This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

\authormark

Chakrabarti and Sen

\corres

*Rituparna Sen, Indian Statistical Institute, Bangalore, KA, India.

Limiting Spectral Distribution of High-dimensional Hayashi-Yoshida Estimator of Integrated Covariance Matrix

Arnab Chakrabarti    Rituparna Sen \orgdivMisra Centre for Financial Markets and Economy, \orgnameIndian Institute of Management, \orgaddressAhmedabad, \stateGJ, \countryIndia \orgdivApplied Statistics Division, \orgnameIndian Statistical Institute, \orgaddressBangalore, \stateKA, \countryIndia ritupar.sen@gmail.com
Abstract

[Summary]In this paper, the estimation of the Integrated Covariance matrix from high-frequency data, for high dimensional stock price process, is considered. The Hayashi-Yoshida covolatility estimator is an improvement over Realized covolatility for asynchronous data and works well in low dimensions. However it becomes inconsistent and unreliable in the high dimensional situation. We study the bulk spectrum of this matrix and establish its connection to the spectrum of the true covariance matrix in the limiting case where the dimension goes to infinity. The results are illustrated with simulation studies in finite, but high, dimensional cases. An application to real data with tick-by-tick data on 50 stocks is presented.

keywords:
Asynchronicity, Integrated Covariance, Realized Variance, Spectral Distribution, High-frequency data.
articletype: Research Articlefootnotetext: Acknowledgement:The authors thank Vikram Sarabhai library for the help with obtaining the data.footnotetext: Codes: All the R codes are available in the following GitHub repository https://github.com/Arnabchakrabarti15/LSD-of-Hayashi-Yoshida-estimator.

1 Introduction

Intraday financial data of multiple stocks are almost always nonsynchronous in nature. If not adjusted appropriately, nonsynchronous trading can affect multivariate stock price data analysis and the resulting inference quite heavily 1. The analysis of (intraday) financial data would fail to capture the reality of the financial market if the effect of asynchronicity is ignored. 2. Despite this problem, intraday data are important to measure the (co)variance of daily log returns of a given set of securities and can offer additional information compared to the estimate obtained from daily financial data. This covariance is called integrated (co)variance (or integrated (co)volatility). For low-dimensional stock-price process, integrated covariance can be accurately estimated by the Hayashi-Yoshida estimator 3. But in case of high-dimensional data, it suffers from the same problems as any sample covariance matrix under high-dimensional set up  4. As a consequence, its eigenvalue spectrum deviates considerably from the population counterpart. In this study, we derive the limiting spectral distribution of the Hayashi-Yoshida estimator for high-dimensional data.

In 1979, T. W. Epps reported that stock return correlations decrease as the sampling frequency of data increases 1. This is one of the earliest manifestations of the problems caused by asynchronicity and is known as the Epps effect. Later the phenomenon has been reported in several studies of different stock markets 5, 6, 7 and foreign exchange markets 8, 9. This is primarily a result of asynchronicity of price observations and the existing lead-lag relation between asset prices 10, 11, 12. Empirical results showed that considering only the synchronous or nearly synchronous ticks mitigates the problem significantly 10. In several studies it was shown that asynchronicity can induce potentially serious bias in the estimates of moments and co-moments of asset returns, such as their means, variances, covariances, betas, and autocorrelation and cross-autocorrelation coefficients 12, 13, 14, 15.

Integrated volatility is defined as the variance of the log return over a day of a given security. In fact, due to its high frequency, intraday financial data has been proven to be more efficient in measuring daily volatility when compared to daily financial data 16. For a single stock, Merton (1980) showed that the variance over a fixed interval can be estimated accurately as sum of squared realization as long as the data are available at sufficiently high sampling frequency 17. This estimator is known as Realized volatility. But often univariate modeling is not sufficient, it is also important to model the correlation dynamics between several assets. Hence one of the parameters of interest, to accurately estimate and infer about, is the integrated co-volatility or integrated covariance matrix. Analogous to the realized volatility, for a multivariate stock price process, the realized covolatility matrix/realized covariance matrix is defined. But the realized covolatility matrix relies upon synchronous observations and can not be readily extended for asynchronous data. Therefore in order to evaluate the realized covolatility, we have to first “synchronize” the data. Fixed clock time and refresh time 18 samplings are two such synchronizing algorithms widely used in practice. But the realized covariance, evaluated on a synchronous grid, is biased 3. 3 proposed an unbiased estimator of the Integrated covolatility that is applicable on intraday data without a need for synchronization. We will call this estimator as the Hayashi-Yoshida estimator. Although in presence of microstructure noise Hayashi-Yoshida estimator is also biased, a bias-corrected version was developed 19.

Hayashi-Yoshida estimator has good asymptotic properties as long as the data comes from an underlying low-dimensional diffusion process. But as the dimension of the data increases the estimator becomes inefficient. Developing a good estimator of high dimensional covolatility is challenging unless we impose some structure. A consistent and positive definite estimator is proposed based on blocking and regularization techniques 20. The central idea is to obtain one large covariance matrix from a series of smaller covariance matrices, each based on different sampling frequency. Shrinkage estimator of the covariance matrix with optimal shrinkage intensity, which is also important for portfolio optimization, also reduces the estimation error significantly 21. Many other modified shrinkage estimator with good asymptotic properties are proposed and applied in financial context 22, 23. The mixed frequency factor models, which uses high-frequency data to estimate factor covariance and low-frequency data to estimate the factor loadings, are also used to estimate high dimensional covariance matrices 24. The composite realized Kernel approach that estimates each entry of ICV matrix optimally (in terms of bandwidth and data loss) has been proposed and asymptotic properties are established 25. High-dimensionality affects the subsequent calculations of many important quantities based on covariance matrix. 26 showed that high-dimensionality affects the solution of Markowitz problem and results in underestimation of risk.

Instead of imposing a structure, an alternative avenue of investigating a high dimensional covariance matrix is to study its spectral distribution. 27 established the limiting spectral distribution of realized covariance matrix obtained from synchronized data. Recently an asymptotic relationship has been established between the limiting spectral distributions of the true sample covariance matrix and noisy sample covariance matrix 28. 29 studied the estimation of integrated covariance matrix based on noisy high-frequency data with multiple transactions using random matrix theory. 30 obtained the limiting spectral distribution of the covariance matrices of time-lagged processes. The limiting spectral distribution of sample covariance matrix was also derived under VARMA(p,q) model assumption 31. In this paper, we establish the limiting spectral distribution for the Hayashi-Yoshida estimator which has not yet been studied. Rest of the paper is organized as follows. In section 2, we discuss the background of the problem. Section 3 deals with a very brief introduction to random matrix theory. In section 4, we determine the limiting spectral distribution of high-dimensional Hayashi-Yoshida estimator. Simulated data analysis results are presented in section 6.1. The summary of this work and a brief discussion on some further directions are given in section 7.

2 Integrated Covariance Matrix and Asynchronicity

Suppose, we have pp stocks, whose price processes are denoted by StjS_{t}^{j} for j=1,,pj=1,...,p. and define the jjth log price process as Xtj:=logStjX_{t}^{j}:=\mathrm{log}S_{t}^{j}. Let Xt=(Xt1,,Xtp)TX_{t}=(X_{t}^{1},...,X_{t}^{p})^{T}. Then we can model XtX_{t} as a pp-dimensional diffusion process described as

dXt=μtdt+σdWtdX_{t}=\mu_{t}dt+\sigma dW_{t} (1)

where μt\mu_{t} is a pp dimensional drift process and σ\sigma is a ppxpp matrix, called instantaneous covolatility process. WtW_{t} is a pp dimensional standard Brownian motion. The Integrated covariance (ICV) matrix, our parameter of interest, is defined by

Σp=01σσT𝑑t.\Sigma_{p}=\int_{0}^{1}\sigma\sigma^{T}dt. (2)

In univariate case, the most widely used estimator of integrated variance is called the Realized variance. For pp stocks, analogous covariance estimator can be defined in the following way.

2.1 Realized covariance

Note that the transactions in each stock occur at random time points. Let nin_{i} be the number of observations for the iith stock. The arrival time of the llth observation of the iith stock is denoted by tlit_{l}^{i} . When the observations are assumed to be synchronous i.e. tli=tlt_{l}^{i}=t_{l} for i\forall i, the Realized Covariance (RCV) matrix can be defined as the following:

ΣpRCV=l=1nΔXlΔXlT,whereΔXl=(ΔXl1ΔXl2..ΔXlp)=(Xtl1Xtl11...XtlpXtl1p).\begin{split}\Sigma_{p}^{RCV}&=\sum_{l=1}^{n}\Delta X_{l}\Delta X_{l}^{T},\text{where}\\ \Delta X_{l}&=\begin{pmatrix}\begin{array}[]{c}\Delta X_{l}^{1}\\ \Delta X_{l}^{2}\\ .\\ .\\ \Delta X_{l}^{p}\end{array}\end{pmatrix}=\begin{pmatrix}\begin{array}[]{c}X_{t_{l}}^{1}-X_{t_{l-1}}^{1}\\ .\\ .\\ .\\ X_{t_{l}}^{p}-X_{t_{l-1}}^{p}\end{array}\end{pmatrix}.\end{split} (3)
t01t_{0}^{1}t11t_{1}^{1}t21t_{2}^{1}t31t_{3}^{1}t41t_{4}^{1}t51t_{5}^{1}t02t_{0}^{2}t12t_{1}^{2}t22t_{2}^{2}t32t_{3}^{2}
(a) Nonsynchronous observations
t01t_{0}^{1}t11t_{1}^{1}t21t_{2}^{1}t31t_{3}^{1}t41t_{4}^{1}t51t_{5}^{1}t02t_{0}^{2}t12t_{1}^{2}t22t_{2}^{2}t32t_{3}^{2}
(b) Synchronized observations illustrated by arrows.
Figure 1: (a) Illustration of asynchronous arrival or transaction times for two stocks are shown. The opens at t01=t02=0t^{1}_{0}=t^{2}_{0}=0 and the corresponding stock price at time 0 can be taken as previous day’s closing price. Transaction times of the first stock are shown by circles. Transaction times of the second stock are shown by rectangles. (b) The synchronized pairs (having the same colour code) are indicated by arrows.

2.2 Hayashi-Yoshida estimator

For asynchronous intraday data, the Realized covariance can not be directly calculated unless we synchronize the data by some ad hoc method. This means that we have to throw away some of the observations such that synchronized vectors of observations can be formed. In Fig. 1, we illustrate this for the bivariate case. Fig. 1(a) shows how nonsynchronous data would look like. The circles and quadrangles represent the transaction times for the first and second stock respectively. Synchronization of transaction times, as indicated by the arrows, are shown in Fig. 1(b). A synchronized dataset can be formed by pairing the stock prices corresponding to the synchronized time points. For example, two consecutive observations can be (Xt211,Xt222)(X^{1}_{t^{1}_{2}},X^{2}_{t^{2}_{2}}) and (Xt511,Xt322)(X^{1}_{t^{1}_{5}},X^{2}_{t^{2}_{3}}). This is equivalent to “pretending” that Xt511X^{1}_{t^{1}_{5}} is observed at t32t^{2}_{3} instead of at t51t^{1}_{5} and similarly, Xt222X^{2}_{t^{2}_{2}} is observed at t21t^{1}_{2} instead of at t22t^{2}_{2}.***For this reason, synchronization methods can be expressed as a problem of choosing a set of sampling times {τ1,τ2,,τn}\{\tau_{1},\tau_{2},...,\tau_{n}\}, from the set 𝒯={t11,,tn11}{t12,,tn22}\mathcal{T}=\{t^{1}_{1},...,t^{1}_{n_{1}}\}\cup\{t^{2}_{1},...,t^{2}_{n_{2}}\}. The corresponding price of each stock at τi\tau_{i} is taken as the price observed previous to τi\tau_{i}. We can see from Fig 1(b) that for the first stock, two observations at time t31t^{1}_{3} and t41t^{1}_{4} are not synchronized with any time point in the second stock and therefore can be excluded from the study. Extending this to pp stocks, let us denote that the number of resulting synchronized vectors is nn. Then it is evident that n{n1,n2,,np}n\leq\{n_{1},n_{2},...,n_{p}\}.

In Fig. 1(b), we have applied a particular synchronization method called refresh time sampling 32, 33.

3 proposed an alternative estimator (ΣpHY\Sigma_{p}^{HY}) of ICV matrix that does not require the dataset to be synchronized and therefore can be directly applied on the asynchronous data. Before defining it for high-dimension, we introduce it for the bivariate case. In the following expression, instead of writing X1X^{1} and X2X^{2} we simply write XX and YY. Now, for two stocks the Hayashi-Yoshida estimator is defined as the following way:

Σ2HY=k,lΔXkΔYl×𝟏{(tk1i,tki)(tl1i,tli)ϕ},\Sigma_{2}^{HY}=\sum_{k,l}\Delta X_{k}\Delta Y_{l}\times{\bf 1}\{(t_{k-1}^{i},t_{k}^{i})\cap(t_{l-1}^{i},t_{l}^{i})\neq\phi\}, (4)

where 𝟏{(tk1i,tki)(tl1i,tli)ϕ}{\bf 1}\{(t_{k-1}^{i},t_{k}^{i})\cap(t_{l-1}^{i},t_{l}^{i})\neq\phi\} is an indicator function that takes value 1 when the condition is satisfied. Fig. 2 illustrates the computation. Δx1×Δy1\Delta x_{1}\times\Delta y_{1} will contribute to the sum in Eq. (4) as (t02,t12)(t^{2}_{0},t^{2}_{1}) and (t01,t11)(t^{1}_{0},t^{1}_{1}) are overlapping intervals. But Δx1×Δy2\Delta x_{1}\times\Delta y_{2} will not contribute to the sum in Eq. (4) as the intervals (t12,t22)(t^{2}_{1},t^{2}_{2}) and (t01,t11)(t^{1}_{0},t^{1}_{1}) are non-overlapping.

t01t_{0}^{1}t11t_{1}^{1}t21t_{2}^{1}t31t_{3}^{1}t41t_{4}^{1}t51t_{5}^{1}t02t_{0}^{2}t12t_{1}^{2}t22t_{2}^{2}t32t_{3}^{2}Δy1\Delta y_{1}Δy2\Delta y_{2}Δy3\Delta y_{3}Δx1\Delta x_{1}Δx2\Delta x_{2}Δx3\Delta x_{3}Δx4\Delta x_{4}Δx5\Delta x_{5}
Figure 2: Illustration of asynchronous arrival or transaction times for two stocks are shown. The opens at t01=t02=0t^{1}_{0}=t^{2}_{0}=0 and the corresponding stock price at time 0 can be taken as previous day’s closing price. Transaction times of the first stock are shown by circles. Transaction times of the second stock are shown by rectangles. The returns on each interarrival (shown by braces) are denoted by Δxk=xkxk1\Delta x_{k}=x_{k}-x_{k-1} and Δyl=ylyl1\Delta y_{l}=y_{l}-y_{l-1}. When the intervals (tk11,tk1)(t^{1}_{k-1},t^{1}_{k}) and (tl12,tl2)(t^{2}_{l-1},t^{2}_{l}) have an nonempty intersection, Δxk×Δyl\Delta x_{k}\times\Delta y_{l} will contribute to the Hayashi-Yoshida covariance.

2.2.1 Hayashi-Yoshida covariance and Refresh-time sampling

Even though the Hayashi-Yoshida estimator does not require prior synchronization of intraday data, we will show that the estimator still throws away some of the data points. Moreover these data-points are exactly the same as thrown by refresh-time sampling. To see this let us consider the case as shown in Fig. 2. The Hayashi-Yoshida estimate is:

σHY(X,Y)=Δx1Δy1+Δx2Δy1+Δx2Δy2+Δx2Δy3+Δx3Δy3+Δx4Δy3+Δx5Δy3=Δx1Δy1+Δx2Δy1+Δx2Δy2+Δx2Δy3+(Δx3+Δx4+Δx5)Δy3\begin{split}\sigma^{HY}(X,Y)&=\Delta x_{1}\Delta y_{1}+\Delta x_{2}\Delta y_{1}+\Delta x_{2}\Delta y_{2}+\Delta x_{2}\Delta y_{3}+\Delta x_{3}\Delta y_{3}+\Delta x_{4}\Delta y_{3}+\Delta x_{5}\Delta y_{3}\\ &=\Delta x_{1}\Delta y_{1}+\Delta x_{2}\Delta y_{1}+\Delta x_{2}\Delta y_{2}+\Delta x_{2}\Delta y_{3}+(\Delta x_{3}+\Delta x_{4}+\Delta x_{5})\Delta y_{3}\end{split} (5)

But Δx3+Δx4+Δx5\Delta x_{3}+\Delta x_{4}+\Delta x_{5} is just the difference between log-price at t51t^{1}_{5} and log-price at t31t^{1}_{3} which doesn’t require any information on stock price anytime time in between. Therefore, although the Hayashi-Yoshida estimator doesn’t require presynchronization, it actually throws away the exact same observation as thrown by refresh-time sampling. As a consequence, the value of Hayashi-Yoshida covariance on full data will be equal to Hayashi-Yoshida estimator on the set of refresh-time pairs. Synchronizing the data using refresh-time sampling before computing the covariance can reduce the computational cost quite significantly.

When we move away from bivariate case to higher dimension, synchronize every pair of variables separately would not be very efficient. It would be preferable to synchronize the data for all the stocks simultaneously. This can be achieved by applying “all refresh method” which results in a synchronous sampled time points {t~1,t~2,,t~n}\{\tilde{t}_{1},\tilde{t}_{2},...,\tilde{t}_{n}\} defined in the following way:

t~j+1=max1iptNi(t~j)i,\begin{split}\tilde{t}_{j+1}=\underset{1\leq i\leq p}{\text{max}}t^{i}_{N_{i}(\tilde{t}_{j})},\end{split} (6)

where Ni(t)N_{i}(t) is the number of observation before time tt 34, 35. In this paper, we will define the Hayashi-Yoshida estimator on refresh time sampling times. The theoretical implication is that- given the synchronized data, we can now assume the number of observations of each stock to be the same. We will denote this common sample size as nn.

For pp stocks the Hayashi-Yoshida estimator is defined as the following :

ΣpHY=k,lΔXkΔXlTI(k,l),whereΔXl=(ΔXl1ΔXl2..ΔXlp)=(Xtl11Xtl111...XtlppXtl1pp)\begin{split}\Sigma_{p}^{HY}&=\sum_{k,l}\Delta X_{k}\Delta X_{l}^{T}\circ I(k,l),\text{where}\\ \Delta X_{l}&=\begin{pmatrix}\begin{array}[]{c}\Delta X_{l}^{1}\\ \Delta X_{l}^{2}\\ .\\ .\\ \Delta X_{l}^{p}\end{array}\end{pmatrix}=\begin{pmatrix}\begin{array}[]{c}X_{t_{l}^{1}}^{1}-X_{t_{l-1}^{1}}^{1}\\ .\\ .\\ .\\ X_{t_{l}^{p}}^{p}-X_{t_{l-1}^{p}}^{p}\end{array}\end{pmatrix}\end{split} (7)

and ‘\circ’ is the Hadamard product and I(k,l)I(k,l) is a p×pp\times p matrix with (i,j)th(i,j)^{th} element is the indicator function involving kthk^{th}interarrival of ithi^{th}stock and lthl^{th}interarrival of jthj^{th} stock: I(IkiIljϕ)I(I_{k}^{i}\cap I_{l}^{j}\neq\phi), where Iki=(tk1i,tki)I_{k}^{i}=(t_{k-1}^{i},t_{k}^{i}). In other words if two interarrivals intersect then product ΔXkiΔXljT\Delta X_{k}^{i}\Delta X_{l}^{j\thinspace T} will contribute to the sum. In Fig. 2, Ik1=(tk11,tk1),k{1,2,3,4,5}I_{k}^{1}=(t_{k-1}^{1},t_{k}^{1}),~{}k\in\{1,2,3,4,5\} and Il2=(tl12,tl2),l{1,2,3}I_{l}^{2}=(t_{l-1}^{2},t_{l}^{2}),~{}l\in\{1,2,3\} are shown.

2.3 Scaled Realized Covariance estimator

In this section we show the “closeness” of the Hayashi-Yoshida estimator with a scaled realized estimator which is motivated from the intraday covariance estimator proposed in 36. We determine the scaling coefficients for the bivariate case. The result will be key in our proof of the Limiting Spectral distribution. For the bivariate case, let us denote the log price for two stocks at a particular time tt as (Xt,Yt)(X_{t},Y_{t}). Following 36, we synchronize the data for two stocks in the following fashion:

Algorithm (𝒜0\mathcal{A}_{0}): 1. For i=1i=1, assign ki1=1k_{i}^{1}=1 and ki2=1k_{i}^{2}=1. 2. While ki1n1k^{1}_{i}\leq n_{1} and ki2n2k^{2}_{i}\leq n_{2}: If tki22>tki11t_{k_{i}^{2}}^{2}>t_{k_{i}^{1}}^{1} then find m=max{j:tj1<tki22}m=\mathrm{max}\{j:\ t_{j}^{1}<t_{k_{i}^{2}}^{2}\}. The iith pair will be (Xtm1,Ytki22)(X_{t_{m}^{1}},Y_{t_{k_{i}^{2}}^{2}}). Modify ki1=mk_{i}^{1}=m. If tki22tki11t_{k_{i}^{2}}^{2}\leq t_{k_{i}^{1}}^{1} then find m=max{j:tj2<tki11}m=\mathrm{max}\{j:\ t_{j}^{2}<t_{k_{i}^{1}}^{1}\}. The iith pair will be (Xtki11,Ytm2)(X_{t_{k_{i}^{1}}^{1}},Y_{t_{m}^{2}}). Modify ki2=mk_{i}^{2}=m Modify i=i+1i=i+1. ki1=ki1+1k_{i}^{1}=k_{i}^{1}+1 and ki2=ki2+1k_{i}^{2}=k_{i}^{2}+1.

The pairs created by this algorithm are identical to the pairs created by refresh time sampling but accommodates more information by retaining the actual transaction times. To see this, note that in Eq.(6)), a common set of synchronized points {t~1,,t~n}\{\tilde{t}_{1},...,\tilde{t}_{n}\} are defined for all stocks. For each stock the last observed stock price prior to τi\tau_{i} was taken to be the price at τi\tau_{i}. Therefore, from the refresh time sampling it is not possible to retrieve the actual transaction times of the synchronized pairs. Algorithm 𝒜0\mathcal{A}_{0}, on the other hand, keep these information. Instead of writing (Xtki11,Ytki22)(X_{t_{k_{i}^{1}}^{1}},Y_{t_{k_{i}^{2}}^{2}}) we shall henceforth write (Xt(ki1),Yt(ki2))(X_{t(k_{i}^{1})},Y_{t(k_{i}^{2})}).

2.3.1 Overlapping and non-overlapping regions for return construction

For two such consecutive synchronized pairs of stock-prices, we can now consider the bivariate return as: {(Xt(ki1)Xt(ki11)),(Yt(ki2)Yt(ki12)):i=1,2,,n}\{(X_{t(k_{i}^{1})}-X_{t(k_{i-1}^{1})}),(Y_{t(k_{i}^{2})}-Y_{t(k_{i-1}^{2})}):i=1,2,...,n\}. Note that, in this bivariate return vector, the first component (for XX) is defined on the interval (t(ki11),t(ki1))\big{(}t(k_{i-1}^{1}),t(k_{i}^{1})\big{)} and the return on the YY is defined on the interval (t(ki12),t(ki2))\big{(}t(k_{i-1}^{2}),t(k_{i}^{2})\big{)}. It can be shown the overlap and nonoverlapping parts of these two intervals play a crucial role in bias of the estimated covariance 36. To define the overlap we first illustrate four possible configurations of the intervals. In Fig. 3 we show four such configurations of intervals corresponding to a particular return vector, constructed from synchronized pairs of observations. More formally, suppose Xt(ki1)Xt(ki11)=i=ml(Xti+1Xti)X_{t(k_{i}^{1})}-X_{t(k_{i-1}^{1})}=\sum_{i=m}^{l}(X_{t_{i+1}}-X_{t_{i}}) for some mm and ll. Then one of these four configurations is true:

[1.Yt(ki2)Yt(ki12)=i=m+1l1(Yti+1Yti)2.Yt(ki2)Yt(ki12)=i=m1l1(Yti+1Yti)3.Yt(ki2)Yt(ki12)=i=m+1l+1(Yti+1Yti)4.Yt(ki2)Yt(ki12)=i=m1l+1(Yti+1Yti)]\left[\begin{array}[]{rcl}1.\quad Y_{t(k_{i}^{2})}-Y_{t(k_{i-1}^{2})}&=&\sum_{i=m+1}^{l-1}(Y_{t_{i+1}}-Y_{t_{i}})\\ 2.\quad Y_{t(k_{i}^{2})}-Y_{t(k_{i-1}^{2})}&=&\sum_{i=m-1}^{l-1}(Y_{t_{i+1}}-Y_{t_{i}})\\ 3.\quad Y_{t(k_{i}^{2})}-Y_{t(k_{i-1}^{2})}&=&\sum_{i=m+1}^{l+1}(Y_{t_{i+1}}-Y_{t_{i}})\\ 4.\quad Y_{t(k_{i}^{2})}-Y_{t(k_{i-1}^{2})}&=&\sum_{i=m-1}^{l+1}(Y_{t_{i+1}}-Y_{t_{i}})\end{array}\right] (8)

Given this set of possible configurations, we define a random variable LiL_{i}, denoting the overlapping time interval of iith interarrivals corresponding to Xt(ki1)Xt(ki12)X_{t(k_{i}^{1})}-X_{t(k_{i-1}^{2})} and Yt(ki2)Yt(ki12)Y_{t(k_{i}^{2})}-Y_{t(k_{i-1}^{2})} as

Li={t(ki2)t(ki12)ifYt(ki2)Yt(ki11)=i=m+1l1(Yti+1Yti)t(ki2)t(ki11)ifYt(ki2)Yt(ki11)=i=m1l1(Yti+1Yti)t(ki1)t(ki12)ifYt(ki2)Yt(ki11)=i=m+1l+1(Yti+1Yti)t(ki1)t(ki11)ifYt(ki2)Yt(ki11)=i=m1l+1(Yti+1Yti)L_{i}=\begin{cases}t(k_{i}^{2})-t(k_{i-1}^{2})&\mathrm{if}\ Y_{t(k_{i}^{2})}-Y_{t(k_{i-1}^{1})}=\sum_{i=m+1}^{l-1}(Y_{t_{i+1}}-Y_{t_{i}})\\ t(k_{i}^{2})-t(k_{i-1}^{1})&\mathrm{if}\ Y_{t(k_{i}^{2})}-Y_{t(k_{i-1}^{1})}=\sum_{i=m-1}^{l-1}(Y_{t_{i+1}}-Y_{t_{i}})\\ t(k_{i}^{1})-t(k_{i-1}^{2})&\mathrm{if}\ Y_{t(k_{i}^{2})}-Y_{t(k_{i-1}^{1})}=\sum_{i=m+1}^{l+1}(Y_{t_{i+1}}-Y_{t_{i}})\\ t(k_{i}^{1})-t(k_{i-1}^{1})&\mathrm{if}\ Y_{t(k_{i}^{2})}-Y_{t(k_{i-1}^{1})}=\sum_{i=m-1}^{l+1}(Y_{t_{i+1}}-Y_{t_{i}})\end{cases} (9)

Fig. 3 illustrates the overlapping regions for all four configurations described in Eq. 8.

t(ki11){\scriptstyle t(k_{i-1}^{1})}t(ki1){\scriptstyle t(k_{i}^{1})}t(ki12){\scriptstyle t(k_{i-1}^{2})}t(ki2){\scriptstyle t(k_{i}^{2})}Li{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}L_{i}} t(ki11){\scriptstyle t(k_{i-1}^{1})}t(ki1){\scriptstyle t(k_{i}^{1})}t(ki12){\scriptstyle t(k_{i-1}^{2})}t(ki2){\scriptstyle t(k_{i}^{2})}Li{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}L_{i}}
(a) (b)
t(ki11){\scriptstyle t(k_{i-1}^{1})}t(ki1){\scriptstyle t(k_{i}^{1})}t(ki12){\scriptstyle t(k_{i-1}^{2})}t(ki2){\scriptstyle t(k_{i}^{2})}Li{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}L_{i}} t(ki11){\scriptstyle t(k_{i-1}^{1})}t(ki1){\scriptstyle t(k_{i}^{1})}t(ki12){\scriptstyle t(k_{i-1}^{2})}t(ki2){\scriptstyle t(k_{i}^{2})}Li{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}L_{i}}
(c) (d)
Figure 3: Four configurations described in Eq.(8). For each case the overlapping interval is shown as LiL_{i}.

The next theorem says that in presence of high-frequency data, the performance of Hayashi-Yoshida covariance would be very similar to a scaled realized covariance. We make the following set of assumptions:

  1. (𝒞1\mathcal{C}_{1}):

    The log return process follows independent and stationary increment property.

  2. (𝒞2\mathcal{C}_{2}):

    The observation times (arrival process) of two stocks are independent Renewal processes and nn\rightarrow\infty as n1,n2n_{1},n_{2}\rightarrow \infty.

  3. (𝒞3\mathcal{C}_{3}):

    Estimation is based on paired data obtained by algorithm A0A_{0}.

Define a scaled (pairwise) realized covariance (SRCV) on a synchronized data {(Xt(ki1),Yt(ki2)):i=1,2,..,n}\big{\{}\big{(}X_{t(k^{1}_{i})},Y_{t(k^{2}_{i})}\big{)}:i=1,2,..,n\big{\}} as follows:

SRCV(X,Y)=iψiΔXt(ki1)ΔYt(ki2),\text{SRCV}(X,Y)=\sum_{i}\psi_{i}\Delta X_{t(k^{1}_{i})}\Delta Y_{t(k^{2}_{i})}, (10)

where

ψi=(t(ki1)t(ki11))(t(ki1)t(ki11))Li.\psi_{i}=\frac{\sqrt{(t(k^{1}_{i})-t(k^{1}_{i-1}))(t(k^{1}_{i})-t(k^{1}_{i-1}))}}{L_{i}}.

The following theorem says that for all practical purposes this estimator performs as good as the Hayashi-Yoshida estimator.

Theorem 2.1.

Under the assumptions 𝒞1𝒞3,\mathcal{C}_{1}-\mathcal{C}_{3}, SRCV is a consistent estimator of the pairwise integrated covariance.

Proof of Theorem 2.1 will be along the same line as in Theorem 1 of 36.

The ψi\psi_{i}’s in Eq.(10) are the scaling coefficients which are functions of the arrival (transaction) times only (does not depend on the stock prices at those time points). The conventional refresh-time periods t~j\tilde{t}_{j}s in Eq. 6 do not allow us to calculate these scaling coefficients. The algorithm 𝒜0\mathcal{A}_{0}, on the other hand, enable us to calculate the scaling coefficients as it preserves the actual arrival times of the synchronized pairs. The importance of this estimator will be evident in

As both Hayashi-Yoshida (pairwise) covariance and SRCV are consistent estimators, in presence of sufficient amount of data, they would be “close” to each other. For pp-dimensional process, the SRVC matrix can be written as the following:

SRCVp=i=1n(ΔXiΔXiTΨi),SRCV_{p}=\sum_{i=1}^{n}(\Delta X_{i}\Delta X_{i}^{T}\circ\Psi_{i}), (11)

where ψi\psi_{i} is a p×pp\times p symmetric matrix consisting of the pairwise scaling coefficients. This matrix will help us to make an important assumption necessary to determine the LSD of Hayashi-Yoshida matrix.

2.4 Inconsistency in high-dimension

From multivariate statistical theory we know that the sample covariance matrix is consistent for the population covariance matrix. In high-dimensional scenario, when the dimension grows at the same or a higher rate as the number of observations, this properties do not hold anymore. It can be shown that under a high-dimensional setup the following is true:

SΣ\nrightarrow0,\|S-\Sigma\|\nrightarrow 0,

where .\|.\| is the operator norm, SS is the usual sample covariance matrix and Σ\Sigma is the population covariance matrix.

But what impact would this inconsistency of sample covariance matrix make on the eigenvalues and the eigenvectors? Weyl’s theorem and Davis-Kahan theorems show that when the sample covariance matrix is not consistent, neither the sample eigenvalues and the eigenvectors are going to converge to their true counterpart 4, 37. Therefore it is worthwhile to study the limiting distribution of sample eigenvalues and its relation to the distribution of true eigenvalues.

For low dimensions, the Hayashi’s estimator is consistent estimator of ICV matrix. But for high dimensional stock price process, when the dimension grows at the same or a higher rate as the number of observations, neither RCV nor Hayashi-Yoshida estimator is consistent anymore. It is evident from the above discussion that the sample spectrum of Hayashi-Yoshida estimator deviates significantly form the true spectrum. In this paper, we study the asymptotic behavior of the distribution of eigenvalues of the Hayashi-Yoshida matrix.

3 Spectral Distribution

The empirical spectral distribution (ESD) of a symmetric (more generally Hermitian) matrix Ap×pA_{p\times p} is defined as

FpA(x)=1p#{jp:λjx}F_{p}^{A}(x)=\frac{1}{p}\#\{j\leq p:\ \lambda_{j}\leq x\}

where λj\lambda_{j}’s are the eigenvalues of the matrix AA and #E\#E denotes the cardinality of set EE. The limit distribution (FF) of ESD is called the limiting spectral distribution (LSD). One commonly used method of finding LSD is through Stieltjes transform.

Stieltjes transform: Let Ap×pA_{p\times p} is a Hermitian matrix and FpAF_{p}^{A} be its ESD. Then the Stieltjes transform of FpAF_{p}^{A} is defined as

sp(z)\displaystyle s_{p}(z) =1xz𝑑FpA(x)\displaystyle=\int\frac{1}{x-z}dF_{p}^{A}(x)
=1ptr(AzI)1\displaystyle=\frac{1}{p}tr(A-zI)^{-1}

where zD={z|𝕀(z)>0}z\in D=\{z\in\mathbb{C}|\mathbb{I}(z)>0\}, 𝕀(z)\mathbb{I}(z) being the imaginary part of zz. The importance of Stieltjes transformation in Random Matrix Theory is due to Theorem B.9 and Theorem B.10 38).

These theorems suggest that in order to determine the LSD, it is enough to obtain the limit of the Stieltjes transform.

4 Spectral Analysis of High Dimensional Hayashi’s Estimator

Based on our model (Eq. 1), the distribution of ΔXi\Delta X_{i} can be written like the following:

ΔXi=σ(tl11tl1𝑑W1tl12tl2𝑑W2..tl1ptlp𝑑Wp)=dσDiYi,\Delta X_{i}=\sigma\left(\begin{array}[]{c}\int_{t_{l-1}^{1}}^{t_{l}^{1}}dW_{1}\\ \int_{t_{l-1}^{2}}^{t_{l}^{2}}dW_{2}\\ .\\ .\\ \int_{t_{l-1}^{p}}^{t_{l}^{p}}dW_{p}\end{array}\right)\stackrel{{\scriptstyle d}}{{=}}\sigma D_{i}Y_{i},

where Di=diag(t(ki1)t(ki11),t(ki2)t(ki12),..,t(kp1)t(ki1p))D_{i}=\text{diag}\Big{(}\sqrt{t(k^{1}_{i})-t(k^{1}_{i-1})},\sqrt{t(k^{2}_{i})-t(k^{2}_{i-1})},..,\sqrt{t(k^{1}_{p})-t(k^{p}_{i-1})}\Big{)} is a p×pp\times p diagonal matrix and YiY_{i} is a pp-dimensional vector with its components being iid standard normal distribution. As a consequence, the Hayashi-Yoshida estimator has the same distribution as that of S0S_{0}:

S0=klσDkYkYlTDlσI(k,l).S_{0}=\sum_{k}\sum_{l}\sigma D_{k}Y_{k}Y_{l}^{T}D_{l}\sigma\circ I(k,l).

Hence, to determine the LSD of Hayashi-Yoshida estimator, it is enough to find the limit of the Stieltjes transform of S0S_{0}.

The set of assumptions (𝒜\mathcal{A}) necessary for determining the limiting spectral distribution of Hayashi-Yoshida estimator are given below:

  1. (𝒜1\mathcal{A}_{1})

    cn=pnc>0c_{n}=\frac{p}{n}\rightarrow c>0 as pp\rightarrow\infty.

  2. (𝒜2\mathcal{A}_{2})

    FΣdHF^{\Sigma}\stackrel{{\scriptstyle d}}{{\Rightarrow}}H (delta\neq\mathrm{delta} measureat 0\mathrm{measure\ at\ 0}) a.s. and HH has a finite second order moment.

  3. (𝒜3\mathcal{A}_{3})

    YljY_{l}^{j}’s (j=1(1)pj=1(1)p) are iid with mean 0, variance 1 and finite moments of all orders.

  4. (𝒜4\mathcal{A}_{4})

    \exists τl+\tau_{l}\in\mathbb{R^{+}} and let S=lτlΣ12YlYlTΣ12S=\sum_{l}\tau_{l}\Sigma^{\frac{1}{2}}Y_{l}Y_{l}^{T}\Sigma^{\frac{1}{2}} such that

    tr(S0S)2=o(p)a.s.\text{tr}(S_{0}-S)^{2}=o(p)~{}\text{a.s.} (12)

    and

    tr([(S0zI)1(SzI)1]2)=O(p)a.s.\text{tr}([(S_{0}-zI)^{-1}(S-zI)^{-1}]^{2})=O(p)~{}\text{a.s.} (13)

    where z=i.vz=i.v with vv is a sufficiently large positive number (S0zI)1(SzI)1(S_{0}-zI)^{-1}(S-zI)^{-1} being positive semidefinite.

  5. (𝒜5\mathcal{A}_{5})

    There exists κ,\kappa, such that maxnmaxl(nτl)κ.\mathrm{max_{n}max_{l}}(n\tau_{l})\leq\kappa. Also there exists a nonnegative cadlag process τt\tau^{\prime}_{t} such that

    limnl=1nl1nln|nτlτt|𝑑t=0.\underset{n\rightarrow\infty}{\text{lim}}\sum_{l=1}^{n}\int_{\frac{l-1}{n}}^{\frac{l}{n}}|n\tau_{l}-\tau^{\prime}_{t}|dt=0. (14)
  6. (𝒜6\mathcal{A}_{6})

    There exists a K<K<\infty and δ<1/6\delta<1/6 such that for all pp, ΣKpδ\|\Sigma\|\leq Kp^{\delta} almost surely.

Before stating the main theorem, we present a brief motivation of the assumption 𝒜4\mathcal{A}_{4}. In Sec. 2.3, we have seen that under a low dimensional setup, SRCV is also a consistent estimator for ICV. Therefore with high frequency data we can expect S0S_{0} and SRCVSRCV to be “close” to each other. The matrix SS replaces the matrix Ψi\Psi_{i} in Eq. (11) with a constant matrix where each element of the matrix is τi\tau_{i}. Assumption 𝒜4\mathcal{A}_{4} claims that under high-dimensional setup, even when both S0S_{0} and SRCVSRCV are inconsistent, upon choosing the τ\tau’s, SS and S0S_{0} have a closeness in the sense expressed by Eq. (12) & Eq. (13).
Now we state our main theorem.

Theorem 4.1.

Under the assumptions (𝒜\mathcal{A}), almost surely, ESD of S0S_{0} converges in distribution to a probability distribution with Stieltjes transform

s(z)=1z11λz01τt1+cτts~(z)𝑑t𝑑H(λ),s(z)=-\frac{1}{z}\int\frac{1}{1-\frac{\lambda}{z}\int_{0}^{1}\frac{\tau^{\prime}_{t}}{1+c\tau^{\prime}_{t}\tilde{s}(z)}dt}dH(\lambda), (15)

where s~(z)\tilde{s}(z) can be solved by the equation:

s~(z)=1zλλ1λz01τt1+cτts~(z)𝑑t𝑑H(λ).\tilde{s}(z)=-\frac{1}{z}\int_{\lambda}\frac{\lambda}{1-\frac{\lambda}{z}\int_{0}^{1}\frac{\tau^{\prime}_{t}}{1+c\tau^{\prime}_{t}\tilde{s}(z)}dt}dH(\lambda). (16)

This theorem establishes the Limiting spectral distribution of Hayashi-Yoshida estimator by determining its spectral distribution through the limiting spectral distribution of Σ\Sigma. In other words the theorem establishes the link between the limiting spectral distributions of Σ\Sigma and Hayashi-Yoshida estimator through its Stieltjes transform.
See Appendix for the lemmas that would lead to the theorem.
Now we are ready to prove Theorem 4.1.

Proof 4.2.

Define

S=Σl=1nτl1+τlal,S^{*}=\Sigma\sum_{l=1}^{n}\frac{\tau_{l}}{1+\tau_{l}a_{l}},

where al=YlΣ12(SlzI)1Σ12Yla_{l}=Y_{l}^{\prime}\Sigma^{\frac{1}{2}}(S_{l}-zI)^{-1}\Sigma^{\frac{1}{2}}Y_{l} and Sl=jlΣ12YjYjΣ12S_{l}=\sum_{j\neq l}\Sigma^{\frac{1}{2}}Y_{j}Y_{j}^{\prime}\Sigma^{\frac{1}{2}}. We denote the Stieltjes transform of FSF^{S} by sns_{n},

sn:=sn(z)=tr((SzI)1)p.s_{n}:=s_{n}(z)=\frac{tr((S-zI)^{-1})}{p}.

Similarly, Stieltjes transform of FS0F^{S_{0}}and Stieltjes transform of FSF^{S^{*}} are denoted by sn0s_{n}^{0} and sns^{*}_{n} respectively.
In order to show the convergence of sn(z)s_{n}(z) it is enough to show for all z=ivz=iv with v>0v>0 sufficiently large and the condition, in Lemma 8, is satisfied.
We will start by showing that sn0s_{n}^{0} and sns^{*}_{n} converge to the same limit.

snsn0\displaystyle s^{*}_{n}-s_{n}^{0} =1ptr[(SzI)1(S0zI)1]\displaystyle=\frac{1}{p}tr[(S^{*}-zI)^{-1}-(S_{0}-zI)^{-1}]
=1ptr[(SzI)1(SS)(SzI)1]+1ptr[(SzI)1(S0S)(S0zI)1]\displaystyle=\frac{1}{p}tr[(S^{*}-zI)^{-1}(S-S^{*})(S-zI)^{-1}]+\frac{1}{p}tr[(S-zI)^{-1}(S_{0}-S)(S_{0}-zI)^{-1}]
tr[(SzI)1(SS)(SzI)1]\displaystyle tr[(S^{*}-zI)^{-1}(S-S^{*})(S-zI)^{-1}]
=tr[(Σl=1nτl1+τlalzI)1[l=1nτlΣ12YlYlΣ12Σl=1nτl1+τlal](SzI)1]\displaystyle=tr[(\Sigma\sum_{l=1}^{n}\frac{\tau_{l}}{1+\tau_{l}a_{l}}-zI)^{-1}[\sum_{l=1}^{n}\tau_{l}\Sigma^{\frac{1}{2}}Y_{l}Y_{l}^{\prime}\Sigma^{\frac{1}{2}}-\Sigma\sum_{l=1}^{n}\frac{\tau_{l}}{1+\tau_{l}a_{l}}](S-zI)^{-1}]
=tr[(Σl=1nτl1+τlalzI)1[l=1nτlΣ12YlYlΣ12](SlzI)11+τlal]\displaystyle=tr[(\Sigma\sum_{l=1}^{n}\frac{\tau_{l}}{1+\tau_{l}a_{l}}-zI)^{-1}[\sum_{l=1}^{n}\tau_{l}\Sigma^{\frac{1}{2}}Y_{l}Y_{l}^{\prime}\Sigma^{\frac{1}{2}}]\frac{(S_{l}-zI)^{-1}}{1+\tau_{l}a_{l}}]
tr[(Σl=1nτl1+τlalzI)1[Σl=1nτl1+τlal](SzI)1]\displaystyle-tr[(\Sigma\sum_{l=1}^{n}\frac{\tau_{l}}{1+\tau_{l}a_{l}}-zI)^{-1}[\Sigma\sum_{l=1}^{n}\frac{\tau_{l}}{1+\tau_{l}a_{l}}](S-zI)^{-1}]
=l=1nτl1+τlal(YlΣ12(SlzI)1(Σl=1nτl1+τlalzI)1Σ12Yl\displaystyle=\sum_{l=1}^{n}\frac{\tau_{l}}{1+\tau_{l}a_{l}}(Y_{l}^{\prime}\Sigma^{\frac{1}{2}}(S_{l}-zI)^{-1}(\Sigma\sum_{l=1}^{n}\frac{\tau_{l}}{1+\tau_{l}a_{l}}-zI)^{-1}\Sigma^{\frac{1}{2}}Y_{l}
tr[Σ12(SzI)1(Σl=1nτl1+τlalzI)1Σ12])\displaystyle-tr[\Sigma^{\frac{1}{2}}(S-zI)^{-1}(\Sigma\sum_{l=1}^{n}\frac{\tau_{l}}{1+\tau_{l}a_{l}}-zI)^{-1}\Sigma^{\frac{1}{2}}])
tr[(SzI)1(S0S)(S0zI)1]\displaystyle tr[(S-zI)^{-1}(S_{0}-S)(S_{0}-zI)^{-1}] tr(S0S)2×tr([(SzI)1(S0zI)1]2)\displaystyle\leq\sqrt{tr(S_{0}-S)^{2}\times tr([(S-zI)^{-1}(S_{0}-zI)^{-1}]^{2})}

According to our assumption, 𝒜6\mathcal{A}_{6}, l=1nτl1+τlalnmax(τl)κ\sum_{l=1}^{n}\frac{\tau_{l}}{1+\tau_{l}a_{l}}\leq n\ \mathrm{max}(\tau_{l})\leq\kappa.
It is enough to show that

Ξ\displaystyle\Xi =1pmaxl(YlΣ12(SlzI)1(Σl=1nτl1+τlalzI)1Σ12Yl\displaystyle=\frac{1}{p}\mathrm{max_{l}}(Y_{l}^{\prime}\Sigma^{\frac{1}{2}}(S_{l}-zI)^{-1}(\Sigma\sum_{l=1}^{n}\frac{\tau_{l}}{1+\tau_{l}a_{l}}-zI)^{-1}\Sigma^{\frac{1}{2}}Y_{l}
tr[Σ12(SzI)1(Σl=1nτl1+τlalzI)1Σ12])0a.s.\displaystyle-tr[\Sigma^{\frac{1}{2}}(S-zI)^{-1}(\Sigma\sum_{l=1}^{n}\frac{\tau_{l}}{1+\tau_{l}a_{l}}-zI)^{-1}\Sigma^{\frac{1}{2}}])\rightarrow 0\ \mathrm{a.s.}

We write Ξ=I+II+III+IV\Xi=I+II+III+IV where

I=\displaystyle I= 1pmaxl(YlΣ12(SlzI)1(Σl=1nτl1+τlalzI)1Σ12Yl\displaystyle\frac{1}{p}\mathrm{max_{l}}(Y_{l}^{\prime}\Sigma^{\frac{1}{2}}(S_{l}-zI)^{-1}(\Sigma\sum_{l=1}^{n}\frac{\tau_{l}}{1+\tau_{l}a_{l}}-zI)^{-1}\Sigma^{\frac{1}{2}}Y_{l}
YlΣ12(SlzI)1(Σjlτj1+τjbjlzI)1Σ12Yl)\displaystyle-Y_{l}^{\prime}\Sigma^{\frac{1}{2}}(S_{l}-zI)^{-1}(\Sigma\sum_{j\neq l}\frac{\tau_{j}}{1+\tau_{j}b_{j}^{l}}-zI)^{-1}\Sigma^{\frac{1}{2}}Y_{l})
II=\displaystyle II= 1pmaxl(YlΣ12(SlzI)1(Σjlτj1+τjbjlzI)1Σ12Yl\displaystyle\frac{1}{p}\mathrm{max_{l}}(Y_{l}^{\prime}\Sigma^{\frac{1}{2}}(S_{l}-zI)^{-1}(\Sigma\sum_{j\neq l}\frac{\tau_{j}}{1+\tau_{j}b_{j}^{l}}-zI)^{-1}\Sigma^{\frac{1}{2}}Y_{l}
tr[Σ12(SlzI)1(Σjlτj1+τjbjlzI)1Σ12])\displaystyle-tr[\Sigma^{\frac{1}{2}}(S_{l}-zI)^{-1}(\Sigma\sum_{j\neq l}\frac{\tau_{j}}{1+\tau_{j}b_{j}^{l}}-zI)^{-1}\Sigma^{\frac{1}{2}}])
III=\displaystyle III= 1pmaxl(tr[Σ12(SlzI)1(Σjlτj1+τjbjlzI)1Σ12]\displaystyle\frac{1}{p}\mathrm{max_{l}}(tr[\Sigma^{\frac{1}{2}}(S_{l}-zI)^{-1}(\Sigma\sum_{j\neq l}\frac{\tau_{j}}{1+\tau_{j}b_{j}^{l}}-zI)^{-1}\Sigma^{\frac{1}{2}}]
tr[Σ12(SlzI)1(Σl=1nτl1+τlalzI)1Σ12])\displaystyle-tr[\Sigma^{\frac{1}{2}}(S_{l}-zI)^{-1}(\Sigma\sum_{l=1}^{n}\frac{\tau_{l}}{1+\tau_{l}a_{l}}-zI)^{-1}\Sigma^{\frac{1}{2}}])
IV=\displaystyle IV= 1pmaxl(tr[Σ12(SlzI)1(Σl=1nτl1+τlalzI)1Σ12]\displaystyle\frac{1}{p}\mathrm{max_{l}}(tr[\Sigma^{\frac{1}{2}}(S_{l}-zI)^{-1}(\Sigma\sum_{l=1}^{n}\frac{\tau_{l}}{1+\tau_{l}a_{l}}-zI)^{-1}\Sigma^{\frac{1}{2}}]
tr[Σ12(SzI)1(Σl=1nτl1+τlalzI)1Σ12])\displaystyle-tr[\Sigma^{\frac{1}{2}}(S-zI)^{-1}(\Sigma\sum_{l=1}^{n}\frac{\tau_{l}}{1+\tau_{l}a_{l}}-zI)^{-1}\Sigma^{\frac{1}{2}}])

where bjl=YjΣ12(Sj,lzI)1Σ12Yjb_{j}^{l}=Y_{j}^{\prime}\Sigma^{\frac{1}{2}}(S_{j,l}-zI)^{-1}\Sigma^{\frac{1}{2}}Y_{j} with Sj,l=ij,lΣ12YiYiΣ12S_{j,l}=\sum_{i\neq j,l}\Sigma^{\frac{1}{2}}Y_{i}Y_{i}^{\prime}\Sigma^{\frac{1}{2}}.

It is sufficient to show that I,II,III,IV0a.s.\mathrm{I},\mathrm{II},\mathrm{III},\mathrm{IV}\rightarrow 0\ \mathrm{a.s.}

Convergence of I\mathrm{I} and II\mathrm{II} are followed by Lemma A.10 and Lemma A.12. Convergence of III\mathrm{III} and IV\mathrm{IV} can be proved similarly.

Now due to Assumption 5, we have

1ptr(l=1nτl1+τlalΣzI)11z11λz01τt1+cτts~(z)𝑑t𝑑H(λ).\frac{1}{p}tr(\sum_{l=1}^{n}\frac{\tau_{l}}{1+\tau_{l}a_{l}}\Sigma-zI)^{-1}\rightarrow-\frac{1}{z}\int\frac{1}{1-\frac{\lambda}{z}\int_{0}^{1}\frac{\tau^{\prime}_{t}}{1+c\tau^{\prime}_{t}\tilde{s}(z)}dt}dH(\lambda).

As

1ptr(l=1nτl1+τlalΣzI)11ptr(SzI)10,\frac{1}{p}tr(\sum_{l=1}^{n}\frac{\tau_{l}}{1+\tau_{l}a_{l}}\Sigma-zI)^{-1}-\frac{1}{p}tr(S-zI)^{-1}\rightarrow 0,

we get

s(z)=1z11λz01τt1+cτts~(z)𝑑t𝑑H(λ).s(z)=-\frac{1}{z}\int\frac{1}{1-\frac{\lambda}{z}\int_{0}^{1}\frac{\tau^{\prime}_{t}}{1+c\tau^{\prime}_{t}\tilde{s}(z)}dt}dH(\lambda). (17)

The fact that 01τt1+cτts~(z)𝑑t0\int_{0}^{1}\frac{\tau^{\prime}_{t}}{1+c\tau^{\prime}_{t}\tilde{s}(z)}dt\neq 0 and Real(01τt1+cs~(z)𝑑t)0Real(\int_{0}^{1}\frac{\tau^{\prime}_{t}}{1+c\tilde{s}(z)}dt)\geq 0 implies |s(z)|<1|z||s(z)|<\frac{1}{|z|} and therefore 1+zs(z)01+zs(z)\neq 0. So from Lemma 5, it is clear that s~(z)0\tilde{s}(z)\neq 0.

Now

01τt1+cs~(z)τt𝑑t\displaystyle\int_{0}^{1}\frac{\tau^{\prime}_{t}}{1+c\tilde{s}(z)\tau^{\prime}_{t}}dt =1cs~(z)(10111+cs~(z)τt𝑑t)\displaystyle=\frac{1}{c\tilde{s}(z)}(1-\int_{0}^{1}\frac{1}{1+c\tilde{s}(z)\tau^{\prime}_{t}}dt)
=1cs~(z)(1(1c(1+zs(z))))\displaystyle=\frac{1}{c\tilde{s}(z)}(1-(1-c(1+zs(z))))
=1+zs(z)s~(z)\displaystyle=\frac{1+zs(z)}{\tilde{s}(z)}

Also, equation 17 implies

1+zs(z)\displaystyle 1+zs(z) =1z01τt1+cτts~(z)𝑑tλλ1λz01τt1+cτts~(z)𝑑t𝑑H(λ)\displaystyle=-\frac{1}{z}\int_{0}^{1}\frac{\tau^{\prime}_{t}}{1+c\tau^{\prime}_{t}\tilde{s}(z)}dt\int_{\lambda}\frac{\lambda}{1-\frac{\lambda}{z}\int_{0}^{1}\frac{\tau^{\prime}_{t}}{1+c\tau^{\prime}_{t}\tilde{s}(z)}dt}dH(\lambda)
01τt1+cs~(z)τt𝑑t\displaystyle\implies\int_{0}^{1}\frac{\tau^{\prime}_{t}}{1+c\tilde{s}(z)\tau^{\prime}_{t}}dt =1zs~(z)01τt1+cτts~(z)𝑑tλλ1λz01τt1+cτts~(z)𝑑t𝑑H(λ)\displaystyle=-\frac{1}{z\tilde{s}(z)}\int_{0}^{1}\frac{\tau^{\prime}_{t}}{1+c\tau^{\prime}_{t}\tilde{s}(z)}dt\int_{\lambda}\frac{\lambda}{1-\frac{\lambda}{z}\int_{0}^{1}\frac{\tau^{\prime}_{t}}{1+c\tau^{\prime}_{t}\tilde{s}(z)}dt}dH(\lambda)
s~(z)\displaystyle\implies\tilde{s}(z) =1zλλ1λz01τt1+cτts~(z)𝑑t𝑑H(λ)\displaystyle=-\frac{1}{z}\int_{\lambda}\frac{\lambda}{1-\frac{\lambda}{z}\int_{0}^{1}\frac{\tau^{\prime}_{t}}{1+c\tau^{\prime}_{t}\tilde{s}(z)}dt}dH(\lambda)

Due to Lemma 6 and Lemma 7, 1zl=1nτl1+alτlQ1-\frac{1}{z}\sum_{l=1}^{n}\frac{\tau_{l}}{1+a_{l}\tau_{l}}\in Q_{1} and 1ptr(SzI)1Q1\frac{1}{p}tr(S-zI)^{-1}\in Q_{1}. Same will be true for their limits. Now to show that s(z)s(z) is not unique it is enough to show that if there are two solutions s1(z)s_{1}(z) and s2(z)s_{2}(z) (and therefore s~1(z),s~2(z)\tilde{s}_{1}(z),\tilde{s}_{2}(z)) then s1=s2s_{1}=s_{2}. If possible let there are two limiting spectral distributions s1s_{1} and s2s_{2} such that s1s2s_{1}\neq s_{2}. To show the contradiction, it is enough to show s~1(z)=s~2(z)\tilde{s}_{1}(z)=\tilde{s}_{2}(z).

Note that,

01τt1+cτts~1(z)𝑑t01τt1+cτts~2(z)𝑑t=01c(τt)2(s~2(z)s~1(z))(1+cτts~1(z))(1+cτts~2(z))𝑑t\int_{0}^{1}\frac{\tau^{\prime}_{t}}{1+c\tau^{\prime}_{t}\tilde{s}_{1}(z)}dt-\int_{0}^{1}\frac{\tau^{\prime}_{t}}{1+c\tau^{\prime}_{t}\tilde{s}_{2}(z)}dt=\int_{0}^{1}\frac{c(\tau^{\prime}_{t})^{2}(\tilde{s}_{2}(z)-\tilde{s}_{1}(z))}{(1+c\tau^{\prime}_{t}\tilde{s}_{1}(z))(1+c\tau^{\prime}_{t}\tilde{s}_{2}(z))}dt

But,

s~1(z)s~2(z)=1zλ[λ1λz01τt1+cτts~1(z)𝑑tλ1λz01τt1+cτts~2(z)𝑑t]𝑑H(λ)\tilde{s}_{1}(z)-\tilde{s}_{2}(z)=-\frac{1}{z}\int_{\lambda\in\mathbb{R}}\Bigg{[}\frac{\lambda}{1-\frac{\lambda}{z}\int_{0}^{1}\frac{\tau^{\prime}_{t}}{1+c\tau^{\prime}_{t}\tilde{s}_{1}(z)}dt}-\frac{\lambda}{1-\frac{\lambda}{z}\int_{0}^{1}\frac{\tau^{\prime}_{t}}{1+c\tau^{\prime}_{t}\tilde{s}_{2}(z)}dt}\Bigg{]}dH(\lambda)

On simplification, this gives

1=cz201(τt)2(1+cτts~1(z))(1+cτts~2(z))𝑑t×01λ2(1λz01τt1+cτts~1(z)𝑑t)(1λz01τt1+cτts~2(z)𝑑t)𝑑H(λ)1=\frac{c}{z^{2}}\int_{0}^{1}\frac{(\tau^{\prime}_{t})^{2}}{\big{(}1+c\tau^{\prime}_{t}\tilde{s}_{1}(z)\big{)}\big{(}1+c\tau^{\prime}_{t}\tilde{s}_{2}(z)\big{)}}dt\times\int_{0}^{1}\frac{\lambda^{2}}{\big{(}1-\frac{\lambda}{z}\int_{0}^{1}\frac{\tau^{\prime}_{t}}{1+c\tau^{\prime}_{t}\tilde{s}_{1}(z)}dt\big{)}\big{(}1-\frac{\lambda}{z}\int_{0}^{1}\frac{\tau^{\prime}_{t}}{1+c\tau^{\prime}_{t}\tilde{s}_{2}(z)}dt\big{)}}dH(\lambda) (18)

As s~1,s~2Q1\tilde{s}_{1},\tilde{s}_{2}\in Q_{1} ,

|01(τt)2(1+cτts~1(z))(1+cτts~2(z))𝑑t|01(τt)2𝑑t<.\Bigg{|}\int_{0}^{1}\frac{(\tau^{\prime}_{t})^{2}}{\Big{(}1+c\tau^{\prime}_{t}\tilde{s}_{1}(z)\Big{)}\Big{(}1+c\tau^{\prime}_{t}\tilde{s}_{2}(z)\Big{)}}dt\Bigg{|}\leq\int_{0}^{1}(\tau^{\prime}_{t})^{2}dt<\infty.

And 1z01τt1+cτts~i(z)𝑑t-\frac{1}{z}\int_{0}^{1}\frac{\tau^{\prime}_{t}}{1+c\tau^{\prime}_{t}\tilde{s}_{i}(z)}dt, i=1,2i=1,2 implies

|01λ2(1λz01τt1+cτts~1(z)𝑑t)(1λz01τt1+cτts~2(z)𝑑t)𝑑H(λ)|λ2𝑑H(λ)<.\Bigg{|}\int_{0}^{1}\frac{\lambda^{2}}{\Big{(}1-\frac{\lambda}{z}\int_{0}^{1}\frac{\tau^{\prime}_{t}}{1+c\tau^{\prime}_{t}\tilde{s}_{1}(z)}dt)(1-\frac{\lambda}{z}\int_{0}^{1}\frac{\tau^{\prime}_{t}}{1+c\tau^{\prime}_{t}\tilde{s}_{2}(z)}dt\Big{)}}dH(\lambda)\Bigg{|}\leq\int\lambda^{2}dH(\lambda)<\infty.

So for z=ivz=iv, with vv sufficiently large, Eq. (18) can not be true. So s(z)s(z) is unique.

The above theorem is true for time-varying instantaneous covolatility process with a little stronger set assumptions. Following 27, we can assume that time varying covolatility process can be decomposed in two parts: a time varying cadlag process and a symmetric matrix that is not varying with time. Formally, σt=γtΣ~12\sigma_{t}=\gamma_{t}\tilde{\Sigma}^{\frac{1}{2}} where Σ~12\tilde{\Sigma}^{\frac{1}{2}} does not depend on time and as mentioned γt\gamma_{t} is a time-varying cadlag process.
If we assume this, then ΔXl\Delta X_{l} has the same distribution as the following-

ΔXl=Σ~12(tl11tl1γs𝑑W1tl12tl2γs𝑑W2..tl1ptlpγs𝑑Wp)=d(Σ~12A)DlYl,\Delta X_{l}=\tilde{\Sigma}^{\frac{1}{2}}\left(\begin{array}[]{c}\int_{t_{l-1}^{1}}^{t_{l}^{1}}\gamma_{s}dW_{1}\\ \int_{t_{l-1}^{2}}^{t_{l}^{2}}\gamma_{s}dW_{2}\\ .\\ .\\ \int_{t_{l-1}^{p}}^{t_{l}^{p}}\gamma_{s}dW_{p}\end{array}\right)\stackrel{{\scriptstyle d}}{{=}}(\tilde{\Sigma}^{\frac{1}{2}}\circ A)D_{l}Y_{l},

where Ylp=(Ylj)1jpY_{l}^{p}=(Y_{l}^{j})_{1\leq j\leq p}; YljY_{l}^{j} ’s are iid normal with mean 0 and variance 11 for p=1,2,p=1,2,... and 1ln1\leq l\leq n and A=(aij)A=(a_{ij}) when aij=tl1itl1jtlitljγs2I((tl1i,tli)I(tl1j,tlj))𝑑sa_{ij}=\int_{t_{l-1}^{i}\vee t_{l-1}^{j}}^{t_{l}^{i}\wedge t_{l}^{j}}\gamma_{s}^{2}I((t_{l-1}^{i},t_{l}^{i})\cap I(t_{l-1}^{j},t_{l}^{j}))ds, and
Dl=diag(tl11tl1γs2𝑑s,tl12tl2γs2𝑑s,,tl1ptlpγs2𝑑s)D_{l}=diag(\sqrt{\int_{t_{l-1}^{1}}^{t_{l}^{1}}\gamma_{s}^{2}ds},\sqrt{\int_{t_{l-1}^{2}}^{t_{l}^{2}}\gamma_{s}^{2}ds},...,\sqrt{\int_{t_{l-1}^{p}}^{t_{l}^{p}}\gamma_{s}^{2}ds}).
Therefore we are interested in the spectral distribution of

S0=kl{(Σ12A)DkxkxlTDl(Σ12A)I(k,l)}S_{0}=\sum_{k}\sum_{l}\{(\Sigma^{\frac{1}{2}}\circ A)D_{k}x_{k}x_{l}^{T}D_{l}(\Sigma^{\frac{1}{2}}\circ A)\circ I(k,l)\}

as both S0S_{0} and ΣpHY\Sigma_{p}^{HY} have the same LSD.
Suppose now we denote the Limiting spectral distribution of Σ~\tilde{\Sigma} as HH. Then along with other assumptions of 𝒜\mathcal{A} we need the following additional assumptions

  1. (𝒜7\mathcal{A}_{7})

    σt=γtΣ~12\sigma_{t}=\gamma_{t}\tilde{\Sigma}^{\frac{1}{2}} where Σ~12\tilde{\Sigma}^{\frac{1}{2}} does not depend on time and γt\gamma_{t} is a time-varying cadlag process.

  2. (𝒜8\mathcal{A}_{8})

    tl1jtljγs2𝑑s\int_{t_{l-1}^{j}}^{t_{l}^{j}}\gamma_{s}^{2}ds are independent of YlY_{l}.

Then the result holds for time-varying covolatility process. How these conditions impose constraints on the time-varying covolatility or more specifically the cadlag process is a separate but interesting question to study. We write the theorem below

Theorem 4.3.

Under the assumptions (𝒜1\mathcal{A}_{1}-𝒜8\mathcal{A}_{8}), almost surely, ESD of S0S_{0} converges in distribution to a probability distribution with Stieltjes transform given by equations (15)-(16).

5 Spectral Analysis of Hayashi’s Estimator when p/n0p/n\rightarrow 0

Due to the fact that Hayashi’s estimator is unbiased we will be concerned about the following matrix (let us call S~\tilde{S}),

S~=np(S0nΣpn).\tilde{S}=\sqrt{\frac{n}{p}}(\frac{S_{0}}{n}-\frac{\Sigma_{p}}{n}).

Like the previous chapter here also we have to take some assumptions, we will call it \mathcal{B}.

  1. (\mathcal{B}1)

    : limpn0\mathrm{lim}~{}\frac{p}{n}\rightarrow 0 as pp\rightarrow\infty and nn\rightarrow\infty.

  2. (\mathcal{B}2)

    : ZijsZ_{ij}^{\prime}s (1ip1\leq i\leq p, 1jn1\leq j\leq n) are iid Gaussian random variables with E(Zij)=0E(Z_{ij})=0, E|Zij|2=1E|Z_{ij}|^{2}=1.

  3. (\mathcal{B}3)

    : FΣpLFH(δ{0})F^{\Sigma_{p}}\stackrel{{\scriptstyle L}}{{\rightarrow}}F^{H}(\neq\delta_{\{0\}}) as pp\rightarrow\infty where FHF^{H} is a distribution function.

  4. (\mathcal{B}4)

    : Define, S¯=np(SnΣpn)\bar{S}=\sqrt{\frac{n}{p}}(\frac{S}{n}-\frac{\Sigma_{p}}{n}). Then S¯S~\bar{S}-\tilde{S}, (S~zI)1(\tilde{S}-zI)^{-1} and (S¯zI)1(\bar{S}-zI)^{-1} are positive definite, tr(S0S)=O(p)tr(S_{0}-S)=O(p), tr(S¯zI)1=O(1)tr(\bar{S}-zI)^{-1}=O(1), tr(S~zI)1=O(1)tr(\tilde{S}-zI)^{-1}=O(1) and (1nl=1nτl)tr(Σ)=O(p)(1-n\sum_{l=1}^{n}\tau_{l})tr(\Sigma)=O(p).

  5. (\mathcal{B}5)

    : diag(τ1,τ2,,τn)||\mathrm{diag}(\tau_{1},\tau_{2},...,\tau_{n})|| is bounded above.

  6. (\mathcal{B}6)

    : 1n(l=1nτl)τ>0\frac{1}{n}(\sum_{l=1}^{n}\tau_{l})\rightarrow\tau>0 and 1n(l=1nτl2)τ¯>0\frac{1}{n}(\sum_{l=1}^{n}\tau_{l}^{2})\rightarrow\bar{\tau}>0 as nn\rightarrow\infty

Now we are ready to state the main theorem.

Theorem 5.1.

If the above assumptions (\mathcal{B}) are true then the empirical spectral distribution of np(1nS01nΣp)\sqrt{\frac{n}{p}}(\frac{1}{n}S_{0}-\frac{1}{n}\Sigma_{p}) almost surely converges weakly to a nonrandom distribution FF as nn\rightarrow\infty, whose Stieltjes transform s(z)s(z) is determined by the following system of equations:

{s(z)=dH(λ)z+τ¯λβ(z)β(z)=λdH(λ)z+τ¯λβ(z)\begin{cases}s(z)=&-\int\frac{dH(\lambda)}{z+\bar{\tau}\lambda\beta(z)}\\ \beta(z)=&-\int\frac{\lambda dH(\lambda)}{z+\bar{\tau}\lambda\beta(z)}\end{cases}

for any z+z\in\mathbb{C}_{+}.

Proof of this theorem will be in the similar path as in 39. Before proving the theorem we will define some quantities and make some observations. Let Σ=UΛU\Sigma=U^{*}\Lambda U be the spectral decomposition of Σ\Sigma. Now define,

W=UΣ12{lτl12Zlel}.W=U\Sigma^{\frac{1}{2}}\{\sum_{l}\tau_{l}^{\frac{1}{2}}Z_{l}e_{l}^{*}\}.

Let wkw_{k} be the kkth row of WW and WkW_{k} be the matrix after deleting the kkth row. Define,

hk=\displaystyle h_{k}= npWkwk,\displaystyle\sqrt{\frac{n}{p}}W_{k}w_{k},
M=\displaystyle M= np(WW(lτl)Λ),\displaystyle\sqrt{\frac{n}{p}}(WW^{*}-(\sum_{l}\tau_{l})\Lambda),
Mk=\displaystyle M_{k}= np(WkW(lτl)Λk),\displaystyle\sqrt{\frac{n}{p}}(W_{k}W^{*}-(\sum_{l}\tau_{l})\Lambda_{k}),
M¯k=\displaystyle\bar{M}_{k}= np(WkWk(lτl)Λk),\displaystyle\sqrt{\frac{n}{p}}(W_{k}W_{k}^{*}-(\sum_{l}\tau_{l})\Lambda_{k}),
tkk=\displaystyle t_{kk}= np(wkwk(lτl)λk),\displaystyle\sqrt{\frac{n}{p}}(w_{k}w_{k}^{*}-(\sum_{l}\tau_{l})\lambda_{k}),

where Λk\Lambda_{k} is the matrix obtained by deleting kkth diagonal element of Λ\Lambda.
Now we will make some remarks. Justifications of the remarks are given in the appendix.

  1. Remark 1:

    M=k=1pek(hk+tkkek)M=\sum_{k=1}^{p}e_{k}(h_{k}+t_{kk}e_{k})^{*} and M=Mk+ek(hk+tkkek)M=M_{k}+e_{k}(h_{k}+t_{kk}e_{k})^{*}.

  2. Remark 2:

    tr(I+z(MzI)1)=k=1p[(hk+tkkek)(MkzI)1ek]/[1+(hk+tkkek)(MkzI)1ek].tr(I+z(M-zI)^{-1})=\sum_{k=1}^{p}\big{[}(h_{k}+t_{kk}e_{k})^{*}(M_{k}-zI)^{-1}e_{k}\big{]}/\big{[}1+(h_{k}+t_{kk}e_{k})^{*}(M_{k}-zI)^{-1}e_{k}\big{]}.

  3. Remark 3:

    (hk+tkkek)(MkzI)1ek=hk(M¯kzI)1hktkk/z.(h_{k}+t_{kk}e_{k})^{*}(M_{k}-zI)^{-1}e_{k}=h_{k}^{*}\bar{(M}_{k}-zI)^{-1}h_{k}-t_{kk}/z.

  4. Remark 4:

    E(hk(M¯kzI)1hk(l=1nτl2)λktr((M¯kzI)1Λk)/np)=0.E(h_{k}^{*}\bar{(M}_{k}-zI)^{-1}h_{k}-(\sum_{l=1}^{n}\tau_{l}^{2})\lambda_{k}tr(\bar{(M}_{k}-zI)^{-1}\Lambda_{k})/np)=0.

Now we are ready to prove Theorem 5.1.

Proof 5.2 (Proof of Theorem 5.1).

Suppose the Stiletjes transform of S¯n\bar{S}_{n} is s¯n(z)\bar{s}_{n}(z). So s¯n(z)=1ptr(S¯nzI)1\bar{s}_{n}(z)=\frac{1}{p}\mathrm{tr}(\bar{S}_{n}-zI)^{-1}.
Observe that, z1I+(S¯nzI)1=z1S¯n(S¯nzI)1.z^{-1}I+(\bar{S}_{n}-zI)^{-1}=z^{-1}\bar{S}_{n}(\bar{S}_{n}-zI)^{-1}. This implies the following:

(S¯nzI)1=\displaystyle(\bar{S}_{n}-zI)^{-1}= z1S¯n(S¯nzI)1z1I\displaystyle z^{-1}\bar{S}_{n}(\bar{S}_{n}-zI)^{-1}-z^{-1}I
s¯n(z)=\displaystyle\implies\hskip 5.69046pt\bar{s}_{n}(z)= z1+z1ptr(S¯n(S¯nzI)1).\displaystyle-z^{-1}+\frac{z^{-1}}{p}tr(\bar{S}_{n}(\bar{S}_{n}-zI)^{-1}). (19)

Notice that,

s¯n(z)s~n(z)\displaystyle\bar{s}_{n}(z)-\tilde{s}_{n}(z) =1p[tr(S¯nzI)1tr(S~nzI)1]\displaystyle=\frac{1}{p}[tr(\bar{S}_{n}-zI)^{-1}-tr(\tilde{S}_{n}-zI)^{-1}]
=1ptr[(S~nS¯n)(S~nzI)1(S¯nzI)1]\displaystyle=\frac{1}{p}tr[(\tilde{S}_{n}-\bar{S}_{n})(\tilde{S}_{n}-zI)^{-1}(\bar{S}_{n}-zI)^{-1}]
1ptr(S~nS¯n)tr(S~nzI)1tr(S¯nzI)1\displaystyle\leq\frac{1}{p}tr(\tilde{S}_{n}-\bar{S}_{n})tr(\tilde{S}_{n}-zI)^{-1}tr(\bar{S}_{n}-zI)^{-1}

Now,

tr(S~nS¯n)\displaystyle tr(\tilde{S}_{n}-\bar{S}_{n}) =tr[np{(S0/nS/n)(1nΣ(l=1nτl)Σ)}]\displaystyle=tr\Big{[}\sqrt{\frac{n}{p}}\Big{\{}(S_{0}/n-S/n)-(\frac{1}{n}\Sigma-(\sum_{l=1}^{n}\tau_{l})\Sigma)\Big{\}}\Big{]}
=1nptr(S0S)1np[1n(l=1nτl)tr(Σ)]\displaystyle=\sqrt{\frac{1}{np}}tr(S_{0}-S)-\sqrt{\frac{1}{np}}\Big{[}1-n(\sum_{l=1}^{n}\tau_{l})tr(\Sigma)\Big{]}
=o(p)a.s.\displaystyle=o(p)~{}\text{a.s.}

The last line of the above derivation is a consequence of Assumption 4\mathcal{B}4. Moreover as tr(S¯zI)1=O(1)tr(\bar{S}-zI)^{-1}=O(1) and tr(S~zI)1=O(1)tr(\tilde{S}-zI)^{-1}=O(1), we have s¯n(z)s~n(z)0,\bar{s}_{n}(z)-\tilde{s}_{n}(z)\rightarrow 0, a.s.
This means that we can derive valuable information about s~n(z)\tilde{s}_{n}(z) by studying the spectral distribution of S¯n(z)\bar{S}_{n}(z).
Note that, according to our definitions

tr(S¯nzI)1\displaystyle tr(\bar{S}_{n}-zI)^{-1} =tr[np(Σ12{lτl12Ylel})(Σ12{lτl12Ylel}(lτl)Σ))zI]1\displaystyle=tr[\sqrt{\frac{n}{p}}(\Sigma^{\frac{1}{2}}\{\sum_{l}\tau_{l}^{\frac{1}{2}}Y_{l}e_{l}^{*}\})(\Sigma^{\frac{1}{2}}\{\sum_{l}\tau_{l}^{\frac{1}{2}}Y_{l}e_{l}^{*}\}^{*}-(\sum_{l}\tau_{l})\Sigma))-zI]^{-1}
=tr[npU(UΣ12{lτl12Ylel})({lτl12Ylel}Σ12UU(lτl)UΛU))zI]1\displaystyle=tr[\sqrt{\frac{n}{p}}U^{*}(U\Sigma^{\frac{1}{2}}\{\sum_{l}\tau_{l}^{\frac{1}{2}}Y_{l}e_{l}^{*}\})(\{\sum_{l}\tau_{l}^{\frac{1}{2}}Y_{l}e_{l}^{*}\}^{*}\Sigma^{\frac{1}{2}}U^{*}U-(\sum_{l}\tau_{l})U^{*}\Lambda U))-zI]^{-1}
=tr[np(WW(lτl)Λ)zI]1.\displaystyle=tr[\sqrt{\frac{n}{p}}(WW^{*}-(\sum_{l}\tau_{l})\Lambda)-zI]^{-1}. (20)

Now from Eq. (19) we have,

s¯n(z)\displaystyle\bar{s}_{n}(z) =z1+z1ptr(np(WW(lτl)Λ)(np((WW(lτl)Λ)zI)1)\displaystyle=-z^{-1}+\frac{z^{-1}}{p}tr(\sqrt{\frac{n}{p}}(WW^{*}-(\sum_{l}\tau_{l})\Lambda)(\sqrt{\frac{n}{p}}((WW^{*}-(\sum_{l}\tau_{l})\Lambda)-zI)^{-1})
=z1+z1ptr{M(MzI)1)\displaystyle=-z^{-1}+\frac{z^{-1}}{p}tr\{M(M-zI)^{-1})
=z1pk=1p11+(hk+tkkek)(MkzI)1ek\displaystyle=-\frac{z^{-1}}{p}\sum_{k=1}^{p}\frac{1}{1+(h_{k}+t_{kk}e_{k})^{*}(M_{k}-zI)^{-1}e_{k}}
=1pk=1p1ztkk+nλk(lτn2)ptr(Λk(M¯kzI)1)+ϵ1k,\displaystyle=-\frac{1}{p}\sum_{k=1}^{p}\frac{1}{z-t_{kk}+\frac{n\lambda_{k}(\sum_{l}\tau_{n}^{2})}{p}tr(\Lambda_{k}\bar{(M}_{k}-zI)^{-1})+\epsilon_{1k}},

Define βn=1ptr(S¯nzI)1Σ\beta_{n}=\frac{1}{p}tr(\bar{S}_{n}-zI)^{-1}\Sigma, then similar derivation will lead to

βn(z)=1pk=1pλkz+hk(M¯kzI)1hktkk.\beta_{n}(z)=-\frac{1}{p}\sum_{k=1}^{p}\frac{\lambda_{k}}{z+h_{k}^{*}\bar{(M}_{k}-zI)^{-1}h_{k}-t_{kk}}. (21)

But again,

βn(z)\displaystyle\beta_{n}(z) =1ptr(S¯nzI)Σ)1\displaystyle=\frac{1}{p}tr(\bar{S}_{n}-zI)\Sigma)^{-1}
βn(z)\displaystyle\implies\beta_{n}(z) =1ptr(S¯nzI)UΛU)1\displaystyle=\frac{1}{p}tr(\bar{S}_{n}-zI)U^{*}\Lambda U)^{-1}
βn(z)\displaystyle\implies\beta_{n}(z) =1ptr(MzI)1Λ)byEq.(20)\displaystyle=\frac{1}{p}tr(M-zI)^{-1}\Lambda)\qquad\mathrm{by~{}Eq.}\ \eqref{eq:2-1}
λkn(l=1nτl2)βn(z)\displaystyle\implies\lambda_{k}n(\sum_{l=1}^{n}\tau_{l}^{2})\beta_{n}(z) =λkn(l=1nτl2)1ptr(Λ(MzI)1).\displaystyle=\lambda_{k}n(\sum_{l=1}^{n}\tau_{l}^{2})\frac{1}{p}tr(\Lambda(M-zI)^{-1}).

By Remark 4, we know that hk(M¯kzI)1hk=n(l=1nτl2)λktr(Λk(M¯kzI)1)/p+ϵ1kh_{k}^{*}\bar{(M}_{k}-zI)^{-1}h_{k}=n(\sum_{l=1}^{n}\tau_{l}^{2})\lambda_{k}tr(\Lambda_{k}\bar{(M}_{k}-zI)^{-1})/p+\epsilon_{1k}. Define, ϵ2k=λkn(l=1nτl2)E(βn(z))hk(M¯kzI)1hk\epsilon_{2k}=\lambda_{k}n(\sum_{l=1}^{n}\tau_{l}^{2})E(\beta_{n}(z))-h_{k}^{*}\bar{(M}_{k}-zI)^{-1}h_{k}. Therefore from Eq. (21),

βn(z)\displaystyle\beta_{n}(z) =1pk=1pλkz+λkn(l=1nτl2)E(βn(z))+\displaystyle=-\frac{1}{p}\sum_{k=1}^{p}\frac{\lambda_{k}}{z+\lambda_{k}n(\sum_{l=1}^{n}\tau_{l}^{2})E(\beta_{n}(z))}+
{1pk=1pλkz+λkn(l=1nτl2)E(βn(z))tkkϵ2k+1pk=1pλkz+λkn(l=1nτl2)E(βn(z))}.\displaystyle\{-\frac{1}{p}\sum_{k=1}^{p}\frac{\lambda_{k}}{z+\lambda_{k}n(\sum_{l=1}^{n}\tau_{l}^{2})E(\beta_{n}(z))-t_{kk}-\epsilon_{2k}}+\frac{1}{p}\sum_{k=1}^{p}\frac{\lambda_{k}}{z+\lambda_{k}n(\sum_{l=1}^{n}\tau_{l}^{2})E(\beta_{n}(z))}\}.

Let us consider the second part of the right hand side. This equals

1pk=1pλk(tkk+ϵ2k)(z+λkn(l=1nτl2)E(βn(z)))2\displaystyle\frac{1}{p}\sum_{k=1}^{p}\frac{-\lambda_{k}(t_{kk}+\epsilon_{2k})}{(z+\lambda_{k}n(\sum_{l=1}^{n}\tau_{l}^{2})E(\beta_{n}(z)))^{2}}
+1pk=1pλk(tkk+ϵ2k)(z+λkn(l=1nτl2)E(βn(z)))(z+λkn(l=1nτl2)E(βn(z))tkkϵ2k)\displaystyle+\frac{1}{p}\sum_{k=1}^{p}\frac{-\lambda_{k}(t_{kk}+\epsilon_{2k})}{(z+\lambda_{k}n(\sum_{l=1}^{n}\tau_{l}^{2})E(\beta_{n}(z)))(z+\lambda_{k}n(\sum_{l=1}^{n}\tau_{l}^{2})E(\beta_{n}(z))-t_{kk}-\epsilon_{2k})}
1pk=1pλk(tkk+ϵ2k)(z+λkn(l=1nτl2)E(βn(z)))2\displaystyle-\frac{1}{p}\sum_{k=1}^{p}\frac{-\lambda_{k}(t_{kk}+\epsilon_{2k})}{(z+\lambda_{k}n(\sum_{l=1}^{n}\tau_{l}^{2})E(\beta_{n}(z)))^{2}}
=1pk=1pλk(tkk+ϵ2k)(z+λkn(l=1nτl2)E(βn(z)))2\displaystyle=-\frac{1}{p}\sum_{k=1}^{p}\frac{\lambda_{k}(t_{kk}+\epsilon_{2k})}{(z+\lambda_{k}n(\sum_{l=1}^{n}\tau_{l}^{2})E(\beta_{n}(z)))^{2}}
+1pk=1pλk(ϵ2k+tkk)(tkk+ϵ2k)(z+λkn(l=1nτl2)E(βn(z)))2(z+λkn(l=1nτl2)E(βn(z))tkkϵ2k)\displaystyle+\frac{1}{p}\sum_{k=1}^{p}\lambda_{k}(\epsilon_{2k}+t_{kk})\frac{-(t_{kk}+\epsilon_{2k})}{(z+\lambda_{k}n(\sum_{l=1}^{n}\tau_{l}^{2})E(\beta_{n}(z)))^{2}(z+\lambda_{k}n(\sum_{l=1}^{n}\tau_{l}^{2})E(\beta_{n}(z))-t_{kk}-\epsilon_{2k})}
=ϵ3k+ϵ4k.\displaystyle=\epsilon_{3k}+\epsilon_{4k}.

We try to argue that |E(ϵ2k)|0|E(\epsilon_{2k})|\rightarrow 0 as pp\rightarrow\infty,

|E(ϵ2k)|\displaystyle|E(\epsilon_{2k})| =|E(λk(l=1nτl2)nE(βn(z))hkM¯k(z)1hk)|\displaystyle=|E(\lambda_{k}\frac{(\sum_{l=1}^{n}\tau_{l}^{2})}{n}E(\beta_{n}(z))-h_{k}^{*}\bar{M}_{k}(z)^{-1}h_{k})|
=|λkl=1nτl2npE{tr((M(z)1Λ)tr(M¯k(z)1(Λλkekek))}ϵ1k|\displaystyle=|\lambda_{k}\frac{\sum_{l=1}^{n}\tau_{l}^{2}}{np}E\{tr((M(z)^{-1}\Lambda)-tr(\bar{M}_{k}(z)^{-1}(\Lambda-\lambda_{k}e_{k}e_{k}^{*}))\}-\epsilon_{1k}|
=|λkl=1nτl2npE{tr((M(z)1M¯k(z)1)Λ)λkz)}ϵ1k|\displaystyle=|\lambda_{k}\frac{\sum_{l=1}^{n}\tau_{l}^{2}}{np}E\{tr((M(z)^{-1}-\bar{M}_{k}(z)^{-1})\Lambda)-\frac{\lambda_{k}}{z})\}-\epsilon_{1k}|
λkl=1nτl2np|Etr((M(z)1M¯k(z)1)Λ)|+|λk2(l=1nτl2)znp|+λkl=1nτl2np|ϵ1k|.\displaystyle\leq\lambda_{k}\frac{\sum_{l=1}^{n}\tau_{l}^{2}}{np}|Etr((M(z)^{-1}-\bar{M}_{k}(z)^{-1})\Lambda)|+|\frac{\lambda_{k}^{2}(\sum_{l=1}^{n}\tau_{l}^{2})}{znp}|+\lambda_{k}\frac{\sum_{l=1}^{n}\tau_{l}^{2}}{np}|\epsilon_{1k}|.

It can be shown that the above expression is o(p)o(p) (see 39).

E(|ϵ2k+tkk|2)\displaystyle E(|\epsilon_{2k}+t_{kk}|^{2}) =E|ϵ2k+tkkE(ϵ2k)|2+|E(ϵ2k)|2\displaystyle=E|\epsilon_{2k}+t_{kk}-E(\epsilon_{2k})|^{2}+|E(\epsilon_{2k})|^{2}
=E|hkM¯k(z)1hk+tkk+(l=1nτl2)λktrE(ΛkM¯k(z)1)/np|2+|E(ϵ2k)|2\displaystyle=E|-h_{k}^{*}\bar{M}_{k}(z)^{-1}h_{k}+t_{kk}+(\sum_{l=1}^{n}\tau_{l}^{2})\lambda_{k}trE(\Lambda_{k}\bar{M}_{k}(z)^{-1})/np|^{2}+|E(\epsilon_{2k})|^{2}
=E|hkM¯k(z)1hk+tkk+(l=1nτl2)λktrE(ΛkM¯k(z)1)/np|2\displaystyle=E|-h_{k}^{*}\bar{M}_{k}(z)^{-1}h_{k}+t_{kk}+(\sum_{l=1}^{n}\tau_{l}^{2})\lambda_{k}trE(\Lambda_{k}\bar{M}_{k}(z)^{-1})/np|^{2}
+E|(l=1nτl2)λktrE(ΛkM¯k(z)1)/np(l=1nτl2)λktr(ΛkM¯k(z)1)/np|2+|E(ϵ2k)|2.\displaystyle+E|(\sum_{l=1}^{n}\tau_{l}^{2})\lambda_{k}trE(\Lambda_{k}\bar{M}_{k}(z)^{-1})/np-(\sum_{l=1}^{n}\tau_{l}^{2})\lambda_{k}tr(\Lambda_{k}\bar{M}_{k}(z)^{-1})/np|^{2}+|E(\epsilon_{2k})|^{2}.

It can be shown that all three terms is o(p)o(p).
We further observe that

|z+λk(l=1nτl2)nE(βn(z))|Im(z+λk(l=1nτl2)nE(βn(z)))Im(z)=v>0|z+\lambda_{k}\frac{(\sum_{l=1}^{n}\tau_{l}^{2})}{n}E(\beta_{n}(z))|\geq Im(z+\lambda_{k}\frac{(\sum_{l=1}^{n}\tau_{l}^{2})}{n}E(\beta_{n}(z)))\geq Im(z)=v>0

and

|z+λk(l=1nτl2)nE(βn(z))tkkϵ2k||z+hkM¯k1hktkk|v.|z+\lambda_{k}\frac{(\sum_{l=1}^{n}\tau_{l}^{2})}{n}E(\beta_{n}(z))-t_{kk}-\epsilon_{2k}|\geq|z+h_{k}^{*}\bar{M}_{k}^{-1}h_{k}-t_{kk}|\geq v.

This implies ϵ3k0\epsilon_{3k}\rightarrow 0 a.s.
With this we showed that

E(βn(z))=1pk=1pλkz+λk(l=1nτl2)nE(βn(z))+dn,E(\beta_{n}(z))=-\frac{1}{p}\sum_{k=1}^{p}\frac{\lambda_{k}}{z+\lambda_{k}\frac{(\sum_{l=1}^{n}\tau_{l}^{2})}{n}E(\beta_{n}(z))}+d_{n},

where dn0d_{n}\rightarrow 0 as nn\rightarrow\infty.
If we replace (x1i,,xpi)(x_{1i},...,x_{pi}) by (x1i,,xpi)(x^{\prime}_{1i},...,x^{\prime}_{pi}) and call the resulting new s¯n\bar{s}_{n} by sns^{\prime}_{n} (similarly βn\beta_{n} by βn(z)\beta^{\prime}_{n}(z)) then it is easy to show that

|s¯n(z)sn(z)|c1pv|\bar{s}_{n}(z)-s^{\prime}_{n}(z)|\leq\frac{c_{1}}{pv}

and

|β¯n(z)βn(z)|c1pv.|\bar{\beta}_{n}(z)-\beta^{\prime}_{n}(z)|\leq\frac{c_{1}}{pv}.

Hence by Lemma 9,

nP(|s¯n(z)sn(z)|>δ)<\sum_{n}P(|\bar{s}_{n}(z)-s^{\prime}_{n}(z)|>\delta)<\infty

and

nP(|β¯n(z)βn(z)|>δ)<.\sum_{n}P(|\bar{\beta}_{n}(z)-\beta^{\prime}_{n}(z)|>\delta)<\infty.

By Borel-Cantelli lemma, s¯n(z)E(s¯n)0\bar{s}_{n}(z)-E(\bar{s}_{n})\rightarrow 0 a.s. and βn(z)E(βn(z))0\beta_{n}(z)-E(\beta_{n}(z))\rightarrow 0 a.s. and thus we also have s~n(z)E(s¯n)0\tilde{s}_{n}(z)-E(\bar{s}_{n})\rightarrow 0. Now as maxk|λk|a0\text{max}_{k}|\lambda_{k}|\leq a_{0} and |z+λkn(l=1nτl2)E(βn(z))|v|z+\lambda_{k}n(\sum_{l=1}^{n}\tau_{l}^{2})E(\beta_{n}(z))|\geq v is bounded, E(βn(z))E(\beta_{n}(z)) is bounded. So by dominated convergence theorem E(βn(z))E(\beta_{n}(z)) converges to β(z)\beta(z). But as βn(z)E(βn(z))0\beta_{n}(z)-E(\beta_{n}(z))\rightarrow 0 a.s., we have βn(z)β(z)\beta_{n}(z)\rightarrow\beta(z) a.s. Similarly for sn(z)s(z)s_{n}(z)\rightarrow s(z) a.s. So sn(z)s_{n}(z) can be evaluated by the following two equations,

s(z)=dHΣ(λ)z+n(lτl2)λE(β(z))s(z)=-\int\frac{dH^{\Sigma}(\lambda)}{z+n(\sum_{l}\tau_{l}^{2})\lambda E(\beta(z))}

β(z)=λdHΣ(λ)z+n(lτl2)λE(β(z)).\beta(z)=-\int\frac{\lambda dH^{\Sigma}(\lambda)}{z+n(\sum_{l}\tau_{l}^{2})\lambda E(\beta(z))}.

6 Data Analysis

6.1 Simulated Data Analysis

Although we are working with a high-dimensional set up, the computational complexity of the Hayashi’s estimator is worth paying attention to. The fact that the time to compute the Hayashi’s estimator is much greater compared to the Realized covariance estimator, restricts us to a moderate dimension and sample size in the simulation study. We simulated 30 stocks with each 500 observations where the spot volatility matrix is taken to be II. The empirical cdf of Hayashi-Yoshida estimator and the cdf of integrated covariance matrix are shown in Fig. 4. The red line is obtained by generating data from the same underlying process on sufficiently fine synchronous grid and calculating the realized covariance for such data. It serves as the proxy of the spectrum of Integrated covariance matrix. One limitation of the simulation study is that we have taken same number of observations for each stock. This is of course not a practical assumption but as discussed in Sec.2.2.1 it corresponds to the refresh time sampling.

We create the nonsynchronous data using the following algorithm.

Algorithm 1. Initialise pp (the number of stocks), n1,n2,n3,,npn_{1},n_{2},n_{3},...,n_{p} (the number of observations in each stock) and DD (the interval [0,D][0,D] represents a day). 2. Draw a sample of size n=nin=\sum n_{i} from a uniform distribution on [0,D][0,D] where [0,D][0,D] represents a day. Denote it by T={t1,,tn}T=\{t_{1},...,t_{n}\}. Assume t0=0t_{0}=0. 3. Generate nn random vectors (𝐱j,j=1(1)n\mathbf{x}_{j},j=1(1)n) from a pp-dimensional distribution. Denote them by 𝐱j,j=1(1)n\mathbf{x}_{j},j=1(1)n. Scale the 𝐱j\mathbf{x}_{j}’s appropriately to represent the increment in returns for the interval (tjtj1t_{j}-t_{j-1}): 𝐱j=𝐱jtjtj1\mathbf{x}_{j}=\thinspace\mathbf{x}_{j}\sqrt{t_{j}-t_{j-1}}. 4. Define 𝐲k=1k𝐱𝐣k{1,2,,n}\mathbf{y}_{k}=\sum_{1}^{k}\mathbf{x_{j}}\quad\forall k\in\{1,2,...,n\} 5. for i=1(1)pi=1(1)p From TT take a random subset of size nin_{i}. Denote it by TiT_{i}. Data for iith stock is {(xi,j,tj),j=1(1)nix_{i,j},t_{j}),j=1(1)n_{i}} Update T=TTiT=T-T_{i} i.e. removing the time points chose for TiT_{i} from TT.

Refer to caption
Figure 4: Plot for empirical cumulative distribution function of Hayashi-Yoshida estimator and cumulative distribution functions of Integrated covariance matrix where the spot volatility matrix is taken to be I for 30 stocks and 500 observations.

We want to see the effect of γ\gamma, that is the ratio of p/np/n on the empirical spectral distribution. So we repeat the simulation with p=30p=30 and n=60n=60. The result is given in the left panel of Fig. 5

Next, we create a similar plot when the stocks are dependent. We have taken a 30-dimensional positive semi-definite covariance matrix with p=30p=30 stocks. As nontrivial high-dimensional covariance matrix is difficult to prefix, we take the 30×3030\times 30 principal sub-matrix of the estimated covariance matrix from real data (Sec. 6.2). The right panel of Fig. 5 shows the distribution of the eigenvalues. We can see that for general covariance matrix Hayashi-Yoshida estimator may not be positive definite.

Refer to caption
Refer to caption
Figure 5: Plot for empirical cumulative distribution function of Hayashi-Yoshida estimator and cumulative distribution functions of Integrated covariance matrix. Left: The spot volatility matrix is taken to be II for 30 variables and sample size 60. Right: Spot volatility matrix Σ\Sigma is different from II with 30 variables and sample size 500.

It is clear from the algorithm that the time points are generated by a Poisson Process with different intensity parameter. For computational convenience we have taken all nin_{i}’s to be same (=n=n).

6.2 Real Data analysis

The limiting spectral distribution is particularly useful to test for deviation from null model, for example, whether the covolatility process is II or not. Spectrum of integrated covariance matrix also helps to understand some of the key properties of the interacting units of the intraday-financial-network 40. The extreme (highest) eigenvalue, for example, gives us significant insight about the market mode or the collective overall response of the market to some external information. Spectral analysis, therefore, reveals broadly three types of fluctuations: (i) common to all stocks (i.e., due to market), (ii) related to a particular business sector (e.g. sectoral) and (iii) limited to an individual stock (i.e., idiosyncratic). These can be captured by simply segregating the network spectrum into the following parts: (i) the extreme eigenvalue (ii) eigenvalues deviating from the theoretical spectral distribution and (iii) bulk of the spectrum (41, 42, 43). Limiting spectral distribution of Hayashi-Yoshida estimator would help us to identify the sectoral mode of intraday financial network.

We collect intraday tick by tick Bloomberg data of equities in Nifty 50 for several days. Here we present the results for three consecutive days starting from 22-12-2020 which are fairly representative. In Fig.6, we have plotted the scree plots of eigenvalues for these 50 stocks for the three days on the left panel. We see that the impact of the market mode makes the largest eigenvalue away from the bulk. On the right panel some of the eigenvectors, for the corresponding days, are plotted. Specifically, these are the eigenvector 3 of day 1, eigenvector 2 of day 2 and eigenvector 3 of day 3. Each of these has high contributions from stocks 2,4,13,14, 36 with same sign. These stocks are all from IT sector and there are no other stocks from IT sector in our dataset. This suggests that the IT sector mean (same sign) is the next big component that drives the market after the overall mean (market mode).

Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 6: Left panel: Scree plots of three consecutive days are plotted. We can see that the highest eigenvalue representing the market mode is away from the bulk. Right panel: Third eigenvector obtained from the data of first day, second eigenvector of second day and third eigenvector of the third day.

7 Conclusion and further directions

In this work we have determined the limiting spectral distribution of high dimensional Hayashi-Yoshida estimator for nonsynchronous intraday data. Limiting spectral distribution can help to construct a shrinkage estimator of high-dimensional integrated covariance matrix (see 44). It can also be used for testing for a particular structure in spot volatility matrix.

In this paper we have only considered asynchronicity but not the presence of microstructure noise, which is also a feature of intraday stock-price data. So a natural direction to extend this work is by adding microstructure noise to it. Significant insights can be obtained from 28 where the same was derived for realized (co)volatility matrix. In presence of noise the spectral distribution may deviate from the ideal situation in significant ways. We have restricted ourselves to the simple Black-Scholes setup. Geometric Brownian motion models are not always very realistic models to describe financial data. One can try to go beyond that and investigate into the spectral analysis of Hayashi’s estimator for more complex models. One can also try to extend the results for a general class of time varying covolatility processes (for more details, see 27, 28). Changes due to leverage effects can also be quite serious and so worth looking into.

References

  • 1 Epps TW. Comovements in stock prices in the very short run. Journal of the American Statistical Association 1979; 74(366a): 291–298.
  • 2 Baumöhl E, Vỳrost T. Stock market integration: Granger causality testing with respect to nonsynchronous trading effects. Finance a Uver 2010; 60(5): 414.
  • 3 Hayashi T, Yoshida N, others . On covariance estimation of non-synchronously observed diffusion processes. Bernoulli 2005; 11(2): 359–379.
  • 4 Pourahmadi M. High-dimensional covariance estimation: with high-dimensional data. 882. John Wiley & Sons . 2013.
  • 5 Tumminello M, Di Matteo T, Aste T, Mantegna R. Correlation based networks of equity returns sampled at different time horizons. The European Physical Journal B-Condensed Matter and Complex Systems 2007; 55(2): 209–217.
  • 6 Zebedee AA, Kasch-Haroutounian M. A closer look at co-movements among stock returns. Journal of Economics and Business 2009; 61(4): 279–294.
  • 7 Bonanno G, Lillo F, Mantegna RN. High-frequency cross-correlation in a set of stocks. 2001.
  • 8 Muthuswamy J, Sarkar S, Low A, Terry E, others . Time variation in the correlation structure of exchange rates: High-frequency analyses. Journal of Futures markets 2001; 21(2): 127–144.
  • 9 Lundin MC, Dacorogna MM, Müller UA. Correlation of high frequency financial time series. 1998.
  • 10 Reno R. A closer look at the Epps effect. International Journal of theoretical and applied finance 2003; 6(01): 87–102.
  • 11 Precup OV, Iori G. A comparison of high-frequency cross-correlation measures. Physica A: Statistical Mechanics and its Applications 2004; 344(1): 252–256.
  • 12 Lo AW, MacKinlay AC. An econometric analysis of nonsynchronous trading. Journal of Econometrics 1990; 45(1-2): 181–211.
  • 13 Campbell JY, Lo AWC, MacKinlay AC. The econometrics of financial markets. princeton University press . 1997.
  • 14 Bernhardt D, Davies RJ. The impact of nonsynchronous trading on differences in portfolio cross-autocorrelations. 2008.
  • 15 Atchison MD, Butler KC, Simonds RR. Nonsynchronous security trading and market index autocorrelation. The Journal of Finance 1987; 42(1): 111–118.
  • 16 Barndorff-Nielsen OE, Shephard N. Econometric analysis of realized covariation: High frequency based covariance, regression, and correlation in financial economics. Econometrica 2004; 72(3): 885–925.
  • 17 Merton RC. On estimating the expected return on the market: An exploratory investigation. Journal of financial economics 1980; 8(4): 323–361.
  • 18 Barndorff-Nielsen OE, Hansen PR, Lunde A, Shephard N. Multivariate realised kernels: consistent positive semi-definite estimators of the covariation of equity prices with noise and non-synchronous trading. Journal of Econometrics 2011; 162(2): 149–169.
  • 19 Voev V, Lunde A. Integrated covariance estimation using high-frequency data in the presence of noise. Journal of Financial Econometrics 2006; 5(1): 68–104.
  • 20 Hautsch N, Kyj LM, Oomen RC. A blocking and regularization approach to high-dimensional realized covariance estimation. Journal of Applied Econometrics 2012; 27(4): 625–645.
  • 21 Ledoit O, Wolf M. Honey, I shrunk the sample covariance matrix. UPF economics and business working paper 2003(691).
  • 22 Ledoit O, Wolf M. A well-conditioned estimator for large-dimensional covariance matrices. Journal of multivariate analysis 2004; 88(2): 365–411.
  • 23 Ledoit O, Wolf M, others . Analytical nonlinear shrinkage of large-dimensional covariance matrices. Annals of Statistics 2020; 48(5): 3043–3065.
  • 24 Bannouh K, Martens M, Oomen RC, Dijk vDJ. Realized mixed-frequency factor models for vast dimensional covariance estimation. ERIM Report Series Reference No. ERS-2012-017-F&A 2012.
  • 25 Lunde A, Shephard N, Sheppard K. Econometric analysis of vast covariance matrices using composite realized kernels. Manuscript, University of Aarhus 2011.
  • 26 El Karoui N, others . High-dimensionality effects in the Markowitz problem and other quadratic programs with linear constraints: Risk underestimation. The Annals of Statistics 2010; 38(6): 3487–3566.
  • 27 Zheng X, Li Y. On the estimation of integrated covariance matrices of high dimensional diffusion processes. 2011.
  • 28 Xia N, Zheng X, others . On the inference about the spectral distribution of high-dimensional covariance matrix based on high-frequency noisy observations. The Annals of Statistics 2018; 46(2): 500–525.
  • 29 Wang M, Xia N. Estimation of high-dimensional integrated covariance matrix based on noisy high-frequency data with multiple observations. Statistics & Probability Letters 2021; 170: 108996.
  • 30 Robert CY, Rosenbaum M. On the limiting spectral distribution of the covariance matrices of time-lagged processes. Journal of multivariate analysis 2010; 101(10): 2434–2451.
  • 31 Wang C, Jin B, Miao B. On limiting spectral distribution of large sample covariance matrices by VARMA (p, q). Journal of Time Series Analysis 2011; 32(5): 539–546.
  • 32 Aït-Sahalia Y, Fan J, Xiu D. High-frequency covariance estimates with noisy and asynchronous financial data. Journal of the American Statistical Association 2010; 105(492): 1504–1517.
  • 33 Fan J, Li Y, Yu K. Vast volatility matrix estimation using high-frequency data for portfolio selection. Journal of the American Statistical Association 2012; 107(497): 412–428.
  • 34 Guo X, Lai TL, Shek H, Wong SPS. Quantitative trading: algorithms, analytics, data, models, optimization. CRC Press . 2017.
  • 35 Barndorff-Nielsen OE, Hansen PR, Lunde A, Shephard N. Subsampling realised kernels. Journal of Econometrics 2011; 160(1): 204–219.
  • 36 Chakrabarti A, Sen R. Copula estimation for nonsynchronous financial data. arXiv preprint arXiv:1904.10182 2019.
  • 37 Yu Y, Wang T, Samworth RJ. A useful variant of the Davis–Kahan theorem for statisticians. Biometrika 2015; 102(2): 315–323.
  • 38 Bai Z, Silverstein JW. Spectral analysis of large dimensional random matrices. 20. Springer . 2010.
  • 39 Wang L, Paul D. Limiting spectral distribution of renormalized separable sample covariance matrices when p/n\rightarrow 0. Journal of Multivariate Analysis 2014; 126: 25–52.
  • 40 Kumar S, Deo N. Correlation and network analysis of global financial indices. Physical Review E 2012; 86(2): 026101.
  • 41 Plerou V, Gopikrishnan P, Rosenow B, Amaral LAN, Stanley HE. Universal and nonuniversal properties of cross correlations in financial time series. Physical review letters 1999; 83(7): 1471.
  • 42 Sinha S, Chatterjee A, Chakraborti A, Chakrabarti BK. Econophysics: an introduction. John Wiley & Sons . 2010.
  • 43 Onatski A. Determining the number of factors from empirical distribution of eigenvalues. The Review of Economics and Statistics 2010; 92(4): 1004–1016.
  • 44 Ledoit O, Wolf M, others . Nonlinear shrinkage estimation of large-dimensional covariance matrices. The Annals of Statistics 2012; 40(2): 1024–1060.
  • 45 Bai Z. Methodologies in spectral analysis of large dimensional random matrices, a review. Statistica Sinica 1999: 611–662.
  • 46 Geronimo JS, Hill TP. Necessary and sufficient condition that the limit of Stieltjes transforms is a Stieltjes transform. Journal of Approximation Theory 2003; 121(1): 54–60.
  • 47 McDiarmid C. Centering sequences with bounded differences. Combinatorics, Probability and Computing 1997; 6(1): 79–86.

Appendix

Appendix A Lemmas

Lemma A.1.

Let w1,w2w_{1,}w_{2}\in\mathbb{C}, with Re(w1)0Re(w_{1})\geq 0 and Re(w2)0Re(w_{2})\geq 0, AA is a p×pp\times p Hermitian nnd matrix, BB being any p×pp\times p matrix, and q𝕡q\in\mathbb{C^{p}}, then

|qB(w1A+I)1qqB(w2A+I)1q||w1w2||q|2BA.|q^{*}B(w_{1}A+I)^{-1}q-q^{*}B(w_{2}A+I)^{-1}q|\leq|w_{1}-w_{2}|~{}|q|^{2}~{}\|B\|~{}\|A\|.

For proof, see 45.

Lemma A.2.

Let zz\in\mathbb{C}, with v=I(z)>0v=I(z)>0, AA and BB are p×pp\times p matrices, with BB being Hermitian, and qpq\in\mathbb{C}^{p}. Then

|tr(((BzI)1(B+θqqzI)1A|Av|tr(((B-zI)^{-1}-(B+\theta qq^{*}-zI)^{-1}A|\leq\frac{||A||}{v}

for all θ.\theta\in\mathbb{R}.

For proof, see 45.

Lemma A.3.

For any Hermitian matrix AA and zz\in\mathbb{C}, with Im(z)=v>0Im(z)=v>0-

(AzI)11/v.||(A-zI)^{-1}||\leq 1/v.

For proof, see 38.

Lemma A.4.

For X=(X1,X2,,Xp)TX=(X_{1},X_{2},...,X_{p})^{T} where XjX_{j}’s are iid random variables such that 𝔼(X1)=0\mathbb{E}(X_{1})=0 , 𝔼|X1|2=1\mathbb{E}|X_{1}|^{2}=1, and 𝔼|X1|2k<\mathbb{E}|X_{1}|^{2k}<\infty for some 2k2\leq k\in\mathbb{N}, there exists Ck0C_{k}\geq 0, depending only on kk, 𝔼|X1|4\mathbb{E}|X_{1}|^{4}and 𝔼|X1|2k\mathbb{E}|X_{1}|^{2k}, such that for any p×pp\times p nonrandom matrix AA,

𝔼|XAXtr(A)|2kCk(tr(AA))kCkpkA2k.\mathbb{E}|X^{*}AX-tr(A)|^{2k}\leq C_{k}(tr(AA^{*}))^{k}\leq C_{k}p^{k}\|A\|^{2k}. (22)

For proof, see 27.

Lemma A.5.

Suppose SS is a matrix defined as assumption 𝒜4\mathcal{A}_{4}, zz\in\mathbb{C} and 1ptr(SnkzI)1s(z)\frac{1}{p}\text{tr}(S_{n_{k}}-zI)^{-1}\rightarrow s(z), then

l=1nkτl1+τlal01τt1+cτts~(z)𝑑t0\sum_{l=1}^{n_{k}}\frac{\tau_{l}}{1+\tau_{l}a_{l}}\rightarrow\int_{0}^{1}\frac{\tau^{\prime}_{t}}{1+c\tau^{\prime}_{t}\tilde{s}(z)}dt\neq 0

where al=YlΣ12(jlΣ12YjYjΣ12zI)1Σ12Yla_{l}=Y_{l}^{\prime}\Sigma^{\frac{1}{2}}\big{(}\sum_{j\neq l}\Sigma^{\frac{1}{2}}Y_{j}Y_{j}^{\prime}\Sigma^{\frac{1}{2}}-zI\big{)}^{-1}\Sigma^{\frac{1}{2}}Y_{l} (YlY_{l} is defined as in assumption 𝒜3\mathcal{A}_{3}) and s~(z)\tilde{s}(z) is the unique solution in Q1Q_{1} to the following equation:

01τt1+cτts~(z)𝑑t=1c(1+zs(z)).\int_{0}^{1}\frac{\tau^{\prime}_{t}}{1+c\tau^{\prime}_{t}\tilde{s}(z)}dt=1-c(1+zs(z)).

For proof, see 27.

Lemma A.6.

Let z=iv,z=iv\in\mathbb{C},with v>0v>0, AA be any p×pp\times p matrix, and BB be a p×pp\times p Hermitian nonnegative definite matrix. Then tr(A(BzI)1A)Q1tr(A(B-zI)^{-1}A^{*})\in Q_{1}.

For proof, See 27.

Lemma A.7.

Let z=ivz=iv\in\mathbb{C} with v>0v>0, A be a p×pp\times p Hermitian nonnegative definite matrix, qp,a>0q\in\mathbb{C}^{p},a>0. Then

1z.11+a.q(AzI)1qQ1={z:Re(z)0,Im(z)0}.-\frac{1}{z}.\frac{1}{1+a.q^{*}(A-zI)^{-1}q}\in Q_{1}=\{z\in\mathbb{C}:\thinspace Re(z)\geq 0,\thinspace Im(z)\geq 0\}. (23)

For proof, see 27.

Lemma A.8.

Suppose that PnP_{n} are real probability measures with Stieltjes transforms sn(z)s_{n}(z) . Let K+K\subset\mathbb{C}_{+} be an infinite set with limit points in +\mathbb{C}_{+}. If limsn(z)=s(z)lims_{n}(z)=s(z) exists for all zKz\in K, then there exists a Probability measure PP with Stieltjes transform m(z)m(z) if and only if

limviv.s(iv)=1lim_{v\rightarrow\infty}iv.s(iv)=-1

in which case PnPP_{n}\rightarrow P.

For proof, see 46.

The next Lemma is known as McDiarmid Inequality 47.

Lemma A.9.

Let Y1,Y2,,YmY_{1},Y_{2},...,Y_{m} be independent random vectors taking values in 𝒳\mathcal{X}. Suppose that f:𝒳kf:\mathcal{X}^{k}\rightarrow\mathbb{R} is a function of Y1,Y2,,YmY_{1},Y_{2},...,Y_{m} satisfying y1,y2,..,ym,yi\forall y_{1},y_{2},..,y_{m},y_{i}^{\prime},

|f(y1,y2,..,yi,..,ym)f(y1,y2,..,yi,..,ym)|ci,|f(y_{1},y_{2},..,y_{i},..,y_{m})-f(y_{1},y_{2},..,y_{i}^{\prime},..,y_{m})|\leq c_{i},

Then for all ϵ>0,\epsilon>0,

P(|f(y1,y2,,ym)f(y1,y2,,ym)|>ϵ)2exp(2ϵ2i=1mci2).P(|f(y_{1},y_{2},...,y_{m})-f(y_{1},y_{2},...,y_{m})|>\epsilon)\leq 2exp(-\frac{2\epsilon^{2}}{\sum_{i=1}^{m}c_{i}^{2}}).
Lemma A.10.
Lemma A.10 and A.12 can be found in Theorem 1 in 27 in slightly different form.

For YlY_{l}, Σ\Sigma, τl\tau_{l} described as Assumption 𝒜\mathcal{A} and z=ivz=iv, define,

M1=YlΣ12(SlzI)1(Σl=1nτl1+τlalzI)1Σ12YlM_{1}=Y_{l}^{\prime}\Sigma^{\frac{1}{2}}(S_{l}-zI)^{-1}(\Sigma\sum_{l=1}^{n}\frac{\tau_{l}}{1+\tau_{l}a_{l}}-zI)^{-1}\Sigma^{\frac{1}{2}}Y_{l}

and

M2=YlΣ12(SlzI)1(Σjlτj1+τjbjlzI)1Σ12YlM_{2}=Y_{l}^{\prime}\Sigma^{\frac{1}{2}}(S_{l}-zI)^{-1}(\Sigma\sum_{j\neq l}\frac{\tau_{j}}{1+\tau_{j}b_{j}^{l}}-zI)^{-1}\Sigma^{\frac{1}{2}}Y_{l}

Then,

1pmaxl(M1M2)0a.s.\frac{1}{p}\mathrm{max_{l}}(M_{1}-M_{2})\rightarrow 0\ a.s.
Proof A.11 (Proof of Lemma A.10).
1pmaxl|YlΣ12(SlzI)1(Σl=1nτl1+τlalzI)1Σ12ZlYlΣ12(SlzI)1(Σjlτj1+τjbjlzI)1Σ12Yl|\displaystyle\frac{1}{p}\mathrm{max_{l}}|Y_{l}^{\prime}\Sigma^{\frac{1}{2}}(S_{l}-zI)^{-1}(\Sigma\sum_{l=1}^{n}\frac{\tau_{l}}{1+\tau_{l}a_{l}}-zI)^{-1}\Sigma^{\frac{1}{2}}Z_{l}-Y_{l}^{\prime}\Sigma^{\frac{1}{2}}(S_{l}-zI)^{-1}(\Sigma\sum_{j\neq l}\frac{\tau_{j}}{1+\tau_{j}b_{j}^{l}}-zI)^{-1}\Sigma^{\frac{1}{2}}Y_{l}|
=1pmaxl|YlΣ12(SlzI)1(1z)(1zΣl=1nτl1+τlal+I)1Σ12Yl\displaystyle=\frac{1}{p}\mathrm{max_{l}}|Y_{l}^{\prime}\Sigma^{\frac{1}{2}}(S_{l}-zI)^{-1}(-\frac{1}{z})(-\frac{1}{z}\Sigma\sum_{l=1}^{n}\frac{\tau_{l}}{1+\tau_{l}a_{l}}+I)^{-1}\Sigma^{\frac{1}{2}}Y_{l}
YlΣ12(SlzI)1(1z)(1zΣjlτj1+τjbjl+I)1Σ12Yl|\displaystyle-Y_{l}^{\prime}\Sigma^{\frac{1}{2}}(S_{l}-zI)^{-1}(-\frac{1}{z})(-\frac{1}{z}\Sigma\sum_{j\neq l}\frac{\tau_{j}}{1+\tau_{j}b_{j}^{l}}+I)^{-1}\Sigma^{\frac{1}{2}}Y_{l}|
1pmaxl|1zlτl1+τlal+1zjlτj1+τjbjl||Σ12Yl|2Cpδ\displaystyle\leq\frac{1}{p}\mathrm{max}_{l}|-\frac{1}{z}\sum_{l}\frac{\tau_{l}}{1+\tau_{l}a_{l}}+\frac{1}{z}\sum_{j\neq l}\frac{\tau_{j}}{1+\tau_{j}b_{j}^{l}}||\Sigma^{\frac{1}{2}}Y_{l}|^{2}Cp^{\delta}

The last inequality is due to Lemma 1 (where q=Σ12Ylq=\Sigma^{\frac{1}{2}}Y_{l}, B=(SlzI)1,B=(S_{l}-zI)^{-1},A=ΣA=\Sigma, w1=1zl=1nτl1+τlalw_{1}=-\frac{1}{z}\sum_{l=1}^{n}\frac{\tau_{l}}{1+\tau_{l}a_{l}}, w2=1zjlτj1+τjbjlw_{2}=-\frac{1}{z}\sum_{j\neq l}\frac{\tau_{j}}{1+\tau_{j}b_{j}^{l}}), and Assumption 6.
Applying Markov Inequality and Lemma 4,

P(|YlΣYltr(Σ)|\displaystyle P(|Y_{l}^{\prime}\Sigma Y_{l}-tr(\Sigma)| pϵ)E|YlΣYltr(Σ)|2k(pϵ)2k\displaystyle\geq p\epsilon)\leq\frac{E|Y_{l}^{\prime}\Sigma Y_{l}-\mathrm{tr}(\Sigma)|^{2k}}{(p\epsilon)^{2k}}
CkpkΣ2k(pϵ)2k\displaystyle\leq\frac{C_{k}p^{k}||\Sigma||^{2k}}{(p\epsilon)^{2k}}
CCkpkp2δk(pϵ)2k\displaystyle\leq\frac{CC_{k}p^{k}p^{2\delta k}}{(p\epsilon)^{2k}}

Choosing k>212δk>\frac{2}{1-2\delta}, and using Borel Cantelli lemma, we get maxl|YlΣYltr(Σ)|/p0\mathrm{max}_{l}|Y_{l}^{\prime}\Sigma Y_{l}-tr(\Sigma)|/p\rightarrow 0 a.s. As a consequence maxl|Σ12Yl|2/p<M\mathrm{max}_{l}|\Sigma^{\frac{1}{2}}Y_{l}|^{2}/p<Mn>n0\forall n>n_{0} for some n0n_{0}.

Define, c=tr(Σ12(SzI)1Σ12)c=tr(\Sigma^{\frac{1}{2}}(S-zI)^{-1}\Sigma^{\frac{1}{2}}), and consider

maxlpδ|lτl1+τlaljlτj1+τjbjl|\mathrm{max}_{l}p^{\delta}|\sum_{l}\frac{\tau_{l}}{1+\tau_{l}a_{l}}-\sum_{j\neq l}\frac{\tau_{j}}{1+\tau_{j}b_{j}^{l}}|
=maxlpδ|lτl1+τlallτl1+τlc+lτl1+τlcjlτj1+τjc+jlτj1+τjcjlτj1+τjbjl|\displaystyle=\mathrm{max}_{l}p^{\delta}|\sum_{l}\frac{\tau_{l}}{1+\tau_{l}a_{l}}-\sum_{l}\frac{\tau_{l}}{1+\tau_{l}c}+\sum_{l}\frac{\tau_{l}}{1+\tau_{l}c}-\sum_{j\neq l}\frac{\tau_{j}}{1+\tau_{j}c}+\sum_{j\neq l}\frac{\tau_{j}}{1+\tau_{j}c}-\sum_{j\neq l}\frac{\tau_{j}}{1+\tau_{j}b_{j}^{l}}|
maxlpδ|lτl1+τlallτl1+τlc|+maxlpδ|lτl1+τlcjlτj1+τjc|+maxl|jlτj1+τjcjlτj1+τjbjl|\displaystyle\leq\mathrm{max}_{l}p^{\delta}|\sum_{l}\frac{\tau_{l}}{1+\tau_{l}a_{l}}-\sum_{l}\frac{\tau_{l}}{1+\tau_{l}c}|+\mathrm{max}_{l}p^{\delta}|\sum_{l}\frac{\tau_{l}}{1+\tau_{l}c}-\sum_{j\neq l}\frac{\tau_{j}}{1+\tau_{j}c}|+\mathrm{max}_{l}|\sum_{j\neq l}\frac{\tau_{j}}{1+\tau_{j}c}-\sum_{j\neq l}\frac{\tau_{j}}{1+\tau_{j}b_{j}^{l}}|
=maxlpδ|lτl1+τlallτl1+τlc|+maxlpδ|τl1+τlc|+maxl|jlτj1+τjcjlτj1+τjbjl|\displaystyle=\mathrm{max}_{l}p^{\delta}|\sum_{l}\frac{\tau_{l}}{1+\tau_{l}a_{l}}-\sum_{l}\frac{\tau_{l}}{1+\tau_{l}c}|+\mathrm{max}_{l}p^{\delta}|\frac{\tau_{l}}{1+\tau_{l}c}|+\mathrm{max}_{l}|\sum_{j\neq l}\frac{\tau_{j}}{1+\tau_{j}c}-\sum_{j\neq l}\frac{\tau_{j}}{1+\tau_{j}b_{j}^{l}}|

maxlpδ|τl1+τlc|maxlpδ|nτln(1+τlc)|κpδn0\mathrm{max}_{l}p^{\delta}|\frac{\tau_{l}}{1+\tau_{l}c}|\leq\mathrm{max}_{l}p^{\delta}|\frac{n\tau_{l}}{n(1+\tau_{l}c)}|\leq\frac{\kappa p^{\delta}}{n}\rightarrow 0

Now we will consider the third part of the above equation,

maxlpδ|jlτj1+τjbjljlτj1+τjc|\displaystyle\mathrm{max}_{l}p^{\delta}|\sum_{j\neq l}\frac{\tau_{j}}{1+\tau_{j}b_{j}^{l}}-\sum_{j\neq l}\frac{\tau_{j}}{1+\tau_{j}c}|
=maxlpδ|jlτj2(cbjl)(1+τjbjl)(1+τjc)|\displaystyle=\mathrm{max}_{l}p^{\delta}|\sum_{j\neq l}\frac{\tau_{j}^{2}(c-b_{j}^{l})}{(1+\tau_{j}b_{j}^{l})(1+\tau_{j}c)}|
=maxlpδ|jlτj2p(cbjl)/p(1+τjbjl)(1+τjc)|\displaystyle=\mathrm{max}_{l}p^{\delta}|\sum_{j\neq l}\frac{\tau_{j}^{2}p(c-b_{j}^{l})/p}{(1+\tau_{j}b_{j}^{l})(1+\tau_{j}c)}|

maxl|jlτj2p(1+τjbjl)(1+τjc)|κ2pn2\mathrm{max}_{l}|\sum_{j\neq l}\frac{\tau_{j}^{2}p}{(1+\tau_{j}b_{j}^{l})(1+\tau_{j}c)}|\leq\frac{\kappa^{2}p}{n^{2}} by Assumption 1 and 6,

maxlmaxjlpϵ|bjl/pc/p|\displaystyle\mathrm{max}_{l}\mathrm{max}_{j\neq l}p^{\epsilon}|b_{j}^{l}/p-c/p|
=maxlmaxjlpϵ|YjΣ12(Sj,lzI)1Σ12Yj/ptr(Σ12(SzI)1Σ12/p)|\displaystyle=\mathrm{max}_{l}\mathrm{max}_{j\neq l}p^{\epsilon}|Y_{j}^{\prime}\Sigma^{\frac{1}{2}}(S_{j,l}-zI)^{-1}\Sigma^{\frac{1}{2}}Y_{j}/p-tr(\Sigma^{\frac{1}{2}}(S-zI)^{-1}\Sigma^{\frac{1}{2}}/p)|
=maxlmaxjlpϵ|YjΣ12(Sj,lzI)1Σ12Yj/ptr(Σ12(Sj,lzI)1Σ12/p)|\displaystyle=\mathrm{max}_{l}\mathrm{max}_{j\neq l}p^{\epsilon}|Y_{j}^{\prime}\Sigma^{\frac{1}{2}}(S_{j,l}-zI)^{-1}\Sigma^{\frac{1}{2}}Y_{j}/p-tr(\Sigma^{\frac{1}{2}}(S_{j,l}-zI)^{-1}\Sigma^{\frac{1}{2}}/p)|
+maxlmaxjlpϵ|tr(Σ12(Sj,lzI)1Σ12/p)tr(Σ12(SzI)1Σ12/p)|\displaystyle+\mathrm{max}_{l}\mathrm{max}_{j\neq l}p^{\epsilon}|tr(\Sigma^{\frac{1}{2}}(S_{j,l}-zI)^{-1}\Sigma^{\frac{1}{2}}/p)-tr(\Sigma^{\frac{1}{2}}(S-zI)^{-1}\Sigma^{\frac{1}{2}}/p)|

Use of Lemma 4, Lemma 3 with Borel Cantelli Lemma will give us for ϵ<12δ\epsilon<\frac{1}{2}-\delta and k>312δ2ϵk>\frac{3}{1-2\delta-2\epsilon} ,

maxlmaxjlpϵ|YjΣ12(Sj,lzI)1Σ12Yj/ptr(Σ12(Sj,lzI)1Σ12/p)|0a.s.\mathrm{max}_{l}\mathrm{max}_{j\neq l}p^{\epsilon}|Y_{j}^{\prime}\Sigma^{\frac{1}{2}}(S_{j,l}-zI)^{-1}\Sigma^{\frac{1}{2}}Y_{j}/p-tr(\Sigma^{\frac{1}{2}}(S_{j,l}-zI)^{-1}\Sigma^{\frac{1}{2}}/p)|\rightarrow 0\ \mathrm{a.s.}

Also,

maxlmaxjlpϵ|1ptr(Σ12(Sj,lzI)1Σ12)1ptr(Σ12(SzI)1Σ12)|\displaystyle\mathrm{max}_{l}\mathrm{max}_{j\neq l}p^{\epsilon}|\frac{1}{p}tr(\Sigma^{\frac{1}{2}}(S_{j,l}-zI)^{-1}\Sigma^{\frac{1}{2}})-\frac{1}{p}tr(\Sigma^{\frac{1}{2}}(S-zI)^{-1}\Sigma^{\frac{1}{2}})|
=maxlmaxjlpϵ|1ptr[Σ12{(Sj,lzI)1(SzI)1}Σ12]|\displaystyle=\mathrm{max}_{l}\mathrm{max}_{j\neq l}p^{\epsilon}|\frac{1}{p}tr[\Sigma^{\frac{1}{2}}\{(S_{j,l}-zI)^{-1}-(S-zI)^{-1}\}\Sigma^{\frac{1}{2}}]|
=maxlmaxjlpϵ|1ptr[{(Sj,lzI)1(SzI)1}Σ]|\displaystyle=\mathrm{max}_{l}\mathrm{max}_{j\neq l}p^{\epsilon}|\frac{1}{p}tr[\{(S_{j,l}-zI)^{-1}-(S-zI)^{-1}\}\Sigma]|
maxlmaxjlpϵ1pΣv\displaystyle\leq\mathrm{max}_{l}\mathrm{max}_{j\leq l}p^{\epsilon}\frac{1}{p}\frac{||\Sigma||}{v}
1pCpδ+ϵv0a.s.\displaystyle\leq\frac{1}{p}\frac{Cp^{\delta+\epsilon}}{v}\rightarrow 0\ \mathrm{a.s.}

The first and second inequalities are result of application of Lemma 2 and Assumption 6 respectively.
This proves out claim.

Lemma A.12.

For YlY_{l}, Σ\Sigma, τl\tau_{l} described as Assumption 𝒜\mathcal{A} and z=ivz=iv, define,

M3=YlΣ12(SlzI)1(Σjlτj1+τjbjlzI)1Σ12YlM_{3}=Y_{l}^{\prime}\Sigma^{\frac{1}{2}}(S_{l}-zI)^{-1}(\Sigma\sum_{j\neq l}\frac{\tau_{j}}{1+\tau_{j}b_{j}^{l}}-zI)^{-1}\Sigma^{\frac{1}{2}}Y_{l}

and

M4=tr[Σ12(SlzI)1(Σjlτj1+τjbjlzI)1Σ12],M_{4}=tr[\Sigma^{\frac{1}{2}}(S_{l}-zI)^{-1}(\Sigma\sum_{j\neq l}\frac{\tau_{j}}{1+\tau_{j}b_{j}^{l}}-zI)^{-1}\Sigma^{\frac{1}{2}}],

then

1pmaxl|M3M4|0a.s.\frac{1}{p}max_{l}|M_{3}-M_{4}|\rightarrow 0\ a.s.
Proof A.13 (Proof of Lemma A.12).

Using Lemma 4 and and Markov Inequality it is easy to show that,

E(YlΣ12(SlzI)1(Σjlτj1+τjbjlzI)1Σ12Yltr[Σ12(SlzI)1(Σjlτj1+τjbjlzI)1Σ12])\displaystyle E(Y_{l}^{\prime}\Sigma^{\frac{1}{2}}(S_{l}-zI)^{-1}(\Sigma\sum_{j\neq l}\frac{\tau_{j}}{1+\tau_{j}b_{j}^{l}}-zI)^{-1}\Sigma^{\frac{1}{2}}Y_{l}-tr[\Sigma^{\frac{1}{2}}(S_{l}-zI)^{-1}(\Sigma\sum_{j\neq l}\frac{\tau_{j}}{1+\tau_{j}b_{j}^{l}}-zI)^{-1}\Sigma^{\frac{1}{2}}])
Ckpkp2δkv2k\displaystyle\leq C_{k}p^{k}\frac{p^{2\delta k}}{v^{2k}}

After choosing appropriate value of kk and using Borel Cantelli Lemma we can get the claim.

Appendix B Proof of the Remarks:

Proof B.1 (Proof of Remark 1).
M\displaystyle M =np(WW(lτl)Λ)\displaystyle=\sqrt{\frac{n}{p}}(WW^{*}-(\sum_{l}\tau_{l})\Lambda)
=np(k=1pekwkW(lτl)Λ)\displaystyle=\sqrt{\frac{n}{p}}(\sum_{k=1}^{p}e_{k}w_{k}^{*}W^{*}-(\sum_{l}\tau_{l})\Lambda)
=npk=1pekwk(Wl+wkek)np(lτl)Λ\displaystyle=\sqrt{\frac{n}{p}}\sum_{k=1}^{p}e_{k}w_{k}^{*}(W_{l}^{*}+w_{k}e_{k}^{*})-\sqrt{\frac{n}{p}}(\sum_{l}\tau_{l})\Lambda
=npk=1pekwkWk+npk=1pekwkwkeknp(lτl)Λ\displaystyle=\sqrt{\frac{n}{p}}\sum_{k=1}^{p}e_{k}w_{k}^{*}W_{k}^{*}+\sqrt{\frac{n}{p}}\sum_{k=1}^{p}e_{k}w_{k}^{*}w_{k}e_{k}^{*}-\sqrt{\frac{n}{p}}(\sum_{l}\tau_{l})\Lambda
=npk=1pek(Wkwk)+npk=1pekwkwkeknp(lτl)k=1pλkekek\displaystyle=\sqrt{\frac{n}{p}}\sum_{k=1}^{p}e_{k}(W_{k}w_{k})^{*}+\sqrt{\frac{n}{p}}\sum_{k=1}^{p}e_{k}w_{k}^{*}w_{k}e_{k}^{*}-\sqrt{\frac{n}{p}}(\sum_{l}\tau_{l})\sum_{k=1}^{p}\lambda_{k}e_{k}e_{k}^{*}
=k=1pekhk+k=1peknp(wkwk(lτl)λk)ek\displaystyle=\sum_{k=1}^{p}e_{k}h_{k}^{*}+\sum_{k=1}^{p}e_{k}\sqrt{\frac{n}{p}}(w_{k}^{*}w_{k}-(\sum_{l}\tau_{l})\lambda_{k})e_{k}^{*}
=k=1pekhk+k=1peknp(wkwk(lτl)λk)ek\displaystyle=\sum_{k=1}^{p}e_{k}h_{k}^{*}+\sum_{k=1}^{p}e_{k}\sqrt{\frac{n}{p}}(w_{k}^{*}w_{k}-(\sum_{l}\tau_{l})\lambda_{k})e_{k}^{*}
=k=1pekhk+k=1pektkkek\displaystyle=\sum_{k=1}^{p}e_{k}h_{k}^{*}+\sum_{k=1}^{p}e_{k}t_{kk}e_{k}^{*}
=k=1pek(hk+tkkek),\displaystyle=\sum_{k=1}^{p}e_{k}(h_{k}+t_{kk}e_{k})^{*},

also it is easy to see that,

M\displaystyle M =np(WW(lτl)Λ)\displaystyle=\sqrt{\frac{n}{p}}(WW^{*}-(\sum_{l}\tau_{l})\Lambda)
=np(WkW+ekwkW(lτl)Λk(lτl)ekekλk)\displaystyle=\sqrt{\frac{n}{p}}(W_{k}W^{*}+e_{k}w_{k}^{*}W^{*}-(\sum_{l}\tau_{l})\Lambda_{k}-(\sum_{l}\tau_{l})e_{k}e_{k}^{*}\lambda_{k})
=np(WkW(lτl)Λk)+np[ekwk(Wk+wkek)(lτl)ekekλk]\displaystyle=\sqrt{\frac{n}{p}}(W_{k}W^{*}-(\sum_{l}\tau_{l})\Lambda_{k})+\sqrt{\frac{n}{p}}[e_{k}w_{k}^{*}(W_{k}^{*}+w_{k}e_{k}^{*})-(\sum_{l}\tau_{l})e_{k}e_{k}^{*}\lambda_{k}]
=Mk+ekhk+npekwkwkeknp(lτl)ekekλk\displaystyle=M_{k}+e_{k}h_{k}^{*}+\sqrt{\frac{n}{p}}e_{k}w_{k}^{*}w_{k}e_{k}^{*}-\sqrt{\frac{n}{p}}(\sum_{l}\tau_{l})e_{k}e_{k}^{*}\lambda_{k}
=Mk+ekhk+eknp(wkwk(lτl)λk)ek\displaystyle=M_{k}+e_{k}h_{k}^{*}+e_{k}\sqrt{\frac{n}{p}}(w_{k}^{*}w_{k}-(\sum_{l}\tau_{l})\lambda_{k})e_{k}^{*}
=Mk+ekhk+tkkekek\displaystyle=M_{k}+e_{k}h_{k}^{*}+t_{kk}e_{k}e_{k}^{*}
=Mk+ek(hk+tkkek).\displaystyle=M_{k}+e_{k}(h_{k}+t_{kk}e_{k})^{*}.
Proof B.2 (Proof of Remark2:).
tr[ek(hk+tkkek)(MkzI)1(hk+tkkek)(MkzI)1ek]\displaystyle tr[e_{k}(h_{k}+t_{kk}e_{k})^{*}(M_{k}-zI)^{-1}(h_{k}+t_{kk}e_{k})^{*}(M_{k}-zI)^{-1}e_{k}]
=(hk+tkkek)(MkzI)1ektr[ek(hk+tkkek)(MkzI)1]\displaystyle=(h_{k}+t_{kk}e_{k})^{*}(M_{k}-zI)^{-1}e_{k}tr[e_{k}(h_{k}+t_{kk}e_{k})^{*}(M_{k}-zI)^{-1}]
={(hk+tkkek)(MkzI)1ek}{(hk+tkkek)(MkzI)1ek}\displaystyle=\{(h_{k}+t_{kk}e_{k})^{*}(M_{k}-zI)^{-1}e_{k}\}\{(h_{k}+t_{kk}e_{k})^{*}(M_{k}-zI)^{-1}e_{k}\}
={(hk+tkkek)(MkzI)1ek}2\displaystyle=\{(h_{k}+t_{kk}e_{k})^{*}(M_{k}-zI)^{-1}e_{k}\}^{2}

and

tr[ek(hk+tkkek)(MkzI)1ek(hk+tkkek)(MkzI)1]\displaystyle tr[e_{k}(h_{k}+t_{kk}e_{k})^{*}(M_{k}-zI)^{-1}e_{k}(h_{k}+t_{kk}e_{k})^{*}(M_{k}-zI)^{-1}]
=tr[(hk+tkkek)(MkzI)1ek(hk+tkkek)(MkzI)1ek]\displaystyle=tr[(h_{k}+t_{kk}e_{k})^{*}(M_{k}-zI)^{-1}e_{k}(h_{k}+t_{kk}e_{k})^{*}(M_{k}-zI)^{-1}e_{k}]
={(hk+tkkek)(MkzI)1ek}2.\displaystyle=\{(h_{k}+t_{kk}e_{k})^{*}(M_{k}-zI)^{-1}e_{k}\}^{2}.

So,

LHS\displaystyle LHS =tr(M(MzI)1)\displaystyle=tr(M(M-zI)^{-1})
=tr{ek(hk+tkkek)(MkzI+ek(hk+tkkek))1}\displaystyle=tr\{e_{k}(h_{k}+t_{kk}e_{k})^{*}(M_{k}-zI+e_{k}(h_{k}+t_{kk}e_{k})^{*})^{-1}\}
=tr[ek(hk+tkkek)(Mk(z)1(MkzI)1ek(hk+tkkek)(MkzI)11+(hk+tkkek)(MkzI)1ek)]\displaystyle=tr[e_{k}(h_{k}+t_{kk}e_{k})^{*}(M_{k}(z)^{-1}-\frac{(M_{k}-zI)^{-1}e_{k}(h_{k}+t_{kk}e_{k})^{*}(M_{k}-zI)^{-1}}{1+(h_{k}+t_{kk}e_{k})^{*}(M_{k}-zI)^{-1}e_{k}})]
=tr[ek(hk+tkkek)((MkzI)1(MkzI)1ek(hk+tkkek)(MkzI)11+(hk+tkkek)(MkzI)1ek)]\displaystyle=tr[e_{k}(h_{k}+t_{kk}e_{k})^{*}((M_{k}-zI)^{-1}-\frac{(M_{k}-zI)^{-1}e_{k}(h_{k}+t_{kk}e_{k})^{*}(M_{k}-zI)^{-1}}{1+(h_{k}+t_{kk}e_{k})^{*}(M_{k}-zI)^{-1}e_{k}})]
=tr[ek(hk+tkkek)((MkzI)1(1+(hk+tkkek)(MkzI)1ek)(MkzI)1ek(hk+tkkek)(MkzI)11+(hk+tkkek)Mk(z)1ek)\displaystyle=tr[e_{k}(h_{k}+t_{kk}e_{k})^{*}(\frac{(M_{k}-zI)^{-1}(1+(h_{k}+t_{kk}e_{k})^{*}(M_{k}-zI)^{-1}e_{k})-(M_{k}-zI)^{-1}e_{k}(h_{k}+t_{kk}e_{k})^{*}(M_{k}-zI)^{-1}}{1+(h_{k}+t_{kk}e_{k})^{*}M_{k}(z)^{-1}e_{k}})
=tr[ek(hk+tkkek)(MkzI)11+(hk+tkkek)(MkzI)1ek]\displaystyle=tr[\frac{e_{k}(h_{k}+t_{kk}e_{k})^{*}(M_{k}-zI)^{-1}}{1+(h_{k}+t_{kk}e_{k})^{*}(M_{k}-zI)^{-1}e_{k}}]
=(hk+tkkek)(MkzI)1ek1+(hk+tkkek)(MkzI)1ek\displaystyle=\frac{(h_{k}+t_{kk}e_{k})^{*}(M_{k}-zI)^{-1}e_{k}}{1+(h_{k}+t_{kk}e_{k})^{*}(M_{k}-zI)^{-1}e_{k}}
Proof B.3 (Proof of Remark 3).

To see this, manipulate the left hand side the following way-

(hk+tkkek)(MkzI)1ek\displaystyle(h_{k}+t_{kk}e_{k})^{*}(M_{k}-zI)^{-1}e_{k} =(hk+tkkek)(np(WkW(lτl)Λk)zI)1ek\displaystyle=(h_{k}+t_{kk}e_{k})^{*}(\sqrt{\frac{n}{p}}(W_{k}W^{*}-(\sum_{l}\tau_{l})\Lambda_{k})-zI)^{-1}e_{k}
=(hk+tkkek)(np(WkWk+Wkwkek(lτl)Λk)zI)1ek\displaystyle=(h_{k}+t_{kk}e_{k})^{*}(\sqrt{\frac{n}{p}}(W_{k}W_{k}^{*}+W_{k}w_{k}e_{k}^{*}-(\sum_{l}\tau_{l})\Lambda_{k})-zI)^{-1}e_{k}
=(hk+tkkek)(np(WkWk(lτl)Λk)zI+npWkwkek)1ek\displaystyle=(h_{k}+t_{kk}e_{k})^{*}(\sqrt{\frac{n}{p}}(W_{k}W_{k}^{*}-(\sum_{l}\tau_{l})\Lambda_{k})-zI+\sqrt{\frac{n}{p}}W_{k}w_{k}e_{k}^{*})^{-1}e_{k}
=(hk+tkkek)(np(WkWk(lτl)Λk)zI+hkek)1ek\displaystyle=(h_{k}+t_{kk}e_{k})^{*}(\sqrt{\frac{n}{p}}(W_{k}W_{k}^{*}-(\sum_{l}\tau_{l})\Lambda_{k})-zI+h_{k}e_{k}^{*})^{-1}e_{k}
=(hk+tkkek)((M¯kzI)+hkek)1ek\displaystyle=(h_{k}+t_{kk}e_{k})^{*}((\bar{M}_{k}-zI)+h_{k}e_{k}^{*})^{-1}e_{k}
=(hk+tkkek)((M¯kzI)1(M¯kzI)1hkek(M¯kzI)11+ek(M¯kzI)1hk)ek\displaystyle=(h_{k}+t_{kk}e_{k})^{*}(\bar{(M}_{k}-zI)^{-1}-\frac{\bar{(M}_{k}-zI)^{-1}h_{k}e_{k}^{*}\bar{(M}_{k}-zI)^{-1}}{1+e_{k}^{*}\bar{(M}_{k}-zI)^{-1}h_{k}})e_{k}
=(hk+tkkek)M¯k(z)1ek(hk+tkkek)(M¯kzI)1hkek(M¯kzI)1ek1+ek(M¯kzI)1hk.\displaystyle=(h_{k}+t_{kk}e_{k})^{*}\bar{M}_{k}(z)^{-1}e_{k}-\frac{(h_{k}+t_{kk}e_{k})^{*}\bar{(M}_{k}-zI)^{-1}h_{k}e_{k}^{*}\bar{(M}_{k}-zI)^{-1}e_{k}}{1+e_{k}^{*}\bar{(M}_{k}-zI)^{-1}h_{k}}.

Observe that hkek=ehk=0h_{k}^{*}e_{k}=e^{*}h_{k}=0 because Wkek=0W_{k}^{*}e_{k}=0. So,

(M¯kzI)ek\displaystyle\bar{(M}_{k}-zI)e_{k} =np(WkWk(lτl)Λk)ekzek\displaystyle=\sqrt{\frac{n}{p}}(W_{k}W_{k}^{*}-(\sum_{l}\tau_{l})\Lambda_{k})e_{k}-ze_{k}
=np(WkWkek(lτl)Λkek)zek\displaystyle=\sqrt{\frac{n}{p}}(W_{k}W_{k}^{*}e_{k}-(\sum_{l}\tau_{l})\Lambda_{k}e_{k})-ze_{k}
=zek.\displaystyle=-ze_{k}.

As a consequence ek(M¯kzI)1hk=ekhk/z=0e_{k}^{*}(\bar{M}_{k}-zI)^{-1}h_{k}=-e_{k}^{*}h_{k}/z=0. And,

(hk+tkkek)(MkzI)1ek\displaystyle(h_{k}+t_{kk}e_{k})^{*}(M_{k}-zI)^{-1}e_{k} =(hk+tkkek)ek/z+(hk+tkkek)(M¯kzI)1hkekek/z\displaystyle=-(h_{k}+t_{kk}e_{k})^{*}e_{k}/z+(h_{k}+t_{kk}e_{k})^{*}\bar{(M}_{k}-zI)^{-1}h_{k}e_{k}^{*}e_{k}/z
=tkk/z+(hk+tkkek)(M¯kzI)1hk/z\displaystyle=-t_{kk}/z+(h_{k}+t_{kk}e_{k})^{*}\bar{(M}_{k}-zI)^{-1}h_{k}/z
=tkk+hk(M¯kzI)1hk+tkkek(M¯kzI)1hkz\displaystyle=\frac{-t_{kk}+h_{k}^{*}\bar{(M}_{k}-zI)^{-1}h_{k}+t_{kk}e_{k}{}^{*}\bar{(M}_{k}-zI)^{-1}h_{k}}{z}
=tkk+hk(M¯kzI)1hkz.\displaystyle=\frac{-t_{kk}+h_{k}^{*}\bar{(M}_{k}-zI)^{-1}h_{k}}{z}.
Proof B.4 (Proof of Remark 4).
hk(M¯kzI)1hk\displaystyle h_{k}^{*}\bar{(M}_{k}-zI)^{-1}h_{k} =npwkWk(M¯kzI)1Wkwk\displaystyle=\frac{n}{p}w_{k}^{*}W_{k}^{*}\bar{(M}_{k}-zI)^{-1}W_{k}w_{k}
=nptr(WkwkwkWk(M¯kzI)1)\displaystyle=\frac{n}{p}tr(W_{k}w_{k}w_{k}^{*}W_{k}^{*}\bar{(M}_{k}-zI)^{-1})
=nλkpntr(Wkdiag(τ112,..,τn12)UYkYkUdiag(τ112,..,τn12)Wk(M¯kzI)1)\displaystyle=\frac{n\lambda_{k}}{pn}tr(W_{k}diag(\tau_{1}^{\frac{1}{2}},..,\tau_{n}^{\frac{1}{2}})UY_{k}Y_{k}^{*}U^{*}diag(\tau_{1}^{\frac{1}{2}},..,\tau_{n}^{\frac{1}{2}})W_{k}^{*}\bar{(M}_{k}-zI)^{-1})
=nλkptr(Wkdiag(τ112,..,τn12)U(YkYkI)Udiag(τ112,..,τn12)Wk(M¯kzI)1)\displaystyle=\frac{n\lambda_{k}}{p}tr(W_{k}diag(\tau_{1}^{\frac{1}{2}},..,\tau_{n}^{\frac{1}{2}})U(Y_{k}Y_{k}^{*}-I)U^{*}diag(\tau_{1}^{\frac{1}{2}},..,\tau_{n}^{\frac{1}{2}})W_{k}^{*}\bar{(M}_{k}-zI)^{-1})
+nλkptr(Wkdiag(τ1,..,τn)Wk(M¯kzI)1)\displaystyle+\frac{n\lambda_{k}}{p}tr(W_{k}diag(\tau_{1},..,\tau_{n})W_{k}^{*}\bar{(M}_{k}-zI)^{-1})
=nλkptr(Wkdiag(τ112,..,τn12)U(YkYkI)Udiag(τ112,..,τn12)Wk(M¯kzI)1)\displaystyle=\frac{n\lambda_{k}}{p}tr(W_{k}diag(\tau_{1}^{\frac{1}{2}},..,\tau_{n}^{\frac{1}{2}})U(Y_{k}Y_{k}^{*}-I)U^{*}diag(\tau_{1}^{\frac{1}{2}},..,\tau_{n}^{\frac{1}{2}})W_{k}^{*}\bar{(M}_{k}-zI)^{-1})
+nλkptr(Λk12UYn(diag(τ1,..,τn))2YnUΛk12(M¯kzI)1)\displaystyle+\frac{n\lambda_{k}}{p}tr(\Lambda_{k}^{\frac{1}{2}}UY_{n}(diag(\tau_{1},..,\tau_{n}))^{2}Y_{n}^{*}U^{*}\Lambda_{k}^{\frac{1}{2}}\bar{(M}_{k}-zI)^{-1})
=nλkptr(Wkdiag(τ112,..,τn12)U(YkYkI)Udiag(τ112,..,τn12)Wk(M¯kzI)1)\displaystyle=\frac{n\lambda_{k}}{p}tr(W_{k}diag(\tau_{1}^{\frac{1}{2}},..,\tau_{n}^{\frac{1}{2}})U(Y_{k}Y_{k}^{*}-I)U^{*}diag(\tau_{1}^{\frac{1}{2}},..,\tau_{n}^{\frac{1}{2}})W_{k}^{*}\bar{(M}_{k}-zI)^{-1})
+nλkptr(Λk12UYn(lτn2elel(lτn2)I)YnUΛk12(M¯kzI)1)\displaystyle+\frac{n\lambda_{k}}{p}tr(\Lambda_{k}^{\frac{1}{2}}UY_{n}(\sum_{l}\tau_{n}^{2}e_{l}e_{l}^{*}-(\sum_{l}\tau_{n}^{2})I)Y_{n}^{*}U^{*}\Lambda_{k}^{\frac{1}{2}}\bar{(M}_{k}-zI)^{-1})
+nλkptr(Λk12UYn(lτn2)YnUΛk12(M¯kzI)1)\displaystyle+\frac{n\lambda_{k}}{p}tr(\Lambda_{k}^{\frac{1}{2}}UY_{n}(\sum_{l}\tau_{n}^{2})Y_{n}^{*}U^{*}\Lambda_{k}^{\frac{1}{2}}\bar{(M}_{k}-zI)^{-1})
=nλkptr(Wkdiag(τ112,..,τn12)U(YkYkI)Udiag(τ112,..,τn12)Wk(M¯kzI)1)\displaystyle=\frac{n\lambda_{k}}{p}tr(W_{k}diag(\tau_{1}^{\frac{1}{2}},..,\tau_{n}^{\frac{1}{2}})U(Y_{k}Y_{k}^{*}-I)U^{*}diag(\tau_{1}^{\frac{1}{2}},..,\tau_{n}^{\frac{1}{2}})W_{k}^{*}\bar{(M}_{k}-zI)^{-1})
+nλkptr(Λk12UXn(lτn2elel(lτn2)I)YnUΛk12(M¯kzI)1)\displaystyle+\frac{n\lambda_{k}}{p}tr(\Lambda_{k}^{\frac{1}{2}}UX_{n}(\sum_{l}\tau_{n}^{2}e_{l}e_{l}^{*}-(\sum_{l}\tau_{n}^{2})I)Y_{n}^{*}U^{*}\Lambda_{k}^{\frac{1}{2}}\bar{(M}_{k}-zI)^{-1})
+nλk(lτn2)ptr(Λk12U(YnYnI)UΛk12(M¯kzI)1)\displaystyle+\frac{n\lambda_{k}(\sum_{l}\tau_{n}^{2})}{p}tr(\Lambda_{k}^{\frac{1}{2}}U(Y_{n}Y_{n}^{*}-I)U^{*}\Lambda_{k}^{\frac{1}{2}}\bar{(M}_{k}-zI)^{-1})
+nλk(lτn2)ptr(Λk(M¯kzI)1).\displaystyle+\frac{n\lambda_{k}(\sum_{l}\tau_{n}^{2})}{p}tr(\Lambda_{k}\bar{(M}_{k}-zI)^{-1}).