ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/xDissertation/ld.tex
(Generate patch)

Comparing trunk/xDissertation/ld.tex (file contents):
Revision 3336 by xsun, Wed Jan 30 16:01:02 2008 UTC vs.
Revision 3383 by xsun, Wed Apr 16 21:56:34 2008 UTC

# Line 1 | Line 1
1   \chapter{\label{chap:ld}LANGEVIN DYNAMICS}
2 +
3 + Recent examples of the usefulness of Langevin simulations include a
4 + study of met-enkephalin in which Langevin simulations predicted
5 + dynamical properties that were large\-ly in agreement with explicit
6 + solvent simulations.\cite{Shen2002} By applying Langevin dynamics with
7 + the UNRES model, Liwo and his coworkers suggest that protein folding
8 + pathways can be explored within a reasonable amount of
9 + time.\cite{Liwo2005}
10 +
11 + The stochastic nature of Langevin dynamics also enhances the sampling
12 + of the system and increases the probability of crossing energy
13 + barriers.\cite{Cui2003,Banerjee2004} Combining Langevin dynamics with
14 + Kramers' theory, Klimov and Thirumalai identified free-energy
15 + barriers by studying the viscosity dependence of the protein folding
16 + rates.\cite{Klimov1997} In order to account for solvent induced
17 + interactions missing from the implicit solvent model, Kaya
18 + incorporated a desolvation free energy barrier into protein
19 + folding/unfolding studies and discovered a higher free energy barrier
20 + between the native and denatured states.\cite{HuseyinKaya07012005}
21 +
22 + In typical LD simulations, the friction and random ($f_r$) forces on
23 + individual atoms are taken from Stokes' law,
24 + \begin{eqnarray}
25 + m \dot{v}(t) & = & -\nabla U(x) - \xi m v(t) + f_r(t) \notag \\
26 + \langle f_r(t) \rangle & = & 0 \\
27 + \langle f_r(t) f_r(t') \rangle & = & 2 k_B T \xi m \delta(t - t') \notag
28 + \end{eqnarray}
29 + where $\xi \approx 6 \pi \eta \rho$.  Here $\eta$ is the viscosity of the
30 + implicit solvent, and $\rho$ is the hydrodynamic radius of the atom.
31 +
32 + The use of rigid substructures,\cite{Chun:2000fj}
33 + coarse-graining,\cite{Ayton01,Golubkov06,Orlandi:2006fk,SunX._jp0762020}
34 + and ellipsoidal representations of protein side
35 + chains~\cite{Fogolari:1996lr} has made the use of the Stokes-Einstein
36 + approximation problematic.  A rigid substructure moves as a single
37 + unit with orientational as well as translational degrees of freedom.
38 + This requires a more general treatment of the hydrodynamics than the
39 + spherical approximation provides.  Also, the atoms involved in a rigid
40 + or coarse-grained structure have solvent-mediated interactions with
41 + each other, and these interactions are ignored if all atoms are
42 + treated as separate spherical particles.  The theory of interactions
43 + {\it between} bodies moving through a fluid has been developed over
44 + the past century and has been applied to simulations of Brownian
45 + motion.\cite{FIXMAN:1986lr,Ramachandran1996}
46 +
47 + In order to account for the diffusion anisotropy of complex shapes,
48 + Fernandes and Garc\'{i}a de la Torre improved an earlier Brownian
49 + dynamics simulation algorithm~\cite{Ermak1978,Allison1991} by
50 + incorporating a generalized $6\times6$ diffusion tensor and
51 + introducing a rotational evolution scheme consisting of three
52 + consecutive rotations.\cite{Fernandes2002} Unfortunately, biases are
53 + introduced into the system due to the arbitrary order of applying the
54 + noncommuting rotation operators.\cite{Beard2003} Based on the
55 + observation the momentum relaxation time is much less than the time
56 + step, one may ignore the inertia in Brownian dynamics.  However, the
57 + assumption of zero average acceleration is not always true for
58 + cooperative motion which is common in proteins. An inertial Brownian
59 + dynamics (IBD) was proposed to address this issue by adding an
60 + inertial correction term.\cite{Beard2000} As a complement to IBD,
61 + which has a lower bound in time step because of the inertial
62 + relaxation time, long-time-step inertial dynamics (LTID) can be used
63 + to investigate the inertial behavior of linked polymer segments in a
64 + low friction regime.\cite{Beard2000} LTID can also deal with the
65 + rotational dynamics for nonskew bodies without translation-rotation
66 + coupling by separating the translation and rotation motion and taking
67 + advantage of the analytical solution of hydrodynamic
68 + properties. However, typical nonskew bodies like cylinders and
69 + ellipsoids are inadequate to represent most complex macromolecular
70 + assemblies. Therefore, the goal of this work is to adapt some of the
71 + hydrodynamic methodologies developed to treat Brownian motion of
72 + complex assemblies into a Langevin integrator for rigid bodies with
73 + arbitrary shapes.
74 +
75 + \subsection{Rigid Body Dynamics}
76 + Rigid bodies are frequently involved in the modeling of large
77 + collections of particles that move as a single unit.  In molecular
78 + simulations, rigid bodies have been used to simplify protein-protein
79 + docking,\cite{Gray2003} and lipid bilayer
80 + simulations.\cite{SunX._jp0762020} Many of the water models in common
81 + use are also rigid-body
82 + models,\cite{Jorgensen83,Berendsen81,Berendsen87} although they are
83 + typically evolved in molecular dynamics simulations using constraints
84 + rather than rigid body equations of motion.
85 +
86 + Euler angles are a natural choice to describe the rotational degrees
87 + of freedom.  However, due to $\frac{1}{\sin \theta}$ singularities, the
88 + numerical integration of corresponding equations of these motion can
89 + become inaccurate (and inefficient).  Although the use of multiple
90 + sets of Euler angles can overcome this problem,\cite{Barojas1973} the
91 + computational penalty and the loss of angular momentum conservation
92 + remain.  A singularity-free representation utilizing quaternions was
93 + developed by Evans in 1977.\cite{Evans1977} The Evans quaternion
94 + approach uses a nonseparable Hamiltonian, and this has prevented
95 + symplectic algorithms from being utilized until very
96 + recently.\cite{Miller2002}
97 +
98 + Another approach is the application of holonomic constraints to the
99 + atoms belonging to the rigid body.  Each atom moves independently
100 + under the normal forces deriving from potential energy and constraints
101 + are used to guarantee rigidity. However, due to their iterative
102 + nature, the SHAKE and RATTLE algorithms converge very slowly when the
103 + number of constraints (and the number of particles that belong to the
104 + rigid body) increases.\cite{Ryckaert1977,Andersen1983}
105 +
106 + In order to develop a stable and efficient integration scheme that
107 + preserves most constants of the motion in microcanonical simulations,
108 + symplectic propagators are necessary.  By introducing a conjugate
109 + momentum to the rotation matrix ${\bf Q}$ and re-formulating
110 + Hamilton's equations, a symplectic orientational integrator,
111 + RSHAKE,\cite{Kol1997} was proposed to evolve rigid bodies on a
112 + constraint manifold by iteratively satisfying the orthogonality
113 + constraint ${\bf Q}^T {\bf Q} = 1$.  An alternative method using the
114 + quaternion representation was developed by Omelyan.\cite{Omelyan1998}
115 + However, both of these methods are iterative and suffer from some
116 + related inefficiencies. A symplectic Lie-Poisson integrator for rigid
117 + bodies developed by Dullweber {\it et al.}\cite{Dullweber1997} removes
118 + most of the limitations mentioned above and is therefore the basis for
119 + our Langevin integrator.
120 +
121 + The goal of the present work is to develop a Langevin dynamics
122 + algorithm for ar\-bi\-trary-shaped rigid particles by integrating an
123 + accurate estimate of the friction tensor from hydrodynamics theory
124 + into a stable and efficient rigid body dynamics propagator.  In the
125 + sections below, we review some of the theory of hydrodynamic tensors
126 + developed primarily for Brownian simulations of multi-particle
127 + systems, we then present our integration method for a set of
128 + generalized Langevin equations of motion, and we compare the behavior
129 + of the new Langevin integrator to dynamical quantities obtained via
130 + explicit solvent molecular dynamics.
131 +
132 + \subsection{\label{ldintroSection:frictionTensor}The Friction Tensor}
133 + Theoretically, a complete friction kernel for a solute particle can be
134 + determined using the velocity autocorrelation function from a
135 + simulation with explicit solvent molecules. However, this approach
136 + becomes impractical when the solute becomes complex.  Instead, various
137 + approaches based on hydrodynamics have been developed to calculate
138 + static friction coefficients. In general, the friction tensor $\Xi$ is
139 + a $6\times 6$ matrix given by
140 + \begin{equation}
141 + \Xi  = \left( \begin{array}{*{20}c}
142 +   \Xi^{tt} & \Xi^{rt}  \\
143 +   \Xi^{tr} & \Xi^{rr}  \\
144 + \end{array} \right).
145 + \end{equation}
146 + Here, $\Xi^{tt}$ and $\Xi^{rr}$ are $3 \times 3$ translational and
147 + rotational resistance (friction) tensors respectively, while
148 + $\Xi^{tr}$ is translation-rotation coupling tensor and $\Xi^{rt}$ is
149 + rotation-translation coupling tensor. When a particle moves in a
150 + fluid, it may experience a friction force ($\mathbf{f}_f$) and torque
151 + ($\mathbf{\tau}_f$) in opposition to the velocity ($\mathbf{v}$) and
152 + body-fixed angular velocity ($\mathbf{\omega}$),
153 + \begin{equation}
154 + \left( \begin{array}{l}
155 + \mathbf{f}_f  \\
156 + \mathbf{\tau}_f  \\
157 + \end{array} \right) =  - \left( \begin{array}{*{20}c}
158 +   \Xi^{tt} & \Xi^{rt}  \\
159 +   \Xi^{tr} & \Xi^{rr}  \\
160 + \end{array} \right)\left( \begin{array}{l}
161 + \mathbf{v} \\
162 + \mathbf{\omega} \\
163 + \end{array} \right).
164 + \end{equation}
165 + For an arbitrary body moving in a fluid, Peters has derived a set of
166 + fluctuation-dissipation relations for the friction
167 + tensors,\cite{Peters:1999qy,Peters:1999uq,Peters:2000fk}
168 + \begin{eqnarray}
169 + \Xi^{tt} & = & \frac{1}{k_B T} \int_0^\infty \left[ \langle {\bf
170 + F}(0) {\bf F}(-s) \rangle_{eq} - \langle {\bf F} \rangle_{eq}^2
171 + \right] ds \\
172 + \notag \\
173 + \Xi^{tr} & = & \frac{1}{k_B T} \int_0^\infty \left[ \langle {\bf
174 + F}(0) {\bf \tau}(-s) \rangle_{eq} - \langle {\bf F} \rangle_{eq}
175 + \langle {\bf \tau} \rangle_{eq} \right] ds \\
176 + \notag \\
177 + \Xi^{rt} & = & \frac{1}{k_B T} \int_0^\infty \left[ \langle {\bf
178 + \tau}(0) {\bf F}(-s) \rangle_{eq} - \langle {\bf \tau} \rangle_{eq}
179 + \langle {\bf F} \rangle_{eq} \right] ds \\
180 + \notag \\
181 + \Xi^{rr} & = & \frac{1}{k_B T} \int_0^\infty \left[ \langle {\bf
182 + \tau}(0) {\bf \tau}(-s) \rangle_{eq} - \langle {\bf \tau} \rangle_{eq}^2
183 + \right] ds
184 + \end{eqnarray}
185 + In these expressions, the forces (${\bf F}$) and torques (${\bf
186 + \tau}$) are those that arise solely from the interactions of the body with
187 + the surrounding fluid. For a single solute body in an isotropic fluid,
188 + the average forces and torques in these expressions ($\langle {\bf F}
189 + \rangle_{eq}$ and $\langle {\bf \tau} \rangle_{eq}$)
190 + vanish, and one obtains the simpler force-torque correlation formulae
191 + of Nienhuis.\cite{Nienhuis:1970lr} Molecular dynamics simulations with
192 + explicit solvent molecules can be used to obtain estimates of the
193 + friction tensors with these formulae. In practice, however, one needs
194 + relatively long simulations with frequently-stored force and torque
195 + information to compute friction tensors, and this becomes
196 + prohibitively expensive when there are large numbers of large solute
197 + particles.  For bodies with simple shapes, there are a number of
198 + approximate expressions that allow computation of these tensors
199 + without the need for expensive simulations that utilize explicit
200 + solvent particles.
201 +
202 + \subsubsection{\label{ldintroSection:resistanceTensorRegular}\textbf{The Resistance Tensor for Regular Shapes}}
203 + For a spherical body under ``stick'' boundary conditions, the
204 + translational and rotational friction tensors can be estimated from
205 + Stokes' law,
206 + \begin{equation}
207 + \label{ldeq:StokesTranslation}
208 + \Xi^{tt}  = \left( \begin{array}{*{20}c}
209 +   {6\pi \eta \rho} & 0 & 0  \\
210 +   0 & {6\pi \eta \rho} & 0  \\
211 +   0 & 0 & {6\pi \eta \rho}  \\
212 + \end{array} \right)
213 + \end{equation}
214 + and
215 + \begin{equation}
216 + \label{ldeq:StokesRotation}
217 + \Xi^{rr}  = \left( \begin{array}{*{20}c}
218 +   {8\pi \eta \rho^3 } & 0 & 0  \\
219 +   0 & {8\pi \eta \rho^3 } & 0  \\
220 +   0 & 0 & {8\pi \eta \rho^3 }  \\
221 + \end{array} \right)
222 + \end{equation}
223 + where $\eta$ is the viscosity of the solvent and $\rho$ is the
224 + hydrodynamic radius.  The presence of the rotational resistance tensor
225 + implies that the spherical body has internal structure and
226 + orientational degrees of freedom that must be propagated in time.  For
227 + non-structured spherical bodies (i.e. the atoms in a traditional
228 + molecular dynamics simulation) these degrees of freedom do not exist.
229 +
230 + Other non-spherical shapes, such as cylinders and ellipsoids, are
231 + widely used as references for developing new hydrodynamic theories,
232 + because their properties can be calculated exactly. In 1936, Perrin
233 + extended Stokes' law to general
234 + ellipsoids,\cite{Perrin1934,Perrin1936} described in Cartesian
235 + coordinates as
236 + \begin{equation}
237 + \frac{x^2 }{a^2} + \frac{y^2}{b^2} + \frac{z^2 }{c^2} = 1.
238 + \end{equation}
239 + Here, the semi-axes are of lengths $a$, $b$, and $c$. Due to the
240 + complexity of the elliptic integral, only uniaxial ellipsoids, either
241 + prolate ($a \ge b = c$) or oblate ($a < b = c$), were solved
242 + exactly. Introducing an elliptic integral parameter $S$ for prolate,
243 + \begin{equation}
244 + S = \frac{2}{\sqrt{a^2  - b^2}} \ln \frac{a + \sqrt{a^2  - b^2}}{b},
245 + \end{equation}
246 + and oblate,
247 + \begin{equation}
248 + S = \frac{2}{\sqrt {b^2  - a^2 }} \arctan \frac{\sqrt {b^2  - a^2}}{a},
249 + \end{equation}
250 + ellipsoids, it is possible to write down exact solutions for the
251 + resistance tensors.  As is the case for spherical bodies, the translational,
252 + \begin{eqnarray}
253 + \Xi_a^{tt}  & = & 16\pi \eta \frac{a^2  - b^2}{(2a^2  - b^2 )S - 2a}. \\
254 + \Xi_b^{tt} =  \Xi_c^{tt} & = & 32\pi \eta \frac{a^2  - b^2 }{(2a^2 - 3b^2 )S + 2a},
255 + \end{eqnarray}
256 + and rotational,
257 + \begin{eqnarray}
258 + \Xi_a^{rr} & = & \frac{32\pi}{3} \eta \frac{(a^2  - b^2 )b^2}{2a - b^2 S}, \\
259 + \Xi_b^{rr} = \Xi_c^{rr} & = & \frac{32\pi}{3} \eta \frac{(a^4  - b^4)}{(2a^2  - b^2 )S - 2a}
260 + \end{eqnarray}
261 + resistance tensors are diagonal $3 \times 3$ matrices. For both
262 + spherical and ellipsoidal particles, the translation-rotation and
263 + rotation-translation coupling tensors are zero.
264 +
265 + \subsubsection{\label{ldintroSection:resistanceTensorRegularArbitrary}\textbf{The Resistance Tensor for Arbitrary Shapes}}
266 + Other than the fluctuation dissipation formulae given by
267 + Peters,\cite{Peters:1999qy,Peters:1999uq,Peters:2000fk} there are no
268 + analytic solutions for the friction tensor for rigid molecules of
269 + arbitrary shape. The ellipsoid of revolution and general triaxial
270 + ellipsoid models have been widely used to approximate the hydrodynamic
271 + properties of rigid bodies. However, the mapping from all possible
272 + ellipsoidal spaces ($r$-space) to all possible combinations of
273 + rotational diffusion coefficients ($D$-space) is not
274 + unique.\cite{Wegener1979} Additionally, because there is intrinsic
275 + coupling between translational and rotational motion of {\it skew}
276 + rigid bodies, general ellipsoids are not always suitable for modeling
277 + rigid molecules.  A number of studies have been devoted to determining
278 + the friction tensor for irregular shapes using methods in which the
279 + molecule of interest is modeled with a combination of
280 + spheres\cite{Carrasco1999} and the hydrodynamic properties of the
281 + molecule are then calculated using a set of two-point interaction
282 + tensors.  We have found the {\it bead} and {\it rough shell} models of
283 + Carrasco and Garc\'{i}a de la Torre to be the most useful of these
284 + methods,\cite{Carrasco1999} and we review the basic outline of the
285 + rough shell approach here.  A more thorough explanation can be found
286 + in Ref. \citen{Carrasco1999}.
287 +
288 + Consider a rigid assembly of $N$ small beads moving through a
289 + continuous medium.  Due to hydrodynamic interactions between the
290 + beads, the net velocity of the $i^\mathrm{th}$ bead relative to the
291 + medium, ${\bf v}'_i$, is different than its unperturbed velocity ${\bf
292 + v}_i$,
293 + \begin{equation}
294 + {\bf v}'_i  = {\bf v}_i  - \sum\limits_{j \ne i} {{\bf T}_{ij} {\bf F}_j }
295 + \end{equation}
296 + where ${\bf F}_j$ is the frictional force on the medium due to bead $j$, and
297 + ${\bf T}_{ij}$ is the hydrodynamic interaction tensor between the two beads.
298 + The frictional force felt by the $i^\mathrm{th}$ bead is proportional to
299 + its net velocity
300 + \begin{equation}
301 + {\bf F}_i  = \xi_i {\bf v}_i  - \xi_i \sum\limits_{j \ne i} {{\bf T}_{ij} {\bf F}_j }.
302 + \label{ldintroEquation:tensorExpression}
303 + \end{equation}
304 + Eq. (\ref{ldintroEquation:tensorExpression}) defines the two-point
305 + hydrodynamic tensor, ${\bf T}_{ij}$.  There have been many proposed
306 + solutions to this equation, including the simple solution given by
307 + Oseen and Burgers in 1930 for two beads of identical radius,
308 + \begin{equation}
309 + {\bf T}_{ij}  = \frac{1}{{8\pi \eta R_{ij} }}\left( {{\bf I} +
310 + \frac{{{\bf R}_{ij} {\bf R}_{ij}^T }}{{R_{ij}^2 }}} \right).
311 + \label{ldintroEquation:oseenTensor}
312 + \end{equation}
313 + Here ${\bf R}_{ij}$ is the distance vector between beads $i$ and
314 + $j$. A second order expression for beads of different hydrodynamic
315 + radii was introduced by Rotne and Prager,\cite{Rotne1969} and improved
316 + by Garc\'{i}a de la Torre and Bloomfield,\cite{Torre1977}
317 + \begin{equation}
318 + {\bf T}_{ij}  = \frac{1}{{8\pi \eta R_{ij} }}\left[ {\left( {{\bf I} +
319 + \frac{{{\bf R}_{ij} {\bf R}_{ij}^T }}{{R_{ij}^2 }}} \right) + \frac{{\rho
320 + _i^2  + \rho_j^2 }}{{R_{ij}^2 }}\left( {\frac{{\bf I}}{3} -
321 + \frac{{{\bf R}_{ij} {\bf R}_{ij}^T }}{{R_{ij}^2 }}} \right)} \right].
322 + \label{ldintroEquation:RPTensorNonOverlapped}
323 + \end{equation}
324 + Both the Oseen-Burgers tensor and
325 + Eq.~\ref{ldintroEquation:RPTensorNonOverlapped} have an assumption
326 + that the beads do not overlap ($R_{ij} \ge \rho_i + \rho_j$).
327 +
328 + To calculate the resistance tensor for a body represented as the union
329 + of many non-overlapping beads, we first pick an arbitrary origin $O$
330 + and then construct a $3N \times 3N$ supermatrix consisting of $N
331 + \times N$ ${\bf B}_{ij}$ blocks
332 + \begin{equation}
333 + {\bf B} = \left( \begin{array}{*{20}c}
334 + {\bf B}_{11} &  \ldots  & {\bf B}_{1N}   \\
335 + \vdots  &  \ddots  &  \vdots   \\
336 + {\bf B}_{N1} &  \cdots  & {\bf B}_{NN}
337 + \end{array} \right)
338 + \end{equation}
339 + ${\bf B}_{ij}$ is a version of the hydrodynamic tensor which includes the
340 + self-contributions for spheres,
341 + \begin{equation}
342 + {\bf B}_{ij}  = \delta _{ij} \frac{{\bf I}}{{6\pi \eta R_{ij}}} + (1 - \delta_{ij}
343 + ){\bf T}_{ij}
344 + \end{equation}
345 + where $\delta_{ij}$ is the Kronecker delta function. Inverting the
346 + ${\bf B}$ matrix, we obtain
347 + \begin{equation}
348 + {\bf C} = {\bf B}^{ - 1}  = \left(\begin{array}{*{20}c}
349 +  {\bf C}_{11} &  \ldots  & {\bf C}_{1N} \\
350 + \vdots  &  \ddots  &  \vdots   \\
351 +  {\bf C}_{N1} &  \cdots  & {\bf C}_{NN}
352 + \end{array} \right),
353 + \end{equation}
354 + which can be partitioned into $N \times N$ blocks labeled ${\bf C}_{ij}$.
355 + (Each of the ${\bf C}_{ij}$ blocks is a $3 \times 3$ matrix.)  Using the
356 + skew matrix,
357 + \begin{equation}
358 + {\bf U}_i  = \left(\begin{array}{*{20}c}
359 +  0 & -z_i & y_i \\
360 + z_i &  0   & - x_i \\
361 + -y_i & x_i & 0
362 + \end{array}\right)
363 + \label{ldeq:skewMatrix}
364 + \end{equation}
365 + where $x_i$, $y_i$, $z_i$ are the components of the vector joining
366 + bead $i$ and origin $O$, the elements of the resistance tensor (at the
367 + arbitrary origin $O$) can be written as
368 + \begin{eqnarray}
369 + \label{ldintroEquation:ResistanceTensorArbitraryOrigin}
370 + \Xi^{tt}  & = & \sum\limits_i {\sum\limits_j {{\bf C}_{ij} } } \notag , \\
371 + \Xi^{tr}  = \Xi _{}^{rt}  & = & \sum\limits_i {\sum\limits_j {{\bf U}_i {\bf C}_{ij} } } , \\
372 + \Xi^{rr}  & = &  -\sum\limits_i \sum\limits_j {\bf U}_i {\bf C}_{ij} {\bf U}_j + 6 \eta V {\bf I}. \notag
373 + \end{eqnarray}
374 + The final term in the expression for $\Xi^{rr}$ is a correction that
375 + accounts for errors in the rotational motion of the bead models. The
376 + additive correction uses the solvent viscosity ($\eta$) as well as the
377 + total volume of the beads that contribute to the hydrodynamic model,
378 + \begin{equation}
379 + V = \frac{4 \pi}{3} \sum_{i=1}^{N} \rho_i^3,
380 + \end{equation}
381 + where $\rho_i$ is the radius of bead $i$.  This correction term was
382 + rigorously tested and compared with the analytical results for
383 + two-sphere and ellipsoidal systems by Garc\'{i}a de la Torre and
384 + Rodes.\cite{Torre:1983lr}
385 +
386 + In general, resistance tensors depend on the origin at which they were
387 + computed.  However, the proper location for applying the friction
388 + force is the center of resistance, the special point at which the
389 + trace of rotational resistance tensor, $\Xi^{rr}$ reaches a minimum
390 + value.  Mathematically, the center of resistance can also be defined
391 + as the unique point for a rigid body at which the translation-rotation
392 + coupling tensors are symmetric,
393 + \begin{equation}
394 + \Xi^{tr}  = \left(\Xi^{tr}\right)^T
395 + \label{ldintroEquation:definitionCR}
396 + \end{equation}
397 + From Eq. \ref{ldintroEquation:ResistanceTensorArbitraryOrigin}, we can
398 + easily derive that the {\it translational} resistance tensor is origin
399 + independent, while the rotational resistance tensor and
400 + translation-rotation coupling resistance tensor depend on the
401 + origin. Given the resistance tensor at an arbitrary origin $O$, and a
402 + vector ,${\bf r}_{OP} = (x_{OP}, y_{OP}, z_{OP})$, from $O$ to $P$, we
403 + can obtain the resistance tensor at $P$ by
404 + \begin{eqnarray}
405 + \label{ldintroEquation:resistanceTensorTransformation}
406 + \Xi_P^{tt}  & = & \Xi_O^{tt}  \notag \\
407 + \Xi_P^{tr}  = \Xi_P^{rt}  & = & \Xi_O^{tr}  - {\bf U}_{OP} \Xi _O^{tt}  \\
408 + \Xi_P^{rr}  & = &\Xi_O^{rr}  - {\bf U}_{OP} \Xi_O^{tt} {\bf U}_{OP}
409 + + \Xi_O^{tr} {\bf U}_{OP}  - {\bf U}_{OP} \left( \Xi_O^{tr}
410 + \right)^{^T} \notag
411 + \end{eqnarray}
412 + where ${\bf U}_{OP}$ is the skew matrix (Eq. (\ref{ldeq:skewMatrix}))
413 + for the vector between the origin $O$ and the point $P$. Using
414 + Eqs.~\ref{ldintroEquation:definitionCR}~and~\ref{ldintroEquation:resistanceTensorTransformation},
415 + one can locate the position of center of resistance,
416 + \begin{eqnarray*}
417 + \left(\begin{array}{l}
418 + x_{OR} \\
419 + y_{OR} \\
420 + z_{OR} \\
421 + \end{array}\right) & = &
422 + \left(\begin{array}{*{20}c}
423 + (\Xi_O^{rr})_{yy} + (\Xi_O^{rr})_{zz} & -(\Xi_O^{rr})_{xy} & -(\Xi_O^{rr})_{xz} \\
424 + -(\Xi_O^{rr})_{xy} & (\Xi_O^{rr})_{zz} + (\Xi_O^{rr})_{xx} & -(\Xi_O^{rr})_{yz} \\
425 + -(\Xi_O^{rr})_{xz} & -(\Xi_O^{rr})_{yz} & (\Xi_O^{rr})_{xx} + (\Xi_O^{rr})_{yy} \\
426 + \end{array}\right)^{-1} \\
427 + & & \left(\begin{array}{l}
428 + (\Xi_O^{tr})_{yz} - (\Xi_O^{tr})_{zy} \\
429 + (\Xi_O^{tr})_{zx} - (\Xi_O^{tr})_{xz} \\
430 + (\Xi_O^{tr})_{xy} - (\Xi_O^{tr})_{yx} \\
431 + \end{array}\right) \\
432 + \end{eqnarray*}
433 + where $x_{OR}$, $y_{OR}$, $z_{OR}$ are the components of the vector
434 + joining center of resistance $R$ and origin $O$.
435 +
436 + For a general rigid molecular substructure, finding the $6 \times 6$
437 + resistance tensor can be a computationally demanding task.  First, a
438 + lattice of small beads that extends well beyond the boundaries of the
439 + rigid substructure is created.  The lattice is typically composed of
440 + 0.25 \AA\ beads on a dense FCC lattice.  The lattice constant is taken
441 + to be the bead diameter, so that adjacent beads are touching, but do
442 + not overlap. To make a shape corresponding to the rigid structure,
443 + beads that sit on lattice sites that are outside the van der Waals
444 + radii of all of the atoms comprising the rigid body are excluded from
445 + the calculation.
446 +
447 + For large structures, most of the beads will be deep within the rigid
448 + body and will not contribute to the hydrodynamic tensor.  In the {\it
449 + rough shell} approach, beads which have all of their lattice neighbors
450 + inside the structure are considered interior beads, and are removed
451 + from the calculation.  After following this procedure, only those
452 + beads in direct contact with the van der Waals surface of the rigid
453 + body are retained.  For reasonably large molecular structures, this
454 + truncation can still produce bead assemblies with thousands of
455 + members.
456 +
457 + If all of the {\it atoms} comprising the rigid substructure are
458 + spherical and non-overlapping, the tensor in
459 + Eq.~(\ref{ldintroEquation:RPTensorNonOverlapped}) may be used directly
460 + using the atoms themselves as the hydrodynamic beads.  This is a
461 + variant of the {\it bead model} approach of Carrasco and Garc\'{i}a de
462 + la Torre.\cite{Carrasco1999} In this case, the size of the ${\bf B}$
463 + matrix can be quite small, and the calculation of the hydrodynamic
464 + tensor is straightforward.
465 +
466 + In general, the inversion of the ${\bf B}$ matrix is the most
467 + computationally demanding task.  This inversion is done only once for
468 + each type of rigid structure.  We have used straightforward
469 + LU-decomposition to solve the linear system and to obtain the elements
470 + of ${\bf C}$. Once ${\bf C}$ has been obtained, the location of the
471 + center of resistance ($R$) is found and the resistance tensor at this
472 + point is calculated.  The $3 \times 1$ vector giving the location of
473 + the rigid body's center of resistance and the $6 \times 6$ resistance
474 + tensor are then stored for use in the Langevin dynamics calculation.
475 + These quantities depend on solvent viscosity and temperature and must
476 + be recomputed if different simulation conditions are required.
477 +
478 + \section{Langevin Dynamics for Rigid Particles of Arbitrary Shape\label{ldLDRB}}
479 +
480 + Consider the Langevin equations of motion in generalized coordinates
481 + \begin{equation}
482 + {\bf M} \dot{{\bf V}}(t) = {\bf F}_{s}(t) +
483 + {\bf F}_{f}(t)  + {\bf F}_{r}(t)
484 + \label{ldLDGeneralizedForm}
485 + \end{equation}
486 + where ${\bf M}$ is a $6 \times 6$ diagonal mass matrix (which
487 + includes the mass of the rigid body as well as the moments of inertia
488 + in the body-fixed frame) and ${\bf V}$ is a generalized velocity,
489 + ${\bf V} =
490 + \left\{{\bf v},{\bf \omega}\right\}$. The right side of
491 + Eq.~\ref{ldLDGeneralizedForm} consists of three generalized forces: a
492 + system force (${\bf F}_{s}$), a frictional or dissipative force (${\bf
493 + F}_{f}$) and a stochastic force (${\bf F}_{r}$). While the evolution
494 + of the system in Newtonian mechanics is typically done in the lab
495 + frame, it is convenient to handle the dynamics of rigid bodies in
496 + body-fixed frames. Thus the friction and random forces on each
497 + substructure are calculated in a body-fixed frame and may converted
498 + back to the lab frame using that substructure's rotation matrix (${\bf
499 + Q}$):
500 + \begin{equation}
501 + {\bf F}_{f,r} =
502 + \left( \begin{array}{c}
503 + {\bf f}_{f,r} \\
504 + {\bf \tau}_{f,r}
505 + \end{array} \right)
506 + =
507 + \left( \begin{array}{c}
508 + {\bf Q}^{T} {\bf f}^{~b}_{f,r} \\
509 + {\bf Q}^{T} {\bf \tau}^{~b}_{f,r}
510 + \end{array} \right)
511 + \end{equation}
512 + The body-fixed friction force, ${\bf F}_{f}^{~b}$, is proportional to
513 + the (body-fixed) velocity at the center of resistance
514 + ${\bf v}_{R}^{~b}$ and the angular velocity ${\bf \omega}$
515 + \begin{equation}
516 + {\bf F}_{f}^{~b}(t) = \left( \begin{array}{l}
517 + {\bf f}_{f}^{~b}(t) \\
518 + {\bf \tau}_{f}^{~b}(t) \\
519 + \end{array} \right) =  - \left( \begin{array}{*{20}c}
520 +   \Xi_{R}^{tt} & \Xi_{R}^{rt} \\
521 +   \Xi_{R}^{tr} & \Xi_{R}^{rr}    \\
522 + \end{array} \right)\left( \begin{array}{l}
523 + {\bf v}_{R}^{~b}(t) \\
524 + {\bf \omega}(t) \\
525 + \end{array} \right),
526 + \end{equation}
527 + while the random force, ${\bf F}_{r}$, is a Gaussian stochastic
528 + variable with zero mean and variance,
529 + \begin{equation}
530 + \left\langle {{\bf F}_{r}(t) ({\bf F}_{r}(t'))^T } \right\rangle  =
531 + \left\langle {{\bf F}_{r}^{~b} (t) ({\bf F}_{r}^{~b} (t'))^T } \right\rangle  =
532 + 2 k_B T \Xi_R \delta(t - t'). \label{ldeq:randomForce}
533 + \end{equation}
534 + $\Xi_R$ is the $6\times6$ resistance tensor at the center of
535 + resistance.  Once this tensor is known for a given rigid body (as
536 + described in the previous section) obtaining a stochastic vector that
537 + has the properties in Eq. (\ref{ldeq:randomForce}) can be done
538 + efficiently by carrying out a one-time Cholesky decomposition to
539 + obtain the square root matrix of the resistance tensor,
540 + \begin{equation}
541 + \Xi_R = {\bf S} {\bf S}^{T},
542 + \label{ldeq:Cholesky}
543 + \end{equation}
544 + where ${\bf S}$ is a lower triangular matrix.\cite{Schlick2002} A
545 + vector with the statistics required for the random force can then be
546 + obtained by multiplying ${\bf S}$ onto a random 6-vector ${\bf Z}$ which
547 + has elements chosen from a Gaussian distribution, such that:
548 + \begin{equation}
549 + \langle {\bf Z}_i \rangle = 0, \hspace{1in} \langle {\bf Z}_i \cdot
550 + {\bf Z}_j \rangle = \frac{2 k_B T}{\delta t} \delta_{ij},
551 + \end{equation}
552 + where $\delta t$ is the timestep in use during the simulation. The
553 + random force, ${\bf F}_{r}^{~b} = {\bf S} {\bf Z}$, can be shown to have the
554 + correct properties required by Eq. (\ref{ldeq:randomForce}).
555 +
556 + The equation of motion for the translational velocity of the center of
557 + mass (${\bf v}$) can be written as
558 + \begin{equation}
559 + m \dot{{\bf v}} (t) =  {\bf f}_{s}(t) + {\bf f}_{f}(t) +
560 + {\bf f}_{r}(t)
561 + \end{equation}
562 + Since the frictional and random forces are applied at the center of
563 + resistance, which generally does not coincide with the center of mass,
564 + extra torques are exerted at the center of mass. Thus, the net
565 + body-fixed torque at the center of mass, $\tau^{~b}(t)$,
566 + is given by
567 + \begin{equation}
568 + \tau^{~b} \leftarrow \tau_{s}^{~b} + \tau_{f}^{~b} + \tau_{r}^{~b} + {\bf r}_{MR} \times \left( {\bf f}_{f}^{~b} + {\bf f}_{r}^{~b} \right)
569 + \end{equation}
570 + where ${\bf r}_{MR}$ is the vector from the center of mass to the center of
571 + resistance. Instead of integrating the angular velocity in lab-fixed
572 + frame, we consider the equation of motion for the angular momentum
573 + (${\bf j}$) in the body-fixed frame
574 + \begin{equation}
575 + \frac{\partial}{\partial t}{\bf j}(t) = \tau^{~b}(t)
576 + \end{equation}
577 + Embedding the friction and random forces into the the total force and
578 + torque, one can integrate the Langevin equations of motion for a rigid
579 + body of arbitrary shape in a velocity-Verlet style 2-part algorithm,
580 + where $h = \delta t$:
581 +
582 + {\tt move A:}
583 + \begin{align*}
584 + {\bf v}\left(t + h / 2\right)  &\leftarrow  {\bf v}(t)
585 +    + \frac{h}{2} \left( {\bf f}(t) / m \right), \\
586 + %
587 + {\bf r}(t + h) &\leftarrow {\bf r}(t)
588 +    + h  {\bf v}\left(t + h / 2 \right), \\
589 + %
590 + {\bf j}\left(t + h / 2 \right)  &\leftarrow {\bf j}(t)
591 +    + \frac{h}{2} {\bf \tau}^{~b}(t), \\
592 + %
593 + {\bf Q}(t + h) &\leftarrow \mathrm{rotate}\left( h {\bf j}
594 +    (t + h / 2) \cdot \overleftrightarrow{\mathsf{I}}^{-1} \right).
595 + \end{align*}
596 + In this context, $\overleftrightarrow{\mathsf{I}}$ is the diagonal
597 + moment of inertia tensor, and the $\mathrm{rotate}$ function is the
598 + reversible product of the three body-fixed rotations,
599 + \begin{equation}
600 + \mathrm{rotate}({\bf a}) = \mathsf{G}_x(a_x / 2) \cdot
601 + \mathsf{G}_y(a_y / 2) \cdot \mathsf{G}_z(a_z) \cdot \mathsf{G}_y(a_y
602 + / 2) \cdot \mathsf{G}_x(a_x /2),
603 + \end{equation}
604 + where each rotational propagator, $\mathsf{G}_\alpha(\theta)$,
605 + rotates both the rotation matrix ($\mathbf{Q}$) and the body-fixed
606 + angular momentum (${\bf j}$) by an angle $\theta$ around body-fixed
607 + axis $\alpha$,
608 + \begin{equation}
609 + \mathsf{G}_\alpha( \theta ) = \left\{
610 + \begin{array}{lcl}
611 + \mathbf{Q}(t) & \leftarrow & \mathbf{Q}(0) \cdot \mathsf{R}_\alpha(\theta)^T, \\
612 + {\bf j}(t) & \leftarrow & \mathsf{R}_\alpha(\theta) \cdot {\bf
613 + j}(0).
614 + \end{array}
615 + \right.
616 + \end{equation}
617 + $\mathsf{R}_\alpha$ is a quadratic approximation to the single-axis
618 + rotation matrix.  For example, in the small-angle limit, the
619 + rotation matrix around the body-fixed x-axis can be approximated as
620 + \begin{equation}
621 + \mathsf{R}_x(\theta) \approx \left(
622 + \begin{array}{ccc}
623 + 1 & 0 & 0 \\
624 + 0 & \frac{1-\theta^2 / 4}{1 + \theta^2 / 4}  & -\frac{\theta}{1+
625 + \theta^2 / 4} \\
626 + 0 & \frac{\theta}{1+ \theta^2 / 4} & \frac{1-\theta^2 / 4}{1 +
627 + \theta^2 / 4}
628 + \end{array}
629 + \right).
630 + \end{equation}
631 + All other rotations follow in a straightforward manner. After the
632 + first part of the propagation, the forces and body-fixed torques are
633 + calculated at the new positions and orientations.  The system forces
634 + and torques are derivatives of the total potential energy function
635 + ($U$) with respect to the rigid body positions (${\bf r}$) and the
636 + columns of the transposed rotation matrix ${\bf Q}^T = \left({\bf
637 + u}_x, {\bf u}_y, {\bf u}_z \right)$:
638 +
639 + {\tt Forces:}
640 + \begin{align*}
641 + {\bf f}_{s}(t + h) & \leftarrow
642 +    - \left(\frac{\partial U}{\partial {\bf r}}\right)_{{\bf r}(t + h)} \\
643 + %
644 + {\bf \tau}_{s}(t + h) &\leftarrow {\bf u}(t + h)
645 +    \times \frac{\partial U}{\partial {\bf u}} \\
646 + %
647 + {\bf v}^{b}_{R}(t+h) & \leftarrow \mathbf{Q}(t+h) \cdot \left({\bf v}(t+h) + {\bf \omega}(t+h) \times {\bf r}_{MR} \right) \\
648 + %
649 + {\bf f}_{R,f}^{b}(t+h) & \leftarrow - \Xi_{R}^{tt} \cdot
650 + {\bf v}^{b}_{R}(t+h) - \Xi_{R}^{rt} \cdot {\bf \omega}(t+h) \\
651 + %
652 + {\bf \tau}_{R,f}^{b}(t+h) & \leftarrow - \Xi_{R}^{tr} \cdot
653 + {\bf v}^{b}_{R}(t+h) - \Xi_{R}^{rr} \cdot {\bf \omega}(t+h) \\
654 + %
655 + Z & \leftarrow  {\tt GaussianNormal}(2 k_B T / h, 6) \\
656 + {\bf F}_{R,r}^{b}(t+h) & \leftarrow {\bf S} \cdot Z \\
657 + %
658 + {\bf f}(t+h) & \leftarrow {\bf f}_{s}(t+h) + \mathbf{Q}^{T}(t+h)
659 + \cdot \left({\bf f}_{R,f}^{~b} + {\bf f}_{R,r}^{~b} \right) \\
660 + %
661 + \tau(t+h) & \leftarrow \tau_{s}(t+h) + \mathbf{Q}^{T}(t+h) \cdot \left(\tau_{R,f}^{~b} + \tau_{R,r}^{~b} \right) + {\bf r}_{MR} \times \left({\bf f}_{f}(t+h) + {\bf f}_{r}(t+h) \right) \\
662 + \tau^{~b}(t+h) & \leftarrow \mathbf{Q}(t+h) \cdot \tau(t+h) \\
663 + \end{align*}
664 + Frictional (and random) forces and torques must be computed at the
665 + center of resistance, so there are additional steps required to find
666 + the body-fixed velocity (${\bf v}_{R}^{~b}$) at this location.  Mapping
667 + the frictional and random forces at the center of resistance back to
668 + the center of mass also introduces an additional term in the torque
669 + one obtains at the center of mass.
670 +
671 + Once the forces and torques have been obtained at the new time step,
672 + the velocities can be advanced to the same time value.
673 +
674 + {\tt move B:}
675 + \begin{align*}
676 + {\bf v}\left(t + h \right)  &\leftarrow  {\bf v}\left(t + h / 2
677 + \right)
678 +    + \frac{h}{2} \left( {\bf f}(t + h) / m \right), \\
679 + %
680 + {\bf j}\left(t + h \right)  &\leftarrow {\bf j}\left(t + h / 2
681 + \right)
682 +    + \frac{h}{2} {\bf \tau}^{~b}(t + h) .
683 + \end{align*}
684 +
685 + \section{Validating the Method\label{ldsec:validating}}
686 + In order to validate our Langevin integrator for arbitrarily-shaped
687 + rigid bodies, we implemented the algorithm in {\sc
688 + oopse}\cite{Meineke2005} and  compared the results of this algorithm
689 + with the known
690 + hydrodynamic limiting behavior for a few model systems, and to
691 + microcanonical molecular dynamics simulations for some more
692 + complicated bodies. The model systems and their analytical behavior
693 + (if known) are summarized below. Parameters for the primary particles
694 + comprising our model systems are given in table \ref{ldtab:parameters},
695 + and a sketch of the arrangement of these primary particles into the
696 + model rigid bodies is shown in figure \ref{ldfig:models}. In table
697 + \ref{ldtab:parameters}, $d$ and $l$ are the physical dimensions of
698 + ellipsoidal (Gay-Berne) particles.  For spherical particles, the value
699 + of the Lennard-Jones $\sigma$ parameter is the particle diameter
700 + ($d$).  Gay-Berne ellipsoids have an energy scaling parameter,
701 + $\epsilon^s$, which describes the well depth for two identical
702 + ellipsoids in a {\it side-by-side} configuration.  Additionally, a
703 + well depth aspect ratio, $\epsilon^r = \epsilon^e / \epsilon^s$,
704 + describes the ratio between the well depths in the {\it end-to-end}
705 + and side-by-side configurations.  For spheres, $\epsilon^r \equiv 1$.
706 + Moments of inertia are also required to describe the motion of primary
707 + particles with orientational degrees of freedom.
708 +
709 + \begin{table*}
710 + \begin{minipage}{\linewidth}
711 + \begin{center}
712 + \caption{PARAMETERS FOR THE PRIMARY PARTICLES IN USE BY THE RIGID BODY
713 + MODELS IN FIGURE \ref{ldfig:models}}
714 + \begin{tabular}{lrcccccccc}
715 + \hline
716 + & & & & & & \multicolumn{3}c{$\overleftrightarrow{\mathsf I}$ (amu \AA$^2$)} \\
717 + & $d$ (\AA) & $l$ (\AA) & $\epsilon^s$ ($\frac{kcal}{mol}$) & $\epsilon^r$ &
718 + $m$ (amu) & $I_{xx}$ & $I_{yy}$ & $I_{zz}$ \\ \hline
719 + Sphere   & 6.5 & $= d$ & 0.8 & 1 & 190 & 802.75 & 802.75  & 802.75  \\
720 + Ellipsoid & 4.6 & 13.8  & 0.8 & 0.2 & 200 & 2105 & 2105 & 421 \\
721 + Dumbbell: & & & & & & & & \\
722 + \quad {\it 2 spheres} & 6.5 & $= d$ & 0.8 & 1   & 190 & 802.75 & 802.75 & 802.75 \\
723 + Banana: & & & & & & & & \\
724 + \quad {\it 3 ellipsoids} & 4.2 & 11.2  & 0.8 & 0.2 & 240 & 10000 & 10000 & 0 \\
725 + Lipid: & & & & & & & & \\
726 + \quad {\it head} & 6.5 & $= d$ & 0.185 & 1 & 196 & & & \\
727 + \quad {\it Tail} & 4.6 & 13.8  & 0.8   & 0.2 & 760 & 45000 & 45000 & 9000 \\
728 + Solvent & 4.7 & $= d$ & 0.8 & 1   & 72.06 & & & \\
729 + \hline
730 + \end{tabular}
731 + \label{ldtab:parameters}
732 + \end{center}
733 + \end{minipage}
734 + \end{table*}
735 +
736 + \begin{figure}
737 + \centering
738 + \includegraphics[width=3in]{./figures/ldSketch}
739 + \caption[Sketch of the model systems]{A sketch of the model systems
740 + used in evaluating the behavior of the rigid body Langevin
741 + integrator.} \label{ldfig:models}
742 + \end{figure}
743 +
744 + \subsection{Simulation Methodology}
745 + We performed reference microcanonical simulations with explicit
746 + solvents for each of the different model system.  In each case there
747 + was one solute model and 1929 solvent molecules present in the
748 + simulation box.  All simulations were equilibrated for 5 ns using a
749 + constant-pressure and temperature integrator with target values of 300
750 + K for the temperature and 1 atm for pressure.  Following this stage,
751 + further equilibration (5 ns) and sampling (10 ns) was done in a
752 + microcanonical ensemble.  Since the model bodies are typically quite
753 + massive, we were able to use a time step of 25 fs.
754 +
755 + The model systems studied used both Lennard-Jones spheres as well as
756 + uniaxial Gay-Berne ellipoids. The Gay-Berne potential is given by
757 + equation~\ref{mdeq:gb}.  For the interaction between nonequivalent
758 + uniaxial ellipsoids (or between spheres and ellipsoids), the spheres
759 + are treated as ellipsoids with an aspect ratio of 1 ($d = l$) and with
760 + an well depth ratio ($\epsilon^r$) of 1 ($\epsilon^e = \epsilon^s$).
761 + A switching function (Eq.~\ref{mdeq:dipoleSwitching}) was applied to
762 + all potentials to smoothly turn off the interactions between a range
763 + of $22$ and $25$ \AA.
764 +
765 + To measure shear viscosities from our microcanonical simulations, we
766 + used the Einstein form of the pressure correlation function,\cite{hess:209}
767 + \begin{equation}
768 + \eta = \lim_{t->\infty} \frac{V}{2 k_B T} \frac{d}{dt} \left\langle \left(
769 + \int_{t_0}^{t_0 + t} P_{xz}(t') dt' \right)^2 \right\rangle_{t_0}.
770 + \label{ldeq:shear}
771 + \end{equation}
772 + which converges much more rapidly in molecular dynamics simulations
773 + than the traditional Green-Kubo formula.
774 +
775 + The Langevin dynamics for the different model systems were performed
776 + at the same temperature as the average temperature of the
777 + microcanonical simulations and with a solvent viscosity taken from
778 + Eq. (\ref{ldeq:shear}) applied to these simulations.  We used 1024
779 + independent solute simulations to obtain statistics on our Langevin
780 + integrator.
781 +
782 + \subsection{Analysis}
783 +
784 + The quantities of interest when comparing the Langevin integrator to
785 + analytic hydrodynamic equations and to molecular dynamics simulations
786 + are typically translational diffusion constants and orientational
787 + relaxation times.  Translational diffusion constants for point
788 + particles are computed easily from the long-time slope of the
789 + mean-square displacement (Eq.~\ref{mdeq:msdisplacement}) of the solute
790 + molecules.  For models in which the translational diffusion tensor
791 + (${\bf D}_{tt}$) has non-degenerate eigenvalues (i.e. any
792 + non-spherically-symmetric rigid body), it is possible to compute the
793 + diffusive behavior for motion parallel to each body-fixed axis by
794 + projecting the displacement of the particle onto the body-fixed
795 + reference frame at $t=0$.  With an isotropic solvent, as we have used
796 + in this study, there may be differences between the three diffusion
797 + constants at short times, but these must converge to the same value at
798 + longer times.  Translational diffusion constants for the different
799 + shaped models are shown in table \ref{ldtab:translation}.
800 +
801 + In general, the three eigenvalues ($D_1, D_2, D_3$) of the rotational
802 + diffusion tensor (${\bf D}_{rr}$) measure the diffusion of an object
803 + {\it around} a particular body-fixed axis and {\it not} the diffusion
804 + of a vector pointing along the axis.  However, these eigenvalues can
805 + be combined to find 5 characteristic rotational relaxation
806 + times,\cite{PhysRev.119.53,Berne90}
807 + \begin{eqnarray}
808 + 1 / \tau_1 & = & 6 D_r + 2 \Delta \\
809 + 1 / \tau_2 & = & 6 D_r - 2 \Delta \\
810 + 1 / \tau_3 & = & 3 (D_r + D_1)  \\
811 + 1 / \tau_4 & = & 3 (D_r + D_2) \\
812 + 1 / \tau_5 & = & 3 (D_r + D_3)
813 + \end{eqnarray}
814 + where
815 + \begin{equation}
816 + D_r = \frac{1}{3} \left(D_1 + D_2 + D_3 \right)
817 + \end{equation}
818 + and
819 + \begin{equation}
820 + \Delta = \left( (D_1 - D_2)^2 + (D_3 - D_1 )(D_3 - D_2)\right)^{1/2}
821 + \end{equation}
822 + Each of these characteristic times can be used to predict the decay of
823 + part of the rotational correlation function when $\ell = 2$,
824 + \begin{equation}
825 + C_2(t)  =  \frac{a^2}{N^2} e^{-t/\tau_1} + \frac{b^2}{N^2} e^{-t/\tau_2}.
826 + \end{equation}
827 + This is the same as the $F^2_{0,0}(t)$ correlation function that
828 + appears in Ref. \citen{Berne90}.  The amplitudes of the two decay
829 + terms are expressed in terms of three dimensionless functions of the
830 + eigenvalues: $a = \sqrt{3} (D_1 - D_2)$, $b = (2D_3 - D_1 - D_2 +
831 + 2\Delta)$, and $N = 2 \sqrt{\Delta b}$.  Similar expressions can be
832 + obtained for other angular momentum correlation
833 + functions.\cite{PhysRev.119.53,Berne90} In all of the model systems we
834 + studied, only one of the amplitudes of the two decay terms was
835 + non-zero, so it was possible to derive a single relaxation time for
836 + each of the hydrodynamic tensors. In many cases, these characteristic
837 + times are averaged and reported in the literature as a single relaxation
838 + time,\cite{Garcia-de-la-Torre:1997qy}
839 + \begin{equation}
840 + 1 / \tau_0 = \frac{1}{5} \sum_{i=1}^5 \tau_{i}^{-1},
841 + \end{equation}
842 + although for the cases reported here, this averaging is not necessary
843 + and only one of the five relaxation times is relevant.
844 +
845 + To test the Langevin integrator's behavior for rotational relaxation,
846 + we have compared the analytical orientational relaxation times (if
847 + they are known) with the general result from the diffusion tensor and
848 + with the results from both the explicitly solvated molecular dynamics
849 + and Langevin simulations.  Relaxation times from simulations (both
850 + microcanonical and Langevin), were computed using Legendre polynomial
851 + correlation functions for a unit vector (${\bf u}$) fixed along one or
852 + more of the body-fixed axes of the model.
853 + \begin{equation}
854 + C_{\ell}(t)  =  \left\langle P_{\ell}\left({\bf u}_{i}(t) \cdot {\bf
855 + u}_{i}(0) \right) \right\rangle
856 + \end{equation}
857 + For simulations in the high-friction limit, orientational correlation
858 + times can then be obtained from exponential fits of this function, or by
859 + integrating,
860 + \begin{equation}
861 + \tau = \ell (\ell + 1) \int_0^{\infty} C_{\ell}(t) dt.
862 + \end{equation}
863 + In lower-friction solvents, the Legendre correlation functions often
864 + exhibit non-ex\-po\-nen\-tial decay, and may not be characterized by a
865 + single decay constant.
866 +
867 + In table \ref{ldtab:rotation} we show the characteristic rotational
868 + relaxation times (based on the diffusion tensor) for each of the model
869 + systems compared with the values obtained via microcanonical and Langevin
870 + simulations.
871 +
872 + \subsection{Spherical particles}
873 + Our model system for spherical particles was a Lennard-Jones sphere of
874 + diameter ($\sigma$) 6.5 \AA\ in a sea of smaller spheres ($\sigma$ =
875 + 4.7 \AA).  The well depth ($\epsilon$) for both particles was set to
876 + an arbitrary value of 0.8 kcal/mol.  
877 +
878 + The Stokes-Einstein behavior of large spherical particles in
879 + hydrodynamic flows with ``stick'' boundary conditions is well known,
880 + and is given in Eqs. (\ref{ldeq:StokesTranslation}) and
881 + (\ref{ldeq:StokesRotation}).  Recently, Schmidt and Skinner have
882 + computed the behavior of spherical tag particles in molecular dynamics
883 + simulations, and have shown that {\it slip} boundary conditions
884 + ($\Xi_{tt} = 4 \pi \eta \rho$) may be more appropriate for
885 + molecule-sized spheres embedded in a sea of spherical solvent
886 + particles.\cite{Schmidt:2004fj,Schmidt:2003kx}
887 +
888 + Our simulation results show similar behavior to the behavior observed
889 + by Schmidt and Skinner.  The diffusion constant obtained from our
890 + microcanonical molecular dynamics simulations lies between the slip
891 + and stick boundary condition results obtained via Stokes-Einstein
892 + behavior.  Since the Langevin integrator assumes Stokes-Einstein stick
893 + boundary conditions in calculating the drag and random forces for
894 + spherical particles, our Langevin routine obtains nearly quantitative
895 + agreement with the hydrodynamic results for spherical particles.  One
896 + avenue for improvement of the method would be to compute elements of
897 + $\Xi^{tt}$ assuming behavior intermediate between the two boundary
898 + conditions.
899 +
900 + In the explicit solvent simulations, both our solute and solvent
901 + particles were structureless, exerting no torques upon each other.
902 + Therefore, there are not rotational correlation times available for
903 + this model system.
904 +
905 + \subsection{Ellipsoids}
906 + For uniaxial ellipsoids ($a > b = c$), Perrin's formulae for both
907 + translational and rotational diffusion of each of the body-fixed axes
908 + can be combined to give a single translational diffusion
909 + constant,\cite{Berne90}
910 + \begin{equation}
911 + D = \frac{k_B T}{6 \pi \eta a} G(s),
912 + \label{ldDperrin}
913 + \end{equation}
914 + as well as a single rotational diffusion coefficient,
915 + \begin{equation}
916 + \Theta = \frac{3 k_B T}{16 \pi \eta a^3} \left\{ \frac{(2 - s^2)
917 + G(s) - 1}{1 - s^4} \right\}.
918 + \label{ldThetaPerrin}
919 + \end{equation}
920 + In these expressions, $G(s)$ is a function of the axial ratio
921 + ($s = b / a$), which for prolate ellipsoids, is
922 + \begin{equation}
923 + G(s) = (1- s^2)^{-1/2} \ln \left\{ \frac{1 + (1 - s^2)^{1/2}}{s} \right\}
924 + \label{ldGPerrin}
925 + \end{equation}
926 + Again, there is some uncertainty about the correct boundary conditions
927 + to use for molecular scale ellipsoids in a sea of similarly-sized
928 + solvent particles.  Ravichandran and Bagchi found that {\it slip}
929 + boundary conditions most closely resembled the simulation
930 + results,\cite{Ravichandran:1999fk} in agreement with earlier work of
931 + Tang and Evans.\cite{TANG:1993lr}
932 +
933 + Even though there are analytic resistance tensors for ellipsoids, we
934 + constructed a rough-shell model using 2135 beads (each with a diameter
935 + of 0.25 \AA) to approximate the shape of the model ellipsoid.  We
936 + compared the Langevin dynamics from both the simple ellipsoidal
937 + resistance tensor and the rough shell approximation with
938 + microcanonical simulations and the predictions of Perrin.  As in the
939 + case of our spherical model system, the Langevin integrator reproduces
940 + almost exactly the behavior of the Perrin formulae (which is
941 + unsurprising given that the Perrin formulae were used to derive the
942 + drag and random forces applied to the ellipsoid).  We obtain
943 + translational diffusion constants and rotational correlation times
944 + that are within a few percent of the analytic values for both the
945 + exact treatment of the diffusion tensor as well as the rough-shell
946 + model for the ellipsoid.
947 +
948 + The translational diffusion constants from the microcanonical
949 + simulations agree well with the predictions of the Perrin model,
950 + although the {\it rotational} correlation times are a factor of 2
951 + shorter than expected from hydrodynamic theory.  One explanation for
952 + the slower rotation of explicitly-solvated ellipsoids is the
953 + possibility that solute-solvent collisions happen at both ends of the
954 + solute whenever the principal axis of the ellipsoid is turning. In the
955 + upper portion of figure \ref{ldfig:explanation} we sketch a physical
956 + picture of this explanation.  Since our Langevin integrator is
957 + providing nearly quantitative agreement with the Perrin model, it also
958 + predicts orientational diffusion for ellipsoids that exceed explicitly
959 + solvated correlation times by a factor of two.
960 +
961 + \subsection{Rigid dumbbells}
962 + Perhaps the only {\it composite} rigid body for which analytic
963 + expressions for the hydrodynamic tensor are available is the
964 + two-sphere dumbbell model.  This model consists of two non-overlapping
965 + spheres held by a rigid bond connecting their centers. There are
966 + competing expressions for the 6x6 resistance tensor for this
967 + model. The second order expression introduced by Rotne and
968 + Prager,\cite{Rotne1969} and improved by Garc\'{i}a de la Torre and
969 + Bloomfield,\cite{Torre1977} is given above as
970 + Eq. (\ref{ldintroEquation:RPTensorNonOverlapped}).  In our case, we use
971 + a model dumbbell in which the two spheres are identical Lennard-Jones
972 + particles ($\sigma$ = 6.5 \AA\ , $\epsilon$ = 0.8 kcal / mol) held at
973 + a distance of 6.532 \AA.
974 +
975 + The theoretical values for the translational diffusion constant of the
976 + dumbbell are calculated from the work of Stimson and Jeffery, who
977 + studied the motion of this system in a flow parallel to the
978 + inter-sphere axis,\cite{Stimson:1926qy} and Davis, who studied the
979 + motion in a flow {\it perpendicular} to the inter-sphere
980 + axis.\cite{Davis:1969uq} We know of no analytic solutions for the {\it
981 + orientational} correlation times for this model system (other than
982 + those derived from the 6 x 6 tensor mentioned above).
983 +
984 + The bead model for this model system comprises the two large spheres
985 + by themselves, while the rough shell approximation used 3368 separate
986 + beads (each with a diameter of 0.25 \AA) to approximate the shape of
987 + the rigid body.  The hydrodynamics tensors computed from both the bead
988 + and rough shell models are remarkably similar.  Computing the initial
989 + hydrodynamic tensor for a rough shell model can be quite expensive (in
990 + this case it requires inverting a 10104 x 10104 matrix), while the
991 + bead model is typically easy to compute (in this case requiring
992 + inversion of a 6 x 6 matrix).  
993 +
994 + \begin{figure}
995 + \centering
996 + \includegraphics[width=2in]{./figures/ldRoughShell}
997 + \caption[Model rigid bodies and their rough shell approximations]{The
998 + model rigid bodies (left column) used to test this algorithm and their
999 + rough-shell approximations (right-column) that were used to compute
1000 + the hydrodynamic tensors.  The top two models (ellipsoid and dumbbell)
1001 + have analytic solutions and were used to test the rough shell
1002 + approximation.  The lower two models (banana and lipid) were compared
1003 + with explicitly-solvated molecular dynamics simulations. }
1004 + \label{ldfig:roughShell}
1005 + \end{figure}
1006 +
1007 +
1008 + Once the hydrodynamic tensor has been computed, there is no additional
1009 + penalty for carrying out a Langevin simulation with either of the two
1010 + different hydrodynamics models.  Our naive expectation is that since
1011 + the rigid body's surface is roughened under the various shell models,
1012 + the diffusion constants will be even farther from the ``slip''
1013 + boundary conditions than observed for the bead model (which uses a
1014 + Stokes-Einstein model to arrive at the hydrodynamic tensor).  For the
1015 + dumbbell, this prediction is correct although all of the Langevin
1016 + diffusion constants are within 6\% of the diffusion constant predicted
1017 + from the fully solvated system.
1018 +
1019 + For rotational motion, Langevin integration (and the hydrodynamic tensor)
1020 + yields rotational correlation times that are substantially shorter than those
1021 + obtained from explicitly-solvated simulations.  It is likely that this is due
1022 + to the large size of the explicit solvent spheres, a feature that prevents
1023 + the solvent from coming in contact with a substantial fraction of the surface
1024 + area of the dumbbell.  Therefore, the explicit solvent only provides drag
1025 + over a substantially reduced surface area of this model, while the
1026 + hydrodynamic theories utilize the entire surface area for estimating
1027 + rotational diffusion.  A sketch of the free volume available in the explicit
1028 + solvent simulations is shown in figure \ref{ldfig:explanation}.
1029 +
1030 +
1031 + \begin{figure}
1032 + \centering
1033 + \includegraphics[width=4in]{./figures/ldExplanation}
1034 + \caption[Explanations of the differences between orientational
1035 + correlation times for explicitly-solvated models and hydrodynamics
1036 + predictions]{Explanations of the differences between orientational
1037 + correlation times for explicitly-solvated models and hydrodynamic
1038 + predictions.   For the ellipsoids (upper figures), rotation of the
1039 + principal axis can involve correlated collisions at both sides of the
1040 + solute.  In the rigid dumbbell model (lower figures), the large size
1041 + of the explicit solvent spheres prevents them from coming in contact
1042 + with a substantial fraction of the surface area of the dumbbell.
1043 + Therefore, the explicit solvent only provides drag over a
1044 + substantially reduced surface area of this model, where the
1045 + hydrodynamic theories utilize the entire surface area for estimating
1046 + rotational diffusion.
1047 + } \label{ldfig:explanation}
1048 + \end{figure}
1049 +
1050 + \subsection{Composite banana-shaped molecules}
1051 + Banana-shaped rigid bodies composed of three Gay-Berne ellipsoids have
1052 + been used by Orlandi {\it et al.} to observe mesophases in
1053 + coarse-grained models for bent-core liquid crystalline
1054 + molecules.\cite{Orlandi:2006fk} We have used the same overlapping
1055 + ellipsoids as a way to test the behavior of our algorithm for a
1056 + structure of some interest to the materials science community,
1057 + although since we are interested in capturing only the hydrodynamic
1058 + behavior of this model, we have left out the dipolar interactions of
1059 + the original Orlandi model.
1060 +
1061 + A reference system composed of a single banana rigid body embedded in
1062 + a sea of 1929 solvent particles was created and run under standard
1063 + (microcanonical) molecular dynamics.  The resulting viscosity of this
1064 + mixture was 0.298 centipoise (as estimated using
1065 + Eq. (\ref{ldeq:shear})).  To calculate the hydrodynamic properties of
1066 + the banana rigid body model, we created a rough shell (see
1067 + Fig.~\ref{ldfig:roughShell}), in which the banana is represented as a
1068 + ``shell'' made of 3321 identical beads (0.25 \AA\ in diameter)
1069 + distributed on the surface.  Applying the procedure described in
1070 + Sec.~\ref{ldintroEquation:ResistanceTensorArbitraryOrigin}, we
1071 + identified the center of resistance, ${\bf r} = $(0 \AA, 0.81 \AA, 0
1072 + \AA).  
1073 +
1074 + The Langevin rigid-body integrator (and the hydrodynamic diffusion
1075 + tensor) are essentially quantitative for translational diffusion of
1076 + this model.  Orientational correlation times under the Langevin
1077 + rigid-body integrator are within 11\% of the values obtained from
1078 + explicit solvent, but these models also exhibit some solvent
1079 + inaccessible surface area in the explicitly-solvated case.
1080 +
1081 + \subsection{Composite sphero-ellipsoids}
1082 +
1083 + Spherical heads perched on the ends of Gay-Berne ellipsoids have been
1084 + used recently as models for lipid
1085 + molecules.\cite{SunX._jp0762020,Ayton01} A reference system composed
1086 + of a single lipid rigid body embedded in a sea of 1929 solvent
1087 + particles was created and run under a microcanonical ensemble.  The
1088 + resulting viscosity of this mixture was 0.349 centipoise (as estimated
1089 + using Eq. (\ref{ldeq:shear})).  To calculate the hydrodynamic properties
1090 + of the lipid rigid body model, we created a rough shell (see
1091 + Fig.~\ref{ldfig:roughShell}), in which the lipid is represented as a
1092 + ``shell'' made of 3550 identical beads (0.25 \AA\ in diameter)
1093 + distributed on the surface.  Applying the procedure described by
1094 + Eq. (\ref{ldintroEquation:ResistanceTensorArbitraryOrigin}), we
1095 + identified the center of resistance, ${\bf r} = $(0 \AA, 0 \AA, 1.46
1096 + \AA).
1097 +
1098 + The translational diffusion constants and rotational correlation times
1099 + obtained using the Langevin rigid-body integrator (and the
1100 + hydrodynamic tensor) are essentially quantitative when compared with
1101 + the explicit solvent simulations for this model system.  
1102 +
1103 + \subsection{Summary of comparisons with explicit solvent simulations}
1104 + The Langevin rigid-body integrator we have developed is a reliable way
1105 + to replace explicit solvent simulations in cases where the detailed
1106 + solute-solvent interactions do not greatly impact the behavior of the
1107 + solute.  As such, it has the potential to greatly increase the length
1108 + and time scales of coarse grained simulations of large solvated
1109 + molecules.  In cases where the dielectric screening of the solvent, or
1110 + specific solute-solvent interactions become important for structural
1111 + or dynamic features of the solute molecule, this integrator may be
1112 + less useful.  However, for the kinds of coarse-grained modeling that
1113 + have become popular in recent years (ellipsoidal side chains, rigid
1114 + bodies, and molecular-scale models), this integrator may prove itself
1115 + to be quite valuable.
1116 +
1117 + \begin{figure}
1118 + \centering
1119 + \includegraphics[width=\linewidth]{./figures/ldGraph}
1120 + \caption[Mean squared displacements and orientational
1121 + correlation functions for each of the model rigid bodies]{The
1122 + mean-squared displacements ($\langle r^2(t) \rangle$) and
1123 + orientational correlation functions ($C_2(t)$) for each of the model
1124 + rigid bodies studied.  The circles are the results for microcanonical
1125 + simulations with explicit solvent molecules, while the other data sets
1126 + are results for Langevin dynamics using the different hydrodynamic
1127 + tensor approximations.  The Perrin model for the ellipsoids is
1128 + considered the ``exact'' hydrodynamic behavior (this can also be said
1129 + for the translational motion of the dumbbell operating under the bead
1130 + model). In most cases, the various hydrodynamics models reproduce
1131 + each other quantitatively.}
1132 + \label{ldfig:results}
1133 + \end{figure}
1134 +
1135 + \begin{table*}
1136 + \begin{center}
1137 + \caption{TRANSLATIONAL DIFFUSION CONSTANTS (D) FOR THE MODEL SYSTEMS
1138 + CALCULATED USING MICROCANONICAL SIM\-U\-LA\-TIONS (WITH EXPLICIT
1139 + SOLVENT), THEORETICAL PREDICTIONS, AND LANGEVIN SIMULATIONS (WITH
1140 + IMPLICIT SOLVENT)}
1141 + \begin{tabular}{lccccccc}
1142 + \hline
1143 + & \multicolumn{2}c{microcanonical} & & \multicolumn{3}c{Theoretical} & Langevin \\
1144 + \cline{2-3} \cline{5-7}
1145 + model & $\eta$ (cP)  & D & & Analytical & method & Hydro &  \\
1146 + \hline
1147 + sphere    & 0.279  & 3.06 & & 2.42 & exact       & 2.42 & 2.33 \\
1148 + ellipsoid & 0.255  & 2.44 & & 2.34 & exact       & 2.34 & 2.37 \\
1149 +          & 0.255  & 2.44 & & 2.34 & rough shell & 2.36 & 2.28 \\
1150 + dumbbell  & 0.308  & 2.06 & & 1.64 & bead model  & 1.65 & 1.62 \\
1151 +          & 0.308  & 2.06 & & 1.64 & rough shell & 1.59 & 1.62 \\
1152 + banana    & 0.298  & 1.53 & &      & rough shell & 1.56 & 1.55 \\
1153 + lipid     & 0.349  & 1.41 & &      & rough shell & 1.33 & 1.32 \\
1154 + \hline
1155 + \end{tabular}
1156 + \begin{minipage}{\linewidth}
1157 + %\centering
1158 + \vspace{2mm}
1159 + Analytical solutions for the exactly-solved hydrodynamics models are
1160 + obtained from: Stokes' law (sphere), and Refs. \citen{Perrin1934} and
1161 + \citen{Perrin1936} (ellipsoid), \citen{Stimson:1926qy} and
1162 + \citen{Davis:1969uq} (dumbbell). The other model systems have no known
1163 + analytic solution.  All diffusion constants are reported in units of
1164 + $10^{-3}$ cm$^2$ / ps (= $10^{-4}$ \AA$^2$ / fs).
1165 + \label{ldtab:translation}
1166 + \end{minipage}
1167 + \end{center}
1168 + \end{table*}
1169 +
1170 + \begin{table*}
1171 + \begin{center}
1172 + \caption{ORIENTATIONAL RELAXATION TIMES ($\tau$) FOR THE MODEL SYSTEMS USING
1173 + MICROCANONICAL SIMULATION (WITH EXPLICIT SOLVENT), THEORETICAL
1174 + PREDICTIONS, AND LANGEVIN SIMULATIONS (WITH IMPLICIT SOLVENT)}
1175 + \begin{tabular}{lccccccc}
1176 + \hline
1177 + & \multicolumn{2}c{microcanonical} & & \multicolumn{3}c{Theoretical} & Langevin \\
1178 + \cline{2-3} \cline{5-7}
1179 + model & $\eta$ (cP) & $\tau$ & & Analytical & method & Hydro &  \\
1180 + \hline
1181 + sphere    & 0.279  &      & & 9.69 & exact       & 9.69 & 9.64 \\
1182 + ellipsoid & 0.255  & 46.7 & & 22.0 & exact       & 22.0 & 22.2 \\
1183 +          & 0.255  & 46.7 & & 22.0 & rough shell & 22.6 & 22.2 \\
1184 + dumbbell  & 0.308  & 14.1 & &      & bead model  & 50.0 & 50.1 \\
1185 +          & 0.308  & 14.1 & &      & rough shell & 41.5 & 41.3 \\
1186 + banana    & 0.298  & 63.8 & &      & rough shell & 70.9 & 70.9 \\
1187 + lipid     & 0.349  & 78.0 & &      & rough shell & 76.9 & 77.9 \\
1188 + \hline
1189 + \end{tabular}
1190 + \begin{minipage}{\linewidth}
1191 + %\centering
1192 + \vspace{2mm}
1193 + All relaxation times are for the rotational correlation function with
1194 + $\ell = 2$ and are reported in units of ps.  The ellipsoidal model has
1195 + an exact solution for the orientational correlation time due to
1196 + Perrin, but the other model systems have no known analytic solution.
1197 + \label{ldtab:rotation}
1198 + \end{minipage}
1199 + \end{center}
1200 + \end{table*}
1201 +
1202 + \section{Application: A rigid-body lipid bilayer}
1203 +
1204 + To test the accuracy and efficiency of the new integrator, we applied
1205 + it to study the formation of corrugated structures emerging from
1206 + simulations of the coarse grained lipid molecular models presented
1207 + above.  The initial configuration is taken from earlier molecular
1208 + dynamics studies on lipid bilayers which had used spherical
1209 + (Lennard-Jones) solvent particles and moderate (480 solvated lipid
1210 + molecules) system sizes.\cite{SunX._jp0762020} the solvent molecules
1211 + were excluded from the system and the box was replicated three times
1212 + in the x- and y- axes (giving a single simulation cell containing 4320
1213 + lipids).  The experimental value for the viscosity of water at 20C
1214 + ($\eta = 1.00$ cp) was used with the Langevin integrator to mimic the
1215 + hydrodynamic effects of the solvent.  The absence of explicit solvent
1216 + molecules and the stability of the integrator allowed us to take
1217 + timesteps of 50 fs. A simulation run time of 30 ns was sampled to
1218 + calculate structural properties.  Fig. \ref{ldfig:bilayer} shows the
1219 + configuration of the system after 30 ns.  Structural properties of the
1220 + bilayer (e.g. the head and body $P_2$ order parameters) are nearly
1221 + identical to those obtained via solvated molecular dynamics. The
1222 + ripple structure remained stable during the entire trajectory.
1223 + Compared with using explicit bead-model solvent molecules, the 30 ns
1224 + trajectory for 4320 lipids with the Langevin integrator is now {\it
1225 + faster} on the same hardware than the same length trajectory was for
1226 + the 480-lipid system previously studied.
1227 +
1228 + \begin{figure}
1229 + \centering
1230 + \includegraphics[width=\linewidth]{./figures/ldBilayer}
1231 + \caption[Snapshot of a bilayer of rigid-body models for lipids]{A
1232 + snapshot of a bilayer composed of 4320 rigid-body models for lipid
1233 + molecules evolving using the Langevin integrator described in this
1234 + work.} \label{ldfig:bilayer}
1235 + \end{figure}
1236 +
1237 + \section{Conclusions}
1238 +
1239 + We have presented a new algorithm for carrying out Langevin dynamics
1240 + simulations on complex rigid bodies by incorporating the hydrodynamic
1241 + resistance tensors for arbitrary shapes into a stable and efficient
1242 + integration scheme.  The integrator gives quantitative agreement with
1243 + both analytic and approximate hydrodynamic theories, and works
1244 + reasonably well at reproducing the solute dynamical properties
1245 + (diffusion constants, and orientational relaxation times) from
1246 + explicitly-solvated simulations.  For the cases where there are
1247 + discrepancies between our Langevin integrator and the explicit solvent
1248 + simulations, two features of molecular simulations help explain the
1249 + differences.
1250 +
1251 + First, the use of ``stick'' boundary conditions for molecular-sized
1252 + solutes in a sea of similarly-sized solvent particles may be
1253 + problematic.  We are certainly not the first group to notice this
1254 + difference between hydrodynamic theories and explicitly-solvated
1255 + molecular
1256 + simulations.\cite{Schmidt:2004fj,Schmidt:2003kx,Ravichandran:1999fk,TANG:1993lr}
1257 + The problem becomes particularly noticable in both the translational
1258 + diffusion of the spherical particles and the rotational diffusion of
1259 + the ellipsoids.  In both of these cases it is clear that the
1260 + approximations that go into hydrodynamics are the source of the error,
1261 + and not the integrator itself.
1262 +
1263 + Second, in the case of structures which have substantial surface area
1264 + that is inaccessible to solvent particles, the hydrodynamic theories
1265 + (and the Langevin integrator) may overestimate the effects of solvent
1266 + friction because they overestimate the exposed surface area of the
1267 + rigid body.  This is particularly noticable in the rotational
1268 + diffusion of the dumbbell model.  We believe that given a solvent of
1269 + known radius, it may be possible to modify the rough shell approach to
1270 + place beads on solvent-accessible surface, instead of on the geometric
1271 + surface defined by the van der Waals radii of the components of the
1272 + rigid body.  Further work to confirm the behavior of this new
1273 + approximation is ongoing.

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines