Fact-checked by Grok 2 weeks ago

Granger causality

Granger causality is a for determining whether one can be used to forecast another, based on the principle that if variable X "Granger-causes" variable Y, then past values of X provide statistically significant information about future values of Y beyond what is available from past values of Y alone. This concept was formally introduced by econometrician W. J. Granger in his seminal paper, for which Granger was awarded the Nobel Memorial Prize in Economic Sciences in 2003 (jointly with ), where he proposed testable definitions of causality and feedback using econometric models and cross-spectral methods to analyze relationships in economic data. The core method involves fitting vector autoregressive (VAR) models to multivariate time series and testing whether the coefficients associated with lagged values of one variable significantly improve the model's predictive accuracy for another variable, often using F-tests or likelihood ratio statistics under the assumption of stationarity. Extensions include conditional Granger causality, which accounts for additional variables to avoid spurious inferences from omitted information, and frequency-domain variants that decompose causal influences across different oscillatory bands. Originally developed for economics to investigate causal directions in phenomena like money supply and income fluctuations, Granger causality has since been widely adopted in fields such as neuroscience for mapping directed interactions in brain networks via electroencephalography (EEG) or functional magnetic resonance imaging (fMRI) data, and in genomics for inferring regulatory relationships among gene expressions. Despite its utility, Granger causality does not imply true philosophical causation but rather predictive precedence, and it is sensitive to limitations such as non-stationarity, instantaneous correlations, and model misspecification, which can lead to false positives if variables are excluded. Recent advances address these issues through robust estimation techniques, nonlinear extensions, and out-of-sample validation to enhance reliability in high-dimensional settings like climate modeling or assessment.

Fundamentals

Intuition

Granger causality refers to a statistical where one , say X, is said to Granger-cause another Y if the past values of X provide statistically significant information for predicting future values of Y, beyond the information already available from the past values of Y alone. This predictive relationship emphasizes temporal precedence and improved forecasting accuracy, rather than implying a direct physical or mechanistic cause. To build intuition, consider a weather forecasting analogy: just as historical wind patterns and temperature data from one region can enhance predictions of rainfall in another area—beyond relying solely on that area's own past weather—Granger causality checks if one variable's history bolsters forecasts for another without assuming an underlying physical process. This approach highlights directionality in time series , where the "cause" precedes the "effect" in a predictive sense. Importantly, Granger causality differs from mere , which detects simultaneous associations without regard to timing or ; instead, it focuses on whether including the potential cause reduces forecast errors, potentially revealing directional influences confounded by omitted variables. For example, in , past values of can Granger-cause by improving forecasts of economic activity beyond using alone.

Historical development

The concept of what would later be known as Granger causality traces its origins to the work of in the , particularly his ideas on prediction theory and , where one is considered causal to another if past values of the former improve the prediction of the latter. Wiener's 1956 formulation emphasized statistical predictability in stationary processes, laying the groundwork for operationalizing in time series analysis. Clive W. J. Granger formalized and extended this notion in his seminal 1969 paper, "Investigating Causal Relations by Econometric Models and Cross-Spectral Methods," published in Econometrica, introducing a testable framework for inferring causality based on predictive information content in econometric contexts. Granger's approach adapted Wiener's ideas to econometric modeling, focusing on whether including lagged values of one variable enhances forecasts of another, thus providing a practical tool for empirical analysis in economics. In the , Granger causality gained early traction in amid macroeconomic debates, such as those on money-income relationships spurred by Christopher Sims' 1972 paper, "Money, Income, and ," which employed the test to challenge traditional econometric models' identification assumptions, arguing that atheoretical vector autoregressions (VARs) better revealed causal directions without imposing restrictive prior structures. This work, along with related analyses of and economic indicators, spurred methodological refinements and broader adoption of causality testing in . By the and , the framework expanded beyond into fields such as and , where it was used to probe directional influences in dynamics and neural signal interactions, respectively. Granger's contributions, including this concept, were recognized with the 2003 Nobel Memorial Prize in Economic Sciences, shared with , for advancing analysis techniques that handle and volatility.

Mathematical Foundations

Bivariate formulation

In the bivariate formulation, Granger causality examines whether one , say X_t, provides statistically significant information for predicting another Y_t beyond what is contained in the past values of Y_t itself. Formally, X is said to Granger-cause Y if the of Y_t obtained from forecasts using only its own past values exceeds the when past values of X_t are also included in the forecasting model. This concept is operationalized within a linear vector autoregressive (VAR) framework assuming stationarity and no instantaneous causation between the series. Consider two stationary time series Y_t and X_t, stacked into a bivariate vector \mathbf{y}_t = \begin{pmatrix} Y_t \\ X_t \end{pmatrix}. The unrestricted VAR(p) model of order p is given by \mathbf{y}_t = \boldsymbol{\nu} + \sum_{i=1}^p \mathbf{A}_i \mathbf{y}_{t-i} + \mathbf{u}_t, where \boldsymbol{\nu} is a constant term vector, each \mathbf{A}_i is a $2 \times 2 coefficient matrix partitioned as \mathbf{A}_i = \begin{pmatrix} a_{11,i} & a_{12,i} \\ a_{21,i} & a_{22,i} \end{pmatrix}, and \mathbf{u}_t is a white noise error vector with zero mean and covariance matrix \boldsymbol{\Sigma}_u. Equivalently, the equation for Y_t from this model is Y_t = \nu_Y + \sum_{i=1}^p \alpha_i Y_{t-i} + \sum_{i=1}^p \beta_i X_{t-i} + \epsilon_t, where \epsilon_t is the error term, the \alpha_i capture autoregressive effects from past Y, and the \beta_i represent potential Granger-causal influences from past X. To test the null hypothesis that X does not Granger-cause Y (i.e., \beta_1 = \beta_2 = \dots = \beta_p = 0), a restricted model is estimated that excludes the lagged X terms: Y_t = \nu_Y + \sum_{i=1}^p \gamma_i Y_{t-i} + \eta_t, where \eta_t is the corresponding error term. The test compares the residual sum of squares from the restricted model (RSS_r) and the unrestricted model (RSS_u). Under the null, the F-statistic for the joint significance of the \beta_i coefficients is F = \frac{(RSS_r - RSS_u)/p}{RSS_u / (n - 2p - 1)}, which follows an F-distribution with p and n - 2p - 1 degrees of freedom, where n is the effective sample size and p is the lag order. Rejection of the null at conventional significance levels indicates Granger causality from X to Y. The validity of this bivariate test relies on key assumptions: both series must be covariance stationary to ensure consistent estimation and valid inference; the errors across equations may be contemporaneously correlated, which can reflect instantaneous relationships between the variables, but the Granger causality test focuses on lagged effects; and the lag length p must be sufficiently large to capture the dynamics but not so large as to overfit the model. The lag order p is typically selected by minimizing an information criterion such as the (AIC), defined for the as AIC(p) = \ln|\hat{\boldsymbol{\Sigma}}_u(p)| + 2(4p)/n, where \hat{\boldsymbol{\Sigma}}_u(p) is the estimated of residuals.

Multivariate formulation

The multivariate formulation of Granger causality extends the bivariate case to systems involving multiple time series, incorporating vector autoregressive (VAR) models to account for interactions among all variables. In this framework, a stationary k-dimensional time series \mathbf{Y}_t is modeled as a VAR(p) process: \mathbf{Y}_t = \sum_{i=1}^p \mathbf{A}_i \mathbf{Y}_{t-i} + \boldsymbol{\epsilon}_t, where \mathbf{Y}_t is a k \times 1 vector, each \mathbf{A}_i is a k \times k coefficient matrix capturing the influence of lagged values, p is the lag order, and \boldsymbol{\epsilon}_t is a k \times 1 white noise vector with mean zero and possibly contemporaneous correlations. This represents the unrestricted full model, where past values of all variables jointly predict the current state. The bivariate case emerges as a special instance when k=2 and no additional controlling variables are present. Granger causality in the multivariate setting tests whether past values of a of variables, say \mathbf{X}, help predict another \mathbf{Y} beyond the provided by the remaining variables \mathbf{Z}. Specifically, \mathbf{X} Granger-causes \mathbf{Y} if the in the blocks of the \mathbf{A}_i matrices linking lagged \mathbf{X} to current \mathbf{Y} are jointly non-zero, after controlling for lags of \mathbf{Y} and \mathbf{Z}. This conditional form of causality contrasts with the unconditional bivariate test, as the full multivariate VAR inherently conditions on all other variables in the , allowing assessment of direct influences while isolating effects from the broader set. For a S (e.g., \mathbf{X}) potentially causing another (e.g., the equations for \mathbf{Y}), the test examines whether restricting the relevant blocks to zero significantly worsens the model's fit. The standard inference procedure adapts the block to compare the restricted model (where lags of S are excluded from the equations for the target variables) against the unrestricted full . Under the of no from S, the is F = \frac{(RSS_r - RSS_u)/q}{RSS_u/(n - m)} \sim F(q, n - m), where RSS_r and RSS_u are the from the restricted and unrestricted models, respectively; q is the number of restricted parameters; n is the effective sample size; and m is the total number of parameters in the unrestricted model. Rejection of the null indicates significant predictive from S to the targets. This block-wise approach enables testing between groups of variables, such as entire sectors in . Selecting the lag order p in multivariate VARs is crucial but challenging due to the rapid growth in parameters (k^2 p) with dimensionality k, leading to degrees-of-freedom issues and potential in high dimensions. Multivariate extensions of information criteria, such as the (AIC) or (BIC), are commonly used: these penalize model complexity by comparing likelihoods across candidate p values, with BIC imposing a stronger penalty for larger models to favor . In practice, sequential testing or cross-validation may supplement these criteria to balance fit and interpretability.

Testing Procedures

Parametric estimation

Parametric estimation of Granger causality relies on fitting linear vector autoregressive () models to data, as outlined in the bivariate and multivariate formulations, and then testing specific coefficient restrictions to assess predictive causality. The process begins with pre-testing the data for stationarity to ensure the validity of the VAR assumptions, typically using unit root tests such as the Augmented Dickey-Fuller () test, which examines the of a unit root against the alternative of stationarity. If the series are non-stationary, cointegration tests like the Johansen are conducted to check for long-run relationships; in the presence of cointegration, a vector () is preferred over a levels VAR to account for the error correction terms. Once stationarity is confirmed (or differences/VECM applied as needed), the VAR model is fitted using ordinary least squares (OLS) estimation for each equation in the system, which is efficient under Gaussian assumptions. The lag order p is selected to balance model fit and parsimony, commonly via information criteria such as the (AIC) or (BIC), or through sequential likelihood ratio (LR) tests that compare nested models up to a maximum lag. These methods help avoid while capturing relevant dynamics, with AIC often favoring more lags for forecasting purposes. To test for Granger causality, restrictions are imposed on the VAR coefficients—specifically, the joint significance of lagged terms of one variable in the equation of another is examined using , likelihood ratio (LR), or , which under standard conditions follow an asymptotic . The is particularly common in finite samples for its simplicity and equivalence to the under normality. For small samples where asymptotic approximations may lack power, bootstrap methods improve inference by resampling the OLS residuals to generate empirical distributions of the and compute p-values, often using a or wild bootstrap to preserve the VAR structure. Post-estimation diagnostic checks are essential to validate the model assumptions. Residual is assessed with the Ljung-Box test, testing the null of no serial correlation up to a specified . of residuals is evaluated using the Jarque-Bera test, which combines and measures against the null of Gaussianity. Heteroskedasticity, often checked via ARCH-LM tests on squared residuals, ensures constant variance; violations may necessitate robust standard errors or GARCH extensions, though standard Granger tests assume homoskedasticity. Software implementations facilitate these steps, with packages like the 'vars' library in providing built-in functions for VAR estimation, lag selection via AIC/BIC/LR, Granger tests (F/Wald), bootstrap procedures, and diagnostics including Ljung-Box and Jarque-Bera. Similar functionality is available in Python's 'statsmodels' and MATLAB's Toolbox, enabling reproducible parametric analysis.

Non-parametric tests

Non-parametric tests for Granger causality relax the assumption of linear parametric models, such as vector autoregressions, by employing flexible estimation techniques that can capture nonlinear dependencies without specifying functional forms. These methods are particularly useful when the underlying data-generating process exhibits nonlinearity or when model misspecification in parametric approaches might lead to biased inferences. Kernel density estimation approaches form a core class of non-parametric tests, utilizing local constant or polynomial regressions to estimate conditional means and densities directly from the data, thereby avoiding the need for explicit specifications in the model structure. This involves estimating joint probability densities of past values of to assess whether the inclusion of one 's history improves predictions of another, capturing both linear and nonlinear causal relationships through smoothed local approximations. For instance, the test evaluates by comparing density estimates with and without the potential causing , providing a robust alternative to benchmarks that assume linearity. In the spectral domain, non-parametric tests decompose the time series via Fourier transforms to examine causality in the frequency domain, where Granger causality holds if the spectral density matrix exhibits significant off-diagonal elements at specific frequencies, indicating directional influence at those scales. This approach estimates the power spectra and cross-spectra non-parametrically, often using methods like the Welch periodogram or multitaper techniques, to quantify how variance in one series at a given frequency is explained by another without relying on time-domain parametric fits. A prominent example is the Diks-Panchenko test introduced in 2006, which employs a -based to detect nonlinear Granger causality by testing for serial in residuals after accounting for linear dependencies, effectively extending the classical linear to nonlinear settings through local dependence measures derived from kernel density estimates. The test aggregates pairwise local correlations weighted by density estimates, yielding an asymptotically under the null of non-causality. These non-parametric methods offer key advantages, including their ability to handle nonlinear dependencies that tests might overlook and the reduced to lag length selection, as smoothing or inherently incorporates temporal structure without fixed-order assumptions. However, they come with trade-offs in implementation, such as the need for selection in methods—typically achieved via cross-validation or rules like e_n = C n^{-2/7} with C around 8 for certain processes—and higher computational intensity due to density or spectral estimations over large datasets.

Limitations and Criticisms

Philosophical and statistical issues

Granger causality, as originally defined, pertains to the ability of one time series to improve the prediction of another, rather than establishing true causal mechanisms involving interventions or manipulations. This predictive focus means that a variable may appear to Granger-cause another without exerting any actual influence, as illustrated by the classic example where a barometer reading "Granger-causes" rainfall predictions due to shared correlation with atmospheric pressure, yet changing the barometer has no effect on the weather. A major statistical issue arises from , where excluding relevant confounders in the model can lead to spurious or missed Granger causal relations. In bivariate analyses, non-causality may stem from an omitted third variable that influences the target series, necessitating multivariate specifications to mitigate this bias and accurately identify causal directions. models underlying Granger causality accommodate contemporaneous effects through correlations in the innovations' ; however, the tests focus on lagged values and do not distinguish instantaneous from lagged influences, which can distort causal inferences by implying without clear directionality when innovations are correlated. This issue requires structural adjustments, such as rotation matrices in frequency-domain analyses, to isolate lagged effects. Directionality in Granger causality tests can be problematic when relations are bidirectional, as the standard formulation yields symmetric results that fail to distinguish feedback loops without additional lags, theoretical priors, or supplementary tests. In such cases, both series equally improve mutual predictions, obscuring the net flow of influence and requiring to resolve ambiguity. Philosophically, Granger causality aligns with probabilistic theories of causation, such as ' framework, which defines prima facie causes through precedence and positive probabilistic dependence, providing a temporal and basis for Granger's predictive criterion. It contrasts with interventionist approaches like Pearl's do-calculus, which emphasizes manipulability and counterfactuals to distinguish causation from mere association, highlighting Granger's limitations in handling or structural changes.

Practical challenges

In applied settings, Granger causality tests, whether or non-parametric, encounter several practical obstacles that can compromise their reliability and implementation. One key challenge is the requirement for adequate sample sizes; reliable typically demands at least 80–100 observations per to achieve sufficient statistical power and minimize Type II errors, where genuine causal relationships fail to be detected due to low power in smaller datasets. For instance, with modest sample lengths like 30–60 periods, post-sample validation becomes difficult, leading to erratic or inconclusive results. Another implementation hurdle involves preprocessing the data. In non-parametric tests, prewhitening is often essential to filter out present in the series, as unaddressed serial correlation can lead to spurious detections. This step helps ensure that cross-correlations reflect true relationships rather than artifacts of . In multivariate formulations, among predictor variables poses a significant issue, as high correlations inflate the variance of estimates in VAR models, thereby reducing the tests' power to detect true Granger causality. Regularization techniques, such as penalties, are often required to stabilize estimates and mitigate this problem in high-dimensional settings. Software availability further complicates practical application, particularly for non-parametric tests. Standard libraries like R's lmtest package offer only parametric Granger causality functions, such as grangertest, while non-parametric alternatives demand specialized or custom code. Similarly, Python's statsmodels provides parametric grangercausalitytests but lacks built-in non-parametric options, often necessitating additional packages or implementations. Finally, arising from reverse causation or bidirectional feedback loops can distort Granger tests, which do not account for contemporaneous dependencies. Addressing this requires shifting to structural VAR models, which impose identifying restrictions based on economic to disentangle causal directions and control for .

Extensions

Time-varying methods

Time-varying methods extend the Granger causality to accommodate evolving relationships in non-stationary data, where causal influences may change due to structural shifts or gradual dynamics. These approaches adapt the multivariate (VAR) model by allowing parameters to vary over time, enabling detection of transient or regime-dependent causality. Such techniques are particularly valuable for analyzing data subject to interventions, economic cycles, or other breaks that violate stationarity assumptions. One common method is the rolling-window VAR estimation, which assesses Granger causality by repeatedly fitting the model to overlapping subsets of the data within a fixed-size window, such as 50 observations, and tracking how causal directions or strengths evolve across windows. This nonparametric approach captures local dynamics without assuming a specific form for parameter changes, though it requires careful selection of window size to balance bias and variance. For instance, in examining -output relationships, rolling-window tests reveal periods where Granger-causes output more strongly, highlighting shifts in over time. Time-varying parameter (TVP) models further formalize this evolution by modeling coefficients as processes, often estimated via the , where parameters update recursively as \beta_t = \beta_{t-1} + \nu_t with \nu_t as a term. This state-space approach allows for smooth or abrupt changes in Granger causal linkages, accommodating heteroskedasticity and non-stationarity in multivariate settings. Applications in , for example, use TVP-based dynamic Granger causality to track frequency-domain connectivity in functional MRI data, revealing time-dependent directed influences between brain regions. Local projections, introduced by Jordà (2005), provide an alternative by directly regressing future outcomes on current shocks and lags using time-local weights or kernel smoothing, bypassing full VAR specification to estimate impulse responses that inform Granger causality. This method projects outcomes at horizon h onto leads and lags of variables, weighted locally around each time point, yielding flexible estimates of dynamic causal effects without assuming global parameter constancy. It proves robust for identifying time-varying responses in macroeconomic data, such as policy shock impacts. To validate the need for time-varying methods, tests for constancy, such as the supremum F (Sup-F) and supremum Wald (Sup-Wald) , assess in VAR coefficients against alternatives of one-time or gradual breaks. These Lagrange multiplier-based tests, applicable to integrated processes, detect structural instability by scanning for maximum deviations from the null of constant parameters. (1992) derives their asymptotic distributions, showing non-standard limiting behavior under the null, which necessitates bootstrap or simulation for critical values in finite samples. These methods are especially useful for analyzing scenarios involving changes or structural breaks, where standard Granger tests may fail due to unmodeled , as demonstrated in studies of and market linkages.

Nonlinear and frequency-domain variants

Traditional linear Granger causality assumes linear relationships between , but many real-world systems exhibit nonlinear dependencies that these methods fail to detect. To address this, nonlinear extensions have been developed using nonparametric approaches. A seminal test for nonlinear Granger causality in bivariate settings was proposed by Baek and Brock, employing correlation-integral statistics to assess whether past values of one series improve predictions of another beyond linear effects. This approach was extended to multivariate cases by Hiemstra and Jones, allowing detection of nonlinear causal links in higher dimensions. More recent nonlinear methods leverage techniques for greater flexibility. Kernel-based approaches map into a , where linear Granger causality is applied to capture arbitrary nonlinear interactions without assuming specific functional forms. Similarly, models, such as multilayer perceptrons or recurrent s with sparsity-inducing penalties, estimate nonlinear predictive improvements, enabling in complex, high-dimensional systems. Frequency-domain variants extend Granger causality to analyze causal influences at specific frequencies, decomposing time-domain effects into spectral components. Geweke introduced a measure of causal influence from series Y to X at frequency \omega, defined as f_{Y \to X}(\omega) = \frac{1}{2\pi} \left[ \log \Sigma_{XX}(\omega) - \log \Sigma_{XX|Y}(\omega) \right], where \Sigma_{XX}(\omega) is the of X, and \Sigma_{XX|Y}(\omega) is the of X conditional on past values of Y. This formulation quantifies how past Y reduces the prediction error variance of X in the , revealing causality strengths across different oscillatory bands. For multivariate systems, partial directed coherence (PDC) provides a graph-theoretic of frequency-specific . Introduced by Baccalá and Sameshima, PDC measures direct causal flows between nodes by normalizing the elements, suppressing indirect paths and highlighting network structures in frequency bands. Testing these frequency-domain measures often involves decompositions. Breitung and Candelon developed a framework to distinguish short- and long-run by examining the of the function, using a simple based on Geweke's measure to assess significance at targeted frequencies. These variants offer advantages in capturing periodic phenomena, such as business cycles, where causal influences may dominate at low frequencies corresponding to long-term fluctuations rather than high-frequency noise.

Applications

Economics and finance

Granger causality has played a pivotal role in macroeconomic research since its introduction, particularly in testing the directional influence of on economic output. This application stems from debates sparked by and Anna Schwartz's 1963 analysis in A Monetary History of the United States, 1867-1960, which highlighted money's leading role in business cycles and challenged Keynesian views of fiscal dominance. Early econometric studies adopted Granger's to empirically verify these claims, finding of unidirectional causality from to output in postwar U.S. data, thereby supporting monetarist arguments that monetary expansions drive real economic fluctuations. For instance, analyses of U.S. from the onward demonstrated that lagged money growth significantly improves forecasts of GDP, reinforcing the policy relevance of controlling money aggregates to stabilize output. In financial markets, Granger causality tests have illuminated relationships between corporate disclosures and asset prices, such as the predictive power of earnings announcements on stock returns. Seminal work on the has shown that innovations in quarterly earnings Granger-cause subsequent stock returns, with the effect strongest during non-bubble periods when fundamental information dominates pricing. This causality implies that past earnings data enhance return forecasts beyond autoregressive models alone, aiding investors in timing entries based on announcement surprises. Complementing this, volatility spillovers across assets have been probed using extensions like combined with generalized autoregressive conditional heteroskedasticity (VAR-GARCH) models, where Granger causality in variances captures how shocks in one market's propagate to others. For example, studies of and markets reveal bidirectional volatility causality during crises, quantifying contagion risks in portfolios. Policy implications of Granger causality extend to central banking, especially in post-1970s studies examining how adjustments influence dynamics amid the Great Inflation era. Research on U.S. data from the through the has tested whether policy rate changes Granger-cause , often finding unidirectional effects where rate hikes precede disinflationary episodes, validating the transmission mechanism of monetary tightening. These findings informed the Volcker-era shift toward aggressive rate targeting to break inflationary inertia, with Granger tests confirming that policy innovations explain significant variance in future price levels. A key contribution comes from Stock and Watson's (2001) framework, which applied Granger causality to the output-inflation nexus, revealing bidirectional predictability that underscores the trade-offs central banks face in stabilizing both variables. Despite its utility, applying Granger causality in encounters challenges with high-frequency , where intraday trading generates vast, noisy series that standard tests assume away. Traditional Granger methods, designed for daily or lower frequencies, struggle with microstructure noise, asynchronous arrivals, and nonstationarities in tick-by-tick prices, often yielding spurious non-causality due to aggregation biases. Adaptations have emerged, such as high-frequency analogs that incorporate realized measures and kernel-based estimators to detect causality between intraday returns and order flows, enhancing detection of short-lived spillovers in environments. These modifications are crucial for high-speed , where failing to account for intraday patterns can mask true causal links in .

Neuroscience

In neuroscience, Granger causality has been adapted to analyze directed functional connectivity from neuroimaging time-series data, such as (fMRI) and (EEG), to infer how activity in one region predicts or influences another. For instance, it has revealed directional influences from the to the during visuomotor tasks, where visual processing signals precede and enhance predictions of motor activity. This approach leverages vector autoregressive models to quantify how past values of one region's BOLD signal or EEG oscillation improve forecasting of another's, distinguishing effective connectivity from mere correlation. Extensions of Granger causality to point processes have enabled its application to neural spike trains, treating discrete neuronal firing events as processes to detect causal influences in spiking activity. Pereda et al. (2005) reviewed nonlinear extensions, including adaptations for multivariate spike train analysis, allowing inference of directed interactions among ensembles of neurons. Further developments incorporate likelihood ratio tests on Hawkes models, which model spike trains as self-exciting point processes with cross-excitations representing causal drives; significant likelihood improvements indicate Granger causality between neuronal populations. These methods are particularly useful for high-resolution electrophysiological data, revealing causal interactions among neurons in the during reaching tasks. In multivariate settings, Granger causality facilitates mapping networks by estimating directed influences across multiple regions simultaneously, often using the directed transfer function (DTF) to decompose total flow into direct causal paths. Introduced by Kaminski and Blinowska (2001), DTF extends Granger causality into frequency-domain graphs of , applied in EEG studies to identify hierarchical influences in cognitive networks, such as prefrontal-to-parietal directions during tasks. Key studies, including Seth (2010), have emphasized its role in effective , integrating Granger measures with to validate causal hypotheses in fMRI data while addressing confounds like hemodynamic delays. However, applications in EEG face limitations from volume conduction, where instantaneous field spread creates spurious zero-lag correlations that inflate bidirectional causality; mitigation strategies include source reconstruction or spectral formulations of Granger causality. Software tools like the in support these analyses for neural data, providing routines for unconditional and conditional Granger tests on multivariate , including decompositions suitable for fMRI and EEG. Designed by et al., the toolbox handles multi-trial data and network inference, enabling researchers to compute DTF-based connectivity graphs from spike or oscillatory signals.

Computing and other fields

In , Granger causality serves as a foundational tool in algorithms for causal discovery from time series data, particularly within pipelines that handle high-dimensional and nonlinear dependencies. One prominent example is the PCMCI (Peter and Clark Momentary ) algorithm, which extends traditional Granger causality by incorporating tests to detect both direct and indirect causal links while controlling for false positives in large datasets. Implemented in open-source libraries like Tigramite, PCMCI integrates seamlessly into workflows, enabling scalable analysis in fields requiring automated discovery of temporal networks. In , Granger causality is integrated into hybrid models that combine vector autoregressive () structures with neural networks for enhanced . For instance, VAR-LSTM hybrids employ Granger tests to verify bidirectional between variables, such as and volumes, before feeding selected lags into the LSTM component to capture nonlinear patterns and residuals from VAR predictions. Similarly, VAR-GRU models use Granger causality to identify and select causal variables via optimal lag determination, improving prediction accuracy on multivariate financial data over standalone neural or linear models. These integrations leverage Granger's to inform , yielding superior performance metrics like MAPE reductions of up to 0.13% in tasks. Beyond computing, Granger causality finds applications in diverse fields such as climate science, where it has been used to establish directed influences like El Niño-Southern Oscillation (ENSO) on global vegetation activity, such as (LAI). Studies from the , including analyses of CMIP6 model outputs, demonstrate robust Granger causality from ENSO indices to LAI, with projected increases in impact under warming scenarios affecting up to 9% of global land area. In epidemiology, Granger causality helps trace exposure factors to disease spread; for example, during the 2018-19 Ebola outbreak in the Democratic Republic of , targeted against medical workers was found to Granger-cause increased Ebola incidence with lags of 8-29 days, disrupting response efforts and amplifying transmission. A key example in cybersecurity involves applying Granger causality to network traffic for attack attribution and detection. The Cybersecurity Granger Causality (CGC) framework analyzes of attack rates from data, constructing directed graphs to reveal bidirectional causal links between IP networks (e.g., 58.76% at /8 resolution), enabling improved and prediction accuracy when incorporating causal helpers at optimal lags like z=4. This approach outperforms non-causal baselines in identifying propagating threats across resolutions up to /24. In genomics, Granger causality is applied to time-series gene expression data to infer directed regulatory relationships, such as how variations in one gene's expression predict changes in others during processes like circadian regulation or cellular stress responses. Despite these advances, a notable gap in Granger causality applications is scalability to big data, addressed through parallel computing techniques like gradient boosting with Spark-based implementations. These methods achieve high accuracy (up to 1.0 on synthetic datasets) while reducing execution time on million-scale rows by factors proportional to worker nodes (e.g., 7:58 minutes for 10M rows with 8 nodes), supporting causality discovery in nonlinear, high-resolution environments.

References

  1. [1]
    Investigating Causal Relations by Econometric Models and Cross ...
    TiE OBJECT of this paper is to throw light on the relationships between certain classes of econometric models involving feedback and the functions arising in.
  2. [2]
    Granger Causality: A Review and Recent Advances
    Introduced more than a half-century ago, Granger causality has become a popular tool for analyzing time series data in many application domains, from economics ...
  3. [3]
    Granger Causality Analysis in Neuroscience and Neuroimaging - PMC
    Feb 25, 2015 · G-causality implements a statistical, predictive notion of causality whereby causes precede, and help predict, their effects. It is defined in ...Missing: original | Show results with:original
  4. [4]
  5. [5]
    Testing for Linear and Nonlinear Granger Causality in the Stock Price
    Linear and nonlinear Granger causality tests are used to examine the dynamic relation between daily Dow Jones stock returns and percentage changes in New.
  6. [6]
    Wiener–Granger Causality: A well established methodology
    Sep 15, 2011 · In 1956 Norbert Wiener introduced the notion that one variable (or time series) could be called 'causal' to another if the ability to predict ...Missing: theory | Show results with:theory
  7. [7]
    Investigating Causal Relations by Econometric Models and Cross ...
    Granger, C W J, 1969. "Investigating Causal Relations by Econometric Models and Cross-Spectral Methods," Econometrica, Econometric Society, vol.
  8. [8]
    [PDF] Money-Income Causality-- A Critical Review of the Literature Since ...
    3.4 Testing for Causality. I see the motivation for the revolution in method introduced by. Granger (1969) and Sims (1972) as residing in the feedback problem.
  9. [9]
    Money, Income, and Causality - jstor
    Granger's definition if, and only if, a or b can be chosen identically 0.12. This result gives us another intuitive handle on Granger causality. If causality.
  10. [10]
    [PDF] Money, Income, and Causality
    Granger has given a definition of a testable kind of causal ordering based on the notion that absence of correlation be- tween past values of one variable X and.
  11. [11]
    [PDF] Time-series Econometrics: Cointegration and Autoregressive ...
    In this section, we describe Clive Granger's contributions that lead up to the concept of cointegration and its applications. We begin by defining the concept.
  12. [12]
    [PDF] Investigating Causal Relations by Econometric Models and Cross ...
    TiE OBJECT of this paper is to throw light on the relationships between certain classes of econometric models involving feedback and the functions arising in.
  13. [13]
    [PDF] New Introduction to Multiple Time Series Analysis
    Feb 3, 2023 · It contains a discussion of structural vector autoregressive and vector error correction models which are by now also standard tools in applied ...
  14. [14]
    Granger Causality: A Review and Recent Advances - PMC
    Granger causality has become a popular tool for analyzing time series data in many application domains, from economics and finance to genomics and neuroscience.
  15. [15]
    Measurement of Linear Dependence and Feedback between ...
    Mar 12, 2012 · The measure of linear dependence is the sum of the measure of linear feedback from the first series to the second, linear feedback from the second to the first ...
  16. [16]
    None
    Nothing is retrieved...<|control11|><|separator|>
  17. [17]
    [PDF] Linking Granger Causality and the Pearl Causal Model with Settable ...
    The causal notions embodied in the concept of Granger causality have been argued to belong to a different category than those of Judea Pearl's Causal Model, ...Missing: book | Show results with:book
  18. [18]
  19. [19]
    A study of problems encountered in Granger causality analysis from ...
    In this paper, we analyze fundamental properties of Granger causality and illustrate statistical and conceptual problems that make Granger causality difficult ...
  20. [20]
    [PDF] 1 Granger Causality.
    Mar 1, 2005 · If the innovation to yt and the innovation to xt are correlated we say there is instantaneous causality. You will usually (or at least often) ...
  21. [21]
    Credible Granger-Causality Inference with Modest Sample Lengths
    Credible Granger-causality analysis appears to require post-sample inference, as it is well-known that in-sample fit can be a poor guide to actual ...
  22. [22]
  23. [23]
    grangertest function - RDocumentation
    ### Summary of grangertest (lmtest version 0.9-40)
  24. [24]
    statsmodels.tsa.stattools.grangercausalitytests - statsmodels 0.14.4
    ### Summary of statsmodels.tsa.stattools.grangercausalitytests
  25. [25]
    [PDF] Causal Search in Structural Vector Autoregressive Models
    The concept of Granger causality has been criticized for failing to capture 'structural causality' (Hoover, 2008). Suppose one finds that a variable A Granger- ...Missing: endogeneity | Show results with:endogeneity
  26. [26]
    [PDF] Granger causality in the presence of structural changeS
    We allow both for a fixed and an abrupt change of the constant terms in the VAR representation and evaluate how the true causal structure emerges in the two dif.Missing: endogeneity | Show results with:endogeneity
  27. [27]
    [PDF] Money and Output Viewed Through a Rolling Window * ABSTRACT
    We examine the extent to which fluctuations in the money stock anticipate (or Granger cause) fluctuations in real output using a variety of rolling window ...
  28. [28]
    [PDF] Testing for time-varying Granger causality - Stata
    The formulation of a test for Granger causality in a VAR(G) system, ... window specifies the number of observations to be included in the rolling windows.
  29. [29]
    Time-varying linear and nonlinear parametric model for Granger ...
    The main objective of this work is to present a linear and nonlinear time-varying parametric modeling and identification approach that can be used to detect ...
  30. [30]
    Dynamic Granger causality based on Kalman filter for evaluation of ...
    We propose an approach to dynamic Granger causality in the frequency domain for evaluating functional network connectivity in fMRI data.
  31. [31]
    Estimation and Inference of Impulse Responses by Local Projections
    This paper introduces methods to compute impulse responses without specification and estimation of the underlying multivariate dynamic system.Missing: Granger causality
  32. [32]
    [PDF] Tests for Parameter Instability in Regressions with I(1) Processes
    This article derives the large-sample distributions of Lagrange multiplier (LM) tests for parameter instability against several alternatives of interest in the ...
  33. [33]
    [PDF] Testing for Parameter Instability and Structural Change in Persistent ...
    We use our new sup-Wald testing procedure to study the stability of the predictive relation be- tween option implied and realized volatility (IV and RV), which ...
  34. [34]
    Granger Causality Tests in the Presence of Structural Changes
    Apr 4, 1998 · This paper illustrates that, if significant structural change occurs, these tests can provide misleading results. The paper then goes on to ...
  35. [35]
    Baek, E.G. and Brock, A.W. (1992) A General Test for Non-Linear ...
    (1992) A General Test for Non-Linear Granger Causality: Bivariate Model. Technical Report, Korean Development Institute and University of Wisconsin-Madison. has ...
  36. [36]
    Testing for Linear and Nonlinear Granger Causality in the Stock ...
    Baek, E., and W. Brock, 1992a, A general test for nonlinear Granger causality: Bivariate model, Working paper, Iowa State University and University of Wisconsin ...
  37. [37]
    [PDF] A note on the power of money-output causality testsy
    Since Friedman and Schwartz (1963) rekindled the research on the effects of money on aggregate output, numerous studies have aimed to characterize and establish ...
  38. [38]
    [PDF] Empirical Macroeconomics: The Effects of Monetary Policy
    Friedman and Schwartz's attribution of causality from money to business ... Granger causality to examine the causal relationship between money and output.
  39. [39]
    Granger Causality between Money, Output, Prices and Interest Rates
    Schwartz, A Monetary History of the United States,. 1867-1960. Princeton 1971. Friedman, Benjamin M., "The Role of Money and Credit in Macroeconomic ...Missing: original | Show results with:original
  40. [40]
    Intrinsic bubbles and Granger causality in the S&P 500: Evid
    We conclude that changes in earnings can predict future stock returns, but only in periods absent bubbles.<|separator|>
  41. [41]
    [PDF] Granger-causal analysis of GARCH models: a Bayesian approach
    The well-known concept of Granger causality (see Granger, 1969; Sims, 1972) describes relations between time series in the forecasting context. One variable ...
  42. [42]
    Measuring and Quantifying Uncertainty in Volatility Spillovers
    In multivariate GARCH-type models, spillover effects are described by Granger-causality in variance (eg Hafner and Herwartz Citation2008). In this case, the ...
  43. [43]
    Does Money Matter for U.S. Inflation? Evidence from Bayesian VARs ...
    Mar 1, 2008 · We use Bayesian estimation techniques to investigate whether money growth Granger-causes inflation in the United States. We test for ...I. Introduction · Ii. Other Related Literature · Iii. Methodology And Data<|separator|>
  44. [44]
    [PDF] Vector Autoregressions - Princeton University
    Standard practice in VAR analysis is to report results from Granger-causality tests, impulse responses and forecast error variance decompositions. These ...
  45. [45]
    Identifying money and inflation expectation shocks on real oil prices
    Aug 3, 2023 · This suggests that US monetary factors, in particular expected inflation, may be playing a role. Research investigated this and found Granger- ...
  46. [46]
    [PDF] Measuring High-Frequency Causality Between Returns, Realized ...
    Using high-frequency data increases the chance to detect causal links since aggregation may make the relationship between returns and volatility simultaneous.Missing: adaptations | Show results with:adaptations
  47. [47]
    Learning Financial Networks with High-Frequency Trade Data
    While using intraday data provides interpretability benefits, high-frequency financial time series also pose significant modeling challenges that are not ...
  48. [48]
    Granger Causality Testing in High-Dimensional VARs: A Post ...
    In this article, we focus on testing Granger causality in mean using linear models, in which setup the VAR model is the natural tool to investigate this problem ...2.1 The Lasso Estimator · 2.3 Tuning Parameter... · 3 Asymptotic Properties
  49. [49]
    Granger Causality Analysis in Neuroscience and Neuroimaging
    Feb 25, 2015 · GCA provides a powerful and generic statistical tool for characterizing directed functional interactions from time-series data.
  50. [50]
    Granger causality analysis of fMRI BOLD signals is invariant to ...
    Granger causality is a method for identifying directed functional connectivity based on time series analysis of precedence and predictability.
  51. [51]
  52. [52]
    A Granger Causality Measure for Point Process Models of Ensemble ...
    This paper proposes a point process framework that enables Granger causality to be applied to point process data such as neural spike trains.
  53. [53]
    Granger causality, directed transfer function and statistical ...
    We study the relation between the directed transfer function (DTF) and the well-accepted Granger causality, and show that DTF can be interpreted within the ...
  54. [54]
    A MATLAB toolbox for Granger causal connectivity analysis
    Feb 15, 2010 · This article describes a freely available MATLAB toolbox – 'Granger causal connectivity analysis' (GCCA) – which provides a core set of methods for performing ...Introduction · Bivariate And Conditional... · Causal Networks And Their...
  55. [55]
    A critical assessment of connectivity measures for EEG data
    Avoiding spurious connectivity caused by volume conduction in Granger-causal analyses. Our study illustrates challenges in the estimation of brain ...
  56. [56]
    The MVGC multivariate Granger causality toolbox - ScienceDirect.com
    The MVGC Matlab Toolbox implements numerical routines for calculating multivariate Granger causality (MVGC) from time series data, both unconditional and ...
  57. [57]
    Detecting and quantifying causal associations in large nonlinear ...
    Nov 27, 2019 · We introduce a novel method that flexibly combines linear or nonlinear conditional independence tests with a causal discovery algorithm to estimate causal ...
  58. [58]
    [PDF] Forecasting of Multivariate Time Series Data of Export and Import ...
    The primary objective of this study is to develop a hybrid forecasting model that integrates the Vector Autoregressive (VAR) model with Long Short-. Term Memory ...
  59. [59]
    VAR-GRU: A Hybrid Model for Multivariate Financial Time Series ...
    Vector Auto-regressive (VAR) models are useful for analyz-ing temporal dependencies among multivariate time series, known as Granger causality. There exist ...
  60. [60]
    Increased impact of the El Niño–Southern Oscillation on global ...
    Sep 2, 2023 · The methods used in this work are based on a multivariate predictive model to assess the null hypothesis of no Granger causality between ENSO ...
  61. [61]
    Analyzing Targeted Violence Against Medical Workers and EVD ...
    Dec 16, 2019 · Results found plausible Granger causality relationships between violent events and case incidence in the study region in both early and later ...
  62. [62]
    [PDF] Characterizing and Leveraging Granger Causality in Cybersecurity
    May 11, 2021 · We propose a framework, dubbed Cybersecurity Granger Causality (CGC), for characterizing the presence of G-causality in cyber attack rate time ...
  63. [63]
    [PDF] Parallel Gradient Boosting based Granger Causality Learning
    We study how to leverage gradient boosting meta machine learning techniques to achieve accurate causality discovery and big data parallel techniques for ...