Upanshu Sharma’s research while affiliated with UNSW Sydney and other places

What is this page?


This page lists works of an author who doesn't have a ResearchGate profile or hasn't added the works to their profile yet. It is automatically generated from public (personal) data to further our legitimate goal of comprehensive and accurate scientific recordkeeping. If you are this author and want this page removed, please let us know.

Publications (20)


Consider the setting of independent and irreducible Markov jump particles on a three-point state space with generator Q:=[[-3,2,1],[1,-3,2],[2,1,-3)]]\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$Q:=[[-3,2,1],[1,-3,2],[2,1,-3)]]$$\end{document} and invariant measure π=(13,13,13)\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$\pi =(\frac{1}{3},\frac{1}{3},\frac{1}{3})$$\end{document}. Phase portrait for the (zero-cost) trajectories ρ(t)\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$\rho (t)$$\end{document} associated to aL(ρ(t),j(t))=0\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$\mathcal {L}(\rho (t),j(t))=0$$\end{document}; bLFsym(ρ(t),j(t))=0\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$\mathcal {L}_{F^{\textrm{sym}}}(\rho (t),j(t))=0$$\end{document}; cLFasym(ρ(t),j(t))=0\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$\mathcal {L}_{F^{\textrm{asym}}}(\rho (t),j(t))=0$$\end{document}. Here ρi\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$\rho _i$$\end{document} is the mass at point i and we do not plot ρ3\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$\rho _3$$\end{document} since ∑iρi=1\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$\sum _i\rho _i=1$$\end{document}. The zero-cost trajectories for LFsym\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$\mathcal {L}_{F^{\textrm{sym}}}$$\end{document} and LFasym\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$\mathcal {L}_{F^{\textrm{asym}}}$$\end{document} follow a purely dissipative and Hamiltonian dynamics respectively
Contour lines of a possible concave function ζ↦Rζ12(ρ)\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$\zeta \mapsto \mathcal {R}^{\frac{1}{2}}_{\zeta }(\rho )$$\end{document} for a fixed ρ\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$\rho $$\end{document}, where the superlevel set {ζ∈Tρ∗W:Rζ12(ρ)≥0}\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$\{\zeta \in T_\rho ^*\mathcal {W}:\mathcal {R}^{\frac{1}{2}}_{\zeta }(\rho )\ge 0\}$$\end{document} is depicted in gray. By Definitions 2.10 and 2.17, F(ρ)\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$F(\rho )$$\end{document} is a maximiser for ζ↦Rζ12(ρ)\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$\zeta \mapsto \mathcal {R}^{\frac{1}{2}}_{\zeta }(\rho )$$\end{document}, and assuming ρ∈Domsymdiss(Fasym)\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$\rho \in {{\,\textrm{Dom}\,}}_\textrm{symdiss}({F^{\textrm{asym}}})$$\end{document}, Lemma 2.23 says that 2F(ρ)\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$2F(\rho )$$\end{document}, 2Fsym(ρ)\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$2{F^{\textrm{sym}}}(\rho )$$\end{document} and 2Fasym(ρ)\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$2{F^{\textrm{asym}}}(\rho )$$\end{document} all lie on the 0-contour line. By the convexity of the superlevel set {Rζ12(ρ)≥0}\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$\{\mathcal {R}^{\frac{1}{2}}_\zeta (\rho )\ge 0\}$$\end{document} (see Proposition 2.18), any convex combination ζ\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$\zeta $$\end{document} between 0 and 2F(ρ)\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$2F(\rho )$$\end{document}, 2Fsym(ρ)\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$2{F^{\textrm{sym}}}(\rho )$$\end{document} or 2Fasym(ρ)\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$2{F^{\textrm{asym}}}(\rho )$$\end{document}, drawn by the three lines, yield non-negative Rζ12(ρ)≥0\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$\mathcal {R}^{\frac{1}{2}}_\zeta (\rho )\ge 0$$\end{document}
For |X|=3\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$|\mathcal {X}|=3$$\end{document}, the trajectories ω(t)\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$\omega (t)$$\end{document} rotate around the π\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$\sqrt{\pi }$$\end{document}-axis, and lie at the intersection of the two-dimensional sphere S2\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$S^2$$\end{document} and a plane perpendicular to the π\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$\sqrt{\pi }$$\end{document}-axis. The transformation ρx=ωx\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$\rho _x=\sqrt{\omega }_x$$\end{document} maps the (octant) sphere to the simplex of Fig. 1(c)
Variational Structures Beyond Gradient Flows: a Macroscopic Fluctuation-Theory Perspective
  • Article
  • Full-text available

January 2024

·

66 Reads

·

9 Citations

Journal of Statistical Physics

·

D. R. Michiel Renger

·

Upanshu Sharma

Macroscopic equations arising out of stochastic particle systems in detailed balance (called dissipative systems or gradient flows) have a natural variational structure, which can be derived from the large-deviation rate functional for the density of the particle system. While large deviations can be studied in considerable generality, these variational structures are often restricted to systems in detailed balance. Using insights from macroscopic fluctuation theory, in this work we aim to generalise this variational connection beyond dissipative systems by augmenting densities with fluxes, which encode non-dissipative effects. Our main contribution is an abstract theory, which for a given flux-density cost and a quasipotential, provides a decomposition into dissipative and non-dissipative components and a generalised orthogonality relation between them. We then apply this abstract theory to various stochastic particle systems—independent copies of jump processes, zero-range processes, chemical-reaction networks in complex balance and lattice-gas models—without assuming detailed balance. For macroscopic equations arising out of these particle systems, we derive new variational formulations that generalise the classical gradient-flow formulation.

Download


Untangling dissipative and Hamiltonian effects in bulk and boundary-driven systems

November 2023

·

7 Reads

·

2 Citations

PHYSICAL REVIEW E

Using the theory of large deviations, macroscopic fluctuation theory provides a framework to understand the behavior of nonequilibrium dynamics and steady states in diffusive systems. We extend this framework to a minimal model of a nonequilibrium nondiffusive system, specifically an open linear network on a finite graph. We explicitly calculate the dissipative bulk and boundary forces that drive the system towards the steady state, and the nondissipative bulk and boundary forces that drive the system in orbits around the steady state. Using the fact that these forces are orthogonal in a certain sense, we provide a decomposition of the large-deviation cost into dissipative and nondissipative terms. We establish that the purely nondissipative force turns the dynamics into a Hamiltonian system. These theoretical findings are illustrated by numerical examples.


FIGURE 2. A sample path of the Kac process with speed = 2 and switching rate = 1 2 .
FIGURE 3. Schematic outline of the article. The variational structure (, ,) and the corresponding solution concepts (as zeros of this structure) for the Kac equation (1.4) and the FC system (1.7) are introduced in Sections 2.2 and 2.3 respectively. Section 2.3 also shows that these two equations are equivalent. The FIR inequality for the Kac equation is introduced in Section 3. The asymptotic parabolic and hyperbolic limits are discussed in Sections 4.1 and 4.2 respectively. In Appendix A we heuristically motivate the variational structure (, ,) from large deviations, and in Appendix B we derive a pre-GENERIC structure ( , ,, ) for the FC system from the variational structure (, ,).
Fourier-Cattaneo equation: stochastic origin, variational formulation, and asymptotic limits

November 2022

·

78 Reads

We introduce a variational structure for the Fourier-Cattaneo (FC) system which is a second-order hyperbolic system. This variational structure is inspired by the large-deviation rate functional for the Kac process which is closely linked to the FC system. Using this variational formulation we introduce appropriate solution concepts for the FC equation and prove an a priori estimate which connects this variational structure to an appropriate Lyapunov function and Fisher information, the so-called FIR inequality. Finally, we use this formulation and estimate to study the diffusive and hyperbolic limits for the FC system.


Untangling Dissipative and Hamiltonian effects in bulk and boundary driven systems

May 2022

·

20 Reads

Using the theory of large deviations, macroscopic fluctuation theory provides a framework to understand the behaviour of non-equilibrium dynamics and steady states in diffusive systems. We extend this framework to a minimal model of non-equilibrium non-diffusive system, specifically an open linear network on a finite graph. We explicitly calculate the dissipative bulk and boundary forces that drive the system towards the steady state, and non-dissipative bulk and boundary forces that drives the system in orbits around the steady state. Using the fact that these forces are orthogonal in a certain sense, we provide a decomposition of the large-deviation cost into dissipative and non-dissipative terms. We establish that the purely non-dissipative force turns the dynamics into a Hamiltonian system. These theoretical findings are illustrated by numerical examples.



Quantitative coarse-graining of Markov chains

January 2022

·

20 Reads

Coarse-graining techniques play a central role in reducing the complexity of stochastic models, and are typically characterised by a mapping which projects the full state of the system onto a smaller set of variables which captures the essential features of the system. Starting with a continuous-time Markov chain, in this work we propose and analyse an effective dynamics, which approximates the dynamical information in the coarse-grained chain. Without assuming explicit scale-separation, we provide sufficient conditions under which this effective dynamics stays close to the original system and provide quantitative bounds on the approximation error. We also compare the effective dynamics and corresponding error bounds to the averaging literature on Markov chains which involve explicit scale-separation. We demonstrate our findings on an illustrative test example.


Stochastic gradient descent and fast relaxation to thermodynamic equilibrium: A stochastic control approach

December 2021

·

22 Reads

·

6 Citations

We study the convergence to equilibrium of an underdamped Langevin equation that is controlled by a linear feedback force. Specifically, we are interested in sampling the possibly multimodal invariant probability distribution of a Langevin system at small noise (or low temperature), for which the dynamics can easily get trapped inside metastable subsets of the phase space. We follow Chen et al. [J. Math. Phys. 56, 113302 (2015)] and consider a Langevin equation that is simulated at a high temperature, with the control playing the role of a friction that balances the additional noise so as to restore the original invariant measure at a lower temperature. We discuss different limits as the temperature ratio goes to infinity and prove convergence to a limit dynamics. It turns out that, depending on whether the lower (“target”) or the higher (“simulation”) temperature is fixed, the controlled dynamics converges either to the overdamped Langevin equation or to a deterministic gradient flow. This implies that (a) the ergodic limit and the large temperature separation limit do not commute in general and that (b) it is not possible to accelerate the speed of convergence to the ergodic limit by making the temperature separation larger and larger. We discuss the implications of these observations from the perspective of stochastic optimization algorithms and enhanced sampling schemes in molecular dynamics.


Variational structures beyond gradient flows: a macroscopic fluctuation-theory perspective

March 2021

·

40 Reads

Macroscopic equations arising out of stochastic particle systems in detailed balance (called dissipative systems or gradient flows) have a natural variational structure, which can be derived from the large-deviation rate functional for the density of the particle system. While large deviations can be studied in considerable generality, these variational structures are often restricted to systems in detailed balance. Using insights from macroscopic fluctuation theory, in this work we aim to generalise this variational connection beyond dissipative systems by augmenting densities with fluxes, which encode non-dissipative effects. Our main contribution is an abstract framework, which for a given flux-density cost and a quasipotential, provides a decomposition into dissipative and non-dissipative components and a generalised orthogonality relation between them. We then apply this abstract theory to various stochastic particle systems -- independent copies of jump processes, zero-range processes, chemical-reaction networks in complex balance and lattice-gas models.


Stochastic gradient descent and fast relaxation to thermodynamic equilibrium: a stochastic control approach

March 2021

·

33 Reads

We study the convergence to equilibrium of an underdamped Langevin equation that is controlled by a linear feedback force. Specifically, we are interested in sampling the possibly multimodal invariant probability distribution of a Langevin system at small noise (or low temperature), for which the dynamics can easily get trapped inside metastable subsets of the phase space. We follow [Chen et al., J. Math. Phys. 56, 113302, 2015] and consider a Langevin equation that is simulated at a high temperature, with the control playing the role of a friction that balances the additional noise so as to restore the original invariant measure at a lower temperature. We discuss different limits as the temperature ratio goes to infinity and prove convergence to a limit dynamics. It turns out that, depending on whether the lower ("target") or the higher ("simulation") temperature is fixed, the controlled dynamics converges either to the overdamped Langevin equation or to a deterministic gradient flow. This implies that (a) the ergodic limit and the large temperature separation limit do not commute in general, and that (b) it is not possible to accelerate the speed of convergence to the ergodic limit by making the temperature separation larger and larger. We discuss the implications of these observation from the perspective of stochastic optimisation algorithms and enhanced sampling schemes in molecular dynamics.


Citations (11)


... Our construction of information geometry for dynamics is heavily based on the idea of [74][75][76][77][78][79][80][81]. We clarified its information-geometric aspects in the context of CRN and thermodynamics in our previous work [82]. ...

Reference:

Information Geometry of Dynamics on Graphs and Hypergraphs
Variational Structures Beyond Gradient Flows: a Macroscopic Fluctuation-Theory Perspective

Journal of Statistical Physics

... In [33] the authors obtain closed reduced models for the overdamped Langevin dynamics for a specific class of collective variables (reaction coordinates) that satisfy a suitable Poisson equation. The papers [30,51,11,31,32,21] deal with more general diffusion processes and collective variables using conditional expectations (see also [23] for a similar work done in the context of Markov chains). Recently, in [9], the first and third authors of the present paper introduced a new approach to derive a closed reduced model for the underdamped Langevin dynamics based on the use of the dynamics invariance principle [16] for the deterministic part and of the fluctuation-dissipation theorem for the noise term. ...

Quantitative Coarse-Graining of Markov Chains
  • Citing Article
  • January 2024

SIAM Journal on Mathematical Analysis

... However, we stress that all of our examples -apart from the lattice gas modelcannot be cast into the GENERIC framework. This work also provides a framework to study physically relevant 'open-boundary' jump-process systems (see a recent application in [17]). ...

Untangling dissipative and Hamiltonian effects in bulk and boundary-driven systems
  • Citing Article
  • November 2023

PHYSICAL REVIEW E

... We investigate adaptive step size selection for the parallel-across-the-steps Block Gauß-Seidel SDC algorithm and the parallel-acrossthe-method diagonal SDC approach. So far, there are only very few studies that explore adaptive step size selection for the Parareal PinT method [14][15][16] and none for SDCflavored PinT algorithms. ...

An Adaptive Parareal Algorithm: Application to the Simulation of Molecular Dynamics Trajectories
  • Citing Article
  • February 2022

SIAM Journal on Scientific Computing

... In [46] the fast variables (which are different from the ones in this paper) are instead used for local exploration in parameter space. Multiscale dyamics have also been used in the context of neural networks [15,34], to smoothen the loss function, and in controlled, non-asymptotic versions of annealing-like procedures [7]. As we have already mentioned, more recently, [36] studied an interacting particle system designed for the same purpose of solving the MMLE problem, and proposed an algorithm called particle gradient descent (PGD). ...

Stochastic gradient descent and fast relaxation to thermodynamic equilibrium: A stochastic control approach
  • Citing Article
  • December 2021

... Other physical interpretations of saddle avoidance, such as anisotropic friction [37], may be interpreted as a multiscale feature. Linking our findings to recent literature on coarsegraining [75,76] could provide further insights into multiscale stochastic systems. ...

Coarse Graining of Nonreversible Stochastic Differential Equations: Quantitative Results and Connections to Averaging
  • Citing Article
  • June 2020

SIAM Journal on Mathematical Analysis

... Other physical interpretations of saddle avoidance, such as anisotropic friction [37], may be interpreted as a multiscale feature. Linking our findings to recent literature on coarsegraining [75,76] could provide further insights into multiscale stochastic systems. ...

Effective dynamics for non-reversible stochastic differential equations: a quantitative study

... One direction that is tailored to allow for non-dissipative effects is the study of so-called FIR inequalities, first introduced for the many-particle limit of Vlasov-type nonlinear diffusions [5], independent particles on a graph [6] and chemical reactions [7,Sec. 5]. ...

An inequality connecting entropy distance, Fisher Information and large deviations

Stochastic Processes and their Applications

... 5]. These inequalities bound the free-energy difference and Fisher information by the large-deviation rate functional, providing a useful tool to study singular-limit problems and to derive error estimates [8,9]. Strictly speaking, these inequalities are not variational structures in the sense that they do not fully determine the macroscopic dynamics. ...

Quantification of coarse-graining error in Langevin and overdamped Langevin dynamics

·

·

·

[...]

·

U Sharma

... Rigorous results in this direction for state-dependent drift terms appear in the work of , Hottovy et al. (2015Hottovy et al. ( , 2012a, Pardoux and Veretennikov (2003). Similar results are also established for Langevin dynamics (Herzog et al. 2016;Duong et al. 2017), finite-dimensional single-particle GLE (Lim and Wehr 2019;Lim et al. 2020), as well as infinite-dimensional single-particle GLE (Nguyen 2018;Shi and Wang 2021). Analogous study for the stochastic wave equation was central in the work of Freidlin (2006a, 2006b), Cerrai et al. (2017, Cerrai and Glatt-Holtz (2020, 2014, Cerrai and Salins (2016), Nguyen (2022). ...

Variational approach to coarse-graining of generalized gradient flows

Calculus of Variations and Partial Differential Equations