cond-mat0411280/rb.tex
1: \tolerance = 10000
2: \documentclass[twocolumn,showpacs,amsmath,amssymb]{revtex4}
3: %\documentclass[showpacs,preprintnumbers,amsmath,amssymb]{revtex4}
4: \usepackage{epsfig,psfrag}
5: \usepackage{dcolumn}% Align table columns on decimal point
6: \usepackage{bm}% bold math
7: \usepackage{graphicx}
8: \usepackage{color}
9: %\documentstyle[pra,aps,epsf,floats]{revtex}
10: \newcommand{\beq}{\begin{equation}}
11: \newcommand{\eeq}{\end{equation}}
12: \newcommand{\s}{{\sigma}}
13: \newcommand{\psib}{{\bar{\psi}}}
14: \newcommand{\rb}{{\bar{\rho}}}
15: \newcommand{\phib}{{\bar{\phi}}}
16: \newcommand{\dt}{{\Delta}}
17: \newcommand{\dva}{{\frac{\vp\times\va}{2\pi}-\rb}}
18: \newcommand{\dvab}{{\frac{\vp\times(\va-\vb)}{4p\pi}}}
19: \newcommand{\w}{{\omega}}
20: \newcommand{\zh}{{\hat{z}}}
21: \newcommand{\qh}{{\hat{q}}}
22: \newcommand{\vA}{{\vec{A}}}
23: \newcommand{\va}{{\vec{a}}}
24: \newcommand{\vrr}{{\vec{r}}}
25: \newcommand{\vj}{{\vec{j}}}
26: \newcommand{\vE}{{\vec{E}}}
27: \newcommand{\vB}{{\vec{B}}}
28: \def\ve{{\varepsilon}}
29: \def\bA{{\mathbf A}}
30: \def\bm{{\mathbf m}}
31: \def\bsig{{\mathbf \sigma}}
32: \def\bB{{\mathbf B}}
33: \def\bp{{\mathbf p}}
34: \def\bI{{\mathbf I}}
35: \def\bn{{\mathbf n}}
36: \def\bM{{\mathbf M}}
37: \def\bq{{\mathbf q}}
38: \def\br{{\mathbf r}}
39: \def\bs{{\mathbf s}}
40: \def\bS{{\mathbf S}}
41: \def\bQ{{\mathbf Q}}
42: \def\bs{{\mathbf s}}
43: \def\bB{{\mathbf B}}
44: \def\bl{{\mathbf l}}
45: \def\bPi{{\mathbf \Pi}}
46: \def\bJ{{\mathbf J}}
47: \def\bR{{\mathbf R}}
48: \def\bz{{\mathbf z}}
49: \def\ba{{\mathbf a}}
50: \def\bk{{\mathbf k}}
51: \def\bP{{\mathbf P}}
52: \def\bg{{\mathbf g}}
53: \def\bX{{\mathbf X}}
54: \def\a{{\alpha}}
55: \def\b{{\beta}}
56: \def\g{{\gamma}}
57: \def\d{{\delta}}
58: \newcommand{\etab}{\mbox{\boldmath $\eta $}}
59: \newcommand{\sigmab}{\mbox{\boldmath $\sigma $}}
60: \newcommand{\vx}{{\vec{x}}}
61: \newcommand{\vq}{{\vec{q}}}
62: \newcommand{\vQ}{{\vec{Q}}}
63: \newcommand{\vd}{{\vec{d}}}
64: \newcommand{\vb}{{\vec{b}}}
65: \newcommand{\vp}{{\vec{\partial}}}
66: \newcommand{\p}{{\partial}}
67: \newcommand{\gr}{{\nabla}}
68: \newcommand{\ra}{{\rightarrow}}
69: \def\lam{\lambda}
70: \def\dd{d^{\dagger}}
71: \def\sb{{\bar{s}}}
72: \def\half{{1\over2}}
73: \def\third{{1\over3}}
74: \def\twof{{2\over5}}
75: \def\threes{{3\over7}}
76: \def\rhob{{\bar \rho}}
77: \def\ua{\uparrow}
78: \def\da{\downarrow}
79: \def\eqa{\begin{eqnarray}}
80: \def\eea{\end{eqnarray}}
81: \def\beqr{\begin{eqnarray}}
82: \def\eeqr{\end{eqnarray}}
83: \parindent=4mm
84: %\addtolength{\textheight}{0.9truecm}
85: 
86: \begin{document}
87: 
88: \title{ Ballistic Quantum Dots with Disorder and Interactions: A numerical study on the Robnik-Berry billiard}
89: \author{    Ganpathy Murthy$^1$, R. Shankar$^2$, and Harsh Mathur$^3$}
90: \affiliation{$^1$Department of Physics and Astronomy, University of Kentucky, Lexington KY 40506-0055\\
91: $^2$ Department of Physics, Yale University, New Haven CT 06520\\ $^3$
92: Physics Department, Case Western Reserve University, Cleveland, OH
93: 44106-7079}
94: \date{\today}
95: %\tightenlines
96: %\widetext
97: %\advance\leftskip by 57pt
98: %\advance\rightskip by 57pt
99: \begin{abstract}
100: { In previous work we have found a regime in ballistic quantum
101: dots where interelectron interactions can be treated
102: asymptotically exactly as the Thouless number $g$ of the dot
103: becomes very large. However, this work depends on some assumptions
104: concerning the renormalization group and various properties of the
105: dot obeying Random Matrix Theory predictions at scales of the
106: order of the Thouless energy. In this work we test the validity of
107: those assumptions by considering a particular ballistic dot, the
108: Robnik-Berry billiard, numerically. We find that almost all of our
109: predictions based on the earlier work are borne out, with the
110: exception of fluctuations of certain matrix elements of
111: interaction operators. We conclude that, at least in the
112: Robnik-Berry billiard, one can trust the results of our previous
113: work at a qualitative and semi-quantitative level. }
114: \end{abstract}
115: \vskip 1cm \pacs{73.50.Jt}
116: \maketitle
117: 
118: 
119: \section{Introduction}
120: The quest for a satisfactory theory of quantum dots is driven not only
121: by their obvious importance as mesoscopic devices revealed by a series
122: of groundbreaking experiments\cite{recent-expts}, but also by their
123: challenge as a unique confluence of disorder, interactions and
124: finite-size effects\cite{reviews}. For weak interactions, the
125: Universal Hamiltonian\cite{H_U,univ-ham} (UH) provides a satisfactory
126: description. For ballistic/chaotic quantum dots, we have
127: espoused\cite{mm,qd-ms1,longpaper} an approach based on the fermionic
128: Renormalization Group\cite{rg-shankar} (RG), $1/N$ expansions and the
129: fact that energy eigenstates around the Fermi energy in disordered
130: systems ought to be described by Random Matrix Theory
131: (RMT)\cite{alt1,RMT}. Our approach not only explains the UH as a fixed
132: point of the RG but also describes the physics outside its basin of
133: attraction. It predicts a phase transition at strong coupling and
134: allows a fairly detailed study\cite{longpaper} of the new phase and
135: the quantum critical region\cite{critical-fan} separating it from that
136: governed by the UH.
137: 
138: Our results, however, were predicated on a variety of RMT and RG
139: assumptions. To test our assumptions and the conclusions deduced from
140: them, we performed a detailed numerical study on a ballistic but
141: chaotic billiard (the Robnik-Berry billiard\cite{robnik-berry}) and we
142: report our findings here. Our expectations are borne out, with one
143: notable exception.
144: 
145: We recall our strategy and assumptions briefly so that the reader may
146: see in advance what sort of ideas are put to test in our study. In the
147: primordial problem of interest to us one has electrons confined to a
148: ballistic dot of size $L$, with no impurities inside, and edges so
149: irregular that classical motion is chaotic. The electrons experience
150: the Coulomb interaction. In momentum space, all momenta within the
151: bandwidth (of order $k_F$, the Fermi momentum) exist.  The
152: semiclassical ergodicization time for an electron within the dot is a
153: few bounces, or $\tau_{erg}\simeq L/v_F$. By the uncertainty principle
154: this leads to an important energy scale, the Thouless energy
155: $E_T\simeq\hbar v_F/L$ which has a dual significance. First, it
156: controls the dimensionless conductance of a dot strongly coupled to
157: leads, as follows. Since the transport through the dot takes place in
158: a time such that energy is uncertain by an amount $E_T$, all single
159: particle states that fit into this band will each contribute a unit of
160: dimensionless conductance. If the average single-particle level
161: spacing is $\delta$, then the dimensionless conductance is
162: $g=E_T/\delta$. Second, in the other limit of dots very weakly coupled
163: to leads (which we focus on in this work), the Thouless band of width
164: $E_T$ centered on $E_F$, marks the scale deep within which RMT should
165: apply to the energies and eigenfunctions\cite{alt1}. In this context
166: $g$ is better denoted the Thouless number.
167: 
168: Since we are only interested in a narrow band of energies of width
169: $E_T <<E_F$, the first step in the program\cite{mm} is to use the RG
170: for fermions\cite{rg-shankar} to get an effective low energy theory by
171: eliminating all momentum states outside $E_T$.  Should we worry that
172: we are not eliminating exact single particle eigenstates (labelled
173: here by $\a$)? No, because the disorder due to the boundaries will mix
174: momentum states at roughly the same energy, and it does not matter
175: whether we eliminate momentum states within any annulus of energy
176: thickness $E_T$ or the single particle states they evolve
177: into. Indeed, even the mixing within $E_T$ is due to the fact that
178: momentum itself not well defined in a finite dot, a point we will
179: elaborate on shortly. However, once we come down to within $E_T$
180: of $E_F$, we cannot eliminate the remaining states in one shot since
181: it is the flow of couplings {\em within} this band that is all
182: important in the RG.
183: 
184:  Now it is
185: known\cite{rg-shankar} that the clean system RG (justified above)
186: leads to Landau's Fermi liquid interaction\cite{agd}
187: %
188: \beq
189: V=\sum_{\bk \bk'} F(\theta_{\bk}-\theta_{\bk'})\delta n(\bk
190: )\delta n(\bk' )
191: \eeq
192: %
193: at an energy scale $E_L$ which is small compared
194: to $E_F$. But since $E_L$ is a bulk scale it can always be made larger
195: than $E_T$ which vanishes as $L\to
196: \infty$. Thus Murthy and Mathur\cite{mm} perform their RG on the
197: hamiltonian (focussing on the spinless case for simplicity)
198: %
199: \beq H=\sum_{\a}c^{\dag}_{\a}c_{\a}\varepsilon_{\a}+
200: \sum\limits_{\a \b \g \d}V_{\a \b \g
201: \d}c^{\dag}_{\a}c^{\dag}_{\b}c_{\g}c_{\d}
202: \eeq
203: %
204: where
205: %
206: \beqr
207:  V_{\a \b
208: \g \d} =&{1\over4} \sum\limits_{ \bk \bk'}F(\theta_{\bk}-\theta_{\bk'})(\phi^{*}_{\a}(\bk )\phi^*_{\b}(\bk')-\phi^*_{\b}(\bk)\phi^*_{\a}(\bk'))\nonumber\\
209: &\times(\phi_\g(\bk')\phi_\d(\bk)-\phi_\d(\bk')\phi_\g(\bk)) \label{wof1}
210: \eeqr
211: %
212: is simply the Landau interaction written in the basis of exact
213: eigenstates, a statement that needs some elaboration. In usual RMT
214: treatments,   $\phi_{\a}(\bk )$ is the exact eigenstate $\a$ written
215: in the infinite dimensional basis of all momentum states. In our
216: version which uses  the RG to reduce the Hilbert space, the states
217: labeled by $\bk$ are approximate momentum states with an uncertainty
218: $\delta k\simeq 1/L$ in both directions. The number of such wave
219: packets that fit into an annulus of radius $k_F$ and thickness
220: $E_T/v_F$ is ${\mathcal{O} } (k_F L)=g$. We call them the
221: Wheel-of-Fortune (WOF) states, see Figure (\ref{wof}). One way to
222: construct such packets is to pick $g$ plane waves of equally spaced
223: momenta on the Fermi circle and to chop them off at the edges of the
224: dot to respect the boundary conditions. This is what we mean by
225: $\bk$ in $\phi_{\a}(\bk )=\langle \bk |\a\rangle$.
226: 
227: 
228: \begin{figure}[ht]
229: %\narrowtext
230: %\epsfxsize=2.4in\epsfysize=2.4in
231: \includegraphics*[width=2.4in,angle=0]{fig1.eps}
232: %\vskip 0.15in
233: \caption{The $g$ wheel-of-fortune states in the Thouless band.
234: They are packets in $\bk$ space with average momenta equally
235: spaced on the Fermi circle.} \label{wof}
236: \end{figure}
237: 
238: 
239:  We are now ready to state our two assumptions.\\
240: {\em {\bf Assumption 1}: We assume that the $g$ approximate
241: momentum states generated as above form a complete basis for the
242: $g$ exact eigenstates in the Thouless shell}.
243: 
244: {\bf Assumption II}: The energy eigenvalues $\varepsilon_{\a}$
245: obey RMT statistics as do the wavefunctions. For example we assume
246: that the ensemble averages (denoted by $\langle \rangle$) obey
247: \beq \langle \phi^{*}_{\mu}(\bk_1 )\phi^{}_{\mu}(\bk_2
248: )\phi^{*}_{\nu}(\bk_3 )\phi^{}_{\nu}(\bk_4) \rangle
249: ={\delta_{12}\delta_{34}\over g^2} +O(1/g^3) \label{4pt}\eeq
250: 
251: To some, Assumption I seems remarkable -- How can we furnish {\em in
252: advance, independent of the dot shape} a basis of $g$ states for
253: expanding the $g$ exact eigenstates within $E_T$? After all, these
254: eigenstates are supposed to resemble those of a random matrix.  The
255: point is that no matter how chaotic the dot, it can only mix states at
256: the same energy. While this sounds like Berry's
257: ansatz\cite{berry-ansatz} it is somewhat different in both scope and
258: content: Berry's ansatz states that every exact eigenstate $\a$ can be
259: expanded in terms of an {\it infinite number} of $\bk$ states (with
260: the same energy $\epsilon_{\a}=\bk^2/2m$)in the bulk, while we claim
261: that only $g$ of WOF states are needed. Secondly, we claim that the
262: {\it same} $g$ WOF states can be used to describe all the states
263: within the Thouless band.
264: 
265: In this work we will show that these states are nearly orthonormal and
266: that the exact state right in the middle of the band has 99.9\%
267: overlap with the WOF states. The success of this extension of Berry's
268: conjecture to a finite dot exceeds our expectations in this
269: regard. However, we find that the $g$ WOF states become less effective
270: at describing states as we move away from $E_F$: the overlap drops to
271: 50\% at $E_{\a}=E_F\pm E_T/2$. This
272: prepares us for the possibility that nonuniversal quantities may be
273: quantitatively inaccurate in our approach.
274: 
275: As for Assumption II, we have verified RMT behavior for the
276: eigenvalues (as have others before
277: us\cite{stone-bruus,alhassid-lewenkopf}) but not the eigenfunctions.
278: What we did instead was to see what extent the solution of a specific
279: dot resembled the picture we drew based on these two assumptions. We
280: begin by describing how one starts from Eqn. (\ref{wof1}), which
281: describes the effective hamiltonian, and use our two assumptions with
282: large-$N$ ideas to make our predictions\cite{qd-ms1,longpaper}. These
283: predictions are asymptotically exact as $g\to\infty$.
284: 
285: First one expands the Landau function as 
286: %
287: \beq F(\theta ) = \sum_m u_m
288: e^{im\theta}.
289: \eeq 
290: %
291: Barring accidents, the phase transition occurs in one channel with
292: some $m$ (recall superconductivity). This allows us to focus on a
293: single $u_m\ne 0$, ignoring all others. Then we carry out a
294: Hubbard-Stratovich transformation on the interaction using a
295: collective field $\sigmab$. We then formally integrate out the
296: fermions and get an effective action $S(\sigmab )$ for $\sigmab$.  In
297: this process we make use of assumption II. The action in terms of
298: $\sigmab$ is obtained by summing one loop Feynman diagrams with
299: varying numbers of external $\sigmab$'s connected to a single fermion
300: line running around the loop. Each diagram is a sum over fermion
301: energy denominators multiplying products of a string of $\phi_{\a}(\bk
302: )$'s.  We are able to show that these products may be replaced by
303: their ensemble averages in the large $g$ limit. In other words the sum
304: over so many terms in each diagrams leads to self-averaging. For the
305: averages we use relations like Eqn.  (\ref{4pt}). When this is done,
306: the effective action can be cast into a form which has a $g^2$ in
307: front of it\cite{qd-ms1,longpaper}, so that the saddle point gives
308: exact answers as $g \to \infty$. 
309: 
310: At this point let us collect all the results and predictions of the
311: RMT + large-$N$ theory\cite{longpaper} with a view to comparing them
312: with similar results without using Assumptions I and II on the
313: Robnik-Berry billiard.
314: \begin{itemize}
315: \item{}  In the large-$g$ limit there is a sharp transition to a
316: phase in which $\sigmab$ acquires a vacuum expectation value. The
317: critical value of $u_m$ in our approximation turns out to be
318: $-1/\log{2}$ in the spinless case and $-1/2\log{2}$ in the spinful
319: case. The true critical value is most likely to be the bulk value
320: $-2$ (spinless) or $-1$ (spinful), as has been found in an
321: explicitly solvable model by Adam, Brouwer, and Sharma
322: recently\cite{nowindow}. This is an example of the nonuniversal
323: quantity alluded to earlier, that we cannot predict exactly in our
324: approach even as $g\to \infty$.
325: 
326: \item{}  For finite $g$, instead of a sharp phase transition, there is
327: a crossover from the weak-coupling regime through a quantum critical
328: regime to a strong-coupling regime. Due to the explicit
329: symmetry-breaking at order $1/g$, there is always some nonzero order
330: parameter, which increases to a number of order $g$ (in the
331: normalization we use here, which is different from that of
332: ref.\cite{longpaper}) in the strong-coupling regime.
333: 
334: \item{} For symmetry-breaking in  odd angular momentum channels there are
335: two exactly degenerate minima for every sample arising from
336: time-reversal invariance.
337: 
338: \item{} The ground state energy at the minimum in the strong-coupling
339: regime is lower than that in the weak-coupling regime by a number of
340: order $g^2\d$.
341: 
342: \item{} The effective potential landscape in the strong-coupling
343: regime is in the approximate shape of a Mexican Hat, with the ripples
344: at the bottom of the hat being of order $g\d$.
345: 
346: \item{} In the quantum-critical and strong-coupling regimes,
347: even low-energy quasiparticles acquire large widths given on average by
348: %
349: \beq
350: \Gamma(\varepsilon)\approx {\d\over\pi}\log(\varepsilon/\d)
351: \eeq
352: %
353: \end{itemize}
354: 
355: We found that most of these predictions are verified by our
356: numerical results on the Robnik-Berry billiard, except that the
357: ripples at the bottom of the Mexican Hat turn out to be much
358: larger than expected for the $m=2$ Landau interaction channel.
359: 
360: %
361: \section{The Robnik-Berry billiard}
362: %
363: In this section we will describe how the dot is chosen and how the
364: single particle energy levels $\varepsilon_{\a}$ and eigenfunctions
365: $\phi_{\a}(\br )$ are determined. We use a trick invented by Robnik
366: and Berry\cite{robnik-berry} and elaborated upon by
367: Stone and Bruus\cite{stone-bruus}. Consider a unit circle $|z|=1$ in the complex
368: plane of $z=x+iy$. The analytic function
369: %
370: \beq
371: w(z) = {z+bz^2+cz^3e^{i\chi}\over \sqrt{1+2b^2+3c^2}}
372: \label{conformal-map}\eeq
373: %
374: defines a map
375: under which the unit circle in $z$ gets mapped into a new shape in
376: $w$, which will be our dot. The shape of the dot can be varied by
377: varying the parameters $b,\ c, \ \mbox{and}\ \chi$. The denominator
378: ensures that the billiard has the same area ($\pi$) as the unit disc.
379: The wavefunction $\phi_{\a}(w,\bar{w}) =\phi_{\a}(u, v) $ is required
380: to vanish at the boundary and obey
381: %
382: \begin{eqnarray}
383: &-&\left( {\p^2\over \p
384: u^2}+{\p^2\over \p v^2}\right) \phi_{\a}(u,v)\nonumber \\
385: &=&-4{\p \over \p w} {\p \over \p \bar{w}}\phi_{\a}(w,\bar{w})
386: =\varepsilon_{\a} \phi_{\a}(w,\bar{w}).
387: \end{eqnarray}
388: %
389: (We have
390: chosen $\hbar = 2m =1$). If we now go the $z$ plane where the
391: wavefunction is $\phi_{\a}(w(z),\bar{w}(\bar{z}))$, the
392: Schr\"{o}dinger equation and boundary condition are
393: %
394: \beq
395: -4{\p
396: \over \p z} {\p \over \p
397: \bar{z}}\phi_{\a}(z,\bar{z})=\varepsilon_{\a}|w'(z)|^2
398: \phi_{\a}(z,\bar{z}) \label{se} \ \ \ \ \phi_{\a}(|z=1|)=0
399: \eeq
400: %
401: where $w'(z)=dw/dz$. This differential equation in the continuum
402: is next converted to a discrete matrix equation by writing
403: %
404: \beq
405: \phi_{\a}(z,\bar{z})\equiv \phi_{\a}(r,\theta)=\sum_j {1\over
406: \g_j}C^{\a}_{j}\psi_j(r,\theta )\label{exp}
407: \eeq
408: %
409: where
410: $\psi_j(r,\theta ) $ is the solution to the  free Schr\"odinger
411: equation in the unit disk vanishing on the boundary:
412: %
413: \beq
414: -\nabla^2 \psi_j(r,\theta )=\g^{2}_{j}\psi_j(r,\theta ).
415: \eeq
416: %
417: (That is, these are Bessel functions in $r$ times angular momentum
418: eigenfunctions in $\theta$. ) Feeding this expansion into Eqn.
419: (\ref{se}) one obtains the matrix equation \beq \sum_j
420: M_{ij}C^{\a}_{j}={1 \over \varepsilon_{\a}}C^{\a}_{i}\eeq where
421: \beq M_{ij}={1\over \g_i}\langle i||w'|^2|j\rangle {1\over \g_j}.
422: \eeq (Without the $1/\g_j$ in the expansion Eqn. (\ref{exp}), $M$
423: would not have been Hermitian. )
424: 
425: In practice one truncates $M$ to a finite size (we used 585
426: states) and expects the lower energy levels and wavefunctions to
427: be unaffected.
428: 
429: The parameters $b,c,\chi$ are chosen to lie in the range where
430: classical behavior is chaotic, and where quantum chaos as reflected in
431: the eigenvalue distribution has been established\cite{stone-bruus}. A
432: value we used repeatedly was $b=c=.2,\chi =.85$. A nonzero $\chi$ ensures
433: that the billiard has no reflection symmetry. This shape is often
434: called the Africa Billiard based on its resemblance to that continent,
435: as seen in Fig \ref{fig1} for our chosen values of parameters.
436: 
437: \begin{figure}[ht]
438: %\narrowtext
439: %\epsfxsize=2.4in\epsfysize=2.4in
440: \includegraphics*[width=2.4in,angle=0]{fig2.eps}
441: %\vskip 0.15in
442: \caption{The shape of the Robnik-Berry billiard for $b=c=0.2$, and $\chi=0.85$.}
443: \label{fig1}
444: \end{figure}
445: 
446: We shall refer to these eigenfunctions and eigenvalues as exact even
447: though they come from solving a truncated problem because we can
448: easily increase the accuracy by increasing the size of the truncated
449: Hilbert space.
450: 
451: \subsection{ Testing Assumptions I and II}
452: 
453: Our ability to solve the Schr\"odinger equation (to high accuracy)
454: implies in principle that we can test our two assumptions.
455: 
456: In the next subsection we will test Assumption I, i.e.,  see in
457: detail how well the WOF states serve a basis within the Thouless
458: band.
459: 
460: As for Assumption II, we and our predecessors
461: \cite{stone-bruus,alhassid-lewenkopf} have shown that the
462: eigenvalues and single eigenfunctions obey the distribution
463: expected by RMT for a GOE. (The ensemble is generated by varying
464: the parameters in $w(z)$.)
465: 
466: Similar information about  wavefunction correlations is not known
467: in the ballistic problem (despite some recent progress using
468: supersymmetry methods\cite{supersymm}). We did not try to do this
469: here since  our computing capabilities did not allow us to
470: generate an ensemble.
471: 
472: Instead we computed the fate of the interacting system without
473: recourse to Assumptions I and II and compare  to our predictions
474: based on these assumptions.
475: 
476: \subsection{Completeness of the WOF basis}
477: Let $E_F$ be the Fermi energy. Then $g\simeq \sqrt{4\pi
478: N}\simeq\sqrt{\pi E_F}$, which we arrive at as follows. The Fermi
479: circle has a circumference $2\pi K_F$ and into this will fit
480: $g=2\pi K_F/(2\pi /L)$ WOF states each of width $2\pi /L$ in the
481: tangential direction. Finally $E_F=K_{F}^{2}/2m = K_{F}^{2}$,
482: $N=k_F^2L^2/4\pi$ and $L=\sqrt{\mbox{Area}}=\sqrt{\pi}$.
483: 
484: As a test case when we picked the Fermi energy to be the 100-th
485: level, we found $g=37$. How well is this state $|F>$ at $E_F$
486: spanned by the $g$ WOF states at the Fermi energy?
487: 
488: First we first take $g$ equally spaced points $\bk_n$ on the Fermi
489: circle and form the WOF states \beq \psi_{WOF-n}(\br ) = {1\over
490: \sqrt{\pi}}e^{i\bk \cdot \br} \Theta (\mbox{dot})\eeq where $\Theta
491: (\mbox{dot})$ is unity inside the dot and zero outside. These states
492: are very close to being orthonormal. For example the overlaps of $n=1$
493: state (with $\bk $ along the $y$-axis) with the others as we go around
494: the circle is shown in Fig. (\ref{fig2a}).
495: 
496: \begin{figure}[ht]
497: %\narrowtext
498: %\epsfxsize=2.4in\epsfysize=2.4in
499: \includegraphics*[width=2.4in,angle=0]{fig3.eps}
500: %\vskip 0.15in
501: \caption{The absolute value of $\langle n|1\rangle$, the inner
502: product of WOF state number 1 with the other$g-1=36$ states. }
503: \label{fig2a}
504: \end{figure}
505: 
506: 
507: Next we ask how much of the  state $|F>$ at the Fermi energy is
508: contained in the WOF states. We find $\sum_{n=1}^{g}|\langle
509: n-WOF|F\rangle |^2 = .9993$. This is a rather remarkable result. It
510: says that $|F>$, which is a vector with 580 components (which was
511: the size of our truncated problem) can be expanded almost completely
512: in terms of $g=37 $ WOF states which are given in advance. In other
513: words as one changes the shape of the dot and works at fixed Fermi
514: energy, the state $|F>$ changes in a random way, but that randomness
515: is only in which particular combination of WOF states describes it,
516: not in the completeness of the WOF basis.
517: 
518: While this is very satisfactory we need more to implement our scheme:
519: we need to be expand all $g$ states in the WOF basis. Here we find
520: that as we move off the center of the Thouless band, the fractional
521: norm captured by the WOF basis drops. In a typical case, with $g=37$,
522: there are roughly 12 states (one third of $g$) where the number lies
523: above 95\%. At band edge this drops to 50\%, as shown in Fig
524: (\ref{fig2b}).  Thus there is inevitably some error in transcribing
525: the Landau interaction written in terms of the WOF states labelled by
526: $\bk$ into the basis of $g$ exact eigenstates labelled by $\a$. This
527: just means that the location of the critical point will not be
528: correctly predicted by our RMT based analysis, as pointed out recently
529: by Adam, Brouwer, and Sharma\cite{nowindow}.
530: 
531: \begin{figure}[ht]
532: %\narrowtext
533: %\epsfxsize=2.4in\epsfysize=2.4in
534: \includegraphics*[width=2.4in,angle=0]{fig4.eps}
535: %\vskip 0.15in
536: \caption{The norm of the projection of the $n^{th}$ exact eigenstate from $E_F$ 
537: on to the subspace spanned by the WOF states. It can be seen that more
538: and more of the exact eigenstate lies outside this subspace as one
539: moves away from $E_F$. } \label{fig2b}
540: \end{figure}
541: 
542: 
543: This concludes our (partial) test of Assumptions I and II. We turn
544: to a comparison of our results based on these assumptions with a
545: direct solution of the problem with no recourse to the
546: assumptions.
547: 
548: \subsection{Hartree-Fock  solution of  interacting problem}
549: 
550: 
551: How can  the knowledge of the "exact" eigenfunctions and
552: eigenvalues in the billiard  help in the solution of the problem
553: with interactions? The tactic will be illustrated in schematic
554: form first. Suppose we have a four-Fermi  interaction added to a
555: free hamiltonian which in first-quantization is given by some
556: differential operator $H_0$. Then the path integral becomes
557: %
558: \beq Z=\int d\psi d\bar{\psi} e^{S}
559: \eeq
560: %
561: where
562: %
563: \beq S= \int
564: d\tau \left[ \bar{\psi} \left( i\partial_{\tau} - H_0 \right) \psi
565: +{u\over 2}  (\bar{\psi}\psi )^2\right] .
566: \eeq
567: %
568: Using a Hubbard-Stratanovich transformation we can write
569: %
570: \beq Z=\int d\psi d\bar{\psi} d\sigma
571: e^{S}
572: \eeq
573: %
574: where
575: %
576: \beq S= \int d\tau \left[ \bar{\psi} \left[
577: i\partial_{\tau} - H_0-\sigma  \right] \psi -{\sigma^2\over 2u}
578: \right] .
579: \eeq
580: %
581: If the fermions are integrated out we will get an effective action
582: $S_{eff}(\sigma )$. {\em To find the minimum we need just the
583: action for static $\sigma$. } In this case it is clear that
584: %
585: \beq
586: \int d\psi
587: d\bar{\psi}\exp \left[ \int d\tau  \bar{\psi} \left[
588: i\partial_{\tau} - H_0+\sigma \right] \psi \right]=
589: e^{-E_0(\sigma) T}
590: \eeq
591: %
592: where $T\to \infty$ is the length of the imaginary time $\tau$- axis
593: and $E_0(\sigma ) $ is the ground state energy of $\psi^{\dag}
594: (H_0+\sigma )\psi$. To find $E_0(\sigma )$ one simply solves for the
595: single particle levels of $ (H_0+\sigma )$ and fills up the ones with
596: negative energy. The effective action for static configurations, which
597: is also the effective potential, is
598: %
599: \beq
600: V_{eff}=E_0(\sigma )+{\sigma^2 \over 2u}.
601: \label{effpot}\eeq
602: %
603: 
604: At this point we have a mean-field theory. We still need to
605: justify its use by showing that fluctuations of the collective
606: field $\sigmab$ around its minimum are small. In our previous
607: work, based on Assumptions I and II we showed that the
608: fluctuations were indeed small in the limit of large $g$, since
609: the $g^2$ in front of the actions  limits fluctuations. In the
610: billiard we will justify the mean field similarly, based on the
611: depth and curvature of the minimum.
612: 
613: When the  Landau  interaction is factorized,  the hamiltonian
614: whose ground state  gives us $E_0(\sigma )$ is
615: %
616: \beq
617: \sum_{\a \b} \psi^{\dag}_{\a} (\delta_{\a \b}{\large
618: \varepsilon}_{\b}+\sigmab\cdot \bM_{\a \b})\psi_{\b}
619: \label{ham-to-diag}\eeq
620: %
621: where, for the case $m=1$, for example,
622: %
623: \beq \bM_{\a \b}=\sum_{\bk} \phi^{*}_{\a}(\bk )\phi_{\b}(\bk )
624: {\bk \over k}
625: \eeq
626: %
627: and $\a , \b, \bk$ are not restricted to the Thouless band. This is
628: because we want to solve the problem without any of the assumptions
629: that led to the effective low energy theory within the Thouless
630: band. Note that $\sigmab$ has two components, because the Landau
631: interaction associated with $u_m$ has two parts:
632: %
633: \beq
634: V_{L}= {u_m\over2}\sum_{\bk \bk'} \delta n_{\bk}
635: \delta n_{\bk'}(\cos m \theta_{\bk} \cos m \theta_{\bk'}+ \sin m
636: \theta_{\bk} \sin m \theta_{\bk'}).
637: \eeq
638: %
639: Once $S_{eff}$ is known (on a grid of points in the $\sigmab$
640: plane) one can ask if and when the minimum moves off the origin.
641: 
642: So far our considerations have been fairly generic, and the Landau
643: interaction has been written in momentum space. However, in testing
644: our approach in the billiard, we will find it more convenient to
645: represent the Landau interaction in real space, since the
646: eigenfunctions are known as linear combinations of Bessel functions
647: whose integrals are best carried out in real space. We have carried
648: out calculations for two Laudau parameters corresponding to $m=1$ and
649: $m=2$. The $m=1$ Landau interaction is chosen to be (in
650: second-quantized notation)
651: %
652: \beqr
653: \half &\int d^2r \Psi^{\dag}(\vrr){1\over(2mH_0)^{1/4}}(-i{\vec\nabla}){1\over(2mH_0)^{1/4}}\Psi(\vrr) \cdot\nonumber\\
654: &\times\int d^2r' \Psi^{\dag}(\vrr'){1\over(2mH_0)^{1/4}}(-i{\vec\nabla}'){1\over(2mH_0)^{1/4}}\Psi(\vrr')
655: \eeqr
656: %
657: The factors of ${1\over(2mH_0)^{1/4}}$ on each side of the $\nabla$
658: have the effect of $1/|\bk|$ in momentum space. Since momentum does
659: not commute with the free Hamiltonian $H_0$, the factors have to be
660: placed symmetrically. Note that this corresponds only to the $\vq=0$
661: part of the Landau interaction. In reality, all values of $\vq$ up to
662: the scale $E_L/v_F$ exist in the Hamiltonian. Depending on the shape
663: of the dot a particular combination of them may break symmetry to give
664: the best energy. Still, we expect that since at large $g$ we are close
665: to the zero-dimensional limit, the best combination will consist
666: largely of very small $\vq$ parts of the Landau interaction. In any
667: case, the energy of the true symmetry-broken state can only be
668: lower than what we calculate, so what we have here is a conservative
669: estimate of symmetry-breaking. Similarly the $m=2$ interaction (also
670: at $\vq=0$) is
671: %
672: \beqr
673: \half&\int d^2r \Psi^{\dag}(\vrr){1\over(2mH_0)^{1/2}}(\nabla_x^2-\nabla_y^2){1\over(2mH_0)^{1/2}}\Psi(\vrr) \cdot\nonumber\\
674: &\times\int d^2r'
675: \Psi^{\dag}(\vrr'){1\over(2mH_0)^{1/2}}((\nabla')_x^2-(\nabla')_y^2){1\over(2mH_0)^{1/2}}\Psi(\vrr')\nonumber\\
676: &+\half\int d^2r
677: \Psi^{\dag}(\vrr){1\over(2mH_0)^{1/2}}2\nabla_x\nabla_y{1\over(2mH_0)^{1/2}}\Psi(\vrr)
678: \nonumber\\
679: &\times\int d^2r'
680: \Psi^{\dag}(\vrr'){1\over(2mH_0)^{1/2}}2(\nabla')_x(\nabla')_y{1\over(2mH_0)^{1/2}}\Psi(\vrr')
681: \eeqr
682: %
683: The integrals are over $(w,\bar{w})$, but can be converted to
684: integrals over the disk by using the conformal mapping of Eq.
685: (\ref{conformal-map}). Of course, the derivative operators must
686: also be transformed in the process. In order to find the matrix
687: elements of $\bM_{\a \b}$ we had to take the matrix elements of
688: the above operators in the basis of exact billiard states. We
689: carried out the angular part of the integrals analytically, but
690: had to turn to numerical integration to evaluate the radial
691: integrals. This is a computationally  intensive calculation, but
692: once the matrix $\bM$ has been constructed, one simply
693: diagonalizes the Hamiltonian of Eq. (\ref{ham-to-diag}) for a mesh
694: of $\sigmab$ in the plane, adds up the energies of the lowest $N$
695: particles to obtain the fermionic ground state energy, and obtains
696: the effective potential landscape from Eq. (\ref{effpot}) for
697: various values of the coupling strength $u$. After this, it is a
698: simple matter to identify the global minimum, which gives us the
699: lowering of ground state energy and the value of the order
700: parameter as a function of $u$.
701: 
702: Let us proceed to the results,  displayed in pictorial form. In
703: Fig. \ref{fig3} we show the absolute value of the order parameter,
704: normalized by the nominal value of $g=\sqrt{4\pi N}$, for three
705: values of the number of particles $N$. The bulk transition happens
706: at $u^*_{bulk}=2$. As can be seen, there is a nonzero order
707: parameter for any nonzero $u$, and it grows smoothly and
708: continuously as $u$ increases. Nothing discontinuous happens at
709: $u=2$ or even beyond, indicating that the instability does not
710: suddenly become first-order at the bulk value of $u^*$. Of course,
711: in these finite systems, the Thouless and bulk scales are related
712: by a factor $g/4\pi$, which is not that large (4.4 for the largest
713: system we considered, with $N=245$). So somewhere between $u=2.25$
714: and $u=2.5$ the instability seems to reach the bulk scale.
715: However, note that the size of systems we have considered
716: correspond quite closely to actual ballistic
717: samples\cite{recent-expts}, which typically have a few hundred
718: electrons. Further, the three curves seem to track each other
719: fairly closely, indicating that the expectation value of
720: $|\sigmab|$ indeed scales with $g$, as predicted by our earlier
721: work based on RMT assumptions.
722: 
723: \begin{figure}[ht]
724: %\narrowtext
725: %\epsfxsize=2.4in\epsfysize=2.4in
726: \includegraphics*[width=2.4in,angle=0]{fig5.eps}
727: %\vskip 0.15in
728: \caption{The absolute value of the order parameter normalized by $g$
729:  as a function of coupling strength $u$ for three values of the number
730:  of particles $N$.  The fact that the curves track each other closely
731:  indicates that the order parameter does indeed scale like
732:  $g$. Furthermore, nothing discontinuous happens at the bulk critical
733:  coupling strength $u^*_{bulk}=2$.}
734: \label{fig3}
735: \end{figure}
736: 
737: In Fig. \ref{fig4} we show the corresponding reduction in ground state
738: energy normalized by $g^2$. Once again, the curves track each other
739: fairly closely, indicating that the energy reduction due to
740: interactions is indeed of order $g^2$, as predicted by our earlier
741: work.
742: 
743: \begin{figure}[ht]
744: %\narrowtext
745: %\epsfxsize=2.4in\epsfysize=2.4in
746: \includegraphics*[width=2.4in,angle=0]{fig6.eps}
747: %\vskip 0.15in
748: \caption{Reduction in ground state energy normalized by $g^2$ for three values of the number of particles $N$.}
749: \label{fig4}
750: \end{figure}
751: 
752: In Fig. \ref{fig5} we show the effective potential landscape for
753: $m=2$, with $N=245$, at a value of $u=2.15$, at which the minimum is
754: well-developed, but the order parameter is still within the nominal
755: Thouless scale and has not reached the bulk scale. The RMT analysis
756: predicted a Mexican Hat landscape with ``small'' ripples (down by
757: $1/g$) in the circle of minima of the Mexican Hat. The landscape we
758: see bears no resemblance to this. Instead, it appears to be an
759: isolated minimum at a nonzero $\sigmab$. Upon close inspection it can
760: be seen that the minimum is shallower in the transverse direction than
761: in the radial direction, but this is the only indication we could find
762: of a (perhaps) incipient Mexican Hat structure.
763: 
764: \begin{figure}[ht]
765: %\narrowtext
766: %\epsfxsize=2.4in\epsfysize=2.4in
767: \includegraphics*[width=2.4in,angle=0]{fig7.eps}
768: %\vskip 0.15in
769: \caption{Effective potential landscape for symmetry-breaking in the
770: $m=2$ channel. Instead of a Mexican Hat minimum structure with small
771: ripples we see an isolated minimum. The minimum does seem shallower in
772: the transverse direction. }
773: \label{fig5}
774: \end{figure}
775: 
776: 
777: Fig. \ref{fig6} shows a similar effective potential landscape for
778: symmetry breaking in the $m=1$, channel, where the two exactly
779: degenerate minima expected from time-reversal invariance
780: considerations can be seen. The landscape also appears more
781: Mexican-Hat-like than in the $m=2$ case.
782: 
783: \begin{figure}[ht]
784: %\narrowtext
785: %\BoundingBox{180,180}
786: \includegraphics*[width=2.4in,angle=0]{fig8.eps}
787: %\vskip 0.15in
788: \caption{Effective potential landscape for symmetry-breaking in the
789: $m=1$ channel for $N=200$ and $u=2.15$. The two exactly degenerate
790: minima required by time-reversal invariance can be seen, as can a
791: Mexican-Hat-like structure.}
792: \label{fig6}
793: \end{figure}
794: 
795: 
796: To trace the origin of this difference in behavior, we investigated
797: the average absolute value $\langle |M^i_{\a\b}|\rangle$ and the rms
798: deviation of the matrix elements from the mean absolute value,
799: $\sqrt{\langle |M^i_{\a\b}|^2\rangle-\langle |M^i_{\a\b}|\rangle^2}$
800: for the two cases $m=1,2$. The results for the $i=1$ (corresponding to
801: $\nabla_x$ for $m=1$ and $\nabla_x^2-\nabla_y^2$ for $m=2$) shown in
802: Fig. \ref{fig7} are an energy average for a particular billiard, with
803: the parameters $b=c=0.20, \d=0.85$. (We have confirmed similar
804: behavior of the matrix elements for other parameter values as well.)
805: Fig.
806: \ref{fig7}  shows these quantities  as a function of the energy
807: difference between the two states $\a$ and $\b$. There are two
808: features that are particularly noteworthy.
809: 
810: \begin{itemize}
811: \item There is a ``hole'' in the $m=1$ matrix element near zero
812: energy difference.
813: 
814: \item The rms deviation of the $m=2$ matrix elements from their
815: mean absolute value is huge. As a rough estimate, if the matrix
816: elements were Gaussian distributed complex numbers, the rms
817: deviation should be roughly half the mean modulus.
818: \end{itemize}
819: 
820: \begin{figure}[ht]
821: %\narrowtext
822: %\BoundingBox{180,180}
823: \includegraphics*[width=2.4in,angle=0]{fig9.eps}
824: %\vskip 0.15in
825: \caption{A plot of the absolute value and the rms deviation  of the matrix elements
826: $M^x_{\a\b}$ from their mean absolute value as a function of
827: $\ve_{\a}-\ve_{\b}$ for the two cases $m=1,2$.}
828: \label{fig7}
829: \end{figure}
830: 
831: However, the $i=2$ component (corresponding to $\nabla_y$ for $m=1$
832: and $2\nabla_x\nabla_y$ for $m=2$) shows very different behavior in
833: Fig. \ref{fig8}. While the $m=1$ case looks similar to the $i=1$
834: component, the fluctuations of the $m=2$ $i=2$ component are strongly
835: suppressed by almost an order of magnitude below the mean.
836: 
837: \begin{figure}[ht]
838: %\narrowtext
839: %\BoundingBox{180,180}
840: \includegraphics*[width=2.4in,angle=0]{fig10.eps}
841: %\vskip 0.15in
842: \caption{A plot of the absolute value and the rms deviation  of the matrix elements
843: $M^y_{\a\b}$ from their mean absolute value as a function of
844: $\ve_{\a}-\ve_{\b}$ for the two cases $m=1,2$.}
845: \label{fig8}
846: \end{figure}
847: 
848: Consider first the "hole" at $E_F$ for $m=1$.  By symmetry
849: considerations alone one can understand that the diagonal matrix
850: element $M_{\a\a}$ for $m=1$ has to be zero sample by sample in the
851: absence of an external magnetic field. Focusing on the $x$ component
852: of the order parameter
853: %
854: \beq
855: M^{x}_{\a\a}=\sum\limits_{\bk} \cos(\theta_{\bk}) \phi^*_{\a}(\bk)\phi_{\a}(\bk)
856: \eeq
857: %
858: By time-reversal invariance $\phi^*_{\a}(\bk)=\phi_{\a}(-\bk)$. Noting
859: that $\theta_{-\bk}=\theta_{\bk}+\pi$, and that the $\cos$ term
860: changes sign, one concludes that $M_{\a\a}=-M_{\a\a}=0$. The reason
861: the "hole'' persists for finite energy differences for the operator
862: ${\vec p}=-i{\vec\nabla}$ can be explained by noting
863: that\cite{barnett-pvt} for a billiard
864: %
865: \beqr
866: {\vec p}=&im[\vrr,H]\nonumber\\
867: \Rightarrow (-i{\vec\nabla})_{\a\b}=&-im(\vrr)_{\a\b} (\ve_\a-\ve_\b)
868: \eeqr
869: %
870: which means that the matrix element must vanish at least linearly
871: with the energy difference. In fact, such ``banded'' matrix
872: elements have been found for many operators in ballistic
873: dots\cite{barnett}.
874: 
875: Consider next the fact that  the distribution of the matrix
876: elements of $M^x_{\a\b}$ for the $m=2$ case is much broader than
877: for the $m=1$ case, while the $M^y_{\a\b}$ matrix elements have a
878: very narrow distribution. The RMT answer would have the rms
879: deviation of $M_{\a\b}$ from the mean to be of the same order as
880: the mean absolute value. This seems to be roughly true for both
881: components of $m=1$ but grossly untrue for the $i=1$ component of
882: $m=2$. Since it is these mesoscopic fluctuations in $M_{\a\a}$
883: which determine the size of the ripples at the bottom of the
884: Mexican Hat in the RMT scenario, this broad distribution of
885: $M_{\a\b}$ seems to be the cause of the failure of the RMT
886: prediction that the ripples should be subdominant by $1/g$. While
887: it is tempting to try to explain this in relation to the shape of
888: the billiard (Fig. \ref{fig1}), which certainly appears to favor
889: an $x^2-y^2$ type of symmetry, a satisfactory explanation  of the
890: broad distribution of the $m=2$, $i=1$ matrix elements eludes us.
891: 
892: 
893: Our knowledge of the eigenfunctions at the global minimum allow us
894: to compute the  effective action for time-dependent $\sigmab$ at
895: that minimum. Since the quasiparticles couple to this collective
896: field, the interaction induces a decay width for the
897: quasiparticles (details can be found in ref.\cite{longpaper}). In
898: Fig. \ref{fig9} we compare the numerically calculated values of
899: the width to the parameter-free theoretical prediction (solid
900: line) based on RMT\cite{longpaper}. On average, the RMT based
901: prediction seems consistent with the numerics, though there is a
902: lot of variation in the widths driven by large variations in the
903: matrix elements coupling the quasiparticle levels to the
904: collective mode.
905: 
906: 
907: \begin{figure}[ht]
908: %\narrowtext
909: %\epsfxsize=2.4in\epsfysize=2.4in
910: \includegraphics*[width=2.4in,angle=0]{fig11.eps}
911: %\vskip 0.15in
912: \caption{A plot of the decay width of quasiparticles induced by their coupling to
913: fluctuations of the collective field $\sigmab$, for $N=245$, $m=2$,
914: and $u=2.15$. The solid line is the theoretical prediction from our
915: previous RMT-based analysis\cite{longpaper}. While the prediction does
916: well on average there are huge variations in the widths due to large
917: variations in how strongly each level couples to the collective mode.}
918: \label{fig9}
919: \end{figure}
920: 
921: 
922: \section{Conclusions}
923: \medskip
924: 
925: In our earlier work, we used a global RG assumption to reduce the
926: problem on the scale of the Thouless energy to that of a disordered
927: noninteracting problem with Fermi-liquid interactions. This is quite
928: plausible for ballistic dots on very general grounds. To proceed
929: further we had to make two further assumptions; (i) That the $g$
930: approximate momentum states at the Fermi energy were a good basis in
931: which to expand the exact disorder eigenstates, and (ii) That the wave
932: functions of the exact eigenstates in the momentum basis obeyed all
933: the statistical properties of RMT. Based on these two assumptions we
934: were able to construct a solution to the problem which was
935: asymptotically exact in the limit $g\to\infty$. This solution led to
936: specific predictions for various physical quantities, including the
937: size of the order parameter, the reduction in energy due to
938: interactions, the shape of the energy landscape, and the size of the
939: quasiparticle decay widths.
940: 
941: In this paper we have carried out a calculation which is still
942: predicated on the validity of the Fermi-liquid form of the
943: interactions on a scale $E_L$ much larger than the Thouless energy
944: $E_T$. In retaining this assumption we are on firm ground, since after
945: all, the Thouless energy can be made as small as one wishes merely by
946: increasing the size of the system. We also assumed that the mean-field
947: description of the Landau Fermi-liquid interactions is valid, which is
948: justified by the fact that the minima in the effective potential
949: landscape are indeed of order $g^2\d$. However, we explicitly eschewed
950: the other two assumptions that we made in previous work, with a view
951: to independently testing their validity. We found that our first
952: assumption, that the approximate momentum states were a good basis in
953: which to expand the exact disorder eigenstates, was extremely good
954: near the Fermi energy, but became increasingly inaccurate as one went
955: to the edge of the Thouless shell. We did not test the second
956: assumption about wavefunction correlations explicitly, but indirectly
957: through its effects on the predictions of our earlier work. We found
958: that most of the predictions held up, with the exception of the shape
959: of the effective potential landscape in the case of symmetry-breaking
960: in the $m=2$ channel. Even here, the minimum is shaped more like a
961: crescent, indicating the possible emergence of the Mexican Hat
962: structure at larger values of $g$ (we went to the largest value of $g$
963: that we could given that we kept only 585 states and had to keep at
964: least half the states empty). We traced the discrepancy back to the
965: anomalously broad distribution (compared to estimates based on a
966: complex gaussian distribution) of the matrix elements
967: $M^x_{\a\b}$. However, we were unable to pin down a physical reason
968: for this broad distribution for the $m=2$ case.
969: 
970: In conclusion, much of the physics we uncovered using our RMT
971: assumptions seems to be valid in the Robnik-Berry billiard. The
972: second-order transition that we uncovered in the $g\to\infty$ limit
973: seems to indeed be broadened into a smooth croossover as expected,
974: belying fears that it may be overtaken by a first-order bulk
975: transition. The question of how large $g$ has to be before RMT becomes
976: fully applicable remains open; another way to phrase the question is
977: to ask what the nonuniversal corrections to RMT are in ballistic
978: systems. Finally, an important open question is whether the broad
979: distributions of matrix elements of interaction operators is a generic
980: feature of ballistic systems, rather than being a special feature of
981: the Robnik-Berry billiard, and if so, what physics determines the
982: width of those distributions. However, our results here give us
983: encouragement that the RMT assumptions can indeed be used with
984: confidence in making predictions in ballistic systems, at a
985: qualitative and semi-quantitative level.
986: %
987: \section{Acknowledgements}
988: %
989: We would like to thank Yoram Alhassid, Alex Barnett, Piet Brouwer,
990: and Doug Stone for illuminating conversations, and the Aspen
991: Center for Physics where part of this work was carried out. We are
992: also grateful to the NSF for partial support under grants DMR
993: 0311761 (GM), DMR 0354517(RS).
994: 
995: \begin{thebibliography}{99}
996: \bibitem{recent-expts} U. Sivan {\it et al}, \prl\  {\bf 77}, 1123 (1996);
997: S. R. Patel {\it et al}, \prl\ {\bf 80}, 4522 (1998); F. Simmel {\it
998: et al}, \prb\ {\bf 59}, 10441 (1999); D. Abusch-Magder {\it et al},
999: Physica E\ {\bf 6}, 382 (2000); F. Simmel, T. Heinzel, and
1000: D. A. Wharam, Eur. Lett. \ {\bf 38}, 123 (1997); J. A. Folk {\it et
1001: al},
1002: \prl\ {\bf 86}, 2102 (2001); S. L\"uscher {\it et al}, \prl\ {\bf 86}, 2118 (2001).
1003: \bibitem{reviews} For recent reviews, see, T. Guhr, A. M\"uller-Groeling,
1004: and H. A. Weidenm\"uller, Phys. Rep. {\bf 299}, 189 (1998);
1005: Y. Alhassid, \rmp\ {\bf 72}, 895 (2000); A. D. Mirlin, Phys. Rep. {\bf
1006: 326}, 259 (2000).
1007: \bibitem{H_U} A.  V.  Andreev and A.
1008: Kamenev, \prl {\bf 81}, 3199 (1998); P.  W.  Brouwer, Y.  Oreg, and B.
1009: I.  Halperin, \prb {\bf 60}, R13977 (1999); H.  U.  Baranger, D.
1010: Ullmo, and L.  I.  Glazman, \prb {\bf 61}, R2425 (2000); I.  L.
1011: Kurland, I.  L.  Aleiner, and B.  L.  Al'tshuler, \prb\ {\bf 62},
1012: 14886 (2000).
1013: \bibitem{univ-ham}I. L. Aleiner, P. W. Brouwer, and L. I. Glazman,
1014: Phys. Rep. {\bf 358}, 309 (2002), and references therein; Y. Oreg, P. W. Brouwer,
1015: X. Waintal, and B. I. Halperin, cond-mat/0109541, and references
1016: therein.
1017: \bibitem{mm} G. Murthy and H. Mathur, \prl\ {\bf 89}, 126804 (2002).
1018: \bibitem{qd-ms1} G. Murthy and R. Shankar, \prl {\bf 90}, 066801 (2003).
1019: \bibitem{longpaper} G. Murthy, R. Shankar, D. Herman, and H. Mathur,
1020: \prb {\bf 69}, 075321 (2004).
1021: \bibitem{rg-shankar} R. Shankar, {\it Physica}\ {\bf A177},
1022: 530 (1991); R.Shankar, { Rev. Mod. Phys.} {\bf 66}, 129 (1994).
1023: \bibitem{alt1} K. B. Efetov, Adv. Phys. {\bf 32}, 53 (1983);
1024: B. L. Al'tshuler ad B. I. Shklovskii, { Sov. Phys. JETP}\
1025: {\bf 64}, 127 (1986).
1026: \bibitem{RMT} M. L. Mehta, {\it Random Matrices}, Academic Press, San
1027: Diego, 1991.
1028: \bibitem{critical-fan} S. Chakravarty, B. I. Halperin, and
1029: D. R. Nelson, \prl\ {\bf 60}, 1057 (1988); \prb\ {\bf 39}, 2344
1030: (1989); For a detailed treatment of the generality of the phenomenon,
1031: see, S. Sachdev, {\it Quantum Phase Transitions}, Cambridge University
1032: Press, Cambridge 1999.
1033: \bibitem{robnik-berry} M. Robnik, J. Phys. A {\bf 17}, 1049 (1984);
1034: M. V. Berry and M. Robnik, J. Phys. A {\bf 19}, 649 (1986).
1035: \bibitem{agd} A. A.  Abrikosov, L. P. Gorkov, and I. E. Dzyaloshinski,
1036: {\it Methods of Quantum Field Theory in Statistical Physics}, Dover
1037: Publications, New York, 1963.
1038: \bibitem{berry-ansatz} M. V. Berry, J. Phys. A {\bf 10}, 2083 (1977).
1039: \bibitem{stone-bruus} A. D. Stone and H. Bruus, Physica B {\bf 189}, 43 (1993);
1040: Surface Science {\bf 305}, 490 (1994).
1041: \bibitem{alhassid-lewenkopf} Y. Alhassid and C. H. Lewenkopf, \prb {\bf 55}, 7749 (1997).
1042: \bibitem{supersymm} K. B. Efetov and V. R. Kogan, \prb\ {\bf 67}, 245312 (2003), 
1043: and references therein. 
1044: \bibitem{nowindow} S. Adam, P. W. Brouwer, and P. Sharma,\prb {\bf 68}, 241311 (2003)
1045: \bibitem{barnett-pvt} A. H. Barnett, private communication (2004).
1046: \bibitem{barnett} A. H. Barnett, ``Asymptotic rate of quantum ergodicity in
1047: chaotic Euclidean billiards'', Courant Institute preprint, 2004.
1048: \end{thebibliography}
1049: \end{document}
1050: