shear viscosity and shear thinning in two-dimensional yukawa...

4
Shear Viscosity and Shear Thinning in Two-Dimensional Yukawa Liquids Z. Donko ´, 1 J. Goree, 2 P. Hartmann, 1 and K. Kutasi 1 1 Research Institute for Solid State Physics and Optics of the Hungarian Academy of Sciences, H-1525 Budapest, P.O. Box 49, Hungary 2 Department of Physics and Astronomy, The University of Iowa, Iowa City, Iowa 52242, USA (Received 2 September 2005; revised manuscript received 6 February 2006; published 13 April 2006) A two-dimensional Yukawa liquid is studied using two different nonequilibrium molecular dynamics simulation methods. Shear viscosity values in the limit of small shear rates are reported for a wide range of Coulomb coupling parameter and screening lengths. At high shear rates it is demonstrated that this liquid exhibits shear thinning; i.e., the viscosity diminishes with increasing shear rate. It is expected that two- dimensional dusty plasmas will exhibit this effect. DOI: 10.1103/PhysRevLett.96.145003 PACS numbers: 52.27.Gr, 52.27.Lw, 82.70.y Many-particle systems characterized by the Yukawa potential include a variety of physical systems, e.g., dusty plasmas, charged colloids, astrophysical objects, and high energy density matter. The Yukawa potential r/ Q expr= D =r models a Coulomb repulsion that is ex- ponentially suppressed with a screening length D . Yukawa systems behave like a liquid when the temperature exceeds a melting point which depends on Q, D , and particle spacing, e.g., [1,2]. Transport parameters of Yukawa systems—the diffu- sion coefficient [3], the shear viscosity [4–6], and the thermal conductivity [6,7]—have mainly been calculated for 3D systems, but there is now an increasing interest in 2D settings. For example, in dusty plasma experiments, charged microspheres suspended as a monolayer in a gas discharge make a 2D Yukawa system. By creating a shear flow in such a particle suspension, the viscosity was mea- sured in recent experiments using 2D suspensions [8] and quasi-2D suspensions consisting of a few monolayers of charged microspheres [9]. The transport properties of such ultrathin liquids are also of interest as macroscopic analogs of molecular flow in nanoscience applications [10]. Transport coefficients are meaningful if they are part of a valid ‘‘constitutive relation’’ between the gradients of local variables and fluxes. For shear viscosity , the con- stitutive relation j y dv x y=dy relates a momen- tum flux j y to the velocity gradient dv x y=dy, which is also termed the shear rate. In a non-Newtonian fluid, may vary with the velocity gradient, whereas in Newtonian fluids it does not. In particular, if diminishes as shear is increased, the fluid is said to exhibit ‘‘shear thinning.’’ This occurs in simple liquids [11], as well as in complex mixtures such as foams, micelles, slurries, pastes, gels, polymer solutions, and granular flows [12]. Recently, ex- perimenters have claimed to observe shear thinning in dusty plasma liquids [9]. These reports motivate our simu- lations to search for the presence of shear thinning in 2D Yukawa liquids. Subsequent to the experimental measurement of viscos- ity in a 2D dusty plasma [8], a 2D molecular dynamics simulation was used to obtain the shear viscosity from the Green-Kubo relations [13]. In this Letter we will go be- yond the results of Ref. [13], which were performed for equilibrium conditions, by using nonequilibrium simula- tions to search for non-Newtonian behavior under condi- tions of a high shear rate. We will also compute the viscosity over a wider range of and . Our simulations use a rectangular cell with edge lengths L x and L y and periodic boundary conditions. The number of particles is between N 990 and 7040. The system is characterized by dimensionless parameters Q 2 =4" 0 ak B T and a= D , where a 1=n 1=2 is the Wigner-Seitz radius, with n being the areal density. Additional parameters include the thermal velocity v 0 2k B T=m 1=2 and the 2D analog of the plasma frequency ! p Q 2 =2" 0 ma 3 1=2 , the shear rate dv x =dy, and its normalized value dv x =dya=v 0 . Two types of molecular dynamics techniques are applied for the studies of the shear viscosity. Method 1 reverses the cause-and-effect picture custom- arily used in nonequilibrium molecular dynamics: the ef- fect—the momentum flux—is imposed, and the cause— the velocity gradient (shear rate) —is measured in the simulation [14]. Momentum in the liquid is introduced in a pair of narrow slabs A and B, which are centered at y L y =4 and 3L y =4, respectively. At regular time intervals we identify the particles in slabs A and B having the highest jv x j in the positive and negative directions, respectively. We then instantaneously exchange the v x velocity compo- nent of these two particles without moving the particles. This artificial transfer of momentum between slabs A and B (which is accomplished without changing the system en- ergy) produces a velocity profile v x y, the slope of which can be controlled by the frequency of the momentum exchange steps. The equations of motion dr i dt p i m ; dp i dt F i ; (1) where r x; y, p p x ;p y are the positions and the momenta of particles, m is their mass, and F i is the force PRL 96, 145003 (2006) PHYSICAL REVIEW LETTERS week ending 14 APRIL 2006 0031-9007= 06=96(14)=145003(4)$23.00 145003-1 © 2006 The American Physical Society

Upload: others

Post on 19-Jul-2020

7 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Shear Viscosity and Shear Thinning in Two-Dimensional Yukawa …dusty.physics.uiowa.edu/~goree/papers/donko-shear... · 2007-02-27 · Shear Viscosity and Shear Thinning in Two-Dimensional

PRL 96, 145003 (2006) P H Y S I C A L R E V I E W L E T T E R S week ending14 APRIL 2006

Shear Viscosity and Shear Thinning in Two-Dimensional Yukawa Liquids

Z. Donko,1 J. Goree,2 P. Hartmann,1 and K. Kutasi11Research Institute for Solid State Physics and Optics of the Hungarian Academy of Sciences, H-1525 Budapest, P.O. Box 49, Hungary

2Department of Physics and Astronomy, The University of Iowa, Iowa City, Iowa 52242, USA(Received 2 September 2005; revised manuscript received 6 February 2006; published 13 April 2006)

0031-9007=

A two-dimensional Yukawa liquid is studied using two different nonequilibrium molecular dynamicssimulation methods. Shear viscosity values in the limit of small shear rates are reported for a wide range ofCoulomb coupling parameter and screening lengths. At high shear rates it is demonstrated that this liquidexhibits shear thinning; i.e., the viscosity � diminishes with increasing shear rate. It is expected that two-dimensional dusty plasmas will exhibit this effect.

DOI: 10.1103/PhysRevLett.96.145003 PACS numbers: 52.27.Gr, 52.27.Lw, 82.70.�y

Many-particle systems characterized by the Yukawapotential include a variety of physical systems, e.g., dustyplasmas, charged colloids, astrophysical objects, andhigh energy density matter. The Yukawa potential ��r� /Q exp��r=�D�=r models a Coulomb repulsion that is ex-ponentially suppressed with a screening length �D.Yukawa systems behave like a liquid when the temperatureexceeds a melting point which depends on Q, �D, andparticle spacing, e.g., [1,2].

Transport parameters of Yukawa systems—the diffu-sion coefficient [3], the shear viscosity [4–6], and thethermal conductivity [6,7]—have mainly been calculatedfor 3D systems, but there is now an increasing interest in2D settings. For example, in dusty plasma experiments,charged microspheres suspended as a monolayer in a gasdischarge make a 2D Yukawa system. By creating a shearflow in such a particle suspension, the viscosity was mea-sured in recent experiments using 2D suspensions [8] andquasi-2D suspensions consisting of a few monolayers ofcharged microspheres [9]. The transport properties of suchultrathin liquids are also of interest as macroscopic analogsof molecular flow in nanoscience applications [10].

Transport coefficients are meaningful if they are part ofa valid ‘‘constitutive relation’’ between the gradients oflocal variables and fluxes. For shear viscosity �, the con-stitutive relation jy � ���dvx�y�=dy� relates a momen-tum flux jy to the velocity gradient dvx�y�=dy, which isalso termed the shear rate. In a non-Newtonian fluid, �may vary with the velocity gradient, whereas in Newtonianfluids it does not. In particular, if � diminishes as shear isincreased, the fluid is said to exhibit ‘‘shear thinning.’’ Thisoccurs in simple liquids [11], as well as in complexmixtures such as foams, micelles, slurries, pastes, gels,polymer solutions, and granular flows [12]. Recently, ex-perimenters have claimed to observe shear thinning industy plasma liquids [9]. These reports motivate our simu-lations to search for the presence of shear thinning in 2DYukawa liquids.

Subsequent to the experimental measurement of viscos-ity in a 2D dusty plasma [8], a 2D molecular dynamics

06=96(14)=145003(4)$23.00 14500

simulation was used to obtain the shear viscosity from theGreen-Kubo relations [13]. In this Letter we will go be-yond the results of Ref. [13], which were performed forequilibrium conditions, by using nonequilibrium simula-tions to search for non-Newtonian behavior under condi-tions of a high shear rate. We will also compute theviscosity over a wider range of � and �.

Our simulations use a rectangular cell with edge lengthsLx and Ly and periodic boundary conditions. The numberof particles is between N � 990 and 7040. The systemis characterized by dimensionless parameters � �Q2=4�"0akBT and � � a=�D, where a � �1=n��1=2 isthe Wigner-Seitz radius, with n being the areal density.Additional parameters include the thermal velocity v0 �

�2kBT=m�1=2 and the 2D analog of the plasma frequency!p � �Q

2=2�"0ma3�1=2, the shear rate � � dvx=dy, and

its normalized value �� � �dvx=dy��a=v0�. Two types ofmolecular dynamics techniques are applied for the studiesof the shear viscosity.

Method 1 reverses the cause-and-effect picture custom-arily used in nonequilibrium molecular dynamics: the ef-fect—the momentum flux—is imposed, and the cause—the velocity gradient (shear rate)—is measured in thesimulation [14]. Momentum in the liquid is introduced ina pair of narrow slabs A and B, which are centered at y �Ly=4 and 3Ly=4, respectively. At regular time intervals �we identify the particles in slabs A and B having the highestjvxj in the positive and negative directions, respectively.We then instantaneously exchange the vx velocity compo-nent of these two particles without moving the particles.This artificial transfer of momentum between slabs A andB(which is accomplished without changing the system en-ergy) produces a velocity profile vx�y�, the slope of whichcan be controlled by the frequency of the momentumexchange steps. The equations of motion

dridt�

pim;

dpidt� Fi; (1)

where r � �x; y�, p � �px; py� are the positions and themomenta of particles, m is their mass, and Fi is the force

3-1 © 2006 The American Physical Society

Page 2: Shear Viscosity and Shear Thinning in Two-Dimensional Yukawa …dusty.physics.uiowa.edu/~goree/papers/donko-shear... · 2007-02-27 · Shear Viscosity and Shear Thinning in Two-Dimensional

-0.3-0.2-0.10.00.10.20.3

0.0 0.2 0.4 0.6 0.8 1.0

0.292

0.296

0.300

0.304

v x / v

0

ωp τ = 40

ωp τ = 20

ωp τ = 10

Ty [

K]

y / LY

FIG. 1. (a) Velocity profiles vx�y� obtained from method 1 fordifferent frequencies (1=�) of momentum exchange steps, and(b) Ty�y� temperature profiles for the same conditions. � � 100,� � 1.

PRL 96, 145003 (2006) P H Y S I C A L R E V I E W L E T T E R S week ending14 APRIL 2006

acting on particle i, are integrated by the velocity Verletalgorithm.

Method 2 simulates a planar Couette flow, which isestablished by the Lees-Edwards periodic boundary con-ditions resulting in a homogeneous streaming flow field inthe simulation box: hvxi � ��y� Ly=2�, where hi denotesa time average. The system is described by the Gaussianthermostated SLLOD equations of motion [15]:

dridt�

~pim� �yix;

d~pidt� Fi � �~pyix� �~pi; (2)

where ~p � �~px; ~py� is the peculiar momentum of particles,x is the unit vector in the x direction, and � is the Gaussianthermostating multiplier. The above set of equations issolved using an operator splitting technique [16].

In contrast to method 1, method 2 results in a homoge-neous shear field and a constant temperature within thewhole simulation box. Thus, arbitrarily high shear ratesmay be established without the need of considering anyeffects of temperature gradients on the viscosity.

In both methods the pairwise Yukawa interparticleforces are summed over a �-dependent cutoff radius, usingthe chaining mesh technique. (The force due to particles atthe cutoff radius is� 10�5 smaller compared to that due tothe nearest neighbors.) Both methods neglect any neutralgas drag, which has been observed to alter the velocityprofiles in experiments [8], i.e., they model an atomicsystem where momentum transfer is dominated byCoulomb collisions [1].

Method 1 has the advantage that resembles more closelythe experimental conditions, although the procedure forapplying shear in the simulation involves no introductionor removal of energy from the system. In the experiment[8] shear is applied via an external introduction of bothmomentum and energy in a boundary slab while energy issimultaneously removed elsewhere by frictional dissipa-tion. Method 2 represents a well-established technique formeasurement of viscosity at arbitrary steady, as well astemporally varying shear rates. Although it has little con-nection to the conditions found in the experiment, it hasbeen demonstrated to be an efficient technique to inves-tigate shear thinning [11,15]. Thus we apply this methodfor the studies of this latter effect.

Near-equilibrium (small �) shear viscosity values havebeen obtained using both techniques. In method 1 this isdone at the lowest practical shear rate, where dvx=dy isuniform between slabs A and B. We calculate �eq from

jjyj � �eqdvx�y�=dy � �p=2tsimLy; (3)

where �p is the total x-directional momentum exchangedbetween slabs A and B during the simulation time tsim [14].In method 2 the off-diagonal element of the pressure tensoris measured during the course of the simulation:

14500

Pxy�t� �XNi�1

�mvixviy �

XNj>i

xijyijrij

ddrij

��rij��; (4)

where rij � ri � rj � �xij; yij�, and the shear viscosity isobtained as

� � limt!1hPxy�t�i=�: (5)

In method 1, the spatial profiles for temperature andvelocity, Fig. 1, develop self-consistently in response tothe perturbation applied by introducing momentum inslabs A and B. We use method 1 only for small perturba-tions, so that the velocity profile has a linear gradient andthe temperature is isotropic, with Tx � Ty, where Tx;y �

�m=NjkB�PNji�1h�vix;y�t� � vjx;y�

2i. The index i runs overthe Nj particles in slab j. We verified that vjy is negligiblysmall.

Obtaining reliable results for � at small � requires asimulation duration of typically !pt 104–105 for bothmethods. The required time step is smallest and the simu-lations are most costly at low �. In method 1, system sizeeffects are expected to appear when (i) particles traversethe simulation box without significant interaction with theothers, or (ii) the compressional sound wave transits thebox in a shorter time than the decay time tc of the velocityautocorrelation function. For (i), for our most demandingcondition (small size N � 990 and high temperature � �1) a particle moving at the thermal velocity would transitthe cell in a time !pt � 57 if it were undeflected bycollisions. We find that the decay time is short enough,!ptc 5–10, even for the smallest � values of interest.Thus we expect no ‘‘ballistic’’ trajectories across the entiresimulation box. For (ii), the sound speed [17] at � � 1 isv � d!=dk a!p, and the wave’s transit time �t across a

3-2

Page 3: Shear Viscosity and Shear Thinning in Two-Dimensional Yukawa …dusty.physics.uiowa.edu/~goree/papers/donko-shear... · 2007-02-27 · Shear Viscosity and Shear Thinning in Two-Dimensional

0.0 0.2 0.4 0.6 0.8 1.00.0

0.2

0.4

0.6

0.8

1.0

0.0 0.2 0.4 0.6 0.8 1.00.0

0.2

0.4

0.6

0.8

1.0

x / LX

y / L

Y

(b)

x / LX

(a)

FIG. 2. (a) Trajectories of particles in the simulation based onmethod 2. (a) � � 10, � � 1 at a shear rate �� � 0:2; time ofrecording: !p�T � 5:0. (b) � � 100, � � 1, �� � 0:05,!p�T � 23:6. The shear field is vx � ��y� Ly=2�, i.e., thereis no flow at y � Ly=2. N � 1020.

0.5 1 10 100 5000.0

0.2

0.4

0.6

0.8

1.0

1.2

0.5 1 10 100 3000.0

0.5

1.0

1.5

2.0

(a)

η eq

/ η 0

Γ

κ N 0.5 1600 0.56 1024 (IOWA EMD) 1.0 990 1.0 3960 1.0 1020 (SLLOD) 2.0 990 2.0 3960

(b)

η eq

/ η E

normalized temperature T ' =TY / T

m

FIG. 3. (a) Shear viscosity at near-equilibrium conditions, �eq,obtained from method 1, at the simulation’s lowest practicalshear rate, and normalized by �0 � mn!pa

2. For comparison,data are shown from the Iowa equilibrium MD simulation [13]and from simulations based on method 2 (SLLOD), in the limitof small shear rates, at � � 1. N is the number of simulationparticles. (b) A scaling law is demonstrated by normalizing thedata in (a) using �E � mn!Ea2 and T0 � Ty=Tm, where Tm isthe melting temperature. The thick line is an empirical fit of form�eq=�E � aT0 � b=T0 � c.

PRL 96, 145003 (2006) P H Y S I C A L R E V I E W L E T T E R S week ending14 APRIL 2006

box with length Ly is !p�t Ly=a � 57 for our mostdemanding case, N � 990 particles. Thus, we find bothcriteria fulfilled for a ‘‘sufficiently large’’ system.Method 2 is known to produce accurate results even forthe small number of particles simulated [15]. We verifiedthat the results obtained from both methods did not dependsignificantly on N.

Figures 2(a) and 2(b) illustrate particle trajectories insimulations based on method 2, for conditions � � 10,� � 1 at a shear rate �� � 0:2, and for � � 100, � � 1,�� � 0:05, respectively.

Our results for�eq as a function of �, for different valuesof � are plotted in Fig. 3(a). We find a good agreement withthe earlier equilibrium molecular-dynamic (MD) simula-tion of Ref. [13]. In contrast with most simple liquids,which have a viscosity that varies monotonically withtemperature, a prominent feature of the viscosity of thepresent system is a minimum (e.g., at � 20 for � � 1),which has been noted previously in both one-componentplasma (OCP) and Yukawa liquids. The shape of the�eq���curve can be explained by the prevailing kinetic and po-tential contributions to the viscosity at low and high valuesof �, respectively. The near-equilibrium shear viscosityvalues obtained with method 2 for � � 1 are also displayedFig. 3(a). We find an excellent agreement between theresults of methods 1 and 2.

Similar to what was observed in [5] for 3D Yukawaliquids, we find that the near-equilibrium viscosity �eq

obeys a scaling law as demonstrated in Fig. 3(b), whereviscosity has been normalized by �E � mn!Ea2. TheEinstein frequency !E depends on �, and we computed itfrom Eq. (7) of Ref. [17] using pair-correlation functionsmeasured from our simulations. The horizontal axis is anormalized temperature T0 � Ty=Tm � �m=�, where Tmand �m are melting-point values reported in Ref. [2]. Usingthese normalizations, the data fall on the same curve,demonstrating the existence of a scaling law for the 0:5 �

14500

� � 2:0 range of the screening parameter. We note that forthis purpose we found !E was more significant than !p.The near-equilibrium viscosity is fit by an empirical form(like in [5] for three dimensions) �eq=�E � aT0 � b=T0 �c with coefficients: a � 0:0093, b � 0:78, and c � 0:098.

A shear-thinning effect is revealed in Fig. 4(a), whichshows that � diminishes significantly as the shear rate �� isincreased. In other two-dimensional systems the reductionin �, as compared to the value at small shear, was observedto vary as the square root of �� [11]. We find that this scalingalso occurs for the Yukawa system, as indicated by datathat fall on nearly straight lines in Fig. 4(a) for �� > 0:2. Atsmaller shear rates, �� < 0:2, however, the shear-thinningeffect is less profound and the liquid is more nearlyNewtonian, especially for large �. Results are shown for�� * 0:01, which we found to be reliable, whereas at lower��, method 2 yielded noisy data even for very longsimulations.

Because viscosity arises from both kinetic and potentialcontributions, Eq. (4), we evaluate which of these contri-

3-3

Page 4: Shear Viscosity and Shear Thinning in Two-Dimensional Yukawa …dusty.physics.uiowa.edu/~goree/papers/donko-shear... · 2007-02-27 · Shear Viscosity and Shear Thinning in Two-Dimensional

0.0 0.2 0.4 0.6 0.8 1.00.05

0.10

0.15

0.20

0.25

0.30

0.35

0.0 0.2 0.4 0.6 0.8 1.0-0.05

0.00

0.05

0.10

0.15

0.20

0.25

0.30

η / η

0

γ 1/2

Γ = 10 Γ = 50 Γ = 100 Γ = 120

(a)

(b)

η KIN

/ η

0 ,

ηP

OT /

η 0

γ 1/2

, Γ = 5, Γ = 20, Γ = 100

FIG. 4. (a) Shear viscosity as a function of normalized shearrate �� for selected values of the coupling parameter �.(b) Potential (solid symbols) and kinetic (open symbols) con-tributions to the shear viscosity. � � 1.

PRL 96, 145003 (2006) P H Y S I C A L R E V I E W L E T T E R S week ending14 APRIL 2006

butions is most responsible for the observed shear-thinningeffect in Fig. 4(b). Recall that for equilibrium conditions,the kinetic term dominates for �� 20 and the potentialterm dominates for � 20. Here, we find that for non-equilibrium conditions, as the shear rate increases thereduction in viscosity is mostly due to a reduction of thekinetic contribution at low � and a reduction of the poten-tial contribution at high �. In other words, at extremevalues of �, it is the same term dominating the equilibriumviscosity that also dominates the shear-thinning effect. Atan intermediate value of � � 20, however, where the equi-librium viscosity has a minimum and the two equilibriumcontributions are comparable (for � � 1), we find that it ismostly a reduction of the kinetic term that accounts for theobserved shear-thinning effect.

In summary, we have calculated the shear viscositycoefficient of 2D Yukawa liquids in a wide domain ofparameters � and � using two different molecular dynam-ics approaches. The small shear rate calculations con-firmed that the two techniques used are consistent and

14500

yielded � values in fair agreement with equilibrium MDcalculations [13]. The small shear rate data were found toobey a universal scaling: � normalized by the Einsteinfrequency was found to depend only on the reduced tem-perature (ratio of the temperature to melting temperature).The high shear rate simulations based on method 2 unam-biguously demonstrated a non-Newtonian behavior of theYukawa liquid: � was significantly reduced for these con-ditions in a manifestation of shear thinning, except at thelowest shear rates where the liquid is more nearlyNewtonian. Regimes of the plasma coupling parameterwere identified to distinguish whether the kinetic or poten-tial contribution to the shear viscosity is primarily respon-sible for the shear-thinning effect.

We thank G. J. Kalman and Bin Liu for useful discus-sions. This work was supported by the Hungarian Fund forScientific Research and the Hungarian Academy ofSciences, OTKA-T-48389, MTA-OTKA-90/46140,OTKA-PD-049991. J. G. was supported by NASA andDOE.

3-4

[1] S. A. Khrapak, A. V. Ivlev, and G. E. Morfill, Phys. Rev. E70, 056405 (2004).

[2] P. Hartmann, G. J. Kalman, Z. Donko, and K. Kutasi,Phys. Rev. E 72, 026409 (2005).

[3] H. Ohta and S. Hamaguchi, Phys. Plasmas 7, 4506 (2000).[4] K. Y. Sanbonmatsu and M. S. Murillo, Phys. Rev. Lett. 86,

1215 (2001).[5] T. Saigo and S. Hamaguchi, Phys. Plasmas 9, 1210 (2002).[6] G. Salin and J.-M. Caillol, Phys. Rev. Lett. 88, 065002

(2002); G. Faussurier and M. S. Murillo, Phys. Rev. E 67,046404 (2003).

[7] Z. Donko and P. Hartmann, Phys. Rev. E 69, 016405(2004).

[8] V. Nosenko and J. Goree, Phys. Rev. Lett. 93, 155004(2004).

[9] C.-L. Chan, W.-Y. Woon, and L. I, Phys. Rev. Lett. 93,220602 (2004).

[10] S. Granick, Phys. Today 52, No. 7, 26 (1999); Subrata Royet al., J. Appl. Phys. 93, 4870 (2003).

[11] D. J. Evans, Phys. Rev. A 23, 1988 (1981).[12] A. Kabla and G. Debregeas, Phys. Rev. Lett. 90, 258303

(2003); J. B. Salmon, A. Colin, S. Manneville, andF. Molino, Phys. Rev. Lett. 90, 228303 (2003);F. Varnik, L. Bocquet, J.-L. Barrat, and L. Berthier,Phys. Rev. Lett. 90, 095702 (2003); Y. Jiang et al.,Phys. Rev. E 59, 5819 (1999).

[13] Bin Liu and J. Goree, Phys. Rev. Lett. 94, 185002 (2005).[14] F. Muller-Plathe, Phys. Rev. E 59, 4894 (1999).[15] D. J. Evans and G. P. Morriss, Statistical Mechanics of

Nonequilibrium Liquids (Academic, New York, 1990).[16] G. Pan, J. F. Ely, C. McCabe, and D. J. Isbister, J. Chem.

Phys. 122, 094114 (2005).[17] G. J. Kalman, P. Hartmann, Z. Donko, and M. Rosenberg,

Phys. Rev. Lett. 92, 065001 (2004).