Skip to main content
Log in

The SIML Estimation of Integrated Covariance and Hedging Coefficient Under Round-off Errors, Micro-market Price Adjustments and Random Sampling

  • Published:
Asia-Pacific Financial Markets Aims and scope Submit manuscript

Abstract

For estimating the integrated volatility and covariance by using high frequency data, Kunitomo and Sato (Math Comput Simul 81:1272–1289, 2011; N Am J Econ Finance 26:289–309, 2013) have proposed the separating information maximum likelihood (SIML) method when there are micro-market noises. The SIML estimator has reasonable finite sample properties and asymptotic properties when the sample size is large when the hidden efficient price process follows a Brownian semi-martingale. We shall show that the SIML estimation is useful for estimating the integrated covariance and hedging coefficient when we have round-off errors, micro-market price adjustments and noises, and when the high-frequency data are randomly sampled. The SIML estimation is consistent, asymptotically normal in the stable convergence sense under a set of reasonable assumptions and it has reasonable finite sample properties with these effects.

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

Similar content being viewed by others

References

  • Ait-Sahalia, Y., & Jacod, J. (2014). High frequency financial econometrics. Princeton: Princeton University Press.

    Book  Google Scholar 

  • Anderson, T. W. (1971). The statistical analysis of time series. New York: Wiley.

    Google Scholar 

  • Barndorff-Nielsen, O., Hansen, P., Lunde, A., & Shephard, N. (2008). Designing realized kernels to measure the ex-post variation of equity prices in the presence of noise. Econometrica, 76–6, 1481–1536.

    Google Scholar 

  • Barndorff-Nielsen, O., Hansen, P., Lunde, A., & Shephard, N. (2011). Multivariate realized kernels : Consistent positive semi-definite estimators of the covariation of equity prices with noise and non-synchronous trading. Journal of Econometrics, 162, 149–169.

    Article  Google Scholar 

  • Billigsley, P. (1995). Probability and measures. New York: Wiley.

    Google Scholar 

  • Christensen, K., Podolskij, M., & Vetter, M. (2013). On covariance estimation for multivariate continuous Ito semimartingales with noise in non-synchronous observation schemes. Journal of Multivariate Analysis, 120, 59–84.

    Article  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, 1–28.

    Article  Google Scholar 

  • Duffie, D. (1989). Futures markets. Englewood Cliffs: Prentice-Hall.

    Google Scholar 

  • Engle, R. F., & Russell, J. R. (1998). Autoregressive conditional duration models: A new model for irregularly spaced transaction data. Econometrica, 66–5, 1127–1162.

    Article  Google Scholar 

  • Engle, R., & Sun, Z. (2007). When is noise not noise: A micro-structure estimate of realized volatility. Working paper

  • Hansbrouck, J. (2007). Empirical market microstructure. Oxford: Oxford University Press.

    Google Scholar 

  • Hayashi, T., & Yoshida, N. (2005). On covariance estimation of non-synchronous observed diffusion processes. Bernoulli, 11(2), 359–379.

    Article  Google Scholar 

  • Hayashi, T., & Yoshida, N. (2008). Asymptotic normality of a covariance estimator for non-synchronously observed diffusion processes. Annals of Institute of Statistical Mathematics, 60, 367–407.

    Article  Google Scholar 

  • Ikeda, N., & Watanabe, S. (1989). Stochastic differential equation and diffusion processes (2nd ed.). Amsterdam: North-Holland.

    Google Scholar 

  • Jacod, J., Li, Y., Mykland, P. A., Podolskij, M., & Vetter, M. (2009). Microstructure noise with the continuous case: The pre-averaging approach. Stochastic Processes and Their Applications, 119, 2249–2276.

    Article  Google Scholar 

  • Jacod, J., & Shiryaev, A. N. (2003). Limit theorems for stochastic processes. Berlin: Springer.

    Book  Google Scholar 

  • Jacod, J., & Protter, P. (2012). Discretization of processes. Berlin: Springer.

    Book  Google Scholar 

  • Gloter, A., & Jacod, J. (2001). Diffusions with measurement errors. I: Local asymptotic optimality. ESAIM: Probability and Statistics, 5, 225–242.

    Article  Google Scholar 

  • Harris, F., McInish, T., Shoesmith, G., & Wood, R. (1995). Cointegration, error correction and price discovery on informationally-linked security markets. Journal of Financial and Quantitative Analysis, 30, 563–581.

    Article  Google Scholar 

  • Kunitomo, N., & Sato, S. (1999). Stationary and non-stationary simultaneous switching autoregressive models with an application to financial time series. Japanese Economic Review, 50–2, 161–190.

    Article  Google Scholar 

  • Kunitomo, N., & Sato, S. (2008). Separating information maximum likelihood estimation of integrated volatility and covariance with micro-market noise. In CIRJE discussion paper F-581. University of Tokyo. http://www.e.u-tokyo.ac.jp/cirje/research/

  • Kunitomo, N., & Sato, S. (2011). The SIML estimation of the realized volatility and hedging coefficient of the Nikkei-225 futures with micro-market noise. Mathematics and Computers in Simulation, 81, 1272–1289.

    Article  Google Scholar 

  • Kunitomo, N., & Sato, S. (2013). Separating information maximum likelihood estimation of integrated volatility and covariance with micro-market noise. North American Journal of Economics and Finance, 26, 289–309.

    Article  Google Scholar 

  • Li, Y., & Mykland, P. A. (2015). Rounding errors and volatility estimation. Journal of Financial Econometrics, 13(2), 478–504.

    Article  Google Scholar 

  • Misaki, H. (2013). An empirical analysis of volatility, covariance and hedging-ratio by the SIML method at the Osaka-Securities Exchanges. Unpublished manuscript.

  • Misaki, H., & Kunitomo, N. (2013). On robust property of the SIML estimation of volatility under micro-market noise and random sampling. Discussion paper CIRJE-F-892, CIRJE, Graduate School of Economics, University of Tokyo, forthcoming in International Review of Economics and Finance, Elsevier.

  • Rosenbaum, M. (2009). Integrated volatility and round-off error. Bernoulli, 15–3, 687–720.

    Article  Google Scholar 

  • Sato, S., & Kunitomo, N. (1996). Some properties of the maximum likelihood estimator in simultaneous switching autoregressive model. Journal of Time Series Analysis, 17, 287–307.

    Article  Google Scholar 

  • Sato, S., & Kunitomo, N. (2015). A robust estimation of integrated volatility and covariance with micro-market adjustments and round-off errors. CIRJE discussion paper F-964, University of Tokyo.

  • Zhang, L., Mykland, P., & Ait-Sahalia, Y. (2005). A tale of two time scales: Determining integrated volatility with noisy high-frequency data. Journal of the American Statistical Association, 100, 1394–1411.

    Article  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Naoto Kunitomo.

Additional information

This work was supported by JSPS KAKENHI Grant Number (A)25245033. It is a revision of Discussion Paper CIRJE-F-893, Graduate School of Economics, University of Tokyo. The present paper has been presented at IMS Tokyo meeting in July 2013 and JAFEE meeting in January 2015. We thank Ryuzo Miura for comments to an earlier version.

Appendix: On Mathematical Derivations

Appendix: On Mathematical Derivations

In this Appendix, we give some details of the proof of Theorem 4.1. The method of our derivations is based on the results reported by Kunitomo and Sato (2013) and Sato and Kunitomo (2015), but we need some extra arguments. We shall use the notations \(K_i\;(i\ge 1)\) as positive constants. Our proof consists of several steps.

(Step-1) The first step is to argue that the effects of the randomness of sampling and the effects of drift terms in the underlying stochastic processes are stochastically negligible under our assumptions. Our proof consists of several steps.

Although \(t_i^n\) and \(n^{*}\) are random variables, which are finite-valued and bounded in [0, 1] for any n. We write \(y_i^a=x_i^a+v_i^a \) for \(a=s\) or f ,  where \( y_i^a=y_a({t_i^n})\) and \( x_i^a=X_a({t_i^n})\) in the basic case. We set \(y_i^n=(y^s(t_i^n),y^f(t_i^n))^{\prime },\) \(x_i^n=(x^s(t_i^n),x^f(t_i^n))^{\prime }\) and also we write the underlying (unobservable) returns in the period \((t_{i-1}^n,t_i^n]\) as

$$\begin{aligned} r_i^{*}= & {} x_i^n-x_{i-1}^n =\int _{t_{i-1}^n}^{t_i^n}\mu _x (s,X(s))ds\nonumber \\&+\int _{t_{i-1}^n}^{t_i^n}\sigma _x (s,X(s))dB_s\,\,(i=1,\ldots ,n^{*}) \end{aligned}$$
(6.1)

and the martingale part as

$$\begin{aligned} r_i^n=\int _{t_{i-1}^n}^{t_i^n} \sigma _x (s,X(s)) dB_s\;\;(i=1,\ldots ,n^{*}) \end{aligned}$$
(6.2)

with \(0=t_0\le t_1^n<\cdots <t_{n^{*}}^n\le 1\). By using Assumptions I and II, we have \(n^{*}/n =1+o_p(1)\),

$$\begin{aligned} \mathbf{E}\left[ \left\| r_j^{*}\right\| ^2\right]= & {} \mathbf{E}\left[ \left\| \int _{t_{i-1}^n}^{t_i^n}\mu _x (s,X(s))ds \right| ^2\right] \\&+2\mathbf{E}\left[ \left( \int _{t_{i-1}^n}^{t_i^n}\mu _x (s,X(s))ds\right) ^{\prime } r_i^n\right] +\mathbf{E}\left[ \left\| r_i \right\| ^2\right] , \end{aligned}$$

and

$$\begin{aligned} \mathbf{E}\left[ \left\| \int _{t_{i-1}^n}^{t_i^n}\sigma _x(s)dB_s -\int _{t_{i-1}^n}^{t_i^n}\sigma _x\left( t_{i-1}^n\right) dB_s\right\| ^2\right] =O\left( \left( \frac{1}{n}\right) ^2\right) . \end{aligned}$$

Then we can evaluate as

$$\begin{aligned} \mathbf{E}\left[ \left\| r_i^{*} \right\| ^2-\int _{t_{i-1}^n}^{t_i^n}\mathrm{tr}(\Sigma _x(s))ds\right] =O\left( \left( \frac{1}{n}\right) ^{3/2}\right) . \end{aligned}$$
(6.3)

Hence we find that the effects of drift terms are negligible for the estimation of integrated volatility function under Assumptions I and II.

Similarly, we can use that \(n^{*}/n\mathop {\rightarrow }\limits ^{p} 1\) and

$$\begin{aligned} \sum _{j=1}^{n^{*}} \int _{t_{j-1}^n}^{t_j^n}\Sigma _x(s) ds -\int _0^1 \Sigma _x(s) ds \mathop {\rightarrow }\limits ^{p} O\;, \end{aligned}$$
(6.4)

and then

$$\begin{aligned} \mathbf{E}\left[ \int _{t_{i-1}^n}^{t_i^n}\Sigma _x(s) ds\right] =O\left( \frac{1}{n}\right) . \end{aligned}$$
(6.5)

We note that the volatility function \(\Sigma _x(s) \;(0\le s \le 1)\) and the integrated volatility \(\Sigma _x =\int _0^1 \Sigma _x (s)ds\) can be stochastic.

By above arguments we can proceed the present proof as if \(n^{*}\) were fixed and was replaced by the corresponding fixed n. For instance we can use the standard method for the random sums (martingales) to the following proof of CLT (central limit theorem). (See Section 35 of Billigsley (1995) for instance.)

Then we shall use the fixed n (and \(m_{n^{*}}\)) as if it were \(n^{*}\) (and \(m_n\)) in the following developments of this Appendix for the sake of the resulting simplicities. (Basically we need to replace n by \(n^{*}\) in each step, and then we need to use many straightforward, but tedious arguments because \(n^{*}\) is stochastic.)

(Step-2) Let \(Z_{in}^{(1)}\) and \(Z_{in}^{(2)}\,(i=1,\ldots ,n)\) be the i-th elements of \(n\times 2\) vectors

$$\begin{aligned} \mathbf{Z}_n^{(1)} =h_n^{-1/2}\mathbf{P}_n\mathbf{C}_n^{-1}(\mathbf{X}_n-{\bar{\mathbf{Y}}}_0),\,\mathbf{Z}_n^{(2)}=h_n^{-1/2}\mathbf{P}_n\mathbf{C}_n^{-1}\mathbf{V}_n, \end{aligned}$$
(6.6)

respectively, where we denote \(\mathbf{X}_n=(x_{i}^{\prime })=(x_i^s,x_i^f)\), \(\mathbf{V}_n=(v_{i}^{\prime })=(v_i^s,v_i^f)\), \(\mathbf{X}_n=(z_{in}^{\prime })\) are \(n\times 2\) vectors with \(z_{in}=z_{in}^{(1)}+z_{in}^{(2)}\) and \(\mathbf{P}_n\) is defined by (2.15).

We write \(z_{kn}^s\) and \(z_{kn}^f\) as the first and second components of \(z_{kn}\), and also we use the notations \(z_{kn}^{s,j}\) and \(z_{kn}^{f,j};(j=1,2)\) for the j-th components of \(z_{kn}^s\) and \(z_{kn}^f\). By following the proof developed by Kunitomo and Sato (2013) for the case of fixed n, we use the arguments for investigating on the asymptotic distribution of \(\sqrt{m_n}[{\hat{\sigma }}_{ss}^{(x)} - \sigma _{ss}^{(x)}]\) and \( \sigma _{ss}^{(x)}=\int _0^1\sigma _{ss}^{(x)} (s) ds\). (We can use the similar arguments for \( \sigma _{ff}^{(x)}=\int _0^1\sigma _{ff}^{(x)} (s) ds\) and \( \sigma _{sf}^{(x)}=\int _0^1\sigma _{sf}^{(x)} (s) ds\), but we omit the details.) We use the decomposition

$$\begin{aligned} \sqrt{m_n}\left[ {\hat{\sigma }}_{ss}^{(x)} - \sigma _{ss}^{(x)} \right]= & {} \sqrt{m_n}\left[ \frac{1}{m_n}\sum _{k=1}^{m_n} \left( z_{kn}^s\right) ^2 -\sigma _{ss}^{(x)}\right] \nonumber \\= & {} \sqrt{m_n}\left[ \frac{1}{m_n}\sum _{k=1}^{m_n}\left( z_{kn}^{s,1}\right) ^2-\sigma _{ss}^{(x)}\right] + \frac{1}{\sqrt{m_n}}\sum _{k=1}^{m_n} \mathbf{E}\left[ \left( z_{kn}^{s,2}\right) ^2\right] \nonumber \\&+\frac{1}{\sqrt{m_n}}\sum _{k=1}^{m_n} \left[ \left( z_{kn}^{s,2}\right) ^2-\mathbf{E}\left[ \left( z_{kn}^{s,2}\right) ^2\right] \right] \nonumber \\&+2\frac{1}{\sqrt{m_n}}\sum _{k=1}^{m_n} \left[ \left( z_{kn}^{s,1}\right) \left( z_{kn}^{s,2}\right) \right] . \end{aligned}$$
(6.7)

Because under Assumption II the effects of differences due to \(n^{*}\) and n are small, we notice that we can ignore the effects of \( \sum _{m_{n^{*}}}^n ( z_{kn}^s )^2 \) and \([m_{n^{*}}-m_n]/m_n\) in the following evaluations.

Then we shall investigate the conditions that three terms except the first one of (6.7) are \(o_p(1)\). If they are satisfied, we could estimate the integrated volatility consistently as if there were no noise terms because other terms can be ignored asymptotically as \(n\rightarrow \infty \).

Let \(\mathbf{b}_k= (b_{kj})= \mathbf{e}_k^{\prime }\mathbf{P}_n\mathbf{C}_n^{-1}=(b_{kj}) \) and \(\mathbf{e}_k^{\prime }=(0,\ldots ,1,0,\ldots )\) be an \(n\times 1\) vector. We write \(z_{kn}^{(2)}=\sum _{j=1}^n b_{kj}v_{j}\) and use the relation

$$\begin{aligned} \left( \mathbf{P}_n\mathbf{C}_n^{-1}\mathbf{C}_n^{'-1}\mathbf{P}_n^{\prime }\right) _{k,k^{\prime }} =\delta (k,k^{\prime })4 \sin ^2 \left[ \frac{\pi }{2n+1}\left( k-\frac{1}{2}\right) \right] . \end{aligned}$$

Then because we have \( n\sum _{j=1}^n b_{kj}b_{k^{\prime }j}=\delta (k,k^{\prime }) a_{kn},\) \( \mathbf{E}(v_{ss}^{(v)})\) are bounded and (2.4), it is straightforward to find \(K_1\) such that

$$\begin{aligned} \mathbf{E}\left[ \left( z_{kn}^{s,2}\right) \right] ^2 = n \mathbf{E}\left[ \sum _{i=1}^n b_{ki}v_i\sum _{j=1}^n b_{kj} v_{j}\right] \le K_1 \times a_{kn}, \end{aligned}$$
(6.8)

where we use the notation \(b_{kj}=0\;(j\le 0)\). (We take \(\rho =\max \{\rho _s,\rho _f\}\) and apply (32) of Kunitomo and Sato (2011).) Also Kunitomo and Sato (2013) have shown that

$$\begin{aligned} \frac{1}{m_n}\sum _{k=1}^{m_n} a_{k n} = \frac{1}{m_n} 2n \sum _{k=1}^{m_n} \left[ 1-\cos \left( \pi \frac{2k-1}{2n+1}\right) \right] = O\left( \frac{m_n^2}{n}\right) \end{aligned}$$
(6.9)

and the second term of (6.7) becomes

$$\begin{aligned} \frac{1}{\sqrt{m_n}}\sum _{k=1}^{m_n} \mathbf{E}\left[ z_{kn}^{s,2}\right] ^2 \le K_1 \frac{1}{\sqrt{m_n}}\sum _{k=1}^{m_n} a_{kn} =O\left( \frac{m_n^{5/2}}{n}\right) . \end{aligned}$$
(6.10)

This term is o(1) if \( 0< \alpha <0.4\). (When we use the MSIML estimation, the effect of this term can be removed from \(\sqrt{m_n}[ \hat{\sigma _{m.ss}^{(x)}} - \sigma _{ss}^{(x)}]\) and we can improve the condition such that \( 0< \alpha < 0.5\) in Theorem 4.1.)

For the fourth term of (6.7),

$$\begin{aligned} \mathbf{E}\left[ \frac{1}{\sqrt{m_n}} \sum _{j=1}^{m_n} z_{kn}^{s,1}z_{kn}^{s,2} \right] ^2= & {} \frac{1}{m_n}\sum _{k,k^{\prime }=1}^{m_n} \mathbf{E} \left[ z_{kn}^{s,1}z_{k^{\prime },n}^{(1)} z_{kn}^{s,2}z_{k^{\prime },n}^{s,2} \right] \\= & {} \mathbf{E}\left[ 2\sum _{j,j^{\prime }=1}^ns_{jk}s_{j^{\prime }k^{\prime }} \mathbf{E}\left( r_jr_{j^{\prime }}\vert {\mathcal {F}}_{\min (j,j^{\prime })}\right) z_{kn}^{s,2}z_{k^{\prime }n}^{s,2} \right] \\= & {} O \left( \frac{m_n^2}{n}\right) , \end{aligned}$$

where

$$\begin{aligned} s_{jk}= \cos \left[ \frac{2\pi }{2n+1}\left( j-\frac{1}{2}\right) \left( k-\frac{1}{2}\right) \right] \end{aligned}$$

for \(j,k=1,2,\ldots ,n\). (See Lemma 1.3 of Kunitomo and Sato (2008).) In the above evaluation we have used the relation

$$\begin{aligned} \left| \sum _{j=1}^n s_{jk}s_{j,k^{\prime }}\right| \le \left[ \sum _{j=1}^n s_{jk}^2\right] =n/2+1/4\,\,\mathrm{for\, any}\,k\ge 1. \end{aligned}$$

For the third term of (6.7), we need to consider the variance of

$$\begin{aligned} \left( z_{kn}^{s,2}\right) ^2-\mathbf{E}\left[ \left( z_{kn}^{s,2}\right) ^2\right] = n \sum _{j,j^{\prime }=1}^n b_{kj}b_{k,j^{\prime }} \left[ v_{j}^s v_{j^{a '}}-\mathbf{E}\left( v_{j}^s v_{j^{\prime }}^s\right) \right] \end{aligned}$$

and then it is enough to evaluate the expectation of \( \left[ (z_{kn}^{s, 2})^2-\mathbf{E}[ (z_{kn}^{s, 2})^2 ]\right] \left[ (z_{k^{\prime },n}^{s, 2})^2-\mathbf{E}[ (z_{k^{\prime },n}^{s, 2})^2 ]\right] \). When \(v_j^s\) are correlated and satisfy (2.4), becuse of the independence assumption on \(w_i\;(i=1,\ldots ,n),\) it is possible to evaluate that there exists a positive constant \(K_3\) such that

$$\begin{aligned} \mathbf{E}\left[ \frac{1}{\sqrt{m_n}}\sum _{j=1}^{m_n}\left( \left( z_{kn}^{s, 2}\right) ^2 -\mathbf{E}\left[ \left( z_{kn}^{s,2}\right) ^2\right] \right) \right] ^2\le & {} \frac{K_3}{m_n}\left( \sum _{k=1}^{m_n} a_{kn}\right) ^2 \nonumber \\&= O\left( \frac{1}{m_n}\times \left( \frac{m_n^2}{n}\right) ^2\right) .\qquad \end{aligned}$$
(6.11)

Thus the third term of (6.7) is negligible if \(0< \alpha <0.4\). (see Section 5 of Kunitomo and Sato (2011). When \(\{ v_j^s\}\) have the moving average (MA) representation as (2.4), the term is \(O(m^5/n^2)\) because of (6.9) and then it can be negligible if if \(0< \alpha <0.4\).)

When \(\{ v_j^s\}\) are mutually independent, it is possible to evaluate that the left-hand side of (6.11) is less that \((K_3/m_n)\sum _{k=1}^{m_n}a_{kn}^2\;,\) which is \(O(m_n^4/n^2)\). Then the third term of (6.7) is negligible if \(0< \alpha <0.5\). (See (6.13) of Kunitomo and Sato (2013). This point the important step for the proof of Part (iii) because then the only concern is on (6.10).)

(Step-3) The third step is to give the asymptotic variance of the first term of (6.7), that is,

$$\begin{aligned} \sqrt{m_n}\left[ \frac{1}{m_n}\sum _{k=1}^{m_n}\left( z_{kn}^{s,(1)}\right) ^2 - \sigma _{ss}^{(x)}\right] \end{aligned}$$
(6.12)

because it is of the order \(O_p(1)\). We write

$$\begin{aligned} z_{kn}^{s,(1)}= \sqrt{ \frac{n}{2n+1} }\sum _{j=1}^n r_j^{s} \left( e^{i \theta _{kj}}+e^{-i \theta _{kj}}\right) , \end{aligned}$$
(6.13)

where \(r_j^{s}\) is the first component of \(r_j^{*}\;(= x_j^n-x_{j-1}^n)\) in (6.1) and \(\theta _{kj}=[2\pi /(2n+1)](k-1/2)(j-1/2)\). By using the relation that \((e^{i \theta _{kj}}+e^{-i \theta _{k j}})^2 =2+e^{2i \theta _{kj}}+e^{-2 i \theta _{kj}}\), we represent

$$\begin{aligned}&\left[ \frac{2n+1}{2n}\right] \left[ \frac{1}{m_n}\sum _{k=1}^{m_n} \left( z_{kn}^{s,(1)}\right) ^2 -\int _{0}^{1}\sigma _{ss}^{(x)}(s) ds \right] \nonumber \\&\quad =\frac{1}{m_n}\sum _{k=1}^{m_n} \left\{ \frac{1}{2}\left[ \sum _{j=1}^n \left( r_j^s\right) ^2 \left( e^{i \theta _{kj}}+e^{-i \theta _{kj}} \right) ^2 -2\left( 1+\frac{1}{2n}\right) \int _{0}^{1}\sigma _{ss}^{(x)}(s)ds\right] \right. \nonumber \\&\quad \quad \left. +\left[ \sum _{j\ne j^{\prime }=1}^{n} r_j^s r_{j^{\prime }}^s \left( e^{i \theta _{kj}}+e^{-i \theta _{kj}}\right) \left( e^{i \theta _{kj^{\prime }}}+e^{-i \theta _{kj^{\prime }}}\right) \right] \right\} \nonumber \\&\quad = 2\sum _{j>j^{\prime }=1}^{n} r_j^s r_{j^{\prime }}^s \left[ \frac{1}{m_n}\sum _{k=1}^{m_n}\left( e^{i \theta _{kj}}+e^{-i \theta _{kj}}\right) \left( e^{i \theta _{kj^{\prime }}}+e^{-i \theta _{kj^{\prime }}}\right) \right] \nonumber \\&\quad \quad +\sum _{j=1}^{n}\left[ \left( r_j^s\right) ^2-\int _{t_{j-1}^n}^{t_j^n}\sigma _{ss}^{(x)}(s)ds\right] \nonumber \\&\quad \quad + \frac{1}{2}\sum _{j=1}^{n}\left[ \left( r_j^s\right) ^2 -\int _{t_{j-1}}^{t_j^n}\sigma _{ss}^{(x)}(s)ds\right] \left[ \sum _{k=1}^{m_n}\left( e^{2i\theta _{kj}}+e^{-2i\theta _{kj}}\right) \right] \nonumber \\&\quad \quad -\frac{1}{2n} \sum _{j=1}^{n}\left[ \int _{t_{j-1}}^{t_j^n}\sigma _{ss}^{(x)}(s) ds\right] \nonumber \\&\quad =(A)+(B)+(C)+(D),(\mathrm{say}). \end{aligned}$$
(6.14)

Then by using the derivations in Kunitomo and Sato (2013) (also Sato and Kunitomo (2015)), it is possible to show that except the first term (A) \(\sqrt{m_n} (B) \mathop {\rightarrow }\limits ^{p} 0,\) \(\sqrt{m_n} (C) \mathop {\rightarrow }\limits ^{p} 0,\) and \(\sqrt{m_n} (D) \mathop {\rightarrow }\limits ^{p} 0\) as \(m_n \rightarrow \infty \;(n \rightarrow \infty )\). Also by using the simple relation

$$\begin{aligned} \left( e^{i\theta _{kj}}+e^{-i\theta _{k j}}\right) \left( e^{i\theta _{kj^{\prime } }}+e^{-i \theta _{kj^{\prime }}}\right)= & {} \left( e^{i\left( \theta _{kj}+\theta _{kj^{\prime }}\right) }+e^{-i\left( \theta _{kj}+\theta _{kj^{\prime }}\right) }\right) \\&+\left( e^{i\left( \theta _{kj}-\theta _{kj^{\prime }}\right) }+e^{-i\left( \theta _{kj}-\theta _{kj^{\prime }}\right) }\right) , \end{aligned}$$

the random quantity \(\sqrt{m_n}\left[ \frac{1}{m_n}\sum _{k=1}^{m_n} (z_{kn}^{s,(1)})^2 -\sigma _{ss}^{(x)} \right] \) is asymptotically equivalent to

$$\begin{aligned} (A)^{\prime }=2\sum _{j> j^{\prime }=1}^{n} r_j^s r_{j^{\prime }}^s h_m(j,j^{\prime }), \end{aligned}$$
(6.15)

where for \(j,j^{\prime }=1,\ldots ,n\)

$$\begin{aligned} h_m(j,j^{\prime })= & {} \frac{1}{2\sqrt{m_n}} \frac{\sin 2m\left( \theta _j+\theta _{j^{\prime }}\right) /2}{\sin \left( \theta _j+\theta _{j^{\prime }}\right) /2} +\frac{1}{2\sqrt{m_n}} \frac{ \sin 2m\left( \theta _j-\theta _{j^{\prime }}\right) /2}{ \sin \left( \theta _j-\theta _{j^{\prime }}\right) /2} \end{aligned}$$

and \(\theta _j=[2\pi /(2n+1)](j-1/2)\).

By extending the evaluation method on F\(\acute{e}\)je-kernel (see Chapter 8 of Anderson (1971)), we can derive the variance of the asymptotic distribution of \((A)^{\prime }\), which is asymptotically equivalent to (6.12).

Lemma 6.1

Under Assumption II, as \(n \rightarrow \infty \) the asymptotic variance of \((A)^{\prime }\) is given by

$$\begin{aligned} V_{ss} = 2 \int _0^1 \left[ \sigma _{ss}^{(x)}( \tau (s)) \right] ^2 d(s)^2 ds. \end{aligned}$$
(6.16)

Proof of Lemma 6.1:

From the representation of (6.15), given the sampling process \(t_j^n\;(j=1,\ldots ,n^{*})\) we have the conditional expectation

$$\begin{aligned}&\mathbf{E}[ ((A)^{\prime })^2 \vert \{ t_j^s\} ]\\&\quad = 2\sum _{j> j^{\prime }=1}^{n^{*}} \sigma _{ss.j}^{(x)}\sigma _{ss.j^{\prime }}^{(x)}\left[ \frac{1}{4m}\right] \\&\quad \quad \times \left[ \frac{ \sin 2m\pi \left( j+j^{\prime }-1\right) /\left( 2^{n^{*}}+1\right) }{ \sin \pi \left( j+j^{\prime }\right) /\left( 2^{n^{*}}+1\right) } +\frac{ \sin 2m\pi \left( j-j^{\prime }-1\right) /\left( 2^{n^{*}}+1\right) }{ \sin \pi \left( j-j^{\prime }\right) /\left( 2^{n^{*}}+1\right) }\right] ^2, \end{aligned}$$

where we use the notation

$$\begin{aligned} \sigma _{ss.j}^{(x)} =\mathbf{E}\left[ \int _{t_{j-1}^n}^{t_j^n}\sigma _{ss}^{(x)}(s) ds \vert t_{j-1}^n,t_j^n\right] . \end{aligned}$$
(6.17)

Under Assumption I by using the standard approximation argument on integrations we find that

$$\begin{aligned} \mathbf{E}\left[ \int _{t_{j-1}^n}^{t_j^n}\sigma _{ss}^{(x)}(s) ds -\int _{t_{j-1}^n}^{t_j^n}\sigma _{ss}^{(x)}\left( t_{j-1}^n\right) ds\right] =o\left( \frac{1}{n}\right) \end{aligned}$$
(6.18)

and

$$\begin{aligned} \int _{t_{j-1}^n}^{t_j^n}\sigma _{ss}^{(x)}(t_{j-1}) ds =\left( t_n^n - t_{j-1}^n\right) \sigma _{ss}^{(x)}\left( t_{j-1}^n\right) =O_p\left( \frac{1}{n}\right) . \end{aligned}$$
(6.19)

Also for any continuous functions a(s) and b(t) in [0, 1]

$$\begin{aligned}&2\int _0^1\int _0^1 \frac{1}{4m} \left[ \frac{\sin m\pi (s-t)}{\sin (\pi /2)(s-t)}+\frac{\sin m\pi (s+t)}{\sin (\pi /2)(s+t)}\right] ^2 a(s)b(t) ds dt\nonumber \\&\quad = 2\int _0^1\int _0^1 \frac{1}{4m}\left\{ \left[ \frac{\sin m\pi (s-t)}{ \sin (\pi /2)(s-t)}\right] ^2+\left[ \frac{ \sin m\pi (s+t)}{ \sin (\pi /2) (s+t)}\right] ^2\right. \nonumber \\&\quad \quad \left. +\left[ \frac{\sin m\pi (s-t)}{\sin (\pi /2)(s-t)}\right] \left[ \frac{\sin m\pi (s+t)}{\sin (\pi /2)(s+t)}\right] \right\} a(s)b(t) ds dt\nonumber \\&\quad =(E)+(F)+(G),(\mathrm{say}). \end{aligned}$$
(6.20)

By changing the order of integrations, as \(m \rightarrow \infty \) we can evaluate the first term as

$$\begin{aligned}&2\int _0^1 \frac{1}{2m}\left[ \frac{\sin ^2(2m)\frac{\pi }{2}u}{\sin ^2(\pi /2)u}\right] \left[ \int _{0}^{1-u}a(u+t)b(t)dt\right] du\nonumber \\&\quad \longrightarrow 2 \lim _{u\rightarrow 0}\int _{0}^{1-u} a(u+t)b(t)dt\nonumber \\&=2 \int _0^1 a(t)b(t) dt. \end{aligned}$$
(6.21)

(See Lemma 8.3.3 of Anderson (1971).) Also by changing the order of integrations, we have found that the second term is negligible as

$$\begin{aligned} (F)= 2\int _0^1 \frac{1}{2m}\left[ \frac{\sin ^2(2m)\frac{\pi }{2}u}{ \sin ^2(\pi /2)u}\right] \left[ \int _0^u a(s)b(u-s)ds\right] du \longrightarrow 0\qquad \end{aligned}$$
(6.22)

as \(m \rightarrow \infty \). By applying the similar argument to the third term and we can find that the third term (G) is also negligible when m is large. Under Assumption I and Assumption II, as \((i(n)-1)/n \rightarrow s\) and \((j(n)-1)/n \rightarrow t\) for n being large while a fixed m, we have \( t_i^n \mathop {\rightarrow }\limits ^{p}\tau (s)\), \(t_j^n\rightarrow \tau (t)\), \(n(t_i^n-t_{i-1}^n) \mathop {\rightarrow }\limits ^{p} d(s)\), and \(n(t_j^n-t_{j-1}^n) \mathop {\rightarrow }\limits ^{p} d(t)\). Then the only non-negligible term in (6.17) corresponds to

$$\begin{aligned} V^{\prime }= \int _0^1\int _0^1 \frac{1}{4m}\left[ \frac{\sin m\pi (s-t)}{\sin (\pi /2)(s-t)}\right] ^2 \sigma _{ss}^{(x)}(\tau (s))\sigma _{ss}^{(x)}(\tau (t))d(s)d(t) ds dt\;.\nonumber \\ \end{aligned}$$
(6.23)

Then by letting \(m\rightarrow \infty ,\) we have the desired result as (6.16). \(\square \)

(Step-4) As the next step we need to show the stable convergence in law of (6.12), but the arguments are quite similar to (Step 4) of the proof of Theorem 3.2 in Sato and Kunitomo (2015), which is based on the method explained by Chapter VIII of Jacod and Shiryaev (2003) or Jacod and Protter (2012). Thus we have omitted the details of our arguments in this version.

(Step-5) Finally we need to deal with the integrated covariance. By modifying the derivations for the proof of the integrated covariance, we use \( \hat{\sigma }_{ss}^{(x)} , \hat{\sigma }_{ff}^{(x)}\) and \( \hat{\sigma }_{sf}^{(x)} \). (It is straightforward to develop the similar arguments as for \(\hat{\sigma }_{ss}^{(x)} \) but they are tedious. Then we have omitted the details.) Then we can evaluate the variance of the asymptotic distributions of the integrated covariance SIML estimator. Then the resulting variance formula becomes

$$\begin{aligned} V_{sf} = \int _0^1 \left[ \sigma _{ss}^{(x)}(\tau (s)) \sigma _{ff}^{(x)}(\tau (s))+\left( \sigma _{sf}^{(x)}(\tau (s))\right) ^2 \right] d(s)^2 ds. \end{aligned}$$
(6.24)

\(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Kunitomo, N., Misaki, H. & Sato, S. The SIML Estimation of Integrated Covariance and Hedging Coefficient Under Round-off Errors, Micro-market Price Adjustments and Random Sampling. Asia-Pac Financ Markets 22, 333–368 (2015). https://doi.org/10.1007/s10690-015-9205-3

Download citation

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10690-015-9205-3

Keywords

Navigation