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

Comparing trunk/tengDissertation/Introduction.tex (file contents):
Revision 2907 by tim, Thu Jun 29 16:57:37 2006 UTC vs.
Revision 2911 by tim, Thu Jun 29 23:56:11 2006 UTC

# Line 281 | Line 281 | space of the system,
281   (q,p,t)dq_1 } ...dq_f dp_1 } ...dp_f }}{{\int { \ldots \int {\rho
282   (q,p,t)dq_1 } ...dq_f dp_1 } ...dp_f }}.
283   \label{introEquation:ensembelAverage}
284 \end{equation}
285
286 There are several different types of ensembles with different
287 statistical characteristics. As a function of macroscopic
288 parameters, such as temperature \textit{etc}, the partition function
289 can be used to describe the statistical properties of a system in
290 thermodynamic equilibrium. As an ensemble of systems, each of which
291 is known to be thermally isolated and conserve energy, the
292 Microcanonical ensemble (NVE) has a partition function like,
293 \begin{equation}
294 \Omega (N,V,E) = e^{\beta TS}. \label{introEquation:NVEPartition}
295 \end{equation}
296 A canonical ensemble (NVT) is an ensemble of systems, each of which
297 can share its energy with a large heat reservoir. The distribution
298 of the total energy amongst the possible dynamical states is given
299 by the partition function,
300 \begin{equation}
301 \Omega (N,V,T) = e^{ - \beta A}.
302 \label{introEquation:NVTPartition}
284   \end{equation}
304 Here, $A$ is the Helmholtz free energy which is defined as $ A = U -
305 TS$. Since most experiments are carried out under constant pressure
306 condition, the isothermal-isobaric ensemble (NPT) plays a very
307 important role in molecular simulations. The isothermal-isobaric
308 ensemble allow the system to exchange energy with a heat bath of
309 temperature $T$ and to change the volume as well. Its partition
310 function is given as
311 \begin{equation}
312 \Delta (N,P,T) =  - e^{\beta G}.
313 \label{introEquation:NPTPartition}
314 \end{equation}
315 Here, $G = U - TS + PV$, and $G$ is called Gibbs free energy.
285  
286   \subsection{\label{introSection:liouville}Liouville's theorem}
287  
# Line 1170 | Line 1139 | V(q,Q) + \frac{1}{2}tr[(QQ^T  - 1)\Lambda ].
1139   V(q,Q) + \frac{1}{2}tr[(QQ^T  - 1)\Lambda ].
1140   \label{introEquation:RBHamiltonian}
1141   \end{equation}
1142 < Here, $q$ and $Q$  are the position and rotation matrix for the
1143 < rigid-body, $p$ and $P$  are conjugate momenta to $q$  and $Q$ , and
1144 < $J$, a diagonal matrix, is defined by
1142 > Here, $q$ and $Q$  are the position vector and rotation matrix for
1143 > the rigid-body, $p$ and $P$  are conjugate momenta to $q$  and $Q$ ,
1144 > and $J$, a diagonal matrix, is defined by
1145   \[
1146   I_{ii}^{ - 1}  = \frac{1}{2}\sum\limits_{i \ne j} {J_{jj}^{ - 1} }
1147   \]
# Line 1182 | Line 1151 | Q^T Q = 1, \label{introEquation:orthogonalConstraint}
1151   \begin{equation}
1152   Q^T Q = 1, \label{introEquation:orthogonalConstraint}
1153   \end{equation}
1154 < which is used to ensure rotation matrix's unitarity. Differentiating
1155 < Eq.~\ref{introEquation:orthogonalConstraint} and using
1187 < Eq.~\ref{introEquation:RBMotionMomentum}, one may obtain,
1188 < \begin{equation}
1189 < Q^T PJ^{ - 1}  + J^{ - 1} P^T Q = 0 . \\
1190 < \label{introEquation:RBFirstOrderConstraint}
1191 < \end{equation}
1192 < Using Equation (\ref{introEquation:motionHamiltonianCoordinate},
1154 > which is used to ensure the rotation matrix's unitarity. Using
1155 > Equation (\ref{introEquation:motionHamiltonianCoordinate},
1156   \ref{introEquation:motionHamiltonianMomentum}), one can write down
1157   the equations of motion,
1158   \begin{eqnarray}
# Line 1198 | Line 1161 | the equations of motion,
1161   \frac{{dQ}}{{dt}} & = & PJ^{ - 1},  \label{introEquation:RBMotionRotation}\\
1162   \frac{{dP}}{{dt}} & = & - \nabla _Q V(q,Q) - 2Q\Lambda . \label{introEquation:RBMotionP}
1163   \end{eqnarray}
1164 + Differentiating Eq.~\ref{introEquation:orthogonalConstraint} and
1165 + using Eq.~\ref{introEquation:RBMotionMomentum}, one may obtain,
1166 + \begin{equation}
1167 + Q^T PJ^{ - 1}  + J^{ - 1} P^T Q = 0 . \\
1168 + \label{introEquation:RBFirstOrderConstraint}
1169 + \end{equation}
1170   In general, there are two ways to satisfy the holonomic constraints.
1171   We can use a constraint force provided by a Lagrange multiplier on
1172 < the normal manifold to keep the motion on constraint space. Or we
1173 < can simply evolve the system on the constraint manifold. These two
1174 < methods have been proved to be equivalent. The holonomic constraint
1175 < and equations of motions define a constraint manifold for rigid
1176 < bodies
1172 > the normal manifold to keep the motion on the constraint space. Or
1173 > we can simply evolve the system on the constraint manifold. These
1174 > two methods have been proved to be equivalent. The holonomic
1175 > constraint and equations of motions define a constraint manifold for
1176 > rigid bodies
1177   \[
1178   M = \left\{ {(Q,P):Q^T Q = 1,Q^T PJ^{ - 1}  + J^{ - 1} P^T Q = 0}
1179   \right\}.
1180   \]
1181 < Unfortunately, this constraint manifold is not the cotangent bundle
1182 < $T^* SO(3)$ which can be consider as a symplectic manifold on Lie
1183 < rotation group $SO(3)$. However, it turns out that under symplectic
1184 < transformation, the cotangent space and the phase space are
1216 < diffeomorphic. By introducing
1181 > Unfortunately, this constraint manifold is not $T^* SO(3)$ which is
1182 > a symplectic manifold on Lie rotation group $SO(3)$. However, it
1183 > turns out that under symplectic transformation, the cotangent space
1184 > and the phase space are diffeomorphic. By introducing
1185   \[
1186   \tilde Q = Q,\tilde P = \frac{1}{2}\left( {P - QP^T Q} \right),
1187   \]
# Line 1281 | Line 1249 | Omelyan1998}. Applying the hat-map isomorphism, we obt
1249   motion. This unique property eliminates the requirement of
1250   iterations which can not be avoided in other methods\cite{Kol1997,
1251   Omelyan1998}. Applying the hat-map isomorphism, we obtain the
1252 < equation of motion for angular momentum on body frame
1252 > equation of motion for angular momentum in the body frame
1253   \begin{equation}
1254   \dot \pi  = \pi  \times I^{ - 1} \pi  + \sum\limits_i {\left( {Q^T
1255   F_i (r,Q)} \right) \times X_i }.
# Line 1294 | Line 1262 | given by
1262   \]
1263  
1264   \subsection{\label{introSection:SymplecticFreeRB}Symplectic
1265 < Lie-Poisson Integrator for Free Rigid Body}
1265 > Lie-Poisson Integrator for Free Rigid Bodies}
1266  
1267   If there are no external forces exerted on the rigid body, the only
1268   contribution to the rotational motion is from the kinetic energy
# Line 1346 | Line 1314 | To reduce the cost of computing expensive functions in
1314   \end{array}} \right),\theta _1  = \frac{{\pi _1 }}{{I_1 }}\Delta t.
1315   \]
1316   To reduce the cost of computing expensive functions in $e^{\Delta
1317 < tR_1 }$, we can use Cayley transformation to obtain a single-aixs
1318 < propagator,
1319 < \[
1320 < e^{\Delta tR_1 }  \approx (1 - \Delta tR_1 )^{ - 1} (1 + \Delta tR_1
1321 < ).
1322 < \]
1323 < The propagator maps for $T_2^r$ and $T_3^r$ can be found in the same
1317 > tR_1 }$, we can use the Cayley transformation to obtain a
1318 > single-aixs propagator,
1319 > \begin{eqnarray*}
1320 > e^{\Delta tR_1 }  & \approx & (1 - \Delta tR_1 )^{ - 1} (1 + \Delta
1321 > tR_1 ) \\
1322 > %
1323 > & \approx & \left( \begin{array}{ccc}
1324 > 1 & 0 & 0 \\
1325 > 0 & \frac{1-\theta^2 / 4}{1 + \theta^2 / 4}  & -\frac{\theta}{1+
1326 > \theta^2 / 4} \\
1327 > 0 & \frac{\theta}{1+ \theta^2 / 4} & \frac{1-\theta^2 / 4}{1 +
1328 > \theta^2 / 4}
1329 > \end{array}
1330 > \right).
1331 > \end{eqnarray*}
1332 > The propagators for $T_2^r$ and $T_3^r$ can be found in the same
1333   manner. In order to construct a second-order symplectic method, we
1334   split the angular kinetic Hamiltonian function into five terms
1335   \[
# Line 1378 | Line 1355 | norm of the angular momentum, $\parallel \pi
1355   function $G$ is zero, $F$ is a \emph{Casimir}, which is the
1356   conserved quantity in Poisson system. We can easily verify that the
1357   norm of the angular momentum, $\parallel \pi
1358 < \parallel$, is a \emph{Casimir}. Let$ F(\pi ) = S(\frac{{\parallel
1358 > \parallel$, is a \emph{Casimir}\cite{McLachlan1993}. Let$ F(\pi ) = S(\frac{{\parallel
1359   \pi \parallel ^2 }}{2})$ for an arbitrary function $ S:R \to R$ ,
1360   then by the chain rule
1361   \[
# Line 1448 | Line 1425 | moving rigid bodies
1425   \begin{eqnarray}
1426   \varphi _{\Delta t}  &=& \varphi _{\Delta t/2,F}  \circ \varphi _{\Delta t/2,\tau }  \notag\\
1427    & & \circ \varphi _{\Delta t,T^t }  \circ \varphi _{\Delta t/2,\pi _1 }  \circ \varphi _{\Delta t/2,\pi _2 }  \circ \varphi _{\Delta t,\pi _3 }  \circ \varphi _{\Delta t/2,\pi _2 }  \circ \varphi _{\Delta t/2,\pi _1 }  \notag\\
1428 <  & & \circ \varphi _{\Delta t/2,\tau }  \circ \varphi _{\Delta t/2,F}  .\\
1428 >  & & \circ \varphi _{\Delta t/2,\tau }  \circ \varphi _{\Delta t/2,F}  .
1429   \label{introEquation:overallRBFlowMaps}
1430   \end{eqnarray}
1431  
# Line 1525 | Line 1502 | differential equations into simple algebra problems wh
1502   differential equations,the Laplace transform is the appropriate tool
1503   to solve this problem. The basic idea is to transform the difficult
1504   differential equations into simple algebra problems which can be
1505 < solved easily. Then, by applying the inverse Laplace transform, also
1506 < known as the Bromwich integral, we can retrieve the solutions of the
1507 < original problems. Let $f(t)$ be a function defined on $ [0,\infty )
1508 < $, the Laplace transform of $f(t)$ is a new function defined as
1505 > solved easily. Then, by applying the inverse Laplace transform, we
1506 > can retrieve the solutions of the original problems. Let $f(t)$ be a
1507 > function defined on $ [0,\infty ) $, the Laplace transform of $f(t)$
1508 > is a new function defined as
1509   \[
1510   L(f(t)) \equiv F(p) = \int_0^\infty  {f(t)e^{ - pt} dt}
1511   \]
# Line 1546 | Line 1523 | L(x_\alpha  ) & = & \frac{{\frac{{g_\alpha  }}{{\omega
1523   p^2 L(x_\alpha  ) - px_\alpha  (0) - \dot x_\alpha  (0) & = & - \omega _\alpha ^2 L(x_\alpha  ) + \frac{{g_\alpha  }}{{\omega _\alpha  }}L(x), \\
1524   L(x_\alpha  ) & = & \frac{{\frac{{g_\alpha  }}{{\omega _\alpha  }}L(x) + px_\alpha  (0) + \dot x_\alpha  (0)}}{{p^2  + \omega _\alpha ^2 }}. \\
1525   \end{eqnarray*}
1526 < By the same way, the system coordinates become
1526 > In the same way, the system coordinates become
1527   \begin{eqnarray*}
1528   mL(\ddot x) & = &
1529    - \sum\limits_{\alpha  = 1}^N {\left\{ { - \frac{{g_\alpha ^2 }}{{m_\alpha  \omega _\alpha ^2 }}\frac{p}{{p^2  + \omega _\alpha ^2 }}pL(x) - \frac{p}{{p^2  + \omega _\alpha ^2 }}g_\alpha  x_\alpha  (0) - \frac{1}{{p^2  + \omega _\alpha ^2 }}g_\alpha  \dot x_\alpha  (0)} \right\}}  \\
# Line 1570 | Line 1547 | x_\alpha (0) - \frac{{g_\alpha  }}{{m_\alpha  \omega _
1547   & & + \sum\limits_{\alpha  = 1}^N {\left\{ {\left[ {g_\alpha
1548   x_\alpha (0) - \frac{{g_\alpha  }}{{m_\alpha  \omega _\alpha  }}}
1549   \right]\cos (\omega _\alpha  t) + \frac{{g_\alpha  \dot x_\alpha
1550 < (0)}}{{\omega _\alpha  }}\sin (\omega _\alpha  t)} \right\}}
1551 < \end{eqnarray*}
1552 < \begin{eqnarray*}
1553 < m\ddot x & = & - \frac{{\partial W(x)}}{{\partial x}} - \int_0^t
1554 < {\sum\limits_{\alpha  = 1}^N {\left( { - \frac{{g_\alpha ^2
1555 < }}{{m_\alpha  \omega _\alpha ^2 }}} \right)\cos (\omega _\alpha
1550 > (0)}}{{\omega _\alpha  }}\sin (\omega _\alpha  t)} \right\}}\\
1551 > %
1552 > & = & -
1553 > \frac{{\partial W(x)}}{{\partial x}} - \int_0^t {\sum\limits_{\alpha
1554 > = 1}^N {\left( { - \frac{{g_\alpha ^2 }}{{m_\alpha  \omega _\alpha
1555 > ^2 }}} \right)\cos (\omega _\alpha
1556   t)\dot x(t - \tau )d} \tau }  \\
1557   & & + \sum\limits_{\alpha  = 1}^N {\left\{ {\left[ {g_\alpha
1558   x_\alpha (0) - \frac{{g_\alpha }}{{m_\alpha \omega _\alpha  }}}
# Line 1653 | Line 1630 | or be determined by Stokes' law for regular shaped par
1630   which is known as the Langevin equation. The static friction
1631   coefficient $\xi _0$ can either be calculated from spectral density
1632   or be determined by Stokes' law for regular shaped particles. A
1633 < briefly review on calculating friction tensor for arbitrary shaped
1633 > brief review on calculating friction tensors for arbitrary shaped
1634   particles is given in Sec.~\ref{introSection:frictionTensor}.
1635  
1636   \subsubsection{\label{introSection:secondFluctuationDissipation}\textbf{The Second Fluctuation Dissipation Theorem}}
# Line 1674 | Line 1651 | And since the $q$ coordinates are harmonic oscillators
1651   \left\langle {q_\alpha  (t)q_\beta  (0)} \right\rangle & = &\delta _{\alpha \beta } \left\langle {q_\alpha  (t)q_\alpha  (0)} \right\rangle  \\
1652   \left\langle {R(t)R(0)} \right\rangle & = & \sum\limits_\alpha  {\sum\limits_\beta  {g_\alpha  g_\beta  \left\langle {q_\alpha  (t)q_\beta  (0)} \right\rangle } }  \\
1653    & = &\sum\limits_\alpha  {g_\alpha ^2 \left\langle {q_\alpha ^2 (0)} \right\rangle \cos (\omega _\alpha  t)}  \\
1654 <  & = &kT\xi (t) \\
1654 >  & = &kT\xi (t)
1655   \end{eqnarray*}
1656   Thus, we recover the \emph{second fluctuation dissipation theorem}
1657   \begin{equation}

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines