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 2881 by tim, Fri Jun 23 20:21:54 2006 UTC vs.
Revision 2898 by tim, Tue Jun 27 03:22:42 2006 UTC

# Line 31 | Line 31 | F_{ij} = -F_{ji}
31   $F_{ji}$ be the force that particle $j$ exerts on particle $i$.
32   Newton's third law states that
33   \begin{equation}
34 < F_{ij} = -F_{ji}
34 > F_{ij} = -F_{ji}.
35   \label{introEquation:newtonThirdLaw}
36   \end{equation}
37
37   Conservation laws of Newtonian Mechanics play very important roles
38   in solving mechanics problems. The linear momentum of a particle is
39   conserved if it is free or it experiences no force. The second
# Line 73 | Line 72 | can only be described in cartesian coordinate systems.
72   \subsection{\label{introSection:lagrangian}Lagrangian Mechanics}
73  
74   Newtonian Mechanics suffers from two important limitations: motions
75 < can only be described in cartesian coordinate systems. Moreover, It
76 < become impossible to predict analytically the properties of the
75 > can only be described in cartesian coordinate systems. Moreover, it
76 > becomes impossible to predict analytically the properties of the
77   system even if we know all of the details of the interaction. In
78   order to overcome some of the practical difficulties which arise in
79   attempts to apply Newton's equation to complex system, approximate
# Line 85 | Line 84 | Hamilton's Principle may be stated as follows,
84  
85   Hamilton introduced the dynamical principle upon which it is
86   possible to base all of mechanics and most of classical physics.
87 < Hamilton's Principle may be stated as follows,
88 <
89 < The actual trajectory, along which a dynamical system may move from
90 < one point to another within a specified time, is derived by finding
91 < the path which minimizes the time integral of the difference between
93 < the kinetic, $K$, and potential energies, $U$.
87 > Hamilton's Principle may be stated as follows: the actual
88 > trajectory, along which a dynamical system may move from one point
89 > to another within a specified time, is derived by finding the path
90 > which minimizes the time integral of the difference between the
91 > kinetic, $K$, and potential energies, $U$.
92   \begin{equation}
93   \delta \int_{t_1 }^{t_2 } {(K - U)dt = 0} ,
94   \label{introEquation:halmitonianPrinciple1}
95   \end{equation}
98
96   For simple mechanical systems, where the forces acting on the
97   different parts are derivable from a potential, the Lagrangian
98   function $L$ can be defined as the difference between the kinetic
# Line 138 | Line 135 | p_i  = \frac{{\partial L}}{{\partial q_i }}
135   p_i  = \frac{{\partial L}}{{\partial q_i }}
136   \label{introEquation:generalizedMomentaDot}
137   \end{equation}
141
138   With the help of the generalized momenta, we may now define a new
139   quantity $H$ by the equation
140   \begin{equation}
# Line 146 | Line 142 | $L$ is the Lagrangian function for the system.
142   \label{introEquation:hamiltonianDefByLagrangian}
143   \end{equation}
144   where $ \dot q_1  \ldots \dot q_f $ are generalized velocities and
145 < $L$ is the Lagrangian function for the system.
146 <
151 < Differentiating Eq.~\ref{introEquation:hamiltonianDefByLagrangian},
152 < one can obtain
145 > $L$ is the Lagrangian function for the system. Differentiating
146 > Eq.~\ref{introEquation:hamiltonianDefByLagrangian}, one can obtain
147   \begin{equation}
148   dH = \sum\limits_k {\left( {p_k d\dot q_k  + \dot q_k dp_k  -
149   \frac{{\partial L}}{{\partial q_k }}dq_k  - \frac{{\partial
# Line 180 | Line 174 | t}}
174   t}}
175   \label{introEquation:motionHamiltonianTime}
176   \end{equation}
183
177   Eq.~\ref{introEquation:motionHamiltonianCoordinate} and
178   Eq.~\ref{introEquation:motionHamiltonianMomentum} are Hamilton's
179   equation of motion. Due to their symmetrical formula, they are also
# Line 227 | Line 220 | this system is a $6f$ dimensional space. A point, $x =
220   momentum variables. Consider a dynamic system of $f$ particles in a
221   cartesian space, where each of the $6f$ coordinates and momenta is
222   assigned to one of $6f$ mutually orthogonal axes, the phase space of
223 < this system is a $6f$ dimensional space. A point, $x = (q_1 , \ldots
224 < ,q_f ,p_1 , \ldots ,p_f )$, with a unique set of values of $6f$
225 < coordinates and momenta is a phase space vector.
226 <
223 > this system is a $6f$ dimensional space. A point, $x = (\rightarrow
224 > q_1 , \ldots ,\rightarrow q_f ,\rightarrow p_1 , \ldots ,\rightarrow
225 > p_f )$, with a unique set of values of $6f$ coordinates and momenta
226 > is a phase space vector.
227   %%%fix me
228 < A microscopic state or microstate of a classical system is
229 < specification of the complete phase space vector of a system at any
237 < instant in time. An ensemble is defined as a collection of systems
238 < sharing one or more macroscopic characteristics but each being in a
239 < unique microstate. The complete ensemble is specified by giving all
240 < systems or microstates consistent with the common macroscopic
241 < characteristics of the ensemble. Although the state of each
242 < individual system in the ensemble could be precisely described at
243 < any instance in time by a suitable phase space vector, when using
244 < ensembles for statistical purposes, there is no need to maintain
245 < distinctions between individual systems, since the numbers of
246 < systems at any time in the different states which correspond to
247 < different regions of the phase space are more interesting. Moreover,
248 < in the point of view of statistical mechanics, one would prefer to
249 < use ensembles containing a large enough population of separate
250 < members so that the numbers of systems in such different states can
251 < be regarded as changing continuously as we traverse different
252 < regions of the phase space. The condition of an ensemble at any time
228 >
229 > In statistical mechanics, the condition of an ensemble at any time
230   can be regarded as appropriately specified by the density $\rho$
231   with which representative points are distributed over the phase
232   space. The density distribution for an ensemble with $f$ degrees of
# Line 304 | Line 281 | thermodynamic equilibrium.
281   statistical characteristics. As a function of macroscopic
282   parameters, such as temperature \textit{etc}, the partition function
283   can be used to describe the statistical properties of a system in
284 < thermodynamic equilibrium.
285 <
286 < As an ensemble of systems, each of which is known to be thermally
310 < isolated and conserve energy, the Microcanonical ensemble (NVE) has
311 < a partition function like,
284 > thermodynamic equilibrium. As an ensemble of systems, each of which
285 > is known to be thermally isolated and conserve energy, the
286 > Microcanonical ensemble (NVE) has a partition function like,
287   \begin{equation}
288   \Omega (N,V,E) = e^{\beta TS} \label{introEquation:NVEPartition}.
289   \end{equation}
# Line 603 | Line 578 | them can be found in systems which occur naturally in
578   The hidden geometric properties\cite{Budd1999, Marsden1998} of an
579   ODE and its flow play important roles in numerical studies. Many of
580   them can be found in systems which occur naturally in applications.
606
581   Let $\varphi$ be the flow of Hamiltonian vector field, $\varphi$ is
582   a \emph{symplectic} flow if it satisfies,
583   \begin{equation}
# Line 617 | Line 591 | is the property that must be preserved by the integrat
591   \begin{equation}
592   {\varphi '}^T J \varphi ' = J \circ \varphi
593   \end{equation}
594 < is the property that must be preserved by the integrator.
595 <
596 < It is possible to construct a \emph{volume-preserving} flow for a
597 < source free ODE ($ \nabla \cdot f = 0 $), if the flow satisfies $
598 < \det d\varphi  = 1$. One can show easily that a symplectic flow will
625 < be volume-preserving.
626 <
627 < Changing the variables $y = h(x)$ in an ODE
594 > is the property that must be preserved by the integrator. It is
595 > possible to construct a \emph{volume-preserving} flow for a source
596 > free ODE ($ \nabla \cdot f = 0 $), if the flow satisfies $ \det
597 > d\varphi  = 1$. One can show easily that a symplectic flow will be
598 > volume-preserving. Changing the variables $y = h(x)$ in an ODE
599   (Eq.~\ref{introEquation:ODE}) will result in a new system,
600   \[
601   \dot y = \tilde f(y) = ((dh \cdot f)h^{ - 1} )(y).
602   \]
603   The vector filed $f$ has reversing symmetry $h$ if $f = - \tilde f$.
604   In other words, the flow of this vector field is reversible if and
605 < only if $ h \circ \varphi ^{ - 1}  = \varphi  \circ h $.
606 <
636 < A \emph{first integral}, or conserved quantity of a general
605 > only if $ h \circ \varphi ^{ - 1}  = \varphi  \circ h $. A
606 > \emph{first integral}, or conserved quantity of a general
607   differential function is a function $ G:R^{2d}  \to R^d $ which is
608   constant for all solutions of the ODE $\frac{{dx}}{{dt}} = f(x)$ ,
609   \[
# Line 646 | Line 616 | smooth function $G$ is given by,
616   which is the condition for conserving \emph{first integral}. For a
617   canonical Hamiltonian system, the time evolution of an arbitrary
618   smooth function $G$ is given by,
649
619   \begin{eqnarray}
620   \frac{{dG(x(t))}}{{dt}} & = & [\nabla _x G(x(t))]^T \dot x(t) \\
621                          & = & [\nabla _x G(x(t))]^T J\nabla _x H(x(t)). \\
622   \label{introEquation:firstIntegral1}
623   \end{eqnarray}
655
656
624   Using poisson bracket notion, Equation
625   \ref{introEquation:firstIntegral1} can be rewritten as
626   \[
# Line 666 | Line 633 | is a \emph{first integral}, which is due to the fact $
633   \]
634   As well known, the Hamiltonian (or energy) H of a Hamiltonian system
635   is a \emph{first integral}, which is due to the fact $\{ H,H\}  =
636 < 0$.
670 <
671 < When designing any numerical methods, one should always try to
636 > 0$. When designing any numerical methods, one should always try to
637   preserve the structural properties of the original ODE and its flow.
638  
639   \subsection{\label{introSection:constructionSymplectic}Construction of Symplectic Methods}
# Line 709 | Line 674 | simpler integration of the system.
674   \label{introEquation:FlowDecomposition}
675   \end{equation}
676   where each of the sub-flow is chosen such that each represent a
677 < simpler integration of the system.
678 <
714 < Suppose that a Hamiltonian system takes the form,
677 > simpler integration of the system. Suppose that a Hamiltonian system
678 > takes the form,
679   \[
680   H = H_1 + H_2.
681   \]
# Line 752 | Line 716 | to its symmetric property,
716   \begin{equation}
717   \varphi _h^{ - 1} = \varphi _{ - h}.
718   \label{introEquation:timeReversible}
719 < \end{equation},appendixFig:architecture
719 > \end{equation}
720  
721   \subsubsection{\label{introSection:exampleSplittingMethod}\textbf{Examples of the Splitting Method}}
722   The classical equation for a system consisting of interacting
# Line 1257 | Line 1221 | $T_{\star}SO(3)$. However, it turns out that under sym
1221   \]
1222  
1223   Unfortunately, this constraint manifold is not the cotangent bundle
1224 < $T_{\star}SO(3)$. However, it turns out that under symplectic
1224 > $T^* SO(3)$ which can be consider as a symplectic manifold on Lie
1225 > rotation group $SO(3)$. However, it turns out that under symplectic
1226   transformation, the cotangent space and the phase space are
1227   diffeomorphic. By introducing
1228   \[
# Line 1295 | Line 1260 | introduced to rewrite the equations of motion,
1260   introduced to rewrite the equations of motion,
1261   \begin{equation}
1262   \begin{array}{l}
1263 < \mathop \Pi \limits^ \bullet   = J^{ - 1} \Pi ^T \Pi  + Q^T \sum\limits_i {F_i (q,Q)X_i^T }  - \Lambda  \\
1264 < \mathop Q\limits^{{\rm{   }} \bullet }  = Q\Pi {\rm{ }}J^{ - 1}  \\
1263 > \dot \Pi  = J^{ - 1} \Pi ^T \Pi  + Q^T \sum\limits_i {F_i (q,Q)X_i^T }  - \Lambda  \\
1264 > \dot Q  = Q\Pi {\rm{ }}J^{ - 1}  \\
1265   \end{array}
1266   \label{introEqaution:RBMotionPI}
1267   \end{equation}
# Line 1326 | Line 1291 | matrix,
1291   \]
1292   Using \ref{introEqaution:RBMotionPI}, one can construct a skew
1293   matrix,
1294 < \begin{equation}
1295 < (\mathop \Pi \limits^ \bullet   - \mathop \Pi \limits^ {\bullet  ^T}
1296 < ){\rm{ }} = {\rm{ }}(\Pi  - \Pi ^T ){\rm{ }}(J^{ - 1} \Pi  + \Pi J^{
1297 < - 1} ) + \sum\limits_i {[Q^T F_i (r,Q)X_i^T  - X_i F_i (r,Q)^T Q]} -
1298 < (\Lambda  - \Lambda ^T ) . \label{introEquation:skewMatrixPI}
1299 < \end{equation}
1294 >
1295 > \begin{eqnarray*}
1296 > (\dot \Pi  - \dot \Pi ^T ){\rm{ }} = {\rm{ }}(\Pi  - \Pi ^T ){\rm{
1297 > }}(J^{ - 1} \Pi  + \Pi J^{ - 1} ) + \sum\limits_i {[Q^T F_i
1298 > (r,Q)X_i^T  - X_i F_i (r,Q)^T Q]}  - (\Lambda  - \Lambda ^T ).
1299 > \label{introEquation:skewMatrixPI}
1300 > \end{eqnarray*}
1301 >
1302   Since $\Lambda$ is symmetric, the last term of Equation
1303   \ref{introEquation:skewMatrixPI} is zero, which implies the Lagrange
1304   multiplier $\Lambda$ is absent from the equations of motion. This
# Line 1463 | Line 1430 | kinetic energy are listed in the below table,
1430   The equations of motion corresponding to potential energy and
1431   kinetic energy are listed in the below table,
1432   \begin{table}
1433 < \caption{Equations of motion due to Potential and Kinetic Energies}
1433 > \caption{EQUATIONS OF MOTION DUE TO POTENTIAL AND KINETIC ENERGIES}
1434   \begin{center}
1435   \begin{tabular}{|l|l|}
1436    \hline

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines