Fact-checked by Grok 2 weeks ago

Decorrelation

Decorrelation is a process in , , and related disciplines that reduces or eliminates correlations between random variables, signals, or data components, transforming them into uncorrelated or forms to simplify and enhance efficiency. In , decorrelation typically involves linear transformations that diagonalize the , such as the Mahalanobis transformation, which applies the inverse of the to center and decorrelate multivariate data. This approach, rooted in the work of , results in a new set of variables with an identity , confirming zero cross-correlations. It is foundational for techniques like and whitening, where the goal is to remove redundancies while preserving variance. In , decorrelation minimizes the function between signals, defined as the normalized integral of their product over time shifts, to values near zero without altering essential signal qualities like spectral content. Common methods include allpass filters for audio applications, which create mutually uncorrelated versions of a signal for spatialization, and pairwise orthonormal transforms for sensor networks, aiding by isolating . In image processing, it removes spatial redundancies between pixels via lossless techniques, improving by reducing . Beyond these core areas, decorrelation plays a key role in by enforcing independence among features or layers to boost and mitigate biases, as in decorrelated , which applies orthogonal constraints during to decorrelate activations across the network. This enhances model efficiency and robustness, particularly in deep neural networks handling high-dimensional data.

Core Principles

Definition

Decorrelation is the process of applying a to a set of correlated random variables to produce a new set of uncorrelated random variables. This removes linear dependencies (correlations) between the variables, simplifying their joint distribution for subsequent analysis or modeling, though it does not necessarily imply full statistical . For jointly Gaussian random variables, however, decorrelation does imply statistical . In essence, decorrelation targets the off-diagonal elements of the , setting them to zero without altering the diagonal elements that represent variances. Pairwise decorrelation addresses the correlation between just two variables, whereas multivariate decorrelation extends this to an entire set, ensuring no correlations exist across all pairs. For pairwise cases, a common approach yields a transformed variable uncorrelated with the original, such as the residual in a linear prediction. A basic example involves two correlated Gaussian random variables X and Y with correlation coefficient \rho > 0. Applying the transformation Z = Y - \rho \frac{\sigma_Y}{\sigma_X} X produces Z that is uncorrelated with X (i.e., \operatorname{Cov}(X, Z) = 0), while maintaining appropriate mean and variance for the new variable. The foundations of decorrelation emerged in the late through Karl Pearson's work on and in statistics.

Measures of Correlation

is a fundamental statistical measure that quantifies the extent to which two random variables vary together, serving as an indicator of their linear dependence. For two random variables X and Y with means \mu_X and \mu_Y, the is defined as \operatorname{Cov}(X, Y) = E[(X - \mu_X)(Y - \mu_Y)], where E[\cdot] denotes the . A positive indicates that as one variable increases, the other tends to increase, while a negative value suggests an inverse relationship; a value of zero implies no linear association, though the variables may still be dependent in other ways. The Pearson correlation coefficient normalizes covariance to provide a standardized measure of linear association between two variables, ranging from -1 to 1. It is computed as \rho = \frac{\operatorname{Cov}(X, Y)}{\sigma_X \sigma_Y}, where \sigma_X and \sigma_Y are the standard deviations of X and Y, respectively. A value of \rho = 1 signifies perfect positive linear correlation, \rho = -1 indicates perfect negative linear correlation, and \rho = 0 denotes no linear correlation. This coefficient is widely used because it is scale-invariant and interpretable in terms of the strength and direction of the linear relationship. Despite their utility, linear measures like and Pearson have significant limitations, as they only detect linear dependencies and may overlook nonlinear relationships. For instance, consider X uniformly distributed on [-1, 1] and Y = X^2; while \operatorname{Cov}(X, Y) = 0, indicating no linear , X and Y are clearly dependent since Y is a deterministic of X. Such cases highlight how linear measures can fail to capture functional dependencies that are not monotonic or straight-line in nature. To address these shortcomings, higher-order measures such as are employed to quantify nonlinear dependencies between variables. I(X; Y) between X and Y is defined as I(X; Y) = \int p(x,y) \log \frac{p(x,y)}{p(x)p(y)} \, dx \, dy, where p(x,y), p(x), and p(y) are the joint and marginal probability density functions, respectively. This measure captures the amount of one variable contains about the other, with I(X; Y) = 0 implying (and thus no of any form), while positive values indicate dependence, including nonlinear forms.

Mathematical Techniques

Linear Methods

Linear methods for decorrelation employ linear transformations to orthogonalize data representations, thereby eliminating linear dependencies as measured by the . These approaches are particularly effective for assuming Gaussian distributions or linear relationships, where uncorrelated components simplify and processing. Foundational techniques include orthogonalization procedures and eigendecomposition-based , which transform correlated variables into an uncorrelated basis while preserving key statistical properties like total variance. The Gram-Schmidt process provides a sequential to orthogonalize a set of linearly independent vectors, yielding an that corresponds to uncorrelated directions when applied to representations of random variables. Given an initial set of vectors \mathbf{v}_1, \mathbf{v}_2, \dots, \mathbf{v}_n in an , the process constructs orthogonal vectors \mathbf{u}_1, \mathbf{u}_2, \dots, \mathbf{u}_n as follows: start with \mathbf{u}_1 = \mathbf{v}_1, then for k = 2 to n, compute \mathbf{u}_k = \mathbf{v}_k - \sum_{j=1}^{k-1} \proj_{\mathbf{u}_j} \mathbf{v}_k, where the projection is \proj_{\mathbf{u}_j} \mathbf{v}_k = \frac{\langle \mathbf{v}_k, \mathbf{u}_j \rangle}{\langle \mathbf{u}_j, \mathbf{u}_j \rangle} \mathbf{u}_j; finally, normalize each \mathbf{u}_k to obtain the orthonormal set \mathbf{q}_k = \mathbf{u}_k / \|\mathbf{u}_k\|. In the context of decorrelation, applying this to basis vectors derived from correlated features ensures the inner products (analogous to covariances under the standard ) between distinct basis elements are zero, producing uncorrelated components. This method is computationally straightforward for small dimensions but can suffer from numerical instability in practice due to error accumulation in sequential subtractions. Diagonalization of the \Sigma achieves global decorrelation by finding an P such that P^T \Sigma P = D, where D is diagonal with entries representing variances along uncorrelated axes. This is accomplished through eigendecomposition: since \Sigma is symmetric and positive semi-definite, it admits a \Sigma = V \Lambda V^T, where V is the of eigenvectors and \Lambda = \diag(\lambda_1, \dots, \lambda_p) contains the eigenvalues \lambda_i \geq 0 ordered decreasingly. To derive this, that the eigenvectors satisfy \Sigma \mathbf{v}_i = \lambda_i \mathbf{v}_i for i = 1, \dots, p, and orthogonality of V (i.e., V^T V = I) implies V^T \Sigma V = V^T (V \Lambda V^T) V = \Lambda, confirming D = \Lambda and P = V. The resulting decorrelates any centered random \mathbf{x} via \mathbf{y} = V^T \mathbf{x}, as \Cov(\mathbf{y}) = V^T \Sigma V = \Lambda, with off-diagonal elements zero. For a centered data matrix X \in \mathbb{R}^{n \times p}, the transformed data is Y = X V. This approach is optimal for preserving the data's second-order statistics and underpins many subsequent methods. Principal Component Analysis (PCA) operationalizes this diagonalization as a decorrelation tool, extracting principal components that are uncorrelated linear combinations of the original variables, ranked by explained variance. Formally, PCA applies the eigendecomposition \Sigma = V \Lambda V^T to the sample \Sigma = \frac{1}{n-1} X^T X (for centered X \in \mathbb{R}^{n \times p}), then projects onto the leading eigenvectors: Y = X V, yielding components with \Cov(Y) = \Lambda (diagonal) and total variance \trace(\Sigma) = \sum \lambda_i preserved under the . The first principal component captures the direction of maximum variance \lambda_1, the second is orthogonal and maximizes remaining variance, and so on, ensuring decorrelation since distinct eigenvectors are orthogonal. PCA thus reduces dimensionality while decorrelating features, with the variance preservation property guaranteeing no information loss in the second moments. For multivariate Gaussian data, fully decorrelates the variables into independent components, as the joint distribution transforms to a product of independent univariate Gaussians under the linear projection. Consider a bivariate zero-mean Gaussian with \Sigma = \begin{pmatrix} 2 & 1 \\ 1 & 2 \end{pmatrix}. The eigendecomposition yields eigenvalues \lambda_1 = [3](/page/3), \lambda_2 = 1 and eigenvectors \mathbf{v}_1 = \frac{1}{\sqrt{2}} \begin{pmatrix} 1 \\ 1 \end{pmatrix}, \mathbf{v}_2 = \frac{1}{\sqrt{2}} \begin{pmatrix} 1 \\ -1 \end{pmatrix}, so V = \frac{1}{\sqrt{2}} \begin{pmatrix} 1 & 1 \\ 1 & -1 \end{pmatrix} and \Lambda = \diag([3](/page/3), 1). Projecting data (x_1, x_2) to y_1 = \frac{x_1 + x_2}{\sqrt{2}}, y_2 = \frac{x_1 - x_2}{\sqrt{2}} results in \Cov(y_1, y_2) = 0, with \Var(y_1) = [3](/page/3) and \Var(y_2) = 1, decorrelating the original pair while the joint density factors as \mathcal{N}(y_1; 0, [3](/page/3)) \cdot \mathcal{N}(y_2; 0, 1).

Nonlinear Methods

Nonlinear methods for decorrelation address scenarios where dependencies between variables exhibit nonlinear structures, which linear techniques cannot fully resolve by merely zeroing covariances. These approaches seek statistical independence rather than mere uncorrelation, often leveraging higher-order statistics or kernel mappings to capture complex relationships. Key techniques include (ICA), kernel-based methods, and direct minimization of , each tailored to scenarios like blind source separation where mixing processes involve nonlinearities. Independent component analysis (ICA) models observed data \mathbf{X} as a linear \mathbf{X} = \mathbf{A} \mathbf{S}, where \mathbf{S} represents statistically source components and \mathbf{A} is an unknown mixing . The is to estimate a demixing matrix \mathbf{W} such that \mathbf{Y} = \mathbf{W} \mathbf{X} yields components \mathbf{Y} that are as independent as possible, assuming non-Gaussian sources. Independence is achieved by maximizing non-Gaussianity, quantified via J(\mathbf{y}_i) = H(\mathbf{z}_\mathrm{gauss}) - H(\mathbf{y}_i), where H denotes and \mathbf{z}_\mathrm{gauss} is a Gaussian with the same variance as \mathbf{y}_i. For super-Gaussian sources, negentropy is approximated as J(\mathbf{y}_i) \approx \sum [k(\mathbf{y}_i)^2 - 2], with k(\cdot) as a nonlinearity like \tanh. The FastICA algorithm implements this efficiently through fixed-point iterations: (1) center and whiten the data to obtain \mathbf{Z} = \mathbf{V} (\mathbf{X} - \mathbb{E}[\mathbf{X}]), where \mathbf{V} diagonalizes the covariance; (2) initialize weight vectors \mathbf{w}; (3) update \mathbf{w}^+ \propto \mathbb{E}[\mathbf{Z} g(\mathbf{w}^T \mathbf{Z})] - \mathbb{E}[g'(\mathbf{w}^T \mathbf{Z})] \mathbf{w}, with g as the derivative of the nonlinearity (e.g., g(u) = \tanh u); (4) orthogonalize and normalize \mathbf{w}; (5) repeat until convergence. This process decorrelates nonlinearly mixed signals by enforcing independence via higher-order statistics, outperforming linear methods in cases of non-quadratic mixing. Kernel-based decorrelation extends linear techniques into high-dimensional spaces using the trick, enabling nonlinear transformations without explicit computation of . In component analysis (), data points \mathbf{x}_i are mapped to \phi(\mathbf{x}_i) in a , where standard is applied to decorrelate the projected data. The covariance in space is \mathbf{K} = \Phi \Phi^T / n, with \Phi = [\phi(\mathbf{x}_1), \dots, \phi(\mathbf{x}_n)], and eigendecomposition of the centered matrix \mathbf{K} yields principal components that are uncorrelated in the nonlinear space. kernels include the k(\mathbf{x}, \mathbf{y}) = \exp(-\|\mathbf{x} - \mathbf{y}\|^2 / 2\sigma^2), allowing capture of nonlinear dependencies like circular manifolds that linear misses. This approach achieves decorrelation by implicitly linearizing nonlinear relations through the mapping. Mutual information minimization directly targets statistical independence by solving the optimization \min I(\mathbf{X}; \mathbf{Y}) subject to constraints preserving data variance, where I is the I(\mathbf{X}; \mathbf{Y}) = \sum p(x,y) \log \frac{p(x,y)}{p(x)p(y)}. Since exact computation is intractable, approximations via on parameterized densities or kernel density estimates are used, often incorporating regularization to avoid . In practice, this frames decorrelation as an information-theoretic problem, with updates derived from the score \nabla_\theta I \approx \mathbb{E}[\log p_\theta(\mathbf{y}|\mathbf{x}) - \log p(\mathbf{y})]. Such methods are particularly effective for continuous variables with nonlinear couplings, providing a principled alternative to contrast functions in ICA. A representative application of ICA is the separation of mixed audio signals, such as two speakers' voices recorded by microphones (the "cocktail party" problem), where the mixing is linear but sources are nonlinearly independent. FastICA recovers the individual speech streams by estimating the demixing matrix, successfully isolating non-Gaussian voices from their superposition and demonstrating how nonlinear methods capture dependencies that linear decorrelation overlooks.

Applications in Engineering

Signal Processing

In , decorrelation techniques are essential for removing redundancies between signal components, enhancing analysis, and improving system performance by transforming correlated signals into uncorrelated ones. This preprocessing step facilitates tasks such as suppression and by assuming signals can be modeled as random processes with known statistical properties. A prominent method is the , which preprocesses signals to produce characteristics—uncorrelated components with equal variance—by applying a linear transformation to the . The transformation is mathematically expressed as \mathbf{Y} = \Sigma^{-1/2} \mathbf{X}, where \mathbf{X} is the original signal vector, \Sigma is its , and \mathbf{Y} is the whitened output with covariance. This approach is particularly valuable in equalization, where it serves as a preprocessing step in blind adaptive equalizers to decorrelate received signals, mitigating and improving convergence in algorithms like the constant modulus algorithm (CMA). In array signal processing, decorrelation is applied to received signals at antenna arrays to resolve multiple sources, especially when signals are coherent due to multipath propagation. A key example is the MUltiple SIgnal Classification (MUSIC) algorithm, which incorporates a spatial decorrelation step, often via spatial smoothing techniques, to estimate the directions of arrival (DOAs) by decorrelating the array covariance matrix and separating signal subspaces from noise. This enables high-resolution source localization even in correlated environments, as demonstrated in seminal work on subspace-based methods. For multichannel audio processing, decorrelation enhances spatial imaging by reducing interchannel correlations that can collapse stereo width or degrade immersion. Techniques such as pairwise correlation subtraction estimate and subtract the correlated portions between channel pairs, preserving perceptual quality while widening the soundstage in stereo reproduction systems. This method is commonly used in audio codecs and upmixing to maintain low interchannel correlation coefficients, improving binaural cues without introducing artifacts. Key applications of decorrelation in advanced in the with adaptive filters for . Pioneering work by Bernard Widrow, who introduced the least mean squares (LMS) algorithm in 1960, applied it in adaptive cancellers. These techniques laid the foundation for modern adaptive systems, with applications expanding in the late to acoustic .

Communications

In communication systems, decorrelation techniques are essential for mitigating and multipath effects that degrade signal quality in both and wired channels. These methods transform correlated signals into uncorrelated ones, enabling reliable data transmission by reducing (ISI) and multiuser interference. By inverting channel correlations, decorrelation enhances and error performance, particularly in environments with and . Zero-forcing equalization represents a foundational decorrelation approach to combat ISI in channels with memory, such as those encountered in wired modems or wireless links. This linear inverse filtering technique designs an equalizer that nullifies the channel's dispersive effects, effectively decorrelating successive symbols. For a channel matrix \mathbf{H}, the zero-forcing equalizer \mathbf{W} is computed as \mathbf{W} = (\mathbf{H}^H \mathbf{H})^{-1} \mathbf{H}^H, where ^H denotes the Hermitian transpose, yielding an output where ISI is eliminated at the expense of potential noise enhancement. This method, widely adopted in early digital communication standards, provides a simple means to achieve orthogonality among symbols, though it assumes perfect channel knowledge and can amplify noise in ill-conditioned channels. In code-division multiple-access (CDMA) systems, decorrelating receivers address multiuser by separating overlapping user signals based on their signature sequences. The decorrelator matrix \mathbf{R}^{-1}, where \mathbf{R} is the matrix of user codes, projects the received signal onto directions orthogonal to interfering users, thereby eliminating multiple-access while preserving the desired signal's amplitude. Proposed in seminal work on linear multiuser detection, this approach achieves near-far resistance in synchronous CDMA scenarios, significantly improving bit error rates compared to conventional matched-filter receivers. It forms the basis for suppression in spread-spectrum systems, balancing computational simplicity with robust performance against correlated user signals. For multiple-input multiple-output () systems, transmit enables decorrelation at the transmitter to simplify receiver-side processing and mitigate spatial correlations induced by arrays or scattering environments. matrices are designed to orthogonalize the effective , such as by inverting the transmit , which preconditions the data streams before transmission. This transmit-side decorrelation reduces the receiver's burden of handling correlated streams, enhancing overall system capacity in correlated fading s. In V-BLAST architectures, combining decorrelation with power allocation further optimizes error rates by allocating resources to decorrelated subchannels. The application of decorrelation techniques has evolved significantly since the 1990s, originating with CDMA standards like IS-95 where decorrelating multiuser detectors were integrated to handle user interference in cellular networks. This foundation extended into 4G LTE through MIMO precoding schemes that employed zero-forcing to manage spatial multiplexing. In 5G massive MIMO, decorrelation via large-scale zero-forcing precoding exploits the channel hardening effect from numerous antennas, substantially reducing computational load at the receiver while supporting higher user densities and mitigating pilot contamination. These advancements have enabled spectral efficiencies exceeding 10 bits/s/Hz in practical deployments, underscoring decorrelation's role in scaling modern wireless systems.

Applications in Science

Neuroscience

In neuroscience, decorrelation refers to the process by which neural systems reduce statistical dependencies between the activities of neurons, thereby enhancing the efficiency of transmission in the . This mechanism is central to the efficient coding hypothesis, originally proposed by Horace Barlow in 1961, which posits that sensory systems evolve to minimize in neural representations of the environment, maximizing the conveyed per while incorporating principles like sparseness—where only a small fraction of neurons are active at any time—to optimize coding efficiency. Barlow's framework suggests that decorrelation transforms correlated sensory inputs into more independent neural outputs, aligning with the statistical structure of natural stimuli to reduce metabolic costs and improve representational capacity. In the , decorrelation begins in the , where retinal ganglion cells (RGCs) transform highly correlated photoreceptor inputs into less correlated spike trains, as demonstrated by electrophysiological recordings in and retinas exposed to natural scenes. This retinal decorrelation, achieved through mechanisms like cell diversity and inhibitory feedback, aligns RGC responses with the power spectrum of natural images, supporting efficient by reducing spatial and temporal —experiments show that RGC correlations drop significantly compared to input luminance fluctuations, with decorrelation indices approaching those predicted by optimal whitening filters. In the primary (), further decorrelation occurs among orientation-selective neurons, where surround inhibition from lateral connections suppresses responses to stimuli outside the classical , thereby reducing redundancy in representations of natural images; single-unit recordings in monkeys viewing natural movies reveal that V1 population activity exhibits near-zero pairwise correlations, contrasting with the high correlations in raw visual inputs, and this decorrelated sparse is tuned to the second-order statistics of natural scenes. Functional MRI studies corroborate these findings, showing decreased inter-neuronal correlations in during naturalistic viewing, indicative of a hierarchical decorrelation process from to .

Statistics and Machine Learning

In statistics and machine learning, decorrelation plays a crucial role in regression analysis by addressing multicollinearity among features, which can lead to unstable coefficient estimates and inflated variances. Principal component analysis (PCA) preprocessing transforms correlated predictors into orthogonal components, thereby removing multicollinearity and stabilizing regression coefficients in principal component regression (PCR). Similarly, ridge regression mitigates the effects of multicollinearity through L2 regularization, shrinking coefficients toward zero without explicitly decorrelating features, which improves model stability in high-dimensional settings. Decorrelation is also essential in clustering and dimensionality reduction tasks, where correlated features can distort distance metrics and hinder algorithm performance. Applying prior to decorrelates the data, providing a more effective initialization for centroids by projecting onto principal directions that capture maximum variance and reduce redundancy. This preprocessing step enhances clustering quality, as uncorrelated features allow k-means to better identify natural groupings without bias from feature dependencies. Recent advancements since 2010 have extended decorrelation to , particularly in decorrelated neural networks to accelerate training and improve . For instance, decorrelated (DBN), introduced in 2018, whitens mini-batch activations using zero-phase component analysis (ZCA), reducing feature and enabling faster compared to standard . A practical example of decorrelation in models involves preprocessing economic indicators, such as GDP components and metrics, which are often highly correlated. Using to extract principal components from a large set of these indicators allows for robust nowcasting and of macroeconomic trends, as demonstrated in diffusion index models that summarize collinear predictors into uncorrelated factors.

Applications in Other Fields

Cryptography

In stream ciphers, linear feedback shift registers (LFSRs) are employed to generate keystreams that exhibit high linear complexity, ensuring the output is statistically decorrelated from the initial state to resist cryptanalytic attacks. This decorrelation is quantified through the linear complexity, which measures the length of the shortest LFSR capable of producing the sequence; a high value makes prediction infeasible. The Berlekamp-Massey algorithm serves as the primary tool for analyzing this property, efficiently determining the minimal LFSR from a keystream segment and revealing any exploitable linear dependencies if the complexity is low. In block ciphers, decorrelation manifests through the diffusion layer, which propagates changes in the input to achieve the , thereby breaking statistical associations between and . For instance, in the (), altering a single input bit results in approximately half of the output bits flipping after a few rounds, typically reaching about 64 out of 128 bits affected, which ensures rapid decorrelation and resistance to differential attacks. Decorrelation provides a formal for proving in block cipher constructions, particularly within the Luby-Rackoff paradigm for building pseudorandom permutations from pseudorandom functions using Feistel networks. Introduced by Serge Vaudenay in the late 1990s, this defines decorrelation oracles that model adversary access to multiple views of the , enabling proofs that sufficient rounds yield indistinguishability from random permutations when the component functions satisfy pairwise or higher-order decorrelation bounds. In modern , decorrelation techniques, such as higher-order masking, are integrated into implementations to resist side-channel attacks by ensuring that intermediate computations are statistically independent, thereby decorrelating observable leaks (e.g., power or electromagnetic traces) from secret keys. This approach is critical for lattice-based schemes like , where masking splits sensitive variables into shares with no single share revealing information, maintaining security against even in resource-constrained environments. As of 2025, research continues into machine learning-enhanced side-channel attacks on post-quantum schemes, emphasizing advanced decorrelation in masking.

Finance

In finance, decorrelation strategies are essential for managing and optimizing by minimizing the impact of interdependent asset movements through of assets with low or zero correlations. These approaches enable investors to reduce overall volatility without necessarily sacrificing expected returns, as uncorrelated assets diversify away idiosyncratic risks. Portfolio diversification often involves constructing a to achieve uncorrelated asset returns and minimize risk. The is obtained by solving the quadratic optimization problem: \min_{\mathbf{w}} \mathbf{w}^T \Sigma \mathbf{w} \quad \text{subject to} \quad \mathbf{w}^T \mathbf{1} = 1, where \mathbf{w} denotes the of portfolio weights, \Sigma is the of asset returns, and \mathbf{1} is a of ones ensuring full . This exploits decorrelation within \Sigma to yield the lowest achievable variance, as pioneered by Markowitz in . Factor models further advance decorrelation by separating systematic market influences from asset-specific noise. The (APT), developed by Ross in 1976, posits that asset returns can be expressed as \mathbf{r} = \mathbf{\beta} \mathbf{F} + \boldsymbol{\epsilon}, where \mathbf{F} represents systematic factors, \mathbf{\beta} are factor loadings, and \boldsymbol{\epsilon} are idiosyncratic residuals assumed uncorrelated with factors and across well-diversified assets. This residual decorrelation allows pricing based on factor exposures while neutralizing unsystematic risks. Risk management employs decorrelation in Value-at-Risk () adjustments for , where scenarios simulate decorrelated or heightened regimes to probe vulnerabilities. perturbs the correlation matrix—often via factor models—within plausible bounds, then recomputes to quantify impacts on potential losses, as in variance-covariance methods. For instance, reverse identifies worst-case shifts that amplify while remaining statistically feasible. During the , attempts to decorrelate stock returns failed as correlations surged across global equities, with average pairwise correlations rising from around 0.4 pre-crisis to over 0.8 in late 2008, amplifying systemic losses. This breakdown highlighted the fragility of decorrelation assumptions under extreme volatility, where even diversified portfolios experienced synchronized declines. As of late 2024, analysts anticipate potential "great decorrelation" in 2025 due to shifting monetary policies, which could enhance diversification benefits if realized.

References

  1. [1]
    [PDF] Decorrelation in Statistics: The Mahalanobis Transformation
    Dec 7, 2000 · The dictionary definition of this term is “a variable whose values are random but whose statistical distribution is known.” Given a set of n ...
  2. [2]
    [PDF] Decorrelation as a By-Product of Granular Synthesis
    Decorrelation is defined here as any technique that reduces the absolute value of the cross-correlation measure between two signals y1 and y2 (not only allpass ...
  3. [3]
    [PDF] Distributed Signal Decorrelation and Detection in Sensor Networks ...
    Accurate anomaly detection requires decorrelation of the background signal [23]. In order to decorrelate the background, we need an accurate estimate of its ...
  4. [4]
    Decorrelation - an overview | ScienceDirect Topics
    Decorrelation is defined as the process that removes spatial redundancy between pixels in images, utilizing lossless image compression techniques, ...
  5. [5]
    [PDF] Machine Learning for Signal Processing Independent Component ...
    • Uncorrelated: Two random variables X and Y are ... – For uncorrelated components, YYT = Diagonal ... Decorrelation of H Interpretation: What does this mean?
  6. [6]
    [PDF] October 3 15.1 Review and Outline 15.2 Simple Linear Regression
    This is in some sense desirable since we assumed this about the population residuals. 3. The residuals are uncorrelated with the predictor, i.e., n. X i=1 b ...
  7. [7]
    VII. Note on regression and inheritance in the case of two parents
    Note on regression and inheritance in the case of two parents. Karl Pearson ... Published:01 January 1895https://doi.org/10.1098/rspl.1895.0041. Abstract.
  8. [8]
    Covariance: Formula, Definition & Example - Statistics By Jim
    Covariance in statistics measures the extent to which two variables vary linearly. The covariance formula reveals whether two variables move in the same or ...
  9. [9]
    Pearson Correlation Coefficient (r) | Guide & Examples - Scribbr
    May 13, 2022 · It is a number between –1 and 1 that measures the strength and direction of the relationship between two variables.What is the Pearson... · Calculating the Pearson... · Testing for the significance of...
  10. [10]
    [PDF] Reminder No. 1: Uncorrelated vs. Independent
    Feb 27, 2013 · The joint distribution of X and Y is not uniform on the rectangle [−1,1] ×. [0,1], as it would be if X and Y were independent (Figure 1). The ...
  11. [11]
    [PDF] A Mathematical Theory of Communication
    In the present paper we will extend the theory to include a number of new factors, in particular the effect of noise in the channel, and the savings possible ...Missing: mutual | Show results with:mutual
  12. [12]
    Analysis of a complex of statistical variables into principal components.
    Hotelling, H. (1933). Analysis of a complex of statistical variables into principal components. Journal of Educational Psychology, 24(7), 498–520.
  13. [13]
    [PDF] Gram--Schmidt Orthogonalization: 100 Years and More - UPenn CIS
    Jun 8, 2010 · The Gram-Schmidt process forms an orthogonal sequence from a linearly independent sequence in an inner-product space, computing an orthonormal ...
  14. [14]
    [PDF] Multivariate Gaussian Distribution - Purdue Engineering
    Gaussian Random Variable Decorrelation. • Consider ˜x = E t x, then. E ... is the sample correlation matrix. Sij = 1 n n. X k=1. XikXjk. • Decompose S as. S ...
  15. [15]
    [PDF] Independent Component Analysis: Algorithms and Applications
    For details, see (Hyvärinen, 1999b). In FastICA, convergence speed is optimized by the choice of the matrices diag(αi) and diag(βi). Another advantage of ...
  16. [16]
    [PDF] Nonlinear Component Analysis as a Kernel Eigenvalue Problem
    This article presented a new technique for nonlinear PCA. To develop this technique, we made use of a kernel method so far used only in supervised learning ...Missing: seminal | Show results with:seminal
  17. [17]
    [PDF] Statistical Consistency of Kernel Canonical Correlation Analysis
    This paper gives a mathematical proof of the statistical convergence of kernel CCA, providing a theoretical justification for the method. The proof uses ...Missing: seminal | Show results with:seminal<|control11|><|separator|>
  18. [18]
    [PDF] Optimal Whitening and Decorrelation - arXiv
    Dec 18, 2016 · In the following, we will make use of a number of covariance matrix identities: the decomposition Σ = V1/2PV1/2 of the covariance matrix into ...<|separator|>
  19. [19]
    CMA adaptive equalization in subspace pre-whitened blind receivers
    In blind channel equalization, the use of whitening transformation (WT) preceding the constant modulus algorithm (CMA), referred to as the pre-whitened CMA ...
  20. [20]
  21. [21]
    (PDF) Analysis of Decorrelation Methods in Multichannel Audio
    Jul 22, 2020 · Decorrelation techniques are applied for the purposes of data compression, unwanted interference removal, and signal perception improvement ...
  22. [22]
    US8239210B2 - Lossless multi-channel audio codec - Google Patents
    The channel pair decorrelation coefficient (ChPairDecorrCoeff) is calculated as the zero-lag cross-correlation estimate divided by the zero-lag auto-correlation ...
  23. [23]
    [PDF] Adaptive Noise Cancelling: Principles and Applications
    In the early and middle 1960's, work on adaptive systems intensified. Hundreds of papers on adaptation, adaptive con- trols, adaptive filtering, and adaptive ...<|control11|><|separator|>
  24. [24]
    [PDF] Topics in Acoustic Echo and Noise Control - ReadingSample - NET
    In contrast, the processing of acoustical echoes necessitates adaptive filters that are extremely demanding with respect to signal processing power. It is, ...
  25. [25]
  26. [26]
    Massive MIMO Networks: Spectral, Energy, and Hardware Efficiency
    ... interference to the channel in Figure 1.6, we obtain the discrete memoryless interference channel in Figure 1.7. The interference is not necessarily ...
  27. [27]
    Psychosis spectrum illnesses as disorders of prefrontal critical ...
    Sep 30, 2022 · Psychosis spectrum illnesses arise due to progressive changes in neural microcircuits that result in disturbances in experience-dependent plasticity.
  28. [28]
    Note on the Use of Principal Components in Regression
    Summary. The use of principal components in regression has received a lot of attention in the literature in the past few years, and the topic is now beginn.
  29. [29]
    Ridge Regression: Biased Estimation for Nonorthogonal Problems
    Ridge Regression: Biased Estimation for Nonorthogonal Problems. Arthur E. Hoerl University of Delaware and E. 1. du Pont de Nemours & Co. &. Robert W. Kennard ...
  30. [30]
    K-means clustering via principal component analysis
    K-means clustering via principal component analysis. Authors: Chris Ding. Chris Ding. Lawrence Berkeley National Laboratory, Berkeley, CA. View Profile. , ...
  31. [31]
    [1804.08450] Decorrelated Batch Normalization - arXiv
    Apr 23, 2018 · In this work, we propose Decorrelated Batch Normalization (DBN), which not just centers and scales activations but whitens them.
  32. [32]
    [PDF] LFSR-based Stream Ciphers - Centre Inria de Paris
    Thus, the Berlekamp-Massey algorithm determines the shortest LFSR that generates an infinite linear recurring sequence s from the knowledge of any 2Λ(s) ...
  33. [33]
    Decorrelation: A Theory for Block Cipher Security
    Abstract. Pseudorandomness is a classical model for the security of block ciphers. In this paper we propose convenient tools in order to study it in ...
  34. [34]
    [PDF] On the Masking-Friendly Designs for Post-Quantum Cryptography
    Masking can provide provably secure countermeasures against side-channel attacks. Nevertheless, due to the duplication of computations, the runtime of a masked ...Missing: decorrelation | Show results with:decorrelation
  35. [35]
    PORTFOLIO SELECTION* - Markowitz - 1952 - The Journal of Finance
    This paper is based on work done by the author while at the Cowles Commission for Research in Economics and with the financial assistance of the Social ...
  36. [36]
    [PDF] The Arbitrage Pricing Theory and Multifactor Models of Asset Returns*
    Sep 30, 1993 · The Arbitrage Pricing Theory (APT) of Ross (1976, 1977), and extensions of that theory, constitute an important branch of asset pricing theory ...Missing: decorrelation | Show results with:decorrelation
  37. [37]
  38. [38]
    [PDF] Evaluating "correlation breakdowns" during periods of market volatility
    ... correlation breakdowns may reflect time- varying volatility of financial markets rather than a change in the relationships between asset returns. Since the ...Missing: decorrelation | Show results with:decorrelation