Fact-checked by Grok 2 weeks ago

Conservative system

In , a conservative system is a on a that has no wandering sets of positive measure. This means that for any measurable set A with positive measure, the iterates under the transformation return to intersect A with positive measure infinitely often, embodying the idea of recurrence without dissipation of measure. Informally, conservative systems model dynamical processes where the "volume" or measure in is preserved and does not "leak away" to infinity or disappear, contrasting with dissipative systems. This property ensures that trajectories remain confined in a way that almost every point recurs arbitrarily close to its starting position, as guaranteed by the for finite-measure spaces. The concept is central to understanding long-term behavior in non-dissipative dynamical systems, such as those arising from under , where volume is incompressible. Common examples include irrational rotations on the circle, which are conservative and but not mixing. In broader terms, conservative systems allow for decompositions like the Hopf decomposition into wandering and conservative parts, facilitating analysis of recurrence and .

Introduction

Informal Description

In dynamical systems, a conservative system describes the evolution of points in a where trajectories tend to revisit neighborhoods of their starting points infinitely often, ensuring that the system's behavior remains confined without permanent loss of "mass" or measure. This contrasts with dissipative systems, where parts of the effectively "leak" away, leading to contraction and irreversible spreading of trajectories toward attractors. A real-world illustrates this: imagine gas particles bouncing indefinitely within a sealed, closed , repeatedly returning near their initial positions over vast timescales, versus particles in a leaky that gradually escape, reducing the effective volume explored. In conservative systems, the volume of occupied by an of points remains unchanged over time, preserving the overall without . This idea traces back to Henri Poincaré's insight in the late , where he recognized that in bounded mechanical systems with finite , nearly all points must recur close to their origins due to the conservation of measure, laying foundational groundwork for understanding long-term recurrence in isolated dynamics. Such systems often relate to measure-preserving transformations, where the mapping maintains the size of sets under .

Historical Development

The concept of conservative systems in traces its origins to Poincaré's foundational work on recurrence in . In his 1890 memoir, Poincaré established the recurrence theorem, stating that in a finite-volume under a measure-preserving transformation, almost every point returns arbitrarily close to its initial position infinitely often, laying the groundwork for understanding systems without dissipative behavior. This insight, derived from the , highlighted the conservative nature of mechanical systems with bounded , influencing later developments in by emphasizing recurrent dynamics over dissipation. The formalization of conservative systems advanced significantly in through Eberhard Hopf's contributions to , particularly for infinite measure spaces. Hopf introduced the canonical decomposition that separates conservative and dissipative components, enabling the analysis of transformations where measure is neither created nor destroyed in a global sense. His work on the geodesic flow on surfaces of constant negative curvature demonstrated in such settings, linking recurrence to conservative properties and extending Poincaré's ideas to non-compact spaces. Parallel developments by and further connected conservativity to mixing and ergodic properties in the early 1930s. Birkhoff's 1931 pointwise ergodic theorem established that time averages converge almost everywhere for measure-preserving transformations on finite spaces, implying strong recurrence akin to conservativity. Von Neumann's 1932 mean ergodic theorem complemented this by proving convergence in L² for unitary operators, providing a framework that underscored the stability of conservative dynamics without wandering sets. These results solidified the role of conservativity in distinguishing recurrent from transient behaviors within . Post-1950 advancements, notably by Ulrich Krengel, refined the theory for infinite measures and extended applications to stochastic processes like . Krengel's work in the 1970s and his 1985 monograph developed ergodic theorems for nonsingular transformations on infinite spaces, clarifying the structure of conservative components and their implications for long-term recurrence in non-probabilistic settings. Key milestones include Poincaré's Sur le problème des trois corps et les équations de la dynamique (1890) and Hopf's Fuchsian groups and (1936), marking the progression from qualitative recurrence to rigorous decompositions.

Core Concepts

Measure-Preserving Transformations

A measure-preserving transformation is a fundamental concept in , serving as the foundational structure for studying dynamical systems that conserve measure. Formally, let (X, \mathcal{B}, \mu) be a , where X is the set, \mathcal{B} is a \sigma-algebra of subsets of X, and \mu: \mathcal{B} \to [0, \infty] is a measure. A measurable map T: X \to X is measure-preserving if for every measurable set A \in \mathcal{B}, \mu(T^{-1}(A)) = \mu(A). This condition ensures that the preimage under T of any measurable set has the same measure as the set itself, preserving the "size" of sets in the space. Key properties follow directly from this definition. Measure-preserving transformations preserve null sets: if \mu(A) = 0, then \mu(T^{-1}(A)) = 0, meaning sets of measure zero are mapped to sets of measure zero. Additionally, for the indicator function $1_A of a measurable set A, the transformation preserves its integrability with respect to \mu, as \int_X 1_A \circ T \, d\mu = \mu(T^{-1}(A)) = \mu(A) = \int_X 1_A \, d\mu. More generally, this extends to Lebesgue integral invariance: for any integrable function f: X \to \mathbb{R} (i.e., \int_X |f| \, d\mu < \infty), \int_X f \circ T \, d\mu = \int_X f \, d\mu. This invariance underpins applications such as Birkhoff's pointwise ergodic theorem, which analyzes time averages in dynamical systems. Classic examples illustrate these concepts. Consider the unit circle \mathbb{T} = \mathbb{R}/\mathbb{Z} equipped with the Lebesgue measure \mu normalized to 1. The rotation map T_\alpha: x \mapsto x + \alpha \pmod{1}, for irrational \alpha, is measure-preserving because it rigidly shifts sets without distortion, satisfying \mu(T_\alpha^{-1}(A)) = \mu(A) for any arc A. Similarly, on the d-dimensional torus \mathbb{T}^d, translations by a fixed vector preserve the product Lebesgue measure, providing a simple invertible example of such a transformation. Measure-preserving transformations represent an exact preservation of measure, distinct from the broader class of non-singular transformations, which only require preservation of null sets but may alter measures of positive sets. This strict invariance plays a crucial role in establishing recurrence properties in conservative systems.

Wandering Sets and Recurrence

In ergodic theory, a wandering set for a measure-preserving transformation T on a measure space (X, \mathcal{B}, \mu) is defined as a measurable set W \in \mathcal{B} such that the forward orbits \{T^n(W)\}_{n=0}^\infty are pairwise disjoint, meaning \mu(T^n(W) \cap T^m(W)) = 0 for all n \neq m with n, m \geq 0. This concept captures regions of the space where the dynamics "escape" without overlapping, preventing recurrent behavior in those sets. The provides a foundational result linking measure preservation to recurrent dynamics. Specifically, for a finite-measure space with \mu(X) < \infty and a measure-preserving transformation T: X \to X, if A \in \mathcal{B} has positive measure \mu(A) > 0, then the set of points x \in A such that T^n(x) \notin A for all n > 0 has measure zero; equivalently, point in A returns to A infinitely often under iterations of T. This theorem, originally formulated by in 1890 and rigorously established in the context of , underscores that finite measure preservation forces recurrent trajectories . A key implication of the is that, in finite-measure systems, there cannot exist wandering sets of positive measure, as any such set would contradict the infinite returns guaranteed for almost every point. In other words, the absence of positive-measure wandering sets ensures that orbits revisit neighborhoods indefinitely, reflecting a form of dynamical "" where measure does not dissipate into disjoint regions. This property extends intuitively to -measure settings, where the lack of wandering sets similarly enforces recurrent behavior, though without the full strength of Poincaré's finite-measure guarantee. The theorem can be stated quantitatively for an U \subset X as follows: if \mu(X) < \infty, then \mu\left(\left\{x \in U : T^n(x) \in U \text{ for infinitely many } n \geq 1\right\}\right) = \mu(U). This equality highlights that the recurrent of U retains the full measure of U, providing a precise measure-theoretic description of return frequencies. Systems lacking wandering sets of positive measure are termed conservative with respect to the measure \mu, as the dynamics preserve the "mass" by confining orbits to recurrent components rather than allowing escape to disjoint wandering regions. This notion of conservativity serves as a precursor to broader classifications in ergodic theory, distinguishing recurrent from dissipative behaviors.

Formal Framework

Definition of Conservative Systems

In ergodic theory, a non-singular transformation T: X \to X on a standard probability space (X, \mathcal{B}, \mu) is defined to be conservative if it admits no wandering set of positive measure. Here, a measurable set W \in \mathcal{B} is wandering for T if the forward iterates \{T^n W : n = 0, 1, 2, \dots \} are pairwise disjoint modulo null sets. This condition ensures that measure is not "dissipated" or lost to disjoint regions under iteration, preserving the incompressibility of the dynamics with respect to the measure class of \mu. An equivalent characterization is that T is conservative if and only if for every measurable set A \in \mathcal{B} with \mu(A) > 0, there exists some integer n \geq 1 such that \mu(A \cap T^{-n} A) > 0. This return condition implies that orbits under T revisit sets of positive measure, extending Poincaré recurrence to the non-singular setting. Another equivalence is incompressibility: T is conservative if and only if for every measurable C \in \mathcal{B} with T^{-1} C \subset C, it holds that \mu(C \setminus T^{-1} C) = 0, meaning no positive measure escapes the set under the inverse dynamics. These formulations highlight the absence of systematic measure leakage in conservative systems. The notion of conservativity generalizes naturally to infinite \sigma-finite measure spaces, where the same criteria apply without requiring finite total measure, as formalized in Hopf's foundational work on . In contrast, dissipative systems are those admitting a wandering set of positive measure, allowing measure to be dispersed into disjoint components without return, which fundamentally differs from the recurrent behavior enforced in conservative dynamics. For instance, finite measure-preserving transformations are always conservative by the , underscoring the role of conservativity in ensuring long-term orbital persistence.

Non-Singular Transformations

In the context of ergodic theory, a measurable transformation T: X \to X on a \sigma-finite measure space (X, \mathcal{B}, \mu) is defined to be non-singular if it preserves null sets under the inverse map, meaning \mu(A) = 0 if and only if \mu(T^{-1}A) = 0 for all A \in \mathcal{B}. This condition ensures that T does not map sets of positive measure to null sets or vice versa, maintaining the equivalence of measures in a transformed sense. The non-singularity of T implies that the measures \mu and \mu \circ T^{-1} (the pushforward T_* \mu) are mutually absolutely continuous. By the Radon-Nikodym theorem, there exists a positive measurable function h: X \to (0, \infty), unique up to \mu-almost everywhere equality, such that d(\mu \circ T^{-1}) = h \, d\mu, meaning \mu(T^{-1}A) = \int_A h \, d\mu for all A \in \mathcal{B}. This h serves as the density of the transformed measure with respect to the original. Locally, for x \in T^{-1}(B) and suitable sets B \in \mathcal{B}, h(x) = \frac{d\mu(T^{-1}B)}{d\mu(B)}, analogous to a Jacobian determinant in the differentiable case, capturing the local scaling of measure under T. Non-singular transformations extend the beyond finite measures, accommodating \sigma-finite (possibly infinite) measures where strict invariance may not hold. They encompass all measure-preserving transformations as a subclass, where h = 1 \mu-, but allow for more general dynamics where measure is quasi-invariant. This generality is essential for analyzing systems on infinite spaces, such as certain flows or infinite-dimensional models. The class of non-singular transformations provides the ambient setting for conservative systems, where the conservativity condition—absence of wandering sets of positive measure—ensures recurrent behavior within this broader measure-theoretic structure, enabling the study of dynamics that mimic dissipation without actual measure loss.

Decompositions and Properties

Hopf Decomposition

In ergodic theory, the Hopf decomposition theorem establishes a fundamental partition of the measure space associated with a non-singular transformation, separating recurrent from transient behaviors. For a non-singular transformation T: X \to X on a \sigma-finite measure space (X, \mathcal{B}, \mu), the theorem asserts that there exist disjoint T-invariant measurable sets C and D such that X = C \cup D up to a \mu-null set, where T is conservative on C and dissipative on D, with the decomposition unique up to null sets. This result, originally due to Eberhard Hopf, applies to transformations preserving the measure class without necessarily preserving the measure itself. The construction of the decomposition iteratively identifies and removes wandering sets to form the dissipative part D. A wandering set W \in \mathcal{B} with \mu(W) > 0 satisfies \mu(T^n W \cap T^m W) = 0 for all distinct integers n, m. If the system admits wandering sets of positive measure, a maximal such set exists by Zorn's lemma, and D is the union of all disjoint translates T^n W over n \in \mathbb{Z}, extended to the largest possible such union covering all dissipative behavior. The complement C = X \setminus D (up to null sets) then forms the conservative part, where no wandering sets of positive measure exist. Both C and D are T-invariant modulo null sets. Key properties highlight the distinct dynamics on each part. On C, T is conservative, meaning for any E \subset C with \mu(E) > 0, the iterates T^n E intersect E for infinitely many n, ensuring recurrent orbits . On D, T is dissipative, as D admits a wandering set W such that the \bigcup_{n \in \mathbb{Z}} T^n W covers D up to a , with \sum_{n \in \mathbb{Z}} \mu(T^n W) = \infty possible in measure spaces. These properties hold regardless of whether \mu(C) or \mu(D) is finite or , provided \mu is \sigma-finite. A proof sketch proceeds by first assuming the existence of wandering sets and constructing a maximal one via the , ensuring additivity of measures under disjointness due to non-singularity. If D were not the full dissipative component, a further wandering set in the putative larger conservative part would contradict maximality, forcing C to lack positive-measure wandering sets. follows from the fact that any two such decompositions must coincide on the of all maximal wandering sets, differing only by sets. This decomposition identifies "eternal" recurrent structures on C, where dynamics persist indefinitely, versus "transient" behaviors on D, where orbits escape to infinity, providing essential structure for infinite-measure non-singular systems beyond strict measure preservation.

Ergodic Decomposition

The ergodic decomposition theorem states that any measure-preserving transformation T on a probability space (X, \mathcal{B}, \mu) can be decomposed into ergodic components, where the invariant measure \mu is expressed as an integral over a family of ergodic T-invariant probability measures \{\mu_y\}_{y \in Y} supported on disjoint invariant sets E_y \subset X, with Y being the factor space X / \mathcal{I}(T) modulo the \sigma-algebra \mathcal{I}(T) of T-invariant sets. Each ergodic component E_y is T-invariant, and T|_{E_y} is ergodic with respect to \mu_y for \nu-almost every y, where \nu is the pushforward measure on Y. In the context of conservative systems, the conservative part identified in the Hopf decomposition further decomposes into ergodic conservative components. Since ergodic measure-preserving transformations satisfy the , each such component is conservative, meaning it contains no wandering sets of positive measure. The ergodic components possess key properties: they represent minimal T-invariant sets in the sense that no proper subset of positive measure is invariant, and on each component E_y, the Birkhoff ergodic theorem ensures that time averages equal space averages for integrable functions, i.e., \lim_{n \to \infty} \frac{1}{n} \sum_{k=0}^{n-1} f(T^k x) = \int_{E_y} f \, d\mu_y almost everywhere with respect to \mu_y. This decomposition is unique up to measure-zero sets and is given by the disintegration of \mu with respect to \mathcal{I}(T). Mathematically, for any integrable f, the over the decomposes as \int_X f \, d\mu = \int_Y \left( \int_{E_y} f \, d\mu_y \right) d\nu(y), where the inner is the onto the component. In cases of \sigma-finite measures, may be applied if components have finite positive measure, yielding \int_X f \, d\mu = \int_Y \left( \frac{1}{\mu(E_y)} \int_{E_y} f \, d\mu_y \right) d\nu(y). This decomposition has significant implications, linking to unique ergodicity—where a single ergodic component implies the system is uniquely ergodic—and to , as the spectrum of the Koopman on the full is the of the spectra on the ergodic components.

Examples

Physical Systems

In , physical systems governed by a function exhibit conservative dynamics where the flow on preserves volume, as stated by . This theorem asserts that the volume occupied by an of trajectories remains constant over time, reflecting the absence of in the . A classic example is an confined to a box, where particles interact via elastic collisions and move under Newtonian forces; the 's evolution maintains the incompressibility of the distribution, ensuring recurrent behavior in bounded domains. Celestial mechanics provides another illustration through planetary orbits under gravitational influence, which follow conservative dynamics in the two-body or restricted . Here, the flow preserves the structure, leading to stable, recurrent trajectories for planets like those in the Solar System, where perturbations do not dissipate energy but instead induce long-term periodic motions. This measure-preserving property underpins the predictability of orbital stability over astronomical timescales, as seen in the near-Keplerian paths of and its moons. In , incompressible flows modeled by the Euler equations represent conservative systems where the velocity field satisfies the divergence-free condition, ensuring volume preservation for fluid parcels. The equations, \partial_t \mathbf{u} + (\mathbf{u} \cdot \nabla) \mathbf{u} = -\nabla p with \nabla \cdot \mathbf{u} = 0, describe inviscid fluids like ideal liquids, maintaining kinetic energy and inducing recurrent patterns in bounded domains such as a toroidal flow. This preservation aligns with the structure of the equations, highlighting their time-reversible nature without energy loss. The particle dynamics in Saturn's rings exemplify a conservative system, where icy bodies under Saturn's in a collisionless, framework that conserves measure. Shepherded by moons like , the rings maintain stable density distributions through gravitational resonances, exhibiting recurrent clustering without net mass loss. In contrast, introducing , as in the Navier-Stokes equations for real fluids, transforms the system into a dissipative one by adding frictional terms that contract volumes and prevent recurrence.

Mathematical Models

One prominent example of a conservative system is the irrational rotation on the torus. Consider the two-dimensional torus \mathbb{T}^2 = \mathbb{R}^2 / \mathbb{Z}^2 equipped with the Lebesgue measure \lambda, which is preserved by the rotation map R_{\alpha, \beta}: (x, y) \mapsto (x + \alpha, y + \beta) \pmod{1}, where \alpha, \beta \in [0,1) are such that $1, \alpha, \beta are linearly independent over the rationals. This transformation is ergodic with respect to \lambda, implying that it is conservative, as every ergodic measure-preserving transformation satisfies the Poincaré recurrence theorem without wandering sets. Another classic illustration arises from Bernoulli shifts on infinite product spaces. The one-sided Bernoulli shift acts on the space X = \{0,1\}^{\mathbb{N}} (or more generally, a product of finite sets) via the shift map \sigma: (x_1, x_2, \dots) \mapsto (x_2, x_3, \dots), preserving the \mu = \prod_{n=1}^\infty \mu_n where each \mu_n is a on the finite alphabet (e.g., Bernoulli with p \in (0,1)). This system is ergodic under \mu, hence conservative, due to the independence of coordinates ensuring that sets have measure 0 or 1. Horocycle flows provide an example in the setting of infinite measure spaces, where conservativity follows from the Hopf decomposition. On the unit tangent bundle T^1(M) of a hyperbolic surface M of finite area (but infinite volume in the flow direction), the horocycle flow \{h_t\}_{t \in \mathbb{R}} preserves the Liouville measure \nu, which is infinite. This flow is conservative, meaning the conservative part in the Hopf decomposition coincides with the entire space modulo null sets, as there are no wandering sets of positive measure; ergodicity further ensures unique infinite invariant measures up to scalar multiples. In the setting, recurrent irreducible Markov chains model conservative systems. Consider a countable state space S with transition kernel P: S \times \mathcal{B}(S) \to [0,1] forming an irreducible chain, meaning every state communicates with every other. If the chain is recurrent (positive or ), it admits an measure \pi (finite for positive recurrence, infinite for ), and the associated (S, P, \pi) is conservative, as recurrence implies that orbits return infinitely often without dissipation to transient components. A contrasting non-example is the one-sided shift on a finite equipped with a dissipative measure. For the full one-sided shift \sigma on \{0,1\}^{\mathbb{N}} with a \sigma-finite but dissipative measure \mu (e.g., a measure where the Hopf decomposition has a positive-measure dissipative part, such as certain weighted products decaying to zero), the system fails to be conservative, as points escape to wandering sets under iteration.

References

  1. [1]
    8.2 Conservative and Non-Conservative Forces - UCF Pressbooks
    The work done by a non-conservative force depends on the path taken. Equivalently, a force is conservative if the work it does around any closed path is zero:.
  2. [2]
    [PDF] Conservative and Non-conservative Forces F - Montgomery College
    In physics, we separate forces into conservative and non-conservative categories. The work done by a conservative force depends only on the beginning.
  3. [3]
    27. 7.4 Conservative Forces and Potential Energy - Open Books
    A conservative force is one for which work depends only on the starting and ending points of a motion, not on the path taken. · We can define potential energy ...
  4. [4]
    [PDF] CONSERVATIVE FORCE SYSTEMS
    In a conservative force system the work done by the force can be expressed as the negative of the change in the potential energy (Wc. = - ΔU). Potential ...Missing: definition | Show results with:definition
  5. [5]
    MATHEMATICA tutorial, Part 2.3: Conservative systems
    We will call any autonomous system that possesses a nontrivial first integral a conservative system. A physical system is conservative when it is driven by ...
  6. [6]
    The Principle of Minimum Potential Energy for Conservative ...
    A conservative system is defined as a system whose energy function is independent of the path between different deformation configurations, while a ...
  7. [7]
    [PDF] 8. Conservative Forces and Potential Energy
    Therefore, we can define, for conservative forces, an associated potential energy that, for a given object, depends only on its location. In particular ...
  8. [8]
    [PDF] Notes on Ergodic Theory.
    Mar 5, 2017 · Definition 5. A system (X,T,B,µ) is called conservative if for every set A ∈ B with. µ(A) > 0 ...
  9. [9]
    Modern ergodic theory - AIP Publishing
    "ergodic theory. ... The time evolution of the system causes x to move in phase space, but since we are assuming our system to be conservative the point x always ...
  10. [10]
    [PDF] Disturbing Implications of a Cosmological Constant - arXiv
    Nov 14, 2002 · A rough analogy would be a history in which all the particles in a box start in one corner and after a given long time return to the corner.
  11. [11]
    [PDF] Categories of Conservative Flows - University of Wisconsin–Madison
    Conservative dynamical systems are ones in which the phase space volume is conserved and the flow is incompressible according to Liouville's theorem.
  12. [12]
    [PDF] ERGODIC THEORY – NOTES Contents 1. Measure preserving ...
    Jul 21, 2020 · A measure preserving system is a quadruple (X, B, µ, T) where (X, B,µ) is a probability space and T : X → X is a measure preserving ...
  13. [13]
    [PDF] 9. Ergodic theory 9.1. Measure-preserving transformations. Let (E, E ...
    Ergodic theory. 9.1. Measure-preserving transformations. Let (E, E,µ) be a measure space. A measurable function θ : E → E is called a measure-preserving ...
  14. [14]
    [PDF] Lecture Notes on Ergodic Theory - Weizmann Institute of Science
    Dynamical systems and ergodic theory. Ergodic theory is a part of the theory of dynamical systems. At its simplest form, a dynamical system is a function T ...
  15. [15]
    [PDF] 1.1 Measure preserving transformations
    In ergodic theory, the discrete dynamical systems f : X → X studied are the ones in which X is a measured space and the transformation f is measure- preserving.
  16. [16]
    [PDF] ERGODIC THEORY AND ENTROPY - UChicago Math
    Definition 2.4. An invertible measure-preserving transformation is a bijective transformation T such that T and T−1 are measure-preserving. Example 2.5.
  17. [17]
  18. [18]
  19. [19]
    [PDF] arXiv:0803.2424v3 [math.DS] 28 Jul 2022
    Jul 28, 2022 · A Borel map T : X → X is a nonsingular transformation of X if for any N ∈ B, µ(T−1N) = 0 if and only if µ(N)=0.
  20. [20]
    [0803.2424] Ergodic Theory: Nonsingular Transformations - arXiv
    Mar 17, 2008 · Abstract:This survey is a 2022 update of the 2008 version, with recent developments and new references. Subjects: Dynamical Systems (math.DS).<|control11|><|separator|>
  21. [21]
    [PDF] AN OVERVIEW OF INFINITE ERGODIC THEORY Jon Aaronson
    We review the basic ergodic theory of non-singular transformations plac- ing special emphasis on those transformations admitting σ-finite, infinite invariant.
  22. [22]
  23. [23]
    [PDF] Notes on ergodic theory
    Jul 5, 2017 · These are notes from an introductory course on ergodic theory given at the He- brew University of Jerusalem in the spring semester of 2017.
  24. [24]
    [PDF] A variational principle for volume-preserving dynamics - arXiv
    It is well known that Hamiltonian dynamics preserves the volume in phase space. (Liouville theorem); they are thus a prominent example of incompressible (or.
  25. [25]
    Stability analysis of planetary systems via second-order Rényi entropy
    An important element in the theory of dynamical systems is Poincaré's recurrence theorem. The essence of this is that in the phase space of any conservative ...
  26. [26]
    [PDF] An Introduction to the Incompressible Euler Equations - UC Davis Math
    Sep 25, 2006 · We derive the incompressible Euler equations for the flow of an inviscid, incompressible fluid, describe some of their basic mathematical ...
  27. [27]
    [PDF] 26.457 Notes on Hamiltonians and Conservative Systems
    A second sort of solar-system gaps occurs in the rings of Saturn. In this system. Saturn is the attractor; the perturber is any of the inner satellites, and ...
  28. [28]
    On the stability of Saturn's rings: a quasi-linear kinetic theory
    A self-consistent system of the Boltzmann kinetic equation and the Poisson equation is used to study the dynamical evolution of Saturn's main A, B and C rings.Missing: conservative preserving
  29. [29]
    [PDF] Governing Equations of Fluid Dynamics
    Inviscid flow is, by definition, a flow where the dissipative, transport phenomena of viscosity, mass diffusion and thermal conductivity are neglected. The ...
  30. [30]
    Irrational rotation factors for conservative torus homeomorphisms
    Oct 14, 2014 · Abstract:We provide an equivalent characterisation for the existence of one-dimensional irrational rotation factors of conservative torus ...Missing: ergodic | Show results with:ergodic
  31. [31]
    [PDF] 1. Ergodic theory 1.1. Bernoulli shifts. Let m be a probability ...
    Theorem 1.1.1. The shift map θ is an ergodic measure-preserving transformation. Proof. The details of showing that θ is measurable and measure-preserving ...
  32. [32]
  33. [33]
    Remarks on conservative Markov processes - SpringerLink
    Foguel,The ergodic theory of Markov processes, to be published at Van Nostrand. ... Markov Chain; Random Walk; Characteristic Function; Markov Process; Measure ...
  34. [34]
    [PDF] Totally dissipative measures for the shift and conformal σ-finite ...
    In this paper we investigate some results of ergodic theory with infinite mea- sures for a subshift of finite type. We give an explicit way to construct σ- ...