1: \documentclass[12pt]{iopart}
2:
3: \usepackage[dvips]{graphicx}
4: \usepackage{iopams}
5: \usepackage{hyperref}
6: \begin{document}
7:
8: \sloppy
9:
10: \title[A new model for simulating colloidal dynamics]
11: {A new model for simulating colloidal dynamics}
12:
13: \author{Vladimir Lobaskin\footnote[1]{
14: To whom correspondence should be addressed
15: (lobaskin@mpip-mainz.mpg.de)}
16: and Burkhard D\"unweg}
17:
18: \address{Max Planck Institute for Polymer Research,
19: Ackermannweg 10,
20: D-55128 Mainz, Germany}
21:
22: \begin{abstract}
23: We present a new hybrid lattice-Boltzmann and Langevin molecular
24: dynamics scheme for simulating the dynamics of suspensions of
25: spherical colloidal particles. The solvent is modeled on the level of
26: the lattice-Boltzmann method while the molecular dynamics is done for
27: the solute. The coupling between the two is implemented through a
28: frictional force acting both on the solvent and on the solute, which
29: depends on the relative velocity. A spherical colloidal particle is
30: represented by interaction sites at its surface. We demonstrate that
31: this scheme quantitatively reproduces the translational and rotational
32: diffusion of a neutral spherical particle in a liquid and show
33: preliminary results for a charged spherical particle. We argue that
34: this method is especially advantageous in the case of charged
35: colloids.
36: \end{abstract}
37:
38: %Uncomment for PACS numbers title message
39: %\pacs{00.00, 20.00, 42.10}
40:
41: % Uncomment for Submitted to journal title message
42: \submitto{\NJP}
43:
44: % Comment out if separate title page not required
45: \maketitle
46:
47: \section{Introduction}
48: \label{sec:intro}
49:
50: Understanding the dynamics of colloidal dispersions has been a
51: long-standing problem in condensed matter physics. Despite the
52: continuous need for accurate theoretical predictions concerning
53: particle mobilities or sedimentation rates, the development of the
54: field meets a nearly unpassable barrier. The many-body character of
55: the hydrodynamic interactions (HI) between the colloidal particles
56: (i.~e. their correlated motion as a result of solvent-mediated fast
57: momentum transport) poses a serious challenge for analytic theory. On
58: the other hand, the rapid growth of available computer power in the
59: last decades has made it more and more feasible to descend to a more
60: basic description of colloidal systems rather than trying to develop
61: an adequate macroscopic description. Such a step was made recently
62: for the statics of charged colloids, where simulations of the
63: primitive electrolyte model, taking the particle character of the
64: charges and fluctuations in the charge distribution fully into
65: acount, resolved some important strong correlation issues and thus
66: have resulted in a significant refinement of the existing theories of
67: screening \cite{Hansen1}. In a similar fashion, many-body effects in
68: the HI of suspensions of only moderate density are quite
69: important. Although their analytical form is, in principle, known in
70: terms of a multipole expansion \cite{Mazur}, it is numerically very
71: cumbersome to take these higher-order many-body terms into account in
72: a Brownian or Stokesian dynamics \cite{Brady} simulation, where only
73: the positions of the colloidal particles enter. For this reason, it is
74: more convenient and for large particle numbers also more efficient,
75: to take these effects into account by simulating the solvent degrees
76: of freedom, and, in particular, the momentum transport through the solvent
77: explicitly. An additional bonus is that retardation effects, if
78: present, are automatically taken into account.
79:
80: Standard Molecular Dynamics (MD), where just Newton's equations of
81: motion are integrated numerically, is of course able to simulate HI
82: correctly. However, it is very time-consuming to simulate the motion
83: of a huge number of particles explicitly in such detail. The time step
84: is governed by the local oscillations of the solvent particles in
85: their temporary ``cages'', which is much faster than the motion of the
86: colloidal particles. This information is however not needed for
87: correctly reproducing HI. One just needs an appropriate mechanism for
88: momentum transfer on somewhat larger time scales (however still small
89: compared to the motion of the colloidal particles). This can be done
90: in various ways: Dissipative Particle Dynamics (DPD) \cite{Warren} is
91: essentially MD, where however the particles are made very soft in
92: order to increase the time step, and where a momentum-conserving
93: Langevin thermostat is added \cite{Thoso}. Another possibility is multi-particle
94: collision dynamics \cite{Malevanets}, where the solvent is modeled as
95: an ideal gas, and collisions are modeled as simple stochastic updating
96: rules which conserve energy and momentum. Grid-based methods can be
97: either the direct solution of the Navier-Stokes equation by a finite
98: difference method, or the Lattice Boltzmann (LB) \cite{Succi}
99: approach, where essentially a linearized Boltzmann equation is solved
100: in a fully discretized version (i.~e. space, time, and velocities are
101: discretized).
102:
103: In our opinion, all of these approaches have their advantages and
104: disadvantages. Nevertheless, it seems that all of them are roughly
105: equivalent with respect to their computational efficiency, and with
106: respect to the level of accuracy of describing the physical
107: phenomena. Therefore, we believe that the choice of method is mainly a
108: matter of convenience.
109:
110: For our model we have chosen the LB route, mainly because this
111: algorithm is particularly simple to implement and parallelize. Moreover, we were inspired by previous highly successful simulations of hard-sphere colloidal suspensions \cite{Ladd0,Ladd1,Ladd2,Ladd3,Ladd4,Frenkel1,Frenkel2,Frenkel3,Frenkel4}. A further advantage of a grid-based method is that thermal fluctuations can be both turned on and off, depending on the physical situation under consideration (in a particle method they are always present). This approach is essentially a hybrid method, where the solvent is modeled via LB, while the colloidal particles are run with MD.
112:
113: The original approach by Ladd \cite{Ladd0,Ladd1,Ladd2,Ladd3,Ladd4}
114: models the colloidal particles as extended hollow spheres, while stick
115: boundary conditions at the surface are implemented (roughly spoken)
116: via bounce-back collision rules. One should note that for moving spheres
117: this involves some minor fluctuations
118: in the fluid mass, which is included within a sphere: The moving
119: shell incorporates some new fluid at the front, while it releases
120: mass at the rear. The fluctuations are a natural effect of the
121: thermal density fluctuations of the fluid, and of the lattice
122: discretization. Using this method, the translational and the rotational
123: dynamics of colloidal spheres were accurately described.
124:
125: This model has been recently extended to the case of charged systems
126: \cite{Horbach1}, where the colloidal particles carry a charge, while
127: the counterions and salt ions are taken into account as LB
128: populations, such that the electrostatics is essentially taken into
129: account on the level of the Poisson-Boltzmann equation. This method
130: has two disadvantages: firstly, the discrete nature of the ions and
131: correlations beyond the Poisson-Boltzmann level are not taken into
132: account; secondly, one cannot avoid a leakage of charge into (and out
133: of) the sphere (as in the case of mass), such that it is hard
134: (if not impossible) to maintain a well-defined Debye layer of charges
135: around it \cite{Horbach2}. For these reasons, it is advisable
136: to take the counterions explicitly into account. The disadvantage,
137: however, is that only rather small size ratios (size of colloidal
138: particle vs. size of ion) are easily accessible. Nevertheless, we
139: believe it is useful to study such a system, with respect to both its
140: equilibrium and its nonequilibrium properties. The purpose of the
141: present paper is to describe first steps in the development and
142: validation of the corresponding model.
143:
144: For the coupling of the small ions to the LB hydrodynamics, one would
145: like to simply model the former as point particles. Fortunately, such
146: a coupling has been recently developed by Ahlrichs and D\"unweg
147: \cite{Ahlrichs1,Ahlrichs2} with the purpose of studying the dynamics
148: of polymer solutions \cite{Ahlrichs3}. Each Brownian point particle is
149: assigned a phenomenological friction coefficient $\zeta$, and the
150: coupling is implemented just as a dissipative force $\vec F = - \zeta
151: (\vec V - \vec u)$ acting on the particle, where $\vec V$ is the
152: particle velocity, while $\vec u$ is the solvent flow velocity at the
153: particle's position, obtained via linear interpolation from the
154: surrounding lattice sites. Furthermore, the LB variables at these
155: sites are adjusted to ensure momentum balance, and thermal
156: fluctuations are added to both types of degrees of freedom. For more
157: details on implementation and validation of this method, see
158: Refs. \cite{Ahlrichs1,Ahlrichs2}.
159:
160: It is then convenient to use this coupling not only for the ions, but
161: also for the colloidal particles, whose larger size is taken into
162: account by modeling them as some arrangement of interaction sites. For
163: reasons of efficiency, we take only points at the surface of the
164: colloid. Furthermore, the coupling constant $\zeta$ is chosen rather
165: large, which approximates stick boundary conditions. It should be
166: noted that the dissipative nature of the coupling does not model any
167: squeezing-out of solvent, hence the sphere is filled with the same
168: amount of solvent regardless if it is hollow or if it were filled with
169: particles. The solvent within the sphere follows the motion of the
170: surrounding shell (with some minor time lag, see below), and therefore
171: one can view the solvent inside the sphere as just belonging to the
172: colloidal particle. Adding further particles in the sphere's volume
173: would have no effect except coupling the fluid within the sphere even
174: tighter to its motion. Within our desired level of accuracy, this
175: turned out not to be necessary. The long-time motion of the sphere is
176: the same as that of a corresponding hard sphere, as we will show in
177: the present paper.
178:
179: We analyze basic dynamic properties of such a model colloid, where we
180: restrict ourselves to the case of a single sphere. A detailed analysis
181: is done for the neutral case, while some preliminary data for the case
182: of a charged sphere are presented. We show that the (neutral) model
183: exhibits the essential dynamical features of a spherical colloidal
184: particle in a liquid. In our view, our method provides comparable
185: efficiency to Ladd's approach, while being quite straightforward to
186: implement. Furthermore, it provides substantial flexibility with
187: respect to the properties of the colloidal surface, namely, deformable,
188: permeable, and non-stick surfaces can be easily simulated.
189:
190: The remainder of this article is organized as follows: In
191: Sec. \ref{sec:model}, we describe our simulation model, while
192: Sec. \ref{sec:results} contains the numerical results on translational
193: and rotational diffusion. Finally, Sec. \ref{sec:summary} concludes
194: with a brief summary.
195:
196:
197: \section{Model}
198: \label{sec:model}
199:
200: Our hybrid simulation method involves two subsystems: the solvent that
201: is modeled via LB with fluctuating stress tensor (i.~e. we run a
202: constant-temperature version of the LB method) and a Langevin MD
203: simulation for the particles immersed in the solvent. The LB
204: simulation is performed using the 18-velocity model \cite{Ladd1},
205: using the protocol described in \cite{Ahlrichs1,Ahlrichs2}. The fluid
206: simulation consists of collision and propagation steps, the former
207: being performed with inclusion of the momentum transfer from the
208: solute particles (surface beads, and, for charged systems, ions).
209:
210: \begin{figure}
211: \begin{center}
212: \vskip 0.1in
213: \includegraphics[clip,width=0.5\textwidth]{berry3}
214: \end{center}
215: \caption{
216: Raspberry-like model of a colloidal sphere. There is a central large
217: bead of radius $R=3$ and charge $Z=10$. The small beads of radius 1
218: are connected with their nearest neighbors on the surface via FENE
219: bonds. A repulsive soft-core potential is also operating between all
220: the monomers. The counterions are moving freely in space and interact
221: with the central bead via the Coulomb potential and with the surface
222: beads via the repulsive LJ potential.
223: \href{http://www.mpip-mainz.mpg.de/~lobaskin/TR6/Z20.mpg}{Movie}
224: }
225: \label{fig:sphere}
226: \end{figure}
227:
228: The colloidal particle is represented by a two-dimensional tethered
229: bead-spring network consisting of $100$ beads, which is wrapped around
230: a ball of a radius $\sigma_{cs}$ (for notation, see below), so that
231: the whole construction resembles a raspberry (see
232: Fig. \ref{fig:sphere}). The network connectivity is maintained via
233: finitely extendable nonlinear elastic (FENE) springs,
234: %
235: \begin{equation}
236: V_{FENE} (r) = - \frac{k R_0^2}{2} \ln
237: \left(1 - \left( \frac{r}{R_0}\right)^{2} \right) ,
238: \label{eq:fene}
239: \end{equation}
240: %
241: where $k$ is the spring constant, and $R_0$ the maximum bond
242: extension. Furthermore, the beads repel each other by a modified
243: Lennard-Jones (LJ) potential
244: %
245: \begin{equation}
246: V_{LJ} (r) = \left\lbrace \begin{array} {ll}
247: 4\epsilon_{ij} \left( \left( \frac{\sigma_{ij}}{r} \right)^{12}
248: - \left( \frac{\sigma_{ij}}{r} \right)^{6} +
249: \frac{1}{4} \right) & r<2^{1/6} \sigma_{ij} \\
250: 0 & r\ge 2^{1/6} \sigma_{ij} . \end{array} \right.
251: \label{eq:lj}
252: \end{equation}
253: %
254: An additional repulsive LJ bead is introduced at the center of the
255: sphere in order to maintain its shape. In Eq. \ref{eq:lj}, $i,j$
256: denote either a central (``c'') or a surface (``s'') bead. The unit
257: system is completely defined by the surface bead parameters by setting
258: $\epsilon_{ss}$, $\sigma_{ss}$, and the surface bead mass $m_s$ to
259: unity. The interaction between the central bead and the surface beads
260: is described by $\sigma_{cs} = 3$, which is thus the sphere radius,
261: and $\epsilon_{cs}=8$. Furthermore, the FENE spring constant for the
262: surface beads is $k=300$ and the maximum bond extension is $R_0=
263: 1.25$. To simulate a charged colloidal particle, we place the charge
264: at the central bead, and add an appropriate number of counterions (LJ
265: beads with ``s'' properties) outside the sphere. The electrostatic
266: interaction is taken into account via the Coulomb potential
267: %
268: \begin{equation}
269: V_{el} (r) = \lambda_B k_B T \frac{q_i q_j}{r}
270: \label{eq:electrostatics}
271: \end{equation}
272: %
273: between the various charges, where the standard Ewald summation
274: technique \cite{Allen} is applied. In Eq. \ref{eq:electrostatics},
275: $\lambda_B=e^2/ \left(4\pi\varepsilon_0 \varepsilon_r k_B T\right)$ is
276: the Bjerrum length, $k_B$ the Boltzmann constant, $q_i$ the charge of
277: species $i$ in units of the elementary charge $e$, and $T$ the
278: temperature.
279:
280: The LB lattice constant is chosen as one (in our LJ unit system), and
281: the fluid is simulated in a cubic box with periodic boundary
282: conditions. The force between the LB fluid and the surface beads
283: (or ions) is given by
284: %
285: \begin{equation}
286: \vec{F}= - \zeta \left( \vec{V} - \vec{u} \right) + \vec{f} .
287: \label{eq:4}
288: \end{equation}
289: %
290: Here, $\zeta$ is the ``bare'' \cite{Ahlrichs2} friction coefficient,
291: $\vec{V}$ and $\vec{u}$ are the velocities of the bead and the fluid
292: (at the position of the bead), respectively, while $\vec{f}$ is a
293: Gaussian white noise force with zero mean, whose strength is given via
294: the standard fluctuation-dissipation theorem
295: \cite{Ahlrichs1,Ahlrichs2} to keep the surface beads and ions at
296: the same temperature as the solvent. The central bead is not
297: coupled to the solvent (here $\zeta = \vec f = 0$); as discussed in the
298: Introduction, the behavior of the model would change only marginally
299: if such a coupling were included. In our simulation we used a friction
300: constant $\zeta = 20$, a temperature $k_B T = 1$, a fluid mass density
301: $\rho = 0.85$, and a kinematic viscosity $\nu=3$, resulting in a
302: dynamic viscosity $\eta=2.55$. At least 20000 MD steps were performed
303: to equilibrate the initial random bead configuration before the
304: interaction with the LB solvent was turned on. Further details on the
305: method can be found in \cite{Ahlrichs1,Ahlrichs2}.
306:
307:
308: \section{Results}
309: \label{sec:results}
310:
311: We test the simulation method against basic relations for an isolated
312: sphere in solvent. First, we look at the center of mass' velocity
313: relaxation. The simplest experiment to perform is a ``kick''. The
314: sphere is placed in a LB fluid at rest (i.~e. without thermal noise),
315: and at time $t = 0$ all particles of the sphere are assigned an
316: identical velocity $\vec V = 1$ in $x$ direction. Fig. \ref{fig:vcf2}
317: monitors the time behavior of the sphere's center of mass velocity,
318: normalized by the initial value. According to linear response theory,
319: this relaxation function must be identical to the normalized
320: center-of-mass velocity autocorrelation function for Brownian motion
321: in thermal equilibrium, if the initial kick is weak enough. This is
322: indeed satisfied, as a comparison of the two curves in
323: Fig. \ref{fig:vcf2} shows. For the experiment in thermal equilibrium,
324: we performed 10 runs with different random number generator
325: initializations, in order to reduce the statistical uncertainty.
326:
327: It is well-known that simulations of Brownian motion in a hydrodynamic
328: solvent are always strongly affected by finite size effects. The
329: diffusion constant, and therefore also the relaxation function, depend
330: on the linear system size $L$ due to hydrodynamic interactions with
331: the periodic images. The diffusion constant exhibits a finite-size
332: correction of order $R / L$ \cite{Ahlrichs2,Burkhard}. Asymptotic
333: behavior can therefore only be expected for $R / L \ll 1$, and this is
334: why we performed the experiment in a rather large box of size $L =
335: 80$. For the same reason, the equivalence between ``kick'' experiment
336: and Brownian motion will only hold if the comparison is done for the
337: same box sizes.
338:
339: \begin{figure}
340: \vskip 0.1in
341: \begin{center}
342: \includegraphics[clip, width=0.75\textwidth]{vcf2a.eps}
343: \end{center}
344: \caption{
345: Normalized translational velocity of the center of mass of the
346: colloidal sphere in a "kick" experiment at $k_B T = 0$, and its
347: normalized velocity autocorrelation function for Brownian motion at
348: $k_B T = 1$. The dotted curve shows the exponential decay derived from
349: the Stokes law and the dashed curve the expected long-time asymptotic
350: behavior.}
351: \label{fig:vcf2}
352: \end{figure}
353:
354: \begin{figure}
355: \begin{center}
356: \vskip 0.1in
357: \includegraphics[clip, width=0.75\textwidth]{msd}
358: \end{center}
359: \caption{
360: Mean-square displacement of the center of mass of a neutral sphere at
361: $k_B T = 1$ and different simulation box sizes. }
362: \label{fig:msd}
363: \end{figure}
364:
365: One clearly sees that the initial decay of the relaxation function is
366: characterized by two relaxation processes, one initial fast decay
367: followed by a somewhat slower relaxation. Qualitatively, this may be
368: explained as follows: A compact sphere of radius $R$ and mass $M$
369: should exhibit a velocity relaxation which, after transient ballistic
370: motion, is initially characterized by an exponential decay $\exp
371: \left( - t / \tau \right)$, with a relaxation time $\tau = M /
372: \zeta_{tot}$. Here, $\zeta_{tot}$ is the total friction
373: coefficient, which we estimate via Stokes' law for stick boundary
374: conditions as $\zeta_{tot} \approx 6 \pi \eta R \approx 144$. However,
375: the effective mass is time-dependent: While initially only the mass of
376: the beads $M = 101$ contributes, at later times the fluid within the
377: sphere is dragged as well, such that then the mass is roughly
378: estimated as $M \approx 101 + 4 \pi \rho R^3 / 3 \approx 214$. This
379: gives rise to the initial and final relaxation times $\tau_{in}
380: \approx 0.7$ and $\tau_{fin} \approx 1.5$. However, the initial
381: relaxation time cannot be observed, since in the extreme short-time
382: regime ballistic effects play a role. Conversely, the decay with
383: $\tau_{fin}$ is clearly visible (see Fig. \ref{fig:vcf2}).
384:
385: \begin{figure}
386: \begin{center}
387: \vskip 0.1in
388: \includegraphics[clip, width=0.75\textwidth]{d1}
389: \end{center}
390: \caption{
391: The long-time self-diffusion coefficient and the rotational diffusion
392: coefficient of the colloidal sphere (normalized by the asymptotic
393: Stokes-Einstein and Stokes-Einstein-Debye values) as functions of the inverse size of the primary simulation cell. }
394: \label{fig:ds}
395: \end{figure}
396:
397: After $t \approx 1$, the famous long-time tail \cite{Alder,Hansen2}
398: (normalized relaxation function $V(t) / V(0) = B t^{-3/2}$) sets
399: in. The physical mechanism of this slow relaxation is the fact that
400: the momentum is conserved and hence transported away diffusively from
401: the particle \cite{Hansen2}. For this reason, it can only be observed
402: in a sufficiently large box. The prefactor of the power law is known
403: in the colloidal limit, $B = (1 / 12) (N m / \rho) (\pi \nu)^{-3/2}$,
404: where $N$ is the number of beads, and $m$ the bead mass
405: \cite{Hinch,Cichocki}. It should be noted that the prefactor of the
406: power-law decay of the {\em unnormalized} velocity autocorrelation
407: function does not depend on the properties of the sphere at all, but
408: only on the temperature, and the hydrodynamic properties of the
409: solvent \cite{Hinch,Cichocki}. The mass dependence of $B$ results only
410: from the normalization, i.~e. the value of $\left< V^2 \right>$
411: according to the equipartition theorem, $\left< V^2 \right> = 3 k_B T
412: / (N m)$. From this consideration it is clear that the {\em
413: short-time} value of the sphere's mass ($N m$) enters (at $t = 0$
414: the shell and the inner fluid are not yet coupled). For our
415: simulation parameters, we find $B = 0.342$. As Fig. \ref{fig:vcf2}
416: shows, the data exhibit the expected behavior very nicely.
417:
418: Figure \ref{fig:msd} illustrates the finite size effect in the diffusive
419: properties by plotting the mean square displacement of the colloidal
420: sphere as a function of time, for two different box sizes $L = 6.33 R$
421: and $L = 20 R$, for Brownian motion at $k_B T = 1$. In the long-time
422: regime, the slope is given by $6 D^S$, where $D^S$ is the
423: self-diffusion coefficient. The figure clearly shows that $D^S$
424: increases with box size. This can be explained in terms of
425: hydrodynamic interactions with the periodic images, or, equivalently,
426: in terms of suppression of long-wavelength hydrodynamic modes
427: \cite{Ahlrichs2,Burkhard}. We have therefore performed a systematic
428: finite-size analysis, and measured $D^S$ for various box sizes by
429: integrating the velocity autocorrelation function over time
430: \cite{Hansen2}. In Fig. \ref{fig:ds} we plot $D^S / D_0$
431: ($D_0$ denoting the asymptotic Stokes-Einstein value $D_0 = k_B T / (6
432: \pi \eta R) = 6.9 \times 10^{-3}$) as a function of $1 / L$; the
433: expected linear behavior \cite{Burkhard} is clearly seen.
434:
435: \begin{figure}
436: \begin{center}
437: \vskip 0.1in
438: \includegraphics[clip, width=0.75\textwidth]{rotnew}
439: \end{center}
440: \caption{
441: Decay of angular velocity of the colloidal sphere. The different
442: curves are marked by the primary simulation box sizes. The solid
443: curves show the expected exponential decay according to the Debye law
444: and the long-time asymptotic behavior.}
445: \label{fig:avcf}
446: \end{figure}
447:
448: We now look at the relaxation of the rotational motion. We performed a
449: similar kick experiment as described above, where now an initial
450: angular velocity $\omega_0 = 1$ was provided to the sphere. The data
451: for the normalized decay function $\omega(t) / \omega (0)$ are
452: presented in Fig. \ref{fig:avcf}. The sphere dynamics shows a
453: characteristic ``raw egg'' (damped) rotation pattern with a fast
454: initial decay and a subsequent slower one. For the rotational
455: relaxation of a hard sphere with moment of inertia $I$ and rotational
456: friction coefficient $8 \pi \eta R^3$, theory predicts a decay
457: according to the Debye law $\omega (t) / \omega (0) =
458: \exp \left( - t / \tau_r \right)$, where the relaxation time is given
459: by $\tau_r = I / (8 \pi \eta R^3)$ \cite{Debye}. Similarly to the
460: effective mass, the effective moment of inertia is expected to
461: increase as a function of time, due to the time-delayed dragging of
462: the fluid. Initially we expect a hollow-sphere value of roughly
463: $I_{in} \approx (2/3) M R^2 \approx 600$, where we used the mass of
464: the outer shell $M = 100$. A more accurate value is obtained by
465: direct summation over the contributing beads, yielding $I_{in} = 546$.
466: In the late-time regime the moment of
467: inertia is expected to be the compact-sphere value $I_{fin} \approx
468: (2/5) M R^2 \approx 770$, where the total mass $M \approx 214$ has
469: been used. This results in relaxation times $\tau_{r,in} = 0.34$ and
470: $\tau_{r,fin} = 0.44$. These values are roughly consistent with a fit
471: to the data in the interval $0.1 < t < 1.0$, which yields a somewhat
472: larger relaxation time $\tau_r = 0.68$ (see Fig. \ref{fig:avcf}).
473: Given the general inaccuracy of the
474: estimates, and the crossover of the Debye relaxation function both to
475: short-time ballistic behavior, and long-time hydrodynamic behavior,
476: this deviation is not too surprising.
477:
478: Similar to the translational motion, the exponential decay is then
479: followed by a power-law long-time tail. Theory predicts $\omega (t) /
480: \omega (0) = (\pi I / \rho) (4 \pi \nu t)^{-5/2}$ \cite{Hauge}. As we
481: have seen before, the long-time tail in the translational case is
482: governed by the short-time mass as a result of normalization.
483: Similarly, the rotational tail must be controlled by the {\em
484: short-time} moment of inertia $I_{in}$, which also governs the mean
485: square fluctuations of the angular velocity via the equipartition
486: theorem $\left< \omega^2 \right> = 3 k_B T / I_{in}$. For this reason,
487: we can determine $I_{in}$ by fitting a $t^{-5/2}$ law to the data,
488: which is much more accurate than our rough geometrical estimate. The
489: fit results in $I_{in} = 533$, which is reasonably consistent. If we
490: insert this value into the relaxation time expression, we obtain
491: $\tau_r = 0.64$, in quite good agreement with the data. It hence seems
492: that fluid dragging effects are not yet very important for the initial
493: Debye relaxation.
494:
495: At longer times, one can again notice a significant finite-size
496: effect: the curves obtained in the smaller simulation box depart from
497: the asymptotic power-law line earlier, i.~e. the long-time diffusion
498: is hindered at the small system sizes. In fact, the power-law regime
499: is inaccessible at $L = 10$ while for $L = 100$ it extends up to $t =
500: 200$. This value characterizes the interval after which the particle
501: starts feeling its own periodic images. The rotational diffusion
502: constant $D^R$ is given by the Green-Kubo integral \cite{Berne1}
503: %
504: \begin{equation}
505: D^R = \frac{1}{3} \int_0^\infty dt\, \left< \vec \omega (t)
506: \cdot \vec \omega (0) \right> .
507: \end{equation}
508: %
509: We can evaluate this by again making use of linear response
510: theory: The correlation function $\left< \vec \omega (t)
511: \cdot \vec \omega (0) \right>$ is identical to the relaxation
512: function presented in Fig. \ref{fig:avcf}, multiplied with the initial
513: value $\left< \omega^2 \right>$, which we know from the equipartition
514: theorem (see above). Using this approach, we have calculated $D^R$ for
515: different box sizes $L$. In an infinite fluid, $D^R$ has the
516: Stokes-Einstein-Debye value $D^R = k_B T / (8 \pi \eta R^3) = 0.58 \times
517: 10^{-3}$, towards which the results indeed converge for $L \to
518: \infty$. Again a $1 / L$ behavior (but weaker than for translational
519: diffusion) is observed, as shown in Fig. \ref{fig:ds}. It should be
520: noted that the accuracy of the data is slightly hampered by the fact
521: that we had to use a somewhat arbitrary criterion for cutting off the
522: integral function, which does not arrive at a constant value but
523: rather ends up with a linear increase originating from a small nonzero
524: constant angular velocity in the long-time limit (as a manifestation
525: of the finite-size effect).
526:
527: \begin{figure}
528: \begin{center}
529: \vskip 0.1in
530: \includegraphics[clip, width=0.75\textwidth]{vcfzz}
531: \end{center}
532: \caption{
533: Center of mass velocity autocorrelation functions for a neutral and
534: a charged colloidal sphere. The Green-Kubo integral function
535: is also shown as solid curve for $Z=0$ and dashed curve for $Z=10$.
536: }
537:
538: \label{fig:zz}
539: \end{figure}
540:
541: Finally, we have started to study the effect of charge on the
542: self-diffusion of the colloidal particle. We performed simulations of
543: a sphere with central charge $Z_C = 10$ in an LB box of size $L =
544: 40$. Ten counterions of charge $-1$ were also added. The Bjerrum
545: length was set to $2$. Technically, the simulation ran without
546: any problems just as well as for the neutral system. In Fig.
547: \ref{fig:zz} we compare the decay of the velocity
548: autocorrelation function for the neutral and charged spheres. The
549: difference between the two is not detectable within our error bars at
550: this charge. We cannot say yet much about the influence of the charges
551: on the dynamic properties; we do however expect that for strongly
552: charged systems the Debye layer will effectively increase the
553: hydrodynamic radius and slow down the diffusion. More work needs to be
554: done to resolve this issue.
555:
556: \section{Summary}
557: \label{sec:summary}
558:
559: In summary, we introduced and tested a new model for simulating
560: colloidal dynamics with inclusion of the hydrodynamic effects on the
561: level of the lattice-Boltzmann equation. The suggested
562: ``raspberry''-like colloidal object exhibits the essential features of
563: the diffusion of a spherical colloidal particle. For the first time we
564: combined a model containing explicit charges with accurate treatment
565: of the hydrodynamics. We expect this method to have some advantages
566: over the previously applied scheme \cite{Horbach1} which builds upon
567: the model of Ladd \cite{Ladd0} and has the problem of charge
568: leakage. In our method: (i) the stick boundary conditions are realized
569: not strictly, i.~e. the particle surface is softly bound to the fluid;
570: (ii) the method is easily adaptable to the multiple timestep method
571: where the ionic action on the colloid can be averaged out somewhat
572: within the integration step for the solvent. We should note that such
573: a hybrid simulation requires still some computational effort --- two
574: million hybrid steps for the colloid consisting of 100 beads in a box of
575: $20 \times 20 \times 20$ LB cells take about 12 hours on a 2 GHz Pentium
576: 4 processor, while it is expected to run a few times faster on more
577: sophisticated hardware. Here, one step is a MD step of the total system
578: of beads, plus one LB update of the whole lattice, where the latter part
579: completely dominates the CPU effort. Some improvement is possible by not updating the LB fluid at every MD step, as it was done here. For dense
580: systems, where the MD part is no longer negligible, one should also
581: optimize the interaction potentials. Finally, we would like to mention
582: that the algorithm can be fully parallelized and we are working on the
583: parallel version.
584:
585: \ack
586: We thank Christian Holm, J\"urgen Horbach and Kurt Kremer
587: for stimulating discussions, and the latter also for critical
588: reading of the manuscript. This work was funded by the SFB
589: TR 6 of the Deutsche Forschungsgemeinschaft.
590:
591:
592: \section{References}
593: \begin{thebibliography}{15}
594: \bibitem{Hansen1}
595: Hansen J-P and L\"owen H 2000 {\it Ann. Rev. Phys. Chem.} {\bf 51} 209
596: \bibitem{Mazur}
597: Mazur P and Van Saarloos W 1982 {\it Physica A} {\bf 115} 21
598: \bibitem{Brady}
599: Brady J F and Bossis G 1998 {\it Ann. Rev. Fluid Mech.} {\bf 20} 111
600: \bibitem{Warren}
601: Groot R and Warren P 1997 {\it J. Chem. Phys.} {\bf 107} 4423
602: \bibitem{Thoso}
603: Soddemann T, D\"unweg B, and Kremer K 2003 {\it Phys. Rev. E} in press
604: \bibitem{Malevanets}
605: Malevanets A and Kapral R 1999 {\it J. Chem. Phys.} {\bf 110} 8605
606: \bibitem{Succi}
607: Succi S 2001 {\it The Lattice Boltzmann Equation for Fluid
608: Dynamics and Beyond} (Oxford: Oxford University Press)
609: \bibitem{Ladd0}
610: Ladd AJC 1993 {\it Phys. Rev. Lett.} {\bf 70} 1339
611: \bibitem{Ladd1}
612: Ladd AJC 1994 {\it J. Fluid Mech.} {\bf 271} 285
613: \bibitem{Ladd2}
614: Ladd AJC 1994 {\it J. Fluid Mech.} {\bf 271} 311
615: \bibitem{Ladd3}
616: Ladd AJC and Verberg R 2001 {\it J. Stat. Phys.} {\bf 104} 1191
617: \bibitem{Ladd4}
618: Ladd AJC, Hu Gang, Zhu JX, and Weitz DA 1995 {\it Phys. Rev. E} {\bf 52} 6550
619: \bibitem{Frenkel1}
620: Hagen MHJ, Frenkel D, and Lowe CP 1998 {\it Physica A} {\bf 109} 275
621: \bibitem{Frenkel2}
622: Lowe CP, Frenkel D, and Masters AJ 1995 {\it J. Chem. Phys.} {\bf 103}
623: 1582
624: \bibitem{Frenkel3}
625: Heemels MW, Hagen MHJ, and Lowe CP 2000 {\it J. Comput. Phys}
626: {\bf 164} 48
627: \bibitem{Frenkel4}
628: Hagen MHJ, Pagonabarraga I, Lowe CP, and Frenkel D 1997 {\it Phys.
629: Rev. Lett.} {\bf 78} 3785
630: \bibitem{Horbach1}
631: Horbach J and Frenkel D 2001 {\it Phys. Rev. E} {\bf 64} 061507-1
632: \bibitem{Horbach2}
633: Horbach J {Private communication}
634: \bibitem{Ahlrichs1}
635: Ahlrichs P and D\"unweg B 1998 {\it Int. J. Mod. Phys. C} {\bf 9} 1429
636: \bibitem{Ahlrichs2}
637: Ahlrichs P and D\"unweg B 1999 {\it J. Chem. Phys.} {\bf 111} 8225
638: \bibitem{Ahlrichs3}
639: Ahlrichs P, Everaers R, and D\"unweg B 2001 {\it Phys. Rev. E} {\bf 64}
640: 040501(R)
641: \bibitem{Allen}
642: Allen M and Tildesley DJ 1987 {\it Computer simulation of liquids}
643: (Oxford: Oxford University Press)
644: \bibitem{Burkhard}
645: D\"unweg B and Kremer K 1993 {\it J. Chem. Phys.} {\bf 99} 6983
646: \bibitem{Alder}
647: Alder B and Wainwright T 1970 {\it Phys. Rev. A} {\bf 1} 18
648: \bibitem{Hansen2}
649: Hansen J-P and McDonald I 1986
650: {\it Theory of Simple Liquids} (London: Academic Press)
651: \bibitem{Hinch}
652: Hinch E J 1975 {\it J. Fluid Mech.} {\bf 72} 499
653: \bibitem{Cichocki}
654: Chichocki B and Felderhof B 1995 {\it Phys. Rev. E} {\bf 51} 5549
655: \bibitem{Debye}
656: Debye P 1929 {\it Polar Molecules} (New York: Dover)
657: \bibitem{Hauge}
658: Hauge EH and Martin-L\"of A 1973 {\it J. Stat. Phys.} {\bf 7} 259
659: \bibitem{Berne1}
660: Kushick J and Berne BJ 1973 {\it J. Chem. Phys.} {\bf 59} 4486
661: \end{thebibliography}
662:
663: \end{document}
664: