Fact-checked by Grok 2 weeks ago

Correlation

In statistics, correlation refers to a measure of the strength and direction of the linear between two continuous variables, quantified by a that ranges from -1 to +1, where values near 1 indicate a strong positive association, near -1 a strong negative association, and near 0 no linear association. The concept originated in the late through the work of , who developed the idea of the to quantify consistent linear s between numeric variables, such as the between the heights of parents and their children in his studies of . later formalized the mathematical formula for the Pearson product-moment in 1895, establishing it as a cornerstone of modern statistical analysis. The most common form, Pearson's correlation coefficient (denoted as r for samples and ρ for populations), assumes normally distributed data and measures linear relationships, with positive values indicating that as one variable increases, the other tends to increase, and negative values showing the opposite. For non-normal or , alternatives like (ρ_s) are used, which assess monotonic relationships by ranking variables and are more robust to outliers. Other variants, such as Kendall's , evaluate ordinal associations based on concordant and discordant pairs, providing another measure of strength. Key properties of correlation coefficients include their dimensionless nature, (the correlation between X and Y equals that between Y and X), and independence from variable scaling, making them versatile for comparing relationships across datasets. However, , as associations may arise from factors, chance, or indirect influences, a limitation emphasized since its early development to prevent misinterpretation in fields like and social sciences. It also only captures linear or monotonic patterns, potentially underestimating nonlinear relationships, and is sensitive to outliers in the case of Pearson's method. Applications of correlation span numerous disciplines, including assessing variable associations in , , , and , often visualized through scatterplots to illustrate patterns before formal computation. In research, it serves as a preliminary tool for generation, informing or experimental design, but requires cautious interpretation alongside significance testing (e.g., p-values) to evaluate reliability.

Fundamentals of Correlation

Definition and Interpretation

Correlation is a statistical measure that quantifies the strength and direction of the linear relationship between two , standardized to range from -1 to +1. A of +1 represents perfect positive linear , where one variable increases proportionally with the other; 0 indicates no linear ; and -1 signifies perfect negative linear , where one variable decreases as the other increases. This measure focuses exclusively on linear dependencies and does not capture nonlinear relationships or imply causation. The term "correlation" was coined by British scientist in 1888, during his studies on and biological , to describe the tendency of traits to vary together. 's ideas were expanded by statistician in 1895, who developed a mathematical framework for quantifying this association, laying the foundation for modern correlational analysis. Interpreting the involves assessing both its sign (positive or negative direction) and magnitude (strength of the linear link). Values close to 0 suggest a weak association, while common guidelines classify |r| < 0.3 as weak, 0.3–0.7 as moderate, and >0.7 as strong; however, these thresholds are subjective and context-dependent, varying across fields like or . For instance, a correlation of 0.8 might indicate a robust linear relationship in social sciences but require cautious interpretation in physics due to differing expectations for effect sizes. Scatterplots provide the essential visual aid for interpreting correlation, plotting paired observations as points on a to reveal patterns. High positive correlation appears as points tightly clustered along an upward-sloping line, negative correlation along a downward-sloping line, and low correlation as a diffuse with no clear linear trend, enabling intuitive assessment of both strength and potential outliers.

Correlation and Independence

In probability theory, two random variables X and Y are defined as uncorrelated if their covariance is zero, that is, \operatorname{Cov}(X, Y) = 0, or equivalently, E[(X - \mu_X)(Y - \mu_Y)] = 0, where \mu_X = E[X] and \mu_Y = E[Y]. This condition implies that there is no linear relationship between the deviations of X and Y from their respective means. Independence of X and Y always implies that they are uncorrelated, since the joint expectation factors under independence: E[XY] = E[X]E[Y], leading to \operatorname{Cov}(X, Y) = 0. However, the converse does not hold in general: zero correlation does not imply statistical . A classic counterexample involves X uniformly distributed on [-1, 1] and Y = X^2. Here, E[X] = 0 and E[XY] = E[X^3] = 0 (since X^3 is an odd function over a symmetric interval), so \operatorname{Cov}(X, Y) = 0, confirming uncorrelatedness. Yet, X and Y are dependent, as the distribution of Y given X = 0 (where Y = 0) differs from the marginal distribution of Y, which is a scaled chi-squared-like on [0, 1]. An important exception occurs for jointly normal distributions. If X and Y follow a bivariate , then zero correlation (\rho_{X,Y} = 0) is equivalent to . This equivalence arises because the joint density factors into the product of marginal normals precisely when the off-diagonal term vanishes. Full details on this property are discussed in the context of bivariate s. In practice, tests of zero correlation, such as those based on the Pearson correlation coefficient, can assess independence only when the normality assumption holds; otherwise, they merely detect the absence of linear dependence, potentially missing nonlinear relationships.

Pearson's Product-Moment Correlation

Mathematical Definition

The Pearson product-moment correlation coefficient for two random variables X and Y, denoted \rho_{X,Y}, is defined as the covariance between X and Y divided by the product of their standard deviations: \rho_{X,Y} = \frac{\operatorname{Cov}(X,Y)}{\sigma_X \sigma_Y}, where \operatorname{Cov}(X,Y) = E[(X - \mu_X)(Y - \mu_Y)], \mu_X = E[X] and \mu_Y = E[Y] are the expected values, \sigma_X = \sqrt{\operatorname{Var}(X)}, and \sigma_Y = \sqrt{\operatorname{Var}(Y)}. This formulation, introduced by Karl Pearson in 1895, quantifies the strength and direction of the linear relationship between the variables, assuming finite variances. The coefficient can be derived from the covariance of standardized variables. Let Z_X = (X - \mu_X)/\sigma_X and Z_Y = (Y - \mu_Y)/\sigma_Y be the standardized versions of X and Y, each with mean zero and variance one. Then, \rho_{X,Y} = E[Z_X Z_Y] = \frac{E[(X - \mu_X)(Y - \mu_Y)]}{\sigma_X \sigma_Y}, which normalizes the covariance to lie within a bounded range, facilitating comparison across different scales. Geometrically, \rho_{X,Y} represents the cosine of the angle between the centered random vectors associated with X and Y in the L^2 space of square-integrable functions, where the inner product is the expectation: \rho_{X,Y} = \frac{E[(X - \mu_X)(Y - \mu_Y)]}{\sqrt{E[(X - \mu_X)^2] E[(Y - \mu_Y)^2]} = \cos \theta. This interpretation highlights the coefficient as a measure of directional alignment in a vector space framework. The value of \rho_{X,Y} satisfies -1 \leq \rho_{X,Y} \leq 1, a consequence of the Cauchy-Schwarz inequality applied to the inner product E[(X - \mu_X)(Y - \mu_Y)]. Equality holds at \rho_{X,Y} = 1 Y = aX + b for some a > 0 and constant b (perfect positive linear relationship), and at \rho_{X,Y} = -1 if a < 0 (perfect negative linear relationship).

Sample Correlation Coefficient

The sample correlation coefficient r, also known as Pearson's r, estimates the population correlation \rho from a finite sample of n paired observations (x_i, y_i) for i = 1, \dots, n. It is calculated as r = \frac{\sum_{i=1}^n (x_i - \bar{x})(y_i - \bar{y})}{\sqrt{\sum_{i=1}^n (x_i - \bar{x})^2 \sum_{i=1}^n (y_i - \bar{y})^2}}, where \bar{x} = \frac{1}{n} \sum_{i=1}^n x_i and \bar{y} = \frac{1}{n} \sum_{i=1}^n y_i are the sample means. This expression, originally formulated by , normalizes the sample covariance by the product of the sample standard deviations, yielding a dimensionless measure bounded between -1 and 1. Although r is consistent for \rho as n \to \infty, it serves as a biased estimator for finite n, systematically underestimating |\rho| when |\rho| > 0, with the expected bias approximately E(r) \approx \rho \left(1 - \frac{1 - \rho^2}{2n}\right). The magnitude of this downward bias increases with |\rho| and decreases with larger n, but it can distort inferences in small samples. To mitigate this bias and stabilize variance for inference, introduced the z-transformation, z = \frac{1}{2} \ln \left( \frac{1 + r}{1 - r} \right) = \artanh(r), which follows approximately a normal distribution with mean \artanh(\rho) and variance $1/(n-3) for n > 3. This transformation is particularly useful for confidence intervals and meta-analyses of correlations, as the near-normality holds even for moderate n. Computationally, the formula for r relies on deviations from the means, d_{x_i} = x_i - \bar{x} and d_{y_i} = y_i - \bar{y}, to center the data and eliminate the need for explicit mean subtraction in subsequent steps after initial calculation. Unlike the unbiased sample covariance, which divides the sum of cross-products by n-1 to account for degrees of freedom, the correlation coefficient avoids this adjustment in its core sums because the n-1 factors in the denominator's standard deviations cancel with the numerator's covariance term, preserving the scale-invariant property. This shortcut simplifies implementation in software and manual calculations, as raw sums of deviations suffice without Bessel's correction at the correlation stage. For hypothesis testing, particularly under the H_0: \rho = 0 (no linear association in the ), the sample r can be assessed using the t = r \sqrt{\frac{n-2}{1 - r^2}}, which follows a with n-2 when the data are bivariate . This test, derived from the of r under H_0, provides an exact for small to moderate n, outperforming normal approximations in finite samples. Rejection of H_0 at a chosen significance level indicates evidence of linear dependence, with the test's power increasing with n and |\rho|.

Properties and Assumptions

The Pearson product-moment correlation coefficient exhibits several key invariance properties that make it a robust measure of linear under certain transformations. Specifically, it remains unchanged under separate affine transformations of the variables, meaning that if the variables X and Y are replaced by aX + b and cY + d respectively, where a > 0, c > 0, and b, d are constants, the population correlation \rho and sample correlation r are . This and invariance ensures that the coefficient focuses solely on the relative positioning of points, independent of units or shifts. Regarding sampling properties, the sample r serves as a of the correlation \rho, converging in probability to \rho as the sample size n increases, provided the variables have finite variances. For large n, the of r is approximately after applying Fisher's z-transformation, z = \frac{1}{2} \ln \left( \frac{1 + r}{1 - r} \right), which stabilizes the variance and facilitates inference such as confidence intervals and tests. This asymptotic holds under the of finite fourth moments, though the raw distribution of r is skewed for small to moderate n. The relies on several fundamental assumptions for its and meaningful . It requires that both variables have finite second moments, i.e., E[X^2] < \infty and E[Y^2] < \infty, ensuring the variances \sigma_X^2 and \sigma_Y^2 are well-defined and positive. Additionally, for \rho (or r) to accurately quantify the strength of , the relationship between X and Y must be linear; the coefficient measures only linear dependence and assumes no substantial deviations from this form. If these assumptions are violated, such as when \sigma_X = 0 or \sigma_Y = 0 (indicating a variable), the coefficient is due to in its formula. A notable limitation arises from its focus on linearity: the Pearson correlation is insensitive to nonlinear relationships, even strong ones. For instance, if Y = X^2 for X uniformly distributed over [-1, 1], the variables are perfectly dependent, but \rho = 0 because the association is rather than . This highlights that a near-zero value does not imply , only the absence of linear correlation.

Illustrative Example

To illustrate the computation of Pearson's product-moment correlation coefficient, consider a hypothetical of (in cm) and weights (in kg) for five adults: heights are 160, 165, 170, 175, 180; corresponding weights are 50, 55, 60, 65, 70. This exhibits a perfect linear , as each increase of 5 cm in height corresponds to an increase of 5 kg in weight. The sample r is calculated using the formula r = \frac{\sum (x_i - \bar{x})(y_i - \bar{y})}{\sqrt{\sum (x_i - \bar{x})^2 \sum (y_i - \bar{y})^2}}, where x_i are the heights, y_i are the weights, and \bar{x}, \bar{y} are their respective means. First, compute the means: \bar{x} = 170 cm and \bar{y} = 60 kg. The deviations from the means, their products, and squared deviations are shown in the table below:
Height (x_i)Weight (y_i)x_i - \bar{x}y_i - \bar{y}Product(x_i - \bar{x})^2(y_i - \bar{y})^2
16050-10-10100100100
16555-5-5252525
1706000000
1756555252525
180701010100100100
Sums250250250
Thus, r = \frac{250}{\sqrt{250 \times 250}} = \frac{250}{250} = 1.0. A value of r = 1 indicates a perfect positive linear relationship, meaning that changes in perfectly predict changes in in this , with increasing proportionally as increases. In a scatterplot of these points, the data would form a straight line with a positive passing exactly through all five points, demonstrating no scatter around . To highlight sensitivity to deviations from perfect , consider perturbing the by changing the final from 70 kg to 65 kg (weights now: 50, 55, 60, 65, 65). The new is \bar{y} = 59 kg. Recalculating the deviations, products, and squared sums yields a product of 200 and a denominator of \sqrt{250 \times 170} \approx 206.16, so r \approx \frac{200}{206.16} \approx 0.97. This slight alteration reduces the correlation to a very strong but imperfect positive linear association, with the scatterplot now showing minor deviation from the line for the final point.

Rank-Based Correlation Coefficients

Spearman's Rank Correlation

Spearman's , denoted as \rho_s or r_s, is a nonparametric measure of the strength and direction of association between two ranked variables, introduced by in as a method to quantify the relationship between variables based on their order rather than magnitude. It is particularly suited for or when the underlying is unknown or non-normal, providing a robust alternative to measures. The coefficient is defined by the formula \rho_s = 1 - \frac{6 \sum_{i=1}^n d_i^2}{n(n^2 - 1)}, where d_i is the difference between the ranks of the i-th pair of observations from the two variables, and n is the number of observations, assuming no ties; ties are handled by assigning average ranks to tied values. This formula arises from applying the Pearson product-moment correlation to the ranked data, making \rho_s mathematically equivalent to the Pearson correlation coefficient computed on the ranks of the original variables. In interpretation, \rho_s ranges from -1 to +1, where a value of +1 indicates a perfect positive monotonic (as one increases, the other does so consistently in rank order), -1 indicates a perfect negative monotonic , and 0 suggests no monotonic association. Unlike Pearson's correlation, which assumes linearity and is sensitive to the scale of measurements, Spearman's \rho_s focuses solely on the monotonic ordering, capturing associations where the is steadily increasing or decreasing without requiring a straight-line pattern. Key advantages of Spearman's rank correlation include its robustness to outliers, as ranking diminishes the influence of extreme values on the overall measure, and its lack of reliance on distributional assumptions beyond the continuity of the variables for exact inference. This makes it ideal for real-world data with non-normal distributions or ordinal scales, such as psychological test scores or socioeconomic rankings, where Pearson's method might yield misleading results due to violations of its assumptions.

Kendall's Rank Correlation

Kendall's rank correlation coefficient, denoted as τ, is a non-parametric measure of the strength and direction of the association between two variables based on their ranks, introduced by Maurice Kendall in 1938. It assesses ordinal dependence by examining the relative ordering of pairs of observations, making it suitable for data that may not satisfy assumptions and particularly advantageous for small sample sizes where it provides consistent estimates of association. Unlike measures based on linear relationships, τ focuses on the number of agreeing (concordant) and disagreeing (discordant) pairs, offering a robust alternative when distributional forms are unknown or violated. This coefficient quantifies how well the rankings of one variable predict the rankings of another, with values interpreted as the probability of concordance minus the probability of discordance for randomly selected pairs. In the absence of tied ranks, Kendall's τ (τ_a) is computed as \tau = \frac{C - D}{\binom{n}{2}} = \frac{2(C - D)}{n(n-1)}, where n is the number of data points, C is the number of concordant pairs—for which the relative order of the s in both variables agrees (i.e., for i < j, \operatorname{[rank](/page/Rank)}(x_i) < \operatorname{[rank](/page/Rank)}(x_j) and \operatorname{[rank](/page/Rank)}(y_i) < \operatorname{[rank](/page/Rank)}(y_j), or both greater)—and D is the number of discordant pairs, where the orders disagree. This formulation normalizes the difference between concordant and discordant pairs by the total possible pairs, yielding a value between -1 and 1: τ = 1 for perfect monotonic agreement, τ = 0 for no association (equal concordant and discordant pairs), and τ = -1 for perfect reversal. The coefficient is invariant to monotonic transformations and handles effectively, though computation involves O(n^2) pairwise comparisons, which is feasible for modest n. When tied ranks occur, as is common in ordinal or discrete data, the standard τ_a underestimates the association by not accounting for incomparable pairs; instead, the adjusted τ_b is preferred: \tau_b = \frac{C - D}{\sqrt{(C + D + T_x)(C + D + T_y)}}, where T_x and T_y represent the number of tied pairs within the x and y variables, respectively. This denominator adjusts for the reduced number of decidable pairs, providing a bias-corrected estimate that maintains the range [-1, 1] and improves interpretability in tied scenarios. τ_b is especially valuable in fields like or , where Likert-scale or ranked responses often include ties, ensuring the measure reflects true ordinal relationships without undue penalization. Compared to Spearman's rank correlation ρ_s, which relies on squared rank differences, Kendall's τ emphasizes pairwise agreements and is asymptotically related to the underlying Pearson correlation ρ by \tau = \frac{2}{\pi} \arcsin(\rho) for large n under bivariate assumptions, while \rho_s \approx \frac{6}{\pi} \arcsin\left(\frac{\rho}{2}\right); since \rho_s \approx \rho under , \tau \approx \frac{2}{\pi} \arcsin(\rho_s). This relation highlights τ's lower variance in certain distributions, and it demonstrates greater efficiency when ties are present, as Spearman's method averages ties by assigning mid-ranks, potentially diluting the signal in sparse data. Both measure monotonicity, but τ's pair-wise focus makes it less sensitive to ranks and more appropriate for confirming in non-parametric tests via its exact for small n.

Alternative Measures of Association

Partial Correlation

measures the degree of between two random after removing the linear effects of one or more additional , known as controlling variables or covariates. For the population partial correlation coefficient between variables X and Y controlling for a third Z, it is defined as \rho_{XY \cdot Z} = \frac{\rho_{XY} - \rho_{XZ} \rho_{YZ}}{\sqrt{(1 - \rho_{XZ}^2)(1 - \rho_{YZ}^2)}} where \rho_{XY}, \rho_{XZ}, and \rho_{YZ} are the respective bivariate Pearson correlation coefficients. This formula adjusts the bivariate correlation \rho_{XY} by subtracting the product of the correlations involving Z and normalizing by the residual variances after accounting for Z. The sample partial correlation coefficient r_{XY \cdot Z} is computed using the analogous formula with sample correlations r_{XY}, r_{XZ}, and r_{YZ} in place of the population parameters. Interpretationally, \rho_{XY \cdot Z} (or r_{XY \cdot Z}) quantifies the direct linear relationship between X and Y that is independent of Z, equivalent to the Pearson correlation between the residuals of X and Y after regressing each on Z. Values range from -1 to 1, where 0 indicates no linear association after controlling for Z, and it is particularly useful in multiple regression analysis to evaluate the unique contribution of one predictor to the outcome while holding others constant. When controlling for multiple variables, partial correlations can be calculated recursively by iteratively applying the single-control formula, starting with one covariate and proceeding to the next using the updated correlations. Alternatively, for a set of p variables with correlation matrix R, the partial correlation between variables i and j controlling for the remaining p-2 variables is given by \rho_{ij \cdot \text{rest}} = -\frac{(R^{-1})_{ij}}{\sqrt{(R^{-1})_{ii} (R^{-1})_{jj}}}, where R^{-1} denotes the inverse of R. This matrix-based approach efficiently yields the full partial correlation matrix and is equivalent to correlating residuals from regressing X_i and X_j on all other variables. In applications, partial correlation serves as a preliminary tool in causal inference by isolating direct associations, notably in path analysis as developed by Sewall Wright in his 1921 paper "Correlation and Causation," where it decomposes observed correlations into direct and indirect effects along specified causal paths in systems like genetic or agricultural models. It is widely employed in fields such as psychology, economics, and epidemiology to control for confounding variables and assess conditional dependencies in observational data.

Categorical and Binary Measures

When dealing with categorical or , standard measures like Pearson's are not directly applicable due to the nature of the variables. Instead, specialized analogs extend the concept of linear to these data types, often by treating outcomes as indicators or assuming underlying continuous latent variables. These measures quantify the strength and direction of in tables or mixed variable types, maintaining interpretability similar to Pearson's , which ranges from -1 to 1. The (φ), also known as the mean square contingency coefficient, measures association between two variables in a 2×2 . For a table with cell counts a, b, c, d where rows and columns represent the binary categories, it is defined as: \phi = \frac{ad - bc}{\sqrt{(a+b)(c+d)(a+c)(b+d)}} This formula arises from applying Pearson's product-moment correlation to the binary indicators of the variables, making φ mathematically equivalent to Pearson's r in this setting. Introduced by , φ ranges from -1 to 1, with values near 0 indicating independence and |φ| = 1 signifying perfect association. It is particularly useful in fields like and for analyzing dichotomous traits, such as presence/absence outcomes. For associations between a binary variable and a continuous variable, the (r_pb) serves as an adaptation. It is computed as: r_{pb} = \frac{M_1 - M_0}{s} \sqrt{p(1-p)} where M_1 and M_0 are the means of the continuous for the two binary groups, s is the standard deviation of the continuous across all observations, and p is the proportion in the first binary group. This measure, a special case of Pearson's r when the binary is coded as 0 and 1, assesses how the continuous differs across the binary categories. The term and explicit formula were formalized by Richardson and Stalnaker, though the underlying derivation traces to Pearson's framework for mixed-scale correlations. Values range from -1 to 1, with significance tested via t-statistics analogous to those for Pearson's r. Tetrachoric and polychoric correlations address limitations of φ and r_pb by assuming the observed or reflect underlying bivariate continuous variables, dichotomized or categorized by . The tetrachoric correlation estimates the Pearson correlation of these latent continuous variables for two observables, derived via maximum likelihood from the 2×2 table proportions under the . Pearson introduced this approach to infer correlations for non-quantifiable characters, such as qualitative traits in evolutionary studies, where direct measurement is impossible. Computation involves integrating the bivariate density, often approximated for practical use. For with more than two categories, the generalizes this, estimating the latent Pearson correlation via maximum likelihood on the full . Developed by and , it accommodates multiple ordered categories, common in surveys or Likert scales, and assumes equal spacing in the unless adjusted. Both measures range from -1 to 1 but can be sensitive to violations of the or , leading to biased estimates if the latent variables are not approximately . For contingency tables larger than 2×2 involving multiple categorical levels, extends the as a normalized measure of association. Defined as: V = \frac{\phi}{\sqrt{\min(k-1, r-1)}} where φ is the from the chi-squared statistic (φ = √(χ²/n)), k and r are the number of columns and rows, and n is the total sample size, V ranges from 0 to 1, with higher values indicating stronger association. Proposed by Harald Cramér, it provides a scale-independent generalization of φ, useful for nominal data in and , and is asymptotically equivalent to the under certain conditions. Unlike φ, V adjusts for table dimensions to ensure comparability across different table sizes.

Non-Linear Dependence Measures

While the effectively captures linear relationships between variables, it fails to detect many forms of non-linear dependence, such as quadratic or periodic associations. Non-linear dependence measures address this limitation by quantifying general associations without assuming , often achieving zero value the variables are . These measures are particularly valuable in and high-dimensional settings where non-linear patterns predominate. One prominent measure is , introduced by Székely, Rizzo, and Bakirov. It is defined for random vectors \mathbf{X} and \mathbf{Y} in spaces as R_d(\mathbf{X}, \mathbf{Y}) = \frac{V_d(\mathbf{X}, \mathbf{Y})}{\sqrt{V_d(\mathbf{X}) V_d(\mathbf{Y})}}, where V_d(\mathbf{X}, \mathbf{Y}) is the distance covariance, a non-negative quantity based on the expected distances between observations: specifically, V_d(\mathbf{X}, \mathbf{Y})^2 = \mathbb{E}[\| \mathbf{X} - \mathbf{X}' \| \| \mathbf{Y} - \mathbf{Y}' \|] + \mathbb{E}[\| \mathbf{X} - \mathbf{X}' \|] \mathbb{E}[\| \mathbf{Y} - \mathbf{Y}' \|] - 2 \mathbb{E}[\| \mathbf{X} - \mathbf{X}' \| \| \mathbf{Y} - \mathbf{Y}'' \|], with \mathbf{X}', \mathbf{X}'', \mathbf{Y}', \mathbf{Y}'' copies. detects any form of dependence, linear or non-linear, and equals zero if and only if \mathbf{X} and \mathbf{Y} are . Its sample version is consistent under the of , making it suitable for testing. The (MIC), proposed by Reshef et al., provides another approach to capturing diverse non-linear associations. Derived from the noise equivalence (MINE) framework, MIC approximates the normalized between continuous variables X and Y by partitioning the data into and selecting the partitioning that maximizes the score. Formally, for a G with B bins along each , the characteristic matrix entry is I(X_\perp Y | G) = \frac{I(X,Y | G)}{\log \min \{ n_x(G), n_y(G) \}}, where I(X,Y | G) is the under G, and MIC is the supremum over grids with bounded column resolution. This grid-based method excels at identifying functional relationships of varying complexity, such as exponentials or sinusoids, while being equitable across association strengths. However, MIC has faced criticism for not fully satisfying its claimed equitability property in detecting associations of equal strength, as demonstrated by subsequent theoretical and empirical analyses. Hoeffding's D, developed by Hoeffding, offers a rank-based measure of general dependence for continuous random variables. It is defined as an over the of the variables: D = 12 \int_0^1 \int_0^1 [C(u, v) - u v]^2 \, du \, dv, where C is the copula function capturing the joint dependence structure. This formulation quantifies deviations from across the entire joint , detecting non-linear as well as linear dependencies, and equals zero the variables are . The empirical facilitates non-parametric testing. Distance correlation has been shown to be particularly effective in high-dimensional settings, with extensions proving its consistency for testing independence among multiple vectors.

Sensitivity and Robustness Issues

Effect of Data Distribution

The Pearson correlation coefficient assumes bivariate normality for optimal properties, but deviations such as skewness in the data distribution can introduce bias in the estimate of the population correlation ρ. In particular, positive skewness tends to inflate the absolute value of the sample correlation r for positive associations, as the asymmetric tail pulls extreme values in a way that exaggerates the linear appearance. Simulation studies demonstrate this effect: for highly skewed distributions (e.g., with skewness of 2.8), the bias in r can reach up to +0.14 relative to ρ, especially in small samples (n=10–40), with similar inflation observed in heavy-tailed distributions. Heteroscedasticity, or varying along the relationship, further complicates interpretation by potentially attenuating the magnitude of r, as the increasing spread dilutes the tight linear fit. This manifests in fan-shaped scatterplots, where residuals widen with increasing predictor values, leading to an underestimation of the true strength despite no in the point estimate under certain models. For instance, when variance grows proportionally with the level of the variables, the overall in r incorporates this heterogeneity, reducing its value compared to a homoscedastic scenario. Such effects primarily undermine , invalidating standard t-tests for ρ=0, as heteroscedasticity can mimic or mask significant correlations. A notable arises with binned or aggregated , known as the , where correlations computed at the group level exceed those at the individual level due to spatial or grouping effects. Robinson illustrated this using 1930 U.S. on foreign-born populations and illiteracy rates: the ecological correlation was -0.62 across states, suggesting a strong negative link, but the individual-level correlation was +0.12, demonstrating how aggregation can not only inflate but also reverse the direction of associations. To mitigate these distributional effects, data transformations like the Box-Cox power transformation can normalize skewed data and stabilize variance, restoring the validity of Pearson's r by approximating the normality assumption. Alternatively, rank-based methods provide robustness without transformation, though they address non-linearity as well.

Impact of Outliers and Robust Alternatives

Pearson's correlation coefficient is particularly sensitive to s, as these extreme values can disproportionately influence the linear association estimate, leading to misleading results. A classic illustration is , comprising four bivariate datasets that yield nearly identical Pearson correlation coefficients of approximately 0.816 and the same least-squares regression line, despite scatter plots revealing stark differences, including nonlinear patterns and influential outliers in some cases. Even a single outlier can drastically alter the correlation's magnitude or reverse its sign, transforming an apparent strong positive relationship into a negative one or vice versa. To counteract this vulnerability, several robust alternatives to Pearson's correlation have been developed. The Winsorized correlation coefficient enhances robustness by first trimming or capping the most extreme observations (typically the top and bottom percentages) in each variable, then applying the standard Pearson formula to the modified data; this approach reduces the impact of outliers while preserving much of the linear structure. Similarly, , which transforms data to ranks before computation, exhibits bounded influence functions and thus greater resistance to outliers compared to Pearson's method. Median-based estimators offer additional protection; for example, the Hodges-Lehmann correlation coefficient derives from the of pairwise estimates, providing a nonparametric robust measure suitable for bivariate associations contaminated by extremes. Outliers in correlation analysis can be detected using diagnostic tools borrowed from , given the close equivalence between Pearson's and the in simple . Leverage values identify points distant from the in the predictor space, potentially exerting undue pull on the fit, while Cook's distance quantifies an observation's overall influence by assessing changes in predicted values when that point is excluded. In multivariate contexts involving correlation matrices, the Minimum (MCD) estimator, introduced by Rousseeuw, achieves high breakdown robustness (up to nearly 50% contamination) by selecting the subset of h observations yielding the smallest , from which a cleaned is derived; this remains a standard in 2025 software implementations like R's robustbase package.

Correlation Matrices

Construction and Properties

A correlation matrix R = [\rho_{ij}] for p random variables is constructed such that its diagonal elements are all 1, reflecting perfect correlation of each variable with itself, while the off-diagonal elements \rho_{ij} (for i \neq j) represent the pairwise correlation coefficients between variables i and j. The matrix is inherently symmetric, as \rho_{ij} = \rho_{ji}, due to the symmetry of the underlying correlation measure. In the population setting, these pairwise correlations are typically Pearson correlations, assuming joint normality or linearity. For a sample of n observations on p variables, the sample correlation matrix is derived from the sample covariance matrix S, where S_{ij} is the sample covariance between variables i and j. Specifically, let D be the diagonal matrix with the diagonal elements of S (i.e., the sample variances), then the sample correlation matrix is R = D^{-1/2} S D^{-1/2}, which standardizes the covariances by the square roots of the variances to yield correlations in [-1, 1]. Correlation matrices possess several fundamental algebraic properties that ensure their validity as representations of linear dependence structures. Foremost, every correlation matrix is positive semi-definite (PSD), meaning all its eigenvalues are non-negative, which follows from the fact that it can be viewed as the of standardized variables (with unit variances), and the \mathbf{a}^T R \mathbf{a} = \mathrm{Var}(\mathbf{a}^T \mathbf{Z}) \geq 0 for any \mathbf{a} and standardized variables \mathbf{Z}. Additionally, each off-diagonal entry satisfies |\rho_{ij}| \leq 1, as correlations measure the strength of linear association and cannot exceed perfect positive or negative alignment. These properties impose constraints on possible values; for instance, in the case of three variables, the correlation \rho_{12} must satisfy |\rho_{13} \rho_{23} - \sqrt{(1 - \rho_{13}^2)(1 - \rho_{23}^2)}| \leq \rho_{12} \leq |\rho_{13} \rho_{23}| + \sqrt{(1 - \rho_{13}^2)(1 - \rho_{23}^2)} to ensure the 3×3 remains PSD, analogous to a triangle inequality in the geometric interpretation of correlations via angles. Further properties arise from the PSD nature: the determinant satisfies \det(R) \geq 0, with equality only if the variables are linearly dependent, and the equals the p, since \mathrm{trace}(R) = \sum_{i=1}^p \rho_{ii} = p. The eigenvalues \lambda_1 \geq \lambda_2 \geq \cdots \geq \lambda_p \geq 0 sum to p and each lies in [0, p], providing a measure of the total and distributed variance in the standardized space. In (PCA), the eigen-decomposition of the correlation matrix plays a central role in . The decomposition R = V \Lambda V^T, where V contains the eigenvectors (principal components) and \Lambda is the of eigenvalues, identifies orthogonal directions of maximum variance; retaining the top k < p components with the largest eigenvalues projects the data onto a lower-dimensional subspace that captures most of the variability, facilitating visualization and without significant information loss.

Nearest Valid Correlation Matrix

In practice, sample correlation matrices derived from real-world often fail to be positive semidefinite () due to issues such as missing observations or errors in estimation, rendering them invalid for applications requiring valid correlation structures. This problem necessitates methods to project such matrices onto the set of valid correlation matrices, which are symmetric, PSD, and have unit diagonal entries. A seminal approach to this is the alternating projections method proposed by Higham in , which minimizes the weighted Frobenius norm distance \|R - A\|_W = \|W^{1/2}(R - A)W^{1/2}\|_F subject to A being PSD and having unit diagonal entries, where W is a positive definite weighting matrix. The algorithm iteratively projects the matrix onto the PSD cone (using spectral decomposition to set negative eigenvalues to zero) and then onto the unit diagonal constraint (by normalizing off-diagonal elements in each row/column), incorporating Dykstra's correction for improved convergence. This method guarantees a unique solution due to the convexity of the feasible set and converges monotonically to the nearest valid correlation matrix. These projection techniques find key applications in imputing missing correlations within incomplete matrices, where partial estimates are adjusted to ensure overall validity. They are also essential in simulations for generating realistic multivariate scenarios, particularly in for modeling portfolio where invalid matrices could lead to erroneous estimates. As of 2025, Higham's method remains the standard for computing nearest correlation matrices, with implementations available in numerical libraries and extensions in statistical software such as statsmodels in , which support factor-structured approximations for efficiency in higher dimensions.

Correlation in Stochastic Processes

Uncorrelated Stochastic Processes

In the of processes, two processes \{X_t\}_{t \in T} and \{Y_t\}_{t \in T}, defined on the same , are said to be uncorrelated if the between any pair of realizations at times s, t \in T is zero, that is, \operatorname{Cov}(X_s, Y_t) = 0 for all s, t. This condition generalizes the notion of uncorrelated random variables to time-indexed families, implying no linear dependence between the processes at any temporal points, though higher-order dependencies may persist. The function, defined as C_{XY}(s,t) = \mathbb{E}[(X_s - \mu_{X_s})(Y_t - \mu_{Y_t})], vanishes entirely under this definition, akin to in the cross-domain. For jointly wide-sense stationary processes, where means and covariances depend only on time differences, uncorrelatedness simplifies to the cross-covariance function \gamma_{XY}(h) = \operatorname{Cov}(X_t, Y_{t+h}) = 0 for all lags h \in \mathbb{Z} (or \mathbb{R} for continuous time). This lag-independent zero cross-covariance ensures that the processes exhibit no linear temporal association at any displacement, facilitating decompositions in time series analysis such as filtering or prediction. Stationarity strengthens the interpretability, as the property holds uniformly across the timeline without varying with absolute positions. A key implication arises in processes with uncorrelated increments, such as the (standard ), where non-overlapping increments W_t - W_s and W_v - W_u (for s < t \leq u < v) satisfy \operatorname{Cov}(W_t - W_s, W_v - W_u) = 0, reflecting the process's memoryless linear structure. However, while increments in the are both uncorrelated and independent due to its Gaussian nature, uncorrelatedness alone does not guarantee independence in general stochastic processes, allowing for nonlinear dependencies that preserve zero . An illustrative example involves two independent Poisson processes \{N_t^{(1)}\} and \{N_t^{(2)}\} with rates \lambda_1 and \lambda_2, respectively; their increments on disjoint intervals are independent, hence uncorrelated, with \operatorname{Cov}(N_t^{(1)} - N_s^{(1)}, N_v^{(2)} - N_u^{(2)}) = 0 for non-overlapping (s,t] and (u,v]. This property underscores how uncorrelated counting processes model superimposed event streams without linear interaction, common in and reliability analysis.

Independence in Stochastic Processes

In stochastic processes, uncorrelated increments or components do not necessarily imply statistical , as dependence can manifest through higher-order moments or nonlinear structures. A prominent example is the (ARCH) model, where the error terms are serially uncorrelated but exhibit dependence in their conditional variances, leading to in financial . A key sufficient condition for independence arises when the processes are jointly Gaussian, meaning that any finite collection of their values follows a . In this case, zero between the processes implies full statistical , due to the specific of the multivariate normal that factorizes under zero correlation. This extends the bivariate normal case to dynamic settings, such as Gaussian Markov random fields, where uncorrelated fields are . For Markov processes, which are defined by the property that the future state is of the past given the present state, uncorrelatedness of the future with the past conditional on the present aligns with this independence under joint Gaussianity. Specifically, in Gaussian Markov processes, conditional uncorrelation suffices to establish , facilitating applications in spatial and temporal modeling. Distinguishing weak white noise (uncorrelated but possibly dependent) from strong white noise (independent and identically distributed) requires specialized testing beyond autocorrelation checks. , via estimation of the operator, can verify weak white noise by confirming flat spectra across lags, while the process in a allows detection of higher-order dependencies through kernel-based tests for strong white noise. These methods are particularly useful in functional , where traditional portmanteau tests may fail. Historically, the risks of mistaking correlation for meaningful dependence in time series were highlighted by Yule's analysis of spurious correlations, where integrated random walks exhibited high correlations despite lacking causal links, underscoring the need for independence assessments in dynamic data.

Common Misconceptions

The principle that "" serves as a fundamental caution in statistical analysis, emphasizing that an observed association between two variables does not necessarily indicate that one causes the other. This concept emerged in the late amid early developments in correlation theory and was notably articulated by British statistician , who highlighted in his 1911 work that mere correlation, even if statistically significant, fails to establish causal direction without additional evidence. The maxim underscores the risks of misinterpretation in fields like , , and social sciences, where overlooking this distinction can lead to flawed policies or scientific claims. A primary behind this is , where a third influences both observed variables, creating an illusory link. For instance, the strong positive correlation between monthly sales and incidents is not due to ice cream consumption causing drownings, but rather both being driven by the confounding factor of warmer summer temperatures, which boost outdoor activities and ice cream demand. Such confounders can inflate or mask true associations, as seen in observational studies where unmeasured factors like or environmental conditions systematically affect outcomes. Spurious correlations represent another pitfall, occurring when unrelated variables coincidentally align due to chance or unrelated trends, yielding high correlation coefficients without any causal or confounding basis. A striking example is the 99.26% correlation (r = 0.9926) between margarine consumption and rates in from 2000 to 2009, a that defies logical and highlights how across disparate datasets can produce misleading patterns. These artifacts often arise in large datasets with many variables, emphasizing the need for theoretical grounding to avoid overinterpreting statistical noise. To mitigate the correlation-causation fallacy, researchers employ experimental and quasi-experimental designs that isolate causal effects. Randomized controlled trials (RCTs) achieve this by randomly assigning participants to or groups, thereby balancing confounders and enabling under ideal conditions. In non-experimental settings, instrumental variables—external factors that affect the but not the outcome directly—help address biases from omitted variables or reverse causation. For time-series data, tests evaluate whether one variable's past values improve predictions of another's future values, providing evidence of predictive precedence as a proxy for causation, though not definitive proof.

Limitations of Linear Correlation

The , denoted as [r](/page/R), quantifies the strength and direction of a linear between two but fails to capture non-linear dependencies. For example, a U-shaped —where one increases or decreases non-monotonically with the other—can yield [r](/page/R) = 0, suggesting no despite evident dependence, as the positive and negative deviations cancel out. This limitation underscores the need for data visualization or alternative measures to detect such patterns, as relying solely on [r](/page/R) may overlook meaningful . Small sample sizes exacerbate the variability of correlation estimates, often leading to inflated or unstable r values that do not reliably reflect parameters. Research indicates that sample sizes below approximately 250 are typically insufficient for stable estimates in common scenarios, with convergence to the true correlation \rho requiring larger n depending on and desired precision. Additionally, high variability within the —exceeding 10% of the variable's range—can reduce the shared variance captured by r by 50% or more, even assuming a perfect underlying , thus artifactually weakening observed correlations. Cherry-picking subsets of can further mislead, as illustrated by , where positive correlations in subgroups (e.g., treatment success rates differing by gender) reverse to negative or null in the aggregate due to uneven group weighting. Computing correlations across multiple variable pairs without adjustment inflates the (FWER), the probability of at least one false positive discovery. The Bonferroni correction addresses this by dividing the significance level \alpha by the number of tests m, rejecting null hypotheses only if p < \alpha / m, thereby controlling FWER at \alpha. However, this conservative approach reduces statistical power, particularly for large m, highlighting the in exploratory analyses scanning numerous pairs. In the social sciences, linear correlation has been historically misused to infer spurious hereditary links, notably in 20th-century debates. , who coined the term "correlation," and applied r to family data on traits like height and intelligence, interpreting coefficients as evidence of genetic while ignoring environmental confounders, which fueled discriminatory policies on and sterilization. Such applications, as in Pearson's studies of Jewish children's intelligence (1925–1928), exemplify how uncritical reliance on correlation perpetuated ethical harms in pseudoscientific contexts.

Correlation in Bivariate Normal Distributions

Joint Properties

The joint probability density function (PDF) of two random variables X and Y following a bivariate normal distribution with means \mu_X and \mu_Y, standard deviations \sigma_X and \sigma_Y, and correlation coefficient \rho is given by f(x,y) = \frac{1}{2\pi \sigma_X \sigma_Y \sqrt{1-\rho^2}} \exp\left\{ -\frac{1}{2(1-\rho^2)} \left[ \frac{(x-\mu_X)^2}{\sigma_X^2} + \frac{(y-\mu_Y)^2}{\sigma_Y^2} - 2\rho \frac{(x-\mu_X)(y-\mu_Y)}{\sigma_X \sigma_Y} \right] \right\}, for -\infty < x, y < \infty and -1 < \rho < 1. This form highlights the role of \rho in the cross-term, which captures the linear dependence between X and Y. The equal-density contours of the bivariate normal distribution are ellipses centered at (\mu_X, \mu_Y), with their shape and orientation determined by \sigma_X, \sigma_Y, and \rho. The correlation \rho tilts these elliptical contours: positive \rho orients the major axis upward to the right, while negative \rho orients it upward to the left; when \rho = 0, the contours align with the coordinate axes, reducing to circles if \sigma_X = \sigma_Y. Regardless of the value of \rho, the marginal distributions of X and Y are univariate , with X \sim N(\mu_X, \sigma_X^2) and Y \sim N(\mu_Y, \sigma_Y^2). This property ensures that the bivariate normal preserves in each variable individually. Given specified means \mu_X and \mu_Y, variances \sigma_X^2 and \sigma_Y^2, and correlation \rho, there exists a unique bivariate normal joint distribution for the pair (X, Y). When \rho = 0, this joint distribution factors into the product of the marginals, implying between X and Y.

Conditional Interpretation

In the bivariate normal distribution, the conditional distribution of one variable given the value of the other is also normal, a property that facilitates regression and prediction tasks. Specifically, if (X, Y) follows a bivariate normal distribution with means \mu_X and \mu_Y, standard deviations \sigma_X and \sigma_Y, and correlation coefficient \rho, then the conditional distribution of Y given X = x is normal: Y \mid X = x \sim \mathcal{N}\left( \mu_Y + \rho \frac{\sigma_Y}{\sigma_X} (x - \mu_X), \ \sigma_Y^2 (1 - \rho^2) \right). This result is derived from the joint probability density function of the bivariate normal by integrating out the conditioning variable. The conditional mean \mu_{Y \mid X = x} = \mu_Y + \rho \frac{\sigma_Y}{\sigma_X} (x - \mu_X) shifts linearly with respect to x, where the slope is modulated by the correlation \rho; positive \rho implies that deviations of x above \mu_X pull the expected y upward, and vice versa for negative \rho. The coefficient \beta_{Y \mid X} = \rho \frac{\sigma_Y}{\sigma_X} quantifies this linear relationship, representing the change in the conditional mean of Y per change in X, directly incorporating the strength and direction of the correlation. The conditional variance \sigma_Y^2 (1 - \rho^2) decreases as |\rho| increases, reflecting reduced uncertainty in Y when X provides more information about it through stronger dependence; for \rho = 0, the variance equals the marginal variance of Y, indicating . This variance governs s, which narrow with higher |\rho|; for instance, the width of a 95% for Y \mid X = x is proportional to \sqrt{1 - \rho^2}, making predictions more precise as correlation strengthens. In the special case where \rho = \pm 1, the is zero, resulting in a where Y \mid X = x is deterministically equal to the line \mu_Y + \rho \frac{\sigma_Y}{\sigma_X} (x - \mu_X), implying perfect linear dependence between X and Y.

References

  1. [1]
    A guide to appropriate use of Correlation coefficient in medical ... - NIH
    In statistical terms, correlation is a method of assessing a possible two-way linear association between two continuous variables. Correlation is measured by a ...
  2. [2]
    Francis Galton (1822-1911)
    To quantify the consistency of a linear relationship between two numeric variables, Galton developed the concept of the correlation coefficient. His work was ...
  3. [3]
    [PDF] Thirteen Ways to Look at the Correlation Coefficient Joseph Lee ...
    Feb 19, 2008 · Pearson first developed the math- ematical formula for this important measure in 1895: This, or some simple algebraic variant, is the usual for-.
  4. [4]
    Correlation: Pearson, Spearman, and Kendall's tau | UVA Library
    May 27, 2025 · Correlation is a widely used method that helps us explore how two variables change together, providing insight into whether a relationship ...
  5. [5]
    7 Correlation: What It Really Means - STAT ONLINE
    We describe the direction of the relationship as positive or negative. A positive relationship means that as the value of the explanatory variable increases, ...
  6. [6]
    Correlation - Statistics Resources - LibGuides at National University
    Oct 27, 2025 · The correlation analysis is used to measure the direction and relationship between two variables. It's important to note that correlation does not equal ...
  7. [7]
    Conducting correlation analysis: important limitations and pitfalls - NIH
    The correlation coefficient is a statistical measure often used in studies to show an association between variables or to look at the agreement between two ...
  8. [8]
    Correlation Coefficients: Appropriate Use and Interpretation
    Correlation in the broadest sense is a measure of an association between variables. In correlated data, the change in the magnitude of 1 variable is ...
  9. [9]
    User's guide to correlation coefficients - PMC - NIH
    A negative r means that the variables are inversely related. The strength of the correlation increases both from 0 to +1, and 0 to −1. When writing a manuscript ...
  10. [10]
    Thirteen Ways to Look at the Correlation Coefficient - jstor
    1895 Karl Pearson, British statistician Defined the (Galton-) Pearson product-moment correlation coefficient. 1920 Karl Pearson Wrote "Notes on the History ...
  11. [11]
    Scatterplots and correlation review (article) | Khan Academy
    A scatterplot is a type of data display that shows the relationship between two numerical variables. Each member of the dataset gets plotted as a point.
  12. [12]
    [PDF] Probability and Random Variables, Lecture 25
    Uncorrelated just means E [(X − E [X ])(Y − E [Y ])] = 0,. i.e., the outcomes where (X − E [X ])(Y − E [Y ]) is positive. (the upper right and lower left ...
  13. [13]
    [PDF] Covariance and Correlation
    Jul 28, 2017 · ρ(X,Y) = 0. absence of linear relationship. If ρ(X,Y) = 0 we say that X and Y are “uncorrelated.” If two variables are independent, then their ...
  14. [14]
    [PDF] Reminder No. 1: Uncorrelated vs. Independent
    Feb 27, 2013 · Uncorrelated variables have zero correlation, while independent variables have joint probability as product of marginal distributions. ...
  15. [15]
    [PDF] Independence, Covariance, and Correlation
    Suppose that the random variable X is uniform on the interval [-1, 1]. Let Y = X2. Then X and Y are uncorrelated, but not independent. (To see that X and Y are ...
  16. [16]
    21.2 - Joint P.D.F. of X and Y | STAT 414
    If and have a bivariate normal distribution with correlation coefficient ρ X Y , then and are independent if and only if ρ X Y = 0 . That "if and only if" ...
  17. [17]
    [PDF] Multivariate normal distributions
    which implies independence of U and V. That is, for random variables with a bivariate nor- mal distribution, zero correlation is equivalent to independence.
  18. [18]
    [PDF] Lecture 11: Correlation and independence
    If Cov(X,Y) = 0, then we say that X and Y are uncorrelated. The correlation is a standardized value of the covariance. Theorem 4.5. 6. If X and Y are random ...
  19. [19]
    18.1 - Pearson Correlation Coefficient | STAT 509
    ... coefficient, r p , is the point estimate of the population Pearson correlation coefficient. ρ p = σ X Y σ X X σ Y Y. The Pearson correlation coefficient ...
  20. [20]
    [PDF] Topic #10: Correlation
    The best known is the Pearson product-moment correlation coefficient, which is obtained by dividing the covariance of the two variables by the product of their ...<|control11|><|separator|>
  21. [21]
    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.
  22. [22]
    [PDF] Frequency Distribution of the Values of the Correlation Coefficient in ...
    Dec 7, 2005 · Frequency Distribution of the Values of the Correlation Coefficient in. Samples from an Indefinitely Large Population. R. A. Fisher. Biometrika, ...
  23. [23]
    [PDF] Probabilistic Inferences for the Sample Pearson Product Moment ...
    Nov 1, 2011 · The Pearson's correlation coefficient is a measure of the ... unbiased, consistent estimator controlling. Type I error. 2. Due to ...
  24. [24]
    What is the correlation if the standard deviation of one variable is 0?
    Nov 14, 2011 · No variable that has standard deviation 0 could possibly be correlated with another (non-constant) variable.Pearson correlation of data sets with possibly zero standard deviation?Correlation with a constant - Cross Validated - Stack ExchangeMore results from stats.stackexchange.com
  25. [25]
    [PDF] Forward Selection via Distance Correlation - Rose-Hulman Scholar
    May 22, 2019 · This is a plot of X = Unif(−1,1) and Y = X2. This is a classic example of where Pearson's correlation is zero, but distance correlation is non- ...
  26. [26]
    Pearson Correlation Coefficient (r) | Guide & Examples - Scribbr
    May 13, 2022 · Calculating the Pearson correlation coefficient · Step 1: Calculate the sums of x and y · Step 2: Calculate x2 and y2 and their sums · Step 3: ...
  27. [27]
    Correlation (Coefficient, Partial, and Spearman Rank) and ... - NCBI
    May 25, 2024 · Partial correlation (ρ): Partial correlation measures the linear relationship between 2 continuous variables while controlling for other ...
  28. [28]
    Partial and Semipartial Correlation
    The formula to compute the partial r from correlations is. In our example, (1 = GPA, 2 = CLEP, 3 = SAT)Missing: definition | Show results with:definition
  29. [29]
    6.3 - Testing for Partial Correlation | STAT 505
    Partial correlation testing involves testing if it equals zero, using a t-statistic with n-2-c degrees of freedom, and comparing it to a critical value.
  30. [30]
    [PDF] Partial Correlation
    A second-order partial correlation is a measure of the relationship between X1 and X2 while controlling for two other variables: X3 and X4. This is noted as.Missing: definition | Show results with:definition
  31. [31]
    [PDF] Multiple correlation and multiple regression - The Personality Project
    To find the matrix of partial correlations, R* where the effect of a number of the Z variables been removed, just express equation 5.9 in matrix form. First ...
  32. [32]
    Mathematical contributions to the theory of evolution. VIII ... - Journals
    This memoir, which was read in November of last year, presented the novel feature of determining correlation between characters which were not capable à priori ...
  33. [33]
    ON POLYCHORIC COEFFICIENTS OF CORRELATION | Biometrika
    KARL PEARSON, F.R.S., EGON S. PEARSON; ON POLYCHORIC COEFFICIENTS OF CORRELATION, Biometrika, Volume 14, Issue 1-2, 1 July 1922, Pages 127–156, https://doi.<|control11|><|separator|>
  34. [34]
    Measuring and testing dependence by correlation of distances
    December 2007 Measuring and testing dependence by correlation of distances. Gábor J. Székely, Maria L. Rizzo, Nail K. Bakirov · DOWNLOAD PDF + SAVE TO MY ...
  35. [35]
    Detecting Novel Associations in Large Data Sets - Science
    Dec 16, 2011 · Here, we describe an exploratory data analysis tool, the maximal information coefficient (MIC), that satisfies these two heuristic properties.Missing: maximal information coefficient paper
  36. [36]
    A Non-Parametric Test of Independence - Project Euclid
    A test is proposed for the independence of two random variables with continuous distribution function (d.f.). The test is consistent with respect to the ...Missing: original | Show results with:original
  37. [37]
    Reducing Bias and Error in the Correlation Coefficient Due to ... - NIH
    One previous report (Zimmerman et al., 2003) found some evidence of a slight positive bias (up to +.05) for the Pearson correlation coefficient under conditions ...
  38. [38]
    Ecological Correlations and the Behavior of Individuals - jstor
    That is, the ecological correlation is the weighted difference between the total indi- vidual correlation and the average of the m within-areas individual ...
  39. [39]
    Graphs in Statistical Analysis: The American Statistician
    Mar 12, 2012 · Cleveland et al. Journal of the American Statistical Association. Volume 79, 1984 - Issue 387. Published online: 12 Mar 2012.
  40. [40]
    Robust Correlation Analyses: False Positive and Power Validation ...
    As designed by Anscombe, Pearson's correlation is fooled by outliers and, for each pair, a significant correlation of r = 0.81 is observed (Table 1; Figure 2). ...
  41. [41]
    Correlation Types
    It is relatively robust to outliers and deals well with data that have many ties. ... Winsorized correlation: Correlation of variables that have been Winsorized ...
  42. [42]
    [PDF] Measuring the Relationship of Bivariate Data Using Hodges ...
    The proposed Hodges Lehmann correlation coefficients are denoted as rHL(med), rHL(MAD) and rHL(MADn) that employed the scale estimator median, MAD and MADn.
  43. [43]
    Detection of Influential Observation in Linear Regression
    A new measure based on confidence ellipsoids is developed for judging the contribution of each data point to the determination of the least squares estimate.
  44. [44]
    [PDF] multivariate estimation with high breakdown point - KU Leuven
    The MCD also has the same breakdown point as the MVE, using the same reasoning as in Proposition 3.1. Both the MVE and the MCD are very drastic, because they.
  45. [45]
    What Is a Correlation Matrix? - Nick Higham
    Apr 14, 2020 · A correlation matrix is a symmetric positive semidefinite matrix with unit diagonal. In other words, it is a symmetric matrix with ones on the diagonal whose ...
  46. [46]
    [PDF] Estimating High Dimensional Covariance Matrices and its Applications
    Once given the cleaned correlation matrix, the cleaned sample covariance is constructed as. S = D1/2 C D1/2. It should be pointed that while C is positive ...
  47. [47]
    Proof: Positive semi-definiteness of the covariance matrix
    Sep 26, 2022 · A positive semi-definite matrix is a matrix whose eigenvalues are all non-negative or, equivalently, Mpos. semi-def. ⇔xTMx≥0for allx∈Rn.
  48. [48]
    Pearson's correlation between three variables; Using students' basic ...
    Aug 6, 2025 · Through a geometric interpretation, we start from two correlation coefficients r AB and r BC and then estimate a range for the third correlation ...
  49. [49]
    PCA, eigen decomposition and SVD
    Ease and value of interpretation, output, and statistics; Computational feasibility of computing eigen decomposition and correlation/covariance matrix on large ...
  50. [50]
    Computing the nearest correlation matrix—a problem from finance
    Higham, Computing the nearest correlation matrix—a problem from finance, IMA Journal of Numerical Analysis, Volume 22, Issue 3, July 2002, Pages 329–343 ...
  51. [51]
    [PDF] Computing the Nearest Correlation Matrix—A Problem from Finance
    Oct 22, 2001 · Abstract. Given a symmetric matrix what is the nearest correlation matrix, that is, the nearest symmetric positive semidefinite matrix with ...
  52. [52]
    Explicit solutions to correlation matrix completion problems, with an ...
    Mar 14, 2018 · Here we are concerned with problems in which the missing values are in the correlation matrix itself. Some of the matrix entries are known, ...
  53. [53]
    [PDF] Stochastic Processes - Earth, Atmospheric, and Planetary Physics
    If x(t) is stationary, its auto-correlation function is given by. Γx(τ) = E[x(t)x(t + τ)T] = Cx(τ) + µx(t)µx(t + τ)T. The auto-correlation function for a scalar ...
  54. [54]
    [PDF] Module 2: Stochastic Processes
    The "Cross Correlation" function of Xt and Yt is given by. Rx, y (s, t) = E [Xs Y₂]. -The "Cross Covariance" function is given by. Cx, y (s, t) = Cov (Xss ...
  55. [55]
    Chapter 2 Stationarity, Spectral Theorem, Ergodic Theorem(Lecture ...
    Jan 7, 2021 · Definition 2.2 (Cross-covariance Function) A useful function for the study of coevolution of two stochastic processes, say X X and Y Y ...
  56. [56]
    [PDF] Lecture 2: Measures of Dependence and Stationarity
    Auto-covariance measures linear dependence between variates in a time series. Auto-correlation, derived from it, is between -1 and 1.
  57. [57]
    [PDF] Ch. 6 Stochastic Process - Dr. Jingxian Wu
    Page 3. DEFINITION. • Stochastic process, or, random process. – A random variable changes with respect to time. • Example: the temperature in the room.
  58. [58]
    [PDF] 2 Brownian Motion - Arizona Math
    In short, Brownian motion is a stochastic process whose increments are independent, stationary and normal, and whose sample paths are continuous. Increments ...
  59. [59]
    [PDF] Introduction to Random Processes and Applications - Rice University
    Hence, sampling does not normally yield uncorrelated amplitudes, meaning ... In waveform processes, the stochastic process is defined by the multivariate ...<|control11|><|separator|>
  60. [60]
    [PDF] Chapter 2 - POISSON PROCESSES - MIT OpenCourseWare
    A Poisson process is a simple and widely used stochastic process for modeling the times at which arrivals enter a system.
  61. [61]
    Basic Concepts of the Poisson Process - Probability Course
    The Poisson process is one of the most widely-used counting processes. It is usually used in scenarios where we are counting the occurrences of certain events.
  62. [62]
    Autoregressive Conditional Heteroscedasticity with Estimates of the ...
    ARCH processes are mean zero, serially uncorrelated processes with nonconstant variances conditional on the past, but constant unconditional variances.
  63. [63]
    5.3.2 Bivariate Normal Distribution - Probability Course
    If X and Y are bivariate normal and uncorrelated, then they are independent. Proof. Since X and Y are uncorrelated, we have ρ(X,Y)=0.
  64. [64]
    [PDF] Markov Processes
    A Markov process has the Markov property, where the future is independent of the past given the present. It is defined by transition probabilities.
  65. [65]
    A general white noise test based on kernel lag-window estimates of ...
    In this paper, we develop a general white noise test based on kernel lag-window estimators of the spectral density operator of a time series with values in a ...
  66. [66]
    [PDF] White noise testing for functional time series
    Abstract: We review white noise tests in the context of functional time series, and compare many of them using a custom developed R package wwntests.
  67. [67]
    Correlation, Causation, and Confusion - The New Atlantis
    In 1911, Karl Pearson, inventor of the correlation ... correlation does not imply causation unless the correlation is statistically significant.
  68. [68]
    Correlation, Causation, and Confounding: Decoding Hidden ...
    Feb 18, 2025 · Consider the popular example: the positive correlation between ice cream sales and drowning incidents. On hot summer days, both ice cream ...
  69. [69]
    Per capita consumption of margarine correlates with The divorce ...
    Our findings provide compelling evidence to support the idea that there is a strong correlation between per capita consumption of margarine and the divorce rate ...
  70. [70]
    Spurious correlations: Margarine linked to divorce? - BBC News
    May 26, 2014 · "Maybe when there's more margarine in the house it's more likely to cause divorce," muses Tyler Vigen, "or there's a link with some of the ...
  71. [71]
    Interpreting Randomized Controlled Trials - PMC - PubMed Central
    This article describes rationales and limitations for making inferences based on data from randomized controlled trials (RCTs).
  72. [72]
    Using instrumental variables to establish causality - IZA World of Labor
    Using instrumental variables helps to address omitted variable bias. Instrumental variables can be used to address simultaneity bias.Missing: mitigating fallacy Granger
  73. [73]
    (PDF) From Correlation to Granger Causality - ResearchGate
    Aug 7, 2025 · The paper focuses on establishing causation in regression analysis in observational settings. Simple static regression analysis cannot establish causality.Missing: mitigating fallacy RCTs
  74. [74]
    What Everyone Should Know about Statistical Correlation
    A U-shaped relationship between two variables may have a linear correlation coefficient of zero, but in that case it does not imply that the variables are ...This Article From Issue · Volume 103, Number 1 · Page 26
  75. [75]
  76. [76]
  77. [77]
    The effects of sample size and variability on the correlation coefficient
    The results indicated that variability in excess of 10% of the range for each variable resulted in a mean reduction of the shared variance by 50% or greater.
  78. [78]
    Simpson's Paradox - Stanford Encyclopedia of Philosophy
    Mar 24, 2021 · Simpson's Paradox is a statistical phenomenon where an association between two variables in a population emerges, disappears or reverses when the population is ...Introduction · Simpson's Paradox and... · What Makes Simpson's... · Applications
  79. [79]
    4.2 - Controlling Family-wise Error Rate | STAT 555
    The most commonly used method which controls FWER at level α α is called Bonferroni's method. It rejects the null hypothesis when p<α/m.
  80. [80]
    Teaching the Difficult Past of Statistics to Improve the Future
    Jul 20, 2023 · Francis Galton is perhaps best known for two key concepts: eugenics and correlation, both terms he coined himself. Galton's statistical work, ...
  81. [81]
  82. [82]
    [PDF] STAT 234 Lecture 12 Condition Distributions Section 5.3
    The joint pdf of the Bivariate Normal Distribution is f(x, y)=. 1. 2πσxσyp1 ... The conditional distribution of Y given X = x is Normal with mean = µy + ...
  83. [83]
    Lesson 21: Bivariate Normal Distributions | STAT 414
    To calculate such a conditional probability, we clearly first need to find the conditional distribution of Y given X = x . That's what we'll do in this lesson, ...
  84. [84]
    [PDF] Introduction to Normal Distribution
    Jan 17, 2017 · Bivariate Normal. Distribution Form. Normal Density Function (Bivariate). Given two variables x,y ∈ R, the bivariate normal pdf is f(x,y) ...Missing: formula | Show results with:formula