cond-mat0510223/man.tex
1: \documentclass[twocolumn,preprintnumbers,amsmath,amssymb]{revtex4}
2: %\usepackage{times}
3: \usepackage{tabularx}
4: \usepackage{rotating}
5: \usepackage{graphicx}
6: \usepackage{epsfig}   
7: \usepackage{amsmath,amssymb,bbm}
8: %\usepackage{doublespace}
9: \usepackage{anysize}
10: \usepackage{subfigure}
11: 
12: 
13: \marginsize{1in}{1in}{1in}{1in}
14: 
15: \newcommand{\parc}[2]{\frac{\partial #1}{\partial #2}} 
16: \newcommand{\units}{\, \mathrm}
17: \renewcommand{\thefootnote}{\fnsymbol{footnote}}
18: 
19: 
20: 
21: 
22: \begin{document}
23: 
24: \title{Adaptive Resolution Molecular Dynamics Simulation:\\ Changing the
25: Degrees of Freedom on the Fly}
26: 
27: 
28: \author{Matej Praprotnik}
29: \altaffiliation{On leave from National Institute of Chemistry, Hajdrihova 19,
30:                    1000 Ljubljana, Slovenia. Electronic Mail: praprot@cmm.ki.si.}
31: \author{Luigi Delle Site}
32: \author{Kurt Kremer}
33: \affiliation{%
34: Max-Planck-Institut f\"ur Polymerforschung, Ackermannweg 10, D-55128 Mainz, Germany
35: }%
36: 
37: 
38: 
39: \date{\today}% It is always \today, today,
40:              %  but any date may be explicitly specified
41: 
42: 
43: \begin{abstract}
44:  We present a new adaptive resolution
45: technique for efficient particle-based multiscale molecular
46: dynamics (MD) simulations. The presented approach is tailor-made
47: for molecular systems where atomistic resolution is required only
48: in spatially localized domains whereas a lower mesoscopic level of
49: detail is sufficient for the rest of the system. Our method allows
50: an on-the-fly interchange between a given molecule's atomic and
51: coarse-grained level of description, enabling us to reach large
52: length and time scales while spatially retaining atomistic details
53: of the system. The new approach is tested on a model system of a
54: liquid of tetrahedral molecules. The simulation box is divided
55: into two regions: one containing only atomistically resolved
56: tetrahedral molecules, the other containing only one particle
57: coarse-grained spherical molecules. The molecules can freely move
58: between the two regions while changing their level of resolution
59: accordingly. The coarse-grained and the atomistically
60: resolved systems have the same statistical properties at the same
61: physical conditions.
62: \end{abstract}
63: 
64: \maketitle
65: 
66: \section{Introduction}
67: Many problems in complex soft matter systems are inherently
68: multiscale in nature, i.e., microscopic interactions are coupled
69: strongly to meso- and macroscopic properties. Despite the
70: increasing computational power and ongoing efforts to enhance the
71: efficiency of molecular dynamics (MD) integration
72: algorithms\cite{Deuflhard:1999,Minary:2004,
73: Praprotnik:2005,Praprotnik:2005:1,Praprotnik:2005:2,Praprotnik:2004},
74: all-atom MD simulations are often incapable to cover the time and
75: length  scales needed in order to reach relaxation in a typical
76: molecular system, such as a polymer solution or melt. In many
77: cases it is also questionable, whether the huge amount of detail
78: information might not even obscure the relevant structural
79: information. On the other hand, details of the chemistry do not
80: affect universal power laws but the prefactors of these power
81: laws, which can vary by several orders of magnitude themselves.
82: Thus even on the more coarse-grained level it is advisable to keep
83: a link to the specific chemistry under
84: investigation\cite{kremer:2001,kremer:2004}. In addition typical
85: soft matter systems can be quite inhomogeneous in a way, that
86: different regions within one system are sufficiently described by
87: more or less detail. A consistent and at the same time highly
88: efficient ansatz to understand modern soft matter systems (both of
89: synthetic as well as biological origin) has to take such
90: considerations into account. One first way to tackle this, was to
91: reduce the number of degrees of freedom by a systematic
92: coarse-graining, which retains only those degrees of freedom that
93: are relevant for the particular property of interest. Examples of
94: molecular systems where the coarse-graining approach has been used
95: with much success are fluids\cite{Ayton:2004}, lipid bilayers
96: \cite{Kranenburg:2003,Nielsen:2004,Marrink:2004,Chang:2005,Cook:2005},
97: and polymer
98: systems\cite{Kremer:1990,Tschop:1998,Tschop:1998:2,Everaers:2004}.
99: 
100: 
101: Since some specific chemical details are usually lost in the
102: coarse-graining procedure, much effort has been devoted recently
103: to the development of multiscale modeling approaches, where
104: different parts of the system are modeled at different levels of
105: detail to account for the local resolution
106: requirement\cite{Chun:2000,Malevanets:2000,Villa:2005,DelleSite:2002,Abrams:2003,DelleSite:2004}.
107: In the dual-resolution modeling approach for studying the behavior
108: of polymers near metal
109: surfaces\cite{DelleSite:2002,Abrams:2003,DelleSite:2004}, for
110: example, polymer chain ends that interact with the metal surface
111: are represented partially atomistically while the remaining parts
112: of the polymers, where lower resolution is adequate, are
113: represented as bead-spring chains. However, the switching between
114: different levels of resolution, i.e, the atomic and mesoscopic, is
115: not allowed during the course of that MD simulation, and therefore
116: the initial level of detail and thereby the number of degrees of
117: freedom in the system remain unchanged. Since the chain ends that
118: interact with the metal typically remain close to the surface and
119: they only contain a small fraction of the whole system, an
120: adaptive on-the-fly change of the molecules' resolution is not
121: strictly required for this class of systems. Another approach
122: reported in the literature concerns the link between quantum
123: mechanical and classical MD simulations. In this QM/MM approach a
124: small subset of the system is defined and considered as quantum
125: mechanical, while the rest is treated by a classical force field
126: simulation. Here the atoms as well as the regions of the
127: different regimes are fixed from the very beginning
128: \cite{Laio:2002}, restricting the application to rather specific
129: cases.
130: 
131: 
132: In contrast, MD simulations, in which the spatially localized
133: atomistic domains frequently exchange particles with the remaining
134: mesoscopic part of the system, would allow for much wider
135: applications. Then it would be possible to define certain areas or
136: develop criteria for certain situations, which ask for a more
137: detailed view, while the rest of the system can be treated on the
138: more coarse-grained level. It is the purpose of this paper to
139: present a first attempt of an MD simulation of that kind. Because
140: of that we here at first restrict ourselves to a most elementary
141: model system. Of course one also could resort to Monte Carlo
142: simulations, where different levels of detail are combined. This
143: actually would be somewhat simpler because of the purely
144: stochastic nature of this simulation method. Since we are
145: eventually aiming at molecular systems, where collective motions
146: are crucial, we decided to stick to the MD approach. Existing
147: hybrid MD methods that concurrently couple different length scales
148: have been developed to study solid state systems, where atomistic
149: MD was either combined with the finite elements
150: method\cite{Rafii:1998,Broughton:1999,Smirnova:1999} or it was
151: linked to a quantum mechanical model\cite{Csanyi:2004}. To our
152: knowledge, however, in pursuit of this objective no adaptive
153: hybrid atomistic/mesoscale particle-based MD method, which would
154: allow to dynamically adjust the level of detail, which means the
155: adjustment of the degrees of freedom in the system, has been
156: developed so far.
157: 
158: 
159: In this paper we present a novel adaptive resolution MD scheme
160: that combines a full atomistic description of a desired region of
161: the system with a mesoscale treatment of the remaining part. The
162: key feature of the new method is that it allows to dynamically
163: adapt the level of a given molecule's resolution according to its
164: position in the system. Hence, the number of degrees of freedom is
165: not a conserved quantity in our MD simulations. Furthermore, the
166: presented method is not restricted to couple only the atomistic
167: and mesoscopic levels of detail but can also be applied to systems
168: with mesoscopic domains that are described at different levels of
169: coarse-graining. Here we present a first test case, showing that
170: such an approach is feasible. Therefore the new approach is tested
171: for a simple model liquid consisting of tetrahedral molecules. The
172: simulation box is divided into two regions: one containing
173: ''atomistically'' resolved tetrahedral molecules, the other
174: containing coarse-grained spherical molecules. Molecules are
175: allowed to freely move between the two regions while changing
176: their level of resolution accordingly. The results show that the
177: statistical properties of the corresponding fully atomistic system
178: are accurately reproduced by using the proposed hybrid scheme. In
179: particular, gradients in the chemical potential across the
180: artificial interface where the resolution changes and
181: corresponding spurious fluxes can be avoided. Although we applied
182: the new method here to a generic test system it should find
183: application in more realistic physical systems. There has been an
184: initial attempt reported in Ref. \cite{Abrams:2004}, where MD
185: simulation of an inhomogeneously coarse-grained system of liquid
186: methane has been described. There the starting point were two
187: established models for methane, a five site atomistic and a one
188: site spherical where the interaction between molecules of
189: different species was derived by standard Lennard-Jones mixing
190: rules with the hydrogens of the atomistic invisible to the
191: spherical molecules. In this case it turned out that an effective
192: flux between the two different regimes occurred. This effect is
193: partially due to the different equilibrium state points described
194: by the two models, but in any case in that approach no effective
195: potential between coarse-grained molecules was derived. Our
196: approach differs from that since here we derive such potentials in
197: such a way that the two regimes are in true thermodynamic
198: equilibrium.
199: 
200: 
201: The organization of the article is as follows. In section 2 the
202: methodology is presented, whereby in the first step a
203: coarse-grained model is derived and parameterized from a fully
204: atomistic system and then, in the second step, the atomic and
205: mesoscopic length scales are systematically coupled in a hybrid
206: atomistic/mesoscopic model. The computational details are given in
207: section 3. The results and discussion are presented in section 4,
208: followed by conclusions in section 5.
209: 
210: \section{Methodology}
211: 
212: In this section, the model systems are described and the
213: methodology for the adaptive multiscale MD simulations is
214: presented.
215: 
216: \subsection*{Models}
217: 
218: \subsubsection*{All-Atom Model}
219: 
220: First, we introduce our reference explicit all-atom (\emph{ex})
221: model. Consider a system of $n$ tetrahedral molecules consisting
222: of $N=4$ atoms of the same mass $m_0$ connected by anharmonic
223: bonds as presented in figure \ref{Fig.1} (a) (consider only the
224: right red molecule). 
225: \begin{figure}[!ht]
226: \centering
227: \subfigure[]{\includegraphics[width=7.5cm]{fig1a.ps}}\\
228: \subfigure[]{\includegraphics[width=7.5cm]{fig1b.ps}}
229: \caption{(a) The on-the-fly interchange between the atomic and coarse-grained
230: levels of description. The middle hybrid molecule is a linear
231:  combination of fully atomistic tetrahedral molecule with an additional center of mass
232:  particle representing the coarse-grained molecule. (b) Snapshot of the hybrid atomistic/mesoscopic
233:  model at $\rho^*=0.1$ and $T^*=1$ (LJ units). The red
234: molecules are the explicit atomistically resolved tetrahedral
235: molecules, the blue molecules are the corresponding one particle
236: coarse-grained molecules.}\label{Fig.1}
237: \end{figure}
238: 
239: All atoms in the system interact according to
240: a purely repulsive shifted $12$-$6$ Lennard-Jones potential with a
241: cutoff at $2^{1/6}\sigma$:
242: \begin{multline}
243:  U_{LJ}^{atom}(r_{i\alpha j\beta})=\\\left\{\begin{array}{rc}
244: 4\varepsilon\bigl[\bigl(\frac{\sigma}{r_{i\alpha
245:       j\beta}}\bigr)^{12}-\bigl(\frac{\sigma}{r_{i\alpha j\beta}}\bigr)^6+\frac{1}{4}\bigr];
246: & r_{i\alpha j\beta}\le 2^{1/6}\sigma\\
247:                           0; & r_{i\alpha j\beta}> 2^{1/6}\sigma
248:                              \end{array}\right.\label{eq.8}
249: \end{multline}
250: where $r_{i\alpha j\beta}$ is the distance between the atom $i\alpha$ of the molecule
251: $\alpha$ and the atom $j\beta$ of the molecule $\beta$. We use
252: $\varepsilon$ as a unit of energy. All atoms have the same excluded
253: volume diameter $\sigma$, where $\sigma$ is the unit of length.
254: The neighboring atoms in a given
255: molecule $\alpha$ are linked via an attractive FENE potential
256: \begin{multline}
257:  U_{FENE}^{atom}(r_{i\alpha j\alpha})=\\\left\{\begin{array}{rc}
258:                   -\frac{1}{2}kR_0^2\ln\bigl[1-\bigl(\frac{r_{i\alpha
259:  j\alpha}}{R_0}\bigl)^2\bigr]; & r_{i\alpha j\alpha}\le R_0\\
260:                        \infty; & r_{i\alpha j\alpha}> R_0
261:                   \end{array}\right.\label{eq.9}
262: \end{multline}
263: with divergence length  $R_0=1.5\sigma$ and stiffness
264: $k=30\varepsilon/\sigma^2$, so that the average bond length is
265: approximately $1.0\sigma$ for $k_BT=\varepsilon$, where $T$ is the
266: temperature of the system and $k_B$ is Boltzmann's constant.
267: The functional form of these potentials and their parameters are
268: the same as usually employed in polymer MD simulations\cite{Kremer:1990}.
269: 
270: \subsubsection*{Coarse-Grained Model}
271: 
272: Next, we map the atomistic model to a coarse-grained (\emph{cg})
273: mesoscopic model. For the latter we have chosen a system composed
274: of $n$ one-particle molecules schematically depicted in figure
275: \ref{Fig.1} (a) (consider only the left blue molecule). A given
276: coarse-grained molecule $\alpha$ in the system has a mass
277: $M_\alpha=4m_0$ equal to the total mass of the explicit
278: tetrahedral molecule. All rotational and vibrational degrees of
279: freedom of atomistically resolved tetrahedral molecules are thus
280: removed, and the number of nonbonded interactions is strongly
281: decreased as well.
282: 
283: We shall now find an effective pair potential between
284: coarse-grained molecules such that the structural properties of
285: the underlying atomistic model are reproduced. There is usually no
286: unique way to coarse-grain to an effective pair potential, which
287: is in general temperature and density dependent
288: \cite{Louis:2002,Curtarolo:2002,Klapp:2004}, just as in
289: statistical mechanics there are different ways to perform a
290: renormalization group step. Here, we extract the effective pair
291: potential from a center-of-mass radial distribution function
292: (RDF$_{cm}$) of the reference atomistic model using the potential
293: of mean force $PMF(r)$ as
294:  \begin{equation}
295:   U^{cm}(r)\approx PMF(r)=-k_BT\log g^{cm}_{ex}(r),\label{eq.0}
296: \end{equation}
297: where $g^{cm}_{ex}(r)$ is the RDF$_{cm}$ of the all-atom system
298: and $U^{cm}(r)$ is the derived effective pair
299: potential\cite{Reith:2003}. The effective potential obtained in
300: this way is correct only in the limit of zero density, where the
301: many-body contributions vanish. For systems with nonzero densities
302: in principle many-body interactions would be needed just as for
303: the classical renormalization group theory in statistical
304: mechanics. In a similar spirit we here resort to the (expected)
305: relevant part in order to gain a significant speed up in our
306: simulations. Because of that we use the PMF as the initial guess
307: for the effective pair potential in systems with nonzero density.
308: Then this is further fine-tuned until the RDF$_{cm}$s and 
309: pressures of the reference atomistic and coarse-grained systems
310: match\cite{Reith:2003}. As it turns out (see the Results and
311: Discussion section) the effective pair potential acting between
312: our coarse-grained molecules is significantly softer than the pair
313: potential between atoms of the resolved molecules in accordance
314: with the results previously found in the
315: literature\cite{Klapp:2004,Bryant:2002}.
316: 
317: 
318: 
319: \subsubsection*{Transition Regime: Hybrid Atomistic/Mesoscopic Model}
320: Let us now introduce a hybrid explicit/coarse-grained
321: (\emph{ex-cg}) model. Consider a box of $n$ molecules where one
322: half of the box is occupied by atomistically resolved four-atom
323: tetrahedral molecules while the other half is filled up with the
324: same number of corresponding coarse-grained one-particle molecules
325: as schematically presented in figure \ref{Fig.1} (b). The two
326: domains exchange molecules which adapt their level of resolution
327: accordingly. To ensure that the transition between the two
328: different levels of description is smooth, i.e., the rotational and
329: vibrational degrees of freedom of a given molecule are gradually
330: 'switched on/off' as it crosses the boundary between the atomistic
331: and coarse-grained domains, we also introduce an interface layer
332: for handshaking between atomistic and mesoscale regions. In this
333: regime also the rotational and vibrational velocities have to be
334: reintroduced in a consistent way. Due to the periodic boundary
335: conditions employed in our simulations there are in fact two such
336: layers as depicted in figure \ref{Fig.1} (b).
337: 
338: Since the total number of molecules $n$ in the system is a
339: conserved quantity in our constant temperature simulations, we
340: sample the phase space from the $nVT$ ensemble. However, the total
341: number of degrees of freedom is not constant in this model.
342: 
343: An alternative way compared to the similarities with
344: renormalization group methods, which probably describes the
345: situation even better, is the comparison to a first order phase
346: transition. The rotational and vibrational part of the free energy
347: then can be viewed as the latent heat at this transition.  At
348: equilibrium, conditions analogous to two phase coexistence,
349: \begin{equation}
350:        \mu_{ex}=\mu_{cg},~~~ p_{ex}=p_{cg},~~~ T_{ex}=T_{cg},\label{eq.1}
351: \end{equation}
352: must be automatically satisfied, where $\mu_{ex}$, $p_{ex}$,
353: $T_{ex}$ and $\mu_{cg}$, $p_{cg}$, $T_{cg}$ are the chemical
354: potentials, pressures, and temperatures of the liquid in the
355: atomistic and coarse-grained domains, respectively. These
356: conditions (\ref{eq.1}) assure that there is no net flux of
357: molecules between the atomistic and coarse-grained regions. To
358: keep this absolute requirement also then defines a central task.
359: This guarantees that the liquid is homogeneous across the box as
360: it is in the reference fully atomistic system. From a
361: molecular point of view, the artificial resolution boundary must
362: be essentially invisible, i.e., the molecules have to cross the border
363: without experiencing any `barrier'. Our approach to reach this
364: objective is presented in the proceeding subsection.
365: 
366: 
367: \subsection*{Adaptive Resolution Scheme}
368: 
369: 
370: To allow a coarse-grained molecule to find an energetically
371: permissible orientation  with respect to its neighboring
372: mo\-le\-cu\-les (when it leaves the coarse-grained domain and is
373: remapped into the atomistically resolved four-atom tetrahedral
374: molecule) we introduce an interface layer between the atomistic
375: and coarse-grained regions, which contains 'hybrid' molecules as
376: presented in figure \ref{Fig.1}. Each hybrid molecule
377: schematically shown in figure \ref{Fig.1}(a) (consider the middle
378: molecule) is composed of a tetrahedral molecule with an additional
379: massless center-of-mass particle serving as an interaction site.
380: This is similar to the flexible TIP4P water
381: model\cite{Lawrence:2003} where apart from the interaction sites
382: on the three atoms of a water molecule an additional interaction
383: site is introduced along the symmetry axis between the hydrogen
384: and oxygen atoms.
385: 
386: Thus, each time a coarse-grained molecule $\alpha$ leaves that
387: domain and enters the interface layer, it is remapped first into a
388: hybrid molecule with the same center-of-mass position and a random
389: orientation in space, where the relative positions of the
390: tetrahedral atoms are taken from a molecular configuration
391: corresponding to a randomly chosen molecule from the atomistic
392: regime. Each of the four explicit tetrahedral atoms in the hybrid
393: molecule gains at this remapping a velocity equal to the velocity
394: of the corresponding coarse-grained molecule to maintain the
395: linear momentum of the molecule. In addition, the tetrahedral
396: atoms are also assigned rotational/vibrational velocities
397: corresponding to atoms of a random molecule from the atomistic
398: region, where we subtract the total linear momentum of the latter
399: molecule. In this way we ensure that the kinetic energy is
400: distributed among all degrees of freedom according to the
401: equipartition principle as $k_BT/2$ of average kinetic energy per
402: quadratic degree of freedom while retaining the linear momentum of
403: the coarse-grained molecule. The center-of-mass interaction site
404: moves obeying the constraints:
405: \begin{eqnarray}
406: {\bf R}_\alpha&=&\frac{\sum_{i\alpha} m_{i\alpha}{\bf
407:     r}_{i\alpha}}{M_\alpha},\label{eq.5}\\
408:  {\bf V}_\alpha&=&\frac{\sum_{i\alpha} m_{i\alpha}{\bf v}_{i\alpha}}{M_\alpha},\label{eq.6}
409: \end{eqnarray}
410: where ${\bf R}_\alpha$ is a center of mass of the molecule $\alpha$,
411: ${\bf r}_{i\alpha}$ is the position vector of the explicit
412: tetrahedral atom $i\alpha$ in the molecule $\alpha$, ${\bf
413: V}_\alpha$ is the center-of-mass velocity of the molecule $\alpha$,
414: ${\bf v}_{i\alpha}$ is the velocity of the explicit tetrahedral
415: atom $i\alpha$, and $M_\alpha=\sum_{i\alpha}m_{i\alpha}$ is the
416: total mass of the molecule $\alpha$. In our case,
417: $m_{i\alpha}=m_0$ and $M_\alpha=4m_0$ for all $i\alpha=1,\dots,4$
418: and $\alpha=1,\dots, n$. Each time a hybrid molecule crosses the
419: boundary into atomistic regime it is remapped into a four-particle
420: tetrahedral molecule with the four tetrahedral atoms retaining
421: their current velocities and positions. In this model also the
422: explicit tetrahedral molecules have the center-of-mass interaction
423: sites, but only for the interactions with the hybrid and
424: coarse-grained molecules. Of course, deep in the atomistic region,
425: where the atomistically resolved molecules do not interact
426: anymore with the hybrid and coarse-grained molecules the
427: center-of-mass interaction site can be omitted.  Every time a
428: four-particle tetrahedral molecule leaves the atomistic region and
429: enters into the transition regime it is mapped into a hybrid
430: molecule with the four tetrahedral atoms retaining instantaneous
431: velocities and positions with the center-of-mass interaction site
432: moving according to Eqs. (\ref{eq.5}) and (\ref{eq.6}). Similarly,
433: as a hybrid molecule crosses a boundary to the coarse-grained
434: region it is mapped into a coarse-grained molecule with a velocity
435: equal to the center-of-mass velocity of the hybrid molecule given by Eq.
436: (\ref{eq.6}).
437: 
438: To couple the atomic and mesoscopic length scales we define in the
439: spirit of thermodynamic perturbation
440: approach\cite{Zwanzig:1954,Leach:2001} the total intermolecular
441: force acting between centers of mass of molecules $\alpha$ and
442: $\beta$ as
443: \begin{multline}
444:  {\bf F}_{\alpha\beta}=\\w(X_\alpha)w(X_\beta){\bf
445:  F}_{\alpha\beta}^{atom}+[1-w(X_\alpha)w(X_\beta)]{\bf
446:  F}_{\alpha\beta}^{cm},\label{eq.4}
447: \end{multline}
448: where
449: \begin{equation}
450: {\bf F}_{\alpha\beta}^{atom}=\sum_{i\alpha, j\beta}{\bf F}_{i\alpha\label{eq.4a}
451:  j\beta}^{atom}
452: \end{equation}
453: is the sum of all pair atom interactions between explicit
454: tetrahedral atoms of the molecule $\alpha$ and explicit tetrahedral
455: atoms of the molecule $\beta$ and
456: \begin{eqnarray}
457:  {\bf F}_{i\alpha j\beta}^{atom}&=&- \parc{U^{atom}}{{\bf r}_{i\alpha j\beta}},\\
458:  {\bf F}_{\alpha\beta}^{cm}&=&-\parc{U^{cm}}{{\bf R}_{\alpha\beta}}.
459: \end{eqnarray}
460: The vector ${\bf r}_{i\alpha j\beta}={\bf r}_{i\alpha}-{\bf
461: r}_{j\beta}$ is the relative position vector of atoms $i\alpha$
462: and $j\beta$, ${\bf R}_{\alpha\beta}={\bf R}_{\alpha}-{\bf
463: R}_{\beta}$ is the relative position vector of the centers of mass
464: of the molecules $\alpha$ and $\beta$, $X_\alpha$ and $X_\beta$ are
465: the $x$ center-of-mass coordinates of the molecules $\alpha$ and
466: $\beta$, respectively, and $w$ is the weighting function that
467: determines the 'identity' of a given molecule. The weighting
468: function $w\in [0,1]$ is defined in such a way that values $0<w<1$
469: correspond to a hybrid molecule with extreme cases $w=1$ and $w=0$
470: corresponding to a four-atom tetrahedral molecule and one-particle
471: coarse-grained molecule, respectively. Hence, as soon as one of
472: the two interacting molecules $\alpha$ and $\beta$ is a
473: coarse-grained molecule with no explicit tetrahedral atoms
474: $w(X_\alpha)w(X_\beta)=0$ and ${\bf F}_{\alpha\beta}={\bf
475: F}_{\alpha\beta}^{cm}$. 
476: 
477: We propose the following functional form of the weighting function
478: $w$:
479: \begin{multline}
480:  w(x)=\\\left\{\begin{array}{rc}
481:                                                    1; & d<x\le \frac{a}{2}-d\\
482:                                                    0; & -\frac{a}{2}+d\le x<-d\\
483:                          \sin^2[\frac{\pi}{4d}(x+d)]; & -d\le x\le d\\
484:              \cos^2[\frac{\pi}{4d}(x-\frac{a}{2}+d)]; & \frac{a}{2}-d<x\le\frac{a}{2}\\
485:              \cos^2[\frac{\pi}{4d}(x+\frac{a}{2}+d)]; & -\frac{a}{2}\le x<-\frac{a}{2}+d
486:             \end{array}\right.\label{eq.7}
487: \end{multline}
488: where $a$ is the box length and $d$ the half-width of the interface
489: layer. The weighting function $w$ is shown in figure \ref{Fig.2}. Our
490: choice, which takes into account the periodic boundary conditions,
491:  is a particularly simple way to ensure an interpolation between
492: $w=0$ and $w=1$ that is monotonic, continuous, differentiable and
493: has zero slope at the boundaries to the atomistic and
494: coarse-grained regions. Apart from these requirements we consider
495: the precise functional form as immaterial.
496: \begin{figure}[!ht]
497: \centering
498: \mbox{\includegraphics[width=7.5cm, angle=0]{fig2.ps}}
499:          \caption{ The weighting function $w(x)\in [0,1]$ defined by
500:   Eq. (\ref{eq.7}). The values $w=1$ and $w=0$ correspond to the atomistic
501:   and coarse-grained regions of the hybrid atomistic/mesoscopic system
502:   with the box length $a$, respectively, while the values $0<w<1$ correspond to
503:   the interface layer. Shown is the example where the half-width $d$ of
504:   the interface layer is $d=a/10$. The vertical lines denote the
505:   boundaries of the interface layers.}\label{Fig.2}
506: \end{figure}
507: 
508: 
509: 
510: Exploiting the analogy with the quantum mechanical mixed state
511: description\cite{Schwabl:1995}, one can consider a hybrid molecule
512: in the interface layer as a normalized linear combination of a
513: four-atom tetrahedral molecule and a corresponding one-particle
514: coarse-grained molecule. As a given molecule moves from the
515: coarse-grained boundary of the interface layer to the atomistic
516: boundary, $w$ is gradually changed from $0$ to $1$ and a
517: coarse-grained molecule with only $3$ translational degrees of
518: freedom gradually turns into an atomistically resolved molecule
519: with additional $3N-3=9$ rotational and vibrational degrees of
520: freedom and a defined spatial orientation. The continuous, i.e.,
521: not instantaneous, 'identity' transition is required since a
522: hybrid molecule is given a random orientation at the
523: coarse-grained boundary, and there can consequently be overlaps of
524: its tetrahedral atoms with the atoms of the neighboring molecules.
525: Since at this boundary $w=0$ and the repulsive potential
526: $U^{atom}$ given in Eq. (\ref{eq.8}) is capped (see the
527: Computational Details section), the forces acting on these atoms
528: cannot diverge, however. While moving towards the atomistic region
529: the hybrid molecule slowly adapts its orientation via the
530: gradually increasing atomistic interactions with the neighboring
531: molecules. Likewise, as presented in figure \ref{Fig.1} (a), as a
532: given molecule moves from the atomistic boundary of the interface
533: layer to the coarse-grained boundary, $w$ is continously changing
534: from $1$ to $0$, the fully atomistically resolved molecule
535: gradually turns into the one particle coarse-grained molecule
536: while omitting all rotational and vibrational degrees of freedom
537: and orientation.
538: 
539: 
540: 
541: To elucidate the definition of the force calculation in our
542: model we rewrite Eq. (\ref{eq.4}) as
543: \begin{multline}
544: {\bf F}_{\alpha\beta}=\\
545: w(X_\alpha)w(X_\beta){\bf F}_{\alpha \beta}^{atom}+[1-w(X_\alpha)][1-w(X_\beta)]{\bf
546:   F}_{\alpha\beta}^{cm}\\+[1-w(X_\alpha)]w(X_\beta){\bf
547:   F}_{\alpha\beta}^{cm}+w(X_\alpha)[1-w(X_\beta)]{\bf F}_{\alpha\beta}^{cm}.\label{eq.3}
548: \end{multline}
549: From Eqs. (\ref{eq.4a}) and (\ref{eq.3}) one can then deduce that
550: the pair force in Eq. (\ref{eq.4}) is defined in such a way that
551: two given atoms $i\alpha$ and $j\beta$ in given explicit molecules
552: $\alpha$ and $\beta$ ($w(X_\alpha)=1$ and $w(X_\beta)=1$) interact
553: via the atomistic potential defined by Eqs. (\ref{eq.8}) and
554: (\ref{eq.9}) while two coarse-grained molecules ($w(X_\alpha)=0$
555: and $w(X_\beta)=0$) interact via the corresponding effective pair
556: potential $U^{cm}$. Furthermore, the coarse-grained molecules
557: 'see' the fully atomistically resolved molecules as coarse-grained
558: molecules. Hence, their intermolecular interaction is defined by
559: the effective pair potential $U^{cm}$. To ensure that the
560: center-of-mass dynamics governed by Eqs. (\ref{eq.5}) and
561: (\ref{eq.6}) of a given explicit or hybrid molecule $\alpha$ is
562: correct, the total intermolecular force ${\bf
563: F}_{\alpha\beta}^{cm}$ between the  atomistically resolved
564: molecule $\alpha$ and a coarse-grained molecule $\beta$ is
565: distributed among the explicit atoms of the molecule $\alpha$ as
566: \begin{equation}
567:  {\bf F}_{i\alpha\beta}=\frac{m_{i\alpha}}{\sum_{i\alpha}m_{i\alpha}}{\bf F}_{\alpha\beta}^{cm},\label{eq.4b}
568: \end{equation}
569: where  ${\bf F}_{i\alpha\beta}$ is the force imposed on the
570: explicit tetrahedral atom $i\alpha$ by the coarse-grained molecule
571: $\beta$. The explicit tetrahedral atoms in a given hybrid molecule
572: interact with other explicit atoms in neighboring explicit and
573: hybrid molecules through atomistic forces, while the then massless
574: center-of-mass particle serves as an effective potential
575: interaction site. The total force on a hybrid molecule is then
576: according to Eq. (\ref{eq.4}) a normalized linear combination of
577: atomistic and effective pair forces. 
578: Using
579: \begin{eqnarray}
580:   {\bf F}_{i\alpha j\beta}^{atom}&=&-{\bf F}_{j\beta i\alpha}^{atom},\\
581:   {\bf F}_{\alpha\beta}^{cm}&=&-{\bf F}_{\beta\alpha}^{cm}
582: \end{eqnarray}
583: we obtain from Eq. (\ref{eq.4})
584: \begin{equation}
585: {\bf F}_{\alpha \beta}=-{\bf F}_{\beta \alpha}.
586: \end{equation}
587: The force definition in Eq. (\ref{eq.4}) hence satisfies Newton's Third
588: Law.
589: 
590: Recall that the effective intermolecular potential is determined in such a way that the 
591: equations of state for the \emph{ex} and \emph{cg} models match around the state
592: point considered. Therefore, following the scheme, as given by Eq.
593: (\ref{eq.4}), Eq. (\ref{eq.1}) is implicitly
594: satisfied, due to which spurious fluxes are avoided at the boundary between
595: the atomistic and coarse-grained regimes.
596: 
597: 
598: To summarize, the new adaptive resolution scheme for the hybrid
599: atomistic/mesoscale MD simulations is a two-stage procedure:
600: \begin{enumerate}
601:  \item{Derive the effective
602: pair potential $U^{cm}$ between coarse-grained molecules on the basis of the
603: reference all-atom system.}
604: \item{Introduce the interface layer containing the hybrid molecules that have
605:   additional interaction sites positioned at their centers of mass. 
606:   Define a weighting function $w$ by Eq. (\ref{eq.7}) and
607:   use Eq. (\ref{eq.4}) for the definition of the intermolecular pair forces.
608:   Allow molecules to adapt their level of resolution according to
609:   their position in the system as explained in the second paragraph of
610:   this subsection.}
611: \end{enumerate}
612: Finally, since the switching of the resolution can be considered as a first
613: order phase transition, the adaptive resolution scheme must
614: necessarily be employed in combination with a thermostat. 
615: Because the latent heat is generated in the
616: transition regime it is important to use a thermostat, which 
617: couples locally to the particle motion, e.g., Langevin or Dissipative Particle Dynamics (DPD) thermostats\cite{Soddemann:2003}.
618: 
619: 
620: \section{Computational Details}
621: 
622: 
623: \subsection*{Temperature Calculation}
624: 
625: In order to treat all $n$ molecules equally regardless of the
626: level of detail, we define, using the equipartition principle, the
627: temperature of a system only from translational degrees of
628: freedom. The 'translational' temperature
629: \begin{equation}
630:  T=\frac{2E_k^{cm}}{3nk_B}=\frac{1}{3nk_B}\sum_\alpha M_\alpha{\bf V}^2_\alpha,\label{eq.12}
631: \end{equation}
632: when averaged gives the temperature of the system
633: \cite{Allen:1987}. Here $E_k^{cm}$ is the total translational
634: kinetic energy of the system.  We also checked the ``particle''
635: temperature in the explicit atomistic region. As to be expected we
636: find the same temperature.
637: 
638: \subsection*{Pressure Calculation}
639: 
640: For the same reason also the pressure calculation is based on the
641: molecular instead of atomic interactions:
642: \begin{equation}
643:  p=\frac{1}{V}\biggl[nk_BT-\frac{1}{3}\sum_{\alpha<\beta}{\bf R}_{\alpha\beta}\cdot{\bf F}_{\alpha\beta}\biggr],\label{eq.12a}
644: \end{equation}
645: where ${\bf F}_{\alpha\beta}$ is given by Eqs.
646: (\ref{eq.4})\cite{Berendsen:1984, Allen:1987}. Moreover, using Eq.
647: (\ref{eq.12a}) for pressure evaluation has two additional
648: advantages compared to the pressure calculation based on atomic
649: interactions:  first, Eq. (\ref{eq.12a}) is also valid in the case
650: that forces on atoms involve internal non-pairwise-additive
651: contributions, and second, even if all interactions are pairwise
652: additive (as in our case), the pressure calculation based on
653: atomic interactions introduces additional fluctuations in the
654: pressure\cite{Berendsen:1984}.
655: 
656: 
657: 
658: 
659: \subsection*{Multiscale Simulation Details}
660: 
661: 
662: For computational convenience during the proof-of-principle stage
663: we replace our \emph{ex-cg} model with a model in which the
664: whole box contains exclusively hybrid molecules with four explicit
665: atoms and a center-of-mass interaction site. This has some
666: technical advantages during the tests of the method. The true
667: level of detail of the molecules is then determined from the value
668: of $w(X_{\alpha})$. For later large scale, however, of course the
669: original \emph{ex-cg} model composed of the explicit,
670: coarse-grained, and hybrid molecules will be employed. 
671: Using Eq. (\ref{eq.7}) in
672: the replacement model for distinguishing between explicit,
673: coarse-grained, and hybrid molecules of our original \emph{ex-cg}
674: model we can capture all definitions of pair forces for all
675: classes of particles in our system in a single expression as
676: \begin{multline}
677:  {\bf F}_{i\alpha j\beta}=\\w(X_\alpha)w(X_\beta){\bf F}_{i\alpha
678:  j\beta}^{atom}+[1-w(X_\alpha)w(X_\beta)]\delta_{\alpha,\beta}{\bf F}_{i\alpha
679:  j\beta}^{atom}\\+[1-w(X_\alpha)w(X_\beta)]\frac{m_{i\alpha}m_{j\beta}}{\sum_{i\alpha}m_{i\alpha}\sum_{j\beta}m_{j\beta}}{\bf F}_{\alpha\beta}^{cm},\label{eq.2}
680: \end{multline}
681: where ${\bf F}_{i\alpha j\beta}$ is the total pair force between
682: the explicit atom $i\alpha$ of the molecule $\alpha$ and the
683: explicit atom $j\beta$ of the molecule $\beta$ and
684: $\delta_{\alpha,\beta}$ is the Kronecker symbol. Summing for
685: $\alpha\ne \beta$
686: \begin{equation}
687: {\bf F}_{\alpha\beta}=\sum_{i\alpha, j\beta}{\bf F}_{i\alpha j\beta}
688: \end{equation}
689: we regain the total force between molecules $\alpha$ and $\beta$
690: given in Eq. (\ref{eq.4}). From Eqs. (\ref{eq.5}), (\ref{eq.6}),
691: and (\ref{eq.2}) follows that in the replacement model with only
692: hybrid molecules a hybrid molecule experiences only translational
693: kicks from other molecules in the coarse-grained region ($w=0$)
694: and hence its center of mass moves exactly as the respective
695: one-particle coarse-grained molecule in the original \emph{ex-cg}
696: model. Similarly, in the explicit region ($w=1$) a hybrid molecule
697: experiences only atomistic forces and hence its explicit atoms
698: move exactly as the explicit atoms in the respective tetrahedral molecule in
699: the original model. Therefore, the model containing only hybrid
700: molecules interacting via the pair force defined by Eq.
701: (\ref{eq.2}) together with the applied Langevin
702: thermostat\cite{Kremer:1990} acting on each particle in the system
703: (to assure that the atom velocities are thermalized in accordance
704: with the equipartition principle) exactly mimics the original
705: \emph{ex-cg} model in which the temperature would also be held
706: constant by the Langevin thermostat. From the methodology
707: development point of view, these two models are therefore
708: identical.
709: 
710: 
711: This yields the Langevin equation of motion
712: \begin{equation}
713:  m_i\frac{d^2{\bf r_i}}{dt^2}={\bf F}_i-m_i\Gamma\frac{d{\bf
714:  r}_i}{dt}+{\bf W}_i(t),
715: \end{equation}
716: where $m_i$ is the mass of particle $i$, ${\bf F}_i$ is the total
717: force acting on the respective particle equal to the sum of pair
718: interactions given by Eq. (\ref{eq.2}), $\Gamma$ is the friction
719: constant, and ${\bf W}$ is the random force of a heat
720: bath\cite{Mann:2004}. We sample the random force from a uniform
721: distribution, since it has been shown that there is no advantage
722: of using Gaussian noise for the Langevin
723: thermostat\cite{Duenweg:1991}.
724: 
725: The value of the friction constant used in our simulations is
726: $\Gamma=0.5\tau^{-1}$ where
727: $\tau=(\varepsilon/m_0\sigma^2)^{-1/2}$. The equations of motion
728: are integrated for each particle of the system using the velocity
729: Verlet algorithm with a $0.005\tau$ time step. Here, again only
730: for the purpose of testing the method, we use only one time step
731: in the whole system. In the coarse-grained regime actually a
732: significantly larger time step could be used. Therefore ultimately
733: one would like to introduce a multiple time step algorithm.
734: Simulations are performed at temperature $T=\varepsilon/k_B$ and
735: number density $\rho=n/V=0.1/\sigma^3$. Here $n=5001$ is the
736: number of molecules in the system, which can be either the
737: explicit, coarse-grained or hybrid. If we roughly estimate the
738: excluded volume diameter of the coarse-grained molecule
739: $\sigma_{CG}$ as the distance, where the repulsive effective pair
740: potential between the coarse-grained molecules in our simulations
741: equals $k_BT$, i.e., $\sigma_{CG}\approx 1.7\sigma$, then the
742: number density
743: $\rho=0.1/\sigma^3=0.1(\sigma_{CG}/\sigma)^3/\sigma_{CG}^3\approx
744: 0.5/\sigma_{CG}^3$ corresponds to a medium dense liquid, which is
745: due to the soft effective repulsive interactions rather weakly
746: correlated\cite{Kremer:2005}.
747: %For
748: %comparison a density of a typical dense Lennard-Jones liquid is $
749: %\rho=0.85/\sigma_{CG}^3$.
750: Periodic boundary conditions and the minimum image
751: convention\cite{Allen:1987} are employed. The interaction range in
752: the system is given by the range of the effective pair potential
753: between molecules  and the geometry of tetrahedral molecules, i.e.,
754: the most outer atoms of two tetrahedral molecules with centers of
755: mass slightly less than $2.31\sigma$ apart still experience the
756: effective potential contribution in Eq. (\ref{eq.2}). Hence, the
757: actual interaction range in the system is approximately
758: $3.5\sigma$.
759: 
760: All molecules are initially randomly placed in a cubic box of size
761: $a=36.845\sigma$. To remove the overlaps between them, a $50\tau$
762: long warm-up run is performed during which the repulsive
763: interparticle potential is capped (see capped interactions in Ref.
764: \cite{Espresso:2005}). Thus, at all interparticle distances,
765: which would lead to larger forces between particles than a
766: prescribed maximal force, the forces defined by the original
767: repulsive pair potential are replaced by repulsive central forces
768: of the maximal force magnitude. The latter is gradually increased
769: from $20\varepsilon/\sigma$ to $110\varepsilon/\sigma$ during this
770: warm-up phase. Afterwards an additional $250\tau$ equilibration
771: run is carried out where we set the maximal force magnitude to
772: $10^{9}\varepsilon/\sigma$, which corresponds to interparticle
773: distance of $0.27\sigma$. The chosen maximal force magnitude value
774: is so high that it has no effect on the dynamics of molecules in
775: the atomistic region because there atoms never come this close
776: together. Therefore, by this force capping we only prevent
777: possible force singularities that could emerge due to overlaps
778: with the neighboring molecules when a given molecule enters the
779: interface layer from the coarse-grained side as explained in the
780: previous section. Production runs with the
781: $10^{9}\varepsilon/\sigma$ force capping are then performed for
782: $7500\tau$, storing configurations at each $5\tau$ time interval
783: for analysis. We performed all our MD simulations using the
784: ESPResSo package\cite{Espresso:2005}, developed at our institute.
785: 
786: 
787: The following reduced Lennard-Jones units\cite{Allen:1987} are
788: used throughout:
789:  $m^*=m/m_0$, $r^*=r/\sigma$, $V^*=V/\sigma^3$,
790:  $T^*=k_BT/\varepsilon$, $U^*=U/\varepsilon$, $p^*=p\sigma^3/\varepsilon$,
791:  $\rho^*=n/V^*$, $t^*=t/\tau$, $D^*=D\sqrt{m_0/\varepsilon}/\sigma$,
792: where $D$ is the self-diffusion constant. Note that in our
793: simulations all atoms have a mass $m^*=1$ while all molecules have
794: a mass $M^*=4$.
795: 
796: 
797: 
798: \section{Results and Discussion}
799: 
800: \subsection*{Determination of the Effective Potential}
801: 
802: We have determined the effective nonbonded pair potential
803: ${U^{cm}}^*$ between coarse-grained molecules illustrated in
804: figure \ref{Fig.3} by using the potential of mean force
805: (PMF$_{ex}$) of the \emph{ex} system, Eq. (\ref{eq.0}), at very
806: low number density $\rho^*=0.0025$ as the initial guess. Then we
807: further adjusted it to obtain the adequate agreement between
808: RDF$_{cm}$s of the \emph{ex} and \emph{cg} systems at the
809: $\rho^*=0.1$.
810: \begin{figure}[!ht]
811: \centering
812: \includegraphics[width=7.5cm, angle=0]{fig3.ps}
813: \caption{ The effective pair potential ${U^{cm}}^*$, Eq.
814: (\ref{eq.11}), between the coarse-grained molecules, where the
815: potential of mean force PMF$_{ex}$ of the explicit system at
816: $\rho^*=0.0025$ and $T^*=1$ was used as the initial guess. The
817: presented function ${U^{cm}}^*$ was determined in such a way that
818: the RDF$_{cm}$s of the explicit (\emph{ex}) and coarse-grained (\emph{cg})
819: systems match at the $\rho^*=0.1$.}\label{Fig.3}
820: \end{figure}
821: We have parameterized the effective potential
822: $U^{cm}$ with the Morse potential
823: \begin{equation}
824:  {U^{cm}}^*(r^*)=\gamma^*\{1-\exp[-\kappa^*(r^*-r^*_0)]\}^2\label{eq.11}
825: \end{equation}
826: with parameters $\gamma^*=\gamma/\varepsilon=0.105$, $\kappa^*=\kappa\sigma=2.4$, and $r^*_0=r_0/\sigma=2.31$.
827: As one can see from figure \ref{Fig.3}, the obtained effective
828: potential is softer than the underlying repulsive interatomic interaction
829: potential given by Eqs. (\ref{eq.8}) and (\ref{eq.9}) since it varies
830: more slowly with the interparticle distance. This is a general feature
831: of effective potentials for polyatomic molecular systems\cite{Klapp:2004}.
832: 
833: The obtained RDF$_{cm}$s of the \emph{ex} and \emph{cg} systems at
834: the temperature $T^*=1$ and number density $0.025\le\rho^*\le
835: 0.175$ using for all number densities the same effective potential
836: given by Eq. (\ref{eq.11}) for the pair interactions between the
837: coarse-grained molecules are depicted in figure \ref{Fig.4}. 
838: \begin{figure}[!ht]
839: \centering
840: \includegraphics[height=10cm,width=7.5cm,angle=0]{fig4.ps}
841: \caption{Center-of-mass radial distribution functions of
842: the explicit (\emph{ex}) and coarse-grained (\emph{cg}) systems at
843: the temperature $T^*=1$ and number density $0.025\le\rho^*\le
844: 2.25$.}\label{Fig.4}
845: \end{figure}
846: The
847: RDF$_{cm}$s are calculated in the range $r^*\in [0,5]$ with
848: $\Delta r^*=0.05$. From results presented in figure \ref{Fig.4}
849: it can be observed that although the effective potential given by
850: Eq. (\ref{eq.11}) was parameterized at the number density
851: $\rho^*=0.1$  the agreement of RDF$_{cm}$s of the \emph{cg} system
852: with the corresponding reference RDF$_{cm}$s of the \emph{ex}
853: system is good also for the lower number densities. In fact, due
854: to weaker many-body interactions the agreement is even better for
855: the systems with lower density. Since the effective potentials are
856: density dependent\cite{Louis:2002}, to obtain a better agreement
857: for higher densities the effective potential should be
858: reparameterized\cite{Reith:2003}. Probably the functional form of
859: the effective potential will also change due to increased
860: contribution from many-body interactions.
861: 
862: As a quantitative measure of accuracy of the evaluated RDF$_{cm}$s
863: we define a penalty function $f_p$ as
864: \begin{equation}
865:  f_p=\int [g(r^*)-g^{cm}_{ex}(r^*)]^2\exp(-r^*)\,dr^*,\label{eq.10}
866: \end{equation}
867: where $g^{cm}_{ex}$, which is taken as a reference, is the
868: RDF$_{cm}$ of the \emph{ex} system. The function $\exp(-r^*)$ is
869: employed to penalize more strongly deviations at small
870: distances\cite{Reith:2003}. The values of $f_p$ for \emph{cg}
871: systems with the number density $0.025\le\rho^*\le 0.175$ are
872: reported in table \ref{Tab.1}.
873: \begin{table}[ht]
874:    \centering
875:   \begin{tabular}{||c|c||}
876:     \hline\hline
877:       $\rho^*$  & $f_p\cdot 10^3$        \\
878:     \hline
879:       $0.025$     & $0.0401$  \\
880:       $0.075$     & $0.1101$  \\
881:       $0.1$       & $0.2067$  \\
882:       $0.125$     & $0.4361$  \\
883:       $0.150$     & $0.9257$  \\
884:       $0.175$     & $1.9025$  \\
885:     \hline\hline
886:    \end{tabular}
887:    \caption{Penalty function $f_p$ defined by Eq. (\ref{eq.10}) as a function of number density
888:    $\rho^*$ for RDF$_{cm}$s $g^{cm}_{cg}(r^*)$ of the coarse-grained
889:    systems in which particles are interacting via the effective
890:    potential ${U^{cm}}^*$ given by
891:    Eq. (\ref{eq.11}). The RDF$_{cm}$s $g^{cm}_{ex}(r^*)$ of all-atom systems at the corresponding $\rho^*$
892:    are taken for the reference RDF$_{cm}$s.} \label{Tab.1}
893:   \end{table}
894:  As expected the $f_p$ grows with the growing density of the system.
895: 
896: 
897: From the RDF$_{cm}$ we can evaluate the average number of
898: neighbors of a given molecule within a sphere with the radius
899: $r^*$ as
900: \begin{equation}
901:   n_n(r^*)=\rho^*\int_0^{r^*} g(r^*)4\pi r^{*^2}\,dr^*.
902: \end{equation}
903: The $n_n(r^*)$ for the \emph{ex} and \emph{cg} systems at the
904: temperature $T^*=1$ and number density $0.025\le\rho^*\le 0.175$
905: are shown in figure \ref{Fig.5}. Despite the deviations between
906: the RDF$_{cm}$s of the \emph{ex} and \emph{cg} systems the
907: corresponding average numbers of neighbors exactly match,
908: indicating, together with the RDF$_{cm}$s presented in figure
909: \ref{Fig.4}, that the \emph{cg} model with the molecules
910: interacting via the effective potential given by Eq. (\ref{eq.11})
911: reproduces well the structure of the underlying \emph{ex} system.
912: \begin{figure}[!ht]
913: \centering
914: \includegraphics[height=10cm,width=7.5cm,angle=0]{fig5.ps}
915: \caption{The average number of neighbors $n_n(r^*)$ of a
916: given molecule as a function of distance for explicit (\emph{ex})
917: and coarse-grained (\emph{cg}) systems at the temperature $T^*=1$
918: and number density $0.025\le\rho^*\le 2.25$.}\label{Fig.5}
919: \end{figure}
920: 
921: To show that the \emph{cg} system with the effective potential
922: from Eq. (\ref{eq.11}) is at the same state point as the original
923: \emph{ex} system at the same temperature and density we also
924: evaluated the pressure in the system. The equations of state for
925: the \emph{ex} and \emph{cg} models are shown in figure
926: \ref{Fig.6}. 
927: \begin{figure}[!ht]
928: \centering
929: \mbox{\includegraphics[width=7.5cm, angle=0]{fig6.ps}}
930:          \caption{Pressure $p^*$ in the explicit (\emph{ex}) and
931: coarse-grained (\emph{cg}) systems at the temperature $T^*=1$ as a
932: function of the number density $\rho^*$ of the system.}\label{Fig.6}
933: \end{figure}
934: The resulting equations of state are similar to the
935: case of colloidal sphere systems\cite{Bryant:2002}. This
936: similarity is to be expected since the tetrahedral molecule as
937: well as the spherical coarse-grained molecule can be considered as
938: spherical colloidal particles with a hard core and a soft coating
939: layer. From figure \ref{Fig.6} we can conclude that the \emph{cg}
940: model with the particles interacting via the effective potential
941: from Eq. (\ref{eq.11}) reproduces the equation of state of the
942: underlying \emph{ex} system up to $\rho^*=0.125$, at which the two
943: pressure curves start deviating from each other, reflecting the
944: fact that the effective potential is density dependent. In order
945: to minimize the artifacts originating from our parameterization of
946: the effective potential while still simulating the liquid state we
947: have decided to perform all our MD simulations of the hybrid
948: atomistic/mesoscopic (\emph{ex-cg}) model at the state point with
949: $T^*=1$ and density $\rho^*=0.1$.
950: 
951: 
952: \subsection*{Statistical Properties}
953: 
954: The new adaptive resolution scheme is tested by comparing the
955: computed statistical properties of the \emph{ex-cg} model with the
956: corresponding properties of the reference fully atomistic \emph{ex}
957: system.
958: 
959: 
960: Figure \ref{Fig.7} (a) displays the RDF$_{cm}$s calculated from
961: center-of-mass positions of all molecules in the box of the
962: \emph{ex} and \emph{ex-cg} systems at $\rho^*=0.1$ and $T^*=1$.
963: \begin{figure}[!ht]
964: \centering
965: \subfigure[]{\includegraphics[width=7.5cm]{fig7a.ps}}\\
966: \subfigure[]{\includegraphics[width=7.5cm]{fig7b.ps}}
967:          \caption{(a) Center-of-mass radial distribution functions
968: for all molecules in the box of the all-atom (\emph{ex}) and hybrid
969: atomistic/mesoscopic (\emph{ex-cg}) systems at $\rho^*=0.1$ and
970: $T^*=1$.  Shown are also the corresponding center-of-mass radial
971: distribution functions for only the explicit molecules from the
972: explicit region (\emph{ex-cg}/\emph{ex}) and for only the
973: coarse-grained molecules from the coarse-grained region
974: (\emph{ex-cg}/\emph{cg}). The width of the interface layer is
975: $2d^*=2.5$. (b) The corresponding average numbers of neighbors
976: $n_n(r^*)$ of a given molecule as functions of distance. The
977: different curves are almost indistinguishable.}\label{Fig.7}
978: \end{figure}
979: Shown is the case with the width of the interface layer
980: $2d^*=2.5$. Depicted are also the corresponding local RDF$_{cm}$s
981: for the explicit  (\emph{ex-cg}/\emph{ex}) and coarse-grained
982: regions (\emph{ex-cg}/\emph{cg}) of the \emph{ex-cg} model. As in
983: all simple liquids, e.g., methane, the RDF$_{cm}$s are zero at
984: short distances between molecules' centers of mass because
985: repulsive forces prevent overlapping of molecules. Then the
986: functions increase rapidly to the first peak. With increasing
987: distance RDF$_{cm}$s reach the limiting value $1$ after few
988: oscillations, indicating that there is no order at long distances.
989: The average numbers of neighbors $n_n(r^*)$ of a given molecule as
990: functions of distance are illustrated in figure \ref{Fig.7} (b).
991: The number of nearest neighbors in the first layer corresponding
992: to the first minimum in the RDF$_{cm}$ is about $11$. For comparison, the corresponding
993: experimental value from X-ray diffraction for liquid methane at
994: $T=92\units{K}$ is approximately $12$\cite{Habenschuss:1981}.
995: 
996: All calculated RDF$_{cm}$s of the \emph{ex-cg} model and
997: $n_n(r^*)$  are in good agreement with the reference RDF$_{cm}$
998: and $n_n(r^*)$ of the \emph{ex} model indicating that the
999: structure of the underlying all-atom system is well reproduced
1000: using the adaptive resolution scheme. This is further confirmed by
1001: the values of the penalty function $f_p$ reported in table
1002: \ref{Tab.2}, evaluated for different interface layer widths. 
1003: \begin{table}[h]
1004:    \centering
1005:    \begin{tabular}{||c|c|c|c||}
1006:     \hline\hline
1007:     \vbox to0pt{\hbox{\lower 0.6\baselineskip\hbox{$2d^*$}}\vss}&
1008:     \multicolumn{3}{c||}{$f_p\cdot 10^3$}\\
1009:     \cline{2-4}
1010:            & ex-cg     & ex-cg/ex   & ex-cg/cg  \\
1011:    \hline
1012:     $2.5$  & 0.0821 &  0.0009 & 0.2109 \\
1013:     $3.0$  & 0.0893 &  0.0008 & 0.2158 \\
1014:     $4.0$  & 0.1058 &  0.0016 & 0.2197 \\
1015:     $5.0$  & 0.1324 &  0.0023 & 0.2315 \\
1016:     $6.0$  & 0.1583 &  0.0033 & 0.2367 \\
1017:     $7.0$  & 0.1944 &  0.0047 & 0.2449 \\
1018:     $8.0$  & 0.2301 &  0.0076 & 0.2616 \\
1019:     $9.0$  & 0.2695 &  0.0082 & 0.2489 \\
1020:    $10.0$  & 0.3159 &  0.0101 & 0.2632 \\
1021:     \hline
1022: 
1023:     \hline\hline
1024:    \end{tabular}
1025:    \caption{Penalty function $f_p$ defined by Eq. (\ref{eq.10}) as a
1026:    function of the interface layer width $2d^*$ for RDF$_{cm}$s
1027:    $g^{cm}_{ex-cg}(r^*)$, $g^{cm}_{ex-cg/ex}(r^*)$,
1028:    $g^{cm}_{ex-cg/cg}(r^*)$ of the hybrid atomistic/mesoscopic model
1029:    at $\rho^*=0.1$ and $T^*=1$.  $g^{cm}_{ex-cg}(r^*)$ is the
1030:    RDF$_{cm}$ of all molecules in the box where all molecules are
1031:    considered indistinguishable, $g^{cm}_{ex-cg/ex}(r^*)$ is the RDF$_{cm}$ of only
1032:    the explicit molecules from the explicit region while
1033:    $g^{cm}_{ex-cg/cg}(r^*)$ is the RDF$_{cm}$ of only
1034:    the coarse-grained molecules from the coarse-grained region. The RDF$_{cm}$ $g^{cm}_{ex}(r^*)$
1035:    of all-atom system at the corresponding $\rho^*$ and $T^*$
1036:    is taken for the reference RDF$_{cm}$.}\label{Tab.2}
1037: \end{table}
1038: From
1039: the results given in table \ref{Tab.2} we can see that the local
1040: structure in the explicit region of the \emph{ex-cg} model is
1041: exactly reproduced independently of the interface layer width. In
1042: contrast, the accuracy of the local structure reproduction in the
1043: coarse-grained region of the \emph{ex-cg} model depends on the
1044: accuracy of the effective potential parameterization in the
1045: \emph{cg} model (see the value of $f_p$ given in table \ref{Tab.1}
1046: for $\rho^*=0.1$). The computed $f_p$s for the total RDF$_{cm}$ of
1047: the \emph{ex-cg} model lie up to $2d^*=8.0$ in between the values
1048: for the local explicit and coarse-grained RDF$_{cm}$s.
1049: 
1050: For narrow interface layers with $2d^*<2.5$ the adaptive
1051: resolution scheme fails to work. The molecules are not given
1052: enough space and time to adapt their orientation to their
1053: environment and the system is not properly equilibrated in this
1054: case. The value $2d^*=2.5$, for which the adaptive resolution
1055: scheme gives the acceptable results,  can be rationalized by the
1056: fact that the interface layer width should at least exceed the
1057: maximal range of interaction, which is the range of the effective
1058: potential, namely $2.31\sigma$.
1059: 
1060: To demonstrate that the thermodynamic properties are correctly
1061: reproduced by the adaptive resolution scheme the temperature and
1062: pressure of the system as a functions of the interface layer width
1063: are given in tables \ref{Tab.3} and \ref{Tab.4}, respectively. 
1064: The calculated temperature profile reported in table \ref{Tab.3} shows
1065: that the system is at the right temperature and that all degrees
1066: of freedom are properly equilibrated in accordance with the
1067: equipartition principle regardless of the interface layer width.
1068: The results given in table \ref{Tab.4} also show that the adaptive
1069: resolution scheme succeeds in reproducing the pressure of the
1070: underlying fully atomistic system.
1071: \begin{table*}[ht]
1072:    \centering
1073:   \begin{tabular}{||c|c|c|c|c|c|c||}
1074:     \hline\hline
1075:       $2d^*$  & $T$           & $T_{ex}$       & $T_{cg}$ & $T_{int}$ & $T_{ex}^{all}$& $T_{int}^{all}$        \\
1076:     \hline
1077:       $0^{ex}$& $1.00\pm 0.01$&  $1.00\pm 0.01$&  $-$     & $-$ &$1.00\pm 0.01$   &       $-$                   \\
1078:       $0^{cg}$&  $1.00\pm 0.01$         &   $-$       &   $1.00\pm 0.01$  &    $-$         &    $-$           &  $-$ \\
1079:       $2.5$   &   $1.00\pm 0.01$         &  $1.00\pm 0.01$          &   $1.00\pm 0.02$        &  $1.00 \pm 0.03$            &   $1.00\pm 0.01$            & $1.00\pm 0.02$  \\
1080:       $3.0$   & $1.00\pm 0.01$ & $1.00 \pm 0.02$ &   $1.00 \pm 0.02$ &   $1.00 \pm 0.03$   &  $1.00 \pm 0.01$ &  $1.00 \pm 0.02$   \\
1081:       $4.0$   &  $1.00\pm 0.01$        &  $1.00\pm 0.01$         &  $1.00 \pm 0.02$        &    $1.00 \pm 0.03$           &    $1.00\pm 0.01$             & $1.00\pm 0.01$   \\
1082:       $5.0$   &  $1.00\pm 0.01$        &   $1.00\pm 0.02$        &  $1.00\pm 0.02$        &    $1.00\pm 0.02$           &     $1.00\pm 0.01$          &  $1.00\pm 0.02$ \\
1083:       $6.0$   &  $1.00\pm 0.01$        &  $1.00\pm 0.02$          &  $1.00\pm 0.02$        &    $1.00\pm 0.02$          &        $1.00\pm 0.02$        &   $1.00\pm 0.02$\\
1084:       $7.0$   &    $1.00\pm 0.01$       &    $1.00\pm 0.02$        &   $1.00\pm 0.02$       &     $1.00\pm 0.02$        &    $1.00\pm 0.02$           &  $1.00\pm 0.01$\\
1085:       $8.0$   &   $1.00\pm 0.01$        & $1.00\pm 0.02$          &   $1.00\pm 0.02$       & $1.00\pm 0.02$            &          $1.00\pm 0.02$     &  $1.00\pm 0.01$\\
1086:       $9.0$   &  $1.00\pm 0.01$        &     $1.00\pm 0.03$      & $1.00\pm 0.03$         &     $1.00\pm 0.02$        &        $1.00\pm 0.01$       &   $1.00\pm 0.01$ \\
1087:       $10.0$  &    $1.00\pm 0.01$      &  $1.00\pm 0.03$         &    $1.00\pm 0.02$        &  $1.00\pm 0.02$           &          $1.00\pm 0.01$        & $1.00\pm 0.02$  \\
1088: 
1089:     \hline\hline
1090:    \end{tabular}
1091:    \caption{Average temperature as a function of the interface
1092:  layer width $2d^*$. $T$, $T_{ex}$, $T_{cg}$, and $T_{int}$ are the average temperatures
1093:  of the total system, the explicit, coarse-grained, and interface
1094:  layer regions, respectively, calculated by Eq. (\ref{eq.12}). $T_{ex}^{all}$ and $T_{int}^{all}$
1095:  are the average temperatures of the explicit and interface layer regions,
1096:  respectively, calculated from total velocities
1097:  (translational+vibrational+rotational) of explicit atoms in
1098:  molecules. $0^{ex}$ and $0^{cg}$ denote the all-atom and
1099:  coarse-grained systems, respectively. } \label{Tab.3}
1100:   \end{table*}
1101: \begin{table}[ht]
1102:    \centering
1103:   \begin{tabular}{||c|c||}
1104:     \hline\hline
1105:       $2d^*$        & $p^*$        \\
1106:     \hline
1107:       $0^{ex}$      &  $0.379\pm 0.009$\\
1108:       $0^{cg}$      &  $0.378\pm 0.004$\\
1109:       $2.5$         &  $0.382\pm 0.007$\\
1110:       $3.0$         &  $0.383\pm 0.006$\\
1111:       $4.0$         &  $0.384\pm 0.006$\\
1112:       $5.0$         &  $0.385\pm 0.007$\\
1113:       $6.0$         &  $0.386\pm 0.006$\\
1114:       $7.0$         &  $0.388\pm 0.004$\\
1115:       $8.0$         &  $0.389\pm 0.006$\\
1116:       $9.0$         &  $0.390\pm 0.005$\\
1117:      $10.0$         &  $0.391\pm 0.006$\\
1118:     \hline\hline
1119:    \end{tabular}
1120:    \caption{Average pressure calculated using Eq. (\ref{eq.12a}) as a function of the interface
1121:  layer width $2d^*$. $0^{ex}$ and $0^{cg}$ denote the all-atom and
1122:  coarse-grained systems, respectively. } \label{Tab.4}
1123:   \end{table}
1124: 
1125: In order to check that the chemical potentials in the atomistic
1126: and coarse-grained regions are equal as required by the condition
1127: (\ref{eq.1}) we report in table \ref{Tab.5} the average number of
1128: molecules in different regions of the system.  In table
1129: \ref{Tab.5} we also give the number of degrees of freedom
1130: $n_{DOF}$ in the system defined as
1131: \begin{equation}
1132:   n_{DOF}=3\sum_\alpha [w_\alpha N+(1-w_\alpha)],\label{eq.13}
1133: \end{equation}
1134: where $N=4$ is the number of explicit tetrahedral atoms in a
1135: molecule and $w_\alpha$ is the value of the weighting function
1136: defined by Eq. (\ref{eq.7}) for the molecule $\alpha$. The
1137: summation in Eq. (\ref{eq.13}) goes over all $n$ molecules of the
1138: system. 
1139: \begin{table}[ht]
1140:    \centering
1141:   \begin{tabular}{||c|c|c|c|c||}
1142:     \hline\hline
1143:       $2d^*$  & $n_{ex}$ & $n_{cg}$ & $n_{int}$ & $n_{DOF}$       \\
1144:     \hline
1145:       $0^{ex}$ &  $5001$       &  $0$          &  $0$         &  $60012$\\
1146:       $0^{cg}$ &    $0$        &  $5001$       &  $0$         &  $15003$\\
1147:        $2.5$   &  $2167\pm 38$ &  $2170\pm 48$ & $663\pm 21$  &  $37446\pm 516$\\
1148:        $3.0$   &  $2100\pm 45$ &  $2103\pm 60$ & $796\pm 37$  &  $37455\pm 493$\\
1149:        $4.0$   &  $1966\pm 58$ &  $1969\pm 47$ & $1064\pm 40$ &  $37467\pm 507$              \\
1150:        $5.0$   &  $1832\pm 47$ &  $1835\pm 56$ & $1332\pm 43$ &  $37470\pm 501$             \\
1151:        $6.0$   &  $1697\pm 57$ &  $1701\pm 54$ & $1601\pm 34$ &  $37467\pm 330$             \\
1152:        $7.0$   &  $1563\pm 45$ &  $1566\pm 29$ & $1871\pm 31$ &  $37479\pm 526$              \\
1153:        $8.0$   &  $1428\pm 48$ &  $1431\pm 27$ & $2141\pm 22$ &  $37479\pm 345$              \\
1154:        $9.0$   &  $1294\pm 22$ &  $1295\pm 23$ & $2411\pm 60$ &  $37488\pm 576$           \\
1155:       $10.0$   &  $1158\pm 33$ &  $1160\pm 16$ & $2682\pm 56$ &  $37482\pm 63$              \\
1156:     \hline\hline
1157:    \end{tabular}
1158:    \caption{Average number of molecules as a function of the interface
1159:    layer width $2d^*$. $n_{ex}$, $n_{cg}$, and $n_{int}$ are the
1160:    average number of molecules in the explicit, coarse-grained, and
1161:    interface layer regions, respectively. $n_{DOF}$ is the average
1162:    number of degrees of freedom defined by Eq. (\ref{eq.13}).
1163:    For orientation: in the system with $2500$ coarse-grained molecules,
1164:    $2500$ four atom explicit molecules, and no hybrid molecules
1165:    $n_{DOF}=37500$. $0^{ex}$ and $0^{cg}$ denote the all-atom and
1166:  coarse-grained systems, respectively.} \label{Tab.5}
1167:   \end{table}
1168: The results show that there is no molecule number bias in
1169: the system and that the \emph{ex-cg} system is homogeneous. Note
1170: that $n_{DOF}$ is greatly reduced employing the adaptive
1171: resolution scheme on the \emph{ex-cg} model in comparison with the
1172: fully atomistic model. The time evolution of the number of
1173: molecules in different regions of the system together with the
1174: $n_{DOF}$ is illustrated in figure \ref{Fig.8}.
1175: \begin{figure}[!ht]
1176: \centering
1177: \mbox{\includegraphics[width=7.5cm, angle=0]{fig8.ps}}
1178:          \caption{ Time evolution of number of molecules in a
1179: explicit ($n_{ex}$) , coarse-grained ($n_{cg}$), and interface
1180: regions ($n_{int}$) in the hybrid atomistic/mesoscopic model with
1181: the $2.5\sigma$ interface layer width. Time evolution of number of
1182: degrees of freedom in the system ($n_{DOF}$) is depicted in the
1183: inset.}\label{Fig.8}
1184: \end{figure}
1185:  The results
1186: clearly demonstrate that the system is in thermodynamical
1187: equilibrium, indicating that the conditions (\ref{eq.1}) are
1188: satisfied by our adaptive resolution scheme.
1189: 
1190: 
1191: Although we have parameterized the effective potential to
1192: reproduce the structural properties of the \emph{ex} system we can
1193: also compare the dynamical properties of the \emph{ex-cg} with the
1194: \emph{ex} model. For that purpose we have computed the
1195: self-diffusion coefficient, which is evaluated from the
1196: center-of-mass displacements using the Einstein relation
1197:  \begin{equation}
1198:   D^*=\frac{1}{6}\lim_{t^*\rightarrow\infty}\frac{\langle|{\bf
1199:   R}^*_\alpha(t^*)-{\bf R}^*_\alpha(0)|^2\rangle}{t^*},\label{eq.14}
1200:  \end{equation}
1201: where ${\bf R}^*_\alpha(t^*)$ is the center-of-mass position of
1202: the molecule $\alpha$ at time $t^*$ and averaging is performed
1203: over all molecules and all choices of time origin. The
1204: self-diffusion coefficient of the \emph{ex} and \emph{cg} models
1205: calculated in the microcanonical ensemble, where the Langevin
1206: thermostat is switched off after the initial warm-up run, are
1207: $0.24$ and $0.30$, respectively. The corresponding values of the
1208: self-diffusion coefficient for the \emph{ex}, \emph{cg}, and
1209: \emph{ex-cg} models with the Langevin thermostat switched on are
1210: $0.12$, $0.14$, and $0.13$, respectively. Since
1211: we use the same time and length scales in all our three models
1212: different values of the self-diffusion coefficient in the
1213: \emph{ex} and \emph{cg} models (with no Langevin thermostat
1214: applied) indicate that the coarse-grained molecules experience a
1215: slightly smaller intermolecular frictional hindrance in their
1216: motion compared to the explicit molecules. This indicates that the
1217: effective pair potential given in Eq. (\ref{eq.11}) introduces an
1218: effective time scale shift in the coarse-grained regime. This is
1219: known from other studies \cite{Tschop:1998,Tschop:1998:2}, where
1220: one actually takes advantage of that in order to reach very long
1221: simulation times\cite{leon:2005}. The apparent self-diffusion
1222: coefficient values of the \emph{ex} and \emph{cg} models are lower
1223: and much closer together when the Langevin thermostat is applied
1224: due to the frictional forces arising from the coupling to the
1225: thermostat unlike to the case of polymers. There typically the
1226: friction of the thermostat is negligible compared to the friction
1227: between monomers. Since the self-diffusion coefficient of the
1228: \emph{ex-cg} model is close to the corresponding values for the
1229: \emph{ex} and \emph{cg} models we can conclude that the
1230: center-of-mass dynamics of the molecules is similar in all three
1231: models.
1232: 
1233: 
1234: As the final test to demonstrate the reliability of the adaptive
1235: resolution scheme we have computed the number density profile of
1236: the \emph{ex-cg} model. 
1237: \begin{figure}[!ht]
1238: \centering
1239: \subfigure[]{\includegraphics[width=7.5cm]{fig9a.ps}}\\
1240: \subfigure[]{\includegraphics[width=7.5cm]{fig9b.ps}}
1241: 
1242:          \caption{(a) Normalized density profile in the $x$
1243: direction of the hybrid atomistic/meso\-scopic model with the
1244: $2.5\sigma$ interface layer width. Vertical lines denote
1245: boundaries between atomistic, coarse-grained and interface regions
1246: of the system. (b) The same as in (a) but for the  $10.0\sigma$
1247: interface layer width.}\label{Fig.9}
1248: \end{figure}
1249: The results for the system with $2d^*=2.5$ and
1250: $2d^*=10.0$ are presented in figures \ref{Fig.9} (a) and (b),
1251: respectively. 
1252: The results in figure \ref{Fig.9} (a) for the case
1253: of $2d^*=2.5$ show that the explicit and coarse-grained regions
1254: have the same homogeneous density as the reference system while
1255: the density in the transition regime undergoes an oscillation
1256: around the reference value $\rho^*_0=0.1$ with a magnitude of
1257: approximately $0.05\rho^*_0$. In the case of $2d^*=10.0$ (figure
1258: \ref{Fig.9} (b)) a $5\%$ drop in the density occurs in the
1259: transition regime, which is compensated by the slight increase of
1260: the density in the explicit and coarse-grained regions.
1261: 
1262: This artifact can be explained by considering the results
1263: displayed in figure \ref{Fig.10} (a), where the pressure of the
1264: system containing only hybrid molecules as a function of the
1265: constant value of the weighting function $w$ (corresponding to the
1266: situation in the interface layer) is illustrated.
1267: \begin{figure}[!ht]
1268: \centering
1269: \subfigure[]{\includegraphics[width=7.5cm]{fig10a.ps}}\\
1270: \subfigure[]{\includegraphics[width=7.5cm]{fig10b.ps}}
1271:          \caption{Artifacts of the adaptive resolution scheme. (a)
1272: Average pressure $p^*$ in the system containing only hybrid
1273: molecules as a function of the constant value of the weighting
1274: function $w$. (b) Center-of-mass radial distribution functions for
1275: the explicit system (ex) and the system containing only hybrid
1276: molecules with $w=const=1/2$ (\emph{ex-cg}(w=const=1/2)) at
1277: $\rho^*=0.0025$ and $T^*=1$. The inset also shows the
1278: corresponding PMF$_{ex}$ and PMF$_{ex-cg(w=const=1/2)}$ determined
1279: from the systems with $\rho^*=0.0025$  using Eq. (\ref{eq.0}).
1280: }\label{Fig.10}
1281: \end{figure}
1282:  The pressure is
1283: increased in comparison to the pressure in the reference system.
1284: Clearly the increase is most prominent for the most 'artificial'
1285: case with $w=1/2$, indicating that there still is a small
1286: 'pressure barrier' in the interface region causing the density
1287: dip. This is also evident from results in figure  \ref{Fig.10}
1288: (b), where the RDF$_{cm}$ and the potential of mean force of the
1289: system with $w=1/2$ are shown. The effective potential in a system
1290: containing only hybrid molecules with constant value of the
1291: weighting function  changes in comparison to the all-atom system.
1292: This means that the hybrid molecules in the interface layers of
1293: the \emph{ex-cg} model experience too strong effective interaction
1294: leading to the pressure variations in the interface layer. Since the
1295: artifact occurs at constant values of $w$ it is an artifact of the
1296: linear combination of forces in Eq. (\ref{eq.4}) and not of the
1297: functional form of the weighting function $w$. It must be
1298: emphasized, however, that this artifact of the proposed adaptive
1299: resolution scheme is within a $5\%$ error and that similar
1300: artifacts, occurring at the boundary of two domains with different
1301: level of detail, are also characteristic for other hybrid
1302: schemes\cite{Cai:2000}.
1303: 
1304: The pressure variations in the interface layer could cause a
1305: spurious reflection of molecules from the boundary. However, the
1306: results presented in figure \ref{Fig.11}, where the time evolution
1307: of two diffusion profiles is monitored for molecules that are
1308: initially localized at the two slabs with $a^*/10$ width
1309: neighboring the interface layer, show that this is not the case.
1310: \begin{figure}[!ht]
1311: \centering
1312: \subfigure[]{\includegraphics[width=7.5cm]{fig11a.ps}}\\
1313: \subfigure[]{\includegraphics[width=7.5cm]{fig11b.ps}}
1314:          \caption{Time evolution of diffusion profiles for the
1315: molecules that are initially, at time $t^*=0$, localized at two
1316: neighboring slabs of the mid interface layer with $2d^*=2.5$ ($n$
1317: is the number of this molecules with the center-of-mass position
1318: at a given coordinate $x^*$). The width of the two slabs is
1319: $a^*/10$. Vertical lines denote boundaries of the interface layer.
1320: (a) The diffusion profile, averaged over $500$ different time
1321: origins, at $t^*=0$, $t^*=10$, and $t^*=50$ for the molecules that
1322: are initially localized at the slab on the coarse-grained side of
1323: the interface region. (b) The same as in (a) but for the molecules
1324: that are initially localized at the slab on the atomistic side of
1325: the interface region.
1326: }\label{Fig.11}
1327: \end{figure}
1328: The molecules initially localized at the two slabs spread out
1329: symmetrically with time. This is because the self-diffusion
1330: coefficients of all models are approximately the same in the case
1331: of the applied Langevin thermostat. Thus, the two distributions
1332: occupy at time $t^*$ regions with mean square radius
1333: \begin{equation}
1334:  \langle |x^*(t^*)-x^*(0)|^2\rangle\simeq 2D^*t^*,
1335: \end{equation}
1336: where $x^*(0)$ is the center of the distribution at time $t^*=0$,
1337: which is equal to $-d^*-a^*/20$  for the left slab (figure
1338: \ref{Fig.11} (a)) and $d^*+a^*/20$ for the right slab (figure
1339: \ref{Fig.11} (b)). Since the diffusion profiles are symmetrical at
1340: any given time we can conclude that the artifact at the interface
1341: layer is too small to have any severe effect on the diffusion of
1342: molecules across the interface layer.
1343: 
1344: \section{Conclusions}
1345: A novel approach for efficient hybrid atomistic/mesoscale
1346: molecular dynamics (MD) simulations has been presented in this
1347: paper. The new adaptive resolution MD simulation scheme
1348: dynamically couples the atomic and mesoscale length scales of the
1349: studied system by allowing an on-the-fly dynamical interchange
1350: between molecules' atomic and mesoscopic levels of description. In
1351: our approach the number of degrees of freedom is allowed to
1352: fluctuate during the course of simulation while the statistical
1353: properties of the underlying all-atom model are properly
1354: reproduced for both levels of detail. Since the purpose of this
1355: paper was to develop the method, we restricted ourselves to a most
1356: simple model system of a molecular liquid with short-range
1357: repulsive interactions. Even on this level a number of questions
1358: still have to be tackled, as there is the application to higher
1359: liquid densities for complex molecular liquids or the proper
1360: analysis and understanding of the the time scale problem, when the
1361: diffusion on the coarse-grained level is faster than on the
1362: all-atom level. For separate runs at the two levels of detail this
1363: is understood, while the occurrence within one simulation box
1364: still poses some conceptual problems. On the other hand the
1365: current density is in between a typical small molecule liquid and
1366: a polymeric fluid. Thus we expect that already this approach can
1367: be generalized and applied to different realistic soft condensed
1368: matter systems where both atomic and mesoscopic length scales have
1369: to be considered. This can be either polymer solutions and melts,
1370: such as a synthetic or biological macromolecule embedded in a
1371: solvent. Similarly our method should also find application for
1372: other polymer systems (same force field), molecular liquids such
1373: as methane (same geometry) or water (tetrahedral clusters), etc.,
1374: enabling to reach much larger length and time scales than for
1375: all-atom MD simulations. In all cases the aim is to treat in a
1376: simulation only as many degrees of freedom as absolutely necessary
1377: for the question considered. In this sense the region of higher
1378: detail can be either given by a geometrical constraint, e.g., close
1379: to a surface, or even be chosen on demand due to specific local
1380: conformations of a (macro-)molecular system. Work along these
1381: lines is underway.
1382: \vspace{0.5cm}
1383: 
1384: \section*{\small ACKNOWLEDGMENTS}
1385: We thank A. Arnold, B.~A. Mann, P. Schravendijk, B. Hess, and N.
1386: van der Vegt for useful discussions. We are also grateful to C.~F.
1387: Abrams for discussions at early stage of this work. This work is
1388: supported in part by the Volkswagen foundation. M.~P. acknowledges
1389: the support of the Ministry of Higher Education, Science and
1390: Technology of Slovenia under grant No. P1-0002.
1391: 
1392: 
1393: 
1394: 
1395: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1396: \begin{thebibliography}{10}
1397: 
1398: \bibitem{Deuflhard:1999}
1399: P.~Deuflhard, J.~Hermans, B.~Leimkuhler, A.~Mark, S.~Reich and R.~Skeel, editors,
1400: \newblock {\em Computational Molecular Dynamics: Challenges, Methods, Ideas},
1401:   volume~4 of {\em Lecture Notes in Computational Science and Engineering},
1402: \newblock Springer Verlag, 1999.
1403: 
1404: \bibitem{Minary:2004}
1405: P.~Minary, M.~E. Tuckerman, and G.~J. Martyna,
1406: \newblock Phys. Rev. Lett. {\bf 93}, 150201 (2004).
1407: 
1408: \bibitem{Praprotnik:2005}
1409: D.~Jane\v{z}i\v{c}, M.~Praprotnik, and F.~Merzel,
1410: \newblock J. Chem. Phys. {\bf 122}, 174101 (2005).
1411: 
1412: \bibitem{Praprotnik:2005:1}
1413: M.~Praprotnik and D.~Jane\v{z}i\v{c},
1414: \newblock J. Chem. Phys. {\bf 122}, 174102 (2005).
1415: 
1416: \bibitem{Praprotnik:2005:2}
1417: M.~Praprotnik and D.~Jane\v{z}i\v{c},
1418: \newblock J. Chem. Phys. {\bf 122}, 174103 (2005).
1419: 
1420: \bibitem{Praprotnik:2004}
1421: M.~Praprotnik, D.~Jane\v{z}i\v{c}, and J.~Mavri,
1422: \newblock J. Phys. Chem. A {\bf 108}, 11056 (2004).
1423: 
1424: \bibitem{kremer:2001}
1425: K.~Kremer, and F.~M\"uller-Plathe,
1426: \newblock MRS Bulletin \textbf{26}, 205 (2001).
1427: 
1428: \bibitem{kremer:2004}
1429: K.~Kremer in "Multiscale  Modelling and Simulation, Lecture Notes
1430: on Computational Science and Engineering, S. Attinger and P.
1431: Koumoutsakis eds, Springer 2004.
1432: 
1433: \bibitem{Ayton:2004}
1434: G.~S. Ayton, H.~L. Tepper, D.~T. Mirijanian, and G.~A. Voth,
1435: \newblock J. Chem. Phys. {\bf 120}, 4074 (2004).
1436: 
1437: \bibitem{Kranenburg:2003}
1438: M.~Kranenburg, M.~Venturoli, and B.~Smit,
1439: \newblock J. Phys. Chem. B {\bf 107}, 11491 (2003).
1440: 
1441: \bibitem{Nielsen:2004}
1442: S.~O. Nielsen, C.~F. Lopez, G.~Srinivas, and M.~L. Klein,
1443: \newblock J. Phys.: Condens. Matter {\bf 16}, R481 (2004).
1444: 
1445: \bibitem{Marrink:2004}
1446: S.~J. Marrink, A.~H. de~Vries, and A.~E. Mark,
1447: \newblock J. Phys. Chem. B {\bf 108}, 750 (2004).
1448: 
1449: \bibitem{Chang:2005}
1450: R.~Chang, G.~S. Ayton, and G.~A. Voth,
1451: \newblock J. Chem. Phys. {\bf 122}, 244716 (2005).
1452: 
1453: \bibitem{Cook:2005}
1454: I.~R. Cooke, K.~Kremer, and M.~Deserno,
1455: \newblock Phys. Rev. E {\bf 72}, 011506 (2005). 
1456: 
1457: \bibitem{Kremer:1990}
1458: K.~Kremer and G.~S. Grest,
1459: \newblock J. Chem. Phys. {\bf 92}, 5057 (1990).
1460: 
1461: \bibitem{Tschop:1998}
1462: W.~Tsch{\" o}p, K.~Kremer, J.~Batoulis, T.~B{\" u}rger, and O.~Hahn,
1463: \newblock Acta Polym. {\bf 49}, 61 (1998).
1464: 
1465: \bibitem{Tschop:1998:2}
1466: W.~Tsch{\" o}p, K.~Kremer, O.~Hahn, J.~Batoulis, and T.~B{\" u}rger,
1467: \newblock Acta Polym. {\bf 49}, 75 (1998).
1468: 
1469: \bibitem{Everaers:2004}
1470: R.~Everaers, S.~K.~Sukumaran, G.~S.~Grest, C.~Svaneborg, A.~Sivasubramanian, and K.~Kremer,
1471: \newblock Science {\bf 303}, 823 (2004).
1472: 
1473: \bibitem{Chun:2000}
1474: H.~M. Chun, C.~E. Padilla, D.~N. Chin, M. Watanabe, V.~I. Karlov,
1475: H.~E. Alper, K. Soosaar, K.~B. Blair, O.~M. Becker, L.~S.~D. Caves,
1476: R. Nagle, D.~N. Haneym, and B.~L. Farmer,
1477: \newblock J. Comput. Chem. {\bf 21}, 159 (2000).
1478: 
1479: \bibitem{Malevanets:2000}
1480: A.~Malevanets and R.~Kapral,
1481: \newblock J. Chem. Phys. {\bf 112}, 7260 (2000).
1482: 
1483: \bibitem{Villa:2005}
1484: E.~Villa, A.~Balaeff, and K.~Schulten,
1485: \newblock PNAS {\bf 102}, 6783 (2005).
1486: 
1487: \bibitem{DelleSite:2002}
1488: L.~Delle Site, C.~F. Abrams, A.~Alavi, and K.~Kremer,
1489: \newblock Phys. Rev. Lett. {\bf 89}, 156103 (2002).
1490: 
1491: \bibitem{Abrams:2003}
1492: C.~F. Abrams, L.~Delle Site, and K.~Kremer,
1493: \newblock Phys. Rev. E {\bf 67}, 021807 (2003).
1494: 
1495: \bibitem{DelleSite:2004}
1496: L.~Delle Site, S.~Leon, and K.~Kremer,
1497: \newblock J. Am. Chem. Soc. {\bf 126}, 2944 (2004).
1498: \bibitem{Laio:2002}
1499: A.~Laio, J.~VandeVondele and U.~Roethlisberger,
1500: \newblock J.Chem.Phys. {\bf 116}, 6941 (2002).
1501: \bibitem{Rafii:1998}
1502: H.~Rafii-Tabar, L.~Hua, and M.~Cross,
1503: \newblock J. Phys.: Condens. Matter {\bf 10}, 2375 (1998).
1504: 
1505: \bibitem{Broughton:1999}
1506: J.~Q. Broughton, F.~F. Abraham, N.~B. Bernstein, and E.~Kaxiras,
1507: \newblock Phys. Rev. B {\bf 60}, 2391 (1999).
1508: 
1509: \bibitem{Smirnova:1999}
1510: J.~A. Smirnova, L.~V. Zhigilei, and B.~J. Garrison,
1511: \newblock Comp. Phys. Commun. {\bf 118}, 11 (1999).
1512: 
1513: \bibitem{Csanyi:2004}
1514: G.~Csanyi, T.~Albaret, M.~C. Payne, and A.~D. Vita,
1515: \newblock Phys. Rev. Lett. {\bf 93}, 175503 (2004).
1516: 
1517: \bibitem{Abrams:2004}
1518: C.~F. Abrams,
1519: \newblock Inhomogenous coarse-graining of polymers and polymer/metal
1520:   interfaces,
1521: \newblock in {\em Computational Soft Matter: {F}rom Synthetic Polymers to
1522:   Proteins}, edited by N.~Attig, K.~Binder, H.~Grubm{\" u}ller, and K.~Kremer,
1523:   volume~23 of {\em {NIC} Series}, pages 275--288, John von {N}eumann
1524:   {I}nstitute for {C}omputing, 2004.
1525: 
1526: \bibitem{Louis:2002}
1527: A.~A. Louis,
1528: \newblock J. Phys.: Condens. Matter {\bf 14}, 9187 (2002).
1529: 
1530: \bibitem{Curtarolo:2002}
1531: S.~Curtarolo and G.~Ceder,
1532: \newblock Phys. Rev. Lett. {\bf 88}, 255504 (2002).
1533: 
1534: \bibitem{Klapp:2004}
1535: S.~H.~L. Klapp, D.~J. Diestler, and M.~Schoen,
1536: \newblock J. Phys.: Condens. Matter {\bf 16}, 7331 (2004).
1537: 
1538: \bibitem{Reith:2003}
1539: D.~Reith, M.~P{\" u}tz, and F.~M{\" u}ller-Plathe,
1540: \newblock J. Comput. Chem. {\bf 24}, 1624 (2003).
1541: 
1542: \bibitem{Bryant:2002}
1543: G.~Bryant, S.~R. Williams, L. Qian, I.~K. Snook, E. Perez, and F. Pincet,
1544: \newblock Phys. Rev. E {\bf 66}, 060501(R) (2002).
1545: 
1546: \bibitem{Lawrence:2003}
1547: C.~P. Lawrence and J.~L. Skinner,
1548: \newblock Chem. Phys. Lett. {\bf 372}, 842 (2003).
1549: 
1550: \bibitem{Zwanzig:1954}
1551: R.~W. Zwanzig,
1552: \newblock J. Chem. Phys. {\bf 22}, 1420 (1954).
1553: 
1554: \bibitem{Leach:2001}
1555: A.~R. Leach,
1556: \newblock {\em Molecular Modelling},
1557: \newblock Pearson Education Limited, Harlow, 2 edition, 2001.
1558: 
1559: \bibitem{Schwabl:1995}
1560: F.~Schwabl,
1561: \newblock {\em Quantum Mechanics},
1562: \newblock Springer-Verlag, Berlin Heidelberg, 2 edition, 1996.
1563: 
1564: 
1565: \bibitem{Soddemann:2003}
1566: T.~Soddemann, B.~D{\" u}nweg, and K. Kremer,
1567: \newblock Phys. Rev. E {\bf 68}, 046702 (2003).
1568: 
1569: \bibitem{Allen:1987}
1570: M.~P. Allen and D.~J. Tildesley,
1571: \newblock {\em Computer Simulation of Liquids},
1572: \newblock Clarendon Press, Oxford, 1987.
1573: 
1574: \bibitem{Berendsen:1984}
1575: H.~Berendsen, J.~Postma, W.~V. Gunsteren, A.~D. Nola, and J.~Haak,
1576: \newblock J. Chem. Phys. {\bf 81}, 3684 (1984).
1577: 
1578: \bibitem{Mann:2004}
1579: B.~A. Mann, R.~Everaers, C.~Holm, and K.~Kremer,
1580: \newblock Europhys. Lett. {\bf 67}, 786 (2004).
1581: 
1582: \bibitem{Duenweg:1991}
1583: B.~D{\" u}nweg and W.~Paul,
1584: \newblock Int. J. Mod. Phys. C {\bf 2}, 817 (1991).
1585: 
1586: 
1587: \bibitem{Kremer:2005}
1588: The physically correct way of course would be
1589: to calculate the second virial coefficient. This here is not that
1590: helpful, because the interaction potential is rather smooth. The
1591: present measure is more appropriate, when the structure of the
1592: liquids is compared in an NVT simulation, as in our case.
1593: 
1594: \bibitem{Espresso:2005}
1595: http://www.espresso.mpg.de.
1596: 
1597: \bibitem{Habenschuss:1981}
1598: A.~Habenschuss, E.~Johnson, and A.~H. Narten,
1599: \newblock J. Chem. Phys. {\bf 74}, 5234 (1981).
1600: 
1601: 
1602: \bibitem{leon:2005}
1603: S.~Leon, N.~van der Vegt, L.~Delle Site, and K.~Kremer,
1604: \newblock Macromolecules in press (2005).
1605: 
1606: \bibitem{Cai:2000}
1607: W.~Cai, M.~de~Koning, V.~V. Bulatov, and S.~Yip,
1608: \newblock Phys. Rev. Lett. {\bf 85}, 3213 (2000).
1609: 
1610: 
1611: 
1612: \end{thebibliography}
1613: 
1614: 
1615: \end{document}
1616: