Skip to main content
Log in

Information loss in volatility measurement with flat price trading

  • Published:
Empirical Economics Aims and scope Submit manuscript

    We’re sorry, something doesn't seem to be working properly.

    Please try refreshing the page. If that doesn't work, please contact support so we can address the problem.

Abstract

A model of financial asset price determination is proposed that incorporates flat trading features into an efficient price process. The model involves the superposition of a Brownian semimartingale process for the efficient price and a Bernoulli process that determines the extent of flat price trading. The approach is related to sticky price modeling and the Calvo pricing mechanism in macroeconomic dynamics. A limit theory for the conventional realized volatility (RV) measure of integrated volatility is developed. The results show that RV is still consistent but has an inflated asymptotic variance that depends on the probability of flat trading. Estimated quarticity is similarly affected, so that both the feasible central limit theorem and the inferential framework suggested in Barndorff-Nielsen and Shephard (J Royal Stat Soc Ser B (Stat Methodol) 64:253–280, 2002) remain valid under flat price trading even though there is information loss due to flat trading effects. The results are related to work by Jacod (J Financ Econom 16:526–569, 2018) and Mykland and Zhang (Ann Stat 34:1931–1963, 2006) on realized volatility measures with random and intermittent sampling, and to ACD models for irregularly spaced transactions data. Extensions are given to include models with microstructure noise. Some simulation results are reported. Empirical evaluations with tick-by-tick data indicate that the effect of flat trading on the limit theory under microstructure noise is likely to be minor in most cases, thereby affirming the relevance of existing approaches.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2

Similar content being viewed by others

Notes

  1. Jacod (2018) is a subsequently published version of a working paper that appeared much earlier and was dated as 1993 in later citations of the paper (Delattre and Jacod 1997; Barndorff-Nielsen and Shephard 2002).

  2. Concerns about how to calculate RV based on business time sampling and transaction time sampling have received a great deal of attention in the RV literature; see Hansen and Lunde (2006); Oomen (2006). In particular, business time sampling relates to stopping time approaches and the connection is achieved via the Dubins-Schwarz Theorem. See Yu and Phillips (2001) for an application of the Dubins-Schwarz Theorem to estimate continuous-time models.

  3. Mykland and Zhang (2006) assume the sampling points \(\tau _{j}\) to be deterministic but note later in their paper that the scheme covers the case where the sampling points are random but independent of the observed process. The stopping time scheme (1.10) is random and allows for dependence on past prices.

  4. Note that the values \(L_{i}=0,1,2,\ldots ,\) correspond to realizations \(\xi _{i}=1,\) \(\left\{ \xi _{i}=1,\xi _{i-1}=0\right\} ,\) \(\left\{ \xi _{i}=1,\xi _{i-1}=0,\xi _{i-2}=0\right\} \) with respective probabilities \(\pi ,\) \(\pi \left( 1-\pi \right) ,\pi \left( 1-\pi \right) ^{2},.....\)

  5. In this case we have \(\left\{ m^{-1}\sum _{\ell \le j}D_{m,\ell }<t\right\} , \) which under D2 is asymptotically equivalent to \(\left\{ \int _{0}^{j/m}\mu _{D}\left( s\right) \textrm{d}s<t\right\} .\) The measure \(\mu \left[ 0,t\right] \) is then given by \(\mu \left[ 0,t\right] =r\left( t\right) \) where \(\int _{0}^{r}\mu _{D}\left( s\right) \textrm{d}s=t\) so that \(\mu _{D}\left( r\right) dr=\textrm{d}t.\)

  6. The datasets and dates are arbitrarily selected but are illustrative of heavily traded stocks.

References

  • Aït-Sahalia Y, Mykland PA, Zhang L (2005) How often to sample a continuous-time process in the presence of market microstructure noise. Rev Financ Stud 18:351–416

    Google Scholar 

  • Aït-Sahalia Y, Mykland PA, Zhang L (2011) Ultra high frequency volatility estimation with dependent microstructure noise. J Econom 160:160–175

    Google Scholar 

  • Andersen TG, Bollerslev T, Diebold FX (2010) “CHAPTER 2 - Parametric and Nonparametric Volatility Measurement,” In: Handbook of financial econometrics: tools and techniques, Aït-Sahalia Y, Hansen LP, San Diego: North-Holland, vol. 1 of Handbooks in Finance, pp 67–137

  • Andersen TG, Bollerslev T, Diebold FX, Ebens H (2001) The distribution of realized stock return volatility. J Financ Econ 61:43–76

    Google Scholar 

  • Andersen TG, Bollerslev T, Diebold FX, Labys P (2001) The distribution of realized exchange rate volatility. J Am Stat Assoc 96:42–55

    Google Scholar 

  • Andersen TG, Bollerslev T, Diebold FX, Labys P (2003) Modeling and Forecasting Realized Volatility. Econometrica 71:579–625

    Google Scholar 

  • Andersen TG, Bollerslev T, Diebold FX, Wu J (2005) A framework for exploring the macroeconomic determinants of systematic risk. Am Econ Rev 95:398–404

    Google Scholar 

  • Andersen TG, Bollerslev T, Meddahi N (2005) Correcting the errors: volatility forecast evaluation using high-frequency data and realized volatilities. Econometrica 73:279–296

    Google Scholar 

  • Back K (1991) Asset pricing for general processes. J Math Econ 20:371–395

    Google Scholar 

  • Bandi FM, Kolokolov A, Pirino D, Renò R (2020) Zeros. Manage Sci 66:3466–3479

    Google Scholar 

  • Bandi FM, Pirino D, Renò R (2017) EXcess idle time. Econometrica 85:1793–1846

    Google Scholar 

  • Bandi FM, Russell JR (2008) Microstructure noise, realized variance, and optimal sampling. Rev Econ Stud 75:339–369

    Google Scholar 

  • Barndorff-Nielsen OE, Graversen SE, Jacod J, Shephard N (2006) Limit theorems for bipower variation in financial econometrics. Economet Theor 22:677–719

    Google Scholar 

  • Barndorff-Nielsen OE, Hansen PR, Lunde A, Shephard N (2008) Designing realized Kernels to measure the ex post variation of equity prices in the presence of noise. Econometrica 76:1481–1536

    Google Scholar 

  • Barndorff-Nielsen OE, Shephard N (2002) Econometric analysis of realized volatility and its use in estimating stochastic volatility models. J Royal Stat Soc Ser B (Statistical Methodology) 64:253–280

    Google Scholar 

  • Bekaert G, Harvey C, Lundblad C (2007) Liquidity and expected returns: lessons from emerging markets. Rev Financ Stud 20:1783–1831

    Google Scholar 

  • Berman SM (1962) A law of large numbers for the maximum in a stationary gaussian sequence. Ann Math Stat 33:93–97

    Google Scholar 

  • Boswijk HP (2001) “Testing for a unit root with near-integrated volatility,” Tinbergen Institute Discussion Paper 01-077/4

  • Boswijk HP (2005) “Adaptive testing for a unit root with nonstationary volatility,” Tinbergen Institute Discussion Paper 2005/07

  • Calvo GA (1983) Staggered prices in a utility-maximizing framework. J Monet Econ 12:383–398

    Google Scholar 

  • Corradi V (2000) Reconsidering the continuous time limit of the GARCH(1,1) process. J Econom 96:145–153

    Google Scholar 

  • Da R, Xiu D (2021) When moving-average models meet high-frequency data: uniform inference on volatility. Econometrica 89:2787–2825

    Google Scholar 

  • Delattre S, Jacod J (1997) A central limit theorem for normalized functions of the increments of a diffusion process, in the presence of round-off errors. Bernoulli 3:1–28

    Google Scholar 

  • Engle RF (2000) The econometrics of ultra-high-frequency data. Econometrica 68:1–22

    Google Scholar 

  • Engle RF, Russell JR (1998) Autoregressive conditional duration: a new model for irregularly spaced transaction data. Econometrica 66:1127–1162

    Google Scholar 

  • Feller W (1951) Two singular diffusion problems. Ann Math 54:173–182

    Google Scholar 

  • Fleming J, Kirby C, Ostdiek B (2003) The economic value of volatility timing using “realized’’ volatility. J Financ Econ 67:473–509

    Google Scholar 

  • Ghysels E, Jasiak J (1998) “GARCH for Irregularly Spaced Financial Data: The ACD-GARCH Model,” Studies in Nonlinear Dynamics and Econometrics, 2

  • Hall P, Heyde CC (1980) Martingale limit theory and its application. Academic Press, New York

    Google Scholar 

  • Hansen PR, Lunde A (2006) Realized variance and market microstructure noise. J Business Econom Stat 24:127–161

    Google Scholar 

  • Heston SL (1993) A closed-form solution for options with stochastic volatility with applications to bond and currency options. Rev Financ Stud 6:327–343

    Google Scholar 

  • Jacod J (2018) Limit of random measures associated with the increments of a Brownian Semimartingale. J Financ Economet 16:526–569

    Google Scholar 

  • Jacod J, Shiryaev AN (1987) Limit theorems for stochastic processes. Springer-Verlag, New York

    Google Scholar 

  • Lesmond DA (2005) Liquidity of emerging markets. J Financ Econ 77:411–452

    Google Scholar 

  • Mankiw NG, Reis R (2002) Sticky information versus sticky prices: a proposal to replace the new Keynesian Phillips curve. Q J Econ 117:1295–1328

    Google Scholar 

  • Mykland PA, Zhang L (2006) ANOVA for diffusions and Itô processes. Ann Stat 34:1931–1963

    Google Scholar 

  • Naveau P (2003) Almost sure relative stability of the maximum of a stationary sequence. Adv Appl Probab 35:721–736

    Google Scholar 

  • Nelson DB (1990) ARCH models as diffusion approximations. J Econom 45:7–38

    Google Scholar 

  • Oomen RCA (2006) Properties of realized variance under alternative sampling schemes. J Business Econom Stat 24:219–237

    Google Scholar 

  • Phillips PCB, Yu J (2007) “Information loss in volatility measurement with flat price trading,” Cowles Foundation Discussion Paper No. 1598, Yale University

  • Protter P (2004) Stochastic integration and differential equations. Springer-Verlag, New York

    Google Scholar 

  • Reis R (2006) Inattentive producers. Rev Econ Stud 73:793–821

    Google Scholar 

  • Schilling MF (1990) The longest run of heads. College Math J 21:196–207

    Google Scholar 

  • Schmidt P (1976) Econometrics. Marcel Dekker, New York

    Google Scholar 

  • Yu J, Phillips PCB (2001) A Gaussian approach for continuous time models of the short-term interest rate. Economet J 4:210–224

    Google Scholar 

  • Zhang L (2006) Efficient estimation of stochastic volatility using noisy observations: A multi-scale approach. Bernoulli 12:1019–1043

    Google Scholar 

  • Zhang L, Mykland PA, Aït-Sahalia Y (2005) A tale of two time scales. J Am Stat Assoc 100:1394–1411

    Google Scholar 

  • Zhou B (1996) High-frequency data and volatility in foreign-exchange rates. J Business Econom Stat 14:45–52

    Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Peter C. B. Phillips.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

An early version of this paper was circulated as a working paper (Phillips and Yu 2007) but never published. This version brings the paper up to date and adds further analysis, simulations and discussion. Thanks go to the Editor and two referees for their comments and suggestions on the paper and to Neil Shephard, Jean Jacod and Sungbae An for helpful comments on aspects of the research. Phillips acknowledges support from a Lee Kong Chian Fellowship at SMU, the Kelly Fund at the University of Auckland, and the NSF under Grant No. SES 18-50860. Yu acknowledges that this research/project is supported by the Ministry of Education, Singapore, under its Academic Research Fund (AcRF) Tier 2 (Award Number MOE-T2EP402A20-0002). He also acknowledges the financial support from the Lee Foundation.

Supplementary Information

Below is the link to the electronic supplementary material.

Supplementary file 1 (pdf 255 KB)

Appendix

Appendix

Proof of Theorem 2.1

The specification of p(t) implies that

$$\begin{aligned} p_{i,m}= & {} p_{i,m}^{*}\xi _{i,m}+p_{i-1,m}(1-\xi _{i,m}), \end{aligned}$$
(1.39)
$$\begin{aligned} p_{i,m}-p_{i-1,m}= & {} (p_{i,m}^{*}-p_{i-1,m})\xi _{i,m}, \end{aligned}$$
(1.40)

and

$$\begin{aligned} p_{i,m}^{*}-p_{i,m}=(p_{i,m}^{*}-p_{i-1,m})(1-\xi _{i,m}). \end{aligned}$$
(1.41)

Taking conditional expectations of both sides of equation (1.39), we have

$$\begin{aligned} E(p_{i,m}|{\mathcal {F}}_{i-1,m})= & {} E(p_{i,m}^{*}|{\mathcal {F}}_{i-1,m})\pi +p_{i-1,m}(1-\pi ) \nonumber \\= & {} E(p_{i-1,m}^{*}|{\mathcal {F}}_{i-1,m})\pi +p_{i-1,m}(1-\pi ) \end{aligned}$$
(1.42)

To compute \(E(p_{i-1,m}^{*}|{\mathcal {F}}_{i-1,m})\), note that if \( p_{i-1,m}\ne p_{i-2,m}\), then \(p_{i-1,m}=p_{i-1,m}^{*}\) and hence \( E(p_{i-1,m}^{*}|{\mathcal {F}}_{i-1,m})=p_{i-1,m}\). If \(p_{i-1,m}=p_{i-2,m}\) but \(p_{i-2,m}\ne p_{i-3,m}\), then \(p_{i-2,m}=p_{i-2,m}^{*}\), \( p_{i-1,m}^{*}=p_{i-2,m}^{*}+\varepsilon _{i-1,m}\), and \({\mathcal {F}} _{i-1,m}={\mathcal {F}}_{i-2,m}\). Hence \(E(p_{i-1,m}^{*}|{\mathcal {F}} _{i-1,m})=E(p_{i-2,m}^{*}+\varepsilon _{i-1,m}|{\mathcal {F}} _{i-2,m})=p_{i-2,m}=p_{i-1,m}\). Similarly, if \(p_{i-1,m}=\cdots =p_{i-K,m}\) but \(p_{i-K,m}\ne p_{i-K=1,m}\), then \(p_{i-K,m}=p_{i-K,m}^{*}\), \( p_{i-1,m}^{*}=p_{i-K,m}^{*}+\varepsilon _{i-K+1,m}+\cdots +\varepsilon _{i-1,m}\) and \({\mathcal {F}}_{i-1,m}=\cdots ={\mathcal {F}}_{i-K,m}\) . Hence \(E(p_{i-1,m}^{*}|{\mathcal {F}}_{i-1,m})=E(p_{i-K,m}^{*}+\varepsilon _{i-K+1,m}+\cdots +\varepsilon _{i-1,m}|{\mathcal {F}} _{i-K,m})=p_{i-K,m}=p_{i-1,m}\). In general, we have \(E(p_{i-1,m}^{*}| {\mathcal {F}}_{i-1,m})=p_{i-1,m},\)and so \(E(p_{i,m}|{\mathcal {F}} _{i-1,m})=p_{i-1,m},\) as required. \(\square \)

Proof of Theorem 2.2

Let \(K_{i}\) be the run time of flat trading prior to \(t_{i,m}.\) As discussed in the paper, the maximum run time, \({\bar{K}} _{m}\), for a sequence of identical Bernoulli draws in a sample of size m has mean \(E\left( {\bar{K}}_{m}\right) =O\left( \log _{1/\pi }\left\{ m\left( 1-\pi \right) \right\} \right) =O\left( \frac{\log \left\{ m\left( 1-\pi \right) \right\} }{\log \frac{1}{\pi }}\right) \) and variance \(Var({\bar{K}} _{m})=\frac{\pi ^{2}}{6\log ^{2}\left( \frac{1}{\pi }\right) }.\) It follows that

$$\begin{aligned} {\bar{K}}_{m}=O_{p}\left( \log m\right) , \end{aligned}$$
(1.43)

and so each \(K_{i}\) is at most \(O_{p}\left( \log m\right) .\) It follows that in equi-spaced sampling when \(t_{i,m}=\frac{i}{m},\) we have a maximum grid size \(h_{{\bar{K}}_{m},m}=\frac{{\bar{K}}_{m}}{m}=O_{p}\left( \frac{\log m}{m} \right) \rightarrow 0\) as \(m\rightarrow \infty .\) In the case of a general grid \(\left\{ t_{i,m}\right\} ,\) the maximum grid size is bounded as follows for large m

$$\begin{aligned} h_{{\bar{K}}_{m},m}=\sup _{i}\left| t_{i,m}-t_{i-{\bar{K}}_{m},m}\right| \le \sup _{i}\left| t_{i,m}-t_{i-m^{\delta },m}\right| \text { for some }\delta >0. \end{aligned}$$

Hence, \(h_{{\bar{K}}_{m},m}\rightarrow 0\) if \(h_{m^{\delta },m}\rightarrow 0\) as \(m\rightarrow \infty \) for some \(\delta >0.\) In both cases, therefore, the grid size tends to zero and standard quadratic variation theory ensures that

$$\begin{aligned} \sum _{i=1}^{m}[p_{i,m}-p_{i-1,m}]^{2}\rightarrow _{p}\int _{0}^{1}\sigma ^{2}\left( t\right) \textrm{d}t. \end{aligned}$$
(1.44)

\(\square \)

The result may be proved directly by writing the left side of (1.44) in terms of the empirical quadratic variation of the efficient price \( RV^{(m)}(p^{*})=\sum _{i=1}^{m}(p_{i,m}^{*}-p_{i-1,m}^{*})^{2}\) and showing that the error converges in probability to zero. The derivation is useful in later arguments so it is given here. In particular, from Eq. (1.40), we have

$$\begin{aligned} \sum _{i=1}^{m}[p_{i,m}-p_{i-1,m}]^{2}= & {} \sum _{i=1}^{m}(p_{i,m}^{*}-p_{i-1,m})^{2}\xi _{i,m}^{2} \nonumber \\= & {} \sum _{i=1}^{m}(p_{i,m}^{*}-p_{i-1,m})^{2}E[\xi _{i,m}^{2}]+\sum _{i=1}^{m}(p_{i,m}^{*}-p_{i-1,m})^{2} \nonumber \\{} & {} \quad \times \left( \xi _{i,m}^{2}-E[\xi _{i,m}^{2}]\right) =\pi \sum _{i=1}^{m}(p_{i,m}^{*}-p_{i-1,m})^{2}\nonumber \\{} & {} \quad +\sum _{i=1}^{m}(p_{i,m}^{*}-p_{i-1,m})^{2}\left( \xi _{i,m}^{2}-E[\xi _{i,m}^{2}]\right) . \end{aligned}$$
(1.45)

Write the sum \(\sum _{i=1}^{m}(p_{i,m}^{*}-p_{i-1,m})^{2}\) in the first term above as follows

$$\begin{aligned} \sum _{i=1}^{m}(p_{i,m}^{*}-p_{i-1,m})^{2}= & {} \sum _{i=1}^{m}(p_{i,m}^{*}-p_{i-1,m}^{*}+p_{i-1,m}^{*}-p_{i-1,m})^{2} \nonumber \\= & {} RV^{(m)}(p^{*})+2\sum _{i=1}^{m}(p_{i,m}^{*}-p_{i-1,m}^{*})\left( p_{i-1,m}^{*}-p_{i-1,m}\right) \nonumber \\{} & {} +\sum _{i=1}^{m}(p_{i-1,m}^{*}-p_{i-1,m})^{2} \nonumber \\= & {} RV^{(m)}(p^{*})+2\sum _{i=1}^{m}\varepsilon _{i,m}\left( p_{i-1,m}^{*}-p_{i-1,m}\right) \nonumber \\{} & {} +\sum _{i=1}^{m}(p_{i-1,m}^{*}-p_{i-2,m})^{2}(1-\xi _{i-1,m})^{2} \nonumber \\= & {} RV^{(m)}(p^{*})+2\sum _{i=1}^{m}\varepsilon _{i,m}\left( p_{i-1,m}^{*}-p_{i-1,m}\right) \nonumber \\{} & {} +\sum _{i=1}^{m}(p_{i-1,m}^{*}-p_{i-2,m})^{2}(1-\pi ) \nonumber \\{} & {} +\sum _{i=1}^{m}(p_{i-1,m}^{*}-p_{i-2,m})^{2}\left\{ (1-\xi _{i-1,m})^{2}-(1-\pi )\right\} .\nonumber \\ \end{aligned}$$
(1.46)

Set \(A_{m}=\sum _{i=1}^{m}(p_{i,m}^{*}-p_{i-1,m})^{2}\). So \( A_{m-1}=A_{m}-(p_{m,m}^{*}-p_{m-1,m})^{2}\). Substituting out \(A_{m-1}\) in Eq. (1.46) gives

$$\begin{aligned} A_{m}= & {} RV^{(m)}(p^{*})+2\sum _{i=1}^{m}\varepsilon _{i,m}\left( p_{i-1,m}^{*}-p_{i-1,m}\right) +A_{m}(1-\pi ) \\{} & {} -(1-\pi )(p_{m,m}^{*}-p_{m-1,m})^{2}+\sum _{i=1}^{m}(p_{i-1,m}^{*}-p_{i-2,m})^{2}\\{} & {} \times \left\{ (1-\xi _{i-1,m})^{2}-(1-\pi )\right\} . \end{aligned}$$

Hence

$$\begin{aligned} A_{m}= & {} \frac{1}{\pi }RV^{(m)}(p^{*})+\frac{2}{\pi }\sum _{i=1}^{m} \varepsilon _{i,m}\left( p_{i-1,m}^{*}-p_{i-1,m}\right) \nonumber \\{} & {} -\frac{1-\pi }{\pi }(p_{m,m}^{*}-p_{m-1,m})^{2}\nonumber \\{} & {} +\frac{1}{\pi } \sum _{i=1}^{m}(p_{i-1,m}^{*}-p_{i-2,m})^{2}\left\{ (1-\xi _{i-1,m})^{2}-(1-\pi )\right\} . \end{aligned}$$
(1.47)

Substituting (1.47) into (1.45) we have

$$\begin{aligned} \sum _{i=1}^{m}[p_{i,m}-p_{i-1,m}]^{2}= & {} RV^{(m)}(p^{*})+2\sum _{i=1}^{m}\varepsilon _{i,m}\left( p_{i-1,m}^{*}-p_{i-1,m}\right) \nonumber \\{} & {} -(1-\pi )(p_{m,m}^{*}-p_{m-1,m})^{2} \nonumber \\{} & {} +\sum _{i=1}^{m}(p_{i-1,m}^{*}-p_{i-2,m})^{2}\left\{ (1-\xi _{i-1,m})^{2}-(1-\pi )\right\} \nonumber \\{} & {} +\sum _{i=1}^{m}(p_{i,m}^{*}-p_{i-1,m})^{2}\left( \xi _{i,m}^{2}-\pi \right) \nonumber \\= & {} RV^{(m)}(p^{*})+\underset{A}{\underbrace{2\sum _{i=1}^{m}\varepsilon _{i,m}\left( p_{i-1,m}^{*}-p_{i-1,m}\right) }}\nonumber \\{} & {} \underset{B}{\underbrace{ -(1-\pi )(p_{m,m}^{*}-p_{m-1,m})^{2}}} \nonumber \\{} & {} \underset{C}{\underbrace{-2\sum _{i=1}^{m}(p_{i,m}^{*}-p_{i-1,m})^{2}\xi _{i,m}(1-\xi _{i,m})}} \nonumber \\= & {} RV^{(m)}(p^{*})+A+B+C. \end{aligned}$$
(1.48)

Since \(\xi _{i,m}\) is a Bernoulli variable, \(\xi _{i,m}\left( 1-\xi _{i,m}\right) =0\) a.s.,  and so \(C=0\). Consider term B. Note that for some duration \(K_{m}\ge 1\) and for which at most \(K_{m}=O_{p}(\log m)\) we have

$$\begin{aligned} p_{m,m}^{*}-p_{m-1,m}=p_{m,m}^{*}-p_{m-K_{m},m}^{*}=\int _{t_{m-K_{m},m}}^{1}\sigma (s)\textrm{d}B(s)=O_{p}\left( \sqrt{h_{{\bar{K}}_{m},m} }\right) , \end{aligned}$$

so that

$$\begin{aligned} B=-(1-\pi )(p_{m,m}^{*}-p_{m-1,m})^{2}=-(1-\pi )O_{p}\left( h_{{\bar{K}} _{m},m}\right) =o_{p}(1), \end{aligned}$$
(1.49)

For term A, note that \(\varepsilon _{i,m}=\int _{t_{i-1,m}}^{t_{i,m}}\sigma (s)\textrm{d}B(s)\) and

$$\begin{aligned} p_{i-1,m}^{*}-p_{i-1,m}= & {} (p_{i-1,m}^{*}-p_{i-2,m})(1-\xi _{i-1,m})\nonumber \\= & {} (p_{i-1,m}^{*}-p_{i-K_{i-1},m}^{*})(1-\xi _{i-1,m}), \end{aligned}$$
(1.50)

for some \(K_{i-1}\ge 2\) and where \(K_{i-1}\) is at most of \(O_{p}(\log m).\) Then

$$\begin{aligned} \sum _{i=1}^{m}\varepsilon _{i,m}\left( p_{i-1,m}^{*}-p_{i-1,m}\right) =\sum _{i=1}^{m}\varepsilon _{i,m}(p_{i-1,m}^{*}-p_{i-K_{i-1},m}^{*})(1-\xi _{i-1,m}). \end{aligned}$$

Now \(\varepsilon _{i,m}\) is independent of \(\xi _{i-1,m},\) and \( E(\varepsilon _{i,m})=0\) and \(Var(\varepsilon _{i,m})\rightarrow 0\), as \( m\rightarrow \infty ,\) because \(h_{{\bar{K}}_{m},m}=o\left( 1\right) ,\) while \( \sum _{i=1}^{m}(p_{i-1,m}^{*}-p_{i-K_{i-1},m}^{*})^{2}(1-\xi _{i-1,m})^{2}\) is bounded as \(m\rightarrow \infty .\) It follows that \( A=o_{p}(1).\) Thus, \(\sum _{i=1}^{m}[p_{i,m}-p_{i-1,m}]^{2}=RV^{(m)}(p^{*})+o_{p}(1),\) as required.

Proof Theorem 2.3

From (1.48) we have

$$\begin{aligned}{} & {} \sqrt{m}\left\{ \sum _{i=1}^{m}[p_{i,m}-p_{i-1,m}]^{2}-IV\right\} \nonumber \\{} & {} \quad =\sqrt{m} \left\{ RV^{(m)}(p^{*})-IV\right\} \nonumber \\{} & {} \qquad +\sqrt{m}2\sum _{i=1}^{m}\varepsilon _{i,m}\left( p_{i-1,m}^{*}-p_{i-1,m}\right) \nonumber \\{} & {} \qquad -\sqrt{m}(1-\pi )(p_{m,m}^{*}-p_{m-1,m})^{2} \nonumber \\{} & {} \qquad +\sqrt{m}\sum _{i=1}^{m}(p_{i-1,m}^{*}-p_{i-2,m})^{2}\left\{ (1-\xi _{i-1,m})^{2}-(1-\pi )\right\} \nonumber \\{} & {} \qquad +\sqrt{m}\sum _{i=1}^{m}(p_{i,m}^{*}-p_{i-1,m})^{2}\left( \xi _{i,m}^{2}-\pi \right) \nonumber \\{} & {} \quad =\sqrt{m}\left\{ RV^{(m)}(p^{*})-IV\right\} +\sqrt{m}A+\sqrt{m}B+\sqrt{ m}C \nonumber \\{} & {} \quad =\sqrt{m}\left\{ RV^{(m)}(p^{*})-IV\right\} +\sqrt{m}A+\sqrt{m}B, \end{aligned}$$
(1.51)

since \(C=0,\) a.s. . Standard theory (Barndorff-Nielsen and Shephard 2002; Barndorff-Nielsen et al. 2006; Jacod 2018) gives the CLT

$$\begin{aligned} \sqrt{m}\left\{ RV^{(m)}(p^{*})-IV\right\} \overset{d}{\rightarrow }\ {\mathcal {N}}\left( 0,2\int _{0}^{1}\sigma ^{4}(t)\textrm{d}t\right) , \end{aligned}$$
(1.52)

stably as \(m\rightarrow \infty .\) We now study the asymptotic behavior of \( \sqrt{m}A,\) and \(\sqrt{m}B\).

For term \(\sqrt{m}A\), from (1.50) we get

$$\begin{aligned} \sqrt{m}A= & {} \sqrt{m}2\sum _{i=1}^{m}\varepsilon _{i,m}(p_{i-1,m}^{*}-p_{i-K_{i-1},m}^{*})(1-\xi _{i-1,m}) \\= & {} 2\sqrt{m}\sum _{i=1}^{m}\int _{t_{i-1,m}}^{t_{i,m}}\sigma (s)\textrm{d}B(s)(p_{i-1,m}^{*}-p_{i-K_{i-1},m}^{*})(1-\xi _{i-1,m}) \\= & {} 2\sqrt{m}\sum _{i=1}^{m}\nu _{i,m}(p_{i-1,m}^{*}-p_{i-K_{i-1},m}^{*}), \end{aligned}$$

where \(\nu _{i,m}=\int _{t_{i-1,m}}^{t_{i,m}}\sigma (s)\textrm{d}B(s)(1-\xi _{i-1,m})\) is uncorrelated with \((p_{i-1,m}^{*}-p_{i-K_{i-1},m}^{*}),\) because of the martingale property, and has mean 0 and conditional variance \( \left( 1-\pi \right) m\int _{t_{i-1,m}}^{t_{i,m}}\sigma (s)^{2}\textrm{d}s.\) So \(\sqrt{ m}A\) is a martingale with conditional variance

$$\begin{aligned} m\left( 1-\pi \right) \sum _{i=1}^{m}(p_{i-1,m}^{*}-p_{i-K_{i-1},m}^{*})^{2}\int _{t_{i-1,m}}^{t_{i,m}}\sigma (s)^{2}\textrm{d}s. \end{aligned}$$

By stochastic Taylor series expansion we have

$$\begin{aligned}{} & {} p_{i-1,m}^{*}-p_{i-K_{i-1},m}^{*}\nonumber \\{} & {} \quad =\int _{t_{i-K_{i-1},m}}^{t_{i-1,m}}\sigma (s)\textrm{d}B(s) \nonumber \\{} & {} \quad =\left\{ \sigma \left( t_{i-K_{i-1},m}\right) +O_{p}\left( \sqrt{K_{i-1}h} \right) \right\} \left( B\left( t_{i-1,m}\right) -B\left( t_{i-K_{i-1},m}\right) \right) \nonumber \\{} & {} \quad =\sigma \left( t_{i-K_{i-1},m}\right) \left( B\left( t_{i-1,m}\right) -B\left( t_{i-K_{i-1},m}\right) \right) +O_{p}\left( K_{i-1}h\right) \nonumber \\{} & {} \quad =\sigma \left( \frac{i-1}{m}\right) \left( B\left( \frac{i-1}{m}\right) -B\left( \frac{i-K_{i-1}}{m}\right) \right) +O_{p}\left( \frac{K_{i-1}}{m} \right) \end{aligned}$$
(1.53)

on the equispaced grid \(\{t_{i,m}=\frac{i}{m}:i=0,\ldots ,m\}\) with \(h=\frac{1}{ m}.\) Then

$$\begin{aligned}{} & {} m\sum _{i=1}^{m}(p_{i-1,m}^{*}-p_{i-K_{i-1},m}^{*})^{2}\int _{t_{i-1,m}}^{t_{i,m}}\sigma (s)^{2}\textrm{d}s \\{} & {} \quad =m\sum _{i=1}^{m}\left\{ \sigma \left( \frac{i-1}{m}\right) \left( B\left( \frac{i-1}{m}\right) -B\left( \frac{i-K_{i-1}}{m}\right) \right) +O_{p}\left( \frac{K_{i-1}}{m}\right) \right\} ^{2} \\{} & {} \qquad \times \left\{ \sigma \left( \frac{i-1}{m}\right) ^{2}+O_{p}\left( \frac{1 }{\sqrt{m}}\right) \right\} \frac{1}{m} \\{} & {} \quad =\sum _{i=1}^{m}\left\{ \sigma \left( \frac{i-1}{m}\right) ^{4}\left( B\left( \frac{i-1}{m}\right) -B\left( \frac{i-K_{i-1}}{m}\right) \right) ^{2} \left[ 1+O_{p}\left( \frac{K_{i-1}}{m}\right) \right] \right\} \\{} & {} \quad =\sum _{i=1}^{m}\left\{ \sigma \left( \frac{i-1}{m}\right) ^{4}\frac{ K_{i-1}-1}{m}\left[ 1+O_{p}\left( \frac{K_{i-1}}{m}\right) \right] \right\} \\{} & {} \quad \rightarrow _{p}\left( \int _{0}^{1}\sigma ^{4}(t)\textrm{d}t\right) E(K_{i-1}-1). \end{aligned}$$

It follows by the martingale central limit theorem (Hall and Heyde 1980, Theorem 3.2) that

$$\begin{aligned} \sqrt{m}A\overset{d}{\rightarrow }A_{\infty }=_{d}2\times \mathcal{M}\mathcal{N}\left( 0,\left( 1-\pi \right) \left( \int _{0}^{1}\sigma ^{4}(t)\textrm{d}t\right) E(K_{i-1}-1)\right) , \end{aligned}$$

stably in the sense that \(\left\{ Z,\sqrt{m}A\right\} \overset{d}{ \rightarrow }\left\{ Z,A_{\infty }\right\} \) for \(Z=\int _{0}^{1}\sigma ^{4}(t)\textrm{d}t.\) Observe that

$$\begin{aligned} K_{i-1}-2=\left\{ \begin{array}{cl} 0 &{} \text { with probability }\pi \\ 1 &{} \text { with probability }\pi (1-\pi ) \\ 2 &{} \text { with probability }\pi (1-\pi )^{2} \\ \vdots &{} \end{array} \right. , \end{aligned}$$

so that \(E(K_{i-1}-2)=\pi (1-\pi )+2\pi (1-\pi )^{2}+\cdots =\frac{1-\pi }{ \pi },\) which implies that \(E(K_{i-1}-1)=\frac{1}{\pi }\). Thus

$$\begin{aligned} \sqrt{m}A\overset{d}{\rightarrow }\mathcal{M}\mathcal{N}\left( 0,4\frac{1-\pi }{\pi } \int _{0}^{1}\sigma ^{4}(t)\textrm{d}t\right) , \end{aligned}$$
(1.54)

stably.

Next consider term \(\sqrt{m}B\). From (1.49) we have

$$\begin{aligned} \sqrt{m}B=-\sqrt{m}O_{p}\left( \frac{\log m}{m}\right) (1-\pi )=o_{p}(1). \end{aligned}$$
(1.55)

Observe that the components of \(\sqrt{m}A\) involve the product \(\varepsilon _{i,m}(p_{i-1,m}^{*}-p_{i-K_{i-1},m}^{*})(1-\xi _{i-1,m})\) whereas \( \sqrt{m}\left\{ RV^{(m)}(p^{*})-IV\right\} \) is a centred quadratic in the increments \(\varepsilon _{i,m}=p_{i,m}^{*}-p_{i-1,m}^{*}\), so that \(\sqrt{m}A\) and \(\sqrt{m}\left\{ RV^{(m)}(p^{*})-IV\right\} \) are asymptotically uncorrelated. By combining (1.54) and (1.52), it follows that

$$\begin{aligned} \sqrt{m}\left\{ RV^{(m)}(p)-IV\right\} \overset{d}{\rightarrow }\mathcal{M}\mathcal{N}\left( 0, \frac{4-2\pi }{\pi }\int _{0}^{1}\sigma ^{4}(t)\textrm{d}t\right) , \end{aligned}$$

giving the required result. \(\square \)

Proof of Lemma 2.4

From Eq. (1.40), we have

$$\begin{aligned} m\sum _{i=1}^{m}[p_{i,m}-p_{i-1,m}]^{4}= & {} \pi m\sum _{i=1}^{m}(p_{i,m}^{*}-p_{i-1,m})^{4}\nonumber \\{} & {} +m\sum _{i=1}^{m}(p_{i,m}^{*}-p_{i-1,m})^{4}\left( \xi _{i,m}^{4}-E[\xi _{i,m}^{4}]\right) . \end{aligned}$$
(1.56)

Consider \(\sum _{i=1}^{m}(p_{i,m}^{*}-p_{i-1,m})^{4}\) in the first term, giving

$$\begin{aligned}{} & {} \sum _{i=1}^{m}(p_{i,m}^{*}-p_{i-1,m})^{4} \nonumber \\{} & {} \quad =\sum (p_{i,m}^{*}-p_{i-1,m}^{*}+p_{i-1,m}^{*}-p_{i-1,m})^{4} \nonumber \\{} & {} \quad =\sum _{i=1}^{m}\left( p_{i,m}^{*}-p_{i-1,m}^{*}\right) ^{4}+4\sum _{i=1}^{m}\left( p_{i,m}^{*}-p_{i-1,m}^{*}\right) ^{3}\left( p_{i-1,m}^{*}-p_{i-1,m}\right) \nonumber \\{} & {} \quad \quad +\,6\sum _{i=1}^{m}\left( p_{i,m}^{*}-p_{i-1,m}^{*}\right) ^{2}\left( p_{i-1,m}^{*}-p_{i-1,m}\right) ^{2} \nonumber \\{} & {} \quad \quad +\,4\sum _{i=1}^{m}\left( p_{i,m}^{*}-p_{i-1,m}^{*}\right) \left( p_{i-1,m}^{*}-p_{i-1,m}\right) ^{3}+\sum _{i=1}^{m}\left( p_{i-1,m}^{*}-p_{i-1,m}\right) ^{4} \nonumber \\{} & {} \quad =\sum _{i=1}^{m}\left( p_{i,m}^{*}-p_{i-1,m}^{*}\right) ^{4}+4\sum _{i=1}^{m}\left( p_{i,m}^{*}-p_{i-1,m}^{*}\right) ^{3}\left( p_{i-1,m}^{*}-p_{i-1,m}\right) \nonumber \\{} & {} \quad \quad +\,6\sum _{i=1}^{m}\left( p_{i,m}^{*}-p_{i-1,m}^{*}\right) ^{2}\left( p_{i-1,m}^{*}-p_{i-1,m}\right) ^{2} \nonumber \\{} & {} \quad \quad +\,4\sum _{i=1}^{m}\left( p_{i,m}^{*}-p_{i-1,m}^{*}\right) \left( p_{i-1,m}^{*}-p_{i-1,m}\right) ^{3}\nonumber \\{} & {} \quad \quad +m\sum _{i=1}^{m}\left( p_{i-1,m}^{*}-p_{i-2,m}\right) ^{4}(1-\pi ) \nonumber \\{} & {} \quad \quad +\,\sum _{i=1}^{m}\left( p_{i-1,m}^{*}-p_{i-2,m}\right) ^{4}((1-\xi _{i-1,m})^{4}-(1-\pi )). \end{aligned}$$
(1.57)

Set \(B_{m}=\sum _{i=1}^{m}(p_{i,m}^{*}-p_{i-1,m})^{4}\). So \( B_{m-1}=B_{m}-(p_{m,m}^{*}-p_{m-1,m})^{4}\). Substituting out \(B_{m-1}\) in Eq. (1.57) and solving for \(B_{m}\), we get

$$\begin{aligned} B_{m}= & {} \frac{1}{\pi }\sum _{i=1}^{m}\left( p_{i,m}^{*}-p_{i-1,m}^{*}\right) ^{4}+\frac{4}{\pi }\sum _{i=1}^{m}\left( p_{i,m}^{*}-p_{i-1,m}^{*}\right) ^{3}\left( p_{i-1,m}^{*}-p_{i-1,m}\right) \nonumber \\{} & {} +\,\frac{6}{\pi }\sum _{i=1}^{m}\left( p_{i,m}^{*}-p_{i-1,m}^{*}\right) ^{2}\left( p_{i-1,m}^{*}-p_{i-1,m}\right) ^{2} \nonumber \\{} & {} +\,\frac{4}{\pi }\sum _{i=1}^{m}\left( p_{i,m}^{*}-p_{i-1,m}^{*}\right) \left( p_{i-1,m}^{*}-p_{i-1,m}\right) ^{3} \nonumber \\{} & {} -\,\frac{1-\pi }{\pi }(p_{m,m}^{*}-p_{m-1,m})^{2}\nonumber \\{} & {} +\,\frac{1}{\pi } \sum _{i=1}^{m}(p_{i-1,m}^{*}-p_{i-2,m})^{4}\left\{ (1-\xi _{i-1,m})^{4}-(1-\pi )\right\} . \end{aligned}$$
(1.58)

Substituting (1.58) into (1.56) we have

$$\begin{aligned} m\sum [p_{i,m}-p_{i-1,m}]^{4}= & {} m\sum _{i=1}^{m}\left( p_{i,m}^{*}-p_{i-1,m}^{*}\right) ^{4}\nonumber \\{} & {} +\,4m\sum _{i=1}^{m}\left( p_{i,m}^{*}-p_{i-1,m}^{*}\right) ^{3}\left( p_{i-1,m}^{*}-p_{i-1,m}\right) \nonumber \\{} & {} +\,6m\sum _{i=1}^{m}\left( p_{i,m}^{*}-p_{i-1,m}^{*}\right) ^{2}\left( p_{i-1,m}^{*}-p_{i-1,m}\right) ^{2} \nonumber \\{} & {} +\,4m\sum _{i=1}^{m}\left( p_{i,m}^{*}-p_{i-1,m}^{*}\right) \left( p_{i-1,m}^{*}-p_{i-1,m}\right) ^{3}\nonumber \\{} & {} -\,(1-\pi )(p_{m,m}^{*}-p_{m-1,m})^{2} \nonumber \\{} & {} +\,m\sum (p_{i-1,m}^{*}-p_{i-2,m})^{4}\left\{ (1-\xi _{i-1,m})^{4}-(1-\pi )\right\} \nonumber \\{} & {} +\,m\sum _{i=1}^{m}(p_{i,m}^{*}-p_{i-1,m})^{4}\left( \xi _{i,m}^{4}-\pi \right) \nonumber \\= & {} m\sum _{i=1}^{m}\left( p_{i,m}^{*}-p_{i-1,m}^{*}\right) ^{4}\nonumber \\{} & {} +\, \underset{A}{\underbrace{4m\sum _{i=1}^{m}\left( p_{i,m}^{*}-p_{i-1,m}^{*}\right) ^{3}\left( p_{i-1,m}^{*}-p_{i-1,m}\right) }} \nonumber \\{} & {} +\,\underset{B}{\underbrace{6m\sum _{i=1}^{m}\left( p_{i,m}^{*}-p_{i-1,m}^{*}\right) ^{2}\left( p_{i-1,m}^{*}-p_{i-1,m}\right) ^{2} }} \nonumber \\{} & {} +\,\underset{C}{\underbrace{4m\sum _{i=1}^{m}\left( p_{i,m}^{*}-p_{i-1,m}^{*}\right) \left( p_{i-1,m}^{*}-p_{i-1,m}\right) ^{3}}} \nonumber \\{} & {} \underset{D}{\underbrace{-(1-\pi )m(p_{m,m}^{*}-p_{m-1,m})^{4}}} \nonumber \\{} & {} +\,\underset{E}{\underbrace{m\sum (p_{i,m}^{*}-p_{i-1,m})^{4}(1-\xi _{i,m})^{2}\xi _{i,m}^{2}}} \nonumber \\= & {} m\sum _{i=1}^{m}\left( p_{i,m}^{*}-p_{i-1,m}^{*}\right) ^{4}+A+B+C+D+E. \nonumber \\ \end{aligned}$$
(1.59)

As in (1.53) we have

$$\begin{aligned} p_{i,m}^{*}-p_{i-1,m}^{*}= & {} \sigma \left( t_{i-1,m}\right) \left( B\left( t_{i,m}\right) -B\left( t_{i,m}\right) \right) +O_{p}\left( \frac{1}{ m}\right) \nonumber \\= & {} \sigma \left( t_{i-1,m}\right) \frac{\epsilon _{i,m}}{\sqrt{m}} +O_{p}\left( \frac{1}{m}\right) , \end{aligned}$$
(1.60)

where \(\epsilon _{i,m}\) is iid \({\mathcal {N}}\left( 0,1\right) .\) Hence

$$\begin{aligned} m\sum _{i=1}^{m}\left( p_{i,m}^{*}-p_{i-1,m}^{*}\right) ^{4}= & {} \sum _{i=1}^{m}\sigma \left( t_{i-1,m}\right) ^{4}\frac{\epsilon _{i,m}^{4} }{m}+O_{p}\left( \frac{1}{m^{1/2}}\right) \nonumber \\= & {} \sum _{i=1}^{m}\sigma \left( t_{i-1,m}\right) ^{4}\frac{E\left( \epsilon _{i,m}^{4}\right) }{m}+O_{p}\left( \frac{1}{m^{1/2}}\right) \nonumber \\{} & {} \rightarrow _{p}\,3\int _{0}^{1}\sigma ^{4}(t)\textrm{d}t. \end{aligned}$$
(1.61)

Hence

$$\begin{aligned} \frac{2}{3}m\sum _{i=1}^{m}\left( p_{i,m}^{*}-p_{i-1,m}^{*}\right) ^{4}\rightarrow _{p}2\int _{0}^{1}\sigma ^{4}(t)\textrm{d}t. \end{aligned}$$
(1.62)

This corresponds with the result obtained in Barndorff-Nielsen and Shephard (2002).

We now consider the limit behavior of terms ABCD,  and E. First, for term E, since \(\left( 1-\xi _{i,m}\right) ^{2}\xi _{i,m}^{2}=0\) almost surely, \(E=0\). Second, for term D, note that

$$\begin{aligned} D=-(1-\pi )m(p_{m,m}^{*}-p_{m-1,m})^{4}=-(1-\pi )m\times O_{p}\left( \frac{\log ^{2}m}{m^{2}}\right) =o_{p}(1). \end{aligned}$$

Next, consider term A, viz.,

$$\begin{aligned}{} & {} m\sum _{i=1}^{m}\left( p_{i,m}^{*}-p_{i-1,m}^{*}\right) ^{3}\left( p_{i-1,m}^{*}-p_{i-1,m}\right) \nonumber \\{} & {} \quad =m\sum _{i=1}^{m}\left( p_{i,m}^{*}-p_{i-1,m}^{*}\right) ^{3}\left( p_{i-1,m}^{*}-p_{i-2,m}\right) (1-\xi _{i-1,m}) \nonumber \\{} & {} \quad =m\sum _{i=1}^{m}\left\{ \sigma \left( t_{i-1,m}\right) \frac{\epsilon _{i,m}}{\sqrt{m}}+O_{p}\left( \frac{1}{m}\right) \right\} ^{3}\left( p_{i-1,m}^{*}-p_{i-K_{i-1},m}^{*}\right) (1-\xi _{i-1,m}) \nonumber \\{} & {} \quad \text {for some }K_{i-1}=O_{p}(\log m) \nonumber \\{} & {} \quad =\frac{1}{\sqrt{m}}\sum _{i=1}^{m}\sigma ^{3}\left( t_{i-1,m}\right) \epsilon _{i,m}^{3}\left( p_{i-1,m}^{*}-p_{i-K_{i-1},m}^{*}\right) (1-\xi _{i-1,m})+o_{p}\left( 1\right) . \end{aligned}$$
(1.63)

The component \(\sigma ^{3}\left( t_{i-1,m}\right) \epsilon _{i,m}^{3}\left( p_{i-1,m}^{*}-p_{i-K_{i-1},m}^{*}\right) (1-\xi _{i-1,m})\) in the sum (1.63) has mean zero and conditional variance \(E\left[ \epsilon _{i,m}^{6}\right] \left( 1-\pi \right) \sigma ^{6}\left( t_{i-1,m}\right) \left( p_{i-1,m}^{*}-p_{i-K_{i-1},m}^{*}\right) ^{2}=O_{p}\left( \frac{\log m}{m}\right) \) since from (1.53)

$$\begin{aligned} p_{i-1,m}^{*}-p_{i-K_{i-1},m}^{*}= & {} \sigma \left( \frac{i-1}{m} \right) \left( B\left( \frac{i-1}{m}\right) -B\left( \frac{i-K_{i-1}}{m} \right) \right) +O_{p}\left( \frac{K_{i-1}}{m}\right) \nonumber \\= & {} \sigma \left( \frac{i-1}{m}\right) \frac{\eta _{K_{i-1}}}{\sqrt{m}} +O_{p}\left( \frac{K_{i-1}}{m}\right) =O_{p}\left( \sqrt{\frac{\log m}{m}} \right) , \end{aligned}$$
(1.64)

where

$$\begin{aligned} \eta _{K_{i-1}}:=B\left( \frac{i-1}{m}\right) -B\left( \frac{i-K_{i-1}}{m} \right) =O_{p}\left( \sqrt{\log m}\right) . \end{aligned}$$

It follows that

$$\begin{aligned} E\left\{ \frac{1}{\sqrt{m}}\sum _{i=1}^{m}\sigma ^{3}\left( t_{i-1,m}\right) \epsilon _{i,m}^{3}\left( p_{i-1,m}^{*}-p_{i-K_{i-1},m}^{*}\right) (1-\xi _{i-1,m})\right\} ^{2}=O_{p}\left( \frac{\log m}{m}\right) , \end{aligned}$$

and so \(A=o_{p}(1)\). Similarly, for term C, \(4m\sum \left( p_{i,m}^{*}-p_{i-1,m}^{*}\right) \left( p_{i-1,m}^{*}-p_{i-1,m}\right) ^{3}\) is \(o_{p}(1)\).

Next, consider term B. Using (1.40) and (1.64), we have

$$\begin{aligned}{} & {} m\sum _{i=1}^{m}\left( p_{i,m}^{*}-p_{i-1,m}^{*}\right) ^{2}\left( p_{i-1,m}^{*}-p_{i-1,m}\right) ^{2} \nonumber \\{} & {} \quad =m\sum _{i=1}^{m}\left( p_{i,m}^{*}-p_{i-1,m}^{*}\right) ^{2}\left( p_{i-1,m}^{*}-p_{i-2,m}\right) ^{2}(1-\xi _{i-1,m})^{2} \nonumber \\{} & {} \quad =m\sum _{i=1}^{m}\left( p_{i,m}^{*}-p_{i-1,m}^{*}\right) ^{2}\left( p_{i-1,m}^{*}-p_{i-K_{i-1},m}^{*}\right) ^{2}(1-\xi _{i-1,m})^{2} \nonumber \\{} & {} \qquad \text { for some }K_{i-1}=O_{p}(\log m) \nonumber \\{} & {} \quad =m\sum _{i=1}^{m}\left\{ \sigma \left( t_{i-1,m}\right) \frac{\epsilon _{i,m}}{\sqrt{m}}+O_{p}\left( \frac{1}{m}\right) \right\} ^{2} \nonumber \\{} & {} \qquad \times \left( \sigma \left( t_{i-1,m}\right) \frac{\eta _{K_{i-1}}}{\sqrt{m}}+O_{p}\left( \frac{ \log m}{m}\right) \right) ^{2}(1-\xi _{i,m})^{2} \nonumber \\{} & {} \quad =\frac{1}{m}\sum _{i=1}^{m}\sigma ^{2}(t_{i,m})\epsilon _{i,m}^{2}\sigma \left( t_{i-1,m}\right) ^{2}\eta _{K_{i-1}}^{2}(1-\xi _{i-1,m})^{2}+O_{p}\left( \sqrt{\frac{\log m}{m}}\right) \nonumber \\{} & {} \quad =\frac{1}{m}\sum _{i=1}^{m}\sigma ^{4}(t_{i,m})E\left[ \epsilon _{i,m}^{2}\eta _{K_{i-1}}^{2}(1-\xi _{i-1,m})^{2}\right] \nonumber \\{} & {} \qquad +\frac{1}{m}\sum _{i=1}^{m}\sigma ^{4}(t_{i,m})\left\{ \epsilon _{i,m}^{2}\eta _{K_{i-1}}^{2}(1-\xi _{i-1,m})^{2}-E\left[ \epsilon _{i,m}^{2}\eta _{K_{i-1}}^{2}(1-\xi _{i-1,m})^{2}\right] \right\} \nonumber \\{} & {} \qquad +O_{p}\left( \sqrt{\frac{\log m}{m}}\right) \nonumber \\{} & {} \quad =\frac{1}{m}\sum _{i=1}^{m}\sigma ^{4}\left( t_{i,m}\right) (1-\pi )E(K_{i-1}-1)+O_{p}\left( \frac{\log m}{\sqrt{m}}\right) \nonumber \\{} & {} \quad \rightarrow _{p} \,\left( \int _{0}^{1}\sigma ^{4}(t)\textrm{d}t\right) (1-\pi )E(K_{i-1}-1)=\frac{(1-\pi )}{\pi }\int _{0}^{1}\sigma ^{4}(t)\textrm{d}t, \end{aligned}$$
(1.65)

since \(\epsilon _{i,m},\) \(\eta _{K_{i-1}},\) and \(\xi _{i-1,m}\) are independent. Therefore,

$$\begin{aligned} m\sum \left[ p_{i,m}-p_{i-1,m}\right] ^{4}= & {} 3\int _{0}^{1}\sigma ^{4}(t)\textrm{d}t+6 \frac{(1-\pi )}{\pi }\int _{0}^{1}\sigma ^{4}(t)\textrm{d}t+o_{p}(1) \\= & {} \frac{6-3\pi }{\pi }\int _{0}^{1}\sigma ^{4}(t)\textrm{d}t+o_{p}(1), \end{aligned}$$

leading to the required result. \(\square \)

Proof Theorem 2.5

This follows directly from Lemma 2.4 and Theorem 2.3. \(\square \)

Proof Theorem 3.1

This proof is given in the Online Supplement to the paper. \(\square \)

Proof Theorem 4.1

As in Zhang et al. (2005) we set \(K=cm^{2/3}\) for some constant \( c>0.\) Then

$$\begin{aligned} \frac{K}{{\bar{m}}}=\frac{K^{2}}{m}=c^{2}m^{1/3}. \end{aligned}$$

The two time scale measures involving the efficient price series are denoted by \(\left[ p^{*}\right] ^{(avg)}\) and \(\left[ p^{*}\right] ^{ {\mathcal {G}}_{m}},\) and the corresponding measures using the actual price series are denoted by \(\left[ p\right] ^{(avg)}\) and \(\left[ p\right] ^{ {\mathcal {G}}_{m}}.\)

The two time scale estimator is \(\widehat{\left[ p^{*}\right] }=\left[ p \right] ^{(avg)}-\frac{{\bar{m}}}{J_{m}}\left[ p\right] ^{{\mathcal {G}}_{m}}.\) Decompose the estimation error as

$$\begin{aligned} \widehat{\left[ p^{*}\right] }-\int _{0}^{1}\sigma ^{2}(t)\textrm{d}t=\left( \widehat{\left[ p^{*}\right] }-\left[ p^{*}\right] ^{(avg)}\right) +\left( \left[ p^{*}\right] ^{(avg)}-\int _{0}^{1}\sigma ^{2}(t)\textrm{d}t\right) . \end{aligned}$$
(1.66)

From the proof of Theorem A.1 and equation (56) in Zhang et al. (2005), we have

$$\begin{aligned} \sqrt{\frac{K}{{\bar{m}}}}\left\{ \widehat{\left[ p^{*}\right] }-\left[ p^{*}\right] ^{(avg)}\right\} \Rightarrow {\mathcal {N}}\left( 0,8\sigma _{u}^{4}\right) , \end{aligned}$$
(1.67)

and their proof applies in the present case with only minor notational changes to account for random \(\tau _{j}\) sampling under Assumption D2 (ii).

The second component of (1.66) is the discretization effect and is the same as that in Zhang et al. (2005), after allowing for the fact that we have random stopping times and adjusting accordingly, as we do below. Theorems 2 and 3 of Zhang et al. (2005) give the following limit theory

$$\begin{aligned} \left( \frac{m}{K}\right) ^{1/2}\left( \left[ p^{*}\right] ^{(avg)}-\int _{0}^{1}\sigma ^{2}(t)\textrm{d}t\right) \Rightarrow \mathcal{M}\mathcal{N}\left( 0,\eta ^{2}\right) , \end{aligned}$$
(1.68)

where \(\eta ^{2}\) is the probability limit of the quantity

$$\begin{aligned} \eta _{m}^{2}=\sum _{j=1}^{J_{m}}\sigma ^{4}\left( \tau _{j}\right) h_{j}\left( \tau _{j}-\tau _{j-1}\right) , \end{aligned}$$
(1.69)

where

$$\begin{aligned} h_{j}=\frac{4}{K\overline{\Delta \tau }}\sum _{i=1}^{\left( K-1\right) \wedge j}\left( 1-\frac{i}{K}\right) ^{2}\left( \tau _{j-i}-\tau _{j-i-1}\right) , \end{aligned}$$
(1.70)

and in the present case \(m\overline{\Delta \tau }=J_{m}^{-1} \sum _{j=1}^{J_{m}}D_{m,j}.\) The limiting forms of \(h_{j}\) and the random (variance) quantity \(\eta ^{2}\) are evaluated below.

We first observe that result (1.68) relies on the following representation, given in the Proof of Theorem 2 of Zhang et al. (2005),

$$\begin{aligned} \left[ p^{*}\right] ^{(avg)}-\left[ p^{*}\right] ^{{\mathcal {G}} _{m}}= & {} 2\sum _{j=1}^{J_{m}-1}\left( p^{*}\left( \tau _{j}\right) -p^{*}\left( \tau _{j-1}\right) \right) \\{} & {} \sum _{i=1}^{\left( K-1\right) \wedge j}\left( 1-\frac{i}{K}\right) \left( p^{*}\left( \tau _{j-i}\right) -p^{*}\left( \tau _{j-i-1}\right) \right) +O_{p}\left( \frac{K}{m}\right) \end{aligned}$$

whose leading term is

$$\begin{aligned}{} & {} 2\sum _{j=1}^{J_{m}-1}\sigma \left( \tau _{j-1}\right) \left( B\left( \tau _{j}\right) -B\left( \tau _{j-1}\right) \right) \\{} & {} \quad \sum _{i=1}^{\left( K-1\right) \wedge j}\left( 1-\frac{i}{K}\right) \sigma \left( \tau _{j-i}\right) \left( B\left( \tau _{j-i}\right) -B\left( \tau _{j-i-1}\right) \right) \\{} & {} \qquad \times \left\{ 1+O_{p}\left( \frac{\max _{j\le m}D_{m,j}}{m}\right) \right\} . \end{aligned}$$

Now

$$\begin{aligned}{} & {} 2\sum _{j=1}^{J_{m}-1}\sigma \left( \tau _{j-1}\right) \left( B\left( \tau _{j}\right) -B\left( \tau _{j-1}\right) \right) \\{} & {} \sum _{i=1}^{\left( K-1\right) \wedge j}\left( 1-\frac{i}{K}\right) \sigma \left( \tau _{j-i}\right) \left( B\left( \tau _{j-i}\right) -B\left( \tau _{j-i-1}\right) \right) \end{aligned}$$

is a local martingale with conditional variance whose leading term, as in the proof of Theorem 2 of Zhang et al. (2005), is

$$\begin{aligned}{} & {} 4\sum _{j=1}^{J_{m}-1}\sigma \left( \tau _{j-1}\right) ^{2}\left( \tau _{j}-\tau _{j-1}\right) \chi _{1j}^{2}\\{} & {} \quad \sum _{i=1}^{\left( K-1\right) \wedge j}\left( 1-\frac{i}{K}\right) ^{2}\sigma \left( \tau _{j-i}\right) ^{2}\left( \tau _{j-i}-\tau _{j-i-1}\right) \chi _{1j-i}^{2} \\{} & {} \quad =4\sum _{j=1}^{J_{m}-1}\sigma \left( \tau _{j-1}\right) ^{2}\left( \tau _{j}-\tau _{j-1}\right) \\{} & {} \quad \sum _{i=1}^{\left( K-1\right) \wedge j}\left( 1- \frac{i}{K}\right) ^{2}\sigma \left( \tau _{j-i}\right) ^{2}\left( \tau _{j-i}-\tau _{j-i-1}\right) \textrm{E}\left( \chi _{1j}^{2}\right) \textrm{E} \left( \chi _{1j-1}^{2}\right) +o_{p}\left( 1\right) \\{} & {} \quad =4\sum _{j=1}^{J_{m}-1}\sigma \left( \tau _{j-1}\right) ^{4}\left( \tau _{j}-\tau _{j-1}\right) \\{} & {} \quad \sum _{i=1}^{\left( K-1\right) \wedge j}\left( 1- \frac{i}{K}\right) ^{2}\left( \tau _{j-i}-\tau _{j-i-1}\right) \left\{ 1+O_{p}\left( \frac{K\max _{j\le m}D_{m,j}}{m}\right) \right\} +o_{p}\left( 1\right) \\{} & {} \quad =K\overline{\Delta \tau }\sum _{j=1}^{J_{m}}\sigma ^{4}\left( \tau _{j}\right) h_{j}\left( \tau _{j}-\tau _{j-1}\right) \left\{ 1+o_{p}\left( 1\right) \right\} =K\overline{\Delta \tau }\left\{ \eta _{m}^{2}+o_{p}\left( 1\right) \right\} , \end{aligned}$$

thereby giving (1.69) by martingale central limit theory since, under D2(ii)\(^{\prime }\) and \(K=cm^{2/3},\)

$$\begin{aligned} \frac{K\max _{j\le m}D_{m,j}}{m}=o_{p}\left( 1\right) . \end{aligned}$$

To find an explicit expression for the limit quantity \(\eta ^{2}\) in (1.68) we proceed as follows. First observe that for the present model

$$\begin{aligned} m\overline{\Delta \tau }=J_{m}^{-1}\sum _{j=1}^{J_{m}}D_{m,j}=J_{m}^{-1} \sum _{j=1}^{J_{m}}\textrm{E}\left( D_{m,j}|{\mathcal {F}}_{\tau _{j-1}}\right) +o_{p}\left( 1\right) \rightarrow _{p}\text { }\int _{0}^{1}\mu _{D}\left( s\right) \textrm{d}s, \end{aligned}$$

so that for \(r\ge 0\)

$$\begin{aligned} h_{j=\left[ Kr\right] }= & {} \frac{4m}{K\int _{0}^{1}\mu _{D}\left( s\right) \textrm{d}s} \sum _{i=1}^{\left( K-1\right) \wedge \left[ Kr\right] }\left( 1-\frac{i}{K} \right) ^{2}\left( \tau _{\left[ Kr\right] -i}-\tau _{\left[ Kr\right] -i-1}\right) \left\{ 1+o_{p}\left( 1\right) \right\} \\= & {} \frac{4}{K\int _{0}^{1}\mu _{D}\left( s\right) \textrm{d}s}\sum _{i=1}^{\left( K-1\right) \wedge \left[ Kr\right] }\left( 1-\frac{i}{K}\right) ^{2}D_{m, \left[ Kr\right] -i}\left\{ 1+o_{p}\left( 1\right) \right\} \\= & {} \frac{4}{K\int _{0}^{1}\mu _{D}\left( s\right) \textrm{d}s}\sum _{i=1}^{\left( K-1\right) \wedge \left[ Kr\right] }\left( 1-\frac{i}{K}\right) ^{2}\textrm{E }\left( D_{m,\left[ Kr\right] -i}|{\mathcal {F}}_{\tau _{\left[ Kr\right] -i}}\right) \left\{ 1+o_{p}\left( 1\right) \right\} \\{} & {} \rightarrow _{p}\,\frac{4\left\{ \int _{0}^{1\wedge r}\left( 1-s\right) ^{2}\mu _{D}\left( 1\wedge r-s\right) \textrm{d}s\right\} }{ \int _{0}^{1}\mu _{D}\left( s\right) \textrm{d}s}\text {.} \end{aligned}$$

Since \(K/J_{m}\rightarrow 0,\) it follows that \(\left[ J_{m}t\right] >K\) as \( m\rightarrow \infty \) for all \(t\in (0,1].\) Hence, for all \(t\in (0,1],\)

$$\begin{aligned} h_{j=\left[ J_{m}t\right] }= & {} \frac{4}{K\overline{\Delta \tau }} \sum _{i=1}^{\left( K-1\right) \wedge \left[ J_{m}t\right] }\left( 1-\frac{i}{ K}\right) ^{2}\left( \tau _{\left[ J_{m}t\right] -i}-\tau _{\left[ J_{m}t \right] -i-1}\right) \\{} & {} \rightarrow _{p}\,\frac{4\left\{ \int _{0}^{1}\left( 1-s\right) ^{2}\mu _{D}\left( 1-s\right) \textrm{d}s\right\} }{ \int _{0}^{1}\mu _{D}\left( s\right) \textrm{d}s}, \end{aligned}$$

and then

$$\begin{aligned} \eta _{m}^{2}=\sum _{j=1}^{J_{m}}\sigma ^{4}\left( \tau _{j}\right) h_{j}\left( \tau _{j}-\tau _{j-1}\right) \rightarrow _{p}\, \frac{4\left\{ \int _{0}^{1}\left( 1-s\right) ^{2}\mu _{D}\left( 1-s\right) \textrm{d}s\right\} }{\int _{0}^{1}\mu _{D}\left( s\right) \textrm{d}s}\int _{0}^{1}\sigma ^{4}\left( t\right) \textrm{d}t. \end{aligned}$$

Observe that when \(\mu _{D}\left( s\right) =\mu _{D}\) \(\ a.s.,\) the right side of the last expression simplifies to

$$\begin{aligned} 4\left\{ \int _{0}^{1}\left( 1-s\right) ^{2}\textrm{d}s\right\} \int _{0}^{1}\sigma ^{4}\left( t\right) \textrm{d}t=\frac{4}{3}\int _{0}^{1}\sigma ^{4}\left( t\right) \textrm{d}t, \end{aligned}$$

as given in Zhang et al. (2005) equation (5). It therefore follows that

$$\begin{aligned}{} & {} \left( \frac{m}{K}\right) ^{1/2}\left( \left[ p^{*}\right] ^{(avg)}-\int _{0}^{1}\sigma ^{2}(t)\textrm{d}t\right) \nonumber \\{} & {} \quad \Rightarrow \mathcal{M}\mathcal{N}\left( 0,\frac{ 4\left\{ \int _{0}^{1}\left( 1-s\right) ^{2}\mu _{D}\left( 1-s\right) \textrm{d}s\right\} }{\int _{0}^{1}\mu _{D}\left( s\right) \textrm{d}s}\int _{0}^{1}\sigma ^{4}\left( t\right) \textrm{d}t\right) . \end{aligned}$$
(1.71)

Next, consider (1.67). Since \(\widehat{\left[ p^{*}\right] }=\left[ p\right] ^{(avg)}-\frac{{\bar{m}}}{J_{m}}\left[ p\right] ^{{\mathcal {G}}_{m}}\), we have

$$\begin{aligned}{} & {} \sqrt{\frac{K}{{\bar{m}}}}\left\{ \widehat{\left[ p^{*}\right] }-\left[ p^{*}\right] ^{(avg)}\right\} \nonumber \\{} & {} \quad =\sqrt{\frac{K}{{\bar{m}}}}\left\{ \left[ p\right] ^{(avg)}-\left[ p^{*} \right] ^{(avg)}-\frac{{\bar{m}}}{J_{m}}\left[ p\right] ^{{\mathcal {G}} _{m}}\right\} \nonumber \\{} & {} \quad =\sqrt{\frac{K}{{\bar{m}}}}\left\{ \left[ p\right] ^{(avg)}-\left[ p^{*} \right] ^{(avg)}-2{\bar{m}}\sigma _{u}^{2}\right\} \nonumber \\{} & {} \qquad -2\sqrt{K{\bar{m}}}\left\{ J_{m}^{-1}\left[ p\right] ^{{\mathcal {G}} _{m}}-\sigma _{u}^{2}\right\} \nonumber \\{} & {} \quad \Rightarrow {\mathcal {N}}\left( 0,8\sigma _{u}^{4}\right) , \end{aligned}$$
(1.72)

as in Theorem A.1 and equation (56) of Zhang et al. (2005). Combining (1.71) and (1.72) as in Theorem 4 of Zhang et al. (2005) and using \(K=cm^{2/3}\) for some constant \(c>0,\) so that \(\sqrt{K/{\bar{m}}}=K/m^{1/2}=cm^{1/6}\) and \(\sqrt{m/K} =m^{1/6}/c^{1/2},\) we then have the limit theory

$$\begin{aligned} m^{1/6}\left\{ \widehat{\left[ p^{*}\right] }-\int _{0}^{1}\sigma ^{2}(t)\textrm{d}t\right\} \Rightarrow \left\{ \frac{8}{c^{2}}\sigma _{u}^{4}+c\eta ^{2}\right\} ^{1/2}{\mathcal {N}}\left( 0,1\right) , \end{aligned}$$

as required. \(\square \)

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Phillips, P.C.B., Yu, J. Information loss in volatility measurement with flat price trading. Empir Econ 64, 2957–2999 (2023). https://doi.org/10.1007/s00181-022-02353-y

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00181-022-02353-y

Keywords

JEL Classification

Navigation