physics0507111/RMT.tex
1: \documentclass[aps]{revtex4}
2: %\documentclass[12pt]{article}
3: %\documentstyle[aps,prl]{revtex}
4: \usepackage{rotate}
5: \usepackage{epsfig}
6: \usepackage{psfrag}
7: %\usepackage{a4}
8: 
9: \newcommand \be  {\begin{equation}}
10: \newcommand \bea {\begin{eqnarray} \nonumber }
11: \newcommand \ee  {\end{equation}}
12: \newcommand \eea {\end{eqnarray}}
13: \newcommand{\subs}[1]{{\mbox{\scriptsize #1}}}
14: \newcommand \Tr {\mbox{Tr}}
15: \newcommand \rmi {\mbox{i}}
16: \newcommand \rmdd {\mbox{d}}
17: 
18: \begin{document}
19: \title{Financial Applications of Random Matrix 
20: Theory: Old Laces and New Pieces
21: }
22: 
23: \author{Marc Potters$^*$, Jean-Philippe Bouchaud$^{\dagger,*}$
24: and Laurent Laloux$^*$}
25: \affiliation{
26: {$^*$ Science \& Finance, Capital Fund Management, 6 Bd Haussmann,}
27: {75009 Paris France}\\
28: {$^\dagger$ Commissariat \`a l'Energie Atomique, Orme des Merisiers}\\
29: {91191 Gif-sur-Yvette {\sc cedex}, France}}
30: 
31: 
32: \begin{abstract}
33: This contribution to the proceedings of the Cracow meeting on `Applications of 
34: Random Matrix Theory' summarizes a series of studies, some old and others 
35: more recent on financial applications of Random Matrix Theory (RMT). We first review 
36: some early results in that field, with particular emphasis on 
37: the applications of correlation cleaning to portfolio optimisation, and discuss
38: the extension of the Mar\v{c}enko-Pastur (MP) distribution to a non trivial `true' 
39: underlying correlation matrix. We then present new results concerning different problems 
40: that arise in a financial context: (a) the 
41: generalisation of the MP result to the case of an empirical correlation matrix (ECM)
42: constructed using exponential moving averages, for which we give a new elegant 
43: derivation (b) the specific dynamics of the `market' eigenvalue and its 
44: associated eigenvector, which defines an
45: interesting Ornstein-Uhlenbeck process on the unit sphere and 
46: (c) the problem of the dependence of ECM's on the observation 
47: frequency of the returns and its interpretation in terms of lagged cross-influences.
48: \end{abstract}
49: 
50: \maketitle
51: 
52: \section{Portfolio theory: basic results}
53: 
54: Suppose one builds a portfolio of $N$ assets with weight $w_i$ on the $i$th asset. 
55: The (daily) variance of the portfolio return is
56: given by:
57: \be
58: R^2=\sum_{ij} w_i \sigma_i C_{ij} \sigma_j w_j,
59: \ee
60: where $\sigma_i^2$ is the (daily) variance of asset $i$ and $C_{ij}$ is the correlation matrix.
61: If one has predicted gains $g_i$, then the expected gain of the portfolio is $G=\sum w_i g_i$.
62: 
63: In order to measure and optimize the risk of this portfolio, one therefore has to come up with a 
64: reliable estimate of the correlation matrix $C_{ij}$. This is difficult in general \cite{prl,book} 
65: since one has to 
66: determine of the order of $N^2/2$ coefficients out of $N$ time series of length $T$, and in general 
67: $T$ is not much larger than $N$ (for example, 4 years of data give 1000 daily returns, and the typical
68: size of a portfolio is several hundred stocks.) We will denote, in the following, $q=N/T$; an accurate
69: determination of the true correlation matrix will require $q \ll 1$.
70: If $r^i_t$ is the daily return of stock $i$ at time $t$, 
71: the empirical variance of each stock is: 
72: \be
73: \sigma^2_i=\frac{1}{T}\sum_{t}^T \left(r^i_t\right)^2,
74: \ee
75: and can be assumed for simplicity to be perfectly known (its relative mean square-error is $(2+\kappa)/T$, where 
76: $\kappa$ is the kurtosis of the stock, known to be typically 
77: $\kappa \approx 3$). In the above definition, we have, as usual, neglected the daily mean return, small compared to daily fluctuations. 
78: The empirical correlation matrix is obtained as:
79: \be\label{defE}
80: E_{ij}=\frac{1}{T}\sum_{t}^T x_t^i x_t^j; \qquad x^i_t\equiv r^i_t/\sigma_i.
81: \ee
82: If $T<N$, $\mathbf{E}$ has rank $T<N$, and has $N-T$ zero eigenvalues. 
83: Assume there is a ``true'' correlation matrix $\mathbf{C}$ from 
84: which past and future $x^i_t$ 
85: are drawn. The risk of a portfolio constructed {\it independently} of the past
86: realized $x^i_t$ is faithfully 
87: measured by:
88: \be
89: \left\langle R^2_E\right\rangle=\frac{1}{T}\sum_{ijt}w_i\sigma_i\left\langle x^i_t x^j_t\right\rangle
90: w_j\sigma_j \approx \sum_{ij} w_i \sigma_i C_{ij} \sigma_j w_j.
91: \ee
92: Because the portfolio is not constructed using $\mathbf{E}$, this estimate is unbiased and the relative 
93: mean square-error is small ($\sim 1/T$). Otherwise, the $w$'s would depend on the observed $x$'s and, as
94: we show now, the result can be very different.
95: 
96: 
97: Problems indeed arise when one wants to estimate the risk of an optimized portfolio, resulting from a 
98: Markowitz optimization scheme, which gives the portfolio with maximum expected return for a given risk or equivalently,
99: the minimum risk for a given return ($G$):
100: \be
101: w_i\sigma_i = G\frac{\sum_jC^{-1}_{ij} g_j/\sigma_j}{\sum_{ij}g_i/\sigma_i
102: C^{-1}_{ij} g_j/\sigma_j}
103: \ee
104: From now on, we drop $\sigma_i$ (which can always be absorbed in $g_i$ and $w_i$).
105: In matrix notation, one has:
106: \be
107: \mathbf{w}_C = G\frac{\mathbf{C}^{-1}\mathbf{g}}{\mathbf{g}^{T}\mathbf{C}^{-1}\mathbf{g}}
108: \ee
109: The question is to estimate the risk of this optimized portfolio, and in particular to understand the 
110: biases of different possible estimates. We define the following three quantities:
111: \begin{itemize} 
112: \item The ``In-sample'' risk, corresponding to the risk of the optimal portfolio over the period used
113: to construct it:
114: \be
115: R^2_\subs{in}=\mathbf{w}_E^{T}\mathbf{E}\mathbf{w}_E = \frac{G^2}{\mathbf{g}^{T}\mathbf{E}^{-1}\mathbf{g}}
116: \ee
117: \item The ``true'' minimal risk, which is the risk of the optimized portfolio in the ideal world where $\mathbf{C}$
118: would be perfectly known:
119: \be
120: R^2_\subs{true}=\mathbf{w}_C^{T}\mathbf{C}\mathbf{w}_C =\frac{G^2}{\mathbf{g}^{T}\mathbf{C}^{-1}\mathbf{g}}
121: \ee
122: \item  The ``Out-of-sample'' risk which is the risk of the portfolio constructed using $\mathbf{E}$, 
123: but observed on the next period of time: 
124: \be
125: R^2_\subs{out}=\mathbf{w}_E^{T}\mathbf{C}\mathbf{w}_E =
126: G^2 \frac{\mathbf{g}^{T}\mathbf{E}^{-1}\mathbf{CE}^{-1}\mathbf{g}}
127: {(\mathbf{g}^{T}\mathbf{E}^{-1}\mathbf{g})^2}
128: \ee
129: \end{itemize}
130: From the remark above, the result is expected to be the same (on average) 
131: computed with $\mathbf{C}$ or with $\mathbf{E}'$, the
132: ECM corresponding to the second time period. Since $\mathbf{E}$ is a noisy estimator of $\mathbf{C}$ 
133: such that $\langle\mathbf{E}\rangle=\mathbf{C}$, one can use a convexity argument for the inverse of 
134: positive definite matrices to show that in general:
135: \be
136: \langle\mathbf{g}^{T}\mathbf{E}^{-1}\mathbf{g}\rangle \geq \mathbf{g}^{T}\mathbf{C}^{-1}\mathbf{g}
137: \ee
138: Hence for large matrices, for which the result is self-averaging:
139: \be
140: R^2_\subs{in} \leq R^2_\subs{true}.
141: \ee
142: By optimality, one clearly has that:
143: \be
144: R^2_\subs{true} \leq R^2_\subs{out}.
145: \ee
146: These results show that the out-of-sample risk of an optimized portfolio is larger (and in practice, much larger,
147: see Fig 1) than the in-sample risk, which itself is an underestimate of the true minimal risk. This is a 
148: general situation: using past returns to optimize a strategy always leads to over-optimistic results because 
149: the optimization adapts to the particular realization of the noise, and is unstable in time. In the case where  
150: the true correlation matrix is $\mathbf{C}=\mathbf{1}$, one can show that
151: \cite{Kondor}:
152: \be
153: R^2_\subs{true} =\frac{G^2}{\mathbf{g}^{T}\mathbf{g}}\qquad
154: \mbox{{ and}}\qquad
155: R^2_\subs{in}={R^2_\subs{true}}\sqrt{1-q}=R^2_\subs{out}{(1-q)}
156: \ee
157: Only in the limit $q \to 0$ will these quantities coincide, which is expected since in this case the measurement
158: noise disappears. In the other limit $q \to 1$, the in-sample risk becomes zero since it becomes possible to find 
159: eigenvectors (portfolios) with zero eigenvalues (risk), simply due to the lack of data.
160: 
161: \begin{figure}
162: \begin{center}
163: \psfig{file=Fig_marko_color.eps,width=7cm} 
164: \end{center}
165: \caption{In sample (left curves) and out of sample (right curves) 
166: portfolio risk along the optimal 
167: curve in the risk return plane. Red lines are with the raw empirical matrix, blue lines 
168: with the cleaned matrix using RMT, showing how the risk underestimation can be
169: reduced by matrix cleaning. From \cite{dublin}.
170: }
171: \label{Fig1}
172: \end{figure}
173: 
174: 
175: \section{Matrix cleaning and RMT}
176: 
177: How should one `clean' the empirical correlation matrix to avoid, as much as possible, such biases in the estimation
178: of future risk? In order to get some general intuition on this problem, 
179: let us rewrite the Markowitz solution 
180: in terms of the eigenvalues $\lambda_k$ and eigenvectors $V_i^k$ of the 
181: correlation matrix:
182: \be
183: w_i \propto \sum_{kj} \lambda_k^{-1} V_i^k V_j^k g_j \equiv 
184: g_i + \sum_{kj} (\lambda_k^{-1}-1) V_i^k V_j^k g_j
185: \ee
186: The first term corresponds to the naive solution: one should invest proportionally to the expected gain 
187: (in units where $\sigma_i=1$). The correction term means that the weights 
188: of eigenvectors with $\lambda > 1$ are suppressed, 
189: whereas the weights of eigenvectors with $\lambda < 1$ are enhanced. Potentially, the optimal Markowitz solution 
190: allocates a very large weight to small eigenvalues, which may be entirely dominated by measurement noise and
191: hence unstable. A very naive way to avoid this is to go for the naive solution $w_i \propto g_i$, but with the 
192: $k^*$ largest eigenvectors projected out:
193: \be
194: w_i \propto  
195: g_i - \sum_{k\leq k^*;j} V_i^k V_j^k g_j,
196: \ee
197: so that the portfolio is market neutral (the largest eigenvector correspond to
198: a collective market mode, $V_i^1 \approx 1/\sqrt{N}$) 
199: and sector neutral (other large eigenvectors 
200: contain sector moves). Since most of the volatility is contained in the 
201: market and sector modes, the above portfolio is already quite good in terms of risk. More elaborated ways aim at retaining all 
202: eigenvalues and eigenvectors, but after having somehow cleaned them. A well known cleaning corresponds to the so-called
203: ``shrinkage estimator'': the empirical matrix is shifted `closer' to the identity matrix. This is
204: a Bayesian method that assumes that the prior empirical matrix is the identity, again only justified if
205: the market mode has been subtracted out. More explicitely:
206: \be
207: \mathbf{E}_c = \alpha \mathbf{E} +(1-\alpha) \mathbf{1} \qquad\mbox{{so}}\qquad
208: \lambda^k_c = 1 + \alpha (\lambda^k-1),
209: \ee
210: where the subscript $c$ corresponds to the cleaned objects. This method involves the parameter $\alpha$
211: which is undetermined, but somehow related to the expected signal to noise ratio. If the signal is 
212: large, $\alpha\to 1$, and $\alpha \to 0$ if the noise is large. Another possible interpretation is through a
213: constraint on the effective number of assets in the portfolio, defined as $(\sum_i w_i^2)^{-1}$ \cite{book}. Constraining this
214: number to be large (ie imposing a minimal level of diversification) is equivalent to choosing $\alpha$ small. 
215: 
216: Another route is eigenvalue cleaning, first suggested in \cite{prl,dublin}, where one replaces all low lying eigenvalues
217: with a unique value, and keeps all the high eigenvalues corresponding to meaningful economical information (sectors):
218: \be
219: \lambda^k_c = 1-\delta \qquad\mbox{{if}}\qquad k>k^*,
220: \qquad
221: \lambda^k_c = \lambda^k_E  \qquad\mbox{{if}}\qquad k\leq k^*,
222: \ee
223: where $k^*$ is the number of meaningful number of sectors and $\delta$
224: is chosen such that the trace of the correlation matrix is exactly
225: preserved. How should $k^*$ be chosen? The idea developed in 
226: \cite{prl,dublin} is to use Random Matrix Theory to determine the theoretical edge of the `random' part of the
227: eigenvalue distribution, and to fix $k^*$ such that $\lambda^{k^*}_E$ is close to this edge.
228: 
229: What is then the spectrum of a random correlation matrix? The answer is known in several cases, due to the work of Mar\v{c}enko and Pastur \cite{marcenkopastur} and others 
230: \cite{baisilverstein,sengupta,burdagorlich}. We briefly recall the results and some
231: elegant methods to derive them, with special emphasis on the problem of the largest eigenvalue, 
232: which 
233: we expand on below. Consider an empirical correlation matrix $\mathbf{E}$ of $N$ assets using $T$ data points, 
234: both very large, with $q=N/T$ finite. Suppose that the true correlations are given by $\langle x^i_t x^j_{t'} \rangle = C_{ij}\delta_{tt'}$. This defines the Wishart ensemble \cite{Wishart}. 
235: In order to find the eigenvalue density, one introduces the resolvent:
236: \be
237: G(z)=\frac{1}{N}\Tr\left[(z\mathbf{I}-\mathbf{E})^{-1}\right]
238: \ee
239: from which one gets:
240: \be\label{im_part}
241: \rho(\lambda) = \lim_{\epsilon \to 0} \frac{1}{\pi} \Im\left( G(\lambda-\rmi\epsilon)
242: \right).
243: \ee
244: The simplest case is when $\mathbf{C}=\mathbf{I}$. Then, $\mathbf{E}$ is a sum 
245: of rotationally invariant 
246: matrices $\delta E^t_{ij}=(x^i_t x^j_t)/T$. The trick in that case is to 
247: consider the so-called Blue function,
248: inverse of $G$, i.e. such that $B(G(z))=z$. The quantity $R(x)=B(x)-1/x$ 
249: is the `R-transform' of $G$, and is known to be
250: additive \cite{marcenkopastur,voiculescu,zee}. Since 
251: $\delta \mathbf{E}^t$ has one eigenvalue 
252: equal to $q$ and $N-1$ zero eigenvalues, one has:
253: \be
254: \delta G_t(z)=\frac{1}{N}\left(\frac{1}{z-q}+\frac{N-1}{z}\right)
255: \ee
256: Inverting $\delta G_t(z)$ to first order in $1/N$, 
257: \be
258: \delta B_t(x)=\frac{1}{x}+\frac{q}{N(1-qx)}\quad\longrightarrow \quad B_E(x)=\frac{1}{x}+\frac{1}{(1-qx)},
259: \ee
260: and finally
261: \be
262: G_E(z)=\frac{(z+q-1)-\sqrt{(z+q-1)^2-4zq}}{2 z q},
263: \ee
264: which finally reproduces the well known Mar\v{c}enko-Pastur (MP) result:
265: \be
266: \rho(\lambda) = \frac{\sqrt{4\lambda q - (\lambda+q-1)^2}}{2\pi \lambda q}
267: \ee
268: 
269: The case of a general $\mathbf{C}$ cannot be directly written as a sum of ``Blue'' functions, but
270: one can get a solution using the Replica trick or by summing planar diagrams, which gives the 
271: following relation between resolvents: \cite{baisilverstein,sengupta,burdagorlich}
272: \be\label{burda}
273: zG_E(z) = Z G_C(Z) \qquad\mbox{{where}}\qquad Z = \frac{z}{1+q(z G_E(z) -1)},
274: \ee 
275: from which one can easily obtain $\rho(\lambda)$ numerically 
276: for any choice of $\mathbf{C}$ \cite{burdagorlich}. [In fact, this result can even be obtained from
277: the original Mar\v{c}enko-Pastur paper by permuting the role of the appropriate matrices]. 
278: The above result does however {\it not} apply when $\mathbf{C}$ has isolated eigenvalues, and
279: only describes continuous parts of the spectrum. For example, if one considers a matrix $\mathbf{C}$
280: with one large eigenvalue that is separated from the `Mar\v{c}enko-Pastur sea', 
281: the statistics of this isolated
282: eigenvalue has recently been shown to be Gaussian \cite{BenArous} (see also below), with a width $\sim T^{-1/2}$, much smaller than 
283: the uncertainty on the bulk eigenvalues ($\sim q^{1/2}$). A naive application of Eq. (\ref{burda}),
284: on the other hand, would give birth to a `mini-MP' distribution around the top eigenvalue. This
285: would be the exact result only if the top eigenvalue of $C$ had a degeneracy proportional to $N$. 
286: 
287: From the point of view of matrix cleaning, however, these results show that: 
288: (i) the expected edge of 
289: the bulk, that determines $k^*$, obviously depends on the prior one has for $\mathbf{C}$. 
290: The simplest case where $\mathbf{C}=\mathbf{I}$ was investigated in particular in 
291: \cite{prl,Plerou}, 
292: with the results shown in Fig 2. 
293: Other choices are however possible and could lead to an improved cleaning algorithm; (ii) the uncertainty
294: on large eigenvalues is much less than that on the bulk eigenvalues, meaning that 
295: the bulk needs a lot of shrinkage, but the bigger eigenvalues less so -- at variance with the naive
296: shrinkage procedure explained above. An alternative route may consist in using the `power mapping' method
297: proposed by Guhr \cite{Guhr} or clustering methods \cite{Lillo}.
298: 
299: \begin{figure}
300: \begin{center}
301: \psfig{file=eigen-density.eps,width=7cm} 
302: \end{center}
303: \caption{Empirical eigenvalue density for 406 stocks from the S\&P 500, and fit using the
304: MP distribution. Note (i) the presence of one very large eigenvalue, corresponding to
305: the market mode (see section IV) and (ii) the MP fit reveals systematic deviations, 
306: suggesting a non trivial structure of the true correlation matrix, even after sector modes
307: have been accounted for (see \cite{burdagorlich,us}).
308: }
309: \label{Fig2}
310: \end{figure}
311: 
312: 
313: %\epsfig{file=eigenvectors_color.eps,height=0.90\textheight}
314: 
315: 
316: \section{EWMA Empirical Correlation Matrices}
317: 
318: Consider now the case where $\mathbf{C}=\mathbf{I}$  but where the Empirical matrix is computed using an exponentially 
319: weighted moving average (EWMA). More precisely:
320: \be
321: E_{ij}=\epsilon\sum_{t'=-\infty}^{t-1}(1-\epsilon)^{t-t'} x_i^{t'} x_j^{t'} 
322: \ee
323: with $\epsilon=1/T$.
324: Such an estimate is standard practice in finance. Now, as an ensemble $E_{ij}$ satisfies $E_{ij}=(1-\epsilon) E_{ij} + 
325: \epsilon x_i^{t} x_j^{t}$. We again invert the resolvent of $\delta \mathbf{E_t}$ 
326: to find the elementary R-function,
327: \be
328: \delta B_t(x)=\frac{1}{x}+R_t(x)\qquad\mbox{{with}}\qquad R_t(x)=\frac{q}{N(1-qx)}
329: \ee
330: Using the scaling properties of $G(z)$ we find for $R(x)$:
331: \be
332: R_{aC}(x)=aR_C(ax).
333: \ee
334: This allows one to write:
335: \be
336: R_E(x)=R_{(1-\epsilon) E}(x) + R_t(x) = (1-q/N) R_E((1-q/N)x)+\frac{q}{N(1-qx)}
337: \ee
338: To first order in $1/N$, one now gets:
339: \be
340: R(x)+xR'(x)+\frac{q}{1-qx}=0 \longrightarrow R(x)=-\frac{\log(1-qx)}{qx}.
341: \ee
342: Going back to the resolvent to find the density, we finally get the result first obtained 
343: in \cite{PKP}:
344: \be
345: \rho(\lambda) = \frac{1}{\pi} \Im G(\lambda)\quad\mbox{{ where $G(\lambda)$ solves }}
346: \quad\lambda q G=q-\log(1-qG)
347: \ee
348: This solution is compared to the standard MP distribution in Fig 3.
349: 
350: Another nice properties of Blue functions is that they can be used to find the edges
351: of the eigenvalue spectrum ($\lambda_\pm$). One has:\cite{zee}
352: \be
353: \lambda_\pm =B(x_\pm)\qquad\mbox{where}\qquad B'(x_\pm)=0
354: \ee
355: In the case at hand, by evaluating $B(x)$ when $B'(x)=0$ we can write 
356: directly an equation whose solutions are the spectrum edges ($\lambda_\pm$)
357: \be
358: \lambda_\pm=\log(\lambda_\pm)+q+1
359: \ee
360: When $q$ is zero, the spectrum is a $\delta$ in $1$ as expected. But as the noise increases (or the characteristic
361: time decreases) the lower edge approach zero very quickly as $\lambda_-\sim\exp(-q)$. Although there are no exact zero eigenvalues for EWMA matrices, 
362: the smallest eigenvalue is very close to zero.  
363: 
364: \begin{figure}
365: \begin{center}
366: \psfig{file=fig-sp_exp+std.eps,width=7cm} 
367: \end{center}
368: \caption{Spectrum of the exponentially weighted
369: random matrix with $q\equiv N\epsilon=1/2$ and the spectrum of the standard
370: Wishart random matrix with $q\equiv N/T=1/3.45$, chosen to have the same upper edge. 
371: From \cite{PKP}.
372: }
373: \label{Fig3}
374: \end{figure}
375: 
376: 
377: \section{Dynamics of the top eigenvalue and eigenvector}
378: 
379: As mentioned above, it is well known that financial covariance matrices
380: are such that the largest eigenvalue is 
381: well separated from the `bulk', where all other eigenvalues reside. 
382: The financial interpretation of this large 
383: eigenvalue is the so-called `market mode': in a first approximation, all stocks move together, up or down. One can state 
384: this more precisely in the context of the one factor model, where the $i$th stock return at time $t$ is written
385: as:
386: \be
387: r_t^i = \beta_i \phi_t + \varepsilon_t^i,
388: \ee
389: where the market mode $\phi_t$ is common to all stocks through their market exposure $\beta^i$ and the  
390: $\varepsilon_t^i$ are idiosyncratic noises, uncorrelated from stock to stock. Within such a model, the covariance 
391: matrix reads:
392: \be
393: C_{ij} = \beta_i \beta_j \sigma_\phi^2 + \sigma_i^2 \delta_{ij}.
394: \ee
395: When all $\sigma_i$'s are equal, this matrix is easily diagonalized; for $N$ stocks, its largest eigenvalue is
396: $\Lambda_0= (\sum_j \beta_j^2) \sigma_\phi^2 + \sigma^2$ and is of order $N$, and all the other $N-1$ eigenvalues
397: $\Lambda_\alpha$ are equal to $\sigma^2$. The largest eigenvalue corresponds to the eigenvector $\beta_i$. More 
398: generally, the largest eigenvalue $\Lambda_0$, normalized by the average square volatility of the stocks, can be seen as a 
399: proxy for the average interstock correlation.
400: 
401: A natural question, of great importance for portfolio management, or 
402: dispersion trading (option strategies based on the implied average 
403: correlation between stocks), 
404: is whether $\Lambda_0$ and the $\beta$'s are stable in time. Of course, 
405: the largest eigenvalue and eigenvector of the
406: empirical correlation matrix will be, as discussed at length above, affected by measurement noise. Can one make 
407: predictions about the fluctuations of both the largest eigenvalue and the corresponding eigenvector induced by 
408: measurement noise? This would help separating a true evolution in time of the average stock correlation 
409: and of the market exposure of each stock from one simply related to measurement noise. We shall see that such a 
410: decomposition seems indeed possible in the limit where $\Lambda_0 \gg \Lambda_\alpha$. 
411: 
412: We will assume, as in the previous section, that the covariance matrix 
413: is measured through an exponential moving average of the returns. 
414: This means that
415: the matrix ${\mathbf E}$ evolves in time as:
416: \be
417: E_{ij,t} = (1 - \epsilon) E_{ij,t-1} + \epsilon r_t^i r_t^j.
418: \ee
419: The {\it true} covariance matrix $C_{ij} = \langle r^i r^j \rangle$ 
420: is assumed to be time independent -- this will give
421: us our benchmark hypothesis -- with its largest eigenvalue $\Lambda_0$ associated to the normalized eigenvector $\Psi_0$. In this section we 
422: deal with covariance
423: matrices instead of correlation matrices for simplicity, but most results 
424: should carry over to this latter case as well.
425: 
426: Denoting as $\lambda_{0t}$ the largest eigenvalue of $E_t$ associated to $\psi_{0t}$, standard perturbation theory, valid
427: for $\epsilon \ll 1$, gives:
428: \be
429: \lambda_{0t} = (1 - \epsilon) \lambda_{0t-1} + \epsilon \langle \psi_{0t-1} | C | \psi_{0t-1} \rangle 
430: + \epsilon \langle \psi_{0t-1} | \eta_{t} | \psi_{0t-1} \rangle,
431: \ee
432: with $\eta_{ij} = r^i r^j - \langle r^i r^j \rangle$. We will suppose for simplicity that the returns are Gaussian, yielding:
433: \be \label{etavar}
434: \langle \eta_{ij}\eta_{k\ell} \rangle = C_{ik}C_{j\ell} + C_{i\ell}C_{jk}.
435: \ee
436: In the limit where $\Lambda_0$ becomes much larger than all 
437: other eigenvectors, the above equation simplifies to:
438: \be\label{OU}
439: \lambda_{0t} \approx (1 - \epsilon) \lambda_{0t-1} 
440: + \epsilon \cos^2 \theta_{t-1} \Lambda_0 \left[ 1 + \xi_t \right],  
441: \ee
442: where $\cos \theta_t \equiv \langle \psi_{0t} | \Psi_0 \rangle$ and $\xi_t$ is a random noise term of mean zero and 
443: variance equal to $2$. This noise becomes Gaussian in the limit of large matrices, leading to a Langevin 
444: equation for $\lambda_0$:
445: \be\label{Orn1}
446: \frac{d\lambda_0}{dt} = \epsilon(\cos^2 \theta \Lambda_0- \lambda_{0}) + 
447: \epsilon \cos^2 \theta \xi_t.
448: \ee
449: We have neglected in the above equation a deterministic term equal to 
450: $\epsilon \sin^2 \theta \Lambda_1$, which will turn out to be a factor
451: $(\Lambda_1/\Lambda_0)^2$ smaller than the terms retained in Eq. (\ref{OU}).
452: 
453: We now need to work out an equation for the projection of the instantaneous eigenvector 
454: $\psi_{0t}$ on the true eigenvector $\Psi_0$. This can again be done using perturbation theory, which gives, in 
455: braket notation:
456: \bea\nonumber
457: | \psi_{0t} \rangle &=& | \psi_{0t-1} \rangle + \epsilon \sum_{\alpha \neq 0} 
458: \frac{\langle \psi_{\alpha t-1} | r_t r_t| \psi_{0t-1} \rangle}{\lambda_{0t-1} - \lambda_{\alpha t-1}}
459: | \psi_{\alpha t-1} \rangle\\ 
460: &\approx& | \psi_{0t-1} \rangle + \epsilon \frac{r_t r_t| \psi_{0 t-1} \rangle}{\lambda_{0t-1}}
461: - \epsilon \frac{\langle \psi_{0t-1} | r_t r_t| \psi_{0t-1} \rangle}{\lambda_{0t-1}}
462: | \psi_{0t-1} \rangle,
463: \eea
464: where we have used the fact that the basis of eigenvectors is complete. It is clear by inspection that the 
465: correction term is orthogonal to $| \psi_{0t-1} \rangle$, so that $| \psi_{0t} \rangle$ is still, to first 
466: order in $\epsilon$, normalized. Let us now decompose the matrix $r_t r_t$ into its average part $C$ and
467: the fluctuations $\eta$, and first focus on the former contribution. Projecting the above equation on $< \Psi_0 |$ leads
468: to the deterministic part of the evolution equation for $\cos \theta_t$:
469: \be
470: \cos \theta_t \approx \cos \theta_{t-1} + \epsilon \cos \theta_{t-1} \frac{\Lambda_0}{\lambda_{0t-1}}
471: - \epsilon \cos^3 \theta_{t-1} \frac{\Lambda_0}{\lambda_{0t-1}},
472: \ee
473: where we have neglected the contribution of the small 
474: eigenvalues compared to $\Lambda_0$, which is again a factor 
475: $(\Lambda_1/\Lambda_0)^2$ smaller. In the continuous 
476: time limit $\epsilon \to 0$, this equation can be rewritten as:
477: \be\label{Orn}
478: \frac{d\theta}{dt} = - \frac{\epsilon \Lambda_0}{2\lambda_{0t}}  \sin 2\theta,
479: \ee
480: and describes a convergence of the angle $\theta$ towards $0$ or $\pi$ -- clearly, $\Psi_0$ and $-\Psi_0$ are 
481: equivalent. It is the noise term $\eta$ that will randomly push the instantaneous eigenvector away from its 
482: ideal position, and gives to $\theta$ a non-trivial probability distribution. Our task is therefore to compute the
483: statistics of the noise, which again becomes Gaussian for large matrices, so that we only need to compute its 
484: variance. Writing $| \psi_{0t} \rangle = \cos \theta_t | \Psi_{0} \rangle + \sin \theta_t | \Psi_{1t} \rangle$, 
485: where $| \Psi_{1t} \rangle$ is in the degenerate eigenspace corresponding to small eigenvalues $\Lambda_1$, 
486: and using Eq. (\ref{etavar}), we find that the noise term $\zeta_t$ to be added to Eq. (\ref{Orn}) is given by:
487: \be
488: \langle \zeta_t^2 \rangle 
489: \approx \frac{\epsilon^2}{\lambda_{0t}^2} \left[2 \Lambda_0^2 
490: \cos^2 \theta_t \sin^2 \theta_t + \Lambda_0 \Lambda_1 \cos^2 2\theta_t \right],
491: \ee
492: where we have kept the second term because it becomes the dominant source of noise when $\theta \to 0$, but neglected a term in $\Lambda_1^2$. The 
493: eigenvector $\psi_0$ therefore undergoes an Ornstein-Uhlenbeck like motion 
494: on the unit sphere.
495: One should also note that the two sources of noise $\xi_t$ and $\zeta_t$ are not independent. Rather, one has, neglecting $\Lambda_1^2$ terms:
496: \be\label{correl}
497: 2\langle \xi_t \zeta_t \rangle \approx \Lambda_0 \cos^2 \theta_t \sin 2\theta_t - \Lambda_1 \sin 4\theta_t  
498: \ee
499: In the continuous time limit, we therefore have two coupled Langevin equations for the top eigenvalue $\lambda_{0t}$
500: and the deflection angle $\theta_t$. In the limit $\Lambda_1 \ll \Lambda_0$, the stationary solution for the angle can be 
501: computed to be:
502: \be
503: P(\theta) = {\cal N} \left[\frac{1 + \cos 2\theta (1 - \frac{\Lambda_1}{\Lambda_0})}
504: {1 - \cos 2\theta (1 - \frac{\Lambda_1}{\Lambda_0})}\right]^{1/4\epsilon}
505: \ee
506: As expected, this distribution is invariant when $\theta \to \pi - \theta$, since $-\Psi_0$ is also a top eigenvector.
507: In the limit $\Lambda_1 \ll \Lambda_0$, one sees that the distribution becomes peaked around $\theta=0$ and $\pi$.
508: For small $\theta$, the distribution becomes Gaussian:
509: \be
510: P(\theta) \approx \frac{1}{\sqrt{2 \pi \epsilon \frac{\Lambda_1}{\Lambda_0}}}
511:  \exp\left(-\frac{\theta^2}{2 \epsilon 
512: \frac{\Lambda_1}{\Lambda_0}}\right),
513: \ee
514: leading to $\langle \cos^2 \theta \rangle 
515: \approx 1 - \epsilon {\Lambda_1}/2{\Lambda_0}$ The angle $\theta$ is less 
516: and less fluctuating as $\epsilon \to 0$ (as expected) but also as $\Lambda_1/\Lambda_0 \to 0$: a large separation of
517: eigenvalues leads to a well determined top eigenvector. In this limit, the
518: distribution of $\lambda_0$ also becomes Gaussian (as expected from general results 
519: \cite{BenArous}) and one finds, to leading order:
520: \be
521: \langle \lambda_0 \rangle \approx \Lambda_0 - \epsilon {\Lambda_1}/2; \qquad 
522: \langle(\delta \lambda_0)^2\rangle \approx \epsilon.
523: \ee
524: Therefore, we have shown that in the limit of large averaging time and one large top eigenvalue (a situation 
525: approximately realized for financial markets), the deviation from the true top eigenvalue 
526: $\delta \lambda_0$ and 
527: the deviation angle $\theta$ are independent 
528: Gaussian variables (the correlation between them indeed becomes 
529: zero as can be seen using Eq. (\ref{correl}) in that limit, both following Ornstein-Uhlenbeck
530: processes.  
531: 
532: \begin{figure}
533: \begin{center}
534: \psfig{file=Variog_lambda.eps,width=7cm,angle=270} 
535: \end{center}
536: \caption{Variogram of the top eigenvalue of the {\it correlation} matrix 
537: $\lambda_0$ (top curve) and (log)-variogram 
538: of the top eigenvector, $\ln\left\langle \langle \psi_{0t+\tau}|\psi_{0t}
539: \rangle \right \rangle$, as a function of the time lag $\tau$, for US stocks,
540: and for an exponential moving average parameter $\epsilon=1/30$ days. 
541: We show for comparison (bottom curve) the Ornstein-Uhlenbeck behaviour expected for both 
542: quantities in the case of a time independent underlying correlation structure.  
543: }
544: \label{Fig4}
545: \end{figure}
546: 
547: From these results one directly obtains the variogram of the top eigenvalue as:
548: \be
549: \langle[\lambda_{t+\tau}-\lambda_t]^2\rangle
550: =2 \epsilon \left(1 - \exp(-\epsilon \tau)\right).
551: \ee
552: This is the result expected in the absence of 
553: a `true' dynamical evolution of the structure
554: of the matrix. From Fig. 4, one sees that there is clearly an additional
555: contribution, hinting at a real evolution of the strength of the 
556: average correlation with time. 
557: One can also work out the average overlap of the top eigenvector
558: with itself as a function of time lag, $E(\langle \psi_{0t+\tau}|\psi_{0t}
559: \rangle)$. Writing again $| \psi_{0t} \rangle = 
560: \cos \theta_t | \Psi_{0} \rangle + \sin \theta_t | \Psi_{1t} \rangle$, one sees that an equation for 
561: the evolution of the transverse component $| \Psi_{1t} \rangle$ is a priori
562: also needed. It is easy to show, following the same method as above, that
563: the evolution of the angle $\phi_t$ made by the component with a fixed
564: direction is a free diffusion with a noise term of order $\epsilon 
565: \Lambda_1/\Lambda_0$. Therefore, on the time scale $\epsilon^{-1}$ over
566: which $\theta_t$ evolves,  $| \Psi_{1t} \rangle$ can be considered to 
567: be quasi-constant, leading to:
568: \be 
569: \left\langle \langle \psi_{0t+\tau}|\psi_{0t}
570: \rangle \right \rangle \approx E(\cos(\theta_t - \theta_{t+\tau})) \approx 
571: 1 - \epsilon \frac{\Lambda_1}{\Lambda_0} (1 - \exp(-\epsilon \tau)).
572: \ee
573: Any significant deviation from the above law would, again, indicate a true 
574: evolution of the market structure. Again, Fig. 4, provides some evidence of
575: such an evolution, although weaker than that of the top eigenvalue $\lambda_0$.
576: 
577: 
578: \section{Frequency dependent correlation matrices}
579: 
580: The very definition of the correlation matrix {\it a priori} depends on the time scale over which returns are 
581: measured. The return on time $\tau$ for stock $i$ is defined as: $r_{i,\tau}(t) = \ln p_i(t+\tau) - \ln p_i(t)$,
582: where $p_i(t)$ is the price at time $t$. The correlation matrix is then defined as:
583: \be
584: C_{ij}(\tau)=\frac{\langle r_{i,\tau}(t) r_{j,\tau}(t)\rangle_c}{\sigma_i \sigma_j}
585: \ee
586: A relatively well known effect is that the average inter-stock correlation grows with the observation time scale --
587: this is the so-called Epps effect \cite{Epps,Reno}. For example, for a collection of stocks from the FTSE index, one finds, in the period 1994-2003:
588: \be
589: \langle C_{i \neq j}(5') \rangle \approx 0.06; \qquad  
590: \langle C_{i \neq j}(1h) \rangle \approx 0.19; \qquad
591: \langle C_{i \neq j}(1d) \rangle \approx 0.29
592: \ee
593: Besides the change of the average correlation level, there is also a change of structure of the correlation matrix: the full eigenvalue distribution distribution 
594: changes with $\tau$. A trivial effect is that by increasing the observation frequency one also increases the number of observations; the parameter $q$ defined
595: above decreases and the noise band is expected to shrink. This, at first sight, appears to be a nice way to
596: get rid of the observation noise in the correlation matrix (see \cite{Iori} for a related discussion). Unfortunately, the problem (or the interesting effect,
597: depending on the standpoint) is that this is accompanied by a true modification of the correlations, for which we will provide a 
598: model below. In particular one observes the 
599: emergence of a larger number of eigenvalues leaking out from the bulk of the 
600: eigenvalue spectrum (and corresponding to `sectors') as the 
601: time scale $\tau$ increases. This effect was also noted by Mantegna \cite{Mantegna}: the structure of the minimal spanning tree
602: constructed from the correlation matrix evolves from a `star like' structure for small $\tau$'s (several minutes),
603: meaning that only the market mode is present, to a fully diversified tree at the scale of a day. Pictorially, the
604: market appears as an embryo which progressively forms and differentiates with time. 
605: 
606: The aim of this section is to introduce a simple model of lagged cross-influences that allows one 
607: to rationalize the mechanism leading to such an evolution of the correlation matrix. Suppose that the 
608: return of stock $i$ at time $t$ is influenced in a causal way by return of stock $j$ at all previous times $t' < t$. 
609: The most general linear model for this reads:
610: \be
611: r_{i,1}(t) = \xi_i(t) + 
612: \sum_j \int_{-\infty}^t {\rm d}t' K_{ij}(t-t') r_{j,1}(t')
613: \qquad \langle  \xi_i(t) \xi_j(t') \rangle = D_i \delta_{ij} \delta(t-t')
614: \ee
615: Here $\tau=1$ is the shortest time scale -- say a few seconds. The kernel $K_{ij}$ is in general non-symmetric and
616: describes how the return of stock $j$ affects, on average, that of stock $i$ a certain time later. 
617: We will define the lagged correlation ${\cal C}_{ij}(t-t')$ by:
618: \be
619: {\cal C}_{ij}(t-t') = \langle r_{i,1}(t) r_{j,1}(t') \rangle.
620: \ee
621: This matrix is, for $t \neq t'$, not symmetric; however, one has obviously ${\cal C}_{ij}(t-t')=
622: {\cal C}_{ji}(t'-t)$. These lagged correlations were already studied in \cite{Kertesz}.
623: Going to Fourier space, one finds the Fourier transform of the covariance matrix ${\cal C}_{ij}(\omega)=
624: {\cal C}_{ji}(-\omega)$:
625: \be
626: {\cal C}_{ij}(\omega) = \sum_k (1 - K(\omega))^{-1}_{ik}
627: (1 - K(-\omega))^{-1}_{jk} D_k 
628: \ee 
629: where $K(\omega)$ is the Fourier transform of $K(\tau)$ with by convention $K(\tau < 0)=0$. 
630: When cross-correlations are small, which is justified provided the $r_{i,1}(t)$ corresponds
631: to residual returns, where the market has been subtracted, the relation between ${\cal C}_{ij}$ and $K_{ij}$ 
632: becomes quite simple and reads, for $\tau > 0$:
633: \be
634: {\cal C}_{ij}(\tau) = D_j K_{ij}(\tau).
635: \ee
636: This equation allows one, in principle, to determine $K_{ij}(\tau)$ from the 
637: empirical observation of the lagged correlation matrix. Suppose for simplicity that the influence kernel takes the form 
638: $K_{ij}(\tau)=K_{ij}^0 e^{-\Gamma_{ij} \tau}$, then $K_{ij}(\omega)=K_{ij}^0/(i\omega + \Gamma_{ij})$.
639: In this model, the primary object is the influence matrix $K$ which has a much richer structure than the 
640: correlation matrix: each element defines an {\it influence strength} $K^0$ and an {\it synchronisation time} $\Gamma^{-1}$.
641: In fact, as shown in Figs. 5 and 6, fitting the empirical data requires that $K_{ij}$ is parameterized by a sum of 
642: at least {\it two} exponentials, one corresponding to a time scale of minutes, and a second one of hours; 
643: sometimes the influence strength corresponding to these two time scales have opposite signs. Pooling together
644: the parameters corresponding to different pairs of stocks, we find, as might have been expected, that 
645: strongly coupled stocks (large $K^0$) have short synchronisation times $\Gamma^{-1}$.
646: 
647: \begin{figure}
648: \begin{center}
649: \psfig{file=Laplace11.eps,width=7cm,angle=270} 
650: \end{center}
651: \caption{Typical self-influence kernel $K_{ii}(E)$ in Laplace space and 
652: fit with the Laplace transforms of the sum of two exponentials.
653: }
654: \label{Fig5}
655: \end{figure}
656: 
657: %\begin{center}
658: %{\epsfig{file=Laplace11.eps,height=0.8\textheight,angle=270}}
659: %Self-influence kernel in Laplace space and fit with two exponentials
660: %\end{center}
661: 
662: \begin{figure}
663: \begin{center}
664: \psfig{file=Laplace1j.eps,width=7cm,angle=270} 
665: \end{center}
666: \caption{Typical cross-influence kernels $K_{ij}(E)$ for three pairs of
667: stocks, and 
668: fit with the Laplace transforms of the sum of two exponentials. Note that
669: the influence amplitudes have different signs, even for the same pair of
670: stock, depending on the time scale.
671: }
672: \label{Fig6}
673: \end{figure}
674: 
675: 
676: %\begin{center}
677: %{\epsfig{file=Laplace1j.eps,height=0.8\textheight,angle=270}}
678: %Cross-influence kernel in Laplace space and fit with two exponentials
679: %\end{center}
680: 
681: Coming back to the observation that the correlation matrix is frequency dependent, one should note that the
682: scale dependent correlation matrix $C_{ij}(\tau)$ is related to ${\cal C}_{ij}(\omega)$ by:
683: \be
684: C_{ij}(\tau)= \langle r_{i,\tau} r_{j,\tau} \rangle_c = \int \ d\omega 
685: \ S^2(\omega \tau) {\cal C}_{ij}(\omega)
686: \ee
687: where $S(.)$ is the form factor (i.e. Fourier transform of the window used to define returns on scale $\tau$, for
688: example a flat window in the simplest case). Therefore, for $\tau$ small one finds that residuals are uncorrelated
689: (i.e. the correlation matrix has no structure beyond the market mode):
690: \be
691: C_{ij}(\tau \to 0) \approx D_i \delta_{ij},
692: \ee
693: whereas on long time scales the full correlation develops as:
694: \be
695: C_{ij}(\tau \to \infty) \approx D_i \delta_{ij} + \int_0^\infty {\rm d}\tau \left[D_j K_{ij}(\tau)+D_i K_{ji}(\tau)\right].
696: \ee
697: The emergence of correlation structure therefore reveals the lagged cross-influences in the market. Note that on
698: long time scales, small $K^0$'s can be counterbalanced by large synchronisation times $\Gamma^{-1}$, and lead to 
699: significant correlations between `weakly coupled' stocks.
700: 
701: We believe that a more systematic empirical study of the influence matrix $K_{ij}$ and the way it should be cleaned,
702: in the spirit of the discussion in section II, is worth investigating in details. 
703: 
704: \vskip 0.3cm
705: \small{
706: We want to thank Pedro Fonseca and Boris Schlittgen for many discussions on the
707: issues addressed in sections IV and V, and Szilard Pafka and Imre Kondor for sharing the
708: results on the EWMA matrices given in section III. We also thank G\'erard 
709: Ben Arous and Jack Silverstein for several clarifying discussions. We also thank the 
710: organisers of the meeting in Cracow for inviting us and for making 
711: the conference a success. 
712: }
713: \begin{thebibliography}{99}
714: 
715: \bibitem{prl}
716: L.\ Laloux, P.\ Cizeau, J.-P.\ Bouchaud and M.\ Potters,
717: Phys.\ Rev.\ Lett.\ {\bf 83}, 1467 (1999);
718: L.\ Laloux, P.\ Cizeau, J.-P.\ Bouchaud and M.\ Potters,
719: Risk {\bf 12}, No.\ 3, 69 (1999).
720: 
721: \bibitem{book}
722: J.-P.\ Bouchaud and M.\ Potters, 
723: {\it Theory of Financial Risk and Derivative Pricing}
724: (Cambridge University Press, 2003).
725: 
726: \bibitem{Kondor} S. Pafka, I. Kondor, Physica {\bf A319} 487 (2003) and 
727: cond-mat/0305475
728: 
729: \bibitem{dublin}
730: L.\ Laloux, P.\ Cizeau, J.-P.\ Bouchaud and M.\ Potters,
731: Int.\ J.\ Theor.\ Appl.\ Finance {\bf 3}, 391 (2000).
732: 
733: \bibitem{marcenkopastur}
734: V.\ A.\ Mar\v{c}enko and L.\ A.\ Pastur, Math.\ USSR-Sb, {\bf 1}, 457-483 (1967).
735: 
736: \bibitem{baisilverstein}
737: J.\ W.\ Silverstein and Z.\ D.\ Bai, Journal of Multivariate Analysis, {\bf 54} 175 (1995). 
738: 
739: \bibitem{sengupta}
740: A.\ N.\ Sengupta and P.\ Mitra, 
741: Phys.\ Rev.\ E {\bf 80}, 3389 (1999).
742: 	
743: \bibitem{burdagorlich}
744: Z. Burda, A. G\"{o}rlich, A. Jarosz and J. Jurkiewicz, Physica A, {\bf 343}, 295-310 (2004).
745: 
746: \bibitem{Wishart} J. Wishart, Biometrika, A{\bf20} 32 (1928).
747: 	
748: \bibitem{Plerou} V. Plerou, P. Gopikrishnan, B. Rosenow, L. N. Amaral, H. E. Stanley, 
749: Phys.\ Rev.\ Lett.\ {\bf 83}, 1471 (1999).
750: 
751: \bibitem{voiculescu} D. V. Voiculescu, K. J. Dykema, A. Nica, {\it Free Random Variables}, AMS,
752: Providence, RI (1992). 
753: 
754: \bibitem{zee} A. Zee, Nuclear Physics B {\bf 474}, 726 (1996)
755: 
756: \bibitem{BenArous} 
757: J. Baik, G. Ben Arous, and S. Peche, {\it Phase transition of the largest eigenvalue for 
758: non-null complex sample covariance matrices}, http://xxx.lanl.gov/abs/math.PR/0403022, 
759: to appear in Ann. Prob, and these proceedings
760: 
761: \bibitem{Guhr} T. Guhr, B. Kalber, J. Phys. A {\bf 36}, 3009  (2003), and these proceedings.
762: 
763: \bibitem{Lillo} F. Lillo, R. Mantegna, {\it Cluster analysis for portfolio optimisation}, 
764: these proceedings, and physics/0507006. 
765: 
766: \bibitem{us} M. Potters, J. P. Bouchaud, {\it Embedded sector ansatz for the true 
767: correlation matrix in stock markets}, in preparation.
768: 
769: \bibitem{PKP} S. Pafka, I. Kondor, M. Potters, {\it Exponential weighting and Random Matrix Theory
770: based filtering of financial covariance matrices}, cond-mat/0402573, to appear in Quantitative 
771: Finance.
772: 
773: \bibitem{Epps} T. Epps, {\it Comovement of stock prices in the very short run}, J. Am. Stat. Assoc. 74, 291 (1979)
774: 
775: \bibitem{Reno} R. Reno, Int. Journ. Th. Appl. Fin, {\bf 6}, 87 (2003)
776: 
777: \bibitem{Iori} O. Precup, G. Iori, {\it A comparison of high frequency cross correlation measures}, preprint.
778: 
779: \bibitem{Mantegna} G. Bonanno, N. Vandewalle and R. N. Mantegna,  
780: Physical Review {\bf E62} R7615-R7618 (2001) 
781: 
782: \bibitem{Kertesz} L. Kullman, J. Kertesz, T. Kaski, Phys. Rev. {\bf E66} 026125 (2002)
783: 
784: \end{thebibliography}
785: \end{document}
786: 
787: 
788: 
789: 
790: 
791: 
792: 
793: 
794: 
795: 
796: 
797: 
798: 
799: 
800: