Fact-checked by Grok 2 weeks ago

Stationary process

In probability theory and statistics, a stationary process, also known as a strictly stationary process, is a stochastic process whose finite-dimensional distributions remain invariant under shifts in time. This means that for any integer k \geq 1 and any n \geq 0, the joint distribution of (X_n, X_{n+1}, \dots, X_{n+k-1}) is identical to that of (X_0, X_1, \dots, X_{k-1}). Strict stationarity captures the full probabilistic structure of the process, ensuring that its statistical properties, including all moments and dependencies, do not evolve over time. A related but weaker concept is weak stationarity (or wide-sense stationarity), which applies primarily to second-order processes and requires only that the mean is constant across time and that the autocovariance function depends solely on the time lag between observations, rather than their absolute positions. Specifically, for a weakly stationary process \{X_t\}, \mathbb{E}[X_t] = \mu for all t, and \text{Cov}(X_t, X_{t+\tau}) = \gamma(\tau) for all t and lag \tau, where \gamma(\tau) is finite and symmetric. Weak stationarity is less restrictive than strict stationarity, as the latter implies the former only if second moments exist, but it suffices for many practical analyses involving linear models and spectral properties. Processes that are Gaussian are strictly stationary if and only if they are weakly stationary, due to the complete characterization of Gaussian distributions by their mean and covariance. Stationary processes form the cornerstone of analysis, enabling the estimation of consistent sample such as means, variances, and correlations, which would otherwise be unreliable in non-stationary data. They simplify modeling and by assuming time-invariance, making techniques like autoregressive (ARMA) models applicable, and are essential for ergodic theorems that equate time averages to ensemble averages in large samples. Beyond , stationary processes underpin applications in , , and physics, where assumptions of stationarity facilitate and prediction of random phenomena like noise in communications or fluctuations in financial markets.

Fundamental Concepts

Overview and Importance

A is a in which the of any collection of its random variables remains invariant under time shifts, meaning its statistical properties, such as mean and variance, do not change over time. This time-invariance distinguishes stationary processes from non-stationary ones, where properties evolve, and encompasses two primary forms: strict-sense stationarity, requiring full distributional invariance, and wide-sense stationarity, focusing on constant mean and . The concept is fundamental in time series analysis, where it simplifies modeling, forecasting, and by enabling the assumption of consistent probabilistic behavior across time periods. Without stationarity, standard techniques like can yield misleading results, as the process cannot be reliably treated as a sequence of independent draws from a fixed . This invariance facilitates the application of tools like autoregressive models and , which rely on stable temporal structures to extract meaningful patterns. Historically, stationary processes emerged in physics around 1900 through studies of , where modeled particle displacements as processes with stationary increments to link microscopic fluctuations to observable . The concept transitioned to in the 1920s, as researchers like G. Udny Yule and demonstrated how non-stationary could produce illusory cycles and correlations, underscoring the need for stationarity in economic modeling. Stationary processes play a critical role in diverse fields, including for filtering in stationary signals, finance for pricing assets under stable volatility assumptions, and climate modeling to distinguish genuine trends from random variations. In these domains, non-stationarity often leads to spurious correlations, such as apparent relationships between unrelated trending series, which can invalidate predictions unless addressed. For instance, in climate science, transforming non-stationary data to stationarity enables reliable simulations of variability.

Historical Development

The concept of stationarity originated in 19th-century physics, particularly in the , where James Clerk Maxwell assumed steady-state distributions for molecular velocities to describe conditions in gaseous systems. In his 1860 work, Maxwell derived the distribution of velocities under the assumption that the system reaches a stable, time-invariant state after collisions, laying early groundwork for notions of unchanging statistical properties over time. The formal introduction of stationarity in stochastic processes occurred in the 1930s through the contributions of Soviet mathematicians. Alexander Khinchin established the correlation theory for stochastic processes in 1934, linking stationarity to by showing that time averages converge to ensemble averages under certain correlation decay conditions. further advanced this in 1941 with his foundational work on and of stationary random sequences, providing rigorous probabilistic frameworks for prediction in such processes. In the realm of time series analysis, stationarity gained practical traction in the late 1920s and early 1930s. George Udny Yule's 1927 paper introduced autoregressive models for investigating periodicities in disturbed series, implicitly relying on stationary assumptions to model data as stable linear dependencies on past values. Gilbert Walker extended this in 1931 by developing models for periodicity in interrelated series, incorporating components that assumed underlying stationarity for weather and economic patterns. Post-World War II advancements emphasized computational aspects through . , collaborating with Ralph Blackman in 1958, promoted wide-sense stationarity in power spectrum estimation, enabling practical applications in communications engineering by focusing on time-invariant means and covariances for efficient . This shift facilitated broader adoption in fields requiring tractable analysis of noisy data. In , the transition from strict to wide-sense stationarity became prominent in the mid-20th century, allowing flexible modeling of economic without full distributional invariance. In the 1970s, the concept extended into non-linear dynamics and , where stationary invariant measures describe long-term behavior on strange attractors despite sensitive dependence on initial conditions. Seminal works, such as the 1971 paper by David Ruelle and Floris Takens, integrated stationarity into chaotic systems to analyze stable probability distributions amid apparent .

Strict-Sense Stationarity

Definition

A stochastic process \{X_t\}_{t \in T}, where T is the index set (typically the integers or real numbers), is defined as strictly stationary if its finite-dimensional distributions are invariant under time shifts. Specifically, for any integer k \geq 1, any t_1, \dots, t_k \in T, and any shift h \in T such that t_i + h \in T for all i, the joint distribution of (X_{t_1}, X_{t_2}, \dots, X_{t_k}) is the same as that of (X_{t_1 + h}, X_{t_2 + h}, \dots, X_{t_k + h}). This definition captures the full probabilistic structure of the process and does not require the existence of moments. Strict stationarity implies wide-sense stationarity if the first and second moments exist and are finite.

Properties and Examples

Strict-sense stationary processes possess several key properties arising from the time-invariance of their finite-dimensional distributions. All moments, including the , variance, and higher-order joint moments, are invariant under time shifts, meaning they depend only on the relative time differences rather than absolute times. Similarly, cumulants, which are derived from the moments via the , exhibit the same invariance, providing a complete of the process's statistical structure independent of time origin. This invariance extends to the marginal distributions, which remain across all time points, ensuring that the univariate distribution of X_t is for every t. Under additional conditions, such as mixing (where dependence between distant observations diminishes), strict-sense stationary processes can be ergodic, allowing time averages from a single realization to converge to ensemble averages, such as the . Furthermore, such processes are preserved under time-invariant transformations, like applying a fixed or to the observations, as long as the operation does not introduce time dependence; this links to higher-order stationarity, where moment preservation holds for all orders. Illustrative examples highlight these properties. An independent and identically distributed (i.i.d.) sequence, such as where each X_t is drawn from the same distribution independently, is strictly stationary because any shift preserves the joint distributions exactly. A constant process, defined by X_t = c for all t and some fixed c, is trivially stationary, as all joint distributions are degenerate and unchanged by shifts. Circularly symmetric processes provide another example, particularly in periodic or angular settings. For instance, a random phase signal X(t) = A \cos(\omega t + \Theta), where A is constant and \Theta is uniformly distributed on [0, 2\pi), is strictly stationary due to the uniform phase ensuring rotational invariance equivalent to time shifts.

Wide-Sense Stationarity

Definition

A stochastic process \{X_t\}_{t \in T}, where T is the (typically the real numbers or integers), is defined as wide-sense stationary if it satisfies two key conditions on its first- and second-order moments. First, the expected value E[X_t] = \mu must be constant for all t \in T, independent of time. Second, the covariance between X_t and X_{t+\tau} must depend solely on the time lag \tau, expressed as \operatorname{Cov}(X_t, X_{t+\tau}) = \gamma(\tau) for all t, \tau \in T, where \gamma(\cdot) is the autocovariance function. This definition presupposes that the second moments are finite, i.e., E[X_t^2] < \infty for all t, ensuring the covariance is well-defined. The associated autocorrelation function is then given by \rho(\tau) = \gamma(\tau) / \gamma(0), which is normalized such that \rho(0) = 1 and remains invariant with respect to time shifts. For complex-valued processes, the definition extends by requiring E[X_t] = \mu (constant) and E[X_t \overline{X_{t+\tau}}] = \gamma(\tau), where \overline{\cdot} denotes the complex conjugate, to account for non-real cases while preserving the lag dependence. This formulation is weaker than strict-sense stationarity, as it focuses solely on these moment conditions rather than full distributional invariance.

Motivation and Applications

Wide-sense stationarity is motivated by its relative ease of verification and computation compared to strict-sense stationarity, as it requires only the invariance of the mean and autocorrelation function rather than all finite-dimensional distributions. This makes it particularly suitable for practical analyses where full distributional properties are difficult or unnecessary to establish, while still capturing essential second-order statistics. Furthermore, wide-sense stationarity suffices for many theoretical and applied contexts involving linear systems and spectral analysis, where higher-order moments beyond the second are not required, enabling efficient characterization via power spectral density. In time series forecasting, wide-sense stationarity underpins autoregressive moving average (ARMA) models, which decompose stationary processes into autoregressive and moving average components for parameter estimation and prediction. The Box-Jenkins methodology, developed in the 1970s, relies on this framework to identify, estimate, and validate ARMA models after transforming data to achieve stationarity through differencing or other means. In signal processing, wide-sense stationarity facilitates the design of linear time-invariant filters for stationary noise, as the output of such a system remains wide-sense stationary when the input is, allowing straightforward computation of output autocorrelation via convolution with the system's impulse response. For econometrics, it plays a key role in cointegration testing, where non-stationary integrated series are examined for linear combinations that yield wide-sense stationary residuals, indicating long-run equilibrium relationships. An additional advantage of wide-sense stationarity lies in asymptotic theory, where central limit theorems hold for partial sums of stationary linear processes under mixingale-type conditions, ensuring normal approximations for large samples without requiring ergodicity or stricter stationarity. A representative example is the first-order moving average process defined as X_t = \varepsilon_t + \theta \varepsilon_{t-1}, where \{\varepsilon_t\} is white noise with variance \sigma^2; this process exhibits constant variance (1 + \theta^2) \sigma^2 and covariance that depends only on the lag (nonzero only at lag 1, equal to \theta \sigma^2), confirming its wide-sense stationarity for any finite \theta.

Higher-Order Stationarity

N-th Order Stationarity

A stochastic process \{X_t\} is said to be n-th order stationary if, for every integer k \leq n, the joint distribution of any k observations is invariant under time shifts. Specifically, for any times t_1, \dots, t_k and any shift \tau, the joint cumulative distribution function satisfies F_{X_{t_1 + \tau}, \dots, X_{t_k + \tau}}(x_1, \dots, x_k) = F_{X_{t_1}, \dots, X_{t_k}}(x_1, \dots, x_k), meaning these distributions depend only on the time differences t_i - t_j rather than absolute time. This condition ensures that the statistical behavior up to order n remains consistent across the process. This notion of n-th order stationarity provides a framework for analyzing processes where full strict stationarity—requiring invariance of all finite-dimensional distributions—may be overly restrictive, yet the distributions up to a specific order n are sufficient for modeling or inference. For instance, in signal processing or time series analysis, second-order properties often suffice for linear predictions, allowing focus on n-th order without assuming higher-order invariance. An illustrative example involves processes with stable low-order joint distributions but time-varying higher ones, such as a stochastic process with time-invariant marginal distributions (first-order stationary) but evolving bivariate joint distributions (not second-order stationary), for example, where the dependence structure, like correlation, changes over time while marginals remain fixed. Similar dynamics can occur for higher orders, where low-order joints are stationary, but higher-order ones exhibit trends in dependence, enabling targeted analysis based on the stable orders. As n \to \infty, the process is strictly stationary, as all finite-dimensional distributions are invariant under shifts. This limiting case underscores how cumulative distributional invariance recovers the full time-invariance of strict stationarity.

Relation to Strict and Wide-Sense

A strictly stationary process has shift-invariant finite-dimensional distributions, which implies n-th order stationarity for every finite n. Conversely, if the process is n-th order stationary for every n, then it is strictly stationary. Second-order stationarity requires that the joint distribution of any two observations is shift-invariant, which implies wide-sense stationarity (constant mean and lag-dependent autocovariance) provided the first and second moments exist. However, wide-sense stationarity does not imply second-order stationarity in general, as it only constrains the moments, not the full distribution; the two coincide for , where the mean and covariance fully characterize the distribution. Higher-order stationarity for n > 2 extends second-order by requiring invariance of joint distributions up to order n, imposing conditions on higher dependencies beyond those captured by wide-sense. For Gaussian processes, second-order stationarity equates to strict stationarity and thus to higher-order stationarity of all orders, since the distributions are fully characterized by the first- and second-order properties. This equivalence does not extend to joint stationarity across multiple processes unless additional conditions are met.

Joint Stationarity

Strict-Sense Joint Stationarity

Strict-sense joint stationarity extends the notion of strict-sense stationarity from a single stochastic process to a collection of two or more processes, ensuring that their combined statistical behavior remains unchanged under time shifts. For two stochastic processes \{X_t\} and \{Y_t\}, they are jointly strictly stationary if all finite-dimensional joint distributions are invariant to time translation. Specifically, for any integers n and m, any times t_1, \dots, t_n and s_1, \dots, s_m, and any shift \tau, the joint probability density function satisfies p_{X(t_1), \dots, X(t_n), Y(s_1), \dots, Y(s_m)}(x_1, \dots, x_n, y_1, \dots, y_m) = p_{X(t_1 + \tau), \dots, X(t_n + \tau), Y(s_1 + \tau), \dots, Y(s_m + \tau)}(x_1, \dots, x_n, y_1, \dots, y_m). This condition implies that the joint cumulative distribution function F of any finite collection of observations from both processes also satisfies F_{(X_{t_1 + h}, Y_{s_1 + h}, \dots)} = F_{(X_{t_1}, Y_{s_1}, \dots)} for all shifts h. Equivalently, the vector-valued process (X_t, Y_t) is strictly stationary as a single multivariate process. This framework is crucial for modeling and analyzing cross-dependencies in multivariate , where interactions between processes, such as in coupled oscillators, must preserve distributional invariance over time to enable reliable inference on joint dynamics. A representative example is bivariate , where pairs (X_t, Y_t) are independent and identically distributed according to a fixed joint distribution, such as a bivariate Gaussian with zero and constant \begin{pmatrix} 1 & \rho \\ \rho & 1 \end{pmatrix} for some \rho \in (-1, 1); this setup ensures all joint finite-dimensional distributions are shift-invariant, capturing potential linear or nonlinear cross-dependencies while maintaining strict stationarity.

Wide-Sense Joint Stationarity

Joint wide-sense stationarity extends the concept of wide-sense stationarity to multiple stochastic processes, focusing on the second-order joint statistics. Consider two stochastic processes \{X_t\} and \{Y_t\}. These processes are jointly wide-sense stationary if each is individually wide-sense stationary—meaning their means are constant (\mathbb{E}[X_t] = \mu_X and \mathbb{E}[Y_t] = \mu_Y for all t) and their autocovariances depend only on the time lag (\gamma_X(\tau) = \mathrm{Cov}(X_t, X_{t+\tau}) and \gamma_Y(\tau) = \mathrm{Cov}(Y_t, Y_{t+\tau}))—and additionally, their cross-covariance function depends solely on the lag \tau: \gamma_{XY}(\tau) = \mathrm{Cov}(X_t, Y_{t+\tau}) = \mathbb{E}[(X_t - \mu_X)(Y_{t+\tau} - \mu_Y)] for all t and \tau. The cross-correlation function between the processes is then defined as \rho_{XY}(\tau) = \frac{\gamma_{XY}(\tau)}{\sqrt{\gamma_X(0) \gamma_Y(0)}}, which normalizes the cross-covariance by the standard deviations and provides a measure of linear dependence that is invariant to time shifts. This setup ensures that the second-order joint structure of the processes remains unchanged under time translations. An important application arises in , where cointegrated () processes model multiple that, while individually non-stationary, exhibit wide-sense stationarity in their error terms or linear combinations, capturing long-run equilibrium relationships such as those between economic indicators like GDP and interest rates. This condition is weaker than strict-sense stationarity, as it requires only second-order moment invariance rather than time-invariance of all probability distributions.

Comparisons Among Stationarity Types

Equivalence Conditions

A strictly stationary process implies wide-sense stationarity whenever the second moments are finite, as the constant and lag-dependent follow directly from the time-invariance of the finite-dimensional distributions under this condition. For Gaussian processes, wide-sense stationarity is equivalent to strict-sense stationarity, since the finite-dimensional distributions are fully determined by the and functions, which are time-invariant in the wide-sense case. Higher-order stationarity for all finite orders n, meaning the joint moments of any n random variables are invariant under time shifts, implies strict-sense stationarity provided the moments determine the underlying finite-dimensional distributions, as in cases satisfying Carleman's moment condition. In the joint stationarity context for multiple processes, joint strict-sense stationarity implies joint wide-sense stationarity when second moments are finite, but the reverse implication fails in general, with counterexamples arising from non-Gaussian coupled processes where cross-covariances are stationary yet higher-order joint distributions vary with time shifts. Within , stationary processes that are also ergodic exhibit asymptotic between time averages of functions of the process and ensemble averages, as established by Birkhoff's ergodic theorem, enabling consistent of statistical from single realizations. In multivariate settings, these conditions extend naturally to vector-valued processes, where joint govern the overall stationarity.

Implications for Stochastic Processes

Stationarity plays a foundational role in the analysis and modeling of processes by enabling key theoretical results and practical methodologies. For ergodic processes, the Birkhoff ergodic theorem guarantees that time averages converge to ensemble averages, allowing inferences about long-term behavior from finite observations. This equivalence underpins much of in time series, where ensures that sample statistics reliably estimate population parameters. In wide-sense stationary processes, stationarity facilitates through the Wiener-Khinchin theorem, which establishes that the power exists as the of the function, providing a frequency-domain representation essential for filtering and prediction tasks. Without stationarity, such decompositions fail, complicating the identification of underlying dynamics. For Gaussian processes, this equivalence between strict-sense and wide-sense stationarity further simplifies analysis, as second-order statistics fully characterize the process. Non-stationarity introduces significant risks in modeling, such as spurious regressions, where independent non-stationary series exhibit misleading correlations due to shared trends, as demonstrated in econometric simulations. Stationarity assumptions are thus critical in algorithms like the , which relies on constant statistical properties for deriving optimal recursive estimators in linear dynamic systems. Extensions of stationarity include almost-periodic processes, which arise as limits of stationary processes under weak convergence and possess covariance functions that are almost periodic, enabling analysis of quasi-periodic phenomena in physics and engineering. However, challenges persist in applications like communications, where cyclostationary processes—non-stationary with periodic statistical variations—offer superior modeling for signals with inherent cycles, such as modulated carriers, outperforming stationary approximations.

Techniques for Achieving Stationarity

Differencing Methods

Differencing is a fundamental technique in time series analysis used to transform non-stationary processes into ones by eliminating trends and stabilizing the mean. differencing involves computing the differences between consecutive observations, defined as \Delta X_t = X_t - X_{t-1}, which effectively removes linear trends and stabilizes variance in integrated processes of order one. This method is particularly effective for processes exhibiting a constant drift, as the resulting differenced series often approximates with constant mean and variance. For processes with higher-degree polynomial trends, such as trends of k-1, k-th differencing is applied iteratively. Second- differencing, for instance, is given by \Delta^2 X_t = \Delta X_t - \Delta X_{t-1} = X_t - 2X_{t-1} + X_{t-2}, which removes trends but is rarely needed beyond the second due to and potential introduction of unnecessary complexity. Higher- differencing assumes the original series follows a trend of the specified and is chosen based on the minimal that achieves stationarity, often assessed via autocorrelation function decay. In the context of models, differencing plays a central role in handling integrated processes denoted as I(d), where d represents the order of differencing required to achieve . The (p, d, q) framework applies d-th order differencing to the original series before fitting an ARMA(p, q) model to the stationary residuals, enabling for non-stationary data like those with unit roots. This approach, pioneered in the Box-Jenkins methodology, ensures the differenced series meets the stationarity assumptions necessary for parameter estimation and model identification. A classic example is the process, defined as X_t = X_{t-1} + \varepsilon_t, where \varepsilon_t is ; this is non-stationary due to its and accumulating variance. Applying first-order differencing yields \Delta X_t = \varepsilon_t, transforming it into stationary with zero mean and constant variance, facilitating straightforward modeling and prediction. Despite its utility, differencing has limitations, including the risk of over-differencing, which occurs when more differences are applied than necessary, introducing an artificial (MA) structure and negative lag-1 autocorrelations near -0.5. Over-differencing can inflate variance and distort model forecasts, so it is preceded by tests such as the Dickey-Fuller test to confirm the presence of a and determine the appropriate differencing order. The Augmented Dickey-Fuller test extends this by accounting for higher-order autoregressive terms, providing a robust statistical basis for deciding on d before differencing.

Surrogate Data Approaches

Surrogate data approaches generate artificial time series that mimic the statistical properties of the original data under a specific null hypothesis, typically that of a stationary linear process, to test for deviations such as non-stationarity or nonlinearity. These methods preserve key features like the power spectrum or amplitude distribution while introducing randomness to create realizations consistent with the null hypothesis of stationarity and linearity, allowing researchers to assess whether observed behaviors arise from deterministic structures or stochastic variability. One foundational technique is phase randomization surrogates, which generate data by applying the to the original series to obtain the amplitude spectrum, randomizing the phases uniformly between 0 and 2π, and then performing the inverse . This process yields surrogate series that retain the original power —ensuring the same content and structure—while being consistent with a stationary Gaussian process under the of linearity. To address limitations in preserving the empirical alongside the , the amplitude-adjusted Fourier transform (AAFT) method refines this approach through an iterative procedure: first, rank-order the original data to match a 's amplitude via shuffling, then apply in the domain, and iteratively adjust amplitudes to converge on both the power and the original . This results in series consistent with a linear that match the distributional properties of the original, making it suitable for non-Gaussian data under the . In hypothesis testing, these evaluate the of and by computing discriminating statistics—such as or predictability measures—on the original series and comparing them to distributions from multiple surrogates; significant deviations reject the null in favor of non-stationary or nonlinear alternatives. For instance, when applied to chaotic time series like the , phase randomization surrogates can reveal hidden periodicities by preserving spectral power while randomizing phases, allowing detection if the original exhibits stronger periodicity than expected under the stationary linear null. Compared to differencing methods for removing linear trends, surrogate approaches like AAFT test for stationarity and linearity by generating data under a stationary linear null that preserves properties such as the and , enabling detection of non-stationarity or nonlinearity without assuming Gaussianity or preprocessing the original data. This facilitates analysis of complex systems to determine if they conform to stationary assumptions.

References

  1. [1]
    [PDF] Lecture 13 : Stationary Stochastic Processes
    1 Stationary stochastic processes. DEF 13.1 (Stationary stochastic process) A real-valued process {Xn}n≥0 is sta- tionary if for every k, m. (Xm,...,Xm+k) ∼ (X0, ...
  2. [2]
    [PDF] Stat 8112 Lecture Notes Stationary Stochastic Processes
    Apr 29, 2012 · A stochastic process is strictly stationary if for each fixed positive integer. k the distribution of the random vector. (Xn+1,...,Xn+k)
  3. [3]
    [PDF] Chapter 6 Stationary Stochastic Processes.
    A stationary stochastic process is a collection {ξn : n ∈ Z} of random vari- ables with values in some space (X, B) such that the joint distribution of.
  4. [4]
  5. [5]
    [PDF] Contents - UMD MATH - University of Maryland
    Chapter 3. Elements of Stationary. Processes. Stationary stochastic processes in discrete time provide a natural framework for the development of the theory ...
  6. [6]
    [PDF] Stationary Stochastic Process
    ✪ Stationary Stochastic Process. The behavior of a stochastic process, or simply a process, z(t) on a domain T is characterized by the probability ...Missing: definition | Show results with:definition
  7. [7]
    Stationarity and differencing of time series data - Duke People
    Another reason for trying to stationarize a time series is to be able to obtain meaningful sample statistics such as means, variances, and correlations with ...
  8. [8]
    [PDF] Econ 582 Univariate Stationary Time Series
    The ergodic theorem says that for a stationary and ergodic sequence { } the time average converges to the ensemble average as the sample size gets large. That ...Missing: mathematics | Show results with:mathematics
  9. [9]
    Basics of Time Series Modeling | METEO 820 - Dutton Institute
    The easiest time series to model is one that is stationary. Stationarity is a requirement to perform many time series analyses.Missing: processes | Show results with:processes
  10. [10]
    [PDF] Introduction to Stochastic Processes - Lecture Notes - UT Math
    Dec 24, 2010 · a stochastic process is stationary if, statistically, it does not evolve. Its probabilistic behavior today is the same as its probabilistic ...
  11. [11]
    [PDF] 1. Stochastic Processes and Stationarity - DSpace@MIT
    Definition 1.1 (Stochastic Process). A stochastic process is a family of random variables {X(t, ω),t ∈. T, ω ∈ Ω} defined on a probability space (Ω, F,P ) .
  12. [12]
    [PDF] DISCRETE TIME STOCHASTIC PROCESSES
    A stochastic process composed of a sequence of i.i.d. random variables is always stationary. The concept of stationarity plays an important role in time series ...
  13. [13]
    111 years of Brownian motion - PMC - PubMed Central - NIH
    We consider the Brownian motion of a particle and present a tutorial review over the last 111 years since Einstein's paper in 1905.2 Pure Diffusion · 3 Langevin Equation · 4 Hydrodynamic Model
  14. [14]
    [PDF] Why do we Sometimes get Nonsense-Correlations between Time ...
    It is fairly familiar knowledge that we sometimes obtain between quantities varying with the time (time-variables) quite high correla- tions to which we cannot ...
  15. [15]
    [PDF] The Summation of Random Causes as the Source of Cyclic Processes
    May 9, 2006 · * Professor Eugen Slutzky's paper of 1927, "The Summation of Random. Causes as the Source of Cyclic Processes," Problems of Economic ...
  16. [16]
    [PDF] An Introduction To Stochastic Processes And Their Applications
    Applications discussed include modeling stock prices in finance, population dynamics in biology, queueing systems in operations research, and signal processing ...
  17. [17]
    Stochastic Methods and Complexity Science in Climate Research ...
    Stochastic models of climate variability are based on the idea that climate can be decomposed into fast fluctuations, i.e., weather disturbances, and slow ...
  18. [18]
    Korrelationstheorie der stationären stochastischen Prozesse
    Korrelationstheorie der stationären stochastischen Prozesse. Published: December 1934. Volume 109, pages 604–615, (1934); Cite this article. Download PDF.
  19. [19]
    VII. On a method of investigating periodicities disturbed series, with ...
    On a method of investigating periodicities disturbed series, with special reference to Wolfer's sunspot numbers. George Udny Yule.
  20. [20]
    On periodicity in series of related terms - Journals
    On periodicity in series of related terms. Gilbert Thomas Walker. Google ... (1994) Chapter 3 Stationary Nonseasonal Models Time Series Modelling of ...
  21. [21]
    [PDF] The Measurement of Power Spectra from the Point of View of ...
    By R. B. BLACKMAN and J. W. TUKEY. (Manuscript received August 28, 1957). The measurement of power spectra is a problem of steadily increasing im- portance ...
  22. [22]
    The History of Continuous-Time Econometric Models - jstor
    wide-sense stationary continuous-time random processes. But the work of these mathematicians was concerned mainly with the properties of the tran- sitional ...
  23. [23]
    A history of chaos theory - PMC - PubMed Central
    Poincaré showed that some dynamical nonlinear systems had unpredictable behaviors. A century later, deterministic chaos, or the chaos theory, is much debated.
  24. [24]
    [PDF] Lecture Notes 7 Stationary Random Processes • Strict-Sense and ...
    Stationarity refers to time invariance of some, or all, of the statistics of a random process, such as mean, autocorrelation, n-th-order distribution.Missing: history | Show results with:history
  25. [25]
    [PDF] Decorrelating Discrete-Time WSS Processes - Henry Pfister
    Apr 19, 2011 · is wide-sense stationary (WSS) if E [Xn] = m and E [X∗. nXn+k] ... A complex random variable X is proper (or circularly symmetric) if E ...
  26. [26]
    [PDF] Moments, cumulants and some applications to stationary processes
    Moments and cumulants find many uses in main stream statistics generally and with random processes particularly. Moments reflect the parameters of distributions ...
  27. [27]
    [PDF] Random Processes
    Strict-sense stationarity: – A process is nth order stationary if the joint distribution of any set of n time samples is independent of the placement of the ...
  28. [28]
    A non-stationary model for spatially dependent circular response ...
    Sep 3, 2022 · In this paper, we consider wrapped Gaussian processes and introduce a spatial model for circular data that allow for non-stationarity in the ...<|control11|><|separator|>
  29. [29]
    [PDF] ARMA MODELS AND THE BOX JENKINS - INSEAD
    ARMA models combine AR and MA schemes. Box-Jenkins methodology popularized ARMA models, providing guidelines for making series stationary.
  30. [30]
    [PDF] Wide-Sense Stationary
    Let Y (t, e) = L [X(t, e)] be the output of a linear system when X(t, e) is the input. Clearly, Y (t, e) is an ensemble of functions selected by e, and is a ...Missing: complex X_t \bar
  31. [31]
    [PDF] Time series. Stationarity, spurious regression and cointegration.
    Weak stationarity (wide-sense stationarity) is satisfied when series ... Econometrics. Time Series (I). Cointegration. 31 / 33. Page 36. Testing cointegration.
  32. [32]
    Central limit theorem for stationary linear processes
    ### Summary of Key Points on Central Limit Theorem for Stationary Linear Processes
  33. [33]
    Moving Average Processes — Econ 114 - Eric Aldrich
    MA(1) Stationarity and Ergodicity​​ Since the mean and autocovariances are independent of time, an MA(1) is weakly stationary. This is true for all values of θ.Missing: wide- sense
  34. [34]
    [PDF] Computationally efficient Bayesian approximation of fractional ...
    A time series process {X(t) | t ∈ T} is weakly stationary of order n if all its joint moments up to order n exist and are time invariant. ... stationary process ...
  35. [35]
    [PDF] Some Contributions to Filtering, Modeling and Forecasting of ...
    A weakly stationary process has constant (time invariant) joint moments up to order n. That is, a second order weakly stationary process has constant mean and ...
  36. [36]
    [PDF] Lecture 13 Time Series: Stationarity, AR(p) & MA(q)
    Stationarity is an invariant property: The statistical characteristics of the time series do not change over time. There different definitions of stationarity, ...
  37. [37]
    On conditions under which a probability distribution is uniquely ...
    Nov 30, 2019 · We study the relationship between the well-known Carleman's condition guaranteeing that a probability distribution is uniquely determined by its ...
  38. [38]
    [PDF] STOCHASTIC PROCESSES, DETECTION AND ESTIMATION
    A weaker notion of stationarity is referred to as Nth-order stationarity, and is defined in terms of a partial statistical characterization of the process.
  39. [39]
    [PDF] UNIT – III RANDOM PROCESSES PART
    A random process is called a strongly stationary process (SSS) or strict sense ... Two random process {X (t )} & {Y (t )} are said to be jointly stationary in the.
  40. [40]
    [PDF] 3. Multivariate time series models - Baruch MFE Program
    Multivariate time series analysis seeks to analyze several time series jointly. The rationale behind this is the possible presence of interdependences ...
  41. [41]
  42. [42]
    [PDF] Introduction to Time Series Analysis. Lecture 24.
    The cross-covariance function of two jointly stationary processes {Xt} and. {Yt} is γxy(h) = E[(Xt+h − µx)(Yt − µy)]. (Jointly stationary = constant means ...
  43. [43]
    10.1.4 Stationary Processes - Probability Course
    A stationary process has statistical properties that do not change by time. For example, X(t) and X(t+Δ) have the same probability distributions.Missing: complex- | Show results with:complex-
  44. [44]
    Sense Stationary Process - an overview | ScienceDirect Topics
    In this book, we consider only two types of stationary processes: strict-sense stationary processes and the wide-sense stationary (WSS) processes. 2.5.1 Strict- ...Missing: transition econometrics history
  45. [45]
    [PDF] VECTOR AUTOREGRESSIONS AND COINTEGRATION* - NYU Stern
    This paper surveys three topics: vector autoregressive (VAR) models with integrated regressors, cointegration, and structural VAR modeling.
  46. [46]
    [PDF] INTRODUCTION TO STOCHASTIC PROCESSES
    A stochastic process Xt is said to be wide sense stationary or covariance sta- ... A strictly stationary process with finite second moments is co- variance ...
  47. [47]
    [PDF] Stochastic Processes
    An important property of Gaussian processes is that wide sense stationarity and strict sense stationarity are equivalent. That is a random process that is wide ...
  48. [48]
    [PDF] Read Full Report - Purdue University
    The ran- dom field (X(t)) will be assumed to possess finite second moments, and its autocovariance function will be denoted as y y(s) = Cov(X(t), X(t + s)) ...
  49. [49]
    [PDF] 5 Birkhoff's Ergodic Theorem
    Birkhoff's Ergodic Theorem extends the validity of Kolmogorov's strong law to the class of stationary sequences of random variables.
  50. [50]
    [PDF] Stationary processes - Penn Engineering
    A process is wide sense stationary (WSS) if it is not stationary but. ⇒ Mean is constant ⇒ µ(t) = µ for all t. ⇒ Autocorrelation is shift invariant ⇒ RX (t1 ...Missing: seminal paper
  51. [51]
    [PDF] The Wiener-Khinchin Theorem - University of Toronto
    Feb 14, 2017 · For a wide-sense stationary random process. X having power signals as sample functions, it makes sense to define the power spectral density ...
  52. [52]
    Spurious regressions in econometrics - ScienceDirect.com
    A comparative study of time series prediction techniques on economic data. Ph.D. Thesis, University of Nottingham, UK (1969)
  53. [53]
    [2208.08240] Almost periodic stationary processes - arXiv
    Aug 17, 2022 · We derive a necessary and sufficient condition for stochastic processes to have almost periodic finite dimensional distributions.
  54. [54]
    [PDF] THE SPECTRAL CORRELATION THEORY OF ...
    Spectral correlation theory for cyclostationary time-series, from periodic phenomena, involves averaging pairs of sinewave components to produce spectral ...
  55. [55]
    9.1 Stationarity and differencing | Forecasting - OTexts
    Differencing can help stabilise the mean of a time series by removing changes in the level of a time series, and therefore eliminating (or reducing) trend and ...Missing: Hyndman | Show results with:Hyndman
  56. [56]
    ARIMA Differencing | Real Statistics Using Excel
    An ARMA(p,q) process with d-order differencing is called an ARIMA(p.d,q) process. Thus, for example, an ARIMA(2,1,0) process is an AR(2) process with first- ...
  57. [57]
    Time Series Analysis | Wiley Series in Probability and Statistics
    Jun 12, 2008 · A modernized new edition of one of the most trusted books on time series analysis. Since publication of the first edition in 1970, Time Series Analysis has ...
  58. [58]
    Identifying the order of differencing in ARIMA models - Duke People
    The first (and most important) step in fitting an ARIMA model is the determination of the order of differencing needed to stationarize the series.Missing: th | Show results with:th
  59. [59]
    Distribution of the Estimators for Autoregressive Time Series with a ...
    Representations for the limit distributions of the estimator of p and of the regression t test are derived. The estimator of p and the regression t test furnish ...
  60. [60]