1: % \documentclass[prb,12pt,showpacs]{revtex4}
2: \documentclass[prb,showpacs]{revtex4}
3: %\documentclass[prb,twocolumn,showpacs]{revtex4}
4: %\documentclass[12pt,fullpage,doublespace]{article}
5:
6: \usepackage{graphicx}
7: \usepackage{epsfig}
8: \usepackage{bm}
9: \usepackage{latexsym}
10: \newcommand{\ket}[1]{| {#1} \rangle}
11: \newcommand{\vev}[1]{\mbox{$\left\langle #1 \right\rangle$}}
12: \newcommand{\etal}{\textit{et al.} }
13: \def\goes{\rightarrow}
14:
15:
16: \makeatletter
17: \def\varddots{\mathinner{\mkern1mu
18: \raise\p@\hbox{.}\mkern2mu\raise4\p@\hbox{.}\mkern2mu
19: \raise7\p@\vbox{\kern7\p@\hbox{.}}\mkern1mu}}
20: \makeatother
21:
22: \makeatletter
23: \renewcommand{\theequation}
24: {\thesection.\arabic{equation}}
25: \@addtoreset{equation}{section}
26: \makeatother
27:
28: \begin{document}
29:
30: \title{Critical Spectra and Wavefunctions of a One-dimensional
31: Quasiperiodic System }
32:
33: \author{ Kazusumi Ino$^1$ and Mahito Kohmoto$^2$ }
34:
35: \affiliation{
36: $^1$Department of Basic Science, University of Tokyo,
37: Komaba 3-8-1, Meguro-ku, Tokyo, 153-8902, Japan\\
38: $^2$Institute of Solid State Physics, University of Tokyo,
39: Kashiwanoha 5-1-5, Kashiwa-shi, Chiba, 277-8581, Japan}
40:
41: %\today
42: \begin{abstract}
43:
44: We numerically study a one dimensional quasiperiodic system
45: obtained from two dimensional electrons on the
46: triangular lattice in a uniform magnetic field aided by the multifractal method.
47: The phase diagram
48: consists of three phases: two metallic phases and
49: one insulating phase separated by critical lines
50: with one bicritical point. Novel transitions between the two metallic phases exist.
51: We examine the spectra and the wavefunctions along the critical lines.
52: Several types of level statistics are obtained.
53: Distributions of the band widths $P_B(w)$ near the origin (in the tail)
54: %around the origin (in the tail)
55: have a form
56: $P_B(w) \sim w^{ \beta}$ ($P_B(w) \sim e^{ -\gamma w }$)
57: ($\beta , \gamma > 0 $),
58: while at the bicritical point $P_B(w) \sim w^{-\beta'}$ ($\beta'>0$).
59: Also distributions of the level spacings follow
60: an inverse power law $P_G(s) \sim s^{- \delta}$ ($\delta > 0$ ).
61: For the wavefunctions at the centers of spectra, scaling exponents and their
62: distribution in terms of the $\alpha$-$f(\alpha)$-curve are obtained.
63: The results
64: in the vicinity of critical points are consistent with
65: the phase diagram.
66: \end{abstract}
67:
68:
69:
70:
71: \pacs{71.30.+h, 71.23.Ft, 05.45.Mt}
72:
73: \maketitle
74:
75:
76: \section{Introduction}
77:
78: A peculiar problem of
79: two-dimensional electrons in a periodic potential with
80: a perpendicular magnetic field
81: has been attracted much attentions
82: since the Hofstadter butterfly \cite{hof}.
83: It appears as the spectrum of the underlying one-dimensional
84: system called the Harper model \cite{harper} which is deduced
85: from electrons on the square lattice in a uniform magnetic field . It is also essential in physics of
86: the integer quantum Hall effect \cite{tknn,kohmoto-chern}.
87:
88:
89: Some of the metal-insulator transitions
90: in one-dimensional quasiperiodic systems have been characterized
91: by multifractal structures of band widths and wavefunctions.
92: See \cite{hiramoto}.
93: A one-dimensional tight-binding model is
94: \begin{equation}
95: t_{i+1}\psi_{i+1}+t_{i-1}\psi_{i-1}+\epsilon_i \psi_i =E\psi_i,
96: \label{2}
97: \end{equation}
98: where $\psi_i $ denotes the value of
99: the wavefunction at the $i$-th site,
100: $t_i$ and $\epsilon_i$ are the hopping matrix element
101: and the site energy at the $i$-th site respectively, either or both of them can
102: be taken to be quasiperiodic.
103: The Harper model is the case where
104: $t_{i}=1$ and $\epsilon_i=\lambda \cos(2\pi\sigma i+\theta)$.
105: When $\sigma$ is an irrational number, it is quasiperiodic.
106: All the eigenstates
107: are extended for $\lambda <2$
108: and are localized for $\lambda >2$ with the metal-insulator
109: transitions at $\lambda=2$ \cite{aubry,kohmoto-prl}. The spectrum has a rich structure
110: (the Hofstadter butterfly).
111: The spectrum as well as the eigenstates becomes multifractal
112: %
113: %$\lambda=2$
114: %
115: \cite{hiramoto-kohmoto}.
116: The total measure of the bands at the critical point $\lambda =2$
117: is zero with a fractal dimension less than one \cite{thouless}.
118: The scaling behaviors of the spectra have been extensively studied
119: \cite{hiramoto-kohmoto}
120: by the multifractal analysis \cite{halsey,kohmoto}.
121: Especially the incommensurate limits of flux per plaquette,
122: such as the inverse of the golden mean $\sigma=\frac{-1+\sqrt{5}}{2}$
123: have been extensively studied.
124:
125:
126: Recently level statistics of some of the quasiperiodic systems
127: have been investigated and turned out to have the
128: behaviors characteristic of criticality \cite{evangelou,takada}.
129: The distributions of the normalized band widths
130: $P_B(w)dw$ have been confirmed that
131: \begin{eqnarray}
132: P_B(w)\sim w^{\beta}
133: \quad (w\to 0),
134: \end{eqnarray}
135: and
136: \begin{eqnarray}
137: P_B(w)\sim e^{-\gamma w}
138: \quad (w \to \infty),
139: \end{eqnarray}
140: %where $\beta \sim 2.5,\gamma \sim 1.4$ \cite{takada}.
141: These laws have also been confirmed for
142: a variant of the Harper model at criticality
143: \cite{takada}. A similar type of statistical law has
144: been confirmed for the Fibonacci model \cite{naka}.
145: Remarkably, the form of the distributions of band widths has
146: a similar form as the distributions of the gaps
147: fluctuations observed at the mobility edge of the random systems \cite{altshuler,shapiro}.
148: Distribution of the energy gaps $P_G(s)$
149: was also examined.
150: It diverges near the origin and
151: follows an inverse power law \cite{machigei,takada,naka}
152: \begin{eqnarray}
153: P_G(s)\sim s^{-\delta} \quad (s\to 0),
154: \end{eqnarray}
155: For example,
156: $\delta \sim 1.5$ for the critical Harper model($\lambda=2$).
157:
158:
159:
160: One of the aims of this paper is to investigate these quantities
161: for the one dimensional quasiperiodic model obtained
162: from two dimensional electrons on the triangular lattice
163: in a uniform magnetic field. This problem was
164: studied by Claro and Wannier \cite{claro} ,
165: but systematic studies have not been achieved since then.
166: % kohmoto deleted from here
167: Although the model of two-dimensional
168: electrons on the square lattice with next-nearest hopping,
169: which includes the case of triangular lattice as a special case,
170: were studied previously \cite{han-thouless,hat-koh},
171: statistical techniques such as the multifractal analysis
172: which have been applied for other quasiperiodic systems
173: have not been applied to the triangular lattice model.
174: % Recently this system has been
175: % realized in an experiment \cite{melinte} and the effect of
176: % singular spectrum is observed.
177: % The effect of disorder in such a system was investigated numerically
178: % \cite{zhou} and it turned out that the singular spectrum is not
179: % completely smeared out by the effect of disorder.
180: %kohmoto deleted to here
181: The same model also
182: appears in the theory of the junction of three wires of Luttinger liquid
183: \cite{chamon}. These situations motivate us to investigate
184: various aspects of the quasiperiodic system obtained
185: from the triangular lattice model.
186:
187:
188: The organization of this paper is as follows. In Sec.{\ref{section;model}},
189: we introduce two dimensional electrons on the triangular lattice
190: in a uniform magnetic field and obtain the one-dimensional
191: quasiperiodic system. We
192: investigate the Aubry and Andr\'{e} duality \cite{aubry} in this model.
193: In Sec.{\ref{section;chara}}, we investigate the classical orbits
194: of the model and discuss the phase diagram.
195: In Sec.{\ref{section;level}}, the distributions of the band widths and
196: the gaps are investigated.
197: In Sec.{\ref{section;multi}}, we give a brief review of
198: the general formulation of the multifractal analysis and
199: apply it to the spectra and the wavefunctions. We confirm
200: the phase diagram conjectured in Sec.\ref{section;model}.
201: Sec.{\ref{section;conclusions}} is the conclusion.
202:
203:
204: \section{\label{section;model}Electrons on the triangular lattice
205: in a uniform magnetic field}
206: \subsection{Hamiltonian in real space}
207: We consider tight-binding electrons on the triangular lattice
208: in a magnetic field ({\bf Fig.\ref{fig;tri}}).
209: We take the lattice spacing to be $1$ for simplicity.
210: The Hamiltonian is
211: \begin{eqnarray}
212: H&=& -t_a\sum_{n,m}c^{\dagger}_{n+1,m}c_{n,m} \exp(iA_{n+1,m;n,m})
213: -t_a\sum_{n,m}c^{\dagger}_{n,m}c_{n+1,m} \exp(iA_{n,m;n+1,m})
214: \nonumber \\
215: & & -t_b\sum_{n,m} c^{\dagger}_{n,m+1}c_{n,m} \exp(iA_{n,m+1;n,m})
216: -t_b\sum_{n,m} c^{\dagger}_{n,m}c_{n,m+1} \exp(iA_{n,m;n,m+1})
217: \nonumber \\
218: & &
219: -t_c\sum_{n,m} c^{\dagger}_{n,m+1}c_{n+1,m} \exp(iA_{n,m+1;n+1,m})
220: -t_c\sum_{n,m} c^{\dagger}_{n+1,m}c_{n,m+1} \exp(iA_{n+1,m;n,m+1})
221: \nonumber \\
222: && \equiv H_a+H_b + H_c
223: \label{eq;model}
224: \end{eqnarray}
225: Here $t_a,t_b$ and $t_c$ are the hopping coefficients for each bond,
226: and $c_{n,m}$($c^{\dagger}_{n,m}$) is the
227: annihilation (creation) operator at site $(n,m)$ :
228: $\{c^{\dagger}_{n,m},c_{k,l} \}=\delta_{k,n}\delta_{lm}$.
229: $A_{n,m;k,l}$, $k=n\pm1, l=m\pm1$
230: is a gauge field on each bond. We impose $A_{n,m;k,l} =-
231: A_{k,l;n,m}$ so that $H$ to be hermitian.
232: A uniform magnetic field penetrates
233: each triangle with a flux $\varphi=\frac{\phi}{2}$.
234: We take the Landau gauge
235: \begin{eqnarray}
236: A_{n+1,m;n,m} =0, A_{n,m+1;n,m} =2\pi \phi \hspace{3mm}
237: {\rm and} \hspace{3mm}
238: A_{n,m+1;n+1,m} =2\pi\phi (n+\frac{1}{2}).
239: \label{t_a-gauge}
240: \end{eqnarray}
241: thus $\sum_{{\rm triangle}} A_{n,m;k,l}=\frac{\phi}{2}$.
242: A state $\ket{\Psi}$ is written
243: \begin{eqnarray}
244: \ket{\Psi} = \sum_{n,m} \Psi_{n,m}c^{\dagger}_{n,m}\ket{0}.
245: \end{eqnarray}
246: The Schr\"odinger equation $H\ket{\Psi}=E\ket{\Psi}$ is
247: \begin{eqnarray}
248: -t_a(\Psi_{n-1,m}+\Psi_{n+1,m})-t_b(e^{2\pi i\phi n}\Psi_{n,m-1}
249: +e^{-2\pi i\phi n}\Psi_{n+1,m}) \nonumber \\
250: -t_c(e^{-2\pi i\phi (n-\frac{1}{2})}\Psi_{n-1,m+1}
251: +e^{2\pi i\phi (n+\frac{1}{2})}\Psi_{n+1,m}) =E\Psi_{n,m}.
252: \end{eqnarray}
253: We take the form of the wavefunction $\Psi_{n,m}=e^{ik_y m}\Psi_n$,
254: then the Schr\"odinger equation becomes
255: \begin{eqnarray}
256: -(t_a+t_ce^{-2\pi i \phi(n-\frac{1}{2})+ik_y})\Psi_{n-1}
257: -(t_a+t_ce^{2\pi i\phi(n+\frac{1}{2}-ik_y)})\Psi_{n+1}
258: -2t_b\cos(2\pi\phi n+k_y)\Psi_n =E\Psi_n.
259: \label{eq;triharper}
260: \end{eqnarray}
261:
262:
263:
264:
265: %Take the Landau gauge
266: %\begin{eqnarray}
267: %A_{n+1,m;n,m} =0, A_{n,m+1;n,m} =2\pi \phi \hspace{3mm}
268: %{\rm and}
269: %\hspace{3mm} A_{n,m+1;n+1,m} =2\pi\phi (n+\frac{1}{2}).
270: %\label{t_a-gauge}
271: %\end{eqnarray}
272: %thus $\sum_{{\rm triangle}} A_{n,m;k,l}=\frac{\phi}{2}$.
273: %\begin{eqnarray}
274: %\ket{\Psi} = \sum_{n,m} \Psi_{n,m}c^{\dagger}_{n,m}\ket{0}.
275: %\end{eqnarray}
276: %The Schr\"odinger equation $H\ket{\Psi}=E\ket{\Psi}$ is
277: % Take the form of the wavefunctions $\Psi_{n,m}=e^{ik_y m}\Psi_n$,
278: %then the Schr\"odinger equation is
279: %\begin{eqnarray}
280: %-(t_a+t_ce^{-2\pi i \phi(n-\frac{1}{2})+ik_y})\Psi_{n-1}
281: %-(t_a+t_ce^{2\pi i\phi(n+\frac{1}{2}-ik_y)})\Psi_{n+1}
282: %-2t_b\cos(2\pi\phi n+k_y)\Psi_n =E\Psi_n.
283: %\label{eq;triharper}
284: %\end{eqnarray}
285: When $\phi=\frac{p}{q}$ ($p$ and $q$ are coprime integers), (\ref{eq;triharper})
286: is periodic with period $q$. The Bloch theorem tells
287: that one can put $\Psi_n = \exp(ik_x n)\psi_n$ where $\psi_n$
288: satisfies $\psi_n=\psi_{n+q}$, which implies that
289: $\Psi_{n+q}=e^{ik_x q}\Psi_n$. Thus, if we introduce
290: a row vector $\Psi=(\Psi_1,\Psi_2,\cdots,\Psi_{q-1},\Psi_q )^t$ ($t$
291: means the transpose of a matrix) and
292: $a_n(k_y)=t_a+t_c\exp(2\pi i \frac{p}{q}(n+\frac{1}{2})-ik_y)$ and
293: $b_n(k_y)=2t_b\cos(2\pi\frac{p}{q}n+k_y)$,
294: (\ref{eq;triharper}) is reduced to
295: an eigenvalue problem of a finite size matrix
296: \begin{eqnarray}
297: H_q(k_x,k_y) = \left(
298: \begin{array}{@{\,}cccccccc@{\,}}
299: b_{1}(k_y) &
300: a_{1}(k_y) &
301: 0 & \cdots & & & 0 &
302: e^{ik_x q} a_{0}(k_y)^{*} \\
303: a_{1}(k_y)^{*} &
304: b_{2}(k_y)&
305: a_{2}(k_y) &
306: 0 & \cdots & & & 0 \\
307: 0 &
308: a_{2}(k_y)^{*} &
309: b_{3}(k_y) &
310: a_{3}(k_y) &
311: 0 & \cdots & & 0 \\
312: & 0 &
313: a_{3}(k_y)^{*} &
314: b_{4}(k_y) &
315: a_{4}(k_y) &
316: 0 & \ldots & 0 \\
317: \vdots & \vdots & \ddots &
318: \ddots & \ddots & \ddots &
319: \ddots & \vdots \\
320: 0 & & \cdots & 0 &
321: a_{q-3}(k_y)^{*} &
322: b_{q-2}(k_y) &
323: a_{q-2}(k_y) &
324: 0 \\
325: 0 & & &\cdots & 0 &
326: a_{q-2}(k_y)^{*} &
327: b_{q-1}(k_y) &
328: a_{q-1}(k_y)
329: \\
330: e^{-ik_xq} a_{0}(k_y) &
331: 0 & & &\cdots & 0 &
332: a_{q-1}(k_y)^{*}&
333: b_{q}(k_y) \\
334: \end{array}
335: \right).
336: \label{eq;matrix}
337: \end{eqnarray}
338: Also, in terms of $\psi_n$,
339: (\ref{eq;triharper}) becomes
340: \begin{eqnarray}
341: -e^{-ik_x}(t_a+t_ce^{-2\pi i\phi(n-\frac{1}{2})+ik_y})\psi_{n-1}
342: -e^{ik_x}(t_a+t_ce^{2\pi i \phi(n+\frac{1}{2})-ik_y})\psi_{n+1}
343: -2t_b\cos(2\pi\phi n+k_y)\psi_n =E\psi_n.
344: \label{eq;triharpera}
345: \end{eqnarray}
346: At $t_c=0$, this is reduced to the Harper equation.
347: We define $\lambda \equiv 2\frac{t_b}{t_a}$
348: and $\mu \equiv \frac{t_c}{t_a}$ then (\ref{eq;triharper}) becomes
349: \begin{eqnarray}
350: -\left[1+\mu e^{-2\pi i \frac{p}{q}(n-\frac{1}{2})+ik_y}\right]\Psi_{n-1}
351: -\left[1+\mu e^{2\pi i \frac{p}{q}(n+\frac{1}{2}-ik_y)}\right]\Psi_{n+1}
352: -\lambda \cos\left(2\pi\frac{p}{q} n+k_y\right)\Psi_n =E\Psi_n.
353: \label{eq;triharperlambdamu}
354: \end{eqnarray}
355:
356:
357:
358:
359:
360:
361: %///////// fig;triangular lattice //////////////
362: \begin{figure}
363: \begin{center}
364: \epsfxsize=8cm
365: \epsfbox{triangular-lattice.eps}
366: \caption{Schematic view of the triangular lattice.}.
367: \label{fig;tri}
368: \end{center}
369: \end{figure}
370:
371:
372: \subsection{Hamiltonian in momentum space}
373: We denote $\boldsymbol{k}=(k_x,k_y)$.
374: The electron annihilation operator $c(\boldsymbol{k})$
375: in momentum space is
376: \begin{eqnarray}
377: c_{n,m} =\frac{1}{(2\pi)^2}
378: \int^{\pi}_{-\pi} dk_x \int^{\pi}_{-\pi} dk_y \exp[ik_x n+ik_y m]
379: c(\boldsymbol{k}).
380: \end{eqnarray}
381: The commutation relation for $c(\boldsymbol{k}), c^{\dagger}(\boldsymbol{k})$ is
382: \begin{eqnarray}
383: \{c(\boldsymbol{k}),c^{\dagger}(\boldsymbol{k}')\} = (2\pi)^2\delta_{Z}(k_x-k_x')\delta_{Z}(k_y-k_y'),
384: \end{eqnarray}
385: where $\delta_Z(k)=\sum_{n \in \boldsymbol{Z}} \delta(k+2\pi n)$.
386: In terms of $c(\boldsymbol{k}), c^{\dagger}(\boldsymbol{k})$,
387: the tight-binding Hamiltonian (\ref{eq;model}) is
388: \begin{eqnarray}
389: H= \frac{1}{(2\pi)^2} \int^{\pi}_{-\pi} dk_x
390: \int^{\pi}_{-\pi} dk_y H(\boldsymbol{k}),
391: \end{eqnarray}
392: with
393: \begin{eqnarray}
394: H(\boldsymbol{k})=&&
395: -2t_a \cos k_x c^{\dagger}(\boldsymbol{k})c(\boldsymbol{k}) \nonumber \\
396: &&-(t_b e^{-ik_y}+t_c e^{-ik_x+ik_y-i\pi \phi} )
397: c^{\dagger}(k_x+2\pi\phi,k_y)c(k_x,k_y) \nonumber \\
398: &&-
399: (t_be^{ik_y}+t_ce^{ik_x-ik_y-i\pi \phi})c^{\dagger}(k_x-2\pi\phi,k_y)c(k_x,k_y).
400: \label{eq;hamiltoniank}
401: \end{eqnarray}
402: When $\phi=p/q$, since $k_x$ couples only to $k_x \pm 2\pi \phi$, we write
403: $k_x$ as $k_x^{0}+2\pi\phi j$ with $j \in \boldsymbol{Z}$.
404: Here $k^{0}_x$ is in the
405: magnetic Brillouin zone
406: \begin{eqnarray}
407: -\frac{\pi}{q} \leq k^{0}_x \leq \frac{\pi}{q}.
408: \end{eqnarray}
409: The Hamiltonian $H(\boldsymbol{k})$ acts on the Hilbert space
410: spanned by
411: \begin{eqnarray}
412: \ket{\Psi} = \sum_{=j}^{q} \widetilde{\psi}_j c^{\dagger}(k_x^{0}+2\pi\phi j,k_y)\ket{0},
413: \end{eqnarray}
414: with $\widetilde{\psi}_{j+q}=\widetilde{\psi}_{j}$.
415: The Schr\"odinger equation
416: $H\ket{\Psi}=E\ket{\Psi}$ is
417: \begin{eqnarray}
418: -(t_be^{-ik_y}+t_ce^{-ik^{0}_x+ik_y-2\pi i \phi(j -\frac{1}{2})})\widetilde{\psi}_{j-1}
419: -(t_be^{ik_y}+t_ce^{ik^{0}_x-ik_y+2\pi i\phi(j+\frac{1}{2})})\widetilde{\psi}_{j+1}
420: -2t_a \cos (2\pi\phi j+k^{0}_x) \widetilde{\psi}_j =E\widetilde{\psi}_j.
421: \nonumber \\
422: \label{eq;triharperk}
423: \end{eqnarray}
424: In (\ref{eq;triharper}) and (\ref{eq;triharperk}),
425: the terms $H_a$ and $H_b$ are diagonal respectively.
426: We can also diagonalize $H_c$ which is proportional to $t_c$
427: by changing the gauge. For example, we take the gauge
428: \begin{eqnarray}
429: A_{n+1,m;n,m} =2\pi\phi (n+\frac{1}{2}),
430: \hspace{3mm} A_{n,m+1;n,m} =2\pi \phi n, \hspace{3mm} {\rm and}
431: \hspace{3mm} A_{n,m+1;n+1,m} =0.
432: \label{t_c-gauge}
433: \end{eqnarray}
434: The gauge transformation which transforms from (\ref{t_a-gauge}) to
435: (\ref{t_c-gauge}) is given by
436: \begin{eqnarray}
437: c_{n,m} &\rightarrow& c_{n,m} \exp(if_{n}) \nonumber \\
438: c^{\dagger}_{n,m} &\rightarrow& c^{\dagger}_{n,m} \exp(-if_{n}) \nonumber \\
439: A_{n,m;n',m'} &\rightarrow& A_{n,m;n',m'} + f_{n,m} - f_{n',m'}, \nonumber \\
440: f_n &=&\phi n(n-1).
441: \end{eqnarray}
442: In this gauge, the Hamiltonian in momentum space is
443: \begin{eqnarray}
444: H= \frac{1}{(2\pi)^2} \int^{\pi}_{-\pi} dk_x
445: \int^{\pi}_{-\pi} dk_y H'(\boldsymbol{k}),
446: \end{eqnarray}
447: with
448: \begin{eqnarray}
449: H'(\boldsymbol{k})=&&
450: -2t_c \cos (k_x-k_y) c^{\dagger}(\boldsymbol{k})c(\boldsymbol{k})
451: \nonumber \\
452: &&-(t_b e^{-ik_y}+t_a e^{ik_x+\pi i \phi} )
453: c^{\dagger}(k_x+2\pi\phi,k_y)c(k_x,k_y) \nonumber \\
454: &&-
455: (t_be^{ik_y}+t_ae^{-ik_x+\pi i \phi})c^{\dagger}(k_x-2\pi\phi,k_y)c(k_x,k_y)
456: \label{eq;hamiltonianck}
457: \end{eqnarray}
458: From (\ref{eq;hamiltonianck}), we get the Schr\"odinger equation :
459: \begin{eqnarray}
460: -(t_be^{ik_y}+t_ae^{-ik^{0}_x-2\pi i \phi (\ell -\frac{1}{2})})\widehat{\psi}_{\ell-1}
461: -(t_be^{-ik_y}+t_ae^{ik^{0}_x+2\pi i \phi(\ell+\frac{1}{2})})\widehat{\psi}_{\ell+1}
462: -2t_c \cos (2\pi\phi \ell +k^{0}_x-k_y) \widehat{\psi_\ell} =E\widehat{\psi}_\ell.
463: \label{eq;triharperc}
464: \end{eqnarray}
465: Apparently, if one exchange $k_x^{0}$ by
466: $k_x^{0}-k_y$ and $t_a$ by $t_c$
467: in (\ref{eq;triharperk}), we get (\ref{eq;triharperc}).
468: This is due to the symmetry of the triangular lattice.
469:
470:
471:
472: \subsection{Duality}
473: At $t_c = 0$, it is known that (\ref{eq;triharpera})
474: has the duality of Aubry and Andr\'{e} \cite{aubry}
475: who showed the existence of a transition
476: between localized and extended states of $\psi_j$ when
477: $\phi$ is an irrational number. When $\lambda > 2$,
478: the states are all localized, and when $\lambda < 2$, the states are
479: all extended. At $\lambda=2$, all the states are critical.
480:
481: In the present case, take $\phi=\frac{p}{q}$ and write
482: \begin{eqnarray}
483: \psi_n = \sum_{l=0}^{q-1} e^{2\pi \phi nl} f_l.
484: \label{eq;fourier}
485: \end{eqnarray}
486: and substitute it into (\ref{eq;triharpera}), then
487: \begin{eqnarray}
488: -(t_be^{ik_y}+t_c e^{ik_x-ik_y+2\pi i\phi(l-\frac{1}{2})})f_{l-1}
489: -(t_be^{-ik_y}+t_c e^{-ik_x+ik_y-2\pi i\phi(l+\frac{1}{2})})f_{l+1}
490: -2t_a\cos(2\pi \phi l+k_x)f_l =Ef_l.
491: \label{eq;dual}
492: \end{eqnarray}
493: When $t_c=0$, (\ref{eq;dual}) becomes (\ref{eq;triharperk}) by
494: substituting $k_x \rightarrow k_x^{0}$ and $ k_y \rightarrow -k_y$ and
495: $\lambda \rightarrow \frac{4}{\lambda}$.
496: This is just the Aubry-Andr\'{e} duality when we
497: take the incommensurate limit of $\phi$. However, when $t_c \neq 0$,
498: (\ref{eq;dual}) and (\ref{eq;triharperk}) are not
499: transformed by (\ref{eq;fourier}) due to the term proportional
500: to $t_c$.
501:
502: Because of the symmetry of the triangular lattice, we can
503: consider duality involving $t_c$ by putting $t_a$ or $t_b$ to be zero.
504: Let us consider small $t_a$ limit. In the limit,
505: (\ref{eq;triharperc}) has the Aubry-Andr\'{e} duality for exchanging
506: $t_b$ and $t_c$. This implies that there is a duality between
507: $\lambda$ and $\mu$ for small $t_a$ limit.
508: It relates a state at $(\lambda,\mu)$ to the one at
509: $(2\mu,\frac{\lambda}{2})$ by the transformation (\ref{eq;fourier}).
510: Thus the phase diagram in $(\lambda,\mu)$ should
511: have a localization transitions on the line $\lambda=\mu$ for
512: small $t_a$ limit i.e. large $\lambda$ and $\mu$.
513:
514:
515: \subsection{Characteristic Polynomial}
516: When $\phi=p/q$ is rational, (\ref{eq;triharper})
517: is reduced to the eigenvalue problem of the matrix (\ref{eq;matrix}).
518: The eigenvalues are determined by the zeroes of
519: the characteristic polynomial
520: \begin{eqnarray}
521: P(E)=\det(E-H_q(k_x,k_y)).
522: \end{eqnarray}
523: which has been studied previously
524: \cite{thouless,hat-koh,han-thouless}. In Ref.\cite{han-thouless},
525: it was shown that $P(E)$ can be written in terms of
526: Chebyshev polynomial of order $q$. In the present case,
527: the characteristic polynomial takes a simple form as follows:
528: \begin{eqnarray}
529: P(E) &=& P_0(E)-Q(k_x,k_y) \\
530: Q(k_x,k_y)&=&(-1)^{q}4
531: (t_a^{q}\cos qk_x +t_b^{q}\cos qk_y)+(-1)^{p}t_c^{q}\cos q(k_x-k_y),
532: \end{eqnarray}
533: where $P_0(E)$ is independent of $k_x$ and $k_y$.
534: The energy bands are determined by the zeroes of the
535: polynomial $P(E)$ as we vary $k_x$ and $k_y$.
536: Especially, the edges of energy bands are determined by
537: the minimum and the maximum of the function $Q(k_x,k_y)$.
538: When $k_y=0$, they are given by $k_x=0, \pi$.
539:
540: In Ref.\cite{han-thouless}, the total band width $W$ of
541: the triangular lattice is
542: estimated when $t_b > t_a > t_c$ as
543: \begin{equation}
544: W \sim (t_b-t_a) g\left(q\frac{t_b-t_a}{t_b} \right),
545: \label{eq;totalband}
546: \end{equation}
547: where the scaling function $g(x)$ behaves as $\frac{9.3299}{x}$
548: when its argument is small. Since $t_c$ does not
549: enter the argument, the scaling of the total band width
550: of the triangular lattice is the same as the square lattice.
551: This suggests that the universality class of
552: the scaling property of the spectral measure of
553: the triangular lattice is the same as that of the square
554: lattice.
555:
556:
557:
558:
559: \section{\label{section;chara}Classical Orbits and Phase Diagram}
560: \subsection{Classical Orbits}
561:
562: %///////////////////// contour plot of classical orbit //////////////
563: \begin{figure}
564: % \includegraphics[angle=-90]
565: % \epsfxsize=8cm
566: \includegraphics[scale=1.0]{cont-1.0-0.5.eps}
567: \includegraphics[scale=1.0]{cont-1.0-1.0.eps}
568: \includegraphics[scale=1.0]{cont-1.0-1.5.eps}
569: \caption{ Contour plots of the classical orbits
570: for
571: $(\lambda , \mu) = (1.0 , 0.5),(1.0,1.0)$ and $(1.0,0.5)$.
572: }
573: \label{fig;contour-1.0-1.0}
574: \end{figure}
575:
576:
577: \begin{figure}
578: \includegraphics[scale=1.0]{cont-2.0-1.0.eps}
579: \includegraphics[scale=1.0]{cont-3.0-0.5.eps}
580: \caption{ Contour plots of the classical orbits
581: for
582: $(\lambda , \mu) = (2.0,1.0)$ and $(3.0,0.5)$.
583: }
584: \label{fig;contour-2.0-1.0}
585: \end{figure}
586:
587: The Hamiltonian (\ref{eq;model}) consists of three terms $H_a$, $H_b$ and $H_c$
588: which are noncommutative each other.
589: They are diagonalized in different bases
590: as in (\ref{eq;triharper}), (\ref{eq;triharperk}) and
591: (\ref{eq;triharperc}).
592: The ``classical'' Hamiltonian is thus
593: \begin{eqnarray}
594: H_{\rm classical} = 2t_a \cos k_x+ 2t_b \cos k_y +2t_c \cos (k_x-k_y).
595: \label{eq;classical}
596: \end{eqnarray}
597: In a magnetic field, $k_y$ is canonically conjugate to $k_x$ and vice versa.
598: Thus to analyze classical orbits, we replace $k_y$ by $x$ and
599: $k_x$ by $y$. Setting $t_a=1$, we plot the contour of
600: \begin{eqnarray}
601: H_{\rm classical} = \cos y + \frac{\lambda}{2} \cos x
602: + \mu \cos (y-x)
603: \label{eq;classical2}
604: \end{eqnarray}
605: in {\bf Fig. \ref{fig;contour-1.0-1.0}} for $(\lambda,\mu)= (1.0,0.5),(1.0,1.0)$ and $(1.0,1.5) $, and in {\bf Fig.\ref{fig;contour-2.0-1.0}} for $(\lambda,\mu)= (2.0,1.0)$ and $(3.0,1.5)$.
606: In {\bf Fig.\ref{fig;contour-1.0-1.0}},
607: we see that all the contours for $(\lambda,\mu)=(1.0,0.5) $ and $ (1.0,1.5)$ are
608: extended in the $x$-direction and localized in the $y$-direction while,
609: for $(\lambda,\mu )=(1.0,1.0)$, there is a separatrix which is
610: extended in both directions.
611: We also see in {\bf Fig.\ref{fig;contour-2.0-1.0}} that
612: the contours for $(\lambda,\mu)=(3.0,0.5)$ are
613: extended in $x$-direction and localized in $y$-direction, while
614: there is a separatrix for $(\lambda,\mu)=(2.0,1.0)$.
615:
616:
617: \subsection{Phase Diagram for irrational $\phi$}
618: From behaviors of the classical orbits shown in the previous section,
619: we may deduce the phase diagram of the equation
620: \begin{eqnarray}
621: -\left[1+\mu e^{-2\pi i \phi(n-\frac{1}{2})+ik_y}\right]\Psi_{n-1}
622: -\left[1+\mu e^{2\pi i \phi(n+\frac{1}{2}-ik_y)}\right]\Psi_{n+1}
623: -\lambda \cos\left(2\pi\phi n+k_y\right)\Psi_n =E\Psi_n.
624: \label{eq;triharperlambdamu2}
625: \end{eqnarray}
626: for irrational limit of $\phi=\frac{p}{q}$.
627: The phase diagram is shown in
628: {\bf Fig.\hspace{-.2cm} \ref{fig;phase_diagram}}.
629: %///////// fig;square lattice with NNN hopping //////////////
630: \begin{figure}
631: \begin{center}
632: \epsfxsize=8cm
633: \epsfbox{phase-diagram.eps}
634: \caption{
635: Phase diagram.
636: In region I and III
637: the wavefunctions (spectra) are extended
638: (absolutely continuous),
639: and in region II
640: the wavefunctions are localized (pure points).
641: On the three boundary lines,
642: the wavefunctions (spectra) are critical (singular continuous).
643: }
644: \label{fig;phase_diagram}
645: \end{center}
646: \end{figure}
647: One intriguing aspect is the
648: transitions between phase I and phase III
649: which are the transitions between metals.
650: Indeed a transition in the quantum case is not
651: characterized by an appearance of a separatrix at a certain
652: energy. For the Harper model, it is known that
653: metal-insulator transitions occur for whole energies
654: at $\lambda=2$. This is generalized to the triangular lattice model
655: we consider.
656:
657:
658: As an example of an incommensurate limit,
659: in the sections hereafter, we perform numerical scaling analysis
660: for the energy spectra and the critical wavefunctions
661: when $\phi=\frac{p}{q}$ approaches the inverse of the golden mean
662: $\frac{1}{\tau}=\frac{\sqrt{5}-1}{2}$. A standard sequence
663: which corresponds to the continued fraction expansion
664: of $\frac{1}{\tau}$ is the Fibonacci series $F_n$,
665: which is defined by $F_0=F_1=1$, $F_n=F_{n-1}+F_{n-2}$.
666: $F_n$ behaves $\sim \tau^{n}$ for large $n$.
667: By taking $p=F_{n-1},q=F_n$, $\phi=\frac{p}{q}$ approaches
668: $\frac{1}{\tau}$. $F_n$ is called a Fibonacci number
669: and $n$ is referred to Fibonacci index.
670:
671:
672: To take this incommensurate limit of (\ref{eq;triharperlambdamu2}),
673: the off-diagonal terms in (\ref{eq;triharperlambdamu2}) need
674: a special care. Namely, when $\mu=1$, these terms can be zero
675: if the exponential becomes $-1$. For the sequence above,
676: this actually happens when $q=F_n$ with $n=3\ell+1$ for some integer $\ell$.
677: In that case, the energy spectrum has no dependence on $k_x$,
678: and the dispersion relation is flat.
679:
680:
681:
682:
683:
684: \section{\label{section;level}Level Statistics}
685:
686: Consider (\ref{eq;triharperlambdamu2}) when $k_y=0$.
687: On the critical lines, the spectral measure
688: and the wavefunctions are expected to show characteristic
689: behaviors of criticality.
690: See {\bf Fig.\ref{fig;phase_diagram}}.
691: In order to obtain the distributions of the band widths, the $q \times q$ matrices (\ref{eq;matrix}) are diagonalized. The normalizations are
692: \begin{eqnarray}
693: \int^{\infty}_0 P_B(w) dw &=&1
694: \nonumber\\
695: \langle w \rangle = \int^{\infty}_0 w P_B(w)dw &=&1.
696: \end{eqnarray}
697: Similarly the distributions of the gaps $P_G(s)$
698: are obtained and normalized by
699:
700: \begin{eqnarray}
701: \int^{\infty}_0 P_G(s) ds &=&1
702: \nonumber\\
703: \langle s \rangle = \int^{\infty}_0 s P_G(s)ds &=&1.
704: \end{eqnarray}
705: As we discussed in Sec.\ref{section;chara}, the edges of the energy bands
706: are found at $k_x=0,\pi$ and $k_y=0,\pi$.
707: Thus, to study the measure of the spectrum of (\ref{eq;triharperlambdamu2}),
708: it is sufficient to study those points in the Brillouin zone.
709: When $\phi=\frac{p}{q}$ is a rational number, the problem is reduced to the
710: eigenvalue problem of the finite size matrix (\ref{eq;matrix}).
711: Furthermore, when $q$ is odd and $k_x=0,\pi$ and $k_y=0,\pi$,
712: the matrix (\ref{eq;matrix}) reduces to a tridiagonal form by
713: the symmetric and antisymmetric eigenstates \cite{thouless}.
714:
715:
716:
717:
718:
719: For $\mu=1$ with $q=F_n$ and $n=3\ell+1$, as we noted above,
720: the hopping term becomes zero at a bond
721: and all the band has zero width. Thus we study only the case of
722: $q=F_n$ with $n=3\ell$ when $\mu=1$.
723:
724:
725:
726: \subsection{ Distributions of Band Widths }
727:
728:
729: %///////////////////// TAIL lambda=2 mu=0.4 //////////////
730: \begin{figure}
731: \includegraphics[scale=0.6,angle=0]{evag-2.0-0.4-fig.eps}
732: \caption{ Distributions of the band widths for
733: $(\lambda , \mu) = (2.0,0.4)$.
734: }
735: \label{fig;bw-fig-2.0-0.4}
736: \end{figure}
737:
738:
739: %///////////////////// TAIL lambda=2 mu=0.4 //////////////
740: \begin{figure}
741: \includegraphics[scale=0.6,angle=0]{evag-2.0-0.4-logfig.eps}
742: \caption{ Semi-log plots of the distributions of the band widths
743: for $(\lambda , \mu) = (2.0,0.4)$. Inset:
744: Log-log plots of the distributions of the band widths
745: for $(\lambda , \mu) = (2.0,0.4)$. As seen in the inset,
746: the convergence of the distributions near zero at large $n$
747: is relatively slow.
748: For each value of $(\lambda,\mu)$,
749: we choose a stable part of the distributions to obtain the
750: exponents. This procedure potentially underestimates the values.
751: }
752: \label{fig;bw-figlog-2.0-0.4}
753: \end{figure}
754:
755:
756: %///////////////////// Origin lambda=1 mu=1 //////////////
757: %\begin{figure}
758: %\includegraphics[scale=0.3,angle=-90]{evag-loglog-2.0-0.4.eps}
759: % \caption{ Log-log plot for the band-width distribution
760: % of $(\lambda , \mu) = (1.0,1.0)$.
761: % }
762: % \label{fig;bw-loglog-2.0-0.4}
763: %\end{figure}
764:
765:
766: Consider the distributions of the band widths
767: along the line $\lambda = 2$ with $\mu=0.2,0.4,0.6$ and $0.8$.
768: In {\bf Fig. \ref{fig;bw-fig-2.0-0.4}}, $P_B(w)$ at $( \lambda , \mu )=( 2.0 , 0.4 )$
769: for $q=F_n$ with $n=25, 27$ and $ 28$ are plotted.
770: It shows convergence to a limit,
771: indicating the existence of a limit of the
772: distributions of the gaps for the incommensurate flux $\varphi$.
773: For $0 \leq \mu < 1$
774: the distributions depend on $\mu$.
775: The semi-log plots
776: of $P_B(w)$ is shown in {\bf Fig. \ref{fig;bw-figlog-2.0-0.4}}. One sees the linear behaviors
777: for large $w$, implying an asymptotic form
778: \begin{equation}
779: P_B(w) \sim e^{-\gamma w},\quad {\rm as}\quad w \rightarrow \infty,
780: \label{tail}
781: \end{equation}
782: where $\gamma >0.$ The optimized values of $\gamma$ are shown
783: in {\bf Table \ref{table;level}} for several $\mu$'s.
784: The inset of {\bf Fig.\ref{fig;bw-figlog-2.0-0.4}}
785: shows $P_B(w)$ near the origin which
786: indicates that the distributions of the band widths $P_B(w)$ are zero at the origin with a power law decay.
787: To characterize this behavior, we make an ansatz
788: \begin{equation}
789: P_B(w) \sim w^{\beta}, \quad {\rm as } \quad w \rightarrow 0.
790: \label{origin}
791: \end{equation}
792: where $\beta >0$.
793: The optimized values of $\beta$ are shown
794: in {\bf Table \ref{table;level}}.
795: One sees that $\beta$ becomes smaller as approaching to
796: $\mu=1$.
797:
798:
799:
800:
801: Next, we investigate $P_B(w)$ on the other lines $\mu=1$ and $\lambda=2\mu$.
802: In {\bf Fig. {\bf \ref{fig;bw-figlog-1.0-1.0} }},
803: the semi-log and the log-log
804: plots of $P_B(w)$ are shown for $(\lambda,\mu)=(1.0,1.0)$.
805: We find similar type of behaviors
806: (\ref{tail}) and (\ref{origin})
807: for the $\lambda=2$ line.
808: We also investigate the critical line $\lambda=2\mu$ and
809: find similar type of behaviors. We collect the values of
810: $\beta$,$\gamma$ in {\bf Table \ref{table;level}}.
811:
812:
813: These behaviors of $P_B(w)$ on these lines
814: are consistent with the behavior of $P_B(w)$ in other quasiperiodic
815: model \cite{evangelou,takada,naka} thus gives a support for
816: the phase diagram of {\bf Fig.\ref{fig;phase_diagram}}.
817:
818: %///////////////////// TAIL lambda=1 mu=1 //////////////
819: \begin{figure}
820: \includegraphics[scale=0.6,angle=0]{evag-1.0-1.0-logfig.eps}
821: \caption{ Semi-log plots of the distributions of the band widths
822: for $(\lambda , \mu) = (1.0,1.0)$. Inset:
823: Log-log plots of the distributions of the band widths
824: for $(\lambda , \mu) = (1.0,1.0)$. }
825: \label{fig;bw-figlog-1.0-1.0}
826: \end{figure}
827: %///////////////////// Origin lambda=1 mu=1 //////////////
828: %\begin{figure}
829: %\includegraphics[scale=0.3,angle=-90]{evag-loglog-1.0-1.0.eps}
830: % \caption{
831: % }
832: % \label{fig;bw-loglog-1.0-1.0}
833: %\end{figure}
834:
835:
836: \subsection{Distributions of Gaps}
837: The distribution of the gaps at $( \lambda , \mu) = ( 2.0 , 0.0 )$
838: has been known to follow an inverse power law \cite{machigei}
839: which diverges at the origin
840: \begin{equation}
841: P_G(s) \sim s^{-\delta}
842: \label{eq;gap_fit}
843: \end{equation}
844: with $\delta \sim 1.5$.
845: In {\bf Fig. \ref{fig;gap}}, the distributions of the gaps
846: for $(\lambda,\mu)=(2.0,0.4), (1.0,1.0)$ and $(2.0,1.0) $ are
847: shown. It is clear that
848: $P_{G}(s)$ shows a power law of the inverse.
849: The estimated value of $\delta$ is $\sim 1.5$ for these cases.
850: We also investigate other points on the lines $\lambda=2, \mu=1$
851: and $\lambda=2\mu$
852: and find a similar behavior with $\delta \sim 1.5$ within
853: statistical error. This behavior of $P_G(s)$ shows
854: that the spectra are singular continuous on these lines,
855: gives a further support for
856: the phase diagram {\bf Fig.\ref{fig;phase_diagram}}.
857: Also the value $\delta \sim 1.5$ seems to be
858: a characteristic quantity for this model.
859:
860:
861: %///////////////////// gap //////////////
862: \begin{figure}
863: \includegraphics[scale=0.3,angle=-90]{gap-loglog-2.0-0.4.eps}
864: \includegraphics[scale=0.3,angle=-90]{gap-loglog-1.0-1.0.eps}
865: \includegraphics[scale=0.3,angle=-90]{gap-loglog-2.0-1.0.eps}
866: \caption{ Log-log plots of the distributions of the gaps
867: for $(\lambda , \mu) = (2.0,0.4),(1.0 , 1.0)$.
868: }
869: \label{fig;gap}
870: \end{figure}
871:
872:
873: \subsection{Bicritical Point}
874: %/////////// about bicritical point //////////////
875: We also investigate the point
876: $( \lambda , \mu) = ( 2.0 , 1.0 )$.
877: The result is shown in
878: {\bf Fig.\ref{fig;bw-2.0-1.0}}.
879: In sharp contrast to other points on the critical lines,
880: it shows the inverse power law
881: \begin{equation}
882: P_B(w) \sim w^{-\beta'},
883: \end{equation}
884: ($\beta' >0$) for whole the range.
885: We estimate the exponent of the law as $\beta' \sim 1.4$.
886: This implies that the spectrum at this point
887: is a qualitively different fractal-like set.
888: On the other hand, the gap distribution is shown
889: in {\bf Fig.\ref{fig;gap-2.0-1.0}}. It is an
890: inverse power law
891: \begin{equation}
892: P_G(s) \sim s^{-\delta},
893: \end{equation}
894: ($\delta>0$)
895: with the exponent $\delta \sim 1.5$,
896: analogous to the ones found for
897: other critical points. Thus the band width distribution
898: gives a finer characterization of the spectra than
899: the gap distribution.
900:
901:
902:
903:
904: %///////////////////// bicritical point //////////////
905: \begin{figure}
906: \includegraphics[scale=0.3,angle=-90]{evag-loglog-2.0-1.0.eps}
907: \caption{ Log-log plots of the distributions of the band widths
908: for $(\lambda , \mu) = (2.0,1.0)$.
909: }
910: \label{fig;bw-2.0-1.0}
911: \end{figure}
912:
913:
914:
915: %///////////////////// gap //////////////
916: \begin{figure}
917: \includegraphics[scale=0.3,angle=-90]{gap-loglog-2.0-1.0.eps}
918: \caption{ Log-log plots of the distributions of the gaps
919: for $(\lambda , \mu) = (2.0,1.0)$.
920: }
921: \label{fig;gap-2.0-1.0}
922: \end{figure}
923:
924:
925: \setlength{\arrayrulewidth}{0.8pt}
926: \begin{table}[tb]
927: \begin{tabular}{ccccc} \hline
928: $\lambda$ & $\mu$ & $\beta $
929: & $\gamma$ & $\delta$ \\ \hline
930: 2.0 & 0.0 & 2.5 & 1.4 & 1.5 \\
931: 2.0 & 0.2 & 2.5 & 1.3 & 1.5 \\
932: 2.0 & 0.4 & 2.5 & 1.2 & 1.5 \\
933: 2.0 & 0.6 & 2.3 & 1.2 & 1.5 \\
934: 2.0 & 0.8 & 2.1 & 0.9 & 1.5 \\
935: \hline
936: 2.5 & 1.25& 2.1 & 0.9 & 1.5 \\
937: 3.0 & 1.5 & 2.3 & 1.1 & 1.5 \\
938: 4.0 & 2.0 & 2.4 & 1.2 & 1.5 \\
939: \hline
940: 0.0 & 1.0 & 2.6 & 1.6 & 1.5 \\
941: 0.5 & 1.0 & 2.6 & 1.5 & 1.5 \\
942: 1.0 & 1.0 & 2.4 & 1.3 & 1.5 \\
943: 1.5 & 1.0 & 2.2 & 1.0 & 1.5 \\
944:
945: \end{tabular}
946: \caption{
947: Estimated exponents on the critical lines.
948: For definitions of $\beta,\gamma$ and $\delta$,
949: see
950: Eqs.(\ref{tail}),
951: (\ref{origin}), and
952: (\ref{eq;gap_fit}) respectively.}
953: \label{table;level}
954: \end{table}
955:
956:
957:
958: \section{\label{section;multi}Multifractal Analysis}
959:
960: We apply the method of multifractal analysis \cite{halsey}
961: to the spectra and the critical wavefunctions. In Ref.\cite{kohmoto},
962: the entropy function was introduced which reformulates
963: the theory along
964: the way that standard statistical mechanics is formulated.
965: We use it in our analysis.
966:
967: \subsection{Review of Multifractal Analysis}
968: We consider quantities $l_i$ and their probability measure $p_i$ of
969: a fractal-like set.
970: Though we shall only consider the cases where
971: $l_i$ or $p_i$ is a constant, a general formulation is reviewed for
972: convenience.
973: It is natural to consider distributions of logarithm of $l_i$
974: \begin{equation}
975: \varepsilon_i = -\frac{\ln l_i}{n}, \quad i.e.\quad l_i=\exp(-n\varepsilon).
976: \label{kohmoto:eq2.1}
977: \end{equation}
978: As $n$ becomes large, $l_i$ approaches zero , but $\epsilon_i$ takes a
979: finite nonzero value for critical points.
980: We introduce a scale index $\alpha_i$ as the exponent
981: of $p_i$ measured by $l_i$ as
982: \begin{equation}
983: p_i = l_i^{\alpha_i}, \hspace{5mm}
984: \alpha_i = -\frac{1}{\epsilon_i}\frac{1}{n} \ln p_i.
985: \label{kohmoto:eq3.1a}
986: \end{equation}
987: We write the number of $l_i$ whose scale index lies between
988: $\varepsilon$ and $\varepsilon+d\varepsilon$, $\alpha$ and
989: $\alpha+d\alpha$ as $\Omega(\varepsilon,\alpha)d\varepsilon d\alpha$.
990: We take an ansatz that $\Omega(\varepsilon,\alpha)$ has
991: the following scaling form for large $n$
992: \begin{equation}
993: \Omega(\varepsilon,\alpha) = \exp[nQ(\varepsilon,\alpha)],
994: \label{kohmoto:eq3.2}
995: \end{equation}
996: where $Q(\varepsilon,\alpha)$ can be seen as a kind of
997: entropy function.
998:
999:
1000: Following \cite{halsey,kohmoto}, we consider the generalized
1001: partition function
1002: \begin{eqnarray}
1003: \Gamma(q,\beta) &=& \sum_i p_i^{q}l_i^{\beta} \\
1004: &=& \sum_i \exp[-n\varepsilon_i(\alpha_iq+\beta)].
1005: \label{kohmoto:eq3.3}
1006: \end{eqnarray}
1007: The generalized free energy is
1008: \begin{eqnarray}
1009: G(q,\beta) = \frac{1}{n} \ln \Gamma(q,\beta),
1010: \label{kohmoto:eq3.4}
1011: \end{eqnarray}
1012: % Obviously, we have the relations
1013: %\begin{eqnarray}
1014: %Z(\beta) = \Gamma(q=0,\beta), \quad F(\beta) = G(q=0,\beta).
1015: %\label{kohmoto:eq3.5}
1016: %\end{eqnarray}
1017: Using $Q(\varepsilon,\alpha)$, (\ref{kohmoto:eq3.3}) is written
1018: \begin{eqnarray}
1019: \Gamma(q,\beta) = \int d\varepsilon \int d\alpha
1020: \exp[n[Q(\varepsilon,\alpha)-(\alpha q+\beta)\varepsilon]].
1021: \label{kohmoto:eq3.6}
1022: \end{eqnarray}
1023: For large $n$, the maximum of the exponent dominates
1024: the integral and gives
1025: \begin{eqnarray}
1026: G(q,\beta) = Q(\vev{\varepsilon}, \vev{\alpha}) - (\vev{\alpha}q+\beta)
1027: \vev{\varepsilon},
1028: \label{kohmoto:eq3.7}
1029: \end{eqnarray}
1030: where $\vev{\varepsilon}$ and $\vev{\alpha}$ give the maximum
1031: of $Q(\varepsilon,\alpha)-(\alpha q +\beta)\varepsilon$, so we have
1032: \begin{eqnarray}
1033: \frac{\partial Q(\varepsilon,\alpha)}{\partial \varepsilon}
1034: |_{\varepsilon=\vev{\varepsilon},\alpha=\vev{\alpha}} = \vev{\alpha}+\beta.
1035: \label{kohmoto:eq3.8}
1036: \end{eqnarray}
1037: and
1038: \begin{eqnarray}
1039: \frac{\partial Q(\varepsilon,\alpha)}{\partial \alpha} = \vev{\varepsilon}q.
1040: \label{kohmoto:eq3.9}
1041: \end{eqnarray}
1042: Thus $G(q,\beta)$ is obtained from $Q(\varepsilon, \alpha)$ using
1043: (\ref{kohmoto:eq3.7})(\ref{kohmoto:eq3.8}) and (\ref{kohmoto:eq3.9}).
1044: From (\ref{kohmoto:eq3.9}),
1045: the maximum of $Q(\varepsilon,\alpha)$ with respect to
1046: $\alpha$ occurs when $q=0$.
1047: On the other hand, once $G(q,\beta)$ is calculated,
1048: $\vev{\varepsilon},\vev{\alpha}$ and $Q(\vev{\epsilon},\vev{\alpha})$
1049: are given by
1050: \begin{eqnarray}
1051: \vev{\varepsilon} = -\frac{\partial}{\partial \beta}G(q,\beta),
1052: \label{kohmoto:eq3.10} \hspace{5mm}
1053: \vev{\alpha}\vev{\varepsilon} = - \frac{\partial}{\partial q} G(q,\beta),
1054: \label{kohmoto:eq3.11}
1055: \end{eqnarray}
1056: and
1057: \begin{eqnarray}
1058: Q(\vev{\varepsilon},\vev{\alpha}) = G(q,\beta)
1059: - q\frac{\partial G(q,\beta)}{\partial q}
1060: - \beta\frac{\partial G(q,\beta)}{\partial \beta}.
1061: \label{kohmoto:eq3.12}
1062: \end{eqnarray}
1063: Since $\vev{\varepsilon}$ and $\vev{\alpha}$ are functions of $q$ and
1064: $\beta$, different regions with scaling indices $\varepsilon$ and
1065: $\alpha$ are explored by changing the values of the parameters $q$ and
1066: $\beta$. Thus $Q(\vev{\varepsilon},\vev{\alpha})$ is implicitly
1067: a function of $q$ and $\beta$.
1068:
1069: %One can introduce the entropy $S(\varepsilon)$ and
1070: %\begin{equation}
1071: %\exp[nS(\varepsilon)] = \int d\varepsilon \exp[nQ(\varepsilon,\alpha)].
1072: %\label{kohmoto:eq3.14}
1073: %\end{equation}
1074:
1075:
1076: The limit of $G(q,\beta)$ for large $n$, may be obtained by
1077: \begin{eqnarray}
1078: G(q,\beta_c(q)) = 0,
1079: \label{kohmoto:eq3.19}
1080: \end{eqnarray}
1081: and $\beta_c(q)$ can be regarded as a set of generalized dimensions. The scaling index $\vev{\varepsilon}_c$ which corresponds to $\beta_c(q)$ could be
1082: considered as being a representative for a particular value of $q$.
1083:
1084: From (\ref{kohmoto:eq3.7})
1085: (\ref{kohmoto:eq3.8}) and (\ref{kohmoto:eq3.19}), we see that
1086: $Q(\vev{\varepsilon},\vev{\alpha})$ at the critical point satisfies the
1087: relation
1088: \begin{eqnarray}
1089: Q(\vev{\varepsilon}_c,\vev{\alpha}_c) &=&
1090: \frac{\partial Q(\epsilon,\vev{\alpha}_c)}{\partial \varepsilon}|_{\varepsilon=\vev{\varepsilon}_c}\vev{\varepsilon}_c = f(\vev{\alpha}_c)\vev{\varepsilon}_c,
1091: \label{kohmoto:eq3.21}
1092: \end{eqnarray}
1093: where $f(\vev{\alpha}_c)$ is given by
1094: \begin{equation}
1095: f(\vev{\alpha}_c) =
1096: \frac{\partial Q(\epsilon,\vev{\alpha}_c)}{\partial \varepsilon}|_{\varepsilon=\vev{\varepsilon}_c}.
1097: \label{kohmoto:eq3.22}
1098: \end{equation}
1099: By substituting (\ref{kohmoto:eq3.22}) into (\ref{kohmoto:eq3.8}) and
1100: (\ref{kohmoto:eq3.9}), we obtain
1101: \begin{equation}
1102: f(\vev{\alpha}_c) = \vev{\alpha}_c q+\beta_c(q)
1103: \label{kohmoto:eq3.23}
1104: \end{equation}
1105: and
1106: \begin{equation}
1107: \frac{df(\vev{\alpha}_c)}{d\alpha}|_{\alpha=\vev{\alpha}_c}
1108: = q,
1109: \label{kohmoto:eq3.24}
1110: \end{equation}
1111: respectively. And (\ref{kohmoto:eq3.23}) and (\ref{kohmoto:eq3.24}) give
1112: \begin{equation}
1113: \vev{\alpha}_c = -\frac{d \beta_c(q)}{dq}.
1114: \label{kohmoto:eq3.25}
1115: \end{equation}
1116: Thus once $\beta_c(q)$ is known by solving (\ref{kohmoto:eq3.19}),
1117: $\vev{\alpha}_c$ and
1118: $f(\vev{\alpha}_c)$ are obtained from (\ref{kohmoto:eq3.23}) and (\ref{kohmoto:eq3.25}).
1119: In terms of $f(\alpha)$, the density function of
1120: $\varepsilon$ and $\alpha$ $\Omega(\varepsilon,\alpha)$
1121: is written, using (\ref{kohmoto:eq3.2}) and (\ref{kohmoto:eq3.21}) as
1122: \begin{eqnarray}
1123: \Omega(\vev{\varepsilon},\vev{\alpha}_c) =\exp[n\vev{\varepsilon}_cf(\vev{\alpha}_c)] = \vev{l}_c^{-f(\vev{\alpha}_c)},
1124: \label{kohmoto:eq3.26}
1125: \end{eqnarray}
1126: where $\vev{l}_c = \exp(-n\vev{\varepsilon}_c)$ is a representative
1127: length, and $\vev{\varepsilon}_c$ and $\vev{\alpha}_c$ are
1128: functions of $q$ [see (\ref{kohmoto:eq3.10}) and (\ref{kohmoto:eq3.25})].
1129: $f(\alpha)$ can be considered to be a set of generalized dimensions.
1130: In numerical approach, we calculate $f(\alpha)$ for a given fractal-like
1131: object for a finite $n$ and extrapolate it to the limit
1132: $n \rightarrow \infty$. From $G(q,\beta)$ (\ref{kohmoto:eq3.4}), $G_n(q,\beta)$ at
1133: finite $n$ should behave as $G_n(q,\beta) \sim G(q,\beta) + O(\frac{1}{n})$.
1134: Thus we should extrapolate $G_n(q,\beta)$ as a function of $\frac{1}{n}$
1135: and estimate the limit for $n \rightarrow \infty$.
1136:
1137:
1138: We denote the support of $f(\alpha)$ by $[\alpha_{\rm min},\alpha_{\rm max}]$
1139: and the value of $\alpha$ which gives the maximum of $f(\alpha)$
1140: by $\alpha_0$.
1141:
1142: \subsection{Spectrum}
1143: In this section,
1144: scaling properties of energy spectra are analyzed by the multifractal analysis.
1145:
1146: \subsubsection{Multifractal Analysis of Spectrum}
1147: We apply the general formulation to characterization of the
1148: energy spectra. Take the band widths as variables $l_i$, and let $p_i$ be
1149: constants
1150: \begin{equation}
1151: p_i=\frac{1}{F_n} \sim \frac{1}{\tau^n},
1152: \label{kohmoto:eq4.7}
1153: \end{equation}
1154: where $\tau$ is the golden mean. As seen from (\ref{kohmoto:eq3.4}),
1155: \begin{eqnarray}
1156: G(q,\beta) = -q\ln \tau +F(\beta), \hspace{5mm} F(\beta)
1157: = G(0,\beta) = \frac{1}{n}\ln \Gamma(0,\beta)
1158: \label{kohmoto:eq4.8}.
1159: \end{eqnarray}
1160: For each $q$, the critical value $\beta$ is determined and
1161: vice versa.
1162: From (\ref{kohmoto:eq3.10}) and (\ref{kohmoto:eq4.8}), the
1163: scaling index $\vev{\varepsilon}$ is witten
1164: \begin{equation}
1165: \vev{\varepsilon} = -\frac{\partial}{\partial \beta} G(q,\beta)
1166: =-\frac{\partial}{\partial \beta} F(\beta).
1167: \label{kohmoto:eq4.9}
1168: \end{equation}
1169: Thus $\vev{\varepsilon}$ depends only on $\beta$.
1170: From (\ref{kohmoto:eq3.11}) and (\ref{kohmoto:eq4.8}),
1171: $\vev{\alpha}$ is related to $\vev{\varepsilon}$ by
1172: \begin{equation}
1173: \vev{\alpha} = \ln \tau / \vev{\varepsilon},
1174: \label{kohmoto:eq4.10}
1175: \end{equation}
1176: and $Q(\varepsilon,\alpha)$ is nonzero only for $\alpha$ satisfying
1177: (\ref{kohmoto:eq4.10}), thus depends only on $\varepsilon$.
1178: We may put $Q(\varepsilon,\alpha)$ as $S(\varepsilon)$, and
1179: from (\ref{kohmoto:eq3.12}) and (\ref{kohmoto:eq3.21}) we get
1180: \begin{eqnarray}
1181: S(\varepsilon)= F(\beta) +\varepsilon\beta=\varepsilon f(\alpha).
1182: \label{kohmoto:eq4.11}
1183: \end{eqnarray}
1184: Then $f(\alpha)$ is calculated from the formula
1185: \begin{equation}
1186: f(\alpha) = \frac{S(\varepsilon)}{\varepsilon}.
1187: \end{equation}
1188:
1189:
1190:
1191: \subsubsection{Numerical Results}
1192: We apply the method above to the spectra of our model.
1193: In {\bf Fig.\ref{fig;alpha-falpha-1.0-1.0}}, we show $\alpha$-$f(\alpha)$
1194: curve for the spectrum at $(\lambda,\mu)=(1.0,1.0)$.
1195: The estimated values of $\alpha_{\rm min}$, $\alpha_{\rm max}$
1196: and $\alpha_0$ are $0.421$, $0.547$ and $0.495$ respectively.
1197: These values coincide with the corresponding values of the Harper
1198: model \cite{hiramoto}. We also investigated other points
1199: on lines $\lambda=2,\mu=1, \lambda=2\mu$ in {\bf Fig.\ref{fig;phase_diagram}}.
1200: Except for $(\lambda,\mu)=(2.0,1.0)$, it turns out that
1201: the estimated values of $\alpha_{\rm min}$, $\alpha_{\rm max}$
1202: and $\alpha_0$ are the same as those of the Harper model.
1203: This implies that the universality class for these lines
1204: is the same as the Harper model. This is
1205: consistent
1206: with the scaling of the total band widths (\ref{eq;totalband})
1207: for $\lambda =2,\mu<1$.
1208: On the other hand, $\alpha$-$f(\alpha)$
1209: curve for the spectrum at $(\lambda,\mu)=(2.0,1.0)$ has a different shape
1210: as shown in {\bf Fig.\ref{fig;alpha-falpha-2.0-1.0}}.
1211: We see that $\alpha_{\rm min}=0.381$, $\alpha_{\rm max}=0.755$ and
1212: $\alpha_0=0.498$. Thus the universality of this point is different from
1213: the Harper model.
1214:
1215:
1216: %///////////////////// alpha-falpha around lam=1.0,mu=1.0 //////////////
1217: \begin{figure}
1218: \includegraphics[scale=0.6,angle=-90]{alpha-falpha-1.0-1.0.eps}
1219: \caption{ $\alpha$-$f(\alpha)$ curves of the spectrum
1220: for $(\lambda , \mu) = (1.0 , 1.0)$.
1221: }
1222: \label{fig;alpha-falpha-1.0-1.0}
1223: \end{figure}
1224:
1225:
1226: %///////////////////// alpha-falpha around lam=1.0,mu=1.0 //////////////
1227: \begin{figure}
1228: \includegraphics[scale=0.6,angle=-90]{alpha-falpha-2.0-1.0.eps}
1229: \caption{ $\alpha$-$f(\alpha)$ curves of the spectrum
1230: for $(\lambda , \mu) = (2.0 , 1.0)$. }
1231: \label{fig;alpha-falpha-2.0-1.0}
1232: \end{figure}
1233:
1234: \subsection{Wavefunctions}
1235:
1236: We investigate
1237: scaling properties of the wavefunctions by the multifractal analysis.
1238: We concentrate on the eigenfunctions at
1239: the centers of the spectra.
1240: This enables us to confirm the phase diagram
1241: {\bf Fig.\ref{fig;phase_diagram}}.
1242:
1243:
1244: \subsubsection{Multifractal Analysis of Wavefunctions}
1245: We apply the general formulation to characterize the wavefunctions. Take squares modulus of
1246: the wavefunctions to be variables $p_i$, while take $l_i$
1247: to be constants
1248: \begin{equation}
1249: l_i=l=\frac{1}{F_n} \sim \frac{1}{\tau^n},\hspace{5mm}
1250: \epsilon = - \frac{1}{n} \ln l \sim \ln \tau.
1251: \end{equation}
1252: Thus $\varepsilon$ is a constant in this case.
1253: From (\ref{kohmoto:eq3.3}) and (\ref{kohmoto:eq3.4}), one has
1254: \begin{eqnarray}
1255: G(q,\beta) = -\beta\varepsilon +G(q,0).
1256: \label{kohmoto:eq4.2}
1257: \end{eqnarray}
1258: Using (\ref{kohmoto:eq3.11}) and ({\ref{kohmoto:eq3.12}),
1259: we obtain the generalized entropy
1260: \begin{eqnarray}
1261: Q(\varepsilon,\alpha) = G(q,0)-q\varepsilon \vev{\alpha}, \hspace{5mm}
1262: \vev{\alpha} = -\frac{1}{\varepsilon} \frac{\partial G(q,0)}{\partial q}
1263: \label{kohmoto:eq4.3}.
1264: \end{eqnarray}
1265: Since $Q(\varepsilon,\alpha)$ is nonzero only for $\vev{\alpha}$,
1266: we write it as $S'(\alpha)$.
1267: From (\ref{kohmoto:eq3.21}), we have
1268: \begin{eqnarray}
1269: f(\alpha) = \frac{S'(\alpha)}{\varepsilon}.
1270: \end{eqnarray}
1271: We calculate $f(\alpha)$ for finite Fibonacci index $n$
1272: by this formula and extrapolate them to $n \rightarrow \infty$.
1273: %We will
1274: %denote the support of $f(\alpha)$ by $[\alpha_{\rm min},\alpha_{\rm max}]$.
1275: Actually only a part of $f(\alpha)$ is required to
1276: distinguish localized, extended and critical states.
1277: For a localized state, $f(\alpha)$ has a point support
1278: and takes nonzero value only at $\alpha_{\rm min} =0$ and
1279: $\alpha_{\rm max}=\infty$ and $f(\alpha_{\rm min})=0$,
1280: $f(\alpha_{\rm max})=1$.
1281: For an extended state, it has
1282: $\alpha_{\rm min}=\alpha_{max}=1$ and $f(\alpha_{\rm min })
1283: =f(\alpha_{\rm max})=1$.
1284: For a critical state, $f(\alpha)$ may
1285: have a finite interval $[\alpha_{\rm min},\alpha_{max}]$ as
1286: a support and $f(\alpha)$ takes various values.
1287: We shall use this method to distinguish states near critical points.
1288:
1289:
1290:
1291:
1292:
1293: %///////////////////// wavefunction around lam=1.0,mu=1.0 //////////////
1294: \begin{figure}
1295: % \includegraphics[angle=-90]
1296: % \epsfxsize=8cm
1297: \includegraphics[scale=0.3,angle=-90]{wav-1.0-1.1.eps}
1298: \includegraphics[scale=0.3,angle=-90]{wav-1.0-1.0.eps}
1299: \includegraphics[scale=0.3,angle=-90]{wav-1.0-0.9.eps}
1300: \caption{ Plots of
1301: the wavefunctions at the centers of the spectra
1302: for (a) $(\lambda , \mu) = (1.0 , 1.1)$ ,
1303: (b) $(1.0,1.0)$ and
1304: (c) $(1.0,0.9)$.
1305: Here $q = 17711$.
1306: }
1307: \label{fig;wav-1.0-1.0}
1308: \end{figure}
1309:
1310:
1311: %///////////////////// wavefunction around bicritical point //////////////
1312: \begin{figure}
1313: \includegraphics[scale=0.3,angle=-90]{wav-2.0-1.0.eps}
1314: \includegraphics[scale=0.3,angle=-90]{wav-2.0-1.1.eps}
1315: \includegraphics[scale=0.3,angle=-90]{wav-1.9-0.9.eps}
1316: \includegraphics[scale=0.3,angle=-90]{wav-2.1-1.0.eps}
1317: \caption{ Plots of the wavefunctions at the centers of the spectra
1318: for (a) $(\lambda , \mu) = (2.0 , 1.0)$ ,
1319: (b) $(2.0,1.1)$,
1320: (c) $(1.9,0.9)$
1321: and (d) $(2.1,1.0)$.
1322: Here $q = 17711$ for (a), (b) and (c), and $q=4181$ for (d).
1323: }
1324: \label{fig;wav-2.0-1.0}
1325: \end{figure}
1326:
1327:
1328:
1329:
1330:
1331: \subsubsection{Numerical Results}
1332: We numerically obtain the wavefunctions at the centers
1333: of the spectra for odd $q$ on the $\lambda=2$, $\mu=1$ and $\lambda=2\mu$
1334: lines in the phase diagram {\bf Fig.\ref{fig;phase_diagram}}.
1335: For $\mu=1$, we investigate $q=F_n$ with $n=3\ell$ as well as
1336: $3\ell+1$. Although the dispersion relations are
1337: flat when $n=3\ell+1$, we find that
1338: the wavefunctions still show a characteristic
1339: behavior of a critical state.
1340:
1341:
1342: In {\bf Fig.\ref{fig;wav-1.0-1.0}},
1343: the square moduli of the wavefunctions at
1344: the centers of spectra for $(\lambda,\mu)=(1.0,1.1) ,
1345: (1.0,1.0)$ and $(1.0,0.9)$ are displayed for $n=21$ and $F_n=17711$.
1346: From these figures, we see that
1347: the wavefunctions are extended for $(1.0,1.1)$
1348: and for $(1.0,0.9)$, and critical for $(1.0,1.0)$
1349: which is in accord with the phase diagram {\bf Fig.\ref{fig;phase_diagram}}.
1350: In {\bf Fig.\ref{fig;wav-2.0-1.0}}, the square modulus
1351: of the wavefunctions at
1352: the band center for $(\lambda,\mu)=(2.0,1.0) ,
1353: (2.0,1.1),(1.9,0.9)$ and $(2.1,1.0)$ i.e. in the vicinity of the
1354: bicritical point of {\bf Fig.\ref{fig;phase_diagram}}
1355: are displayed for $n=21$ and $F_n=17711$
1356: ($n=18$ and $F_n=4181$ for $(2.1,1.0)$).
1357: It is rather clear that
1358: the wavefunction is extended for $(2.0,1.1)$ and $(1.9,0.9)$,
1359: localized for $(1.0,0.9)$ and critical for $(2.0,1.0)$.
1360: To draw convincing conclusions, however,
1361: it is necessary to study the scaling properties by multifractal analysis.
1362:
1363:
1364:
1365:
1366: We plot $\alpha_{\rm min}$ for $(\lambda,\mu)=(1.0,1.1),
1367: (1.0,1.0)$ and $(1.0,0.9)$ in {\bf Fig.\ref{fig;alpha-1.0-1.0}}.
1368: For $(\lambda,\mu)=(1.0,0.9)$ and $(1.0,1.1)$,
1369: it is clearly seen that $\alpha_{\rm min}$ extrapolates to $1$ for
1370: $n \rightarrow \infty$. On the other hand, $\alpha_{\rm min}$
1371: is extrapolated to $0.358$ for $(\lambda,\mu)=(1.0,1.0)$.
1372: This value of $\alpha_{\rm min}$ is actually the same as
1373: the one found in the Harper model \cite{hiramoto} within
1374: statistical error.
1375: As shown in {\bf Fig.\ref{fig;falpha-1.0-1.0}}
1376: $f(\alpha_{\rm min})$ extrapolates to $1$ for $(\lambda,\mu)=(1.0,1.1)$
1377: and $(1.0,0.9) $, and $0$ for $(1.0,1.0)$.
1378: The behaviors of $\alpha_{\rm min}$ and $f(\alpha_{\rm min})$ in
1379: {\bf Fig.\ref{fig;alpha-1.0-1.0} -\ref{fig;falpha-1.0-1.0} }
1380: indicate that the state is extended
1381: for $(\lambda,\mu)=(1.0,1.1)$ and $(1.0,0.9) $, and critical for $(1.0,1.0)$.
1382: This confirms a part of
1383: the phase diagram in {\bf Fig.\ref{fig;phase_diagram}},
1384: especially the metal-metal transitions at $\mu=1.0$.
1385:
1386:
1387: %///////////////////// alpha_min at lam=1.0 mu=1.0 //////////////
1388: \begin{figure}
1389: \includegraphics[scale=0.3,angle=-90]{alpha-1.0-1.0.eps}
1390: \caption{ Plots of $\alpha_{\rm min}$ vs. $\frac{1}{n}$
1391: near $(\lambda,\mu)=(1.0,1.0)$ with
1392: $n=12,13,15,16,17,18,19,21$ and $22$. }
1393: \label{fig;alpha-1.0-1.0}
1394: \end{figure}
1395:
1396:
1397: %///////////////////// f_alpha_min at lam=1.0 mu=1.0 //////////////
1398: \begin{figure}
1399: \includegraphics[scale=0.3,angle=-90]{falpha-1.0-1.0.eps}
1400: \caption{ Plots of $f(\alpha_{\rm min})$ against $\frac{1}{n}$ near
1401: $(\lambda,\mu)=(1.0,1.0)$. }
1402: \label{fig;falpha-1.0-1.0}
1403: \end{figure}
1404:
1405:
1406:
1407: %///////////////////// alpha_min at lam=2.0 mu=1.0 //////////////
1408: \begin{figure}
1409: \includegraphics[scale=0.3,angle=-90]{alpha-2.0-1.0.eps}
1410: \caption{ Plots of $\alpha_{\rm min}$ against $\frac{1}{n}$
1411: near $(\lambda,\mu)=(2.0,1.0)$. }
1412: \label{fig;alpha-2.0-1.0}
1413: \end{figure}
1414:
1415:
1416:
1417:
1418: %///////////////////// f_alpha_min at lam=2.0 mu=1.0 //////////////
1419: \begin{figure}
1420: \includegraphics[scale=0.3,angle=-90]{falpha-2.0-1.0.eps}
1421: \caption{ Plots of $f(\alpha_{\rm min})$ vs. $\frac{1}{n}$
1422: near $(\lambda,\mu)=(2.0,1.0)$. }
1423: \label{fig;falpha-2.0-1.0}
1424: \end{figure}
1425:
1426:
1427:
1428: %///////////////////// alpha_min at lam=3.0 mu=1.5 //////////////
1429: \begin{figure}
1430: \includegraphics[scale=0.3,angle=-90]{alpha-3.0-1.5.eps}
1431: \caption{ Plots of $\alpha_{\rm min}$ against $\frac{1}{n}$
1432: near $(\lambda,\mu)=(3.0,1.5)$. }
1433: \label{fig;alpha-3.0-1.5}
1434: \end{figure}
1435:
1436:
1437:
1438: %///////////////////// f_alpha_min at lam=3.0 mu=1.5 //////////////
1439: \begin{figure}5
1440: \includegraphics[scale=0.3,angle=-90]{falpha-3.0-1.5.eps}
1441: \caption{ Plots of $f(\alpha_{\rm min})$ against $\frac{1}{n}$
1442: near $(\lambda,\mu)=(3.0,1.5)$. }
1443: \label{fig;falpha-3.0-1.5}
1444: \end{figure}
1445:
1446:
1447:
1448:
1449: In {\bf Fig. \ref{fig;alpha-2.0-1.0}}, $\alpha_{\rm min}$'s are shown for the
1450: states near the bicritical point $(\lambda,\mu)=(2.0,1.0),
1451: (2.0,1.1),(1.9,0.9)$ and $(2.1,1.0)$. Also
1452: {\bf Fig. \ref{fig;falpha-2.0-1.0}} shows $f(\alpha_{\rm min})$'s for
1453: them.
1454: For $(2.0,1.1)$ and $(1.9,0.9)$, both $\alpha_{\rm min}$ and
1455: $f(\alpha_{\rm min})$ are extrapolated to $1$, telling that
1456: the state is extended. For $(2.1,1.0)$, both $\alpha_{\rm min}$ and
1457: $f(\alpha_{\rm min})$ are extrapolated to $0$, which means that
1458: the state is localized. At $(2.0,1.0)$, the convergence of $\alpha_{\rm min}$ seems slow
1459: but the plots show a tendency to converge to a finite value
1460: near $0.47$. Similarly $f(\alpha_{\rm min})$ converges
1461: to zero. Thus we conclude that the states are critical at
1462: $(2.0,1.0)$ and it is the bicritical point of
1463: metal-insulator and metal-metal transitions. See
1464: {\bf Fig.{\ref{fig;phase_diagram}}}.
1465:
1466: Next {\bf Fig.{\ref{fig;alpha-3.0-1.5}}} and
1467: {\bf Fig.{\ref{fig;falpha-3.0-1.5}}} show
1468: the scaling of $\alpha_{\rm min}$ and $f(\alpha_{\rm min})$ respectively
1469: near $(\lambda,\mu)=(3.0,1.5)$. The extrapolated values of
1470: $\alpha_{\rm min}$ and $f(\alpha_{\rm min})$ are consistent with
1471: the phases diagram {\bf Fig.{\ref{fig;phase_diagram}}}.
1472: We investigate $\alpha_{\rm min}$ and $f(\alpha_{\rm min})$
1473: for other points on the critical lines and the results are consistent with the $(\lambda,\mu)$- phase diagram
1474: {\bf Fig.{\ref{fig;phase_diagram}}}
1475:
1476:
1477: %///////////////////// a-fa curve at lam=2.0 mu=0.4 //////////////
1478: \begin{figure}
1479: \includegraphics[scale=0.6,angle=-90]{alpha-falpha-2.0-0.4.eps}
1480: \caption{ $\alpha$-$f(\alpha)$ curve for $(\lambda,\mu)=(2.0,0.4)$. }
1481: \label{fig;alpha-falpha-2.0-0.4}
1482: \end{figure}
1483:
1484:
1485: Let us next turn to whole $\alpha$-$f(\alpha)$ curve.
1486: In {\bf Fig.\ref{fig;alpha-falpha-2.0-0.4}}
1487: the $\alpha$-$f(\alpha)$ curve at
1488: $(\lambda,\mu)=(2.0,0.4)$ is shown. Within statistical error, $\alpha_{\rm min}$ is
1489: $0.358$.
1490: This value of $\alpha_{\rm min}$ holds
1491: on the critical lines except
1492: in the vicinity of the bicritical point,
1493: where $\alpha_{\rm min}$ slightly changes about $0.05$.
1494: %It is the same as the Harper model $(\lambda,\mu)=(2.0,0.0)$.
1495: In {\bf Fig.\ref{fig;alpha-falpha-2.0-0.4}},
1496: the value of $\alpha_0$ which gives the maximum
1497: of $f(\alpha)$ is observed to be $1.31$ which is also
1498: the same as the Harper model \cite{hiramoto}.
1499: It is the same
1500: for the other points on the critical lines $\lambda=2.0$
1501: and $\lambda=2\mu$ within
1502: statistical error, except in the vicinity of the bicritical point
1503: where $\alpha_0$ slightly changes within $0.05$.
1504: We may interpret these slight changes of $\alpha_{\rm min}$ and $\alpha_0$
1505: in the vicinity of the bicritical point as an effect of
1506: slow convergence there (see below).
1507: For $f(\alpha)$ for $\alpha > \alpha_0$,
1508: the convergence of $f(\alpha)$ is not good, especially
1509: for $\alpha_{\rm max}$.
1510:
1511:
1512: For the critical line $\mu=1$, $\alpha_0$ splits for $n=3 \ell$
1513: and $n=3 \ell +1 $. In {\bf Fig.\ref{fig;alpha0-1.0-1.0}},
1514: plots of $\alpha_0$ and $\alpha_{\rm min}$ for
1515: $(\lambda,\mu) =(1.0,1.0)$ are shown.
1516: It is seen that $\alpha_0$ goes to different
1517: values for $n=3\ell$ and $n=3\ell+1$. Similar
1518: behavior of $\alpha_0$ is observed for other points
1519: on the $\mu=1.0$ critical line.
1520: This implies that the universality class is different
1521: for these two series.
1522: Also the $\alpha$-$f(\alpha)$ curves for $n=3\ell$ and $ 3\ell+1$ are
1523: shown in {\bf Fig.\ref{fig;alpha-falpha-1.0-1.0-wav}}.
1524: The estimated values of $\alpha_{\rm min}$ and $\alpha_0$
1525: for $n=3\ell$ and $3\ell+1$ with $\mu=1$ are shown in {\bf Table \ref{table;wav-alpha}} .
1526:
1527: For the bicritical point $(\lambda,\mu)=(2.0,1.0)$,
1528: the split of $3\ell$ and $3\ell+1$ is not so obvious in
1529: the numerical data. For example,
1530: {\bf Fig. \ref{fig;alpha-2.0-1.0}} shows
1531: $\alpha_{\rm min}$ for $n=12,13,\cdots$ and they don't
1532: clearly split into two series. Also the convergence
1533: is not as good as those for other points.
1534: The numeric for $(\lambda,\mu)=(2.0,1.0)$ in
1535: {\bf Table \ref{table;wav-alpha}} is obtained based
1536: on both series.
1537:
1538:
1539:
1540:
1541:
1542:
1543:
1544:
1545:
1546:
1547: %///////////////////// a0 curve at lam=1.0 mu=1.0 //////////////
1548: \begin{figure}
1549: \includegraphics[scale=0.6]{alpha0-1.0-1.0.eps}
1550: %\includegraphics[scale=0.5,angle=-90]{falpha0-1.0-1.0.eps}
1551: \caption{ The upper two series are plots of $\alpha_0$ and
1552: the lower two series are plots of $\alpha_{\rm min}$
1553: $(\lambda,\mu)=(1.0,1.0)$. }
1554: \label{fig;alpha0-1.0-1.0}
1555: \end{figure}
1556:
1557:
1558: %///////////////////// a-fa curve at lam=1.0 mu=1.0 //////////////
1559: \begin{figure}
1560: \includegraphics[scale=0.5,angle=-90]{alpha-falpha-1.0-1.0o.eps}
1561: \includegraphics[scale=0.5,angle=-90]{alpha-falpha-1.0-1.0e.eps}
1562: \caption{ $\alpha$-$f(\alpha)$ curve at $(\lambda,\mu)=(1.0,1.0)$,
1563: (a) for index=12,15,18 and 21 and (b) for index=13,16,19 and 22. }
1564: \label{fig;alpha-falpha-1.0-1.0-wav}
1565: \end{figure}
1566:
1567:
1568:
1569:
1570: \setlength{\arrayrulewidth}{0.8pt}
1571: \begin{table}[tb]
1572: \begin{tabular}{ccccc} \hline
1573: index & $\lambda$ & $\mu$ & $\alpha_{\rm min}$ & $\alpha_0 $ \\ \hline
1574: $3\ell$ &0.0 & 1.0 & 0.358 & 1.82 \\
1575: & 0.5 & 1.0 & 0.358 & 1.57 \\
1576: &1.0 & 1.0 & 0.357 & 1.55 \\
1577: &1.5 & 1.0 & 0.357 & 1.54 \\
1578: \hline
1579: $3\ell+1$& 0.0 & 1.0 & 0.358 & 1.65 \\
1580: &0.5 & 1.0 & 0.358 & 1.77 \\
1581: &1.0 & 1.0 & 0.358 & 1.77 \\
1582: &1.5 & 1.0 & 0.358 & 1.73 \\
1583: \hline
1584: &2.0&1.0& 0.47 & 1.4
1585: \end{tabular}
1586: \caption{
1587: $\alpha_{\rm min}$ and $\alpha_0$ on the critical line
1588: $\mu=1.0$.}
1589: \label{table;wav-alpha}
1590: \end{table}
1591:
1592:
1593:
1594: \section{\label{section;conclusions}Conclusions}
1595: We study two dimensional electrons on the
1596: triangular lattice in a uniform magnetic field and
1597: the one dimensional quasiperiodic system obtained from it.
1598: We conjectured a phase diagram of the one dimensional model
1599: as in {\bf Fig. \ref{fig;phase_diagram}}. As a typical example,
1600: we investigated the incommensurate limit of the golden mean
1601: via
1602: the level statistics, namely the distributions of the band widths and the gaps,
1603: and scaling properties of spectra and wavefunctions on
1604: the conjectured critical lines. For level statistics,
1605: we find the characteristic behaviors similar to the ones
1606: previously found for other quasiperiodic models.
1607: We also obtain $\alpha$-$f(\alpha)$ curve for
1608: spectra and the wavefunctions at the centers of the spectra. As for the spectra,
1609: $\alpha$-$f(\alpha)$ curve is the same as one in the Harper model
1610: on the critical lines except for the bicritical point.
1611: For the wavefunctions, we find that
1612: $\alpha_{\rm min}$ is the same as the Harper
1613: model except near the bicritical point. For the line $\lambda=2$,
1614: $\alpha$-$f(\alpha)$ curves are the same as
1615: the Harper model.
1616: For $\mu$=1, the Fibonacci sequence $F_n$ splits into two classes
1617: $n=3\ell$ and $n=3\ell+1$
1618: according to the appearance of zero of the hopping terms.
1619: The dispersion relation is flat for $n=3\ell+1$.
1620: Also the $\alpha$-$f(\alpha)$ curve of the wavefunction is
1621: different for $n=3\ell$ and $n=3\ell+1$.
1622: At the bicritical point where the triangular lattice
1623: symmetry retains, both level statistics and multifractal analysis
1624: show qualitively different behaviors from those of other critical points.
1625:
1626:
1627: {\it Acknowledgement}
1628: K. I. would like to thank Y. Takada for collaboration at the
1629: initial stage of this work.
1630: K.I. was partially supported by
1631: the Grand-in-Aid for Science(B), No.1430114 of JSPS.
1632:
1633:
1634: \begin{thebibliography}{99}
1635:
1636:
1637: \bibitem{hof}
1638: D.R. Hofstadter,
1639: Phys. Rev. B {\bf 14}, 2239 (1976).
1640:
1641: \bibitem{harper}
1642: P.G. Harper
1643: Proc. Phys. Soc. London, Sect. A {\bf 68}, 874 (1955).
1644:
1645: \bibitem{tknn}
1646: D. Thouless, M. Kohmoto, M. P. Nightingale and M. denNijs,
1647: Phys. Rev. Lett. {\bf 49}, 405 (1982).
1648:
1649: \bibitem{kohmoto-chern}
1650: M. Kohmoto, Ann. Phys.(N.Y.) {\bf 160},355(1985).
1651:
1652: \bibitem{hiramoto}
1653: See, for example, H. Hiramoto and M. Kohmoto,
1654: Int. J. Mod. Phys. {\bf B 6} 281(1992).
1655:
1656: \bibitem{aubry}
1657: S. Aubry and G. Andr\'{e},
1658: Ann. Israel Phys. Soc. {\bf 3}, 133 (1980).
1659:
1660: \bibitem{kohmoto-prl}
1661: M. Kohmoto, Phys. Rev. Lett.{\bf 51} 1198(1983).
1662:
1663: \bibitem{hiramoto-kohmoto}
1664: H. Hiramoto and M. Kohmoto
1665: Phys. Rev. {\bf B40}, 8225 (1989).
1666:
1667:
1668: \bibitem{thouless}
1669: D.J. Thouless,
1670: Phys. Rev. B {\bf 28}, 4272 (1983);
1671: D.J. Thouless, Commun. Math. Phys. {\bf 127}, 187 (1990).
1672:
1673: \bibitem{halsey}
1674: T. C. Halsey et al., Phys. Rev. {\bf A33},1141(1986).
1675:
1676: \bibitem{kohmoto}
1677: M. Kohmoto, Phys. Rev. {\bf A37},1345(1988).
1678:
1679:
1680:
1681: \bibitem{evangelou}
1682: S.N. Evangelou and J.-L. Pichard,
1683: Phys. Rev. Lett. {\bf 84}, 1643 (2000).
1684:
1685: \bibitem{takada}
1686: Y. Takada, K. Ino and M. Yamanaka,
1687: Phys. Rev. {\bf E 70}, 066203 (2004).
1688:
1689:
1690: \bibitem{naka} M. Naka, K.Ino and M. Kohmoto,
1691: Phys. Rev. {\bf B 71}, 245120 (2005).
1692:
1693:
1694: \bibitem{altshuler}
1695: B.L. Altshuler, I.Kh. Zharekeshev,
1696: S.A. Kotochigova and B.I. Shklovskii,
1697: Sov. Phys. JETP {\bf 67}, 625 (1988).
1698:
1699:
1700: \bibitem{shapiro}
1701: B.I. Shklovskii, B. Shapiro, B.R. Sears,
1702: P. Lambrianides, and H.B. Shore,
1703: Phys. Rev. B {\bf 47}, 11487 (1993).
1704:
1705:
1706: \bibitem{machigei}
1707: K. Machida and M. Fujita,
1708: Phys. Rev. {\bf B34}, 7367 (1986);
1709: T. Geisel, R. Ketzmerick and G. Petschel,
1710: Phys. Rev. Lett. {\bf 66}, 1651 (1991).
1711:
1712: \bibitem{claro} F.H. Claro and G.H. Wannier,
1713: Phys. Rev. {\bf B19}, 6068(1979).
1714:
1715: \bibitem{han-thouless}
1716: J.H. Han, D.J. Thouless, H. Hiramoto, and M. Kohmoto,
1717: Phys. Rev. B {\bf 50}, 11365 (1994).
1718:
1719: \bibitem{hat-koh}
1720: Y. Hatsugai and M. Kohmoto, Phys. Rev. B {\bf 42}, 8282 (1990).
1721:
1722:
1723:
1724: \bibitem{chamon}
1725: C. Chamon, M. Oshikawa and I. Affleck, Phys. Rev. Lett {\bf 91}, 206403 (2003).
1726:
1727:
1728:
1729:
1730: \end{thebibliography}
1731:
1732:
1733:
1734: \end{document}
1735: