1:
2:
3: \section{Resonance assisted tunneling in the kicked Harper model}
4: \label{sec:kh}
5:
6:
7: After the general discussion in the previous sections, we shall now
8: consider in more detail a particular system in the nearly integrable
9: regime, namely the kicked Harper model. The purpose of this section
10: will be to check, for this particular case, the accuracy of the final
11: semiclassical expression Eq.~(\ref{globresonsplit}) -- which gives the
12: splitting between the quasi-energy of a symmetric Floquet mode and
13: its antisymmetric counterpart -- as well as to verify the degree of validity
14: of the various hypotheses that were made along the way of its derivation.
15:
16:
17: The classical dynamics of the kicked Harper is governed by the
18: hamiltonian Eq.~(\ref{eq:kh_H}), yielding the stroboscopic map
19: Eq.~(\ref{eq:kh_map}). In the limit $\tau \to 0$, this dynamics is equivalent to
20: the one generated by the time-independent (integrable) Harper
21: hamiltonian
22: \begin{equation}
23: H_0(p,q) = \cos{p} + \cos{q} \; .
24: \label{eq:harper_ham}
25: \end{equation}
26: In Eq.~(\ref{eq:kh_H}), $\tau$ is thus both the period of the kick and the
27: perturbation parameter (i.e.\ $\epsilon \equiv \tau$).
28:
29:
30:
31:
32: Quantum mechanically, the map Eq.~(\ref{eq:kh_map}) can be associated
33: with the evolution operator
34: \begin{equation}
35: \widehat{U} = \exp\left(\frac{i\tau\cos\widehat{p}}{\hbar}\right)
36: \exp\left(\frac{i\tau\cos\widehat{q}}{\hbar}\right).
37: \end{equation}
38: The $2\pi$ periodicity in $\hat q$ and $\hat p$ makes the quantum treatment of
39: the kicked Harper particularly easy if
40: \begin{equation}
41: \hbar = \frac{2\pi}{N}
42: \end{equation}
43: with integer $N > 0$.
44: For these particular values of $\hbar$, the eigenfunctions $\psi$ of $\widehat{U}$
45: can be written as Bloch functions in both position and momentum -- i.e.,
46: \begin{eqnarray*}
47: \langle q+2\pi | \psi \rangle
48: & = &\exp(i\varphi_q) \langle q |
49: \psi\rangle \; ,\\
50: \langle p+2\pi | \psi \rangle & = &\exp(i\varphi_p) \langle p |
51: \psi\rangle
52: \end{eqnarray*}
53: for some pair of Bloch phases $0 \leq \varphi_q, \varphi_p < 2 \pi$, where $| q \rangle$ and
54: $| p \rangle$ denote the eigenfunctions of the position and momentum operator,
55: respectively.
56: For each pair $(\varphi_q,\varphi_p)$ of Bloch phases, the corresponding subspace of the
57: Hilbert space is {\em finite} dimensional and contains $N$ linearly
58: independent wave functions, spanned, e.g., by the basis states
59: \begin{equation}
60: |n\rangle = \sum_{l=-\infty}^{\infty}\exp(i(l+n/N)\varphi_q) \ |q \equiv ( 2 \pi n + \varphi_p ) / N + 2 \pi l \rangle
61: \end{equation}
62: for $0 \leq n < N$ \cite{Leboeuf90}.
63: The eigenvectors $|\psi_k\rangle$ of $\widehat{U}$ and their eigenphases $\phi_k$ can
64: therefore be computed up to numerical (quadruple) precision, by diagonalizing
65: the $N$$\times$$N$ matrix $\langle n | \widehat{U} | n' \rangle$.
66: %The classical kicked Harper Hamiltonian is a $2\pi$-periodic function
67: %of $p$ and $q$. This implies that the Hilbert space, and thus the
68: %eigenphases $\phi_k$ and eigenfunctions $\phi_k$ of
69: %$\widehat{U}(0,\tau)$, are divided in independent sectors
70: %characterized by a pair of Bloch phases $\varphi_q$ and $\varphi_p$
71: %such that for a wave function $ |\psi \rangle $ in the sector
72: %$(\varphi_q,\varphi_p)$
73: %\begin{eqnarray*}
74: % \langle q+2\pi | \psi \rangle
75: % & = &\exp(i\varphi_q) \langle q |
76: % \psi\rangle \; ,\\
77: % \langle p+2\pi | \psi \rangle & = &\exp(i\varphi_p) \langle p |
78: % \psi\rangle \; .
79: %\end{eqnarray*}
80: %Note that for these two conditions to be compatible, it is required that
81: %\begin{equation}
82: % \hbar = \frac{2\pi}{N}
83: %\end{equation}
84: %with the $N$ the number of states.
85: In the following, we shall consider only the two
86: pairs $(0,0)$ and $(0,\pi)$ of Bloch phases, corresponding to periodic
87: boundary conditions in momentum, and periodic or anti-periodic
88: boundary conditions in position. This choice is equivalent
89: to restricting $p$ to the interval $[-\pi,\pi]$ and $q$ to the interval
90: $[-\pi,3\pi]$ with periodic boundary conditions (see Fig.~\ref{fig:khmap}), and
91: to consider the even and odd symmetry classes with respect to the
92: inversion $q \to -q$.
93:
94:
95:
96:
97: The calculation of the integrable approximation $\widetilde{H}$ for
98: the kicked Harper is performed straightforwardly by applying the
99: formalism of Appendix \ref{app:lie}. One obtains for instance as
100: zeroth order coefficient the Harper Hamiltonian
101: Eq.~(\ref{eq:harper_ham}), and Eq.~(\ref{Hint}) for the approximation
102: of order three (recall that $\epsilon \equiv \tau$). In principle,
103: one may construct $\widetilde{H}^{(n)}$ up to orders as high as $n =
104: 20$ fairly easily with symbolic programs such as {\tt MAPLE}. As
105: mentioned in Sec.~\ref{sec:quasiint}, however, the series Eq.\
106: (\ref{seriesH}) of $\widetilde{H}^{(n)}$ tends to re-diverge beyond an
107: optimal order $n_0$, which, for $\tau = 1.$, is generally found around
108: $n_0 \simeq 6$. This is illustrated in Fig.~\ref{fig:asympH_n}: For
109: various orders $n$ of the integrable approximation, $40$ randomly
110: distributed initial phase space points have been
111: propagated during a given time by means of the kicked Harper map as
112: well as by its integrable approximation $\widetilde{H}^{(n)}$, and the
113: distance in phase space between the two resulting sets of final points
114: is plotted
115: as a function of $n$, yielding a minimum at rather moderate values ($n
116: \simeq 6$ in this particular example). We shall therefore mainly use
117: $\widetilde{H}^{(n)}$ with $n = 6$ in the following.
118:
119:
120: \subsection{Resonances parameters}
121:
122: Fig.~\ref{fig:khmap} and Fig.~\ref{fig:Htildemap} compare the phase space
123: portraits of the kicked Harper and of its integrable (6th order) approximation
124: in the near-integrable regime at $\tau = 1.0$. In fact, one observes that the
125: only significant difference between the two Poincar{\'e} sections is the
126: presence of the resonances. One may further note the relative importance
127: of $r$:$1$ resonances with $r = 10$ and $14$ as compared to the $8$:$1$ and
128: the $12$:$1$ resonances
129: (the absence of resonances with odd $r$ is an obvious consequence of the
130: rectangular symmetry of the kicked Harper).
131: As a matter of fact, these latter resonances, with $r$ a multiple of $4$,
132: are rather weakly developed at $\tau = 1$ and systematically exhibit $2 r$
133: (instead of $r$) islands in the Poincar{\'e} surface of section.
134: We conjecture that this behavior is a consequence of the initial square
135: symmetry of the Harper Hamiltonian, which is still relevant for small
136: values of $\tau$.
137: As the period in the center of the regular region is already larger than $6$
138: and monotonously increases when moving towards the separatrix, $r$:$s$
139: resonances with $r \leq 6$ do not exist at $\tau = 1$.
140:
141: To obtain a quantitative prediction for the tunneling rates, it is
142: necessary to characterize the resonances through the Fourier
143: coefficients $V^{r:s}_{r.m}$. This is done in practice by a direct
144: application of Eq.~(\ref{couplTPS}), i.e.\ by Fourier transforming the
145: function $\delta I_{r:s}(\theta) \equiv \delta I_{r:s}(p(I_{r:s},\theta),q(I_{r:s},\theta))$
146: %where $(p,q)(\theta) = (p,q)(\theta,I_{r:s})$, and
147: where $I_{r:s}$ is the
148: action of the resonant torus ${\Gamma}_{r:s}$.
149: %For a one-dimensional
150: %system,
151: On this torus, the angle variable is given by $\theta = \Omega_{r:s} t$, with
152: $\Omega_{r:s} = \partial H / \partial I (I_{r:s}) = 2\pi s/ (r \tau) $.
153: %Once
154: %an origin $(p_0,q_0)$ has been chosen on
155: %${\Gamma}_{r:s}$ , the phase space point $(p,q)(\theta)$ is obtained
156: %by classical propagation of $(p_0,q_0)$ during a time $t = \theta /
157: %\Omega_{r:s}$.
158: For a given $\theta$, $\delta I_{r:s}$ is computed
159: through the following successive steps:
160: i) Choose once for all a
161: reference point $(p_0,q_0)$ on the resonant torus ${\Gamma}_{r:s}$
162: of $\widetilde{H}^{(n_0)}$.
163: ii) Propagate $(p_0,q_0)$ under $\widetilde{H}^{(n_0)}$ dynamics during the
164: time $t = (r/s) (\theta /2\pi) \tau$.
165: iii) Apply the time reverse of the Poincar{\'e} map Eq.~(\ref{eq:kh_map}) on the
166: resulting point.
167: iv) Compute the difference between the action of this iterated point and
168: the action $I_{r:s}$ of ${\Gamma}_{r:s}$.
169: The values obtained
170: in this way for the $10$:$1$ resonance are plotted on
171: Fig.~\ref{fig:Vk_prl_tps}, for various orders $n$ of the
172: integrable approximation, showing that for $3 \leq n \leq 6$ the coefficients
173: $V^{r:s}_{r.m}$ do not depend sensitively on $n$. Also shown
174: in this figure are the values obtained by the method introduced in
175: \cite{prlbsu}, which is based on a Fourier analysis of the (pseudo-)separatrix
176: structure that is associated with the resonance.
177:
178:
179:
180:
181: Within our setting for the kicked Harper, the tunnel splitting is
182: defined as the difference
183: \begin{equation}
184: \delta \phi_k = |\phi_k(\varphi_q=0) - \phi_k(\varphi_q=\pi)|.
185: \end{equation}
186: As already stated, the exact quantum values of $\delta \phi_k$ can be calculated up to
187: numerical precision.
188: %The exact values of the $\delta \phi_k$ can be obtained by
189: %diagonalization of the matrix $\langle n | \widehat{U}(0,\tau) | m
190: %\rangle$, for instance in the basis $|m\rangle$ of the functions
191: %proportional to the Dirac distributions $\delta(q - (2\pi m +
192: %\varphi_p )/N)$ \cite{Leboeuf90}.
193: Using the coefficients $V^{r:s}_{r.m}$ obtained in the above way, as well
194: as the unperturbed energies $\tilde{E}_k$, the periods $T_{k}$ and the
195: tunneling actions $\sigma_{k}$ which are straightforwardly calculated from the
196: integrable approximation $\widetilde{H}^{(n)}$ of the kicked Harper,
197: these exact splittings can be compared with the ones derived from our
198: semiclassical expression Eq.~(\ref{globresonsplit}) based on the
199: resonance-assisted tunneling mechanism.
200:
201:
202:
203: Before performing this comparison,
204: %it can be however usefull to
205: let us first verify that the qualitative description of the tunneling
206: mechanism we gave in Sec.~\ref{sec:pendulum} and \ref{sec:mechanism}
207: actually applies in this particular example. To start with, we can
208: check that all the resonance involved in the tunneling process are
209: well within the quantum perturbative regime. Indeed, for the value of
210: the perturbation parameter we consider, $\tau=1$, the largest Fourier
211: coefficients for the resonances coming into play are $V^{8:1}_{16}
212: \simeq 9.0 \cdot 10^{-7}$, (as already stated, the $8$:$1$ resonance
213: exhibits 16 islands), $V^{10:1}_{10} \simeq 2.5 \cdot 10^{-4}$,
214: $V^{14:1}_{14} \simeq 9 \cdot 10^{-4}$, while, in the range of $\hbar$
215: we consider, the energy difference between quasi-degenerate states
216: with respect to the resonance is typically of the order of $\simeq
217: 10^{-2}$. Furthermore, taking into account the actual values of the
218: $V^{r:s}_{m.r}$ we observe that as $\hbar=2\pi/N$ gets smaller, higher
219: orders of the quantum perturbation theory become dominant in the
220: calculation of the transition amplitudes ${\mathcal
221: A}^{r:s}_{k,m}$. This can be specifically verified for the $10$:$1$
222: resonance: For this resonance, the $k \to k+20$ transitions are of
223: order one -- i.e., are dominated by the first-order perturbative
224: coupling terms -- for $N \leqsim 38$, but involve perturbation theory
225: of order two for $N \geqsim 38$. Similarly, we find that the $k \to k+30$
226: transitions are of order one for $N \leqsim 38$, of order two for $38
227: \leqsim N \leqsim 127$ and involve higher terms beyond ($k \to k+10$
228: transitions are, of course, always of order one).
229: %considers for instance couplings due to the $10$:$1$ resonance, it can
230: %be checked that the transitions $k \to k+10$ are, of course order one, the
231: %transition $k \to k+20$ are order one for $N$ smaller than $38$, but
232: %order two beyond, and for $k \to k+30$, order one dominates up to
233: %$N=38$, and order two up to $N=127$.
234: Effectively, one finds here the (possibly unusual) situation which will
235: generally be encountered in the semiclassical limit -- namely that the
236: lowest order terms of the perturbative expansion (which converges nevertheless
237: well) are not the dominating ones.
238:
239: Figures~\ref{fig:split1} and \ref{fig:split2} show for a varying value of
240: $\hbar$, i.e., a varying total number $N = 2 \pi / \hbar$ of states, the eigenphase
241: splittings of the eigenmode of $\widehat{U}$ that corresponds to a
242: fixed classical torus, with action $I=\pi/4$ in Fig.~\ref{fig:split1} and with
243: action $I=\pi/6$ in Fig.~\ref{fig:split2}.
244: Evidently, these splittings can be calculated only for particular values of
245: $N$, namely for $N = 4 ( 2 k + 1 )$ and $N = 6 ( 2 k + 1 )$ with $k = 0,1,\ldots$
246: in Fig.~\ref{fig:split1} and Fig.~\ref{fig:split2}, respectively, for which
247: this torus is selected by semiclassical quantization and supports the
248: $k$th excited quasi-mode.
249: % but for a varying value of $\hbar$, which means
250: %a varying value of the number of states $N$. Fig.~\ref{fig:split1}
251: %correspond to the classical torus of classical action $I=\pi/4$, and
252: %figure \ref{fig:split2} with the classical torus $I=\pi/6$.
253: In both cases, the perturbation parameter $\tau$ equals $1.0$.
254: The resonance involved are the $8$:$1$, $10$:$1$, and $14$:$1$.
255: We observe that the agreement between the
256: quantum and semiclassical results is extremely nice. For the
257: moderately small values of $\hbar$ that we consider, it is possible to
258: try all the possible coupling paths that participate at the tunneling
259: process, and in Figs.~\ref{fig:split1} and \ref{fig:split2}, the
260: semiclassical prediction is obtained by summing up all these
261: contributions. However, as shown on the lower panel of
262: Fig.~\ref{fig:split2}, where the action coordinates of the
263: intermediate states that participate at the dominant tunneling path are
264: displayed, we see here that, as discussed at the end of
265: section~\ref{sec:mechanism}, this dominant path is always such that
266: the number of steps is as large as possible, taking into account the
267: contraints due to $\hbar$.
268:
269:
270: Finally we show on Fig.~\ref{fig:split-ki} a comparison, for a fixed
271: value of $\hbar$ and a variable initial torus, between the exact quantum
272: mechanical splitting and the one calculated from the expression
273: corresponding to integrable tunneling, with no resonance coupling. We
274: observe on this figure that, although the two curves strongly differ in
275: the interior of the regular region, they match perfectly as
276: one gets close to the separatrix. This shows that the presence
277: of the separatrix does not introduce any additional effect (e.g.\ from a
278: small chaotic layer) to the tunneling mechanism.
279:
280:
281:
282:
283:
284: \subsection{Singularities of the invariant manifold of the integrable
285: approximation}
286:
287:
288: In addition to the numerical values of the coefficients $V^{r:s}_{r.m}$,
289: needed to obtain quantitative prediction for the tunneling rates, a
290: qualitative understanding of their behaviour, and in particular their
291: asymptotic properties for large $m$, is, as seen for instance in
292: section~\ref{sec:mechanism}, also required to guaranty that the
293: tunneling mechanism we propose is indeed the dominating one. Since
294: the $V^{r:s}_{r.m}$ are proportional to the Fourier coefficients of the
295: function $\delta I_{r:s} (\theta)$, their asymptotic behaviour is
296: related to the singularities of this function, for complex values of
297: the angle $\theta$.
298:
299:
300: Let us consider, more generally, for fixed values of the energy $E$
301: and the order $n$ of the integrable approximation, the invariant
302: manifold ${\Gamma}$ of $\widetilde{H}^{(n)}$, defined by the equation
303: $\widetilde{H}^{(n)}(p,q) = E$ and characterized by the angular frequency
304: $\Omega_0$.
305: Let a function $f(\theta)$ be defined on ${\Gamma}$ as
306: \begin{equation}
307: f(\theta) = \hat f (p(\theta),q(\theta)) \; ,
308: \end{equation}
309: where $\hat f(p,q)$ is an entire function of the phase space variables.
310: %Because $\hat f$ is an entire function,
311: As a consequence,
312: the singularities of
313: $f(\theta)$ are the ones of $(p,q)(\theta)$.
314: What we therefore need to study are the singularities of the analytic
315: continuation of $(p,q)(\theta)$ for complex angles $\theta$. Due to the
316: linear relation between $\theta$ and $t$, this analytic continuation
317: is straightforwardly constructed by propagation (under
318: $\widetilde{H}^{(n)}$) of some real initial point $(p_0,q_0)$ on ${\Gamma}$, taken
319: as the origin of the angle axis, over complex time $t$.
320: %$\widetilde{H}^{(n)}$ and for complex time of the point
321: %$(p_0,q_0)$ on ${\Gamma}$, assumed real, taken as the origin of the angle.
322: A singularity of $(p,q)(\theta)$ is an angle $\theta^\xi$ such that for the
323: time $t^\xi = \theta^\xi / \Omega$ the point $g^{\widetilde H}_{t} (p_0,q_0)$ goes to
324: infinity. Note that because of the existence of these
325: singularities, $(p,q)(\theta)$ actually depend not only on the final
326: time $\theta / \Omega_0$, but also on the homotopy class of the
327: path joining $t=0$ to $\theta / \Omega_0$ in complex the time plane. In other words,
328: $(p,q)(\theta)$ is a priori a multivalued function of $\theta$.
329:
330: To search for the singularities of $(p,q)(\theta)$, the first step will
331: consist in finding asymptotic expression describing the manifold
332: $\widetilde{H}^{(n)} (p,q) = E$ when the imaginary part of $p$ and/or $q$ goes
333: to infinity. For this purpose, we introduce the variables
334: \begin{eqnarray}
335: X & = & \exp(ip) \; , \\
336: Y & = & \exp(iq) \; .
337: \end{eqnarray}
338: In these new variables, the integrable approximation of the kicked Harper
339: Hamiltonian takes the polynomial form
340: \begin{equation}
341: \label{eq:HXY}
342: \widetilde{H}^{(n)} = \sum_{i,j=-(n+1)}^{+(n+1)} a_{ij}^{(n)} X^i Y^j \; ,
343: \end{equation}
344: with known real coefficient $a_{ij}^{(n)}$. For $H^{(0)}$ for
345: instance, the non-zero coefficients are $a_{-10}^{(0)} = a_{10}^{(0)}
346: = a_{0-1}^{(0)} = a_{01}^{(0)} = 1/2$.
347:
348:
349:
350: The manifold ${\Gamma}$ is invariant under the symmetries $s_O: (X
351: \mapsto 1/X, Y \mapsto 1/Y)$, $s_\Delta: (X \leftrightarrow Y)$, and
352: $\bar s: (X \mapsto 1/\bar X, Y \mapsto 1/\bar Y)$. Moreover, one can
353: check easily that if $\widetilde{H}^{(n)} [\tau] (X,Y) = E$, then
354: $\widetilde{H}^{(n)} [-\tau] (1/X,Y) = E$. We shall call $s_\tau$
355: this transformation, although this is not properly speaking a symmetry
356: of $\Gamma$. The asymptotic regions of ${\Gamma}$ -- i.e., the neighborhood of points
357: at infinity on ${\Gamma}$ -- can be obtained by application of one of the above
358: transformation from one region such that $\Im [p] \to +\infty$, (i.e.~ $X \to 0$ )
359: and $\Im [q]$ is either bounded or goes to $+\infty$ (i.e.~ $Y$ bounded).
360: For such regions, one can assume an asymptotic expression of the form
361: \begin{equation}
362: \label{eq:asymptot}
363: Y^\xi(X) = \gamma^\xi_{0} + \gamma^\xi_{1} X +
364: \gamma^\xi_{2} X^{2} + \cdots \; .
365: \end{equation}
366: where $\xi$ label the asymptotic region.
367: Introducing Eq.~(\ref{eq:asymptot}) in the expression
368: Eq.~(\ref{eq:HXY}) of the Hamiltonian to solve the equation
369: $\widetilde{H}^{(n)}(X,Y) = E$ yields a series of polynomial equations
370: for the coefficients $\gamma^\xi_l$, which can be solved order by order
371: to determine successively $\gamma^\xi_{0}$, $\gamma^\xi_{1}$,
372: $\gamma^\xi_{2}$, etc.. Again for the zeroth order Hamiltonian
373: $H^{(0)}$, the set of equations obtained in this way are
374: \begin{eqnarray*}
375: \gamma_{0} & = & 0 \\
376: \frac{1}{2} \left( \gamma_1 + 1 + \gamma_{0} \right)
377: & = & E \gamma_{0} \\
378: \gamma_2 / 2 + \gamma_1 \gamma_{0} + \gamma_{0}/2
379: & = & E \gamma_1 \\
380: \cdots && \cdots \; ,
381: \end{eqnarray*}
382: yielding $\gamma_{0} = 0$, $\gamma_{1} = -1$, $\gamma_{2} = -2E$,
383: $\cdots$. In other words, for small $X$, the manifold defined by the
384: implicit expression $\widetilde{H}^{(0)}(X,Y) = E$ admits the explicit
385: asymptotic expression
386: \begin{equation} \label{eq:asym_H0}
387: Y(X) = -X - 2E X^2 + \cdots \; .
388: \end{equation}
389:
390:
391: Using the above equation with $X$ small enough allows to find a point
392: with a large imaginary part for $p$, such that
393: $\widetilde{H}^{(0)}(p,q)$ is very close to $E$. This point can be
394: brought back to the energy $E$ by following the gradient of the
395: Hamiltonian, giving a point $(p^\xi,q^\xi)$ on the
396: $\widetilde{H}^{(0)}(p,q) = E$ manifold and in the asymptotic region
397: of large $\Im [p]$. From this point, we integrate Hamilton's
398: equations of motion choosing the path in the complex time in two
399: different ways: i) First we take a purely imaginary direction, until
400: $t=i t_I$ such that the trajectory crosses the real manifold
401: ${\Gamma}_{\mathbb R} = {\Gamma} \cap {\mathbb R}^2$. The imaginary
402: part of the angle coordinate of $(p^\xi,q^\xi)$ it then given by $-t_I
403: / \Omega$. ii) Then we start again from $(p^\xi,q^\xi)$ and choose
404: the complex phase of each time step $dt$ in such a way that the
405: imaginary part of $p$ remains constant. The time $t$ describes then a
406: small loop in the complex time plane that contains the singularity.
407: This gives the order of magnitude of the time distance between
408: $(p^\xi,q^\xi)$ and the singularity, which is in practice extremely
409: small as soon as $\Im(p)$ is taken reasonably large. For $n=0$, there
410: is only one independent (i.e. up to symmetries) singularity, and the
411: imaginary part of its time coordinate is just half of $t^\sigma$, the
412: imaginary time required to go from ${\Gamma}_{\mathbb R} \cap
413: [-\pi,\pi] \times [-\pi,\pi]$ to ${\Gamma}_{\mathbb R} \cap [\pi,3\pi
414: ] \times [-\pi ,\pi]$.
415:
416:
417:
418:
419: Such a procedure can be reproduced for various orders $n$ of the
420: integrable Hamiltonian, and we have performed it explicitly up to
421: $n=3$. Although the method we apply is basically the same, a few
422: important differences may be noticed
423: \begin{itemize}
424: \item[i)] The number of singularities (i.e. more precisely, of
425: asymptotic regions of the manifold) increases with the order of the
426: Hamiltonian. Counting only the number of independent singularities,
427: that is the ones that cannot be deduced one from each other by a
428: symmetry, there is only one for $n=0$, but $(2n+2)$ for $n = 1, 2,3$.
429: \item[ii)] If one starts form a point $(p^\xi,q^\xi)$ in an asymptotic
430: region such as Eq.~(\ref{eq:asymptot}) and propagates along a time path that
431: describes a small closed loop of infinitesimal radius around the
432: singularity in time plane, one can show that the real part of the
433: resulting momentum is not Re$[p^\xi]$, but Re$[p^\xi] + 2\pi/\ell^\xi$,
434: where $\ell^\xi$
435: is an integer which depends on the order $n$ of the integrable
436: approximation and on the singularity under consideration
437: %\cite{olivier:thesis}
438: ($\ell^\xi$ is equal to one for $H^{(0)}$ and
439: $H^{(1)}$, to two for $H^{(3)}$ and four of the singularities of
440: $H^{(2)}$, but to three for the two remaining ones). If one identifies
441: $p$ and $p+ 2\pi$, this means that for $n \neq 0,1$, $(p,q)(\theta)$
442: are not meromorphic functions.
443: Instead, the singularities are of logarithmic
444: type. More precisely, there are $\ell^\xi$ distinct sheets of the
445: manifold $(p,q)(\theta)$ around each singularity.
446: \item[iii)] As a consequence, when one tries to reach the complex torus
447: from the neighborhood of a singularity, one should specify on what
448: sheet one places oneself. Moreover, this implies that not all
449: singularities are ``visible'' from the real torus: assuming the best way
450: to compute the Fourier integral Eq.~(\ref{couplTPS}) is to shift the
451: integration contour in the imaginary direction, the only singularities
452: that will be encountered in this way are the ones that can be reached
453: by purely imaginary time propagation from the real manifold. For
454: $n=3$, only four out of the eight singularities are ``visible'' from the
455: real torus.
456: \item[iv)] Starting from the neighborhood of a ``visible'' singularity
457: and following the Hamiltonian flow, one may, depending on whether time
458: runs in the positive or negative imaginary direction, and depending
459: also on the chosen sheet of the manifold, cross the real manifold
460: ${\Gamma}_R$ in different cell $[(2l- 1) \pi, (2l+1)\pi] \times [(2l'- 1)\pi,
461: (2l'+1)\pi]$. Depending on the final cell, the time can be $\pm i
462: t^\xi$ or $\pm i (t^\sigma - t^\xi)$
463: \end{itemize}
464:
465:
466:
467: On Figs.~\ref{fig:txi_vs_tau:2} and \ref{fig:txi_vs_tau:3}, we plot,
468: for the resonant torus $10$:$1$ and as a function of the perturbation
469: parameter $\tau$, the imaginary part of the time coordinate of the
470: ``visible'' singularities of $\widetilde{H}^{(n)}$ for $n = 2$ and $3$
471: respectively. What we are waiting for is that the $k$ dependence of
472: the $V^{r:s}_{k}$ (for $k=rm$ as well as $k \neq rm$) is given by an
473: expression like Eq.~(\ref{varV}), with $t_{r:s}$ the imaginary part of
474: the time coordinate of the singularity closest to the real torus. On
475: Fig~\ref{fig:txi_vs_tau:2} and \ref{fig:txi_vs_tau:3} are also shown
476: the values $\theta_\xi$ obtained by fitting the numerically obtained
477: $V^{r:s}_{k}$ with the expression Eq.~(\ref{varV}). We observe that
478: for $n=2,3$ the variation of the fitted values follows nicely the
479: predicted ones. For higher $n$, and up to $n=6$, the $V^{r:s}_{k}$
480: are insensitive to the variation of the order of the approximation,
481: and therefore the fitted values remains on the curve corresponding to
482: the $n=3$ closest singularity.
483:
484:
485:
486: %\subsection{Discussion}
487:
488: The data shown in Fig~\ref{fig:txi_vs_tau:2} and
489: \ref{fig:txi_vs_tau:3} give a pretty convincing picture, which
490: justifies to use confidently the expression Eq.~(\ref{varV}) to
491: describe the behavior of the $V^{r:s}_{r.m}$ coefficients. Although we
492: believe this to be true from a practical point of view, one should, however,
493: resist the temptation to oversimplify this issue. Indeed, it is, to start
494: with, a priori not obvious to justify on a rigorous basis the form
495: $(mr)^{\gamma} V^{r:s}_{\xi_0}$ we have written for the prefactor, and
496: this can only be taken as a sensible parameterization. Moreover, even
497: if we did not extend the analysis of the location of all singularities
498: for $n$ greater than three because the approach described above
499: becomes somewhat cumbersome, it is still possible to locate the
500: closest singularity by a brute force search in the complex $\theta$
501: plane. Doing this for $n=4,5,6$ for the the torus $10$:$1$ at $\tau =
502: 1$ shows that this closest singularity slightly drifts as $n$
503: increase, and that its imaginary part goes from $0.8$ for $n=3$ to
504: $0.6$ for $n=6$,
505: %{\sf [caution, one should put the numbers here in
506: %concordance with what is shown on the figure (ie with or without the
507: %$\Omega$ factor)]},
508: in spite of the $V^{r:s}_{r.m}$ being not affected
509: by this change. This drift, although moderate, is still larger than
510: the numerical precision of our fit.
511: In accordance, if we take $\hat
512: f(p,q) \equiv \cos(p)$ as done of Fig.~\ref{fig:fk}, we see that,
513: contrary to $\delta I_{r:s}$, the Fourier coefficient of $f(\theta)$
514: changes with $n$ even when this latter is greater than three, and in
515: particular follows the asymptotic slope $0.6$ for $n=6$. This
516: indicates that although the basic picture we gave to interpret the
517: asymptotic behavior of the $V^{r:s}_{r.m}$ is certainly
518: correct, the complete description is presumably more complicated and
519: might involve, for instance, the link between the kicked Harper map
520: ${\mathcal T}$ and the integrable Hamiltonians $\widetilde{H}_{(n)}$, as
521: well as a more careful analysis of the different ranges in the
522: asymptotic behavior of the $V^{r:s}_{r.m}$.
523: