Quantitative Finance ISSN: 1469-7688 (Print) 1469-7696 (Online) Journal homepage: www.tandfonline.com/journals/rquf20 Modeling variance risk in financial markets using power-laws: new evidence from the Garman-Klass variance estimator Masoumeh Fathi & Klaus Grobys To cite this article: Masoumeh Fathi & Klaus Grobys (07 Aug 2025): Modeling variance risk in financial markets using power-laws: new evidence from the Garman-Klass variance estimator, Quantitative Finance, DOI: 10.1080/14697688.2025.2529485 To link to this article: https://doi.org/10.1080/14697688.2025.2529485 © 2025 The Author(s). Published by Informa UK Limited, trading as Taylor & Francis Group Published online: 07 Aug 2025. Submit your article to this journal Article views: 110 View related articles View Crossmark data This article has been awarded the Centre for Open Science 'Open Data' badge. Full Terms & Conditions of access and use can be found at https://www.tandfonline.com/action/journalInformation?journalCode=rquf20 Quantitative Finance, 2025 https://doi.org/10.1080/14697688.2025.2529485 Modeling variance risk in financial markets using power-laws: new evidence from the Garman-Klass variance estimator MASOUMEH FATHI†* and KLAUS GROBYS†‡∗∗ †Finance Research Group, School of Accounting and Finance, University of Vaasa, Wolffintie 34, Vaasa 65200, Finland ‡Innovation and Entrepreneurship (InnoLab), University of Vaasa, Wolffintie 34, Vaasa 65200, Finland (Received 2 January 2025; accepted 26 June 2025 ) This study examines the range-based variance risk of five key financial asset markets—S&P 500, gold, crude oil, the USD/GBP exchange rate, and Bitcoin—using the noise-efficient Garman-Klass variance estimator. Our findings corroborate previous research by demonstrating that range-based asset variances adhere to power-law behavior generating variance behavior that is effectively infi- nite in practical terms. Furthermore, we provide novel evidence that the widely accepted log-normal model is unequivocally rejected for all range-based asset variances, underscoring its inadequacy in capturing the statistical properties of financial asset variances. A key contribution of this study is the discovery that a power-law function with α ≈ 2.8 represents a universal law governing the cross-sectional variances of otherwise unrelated asset markets. These findings have significant impli- cations for risk management frameworks that rely on models developed within the mean-variance space, as they highlight the limitations of traditional approaches in assessing and managing financial risks. Keywords: Garman-Klass estimator; Power-laws; Range-based variance; Second moment; Variance of variance JEL Classification: C18, G10, G12, G14 1. Introduction The seminal study by Markowitz (1952) laid the ground- work for portfolio analysis of financial assets, introducing the mean-variance framework as the cornerstone for opti- mizing portfolios of risky assets. Building on this founda- tion, Treynor (1962), Sharpe (1964), Lintner (1965), and Mossin (1966) developed the Capital Asset Pricing Model (CAPM), which remains a foundational concept in asset pric- ing literature. However, empirical failures of the CAPM have inspired the development of various factor models, culminat- ing in what Feng et al. (2020) term a ‘factor zoo.’ Among these, the Fama-French factor models (Fama and French 1992, 1993, 2015, 2017, 2018, 2020) are widely recognized ∗Corresponding author. Email: masoumeh.fathi@uwasa.fi ∗∗Current address: Department of Monetary Economics and International Finance, Christian-Albrechts University (CAU) of Kiel, Wilhelm-Seelig Platz 1, 24118 Kiel, Germany as benchmark models in empirical investigations aimed at addressing CAPM’s limitations. It is well understood that ‘true’ volatility is unobserv- able, and the literature offers various proxies to estimate it. These include the sample variance, conditional variance mod- eled through GARCH or stochastic volatility frameworks, and realized and range-based volatility derived from high- frequency or intraday price data. Range-based estimators, such as those proposed by Parkinson (1980) and Garman and Klass (1980), have been shown to be particularly efficient and robust, especially in the presence of market microstruc- ture noise, and are widely used as nonparametric measures of realized variance (Chou et al. 2010). While these factor models offer an appealing theoretical framework, their reliance on the mean-variance space presents critical challenges if the variance is statistically undefined. In his commentary on Mandelbrot’s (1963) ‘infinite variance hypothesis,’ which profoundly disrupted traditional finance theory, Fama (1963) cautioned that statistical tools assum- ing finite variance—such as least-squares regression—could © 2025 The Author(s). Published by Informa UK Limited, trading as Taylor & Francis Group This is an Open Access article distributed under the terms of the Creative Commons Attribution-NonCommercial-NoDerivatives License (http://creativecommons.org/licenses/ by-nc-nd/4.0/), which permits non-commercial re-use, distribution, and reproduction in any medium, provided the original work is properly cited, and is not altered, transformed, or built upon in any way. The terms on which this article has been published allow the posting of the Accepted Manuscript in a repository by the author(s) or with their consent. 2 M. Fathi and K. Grobys yield misleading results if variance is infinite. Subsequent evidence suggests that many proposed asset pricing factors may suffer from sample-specificity (Chan et al. 2000, Hirsh- leifer 2001, Schwert 2003). Extending this discourse, Grobys (2021) posited that replication failures in financial economics could stem from infinite variances or infinite variance of variances. Importantly, the qualitative implications of these phenomena are effectively identical. This study aims to address a fundamental question: does the population mean of range-based variances in financial mar- kets exist? Here, we focus on range-based realized variances, which serve as ex-post estimators of volatility and approxi- mate the latent conditional variance of returns. Unlike model- based residual variance, which reflects unexplained variation after regression and depends on model assumptions, range- based variance is directly observable and model-free. If the distribution of these range-based variances exhibits extremely heavy tails, this implies that variance estimates themselves are unstable and potentially without finite moments. Conse- quently, the residual variance in linear regression models (e.g. CAPM or Fama-French), which requires the theoretical vari- ance of the inputs to be statistically well-defined, may itself become undefined, thereby invalidating classical inference tools such as standard errors, t-statistics, and R2. This makes the investigation of tail behavior in, for example, range-based variance distributions essential for assessing the reliability of asset pricing models built on OLS regression. To this end, we investigate range-based variances in five major financial asset markets—S&P 500, gold, crude oil, the USD/GBP exchange rate, and Bitcoin—using intraday data. Range-based variances are computed using the Garman-Klass range-based variance estimator. We also fit power-law models to the data using maximum likelihood estimation (MLE) to examine the statistical properties of extreme variance realiza- tions, with a specific focus on the tail behavior of the distribu- tion beyond the range of typical observations. To assess model suitability, we conduct comprehensive goodness-of-fit tests, comparing the power-law hypothesis against other plausi- ble distributions, such as log-normal and exponential models. Given the dependency structures in financial variance data (e.g. volatility clustering), we orthogonalize the range-based variance data through autoregressive modeling and reapply goodness-of-fit tests to the resulting innovation processes. Furthermore, we test for commonalities in the range-based variances of unrelated asset markets by examining whether power-law behavior is statistically invariant across these mar- kets. Finally, we conduct an extensive set of robustness checks to validate our findings, including sample-split tests, the use of alternative volatility measures (e.g. realized variance), the examination of additional asset classes (e.g. bonds), the appli- cation of Bayesian inference to estimate tail exponents and cutoffs, model comparisons using Akaike Information Crite- rion (AIC) alongside the Kolmogorov–Smirnov test, rolling- window estimations, and the fitting of tempered stable and Generalized Beta of the Second Kind (GB2) distributions to assess the stability and generality of the observed power-law behavior. This study makes several significant contributions to the literature, addressing critical methodological, theoreti- cal, and empirical gaps. These contributions advance our understanding of range-based variances in financial markets and their implications for asset pricing and risk management. First, previous research, such as Grobys (2021), relied on the Parkinson variance estimator (Parkinson 1980) to com- pute range-based variances. While this approach has merits, Molnár (2012) highlighted that range-based variance estima- tors, such as those proposed by Garman and Klass (1980), outperform alternatives when accounting for noise in high- frequency data. Chou et al. (2010) also emphasized the informational advantage of intraday price ranges over clos- ing prices. By adopting the Garman-Klass variance estimator, this study mitigates potential biases inherent in the use of inflated variance estimators, as in the study of Grobys (2021). Our methodological adjustment ensures greater precision in capturing range-based variances, offering a robust founda- tion for further analysis. This refinement is essential to derive reliable insights, particularly when assessing heavy-tailed dis- tributions like power laws, and aligns with calls from Molnár (2012) for a more rigorous treatment of variance estimation methods. Second, realized volatility has long been theorized as log-normal, with Andersen et al. (2001) providing seminal empirical support for this assumption. However, evidence challenging this perspective has been mounting. Renó and Rizza (2003), for instance, demonstrated that volatility in Italian futures markets aligns more closely with a Pareto dis- tribution, necessitating the rejection of log-normality. Fathi et al. (2024) further confirmed power law distributions as superior in capturing realized variances in G10 currencies. Our study expands on this work by rigorously comparing the log-normal, exponential, and power-law models using comprehensive goodness-of-fit tests. We provide robust evi- dence that power laws offer the most plausible framework for modeling daily range-based variances too. By reconcil- ing theoretical assumptions with observed data, this study extends Grobys’ (2021) findings, which focused exclusively on power-law validity. Third, Grobys (2021) used a relatively narrow sample, such as range-based variances for gold from August 30, 2000, to March 31, 2021, missing critical periods such as the gold price bubbles of the 1980s. Data omissions of this nature can lead to upward-biased estimates of power-law exponents, particu- larly if extreme events are underrepresented. To address this limitation, we extend the dataset for gold to cover February 7, 1980, to July 31, 2024, more than doubling the observation count to 11 196. This comprehensive data expansion ensures that periods of significant market stress are incorporated, mit- igating biases and strengthening the reliability of our findings. By filling this empirical gap, the study builds a robust foun- dation for analyzing power-law behavior of range-based asset market variances. Fourth, studies have explored common power-law behavior of range-based variances within related asset categories, such as equity factors (Fathi et al. 2025) and foreign exchange mar- kets (Grobys 2024). These works documented universal expo- nents (e.g. α ≈ 2.6 for range-based variances of G10 curren- cies) that imply risk homogeneity within specific asset classes. However, no prior research has tested for such commonality across unrelated markets. This study pioneers cross-sectional risk analysis by examining whether the range-based variances Modeling variance risk in financial markets using power-laws 3 of diverse financial markets, including equities, commodi- ties, exchange rates, and cryptocurrencies are governed by a common power law. The existence of such a commonal- ity would suggest a shared underlying risk dynamic, likely driven by similar systemic or behavioral forces, such as herd- ing behavior. The identification of a cross-market universal exponent would underscore the need for unified risk modeling approaches and contributes a novel dimension to the literature on power-law behavior. Finally, robustness is a critical aspect of econometric anal- ysis, especially when dealing with heavy-tailed distributions. Unlike Grobys (2021), which relied primarily on raw data, we employ autoregressive modeling to account for volatility clustering, a well-documented phenomenon in financial mar- kets. Analyzing the resulting innovation processes confirms that power-law behavior persists even after accounting for dynamic dependencies. Furthermore, our sample-split tests validate the stability of power-law exponents across time periods and subsamples. These methodological advancements align with calls for more robust analytical frameworks in the financial econometrics literature (Mandelbrot 1963, Grobys 2021, Fathi et al. 2024) and enhance the generalizability of our findings. Through these contributions, the study directly engages with and builds upon prior literature, addressing gaps in methodologies (Molnár 2012), theoretical assumptions (Andersen et al. 2001, Renó and Rizza 2003), and empir- ical scope (Grobys 2021, Fathi et al. 2024). By leverag- ing improved variance estimators, expanding datasets, and incorporating cross-sectional and dynamic robustness analy- ses, the research advances our understanding of range-based variances across financial markets. It also underscores the limitations of traditional models, advocating for the integra- tion of power-law frameworks to better capture the realities of extreme market risks. These findings hold critical impli- cations for both academic research and practical applica- tions in risk management, portfolio optimization, and asset pricing. The results reveal that Garman-Klass range-based vari- ances exhibit heavy tails across all five markets, strongly supporting the power-law hypothesis. The MLE analysis of tail exponents confirms that values consistently fall within 2 < α < 3, implying infinite variance. This corroborates ear- lier findings (e.g. Grobys 2021, 2024, Fathi et al. 2024, 2025) and raises critical concerns about traditional econo- metric techniques relying on finite-variance assumptions. Goodness-of-fit tests firmly reject alternative distributions, including log-normal and exponential models, highlighting the systematic underestimation of tail risks by finite-moment frameworks. Moreover, striking evidence supports the exis- tence of a universal cross-sectional power-law exponent (α ≈ 2.8) governing range-based variances across unrelated mar- kets. Robustness checks using autoregressive modeling and sample-split tests confirm these findings, underscoring their consistency across time periods and market conditions. This study provides critical insights into the statistical properties of range-based variances across key financial mar- kets, offering important implications for finance theory and practice. First, the evidence of practically infinite variance challenges traditional econometric methods, such as point estimation and hypothesis testing, which assume finite vari- ance, necessitating the development of alternative method- ologies. Second, the rejection of the log-normal model in favor of power-law distributions highlights the systematic underestimation of tail risks by finite-moment frameworks, emphasizing the need for power-law-based approaches in assessing and managing extreme financial risks. Third, the identification of a universal power-law exponent across unre- lated markets implies limited diversification benefits, as com- mon dynamics across asset variances may amplify systemic risks, urging the adoption of unified risk modeling strategies. Fourth, the robustness of power-law exponents across time periods and subsamples, even after accounting for volatility clustering, confirms the generality and consistency of these findings, supporting their applicability to a wide range of financial contexts. Collectively, these results challenge pre- vailing assumptions underpinning modern portfolio theory and traditional risk management frameworks, reinforcing the necessity for alternative approaches that better capture the realities of extreme financial risks. By addressing method- ological, theoretical, and empirical gaps, this study paves the way for more accurate risk modeling, portfolio optimiza- tion, and the development of financial instruments resilient to market extremes. This study is organized as follows: Section 2 reviews the relevant literature; Section 3 describes the data; Section 4 out- lines the methodology; Section 5 presents the empirical results and key findings; Section 6 discusses their implications; and Section 7 concludes the study. 2. Literature review The modeling of volatility in financial markets has evolved into a rich and multifaceted area of research, resulting in a variety of theoretical and empirical frameworks. These models are designed to capture different aspects of volatil- ity behavior, ranging from short-term clustering to long- memory and heavy-tailed properties. This literature can be grouped into six major research streams: GARCH-type mod- els, stochastic volatility models, realized volatility models, range-based volatility models, multifractal and scaling mod- els, and power-law (namely, tail-index) approaches. 2.1. Generalized Autoregressive Conditional Heteroskedasticity (GARCH) models The Generalized Autoregressive Conditional Heteroskedas- ticity (GARCH) model, introduced by Bollerslev (1986), and its variants have become foundational in volatility model- ing. These models describe volatility as a conditional pro- cess that evolves based on past squared returns and past variances. Key extensions include the Exponential GARCH (EGARCH) model by Nelson (1991) and the Threshold GARCH (TGARCH) model by Zakoian (1994), which allow for asymmetry in volatility responses. Although highly flex- ible, these models rely on parametric assumptions and are typically calibrated on daily or lower-frequency data. They perform well in capturing volatility clustering but struggle with representing extreme tail behavior unless augmented 4 M. Fathi and K. Grobys with fat-tailed distributions for example, Student’s t (Hansen 1994). A key limitation of GARCH-type models is that their parameter estimates can change substantially when under- lying distributional assumptions are relaxed—for example, when the innovation process is altered from a normal to a Stu- dent’s t distribution, or vice versa. In this regard, Mandelbrot (2008) criticized the tendency of researchers to ‘fix’ GARCH models when they failed to perform adequately, instead of abandoning them altogether. 2.2. Stochastic volatility (SV) models In contrast to GARCH models, stochastic volatility (SV) models treat volatility as a latent stochastic process, indepen- dent from return innovations (Taylor 1982, Shephard 2005). These models provide greater flexibility in capturing volatil- ity persistence and allow for richer dynamic structures. Recent developments include score-driven SV models (Creal et al. 2013, Harvey and Palumbo 2019), which use the score of the conditional likelihood to update latent volatility components. SV models are often computationally intensive but can better accommodate long-memory and structural breaks. 2.3. Realized volatility models Realized volatility (RV) models utilize high-frequency intra- day returns to construct ex-post volatility measures, offering a data-driven alternative to conditionally specified models. Andersen et al. (2001) played a pioneering role by formaliz- ing the concept of realized volatility using the sum of squared intraday returns. This approach avoids strong distributional assumptions and allows researchers to observe and model actual volatility over fixed intervals, for example, daily or weekly periods. Subsequent studies have expanded upon this framework to examine properties for example, the distribu- tional characteristics of RV (Barndorff-Nielsen and Shephard 2002), the impact of microstructure noise and asynchronous trading (Zhang et al. 2005), and the decomposition of volatil- ity into continuous and jump components (Barndorff-Nielsen and Shephard 2004). Realized volatility is now a central tool in empirical finance, particularly in measuring and forecasting short-term volatility with high precision. 2.4. Range-based volatility models Range-based models use daily high, low, open, and close prices to construct volatility estimators. The Parkinson (1980) and Garman and Klass (1980) estimators are seminal exam- ples. These methods are shown to be more efficient than squared return-based measures, especially in the presence of microstructure noise and thin trading. Molnár (2012) and Chou et al. (2010) further highlight the robustness of range- based estimators, making them suitable for low-frequency but high-accuracy volatility measurement. 2.5. Multifractal and scaling models Multifractal models attempt to capture the self-similar, long-memory, and scaling properties observed in financial volatility. The Markov-Switching Multifractal (MSM) model developed by Calvet and Fisher (2004) models volatility as a product of volatility components switching between different regimes at multiple frequencies. Earlier work by Mandelbrot et al. (1997) and Müller et al. (1997) emphasizes the hier- archical and fractal nature of market volatility, linking it to turbulence-like behavior in financial time series. 2.6. Power-law and tail-index approaches A distinct stream focuses on the tail behavior of the volatil- ity distribution. Rather than modeling conditional volatility, these approaches examine whether realized volatility or vari- ance follows a power-law distribution with infinite mean or variance. Clauset et al. (2009) offer robust tools for tail- index estimation and goodness-of-fit testing. Recent empir- ical work (e.g. Grobys 2021, 2024) applies these methods to range-based variance data, exploring the implications of heavy-tailed volatility for the validity of classical econometric models for example, OLS regression. While all these models aim to capture features of financial market volatility, they differ significantly in their assump- tions and goals. GARCH and SV models are designed for conditional volatility forecasting, relying on parametric or semi-parametric methods. Realized and range-based models use empirical data to estimate volatility more accurately but often without strong assumptions about underlying distribu- tions. Multifractal models and power-law approaches focus on structural properties like scaling laws and tail behavior. Notably, most traditional models do not explicitly address extremely fat-tailed distributions or long-range dependence in volatility, unless such features are specifically built into the model. The power-law approach is unique in explicitly testing for the existence (or lack) of moments, which has profound implications for econometric modeling and asset pricing. 2.7. Alternative approaches to address heavy tails and long memory in volatility Notably, the literature on financial volatility offers a range of alternative approaches that contribute to understanding heavy-tailed behavior beyond traditional power-law model- ing. For instance, Corsi (2009) introduced the Heterogeneous Autoregressive (HAR) model, which captures long-memory effects in realized volatility by aggregating volatility com- ponents across different time horizons. This approach has proven effective in modeling the persistent nature of volatility without requiring high-frequency data. Similarly, Harvey and Palumbo (2019) proposed a score-driven model for realized volatility that dynamically adapts to changes in the underly- ing data-generating process using a state-space framework. These methods highlight the growing emphasis on flexi- ble and robust volatility modeling frameworks. Furthermore, Barndorff-Nielsen and Shephard (2002) introduced stochas- tic volatility models driven by Lévy processes, allowing for jumps and capturing the non-Gaussian nature of returns. Similarly, Müller et al. (1997) proposed a time-deformation model using multiplicative cascades, revealing scaling laws and volatility components at different frequencies. Engle and Modeling variance risk in financial markets using power-laws 5 Rangel (2008) addressed long-term volatility trends through a spline-GARCH model that combines smooth structural components with short-term GARCH effects. Calvet and Fisher (2004) developed the MSM model, which generates multifractal volatility patterns and long memory via regime- switching cascades. Bollerslev and Todorov (2011) empha- sized jump-tail risk by integrating extreme events into asset price and volatility dynamics, accounting for the role of rare but impactful shocks. Collectively, these models offer alternative perspectives on how volatility evolves over time under non-Gaussian con- ditions, and they share common objectives: capturing fat- tailed returns, long-range dependence, volatility clustering, and asymmetric responses to market shocks. While traditional GARCH-type models, including their extensions, are typi- cally parametric and rely on specific distributional assump- tions, several of the above frameworks (for example, MSM and multiplicative cascades) adopt semi- or non-parametric approaches, allowing for greater flexibility in capturing the empirical irregularities of financial time series. Moreover, models based on realized measures, for example, range- based or high-frequency volatility, circumvent strong distri- butional assumptions by leveraging more granular data. We have now referenced these studies in the revised manuscript to better position our research within this broader aca- demic discourse and to clarify how our focus on power- law behavior and extreme variance dynamics complements, instead of contradicts, the insights offered by these alternative models. While the existing literature offers valuable insights into the dynamics and persistence of volatility, our study diverges by explicitly focusing on the tail behavior of the volatility distribution—measured in terms of range-based variances— itself. Rather than modeling the full distribution or volatility path, this study investigates whether the extreme realiza- tions of the data generating process—often dismissed as noise—may actually be governed by a power-law distribu- tion, suggesting that the tail, instead of the center, contains the meaningful signal. This tail-centric perspective enables us to characterize the statistical properties of extreme vari- ance events, a topic that remains underexplored in much of the existing volatility modeling literature. 3. Data The data for this study were collected from Yahoo Finance. We retrieved daily observations including open, high, low, and closing prices on S&P 500, gold, crude oil, USD/GBP exchange rate, and Bitcoin. Due to data availability, the data sample for the S&P 500 spans from February 19, 1980, to July 31, 2024, providing 11 207 observations. This exten- sive period captures over four decades of market activity, including significant events such as the 1987 market crash, the dot-com bubble, the 2008 financial crisis, and the COVID- 19 pandemic. For gold, the dataset covers a similarly long timeframe, from February 7, 1980, to July 31, 2024, result- ing in 11 196 observations. This period reflects both periods of stability and heightened volatility, such as 2008 financial crisis. The crude oil dataset begins on January 2, 1991, and runs until July 31, 2024, including 8556 observations. Specif- ically, in our study we use data on West Texas Intermediate (WTI) which is one of the main global benchmark prices for crude oil. This period includes major fluctuations crude oil prices due to events such as the Gulf War, the shale revolution, and the COVID-19-related demand shock. The USD/GBP exchange rate data is available from July 25, 2000, to July 31, 2024, yielding 6263 observations. This timeframe captures significant currency market events, including the global finan- cial crisis and Brexit. Finally, Bitcoin data is available from September 17, 2014, to July 31, 2024, with 3603 observa- tions. This relatively short period reflects the cryptocurrency’s promising stage and its rapid growth as a speculative and alternative asset, marked by extreme volatility. To compute the Garman-Klass variance estimator, which combines open and close prices along with daily highs and lows, we follow the methodology proposed by Garman and Klass (1980) and compute daily range-based variances as: σˆ 2GK = 0.5 [ ln ( Ht Lt )]2 − [2ln2 − 1] [ ln ( Ct Ot )]2 , (1) where H represents the daily high price, L denotes the daily low price, O is the daily opening price, and C refers the daily closing price of each financial asset. GARCH models typically estimate conditional variances based on closing prices and thus may fail to fully capture the intraday dynamics of price movements (e.g. Andersen et al. 2003, 2004, Bubák et al. 2011). In contrast, range-based variances—particularly those derived from range-based esti- mators such as the Garman-Klass method—leverage intraday data (high, low, open, and close prices) to provide a more efficient and information-rich measure of daily volatility. As shown by Garman and Klass (1980) and reinforced by Molnár (2012), range-based estimators outperform return-based esti- mators, especially in the presence of microstructure noise and market frictions. Given our focus on modeling tail behavior and extreme variance risk, range-based variances offer clear advantages: they better capture the heavy-tailed nature of financial markets and are less prone to distributional misspec- ification. This allows for a more robust empirical foundation when testing power-law behavior and modeling the variance of variance—a key objective of our study. Table 1 reports the descriptive statistics for the annual- ized daily range-based variances of these assets. The statistics show significant variations in the mean, standard deviation, skewness, and kurtosis, indicating the existence of heavy tails in the distribution of range-based variances, which could indi- cate power-law behavior. The mean of Garman-Klass range- based variance ranges from 0.95 for the USD/GBP exchange rate to 36.18 for Bitcoin, indicating substantial variation in the average level of volatility. Bitcoin’s high Garman-Klass range-based variance mean value reflects its nature as a spec- ulative and emergent asset class, characterized by extreme price fluctuations. Similarly, crude oil exhibits a high mean variance of 14.46. Conversely, traditional asset classes like gold (2.79) and the S&P 500 (2.18) show relatively mod- erate average variances, aligning with their roles as more stable investment assets. The USD/GBP exchange rate, with a 6 M. Fathi and K. Grobys mean of 0.95, suggests that currency markets involving major economies tend to exhibit lower volatility. In addition, all assets exhibit significant positive skewness and excess kur- tosis, suggesting non-normality in their range-based variance distributions. Crude oil exhibits the highest skewness and kur- tosis (skewness: 88.83, kurtosis: 8095.44), while Bitcoin has the lowest skewness and kurtosis (skewness: 9.50, kurtosis: 130.54) among the analyzed assets. 4. Methodology 4.1. Main analysis 4.1.1. Estimating power-law exponents via MLE and comparing model fit. Following earlier studies (Grobys 2021, 2023, 2024, Fathi et al. 2024, 2025), we model the Garman-Klass range-based variances of each asset in our sample using a power-law function, expressed as: p(x) = Cx−α , (2) where C = (α − 1)xα−1MIN with α ∈ {R+|α > 1}, and x = σ 2j,t denotes the respective range-based Garman-Klass variance of financial asset j at time t. Note that x ∈ {R+|xMIN ≤ x < ∞}, where xMIN is the minimum threshold for variance values gov- erned by the power-law distribution, and α represents the tail exponent. For range-based variances exceeding xMIN (X > xMIN ), the conditional expected first moment is defined as: E[X |X > xMIN ] = ∫ ∞ xMIN xp(x)dx = (α − 1) (α − 2)xMIN . (3) The conditional expected second moment, or the variance of range-based variances for X > xMIN , is expressed as: E[X 2|X > xMIN ] = ∫ ∞ xMIN x2p(x)dx = (α − 1) (α − 3)x 2 MIN . (4) Consequently, higher conditional expected moments of order k follow the general form: E[X k|X > xMIN ] = (α − 1) (α − 1 − k)x k MIN . (5) From Equations (3) and (4), it can be concluded that the theoretical mean of range-based Garman-Klass variances is defined only when α > 2, while the existence of the variance for range-based Garman-Klass variances requires α > 3. Next, as suggested by White et al. (2008) and Clauset et al. (2009), we utilized MLE to estimate the power-law tail exponents. The MLE estimator is given by: αˆ = 1 + N ( N∑ i=1 ln ( xi xMIN ))−1 , (6) where αˆ is the estimated tail exponent, N is the number of observations exceeding xMIN , and other terms maintain their previously defined meanings. The selection of xMIN is determined by minimizing the distance between the empirical data and the power-law model (Clauset et al. 2009). This is achieved using the Kolmogorov–Smirnov (KS) distance, D, which measures the maximum difference between the cumu- lative distribution functions (CDFs) of the observed data and the fitted model. The distance D is given by: D = MAXx≥xMIN |S(x) − P(x)|, (7) where S(x) represents the CDF of the empirical data for obser- vations x ≥ xMIN , and P(x) is the CDF of the fitted power- law model. The optimalxMIN is the value that minimizes D. Additionally, as described by Clauset et al. (2009), the standard deviation of the estimated power-law exponents is expressed as: σ = αˆ − 1√ N + O ( 1 N ) . (8) To verify the appropriateness of a power-law fit, we con- ducted a goodness-of-fit test based on the KS distance, D, following the methodology derived in Clauset et al. (2009) which compares the empirical data to synthetic data gener- ated from a power-law distribution using specific xMIN and αˆ values. The null hypothesis of goodness-of-fit test is that the empirical data and the synthetic data generated from a power-law distribution using the estimated xMIN and αˆ val- ues belong to the same distribution, suggesting the power-law model is a plausible fit. While the alternative hypothesis sug- gests that the empirical and synthetic data originate from different distributions, implying that other models should be considered. Thus, we generated synthetic datasets using the estimated parameters xMIN and αˆ in the prior analysis. For each synthetic dataset, the distance (DS) was calculated and compared to the distance (D) from the original dataset. The p-value for the goodness-of-fit test is defined as the proportion of synthetic datasets where DS > D, denoted as #(DS > D)/K, where K is the total number of synthetic datasets. A significance level of 5% was applied, meaning the power-law model is not rejected if #DS > D/K > 0.05. In addition, to ensure robust results, 1000 synthetic datasets were simulated for each financial asset variance. However, non-rejected power-law null models ascertained by means of Clauset et al.’s (2009) goodness-of-fit test do not necessarily mean that power-law distributions are the best fit for the data, as alternative distributions could potentially offer a better match. In addition, as mentioned earlier, prior stud- ies such as Andersen et al. (2001), Andersen et al. (2001a, 2001b) have suggested that realized volatility often aligns closely with a log-normal distribution. To address this, we compare the power-law distributions with the log-normal dis- tributions for each range-based asset variance to determine the most suitable model for the data. As noted by Clauset et al. (2009), comparing p-values of power-law models with those of alternative distributions, such as log-normal or exponen- tial, provides a robust basis for evaluating the appropriateness of the power-law model. Thus, following the methodology of Clauset et al. (2009), we use the same goodness-of-fit test approach applied to the power-law distribution for assess- ing the plausibility of other competing distributions (viz., Modeling variance risk in financial markets using power-laws 7 Table 1. Descriptive statistics. Statistic S&P 500 Gold Crude oil USD/GBP BTC Mean 2.18 2.79 14.46 0.95 36.18 Median 0.95 1.42 6.91 0.60 13.66 Maximum 161.31 200.71 15901.67 135.29 1809.48 Minimum 0.00 0.00 0.00 0.00 0.00 Std. Dev. 4.81 5.54 174.13 2.32 90.06 Skewness 14.32 11.90 88.83 35.59 9.50 Kurtosis 342.75 265.07 8095.44 1864.48 130.54 Observations 11207 11196 8556 6263 3603 Period (MM/DD/YYYY) 02/19/1980 02/07/1980 01/02/1991 07/25/2000 09/17/2014 07/31/2024 07/31/2024 07/31/2024 07/31/2024 07/31/2024 Note: This table presents the descriptive statistics of the annualized daily range-based variance for various assets, including the S&P 500, gold, crude oil, USD/GBP exchange rate, and Bitcoin. The annualized daily range-based variances for each asset calculated using the Garman-Klass methodology (Garman and Klass 1980). The statis- tics reported in this table are sample statistics, which summarize empirical properties observed in the data. While our analysis later demonstrates that the variance of variance is theoretically infinite due to heavy tails (α ∈ (2,3)), these sample-based measures remain meaningful for descriptive purposes. They do not imply the existence of finite theoretical moments. log-normal, exponential). In this case, the null hypothesis assumes that the alternative distribution is a plausible fit for the data. 4.1.2. Assessing robust standard errors and testing for a common power-law component. Motivated by recent research documenting that range-based foreign exchange rate variances are governed by a universal power-law exponent (Grobys 2024), we explore this issue for the range-based variances of our five asset markets. An important difference between Grobys’ (2024) study and our research is that the underlying asset markets in our study are otherwise unrelated, whereas foreign exchange rates are not. Therefore, a common power-law exponent governing the range-based variances of otherwise unrelated asset markets would be a surprising find- ing. To investigate this issue, we employ the test procedure proposed in the study of Grobys (2024). While the MLE proposed by Clauset et al. (2009) remains asymptotically unbiased even under dependent data, the pres- ence of autocorrelation—manifested, for example, through volatility clustering—can lead to an underestimation of stan- dard errors when standard i.i.d. assumptions are applied. This would result in overly narrow confidence intervals and poten- tially misleading inference. To address this issue, we imple- ment a block bootstrap method, as recommended by Godfrey (2009) and employed by Grobys (2024), which preserves tem- poral dependencies within blocks. This adjustment produces more realistic, wider standard errors and thereby ensures that hypothesis testing related to tail exponents is appro- priately conservative; that is, our block bootstrap approach mitigates the issue of downward-biased standard errors due to serial dependence. Hence, to address this issue, we first estimate robust standard errors for each range-based asset variance. This is achieved by employing the blocks bootstrap methodology, as outlined in the study of Grobys (2024). Specifically, a block bootstrap procedure is implemented with the block length m, where E[m] = T1/3 following the recommendation by Godfrey (2009) for variance estimation. From a given data vector of range-based variances, x, blocks of size m are randomly selected, where m follows a geomet- ric distribution, m ∼ GEO(p) with an expected value E[m] = (1 − p)/p. Thus, in the overall data sample, we use p = 0.0435, p = 0.0434, p = 0.0476, p = 0.0526 and p = 0.0625 for the range-based variances of the S&P 500, gold, crude oil, USD/GBP exchange rate, and Bitcoin, respectively. Using this bootstrap method, blocks drawn from the data vector x have varying lengths. These randomly selected blocks m are then stacked into a new vector xb as follows: xbi = ⎡⎢⎢⎢⎣ m1 m2 m3 . . . ⎤⎥⎥⎥⎦ . (9) Again, it is important to note that the lengths of the blocks m1, m2, . . . vary. The process continues until the length of the constructed vector xb exceeds T. Then, we cut off any observations exceeding T , ensuring that the artificial data vec- tor xb matches the length of the original data vector x. By applying this blocks bootstrap, 1000 artificial data vectors[ x1 x2 · · · xB], are generated for each given original data vector. Next, the point estimates αˆ for each bootstrap data vector [ x1 x2 · · · xB], are calculated and stored in a vec- tor [ αˆ1 αˆ2 · · · αˆB] following the method described by Clauset et al. (2009). As a final step, the bootstrapped stan- dard error σˆBOOT is calculated from [ αˆ1 αˆ2 · · · αˆB] for each given data vector of range-based asset variances as: σˆBOOT = √√√√ 1 B B∑ b=1 ( αˆb − ( 1 B B∑ b=1 αˆb ))2 . (10) After obtaining the robust standard errors, we can con- duct the joint test to examine whether the Garman-Klass range-based variances of different assets share a common component, represented by a common power-law exponent. 8 M. Fathi and K. Grobys Identifying such a shared exponent would suggest a unified risk of these assets. It is worth noting that our bootstrap- ping method ensures COV(αˆi, αˆj) = 0∀i = j. Based on this, the covariance matrix for estimated power law exponents is defined as: Σˆαˆ = ⎛⎜⎜⎜⎜⎜⎜⎜⎜⎝ σˆ 2 αˆboot ,S&P500 0 0 0 0 0 σˆ 2 αˆboot ,Gold 0 0 0 0 0 σˆ 2 αˆboot ,Crudeoil 0 0 0 0 0 σˆ 2 αˆboot ,USD/GBP 0 0 0 0 0 σˆ 2 αˆboot ,BTC ⎞⎟⎟⎟⎟⎟⎟⎟⎟⎠ , (11) where, σˆ 2 αˆboot ,S&P500, σˆ 2 αˆboot ,Gold,σˆ 2 αˆboot ,Crudeoil, σˆ 2 αˆboot ,USD/GBP and σˆ 2 αˆboot ,BTC represent the corresponding bootstrapped variances (viz., σˆ 2BOOT ). Then, consistent with Grobys (2024), we define the test statistic λˆ as follows: λˆ = (αˆ − q1)′Σˆ−1αˆ (αˆ − q1), (12) where the covariance matrix Σˆ αˆ is 5 × 5, αˆ represents a 5 × 1 vector of estimated power-law exponents, 1 is a 5 × 1 vector of ones, and q denoted the hypothesized common power-law exponent. The estimated test statistic λˆ in Eq. (10) follows a χ2(5) distribution under the null hypothesis, with the corresponding critical value at the 5% significance level being χ20.95(5) = 11.07. To evaluate whether the Garman- Klass variances of our five asset markets share a common power-law exponent, the test statistic λˆ is iteratively com- puted across the economically significant range of q values, spanning from 1.2 to 3.2 in increments of 0.1. This iterative computation ensures that the analysis may capture the poten- tial existence of a common power-law exponent governing variance risks across the selected assets. 4.2. Additional analysis 4.2.1. Investigating the impact of autocorrelation on tail index estimates in range-based variances. A potential con- cern could be whether the observed power-law behavior in the Garman-Klass range-based variance distributions of our assets is genuine or a statistical effect resulting from autocor- relation. It is well-recognized that time-varying volatility can create patterns in the tails of return distributions that resem- ble power-laws. This phenomenon could lead to inaccurate conclusions, as standard statistical techniques for estimat- ing power laws might incorrectly attribute thick tails to a power-law structure when they are actually produced by a combination of exponential-tail distributions rather than an actual power law. This raises the important question as to whether similar issues affect the estimation of power-law behavior in the range-based variances of our assets? To examine this, we fit autoregressive models of order p to each asset’s Garman-Klass range-based variance. The models are defined as: σˆ 2j,t = β0,j + β1,jσˆ 2j,t−1 + β2,jσˆ 2j,t−2 + . . . + βp,jσˆ 2j,t−p + εj,t, (13) where σˆ 2j,t defines the Garman-Klass range-based variance of asset j at time t, εj,t, represents the innovation pro- cess, and p is the lag-order. The optimal lag order p is determined by analyzing the partial autocorrelation func- tion. This approach is similar to ARCH-type models, where the conditional variance is modeled using the lagged values of range-based variances directly. Next, we use the abso- lute values of the estimated innovation processes of each asset, |εˆS&P500|, |εˆGold |, |εˆCrudeoil|, |εˆUSD/GBP|, |εˆBTC|, to re- evaluate the power-law exponents for each asset’s orthogo- nalized range-based variance process. Then, we carried out Clauset et al.’s (2009) goodness-of-fit test, following the methodology outlined earlier for the absolute values of the innovation processes for each asset variance. This procedure helps to evaluate whether the observed power-law behav- ior persists after accounting for autocorrelation, providing a robust check on the validity of the power-law model for our range-based variances. 4.2.2. Evaluating the reliability of power-law exponents: evidence from sample-split tests. To further validate the robustness of the observed power-law behavior in the inno- vation processes of the range-based variances, we per- form a sample-split test. We divide the residual series for each asset, obtained from the autoregressive models (viz., |εˆS&P500|, |εˆGold |, |εˆCrudeoil|, |εˆUSD/GBP|, |εˆBTC|), into two non-overlapping subsamples of equal length. By esti- mating the power-law exponents for the absolute values of the residuals in each subsample separately, we aim to assess the consistency of the power-law characteristics across dif- ferent sample episodes. This approach allows us to examine whether the power-law behavior persists independently within each subsample, mitigating concerns about potential sample- specificity. Additionally, we applied the goodness-of-fit tests to the absolute values of the residuals in each subsample to evaluate the validity of the power-law model in both cases. 4.2.3. Alternative measures of asset market variance: A realized variance approach. While the main analysis in this paper is based on the Garman-Klass range-based estimator, concerns can be raised about the potential limitations of rely- ing on a single volatility measure. Specifically, it has been noted that different volatility estimators may capture different aspects of return dynamics, and even the most efficient esti- mators remain noisy approximations of true volatility (Shu and Zhang 2006, Molnár 2012, Be˛dowska-Sójka and Kliber 2021). To address this concern and assess the robustness of our findings, we extend our analysis by computing weekly realized variances based on the sum of squared daily returns over five-day intervals. This return-based volatility measure offers a conventional benchmark and helps assess whether the heavy-tailed behavior observed under the Garman-Klass esti- mator persists under a different volatility estimation frame- work. The daily return of asset market i at time t is computed as follows: RETi,t = 100 (Pi,t+1 − Pi,t)Pi,t , (14) Modeling variance risk in financial markets using power-laws 9 where Pi,t denotes the corresponding daily closing price. Next, we compute the weekly realized variance of each five assets as the sum of the squared returns over five business days according to the following Equation: RVj,t = σ 2i,j = ∑ t∈j (RETi,t)2, (15) where j indicates the week and t ∈ j indicates the correspond- ing trading days in the respective week. We then estimate power-law exponents for the tail behavior of the weekly real- ized variances of each asset as described in section 4.1.1. Finally, to evaluate the appropriateness of the power-law model relative to alternative specifications, we compare the empirical data with synthetic data generated from power-law, log-normal, and exponential distributions for each realized asset variance to determine the most suitable model for the weekly data as described in section 4.1.2. 4.2.4. Analysis of additional asset classes: evidence from bonds. To test the generalizability of our findings across asset classes, we extend the analysis to U.S. government bond yields. Specifically, we include range-based bond yield vari- ance data for the 2-year, 5-year, and 10-year Treasury yields, computed using the Garman-Klass estimator on daily high, low, open, and close prices. This addition allows us to explore whether power-law behavior is also present in fixed income markets, which are often considered less volatile than equi- ties or cryptocurrencies. Following the methodology outlined in Section 4.1.1, we estimate power-law exponents for each bond series and conduct comparative goodness-of-fit tests using synthetic data from power-law, log-normal, and expo- nential distributions. This enables a robust evaluation of the tail behavior in bond market volatility. 4.2.5. Alternative estimation of power-law exponents and cutoffs: A Bayesian inference approach. While our main results rely on the widely used Kolmogorov–Smirnov-based method for estimating power-law parameters, one might be concerned that the selection of the threshold xmin—a crucial modeling choice—could influence the inferred tail behavior. To address this potential sensitivity and enhance the credibil- ity of our findings, we perform a robustness check using an alternative, Bayesian estimation framework. This approach follows the methodology introduced by Virkar and Clauset (2014) and Grigaityte and Atwal (2019), which allows for the simultaneous estimation of the power-law exponent αˆ and the lower-bound threshold xmin. Unlike frequentist meth- ods that rely on point estimates and deterministic threshold selection, the Bayesian approach yields posterior distribu- tions for both parameters. Using Markov Chain Monte Carlo (MCMC) sampling, we generate estimates of αˆ and xmin. This method not only enhances transparency around thresh- old selection but also enables a more informative inference by incorporating parameter uncertainty. To formally describe the Bayesian approach, we consider a sample of continuous observations x1, x2, . . . , xN , each satisfying xi ≥ xmin. The con- tinuous power-law distribution is defined by the following probability density function: p(x|α, xmin) = α − 1 xmin ( x xmin )−α , for x ≥ xmin, α > 1. (16) Given N observations above a threshold xmin, the likelihood (L) of the data is: L(α | x, xmin) = N∏ i=1 α − 1 xmin ( xi xmin )−α . (17) Taking the log, the log-likelihood becomes: log L(α | x, xmin) = Nlog(α − 1) − Nlogxmin − α N∑ i=1 log ( xi xmin ) . (18) To complete the Bayesian model, we specify prior distribu- tions where for the exponent α, we use a weakly informative uniform prior: α ∼ Uniform(1.01, 5). (19) In addition, for the threshold xmin, we assume a discrete uniform prior over a range of candidate values: xmin ∈ {x(1), x(2), . . . , x(k)}, P(xmin) = 1k , (20) where, P(xmin) denotes the prior probability assigned to each candidate threshold, reflecting an assumption of equal plau- sibility before observing the data. In this setting, the joint posterior distribution over α and xmin is given by: P(α, xmin|x) ∝ L(α|x, xmin) · P(α) · P(xmin), (21) where P(α) is the prior distribution over the power-law expo- nent α. Because xmin is discrete and α is continuous, we marginalize over the possible values of xmin by computing: (1) The posterior distribution P(α|x, xmin) for each candi- date xmin. (2) The marginal likelihood (also known as the model evidence): Z(xmin) = ∫L(α|x, xmin) · P(α)dα. (22) (3) The posterior over xmin using Bayes’ rule: P(xmin|x) = Z(xmin)∑K k=1Z(xk) . (23) The overall posterior for α can be computed as a weighted mixture of posteriors conditioned on each xmin, using the pos- terior probabilities P(xmin, x) as weights. Finally, the power- law exponent and threshold are being estimated by performing MCMC sampling for each candidate xmin(k), compute their marginal likelihoods, derive posterior probabilities for each threshold, and calculate the weighted posterior mean. 10 M. Fathi and K. Grobys 4.2.6. An alternative model comparison metric: the Akaike Information Criterion. In addition to the previously reported goodness-of-fit tests based on Clauset et al. (2009), we incorporated a formal model comparison framework to strengthen the case for power-law behavior relative to alter- native models. Specifically, we compare the power-law, log- normal, and exponential distributions using the AIC. Unlike the standalone goodness-of-fit tests, AIC provides a system- atic approach that balances model fit and complexity through the MLE. For each candidate distribution, model parameters are estimated via MLE. The power-law parameters—the scal- ing exponent α̂ and the lower bound xˆmin—are determined following Clauset et al. (2009). The log-likelihood for each model is computed based on the corresponding probability density function and the AIC is calculated as: AIC = 2k − 2log(Lˆ), (24) where k represents the number of parameters (with k = 1 for the power-law and exponential models, and k = 2 for the log- normal model) and log(Lˆ), is the maximum log-likelihood of the fitted model. 4.2.7. Rolling window estimations for retrieving time- Varying power-law exponents. One could argue that finan- cial volatility may be subject to structural changes over time, potentially challenging the validity of our results from the sample-split tests. In this context, implementing rolling win- dow estimations offers a complementary approach to assess the stability of the estimated power-law exponents. To address this issue, we performed rolling window estimations as fol- lows: for each series of range-based variances, we applied a rolling window of 500 observations and estimated the MLE- based power-law exponent as outlined in Section 4.1.1. The window advances in steps of 50 observations, starting from t = 500 until t = T. 4.2.8. Estimation of tempered stable distributions. A finding of a shared power-law exponent governing all range- based asset variances would represent a significant theoretical insight. While it may be argued that certain financial assets exhibit distinct distributional characteristics—particularly in response to varying market conditions—such evidence would motivate deeper investigation. Therefore, in this robustness check, we assess whether the tempered stable distribution can offer a more nuanced understanding of the common dynamics underlying range-based variances. A tempered sta- ble distribution generalizes classical stable distributions (e.g. Lévy) by introducing exponential tempering to the heavy tails. This modification allows for finite moments while preserving power-law behavior over intermediate ranges. Formally, the tail decays as P(X > x) ∼ x−αe−λx, (25) where α ∈ (0, 2) controls the tail heaviness and λ > 0 governs the rate of tempering. This construction retains the scaling properties of stable laws while ensuring statistical regularity in the extreme tails. Tempered stable distributions are particu- larly useful for modeling empirical phenomena that exhibit approximate power-law scaling but experience faster-than- power decay in the tails, as frequently observed in financial, physical, and biological systems. In applications involving multiple datasets that share a common power-law exponent yet display distinct truncation or decay rates, the tempered stable framework provides a coherent modeling strategy. By fixing a common α across data series and allowing data series- specific tempering parameters λi, one can jointly capture universal scaling behavior and dataset-specific tail dynamics. This approach is particularly suited for comparative studies or hierarchical modeling contexts such as volatility dynamics. To estimate the parameters via maximum likelihood, we numer- ically invert the characteristic function, as the probability density function lacks a closed-form expression. Parameters are then estimated by maximizing the log-likelihood using numerical optimization techniques (BFGS). 4.2.9. Comparison between power-law distributions and the generalized beta of the second kind (GB2) family. One might argue that the comparison of power-law distributions with the exponential and log-normal distributions introduced in Section 4.1.1 does not provide an informative benchmark for evaluating the appropriateness of the power-law model. As a consequence, we consider the Generalized Beta of the Second Kind (GB2) family as a more flexible alternative, as it accommodates both fat tails and finite second moments across a wide range of parameterizations. The GB2 is a highly flexible, four-parameter family of continuous proba- bility distributions that can represent a wide variety of shapes, including both heavy-tailed and light-tailed forms. It is fre- quently employed in the analysis of skewed or heavy-tailed data. The GB2 distribution is defined as: f (x; a; b; p; q) = a(x/b) ap−1 bB(p, q)[1 + (x/b)a]p+q , (26) where B(p, q) denotes the Beta function, a > 0 is the tail parameter controlling the tail heaviness, b > 0 is a scale parameter, and p > 0, q > 0 determine the shape of the dis- tribution, including skewness. In Table A.5 of the appendix, we list several important subfamilies of the GB2 distribution, highlighting its role as a ‘distributional umbrella’ that encom- passes many well-known distributions. Asymptotically, the GB2 exhibits power-law decay under certain conditions: P(X > x) ∼ x−aq, (27) as x → ∞, indicating that the product aq serves as an effec- tive tail exponent, thereby linking the GB2 to power-law modeling. For a valid comparison between the GB2 and power-law models, both must be estimated using maximum likelihood, and the likelihoods must be evaluated over the same data range—specifically, the range above the power- law cutoff xMIN . Accordingly, we restrict the GB2 model to the same data ranges used for the power-law estimation, as reported in table 3. We then compute the log-likelihood values and the corresponding AIC for each model. Modeling variance risk in financial markets using power-laws 11 5. Results 5.1. Main results This section presents the findings from our analysis, applying the methodological approach outlined in Section 4.1 to the selected financial assets: S&P 500, gold, crude oil, USD/GBP exchange rate, and Bitcoin. The analysis begins by modeling the Garman-Klass range- based variances using the power-law function described in Section 4.1.1. Table 2 highlights the share of cumulative total variance contributed by the top 1% and top 20% of obser- vations for each asset, revealing a significant concentration. A small proportion of extreme observations accounts for a disproportionate share of total variance, consistent with the heavy-tailed behavior expected under a power-law distribu- tion. For instance, among the top 1% of observations, crude oil exhibits the highest concentration at 27.23%, followed by Bitcoin (20.03%), S&P 500 (16.10%), gold (15.05%), and USD/GBP exchange rate (14.39%). Similarly, for the top 20%, Bitcoin leads with 72.03%, followed by crude oil (66.15%), S&P 500 (65.91%), gold (63.00%), and USD/GBP exchange rate (55.45%). These results underscore the pres- ence of extreme events driving range-based variances, partic- ularly in crude oil and Bitcoin, aligning with the Pareto 80/20 rule. Next, a power-law distribution is fitted to the annual- ized daily range-based variances of each asset using MLE, as detailed in Section 4.1.1 and proposed by White et al. (2008) and Clauset et al. (2009). Table 3 reports the tail exponents, ranging from αˆ = 2.62 for crude oil to αˆ = 2.94 for the S&P 500, indicating that variance of variance is infinite for all assets. Such heavy tails invalidate founda- tional assumptions underlying conventional statistical tools like t-tests, which depend on finite variances. Standard met- rics derived from ordinary least squares regression, such as t-statistics, become unreliable due to their sample-specific dependency. The evidence confirms that heavy-tailed distri- butions render traditional econometric methods inadequate, consistent with critiques by Grobys (2021) and Fathi et al. (2025). To further validate the plausibility of the power-law model, goodness-of-fit tests (Clauset et al. 2009) are applied to assess the plausibility of power-law and alternative distributions, including log-normal and exponential models. Table 4 reports p-values for these tests, demonstrating strong support for the power-law model, as all p-values for the null hypothe- sis of a power-law distribution exceed 5%. In contrast, the log-normal model is rejected outright (p = 0.00) for all assets except Bitcoin, where the power-law model (p = 0.66) out- performs log-normal (p = 0.52). The exponential model is conclusively ruled out, with p-values equating to zero for all cases. These results collectively confirm the superior fit of the power-law model to the data. A joint test, described in Section 4.1.2, examines whether the Garman-Klass range-based variances exhibit a shared component characterized by a common power-law exponent. To account for dependencies such as volatility clustering, block bootstrap methods are employed, yielding robust stan- dard errors. Table 5 illustrates the increased standard devia- tions derived from bootstrapping. For instance, the standard deviation for S&P 500 increases from 0.05 to 0.20, and for gold, from 0.06 to 0.14. These adjustments underscore the critical importance of incorporating dependency structures in estimating tail exponents. Subsequently, the joint test evaluates whether the assets share a common power-law exponent. From table 6, the test statistic λˆ, computed iteratively over a range of α-values Table 2. Share of the top 1% and top 20%. Metric S&P 500 Gold Crude oil USD/GBP BTC Top 1% Share 16.10% 15.05% 27.23% 14.39% 20.03% Top 20% Share 65.91% 63.00% 66.15% 55.45% 72.03% Note: This table displays the share of the cumulative total represented by the top 1% and top 20% of the distribution for the annualized daily Garman-Klass range-based variance across the S&P 500, gold, crude oil, the USD/GBP exchange rate, and Bitcoin. Table 3. Estimation of power-law exponents for annualized daily range-based variance. Metric S&P 500 Gold Crude oil USD/GBP BTC αˆ 2.94 2.74 2.62 2.78 2.84 xˆMIN 4.59 6.62 23.97 1.09 187.71 D 0.02 0.02 0.02 0.01 0.04 NPL 11.93% 8.33% 10.30% 22.77% 3.41% Std. Dev 0.05 0.06 0.05 0.04 0.17 Observations 11207 11196 8556 6263 3603 Period (MM/DD/YYYY) 02/19/1980 02/07/1980 01/02/1991 07/25/2000 09/17/2014 07/31/2024 07/31/2024 07/31/2024 07/31/2024 07/31/2024 Note: This table presents the estimation results for power-law exponents applied to the annualized daily range-based variances of the S&P 500, gold, crude oil, the USD/GBP exchange rate, and Bitcoin. The parameter αˆ signifies the estimated tail exponent, while σˆ denotes the estimated standard deviation. The lower threshold xˆMIN is identi- fied through the optimized Kolmogorov-Smirnov distance (D), following the procedure described by Clauset et al. (2009). The reported xˆMIN corresponds to the optimal distance D. Additionally, the fraction of observations obtained by the power-law process, denoted as NPL is provided for each asset. 12 M. Fathi and K. Grobys Table 4. Goodness-of-fit tests for the annualized daily range-based variance. Asset Power-law Log-normal Exponential S&P 500 0.25 0.00 0.00 Gold 0.39 0.00 0.00 Crude oil 0.40 0.00 0.00 USD/GBP 0.46 0.00 0.00 BTC 0.66 0.52 0.00 Note: This table presents the results of the goodness-of-fit tests conducted using the Kolmogorov-Smirnov (KS) method to assess whether the empirical data and synthetic data generated from a power-law distribution, specified by xmin and α, originate from the same underlying distribution (column 2). Additionally, the table includes the results of goodness-of-fit tests for log-normal and exponential distributions in columns 3 and 4, respectively. The null hypothesis for these tests states that the empirical data and the synthetic data generated from the specified distributions (log-normal and exponential) are consistent, suggesting that each distribution provides an acceptable fit to the observed data. from 1.2 to 3.2, fails to reject the null hypothesis of a shared exponent within the economically important range of 2.5 < α < 3.1. This finding suggests a unified component influencing variances. Notably, our results suggest that αˆ ≈ 2.8 produces the highest p-value (p = 0.87), which indicates that the second moment of the common power law govern- ing the variance risk of unrelated financial asset markets is, statistically, infinite. 5.2. Results from robustness checks To ensure the observed power-law behavior is not an artifact of statistical effects such as autocorrelation, robustness checks are performed. As discussed in Section 4.2.1, autoregressive models of order p are fitted to each asset’s Garman-Klass range-based variance. Table 7 reports the point estimates, confirming significant autocorrelation across all assets, with partial autocorrelation functions indicating orders of p = 2, 6, 1, 4 for S&P 500, gold, crude oil, USD/GBP exchange rate, and Bitcoin, respectively. Descriptive statistics for the result- ing innovation processes, detailed in table 8, reveal extremely heavy tails. For instance, kurtosis values range from 133.83 for Bitcoin to 8053.8 for crude oil, reinforcing the persistence of extreme values even after autoregressive adjustments. Table 9 presents power-law exponents derived from the absolute innovation processes, closely matching those in table 3, further substantiating the presence of genuine power-law behavior. The goodness-of-fit tests applied to the innova- tion processes (table 10) corroborate these findings, failing to reject the power-law null hypothesis for all assets except gold, where the evidence is slightly weaker. Log-normal and exponential distributions are once again rejected outright (p = 0.00) across all assets. Next, sample-split tests are conducted to validate the robustness of power-law characteristics across subsamples. Table 11 provides power-law exponent estimates for non- overlapping subsamples, demonstrating consistent statistical Table 5. Descriptive statistics for the power-law exponents estimated using block bootstrap methods. Statistic αˆbootS&P 500 αˆ boot Gold αˆ boot Crudeoil αˆ boot USD/GBP αˆ boot BTC Mean 2.95 2.74 2.61 2.81 2.33 Median 2.95 2.74 2.59 2.79 2.27 Maximum 3.57 3.37 3.48 3.75 3.44 Minimum 2.38 2.37 2.26 2.38 1.91 Std. Dev 0.20 0.14 0.17 0.18 0.25 Skewness 0.01 0.31 1.01 0.88 1.63 Kurtosis 2.80 3.42 4.94 4.51 5.93 Jarque-Bera (JB) 1.70 23.08 326.81 225.13 802.29 (p-value) JB 0.41 0.00 0.00 0.00 0.00 Observations 11207 11196 8557 6263 3603 Note: This table reports the block bootstrap methodology test to the annualized daily range-based variance for the S&P 500, gold, crude oil, USD/GBP exchange rate, and Bitcoin. The block bootstrap procedure is implemented with the block length m, where E[m] = T1/3 following the recommendation by Godfrey (2009) for variance estimation. From the given data vector x, blocks of size m are randomly selected, where m follows a geometric distribution m ∼ GEO(p) with an expected value E[m] = (1−p)p . Thus, in the overall data sample, we use p = 0.0435, p = 0.0434, p = 0.0476, p = 0.0526 and p = 0.0625 for the range-based variances of the S&P 500, gold, crude oil, USD/GBP exchange rate, and Bitcoin, respectively. Using this bootstrap method, blocks drawn from the data vector x have varying lengths. The randomly selected blocks, m, are then stacked into a new vector xbi = ⎡⎢⎢⎣ m1 m2 m3 . . . ⎤⎥⎥⎦ where the lengths of the blocks m1, m2, . . . vary. The process continues until the length of the constructed vector xb exceeds T. Then, we cut off any observations exceeding T , ensuring that the artificial data vector xb matches the length of the original data vector x. By applying this blocks bootstrap, 1000 artificial data vectors [ x1 x2 · · · xB], are generated for each given original data vector. Next, the point estimates αˆ for each bootstrap data vector [ x1 x2 · · · xB], are calculated and stored in a vector [ αˆ1 αˆ2 · · · αˆB] following the method described by Clauset et al. (2009). The block bootstrap method is described in detail in Section 3.1.2. Modeling variance risk in financial markets using power-laws 13 Table 6. Testing for a common power-law exponent governing from annualized daily range-based variance. q λˆ p-value 1.2 390.68 0.00 1.3 342.52 0.00 1.4 297.53 0.00 1.5 255.73 0.00 1.6 217.11 0.00 1.7 181.68 0.00 1.8 149.42 0.00 1.9 120.34 0.00 2 94.45 0.00 2.1 71.74 0.00 2.2 52.21 0.00 2.3 35.87 0.00 2.4 22.70 0.00 2.5 12.72 0.03 2.6 5.91 0.31 2.7 2.29 0.81 2.8 1.85 0.87 2.9 4.60 0.47 3 10.52 0.06 3.1 19.63 0.00 3.2 31.92 0.00 Note: This table presents the results of a joint test designed to assess whether the range-based variances of S&P 500, gold, crude oil, USD/GBP exchange rate, and Bitcoin share a common component. To investigate whether a common component underlies the power-law behavior of these assets, the following test statistic is utilized: λˆ = (αˆ − q1)′Σˆ−1αˆ (αˆ − q1), where Σˆαˆ is the estimated 5 × 5 covariance matrix, αˆ is a 5 × 1 vector of estimated power-law exponents, 1 is a 5 × 1 vector of ones, and q represents the hypothesized power-law exponent. The test statistic λˆ follows a χ2(5) distribution under the null hypothesis. The statistic is iteratively calculated across the economically significant range of q = (1.2, 1.3, . . . , 3.1, 3.2). Bolded values in the results represents statistical significance at the 5% level. properties across both panels. Exceptions include the S&P 500 in Panel A and gold in Panel B, where the goodness-of-fit test rejects the power-law hypothesis. Nevertheless, the log- normal and exponential models remain consistently rejected across all subsamples (p = 0.00). This consistency across tests and subsamples strengthens the conclusion that gen- uine power-law behavior governs the range-based variances of these financial assets. Building on the robustness analysis outlined in Section 4.2.3, we report the results derived from the weekly real- ized variance estimator in this section. Table A.1 presents the descriptive statistics for the weekly realized variances across the five asset classes. Consistent with our earlier findings, the realized variance distributions exhibit substantial skew- ness and excess kurtosis—hallmarks of heavy-tailed behavior. Weekly realized variances for crude oil and Bitcoin, in par- ticular, display considerable variation and extreme upper-tail values, further supporting the presence of volatility cluster- ing and rare but impactful events. We then estimate power- law exponents for the tail behavior of the weekly realized variances using the MLE approach proposed by Clauset et al. (2009). As reported in Table A.2 the estimated expo- nents all lie within the critical range 2 < αˆ < 3, confirming that the variance of variance remains theoretically unde- fined for all asset markets. These results align closely with those obtained using the Garman-Klass estimator, indicating that our primary conclusions are not sensitive to the choice of volatility measure. To further validate the plausibility of the power-law model under this alternative volatility mea- sure, we conduct KS goodness-of-fit tests for power-law, log-normal, and exponential distributions (Table A.3). Based on the results in Table A.3, the power-law model remains statistically plausible for all assets (p-values > 0.05), except gold, while the exponential model is decisively rejected across all the assets. Additionally, although the log-normal model is also accepted for all assets (p-values > 0.05), the power-law model generally provides a superior fit, as reflected in higher p-values—except in the case of gold. In summary, the persistence of power-law behavior under both range-based and return-based volatility estimators con- firms the robustness of our core results. These findings also underscore the broader insight that volatility risk in financial markets may be governed by a common, heavy- tailed structure—regardless of the specific estimation method employed. Additionally, as described in Section 4.2.4, we extended the analysis to incorporate government bond markets in order to examine whether the statistical properties of realized vari- ances also apply to U.S. Treasury yields with 2-year, 5-year, and 10-year maturities. This extension allows us to assess the broader applicability of power-law behavior to fixed income markets. The results of this extended analysis are presented in tables 12–14. As shown in table 12, the descriptive statis- tics for the range-based variances of the 2-year, 5-year, and 14 M. Fathi and K. Grobys Table 7. Point estimates for autoregressive models of for annualized daily range-based variance. βˆ0,j βˆ1,j βˆ2,j βˆ3,j βˆ4,j βˆ5,j βˆ6,j R2 σˆ 2S&P500,t 0.49 ∗∗∗ 0.42∗∗∗ 0.10 0.08∗∗∗ 0.18∗∗∗ 0.42 (12.27) (44.89) (10.33) (7.81) (18.90) σˆ 2Gold,t 0.74 ∗∗∗ 0.36∗∗∗ 0.07∗∗ 0.03∗ 0.11∗∗∗ 0.08∗∗∗ 0.09∗∗∗ 0.30 (7.33) (8.71) (2.53) (1.69) (2.64) (2.95) (3.58) σˆ 2Crudeoil,t 13.34 ∗∗∗ 0.08∗∗∗ 0.01 (6.09) (3.37) σˆ 2USD/GBP,t 0.48 ∗∗∗ 0.16∗∗∗ 0.13∗∗ 0.09∗∗∗ 0.11∗∗∗ 0.10 (3.52) (3.11) (2.11) (3.29) (3.12) σˆ 2BTC,t 20.35 ∗∗∗ 0.44∗∗∗ 0.19 (10.73) (9.93) ∗ , ∗∗ and ∗∗∗ indicate statistical significance at the 10%, 5% and 1% levels, respectively. Note: This table presents the estimated coefficients from autoregressive models of order P fitted to the annu- alized daily range-based variance for the S&P 500, gold, crude oil, the USD/GBP exchange rate, and Bitcoin (BTC). The models are specified as σˆ 2j,t = β0,j + β1,jσˆ 2j,t−1 + β2,jσˆ 2j,t−2 + . . . + βp,jσˆ 2j,t−p + εj,t, where εj,t is the innovation process, j refers to the respective asset. The optimal lag order P is determined by the partial autocorrelation function. The reported coefficients include the intercept βˆ0,j and autoregressive terms (β0,j, β1,j, β2,j, . . . , βp,j). The t-statistics are reported in parentheses. Table 8. Descriptive statistics for the innovations process of annualized daily range-based variance. Statistic εˆS&P500 εˆGold εˆCrudeoil εˆUSD/GBP εˆBTC Mean 0.00 0.00 0.00 0.00 0.00 Median − 0.40 − 0.63 − 7.14 − 0.24 − 14.29 Maximum 139.02 163.67 15818.40 132.17 1750.00 Minimum − 70.72 − 56.60 − 1048.3 − 21.49 − 676.04 Std. Dev 3.68 4.64 173.62 2.20 80.99 Skewness 11.83 10.18 88.46 38.32 8.33 Kurtosis 377.28 250.67 8053.8 2145.90 133.83 Observations 11203 11190 8555 6259 3602 Note: This table summarizes the descriptive statistics for the innovation process εi,t of the annualized daily range-based variance for the S&P 500, gold, crude oil, the USD/GBP exchange rate, and Bitcoin (BTC), as reported in Table 3. The innovation process of annual- ized daily range-based variances for all five key assets is estimated by running autoregressive models of order p: σˆ 2j,t = β0,j + β1,jσˆ 2j,t−1 + β2,jσˆ 2j,t−2 + . . . + βp,jσˆ 2j,t−p + εj,t, where εj,t is the innovation process, p is the lag-order, j refers to the respective asset. The optimal lag order p is determined through an analysis of the partial autocorrelation function. 10-year Treasury yields exhibit high levels of skewness and kurtosis, indicating heavy-tailed distributions similar to those observed in the five major financial asset markets. The power- law estimation results (table 13) show tail exponents of U.S. Treasury yields with 2-year, 5-year, and 10-year maturities, all falling within the critical interval 2 < αˆ < 3. This implies that the second moment of range-based variances for these bonds remains undefined, in line with our findings for major financial asset markets. In addition, goodness-of-fit test results in table 14 further confirm the plausibility of the power- law model for bond market variances. The null hypothesis of a power-law distribution is not rejected for any of the three bond series (p-values: 0.94, 0.54, and 0.58, respec- tively). In contrast, the log-normal and exponential models are decisively rejected for all three bond series. These results provide compelling evidence that power-law behavior is a persistent and robust characteristic of range-based variances across diverse financial markets—including government bond markets. This suggests that heavy-tailed variance risk may be a universal feature of modern financial systems, further rein- forcing the need to reconsider traditional models based on finite-variance assumptions. Next, to complement our primary estimation strategy and assess the robustness of the results, we apply a Bayesian inference approach to estimate the power-law parameters across all assets as described in section 4.2.5. Table 15 reports the Bayesian estimates of the power-law exponent αˆ, the inferred lower-bound xˆmin. The results confirm that all Modeling variance risk in financial markets using power-laws 15 Table 9. Estimating power-law exponents for the absolute amount of the innovation processes of the annualized daily range-based variance. Metric |εˆS&P500| |εˆGold | |εˆCrudeoil| |εˆUSD/GBP| |εˆBTC| αˆ 2.54 2.46 2.41 2.40 2.30 xˆMIN 3.71 2.81 40.90 1.12 56.55 D 0.02 0.02 0.03 0.03 0.04 NPL 5.51% 14.59% 2.58% 6.65% 9.69% Std. Dev 0.06 0.04 0.09 0.07 0.07 Observations 11203 11190 8555 6259 3602 Note: This table presents the estimation results for the power-law exponents applied to the absolute amount of the residual series of annualized daily range-based variances for the S&P 500, gold, crude oil, the exchange rate of the U.S. dollar and Bitcoin (BTC). The innovation process for each asset is obtained by using the following autoregressive model: σˆ 2j,t = β0,j + β1,jσˆ 2j,t−1 + β2,jσˆ 2j,t−2 + . . . + βp,jσˆ 2j,t−p + εj,t where εj,t is the innovation process, p is the lag-order, and j refers to the respective asset. The parameter αˆ represents the estimated tail exponent, while σˆ denotes the estimated standard deviation. The lower threshold xˆMIN is determined using the optimized Kolmogorov- Smirnov distance (D) approach, as outlined by Clauset et al. (2009). The reported xˆMIN corresponds to the optimal distance D. Additionally, the fraction of observations governed by the power-law process, denoted as NPL is provided for each asset. Table 10. Goodness-of-fit tests for the innovation pro- cesses of annualized daily range-based variance. Power-law Log-normal Exponential |εˆS&P500| 0.82 0.00 0.00 |εˆGold | 0.02 0.00 0.00 |εˆCrudeoil| 0.88 0.00 0.00 |εˆUSD/GBP| 0.22 0.00 0.00 |εˆBTC| 0.07 0.00 0.00 Note: This table presents the results of the goodness-of- fit tests conducted using the Kolmogorov-Smirnov (KS) method to assess whether the empirical data and synthetic data generated from a power-law distribution, specified by xmin and α, originate from the same underlying dis- tribution (column 2). Additionally, the table includes the results of goodness-of-fit tests for log-normal and exponen- tial distributions in columns 3 and 4, respectively. The null hypothesis for these tests states that the empirical data and the synthetic data generated from the specified distributions (log-normal and exponential) are consistent, suggesting that each distribution provides an acceptable fit to the observed data. estimated exponents fall within the range typically associ- ated with heavy-tailed behavior 2 < αˆ < 3, consistent with our main findings based on Clauset et al.’s (2009) approach. Importantly, the 95% credible intervals for αˆ across all assets are relatively narrow, indicating precise and stable estimates of the power-law exponent under the Bayesian framework. In addition, figure 1 visualizes the posterior distribution of the power-law exponent αˆ for each asset, illustrating the concen- tration of posterior mass around the mean and the precision of the Bayesian estimates. This robustness reinforces the cred- ibility of our earlier conclusions regarding tail behavior in financial range-based variances. To complement the goodness-of-fit tests, we conducted a formal model comparison using the AIC, as described in Section 4.2.6. The AIC accounts for both model fit and complexity, with lower values indicating a more parsimo- nious and better-fitting model. Each distribution—power-law, log-normal, and exponential—was fitted to the empirical data using MLE. Table 16 reports the AIC values for each model across the five asset classes. In all cases, the power-law dis- tribution produces the lowest AIC, indicating a superior fit to the extreme tails of the range-based variance distributions when compared to the log-normal and exponential alterna- tives. These findings provide consistent and robust evidence that power-law models are better suited to capture the heavy- tailed nature of range-based asset variance distributions. The results reinforce the conclusions drawn from the goodness-of- fit tests and underscore the relevance of power-law behavior in modeling extreme financial events. Furthermore, the results of the rolling-window estimations outlined in Section 4.2.7 are presented in Table A.4 while Figure A.1 visualizes the evolution of the time-varying power- law exponents over the sample period. Notably, the S&P 500 dataset contains T = 11 207 observations of range-based variance, allowing for the estimation of N = 205 power-law exponents. In contrast, for Bitcoin, only N = 53 exponents could be estimated due to the shorter sample. A comparison between the full-sample estimates (table 3) and the rolling window medians (Table A.4) reveals close alignment. For example, Table A.4 reports median power-law exponents of 2.73 and 2.77 for the S&P 500 and Gold, respectively, while the corresponding full-sample point estimates in table 3 are 2.94 and 2.74. Similar consistency is observed across other range-based variance series. Finally, a visual inspection of Figure A.1 suggests that the time-varying power-law expo- nents derived from the rolling window estimation appear sta- tionary, reinforcing the conclusion that the exponent estimates are stable—that is, converging. Next, we fit tempered stable distributions, as outlined in Section 4.2.8, to the range-based variances of the S&P 500, USD/GBP exchange rate, gold, crude oil, and Bitcoin, initially fixing the power-law exponent αi = α = 2.8. The results 16 M. Fathi and K. Grobys Table 11. Sample split analysis of power-law exponents for the innovation processes of annual- ized daily range-based variance. Panel A: Estimation results for the first subsample of power-law exponents and related metrics. Metric S&P 500 Gold Crude oil USD/GBP BTC αˆ 2.18 2.48 2.29 2.30 2.44 xˆMIN 0.65 2.82 33.07 1.24 26.81 (p-value)power−law 0.00 0.27 0.28 0.96 0.10 D 0.02 0.02 0.04 0.03 0.04 NPL 48.79 16.93 5.10 5.08 18.17 Std. Dev 0.02 0.05 0.09 0.10 0.08 (p-value)log−normal 0.00 0.00 0.00 0.00 0.00 (p-value)exponential 0.00 0.00 0.00 0.00 0.00 Observations 5601 5595 4277 3129 1801 Panel B: Estimation results for the second subsample of power-law exponents and related metrics. Metric S&P 500 Gold Crude oil USD/GBP BTC αˆ 2.76 2.31 2.49 2.47 2.21 xˆMIN 3.13 1.04 13.42 0.92 49.73 (p-value)power−law 0.85 0.00 0.39 0.12 0.40 D 0.02 0.02 0.04 0.04 0.03 NPL 7.10 43.66 5.35 9.90 14.71 Std. Dev 0.09 0.03 0.10 0.08 0.07 (p-value)log−normal 0.00 0.00 0.00 0.00 0.00 (p-value)exponential 0.00 0.00 0.00 0.00 0.00 Observations 5602 5595 4278 3130 1801 Note: This table presents the sample-split test of power-law exponents applied to the residual series of annualized daily range-based variances, with results divided into two subsamples: Panel A reports the first subsample, and Panel B reports the second subsample. The parameter αˆ represents the estimated tail exponent, while σˆ denotes the estimated standard deviation. The lower threshold xˆMIN is determined using the optimized Kolmogorov-Smirnov distance (D) approach, as outlined by Clauset et al. (2009). The reported xˆMIN corresponds to the optimal distance D. Additionally, the fraction of observations obtained by the power-law process, denoted as NPL is provided for each asset. Table 12. Descriptive statistics of range-based U.S. Treasury yield variances. Statistic UST 2Y UST 5Y UST 10Y Mean 51.41 35.47 21.70 Median 14.50 12.28 6.97 Maximum 3042.13 2975.99 5041.98 Minimum 0.00 0.00 0.00 Std. Dev. 157.46 114.40 120.87 Skewness 10.02 12.83 28.73 Kurtosis 133.63 233.29 1087.59 Observations 3010 3033 2818 Period (MM/DD/YYYY) 10/26/2012 10/26/2012 08/26/2013 07/31/2024 07/31/2024 07/31/2024 Note: This table presents the descriptive statistics of the annualized daily range- based variances for U.S. Treasury yields with maturities of 2 years (UST 2Y), 5 years (UST 5Y), and 10 years (UST 10Y). The variances are calculated using the Garman-Klass estimator, based on daily high, low, open, and close yields. reveal that λi could be estimated only for the range-based vari- ance for the USD/GBP exchange rate, while the estimation failed to converge for the remaining series. To validate this finding, we repeated the estimation without constraining α. Again, both αi and λi could be reliably estimated only for the range-based variance for the USD/GBP series. This pat- tern is highly informative. It suggests that, for most assets, the data do not exhibit statistically significant exponential trunca- tion in the tails. Rather, the data appear to conform closely to a pure power law over the observed range, rendering the inclusion of a tempering parameter either superfluous or numerically unstable. This interpretation aligns with earlier empirical evidence of strong power-law decay in these series. The consistent failure to estimate λi even when αi is uncon- strained further supports the conclusion that the tail behavior is well captured by a standard power law without the need for exponential tempering. Finally, restricting the data to x ≥ xMIN , we fit GB2 mod- els to the range-based variances and computed the AIC values, as outlined in Section 4.2.9. Table 17 presents the Modeling variance risk in financial markets using power-laws 17 Table 13. Estimation of power-law exponents for range-based U.S. Treasury yield variances. Metric UST 2Y UST 5Y UST 10Y αˆ 2.30 2.31 2.32 xˆMIN 87.42 56.62 27.45 D 0.02 0.03 0.02 NPL 12.82% 12.43% 14.05% Std. Dev 0.07 0.07 0.07 Observations 3010 3033 2818 Period (MM/DD/YYYY) 02/19/1980 02/07/1980 01/02/1991 07/31/2024 07/31/2024 07/31/2024 Note: This table presents the estimation results for power-law exponents applied to the annualized daily range-based variances of U.S. Treasury yields with 2- year (UST 2Y), 5-year (UST 5Y), and 10-year (UST 10Y) maturities. The parameter αˆ signifies the estimated tail exponent, while σˆ denotes the estimated standard deviation. The lower threshold xˆMIN is identified through the opti- mized Kolmogorov-Smirnov distance (D), following the procedure described by Clauset et al. (2009). The reported xˆMIN corresponds to the optimal distance D. Additionally, the fraction of observations obtained by the power-law process, denoted as NPL is provided for each bond. Figure 1. Posterior distribution of the power-law exponent αˆ, estimated using Bayesian inference. Note: This figure shows the posterior distribution of the power-law exponent αˆ, estimated using Bayesian inference for values greater than the threshold xˆmin. The red dashed line indicates the posterior mean of αˆ. 18 M. Fathi and K. Grobys Table 14. Goodness-of-fit tests for range-based U.S. Treasury yield variances. Asset Power-law Log-normal Exponential UST 2Y 0.94 0.04 0.00 UST 5Y 0.54 0.00 0.00 UST 10Y 0.58 0.00 0.00 Note: This table reports the results (viz., p-values) of the goodness-of-fit tests for the annualized daily range-based variances of U.S. Treasury yields with maturities of 2 years (UST 2Y), 5 years (UST 5Y), and 10 years (UST 10Y). The tests are conducted using the Kolmogorov- Smirnov (KS) method to assess whether the empirical data and the synthetic data—generated from a power- law distribution specified by xmin and α —originate from the same underlying distribution (column 2). Addition- ally, the table includes the results of goodness-of-fit tests for log-normal and exponential distributions in columns 3 and 4, respectively. The null hypothesis for these tests states that the empirical data and the synthetic data gen- erated from the specified distributions (log-normal and exponential) are consistent, suggesting that each distri- bution provides an acceptable fit to the observed data. log-likelihoods and AIC values for both the power-law models—using the parameterizations from table 3—and the GB2 models restricted to the equivalent data ranges. Param- eter estimates for the GB2 models are provided in Table A.6 in the appendix. The results in table 17 indicate that the AIC values are consistently lower for the power-law models across all range-based variance series, regardless of the underlying asset class. This finding reinforces the results of our main analysis, further supporting the conclusion that power-law models are particularly well suited for capturing the dynamics of range-based variances in financial assets. 6. Discussion In this study we explored whether the Garman-Klass range- based variances of five key financial assets—S&P 500, gold, crude oil, USD/GBP exchange rate, and Bitcoin— follow power-law distributions and examined the broader implications of these findings for financial modeling and risk analysis. By utilizing the Garman-Klass variance esti- mator and conducting robust statistical tests, we provided a comprehensive analysis of power-law behavior in range- based variances of otherwise unrelated asset markets. Unlike previous studies that primarily focused on absolute returns or single-asset analyses, as highlighted by Lux and Alfarano (2016), we employ the Garman-Klass range-based vari- ances, which provide greater sensitivity to intraday price dynamics and mitigate estimation bias. This methodolog- ical improvement enables us to capture the complexity of volatility structures across different asset classes more effectively. First, we confirmed the heavy-tailed nature of range-based variances across all five assets, consistent with power-law behavior. Data on range-based asset variances exhibited Ta bl e 15 . B ay es ia n es tim at io n fo rp ow er - la w ex po ne nt s. M et ric S& P 50 0 G ol d Cr ud e o il U SD /G BP B TC αˆ 2. 89 2. 74 2. 60 2. 73 2. 58 xˆ M IN 7. 06 8. 86 37 .0 2 2. 49 13 5. 77 St d. D ev 0. 08 0. 07 0. 08 0. 09 0. 12 CI (αˆ ) 9 5% (2. 73 ,3 .0 5) (2. 60 ,2 .8 8) (2. 44 ,2 .7 6) (2. 55 ,2 .9 1) (2. 34 ,2 .8 2) O bs er va tio ns 11 20 7 11 19 6 85 56 62 63 36 03 Pe rio d (M M /D D/ YY YY ) 02 /1 9/ 19 80 02 /0 7/ 19 80 01 /0 2/ 19 91 07 /2 5/ 20 00 09 /1 7/ 20 14 07 /3 1/ 20 24 07 /3 1/ 20 24 07 /3 1/ 20 24 07 /3 1/ 20 24 07 /3 1/ 20 24 N ot e: Th is ta bl e pr es en ts th e es tim at io n re su lts fo r po w er - la w ex po ne nt s u sin g B ay es ia n ap pr oa ch ap pl ie d to th e an n u al iz ed da ily ra n ge -b as ed v ar ia nc es o f th e S& P 50 0, go ld ,c ru de o il, th e U SD /G BP ex ch an ge ra te ,a n d B itc oi n (B TC ). Th e ta bl e in cl ud es th e po ste rio rm ea n s o ft he ex po ne nt αˆ th e es tim at ed lo w er - bo un d th re sh ol d xˆ m in , th e st an da rd de vi at io ns o ft he po ste rio r di str ib ut io ns fo rαˆ , an d th e 95 % cr ed ib le in te rv al sf or αˆ . Th e n u m be ro fo bs er va tio ns an d th e co rr es po nd in g sa m pl e pe rio d fo re ac h as se ta re al so pr ov id ed . Modeling variance risk in financial markets using power-laws 19 Table 16. AIC model comparison. Asset Power-law∗ Log-normal Exponential S&P 500 6346.79 36394.16 39929.06 Gold 5433.58 42811.29 45326.74 Crude oil 7592.74 58142.31 62010.57 USD/GBP 3044.78 10295.22 11925.14 BTC 1518.62 31259.29 33017.41 ∗Bold values indicate the best-fitting model by AIC. Note: This table presents the AIC values for power-law, log-normal, and exponential distributions fitted to the annu- alized range-based variances for each five assets, including the S&P 500, gold, crude oil, USD/GBP exchange rate, and Bitcoin (BTC). Each model is estimated using maximum likelihood estimation methods. The AIC is calculated as AIC = 2k − 2log(Lˆ), where k is the number of estimated parameters and log(Lˆ) is the maximum log-likelihood of the model. Lower AIC values indicate a better trade-off between model fit and complexity. Table 17. Comparison between power-law and GB2 functions. Range-based variance Log-likelihood Power law (AIC) Log-likelihood GB2 (AIC) S&P 500 –1134.28 –3176.18 (2268.57) (6360.37) Gold –953.78 –2714.71 (1907.56) (5437.42) Crude oil –996.82 –3805.39 (1993.63) (7618.79) USD/GBP –1397.88 –1528.31 (2795.76) (3064.61) Bitcoin –114.42 –782.70 (228.85) (1573.40) Note: We compare the GB2 distribution with our power-law models by fitting the GB2 function to the data range x ≥ xmin for each respective data vector, as reported in Table 3. The values of the log-likelihood functions are computed, along with the cor- responding AIC. This table reports the log-likelihood and AIC values for both the power-law models—using the parameteriza- tions documented in Table 3—and the GB2 models, which are estimated over the same data ranges defined by the respective power-law cutoffs. significant skewness and kurtosis, with extreme values con- centrated in the upper tails of the variance distributions. For example, Bitcoin exhibited the highest mean value of Garman-Klass range-based variance, reflecting its speculative and volatile nature, while the USD/GBP exchange rate dis- played relatively lower mean value of Garman-Klass range- based variance. Second, power-law fitting through MLE verified that tail exponents for all assets fell within the range 2 < α < 3, implying that the variance-of-variance is infinite. Interest- ingly, these findings contrast with literature from the early 1990s, as discussed by Lux and Alfarano (2016), which focused on the tail behavior of distributions and suggested that financial asset returns exhibit power-law behavior with finite variances. This result has profound implications, as it chal- lenges traditional econometric assumptions that rely on finite variances for hypothesis testing. The estimated power-law exponent α has direct implications for economic risk. Lower values of α indicate heavier tails in the return distribution, implying a higher probability of extreme losses. This sug- gests a more volatile and risk-prone market environment. For instance, when α lies between 2 and 3, the theoretical mean of the variance process is defined but higher-order moments such as variance or skewness are undefined, posing chal- lenges for traditional risk models. When α approaches 2 or below, the theoretical mean of the variance process becomes infinite, highlighting severe tail risk. Hence, the power-law exponent serves as a quantifiable indicator of the intensity of tail events, informing asset pricing, risk management, and regulatory capital frameworks. This finding also aligns with the previous literature (Grobys 2021, 2023, 2024, Fathi et al. 2024, 2025) which documented tail exponents for realized variances or range-based variances are within the range 2 < α < 3, indicating that the variance- of-variances is infinite. In the presence of infinite variance-of- variance, extreme observations can disproportionately influ- ence the variance estimate, leading to results that may appear significant in one sample but fail to replicate in others. As highlighted by Fama (1963), if variances are undefined, results derived from traditional statistical methods based on the concept of correlation (i.e. t-statistics) become inherently sample-dependent. According to Grobys (2021), unstable sec- ond moments lead to t-statistics that are highly dependent on the specific sample used. Thus, this study, in line with Grobys (2021), argues that the replication failures observed in financial research may stem from the infinite variance of vari- ances, which leads to traditional statistical methods becoming unreliable. Third, our results derived from goodness-of-fit tests strongly supported power-law distributions as plausible mod- els for the range-based variances of these assets. These findings align with Mandelbrot’s (1963) early insights that financial asset returns may lack finite variances, and strength- ens the notion that traditional statistical tools, such as OLS and other moment-dependent methods, may be inadequate for modeling financial data governed by power-law distributions. Fourth, our analysis comprehensively compared power- law distributions with alternative models (viz., log-normal and exponential). The goodness-of-fit test strongly supported power-law models as the most plausible representation of the data for all assets, with alternative distributions being consistently rejected. This finding strengthens the case for power-law distributions as a fundamental characteristic of financial asset variances. The rejection of log-normal and exponential models aligns with Taleb’s (2020) view on the necessity of adopting power-law frameworks for analyzing financial data, particularly in the context of extreme events and tail risks. Furthermore, these findings contrast with the literature suggesting that realized asset volatility follows a log-normal distribution (e.g. Andersen et al. 2001, 2001a, 2001b), but they align with studies proposing that a Pareto dis- tribution better describes volatility in financial markets (Renò and Rizza 2003, Grobys 2021, 2023, 2024, Fathi et al. 2024, 2025). The presence of power-law behavior in financial returns and volatility carries profound implications for risk measure- ment frameworks such as Value-at-Risk (VaR) and stress test- ing. Traditional models, including those based on log-normal 20 M. Fathi and K. Grobys or Gaussian assumptions, underestimate the probability and magnitude of extreme events due to their thin-tailed nature. Empirical evidence from Bouchaud (2001) and Gencay et al. (2003) demonstrates that financial markets often exhibit heavy-tailed distributions, particularly in the tails of return and variance distributions. This tail heaviness implies that large losses occur far more frequently than predicted by nor- mal models, which can severely compromise the reliability of VaR estimates. Gencay and Selçuk (2004) show that applying Extreme Value Theory (EVT) and the Generalized Pareto Dis- tribution (GPD) provides a more robust framework for model- ing the tails of financial distributions. These approaches allow for better estimation of risk at high quantiles and facilitate more realistic stress testing scenarios. Moreover, power-law scaling properties are indicative of long memory and clustered volatility, which further challenge the assumptions of inde- pendence and stationarity in standard risk models. As such, incorporating power-law features into risk modeling frame- works is essential for accurately capturing the true nature of financial market risk, particularly in the context of systemic shocks. Fifth, to assess whether the range-based variances of the five assets share a common component, we implemented a joint test hypothesizing a common power-law exponent. Rec- ognizing the limitations of standard errors derived under inde- pendence assumptions of Clauset et al. (2009), we employed block bootstrap techniques following Grobys (2024) to com- pute robust standard errors. In line with the study of Grobys (2024), the bootstrap procedure showed that robust standard errors were significantly larger than the ones derived in the study of Clauset et al. (2009). Using the robust standard error, we carried out conjoint tests, iteratively computed across a range of hypothesized exponents (q). Our findings suggest strong evidence for a shared power-law exponent governing the range-based variances of these assets. Specifically, the null hypothesis of a shared exponent could not be rejected for the range 2.5 < q < 3.1, suggesting a unified compo- nent influencing variance risks across the S&P 500, gold, crude oil, USD/GBP exchange rate, and Bitcoin. This finding aligns with theoretical expectations of power-law behavior and underscores the interconnected nature of these assets within a broader market framework. Next, to address the potential impact of autocorrelation on tail index estimates, we applied autoregressive models of order p to the Garman-Klass range-based variances of each asset. The results indicate significant autocorrelation patterns in the processes generating range-based variances. The descriptive statistics indicates that even after accounting for higher-order autocorrelation, the innovation processes of the annualized daily range-based variances for all the assets display extremely heavy tails. We then used the absolute values of the estimated inno- vation processes for each asset to estimate the power-law exponents. Notably, the estimated exponents closely align with those observed for the original data, providing strong evi- dence for the presence of genuine power-law behavior. Next, we performed Clauset et al.’s (2009) goodness-of-fit test, for the absolute values of the innovation processes of each asset. This procedure helps to evaluate whether the observed power- law behavior persists after accounting for autocorrelation, providing a robust check on the validity of the power-law estimates for our range-based variances. The goodness-of-fit test indeed indicated that the power-law null hypothesis holds for all innovation processes except gold, providing additional confirmation of the consistent power-law characteristics in the range-based variances. Additionally, the goodness-of-fit test results, with zero p-values across all cases, reject log-normal and exponential distributions, favoring the power-law distri- bution instead. These results support recent findings of Fathi et al. (2025) documenting that realized variances of Fama- French factors are subject to power-law behavior even after controlling for autocorrelations. Finally, to validate the robustness of the power-law expo- nents, we performed sample-split tests by dividing the resid- ual series for each asset into two equal-length subsamples. The consistency of tail exponents across both subsamples demonstrated that power-law behavior persists independently of sample-specific effects. Additionally, the goodness-of-fit test supports the power-law model in both subsamples for most assets, further enhancing the reliability of our findings. A reader could be concerned about the robustness of tail estimation due to limited data in the extreme upper tail. How- ever, the tail sample sizes in our study consistently exceed the empirical threshold suggested by Clauset et al. (2009), who argue that reliable estimation of power-law parameters typically requires a minimum of approximately 50 tail obser- vations. For instance, in the case of Bitcoin, the number of observations in the power-law tail constitutes at least 3.41% of the total sample size of 3603—yielding more than 120 observations. Thus, the sample size in the tail comfortably meets the minimum requirement for credible tail exponent estimation using maximum likelihood techniques. Another potential concern relates to whether our range- based volatility estimates sufficiently address market micro structure noise. In this regard, we note that our analysis is based on daily OHLC data and employs the Garman– Klass estimator, which has been identified as one of the most efficient and least noisy among range-based volatility measures (Molnár 2012). Importantly, microstructure noise is particularly problematic in high-frequency data, whereas range-based estimators constructed from daily prices are con- siderably less susceptible to such distortions. Thus, while no volatility estimator is entirely free from noise, the estimator we use represents a methodologically sound and practically robust choice that mitigates the typical concerns associated with market microstructure effects. Finally, the potential influence of structural market fea- tures, such as liquidity and trading hours, on tail behavior merits attention. However, our results suggest that markets with widely varying characteristics—such as Bitcoin (traded 24/7 with relatively lower trading volume) and the S&P 500 (traded during regular hours with much higher volume)— exhibit statistically indistinguishable power-law exponents in range-based volatility. This finding is robust across asset classes, including commodities and FX. Additionally, trad- ing volume in the S&P 500 has shown a clear upward trend over the sample period (see Figure A.2). As such, any attempt to split the sample into low- and high-liquidity regimes would effectively mirror our existing time-based splits (as in table 11), which already reveal stable tail exponent estimates Modeling variance risk in financial markets using power-laws 21 across subperiods. These observations suggest that structural differences such as trading hours and volume have limited impact on the tail behavior we observe. 7. Conclusion This study presents novel evidence on the variance risk of financial markets by modeling range-based variances using the Garman-Klass variance estimator and analyzing their properties under hypothesized power-law distributions. Focusing on five key financial assets—S&P 500, gold, crude oil, the USD/GBP exchange rate, and Bitcoin—our find- ings reveal that range-based variances exhibit heavy tails, aligning closely with power-law distributions. These findings carry profound implications for financial modeling and risk management. First, the estimated tail exponents for range-based vari- ances consistently fall within the range 2 < α < 3, indicating characteristics of infinite variance-of-variance. As noted by Mandelbrot (1967), such variance can be considered ‘so large that it may in practice be assumed infinite.’ This challenges the foundational assumption of finite variances in traditional econometric models and highlights the limitations of statisti- cal tools like t-tests, which are derived from the ordinary least squares (OLS) regression framework. The evidence of sam- ple specificity, as highlighted by Fama’s (1963) reflections on Mandelbrot’s infinite variance hypothesis, underscores the inadequacy of moment-dependent statistical methods for handling the heavy-tailed distributions common in financial markets. Second, goodness-of-fit tests consistently reject alternative distributions, including the widely used log-normal model, in favor of power-law distributions, emphasizing the inade- quacy of conventional models in capturing the tail risks that define financial markets. Additionally, evidence derived from blocks bootstrap underscores the importance of accounting for dependency structures when estimating the standard devia- tions of tail exponents. Our joint test for a common power-law exponent reveals statistical invariance in power-law behavior for the range (2.5 < q < 3.1) across all five assets, suggest- ing a universal factor governing variance risk. Autoregressive modeling of range-based variances further corroborates the persistence of power-law behavior even after adjusting for autocorrelation. Lastly, sample-split tests confirm the sta- bility of power-law exponents over different time periods, reinforcing the robustness of the findings. The empirical finding that a common power-law expo- nent governs realized variances across otherwise unre- lated asset markets finds theoretical support in Gabaix (2016). In his comprehensive review, Gabaix (2016) outlines mechanisms—such as proportional random growth processes and universality—that explain why power-law distributions emerge across diverse economic systems. These mechanisms suggest that similar tail behaviors can arise in structurally dif- ferent markets due to shared statistical properties, rather than asset-specific dynamics. Gabaix (2016) also emphasizes the concept of universality, where systems governed by different microfoundations may exhibit identical scaling laws at the macro level. This perspective aligns with the observation in our study that assets as diverse as crude oil, gold, and Bitcoin exhibit similar tail exponents. Furthermore, his discussion of aggregation effects and macroeconomic granularity provides a theoretical foundation for interpreting power-law regular- ities as emergent from complex but generalizable dynam- ics. Hence, our finding is consistent with and theoretically grounded in the framework developed by Gabaix (2016). Furthermore, our results underscore the necessity of adopt- ing alternative statistical frameworks for analyzing financial data and managing extreme risks. Effectively infinite vari- ances and the rejection of log-normality suggests that reliance on traditional tools can significantly underestimate tail risks, particularly in environments characterized by extreme mar- ket events. The heavy-tailed nature of range-based variances has substantial implications for portfolio optimization, asset pricing, and risk management, necessitating frameworks that effectively account for the inherent extremity in financial data. By advancing a robust foundation for understanding variance risks across diverse asset classes, this study contributes to a more accurate and unified approach to financial modeling, particularly under infinite variance phenomena. Despite its significant contributions, this study has limi- tations that merit consideration. For example, the analysis is limited to five financial assets—S&P 500, gold, crude oil, USD/GBP exchange rate, and Bitcoin—which constrains the generalizability of the findings to other asset classes or broader market conditions. Expanding the analysis to a wider range of assets and markets would enhance the applicability and breadth of the results. 8. Open Scholarship This article has earned the Center for Open Science badge for Open Data. The data are openly accessible at (https://zenodo.org/records/15707465 (DOI 10.5281/zen- odo.15707464). Acknowledgments The authors are grateful to the three anonymous referees for their valuable comments. Data availability statement The dataset has been deposited in Zenodo and is pub- licly available at: https://zenodo.org/records/15707465 (DOI 10.5281/zenodo.15707464). Disclosure statement No potential conflict of interest was reported by the author(s). 22 M. Fathi and K. Grobys References Andersen, T. G., Bollerslev, T., Diebold, F. X. and Ebens, H., The dis- tribution of realized stock return volatility. J. financ. econ., 2001, 61(1), 43–76. Andersen, T. G., Bollerslev, T., Diebold, F. X. and Labys, P., Mod- eling and forecasting realized volatility. Econometrica, 2001a, 71(2), 579–625. Andersen, T. G., Bollerslev, T., Diebold, F. X. and Labys, P., The dis- tribution of realized exchange rate volatility. J. Am. Stat. Assoc., 2001b, 96(453), 42–55. Andersen, T. G., Bollerslev, T. and Meddahi, N., Analytical evalu- ation of volatility forecasts. Int. Econ. Rev. (Philadelphia), 2004, 45, 1079–1110. Andersen, T. G., Bollerslev, T., Diebold, F. X. and Labys, P., Mod- eling and forecasting realized volatility. Econometrica, 2003, 71, 579–625. Barndorff-Nielsen, O. E. and Shephard, N., Econometric analysis of realized volatility and its use in estimating stochastic volatil- ity models. J Roy Statist Soc Series B: Stat Method, 2002, 64(2), 253–280. Barndorff-Nielsen, O. E. and Shephard, N., Power and bipower vari- ation with stochastic volatility and jumps. Journal of Financial Econometrics, 2004, 2(1), 1–37. Bedowska-Sojka, B. and Kliber, A., Information content of liquidity and volatility measures. Physica A, 2021, 563, 125436. Bollerslev, T., Generalized autoregressive conditional heteroskedas- ticity. J. Econom., 1986, 31(3), 307–327. Bollerslev, T. and Todorov, V., Tails, fears, and risk premia. J Fin, 2011, 66(6), 2165–2211. Bouchaud, J. P., Power laws in economics and finance: Some ideas fromphysics. Quant Fin, 2001, 1(1), 105. Bubák, V., Kocˇenda, E. and Žikeš, F., Volatility transmission in emerging European foreign exchange markets. J Bank Fin, 2011, 35, 2829–2841. Calvet, L. E. and Fisher, A. J., How to forecast long-run volatility: Regime switching and the estimation of multifractal processes. J Fin Econom, 2004, 2(1), 49–83. Chan, L. K. C., Karceski, J. and Lakonishok, J., New paradigm or same old hype in equity investing? Fin Anal J, 2000, 56, 23–36. Chou, R. Y., Chou, H. C. and Liu, N., Range volatility models and their application in finance. In Handbook of Quantitative Finance and Risk Management, edited by C. F. Lee, A. C. Lee and J. Lee, pp. 1273–1281, 2010 (Springer: Boston, MA). Clauset, A., Shalizi, C. R. and Newman, M. E. J., Power-law distributions in empirical data. SIAM Rev., 2009, 51, 661–703. Corsi, F., A simple approximate long-memory model of realized volatility. J Fin Econ, 2009, 7(2), 174–196. Creal, D., Koopman, S. J. and Lucas, A., Generalized autoregressive score models with applications. J. Appl. Econ., 2013, 28(5), 777– 795. Engle, R. F. and Rangel, J. G., The spline-GARCH model for low- frequency volatility and its global macroeconomic causes. Rev Fin Stud, 2008, 21(3), 1187–1222. Fama, E. F., Mandelbrot and the stable Paretian hypothesis. J. Bus., 1963, 36(4), 420–429. Fama, E. F. and French, K. R., The cross-section of expected stock returns. J Fin, 1992, 47, 427–465. Fama, E. F. and French, K. R., Common factors in the returns on stocks and bonds. J. Financ. Econ., 1993, 33, 3–56. Fama, E. F. and French, K. R., International tests of a five-factor asset pricing model. J. Financ. Econ., 2017, 123(3), 441–463. Fama, E. F. and French, K. R., Choosing factors. J. Financ. Econ., 2018, 128(2), 234–252. Fama, E. F. and French, K. R. The value premium. Fama-Miller Working Paper No. 20–01, 2020. Fama, E. F. and French, K. R., A five-factor asset pricing model. J. Financ. Econ., 2015, 116, 1–22. Fathi, M., Grobys, K. and Äijö, J., A common component of Fama and French factor variances. Nor Am J Econ Fin, 2025, 75, 102292. Fathi, M., Grobys, K. and Kolari, J. W., On the realized risk of for- eign exchange rates: A fractal perspective. J Risk Fin Manag, 2024, 17(2), 79. Feng, G., Giglio, S. and Xiu, D., Taming the factor zoo: A test of new factors. J Fin, 2020, 75, 1327–1370. Gabaix, X., Power laws in economics: An introduction. J Econ Pers, 2016, 30(1), 185–206. Garman, M. B. and Klass, M. J., On the estimation of security price volatilities from historical data. J. Bus., 1980, 53(1), 67–78. Gencay, R. and Selçuk, F., Extreme value theory and value-at-risk: relative performance in emerging markets. Int. J. Forecast., 2004, 20(2), 287–303. Gençay, R., Selçuk, F. and Ulugülyag˘ci, A., High volatility, thick tails and extreme value theory in value-at-risk estimation. Ins: Mathe Econ, 2003, 33(2), 337–356. Godfrey, L., Bootstrap Tests for Regression Models, 2009 (Palgrave Macmillan: Basingstoke). Grigaityte, K. and Atwal, G. Bayesian inference of power law distributions. bioRxiv, 2019. Grobys, K., What do we know about the second moment of financial markets? Int Rev Fin Anal, 2021, 78, 101891. Grobys, K., Correlation versus co-fractality: Evidence from foreign exchange rate variances. Int Rev Fin Analy, 2023, 86, 102531. Grobys, K., A universal exponent governing foreign exchange rate risks. Int Rev Fin Anal, 2024, 95(Part B), 103422. Hansen, B. E., Autoregressive conditional density estimation. Int. Econ. Rev. (Philadelphia), 1994, 35(3), 705–730. Harvey, A. and Palumbo, D. Score-driven models for realized volatil- ity. Cambridge Working Papers in Economics: 1950, 2019. Hirshleifer, D., Investor psychology and asset pricing. J Fin, 2001, 56, 1533–1597. Lintner, J., The valuation of risk assets and the selection of risky investments in stock portfolios and capital budgets. Rev. Econ. Stat., 1965, 47(1), 13–37. Lux, T. and Alfarano, S., Financial power-laws: Empirical evidence, models, and mechanisms. Chaos, Solitons Fractals, 2016, 88, 3– 18. Mandelbrot, B., The variation of certain speculative prices. J. Bus., 1963, 36, 394–419. Mandelbrot, B., The variation of some other speculative prices. J. Bus., 1967, 40, 393–413. Mandelbrot, B. B., The (Mis)behavior of Markets: A Fractal View of Risk, Ruin, and Reward ., 2008 (Profile Books). Mandelbrot, B. B., Fisher, A. and Calvet, L. A multifractal model of asset returns. Cowles Foundation Discussion Paper No. 1164, 1997. Markowitz, H., Portfolio selection. J. Finance., 1952, 7(1), 77–91. Molnár, P., Properties of range-based volatility estimators. Int Rev Fin Anal, 2012, 23, 20–29. Mossin, J., Equilibrium in a capital asset market. Econometrica, 1966, 35, 768–783. Müller, U. A., Dacorogna, M. M., Davé, R. D., Olsen, R. B., Pictet, O. V. and Von Weizsäcker, J. E., Volatilities of different time resolutions—analyzing the dynamics of market components. J Emp Fin, 1997, 4(2-3), 213–239. Nelson, D. B., Conditional heteroskedasticity in asset returns: A new approach. Econometrica, 1991, 59(2), 347–370. Parkinson, M., The extreme value method for estimating the variance of the rate of return. J. Bus., 1980, 53(1), 61–65. Renò, R. and Rizza, R., Is volatility lognormal? Evidence from Italian futures. Physica A, 2003, 322, 620–628. Schwert, G. W., Anomalies and market efficiency. In Handbook of the Economics of Finance, edited by G. M. Constantinides, M. Harris and R. M. Stulz, Vol. 1, pp. 939–974, 2003 (Elsevier Science B.V.: Amsterdam). Shu, J. and Zhang, J. E., Testing range estimators of historical volatility. J Fut Mark, 2006, 26, 297–313. Sharpe, W. F., Capital Asset Prices: A Theory of Market Equilibrium under Conditions of Risk. J. Finance., 1964, 19(3), 425–442. Shephard, N., Stochastic Volatility: Selected Readings, 2005 (Oxford University Press: Oxford). Modeling variance risk in financial markets using power-laws 23 Taleb, N. N., Statistical Consequences of Fat Tails: Real World Preasymptotics, Epistemology, and Applications. Papers and Commentary, 2020 (STEM Academic Press: New York). Taylor, S. J., Financial returns modelled by the product of two stochastic processes—A study of daily sugar prices 1961–79. In Time Series Analysis: Theory and Practice, edited by O. D. Anderson, Vol. 1, pp. 203–226, 1982 (North-Holland). Treynor, J. L. Toward a theory of market value of risky assets. Available at SSRN: https://ssrn.com/abstract=628187, 1962. Virkar, Y. S. and Clauset, A., Power-law distributions and binned empirical data. Ann. Appl. Stat., 2014, 8(1), 89–119. White, E., Enquist, B. and Green, J. L., On estimating the exponent of power-law frequency distributions. Ecology, 2008, 89, 905–912. Zakoian, J. M., Threshold heteroskedastic models. J Econ Dyn Cont, 1994, 18(5), 931–955. Zhang, L., Mykland, P. A. and Aït-Sahalia, Y., A tale of two time scales: Determining integrated volatility with noisy high- frequency data. J. Am. Stat. Assoc., 2005, 100(472), 1394–1411. 24 M. Fathi and K. Grobys Appendix Figure A1. Time-varying estimated power-law exponents based on annualized daily range-based variances for various assets. Note: This figure plots the estimated power-law exponents derived from rolling window estimations based on annualized daily range-based variances for various assets, including the S&P 500, gold, crude oil, the USD/GBP exchange rate, and Bitcoin (BTC). The annualized daily range-based variances for each asset are calculated using the Garman-Klass methodology. The power-law exponents are estimated using the maximum likelihood estimation approach. Figure A2. Trading volume of the S&P 500 over the sample period. Note: The figure displays the daily dollar trading volume over the sample period from February 19, 1980 to July 31, 2024. Modeling variance risk in financial markets using power-laws 25 Table A1. Descriptive statistics for weekly realized variance. Statistic S&P 500 Gold Crude oil USD/GBP BTC Mean 6.49 6.54 34.81 1.67 74.28 Median 2.99 3.35 15.36 1.09 35.46 Maximum 794.29 265.27 8330.27 78.36 4029.48 Minimum 0.02 0.02 0.14 0.01 0.00 Std. Dev. 22.28 12.33 220.41 3.01 189.70 Skewness 23.90 9.25 32.64 15.04 14.63 Kurtosis 758.30 142.91 1186.98 345.92 277.88 Observations 2242 2239 1711 1253 721 Period (MM/DD/YYYY) 02/19/1980 02/07/1980 01/02/1991 07/25/2000 09/17/2014 07/31/2024 07/31/2024 07/31/2024 07/31/2024 07/31/2024 Note: This table presents the descriptive statistics of the weekly realized variance for various assets, including the S&P 500, gold, crude oil, USD/GBP exchange rate, and Bitcoin (BTC). The weekly realized variances for each asset calculated using the sum of the squared returns over five business days. Table A2. Estimation of power-law exponents for weekly realized variance. Metric S&P 500 Gold Crude oil USD/GBP BTC αˆ 2.51 2.24 2.43 2.84 2.93 xˆMIN 10.51 4.13 23.42 1.96 162.39 D 0.02 0.04 0.03 0.03 0.05 NPL 12.53% 43.19% 33.26% 24.18% 10.40% Std. Dev 0.09 0.04 0.06 0.11 0.22 Observations 2242 2239 1711 1253 721 Period (MM/DD/YYYY) 02/19/1980 02/07/1980 01/02/1991 07/25/2000 09/17/2014 07/31/2024 07/31/2024 07/31/2024 07/31/2024 07/31/2024 Note: This table presents the estimation results for power-law exponents applied to the weekly realized variances of the S&P 500, gold, crude oil, the USD/GBP exchange rate, and Bitcoin. The weekly realized variances for each asset calculated using the sum of the squared returns over five business days. The parameter αˆ signifies the estimated tail exponent, while σˆ denotes the estimated standard deviation. The lower threshold xˆMIN is identified through the optimized Kolmogorov-Smirnov distance (D), following the procedure described by Clauset et al. (2009). The reported xˆMIN corresponds to the optimal distance D. Additionally, the fraction of observations obtained by the power-law process, denoted as NPL is provided for each asset. Table A3. Goodness-of-fit tests for the weekly realized variance. Asset market Power-law Log-normal Exponential S&P 500 0.99 0.58 0.00 Gold 0.04 0.65 0.00 Crude oil 0.34 0.11 0.00 USD/GBP 0.67 0.31 0.00 BTC 0.73 0.31 0.00 Note: This table presents the results (viz., p-values) of the goodness- of-fit tests for weekly realized variance, conducted using the Kolmogorov-Smirnov (KS) method to assess whether the empirical data and the synthetic data—generated from a power-law distribu- tion specified by xmin and α —originate from the same underlying distribution (column 2). Additionally, the table includes the results of goodness-of-fit tests for log-normal and exponential distributions in columns 3 and 4, respectively. The null hypothesis for these tests states that the empirical data and the synthetic data generated from the specified distributions (log-normal and exponential) are consis- tent, suggesting that each distribution provides an acceptable fit to the observed data. The weekly realized variances for each asset mar- ket are calculated using the sum of the squared returns over five business days. 26 M. Fathi and K. Grobys Table A4. Descriptive statistics of estimated power-law exponents derived from rolling window estimations. Statistic S&P 500 Gold Crude Oil USD/GBP BTC Minimum 1.83 1.91 1.79 1.76 1.98 5% Qnt. 1.99 2.17 1.92 2.28 2.04 10% Qnt. 2.10 2.31 2.22 2.54 2.05 Median 2.73 2.77 3.11 3.24 2.54 90% Qnt. 3.88 3.34 4.04 7.01 2.84 95% Qnt. 4.59 3.55 4.27 8.41 2.91 Maximum 7.61 4.63 5.33 11.48 3.38 Mean 2.96 2.79 3.09 3.78 2.45 Std. Dev 0.92 0.44 0.68 1.82 0.31 Excess Kurtosis 7.48 2.67 0.32 4.19 − 0.06 Skewness 2.26 1.15 0.39 2.17 0.35 N 205 204 152 106 53 Note: This table reports descriptive statistics for estimated power-law exponents derived from rolling window estimations based on annualized daily range-based variances for var- ious assets, including the S&P 500, gold, crude oil, the USD/GBP exchange rate, and Bitcoin (BTC). The annualized daily range-based variances for each asset are calculated using the Garman-Klass methodology. The power-law exponents are estimated using the maximum likelihood estimation approach. Quintiles (Qnt) are reported from the 5th to the 95th percentile. Table A5. Important subfamilies of the GB2 distribution. Subfamily Condition Generalized Gamma q → ∞ Beta Prime a = 1 F-distribution a = 1, b = 1, p = d12 , q = d22 Singh–Maddala p = 1 Dagum distribution q = 1 Pareto type IV p = 1, b = 1 Note: The Generalized Beta of the Second Kind (GB2) is a four-parameter family of continuous probability distributions that encompasses a wide variety of shapes. The GB2 distribution is defined as: f (x; a; b; p; q) = a(x/b) ap−1 bB(p, q)[1 + (x/b)a]p+q , where B(p, q) is the Beta function, a > 0 is the tail parameter controlling the tail heaviness, b > 0 is a scale parameter, p > 0, q > 0 are shape parameters that also govern skewness. This table reports important subfamilies of the GB2 distribution. Table A6. Estimated parameters of the GB2 distribution. Range-based variance a b p q S&P 500 125.30 4.44 67.73 0.016 Gold 114.95 6.36 116.26 0.015 Crude oil 73.95 22.47 113.15 0.022 USD/GBP 108.31 1.05 105.28 0.016 BTC 223.86 243.08 0.0255 0.0094 Note: The Generalized Beta of the Second Kind (GB2) is a four-parameter family of continuous probability distributions that encompasses a wide variety of shapes. The GB2 distribution is defined as: f (x; a; b; p; q) = a(x/b) ap−1 bB(p, q)[1 + (x/b)a]p+q , where B(p, q) is the Beta function, a > 0 is the tail parameter controlling the tail heaviness, b > 0 is a scale parameter, p > 0, q > 0 are shape parameters that influence the distribution’s form, including its skewness. This table reports the parameter estimates for GB2 distributions fitted to the data ranges defined by the power-law cutoffs listed in table 3.