1: %%%%%%%%%%%%%%%%% spd05.tex %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2: %% REVTeX file for the metallic phase in QHE (long) Paper %%%
3: %%%% including scaling analysis %%%%%%%%%%%%%%%%%%%%%%%%%%%%%
4: %
5: %\documentstyle[twocolumn,prl,aps,epsfig,graphicx]{revtex4}
6: \documentclass[twocolumn,aps,prb,showpacs,superscriptaddress,amsfonts,amssymb,amsmath]{revtex4}
7: %\documentclass[aps,prb,preprint,showpacs,superscriptaddress,amsfonts,amssymb,amsmath]{revtex4}
8: \usepackage{amsmath}
9: \usepackage{graphicx}
10: \begin{document}
11: \title{Possible existence of a band of extended states induced by inter-Landau-band mixing in quantum Hall system}
12: \author{Gang Xiong}
13: \affiliation{Physics Department, Beijing Normal University,
14: Beijing 100875, P. R. China}
15: \author{Shi-Dong Wang}
16: \affiliation{Physics Department, Hong Kong University of Science
17: and Technology, Clear Water Bay, Hong Kong SAR, China}
18: \author{Qian Niu}
19: \affiliation{Physics Department, The University of Texas at
20: Austin, Austin, Texas 78712-1081}
21: \affiliation{International Center for Quantum Structures,
22: Institute of Physics, Chinese Academy of Sciences , Beijing
23: 100080, P. R. China}
24: \author{Yupeng Wang}
25: \affiliation{International Center for Quantum Structures,
26: Institute of Physics, Chinese Academy of Sciences , Beijing
27: 100080, P. R. China}
28: \author{X. C. Xie}
29: \affiliation{Physics Department, Oklahoma State University ,
30: Stillwater, OK 74078}
31: \affiliation{International Center for Quantum Structures,
32: Institute of Physics, Chinese Academy of Sciences , Beijing
33: 100080, P. R. China}
34: \author{De-Cheng Tian}
35: \affiliation{Physics Department, Wuhan
36: University, Wuhan 430072, P. R. China}
37: \author{X. R. Wang}
38: \affiliation{Physics Department, Hong Kong
39: University of Science and Technology, Clear Water Bay,
40: Hong Kong SAR, China}
41:
42: \date{Draft on \today}
43:
44: \begin{abstract}
45: The mixing of states with {\it opposite chirality} in quantum Hall
46: system is shown to have delocalization effect. It is possible that
47: extended states may form bands because of this mixing, as is shown
48: through a numerical calculation on a two-channel network model.
49: Based on this result, a new phase diagram with a narrow {\it
50: metallic} phase separating two adjacent QH phases and/or
51: separating a QH phase from the insulating phase is proposed. The
52: data from recent non-scaling experiments is reanalyzed and shown
53: that they seem to be {\it consistent} with the new phase diagram.
54: However, due to finite-size effects, further study on large system
55: size is still needed to conclude whether there are extended state
56: bands in thermodynamic limit.
57: \end{abstract}
58:
59: \pacs{73.40.Hm, 71.30.+h, 73.20.Jc} \maketitle
60:
61: \section{Introduction}
62: According to the scaling theory of localization\cite {abrahams},
63: all electrons in a disordered two-dimensional system are localized
64: in the absence of a magnetic field. In the presence of a strong
65: magnetic field $B$, a series of disorder-broadened Landau bands
66: (LBs) will appear, and an extended state resides at the center of
67: each band while states at other energies are
68: localized\cite{pruisken}. The integrally quantized Hall plateaus
69: (IQHP) are observed when the Fermi level lies in localized states,
70: with the value of the Hall conductance, $\sigma_{xy}=ne^2/h$,
71: related to the number of occupied extended states($n$). Many
72: previous studies\cite{wei,pruisken2,kivelson,dzliu,wang,yang,
73: galstyan,sheng1,sheng2,haldane,jiang,shahar1,tkwang,glozman,kravchenko,
74: song,schaijk,xrw,hilke,baba,shash} have been focused on so-called
75: plateau transitions. The issue there is how the Hall conductance
76: jumps from one quantized value to another when the Fermi level
77: crosses an extended state. There are two competing proposals. One
78: is the global phase diagram\cite{kivelson} based on the levitation
79: of extended states conjectured by Khmelnitskii\cite{khme} and
80: Laughlin\cite{laughlin}. A crucial prediction of this phase
81: diagram is that an integer quantum Hall effect (IQHE) state $n$ in
82: general can only go into another IQHE states $n\pm 1$, and that a
83: transition into an insulating state is allowed only from the $n=1$
84: state. The other is so-called direct transition phase
85: diagram\cite{dzliu} in which transitions from any IQHE state to
86: the insulating phase are allowed when the disorder is increased at
87: fixed $B$. So far, most experiments\cite{kravchenko,song} are
88: consistent with the direct transition phase diagram although the
89: early experiments were interpreted in terms of the global phase
90: diagram.
91:
92: One important yet overlooked issue regarding IQHE is the {\it
93: nature} of both plateau-plateau and plateau-insulator transitions.
94: In all existing theoretical studies, these transitions are assumed
95: to be continuous quantum phase transitions. This assumption is
96: mainly due to the early scaling experiments\cite{wei,pruisken2}.
97: The fingerprint of a continuous phase transition is scaling laws
98: around the transition point. In the case of IQHE, it means
99: algebraic divergence of longitudinal resistance slope and
100: algebraic shrinkage of longitudinal resistance peak width in
101: temperature $T$ at the transition point. However, some experiments
102: show that longitudinal resistance slope remains finite\cite{hilke}
103: and resistance peak width remains nonzero\cite{shash,baba} when
104: extrapolated to zero temperature. This implies a {\it non-scaling}
105: behavior around a transition point, contradicting the expectation
106: of continuous quantum phase transitions suggested by the theories.
107: Thus the nature of these transitions should be re-examined.
108:
109: The samples used in the non-scaling experiments\cite{hilke} are
110: relatively dirty, and strong disorders should lead to a strong
111: inter-Landau-band mixing. In a recent Letter\cite {xiong}, we
112: showed that the single extended state at each LB center broadens
113: into a narrow band of extended states when the interband mixing of
114: {\it opposite chirality} is taken into account. A narrow metallic
115: phase exists between two adjacent IQHE phases and between an IQHE
116: phase and the insulating phase. A plateau-plateau or
117: plateau-insulator transition corresponds to two consecutive
118: quantum phase transitions instead of one as suggested by existing
119: theories. This possibility is usually overlooked in previous
120: studies where each longitudinal conductance peak is related to a
121: single extended state\cite{glozman}. In this paper we shall
122: present the detailed description of this study.
123:
124: The paper is organized as follows. The semiclassical network model
125: for two coupled LBs is illustrated in Sec. II. It is shown that
126: mixing of localized states of {\it opposite chirality} tends to
127: delocalize a state while mixing of states of the same chirality
128: does not. The similarities and differences between the case we
129: consider and models for double layer systems and spin-degenerate
130: systems are also discussed. Our approach, level-statistics
131: technique, is described in Sec. III. Sec. IV has four subsections
132: containing our numerical results, conclusions and discussions.
133: Sec. IV A provides our main numerical results and conclusions. In
134: Sec. IV B, possible finite size effects and a theoretical
135: understanding for non-scaling behaviors based on standard scaling
136: theory\cite{pruisken3,huckstein} are discussed. Sec. IV C gives
137: further numerical results which directly show a narrow band of
138: extended states in several cases, and a new phase diagram of
139: quantum Hall systems is proposed. To further support our results,
140: in Sec. IV D we reanalyze the original data from the non-scaling
141: experiments\cite{hilke} and show that the non-scaling behaviors in
142: each IQHP-insulator transition can be attributed to two quantum
143: phase transition points. Recent
144: experiments\cite{hohls1,hohls2,hohls3,ponomarenko,li}, in which
145: scaling behaviors are observed, are also discussed and shown to be
146: not inconsistent with early non-scaling experiments and our
147: numerical results. The conclusions of this paper are summarized in
148: Sec. V.
149:
150: \section{The semiclassical model including
151: inter-Landau-band mixing}
152: According to the semiclassical
153: theory\cite{chalker}, the motion of an electron in a strong
154: magnetic field and in a smooth random potential can be decomposed
155: into a rapid cyclotron motion and a slow drifting motion of its
156: guiding center. The kinetic energy of the cyclotron motion is
157: quantized by $E_n=(n+1/2)\hbar\omega_c$, where $\omega_c$ is the
158: cyclotron frequency and $n$ the LB index. The trajectory of the
159: drifting motion of the guiding center is thus along an
160: equipotential contour of value $V_0=E-E_n$, where $E$ is the total
161: energy of the electron. The local drifting velocity
162: $\vec{v}(\vec{r})$ is determined by (in SI unit)
163: \begin{equation}
164: \vec{v}(\vec{r})=\bigtriangledown V
165: (\vec{r})\times\vec{B}/(eB^2)
166: \label{drift}
167: \end{equation}
168: where $\bigtriangledown V(\vec{r})$ is the local potential
169: gradient. The equipotential contour consists of many closed loops.
170: Neglect quantum tunnelling effects, each loop corresponds to
171: trajectory of one eigenstate. The motion of electrons are thus
172: confined around these loops with deviations typically order of the
173: cyclotron radius $l_c=\sqrt{\hbar/(eB)}$.
174: \begin{figure}[ht]
175: \begin{center}
176: % \includegraphics[height=5cm, width=9cm]{fig/lfig1.eps}
177: \includegraphics[height=6cm, width=6cm]{lfig1a.eps}
178: \includegraphics[height=5.5cm, width=4cm]{lfig1b.eps}
179: \end{center}
180: \caption{(a) Four neighboring loops in a one-band model for the
181: case of $V_0<0$. Dashed lines denote quantum tunnellings. The
182: arrows indicate the drifting direction. (b) Loops localized around
183: a valley and a peak, respectively. The arrows inside a loop show
184: the directions of local potential gradient around the peak or
185: valley. The arrows on a loop indicate the drifting direction.}
186: \label{1channel}
187: \end{figure}
188:
189: To illustrate this semiclassical picture, let us think of the
190: smooth random potential as a landscape of many peaks and valleys
191: distributed randomly in the plane. Imagine that the landscape is
192: filled with water up to a height of value $V_0$.
193: The equipotential contour of value $V_0$ is the boundary of
194: land and water. According to the percolation
195: theory\cite{stauff}, the percolation threshold of a
196: two-dimensional (2D) continuum model is $p_c=1/2$, where $p_c$ is
197: the occupation probability of the medium (the land or the water).
198: For simplicity, we suppose that the distribution of the random
199: potential is symmetric around zero. By symmetry the percolation
200: point of both the land and the water is at $V_0=0$ in this case.
201: When $V_0<0$, the occupation probability of land is above $1/2$.
202: Thus the land percolates and the water forms isolated lakes. These
203: lakes are around valleys and their boundaries correspond to
204: trajectories of localized states. In the case of $V_0>0$, the
205: water forms a percolating sea and the land becomes isolated
206: islands around potential peaks. The boundary of each island is an
207: electronic state. In short, semiclassical electronic states in a
208: QH system are equipotential loops. These loops are localized
209: around potential peaks for $V_0>0$ and around potential valleys
210: for $V_0<0$. The drifting direction of each loop is {\it
211: unidirectional}. This means that they are chiral states. From Eq.
212: \ref{drift} one can see that states around peaks have {\it
213: opposite chirality} from states around valleys because the
214: directions of the local potential gradient around a peak is
215: opposite to that around a valley. If one views the plane from the
216: direction opposite to the magnetic field, the drifting is
217: clockwise around valleys and counter-clockwise around peaks, as
218: shown in Fig.\ref{1channel}. Right at $V_0=0$ both the land and
219: the water percolate, and the intersection between them is the
220: trajectory of an extended state. It means that there is only one
221: extended state at $V_0=0$ for each LB. As $V_0$ approaches zero
222: from both sides, the localization length $\xi$ of the system tends
223: to diverge as
224: \begin{equation}
225: \xi\propto |V_0|^{-\nu}
226: \label{ll}
227: \end{equation}
228: where the critical exponent $\nu=4/3$ according to the classical
229: percolation theory. Quantum effects are ignored in the above
230: semiclassical argument. When two spatially separated loops on the
231: same equipotential contour come close at saddle points of the
232: random potential, quantum tunnellings should be considered. An
233: example in the case of $V_0<0$ is shown in Fig.\ref{1channel}(a).
234: In the absence of interband mixing, numerical calculations have
235: suggested that there is still only one extended state in each LB
236: while the value of the critical exponent $\nu$ is modified to be
237: around $7/3$\cite{galstyan}.
238: \begin{figure}[ht]
239: \begin{center}
240: % \includegraphics[height=7cm,width=8cm]{fig/lfig2.eps}
241: \includegraphics[height=7cm,width=8cm]{lfig2.eps}
242: \end{center}
243: \caption{(a) Two adjacent Landau bands in the case when the
244: disorder broadened band width is comparable with the Landau gap.
245: $D(E)$ is the density of states. $E_{\it u}$ and $E_{\it l}$
246: denote the centers of the two bands. (b) Schematic plot of the two
247: sets of equipotential loops on the two-dimensional random
248: potential for electronic states of energy $E$ shown in (a). It is
249: the projection of a three-dimensional landscape on a 2D vertical
250: plane. The solid curve represents the schematic plot of the 2D
251: random potential. Two dashed horizontal lines indicate two
252: constant potential planes $V(\vec{r})\equiv V_{0}$, one for the
253: lower band with $V_0= V_{\it l}=E-E_{\it u}>0$ and the other for
254: the upper band with $V_0=V_{\it u}=E-E_{\it l}<0$. The ellipses
255: denote the loops where the two constant potential planes intersect
256: with the 2D random potential. Arrows on the loops show drifting
257: directions. $\vec{B}$ is the magnetic field.} \label{twoband}
258: \end{figure}
259:
260: % 2. The model including interband mixing
261: In the case when the width of the LBs is comparable with the
262: spacing between adjacent LBs (the Landau gap), and
263: inter-Landau-band mixing should no longer be ignored. In order to
264: investigate the consequences of inter-Landau-band mixing, we shall
265: consider a simple system of two adjacent LBs. Since we are
266: interested in interband-mixing of opposite chirality, we consider
267: those states with energy between the lower and the upper bands
268: whose centers are at $E_l$ and $E_u$, respectively, as shown in
269: Fig.\ref{twoband}(a). Thus, equipotential loops are $V_l=E-E_l>0$
270: and $V_u=E-E_u<0$ for the lower and the upper LBs, respectively.
271: Using the semiclassical theory described in the previous
272: paragraphs, states from the upper band should move along
273: equipotential loops around potential valleys while those from the
274: lower band around potential peaks as shown in
275: Fig.\ref{twoband}(b). The loops for the upper band drift in
276: clockwise direction, and those for the lower band in the
277: counter-clockwise direction. These two sets of loops are thus
278: spatially separated and have {\it opposite chirality}. If we
279: assume that peaks and valleys of random potential form two coupled
280: square lattices, the loops can be arranged as shown in Fig.
281: \ref{network}(a), where {\it P} and {\it V} denote peaks and
282: valleys, respectively. In the absence of interband mixing, the
283: model is reduced to two decoupled single-band models and all
284: electronic states between the two LBs are localized. If we
285: introduce interband mixing, the localized loops may become less
286: localized. To see this, let us consider an extreme case with no
287: tunnellings at saddle points, but with such a strong interband
288: mixing that an electron will move from a loop around a valley to
289: its neighboring loop around a peak and vice versa, as shown by
290: $B\to C$ in Fig.\ref{network}(a). Follow the trajectory of an
291: electron starting at A, it will be $A\to B\to C\to D\to E \cdot
292: \cdot \cdot$. The electron is no longer confined on a closed loop,
293: but delocalized!
294:
295: Before going on, we would like to make a short discussion on the
296: relation between large disorder magnitude and weak magnetic field.
297: The above semiclassical picture for the network model is valid
298: only when the typical fluctuation length of random potential,
299: denoted by $L_F$, is much larger than the magnetic length
300: $l_c=\sqrt{\hbar/(eB)}$. Let us take $L_F$ as a large fixed value.
301: Then, the effect of inter-band mixing is controlled mainly by two
302: parameters. One is the ratio of disorder magnitude $W$ and Landau
303: gap $\hbar\omega_c$, which is proportional to $W/B$ and determines
304: whether inter-band mixing needs to be considered. The other is the
305: ratio of the magnetic length $l_c$ and the typical distance
306: between an upper-band loop and its neighboring lower-band loop
307: $L_F\frac{\hbar\omega_c}{W}$, which is proportional to $W/B^{3/2}$
308: and determines the typical strength of inter-band mixing in the
309: network model. Therefore, the cases of strong disorder $W$ and
310: weak magnetic field $B$ are, to some extent, equivalent in the
311: network model since both of them lead to large inter-band mixing.
312: Of cause, values of local potential gradients are also important,
313: since they determine the detailed distribution of inter-band
314: mixing in the model.
315:
316: %3. Difference between one-band and two-band models
317: In the one-band model, an electron can also hop from one loop to
318: its neighboring loops by quantum tunnellings. At a first glance,
319: this effect seems similar to that of interband mixing. However,
320: they are fundamentally different. In the one-band model,
321: electronic states for a given $V_0$ have {\it the same chirality}.
322: Thus the drifting direction of an electron will be inverted when
323: it tunnels into neighboring loops. This means that a strong
324: tunnelling in a one-band model will induce an effective
325: backward-scattering that tends to also localize electrons. We can
326: understand this by considering a small part of the one-band model
327: as shown in Fig.\ref{1channel}(a) where all loops are moving in
328: clockwise direction. Without tunnelling, the trajectory of an
329: electron starting from point A is $A\to B\to I \to J\to A$, a
330: clockwise closed loop. With strong tunnellings, the trajectory
331: will tend to be $A\to B\to C\to D \cdot \cdot \cdot \to H \to A$,
332: a counter-clockwise closed loop. Thus, the tunnellings between
333: loops of the same chirality cannot delocalize states.
334: \begin{figure}[h]
335: \begin{center}
336: \includegraphics[height=7.5cm, width=7.5cm]{lfig3a.eps}
337: \includegraphics[height=7.5cm, width=7.5cm]{lfig3b.eps}
338: % \includegraphics[height=4.5cm, width=4cm]{fig/lfig3a.eps}
339: % \includegraphics[height=4.5cm, width=4cm]{fig/lfig3b.eps}
340: \end{center}
341: \caption{(a) Topological plot of the trajectories of the drifting
342: motion of guiding centers (rhombus). The drifting motion around a
343: potential peak (valley) is denoted by P (V), and their directions
344: are indicated by the arrows. Dashed lines stand for interband
345: mixing, and dash-dotted lines for tunnelling at saddle points. The
346: thick line (A to I) describes the trajectory of an electron due to
347: a strong interband mixing. (b) The equivalent two-channel network
348: model of (a). Solid and dashed lines on each link denote two
349: channels from two LBs. Squares stand for saddle points. P, V and
350: arrows have the same meaning as those in (a).} \label{network}
351: \end{figure}
352:
353: Two-channel CC model can also be used to simulate both
354: spin-degenerated and bilayer systems. The model for
355: spin-degenerate system is the same as we use. The only difference
356: is that loops for spin-up and spin-down states present at the same
357: positions in real space, and the chiralities of loops of spin-up
358: and spin-down states are the same. While the main difference
359: between the model for bilayer system and ours is that the real
360: space distributions of random potentials in the two layers are
361: different.
362:
363: %4. The reason why only states between LBs are considered.
364: It is worthwhile to explain why we consider only those states
365: between two LB centers. For states outside this range, both sets
366: of loops are localized around either valleys or peaks. This means
367: that interband mixing mainly occurs between two loops localized
368: around the same position, and this mixing will not delocalize a
369: state. In fact, as explained in the previous paragraph, the mixing
370: of the same chirality does not help delocalizing an electron. This
371: is why we shall consider mixing between spatially separated states
372: with opposite chirality. Of course, it does not mean that the
373: mixing of the same chirality has no effect at all. As it was found
374: in some previous works\cite{wang}, this kind of mixing may shift
375: an extended state from its LB center. Level shifting due to mixing
376: between states of the same chirality may distort the shape of the
377: phase diagram, but should not alter its topology. The emergence of
378: the bands of extended states is exclusively due to the mixing
379: between states of opposite chirality.
380:
381:
382: Now, we describe our two-channel network model in detail. Assume
383: that tunnellings of two neighboring localized states (loops) of
384: the same band occur around saddle points, and interband mixing
385: takes place only on the links, Fig.\ref{network}(a) is
386: topologically equivalent to the model shown in
387: Fig.\ref{network}(b). Fig.\ref{network}(b) is the schematic
388: illustration of our {\it two-channel Chalker-Coddington} network
389: model. It is similar to the model studied in previous
390: publications\cite{lee, kagalovsky}. There are two channels on each
391: link. One, denoted by a solid line, is from the lower LB around a
392: potential peak. The other (dashed line) is from the upper LB
393: moving around a potential valley. The arrows indicate the drifting
394: direction of the two sets of states. At each node, the tunnelling
395: between two neighboring states of the same LB occurs. As shown in
396: Fig.\ref{nodelink}(a), let $Z_{u(l)}^{in,1}$ and $Z_{u(l)}^{in,2}$
397: be the incoming wave amplitudes of states 1 and 2 from upper
398: (lower) LB, respectively, and $Z_{u(l)}^{out,1}$ and
399: $Z_{u(l)}^{out,2}$ be the outgoing wave amplitudes of the two
400: states. The tunnelling is described by an SO(4) matrix
401: \begin{equation}
402: \label{smatrix}
403: \left (
404: \begin{array}{l}
405: Z_u^{out,1}\\
406: Z_u^{out,2}\\
407: Z_l^{out,1}\\
408: Z_l^{out,2}
409: \end{array}
410: \right )
411: =
412: \left (
413: \begin{array}{llll}
414: s_u^R & s_u^L & 0 & 0 \\
415: -s_u^L & s_u^R & 0 & 0 \\
416: 0 & 0 & s_l^R & s_l^L \\
417: 0 & 0 & -s_l^L & s_l^R
418: \end{array}
419: \right )
420: \left (
421: \begin{array}{l}
422: Z_u^{in,1}\\Z_u^{in,2}\\Z_l^{in,1}\\Z_l^{in,2}
423: \end{array}
424: \right ) ,
425: \end{equation}
426: where the subscripts $u$ and $l$ denote the upper and the lower
427: bands, respectively. The elements $s_{u(l)}^L$ and $s_{u(l)}^R$
428: are tunnelling coefficients of an incoming wave-function in the
429: upper (lower) band being scattered into outgoing channels at its
430: left-hand and right-hand sides, respectively. $s_{u(l)}^R$ and
431: $s_{u(l)}^L$ are related to each other as $s_{u(l)}^R=\sqrt
432: {1-(s_{u(l)}^L)^2}$ due to the orthogonality of the matrix. Under
433: quadratic potential barrier approximation, ---i.e.,
434: $V(x,y)=-Ux^2+Uy^2+V_c$ around a saddle point, where $U$ is a
435: constant describing the strength of potential fluctuation and
436: $V_c$ is the potential barrier at the point,--- one can show that
437: the left-hand scattering amplitude is given by\cite{fertig1}
438: \begin{equation}
439: s_{u(l)}^L=[1+\exp(-\pi\epsilon_{u(l)})]^{-1/2},
440: \end{equation}
441: where $\epsilon_{u(l)}=[E+V_c-(n_{u(l)}+1/2)E_2]/E_1$ with $E$
442: being electron energy, $E_1=\frac{\hbar\omega_c}{2\sqrt{2}}
443: \sqrt{K-1}$, $E_2=\frac{\hbar\omega_c}{\sqrt{2}}\sqrt{K+1}$,
444: and $K=\sqrt{\frac{64U^2}{m^2\omega_c^4}+1}$. The energies of
445: the cyclotron motion in the two bands are $(n_u+1/2)E_2$ and
446: $(n_l+1/2)E_2$, respectively, where $n_{u(l)}$ are the band
447: indices and $\Delta n=n_u-n_l=1$. The dimensionless ratio
448: $E_r=E_2/E_1=2\sqrt {1+\frac{2}{K-1}}$ approaches $2$ from above
449: as $U$ or the inverse of $\omega_c$ increases\cite{fertig1},
450: i.e., strong disorder regime or a weak magnetic field.
451: In our calculations, we choose it to be 2.2 since this is
452: the regime we are interested in. For convenience, we choose
453: $E_2$ as the energy unit and the cyclotron energy of the lower
454: band as the reference point. The energy regime between the two
455: band centers is thus $E\in[0,1]$.
456:
457: Inter-band mixing between two channels on a link as shown in
458: Fig.\ref{nodelink}(b) is described by a U(2) matrix
459: \begin{equation}
460: \label{mmatrix1}
461: \left (
462: \begin{array}{l} Z^{out}_l\\Z^{out}_u \end{array}
463: \right )=M
464: \left (
465: \begin{array}{l} Z^{in}_l\\Z^{in}_u \end{array}
466: \right ),
467: \end{equation}
468: \begin{equation}
469: \label{mmatrix2}
470: M=\left (
471: \begin{array}{ll} e^{i\phi_1} & 0 \\ 0 & e^{i\phi_2}
472: \end{array}
473: \right )
474: \left (
475: \begin{array}{ll} \cos\theta & \sin\theta \\ -
476: \sin\theta & \cos\theta \end{array}
477: \right )
478: \left (
479: \begin{array}{ll} e^{i\phi_3} & 0 \\ 0 & e^{i\phi_4}
480: \end{array}
481: \right ),
482: \end{equation}
483: where $\sin\theta$ describes the interband mixing. $\phi_i(i=1\sim
484: 4)$ are random Aharonov-Bohm phases accumulated along propagation
485: paths. In our calculations, we shall assume that they are
486: uniformly distributed in $[0,2\pi]$\cite{chalker}. In the
487: following discussion, a parameter $J$, defined as $\sqrt{J/(1+
488: J)}=\sin\theta$, is used to characterize the mixing strength.
489: $J$ will take the same value for all links in our calculations.
490: We hope that this simplification will not affect the physics.
491:
492: \begin{figure}[ht]
493: \begin{center}
494: % \includegraphics[height=5.0cm,width=8cm]{fig/lfig4.eps}
495: \includegraphics[height=5.0cm,width=8cm]{lfig4.eps}
496: \end{center}
497: \caption{(a) A node with four incoming and four outgoing channels.
498: $Z_{u(l)}^{in,i}$ is the wavefunction amplitude of the i$^{th}$
499: incoming wave from the upper (lower) LB. $Z_{u(l)}^{out,i}$ is
500: that of outgoing wavefunction amplitude. (b) A link with two
501: channels. $Z_{u(l)}^{in (out)}$ is the incoming (outgoing)
502: wavefunction amplitude of the upper (lower) LB.} \label{nodelink}
503: \end{figure}
504:
505:
506:
507: \section{The application of level-statistics technique
508: on the network model}
509:
510: Electron localization length is often obtained from the transfer
511: matrix method. For a two-dimensional system, however, it is well
512: known that this quantity alone does not provide conclusive answers
513: to questions related to the metal-insulator transition
514: (MIT)\cite{xie}. On the other hand, level-statistics analysis has
515: been used in studying MIT\cite{kramer,schweitzer}.
516: Level-statistics analysis is based on random matrix theory
517: (RMT)\cite{mehta}. The basic idea is that the localization
518: property of an electronic state can be determined by the
519: statistical distribution function $P(s)$ of the spacing $s$ of two
520: neighboring levels. For localized states, the distribution is
521: Poissonian $P_{PE}(s)=exp(-s)$, called `Poissonian ensemble (PE)'.
522: In the case of extended states, the nearest neighbor level spacing
523: distribution has the following form\cite{mehta}
524: \begin{equation}
525: P(s)=C_1(\beta)s^{\beta}exp[-C_2(\beta)s^2]
526: \end{equation}
527: where $C_1(\beta)$ and $C_2(\beta)$ are normalization factors
528: determined by $\int P(s)ds=1$ and $\int sP(s)ds=1$. The parameter
529: $\beta$ is determined by the dynamical symmetry of the system. The
530: case of $\beta=1$ is for systems with time-reversal symmetry and
531: an integer total angular momentum and is referred as `Gaussian
532: orthogonal ensemble'. Systems with time-reversal symmetry and a
533: half-integer total angular momentum belong to the case of
534: $\beta=4$, called `Gaussian symplectic ensemble'. For systems
535: without time-reversal symmetry $\beta=2$, and it is called
536: `Gaussian unitary ensemble (GUE)'. A fundamental difference
537: between level statistics property of extended states and localized
538: states is that $P(s)$ at $s=0$ is zero for extended states and one
539: for localized states. The physical reason of this difference is
540: the so-called `level repulsion' of extended states. Two extended
541: states with the same `bare energy' will overlap in real
542: space(since they are extended in real space) and form two new
543: extended states with different energies. While localized states
544: can have the same energy staying in different regions of real
545: space. This approach is appropriate for the network model of
546: quantum Hall systems because localized states in the model are
547: loops at different regions of real space while extended states in
548: the model are formed by quantum percolation of such localized
549: loops.
550:
551: We shall follow the approach proposed by Klesse and
552: Metzler\cite{klesse}. A quantum state of a network model can be
553: expressed by a vector whose components are electronic
554: wave-function amplitudes on the links. In our case, the vector can
555: be written as $\Phi=(\{Z_u^i,Z_l^i\})$, where $Z_u^i$ and $Z_l^i$
556: are the electron wave-function amplitudes of the upper band ({\it
557: u}) and the lower band ({\it l}) on the $i$-th link, respectively.
558: As shown by Fertig\cite{fertig2}, the network model can be
559: described by an {\it evolution operator} $\hat{U}(E)$, an
560: $E$-dependent matrix determined by the scattering properties of
561: nodes and links in the model. (As an example, the evolution
562: operator of a two-channel network of size ${\it L}=2$ with
563: periodic boundaries on both directions is constructed explicitly
564: in the Appendix.) In general, the eigenvalue equation of the
565: evolution operator is
566: \begin{equation}
567: \hat{U}(E)\Phi_{\alpha}(E)=
568: e^{i\omega_{\alpha}(E)}\Phi_{\alpha}(E),
569: \end{equation}
570: where $\alpha$ is the eigenstate index of $\hat U$. The true
571: eigenenergies $\{E_n\}$ of the system are those energies at which
572: $\omega_\alpha(E)$ is an integer multiple of $2\pi$. It has been
573: shown by Klesse and Metzler\cite{klesse} that the set of {\it
574: quasi-energies} $\{\omega_{\alpha}(E)\}$ corresponds to the
575: excitation spectrum of the stationary state with energy $E$.
576: Therefore, the statistics property of the set of {\it
577: quasi-energies} $\{\omega_{\alpha}(E)\}$ at $E$ is the same as
578: that of true eigenenergies $\{E_n\}$ around $E$, and the
579: localization property of an electronic state with an energy $E$
580: can be obtained by the quasi-energies. The advantage of this
581: approach is that all the quasi-energies can be used in the
582: analysis so that better statistics can be obtained.
583:
584: Chalker and Coddington\cite{chalker} showed numerically that an
585: open boundary condition along one direction creates extended edge
586: states along the other direction. In order to get rid of the edge
587: states, we employ a periodic boundary along both directions in our
588: calculation. For a two-channel network model of ${\it L}\times
589: {\it L}$ nodes with periodic boundaries along both directions,
590: there are $4{\it L}^2$ components in $\Phi$. $\hat{U}$ is thus a
591: $(4{\it L}^2)\times (4{\it L}^{2})$ matrix. However, there is a
592: special property of the network model\cite{metzler}: the nodes
593: scatter electrons only from vertical channels into horizontal
594: channels and {\it vice versa}. If one separates $\Phi$ into the
595: set of wavefunction amplitudes on the horizontal links $\Phi_H$
596: and the set of wavefunction amplitudes on the vertical links
597: $\Phi_V$, the evolution equation in one time step can be written
598: in the following form
599: \begin{equation}
600: \label{evolution}
601: \left (
602: \begin{array}{l}
603: \Phi_H(t+1) \\ \Phi_V(t+1)
604: \end{array}
605: \right )
606: =
607: \left (
608: \begin{array}{ll}
609: \hat{0} & \hat{U}_{V\to H} \\
610: \hat{U}_{H\to V} & \hat{0}
611: \end{array}
612: \right )
613: \left (
614: \begin{array}{l}
615: \Phi_H(t) \\ \Phi_V(t)
616: \end{array}
617: \right ),
618: \end{equation}
619: where $\hat{0}$ is the $(2{\it L}^2)\times (2{\it L}^{2})$ zero
620: matrix. $\hat{U}_{V\to H}$ describes how wavefunction on vertical
621: links evolves into that on the horizontal links. Similarly,
622: $\hat{U} _{H\to V}$ describes that from horizontal to vertical
623: links. For the detail derivation, we refer readers to the example
624: shown in the Appendix. The evolution equation in two time steps is
625: given as
626: \begin{eqnarray}
627: \Phi_H(t+2)=
628: \hat{U}_{V\to H}\hat{U}_{H\to V}\Phi_H(t) \\
629: \Phi_V(t+2)=
630: \hat{U}_{H\to V}\hat{U}_{V\to H}\Phi_V(t)
631: \end{eqnarray}
632: Therefore, the evolution matrix in two time steps is
633: block-diagonal and the two blocks have essentially the same
634: statistical property. We thus need only consider one of them.
635:
636: We study the model for {\it L}=8, 12, 16, 20 and 24. The
637: calculation procedure is as follows. Take a realization of the
638: random phases, construct the evolution matrix and obtain the
639: quasi-energies $\{\omega_{i}\}$. Put them in descending order and
640: calculate the level spacings $s_i=(\omega_i-\omega_{i-1})
641: /\delta$, where $\delta$ is the average of ${s_i}$. Repeat this
642: procedure for sufficient times so that more than $5\times10^4$
643: level spacings are collected for a given $E$ and $J$. The
644: level-spacing distribution function $P(s)$ is thus obtained
645: numerically.
646:
647: \section{Numerical results and discussions}
648: %IV. Results, discussions and conclusions
649: % 1. Numerical results
650: %(1) P(s) shown in global shape
651: \subsection{Analysis of the level-spacing spectrum}
652:
653: We shall analyze the numerical results of the level-spacing
654: distribution function $P(s)$ in order to provide evidences for the
655: existence of extended state bands in our model. Due to the
656: chirality nature of the drifting motion, time-reversal symmetry is
657: absent from our semiclassical network model. Then, according to
658: the RMT\cite{mehta}, $P(s)$ should be the GUE distribution,
659: $P_{GUE}(s)=32\pi^{-2}s^2exp(-4s^2/\pi)$, for extended states, and
660: the PE distribution $P_{PE}(s)$ for localized states. Since the
661: overall shape of the GUE distribution is quite different from that
662: of the PE distribution, one may use $P(s)$ to distinguish an
663: extended state from a localized state. Fig.\ref{ps_global1} is
664: $P_{GUE}(s)$ and $P(s)$ for $(E=0,J=0.1)$ (a), $(E=0.02, J=0.1)$
665: (b), and $(E=0.5,J=1.5)$ (c) with ${\it L}=8, 12,16,20,24$.
666: Fig.\ref{ps_global2} is $P(s)$ for $(E=0.0,J=0.7)$ (a),
667: $(E=0.02,J=0.7)$ (b) and $(E=0.5,J=0.5)$ (c).
668: \begin{figure}[ht]
669: \begin{center}
670: % \includegraphics[width=8cm,height=4.9cm]{fig/lfig5a.eps}
671: \includegraphics[width=8cm,height=4.9cm]{lfig5a.eps}
672: \end{center}
673: \begin{center}
674: \includegraphics[width=8cm,height=4.9cm]{lfig5b.eps}
675: % \includegraphics[width=8cm,height=4.9cm]{fig/lfig5b.eps}
676: \end{center}
677: \begin{center}
678: \includegraphics[width=8cm,height=4.9cm]{lfig5c.eps}
679: % \includegraphics[width=8cm,height=4.9cm]{fig/lfig5c.eps}
680: \end{center}
681: \caption{$P(s)$ vs. $s$ for ${\it L}=8,24$ and $P_{GUE}(s)$. (a)
682: $E=0$ and $J=0.1$; (b) $E=0.02$ and $J=0.1$; (c) $E=0.5$ and
683: $J=1.5$.} \label{ps_global1}
684: \end{figure}
685: \begin{figure}[ht]
686: \begin{center}
687: % \includegraphics[width=8cm,height=4.9cm]{fig/lfig6a.eps}
688: \includegraphics[width=8cm,height=4.9cm]{lfig6a.eps}
689: \end{center}
690: \begin{center}
691: % \includegraphics[width=8cm,height=4.9cm]{fig/lfig6b.eps}
692: \includegraphics[width=8cm,height=4.9cm]{lfig6b.eps}
693: \end{center}
694: \begin{center}
695: % \includegraphics[width=8cm,height=4.9cm]{fig/lfig6c.eps}
696: \includegraphics[width=8cm,height=4.9cm]{lfig6c.eps}
697: \end{center}
698: \caption{$P(s)$ vs. $s$ for ${\it L}=8,24$ and $P_{GUE}(s)$. (a)
699: $E=0$ and $J=0.7$; (b) $E=0.02$ and $J=0.7$; (c) $E=0.5$ and
700: $J=0.5$.} \label{ps_global2}
701: \end{figure}
702: \begin{figure}[ht]
703: \begin{center}
704: % \includegraphics[width=8cm,height=4.9cm]{fig/lfig7a.eps}
705: \includegraphics[width=8cm,height=4.9cm]{lfig7a.eps}
706: \end{center}
707: \begin{center}
708: % \includegraphics[width=8cm,height=4.9cm]{fig/lfig7b.eps}
709: \includegraphics[width=8cm,height=4.9cm]{lfig7b.eps}
710: \end{center}
711: \begin{center}
712: % \includegraphics[width=8cm,height=4.9cm]{fig/lfig7c.eps}
713: \includegraphics[width=8cm,height=4.9cm]{lfig7c.eps}
714: \end{center}
715: \caption{$I_0$ vs. $J$ for ${\it L}=8, 12, 16, 20, 24$, (a) $E=0$;
716: (b) $E=0.02$; (c) $E=0.5$.} \label{data}
717: \end{figure}
718: \begin{figure}[ht]
719: \begin{center}
720: % \includegraphics[width=8cm,height=4.9cm]{fig/lfig8a.eps}
721: \includegraphics[width=8cm,height=4.9cm]{lfig8a.eps}
722: \end{center}
723: \begin{center}
724: % \includegraphics[width=8cm,height=4.9cm]{fig/lfig8b.eps}
725: \includegraphics[width=8cm,height=4.9cm]{lfig8b.eps}
726: \end{center}
727: \begin{center}
728: % \includegraphics[width=8cm,height=4.9cm]{fig/lfig8c.eps}
729: \includegraphics[width=8cm,height=4.9cm]{lfig8c.eps}
730: \end{center}
731: \caption{$I_P(s)$ vs. $s$ for ${\it L}=8, 12, 16, 20, 24$ and that
732: for $P_{GUE}(s)$. (a) $E=0$ and $J=0.1$; (b) $E=0.02$ and $J=0.1$;
733: (c) $E=0.5$ and $J=1.5$.} \label{ps_small1}
734: \end{figure}
735: \begin{figure}[ht]
736: \begin{center}
737: % \includegraphics[width=8cm,height=4.9cm]{fig/lfig9a.eps}
738: \includegraphics[width=8cm,height=4.9cm]{lfig9a.eps}
739: \end{center}
740: \begin{center}
741: % \includegraphics[width=8cm,height=4.9cm]{fig/lfig9b.eps}
742: \includegraphics[width=8cm,height=4.9cm]{lfig9b.eps}
743: \end{center}
744: \begin{center}
745: % \includegraphics[width=8cm,height=4.9cm]{fig/lfig9c.eps}
746: \includegraphics[width=8cm,height=4.9cm]{lfig9c.eps}
747: \end{center}
748: \caption{$I_P(s)$ vs. $s$ for ${\it L}=8, 12, 16, 20, 24$ and that
749: for $P_{GUE}(s)$, (a) $E=0$ and $J=0.7$; (b) $E=0.02$ and $J=0.7$;
750: (c) $E=0.5$ and $J=0.5$.} \label{ps_small2}
751: \end{figure}
752: \begin{figure}[ht]
753: \begin{center}
754: % \includegraphics[width=8cm,height=4.9cm]{fig/lfig10a.eps}
755: \includegraphics[width=8cm,height=4.9cm]{lfig10a.eps}
756: \end{center}
757: \begin{center}
758: % \includegraphics[width=8cm,height=4.9cm]{fig/lfig10b.eps}
759: \includegraphics[width=8cm,height=4.9cm]{lfig10b.eps}
760: \end{center}
761: \begin{center}
762: % \includegraphics[width=8cm,height=4.9cm]{fig/lfig10c.eps}
763: \includegraphics[width=8cm,height=4.9cm]{lfig10c.eps}
764: \end{center}
765: \caption{$I_P(s=0.5,J)$ vs. $J$ for ${\it L}=8, 12, 16, 20, 24$,
766: (a) $E=0$; (b) $E=0.02$; (c) $E=0.5$.} \label{ps_small3}
767: \end{figure}
768: \begin{figure}[ht]
769: \begin{center}
770: % \includegraphics[width=8cm,height=4.9cm]{fig/lfig11a.eps}
771: \includegraphics[width=8cm,height=4.9cm]{lfig11a.eps}
772: \end{center}
773: \begin{center}
774: % \includegraphics[width=8cm,height=4.9cm]{fig/lfig11b.eps}
775: \includegraphics[width=8cm,height=4.9cm]{lfig11b.eps}
776: \end{center}
777: \begin{center}
778: % \includegraphics[width=8cm,height=4.9cm]{fig/lfig11c.eps}
779: \includegraphics[width=8cm,height=4.9cm]{lfig11c.eps}
780: \end{center}
781: \caption{$F(s)$ vs. $s$ for ${\it L}=8, 12, 16, 20, 24$, and that
782: for $P_{GUE}(s)$. (a) $E=0$ and $J=0.1$; (b) $E=0.02$ and $J=0.1$;
783: (c) $E=0.5$ and $J=1.5$.} \label{ps_tail1}
784: \end{figure}
785: \begin{figure}[ht]
786: \begin{center}
787: % \includegraphics[width=8cm,height=4.9cm]{fig/lfig12a.eps}
788: \includegraphics[width=8cm,height=4.9cm]{lfig12a.eps}
789: \end{center}
790: \begin{center}
791: % \includegraphics[width=8cm,height=4.9cm]{fig/lfig12b.eps}
792: \includegraphics[width=8cm,height=4.9cm]{lfig12b.eps}
793: \end{center}
794: \begin{center}
795: % \includegraphics[width=8cm,height=4.9cm]{fig/lfig12c.eps}
796: \includegraphics[width=8cm,height=4.9cm]{lfig12c.eps}
797: \end{center}
798: \caption{$F(s)$ vs. $s$ for ${\it L}=8, 12, 16, 20, 24$ and that
799: for $P_{GUE}(s)$. (a) $E=0$ and $J=0.7$; (b) $E=0.02$ and $J=0.7$;
800: (c) $E=0.5$ and $J=0.5$.} \label{ps_tail2}
801: \end{figure}
802: \begin{figure}[ht]
803: \begin{center}
804: % \includegraphics[width=8cm,height=4.9cm]{fig/lfig13a.eps}
805: \includegraphics[width=8cm,height=4.9cm]{lfig13a.eps}
806: \end{center}
807: \begin{center}
808: % \includegraphics[width=8cm,height=4.9cm]{fig/lfig13b.eps}
809: \includegraphics[width=8cm,height=4.9cm]{lfig13b.eps}
810: \end{center}
811: \begin{center}
812: % \includegraphics[width=8cm,height=4.9cm]{fig/lfig13c.eps}
813: \includegraphics[width=8cm,height=4.9cm]{lfig13c.eps}
814: \end{center}
815: \caption{$F(s=2,J)$ vs. $J$ for ${\it L}=8, 12, 16, 20, 24$, (a)
816: $E=0$; (b) $E=0.02$; (c) $E=0.5$.} \label{ps_tail3}
817: \end{figure}
818: The overall shape of these curves has some common features. All
819: curves have a vanishing value when $s$ goes to zero. At small $s$
820: they increase with $s$ and reach a peak at some intermediate $s$.
821: Then they decrease monotonically to zero with increasing $s$.
822: These features are the same as those for
823: $P_{GUE}(s)$\cite{metzler}. Thus they all look like to be
824: $P_{GUE}(s)$ at first glance. This raises the question of how to
825: distinguish numerically extended states from localized states. As
826: a simple way, it is natural to expect that $P(s)$ for an extended
827: state approaches $P_{GUE}(s)$ while that for a localized state
828: should deviate from $P_{GUE}(s)$ as ${\it L}$ increases. Indeed,
829: as ${\it L}$ increases, curves in each sub-figure of
830: Fig.\ref{ps_global1} approach $P_{GUE}(s)$ while those in
831: Fig.\ref{ps_global2} show the opposite tendency. Thus we can use
832: this different {\it tendency} of $P(s)$ to distinguish extended
833: states from localized states. We shall show quantitatively that
834: such opposite tendency for extended states and localized states
835: exists also in several other characteristic quantities.
836:
837: %(2) I_0=\int s^2P(s)ds/2
838: Let us first consider a characteristic quantity $I_0$ defined by
839: $I_0=\int s^2P(s)ds/2$. It is commonly used to characterize the
840: overall shape of $P(s)$ and to examine the localization
841: property\cite{metzler}. It is well-known that $I_0=1$ for
842: localized states while $I_0<1$ for extended states\cite{mehta}.
843: Thus, the following simple criteria is employed: A state is
844: localized if its $I_0$ increases and approaches $1$ as ${\it L}$
845: increases. Otherwise, it is extended. Curves in Fig.\ref{data} are
846: $I_0$ vs. mixing strength $J$ for $E=0$ (a); 0.02 (b); and 0.5 (c)
847: for ${\it L}=8, 12, 16, 20, 24$. Fig.\ref{data}(b) shows that the
848: state of $E=0.02$ is localized at zero mixing and extended at
849: small $J$. Then it is localized again after $J$ passes a
850: particular $J_c$ where $I_0$ of different ${\it L}$ cross. For the
851: state of $E=0$ at the lower band center shown in
852: Fig.\ref{data}(a), it is extended at zero mixing. Then, it shows
853: the same feature as the state of $E=0.02$ at small and large $J$.
854: Fig.\ref{data}(c) shows that state of $E=0.5$ is always localized
855: at small $J$ and extended only for large $J(>1)$ where all curves
856: of different system sizes tend to merge together.
857: %(3) P_I(s_1,s_2)=\int_{s_1}^{s_2}P(s')ds' at small s
858:
859: A fundamental difference between $P(s)$ for a localized state and
860: that for an extended states is its behavior at small $s$. As $s$
861: goes to zero, $P(s)$ vanishes for extended states due to
862: level-repulsion while it approaches $1$ for localized states due
863: to level-aggregation\cite{mehta}. Thus we need to consider the
864: behavior of $P(s)$ at small $s$ for further test of the results in
865: the last paragraph. It is convenient to consider a function of
866: integrated level-spacing distribution at small $s$ defined by
867: $I_P(s)=\int_{0}^{s}P(s^{\prime})ds^{\prime}$. $I_P(s)$ is the
868: fraction of level-spacing smaller than $s$. Although $P(s)$ in
869: most cases of our numerical results is close to the GUE
870: distribution, level-repulsion of extended states and
871: level-aggregation of localized states should still be expected at
872: small $s$. This leads to the following criteria: $I_P(s)$ at small
873: $s$ should increase (decrease) with ${\it L}$ for localized
874: (extended) states. Thus the behavior of $I_P(s)$ at small $s$ can
875: serve as another way of distinguishing extended states from
876: localized ones. Fig.\ref{ps_small1} shows $I_P(s)$ for
877: $(E=0,J=0.1)$ (a), $(E=0.02,J=0.1)$ (b) and $(E=0.5,J=1.5)$ (c)
878: for ${\it L}=8, 12, 16, 20,24$ and comparison with $I_P(s)$ of
879: $P_{GUE}(s)$. Fig.\ref{ps_small2} is for $(E=0,J=0.7)$ (a),
880: $(E=0.02,J=0.7)$ (b) and $(E=0.5,J=0.5)$ (c). One can see clearly
881: that states in Fig.\ref{ps_small1} show the feature of extended
882: states while states in Fig.\ref{ps_small2} are localized. In order
883: to examine an electronic state of fixed energy in the whole range
884: of mixing, we consider $I_P(s=0.5)$, the fraction of the
885: level-spacings less than $0.5$. We plot the results of
886: $I_P(s=0.5)$ vs. $J$ at $E=0,0.02$ and $0.5$ for {\it L}=8, 12,
887: 16, 20, 24 in Fig.\ref{ps_small3}. Similar to the criteria for
888: $I_P(s)$, we use the following ones. If $I_P(0.5)$ of a state
889: increases with ${\it L}$, the state is localized. Otherwise, it is
890: extended. According to this criteria, curves in
891: Fig.\ref{ps_small3} give essentially the same results as those
892: obtained from $I_0$ in Fig.\ref{data}.
893:
894: Let us now turn to the region of large $s$. Since $P_{GUE}(s)$
895: decays faster than $P_{PE}(s)$ at large $s$, the behavior of
896: $P(s)$ in this region can also be used to differentiate extended
897: and localized states. In this region it is convenient to consider
898: another quantity defined by $F(s)=\int_{s}^{\infty}
899: P(s)ds=1-I_P(s)$. The meaning of $F(s)$ is the integrated fraction
900: of level-spacings larger than $s$. Since $F(s)$ of $P_{GUE}(s)$ is
901: less than that of $P_{PE}(s)$ at large $s$, we may expect that
902: $F(s)$ at larger $s$ decreases (increases) with ${\it L}$ for
903: extended (localized) states. Fig.\ref{ps_tail1} is $F(s)$ for
904: $P_{GUE}(s)$ and that for $(E=0,J=0.1)$ (a), $(E=0.02,J=0.1)$ (b),
905: and $(E=0.5,J=1.5)$ (c) with ${\it L}=8, 12, 16, 20, 24$.
906: Fig.\ref{ps_tail2} is $F(s)$ vs. $s$ for $(E=0, J=0.7)$ (a),
907: $(E=0.02,J=0.7)$ (b), and $(E=0.5,J=0.5)$ (c). In view of
908: Fig.\ref{ps_small1} and Fig.\ref{ps_small2}, it is clear that the
909: results of $F(s)$ coincide with those of $I_P (s)$ concerning the
910: localization property. We also calculate $F(s=2)$ for fixed energy
911: states in the whole range of mixing. As shown above, the same
912: criteria as that for $I_0$ and $I_P(s=0.5)$ can be employed. The
913: curves of $F(s=2)$ vs. $J$ are plotted in Fig.\ref{ps_tail3} for
914: $E=0$ (a), $E=0.02$ (b) and $E=0.5$ (c). One can see that they are
915: consistent with the results of $I_0$ (Fig.\ref{data}) and
916: $I_P(s=0.5)$ (Fig.\ref{ps_small3}).
917:
918:
919: \subsection{Discussion of finite-size effect}
920:
921: In this subsection, we shall consider possible influence of
922: finite-size effect on our numerical results. It is known that
923: localization lengths for 2D models can exceed $3\times10^4$
924: lattice spacings, a thousand times larger than the maximum lattice
925: size ${\it L}=24$ in our numerical calculation. Therefore, one
926: should worry about finite-size effect, and question the
927: unsuitability of our simple criteria for localization property. In
928: the following discussion, we shall only examine the results of the
929: quantity $I_0$. Essentially the same discussions can be made for
930: other quantities such as $I_P(s)$ and $F(s)$.
931:
932: Let us first consider the results for $E=0$ and $E=0.02$ in
933: Fig.\ref{data}. The two cases are quite similar. Curves for
934: different ${\it L}$ cross at a single point $J_c$. As ${\it L}$
935: increases, $I_0$ decreases and approaches the value for extended
936: states when $J<J_c$, while it increases and approaches the value
937: for localized states when $J>J_c$. A straightforward way of
938: interpreting this behavior is that the state exhibits a transition
939: from an extended state in $J<J_c$ to a localized one in $J>J_c$.
940: The correlation length diverges at critical point $J=J_c$ and
941: decreases sharply when $J$ is slightly away from $J_c$. (In a
942: metallic phase the correlation length is small while the
943: localization length is divergent.) Fluctuations at length scale of
944: order of the correlation length cause $I_0$ to deviate from its
945: thermodynamic-limit value, but $I_0$ approaches its
946: thermodynamic-limit values for both $J<J_c$ and $J>J_c$ as lattice
947: size increases. This is a natural explanation of the results. The
948: only finite-size effect in this explanation is rounding behavior
949: in a region close to the critical point $J=J_c$. This region is
950: normally very narrow because of the sharp drop of the correlation
951: length near $J=J_c$.
952:
953: Another possible interpretation is to assume that states are
954: always localized in the whole region except at $J=J_c$. In this
955: case, the localization length equals to the correlation length.
956: Then, in order to explain the above behavior, we have to assume
957: the following features of the localization length: (a) In the
958: region $J<J_c$, the ratio of localization length $\xi({\it L})$
959: and system size ${\it L}$ should increase with ${\it L}$ for small
960: ${\it L}$, leading to a false metallic behavior, while this ratio
961: decreases with ${\it L}$ for large ${\it L}$, a behavior for the
962: localized state; (b) in the region $J>J_c$, $\xi({\it L})/{\it L}$
963: should always decrease with ${\it L}$. In principle, one cannot
964: rule out this possibility without doing calculations for large
965: lattice sizes. However, (a) is too strange to justify.
966: Furthermore, there is no reason to believe why (a) occurs in
967: region $J<J_c$, but not in region $J>J_c$.
968:
969: For $E=0.5$ in Fig.\ref{data}, localized behavior is clearly
970: seen for small and intermediate $J$ while the curves of
971: different system size tend to merge at large $J$.
972: Here the finite-size effect should be considered seriously.
973: There are two ways to explain the merging behavior. One is
974: that a line of critical points exists for large $J$ where the
975: correlation length is always divergent. The other is that the
976: correlation length at the thermodynamic limit is very large
977: yet {\it finite} and the merging behavior is just a
978: finite-size effect. Unlike the case of $E=0$ and $E=0.02$,
979: both explanations in this case are reasonable. The only way
980: to make an unambiguous conclusion is to do calculations
981: for large sizes.
982:
983: However, we can propose a physical picture for the existence of
984: new extended states at $E\sim0.5$ in the case of strong interband
985: mixing. Assuming that the intra-band tunnelling at nodes are
986: negligibly weak for states of $E\sim0.5$, we saw already from Fig.
987: \ref{network}(a) that the maximum interband mixing
988: ($\sin\theta=1$) delocalizes the state, which is localized at zero
989: interband mixing. If one views $p=\sin^2 \theta$ as connection
990: probability of two neighboring loops of opposite chirality, our
991: two-channel model without intra-band tunnellings at nodes is
992: analogous to a bond-percolation problem. It is well-known that a
993: percolation cluster exists at $p\ge p_c=1/2$ or $J\ge J_c=1$ for a
994: square lattice\cite{stauff}. Therefore, an extended state is
995: formed by strong mixing. One hopes that the intra-band tunnellings
996: at nodes will only modify the threshold value of the mixing
997: strength. If this picture is correct, the finite-size effect only
998: affects our efforts to determine the accurate value of $J_c$ yet
999: it does not influence the existence of such a critical point.
1000:
1001: It should be noted that all above discussions are based on the
1002: single parameter scaling argument. Suppose that the region of
1003: extended states in thermodynamic limit is vanishing instead of
1004: remaining finite, our above numerical results can also be
1005: explained by introducing irrelevant length scales and considering
1006: their corrections to scaling, as has been pointed out by
1007: Pruisken\cite{pruisken2} and Huckstein\cite{huckstein}. Thus both
1008: our picture of finite region of extended states and the idea of
1009: corrections to scaling are alternative explanations for
1010: non-scaling behaviors. Therefore, calculations for larger sizes of
1011: system are necessary to make unambiguous conclusion for the width
1012: of extended state region in thermodynamic limit. In a recent
1013: theoretical study, Pruisken et. al.\cite{pruisken3} have developed
1014: a microscopic theory based on nonlinear $\sigma$ model to explain
1015: the non-scaling behaviors within the assumption of a single
1016: critical point. However, in their consideration of the case of
1017: long-range correlated disorder which corresponds to the network
1018: model, interband mixing, the physical reason for the existence of
1019: finite region of extended states in our picture, is neglected.
1020: Therefore, their study does not rule out the possibility of finite
1021: region of extended states.
1022:
1023: \subsection{Discussion of localization property}
1024:
1025: In the past two subsections, we analyzed the global shape of
1026: $P(s)$ and its behavior at small and large $s$ by considering
1027: $I_0$, $I_P(s)$ and $F(s)$, respectively, and discuss possible
1028: influence of finite-size effect. Analysis of all these quantities
1029: leads to essentially the same conclusion concerning the
1030: localization property, as follows. For zero interband mixing, only
1031: states at the two LB centers are extended. In the presence of
1032: interband mixing, new extended states emerge. States near the LB
1033: centers,--- i.e., $E\sim0$,--- are delocalized by weak interband
1034: mixing and localized by strong mixing, with a transition point at
1035: some intermediate mixing $J_c$. For states near the region between
1036: two LBs,---i.e., $E\sim0.5$,--- they are localized at both weak
1037: and intermediate mixing and delocalized by strong mixing.
1038:
1039: \begin{figure}[ht]
1040: \begin{center}
1041: % \includegraphics[width=8cm,height=4.9cm]{fig/lfig14a.eps}
1042: \includegraphics[width=8cm,height=4.9cm]{lfig14a.eps}
1043: \end{center}
1044: \begin{center}
1045: % \includegraphics[width=8cm,height=4.9cm]{fig/lfig14b.eps}
1046: \includegraphics[width=8cm,height=4.9cm]{lfig14b.eps}
1047: \end{center}
1048: \begin{center}
1049: % \includegraphics[width=8cm,height=4.9cm]{fig/lfig14c.eps}
1050: \includegraphics[width=8cm,height=4.9cm]{lfig14c.eps}
1051: \end{center}
1052: \caption{$I_0$ vs. $E$ for ${\it L}=8, 12, 16, 20$, (a) $J=0.1$;
1053: (b) $J=0.7$; (c) $J=1.5$.} \label{fixed_J}
1054: \end{figure}
1055:
1056: %added paragraphs-------------------------------------------------
1057: In order to show explicitly the existence of a narrow band of
1058: extended states and its evolution with increasing mixing, curves
1059: of $I_0$ vs. $E$ are plotted for three values of $J$ in
1060: Fig.\ref{fixed_J}. A band of extended states is formed around the
1061: LB center $E\in[0,0.1]$ for $J=0.1$. When $J$ is increased to an
1062: intermediate value $0.7$, this band of extended states is lifted
1063: up to $E\in[0.8,1.6]$. For strong mixing, it is further shifted to
1064: $E\in[0.4,0.5]$. Thus the band of extended states in the lower LB
1065: tends to float up in energy while that of the upper tends to dive
1066: down in energy as mixing strength increases. The two bands should
1067: finally meet at the middle energy region in the case of strong
1068: mixing.
1069:
1070: The above results are restricted to the case of two LBs. However,
1071: there are infinite number of LBs in a realistic system. In order
1072: to conjecture the situation when infinite number of LBs is
1073: incorporated in our result, we take into account the
1074: float-up-merge picture proposed by Sheng et. al.\cite{sheng2}. We
1075: shall expect that a narrow extended band appears in each LB center
1076: for weak interband mixing. Increasing mixing, i.e., increasing
1077: disorders or decreasing magnetic field, the extended band in the
1078: lowest LB floats up and finally merges with that in the second
1079: lowest LB. Then, this extended band will further shift up and
1080: merge into that in the third lowest LB, and so on so forth.
1081: %added paragraphs---------------------------------------------------
1082:
1083: To express our numerical results in the plane of energy and
1084: interband mixing, a topological phase diagram shown in
1085: Fig.\ref{phase}(a) is obtained. In the absence of interband
1086: mixing, only the singular energy level at each LB center is
1087: extended. In the presence of interband mixing of opposite
1088: chirality, there are two regimes. At weak mixing, each of the
1089: extended states broadens into a narrow band of extended states
1090: near the LB centers. With increased mixing, the extended states in
1091: the lowest LB shift from the LB center(see Fig.\ref{fixed_J}).
1092: These extended states will eventually merge with those from the
1093: higher LBs. This shifting of bands of extended states is similar
1094: to the shifting of single extended states at LB centers observed
1095: in previous studies \cite{kivelson} where emerging of extended
1096: bands is missing. At strong mixing, a band of extended states
1097: exists between neighboring LBs where all states are localized
1098: without the mixing.
1099:
1100: %3. Discussions
1101: Let us look at the consequences of the above results. For weakly
1102: disordered systems in IQHE regime, the Landau gap is larger than
1103: the LB bandwidth. Thus there is no overlap between adjacent LBs.
1104: According to the semiclassical picture, electronic states between
1105: the two adjacent LBs should be from either the upper or the lower
1106: bands with the same chirality in this case. It means that no
1107: interband mixing occurs and there is only one extended state in
1108: each LB. This may explain why scaling behaviors were observed for
1109: plateau transitions in early experiments on clean samples.
1110: Interband mixing occurs when the Landau gap is less than the LB
1111: bandwidth. Systems of relatively strong disorders in IQHE regime
1112: should correspond to this case. As the single extended state at
1113: each LB center broadens into a narrow extended band, a narrow
1114: metallic phase emerges between two neighboring IQHE phases. Thus
1115: each plateau transition contains two consecutive quantum phase
1116: transitions for strongly disordered systems. The bands of extended
1117: states will merge together in strong mixing. This strong mixing
1118: regime corresponds to the case when the Landau gap is much smaller
1119: than the bandwidth. Since the Landau gap is proportional to the
1120: magnetic field, the disordered system should always enter the
1121: strong mixing regime before it reaches the weak field insulating
1122: phase, regardless of how weak the disorders are. In terms of QH
1123: plateau transitions, a direct transition occurs because a narrow
1124: metallic phase exists between two QH phases in a weak field. Thus,
1125: we propose that a direct transition from an IQHE phase to the
1126: insulating phase at weak field is realized by passing through a
1127: metallic phase, and it should hold for both weak and strong
1128: disordered systems.
1129: \begin{figure}[ht]
1130: \begin{center}
1131: \includegraphics[width=4cm,height=4cm]{lfig15a.eps}
1132: \includegraphics[width=4cm,height=4cm]{lfig15b.eps}
1133: % \includegraphics[width=4cm,height=4cm]{fig/lfig15a.eps}
1134: % \includegraphics[width=4cm,height=4cm]{fig/lfig15b.eps}
1135: \end{center}
1136: \caption{(a) Topological phase diagram of electron localization in
1137: $E-J$ plane. The shadowed regime is for extended states (metallic
1138: phase). (b) Topological QH phase diagram in $W-B$ plane. $W$
1139: stands for the disorder strength, and $B$ for the magnetic field.
1140: The shadowed regime is for the metallic phase. The area indicated
1141: by the symbol $n$ is the $n$-plateau IQHE phase. The rest area is
1142: the insulating phase.} \label{phase}
1143: \end{figure}
1144:
1145: %4. The phase diagram in disorder vs magnetic field (Figure)
1146: Plotting above results in the plane of disorder and the magnetic
1147: field, we obtain a new topological QH phase diagram as shown in
1148: Fig.\ref{phase}(b). This is similar to the empirical diagram
1149: obtained experimentally in Ref. 15. The origin ($W=0, \ B=0$) is a
1150: singular point. According to the weak localization
1151: theory\cite{abrahams}, no extended state exists at this point.
1152: Differing from existing theories, there exists a narrow metallic
1153: phase between two adjacent IQHE phases and between an IQHE phase
1154: and an insulating phase.
1155:
1156: \subsection{Comparison with previous studies}
1157: In this subsection, we shall compare our results with previous
1158: studies. We shall show that the new phase diagram in
1159: Fig.\ref{phase} is consistent with the non-scaling
1160: experiments\cite{hilke} where samples are relatively dirty, and
1161: interband mixing is strong, corresponding to a process along line
1162: $a$ in Fig.\ref{phase}(b). The system undergoes two quantum phase
1163: transitions each time when it moves from the QH insulating phase
1164: to IQHE phase of $n=1$ and back to the weak field insulating phase
1165: as the magnetic field decreases.
1166: \begin{figure}[ht]
1167: \begin{center}
1168: %\includegraphics[width=8cm,height=5cm]{fig/lfig16.eps}
1169: \includegraphics[width=8cm,height=5cm]{lfig16.eps}
1170: \end{center}
1171: \caption{\label{experiment} The original experimental data of
1172: $\ln[R_{xx}(f,T)/R_{xx}(0.647,40mK)]$ in Ref. 18.
1173: %\cite{hilke}.
1174: $f$ is the filling factor of LBs and $T$ is the temperature.}
1175: \end{figure}
1176: To verify this claim, we analyzed the original experimental data in
1177: Ref. 18 according to the assumption of two quantum %\cite{hilke}
1178: phase transition points. The experiment data of the logarithm of
1179: the longitudinal resistance $ln[R_{xx}(f,T)]$ are shown in
1180: Fig.\ref{experiment} where $f$ is the filling factor of LBs and
1181: $T$ is the temperature. According to the theory of continuous
1182: transitions, one should obtain
1183: \begin{equation}
1184: ln[R_{xx}(\nu,T)]=F_1(S_1(f)/T)
1185: \end{equation}
1186: with $S_1(f)\sim(f_{c1}-f)^{z_1\nu_1}$ for the region of
1187: $f<f_{c1}$ while
1188: \begin{equation}
1189: ln[R_{xx}(\nu,T)]=F_2(S_2(f)/T)
1190: \end{equation}
1191: with $S_2(f)\sim(f-f_{c2})^{z_2\nu_2}$ for the region of
1192: $f>f_{c2}$. Previous theories predict one single critical point,
1193: ---i.e., $f_{c1}=f_{c2}$ and $z_1\nu_1=z_2\nu_2$. But our results
1194: suggest two distinct critical points. By standard scaling
1195: analysis, two good scaling behaviors are obtained for two close
1196: critical filling factors of $f_{c1}=0.6453$ and $f_{c2}=0.6477$ as
1197: shown in Fig.\ref{twopoint}. The critical exponents in both the
1198: left side and the right side of the transition region are equal to
1199: the value $z\nu=2.33\pm0.01$. On the other hand, the fit for one
1200: single critical point fails. Fig.\ref{onepoint} shows the result
1201: of a single critical point at $\nu_c=0.646$. It is the best
1202: fitting result for a single critical point if we require that the
1203: two critical exponents are approximately equal and the scaling law
1204: is optimally obeyed. The two critical exponents are
1205: $z_1\nu_1=2.58\pm0.02$ and $z_2 \nu_2 = 2.60 \pm 0.02$, deviating
1206: from the theoretical results $z\nu \sim 2.33$. One can also see
1207: clearly systematic deviations from the scaling law in the region
1208: close to the critical point at both sides in Fig.\ref{onepoint}.
1209: This implies that the transition process is governed by two
1210: separated critical points instead of one. The regime between the
1211: two critical points should correspond to the metallic phase.
1212: %Our fitting shows that the width of this regime is about
1213: %$5\times10^{-3} tesla$ while the value of the magnetic field was
1214: %increased by $1\sim2\times10^{-3} tesla$ each time in the
1215: %experiments. This may explain why the metallic phase was
1216: %overlooked.
1217: \begin{figure}
1218: \begin{center}
1219: %\includegraphics[width=8cm,height=5cm]{fig/lfig17.eps}
1220: \includegraphics[width=8cm,height=5cm]{lfig17.eps}
1221: \end{center}
1222: \caption{The fitting result of two critical points at the left and
1223: the right side. The two straight lines show coincidence with the
1224: scaling law. The critical filling factors are $f_{c1}=0.6453$ and
1225: $f_{c2}=0.6477$. The two critical exponents are equal to the value
1226: $z\nu=2.33\pm0.01$.} \label{twopoint}
1227: \end{figure}
1228: \begin{figure}
1229: \begin{center}
1230: %\includegraphics[width=8cm,height=5cm]{fig/lfig18.eps}
1231: \includegraphics[width=8cm,height=5cm]{lfig18.eps}
1232: \end{center}
1233: \caption{The best fitting result of one single critical point at
1234: the left and right side. The critical filling factor is
1235: $f_{c}=0.646$. The two straight lines illustrate systematic
1236: deviations from the scaling law at regions close to the critical
1237: point. The average values of the two critical exponents are
1238: $z_1\nu_1=2.58\pm0.02$ and $z_2\nu_2=2.60\pm0.02$, respectively.}
1239: \label{onepoint}
1240: \end{figure}
1241:
1242: It is worth noting that there is another puzzle in the non-scaling
1243: experiment which may be solved by our two-critical-point picture.
1244: As an example, we consider the experimental data for the transition
1245: between the QH insulating phase and the $n=1$ IQHE phase.
1246: It was shown\cite{hilke} that the logarithm of the longitudinal
1247: resistance $ln[R_{xx}(f,T)]$ can be fitted by a linear function
1248: of the filling factor $f$ (see Fig.\ref{experiment})
1249: \begin{equation}
1250: ln[R_{xx}(f,T)]=ln[R_{xx}(f_c,T)]-(f-f_c)/(\alpha+\beta T)]
1251: \end{equation}
1252: where $\alpha$ and $\beta$ are positive constants, $f_c$ is the
1253: filling factor where curves of different temperature $T$ cross
1254: approximately. Since $\alpha$ is non-zero\cite{hilke}, it leads to
1255: the conclusion that $R_{xx}(f,T)$ at the limit of $T=0$ remains
1256: finite {\it for every $f$}. This is puzzling because it is
1257: inconsistent with the theoretical requirement that
1258: $R_{xx}(T=0)=\infty$ in the QH insulating phase, i.e., $f<f_c$,
1259: and $R_{xx}(T=0)=0$ in the $n=1$ IQHE phase, i.e., $f>f_c$. This
1260: puzzle may be solved as follows. Combine the linear relationship
1261: between $ln[R_{xx}(f,T)]$ and $f$ for fixed $T$ with our picture
1262: of two critical points $f_{c1}<f_{c2}$, we expect
1263: \begin{equation}
1264: ln[R_{xx}(f,T)]=ln[R_{xx}(f_{c1},T)]-(f-f_{c1})/(A_1 T^{z\nu})
1265: \end{equation}
1266: in the QH insulating phase, i.e., $f<f_{c1}$, while
1267: \begin{equation}
1268: ln[R_{xx}(f,T)]=ln[R_{xx}(f_{c2},T)]-(f-f_{c2})/(A_2 T^{z\nu})
1269: \end{equation}
1270: in the n=1 IQHE phase, i.e., $f>f_{c2}$, where $A_1$ and $A_2$ are
1271: positive constants, and $z$ and $\nu$ are critical exponents. It
1272: is clear that both $R_{xx}(f,T=0)=\infty$ in $f<f_{c1}$ and
1273: $R_{xx}(f,T=0)=0$ in $f>f_{c2}$ are recovered. While a finite
1274: value of $R_{xx}(f,T=0)$ in the region $f_{c1}<f<f_{c2}$ is
1275: consistent with our prediction of a metallic phase between the two
1276: critical points.
1277:
1278: Scaling behaviors of plateau-plateau transitions are observed in
1279: recent experiments\cite{hohls1,hohls2,hohls3,ponomarenko,li}. At a
1280: first glance, it seems that these results conflict with both
1281: non-scaling experiments and our numerical results. However, there
1282: are two important differences between recent scaling experiments
1283: and non-scaling experiments. One is that some scaling experiments
1284: \cite{hohls1,hohls2,hohls3} were done deep inside QH plateau, or
1285: far from the plateau transition point, while the non-scaling
1286: behavior was obtained by using data very close to the transition
1287: points. In the non-scaling experiments\cite{hilke}, it is known
1288: that data not too close to a transition point follow a scaling
1289: law. The other is that the samples used in all scaling
1290: experiments\cite{hohls1,hohls2,hohls3,ponomarenko,li} are clean
1291: with very high mobility while non-scaling behavior was observed in
1292: relative dirty samples\cite{hilke,shash,baba}. In fact, the
1293: mobility in recent scaling experiments is at least one order of
1294: magnitude larger than that in early non-scaling experiments. This
1295: means that the scaling and non-scaling experiments correspond to
1296: regions of vanishing and relatively strong inter-band mixing,
1297: respectively. Thus, there is no real conflict between the recent
1298: scaling experiments and those early non-scaling experiments. Since
1299: our model is valid only when inter-band mixing is considerable,
1300: there is also no real conflict between recent scaling experiments
1301: and our numerical results.
1302:
1303: Two-channel CC model has been used to simulate two degenerated or
1304: nearly-degenerated spin-resolved Landau subbands with strong
1305: interband mixing by Wang {\it et al.}\cite{wang}. They found
1306: two distinct critical points which were related to mixing-induced
1307: repulsion\cite{wang}. For the nearly-degenerated case, they did
1308: not consider the states between the two LB centers. For the
1309: degenerated case, their study could not discern whether the states
1310: between the two critical points are extended or localized.
1311: According to our results, a band of extended states is formed
1312: around the degenerated LB center for the degenerated case at
1313: strong mixing, while for the nearly-degenerated case electronic
1314: states between the two LB centers are delocalized by strong mixing.
1315: Thus our results are consistent with their results.
1316:
1317: One should also notice that two types of metallic phases have been
1318: studied extensively in the QH system. One is the composite Fermion
1319: state at the half-filling in the lowest Landau level (LL) and the
1320: other is the stripe state at the half-filled higher LLs. These
1321: states are formed by the Coulomb interaction effect in the high
1322: mobility samples. They are different from our metallic phase
1323: caused by level mixing. Although we have not considered the
1324: electron-electron interactions in our study, there is no reason
1325: why the delocalization effect of level mixing will be diminished
1326: by the Coulomb interaction. Of course, the interaction could
1327: lead to a level shift, thus it may modify the band width.
1328:
1329: \section{Conclusions}
1330: % V. Discussions and the conclusion
1331: % 1) The picture of two-step transition
1332: % 2) The conclusion
1333: In conclusion, we find by numerical calculations within the
1334: network model that it is possible that the single extended state
1335: at each LB center in the absence of interband mixing may broaden
1336: into a narrow band of extended states when the effect of mixing of
1337: states of {\it opposite chirality} is taken into account. We also
1338: provide a physical picture to show how the mixing of states of
1339: {\it opposite chirality} may possibly lead to the existence of
1340: extended state bands. Based on above results, we propose a new
1341: phase diagram in which a narrow metallic phase exists between two
1342: neighboring IQHE phases and between an IQHE phase and an
1343: insulating phase. This new phase diagram is consistent with
1344: non-scaling behaviors observed in recent experiments. A standard
1345: scaling analysis on nonscaling experiment data \cite{hilke}
1346: supports our results. However, due to finite-size effects, our
1347: numerical results can also be explained based on the assumption of
1348: a single critical point. Thus further study on large system size
1349: is still needed to conclude whether there are extended state bands
1350: in quantum Hall systems in thermodynamic limit.
1351:
1352: \begin{acknowledgments}
1353: This work was substantially supported by a grant from the Research
1354: Grant Council of HKSAR, China (Project No. HKUST6153/99P, and
1355: HKUST6149/00P). GX acknowledges the support of CNSF under grant
1356: No. 10347101 and the Grant from Beijing Normal University. GX and
1357: YW also acknowledge the support of CNSF under grant No. 90103024.
1358: \end{acknowledgments}
1359:
1360: \appendix*
1361: \section{}
1362: \begin{figure}[ht]
1363: \begin{center}
1364: % \includegraphics[height=7cm, width=8cm]{fig/lfig19.eps}
1365: \includegraphics[height=7cm, width=8cm]{lfig19.eps}
1366: \end{center}
1367: \caption{A two-channel network model of $2\times2$ nodes with
1368: periodic boundaries along both directions. $Z$-s are the
1369: wavefunction amplitudes on links. The notations are as follows. H
1370: and V stand for horizontal and vertical links, respectively. $u$
1371: ($l$) is for the upper (lower) LB. $\hat{S}_i$ are SO(4) matrices
1372: describing tunnelling at nodes, and $\hat{M}_i$ are $U(2)$
1373: matrices for interband mixing.} \label{example}
1374: \end{figure}
1375:
1376: In this appendix, we explicitly construct the evolution matrix
1377: $\hat U$ for a $2\times 2$ two-channel CC-network model as shown
1378: in Fig.\ref{example}. Periodical boundary conditions in both
1379: directions are imposed as explained in section III. $Z$-s are the
1380: wavefunction amplitudes on links. The notations are as follows. H
1381: and V stand for horizontal and vertical links, respectively. $u$
1382: ($l$) is for the upper (lower) LB. $\hat{S}_i$ are SO(4) matrices
1383: defined in Eq.(\ref{smatrix}) describing the tunnelling on nodes,
1384: and $\hat{M}_i$ are $U(2)$ matrices defined in Eqs. \ref{mmatrix1}
1385: and \ref{mmatrix2} describing interband mixing. From
1386: Fig.\ref{example} we can obtain
1387: \begin{equation}
1388: \left (
1389: \begin{array}{l}
1390: Z_{u,H}^{1}(t+1) \\ Z_{l,H}^{1}(t+1) \\
1391: Z_{l,H}^{2}(t+1) \\ Z_{u,H}^{2}(t+1)
1392: \end{array}
1393: \right ) =
1394: \hat{H}_1
1395: \left (
1396: \begin{array}{l}
1397: Z_{l,V}^{1}(t)\\ Z_{u,V}^{1}(t)\\ Z_{u,V}^{2}(t) \\
1398: Z_{l,V}^{2}(t)
1399: \end{array}
1400: \right )
1401: \end{equation}
1402: \begin{equation}
1403: \left (
1404: \begin{array}{l}
1405: Z_{u,H}^{3}(t+1) \\ Z_{l,H}^{3}(t+1) \\
1406: Z_{l,H}^{4}(t+1) \\ Z_{u,H}^{4}(t+1)
1407: \end{array}
1408: \right ) =
1409: \hat{H}_2
1410: \left (
1411: \begin{array}{l}
1412: Z_{l,V}^{3}(t) \\ Z_{u,V}^{3}(t) \\ Z_{u,V}^{4}(t)
1413: \\ Z_{l,V}^{4}(t)
1414: \end{array}
1415: \right )
1416: \end{equation}
1417: \begin{equation}
1418: \left (
1419: \begin{array}{l}
1420: Z_{l,V}^{1}(t+1) \\ Z_{u,V}^{1}(t+1) \\
1421: Z_{u,V}^{2}(t+1) \\ Z_{l,V}^{2}(t+1)
1422: \end{array}
1423: \right ) =
1424: \hat{H}_3
1425: \left (
1426: \begin{array}{l}
1427: Z_{u,H}^{3}(t) \\ Z_{l,H}^{3}(t) \\ Z_{l,H}^{4}(t)
1428: \\ Z_{u,H}^{4}(t)
1429: \end{array}
1430: \right )
1431: \end{equation}
1432: \begin{equation}
1433: \left (
1434: \begin{array}{l}
1435: Z_{l,V}^{3}(t+1) \\ Z_{u,V}^{3}(t+1) \\
1436: Z_{u,V}^{4}(t+1) \\ Z_{l,V}^{4}(t+1)
1437: \end{array}
1438: \right ) =
1439: \hat{H}_4
1440: \left (
1441: \begin{array}{l}
1442: Z_{u,H}^{1}(t) \\ Z_{v,H}^{1}(t) \\ Z_{l,H}^{2}(t)
1443: \\ Z_{u,H}^{2}(t)
1444: \end{array}
1445: \right ),
1446: \end{equation}
1447: with
1448: \begin{equation}
1449: \hat{H}_1=
1450: \left (
1451: \begin{array}{ll}
1452: \hat{1} & \hat{0} \\
1453: \hat{0} & \hat{M}_1
1454: \end{array}
1455: \right )
1456: \hat{S}_1; \quad
1457: \hat{H}_2=
1458: \left (
1459: \begin{array}{ll}
1460: \hat{M}_2 & \hat{0} \\
1461: \hat{0} & \hat{1}
1462: \end{array}
1463: \right )
1464: \hat{S}_2; \nonumber
1465: \end{equation}
1466: \begin{equation}
1467: \hat{H}_3=
1468: \left (
1469: \begin{array}{ll}
1470: \hat{1} & \hat{0} \\
1471: \hat{0} & \hat{M}_3
1472: \end{array}
1473: \right )
1474: \hat{S}_3; \quad
1475: \hat{H}_4=
1476: \left (
1477: \begin{array}{ll}
1478: \hat{M}_4 & \hat{0} \\
1479: \hat{0} & \hat{1}
1480: \end{array}
1481: \right )
1482: \hat{S}_4, \nonumber
1483: \end{equation}
1484: where $\hat{1}$ and $\hat{0}$ are the $2\times 2$ identity
1485: and zero matrices, respectively.
1486: If we define
1487: \begin{eqnarray}
1488: \phi_{H} = \left (
1489: \begin{array}{l}
1490: Z_{u,H}^{1}\\ Z_{l,H}^{1}\\
1491: Z_{l,H}^{2}\\ Z_{u,H}^{2}\\
1492: Z_{u,H}^{3}\\ Z_{l,H}^{3}\\
1493: Z_{l,H}^{4}\\ Z_{u,H}^{4}\\
1494: \end{array}
1495: \right ); \quad
1496: \phi_{V} = \left (
1497: \begin{array}{l}
1498: Z_{l,V}^{1}\\ Z_{u,V}^{1}\\
1499: Z_{u,V}^{2}\\ Z_{l,V}^{2}\\
1500: Z_{l,V}^{3}\\ Z_{u,V}^{3} \\
1501: Z_{u,V}^{4}\\ Z_{l,V}^{4} \\
1502: \end{array}
1503: \right ), \nonumber
1504: \end{eqnarray}
1505: then the evolution equation is
1506: \begin{equation}
1507: \left (
1508: \begin{array}{l}
1509: \phi_H(t+1) \\ \phi_V(t+1)
1510: \end{array}
1511: \right )
1512: =
1513: \hat{U}
1514: \left (
1515: \begin{array}{l}
1516: \phi_H(t) \\ \phi_V(t)
1517: \end{array}
1518: \right ).
1519: \end{equation}
1520: The evolution operator $\hat{U}$ is
1521: \begin{equation}
1522: \hat{U}
1523: =
1524: \left (
1525: \begin{array}{llll}
1526: \hat{0} & \hat{0} & \hat{0} & \hat{H}_1 \\
1527: \hat{0} & \hat{0} & \hat{H}_2 & \hat{0} \\
1528: \hat{H}_3 & \hat{0} & \hat{0} & \hat{0} \\
1529: \hat{0} & \hat{H}_4 & \hat{0} & \hat{0}
1530: \end{array}
1531: \right ),
1532: \end{equation}
1533: where $\hat{0}$ is the $4\times 4$ zero matrix. It has the
1534: structure of Eq.(\ref{evolution}).
1535:
1536: \begin{thebibliography}{}
1537: \bibitem{abrahams} E. Abrahams, P. W. Anderson, D. C. Licciardello,
1538: and T. V. Ramakrishnan, Phys. Rev. Lett. {\bf 42}, 673 (1979).
1539: \bibitem{pruisken} A. M. M. Pruisken, in `{\it The Quantum Hall Effect}',
1540: ed. by R. E. Prange and S. M. Girvin (Springer Verlag, New York, 1987).
1541: \bibitem{wei} H. P. Wei, D. C. Tsui, M. P. Paalanen, and A. M. M.
1542: Pruisken, Phy. Rev. Lett. {\bf 61}, 1294 (1988).
1543: \bibitem{pruisken2} A. M. M. Pruisken, Phys. Rev. Lett. {\bf 61},
1544: 1297(1988).
1545: \bibitem{kivelson} S. Kivelson, D. H. Lee and S. C. Zhang,
1546: Phys. Rev. B {\bf 46}, 2223 (1992).
1547: \bibitem{dzliu} D.Z. Liu, X.C. Xie, and Q. Niu,
1548: Phys. Rev. Lett. {\bf 76}, 975 (1996); X.C. Xie, et al.,
1549: Phys. Rev. B{\bf 54}, 4966 (1996).
1550: \bibitem{wang} Z. Q. Wang, D. H. Lee and X. G. Wen, Phys.
1551: Rev. Lett {\bf 72}, 2454 (1994)
1552: \bibitem{yang} K. Yang and R. N. Bhatt, Phys. Rev. Lett.
1553: {\bf 76}, 1316 (1996).
1554: \bibitem{shahar1} D. Shahar, D. C. Tsui, J. E. Cunningham,
1555: Phys. Rev. B {\bf 52}, R14372(1995); S.-H. Song, D. Shahar, D. C.
1556: Tsui, Y. H. Xie, Don Monroe, Phys. Rev. Lett. {\bf 78},
1557: 2200(1997).
1558: \bibitem{galstyan} A. G. Galstyan and M. E. Raikh,
1559: Phys. Rev. {\bf B56}, 1422 (1997).
1560: \bibitem{sheng1} D. N. Sheng and Z. Y. Weng, Phys. Rev. Lett.
1561: {\bf 78}, 318 (1997); {\it ibid} {\bf 80}, 580 (1998).
1562: \bibitem{sheng2} D. N. Sheng, Z. Y. Weng, and X. G. Wen, cond-mat/0003117.
1563: \bibitem{haldane} F. D. M. Haldane and K. Yang, Phys. Rev. Lett.
1564: {\bf 78}, 298 (1997).
1565: \bibitem{jiang} H. W. Jiang et al., Phys. Rev. Lett. {\bf 71}, 1439 (1993).
1566: \bibitem{tkwang} T. K. Wang et al., Phys. Rev. Lett. {\bf 72}, 709 (1994).
1567: \bibitem{glozman} I. Glozman, C. E. Johnson, and H. W. Jiang,
1568: Phys. Rev. Lett. {\bf 74}, 594 (1995).
1569: \bibitem{kravchenko} S. V. Kravchenko, W. E. Mason, J. E.
1570: Fureaux, V. M. Pudalov, Phys. Rev. Lett. {\bf 75}, 910 (1995).
1571: \bibitem{song} S.-H. Song, D. Shahar, D. C. Tsui, Y. H. Xie,
1572: Don Monroe, Phys. Rev. Lett. {\bf 78}, 2200 (1997).
1573: \bibitem{schaijk} R. T. F. van Schaijk, A. de Visser, S. M. Olsthoorn,
1574: H. P. Wei, and A. M. M. Pruisken, Phys. Rev. Lett. {\bf 84},
1575: 1567(2000).
1576: \bibitem{xrw} X. R. Wang, X. C. Xie, Q. Niu, and J.
1577: Jain, cond-mat/0008411.
1578: \bibitem{hilke} M. Hilke, D. Shahar, S. H. Song, D. C. Tsui,
1579: Y. H. Xie and D. Monroe, Phy. Rev. {\bf B56}, 15545 (1997);
1580: D. Shahar, M. Hilke, C. C. Li, D. C. Tsui,
1581: S. L. Sondhi, J. E. Cunningham and M. Razeghi,
1582: Solid State Commun. {\bf107}, 19 (1998).
1583: \bibitem{baba} N. Q. Balaban, U. Meirav, I. Bar-Joseph, Phys.
1584: Rev. Lett. {\bf 81}, 4967(1998).
1585: \bibitem{shash} A. A. Shashkin, V. T. Dolgopolov, and G. V.
1586: Kravchenko, Phys. Rev. B {\bf 49}, 14486(1994).
1587: \bibitem{khme} D. E. Khmelnitskii, Phys. Lett. {\bf 106A}, 182 (1984).
1588: \bibitem{laughlin} R. B. Laughlin, Phys. Rev. Lett. {\bf 52}, 2304 (1984).
1589: \bibitem{xiong} G. Xiong, S.D. Wang, Q. Niu, D.C. Tian, and
1590: X. R. Wang, Phys. Rev. Lett., {\bf 87}, 216802 (2001).
1591: \bibitem{pruisken3} A. M. M Pruisken, B. Skoric, and M. A.
1592: Baranov, Phys. Rev. {\bf B60}, 16838(1999).
1593: \bibitem{huckstein} B. Huckstein, Phys. Rev. Lett. {\bf 84},
1594: 3141(2000); Rev. Mod. Phys. 67,357(1995) .
1595: \bibitem{hohls1} F. Hohls, U. Zeitler, and R. J. Haug,
1596: Phys. Rev. Lett. {\bf 86}, 5124(2001).
1597: \bibitem{hohls2} F. Hohls, U. Zeitler, and R. J. Haug,
1598: Phys. Rev. Lett. {\bf 88}, 036802(2002).
1599: \bibitem{hohls3} F. Hohls, U. Zeitler, R. J. Haug, R. Meisels, K.
1600: Dybko, and F. Kuchar, Phys. Rev. Lett. {\bf 89}, 276801(2002).
1601: \bibitem{ponomarenko} Ponomarenko et. al., Physica {\bf E22},
1602: 236(2004).
1603: \bibitem{li} Wanli Li et. al., Phys. Rev. Lett. {\bf 94},
1604: 206807(2005).
1605: \bibitem{chalker} J. T. Chalker and P. D. Coddington,
1606: J.Phys. C: Solid State Phys.{\bf 21}, 2665 (1988).
1607: \bibitem{stauff} D. Stauff and A. Aharony,
1608: `{\it Introduction to Percolation Theory}'
1609: (Taylor and Francis, London, 1994).
1610: \bibitem{lee} D. K. K. Lee and J. T. Chalker,
1611: Phys. Rev. Lett {\bf 72}, 1510 (1994).
1612: \bibitem{kagalovsky} V. Kagalovsky, B. Horovitz and Y. Avishiai,
1613: Phys. Rev. {\bf B55}, 7761 (1997).
1614: \bibitem{fertig1} H. A. Fertig and B. I. Halperin,
1615: Phys. Rev. {\bf B36}, 7969 (1987).
1616: \bibitem{xie} X. C. Xie, X. R. Wang, and D. Z. Liu, Phys.
1617: Rev. Lett. {\bf 80}, 3563 (1998).
1618: \bibitem{kramer} I. Kh. Zharekeshev and B. Kramer, Phys. Rev.
1619: Lett. {\bf 79}, 717 (1997); H. Potempa and L. Schweitzer, Ann.
1620: Phys.(Leipzig) {\bf 8}, Spec. Issue, Si209 (1999); and references
1621: therein.
1622: \bibitem{schweitzer} M. Batch and L. Schweitzer,
1623: cond-mat/9608148; Y. Ono, T. Ohtsuki and B. Kramer,
1624: J. Phys. Soc. Jpn. {\bf 65}, 1734 (1996).
1625: \bibitem{mehta} M. L. Mehta,
1626: `{\it Random Matrices}' , 2nd ed. (Academic Press, 1991).
1627: \bibitem{klesse} R. Klesse and M. Metzler,
1628: Phys. Rev. Lett. {\bf 79}, 721 (1997).
1629: \bibitem{fertig2} H. A. Fertig, Phys. Rev. {\bf B38}, 996 (1988).
1630: \bibitem{metzler} M. Metzler and I. Varga,
1631: J. Phys. Soc. Jpn. {\bf 67}, 1856 (1998).
1632: \end{thebibliography}
1633: \end{document}
1634: