Nothing Special   »   [go: up one dir, main page]

You seem to have javascript disabled. Please note that many of the page functionalities won't work as expected without javascript enabled.
 
 
entropy-logo

Journal Browser

Journal Browser

Thermodynamics and Statistical Mechanics of Small Systems

A special issue of Entropy (ISSN 1099-4300). This special issue belongs to the section "Statistical Physics".

Deadline for manuscript submissions: closed (28 February 2018) | Viewed by 102997

Printed Edition Available!
A printed edition of this Special Issue is available here.

Special Issue Editors


E-Mail Website
Guest Editor
Consiglio Nazionale delle Ricerche (CNR), Istituto dei Sistemi Complessi (ISC), c/o Dipartimento di Fisica, Universita' Sapienza Roma, p.le A. Moro 2, 00185 Roma, Italy
Interests: granular materials; non-equilibrium statistical mechanics; computational cognitive science

E-Mail Website
Guest Editor
Dipartimento di Ingegneria, Università della Campania "L. Vanvitelli", Aversa (CE), Italy
Interests: nonequilibrium statistical mechanics; granular systems; anomalous diffusion
Special Issues, Collections and Topics in MDPI journals

E-Mail Website
Guest Editor
Dipartimento di Fisica Università degli studi di Roma "La Sapienza", Piazzale A. Moro, 5 00185 Roma, Italy
Interests: chaos and complexity in dynamical systems; non-equilibrium statistical mechanics; transport and reaction/diffusion

Special Issue Information

Dear Colleagues,

A challenging frontier in statistical physics concerns systems with a small number N of degrees of freedom, far from the thermodynamic limit: such an interest is motivated by the recent increase of resolution in the observation and in the manipulation of the micro-nano world. The peculiar feature of small systems is the relevance of fluctuations, which cannot be neglected. The study of fluctuations of thermodynamics quantities such as energy or entropy goes back to Einstein, Onsager and Kubo: more recently it has taken an acceleration with the establishing of new results in response theory and in the so-called stochastic thermodynamics. Such a turning point has received a great impulse from the study of systems which are far from thermodynamic equilibrium. Applications of the thermodynamics and statistical mechanics of small systems range from molecular biology to micromechanics, including, among others, models of nanotransport, of Brownian motors and of (living or artificial) self-propelled organisms.

Prof. Dr. Andrea Puglisi
Dr. Alessandro Sarracino
Prof. Dr. Angelo Vulpiani
Guest Editors

Manuscript Submission Information

Manuscripts should be submitted online at www.mdpi.com by registering and logging in to this website. Once you are registered, click here to go to the submission form. Manuscripts can be submitted until the deadline. All submissions that pass pre-check are peer-reviewed. Accepted papers will be published continuously in the journal (as soon as accepted) and will be listed together on the special issue website. Research articles, review articles as well as short communications are invited. For planned papers, a title and short abstract (about 100 words) can be sent to the Editorial Office for announcement on this website.

Submitted manuscripts should not have been published previously, nor be under consideration for publication elsewhere (except conference proceedings papers). All manuscripts are thoroughly refereed through a single-blind peer-review process. A guide for authors and other relevant information for submission of manuscripts is available on the Instructions for Authors page. Entropy is an international peer-reviewed open access monthly journal published by MDPI.

Please visit the Instructions for Authors page before submitting a manuscript. The Article Processing Charge (APC) for publication in this open access journal is 2600 CHF (Swiss Francs). Submitted papers should be well formatted and use good English. Authors may use MDPI's English editing service prior to publication or during author revisions.

Keywords

  • Statistical Mechanics
  • Small Systems
  • Stochastic Thermodynamics
  • Non-Equilibrium Fluctuations
  • Large Deviations

Benefits of Publishing in a Special Issue

  • Ease of navigation: Grouping papers by topic helps scholars navigate broad scope journals more efficiently.
  • Greater discoverability: Special Issues support the reach and impact of scientific research. Articles in Special Issues are more discoverable and cited more frequently.
  • Expansion of research network: Special Issues facilitate connections among authors, fostering scientific collaborations.
  • External promotion: Articles in Special Issues are often promoted through the journal's social media, increasing their visibility.
  • e-Book format: Special Issues with more than 10 articles can be published as dedicated e-books, ensuring wide and rapid dissemination.

Further information on MDPI's Special Issue polices can be found here.

Published Papers (21 papers)

Order results
Result details
Select all
Export citation of selected articles as:

Editorial

Jump to: Research, Review

4 pages, 173 KiB  
Editorial
Thermodynamics and Statistical Mechanics of Small Systems
by Andrea Puglisi, Alessandro Sarracino and Angelo Vulpiani
Entropy 2018, 20(6), 392; https://doi.org/10.3390/e20060392 - 23 May 2018
Cited by 8 | Viewed by 3406
Abstract
A challenging frontier in modern statistical physics is concerned with systems with a small number of degrees of freedom, far from the thermodynamic limit.[...] Full article
(This article belongs to the Special Issue Thermodynamics and Statistical Mechanics of Small Systems)

Research

Jump to: Editorial, Review

21 pages, 403 KiB  
Article
Exact Expressions of Spin-Spin Correlation Functions of the Two-Dimensional Rectangular Ising Model on a Finite Lattice
by Tao Mei
Entropy 2018, 20(4), 277; https://doi.org/10.3390/e20040277 - 12 Apr 2018
Cited by 1 | Viewed by 3860
Abstract
We employ the spinor analysis method to evaluate exact expressions of spin-spin correlation functions of the two-dimensional rectangular Ising model on a finite lattice, special process enables us to actually carry out the calculation process. We first present some exact expressions of correlation [...] Read more.
We employ the spinor analysis method to evaluate exact expressions of spin-spin correlation functions of the two-dimensional rectangular Ising model on a finite lattice, special process enables us to actually carry out the calculation process. We first present some exact expressions of correlation functions of the model with periodic-periodic boundary conditions on a finite lattice. The corresponding forms in the thermodynamic limit are presented, which show the short-range order. Then, we present the exact expression of the correlation function of the two farthest pair of spins in a column of the model with periodic-free boundary conditions on a finite lattice. Again, the corresponding form in the thermodynamic limit is discussed, from which the long-range order clearly emerges as the temperature decreases. Full article
(This article belongs to the Special Issue Thermodynamics and Statistical Mechanics of Small Systems)
Show Figures

Figure 1

Figure 1
<p>The domain of the integration in (35).</p>
Full article ">
10 pages, 1678 KiB  
Article
Information Dynamics of a Nonlinear Stochastic Nanopore System
by Claire Gilpin, David Darmon, Zuzanna Siwy and Craig Martens
Entropy 2018, 20(4), 221; https://doi.org/10.3390/e20040221 - 23 Mar 2018
Cited by 4 | Viewed by 4802
Abstract
Nanopores have become a subject of interest in the scientific community due to their potential uses in nanometer-scale laboratory and research applications, including infectious disease diagnostics and DNA sequencing. Additionally, they display behavioral similarity to molecular and cellular scale physiological processes. Recent advances [...] Read more.
Nanopores have become a subject of interest in the scientific community due to their potential uses in nanometer-scale laboratory and research applications, including infectious disease diagnostics and DNA sequencing. Additionally, they display behavioral similarity to molecular and cellular scale physiological processes. Recent advances in information theory have made it possible to probe the information dynamics of nonlinear stochastic dynamical systems, such as autonomously fluctuating nanopore systems, which has enhanced our understanding of the physical systems they model. We present the results of local (LER) and specific entropy rate (SER) computations from a simulation study of an autonomously fluctuating nanopore system. We learn that both metrics show increases that correspond to fluctuations in the nanopore current, indicating fundamental changes in information generation surrounding these fluctuations. Full article
(This article belongs to the Special Issue Thermodynamics and Statistical Mechanics of Small Systems)
Show Figures

Figure 1

Figure 1
<p>Potential at fixed positive values of <span class="html-italic">y</span> (<b>A</b>) and fixed negative values of <span class="html-italic">y</span> (<b>B</b>). The <math display="inline"> <semantics> <mrow> <mi>y</mi> <mo>=</mo> <mn>0</mn> </mrow> </semantics> </math> configuration is shown on both plots. As <span class="html-italic">y</span> becomes more positive, it causes the potential to skew towards a transition to negative <span class="html-italic">x</span>. As <span class="html-italic">y</span> becomes more negative, it causes the potential to skew towards a transition to positive <span class="html-italic">x</span>. Physically, positive values of <span class="html-italic">x</span> in this graph correspond to positive current values.</p>
Full article ">Figure 2
<p><b>Top</b>: the nanopore current, with each orange dot representing a measurement. <b>Middle</b>: the estimate of the local entropy rate (LER) of the nanopore system as a function of time. <b>Bottom</b>: the estimate of the specific entropy rate (SER) of the nanopore system as a function of time. This is a representative excerpt from a 40,000 point time series containing on the order of 100 transitions.</p>
Full article ">Figure 3
<p>A projection of the reconstructed state space for the nanopore system shaded by the estimates of the LER (<b>left</b>) and SER (<b>right</b>) associated with the overall state. The plots reveal a clear trajectory in the reconstructed state space, and the arrows indicate the direction along the transitions between open and closed states. Along this trajectory, regions of relatively low surprise (LER) and low uncertainty (SER) occur when the system is in an open/closed state. Conversely, in the central regions, corresponding to transitions, we see increases in both the LER and SER. Anti-symmetry is noted in the onset of increase in SER, which shows that uncertainty is highest at the beginning of a transition and decreases as the transition proceeds to completion. (<b>a</b>) LER; (<b>b</b>) SER.</p>
Full article ">Figure 4
<p>A 2D cross section of the reconstructed state space constructed from points within <math display="inline"> <semantics> <mrow> <mi>ϵ</mi> <mo>=</mo> <mo>±</mo> <mspace width="3.33333pt"/> <mn>0.05</mn> </mrow> </semantics> </math> of the <math display="inline"> <semantics> <mrow> <msub> <mi>x</mi> <mi>t</mi> </msub> <mo>=</mo> <mn>0</mn> </mrow> </semantics> </math> (<b>left</b>) and <math display="inline"> <semantics> <mrow> <msub> <mi>x</mi> <mrow> <mi>t</mi> <mo>−</mo> <mn>1</mn> </mrow> </msub> <mo>=</mo> <mn>0</mn> </mrow> </semantics> </math> (<b>right</b>) planes for each plot, respectively. These plots show that information is generated most heavily around atypical transition events (the highest LER visible on the periphery of the transition tubes in the LER plot), and there is relatively uniform, high uncertainty for all transitions in the SER plot. (<b>a</b>) LER; (<b>b</b>) SER.</p>
Full article ">
768 KiB  
Article
Information Landscape and Flux, Mutual Information Rate Decomposition and Connections to Entropy Production
by Qian Zeng and Jin Wang
Entropy 2017, 19(12), 678; https://doi.org/10.3390/e19120678 - 11 Dec 2017
Cited by 12 | Viewed by 4140
Abstract
We explored the dynamics of two interacting information systems. We show that for the Markovian marginal systems, the driving force for information dynamics is determined by both the information landscape and information flux. While the information landscape can be used to construct the [...] Read more.
We explored the dynamics of two interacting information systems. We show that for the Markovian marginal systems, the driving force for information dynamics is determined by both the information landscape and information flux. While the information landscape can be used to construct the driving force to describe the equilibrium time-reversible information system dynamics, the information flux can be used to describe the nonequilibrium time-irreversible behaviors of the information system dynamics. The information flux explicitly breaks the detailed balance and is a direct measure of the degree of the nonequilibrium or time-irreversibility. We further demonstrate that the mutual information rate between the two subsystems can be decomposed into the equilibrium time-reversible and nonequilibrium time-irreversible parts, respectively. This decomposition of the Mutual Information Rate (MIR) corresponds to the information landscape-flux decomposition explicitly when the two subsystems behave as Markov chains. Finally, we uncover the intimate relationship between the nonequilibrium thermodynamics in terms of the entropy production rates and the time-irreversible part of the mutual information rate. We found that this relationship and MIR decomposition still hold for the more general stationary and ergodic cases. We demonstrate the above features with two examples of the bivariate Markov chains. Full article
(This article belongs to the Special Issue Thermodynamics and Statistical Mechanics of Small Systems)
593 KiB  
Article
Magnetic Engine for the Single-Particle Landau Problem
by Francisco J. Peña, Alejandro González, Alvaro S. Nunez, Pedro A. Orellana, René G. Rojas and Patricio Vargas
Entropy 2017, 19(12), 639; https://doi.org/10.3390/e19120639 - 25 Nov 2017
Cited by 13 | Viewed by 5195
Abstract
We study the effect of the degeneracy factor in the energy levels of the well-known Landau problem for a magnetic engine. The scheme of the cycle is composed of two adiabatic processes and two isomagnetic processes, driven by a quasi-static modulation of external [...] Read more.
We study the effect of the degeneracy factor in the energy levels of the well-known Landau problem for a magnetic engine. The scheme of the cycle is composed of two adiabatic processes and two isomagnetic processes, driven by a quasi-static modulation of external magnetic field intensity. We derive the analytical expression of the relation between the magnetic field and temperature along the adiabatic process and, in particular, reproduce the expression for the efficiency as a function of the compression ratio. Full article
(This article belongs to the Special Issue Thermodynamics and Statistical Mechanics of Small Systems)
Show Figures

Figure 1

Figure 1
<p>Pictorial description for the novel-magnetic engine represented as an entropy versus a magnetic field diagram.</p>
Full article ">Figure 2
<p>Behavior of the magnetic field versus the temperature for the case without the degeneracy factor (<b>a</b>) and the case with the degeneracy factor <math display="inline"> <semantics> <mfrac> <mrow> <mi mathvariant="sans-serif">Φ</mi> <mo>(</mo> <mi>B</mi> <mo>)</mo> </mrow> <mrow> <mn>2</mn> <msub> <mi mathvariant="sans-serif">Φ</mi> <mn>0</mn> </msub> </mrow> </mfrac> </semantics> </math> (<b>b</b>). We select the factor <math display="inline"> <semantics> <mrow> <mfrac> <mi>A</mi> <mrow> <mn>2</mn> <msub> <mi mathvariant="sans-serif">Φ</mi> <mn>0</mn> </msub> </mrow> </mfrac> <mo>∝</mo> <msup> <mn>10</mn> <mn>8</mn> </msup> </mrow> </semantics> </math> T<sup>−1</sup> for this example.</p>
Full article ">Figure 3
<p>The isoentropic trajectories behavior for the two cases under discussion. In (<b>a</b>) we plot the non-degenerate case <math display="inline"> <semantics> <mrow> <mi>S</mi> <mo>(</mo> <mi>T</mi> <mo>,</mo> <mi>B</mi> <mo>)</mo> <mo>=</mo> <mi>S</mi> <mo>(</mo> <mn>10</mn> <mo>,</mo> <mn>1</mn> <mo>)</mo> </mrow> </semantics> </math> and in (<b>b</b>) the degenerate case <math display="inline"> <semantics> <mrow> <msub> <mi>S</mi> <mi>L</mi> </msub> <mrow> <mo>(</mo> <mi>T</mi> <mo>,</mo> <mi>B</mi> <mo>,</mo> <msup> <mn>10</mn> <mn>8</mn> </msup> <mo>)</mo> </mrow> <mo>=</mo> <msub> <mi>S</mi> <mi>L</mi> </msub> <mrow> <mo>(</mo> <mn>10</mn> <mo>,</mo> <mn>1</mn> <mo>,</mo> <msup> <mn>10</mn> <mn>8</mn> </msup> <mo>)</mo> </mrow> </mrow> </semantics> </math>.</p>
Full article ">Figure 4
<p>Total work (<b>a</b>) and input heat (<b>b</b>) versus the <span class="html-italic">r</span> parameter along the cycle for the case with degeneracy (dot dashed line) and without degeneracy (dashed line).</p>
Full article ">Figure 5
<p>Efficiency for different cases of interest. For this case, the dotted red line corresponds to the value of Carnot cycle for a machine operating between the two temperatures <math display="inline"> <semantics> <mrow> <msub> <mi>T</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>4</mn> </mrow> </semantics> </math> K and <math display="inline"> <semantics> <mrow> <msub> <mi>T</mi> <mn>3</mn> </msub> <mo>=</mo> <mn>10</mn> </mrow> </semantics> </math> K.</p>
Full article ">Figure 6
<p>Magnetization as a function of <span class="html-italic">B</span> and <span class="html-italic">T</span> along the adiabatic trajectories.</p>
Full article ">Figure 7
<p>Magnetization along the first iso-magnetic trajectory as a function of <span class="html-italic">T</span> in the range of 4 K to 10 K. We selected the different values for <math display="inline"> <semantics> <msub> <mi>B</mi> <mn>2</mn> </msub> </semantics> </math> that we found from numerical calculations.</p>
Full article ">Figure A1
<p>A parametric solution of the differential equation along the adiabatic trajectories for the Landau case. The dotted line represents the exact solution and the dot-dashed line the asymptotic case for <math display="inline"> <semantics> <mrow> <mi>u</mi> <mo>&lt;</mo> <mo>&lt;</mo> <mn>1</mn> </mrow> </semantics> </math>. We can clearly see the constant value 0.5 for the solution in the case of <math display="inline"> <semantics> <mrow> <mi>u</mi> <mo>&gt;</mo> <mo>&gt;</mo> <mn>1</mn> </mrow> </semantics> </math> from the dotted line in the figure. The solid line represents the proposal curve given by Equation (<a href="#FD52-entropy-19-00639" class="html-disp-formula">A13</a>) showing a good fit for the problem under study.</p>
Full article ">
275 KiB  
Article
On Work and Heat in Time-Dependent Strong Coupling
by Erik Aurell
Entropy 2017, 19(11), 595; https://doi.org/10.3390/e19110595 - 7 Nov 2017
Cited by 24 | Viewed by 3932
Abstract
This paper revisits the classical problem of representing a thermal bath interacting with a system as a large collection of harmonic oscillators initially in thermal equilibrium. As is well known, the system then obeys an equation, which in the bulk and in the [...] Read more.
This paper revisits the classical problem of representing a thermal bath interacting with a system as a large collection of harmonic oscillators initially in thermal equilibrium. As is well known, the system then obeys an equation, which in the bulk and in the suitable limit tends to the Kramers–Langevin equation of physical kinetics. I consider time-dependent system-bath coupling and show that this leads to an additional harmonic force acting on the system. When the coupling is switched on and switched off rapidly, the force has delta-function support at the initial and final time. I further show that the work and heat functionals as recently defined in stochastic thermodynamics at strong coupling contain additional terms depending on the time derivative of the system-bath coupling. I discuss these terms and show that while they can be very large if the system-bath coupling changes quickly, they only give a finite contribution to the work that enters in Jarzynski’s equality. I also discuss that these corrections to standard work and heat functionals provide an explanation for non-standard terms in the change of the von Neumann entropy of a quantum bath interacting with a quantum system found in an earlier contribution (Aurell and Eichhorn, 2015). Full article
(This article belongs to the Special Issue Thermodynamics and Statistical Mechanics of Small Systems)
468 KiB  
Article
Equilibration in the Nosé–Hoover Isokinetic Ensemble: Effect of Inter-Particle Interactions
by Shamik Gupta and Stefano Ruffo
Entropy 2017, 19(10), 544; https://doi.org/10.3390/e19100544 - 14 Oct 2017
Cited by 1 | Viewed by 4336
Abstract
We investigate the stationary and dynamic properties of the celebrated Nosé–Hoover dynamics of many-body interacting Hamiltonian systems, with an emphasis on the effect of inter-particle interactions. To this end, we consider a model system with both short- and long-range interactions. The Nosé–Hoover dynamics [...] Read more.
We investigate the stationary and dynamic properties of the celebrated Nosé–Hoover dynamics of many-body interacting Hamiltonian systems, with an emphasis on the effect of inter-particle interactions. To this end, we consider a model system with both short- and long-range interactions. The Nosé–Hoover dynamics aim to generate the canonical equilibrium distribution of a system at a desired temperature by employing a set of time-reversible, deterministic equations of motion. A signature of canonical equilibrium is a single-particle momentum distribution that is Gaussian. We find that the equilibrium properties of the system within the Nosé–Hoover dynamics coincides with that within the canonical ensemble. Moreover, starting from out-of-equilibrium initial conditions, the average kinetic energy of the system relaxes to its target value over a size-independent timescale. However, quite surprisingly, our results indicate that under the same conditions and with only long-range interactions present in the system, the momentum distribution relaxes to its Gaussian form in equilibrium over a scale that diverges with the system size. On adding short-range interactions, the relaxation is found to occur over a timescale that has a much weaker dependence on system size. This system-size dependence of the timescale vanishes when only short-range interactions are present in the system. An implication of such an ultra-slow relaxation when only long-range interactions are present in the system is that macroscopic observables other than the average kinetic energy when estimated in the Nosé–Hoover dynamics may take an unusually long time to relax to its canonical equilibrium value. Our work underlines the crucial role that interactions play in deciding the equivalence between Nosé–Hoover and canonical equilibrium. Full article
(This article belongs to the Special Issue Thermodynamics and Statistical Mechanics of Small Systems)
Show Figures

Figure 1

Figure 1
<p>Comparison of Nosé–Hoover and canonical equilibrium results for model (<a href="#FD18-entropy-19-00544" class="html-disp-formula">18</a>) with <math display="inline"> <semantics> <mrow> <mi>J</mi> <mo>=</mo> <mn>0.0</mn> <mo>,</mo> <mi>K</mi> <mo>=</mo> <mn>1.0</mn> </mrow> </semantics> </math> (that is, with only short-range interactions). (<b>a</b>) variation of the average kinetic energy density with time. The black line denotes the value <math display="inline"> <semantics> <mrow> <msub> <mi>T</mi> <mi>target</mi> </msub> <mo>/</mo> <mn>2</mn> </mrow> </semantics> </math>; (<b>b</b>) variation of the internal energy density with time. The black line denotes the average internal energy density within the canonical ensemble given by Equation (<a href="#FD44-entropy-19-00544" class="html-disp-formula">44</a>); (<b>c</b>) stationary single-particle momentum distribution obtained from momentum values measured at time <math display="inline"> <semantics> <mrow> <mi>t</mi> <mo>=</mo> <mn>5000</mn> </mrow> </semantics> </math>. The black line denotes a Gaussian distribution with zero mean and width equal to <math display="inline"> <semantics> <msub> <mi>T</mi> <mi>target</mi> </msub> </semantics> </math>; (<b>d</b>) caloric curve for two system sizes, <math display="inline"> <semantics> <mrow> <mi>N</mi> <mo>=</mo> <mn>128</mn> </mrow> </semantics> </math> and <math display="inline"> <semantics> <mrow> <mi>N</mi> <mo>=</mo> <mn>512</mn> </mrow> </semantics> </math>. The black line shows the caloric curve within the canonical ensemble given by Equation (<a href="#FD44-entropy-19-00544" class="html-disp-formula">44</a>). The data for the Nosé–Hoover dynamics are generated by integrating the equations of motion (<a href="#FD21-entropy-19-00544" class="html-disp-formula">21</a>) using a fourth-order Runge–Kutta method with timestep equal to 0.01. The initial condition corresponds to the <math display="inline"> <semantics> <msub> <mi>θ</mi> <mi>j</mi> </msub> </semantics> </math>’s independently and uniformly distributed in <math display="inline"> <semantics> <mrow> <mo stretchy="false">[</mo> <mn>0</mn> <mo>,</mo> <mn>2</mn> <mi>π</mi> <mo stretchy="false">)</mo> </mrow> </semantics> </math> and the <math display="inline"> <semantics> <msub> <mi>p</mi> <mi>j</mi> </msub> </semantics> </math>’s independently sampled from a Gaussian distribution with zero mean and width equal to 0.5. The initial value of the parameter <math display="inline"> <semantics> <mi>ζ</mi> </semantics> </math> is 2, while we have taken <math display="inline"> <semantics> <mrow> <mi>τ</mi> <mo>=</mo> <mn>0.01</mn> </mrow> </semantics> </math>.</p>
Full article ">Figure 2
<p>Comparison of Nosé–Hoover and canonical equilibrium results for the model (<a href="#FD18-entropy-19-00544" class="html-disp-formula">18</a>) with <math display="inline"> <semantics> <mrow> <mi>J</mi> <mo>=</mo> <mn>1.0</mn> <mo>,</mo> <mi>K</mi> <mo>=</mo> <mn>0.0</mn> </mrow> </semantics> </math> (that is, with only long-range interactions); (<b>a</b>) variation of the average kinetic energy density with time. The black line denotes the value <math display="inline"> <semantics> <mrow> <msub> <mi>T</mi> <mi>target</mi> </msub> <mo>/</mo> <mn>2</mn> </mrow> </semantics> </math>; (<b>b</b>) variation of the internal energy density with time. The black line denotes the average internal energy density within the canonical ensemble given by Equation (<a href="#FD29-entropy-19-00544" class="html-disp-formula">29</a>); (<b>c</b>) stationary single-particle momentum distribution obtained from momentum values measured at time <math display="inline"> <semantics> <mrow> <mi>t</mi> <mo>=</mo> <mn>5000</mn> </mrow> </semantics> </math>. The black line denotes a Gaussian distribution with zero mean and width equal to <math display="inline"> <semantics> <msub> <mi>T</mi> <mi>target</mi> </msub> </semantics> </math>; (<b>d</b>) caloric curve for two system sizes, <math display="inline"> <semantics> <mrow> <mi>N</mi> <mo>=</mo> <mn>128</mn> </mrow> </semantics> </math> and <math display="inline"> <semantics> <mrow> <mi>N</mi> <mo>=</mo> <mn>512</mn> </mrow> </semantics> </math>. The black line shows the caloric curve within the canonical ensemble given by Equation (<a href="#FD29-entropy-19-00544" class="html-disp-formula">29</a>). The data for the Nosé–Hoover dynamics are generated by integrating the equations of motion (<a href="#FD21-entropy-19-00544" class="html-disp-formula">21</a>) using a fourth-order Runge–Kutta method with timestep equal to 0.01. The initial condition corresponds to the <math display="inline"> <semantics> <msub> <mi>θ</mi> <mi>j</mi> </msub> </semantics> </math>’s independently and uniformly distributed in <math display="inline"> <semantics> <mrow> <mo stretchy="false">[</mo> <mn>0</mn> <mo>,</mo> <mn>2</mn> <mi>π</mi> <mo stretchy="false">)</mo> </mrow> </semantics> </math> and the <math display="inline"> <semantics> <msub> <mi>p</mi> <mi>j</mi> </msub> </semantics> </math>’s independently sampled from a Gaussian distribution with zero mean and width equal to 0.5. The initial value of the parameter <math display="inline"> <semantics> <mi>ζ</mi> </semantics> </math> is 2, while we have taken <math display="inline"> <semantics> <mrow> <mi>τ</mi> <mo>=</mo> <mn>0.01</mn> </mrow> </semantics> </math>.</p>
Full article ">Figure 3
<p>Relaxation properties of the Nosé–Hoover dynamics for model (<a href="#FD18-entropy-19-00544" class="html-disp-formula">18</a>) with <math display="inline"> <semantics> <mrow> <mi>J</mi> <mo>=</mo> <mn>1.0</mn> <mo>,</mo> <mi>K</mi> <mo>=</mo> <mn>0.1</mn> </mrow> </semantics> </math>. (<b>a</b>) variation of the average kinetic energy density with time, for four different system sizes. The black line denotes the value <math display="inline"> <semantics> <mrow> <msub> <mi>T</mi> <mi>target</mi> </msub> <mo>/</mo> <mn>2</mn> </mrow> </semantics> </math>; (<b>b</b>) variation of the ratio <math display="inline"> <semantics> <mrow> <mrow> <mo>〈</mo> <msup> <mi>p</mi> <mn>4</mn> </msup> <mo>〉</mo> </mrow> <mo>/</mo> <msup> <mrow> <mo>〈</mo> <msup> <mi>p</mi> <mn>2</mn> </msup> <mo>〉</mo> </mrow> <mn>2</mn> </msup> </mrow> </semantics> </math> with time, for four different system sizes. The black line denotes the value 3 corresponding to a Gaussian distribution; (<b>c</b>) variation of the magnetization with time, again for four different system sizes. The black line denotes the canonical equilibriu m value obtained by the method described in <a href="#sec3dot2dot2-entropy-19-00544" class="html-sec">Section 3.2.2</a>; (<b>d</b>) single-particle momentum distribution as a function of time, for system size <math display="inline"> <semantics> <mrow> <mi>N</mi> <mo>=</mo> <mn>512</mn> </mrow> </semantics> </math>. The black line denotes a Gaussian distribution with zero mean and width equal to <math display="inline"> <semantics> <msub> <mi>T</mi> <mi>target</mi> </msub> </semantics> </math>, Equation (<a href="#FD17-entropy-19-00544" class="html-disp-formula">17</a>). The data for the Nosé–Hoover dynamics are generated by integrating the equations of motion (<a href="#FD21-entropy-19-00544" class="html-disp-formula">21</a>) using a fourth-order Runge–Kutta method with timestep equal to 0.01. The initial condition corresponds to the <math display="inline"> <semantics> <msub> <mi>θ</mi> <mi>j</mi> </msub> </semantics> </math>’s independently and uniformly distributed in <math display="inline"> <semantics> <mrow> <mo stretchy="false">[</mo> <mn>0</mn> <mo>,</mo> <mn>2</mn> <mi>π</mi> <mo stretchy="false">)</mo> </mrow> </semantics> </math> and the <math display="inline"> <semantics> <msub> <mi>p</mi> <mi>j</mi> </msub> </semantics> </math>’s independently and uniformly distributed in <math display="inline"> <semantics> <mrow> <mo>[</mo> <mo>−</mo> <msqrt> <mrow> <mn>1.5</mn> </mrow> </msqrt> <mo>,</mo> <msqrt> <mrow> <mn>1.5</mn> </mrow> </msqrt> <mo>]</mo> </mrow> </semantics> </math>. The initial value of the parameter <math display="inline"> <semantics> <mi>ζ</mi> </semantics> </math> is 2, while we have taken <math display="inline"> <semantics> <mrow> <mi>τ</mi> <mo>=</mo> <mn>1.0</mn> </mrow> </semantics> </math>.</p>
Full article ">Figure 4
<p>Relaxation properties of the Nosé–Hoover dynamics for the model (<a href="#FD18-entropy-19-00544" class="html-disp-formula">18</a>) with <math display="inline"> <semantics> <mrow> <mi>J</mi> <mo>=</mo> <mn>1.0</mn> <mo>,</mo> <mi>K</mi> <mo>=</mo> <mn>0.0</mn> </mrow> </semantics> </math> (that is, with only long-range interactions). (<b>a</b>) variation of the average kinetic energy density with time, for four different system sizes. The black line denotes the value <math display="inline"> <semantics> <mrow> <msub> <mi>T</mi> <mi>target</mi> </msub> <mo>/</mo> <mn>2</mn> </mrow> </semantics> </math>; (<b>b</b>) variation of the ratio <math display="inline"> <semantics> <mrow> <mrow> <mo>〈</mo> <msup> <mi>p</mi> <mn>4</mn> </msup> <mo>〉</mo> </mrow> <mo>/</mo> <msup> <mrow> <mo>〈</mo> <msup> <mi>p</mi> <mn>2</mn> </msup> <mo>〉</mo> </mrow> <mn>2</mn> </msup> </mrow> </semantics> </math> with time, for four different system sizes. The black line denotes the value 3 corresponding to a Gaussian distribution; (<b>c</b>) variation of the magnetization with time, again for four different system sizes. The black line denotes the canonical equilibrium value given by Equation (<a href="#FD28-entropy-19-00544" class="html-disp-formula">28</a>); (<b>d</b>) single-particle momentum distribution as a function of time, for system size <math display="inline"> <semantics> <mrow> <mi>N</mi> <mo>=</mo> <mn>512</mn> </mrow> </semantics> </math>. The black line denotes a Gaussian distribution with zero mean and width equal to <math display="inline"> <semantics> <msub> <mi>T</mi> <mi>target</mi> </msub> </semantics> </math>, Equation (<a href="#FD17-entropy-19-00544" class="html-disp-formula">17</a>). The data for the Nosé–Hoover dynamics are generated by integrating the equations of motion (<a href="#FD21-entropy-19-00544" class="html-disp-formula">21</a>) using a fourth-order Runge–Kutta method with timestep equal to 0.01. The initial condition corresponds to the <math display="inline"> <semantics> <msub> <mi>θ</mi> <mi>j</mi> </msub> </semantics> </math>’s independently and uniformly distributed in <math display="inline"> <semantics> <mrow> <mo stretchy="false">[</mo> <mn>0</mn> <mo>,</mo> <mn>2</mn> <mi>π</mi> <mo stretchy="false">)</mo> </mrow> </semantics> </math> and the <math display="inline"> <semantics> <msub> <mi>p</mi> <mi>j</mi> </msub> </semantics> </math>’s independently and uniformly distributed in <math display="inline"> <semantics> <mrow> <mo>[</mo> <mo>−</mo> <msqrt> <mrow> <mn>1.5</mn> </mrow> </msqrt> <mo>,</mo> <msqrt> <mrow> <mn>1.5</mn> </mrow> </msqrt> <mo>]</mo> </mrow> </semantics> </math>. The initial value of the parameter <math display="inline"> <semantics> <mi>ζ</mi> </semantics> </math> is 2, while we have taken <math display="inline"> <semantics> <mrow> <mi>τ</mi> <mo>=</mo> <mn>1.0</mn> </mrow> </semantics> </math>.</p>
Full article ">Figure 5
<p>Relaxation properties of the Nosé–Hoover dynamics for the model (<a href="#FD18-entropy-19-00544" class="html-disp-formula">18</a>) with <math display="inline"> <semantics> <mrow> <mi>J</mi> <mo>=</mo> <mn>0.0</mn> <mo>,</mo> <mi>K</mi> <mo>=</mo> <mn>1.0</mn> </mrow> </semantics> </math> (that is, with only short-range interactions). (<b>a</b>) variation of the average kinetic energy density with time, for four different system sizes. The black line denotes the value <math display="inline"> <semantics> <mrow> <msub> <mi>T</mi> <mi>target</mi> </msub> <mo>/</mo> <mn>2</mn> </mrow> </semantics> </math>; (<b>b</b>) variation of the ratio <math display="inline"> <semantics> <mrow> <mrow> <mo>〈</mo> <msup> <mi>p</mi> <mn>4</mn> </msup> <mo>〉</mo> </mrow> <mo>/</mo> <msup> <mrow> <mo>〈</mo> <msup> <mi>p</mi> <mn>2</mn> </msup> <mo>〉</mo> </mrow> <mn>2</mn> </msup> </mrow> </semantics> </math> with time, for four different system sizes. The black line denotes the value 3 corresponding to a Gaussian distribution; (<b>c</b>) variation of the magnetization with time, again for four different system sizes. The equilibrium magnetization goes to zero with increase of <span class="html-italic">N</span> as <math display="inline"> <semantics> <mrow> <msup> <mi>m</mi> <mi>eq</mi> </msup> <mo>∼</mo> <mn>1</mn> <mo>/</mo> <msqrt> <mi>N</mi> </msqrt> </mrow> </semantics> </math>; (<b>d</b>) single-particle momentum distribution as a function of time, for system size <math display="inline"> <semantics> <mrow> <mi>N</mi> <mo>=</mo> <mn>512</mn> </mrow> </semantics> </math>. The black line denotes a Gaussian distribution with zero mean and width equal to <math display="inline"> <semantics> <msub> <mi>T</mi> <mi>target</mi> </msub> </semantics> </math>, Equation (<a href="#FD17-entropy-19-00544" class="html-disp-formula">17</a>). The data for the Nosé–Hoover dynamics are generated by integrating the equations of motion (<a href="#FD21-entropy-19-00544" class="html-disp-formula">21</a>) using a fourth-order Runge–Kutta method with timestep equal to 0.01. The initial condition corresponds to the <math display="inline"> <semantics> <msub> <mi>θ</mi> <mi>j</mi> </msub> </semantics> </math>’s independently and uniformly distributed in <math display="inline"> <semantics> <mrow> <mo stretchy="false">[</mo> <mn>0</mn> <mo>,</mo> <mn>2</mn> <mi>π</mi> <mo stretchy="false">)</mo> </mrow> </semantics> </math> and the <math display="inline"> <semantics> <msub> <mi>p</mi> <mi>j</mi> </msub> </semantics> </math>’s independently and uniformly distributed in <math display="inline"> <semantics> <mrow> <mo>[</mo> <mo>−</mo> <msqrt> <mrow> <mn>1.5</mn> </mrow> </msqrt> <mo>,</mo> <msqrt> <mrow> <mn>1.5</mn> </mrow> </msqrt> <mo>]</mo> </mrow> </semantics> </math>. The initial value of the parameter <math display="inline"> <semantics> <mi>ζ</mi> </semantics> </math> is 2, while we have taken <math display="inline"> <semantics> <mrow> <mi>τ</mi> <mo>=</mo> <mn>1.0</mn> </mrow> </semantics> </math>.</p>
Full article ">
4976 KiB  
Article
Kovacs-Like Memory Effect in Athermal Systems: Linear Response Analysis
by Carlos A. Plata and Antonio Prados
Entropy 2017, 19(10), 539; https://doi.org/10.3390/e19100539 - 13 Oct 2017
Cited by 12 | Viewed by 4926
Abstract
We analyze the emergence of Kovacs-like memory effects in athermal systems within the linear response regime. This is done by starting from both the master equation for the probability distribution and the equations for the physically-relevant moments. The general results are applied to [...] Read more.
We analyze the emergence of Kovacs-like memory effects in athermal systems within the linear response regime. This is done by starting from both the master equation for the probability distribution and the equations for the physically-relevant moments. The general results are applied to a general class of models with conserved momentum and non-conserved energy. Our theoretical predictions, obtained within the first Sonine approximation, show an excellent agreement with the numerical results. Furthermore, we prove that the observed non-monotonic relaxation is consistent with the monotonic decay of the non-equilibrium entropy. Full article
(This article belongs to the Special Issue Thermodynamics and Statistical Mechanics of Small Systems)
Show Figures

Figure 1

Figure 1
<p>Scheme of the Kovacs experiment described in the text. The dashed curve on the right, labeled by <math display="inline"> <semantics> <mrow> <mi>ϕ</mi> <mo>(</mo> <mi>t</mi> <mo>)</mo> </mrow> </semantics> </math>, represents the direct relaxation from <math display="inline"> <semantics> <msub> <mi>T</mi> <mn>0</mn> </msub> </semantics> </math> to <span class="html-italic">T</span>. The dashed curve on the left stands for the part of the relaxation from <math display="inline"> <semantics> <msub> <mi>T</mi> <mn>0</mn> </msub> </semantics> </math> to <math display="inline"> <semantics> <msub> <mi>T</mi> <mn>1</mn> </msub> </semantics> </math> that is interrupted at <math display="inline"> <semantics> <mrow> <mi>t</mi> <mo>=</mo> <msub> <mi>t</mi> <mi>w</mi> </msub> </mrow> </semantics> </math> by the second temperature jump, changing abruptly the temperature from <math display="inline"> <semantics> <msub> <mi>T</mi> <mn>1</mn> </msub> </semantics> </math> to <span class="html-italic">T</span>. After this second jump, the system follows the non-monotonic response <math display="inline"> <semantics> <mrow> <mi>K</mi> <mo>(</mo> <mi>t</mi> <mo>)</mo> </mrow> </semantics> </math> (solid line), which reaches a maximum at <math display="inline"> <semantics> <mrow> <mi>t</mi> <mo>=</mo> <msub> <mi>t</mi> <mi>k</mi> </msub> </mrow> </semantics> </math> and, afterwards, approaches <math display="inline"> <semantics> <mrow> <mi>ϕ</mi> <mo>(</mo> <mi>t</mi> <mo>)</mo> </mrow> </semantics> </math> for very long times.</p>
Full article ">Figure 2
<p>(Color online) Direct relaxation of the granular temperature <span class="html-italic">T</span> for different final noise amplitudes. All curves start from the stationary state corresponding to <math display="inline"> <semantics> <mrow> <msub> <mi>χ</mi> <mn>0</mn> </msub> <mo>=</mo> <mn>1</mn> </mrow> </semantics> </math>. We compare Monte Carlo simulation results for a system of <math display="inline"> <semantics> <mrow> <mi>N</mi> <mo>=</mo> <mn>100</mn> </mrow> </semantics> </math> sites (symbols) with the numerical solution of the first Sonine approximation, Equation (38) (solid lines), and the analytic solution of the linear response system, Equation (<a href="#FD49-entropy-19-00539" class="html-disp-formula">44</a>) (dashed lines).</p>
Full article ">Figure 3
<p>(Color online) Kovacs hump in linear response. We have fixed the initial and final drivings, <math display="inline"> <semantics> <mrow> <msub> <mi>χ</mi> <mn>0</mn> </msub> <mo>=</mo> <mn>1.05</mn> </mrow> </semantics> </math> and <math display="inline"> <semantics> <mrow> <mi>χ</mi> <mo>=</mo> <mn>1</mn> </mrow> </semantics> </math>, and considered four values for the intermediate driving <math display="inline"> <semantics> <mrow> <msub> <mi>χ</mi> <mn>1</mn> </msub> <mo>=</mo> <mrow> <mo>{</mo> <mn>0.5</mn> <mo>,</mo> <mn>0.8</mn> <mo>,</mo> <mn>0.95</mn> <mo>,</mo> <mn>0.99</mn> <mo>}</mo> </mrow> </mrow> </semantics> </math>. The linear response theory result (53) (solid line) perfectly agrees with the numerical solution of the first Sonine approximation (symbols), Equation (42). Furthermore, the theoretical prediction for the maximum (<a href="#FD61-entropy-19-00539" class="html-disp-formula">54</a>), which again agrees with the numerical results, is plotted (dotted line).</p>
Full article ">Figure 4
<p>(Color online) Kovacs hump out of the linear regime. The initial driving is much higher than that in <a href="#entropy-19-00539-f003" class="html-fig">Figure 3</a>, <math display="inline"> <semantics> <mrow> <msub> <mi>χ</mi> <mn>0</mn> </msub> <mo>=</mo> <mn>10</mn> </mrow> </semantics> </math>, whereas the final and intermediate values of the driving are again <math display="inline"> <semantics> <mrow> <mi>χ</mi> <mo>=</mo> <mn>1</mn> </mrow> </semantics> </math> and <math display="inline"> <semantics> <mrow> <msub> <mi>χ</mi> <mn>1</mn> </msub> <mo>=</mo> <mrow> <mo>{</mo> <mn>0.5</mn> <mo>,</mo> <mn>0.8</mn> <mo>,</mo> <mn>0.95</mn> <mo>}</mo> </mrow> </mrow> </semantics> </math>. The linear response theoretical expression (<a href="#FD54-entropy-19-00539" class="html-disp-formula">48</a>) (solid line) remains quite below the numerical solutions of the first Sonine approximation (42) (symbols). The theoretical expression for the maximum in linear response (<a href="#FD61-entropy-19-00539" class="html-disp-formula">54</a>) (dotted line) still gives a good description thereof; see also <a href="#entropy-19-00539-f005" class="html-fig">Figure 5</a>.</p>
Full article ">Figure 5
<p>(Color online) Kovacs hump in the nonlinear regime and prediction of the perturbative expansion in <math display="inline"> <semantics> <msubsup> <mi>a</mi> <mrow> <mn>2</mn> </mrow> <mi mathvariant="normal">s</mi> </msubsup> </semantics> </math>. We have considered the following values of the drivings: <math display="inline"> <semantics> <mrow> <msub> <mi>χ</mi> <mn>0</mn> </msub> <mo>=</mo> <mrow> <mo>{</mo> <mn>2</mn> <mo>,</mo> <mn>10</mn> <mo>,</mo> <mn>50</mn> <mo>}</mo> </mrow> </mrow> </semantics> </math>, <math display="inline"> <semantics> <mrow> <msub> <mi>χ</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>0</mn> </mrow> </semantics> </math> and <math display="inline"> <semantics> <mrow> <mi>χ</mi> <mo>=</mo> <mn>1</mn> </mrow> </semantics> </math>. Symbols stand for the numerical solutions of the first Sonine approximation (42), whereas lines correspond to the theoretical expression (<a href="#FD58-entropy-19-00539" class="html-disp-formula">52</a>). For the solid line, <math display="inline"> <semantics> <mrow> <msubsup> <mi>a</mi> <mrow> <mn>2</mn> </mrow> <mi>ini</mi> </msubsup> <mo>=</mo> <msubsup> <mi>a</mi> <mrow> <mn>2</mn> </mrow> <mi>HCS</mi> </msubsup> </mrow> </semantics> </math>, while for the dashed lines, we have used the value of <math display="inline"> <semantics> <msubsup> <mi>a</mi> <mrow> <mn>2</mn> </mrow> <mi>ini</mi> </msubsup> </semantics> </math> in the numerical solution. A perfect agreement is observed. Finally, we have plotted the theoretical expression for the maximum position in nonlinear response (dotted line), Equation (<a href="#FD62-entropy-19-00539" class="html-disp-formula">55</a>), which also shows an excellent agreement with numerics.</p>
Full article ">Figure 6
<p>(Color online) Time evolution of the <span class="html-italic">H</span>-functional. The relaxation of <span class="html-italic">H</span> is shown to be monotonic even for Kovacs-like experiments. The left and right panels correspond to the protocols in <a href="#entropy-19-00539-f003" class="html-fig">Figure 3</a> (filled symbols) and <a href="#entropy-19-00539-f004" class="html-fig">Figure 4</a> (open symbols), that is to the linear and nonlinear regimes. The vertical dotted line marks the theoretical position of the maximum in the corresponding regime.</p>
Full article ">
773 KiB  
Article
Hydrodynamics of a Granular Gas in a Heterogeneous Environment
by Francisco Vega Reyes and Antonio Lasanta
Entropy 2017, 19(10), 536; https://doi.org/10.3390/e19100536 - 11 Oct 2017
Cited by 5 | Viewed by 3746
Abstract
We analyze the transport properties of a low density ensemble of identical macroscopic particles immersed in an active fluid. The particles are modeled as inelastic hard spheres (granular gas). The non-homogeneous active fluid is modeled by means of a non-uniform stochastic thermostat. The [...] Read more.
We analyze the transport properties of a low density ensemble of identical macroscopic particles immersed in an active fluid. The particles are modeled as inelastic hard spheres (granular gas). The non-homogeneous active fluid is modeled by means of a non-uniform stochastic thermostat. The theoretical results are validated with a numerical solution of the corresponding the kinetic equation (direct simulation Monte Carlo method). We show a steady flow in the system that is accurately described by Navier-Stokes (NS) hydrodynamics, even for high inelasticity. Surprisingly, we find that the deviations from NS hydrodynamics for this flow are stronger as the inelasticity decreases. The active fluid action is modeled here with a non-uniform fluctuating volume force. This is a relevant result given that hydrodynamics of particles in complex environments, such as biological crowded environments, is still a question under intense debate. Full article
(This article belongs to the Special Issue Thermodynamics and Statistical Mechanics of Small Systems)
Show Figures

Figure 1

Figure 1
<p>DSMC simulation data (<math display="inline"> <semantics> <mrow> <mi>d</mi> <mo>=</mo> <mn>3</mn> </mrow> </semantics> </math>, spheres) for temperature profile for two different values of inelasticity: <math display="inline"> <semantics> <mrow> <mi>α</mi> <mo>=</mo> <mn>0.9</mn> </mrow> </semantics> </math> (solid symbols) and <math display="inline"> <semantics> <mrow> <mi>α</mi> <mo>=</mo> <mn>0.7</mn> </mrow> </semantics> </math> (open symbols) and with the same boundary condition for wall temperature difference (<math display="inline"> <semantics> <mrow> <msub> <mi>T</mi> <mo>+</mo> </msub> <mo>/</mo> <msub> <mi>T</mi> <mo>−</mo> </msub> <mo>=</mo> <mn>5</mn> </mrow> </semantics> </math>, <math display="inline"> <semantics> <mrow> <mo>Δ</mo> <mi>L</mi> <mo>=</mo> <mn>15</mn> <mover> <mi>λ</mi> <mo>¯</mo> </mover> </mrow> </semantics> </math>) with <math display="inline"> <semantics> <mrow> <mover> <mi>λ</mi> <mo>¯</mo> </mover> <mo>=</mo> <msqrt> <mn>2</mn> </msqrt> <mi>π</mi> <mover> <mi>n</mi> <mo>¯</mo> </mover> <msup> <mi>σ</mi> <mrow> <mi>d</mi> <mo>−</mo> <mn>1</mn> </mrow> </msup> </mrow> </semantics> </math>, <math display="inline"> <semantics> <mover> <mi>n</mi> <mo>¯</mo> </mover> </semantics> </math> being the average density in the system. There is excellent agreement with the theoretical prediction of <math display="inline"> <semantics> <mrow> <mover accent="true"> <mi>T</mi> <mo stretchy="false">^</mo> </mover> <mrow> <mo>(</mo> <mover accent="true"> <mi>y</mi> <mo stretchy="false">^</mo> </mover> <mo>)</mo> </mrow> <mo>∝</mo> <msup> <mi>y</mi> <mrow> <mn>2</mn> <mo>/</mo> <mn>3</mn> </mrow> </msup> </mrow> </semantics> </math> (solid line stands for the NS theoretical profile), for both <math display="inline"> <semantics> <mi>α</mi> </semantics> </math> values.</p>
Full article ">Figure 2
<p>(<b>a</b>) Viscosity vs. coefficient of restitution <math display="inline"> <semantics> <mi>α</mi> </semantics> </math>. (<b>b</b>) Effective thermal conductivity vs. <math display="inline"> <semantics> <mi>α</mi> </semantics> </math>. In both panels, lines represent the results for the theoretical NS coefficient (solid for standard Sonine approximation and dashed for improved Sonine approximation) whereas the symbols stand for DSMC simulations. Several DSMC series have been represented, with increasing relative wall velocity. Solid symbols represent the no-shear case and open symbols are the sheared states for: <math display="inline"> <semantics> <mrow> <mo>Δ</mo> <mi>U</mi> <mo>=</mo> <mn>0.25</mn> </mrow> </semantics> </math> with <math display="inline"> <semantics> <mrow> <mi>a</mi> <mo>≃</mo> <mn>0.019</mn> </mrow> </semantics> </math> (black), <math display="inline"> <semantics> <mrow> <mo>Δ</mo> <mi>U</mi> <mo>=</mo> <mn>1</mn> </mrow> </semantics> </math> with <math display="inline"> <semantics> <mrow> <mi>a</mi> <mo>≃</mo> <mn>0.04</mn> </mrow> </semantics> </math> (blue) , <math display="inline"> <semantics> <mrow> <mo>Δ</mo> <mi>U</mi> <mo>=</mo> <mn>3</mn> </mrow> </semantics> </math> with <math display="inline"> <semantics> <mrow> <mi>a</mi> <mo>≃</mo> <mn>0.10</mn> </mrow> </semantics> </math> (red), <math display="inline"> <semantics> <mrow> <mo>Δ</mo> <mi>U</mi> <mo>=</mo> <mn>5</mn> </mrow> </semantics> </math> with <math display="inline"> <semantics> <mrow> <mi>a</mi> <mo>≃</mo> <mn>0.18</mn> </mrow> </semantics> </math> (green).</p>
Full article ">Figure 3
<p>(<b>a</b>) DSMC data for reduced normal stresses <math display="inline"> <semantics> <mrow> <msub> <mi>θ</mi> <mi>x</mi> </msub> <mo>≡</mo> <msub> <mi>P</mi> <mrow> <mi>x</mi> <mi>x</mi> </mrow> </msub> <mo>/</mo> <mi>p</mi> <mo>,</mo> <msub> <mi>θ</mi> <mi>y</mi> </msub> <mo>≡</mo> <msub> <mi>P</mi> <mrow> <mi>y</mi> <mi>y</mi> </mrow> </msub> <mo>/</mo> <mi>p</mi> </mrow> </semantics> </math> vs. coefficient of restitution <math display="inline"> <semantics> <mi>α</mi> </semantics> </math>. Square symbols stand for <math display="inline"> <semantics> <msub> <mi>θ</mi> <mi>x</mi> </msub> </semantics> </math> and triangles stand for <math display="inline"> <semantics> <msub> <mi>θ</mi> <mi>y</mi> </msub> </semantics> </math>. (<b>b</b>) DSMC data for thermal cross conductivity vs. <math display="inline"> <semantics> <mi>α</mi> </semantics> </math>. In both panels, solid symbols represent the no-shear case and open symbols are the sheared states for: <math display="inline"> <semantics> <mrow> <mo>Δ</mo> <mi>U</mi> <mo>=</mo> <mn>1</mn> </mrow> </semantics> </math> (black) with <math display="inline"> <semantics> <mrow> <mi>a</mi> <mo>≃</mo> <mn>0.04</mn> </mrow> </semantics> </math>, <math display="inline"> <semantics> <mrow> <mo>Δ</mo> <mi>U</mi> <mo>=</mo> <mn>3</mn> </mrow> </semantics> </math> with <math display="inline"> <semantics> <mrow> <mi>a</mi> <mo>≃</mo> <mn>0.10</mn> </mrow> </semantics> </math> (blue), <math display="inline"> <semantics> <mrow> <mo>Δ</mo> <mi>U</mi> <mo>=</mo> <mn>5</mn> </mrow> </semantics> </math> with <math display="inline"> <semantics> <mrow> <mi>a</mi> <mo>≃</mo> <mn>0.18</mn> </mrow> </semantics> </math> (red).</p>
Full article ">
339 KiB  
Article
Participation Ratio for Constraint-Driven Condensation with Superextensive Mass
by Giacomo Gradenigo and Eric Bertin
Entropy 2017, 19(10), 517; https://doi.org/10.3390/e19100517 - 26 Sep 2017
Cited by 14 | Viewed by 3532
Abstract
Broadly distributed random variables with a power-law distribution f ( m ) m - ( 1 + α ) are known to generate condensation effects. This means that, when the exponent α lies in a certain interval, the largest variable in a [...] Read more.
Broadly distributed random variables with a power-law distribution f ( m ) m - ( 1 + α ) are known to generate condensation effects. This means that, when the exponent α lies in a certain interval, the largest variable in a sum of N (independent and identically distributed) terms is for large N of the same order as the sum itself. In particular, when the distribution has infinite mean ( 0 < α < 1 ) one finds unconstrained condensation, whereas for α > 1 constrained condensation takes places fixing the total mass to a large enough value M = i = 1 N m i > M c . In both cases, a standard indicator of the condensation phenomenon is the participation ratio Y k = i m i k / ( i m i ) k ( k > 1 ), which takes a finite value for N when condensation occurs. To better understand the connection between constrained and unconstrained condensation, we study here the situation when the total mass is fixed to a superextensive value M N 1 + δ ( δ > 0 ), hence interpolating between the unconstrained condensation case (where the typical value of the total mass scales as M N 1 / α for α < 1 ) and the extensive constrained mass. In particular we show that for exponents α < 1 a condensate phase for values δ > δ c = 1 / α - 1 is separated from a homogeneous phase at δ < δ c from a transition line, δ = δ c , where a weak condensation phenomenon takes place. We focus on the evaluation of the participation ratio as a generic indicator of condensation, also recalling or presenting results in the standard cases of unconstrained mass and of fixed extensive mass. Full article
(This article belongs to the Special Issue Thermodynamics and Statistical Mechanics of Small Systems)
Show Figures

Figure 1

Figure 1
<p>Phase diagram for the values of the participation rations in the (<math display="inline"> <semantics> <mi>α</mi> </semantics> </math>, <math display="inline"> <semantics> <mi>ρ</mi> </semantics> </math>) plane in the presence of an extensive constraint on the total value of the mass <math display="inline"> <semantics> <mrow> <msubsup> <mo>∑</mo> <mrow> <mi>i</mi> <mo>=</mo> <mn>1</mn> </mrow> <mi>N</mi> </msubsup> <msub> <mi>m</mi> <mi>i</mi> </msub> <mo>=</mo> <mi>ρ</mi> <mi>N</mi> </mrow> </semantics> </math>. The (red) continuous line marks the separation of the condensed phase (<math display="inline"> <semantics> <mrow> <msub> <mo movablelimits="true" form="prefix">lim</mo> <mrow> <mi>N</mi> <mo>→</mo> <mo>∞</mo> </mrow> </msub> <msub> <mi>Y</mi> <mi>k</mi> </msub> <mo>&gt;</mo> <mn>0</mn> </mrow> </semantics> </math>) from the homogeneous phase (<math display="inline"> <semantics> <mrow> <msub> <mo movablelimits="true" form="prefix">lim</mo> <mrow> <mi>N</mi> <mo>→</mo> <mo>∞</mo> </mrow> </msub> <msub> <mi>Y</mi> <mi>k</mi> </msub> <mo>=</mo> <mn>0</mn> </mrow> </semantics> </math>). The vertical dotted (black) line marks the critical value <math display="inline"> <semantics> <mrow> <msub> <mi>α</mi> <mi>c</mi> </msub> <mo>=</mo> <mn>1</mn> </mrow> </semantics> </math> where the critical density <math display="inline"> <semantics> <msub> <mi>ρ</mi> <mi>c</mi> </msub> </semantics> </math> for condensation diverges.</p>
Full article ">Figure 2
<p><span class="html-italic">Main</span>: Phase diagram for the values of the participation ratio in the (<math display="inline"> <semantics> <mi>α</mi> </semantics> </math>, <math display="inline"> <semantics> <mi>δ</mi> </semantics> </math>) plane in the presence of a super-extensive constraint on the total value of the mass <math display="inline"> <semantics> <mrow> <msubsup> <mo>∑</mo> <mrow> <mi>i</mi> <mo>=</mo> <mn>1</mn> </mrow> <mi>N</mi> </msubsup> <msub> <mi>m</mi> <mi>i</mi> </msub> <mo>=</mo> <mover accent="true"> <mi>ρ</mi> <mo>˜</mo> </mover> <msup> <mi>N</mi> <mrow> <mn>1</mn> <mo>+</mo> <mi>δ</mi> </mrow> </msup> </mrow> </semantics> </math>. The (red) continuous line marks the separation of the condensed phase (<math display="inline"> <semantics> <mrow> <msub> <mo movablelimits="true" form="prefix">lim</mo> <mrow> <mi>N</mi> <mo>→</mo> <mo>∞</mo> </mrow> </msub> <msub> <mi>Y</mi> <mi>k</mi> </msub> <mo>&gt;</mo> <mn>0</mn> </mrow> </semantics> </math>) and the homogeneous phase (<math display="inline"> <semantics> <mrow> <msub> <mo movablelimits="true" form="prefix">lim</mo> <mrow> <mi>N</mi> <mo>→</mo> <mo>∞</mo> </mrow> </msub> <msub> <mi>Y</mi> <mi>k</mi> </msub> <mo>=</mo> <mn>0</mn> </mrow> </semantics> </math>). <span class="html-italic">Inset</span>: schematic representation of the marginal probability distribution of the local mass, <math display="inline"> <semantics> <mrow> <mi>p</mi> <mo>(</mo> <mi>m</mi> <mo>)</mo> </mrow> </semantics> </math>, in the presence of condensation, namely in the whole region <math display="inline"> <semantics> <mrow> <mi>δ</mi> <mo>≥</mo> <msub> <mi>δ</mi> <mi>c</mi> </msub> </mrow> </semantics> </math>.</p>
Full article ">
794 KiB  
Article
Far-From-Equilibrium Time Evolution between Two Gamma Distributions
by Eun-jin Kim, Lucille-Marie Tenkès, Rainer Hollerbach and Ovidiu Radulescu
Entropy 2017, 19(10), 511; https://doi.org/10.3390/e19100511 - 22 Sep 2017
Cited by 14 | Viewed by 5162
Abstract
Many systems in nature and laboratories are far from equilibrium and exhibit significant fluctuations, invalidating the key assumptions of small fluctuations and short memory time in or near equilibrium. A full knowledge of Probability Distribution Functions (PDFs), especially time-dependent PDFs, becomes essential in [...] Read more.
Many systems in nature and laboratories are far from equilibrium and exhibit significant fluctuations, invalidating the key assumptions of small fluctuations and short memory time in or near equilibrium. A full knowledge of Probability Distribution Functions (PDFs), especially time-dependent PDFs, becomes essential in understanding far-from-equilibrium processes. We consider a stochastic logistic model with multiplicative noise, which has gamma distributions as stationary PDFs. We numerically solve the transient relaxation problem and show that as the strength of the stochastic noise increases, the time-dependent PDFs increasingly deviate from gamma distributions. For sufficiently strong noise, a transition occurs whereby the PDF never reaches a stationary state, but instead, forms a peak that becomes ever more narrowly concentrated at the origin. The addition of an arbitrarily small amount of additive noise regularizes these solutions and re-establishes the existence of stationary solutions. In addition to diagnostic quantities such as mean value, standard deviation, skewness and kurtosis, the transitions between different solutions are analysed in terms of entropy and information length, the total number of statistically-distinguishable states that a system passes through in time. Full article
(This article belongs to the Special Issue Thermodynamics and Statistical Mechanics of Small Systems)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) shows the result of switching <math display="inline"> <semantics> <mrow> <mi>γ</mi> <mo>=</mo> <mn>0.5</mn> <mo>→</mo> <mn>0.05</mn> </mrow> </semantics> </math>; (<b>b</b>) <math display="inline"> <semantics> <mrow> <mi>γ</mi> <mo>=</mo> <mn>0.05</mn> <mo>→</mo> <mn>0.5</mn> </mrow> </semantics> </math>, both at fixed <math display="inline"> <semantics> <mrow> <mo>ϵ</mo> <mo>=</mo> <mn>1</mn> </mrow> </semantics> </math> and <math display="inline"> <semantics> <mrow> <mi>D</mi> <mo>=</mo> <mn>0.02</mn> </mrow> </semantics> </math>. The initial (red) and final (blue) gamma distributions are shown as heavy lines. The four intermediate lines are when the time-dependent solutions have <math display="inline"> <semantics> <mrow> <mo stretchy="false">〈</mo> <mi>x</mi> <mo stretchy="false">〉</mo> <mo>=</mo> <mn>0.1</mn> <mo>,</mo> <mspace width="4pt"/> <mn>0.2</mn> <mo>,</mo> <mspace width="4pt"/> <mn>0.3</mn> <mo>,</mo> <mspace width="4pt"/> <mn>0.4</mn> </mrow> </semantics> </math>. The arrows are a reminder of the direction of motion, inward on the left and outward on the right.</p>
Full article ">Figure 2
<p>(<b>a</b>) shows <math display="inline"> <semantics> <mrow> <mo stretchy="false">〈</mo> <mi>x</mi> <mo stretchy="false">〉</mo> </mrow> </semantics> </math> as a function of time; (<b>b</b>) shows <math display="inline"> <semantics> <mrow> <mi mathvariant="script">E</mi> <mo>·</mo> <mi>D</mi> </mrow> </semantics> </math> (to indicate the <math display="inline"> <semantics> <mrow> <mi mathvariant="script">E</mi> <mo>∼</mo> <msup> <mi>D</mi> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </msup> </mrow> </semantics> </math> scaling); (<b>c</b>) shows <math display="inline"> <semantics> <mrow> <mi mathvariant="script">L</mi> <mo>·</mo> <msup> <mi>D</mi> <mrow> <mn>1</mn> <mo>/</mo> <mn>2</mn> </mrow> </msup> </mrow> </semantics> </math> (to indicate the <math display="inline"> <semantics> <mrow> <mi mathvariant="script">L</mi> <mo>∼</mo> <msup> <mi>D</mi> <mrow> <mo>−</mo> <mn>1</mn> <mo>/</mo> <mn>2</mn> </mrow> </msup> </mrow> </semantics> </math> scaling). Solid lines denote <math display="inline"> <semantics> <mrow> <mi>γ</mi> <mo>=</mo> <mn>0.5</mn> <mo>→</mo> <mn>0.05</mn> </mrow> </semantics> </math>, dashed lines the reverse. Each solid or dashed “line” is in fact three, occasionally barely distinguishable, lines with <math display="inline"> <semantics> <mrow> <mi>D</mi> <mo>=</mo> <mn>0.01</mn> <mo>,</mo> <mspace width="4pt"/> <mn>0.02</mn> <mo>,</mo> <mspace width="4pt"/> <mn>0.04</mn> </mrow> </semantics> </math>. The dots on the <math display="inline"> <semantics> <mrow> <mo stretchy="false">〈</mo> <mi>x</mi> <mo stretchy="false">〉</mo> </mrow> </semantics> </math> curves correspond to the PDFs shown in <a href="#entropy-19-00511-f001" class="html-fig">Figure 1</a>.</p>
Full article ">Figure 3
<p>(<b>a</b>) shows <math display="inline"> <semantics> <mrow> <mi mathvariant="script">L</mi> <mo>·</mo> <msup> <mi>D</mi> <mrow> <mn>1</mn> <mo>/</mo> <mn>2</mn> </mrow> </msup> </mrow> </semantics> </math> and (<b>b</b>) entropy, both as functions of <math display="inline"> <semantics> <mrow> <mo stretchy="false">〈</mo> <mi>x</mi> <mo stretchy="false">〉</mo> </mrow> </semantics> </math>. Solid lines denote <math display="inline"> <semantics> <mrow> <mi>γ</mi> <mo>=</mo> <mn>0.5</mn> <mo>→</mo> <mn>0.05</mn> </mrow> </semantics> </math>, dashed lines the reverse. Numbers besides curves indicate <math display="inline"> <semantics> <mrow> <mi>D</mi> <mo>=</mo> <mn>0.01</mn> <mo>,</mo> <mspace width="4pt"/> <mn>0.02</mn> <mo>,</mo> <mspace width="4pt"/> <mn>0.04</mn> </mrow> </semantics> </math>. The arrows on the entropy plot are a reminder of the direction of inward/outward motion.</p>
Full article ">Figure 4
<p>(<b>a</b>) <math display="inline"> <semantics> <mrow> <mi>σ</mi> <mo>/</mo> <msup> <mi>D</mi> <mrow> <mn>1</mn> <mo>/</mo> <mn>2</mn> </mrow> </msup> </mrow> </semantics> </math>, (<b>b</b>) (skewness <math display="inline"> <semantics> <mrow> <mo>/</mo> <msup> <mi>D</mi> <mrow> <mn>1</mn> <mo>/</mo> <mn>2</mn> </mrow> </msup> </mrow> </semantics> </math>) and (<b>c</b>) (kurtosis <math display="inline"> <semantics> <mrow> <mo>/</mo> <mi>D</mi> </mrow> </semantics> </math>), as functions of <math display="inline"> <semantics> <mrow> <mo stretchy="false">〈</mo> <mi>x</mi> <mo stretchy="false">〉</mo> </mrow> </semantics> </math>. Solid lines denote <math display="inline"> <semantics> <mrow> <mi>γ</mi> <mo>=</mo> <mn>0.5</mn> <mo>→</mo> <mn>0.05</mn> </mrow> </semantics> </math>, dashed lines the reverse. Numbers besides curves indicate <math display="inline"> <semantics> <mrow> <mi>D</mi> <mo>=</mo> <mn>0.01</mn> <mo>,</mo> <mspace width="4pt"/> <mn>0.02</mn> <mo>,</mo> <mspace width="4pt"/> <mn>0.04</mn> </mrow> </semantics> </math>. The heavy green curves are <math display="inline"> <semantics> <msqrt> <mrow> <mo stretchy="false">〈</mo> <mi>x</mi> <mo stretchy="false">〉</mo> </mrow> </msqrt> </semantics> </math>, <math display="inline"> <semantics> <mrow> <mn>2</mn> <mo>/</mo> <msqrt> <mrow> <mo stretchy="false">〈</mo> <mi>x</mi> <mo stretchy="false">〉</mo> </mrow> </msqrt> </mrow> </semantics> </math> and <math display="inline"> <semantics> <mrow> <mn>6</mn> <mo>/</mo> <mrow> <mo stretchy="false">〈</mo> <mi>x</mi> <mo stretchy="false">〉</mo> </mrow> </mrow> </semantics> </math>, respectively, and indicate the behaviour expected for exact gamma distributions.</p>
Full article ">Figure 5
<p>(<b>a</b>) shows the difference (13) between the actual PDF and the equivalent gamma distribution, as functions of <math display="inline"> <semantics> <mrow> <mo stretchy="false">〈</mo> <mi>x</mi> <mo stretchy="false">〉</mo> </mrow> </semantics> </math>. Solid lines denote <math display="inline"> <semantics> <mrow> <mi>γ</mi> <mo>=</mo> <mn>0.5</mn> <mo>→</mo> <mn>0.05</mn> </mrow> </semantics> </math>, dashed lines the reverse, with arrows also indicating the direction of motion. The dots at <math display="inline"> <semantics> <mrow> <mo stretchy="false">〈</mo> <mi>x</mi> <mo stretchy="false">〉</mo> <mo>=</mo> <mn>0.3</mn> </mrow> </semantics> </math> for <math display="inline"> <semantics> <mrow> <mi>γ</mi> <mo>=</mo> <mn>0.05</mn> <mo>→</mo> <mn>0.5</mn> </mrow> </semantics> </math> and <math display="inline"> <semantics> <mrow> <mo stretchy="false">〈</mo> <mi>x</mi> <mo stretchy="false">〉</mo> <mo>=</mo> <mn>0.1</mn> </mrow> </semantics> </math> for <math display="inline"> <semantics> <mrow> <mi>γ</mi> <mo>=</mo> <mn>0.5</mn> <mo>→</mo> <mn>0.05</mn> </mrow> </semantics> </math>, correspond to the other two panels: (<b>b</b>) compares the <math display="inline"> <semantics> <mrow> <mi>γ</mi> <mo>=</mo> <mn>0.05</mn> <mo>→</mo> <mn>0.5</mn> </mrow> </semantics> </math> PDF with its equivalent gamma distribution; (<b>c</b>) compares the <math display="inline"> <semantics> <mrow> <mi>γ</mi> <mo>=</mo> <mn>0.5</mn> <mo>→</mo> <mn>0.05</mn> </mrow> </semantics> </math> PDF with its equivalent gamma distribution. The actual PDFs in each case are solid (red), and the equivalent gamma distributions are dashed (blue). <math display="inline"> <semantics> <mrow> <mi>D</mi> <mo>=</mo> <mn>0.04</mn> </mrow> </semantics> </math> for both sets.</p>
Full article ">Figure 6
<p>The initial condition is a gamma distribution with <math display="inline"> <semantics> <mrow> <mi>γ</mi> <mo>=</mo> <mn>0.5</mn> </mrow> </semantics> </math>, <math display="inline"> <semantics> <mrow> <mo>ϵ</mo> <mo>=</mo> <mn>1</mn> </mrow> </semantics> </math> and <math display="inline"> <semantics> <mrow> <mi>D</mi> <mo>=</mo> <msup> <mn>10</mn> <mrow> <mo>−</mo> <mn>3</mn> </mrow> </msup> </mrow> </semantics> </math>; <math display="inline"> <semantics> <mi>γ</mi> </semantics> </math> is then switched to zero, and the solution is evolved according to Equation (<a href="#FD3-entropy-19-00511" class="html-disp-formula">3</a>). Numbers besides curves indicate time, from the initial condition at <math display="inline"> <semantics> <mrow> <mi>t</mi> <mo>=</mo> <mn>0</mn> </mrow> </semantics> </math> to the final time 1000. The dashed curves indicate the equivalent gamma distributions having the same <math display="inline"> <semantics> <mrow> <mo stretchy="false">〈</mo> <mi>x</mi> <mo stretchy="false">〉</mo> </mrow> </semantics> </math> and <math display="inline"> <semantics> <mi>σ</mi> </semantics> </math>.</p>
Full article ">Figure 7
<p>As in <a href="#entropy-19-00511-f002" class="html-fig">Figure 2</a>, (<b>a</b>) shows <math display="inline"> <semantics> <mrow> <mo stretchy="false">〈</mo> <mi>x</mi> <mo stretchy="false">〉</mo> </mrow> </semantics> </math>; (<b>b</b>) shows <math display="inline"> <semantics> <mrow> <mi mathvariant="script">E</mi> <mo>·</mo> <mi>D</mi> </mrow> </semantics> </math>; and (<b>c</b>) <math display="inline"> <semantics> <mrow> <mi mathvariant="script">L</mi> <mo>·</mo> <msup> <mi>D</mi> <mrow> <mn>1</mn> <mo>/</mo> <mn>2</mn> </mrow> </msup> </mrow> </semantics> </math>. Solid lines denote <math display="inline"> <semantics> <mrow> <mi>γ</mi> <mo>=</mo> <mn>0.5</mn> <mo>→</mo> <mn>0</mn> </mrow> </semantics> </math> for <math display="inline"> <semantics> <mrow> <mi>D</mi> <mo>=</mo> <msup> <mn>10</mn> <mrow> <mo>−</mo> <mn>3</mn> </mrow> </msup> </mrow> </semantics> </math>, dashed lines the previous <math display="inline"> <semantics> <mrow> <mi>γ</mi> <mo>=</mo> <mn>0.5</mn> <mo>→</mo> <mn>0.05</mn> </mrow> </semantics> </math> for <math display="inline"> <semantics> <mrow> <mi>D</mi> <mo>=</mo> <mn>0.01</mn> </mrow> </semantics> </math>. Note how the scalings of <math display="inline"> <semantics> <mi mathvariant="script">E</mi> </semantics> </math> and <math display="inline"> <semantics> <mi mathvariant="script">L</mi> </semantics> </math> with <span class="html-italic">D</span> are still preserved even when <span class="html-italic">D</span> is changed by a factor of 10.</p>
Full article ">Figure 8
<p>(<b>a</b>) Entropy, (<b>b</b>) <math display="inline"> <semantics> <mrow> <mi>σ</mi> <mo>/</mo> <msup> <mi>D</mi> <mrow> <mn>1</mn> <mo>/</mo> <mn>2</mn> </mrow> </msup> </mrow> </semantics> </math>, (<b>c</b>) (skewness <math display="inline"> <semantics> <mrow> <mo>/</mo> <msup> <mi>D</mi> <mrow> <mn>1</mn> <mo>/</mo> <mn>2</mn> </mrow> </msup> </mrow> </semantics> </math>) and (<b>d</b>) (kurtosis <math display="inline"> <semantics> <mrow> <mo>/</mo> <mi>D</mi> </mrow> </semantics> </math>), as functions of <math display="inline"> <semantics> <mrow> <mo stretchy="false">〈</mo> <mi>x</mi> <mo stretchy="false">〉</mo> </mrow> </semantics> </math>, for the <math display="inline"> <semantics> <mrow> <mi>γ</mi> <mo>=</mo> <mn>0.5</mn> <mo>→</mo> <mn>0</mn> </mrow> </semantics> </math> calculation from <a href="#entropy-19-00511-f006" class="html-fig">Figure 6</a>. The heavy green curves in the last three panels are <math display="inline"> <semantics> <msqrt> <mrow> <mo stretchy="false">〈</mo> <mi>x</mi> <mo stretchy="false">〉</mo> </mrow> </msqrt> </semantics> </math>, <math display="inline"> <semantics> <mrow> <mn>2</mn> <mo>/</mo> <msqrt> <mrow> <mo stretchy="false">〈</mo> <mi>x</mi> <mo stretchy="false">〉</mo> </mrow> </msqrt> </mrow> </semantics> </math> and <math display="inline"> <semantics> <mrow> <mn>6</mn> <mo>/</mo> <mrow> <mo stretchy="false">〈</mo> <mi>x</mi> <mo stretchy="false">〉</mo> </mrow> </mrow> </semantics> </math>, respectively, and indicate the behaviour expected for exact gamma distributions.</p>
Full article ">Figure 9
<p>(<b>a</b>) shows the result of switching <math display="inline"> <semantics> <mrow> <mi>γ</mi> <mo>=</mo> <mn>0.5</mn> <mo>→</mo> <mn>0.1</mn> </mrow> </semantics> </math>, (<b>b</b>) <math display="inline"> <semantics> <mrow> <mi>γ</mi> <mo>=</mo> <mn>0.1</mn> <mo>→</mo> <mn>0.5</mn> </mrow> </semantics> </math>, both at fixed <math display="inline"> <semantics> <mrow> <mo>ϵ</mo> <mo>=</mo> <mn>1</mn> </mrow> </semantics> </math>, <math display="inline"> <semantics> <mrow> <mi>D</mi> <mo>=</mo> <msup> <mn>10</mn> <mrow> <mo>−</mo> <mn>3</mn> </mrow> </msup> </mrow> </semantics> </math> and <math display="inline"> <semantics> <mrow> <mi>Q</mi> <mo>=</mo> <msup> <mn>10</mn> <mrow> <mo>−</mo> <mn>5</mn> </mrow> </msup> </mrow> </semantics> </math>. The initial (red) and final (blue) gamma distributions are shown as heavy lines. The three intermediate lines are when the time-dependent solutions have <math display="inline"> <semantics> <mrow> <mo stretchy="false">〈</mo> <mi>x</mi> <mo stretchy="false">〉</mo> <mo>=</mo> <mn>0.2</mn> <mo>,</mo> <mspace width="4pt"/> <mn>0.3</mn> <mo>,</mo> <mspace width="4pt"/> <mn>0.4</mn> </mrow> </semantics> </math>. <math display="inline"> <semantics> <mrow> <msub> <mi mathvariant="script">L</mi> <mo>∞</mo> </msub> <mo>=</mo> <mn>25</mn> </mrow> </semantics> </math> on the left and 16 on the right.</p>
Full article ">Figure 10
<p>(<b>a</b>) the result of switching <math display="inline"> <semantics> <mrow> <mi>γ</mi> <mo>=</mo> <mn>0.5</mn> <mo>→</mo> <mn>0</mn> </mrow> </semantics> </math>, (<b>b</b>) <math display="inline"> <semantics> <mrow> <mi>γ</mi> <mo>=</mo> <mn>0</mn> <mo>→</mo> <mn>0.5</mn> </mrow> </semantics> </math>, both at fixed <math display="inline"> <semantics> <mrow> <mo>ϵ</mo> <mo>=</mo> <mn>1</mn> </mrow> </semantics> </math>, <math display="inline"> <semantics> <mrow> <mi>D</mi> <mo>=</mo> <msup> <mn>10</mn> <mrow> <mo>−</mo> <mn>3</mn> </mrow> </msup> </mrow> </semantics> </math> and <math display="inline"> <semantics> <mrow> <mi>Q</mi> <mo>=</mo> <msup> <mn>10</mn> <mrow> <mo>−</mo> <mn>5</mn> </mrow> </msup> </mrow> </semantics> </math>. The initial (red) and final (blue) gamma distributions are shown as heavy lines. The four intermediate lines are when the time-dependent solutions have <math display="inline"> <semantics> <mrow> <mo stretchy="false">〈</mo> <mi>x</mi> <mo stretchy="false">〉</mo> <mo>=</mo> <mn>0.1</mn> <mo>,</mo> <mspace width="4pt"/> <mn>0.2</mn> <mo>,</mo> <mspace width="4pt"/> <mn>0.3</mn> <mo>,</mo> <mspace width="4pt"/> <mn>0.4</mn> </mrow> </semantics> </math>. <math display="inline"> <semantics> <mrow> <msub> <mi mathvariant="script">L</mi> <mo>∞</mo> </msub> <mo>=</mo> <mn>35</mn> </mrow> </semantics> </math> on the left and 9.5 on the right.</p>
Full article ">Figure 11
<p>The <math display="inline"> <semantics> <mrow> <mi>γ</mi> <mo>=</mo> <mn>0</mn> <mo>→</mo> <mn>0.5</mn> </mrow> </semantics> </math> process as in <a href="#entropy-19-00511-f010" class="html-fig">Figure 10</a>, but now shown in more detail. The dashed (magenta) curves are the gamma distributions that best fit the two thicker curves at intermediate times. Note how even a “best-fit” is a rather poor approximation to the actual PDFs.</p>
Full article ">
568 KiB  
Article
Thermodynamics of Small Magnetic Particles
by Eugenio E. Vogel, Patricio Vargas, Gonzalo Saravia, Julio Valdes, Antonio Jose Ramirez-Pastor and Paulo M. Centres
Entropy 2017, 19(9), 499; https://doi.org/10.3390/e19090499 - 15 Sep 2017
Cited by 3 | Viewed by 4862
Abstract
In the present paper, we discuss the interpretation of some of the results of the thermodynamics in the case of very small systems. Most of the usual statistical physics is done for systems with a huge number of elements in what is called [...] Read more.
In the present paper, we discuss the interpretation of some of the results of the thermodynamics in the case of very small systems. Most of the usual statistical physics is done for systems with a huge number of elements in what is called the thermodynamic limit, but not all of the approximations done for those conditions can be extended to all properties in the case of objects with less than a thousand elements. The starting point is the Ising model in two dimensions (2D) where an analytic solution exits, which allows validating the numerical techniques used in the present article. From there on, we introduce several variations bearing in mind the small systems such as the nanoscopic or even subnanoscopic particles, which are nowadays produced for several applications. Magnetization is the main property investigated aimed for two singular possible devices. The size of the systems (number of magnetic sites) is decreased so as to appreciate the departure from the results valid in the thermodynamic limit; periodic boundary conditions are eliminated to approach the reality of small particles; 1D, 2D and 3D systems are examined to appreciate the differences established by dimensionality is this small world; upon diluting the lattices, the effect of coordination number (bonding) is also explored; since the 2D Ising model is equivalent to the clock model with q = 2 degrees of freedom, we combine previous results with the supplementary degrees of freedom coming from the variation of q up to q = 20 . Most of the previous results are numeric; however, for the case of a very small system, we obtain the exact partition function to compare with the conclusions coming from our numerical results. Conclusions can be summarized in the following way: the laws of thermodynamics remain the same, but the interpretation of the results, averages and numerical treatments need special care for systems with less than about a thousand constituents, and this might need to be adapted for different properties or devices. Full article
(This article belongs to the Special Issue Thermodynamics and Statistical Mechanics of Small Systems)
Show Figures

Figure 1

Figure 1
<p>Magnetic switches. (<b>a</b>) Top: Under <math display="inline"> <semantics> <msup> <mi>T</mi> <mo>*</mo> </msup> </semantics> </math> the magnetic nanoparticle (nm) sticks to the ferromagnetic material (F) even against a restoring elastic force represented by a zig-zag line. Bottom: As temperature <span class="html-italic">T</span> goes over <math display="inline"> <semantics> <msup> <mi>T</mi> <mo>*</mo> </msup> </semantics> </math>, nm loses enough magnetization, so the elastic force dominates switching off the contact; (<b>b</b>) Top: Under <math display="inline"> <semantics> <msup> <mi>T</mi> <mrow> <mo>*</mo> <mo>*</mo> </mrow> </msup> </semantics> </math>, a weak magnet labeled NS, with high enough <math display="inline"> <semantics> <msub> <mi>T</mi> <mi>C</mi> </msub> </semantics> </math>, attracts the north pole (painted black) of a rod-like nanomagnet made of a material with lower <math display="inline"> <semantics> <msub> <mi>T</mi> <mi>C</mi> </msub> </semantics> </math>; the black part has a conducting element that closes a circuit. Bottom: Over <math display="inline"> <semantics> <msup> <mi>T</mi> <mrow> <mo>*</mo> <mo>*</mo> </mrow> </msup> </semantics> </math>, an internal magnetization reversal occurs in the nanomagnet; the north pole is now in the sector painted white; the rod rotates with respect to an axis going through the center of the rod and perpendicular to the figure; in this new position, the circuit is switched off.</p>
Full article ">Figure 2
<p>Magnetization for a lattice <math display="inline"> <semantics> <mrow> <mi>L</mi> <mo>=</mo> <mn>4</mn> </mrow> </semantics> </math> obtained using the analytic expressions (<a href="#FD16-entropy-19-00499" class="html-disp-formula">16</a>) (up blue triangles) and (<a href="#FD17-entropy-19-00499" class="html-disp-formula">17</a>) (down yellow triangles) to compare with the results obtained from numeric simulations using PBC (empty magenta diamonds) and FBC (empty black squares); the Onsager solution is also plotted to mark the expected <math display="inline"> <semantics> <msub> <mi>T</mi> <mi>C</mi> </msub> </semantics> </math> in the thermodynamic limit (TL) (solid red circles).</p>
Full article ">Figure 3
<p>Solid (black) squares: last instantaneous magnetization <math display="inline"> <semantics> <mrow> <mi>μ</mi> <mo stretchy="false">(</mo> <mi>T</mi> <mo>,</mo> <mn>2</mn> <mi>τ</mi> </mrow> </semantics> </math>) measured after <math display="inline"> <semantics> <mrow> <mn>2</mn> <mo>×</mo> <msup> <mn>10</mn> <mn>6</mn> </msup> </mrow> </semantics> </math> Monte Carlo steps (MCSs) for each temperature; empty circles (red) represent the average magnetization over <math display="inline"> <semantics> <mrow> <mi>ν</mi> <mo>=</mo> </mrow> </semantics> </math> 50,000 instantaneous measurements of magnetization at intervals of 20 MCSs after <math display="inline"> <semantics> <msup> <mn>10</mn> <mn>6</mn> </msup> </semantics> </math> MCSs of equilibration.</p>
Full article ">Figure 4
<p>Simulations on <math display="inline"> <semantics> <mrow> <mn>16</mn> <mo>×</mo> <mn>16</mn> </mrow> </semantics> </math> lattices for different effective coordination numbers <math display="inline"> <semantics> <mi>κ</mi> </semantics> </math> whose values are given in the inset: 2.000 corresponds to PBC; 1.875 to FBC; 1.75 to the dilution of the lattice according to the decoration proposed in Figure 5 where four pieces <math display="inline"> <semantics> <mrow> <mi>ℓ</mi> <mo>×</mo> <mi>ℓ</mi> </mrow> </semantics> </math> with <math display="inline"> <semantics> <mrow> <mi>ℓ</mi> <mo>=</mo> <mn>4</mn> </mrow> </semantics> </math> have been withdrawn; 1.500 corresponds to a similar decoration with <math display="inline"> <semantics> <mrow> <mi>ℓ</mi> <mo>=</mo> <mn>6</mn> </mrow> </semantics> </math>; 1.00 corresponds to a similar decoration with <math display="inline"> <semantics> <mrow> <mi>ℓ</mi> <mo>=</mo> <mn>7</mn> </mrow> </semantics> </math>. <span class="html-italic">B</span> means <span class="html-italic">P</span> when <math display="inline"> <semantics> <mrow> <mi>κ</mi> <mo>=</mo> <mn>2</mn> <mo>.</mo> <mn>000</mn> </mrow> </semantics> </math>, and it represents <span class="html-italic">F</span> for all of the other cases. For clarity, not all curves are shown with error bars.</p>
Full article ">Figure 5
<p>Diluted <math display="inline"> <semantics> <mrow> <mn>16</mn> <mo>×</mo> <mn>16</mn> </mrow> </semantics> </math> lattice upon decoration: four sectors <math display="inline"> <semantics> <mrow> <mn>4</mn> <mo>×</mo> <mn>4</mn> </mrow> </semantics> </math> are removed to leave <math display="inline"> <semantics> <mrow> <mi>N</mi> <mo>=</mo> <mn>192</mn> </mrow> </semantics> </math> spins with a total of <math display="inline"> <semantics> <mrow> <mi>B</mi> <mo>=</mo> <mn>336</mn> </mrow> </semantics> </math> bonds; FBC are imposed to get <math display="inline"> <semantics> <mrow> <mi>κ</mi> <mo>=</mo> <mn>1</mn> <mo>.</mo> <mn>750</mn> </mrow> </semantics> </math>. Other similar decorations are defined in the text and in the previous figure.</p>
Full article ">Figure 6
<p>Average absolute magnetization for lattices of different sizes showing the tails for the artificial remnant magnetization induced by forced ergodicity breaking. The inset shows the decrease with the size of the number of magnetization reversal within 50,000 measurements after equilibration. For clarity, error bars are given for <math display="inline"> <semantics> <mrow> <mi>L</mi> <mo>=</mo> <mn>16</mn> </mrow> </semantics> </math> only.</p>
Full article ">Figure 7
<p>Instantaneous magnetization after <math display="inline"> <semantics> <mrow> <mi>τ</mi> <mo>=</mo> <msup> <mn>10</mn> <mn>6</mn> </msup> </mrow> </semantics> </math> MCSs of equilibration for the lattice <math display="inline"> <semantics> <mrow> <mn>16</mn> <mo>×</mo> <mn>16</mn> </mrow> </semantics> </math>, <math display="inline"> <semantics> <mrow> <mi>κ</mi> <mo>=</mo> <mn>1</mn> <mo>.</mo> <mn>0</mn> </mrow> </semantics> </math>, with FBC at <math display="inline"> <semantics> <mrow> <mi>T</mi> <mo>=</mo> <mn>0</mn> <mo>.</mo> <mn>6</mn> </mrow> </semantics> </math>. Measurements are separated by 20 MCSs.</p>
Full article ">Figure 8
<p>Magnetization for different systems based on <math display="inline"> <semantics> <mrow> <mn>4</mn> <mo>×</mo> <mn>4</mn> </mrow> </semantics> </math> layers where the effective coordination number <math display="inline"> <semantics> <mi>κ</mi> </semantics> </math> varies as given in the inset: 1.50 (one layer ; B = F); 2.0 (two layers <math display="inline"> <semantics> <mrow> <mn>4</mn> <mo>×</mo> <mn>4</mn> </mrow> </semantics> </math>; B = F); 2.167 (three layers <math display="inline"> <semantics> <mrow> <mn>4</mn> <mo>×</mo> <mn>4</mn> </mrow> </semantics> </math>; B = F); 2.25 (four layers <math display="inline"> <semantics> <mrow> <mn>4</mn> <mo>×</mo> <mn>4</mn> </mrow> </semantics> </math>; B = F); 2.50 (four layers <math display="inline"> <semantics> <mrow> <mn>4</mn> <mo>×</mo> <mn>4</mn> </mrow> </semantics> </math>; B = 2F + 1P); 2.75 (four layers <math display="inline"> <semantics> <mrow> <mn>4</mn> <mo>×</mo> <mn>4</mn> </mrow> </semantics> </math>; B = 1F + 2P); 3.0 (four layers <math display="inline"> <semantics> <mrow> <mn>4</mn> <mo>×</mo> <mn>4</mn> </mrow> </semantics> </math>; B = P) (the first curve is truly 2D, and part of it was already included in <a href="#entropy-19-00499-f002" class="html-fig">Figure 2</a>; it is given here for completeness and comparison purposes).</p>
Full article ">Figure 9
<p>Sensitive temperature <math display="inline"> <semantics> <msup> <mi>T</mi> <mo>*</mo> </msup> </semantics> </math> from a magnetic ordered to a disordered phase for the <span class="html-italic">q</span>-states clock model as a function of <math display="inline"> <semantics> <mrow> <mn>1</mn> <mo>/</mo> <mi>q</mi> </mrow> </semantics> </math>. A square lattice <math display="inline"> <semantics> <mrow> <mn>8</mn> <mo>×</mo> <mn>8</mn> </mrow> </semantics> </math> with FBC is used. Measurements are obtained after <math display="inline"> <semantics> <mrow> <mn>5</mn> <mo>×</mo> <msup> <mn>10</mn> <mn>6</mn> </msup> </mrow> </semantics> </math> MCSs for averaging.</p>
Full article ">
4159 KiB  
Article
A Chain, a Bath, a Sink, and a Wall
by Stefano Iubini, Stefano Lepri, Roberto Livi, Gian-Luca Oppo and Antonio Politi
Entropy 2017, 19(9), 445; https://doi.org/10.3390/e19090445 - 25 Aug 2017
Cited by 22 | Viewed by 4429
Abstract
We numerically investigate out-of-equilibrium stationary processes emerging in a Discrete Nonlinear Schrödinger chain in contact with a heat reservoir (a bath) at temperature T L and a pure dissipator (a sink) acting on opposite edges. Long-time molecular-dynamics simulations are performed by evolving the [...] Read more.
We numerically investigate out-of-equilibrium stationary processes emerging in a Discrete Nonlinear Schrödinger chain in contact with a heat reservoir (a bath) at temperature T L and a pure dissipator (a sink) acting on opposite edges. Long-time molecular-dynamics simulations are performed by evolving the equations of motion within a symplectic integration scheme. Mass and energy are steadily transported through the chain from the heat bath to the sink. We observe two different regimes. For small heat-bath temperatures T L and chemical-potentials, temperature profiles across the chain display a non-monotonous shape, remain remarkably smooth and even enter the region of negative absolute temperatures. For larger temperatures T L , the transport of energy is strongly inhibited by the spontaneous emergence of discrete breathers, which act as a thermal wall. A strongly intermittent energy flux is also observed, due to the irregular birth and death of breathers. The corresponding statistics exhibit the typical signature of rare events of processes with large deviations. In particular, the breather lifetime is found to be ruled by a stretched-exponential law. Full article
(This article belongs to the Special Issue Thermodynamics and Statistical Mechanics of Small Systems)
Show Figures

Figure 1

Figure 1
<p>Phase diagram of the DNLS equation in the (<math display="inline"> <semantics> <mrow> <mi>a</mi> <mo>−</mo> <mi>h</mi> </mrow> </semantics> </math>) plane of, respectively, energy and mass densities. The positive-temperature region extends between the ground state <math display="inline"> <semantics> <mrow> <mi>β</mi> <mo>=</mo> <mo>+</mo> <mo>∞</mo> </mrow> </semantics> </math> line (solid blue lower curve) and the <math display="inline"> <semantics> <mrow> <mi>β</mi> <mo>=</mo> <mn>0</mn> </mrow> </semantics> </math> isothermal (red dashed curve). Purple circles show the <math display="inline"> <semantics> <mrow> <mi>μ</mi> <mo>=</mo> <mn>0</mn> </mrow> </semantics> </math> line, which has been determined numerically through equilibrium simulations (data are taken from Reference [<a href="#B13-entropy-19-00445" class="html-bibr">13</a>]). Black curves refer to nonequilibrium profiles obtained by employing a heat bath with parameters <math display="inline"> <semantics> <mrow> <msub> <mi>T</mi> <mi>L</mi> </msub> <mo>=</mo> <mn>3</mn> </mrow> </semantics> </math> and <math display="inline"> <semantics> <mrow> <msub> <mi>μ</mi> <mi>L</mi> </msub> <mo>=</mo> <mn>0</mn> </mrow> </semantics> </math> and a pure dissipator located at the left and right edges of the chain, respectively. Dotted, dot-dashed, dot-dot-dashed and solid curves refer to chain sizes <math display="inline"> <semantics> <mrow> <mi>N</mi> <mo>=</mo> <mn>511</mn> </mrow> </semantics> </math>, 1023, 2047, and 4095, respectively. Upon increasing <span class="html-italic">N</span>, these above profiles tend to enter the negative temperature region. Simulations are performed by evolving the DNLS chain over <math display="inline"> <semantics> <msup> <mn>10</mn> <mn>7</mn> </msup> </semantics> </math> time units after a transient of <math display="inline"> <semantics> <mrow> <mn>4</mn> <mo>×</mo> <msup> <mn>10</mn> <mn>7</mn> </msup> </mrow> </semantics> </math> units. For the system size <math display="inline"> <semantics> <mrow> <mi>N</mi> <mo>=</mo> <mn>4095</mn> </mrow> </semantics> </math>, a further average over 10 independent trajectories is performed.</p>
Full article ">Figure 2
<p>Average profiles of the inverse temperature <math display="inline"> <semantics> <mrow> <mi>β</mi> <mo>(</mo> <mi>x</mi> <mo>)</mo> </mrow> </semantics> </math> (panel (<b>a</b>)) and mass-density <math display="inline"> <semantics> <mrow> <mi>a</mi> <mo>(</mo> <mi>x</mi> <mo>)</mo> </mrow> </semantics> </math> (panel (<b>b</b>)) for a DNLS chain with <math display="inline"> <semantics> <mrow> <mi>N</mi> <mo>=</mo> <mn>4095</mn> </mrow> </semantics> </math> and different temperatures <math display="inline"> <semantics> <msub> <mi>T</mi> <mi>L</mi> </msub> </semantics> </math> of the reservoir acting at the left edge, where <math display="inline"> <semantics> <mrow> <mi>μ</mi> <mo>=</mo> <mn>0</mn> </mrow> </semantics> </math>. The profile <math display="inline"> <semantics> <mrow> <mi>β</mi> <mo>(</mo> <mi>x</mi> <mo>)</mo> </mrow> </semantics> </math> is computed making use of the microcanonical definition of temperature. Simulations are performed evolving the DNLS chain over <math display="inline"> <semantics> <msup> <mn>10</mn> <mn>7</mn> </msup> </semantics> </math> time units after a transient of <math display="inline"> <semantics> <mrow> <mn>4</mn> <mo>×</mo> <msup> <mn>10</mn> <mn>7</mn> </msup> </mrow> </semantics> </math> units. In order to obtain a reasonable smoothing of these time-averaged profiles we have further averaged each of them over 10 independent trajectories.</p>
Full article ">Figure 3
<p>Average profiles of the temperature <math display="inline"> <semantics> <mrow> <mi>T</mi> <mo>(</mo> <mi>x</mi> <mo>)</mo> </mrow> </semantics> </math> for <math display="inline"> <semantics> <mrow> <msub> <mi>T</mi> <mi>L</mi> </msub> <mo>=</mo> <mn>1</mn> </mrow> </semantics> </math> and different system sizes <span class="html-italic">N</span>. The profile <math display="inline"> <semantics> <mrow> <mi>T</mi> <mo>(</mo> <mi>x</mi> <mo>)</mo> </mrow> </semantics> </math> is computed by means of the microcanonical definition of temperature. The inset shows the boundary inverse temperature <math display="inline"> <semantics> <msub> <mi>β</mi> <mi>R</mi> </msub> </semantics> </math> (black circles) and chemical potential <math display="inline"> <semantics> <msub> <mi>μ</mi> <mi>R</mi> </msub> </semantics> </math> (red squares) close to the dissipator side as a function of the system size <span class="html-italic">N</span>. The data refers to the microcanonical definitions of temperature and chemical potential computed on the last 10 sites of the chain. Simulations are performed evolving the system over <math display="inline"> <semantics> <msup> <mn>10</mn> <mn>7</mn> </msup> </semantics> </math> time units after a transient of <math display="inline"> <semantics> <mrow> <mn>4</mn> <mo>×</mo> <msup> <mn>10</mn> <mn>7</mn> </msup> </mrow> </semantics> </math> units. For the system size <math display="inline"> <semantics> <mrow> <mi>N</mi> <mo>=</mo> <mn>4095</mn> </mrow> </semantics> </math> a further average over 10 independent trajectories has been performed.</p>
Full article ">Figure 4
<p>Panel (<b>a</b>): average profiles of inverse temperature <math display="inline"> <semantics> <mi>β</mi> </semantics> </math> for <math display="inline"> <semantics> <mrow> <msub> <mi>T</mi> <mi>L</mi> </msub> <mo>=</mo> <mn>3</mn> </mrow> </semantics> </math> and different system sizes <span class="html-italic">N</span>. The profile <math display="inline"> <semantics> <mi>β</mi> </semantics> </math> is computed by means of the microcanonical definition of temperature. In the inset, an alternative scaling by <math display="inline"> <semantics> <mrow> <mn>1</mn> <mo>/</mo> <msqrt> <mi>N</mi> </msqrt> </mrow> </semantics> </math> is proposed. For the same setup, panels (<b>b</b>,<b>c</b>) show the behavior of the energy profile <span class="html-italic">h</span> and the mass profile <span class="html-italic">a</span>, respectively (again scaling the position by <math display="inline"> <semantics> <msqrt> <mi>N</mi> </msqrt> </semantics> </math>). Simulations are performed evolving the DNLS chain over <math display="inline"> <semantics> <msup> <mn>10</mn> <mn>7</mn> </msup> </semantics> </math> time units after a transient of <math display="inline"> <semantics> <mrow> <mn>4</mn> <mo>×</mo> <msup> <mn>10</mn> <mn>7</mn> </msup> </mrow> </semantics> </math> units. For the system size <math display="inline"> <semantics> <mrow> <mi>N</mi> <mo>=</mo> <mn>4095</mn> </mrow> </semantics> </math>, a further average over ten independent trajectories is performed.</p>
Full article ">Figure 5
<p>Average mass flux <math display="inline"> <semantics> <msub> <mi>j</mi> <mi>a</mi> </msub> </semantics> </math> versus system size <span class="html-italic">N</span> for different reservoir temperatures <math display="inline"> <semantics> <msub> <mi>T</mi> <mi>L</mi> </msub> </semantics> </math>. The black dotted line refers to a power-law decay <math display="inline"> <semantics> <mrow> <msub> <mi>j</mi> <mi>a</mi> </msub> <mo>∼</mo> <msup> <mi>N</mi> <mrow> <mo>−</mo> <mn>1</mn> <mo>/</mo> <mn>2</mn> </mrow> </msup> </mrow> </semantics> </math>. The inset shows the dependence of <math display="inline"> <semantics> <msub> <mi>j</mi> <mi>a</mi> </msub> </semantics> </math> on the reservoir temperature <math display="inline"> <semantics> <msub> <mi>T</mi> <mi>L</mi> </msub> </semantics> </math> for the system size <math display="inline"> <semantics> <mrow> <mi>N</mi> <mo>=</mo> <mn>4095</mn> </mrow> </semantics> </math>. Simulations are performed evolving the DNLS chain over <math display="inline"> <semantics> <msup> <mn>10</mn> <mn>7</mn> </msup> </semantics> </math> time units after a transient of <math display="inline"> <semantics> <mrow> <mn>4</mn> <mo>×</mo> <msup> <mn>10</mn> <mn>7</mn> </msup> </mrow> </semantics> </math> units. For the system size <math display="inline"> <semantics> <mrow> <mi>N</mi> <mo>=</mo> <mn>4095</mn> </mrow> </semantics> </math> we have averaged over 10 independent trajectories.</p>
Full article ">Figure 6
<p>Profiles of heat flux (black lines) for <math display="inline"> <semantics> <mrow> <msub> <mi>T</mi> <mi>L</mi> </msub> <mo>=</mo> <mn>1</mn> </mrow> </semantics> </math> (panel (<b>a</b>)), <math display="inline"> <semantics> <mrow> <msub> <mi>T</mi> <mi>L</mi> </msub> <mo>=</mo> <mn>3</mn> </mrow> </semantics> </math> (panel (<b>b</b>)) and <math display="inline"> <semantics> <mrow> <msub> <mi>T</mi> <mi>L</mi> </msub> <mo>=</mo> <mn>9</mn> </mrow> </semantics> </math> (panel (<b>c</b>)) in a chain with <math display="inline"> <semantics> <mrow> <mi>N</mi> <mo>=</mo> <mn>4095</mn> </mrow> </semantics> </math> lattice sites. For each boundary temperature, we find the following values of mass and energy fluxes: panel (<b>a</b>) <math display="inline"> <semantics> <mrow> <msub> <mi>j</mi> <mi>a</mi> </msub> <mo>=</mo> <mn>3.5</mn> <mo>×</mo> <msup> <mn>10</mn> <mrow> <mo>−</mo> <mn>3</mn> </mrow> </msup> </mrow> </semantics> </math>, <math display="inline"> <semantics> <mrow> <msub> <mi>j</mi> <mi>h</mi> </msub> <mo>=</mo> <mo>−</mo> <mn>3.3</mn> <mo>×</mo> <msup> <mn>10</mn> <mrow> <mo>−</mo> <mn>4</mn> </mrow> </msup> </mrow> </semantics> </math>; panel (<b>b</b>) <math display="inline"> <semantics> <mrow> <msub> <mi>j</mi> <mi>a</mi> </msub> <mo>=</mo> <mn>2.5</mn> <mo>×</mo> <msup> <mn>10</mn> <mrow> <mo>−</mo> <mn>3</mn> </mrow> </msup> </mrow> </semantics> </math>, <math display="inline"> <semantics> <mrow> <msub> <mi>j</mi> <mi>h</mi> </msub> <mo>=</mo> <mn>3.4</mn> <mo>×</mo> <msup> <mn>10</mn> <mrow> <mo>−</mo> <mn>4</mn> </mrow> </msup> </mrow> </semantics> </math>; panel (<b>c</b>) <math display="inline"> <semantics> <mrow> <msub> <mi>j</mi> <mi>a</mi> </msub> <mo>=</mo> <mn>1.2</mn> <mo>×</mo> <msup> <mn>10</mn> <mrow> <mo>−</mo> <mn>4</mn> </mrow> </msup> </mrow> </semantics> </math>, <math display="inline"> <semantics> <mrow> <msub> <mi>j</mi> <mi>h</mi> </msub> <mo>=</mo> <mn>1.5</mn> <mo>×</mo> <msup> <mn>10</mn> <mrow> <mo>−</mo> <mn>4</mn> </mrow> </msup> </mrow> </semantics> </math>. The red dashed line in panel (<b>b</b>) refers to the rescaled profile of the inverse temperature <math display="inline"> <semantics> <mrow> <msup> <mi>β</mi> <mo>′</mo> </msup> <mrow> <mo>(</mo> <mi>x</mi> <mo>)</mo> </mrow> <mo>=</mo> <mi>β</mi> <mrow> <mo>(</mo> <mi>x</mi> <mo>)</mo> </mrow> <mo>/</mo> <mn>5</mn> </mrow> </semantics> </math> measured along the chain (see <a href="#entropy-19-00445-f002" class="html-fig">Figure 2</a>). Panels (<b>d</b>–<b>f</b>) show the profiles of entropy flux for the same temperatures: <math display="inline"> <semantics> <mrow> <msub> <mi>T</mi> <mi>L</mi> </msub> <mo>=</mo> <mn>1</mn> </mrow> </semantics> </math>, <math display="inline"> <semantics> <mrow> <msub> <mi>T</mi> <mi>L</mi> </msub> <mo>=</mo> <mn>3</mn> </mrow> </semantics> </math> and <math display="inline"> <semantics> <mrow> <msub> <mi>T</mi> <mi>L</mi> </msub> <mo>=</mo> <mn>9</mn> </mrow> </semantics> </math>, respectively. Other simulation details are the same as given in <a href="#entropy-19-00445-f002" class="html-fig">Figure 2</a>.</p>
Full article ">Figure 7
<p>(<b>a</b>) Qualitative DB trajectory in a stationary state with <math display="inline"> <semantics> <mrow> <mi>N</mi> <mo>=</mo> <mn>511</mn> </mrow> </semantics> </math> and <math display="inline"> <semantics> <mrow> <msub> <mi>T</mi> <mi>L</mi> </msub> <mo>=</mo> <mn>10</mn> </mrow> </semantics> </math>. Each point of the curve corresponds to the position of the maximum average amplitude of the chain in a temporal window of <math display="inline"> <semantics> <mrow> <mn>5</mn> <mo>×</mo> <msup> <mn>10</mn> <mn>3</mn> </msup> </mrow> </semantics> </math> time units. (<b>b</b>) Temporal evolution of the outgoing mass flux <math display="inline"> <semantics> <msubsup> <mi>j</mi> <mi>a</mi> <mo>∗</mo> </msubsup> </semantics> </math> through the dissipator edge during the same dynamics of panel (a). The flux <math display="inline"> <semantics> <msubsup> <mi>j</mi> <mi>a</mi> <mo>∗</mo> </msubsup> </semantics> </math> is computed every 20 time units as the average amount of mass flowing to the dissipator during such time interval. Higher peaks typically correspond to the breakdown of one or more DBs. Notice that the boundary mass flux can take only positive values because the chain interacts with a pure dissipator.</p>
Full article ">Figure 8
<p>Normalized boundary flux distributions through the dissipator for a stationary state with <math display="inline"> <semantics> <mrow> <msub> <mi>T</mi> <mi>L</mi> </msub> <mo>=</mo> <mn>11</mn> </mrow> </semantics> </math>. Panel (<b>a</b>) shows the mass flux distribution <math display="inline"> <semantics> <mrow> <mi>P</mi> <mfenced separators="" open="(" close=")"> <msubsup> <mi>j</mi> <mi>a</mi> <mo>∗</mo> </msubsup> </mfenced> </mrow> </semantics> </math>, while panel (<b>b</b>) refers to the energy flux distribution <math display="inline"> <semantics> <mrow> <mi>P</mi> <mfenced separators="" open="(" close=")"> <msubsup> <mi>j</mi> <mi>h</mi> <mo>∗</mo> </msubsup> </mfenced> </mrow> </semantics> </math>. Black circles, red squares, and blue triangles refer to system sizes <math display="inline"> <semantics> <mrow> <mi>N</mi> <mo>=</mo> <mn>511</mn> <mo>,</mo> <mn>1023</mn> </mrow> </semantics> </math>, and 2047, respectively. Power-law fits on the largest size <math display="inline"> <semantics> <mrow> <mi>N</mi> <mo>=</mo> <mn>2047</mn> </mrow> </semantics> </math> (see black dashed lines) give <math display="inline"> <semantics> <mrow> <mi>P</mi> <mrow> <mo>(</mo> <msubsup> <mi>j</mi> <mi>a</mi> <mo>∗</mo> </msubsup> <mo>)</mo> </mrow> <mo>∼</mo> <msup> <mrow> <msubsup> <mi>j</mi> <mi>a</mi> <mo>∗</mo> </msubsup> </mrow> <mrow> <mo>−</mo> <mn>4.32</mn> </mrow> </msup> </mrow> </semantics> </math> and <math display="inline"> <semantics> <mrow> <mi>P</mi> <mrow> <mo>(</mo> <msubsup> <mi>j</mi> <mi>h</mi> <mo>∗</mo> </msubsup> <mo>)</mo> </mrow> <mo>∼</mo> <msup> <mrow> <msubsup> <mi>j</mi> <mi>h</mi> <mo>∗</mo> </msubsup> </mrow> <mrow> <mo>−</mo> <mn>3.17</mn> </mrow> </msup> </mrow> </semantics> </math>. Boundary fluxes are sampled by evolving the DNLS chain for a total time <math display="inline"> <semantics> <msub> <mi>t</mi> <mi>f</mi> </msub> </semantics> </math> after a transient of <math display="inline"> <semantics> <mrow> <mn>4</mn> <mo>×</mo> <msup> <mn>10</mn> <mn>7</mn> </msup> </mrow> </semantics> </math> temporal units and averaged over time windows of five temporal units. For the sizes <math display="inline"> <semantics> <mrow> <mi>N</mi> <mo>=</mo> <mn>511</mn> </mrow> </semantics> </math> and <math display="inline"> <semantics> <mrow> <mi>N</mi> <mo>=</mo> <mn>1023</mn> </mrow> </semantics> </math> we have considered a single trajectory with <math display="inline"> <semantics> <mrow> <msub> <mi>t</mi> <mi>f</mi> </msub> <mo>=</mo> <msup> <mn>10</mn> <mn>8</mn> </msup> </mrow> </semantics> </math>. For the size <math display="inline"> <semantics> <mrow> <mi>N</mi> <mo>=</mo> <mn>2047</mn> </mrow> </semantics> </math> the distributions are extracted from five independent trajectories with <math display="inline"> <semantics> <mrow> <msub> <mi>t</mi> <mi>f</mi> </msub> <mo>=</mo> <mn>2</mn> <mo>×</mo> <msup> <mn>10</mn> <mn>7</mn> </msup> </mrow> </semantics> </math>.</p>
Full article ">Figure 9
<p>(<b>a</b>) Probability distribution of the duration of bursts <math display="inline"> <semantics> <msub> <mi>τ</mi> <mi>b</mi> </msub> </semantics> </math>. Note the logarithmic scale on the vertical axis. From an exponential fit <math display="inline"> <semantics> <mrow> <mi>P</mi> <mo>∼</mo> <msup> <mi mathvariant="normal">e</mi> <mrow> <mo>−</mo> <mi>γ</mi> <msub> <mi>τ</mi> <mi>b</mi> </msub> </mrow> </msup> </mrow> </semantics> </math>, we find a decay constant <math display="inline"> <semantics> <mrow> <mi>γ</mi> <mo>=</mo> <mn>5</mn> <mo>×</mo> <msup> <mn>10</mn> <mrow> <mo>−</mo> <mn>4</mn> </mrow> </msup> </mrow> </semantics> </math> (red dashed line). The inset shows the relation between the duration <math display="inline"> <semantics> <msub> <mi>τ</mi> <mi>b</mi> </msub> </semantics> </math> and the amount of mass <math display="inline"> <semantics> <mrow> <mi>δ</mi> <mi>A</mi> </mrow> </semantics> </math> (black dots) and energy <math display="inline"> <semantics> <mrow> <mi>δ</mi> <mi>E</mi> </mrow> </semantics> </math> (red dots) released to the dissipator. (<b>b</b>) Probability distribution of DB lifetimes. In the inset we show that this is compatible with a stretched exponential law <math display="inline"> <semantics> <mrow> <mi>P</mi> <mo>=</mo> <msub> <mi>P</mi> <mn>0</mn> </msub> <msup> <mi mathvariant="normal">e</mi> <mrow> <mo>−</mo> <msubsup> <mi>τ</mi> <mi>l</mi> <mi>σ</mi> </msubsup> </mrow> </msup> </mrow> </semantics> </math>, where <math display="inline"> <semantics> <mrow> <mi>σ</mi> <mo>≃</mo> <mn>0.5</mn> </mrow> </semantics> </math> (red dashed line) and <math display="inline"> <semantics> <msub> <mi>P</mi> <mn>0</mn> </msub> </semantics> </math> is the maximum value of the distribution.</p>
Full article ">
777 KiB  
Article
Parameterization of Coarse-Grained Molecular Interactions through Potential of Mean Force Calculations and Cluster Expansion Techniques
by Anastasios Tsourtis, Vagelis Harmandaris and Dimitrios Tsagkarogiannis
Entropy 2017, 19(8), 395; https://doi.org/10.3390/e19080395 - 1 Aug 2017
Cited by 17 | Viewed by 5999
Abstract
We present a systematic coarse-graining (CG) strategy for many particle molecular systems based on cluster expansion techniques. We construct a hierarchy of coarse-grained Hamiltonians with interaction potentials consisting of two, three and higher body interactions. In this way, the suggested model becomes computationally [...] Read more.
We present a systematic coarse-graining (CG) strategy for many particle molecular systems based on cluster expansion techniques. We construct a hierarchy of coarse-grained Hamiltonians with interaction potentials consisting of two, three and higher body interactions. In this way, the suggested model becomes computationally tractable, since no information from long n-body (bulk) simulations is required in order to develop it, while retaining the fluctuations at the coarse-grained level. The accuracy of the derived cluster expansion based on interatomic potentials is examined over a range of various temperatures and densities and compared to direct computation of the pair potential of mean force. The comparison of the coarse-grained simulations is done on the basis of the structural properties, against detailed all-atom data. On the other hand, by construction, the approximate coarse-grained models retain, in principle, the thermodynamic properties of the atomistic model without the need for any further parameter fitting. We give specific examples for methane and ethane molecules in which the coarse-grained variable is the centre of mass of the molecule. We investigate different temperature (T) and density ( ρ ) regimes, and we examine differences between the methane and ethane systems. Results show that the cluster expansion formalism can be used in order to provide accurate effective pair and three-body CG potentials at high T and low ρ regimes. In the liquid regime, the three-body effective CG potentials give a small improvement over the typical pair CG ones; however, in order to get significantly better results, one needs to consider even higher order terms. Full article
(This article belongs to the Special Issue Thermodynamics and Statistical Mechanics of Small Systems)
Show Figures

Figure 1

Figure 1
<p>Visualization of the partition in (<a href="#FD19-entropy-19-00395" class="html-disp-formula">19</a>) for non-intersecting sets <math display="inline"> <semantics> <mrow> <msub> <mi>V</mi> <mn>1</mn> </msub> <mo>=</mo> <mrow> <mo>{</mo> <mn>2</mn> <mo>,</mo> <mn>3</mn> <mo>,</mo> <mn>4</mn> <mo>,</mo> <mn>5</mn> <mo>}</mo> </mrow> </mrow> </semantics> </math>, <math display="inline"> <semantics> <mrow> <msub> <mi>V</mi> <mn>2</mn> </msub> <mo>=</mo> <mrow> <mo>{</mo> <mn>6</mn> <mo>,</mo> <mn>7</mn> <mo>}</mo> </mrow> </mrow> </semantics> </math>, <math display="inline"> <semantics> <mrow> <msub> <mi>V</mi> <mn>3</mn> </msub> <mo>=</mo> <mrow> <mo>{</mo> <mn>9</mn> <mo>,</mo> <mn>10</mn> <mo>,</mo> <mn>11</mn> <mo>}</mo> </mrow> </mrow> </semantics> </math> in each of which we display by solid lines the connected graphs <math display="inline"> <semantics> <mrow> <msub> <mi>g</mi> <mi>i</mi> </msub> <mo>∈</mo> <msub> <mi mathvariant="script">C</mi> <msub> <mi>V</mi> <mi>i</mi> </msub> </msub> </mrow> </semantics> </math>, <math display="inline"> <semantics> <mrow> <mi>i</mi> <mo>=</mo> <mn>1</mn> <mo>,</mo> <mn>2</mn> <mo>,</mo> <mn>3</mn> </mrow> </semantics> </math>.</p>
Full article ">Figure 2
<p>Snapshot of model systems in atomistic and coarse-grained description. (<b>a</b>,<b>b</b>) Two and three methanes used for the estimation of the coarse-graining (CG) effective potential from isolated molecules; (<b>c</b>) bulk methane liquid.</p>
Full article ">Figure 3
<p>Representation of the two-body PMF, for two isolated molecules, as a function of distance <span class="html-italic">r</span>, through different approximations: geometric averaging, (constrained) force matching and inversion of <math display="inline"> <semantics> <mrow> <mi>g</mi> <mo>(</mo> <mi>r</mi> <mo>)</mo> </mrow> </semantics> </math>. (<b>a</b>) CH<sub>4</sub> at <span class="html-italic">T</span> = 100 K; (<b>b</b>) CH<sub>3</sub>–CH<sub>3</sub> at <span class="html-italic">T</span> = 150 K. For the methane, the corresponding <math display="inline"> <semantics> <mrow> <mi>g</mi> <mo>(</mo> <mi>r</mi> <mo>)</mo> </mrow> </semantics> </math> curve is also shown.</p>
Full article ">Figure 4
<p>Relation of the PMF through cluster expansions and energy averaging at high temperatures, i.e., <math display="inline"> <semantics> <mrow> <msup> <mi>W</mi> <mrow> <mo>(</mo> <mn>2</mn> <mo>)</mo> </mrow> </msup> <mrow> <mo>(</mo> <msub> <mi>r</mi> <mn>1</mn> </msub> <mo>,</mo> <msub> <mi>r</mi> <mn>2</mn> </msub> <mo>)</mo> </mrow> </mrow> </semantics> </math> and <math display="inline"> <semantics> <mrow> <msup> <mi>W</mi> <mrow> <mo>(</mo> <mn>2</mn> <mo>)</mo> <mo>,</mo> <mi>full</mi> </mrow> </msup> <mrow> <mo>(</mo> <msub> <mi>r</mi> <mn>1</mn> </msub> <mo>,</mo> <msub> <mi>r</mi> <mn>2</mn> </msub> <mo>)</mo> </mrow> </mrow> </semantics> </math>, through expansion over <math display="inline"> <semantics> <mi>β</mi> </semantics> </math> for (<b>a</b>) CH<sub>4</sub> at <span class="html-italic">T</span> = 300 K; (<b>b</b>) CH<sub>3</sub>–CH<sub>3</sub> at <span class="html-italic">T</span> = 650 K. As expected from the analytic form and the relation between the two formulas, <math display="inline"> <semantics> <msup> <mi>W</mi> <mrow> <mo>(</mo> <mn>2</mn> <mo>)</mo> </mrow> </msup> </semantics> </math> and <math display="inline"> <semantics> <msup> <mi>W</mi> <mrow> <mo>(</mo> <mn>2</mn> <mo>)</mo> <mo>,</mo> <mi>full</mi> </mrow> </msup> </semantics> </math> tend to converge to the same effective potential.</p>
Full article ">Figure 5
<p>(<b>a</b>) PMF through cluster expansions, using (<a href="#FD22-entropy-19-00395" class="html-disp-formula">22</a>) and (<a href="#FD25-entropy-19-00395" class="html-disp-formula">25</a>) for different temperatures for the CH<sub>4</sub> model; (<b>b</b>) PMF through cluster expansions and energy averaging, i.e., <math display="inline"> <semantics> <mrow> <msup> <mi>W</mi> <mrow> <mo>(</mo> <mn>2</mn> <mo>)</mo> </mrow> </msup> <mrow> <mo>(</mo> <msub> <mi>r</mi> <mn>1</mn> </msub> <mo>,</mo> <msub> <mi>r</mi> <mn>2</mn> </msub> <mo>)</mo> </mrow> </mrow> </semantics> </math> and <math display="inline"> <semantics> <mrow> <msup> <mi>W</mi> <mrow> <mo>(</mo> <mn>2</mn> <mo>)</mo> <mo>,</mo> <mi>full</mi> </mrow> </msup> <mrow> <mo>(</mo> <msub> <mi>r</mi> <mn>1</mn> </msub> <mo>,</mo> <msub> <mi>r</mi> <mn>2</mn> </msub> <mo>)</mo> </mrow> </mrow> </semantics> </math> through expansion over <math display="inline"> <semantics> <mi>β</mi> </semantics> </math> for CH<sub>4</sub> at <span class="html-italic">T</span> = 150 K. The expansion is not valid at this temperature.</p>
Full article ">Figure 6
<p>Potential of mean force at different temperatures (geometric averaging). Two CH<sub>4</sub> molecules at <span class="html-italic">T</span> = 80 K, 100 K, 300 K.</p>
Full article ">Figure 7
<p>RDF from atomistic and CG using pair potential, <math display="inline"> <semantics> <msup> <mi>W</mi> <mrow> <mo>(</mo> <mn>2</mn> <mo>)</mo> </mrow> </msup> </semantics> </math>, for (<b>a</b>) CH<sub>4</sub> at <span class="html-italic">T</span> = 80 K and (<b>b</b>) CH<sub>3</sub>–CH<sub>3</sub> at <span class="html-italic">T</span> = 150 K. Spherical CG approximation to the non-symmetric ethane molecule induces discrepancy and implies that there is more room for improvement.</p>
Full article ">Figure 8
<p>RDF of methane from atomistic data and CG models using pair potential at different temperatures: (<b>a</b>) <span class="html-italic">T</span> = 300 K; (<b>b</b>) <span class="html-italic">T</span> = 900 K. In both cases, the density is <math display="inline"> <semantics> <mrow> <msub> <mi>ρ</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>0.399</mn> <mfrac> <mi mathvariant="normal">g</mi> <msup> <mrow> <mi>cm</mi> </mrow> <mn>3</mn> </msup> </mfrac> </mrow> </semantics> </math>.</p>
Full article ">Figure 9
<p>RDF of methane from atomistic and CG using pair potential at different densities <math display="inline"> <semantics> <mrow> <msub> <mi>ρ</mi> <mn>1</mn> </msub> <mo>&gt;</mo> <msub> <mi>ρ</mi> <mn>2</mn> </msub> </mrow> </semantics> </math>. (<b>a</b>) <math display="inline"> <semantics> <mrow> <mi>T</mi> <mo>=</mo> <mn>300</mn> <mi>K</mi> </mrow> </semantics> </math>; (<b>b</b>) <span class="html-italic">T</span> = 900 K. For this model, the pair approximation is sufficient, and in low <math display="inline"> <semantics> <mi>ρ</mi> </semantics> </math>, high <span class="html-italic">T</span> conditions, <math display="inline"> <semantics> <msup> <mi>W</mi> <mrow> <mo>(</mo> <mn>2</mn> <mo>)</mo> </mrow> </msup> </semantics> </math> converges to the reference <math display="inline"> <semantics> <mrow> <mi>g</mi> <mo>(</mo> <mi>r</mi> <mo>)</mo> </mrow> </semantics> </math>.</p>
Full article ">Figure 10
<p>Effective potential comparison between the <math display="inline"> <semantics> <msup> <mi>W</mi> <mrow> <mo>(</mo> <mn>3</mn> <mo>)</mo> <mo>,</mo> <mi>full</mi> </mrow> </msup> </semantics> </math> three-body and <math display="inline"> <semantics> <mrow> <mo>∑</mo> <msup> <mi>W</mi> <mrow> <mo>(</mo> <mn>2</mn> <mo>)</mo> <mo>,</mo> <mi>full</mi> </mrow> </msup> </mrow> </semantics> </math> simulations (geometric averaging) for CH<sub>4</sub> at <span class="html-italic">T</span> = 80 K for different fixed centre of mass (COM) distances. (<b>a</b>) <math display="inline"> <semantics> <mrow> <msub> <mi>r</mi> <mn>12</mn> </msub> <mo>=</mo> <mspace width="3.33333pt"/> <mn>3</mn> <mo>.</mo> <mn>9</mn> <mo>,</mo> <msub> <mi>r</mi> <mn>13</mn> </msub> <mo>=</mo> <mn>3.9</mn> </mrow> </semantics> </math>; (<b>b</b>) <math display="inline"> <semantics> <mrow> <msub> <mi>r</mi> <mn>12</mn> </msub> <mo>=</mo> <mn>4.0</mn> <mo>,</mo> <msub> <mi>r</mi> <mn>13</mn> </msub> <mo>=</mo> <mn>4.0</mn> </mrow> </semantics> </math>; (<b>c</b>) <math display="inline"> <semantics> <mrow> <msub> <mi>r</mi> <mn>12</mn> </msub> <mo>=</mo> <mn>4.3</mn> <mo>,</mo> <msub> <mi>r</mi> <mn>13</mn> </msub> <mo>=</mo> <mn>4.0</mn> </mrow> </semantics> </math>; (<b>d</b>) <math display="inline"> <semantics> <mrow> <msub> <mi>r</mi> <mn>12</mn> </msub> <mo>=</mo> <mn>3.8</mn> <mo>,</mo> <msub> <mi>r</mi> <mn>13</mn> </msub> <mo>=</mo> <mn>5.64</mn> </mrow> </semantics> </math>.</p>
Full article ">Figure 11
<p>RDF from atomistic and CG using pair, <math display="inline"> <semantics> <msup> <mi>W</mi> <mrow> <mo>(</mo> <mn>2</mn> <mo>)</mo> <mo>,</mo> <mi>full</mi> </mrow> </msup> </semantics> </math>, and three-body, <math display="inline"> <semantics> <msup> <mi>W</mi> <mrow> <mo>(</mo> <mn>3</mn> <mo>)</mo> <mo>,</mo> <mi>full</mi> </mrow> </msup> </semantics> </math>, potential for CH<sub>4</sub> (<span class="html-italic">T</span> = 80 K). The three-dimensional cubic polynomial was used for the fitting.</p>
Full article ">
550 KiB  
Article
An Application of Pontryagin’s Principle to Brownian Particle Engineered Equilibration
by Paolo Muratore-Ginanneschi and Kay Schwieger
Entropy 2017, 19(7), 379; https://doi.org/10.3390/e19070379 - 24 Jul 2017
Cited by 10 | Viewed by 4334
Abstract
We present a stylized model of controlled equilibration of a small system in a fluctuating environment. We derive the optimal control equations steering in finite-time the system between two equilibrium states. The corresponding thermodynamic transition is optimal in the sense that it occurs [...] Read more.
We present a stylized model of controlled equilibration of a small system in a fluctuating environment. We derive the optimal control equations steering in finite-time the system between two equilibrium states. The corresponding thermodynamic transition is optimal in the sense that it occurs at minimum entropy if the set of admissible controls is restricted by certain bounds on the time derivatives of the protocols. We apply our equations to the engineered equilibration of an optical trap considered in a recent proof of principle experiment. We also analyze an elementary model of nucleation previously considered by Landauer to discuss the thermodynamic cost of one bit of information erasure. We expect our model to be a useful benchmark for experiment design as it exhibits the same integrability properties of well-known models of optimal mass transport by a compressible velocity field. Full article
(This article belongs to the Special Issue Thermodynamics and Statistical Mechanics of Small Systems)
Show Figures

Figure 1

Figure 1
<p>First <a href="#entropy-19-00379-f001" class="html-fig">Figure 1</a>a and second law <a href="#entropy-19-00379-f001" class="html-fig">Figure 1</a>b of thermodynamics for the same transition between Gaussian states as in [<a href="#B4-entropy-19-00379" class="html-bibr">4</a>]. The initial state is a normal distribution with variance <math display="inline"> <semantics> <msup> <mi>β</mi> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </msup> </semantics> </math>. The final distribution is Gaussian with variance <math display="inline"> <semantics> <mrow> <msup> <mi>β</mi> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </msup> <mo>/</mo> <mn>2</mn> </mrow> </semantics> </math>. The condition <math display="inline"> <semantics> <mrow> <mi>K</mi> <mo>(</mo> <mi mathvariant="italic">q</mi> <mo>)</mo> <mo>∝</mo> <mo>|</mo> <mi>ℓ</mi> <mo>(</mo> <mi mathvariant="italic">q</mi> <mo>)</mo> <mo>−</mo> <mi mathvariant="italic">q</mi> <mo>|</mo> </mrow> </semantics> </math> ensures that the probability density remains Gaussian at any time in the control horizon <math display="inline"> <semantics> <mrow> <mi>t</mi> <mo>∈</mo> <mo>[</mo> <mn>0</mn> <mo>,</mo> <mn>1</mn> <mo>]</mo> </mrow> </semantics> </math>. The proportionality factor is chosen such that <math display="inline"> <semantics> <mrow> <msub> <mi>t</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>0</mn> <mo>.</mo> <mn>3</mn> </mrow> </semantics> </math> in (<a href="#FD35-entropy-19-00379" class="html-disp-formula">32</a>). The behavior of the variance (inset of <a href="#entropy-19-00379-f001" class="html-fig">Figure 1</a>a) is qualitatively identical to the one observed in [<a href="#B4-entropy-19-00379" class="html-bibr">4</a>] (<a href="#entropy-19-00379-f002" class="html-fig">Figure 2</a>). The behavior of the average work and heat also reproduces the one of <a href="#entropy-19-00379-f003" class="html-fig">Figure 3</a> of [<a href="#B4-entropy-19-00379" class="html-bibr">4</a>]. (<b>a</b>) Work (continuous curve, blue on-line) and heat release (dashed curve, yellow on-line) during the control horizon. Inset: time evolution of the variance of the process; (<b>b</b>) Entropy production (continuous curve, blue on-line) and heat release (dashed curve, yellow on-line) during the control horizon.</p>
Full article ">Figure 2
<p>Initial (solid curve, blue on-line) and final (dashed curve, blue on-line) probability distribution of the state of the system for <math display="inline"> <semantics> <mrow> <mi>β</mi> <mo>=</mo> <mn>112</mn> </mrow> </semantics> </math><math display="inline"> <semantics> <mrow> <mi>σ</mi> <mo>=</mo> <mn>1</mn> </mrow> </semantics> </math> and <math display="inline"> <semantics> <mrow> <mover accent="true"> <mi mathvariant="italic">q</mi> <mo stretchy="false">¯</mo> </mover> <mo>=</mo> <mn>1</mn> <mo>/</mo> <mn>2</mn> </mrow> </semantics> </math>. The evaluation of the Lagrangian map occasions numerical stiffness in the region in between the two minima. (<b>a</b>) Boundary conditions for the nucleation problem; (<b>b</b>) Lagrangian map; (<b>c</b>) Numerical derivative of the Lagrangian map.</p>
Full article ">Figure 3
<p>Probability density snapshots at different times within the control horizon. The plots are for<math display="inline"> <semantics> <mrow> <mi mathvariant="italic">T</mi> <mo>=</mo> <mn>1</mn> </mrow> </semantics> </math> and switching time <math display="inline"> <semantics> <mrow> <msub> <mi>t</mi> <mn>1</mn> </msub> <mspace width="3.33333pt"/> <mo>=</mo> <mspace width="3.33333pt"/> <msup> <mn>10</mn> <mrow> <mo>−</mo> <mn>6</mn> </mrow> </msup> </mrow> </semantics> </math> (dashed interpolation curve, yellow on-line) and <math display="inline"> <semantics> <mrow> <msub> <mi>t</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>0</mn> <mo>.</mo> <mn>3</mn> </mrow> </semantics> </math> (continuous interpolation curve, blue on-line) <math display="inline"> <semantics> <mrow> <mover accent="true"> <mi mathvariant="italic">q</mi> <mo stretchy="false">¯</mo> </mover> <mo>=</mo> <mn>0</mn> <mo>.</mo> <mn>5</mn> </mrow> </semantics> </math>, <math display="inline"> <semantics> <mrow> <mi>σ</mi> <mo>=</mo> <mn>1</mn> </mrow> </semantics> </math> and <math display="inline"> <semantics> <mrow> <mi>β</mi> <mo>=</mo> <mn>112</mn> </mrow> </semantics> </math>. We plot the Lagrangian map in the interval <math display="inline"> <semantics> <mrow> <mover accent="true"> <mi mathvariant="italic">q</mi> <mo stretchy="false">¯</mo> </mover> <mo>∈</mo> <mo>[</mo> <mrow> <mo>−</mo> <mn>2</mn> </mrow> <mo>,</mo> <mn>2</mn> <mo>]</mo> </mrow> </semantics> </math>.</p>
Full article ">Figure 4
<p>Current velocity snapshots at different times within the control horizon. The plots are for <math display="inline"> <semantics> <mrow> <mi mathvariant="italic">T</mi> <mo>=</mo> <mn>1</mn> </mrow> </semantics> </math> and switching time <math display="inline"> <semantics> <mrow> <msub> <mi>t</mi> <mn>1</mn> </msub> <mspace width="3.33333pt"/> <mo>=</mo> <mspace width="3.33333pt"/> <msup> <mn>10</mn> <mrow> <mo>−</mo> <mn>6</mn> </mrow> </msup> </mrow> </semantics> </math> (continuous interpolation, yellow on-line) and <math display="inline"> <semantics> <mrow> <mi>t</mi> <mo>=</mo> <mn>0</mn> <mo>.</mo> <mn>3</mn> </mrow> </semantics> </math> (points, blue on-line).</p>
Full article ">Figure 5
<p>First and second law of thermodynamics for the optimally-controlled nucleation transition. All parameters are as in <a href="#entropy-19-00379-f002" class="html-fig">Figure 2</a>. The qualitative picture is the same as in the Gaussian case, <a href="#entropy-19-00379-f001" class="html-fig">Figure 1</a>, with the running average work above the running average heat. The numerical values yield, however, almost overlapping curves. The running average entropy production in <a href="#entropy-19-00379-f005" class="html-fig">Figure 5</a>b is strictly monotonic in the control horizon. The entropy production rate vanishes at the boundary highlighting the reaching of an equilibrium state when the switching time is <math display="inline"> <semantics> <mrow> <msub> <mi>t</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>0</mn> <mo>.</mo> <mn>3</mn> </mrow> </semantics> </math>. (<b>a</b>) First law of thermodynamics for the optimally-controlled nucleation. Continuous curve (blue on-line) running average work. Dashed curve (yellow on-line) running average heat; (<b>b</b>) Running average entropy production. The continuous curve (blue on-line) is obtained for switching time at <math display="inline"> <semantics> <mrow> <msub> <mi>t</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>0</mn> <mo>.</mo> <mn>3</mn> </mrow> </semantics> </math>, the dashed curve (yellow on-line) for <math display="inline"> <semantics> <mrow> <msub> <mi>t</mi> <mn>1</mn> </msub> <mspace width="3.33333pt"/> <mo>=</mo> <mspace width="3.33333pt"/> <msup> <mn>10</mn> <mrow> <mo>−</mo> <mn>6</mn> </mrow> </msup> </mrow> </semantics> </math>.</p>
Full article ">Figure 6
<p>Qualitative comparison of universal part of the running Lagrangian maps (<a href="#FD35-entropy-19-00379" class="html-disp-formula">32</a>) (continuous curve, blue on line) and (<a href="#FD43-entropy-19-00379" class="html-disp-formula">40</a>) (dashed curve, orange on line), <a href="#entropy-19-00379-f006" class="html-fig">Figure 6</a>a. In (<a href="#FD43-entropy-19-00379" class="html-disp-formula">40</a>), we choose <math display="inline"> <semantics> <mrow> <mi>τ</mi> <mo>=</mo> <mn>1</mn> </mrow> </semantics> </math>, <math display="inline"> <semantics> <mrow> <mi>ε</mi> <mo>=</mo> <mn>0</mn> <mo>.</mo> <mn>3</mn> </mrow> </semantics> </math>. <a href="#entropy-19-00379-f006" class="html-fig">Figure 6</a>b evinces, as to be expected, the qualitatively equivalent behaviors of the entropy production for finite value (<math display="inline"> <semantics> <mrow> <msub> <mi>t</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>0</mn> <mo>.</mo> <mn>3</mn> </mrow> </semantics> </math>) of the switching time. The dashed green line is computed from (<a href="#FD43-entropy-19-00379" class="html-disp-formula">40</a>). The continuous blue line is the lower bound for the transition as predicted by [<a href="#B7-entropy-19-00379" class="html-bibr">7</a>]. (<b>a</b>) <math display="inline"> <semantics> <mfrac> <mrow> <msub> <mi>χ</mi> <mi>t</mi> </msub> <mo>−</mo> <mi mathvariant="italic">q</mi> </mrow> <mrow> <mi>ℓ</mi> <mo>(</mo> <mi mathvariant="italic">q</mi> <mo>)</mo> <mo>−</mo> <mi mathvariant="italic">q</mi> </mrow> </mfrac> </semantics> </math>; (<b>b</b>) Running entropy production.</p>
Full article ">
269 KiB  
Article
Clausius Relation for Active Particles: What Can We Learn from Fluctuations
by Andrea Puglisi and Umberto Marini Bettolo Marconi
Entropy 2017, 19(7), 356; https://doi.org/10.3390/e19070356 - 13 Jul 2017
Cited by 44 | Viewed by 5086
Abstract
Many kinds of active particles, such as bacteria or active colloids, move in a thermostatted fluid by means of self-propulsion. Energy injected by such a non-equilibrium force is eventually dissipated as heat in the thermostat. Since thermal fluctuations are much faster and weaker [...] Read more.
Many kinds of active particles, such as bacteria or active colloids, move in a thermostatted fluid by means of self-propulsion. Energy injected by such a non-equilibrium force is eventually dissipated as heat in the thermostat. Since thermal fluctuations are much faster and weaker than self-propulsion forces, they are often neglected, blurring the identification of dissipated heat in theoretical models. For the same reason, some freedom—or arbitrariness—appears when defining entropy production. Recently three different recipes to define heat and entropy production have been proposed for the same model where the role of self-propulsion is played by a Gaussian coloured noise. Here we compare and discuss the relation between such proposals and their physical meaning. One of these proposals takes into account the heat exchanged with a non-equilibrium active bath: such an “active heat” satisfies the original Clausius relation and can be experimentally verified. Full article
(This article belongs to the Special Issue Thermodynamics and Statistical Mechanics of Small Systems)
1313 KiB  
Article
Fourier’s Law in a Generalized Piston Model
by Lorenzo Caprini, Luca Cerino, Alessandro Sarracino and Angelo Vulpiani
Entropy 2017, 19(7), 350; https://doi.org/10.3390/e19070350 - 11 Jul 2017
Cited by 6 | Viewed by 3856
Abstract
A simplified, but non trivial, mechanical model—gas of N particles of mass m in a box partitioned by n mobile adiabatic walls of mass M—interacting with two thermal baths at different temperatures, is discussed in the framework of kinetic theory. Following an [...] Read more.
A simplified, but non trivial, mechanical model—gas of N particles of mass m in a box partitioned by n mobile adiabatic walls of mass M—interacting with two thermal baths at different temperatures, is discussed in the framework of kinetic theory. Following an approach due to Smoluchowski, from an analysis of the collisions particles/walls, we derive the values of the main thermodynamic quantities for the stationary non-equilibrium states. The results are compared with extensive numerical simulations; in the limit of large n, m N / M 1 and m / M 1 , we find a good approximation of Fourier’s law. Full article
(This article belongs to the Special Issue Thermodynamics and Statistical Mechanics of Small Systems)
Show Figures

Figure 1

Figure 1
<p>Sketch of the system.</p>
Full article ">Figure 2
<p>Piston temperatures for a 3-piston system, with parameters <math display="inline"> <semantics> <mrow> <msub> <mi>T</mi> <mi>A</mi> </msub> <mo>=</mo> <mn>15</mn> </mrow> </semantics> </math>, <math display="inline"> <semantics> <mrow> <mi>N</mi> <mo>=</mo> <mi>M</mi> <mo>=</mo> <mn>100</mn> </mrow> </semantics> </math>, <math display="inline"> <semantics> <mrow> <mi>m</mi> <mo>=</mo> <mn>1</mn> </mrow> </semantics> </math>, <math display="inline"> <semantics> <mrow> <mi>L</mi> <mo>=</mo> <mn>1</mn> </mrow> </semantics> </math>.</p>
Full article ">Figure 3
<p>Temperature profiles for a 22-piston system with <math display="inline"> <semantics> <mrow> <mi>R</mi> <mo>=</mo> <mn>1</mn> </mrow> </semantics> </math> and <math display="inline"> <semantics> <mrow> <mi>R</mi> <mo>=</mo> <mn>10</mn> </mrow> </semantics> </math>. Other parameters are <math display="inline"> <semantics> <mrow> <msub> <mi>T</mi> <mi>A</mi> </msub> <mo>=</mo> <mn>15</mn> <mo>,</mo> <msub> <mi>T</mi> <mi>B</mi> </msub> <mo>=</mo> <mn>30</mn> <mo>,</mo> <mi>M</mi> <mo>=</mo> <mn>50</mn> <mo>,</mo> <mi>m</mi> <mo>=</mo> <mn>1</mn> <mo>,</mo> <mi>L</mi> <mo>=</mo> <mn>1</mn> </mrow> </semantics> </math>.</p>
Full article ">Figure 4
<p>Probability distribution function of the velocities <math display="inline"> <semantics> <mrow> <msub> <mi>ϕ</mi> <mi>L</mi> </msub> <mrow> <mo>(</mo> <mi>v</mi> <mo>)</mo> </mrow> </mrow> </semantics> </math> (<math display="inline"> <semantics> <mrow> <msub> <mi>ϕ</mi> <mi>R</mi> </msub> <mrow> <mo>(</mo> <mi>v</mi> <mo>)</mo> </mrow> </mrow> </semantics> </math>) of the particles colliding with the first piston from the left (right), for a system with <math display="inline"> <semantics> <mrow> <mi>n</mi> <mo>=</mo> <mn>2</mn> </mrow> </semantics> </math>. The curves, red and blue, are respectively: <math display="inline"> <semantics> <mrow> <msub> <mi>ϕ</mi> <mi>L</mi> </msub> <mo>=</mo> <mfrac> <mi>m</mi> <mrow> <msub> <mi>k</mi> <mi>B</mi> </msub> <msub> <mi>T</mi> <mi>L</mi> </msub> </mrow> </mfrac> <msup> <mi>e</mi> <mrow> <mo>−</mo> <msup> <mi>v</mi> <mn>2</mn> </msup> <mo>/</mo> <msub> <mi>k</mi> <mi>B</mi> </msub> <msub> <mi>T</mi> <mi>L</mi> </msub> </mrow> </msup> <mi>v</mi> <mspace width="0.166667em"/> <mi>θ</mi> <mrow> <mo>(</mo> <mo>−</mo> <mi>v</mi> <mo>)</mo> </mrow> </mrow> </semantics> </math> and <math display="inline"> <semantics> <mrow> <msub> <mi>ϕ</mi> <mi>R</mi> </msub> <mo>=</mo> <mfrac> <mi>m</mi> <mrow> <msub> <mi>k</mi> <mi>B</mi> </msub> <msub> <mi>T</mi> <mi>R</mi> </msub> </mrow> </mfrac> <msup> <mi>e</mi> <mrow> <mo>−</mo> <msup> <mi>v</mi> <mn>2</mn> </msup> <mo>/</mo> <msub> <mi>k</mi> <mi>B</mi> </msub> <msub> <mi>T</mi> <mi>R</mi> </msub> </mrow> </msup> <mi>v</mi> <mspace width="0.166667em"/> <mi>θ</mi> <mrow> <mo>(</mo> <mi>v</mi> <mo>)</mo> </mrow> </mrow> </semantics> </math>, where <math display="inline"> <semantics> <mrow> <msub> <mi>T</mi> <mi>L</mi> </msub> <mo>=</mo> <mn>15.0</mn> </mrow> </semantics> </math> and <math display="inline"> <semantics> <mrow> <msub> <mi>T</mi> <mi>R</mi> </msub> <mo>=</mo> <mn>21.6</mn> </mrow> </semantics> </math> are the temperatures of the gas on the left and on the right with respect to the first piston. Other parameters are: <math display="inline"> <semantics> <mrow> <msub> <mi>T</mi> <mi>A</mi> </msub> <mo>=</mo> <mn>15</mn> </mrow> </semantics> </math>, <math display="inline"> <semantics> <mrow> <msub> <mi>T</mi> <mi>B</mi> </msub> <mo>=</mo> <mn>30</mn> </mrow> </semantics> </math>, <math display="inline"> <semantics> <mrow> <mi>N</mi> <mo>=</mo> <mn>100</mn> </mrow> </semantics> </math>, <math display="inline"> <semantics> <mrow> <mi>M</mi> <mo>=</mo> <mn>100</mn> </mrow> </semantics> </math>, <math display="inline"> <semantics> <mrow> <mi>m</mi> <mo>=</mo> <mn>1</mn> </mrow> </semantics> </math>, <math display="inline"> <semantics> <mrow> <mi>n</mi> <mo>=</mo> <mn>2</mn> </mrow> </semantics> </math>, <math display="inline"> <semantics> <mrow> <mi>L</mi> <mo>=</mo> <mn>1</mn> </mrow> </semantics> </math>.</p>
Full article ">Figure 5
<p>Piston temperatures for a 4-piston system as a function of <span class="html-italic">R</span>. Dashed lines are the theoretical predictions. Other parameters are <math display="inline"> <semantics> <mrow> <msub> <mi>T</mi> <mi>A</mi> </msub> <mo>=</mo> <mn>15</mn> <mo>,</mo> <msub> <mi>T</mi> <mi>B</mi> </msub> <mo>=</mo> <mn>30</mn> <mo>,</mo> <mi>m</mi> <mo>=</mo> <mn>1</mn> <mo>,</mo> <mi>L</mi> <mo>=</mo> <mn>1</mn> </mrow> </semantics> </math>.</p>
Full article ">
1575 KiB  
Article
Information-Theoretic Bound on the Entropy Production to Maintain a Classical Nonequilibrium Distribution Using Ancillary Control
by Jordan M. Horowitz and Jeremey L. England
Entropy 2017, 19(7), 333; https://doi.org/10.3390/e19070333 - 4 Jul 2017
Cited by 8 | Viewed by 4224
Abstract
There are many functional contexts where it is desirable to maintain a mesoscopic system in a nonequilibrium state. However, such control requires an inherent energy dissipation. In this article, we unify and extend a number of works on the minimum energetic cost to [...] Read more.
There are many functional contexts where it is desirable to maintain a mesoscopic system in a nonequilibrium state. However, such control requires an inherent energy dissipation. In this article, we unify and extend a number of works on the minimum energetic cost to maintain a mesoscopic system in a prescribed nonequilibrium distribution using ancillary control. For a variety of control mechanisms, we find that the minimum amount of energy dissipation necessary can be cast as an information-theoretic measure of distinguishability between the target nonequilibrium state and the underlying equilibrium distribution. This work offers quantitative insight into the intuitive idea that more energy is needed to maintain a system farther from equilibrium. Full article
(This article belongs to the Special Issue Thermodynamics and Statistical Mechanics of Small Systems)
Show Figures

Figure 1

Figure 1
<p>Illustration of three types of control: (<b>Top</b>) Graph representation of the system’s configuration space without control. Mesoscopic configurations are represented as vertices (or nodes) with edges signifying allowed transitions; (<b>Bottom</b>) Control is implemented by adding additional edges (red dashed) or nodes (red dots) in order to drive the system into a nonequilibrium distribution. From Left to Right: Edge control, Node control, and Auxiliary control.</p>
Full article ">Figure 2
<p>Illustration of chemical control: Two species <math display="inline"> <semantics> <msub> <mi>X</mi> <mn>1</mn> </msub> </semantics> </math> and <math display="inline"> <semantics> <msub> <mi>X</mi> <mn>2</mn> </msub> </semantics> </math> interconvert through a single chemical reaction <math display="inline"> <semantics> <mrow> <msub> <mi>X</mi> <mn>1</mn> </msub> <mo>⇌</mo> <msub> <mi>X</mi> <mn>2</mn> </msub> </mrow> </semantics> </math> that conserves particle number, depicted as solid black lines. As a result, the dynamical evolution of this chemical reaction network is restricted to a single diagonal subspace of the network of states. Control can be implemented by chemostating one of the species, say <math display="inline"> <semantics> <msub> <mi>X</mi> <mn>1</mn> </msub> </semantics> </math>, by allowing <math display="inline"> <semantics> <msub> <mi>X</mi> <mn>1</mn> </msub> </semantics> </math> molecules to be added and subtracted from the reaction volume through the reaction <math display="inline"> <semantics> <mrow> <msub> <mi>X</mi> <mn>1</mn> </msub> <mo>⇌</mo> <mi>ϕ</mi> </mrow> </semantics> </math>, depicted as red dashed lines. As this reaction breaks the particle number conservation law, it extends the possible configurations the system can dynamically explore.</p>
Full article ">
545 KiB  
Article
Lyapunov Spectra of Coulombic and Gravitational Periodic Systems
by Pankaj Kumar and Bruce N. Miller
Entropy 2017, 19(5), 238; https://doi.org/10.3390/e19050238 - 20 May 2017
Cited by 3 | Viewed by 5167
Abstract
An open question in nonlinear dynamics is the relation between the Kolmogorov entropy and the largest Lyapunov exponent of a given orbit. Both have been shown to have diagnostic capability for phase transitions in thermodynamic systems. For systems with long-range interactions, the choice [...] Read more.
An open question in nonlinear dynamics is the relation between the Kolmogorov entropy and the largest Lyapunov exponent of a given orbit. Both have been shown to have diagnostic capability for phase transitions in thermodynamic systems. For systems with long-range interactions, the choice of boundary plays a critical role and appropriate boundary conditions must be invoked. In this work, we compute Lyapunov spectra for Coulombic and gravitational versions of the one-dimensional systems of parallel sheets with periodic boundary conditions. Exact expressions for time evolution of the tangent-space vectors are derived and are utilized toward computing Lypaunov characteristic exponents using an event-driven algorithm. The results indicate that the energy dependence of the largest Lyapunov exponent emulates that of Kolmogorov entropy for each system for a given system size. Our approach forms an effective and approximation-free instrument for studying the dynamical properties exhibited by the Coulombic and gravitational systems and finds applications in investigating indications of thermodynamic transitions in small as well as large versions of the spatially periodic systems. When a phase transition exists, we find that the largest Lyapunov exponent serves as a precursor of the transition that becomes more pronounced as the system size increases. Full article
(This article belongs to the Special Issue Thermodynamics and Statistical Mechanics of Small Systems)
Show Figures

Figure 1

Figure 1
<p>Full spectra of Lyapunov characteristic exponents (LCEs) plotted against per-particle energy for (<b>a</b>) Coulombic system, and (<b>b</b>) gravitational system, with <math display="inline"> <semantics> <mrow> <mi>N</mi> <mo>=</mo> <mn>11</mn> </mrow> </semantics> </math>. The topmost curve shows <math display="inline"> <semantics> <msub> <mi>λ</mi> <mn>1</mn> </msub> </semantics> </math>, the second to top curve shows <math display="inline"> <semantics> <msub> <mi>λ</mi> <mn>2</mn> </msub> </semantics> </math>, and so on all the way to the curve on the very bottom representing <math display="inline"> <semantics> <msub> <mi>λ</mi> <mn>22</mn> </msub> </semantics> </math> in (<b>a</b>,<b>b</b>). The central solid (blue) line in each plot shows the sum of LCEs. <math display="inline"> <semantics> <msub> <mi mathvariant="script">H</mi> <mi>c</mi> </msub> </semantics> </math> and <math display="inline"> <semantics> <msub> <mi mathvariant="script">H</mi> <mi>g</mi> </msub> </semantics> </math> are expressed in units of <math display="inline"> <semantics> <mrow> <mfrac> <mrow> <mn>2</mn> <mi>L</mi> </mrow> <mi>N</mi> </mfrac> <mrow> <mo>|</mo> <mi>κ</mi> <mo>|</mo> </mrow> </mrow> </semantics> </math>. <math display="inline"> <semantics> <msub> <mi>λ</mi> <mi>i</mi> </msub> </semantics> </math> are expressed in units of (<b>a</b>) <math display="inline"> <semantics> <mi>ω</mi> </semantics> </math>, and (<b>b</b>) <math display="inline"> <semantics> <mi mathvariant="sans-serif">Λ</mi> </semantics> </math>.</p>
Full article ">Figure 2
<p>Energy dependence of (<b>a</b>) the largest LCE, and (<b>b</b>) Kolmogorov-entropy density for Coulombic system with different degrees of freedom. <math display="inline"> <semantics> <msub> <mi>λ</mi> <mn>1</mn> </msub> </semantics> </math> and <math display="inline"> <semantics> <msub> <mi>λ</mi> <mi>S</mi> </msub> </semantics> </math> are expressed in units of <math display="inline"> <semantics> <mi>ω</mi> </semantics> </math>, whereas <math display="inline"> <semantics> <msub> <mi mathvariant="script">H</mi> <mi>c</mi> </msub> </semantics> </math> is expressed in units of <math display="inline"> <semantics> <mrow> <mfrac> <mrow> <mn>2</mn> <mi>L</mi> </mrow> <mi>N</mi> </mfrac> <mrow> <mo>|</mo> <mi>κ</mi> <mo>|</mo> </mrow> </mrow> </semantics> </math>.</p>
Full article ">Figure 3
<p>Energy dependence of (<b>a</b>) the largest LCE, and (<b>b</b>) Kolmogorov-entropy density for gravitational system with different degrees of freedom. <math display="inline"> <semantics> <msub> <mi>λ</mi> <mn>1</mn> </msub> </semantics> </math> and <math display="inline"> <semantics> <msub> <mi>λ</mi> <mi>S</mi> </msub> </semantics> </math> are expressed in units of <math display="inline"> <semantics> <mi mathvariant="sans-serif">Λ</mi> </semantics> </math>, whereas <math display="inline"> <semantics> <msub> <mi mathvariant="script">H</mi> <mi>g</mi> </msub> </semantics> </math> is expressed in units of <math display="inline"> <semantics> <mrow> <mfrac> <mrow> <mn>2</mn> <mi>L</mi> </mrow> <mi>N</mi> </mfrac> <mrow> <mo>|</mo> <mi>κ</mi> <mo>|</mo> </mrow> </mrow> </semantics> </math>.</p>
Full article ">Figure 4
<p>Energy dependence of the normalized values of the largest LCE and Kolmogorov-entropy density for (<b>a</b>) Coulombic system, and (<b>b</b>) gravitational system for <math display="inline"> <semantics> <mrow> <mi>N</mi> <mo>=</mo> <mn>11</mn> </mrow> </semantics> </math>. <math display="inline"> <semantics> <msub> <mi mathvariant="script">H</mi> <mi>c</mi> </msub> </semantics> </math> and <math display="inline"> <semantics> <msub> <mi mathvariant="script">H</mi> <mi>g</mi> </msub> </semantics> </math> are expressed in units of <math display="inline"> <semantics> <mrow> <mfrac> <mrow> <mn>2</mn> <mi>L</mi> </mrow> <mi>N</mi> </mfrac> <mrow> <mo>|</mo> <mi>κ</mi> <mo>|</mo> </mrow> </mrow> </semantics> </math>, whereas <math display="inline"> <semantics> <mover accent="true"> <mi>λ</mi> <mo stretchy="false">^</mo> </mover> </semantics> </math> is dimensionless.</p>
Full article ">
301 KiB  
Article
Stochastic Stirling Engine Operating in Contact with Active Baths
by Ruben Zakine, Alexandre Solon, Todd Gingrich and Frédéric Van Wijland
Entropy 2017, 19(5), 193; https://doi.org/10.3390/e19050193 - 27 Apr 2017
Cited by 61 | Viewed by 10158
Abstract
A Stirling engine made of a colloidal particle in contact with a nonequilibrium bath is considered and analyzed with the tools of stochastic energetics. We model the bath by non Gaussian persistent noise acting on the colloidal particle. Depending on the chosen definition [...] Read more.
A Stirling engine made of a colloidal particle in contact with a nonequilibrium bath is considered and analyzed with the tools of stochastic energetics. We model the bath by non Gaussian persistent noise acting on the colloidal particle. Depending on the chosen definition of an isothermal transformation in this nonequilibrium setting, we find that either the energetics of the engine parallels that of its equilibrium counterpart or, in the simplest case, that it ends up being less efficient. Persistence, more than non-Gaussian effects, are responsible for this result. Full article
(This article belongs to the Special Issue Thermodynamics and Statistical Mechanics of Small Systems)
Show Figures

Figure 1

Figure 1
<p>Schematic diagram of the Stirling cycle in stiffness-position space. Unlike its thermodynamic counterpart, the cycle is run counter-clockwise but is nevertheless an engine cycle.</p>
Full article ">Figure 2
<p>Log of the probability of the colloid’s position as a function of position (for a unit <span class="html-italic">a</span>), in equilibrium with Gaussian statistics (red at <math display="inline"> <semantics> <mrow> <mi>T</mi> <mo>/</mo> <mi>k</mi> <mo>=</mo> <mn>2</mn> </mrow> </semantics> </math>, green at <math display="inline"> <semantics> <mrow> <mi>T</mi> <mo>/</mo> <mi>k</mi> <mo>=</mo> <mn>4</mn> </mrow> </semantics> </math>) or out of equilibrium as given by <math display="inline"> <semantics> <msub> <mi>P</mi> <mi>ss</mi> </msub> </semantics> </math> (blue at <math display="inline"> <semantics> <mrow> <mi>T</mi> <mo>/</mo> <mi>k</mi> <mo>=</mo> <mn>2</mn> </mrow> </semantics> </math>, orange at <math display="inline"> <semantics> <mrow> <mi>T</mi> <mo>/</mo> <mi>k</mi> <mo>=</mo> <mn>4</mn> </mrow> </semantics> </math>).</p>
Full article ">

Review

Jump to: Editorial, Research

16 pages, 899 KiB  
Review
Fluctuations, Finite-Size Effects and the Thermodynamic Limit in Computer Simulations: Revisiting the Spatial Block Analysis Method
by Maziar Heidari, Kurt Kremer, Raffaello Potestio and Robinson Cortes-Huerto
Entropy 2018, 20(4), 222; https://doi.org/10.3390/e20040222 - 24 Mar 2018
Cited by 27 | Viewed by 6155
Abstract
The spatial block analysis (SBA) method has been introduced to efficiently extrapolate thermodynamic quantities from finite-size computer simulations of a large variety of physical systems. In the particular case of simple liquids and liquid mixtures, by subdividing the simulation box into blocks of [...] Read more.
The spatial block analysis (SBA) method has been introduced to efficiently extrapolate thermodynamic quantities from finite-size computer simulations of a large variety of physical systems. In the particular case of simple liquids and liquid mixtures, by subdividing the simulation box into blocks of increasing size and calculating volume-dependent fluctuations of the number of particles, it is possible to extrapolate the bulk isothermal compressibility and Kirkwood–Buff integrals in the thermodynamic limit. Only by explicitly including finite-size effects, ubiquitous in computer simulations, into the SBA method, the extrapolation to the thermodynamic limit can be achieved. In this review, we discuss two of these finite-size effects in the context of the SBA method due to (i) the statistical ensemble and (ii) the finite integration domains used in computer simulations. To illustrate the method, we consider prototypical liquids and liquid mixtures described by truncated and shifted Lennard–Jones (TSLJ) potentials. Furthermore, we show some of the most recent developments of the SBA method, in particular its use to calculate chemical potentials of liquids in a wide range of density/concentration conditions. Full article
(This article belongs to the Special Issue Thermodynamics and Statistical Mechanics of Small Systems)
Show Figures

Figure 1

Figure 1
<p>Snapshot of the simulation box for a system of particles interacting via a TSLJ potential at density <math display="inline"> <semantics> <mrow> <mi>ρ</mi> <msup> <mi>σ</mi> <mn>3</mn> </msup> <mo>=</mo> <mn>0.1</mn> </mrow> </semantics> </math> and temperature <math display="inline"> <semantics> <mrow> <msub> <mi>k</mi> <mi mathvariant="normal">B</mi> </msub> <mi>T</mi> <mo>=</mo> <mn>1.2</mn> <mi>ϵ</mi> </mrow> </semantics> </math>. In this particular example, a box of linear size <math display="inline"> <semantics> <msub> <mi>L</mi> <mn>0</mn> </msub> </semantics> </math> has been subdivided into blocks of linear dimension <math display="inline"> <semantics> <mrow> <mi>L</mi> <mo>=</mo> <msub> <mi>L</mi> <mn>0</mn> </msub> <mo>/</mo> <mn>5</mn> </mrow> </semantics> </math> as indicated by the different color shades. The figure has been rendered with the Visual Molecular Dynamics (VMD) program [<a href="#B31-entropy-20-00222" class="html-bibr">31</a>].</p>
Full article ">Figure 2
<p>Fluctuations of the number of particles <math display="inline"> <semantics> <mrow> <msub> <mi>χ</mi> <mi>T</mi> </msub> <mrow> <mo stretchy="false">(</mo> <mi>L</mi> <mo>,</mo> <msub> <mi>L</mi> <mn>0</mn> </msub> <mo stretchy="false">)</mo> </mrow> </mrow> </semantics> </math> as a function of <math display="inline"> <semantics> <mrow> <mi>σ</mi> <mo>/</mo> <mi>L</mi> </mrow> </semantics> </math> for systems described by a TSLJ potential with <math display="inline"> <semantics> <mrow> <msub> <mi>r</mi> <mi>c</mi> </msub> <mo>/</mo> <mi>σ</mi> <mo>=</mo> <msup> <mn>2</mn> <mrow> <mn>1</mn> <mo>/</mo> <mn>6</mn> </mrow> </msup> </mrow> </semantics> </math>. Data corresponding to system sizes <math display="inline"> <semantics> <mrow> <msub> <mi>N</mi> <mn>0</mn> </msub> <mo>=</mo> <msup> <mn>10</mn> <mn>4</mn> </msup> <mo>,</mo> <msup> <mn>10</mn> <mn>5</mn> </msup> </mrow> </semantics> </math> and <math display="inline"> <semantics> <msup> <mn>10</mn> <mn>6</mn> </msup> </semantics> </math> are presented using red squares, blue triangles and green circles, respectively. The vertical lines indicate the limit <math display="inline"> <semantics> <mrow> <mi>σ</mi> <mo>/</mo> <msub> <mi>L</mi> <mn>0</mn> </msub> </mrow> </semantics> </math> at which fluctuations become zero. The black horizontal dashed line indicates the value <math display="inline"> <semantics> <mrow> <msubsup> <mi>χ</mi> <mrow> <mi>T</mi> </mrow> <mo>∞</mo> </msubsup> <mo>=</mo> <mi>ρ</mi> <msub> <mi>k</mi> <mi mathvariant="normal">B</mi> </msub> <mi>T</mi> <msub> <mi>κ</mi> <mi>T</mi> </msub> <mo>=</mo> <mn>0.0295</mn> </mrow> </semantics> </math> with <math display="inline"> <semantics> <msub> <mi>κ</mi> <mi>T</mi> </msub> </semantics> </math> the bulk compressibility obtained with the method described in [<a href="#B6-entropy-20-00222" class="html-bibr">6</a>].</p>
Full article ">Figure 3
<p>Fluctuations of the number of particles <math display="inline"> <semantics> <mrow> <msub> <mi>χ</mi> <mi>T</mi> </msub> <mrow> <mo stretchy="false">(</mo> <mi>L</mi> <mo>,</mo> <msub> <mi>L</mi> <mn>0</mn> </msub> <mo stretchy="false">)</mo> </mrow> </mrow> </semantics> </math> as a function of the ratio <math display="inline"> <semantics> <mrow> <mi>L</mi> <mo>/</mo> <msub> <mi>L</mi> <mn>0</mn> </msub> </mrow> </semantics> </math> for systems described by a TSLJ potential with <math display="inline"> <semantics> <mrow> <msub> <mi>r</mi> <mi>c</mi> </msub> <mo>/</mo> <mi>σ</mi> <mo>=</mo> <msup> <mn>2</mn> <mrow> <mn>1</mn> <mo>/</mo> <mn>6</mn> </mrow> </msup> </mrow> </semantics> </math>. Results corresponding to systems of <math display="inline"> <semantics> <mrow> <msub> <mi>N</mi> <mn>0</mn> </msub> <mo>=</mo> <msup> <mn>10</mn> <mn>5</mn> </msup> </mrow> </semantics> </math> particles with densities <math display="inline"> <semantics> <mrow> <mi>ρ</mi> <msup> <mi>σ</mi> <mn>3</mn> </msup> <mo>=</mo> <mn>0.1</mn> <mo>,</mo> <mspace width="0.166667em"/> <mn>0.2</mn> </mrow> </semantics> </math> and 0.3 are presented using red squares, blue triangles and green circles, respectively. The theoretical prediction presented in the text is plotted using the corresponding value for <math display="inline"> <semantics> <msubsup> <mi>χ</mi> <mrow> <mi>T</mi> </mrow> <mo>∞</mo> </msubsup> </semantics> </math>, obtained as described in [<a href="#B6-entropy-20-00222" class="html-bibr">6</a>], and solid-line curves with the same color code.</p>
Full article ">Figure 4
<p>Fluctuations of the number of particles <math display="inline"> <semantics> <mrow> <msub> <mi>χ</mi> <mi>T</mi> </msub> <mrow> <mo stretchy="false">(</mo> <mi>L</mi> <mo>,</mo> <msub> <mi>L</mi> <mn>0</mn> </msub> <mo stretchy="false">)</mo> </mrow> </mrow> </semantics> </math> as a function of the ratio <math display="inline"> <semantics> <mrow> <mi>L</mi> <mo>/</mo> <msub> <mi>L</mi> <mn>0</mn> </msub> </mrow> </semantics> </math> for systems described by a TSLJ potential with <math display="inline"> <semantics> <mrow> <msub> <mi>r</mi> <mi>c</mi> </msub> <mo>/</mo> <mi>σ</mi> <mo>=</mo> <msup> <mn>2</mn> <mrow> <mn>1</mn> <mo>/</mo> <mn>6</mn> </mrow> </msup> </mrow> </semantics> </math>. Results corresponding to sizes <math display="inline"> <semantics> <mrow> <msub> <mi>N</mi> <mn>0</mn> </msub> <mo>=</mo> <msup> <mn>10</mn> <mn>4</mn> </msup> <mo>,</mo> <mspace width="0.166667em"/> <msup> <mn>10</mn> <mn>5</mn> </msup> </mrow> </semantics> </math> and <math display="inline"> <semantics> <msup> <mn>10</mn> <mn>6</mn> </msup> </semantics> </math>, with density <math display="inline"> <semantics> <mrow> <mi>ρ</mi> <msup> <mi>σ</mi> <mn>3</mn> </msup> <mo>=</mo> <mn>0.864</mn> </mrow> </semantics> </math>, using red squares, blue triangles and green circles, respectively. The theoretical prediction presented in the text is plotted as the black dashed curve using <math display="inline"> <semantics> <mrow> <msubsup> <mi>χ</mi> <mrow> <mi>T</mi> </mrow> <mo>∞</mo> </msubsup> <mo>=</mo> <mn>0.0295</mn> </mrow> </semantics> </math>.</p>
Full article ">Figure 5
<p>Scaled fluctuations of the number of particles <math display="inline"> <semantics> <mrow> <mi>λ</mi> <msub> <mi>χ</mi> <mi>T</mi> </msub> <mrow> <mo stretchy="false">(</mo> <mi>L</mi> <mo>,</mo> <msub> <mi>L</mi> <mn>0</mn> </msub> <mo stretchy="false">)</mo> </mrow> </mrow> </semantics> </math>, minus <math display="inline"> <semantics> <mrow> <mi>c</mi> <mo>/</mo> <msub> <mi>L</mi> <mn>0</mn> </msub> </mrow> </semantics> </math>, versus the ratio <math display="inline"> <semantics> <mrow> <mi>λ</mi> <mo>=</mo> <mi>L</mi> <mo>/</mo> <msub> <mi>L</mi> <mn>0</mn> </msub> </mrow> </semantics> </math> for systems described by a TSLJ potential with <math display="inline"> <semantics> <mrow> <msub> <mi>r</mi> <mi>c</mi> </msub> <mo>/</mo> <mi>σ</mi> <mo>=</mo> <msup> <mn>2</mn> <mrow> <mn>1</mn> <mo>/</mo> <mn>6</mn> </mrow> </msup> </mrow> </semantics> </math>. Results corresponding to sizes <math display="inline"> <semantics> <mrow> <msub> <mi>N</mi> <mn>0</mn> </msub> <mo>=</mo> <msup> <mn>10</mn> <mn>4</mn> </msup> <mo>,</mo> <mspace width="0.166667em"/> <msup> <mn>10</mn> <mn>5</mn> </msup> </mrow> </semantics> </math> and <math display="inline"> <semantics> <msup> <mn>10</mn> <mn>6</mn> </msup> </semantics> </math>, with density <math display="inline"> <semantics> <mrow> <mi>ρ</mi> <msup> <mi>σ</mi> <mn>3</mn> </msup> <mo>=</mo> <mn>0.864</mn> </mrow> </semantics> </math>, using red squares, blue triangles and green circles, respectively. The theoretical prediction Equation (<a href="#FD7-entropy-20-00222" class="html-disp-formula">7</a>) presented in the text is plotted as the black solid curve using <math display="inline"> <semantics> <mrow> <msubsup> <mi>χ</mi> <mrow> <mi>T</mi> </mrow> <mo>∞</mo> </msubsup> <mo>=</mo> <mn>0.0295</mn> </mrow> </semantics> </math> and <math display="inline"> <semantics> <mrow> <mi>c</mi> <mo>=</mo> <mn>0.415</mn> <mi>σ</mi> </mrow> </semantics> </math>.</p>
Full article ">Figure 6
<p>Ratio <math display="inline"> <semantics> <mrow> <msubsup> <mi>χ</mi> <mrow> <mi>T</mi> </mrow> <mo>∞</mo> </msubsup> <mo>=</mo> <msub> <mi>κ</mi> <mi>T</mi> </msub> <mo>/</mo> <msubsup> <mi>κ</mi> <mrow> <mi>T</mi> </mrow> <mrow> <mi>I</mi> <mi>G</mi> </mrow> </msubsup> </mrow> </semantics> </math> at <math display="inline"> <semantics> <mrow> <msub> <mi>k</mi> <mi mathvariant="normal">B</mi> </msub> <mi>T</mi> <mo>=</mo> <mn>1.2</mn> <mi>ϵ</mi> </mrow> </semantics> </math> as a function of the density for systems described by a TSLJ potential with <math display="inline"> <semantics> <mrow> <msub> <mi>r</mi> <mi>c</mi> </msub> <mo>/</mo> <mi>σ</mi> <mo>=</mo> <msup> <mn>2</mn> <mrow> <mn>1</mn> <mo>/</mo> <mn>6</mn> </mrow> </msup> </mrow> </semantics> </math>, with <math display="inline"> <semantics> <mrow> <msubsup> <mi>κ</mi> <mrow> <mi>T</mi> </mrow> <mrow> <mi>I</mi> <mi>G</mi> </mrow> </msubsup> <mo>=</mo> <msup> <mrow> <mo stretchy="false">(</mo> <mi>ρ</mi> <msub> <mi>k</mi> <mi mathvariant="normal">B</mi> </msub> <mi>T</mi> <mo stretchy="false">)</mo> </mrow> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </msup> </mrow> </semantics> </math> the isothermal compressibility of the ideal gas. The red curve is a guide to the eye.</p>
Full article ">Figure 7
<p>Excess chemical potential <math display="inline"> <semantics> <mrow> <msup> <mi>μ</mi> <mrow> <mi>e</mi> <mi>x</mi> </mrow> </msup> <mo>/</mo> <mi>ϵ</mi> </mrow> </semantics> </math> at <math display="inline"> <semantics> <mrow> <msub> <mi>k</mi> <mi mathvariant="normal">B</mi> </msub> <mi>T</mi> <mo>=</mo> <mn>1.2</mn> <mi>ϵ</mi> </mrow> </semantics> </math> as a function of the density for systems described by a TSLJ potential with <math display="inline"> <semantics> <mrow> <msub> <mi>r</mi> <mi>c</mi> </msub> <mo>/</mo> <mi>σ</mi> <mo>=</mo> <msup> <mn>2</mn> <mrow> <mn>1</mn> <mo>/</mo> <mn>6</mn> </mrow> </msup> </mrow> </semantics> </math>. Red squares indicate the data obtained with the spatially-resolved thermodynamic integration (SPARTIAN) method [<a href="#B36-entropy-20-00222" class="html-bibr">36</a>], and the blue triangles are the data points obtained with the method outlined in the text.</p>
Full article ">Figure 8
<p>Reduced fluctuations as a function of <math display="inline"> <semantics> <mi>λ</mi> </semantics> </math> for systems described by a TSLJ potential with <math display="inline"> <semantics> <mrow> <msub> <mi>r</mi> <mi>c</mi> </msub> <mo>/</mo> <mi>σ</mi> <mo>=</mo> <mn>2.5</mn> </mrow> </semantics> </math> with density <math display="inline"> <semantics> <mrow> <mi>ρ</mi> <msup> <mi>σ</mi> <mn>3</mn> </msup> <mo>=</mo> <mn>0.3</mn> </mrow> </semantics> </math> at temperatures <math display="inline"> <semantics> <mrow> <msub> <mi>k</mi> <mi mathvariant="normal">B</mi> </msub> <mi>T</mi> <mo>=</mo> <mn>2.00</mn> <mi>ϵ</mi> </mrow> </semantics> </math> and <math display="inline"> <semantics> <mrow> <mn>1.15</mn> <mi>ϵ</mi> </mrow> </semantics> </math>. For the latter case, it is apparent that the contribution proportional to <math display="inline"> <semantics> <msup> <mi>λ</mi> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </msup> </semantics> </math> is not negligible. The inset shows the full range <math display="inline"> <semantics> <mrow> <mn>0</mn> <mo>&lt;</mo> <mi>λ</mi> <mo>&lt;</mo> <mn>1</mn> </mrow> </semantics> </math>. The black curves are the result of fitting the data to Equation (<a href="#FD22-entropy-20-00222" class="html-disp-formula">22</a>).</p>
Full article ">Figure 9
<p>Bulk isothermal compressibility <math display="inline"> <semantics> <msub> <mi>κ</mi> <mi>T</mi> </msub> </semantics> </math> as a function of the density <math display="inline"> <semantics> <mi>ρ</mi> </semantics> </math> at <math display="inline"> <semantics> <mrow> <msub> <mi>k</mi> <mi mathvariant="normal">B</mi> </msub> <mi>T</mi> <mo>=</mo> <mn>1.15</mn> <mi>ϵ</mi> </mrow> </semantics> </math> (red circles) and <math display="inline"> <semantics> <mrow> <msub> <mi>k</mi> <mi mathvariant="normal">B</mi> </msub> <mi>T</mi> <mo>=</mo> <mn>2.00</mn> <mi>ϵ</mi> </mrow> </semantics> </math> (green squares) for systems described by a TSLJ potential with <math display="inline"> <semantics> <mrow> <msub> <mi>r</mi> <mi>c</mi> </msub> <mo>/</mo> <mi>σ</mi> <mo>=</mo> <mn>2.5</mn> </mrow> </semantics> </math>. The vertical black line indicates the location of the critical density <math display="inline"> <semantics> <mrow> <mi>ρ</mi> <msup> <mi>σ</mi> <mn>3</mn> </msup> <mo>=</mo> <mn>0.319</mn> </mrow> </semantics> </math> [<a href="#B38-entropy-20-00222" class="html-bibr">38</a>].</p>
Full article ">Figure 10
<p>Scaled finite-size Kirkwood–Buff integrals <math display="inline"> <semantics> <mrow> <mi>λ</mi> <msub> <mi>G</mi> <mrow> <mi>i</mi> <mi>j</mi> </mrow> </msub> <mrow> <mo stretchy="false">(</mo> <mi>λ</mi> <mo stretchy="false">)</mo> </mrow> </mrow> </semantics> </math> as a function of <math display="inline"> <semantics> <mi>λ</mi> </semantics> </math> for different mole fractions: (<b>a</b>) <math display="inline"> <semantics> <mrow> <msub> <mi>x</mi> <mi>A</mi> </msub> <mo>=</mo> <mn>0.20</mn> </mrow> </semantics> </math>; (<b>b</b>) <math display="inline"> <semantics> <mrow> <msub> <mi>x</mi> <mi>A</mi> </msub> <mo>=</mo> <mn>0.30</mn> </mrow> </semantics> </math>; (<b>c</b>) <math display="inline"> <semantics> <mrow> <msub> <mi>x</mi> <mi>A</mi> </msub> <mo>=</mo> <mn>0.50</mn> </mrow> </semantics> </math> and (<b>d</b>) <math display="inline"> <semantics> <mrow> <msub> <mi>x</mi> <mi>A</mi> </msub> <mo>=</mo> <mn>0.80</mn> </mrow> </semantics> </math>, for mixtures described by a TSLJ potential with <math display="inline"> <semantics> <mrow> <msub> <mi>r</mi> <mi>c</mi> </msub> <mo>/</mo> <mi>σ</mi> <mo>=</mo> <msup> <mn>2</mn> <mrow> <mn>1</mn> <mo>/</mo> <mn>6</mn> </mrow> </msup> </mrow> </semantics> </math>. For clarity, only the cases <math display="inline"> <semantics> <msub> <mi>G</mi> <mrow> <mi>A</mi> <mi>A</mi> </mrow> </msub> </semantics> </math> (red squares) and <math display="inline"> <semantics> <msub> <mi>G</mi> <mrow> <mi>A</mi> <mi>B</mi> </mrow> </msub> </semantics> </math> (green circles) are plotted. In the asymptotic case <math display="inline"> <semantics> <mrow> <mi>λ</mi> <mo>→</mo> <mn>1</mn> </mrow> </semantics> </math>, <math display="inline"> <semantics> <mrow> <msub> <mi>G</mi> <mrow> <mi>A</mi> <mi>B</mi> </mrow> </msub> <mo>→</mo> <mn>0</mn> </mrow> </semantics> </math> and <math display="inline"> <semantics> <mrow> <msub> <mi>G</mi> <mrow> <mi>A</mi> <mi>A</mi> </mrow> </msub> <mo>→</mo> <mn>1</mn> <mo>/</mo> <msub> <mi>ρ</mi> <mi>A</mi> </msub> </mrow> </semantics> </math>, as indicated by the horizontal green and red lines, respectively. The black curves correspond to Equation (<a href="#FD26-entropy-20-00222" class="html-disp-formula">26</a>) with <math display="inline"> <semantics> <msubsup> <mi>G</mi> <mrow> <mi>i</mi> <mi>j</mi> </mrow> <mo>∞</mo> </msubsup> </semantics> </math> and <math display="inline"> <semantics> <msub> <mi>α</mi> <mrow> <mi>i</mi> <mi>j</mi> </mrow> </msub> </semantics> </math> obtained from a simple regression analysis in the interval <math display="inline"> <semantics> <mrow> <mi>λ</mi> <mo>&lt;</mo> <mn>0.3</mn> </mrow> </semantics> </math>.</p>
Full article ">Figure 11
<p>Isothermal compressibility at <math display="inline"> <semantics> <mrow> <msub> <mi>k</mi> <mi mathvariant="normal">B</mi> </msub> <mi>T</mi> <mo>=</mo> <mn>1.20</mn> <mi>ϵ</mi> </mrow> </semantics> </math> and <math display="inline"> <semantics> <mrow> <mi>P</mi> <msup> <mi>σ</mi> <mn>3</mn> </msup> <mo>/</mo> <mi>ϵ</mi> <mo>=</mo> <mn>9.8</mn> </mrow> </semantics> </math> as a function of the mole fraction of type-<span class="html-italic">A</span> particles <math display="inline"> <semantics> <msub> <mi>x</mi> <mi>A</mi> </msub> </semantics> </math> for mixtures described by a TSLJ potential with <math display="inline"> <semantics> <mrow> <msub> <mi>r</mi> <mi>c</mi> </msub> <mo>/</mo> <mi>σ</mi> <mo>=</mo> <msup> <mn>2</mn> <mrow> <mn>1</mn> <mo>/</mo> <mn>6</mn> </mrow> </msup> </mrow> </semantics> </math>. The horizontal black lines indicate the compressibility for a pure system of type-<span class="html-italic">A</span> particles <math display="inline"> <semantics> <mrow> <msubsup> <mi>κ</mi> <mrow> <mi>T</mi> </mrow> <mi>A</mi> </msubsup> <mi>ϵ</mi> <mo>/</mo> <msup> <mi>σ</mi> <mn>3</mn> </msup> <mo>=</mo> <mn>0.012</mn> <mrow> <mo stretchy="false">(</mo> <mn>1</mn> <mo stretchy="false">)</mo> </mrow> </mrow> </semantics> </math> and for a pure system of type-<span class="html-italic">B</span> particles <math display="inline"> <semantics> <mrow> <msubsup> <mi>κ</mi> <mrow> <mi>T</mi> </mrow> <mi>B</mi> </msubsup> <mi>ϵ</mi> <mo>/</mo> <msup> <mi>σ</mi> <mn>3</mn> </msup> <mo>=</mo> <mn>0.0281</mn> <mrow> <mo stretchy="false">(</mo> <mn>8</mn> <mo stretchy="false">)</mo> </mrow> </mrow> </semantics> </math>. The red line is a guide to the eye. The ideal case corresponds to <math display="inline"> <semantics> <mrow> <msub> <mi>κ</mi> <mi>T</mi> </msub> <mo>=</mo> <mrow> <mo stretchy="false">(</mo> <mn>1</mn> <mo>−</mo> <msub> <mi>x</mi> <mi>A</mi> </msub> <mo stretchy="false">)</mo> </mrow> <msubsup> <mi>κ</mi> <mrow> <mi>T</mi> </mrow> <mi>B</mi> </msubsup> <mo>+</mo> <msub> <mi>x</mi> <mi>A</mi> </msub> <msubsup> <mi>κ</mi> <mrow> <mi>T</mi> </mrow> <mi>A</mi> </msubsup> </mrow> </semantics> </math>.</p>
Full article ">Figure 12
<p>Excess chemical potential of type-<span class="html-italic">A</span> particles as a function of the mole fraction <math display="inline"> <semantics> <msub> <mi>x</mi> <mi>A</mi> </msub> </semantics> </math> for mixtures described by a TSLJ potential with <math display="inline"> <semantics> <mrow> <msub> <mi>r</mi> <mi>c</mi> </msub> <mo>/</mo> <mi>σ</mi> <mo>=</mo> <msup> <mn>2</mn> <mrow> <mn>1</mn> <mo>/</mo> <mn>6</mn> </mrow> </msup> </mrow> </semantics> </math> at <math display="inline"> <semantics> <mrow> <msub> <mi>k</mi> <mi mathvariant="normal">B</mi> </msub> <mi>T</mi> <mo>=</mo> <mn>1.2</mn> <mi>ϵ</mi> </mrow> </semantics> </math> and <math display="inline"> <semantics> <mrow> <mi>P</mi> <msup> <mi>σ</mi> <mn>3</mn> </msup> <mo>/</mo> <mi>ϵ</mi> <mo>=</mo> <mn>9.8</mn> </mrow> </semantics> </math>. Data points obtained with the method in [<a href="#B36-entropy-20-00222" class="html-bibr">36</a>], in particular for <math display="inline"> <semantics> <mrow> <msub> <mi>x</mi> <mi>A</mi> </msub> <mo>=</mo> <mn>0.3</mn> </mrow> </semantics> </math>, are used as a reference for the data points obtained with Equations (<a href="#FD30-entropy-20-00222" class="html-disp-formula">30</a>) and (<a href="#FD31-entropy-20-00222" class="html-disp-formula">31</a>).</p>
Full article ">
Back to TopTop