. = → N (0, 1), (27) V where    2   l1 +r1  l1 +r1 l1 + r1 l1 + r1 n l2 +r2 −1 n 2 l2 +r2 −1 ˇ ˇ V = Σ11 − 2 BV (l2 , r2 ) Σ12 + Σ22 . (BV (l2 , r2 ) ) l2 + r2 l2 + r2

(28)

Note that the convergence in (27) holds only under H0 , while under Ha it follows from (26) that l1 +r1  ˇ (l1 , r1 )n −(BV ˇ (l2 , r2 )n ) l2 +r2 → ∞. This way we can determine if X d has homoskedastic or n1/4 BV heteroskedastic volatility with asymptotically correct size and power tending to one, as n → ∞. To render the test feasible, we propose a consistent estimator of Σ in Section 4, which can be plugged into (28). It is both inherently robust to heteroskedasticity and positive semi-definite.

3

A local estimator of diurnal variance

In the previous section, we pretended the diurnal component of volatility was available to deflate the noisy log-return series (i.e., (7)). In practice, σu is unobserved. We here propose a nonparametric jump- and noise-robust estimator of it and state appropriate conditions, under which the sampling error—induced by this estimation—is asymptotically negligible, so that it does not thwart the results in Section 2 (and 4). It turns out to be impossible to recover the latent diurnal variance on a fixed time interval. We thus resort to a long-span asymptotic theory, which extracts information about it by pooling high-frequency data across days. 11

As above, we suppose that on day t we record Y at equidistant time points ti = t − 1 + i/n, for i = 0, 1, . . . , n and write the associated n-frequency log-returns as: ∆n(t−1)n+i Y ≡ Yt−1+i/n − Yt−1+(i−1)/n ,

for t = 1, . . . , T and i = 1, . . . , n.

(29)

As in Zhang, Mykland, and A¨ıt-Sahalia (2005), we operate within a two time scale framework, where the “slow” scale uses a coarser set of m-frequency returns, where m < n, i.e. ∆m (t−1)m+j Y = Yt−1+j/m − Yt−1+(j−1)/m , for t = 1, . . . , T and j = 1, . . . , m, which is reserved for diurnal variance estimation, while the “fast” scale is based on all observed n-frequency returns and is intended for n a bias-correction. Throughout, we assume m is a divisor of n, so that {j/m}m j=0 ⊆ {i/n}i=0 . In the following, we say a process (bt )t≥0 is bounded in Lp , if   sup E |bt |p < ∞.

(30)

t∈R+

Assumption (D4):

(at )t≥0 , (˜ at )t≥0 , (˜ σt )t≥0 and (˜ vt )t≥0 are bounded in L4 .

Assumption (D4) adds some regularity to the driving processes in X, which is necessary here as T → ∞ in the asymptotic theory, and so we cannot appeal to the standard “localization” procedure (e.g., Jacod, 2008) to bound various terms in the proofs. Now, we set: 2 σ ˆu,s

T T X 1X√ m 2 m ( m∆(t−1)m+j Y ) − [var( ˆ t−1+(j−1)/m )+var( ˆ t−1+j/m )], = T t=1 T t=1

for s ∈ [(j−1)/m, j/m), (31)

where var( ˆ t−1+(j−1)/m ) is a consistent estimator of var(t−1+(j−1)/m ), which has to converge at a rate faster than m−1 , e.g. n/m  T X   X 1 1 n n var( ˆ t−1+(j−1)/m ) = − ∆i+[t−1+ j−1 ]n Y ∆i−1+[t−1+ j−1 ]n Y . m m T n/m − 1 t=1 i=1

(32)

2 As readily seen, σ ˆu,s is based directly on the raw noisy high-frequency data. It does not require

jump-truncation nor pre-averaging and is therefore trivial to compute.14 Due to its reliance on the squared normalized noisy high-frequency increment, however, it accumulate a bias from the microstructure noise, which the second term in (31) cancels out by computing a local block-wise estimator of the noise variance. 2 While it appears counterintuitive, σ ˆu,s is also jump-robust, as we show below. The intuition is

that in our model, there are no fixed points of discontinuity in X, so that the influence of any jumps is intrinsically averaged away, as m → ∞, T → ∞ and n → ∞. Proposition 3.1 Assume that X is given by (1). Moreover, we suppose Assumption (D1) – (D4), 14

It is naturally also possible to pre-average and follow up with jump-truncation. This may lead to a better rate of convergence for the diurnal variance estimator. Here, we do not follow this line of thought, as it requires extra tuning parameters, and because the current setup appears to work reasonably well in practice.

12

(V), (V’), (A) and (A’). If m → ∞ and T → ∞, as n → ∞, then it holds that 2 2 σ ˆu,s = σu,s + OP (mT −1/2 ) + OP (m3/2 n−1/2 T −1/2 ).

(33)

q q  1 2 2 2 − 2 = σ σ ˆu,s σu,s ˆu,s − σu,s ' p 2 σ ˆu,s − σu,s , 2 σu,s

(34)

Next, we note that:

2 bounded away from 0. Thus, we can write: with σu,s

σ ˆu,s = σu,s + OP (mT −1/2 ) + OP (m3/2 n−1/2 T −1/2 ).

(35)

It therefore follows that if m ∝ nδ2

T ∝ nδ 3 ,

and

(36)

where δ2 ∈ (0, 1/2] and δ3 > 1/2+2δ2 (for a fixed value of δ2 ), then the error induced from estimating σu does not alter the analysis in Section 2 and 4. That is, neither Theorem 2.1 or (27) are affected, nor is the bootstrap applied to yˇ(l, r)ni = |∆ni Y¯ d |l 1{|∆n Y¯ d |<υn } |∆ni+kn Y¯ d |r 1{|∆n Y¯ d |<υn } , where (with i

a slight abuse of notation) we redefine  kX n −1  j n¯d ∆i Y = g ∆ni+j−1 Y d , k n j=1

i+kn

i = 1, . . . , n − kn + 2,

(37)

to be based on: ∆ni+j−1 Y d =

∆ni+j−1 Y , σ ˆu, i+j−1

(38)

n

with σ ˆu, i+j−1 from (31). n

Remark 1 If the noise is autocorrelated but not heteroskedastic, ω ˆ 2 = var( ˆ t−1+(j−1)/m ) given by (32) is no longer a consistent estimator of ω 2 = var(). Indeed, when the noise is a stationary qdependent sequence (for known q > 0), the statistic defined in (32) estimates the quantity 2 ρ(0) −  ρ(1) , where ρ(k) = cov(1 , 1+k ). Hautsch and Podolskij (2013, Lemma 2) discuss an estimator of ρ(k), k = 0, . . . , q + 1, which is obtained from a simple recursion formula. Building on their result, we can deduce an estimator of ω 2 in this alternative setup: 2

ω ˆ =−

q+1 X

kˆ γ (k),

(39)

k=1

where T n/m−k   X X  1 1 n n γˆ (k) = ∆i+[t−1+ j−1 ]n Y ∆i+k+[t−1+ j−1 ]n Y , for k = 0, . . . , q + 1. m m T n/m − k t=1 i=1

(40)

Then, p

ω ˆ 2 → ω 2 = ρ(0).

13

(41)

The Monte Carlo and empirical analysis is based on 2 σ ˆu,s

T 1X√ 2 ≡ ω2, ( m∆m j+(t−1)m Y ) − 2mˆ T t=1

for s ∈ [(j − 1)/m, j/m),

(42)

where ω ˆ 2 is (39) with q = 3.

4

The bootstrap

In this section, we improve the quality of inference in our test of heteroskedasticity in the noisy jumpdiffusion setting by relying on the bootstrap, when computing critical values for the t-statistic. This is warranted by the Monte Carlo in Section 5, which reveals that in small samples, the feasible version of (27) (cf. (66)) is poorly approximated by the standard normal. Next, we propose a bootstrap ˇ (l1 , r1 )n , BV ˇ (l2 , r2 )n )| , estimator of the conditional covariance matrix of the limiting process n1/4 BV i.e. Σ. As the bootstrap estimator is positive semi-definite by construction, it renders our test implementable. We build on a series of papers in the high-frequency volatility area. In particular, Gon¸calves and Meddahi (2009) propose the wild bootstrap for realized variance, in a framework where the asset price is observed without error (see also Hounyo, 2018). Gon¸calves, Hounyo, and Meddahi (2014) and Hounyo, Gon¸calves, and Meddahi (2017) extend their work to accommodate noise. The latter studies the pre-averaged realized variance estimator—i.e., BV (2, 0)n —proposed by Jacod, Li, Mykland, Podolskij, and Vetter (2009), where the pre-averaged returns are both overlapping and heteroskedastic due to stochastic volatility. In this context, a block bootstrap applied to (∆ni Y¯ d )n−kn +2 appears natural. i=1

Nevertheless, such a scheme is only consistent if σsv,t is constant. As shown by Hounyo, Gon¸calves, and Meddahi (2017), the problem is that |∆ni Y¯ d |2 are heterogeneously distributed under time-varying volatility.15 In particular, their mean and variance are unequal. This creates a bias term in the blocks of blocks bootstrap variance estimator. To cope with both dependence and heterogeneity of |∆ni Y¯ d |2 , they combine the wild bootstrap with the blocks of blocks bootstrap. The procedure exploits that heteroskedasticity can be handled by the former, while the latter can replicate serial dependence in the data. Hounyo (2017) generalizes Hounyo, Gon¸calves, and Meddahi (2017) to a broad class of covariation estimators in a general setting that accommodates jumps, microstructure noise, irregularly spaced high-frequency data and non-synchronous trading. Also, Dovonon, Gon¸calves, Hounyo, and Meddahi (2017) develop a new local Gaussian bootstrap for high-frequency jump testing, but market microstructure noise is supposed to be absent. Here, we allow for noise and concentrate on heteroskedasticity. 15

This feature is highlighted by the asymptotic distribution of ∆ni Y¯ d in (90) below.

14

ˇ (l, r)n is The bootstrap version of BV Nn X −1 1 ˇ (l, r)n∗ = n l+r 4 yˇ(l, r)n∗ BV i , µl µr i=1

(43)

Nn n where (ˇ y (l, r)n∗ y (l, r)ni )N i )i=1 is a bootstrap sample from (ˇ i=1 .

We apply a bootstrap to yˇ(l, r)ni , which replicates their dependence and heterogeneity. As suggested by Hounyo, Gon¸calves, and Meddahi (2017), we merge the wild bootstrap with blockbased resampling. However, our bootstrap is new, and it can be viewed as an overlapping version of their algorithm. We name it “the overlapping wild blocks of blocks bootstrap.” We note that the degree of overlap among the blocks to be bootstrapped plays a major role in efficiency: the nonoverlapping block-based approach is less efficient than a partial or full-overlap block (e.g., Dudek, Le´skow, Paparoditis, and Politis, 2014). To describe this approach, let bn be a sequence of integers, which will denote the bootstrap block size, such that for some δ1 ∈ (0, 1):  bn = O nδ1 .

(44)

n We divide (ˇ y (l, r)ni )N i=1 into overlapping blocks of size bn . The total number of such blocks is

Nn − bn + 1. The bootstrap is based on Nn − 2bn + 2 of them. In particular, we look at overlapping n −bn blocks within the set (ˇ y (l, r)ni )N (there is Jn = Nn − 2bn + 1 many such blocks) and the last i=1

block containing the elements yˇ(l, r)nNn −bn +1 , . . . , yˇ(l, r)nNn . The bootstrap sample is constructed by properly combining the first Jn blocks. To explain this setup and avoid confusion, note that the main ingredient behind the theoretical validity of the suggested resampling scheme is that we center all bootstrap draws from a block of bn consecutive observations, say the jth that holds yˇ(l, r)nj , . . . , yˇ(l, r)nj+bn −1 , around a local average of data in the (j + bn )th block (which is thus shifted to the right and consists of ¯j+bn in (45) below. This principle is no longer applica), as given by B yˇ(l, r)n , . . . , yˇ(l, r)n j+bn

j+2bn −1

ble starting with the block that covers the elements yˇ(l, r)nNn −2bn +2 , . . . , yˇ(l, r)nNn −bn +1 , because the centering here demands a local average to be computed from yˇ(l, r)nNn −bn +2 , . . . , yˇ(l, r)nNn +1 , and the last observation is not available. Let u1 , . . . , uJn +1 be i.i.d. random variables, whose distribution is independent of the original  sample. We denote by µ∗q = E ∗ uqj its qth order moments.16 Then, bn 1 X ¯ Bj = yˇ(l, r)ni−1+j , bn i=1

j = 1, . . . , Nn − bn + 1,

(45)

As usual in the bootstrap literature, P ∗ (E ∗ and var∗ ) denotes the probability measure (expected value and variance) induced by the resampling, conditional on a realization of the original time series. In addition, for a 16

p∗

sequence of bootstrap statistics Zn∗ , we write (i) Zn∗ = op∗ (1) or Zn∗ → 0, as n → ∞, if for any ε > 0, δ > 0, limn→∞ P [P ∗ (|Zn∗ | > δ) > ε] = 0, (ii) Zn∗ = Op∗ (1) as n → ∞, if for all ε > 0 there exists an Mε < ∞ such that d∗

limn→∞ P [P ∗ (|Zn∗ | > Mε ) > ε] = 0, and (iii) Zn∗ → Z as n → ∞, if conditional on the sample Zn∗ converges weakly to Z under P ∗ , for all samples contained in a set with probability P converging to one.

15

is the average of the data in the jth block consisting of yˇ(l, r)nj , . . . , yˇ(l, r)nj+bn −1 . Next, we generate the overlapping wild blocks of blocks bootstrap observations by:   1 Pm n ¯ √ − B uj , y ˇ (l, r) b +j n  m j=1 bn     P bn  ¯m+j um+j−bn ,  √1 ˇ(l, r)nm − B j=1 y bn n∗ Nn ¯ yˇ(l, r)m − B =  PNn −bn +1−m n  ¯Jn +1−j+bn uJn +1−j , √1 y ˇ (l, r) − B  m j=1  b n     √1 n ¯Nn −bn +1 uJn +1 , − B y ˇ (l, r) m bn

if m ∈ I1n , if m ∈ I2n , if m ∈ I3n ,

(46)

if m ∈ I4n ,

where Nn 1 X Nn ¯ B = yˇ(l, r)ni , Nn i=1

(47)

and I1n = {1, . . . , bn − 1},

I2n = {bn , . . . , Jn },

I3n

I4n

(48) = {Jn + 1, . . . , Nn − bn },

= {Nn − bn + 1, . . . , Nn }.

¯ Nn , instead of It is interesting to note that if we were to center yˇ(l, r)nm around the grand mean B ¯j+m , it would yield a bootstrap observation the localized block average B   n∗ Nn n Nn ¯ ¯ yˇ(l, r)m − B = yˇ(l, r)m − B ηm , (49) for m ∈ I2n (the main set), where ηm =

√1 bn

Pbn

j=1

um+j−bn . Therefore, under the assumption

that E(uj ) = 0 and var(uj ) = 1, we find that E(ηm ) = 0, var(ηm ) = 1, and cov(ηm , ηm−k ) =  1 − bkn 1{k≤bn } . Thus, our approach is related to the dependent wild bootstrap of Shao (2010) (see also, e.g., Hounyo (2014)), who extends the traditional wild bootstrap of Wu (1986); Liu (1988) to the time series setting, and it is the special case, where the kernel function is assumed to be Bartlett (see Assumption 2.1 in Shao, 2010). ¯j+m is to deal with the mean heterogeneity of yˇ(l, r)n . As The idea of the new centering B m shown by Hounyo, Gon¸calves, and Meddahi (2017), for the case of squared pre-averaged returns y(2, 0)nm , centering the non-overlapping wild blocks of blocks bootstrap around the corresponding P n n grand mean Nn−1 N i=1 y(2, 0)i does not work, when σsv,t is time-varying. In this paper, we show that generating the bootstrap observations as in (46) does yield an asymptotically valid bootstrap ˇ (l1 , r1 )n , BV ˇ (l2 , r2 )n )| , even if σsv,t is not constant. for (BV As in Shao (2010) and Hounyo (2014), the dependence between neighboring observations yˇ(l, r)nm and yˇ(l, r)nm0 is not only preserved, if m and m0 belong to a particular block, as typical in block-based resampling. Indeed, if |m − m0 | < bn , yˇ(l, r)n∗ ˇ(l, r)n∗ m and y m0 are conditionally dependent (except for the last bn data). A common feature of the block-based bootstrap, in particular the non-overlapping wild blocks of blocks approach of Hounyo, Gon¸calves, and Meddahi (2017), is that if the sample size Nn is not a multiple of bn , then one has to either take a shorter bootstrap sample or use a fraction of the last resampled block. This leads to some inaccuracy, when bn is large relative to Nn . In contrast, for 16

the overlapping version proposed in this paper, the size of the bootstrap sample is always equal to the original sample size.17 Write Nn 1 X Nn ∗ ¯ B = yˇ(l, r)n∗ i , Nn i=1

(50)

¯ Nn ∗ suggests that we as the average value of the bootstrap observations. A closer inspection of B ¯ Nn ∗ − B ¯ Nn as follows can rewrite the centered bootstrap sample mean B Nn



Jn   1 X Nn ∗ Nn ¯ ¯ ¯j − B ¯j+bn uj . bn B B −B =√ bn j=1

(51)

Thus, Jn  1 X −1 1 ˇ (l, r)n∗ = BV ˇ (l, r)n + n l+r ¯j − B ¯j+bn uj 4 √ BV bn B µl µr bn j=1

(52) Jn 1 X n ˇ ˇ ∆B(l, r)nj uj , = BV (l, r) − √ bn j=1

where ˇ r)nj , ˇ ˇ r)nj+b − B(l, ∆B(l, r)nj = B(l, n

(53)

such that bn X −1 1 ˇ r)nj = n l+r 4 B(l, yˇ(l, r)ni−1+j . (54) µl µr i=1   ˇ (l1 , r1 )n∗ BV 1/4 We can now derive the first and second bootstrap moment of n ˇ (l2 , r2 )n∗ . The following BV Lemma states the formulas.

Lemma 4.1 Assume that yˇ(l, r)n∗ m are generated as in (46). Then, it follows that Jn  1 X n∗ n ˇ ˇ ˇ E BV (l, r) = BV (l, r) − √ ∆B(l, r)nj E ∗ (uj ), bn j=1 ∗

(55)

Also, for 1 ≤ i, j ≤ 2, ∗

1/4

cov n

ˇ (li , ri ) , n BV n∗

1/4

ˇ (lj , rj ) BV

n∗



√ X J n n ˇ ˇ j , rj )nk var∗ (uk ). = ∆B(li , ri )nk ∆B(l bn k=1

(56)

ˇ (l, r)n∗ is an unbiased estimator of Equation (55) of Lemma 4.1 implies that with E ∗ (uj ) = 0, BV  ˇ (l, r)n , i.e. E ∗ BV ˇ (l, r)n∗ = BV ˇ (l, r)n . The second part shows that the bootstrap covariance BV 17

We compared the rejection rate of our test based on the overlapping bootstrap suggested here to the nonoverlapping version of Hounyo, Gon¸calves, and Meddahi (2017). We found our procedure has better size control, even if bn is a multiple of Nn . Thus, both from a theoretical and practical viewpoint our approach is preferable.

17

ˇ (li , ri )n∗ and n1/4 BV ˇ (lj , rj )n∗ depends on the variance of u. In particular, if we select of n1/4 BV var∗ (u) = 1/2 as in Hounyo, Gon¸calves, and Meddahi (2017): !! ˇ (l1 , r1 )n∗ BV ˇ n, var∗ n1/4 =Σ n∗ ˇ BV (l2 , r2 )  ˇn = Σ ˇ l1 ,r1 ,l2 ,r2 ,n where Σ and ij 1≤i,j≤2 ˇ li ,ri ,lj ,rj ,n = Σ ij

(57)

√ X J n n ˇ ˇ j , rj )n . ∆B(li , ri )nk ∆B(l k 2bn k=1

(58)

ˇ n as Note that based on (58), we can rewrite Σ √ X J n n ˇ ˇ| n ˇ Σ = ξj ξ , 2bn j=1 j

(59)

|  ˇ 2 , r2 )n . It follows that if the external random variable u is selected ˇ 1 , r1 )n , ∆B(l where ξˇj ≡ ∆B(l j j as above, the overlapping wild blocks of blocks bootstrap variance estimator is consistent for the  ˇ (l1 , r1 )n , BV ˇ (l2 , r2 )n | provided Σ ˇ n is a consistent estimator of Σ, asymptotic variance of n1/4 BV ˇ n is related to recent work on asymptotic variance as proved in Theorem 4.1 below. Note that Σ estimation by Mykland and Zhang (2017); see also, e.g., Christensen, Podolskij, Thamrongrat, and Veliyev (2016); Jacod and Todorov (2009); Mancini and Gobbi (2012). ˇ n as follows: Remark 2 Note that from (59), we can also rewrite Σ bn X ˇn = 1 ˇn , Σ Σ bn m=1 m

(60)

where √ ˇ nm Σ

=

n 2

bNn /bn c−2

X

  | ˇ l1 ,r1 ,l2 ,r2 ,n ξˇjbn +m ξˇjb = Σ ij,m n +m

j=0

.

(61)

1≤i,j≤2

1 ,r1 ,l2 ,r2 ,n 1 ,r1 ,l2 ,r2 ,n ˇ n , i.e. Σ ˇ l22,m ˇ l11,m We deduce that the diagonal elements of Σ and Σ are nothing else m ˇ (l1 , r1 )n and than the consistent bootstrap variance estimators of the asymptotic variance of n1/4 BV ˇ (l2 , r2 )n , as proposed by Hounyo (2017). n1/4 BV

ˇ n converges in The next result shows that under some regularity conditions, the estimator Σ probability to Σ in a general Itˆo semimartingale context. Theorem 4.1 Assume that X fulfills Assumption (J) for some β ∈ [0, 2]. Furthermore, suppose that the conditions of Theorem B.1 in Appendix B hold true, when X is continuous (i.e., X follows (A.7)), and also if X has jumps (i.e., X follows (1)) with either l1 + r1 + l2 + r2 ≤ 4(1 − δ1 ),

18

0 ≤ β < 4(1 − δ1 ),

(62)

or l1 + r1 + l2 + r2 > 4(1 − δ1 ),

0 ≤ β < 4(1 − δ1 ),

l1 + r1 + l2 + r2 − 4(1 − δ1 ) 1 ≤ $ < . (63) 2(l1 + r1 + l2 + r2 − β) 2

Then, as n → ∞, it holds that p ˇn → Σ Σ,

(64)

where Σ is defined in Appendix A. In our Monte Carlo studies and empirical application, we take l1 = r1 = 2 and l2 = r2 = 1. Here, (63) holds provided β < 4(1 − δ1 ). As 1/2 < δ1 < 2/3 by assumption (i.e, 4/3 < 4(1 − δ1 ) < 2), it therefore suffices that β ∈ [0, 4/3). Theorem 4.1 implies that in finite samples, we get a consistent and nonnegative estimator of V :  2      ll1 +r  l1 +r1 1 −1 l + r l + r 1 1 1 1 n n n 2 l2 +r2 −1 ˇ n n n +r2 ˇ ˇ ˇ ˇ ˇ 2 BV (l2 , r2 ) Σ12 + BV (l2 , r2 ) Σ22 . (65) V = Σ11 − 2 l2 + r2 l2 + r2 Corollary 4.1 Assume that the conditions from Theorem 4.1 hold true. If X is given by (10), 1 3 such that Assumption (J) holds for some β ∈ [0, 1) and 2(2−β) ∨ 2(4−β) ≤ $ < 1/2. Then, if l1 + r1 > l2 + r2 and as n → ∞,    l1 +r1 n l2 +r2 n 1/4 ˇ ˇ BV (l1 , r1 ) − BV (l2 , r2 ) n d n √ → N (0, 1). T ≡ Vˇ n

(66)

Corollary 4.1 delivers the asymptotic normality of the studentized statistic T n ; the feasible ˇ (l1 , r1 )n − version of (27). Note that under the alternative presence of heteroskedasticity, BV l1 +r1  ˇ (l2 , r2 )n l2 +r2 converges to a strictly positive random variable. Moreover, as Vˇ n was shown BV to be a robust estimator of V even in the presence of stochastic volatility, jumps and noise, we can conclude that the statistic T n → ∞ if the realization of X d has a heteroskedastic volatility path. Therefore, appealing to the properties of stable convergence, we deduce that  lim P T n > z1−α | ΩH0 = α, n→∞

 lim P T n > z1−α | ΩHa = 1

n→∞

(67) (68)

where zα is the α-quantile of a standard normal distribution. The implication is that we reject H0 , if T n is significantly positive. While the alternative inference procedure based on (66) does not require any resampling, it possesses inferior finite sample properties, as shown in Section 5. Remark 3 The results from Jacod, Podolskij, and Vetter (2010) and Podolskij and Vetter (2009a) indicate that some assumptions can be relaxed. In particular, in Corollary 4.1, if all the powers are even numbers (e.g., l1 = 4, r1 = 0, l2 = 2 and r2 = 0), we can prove the results in the general setting of Jacod, Podolskij, and Vetter (2010) with heteroskedastic noise, which is of a general form   2 E 2t | X = ωt2 and not necessarily restricted to E 2t | X = ω 2 σu,t . Here, the null is modified as   H0 : ω ∈ ΩH0 ∩ ω : t 7−→ var dt | X is constant on [0, 1] , (69) 19

ˇ (1, 1)n . Figure 1: Proportion of microstructure noise in BV Panel A: θ = 1/3.

Panel B: θ = 1.

1.25

1.25 SPY cross−sectional average (rest)

SPY cross−sectional average (rest)

1.00

proportion (in percent)

proportion (in percent)

1.00

0.75

0.50

0.25

0.00 2010

0.75

0.50

0.25

2011

2012 time

2013

2014

0.00 2010

2011

2012 time

2013

2014

ˇ (1, 1)n that is due to residual variation (after pre-averaging) in the microstructure noise. Note. We plot the proportion of BV kn n ˇ BV (1, 1) is rescaled by θψ2 to provide an estimate of the integrated variance up to a bias term ψ1kn ω 2 /(θ2 ψ2kn ), see (21) and ˇ (1, 1)n estimate over time for the ticker symbols included in Theorem 2.1. The figure shows the ratio of the bias to the total BV our empirical analysis. ω 2 is replaced by the robust estimator ω ˆ 2 in (39) computed daily with q = 3.

where dt ≡ t /σu,t . This suggests that in the presence of heteroskedastic noise of a general form, H0 is a joint statement about the constancy of both diffusive and the rescaled noise variance. Such information should be useful in practice, because it delivers knowledge about the presence of heteroskedasticity irrespective of its origin. Meanwhile, Figure 1 shows that for our empirical highˇ (1, 1)n is almost exclusively frequency data—and the two choices of θ adopted in the paper—BV composed by diffusive volatility. This implies that very little residual noise is left in the data after pre-averaging, which indicates that any rejection of the null is more likely due to genuine timevariation in σsv,t . We note that the dampening of the noise is naturally much weaker for θ = 1/3, which is therefore more susceptible to reject H0 on this ground.

Corollary 4.2 Assume that the conditions of Theorem 4.1 hold true and the external random  i.i.d. variable is chosen as uj ∼ E ∗ (uj ), var∗ (uj ) , such that var∗ (uj ) = 1/2. Then, as n → ∞, !! ˇ (l1 , r1 )n∗ BV p ˇn → Σ, (70) var∗ n1/4 =Σ ˇ (l2 , r2 )n∗ BV both in model (A.7) and (1), where Σ is defined in Appendix A. Given the consistency of the bootstrap variance estimator, we now prove the associated convergence  ˇ (l1 , r1 )n∗ , BV ˇ (l2 , r2 )n∗ | . of the bootstrap distribution of n1/4 BV 20

2+δ Theorem 4.2 Assume that all conditions from Corollary 4.2 hold true and that E ∗ |uj | <∞ for some δ > 0. Then, as n → ∞, ˇ Σ

 n −1/2

ˇ (l1 , r1 )n∗ − E ∗ BV ˇ (l1 , r1 )n∗ BV

n1/4

 !

 ˇ (l2 , r2 )n∗ − E ∗ BV ˇ (l2 , r2 )n∗ BV

d∗

→ N (0, I2 ),

(71)

in probability-P , both in model (A.7) and (1). Moreover, let # "  +r1    ll1 +r l1 +r1    ˇ (l1 , r1 )n∗ − E ∗ BV ˇ (l2 , r2 )n∗ 2 2 ˇ (l1 , r1 )n∗ − BV ˇ (l2 , r2 )n∗ l2 +r2 − E ∗ BV n1/4 BV S n∗ =



,

V

(72) where l1 + r1 > l2 + r2 . It holds that     l1 +r1  p 1/4 n∗ ∗ n∗ n∗ l2 +r2 ˇ ˇ V ≡ var n BV (l1 , r1 ) − BV (l2 , r2 ) → V,

(73)

and d∗

S n∗ → N (0, 1),

(74)

in probability-P , both in model (10) and (1). Theorem 4.2 shows that the normalized statistic S n∗ is asymptotically normal both in model (10) d∗

and (1). This implies, independently of whether H0 or Ha is true, S n∗ → N (0, 1), in probability-P . This ensures that the following bootstrap test both controls the size and is consistent under the alternative. Let h  l1 +r1    l1 +r1 i ˇ (l1 , r1 )n∗ − BV (l2 , r2 )n∗ l2 +r2 − E ∗ BV (l1 , r1 )n∗ − E ∗ (BV (l2 , r2 )n∗ ) l2 +r2 Z n∗ ≡ n1/4 BV (75) and n

Z ≡n

1/4

  +r1  ll1 +r n n ˇ (l2 , r2 ) 2 2 . ˇ (l1 , r1 ) − BV BV

(76)

Remark 4 We reject H0 at level α, if Z n > p∗1−α , where p∗1−α is the (1 − α)-percentile of the d∗

bootstrap distribution of Z n∗ . Under the conditions of Theorem 4.2, the statistic Z n∗ → N (0, V ), in d∗

d

st probability-P . Note that as Z n → N (0, V ) on ΩH0 , the fact that Z n∗ → N (0, V ), in probability-P ,

ensures that the test has correct size, as n → ∞. On the other hand, under the alternative (i.e. on d∗ ΩHa ), as Z n diverges at rate n1/4 , but we still have that Z n∗ → N (0, V ) = Op∗ (1), the test has unit power asymptotically. The above bootstrap test is convenient, as it does not require estimation of the asymptotic variance-covariance matrix Σ, but it may not lead to asymptotic refinements. In order to attain such improvement (or at least be able to prove it), we should base the bootstrap on an asymptotically pivotal t-statistic. To this end, we propose a consistent bootstrap estimator of   n ∗ 1/4 ˇ ˇ (l1 , r1 )n∗ , BV ˇ (l2 , r2 )n∗ | . We look at the following adjusted bootstrap version Σ = var n BV 21

  ˇ n given by Σ ˇ n∗ = Σ ˇ l1 ,r1 ,l2 ,r2 ,n∗ of Σ ij

ˇ n∗ are , where the individual entries of Σ 1≤i,j≤2



ˇ li ,ri ,lj ,rj ,n∗ Σ ij

J

n n var∗ (u) X ˇ i , ri )n∗ ∆B ˇ ∗ (lj , rj )n , ∆B(l = k k ∗ 2 bn E (u ) k=1

(77)

with n ˇ ˇ ∆B(l, r)n∗ j = ∆B(l, r)j uj ,

(78)

n ˇ where ∆B(l, r)nj is from (53) and (uj )Jj=1 are the external random variables used to generate the

bootstrap observations in (46). We can also write √ Jn n var∗ (u) X | n∗ ˇ ξˇj∗ ξˇj∗ , Σ = ∗ 2 bn E (u ) j=1

(79)

 | ˇ 1 , r1 )n , ∆B(l ˇ 2 , r2 )n . We can show that Σ ˇ n∗ consistently estimates Σ ˇ n for where ξˇj∗ ≡ uj ∆B(l j j  ˇ n∗ we construct a any choice of external random variable u with E ∗ |uj |4 < ∞. Next, based on Σ bootstrap studentized variant of (66): Z n∗ , T n∗ ≡ √ Vˇ n∗

(80)

where    2    l1 +r1  l1 +r1 l1 + r1 l1 + r1 n∗ n∗ n l2 +r2 −1 ˇ n∗ n 2 l2 +r2 −1 ˇ n∗ ˇ ˇ ˇ ˇ V = Σ11 − 2 BV (l2 , r2 ) BV (l2 , r2 ) Σ22 . Σ12 + l2 + r2 l2 + r2

(81)

Theorem 4.3 Assume that the conditions of Corollary 4.2 are true and the external random vari  i.i.d. able is chosen as uj ∼ E ∗ (uj ), var∗ (uj ) , such that E ∗ |uj |4+δ < ∞ for some δ > 0. Then, as n → ∞,  ˇ n∗ −1/2 1/4

Σ

n

ˇ (l1 , r1 )n∗ ˇ (l1 , r1 )n∗ − E ∗ BV BV

 !

 ˇ (l2 , r2 )n∗ ˇ (l2 , r2 )n∗ − E ∗ BV BV

d∗

→ N (0, I2 ),

(82)

in probability-P , both in model (A.7) and (1). Also, d∗

T n∗ → N (0, 1),

(83)

in probability-P , both in model (10) and (1). Theorem 4.3 shows the asymptotic normality of the studentized statistic T n∗ . An implication ∗ ∗ of results in Theorem 4.3 is that we reject H0 at significance level α, if T n > q1−α , where q1−α is

the (1 − α)-percentile of the bootstrap distribution of T n∗ .

5

Monte Carlo analysis

We here assess the properties of the nonparametric noise- and jump-robust test of deterministic versus stochastic variation in the intraday volatility coefficient that was proposed in Section 2. We 22

also highlight the refinements that can potentially be offered by the bootstrap, as outlined in Section 4, in sample sizes that resemble those, we tend to encounter in practice. We do so via detailed and realistic Monte Carlo simulations, and we start by describing the design of the study. To simulate the efficient log-price Xt , we adopt the model: dXt = at dt + σt dWt + dJt ,

(84)

where X0 = 0, at = 0.03 (per annum) and the other components are defined below. As above, σt = σsv,t σu,t . To describe σu,t , we follow earlier work of Hasbrouck (1999) and Andersen, Dobrev, and Schaumburg (2012) by using the specification: σu,t = C + Ae−a1 t + Be−a2 (1−t) .

(85)

We set A = 0.75, B = 0.25, C = 0.88929198 and a1 = a2 = 10.18 This produces a pronounced, asymmetric reverse J-shape in σu,t with a value of about 1.8 (1.1) times higher at the start (end) of each simulation compared to the observations in the middle. This is also a good description of the actual intraday volatility pattern observed in our empirical high-frequency data (cf. Panel B in Figure 5). We assume that σsv,t follows a stochastic volatility two-factor structure (SV2F):19 σsv,t = s-exp (β0 + β1 τ1,t + β2 τ2,t ) ,

(86)

where dτ1,t = α1 τ1,t dt + dB1,t ,

dτ2,t = α2 τ2,t dt + (1 + φτ2,t )dB2,t .

(87)

Here, B1,t and B1,t are two independent standard Brownian motions with E (dWt dB1,t ) = ρ1 dt and E (dWt dB2,t ) = ρ2 dt. We follow Huang and Tauchen (2005) and use the parameters β0 = −1.2, β1 = 0.04, β2 = 1.5, α1 = −0.00137, α2 = −1.386, φ = 0.25 and ρ1 = ρ2 = −0.3.20 This means that the first factor becomes a slowly-moving component, which generates persistence in volatility, while the second is a fast mean-reverting process that allows for a sufficient amount of volatility-of-volatility. At the start of each simulation, we initialize τ1 at random from its stationary distribution, i.e. τ1,0 ∼  N 0, −[2α1 ]−1 . Meanwhile, τ2 is started at τ2,0 = 0 (e.g., Barndorff-Nielsen, Hansen, Lunde, and Shephard, 2008). In absence of stochastic volatility, i.e. under the null hypothesis of deterministic diurnal variation, we freeze σsv,t at a value equal to the unconditional expectation implied by the above SV2F 2 model, i.e. σ 2 = E(σsv,t ). R1 2 The calibration of C ensures that 0 σu,t dt = 1. 19 The s-exp function is used to denote the exponential function that has been spliced with a polynomial of linear x0 growth at high values of its argument, i.e. s-exp(x) = ex if x ≤ x0 and s-exp(x) = √ e 2 2 , if x > x0 . As 18

x0 −x0 +x

advocated by Chernov, Gallant, Ghysels, and Tauchen (2003), we set x0 = ln(1.5). 20 Note that these parameters are annualized. We assume there are 250 trading days in a year.

23

Jt is a symmetric tempered stable process with L´evy measure: e−λx dx, (88) x1+β where c > 0, λ > 0, and β ∈ [0, 2) measures the degree of jump activity. We assume λ = 3 and ν(dx) = c

β = 0.5. The choice of β produces an infinite-activity, finite-variation process and is consistent with Theorem 4.1. The idea is to subdue Xt to a stream of small jumps that, in contrast to the large ones, are typically difficult to filter via truncation, and which can be confused by the t-statistic with stochastic volatility. We therefore anticipate that this setup induces some size distortions in the test. c is calibrated so that Jt accounts for 20% of the quadratic variation. This parameterization aligns with other papers (e.g., A¨ıt-Sahalia, Jacod, and Li, 2012; A¨ıt-Sahalia and Xiu, 2015; Huang and Tauchen, 2005). We approximate the continuous time representation of σsv,t using an Euler scheme, while Jt is generated as the difference between two spectrally positive tempered stable processes, which are simulated using the acceptance-rejection algorithm of Baeumer and Meerschaert (2010), as described in Todorov, Tauchen, and Grynkiv (2014).21 Note that the latter is exact, if β < 1, as is the case here. We simulate data for t ∈ [0, 1] (this is thought of as corresponding to a trading session on a US stock exchange, which spans 6.5 hours), where the discretization step is ∆t = 1/23, 400 (i.e., time runs on a one second grid). In Figure 2, we provide an illustration of a realization of the volatility and jump process from the full model. A total of T = 1, 000 Monte Carlo replica is generated. In each simulation, we pollute the efficient price with an additive noise term by setting Yi/n = Xi/n + i/n . To capture the wellknown negative serial correlation in log-returns induced by bid-ask bounce in transaction prices and apparent second-order effects present in our real data (cf. Panel A in Figure 5), we follow Kalnina (2011) and model i/n (for a given observation frequency n) as an MA(1):   ω2 i.i.d. 0 0 0 , i/n = i/n + ϕ(i−1)/n , where i/n | (σt )t∈[0,1] ∼ N 0, 1 + ϕ2

(89)

so that var() = ω 2 . To gauge how the strength of autocorrelation in  affects our results, we consider ϕ = 0, −0.3, −0.5, and −0.9. Of course, theq first value corresponds to the i.i.d. noise case. To model the R1 4 magnitude of , we set ω 2 = ξ 2 σ dt, such that the variance of the market microstructure 0 t component scales with volatility (e.g., Bandi and Russell, 2006; Kalnina and Linton, 2008). As in Barndorff-Nielsen, Hansen, Lunde, and Shephard (2008), we fix ξ 2 = 0.0001, 0.001 and 0.01, as motivated by the empirical work of Hansen and Lunde (2006), who find these to be typical sizes of noise contamination for the 30 stocks in the Dow Jones Industrial Average index (see also, e.g., 21

We thank Viktor Todorov for sharing Matlab code to simulate a tempered stable process.

24

Figure 2: Illustration of a simulation. Panel A: volatility.

Panel B: jump process.

2

0.5 σt = σsv,t σu,t

dJ

t

σ

sv,t

1.8

σu,t

0

1.6

jump size (in percent)

−0.5 spot volatility

1.4

1.2

1

−1

−1.5 0.8 −2 0.6

0.4

0

0.1

0.2

0.3

0.4

0.5 time

0.6

0.7

0.8

0.9

−2.5

1

0

0.1

0.2

0.3

0.4

0.5 time

0.6

0.7

0.8

0.9

1

Note. The figure shows a sample path of the two main ingredients in the jump-diffusion model (from the first simulation of 1,000 replica in total). In Panel A, the volatility is measured relative to its unconditional average. In Panel B, as barely noticeable, the tempered stable jump process has many small increments that are close, but not equal, to zero.

A¨ıt-Sahalia and Yu, 2009). d d at sampling frequency n = 390, 780, 1560, 4680, 7800, − Y(i−1)/n We construct ∆ni Y d ≡ Yi/n

11700 and 23400, thereby varying the sample size across a broad range of selections. With the above interpretation of time, the smallest (largest) value of n amounts to observing a new price every minute (second). Such a number of trade arrivals is not unrealistic compared to real highfrequency data, as reported in Section 6. The estimator σ ˆu,t in (31) is applied to deflate ∆ni Y , as in (38).22 We compute σ ˆu,t based on the T simulations with the associated value of n and m = 78 (i.e., 5-minute data).23 We bias-correct using ω ˆ 2 from (39) with q = 3 (even if the noise is at most 1-dependent here). This matches the empirical implementation, where this value of q suffices to capture the autocorrelation found in our real high-frequency data, as evident from Panel A in Figure 5. In Figure 3, we plot σ ˆu,t alongside σu,t as estimated both under H0 and Ha . As seen, σ ˆu,t is roughly unbiased, but it exhibits higher 22

In unreported results (available upon request), we found that replacing σu,t with σ ˆu,t does not modify the properties of the t-statistic much. In particular, the rejection rates are only marginally higher with the latter. An increase was to be expected due to the inherent sampling error in the estimator. This means σ ˆu,t does not completely control for the true seasonal pattern—to the extend it deviates from σu,t —leaving ∆ni Y d modestly heteroskedastic even under H0 . The effect is rather benign, which is remarkable given our naive configuration of σ ˆu,t with a relatively small choice of m even for larger sample sizes n. 23 The noise plays a limited role at this sampling frequency, but we prefer a low value of m to accommodate all the choices of n in the simulation design. In our empirical work, we exploit 30-second data (i.e., m = 780), as the instruments analyzed there are very liquid and, as a result, sample sizes are generally large.

25

Figure 3: Estimation of the diurnal component. Panel A: H0 : σt = σσu,t .

Panel B: Ha : σt = σsv,t σu,t .

1.6

1.6 point estimate true value

1.5

1.5

1.4

1.4

1.3

1.3

σu,t

σu,t

point estimate true value

1.2

1.2

1.1

1.1

1

1

0.9

0.9

0.8

0

0.1

0.2

0.3

0.4

0.5 time

0.6

0.7

0.8

0.9

0.8

1

0

0.1

0.2

0.3

0.4

0.5 time

0.6

0.7

0.8

0.9

1

Note. We plot σu,t from (85). An estimate is recovered from σ ˆu,t proposed in (31) based on m = 78. The bias-correction is done via ω ˆ 2 in (39) with q = 3. Panel A depicts estimates under H0 : σt = σσu,t , while Panel B is for Ha : σt = σsv,t σu,t . Each point on a vertical line corresponds to an estimate in that time interval for some combination of n, ξ 2 and ϕ.

variation around σu,t in the presence of stochastic volatility, which adds measurement error to the calculations.24 √ We pre-average using (17) and (37), which we do locally on a window of size kn = [θ n]. We work with θ = 1/3 and θ = 1 (as also done in, e.g., Christensen, Kinnebrock, and Podolskij, 2010).25 As standard, the weight function is g(x) = min(x, 1 − x). ˇ (l, r)n with l1 = r1 = 2 and l2 = r2 = 1. To truncate, The t-statistic is based on comparing BV we set vn = cu$ n with un = kn /n in (24), which is adapted to an estimate of the local spot volatility. As in, e.g., Li, Todorov, and Tauchen (2013, 2016), we fix the “rate” parameter $ = 0.49, while p we determine the “scale” dynamically as c = Φ(0.999) BV (1, 1)n , where Φ(0.999) is the 99.9%quantile from the standard normal distribution and BV (1, 1)n is the non-truncated estimator in (20). The intuition behind this construction is as follows. Assume there are no jumps in the interval [i/n, (i + kn )/n]. Then, under mild regularity conditions:   1 1/4 n ¯ d a 2 2 n ∆i Y ∼ N 0, θψ2 σsv,i/n + ψ1 ω . (90) θ p  p R1 2 It follows from (21) that BV (1, 1)n → 0 θψ2 σsv,s + 1θ ψ1 ω 2 ds, so that BV (1, 1)n is a (jump24

To illustrate the robustness of our approach, we do not filter σ ˆu,t , although it appears natural to exploit the smoothness condition in Assumption (D3) to reduce sampling errors further. √ 25 As this introduces a small rounding effect in the relation between θ and kn , we therefore reset θ = kn / n following the determination of kn . We apply this “effective” θ in all the subsequent computations, as also advocated in Jacod, Li, Mykland, Podolskij, and Vetter (2009).

26

robust) measure of the average dispersion (i.e., standard deviation) of the sequence ∆ni Y¯ d , while Φ(·) controls how far out in the tails of the distribution truncation is enforced.26 On the other  hand, while ∆ni Y¯ d is of order Op n−1/4 , $ ∈ (0, 1/2) implies that u$ n shrinks at a slower pace n¯d than ∆i Y . Therefore, purely “continuous” returns fall within the boundary of the threshold asymptotically. In contrast, if there are jumps in [i/n, (i + kn )/n], ∆n Y¯ d usually has order Op (1), i

and such “discontinuous” returns are, eventually, discarded. The bootstrap inference is done as follows. We resample the pre-averaged high-frequency data B = 999 times for each Monte Carlo replication. Application of our bootstrap also requires the selection of the external random variable u. This is an important choice in practice, and consistent with previous work (e.g., Hounyo, 2017; Hounyo, Gon¸calves, and Meddahi, 2017) we examine the robustness of our approach by adopting two candidate distributions:27 (1.) uj ∼ N (0, 1/2). (2.)

 √ √ ! 1 5 5+1 1 −   √ , with probability p = √   2 2 2 5 ! uj = √ √   1 1 + 5 5−1   , with probability 1 − p = √ . √ 2 2 2 5

(91)

In both cases, E ∗ (uj ) = 0 and var∗ (uj ) = 1/2, so these are asymptotically valid choices of uj for the purpose of constructing a bootstrap test based on studentized and unstudentized statistics. The two-point distribution in (2.) was originally proposed by Mammen (1993), and here we just scale it such that its variance is a half. Estimation of the asymptotic variance-covariance matrix Σ depends on the block size bn = O(nδ ) with 1/2 < δ < 2/3. Of course, this means nothing other than eventually bn = cnδ , for some constant c. There is no available theory, which can help us find optimal choices of c and δ (e.g., via a MSE criterion). Moreover, in finite samples any fixed block size bn can be achieved from many combinations of c and δ. Set against this upshot, we propose the following heuristic. We fix δ = 2/3 at the upper bound (again, the inequality constraint is only binding in the limit). We set bmin = [2nδ ] and bmax = [min(3nδ , Nn /2)]. The first choice is motivated, since we need at least n n n bn ≥ 2kn for the estimator to capture the dependence in (ˇ y (l, r)ni )N i=1 , while the latter amounts to

max saying bn should also not be too large compared to Nn . We partition [bmin n , bn ] into 30 equidistant

subintervals and loop bn over the integers that are closest to the endpoints. We then select a value 26

While the pre-averaged (1,1)-bipower variation is robust to the presence of jumps in the p-lim, as n → ∞, in practice it tends to be slightly upward biased for a finite value of n, because the jumps are not completely eliminated, see, for example, Christensen, Oomen, and Podolskij (2014). 27 We also experimented with a third external random variable using an alternative formulation of the two-point distribution, where uj = ±1 with probability p = 1 − p = 1/2. The outcome was more or less identical to the results we report based on (2.), so we decided to exclude these results to save space.

27

of bn by using the minimum variance criterion of Politis, Romano, and Wolf (1999) with a two-sided averaging window of length d = 2. In Table 1 – 2, we report the rejection rates—averaged across simulations—of the above jumpand noise-robust test of H0 at the 5% level of significance. The critical value in each test is found either via the 95% quantile of the standard normal distribution function (labeled CLT), as motivated by the asymptotic theory in Corollary 4.1, or with the help of the bootstrap-based percentile and percentile-t approach—with the headings zwb· and twb· —for the two external random variates u introduced above. Throughout, we highlight the setting with ϕ = −0.5, while noting the simulated size and power for other values of ϕ are generally within ±1%-point of the numbers reported here (the latter are omitted, but available at request). This is also true for the noise variance parameter, ξ, which changes the results in a limited way, if at all, as gauged by inspection of Panel A – C in each table. As such, neither of the parameters associated to noise has a material effect on the outcome of the t-statistic, illustrating its robustness to market frictions. The only exception to this rule is for θ = 1/3 and ξ 2 = 0.01, where a visible drop in the rejection rate under Ha is noticed, suggesting that narrow pre-averaging is inadequate to counter the microstructure contamination in the presence of an (abnormally) large noise. As expected, the block size bn increases monotonically with n. On the other hand, by comparing the two tables we note the pre-averaging window itself, via θ, has a more significant impact on the test, though mostly in small samples. We comment further on that below. Turning to the analysis of the rejection rates under H0 of deterministic volatility (size), the tables show the test is oversized. In particular, the CLT-based approach has a pronounced distortion in finite samples, starting at about 25.3% (20.5%) for θ = 1/3 (θ = 1). This is more than five (four) times larger than the nominal level. These rates improve and decline towards 5% as n increases, but remain notably elevated even in fairly large samples. In contrast, the bootstrap-based approaches are much less biased relative to inference with the asymptotic critical value. The refinement brought about by bootstrapping is often substantial, when the sample size is limited, and the rejection rates are closer to the significance level across the board, albeit they are also mildly inflated initially. The percentile approach appears to possess better size properties compared to the percentile-t, and it settles around 5 – 6% fast. As noted above, the former procedure has the added advantage that it does not require the user to input a—potentially imprecise—estimate of Σ. This also helps to make it slightly less computationally intensive, so as a practical choice we advocate the percentile approach. It is interesting to see that the difference between the two external random variables, in terms of controlling size, is negligible, perhaps with a very weak preference for the one based on the discrete two-point distribution. In the empirical application below, we base our investigation on zwb2 . Next, we look at the simulation results under Ha with stochastic volatility (power). The power exhibited by the various tests is not overwhelming for small n, but it improves steadily towards 28

29

= 0.0100 24.7 6.6 18.0 3.6 15.0 5.1 12.3 6.2 10.6 5.4 9.4 6.1 9.9 6.4

Panel C: ξ 2 n = 390 780 1560 4680 7800 11700 23400

6.7 10.1 10.6 136 131 130 131 131 4.0 5.8 6.0 216 211 210 211 211 5.1 7.7 7.8 343 335 334 334 335 6.4 7.8 8.0 706 702 699 699 697 4.8 6.5 6.4 999 978 989 979 985 6.3 7.0 6.9 1,305 1,287 1,291 1,285 1,293 6.3 8.4 8.1 2,050 2,033 2,051 2,026 2,041

6.6 10.7 11.1 136 131 131 132 131 5.7 8.0 8.9 215 211 211 211 210 4.5 6.3 6.2 343 336 336 335 335 5.9 7.1 7.4 705 699 700 702 696 6.8 8.3 8.1 998 982 985 983 981 6.8 8.3 7.8 1,311 1,285 1,292 1,287 1,289 6.5 7.8 8.4 2,057 2,039 2,041 2,049 2,038

7.3 10.1 10.1 136 130 131 130 131 5.5 8.0 8.3 215 210 211 212 211 4.9 6.4 6.3 342 335 335 334 334 5.9 6.5 7.2 712 698 701 700 700 7.4 9.1 9.2 999 978 984 984 981 6.7 8.4 8.2 1,297 1,284 1,283 1,283 1,295 6.9 8.0 7.8 2,061 2,052 2,049 2,047 2,051

twb2

49.0 52.3 50.2 61.9 65.9 69.8 75.5

61.6 63.7 64.6 70.6 75.0 79.5 84.5

62.0 64.2 66.7 71.5 76.4 79.5 85.7

20.6 26.3 31.3 48.4 57.7 62.4 69.4

33.3 41.3 48.5 60.9 66.6 72.8 81.0

34.9 43.2 50.5 61.6 68.3 73.3 81.2

21.1 25.1 30.7 48.1 57.0 61.4 69.1

33.1 39.9 48.1 60.2 66.6 72.8 80.6

35.7 41.9 50.1 61.9 68.2 72.8 82.1

26.3 31.3 35.2 49.7 58.0 60.7 69.9

39.2 45.6 51.7 60.5 66.9 71.8 80.5

42.1 46.2 52.5 62.3 68.4 73.5 82.1

twb2

26.7 135 131 131 131 131 31.6 216 210 211 209 212 35.0 341 339 337 337 336 49.6 707 702 699 703 700 58.1 997 982 981 984 980 61.0 1,301 1,293 1,291 1,279 1,286 70.3 2,058 2,032 2,036 2,038 2,041

40.2 135 131 131 131 131 44.9 216 211 212 211 210 51.6 341 336 337 334 336 60.3 708 702 702 699 702 66.2 998 984 983 978 973 72.5 1,306 1,283 1,293 1,285 1,295 81.2 2,056 2,040 2,054 2,037 2,044

41.1 136 131 131 132 131 47.0 215 210 211 210 208 52.4 341 335 334 335 335 62.1 707 700 698 699 700 68.6 995 979 978 978 982 73.2 1,308 1,291 1,293 1,288 1,284 82.3 2,062 2,043 2,042 2,042 2,049

Ha : stochastic volatility power avg. block length CLT zwb1 zwb2 twb1 twb2 CLT zwb1 zwb2 twb1

Note. We simulate from a model with drift, volatility, infinite-activity jumps and microstructure noise. We test the hypothesis that σt = σσu,t is a deterministic function of time (induced by diurnal variation) and report rejection rates both under H0 (size) and H √a (power). In the latter, σt = σsv,t σu,t is also time-varying due to a two-factor SV structure. θ is a tuning parameter that is used to compute the pre-averaging window kn = [θ n], ϕ is the MA(1) coefficient in the noise process, n is the sample size, and ξ 2 controls the magnitude of noise relative to volatility. CLT is for the asymptotic theory from (66), while zwb· and twb· are rejection rates based on the percentile and percentile-t bootstrap test for two choices of the external random variable u. We made 1,000 Monte Carlo trials with 999 bootstrap replica in each simulation. Further details can be found in Section 5.

= 0.0010 24.6 7.0 19.0 5.4 14.6 4.8 12.3 5.8 11.8 6.6 11.1 6.4 10.1 6.3

CLT zwb1 = 0.0001 25.3 6.9 18.5 5.6 13.5 5.1 12.0 5.7 11.7 7.5 11.2 6.4 10.7 7.1

Panel B: ξ 2 n = 390 780 1560 4680 7800 11700 23400

Panel A: n = 390 780 1560 4680 7800 11700 23400

ξ2

H0 : deterministic volatility size avg. block length zwb2 twb1 twb2 CLT zwb1 zwb2 twb1

Table 1: Rejection rate at 5% level of significance with θ = 0.333 and ϕ = −0.5.

30

= 0.0100 20.6 8.5 14.4 6.0 14.3 6.8 10.0 6.3 9.9 6.2 10.1 5.8 9.9 7.0

Panel C: ξ 2 n = 390 780 1560 4680 7800 11700 23400

8.2 12.2 12.2 135 130 130 131 130 5.8 7.9 8.4 215 210 211 210 211 7.0 8.8 8.5 342 335 337 334 334 6.0 7.4 7.7 708 698 700 699 695 6.0 7.2 7.2 996 983 984 986 983 6.3 7.2 7.2 1,296 1,287 1,286 1,284 1,288 7.3 8.3 8.3 2,061 2,040 2,045 2,047 2,039

8.5 11.3 11.3 135 130 130 131 130 5.3 7.4 8.1 215 209 210 209 211 6.3 8.3 8.7 339 335 337 336 334 6.1 8.1 8.0 708 700 700 696 699 5.7 6.9 7.3 999 984 987 983 985 6.3 7.2 7.6 1,295 1,288 1,291 1,287 1,285 7.6 7.8 7.8 2,056 2,044 2,057 2,048 2,047

8.6 11.6 11.9 135 131 130 131 130 5.0 7.4 7.5 215 211 210 210 209 6.5 7.9 8.1 341 335 337 335 334 6.1 8.3 8.1 710 703 698 700 699 5.8 7.0 7.4 996 983 985 982 975 7.0 7.7 7.9 1,292 1,295 1,296 1,294 1,291 8.0 8.0 7.8 2,066 2,053 2,045 2,064 2,038

twb2

49.2 41.7 47.4 53.1 57.1 61.1 67.1

49.9 44.9 48.1 53.2 60.2 63.2 68.6

49.1 42.8 48.4 54.1 60.3 64.2 69.0

25.4 24.2 33.2 42.6 47.3 52.7 60.2

24.5 25.6 34.6 44.2 50.5 55.8 63.6

24.0 24.9 35.3 44.3 51.4 55.6 63.6

25.3 24.6 33.2 41.7 45.9 52.5 61.5

24.3 25.3 34.6 44.2 49.9 54.9 63.4

23.7 25.1 34.8 44.2 50.6 56.1 63.9

29.3 27.8 34.4 43.8 47.0 52.8 60.5

28.3 28.8 37.3 44.4 49.6 55.2 62.6

28.7 27.6 37.0 44.5 51.1 56.2 63.0

twb2

29.3 136 131 130 131 130 27.3 216 210 210 210 211 34.5 340 336 334 335 333 43.8 708 698 702 693 701 46.9 994 976 980 977 982 52.8 1,297 1,284 1,287 1,287 1,279 60.6 2,069 2,039 2,042 2,050 2,058

29.1 135 131 130 131 130 29.2 215 211 211 211 210 36.9 341 335 333 335 334 44.7 706 700 699 700 702 50.0 996 979 991 975 983 54.6 1,302 1,293 1,291 1,285 1,291 62.6 2,074 2,056 2,052 2,056 2,047

28.4 136 131 131 132 130 27.3 216 210 209 209 211 36.9 342 334 335 336 334 44.4 710 698 700 694 695 51.8 990 981 979 981 976 55.7 1,302 1,284 1,281 1,285 1,292 63.0 2,066 2,046 2,047 2,049 2,036

Ha : stochastic volatility power avg. block length CLT zwb1 zwb2 twb1 twb2 CLT zwb1 zwb2 twb1

Note. We simulate from a model with drift, volatility, infinite-activity jumps and microstructure noise. We test the hypothesis that σt = σσu,t is a deterministic function of time (induced by diurnal variation) and report rejection rates both under H0 (size) and H √a (power). In the latter, σt = σsv,t σu,t is also time-varying due to a two-factor SV structure. θ is a tuning parameter that is used to compute the pre-averaging window kn = [θ n], ϕ is the MA(1) coefficient in the noise process, n is the sample size, and ξ 2 controls the magnitude of noise relative to volatility. CLT is for the asymptotic theory from (66), while zwb· and twb· are rejection rates based on the percentile and percentile-t bootstrap test for two choices of the external random variable u. We made 1,000 Monte Carlo trials with 999 bootstrap replica in each simulation. Further details can be found in Section 5.

= 0.0010 21.0 9.6 14.4 5.1 14.1 6.4 11.5 7.2 10.2 5.6 10.7 5.8 9.4 7.1

CLT zwb1 = 0.0001 20.5 8.9 14.4 5.2 14.2 6.4 11.7 6.9 10.9 6.5 10.4 6.2 10.5 7.8

Panel B: ξ 2 n = 390 780 1560 4680 7800 11700 23400

Panel A: n = 390 780 1560 4680 7800 11700 23400

ξ2

H0 : deterministic volatility size avg. block length zwb2 twb1 twb2 CLT zwb1 zwb2 twb1

Table 2: Rejection rate at 5% level of significance with θ = 1.000 and ϕ = −0.5.

Figure 4: Simulation properties of t-statistic and H-index. ˆ Panel B: relative bias of H-index.

Panel A: assessment of power. 1.4

1

θ = 1/3 θ=1 sample average

0.9

1.2 0.8

1

0.6

0.8 density

outcome of t−statistic

0.7

0.5

0.6 0.4

0.3

0.4

0.2

0.2 0.1

0

θ = 1/3 θ=1 0

0.1

0.2

0.3

0.4

0 −0.5

0.5

H−index

0

0.5

1

1.5 relative bias

2

2.5

3

Note. In Panel A, we create an indicator variable I, which equals to one if the t-statistic (based on zwb2 ) is significant at the 5% nominal level, zero otherwise. We plot it against the true H-index from (92) (small symbol). The curve is from a logistic regression between the two. Also shown are local averages of I (large symbol). In Panel B, we plot the distribution of b H-index—defined in (93)—scaled by the H-index. Throughout, the setting is n = 23, 400, ξ 2 = 0.001 and ϕ = −0.5.

100% as n grows large. Still, it stays somewhat less than unity even for n = 23, 400. It appears the CLT-based test has good power, but this is largely due to the sheer amount of Type I errors committed with this statistic. We observe a notable drop in the rejection rates going from Table 1 (with θ = 1/3) vis-`a-vis to Table 2 (with θ = 1), caused by the heavy increase in pre-averaging. While this generally renders our testing procedures more resilient to the detrimental effects of microstructure noise, it also smooths out the underlying volatility path and thereby diminishes our ability to uncover genuine heteroskedasticity in the data. It thus highlights a crucial trade-off in practice in terms of selecting θ. The above suggests our test is not always powerful enough to pick up variation in σsv,t . There are several possible explanations of this finding. Firstly, the choice of the tuning parameter θ has a significant impact, as we highlighted above and inspect further below. Secondly, the problem is not trivial. It may just be hard to detect fluctuations in σsv,t from noisy high-frequency data, leaving the jump distortion aside. Thirdly, and although σsv,t is time-varying under the alternative, it may be so persistent that its sample path—which of course differs between simulations—moves about so little (in relative terms) that it appears essentially homoskedastic on an intraday time frame. While this feature partially justifies regarding σsv,t as “almost constant,” it also makes it hard for the test to discriminate Ha from natural sampling variation under H0 (which it rightfully should do here), at least for the simulated sample sizes.

31

To shed light on these aspects, we compute: R1 H-index = 1 −

R01 0

2 ds σsv,s

2

4 ds σsv,s

.

(92)

The H-index compares the square of integrated variance to the integrated quarticity of X d . It has the intuitive interpretation that it describes how much σsv,t deviates from H0 in percent, see, e.g., Podolskij and Wasmuth (2013).28 We note that H-index ∈ [0, 1] by construction and that it equals zero if and only if σsv,t = σ is constant. Strictly positive values imply σsv,t is to some extent time-varying (not necessarily random, though). The H-index is therefore a natural measure of heteroskedasticity in our framework.29 The two-factor stochastic volatility process used in this paper has an average H-index value of about 0.20 (based on a large number of paths drawn from the model). It falls below 0.10 20% of the times, while it is rarely smaller than 0.05. In Panel A of Figure 4, we report the outcome of the t-statistic both for the set of experiments with deterministic and stochastic volatility. We define an indicator variable I, which takes the value one if H0 was rejected (on the basis of zwb2 ), and zero otherwise. The figure is a scatter plot of I versus the H-index. The fitted line originates from a logit regression of I on the H-index, which can be interpreted as the power of the test, conditional on H-index. As expected, the tendency to discard H0 is an increasing function of the H-index. When the deviation from the null is 0.15, the t-statistic is significant about 80% of the time for θ = 1/3, which falls down to 50% for θ = 1. Meanwhile, an H-index above 0.35 (θ = 1/3) – 0.45 (θ = 1) implies it more or less always lies in the rejection region. It thus requires rather convincing evidence against the null to firmly reject it, more so for θ = 1. In practice, we estimate the H-index based on c2 IV ˆ H-index = 1 − , c IQ

(93)

where n c = 1 BV ˇ (1, 1)n − ψ1 ρˆ2 , IV θψ2n θ2 ψ2n

n 1 c= ˇ (2, 2)n − 2 ψ1 ρˆ2 IV c− IQ BV (θψ2n )2 θ2 ψ2n



ψ1n θ2 ψ2n

2

ρˆ4 ,

(94)

and 2

ρˆ = ρˆ(0) + 2

q X

ρˆ(k)

(95)

k=1 28

The statistic has been used in earlier work to test for the parametric form of volatility (e.g., Dette, Podolskij, and Vetter, 2006; Vetter and Dette, 2012). In contrast to our paper, the former operate with continuous X. Moreover, the ratio appears—sometimes in a different format—in other strands of the literature, for example asymptotic variance reduction (e.g., Clinet and Potiron, 2017), estimation of integrated variation (e.g., Andersen, Dobrev, and Schaumburg, 2014; Barndorff-Nielsen, Hansen, Lunde, and Shephard, 2008; Xiu, 2010), or in the context of jumptesting (e.g., Barndorff-Nielsen and Shephard, 2006; Kolokolov and Ren`o, 2016). 29 While it is possible to base the t-statistic on the H-index by transforming the CLT in (66) via the delta rule, we refrain from doing so due to the severe amount of time it takes to run the code.

32

ˇ (l, r)n in the presence of q-dependent measurement error (e.g., estimates the asymptotic bias in BV Hautsch and Podolskij, 2013, Lemma 2), q−k+1

ρˆ(k) = −

X

jˆ γ (k + j),

n−k 1 X n d n γˆ (k) = ∆ Y ∆i+k Y d , n − k i=1 i

and

j=1

(96)

for k = 0, . . . , q + 1. b In Panel B of Figure 4, we plot the distribution of the relative bias H-index/H-index as a function of θ across simulations under the alternative. Note that an unbiased estimator has the distribution b centered at one. As apparent, H-index is slightly downward biased both for θ = 1/3 and θ = 1, which is mainly caused by a modest underestimation of IQ.30 This leads to conservative statements about the true level of heteroskedasticity in the data, thereby reducing the rejection rate. The distribution is more dispersed and has a higher probability of being close to zero or even outright negative, when θ = 1. This, in part, can help to explain why the simulated power is smaller for θ = 1. Overall, our noise- and jump-robust test of heteroskedasticity in diurnally-corrected diffusive volatility implemented via the bootstrap percentile-approach has good properties. In contrast to the CLT-based version of the test, it is almost unbiased, also for very small values of n, while it has decent—albeit not perfect—power under the presence of stochastic volatility.

6

Empirical application

We apply our framework to a large cross-sectional panel of US equity high-frequency data. It includes the 30 stocks of the Dow Jones Industrial Average—following the update of its constituent list on March 18, 2015—and the SPDR S&P 500 trust. The latter is an ETF with a price of about 1/10 the cash market value of the S&P 500 index. The sample period is January 4, 2010 through December 31, 2013 for a total of T = 1, 006 official exchange trading days. Table 3 presents a list of ticker symbols along with a few summary statistics.31 The data were extracted from the TAQ database and comprise a complete transaction record for each stock. They were cleaned with the algorithm developed by Christensen, Oomen, and Podolskij (2014), who build on earlier work of Brownless and Gallo (2006) and Barndorff-Nielsen, Hansen, Lunde, and Shephard (2009). It is a standard way of preparing data for analysis in the high-frequency volatility literature. To compute σ ˆu,t , we create an equidistant log-price series for each asset pre-ticked to a 5-second resolution, i.e. n = 4, 680. We then set m = 780—or n/m = 6—to retrieve a local estimate σ ˆu,t that covers a 30-second interval.32 It ensures that we recover a detailed view of the diurnality in 30

c There is also an attenuation effect induced by the non-linear transformation of IV. 2 ˆ Notice that the variance of the noise, as captured by ξ , is generally smaller than what we assumed in the simulations. This is consistent with the notion that the noise has decreased over time. 32 To estimate σu,t , we further delete a few outliers from the sample. Firstly, the Flash Crash of May 6, 2010. 31

33

34

n 14,576 6,843 6,486 8,510 11,884 8,568 6,521 7,422 12,990 7,940 7,798 7,193 12,515 8,766 12,141 7,991 7,192 5,466 8,455 12,814 4,847 11,253 7,967 4,779 6,730 5,938 6,213 9,191 8,085 10,693 18,154

σ ˆ 18.9 19.5 18.7 21.7 16.5 16.1 18.4 17.1 16.5 22.2 17.0 13.9 17.3 11.3 21.9 12.2 12.2 14.9 14.8 16.2 18.0 14.6 11.9 14.9 20.6 16.1 19.3 13.4 12.4 14.7 10.4

ρˆ1 -0.12 -0.07 -0.10 -0.09 -0.38 -0.05 -0.11 -0.14 -0.41 -0.10 -0.16 -0.14 -0.37 -0.18 -0.14 -0.21 -0.13 -0.08 -0.25 -0.34 -0.07 -0.38 -0.16 -0.11 -0.09 -0.09 -0.12 -0.29 -0.17 -0.07 -0.05

kn 40 27 27 30 36 31 27 29 38 29 29 28 37 31 37 30 28 24 31 38 23 35 30 23 27 25 26 32 30 34 45

zwb2 0.50 0.39 0.44 0.34 0.88 0.31 0.41 0.52 0.89 0.42 0.54 0.37 0.87 0.59 0.58 0.61 0.43 0.37 0.69 0.87 0.41 0.90 0.53 0.42 0.59 0.42 0.48 0.78 0.60 0.38 0.56

> 0.11 0.12 0.15 0.07 0.76 0.07 0.14 0.24 0.82 0.09 0.27 0.11 0.74 0.37 0.22 0.36 0.16 0.10 0.48 0.74 0.10 0.81 0.30 0.16 0.26 0.13 0.18 0.56 0.33 0.15 0.32

d ˆ ˆ H-index H-index 0.23 0.12 0.23 0.15 0.27 0.17 0.22 0.11 0.41 0.35 0.20 0.12 0.24 0.15 0.26 0.18 0.42 0.36 0.24 0.14 0.28 0.19 0.23 0.15 0.39 0.32 0.28 0.21 0.26 0.16 0.30 0.22 0.26 0.16 0.24 0.15 0.33 0.26 0.37 0.31 0.27 0.16 0.43 0.38 0.27 0.20 0.26 0.17 0.31 0.19 0.26 0.16 0.28 0.19 0.36 0.28 0.29 0.21 0.19 0.14 0.21 0.16

θ = 1/3 p∗1−α/T d zwb2 kn 120 82 80 91 108 92 80 86 114 88 88 84 112 93 110 89 84 73 92 113 69 106 89 68 81 76 78 95 89 103 135

zwb2 0.17 0.12 0.19 0.15 0.42 0.10 0.14 0.14 0.44 0.16 0.17 0.13 0.37 0.17 0.18 0.21 0.17 0.15 0.24 0.38 0.18 0.48 0.15 0.17 0.22 0.15 0.19 0.28 0.17 0.11 0.13

#Z n >

θ=1

0.01 0.02 0.03 0.00 0.22 0.01 0.02 0.02 0.25 0.01 0.03 0.01 0.14 0.03 0.02 0.04 0.03 0.02 0.07 0.13 0.01 0.25 0.03 0.03 0.04 0.02 0.03 0.06 0.03 0.01 0.02

p∗1−α/T d zwb2

d ˆ ˆ H-index H-index 0.18 0.06 0.17 0.09 0.23 0.12 0.18 0.05 0.29 0.21 0.15 0.07 0.18 0.09 0.18 0.09 0.30 0.22 0.19 0.08 0.21 0.11 0.18 0.10 0.29 0.18 0.20 0.12 0.19 0.08 0.21 0.12 0.20 0.11 0.19 0.10 0.23 0.14 0.27 0.18 0.22 0.11 0.31 0.23 0.19 0.12 0.19 0.12 0.25 0.12 0.21 0.11 0.24 0.14 0.25 0.16 0.21 0.11 0.14 0.07 0.14 0.08

Note. This table reports descriptive statistics for our TAQ high-frequency data computed daily q and averaged over time. The sample covers January 4, 2010 through December 31, 2013 c is an annualized jump-robust measure of volatility based on (94) (with θ = 1/3), for a total of T = 1, 006 days. n is the number of transaction data available after filtering, σ b = 256 × IV n 2 4 ˆ ρˆ1 is the first-order autocorrelation of ∆i Y , and ξ is the noise level ×10 . kn is the pre-averaging window, while #Z n > p∗1−α/T is the fraction of t-statistics for testing H0 larger than ˆ the (1 − α/T )-quantile of the bootstrap distribution of Z n∗ (based on zwb2 ) with α = 0.05. H-index is the heteroscedasticity measure defined in (93). The latter three are computed for d both θ = 1/3 and θ = 1. A superscript d refers to the average value of a statistic based on ∆n ˆu,t from (31). i Y , where the rescaling is implemented via σ

ticker AAPL AXP BA CAT CSCO CVX DD DIS GE GS HD IBM INTC JNJ JPM KO MCD MMM MRK MSFT NKE PFE PG TRV UNH UTX V VZ WMT XOM SPY

ξˆ2 × 104 0.12 0.29 0.39 0.18 1.74 0.29 0.44 0.34 2.02 0.22 0.27 0.38 1.24 0.32 0.11 0.46 0.28 0.64 0.49 0.72 0.35 2.12 0.28 0.88 0.28 0.47 0.53 0.68 0.25 0.17 0.10

#Z n

Table 3: Descriptive statistics of TAQ high-frequency data.

Figure 5: Properties of equity high-frequency data. Panel A: ACF of ∆ni Y .

Panel B: σ ˆu,t .

0.1

5.5 maximum average minimum fitted

5

4.5

4 0

σ ˆu,t

correlation

3.5

3

2.5 −0.1 2

1.5

1

estimate ± 2 s.e. −0.2

1

2

3 lag

4

0.5 9:30

5

10:00

11:00

12:00pm

13:00 time

14:00

15:00

16:00

Note. In Panel A, we plot the ACF of our TAQ high-frequency equity data averaged across assets and over time. In Panel B, we present our estimator σ ˆu,t of within-day volatility. The cross-sectional average is reported, along with the highest and smallest point estimate. As a comparison, we also fit the parametric form of diurnal variation in (85) via non-linear least squares.

volatility, while still being able to purge the associated noise with decent accuracy. On each block, we bias-correct with the robust estimator in (39) using q = 3.33 This is motivated by Panel A in Figure 5, which reports the autocorrelation function (ACF) of ∆ni Y . As shown, while the first few autocorrelations are significant, the ACF dies out fast and is generally insignificant after lag three and negligible beyond lag five (not shown in figure), so this choice of q suffices to capture the observed serial dependence in the noise. The cross-sectional average of σ ˆu,t is reported along with the minimum and maximum value in Panel B of Figure 5. We note σ ˆu,t features the reverted J-shape as reported in prior work, but also that it is very rough with sharp increases around pre-scheduled macroeconomic announcements (e.g., at 10:00am or 2:00pm). The latter is consistent with empirical findings in, e.g., Todorov and Tauchen (2011), who note that jumps in volatility are strongly correlated with large moves in market prices.34 There are also some notable spikes in within-day volatility prior to the close of the Secondly, for each 30-second interval we remove the top 1% of data for each stock—as measured by |∆m i Y |—to filter out observations typically associated with idiosyncratic news announcements. These events exert an unduly influence on the estimates due to our relatively small value of T . In a larger sample this should not be necessary. 33 2 ω is sometimes estimated to be negative, due to the sampling distribution of ω ˆ 2 . This cannot be true, of course, and we therefore truncate ω ˆ 2 at zero throughout the paper. This happens more often if the noise is really small relative to the underlying volatility of the asset, as demonstrated by the index-tracker SPY in Figure 1. 34 It also suggests that σu,t may not be as smooth as stipulated by Assumption (D3). Note, however, that although volatility peaks at the announcement, it does not necessarily jump. In our data it actually starts to increase around 1-minute to 30-seconds prior to the time, where the numbers are officially slated for release. This is consistent with the findings of Jiang, Lo, and Verdelhan (2011) from the U. S. Treasury market.

35

ˆ Figure 6: The distribution of the H-index. Panel A: before correction.

Panel B: after correction.

4

4 θ = 1/3 θ=1 sample average

3.5

3

3

2.5

2.5 density

density

3.5

2

2

1.5

1.5

1

1

0.5

0.5

0 −0.2

−0.1

0

0.1

0.2 0.3 H−index

0.4

0.5

0.6

θ = 1/3 θ=1 sample average

0.7

0 −0.2

−0.1

0

0.1

0.2 0.3 H−index

0.4

0.5

0.6

0.7

ˆ Note. We plot the cross-sectional distribution of the H-index before and after diurnal correction with our nonparametric estimator from (31), also shown in Panel B of Figure 5. The dashed vertical line at zero indicates the theoretical lower bound.

exchange. Overall, diurnal variation is remarkably constant across assets. To assess the parametric model used in the simulations, we estimate the parameters of (85) via non-linear least squares based on the cross-sectional average. The fitted equation σu,t = 0.78 + 1.71e11.85t + 0.30e14.17(1−t) is a good approximation to our nonparametric estimates, although it does not track the sharp initial decay in early trading. On this basis, we construct the deflated log-returns and compute the test for each stock and day ˇ (2, 2)n and BV ˇ (1, 1)n with θ = 1/3 and in the sample. As above, we base the investigation on BV θ = 1. The bootstrap percentile approach is applied to evaluate the significance of our t-statistic, i.e. zwb2 . We apply a standard Bonferroni-type correction to account for multiple testing and control the family-wise error rate. That is, we run each individual test at significance level α/T with α = 0.05. In the right-hand side of Table 3, we report the average rejection rate of H0 and associated Hindex measurement. As a comparison, we also retrieve the corresponding results from the raw data prior to diurnal correction. Looking at the table, we observe that H0 is discarded about half of the times for the typical stock if θ = 1/3 and there is no seasonal adjustment. The levels are lower for θ = 1, where about one-third of the tests are rejected. This is consistent with the decrease in power if θ is higher, as uncovered in the simulation section. On the other hand, there is also evidence of the caveat raised in Remark 3, i.e. the pre-averaging estimator is affected harder by residual microstructure noise if θ is small, so that here the test is prone to discredit H0 in the presence of general forms of heteroscedasticity in the variance of the noise. Indeed, there is a tendency for 36

ˆ Figure 7: Empirical properties of t-statistic and H-index. Panel B: after correction.

1

1

0.9

0.9

0.8

0.8

0.7

0.7 outcome of t−statistic

outcome of t−statistic

Panel A: before correction.

0.6

0.5

0.4

0.6

0.5

0.4

0.3

0.3

0.2

0.2

0.1

0 −0.1

0.1

θ = 1/3 θ=1 −0.05

0

0.05

0.1

0.15 H−index

0.2

0.25

0.3

0.35

0.4

0 −0.1

θ = 1/3 θ=1 −0.05

0

0.05

0.1

0.15 H−index

0.2

0.25

0.3

0.35

0.4

Note. We create an indicator variable I, which equals one if the t-statistic is significant at the 5% nominal level, zero otherwise. d d ˆ ˆ ˆ We plot I against the H-index before and after diurnal correction (i.e., based on (H-index, zwb2 ) or (H-index , zwb2 )) and as a function of θ. The curve is from a logistic regression between the two. We also show local averages of I.

stocks with very negative first-order return autocorrelation and large levels of noise—as measured by ξˆ2 — to reject more frequently. If we control for diurnal variation, the rejection rate drops to about 30% (for θ = 1/3) to almost 5% (for θ = 1). This implies the diurnal pattern is a first-order effect, which captures a large fraction of variation in intraday volatility. As readily seen, however, important and potent sources of heteroskedasticity remain present in the data. ˆ This is corroborated by the H-index, for which we also plot the cross-sectional distribution before and after diurnal correction in Figure 6. The distribution shifts to the left and displays less sampling variation after diurnal correction, while we again notice a slight increase in the dispersion by moving θ up, although the effect is weak. As measured by the cross-sectional average shown in Panel B of the figure, the strength of residual heteroskedasticity present in ∆ni Y d is broadly comparable to that of the two-factor stochastic volatility model from Section 5 for θ = 1/3, while it is somewhat less for θ = 1. At last, in Figure 7 we model the empirical rejection rate of the t-statistic (with no Bonferroni correction here to be comparable with Figure 4 and improve the fit of the logistic regression) as a ˆ function of the H-index. The logit fit is consistent with Section 5 in that it takes a relatively high ˆ reading of H-index to confidently reject H0 . Note that the rejection rates are close to each other d ˆ ˆ if H-index ' H-index , as can be gauged by comparing Panel A and B. The intuitive explanation is that the power of the test depends only on the level of heteroskedasticity in the data, which is

37

captured by the H-index, and not as such on whether one has diurnally-corrected it or not.

7

Conclusion

In this paper, we study a new approach to determine if changes in intraday spot volatility of a discretely sampled noisy jump-diffusion model can be attributed solely to a deterministic cyclical component (i.e., the so-called U- or reverse J-shape) against an alternative of further variation induced by a stochastic process. We propose to construct a test of this hypothesis from an asset return series, which has been deflated by the diurnal component and, as such, is homoskedastic under the null. The t-statistic diverges to infinity, if the deflated return series is heteroskedastic, and it has a standard normal distribution otherwise. To get a feasible test, we develop a—surprisingly robust—nonparametric estimator of unobserved diurnal volatility, which (in contrast to the test itself) is computed directly from noisy high-frequency data without pre-averaging or jump-truncation. It requires only a trivial bias-correction to eliminate the noise variance. Our estimator is consistent and has a sampling error, which is of small enough order that replacing the true diurnal factor with it does not alter the asymptotic theory. We inspect the properties of the test in a Monte Carlo simulation. We note the theory-based version has gross size distortions in the presence of infinite-activity price jumps, thus motivating a bootstrap. We validate the bootstrap and confirm it helps to improve inference by making the test almost correctly sized. The test also has acceptable power, but can fail to reject the null even in large samples, if a wide pre-averaging window is applied. The estimation of the diurnal factor has a limited impact, but it raises the rejection rate slightly. We implement our nonparametric estimator of diurnal variance and test of heteroskedasticity on real high-frequency data. The diurnal pattern explains a sizable portion of within-day variation in the volatility in practice, as inferred by the notable drop in the rejection rate of the test and the reduction in the H-index—a descriptive statistic that measures the strength of time-varying volatility—once we control for intraday seasonality. This suggests that the rescaled log-returns are often close to homoskedastic. Still, important sources of variation remain present in the data. So the answer to the title of the paper appears to be “no.” The diurnal pattern does not explain all intraday variation in volatility, but it does capture a rather significant portion of it.

38

A

The explicit form of Σ

ˇ n is consistent for the asymptotic covariance In Section 2, we show that our proposed estimator Σ  | matrix of n1/4 BV (Y d , l1 , r1 )n , BV (Y d , l2 , r2 )n , i.e. Σ appearing in (22), Theorem B.1 and Theoˇ n∗ , in Section 4. In this rem 4.1. We also prove a corresponding result for the bootstrap version, Σ short appendix, we derive an explicit expression for Σ, which was not put in the main text. We follow Podolskij and Vetter (2009a) by first defining:   hij (a, b, c) = cov |H1 |li |H2 |ri , |H3 |lj |H4 |rj , where a is a real number, b and c are a two- and four-dimensional vector. Moreover, (H1 , . . . , H4 ) follows a multivariate normal distribution with: 1. E(Hl ) = 0 and var(Hl ) = b1 a2 + b2 ω 2 , 2. H1 ⊥ H2 , H1 ⊥ H4 , and H3 ⊥ H4 , 3. cov(H1 , H3 ) = cov(H2 , H4 ) = c1 a2 + c2 omega2 and cov(H2 , H3 ) = c3 a2 + c3 ω 2 .   1 We set t = ψ1 , θψ2 and define: θ 1 1 f1 (s) = φ1 (s), f2 (s) = θφ2 (s), f3 (s) = θφ3 (s), f4 (s) = φ4 (s), θ θ for s ∈ [0, 2], where Z 1−s Z 1−s 0 0 φ1 (s) = g (u)g (u + s)du, φ2 (s) = g(u)g(u + s)du, 0

Z

0 2−s 0

Z

0

g (u)g (u + s − 1)du

φ3 (s) =

and

2−s

g(u)g(u + s − 1)du.

φ4 (s) = 0

0

We note that both f1 and f2 are 0 for s ∈ [1, 2], according to the assumptions imposed on g. We | next let f (s) = f1 (s), f2 (s), f3 (s), f4 (s) . At last, we get that   l1 ,r1 ,l2 ,r2 Σ = Σij 1≤i,j≤2

Z = 0

1

l1 ,r1 ,l2 ,r2 l1 ,r1 ,l2 ,r2 w11 w12 l1 ,r1 ,l2 ,r2 l1 ,r1 ,l2 ,r2 w21 w22

! (σsv,u )du,

where l1 ,r1 ,l2 ,r2 wij (σsv,u )

Z = 2θ 0

39

2

 hij σsv,u , t, f (s) ds.

B

Proofs

In this appendix, K denotes a generic constant, which changes from line to line. Also, as in Jacod and Protter (2012), we assume that a, σ, δ and X are bounded. As Jacod, Podolskij, and Vetter (2010) explain, this follows by a standard localization procedure, described in Jacod (2008), and does not lose generality. Formally, we derive our results under the assumption: Assumption (G):

X follows (1) with a and σ are adapted, c`adl`ag processes such that a, σ, δ

and X are bounded, so that for some constant K and nonnegative deterministic function γ˜ : Z kat (ω)k ≤ K, kσt (ω)k ≤ K, kXt (ω)k ≤ K, kδ(ω, t, x)k ≤ γ˜ (x) ≤ K, γ˜ (x)β λ(dx) ≤ K. R

Throughout the appendix, it will be convenient to define the continuous part of X by X 0 and the discontinuous martingale part by X 00 , i.e. Z t Z t 0 0 Xt = X0 + as ds + σs dWs , 0

Xt00 = Xt − Xt0 ,

(A.1)

0

where, according to the value of β, we set   a − δ1 ? ν , if β ≤ 1 s {|δ|≤1}  t . a0s = as + δ1{|δ|>1} ? ν t , if β > 1 Then, we can write Yt = Yt0 + Yt00 ,

(A.2)

ˇ (l, r)n , B(l, r)n , where Yt0 = Xt0 + t and Yt00 = Xt00 . As in the main text, if we write BV (l, r)n , BV i d n n n ˇ n ˇ ∆B(l, r) , yˇ(l, r) , B(l, r) or ∆B(l, r) , we assume they are defined with respect to Y . i

i

i

i

Proof of Theorem 2.1. Here, we more or less follow the techniques applied in the proof of Theorem 4.1 in Hounyo (2017). Firstly, we introduce the pre-averaged return ∆ni Y¯ computed on the raw unscaled high-frequency returns:  kX n −1  j n¯ ∆i Y = g ∆ni+j−1 Y, kn j=1

i = 1, . . . , n − kn + 2.

ˇ (Y, l, r)n , BV ˇ (Y 0 , l, r)n , together with the maintained Next, for the associated definitions of BV assumptions appearing in the main text and if σu,t = 1, the central limit theorem in Theorem 3 of Podolskij and Vetter (2009a) implies that, as n → ∞, n1/4

ˇ (Y 0 , l1 , r1 )n − BV (l1 , r1 ) BV ˇ (Y 0 , l2 , r2 )n − BV (l2 , r2 ) BV

! d

→s M N (0, Σ).

(A.3)

A careful inspection of the proof of this result shows that the stable convergence in (A.3) remains valid, when the pre-averaged return is given by ∆ni Y¯ d in (17) and σt = σsv,t σu,t . Indeed, the main

40

ingredient is the weak convergence: n

1/4

d ∆ni d →

 1 2 N 0, ψ1 ω , θ 

and 1/4

n

a ∆ni Y¯ 0d ∼

  1 2 2 N 0, θψ2 σsv,i/n + ψ1 ω , θ

which follows from (5), as σu,t > 0 for all t ≥ 0 and locally bounded. Thus, it suffices to prove that for any l, r > 0,  p ˇ (Y d , l, r)n − BV ˇ (Y 0d , l, r)n → n1/4 BV 0.

(A.4)

To show (A.4), we let Fu (x) = F (x)1{|x1 | 0, where F (x) = |x1 |l |x2 |r with | x = x1 x2 . As in the line of thought on page 385 in Jacod and Protter (2012), we can show √ that for wn = υn / un with un = kn /n:  2   2 − 1−2ω l+r+ 1−2ω l+r l+r l+r kxk + 1 + kxk |Fwn (x + y) − Fwn (x)| ≤ wn kyk ∧ 1 + kyk ∧ wn . | √ | √ Next, let x = ∆ni Y¯ d ∆ni+kn Y¯ d / un and y = ∆ni Y¯ 00d ∆ni+kn Y¯ 00d / un . According to (16.4.9) in Jacod and Protter (2012) in conjunction with results in part 3 in the proof of Lemma 16.4.5 in that book, for some l + r > 0:    E kxkl+r ≤ K, E kyk ∧ 1 ≤ Ku1−β/2 φn and E kyk2 ∧ wn2 ≤ Kuω(2−β) φn , n n

(A.5)

where φn → 0 as n → 0. In addition, from (A.5) and the inequality (kyk∧wn )p ≤ wnp−m (kyk∧wn )m , for 0 < m < p, it is found that  2 ω(l+r−β)− 21 (l+r−2) E kykl+r ∧ wnl+r ≤ Kwnl+r−2 E kyk ∧ wn ≤ Kun φn ,

(A.6)

where again φn → 0 as n → 0. Thus, from the above inequalities together with the definition  ˇ (Y d , l, r)n − BV ˇ (Y 0d , l, r)n n1/4 BV l+r−3

n 4 = µl µr

n−2k n +2  X

|∆ni Y¯ d |l |∆ni+kn Y¯ d |r



|∆ni Y¯ 0d |l |∆ni+kn Y¯ 0d |r



1{|∆ni Y¯ d |<υn } 1{|∆ni+k

n

Y¯ d |<υn } ,

i=1

it follows that n

l+r−3 4

1 µl µr

n−2k n +2 X

    0d l n 0d r n¯d l n d r n E |∆i Y | |∆i+kn Y¯ | − |∆i Y¯ | |∆i+kn Y¯ | 1{|∆ni Y¯ d |<υn } 1{|∆ni+k Y¯ d |<υn } n

i=1

≤ Kn

l+r−3 4

1 4



l+r

n · un2

−1/2



ω(l+r−β)− 12 (l+r−2)

un + u1−r/2 φn + un n

β−2 4

(l+r−2)−2ω(l+r−β) 4

φn





+ n φn + n φn ≤ Kn n   (β−1) (l+r−1)−2ω(l+r−β) −1/4 4 4 ≤K n +n φn + n φn . Thus, if β < 1 and

l+r−1 2(l+r−β)

 ˇ (Y d , l, r)n − BV ˇ (Y 0d , l, r)n )| → 0 and ≤ $ < 1/2, then E |n1/4 (BV 41

p ˇ (Y d , l, r)n − BV ˇ (Y 0d , l, r)n ) → therefore n1/4 (BV 0. This completes the proof of Theorem 2.1.



Next, we establish the following result (under no jumps) since it will be useful later in the proof of Theorem 4.1. Theorem B.1 Let l1 , r1 , l2 and r2 be four positive real numbers and X given by Z t Z t as ds + σs dWs . X t = X0 + 0

(A.7)

0

We define: √ ˆn = Σ where ξi ≡

(∆B(l1 , r1 )ni , ∆B(l2 , r2 )ni )| ,

n 2bn

Nn −2b Xn +1

ξi ξi| ,

(A.8)

i=1

such that

∆B(l, r)nj = B(Y d , l, r)nj+bn − B(Y d , l, r)nj ,

(A.9)

with B(l, r)nj

=n

l+r −1 4

bn 1 X y(l, r)ni−1+j . µl µr i=1

(A.10)

Furthermore, we assume (V), (A), and impose the moment condition E(|t |s ) < ∞, for some s > (3 ∧ 2(r1 + l1 ) ∧ 2(r2 + l2 )). If any li or ri is in (0, 1], we postulate (V0 ), otherwise either (V0 ) or (A0 ). In addition, suppose that kn → ∞ as n → ∞ such that (19) holds, and the block size bn fulfills (44) for some 1/2 < δ1 < 2/3. Then, as n → ∞, p ˆn → Σ Σ,

(A.11)

where Σ is defined in Appendix A. Proof of Theorem B.1. Here, recall that X follows (A.7) and note that given (A.8), we can ˆ n as follows: rewrite Σ bn 1 X n ˆn , ˆ Σ Σ = bn m=1 m

(A.12)

n −2 c √ b NbX n   n | n ˆ ˆ l1 ,r1 ,l2 ,r2 ,n Σ . Σm = ξkbn +m ξkb = ij,m n +m 2 1≤i,j≤2 k=0

(A.13)

where

p ˆn → Thus, it suffices that Σ Σ, uniformly in m. Thus, the proof is reduced to showing that m

ˆ l1 ,r1 ,l2 ,r2 ,n = Σl1 ,r1 ,l2 ,r2 , p-lim Σ ij,m ij

1 ≤ i, j ≤ 2,

n→∞

ˆ l1 ,r1 ,l2 ,r2 ,n as uniformly in m. Note that we can rewrite Σ ij,m ˆ l1 ,r1 ,l2 ,r2 ,n Σ ij,m

n −2 c √ b NbX n n = ∆B(Y d , li , ri )nkbn +m ∆B(Y d , lj , rj )nkbn +m . 2 k=0

42

(A.14)

Then, given the definition of ∆B(Y d , l, r)nm given in (A.9), by adding and subtracting appropriately, it follows that   n −2 2B(Y d , li , ri )n(k+1)bn +m B(Y d , lj , rj )n(k+1)bn +m c √ b NbX n   −B(Y d , l , r )n d n ˆ l1 ,r1 ,l2 ,r2 ,n = n Σ  i i (k+1)bn +m B(Y , lj , rj )kbn +m  ij,m 2 k=0 −B(Y d , li , ri )nkbn +m B(Y d , lj , rj )n(k+1)bn +m   B(Y d , li , ri )nm B(Y d , lj , rj )nm √ +B(Y d , li , ri )n Nn B(Y d , lj , rj )n Nn −1 b +m  −1)bn +m n (b c (b bn c ) n  b n   + d n d n  2 −B(Y , li , ri )(b Nn c−2)bn +m B(Y , lj , rj )(b Nn c−1)bn +m   bn bn d n d n −B(Y , li , ri ) Nn −1 b +m B(Y , lj , rj ) Nn −2 b +m (b bn c ) n (b bn c ) n l1 ,r1 ,l2 ,r2 ,n l1 ,r1 ,l2 ,r2 ,n = Mij,m (Y d ) + Rij,m (Y d ),

where the remainder term l1 ,r1 ,l2 ,r2 ,n Rij,m (Y d )

=



3 Op n− 2 b2n



= op (1),

uniformly in m, so long as δ1 < 3/4, where we apply the definition of B(Y d , l, r)nm in (A.10), the  Cauchy-Schwartz inequality, and the fact that E |∆ni Y¯ d |l ≤ Kn−l/4 (cf., Lemma 1 of Podolskij and Vetter, 2010). Next, we show the main term is such that l1 ,r1 ,l2 ,r2 ,n p-lim Mij,m (Y d ) = Σij ,

1 ≤ i, j ≤ 2,

(A.15)

n→∞

uniformly in m. We prove the result for the following unsymmetrized estimator: n −1 b NbX c n √ l1 ,r1 ,l2 ,r2 ,n d ˜ Mij,m (Y ) = n

k=1

B(Y d , li , ri )nkbn +m B(Y d , lj , rj )nkbn +m −B(Y d , li , ri )nkbn +m B(Y d , lj , rj )n(k−1)bn +m

! .

(A.16)

We introduce two approximations of B(Y d , l, r)njbn +m : bn X −1 1 ˜ d , l, r)njb +m = n l+r 4 B(Y y˜(l, r)ni−1+jbn +m , n µl µr i=1 bn X l+r −1 1 ¯ d , l, r)n 4 B(Y = n y˜(l, r)ni−1+(j−1)bn +m , jbn +m µl µr i=1

l r ¯ , for jbn + m ≤ i ≤ where y˜(Y d , l, r)i = ∆ni Y˜ d ∆ni+kn Y˜ d with ∆ni Y˜ d = ∆ni ¯d + σsv, jbn ∆ni W Nn (j + 1)bn + m − 1. We then show that the error due to replacing ∆ni Y¯ d by ∆ni Y˜ d is small enough to be ignored and, hence, does not affect our theoretical results. This is true, because σsv is assumed

43

to be an Itˆo semimartingale itself, so that !   Z i+j   Z i+j  kn kn X    X n n j j E ∆ni Y¯ d − ∆ni Y˜ d = E ads ds + g σsv,s −σsv, jbn dWs g Nn i+j−1 i+j−1 k k n n j=1 j=1 n n   2 !1/2    Z i+j  kn  X n kn j  ≤K + E σsv,s − σsv, jbn dWs g2 Nn i+j−1 n k n j=1 n !  1/2 kn kn bn (kn bn )1/2 ≤K + ≤K . n n n n  Note that E |B(Y d , l, r)nm | ≤ K bnn uniformly in m, and so !   (l+r) −1   1/2 4 (k b ) 1 n n ˜ d , l, r)njb +m ≤ Kbn √ E B(Y d , l, r)njbn +m − B(Y n n kn  3/2 bn . ≤K n  3/2   n bn d d n d n ¯ ¯ As for B Y , l, r jbn +m , we find that E B(Y , l, r)jbn +m − B(Y , l, r)jbn +m ≤ K . And n ˜ l1 ,r1 ,l2 ,r2 ,n (Y d ) − M ¯ l1 ,r1 ,l2 ,r2 ,n (Y d ) = op (1), uniformly in m, where because δ < 2/3, we deduce that M ij,m ij,m ¯ l1 ,r1 ,l2 ,r2 ,n (Y d ) = M ij,m



n

n −1 b NbX c  n

 n ˆn Bkb − B kbn +m , n +m

k=1

such that n n n n ¯ d , l1 , r1 )n ¯ d ˆn ¯ d ˜ d Bkb = B(Y kbn +m B(Y , l2 , r2 )kbn +m and Bkbn +m = B(Y , l1 , r1 )kbn +m B(Y , l2 , r2 )(k−1)bn +m . n +m

Then, b Nn c−1  3/2 n   √ bX bn n n n n E Bkbn +m − E Bkbn +m | F (k−1)bn +m ≤ K , n Nn k=1

b Nn c−1  3/2 n   √ bX bn n n n ˆ ˆ n E Bkbn +m − E Bkbn +m | F (k−1)bn +m ≤ K , n Nn k=1

by conditional independence, and now we are left with n −1 b NbX c n   √ l ,r ,l ,r ,n n n n d 1 1 2 2 ˆ ¯ E Bkbn +m − Bkbn +m | F (k−1)bn +m + op (1), Mij,m (Y ) = n Nn

k=1

uniformly in m. As in Podolskij and Vetter (2010) and using δ > 1/2, we note that   Z kbn Z 2  Nn  √  n bn n n ˆ nE Bkbn +m − Bkbn +m | F (k−1)bn +m = 2θ hij σsv,u , t, f (s) dsdu + o , (k−1)bn N Nn n 0 N n

44

uniformly in k and m, and thus ¯ l1 ,r1 ,l2 ,r2 ,n (Y d ) M ij,m

Z

1

Z

= 2θ 0

Z

2

 hij σsv,u , t, f (s) dsdu + op (1)

0

1

=

 l1 ,r1 ,l2 ,r2 σsv,u du + op (1), wij

0

uniformly in m, and the proof is complete.



Proof of Theorem 4.1. We prove (64) solely in model (1), which is enough, as it is the most ˇ l1 ,r1 ,l2 ,r2 ,n (Y d ), general and nests (A.7). Now, under the stated assumptions, the definitions of Σ ij l1 ,r1 ,l2 ,r2 ,n 0d ˇ Σ (Y ), and the limiting result in Theorem B.1, we deduce that, as n → ∞, ij

ˆ l1 ,r1 ,l2 ,r2 ,n (Y 0d ) = Σl1 ,r1 ,l2 ,r2 , p-lim Σ ij,m ij

for 1 ≤ i, j ≤ 2,

n→∞

uniformly in m. Thus, to get the desired result, it suffices to show that   ˇ l1 ,r1 ,l2 ,r2 ,n (Y d ) − Σ ˇ l1 ,r1 ,l2 ,r2 ,n (Y 0d ) = 0, for 1 ≤ i, j ≤ 2, p-lim Σ ij,m ij,m

(A.17)

n→∞

ˇ l1 ,r1 ,l2 ,r2 ,n (Y d ) and Σ ˇ l1 ,r1 ,l2 ,r2 ,n (Y 0d ), it holds that uniformly in m. Inserting the definition of Σ ij ij  2  ˇ l1 ,r1 ,l2 ,r2 ,n d ˇ l1 ,r1 ,l2 ,r2 ,n (Y 0d ) √ Σ (Y ) − Σ ij,m n ij,m =

n −2 b NbX c  n

d d 0d 0d ˇ ˇ ˇ ˇ ∆B(Y , li , ri )nkbn +m ∆B(Y , lj , rj )nkbn +m − ∆B(Y , li , ri )nkbn +m ∆B(Y , lj , rj )nkbn +m



k=0

c−2  bX Nn bn

=

ˇ d , li , ri )n(k+1)b +m B(Y ˇ d , lj , rj )n(k+1)b +m − B(Y ˇ 0d , li , ri )n(k+1)b +m B(Y ˇ 0d , lj , rj )n(k+1)b +m B(Y n n n n

k=0



 d n d n 0d n 0d n ˇ ˇ ˇ ˇ − B(Y , li , ri )(k+1)bn +m B(Y , lj , rj )kbn +m − B(Y , li , ri )(k+1)bn +m B(Y , lj , rj )kbn +m   d n d n 0d n 0d n ˇ ˇ ˇ ˇ − B(Y , li , ri )kbn +m B(Y , lj , rj )(k+1)bn +m − B(Y , li , ri )kbn +m B(Y , lj , rj )(k+1)bn +m   d n 0d n 0d n d n ˇ ˇ ˇ ˇ + B(Y , li , ri )kbn +m B(Y , lj , rj )kbn +m − B(Y , li , ri )kbn +m B(Y , lj , rj )kbn +m where bn X −1 1 ˇ d , l, r)n = n l+r 4 B(Y yˇ(Y d , l, r)ni−1+j . j µl µr i=1

In the following, we define: l1 ,r1 ,l2 ,r2 ,n πk,k (Y d , Y 0d ) = yˇ(Y d , li , ri )nk yˇ(Y d , lj , rj )nk0 − yˇ(Y 0d , li , ri )nk yˇ(Y 0d , lj , rj )nk0 0  = |∆nk Y¯ d |l1 |∆nk+kn Y¯ d |r1 |∆nk0 Y¯ d |l2 |∆nk0 +kn Y¯ d |r2  − |∆nk Y¯ 0d |l1 |∆nk+kn Y¯ 0d |r1 |∆nk0 Y¯ 0d |l2 |∆nk0 +kn Y¯ 0d |r2 1Ck,k0 ,

45

(A.18)



    where Ck,k0 = |∆nk Y¯ d | < υn ∩ |∆nk+kn Y¯ d | < υn ∩ |∆nk0 Y¯ 0d | < υn ∩ |∆nk0 +kn Y¯ 0d | < υn . Then, from (A.18) it follows that ˇ l1 ,r1 ,l2 ,r2 ,n (Y d ) − Σ ˇ l1 ,r1 ,l2 ,r2 ,n (Y 0d ) Σ ij,m ij,m =

l1 +r1 +l2 +r2 −6 4

n 2µl1 µr1 µl1 µr2

n −2 c X b NbX bn X bn  n l1 ,r1 ,l2 ,r2 ,n l1 ,r1 ,l2 ,r2 ,n πk−1+(j+1)b (Y d , Y 0d ) − πk−1+(j+1)b (Y d , Y 0d ) 0 0 n +m,k −1+(j+1)bn +m n +m,k −1+jbn +m

j=0

k=1 k0 =1

l1 ,r1 ,l2 ,r2 ,n l1 ,r1 ,l2 ,r2 ,n − πk−1+jb (Y d , Y 0d ) + πk−1+jb (Y d , Y 0d ) 0 0 n +m,k −1+jbn +m n +m,k −1+(j+1)bn +m



ˇ (1),l1 ,r1 ,l2 ,r2 ,n (Y d , Y 0d ) − Σ ˇ (2),l1 ,r1 ,l2 ,r2 ,n (Y d , Y 0d ) − Σ ˇ (3),l1 ,r1 ,l2 ,r2 ,n (Y d , Y 0d ) + Σ ˇ (4),l1 ,r1 ,l2 ,r2 ,n (Y d , Y 0d ). ≡Σ ij,m ij,m ij,m ij,m The statement in (A.17) is therefore reduced to showing that p ˇ (k),l1 ,r1 ,l2 ,r2 ,n (Y d , Y 0d ) → Σ 0, ij,m

(A.19)

for k = 1, . . . , 4. The convergence in probability to zero of the four terms is proven with identical techniques. It is therefore sufficient to show it for a single k, so we do it with k = 1. To this end, let Fu (x) = F (x)1{|x1 | 0, where F (x) = |x1 |l1 |x2 |r1 |x3 |l2 |x4 |r2 with | x = x1 x2 x3 x4 . Following the line of thought used also in the proof of Theorem 2.1, we can √ show that for wn = υn / un with un = kn /n:  −2  2 |Fwn (x + y) − Fwn (x)| ≤ wn1−2ω kxkp+ 1−2ω + (1 + kxkp ) kyk ∧ 1 + (kyk ∧ wn )p , | √ where p = l1 + r1 + l2 + r2 . Next, set x = ∆nk Y¯ d ∆nk+kn Y¯ d ∆nk0 Y¯ d ∆nk0 +kn Y¯ d / un , y = | √ ∆nk Y¯ 00d ∆nk+kn Y¯ 00d ∆nk0 Y¯ 00d ∆nk0 +kn Y¯ 00d / un . As in (A.5) – (A.6), it holds true that  E kxkp ≤ K,

 E kyk ∧ 1 ≤ Ku1−β/2 φn n

and

(p−2)  ω(p−β)− 2 E (kyk ∧ wn )p ≤ Kun φn , (A.20)

where φn → 0 as n → 0. Therefore, l1 +r1 +l2 +r2 −6 4

n 2µl1 µr1 µl1 µr2

n −2 b NbX c X bn bn X n

j=0

  l1 ,r1 ,l2 ,r2 ,n d 0d E πk−1+(j+1)bn +m,k0 −1+(j+1)bn +m (Y , Y ) | {z }! k=1 k0 =1 l1 +r1 +l2 +r2  2 = O un un + u1−r/2 φn + uω(4−r)−1 φn n n

 1  4δ1 −2 l1 +r1 +l2 +r2 −2−2ω(l1 +r1 +l2 +r2 −β) β−2 4 ≤ Kn 4 n− 2 + n 4 φn + n φn   4δ1 −4+β 4δ1 −4+l1 +r1 +l2 +r2 −2ω(l1 +r1 +l2 +r2 −β) 4 ≤ K nδ1 −1 + n 4 φn + n φn → 0, which concludes the proof of (A.17) and, hence, Theorem 4.1.

46



Proof of Lemma 4.1. The linearity of the expectation operator implies that   Jn  1 X ∗ n ∗ n∗ n ˇ ˇ ˇ = E BV (l, r) − √ E BV (l, r) ∆B(l, r)j uj bn j=1 Jn X ˇ (l, r)n − √1 ˇ = BV ∆B(l, r)nj E ∗ (uj ). bn j=1  ˇ (l, r)n . The second part of the lemma ˇ (l, r)n∗ = BV Then, if E ∗ (uj ) = 0, it follows that E ∗ BV

follows from (46) and (52), as for 1 ≤ i, j ≤ 2,   ˇ (li , ri n∗ , n1/4 BV ˇ (lj , rj )n∗ cov∗ n1/4 BV   Jn Jn X X √ 1 1 n ∗ n n n ˇ (lj , rj ) − √ ˇ (li , ri ) − √ ˇ i , ri )k uk , BV ˇ j , rj )k uk = n cov BV ∆B(l ∆B(l bn k=1 bn k=1 √ X J n n ˇ ˇ j , rj )nk var∗ (uk ). ∆B(li , ri )nk ∆B(l = bn k=1 Thus, if var∗ (uk ) = 1/2, we find that ˇ (li , ri cov∗ n1/4 BV where ˇ l1 ,r1 ,l2 ,r2 ,n Σ ij

n∗

 ˇ (lj , rj )n∗ = Σ ˇ l1 ,r1 ,l2 ,r2 ,n , , n1/4 BV ij

√ X J n n ˇ ˇ j , rj )n . ∆B(li , ri )nk ∆B(l = k 2bn k=1 

Proof of Corollary 4.1. Given (25), (27), and (65) the results follows from the properties of stable convergence.



ˇ l1 ,r1 ,l2 ,r2 Proof of Corollary 4.2. The result follows directly given (57) and the consistency result of Σ in Theorem 4.1.



Proof of Theorem 4.2. We again prove the theorem in model (1) only, noting that this is enough, as it nests both (A.7) and (10). Write Z

n∗

ˇn = Σ

−1/2

n

1/4

Jn X

Dj e∗j

1/4

≡n

j=1

Jn X

zj∗ ,

j=1

 ˇ n −1/2 Dj e∗ , with zj∗ ≡ Σ j Dj =

ˇ 1 , r1 )nj ∆B(l 0 ˇ 2 , r2 )nj , 0 ∆B(l

! and

e∗j

 =

 uj − E ∗ (uj ) uj − E ∗ (uj )

where uj are i.i.d. with var∗ (uj ) = 1/2. Note that e∗j is an i.i.d. zero mean vector. We follow Pauly (2011) and use a modified Cramer-Wold device to establish the bootstrap CLT. Let D = {λk : k ∈ 47

N } be a countable dense subset of the unit circle on R2 . The proof follows by showing that for any d∗

λ ∈ D, λ| Zn∗ → N (0, 1), in probability-P , as n → ∞. We note that |

λ

Zn∗

=n

1/4

Jn X

λ| zj∗ .

j=1

It follows from Lemma 4.1 and Corollary 4.2 that E ∗ (λ| Zn∗ ) = 0 and var∗ (λ| Zn∗ ) = 1 for all n. To conclude, it thus remains to prove that λ| Zn∗ is asymptotically normally distributed, conditionally n on the original sample and with probability P approaching one. As (zj∗ )Jj=1 forms an independent array—conditionally on the sample—by (e.g., Katz (1963)), for some small thePBerry-Esseen bound  P n ∗ 1/4 | ∗ 2+ε ∗ Jn 1/4 | ∗ E n λ zj . λ zj ≤ x − Φ(x) ≤ K Jj=1 ε > 0 and a constant K > 0, supx∈R P j=1 n PJn ∗ 1/4 | ∗ 2+ε Next, we show that j=1 E n λ zj = op (1). First, for a constant K independent of n (note ∗ that the moments of ej do not depend on n) and any 1 ≤ j ≤ Jn by the cr -inequality:

| ∗ 2+ε

ˇ n −1/2 2+ε λ zj ≤ kλk2+ε Σ

kDj k2+ε ke∗j k2+ε . Thus, E



|λ| zj∗ |2+ε



≤ kλk

≤ K

2+ε   n −1/2 ˇ

Σ

kDj k2+ε E ∗ ke∗j k2+ε

2+ε  ˇ n −1/2 Σ

kDj k2+ε ,

2+ε

implying that Jn X

E



|n1/4 λ| zj∗ |2+ε

≤ Kn

2+ε 4

j=1

Jn

ˇ n −1/2 2+ε X kDj k2+ε

Σ

j=1

≤ Kn

2+ε 4

Jn 

  

ˇ n −1/2 2+ε X ˇ 1 , r1 )nj 2+ε + ∆B(l ˇ 2 , r2 )n 2+ε ∆B(l

Σ

j j=1

≤ Kn

2+ε 4

Jn



ˇ n −1/2 2+ε X ˇ ˇ 1 , r1 )nj 2+ε B(l1 , r1 )nj+bn − B(l Σ

j=1

+ Kn

2+ε 4

Jn



ˇ n −1/2 2+ε X ˇ ˇ 2 , r2 )nj 2+ε B(l2 , r2 )nj+bn − B(l

Σ j=1

≤ Kn

2+ε 4

Jn 

2+ε  

ˇ n −1/2 2+ε X ˇ ˇ 2 , r2 )n 2+ε , B(l1 , r1 )nj + B(l

Σ

j

(A.21)

j=1

ˇ 1 , r1 )n where the second inequality is due to that, for any j, kDj k2 = ∆B(l j

48

2

 ˇ 2 , r2 )n 2 , + ∆B(l j

ˇ ˇ r)n : while the third is by expression of ∆B(l, r)nj . Next, note that by definition of B(l, j Jn X

ˇ r)n B(l, j

2+ε

=

j=1

Jn  X

n

l+r −1 4

j=1

2+ε bn 1 X n yˇ(l, r)i−1+j µl µr i=1

l+r ≤ Kn( 4 −1)(2+ε) b1+ε n

Jn X bn X

yˇ(l, r)ni−1+j

2+ε

j=1 i=1

= Op n

(δ1 −1)(1+ε)



.

We can therefore write (A.21) as follows: Jn X

 2+ε   E ∗ |n1/4 λ| zj∗ |2+ε = Op n 4 n(δ1 −1)(1+ε)

j=1

= op (1), where the last equality follows as for ε > 2, so long as 1/2 < δ1 < 2/3, (δ1 − 1)(1 + ε) +

2+ε 4

This completes the proof of (71). The last results then follow by application of the delta rule.

< 0. 

Proof of Theorem 4.3. First, we define: H

n∗

 ˇ n∗ −1/2 1/4

= Σ

ˇ n∗ ≡ Σ

n

−1/2

ˇn Σ

ˇ (l1 , r1 )n∗ ˇ (l1 , r1 )n∗ − E ∗ BV BV

 !

ˇ (l2 , r2 )n∗ ˇ (l2 , r2 )n∗ − E ∗ BV BV



1/2

Z n∗ ,

where Z

n∗

 ˇ n −1/2 1/4

= Σ

n

ˇ (l1 , r1 )n∗ ˇ (l1 , r1 )n∗ − E ∗ BV BV

 !

ˇ (l2 , r2 )n∗ ˇ (l2 , r2 )n∗ − E ∗ BV BV



.

d∗

It follows from Theorem 4.2 that Z n∗ → N (0, I2 ). Thus, the central limit theory for H n∗ is   p∗ ˇ n −1 Σ ˇ n∗ → ˇ n∗ −1 Σ ˇn = Σ established, if we can show that Σ I2 . To do this, we prove that i ∗ h i ∗ h   p p ˇ n∗ → ˇ n −1 Σ ˇ n∗ → ˇ n −1 Σ 0. (A.22) E∗ Σ I2 and var∗ Σ ˇ n and Σ ˇ n∗ . Next, again by definition: The first equation in (A.22) holds by the definition of Σ ! √ Jn h i h i ∗ X     n var (u) ˇ n −1 Σ ˇ n∗ = Σ ˇ n −1 ⊕ Σ ˇ n −1 var∗ var∗ Σ ξˇj ξˇ| u2 bn E ∗ (u2 ) j=1 j j !2 √ Jn  iX  n var∗ (u) h ˇ n −1 n −1 ∗ ˇ ˇ| 2 ˇ = Σ ⊕ Σ var ξ ξ u j j j bn E ∗ (u2 ) j=1 var∗ (u) = var (u ) E ∗ (u2 ) ∗

2



2 h

49

ˇn Σ

−1

ˇn ⊕ Σ

Jn −1 i n X   ξˇj ξˇj| ⊕ ξˇj ξˇj| . 2 bn j=1

As in the proof of Theorem 4.2 in Hounyo (2017): r  r 

!

Jn Jn X

4  

n X   n | | n ˇj ξˇ ⊕ ξˇj ξˇ ≤ K ˇ 1 , r1 ) ˇ 2 , r2 )n 4 B(l B(l ξ E E E j j j j

b2 b2n j=1 n j=1 r  r  4   ˇ 1 , r1 )n ˇ 2 , r2 )n 4 + E B(l E B(l j

j+bn

r  r  4   ˇ 1 , r1 )n ˇ 2 , r2 )n 4 B(l + E B(l E j j+bn r  ˇ 1 , r1 )n + E B(l

j+bn

4 

r  ˇ 2 , r2 )n E B(l

j+bn

4 

! ≤K

b2n → 0, n2

r   l+r−1 ˇ r)n 4 ≤ ≤ $ < 1/2 means that E B(l, where the last inequality follows, because j 2(l + r − β) 2  b K n2 . As Jn = O(n) and bn = O nδ1 such that 1/2 < δ1 < 2/3 from (44), it follows that n h i ∗  p ˇ n −1 Σ ˇ n∗ → var∗ Σ 0. This finishes the proof of the first part Theorem 4.3. The last result again follows by a direct application of the delta rule.



Proof of Proposition 3.1. To begin with, notice that 2 2 √ √  m  2 √ √ m m∆m m∆m = + m m∆m (t−1)m+i Y (t−1)m+i X (t−1)m+i X ∆(t−1)m+i  + m ∆(t−1)m+i  . Thus, for s ∈ [t − 1 + (j − 1)/m, t − 1 + j/m), where j = 1, . . . , m and t = 1, . . . , T , √ X T T 2  m  √ m 1X √ m 2 m∆(t−1)m+j X + m∆m σ ˆu,s = (t−1)m+j X ∆(t−1)m+j  T t=1 T t=1 T T 2 m X    mX m + ∆(t−1)m+j  − var ˆ (t−1)+(j−1)/m + var ˆ (t−1)+j/m . T t=1 T t=1

(A.23)

The proof now proceeds in three steps, where we show that: T 2  1X √ 2 + OL2 T −1/2 m1/4 , m∆m = σu,s (t−1)m+j X T t=1

(A.24)



T p   m  mX √ m m∆(t−1)m+j X ∆(t−1)m+j  = OP m/T , T t=1

(A.25)

and T T 2 m X    mX m ∆(t−1)m+j  = var t−1+(j−1)/m + var t−1+j/m ) + OP mT −1/2 . T t=1 T t=1

50

(A.26)

To show step 1, we define: √ m α(t−1)m+j ≡ mσt−1+(j−1)/m ∆m (t−1)m+j W

and χm (t−1)m+j ≡

√ m m∆m (t−1)m+j X − α(t−1)m+j .

We also set 2 σ ˜u,s

T 1X 2 σ . ≡ T t=1 t−1+(j−1)/m

Now, the proof is complete, if we can show that:  2 2 σ ˜u,s − σu,s = OL2 T −1/2 ,

(A.27)

T  1X m 2 |α(t−1)m+j |2 − σ ˜u,s = OL2 T −1/2 , T t=1

(A.28)

T  2  1 X √ m 2 −1/2 1/4 2 T m∆m X − |α | = O m . L (t−1)m+j (t−1)m+j T t=1

(A.29)

and

As for (A.27), note that 2 σ ˜u,s

T T 2 X σu,(j−1)/m 1X 2 2 σ = σsv,t−1+(j−1)/m . ≡ T t=1 t−1+(j−1)/m T t=1

Now, by Assumption (D2) we deduce that   ∞ X  K  2 2 2 var σ ˜u,s ≤ cov σsv,1 , σsv,1+k . 1+2 T k=0 Hence (A.27) follows. Next, (A.28) can be verified by martingale techniques. First, we write T T    1X m 1X m 2 2 2 m 2 |α |α(t−1)m+j | − E |α(t−1)m+j | | Ft−1+ j−1 . | −σ ˜u,s = m T t=1 (t−1)m+j T t=1 Then, "

T 1X m 2 |2 − σ ˜u,s |α E T t=1 (t−1)m+j

#2

 T  2 1 X 2 m 2 m = 2 E |α(t−1)m+j | − E |α(t−1)m+j | | Ft−1+ j−1 m T t=1 T 2 X 4 = 2 σ . T t=1 t+(j−1)/m

Treating the error T

−1

T   X 2 √ m 2 − |α | from (A.29)) is the hardest one. m∆m X (t−1)+j (t−1)m+j t=1

In the following, we denote Z t Z t 0 00 σs dWs Xt = X0 + as ds + 0

0

  and Xt00 = Xt − Xt0 = δ1{|δ|≤1} ? µt − ν t ,

51

where  a00t = as + δ1{|δ|>1} ? ν t is bounded.  Also, note that an error OL2 T −1/2 m1/4 appears, when we are dealing with terms that do not have a martingale structure (i.e., the jump component of X). Here, Corollary 2.1.9 of Jacod and Protter (2012) plays a key role in the proof, because it turns out the discontinuous part is asymptotically negligible in front of its Brownian part on a vanishing interval of the form [t − 1 + (i − 1)/m, t − 1 + i/m). The definition of χm (t−1)m+j and Assumption (V) yields the decomposition: √ m χm m∆m (t−1)m+j = (t−1)m+j X − α(t−1)m+j √ √ 0 00 m = m∆m m∆m (t−1)m+j X + (t−1)m+j X − α(t−1)m+j ! Z t−1+ j Z t−1+ j   m m √ 00 a00s ds + σs − σt−1+ j−1 dWs + ∆m = m (t−1)m+j X t−1+ j−1 m

m

t−1+ j−1 m

m m ≡ χm (t−1)m+j (1) + χ(t−1)m+j (2) + χ(t−1)m+j (3),

where χm (t−1)m+j (1) = χm (t−1)m+j (2)

=





1 00 m a j−1 + m t−1+ m Z

j t−1+ m

m



t−1+ j−1 m

√ + m χm (t−1)m+j (3) =



Z

j t−1+ m

t−1+ j−1 m



a00s

Z

j t−1+ m







σ ˜t−1+ j−1 Ws − Wt−1+ j−1 + v˜s Bs − Bt−1+ j−1 m

t−1+ j−1 m





a00t−1+ j−1 m



m

Z

j t−1+ m

"Z

#

t−1+ j−1 m

t−1+ j−1 m

"Z



σ ˜u − σ ˜t−1+ j−1 dWs + m

s



t−1+ j−1 m

m

s

ds +



! dWs ,

!

a ˜s du dWs 

#

!

v˜u − v˜t−1+ j−1 dBs dWs , m

00 m∆m (t−1)m+j X .

Together with the assumptions behind Proposition 3.1, we can then appeal to the Burkholder and Cauchy-Schwarz inequality (first and second expression) and Lemma 2.1.5 in Jacod and Protter (2012) with p = 4 (last estimate) to deduce that:   −1 −1/2 , χm , χm (t−1)m+j (2) = OL4 m (t−1)m+j (1) = OL4 m

 1/4 χm . (t−1)m+j (3) = OL4 m

(A.30)

Next, let f (x) = x2 so that f 0 (x) = 2x. Then, by Taylor expansion: T   2 1 X √ −1/2 m 2 m m , m∆m X −|α | = Am (t−1)m+j (t−1)m+j (t−1)m+j (1)+A(t−1)m+j (2)+A(t−1)m+j (3)+OL2 T T t=1

where Am (t−1)m+j (k)

T  2X m = α(t−1)m+j χm (k) , (t−1)m+j T t=1

52

for k = 1, 2 and 3. Note the following martingale difference property:   m m j−1 = 0, Am (k) = E α χ (k) | F (t−1)m+j (t−1)m+j (t−1)m+j t−1+ m

for k = 1, 2 and 3. Thus, from the Cauchy-Schwarz inequality T   m  2i 4 X h m 2 E |A(t−1)m+j (1)| = 2 E | α(t−1)m+j χm (1) | (t−1)m+j T t=1 T  4 1/2  4 1/2 4 X m ≤ 2 E| αj+tm | E| χm . j+tm (1) | T t=1 m Thus, given (A.30) and the fact that α(t−1)m+j = OL4 (1), it follows that    −1/2 1/4 −1/2 −1 −1/2 −1/2 . m , Am m , and Am Am m (t−1)m+j (3) = OL2 T (t−1)m+j (2) = OL2 T (t−1)m+j (1) = OL2 T

This shows (A.29), and then we conclude that: T 2     1X √ 2 m∆m = σu,s + OL2 T −1/2 + OL2 T −1/2 m−1/2 + OL2 T −1/2 m−1 + OL2 T −1/2 m1/4 (t−1)m+j X T t=1  2 = σu,s + OL2 T −1/2 m1/4 ,

which completes the entire proof of step 1. We move forward to step 2. To deduce (A.25), we write: T T T X X  m  X   √ √ √ m m m∆(t−1)m+j X ∆(t−1)m+j  = m∆(t−1)m+j X t−1+j/m − m∆m (t−1)m+j X t−1+(j−1)/m . t=1

t=1

t=1

Note that from (5) and as t is independently distributed with X ⊥⊥ : "√ # "√ # T T X   √ mX √ m E m∆m m∆m (t−1)m+j X t−1+j/m | X = 0 and E (t−1)m+j X t−1+(j−1)/m | X = 0. T t=1 T t=1 Then, we get "√ " # # 2 T T X   √ mX √ 1 m 2 2 E m∆m m∆m | X = ω2 (t−1)m+j X t−1+k/m (t−1)m+j X σu,t−1+k/m , T t=1 T T t=1 for k = j and j − 1. And from T

−1

T X 2 √ 2 m∆m X = OP (1) and σu,t being bounded: (t−1)m+j t=1



T   mX √ −1/2 1/2 m∆m m , (t−1)m+j X t−1+k/m = OP T T t=1

2 for k = j and j − 1. This establishes (A.25). To show (A.26), note that because ∆m  is a (t−1)m+j h 2 i   1-dependent sequence: E ∆m  = var t−1+(j−1)/m + var t−1+j/m , and (t−1)m+j " # T 2  1 X m K var ∆(t−1)m+j  − var(t−1+(j−1)/m ) + var(t−1+j/m ) ≤ . T t=1 T 53

Hence, T T 2   1X m 1X ∆(t−1)m+j  = var(t−1+(j−1)/m ) + var(t−1+j/m ) + OP T −1/2 . T t=1 T t=1

It follows that T 2  m X m ∆(t−1)m+j  − var( ˆ t−1+(j−1)/m ) + var( ˆ t−1+j/m ) T t=1 T  i m Xh =− var( ˆ t−1+(j−1)/m ) − var(t−1+(j−1)/m ) + var( ˆ t−1+j/m ) − var(t−1+j/m ) T t=1  + OP mT −1/2 ,

so (A.26) holds. Inserting (A.24) – (A.26) into (A.23): T h  i  mX = var( ˆ t−1+(j−1)/m )−var(t−1+(j−1)/m ) + var( ˆ t−1+j/m )−var(t−1+j/m ) . mT − T t=1  At last, we note that when var ˆ t−1+(j−1)/m is given by (32), it holds that 2 σ ˆu,s

2 σu,s +OP

−1/2

T   i −1/2 −1/2  1 Xh var ˆ t−1+(j−1)/m + var ˆ t−1+j/m = OP n/m T . T t=1

Hence, (31) reduces to:   2 2 σ ˆu,s = σu,s + OP mT −1/2 + OP m3/2 n−1/2 T −1/2 , which completes the proof.



54

References A¨ıt-Sahalia, Y., and J. Jacod, 2012a, “Analyzing the spectrum of asset returns: Jump and volatility components in high frequency data,” Journal of Economic Literature, 50(4), 1007–1050. , 2012b, “Is Brownian motion necessary to model high-frequency data?,” Annals of Statistics, 38(5), 3093–3128. A¨ıt-Sahalia, Y., J. Jacod, and J. Li, 2012, “Testing for jumps in noisy high frequency data,” Journal of Econometrics, 168(2), 207–222. A¨ıt-Sahalia, Y., and D. Xiu, 2015, “Increased correlation among asset classes: Are volatility or jumps to blame, or both?,” Journal of Econometrics, 194(2), 205–219. A¨ıt-Sahalia, Y., and J. Yu, 2009, “High frequency market microstructure noise estimates and liquidity measures,” Annals of Applied Statistics, 3(1), 422–457. Andersen, T. G., and T. Bollerslev, 1997, “Intraday periodicity and volatility persistence in financial markets,” Journal of Empirical Finance, 4(2), 115–158. , 1998, “Deutsche Mark-Dollar volatility: Intraday activity patterns, macroeconomic announcements, and longer run dependencies,” Journal of Finance, 53(1), 219–265. Andersen, T. G., T. Bollerslev, and A. Das, 2001, “Variance-ratio statistics and high-frequency data: Testing for changes in intraday volatility patterns,” Journal of Finance, 56(1), 305–327. Andersen, T. G., T. Bollerslev, and D. Dobrev, 2007, “No-arbitrage semi-martingale restrictions for continuous-time volatility models subject to leverage effects, jumps and i.i.d. noise: Theory and testable distributional implications,” Journal of Econometrics, 138(1), 125–180. Andersen, T. G., D. Dobrev, and E. Schaumburg, 2012, “Jump-robust volatility estimation using nearest neighbour truncation,” Journal of Econometrics, 169(1), 75–93. , 2014, “A robust neighborhood truncation approach to estimation of integrated quarticity,” Econometric Theory, 30(1), 3–59. Baeumer, B., and M. M. Meerschaert, 2010, “Tempered stable L´evy motion and transient super-diffusion,” Journal of Computational and Applied Mathematics, 233(10), 2438–2448. Bandi, F. M., and J. R. Russell, 2006, “Separating microstructure noise from volatility,” Journal of Financial Economics, 79(3), 655–692. Barndorff-Nielsen, O. E., S. E. Graversen, J. Jacod, M. Podolskij, and N. Shephard, 2006, “A central limit theorem for realized power and bipower variations of continuous semimartingales,” in From Stochastic Calculus to Mathematical Finance: The Shiryaev Festschrift, ed. by Y. Kabanov, R. Lipster, and J. Stoyanov. Springer-Verlag, Heidelberg, pp. 33–68. Barndorff-Nielsen, O. E., P. R. Hansen, A. Lunde, and N. Shephard, 2008, “Designing realized kernels to measure the ex post variation of equity prices in the presence of noise,” Econometrica, 76(6), 1481–1536. , 2009, “Realized kernels in practice: trades and quotes,” Econometrics Journal, 12(3), 1–32. Barndorff-Nielsen, O. E., and N. Shephard, 2004, “Power and bipower variation with stochastic volatility and jumps,” Journal of Financial Econometrics, 2(1), 1–48.

55

, 2006, “Econometrics of testing for jumps in financial economics using bipower variation,” Journal of Financial Econometrics, 4(1), 1–30. Boudt, K., C. Croux, and S. Laurent, 2011, “Robust estimation of intraweek periodicity in volatility and jump detection,” Journal of Empirical Finance, 18(2), 353–367. Brownless, C. T., and G. M. Gallo, 2006, “Financial econometric analysis at ultra-high frequency: Data handling concerns,” Computational Statistics and Data Analysis, 51(4), 2232–2245. Chernov, M., A. R. Gallant, E. Ghysels, and G. Tauchen, 2003, “Alternative models for stock price dynamics,” Journal of Econometrics, 116(1-2), 225–257. Christensen, K., S. Kinnebrock, and M. Podolskij, 2010, “Pre-averaging estimators of the ex-post covariance matrix in noisy diffusion models with non-synchronous data,” Journal of Econometrics, 159(1), 116–133. Christensen, K., R. C. A. Oomen, and M. Podolskij, 2014, “Fact or friction: Jumps at ultra high frequency,” Journal of Financial Economics, 114(3), 576–599. Christensen, K., M. Podolskij, N. Thamrongrat, and B. Veliyev, 2016, “Inference from high-frequency data: A subsampling approach,” Journal of Econometrics, 197(2), 245–272. Christensen, K., M. Podolskij, and M. Vetter, 2013, “On covariation estimation for multivariate continuous Itˆo semimartingales with noise in non-synchronous observation schemes,” Journal of Multivariate Analysis, 120(1), 59–84. Clinet, S., and Y. Potiron, 2017, “Efficient asymptotic variance reduction when estimating volatility in high frequency data,” Working paper, University of Tokyo. Corsi, F., D. Pirino, and R. Ren` o, 2010, “Threshold bipower variation and the impact of jumps on volatility forecasting,” Journal of Econometrics, 159(2), 276–288. Delbaen, F., and W. Schachermayer, 1994, “A general version of the fundamental theorem of asset pricing,” Mathematische Annalen, 300(1), 463–520. Dette, H., V. Golosnoy, and J. Kellermann, 2016, “The effect of intraday periodicity on realized volatility measures,” Working paper, Ruhr-Universit¨at Bochum. Dette, H., and M. Podolskij, 2008, “Testing the parametric form of the volatility in continuous time diffusion models–A stochastic process approach,” Journal of Econometrics, 143(1), 56–73. Dette, H., M. Podolskij, and M. Vetter, 2006, “Estimation of integrated volatility in continuous-time financial models with applications to goodness-of-fit testing,” Scandinavian Journal of Statistics, 33(2), 259–278. Dovonon, P., S. Gon¸calves, U. Hounyo, and N. Meddahi, 2017, “Bootstrapping high-frequency jump tests,” Journal of the American Statistical Association, (Forthcoming). Dudek, A. E., J. Le´skow, E. Paparoditis, and D. N. Politis, 2014, “A generalized block bootstrap for seasonal time series,” Journal of Time Series Analysis, 35(2), 89–114. Engle, R. F., and M. E. Sokalska, 2012, “Forecasting intraday volatility in the US equity market. Multiplicative component GARCH,” Journal of Financial Econometrics, 10(1), 54–83. Fama, E. F., 1965, “The behavior of stock-market prices,” Journal of Business, 38(1), 34–105.

56

Gon¸calves, S., U. Hounyo, and N. Meddahi, 2014, “Bootstrap inference for pre-averaged realized volatility based on non-overlapping returns,” Journal of Financial Econometrics, 12(4), 679–707. Gon¸calves, S., and N. Meddahi, 2009, “Bootstrapping realized volatility,” Econometrica, 77(1), 283–306. Hansen, P. R., and A. Lunde, 2006, “Realized variance and market microstructure noise,” Journal of Business and Economic Statistics, 24(2), 127–161. Harris, L., 1986, “A transaction data study of weekly and intradaily patterns in stock returns,” Journal of Financial Economics, 16(1), 99–117. Hasbrouck, J., 1999, “The dynamics of discrete bid and ask quotes,” Journal of Finance, 54(6), 2109–2142. Hautsch, N., and M. Podolskij, 2013, “Pre-averaging based estimation of quadratic variation in the presence of noise and jumps: Theory, implementation, and empirical evidence,” Journal of Business and Economic Statistics, 31(2), 165–183. Hecq, A., S. Laurent, and F. C. Palm, 2012, “Common intraday periodicity,” Journal of Financial Econometrics, 10(2), 325–353. Hounyo, U., 2014, “The wild tapered block bootstrap,” Working paper, CREATES, Aarhus University. , 2017, “Bootstrapping integrated covariance matrix estimators in noisy jump-diffusion models with non-synchronous trading,” Journal of Econometrics, 197(1), 130–152. , 2018, “A local Gaussian bootstrap method for realized volatility and realized beta,” Econometric Theory, (Forthcoming). Hounyo, U., S. Gon¸calves, and N. Meddahi, 2017, “Bootstrapping pre-averaged realized volatility under market microstructure noise,” Econometric Theory, 33(4), 791–838. Huang, X., and G. Tauchen, 2005, “The relative contribution of jumps to total price variance,” Journal of Financial Econometrics, 3(4), 456–499. Jacod, J., 2008, “Asymptotic properties of realized power variations and related functionals of semimartingales,” Stochastic Processes and their Applications, 118(4), 517–559. Jacod, J., Y. Li, P. A. Mykland, M. Podolskij, and M. Vetter, 2009, “Microstructure noise in the continuous case: The pre-averaging approach,” Stochastic Processes and their Applications, 119(7), 2249–2276. Jacod, J., M. Podolskij, and M. Vetter, 2010, “Limit theorems for moving averages of discretized processes plus noise,” Annals of Statistics, 38(3), 1478–1545. Jacod, J., and P. E. Protter, 2012, Discretization of Processes. Springer-Verlag, Berlin, 2nd edn. Jacod, J., and V. Todorov, 2009, “Testing for common arrival of jumps in discretely-observed multidimensional processes,” Annals of Statistics, 37(4), 1792–1838. Jiang, G. J., I. Lo, and A. Verdelhan, 2011, “Information shocks, liquidity shocks, jumps, and price discovery: Evidence from the U.S. Treasury market,” Journal of Financial and Quantitative Analysis, 46(2), 527–551. Jing, B.-Y., Z. Liu, and X.-B. Kong, 2014, “On the estimation of integrated volatility with jumps and microstructure noise,” Journal of Business and Economic Statistics, 32(3), 457–467. Kalnina, I., 2011, “Subsampling high frequency data,” Journal of Econometrics, 161(2), 262–283.

57

Kalnina, I., and O. Linton, 2008, “Estimating quadratic variation consistently in the presence of endogenous and diurnal measurement error,” Journal of Econometrics, 147(1), 47–59. Katz, M. L., 1963, “Note on the Berry-Esseen theorem,” Annals of Mathematical Statistics, 34(3), 1107– 1108. Kolokolov, A., and R. Ren` o, 2016, “Efficient multipowers,” Working paper, University of Verona. , 2017, “Jumps or flatness?,” Working paper, University of Verona. Lee, S. S., and P. A. Mykland, 2008, “Jumps in financial markets: A new nonparametric test and jump dynamics,” Review of Financial Studies, 21(6), 2535–2563. Li, J., V. Todorov, and G. Tauchen, 2013, “Volatility occupation times,” Annals of Statistics, 41(4), 1865–1891. , 2016, “Estimating the volatility occupation time via regularized Laplace inversion,” Econometric Theory, 32(5), 1253–1288. Liu, R. Y., 1988, “Bootstrap procedures under some non-i.i.d. models,” Annals of Statistics, 16(4), 1696– 1708. Mammen, E., 1993, “Bootstrap and wild bootstrap for high dimensional linear models,” Annals of Statistics, 21(1), 255–285. Mancini, C., 2009, “Non-parametric threshold estimation for models with stochastic diffusion coefficient and jumps,” Scandinavian Journal of Statistics, 36(2), 270–296. Mancini, C., and F. Gobbi, 2012, “Identifying the Brownian covariation from the co-jumps given discrete observations,” Econometric Theory, 28(2), 249–273. Mandelbrot, B. B., 1963, “The variation of certain speculative prices,” Journal of Business, 36(4), 394–419. Mykland, P. A., and L. Zhang, 2009, “Inference for continuous semimartingales observed at high frequency: A general approach,” Econometrica, 77(5), 1403–1445. , 2017, “Assessment of uncertainty in high frequency data: The observed asymptotic variance,” Econometrica, 85(1), 197–231. Pauly, M., 2011, “Weighted resampling of martingale difference arrays with applications,” Electronic Journal of Statistics, 5(1), 41–52. Podolskij, M., and M. Vetter, 2009a, “Bipower-type estimation in a noisy diffusion setting,” Stochastic Processes and their Applications, 119(9), 2803–2831. , 2009b, “Estimation of volatility functionals in the simultaneous presence of microstructure noise and jumps,” Bernoulli, 15(3), 634–658. , 2010, “Understanding limit theorems for semimartingales: A short survey,” Statistica Neerlandica, 64(3), 329–351. Podolskij, M., and K. Wasmuth, 2013, “Goodness-of-fit testing for fractional diffusions,” Statistical Inference for Stochastic Processes, 16(2), 147–159. Politis, D. N., J. P. Romano, and M. Wolf, 1999, Subsampling, vol. 1. Springer-Verlag.

58

Shao, X., 2010, “The dependent wild bootstrap,” Journal of the American Statistical Association, 105(489), 218–235. Taylor, S. J., and X. Xu, 1997, “The incremental volatility information in one million foreign exchange quotations,” Journal of Empirical Finance, 4(4), 317–340. Todorov, V., and T. Bollerslev, 2010, “Jumps and betas: A new framework for disentangling and estimating systematic risks,” Journal of Econometrics, 157(2), 220–235. Todorov, V., and G. Tauchen, 2010, “Activity signature functions for high-frequency data analysis,” Journal of Econometrics, 154(2), 125–138. , 2011, “Volatility jumps,” Journal of Business and Economic Statistics, 29(3), 356–371. , 2012, “The realized Laplace transform of volatility,” Econometrica, 80(3), 11051127. Todorov, V., G. Tauchen, and I. Grynkiv, 2014, “Volatility activity: Specification and estimation,” Journal of Econometrics, 178(1), 180–193. Vetter, M., 2008, “Estimation methods in noisy diffusion models,” Ph.D. thesis, Ruhr-Universit¨at Bochum. , 2010, “Limit theorems for bipower variation of semimartingales,” Stochastic Processes and their Applications, 120(1), 22–38. Vetter, M., and H. Dette, 2012, “Model checks for the volatility under microstructure noise,” Bernoulli, 18(4), 1421–1447. Wood, R. A., T. H. McInish, and J. K. Ord, 1985, “An investigation of transactions data for NYSE stocks,” Journal of Finance, 40(3), 723–739. Wu, C. F. J., 1986, “Jackknife, bootstrap and other resampling methods in regression analysis,” Annals of Statistics, 14(4), 1261–1295. Xiu, D., 2010, “Quasi-maximum likelihood estimation of volatility with high frequency data,” Journal of Econometrics, 159(1), 235250. Zhang, L., 2006, “Efficient estimation of stochastic volatility using noisy observations: A multi-scale approach,” Bernoulli, 12(6), 1019–1043. Zhang, L., P. A. Mykland, and Y. A¨ıt-Sahalia, 2005, “A tale of two time scales: determining integrated volatility with noisy high-frequency data,” Journal of the American Statistical Association, 100(472), 1394–1411.

59

Is the diurnal pattern sufficient to explain intraday ...

We show that replacing the true diurnal factor with a model- free jump- and .... highly active (e.g., Aıt-Sahalia and Jacod, 2012a).4 Moreover, even if stochastic volatility is truly present, in practice its ...... The intuition is that in our model, there are no fixed points of discontinuity in X, so that the influence of any jumps.

1MB Sizes 0 Downloads 89 Views

Recommend Documents

Is the diurnal pattern sufficient to explain the intraday ...
A recent strand of work, fueled by access to high-frequency data and ..... In Section 3, we use this idea to recover σu from noisy high-frequency data, so that we ...

Meaning Through Syntax Is Insufficient to Explain ...
Keywords: lexical semantics, sentence processing, meaning through syntax ... between the linguistic analysis of a sentence and how it is pro- cessed. Rather ...

Very Limited Wealth Is Sufficient For You -
Mua'wiyah came to visit him, and Abû Hashim was crying. So Mua'wiyah asked: "What is it that makes you cry, O Uncle?! Is it the pain or is it for the world (that ...

How professors use the language of “perception” to explain ...
Sep 15, 2016 - For example, a male associate professor said, “I guess there's a perception among a lot of women that a career in engineering isn't desirable.

On Sufficient Conditions for Starlikeness
zp'(z)S@Q)) < 0(q(r)) * zq'(r)6@Q)), then p(z) < q(z)and q(z) i,s the best domi ..... un'iualent i,n A and sati,sfy the follow'ing condit'ions for z e A: .... [3] Obradovia, M., Thneski, N.: On the starlike criteria defined Silverman, Zesz. Nauk. Pol

G-is-for-Goat-pattern-crystal-and-company.pdf
Page 1 of 1. G-is-for-Goat-pattern-crystal-and-company.pdf. G-is-for-Goat-pattern-crystal-and-company.pdf. Open. Extract. Open with. Sign In. Main menu.

Sufficient Cohesion over atomic toposes
lois topos of Barr-atomic sheaves on finite extensions of the ground field. What does .... primitive Nullstellensatz if for every C in C there exists a map ιD → C for.

WIRELESS HOUSE, SELF - SUFFICIENT AND ...
Keywords: self sufficiency, renewable, solar, energy, appropriate technology, stirling engine, CSP, ... 'Wireless' in this context describes the independence ... For comparison, three different consumption scenarios have been defined: Scenario ...

Necessary and sufficient conditions in the problem of ...
Savings account with zero interest rate. M - family of martingale ... Primal Problem u(x) = sup c,XT : X0=x,X≥0. E. [∫ T. 0. U1(t, ω,ct )dt + U2(ω,XT ). ] . Is u a utility ...

sufficient-provision-for-the-seekers-of-the-path.pdf
sufficient-provision-for-the-seekers-of-the-path.pdf. sufficient-provision-for-the-seekers-of-the-path.pdf. Open. Extract. Open with. Sign In. Main menu.

Likelihood-based Sufficient Dimension Reduction
Sep 25, 2008 - If SY |X = Rp (d = p) then the log likelihood (1) reduces to the usual log ...... Figure 5: Plot of the first two LAD directions for the birds-planes-cars ...

WIRELESS HOUSE, SELF - SUFFICIENT AND ...
medium temperature (MT) storage for electricity generation, cooling and cooking .... as a biomass back-up system) may be necessary in order to assure sufficient.

The Impact of Competition and Information on Intraday ...
Nov 6, 2013 - Such an improvement can occur, for instance, when a company adopts or a regulator .... that are conditional on the net order flow (buy and sell orders). The market ...... Journal of Banking and Finance 24, 1323–1357. Back, K.

can immanence explain
The contract o_/'rtss()cicttio1t is intrinsic to arid it'tse.purob/e. _/'rot/n the contract ofsttbjtigution. [Empire 84] .... Empire deals with the distinction strategy/tactics.

FREE* Download Men Explain Things to Me *Ebook ...
misogyny, and the essay "Cassandra Syndrome." This book is also available in hardcover. Writer, historian, and activist Rebecca Solnit is the author of fourteen ...

On Sufficient Conditions for Starlikeness
E-mail: [email protected]. AMS Mathematics Subject Classification ..... zf'(z) l,o_6)'{',?) +t(. -'f"(')\l - CI(r-)'* 4:- f a Sru r\z) \'* f,@ )l -'\1-'/ - (t - zy' then f (z) € ,S* .