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 2718 by tim, Tue Apr 18 04:11:56 2006 UTC vs.
Revision 2720 by tim, Wed Apr 19 19:46:53 2006 UTC

# Line 822 | Line 822 | q(\Delta t)} \right]. %
822   %
823   q(\Delta t) &= q(0) + \frac{{\Delta t}}{2}\left[ {\dot q(0) + \dot
824   q(\Delta t)} \right]. %
825 < \label{introEquation:positionVerlet1}
825 > \label{introEquation:positionVerlet2}
826   \end{align}
827  
828   \subsubsection{\label{introSection:errorAnalysis}Error Analysis and Higher Order Methods}
# Line 883 | Line 883 | As a special discipline of molecular modeling, Molecul
883  
884   \section{\label{introSection:molecularDynamics}Molecular Dynamics}
885  
886 < As a special discipline of molecular modeling, Molecular dynamics
887 < has proven to be a powerful tool for studying the functions of
888 < biological systems, providing structural, thermodynamic and
889 < dynamical information.
890 <
891 < \subsection{\label{introSec:mdInit}Initialization}
892 <
893 < \subsection{\label{introSec:forceEvaluation}Force Evaluation}
886 > As one of the principal tools of molecular modeling, Molecular
887 > dynamics has proven to be a powerful tool for studying the functions
888 > of biological systems, providing structural, thermodynamic and
889 > dynamical information. The basic idea of molecular dynamics is that
890 > macroscopic properties are related to microscopic behavior and
891 > microscopic behavior can be calculated from the trajectories in
892 > simulations. For instance, instantaneous temperature of an
893 > Hamiltonian system of $N$ particle can be measured by
894 > \[
895 > T(t) = \sum\limits_{i = 1}^N {\frac{{m_i v_i^2 }}{{fk_B }}}
896 > \]
897 > where $m_i$ and $v_i$ are the mass and velocity of $i$th particle
898 > respectively, $f$ is the number of degrees of freedom, and $k_B$ is
899 > the boltzman constant.
900  
901 < \subsection{\label{introSection:mdIntegration} Integration of the Equations of Motion}
901 > A typical molecular dynamics run consists of three essential steps:
902 > \begin{enumerate}
903 >  \item Initialization
904 >    \begin{enumerate}
905 >    \item Preliminary preparation
906 >    \item Minimization
907 >    \item Heating
908 >    \item Equilibration
909 >    \end{enumerate}
910 >  \item Production
911 >  \item Analysis
912 > \end{enumerate}
913 > These three individual steps will be covered in the following
914 > sections. Sec.~\ref{introSec:initialSystemSettings} deals with the
915 > initialization of a simulation. Sec.~\ref{introSec:production} will
916 > discusses issues in production run, including the force evaluation
917 > and the numerical integration schemes of the equations of motion .
918 > Sec.~\ref{introSection:Analysis} provides the theoretical tools for
919 > trajectory analysis.
920  
921 + \subsection{\label{introSec:initialSystemSettings}Initialization}
922 +
923 + \subsubsection{Preliminary preparation}
924 +
925 + When selecting the starting structure of a molecule for molecular
926 + simulation, one may retrieve its Cartesian coordinates from public
927 + databases, such as RCSB Protein Data Bank \textit{etc}. Although
928 + thousands of crystal structures of molecules are discovered every
929 + year, many more remain unknown due to the difficulties of
930 + purification and crystallization. Even for the molecule with known
931 + structure, some important information is missing. For example, the
932 + missing hydrogen atom which acts as donor in hydrogen bonding must
933 + be added. Moreover, in order to include electrostatic interaction,
934 + one may need to specify the partial charges for individual atoms.
935 + Under some circumstances, we may even need to prepare the system in
936 + a special setup. For instance, when studying transport phenomenon in
937 + membrane system, we may prepare the lipids in bilayer structure
938 + instead of placing lipids randomly in solvent, since we are not
939 + interested in self-aggregation and it takes a long time to happen.
940 +
941 + \subsubsection{Minimization}
942 +
943 + It is quite possible that some of molecules in the system from
944 + preliminary preparation may be overlapped with each other. This
945 + close proximity leads to high potential energy which consequently
946 + jeopardizes any molecular dynamics simulations. To remove these
947 + steric overlaps, one typically performs energy minimization to find
948 + a more reasonable conformation. Several energy minimization methods
949 + have been developed to exploit the energy surface and to locate the
950 + local minimum. While converging slowly near the minimum, steepest
951 + descent method is extremely robust when systems are far from
952 + harmonic. Thus, it is often used to refine structure from
953 + crystallographic data. Relied on the gradient or hessian, advanced
954 + methods like conjugate gradient and Newton-Raphson converge rapidly
955 + to a local minimum, while become unstable if the energy surface is
956 + far from quadratic. Another factor must be taken into account, when
957 + choosing energy minimization method, is the size of the system.
958 + Steepest descent and conjugate gradient can deal with models of any
959 + size. Because of the limit of computation power to calculate hessian
960 + matrix and insufficient storage capacity to store them, most
961 + Newton-Raphson methods can not be used with very large models.
962 +
963 + \subsubsection{Heating}
964 +
965 + Typically, Heating is performed by assigning random velocities
966 + according to a Gaussian distribution for a temperature. Beginning at
967 + a lower temperature and gradually increasing the temperature by
968 + assigning greater random velocities, we end up with setting the
969 + temperature of the system to a final temperature at which the
970 + simulation will be conducted. In heating phase, we should also keep
971 + the system from drifting or rotating as a whole. Equivalently, the
972 + net linear momentum and angular momentum of the system should be
973 + shifted to zero.
974 +
975 + \subsubsection{Equilibration}
976 +
977 + The purpose of equilibration is to allow the system to evolve
978 + spontaneously for a period of time and reach equilibrium. The
979 + procedure is continued until various statistical properties, such as
980 + temperature, pressure, energy, volume and other structural
981 + properties \textit{etc}, become independent of time. Strictly
982 + speaking, minimization and heating are not necessary, provided the
983 + equilibration process is long enough. However, these steps can serve
984 + as a means to arrive at an equilibrated structure in an effective
985 + way.
986 +
987 + \subsection{\label{introSection:production}Production}
988 +
989 + \subsubsection{\label{introSec:forceCalculation}The Force Calculation}
990 +
991 + \subsubsection{\label{introSection:integrationSchemes} Integration
992 + Schemes}
993 +
994 + \subsection{\label{introSection:Analysis} Analysis}
995 +
996   \section{\label{introSection:rigidBody}Dynamics of Rigid Bodies}
997  
998   Rigid bodies are frequently involved in the modeling of different
# Line 927 | Line 1026 | rotation matrix $A$ and re-formulating Hamiltonian's e
1026   The break through in geometric literature suggests that, in order to
1027   develop a long-term integration scheme, one should preserve the
1028   symplectic structure of the flow. Introducing conjugate momentum to
1029 < rotation matrix $A$ and re-formulating Hamiltonian's equation, a
1029 > rotation matrix $Q$ and re-formulating Hamiltonian's equation, a
1030   symplectic integrator, RSHAKE, was proposed to evolve the
1031   Hamiltonian system in a constraint manifold by iteratively
1032 < satisfying the orthogonality constraint $A_t A = 1$. An alternative
1032 > satisfying the orthogonality constraint $Q_T Q = 1$. An alternative
1033   method using quaternion representation was developed by Omelyan.
1034   However, both of these methods are iterative and inefficient. In
1035   this section, we will present a symplectic Lie-Poisson integrator
# Line 1136 | Line 1235 | To reduce the cost of computing expensive functions in
1235     0 & { - \sin \theta _1 } & {\cos \theta _1 }  \\
1236   \end{array}} \right),\theta _1  = \frac{{\pi _1 }}{{I_1 }}\Delta t.
1237   \]
1238 < To reduce the cost of computing expensive functions in e^{\Delta
1239 < tR_1 }, we can use Cayley transformation,
1238 > To reduce the cost of computing expensive functions in $e^{\Delta
1239 > tR_1 }$, we can use Cayley transformation,
1240   \[
1241   e^{\Delta tR_1 }  \approx (1 - \Delta tR_1 )^{ - 1} (1 + \Delta tR_1
1242   )
1243   \]
1244 <
1146 < The flow maps for $T_2^r$ and $T_2^r$ can be found in the same
1244 > The flow maps for $T_2^r$ and $T_3^r$ can be found in the same
1245   manner.
1246  
1247   In order to construct a second-order symplectic method, we split the
# Line 1213 | Line 1311 | Moreover, \varphi _{\Delta t/2,V} can be divided into
1311   \varphi _{\Delta t}  = \varphi _{\Delta t/2,V}  \circ \varphi
1312   _{\Delta t,T}  \circ \varphi _{\Delta t/2,V}.
1313   \]
1314 < Moreover, \varphi _{\Delta t/2,V} can be divided into two sub-flows
1315 < which corresponding to force and torque respectively,
1314 > Moreover, $\varphi _{\Delta t/2,V}$ can be divided into two
1315 > sub-flows which corresponding to force and torque respectively,
1316   \[
1317   \varphi _{\Delta t/2,V}  = \varphi _{\Delta t/2,F}  \circ \varphi
1318   _{\Delta t/2,\tau }.
1319   \]
1320   Since the associated operators of $\varphi _{\Delta t/2,F} $ and
1321   $\circ \varphi _{\Delta t/2,\tau }$ are commuted, the composition
1322 < order inside \varphi _{\Delta t/2,V} does not matter.
1322 > order inside $\varphi _{\Delta t/2,V}$ does not matter.
1323  
1324   Furthermore, kinetic potential can be separated to translational
1325   kinetic term, $T^t (p)$, and rotational kinetic term, $T^r (\pi )$,
# Line 1251 | Line 1349 | generalized Langevin Dynamics will be given first. Fol
1349   mimics a simple heat bath with stochastic and dissipative forces,
1350   has been applied in a variety of studies. This section will review
1351   the theory of Langevin dynamics simulation. A brief derivation of
1352 < generalized Langevin Dynamics will be given first. Follow that, we
1352 > generalized Langevin equation will be given first. Follow that, we
1353   will discuss the physical meaning of the terms appearing in the
1354   equation as well as the calculation of friction tensor from
1355   hydrodynamics theory.
1356  
1357 < \subsection{\label{introSection:generalizedLangevinDynamics}Generalized Langevin Dynamics}
1357 > \subsection{\label{introSection:generalizedLangevinDynamics}Derivation of Generalized Langevin Equation}
1358  
1359 + Harmonic bath model, in which an effective set of harmonic
1360 + oscillators are used to mimic the effect of a linearly responding
1361 + environment, has been widely used in quantum chemistry and
1362 + statistical mechanics. One of the successful applications of
1363 + Harmonic bath model is the derivation of Deriving Generalized
1364 + Langevin Dynamics. Lets consider a system, in which the degree of
1365 + freedom $x$ is assumed to couple to the bath linearly, giving a
1366 + Hamiltonian of the form
1367   \begin{equation}
1368   H = \frac{{p^2 }}{{2m}} + U(x) + H_B  + \Delta U(x,x_1 , \ldots x_N)
1369 < \label{introEquation:bathGLE}
1370 < \end{equation}
1371 < where $H_B$ is harmonic bath Hamiltonian,
1369 > \label{introEquation:bathGLE}.
1370 > \end{equation}
1371 > Here $p$ is a momentum conjugate to $q$, $m$ is the mass associated
1372 > with this degree of freedom, $H_B$ is harmonic bath Hamiltonian,
1373   \[
1374 < H_B  =\sum\limits_{\alpha  = 1}^N {\left\{ {\frac{{p_\alpha ^2
1375 < }}{{2m_\alpha  }} + \frac{1}{2}m_\alpha  w_\alpha ^2 } \right\}}
1374 > H_B  = \sum\limits_{\alpha  = 1}^N {\left\{ {\frac{{p_\alpha ^2
1375 > }}{{2m_\alpha  }} + \frac{1}{2}m_\alpha  \omega _\alpha ^2 }
1376 > \right\}}
1377   \]
1378 < and $\Delta U$ is bilinear system-bath coupling,
1378 > where the index $\alpha$ runs over all the bath degrees of freedom,
1379 > $\omega _\alpha$ are the harmonic bath frequencies, $m_\alpha$ are
1380 > the harmonic bath masses, and $\Delta U$ is bilinear system-bath
1381 > coupling,
1382   \[
1383   \Delta U =  - \sum\limits_{\alpha  = 1}^N {g_\alpha  x_\alpha  x}
1384   \]
1385 < Completing the square,
1385 > where $g_\alpha$ are the coupling constants between the bath and the
1386 > coordinate $x$. Introducing
1387   \[
1388 < H_B  + \Delta U = \sum\limits_{\alpha  = 1}^N {\left\{
1389 < {\frac{{p_\alpha ^2 }}{{2m_\alpha  }} + \frac{1}{2}m_\alpha
1390 < w_\alpha ^2 \left( {x_\alpha   - \frac{{g_\alpha  }}{{m_\alpha
1391 < w_\alpha ^2 }}x} \right)^2 } \right\}}  - \sum\limits_{\alpha  =
1392 < 1}^N {\frac{{g_\alpha ^2 }}{{2m_\alpha  w_\alpha ^2 }}} x^2
1281 < \]
1282 < and putting it back into Eq.~\ref{introEquation:bathGLE},
1388 > W(x) = U(x) - \sum\limits_{\alpha  = 1}^N {\frac{{g_\alpha ^2
1389 > }}{{2m_\alpha  w_\alpha ^2 }}} x^2
1390 > \] and combining the last two terms in Equation
1391 > \ref{introEquation:bathGLE}, we may rewrite the Harmonic bath
1392 > Hamiltonian as
1393   \[
1394   H = \frac{{p^2 }}{{2m}} + W(x) + \sum\limits_{\alpha  = 1}^N
1395   {\left\{ {\frac{{p_\alpha ^2 }}{{2m_\alpha  }} + \frac{1}{2}m_\alpha
1396   w_\alpha ^2 \left( {x_\alpha   - \frac{{g_\alpha  }}{{m_\alpha
1397   w_\alpha ^2 }}x} \right)^2 } \right\}}
1398   \]
1289 where
1290 \[
1291 W(x) = U(x) - \sum\limits_{\alpha  = 1}^N {\frac{{g_\alpha ^2
1292 }}{{2m_\alpha  w_\alpha ^2 }}} x^2
1293 \]
1399   Since the first two terms of the new Hamiltonian depend only on the
1400   system coordinates, we can get the equations of motion for
1401   Generalized Langevin Dynamics by Hamilton's equations
1402   \ref{introEquation:motionHamiltonianCoordinate,
1403   introEquation:motionHamiltonianMomentum},
1404 < \begin{align}
1405 < \dot p &=  - \frac{{\partial H}}{{\partial x}}
1406 <       &= m\ddot x
1407 <       &= - \frac{{\partial W(x)}}{{\partial x}} - \sum\limits_{\alpha  = 1}^N {g_\alpha  \left( {x_\alpha   - \frac{{g_\alpha  }}{{m_\alpha  w_\alpha ^2 }}x} \right)}
1408 < \label{introEquation:Lp5}
1409 < \end{align}
1410 < , and
1411 < \begin{align}
1412 < \dot p_\alpha   &=  - \frac{{\partial H}}{{\partial x_\alpha  }}
1413 <                &= m\ddot x_\alpha
1414 <                &= \- m_\alpha  w_\alpha ^2 \left( {x_\alpha   - \frac{{g_\alpha}}{{m_\alpha  w_\alpha ^2 }}x} \right)
1415 < \end{align}
1404 > \begin{equation}
1405 > m\ddot x =  - \frac{{\partial W(x)}}{{\partial x}} -
1406 > \sum\limits_{\alpha  = 1}^N {g_\alpha  \left( {x_\alpha   -
1407 > \frac{{g_\alpha  }}{{m_\alpha  w_\alpha ^2 }}x} \right)},
1408 > \label{introEquation:coorMotionGLE}
1409 > \end{equation}
1410 > and
1411 > \begin{equation}
1412 > m\ddot x_\alpha   =  - m_\alpha  w_\alpha ^2 \left( {x_\alpha   -
1413 > \frac{{g_\alpha  }}{{m_\alpha  w_\alpha ^2 }}x} \right).
1414 > \label{introEquation:bathMotionGLE}
1415 > \end{equation}
1416  
1417 < \subsection{\label{introSection:laplaceTransform}The Laplace Transform}
1417 > In order to derive an equation for $x$, the dynamics of the bath
1418 > variables $x_\alpha$ must be solved exactly first. As an integral
1419 > transform which is particularly useful in solving linear ordinary
1420 > differential equations, Laplace transform is the appropriate tool to
1421 > solve this problem. The basic idea is to transform the difficult
1422 > differential equations into simple algebra problems which can be
1423 > solved easily. Then applying inverse Laplace transform, also known
1424 > as the Bromwich integral, we can retrieve the solutions of the
1425 > original problems.
1426  
1427 + Let $f(t)$ be a function defined on $ [0,\infty ) $. The Laplace
1428 + transform of f(t) is a new function defined as
1429   \[
1430 < L(x) = \int_0^\infty  {x(t)e^{ - pt} dt}
1430 > L(f(t)) \equiv F(p) = \int_0^\infty  {f(t)e^{ - pt} dt}
1431   \]
1432 + where  $p$ is real and  $L$ is called the Laplace Transform
1433 + Operator. Below are some important properties of Laplace transform
1434 + \begin{equation}
1435 + \begin{array}{c}
1436 + L(x + y) = L(x) + L(y) \\
1437 + L(ax) = aL(x) \\
1438 + L(\dot x) = pL(x) - px(0) \\
1439 + L(\ddot x) = p^2 L(x) - px(0) - \dot x(0) \\
1440 + L\left( {\int_0^t {g(t - \tau )h(\tau )d\tau } } \right) = G(p)H(p) \\
1441 + \end{array}
1442 + \end{equation}
1443  
1444 + Applying Laplace transform to the bath coordinates, we obtain
1445   \[
1446 < L(x + y) = L(x) + L(y)
1446 > \begin{array}{c}
1447 > p^2 L(x_\alpha  ) - px_\alpha  (0) - \dot x_\alpha  (0) =  - \omega _\alpha ^2 L(x_\alpha  ) + \frac{{g_\alpha  }}{{\omega _\alpha  }}L(x) \\
1448 > L(x_\alpha  ) = \frac{{\frac{{g_\alpha  }}{{\omega _\alpha  }}L(x) + px_\alpha  (0) + \dot x_\alpha  (0)}}{{p^2  + \omega _\alpha ^2 }} \\
1449 > \end{array}
1450   \]
1451 <
1451 > By the same way, the system coordinates become
1452   \[
1453 < L(ax) = aL(x)
1453 > \begin{array}{c}
1454 > mL(\ddot x) =  - \frac{1}{p}\frac{{\partial W(x)}}{{\partial x}} \\
1455 >  - \sum\limits_{\alpha  = 1}^N {\left\{ { - \frac{{g_\alpha ^2 }}{{m_\alpha  \omega _\alpha ^2 }}\frac{p}{{p^2  + \omega _\alpha ^2 }}pL(x) - \frac{p}{{p^2  + \omega _\alpha ^2 }}g_\alpha  x_\alpha  (0) - \frac{1}{{p^2  + \omega _\alpha ^2 }}g_\alpha  \dot x_\alpha  (0)} \right\}}  \\
1456 > \end{array}
1457   \]
1458  
1459 + With the help of some relatively important inverse Laplace
1460 + transformations:
1461   \[
1462 < L(\dot x) = pL(x) - px(0)
1462 > \begin{array}{c}
1463 > L(\cos at) = \frac{p}{{p^2  + a^2 }} \\
1464 > L(\sin at) = \frac{a}{{p^2  + a^2 }} \\
1465 > L(1) = \frac{1}{p} \\
1466 > \end{array}
1467   \]
1468 <
1330 < \[
1331 < L(\ddot x) = p^2 L(x) - px(0) - \dot x(0)
1332 < \]
1333 <
1334 < \[
1335 < L\left( {\int_0^t {g(t - \tau )h(\tau )d\tau } } \right) = G(p)H(p)
1336 < \]
1337 <
1338 < Some relatively important transformation,
1339 < \[
1340 < L(\cos at) = \frac{p}{{p^2  + a^2 }}
1341 < \]
1342 <
1343 < \[
1344 < L(\sin at) = \frac{a}{{p^2  + a^2 }}
1345 < \]
1346 <
1347 < \[
1348 < L(1) = \frac{1}{p}
1349 < \]
1350 <
1351 < First, the bath coordinates,
1352 < \[
1353 < p^2 L(x_\alpha  ) - px_\alpha  (0) - \dot x_\alpha  (0) =  - \omega
1354 < _\alpha ^2 L(x_\alpha  ) + \frac{{g_\alpha  }}{{\omega _\alpha
1355 < }}L(x)
1356 < \]
1357 < \[
1358 < L(x_\alpha  ) = \frac{{\frac{{g_\alpha  }}{{\omega _\alpha  }}L(x) +
1359 < px_\alpha  (0) + \dot x_\alpha  (0)}}{{p^2  + \omega _\alpha ^2 }}
1360 < \]
1361 < Then, the system coordinates,
1468 > , we obtain
1469   \begin{align}
1363 mL(\ddot x) &=  - \frac{1}{p}\frac{{\partial W(x)}}{{\partial x}} -
1364 \sum\limits_{\alpha  = 1}^N {\left\{ {\frac{{\frac{{g_\alpha
1365 }}{{\omega _\alpha  }}L(x) + px_\alpha  (0) + \dot x_\alpha
1366 (0)}}{{p^2  + \omega _\alpha ^2 }} - \frac{{g_\alpha ^2 }}{{m_\alpha
1367 }}\omega _\alpha ^2 L(x)} \right\}}
1368 %
1369 &= - \frac{1}{p}\frac{{\partial W(x)}}{{\partial x}} -
1370 \sum\limits_{\alpha  = 1}^N {\left\{ { - \frac{{g_\alpha ^2 }}{{m_\alpha  \omega _\alpha ^2 }}\frac{p}{{p^2  + \omega _\alpha ^2 }}pL(x)
1371 - \frac{p}{{p^2  + \omega _\alpha ^2 }}g_\alpha  x_\alpha  (0)
1372 - \frac{1}{{p^2  + \omega _\alpha ^2 }}g_\alpha  \dot x_\alpha  (0)} \right\}}
1373 \end{align}
1374 Then, the inverse transform,
1375
1376 \begin{align}
1470   m\ddot x &=  - \frac{{\partial W(x)}}{{\partial x}} -
1471   \sum\limits_{\alpha  = 1}^N {\left\{ {\left( { - \frac{{g_\alpha ^2
1472   }}{{m_\alpha  \omega _\alpha ^2 }}} \right)\int_0^t {\cos (\omega
# Line 1392 | Line 1485 | t)\dot x(t - \tau )d} \tau }  + \sum\limits_{\alpha  =
1485   (\omega _\alpha  t)} \right\}}
1486   \end{align}
1487  
1488 + Introducing a \emph{dynamic friction kernel}
1489   \begin{equation}
1396 m\ddot x =  - \frac{{\partial W}}{{\partial x}} - \int_0^t {\xi
1397 (t)\dot x(t - \tau )d\tau }  + R(t)
1398 \label{introEuqation:GeneralizedLangevinDynamics}
1399 \end{equation}
1400 %where $ {\xi (t)}$ is friction kernel, $R(t)$ is random force and
1401 %$W$ is the potential of mean force. $W(x) =  - kT\ln p(x)$
1402 \[
1490   \xi (t) = \sum\limits_{\alpha  = 1}^N {\left( { - \frac{{g_\alpha ^2
1491   }}{{m_\alpha  \omega _\alpha ^2 }}} \right)\cos (\omega _\alpha  t)}
1492 < \]
1493 < For an infinite harmonic bath, we can use the spectral density and
1494 < an integral over frequencies.
1495 <
1409 < \[
1492 > \label{introEquation:dynamicFrictionKernelDefinition}
1493 > \end{equation}
1494 > and \emph{a random force}
1495 > \begin{equation}
1496   R(t) = \sum\limits_{\alpha  = 1}^N {\left( {g_\alpha  x_\alpha  (0)
1497   - \frac{{g_\alpha ^2 }}{{m_\alpha  \omega _\alpha ^2 }}x(0)}
1498   \right)\cos (\omega _\alpha  t)}  + \frac{{\dot x_\alpha
1499 < (0)}}{{\omega _\alpha  }}\sin (\omega _\alpha  t)
1500 < \]
1501 < The random forces depend only on initial conditions.
1499 > (0)}}{{\omega _\alpha  }}\sin (\omega _\alpha  t),
1500 > \label{introEquation:randomForceDefinition}
1501 > \end{equation}
1502 > the equation of motion can be rewritten as
1503 > \begin{equation}
1504 > m\ddot x =  - \frac{{\partial W}}{{\partial x}} - \int_0^t {\xi
1505 > (t)\dot x(t - \tau )d\tau }  + R(t)
1506 > \label{introEuqation:GeneralizedLangevinDynamics}
1507 > \end{equation}
1508 > which is known as the \emph{generalized Langevin equation}.
1509 >
1510 > \subsubsection{\label{introSection:randomForceDynamicFrictionKernel}Random Force and Dynamic Friction Kernel}
1511 >
1512 > One may notice that $R(t)$ depends only on initial conditions, which
1513 > implies it is completely deterministic within the context of a
1514 > harmonic bath. However, it is easy to verify that $R(t)$ is totally
1515 > uncorrelated to $x$ and $\dot x$,
1516 > \[
1517 > \begin{array}{l}
1518 > \left\langle {x(t)R(t)} \right\rangle  = 0, \\
1519 > \left\langle {\dot x(t)R(t)} \right\rangle  = 0. \\
1520 > \end{array}
1521 > \]
1522 > This property is what we expect from a truly random process. As long
1523 > as the model, which is gaussian distribution in general, chosen for
1524 > $R(t)$ is a truly random process, the stochastic nature of the GLE
1525 > still remains.
1526 >
1527 > %dynamic friction kernel
1528 > The convolution integral
1529 > \[
1530 > \int_0^t {\xi (t)\dot x(t - \tau )d\tau }
1531 > \]
1532 > depends on the entire history of the evolution of $x$, which implies
1533 > that the bath retains memory of previous motions. In other words,
1534 > the bath requires a finite time to respond to change in the motion
1535 > of the system. For a sluggish bath which responds slowly to changes
1536 > in the system coordinate, we may regard $\xi(t)$ as a constant
1537 > $\xi(t) = \Xi_0$. Hence, the convolution integral becomes
1538 > \[
1539 > \int_0^t {\xi (t)\dot x(t - \tau )d\tau }  = \xi _0 (x(t) - x(0))
1540 > \]
1541 > and Equation \ref{introEuqation:GeneralizedLangevinDynamics} becomes
1542 > \[
1543 > m\ddot x =  - \frac{\partial }{{\partial x}}\left( {W(x) +
1544 > \frac{1}{2}\xi _0 (x - x_0 )^2 } \right) + R(t),
1545 > \]
1546 > which can be used to describe dynamic caging effect. The other
1547 > extreme is the bath that responds infinitely quickly to motions in
1548 > the system. Thus, $\xi (t)$ can be taken as a $delta$ function in
1549 > time:
1550 > \[
1551 > \xi (t) = 2\xi _0 \delta (t)
1552 > \]
1553 > Hence, the convolution integral becomes
1554 > \[
1555 > \int_0^t {\xi (t)\dot x(t - \tau )d\tau }  = 2\xi _0 \int_0^t
1556 > {\delta (t)\dot x(t - \tau )d\tau }  = \xi _0 \dot x(t),
1557 > \]
1558 > and Equation \ref{introEuqation:GeneralizedLangevinDynamics} becomes
1559 > \begin{equation}
1560 > m\ddot x =  - \frac{{\partial W(x)}}{{\partial x}} - \xi _0 \dot
1561 > x(t) + R(t) \label{introEquation:LangevinEquation}
1562 > \end{equation}
1563 > which is known as the Langevin equation. The static friction
1564 > coefficient $\xi _0$ can either be calculated from spectral density
1565 > or be determined by Stokes' law for regular shaped particles.A
1566 > briefly review on calculating friction tensor for arbitrary shaped
1567 > particles is given in Sec.~\ref{introSection:frictionTensor}.
1568  
1569   \subsubsection{\label{introSection:secondFluctuationDissipation}The Second Fluctuation Dissipation Theorem}
1570 < So we can define a new set of coordinates,
1570 >
1571 > Defining a new set of coordinates,
1572   \[
1573   q_\alpha  (t) = x_\alpha  (t) - \frac{1}{{m_\alpha  \omega _\alpha
1574   ^2 }}x(0)
1575 < \]
1576 < This makes
1575 > \],
1576 > we can rewrite $R(T)$ as
1577   \[
1578 < R(t) = \sum\limits_{\alpha  = 1}^N {g_\alpha  q_\alpha  (t)}
1578 > R(t) = \sum\limits_{\alpha  = 1}^N {g_\alpha  q_\alpha  (t)}.
1579   \]
1580   And since the $q$ coordinates are harmonic oscillators,
1581   \[
1582 < \begin{array}{l}
1582 > \begin{array}{c}
1583 > \left\langle {q_\alpha ^2 } \right\rangle  = \frac{{kT}}{{m_\alpha  \omega _\alpha ^2 }} \\
1584   \left\langle {q_\alpha  (t)q_\alpha  (0)} \right\rangle  = \left\langle {q_\alpha ^2 (0)} \right\rangle \cos (\omega _\alpha  t) \\
1585   \left\langle {q_\alpha  (t)q_\beta  (0)} \right\rangle  = \delta _{\alpha \beta } \left\langle {q_\alpha  (t)q_\alpha  (0)} \right\rangle  \\
1586 + \left\langle {R(t)R(0)} \right\rangle  = \sum\limits_\alpha  {\sum\limits_\beta  {g_\alpha  g_\beta  \left\langle {q_\alpha  (t)q_\beta  (0)} \right\rangle } }  \\
1587 +  = \sum\limits_\alpha  {g_\alpha ^2 \left\langle {q_\alpha ^2 (0)} \right\rangle \cos (\omega _\alpha  t)}  \\
1588 +  = kT\xi (t) \\
1589   \end{array}
1590   \]
1591 <
1435 < \begin{align}
1436 < \left\langle {R(t)R(0)} \right\rangle  &= \sum\limits_\alpha
1437 < {\sum\limits_\beta  {g_\alpha  g_\beta  \left\langle {q_\alpha
1438 < (t)q_\beta  (0)} \right\rangle } }
1439 < %
1440 < &= \sum\limits_\alpha  {g_\alpha ^2 \left\langle {q_\alpha ^2 (0)}
1441 < \right\rangle \cos (\omega _\alpha  t)}
1442 < %
1443 < &= kT\xi (t)
1444 < \end{align}
1445 <
1591 > Thus, we recover the \emph{second fluctuation dissipation theorem}
1592   \begin{equation}
1593   \xi (t) = \left\langle {R(t)R(0)} \right\rangle
1594 < \label{introEquation:secondFluctuationDissipation}
1594 > \label{introEquation:secondFluctuationDissipation}.
1595   \end{equation}
1596 + In effect, it acts as a constraint on the possible ways in which one
1597 + can model the random force and friction kernel.
1598  
1599   \subsection{\label{introSection:frictionTensor} Friction Tensor}
1600   Theoretically, the friction kernel can be determined using velocity
# Line 1622 | Line 1770 | where \delta _{ij} is Kronecker delta function. Invert
1770   B_{ij}  = \delta _{ij} \frac{I}{{6\pi \eta R}} + (1 - \delta _{ij}
1771   )T_{ij}
1772   \]
1773 < where \delta _{ij} is Kronecker delta function. Inverting matrix
1773 > where $\delta _{ij}$ is Kronecker delta function. Inverting matrix
1774   $B$, we obtain
1775  
1776   \[
# Line 1666 | Line 1814 | translation-rotation coupling resistance tensor do dep
1814   Form Equation \ref{introEquation:ResistanceTensorArbitraryOrigin},
1815   we can easily find out that the translational resistance tensor is
1816   origin independent, while the rotational resistance tensor and
1817 < translation-rotation coupling resistance tensor do depend on the
1817 > translation-rotation coupling resistance tensor depend on the
1818   origin. Given resistance tensor at an arbitrary origin $O$, and a
1819   vector ,$r_{OP}(x_{OP}, y_{OP}, z_{OP})$, from $O$ to $P$, we can
1820   obtain the resistance tensor at $P$ by

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines