\(\newcommand{\abs}[1]{\left\lvert#1\right\rvert}\) \(\newcommand{\norm}[1]{\left\lVert#1\right\rVert}\) \(\newcommand{\as}{\overset{a.s.}{\to}}\) \(\DeclareMathOperator*{\E}{\mathbb{E}}\)

Classical vs High-dimensional

L

et $n$ be # of observations, $p$ be # of variables. The classical regime allows $n$ to diverge, but assumes $p$ fixed. In contrast, the high-dimensional regime permits both $n$ and $p$ to diverge, $ p/n \to \gamma > 0$. Many of the classical results break down in that case. Here I consider eigenvalues and eigenvectors of a high-dimensional covariance matrix. This has immediate implications for covariance estimation, but also for all the statistical tools based on covariance estimates: PCA, GLS, GMM, classification, portfolio optimization, etc.

Consider a simple case   $X_i \overset{iid}{\sim} \mathcal{N}_p(\mathbf{0}, \Sigma),\quad i=1,\ldots, n.$ How to estimate $\Sigma$?

Some notation:

Sample covariance estimator   $S = \frac{1}{n}\sum_i^n X_iX_i’ = \frac{1}{n} X’X.$ Eigendecompositions   $\Sigma = ULU’ = \sum_j^p \ell_j \mathrm{u}_j \mathrm{u}_j’, \quad S = V\Lambda V’ = \sum_j^p \lambda_j \mathrm{v}_j \mathrm{v}_j’.$ Eigenvalues distinct, sorted in decreasing order. Eigenvectors chosen with the first element positive.

Clasical Regime

In a classical regime, $S$ is a very good estimator (Anderson 1963, Van der Vaart 2000): Unbiased   $\E(S) = \Sigma.$
Consistent   $S \as \Sigma$ as $n\to\infty.$
Asymptotically normal eigenvalues   \(\sqrt{n}(\lambda_i-\ell_i) \overset{d}{\to} \mathcal{N}(0,2\ell_i^2), \quad j=1,\ldots,p.\)
Is invertible.

It gets trickier in high dimensions It is especially interesting what happens to eigenvalues and eigenvectors in high dimensions. There are three key features: eigenvalue spreading, eigenvalue bias and eigenvectors inconsistency.


High-dimensional Regime

Eigenvalue spreading

Marchenko-Pastur (1967)

In high dimensions, sample eigenvalues $\lambda_j$ are more spread out than their population counterparts $\ell_j.$ In fact, the higher the dimension, the more is the spreading.

Consider the case when $\Sigma = I_p,$ i.e. $\ell_1 = \ldots = \ell_p = 1,$ and $p/n \to \gamma \le 1.$

Empirical d’n of eigenvalues of sample covariance   \(F_p(x) := \frac{1}{p} \# \{ \lambda_j\le x \}\)

Ukranian mathematicians Marchenko & Pastur (MP) showed that this empirical d’n converges $F_p(x) \to F(x),$ with the limit pdf given by:

\[f^{MP}(x) = \frac{\sqrt{(\lambda_+-x)(x-\lambda_-)}}{2\pi x \gamma}, \quad \lambda_+ = (1+\sqrt{\gamma})^2, \quad \lambda_- = (1-\sqrt{\gamma})^2.\]
\[\begin{split} F(x) = & \frac{1}{2} + \frac{1}{2\pi \gamma} \Big[\sqrt{(\lambda_+-x)(x-\lambda_-)} \\ & + (1+\gamma)\arcsin(\frac{x-1-\gamma}{2\sqrt{\gamma}}) + (1-\gamma)\arcsin(\frac{(1-\gamma)^2-(1+\gamma)x}{2x\sqrt{\gamma}})\Big]. \end{split}\]

Some properties: Mean $1,$
Mode $\frac{(1-\gamma)^2}{1+\gamma},$
Median \(m(\gamma),\) with \(1 - (\sqrt{2}-1)\gamma < m(\gamma) < 1\) and \(\underset{\gamma\to 0}{\lim} m(\gamma) = 1-\frac{\gamma}{3} + \mathcal{o}(\gamma).\)

Quarter circle Law

An interesting special case is when $\gamma = 1.$ Then the d’n of normalized sample singular values of $X,$ $s_i/\sqrt{n},$ converges to the “quarter circle” law:

\[f^{Q}(x) = \frac{\sqrt{(4-x^2)}}{\pi}, \quad 0\le x \le 2,\]

that is, the singular values of a random normal square matrix lie on a quarter circle. Moreover, its moments are Catalan numbers.

Bai & Yin’s (1993) Law

Also when $\Sigma = I_p$ and $\gamma \le 1$, the largest and smallest eigenvalues converge almost surely to the corresponding boundaries of the support,

\[\lambda_1 \as \lambda_+ \quad \text{and} \quad \lambda_p \as \lambda_-.\]

Notice that the larger is $\gamma$, the wider is the spreding and the stronger is the eigenvalues bias! This phenomenon is very general and is not limited to the identity case.


If $\gamma>1$, then the sample covariance has only $n$ positive eigenvalues, while the remaining $p-n$ equal zero. In that case the limit distribution has a differential form and an isolated point zero is added to the support:

\[F(dx) = (1-1/\gamma) \delta_0(dx) + f^{MP}(x)dx,\]

where $\delta_0$ is the Dirac delta at $0$.

Eigenvalue bias

Let’s consider a covariance with a few “spiked” eigenvalues.

BBP (2005) Phase transition

• \(X_i \overset{iid}{\sim} \mathcal{N}_p(0,\Sigma), \quad i=1,\ldots,n,\)
• \(p/n \to \gamma, \quad 0< \gamma \le 1,\;\) as \(\; n\to\infty,\)
• \(\Sigma = diag(\ell_1, \ldots, \ell_r, 1,\ldots,1), \quad \ell_r \ge 1\)

Top $r$ sample eigenvalues will converge, but not to their true counterparts. Depending on where the true counterparts are positioned wrt to the so-called Baik-Ben Arous-Peche (BBP) transition point \(\lambda_+^{1/2}\),

\[\lambda_j \overset{as}{\longrightarrow} \begin{cases} \lambda_+, \; & \ell_j < \lambda_+^{1/2}, \\ \ell_j + \gamma \frac{\ell_j}{\ell_j-1}, \; & \ell_j > \lambda_+^{1/2}, \end{cases} \quad \text{for} \quad j = 1,\ldots, r.\]

$\lambda_1$ is asymptotically upward biased, while $\lambda_p$ will be downward biased


Tracy-Widom (1996)

The exact asymptotic d’n is also known for both cases! Below the BBP transition point the top eigenvalues are distributed with Tracy-Widom d’n with rate $n^{2/3}$, above with Normal with rate $n^{1/2}:$

\[n^{2/3}\lambda_1 \overset{d}{\to} TW_1 \left(\lambda_+, \; \left(\frac{\lambda_+}{\gamma^{1/4}}\right)^{4/3} \right) \quad \text{if} \; \ell_j < \lambda_+^{1/2}\]
\[n^{1/2} \lambda_1 \overset{d}{\to} \mathcal{N}\left( \ell_j + \gamma \frac{\ell_j}{\ell_j-1}, \; 2\ell_j^2 \left(1-\frac{\gamma}{(\ell_j-1)^2} \right)\right) \quad \text{if} \; \ell_j > \lambda_+^{1/2}\]

That is, if the true spikes are not large enough, the sample eigendistribution will look like that of $\Sigma = I_p$, i.e. according to MP d’n. In the opposite case, the spiked sample eigenvalues will overshoot the true counterparts and lie above the MP sea.

Eigenvector inconsistency

Paul (2007) showed that when \(p/n \to \gamma \in (0,\infty)\), the sample eigenvectors are not consistent and hence PCA would generally be inconsistent. Their Theorem 4 characterizes precisely how bad this inconsistency is

\[\langle \mathrm{v}_j, \mathrm{u}_j \rangle^2 \as \begin{cases} 0, \quad & \ell_j < \lambda_+^{1/2}, \\ \frac{1-\gamma/(\ell_j-1)^2}{1+\gamma/(\ell_j-1)}, \quad & \ell_j > \lambda_+^{1/2}, \end{cases}\\ \abs{\langle \mathrm{v}_j, \mathrm{u}_k \rangle} \as 0, \quad \text{for} \quad j\ne k.\quad\quad\quad\quad\quad\;\]

In the special case where \(\Sigma = I\) and the \(X_{ij}\) are iid standard (real or complex) Gaussian random variables, it is known that the matrix of sample eigenvectors is Haar distributed.


PCA in high dimensions

Johnstone Lu (2009), Thm 1

Assume a $p$-dimensional one-factor model

\[\mathrm{x}_i = v_i\rho + \sigma z_i, \quad i=1,\ldots,n,\]

and that \(\frac{p}{n} \to c\) and \(\frac{\|\rho\|^2}{\sigma^2} \to \omega > 0\) and define the normalized inner product (cos of the angle)

\[R(\hat{\rho},\rho) = \frac{\hat{\rho}'\rho}{\|\hat{\rho}\|\|\rho\|}.\]

Then

\[\lim R^2(\widehat{\rho}, \rho) = \frac{(\omega^2-c)_+}{\omega^2 +c\omega} \quad a.s.\]

i.e. PCA eigenvector estimate is consistent iff \(p/n \to 0\).

Paul (2007) shows that this is also true for spiked covariance.

Luckily, consistency can be recovered if there exists a sparse representation in some basis. In that case, PCA on a subset of variables with sufficiently high variability can yield consistent estimates.