Superdiffusion in two-dimensional Yukawa liquids

Similar documents
arxiv:cond-mat/ v1 8 Nov 2005

Non-Gaussian statistics and superdiffusion in a driven-dissipative dusty plasma

arxiv: v2 [physics.plasm-ph] 10 Jan 2013

Experimental test of an expression for the decay of an autocorrelation function

Experimental demonstration that a strongly coupled plasma obeys the fluctuation theorem for entropy production

First-principle results for the radial pair distribution function in strongly coupled one-component plasmas

Nonlinear interaction of compressional waves in a 2D dusty. plasma crystal. Abstract

Molecular dynamics simulations of strongly coupled plasmas

arxiv: v1 [cond-mat.stat-mech] 26 Jan 2008

766 Liu Bin et al Vol. 12 melting transition of the plasma crystal. Experimental investigations showed the melting transition from a solid-state struc

Diagnostics for transport phenomena in strongly coupled dusty plasmas

Nonlinear longitudinal waves in a two-dimensional screened Coulomb crystal

The Kawasaki Identity and the Fluctuation Theorem

arxiv: v2 [physics.plasm-ph] 6 Oct 2012

arxiv:cond-mat/ v2 7 Dec 1999

Thermal Conductivity of Complex Plasmas Using Novel Evan-Gillan Approach

STRONGLY coupled plasma layers can be created in complex

Theory of fractional Lévy diffusion of cold atoms in optical lattices

4. The Green Kubo Relations

The Truth about diffusion (in liquids)

arxiv: v1 [cond-mat.soft] 12 Sep 2007

F ij (t) = Φ ij (t) (3) Φ ij (t) = , 4πϵ 0 r ij (t)

Phonons in a one-dimensional Yukawa chain: Dusty plasma experiment and model

arxiv:nlin/ v1 [nlin.cd] 8 May 2001

arxiv: v1 [physics.plasm-ph] 18 Apr 2011

arxiv: v4 [physics.plasm-ph] 9 Nov 2017

Normal modes of a small bilayer system of binary classical charged particles trapped in a parabolic confinement potential

Vortex glass scaling in Pb-doped Bi2223 single crystal

Molecular Dynamics Simulation Study of Transport Properties of Diatomic Gases

Self-diffusion in two-dimensional quasi-magnetized rotating dusty plasmas

Direct observation of aggregative nanoparticle growth: Kinetic modeling of the size distribution and growth rate

Supporting Information. Even Hard Sphere Colloidal Suspensions Display. Fickian yet Non-Gaussian Diffusion

Superfluidity and Superconductivity

2 Lyapunov Exponent exponent number. Lyapunov spectrum for colour conductivity of 8 WCA disks.

APMA 2811T. By Zhen Li. Today s topic: Lecture 2: Theoretical foundation and parameterization. Sep. 15, 2016

The dynamics of small particles whose size is roughly 1 µmt or. smaller, in a fluid at room temperature, is extremely erratic, and is

arxiv: v1 [physics.plasm-ph] 23 May 2018

MD Thermodynamics. Lecture 12 3/26/18. Harvard SEAS AP 275 Atomistic Modeling of Materials Boris Kozinsky

The phonon wake behind a charge moving relative to a two-dimensional plasma crystal

Thermal fluctuations, mechanical response, and hyperuniformity in jammed solids

Statistical Mechanics of Nonequilibrium Liquids

arxiv: v1 [nlin.ps] 9 May 2015

Lecture notes for /12.586, Modeling Environmental Complexity. D. H. Rothman, MIT September 24, Anomalous diffusion

Diffusing-wave-spectroscopy measurements of viscoelasticity of complex fluids

In-depth analysis of viscoelastic properties thanks to Microrheology: non-contact rheology

Potentials of Unbalanced Complex Kinetics Observed in Market Time Series

CHAPTER V. Brownian motion. V.1 Langevin dynamics

Excitation of Ion Cyclotron Instability in. Inhomogeneous Rotating Dusty Plasma

PHYSICAL REVIEW E 72,

Entropic Crystal-Crystal Transitions of Brownian Squares K. Zhao, R. Bruinsma, and T.G. Mason

Determination of interaction between a dust particle pair in complex plasmas

Evaluating Diffusion Coefficients in LAMMPS

MEASUREMENT OF THE ION DRAG FORCE IN A COMPLEX DC- PLASMA USING THE PK-4 EXPERIMENT

Anomalous Lévy diffusion: From the flight of an albatross to optical lattices. Eric Lutz Abteilung für Quantenphysik, Universität Ulm

Analysis of the simulation

FINITE COULOMB CRYSTAL FORMATION

Microscopic dynamics in two-dimensional strongly-coupled dusty plasmas

Superdiffusive and subdiffusive transport of energetic particles in astrophysical plasmas: numerical simulations and experimental evidence

Dusty plasma as a unique object of plasma physics

Anomalous Transport and Fluctuation Relations: From Theory to Biology

Dynamic properties of Lennard-Jones fluids and liquid metals

Correlation between local structure and dynamic heterogeneity in a metallic glass-forming liquid

Author's personal copy

Radiation pressure and gas drag forces on a melamine-formaldehyde microsphere in a dusty plasma

Pair diffusion in quasi-one- and quasi-two-dimensional binary colloid suspensions

Lagrangian Statistics. of 3D MHD Convection. J. Pratt, W.-C. Müller. Boussinesq Simulation. Lagrangian. simulation. March 1, 2011

Anomalous diffusion in biology: fractional Brownian motion, Lévy flights

Normal Mode Analysis of Chain Structures in Complex Plasma

Disordered Hyperuniformity: Liquid-like Behaviour in Structural Solids, A New Phase of Matter?

Contents. 1.1 Prerequisites and textbooks Physical phenomena and theoretical tools The path integrals... 9

Analysis of Pair Interparticle Interaction in Nonideal Dissipative Systems

ON ALGORITHMS FOR BROWNIAN DYNAMICS COMPUTER SIMULATIONS

Collective behavior, from particles to fields

From time series to superstatistics

Experimental Colloids I (and I)

Self-consistent particle tracking in a simulation of the entropy mode in a Z pinch

Hexatic and microemulsion phases in a 2D quantum Coulomb gas

Weak Ergodicity Breaking. Manchester 2016

Defense Technical Information Center Compilation Part Notice

How DLS Works: Interference of Light

Modeling and Simulating Gold Nanoparticle Interactions on a Liquid-Air Interface

Aging in laponite water suspensions. P. K. Bhattacharyya Institute for Soldier Nanotechnologies Massachusetts Institute of Technology

Active microrheology: Brownian motion, strong forces and viscoelastic media

Cover times of random searches

Natalia Tronko S.V.Nazarenko S. Galtier

On the Asymptotic Convergence. of the Transient and Steady State Fluctuation Theorems. Gary Ayton and Denis J. Evans. Research School Of Chemistry

Noisy Lévy walk analog of two-dimensional DNA walks for chromosomes of S. cerevisiae

arxiv:cond-mat/ v1 [cond-mat.stat-mech] 4 Jan 2007

Generalizations of a nonlinear fluid model for void formation in dusty plasmas

Polymer-specific effects of bulk relaxation and stringlike correlated motion in the dynamics of a supercooled polymer melt

Determination of Refractive Index Gradient and Diffusion Coefficient of Salt Solution from Laser Deflection Measurement (10 points)

Three-dimensional structure in a crystallized dusty plasma

Voids in Dusty Plasma of a Stratified DC Glow Discharge in Noble Gases

Measurement of the ion drag force in a collisionless plasma with strong ion-grain coupling

On the Dynamics and Disentanglement in Thin and Two-Dimensional Polymer Films

Equilibrium correlations and heat conduction in the Fermi-Pasta-Ulam chain.

Non-equilibrium phenomena and fluctuation relations

4 Results of the static and dynamic light scattering measurements

Vlasov simulations of electron holes driven by particle distributions from PIC reconnection simulations with a guide field

Ab Ini'o Molecular Dynamics (MD) Simula?ons

Transcription:

Superdiffusion in two-dimensional Yukawa liquids Bin Liu and J. Goree Department of Physics and Astronomy, The University of Iowa, Iowa City, Iowa 52242, USA Received 31 May 2006; published 18 January 2007 Superdiffusion of two-dimensional 2D liquids was studied using an equilibrium molecular dynamics simulation. At intermediate temperatures, the mean-squared displacement, probability distribution function PDF, and velocity autocorrelation function VACF all indicate superdiffusion; the VACF has a long-time tail; and the PDF indicates no Lévy flights. These effects are predicted to occur in 2D dusty plasmas and other 2D liquids that can be modeled with a long-range repulsive potential. DOI: 10.1103/PhysRevE.75.016405 PACS number s : 52.27.Gr, 66.10.Cb, 52.27.Lw, 82.70.Dd I. INTRODUCTION There are many two-dimensional 2D physical systems, some of them at a macroscopic scale, that can be in an ordered solid or disordered state liquid. These include a Wigner lattice of electrons on the surface of liquid helium 1, an array of vortices in the mixed state of type-ii superconductors 2,3, colloidal suspensions 4, ions confined magnetically in a Penning trap 5, and dusty plasmas levitated in a monolayer 6. There are also physical systems with a monolayer of particles or clusters that interact not only amongst themselves, but also with an underlying substrate, such as an atomically flat surface 7. Brown s original observation of Brownian motion was for a 2D system of small particles on a water surface 8. Here we investigate random motion in 2D equilibrium liquids. We are motivated by a long-standing controversy over diffusion in 2D liquids, with conflicting theoretical and simulation results. In general, motion has been predicted theoretically to be nondiffusive in 2D liquids, regardless of the pair potential 9,10 ; and early molecular dynamics MD simulations 11 with hard-disks came to a similar prediction. However, normal diffusion was detected in more recent simulations with softer potentials 12 15. We are also motivated by recent experimental results for diffusion in 2D dusty plasma. One experimental group 16 using a quasi-2d dusty plasma claimed to observe superdiffusion, as did a second group 17 using a monolayer dusty plasma cluster in a partly ordered state that allows hydrodynamical vortical flows. However, a third group 6 with a 2D monolayer dusty plasma claimed to observe no superdiffusive motion. Superdiffusion is random motion much like normal diffusion, but it differs from normal diffusion by having larger displacements as well as distinctive signatures in correlation functions and probability distribution functions, as explained later. The two experiments with a monolayer 6,17 have a pair potential that is Yukawa 18. Here we report results from an equilibrium MD simulation for particles interacting with a Yukawa potential and constrained to moving on a plane. We characterize the nondiffusive motion, and we find that it is superdiffusion that does not arise from Lévy flights; instead, it seems to arise from hydrodynamic modes, as indicated by a long-time tail in the velocity autocorrelation function VACF. We note that the Yukawa system in 3D, unlike in 2D, has only been reported to exhibit diffusive motion 19. Our results will be helpful in understanding recent experiments with 2D dusty plasmas 6,17. They will be helpful also for gaining an understanding of other experimental systems with long-range repulsive potentials, including pure-ion plasmas, electrons on the surface of liquid helium, and microspheres in a colloidal suspension. II. SIMULATIONS We performed equilibrium MD simulations. We used the Yukawa pair potential, r =Q 2 exp r/ D /4 0 r, where Q is particle charge and D is screening length. Equilibrium Yukawa systems can be classified completely by the values of and 19,20. Here, =Q 2 /4 0 ak B T and a/ D, where T is particle kinetic temperature, a n 1 is the Wigner-Seitz radius 21, and n is the areal number density. 1 Time scales of interest are characterized by pd = Q 2 /2 0 ma 3 1/2, where m is particle mass. Another parameter of interest is T/T m, where the melting temperature T m depends on and has a value of T m 145 for =0.56 22. In this paper, we investigate the intermediatetemperature regime 5 88, identified in 23 as having nondiffusive motion. We assume =0.56, comparable to the value in 2D dusty plasma experiments 6. Our simulation is similar to the monolayer dusty plasma experiments 6,17 in the use of a 2D monolayer with a Yukawa potential and similar values of parameters and, but it differs from these experiments in three ways. It describes a frictionless atomic system, it uses periodic boundary conditions to model an infinite system, and it models an equilibrium situation without instabilities driven by ion flow. In our simulation, we integrate the equation of motion of N particles. We used N=1024, and we verified that our results are independent of N over a range from 1024 to 16 384. We repeated our simulations for six different initial configurations of the particle positions, averaging these results to improve the statistics. A Nosé-Hoover thermostat is applied to control the temperature. Further details of the simulation method, along with further tests of its validity, were reported in 23,24. Among other considerations, a simulation must use a sufficiently short time step to conserve energy and an adequately large box size to allow computing meaningful quantity at long times. The required box size is generally the desired maximum time multiplied by the sound speed for 1539-3755/2007/75 1 /016405 5 016405-1 2007 The American Physical Society

BIN LIU AND J. GOREE ordered liquids or ballistic particle speed for gases. III. SELF-DIFFUSION We use three diagnostics to characterize particle motion. We found they are consistent with one another when classifying the motion as diffusive or superdiffusive. However, investigators may differ in their classification, depending on their system size, which is related to the time range where data can be evaluated, and qualitative judgments of the appearance of a curve. The first diagnostic, mean-squared displacement MSD, r i t r i 0 2, is the most common. Here, r i 0 is the initial position of particle i. Normal diffusion, i.e., Fickian diffusion, can be identified when the MSD scales linearly with time t, while for superdiffusion MSD increases more rapidly. When graphed as a function of time in a log-log plot, if the curve fits a straight line with a slope, then the MSD obeys a power low t. Normal diffusion is characterized by =1, while superdiffusion and subdiffusion have larger and smaller slopes, respectively. Most authors mention =1 as a criterion for classifying motion as diffusive, but in practice this is not useful because actual empirical data will never yield exactly a slope of unity. While most investigators do not apply a numerical criterion, we follow the prescription of Feder et al. 25 : a range of 0.9 1.1 is classified as diffusive motion. Here, we also make a stricter test of the MSD curve using the prediction of the Stokes-Einstein relation. We combine three relations: the Stokes-Einstein relation D=k B T/1.69 23, which relates the shear viscosity coefficient and diffusion constant D; the Einstein relation lim t r t r 0 2 =4Dt; and the Green-Kubo relation for computing from simulation data. This method assumes a valid viscosity coefficient exists, which has been verified in 24,26. Thus, the value of D predicted in this way can be used to compute predicted MSD data, which can be compared to the simulation MSD data to determine whether the value of D is valid and therefore determine whether the motion is actually diffusive. This test is stricter than relying solely on the slope because it makes use of both the slope and the intercept. The second diagnostic, PDF, is the probability distribution function, which is a histogram of the squared displacement r 2 made by particles as compared to their initial positions. Diffusive motion is characterized by a PDF that is Gaussian; in contrast, motion that obeys Lévy statistics is characterized by a PDF that has an inverse power law 27. The PDF is therefore useful for identifying whether superdiffusion is due to Lévy statistics, as was claimed for the experiment of 17. The third diagnostic, VACF, is the ensemble average of the product of particle velocities at time t and an earlier time 0. The VACF is computed as v i t v i 0 /2. If the motion is normal diffusion, the VACF will decay much faster than 1/t, and a valid diffusion coefficient can be computed as the integral, i.e., the area under the VACF, using the Green-Kubo relation 28. On the other hand, if motion is superdiffusive, the VACF will decay much more slowly, typically as 1/t, and the integral diverges due to this long-time tail. If the VACF FIG. 1. Pair correlation function g r from our MD simulation, for some of the values of reported in this paper. The states varied from a nonideal gas at =1 to a strongly coupled liquid at =136. Temperatures here are normalized by T m as reported by Hartmann et al. 22. oscillates and has an integral that tends to zero, the motion is identified as subdiffusive. The VACF is most useful at higher temperatures, well above melting, whereas at lower temperatures it has oscillations that hinder identifying the scaling of a long-time tail. For this reason, at low temperatures the MSD or PDF can be more useful than the VACF. 016405-2

DIMENSIONAL YUKAWA LIQUIDS FIG. 2. Mean-squared displacement MSD. Simulation results are shown as a solid line. Superdiffusion is indicated by a slope 1, and by an MSD below the dashed line except at very long time. The dashed line is a prediction based on a combination of the Stokes-Einstein relation, the Einstein relation, and the viscosity computed in 24 from the Green-Kubo relation. Note that the system shown here at =18 and =0.56 is a disordered liquid, as shown by particle trajectories in the inset for a period of 0 pd t 10. Distance is normalized by the Wigner-Seitz radius a. IV. RESULTS This simulation studies disordered Yukawa liquids, covering from nonideal gases to near disordering transition strongly coupled liquid. As a measurement of the order or disorder of the liquid, we present the pair correlation function, g r, in Fig. 1. This is useful to allow the reader to compare our results, for a given value of and, to other authors results for different values of these parameters or even a different potential provided that g r appears similar. A. Superdiffusive case, =18 Results for the MSD, in Fig. 2, show that the motion is superdiffusive, for =18. Here we are interested in motion for pd t 1. At shorter times the motion is ballistic, as a particle moves inside the cage formed by neighboring particles. We find that the MSD grows faster than linearly, as compared to the MSD calculated by using the Stokes- Einstein relation. First, in the most common test of superdiffusion, here we find that the slope is greater than unity. Calculating the exponent, = d log r i t r i 0 2 1 d log t and averaging this value in the range 30 pd t 70, we find =1.30 for =18. Moreover, comparing the simulation results to the dashed line in Fig. 2 predicted by using Stokes- Einstein relation, we find a noticeable deviation. Both of these findings, 1 and an MSD curve that deviates from the prediction of the Stokes-Einstein relation, indicate that the motion is superdiffusion. Results for the PDF, Fig. 3, indicate that the motion is superdiffusive. Instead of plotting the PDF vs r 2, here we plot it vs r 2 /t. If motion were diffusive, the data would fall on a universal curve. For the intermediate-temperature regime studied here, we found that the data never fall on a FIG. 3. Probability distribution function, PDF. This is a histogram of the squared displacement r 2 made by a particle during a measurement interval t. A straight line in this semilogarithmic graph indicates a Gaussian PDF, which rules out Lévy statistics. a In the intermediate-temperature regime, the width of the Gaussian increases faster than t, indicating superdiffusion. b In a cold liquid, data for pd t 10 fall on a universal curve, indicating diffusive motion or much less extreme superdiffusion. Motion for pd t 10 corresponds to ballistic motion, which is not of interest when classifying random motion as diffusion or superdiffusion. universal curve, so that we identify the motion as nondiffusive. In particular, the data fall on a line with a slope that becomes less steep with time, as in Fig. 3 a for =18, and this is an indication of superdiffusion. For comparison, we also show in Fig. 3 b the PDF for the different regime of a cold liquid, =106, where the data fall on a universal curve for pd t 10. The results for the PDF also indicate that the superdiffusive motion we observed is not attributable to Lévy statistics. As plotted in Fig. 3, the data fall on a straight line when graphed with a logarithmic vertical axis and a horizontal axis representing squared displacement. This result shows that the PDF is Gaussian. Lévy statistics, where large displacements happen frequently and can be a possible cause of superdiffusion, can be excluded here because the PDF is not a power law as would be expected for Lévy statistics. One of the 016405-3

BIN LIU AND J. GOREE FIG. 4. Velocity autocorrelation function, VACF, shown with linear axes in a and log-log axes in b. For the intermediatetemperature regime, the VACF decays as slowly as 1/t, i.e., it has a long-time tail. This long-time tail indicates superdiffusion, and is considered to be a signature of mode coupling. experimental groups claiming to observe superdiffusion in a monolayer dusty plasma attributed their result to Lévy flights arising from a rotation of domains 17, which can occur under nonequilibrium conditions. This apparently did not occur in our equilibrium simulation. Results for the VACF, Fig. 4, show a long-time tail in the VACF, indicating nondiffusive motion. The results are again for =18. We found that for large times, after the initial damped oscillations due to caged particle oscillating motion, the VACF decays as slowly as 1/t. The long-time tail causes a divergence when attempting to calculate a diffusion coefficient using the Green-Kubo relation 28. This divergence indicates superdiffusion. B. Dependence on Varying the temperature, we found the dependence on of the MSD results shown in Fig. 5. We always used the same time range, except for the two highest temperatures. The exponent has its maximum value, indicating the most extreme superdiffusion, at 18. This result is significant because this happens at the same value of where the viscosity was previously shown to have a minimum 24,26.We performed our simulations for only one value of ; in general, the value of at the minimum viscosity depends slightly on 20. Using the VACF, we found that at high temperatures the decay was much faster than 1/t, indicating that either the FIG. 5. Exponent, computed from the MSD curve using Eq. 1. A value 1 indicates superdiffusion, which occurs prominently in the intermediate-temperature regime. Data are also shown for colder and hotter regimes. Maximum superdiffusion occurs approximately at =18, where the viscosity happens to have a minimum 24,26. The two data points indicated by the asterisk are less reliable than the others because the value of is found by fitting a portion of the MSD curve which, for our simulation box size, is very limited. For low, it is better to rely on the VACF. motion is diffusive or at least that any superdiffusion is not extreme. This diagnostic is especially useful at high temperatures because the portion of the MSD curve used for fitting the exponent is limited for our simulation box size. Near the disordering transition the VACF is more difficult to interpret because it exhibits oscillations. V. DISCUSSION A1/t long-time tail, as we have observed, for example, at =18, is a signature of mode-coupling in 2D equilibrium liquids. According to the mode-coupling theory, fluctuations decay into pairs of hydrodynamic modes 9. The hydrodynamic modes affect the tail of the VACF. Our finding a lack of Lévy statistics indicates that the modes do not consist primarily of zero-frequency counter-rotating vortices. Additionally, we report the result that the superdiffusion and long-time tails are most extreme for the condition where the viscosity has a minimum. For all liquids a coupling between diffusion and viscosity is expected, and for many liquids this is manifested by a Stokes-Einstein relation 23, although the Stokes-Einstein relation does not occur here because the motion is nondiffusive. Nevertheless, a relationship between random motion and viscosity is revealed here, by the coincidence of a minimum viscosity and maximum superdiffusion. We thank Z. Donkó, V. Nosenko, and F. Skiff for helpful discussions. This work was supported by NASA and DOE. 016405-4

DIMENSIONAL YUKAWA LIQUIDS 1 C. C. Grimes and G. Adams, Phys. Rev. Lett. 42, 795 1979. 2 P. L. Gammel, D. J. Bishop, G. J. Dolan, J. R. Kwo, C. A. Murray, L. F. Schneemeyer, and J. V. Waszczak, Phys. Rev. Lett. 59, 2592 1987. 3 A. A. Abrikosov, Sov. Phys. JETP 5, 1174 1957. 4 C. A. Murray, W. O. Sprenger, and R. A. Wenk, Phys. Rev. B 42, 688 1990. 5 T. B. Mitchell, J. J. Bollinger, X.-P. Huang, W. M. Itano, and D. H. E. Dubin, Phys. Plasmas 6, 1751 1999. 6 S. Nunomura, D. Samsonov, S. Zhdanov, and G. Morfill, Phys. Rev. Lett. 96, 015003 2006. 7 Y. Maruyama and J. Murakami, Phys. Rev. B 67, 085406 2003. 8 R. Brown, Edinburgh New Philosophical Journal 5, 358 1828. 9 M. H. Ernst, E. H. Hauge, and J. M. J. Van Leeuwen, Phys. Rev. Lett. 25, 1254 1970. 10 J. R. Dorfman and E. G. D. Cohen, Phys. Rev. Lett. 25, 1257 1970. 11 B. J. Alder and T. E. Wainwright, Phys. Rev. A 1, 18 1970. 12 R. Zangi and S. A. Rice, Phys. Rev. Lett. 92, 035502 2004. 13 D. N. Perera and P. Harrowell, Phys. Rev. Lett. 81, 120 1998. 14 M. M. Hurley and P. Harrowell, Phys. Rev. E 52, 1694 1995. 15 P. J. Camp, Phys. Rev. E 71, 031507 2005. 16 W. T. Juan and Lin I, Phys. Rev. Lett. 80, 3073 1998. 17 S. Ratynskaia, K. Rypdal, C. Knapek, S. Khrapak, A. V. Milovanov, A. Ivlev, J. J. Rasmussen, and G. E. Morfill, Phys. Rev. Lett. 96, 105010 2006. 18 U. Konopka, G. E. Morfill, and L. Ratke, Phys. Rev. Lett. 84, 891 2000. 19 H. Ohta and S. Hamaguchi, Phys. Plasmas 7, 4506 2000. 20 K. Y. Sanbonmatsu and M. S. Murillo, Phys. Rev. Lett. 86, 1215 2001. 21 G. J. Kalman, P. Hartmann, Z. Donkó, and M. Rosenberg, Phys. Rev. Lett. 92, 065001 2004. 22 P. Hartmann, G. J. Kalman, Z. Donkó, and K. Kutasi, Phys. Rev. E 72, 026409 2005. 23 B. Liu, J. Goree, and O. S. Vaulina, Phys. Rev. Lett. 96, 015005 2006. 24 B. Liu and J. Goree, Phys. Rev. Lett. 94, 185002 2005. 25 T. J. Feder, I. Brust-Mascher, J. P. Slattery, B. Baird, and W. W. Webb, Biophys. J. 70, 2767 1996. 26 Z. Donkó, J. Goree, P. Hartmann, and K. Kutasi, Phys. Rev. Lett. 96, 145003 2006. 27 B. J. West and W. Deering, Phys. Rep. 246, 1 1994. 28 J.-P. Hansen and I. R. McDonald, Theory of Simple Liquids, 2nd ed. Elsevier-Academic, San Diego, 1986. 016405-5