nlin0408038/ch2.tex
1: \chapter[\protect\uppercase{Background Information}]{Background Information}\label{ch.ch2}%
2: \addtocontents{toc}{\protect\vspace{0.25in}}
3: 
4: \section[Effective Quantum Hamiltonian and Polyads]{\underline{Effective Quantum Hamiltonian and Polyads}}
5: \addtocontents{toc}{\protect\vspace*{7pt}}
6: 
7: The molecular vibrational spectrum is often modeled by an effective Hamiltonian obtained from fitting the spectral levels \fn{The fit is to either the resolved experimental spectra or the spectra from theoretical calculations \cite{HCP}.}.  Taking $\hbar=1$, the general form of an effective Hamiltonian is: 
8: \begin{eqnarray}
9: \hat{H}_{eff} = \hat{H}_0+\sum \hat{V}_{2}^{ij}+\sum \hat{V}_{3}^{ijk}+...  \label{generalHeff} 
10: \end{eqnarray}
11: \noindent with
12: \begin{align}
13: \hat{H}_0 =& \sum_i \omega_i \left( \hat{n}_i + \frac{d_i}{2} \right) + \sum_{i,j;i \leq j} x_{ij} \left( \hat{n}_i + \frac{d_i}{2} \right) \left( \hat{n}_j + \frac{d_j}{2} \right) + \cdots \label{QH0} \\
14: \hat{V}_{2}^{ij}= & V_{ij}[(\hat{a}_i^\dagger)^m(\hat{a}_j)^n+(\hat{a}_j^\dagger)^n(\hat{a}_i)^m]  \\
15: \hat{V}_{3}^{ijk}=&V_{ijk}[(\hat{a}_i^\dagger)^m (\hat{a}_j)^n (\hat{a}_k)^p + (\hat{a}_j^\dagger)^n (\hat{a}_k^\dagger)^p (\hat{a}_i)^m]  \label{QHeff}
16: \end{align}
17: 
18: %=cite: Herzberg (II.281)
19: The $\hat{H}_0$ term is in the form of a Dunham expansion \cite{Herzberg}.  $\hat{n}_i$ is a zero-order mode (e.g. normal or local modes) number operator, whose eigenvalue is $n_i$.  $d_i$ is the degeneracy of mode $i$: $1$ for non-degenerate modes and $2$ for doubly degenerate modes (such as the bending of a linear molecule).  Each $\hat{V}_{2}^{ij}$ term in eqn. (\ref{generalHeff}) represents a resonance that couples the modes $i$ and $j$.  It exchanges $m$ quanta in mode $i$ and $n$ quanta in mode $j$.  $\hat{V}_{3}^{ijk}$ acts in a similar manner among three modes $i$, $j$ and $k$.  The operators $\hat{a}_i^\dagger$, $\hat{a}_i$ have matrix elements identical to those of harmonic raising and lowering operators, i.e.
20: \begin{align}
21: \hat{a}_i^\dagger | n_i \rangle &= \sqrt{n_i+1} \, | n_i+1 \rangle  \label{raising} \\
22: \hat{a}_i | n_i \rangle  &= \sqrt{n_i} \, | n_i-1 \rangle        \label{lowering}  \\
23: \hat{a}_i^\dagger \hat{a}_i  |n_i \rangle & =\hat{n}_i|n_i \rangle = n_i \, | n_i \rangle
24: %|\hat{a}_i^\dagger, \hat{a}_i | &= 1 
25: \end{align}
26: 
27: \noindent $|n_1, n_2, \cdots, n_N \rangle$ comprise a set of eigenstates of $\hat{H}_0$.  They are referred to as the {\it Zero Order States} (ZOS).  In the basis spanned by the ZOS, the matrix form of $\hat{H}_{eff}$ is obtained from eqns. (\ref{raising},\ref{lowering}).  Diagonalization of this matrix yields quantum eigenfunctions in terms of the ZOS.  In order to compare these eigenfunctions to the molecular coordinate space (such as bond length and angle), one needs to assume for each $n_i$ an oscillator model, e.g. of harmonic \cite{MartensEzra} or Morse \cite{RankinMiller} type.  
28: 
29: %== Polyad Numbers
30: 
31: The resonance coupling terms $\hat{V}$ cause the ZOS to mix in the eigenfunctions.  The quantum numbers $n_i$ then are no longer good quantum numbers.  However, certain linear combinations of them, known as the {\it polyad numbers}, may remain conserved in the fitting Hamiltonian (and approximately conserved in the exact molecular Hamiltonian).  In triatomic molecules like H$_2$O, there is often an approximate 2:1 frequency ratio between one normal stretching mode $n_1$ and the bending mode $n_2$ \cite{JaffeBrumer} \fn{Meanwhile, the other non-interacting stretching normal mode $n_3$ can be regarded as a ``spectator".  Because the number of quanta in it is constant, it is absorbed into the other parameters when we only consider a specific $n_3$ manifold.}.  This ratio leads to the inclusion of a {\it Fermi resonance} term in $\hat{H}_{eff}$.  The Fermi resonance (1) takes  one quantum out of $n_1$ and adds two quanta to $n_2$, and (2) takes two quanta out of $n_2$ and adds one quantum to $n_1$.  When the equilibrium configuration of the molecule is non-linear, the bending $n_2$ mode is singly degenerate ($d_2=1$).  The effective two-mode Hamiltonian with Fermi resonance is
32: \begin{align}
33: \hat{H}_{Fermi}=& \omega_1 (\hat{n}_1+\frac{1}{2}) +\omega_2 (\hat{n}_2+\frac{1}{2}) +x_{11} (\hat{n}_1+\frac{1}{2})^2 + x_{12}(\hat{n}_1+\frac{1}{2}) (\hat{n}_2+\frac{1}{2}) \nonumber\\
34: & \quad + x_{22}(\hat{n}_2+\frac{1}{2})^2 +V_{Fermi} [ \hat{a}_1^\dagger(\hat{a}_2)^2 + (\hat{a}_2^\dagger )^2 \hat{a}_1 ] \label{FermiExample}
35: \end{align}
36: \noindent The polyad number  
37: \begin{eqnarray}
38: \hat{P}=2\hat{n}_1 + \hat{n}_2 
39: \end{eqnarray}
40: \noindent remains conserved since it commutes with the Hamiltonian
41: \begin{eqnarray}
42: \mbox{[} \hat{P}, \hat{H}_{Fermi} \mbox{]} = \hat{P} \hat{H}_{Fermi}-\hat{H}_{Fermi} \hat{P} =0   \label{commu}
43: \end{eqnarray}
44: 
45: In the quantum Hamiltonian, the presence of $\hat{P}$ means that the resonance coupling only couples ZOS with the same polyad number.  For example, there are four ZOS $\vert n_1, n_2 \rangle$ interacting within $P=3$:
46: \begin{eqnarray}
47: \vert 3,0 \rangle \leftrightarrow \vert 2,2 \rangle \leftrightarrow \vert 1,4 \rangle \leftrightarrow \vert 0,6 \rangle \nonumber 
48: \end{eqnarray}
49: \noindent States belong to the same polyad appear in clusters in the spectra, as illustrated in Fig.~\ref{polyadclusters}.  Experimentally, it is often the observation of such clustering that leads to the adoption of a polyad model \cite{FieldPolyadRecog}.
50: 
51: \newpage \begin{figure}[hbtp]
52: %==\vspace{3.34in}
53: \begin{center} \includegraphics[width=3.46in]{polyadclusters.eps}\end{center}
54: \cpn{Polyad structures in spectra} {Polyad structures in spectra, a schematic illustration.  The states with the same polyad number $P$ may strongly interact with each other due to their closeness.  Interpolyad couplings, on the other hand, are relatively weak due to the large energy spacing between polyads. \label{polyadclusters}}
55: \end{figure}  
56: \newpage 
57: 
58: Under the polyad model there is no coupling between one polyad and another.  The Hamiltonian matrix therefore can be separated into blocks, each containing the intra-polyad couplings.  The blocks can be individually diagonalized, which substantially reduces the amount of computation involved. In the time domain, the existence of polyad(s) imposes an approximate restriction on the energy flow: IVR can only occur within the same polyad \cite{SmithWinn1}.  
59: 
60: A systematic method to locate polyad numbers was devised by Kellman \cite{KellmanVector}.  The method is closely related to an earlier van Vleck perturbation study by Fried and Ezra \cite{FriedEzra}.  Let there be $N$ zero-order vibrational modes, excluding the spectator ones.  Each resonance term $\hat{V}$ can be represented by a {\it resonance vector} in the $N$-dimensional linear space $\vec{V}_i = \{ n_1, n_2, \cdots n_N \}$.  All the resonance vectors $\vec{V}_i$ in $\hat{H}_{eff}$, taken together, form a linear subspace with $M$ dimensions ($M \leq N$).  Orthogonal to this subspace is another ($N-M$) dimensional subspace from which the ($N-M$) polyad numbers are found.  This is graphically illustrated in Fig.~\ref{polyads} for $N=3, M=2$.  The coefficients in the polyad  number can be taken to be any set of linearly independent vectors $\vec{P}_j$, which span the ($N-M$) dimensional subspace.  In the Fermi resonance system, for example, the resonance vector $\vec{V}_{Fermi}=\{1, -2 \}$ gives the polyad number $P=2n_1+n_2$, in accordance with the vector $\vec{P}=\{2, 1 \}$ orthogonal to $\vec{V}_{Fermi}$. 
61: 
62: \newpage  \begin{figure}[hbtp]
63: %==\vspace{2.94in}
64: \begin{center}\includegraphics[width=2.65in]{polyads.eps}\end{center}
65: \cpn{Locating a polyad number from resonance vectors} {Locating a polyad number from resonance vectors, a schematic illustration adapted from Fig.~1 of \cite{KellmanAnnRev}. \label{polyads}}
66: \end{figure} 
67: \newpage 
68: 
69: Although the total number of polyad numbers is fixed, each of them is not uniquely defined.  There is the liberty of multiplying $\vec{P}_j$ by an  arbitrary factor or linearly combining any number of $\vec{P}_j$.  The usual choice is to match their coefficients to the approximate integer ratios among the zero-order mode frequencies, as these ratios lead to the inclusion of the respective resonance terms in $\hat{H}_{eff}$ in the first place \cite{ReinhardtHynes}. 
70: 
71: Even when the more comprehensive Potential Energy Surface (PES) is available, effective Hamiltonians are often constructed from the PES with perturbative methods \cite{SibertC2H22,JoueuxSibertIso} in order to simplify the subsequent analysis.  The effective Hamiltonian is not only a more reliable model (compared to PES) for the highly excited states in triatomic or larger molecules, but also easily gives the useful insight of the polyad structure.   
72: 
73: The conservation of polyad numbers is never exact \fn{Physical laws of rigorous conservation are based on fundamental symmetry, known as Noether's Theorem (see Chapter 12.7 of \cite{Goldstein}). As an example, the conservation of energy and linear, angular momentum result from the homogeneity of time, space and the isotropy of space, respectively.  This is not the case with polyad numbers.}.  The degree of their conservation can be estimated by the uncertainty relationship $\Delta E \cdot \Delta t \geq \hbar$.  Spectral data recorded at low frequency resolution (larger $\Delta E$) decodes dynamics at shorter timescale (smaller $\Delta t$), and vice versa.  In molecules, spectral peaks well described by a polyad Hamiltonian may break into finer structures when scrutinized at finer frequency resolution.  This is caused by the small coupling terms not included in the effective Hamiltonian, which exercise their effects (including the breaking down of the polyad number) at longer time scales \cite{SEPReview}.  As an example, the acetylene pure bending Hamiltonian of Chapter 4 is fitted to spectra recorded with a resolution of 2 cm$^{-1}$ or finer.  The corresponding uncertainty $\Delta t=2.6$ picosecond is much longer than the bending vibration period (50 femtoseconds).  Since the polyad structure is still present at $2.6 ps$ time scale, the polyad numbers predicted from the effective Hamiltonian can be assumed valid at the same time scale or longer.
74: 
75: \section[Basic Concepts in Classical Mechanics]{\underline{Basic Concepts in Classical Mechanics}}
76: \addtocontents{toc}{\protect\vspace*{7pt}}
77: 
78: Although microscopic systems are governed by quantum mechanics, classical mechanics continues to be an important tool in understanding molecular processes due to the following fundamental and empirical reasons. (1) Quantum mechanics is built upon classical mechanics, as opposed to being a self-consistent theory.  Various semiclassical methods serve as a bridge between the quantum and classical worlds.  (2) Even when the classical description is not exact, it provides an intuitive tool for the human researcher, whose perceptions are unfortunately macroscopic and therefore classical, to understand the microscopic phenomena.  (3) In large and/or highly excited systems, treating the whole system quantum mechanically can become challenging, making classical and semiclassical methods useful supplements.
79: 
80: Below we discuss some basic concepts in classical mechanics (of the Hamiltonian formulation) that are pertinent to the topic of this thesis.  The reader is referred to Tabor \cite{Tabor} for a general introduction to classical mechanics with emphasis on the nonlinear dynamics.  The textbook by Goldstein \cite{Goldstein} may serve as a more comprehensive reference.
81:  
82: \subsection{2.2.1 Heisenberg's Correspondence Principle}
83: \addtocontents{toc}{\protect\vspace*{5pt}}
84: 
85: Heisenberg's Correspondence Principle provides an important connection between the quantum and classical worlds \cite{Heisenberg}.  It relates raising and lowering operators in quantum mechanics to Fourier components of action-angle variables in classical mechanics \cite{Child}:
86: \begin{align}
87: \hat{a}_i^\dagger & \rightarrow  \sqrt{n_i+\frac{d_i}{2}} e^{i\phi_i}= \sqrt{I_i} e^{i\phi_i}  \nonumber\\
88: \hat{a}_i & \rightarrow  \sqrt{n_i+\frac{d_i}{2}} e^{-i\phi_i} = \sqrt{I_i} e^{-i\phi_i}  \label{Heisenberg}
89: \end{align}
90: 
91: The $N$-mode quantum Hamiltonian $\hat{H}_{eff}$ is mapped to an $N$ DOF classical Hamiltonian $H_{eff}$ in canonical variables $(I_i, \phi_i)$. The $\hat{n}_i$ terms in $\hat{H}_0$ transform as
92: \begin{eqnarray}
93: I_i = n_i + \frac{d_i}{2}  \label{nitrans}
94: \end{eqnarray}
95: \noindent Hence, substitution of eqn. (\ref{Heisenberg}) into (\ref{FermiExample}) yields  
96: \begin{eqnarray}
97: H_{Fermi}=\omega_1 I_1+\omega_2 I_2 +x_{11} I_1^2+x_{12} I_1 I_2+x_{22} I_2^2 +2V_{Fermi}\sqrt{I_1^2 I_2} \cos[\phi_1-2\phi_2]   \label{Classical1}
98: \end{eqnarray}
99: 
100: Classically, the action-angle variables are defined as the conserved action $I_i$ and conjugate angle $\phi_i$ of the zero-order oscillator of mode $i$.  In particular, when the oscillator is 1-dimensional and harmonic, they can be related to the Cartesian coordinate and momentum by
101: \begin{align}  q_i  &= \sqrt{I_i} \cos \phi_i , & p_i & = \sqrt{I_i} \sin \phi_i  \end{align}
102: 
103: \subsection{2.2.2 Hamiltonian Classical Dynamics}
104: \addtocontents{toc}{\protect\vspace*{5pt}}
105: 
106: The $(I_i, \phi_i)$ variables form a set of canonically conjugate coordinates in Hamiltonian mechanics.  Their time evolution (also known as equations of motion) has the elegant form 
107: \begin{align}
108: \dot{I_i} =\frac{d I_i}{d t}  & = -\frac{\partial H}{\partial \phi_i} \label{EOM1} \\
109: \dot{\phi_i} = \frac{d \phi_i}{d t} &= \frac{\partial H}{\partial I_i}  \label{EOM2}
110: \end{align}
111: \noindent Once the initial condition $\{I_{i}(0), \phi_{i}(0) \}$ is given, the subsequent solution $\{I_i (t), \phi_i(t) \}$ is determined by integrating eqns. (\ref{EOM1},\ref{EOM2}).  $\{I_i (t), \phi_i(t) \}$ is known as a {\it phase space trajectory} or trajectory.  When the Hamiltonian $H$ is independent of time, a trajectory cannot intersect with itself in phase space, although it could retrace the same closed orbit repeatedly.
112: 
113: The equations of motion remain formally invariant.  After a transformation between two sets of canonical coordinates (called a {\it canonical transformation}), e.g. $ (I_i, \phi_i) \rightarrow (J_i, \Phi_i)$, we have:
114: \begin{align}
115: \dot{J}_i = & -\frac{\partial H}{\partial \Phi_i} \\
116: \dot{\Phi}_i = & \frac{\partial H}{\partial J_i}  
117: \end{align}
118: 
119: \subsection{2.2.3 Constants of Motion}
120: \addtocontents{toc}{\protect\vspace*{5pt}}
121: 
122: In a time-independent Hamiltonian, the energy is conserved, i.e. is a {\it constant of motion}.  Additional constants of motion may be present due to the polyad numbers.  In the Fermi resonance system, the polyad number $\hat{P}= 2\hat{n}_1+\hat{n}_2$ corresponds to a constant of motion through eqn. (\ref{nitrans}): 
123: \begin{eqnarray}
124: I = 2I_1+I_2 = P+ \frac{3}{2}
125: \end{eqnarray}
126: \noindent Like the quantum commutator in eqn. (\ref{commu}), the Poisson bracket between  $I$ and $H_{Fermi}$ also vanishes.  In order for the angle $\theta$ conjugate to $I$ to satisfy
127: \begin{eqnarray}
128: \dot{I} = -\frac{\partial H_{Fermi}}{\partial \theta} = 0
129: \end{eqnarray}
130: \noindent $\theta$ does not appear explicitly in the Hamiltonian.  Such a variable is known as a {\it cyclic angle} \cite{Tabor}.  This property leads  naturally to a canonical transformation $$(I_1, \phi_1, I_2, \phi_2) \rightarrow (I, \theta, I_z, \Psi)$$
131: \noindent with 
132: \begin{align}
133: I = & \frac{2I_1 +I_2}{2}, & \theta = & \phi_1+ 2\phi_2 \\
134: I_z = & \frac{2I_1-I_2}{2},  & \Psi = & \phi_1-2\phi_2
135: \end{align}
136: \noindent In the new coordinates, the classical Hamiltonian in eqn. (\ref{Classical1}) is expressed  as
137: \begin{align}
138: H_{Fermi}= & \omega_1(I+I_z)+\omega_2(I-I_z) +x_{11}(I+I_z)^2+x_{12}(I^2 - I_z^2)+x_{22}(I - I_z)^2 \nonumber\\
139: &+2V_{Fermi}\sqrt{(I+I_z)^2(I-I_z)} \cos \Psi   \label{FermiClassical}
140: \end{align}
141: 
142: \noindent Since its value does not change with time, $I$ can be regarded as an external parameter.  $\theta$ is absent from the Hamiltonian, and has limited physical significance.  The non-trivial part of the dynamics is captured in a 2-dimensional phase space $(I_z, \Psi)$, called the {\it reduced phase space}.  In general, in an $N$ DOF Hamiltonian with $(N-M)$ constants of motion, the phase space can be reduced from $2N$ to $2M$ dimensions by a similar transformation.  The details of such transformations are discussed in Appendix A.
143: 
144: A system is called {\it integrable} is there are as many constants of motion as the number of DOF.  In an integrable system, it is possible to express the Hamiltonian in $N$ constants of motion only (without their cyclic angles).  Then the equations of motion can be solved analytically without recourse to numerical integration.  A Hamiltonian with 1 DOF is always integrable when the energy is conserved.    
145: 
146: \subsection{2.2.4 Invariant Phase Space Structures}
147: \addtocontents{toc}{\protect\vspace*{5pt}}
148: 
149: An {\it invariant phase space structure} is defined as any lower-dimensional subset of the phase space that is mapped onto itself by the equations of motion.  These structures are the ``landmarks" that delineate regions in phase space with different kinds of dynamics.  The qualitative description of all these regions is called the {\it phase portrait} \cite{Kuznetsov}.
150: 
151: We are primarily interested in non-integrable systems, whose phase spaces are at least 4-dimensional and not easy to visualize directly.  In order to illustrate the role played by these landmarks, the following analogy seems appropriate.  Construction of the phase portrait of a new dynamical system can be compared to mapping out the hydrologic flow on an unknown continent \fn{This comparison certainly should not be taken literally.  For example, the autonomous flow in the phase space has neither ``sources" nor ``sinks".}.   Invariant phase space structures then serve purposes similar to the watersheds, rivers and lakes et cetera in the exploration of the continental water system.  In the cartoon map of Fig.~\ref{landmark}, the general flow trends on the surface is well characterized by the landmarks in the map, even though they do not account for the fate of every single raindrop, and whole regions may be left out of the picture (the areas marked with ``?").   When there is no prior knowledge about the system, finding these landmarks is the first step of a systematic exploration.
152: 
153: \newpage  \begin{figure}[hbtp]
154: %==\vspace{3.50in}
155: \begin{center}\includegraphics[width=3.13in]{atlantis.eps}\end{center}
156: \cpn{Hydrologic landmarks on an unknown continent} {Hydrologic landmarks on an unknown continent, a schematic illustration.  Regions with different hydrologic dynamics are summarized by a map with the watershed (peaks in the center), rivers and lakes.     \label{landmark}}
157: \end{figure}  
158: \newpage 
159: 
160: \noindent \underline {Critical Points} \,\,\,  The simplest invariant phase space structure is a {\it critical point} \cite{WigginsBook}.  These are defined as points where the equations of motion (\ref{EOM1}, \ref{EOM2}) vanish:
161: \begin{align}
162: \dot{I}_i = \dot{\phi}_i = - \frac{\partial H}{\partial \phi_i} = \frac{\partial H}{\partial I_i}=0   \label{CP}
163: \end{align}
164: \noindent Note that the term ``critical points" has also been used referring to where the gradient of a given function vanishes \cite{Weissen}, and the function may not be related to any dynamical property.  However, in Hamiltonian systems this definition coincides with the one defined above in eqn. (\ref{CP}).
165: 
166: Critical points are the simplest invariant structure because they have the lowest dimensionality 0, and because they can be exactly solved for as roots of simultaneous equations.
167: 
168: The {\it stability} of a critical point intuitively refers to the dynamics of trajectories in its neighborhood.  Near a stable (unstable) critical point, the dynamics can be compared to that near the minimum (maximum) of a classical potential.   In the illustration of Fig.~\ref{EH} (a), trajectories near a stable critical point are confined to the neighborhood, oscillating with small amplitude.  In panel (b), an unstable critical point behaves locally like a saddle point, with nearby trajectories deviating exponentially.  Mathematically, the stability of a critical point (or other invariant structures) has been defined in various ways to suit different purposes.  The most important ones are {\it Lyapunov}, {\it linear} and {\it spectral} stabilities.  Lyapunov stability implies linear stability, which in turn implies spectral stability \cite{Stability2}.  The linear stability has been widely used, as it is easy to calculate.  It will be the definition used in this thesis to characterize critical points; the exact derivation will be discussed in $\S$ 3.3.1.  
169: 
170: \newpage \begin{figure}[hbtp]
171: %==\vspace{2.5in}
172: \begin{center}  \includegraphics[width=4.68in]{e-h.eps}\end{center}  
173: \cpn{Dynamics near linearly stable and unstable critical points} {Dynamics near linearly stable and unstable critical points in a 1 DOF system.  Panel (a): trajectories near a linearly stable critical point. (b) trajectories near a linearly unstable critical point. \label{EH}}
174: \end{figure}  
175: \newpage 
176: 
177: \noindent \underline {Periodic Orbits} \,\,\,  A {\it Periodic Orbit} (PO) is a trajectory that retraces itself with a finite period $T$:
178: \begin{eqnarray}
179: \{I_{i} (n T), \phi_{i} (n T) \} = \{I_i (0) , \phi_i (0) \} \mbox{\,\,\,\,\,\,with\,\,} n=1,2,3, \ldots  \label{PODefinition}
180: \end{eqnarray}
181: A PO is 1-dimensional invariant structure in phase space.   Unlike the critical points, the only general way to locate a PO is through an iterative numerical search \cite{FarantosReview}.  
182: 
183: \noindent \underline {Invariant Tori} \,\,\,   Another example of invariant phase space structures is the {\it invariant torus}.   In an $N$ DOF integrable system, if $I_i$ are the $N$ constants of motion, then their conjugate angles $\theta_i$ evolve at constant frequencies $\dot{\theta}_i=\frac{\partial H}{\partial I_i}$ for any initial condition.   An example with $N=2$ is illustrated below in Fig.~\ref{invarianttorus}.  If $\dot{\theta}_1 : \dot{\theta}_2 $ happens to be an integer ratio $p:q$ (this condition is known as being {\it commensurable}), a trajectory will close on itself after $q$ periods in direction $\theta_1$ and $p$ periods in direction $\theta_2$.  The surface of the torus is therefore covered by a family of PO \fn{See the 3D model on the accompanying CD-ROM.}.   
184: 
185: If the frequencies are not commensurable, a trajectory gradually fills the entire toroidal surface without closing on itself for any finite time.  These {\it quasiperiodic} trajectories form a set of nesting N-dimensional tori, filling the entire phase space.   
186: 
187: \newpage \begin{figure}[hbtp]
188: %==\vspace{3.62in}
189: \begin{center} \includegraphics[width=3.6in]{invtorus.eps}\end{center}
190: \cpn{Trajectory on an invariant 2-torus.} {Trajectory on an invariant 2-torus. \label{invarianttorus}}
191: \end{figure} 
192: \newpage 
193: 
194: When small perturbations are added to an integrable Hamiltonian, some of its invariant tori are destroyed and others become deformed.  This is the conclusion according to the Kolmogorov-Arnold-Moser theorem \cite{Tabor}.  The existence of invariant tori in a non-integrable system indicates regions where the local dynamics resembles an integrable one.
195: 
196: In higher dimensions, there are also the {\it normal hyperbolic invariant manifolds} \cite{EzraH2O}, which act as impenetrable barriers in the phase space \cite{PhaseSpaceBarriers}.   In an isolated system, the ($2N-1$)-dimensional constant energy shell is also an invariant structure.
197: 
198: \subsection{2.2.5 Bifurcations}
199: \addtocontents{toc}{\protect\vspace*{5pt}}
200: 
201: A bifurcation generally refers to any qualitative change in the phase portrait, as some external control parameters are being varied \cite{Kuznetsov}.  The ``qualitative change" is typically labeled by the change in the number and/or stability of the invariant phase space structures.  The ``external control parameters" may be either variable physical quantities (such as the energy), or the coefficients of the Hamiltonian.
202: 
203: Fig.~\ref{dbwellbifur} illustrates the {\it pitchfork bifurcation} of critical points in a 1 DOF system \cite{Kuznetsov}.    As the potential $V(x)$ in the Hamiltonian is continuously deformed, suddenly the single well (stable critical point) lifts to a barrier (unstable critical point), and two additional wells are born.   The overall phase portrait changes accordingly, adding two zones corresponding to trajectories ``trapped" in the two new minima.
204: 
205: \newpage \begin{figure}[hbtp]
206: %==\vspace{3.9in}
207: \begin{center}  \includegraphics[width=4.98in]{pitchfork.eps}\end{center}  
208: \cpn{Pitchfork bifurcation in 1 DOF system} {Pitchfork bifurcation in 1 DOF system and the associated phase space change.  The right, middle and left panels are before, at and after the bifurcation point. \label{dbwellbifur}}
209: \end{figure}  
210: \newpage 
211: 
212: In this thesis, qualitative changes in the classical phase space are tracked by bifurcations of critical points in the reduced phase space.  The parameters in the effective Hamiltonian are regarded as given, and the polyad number(s) is the variable control parameter.
213: 
214: \subsection{2.2.6 Poincar\'{e} Surface of Section}
215: \addtocontents{toc}{\protect\vspace*{7pt}}
216: 
217: The phase space of a 2 DOF Hamiltonian is 4-dimensional and cannot be directly graphed like in Fig.~\ref{dbwellbifur}.  However, it may be visualized as a series of 2-dimensional slices.  This technique is called Poincar\'{e} {\it Surface Of Section} (SOS) \cite{Tabor}.  
218: 
219: The typical construction of an SOS proceeds as follows.  Let the 4 canonical coordinates be ($I_1, \phi_1, I_2, \phi_2$).  First, an energy value of interest is determined \fn{It can also be some other constant of motion that is held fixed instead of energy-- see the footnote in $\S$ 3.1.2.}, as well as a 2-dimensional dividing surface (e.g. by setting $\phi_2$ at a constant value).  An ensemble of trajectories at this energy and starting on the dividing surface is then calculated.  Their intersections with the dividing surface are recorded by two of the other independent coordinates (e.g. $I_1, \phi_1$), as illustrated in Fig.~\ref{SOS}.  Due to time-reversal symmetry, it is sufficient to record crossings in one direction only, such as by letting ${d \phi _2}/{d t} >0$. 
220: \newpage  \begin{figure}[hbtp]
221: %==\vspace{2.73in}
222: \begin{center}  \includegraphics[width=2.77in]{sos.eps}\end{center}  
223: \cpn{The construction of an SOS} {The construction of an SOS.  \label{SOS}}
224: \end{figure} 
225: \newpage 
226: 
227: The classical dynamics at this energy is reflected in patterns on the SOS.   Displayed in Fig.~\ref{fig2.2} are 4 SOS for the Henon-Heiles Hamiltonian \cite{HenonHeiles}, which consists of two coupled 1-dimensional oscillators
228: \begin{eqnarray}
229: H_{HH} = \frac{p_x^2+p_y^2}{2} +\frac{x^2+y^2}{2} + xy^2 -\frac{x^3}{3} 
230: \end{eqnarray}
231: 
232: \noindent In panel (a) there are two distinct types of trajectories.  For  each of the red, black, blue and green trajectories, the marks remain on a pair of closed curves.  It is because each of these trajectories lies on an invariant torus, and the two curves on the SOS are the result of slicing the torus with a plane (e.g. one that cuts along diameter $I_1$ in Fig.~\ref{invarianttorus}).  This kind of quasiperiodic motion is also known as {\it regular}.  The magenta trajectory, on the other hand, randomly fills an area complementary to the regular areas, indicating a lack of periodicity.  The corresponding random-looking trajectory is also known as {\it chaotic}.
233: 
234: \newpage  \begin{figure}[hbtp]  
235: %==\vspace{3.54in}
236: \begin{center}  \includegraphics[width=5.59in]{hhsos.eps}\end{center}  
237: \cpn{Regular, mixed and chaotic dynamics from an SOS} {Regular, mixed and chaotic dynamics from an SOS of the Henon-Heiles Hamiltonian.  Panel (a) was created in the online program at \cite{HHJava}.  It displays the marks from 5 individual trajectories coded by color.   Panels (b)-(d), adapted from \cite{Tabor}, are SOS taken at increasing energy.   The teardrop-shaped boundary in all panels is determined by the conservation of energy. \label{fig2.2}}
238: \end{figure}  
239: \newpage 
240: 
241: Panels (b)-(d) of Fig.~\ref{fig2.2} are of the same Hamiltonian with increasing energy.  At the lowest energy (panel b), most of the area on the SOS has a regular pattern.  In panel (c) there are both regular and chaotic regions.  At the highest energy (d), most of the SOS is filled with chaotic trajectories.   This regular-to-chaotic trend is typical for non-integrable systems.   In this thesis, we focus on systems with mixed dynamics like that depicted in panel (c).  The important question here is how to distinguish the regular dynamics from a sea of chaos.
242: 
243: \section[Quantum-Classical Correspondence]{\underline{Quantum-Classical Correspondence}}
244: \addtocontents{toc}{\protect\vspace{0.15in}}
245: 
246: According to the Bohr correspondence principle, the behavior of a quantum system converges to that of the corresponding classical system when Planck's constant $\hbar \rightarrow 0$, or when the quantum number approaches infinity.   A more recent theorem by Helton and Tabor indicates that in the $\hbar \rightarrow 0$ limit, quantum eigenstates must localize into phase space regions supporting an ``invariant measure", i.e. the invariant phase space structures \cite{HeltonTabor}.    
247: 
248: The correspondence between classical invariant phase space structure and quantum wavefunction has also been observed at finite $\hbar$ in a phenomenon called {\it localization} \cite{QMlocalization}.  A regular region of classical phase space corresponds to eigenfunctions with nodes positioned according to the invariant phase space structures.  This alignment is illustrated in Fig.~\ref{HCPWF}.  The nodal backbones of wavefunctions closely follow the PO labeled $\lbrack r \rbrack$, $\lbrack B \rbrack$,  and $\lbrack SN \rbrack$.  Wavefunction 4 corresponds to a combination of modes $\lbrack r \rbrack$ and  $\lbrack B \rbrack$, and the nodes form a rectangular grid distorted along these directions.  
249: 
250: \newpage \begin{figure}[hbtp]  
251: %==\vspace{3.0in}
252: \begin{center}  \includegraphics[width=3.0in]{fig04.eps}\end{center}  
253: \cpn{Localization of semiclassical wavefunctions} {Localization of semiclassical wavefunctions, adapted from \cite{HCP}.  The curves/lines labeled $\lbrack B \rbrack$, $\lbrack SN \rbrack$ and $\lbrack r \rbrack$ (which coincides with the left edge of upper left panel) are the periodic orbits that form the backbones of the wavefunctions.  The axes are the harmonic normal mode coordinates. \label{HCPWF}}  \end{figure}  
254: \newpage 
255: 
256: In a classically chaotic system, most eigenfunctions lack such regular patterns \cite{ReinhardtRegularChaotic}.  However, invariant phase space structures such as PO have been observed to have important influence, even when the overall dynamics is chaotic.   Examples with both stable and unstable PO are known.  The latter is called ``scarring" by Heller and co-workers \cite{HellerScarring}.   The wavefunctions could be assigned to new anharmonic modes by the PO.  
257: 
258: These intriguing observations of quantum-classical correspondence have led to renewed interest in using classical mechanics to understand molecular dynamics \cite{PollakReview}.  The goal is not only finding appropriate phase space structure to explain dynamics in an {\it a posteriori} manner, but also actively predicting the dynamics from analytically detected invariant phase space structures.
259: